nlin0207009/bch.tex
1: \documentclass{elsart}
2: %\documentclass[doublespacing]{elsart}
3: %\newcommand{\boite}[2]{\centering{\fbox{\rule{#1}{0cm}\rule{0cm}{#2}}}}
4: %\usepackage[T1]{fontenc}
5: %\usepackage[latin1]{inputenc}
6: \usepackage{amsmath,amssymb}
7: \usepackage{longtable}
8: \newif\iffigs
9: \figstrue
10: %\figsfalse
11: 
12: %\def\ind{{_E}}
13: %\def\EV{\nu_\ind} initial noise amplitude and the
14: 
15: %\textwidth=15.5cm
16: %\textheight=23cm
17: %\evensidemargin=-1.0cm
18: %\oddsidemargin=-0.0cm
19: % Macros locales
20: \def\pt{{\partial_t}}
21: \def\pdm{{\partial_x^{-4}}}
22: \def\pem{{\partial_x^{-5}}}
23: \def\pfm{{\partial_x^{-6}}}
24: \def\pcm{{\partial_x^{-3}}}
25: \def\pxxm{{\partial_x^{-2}}}
26: \def\pxm{{\partial_x^{-1}}}
27: \def\px{{\partial_x}}
28: \def\pxx{{\partial_x^2}}
29: \def\pc{{\partial_x^3}}
30: \def\pd{{\partial_x^4}}
31: \newcommand{\dps}[1]{\displaystyle #1}
32: %\newcommand{\f}[2]{\frac{#1}{#2}}
33: \def\lc{{\lambda}}
34: %\newcommand{\be}{\begin{equation}}
35: %\newcommand{\ee}{\end{equation}}
36: \def\f#1#2{{\frac{#1}{#2}}}
37: %\def\b#1#2{\left( {#1 \over #2} \right)}
38: \def\pa#1{{\partial #1}}
39: \def\pr#1{{\left( #1 \right)}}
40: \def\va{{v_a}}
41: \def\vb{{v_b}}
42: \def\va#1{{v_{a #1}}}
43: \def\vb#1{{v_{b #1}}}
44: \def\vfa{{\varphi_a}}
45: \def\vfb{{\varphi_b}}
46: \def\vpa#1{{\varphi_{a #1}}}
47: \def\vpb#1{{\varphi_{b #1}}}
48: \def\vvv{{\overline{v}}}
49: \def\vv#1{{\overline{v}^{(#1)}}}
50: \def\vur{{\overline{v_r}^{(1)}}}
51: \def\vub{{\overline{v_\beta}^{(1)}}}
52: \def\c#1{{c_{#1}}}
53: \def\Q#1{{Q^{(#1)}}}
54: \def\Gr#1{{\mathcal{G}_{#1}}}
55: \def\FGr#1{{\widehat{\mathcal{G}}_{#1}}}
56: \def\tv{{\tilde{v}}}
57: \def\Lr#1{{\mathcal{L}_{#1}}}
58: 
59: \iffigs
60: \usepackage{graphicx}
61: \graphicspath{{figs/}}
62: \fi
63: 
64: \begin{document}
65: \bibliographystyle{elsart-num}
66: \runauthor{Legras and Villone}
67: \begin{frontmatter}
68: \title{Dispersive and friction-induced stabilization of an inverse
69: cascade. The theory for the Kolmogorov flow in the slightly 
70: supercritical regime}
71: \author[LMD]{Bernard Legras}
72: \ead{legras@lmd.ens.fr}
73: \author[ICG]{Barbara Villone}
74: \address[LMD]{Laboratoire de M\'et\'eorologie Dynamique, UMR8539,
75: 24 rue Lhomond, 75231 Paris Cedex 5, France}
76: \address[ICG]{Istituto Cosmogeofisica, CNR, Corso Fiume 4,
77: 10133 Torino, Italy}
78: \date{22 May 2002}
79: 
80: \begin{abstract} 
81:   We discuss the stabilisation of the inverse cascade in the large
82:   scale instability of the Kolmogorov flow described by the complete
83:   Cahn-Hilliard equation with inclusion of $\beta$ effect, large-scale
84:   friction and deformation radius.  The friction and the $\beta$
85:   values halting the inverse cascade at the various possible
86:   intermediate states are calculated by means of singular
87:   perturbation techniques and compared to the values resulting from
88:   numerical simulation of the complete Cahn-Hilliard equation. The
89:   excellent agreement validates the theory. Our main result is that 
90:   the critical values of friction or $\beta$ halting the inverse cascade
91:   scale exponentially as a function of the jet separation in the final
92:   flow, contrary to previous theories and phenomenological approach.
93:  
94:   
95: \end{abstract}
96: \begin{keyword}
97: Cahn-Hilliard equation, Kolmogorov flow, inverse cascade, large-scale 
98: instability, nonlinear instability \\
99: \emph{PACS numbers}\,: 47.10.+g, 92.60.Ek, 47.27.-i, 47.27.Ak,
100:   47.35.+i
101: \end{keyword}
102: \end{frontmatter}
103: 
104: \section{Introduction}
105: \label{s:intro}
106: 
107: Inverse cascades are a common feature in the large-scale velocity and
108: magnetic fields of geophysical, planetary and astrophysical
109: two-dimensional flows.  Their halting by spontaneous formation of
110: zonal jets has been object of great interest and considerable work by
111: many scientists, see among other
112: Refs~\cite{Rhines:75,Williamson:78,Rhines:94,Vallis:93,Manfroi:99}.
113: The phenomenon of an halted inverse cascade could play a role in the 
114: atmosphere of Jupiter and other Jovian
115: planets which exhibit jet streams of east-west and west-east
116: circulation.
117: 
118: Frisch et al. \cite{Frisch:96} showed that the inverse cascade of the
119: large-scale nonlinear instability of the Kolmogorov flow described by
120: the Cahn--Hilliard equation
121: \cite{Meshalkin:61,Sivashinsky:85,Nepomnyashchy:76} may be stopped by
122: the dispersive Rossby waves, i.e. by the so-called the $\beta$-effect. We
123: recall here that in the absence of any stabilizing effect the inverse
124: cascade proceeds by visiting a family of metastable states with
125: increasing scale until the final largest scale is reached
126: \cite{She:87}.  In a later paper, Legras et al. \cite{Legras:99} have
127: used singular perturbation techniques to calculate the range of the
128: $\beta$ values stopping the inverse cascade in one of the otherwise
129: (if $\beta$ = 0) metastable state.  Their result was different from
130: that obtained by the standard phenomenology based on dimensional
131: arguments \cite{Rhines:75} which fails because it does not take into
132: account the strong suppression of non linearities in the metastable
133: states of the inverse cascade.  
134: 
135: In Refs.~\cite{Frisch:96,Legras:99} the forcing maintaining the basic
136: flow was chosen parallel to the planetary vorticity contours and there
137: was no friction or advection. In a more realistic setup, Manfroi and
138: Young \cite{Manfroi:99} studied the stability of a forced meridian
139: flow on a $\beta$-plane pushing the fluid across the planetary
140: vorticity contours, and including both friction and advection by a
141: mean flow.  They obtained a complete amplitude equation for the
142: leading order perturbation, from which, with some formal modifications
143: and with a slightly different interpretation of the parameters, the
144: amplitude equation of Ref.~\cite{Frisch:96} can be recovered (see
145: Section.\ref{s:basiceq}).  Using this equation, but disregarding the
146: dispersive contribution from $\beta$-effect, they showed that random
147: initial perturbations rapidly reorganize into a set of fast and narrow
148: eastward jets separated by slower and broader westward jets, followed
149: then by a much slower adjustment of the jets, involving gradual
150: migration and merger. The stabilization discussed in
151: Ref.~\cite{Manfroi:99} is only due to friction and not to dispersive
152: effects.
153: 
154: In this paper we present a systematic approach to the study of
155: stabilization of the inverse cascade in the supercritical regime of
156: the large-scale Kolmogorov flow provided both by friction and $\beta$
157: effect.  The stabilization by $\beta$ effect already discussed in
158: \cite{Legras:99} will be here presented in a greater detail and
159: stabilization by friction will be discussed in the same mathematical
160: approach as for the $\beta$ case.  The mathematical framework is based
161: on the kink dynamics introduced by Kawasaki and Ohta
162: \cite{Kawasaki:82} to describe the solutions to the Cahn--Hilliard
163: equation; singular perturbation technique is used to study the
164: stability of these solutions with respect to small perturbations due
165: to friction and $\beta$ effect. The perturbative calculations are
166: performed analytically for large wavenumbers and numerically for all
167: cases. The results are compared together and with direct numerical
168: stability and time-dependent solutions of the amplitude equation.
169: 
170: The paper is organized as follows: in Section~\ref{s:compCH} the
171: Cahn--Hilliard equation in its {\em complete} form is presented; by
172: complete in this context we mean that $\beta$, friction, advection
173: velocity and deformation terms have been added to the standard
174: Cahn--Hilliard equation.  The perturbation to the steady metastable
175: solution of the Cahn--Hilliard equation by small $\beta$ and a
176: friction terms is presented in Section~\ref{s:persta}.
177: Section~\ref{s:stab} discusses the stability of the steady metastable
178: solutions and provides the main results of this work.
179: Section~\ref{s:numerical} presents the techniques used to solve
180: numerically the perturbation and the time-dependent problem, and
181: compares the results of those calculations and the analytical results.
182: Section~\ref{s:conclu} offers a summary and conclusions.
183: 
184: We do not give the detailed presentation of the mean advection effect,
185: which introduces considerable additional algebraic complications, in
186: order to focus here on the essential mechanisms that stabilizes the
187: Cahn-Hilliard cascade. The main results in presence of mean advection
188: are, however, given without demonstration in Appendix~\ref{s:meanadv}.
189: 
190: %% SECTION II
191: \section{The complete Cahn--Hilliard equation}
192: \label{s:compCH}
193: 
194: \subsection{Basic equation}
195: \label{s:basiceq}
196: 
197: Our starting point is the large-scale Kolmogorov flow in its slightly
198: supercritical regime, which is described by the Cahn-Hilliard equation
199: \cite{Meshalkin:61,Sivashinsky:85,Nepomnyashchy:76}.  We recall here
200: that the basic Kolmogorov flow, ${\bf u} = (\cos y,0)$, is maintained
201: by a force ${\bf f}= \nu(\cos y,0)$ against viscous dissipation.  This
202: flow exhibits a large-scale instability of the negative eddy viscosity
203: type when the kinematic viscosity $\nu$ is slightly below the critical
204: value $\nu_c=1/\sqrt2$. 
205: 
206: Taking further into account $\beta$ effect, friction $r$, external
207: deformation radius $1/S$ \footnote{The external deformation radius
208:   accounts for the inertial large-scale effect of a free surface in
209:   geophysical flows \cite{Pedlosky:87}.} and advection effect due to
210: a non zero mean velocity $\gamma$ , we obtain by multi-scale techniques
211: the following adimensional equation for the leading order large-scale
212: perturbation :
213: \begin{equation}
214:    \pt (1 - S^2 \pxxm) (v-\gamma) = \lc \pxx W'(v) - \lc \pd v 
215:    -\beta \pxm (v-\gamma) - r v,
216:    \label{vbetach}
217: \end{equation}
218: where $\pxm$ denotes the integration in $x$ defined for the family of
219: functions with zero average over the interval $[0,L]$.  The constants
220: in (\ref{vbetach}) are
221: \begin{equation}
222:    s=\frac{1}{\sqrt{3}},\qquad
223:    \Gamma=\sqrt{\frac{3}{2}},\qquad 
224:    \lambda_3=\frac{3}{\sqrt2}.
225:    \label{deflam123}
226: \end{equation}
227:  and the potential $W(v)$ is
228: \begin{equation} 
229:    W(v)= \frac{s^2}{2 \Gamma^2} v^{4} - s^2 v^2 
230: \label{Wdef} 
231: \end{equation}
232: Equation (\ref{vbetach}) was derived in \cite{Frisch:96} with $S=r=\gamma=0$.
233: In this derivation, the Kolmogorov basic flow is oriented in the zonal 
234: direction on the $\beta$-plane and therefore the large amplitude flow 
235: develops in the meridional direction. Using a more realistic setting
236: where the Kolmogorov basic flow is oriented in the meridional 
237: direction as an idealized baroclinic perturbation, and introducing friction 
238: and  mean advection velocity, 
239: Manfroi and Young~\cite{Manfroi:99} derived the following amplitude equation
240: \begin{multline}
241:    \partial_\tau A = - r A - (2-{\gamma^*}^2) 
242:    \partial_{\eta^2} A - 3 \partial_{\eta^4} A \\ 
243:    + 2 \gamma^* \partial_\eta(\partial_\eta A^2)
244:    + \frac{2}{3} \partial_\eta(\partial_\eta A^3) 
245:    - \beta \partial_{\eta^{-1}} A 
246:    \label{eq:mf}
247: \end{multline}
248: where $\tau$ and $\eta$ are the temporal and spatial variables.  
249: In their study, the last term on the right-hand side, which is
250: the only dispersicve term in the equation, was set to zero.
251:  \footnote{The coefficient $\beta$ in the last term of the right hand side
252:   of (\ref{eq:mf}) is the product of the planetary vorticity gradient
253:   per the small angle between the Kolmogorov flow direction and the
254:   planetary vorticity gradient.  Therefore it can be either positive
255:   or negative unlike in the derivation of Ref.~\cite{Frisch:96} where it is
256:   always positive}.  By the change of variable $\px A=w-\gamma$ and
257: the rescaling $\tau= a t$, $\eta=bx$, $w=cv$ and $\gamma^*=c \gamma$
258: where $a=12 s^4 \lc / (2+{\gamma^*}^2)^2$,
259: $b=(6s^2/(2+{\gamma^*}^2))^{1/2}$ and $c=(3(2+{\gamma^*}^2)/(2
260: \Gamma^2))^{1/2}$, this equation is easily transformed into
261: (\ref{vbetach}) up to the term in $S$ which is only a trivial
262: modification \cite{Pedlosky:87}.  We see that the quadratic term in
263: (\ref{eq:mf}) disappears in this transformation.  The $\gamma$ term
264: represents the net advection velocity (see \cite{Manfroi:99}) and is also
265: the average value of $v$.
266: 
267: In a recent work \cite{Manfroi:02}, it is claimed that the Kolmogorov 
268: instability exhibits a singular limit when $\beta \leftarrow O$. This result
269: is, however, established for a very special situation which does 
270: not arise here \cite{Legras:02a}. 
271: 
272: \subsection{Kinks and antikinks}
273: 
274: The pure Cahn--Hilliard equation, is recovered from (\ref{vbetach}) by
275: setting $\beta=r=S=\gamma=0$. It admits a Lyapunov functional and is
276: therefore integrable \cite{Sivashinsky:85}.  
277: This property is preserved in the presence of
278: friction but is lost in the presence of $\beta$ 
279: \footnote{A Lyapunov formulation for the approximated equation is
280:   recovered, however, in the limit of large $\beta$ \cite{Frisch:96}.}. 
281: The solutions to the
282: pure Cahn--Hilliard equation live essentially, albeit some initial
283: transients, within a slow manifold of soliton-like solutions with an
284: alternation of plateaus $v=\pm \Gamma$, separated by alternating
285: positive and negative kinks, that we will call respectively kinks and
286: antikinks \cite{Bates:95} in the following. For large enough
287: separation between adjacent kinks, the kink centered in $x=x_j$ is
288: locally given by
289: \begin{equation}
290:    M_j(x)= \epsilon_j M(x-x_j)=\epsilon_j \Gamma \tanh s (x-x_j) \,,
291:    \label{kink}
292: \end{equation}
293: where 
294: $\epsilon _j = 1$ for a kink and $\epsilon _j = -1$ 
295: for an antikink \cite{Kawasaki:82}.
296: This solution satisfies the equation
297: \[
298:    -\pxx M_j + W'(M_j) = 0 \, .
299: \]
300: 
301: Within a $x$ periodic domain of period $L$, the Cahn--Hilliard equation 
302: exhibits
303: stationary metastable solutions of period $\Lambda=L/N$ with $N$ pairs
304: of alternating and equally spaced kinks and antikinks.  These fixed
305: points are unstable saddle points of the Lyapunov functional, except
306: for $N=1$ which corresponds to an absolute stable minimum.  The
307: temporal evolution characterized by a growing total energy is a
308: cascade of annihilations of kink-antikink pairs 
309: (see Fig.~\protect\ref{f:ANKINK}) leading
310: eventually to the gravest mode $N=1$ \cite{Kawasaki:82}.
311: It is shown in Appendix~\ref{aCH} that the local solution 
312: (\ref{kink}) is modified by terms of order $\exp(-s\Lambda)$ when
313: the periodicity is taken into account.
314: 
315: \begin{figure}
316:   \iffigs
317:     \centering
318:     \includegraphics*[width=10cm]{ANKINK.eps}
319:   \else
320:     \drawing 100 10 {Kink-antikink annihilation}
321:   \fi
322:   \caption {Kink-antikink annihilation in a numerical simulation
323:   of the pure Cahn-Hilliard equation with $L$=76.95 .}
324:   \protect\label{f:ANKINK}
325: \end{figure}
326: 
327: %% SECTION III
328: \section{Perturbation of stationary solution under the
329: action of $\beta$ and friction}
330: \label{s:persta}
331: 
332: In order to study the modification of the stationary solutions to the
333: pure Cahn-Hilliard equation under the effect of a small $\beta$ or friction,
334: we multiply both terms by a small control parameter, $\varepsilon$, 
335: keeping $\beta$ and $r$ as O(1).
336: 
337: We put $\gamma = 0$, leaving the case $\gamma \ne 0$, which is 
338: technically much more involved, for Appendix~\ref{s:meanadv}.
339: We only notice here that $\gamma > 0$ breaks one symmetry and generates 
340: narrow intense westerlies and broad narrow easterlies \cite{Manfroi:99}.
341:   
342: It is convenient to integrate (\ref{vbetach}) twice in $x$ to obtain
343: \begin{equation}
344:    - \frac{1}{\lc} \pxxm (1 - S^2 \pxxm)
345:    \pt v - \varepsilon \frac{\beta}{\lc} \pcm v 
346:    - \varepsilon \frac{r}{\lc} \pxxm  v = \pxx v - W'(v) + h(t) \,,
347:    \label{bCH} 
348: \end{equation}
349: where $h(t)$ arises from integration in $x$.  Then the perturbed
350: solution is defined as $\vvv=\vv0 +\varepsilon \vv1 +
351: \varepsilon^2\vv2 + O(\varepsilon^3)$, where $\vv0$ satisfies the
352: stationary CH equation $-\pxx \vv0 + U'(\vv0)=0$.  We will distinguish
353: two cases for $S$: (i) $S=0$, associated with synoptic and subsynoptic
354: dynamics, and (ii) $S=O(1)$, associated with planetary motion (cf.
355: \cite{Pedlosky:87}).  When $\beta \ne 0$, we introduce the phase velocity
356: $c = \varepsilon \c1 + \varepsilon^2 \c2 + O(\varepsilon^{3})$ of the
357: traveling framework in which $\vvv$ is stationary.
358: 
359: %SUBSECTION III 1
360: \subsection{Order 1 perturbation}
361: \medskip
362: 
363: We first treat the case $S=0$.
364: The first order perturbation $\vv1$ satisfies the linear equation 
365: \begin{equation}
366:    \mathcal{F}\pr{\vv1} = \Q0      
367:    \label{baseper}
368: \end{equation}
369: with
370: \begin{align*}
371:    \mathcal{F}\pr{g}  &= \pxx g - W''_{0} g
372:    \, , \\
373:    W''_{0} &= W'' (\vv0) \, , \\
374:    \Q0 &= \frac{1}{\lc} (\c1 \pxm \vv0 - 
375:    \beta \pcm \vv0 - r \pxxm \vv0 ) 
376:    \, .
377: \end{align*}
378:  
379: $\px \vv0$ belongs to the kernel 
380: of $\mathcal{F}$: this can be easily verified 
381: by multiplying (\ref{baseper}) by $\px \vv0$ and integrating within 
382: the domain. 
383: After integration over the spatial period,
384: the  solvability condition for (\ref{baseper}) gives the first order 
385: contribution to the phase velocity
386: \begin{equation}
387:    \c1 = - \beta \frac{\int_0^L \pr{\pxm \vv0}^2 dx}
388:    {\int_0^L \pr{\vv0}^2 dx} \, .
389:    \label{speed} 
390: \end{equation}
391: 
392: The symmetry of the stationary Cahn--Hilliard equation with respect to $x$
393: reversal is inherited by $\vv0$. The solution is
394: antisymmetric with respect to kinks locations and symmetric with
395: respect to the middles of the plateaus. The perturbation $\vv1$ is
396: the sum of two parts
397: \[   \vv1 = \beta \vub + r \vur  \,, \]
398: where $\vur$ has the same symmetry as $\vv0$ and $\vub$ has opposite symmetry. 
399: 
400: It is shown in Appendix~\ref{aCH} that, up to errors of
401: $O(e^{-s\Lambda/2})$ the basic solution $\vv0$ can be approximated by a
402: series of jumps locally described by (\ref{kink}). Over each interval
403: $[x_j-\Lambda/4,x_j+\Lambda/4]$ centered on a kink in $x_j$ (see
404: (\ref{kink})), we have
405: \[ 
406:    \pxm \vv0 =  \epsilon_j \Gamma 
407:    \ln \pr{\frac{\cosh s x}{\cosh \f14 s \Lambda}} + O(e^{-s \Lambda/2})\,, 
408: \]
409: from which the velocity $\c1$ is readily calculated using (\ref{speed}).
410: After some algebra, we get
411: \begin{equation}
412:    \c1 = - \beta \pr{\frac{\Lambda^2}{48} - \frac{\pi^2}{12 s^2} +
413:    \frac{A + \f32 \zeta(3)}{\Lambda s^3}}\pr{
414:    1 - \frac{4}{\Lambda s}}^{-1} + O(e^{-s \Lambda/2}) \,,
415:    \label{c1}
416: \end{equation}
417: where $\zeta$ is the Riemann zeta function and $A=\int_0^\infty
418: \ln^2 (1 + e^{2 x}) dx = 0.150257 \cdots$.  The phase velocity is
419: directed to the left when $\beta>0$ and increases with the wavelength
420: $\Lambda$. The fact that the error in (\ref{c1}) is exponentially
421: small makes this expression very accurate even for not so large values
422: of $\Lambda$ as we shall check below.  By straightforward algebra, one
423: gets 
424: \begin{multline} 
425:   \Q0 = - \frac{\beta \Gamma}{2 \lc} |x| \pr{\frac{x^2}{3}
426:   -\f14 \Lambda |x| + \frac{1}{24} \Lambda^{2}} \\
427:   - \frac{r \Gamma}{2 \lc} 
428:   |x| \pr{ |x| - \f12 \Lambda } + O(\beta \Lambda^2 + r \Lambda)
429:   \,.  
430: \end{multline}
431: At distance from the kinks the ratio between the derivative and the
432: potential term in $\mathcal{F}$ is $O(1/\Lambda^2)$.  At leading order
433: in $1/\Lambda$:
434: \begin{align}
435:    \vub &= \frac{\Gamma}{24 s^2 \lc} |x| \pr{|x| - \f12 \Lambda}
436:    \pr{|x| - \f14 \Lambda} \, ,  \label{vub} \\
437:    \vur &= \frac{\Gamma}{8 s^2 \lc} |x| \pr{|x| - \f12 \Lambda}
438:    \, . \label{vur}
439: \end{align} 
440: These expressions are valid at $O(\Lambda)$ distance from the kinks.
441: It can be shown that at $O(1)$ distance from the kinks, $\vub$ is
442: $O(\Lambda^2)$, while $\vur$ is $O(\Lambda)$.
443: 
444: For $S=O(1)$, the leading contribution of friction is
445: unchanged, but the effect of $\beta$ is deeply affected. $\Q0$ is
446: modified as
447: \[
448:    Q^{(0)}_S = \frac{1}{\lc} \left(\c1 \pxm \vv0 - 
449:    (\beta + \c1 S^2) \pcm \vv0 - r \pxxm \vv0 \right) \, .
450: \]
451: Therefore, the phase speed is now 
452: \[
453:    \c1_S = - \frac{\beta \c1}{S^2 \c1 + \beta},
454: \]
455: where $\c1$ is given by (\ref{c1}), that is 
456: \begin{equation}
457:   \c1_S =  -\frac{\beta}{S^{2}} + \frac{48 \beta}{\Lambda^2 S^4} + 
458:   O\pr{\frac{\beta}{\Lambda^{3}}} \, .
459:   \label{c1S}
460: \end{equation}
461: Unlike the infinite radius case, the phase speed varies weakly 
462: with $\Lambda$. Reporting (\ref{c1S}) in $Q^{(0)}_S$ 
463: and solving for $\vub$, we obtain
464: \begin{equation}
465:    \vub = \frac{\Gamma}{s^2 S^2 \Lambda \lc} |x| \pr{|x| - \f14 \Lambda}
466:    \pr{1 - 2 \frac{|x|}{\Lambda}} + O(1)
467:    \label{vubS}
468: \end{equation}
469: at distance from the kinks. The correction to the stationary solution
470: scales as $\Lambda$ and is considerably reduced with respect to the 
471: case $S=0$, for which it scales as $\Lambda^3$.
472: 
473: %% SECTION IV
474: \section{Stability}
475: \label{s:stab}
476: 
477: \subsection{Stability of the Cahn--Hilliard equation}
478: \label{s:CHstab}
479: 
480: The stability of the solution $v=0$ is easily obtained by linearizing
481: (\ref{vbetach}) in the Fourier domain.  When $r=0$, the solution is
482: unstable to all Fourier modes with wavenumber $0<k<k_m=\sqrt{2} s =
483: \sqrt{2/3}$. This result does not depend on the values of $\beta$ and
484: $S$. When $r>0$ the modes near $k=0$ and $k_m$ are stabilized and the
485: instability band in $k$-space shrinks as $r$ grows. The solution $v=0$
486: is stable for $r> r_0 = \lc s^4 = (3 \sqrt{2})^{-1}$.  The vicinity of
487: this value has been studied in \cite{Manfroi:99}.  We are here
488: interested in the limit of small $r$.
489: 
490: The stability of the non-zero stationary solutions to the Cahn--Hilliard
491: equation can be studied using the equation for kink motion derived in
492: Appendix~\ref{s:CH}. For an arbitrary perturbation, fast transients
493: dissipate rapidly, leaving only after a short time the part of the
494: perturbation that projects onto kink displacement.  The $j$th kink
495: being displaced by $\delta x_j$, the perturbation to $\vv0$ is
496: \begin{equation} 
497:    \delta v = - \sum_{\ell=0}^{2N-1} \px
498:    M_{\ell}\, \delta x_{\ell} \, .
499:    \label{deltav}
500: \end{equation}
501: 
502: Then using (\ref{motok}), the equation for the displacements is 
503: \begin{multline}
504:     -\frac{4 \Gamma^2}{\lc} \sum_{\ell=0}^{2N-1}
505:     \pr{(-1)^{j-\ell}\Gr2(x_j-x_{\ell})+ (-1)^{j-\ell} \frac{\pi^2}{12 L s^2}
506:     - \frac{1}{2 s}\delta_{j-\ell}} \delta \dot{x}_{\ell} \\
507:     = 64 s^3 \Gamma^2 e^{-s \Lambda} [ 2 \delta x_j - \delta x_{j+1}
508:     - \delta x_{j-1}] - 2 \Gamma \epsilon_j \delta h \,.
509:     \label{eq:pertu}
510: \end{multline}
511: 
512: It is convenient to use Fourier components defined as
513: \[ \delta x_j = \sum_{m=0}^{2N-1} \psi_m e^{i \pi\frac{mj}{N}} \,,\]
514: with $m \in [0,2N-1]$ and $\psi_{2N-m}= \overline{\psi_m}$.  The
515: equation for $\psi_m$ is obtained by multiplying (\ref{eq:pertu}) by
516: $\frac{1}{2N} e^{-i \pi\frac{mj}{N}}$ and summing over the $j$ from
517: $0$ to $2N-1$. Taking into account the regular alternation of kinks and
518: antikinks separated by intervals of length $L/2N$ in the basic
519: solution and using the Fourier transform of the Green function given
520: by (\ref{sg}), one obtains after some algebra:
521: \begin{multline}
522:    \label{eq:psim}
523:    \frac{1}{\lc}\pr{\frac{\Lambda}{1+\cos \theta_m} - \frac{2}{s}}
524:    \dot{\psi}_m
525:    = 128 s^3  e^{-s \Lambda} (1 - \cos \theta_m) \psi_m \\
526:    - 4\frac{N}{\Gamma} \delta h \; \delta(N-m) \, ,
527: \end{multline}
528: with $\theta_{m}= \pi m /N$.  The leading order form of
529: (\ref{eq:psim}) was given by Kawasaki and Ohta~\cite{Kawasaki:82} with a
530: factor 2 error (see Appendix \ref{s:CH}).
531: 
532: The first term in the right hand side of (\ref{eq:psim}) is
533: destabilizing the stationary solution with the eigenvalue
534: \begin{equation}
535:    \sigma_{0} = \frac{128 s^{3} \lc e^{-s \Lambda}}{\Lambda} \sin^2 
536:    \theta_m \pr{1 - \frac{2(1 + \cos \theta_m)}{s \Lambda}}^{-1} \, .
537:    \label{s0}
538: \end{equation}
539: This instability is responsible of the inverse cascade in the 
540: CH equation. Each value of $m$ is associated with a real eigenvalue of 
541: $\mathcal{F}$ and a dimension 2 eigenspace. It turns out that an appropriate 
542: basis of this eigenspace is provided by the couple of orthogonal vectors
543: \begin{align}
544:    \va (x) &= \sum_{j=0}^{2N-1} (-1)^j \cos j \theta_m \px M 
545:    (x-x_{j})  \, ,
546:    \label{defva} \\
547:    \vb (x) &= \sum_{j=0}^{2N-1} (-1)^j \sin j \theta_m \px M 
548:    (x-x_{j}) \, ,
549:    \label{defvb}
550: \end{align}
551: which are here given up to an error $O(e^{-s \Lambda/2})$. 
552: These expressions agree very well with the numerical solution shown in
553: figure~\ref{fig:vavb}.
554: 
555: \begin{figure}
556:   \iffigs
557:      \centering
558:      \includegraphics*[angle=-90,width=6cm]{statiog-4.1-a.epsi}
559:      \hspace{1.cm}
560:      \includegraphics*[angle=-90,width=6cm]{statiog-4.1-b.epsi}
561:   \else
562:      \drawing 100 10 {Eigenvectors va and vb}
563:   \fi
564:   \caption{Numerical calculation of $v_a$ and $v_b$ by discretization of 
565:   the eigenvalue problem for $\mathcal{F}$ (cf Section
566:   \ref{s:numpert}) with 128 points for $L=76.953$ and $m=4$.  (a) thin
567:   solid: $\vv0$; solid: $v_a$, dash: $v_b$; (b) thin solid: $\px
568:   \vv0$; solid: $v_a/\px \vv0$, dash: $v_b/\px \vv0$.  The scale is
569:   arbitrary for $v_a$ and $v_b$.  Even if the plateaus in $\vv0$ are
570:   very short for the chosen values of $L$ and $m$, the ratios in (b)
571:   show the staircase structure of $v_a$ and $v_b$ over the kinks as
572:   given in \protect{(\ref{defva}, \ref{defvb})} except where $\px
573:   \vv0$ vanishes in the middle of the plateaus.
574:   \label{fig:vavb}}
575: \end{figure}
576: 
577: The $\delta h$ contribution vanishes but on the mode $m=N$. The 
578: solution $v(x,t)$ must average to zero within the periodic interval
579: $[0,L]$. In terms of kink motion, this imposes the constraint
580: \[ \sum_{\ell=0}^{2N-1} (-1)^{\ell} \delta \dot{x}_{\ell} = 0 \,, \]
581: and thus, $\dot{\psi}_{N} = 0$.
582: The presence of $\delta h(t)$ in (\ref{eq:psim}) is required to impose 
583: this condition. 
584: 
585: \subsection{Stability of $\beta$-CH equation with infinite radius of 
586: deformation}
587: 
588: \subsubsection{Formulation of the problem}
589: The equation governing the perturbation $\delta v$ to $\vv0$ can be 
590: written conveniently for $\varphi = \pxm \delta v$
591: \begin{equation}
592:    \pt \varphi = \mathcal{L} \varphi \, ,
593:    \label{eq:stab}
594: \end{equation}
595: with 
596: \begin{equation}
597:    \mathcal{L}  =   - \lc \px (\pxx - W''(\overline{v})) \px 
598:    - c \px - \varepsilon \beta \pxm  - \varepsilon r \, .
599:    \label{L:def}
600: \end{equation}
601: The reason of using $\varphi$ instead of $\delta v$ is that $\mathcal{L}$ is 
602: auto-adjoint while the corresponding operator for $\delta v$ is not.
603: 
604: Unlike the case $\beta = 0$, the slow component perturbation to the
605: stationary solution of the $\beta$-CH equation does not reduce to the
606: simple motion of kinks. One has also to take into account the
607: dispersive effect of the $\beta$ term modifying the shape of the slow
608: modes and contributing to the stability.  Therefore, we expand
609: $\mathcal{L}$ as
610: \begin{equation}
611:    \mathcal{L} = \Lr0 + \varepsilon \Lr1 + \varepsilon^2 \Lr2 
612:    + O(\varepsilon^3) 
613:     \, , 
614: \end{equation}
615: with
616: \begin{align}
617:    \Lr0 &=  - \lc \px ( \pxx - U''_{0}) \px \, , 
618:    \label{L0:def} \\
619:    \Lr1 &=  \lc \px (W'''_{0} \vv1 \px) + \c1 \px - \beta \pxm - r
620:    \, , \label{L1:def}  \\
621:    \Lr2 &=  \lc \px (W'''_{0} \vv2 + \f12 W^{IV}_{0} \vv1^2 ) \px + 
622:    \c2 \px \, .
623:    \label{L2:def}
624: \end{align}
625: 
626: The eigenvalues of (\ref{eq:stab}) are perturbations of the 
627: eigenvalues of (\ref{eq:pertu}). For a given $m \neq N$, we obtain
628: \begin{equation}
629:    \sigma = \sigma_{0} + \varepsilon \sigma _{1} + i \varepsilon 
630:    \mu_{1} + \varepsilon \sigma_{2} + i \varepsilon \mu_{2} 
631:    + O(\varepsilon^{3}) \, .
632: \end{equation}
633: The functions $\vfa = \pxm \va{} $ and $\vfb = \pxm \vb{} $ are
634: orthogonal eigenmodes of $\mathcal{L}_0$ and, it turns out, an
635: appropriate Jordan basis for the perturbation problem. They are
636: respectively modified as $\vfa + \varepsilon \vpa1 + \varepsilon^2
637: \vpa2 + O(\varepsilon^3)$ and $\vfb + \varepsilon \vpb1 +
638: \varepsilon^2 \vpb2 + O(\varepsilon^3)$. The hierarchy of linear
639: problems is
640: \begin{align}
641:    \Lr0 \vfa & = \sigma_0 \vfa \, ,
642:    \label{eq:L0}  \\
643:    \Lr0 \vpa1 + \Lr1 \vfa &=  \sigma_0 \vpa1 + \sigma_1 
644:    \vpa - \mu_{1} \vfb \, ,
645:    \label{eq:L1}  \\
646:    \Lr0 \vpa2+ \Lr1 \vpa1 + \Lr2 \vfa &= \sigma_0 \vpa2
647:    + \sigma_{1} \vpa1 + \sigma_{2} \vfa - \mu_{1} \vpb1 - \mu_2     
648:    \vfb \, .
649:    \label{eq:L2}
650: \end{align}
651: and similar equations for $\vpb{i}$. 
652: 
653: \subsubsection{Stabilization by friction at first order}
654: 
655: The first order corrections of the eigenvalue $\sigma$ are obtained as 
656: solvability conditions of (\ref{eq:L1}) by
657: \begin{align}
658:    <\vfa,\Lr1 \vfa> &=  \sigma_{1}  <\vfa,\vfa> \, ,
659:    \label{eq:s1}  \\
660:    <\vfb,\Lr1 \vfa> &= - \mu_{1}  <\vfb,\vfb> \, ,
661:    \label{eq:m1}
662: \end{align}
663: where $<f,g> \equiv \frac{1}{L}\int_{0}^{L} f(x) g(x) dx$.
664: In $<\vfa,\Lr1 \vfa>$, the contributions from $\beta$ immediately 
665: vanish by integration. We have, after integration by part
666: \begin{equation} 
667:    <\vfa,\Lr1 \vfa> =  -\lc < \va{}^2, W'''_0 \vv1 > 
668:    + r <\va{}, \pxxm \va{}> 
669:    \, . \label{paL1pa}
670: \end{equation}
671: The first contribution to the right hand side of (\ref{paL1pa}) can be 
672: reduced to an integral over a single kink by summing the trigonometric 
673: factors arising from $\va{}^2$:
674: \begin{equation} 
675:    < \va{}^2, W'''_{0} \vv1 > = \frac{r}{\Lambda} \int W'''_{0} 
676:    \vur (\px M_{j})^2 dx \, .
677: \end{equation}
678: Here $j$ labels an arbitrary kink and the integral bounds do not need 
679: to be specified since $\px M_{j}$ decays exponentially on both side. 
680: This contribution is further transformed using
681: \[
682:    \int W'''_{0} \vur (\px M_j)^2 dx =   
683:    \frac{1}{\lc} \int \pxm \vv0 \px  M_j dx \, ,
684: \]
685: which is valid up to exponentially small errors. Finally the 
686: contribution is reduced to a non local integral by part,
687: \[
688:    \frac{1}{\Lambda} \int \pxm \vv0 \px  M_j dx = \frac{1}{2 \Lambda} 
689:    \int_{x_{j}- \f12 \Lambda}^{x_{j}+ \f12 \Lambda} \px \vv0 \pxm \vv0 
690:    dx =  -\f12 \Gamma^2 + \frac{2 \Gamma^{2}}{\Lambda s}\, . 
691: \]
692: Combining this with (\ref{vanva}) and (\ref{sg}), we obtain
693: \begin{multline} 
694:   \sigma_{1} = -r \sin^2 \f12 \theta_{m} \\
695:   + \frac{4r}{\Lambda s} \left(1 + \cos^{2} \f12 \theta_{m}\right)
696:   \left(1 - \frac{4}{\Lambda s } \cos^{2} \f12 \theta_{m}\right)^{-1}
697:   + O(e^{-s \Lambda/2}) \, .
698:    \label{s1}
699: \end{multline}
700: 
701: In a similar way, we have
702: \begin{multline}
703:    <\vfb,\Lr1 \vfa> = - \lc <\va{} \vb{} , W'''_0 \vv1> \\
704:    + \c1  <\va{},\pxm \vb{}> - \beta <\va{} , \pcm \vb{}> \, .
705:    \label{pbL1pa}
706: \end{multline}
707: The first term in the right hand side of (\ref{pbL1pa}) vanishes after
708: the trigonometric summation. Using (\ref{vanva}-\ref{vanvb}) and
709: (\ref{sf}-\ref{sh}), we obtain
710: \begin{multline}
711:   \mu_1 = \left(-\frac{2 \c1 \Gamma^2}{\Lambda} t - 
712:   \frac{\Gamma^2 \Lambda}{8} t(1+t^2)
713:   + \frac{\pi^2 \Gamma^2}{6 s^2 \Lambda} t \right)\\
714:   \times \left( \frac{\Gamma^2}{2} (1+t^2) - \frac{2 \Gamma^2}
715:   {s\Lambda } \right)^{-1} + O(e^{-s \Lambda/2}) \, ,
716:   \label{m1a} 
717: \end{multline}
718: or, retaining only the first three orders of the 
719: expansion,
720: \begin{equation}
721:    \mu_{1} = \beta \pr{- \frac{\Lambda t (2 + 3 t^2)}{12 (1+ t^2)} 
722:    - \frac{t (1+2t^{2})}{3s(1+t^{2})^{2}}+ \frac{4}{3\Lambda s^{2}}
723:    \frac{t^{5}}{(1+t^{2})3}} +  O \pr{\frac{\beta}{\Lambda^{2}}}\, ,
724:    \label{m1b}
725: \end{equation}
726: with $t= \tan \pi m/ 2 N$. 
727: 
728: It is interesting to notice that the nonlinear contribution $<
729: \va{}^2, U'''_{0} \vv1 >$ is destabilizing the stationary solution.
730: However, the direct linear damping by friction remains larger and the
731: total effect of friction is always stabilizing. Therefore, the
732: $m$-mode perturbation to the stationary solution is stabilized by
733: friction for
734: \begin{equation} 
735:    r > r_{c} = 512 \frac{e^{-s \Lambda}}{\Lambda} s^3
736:    \lc \cos^2 \frac{\pi m}{2 N} \, ,
737:    \label{critr} 
738: \end{equation}
739: at leading order.
740: 
741: The corresponding result when $\gamma \neq 0$ is given in 
742: (\ref{eq:rcadv}). The mean advection decreases the value
743: of friction necessary to stabilize a given wavenumber.
744: 
745: \subsubsection{Stabilization by $\beta$-effect at second order}
746: 
747: At second order in $\varepsilon$, $\sigma_{2}$ is solution of the 
748: solvability condition. Here we set $r$ to zero  for simplification
749: as stabilization by $r$ is already obtained at first order.
750: \begin{equation}
751:    <\vfa, \Lr1 \vpa1 > + < \vfa, \Lr2 \vfa >\; = \sigma_{2} <\vfa, 
752:    \vfa > - \mu_{1} <\vfa,\vpb1>  \, .
753:    \label{eq:s2}
754: \end{equation}
755: The second term on the left hand side of (\ref{eq:s2}) expands as
756: \begin{equation}
757:    < \vfa, \Lr2 \vfa >\; = 
758:    - \lc < \va{}^2, W'''_0 \vv2 > 
759:    - \f12 \lc <\va{}^2, W^{IV}_0 (\vv1)^2 >  \, .
760:    \label{faL2fa}
761: \end{equation}
762: The second term is $O(\beta^2 \Lambda^{3})$ while the first term is
763: $O(\beta^2 \Lambda^{4})$ and dominates at leading order.  This term
764: can be transformed in the same way as above for $<\vfa,\Lr1 \vfa>$,
765: leading to
766: \[
767:    < \vfa, \Lr2 \vfa > = - \frac{\lc}{\Lambda} \int \Q1 \pxx M_{j} 
768:    dx + O(\beta^2 \Lambda^{3}) \, ,
769: \]
770: where $\Q1$ is the right hand side for the second order version of
771: (\ref{baseper}).
772: After dropping out all contributions that vanish owing to 
773: the symmetries, we obtain
774: \begin{multline}
775:    < \vfa, \Lr2 \vfa > = \frac{\beta^2}{\Lambda} \int \pcm \vub \pxx 
776:    M_{j} dx  \\
777:    +\frac{\lc}{2 \Lambda} \int W'''_{0} (\vv1)^2 \pxx M_{j } dx 
778:    + O(\beta^2 \Lambda^{3}) \, .
779:    \label{faL2fa2}
780: \end{multline}
781: The second term on the right hand side of (\ref{faL2fa2}) is negligible  
782: relatively to the first. This latter can be calculated using 
783: \[
784:    \frac{1}{\Lambda} \int \pcm \vub \pxx M_{j} dx = 
785:    \frac{1}{2 \Lambda} \int_{x_{j}-\f12 \Lambda}^{x_{j}+\f12 \Lambda} 
786:    \pcm \vub \pxx \vv0 dx = 
787:    \frac{1}{2 \Lambda} \int_{x_{j}-\f12 \Lambda}^{x_{j}+\f12 \Lambda}
788:    \pxm \vub \vv0  \, 
789: \]
790: and (\ref{vub}). We obtain
791: \begin{equation} 
792:    < \vfa, \Lr2 \vfa > = \frac{\beta^2 \Lambda^4 \Gamma ^2}{92,\!160\; s^2 
793:    \lc}  + O(\beta^2 \Lambda^3) \, .
794:    \label{faL2fa3}
795: \end{equation}
796: 
797: The other contributions in (\ref{eq:s2}) involve $\vpa1$ and $\vpb1$. 
798: These quantities can be approximated in the same way as $\vv1$ at 
799: distance from the kinks. We have
800: \begin{equation}
801:    4 s^2 \lc \vpa1 = - \lc \pxm (W'''_{0} \vv1 \va{}) - \c1 \pxm \vfa 
802:    + \beta \pcm \vfa - \mu_{1} \pxxm \vfb \, ,
803:    \label{vpa1}
804: \end{equation}
805: and a similar expression for $\vpb1$. Then,
806: after some algebra and dropping the terms which do not 
807: contribute to the leading order, we have
808: \begin{multline*}
809:    4 s^{2 } \lc <\vfa, \Lr1 \vpa1> = \c1^2 <\va{}, \pxxm \va{}> 
810:    + \beta^2 <\va{}, \pfm \va{}> \\ - 2 \beta \c1 <\va{}, \pdm \va{}>
811:    - \mu_{1} \beta <\va{}, \pem \vb{}> + \mu_{1} \c1 <\va{}, \pcm \vb{}>
812:    + O(\beta^2 \Lambda^{3}) \, .
813: \end{multline*}
814: Using the results of Appendix~\ref{s:Green}, we obtain
815: \[
816:    <\vfa, \Lr1 \vpa1> = - \frac{\beta^2 \Gamma^2 \Lambda^4}{ 92,\!160\ s^2
817:    \lc} (1 + 6 t^2 + 15 t^4) + O( \beta^2 \Lambda^{3})\, .
818: \]
819: and
820: \[
821:    < \vfa, \vpb1 > = - \frac{\beta \Gamma^2 \Lambda^3}{4,\!608\;  s^2 \lc} 
822:    (t + 3 t^3) + O(\beta \Lambda^2) \, .
823: \]
824: 
825: Finally
826: \begin{equation}
827:    \sigma_2 = - \frac{\beta^2 \Lambda^4}{69,\!120\; s^2 \lc}
828:    \frac{t^2(4+9t^2)}{(1+t^2)^2} + O(\beta^2 \Lambda^{3}) \, .
829:    \label{s2r}
830: \end{equation}
831: The contribution from $r^{2}$ is a correction to the first order 
832: stabilization obtained in $\sigma_{1}$. The $\beta$ term does not 
833: appear at first order and is stabilizing in (\ref{s2r}). Though the 
834: effect is small, it increases algebraically with $\Lambda$ while the 
835: non linear coupling of kinks decreases exponentially in  (\ref{s0}).
836: Therefore, if $r=0$, stabilization of the $m$-mode perturbation to the 
837: stationary solution is obtained at leading order for 
838: \begin{equation}
839:    \beta > \beta_{c} = \left( 35,\!389,\!440\, 
840:    \frac{e^{-s \Lambda}}{\Lambda^5} s^5 
841:    \lc^2 \frac{1 }{4 + 9 t^2} \right)^{1/2} \, .
842:    \label{betac}
843: \end{equation}
844: The condition is the most restrictive for $m=1$, that is $t= \tan 
845: \pi/(2N)$.
846: 
847: \subsection{Stability of $\beta$-CH equation with finite radius of 
848: deformation}
849: 
850: In this case $\Lr1$ and $\Lr2$ are modified as 
851: \begin{align}
852:    \Lr1 &= \lc \px W'''_{0} \vv1 \px + \c1 \px - 
853:    (\beta + \c1 S^{2})\pxm - r
854:    \, , \label{L1S:def}  \\
855:    \Lr2 &=  \lc \px (W'''_{0} \vv2 + \f12 U^{IV}_{0} \vv1^2 ) \px + 
856:    \c2 (\px - S^2 \pxm) \, .
857:    \label{L2S:def}
858: \end{align}
859: 
860: At first order in $\varepsilon$, the imaginary part of the eigenvalue is
861: \begin{equation}\begin{split}
862:    \mu_{1 } &= -\c1\frac{<\va{}, \pxm \vb{}>}{<vfb, \pxxm \vfb} +
863:    (\beta + S^2 \c1) \frac{<\va{}, \pcm \vb{}>}{<vfb, \pxxm \vfb}
864:    \\
865:    &= - \frac{4 \beta t (2 + 3 t^2)}{\Lambda S^2 (1 + t^2)} + 
866:    O\pr{\frac{\beta}{\Lambda^2}}
867: \end{split}\end{equation}
868: 
869: At second order in $\varepsilon$, $\sigma_{2}$ is still given by 
870: (\ref{eq:s2}). The contribution $<\vfa, \Lr2 \vfa>$ is now 
871: \[
872:    <\vfa, \Lr2 \vfa> = \frac{\beta (\beta + \c1 S^2)}{\Lambda} \int \pcm 
873:    \vub \pxx M_{j} dx + O\pr{\frac{\beta^2}{\Lambda}} \, .
874: \]
875: After integration by part and using (\ref{vubS}), we obtain 
876: \[ 
877:    <\vfa, \Lr2 \vfa> = - \frac{\beta^2 \Gamma^2}{40\; s^2 S^4 \lc} 
878:    + \pr{\frac{\beta^2}{\Lambda}} \, .
879: \]
880: Similarly, we have
881: \begin{align*}
882:    <\vfa, \Lr1 \vpa1 > &= - \frac{\beta^2 \Gamma^2}{40\; s^2 S^4 \lc}
883:    (1+6 t^2 + 15 t^4) 
884:    + O\pr{\frac{\beta^{2}}{\Lambda}} \, , \\
885:    <\vfa,\vpb1> &= - \frac{\Gamma^2 \Lambda}{96\; s^6 S^{2} \lc} t(1+3t^2)
886:    +  O(1) \, .
887: \end{align*}
888: 
889: Finally, we obtain $\sigma_{2}$ as
890: \begin{equation} 
891:    \sigma_{2} = - \frac{\beta^2}{30\; s^2 S^4 \lc}
892:    \frac{3 + 7 t^2 + 9 t^4}{(1 + t^2)^2} 
893:    + O\pr{\frac{\beta^2}{\Lambda}} \, .
894:    \label{s2S}
895: \end{equation}
896: 
897: The stability crossover for $\beta$ is now
898: \begin{equation}
899:    \beta_{c} = \left( 15,\!360\, s^{5} S^4 \lc^{2} \frac{3 + 7 t^2 + 9 
900:    t^4}{t^2}\frac{e^{-s \Lambda}}{\Lambda} \right)^{1/2} \, .
901: \end{equation}
902: Therefore, the stabilizing effect of $\beta$ is much reduced compared 
903: to that with infinite radius of deformation.
904: 
905: All the perturbative calculations of Sections \ref{s:persta}
906: and \ref{s:stab} have been checked with Mathematica.
907: 
908: %% SECTION V
909: \section{Numerical approach}
910: \label{s:numerical}
911: 
912: The analytic results established in Section~\ref{s:stab} are valid in
913: the double limit of small $\epsilon$ and large $\Lambda$. These
914: asymptotic results are complemented and compared with three types of
915: numerical calculations: (i) numerical solution of the perturbative
916: problem for several values of $\Lambda$, (ii) direct numerical
917: simulation of the Cahn--Hilliard equation in the Fourier space and
918: (iii) direct stability calculation.
919: 
920: \subsection{Numerical solution of the perturbative problem}
921: \label{s:numpert}
922: 
923: We relax here the hypothesis on the large value of $\Lambda$ by
924: solving numerically the perturbative equations for $\vv1$, $\vv2$,
925: $\vpa1$ and $\vpb1$, and numerically evaluating the solvability
926: conditions.
927: 
928: The calculation is performed according to the following algorithm
929: \begin{enumerate}
930: \item The basic solution $\vv0$ is defined by the approximate form
931:   given in appendix~\ref{aCH}.
932: \item The inverse derivatives $\pxm \vv0$, $\pxxm \vv0$ and $\pcm
933:   \vv0$ are calculated by Fourier transform and tabulated to obtain
934:   $\Q0$.
935: \item $\c1$ is calculated by discrete evaluation of (\ref{speed}).
936: \item (\ref{baseper}) is discretized as a tridiagonal problem and
937:   solved for $\vub$ and $\vur$.
938: \item $\Q1$ is built in the same way as $\Q0$ from the inverse
939:   derivatives of $\vub$ and $\vur$.
940: \item $\vv2$ is calculated by inverting a tridiagonal discretized
941:   problem.
942: \item The eigenvectors $\va{}$ and $\vb{}$ are defined using (\ref{defva})
943:   and (\ref{defvb}).
944: \item $\vfa$ and $\vfb$ are calculated by Fourier transform and
945:   tabulated.
946: \item $\sigma_1$ and $\mu_1$ are calculated according to
947:   (\ref{eq:s1}), (\ref{eq:m1}) and (\ref{L1:def}).
948: \item $\vpa1$ and $\vpb1$ are obtained using (\ref{eq:L1}) and solving
949:   a tridiagonal discretized problem.
950: \item $\sigma_2$ is obtained from (\ref{eq:s2}), (\ref{L1:def}) and
951:   ((\ref{L2:def}).
952: \end{enumerate}
953: 
954: This algorithm admits $O(e^{-s \Lambda/2})$ errors, but all the steps
955: generating errors of algebraic order in the asymptotic expansion are
956: here solved numerically. The implementation has been done as a
957: Mathematica notebook available from the authors. The number of grid
958: points and Fourier modes has been adjusted as a function of $L$ and
959: $N$ in order to provide at least 3 digits of accuracy in the results.
960: The symmetries have been exploited to distinguish the solutions and
961: reduce the number of points.
962: 
963: \subsection{Numerical solution of the complete Cahn--Hilliard equation}
964: 
965: We have done numerical simulations of the complete Cahn--Hilliard
966: equation (1) for the case $\gamma=S=0$. For practical convenience, the
967: spatial period has been kept fixed to $2 \pi$ by rescaling $x$ as $x
968: \rightarrow p x$. The complete Cahn-Hilliard equation then reads
969: \begin{equation}
970:    \partial_T v=
971:    \frac{\lambda_1}{3 p^2}\partial^2_x v^3-\frac{\lambda_2}{p^2}
972:    \partial^2_x v - \frac{\lambda_3}{p^4} \partial^4_x v - p \beta 
973:    \partial_x^{-1} v - r v  \, .
974:    \label{rbetach}
975: \end{equation}
976: Since Fourier modes are discretized by the periodicity condition,
977: the number $n$ of unstable modes for $v=0$ is the integer part of
978: $p(2/3)^{1/2}$.
979: 
980: \subsubsection{Time-dependent simulations}
981: \label{s:tdep}
982: 
983: The simulations are performed using a standard semi-spectral method
984: where the number of retained real Fourier modes is 256. The
985: collocation grid in the spatial domain has 512 points in order to
986: fully remove the aliasing due to the cubic term in (\ref{rbetach}). We
987: have checked that using higher resolution does not modify the results
988: within the explored parameter range. The temporal integration is
989: performed with an Adams-Bashforth second-order scheme.  Initial
990: conditions are a random white noise in the spatial domain. We have
991: made a large number of runs by varying the initial conditions
992: (changing the seed of the pseudo-random number generator and the
993: amplitude of the noise), the values of $r$ and $n$.
994: 
995: Section \ref{s:stab} shows the existence of multiple stable solutions
996: induced by $\beta$ and friction but does not provide indications about
997: the attraction basins of these solutions. In the inverse cascade of
998: the standard Cahn-Hilliard equation, the interaction of a pair of
999: neighbor kink and antikink scales as $\exp(-s \Delta x)$, where
1000: $\Delta x$ is the distance between the kink and the antikink (see
1001: Appendix B).  When friction $r$ is just above the critical value
1002: $r_c(N)$ stabilizing the solution with $N$ pairs, we may conjecture
1003: that the stabilizing effect is of the order $O((r-r_c(N))\delta x)$,
1004: where $\delta_x$ is the departure of a kink from its equilibrium
1005: position (this is clear from the shape of the eigenmodes $v_a$ and
1006: $v_b$ (cf. ection \ref{s:CHstab})). This effect, however, cannot
1007: extend very far in $\delta x$ as the attraction to the neighbor
1008: antikink grows as $\exp(s \delta x)$.  Moreover, the time-dependent
1009: solutions do not need to pass in the vicinity of the $N$-pair fixed
1010: point during the cascade of kink-antikink annihilations.  Therefore,
1011: we expect that the fraction of the solutions halting on the $N$-pair
1012: stable state will be small in the vicinity of the critical $r_c$ and
1013: will be significant only when the stabilizing effect is felt over a
1014: distance of order $\Lambda$. Once this is obtained, very few solutions
1015: should jump to lower $N$ states.
1016: 
1017: Figure~\ref{f:FRISIM} shows how the inverse cascade evolves as a
1018: function of $r$ for the same initial conditions, with the
1019: corresponding steady states shown in figure~\ref{f:FRISIM_PH}. The
1020: final state wavenumber increases with $r$ and stays bounded by the
1021: value of the most unstable eigenmode of the Kolmogorov flow $k=
1022: (2/3)^{1/2}n$ which is the unique mode excited when $r$ approaches $r_0$.
1023: The total energy calculated as the sum of the various energies E(k) of
1024: the single modes $k$, decreases as the friction increases.
1025: 
1026: Figure~\ref{f:DUOFB} compares the halting of the inverse cascade by
1027: friction and by $\beta$ effect for the same initial conditions.  It is
1028: apparent from figure~\ref{f:FRISIM} that friction simply halts the
1029: cascade by stabilizing one of the intermediate steps.  The effect of
1030: the $\beta$-term is more complex: oscillatory transients are excited
1031: and, paradoxically, the cascade is accelerated. In Fig.~\ref{f:DUOFB}
1032: the transitions to $N=3$ and to $N=2$ occur much earlier than in the
1033: absence of $beta$. Other examples can be found in
1034: \cite{Frisch:95,Legras:99}.  In the final state, the $\beta$ effect
1035: breaks the symmetry between the kinks and the antikinks which is
1036: preserved by friction.
1037: 
1038: \begin{figure}
1039:   \iffigs 
1040:      \centering \includegraphics*[width=10cm]{FRISIMtest.eps}
1041:   \else 
1042:      \drawing 100 10 {Friction simulations} 
1043:   \fi
1044:   \caption {Temporal evolution of the energies $E(k)$ of the Fourier modes
1045:   and of their sum $E_\mathrm{tot}$.  The three shown cases, are for $r=0$,
1046:   $r=3\,10^6$ and $r=10^5$, with $\beta=0$, $n=10$ and the same
1047:   realization of white noise as initial condition.  For $r=0$ the
1048:   inverse cascade is complete to $N=1$; for the other values the
1049:   cascade stops respectively on $N=2$ and on $N=3$. Note that $E_tot$
1050:   is constant between two annihilation events. Increasing further $r$
1051:   enables to stop the cascade on larger $N$ configurations (not
1052:   shown).
1053:   \label{f:FRISIM}}
1054: \end{figure}
1055: \begin{figure}
1056:    \iffigs \centering \includegraphics*[width=10cm]{FRISIM_PHtest.eps}
1057:    \else \drawing 100 10 {Friction ph simulations}
1058:    \fi
1059:    \caption {The corresponding asymptotic velocity profile 
1060:      for the three cases presented in Fig.~\ref{f:FRISIM}.  Note that the
1061:      kinks-antinkinks pairs are equidistant, unlike the $N=2$ or $N=3$
1062:      configurations of Fig.~\ref{f:ANKINK} which are only metastable.
1063:    \label{f:FRISIM_PH}}
1064: \end{figure}
1065: \begin{figure}
1066:   \iffigs \centering \includegraphics*[width=10cm]{DUOFBtesttest.eps}
1067:   \else \drawing 100 10 {Comparison $\beta$ and friction simulations}
1068:   \fi
1069:   \caption {Comparison between two temporal evolution of 
1070:      $E_\mathrm{tot}$ and $E(k)$, for the same initial conditions 
1071:      and for very similar final solutions with same $N$ and final
1072:      energy $E_\mathrm{tot}\approx0.5535$.
1073:      Top: $r=10^{-5}$ and $\beta=0$. Bottom: $r=0$ and $\beta=9.7\,10^{-3}$.
1074:    \label{f:DUOFB}}
1075: \end{figure}
1076: 
1077: \subsubsection{Numerical stability calculations}
1078: \label{s:numstab}
1079: 
1080: In order to compare the results from the analytic and numerical
1081: perturbation approaches to the direct simulation of the complete
1082: Cahn--Hilliard equation, we have developed a direct analysis of
1083: stability by the same technique as for the time-dependent
1084: simulations.  The core of this analysis is to calculate the Jacobian
1085: matrix of the right-hand side of (\ref{rbetach}) linearized around a
1086: given state $v$, with respect to each of the Fourier component.  This
1087: is done by differentiating (\ref{rbetach}) and calculating the columns
1088: of the Jacobian matrix by the semi-spectral method applied to the
1089: differentiated equations for each Fourier component.
1090: 
1091: The stability calculation is performed as follows.  First an estimate
1092: of the stationary solution based on Appendix~\ref{aCH} is refined by a
1093: Newton-Raphson algorithm which usually converges within a few steps.
1094: The degeneracy due to the $x$-invariance of the complete Cahn-Hilliard
1095: equation is removed by setting to zero the imaginary part of the
1096: dominating Fourier mode and removing the corresponding row and line
1097: from the Jacobian matrix. In the case $\beta \ne 0$, the solution is
1098: stationary in a frame traveling at a velocity $c$. This phase velocity
1099: is treated as an additional unknown increasing the dimension of the
1100: Jacobian matrix by one. The stability of this numerical solution is
1101: then found by finding the eigenvalues and the eigenmodes of the
1102: Jacobian matrix using a standard QR algorithm from LINPACK. In this
1103: procedure, we find both the slow components associated to kink
1104: dynamics and the highly damped modes associated with fast relaxing
1105: transients. The conditioning of the eigenvalue problem gets very bad
1106: as $\Lambda$ increases as a result of the large separation between
1107: slow and fast eigenvalues, thus limiting the parameter range for the
1108: numerical calculation of stability. There is enough overlap, though,
1109: with the validity domain of the perturbative theory to
1110: provide detailed comparison. The number of Fourier modes used in this
1111: analysis has been 128, 256 or 384, depending on the values of $L$ and
1112: $N$.
1113: 
1114: \subsection{Comparison of stability results}
1115: 
1116: Table~\ref{t:fricbeta} compares the critical values of friction and
1117: $\beta$ estimated from the analytic perturbative expansion, the
1118: numerical solution to the perturbation problem and the numerical
1119: stability analysis. In the numerical stability analysis the value of
1120: the parameter is adjusted by dichotomy from two values bracketing the
1121: transition.  There is an excellent agreement between the three values
1122: of critical friction when $\Lambda$ is large. At the largest values,
1123: however, the QR algorithm fails to converge and no results are
1124: obtained for the numerical stability.  When $\Lambda$ is not large,
1125: that is when the kinks are not distant enough to neglect the
1126: contribution of $O(e^{-s \Lambda/2})$, significant discrepancies occur
1127: between the different estimates. We see from the table that this
1128: occurs when $e^{-s \Lambda/2} \gtrapprox 3\,10^{-2}$.
1129: 
1130: There is also an excellent agreement between $\beta_c^\mathrm{pert}$
1131: and $\beta_c^\mathrm{num}$ for the same range of $\Lambda$ values as
1132: for $r_c$.  The analytical prediction (\ref{betac}), however, provides
1133: only an order of magnitude and is wrong by at least a factor two when
1134: the two other quantities agree by four digits. The reason is that the
1135: error in (\ref{betac}) depends algebraically on $\Lambda$ unlike the
1136: error in (\ref{critr}) where the error exhibits an exponential
1137: dependence. Very large values of $\Lambda$ are required to make
1138: (\ref{betac}): for $n=200$ and $N=5$, we obtain
1139: $\beta_c^\mathrm{pert}=2.24\,10^{-42}$ and
1140: $\beta_c^\mathrm{num}=2.35\,10^{-42}$, while for $N=2$ we obtain
1141: $\beta_c^\mathrm{pert}=1.842\,10^{-101}$ and
1142: $\beta_c^\mathrm{num}=1.808\,10^{-101}$.  This difficulty with $\beta$
1143: is entirely due to the need to solve completely the perturbation at
1144: order 1. The phase speed $c_1$ and the frequency $\mu_1$, which are
1145: obtained as solvability conditions at order one, are known with the
1146: same accuracy as $r_c$, as can be checked in Table~\ref{t:cmu}.
1147: 
1148: \subsection{Comparison of stability properties and time-dependent solutions}
1149: 
1150: In order to assess the distribution of final states among
1151: the multiple stable steady states we have performed ensemble
1152: simulations for a number of values of $r$ and $\beta$. 
1153: Each numerical integration of the time-dependent numerical model 
1154: described in Section~\ref{s:tdep}
1155: is characterized by $n$, $r$, $\beta$ and the initial condition.
1156: We choose for this latter a white noise with amplitude $A$ in the 
1157: spatial domain. For each value of the parameters and for two
1158: values of the amplitude, $A=0.1$ and $A=1.$, we performed an
1159: ensemble of 100 independent simulations by varying the seed of the random 
1160: number generator. After some time, all simulations 
1161: converge to a final stationary or uniformly traveling state (if $\beta=0$
1162: or $\beta \neq 0$, respectively). The non convergent cases are due
1163: to fast transients leading to nonlinear numerical instabilities.
1164: 
1165: Table~\ref{t:r} shows the dependence on $r$ when $\beta=0$ and $n=20$.
1166: The distribution of final states agree qualitatively with the analysis
1167: presented in Section~\ref{s:tdep}. Stationary states associated to
1168: given value of $N$ are only reached for $r$ larger than the critical
1169: value $r_c(N)$.  The proportion of solutions reaching these stationary
1170: states is about 15\% for $A=0.1$ and 10\% for $A=1.$ when
1171: $r/r_c(N)\approx 3.$ This proportion grows rapidly as $r$ increases
1172: further while the number of states reaching solutions with smaller $N$
1173: falls dramatically. In practice, it is difficult to find values of $r$
1174: where more than 3 steady states are obtained. It is also visible that
1175: larger amplitude of the initial conditions favor smaller final $N$ and
1176: more dispersion of the final states.
1177: 
1178: Table~\ref{t:beta} shows the corresponding results as a function of
1179: $\beta$ when $r=0$. They are qualitatively similar to the precedings
1180: although up to 7 different final states are now observed for
1181: $\beta=1.$
1182: 
1183: \section{Summary and conclusion}
1184: \label{s:conclu}
1185: 
1186: We have investigated the stabilization induced by friction and $\beta$
1187: effect in the inverse cascade of the large-scale instability of the
1188: Kolmogorov flow.  This problem has been treated by solving
1189: the unidimensional complete Cahn-Hilliard equation
1190: using both numerical simulations and perturbation techniques
1191: 
1192: In the standard Cahn-Hilliard equation, the kinks and antikinks are
1193: coupled by interactions decreasing exponentially as function of their
1194: separation. The number of kinks decreases with time and their average
1195: separation increases as a result of the inverse cascade. Therefore the
1196: Cahn-Hilliard coupling also decreases with time.
1197: 
1198: The basic effect of friction is to damp uniformly the motion of the
1199: kinks and antikinks. Nonlinear effects due to the deformation of the
1200: kinks tend to reduce this damping but cannot invert it.  When the
1201: separation is large enough, the damping overcomes the destabilizing
1202: Cahn-Hilliard coupling halting the inverse cascade before it reaches
1203: the gravest mode. The dependence of the critical friction $r_c$ upon
1204: the kink separation $\Lambda/2$ scales as $r_c \sim (e^{-s \Lambda}
1205: /\Lambda)$. The perturbative approach yields a very accurate
1206: analytical expression of $r_c$ because the leading contribution to
1207: $r_c e^{s \Lambda}$, which is algebraic in $\Lambda$ can be obtained
1208: exactly leaving out only terms which are exponentially small in
1209: $\Lambda$.
1210: 
1211: Stabilization by $\beta$-effect is more complex. It does not contribute 
1212: to the linearization of Cahn-Hilliard equation around a steady state and
1213: appears only at the second order of the perturbative expansion in $\beta$. 
1214: As the first-order step of the perturbation expansion is only solved 
1215: for the leading contribution in a $1/\Lambda$ expansion, the analytic
1216: expression of the critical $\beta_c$ is fairly inaccurate when compared
1217: to numerical solution of the perturbative problem or to direct stability
1218: calculations. Its scaling $\beta_c \sim (e^{-s \Lambda} /\Lambda^5)^{1/2}$ 
1219: is, however, correctly predicted.
1220: 
1221: The presence of a finite radius of deformation $1/S$ , which slows
1222: down the fastest Rossby waves, is to provide less efficient
1223: stabilization by the $\beta$ effect.  The critical $\beta_c$ then
1224: scales as $\beta_c \sim (e^{S^4 -s \Lambda} /\Lambda)^{1/2}$.
1225: 
1226: In the presence of mean advection, as in Ref.~\cite{Manfroi:99}, the
1227: eastward jets are narrow and strong while the westward jets are broad
1228: and slow.  The critical value of friction is reduced (see
1229: (\ref{eq:rcadv})) and scales as $r_c \sim (e^{-s
1230:   \Lambda(1-\gamma/\Gamma)} /\Lambda)$
1231: 
1232: Stabilization is demonstrated near the modified steady states of the
1233: Cahn-Hilliard equation. The attraction basin of these steady states
1234: depends on the stability of time-dependent solutions and has been
1235: investigated numerically. The results suggest a fairly simple pattern
1236: where, for most values of the parameters, the phase-space is filled by
1237: the attraction basins of only 2 or 3 stationary solutions. The
1238: boundary of these basins might be very complicated, even fractal.
1239: 
1240: We have observed in the numerical simulations of the pure
1241: Cahn-Hilliard equation that the inverse cascade does not always begin
1242: by the most unstable state $N=k_m$; however it is in principle always
1243: possible to stop the cascade at such a state by enhancing the
1244: friction. The same is not true for $\beta$, which is not always able
1245: to stop the cascade at the scale corresponding to the most unstable
1246: state. As already noticed, $\beta$ dispersive effect is more complex
1247: than friction effect. While the inverse cascade prior to the halting
1248: by friction does not differ from the pure Cahn-Hilliard case, the
1249: $\beta$ effect is paradoxically accelerating the cascade before
1250: halting. We speculate that this is due to the propagation of fast
1251: Rossby waves superimposed to the kinks and increasing their coupling.
1252: 
1253: Our result for the critical friction $r_c$ differs from that given in
1254: Ref.~\cite{Manfroi:99} where the authors found $r_c \sim
1255: \Lambda^{-3}$.  Their reasoning was based on varying and minimizing
1256: the Lyapunov functional with respect to $\Lambda$. Our interpretation
1257: is that this is questionablesince the Lyapunov functional can only be
1258: minimized with respect to the solution, not with respect to the
1259: parameters.
1260: 
1261: The scaling of the critical $\beta_c$ also differs from the scaling
1262: $\beta_c \sim \Lambda^{-3}$ which would arise from standard
1263: phenomenology \cite{Rhines:75} by balancing the nonlinear and the
1264: dispersive term in (\ref{vbetach}). The reason lies in the suppression
1265: of nonlinearities in the slow manifold for solutions of the complete
1266: Cahn-Hilliard equation once the initial transients have been
1267: dissipated. In more realistic two-dimensional or quasigeostrophic
1268: flows, the presence of strong dominating jets and/or coherent eddies
1269: is similarly inducing a reduction of nonlinearities with respect to a
1270: plain dimensional estimate. This is why the prediction of a $k^{-3}$
1271: energy spectra \cite{Kraichnan:67} is usually not observed in forced
1272: two-dimensional flows \cite{Legras:88} with the possible exceptions of
1273: the smallest quasi-passive scales of the motion.
1274: 
1275: 
1276: \section*{Acknowledgments}
1277: 
1278: We thank Joanne Deval for her careful checking of the perturbation 
1279: calculations. We thank Uriel Frisch for his encouragements and numerous
1280: discussions.
1281: 
1282: %%% APPENDIX  -----
1283: \newpage
1284: \appendix
1285: 
1286: \setcounter{section}{0}
1287: \section{Approximate solution of the Cahn-Hilliard equation}
1288: \label{aCH}
1289: \setcounter{equation}{0}
1290: 
1291: The stationary form of the Cahn--Hilliard equation $\pxx v - U'(v)=0$
1292: can be recasted as
1293: \begin{equation}
1294:    \px U = \f12 \px (\px v)^{2} \,,
1295:    \label{statiobase}
1296: \end{equation}
1297: from which the solution is obtained by quadrature.
1298: By integrating (\ref{statiobase}) once, we obtain
1299: \begin{equation}
1300: U = \f12 (\px v)^{2} -C \,,
1301: \label{statioquad}
1302: \end{equation}
1303: where $C$ is a constant which determines the value of $\px v$ on the 
1304: kinks and the periodicity of the solution. For the single-kink 
1305: solution (\ref{kink}), we have $C= \f12 s^{2} \Gamma^{2}$ and the 
1306: asymptotic value of $U$ is $-C$. 
1307: Let us now assume 
1308: \[ C = \f12 s^2 \Gamma^2 (1 - \mu) \, , \]  
1309: where $\mu$ is assumed a small perturbation and try a 
1310: solution under the form
1311: \begin{equation} 
1312:    v = \Gamma \tanh s x + \mu \tv \, .
1313:    \label{approxv}
1314: \end{equation}
1315: By replacing in (\ref{statioquad}) and expanding, we obtain, at first 
1316: order in $\mu$,
1317: \begin{equation}
1318:    2 \px \tv = -4 s \tv \tanh s x - \Gamma s \cosh^2 sx \, .
1319:    \label{vtequ}
1320: \end{equation}
1321: This equation can be solved as
1322: \begin{equation}
1323:    \tv = - \frac{\Gamma}{16} \pr{(3 + \cosh^{2} s x)\tanh s x + \frac{3 s 
1324:    x}{\cosh^2 s x}} \, ,
1325:    \label{vtsol}
1326: \end{equation}
1327: where the condition $\tv(0) = 0$ has been used. We can relate 
1328: $\mu$ to the period of the solution by using $\px v = 0$ for $x = 
1329: \Lambda/4$, leading to
1330: \begin{multline}
1331:    \mu\left\{\f18 s \Gamma \pr{\tanh^{2} sx(3+2\cosh^{2}sx) + 
1332:    \frac{3s \tanh s x}{\cosh^{2} sx}} \right. \\ \left.
1333:    - \f12 \Gamma s \cosh^{2} s x \right\} + s \frac{s 
1334:    \Gamma}{\cosh^2 s x} = 0 \, .
1335:    \label{musol0} 
1336: \end{multline}
1337: When $\Lambda$ is large the main contribution is 
1338: \begin{equation}
1339:    \mu = 64 e^{-s \Lambda} \,.
1340:    \label{musol}
1341: \end{equation}
1342: 
1343: Comparison with numerical solutions of (\ref{vbetach}) shows that
1344: (\ref{approxv}) with (\ref{vtsol}) and (\ref{musol}) approximates
1345: periodic stationary stationary to less than 0.2\% for $\Lambda>10$.
1346: One can build a solution over the whole domain by using
1347: (\ref{approxv}) over contiguous intervals containing a kink matched at
1348: mid-distance between adjacent kinks.  The approximate solution is
1349: continuous and the discontinuity of its derivative at matching points
1350: is $O(\exp(-s |x_{AK}-x_K|))$ where $|x_{AK}-x_K|$ is the distance
1351: between adjacent kinks.
1352: 
1353: \setcounter{section}{1}
1354: \section{Kink motion in the Cahn-Hilliard equation}
1355: \label{s:CH}
1356: \setcounter{equation}{0}
1357: 
1358: This Section is adapted from Kawasaki \& Ohta~\cite{Kawasaki:82} and
1359: corrects one error found in this paper.
1360: 
1361: We assume that the solution is a combination of kinks which are
1362: individually described by (\ref{kink}). As $\pxx M_{\ell}(x)$ and
1363: $W'(M_{\ell})$ decay rapidly away from $x=x_{\ell}$, the solution in
1364: the vicinity of the $j$-th kink is the sum of $M_j(x)$ and of small
1365: contributions from adjacent kinks which are the small deviations from
1366: their asymptotic values. We write
1367: \begin{equation}
1368:    v(x,t) = M_j(x) + \tilde{v}_j(x,t) \, ,         
1369: \end{equation}
1370: where $\tilde{v}_j$ is small in the vicinity of the $j$-th kink, but
1371: takes finite value at distance. A valid expression in the vicinity of
1372: neighbor kinks is
1373: \begin{equation}
1374:    \tilde{v}_j(x,t) = \sum_{\ell<j} (M_{\ell}(x) - M_{\ell}(+ \infty))
1375:    + \sum_{\ell>j} (M_{\ell}(x) - M_{\ell}(- \infty)) \, ,
1376: \end{equation}
1377: where $M_{\ell}(+ \infty)$ and $M_{\ell}(- \infty)$ are the 
1378: asymptotic values at infinity for the basic kink profile.
1379: Here we assume that kinks and antikinks alternate(i.e. $\epsilon_j 
1380: \epsilon_{j+1} = -1$), and that they are numbered from $0$ to $2N-1$ 
1381: within the periodic interval $[0,L]$. The time dependence is entirely 
1382: contained within the positions of the kinks $\{x_j(t)\}$.
1383: 
1384: The temporal evolution of the solution is then given by:
1385: \begin{equation}
1386:    \pt v(x,t) = - \sum_{\ell=0}^{2N-1} \dot{x}_{\ell}(t) \px M_{\ell}(x) \,,
1387:    \label{vtemp}
1388: \end{equation}
1389: where $v(x,t)$ is governed by 
1390: \begin{equation}
1391:    - \frac{1}{\lc} \pxxm \pt v = 
1392:    \pxx v - W'(v) + h(t) 
1393:    \label{CH}\, .
1394: \end{equation}
1395: 
1396: The function $h(t)$ arises from the integration in $x$. The other 
1397: terms arising from the integration vanish owing to the periodicity 
1398: in $x$.
1399: 
1400: In order to estimate the motion of the $j$-th kink, 
1401: we use $M'_j$ as a test function by multiplying (\ref{CH}) 
1402: and integrating over the domain. Then we obtain:
1403: \begin{equation}
1404:    \begin{array}{ccccc}
1405:       \underbrace{\int_0^L \frac{1}{\lc} \sum_{\ell=0}^{2N-1} 
1406:       \dot{x}_{\ell} \px M_j \pxm M_{\ell} dx} &
1407:       = & \underbrace{\int_0^L ( \pxx v - W'(v)) \px M_j dx} & + 
1408:       &\underbrace{\int_0^L  h \px M_j dx} \, . \\
1409:       A&&B&&C 
1410:    \end{array}
1411:    \label{kmoto}
1412: \end{equation}
1413: 
1414: Contribution $A$ can be written as
1415: \begin{equation} 
1416:    A = -\frac{1}{\lc} 
1417:    \sum_{\ell=0}^{2N-1} \dot{x}_{\ell} \int_0^L \int_0^L \px M_j(x) 
1418:    \Gr2(x-x') \px M_{\ell}(x') dx dx'\,,
1419:    \label{A0} 
1420: \end{equation}
1421: where $\Gr2$ is the Green function solution of 
1422: \[ - \pxx \Gr2(x) = \delta(x) \, .\]
1423: $\px M_j$ and $\px M_{\ell}$ are two well separated functions which 
1424: contribute to the integral in (\ref{A0}) respectively in the close vicinity of 
1425: $x_j$ and $x_{\ell}$. 
1426: By expanding $\Gr2(x-x')$ near $\Gr2(x_j-x_{\ell})$ and summing local 
1427: contributions using 
1428: $\int (x-x_j)^2 \px M_j dx = (-1)^j \pi^2 \Gamma / 6 s^2$ and
1429: $\int\int |x-x'| \px M_j \px M_{\ell} dx dx' = 4 \Gamma^2 /s$, we obtain
1430: \begin{equation}
1431:    A = -\frac{4 \Gamma^2}{\lc} 
1432:    \sum_{\ell=0}^{2N-1} \dot{x}_{\ell}
1433:    \pr{(-1)^{j-\ell}\Gr2(x_j-x_{\ell})+ (-1)^{j-\ell} \frac{\pi^2}{12 L s^2}
1434:    - \frac{1}{2 s}\delta_{j-\ell}} \, .
1435: \end{equation}  
1436: Using the expression for $\Gr2$ within the interval $[0,L]$, one
1437: obtains
1438: \begin{multline}
1439:    A = -\frac{2 L \Gamma^2}{\lc} \sum_{\ell=0}^{2N-1} (-1)^{j-\ell}
1440:    \dot{x}_{\ell}
1441:    \left( \frac{((x_j-x_{\ell})[L])^2}{L^2} - 
1442:    \frac{(x_j-x_{\ell})[L]}{L} \right. \\ \left. 
1443:    + \f16 + \frac{\pi^2}{24 L^{2} s^2}
1444:    - \frac{1}{4 L s} (-1)^{j-\ell} \delta_{j-\ell} \right) \,.
1445:    \label{A2}
1446: \end{multline}
1447: 
1448: Contribution $B$ is expanded using 
1449: \begin{equation}
1450:    W'(v) = W'(M_{j}) + W''(M_{j}) \tilde{v}_{j} + 
1451:    W'_{NL}(M_{j},\tilde{v}_{j}) \, .
1452: \end{equation}
1453: Like $\tilde{v}_{j}$, $W'_{NL}$ is small in the vicinity of the $j$-th 
1454: kink but finite at distance. We have
1455: \begin{multline}
1456:    B = - \int_{O}^{L} W'_{NL} \px M_{j} dx
1457:    + \int_{O}^{L}  (\pxx M_{j} - W'(M_{j})) \px M_{j} dx \\
1458:    + \int_{O}^{L} (\pxx \tilde{v}_{j} - W''(M_{j}) \tilde{v}_{j}) 
1459:    \px M_{j} dx \, .
1460:    \label{eq:B}
1461: \end{multline}
1462: The second integral in (\ref{eq:B}) vanishes and the third one 
1463: vanishes also after integration of its first term by part. Therefore, 
1464: we are left with
1465: \begin{equation} 
1466:    B = - \int_{O}^{L}  W'_{NL} \px M_{j} dx \, .
1467:    \label{eq:B2}
1468: \end{equation}
1469: Using (\ref{Wdef}) we find
1470: \[ W'_{NL} = \frac{2 s^2}{\Gamma^2} \tilde{v}_{j}^2 (3 M_{j}  + 
1471: \tilde{v}_{j}) \, . \]
1472: In the vicinity of $x_{j}$, $\tilde{v}_{j}$ is of the order of the 
1473: tails of $M_{j+1}(x_{j})-M_{j+1}(- \infty)$ and 
1474: $M_{j-1}(x_{j})-M_{j-1}(\infty)$, that is $O(\max(e^{-2s (x_{j+1}-x_{j})},
1475: e^{-2s (x_{j}-x_{j-1})})$.
1476: In the vicinity of $x_{j+1}$ and $x_{j-1}$, $\tilde{v}_{j}$ is $O(1)$ 
1477: while $\px M_{j}$ is $O(e^{-2s (x_{j+1}-x_{j})})$ and $O(e^{-2s (x_{j}-x_{j-1})})$.
1478: Therefore, the two main 
1479: contributions to B in (\ref{eq:B2}) arise from the vicinities of 
1480: $x_{j+1}$ and $x_{j-1}$. In the vicinity of $x_{j+1}$ we use
1481: \[ \tilde{v}_{j} = M_{j+1} + \epsilon_{j+1} \Gamma \]
1482: so that 
1483: \[ \tilde{v}_{j}^2 (3 M_{j}  + \tilde{v}_{j}) = \epsilon_{j} \Gamma^3
1484: (2 - \tanh s (x-x_{j+1}))(1+\tanh s (x-x_{j+1}))^2 \, .
1485: \]
1486: Using also
1487: \[ \px M_{j} = 4 s \epsilon_{j} \Gamma e^{-2s(x-x_{j})} \, , \]
1488: and replacing in (\ref{eq:B2}) with similar contributions from the 
1489: vicinity of $x_{j-1}$, we obtain 
1490: \begin{equation}
1491:    B = - 32 s^2 \Gamma^2 (e^{-2s(x_{j+1}-x_{j})} 
1492:    -e^{-2s(x_{j}-x_{j-1})}) \, .
1493: \end{equation}
1494:         
1495: Finally, contribution $C$ gives
1496: \begin{equation}
1497:    C = 2 \epsilon_j \Gamma h(t) \, .
1498: \end{equation} 
1499: 
1500: Summarizing the results, one gets
1501: \begin{multline}
1502:    \label{motok}
1503:    \frac{2 L \Gamma^2}{\lc} \sum_{\ell=0}^{2N-1} (-1)^{j-\ell}
1504:    \left( \frac{((x_j-x_{\ell})[L])^2}{L^2} - 
1505:    \frac{(x_j-x_{\ell})[L]}{L} \right. \\ \left. 
1506:    + \f16 + \frac{\pi^2}{24 L^2 s^2}
1507:    - \frac{1}{4 L s} (-1)^{j-\ell} \delta_{j-\ell} \right) 
1508:    \dot{x}_{\ell} \\ =
1509:    32 s^{2} \Gamma^2 \pr{e^{-2s(x_{j+1}-x_j)} - 
1510:    e^{-2s(x_j-x_{j-1})}} - 2 \epsilon_j \Gamma h(t) \,.
1511: \end{multline}
1512: 
1513: Eq. (\ref{motok}) shows that two neighbor kink and antikink attracts
1514: themselves A stationary solution is obtained when $B$ vanishes for all
1515: values of $j$. This condition is satisfied if the kinks and antikinks
1516: are equispaced over the interval $[0,L]$. Then, $h(t)=0$.
1517: 
1518: Eq. (\ref{motok}) can be used to determine the motion of the kinks
1519: far from the equilibrium, under the condition that the kinks and 
1520: antikinks remain far enough to satisfy the approximations of the above 
1521: calculation.
1522: 
1523: The calculation of Kawasaki \& Ohta ~\cite{Kawasaki:82} slightly differs from 
1524: our own and is limited to the leading order. They 
1525: fail to take into account the exponential variation of 
1526: $\tilde{v}_{j}$ near $x_{j+1}$ and $x_{j-1}$. Therefore, their result 
1527: for the leading order of $B$ contains an error, being too small by a factor 2. 
1528: 
1529: \setcounter{section}{2}
1530: \section{Green function in the periodic domain and calculations of 
1531: coupling coefficient}
1532: \label{s:Green}
1533: \setcounter{equation}{0}
1534: Within the periodic domain $[0,L]$, the $\delta$ function is made periodic
1535: by adding a constant value $-1/L$ everywhere but at the origin.
1536: One can also use
1537: \[ 
1538:    \delta(x) = \f1L \sum_{n \neq 0} \exp{\left(i \frac{2 \pi n}{L}x\right)} 
1539: \, . \]
1540: The solution to 
1541: \[
1542:    \partial^n_{x} \mathcal{G}_n(x) = - \delta(x)
1543: \]
1544: is 
1545: \[    
1546:    \Gr{n}(x)  =  L^{n-1} g_n \pr{\f{x}{L}}  \, ,
1547: \]
1548: where
1549: \begin{align*}
1550: %\label{fgh}
1551:     g_1(x) &= x[1] - \f12 \,, \\
1552:     g_2(x) &=  \f12 ((x[1])^2 - x[1] + \f16)  \,, \\
1553:     g_3(x) &= \f14\pr{\f23 (x[1])^3 - (x[1])^2 +\f13 x[1]} \,, \\
1554:     g_4(x) &=  \f1{24}\pr{ (x[1])^4 -2 (x[1])^3 + 
1555:       (x[1])^{2} -\f1{30}} \,, \\
1556:     g_5(x) &= \f1{720} \pr{6( x[1])^5 - 15 (x[1])^4 + 10(x[1])^3
1557:       -x[1]} \,, \\
1558:     g_6(x) &= \f1{720}\pr{ (x[1])^6 -3 (x[1])^5 + \f52 (x[1])^4 
1559:       -\f12  (x[1])^{2} + \f1{42}} \,,
1560: \end{align*}
1561: where $x[1]$ means $x$ modulo 1.
1562: 
1563: The calculation of the perturbed motion requires to calculate the
1564: Fourier transform
1565: \[        
1566:    \FGr{n}(m) \equiv \sum_{j=0}^{2N-1} (-1)^{j} \Gr{n} 
1567:    \pr{\frac{j \Lambda}{2}} e^{-i \pi \frac{jm}{N}}\, , 
1568: \]
1569: which can be written as a function of
1570: \[        
1571:    s_m(p) \equiv \sum_{j=0}^{2N-1} (-1)^{j} \pr{\frac{j}{2N}}^p
1572:    e^{-i \pi \frac{jm}{N}} \,.
1573: \]
1574: For $m \neq N$, $s_m(p)$ can be calculated using the following 
1575: relations :
1576: \[ 
1577:    P(z,x,J,p) \equiv \sum_{j=0}^{J-1} j^p z^{jx} = \pr{\frac{1}{\ln z}}^p 
1578:    \partial_{x}^p \sum_{j=0}^{J-1} z^{jx} \, , \]
1579:    \[ s_m(p) = \pr{\frac{1}{2N}}^p P(-e^{-i\frac{\pi m}{N}},1,2 N,p) \,. 
1580: \]
1581: We get
1582: \begin{align*}
1583:    s_m(0) &=   0 \,,\\
1584:    s_m(1) &=  -\frac{1}{1+a}  \,,\\
1585:    s_m(2) &=  -\frac{1}{1+a} + \frac{a}{N(1+a)^2}  \,,\\
1586:    s_m(3) &=  -\frac{1}{1+a} + \frac{3a}{2N(1+a)^2}
1587:      + \frac{3a(1-a)}{4N^2(1+a)^3} \,, \\
1588:    s_m(4) &=  -\frac{1}{1+a} + \frac{2a}{N(1+a)^2}
1589:      + \frac{3a(1-a)}{2N^2(1+a)^3} 
1590:      + \frac{a(1-4a+a^2)}{2N^3(1+a)^4} \,, \\
1591:    s_m(5) &=  \cdots + \frac{5a(1-11a+11a^2-a^3)}{16 
1592:      N^{4}(1+a)^{5}} \,, \\
1593:    s_m(6) &=  \cdots + \frac{3a(1-26a+66a^2-26a^3+a^4)}
1594:      {16 N^5(1+a)^6} \,,
1595: \end{align*}
1596: where $a= \exp(-i \theta_m)$ with $\theta_m = \pi m/N$.
1597: We also have
1598: \begin{align*}
1599:    s_N(0) &=  2N \,,\\
1600:    s_N(1) &=  N - \f12  \,,\\
1601:    s_N(2) &=  \frac{1}{12N}(2N-1)(4N-1) \,,\\
1602:    s_N(3) &=  \frac{1}{8N}(2N-1)^2 \,, \\
1603:    s_N(4) &=  \frac{1}{240N^3}(2N-1)(4N-1)(12N^2 -6N -1) \,, \\
1604:    s_N(5) &=  \frac{1}{96N^3}(2N-1)^2(8N^2 -4N -1)  \,, \\
1605:    s_N(6) &=  \frac{1}{1344N^5}(2N-1)(4N-1)(48N^4 -48N^3+6N+1) \,.
1606: \end{align*}
1607: Using these relations, the Fourier transforms are readily calculated.
1608: For $m \neq N$, we have
1609: \begin{align}
1610:    \FGr1(m) 
1611:    &= s_m(1) - \f12 s_m(0) + \f12 \; = \; - \f12 i t
1612:    \label{sf} \,,\\
1613:    \FGr2(m) 
1614:    &=  \frac{L}{2} \pr{ s_m(2) - s_m(1) + \f16 s_m(0)}
1615:    \; = \; \frac{\Lambda}{8} (1 + t^2)
1616:    \label{sg} \,,\\
1617:    \FGr3(m) 
1618:    &=  \frac{L^2}{4} \pr{ \f23 s_m(3) - s_m(2) + \f13 s_m(1)}
1619:    \; = \;  i \frac{\Lambda^2}{32} t (1+t^2)
1620:    \label{sh} \,,
1621: \end{align}
1622: with $t = \tan \pi m / 2 N $.
1623: Since the Fourier transform $\FGr{n}$ scales as $\Lambda^{n-1}$, it 
1624: depends only on the $O(1/N^{n-1})$ term in $s_{m}(n)$. Therefore, the
1625: higher order transforms are, for $m<N$ :
1626: \begin{align}
1627:    \FGr4(m) 
1628:    &=  - \frac{\Lambda^3}{384} (1+t^2) (1+3t^2)
1629:    \label{si} \,,\\
1630:    \FGr5(m) 
1631:    &=  -i \frac{\Lambda^4}{1,\!536} t(1+t^2) (2+3t^2)
1632:    \label{sj} \,,\\
1633:    \FGr6(m) 
1634:    &=  \frac{\Lambda^5}{30,\!720} (1+t^2) (2+15t^2+15t^4)
1635:    \label{sk} \,.
1636: \end{align}
1637: 
1638: Using this formalism, it is possible to calculate $<\va{}, 
1639: \partial_{x}^{-n} \va{}>$ for even $n$ as
1640: \begin{multline} \label{vanvab}
1641:    < \va{}, \partial_{x}^{-n} \va{}> 
1642:    \, = \, - \frac{1}{L} \int_{0}^L \int_{0}^L \va{}(x) 
1643:    \Gr{n}(x-x') \va{}(x') dx dx'  \\
1644:    = - \frac{1}{L} \sum_{j=0}^{2N-1} \sum_{l=0}^{2N-1}
1645:    \cos j \theta_{m} \cos l \theta_{m}
1646:    \int _{0}^L  \int_{0}^L  \px M_{j}(x) \px M_{l}(x') 
1647:    \Gr{n}(x-x') dx dx' \,.  
1648: \end{multline}
1649: This integral contains two types of contributions. Type I arises from 
1650: the interaction of distant kinks and type II is a local correction 
1651: arising from the autocoupling of a given kink and taking into account 
1652: the discontinuity of $\Gr{n}(x)$ or its derivatives in $x=0$. At first, 
1653: we consider only the type I contribution for which $\Gr{n}(x-x')$ can 
1654: be developed as a Taylor series around $x_{j}-x_{l}$ in order to 
1655: separate the double integration in two independent integrations 
1656: around the kinks
1657: \begin{multline*}
1658:    \Gr{n}(x-x') = \Gr{n}\pr{\frac{(j-l) \Lambda}{2}} + 
1659:    \Gr{n-1}\pr{\frac{(j-l) \Lambda}{2}}
1660:    ((x-x_{j})-(x'-x_{l})) \\ 
1661:    + \f12 \Gr{n-2}\pr{\frac{(j-l) \Lambda}{2}}
1662:    ((x-x_{j})-(x'-x_{l}))^2 + \cdots \, .
1663: \end{multline*}
1664: Now we replace in (\ref{vanvab}) and calculate the local contributions 
1665: using
1666: \begin{align*}
1667:    \int \px M_{j} dx  &= 2 \Gamma (-1)^j  \,,\\
1668:    \int (x-x_{j})\px M_{j}  dx &= 0  \,,\\
1669:    \int (x-x_{j})^2 \px M_{j} dx &= \frac{\pi^2 \Gamma}{6 s^2} (-1)^j \,.
1670: \end{align*}
1671: We also expand the trigonometric factor and find that the only 
1672: non vanishing contribution depends on $j-l$. After relabeling, we 
1673: obtain
1674: \begin{multline*}
1675:    <\va{}, \px^{-n} \va{}>_{I} = - \frac{4 \Gamma^{2}}{\Lambda}
1676:    \sum_{j=0}^{2N-1} \cos \frac{\pi m j}{N} (-1)^j \Gr{n}\pr{\frac{j 
1677:    \Lambda}{2}} \\
1678:    - \frac{\pi^2 \Gamma^{2}}{3 s^2 \Lambda}
1679:    \sum_{j=0}^{2N-1} \cos \frac{\pi m j}{N} (-1)^j \Gr{n-2}\pr{\frac{j 
1680:    \Lambda}{2}} \,,
1681: \end{multline*}
1682: that is
1683: \begin{equation}
1684:    <\va{}, \px^{-n} \va{}>_{I} = - \frac{4 \Gamma^{2}}{\Lambda} 
1685:    \FGr{n}(m) - \frac{\pi^2 \Gamma^{2}}{3 s^2 \Lambda} \FGr{n-2}(m) 
1686:    + O (\Lambda^{n-6})
1687:    \label{vanva}
1688: \end{equation}
1689: for even $n$.
1690: A similar relation is obtained for odd $n$ :
1691: \begin{equation}
1692:    <\va{}, \px^{-n} \vb{}>_{I} = - \frac{4 \Gamma^{2}}{\Lambda} 
1693:    \f1i \FGr{n}(m)  - \frac{\pi^2 \Gamma^{2}}{3 s^2 \Lambda} \f1i \FGr{n-2}(m) 
1694:    + O (\Lambda^{n-6}) \,.
1695:    \label{vanvb}
1696: \end{equation}
1697: 
1698: The local type-II contributions must be examined case by case. For 
1699: $n=2$, $\Gr2(x)$ is continuous in $x=0$ but its derivative is not. 
1700: Near $x=0$, we have 
1701: \[ 
1702:    \Gr2(x) = \frac{L}{2} \pr{\f16 - \left| \frac{x}{L}\right| + 
1703:    \frac{x^2}{L^2}} 
1704: \]
1705: The two terms $\f16 + \frac{x^2}{L^2}$ are taken into account in type-I 
1706: contribution. The complementary contribution is
1707: \[
1708:    <\va{}, \pxxm \va{}>_{II} =  \frac{1}{2 \Lambda} \int \int 
1709:    \px M(x) \px M(x') | x-x'| dx dx' \,.
1710: \]
1711: After a bit of algebra, we obtain 
1712: \[
1713:    <\va{}, \pxxm \va{}>_{II} =  \frac{2 \Gamma^2}{\Lambda s}
1714: \]
1715: and finally
1716: \begin{equation}
1717:    <\va{}, \pxxm \va{}> = - \frac{\Gamma^2}{2}(1+t^{2}) + 
1718:    \frac{2 \Gamma^2}{\Lambda s} \, .
1719: \end{equation}
1720: For $n=4$, $\Gr4(x)$ has a discontinuity on its third derivative in 
1721: $x=0$. It brings an $O(1/\Lambda)$ correction in $<\va, \px^{-4} 
1722: \va >$, which is of higher order than the terms in (\ref{vanvb}).
1723: 
1724: For odd orders, including $n=1$, the sine factor cancels the type-II
1725: contribution. Notice that for $n=1$, $\Gr1(x)$ is discontinuous in
1726: $x=0$.  We have arbitrarily assumed that $\Gr1(0)=0$ in (\ref{vanvab})
1727: but this is not important since only the imaginary part of $\FGr1(m)$
1728: is used in the sine transform.
1729: 
1730: Some other quantities may need to be calculated, of the type $<f
1731: \va{}, \px^{-n} \va{}>$ where $n$ is odd and $f$ is a period-$\Lambda$
1732: non localized function which is odd over the kinks. In this case, the
1733: non vanishing contributions are those arising from the odd derivatives
1734: of $\Gr{n}$. At leading order, we obtain
1735: \begin{equation}
1736:    <f \va{}, \px^{-n} \va{}> = - \frac{2 \Gamma}{\Lambda} \FGr{n-1} (m) \int x 
1737:    f(x) \px M(x) dx + \cdots \, .
1738:    \label{vanfva}
1739: \end{equation}
1740: The same expression holds when $\va{}$ is replaced by $\vb{}$.
1741: The case $n=1$ is special, we have
1742: \[
1743:    <f \va{}, \pxm \va{}> = \frac{1}{2 \Lambda} \int  
1744:    f(x) \px M^2(x) dx \, .
1745: \]
1746: 
1747: \setcounter{section}{3}
1748: \section{The effect of mean advection}
1749: \label{s:meanadv}
1750: \setcounter{equation}{0}
1751: 
1752: The case $\gamma \neq 0$ has been studied numerically and near the
1753: linear limit $r_0$ (see Section \ref{s:CHstab}) in
1754: Ref.~\cite{Manfroi:99}. It is possible to study the stability
1755: properties of steady solutions for small $r$ in the same way as for
1756: $\gamma=0$ but this is to the price of a considerable increase of
1757: complexity in the algebra. It is beyond the scope of this manuscript
1758: to describe the details of these cumbersome calculations.  We provide
1759: here the results without demonstration.
1760: 
1761: When $\gamma \neq 0$, the equilibrium positions of the kinks and antikinks
1762: are given by
1763: \begin{align*}
1764:   x_{2p} &= p \Lambda - \frac{\Delta}{4}  \, , \\
1765:   x_{2p+1} &= (2p+1) \frac{\Lambda}{2} + \frac{\Delta}{4} \, , 
1766: \end{align*}
1767: where $\Delta$ is related to $\gamma$ by
1768: \begin{equation} 
1769:   16 \Gamma e^{-s \Delta} \sinh s \Delta = - \frac{\Delta \Gamma}{\Lambda} 
1770:   + \gamma \, .
1771: \end{equation}
1772: We define also $d=\Delta/\Lambda$.
1773: 
1774: Similarly to Section~\ref{s:CHstab}, it is convenient to expand the
1775: displacements of the kinks with respect to the equilibrium in terms
1776: of Fourier components. One has now to separate the kinks and
1777: antikinks as 
1778: \begin{align*}
1779:  \delta x_{2p}      &= \sum_{m=0}^{N-1} \psi_m^- e^{i \pi\frac{2mp}{N}} \, ,\\
1780:  \delta x_{2p\!+\!1}&= \sum_{m=0}^{N-1} \psi_m^+ e^{i \pi\frac{(2p\!+\!1)m}{N}}
1781:  \, .
1782: \end{align*}
1783: By combining these components into 
1784: \[
1785:   \Phi_m = \frac{1}{\sqrt{2}} 
1786:   \left( \begin{array}{cc}
1787:     1 & 1 \\
1788:     e^{-i \theta_m} & e^{i \theta_m}
1789:   \end{array} \right)
1790:   \left( \begin{array}{c}
1791:     \psi_m^-\\
1792:     \psi_m^+
1793:   \end{array} \right) \, ,
1794: \]
1795: we obtain
1796: \begin{equation}
1797:   \dot{\Phi_m}=                           
1798:   \frac{\mathcal{A}\sin^2\theta_m}{\mathcal{E}_m} 
1799:   \left(\begin{array}{cc}
1800:     1-\mathcal{C}\cosh s\Delta - \mathcal{Q}_m & 0 \\
1801:     0 & 1-\mathcal{C}\cosh s\Delta +\mathcal{Q}_m
1802:   \end{array} \right) \Phi_m \,
1803:   \label{eq:Phieq}
1804: \end{equation}
1805: with
1806: \begin{align*}
1807:   \mathcal{A}  &=\frac{128\lc \,e^{-s\Lambda}}{(1-d^2)\,\Lambda} D , \\
1808:   \mathcal{E}_m&=1-\frac{4}{s\Lambda(1-d^2)} +\frac{4\sin^2\theta_m}
1809:     {s^2\Lambda^2(1-d^2)} , \\
1810:   \mathcal{B}  &=\frac{d\cosh s\Delta-\sinh s\Delta}{D} , \\
1811:   \mathcal{C}  &=\frac{2}{s\Lambda D} , \\
1812:   D &= \cosh s\Delta - d \sinh s\Delta , \\
1813:   \mathcal{Q}_m &=\left(\mathcal{B}+2\mathcal{C} \sinh s\Delta 
1814:     +\frac{1}{2}\mathcal{C}^2(\cosh 2s\Delta+\cos 2\theta_m)\right)^{1/2} .
1815: \end{align*}
1816: This result generalizes (\ref{s0}).
1817: 
1818: Now, the stabilization effect by friction is still given at first order 
1819: of the perturbative expansion for small $r$. 
1820: We first need to define
1821: \[
1822:   \rho_m = 2(\mathcal{B} \cos^2 \theta_m + \mathcal{Q}_m \sin^2 \theta_m)^{1/2}
1823: \]
1824: \[\begin{array}{ll}
1825:   \cos \phi_{1m} = \frac{2}{\rho_m} (\mathcal{B} \cos \theta_m) ,
1826: & \sin \phi_{1m} = \frac{2}{\rho_m} (\mathcal{Q}_m \sin \theta_m) , \\
1827:   \cos \phi_{2m} = \frac{2}{\rho_m}(\mathcal{B} + \mathcal{C} \sinh s \Delta
1828:     \sin^2 \theta_m) ,
1829: & \sin \phi_{2m} = -\frac{1}{\rho_m}
1830:     (\mathcal{C} \sin 2 \theta_m \cosh s \Delta).
1831: \end{array}\]
1832: The stabilization effect is 
1833: \begin{multline}
1834:   \sigma_1 = -r + \frac{r \sin^2 \theta_m}{2}
1835:   \left(1-d^2 - \frac{4}{s \Lambda}\right) \\ \times
1836:   \left(1 + d \sin \theta_m \sin (\phi_{2m} - \phi_{1m})
1837:   - \cos \theta_m \cos (\phi_{2m} - \phi_{1m}) 
1838:   - \frac{2 \sin^2 \theta_m}{s \Lambda}\right)^{-1} \, .
1839:   \label{eq:s1adv}
1840: \end{multline}
1841: 
1842: These results have been checked numerically with the same accuracy as those
1843: presented for $\gamma=0$.
1844: 
1845: In the case of large $\Lambda$ and when $d$ is not small, that is when
1846: the asymmetry is strong, we can neglect all terms $O(e^{-s \Delta})$ in
1847: front of terms which are $O(1)$ or larger. Equations (\ref{eq:Phieq})
1848: and (\ref{eq:s1adv}) then simplify considerably. We obtain
1849: \begin{equation}
1850:   \dot{\Phi_m}=                           
1851:   \frac{128 s^3 \lambda e^{-s(\Lambda-\Delta)}}{(1+d)\mathcal{E}_m \Lambda} 
1852:   \left(\begin{array}{cc}
1853:     \left((1-d)s\Lambda\right)^{-1} & 0 \\
1854:     0 & 1
1855:   \end{array} \right) \Phi_m \,,
1856:   \label{eq:Phieq2}
1857: \end{equation}
1858: and
1859: \begin{equation}
1860:   \sigma_1 = -r + \frac{r}{2} \left((1-d^2)-\frac{4}{s \Lambda} \right)
1861:     \left( 1+d-\frac{2}{s \Lambda}\right)^{-1} \, .
1862: \end{equation}
1863: 
1864: As a consequence, the critical value for friction is, at leading order,
1865: \begin{equation}
1866:   r_c = \frac{256 s^3 \lambda e^{-s(\Lambda-\Delta)}}{(1+d)^2 \Lambda} \, .
1867:   \label{eq:rcadv}
1868: \end{equation}
1869: Notice that this relation is not valid for small $d$ and does not 
1870: match (\ref{critr}) for $d=\Delta=0$.
1871: 
1872: \bibliography{../gfd}
1873: 
1874: \setcounter{table}{0}
1875: \renewcommand{\thetable}{\arabic{table}}
1876: \newpage
1877: \input{table-fricbeta.tex}
1878: 
1879: \newpage
1880: \input{table-cmu.tex}
1881: 
1882: \newpage
1883: \input{table-r.tex}
1884: 
1885: \newpage
1886: \input{table-beta.tex}
1887: 
1888: \end{document}
1889: 
1890: 
1891: 
1892: 
1893: 
1894: 
1895: 
1896: 
1897: 
1898: 
1899: 
1900: 
1901: 
1902: 
1903: 
1904: 
1905: 
1906: 
1907: 
1908: 
1909: 
1910: