nlin0208001/lan.tex
1: \documentclass{elsart}
2: %\documentclass[doublespacing]{elsart}
3: \usepackage{amssymb}
4: \usepackage{amsmath}
5: \usepackage{graphicx}
6: \newtheorem{theorem}{Theorem}[section]
7: \newtheorem{defit}{Definition}[section]
8: 
9: \begin{document}
10: 
11: \begin{frontmatter}
12: 
13: \title{Stationary modulated-amplitude waves in the 
14: 	1-$D$ complex Ginzburg-Landau equation}
15: \author{Yueheng Lan, Nicolas Garnier and Predrag Cvitanovi\'c}
16: \address{Center for Nonlinear Science, School of Physics,
17: 	Georgia Institute of Technology \\
18: 	837 State Street, Atlanta, GA 30332-0430, USA}
19: \date{July 29th 2002}
20: \maketitle
21: 
22: \begin{abstract}
23: 
24: We reformulate the one-dimensional complex Ginzburg-Landau equation as a
25: fourth order ordinary differential equation in order to find stationary
26: spatially-periodic solutions. Using this formalism, we prove the
27: existence and stability of stationary modulated-amplitude wave
28: solutions. Approximate analytic expressions and a comparison with
29: numerics are given.
30: 
31: \end{abstract}
32: 
33: \begin{keyword}
34: complex Ginzburg-Landau equation \sep coherent structures
35: \PACS 05.45.-a \sep 47.54.+r \sep 05.45.Jn
36: \end{keyword}
37: 
38: \end{frontmatter}
39: 
40: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
41: \section*{Introduction}
42: 
43: The cubic complex Ginzburg-Landau equation (CGLe) is a generic amplitude
44: equation describing Hopf bifurcation in spatially extended systems,
45: {i.e.}, $I_o$ systems~\cite{patt}, with reflection
46: symmetry~\cite{sup,sub,amp}. It is of great interest due to its
47: genericity and applications to onset of wave pattern-forming
48: instabilities~\cite{patt} in various physical systems such as fluid
49: dynamics, optics, chemistry and biology. It exhibits rich dynamics and
50: has become a paradigm for the transition to spatio-temporal chaos.
51: 
52: We consider the one-dimensional CGLe for the complex amplitude field
53: $A(x,t)$:
54: % 
55: \begin{eqnarray} 
56: A_t=\mu A+(1+i\alpha)A_{xx}-(1+i\beta)|A|^2 A 
57: \label{cgl} 
58: \end{eqnarray} 
59: % 
60: where $A(x,t):\mathbb{R}^2 \mapsto \mathbb{C}$, and $\mu,\alpha,\beta
61: \in \mathbb{R}$, $x \in {\mathcal D}$. $\mathcal D$ is the spatial
62: domain on which the equation is defined. Interesting domains for us are
63: either the whole real axis or a finite box of length $L$ with periodic
64: boundary conditions. $\mu$ is the control parameter. Only $\mu>0$ is
65: considered because we study the supercritical Ginzburg-Landau equation;
66: one could set $\mu=1$ by appropriate rescaling of the time, space and
67: amplitude, but we keep it as a parameter for closer connection with
68: experimental results and previous literature. Coefficients $\alpha$ and
69: $\beta$ parametrize the linear and nonlinear dispersion. 
70: 
71: If both $\alpha$ and $\beta$ are set to $0$, we recover the real
72: Ginzburg-Landau equation (RGLe) in which only the diffusion term and the
73: stabilizing cubic term compete with each other and the linear term. A
74: Lyapunov functional exists in that case~\cite{patt} and the RGLe behaves
75: like a gradient system. Another limit ---~the nonlinear Schr\"{o}dinger
76: equation~--- results from setting $\alpha,\beta \rightarrow \infty$; we
77: then have an integrable nonlinear PDE. For parameter values in the
78: intermediate range, long-time behavior of the CGLe can vary from
79: stationary to periodic and to spatiotemporal chaos~\cite{Shraiman:92}.
80: In this paper, we concentrate on the stationary solutions of the CGLe in
81: a finite box of length $L$ with periodic boundary conditions, and the
82: case $\alpha \neq \beta$. Stationary solutions are the simplest
83: non-trivial solutions, related to propagating solutions by an
84: appropriate change of frame of reference $(x,t) \mapsto (x-vt,t)$ with
85: fixed $v \in \mathbb{R}$. 
86: 
87: Searching for coherent structures allows one to reduce a partial
88: differential equation into an ordinary one, and such solutions of the
89: CGLe are believed to be extremely important in many regimes, including
90: the spatiotemporal chaos~\cite{hohen}. Recently, numerical integrations
91: of the CGLe have focused on a class of solutions called
92: modulated-amplitude waves (MAWs) and their role in the nonlinear
93: evolution of the Eckhaus instability of initially homogeneous plane
94: waves~\cite{maw,defect}. 
95: 
96: MAWs can bifurcate from the trivial solution $A=0$ (case I) or plane
97: wave solutions of zero wavenumber (case II). Analytical aspects of
98: modulated solutions of the CGLe have been addressed by Newton and
99: Sirovich who have applied a perturbation analysis to study the
100: bifurcation in case II~\cite{quarter1}, and discussed the secondary
101: bifurcation of those MAWs~\cite{quarter2}. Tak\'a\v{c}~\cite{pt} proved
102: the existence of MAW solutions using a standard bifurcation analysis in
103: the infinite-dimensional phase space of the CGLe, in both cases I and
104: II, together with a stability analysis in case I by means of the center
105: manifold theorem.
106: 
107: In this article we reformulate the CGLe equation assuming a coherent
108: structure form for the solutions, and obtain a fourth-order ordinary
109: differential equation (ODE) with a consistency condition. This form is
110: algebraically convenient, because the deduced system of four first-order
111: ODEs contains only quadratic non-linearity. In the Benjamin-Feir-Newell
112: regime, where plane waves solutions are always unstable, we give a proof
113: of existence of MAWs in both case I and II using our ODE. For weak
114: perturbations in case I or II, we write approximate analytic solutions
115: in the ODE phase space. Coming back to the full CGLe, we then prove the
116: stability of those MAWs in a finite box in case II, and prove that the
117: bifurcation is supercritical, as suggested by recent numerical
118: work~\cite{maw}.
119: 
120: In the next section, we discuss symmetries and solutions of the CGLe. In
121: section~\ref{sec:ODE} we transform the steady CGLe for MAWs into an
122: equivalent ODE, and give the sufficient condition to identify the
123: solutions of these two equations. In section~\ref{sec:existence} this
124: ODE is used to construct a 4-$D$ dynamical system and prove the existence
125: of symmetric stationary solutions of the CGLe in the two cases I and II.
126: In section~\ref{sec:approximate} the approximate analytic form of the
127: solutions is given and compared to numerical calculations, and the
128: stability of MAWs in case II is proved. Several theorems needed in the
129: proofs are reproduced in appendix~\ref{sec:theorems}.
130: 
131: 
132: 
133: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
134: \section{Basic properties of the CGLe}
135: 
136: \subsection{Symmetries}
137: 
138: The equation~(\ref{cgl}) is invariant under temporal and spatial
139: translations. Moreover, it is invariant under a global gauge
140: transformation $A\rightarrow A \exp({i\phi})$, where $\phi \in \mathbb{R}$,
141: and under $x \rightarrow -x$ reflection. As a consequence, it preserves
142: parity of A, i.e., if $A(-x,0)=\pm A(x,0) $, then $A(-x,t)=\pm A(x,t)$
143: for any later time $t>0$. If $A(x,t)$ has no parity, then $A(-x,t)$
144: gives another solution.
145: 
146: %{\tt Lan, where to put the following further in the text ?}
147: %
148: %The one-parameter family of solutions initially derived by Nozaki and
149: %Bekki~\cite{bk} strongly suggests the existence of a hidden symmetry of
150: %the CGLe \cite{hohen}. Including higher order terms like quintic term
151: %$\delta |A|^4A$ will send this family away~\cite{symm}. So, the family
152: %is structurally unstable and this hidden symmetry is possessed only by the
153: %cubic CGLe. In the stationary case, the existence of the Nozaki-Bekki
154: %solution and more complicated solutions could be inferred by applying
155: %the counting argument associated with the reflection symmetry~\cite{tb}
156: %and they are structurally stable. So their existence does not relate to
157: %the hidden symmetry.
158: 
159: 
160: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
161: \subsection{Stokes solutions and their stability}
162: 
163: The global phase invariance implies that the CGLe has nonlinear plane
164: wave solutions of form
165: %
166: \begin{equation}
167: A(x,t)=R_0 \exp ({i(qx-\omega t)}) \,,
168: \label{eq:Stokes}
169: \end{equation}
170: %
171: where $R_0^2=\mu-q^2$ is the amplitude squared, $\omega=\mu \beta +
172: (\alpha-\beta)q^2$ is the frequency, and $q \in \mathbb{R}, q^2 \leq
173: \mu$ is the wavenumber. They are called Stokes solutions~\cite{WoCGLe}
174: and are parametrized by the wavenumber $q$. The two limit cases of
175: interest to us are highlighted on figure~\ref{fig:diagram}: a plane wave
176: of wavenumber $\mu^{1/2}$ and of vanishing amplitude (case I), and the
177: wave with zero wavenumber and maximum amplitude $\mu$ (case II). In case
178: II, the solution oscillates uniformly in time; we call it the
179: homogeneously oscillating state (HOS).
180: 
181: For the infinite system, the Benjamin-Feir-Newell~\cite{Newell:74}
182: criterion states that all plane wave solutions are unstable with respect
183: to long wavelength perturbations (i.e., of wavenumber $k \rightarrow 0$)
184: if $1+\alpha \beta<0$. If $1+\alpha \beta>0$, we have to consider the
185: Eckhaus instability criterion; only a band of wavenumbers are stable
186: against long wavelength perturbations (figure~\ref{fig:diagram}):
187: 
188: \begin{equation}
189: q^2 < q_{\rm E}^2 \equiv 
190: \frac{(1+\alpha\beta) \mu}{ 3 + \alpha \beta + 2\beta^2} \,.
191: \label{eq:Eckhaus}
192: \end{equation}
193: 
194: For a finite periodic system the wavenumbers for both the original
195: states and the perturbations are quantized. These criteria have been
196: reexamined by Matkowsky and Volpert using linear stability
197: analysis~\cite{stp}.
198: 
199: \begin{figure}\begin{center}
200: \includegraphics[width=8cm]{diagram}
201: \setlength\unitlength{1cm}
202: \begin{picture}(0,5)
203: \put(-4.5,5.5){$\mu$}
204: \put(-0.2,0.2){$q$}
205: \put(-0.8,5  ){$MS$}
206: \put(-2.5,5  ){$E$}
207: \put(-4.6,0.1){0}
208: \put(-4.2,2.7){II}
209: \put(-1.8,2.7){I}
210: \end{picture}
211: \end{center}
212: \caption{Marginal stability curve (MS) and Eckhaus instability 
213: 	curve (E) defining regions where plane waves solutions 
214: 	exist in the CGLe (inside (MS)), and are stable when $1+\alpha\beta>0$
215: 	(inside (E)). Case I corresponds to the wave of maximal
216: 	possible wavenumber and Case II to the wave with $q=0$ and 
217: 	the maximal amplitude.}
218: \label{fig:diagram}
219: \end{figure}
220: 
221: 
222: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
223: \subsection{Coherent structures and MAWs}
224: \label{sec:intro:MAWs}
225: 
226: Coherent structures play a very important role in the study of pattern
227: formation and dynamical properties of the CGLe~\cite{hohen}. They are
228: uniformly propagating structures of the form
229: \[
230: A(x,t)=R(x-vt)\exp({i\phi (x-vt)}e^{-i\omega t})
231: \]
232: which can be expressed as solutions of a 3-$D$ nonlinear dynamical system
233: obtained by substituting the above ansatz into the CGLe. There are two
234: free parameters: the frequency $\omega$ and the group velocity $v$.
235: 
236: The fixed points of the 3-$D$ system are the plane waves described in
237: the previous section. The homoclinic~\cite{build} and
238: heteroclinic~\cite{hohen} connections between the fixed points
239: correspond to localized coherent structures. The Nozaki-Bekki
240: solutions~\cite{bk} belong to this category; they connect asymptotic
241: plane waves with different wavenumbers. In numerical simulations in
242: large domains, nearly coherent structures are frequently observed in
243: chaotic regimes, thus suggesting those objects are also relevant to
244: spatiotemporally chaotic dynamics.
245: 
246: Recent numerical studies reveal another kind of coherent structure:
247: modulated amplitude waves (MAWs) for the CGLe~\cite{maw}. They
248: correspond to limit cycles of the 3-$D$ nonlinear system. When $v=0$,
249: MAWs are stationary. The formation of MAWs is the first instability
250: encountered when a plane wave state crosses the Eckhaus or Benjamin-Feir
251: stability line. The MAW structure is frequently observed in
252: experiments~\cite{sup,garnier} and considered as a key to interpretation
253: of patterns and bifurcations exhibited during the system's transition to
254: spatio-temporal chaos~\cite{defect}. Traveling MAWs have been observed
255: in numerical simulations of the CGLe in periodic boxes, with parameter
256: $q$ between 0 and $\mu^{1/2}$, i.e., in between cases I and II; we are
257: interested here only in stationary MAWs that appear either in case I or
258: case II.
259: 
260: In this paper, we propose a new real-valued ODE to describe steady
261: solutions of the CGLe. A 4-$D$ dynamical system derived from this ODE
262: enables us to apply the successive approximation method~\cite{jhale},
263: to prove the existence of stationary MAWs and to give the analytical 
264: form of the approximate solutions in both case I and case II. Numerical
265: integrations of the exact CGLe are then compared to the approximate
266: analytic result. Furthermore, we show non-analyticity at discrete points
267: of solutions in case I, and prove the stability of the MAWs in case II.
268: Some theorems needed in our proof are reproduced in the
269: appendix~\ref{sec:theorems}. In what follows, $\mathrm{diag}(\cdots)$
270: denotes a (block) diagonal matrix and $\mathrm{col}(\cdots)$ a column
271: vector. 
272: 
273: 
274: 
275: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
276: \section{Stationary case}
277: \label{sec:ODE}
278: 
279: Since we are only interested in the steady solutions of the CGLe, we
280: substitute the ansatz
281: %
282: \begin{equation}
283:  A(x,t)=R(x) \exp({i\phi(x)-i\omega t}), 
284: 	\qquad (R,\phi) \in \mathbb{R}^2
285: \label{eq:antsatz}
286: \end{equation}
287: %
288: into (\ref{cgl}). We then have    
289: %
290: \begin{eqnarray}
291: (1+\alpha^2)G_x & = K & \equiv (\beta-\alpha) R^4
292: 	-(\omega-\mu \alpha)R^2  \label{or1}\\ 
293: (1+\alpha^2)G^2 & = M & \equiv (1+\alpha^2)R^3 R_{xx}
294: 	+(\alpha \omega +\mu)R^4 -(1+\alpha \beta) R^6  \,.  \label{or2} 
295: \end{eqnarray}
296: %
297: where $G \equiv \phi_x R^2$ is reminiscent of ``angular momentum''. Note
298: that if $\alpha=\beta$, this ``angular momentum'' is conserved ---~it is
299: constant in space~--- provided that $\omega=\mu \alpha$. In that case,
300: (\ref{or2}) can be solved in terms of elliptic functions~\cite{ellip}.
301: We will only consider the case $\alpha \neq \beta$ in the following.
302: Equations (\ref{or1}) and (\ref{or2}) are invariant under $(G,x)
303: \rightarrow (-G,-x)$. Note that for plane waves, $K=0$ and $G$ is a
304: constant. If $K$ is not always zero, differentiating (\ref{or2}) and
305: dividing the result by (\ref{or1}) gives
306: %
307: \begin{equation}
308: 	2G = M_x/K \,,  \label{dg}
309: \end{equation} 
310: %
311: and by (\ref{or2})
312: \begin{eqnarray}
313: 	M = \frac{1+\alpha^2}{4}\frac{M_x^2}{K^2} \,.\label{sqr}
314: \end{eqnarray} 
315: 
316: Furthermore, we can factorize $R^2$ from $M_x$ and $K$ and write
317: $M_x = R^2 N $ and $K=R^2 P$, where
318: \begin{eqnarray}
319: 	N & \equiv & (1+\alpha^2)\frac{1}{2}(R^2)_{xxx}+(\alpha
320: 		\omega+\mu)2(R^2)_x -(1+\alpha \beta)3R^2(R^2)_x 	
321: 							\nonumber\\ 
322: 	P & \equiv & (\beta-\alpha)R^2-(\omega-\mu \alpha) \,.		
323: 							\label{eq:introP}
324: \end{eqnarray}
325: %
326: The last relation can be used to express $R^2$ in terms of $P$:
327: \begin{eqnarray}
328: 	R^2=\frac{(\omega-\mu \alpha)+P}{\beta-\alpha}=e+dP
329: 	= R_0^2 + \frac{P}{\beta-\alpha} \,,		\label{dR}
330: \end{eqnarray}
331: %
332: where $d\equiv 1/(\beta-\alpha)$ and 
333: $e \equiv {(\omega-\mu \alpha)}/{(\beta-\alpha)}$.
334: 
335: Note that $e=R_0^2$ is the square of the homogeneous amplitude
336: $R_0(q,\omega)$ of the Stokes plane wave solution (\ref{eq:Stokes}) of
337: frequency $\omega$ and wavevector $q(\omega)$. $P$ then appears as the
338: modulation of the amplitude squared with respect to the Stokes solution,
339: and so it is an appropriate variable to describe a MAW.
340: 
341: Substituting $K$ and $M_x$ into (\ref{sqr}), we have
342: %
343: \begin{equation}
344: 	\frac{1+\alpha^2}{4} \frac{N^2}{P^2}=M \,. \label{or3}
345: \end{equation} 
346: %
347: If $P \neq 0$ (\ref{or3}) is equivalent to (\ref{or1}) and
348: (\ref{or2}). It is easy to check that if we regard
349: (\ref{dg}) as a definition of $G$, and use $K,M,N,P$ expressed in terms
350: of $R$, equation (\ref{or1}) and (\ref{or2}) will be recovered as a
351: result of (\ref{sqr}) and (\ref{or3}). Differentiating both sides of
352: (\ref{or3}) results in
353: %
354: \begin{equation}
355: 	\frac{1+\alpha^2}{2}(PN_x-NP_x)  =  R^2 P^3 \,.		\label{rd}
356: \end{equation}
357: 
358: In this step we have extended the solution set of (\ref{or3}), because as 
359: we integrate (\ref{rd}) back, we get
360: %
361: \begin{equation}
362: 	\frac{1+\alpha^2}{4} \frac{N^2}{P^2}= M + {\mathcal C} \,, 	\label{mrd}
363: \end{equation} 
364: %
365: where ${\mathcal C}$ is an integration constant. Only when ${\mathcal
366: C}=0$, a solution of (\ref{rd}) is a solution of (\ref{or3}). For this
367: reason, when obtaining solutions of (\ref{rd}), we have to check the
368: {\em consistency condition}
369: %
370: \begin{equation}
371: 	\frac{1+\alpha^2}{4} \frac{N^2}{P^2}-M=0 		\label{cc1}
372: \end{equation}  
373: %
374: to make sure that we have a solution of (\ref{or3}), thus a solution of
375: (\ref{or1}) and (\ref{or2}). Note that if $K$ vanishes we have to go
376: back to (\ref{or1}) and (\ref{or2}), since in that case (\ref{or3}) is
377: not well defined. Let us rewrite $N$ in terms of $P$:
378: %
379: \begin{eqnarray}
380: N= \displaystyle \frac{2}{1+\alpha^2} \left(aP_{xxx}+bP_x+cPP_x \right)	\,, \label{dC}
381: \end{eqnarray}
382: %
383: where $a,b,c$ are constants
384: %
385: \begin{eqnarray}
386: a & \equiv & \frac{(1+\alpha^2)^2}{4(\beta-\alpha)} \nonumber \\
387: b & \equiv & \frac{1+\alpha^2}{2}\left(\frac{2(\alpha \omega+\mu)}{\beta-\alpha}-
388: \frac{3(1+\alpha \beta)(\omega-\mu \alpha) }{(\beta-\alpha)^2}\right) \label{eq:abcdefs} \\
389: c & \equiv & -\frac{3(1+\alpha \beta)(1+\alpha^2)}{2(\beta-\alpha)^2} \,. \nonumber
390: \end{eqnarray}
391: %
392: After some algebra (here relegated to appendix~\ref{sec:algebra}),
393: we get an equation for $P$ only:
394: %
395: \begin{equation}
396: \left(\frac{\tilde{M}_x}{P}\right)_x = \frac{\lambda}{a}\tilde{M} + kP \,,
397: \qquad \tilde{M} \equiv \lambda P_{xx}+dP^2+\tilde{e}P \,.
398: \label{ode1}
399: \end{equation} 
400: 
401: $\lambda$ is a fixed real constant that depends on $\alpha$ and $\beta$
402: only, and that takes two different values given in
403: appendix~\ref{sec:algebra}. $\lambda$ is a transient variable used in
404: the proof and derivation but our solutions to the CGLe do not depend on
405: $\lambda$ and do not distinguish the two values of $\lambda$ (see
406: section~\ref{sec:approximate}). $\tilde{e}$ and $k$ are real parameters
407: introduced as $\tilde{e}+\frac{a}{\lambda}k=e$. So (\ref{ode1}) has two
408: free parameters: $\omega$, introduced by the ansatz~(\ref{eq:antsatz})
409: as the carrier frequency of the solution, and $k$. the {\em consistency
410: condition} (\ref{cc1}) fixes one parameter.
411: 
412: 
413: 
414: 
415: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
416: \section{4-$D$ dynamical system and the existence of periodic solutions}
417: \label{sec:existence}
418: 
419: Let us take $\tau$ as the spatial variable, $P=P(\tau)$ in
420: (\ref{eq:introP}), and rewrite (\ref{ode1}) as a system of first order
421: equations in $\tau$. With $\tilde{N} = {\tilde{M}_\tau}/{P}$ and
422: $Q=P_\tau$, from (\ref{ode1}) we have
423: %
424: \begin{equation}
425: \left\{\begin{array}{rcl}
426:        \dot{\tilde{M}} &=& \tilde{N}P \\
427:        \dot{\tilde{N}} &=& \frac{\lambda}{a}\tilde{M}+kP \\
428:        \dot{P}         &=& Q \\
429:        \dot{Q}         &=& \frac{1}{\lambda}(\tilde{M}-dP^2-\tilde{e}P)  
430:        \end{array}\right. \,,
431: \label{eq:4d}
432: \end{equation}
433: %
434: where the dot represents the derivation with respect to the spatial
435: variable $\tau$.
436: 
437: It is easy to check that $P=0$ is a solution of the original equations
438: (\ref{or1}) and (\ref{or2}), corresponding to the plane wave solution of
439: the CGLe with frequency $\omega$. We will study the behavior near $P=0$
440: and prove the existence of periodic solutions for small $P$. In the CGLe,
441: this corresponds to a weakly modulated amplitude wave which bifurcates
442: from a plane wave solution. If $P \sim \epsilon$, where $\epsilon$ is a
443: small parameter, so are $\tilde{M},\tilde{N},Q$ by their
444: definitions. Write
445: %
446: \begin{equation*}
447: (\tilde{M}, \tilde{N}, P, Q)
448: = (\epsilon x, \epsilon y, \epsilon z, \epsilon w)
449: \end{equation*}
450: %
451: and set $k=k_1+\epsilon k_2$. Substituting these into the 4-$D$ system, we have
452: %
453: \[
454: \frac{d}{d\tau}\left(\begin{array}{c}
455:                     x \\ y \\ z \\ w
456:                   \end{array} \right)=A\left(\begin{array}{c}
457:         x \\ y \\ z \\ w
458:       \end{array} \right)+\epsilon \left(\begin{array}{c}
459:                                          y\,z \\ k_2\, z \\0\\ - \frac{d}{\lambda}z^2 
460:                                          \end{array} \right),\mbox{ where }
461: A=\left(\begin{array}{cccc}
462:                                              0 & 0 & 0                  & 0 \\
463:                              \frac{\lambda}{a} & 0 & k_1                & 0 \\
464:                                              0 & 0 & 0                  & 1 \\
465:                              \frac{1}{\lambda} & 0 & \frac{-\tilde{e}}{\lambda} & 0
466:                                             \end{array}\right). 
467: \]
468: 
469: The linear part $A$ describes the behavior of the system in the
470: neighborhood of the trivial fixed point $(0,0,0,0)$. Note that the
471: system is invariant under $(t,y,w) \rightarrow -(t,y,w)$. We use this
472: property to simplify our analysis. Moreover, this system defines an
473: incompressible flow since $\nabla \cdot \vec{X}=0$, where
474: $\vec{X}=(x,y,z,w)$. It follows from (\ref{mrd}) that the system
475: has one integration constant ${\mathcal C}$. This constant induces a
476: foliation of the phase space into three-dimensional manifolds. Physical
477: solutions, i.e., the solutions of the original CGLe, are restricted to
478: ${\mathcal C}=0$, the manifold that satisfies the {\em consistency
479: condition} (\ref{cc1}). 
480: 
481: These properties strongly restrict the possible distribution of
482: eigenvalues of $A$. We restrict our analysis to the case
483: ${\tilde{e}}/{\lambda}>0$, then $A$ has eigenvalues
484: $\left\{ 0, 0, i \omega_1, -i \omega_1 \right\}$ with $\omega_1 =
485: \sqrt{{\tilde{e}}/{\lambda}}$. In that case, periodic solutions or
486: MAWs can exist as we will prove in the following. 
487: The evolution of the system along either
488: of the two degenerate eigenvalue 0 directions respects the constant
489: ${\mathcal C}$ foliation: if the solution is on a constant ${\mathcal
490: C}$ manifold at initial time, it remains there for any later time.
491: 
492: We now discuss the condition $\tilde{e}/\lambda>0$ in terms of an
493: instability of the underlying plane wave. We can rewrite
494: $\tilde{e}/\lambda$ using (\ref{eq:abcdefs}) and (\ref{condi}). Assuming
495: that the solution we are searching for is close to a plane wave, we can
496: use the wavenumber $q$ instead of the frequency $\omega$, using the
497: dispersion relation (\ref{eq:Stokes}) for plane waves:
498: %
499: \begin{eqnarray}
500: \frac{\tilde{e}}{\lambda} = \frac{b}{a} 
501: 	& = & \frac{2}{1+\alpha^2}\left[2(\alpha \omega +\mu)
502: 	 	 -\frac{3(1+\alpha \beta)(\omega-\mu \alpha)}
503: 		{\beta-\alpha}\right]			\label{domega1} \\
504: 	& = & \frac{2}{1+\alpha^2}\left[ (3+\alpha\beta+2\alpha^2)q^2
505: 	  	-(1+\alpha\beta)\mu \right] \,. \nonumber 
506: \end{eqnarray}
507: %
508: If we write
509: \begin{equation}
510: q_{\rm M}^2 \equiv \frac{(1+\alpha\beta) \mu}{ 3 + \alpha \beta + 2\alpha^2} \,,
511: \label{eq:defqm}
512: \end{equation}
513: %
514: we have
515: %
516: \begin{eqnarray}
517: \frac{\tilde{e}}{\lambda} > 0 & \Leftrightarrow & 
518: \left| \begin{array}{rl}
519: q^2 > q_{\rm M}^2 
520: 	& \quad {\rm if} \quad (1+\alpha\beta)>0 \\
521: q^2 < q_{\rm M}^2
522:  	& \quad {\rm if} \quad (1+\alpha\beta)<-2(1+\alpha^2) \\
523: \forall q \in [-\sqrt{\mu}, \sqrt{\mu}]
524: 	& \quad {\rm if} \quad -2(1+\alpha^2) < 1+\alpha\beta < 0
525: \end{array} \right. .
526: \label{eq:positivity}
527: \end{eqnarray}
528: 
529: \begin{figure}
530: \begin{center}
531: \includegraphics[width=8cm]{phases2} \qquad
532: \includegraphics[width=3.5cm]{Eck} 
533: \end{center}
534: \caption{Left: wavenumber distribution
535: 	of stationary MAWs in the $(\alpha,\beta)$ plane. 
536: 	In (BFS), MAWs exist if $q^2>q_{\rm M}^2$.
537: 	In (M1), MAWs exist $\forall q$.
538: 	In (M2), MAWs exist if $q^2<q_{\rm M}^2$.
539: 	Right: regions of existence of MAWs in the $(q,\mu)$ 
540: 	plane in the Benjamin-Feir-Newell stable regime
541: 	((BFS) region). (MS) is the marginal stability curve, 
542: 	(E) is the Eckhaus instability curve and (M) is existence 
543: 	curve defined by~(\ref{eq:defqm}). 
544: 	Stationary MAWs exist outside (M).}
545: \label{fig:phases}
546: \end{figure}
547: 
548: The corresponding regions are illustrated on Fig.~\ref{fig:phases}.
549: Note that $q_{\rm M}(\alpha,\beta,\mu)=q_{\rm E}(\beta,\alpha,\mu)$. If
550: $|\alpha|=|\beta|$, the positivity of $\tilde{e}/\lambda$ is assured
551: when the corresponding plane wave is Eckhaus unstable. If
552: $|\alpha|\neq|\beta|$, the positivity does not coincide anymore with the
553: Eckhaus criterion; this is not surprising considering that we do not
554: restrict our analysis to long wavelength perturbations of plane waves,
555: but that the solutions we are seeking may have any wavenumber.
556: 
557: In the following we distinguish two cases. In the first case eigenvalue
558: $0$ has a simple elementary divisor, {\em i.e.}, has two distinct
559: eigenvectors~cite{jhale}. This coincides with case I: the MAW solution
560: bifurcates from the $A=0$ state, with $\omega\sim \mu \alpha$ and hence
561: ${\tilde{e}}/{\lambda}\sim 4 \mu >0$, for $\mu>0$. In the second case,
562: eigenvalue $0$ has only one eigenvector. This coincides with case II:
563: the MAW is superimposed over a plane wave with $\omega \simeq \mu \beta$,
564: so $q \simeq 0$, and
565: %
566: \[
567: \frac{\tilde{e}}{\lambda} \simeq -\frac{2\mu(1+\alpha \beta)}{1+\alpha^2}>\,0,
568: \]
569: %
570: The positivity is insured if the system is Benjamin-Feir-Newell 
571: unstable, $(1+\alpha\beta)<0$.
572: 
573: In terms of $\tilde{M}, \tilde{N}, P, Q$, the {\em consistency
574: condition} (\ref{cc1}) can be written as
575: %
576: \begin{equation}
577: (1+\alpha^2) M = \left( \frac{a}{\lambda}\tilde{N} - \lambda Q \right)^2 
578: 							\label{cc2}
579: \end{equation}
580: %
581: where in new variables
582: %
583: \begin{eqnarray*}
584: M & = & \frac{d(1+\alpha^2)}{2 \lambda}(d\,P+e)(\tilde{M}-d\,P^2-\tilde{e}\,P)-
585: \frac{d^2(1+\alpha^2)}{4}\, Q^2 \\
586:   &   & (\alpha \omega+\mu)(d \, P+e)^2-(1+\alpha \, \beta)(d \, P+e)^3 \,.
587: \end{eqnarray*}
588: %
589: Recalling (\ref{or2}), we may  express $G$ by
590: %
591: \begin{equation}
592: G  =  \frac{\frac{a}{\lambda}\tilde{N}-\lambda Q}{1+\alpha^2}  \,.
593: 							\label{ag}
594: \end{equation}
595: %
596: Here we are allowed to fix the sign of the right hand side expression
597: because of the $(G,x)\mapsto(-G,-x)$ reflection symmetry of (\ref{or1})
598: and (\ref{or2}). 
599: 
600: 
601: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
602: \subsection{Case I}
603: 
604: We want the eigenvalue $0$ to have non-degenerate eigenvectors, for
605: this, we set
606: %
607: \[
608:  \frac{\frac{\lambda}{a}}{\frac{1}{\lambda}}
609: =\frac{k_1}{-\frac{\tilde{e}}{\lambda}},\mbox{~i.e.,~}
610:   k_1=-\frac{\lambda \tilde{e}}{a}
611: \]
612: %
613: Consequently, we have 
614: %
615: \begin{equation}
616: e=\tilde{e}+\frac{a}{\lambda}k=\frac{\epsilon a}{\lambda}k_2 \,.
617: 							\label{2nde}
618: \end{equation}
619: 
620: Notice that $e\sim 0$ to the zeroth order, so $R_0 \sim 0$ and $\omega
621: \sim \mu \alpha$, which means that the solution to be considered
622: bifurcates from the zero solution $A=0$, corresponding to a plane wave
623: around the marginal stability curve, with wavenumber $q \sim \pm \mu^{1/2}$. 
624: This solution is therefore outside the Eckhaus stability
625: region when $1+\alpha\beta>0$.
626: 
627: The four eigenvectors of $A$ are:
628: %
629: \[
630: \left(\begin{array}{c} 0\\1\\0\\0\end{array}\right) \qquad
631: \left(\begin{array}{c} \tilde{e}\\0\\1\\0\end{array}\right) \qquad
632: \left(\begin{array}{c} 0\\ -i k_1 \omega_1^{-1}\\ 1\\
633:       i\omega_1 \end{array}\right) \qquad
634: \left(\begin{array}{c}  0\\ i k_1 \omega_1^{-1}\\ 1\\
635:       -i\omega_1 \end{array}\right) \,.
636: \]
637: %
638: Let
639: %
640: \[
641: D=\left(\begin{array}{rccl}
642:         0 & \tilde{e} & 0 & 0 \\
643:         1 & 0 & 0 & \frac{\lambda^2}{a} \\
644:         0 & 1 & 1 & 0 \\
645:         0 & 0 & 0 & 1
646:         \end{array} \right),
647: D^{-1}=\left(\begin{array}{rccl}
648:         0 & 1 & 0 & -a^{-1}{\lambda^2} \\
649:         {\tilde{e}^{-1}}  & 0 & 0 & 0 \\
650:         -{\tilde{e}^{-1}} & 0 & 1 & 0 \\
651:         0 & 0 & 0 & 1
652:         \end{array} \right).
653: \]
654: %
655: and $\vec{\tilde{X}} \equiv (\tilde{x},\tilde{y},\tilde{z},\tilde{w})
656: = D^{-1}\vec{X}$. The dynamical equations for the new variables become
657: %
658: \[
659: \frac{d}{d\tau}\vec{\tilde{X}}=M(\omega_1)\vec{\tilde{X}}
660: +\epsilon\left(\begin{array}{c}
661: k_2(\tilde{y}+\tilde{z})+\frac{\lambda d}{a}(\tilde{y}+\tilde{z})^2\\
662: \frac{1}{\tilde{e}}(\tilde{x}+\frac{\lambda^2}{a}\tilde{w})(\tilde{y}+\tilde{z})\\
663: -\frac{1}{\tilde{e}}(\tilde{x}+\frac{\lambda^2}{a}\tilde{w})(\tilde{y}+\tilde{z})\\
664: -\frac{d}{\lambda}(\tilde{y}+\tilde{z})^2
665: \end{array}\right) \,,
666: \]
667: %
668: where
669: %
670: \[
671: M(\omega_1)=D^{-1}AD=\left(\begin{array}{rccl}
672:         0 & 0 & 0 & 0 \\
673:         0 & 0 & 0 & 0 \\
674:         0 & 0 & 0 & 1 \\
675:         0 & 0 & -\omega_1^2 & 0
676:                      \end{array}\right) \,.
677: \]
678: 
679: The angular frequency of the solution $\Omega$ should be close to
680: $\omega_1$, $\Omega^2=\omega_1^2+\epsilon \gamma$, with the shift
681: $\gamma$ to be determined later. Next, we change variables to:
682: %
683: \begin{equation}
684: \left\{\begin{array}{rcl}
685: \tilde{x} & = & x_1 \\
686: \tilde{y} & = & x_2 \\
687: \tilde{z} & = & z_1\, \sin\Omega\tau +z_2\, \cos \Omega\tau \\
688: \tilde{w} & = & \Omega z_1\, \cos \Omega\tau -\Omega z_2\, \sin\Omega\tau
689:        \end{array}\right. 
690: \label{eq:transfo}
691: \end{equation}
692: %
693: The 4-$D$ system of equations then takes form:
694: %
695: \begin{small}
696: \begin{eqnarray}
697: \dot{x_1} &=& \epsilon\,\left[ k_2 (x_2 + z_1\sin\Omega \tau + z_2\cos \Omega \tau)
698: 		+\frac{\lambda d}{a} 
699: 		(x_2 + z_1\sin\Omega \tau + z_2\cos \Omega \tau)^2  \right]
700: 								\nonumber\\
701: \dot{x_2} &=& \frac{\epsilon}{\tilde{e}} \left[ x_1 
702: 		+\frac{\Omega \lambda^2}{a} 
703: 		(z_1 \cos\Omega \tau - z_2 \sin \Omega \tau) \right] 
704: 		(x_2 + z_1 \sin \Omega \tau + z_2 \cos \Omega \tau) 
705: 								\nonumber\\
706: \dot{z_1} &=& \frac{\epsilon}{\Omega} \left[ -\frac{d}{\lambda} 
707: 		(x_2 + z_1 \sin\Omega\tau + z_2\cos\Omega\tau)^2
708: 		\cos\Omega\tau \right.				\nonumber \\
709: 	  & & \quad	+ \gamma (z_1 \sin\Omega\tau + z_2 \cos\Omega\tau) 
710: 		\cos\Omega\tau 
711: 								\nonumber \\
712: 	  & & \quad \left. - \frac{\Omega}{\tilde{e}} \left( x_1 
713: 		+ \frac{\Omega \lambda^2}{a} (z_1 \cos\Omega\tau 
714: 		- z_2 \sin\Omega\tau)\right) (x_2 + z_1 \sin\Omega\tau 
715: 		+ z_2 \cos\Omega\tau) \sin\Omega \tau \right]
716: 								\nonumber\\
717: \dot{z_2} &=& \frac{\epsilon}{\Omega} \left[ \frac{d}{\lambda} 
718: 		(x_2 + z_1 \sin\Omega\tau + z_2 \cos\Omega\tau)^2 
719: 		\sin\Omega\tau \right.				\label{sys2} \\
720: 	  & & \quad -\gamma (z_1 \sin\Omega\tau + z_2 \cos\Omega\tau) 
721: 		\sin\Omega\tau 					\nonumber \\
722: 	  & & \quad \left. - \frac{\Omega}{\tilde{e}} \left( x_1 
723: 		+ \frac{\Omega \lambda^2}{a} (z_1 \cos\Omega\tau 
724: 		- z_2 \sin\Omega\tau)\right)
725: 		(x_2 + z_1 \sin\Omega\tau + z_2 \cos\Omega\tau) 
726: 		\cos\Omega\tau \right]   
727: 								\nonumber
728: \end{eqnarray}
729: \end{small}
730: 
731: The proof of the existence of weak MAWs close to $P=0$ relies
732: on a series of theorems from J. Hale's monograph~\cite{jhale}.
733: We reproduce the relevant theorems in appendix~\ref{sec:theorems},
734: and refer to them as the need arises.
735: 
736: Note that the transformation $(\tau, x_1, x_2, z_1, z_2) \rightarrow 
737: (-\tau,-x_1, x_2 , -z_1, z_2)$ leaves the system (\ref{sys2}) invariant. 
738: So, by definition \ref{def1} of appendix~\ref{sec:theorems} the system 
739: has the property $E$ with respect to $Q$, with 
740: %
741: \[
742: Q=\mathrm{diag}(-1,1,-1,1) \,.
743: \]
744: 
745: As we are interested only in the solutions with
746: definite parity, we may start the iteration with the vector
747: %
748: \[
749: \vec{X_0}=(0,a_2,0,a_4) \,.
750: \]
751: %
752: According to Theorem~\ref{sym1}, our solution 
753: $z(\tau,\vec{X_0},\epsilon)$ has the property
754: %
755: \[
756: Qz(-\tau,\vec{X_0},\epsilon)=z(\tau,\vec{X_0},\epsilon) \,,
757: \]
758: %
759: which means that our solutions are either symmetric or antisymmetric.
760: According to Theorem~\ref{sym2}, the second and the fourth determining
761: equations are always zero for this starting vector. For the first and
762: the third determining equations, the zeroth order solution of
763: $\vec{\tilde{X}}$ , {\em i.e.} $\vec{X_0}$, may be substituted, and we get
764: %
765: \begin{eqnarray}
766: k_2 a_2+\frac{\lambda d}{a}(a_2^2+\frac{1}{2}a_4^2) & = & 0 
767: 							\label{nd1}\\
768: \frac{\gamma}{2\Omega}a_4+\frac{\lambda^2 \Omega a_2 a_4}{2a \tilde{e}}
769: -\frac{d} {\lambda \Omega}\, a_2 a_4 		    & = & 0  \,.
770: 							\label{nd2} 
771: \end{eqnarray}
772: %
773: From (\ref{nd2}), we have two possibilities: either $a_4 = 0$ or
774: %
775: \begin{equation}
776: \gamma + a_2 \left( \frac{\lambda^2 \Omega^2}{a \tilde{e}}
777: 	-\frac{2 d}{\lambda} \right) = 0 \,.		\label{nd22}
778: \end{equation}
779: 
780: When $a_4=0$, using $\vec{X_0}=(0,a_2,0,0)$ in~(\ref{sys2}) leads to a
781: trivial constant solution. In the following, we consider only the second
782: case~(\ref{nd22}). We can solve (\ref{nd1}) and (\ref{nd22}) for
783: $\gamma$ and $a_4$ and prove that the system (\ref{sys2}) has periodic
784: solutions. Note that we have three free parameters $\epsilon, a_2, k_2$.
785: But as we will see further, $\epsilon$ and $a_2$ are always combined as
786: $\epsilon a_2$ in the first approximation controlling the amplitude and
787: the period of the solution, and the combination will therefore be
788: regarded here as one single free parameter. For general periodic
789: solutions, $a_2$ can be interpreted as a phase control parameter, {\em
790: i.e.}, a parameter giving the initial location on the periodic orbit at
791: $\tau=\tau_0$. Here, because we only consider symmetric solutions, the
792: translational symmetry of the autonomous system is broken, and that is
793: the reason why $\epsilon$ and $a_2$ combine into a single parameter. The
794: remaining parameter $k_2$ can be chosen freely, for example as to
795: satisfy the {\em consistency condition} (\ref{cc2}), which, when the zeroth
796: order solution is substituted, becomes at order $(\epsilon^2)$:
797: %
798: \begin{equation}
799: -\frac{d^2}{4}\,\Omega^2\,a_4^2+\mu \left(d a_2 
800: 	+\frac{k_2 a}{\lambda}\right)^2=0 \,.		\label{nd3}
801: \end{equation}
802: 
803: At zeroth order, $\Omega^2 = \omega_1^2 = 4\mu$ and $\tilde{e}=4\mu \lambda$. 
804: Solving the system of equations (\ref{nd1}),(\ref{nd22}) and (\ref{nd3}), 
805: we get
806: %
807: \[
808: \left\{\begin{array}{rcl} 
809: k_2 & =- &\frac{3\lambda}{a} d a_2 \\
810: \gamma & = & \frac{c}{a} a_2\\
811: a_4 & = & \pm 2 a_2 \end{array} \right. \,.
812: \]
813: %
814: We can write out the Jacobian for those three equations explicitly:
815: %
816: \[
817: J=\left(\begin{array}{ccc}
818: a_2	&  0  & \frac{\lambda d}{a} a_4 	\\ 
819: 0	&  1  & 0 				\\
820: \frac{2\mu a}{\lambda}(d a_2+\frac{k_2 a}{\lambda}) 
821: 	&  0  & -\frac{d^2}{2} \Omega^2 a_4
822: \end{array}\right) \,.
823: \]
824: %
825: The determinant of this Jacobian is
826: %
827: \[
828: \det J=\frac{1}{2} d^2 \Omega^2 a_2 a_4 \neq 0 \qquad a_2 \neq 0 \,.
829: \]
830: 
831: We now invoke theorem \ref{ther2}, reproduced in
832: appendix~\ref{sec:theorems}, and conclude our proof that system
833: (\ref{or1}) and (\ref{or2}) has periodic solutions near $P=0$. We shall
834: give approximate solutions in section~\ref{sec:approximate}, and show
835: that in this case they contain defects.
836: 
837: 
838: 
839: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
840: \subsection{Case II}
841: 
842: Eigenvalue 0 has only one eigenvector. In this case, we assume that
843: $\frac{\lambda}{a} \tilde{e}+k_1 \neq 0$ to the zeroth order in
844: $\epsilon$, so without loss of generality we can choose $k_2=0$. Then
845: $\frac{\lambda}{a}e=\frac{\lambda}{a} \tilde{e}+k_1$. Implementing the
846: transformation $\vec{X}=D\vec{\tilde{X}}$ with
847: %
848: \[
849: D=\left(\begin{array}{cccc}
850:         0 & \tilde{e} & 0 & 0 \\
851:         1 & 0 & 0 & - k_1\omega_1^{-2} \\
852:         0 & 1 & 1 & 0 \\
853:         0 & 0 & 0 & 1
854:         \end{array} \right)
855: \]
856: %
857: we have
858: %
859: \[
860: \frac{d}{d\tau}\vec{\tilde{X}} = 
861: 	M(\omega_1)\vec{\tilde{X}}+\epsilon\left(\begin{array}{c}
862: 	-\frac{d}{\tilde{e}} k_1 (\tilde{y}+\tilde{z})^2 \\
863: \frac{1}{\tilde{e}}(\tilde{x}-k_1
864: 	\frac{\lambda}{\tilde{e}}\tilde{w})(\tilde{y}+\tilde{z}) \\
865: -\frac{1}{\tilde{e}}(\tilde{x}-k_1
866: 	\frac{\lambda}{\tilde{e}}\tilde{w})(\tilde{y}+\tilde{z}) \\
867: -\frac{d}{\lambda}(\tilde{y}+\tilde{z})^2
868: \end{array} \right) \,,
869: \]
870: %
871: where
872: %
873: \[
874: M(\omega_1)=D^{-1}AD=\left(\begin{array}{rccl}
875:         0 & \frac{\lambda e}{a} & 0 & 0 \\
876:         0 & 0 & 0 & 0 \\
877:         0 & 0 & 0 & 1 \\
878:         0 & 0 & -\omega_1^2 & 0
879:                      \end{array}\right) \,.
880: \]
881: 
882: As in case I, let $\Omega^2=\omega_1^2+\epsilon \gamma$ and perform the
883: same transformation~(\ref{eq:transfo}) into variables $x_1,x_2,z_1,z_2$.
884: We then obtain a 4-$D$ system similar to (\ref{sys2}). However, in the
885: equation for $\dot{x_1}$, there is an $\epsilon$-free term. In order to
886: use the successive approximation method, further transformations are
887: required. Let $\rho \in \mathbb{R}$ such that $\rho^2=\epsilon$. With
888: the transformation $x_2 \rightarrow \rho x_2, \epsilon \rightarrow
889: \rho^2$ we recover the standard form
890: 
891: \begin{small}
892: \begin{eqnarray}
893: \dot{x_1} & = & \frac{\rho \lambda e}{a} x_2-\rho^2 k_1\frac{d}{\tilde{e}}
894: 		(\rho x_2+z_1\, \sin\Omega \tau+z_2 \,\cos \Omega \tau)^2 
895: 								\nonumber\\
896: \dot{x_2} & = &\frac{\rho}{\tilde{e}}
897: 		\left[ x_1-k_1 \frac{\Omega \lambda}{\tilde{e}} 
898: 		(z_1\, \cos\Omega\tau - z_2 \sin\Omega\tau) \right]
899: 		(\rho x_2+z_1 \sin\Omega\tau + z_2 \cos\Omega\tau) 
900: 								\nonumber\\
901: \dot{z_1} & = & \frac{\rho^2}{\Omega} 
902: 		\left[ \frac{-d}{\lambda} 
903: 		(\rho x_2 + z_1 \sin\Omega\tau + z_2\cos\Omega\tau)^2
904: 		\cos\Omega\tau \right. \\			\nonumber \\
905: 	  &   & + \gamma (z_1\sin \Omega \tau+z_2 \cos \Omega \tau) \cos\Omega\tau
906: 								\label{sys33} \\
907: 	  &   & - \frac{\Omega}{\tilde{e}}  \left.
908: 		\left(x_1-\frac{\Omega \lambda}{\tilde{e}} 
909: 		k_1\,(z_1\, \cos\Omega\tau - z_2 \sin\Omega\tau)\right)
910: 		(\rho x_2+z_1 \sin\Omega\tau + z_2 \cos\Omega\tau) 
911: 		\sin\Omega\tau \right]
912: 			 					\nonumber \\
913: \dot{z_2} & = & \frac{\rho^2}{\Omega} \left[ \frac{d}{\lambda} 
914: 		(\rho x_2+z_1 \sin\Omega\tau + z_2 \cos\Omega\tau )^2 
915: 		\sin\Omega\tau \right.
916: 								\nonumber \\
917: 	  &   & - \gamma (z_1\sin\Omega\tau + z_2\cos\Omega\tau) \sin\Omega\tau 
918: 								\nonumber \\
919: 	  &   & - \frac{\Omega}{\tilde{e}} \left.
920: 		\left( x_1 -\frac{\Omega \lambda}{\tilde{e}} k_1 
921: 		(z_1\, \cos\Omega\tau -z_2\sin\Omega\tau) \right) 
922: 		(\rho x_2 + z_1 \sin\Omega\tau + z_2\cos\Omega\tau) \cos\Omega\tau 
923: 		\right] \,,
924: 								\nonumber\\
925: \end{eqnarray}
926: \end{small}
927: %
928: The system (\ref{sys33}) has the same symmetry as identified in the case
929: I. If we are only interested in solutions with definite parity, we may
930: again start the iteration with $\vec{X_0}=(0,a_2,0,a_4)$. To the second
931: order $(\rho^2)$, the determining equations are:
932: %
933: \begin{eqnarray}
934: a_2 \frac{\lambda e}{a}-\rho \frac{d a_4^2 k_1}{2\tilde{e}}	
935: 		+0(\rho^3)&=&0			\label{dd1} \\
936: \rho \frac{\gamma a_4}{2 \Omega}-\rho^2\left(\frac{d a_2 a_4}{\lambda \Omega}
937: 		+\frac{\lambda \Omega a_2 a_4 k_1} {2\tilde{e}^2}\right)
938: 		+0(\rho^3)&=&0 \,.		\label{dd2}
939: \end{eqnarray}
940: %
941: From the second equation we obtain either $a_4=0$ (trivial for
942: our purposes, as discussed above) or
943: %
944: \begin{equation}
945: \gamma-\rho a_2\left(\frac{2 d}{\lambda}
946: +\frac{\lambda \Omega^2 k_1}{\tilde{e}^2}\right)+0(\rho^2)=0 \,.
947: 						\label{dd22}
948: \end{equation}
949: %
950: If we backtrack the transformations made, it is clear that 
951: the {\em consistency condition} requires that we keep terms up to the 
952: fourth order $(\rho^4)$. We found that with the substitution
953: %
954: \[
955: e=\frac{\alpha \omega + \mu}{1+\alpha \beta}
956: +\rho^2(\rho^2 \omega_3-\rho d a_2) \,,
957: \]
958: %
959: where $\omega_3$ is a new parameter, only the fourth or higher order
960: terms are left in the {\em consistency condition}. From the definition $e =
961: R_0^2 = (\omega-\mu \alpha)/(\beta-\alpha)$ and the above equation, we
962: get $\omega \sim \mu \beta$ and then $e \sim \mu$ to the zeroth order.
963: So $R_0 \sim \sqrt{\mu}$, $q \sim 0$, which means that this solution
964: bifurcates from the HOS $A=\sqrt{\mu}\exp({-i \omega t})$. To the leading
965: order ($\rho^4$), we are allowed to use the following substitutions in
966: the {\em consistency condition} (\ref{cc2}):
967: %
968: \begin{eqnarray}
969: a_2 \rightarrow 0 \quad 
970: \omega \rightarrow \mu \beta \quad 
971: \Omega \rightarrow \sqrt{-\frac{2\mu(1+\alpha \beta)}{1+\alpha^2}} ,
972: 							\nonumber \\ 
973: k_1 \rightarrow \frac{\mu \lambda}{a}
974: \left( 1+\frac{2 \lambda (1+\alpha \beta)}{1+\alpha^2}\right) \quad
975: \tilde{e} \rightarrow -\frac{2 \mu \lambda (1+\alpha \beta)} {1+\alpha^2} \,.
976: 							\label{app}
977: \end{eqnarray}
978: %
979: The resulting equation is of a relatively simple form:
980: %
981: \begin{equation}
982: a_4^2(-\lambda+d^2(1+\alpha \beta)(1+\alpha^2+\lambda
983: +\lambda \alpha \beta))+4(1+\alpha \beta)^2 \lambda \mu \omega_3 = 0 \,.
984: 							\label{dd3}
985: \end{equation}
986: %
987: From (\ref{dd1}) it follows that $a_2$ is of order $\rho$, and from 
988: (\ref{dd22}) that $\gamma \sim 0(\rho^2)$. After a change of 
989: variable $a_2=\rho \, a_{22}$ and keeping only the highest 
990: order for the equations, we can rewrite (\ref{dd1}) and (\ref{dd22})
991: as
992: %
993: \begin{eqnarray}
994: a_{22} \frac{\lambda e}{a}-\frac{k_1 d a_4^2}{2 \tilde{e}} & = & 0  \label{ndd1} \\
995: \gamma 							   & = & 0 .\label{ndd2}
996: \end{eqnarray}
997: %
998: For $e, \tilde{e}, k_1$ we use the values in (\ref{app}). 
999: From (\ref{ndd1}), (\ref{ndd2}) and (\ref{dd3}), we can solve for
1000: $a_{22}, \gamma, \omega_3$. The Jacobian of those equations is
1001: %
1002: \[
1003: J=\left(\begin{array}{ccc}
1004:     \frac{\lambda e}{a} & 0  & 0 \\
1005:        0                & 1  & 0 \\ 
1006:        0 & 0 & 4(1+\alpha \beta)^2 \lambda \mu
1007:    \end{array} \right), 
1008: \]
1009: %
1010: So, $\det J = 4(1+\alpha \beta)^2 \lambda^2 \mu e / a \neq 0$ 
1011: for $1+\alpha \beta \neq 0$. According to Theorem~\ref{ther2}, 
1012: we have proved that equations (\ref{or1}) and (\ref{or2})
1013: possess periodic solutions.
1014: 
1015: 
1016: 
1017: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1018: \section{Analytic form of periodic solutions, stability analysis
1019: 	and numerical tests}
1020: \label{sec:approximate}
1021: 
1022: We have proved in the preceding section the existence of symmetric
1023: periodic solutions in case I and II. In both cases, a small parameter
1024: $\epsilon$ or $\rho$ ensures the convergence of successive
1025: approximations. However, we did not give a bound on the highest value of
1026: this parameter, nor did we show that the solutions which we obtain are
1027: the ones observed in numerical simulations. In this section we give the
1028: approximate analytical form of periodic solutions. We 
1029: compare them with direct numerical integration of the CGLe
1030: in case II. 
1031: 
1032: The solutions are shown to be independent of $\lambda$ 
1033: to order $\epsilon$ in case I
1034: and to order $\epsilon^2$ in case II. In addition, these solutions
1035: should also satisfy the 3-$D$ ODE mentioned in
1036: section~\ref{sec:intro:MAWs} which do not contain $\lambda$, so they can
1037: be matched with the solutions of the 3-$D$ system in a unique way,
1038: independent of the value of $\lambda$. Hence, we conclude that to all
1039: orders the physical solutions are identical for the two values of
1040: $\lambda$.
1041: 
1042: The two cases are taken separately. In this section, we reinstate
1043: $x$ as the spatial variable, $R=R(x)$.
1044: 
1045: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1046: \subsection{Case I}
1047: 
1048: Using (\ref{dR}), (\ref{ag}) and the case I calculations of the
1049: preceding section, we have after some algebra:
1050: %
1051: \begin{eqnarray}
1052: R^2 & = & -2 \epsilon d a_2(1\pm \cos \Omega x) 
1053: 						\label{rphi} \\
1054: \phi_x & = & -\frac{\epsilon a_2}{2(1+\alpha^2)\Omega}
1055: \frac{\sin2\Omega x \pm 2\sin\Omega x}{1\pm \cos \Omega x} \,. 
1056: 						\nonumber
1057: \end{eqnarray}
1058: %
1059: To the first order of $\epsilon$, $R$ and $\phi_x$ 
1060: are independent of $\lambda$. The
1061: $\pm$ sign selects two solutions which transform into each other by
1062: translating by a half period. This is reminiscent of the spatial
1063: translational invariance in the symmetric solution space. From the
1064: definitions of $e, \Omega$ and from (\ref{2nde}), (\ref{domega1}), we
1065: get to the first order:
1066: %
1067: \begin{eqnarray}
1068: \omega &=& \mu \alpha-3 \epsilon  a_2 \nonumber\\
1069: \omega_1^2 &=& 4 \mu+\frac{6 \epsilon d a_2}{1+\alpha^2}(\alpha \beta +2 \alpha^2+3)
1070: 	\nonumber\\
1071: \Omega &=& \omega_1+\frac{\epsilon \gamma}{2 \omega_1} \,.
1072: 							\label{sln1}
1073: \end{eqnarray}
1074: 
1075: We see that $\omega$ and $\Omega$ are independent of $\lambda$. On the
1076: other hand, for periodic boundary conditions, we can use Fourier modes
1077: directly to transform the PDE (\ref{cgl}) to a finite set of approximate
1078: ODE's by Galerkin truncation. Then the stationary solution can be
1079: obtained by solving a set of nonlinear algebraic equations. 
1080: 
1081: 
1082: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1083: \subsubsection*{Numerical comparison}
1084: 
1085: If we take as an example the following parameter values (previously
1086: used in~\cite{Lega:holes}) for which defect chaos is expected:
1087: %
1088: \[
1089: \alpha=1.5,\quad  \beta=-1.2
1090: \]
1091: %
1092: and fix the size of the domain to $L=24$, then at $\mu=0.072644$,
1093: $\omega=0.097879$, a periodic solution of period $L/2$ is found. This
1094: solution has $R_{max} \simeq 0.0750$. On the other hand, if we use the
1095: same $\alpha,\beta,\mu$ and search for $R_{max} \simeq 0.075$ by
1096: adjusting $\epsilon$ (we always keep $a_2=1$), we find that 
1097: %
1098: \[
1099: \epsilon \sim 0.00380,\quad \omega=0.097566, 
1100: \quad \mbox{ period } \frac{2\pi}{\Omega}=12.0102 \,.
1101: \]
1102: 
1103: The approximate analytic solution and the numerical solution of the
1104: exact CGLe agree very well. The profile of $R$ from our successive
1105: approximation is shown in Fig.~\ref{fig1}.
1106: 
1107: \begin{figure}[ht] 
1108: \begin{center}  \includegraphics{fcase1.eps} \end{center}
1109: \caption{Spatial profile of the amplitude $R(x)$ at $\mu=0.072644$,
1110: 	from (\ref{rphi}) with $R_{max} = 0.075$. 
1111: \label{fig1}} \end{figure}
1112: 
1113: 
1114: 
1115: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1116: \subsubsection*{Structure near the defect}
1117: 
1118: It is easy to see from (\ref{rphi}) that only $\epsilon a_2>0$ is the
1119: physically interesting combination. However, we may wonder whether it is
1120: really true that $R^2=dP+e$ remains non-negative everywhere while
1121: touching zero at some points. Fig.~\ref{fig1} and the first equation of
1122: (\ref{rphi}) suggest a positive answer to this question. But since we
1123: have only an approximate solution, further justification is needed.
1124: Suppose at some instant $x_0$, we have $dP+e=0$ on the periodic orbit.
1125: From the {\em consistency condition} (\ref{cc2}), at this transition point
1126: %
1127: \[
1128: \frac{d^2(1+\alpha^2)}{4}Q^2 
1129: + \left(\frac{a}{\lambda}\tilde{N}-\lambda Q\right)^2=0,
1130: \]
1131: %
1132: so, $Q=\tilde{N}=0$. According to (\ref{eq:4d}), $\dot{\tilde{M}}=0$ and
1133: $\dot{P}=0$. Assume that $\dot{Q}=0$, then $\dot{\tilde{N}}\neq 0$ since
1134: the point is not an equilibrium. At next instant $x_0+\delta x$, the
1135: {\em consistency condition} can not be satisfied as the two sides of
1136: (\ref{cc2}) have different orders of $\delta x$. So we conclude that
1137: $\dot{Q}\neq 0$ at the point $x_0$, which means that $Q(x_0+\delta x)$
1138: has negative sign to that of $Q(x_0-\delta x)$. Thus, after touching the
1139: zero value plane, $dP+e$ returns to the positive half space again. The
1140: turning happens exactly on the $dP+e=0$ plane. We claim that $dP+e \geq
1141: 0$ always holds and the equality holds periodically. From (\ref{rphi}),
1142: in the neighborhood of $R=0$ at $x=x_0$ on the periodic orbit, $R$
1143: behaves like 
1144: %
1145: \[
1146: R \sim \left(\frac{d \dot{Q}}{2} \right)^{1/2} |x-x_0| \,,
1147: \]
1148: %
1149: and is manifestly a non-analytic function of $x$.
1150: 
1151: We do not discuss the stability of the solutions in case I, as
1152: this has already been accomplished by Tak\'{a}\v{c}~\cite{pt} who has
1153: proven that these solutions are unstable.
1154: 
1155: 
1156: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1157: \subsection{Case II}
1158: 
1159: To the first order of $\epsilon$, the solutions are
1160: %
1161: \[ \left \{ \begin{array}{rcl}
1162:     x_1 &=&  -\frac{\epsilon k_1 a_4^2}{8\tilde{e}^2 \Omega}(2d\tilde{e}
1163: 		+\lambda k_3)\sin2 \Omega x \\
1164:     x_2 &=& \epsilon (a_2^2-\frac{\lambda k_1 a_4^2}{4\tilde{e}^2}\cos2 \Omega x) \\
1165:     z_1 &=& -\frac{\epsilon a_4^2}{12 \tilde{e}^2 \lambda \Omega^2}
1166: 		(3(3d\tilde{e}^2+\lambda^2 \Omega^2 k_1)\sin \Omega x
1167: 		+(d\tilde{e}^2-\lambda^2 \Omega^2 k_1)\sin 3\Omega x) \\
1168:     z_2 &=& a_4+\frac{\epsilon a_4^2}{12\tilde{e}^2\lambda \Omega^2}
1169: 		(3(\lambda^2 \Omega^2 k_1-d\tilde{e}^2)\cos\Omega x
1170: 		+(\lambda^2 \Omega^2 k_1-d\tilde{e}^2)\cos 3\Omega x) \,, \\
1171:    \end{array}\right.
1172: \]
1173: %
1174: where $\epsilon=\rho^2>0$, and $a_4$ is a free parameter. In the
1175: following, we will see that $\epsilon$ and $a_4$ always emerge 
1176: in the combination $\epsilon \, a_4$. To the second order, 
1177: $\omega$ is
1178: %
1179: \[
1180: \omega=\mu \beta +\frac{\epsilon^2 a_4^2}{4\mu(1+\alpha^2)}
1181: 	\left(\frac{1+\alpha \,\beta}{\beta-\alpha} 
1182: 	+\frac{\beta-\alpha}{1+\alpha \,\beta}\right) \,.
1183: \] 
1184: %
1185: It is independent of $\lambda$, and therefore $e,b,\Omega$ are also
1186: independent of $\lambda$. $R$ and $\phi_x$ can also be calculated 
1187: to the second order:
1188: %
1189: \begin{eqnarray}
1190: R^2 &=& -\frac{d^2}{2\mu}(\epsilon a_4)^2 + d \epsilon a_4 \cos\Omega x
1191: 	+\frac{d(\epsilon a_4)^2}{12 \Omega^2} 
1192: 	\left( \frac{c}{a}+\frac{e}{b}\right) \cos2\Omega x + e 
1193: 						\label{sol:case2} \\
1194: \phi_x &=& \frac{\epsilon a_4}{\mu \Omega (1+\alpha^2)}
1195: 	\left[e \sin\Omega x-\frac{\epsilon a_4}{24\Omega^2}
1196: 	\left(6d\Omega^2+\frac{7e^2}{a\Omega^2}+\frac{7ce}{a}\right)
1197: 	\sin2 \Omega x\right]
1198: 						\nonumber 
1199: \end{eqnarray}
1200: 
1201: So clearly $R$ and $\phi_x$ are independent of $\lambda$. Similarly, the
1202: different signs of $a_4$ will give the same solution up to a half-period
1203: translation. This solution is the one observed in the numerics when
1204: passing the Eckhaus instability for underlying wavevector $q=0$. Linear
1205: stability analysis reveals~\cite{stp} that the $q=0$ state, the most
1206: stable state under the long wavelength perturbations, becomes unstable
1207: when the size of the system is such that the smallest possible nonzero
1208: wavenumber $k$ satisfies 
1209: %
1210: \[
1211: k^2 < -\frac{2\mu(1+\alpha \beta)}{1+\alpha^2}\equiv \kappa^2 \,.
1212: \]
1213: %
1214: It is easy to see that $\kappa^2=\omega_1^2$ up to order $(\rho^4)$. 
1215: 
1216: For our parameter choices $\mu=1, \alpha=1.5, \beta=-1.2$,
1217: the bifurcation size of the system is $L_0=\frac{2\pi}{\kappa}=8.95492$.
1218: In the following, we will first prove the stability of our solutions 
1219: near the bifurcation point. Then we will compare them with the stable
1220: solutions observed in numerics.
1221: 
1222: \subsubsection*{Stability analysis: presentation}
1223: 
1224: Assume that $A=R\exp({i\phi})$ where $R, \phi \in \mathbb{R}$ is 
1225: an exact solution of (\ref{cgl}). The perturbed solution is assumed to be 
1226: $\bar{A}=(R+r)\exp({\phi+\theta})$, where $r, \theta \in
1227: \mathbb{R}$ is the perturbation on the amplitude and phase, separately. 
1228: Substitute it into (\ref{cgl}), keeping only the linear terms in $r$ 
1229: and $\theta$. We have 
1230: %
1231: \begin{eqnarray}
1232: r_t & = & (\mu-\phi_x^2-\alpha \phi_{xx}-3 R^2)r
1233: 		+ r_{xx}-2 \alpha \phi_x r_x \nonumber \\
1234:     &   & -(2 R \phi_x+2 \alpha R_x)\theta_x-\alpha R \theta_{xx} 
1235: 						\label{st1} \\
1236: R \theta_t & = & (\omega-\alpha \phi_x^2+\phi_{xx}
1237: 		-3 \beta R^2)r+\alpha r_{xx}+2 \phi_x r_x \nonumber \\
1238:     &   & + (2 R_x-2 \alpha R \phi_x)\theta_x+R \theta_{xx}  \,,
1239: 						\label{st2} 
1240: \end{eqnarray}
1241: %
1242: where in (\ref{st2}) we have used $\phi_t=-\omega$. To study the
1243: stability of the starting solution $A$, we treat these equations as an
1244: eigenvalue problem for a two components vector, {\em i.e.}, we let
1245: $r_t=\sigma r$, $\theta_t=\sigma \theta$ and we investigate the spectra
1246: $\sigma$ of the linear operator resulting from (\ref{st1}) and
1247: (\ref{st2}) in the $C^1$ continuous periodic function space. As the CGLe
1248: has global phase invariance, the eigenvalue equations always have
1249: solution $(r,\theta)=(0,\theta_0)$ with eigenvalue
1250: $\sigma=0$. At the same time, spatial translational invariance implies
1251: that another eigenmode has $\sigma=0$. As a result, saying that the
1252: solution is stable means that it is stable up to a phase and a spatial
1253: translation, and that all other eigenmodes have eigenvalues with
1254: negative real parts.
1255: 
1256: Invoking the expression for $R,\phi_x$ to the second order of
1257: $\epsilon$, the coefficients of various terms of $r,\theta$ and their
1258: derivatives in (\ref{st1}) and (\ref{st2}) become explicit functions of
1259: $x$. The resulting linear operator on $(r,\theta)$ has even parity due
1260: to the symmetry of our solution, and we can consider the even and odd
1261: solutions of $r,\theta$ separately. If we set $\epsilon=0$, i.e., the
1262: starting state $A$ is a plane wave state, then $\cos(n\Omega x)$ and
1263: $\sin(n\Omega x)$ are the eigenfunctions of the unperturbed linear
1264: operator. They give the stability spectrum of the plane waves. Now, let
1265: us move a little (to the order of $\epsilon$) beyond the bifurcation
1266: point. The eigenfunctions are still $\cos(n\Omega x)$ and $\sin(n\Omega
1267: x)$ up to $\epsilon$ corrections. For example, if the even solutions are
1268: considered first, we assume that to the first order the eigenfunctions
1269: are (the time dependence for $r,\theta$ has been suppressed):
1270: %
1271: \begin{eqnarray}
1272: r & = & m_1\cos(n\Omega x)+\epsilon (m_0\cos((n-1)\Omega x)
1273: 	+ m_2\cos((n+1)\Omega x)) 	\label{eig1} \\
1274: \theta & = & n_1\cos(n\Omega x)+\epsilon (n_0\cos((n-1)\Omega x))
1275: 	+ n_2\cos((n+1)\Omega x)) \,,	\label{eig2}
1276: \end{eqnarray}
1277: %
1278: where $n$ is a non-negative integer. Note that we do not include the
1279: terms such as $\epsilon^2 \cos((n\pm 2)\Omega x)$ in the above
1280: expressions because they induce corrections of order $\epsilon^3$ or
1281: higher in the eigenvalues. Now if we substitute (\ref{eig1}) and
1282: (\ref{eig2}) into the eigenvalue equations and identify the coefficients
1283: of $\cos(n\Omega x), \cos((n-1)\Omega x) \mbox{ and } \cos((n+1)\Omega
1284: x)$, a set of six homogeneous linear equations for
1285: $m_0,m_1,m_2,n_0,n_1,n_2$ can be derived. The determinant of the
1286: coefficient matrix will give an eigenvalue equation for $\sigma$. The
1287: resulting expression is too complicated to merit being displayed here.
1288: 
1289: Before bifurcation, the HOS is stable. The first instability occurs for
1290: $n=1$ mode, one eigenvalue of which is very close to $0$ near the
1291: bifurcation point, being negative before and positive after. Meanwhile,
1292: for $n>1$ modes, the corresponding eigenvalues have negative real parts
1293: bounded away from zero. As the bifurcating solution emerges continuously
1294: from the HOS, near the bifurcation point $(\epsilon \ll 1)$ the
1295: perturbed linear operator has all the eigenvalues with negative real
1296: parts away from $0$ for $n>1$ and one eigenvalue close to $0$ for $n=1$.
1297: So, we only need to check the stability of our solutions for $n=1$. 
1298: 
1299: For convenience, we can fix parameters $\alpha$ and $\beta$ to any
1300: values allowed by~(\ref{eq:positivity}) and perform the above 
1301: stability analysis of the solution.
1302: 
1303: 
1304: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%5
1305: \subsubsection*{Stability analysis: numerical checks}
1306: 
1307: The numerical values we used are 
1308: $\mu=1.0, \alpha=1.5,\beta=-1.2,a_4=1$. The eigenvalue equation is then
1309: %
1310: \begin{eqnarray*}
1311:             7.9860 \epsilon^2 \sigma 
1312: + (56.423 - 63.394 \epsilon^2)\sigma^2 
1313: + (82.564 - 75.135 \epsilon^2)\sigma^3 & & \\
1314: + (45.022 - 28.859 \epsilon^2)\sigma^4 
1315: + (10.923 - 4.3059 \epsilon^2)\sigma^5 & & \\
1316: + (1.0    - 0.18864\epsilon^2)\sigma^6 &=& 0 \,.
1317: \end{eqnarray*}
1318: %
1319: $\sigma=0$ corresponds to the neutral mode associated with the global
1320: phase invariance. All others solutions have negative real parts. The
1321: $\sigma_{-}=-0.14154 \epsilon^2$ solution is the interesting one. If we
1322: use the same parameter values to calculate the stability of the HOS, the
1323: eigenvalue equation for $n=1$ is
1324: %
1325: \[
1326: \epsilon^2(-0.21122 - 0.26402\sigma) + 2.98462 \sigma + \sigma^2=0 \,.
1327: \] 
1328: 
1329: To the second order in $\epsilon$, we have $\sigma=-2.98462- 0.19325
1330: \epsilon^2$ or $\sigma_{+}=0.07077\epsilon^2$. The later positive
1331: eigenvalue indicates that the plane wave solution is not stable. We note
1332: that $2\sigma_{+}=-\sigma_{-}$ to order $\epsilon^2$ which
1333: indicates a supercritical pitchfork bifurcation. We have proved
1334: that this equality holds exactly at the bifurcation point
1335: for any values of $\alpha$ and $\beta$, and this justifies the 
1336: above numerical checks.
1337: Under perturbation the HOS will evolve to the
1338: modulated amplitude solution given above. When the instability is
1339: saturated, the corresponding eigenvalue for the MAW is negative. If we
1340: change the sign of $a_4$ or use the other value of $\lambda$, the
1341: eigenvalue does not change, as expected. 
1342: 
1343: If we alternatively consider
1344: the odd-parity function space $\{\sin(n\Omega x)\}_{n \in \mathbb{N}}$,
1345: we obtain the following eigenvalue equation:
1346: %
1347: \begin{eqnarray*}
1348:  (28.2115-33.6568 \epsilon^2)\sigma + (27.1763-21.2703\epsilon^2)\sigma^2 
1349: + (8.92308 - \\
1350:  4.04125\epsilon^2)\sigma^3 + (1 -0.19287\epsilon^2)\sigma^4 =0 \,.
1351: \end{eqnarray*}
1352: %
1353: This equation is quartic because for $n=1$ only two modes $\sin \Omega
1354: x$ and $\sin 2\Omega x$ are used. Now $\sigma=0$ corresponds to the
1355: neutral mode associated with the spatial translation of the CGLe. Other
1356: eigenvalues of the equation have negative real parts bounded away 
1357: from $0$. 
1358: 
1359: To summarize, our solution is stable in the whole phase space of 
1360: the CGLe, up to a phase and a spatial translation.
1361: 
1362: In ref~\cite{sup}, B. Janiaud {\it et al.} have investigated the
1363: stability of traveling waves near the Eckhaus instability in
1364: Benjamin-Feir stable regime. They derived a necessary condition for the
1365: bifurcation to be supercritical and located the corresponding regions as
1366: two strips in the $\alpha, \beta$ parameter space. We have studied the
1367: stationary MAWs in the Benjamin-Feir unstable regime and found that the
1368: bifurcation from the HOS to MAWs is always supercritical, even when
1369: parameter values lay outside of the region given in ref~\cite{sup}.
1370: 
1371: In ref~\cite{quarter2}, application of the perturbation method to the
1372: zeroth order $(\epsilon^0)$ equation gave nonzero eigenvalue
1373: $\lambda_0=2/\beta$. This can not be correct since the zeroth order
1374: equation just gives the stability of the unstable HOS. Furthermore, in
1375: the Galerkin projection, somewhat surprisingly the $N=1$ truncation was
1376: found to give a better result than the $N=2$ truncation. In our case, if
1377: we use only the first order expressions for $R,\phi_x$ in (\ref{st1})
1378: and (\ref{st2}), we cannot get the correct eigenvalues even near the
1379: bifurcation point, not to mention that it would not be possible to
1380: extend the result to the next bifurcation.
1381: 
1382: \subsubsection*{Comparaison with numerical integration of the CGLe}
1383: 
1384: In our numerical simulations we employed a pseudo-spectral method to
1385: evolve equation (\ref{cgl}) using $128$ modes. For system size $L<L_0$,
1386: we always recover the HOS ($q=0$). For $L$ slightly larger than $L_0$,
1387: however, the solution relaxes to the modulated amplitude solution given
1388: irrespective of the initial condition. Figure~\ref{fig2} depicts the
1389: stable steady solutions given by the two methods.
1390: 
1391: \begin{figure}[ht] 
1392: \begin{center} \includegraphics{fcase2.eps} \end{center} 
1393: \caption{Spatial profiles of the amplitude $R$ for $\mu=1, \alpha=1.5, 
1394: 	\beta=-1.2, L=8.958$ from numerical simulation (dots) and 
1395: 	the approximate solution~(\ref{sol:case2}) (solid line). The 
1396: 	agreement is good, with the discrepancy mainly due to the long 
1397: 	relaxation time close to the bifurcation. 
1398: \label{fig2}} 
1399: \end{figure}
1400: 
1401: 
1402: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1403: \section{Conclusion}
1404: 
1405: We have reformulated the stationary one-dimensional CGLe in a finite box
1406: with periodic boundary conditions as a fourth-order ODE for a variable
1407: $P$ that can be interpreted as the modulation of the amplitude squared
1408: of a plane wave solution. This reformulation enabled us to prove the
1409: existence of stationary MAW solutions in the two limit cases
1410: corresponding to the bifurcation of the trivial solution $A=0$ (case I),
1411: or to the bifurcation of the plane wave solution of zero wavenumber
1412: (case II), when those solutions are within the Benjamin-Feir-Newell
1413: regime, or more generally in a region of instability defined
1414: by~(\ref{eq:positivity}). That region coincides with the Eckhaus domain
1415: if $|\alpha|=|\beta|$, but it is different otherwise. We proved the
1416: stability of MAW solutions for the full CGLe in a finite box with
1417: periodic boundary conditions in case II, where a homogeneous plane wave
1418: becomes unstable. We tested our analytical results by comparison of
1419: numerical integrations of the full CGLe with our approximate analytical
1420: solutions.
1421: 
1422: In case I, unstable periodic hole solutions were shown to exist. This
1423: could not be inferred from any phase equation: around the defect point
1424: $A=0$, the amplitude behaves non-analytically, namely piecewise
1425: affinely, and the phase is not defined. In case II we found the
1426: symmetric stable solutions observed in the numerical integrations of the
1427: CGLe just beyond the bifurcation point, using the box size $L$ as the
1428: bifurcation parameter. This bifurcation was shown to be always
1429: supercritical in the Benjamin-Feir-Newell unstable regime. The MAWs
1430: continue to exist when the size $L$ is increased.
1431: 
1432: The analysis of MAWs bifurcating from a plane wave with wavenumber
1433: $0<q<1$ should be similar to the study of case II. It would be
1434: interesting to study the higher order instabilities of MAWs when the
1435: system size is increased beyond the region in which our analysis takes
1436: place. It has been observed that stationary symmetrical MAWs bifurcate
1437: into uniformly-propagating asymmetrical ones via a drift-pitchfork
1438: bifurcation. This happens when $L$ is increased as a consequence of the
1439: growth of the amplitude of the modulation, and the increase of the
1440: spectral richness of the MAW solution. Moreover, MAWs are expected to be
1441: the building blocks of phase turbulence, and the analytical analysis of
1442: their global stability may lead to a characterization of the suspected
1443: transition between phase and defect chaos in the CGLe~\cite{maw,defect}.
1444: 
1445: 
1446: {\bf Acknowledgments} The authors thank Georgia Tech. Center
1447: for Nonlinear Science and G.P. Robinson for support. Conversations
1448: with J. Lega are gratefully acknowledged.
1449: 
1450: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1451: \appendix
1452: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1453: \section{Derivation of the governing equation}
1454: \label{sec:algebra}
1455: 
1456: We use (\ref{dC}) and (\ref{dR}) to rewrite (\ref{rd}) using only $P$ and 
1457: its spatial derivatives:
1458: %
1459: \begin{equation}
1460: 	P(aP_{xxx}+bP_x+cPP_x)_x-P_x(aP_{xxx}+bP_x+cPP_x) = (dP+e)P^3  \label{eq:*}
1461: \end{equation}
1462: %
1463: Note that this equation contains even numbers of derivatives of $P$ in each
1464: term in parenthesis, and also that the powers of $P$ increase while the
1465: derivatives decrease. We now rewrite the equation in a form which take
1466: advantage of this structure. For example, the following equation is
1467: equivalent to (\ref{eq:*}) for any real $\lambda$:
1468: %
1469: \begin{eqnarray*}
1470: 	P(aP_{xxx}+bP_x+(c+\lambda)PP_x)_x-P_x(aP_{xxx}+bP_x+(c+\lambda)PP_x)\\
1471: 	=P^2(\lambda P_{xx}+dP^2+eP) \,,
1472: \end{eqnarray*}  
1473: %
1474: or, put in another form and introducing another real parameter $k$:
1475: %
1476: \begin{eqnarray*}
1477: \left(\frac{(aP_{xx}+bP+\frac{c+\lambda}{2}P^2 )_x}{P}\right)_x
1478: =\lambda P_{xx}+dP^2+\tilde{e}P+\frac{a}{\lambda}kP \,,
1479: \end{eqnarray*} 
1480: %
1481: where we have written $\tilde{e}+\frac{a}{\lambda}k=e$.
1482: 
1483: In this equation, we have three free parameters: besides $\omega$,
1484: introduced by the ansatz~(\ref{eq:antsatz}) as the carrier frequency 
1485: of the solution, we have introduced free parameters $\lambda$ and $k$. 
1486: We now fix $\lambda$ by imposing the condition
1487: %
1488: \begin{eqnarray}
1489: \frac{a}{\lambda}=\frac{b}{\tilde{e}}=\frac{c+\lambda}{2d} \label{condi} \, ,
1490: \end{eqnarray} 
1491: %
1492: which allows us to write the equation in a more suggestive form:
1493: %
1494: \begin{equation*}
1495: \left(\frac{(\lambda P_{xx}+dP^2+\tilde{e}P )_x}{P}\right)_x
1496: =\frac{\lambda}{a}(\lambda P_{xx}+dP^2+\tilde{e}P)+kP \,, 
1497: \end{equation*} 
1498: %
1499: the equation (\ref{ode1}) that leads to the 4-$D$ ODE of
1500: section~\ref{sec:existence}.
1501: 
1502: $\lambda$ is determined by (\ref{condi}):
1503: %
1504: \begin{equation}
1505: \lambda^2+c\lambda-2ad=0 \label{l1} \,.
1506: \end{equation}
1507: %
1508: The discriminant of (\ref{l1}) is
1509: \begin{eqnarray*}
1510: \Delta & = & c^2+8ad \\
1511:        & = & \left(\frac{1+\alpha^2}{2(\beta-\alpha)}\right)^2\left(\frac{9(1+\alpha \beta)^2}
1512: {(\beta-\alpha)^2}+8\right) \,.
1513: \end{eqnarray*}
1514: %
1515: So $\Delta>0$ for any real values of $\alpha$ and $\beta$, and the quadratic equation
1516: (\ref{l1}) always has two real roots
1517: %
1518: \begin{eqnarray}
1519: \lambda=\frac{3(1+\alpha \beta)(1+\alpha^2)}{4\,(\beta-\alpha)^2} \pm \frac{1+\alpha^2}
1520: {4(\beta-\alpha)^2} \sqrt{9(1+\alpha \beta)^2+8 (\beta-\alpha)^2} \label{lam}
1521: \end{eqnarray}
1522: 
1523: Note that $\lambda$ is a function of $\alpha$ and $\beta$ only. In some
1524: applications~\cite{per}, the two values of $\lambda$ correspond to two
1525: distinct solutions of the CGLe. In our case, $\lambda$ is an
1526: intermediate variable used in the derivation and the proofs, but our
1527: solutions to the CGLe do not distinguish the two values of $\lambda$. 
1528: 
1529: 
1530: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1531: \section{Theorems used in the proofs}
1532: \label{sec:theorems}
1533: 
1534: We use successive approximation method to prove the existence of
1535: modulated amplitude waves. Below are listed several theorems from the
1536: theory of nonlinear oscillations taken from Hale's
1537: monograph~\cite{jhale}. 
1538: 
1539: Consider the system of equations
1540: %
1541: \begin{equation}
1542: \dot{z}=Az+\epsilon Z(\tau,z,\epsilon) \label{sys1}
1543: \end{equation}
1544: %
1545: where $A$ is a constant matrix, $\epsilon,\tau \in \mathbb{R}$, and $z,Z
1546: \in \mathbb{R}^n$. $Z$ is a continuous function of $\tau,z,\epsilon$,
1547: periodic in $\tau$ of period $T$. In the following, we only consider the
1548: case that $Z$ is a smooth function. Without loss of generality, $A$ can
1549: always be assumed to have the standard form
1550: \[
1551: A=\mathrm{diag}(0_p,B),
1552: \]
1553: Where $0_p$ is a $p\times p$ zero matrix and $B$ is a constant matrix
1554: with the property that the equation $\dot{y}=By$ has no nontrivial
1555: periodic solution of period $T$. Under these settings, if the successive
1556: approximation is applied to (\ref{sys1}), we have
1557: %
1558: \begin{theorem} \label{ther1}
1559: Given $d>b>0$, there is an $\epsilon_1>0$ such that for any given 
1560: constant $p$ vector $a,\| a \| <b$ and real $\epsilon, 
1561: |\epsilon|<\epsilon_1$, there is a unique function
1562: \[
1563: z^*(\tau)=z(\tau,a,\epsilon), \mbox{with } \underset{\tau}{\sup}\|z^*(\tau)\|<d
1564: \] 
1565: which has continuous first derivative with respect to $\tau$ and satisfies
1566: \[
1567: \dot{z^*}=Az^*+\epsilon Z(\tau,z^*,\epsilon)-\epsilon P_0 Z(\tau,z^*,\epsilon).
1568: \]
1569: Furthermore, $z(\tau,a,0)=a^*$, $a^*=\mathrm{col}(a,0)$, $P_0(z^*)=a^*$, 
1570: and $z(\tau,a,\epsilon)$ has continuous first derivatives with respect 
1571: to $a, \epsilon$.  
1572: \end{theorem}
1573: %
1574: $P_0$ is defined as a projection operator on the Banach space $S$
1575: of continuous periodic functions of period T. If $f\in S$, write
1576: $f=\mathrm{col}(g,h)$ where $g$ is a $p$ vector and $h$ is a $n-p$ 
1577: vector, then
1578: %
1579: \[
1580: P_0(f)=\mathrm{col}\left(T^{-1}\int_0^T g(t)\, dt,0 \right)
1581: \]
1582: %
1583: So, $P_0$ brings an element $f$ in $S$ to a constant vector which
1584: has the average values of $g$ over one period as the first $p$ 
1585: components and zeros as the rest components. The equation satisfied 
1586: by $z^*$ is different from (\ref{sys1}) by a constant vector. By 
1587: a proper choice of the starting vector $a$, we may make this 
1588: constant vector zero to obtain a solution for the system (\ref{sys1}). 
1589: The mathematical statement is give by the following theorem.
1590: %
1591: \begin{theorem} \label{ther2}
1592: Let $z(\tau,a,\epsilon)$ be the function given by the 
1593: Theorem~\ref{ther1} for all $\|a\| \leq b <d, 
1594: |\epsilon| \leq \epsilon_1$. If there exist an 
1595: $\epsilon_2 \leq \epsilon_1$ and a continuous function 
1596: $a(\epsilon)$ such that 
1597: %
1598: \begin{equation}
1599: P_0Z(\tau,z(\tau,a(\epsilon),\epsilon),\epsilon)=0, 
1600: \;\mbox{with }\|a(\epsilon)\| \leq b 
1601: \mbox{   for } |\epsilon| \leq \epsilon_2 \label{cond1} 
1602: \end{equation}
1603: %
1604: then $z(\tau,a(\epsilon),\epsilon)$ is a periodic solution 
1605: of system (\ref{sys1}) for $\|\epsilon\| \leq \epsilon_2$. 
1606: Conversely, if system (\ref{sys1}) has a periodic solution 
1607: $\bar{z}(\tau,\epsilon)$, of period $T$, $\|\bar{z} (\tau,\epsilon)\| 
1608: \leq d, |\epsilon| \leq \epsilon_2$, then $\bar{z} (\tau,\epsilon)
1609: =z(\tau,a(\epsilon),\epsilon)$. 
1610: \end{theorem}
1611: %
1612: Therefore, the existence of a continuous function $a(\epsilon)$ 
1613: satisfying (\ref{cond1}) is a necessary and sufficient condition 
1614: for the existence of a periodic solution of system (\ref{sys1}) 
1615: of period $T$. As we do not know the exact functional form of 
1616: the periodic solution, the condition (\ref{cond1}) could not be 
1617: solved explicitly. But by using implicit function theorem, we 
1618: can show that the substitution into (\ref{cond1}) of a proper
1619: approximate function of $z(\tau,a,\epsilon)$ leads to the 
1620: existence condition for periodic solutions. 
1621: %
1622: \begin{theorem} \label{ther3}
1623: In the system (\ref{sys1}), let
1624: \[
1625: Z=\mathrm{col}(X,Y), \quad z=\mathrm{col}(x,y)
1626: \]
1627: where $X,x$ are $p$ vectors and define
1628: \[
1629: X_0(x,y,\epsilon)=\frac{1}{T}\int_0^T X(\tau,x,y,\epsilon) d\tau.
1630: \]
1631: If there is a $p$ vector $a_0, \|a_0\|<d $, such that
1632: %
1633: \begin{equation}
1634: X_0(a_0,0,0)=0, \quad \det \left[\frac{\partial X_0(a_0,0,0)}{\partial x}\right]
1635: \neq 0  \label{cond2}
1636: \end{equation}
1637: %
1638: then there exists an $\epsilon_1>0$ and a periodic function 
1639: $z(\tau,\epsilon), |\epsilon| \leq \epsilon_1$, of system 
1640: (\ref{sys1}) of period $T$ with $z(\tau,0)=\mathrm{col}(a_0,0)$.
1641: \end{theorem} 
1642: 
1643: If we need to determine other parameters as a function of $\epsilon$ in
1644: practical applications, similar theorems could be derived. Specifically,
1645: in the main text we consider the period $T$ as a function of $\epsilon$.
1646: It is clear that theorem~\ref{ther3} applies if we suppose $T(\epsilon)$
1647: is continuous in $\epsilon$ and bounded for $|\epsilon| \leq
1648: \epsilon_1$. Furthermore, despite the use of the zeroth approximation in
1649: the above theorem, the $n$th approximation could be used instead. If
1650: {\em simple} (non-vanishing determinant) solutions to the determining
1651: equations can be found for $\epsilon$ in the neighborhood of 0
1652: then system (\ref{sys1}) has a periodic solution.
1653: 
1654: If the system which we are studying possesses certain symmetries, we can
1655: prove the existence of particular symmetric solutions by a simplified
1656: version of determining equations. Let us define the symmetry first.
1657: 
1658: \begin{defit}
1659: Let $\dot{z}=f(\tau,z)$, where $z,f \in \mathbb{R}^n$, be a system of
1660: differential equations. It is said to have the property $E$ with respect
1661: to $Q$ if there exists a nonsingular matrix $Q$ such that
1662: %
1663: \[
1664: Q^2=I \quad Qf(-\tau,Qz)=-f(\tau,z) \quad QP_0=P_0 Q 
1665: \]
1666: where $P_0$ is the projection operator defined before. \label{def1}
1667: \end{defit}     
1668: 
1669: Under this symmetry assumption the following theorems apply:
1670: %
1671: \begin{theorem} \label{sym1}
1672: Suppose $Q=\mathrm{diag}(Q_1,Q_2)$ where $Q_1$ is a 
1673: $p \times p$ matrix. If system (\ref{sys1}) has property $E$ 
1674: with respect to this $Q$ for all $\epsilon$. If $a,\|a\| \leq b,$ 
1675: is a $p$ vector and $a^*=\mathrm{col}(a,0)$ is a $n$ vector, 
1676: chosen in such a way that $Q a^*=a^*$, then the solution 
1677: $z(\tau,a,\epsilon)$ satisfies the relation
1678: \[
1679: Qz(-\tau,a,\epsilon)=z(\tau,a,\epsilon)
1680: \] 
1681: and consequently,
1682: \[
1683: Z(-\tau,z(-\tau,a,\epsilon),\epsilon)=-Qz(\tau,z(\tau,a,\epsilon),\epsilon)
1684: \]
1685: \end{theorem}
1686: %
1687: %
1688: \begin{theorem} 
1689: \label{sym2}
1690: If the $j$-th element of the diagonal of the matrix $Q_1$ in 
1691: Theorem~\ref{sym1} is $+1$, then the $j$-th equation in the determining
1692: equations is equal to zero for every vector $a^*$ in Theorem~\ref{sym1}.
1693: \end{theorem}
1694: 
1695: The system (\ref{eq:4d}) derived here from the 1-$D$ CGLe has this
1696: symmetry, so the number of determining equations can be reduced using
1697: these two theorems.
1698: 
1699: 
1700: 
1701: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1702: \begin{thebibliography}{99}
1703: 
1704: \bibitem{patt}
1705:     M.C.~Cross and P.C.~Hohenberg,
1706:     {Pattern formation outside of equilibrium},
1707:     {\em Rev. Mod. Phys.} {\bf 65}, 851 (1993).
1708: 
1709: \bibitem{WoCGLe}
1710:     I. Aranson and L. Kramer,
1711:     {The world of the complex Ginzburg-Landau equation},
1712:     {\em Rev. Mod. Phys.} {\bf 74}, 99 (2002).
1713: 
1714: \bibitem{sub}
1715:      W.~Sch\"{o}pf and L.~Kramer,
1716:     {Small-Amplitude Periodic and Chaotic Solutions of the
1717: 	Complex Ginzburg-Landau Equation for a Subcritical Bifurcation},
1718:     {\em Phys. Rev. Lett.} {\bf 66} (18), 2316 (1991).
1719: 
1720: \bibitem{amp}
1721:      M. van Hecke, P.C. Hohenberg and W. van Saarloos,
1722:     {Amplitude equations for pattern forming systems}, 
1723:     in H. van Beijeren and M. H. Ernst, eds,
1724:     {\em Fundamental Problems in Statistical Mechanics VIII}
1725:     (North-Holland, Amsterdam, 1994).
1726: 
1727: \bibitem{Shraiman:92}
1728:     B.I.~Shraiman, A. Pumir H. Chat\'e,
1729:     {Spatiotemporal chaos in the one-dimensional complex Ginzburg-Landau equation}
1730:     {\em Physica D} {\bf 57}, 241 (1992).
1731: 
1732: \bibitem{sup}
1733:     B.~Janiaud, A. Pumir, D. Bensimon, V. Croquette, H. Richter and L. Kramer,
1734:     {The Eckhaus instability for traveling waves},
1735:     {\em Physica D} {\bf 55}, 269 (1992).
1736: 
1737: \bibitem{bk}
1738:      N.Bekki and K. Nozaki,
1739:     {Formations of spatial patterns and holes in the generalized Ginzburg-Landau equation},
1740:     {\em Phys. Lett.} {\bf 110A}, 133 (1985).
1741: 
1742: \bibitem{jhale}
1743:     Jack K. Hale,
1744:     {\em Oscillations in Nonlinear Systems} 
1745:     (Mc Graw-Hill, New-York, 1963).
1746: 
1747: \bibitem{hohen}
1748:     W. Van Saarloos and P.C. Hohenburg,
1749:     {Fronts, pulses, sources and sinks in generalized complex Ginzburg-Landau equations},
1750:     {\em Physica D} {\bf 56}, 303 (1992).
1751: 
1752: \bibitem{Newell:74}
1753:     A. Newell,
1754:     {Enveloppe equations},
1755:     {\em Lectures in Applied Math.} {\bf 15}, 157 (1974).
1756: 
1757: %\bibitem{symm}
1758: %    A.~Doelman,
1759: %    {Breaking the hidden symmetry in the Ginzburg-Landau equation},
1760: %    {\em Physica D} {\bf 97}, 398 (1996).
1761: 
1762: %\bibitem{tb}
1763: %    M.~Bazhenov, T.~Bohr, K.~Gorshkov and M.~Rabinovich,
1764: %    {The diversity of steady state solutions of the complex Ginzburg-Landau equation},
1765: %    {\em Phys. Lett. A} {\bf 17}, 104 (1996).
1766: 
1767: \bibitem{build}
1768:     M.~van~Hecke,
1769:     {Building Blocks of Spatiotemporal Intermittency},
1770:     {\em Phys. Rev. Lett.} {\bf 80}, 1896 (1998).
1771: 
1772: \bibitem{maw}
1773:      L.~Brusch, M.G.~Zimmermann, M.~van Hecke, M.~B\"{a}r and A.~Torcini,
1774:     {Modulated amplitude waves and the transition from phase to defect chaos},  
1775:     {\em Phys. Rev. Lett.} {\bf 85}, 86 (2000).
1776: 
1777: \bibitem{defect}
1778:      L.~Brusch, A.~Torcini, M.~van~Hecke, M.G.~Zimmermann and
1779:      M. B\"{a}r,
1780:     {Modulated amplitude waves and defect formation in the one-dimensional 
1781:      complex Ginzburg-Landau equation},
1782:     {\em Physica D} {\bf 160}, 127 (2001).
1783: 
1784: \bibitem{quarter1}
1785:      P.K.~Newton and L.~Sirovich,
1786:      {Instabilities of Ginzburg-Landau equation: periodic solutions},
1787:      {\em Quart. of Appl. Math.} {\bf XLIV}, 49 (1986).
1788: 
1789: \bibitem{quarter2}
1790:      P.K.~Newton and L.~Sirovich,
1791:      {Instabilities of Ginzburg-Landau equation: secondary bifurcation},
1792:      {\em Quart. of Appl. Math.} {\bf XLIV}, 367 (1986).
1793: 
1794: \bibitem{pt}
1795:      P.~Tak\'{a}\v{c},
1796:      {Invariant 2-tori in the time-dependent Ginzburg-Landau equation},
1797:      {\em Nonlinearity} {\bf 5}, 289 (1992).
1798: 
1799: \bibitem{ellip}
1800:      D.F.~Lawden,
1801:      {\em Elliptic Functions and Applications} 
1802:      Appl. Math. Sci. {\bf 80}
1803:      (Springer, New-York, 1989).
1804: 
1805: \bibitem{stp}
1806:      B.J.~Matkowsky and V.~Volpert,
1807:      {Stability of Plane Wave Solutions of Complex Ginzburg-Landau Equations},
1808:      {\em Quart. of Appl. Math.} {\bf LI} (2), 265 (1993). %265-281
1809: 
1810: \bibitem{per}
1811:      A.V.~Porubov and M.G.~Velarde,
1812:      {Exact periodic solutions of the complex Ginzburg-Landau equation},
1813:      {\em J. Math. Phys.} {\bf 40}, 884 (1999).
1814: 
1815: \bibitem{garnier}
1816:      N.~Garnier, A.~Chiffaudel and F. Daviaud,
1817:      {Nonlinear dynamics of waves and modulated waves in 1D thermocapillary 
1818: 	flows: I. periodic solutions},
1819:      {\em Physica D}, to appear (2002).
1820: 
1821: \bibitem{Lega:holes}
1822:      J. Lega,
1823:      {Traveling hole solutions of the complex Ginzburg Landau equation: 
1824: 	a review},
1825:      {\em Physica D} {\bf 152-153}, 269 (2001).
1826: 
1827: \end{thebibliography}
1828: 
1829: 
1830: \end{document}
1831: