math0703223/PR1.tex
1: \documentclass[twoside,onecolumn]{article}
2: \usepackage{amsmath,amssymb,epsfig}
3: \setlength\textheight{20cm}
4: 
5: \markboth{B. Bid\'egaray-Fesquet \& J.-C. Saut}{
6: On the propagation of an optical wave in a photorefractive medium}
7: 
8: \title{\textsc{On the propagation of an optical wave in a photorefractive medium}}
9: 
10: \author{\small B. BID\'EGARAY-FESQUET\\
11: \it\small Laboratoire Jean Kuntzmann, \\
12: \it\small CNRS UMR 5224 et Grenoble Universit\'es, \\ 
13: \it\small B.P. 53, 38041 Grenoble Cedex 9, France\\
14: \it\small Brigitte.Bidegaray@imag.fr\\[5mm]
15: \small J.-C. SAUT\\
16: \it\small UMR de Math\'ematiques, Universit\'e Paris-Sud,\\ 
17: \it\small 91405 Orsay, France\\
18: \it\small Jean-Claude.Saut@math.u-psud.fr}
19: 
20: \date{}
21: 
22: \begin{document}
23: 
24: \maketitle
25: 
26: \begin{abstract}
27: The aim of this paper is first to review the derivation of a model describing
28: the propagation of an optical wave in a photorefractive medium and to present
29: various mathematical results on this model: Cauchy problem, solitary waves.
30: \end{abstract}
31: 
32: \noindent {\it Keywords:} 
33: Photorefractive media; Cauchy problem; solitary waves.
34: 
35: \noindent AMS Subject Classification: 35A05, 35A07, 35Q55, 35Q60, 78A60
36: 
37: \renewcommand{\thefootnote}{\fnsymbol{footnote}}
38: 
39: \footnote[0]{Electronic version of an article published as \\	
40: Mathematical Models and Methods in Applied Sciences (M3AS) \\
41: Volume 17, Issue 11, 2007, 1883--1904, DOI 10.1142/S0218202507002509 \\
42: \copyright World Scientific Publishing Company \\
43: {\texttt{http://www.worldscinet.com/cgi-bin/details.cgi?id=voliss:m3as{\textunderscore}1711\&type=toc}}}
44: 
45: %%%%%%%%%%%%%% AUTHORS' MACROS %%%%%%%%%%%%%% 
46: 
47: \def\ds{\displaystyle}
48: \def\ph{\varphi}
49: \def\eps{\varepsilon}
50: \def\bfE{\mathbf{E}}
51: \def\bfJ{\mathbf{J}}
52: \def\bfc{\mathbf{c}}
53: \def\bfe{\mathbf{e}}
54: \def\bfk{\mathbf{k}}
55: \def\bfr{\mathbf{r}}
56: \def\bfx{\mathbf{x}} 
57: \def\bbR{\mathbb{R}}
58: \def\calC{\mathcal{C}}
59: \def\calD{\mathcal{D}}
60: \def\calF{\mathcal{F}}
61: \def\calS{\mathcal{S}}
62: \renewcommand{\div}{\operatorname{div}}
63: \newcommand{\tens}[1]{\widehat{#1}}
64: 
65: \newtheorem{theorem}{Theorem}
66: \newtheorem{lemma}[theorem]{Lemma}
67: \newtheorem{corollary}[theorem]{Corollary}
68: \newtheorem{proposition}[theorem]{Proposition}
69: 
70: \def\theequation{\arabic{section}.\arabic{equation}}
71: \newcommand{\refcite}[1]{\ \cite{#1}}
72: \def\proof{\noindent {\bf Proof.\ }}
73: \def\endproof{$\Box$}
74: 
75: \def\remark{\noindent\textbf{Remark.}\ }
76: 
77: \newlength{\fntxvi} \newlength{\fntxvii}
78: \newcommand{\chemical}[1]
79: {{\fontencoding{OMS}\fontfamily{cmsy}\selectfont
80:   \fntxvi\the\fontdimen16\font
81:   \fntxvii\the\fontdimen17\font
82:   \fontdimen16\font=3pt \fontdimen17\font=3pt
83:   $\mathrm{#1}$
84:   \fontencoding{OMS}\fontfamily{cmsy}\selectfont
85:   \fontdimen16\font=\fntxvi \fontdimen17\font=\fntxvii }}
86: 
87: %%%%%%%%%%%%%% END AUTHORS' MACROS %%%%%%%%%%%%%% 
88: 
89: \maketitle
90: 
91: \section{Introduction}
92: 
93: A modification of the refraction index in \chemical{LiNbO_3} or
94: \chemical{LiTaO_3} crystals has been observed in the 1960s and first 
95: considered as a drawback. This photo-induced varia\-tion of the index is called 
96: the photorefractive effect and occurs in any electro-optical or photoconductive 
97: crystal. Applications have been found in the 1970s--1980s to real-time 
98: signal processing, phase conjugation, or amplification of beams or images. 
99: 
100: In this paper we are interested in deriving a not-too-simple but tractable 
101: math\-ematical model for the propagation of light in such materials. Solitonic 
102: propagation is one of our concern but we focus here on initial value 
103: problems. A very complete review of solitonic propagation in photorefractive 
104: media may be found in \refcite{delre-crosignani-diporto}. Our derivation 
105: follows the same guidelines as theirs but point out the different 
106: approximations made for future mathe\-matical studies. 
107: 
108: The outline of the paper is the following. In Sec. \ref{Sec_Model} we 
109: first derive the Kukhtarev model for the material and then couple it to a wave 
110: propagation model for light to obtain a complete set of equations. A 1D model
111: is obtained keeping only one of the two transverse space variables. This is a 
112: saturated nonlinear Schr\"odinger equation, the mathematical theory of which 
113: is addressed in Sec. \ref{Sec_saturated} in arbitrary dimension. Section 
114: \ref{Sec_Cauchy} is devoted to the study of the full 2D model with emphasis on
115: the Cauchy problem and the solitary wave solutions. 
116: 
117: \section{Derivation of the Model}
118: \label{Sec_Model}
119: 
120: \subsection{The photorefractive effect}
121: 
122: The propagation of an optical wave in insulating or semi-insulating 
123: electro-optical crystals induces a charge transfer. The new distribution of 
124: charges induces in turn an electric field which produces a variation of the 
125: refraction index. The main characteristics of this effect are the following: 
126: (1) Sensibility to energy (and not to the electric field), 
127: (2) Nonlocal effect (charge distributions and the electric field are not 
128: located at the same position),
129: (3) Inertia (charges need a certain time to move), 
130: (4) Memory and reversibility (in the dark the space charge, and therefore
131: the index variation, is persistent but an uniform light redistributes 
132: uniformly all charges --- this yields applications to holography). 
133: 
134: The sensibility to energy reminds us of Kerr media yielding the classical 
135: cubic nonlinear Schr\"odinger (NLS) equation. The nonlocal effects will of 
136: course complicate the mathematical analysis compared to NLS equations, but the 
137: general ideas will be the same. In our final model, inertial effects will be 
138: neglected since time is removed from the material equations. Memory and 
139: reversibility effects involve ion displacement in materials like 
140: \chemical{Bi_2TeO_5}, which we will not take into account in the present 
141: study. 
142: 
143: \subsection{The Kukhtarev model}
144: 
145: The physical modeling of the photorefractive effect assumes that charges are 
146: trapped in impurities or defaults of the crystal mesh. We chose here to derive 
147: the model only in the case when charges are electrons. Some materials like 
148: semi-conductors necessitate to model both electrons and holes. Therefore we 
149: restrict our study to insulating media. 
150: 
151: \subsubsection{Charge equation}
152: 
153: Electrons come from donor sites with density $N_{\rm D}$. This density is 
154: supposed to be much greater than that of the acceptor sites (impurities) which 
155: we denote by $N_{\rm A}$. The density of donor sites which are indeed ionized 
156: is $N_{\rm D}^+$ and we of course have 
157: $N_{\rm D}^+\leq N_{\rm A} \ll N_{\rm D}$. Local neutrality, i.e. no electrons 
158: in the conduction band, corresponds to the relation $N_{\rm D}^+=N_{\rm A}$. 
159: The total charge is given by
160: \begin{equation}
161: \label{Eq_Charge} %2.1
162: \rho = e(N_{\rm D}^+-N_{\rm A}-n_{\rm e}),
163: \end{equation}
164: where $e$ is the electron charge and $n_{\rm e}$ the electron density.
165: 
166: 
167: \subsubsection{Evolution of ionized donor sites}
168: 
169: Photoionization and recombination affect the density of ionized donor sites.
170: Photoionization is proportional to the density of not ionized donor sites 
171: ($N_{\rm D}-N_{\rm D}^+$). In the dark it is proportional to a thermal 
172: excitation rate $\beta$ but is also sensitive to light intensity $I_{\rm em}$ 
173: with a photoexcitation coefficient $s$. Recombination is proportional to the 
174: density of electrons and occurs over a time scale 
175: $\tau=1/(\gamma_{\rm r}N_{\rm D}^+)$ which does not depend on $n_{\rm e}$ if 
176: the excitation rate is low, therefore the total evolution of ionized donor 
177: sites is
178: \begin{equation}
179: \label{Eq_Donor} %2.2
180: \partial_t N_{\rm D}^+ 
181: = (\beta+sI_{\rm em})(N_{\rm D}-N_{\rm D}^+) 
182: - \gamma_{\rm r} n_{\rm e} N_{\rm D}^+.
183: \end{equation}
184: 
185: \subsubsection{Charge transport}
186: 
187: Now the main point is to describe the three phenomena which contribute to the 
188: charge transport or current density. The first phenomenon is isotropic and is 
189: due to thermal diffusion. It is proportional to the gradient of the electron 
190: density. The electron mobility is denoted by $\mu$, $T$ is the temperature and 
191: $k_{\rm B}$ the Boltzmann constant. The second phenomenon is drift and is 
192: collinear to the electric field $\bfE_{\rm tot}$. Finallly, the photovoltaic 
193: effect is collinear to the optical axis $\bfc$ and proportional to the 
194: non-ionized donor density and the field intensity with a photovoltaic 
195: coefficient $\beta_{\rm ph}$. The total current density is therefore
196: \begin{equation}
197: \label{Eq_Current} %2.3
198: \bfJ 
199: = e\mu n_{\rm e}\bfE_{\rm tot} 
200: + \mu k_{\rm B} T\nabla n_{\rm e} 
201: + \beta_{\rm ph}(N_{\rm D}-N_{\rm D}^+)\bfc I_{\rm em}.
202: \end{equation}
203: 
204: \subsubsection{Closure of the model}
205: 
206: The closure of the model is first based on charge conservation and the Poisson
207: equation:
208: \begin{equation}
209: \label{Eq_Conservation} %2.4
210: \partial_t \rho + \nabla\cdot\bfJ = 0, 
211: \end{equation}
212: \begin{equation}
213: \label{Eq_Poisson} %2.5
214: \nabla\cdot(\eps_0\tens\eps\bfE_{\rm sc}) = \rho.
215: \end{equation}
216: The crystal is anisotropic and this is accounted for in the relative 
217: permittivity~$\tens\eps$ which is a tensor. A careful analysis of the 
218: different electric fields has to be done. In  Poisson equation 
219: \eqref{Eq_Poisson}, $\bfE_{\rm sc}$ is the space charge field which is induced 
220: by the charge density. The total field $\bfE_{\rm tot}$ only occurs in the 
221: equations through its gradient (Eqs. \eqref{Eq_Conservation} and 
222: \eqref{Eq_Current}). Two fields are constant and disappear in the final 
223: equations: the photovoltaic field 
224: $\bfE_{\rm ph}=\beta_{\rm ph}\gamma_{\rm r}N_{\rm A}\bfc/e\mu s=E_{\rm ph}\bfc$,
225: and an external field $\bfE_{\rm ext}$ which is often applied in one transverse 
226: direction on the faces of the crystal. A last contribution to the total field 
227: is $\bfE$, connected to the light which propagates in the crystal and its 
228: description is given in Sec. \ref{Subsec_NLS}.
229: 
230: The set of five equations \eqref{Eq_Charge}--\eqref{Eq_Poisson} is called the 
231: Kukhtarev model and was first given in \refcite{kukhtarev-markow-odoluv-soskin-vinetskii}.
232: 
233: \subsection{Propagation of the light wave in the crystal}
234: \label{Subsec_NLS}
235: 
236: We have already introduced the relative permittivity tensor $\tens\eps$ which
237: plays a r\^ole in the description of the propagation of a light wave in the 
238: crystal via the wave equation:
239: \begin{equation*}
240: \partial_t^2 (\tens\eps \bfE) - c^2 \nabla^2\bfE = 0.
241: \end{equation*}
242: In a non-centrosymmetric crystal the preponderant nonlinear effect is the Pockels 
243: effect which yields the following $\bfE$-dependence for the permittivity 
244: tensor:
245: \begin{equation*}
246: \tens{\eps(\bfE)} 
247: = \tens{\eps(0)} - \tens\eps \tens{(\tens\bfr\cdot\bfE)} \tens\eps 
248: = n^2 - \tens\eps \tens{(\tens\bfr\cdot\bfE)} \tens\eps, 
249: \end{equation*} 
250: where $\tens\bfr$ is the linear electro-optic tensor and $n$ the mean 
251: refraction index. We now suppose that $\bfE$ is a space perturbation of a 
252: plane wave (paraxial approximation) of frequency $\omega$, wave vector $\bfk$ 
253: and polarization $\bfe$:
254: \begin{equation*}
255: \bfE(t,\bfx) = A(\bfx) \exp(i(\omega t-\bfk\cdot\bfx)) \bfe.
256: \end{equation*}
257: Such a wave with polarization $\bfe$ only "sees" a part of tensor 
258: $\tens{\eps(\bfE)}$, or equivalently a variation $\delta n$ of the refraction 
259: index $n$:
260: \begin{equation*}
261: \delta n 
262: = -\frac1{2n} [\bfe \tens\eps\ \tens{\tens{\bfr}\bfe}\ \tens\eps \bfe^*] \bfE.
263: \end{equation*}
264: Now we can write an equation for the amplitude $A$ which takes into account 
265: the dispersion relation $c^2|\bfk|^2 = n^2\omega^2$ and the slowly varying 
266: envelope approximation in the $\bfk$ direction. We denote by $\nabla_\perp$ 
267: the gradient in the perpendicular directions to $\bfk$ and
268: \begin{equation}
269: \label{Eq_Envelope} %2.6
270: \left[\nabla_\perp^2 - 2i\bfk\cdot\nabla + 2 |\bfk|^2 \frac{\delta n}{n}\right]
271: A(\bfx) \bfe = 0.
272: \end{equation}
273: Of course, we can consider the superposition of such waves to describe for 
274: example pump and probe experiments.
275: 
276: The system is now closed but it is impossible to solve Eqs. 
277: \eqref{Eq_Charge}--\eqref{Eq_Envelope}. We have to simplify them taking into 
278: account characteristic scales. Our description follows (or more precisely 
279: makes explicit the assumptions in \refcite{zozulya-anderson}) and is purely 
280: formal. The rigorous justification is certainly difficult and should include 
281: the approximations made in Sec. \ref{Subsec_NLS}.
282: 
283: \subsection{Characteristic values}
284: 
285: We first want to define a characteristic electron intensity $n_0$ by 
286: considering uniform solutions in space and time. Equations \eqref{Eq_Charge}, 
287: \eqref{Eq_Donor} and \eqref{Eq_Poisson} yield
288: \begin{equation*}
289: \rho = e(N_{\rm D}^+-N_{\rm A}-n_{\rm e}) = 0 \textrm{ and } 
290: 0 = (\beta+sI)(N_{\rm D}-N_{\rm D}^+) - \gamma_{\rm r} n_{\rm e} N_{\rm D}^+.
291: \end{equation*}
292: With a characteristic intensity $I_0$, neglecting $\beta$ and assuming 
293: $n_{\rm e}\ll N_{\rm A}$, we have 
294: $n_0 = sI_0 (N_{\rm D}-N_{\rm A})/\gamma_{\rm r} N_{\rm A}$.
295: 
296: There are three characteristic times: 
297: (1) the characteristic lifetime of an electron (in the dark) 
298: $\tau_{\rm e}=1/\gamma_{\rm r}N_{\rm A}$, 
299: (2) the characteristic evolution time of ionized donors 
300: $\tau_{\rm d}=1/\gamma_{\rm r}n_0$ (and a consequence of $n_0\ll N_{\rm A}$ is 
301: $\tau_{\rm e}\ll\tau_{\rm d}$),
302: (3) the characteristic relaxation time of the electric field 
303: $t_0=\eps_0\eps_{\rm c}/e\mu n_0$, where $\eps_{\rm c}$ is the characteristic 
304: value of $\tens\eps$ along the $\bfc$ direction. It is obtained combining 
305: Eqs. \eqref{Eq_Current} and \eqref{Eq_Poisson} assuming there is only 
306: drift. If a timescale has to be kept, it is $t_0$, but we do not detail 
307: this point since we neglect time-dependence in the final equations. 
308: 
309: The Debye length $L_{\rm D}$ is the characteristic value of the field space 
310: variation. It is determined together with the characteristic field $E_0$. The 
311: Poisson equation \eqref{Eq_Poisson} yields $L_{\rm D}=\eps_0\eps_{\rm c}E_0/eN_{\rm A}$. If 
312: drift and isotropic diffusion have the same order, $E_0=k_{\rm B}T/eL_{\rm D}$ and 
313: therefore
314: \begin{equation*}
315: L_{\rm D} = \left(\frac{k_{\rm B}T\eps_0\eps_{\rm c}}{e^2N_{\rm A}}\right)^{1/2},\ \
316: E_0 = \left(\frac{k_{\rm B}TN_{\rm A}}{\eps_0\eps_{\rm c}}\right)^{1/2},
317: \textrm{ \ and \ } I_0 = \frac{k_{\rm B}TN_{\rm A}}{\eps_0\eps_{\rm c}}.
318: \end{equation*} 
319: 
320: \subsection{The Zozulya--Anderson model}
321: \label{Subsec_ZA}
322: 
323: Zozulya--Anderson model\cite{zozulya-anderson} is obtained using the above 
324: characteristic values and for a specific material (\chemical{LiNbO_3}) which 
325: imposes certain symmetries. The adiabatic assumption allows to get rid of the 
326: time-dependence and an asymptotic formal analysis which accounts for the 
327: very large donors density $N_{\rm D}$ ends the derivation. 
328: 
329: Dimensionless equations are obtained using $n_0$ and $N_{\rm A}$ for electron 
330: and ion densities respectively, $I_0$ for intensities, $E_0$ for fields and 
331: $\eps_{\rm c}$ for the permittivity tensor. Coefficient $\beta$ is normalized 
332: as a dark intensity $I_{\rm d}=\beta/sI_0$. We keep all the other notations but 
333: they now denote the normalized variables. The total intensity is 
334: $I=I_{\rm em}+I_{\rm d}$. We assume that the space charge field $\bfE_{\rm sc}$ 
335: derives from a potential: $L_{\rm D}\nabla\ph=-\bfE_{\rm sc}$. In the adiabatic 
336: assumption matter equations reduce to
337: \begin{eqnarray*}
338: & \ds I \frac{1-N_{\rm D}^+ N_{\rm A}/N_{\rm D}}{1-N_{\rm A}/N_{\rm D}} 
339:    = n_{\rm e} N_{\rm D}^+, \\[2mm]
340: & \ds L_{\rm D} \nabla \cdot \left\{n_{\rm e} \bfE_{\rm tot} 
341:    + L_{\rm D} \nabla n_{\rm e} 
342:    + \bfE_{\rm ph} I_{\rm em} \frac{1-N_{\rm D}^+ N_{\rm A}/N_{\rm D}}
343:                                    {1-N_{\rm A}/N_{\rm D}}\right\} = 0, \\[2mm]
344: & \ds -L_{\rm D}^2 \nabla \cdot (\tens\eps \nabla\ph) 
345:    = N_{\rm D}^+ -1 - \frac{n_0}{N_{\rm A}}n_{\rm e}.
346: \end{eqnarray*}
347: In \chemical{LiNbO_3}, $N_{\rm A}/N_{\rm D}\sim10^{-3}$ and 
348: $n_0/N_{\rm D}\sim10^{-6}$ and we neglect them. Finally, we make different 
349: assumptions on the fields: first the beam is not too thin, the photogalvanic 
350: and the external applied fields are not too large and therefore we may neglect 
351: $-L_{\rm D}^2 \nabla \cdot (\tens\eps \nabla\ph)$; second the propagation field 
352: amplitude is relatively small and we assimilate $\bfE_{\rm tot}$ and 
353: $\bfE_{\rm sc}$. This implies $n_{\rm e}=I$ and $N_{\rm D}^+=1$ and we 
354: have only one matter equation, namely
355: \begin{equation*}
356: \nabla I \cdot \nabla\ph + I\nabla^2\ph - \nabla^2I 
357: - k_{\rm D} E_{\rm ph} \bfc \cdot \nabla I = 0,
358: \end{equation*}
359: where $k_{\rm D}=1/L_{\rm D}$. To obtain a "simpler" equation, in physics 
360: papers the variable $U=\ph-\ln I$ is often used. This variable seems however 
361: to lack physical meaning. 
362: 
363: \noindent
364: The final matter equation is
365: \begin{equation}
366: \label{Eq_U} %2.7
367: \nabla U \cdot \nabla\ph + \nabla^2U - k_{\rm D} E_{\rm ph} \bfc \cdot \nabla I = 0.
368: \end{equation}
369: 
370: We now fix different space directions. Propagation is supposed to take place 
371: in the $z$-direction and $\bfk=k\bfe_z$. The two transverse directions are 
372: therefore $x$ and $y$. The $\bfe_x$ direction is chosen as both $\bfc$ and 
373: $\bfe$. If an external field is applied, it will be along $\bfe_x$ as well.
374: In the matter equation \eqref{Eq_U}, the quantity $\bfc \cdot \nabla I$ simply
375: reads $\partial_x I$. In \chemical{LiNbO_3}, $r=r_{xxx}$ is responsible
376: for the change of refractive index (it is $r_{xxy}$ in some other
377: materials) and we approximate $\tens\eps$ by $n^2$ in the expression
378: for $\delta n$ which becomes 
379: $\delta n = \frac12 n^3 r E_0 L_{\rm D}\partial_x \ph$. 
380: Together with Eq. \eqref{Eq_Envelope} the envelope equation now reads 
381: \begin{equation*}
382: \left[\partial_z + \frac{i}{2k} \nabla_\perp^2\right]A(\bfx) 
383: = - i\frac k2 n^2 r E_0 L_{\rm D}\partial_x \ph A(\bfx).
384: \end{equation*}
385: 
386: The last step is to have dimensionless space variables. We set 
387: $\alpha = \frac k2 n^2 r E_0$ which has the dimension of the inverse of a 
388: space variable. We denote $z'=|\alpha|z$, $(x',y')=\sqrt{k|\alpha|}(x,y)$, 
389: $A'= A/\sqrt{I_0I_{\rm d}}$, $\ph'=\sqrt{k|\alpha|}\ph/k_{\rm D}$ and 
390: $U'=\sqrt{k|\alpha|}U/k_{\rm D}$. 
391: The last approximations are now $U'=\ph'$, $k\gg|\alpha|$ and 
392: $E_{\rm ph}\sim E_0$, and omitting primes:
393: \begin{eqnarray*}
394: \left[\partial_z-\frac i2 \nabla_\perp^2\right]A & = & -iA\partial_x\ph, \\[2mm]
395: \nabla_\perp^2 \ph + \nabla_\perp \ln(1+|A|^2)\cdot \nabla_\perp \ph 
396: & = & \partial_x\ln(1+|A|^2).
397: \end{eqnarray*}
398: These equations are usually referred to as a model derived in
399: \refcite{zozulya-anderson} but only seeds of these equations are derived 
400: there usually including many other terms and especially time derivatives.
401: 
402: In the wide literature devoted to photorefractive media, many equations are 
403: written which resemble those above but with different choices of asymptotic 
404: approx\-imations. In particular numerical results are very often obtained keeping 
405: the time in the matter equations (see \refcite{stepken-kaiser-belic-krolikowski} or
406: \refcite{wolfersberger-fressengeas-maufroy-kugel}).
407: 
408: \subsection{Mathematical setting}
409: 
410: If we look at a wider class of materials we may have different signs for the 
411: nonlinearity (in reference to the cubic nonlinear Schr\"odinger equation, 
412: the case $a=1$ is classically called the focusing case, and $a=-1$ the 
413: defocusing case). Besides mathematicians are more accustomed to use $t$ as the 
414: evolution variable. We will therefore consider the system
415: \begin{equation}
416: \label{Eq_Aln} %2.8
417: \left\{
418: \begin{array}{ll}
419: i\partial_t A + \Delta A = -aA\partial_x\ph,              & a=\pm1, \\
420: \Delta \ph + \nabla \ln(1+|A|^2) \cdot \nabla \ph  = \partial_x\ln(1+|A|^2),
421: \end{array}
422: \right.
423: \end{equation}
424: where $\Delta = \partial_x^2+\partial_y^2$ or $\Delta=\partial_x^2$. 
425: 
426: These expressions with logarithms are widely used in the physics literature, 
427: maybe because they are the starting point of solitonic studies and logarithms 
428: appear
429: naturally in the expression of solitary waves (see Sec. \ref{Sec_1D}). This 
430: form is however cumbersome to handle for the mathematical analysis, and it is 
431: much more convenient to cast \eqref{Eq_Aln} as
432: \begin{equation}
433: \label{Eq_Adiv} %2.9
434: \left\{
435: \begin{array}{l}
436: i\partial_t A + \Delta A = -aA\partial_x\ph, \\[2mm]
437: \div \left((1+|A|^2) \nabla \ph\right) = \partial_x(|A|^2),
438: \end{array}
439: \right.
440: \end{equation}
441: which is closer to the original Kukhtarev equations.
442: 
443: We have seen that the main effects take place in the $t$- (propagation) and the 
444: $x$-directions (drift, anisotropic diffusion, external field, polarization). 
445: It is therefore natural to study the equations with no dependence in the $y$ 
446: variable. In the one-dimensional case, we infer immediately from the last 
447: equation in System \eqref{Eq_Adiv} that 
448: $(1+|A|^2) \partial_x \ph = |A|^2 - C(t)$ where the constant $C(t)$ is given by 
449: the boundary conditions. If no external field is applied $C(t)\equiv0$. This is 
450: the case for bright solitary waves (see \refcite{mamaev-saffman-zozulya2}). 
451: In the case of dark solitary waves $C(t) = \lim_{x\to\pm\infty} |A|^2$ 
452: (see \refcite{mamaev-saffman-zozulya1}), which does not depend on $t$ 
453: either. In both cases, System \eqref{Eq_Adiv} reduces to the saturated NLS 
454: equation
455: \begin{equation}
456: \label{Eq_SaturatedNLS} %2.10
457: i \partial_t A + \partial_x^2 A = -a \frac{|A|^2 - |A_\infty|^2}{1+|A|^2}A.
458: \end{equation}
459: In the sequel we will mainly consider the case when $A_\infty=0$ and show 
460: that, in some sense, the dynamics of \eqref{Eq_Adiv} is similar to that of 
461: \eqref{Eq_SaturatedNLS} which we will recall in Sec. \ref{Sec_saturated}.
462: 
463: In the two-dimensional case \eqref{Eq_Adiv} can be viewed as a saturated version 
464: of a Davey--Stewartson system. Namely, replacing $1+|A|^2$ by 1 in the L.H.S. 
465: of \eqref{Eq_Adiv} yields
466: \begin{equation*}
467: \left\{
468: \begin{array}{rcl}
469: i\partial_t A + \Delta A & = & -a A\partial_x\ph, \\[2mm]
470: \Delta \ph & = & \partial_x(|A|^2),
471: \end{array}
472: \right.
473: \end{equation*}
474: which is the Davey--Stewartson system of the elliptic--elliptic type (see 
475: \refcite{ghidaglia-saut1}). 
476: 
477: \section{The Saturated NLS Equation}
478: \label{Sec_saturated}
479: 
480: We review here some mathematical facts, more or less known, on the saturated 
481: NLS equation
482: \begin{equation}
483: \label{Eq_NLSCauchy} %3.1
484: \left\{
485: \begin{array}{ll}
486: \ds i\partial_t A + \Delta A = - a\frac{|A|^2A}{1+|A|^2},\ & a=\pm1, \\[2mm]
487: A(\bfx,0) = A_0(\bfx),
488: \end{array}
489: \right.
490: \end{equation}
491: where $A=A(\bfx,t)$ and $\bfx\in\bbR^d$. We have derived this equation for 
492: $d=1$, but give here results for a general $d$. This equation is also derived
493: in other contexts, for example the propagation of a laser beam in gas vapors
494: \cite{tikhonenko-christou-luther-davies}.
495: 
496: \subsection{The Cauchy problem}
497: 
498: The Cauchy problem \eqref{Eq_NLSCauchy} can be solved in $L^2$ and in the energy 
499: space $H^1$.
500: 
501: \begin{theorem}
502: \label{Th_NLSCauchy}
503: {\em (i)} Let $A_0\in L^2(\bbR^d)$. Then there exists a unique solution 
504: $A\in\calC(\bbR;L^2(\bbR^d))$ of \eqref{Eq_NLSCauchy} which satisfies 
505: furthermore
506: \begin{equation}
507: \label{Eq_NLSL2Claw}
508: \int_{\bbR^d} |A(\bfx,t)|^2 d\bfx = \int_{\bbR^d} |A_0(\bfx)|^2 d\bfx, 
509: \hspace{1cm} t\in\bbR. 
510: \end{equation}
511: {\em (ii)} Let $A_0\in H^1(\bbR^d)$. Then the solution above satisfies 
512: $A\in\calC(\bbR;H^1(\bbR^d))$ and
513: \begin{eqnarray}
514: && \int_{\bbR^d} 
515: \left[|\nabla A(\bfx,t)|^2 d\bfx + a \ln (1+|A(\bfx,t)|^2) \right]
516: d\bfx \nonumber \\
517: \label{Eq_NLSH1Claw}
518: && \hspace{1cm} = \int_{\bbR^d} 
519: \left[|\nabla A_0(\bfx)|^2 d\bfx + a \ln (1+|A_0(\bfx)|^2) \right]
520: d\bfx, \hspace{1cm} t\in\bbR. 
521: \end{eqnarray}
522: \end{theorem}
523: 
524: \proof
525: The norm conservations \eqref{Eq_NLSL2Claw} and \eqref{Eq_NLSH1Claw} result 
526: from multiplying \eqref{Eq_NLSCauchy} by $\bar A$ and $\partial_t \bar A$ 
527: respectively and integrating the complex and real parts respectively. This 
528: formal proof is justified by the standard truncation process. 
529: 
530: Let $S(t)$ be the group operator associated to the linear Schr\"odinger equation 
531: $i\partial_t A + \Delta A =0$. Then the Duhamel formula for \eqref{Eq_NLSCauchy} 
532: reads
533: \begin{equation}
534: \label{Eq_3.4}
535: A(\bfx,t) = S(t) A_0(\bfx) 
536: - a \int_0^t S(t-s) \frac{|A(\bfx,s)|^2}{1+|A(\bfx,s)|^2}A(\bfx,s)\ ds.
537: \end{equation}
538: Since $x\mapsto x/(1+x)$ is Lipschitz, we easily infer that the R.H.S. of 
539: \eqref{Eq_3.4} defines a contraction on a suitable ball of 
540: $\calC([0,T];L^2(\bbR^d))$ for some $T>0$. The local well-posedness in 
541: $L^2(\bbR^d))$ follows. Global well-posedness is derived from the conservation
542: law \eqref{Eq_NLSL2Claw}.
543: 
544: The $H^1$ theory follows the same argument, noticing that
545: \begin{equation*}
546: \left|\nabla\left(\frac{|A|^2}{1+|A|^2}A\right)\right| 
547: = \left|\frac{A^2}{(1+|A|^2)^2}\nabla \bar A 
548:         + \frac{|A|^2}{(1+|A|^2)^2}\nabla A\right|
549: \leq \frac12|\nabla A|.
550: \end{equation*}
551: \endproof
552: 
553: \remark As a consequence of \eqref{Eq_NLSL2Claw}, \eqref{Eq_NLSH1Claw} and 
554: $\ln(1+|A|^2)\leq|A|^2$, we obtain the uniform bound 
555: \begin{equation}
556: \label{Eq_NLSH1Est}
557: \int_{\bbR^d} |\nabla A(\bfx,t)|^2 d\bfx 
558: \leq \int_{\bbR^d} |A_0(\bfx)|^2 d\bfx 
559: + \int_{\bbR^d} |\nabla A_0(\bfx)|^2 d\bfx, 
560: \hspace{1cm} t\in\bbR. 
561: \end{equation}
562: Contrarily to the context of the usual nonlinear cubic Schr\"odinger equation,
563: this bound does not depend on the sign of $a$ and in particular saturation 
564: means that no blow-up is occurs. 
565: 
566: \subsection{Solitary waves --- one-dimensional results}
567: \label{Sec_1D}
568: 
569: In the one-dimensional case, it is possible to compute first integral 
570: formulations of the solitary waves. \\
571: 
572: Bright solitary waves are sought for in the form $A(x,t) = e^{i\omega t}u(x)$ 
573: (see \refcite{mamaev-saffman-zozulya2}), where $A$ is a solution to 
574: \eqref{Eq_SaturatedNLS} with $A_\infty=0$. The function $u$ is supposed to have
575: a maximum at $x=0$ ($u(0)=u_m>0$ and $u'(0)=0$), therefore 
576: \begin{equation*}
577: [u'(x)]^2 = (\omega-a)[u^2(x)-u_m^2] + a [\ln(1+u^2)-\ln(1+u_m^2)].
578: \end{equation*}
579: We furthermore want that for $x\to\infty$, $u(x)\to0$ and $u'(x)\to0$. This
580: yields a unique possible frequency for the solitary wave, namely
581: \begin{equation*}
582: \omega = a\left(1-\frac{\ln(1+u_m^2)}{u_m^2}\right)
583: \end{equation*}
584: and
585: \begin{equation*}
586: [u'(x)]^2 = a \left(-\frac{u^2(x)}{u_m^2}\ln(1+u_m^2)+\ln(1+u^2)\right).
587: \end{equation*}
588: Since $u_m$ is supposed to be the maximum of $u$, this quantity is positive 
589: only if $a=1$ (focusing case) and the bright soliton is solution to the 
590: first order equation:
591: \begin{equation*}
592: u'(x) = -\textrm{sign}(x) \sqrt{\ln(1+u^2)-\frac{u^2}{u_m^2}\ln(1+u_m^2)} 
593: \textrm{ with } \omega = 1-\frac{\ln(1+u_m^2)}{u_m^2}.
594: \end{equation*}
595: 
596: Dark solitary waves are sought for in the form $A(x,t) = u(x)$ (see 
597: \refcite{mamaev-saffman-zozulya1}) where $A$ is solution to
598: \eqref{Eq_SaturatedNLS}. There is no time-dependence. We assume that 
599: $\lim_{x\to\pm\infty}u'(x)=0$ and consistently with $A_\infty\neq0$,  
600: \begin{equation*}
601: \lim_{x\to+\infty}u(x)=-\lim_{x\to-\infty}u(x)=u_\infty.
602: \end{equation*}
603: Then 
604: \begin{equation*}
605: [u'(x)]^2 
606: = a \left( - (u^2-u_\infty^2) 
607: + (1+u_\infty^2)\ln\left(\frac{1+u^2}{1+u_\infty^2}\right) \right).
608: \end{equation*}
609: At the origin $u(0)=0$ and we want more generally that $|u(x)|\leq|u_\infty|$.
610: Therefore, dark solitary waves only exist if $a=-1$ (defocusing case). In this 
611: context $u(x)$ is a monotonous function and is solution to the first order 
612: equation:
613: \begin{equation*}
614: u'(x) = \textrm{sign} (u_\infty) \sqrt{u^2-u_\infty^2 
615: - (1+u_\infty^2)\ln\left(\frac{1+u^2}{1+u_\infty^2}\right)}.
616: \end{equation*}
617: For both bright and dark solitary waves, no explicit solution is known.
618: 
619: \subsection{Solitary waves --- \textit{a priori} estimates and non existence}
620: 
621: Consider now the solitary wave solutions of \eqref{Eq_NLSCauchy} in any dimension 
622: $d$, that is solutions of the type $A(\bfx,t)=e^{i\omega t}U(\bfx)$, where 
623: $U\in H^1(\bbR^d)$ (we thus are only concerned with "bright" solitary waves). 
624: A solitary wave is a solution of the elliptic equation
625: \begin{equation}
626: \label{Eq_NLSSW}
627: -\Delta U + \omega U = a \frac{|U|^2U}{1+|U|^2}, \hspace{1cm} U\in H^1(\bbR^d).
628: \end{equation}
629: A trivial solution is $U\equiv0$. We seek other nontrivial solutions.
630: 
631: \begin{lemma}
632: Any $H^1(\bbR^d)$ solitary wave satisfies
633: \begin{equation}
634: \label{Eq_NLSSWEnergy}
635: \int_{\bbR^d} \left[ |\nabla U|^2 
636: + \left( \omega - a \frac{|U|^2}{1+|U|^2} \right) |U|^2 \right] d\bfx =0
637: \end{equation}
638: (energy identity)
639: \begin{equation}
640: \label{Eq_NLSSWPohozaev}
641: (d-2) \int_{\bbR^d} |\nabla U|^2 d\bfx + d\omega \int_{\bbR^d} |U|^2 d\bfx 
642: - ad  \int_{\bbR^d} \left[ |U|^2 - \ln (1+|U|^2) \right] d\bfx =0
643: \end{equation}
644: (Pohozaev identity).
645: \end{lemma}
646: 
647: \proof
648: As for Theorem \ref{Th_NLSCauchy}, \eqref{Eq_NLSSWEnergy} results from 
649: multiplying \eqref{Eq_NLSSW} by $\bar U$ and inte\-grating. To get 
650: \eqref{Eq_NLSSWPohozaev}, one multiplies \eqref{Eq_NLSSW} by 
651: $x_k\partial_{x_k}\bar U_k$, integrates the real part, and sums from 1 to $d$. 
652: This is justified by a standard truncation argument.
653: \endproof
654: 
655: \begin{corollary}
656: \label{Cor_NLSnoSW}
657: No nontrivial solitary wave (solution of \eqref{Eq_NLSSW}) exists when
658: \setlength\leftmargini{2pc}
659: \begin{itemize}
660: \item[{\em (i)}] $a=-1$ {\em(}defocusing case{\em)}, for $\omega\geq0$.
661: \item[{\em (ii)}] $a=1$ {\em(}focusing case{\em)} and $\omega\geq1$.
662: \item[{\em (iii)}] $a=\pm1$ if $\omega<0$ provided 
663: $|U|^2/(1+|U|^2) = O(1/|\bfx|^{1+\eps})$, $\eps>0$ as $|\bfx|\to+\infty$.
664: \end{itemize}
665: \end{corollary}
666: 
667: \proof
668: Identity \eqref{Eq_NLSSWEnergy} implies that no solitary wave may exist when 
669: $a=-1$ and $\omega\geq0$ or $a=1$ and $\omega\geq1$. When $d=1,2$, Eq. 
670: \eqref{Eq_NLSSWPohozaev} implies that no solitary wave exist when $\omega\leq0$ 
671: and $a=1$. Recall $d=2$ is the physical case. The remaining cases ($\omega<0$, 
672: $a=-1$ or $a=1$, $d\geq3$) follow from the classical result of Kato\cite{kato} 
673: on the absence of embedded eigenvalues. Indeed, we can write \eqref{Eq_NLSSW} as
674: \begin{equation*}
675: \Delta U + (-\omega + V(\bfx)) U = 0, \hspace{1cm} V(\bfx)=a\frac{|U|^2}{1+|U|^2},
676: \end{equation*}
677: assuming furthermore that $V(\bfx)=O(1/|\bfx|^{1+\eps})$, $\eps>0$, as 
678: $|\bfx|\to\infty$. A proof for $d=3,4$ or $d\geq5$, $\omega\leq-(d-2)/2$ with no
679: decaying assumption is given in Appendix.
680: \endproof
681: 
682: \begin{corollary}
683: \label{Cor_NLSSWExist}
684: Solitary waves may exist only when $a=1$ and $0<\omega<1$. 
685: \end{corollary}
686: 
687: Corollary \ref{Cor_NLSSWExist} is consistent with the one-dimensional "explicit" 
688: result. We first have a classical regularity and decay result. 
689: 
690: \begin{proposition}
691: \label{Prop_NLSSWDecay}
692: Let $a=1$ and $0<\omega<1$. Then any $U\in H^1(\bbR^d)$ solution of 
693: \eqref{Eq_NLSSW} satisfies
694: \begin{equation}
695: \label{Eq_NLSSWHinfty}
696: U \in H^\infty(\bbR^d),
697: \end{equation}
698: \begin{equation}
699: \label{Eq_NLSSWDecay}
700: e^{\delta|\bfx|} U \in L^\infty(\bbR^d) \textrm{ for any } \delta<\omega/2.
701: \end{equation}
702: \end{proposition}
703: 
704: \proof
705: $U\in H^\infty(\bbR^d)$ results trivially from a bootstrapping argument using 
706: $|U|^2/(1+|U|^2)<1$. To prove \eqref{Eq_NLSSWDecay}, we first derive the estimate
707: \begin{equation}
708: \label{Eq_NLSSWH1FiniteWeight}
709: \int_{\bbR^d} e^{\omega|\bfx|} \left[ |\nabla U|^2 + |U|^2 \right] d\bfx 
710: < +\infty.
711: \end{equation}
712: In fact, as in Cazenave\cite{cazenave1,cazenave2}, we multiply \eqref{Eq_NLSSW} 
713: by $e^{\omega|\bfx|}\bar U$ and integrate the real part (this formal argument is 
714: made rigorous by replacing $e^{\omega|\bfx|}$ by 
715: $e^{\omega|\bfx|/(1+\eps|\bfx|)}$, $\eps>0$, $\eps\to0$) to get
716: \begin{equation}
717: \label{Eq_NLSSWH1Weight}
718: \int_{\bbR^d} e^{\omega|\bfx|} \left[ |\nabla U|^2 + \omega |U|^2 \right] d\bfx
719: \leq \int_{\bbR^d} e^{\omega|\bfx|} \frac{|U|^4}{1+|U|^2} d\bfx 
720: + \omega \int_{\bbR^d} e^{\omega|\bfx|} |U||\nabla U| d\bfx.  
721: \end{equation}
722: By \eqref{Eq_NLSSWHinfty} there exists $R>0$ such that $|U|^2/(1+|U|^2)<\omega/4$ on 
723: $\bbR^d\setminus B_R$. Thus we infer from \eqref{Eq_NLSSWH1Weight} that
724: \begin{eqnarray*}
725: \int_{\bbR^d} e^{\omega|\bfx|} \left[ |\nabla U|^2 + \omega |U|^2 \right] d\bfx
726: & \leq & \int_{B_R} e^{\omega|\bfx|} \frac{|U|^4}{1+|U|^2} d\bfx 
727: + \frac\omega4 \int_{\bbR^d} e^{\omega|\bfx|} |U|^2 d\bfx \\ 
728: && + \frac\omega2 \left(\int_{\bbR^d}  e^{\omega|\bfx|}|U|^2 d\bfx
729: + \int_{\bbR^d}  e^{\omega|\bfx|}|\nabla U|^2 d\bfx \right) 
730: \end{eqnarray*}
731: which implies \eqref{Eq_NLSSWH1FiniteWeight}.
732: 
733: Now we write $U$ as a convolution
734: \begin{equation}
735: \label{Eq_NLSSWCovolution}
736: U(\bfx) = H_\omega \star \frac{|U|^2U}{1+|U|^2},
737: \hspace{1cm}\textrm{ where }
738: H_\omega = \calF^{-1} \left(\frac1{\omega+|\xi|^2}\right).
739: \end{equation}
740: As it is well known (\refcite{abramowitz-steglun}), 
741: $H_\omega(\bfx)=\omega^{(d-2)/2} G_1(\omega^{1/2}\bfx)$ where
742: \begin{equation*}
743: G_1(z) = 
744: \begin{cases}
745: C|z|^{(2-d)/2} K_{(d-2)/2} (|z|), & d\geq3, \\[2mm]
746: K_0 (|z|), & d=2,
747: \end{cases}
748: \end{equation*}
749: where $K_\nu$ is the modified Bessel function of order $\nu$. Furthermore (see 
750: \refcite{abramowitz-steglun}), one has the asymptotic behavior:
751: \begin{equation}
752: \label{Eq_Bessel}
753: \left\{
754: \begin{array}{rcll}
755: K_\nu (z) & \sim & \frac12 \Gamma(\nu) \left(\frac12|z|\right)^{-\nu}, 
756: & \textrm{ for } \nu>0, \textrm{ as } |z|\to0, \\[2mm]
757: K_0 (z) & \sim & - \ln(|z|), & \textrm{ as } |z|\to0, \\[2mm]
758: K_\nu (z) & \sim & C|z|^{-1/2} e^{-|z|}, & \textrm{ as } |z|\to\infty.
759: \end{array}
760: \right.
761: \end{equation}
762: We infer from \eqref{Eq_NLSSWCovolution} that
763: \begin{equation}
764: \label{Eq_NLSSWCovolutionEst}
765: e^{\delta|\bfx|}|U(\bfx)| 
766: \leq \int_{\bbR^d} e^{\delta|\bfx-\bfx'|}H_\omega(\bfx-\bfx')e^{\delta|\bfx'|} 
767: \frac{|U|^3}{1+|U|^2}(\bfx')d\bfx'.
768: \end{equation}
769: Since by \eqref{Eq_Bessel} $e^{\delta|\bfx|}H_\omega(\bfx)\in L^2(\bbR^d)$ for
770: $0<\delta<\omega^{1/2}$, and by \eqref{Eq_NLSSWH1FiniteWeight} 
771: $e^{\delta|\bfx|} |U|^3/(1+|U|^2) \in L^2(\bbR^d)$ for 
772: $\delta\leq\omega/2<\omega^{1/2}$, we deduce from \eqref{Eq_NLSSWCovolutionEst} 
773: that $e^{\delta|x|}U\in L^\infty(\bbR^d)$ for $0<\delta<\omega/2$.
774: \endproof
775: 
776: \remark 
777: Actually, the saturated cubic NLS equation should involve a small 
778: parameter $\eps>0$, namely, in the focusing case, we should consider instead 
779: of \eqref{Eq_NLSCauchy} 
780: \begin{equation}
781: \label{Eq_NLSeps}
782: i\partial_t A^\eps + \Delta A^\eps = - \frac{|A^\eps|^2A^\eps}{1+\eps|A^\eps|^2}.
783: \end{equation}
784: Theorem \ref{Th_NLSCauchy} is of course still valid for a fixed $\eps>0$, but 
785: \eqref{Eq_NLSH1Claw} and \eqref{Eq_NLSH1Est} should be replaced by
786: \begin{equation*}
787: \begin{aligned}
788: \int_{\bbR^d} &
789: \left[|\nabla A(\bfx,t)|^2 d\bfx + \frac1{\eps^2} \ln (1+|A(\bfx,t)|^2 \right]
790: d\bfx \\
791: & = \int_{\bbR^d} 
792: \left[|\nabla A_0(\bfx)|^2 d\bfx + \frac1{\eps^2} \ln (1+|A_0(\bfx)|^2 \right]
793: d\bfx,
794: \end{aligned}
795: \end{equation*}
796: \begin{equation*}
797: \int_{\bbR^d} |\nabla A(\bfx,t)|^2 d\bfx 
798: \leq \frac1\eps \int_{\bbR^d} |A_0(\bfx)|^2 d\bfx 
799: + \int_{\bbR^d} |\nabla A_0(\bfx)|^2 d\bfx. 
800: \end{equation*}
801: For solitary waves $A^\eps(\bfx,t)=e^{i\omega t} U(\bfx)$, \eqref{Eq_NLSeps} 
802: reduces to the elliptic equation
803: \begin{equation*}
804: -\Delta U + \omega U = \frac{|U|^2U}{1+\eps|U|^2}.
805: \end{equation*}
806: Setting $V=\eps^{1/2}U$, one obtains
807: \begin{equation*}
808: -\Delta V + \omega V = \frac1\eps\ \frac{|V|^2V}{1+|V|^2}.
809: \end{equation*}
810: The only possible range for the existence of nontrivial $H^1$ solitary waves
811: is $\omega\in]0,1/\eps[$. Proposition \ref{Prop_NLSSWDecay} is still valid for 
812: $\omega$ in this range.
813: 
814: \subsection{Solitary waves --- existence results}
815: 
816: We now turn to the existence of non-trivial $H^2$ solutions of
817: \begin{equation*}
818: -\Delta U + \omega U = \frac{|U|^2U}{1+|U|^2}
819: \end{equation*}
820: when $0<\omega<1$. We will look for real radial solutions
821: $U(\bfx)=u(|\bfx|)\equiv u(r)$ and thus consider 
822: the ODE problem
823: \begin{equation}
824: \label{Eq_NLSSWRadial}
825: \left\{
826: \begin{array}{l}
827: \ds - u'' - \frac{d-1}{r} u' + \omega u = \frac{u^3}{1+u^2}, \\
828: u \in H^2(]0,\infty[), \hspace{1cm} u'(0)=0.
829: \end{array}
830: \right.
831: \end{equation}
832: We recall a classical result of Berestycki \textit{et al.}% 
833: \cite{berestycki-lions-peletier}
834: 
835: \begin{theorem}[\refcite{berestycki-lions-peletier}, p. 143]
836: \label{Th_BLP}
837: Let $g$ be a locally Lipschitz continuous function on $\bbR_+=[0,+\infty[$ 
838: such that $g(0)=0$, satisfying the following hypotheses.
839: \setlength\leftmargini{2pc}
840: \begin{itemize}
841: \item[{\em (H1)}] $\alpha=\inf\{\zeta>0,\ g(\zeta) \geq 0\}$ exists and $\alpha>0$.
842: \item[{\em (H2)}] Let $G(t) = \int_0^t g(s) ds$. There exists $\zeta>0$ such that
843: $G(\zeta)>0$. \\
844: Let $\zeta_0=\inf\{\zeta>0,\ G(\zeta) \geq 0\}$. In view of {\em (H1)} and 
845: {\em (H2),} $\zeta_0$ exists and $\zeta_0>\alpha$. 
846: \item[{\em (H3)}] $\lim_{s\searrow\alpha} g(s)/(s-\alpha)>0$
847: \item[{\em (H4)}] $g(s)>0$ for $s\in]\alpha,\zeta_0]$. \\ 
848: Let $\beta=\inf\{\zeta>\zeta_0,\ G(\zeta) \geq 0\}$. In view of {\em (H4),} 
849: $\zeta_0<\beta\leq+\infty$.
850: \item[{\em (H5)}] If $\beta=+\infty$, then $\lim_{s\to+\infty} g(s)/s^l=0$ 
851: for some $l<(d+2)/(d-2)$ (if $d=2$, we may choose for $l$ any finite real 
852: number).
853: \end{itemize}
854: Let us consider the Cauchy problem
855: \begin{equation}
856: \label{Eq_BPL}
857: \left\{
858: \begin{array}{ll}
859: \ds - u'' - \frac{d-1}{r} u' = g(u),\ & r>0, \\
860: u(0)=\zeta, & u'(0)=0.
861: \end{array}
862: \right.
863: \end{equation}
864: Then there exists $\zeta\in]\zeta_0,\beta[$ such that \eqref{Eq_BPL} has a 
865: unique solution satisfying $u(r)>0$ for $r\in]0,+\infty[$, $u'(r)<0$ for 
866: $r\in]0,+\infty[$ and $\lim_{r\to+\infty}u(r)=0$. If in addition 
867: $\limsup_{s\searrow0} g(s)/s<0$, then there exists $C>0$ and $\delta>0$ 
868: such that $0<u(r)\leq Ce^{-\delta r}$, for $0\leq r<+\infty$.
869: \end{theorem}
870: 
871: \begin{theorem}
872: If $a=1$ and $0<\omega<1$, there exists a nontrivial positive solution of 
873: \eqref{Eq_NLSSWRadial}.
874: \end{theorem}
875: 
876: \proof
877: The case $d=1$ has been addressed in Sec. \ref{Sec_1D}. Consider now 
878: $d\geq2$. We apply Theorem \ref{Th_BLP} with 
879: \begin{equation*}
880: g(u)=-\omega u + \frac{u^3}{1+u^2},
881: \end{equation*} 
882: which graph is displayed in Fig. \ref{Fig_1}.
883: 
884: \begin{figure}[b]
885: \centerline{\psfig{file=Fig1.eps,width=5cm}}
886: \vspace*{8pt}
887: \caption{\label{Fig_1}Graph of function $g$}
888: \end{figure}
889: 
890: Note that $\alpha=\sqrt{\omega/(1-\omega)}$ which yields (H1). Setting 
891: $u=\sqrt{\omega/(1-\omega)}+\eps$, one easily checks that
892: \begin{equation*}
893: \frac{g(u)}{u-\alpha} = 2\omega(1-\omega) + O(\eps),
894: \end{equation*}
895: and (H3) is satisfied. One computes $G(u)=(1-\omega)u^2-\frac12 \ln(1+u^2)$, 
896: which obviously satisfies (H2) and (H4) with $\beta=+\infty$. Last (H5) holds 
897: true (for $l>1$).
898: \endproof
899: 
900: \remark 
901: $u$ satisfies the decay rate of Proposition \ref{Prop_NLSSWDecay}.
902: 
903: \section{The Zozulya--Anderson System}
904: \label{Sec_Cauchy}
905: 
906: \subsection{Estimate on the potential}
907: 
908: We now restrict to the space-dimension $d=2$ which is the context of the
909: derivation. To mimic the proof for the Cauchy problem in the one-dimensional 
910: case, we would like to express $\ph$ in terms of $A$ for say 
911: $A\in L^2(\bbR^2)$. With such a data $A$, we indeed have a unique $\ph$ in some 
912: convenient space but no Lipschitz regularity for the mapping $A\mapsto\ph$,
913: which is required to perform some fixed point procedure. To ensure this we 
914: will have to assume $A\in H^2(\bbR^2)$. 
915: 
916: To derive the first estimates, we consider time as a parameter and do not 
917: express it. We therefore introduce the weighted homogeneous Sobolev space 
918: \begin{equation*}
919: H = \{\ph\in\calS'(\bbR^d),\ (1+|A|^2)^{1/2}\nabla\ph\in L^2(\bbR^d)\}\slash\bbR
920: \end{equation*}
921: together with its natural Hilbertian structure.
922: 
923: \begin{lemma}
924: \label{Lm_Invdivgrad}
925: {\em (i)} Let $A\in L^2(\bbR^2)$. There exists a unique $\ph\in H$ solution of
926: \begin{equation}
927: \label{Eq_ZAPotential}
928: \div ((1+|A|^2)\nabla\ph) = \partial_x (|A|^2) \textrm{ in } \calD'(\bbR^2)
929: \end{equation}
930: such that
931: \begin{equation}
932: \label{Eq_ZAH1Est}
933: \int_{\bbR^2} (1+\frac12|A|^2)|\nabla\ph|^2 d\bfx 
934: \leq \frac12 \int_{\bbR^2} |A|^2 d\bfx. 
935: \end{equation}
936: {\em (ii)} If furthermore $A\in H^2(\bbR^2)$, then $\nabla\ph\in H^2(\bbR^2)$ and 
937: there exists a polynomial $P$ vanishing at 0 such that
938: \begin{equation}
939: \label{Eq_P}
940: \|\nabla \ph\|_{H^2(\bbR^2)} \leq P(\|A\|_{H^2(\bbR^2)}).
941: \end{equation}
942: \end{lemma}
943: 
944: \proof
945: (i) We define a smoothing sequence $(\theta_\eps)_{\eps>0}$ with
946: $\int_{\bbR^2}\theta_\eps d\bfx=1$ and $A_\eps=A\star\theta_\eps$ is such that 
947: $A_\eps\to A\in L^2(\bbR^2)$. In particular
948: \begin{equation}
949: \label{Eq_ZAUnif}
950: \|A_\eps\|_{L^2(\bbR^2)} \leq \|A\|_{L^2(\bbR^2)}. 
951: \end{equation}
952: By Riesz theorem there exists a unique solution to 
953: \begin{equation}
954: \label{Eq_ZAPotentialeps}
955: \div ((1+|A_\eps|^2)\nabla\ph_\eps) = \partial_x (|A_\eps|^2),
956: \end{equation}
957: i.e. 
958: \begin{equation}
959: \label{Eq_ZAEllipticeps}
960: -\Delta \ph_\eps - \div(|A_\eps|^2\nabla\ph_\eps) = \partial_x (|A_\eps|^2)
961: \end{equation}
962: after noticing that the R.H.S. of Eq. \eqref{Eq_ZAPotentialeps} defines a 
963: linear continuous form on $H$ given by
964: \begin{equation*}
965: \langle \partial_x(|A_\eps|^2),\psi \rangle 
966: = \int_{\bbR^2} |A_\eps|^2 \partial_x\psi d\bfx.
967: \end{equation*}
968: 
969: \noindent
970: Now we get from \eqref{Eq_ZAPotentialeps}
971: \begin{eqnarray*}
972: \int_{\bbR^2} (1+|A_\eps|^2)|\nabla\ph_\eps|^2 d\bfx 
973: & = & - \int_{\bbR^2} |A_\eps|^2 \partial_x\ph_\eps d\bfx \\
974: & \leq & \frac12 \int_{\bbR^2} |A_\eps|^2  d\bfx 
975: + \frac12  \int_{\bbR^2} |A_\eps|^2 (\partial_x\ph_\eps)^2 d\bfx,
976: \end{eqnarray*}
977: which yields (together with \eqref{Eq_ZAUnif})
978: \begin{equation}
979: \label{Eq_ZAH1Esteps}
980: \int_{\bbR^2} (1+\frac12|A_\eps|^2)|\nabla\ph_\eps|^2 d\bfx 
981: \leq \frac12 \int_{\bbR^2} |A_\eps|^2 d\bfx
982: \leq \frac12 \int_{\bbR^2} |A|^2 d\bfx. 
983: \end{equation}
984: Up to the extraction of a sub-sequence, we have $\nabla \ph_\eps\to\nabla\ph$
985: and $\partial_x (|A_\eps|^2)\to\partial_x (|A|^2)$ in $\calD'(\bbR^2)$. From 
986: Eq. \eqref{Eq_ZAH1Esteps}, $A_\eps\nabla\ph_\eps\rightharpoonup B$ weakly
987: in $L^2$ and for all $\psi\in\calD$, 
988: \begin{equation*}
989: \int_{\bbR^2} A_\eps\nabla\ph_\eps\cdot\nabla\psi d\bfx 
990: \to
991: \int_{\bbR^2} A\nabla\ph\cdot\nabla\psi d\bfx, 
992: \end{equation*}
993: therefore $B=A\nabla\ph$.  Since
994: $||A_\eps|^2\nabla\ph_\eps|=|A_\eps||A_\eps\nabla\ph_\eps|$,
995: $|A_\eps|^2\nabla\ph_\eps\to|A|^2\nabla\ph$ in $\calD'$. We can pass to the
996: limit in Eq. \eqref{Eq_ZAEllipticeps} and obtain 
997: \begin{equation*}
998: -\Delta \ph - \div(|A|^2\nabla\ph) = \partial_x (|A|^2) 
999: \textrm{ in } \calD'(\bbR^2),
1000: \end{equation*}
1001: i.e. $\div ((1+|A|^2)\nabla\ph) = \partial_x (|A|^2)$ and deduce estimate 
1002: \eqref{Eq_ZAH1Est} from \eqref{Eq_ZAH1Esteps}. This yields the existence of
1003: $\ph\in H$. The uniqueness is straightforward: two solutions $\ph_1$ and
1004: $\ph_2$ would satisfy
1005: \begin{equation*}
1006: \int_{\bbR^2} (1+\frac12|A|^2)|\nabla(\ph_1-\ph_2)|^2 d\bfx = 0, 
1007: \textrm{ i.e. }\nabla(\ph_1-\ph_2)=0\textrm{ a.e. }
1008: \end{equation*}
1009: and hence be equal in $H$. \\[2mm]
1010: (ii) We first notice that $|A|^2\Delta\ph$ is meaningful in 
1011: $H^{-1}(\bbR^2)$. Actually, for any $\psi\in H^1(\bbR^2)$, one defines
1012: \begin{equation*}
1013: \langle |A|^2\Delta\ph,\psi\rangle_{H^{-1}(\bbR^2),H^1(\bbR^2)}
1014: = \langle \Delta\ph,|A|^2\psi\rangle_{H^{-1}(\bbR^2),H^1(\bbR^2)},
1015: \end{equation*}
1016: which makes sense since $|A|^2\psi\in H^1(\bbR^2)$ for $A\in H^2(\bbR^2)$,
1017: $\psi\in H^1(\bbR^2)$. Thus we can write 
1018: \eqref{Eq_ZAPotential} as
1019: \begin{equation*}
1020: (1+|A|^2)\Delta\ph = - \nabla|A|^2\cdot\nabla\ph + \partial_x(|A|^2),
1021: \end{equation*}
1022: and
1023: \begin{equation*}
1024: \Delta\ph = - \frac{\nabla|A|^2}{1+|A|^2}\cdot\nabla\ph 
1025: + \frac{\partial_x(|A|^2)}{1+|A|^2} =: F.
1026: \end{equation*}
1027: We claim that $F\in L^r(\bbR^2)$, for any $r\in(1,2)$, with 
1028: \begin{equation*}
1029: \|F\|_{L^r(\bbR^2)} \leq C \|A\|_{L^2(\bbR^2)} \|A\|_{H^2(\bbR^2)}.
1030: \end{equation*}
1031: First, $|\nabla|A|^2\cdot\nabla\ph| \leq 2|\nabla A||A\nabla\ph|$ and by
1032: H\"older
1033: \begin{equation*}
1034: \| \nabla|A|^2\cdot\nabla\ph \|_{L^r(\bbR^2)} 
1035: \leq 2 \|\nabla A\|_{L^p(\bbR^2)} \|A\nabla \ph\|_{L^2(\bbR^2)}
1036: \end{equation*}
1037: for any $1<r<2$ and $p=2r/(2-r)\in(2,\infty)$. Since
1038: $\|A\nabla\ph\|_{L^2(\bbR^2)} \leq \|A\|_{L^2(\bbR^2)}$ and 
1039: $H^1(\bbR^2)\subset L^q(\bbR^2)$ for all $q>2$, we obtain that
1040: \begin{equation*}
1041: \left\|\frac{\nabla|A|^2}{1+|A|^2}\cdot\nabla\ph\right\|_{L^r(\bbR^2)}
1042: \leq C \|A\|_{H^2(\bbR^2)} \|A\|_{L^2(\bbR^2)}, \hspace{1cm} 1<r<2.  
1043: \end{equation*}
1044: Similarly
1045: \begin{eqnarray*}
1046: \left\|\frac{\partial_x|A|^2}{1+|A|^2}\right\|_{L^r(\bbR^2)}
1047: & \leq & 2 \|A \partial_x A\|_{L^r(\bbR^2)} 
1048: \leq 2 \|A\|_{L^2(\bbR^2)} \|\partial_x A\|_{L^p(\bbR^2)} \\[2mm]
1049: & \leq & C \|A\|_{H^2(\bbR^2)} \|A\|_{L^2(\bbR^2)}, \hspace{1cm} 1<r<2.  
1050: \end{eqnarray*}
1051: 
1052: By elliptic regularity, we infer thus that for any $r$, $1<r<2$,
1053: \begin{equation*}
1054: \|\nabla\ph\|_{W^{1,r}(\bbR^2)} \leq C \|A\|_{H^2(\bbR^2)} \|A\|_{L^2(\bbR^2)}.  
1055: \end{equation*}
1056: By Sobolev embedding,
1057: \begin{equation*}
1058: \|\nabla\ph\|_{L^q(\bbR^2)} 
1059: \leq C \|\nabla\ph\|_{W^{1,r}(\bbR^2)} 
1060: \leq C \|A\|_{H^2(\bbR^2)} \|A\|_{L^2(\bbR^2)}.  
1061: \end{equation*}
1062: for $\frac 1q = \frac1r - \frac12$, i.e. $q=2r/(2-r)$ for all $r$, $1<r<2$. Thus
1063: for any $p>2$
1064: \begin{eqnarray*}
1065: \left\|\frac{\nabla|A|^2}{1+|A|^2}\cdot\nabla\ph\right\|_{L^p(\bbR^2)}
1066: & \leq & \|\nabla \ph\|_{L^{2p}(\bbR^2)} \|\nabla |A|^2\|_{L^{2p}(\bbR^2)} \\[2mm]
1067: & \leq & C \|A\|_{H^2(\bbR^2)} \|A\|_{L^2(\bbR^2)} \|A\|^2_{H^2(\bbR^2)} \\[2mm]
1068: & = & C \|A\|_{L^2(\bbR^2)} \|A\|^3_{H^2(\bbR^2)}
1069: \end{eqnarray*}
1070: (we have used the fact that $H^2(\bbR^2)$ is an algebra and the embedding 
1071: $H^1(\bbR^2)\subset L^q(\bbR^2)$ for all $q>2$). 
1072: 
1073: Similarly, for any $p>2$
1074: \begin{equation*}
1075: \left\|\frac{\partial_x(|A|^2)}{1+|A|^2}\right\|_{L^p(\bbR^2)}
1076: \leq 2 \|A\|_{L^{2p}(\bbR^2)} \|\partial_x A\|_{L^{2p}(\bbR^2)} \\[2mm]
1077: \leq C \|A\|_{H^1(\bbR^2)} \|A\|_{H^2(\bbR^2)}.
1078: \end{equation*}
1079: Finally for any $p>2$ 
1080: \begin{equation*}
1081: \|F\|_{L^p(\bbR^2)} 
1082: \leq C \|A\|^2_{H^2(\bbR^2)} (1+\|A\|_{L^2(\bbR^2)}\|A\|_{H^2(\bbR^2)}).  
1083: \end{equation*}
1084: and by elliptic regularity
1085: \begin{equation*}
1086: \|\nabla\ph\|_{W^{1,p}(\bbR^2)} 
1087: \leq C \|A\|^2_{H^2(\bbR^2)} (1+\|A\|_{L^2(\bbR^2)}\|A\|_{H^2(\bbR^2)}),
1088: \hspace{1cm} \forall p>2.   
1089: \end{equation*}
1090: We now check that 
1091: $\nabla\ph\cdot\nabla|A|^2/(1+|A|^2)\in H^1(\bbR^2)$. This easily reduces to
1092: showing that $\nabla(\nabla\ph\cdot\nabla|A|^2)\in L^2(\bbR^2)$. For 
1093: $(i,j)\in\{1,2\}$, $\partial_{x_i}\partial_{x_j}\ph\in L^p(\bbR^2)$ since 
1094: \begin{equation*}
1095: \widehat{\partial_{x_i}\partial_{x_j}\ph} =
1096: \frac{\xi_i\xi_j}{|\xi|^2}\widehat{\Delta\ph} 
1097: \textrm{ and } \Delta\ph\in L^p(\bbR^2),\ p>2. 
1098: \end{equation*}
1099: Thus $\partial_{x_i}\partial_{x_j}\ph \nabla|A|^2 \in L^2(\bbR)$
1100: ($\nabla|A|^2\in H^1(\bbR^2)\subset L^q(\bbR^2)$, $\forall q>2$). 
1101: 
1102: On the other hand, taking $p>2$ we see that $\nabla\ph\in L^\infty(\bbR^2)$ and
1103: thus $\nabla\ph \partial_{x_i}\partial_{x_j} |A|^2 \in L^2(\bbR^2)$. 
1104: 
1105: It is also easy to check that $\partial_x(|A|^2)/(1+|A|^2)\in H^1(\bbR^2)$. 
1106: 
1107: \noindent
1108: Finally, $\Delta\ph = F \in H^1(\bbR^2)$, proving that 
1109: $\nabla\ph\in H^2(\bbR^2)$ with an estimate of the form 
1110: \begin{equation*}
1111: \|\nabla \ph\|_{H^2(\bbR^2)} \leq P(\|A\|_{H^2(\bbR^2)}),
1112: \end{equation*} 
1113: where $P$ is a polynomial vanishing at 0, which proves \eqref{Eq_P}.
1114: \endproof
1115: 
1116: \remark
1117: All above estimates are therefore uniform in time, and if
1118: $A\in\calC([0,T];H^2(\bbR^2))$ for some $T>0$, one has
1119: \begin{equation*}
1120: \|\nabla \ph\|_{\calC([0,T];H^2(\bbR^2))} 
1121: \leq P(\|A\|_{\calC([0,T];H^2(\bbR^2))}).
1122: \end{equation*}
1123: 
1124: \subsection{Solitary waves --- non existence results}
1125: 
1126: We now look for solitary wave solutions of \eqref{Eq_Adiv}, that is solutions of 
1127: the form $(e^{i\omega t}U(x),\phi(x))$ with $x\in\bbR^d$, $\omega\in\bbR$, 
1128: $U\in H^1(\bbR^d)$, and $\phi\in H$. Thus $(U,\phi)$ should satisfy the
1129: system
1130: \begin{equation}
1131: \label{Eq_ZASW}
1132: \left\{
1133: \begin{array}{l}
1134: - \Delta U + \omega U = a U \partial_x\phi, \\[2mm]
1135: \div ( (1+|U|^2) \nabla \phi ) = \partial_x(|U|^2).
1136: \end{array}
1137: \right.
1138: \end{equation}
1139: The existence of nontrivial solutions of \eqref{Eq_ZASW} is an open problem. 
1140: Note that \eqref{Eq_ZASW} does not seem to be the Euler--Lagrange equation 
1141: associated to a variational problem. We have however:
1142: \begin{proposition}
1143: {\em (i)} Let $a=-1$ {\em (}defocusing case{\em )}. Then no nontrivial solution  
1144: of \eqref{Eq_ZASW} exists for $\omega\geq0$. \\
1145: {\em (ii)} Let $a=1$ {\em (}focusing case{\em )}. No nontrivial solution of 
1146: \eqref{Eq_ZASW} exists for $\omega\geq1$. \\
1147: {\em (iii)} Let $a=\pm1$. No nontrivial solution of \eqref{Eq_ZASW} exists 
1148: if $\omega<0$ provided $\partial_x\phi = O(1/|\bfx|^{1+\eps})$, $\eps>0$ as 
1149: $|\bfx|\to+\infty$.
1150: \end{proposition}
1151: 
1152: \proof
1153: From \eqref{Eq_ZASW} we have
1154: \begin{equation*}
1155: \int_{\bbR^d} |\nabla U|^2 d\bfx + \omega \int_{\bbR^d} |U|^2 d\bfx 
1156: = a \int_{\bbR^d} |U|^2 \partial_x\phi d\bfx,
1157: \end{equation*}
1158: \begin{equation*}
1159: \int_{\bbR^d} (1+|U|^2)|\nabla\phi|^2 d\bfx 
1160: = - \int_{\bbR^d} |U|^2 \partial_x\phi d\bfx,
1161: \end{equation*}
1162: and 
1163: \begin{equation}
1164: \label{Eq_ZASWH1Claw}
1165: \int_{\bbR^d} |\nabla U|^2 d\bfx + \omega \int_{\bbR^d} |U|^2 d\bfx 
1166: - a \int_{\bbR^d} (1+|U|^2)|\nabla\phi|^2 d\bfx = 0,
1167: \end{equation}
1168: which proves (i). Now independent of the sign of $a$, and from 
1169: \eqref{Eq_ZAH1Est} and \eqref{Eq_ZASWH1Claw},
1170: \begin{equation*}
1171: \int_{\bbR^d} |\nabla U|^2 d\bfx + \omega \int_{\bbR^d} |U|^2 d\bfx 
1172: \leq \int_{\bbR^d} |U|^2 d\bfx. 
1173: \end{equation*}
1174: Thus
1175: \begin{equation*}
1176: \int_{\bbR^d} |\nabla U|^2 d\bfx + (\omega-1) \int_{\bbR^d} |U|^2 d\bfx \leq0,
1177: \end{equation*}
1178: which proves (ii). Part (iii) results from \refcite{kato}.
1179: \endproof
1180: 
1181: \subsection{The Cauchy problem}
1182: 
1183: We consider the system 
1184: \begin{equation}
1185: \label{Eq_ZACauchy}
1186: \left\{
1187: \begin{array}{l}
1188: i\partial_t A + \Delta A = -aA\partial_x\ph, \\[2mm]
1189: \div \left((1+|A|^2) \nabla \ph\right) = \partial_x(|A|^2), \\[2mm]
1190: A(\cdot,0) = A_0.
1191: \end{array}
1192: \right.
1193: \end{equation}
1194: 
1195: \begin{theorem}
1196: \label{Th_ZACauchy}
1197: Let $A_0\in H^2(\bbR^2)$. \\
1198: Then there exists $T_0>0$ and a unique solution $(A,\nabla\ph)$ of 
1199: \eqref{Eq_ZACauchy} such that $A\in\calC([0,T_0];H^2(\bbR^2))$ and 
1200: $\nabla\ph\in\calC([0,T_0];H^2(\bbR^2))$. Moreover
1201: \begin{equation*}
1202: \|A(\cdot,t)\|_{L^2(\bbR^2)} = \|A_0\|_{L^2(\bbR^2)}, \hspace{1cm}0\leq t\leq T_0
1203: \end{equation*}
1204: and
1205: \begin{equation*}
1206: \int_{\bbR^2} (1+\frac12|A|^2) |\nabla\ph|^2 d\bfx 
1207: \leq \frac12 \int_{\bbR^2} |A_0|^2 d\bfx, \hspace{1cm}0\leq t\leq T_0.
1208: \end{equation*}
1209: 
1210: \end{theorem}
1211: 
1212: \proof
1213: \textit{Uniqueness.} Let $(A,\nabla\ph)\in L^\infty(0,T;H^2(\bbR^2))$ and  
1214: $(B,\nabla\psi)\in L^\infty(0,T;H^2(\bbR^2))$ two solutions of
1215: \eqref{Eq_ZACauchy} with $A(\cdot,0)=B(\cdot,0)$. Then from
1216: \eqref{Eq_ZACauchy}$_2$ one gets
1217: \begin{equation*}
1218: \Delta (\ph-\psi) + \div(|A|^2\nabla\ph-|B|^2\nabla\psi)
1219: = \partial_x(|A|^2) - \partial_x(|B|^2),
1220: \end{equation*}
1221: yielding
1222: \begin{equation}
1223: \label{Eq_ZACauchy_unique}
1224: \begin{aligned}
1225: \int_{\bbR^2} & |\nabla (\ph-\psi)|^2 d\bfx 
1226: + \int_{\bbR^2} |A|^2 |\nabla(\ph-\psi)|^2  d\bfx \\
1227: & = \int_{\bbR^2} (|A|^2-|B|^2)\partial_x (\ph-\psi) d\bfx 
1228: - \int_{\bbR^2} (|A|^2-|B|^2) \nabla\psi\cdot\nabla(\ph-\psi) d\bfx .
1229: \end{aligned}
1230: \end{equation}
1231: Observing that $|A|^2-|B|^2=A(A-\bar B)+\bar B(A-B)$, the R.H.S. of
1232: \eqref{Eq_ZACauchy_unique} is majorized by
1233: \begin{equation*}
1234: \begin{aligned}
1235: \frac14 \int_{\bbR^2} |\partial_x & (\ph-\psi)|^2 d\bfx 
1236: + (\|A\|_{L^\infty(\bbR^2)}+\|B\|_{L^\infty(\bbR^2)}) 
1237: \int_{\bbR^2} |A-B|^2 d\bfx \\
1238: & + \frac14 \int_{\bbR^2} |\nabla (\ph-\psi)|^2 d\bfx \\
1239: & + (\|A\|_{L^\infty(\bbR^2)}+\|B\|_{L^\infty(\bbR^2)})\|
1240: \nabla\psi\|_{L^\infty(\bbR^2)} \int_{\bbR^2} |A-B|^2 d\bfx 
1241: \end{aligned}
1242: \end{equation*}
1243: 
1244: \noindent
1245: and by Sobolev embedding
1246: \begin{equation}
1247: \label{Eq_ZACauchy_nablapsi}
1248: \|\nabla(\ph-\psi)\|_{L^2(\bbR^2)} 
1249: \leq C(\|A\|_{H^2(\bbR^2)},\|B\|_{H^2(\bbR^2)},\|\nabla\psi\|_{H^2(\bbR^2)})
1250: \|A-B\|_{L^2(\bbR^2)}.
1251: \end{equation}
1252: On the other hand, we obtain readily from \eqref{Eq_ZACauchy}$_1$ that
1253: \begin{equation*}
1254: \frac12 \frac{d}{dt} \int_{\bbR^2} |A-B|^2 d\bfx 
1255: \leq \int_{\bbR^2} |A-B|^2 |\partial_x\ph| d\bfx
1256: + \int_{\bbR^2} |B||\partial_x(\ph-\psi)||A-B| d\bfx
1257: \end{equation*}
1258: which together with \eqref{Eq_ZACauchy_nablapsi} and the Cauchy-Schwarz lemma
1259: yields
1260: \begin{equation*}
1261: \begin{aligned}
1262: \frac12 \frac{d}{dt} & \int_{\bbR^2} |A-B|^2 d\bfx \\
1263: & \leq C(\|A\|_{H^2(\bbR^2)},\|B\|_{H^2(\bbR^2)},\|\nabla\ph\|_{H^2(\bbR^2)},
1264: \|\nabla\psi\|_{H^2(\bbR^2)}) \|A-B\|_{L^2(\bbR^2)}
1265: \end{aligned}
1266: \end{equation*}
1267: and $A=B$ by Gronwall lemma. \\[2mm]
1268: \textit{$H^2$ a priori estimate.} We derive a (formal) $H^2$ 
1269: \textit{a priori} estimate on the solution of \eqref{Eq_ZACauchy}. Since 
1270: $H^2(\bbR^2)$ is an algebra, we deduce from Lemma~\ref{Lm_Invdivgrad} that 
1271: \begin{equation}
1272: \label{Eq_EstH2_Aphx}
1273: \|A\partial_x \ph\|_{\calC([0,T];H^2(\bbR^2))} 
1274: \leq \|A\|_{\calC([0,T];H^2(\bbR^2))} P(\|A\|_{\calC([0,T];H^2(\bbR^2))}), 
1275: \end{equation}
1276: where $P$ was introduced in \eqref{Eq_P}. From the energy estimate
1277: \begin{equation*}
1278: \frac12 \frac{d}{dt} \|A(\cdot,t)\|^2_{H^2(\bbR^2)} 
1279: \leq C \|A\partial_x \ph (\cdot,t)\|_{H^2(\bbR^2)}\|A(\cdot,t)\|_{H^2(\bbR^2)},
1280: \end{equation*}
1281: we infer with \eqref{Eq_EstH2_Aphx} the local $H^2$ bound
1282: \begin{equation}
1283: \label{Eq_EstH2_A}
1284: \|A(\cdot,t)\|_{H^2(\bbR^2)} \leq C \left(\|A_0\|_{H^2(\bbR^2)} \right)
1285: \textrm{ for } 0<t<T_0, 
1286: \end{equation}
1287: $T_0<T$ sufficiently small. \\[2mm]
1288: \textit{Approximation of \eqref{Eq_ZACauchy}.} The strategy is now to implement
1289: a compactness method using the (justified) \textit{a priori} estimate 
1290: \eqref{Eq_EstH2_A}. For $\eps>0$, we consider the system
1291: \begin{eqnarray}
1292: \label{Eq_ZACauchy_epsA}
1293: i\partial_t A^\eps + \Delta A^\eps & = & -aA^\eps\partial_x\ph^\eps, \\[2mm]
1294: \label{Eq_ZACauchy_epsph}
1295: \div \left((1+\eps \Delta^2+ |A^\eps|^2) \nabla \ph^\eps\right) 
1296: & = & \partial_x(|A^\eps|^2), \\[2mm]
1297: \label{Eq_ZACauchy_epsA0}
1298: A^\eps(\cdot,0) & = & A_0. 
1299: \end{eqnarray}
1300: Solving $\nabla\ph^\eps$ in terms of $A^\eps$, we obtain from
1301: \eqref{Eq_ZACauchy_epsph} that $\nabla\ph^\eps$ satisfies
1302: \begin{equation}
1303: \label{Eq_ZACauchy_uniform}
1304: \eps \int_{\bbR^2} |\Delta\nabla\ph^\eps|^2 d\bfx
1305: + \int_{\bbR^2} (1+\frac12|A^\eps|^2)|\nabla\ph^\eps|^2 d\bfx
1306: \leq \frac12 \int_{\bbR^2} |A^\eps|^2 d\bfx.
1307: \end{equation}
1308: 
1309: \noindent
1310: \textit{Well-posedness of approximate system.} We now check that the Cauchy 
1311: problem \eqref{Eq_ZACauchy_epsA}--\eqref{Eq_ZACauchy_epsA0} is globally 
1312: well-posed in $H^2(\bbR^2)$. Let first $A^\eps$, $B^\eps\in H^2(\bbR^2)$ and 
1313: $\ph^\eps$, $\psi^\eps$ the corresponding solutions of
1314: \eqref{Eq_ZACauchy_epsph}. Proceeding as in the uniqueness proof above, one 
1315: gets 
1316: \begin{equation}
1317: \label{Eq_ZACauchy_unifeps}
1318: \begin{aligned}
1319: \eps \int_{\bbR^2} & |\nabla\Delta (\ph^\eps-\psi^\eps)|^2 d\bfx 
1320: + \int_{\bbR^2} |\nabla(\ph^\eps-\psi^\eps)|^2 d\bfx \\
1321: & \leq C(\|A^\eps\|_{H^2(\bbR^2)},\|B^\eps\|_{H^2(\bbR^2)},
1322: \|\nabla\ph^\eps\|_{H^2(\bbR^2)},\|\nabla\psi^\eps\|_{H^2(\bbR^2)}) 
1323: \|A^\eps-B^\eps\|_{L^2(\bbR^2)}.
1324: \end{aligned}
1325: \end{equation}
1326: Denoting $\partial_x\ph^\eps$ by $F^\eps(A^\eps)$ we write 
1327: \eqref{Eq_ZACauchy_epsA} on the Duhamel form with $S(t)=\exp(it\Delta)$,
1328: \begin{equation}
1329: \label{Eq_ZACauchy_Duhameleps}
1330: A^\eps(t) = S(t) A_0 - a \int_0^t S(t-s)A^\eps F^\eps(A^\eps) ds.
1331: \end{equation}
1332: Using \eqref{Eq_ZACauchy_unifeps} and the unitarity of $S(t)$ in $H^s(\bbR^2)$,
1333: we deduce that the R.H.S. of \eqref{Eq_ZACauchy_Duhameleps} defines a
1334: contraction in $\calC([0,T_\eps];H^2(\bbR^2))$ for some $T_\eps>0$. 
1335: 
1336: This implies the local well-posedness of 
1337: \eqref{Eq_ZACauchy_epsA}--\eqref{Eq_ZACauchy_epsA0} in $H^2(\bbR^2)$. Using the 
1338: $H^2$ bound \eqref{Eq_ZACauchy_unifeps} on $\nabla\ph^\eps$, we infer from
1339: \eqref{Eq_ZACauchy_epsA} an \textit{a priori} bound in 
1340: $\calC([0,T];H^2(\bbR^2))$ for $A^\eps$ and for all $T>0$. This proves that the
1341: Cauchy problem \eqref{Eq_ZACauchy_epsA}--\eqref{Eq_ZACauchy_epsA0} is globally
1342: well-posed, for any fixed $\eps>0$.\\[2mm]
1343: \textit{Limit $\eps\to0$.} Now we have the bounds \eqref{Eq_ZACauchy_uniform} 
1344: and 
1345: \begin{equation}
1346: \label{Eq_ZACauchy_boundeps}
1347: \|\nabla\ph^\eps(\cdot,t)\|_{L^2(\bbR^2)} 
1348: + \sqrt\eps \|\Delta\nabla\ph^\eps(\cdot,t)\|_{L^2(\bbR^2)} \leq C, 
1349: \hspace{1cm} 0\leq t\leq T,
1350: \end{equation}
1351: where $C$ and $T$ do not depend on $\eps$. Moreover, from
1352: \eqref{Eq_ZACauchy_epsA} and \eqref{Eq_ZACauchy_boundeps} we have a bound on
1353: $\partial_t A^\eps$ which is independent of $\eps$: 
1354: \begin{equation*}
1355: \|\partial_t A^\eps(\cdot,t)\|_{L^2(\bbR^2)} \leq C, 
1356: \hspace{1cm} 0\leq t\leq T.
1357: \end{equation*}
1358: It is now standard to pass to the limit as $\eps\to0$ (see \refcite{lions}). 
1359: By the Aubin--Lions compactness lemma, we obtain a subsequence
1360: $(A^\eps,\nabla\ph^\eps)$ such that $A^\eps\to A$ in $L^\infty(0,T;H^2(\bbR^2))$
1361: weak-star and $L^2(0,T;H^1_{\textrm{loc}}(\bbR^2))$ strongly, 
1362: $\nabla\ph^\eps\to \nabla\ph$ in $L^\infty(0,T;H^2(\bbR^2))$
1363: weak-star and $L^2([0,T]\times\bbR^2)$ weakly. The limit 
1364: $(A,\nabla\ph)$ belongs to $(L^\infty(0,T;H^2(\bbR^2)))^2$ and satisfies
1365: \eqref{Eq_ZACauchy}. In fact \eqref{Eq_ZACauchy}$_1$ is satisfied in
1366: $L^2(\bbR^2)$ and \eqref{Eq_ZACauchy}$_2$ is satisfied in $H^1(\bbR^2)$.
1367: 
1368: The fact that $(A,\nabla\ph)\in (\calC(0,T;H^2(\bbR^2)))^2$ results from the
1369: Bona--Smith approximation (see \refcite{bona-smith}).
1370: \endproof
1371: 
1372: \remark 
1373: We do not know whether the local solution obtained in Theorem \ref{Th_ZACauchy}
1374: is global or not.
1375: 
1376: \section{Conclusion}
1377: 
1378: We have given a full description of how to derive from the Kukhtarev equations
1379: an asymptotic model for the propagation of light in a photorefractive
1380: medium. This derivation is only heuristic insofar as asymptotics are not
1381: justified, which would be out of reach now. Some properties of photorefractive
1382: media such as memory have also been neglected.
1383: 
1384: The 1D asymptotic model is a saturated nonlinear Schr\"odinger equation the 
1385: Cauchy problem of which is studied (in any space dimension) in $L^2$ and $H^1$. 
1386: We also prove the existence of solitary waves in one and higher dimensions. An 
1387: interesting and open issue would be to study the transverse stability of the 
1388: 1D solitary waves in the framework of the asymptotic model.
1389: 
1390: For the 2D asymptotic model (the Zozulya--Anderson model) we also have studied 
1391: the Cauchy problem and the non-existence of solitary waves. The question of
1392: imposing other boundary conditions, not vanishing in one space direction, can 
1393: also be addressed to treat a wider range of experimental applications.   
1394: 
1395: \appendix
1396: 
1397: \section{Non-Existence of Solitary Waves in Non-Physical Cases}
1398: 
1399: The goal is here to complete the results of Corollary \ref{Cor_NLSnoSW} for 
1400: $\omega<0$ with no decaying assumption. We have already seen that Eq.
1401: \eqref{Eq_NLSSWPohozaev} implies that no solitary wave may exist for $d=1,2$ and
1402: $a=1$ (focusing case). 
1403: 
1404: To go further, let us use both Eqs. \eqref{Eq_NLSSWEnergy} and
1405: \eqref{Eq_NLSSWPohozaev} to obtain
1406: \begin{equation*}
1407: \int_{\bbR^d} \left(2\omega + \frac{(d-2)a|U|^2}{1+|U|^2}\right) |U|^2 d\bfx 
1408: - ad  \int_{\bbR^d} \left[ |U|^2 - \ln (1+|U|^2) \right] d\bfx = 0.
1409: \end{equation*}
1410: We set 
1411: \begin{equation*}
1412: F(X) = \left(2\omega + \frac{(d-2)aX}{1+X}\right) X - ad (X - \ln (1+X)),
1413: \end{equation*}
1414: and we know that $\int_{\bbR^d} F(|U|^2) d\bfx = 0$. Now $F(0)=0$ and 
1415: \begin{equation*}
1416: F'(X)= \frac{2X^2(\omega-a)+X(4\omega-(4-d)a)+2\omega}{(1+X)^2} < 0, 
1417: \end{equation*}
1418: if $\omega<0$, $a=1$ and $d=3,4$.
1419: Therefore $F(|U|^2)=0$ a.e. By a bootstrapping argument, we notice that any
1420: $H^1$ solution to Eq. \eqref{Eq_NLSSW} is indeed in $H^k$ for all $k$ and
1421: therefore continuous. Hence $F(|U|^2)=0$ on $\bbR^d$. Since $F'(X)<0$ the only
1422: possible value for $U$ is $U=0$ on $\bbR^d$.
1423: 
1424: We can refine this result, finding other parameter ranges for which 
1425: $2X^2(\omega-a)+X(4\omega-(4-d)a)+2\omega<0$. If $d\geq5$ and $a=1$, this holds
1426: for $\omega\leq -(d-4)/4$. Moreover,
1427: \begin{equation*}
1428: \begin{aligned}
1429: 2X^2(\omega-a)&+X(4\omega-(4-d)a)+2\omega \\
1430: & = 2\left[(\omega-a)(X-1)^2+X(4\omega-4a+\frac d2 a) + \omega+a\right].
1431: \end{aligned}
1432: \end{equation*}
1433: No solitary wave can exist for $a=-1$ and $\omega\leq-1$. Hence we complete 
1434: Corollary \ref{Cor_NLSnoSW} with
1435: 
1436: \def\thecorollary{A.\arabic{corollary}}
1437: \setlength\leftmargini{2pc}
1438: \begin{corollary}
1439: No non-trivial solitary wave (solution of \eqref{Eq_NLSSW}) of the saturated NLS
1440: equation exists when
1441: \begin{itemize}
1442: \item[{\em (i)}] $a=-1$ {\em (}defocusing case{\em )}, for $\omega\leq-1$.
1443: \item[{\em (ii)}] $a=1$ {\em (}focusing case{\em )}, for $\omega\leq0$, if $d=3,4$ and
1444:   $\omega\leq-(d-4)/4$ if $d\geq5$.
1445: \end{itemize}
1446: \end{corollary}
1447: 
1448: % \section*{References}
1449: 
1450: {\small
1451: \begin{thebibliography}{10}
1452: 
1453: \bibitem{abramowitz-steglun}
1454: M. Abramowitz and I. A. Steglun,
1455: \newblock {\em Handbook of Mathematical Functions}
1456: (National Bureau of Standards, 1964).
1457: 
1458: \bibitem{berestycki-lions-peletier}
1459: H.~Berestycki, P.-L. Lions and L. A. Peletier,
1460: \newblock An {ODE} approach to the existence of positive solutions for
1461:   semilinear problems in {$\bbR^N$},
1462: \newblock {\em Indiana Univ. Math. J.} \textbf{30} (1981) 141--157.
1463: 
1464: \bibitem{bona-smith}
1465: J. L. Bona and R. Smith,
1466: \newblock The initial value problem for the {K}orteweg--de {V}ries equation,
1467: \newblock {\em Philos. Trans. Royal. Soc. London A} \textbf{278} (1975) 
1468: 555--601.
1469: 
1470: \bibitem{cazenave1}
1471: T. Cazenave,
1472: \newblock {\em An Introduction to Nonlinear {S}chr\"odinger Equations,} 
1473: 3rd edition, Textos de Metodos Matematicos. Vol. 26 (Instituto de Matematica, 
1474: Universidade Federal do Rio de Janeiro, 1996).
1475: 
1476: \bibitem{cazenave2}
1477: T. Cazenave,
1478: \newblock {\em Semilinear {S}chr\"odinger Equations}, 
1479: Courant Lecture Notes in Mathematics. Vol. 10 (Courant Institute of 
1480: Mathematical Sciences, New York, 2003).
1481: 
1482: \bibitem{delre-crosignani-diporto}
1483: E. DelRe, B. Crosignani and P. {Di Porto},
1484: \newblock Photorefractive spatial solitons,
1485: \newblock in {\em Spatial Solitons,} S. Trillo and W. Torruellas (eds), 
1486: Springer Series in Optical Sciences (Springer, 2001), pp. 61--85.
1487: 
1488: \bibitem{ghidaglia-saut1}
1489: J.-M. Ghidaglia and J.-C. Saut,
1490: \newblock On the initial value problem for the {D}avey--{S}tewartson systems,
1491: \newblock {\em Nonlinearity} \textbf{3} (1990) 475--506.
1492: 
1493: \bibitem{kato}
1494: T. Kato,
1495: \newblock Growth properties of solutions of the reduced wave equation with a
1496:   variable coefficient,
1497: \newblock {\em Commun. Pure Appl. Math.} \textbf{12} (1959) 403--425.
1498: 
1499: \bibitem{kukhtarev-markow-odoluv-soskin-vinetskii}
1500: N. V. Kukhtarev, V. B. Markow, S. G. Odoluv, M. S. Soskin  and V. L. Vinetskii,
1501: \newblock Holographic storage in electrooptic crystals,
1502: \newblock {\em Ferroelectrics} \textbf{22} (1979) 949--960.
1503: 
1504: \bibitem{lions}
1505: J.-L. Lions,
1506: \newblock {\em Quelques m\'ethodes de r\'esolution des probl\`emes aux limites
1507:   non lin\'eaires} (Dunod, Paris, 1969).
1508: 
1509: \bibitem{mamaev-saffman-zozulya2}
1510: A. V. Mamaev, M. Saffman and A. A. Zozulya,
1511: \newblock Break-up of two-dimensional bright spatial solitons due to transverse
1512:   modulational instability,
1513: \newblock {\em Europhys. Lett.} \textbf{35} (1996) 25--30.
1514: 
1515: \bibitem{mamaev-saffman-zozulya1}
1516: A. V. Mamaev, M. Saffman  and A. A. Zozulya,
1517: \newblock Propagation of dark stripe beams in nonlinear media: {S}nake
1518:   instability and creation of optical vertices, 
1519: \newblock {\em Phys. Rev. Lett.} \textbf{76} (1996) 2262--2265.
1520: 
1521: \bibitem{stepken-kaiser-belic-krolikowski}
1522: A. Stepken, F. Kaiser, M. R. Beli\'c  and W. Kr\'olikowski,
1523: \newblock Interaction of incoherent two-dimensional photorefractive solitons,
1524: \newblock {\em Phys. Rev. E} \textbf{58} (1998) R4112--R4115.
1525: 
1526: \bibitem{tikhonenko-christou-luther-davies}
1527: V. Tikhonenko, J. Christou and B. Luther-Davies,
1528: \newblock Three-dimensional bright spatial soliton collision and fusion in a
1529:   saturable nonlinear medium,
1530: \newblock {\em Phys. Rev. Lett.} \textbf{76} (1996) 2698--2701.
1531: 
1532: \bibitem{wolfersberger-fressengeas-maufroy-kugel}
1533: D. Wolfersberger, N. Fressengeas, J. Maufroy and G. Kugel,
1534: \newblock Self-focusing of a single laser pulse in a photorefractive medium,
1535: \newblock {\em Phys. Rev. E} \textbf{62} (2000) 8700--8704.
1536: 
1537: \bibitem{zozulya-anderson}
1538: A. A. Zozulya and D. Z. Anderson,
1539: \newblock Propagation of an optical beam in a photorefractive medium in the
1540:   presence of a photogalvanic nonlinearity or an externally applied electric
1541:   field,
1542: \newblock {\em Phys. Rev. A} \textbf{51} (1995) 1520--1531.
1543: 
1544: \end{thebibliography}
1545: }
1546: 
1547: \end{document}