1: \documentclass[a4paper,pre,twocolumn,showpacs,amsmath,amssymb]{revtex4}
2:
3: \usepackage{graphicx}
4: \usepackage{amsmath}
5:
6: \DeclareMathOperator{\sech}{sech}
7: \DeclareMathOperator{\csch}{cosech}
8: \DeclareMathOperator{\Real}{Re}
9: \DeclareMathOperator{\Imag}{Im}
10: \newcommand{\dx}{\, dx}
11: \newcommand{\conj}[1]{{#1}^{\ast}}
12: \newcommand{\conjl}[1]{\overline{#1}}
13: \newcommand{\order}[1]{{\cal O}\left({#1}\right)}
14:
15: \raggedbottom
16:
17: \usepackage[colorlinks,citecolor=blue,linkcolor=blue,filecolor=blue,urlcolor=blue]{hyperref}
18:
19: \begin{document}
20:
21: \title{Moving solitons in the discrete nonlinear Schr\"odinger equation}
22: \author{O. F. Oxtoby}
23: \email{Oliver.Oxtoby@uct.ac.za}
24: \author{I. V. Barashenkov}
25: \email{Igor.Barashenkov@uct.ac.za}
26: \affiliation{Department of Maths and Applied Maths, University of Cape
27: Town, Rondebosch 7701, South Africa}
28: \date{17 July 2007}
29:
30: \begin{abstract}
31: Using the method of asymptotics beyond all orders,
32: we evaluate the amplitude of radiation from a moving small-amplitude
33: soliton in the discrete nonlinear Schr\"odinger equation.
34: When the nonlinearity is of the cubic type, this amplitude
35: is shown to be nonzero for all velocities and therefore small-amplitude
36: solitons moving without emitting radiation do not exist.
37: In the case of a \emph{saturable} nonlinearity, on the other hand,
38: the radiation is found to be completely suppressed when the soliton moves
39: at one of certain
40: isolated
41: `sliding velocities'.
42: We show that a discrete soliton moving at a general speed will
43: experience radiative deceleration until it either stops and remains
44: pinned to the lattice, or---in the saturable case---%
45: locks, metastably, onto one of the sliding velocities.
46: When the soliton's amplitude is small, however, this deceleration is extremely slow; hence,
47: despite losing energy to radiation, the
48: discrete soliton may spend an exponentially long time travelling
49: with virtually unchanged amplitude and speed.
50:
51: \end{abstract}
52:
53: \pacs{05.45.Yv, 42.65.Tg, 63.20.Pw, 05.45.Ra}
54:
55: \maketitle
56:
57: \section{Introduction}
58:
59: This paper deals with moving solitons of the discrete
60: nonlinear Schr\"odinger (DNLS) equation.
61: The earliest applications of the cubic DNLS equation were to the
62: self-trapping of electrons in lattices (the polaron problem)
63: and energy transfer in biological chains (Davydov solitons)
64: -- see the reviews \cite{hennig,braun,scott} for references.
65: Relatedly, the equation arises in the description of
66: small-amplitude breathers in Frenkel-Kontorova chains
67: with weak coupling \cite{braun}.
68: In optics the equation describes
69: light-pulse propagation in nonlinear waveguide arrays in the
70: tight-binding limit \cite{christodoulides,campbell}.
71: Most recently the DNLS equation has been used to model
72: Bose-Einstein condensates in optically induced lattices
73: \cite{trambettoni}.
74:
75: The question of the existence of moving solitons in the
76: DNLS equation has been the
77: subject of debate for some time
78: \cite{duncan,flach,eilbeck,ablowitz,Kundu,pelinovsky,kevrekidis,
79: Feddersen,FKZ,Flach_Kladko,
80: Malomed_collisions_1,Malomed_collisions_2,Malomed_nonlinearity_management}.
81: Recently, G\'omez-Garde\~nes, Flor\'ia, Falo, Peyrard and Bishop
82: \cite{gomez1,gomez}
83: have demonstrated that the stationary motion of pulses in the cubic one-site DNLS
84: (the `standard' DNLS) is only possible over an oscillatory
85: background consisting of a superposition of plane waves.
86: This result was
87: obtained by numerical continuation of the moving
88: Ablowitz-Ladik breather with two commensurate time scales.
89: In our present paper, we study the travelling discrete solitons analytically
90: and independently of any reference models.
91: Consistently with the conclusions of \cite{gomez1,gomez},
92: we will show that solitons cannot freely move in the cubic DNLS equation;
93: they emit radiation, decelerate and eventually become
94: pinned by the lattice.
95: We shall show, however,
96: that this radiation is exponentially small in the soliton's
97: amplitude, so that broad, small-amplitude pulses are highly mobile and
98: are for all practical purposes indistinguishable from
99: freely moving solitons.
100:
101:
102:
103: In the context of optical waveguide arrays---important
104: not only in themselves but also as a first step
105: to understanding more complicated
106: optical systems such as photonic crystals---%
107: interest among experimentalists \cite{efremidis,fleischer,chen}
108: has recently shifted away from media
109: with pure-Kerr nonlinearity (which gives rise to a cubic term in the DNLS
110: equation) and towards photorefractive media, which exhibit a
111: saturable nonlinearity \cite{cuevas,fleischer-nature,
112: vicencio,hadzievski,stepic}. In practice such arrays may
113: be optically induced in a photorefractive material
114: \cite{efremidis,fleischer} or fabricated permanently -- see
115: \cite{chen}, for example.
116: The study of solitons in \emph{continuous} optical
117: systems with saturable nonlinearitites has a
118: long history; interesting phenomena here
119: include
120: bistability \cite{gatz1,gatz2}, fusion \cite{tikhonenko}
121: and radiation effects \cite{vidal} which do not arise in
122: the cubic equation. As for the \emph{discrete} case, the work of
123: Khare, Rasmussen, Samuelsen and Saxena \cite{khare}
124: suggests that the saturable one-site
125: DNLS may be \emph{exceptional} in the sense of ref.\,\cite{BOP};
126: that is, despite not
127: being a translation invariant system, it supports translationally
128: invariant stationary solitons.
129: This property is usually
130: seen as a prerequisite for undamped motion in
131: discrete equations (see e.g.\ \cite{OPB}) and indeed,
132: the numerical experiments of
133: Vicencio and Johansson \cite{vicencio} have revealed that soliton
134: mobility is enhanced in the saturable DNLS equation.
135:
136:
137: The DNLS equation with a saturable nonlinearity is the second object of
138: our analysis here; our conclusions will turn out to be in agreement with
139: the numerical observations of ref.\,\cite{vicencio}.
140: We will show that for nonlinearities which
141: saturate at a low enough intensity, solitons can
142: \emph{slide}---that is, move without
143: radiative deceleration---at certain isolated velocities.
144: These `sliding' solitons are examples of embedded solitons.
145:
146:
147:
148: The usual saturable DNLS equation is the
149: discrete form of the Vinetskii-Kukhtarev model \cite{vinetskii}:
150: \begin{equation}
151: \label{VK}
152: i\dot{\Phi}_n + \Phi_{n+1} - 2\Phi_n + \Phi_{n-1} -
153: \gamma\frac{1}{1+|\Phi_n|^2} \Phi_n = 0.
154: \end{equation}
155: In order to encompass both cubic and saturable nonlinearities in a single
156: model, we shall instead consider the equation
157: \begin{equation}
158: \label{DNLS}
159: i\dot{\phi}_n + \phi_{n+1} + \phi_{n-1}
160: + \frac{2|\phi_n|^2}{1+\mu|\phi_n|^2}\phi_n = 0,
161: \end{equation}
162: obtained from eq.\,\eqref{VK} by making the transformation
163: $\Phi_n = \sqrt{2/\gamma}e^{-i(2+\gamma)t} \phi_n$ and letting $\mu=2/\gamma$.
164: In the form above, $1/\mu$ represents the saturation threshold of
165: the medium \cite{gatz2}, which tends to infinity as one approaches
166: the pure Kerr (cubic) case of $\mu = 0$. The higher the value of $\mu$,
167: the lower is the intensity at which the nonlinearity saturates.
168:
169:
170:
171: This paper is structured in the following way.
172: In section \ref{section2}, we construct
173: a small-amplitude, broad
174: travelling pulse as an asymptotic
175: series in powers of $\epsilon$, its amplitude. The velocity and frequency of this soliton
176: are obtained as explicit functions of $\epsilon$ and
177: its carrier-wave wavenumber.
178: Then in section \ref{section3}, the main section
179: of this paper, we derive an expression for
180: the soliton's radiation tails and measure their amplitude using the method of asymptotics
181: beyond all orders. In section \ref{section4}, we investigate the influence of
182: this exponentially weak radiation on the
183: soliton's amplitude and speed. Finally, in section
184: \ref{section5}, we summarise our work and make comparisons with some earlier results.
185:
186:
187: \section{Asymptotic Expansion}
188: \label{section2}
189:
190: \subsection{Leading order}
191: \label{section2_A}
192:
193: We begin by seeking solutions of the form
194: \begin{equation}
195: \label{ansatz}
196: \phi_n(t) = \psi(X)e^{ikn+i\omega t},
197: \end{equation}
198: where
199: \begin{equation} X = \epsilon(n-vt)
200: \label{eps_wid}
201: \end{equation} and $\epsilon$ is a parameter.
202: By analogy with the soliton of the continuous NLS,
203: we expect the discrete
204: soliton to be uniquely characterised by two parameters---%
205: e.g.\ $\epsilon$ and $k$---%
206: while the other two ($\omega$ and $v$)
207: are expected to be expressible through
208: $\epsilon$ and $k$.
209: Substituting the Ansatz \eqref{ansatz},\,\eqref{eps_wid} into
210: \eqref{DNLS} gives a differential
211: advance-delay equation
212: \begin{multline}
213: \label{maineqn}
214: \psi(X+\epsilon)e^{ik}+\psi(X-\epsilon)e^{-ik}-\omega\psi(X) \\
215: -i\epsilon v\psi'(X) +
216: \frac{2|\psi(X)|^2} {1+\mu|\psi(X)|^2} \psi(X) = 0.
217: \end{multline}
218: This can be written as an ordinary differential equation of an infinite
219: order:
220: \begin{multline}
221: \label{ode}
222: e^{ik}\sum_{n=0}^{\infty}\epsilon^n\frac{1}{n!}\psi^{(n)}
223: + e^{-ik}\sum_{n=0}^{\infty}\epsilon^n\frac{(-1)^n}{n!}\psi^{(n)}
224: -\omega\psi\\-i\epsilon v\psi' +
225: \frac{2|\psi|^2} {1+\mu|\psi|^2} \psi= 0,
226: \end{multline}
227: where $\psi^{(n)}= d^n \psi/dX^n$.
228:
229: From now on we assume that $\epsilon$ is small.
230: Our aim in this section is to find an approximate solution to
231: eq.\,\eqref{ode}---and hence eq.\,\eqref{DNLS}---with $\psi={\cal O} (\epsilon)$.
232: (That is, we are
233: looking for small, broad
234: pulses which modulate a periodic carrier wave.)
235: To this end, we expand
236: $\psi$, $\omega$ and $v$ as power series in $\epsilon$:
237: \begin{subequations}
238: \label{expansions}
239: \begin{align}
240: \psi &= \epsilon(\psi_0 + \epsilon\psi_1 + \cdots),
241: \label{exp_psi} \\
242: \omega &= \omega_0 + \epsilon^2 \omega_2 + \cdots,
243: \label{exp_om} \\
244: v &= v_0 + \epsilon^2 v_2 + \cdots.
245: \label{exp_v}
246: \end{align}
247: \end{subequations}
248: (We are not expanding $k$ as we consider it, along with $\epsilon$,
249: as one of the two independent parameters characterising our solution.)
250: Substituting these expansions into eq.\,\eqref{ode} gives a hierarchy
251: of equations to be satisfied at each power of $\epsilon$
252: by choosing $\omega_n$ and $v_n$ properly.
253: In nonlinear
254: oscillations, this perturbation
255: procedure is known as Lindstedt's method
256: \cite{Lindstedt}.
257:
258:
259: At the order $\epsilon^1$ we obtain
260: \begin{equation}
261: \omega_0 = 2\cos k,
262: \label{alpha}
263: \end{equation}
264: while the order $\epsilon^2$ gives
265: \begin{equation}
266: v_0=2\sin k.
267: \label{beta}
268: \end{equation}
269: These two relations correspond to the
270: dispersion of linear waves.
271: At the power $\epsilon^3$ we obtain the following nonlinear equation
272: for $\psi_0$:
273: \begin{equation*}
274: \cos k \, \psi''_0 - \omega_2\psi_0
275: +2|\psi_0|^2\psi_0 = 0.
276: \end{equation*}
277: This is the stationary form of the NLS equation, which has
278: the homoclinic solution
279: \[\psi_0 = a\sqrt{\cos k}\sech(aX)\]
280: with
281: \[a^2 = \frac{\omega_2}{\cos k}.\]
282: Returning to the original variable $\psi$, we note that
283: the amplitude $a$ can always be absorbed into
284: $\epsilon$, the parameter in eq.\,(\ref{eps_wid}) and
285: in front of $\psi_0$ in
286: (\ref{exp_psi}). That is, there is no loss of generality
287: in setting $a = 1$ and letting $\epsilon$ describe
288: the amplitude (and inverse width) of the pulse instead.
289: This allows us to set
290: \begin{equation}
291: \omega_2 = \cos k.
292: \label{gamma}
293: \end{equation}
294:
295:
296: Note that the coefficients in eq.\,\eqref{maineqn} are periodic functions
297: of the parameter $k$ with period $2 \pi$; therefore it
298: is sufficient to consider $k$ in the
299: interval $(-\pi, \pi)$.
300: Also, \eqref{maineqn} is invariant with
301: respect to the transformation $k \rightarrow -k$,
302: $ \epsilon \rightarrow -\epsilon$, $v \rightarrow -v$;
303: hence it is sufficient to consider positive $k$
304: only. Finally, our perturbative solution
305: does not exist if $\cos k$ is negative.
306: Thus, from now on we shall assume that $0 \leq k \leq \pi/2$.
307:
308:
309: \subsection{Higher orders}
310:
311: At the order $\epsilon^{n+3}$, where $n \ge 1$, we arrive at the following
312: equations for the real and imaginary parts of $\psi_n$:
313: \begin{subequations}
314: \label{genlinear}
315: \begin{align}
316: \label{genlinearreal}
317: {\cal L}_1\Real\psi_n &= \frac{\Real f_{n-1}(X)}{\cos k} \\
318: \label{genlinearimag}
319: {\cal L}_0\Imag\psi_n &= \frac{\Imag f_{n-1}(X)}{\cos k},
320: \end{align}
321: \end{subequations}
322: where
323: \begin{align*}
324: {\cal L}_0 &= -d^2/dX^2 + 1 - 2\sech^2X, \\
325: {\cal L}_1 &= -d^2/dX^2 + 1 - 6\sech^2X
326: \end{align*}
327: and
328: \begin{widetext}
329: \begin{multline}
330: f_{n-1}(X) = \sum_{j=1}^{[n/2]}\left(\frac{2\cos k}{(2j+2)!}\psi_{n-2j}^{(2j+2)}-\omega_{2j+2}\psi_{n-2j}\right)
331: +i\sum_{j=1}^{[\frac{n+1}{2}]}\left(\frac{2\sin k}{(2j+1)!}\psi_{n-2j+1}^{(2j+1)}-v_{2j}\psi'_{n-2j+1}\right) \\
332: +\sum_{m=1}^{n-1}2\psi_0(\psi_m\conj{\psi}_{n-m}-\psi_{n-m}\conj{\psi}_m)
333: +\sum_{m=1}^{n-1}\sum_{\ell=0}^{n-1-m}2\psi_{n-m-\ell}\psi_m\conj{\psi}_{\ell}\\
334: +\mu\sum_{m=0}^{n-1}\sum_{\ell=0}^{n-1-m}\Bigg[
335: \sum_{j=1}^{[\frac{n-m-\ell}{2}]}\left(\frac{2\cos k}{(2j)!}\psi^{(2j)}_{n-m-\ell-2j} -\omega_{2j}\psi_{n-m-\ell-2j}\right)\\
336: +i\sum_{j=1}^{[\frac{n-m-\ell-1}{2}]}\left(\frac{2\sin k}{(2j+1)!}\psi^{(2j+1)}_{n-m-\ell-2j-1} -v_{2j}\psi'_{n-m-\ell-2j-1}\right)
337: \Bigg]\psi_m\conj{\psi}_{\ell}.
338: \label{all_p}
339: \end{multline}
340: \end{widetext}
341: The linear nonhomogeneous ordinary differential equations \eqref{genlinear} must
342: be solved subject to a boundedness condition.
343:
344: The bounded homogeneous solutions of eqs \eqref{genlinearreal} and
345: \eqref{genlinearimag} ($\sech X\tanh X$ and $\sech X$, respectively)
346: correspond to the translation- and U(1)-invariances of eq.\,\eqref{maineqn}.
347: Including these zero modes in the full solution of eqs \eqref{genlinear}
348: would amount just to the translation of $\psi(X)$ by a constant distance in $X$
349: and its multiplication by a
350: constant phase factor. These deformations are trivial,
351: and hence we can safely discard
352: the homogeneous solutions at each order of $\epsilon$.
353:
354: As $\psi_0$'s real part is even and its imaginary part is odd (zero),
355: $\psi_1$'s real and imaginary parts are also even and odd, respectively.
356: The same holds, by induction, to all orders of the perturbation theory.
357: Indeed, assume
358: that $\psi_0, \psi_1, \ldots, \psi_{n-1}$ have even real parts and odd
359: imaginary parts. Then it is not difficult to verify
360: that the function ${\rm Re} f_{n-1}(X)$ is even and ${\rm Im} f_{n-1}(X)$
361: is odd. Since the operators ${\cal L}_0$
362: and ${\cal L}_1$ are parity-preserving, and since
363: we have excluded the corresponding homogeneous solutions, this means
364: that $\psi_n$ has an even real part and an odd imaginary part.
365: Finally, the homogeneous solution of \eqref{genlinearimag} being even
366: and that of \eqref{genlinearreal} being odd, the
367: corresponding solvability conditions are satisfied
368: at any order.
369:
370:
371:
372:
373: Note that since the solvability conditions do not impose any constraints
374: on $v_n$ and $\omega_n$, the coefficients
375: $v_n$ with $n \geq 2$ and $\omega_n$
376: with $n \geq 4$ can be chosen completely arbitrarily.
377:
378:
379: \subsection{Explicit perturbative solution to order $\epsilon^3$}
380:
381: Solving eqs \eqref{genlinear} successively,
382: we can obtain the discrete soliton \eqref{exp_psi} to
383: any desired accuracy. Here, we restrict ourselves
384: to corrections up to the cubic power in $\epsilon$.
385: The order $\epsilon^3$ is the lowest order
386: at which the saturation parameter $\mu$ appears in the solution.
387: On the other hand, it is high enough to exemplify
388: and motivate
389: our choice of the coefficients in \eqref{exp_om} and \eqref{exp_v}.
390:
391:
392:
393:
394: Letting $n=1$ in eq.\,\eqref{all_p}, we have
395: \begin{equation*}
396: f_0(X) = \frac{2i}{3!} \sin k \psi_0''' - iv_2\psi'_0.
397: \end{equation*}
398: The corresponding solution of
399: eq.\,\eqref{genlinear} is
400: \begin{align*}
401: \psi_1 = \, &\frac{i}{\sqrt{\cos k}} \Bigg[ \frac{1}{2}\sin k\sech X\tanh X \\
402: &\;\;\;\;+\frac{1}{2}\left(v_2-\frac{1}{3}\sin k \right)X\sech X \Bigg].
403: \end{align*}
404: Here the term proportional to $X\sech X$
405: decays to zero as $|X| \to \infty$; however, it becomes greater than
406: $\psi_0(X)$ for sufficiently large $|X|$, leading to
407: nonuniformity of the expansion \eqref{exp_psi}.
408: In order to obtain a uniform
409: expansion, the term in question
410: should be eliminated. Being free to choose the coefficients
411: $v_n$ with $n \geq 2$, we use this freedom to set
412: \[v_2 = \frac{1}{3}\sin k. \]
413: This leaves us with
414: \[\psi_1 = \frac{i}{2} \sqrt{\cos k}\tan k\sech X\tanh X. \]
415: After `distilling' in a similar way the correction
416: $\psi_2$,
417: where we fix $$ \omega_4 = \frac{1}{12} \cos k $$
418: to eliminate a term proportional to $X \sech X$, we obtain
419: \begin{align}
420: \label{regularpert}
421: \psi &= \epsilon\sqrt{\cos k}\Big\{ \sech X + \frac{i}{2} \epsilon\tan k
422: \sech X\tanh X \nonumber\\
423: &\;\;+ \frac{1}{12}\epsilon^2\big[4\sech^3X-3\sech X \nonumber\\
424: &\;\;\;\;+ \frac{1}{2}\tan^2k(14\sech^3X-13\sech X) \nonumber\\
425: &\;\;\;\;\;\;+ 4\mu\cos k(2\sech X - \sech^3X)\big]
426: + {\cal O}(\epsilon^3)\Big\},
427: \end{align}
428: where
429: \begin{subequations}
430: \begin{align}
431: \omega &=
432: 2\left[1+\frac{1}{2!}\epsilon^2+\frac{1}{4!}\epsilon^4+
433: {\cal O}(\epsilon^6)\right]\cos k,
434: \label{omegadef} \\
435: v &=
436: 2\left[1+\frac{1}{3!}\epsilon^2+{\cal O}(\epsilon^4)\right]\sin k.
437: \label{cdef}
438: \end{align}
439: \end{subequations}
440:
441:
442:
443:
444:
445:
446:
447:
448: \subsection{Velocity and frequency of the discrete soliton}
449:
450: In the previous subsection we have shown that fixing
451: suitably the coefficients $\omega_n$ and $v_n$ can lead to a uniform
452: expansion of $\psi$ to order $\epsilon^3$. Our approach
453: was based on solving for $\psi_n$ explicitly and then setting
454: the coefficients in front of the `secular' terms $X \sech X$ to zero.
455: Here, we show that the secular terms can be eliminated
456: to all orders -- and
457: without appealing to explicit solutions.
458:
459: Assume that the secular terms have been
460: suppressed in all nonhomogeneous solutions
461: $\psi_m$ with $m$ up to $n-1$; that is, let
462: \begin{align}
463: \psi_m(X) &\to C_m e^X + o(e^X) \nonumber\\
464: &\quad\mbox{as} \ X \to -\infty \quad
465: (m=0,1, \ldots, n-1).
466: \label{as_m}
467: \end{align}
468: (Since we know that the real part of $\psi_m$ is even
469: and the imaginary part odd, it is sufficient to consider the
470: asymptotic behaviour at one infinity only.)
471: The constant $C_m$ may happen to be zero for some $m$
472: in which case the decay of $\psi_m$ will be faster
473: than $e^X$.
474: Our objective is to choose $\omega_n$ and $v_n$ in
475: such a way that $\psi_n(X)$ will also satisfy \eqref{as_m}.
476: To this end, we consider the function \eqref{all_p}.
477: All terms which are trilinear in $\psi_0,\ldots,\psi_{n-1}$
478: and derivatives of these functions decay as $e^{3X}$ or faster; these
479: terms in $f_{n-1}$
480: cannot give rise to the secular terms
481: proportional to $X \sech X$ in
482: $\psi_n$. On the other hand, the terms making up the first
483: line in \eqref{all_p} tend to
484: \begin{align*}
485: &e^X \sum_{j=1}^{[n/2]} \left[
486: \frac{2\cos k}{(2j+2)!}
487: -\omega_{2j+2}\right] C_{n-2j} \\
488: &\quad +i e^X \sum_{j=1}^{[\frac{n+1}{2}]}\left[
489: \frac{2\sin k}{(2j+1)!}
490: -v_{2j}\right] C_{n-2j+1}
491: \end{align*}
492: as $X \to -\infty$. These are `asymptotically resonant' terms --
493: in the sense that
494: their asymptotics are proportional to the asymptotics of the
495: homogeneous solutions of \eqref{genlinearreal} and
496: \eqref{genlinearimag}.
497: It is these terms on the right-hand sides
498: of \eqref{genlinearreal} and
499: \eqref{genlinearimag} that give rise to the secular terms in the
500: solution $\psi_n$.
501: The resonant terms will be suppressed if
502: we let
503: \begin{equation}
504: \omega_{2j+2} = \frac{2 \cos k} { (2 j+2)!}, \quad
505: v_{2j} =
506: \frac{2 \sin k} { (2 j+1)!}, \quad
507: j \ge 1.
508: \label{om_and_v}
509: \end{equation}
510: After the resonant terms have been eliminated, the bounded solution
511: to \eqref{genlinear} will have the asymptotic behaviour $\psi_n \to C_n e^X$
512: as $X \to -\infty$. By induction, this result extends to all $n \ge 0$.
513:
514: Substituting \eqref{om_and_v}, together with
515: \eqref{alpha}--\eqref{gamma}, into
516: eqs \eqref{exp_om} and \eqref{exp_v}, and summing
517: up the series, we obtain
518: \begin{equation}
519: \omega= 2 \cos k \cosh \epsilon, \quad
520: v= 2 \sin k \frac{\sinh \epsilon}{\epsilon},
521: \label{om_and_v_via_eps}
522: \end{equation}
523: the frequency and velocity of the discrete soliton
524: parameterised
525: in terms of $k$ and $\epsilon$.
526:
527:
528:
529: Equations \eqref{om_and_v_via_eps}
530: coincide with the expressions \cite{AL} of the Ablowitz-Ladik
531: soliton's velocity and frequency in terms of its amplitude and wavenumber.
532: The difference between the two sets of answers is in that
533: eqs \eqref{om_and_v_via_eps} pertain to small-amplitude
534: solitons only, whereas Ablowitz and Ladik's formulas
535: are valid for arbitrarily large amplitudes.
536:
537: Note, also, that the velocity and frequency
538: \eqref{om_and_v_via_eps}
539: do not depend on the saturation parameter $\mu$.
540: This is in contrast to the stationary ($v=0$) soliton of
541: eqs \eqref{maineqn} and \eqref{ode}
542: obtained by Khare, Rasmussen, Samuelsen and Saxena \cite{khare}.
543: The soliton of ref.\,\cite{khare} has its amplitude and
544: frequency uniquely determined by $\mu$.
545:
546:
547:
548:
549: % We close this section by commenting on the case of stationary
550: % solitons.
551: % From eq.\,\eqref{om_and_v_via_eps} it follows that
552: % $v=0$ implies $k=0$, and so all terms with $i$ in front are absent from the
553: % expansion \eqref{all_p}.
554: % As a result,
555: % both $f_{n-1}(X)$ and $\psi_n(X)$ are real in this case,
556: % for all $n$ --- this can be easily shown by induction.
557: % Since real parts of all $\psi_n(X)$ are even, we conclude that
558: % for those $\epsilon$ and $\mu$ for which the
559: % series \eqref{exp_psi} converges, it converges to an even solution.
560:
561:
562:
563: \section{Terms beyond all orders of the perturbation theory}
564: \label{section3}
565:
566: \subsection{Dispersion relation for linear waves}
567:
568:
569:
570: As $X \to -\infty$, the series \eqref{exp_psi} reduces to
571: $\psi(X)= \sum_{n=0}^\infty \epsilon^{n+1}C_n e^X$.
572: The convergence of the series $\sum \epsilon^{n+1} \psi_n(X)$
573: for all $X$ would imply, in particular, the convergence of the
574: series $\sum \epsilon^{n+1}C_n $. Therefore, if the series
575: \eqref{exp_psi} converged, the solution $\psi(X)$ would be
576: decaying to zero as $X \to -\infty$. However, although
577: we have shown that the series $\sum \epsilon^{n+1} \psi_n(X)$
578: is asymptotic
579: to all orders, it does not have to be convergent.
580: For instance, it is easy to see that the series
581: cannot converge for $v=\mu=0$ and any $\epsilon$; since the
582: advance-delay equation \eqref{maineqn} is
583: translation invariant, this would imply that we have constructed
584: a family of stationary solitons
585: with an arbitrary position relative to the lattice.
586: This, in turn, would contradict the well established fact that the
587: `standard' cubic DNLS solitons can only be centered on a
588: site or midway between two adjacent sites \cite{Laedke_Kluth_Spatschek}
589: (that is, that the `standard' discretisation
590: of the cubic NLS is not exceptional \cite{pelinovsky-exceptional}).
591: Thus, we expect that
592: the perturbative solution \eqref{exp_psi} satisfies
593: $\psi(X) \to 0$ as $|X| \to \infty$ only for some special
594: choices of $v$, $\epsilon$ and $\mu$.
595:
596:
597: Can eqs \eqref{maineqn} and \eqref{ode} have a bounded solution
598: despite the divergence of the corresponding series $\sum
599: \epsilon^{n+1}C_n $? To gain some insight into this matter, we linearise
600: eq.\,\eqref{maineqn} about $\psi=0$ and
601: find nondecaying solutions of the form $\psi = e^{i{\cal Q} X/\epsilon}$
602: where ${\cal Q}$ is a root of the dispersion relation
603: \begin{equation}
604: \label{disprel}
605: \omega = 2\cos(k+{\cal Q})+ {\cal Q} v.
606: \end{equation}
607: It is easy to check that there is at least one such harmonic solution
608: [i.e.\ eq.\,\eqref{disprel} has at least one root] if $v \neq 0$.
609: These harmonic waves can form a radiation background over
610: which the soliton propagates (as suggested by the numerics of
611: \cite{gomez1,gomez} and the analysis
612: of a similar problem for the $\phi^4$ kinks in \cite{OPB}). Being
613: nonanalytic in $\epsilon$, such backgrounds
614: cannot be captured by any order of the
615: perturbation expansion.
616:
617: As we will show below, not only the wavenumbers but also the amplitudes
618: of the harmonic waves are nonanalytic in $\epsilon$.
619: This phenomenon was first encountered
620: in the context of the breather of the continuous
621: $\phi^4$ model, where Eleonskii, Kulagin, Novozhilova,
622: and Silin \cite{eleonskii}
623: suggested that the radiation from
624: the breather could be
625: exponentially weak. Segur and Kruskal \cite{sk,sk2}
626: then developed the method of `asymptotics beyond all orders' to
627: demonstrate that, in the limit
628: $\epsilon \rightarrow 0$, such radiation does exist.
629: We will use Segur and Kruskal's method to measure
630: the magnitude of the radiation background of the travelling discrete soliton.
631:
632:
633: Qualitatively, the fact that the radiation
634: is not excited at any order of the perturbation expansion
635: is explained by the fact that the soliton exists on the
636: long length scale $X$, whereas the radiation has the shorter
637: scale $X/\epsilon$. To all orders, the two are uncoupled.
638:
639:
640:
641: Using the relations \eqref{om_and_v_via_eps},
642: we can rewrite \eqref{disprel} as
643: \begin{equation}
644: \label{nicedisprel}
645: \frac{\cosh\epsilon-\cos {\cal Q}}
646: {(\sinh\epsilon/ \epsilon){\cal Q} -\sin {\cal Q}} = \tan
647: k.
648: \end{equation}
649: The left hand side is plotted in fig.\,\ref{nicedisprelfig}. For $\epsilon =
650: 0$ it has minima at multiples of $2\pi$ where the curve is tangent to the horizontal axis.
651: For nonzero $\epsilon$, the minima of the curve are lifted off the
652: ${\cal Q}$ axis
653: slightly.
654: The minima with larger values of ${\cal Q}$ have smaller elevations above the
655: horizontal axis; i.e.,
656: the minima come closer and closer to the ${\cal Q}$
657: axis as ${\cal Q}$ grows.
658: We see from the figure that for $k >k^{(1)}_{\rm max}$,
659: where $k^{(1)}_{\rm max} \approx 0.22$, there is only one
660: radiation mode. Note also that the left-hand side of
661: eq.\,\eqref{nicedisprel} is negative for negative ${\cal Q}$; since
662: we have assumed that $0 \le k \le \pi/2$, this implies
663: that eq.\,\eqref{nicedisprel} cannot have negative roots.
664:
665: \begin{figure}[btp]
666: \includegraphics{disprel.eps}
667: \caption{\label{nicedisprelfig}The left hand side of eq.\,\eqref{nicedisprel} for
668: two values of $\epsilon$. The root(s) ${\cal Q}_n$ of the
669: dispersion relation \eqref{disprel} are located where this graph
670: is intersected by a horizontal straight line with ordinate
671: equal to $\tan k$.}
672: \end{figure}
673:
674:
675:
676:
677:
678:
679: \subsection{Radiating solitons}
680: \label{Radi_Soli}
681:
682: Although we originally constructed the
683: expansion \eqref{exp_psi} as an asymptotic
684: approximation to a solution which is stationary
685: in the frame of reference moving with the velocity $v$, it can
686: also represent an approximation to a time-dependent solution
687: $\psi(X,t)$. Here $\psi(X,t)$ is related to $\phi_n(t)$,
688: the discrete variable in eq.\,\eqref{DNLS},
689: by the substitution \eqref{ansatz}:
690: \begin{equation}
691: \label{transformation}
692: \phi_n(t) = \psi(X,t)e^{ikn+i\omega t}.
693: \end{equation}
694: The coefficients $\psi_n$ in the asymptotic expansion of $\psi(X,t)$ will
695: coincide with the coefficients in the expansion of the stationary
696: solution $\psi(X)= \sum
697: \epsilon^{n+1} \psi_n $ if
698: the time derivatives $\partial_t\psi_n$
699: lie beyond all orders of $\epsilon$ and hence the time evolution
700: of the free parameters $k$ and $\epsilon$ occurs on a time scale
701: longer than
702: any power of $\epsilon^{-1}$.
703: Physically, one such solution represents
704: a travelling soliton slowing down and attenuating as the Cherenkov
705: radiation left in its wake carries momentum and energy away from its
706: core.
707:
708:
709: Substituting \eqref{transformation} into
710: eq.\,\eqref{DNLS}, gives
711: \begin{equation}
712: \label{timedepeqn}
713: i\psi_t+\psi^+e^{ik}+\psi^-e^{-ik}-\omega\psi-i\epsilon v\psi_X
714: +\frac{2|\psi|^2}{1+\mu|\psi|^2} \psi = 0,
715: \end{equation}
716: where $\psi^{\pm} = \psi(X\pm\epsilon,t)$.
717:
718: We consider two solutions of this equation which both have the same
719: asymptotic expansion \eqref{exp_psi}, denoted $\psi_s(X,t)$
720: and $\psi_u(X,t)$, such that $\psi_s(X,t)\to 0$ as $X \to +\infty$ and
721: $\psi_u(X,t) \to 0$ as $X \to -\infty$.
722: Since the difference $\Psi \equiv \psi_s-\psi_u$ is small
723: (lies beyond all orders of $\epsilon$), and since the solution
724: $\psi_s$ can be regarded as a perturbation of $\psi_u$,
725: $\Psi$ obeys the linearisation of eq.\,\eqref{timedepeqn}
726: about $\psi_u$ to a good approximation. That is,
727: \begin{multline}
728: \label{linearised}
729: i\Psi_t + \Psi^+e^{ik} + \Psi^-e^{-ik} - \omega\Psi
730: -iv\epsilon\Psi_X \\
731: + \frac{4|\psi_u|^2\Psi +2\psi_u^2\conj{\Psi}}{1+\mu|\psi_u|^2}
732: + \frac{2\mu|\psi_u|^2(|\psi_u|^2\Psi +\psi_u^2\conj{\Psi})}{(1+\mu|\psi_u|^2)^2} = 0.
733: \end{multline}
734: Since $\psi_u = \order{\epsilon}$, we can solve eq.\,\eqref{linearised}
735: to leading order in $\epsilon$ by ignoring the last two terms in it; the
736: resulting solutions are
737: exponentials of the form $e^{i\mathcal{Q}X/\epsilon-i\Omega t}$. We make a
738: preemptive simplification by setting $\Omega = 0$.
739: [That $\Omega$ has to be set equal to zero follows from
740: matching these exponentials to the far-field
741: asymptotes of the stationary `inner' solution;
742: see section \ref{IIID} below.
743: Physically, $\Omega = 0$ implies that the travelling pulse will
744: only excite the radiation
745: with its own (zero) frequency in the comoving frame.]
746: The leading-order solution of eq.\,\eqref{linearised} is therefore
747: \begin{equation}
748: \label{stable-minus-unstable-outer}
749: \Psi = \sum_n A_n e^{i {\cal Q}_nX/\epsilon} + \order{\epsilon^1},
750: \end{equation}
751: where ${\cal Q}_n$ ($n=1,2,\ldots$) are the roots, numbered in order from smallest to
752: largest, of the dispersion relation \eqref{disprel}.
753: (Recall that since we have taken $k$ in the interval
754: $[0, \pi/2]$, all the roots ${\cal Q}_n$ are positive.)
755: For $k >k_{\rm max}^{(1)} \approx 0.22$, there is only one root, ${\cal Q}_1$.
756:
757: Higher-order corrections to the solution \eqref{stable-minus-unstable-outer}
758: can be found
759: by substituting the ansatz
760: \begin{multline}
761: \label{radansatz}
762: \Psi = \sum_n A_n[1+\epsilon f_1^{(n)}(X)
763: +\epsilon^2f_2^{(n)}(X)+\cdots]e^{i {\cal Q}_n X/\epsilon} \\
764: + \sum_n A_n^* [\epsilon^2 g_2^{(n)}(X)
765: +\epsilon^3 g_3^{(n)}(X)+\cdots]e^{-i{\cal Q}_n X/\epsilon}
766: \end{multline}
767: into eq.\,\eqref{linearised}, expanding the advance/delay terms
768: $f^{(n)}_{1,2,\ldots}(X\pm\epsilon)$ and
769: $g^{(n)}_{2,3,\ldots}(X\pm\epsilon)$ in Taylor series in $\epsilon$,
770: and making use of
771: the asymptotic expansion \eqref{exp_psi},\,\eqref{regularpert} for $\psi_u$.
772: For instance, the first few corrections are found to be
773: \begin{equation}
774: \begin{split}
775: %\label{rad2ndorder}
776: f_1^{(n)}(X) &= \frac{4i\cos k}{2\sin(k+{\cal Q}_n)-v}\tanh X,
777: \\
778: g_2^{(n)}(X) &= \frac{2 \cos k}{\omega+ vq_n-2\cos(k-{\cal Q}_n)}\sech^2 X.
779: \label{f1g2}
780: \end{split}
781: \end{equation}
782:
783:
784:
785: Since $\psi_u \to 0$ as $X \to -\infty$, it follows that
786: $\psi_s \to \Psi$ as $X \to -\infty$, and hence, once we know
787: the amplitudes $A_n$, we know the asymptotic behaviour of $\psi_s$.
788: We shall now employ the method of asymptotics beyond all orders to
789: evaluate these amplitudes.
790:
791:
792: \medskip
793:
794: \subsection{`Inner' equations}
795:
796: Segur and Kruskal's method allows one to measure the amplitude of
797: the exponentially small radiation
798: by continuing the solution analytically into the complex plane.
799: The leading-order term
800: of $\psi$, $\epsilon\sqrt{\cos k}\sech X$, has singularities at
801: $X = \frac{i\pi}{2} + i\pi n$, $n=0, \pm 1, \pm 2, \ldots$.
802: In the vicinity of these points,
803: the radiation
804: becomes significant; the qualitative explanation for this is that the
805: $\sech$ function forms a sharp spike with a
806: short length scale near the singularity point,
807: and hence there is a strong coupling to
808: the radiation modes, unlike on
809: the real axis. The radiation, which is exponentially small
810: on the real axis, becomes large enough to be measured near the
811: singularities.
812:
813:
814:
815: We define a new complex variable $y$, such that $\epsilon y$ is
816: small in absolute value
817: when $X$ is near the lowest singularity
818: in the upper half-plane:
819: \begin{equation}
820: \label{inner-variable}
821: \epsilon y = X-\frac{i\pi}{2}.
822: \end{equation}
823: The variables $y$ and $X$ are usually
824: referred to as the
825: `inner' and `outer' variables, respectively -- the transformation
826: to $y$ effectively `zooms in' on the singularity at
827: $X = \frac{i\pi}{2}$.
828: We also define $u(y) \equiv \psi(X)$ and $w(y) \equiv \psi^*(X)$.
829: Continuing eq.\,\eqref{maineqn}
830: to the complex plane---%
831: i.e.\ substituting $u(y)$ for $\psi(X)$ and $w(y)$
832: for $\psi^*(X)$---we obtain, in the limit $\epsilon\to 0$,
833: \begin{widetext}
834: \begin{subequations}
835: \label{inner-equations}
836: \begin{align}
837: & e^{ik}u(y+1)+e^{-ik}u(y-1)-2\cos k \, u(y) - 2i\sin k \, u'(y) +
838: \frac{2u^2w}{1+\mu uw} = 0, \\
839: & e^{-ik}w(y+1)+e^{ik}w(y-1)-2\cos k \, w(y) + 2i\sin k \, w'(y) +
840: \frac{2w^2u}{1+\mu wu} = 0.
841: \end{align}
842: \end{subequations}
843: Here we have used the fact that
844: $\omega \to 2\cos k$
845: and $v \to 2\sin k$ as $\epsilon \rightarrow 0$. Equations \eqref{inner-equations}
846: are our `inner
847: equations'; they are valid in the `inner region' $-\infty< \text{Re}
848: \, y
849: < \infty$,
850: $\text{Im} \, y<0$. (The solution cannot be continued up from the real $X$ axis
851: past the singularity at $y = 0$.)
852:
853: Solving the system
854: \eqref{inner-equations} order by order,
855: we can find solutions in the form of
856: a series in powers of $y^{-1}$.
857: Alternatively, we can
858: make the change of variables \eqref{inner-variable}
859: in the asymptotic expansion %\eqref{exp_psi},
860: \eqref{regularpert} and send
861: $\epsilon \rightarrow 0$. This gives, for the first few terms,
862: \begin{subequations}
863: \label{pertexpy}
864: \begin{align}
865: \hat{u} &= \sqrt{\cos k}\left\{ -\frac{i}{y} + \tan k\frac{1}{2y^2}
866: + \left[\frac{1}{3}(1-\mu\cos k)+\frac{7}{12}\tan^2k\right]\frac{i}{y^3}
867: + \mathcal{O}(y^{-4})\right\}, \\
868: \hat{w} &= \sqrt{\cos k}\left\{ -\frac{i}{y} - \tan k\frac{1}{2y^2}
869: + \left[\frac{1}{3}(1-\mu\cos k)+\frac{7}{12}\tan^2k\right]\frac{i}{y^3}
870: + \mathcal{O}(y^{-4})\right\}.
871: \end{align}
872: \end{subequations}\end{widetext}
873: We are using hats over $u$ and $w$ to
874: distinguish the series solution \eqref{pertexpy}
875: from other solutions of eq.\,\eqref{inner-equations}
876: that will appear in the next section. The asymptotic series \eqref{pertexpy}
877: may or may not converge.
878: We note a symmetry $\hat{u}(-y)=-\hat{w}(y)$ of the power-series
879: solution.
880:
881:
882: \subsection{Exponential expansion}
883: \label{IIID}
884:
885: In order to obtain an expression for the terms which lie beyond
886: all orders of $y^{-1}$, we substitute $(u,w) = (\hat{u}, \hat{w}) +
887: (\delta u, \delta w)$ into eqs \eqref{inner-equations}.
888: Since $\hat{u}$ and $\hat{w}$ solve
889: the equations to all orders in $y^{-1}$, then provided $\delta u$ and
890: $\delta w$ are small, they will solve the linearisation of eqs
891: \eqref{inner-equations} about $(\hat{u},\hat{w})$ for large $|y|$.
892:
893: Formal solutions to the linearised system can
894: be constructed as series in powers of $y^{-1}$.
895: Because $\hat{u}$ and $\hat{w}$ are both $\order{y^{-1}}$, the
896: leading-order expressions for $\delta u$ and $\delta w$ as $y \to
897: \infty$ are obtained by substituting zero for $\hat{u}$ and $\hat{w}$
898: in the linearised equations. This gives
899: \begin{equation}
900: \begin{split}
901: \label{leading-order-inner}
902: \delta u &\to \sum_n J_n \exp(iq_ny), \\
903: \delta w &\to \sum_n K_n \exp(-iq_ny) \quad \mbox{as} \ y \to \infty,
904: \end{split}
905: \end{equation}
906: where $q_n$ ($n=0,1,2,\ldots$)
907: are the roots of
908: \begin{equation}
909: \label{leading-order-disp}
910: \cos(k+q)-\cos k + q\sin k = 0.
911: \end{equation}
912: Note that the roots $q_n$ with $n \geq 1$ are given by the $\epsilon \to 0$
913: limits of the roots of
914: the dispersion equation
915: \eqref{nicedisprel}:
916: $q_n=\lim_{\epsilon \to 0} {\cal Q}_n$.
917: In addition, there is a root $q_0=0$ which does not have a $\mathcal{Q}_0$ counterpart.
918:
919: The full solutions (i.e.\ solutions including corrections
920: to all orders in $y^{-1}$) will result if we use the full inverse-power series
921: \eqref{pertexpy}
922: for $\hat u$ and $\hat w$; these solutions should have the form
923: \begin{subequations}
924: \label{expexpansion}
925: \begin{align}
926: \label{expexpansiona}
927: \delta u &= \sum_n K_n\sum_{m=1}^{\infty}
928: \frac{d_m^{(n)}}{y^m} \exp(-iq_ny), \\
929: \label{expexpansionb}
930: \delta w &= \sum_n K_n\left[1+\sum_{m=1}^{\infty}
931: \frac{c_m^{(n)}}{y^m}\right] \exp(-iq_ny).
932: \end{align}
933: \end{subequations}
934: Note that we have excluded the terms proportional to $e^{iq_ny}$
935: from this ansatz (i.e.\ set the amplitudes $J_n$ to zero) as they would become
936: exponentially large
937: on the real $X$ axis. [One can readily verify this by making the change of
938: variables \eqref{inner-variable} in eq.\,\eqref{leading-order-inner}.]
939: The coefficients $c_1^{(n)}$, $c_2^{(n)}$, \ldots and $d_1^{(n)}$,
940: $d_2^{(n)}$, \ldots
941: are found recursively when the ansatz \eqref{expexpansion}
942: is substituted into the linearised
943: equations and like powers of $y^{-1}$ collected.
944: In particular, the first few coefficients are
945: \begin{subequations}
946: \begin{align}
947: c_1^{(n)} &= -\frac{2i\cos k}{\sin(k+q_n)-\sin k}, \nonumber\\
948: \label{c1c2}
949: c_2^{(n)} &= \frac{[\cos(k+q_n)-2\cos k]\cos k}{[\sin(k+q_n)
950: -\sin k]^2}
951: \end{align}
952: and
953: \begin{equation}
954: d_1^{(n)} = 0, \quad d_2^{(n)} = \frac{\cos k}{\cos(k-q_n)
955: -\cos k-q_n\sin k},
956: \label{d1n}
957: \end{equation}
958: \end{subequations}
959: where $n=1,2, \ldots$.
960:
961: Having restricted ourselves to considering the linearised
962: equations for $\delta u$ and $\delta w$, we have
963: only taken into account the simple harmonics in
964: \eqref{leading-order-inner} and \eqref{expexpansion}.
965: Writing $\delta u= \varepsilon^1 U_1+ \varepsilon^2 U_2+\cdots$ and
966: $\delta w=\varepsilon^1 W_1+ \varepsilon^2 W_2+\cdots$, where
967: $\varepsilon$ is an auxiliary small parameter
968: (not to be confused with our `principal' small parameter $\epsilon$);
969: substituting $u=\hat{u}+\delta u$ and
970: $w=\hat{w}+\delta w$ in eqs \eqref{inner-equations},
971: and solving order-by-order the resulting hierarchy
972: of nonhomogeneous linear equations, we can recover all nonlinear
973: corrections to $\delta u$ and $\delta w$.
974: The $\varepsilon^2$-corrections
975: will be proportional to
976: $e^{-i(q_n+q_m)y}$;
977: higher-order corrections will introduce harmonics
978: with higher combination wavenumbers. Later in this section
979: it will become clear that $\varepsilon$ is actually of the order $\exp(-\pi
980: \mathcal{Q}_1 / 2 \epsilon)$ [see eq.\,\eqref{symb} below]; hence the
981: amplitudes of the combination harmonics will
982: be exponentially smaller than that of $\exp(-i q_1 y)$.
983:
984: Now we return to the object that is of ultimate interest to us in
985: this work -- that is, to the function $\Psi$ of section \ref{Radi_Soli}
986: representing the radiation of the moving soliton. We
987: need to match $\Psi$ to the corresponding object in the
988: inner region. To this end, we recall that $\Psi=\psi_s-\psi_u$, where
989: $\psi_s$ and $\psi_u$ are two solutions of the outer equation
990: \eqref{maineqn} which share the same asymptotic expansion
991: to all orders. In the limit $\epsilon \to 0$, the
992: corresponding functions
993: \begin{equation}
994: \begin{aligned}
995: u_s(y,t) &\equiv \psi_s(X,t),
996: &w_s(y,t) &\equiv \psi_s^*(X,t), \\
997: u_u(y,t) &\equiv \psi_u(X,t),
998: &w_u(y,t) &\equiv \psi_u^*(X,t) \label{continua}
999: \end{aligned}
1000: \end{equation}
1001: solve eqs \eqref{inner-equations}
1002: and share the same inverse-power expansions. We express this fact by writing
1003: \begin{align*}
1004: u_s(y,t) &\sim \hat{u}(y),\quad
1005: &w_s(y,t) &\sim \hat{w}(y),\qquad \\
1006: u_u(y,t) &\sim \hat{u}(y),\quad
1007: &w_u(y,t) &\sim \hat{w}(y).\qquad
1008: \end{align*}
1009: \begin{widetext}
1010: Therefore, the difference $u_s-u_u$ (which
1011: results from the analytic continuation of the function $\Psi$)
1012: can be identified with $\delta u$ and $w_s-w_u$ with $\delta w$.
1013: Continuing eq.\,\eqref{radansatz} and its
1014: complex conjugate gives,
1015: as $\epsilon \to 0$,
1016: \begin{subequations} \label{cont12}
1017: \begin{align} u_s-u_u = & \sum_n
1018: \lim_{\epsilon \to 0}
1019: A_n(\epsilon) \exp \left(-\frac{\pi \mathcal{Q}_n}{2 \epsilon} \right)
1020: \left[ 1+\order{\frac1y} \right]
1021: \exp(iq_ny) \nonumber\\
1022: &- \sum_n
1023: \lim_{\epsilon \to 0}
1024: A_n^*(\epsilon) \exp \left(\frac{\pi \mathcal{Q}_n}{2 \epsilon} \right)
1025: \left[\frac{2 \cos k}{\omega + vq_n- 2 \cos(k-q_n)}
1026: \frac{1}{y^2}+\order{\frac{1}{y^3}}\right]
1027: \exp(-iq_ny)
1028: \label{cont1}
1029: \end{align}
1030: and
1031: \begin{align}
1032: w_s-w_u = &\sum_n
1033: \lim_{\epsilon \to 0}
1034: A_n^*(\epsilon) \exp \left(\frac{\pi \mathcal{Q}_n}{2 \epsilon} \right)
1035: \left[ 1+\order{\frac1y} \right]
1036: \exp(-iq_ny) \nonumber\\
1037: &- \sum_n
1038: \lim_{\epsilon \to 0}
1039: A_n(\epsilon) \exp \left(-\frac{\pi \mathcal{Q}_n}{2 \epsilon} \right)
1040: \left[\frac{2 \cos k}{\omega + vq_n- 2 \cos(k-q_n)}
1041: \frac{1}{y^2}+\order{\frac{1}{y^3}}\right]
1042: \exp(iq_ny).
1043: \label{cont2}
1044: \end{align}
1045: \end{subequations}
1046: \end{widetext}
1047: [Here we have used eq.\,\eqref{f1g2}.]
1048: Matching \eqref{cont1} to \eqref{expexpansiona} and \eqref{cont2}
1049: to \eqref{expexpansionb} yields then $K_0=0$ and
1050: \begin{subequations}
1051: \begin{eqnarray}
1052: \lim_{\epsilon \to 0} A_n(\epsilon) \exp\left(-\frac{\pi \mathcal{Q}_n}{2
1053: \epsilon} \right)=0,
1054: \label{Anex1} \\
1055: \lim_{\epsilon \to 0} A_n^*(\epsilon) \exp\left(\frac{\pi \mathcal{Q}_n}{2
1056: \epsilon} \right)=K_n
1057: \label{Anex2}
1058: \end{eqnarray}
1059: \end{subequations}
1060: for $n=1,2, \ldots $. We note that eq.\,\eqref{Anex1} follows from eq.\,\eqref{Anex2},
1061: while the latter equation can be written, symbolically,
1062: as
1063: \begin{equation}
1064: A_n(\epsilon) \longrightarrow
1065: K_n^* \exp\left(-\frac{\pi \mathcal{Q}_n}
1066: {2\epsilon} \right) \quad \text{as} \ \epsilon \to 0.
1067: \label{symb}
1068: \end{equation}
1069: Our subsequent efforts will focus on the evaluation of the
1070: constants $K_n$.
1071:
1072:
1073: For $k$ greater than $k^{(1)}_{\text{max}}$ (approximately $0.22$),
1074: there is only one radiation
1075: mode and therefore only one pre-exponential factor, $K_1$. For smaller
1076: $k$ we note that the
1077: amplitude of the $n$th radiation mode, $A_n$,
1078: is a factor of $\exp\left\{
1079: \frac{\pi}{2\epsilon} (\mathcal{Q}_n-\mathcal{Q}_1)\right\}$
1080: smaller than $A_1$, the amplitude of the first mode.
1081: Referring to fig.\,\ref{nicedisprelfig}, it is clear that for $n \geq 3$,
1082: the difference
1083: $\mathcal{Q}_n-\mathcal{Q}_1$ will be no smaller than $\pi$.
1084: As for the second mode, it becomes as significant as the first one
1085: only when $k=\order{\epsilon^2}$ in which case
1086: $(\mathcal{Q}_2-\mathcal{Q}_1)/\epsilon =\order{1}$.
1087: But in our asymptotic expansion of section
1088: \ref{section2} we assumed, implicitly, that
1089: $k$ is of order $1$ and so the case of $k =\order{\epsilon^2}$
1090: is beyond the scope of
1091: our current analysis. Therefore, for our purposes all
1092: the radiation modes with
1093: $n \geq 2$ (when they exist)
1094: will have
1095: negligible amplitudes compared to that of the first mode,
1096: provided $K_1$ is nonzero and $\epsilon$ is small.
1097: For this reason, we shall
1098: only attempt to evaluate $K_1$ in this paper.
1099:
1100:
1101: \subsection{Borel summability of the asymptotic series}
1102: \label{BLT}
1103:
1104: Pomeau, Ramani, and Grammaticos~\cite{pomeau}
1105: have shown that the radiation
1106: can be measured using the technique of
1107: Borel summation rather than by solving differential equations
1108: numerically,
1109: as in Segur and Kruskal's original approach. The method has been refined by
1110: (among others)
1111: Grimshaw and Joshi \cite{GrimshawJoshi, GrimshawNLS} and Tovbis,
1112: Tsuchiya, Jaff\'e
1113: and Pelinovsky \cite{Tovbis1, Tovbis2, Tovbis3, TovPel}, who have applied it to
1114: difference equations. Most recently it has been applied to
1115: differential-difference equations
1116: in the context of moving kinks in $\phi^4$ models \cite{OPB}.
1117: This is the approach that we will be pursuing here.
1118:
1119: Expressing $u(y)$ and $w(y)$ as Laplace transforms
1120: \begin{subequations}
1121: \label{laplace}
1122: \begin{equation}
1123: u(y) = \int_{\gamma}U(p)e^{-py}dp
1124: \end{equation}
1125: and
1126: \begin{equation}
1127: \quad w(y) = \int_{\gamma}W(p)e^{-py}dp,
1128: \end{equation}
1129: \end{subequations}
1130: where $\gamma$ is a contour extending
1131: from the origin to infinity in the complex $p$ plane,
1132: the inner equations \eqref{inner-equations} are cast in the
1133: form of integral equations
1134: \begin{subequations}
1135: \label{integral-equations}
1136: \begin{gather}
1137: f(p)U + \mu [f(p)U]\ast U\ast W + U\ast U\ast W = 0,
1138: \label{fpU} \\
1139: f(-p)W + \mu [f(-p)W]\ast W\ast U + W\ast W\ast U = 0,
1140: \label{gpU}
1141: \end{gather}
1142: \end{subequations}
1143: where
1144: \[
1145: f(p) = (\cosh p - 1)\cos k - i(\sinh p - p)\sin k.
1146: \]
1147: The asterisk $\ast$ denotes the convolution integral,
1148: \[ U(p) \ast W(p) = \int_0^p U(p-p_1) W(p_1) d p_1, \]
1149: where the integration is performed from the origin to the point $p$ on the
1150: complex plane, along the contour $\gamma$.
1151: In deriving eqs \eqref{integral-equations}, we
1152: have made use of the convolution theorem for the Laplace transform
1153: of the form \eqref{laplace}, where the integration is over
1154: a contour in the complex plane rather than a positive
1155: real axis. The theorem states that
1156: \begin{equation}
1157: u(y) w(y)= \int_\gamma [ U(p) * W(p) ] e^{-py} dp.
1158: \label{conv_thm}
1159: \end{equation}
1160: The proof of this theorem is provided in appendix \ref{A}.
1161:
1162:
1163:
1164: We choose the contour so that $\arg p \to \pi/2$ as $|p| \to
1165: \infty$ along $\gamma$. In this case we have
1166: $e^{-py} \to 0$ as $|p| \to \infty$ for all $y$ along any line
1167: $-\infty< \text{Re} \, y < \infty$ with $\text{Im} \, y<0$.
1168: Therefore, the integrals in \eqref{laplace} converge
1169: for all $y$ along this line
1170: and any bounded $U(p)$ and $W(p)$.
1171:
1172:
1173:
1174:
1175: The function $U(p)$ will have singularities
1176: at the points
1177: where $f(p)$ vanishes while the
1178: sum of the double-convolution terms in \eqref{fpU} does not. Similarly,
1179: $W(p)$ will have a singularity wherever $f(-p)$ vanishes
1180: [while the
1181: sum of the double-convolution terms in \eqref{gpU} does not].
1182: Therefore, $U$ and $W$ may have singularities at the points where
1183: \begin{equation}
1184: \label{singularities}
1185: \cosh p - 1 = \pm i \tan k \, (\sinh p - p),
1186: \end{equation}
1187: with the top and bottom signs referring to $U(p)$ and $W(p)$,
1188: respectively.
1189: The imaginary roots of eq.\,\eqref{singularities}
1190: with the top sign are at $p = -iq_n$
1191: and those of eq.\,\eqref{singularities}
1192: with the bottom sign at $p = iq_n$,
1193: where $q_n$ are the real roots of
1194: \eqref{leading-order-disp}.
1195: (We remind the reader that all roots $q_n$
1196: are positive.) The point $p = 0$ is not a singularity
1197: as both double-convolution terms in each line of
1198: \eqref{integral-equations} vanish here.
1199: There is always at least one pure imaginary root of eq.\,\eqref{singularities}
1200: (and \emph{only} one if $k > k^{(1)}_{\rm max}$, where $k^{(1)}_{\rm max} \approx 0.22$).
1201:
1202: In addition,
1203: there are infinitely many complex roots. The complex singularities of $U$
1204: [complex roots of the top-sign equation in \eqref{singularities}] are at the
1205: intersections of the curve given by
1206: \begin{subequations}
1207: \begin{equation}
1208: q = \frac{\cosh \kappa}{\sin k}
1209: \sqrt{1-\sin^2 k \left( \frac{\kappa}{\sinh \kappa} \right)^2 } -\cot k
1210: \label{curve1a}
1211: \end{equation}
1212: with the family of curves described by
1213: \begin{align}
1214: \label{curve2a}
1215: q &= k - {\rm arcsin} \, \left( \frac{\kappa}{\sinh \kappa} \sin k \right)
1216: + 2 \pi n, \\
1217: &\quad n=1,2, \ldots \, ,\nonumber
1218: \end{align}
1219: \end{subequations}
1220: and at the
1221: intersections of the curve
1222: \begin{subequations}
1223: \begin{equation}
1224: q = -\frac{\cosh \kappa}{\sin k}
1225: \sqrt{1- \sin^2 k \left(\frac{\kappa}{\sinh \kappa} \right)^2 } -\cot k
1226: \label{curve1b}
1227: \end{equation}
1228: with the family of curves described by
1229: \begin{align}
1230: \label{curve2b}
1231: q &= k + {\rm arcsin} \, \left( \frac{\kappa}{\sinh \kappa} \sin k \right)
1232: + \pi (2n+1), \\
1233: &\quad n=-2,-3,-4, \ldots \, .\nonumber
1234: \end{align}
1235: \end{subequations}
1236: Here $\kappa$ and $q$ are the real and imaginary part of $p$:
1237: $p=\kappa+iq$.
1238: The curve \eqref{curve1a} looks like a parabola
1239: opened upwards, with the vertex at $\kappa=q=0$, and the
1240: curve \eqref{curve1b} like a parabola opened
1241: downwards, with the vertex at $\kappa=0$, $q=-2 \cot k$. The curves
1242: \eqref{curve2a} and \eqref{curve2b},
1243: on the other hand, look like parabolas for small $\kappa$
1244: but then flatten out
1245: and approach horizontal straight lines as
1246: $\kappa \to \pm \infty$.
1247: [In compiling the list of these `flat' curves in \eqref{curve2a} and
1248: \eqref{curve2b}, we have taken into account
1249: that the curve \eqref{curve2a} with $n=0$ does not have
1250: any intersections with the parabola \eqref{curve1a}, and
1251: the curve \eqref{curve2b} with $n=-1$ does not have
1252: any intersections with the parabola \eqref{curve1b}.]
1253: As $k$ is reduced, the vertex of the
1254: parabola
1255: \eqref{curve1b} moves down along the $q$ axis; the
1256: intersections of this parabola with the `flat' curves
1257: \eqref{curve2b} approach, pairwise, the $q$ axis. After colliding on
1258: the $q$ axis, pairs of complex roots move away from
1259: each other along it.
1260: [The parabola \eqref{curve1a}
1261: does not move as $k$ is reduced, but only steepens,
1262: which results in the singularities in the upper half-plane approaching the
1263: imaginary axis but not reaching it until $k = 0$.]
1264: In a similar way, the complex singularities
1265: of $W(p)$ move onto the imaginary axis as $k$ is decreased.
1266: In section \ref{BLT} below we will use the fact that the
1267: distance from any complex
1268: singularity to the origin is larger than $2 \pi$;
1269: this follows from the observation that
1270: the closest points of the curves \eqref{curve2a} and
1271: \eqref{curve2b} to the origin are their intersections with
1272: the $q$ axis. These are further away than $2 \pi$ from the origin.
1273:
1274:
1275:
1276: In addition to singularities at $p= -iq_n$, the function $U(p)$ will
1277: have singularities at points $p=iq_n$, $n=1,2,\ldots$. These are
1278: induced by the cubic terms in eq.\,\eqref{fpU}; for instance,
1279: the singularity at $p=iq_1$ arises from the convolution of
1280: the term proportional to $p$ in $U*U$ with
1281: the function $W$ which has a
1282: singularity at $p=iq_1$. [That $U(p)$ has singularities at
1283: $p=iq_n$ can also be seen directly from eq.\,\eqref{expexpansiona}.]
1284: Similarly, the function $W(p)$ will
1285: have singularities at points $p=-iq_n$, $n=1,2,\ldots$.
1286: By virtue of the nonlinear terms there will also be singularities
1287: at the `combination points' $i(\pm q_n \pm q_m)$, $i(\pm q_n \pm q_m \pm
1288: q_j)$, etc.
1289:
1290: The formal inverse-power series \eqref{pertexpy}, which
1291: we can write as
1292: \begin{equation}
1293: \hat{u}(y)= \sum_{\ell=0}^\infty \rho_\ell \, \frac{\ell !}{y^{\ell+1}},
1294: \quad
1295: \hat{w}(y)= \sum_{\ell=0}^\infty \nu_\ell \,
1296: \frac{\ell !}{y^{\ell+1}},
1297: \label{uywy}
1298: \end{equation}
1299: result from
1300: the Laplace transformation of power series for $U(p)$ and $W(p)$:
1301: \begin{equation}
1302: U(p)= \sum_{\ell=0}^\infty \rho_\ell \, p^\ell,
1303: \quad
1304: W(p)= \sum_{\ell=0}^\infty \nu_\ell \, p^\ell.
1305: \label{UpWp}
1306: \end{equation}
1307: The series \eqref{UpWp} converge in the disk of radius
1308: $q_1$, centred at the origin, and hence can be integrated term by term
1309: only over the portion of the contour $\gamma$ which lies within that disk.
1310: However, by Watson's lemma, the remaining part of the contour
1311: makes an exponentially small contribution to the integral and
1312: the resulting series \eqref{uywy} are asymptotic as $y \to \infty$.
1313: The functions
1314: $u(y)$ and $w(y)$ defined by eqs \eqref{laplace} give the Borel
1315: sums of the series $\hat{u}(y)$ and $\hat{w}(y)$.
1316:
1317:
1318: Consider now some horizontal line in the inner region; that is, let
1319: $\text{Im} \,y<0$ be fixed and $\text{Re} \,y$ vary from $-\infty$
1320: to $\infty$.
1321: If the integration contour $\gamma$ is chosen to lie
1322: in the first quadrant of the complex $p$ plane, the functions $u(y)$ and $w(y)$ generated
1323: by eqs \eqref{laplace} will tend to zero as $\Real y\to +\infty$
1324: along this line. Similarly,
1325: if it is chosen to lie in the second quadrant, they will tend
1326: to zero as $\Real y\to -\infty$. Suppose there were no singularities
1327: between two such contours: then the one could be continously
1328: deformed to the other
1329: without any singularity crossings; i.e.\ they would
1330: generate the same solution which,
1331: therefore, would decay to zero at
1332: both infinities. [That is, the oscillatory tails in
1333: \eqref{expexpansion} would have zero amplitudes, $K_n=0$.]
1334: In general, however, $U(p)$ and $W(p)$ have singularities
1335: both on and away from the imaginary axis. In order to minimise the
1336: number of singularities to be crossed in the deformation
1337: of one contour to the other, we choose the contours to lie above
1338: all singularities with nonzero real part.
1339: (That this is possible,
1340: is shown in appendix \ref{B}.)
1341: Note also that the imaginary part of the singularity
1342: grows faster than its real part and
1343: hence $\arg p$ should tend to $\pi/2$ as $|p| \to
1344: \infty$ along $\gamma$; this was precisely our choice for
1345: the direction of the contours $\gamma$ in the beginning
1346: of section \ref{BLT}.
1347:
1348:
1349:
1350:
1351: \begin{figure}[btp]
1352: \includegraphics{contours.eps}
1353: \caption{\label{integration-contours}The integration contours
1354: $\gamma_s$ and $\gamma_u$ used to generate the solutions
1355: $(u_s, w_s)$ and $(u_u, w_u)$ respectively via eqs \eqref{laplace}.
1356: The dots are singularities of $W(p)$. Shown is the
1357: situation where the linear dispersion relation \eqref{leading-order-disp}
1358: has only one real root, $q_1$.}
1359: \end{figure}
1360:
1361: Let $\gamma_s$ and $\gamma_u$ be two contours chosen in this
1362: way, with
1363: $\gamma_s$ lying in the first quadrant and $\gamma_u$ in the
1364: second quadrant.
1365: The solutions $u(y), w(y)$ generated by eqs \eqref{laplace} with the contour
1366: $\gamma_s$ will
1367: tend to zero as $\Real y \to \infty$
1368: (with $\text{Im} \, y <0$ fixed). Hence they can be identified with solutions
1369: $u_s, w_s$ obtained by the continuation of the outer solution $\psi_s$
1370: which has the
1371: same asymptotic behaviour.
1372: Similarly, the solutions generated by eqs \eqref{laplace} with the contour $\gamma_u$
1373: coincide with solutions $u_u$, $w_u$ -- like $u_u$, $w_u$, the solutions
1374: generated by eqs \eqref{laplace} tend to zero
1375: as $\Real y \to -\infty$ (with fixed $\Imag y <0$).
1376:
1377: Consider, first, solutions $w_s$ and
1378: $w_u$.
1379: Since the contours $\gamma_s$ and $\gamma_u$
1380: are separated by singularities of $W(p)$ on
1381: the positive imaginary axis, they
1382: cannot be
1383: continuously deformed to each other
1384: without singularity crossings and so
1385: the solution $w_s$ does not coincide with $w_u$,
1386: unless the residue at the singularity happens to be zero.
1387: If we deform $\gamma_s$ to $\gamma_s'$ and $\gamma_u$ to
1388: $\gamma_u'$ as shown in fig.\,\ref{integration-contours}, without
1389: crossing any singularities, then the only difference between
1390: the two contours is that $\gamma_s'$ encircles the singularities,
1391: whereas $\gamma_u'$ does not. Therefore, the difference $w_s-w_u$
1392: can be deduced exclusively from the leading-order behaviour of $W(p)$
1393: near its singularities.
1394: There can be two contributions to this difference:
1395: the first arises from integrating around the poles and
1396: is a sum of residues, while the second arises if the singularity is a
1397: branch point.
1398:
1399: To find the singularity structure of the function $W(p)$, we equate
1400: \begin{equation}
1401: w_s-w_u= \int_{\gamma_s'} W(p) e^{-py} dp- \int_{\gamma_u'} W(p) e^{-py} dp
1402: \label{wswu}
1403: \end{equation}
1404: to the expansion \eqref{expexpansionb}. The
1405: first term in \eqref{expexpansionb}, $e^{-i q_ny}$, arises
1406: from the integration of a term $(2 \pi i)^{-1} (p-i q_n)^{-1}$ in
1407: $W(p)$. For such a term, the difference of the two integrals in
1408: \eqref{wswu} reduces to an integral around a circle
1409: centered on the point $p=iq_n$:
1410: \[
1411: \frac{1}{2 \pi i} \oint \frac{1}{p-iq_n} e^{-py} dp=
1412: \text{res} \left\{ \frac{e^{-py}}{p-iq_n}, iq_n \right\}.
1413: \]
1414: The term $y^{-m}e^{-iqy_n}$ in \eqref{expexpansionb}, with $m=1,2,\ldots$,
1415: arises from the integration of a term
1416: \[
1417: \frac{1}{2 \pi i} \frac{(p-iq_n)^{m-1}}{(m-1)!} \ln (p-iq_n)
1418: \]
1419: in $W(p)$. This time, $p=iq_n$ is a branch point. After going around
1420: this point along the circular part of $\gamma_s'$, the logarithm
1421: increases by $2 \pi i$ and the difference between the integrals in
1422: \eqref{wswu} is given by
1423: \begin{align*}
1424: &\frac{1}{(m-1)!} \int_C (p-iq_n)^{m-1} e^{-py} dp \\
1425: &\quad = \frac{e^{-i q_n y}}{(m-1)!} \int_0^\infty z^{m-1} e^{-zy}dz,
1426: \end{align*}
1427: where $C$ is the part of $\gamma_s'$ extending from
1428: $p=iq_n$ to infinity. This equals exactly
1429: $y^{-m} e^{-i q_n y}$.
1430:
1431: Thus, in order to generate the full series \eqref{expexpansionb}
1432: we must have
1433: \begin{multline}
1434: \label{Wsingularities}
1435: W(p) =
1436: \frac{1}{2 \pi i} \sum_n K_n
1437: \Bigg[
1438: \frac{1}{p-iq_n}
1439: + \sum_{m=1}^\infty \frac{c_m^{(n)}}{(m-1)!}\\
1440: \times (p-iq_n)^{m-1} \ln(p-iq_n) \Bigg]
1441: + W_{\rm reg}(p),
1442: \end{multline}
1443: where $W_{\rm reg}$ denotes the part of $W$ which is regular
1444: at $p=iq_n$, $n=1,2,\ldots$.
1445: By the same process, matching $u_s-u_u$ to $\delta u$ in
1446: \eqref{expexpansiona} yields
1447: \begin{multline}
1448: \label{Usingularities}
1449: U(p) =
1450: \frac{1}{2 \pi i} \sum_n K_n
1451: \sum_{m=1}^\infty \frac{d_m^{(n)}}{(m-1)!}\\
1452: \times (p-iq_n)^{m-1} \ln(p-iq_n)
1453: + U_{\text{reg}}(p).
1454: \end{multline}
1455:
1456: The solution to eqs \eqref{integral-equations} is nonunique;
1457: for instance, if $\{ U(p), W(p) \}$ is a solution,
1458: then so is $\{ e^{py_0+ \zeta_0} U(p), e^{py_0 - \zeta_0}W(p) \}$
1459: with any complex $y_0$ and $\zeta_0$. Also,
1460: if $\{ U(p), W(p) \}$ is a solution, $\{ W(-p), U(-p) \}$ is
1461: another one. We will impose the constraint
1462: \begin{equation}
1463: U(p)= W(-p);
1464: \label{constraint}
1465: \end{equation}
1466: this constraint is obviously compatible with
1467: eqs \eqref{integral-equations}.
1468: It is not difficult to see that the reduction \eqref{constraint}
1469: singles out a unique solution of eqs \eqref{integral-equations}.
1470: The motivation for imposing the constraint \eqref{constraint}
1471: comes from the symmetry $\hat{u}(-y)=-\hat{w}(y)$ of the power-series
1472: solution of eq.\,\eqref{inner-equations}. Using this symmetry in
1473: eq.\,\eqref{uywy}, we get $\rho_\ell= (-1)^\ell \nu_\ell$ and
1474: then eq.\,\eqref{UpWp} implies \eqref{constraint}.
1475:
1476: In view of \eqref{constraint}, the singularities of $U(p)$ in the upper
1477: half-plane are singularities of $W(p)$ in the lower half-plane,
1478: which fall within $W_{\text{reg}}(p)$, and vice versa.
1479: Thus we have, finally,
1480: \begin{multline}
1481: \label{allWsingularities}
1482: W(p) =
1483: \frac{1}{2 \pi i} \sum_n
1484: \frac{K_n }{p-iq_n}
1485: + \frac{1}{2 \pi i} \sum_n K_n
1486: \sum_{m=1}^\infty \frac{1}{(m-1)!} \\
1487: \shoveleft{\times\Big[ c_m^{(n)}(p-iq_n)^{m-1} \ln(p-iq_n)}\\
1488: -(-1)^m d_m^{(n)}
1489: (p+iq_n)^{m-1} \ln(p+iq_n) \Big]
1490: + \tilde{W}_{\text{reg}}(p),
1491: \end{multline}
1492: where $\tilde{W}_{\text{reg}}(p)$ is regular at $p = \pm iq_n$.
1493: We also mention an equivalent representation for
1494: \eqref{allWsingularities} which turns out to be
1495: computationally advantageous:
1496: \begin{multline}
1497: \label{allWsingularities_int}
1498: W(p) =
1499: \frac{1}{2 \pi i} \sum_n K_n
1500: \sum_{m=0}^\infty
1501: D^{-m} \left[\frac{c_m^{(n)}}{p-iq_n} -
1502: \frac{ (-1)^m d_m^{(n)}}{p+iq_n}
1503: \right] \\
1504: + \check{W}_{\text{reg}}(p).
1505: \end{multline}
1506: Here $D^{-1}$ is an integral map:
1507: \[
1508: D^{-1} f(p) \equiv \int_0^p f(p_1) dp_1;
1509: \]
1510: the notation $D^{-m} f(p)$ should be understood as
1511: \[
1512: D^{-m} f(p) \equiv \int_0^p \! dp_1 \int_0^{p_1} \! dp_2 \int_0^{p_2}
1513: \! dp_3 \cdots \int_0^{p_{m-1}} \! dp_m f(p_m).
1514: \]
1515: We have also introduced $c_0^{(n)} = 1$ and
1516: $d_0^{(n)} = 0$ for economy of notation.
1517: The only difference between eqs \eqref{allWsingularities} and
1518: \eqref{allWsingularities_int} is that the double-sum term on the right-hand
1519: side of \eqref{allWsingularities_int} includes some terms
1520: which are regular at $p=\pm iq_n$, whereas in
1521: eq.\,\eqref{allWsingularities}, all regular terms are contained in
1522: $\tilde{W}_{\text{reg}}(p)$.
1523:
1524: The residues $K_n$ at the poles of $W(p)$ are known
1525: as the Stokes constants. The
1526: leading-order Stokes constant $K_1$
1527: can be related to the behaviour of the
1528: coefficients in the power-series expansion of $W(p)$.
1529: Indeed, the coefficients in the power series \eqref{UpWp} satisfy
1530: \begin{equation}
1531: \label{nudef}
1532: \nu_{\ell} \longrightarrow K_1 \sum_{m=0}^{\ell}\frac{c^{(1)}_m +
1533: (-1)^{\ell}d^{(1)}_m}{2\pi q_1
1534: (i q_1)^{\ell-m}}\frac{(\ell-m)!}{\ell!}
1535: \quad \text{as } \ell \to \infty.
1536: \end{equation}
1537: This is obtained by expanding the
1538: singular part of the expression \eqref{allWsingularities_int}
1539: in powers of $p$. (Coefficients of the regular part become
1540: negligible in the limit $\ell\to\infty$ compared to those of the
1541: singular part.) Note that we have ignored singularities
1542: with nonzero real part and singularities on the imaginary axis
1543: other than at $p = \pm iq_1$. The reason is that all these
1544: singularities are further away from the origin than the points
1545: $\pm iq_1$
1546: (in particular all complex singularities are separated
1547: from the origin by a distance greater than $2 \pi$),
1548: and their contribution to $\nu_{\ell}$ becomes
1549: vanishingly small as $\ell \to \infty$.
1550: We have also neglected singularities at the `combination points'
1551: because of their exponentially small residues.
1552: The coefficients $\nu_{\ell}$ can be calculated numerically; once they
1553: are known, it follows from \eqref{nudef} that
1554: \begin{equation}
1555: \label{limit}
1556: K_1 = 2\pi q_1\lim_{\ell\to\infty}(iq_1)^{\ell}\nu_{\ell}
1557: \end{equation}
1558: (where we have recalled that $c_0^{(1)} = 1$ and $d_0^{(1)} = 0$).
1559:
1560: We now turn to the numerical calculation of the coefficients $\nu_{\ell}$.
1561:
1562: \subsection{Recurrence relation}
1563:
1564:
1565: To make our forthcoming
1566: numerical procedure more robust,
1567: we normalise the coefficients in the
1568: power series
1569: \eqref{UpWp} by writing
1570: \begin{equation}
1571: \nu_\ell = -i\frac{\delta_\ell}{(iq_1)^\ell}.
1572: \label{delta}
1573: \end{equation}
1574: Substituting the expansions \eqref{UpWp} with eq.\,\eqref{delta}
1575: as well as the constraint \eqref{constraint} into
1576: either of equations
1577: \eqref{integral-equations} and equating coefficients of
1578: like powers of $p$, yields the following
1579: recurrence relation for the numbers $\delta_n$
1580: ($n \ge 0$):
1581: \begin{multline}
1582: \sum_{m=0}^n \frac{q_1^m\delta_{n-m}}{(m+2)!}{\cal R}_m
1583: = \frac{1}{(n+2)(n+1)}\\
1584: \shoveleft{\times \sum_{m=0}^n\left[\delta_{n-m}
1585: +\mu\sum_{j=2}^{n-m}\frac{q_1^j\delta_{n-m-j}}{j!}
1586: {\cal R}_j \right]} \\
1587: \times\left(\sum_{j=0}^m (-1)^{m-j}\delta_{m-j}
1588: \delta_j\frac{j!(m-j)!}{m!}\right)\frac{m!(n-m)!}{n!}.
1589: \label{recu}
1590: \end{multline}
1591: Here
1592: \[
1593: {\cal R}_m= \Bigg\{
1594: \begin{array}{ll}
1595: (-1)^{\frac{m}{2}}\cos k, & \text{for } m \text{ even}, \\
1596: (-1)^{\frac{m-1}{2}}\sin k, & \text{for } m \text{ odd}.
1597: \end{array}
1598: \]
1599: Solving eq.\,\eqref{recu} with $n = 0$ gives $\delta_0 = \sqrt{1-c^2}$.
1600: Thereafter it can be solved for each member of the sequence
1601: $\{ \delta_n \}$ in terms of the preceding ones, and thus each $\delta_n$
1602: can be calculated in turn.
1603: [Since all the coefficients in the recurrence relation \eqref{recu}
1604: are real, the sequence $\{\delta_n\}$ turns out to be a sequence
1605: of real numbers.]
1606:
1607:
1608: Once the sequence $\{ \delta_n \}$ has been generated,
1609: expression \eqref{limit} can be
1610: used to calculate the Stokes constant $K_1$:
1611: \begin{equation}
1612: K_1 = -2 \pi i q_1 \delta_\ell,
1613: \label{limmo}
1614: \end{equation}
1615: for sufficiently large $\ell$.
1616: Unfortunately, the convergence of the sequence $\{\delta_n\}$
1617: is slow and thus
1618: the above procedure is computationally expensive.
1619: The convergence can be accelerated by
1620: expanding eq.\,\eqref{nudef} in powers of
1621: small $1/\ell$:
1622: \[
1623: \delta_{\ell} = \frac{iK_1}{2\pi q_1}\!\!\left[1 +iq_1 \frac{c_1^{(1)}}{\ell}
1624: - q_1^2\frac{c_2^{(1)}+(-1)^\ell d_2^{(1)}}{\ell^2} +
1625: \order{\frac{1}{\ell^3}}\right],
1626: \]
1627: whence
1628: \begin{multline}
1629: K_1 = -2\pi i q_1 \delta_{\ell} \Bigg[1 -iq_1 \frac{c_1^{(1)}}{\ell} \\
1630: + q_1^2\frac{ c_2^{(1)}-(c_1^{(1)})^2 +(-1)^\ell d_2^{(1)} }{\ell^2}
1631: + \order{\frac{1}{\ell^3}}\Bigg]. \label{K1}
1632: \end{multline}
1633: According to eq.\,\eqref{K1}, eq.\,\eqref{limmo} gives $K_1$ with
1634: a relatively large error of order $1/\ell$.
1635: On the other hand, a two-term approximation
1636: \begin{equation}
1637: \label{betterlimit}
1638: K_1 = -2\pi i q_1 \delta_{\ell}\left(1 -iq_1 \frac{c_1^{(1)}}{\ell} \right)
1639: \end{equation}
1640: is correct to $\mathcal{O} (1/\ell^2)$. More precisely, the
1641: relative error associated with the answer \eqref{betterlimit} is given by
1642: \begin{equation}
1643: \frac{{\mathcal E}}{K_1} =q_1^2 \frac{c_2^{(1)}-(c_1^{(1)})^2 +(-1)^\ell
1644: d_2^{(1)}}{\ell^2}.
1645: \label{relati}
1646: \end{equation}
1647: In our calculations, we set $\mathcal{E}/K_1=10^{-5}$.
1648: Since $c_1^{(1)}$, $c_2^{(1)}$
1649: and $d_2^{(1)}$ are known constants [given by
1650: \eqref{c1c2} and \eqref{d1n}], eq.\,\eqref{relati} tells us what
1651: $\ell$ we should take -- i.e.\ how many members of the sequence $\{\delta_n\}$
1652: we should calculate in order to achieve the set accuracy.
1653: Figure \ref{series} illustrates the
1654: convergence of the
1655: approximate values of $K_1$ calculated using \eqref{limmo} and
1656: \eqref{betterlimit} as $\ell$ is increased. Note the drastic
1657: acceleration of convergence
1658: in the latter case.
1659:
1660: \begin{figure}[tbp]
1661: \includegraphics{dnlseg.eps}
1662: \caption{\label{series}Convergence of the
1663: sequence on the right-hand side of \eqref{limmo} (dashed line) and
1664: the `accelerated' sequence defined by
1665: the right-hand side of \eqref{betterlimit}
1666: (solid line).
1667: Shown are the $\ell$th approximations to the Stokes constant $K_1$
1668: [the $\ell$th members of the sequences \eqref{limmo}
1669: and \eqref{betterlimit}]
1670: divided by $i$ to get a real value.
1671: In this plot, $\mu = 0$ and $k = 0.5$.
1672: }
1673: \end{figure}
1674:
1675: \begin{figure}[tbp]
1676: \includegraphics{dnlssat.eps}
1677:
1678: \medskip
1679: \includegraphics{dnlssatmulti.eps}
1680: \caption{\label{stokes}The Stokes constant $K_1$ for various
1681: values of $\mu$.
1682: Note the logarithmic scale on the vertical axis.
1683: The
1684: downward spikes extend all the way
1685: to $-\infty$; hence each spike corresponds to a zero crossing.
1686: Each panel shows only a portion of the full range
1687: $0 \leq k \leq \pi/2$; there
1688: are no additional zero crossings in the part which is not shown.
1689: }
1690: \end{figure}
1691:
1692: Figure \ref{stokes}(a) shows the calculated Stokes constant as a
1693: function of $k$ for various values of
1694: the saturation parameter $\mu$.
1695: First of all, $K_1(k)$
1696: does not have any zeros
1697: in the case of the cubic nonlinearity ($\mu=0$).
1698: This means that solitons of the
1699: cubic one-site discrete NLS equation [eq.\,\eqref{DNLS} with $\mu=0$]
1700: cannot propagate without losing energy to radiation.
1701: For $\mu = 3$ the Stokes constant
1702: does have a zero, but at a value of $k$ smaller than $k^{(1)}_{\rm max}$
1703: (where $k^{(1)}_{\rm max}\approx0.22=0.07\pi$).
1704: Since higher radiation modes do exist in this range of $k$,
1705: there will still be radiation from the soliton -- unless the
1706: `higher' Stokes constants $K_2(k)$, $K_3(k)$,\ldots, happen to be zero
1707: at the same value of $k$.
1708: Finally,
1709: for $\mu = 4$ the zero is seen to have moved just above $k^{(1)}_{\rm max}$ and
1710: for $\mu = 6$ it has an even higher value.
1711: There are no $\mathcal{Q}_2, \mathcal{Q}_3,\ldots$ radiations
1712: for these $k$; hence the zeros of $K_1(k)$ define
1713: the carrier wavenumbers at which the soliton `slides' --
1714: i.e.\ travels without emitting any radiation. Equation \eqref{om_and_v_via_eps}
1715: then gives the corresponding sliding velocities, for each $\epsilon$.
1716:
1717: Figure \ref{stokes}(b) shows the Stokes constant $K_1(k)$
1718: for higher values of the
1719: parameter $\mu$. For $\mu = 12$ a second zero of the
1720: Stokes constant has appeared while for $\mu = 25$,
1721: the function $K_1(k)$ has three zeros.
1722: As $\mu$ is increased, the existing zeros move to
1723: larger values of $k$ while new
1724: ones emerge at the origin of the $k$ axis.
1725:
1726:
1727:
1728: \subsection{Radiation waves}
1729:
1730: For not very large $|X|$, the solution $\psi_s$ is close to the localised pulse
1731: found by means of the perturbation expansion
1732: in section \ref{section2}. As $X \to +\infty$, it tends
1733: to zero, by definition, while the $X \to -\infty$ asymptotic
1734: behaviour is found from $\psi_s = \psi_u + \Psi$.
1735: Here the solution $\psi_u$ decays to zero as $X \to
1736: -\infty$ and hence
1737: $\psi_s$ approaches the oscillatory waveform
1738: $\Psi$
1739: given by eqs \eqref{radansatz}, \eqref{f1g2}, and \eqref{symb}:
1740: \begin{multline}
1741: \label{final-solution}
1742: \psi_s(X) \to \sum_n K_n^*e^{-\pi \mathcal{Q}_n/2\epsilon} \\
1743: \times\left[1 +\epsilon
1744: \frac{4 i\cos k\tan X}{2\sin(k+\mathcal{Q}_n)-v} +
1745: \order{\epsilon^2}\right]e^{i\mathcal{Q}_n X/\epsilon}
1746: \end{multline}
1747: as $X\to -\infty$.
1748: Equation \eqref{final-solution} describes a radiation background over which
1749: the soliton is superimposed.
1750: As we have explained, we can ignore all but the first term in the
1751: sum.
1752:
1753: To determine whether the radiation is emitted by the soliton
1754: or being fed into it from outside sources,
1755: we
1756: consider a harmonic solution $\psi = e^{i\mathcal{Q}X/\epsilon-i\Omega t}$
1757: of the linearised eq.\,\eqref{timedepeqn}; the
1758: corresponding dispersion relation is
1759: \begin{equation*}
1760: \Omega(\mathcal{Q}) = -2\cos(k+\mathcal{Q}) + \omega - \mathcal{Q}v.
1761: \end{equation*}
1762: The radiation background $\Psi$ consists of
1763: harmonics with
1764: $\Omega = 0$ and $\mathcal{Q} = \mathcal{Q}_n$
1765: where $\mathcal{Q}_n$ are roots of eq.\,\eqref{disprel}.
1766: The group velocities of these harmonic waves are given by
1767: \begin{equation}
1768: \Omega'(\mathcal{Q}_n) = 2\sin(k+\mathcal{Q}_n)-v.
1769: \label{vgr}
1770: \end{equation}
1771: The $\mathcal{Q}_n$'s are zeros of the function $\Omega(\mathcal{Q})$
1772: and the group velocities are the slopes of this function at its zeros;
1773: thus the group velocities
1774: $\Omega'(\mathcal{Q}_1)$, $\Omega'(\mathcal{Q}_2)$,\ldots,
1775: have alternating signs. The first one,
1776: which is the only one that concerns us in this work,
1777: must be negative. Indeed, the value
1778: \[
1779: \Omega(0) = -2\cos k+\omega = 2\cos k (\cosh\epsilon-1)
1780: \]
1781: is positive, and hence the
1782: slope of the
1783: function $\Omega(\mathcal{Q})$ as it crosses the $\mathcal{Q}$ axis
1784: at $\mathcal{Q}_1>0$ is negative.
1785: Therefore, the first radiation
1786: mode, extending to $-\infty$, carries energy {\it away\/} from the
1787: soliton.
1788:
1789: The even-numbered radiation modes (where present) in our
1790: asymptotic solution \eqref{final-solution} have positive group
1791: velocities
1792: and hence describe the flux of energy fed into
1793: the system at the left infinity.
1794: A more interesting situation is obviously the one with
1795: no incoming radiation; the corresponding solution is obtained by
1796: subtracting off the required multiple of the solution of the
1797: linearised equation, e.g.\ $e^{i\mathcal{Q}_2X/\epsilon}$. One would then
1798: have a pulse leaving the odd modes in its wake and sending even
1799: modes ahead of it.
1800:
1801: If the first Stokes constant $K_1(k)$ has a zero at
1802: some $k=k_1$ while $\mathcal{Q}_1$ is the only
1803: radiation mode available (as happens in our
1804: saturable model with $\mu$ greater than
1805: approximately $4$), then according to
1806: eq.\,\eqref{final-solution}, the radiation from
1807: the soliton with the carrier-wave wavenumber $k_1$ is
1808: suppressed completely.
1809:
1810:
1811:
1812: \section{Time evolution of a radiating soliton}
1813: \label{section4}
1814:
1815: \subsection{Amplitude-wavenumber dynamical system}
1816:
1817: To find the radiation-induced evolution of the
1818: travelling soliton, we use conserved quantities of the advance-delay
1819: equation associated with eq.\,\eqref{DNLS}.
1820: In the reference frame moving at the
1821: soliton velocity $v$ this equation reads
1822: \begin{equation}
1823: \label{timedepeqn2}
1824: i\varphi_t +\varphi(x+1,t) + \varphi(x-1,t) -iv \varphi_x
1825: +\frac{2|\varphi|^2\varphi}{1+\mu|\varphi|^2} = 0.
1826: \end{equation}
1827: The discrete variable $\phi_n(t)$ in eq.\,\eqref{DNLS} is related to
1828: the value of the continuous variable $\varphi(x,t)$ at the
1829: point $x=n-vt$: $\phi_n(t) = \varphi(n-vt,t)$.
1830: For future use, we also mention the relation between $\varphi(x,t)$ and
1831: the corresponding solution of eq.\,\eqref{timedepeqn}:
1832: \begin{equation}
1833: \varphi(x,t)= \psi(X,t) e^{ik(x+vt)+ i \omega t}.
1834: \label{corresp}
1835: \end{equation}
1836:
1837:
1838: We first consider the number of particles integral:
1839: \[ N = \int_{a}^{b}|\varphi|^2dx. \]
1840: Multiplying eq.\,\eqref{timedepeqn2}
1841: by $\conj{\varphi}$, subtracting the complex
1842: conjugate and integrating yields the rate of change of the integral $N$:
1843: \begin{align}
1844: \label{mainN}
1845: i\frac{dN}{dt} &= \int_{a-1}^{a} (\varphi^+\conj{\varphi} - c.c.) dx \notag\\
1846: &\quad + \int_b^{b+1} (\varphi^-\conj{\varphi} - c.c.) dx
1847: + iv|\varphi|^2 \Big|_{a}^b.
1848: \end{align}
1849: In eq.\,\eqref{mainN}, $\varphi^\pm \equiv \varphi(x \pm 1,t)$
1850: and $c.c.$ stands for the complex conjugate of the
1851: immediately preceding term.
1852: The integration limits $a<0$ and $b>0$ are
1853: assumed to be large ($|a|,b \gg \epsilon^{-1}$)
1854: but finite; for example one can take $a,b= {\cal O}(\epsilon^{-2})$.
1855:
1856:
1857: Consider the soliton moving with a positive velocity
1858: and leaving radiation in its wake. This
1859: configuration is described by the
1860: solution $\psi_s$ of eq.\,\eqref{timedepeqn}; the corresponding
1861: solution of eq.\,\eqref{timedepeqn2} has the asymptotic behaviour
1862: $\varphi_s \to 0$ as $x \to +\infty$.
1863: Substituting the leading-order expression \eqref{stable-minus-unstable-outer}
1864: for the soliton's radiation tail into eq.\,\eqref{corresp} yields
1865: the asymptotic behaviour at the other infinity:
1866: \begin{equation}
1867: \varphi_s(x,t) \to \sum_n A_n[1+\order{\epsilon}]e^{i(k+{\cal Q}_n)x+i(\omega+kv)t}
1868: \label{asympie}
1869: \end{equation}
1870: as $x \to -\infty$.
1871: Since, as we have explained above, the $\mathcal{Q}_1$ radiation is dominant, it
1872: is sufficient to keep only the $n=1$ term in \eqref{asympie}.
1873:
1874: Substituting \eqref{asympie} into \eqref{mainN} and evaluating the
1875: integral over the region $(a-1,a)$ in the soliton's wake,
1876: we obtain
1877: \begin{equation}
1878: \label{Ndot}
1879: \frac{dN}{dt} =
1880: |K_1|^2e^{-\pi \mathcal{Q}_1/\epsilon}[2\sin(k+\mathcal{Q}_1) - v],
1881: \end{equation}
1882: where we have used eq.\,\eqref{symb}.
1883: Note that $[2\sin(k+\mathcal{Q}_1) - v]$ is the group velocity
1884: of the first radiation
1885: mode, $\Omega'(\mathcal{Q}_1)$, which, as we have established, is negative;
1886: hence ${\dot N} \leq 0$.
1887:
1888: We now turn to the momentum integral,
1889: \[
1890: P =
1891: \frac{i}{2}\int_{a}^b (\conj{\varphi}_x\varphi-\varphi_x\conj{\varphi}) dx.
1892: \]
1893: Multiplying \eqref{timedepeqn2} by $\conj{\varphi}_x$, adding its complex
1894: conjugate and integrating, gives the rate equation
1895: \begin{multline}
1896: \label{mainP}
1897: \frac{dP}{dt} =
1898: \int_{a-1}^{a} (\varphi^+\conj{\varphi}_x + c.c.) dx
1899: -\int_b^{b+1} (\varphi^-\conj{\varphi}_x + c.c.) dx \\
1900: + \frac{i}{2} (\varphi \conj{\varphi_t}
1901: -\conj{\varphi} \varphi_t) \Big|_{a}^b
1902: - (\varphi^+)^* \varphi \Big|_{a-1}^b
1903: - (\varphi^-)^* \varphi \Big|_{a}^{b+1} \\
1904: -\left[\frac{2}{\mu}|\varphi|^2
1905: -\frac{2}{\mu^2}\ln\left(1+\mu|\varphi|^2\right)\right]_{a}^b.
1906: \end{multline}
1907: Evaluating the right-hand side of
1908: \eqref{mainP} similarly to the
1909: way we obtained \eqref{Ndot} and substituting eq.\,\eqref{disprel}
1910: for $\omega$,
1911: produces
1912: \begin{equation}
1913: \label{Pdot}
1914: \frac{dP}{dt} =
1915: |K_1|^2 e^{-\pi \mathcal{Q}_1/\epsilon}(k+\mathcal{Q}_1)[2\sin(k+\mathcal{Q}_1) - v].
1916: \end{equation}
1917:
1918: Using the leading-order term of the perturbative solution
1919: \eqref{regularpert} and eq.\,\eqref{corresp},
1920: we can express $N$ and $P$
1921: via $\epsilon$ and $k$:
1922: \begin{eqnarray*}
1923: N =2\epsilon\cos k + \order{\epsilon^2}, \\
1924: P = 2k \epsilon\cos k + \order{\epsilon^2}.
1925: \end{eqnarray*}
1926: In calculating $N$ and $P$
1927: we had to integrate from $X =\epsilon a$ to $X=\epsilon b$.
1928: Since the integrands decay exponentially, and since
1929: $a<0$ and $b>0$ were assumed to be much larger than $\epsilon^{-1}$
1930: in absolute value, it was legitimate to replace these
1931: limits with $-\infty$ and $\infty$, respectively.
1932: The error introduced in this way is exponentially small in $\epsilon$.
1933:
1934: Taking time derivatives of $N$ and $P$ above
1935: and discarding $\epsilon^1$ corrections to $\epsilon^0$ terms,
1936: we deduce that
1937: \begin{align*}
1938: {\dot \epsilon} & = \frac{\dot{N}+\tan k(\dot{P}-k\dot{N})}{2\cos k},
1939: \\
1940: {\dot k} & =\frac{\dot{P}-k\dot{N}}{2\epsilon\cos k}.
1941: \end{align*}
1942: Finally,
1943: substituting for $\dot{N}$ and $\dot{P}$ from eqs \eqref{Ndot} and \eqref{Pdot},
1944: we arrive at the dynamical system
1945: \begin{subequations}
1946: \label{paramev}
1947: \begin{align}
1948: \dot{\epsilon} &= |K_1(k)|^2e^{-\pi \mathcal{Q}_1/\epsilon} \,
1949: \Omega'(\mathcal{Q}_1) \,
1950: \frac{1+\mathcal{Q}_1\tan k}{2\cos k},
1951: \label{param_top} \\
1952: \dot{k} &= |K_1(k)|^2e^{-\pi \mathcal{Q}_1/\epsilon} \,
1953: \Omega'(\mathcal{Q}_1) \,
1954: \frac{\mathcal{Q}_1}{2\epsilon\cos k}, \label{param_bottom}
1955: \end{align}
1956: \end{subequations}
1957: where the group
1958: velocity $\Omega'(\mathcal{Q}_1)=2 \sin(k+\mathcal{Q}_1)-v$ with $v=2 (\sinh \epsilon / \epsilon)
1959: \sin k$,
1960: and $\mathcal{Q}_1=\mathcal{Q}_1(\epsilon,k)$ is the
1961: smallest root of eq.\,\eqref{nicedisprel}.
1962:
1963: \begin{figure}[b]
1964: \includegraphics{phaseplane.eps}
1965:
1966: \medskip
1967: \includegraphics{phaseplane2.eps}
1968: \caption{\label{phaseplane} The phase portrait of the system \eqref{paramev}
1969: in the case (a) where the Stokes constant $K_1(k)$ does not have zeros, and (b)
1970: where $K_1(k)$ has one zero. In (b), the dashed line is the line of nonisolated
1971: fixed points $k=k_1$. Note that in (b), the phase portrait has been replotted on
1972: the $(\epsilon, v)$ plane; hence the dashed line gives the value of the sliding
1973: velocity for each value of $\epsilon$. In (a), $\mu=0$; in (b), $\mu=6$.}
1974: \end{figure}
1975:
1976:
1977:
1978: \subsection{Soliton's deceleration and sliding velocities}
1979:
1980: The vector field \eqref{paramev} is defined for $k \geq k^{(1)}_{\rm min}$,
1981: where $k^{(1)}_{\rm min}$ is the value of $k$
1982: for which the roots $\mathcal{Q}_1$ and $\mathcal{Q}_2$ merge in
1983: fig.\,\ref{nicedisprelfig} (i.e.\ the smallest value of $k$ for
1984: which the roots $\mathcal{Q}_1$ and $\mathcal{Q}_2$
1985: still exist). When $k = k^{(1)}_{\rm min}$, the group velocity $\Omega'(\mathcal{Q}_1)$
1986: becomes equal to zero; therefore,
1987: the equation $k=k^{(1)}_{\rm min}(\epsilon)$ defines a line of
1988: (nonisolated) fixed points of the system
1989: \eqref{paramev}.
1990: Assume, first, that
1991: the saturation parameter $\mu$ is such that $K_1(k)$ does not vanish for
1992: any $k$.
1993: For $k > k^{(1)}_{\rm min}$,
1994: the factor $\Omega'(\mathcal{Q}_1)$ in \eqref{param_bottom} is negative; since
1995: $\mathcal{Q}_1$ and $\cos k$ are
1996: positive for all $0< k \leq \pi/2$, the derivative
1997: $\dot{k}$ satisfies
1998: $\dot{k} \leq 0$ for all times. Hence, all trajectories are flowing towards the
1999: line $k=k^{(1)}_{\rm min}(\epsilon)$ from above.
2000: It is worth
2001: noting that the derivative
2002: $\dot{\epsilon}$ is also negative, and that the $k$ axis is also
2003: a line of nonisolated fixed points. However, no
2004: trajectories will end there as follows from the equation
2005: \[
2006: \frac{d \epsilon}{dk} = \epsilon \left[ \tan k +
2007: \frac{1}{\mathcal{Q}_1(k,\epsilon)} \right],
2008: \]
2009: which is a consequence of \eqref{param_top} and \eqref{param_bottom}.
2010:
2011:
2012:
2013:
2014:
2015:
2016: Representative trajectories are plotted in fig.\,\ref{phaseplane}(a).
2017: Since $k^{(1)}_{\rm min}(\epsilon) \approx \epsilon^2/4 \pi$
2018: is very small for small $\epsilon$,
2019: the line $k=k^{(1)}_{\rm min}(\epsilon)$ is practically indistinguishable
2020: from the horizontal axis. Therefore, we can assert that, to a good
2021: accuracy,
2022: $k \to 0$ as $t \to \infty$.
2023: This means that the soliton stops moving --
2024: and stops decaying at the same time.
2025:
2026: For $k$ smaller than $k^{(1)}_{\rm min}$, the vector field \eqref{paramev} is
2027: undefined and we cannot use it to find out what happens
2028: to the soliton after $k$
2029: has reached $k^{(1)}_{\rm min}$. The reason
2030: for this is that the soliton stops radiating
2031: at the wavenumber $\mathcal{Q}_1$
2032: as $k$ drops below $k^{(1)}_{\rm min}$. In fact our analysis becomes invalid
2033: as soon as $k$ becomes $\mathcal{O}(\epsilon^2)$---i.e.\ even before $k$
2034: reaches $k^{(1)}_{\rm min}$---because we can no longer disregard the
2035: $n=2$ radiation here.
2036: At the qualitative level it
2037: is obvious, however, that the parameter $k$ should continue
2038: to decay all the way to zero, in a cascade way.
2039: First, the $n=2$ radiation will
2040: become as intense as the $n=1$ mode when $k$ approaches $\epsilon^2/4\pi$.
2041: Subsequently---i.e.\ for $k$ smaller than $\epsilon^2/4\pi$---the $n=3$ harmonic will
2042: replace the radiation with the wavenumbers $\mathcal{Q}_1$ and
2043: $\mathcal{Q}_2$ as a dominant mode.
2044: The $n=4$ mode will become equally intense
2045: near $k=k^{(2)}_{\rm min} \approx \epsilon^2/8\pi$; as $k$ drops below $\epsilon^2/8\pi$,
2046: both $\mathcal{Q}_3$ and $\mathcal{Q}_4$ will cede to
2047: $\mathcal{Q}_5$, and so on.
2048:
2049:
2050: If $\mu$ is such that
2051: the Stokes constant $K_1(k)$ vanishes at one or more values of $k$,
2052: the system has one or more lines of nonisolated fixed points, $k=k_i$.
2053: The corresponding values of $v$, $v=v_i(\epsilon)
2054: \equiv 2 (\sinh\epsilon /\epsilon ) \sin
2055: k_i$, define the {\it sliding\/} velocities of the soliton --
2056: i.e.\ velocities at which the soliton moves without radiative friction.
2057: One such velocity is shown by the dashed line in fig.\,\ref{phaseplane}(b).
2058: The fixed points $(\epsilon, k_i)$
2059: are semistable; for $k$ above
2060: $k_i$, the
2061: flow is towards the line $k=k_i$ but when $k$ is below $k_i$,
2062: the flow is directed away from this straight line [see fig.\,\ref{phaseplane}(b)].
2063: Since both $\dot{\epsilon}$ and $\dot{k}$ are negative,
2064: the soliton's velocity $v = 2 (\sinh\epsilon /\epsilon ) \sin k$
2065: will generally be decreasing -- until it hits the nearest underlying
2066: sliding velocity $v_i$ and locks on to it.
2067: The ensuing sliding motion will be unstable; a small
2068: perturbation will be sufficient to take the soliton out of the sliding regime
2069: after which it will resume its radiative deceleration.
2070: However, since $\dot{k}$ is proportional to the {\it square\/} of the
2071: Stokes constant and not to $K_1(k)$ itself, small perturbations
2072: $\delta k$ will be growing linearly, not exponentially, in $t$.
2073: As a result,
2074: the soliton may spend a fairly long time sliding at the velocity $v_i$.
2075: It is therefore not unreasonable to classify this sliding motion
2076: as {\it metastable\/}.
2077:
2078:
2079: We conclude
2080: that the soliton becomes pinned (i.e.\ $k \rightarrow 0$ and so $v
2081: \to 0$) before it has
2082: decayed fully (i.e.\ before the amplitude $\epsilon$ has decreased to zero).
2083: The exponential dependence
2084: of $\dot \epsilon$ and $\dot k$ on $1/\epsilon$ implies that
2085: tall, narrow pulses will be pinned very quickly while short, broad ones
2086: will travel for a very long time before they have slowed appreciably.
2087: Next, if $\mu$ is such that there is one or more sliding velocity
2088: available in the system,
2089: and if the soliton is initially moving faster than some of these,
2090: its deceleration will be interrupted by long periods of undamped
2091: motion at the corresponding sliding velocities.
2092:
2093: \begin{figure}
2094: \hspace*{-2.5em}\includegraphics{decayrate.eps}
2095: \caption{\label{decayrate} The soliton's decay rate as a function of its velocity, for fixed $\epsilon=0.1$.
2096: (Here $\mu=0$.) In the inset, the same curve is replotted using a logarithmic scale
2097: on the vertical axis.}
2098: \end{figure}
2099:
2100:
2101: It is worth reemphasising here that if the
2102: amplitude $\epsilon$ is small, then
2103: even if the soliton is {\it not\/}
2104: sliding, its deceleration will be so slow that it will spend an exponentially long time travelling
2105: with virtually unchanged amplitude and speed.
2106: The deceleration rate $-{\dot v}/v$ is shown in fig.\,\ref{decayrate}, as a function of
2107: the soliton's velocity $v$, for fixed amplitude $\epsilon$.
2108: Note that the decay rate drops, exponentially, as the velocity is decreased;
2109: this drop is due to the exponential factor $e^{-\pi \mathcal{Q}_1/ \epsilon}$ in eq.\,\eqref{paramev}.
2110: As the velocity $v$
2111: (and hence the wavenumber $k$) decreases, the root $\mathcal{Q}_1(\epsilon,k)$ grows
2112: towards the limit value of approximately $2 \pi$ (see fig.\,\ref{nicedisprelfig}). This
2113: variation in $\mathcal{Q}_1$ is amplified by the division by small $\epsilon$
2114: and exponentiation in $e^{-\pi \mathcal{Q}_1/ \epsilon}$.
2115:
2116:
2117:
2118:
2119: \section{Concluding remarks}
2120: \label{section5}
2121:
2122: \subsection{Summary}
2123:
2124: In this paper, we have constructed the moving discrete soliton of the
2125: saturable NLS equation \eqref{DNLS} [and hence \eqref{VK}]
2126: as an asymptotic expansion in powers of its amplitude.
2127: The saturable nonlinearity includes the cubic NLS as a particular case
2128: (for which $\mu=0$). Our perturbation procedure is a variant of the Lindstedt--Poincar\'e
2129: technique where corrections to the parameters of the solution are
2130: calculated along with the calculation of the solution itself.
2131: Although the resulting asymptotic series \eqref{regularpert} for the soliton
2132: is generally not convergent,
2133: the associated expansions for its frequency
2134: and velocity sum up to exact explicit expressions
2135: \eqref{om_and_v_via_eps}.
2136:
2137: From the divergence of the asymptotic series \eqref{regularpert}
2138: it follows, in particular,
2139: that the travelling discrete soliton does not decay to zero at least at one of the
2140: two infinities. Instead, the soliton approaches, as
2141: $X \to \infty$ or $X \to -\infty$, an oscillatory resonant background
2142: \eqref{stable-minus-unstable-outer}
2143: where the amplitudes $A_n$ of its constituent harmonic waves
2144: lie beyond all orders of $\epsilon$. For the soliton moving with a
2145: positive velocity and approaching zero as $X \to \infty$, the oscillatory
2146: background at the left infinity represents
2147: the Cherenkov radiation left in the soliton's wake. To
2148: evaluate the amplitudes of the harmonic waves
2149: arising as $X \to -\infty$, we have
2150: continued the radiating soliton
2151: into the complex plane, where it exhibits
2152: singularities. We then matched the asymptotic expansion \eqref{cont12}
2153: of the background
2154: near the lowest singularity on the imaginary $X$ axis to the far-field asymptotic
2155: expansion \eqref{expexpansion} of
2156: the background solution of the `inner' equation -- i.e.\ of the
2157: advance-delay equation `zoomed in' on this singularity. The asymptotic
2158: expansions here are in inverse powers of the zoomed variable, $y$.
2159: The amplitudes of the radiation waves were found to be exponentially
2160: small in $\epsilon$, with the pre-exponential factors (the so-called Stokes
2161: constants) being dependent only on the soliton's carrier-wave wavenumber.
2162: Representing solutions to the inner equation % \eqref{inner-equations}
2163: as Borel sums
2164: of their asymptotic expansions,
2165: the Stokes constants can be related to the expansion coefficients;
2166: we have calculated these coefficients numerically, using algebraic recurrence
2167: relations.
2168:
2169: The upshot of the calculation of the leading Stokes constant $K_1$ is that
2170: in the case of the cubic nonlinearity (i.e.\ for $\mu=0$), $K_1(k)$ does not vanish
2171: for any $k$. This means that the cubic discrete soliton cannot `slide' --
2172: i.e.\ cannot move without radiative friction. The saturable
2173: solitons, on the other hand, {\it can\/} slide provided the saturation
2174: parameter $\mu$ is large enough.
2175: This is because, for a sufficiently large $\mu$, the
2176: Stokes constant $K_1(k)$ is found to have one or more zeros $k_1, k_2, \ldots$.
2177: Since the soliton with wavenumber $k>0.22$ can have no higher-order
2178: resonances,
2179: those zeros which satisfy $k_i>0.22$ do define the wavenumbers at which sliding motion
2180: occurs.
2181: For each
2182: value of the soliton's amplitude $\epsilon$, the formula \eqref{om_and_v_via_eps}
2183: then gives the sliding velocities $v(k_i, \epsilon)$.
2184:
2185: The calculation of asymptotics beyond all orders is useful not only for
2186: determining the sliding velocities. Knowing the radiation amplitudes has
2187: allowed us to derive a two-dimensional dynamical system
2188: \eqref{paramev} for the soliton's parameters. Trajectories
2189: of this dynamical system describe the evolution of the soliton
2190: travelling at a generic speed. The evolution turns out to be
2191: simple:
2192: If $\mu$ is such that the Stokes constant
2193: $K_1(k)$ does not vanish anywhere in the region $k> 0.22$,
2194: the soliton decelerates, although very slowly.
2195: Eventually it becomes pinned to the lattice
2196: with a decreased but finite amplitude.
2197: However, if $\mu$ is such that there are sliding velocities
2198: in the system, and if the soliton starts its motion
2199: with the velocity higher than some of these then, although
2200: it will finally become pinned to the lattice, its
2201: deceleration will be interrupted by long periods of metastable sliding
2202: motion at these isolated velocities.
2203:
2204:
2205:
2206: \subsection{Concluding remarks}
2207:
2208: It is interesting to tie up the above results with our
2209: previous work on exceptional discretisations of the $\phi^4$
2210: model \cite{OPB}. In that case we discovered that for some
2211: exceptional models (in which the stationary soliton possesses an
2212: effective translational symmetry), sliding velocities, at which
2213: the radiation disappears, do exist.
2214: The fact that all these models involve a
2215: complicated nonlocal discretisation of the nonlinearity
2216: leads one to wonder whether the nonlinearity
2217: {\it has\/} to be discretised nonlocally
2218: in order for sliding solitons to exist.
2219: This question is answered%
2220: ---negatively---by
2221: our present work which gives a counterexample of
2222: a simple, local and physically motivated
2223: nonlinearity supporting sliding solitons.
2224:
2225: A remaining open issue is whether exceptionality of the system is a
2226: prerequisite for the existence of sliding solitary waves. Indeed, we
2227: have yet to encounter a nonexceptional discrete system
2228: permitting sliding motion.
2229:
2230: We conclude this section by placing our results in the context of
2231: earlier studies of discrete solitons.
2232:
2233: Moving solitons in the cubic DNLS equation
2234: have previously been studied by Ablowitz, Musslimani, and
2235: Biondini \cite{ablowitz}
2236: (among others), using a numerical technique based on discrete
2237: Fourier transforms
2238: as well as perturbation expansions for small velocities. As
2239: we have done, these authors suggest that sliding (radiationless)
2240: solitons may not exist. They also observe,
2241: with the aid of numerical simulations, that strongly
2242: localised pulses are pinned quickly to the lattice, while
2243: broader ones are more mobile -- results which are in agreement
2244: with ours.
2245:
2246: Duncan, Eilbeck, Feddersen and Wattis \cite{Feddersen,duncan} have studied the
2247: bifurcation of periodic travelling waves from constant
2248: solutions in the cubic DNLS equation, using numerical
2249: path-following techniques. They find that the paths terminate
2250: when the soliton's amplitude reaches a certain limit value,
2251: and in the light of our results it seems likely that this
2252: is the point at which the radiation becomes large enough to
2253: have an effect on the numerics.
2254:
2255: Solitary waves which decay to constant values at the spatial infinities,
2256: despite the fact that the generic asymptotic
2257: behaviour in the underlying model is oscillatory,
2258: are commonly referred to as embedded solitons. An example is given
2259: by the sliding solitons reported in this paper as well as the
2260: sliding kinks of \cite{OPB}; both types of solitary waves propagate without
2261: exciting the resonant background oscillations.
2262: Embedded solitons have been studied for some time in continuous systems
2263: \cite{embedded_review}, but their history in lattice equations is
2264: younger. Recently, Malomed, Gonz\'alez-P\'erez-Sandi, Fujioka,
2265: Espinosa-Cer\'on and Rodr\'iguez have considered certain lattice equations
2266: with next-to-nearest-neighbour couplings and shown (by means of explicit
2267: solutions) that both stationary
2268: \cite{Malomed_embedded_0} and moving \cite{Malomed_embedded} embedded solitons exist. Stationary embedded
2269: solitons in discrete waveguide arrays have also been analysed by
2270: Yagasaki, Champneys and Malomed \cite{Malomed_embedded_2}.
2271:
2272:
2273:
2274: Finally, while preparing the revised version of this paper,
2275: we learnt that Melvin, Champneys, Kevrekidis, and Cuevas
2276: have obtained results
2277: very similar to ours. Using a combination of
2278: intuitive arguments
2279: and numerical computations, they have found
2280: sliding solitons for
2281: certain values of the
2282: parameters of a saturable DNLS model \cite{melvin}.
2283:
2284:
2285:
2286: \begin{acknowledgments}
2287: The authors would like to thank Dmitry Pelinovsky and Alexander Tovbis
2288: for helpful discussions relating to this work.
2289: O.O. was supported by the National Research Foundation of South Africa
2290: and the University of Cape Town.
2291: I.B. was supported by the NRF under
2292: grant 2053723 and by the URC of UCT.
2293:
2294: \end{acknowledgments}
2295:
2296:
2297: \appendix
2298: \section{Convolution theorem for the Laplace transform in the complex plane}
2299: \label{A}
2300:
2301: This appendix deals with the convolution property
2302: of the modified Laplace
2303: transform of the form \eqref{laplace}, where the integration is
2304: along an infinite contour in the complex plane rather than the
2305: positive real axis. In the context of asymptotics beyond
2306: all orders, this transform was pioneered by Grimshaw and Joshi
2307: \cite{GrimshawJoshi}. Since no proof of the convolution result \eqref{conv_thm}
2308: is available in literature, and since it requires a nontrivial
2309: property (concavity) of the integration contour, we produce such
2310: a proof here.
2311:
2312:
2313: We wish to show that
2314: \begin{align}
2315: \label{rtp}
2316: &\int_{\gamma}e^{-pz}F(p)dp \int_{\gamma}e^{-p'z}G(p')dp' \nonumber\\
2317: = &\int_{\gamma}e^{-pz}
2318: \left[ \int_0^pF(p_1)G(p-p_1)dp_1 \right] dp,
2319: \end{align}
2320: where $\int_0^p$ stands for an integral along the curve $\gamma$
2321: from the origin to the point $p$ on that curve.
2322: We assume that the curve
2323: $\gamma$ extends from the origin to infinity on the complex plane; lies in
2324: its first quadrant ($\Real p>0$, $\Imag p>0$);
2325: is described as a graph of a single-valued function $\Imag p= f( \Real p)$
2326: (i.e.\ never turns back on itself),
2327: and is concave-up everywhere:
2328: \[
2329: \frac{d^2 f}{d(\Real p)^2} > 0 \quad \text{for} \ p \ \text{on} \, \gamma.
2330: \]
2331: We also assume that the function $G(p)$ is analytic in the region between the contour
2332: $\gamma$ and the imaginary axis.
2333:
2334: \begin{figure}
2335: \includegraphics{conv.eps}
2336: \caption{\label{conv-paths}Deformation of the integration contour $\Gamma_p$ to $\Gamma_p'$.}
2337: \end{figure}
2338:
2339: We begin by writing the left-hand side of \eqref{rtp} as
2340: \begin{align}
2341: &\int_{\gamma}\int_{\gamma}e^{-(p+p')z}F(p)G(p')dp'dp
2342: \nonumber \\
2343: \label{undeformed-integral}
2344: = &\int_{\gamma}F(p) \left[ \int_{\Gamma_p}e^{-rz}G(r-p)dr \right] dp,
2345: \end{align}
2346: where $r = p' + p$. The curve $\Gamma_p$ is the path traced out
2347: by $r$ as $p'$ traces out the curve $\gamma$, for a given
2348: $p$ (which also lies on $\gamma$) -- this is depicted in
2349: fig.\ \ref{conv-paths}. The path $\Gamma_p$ is the same
2350: as the path $\gamma$ but translated from the origin to
2351: the point $p$.
2352:
2353: For each given $p$ we deform the path $\Gamma_p$ so that
2354: it now lies on $\gamma$, still starting at the point $p$.
2355: We call this deformed path $\Gamma_p'$. The point $r-p = p'$, which
2356: lay on the path $\gamma$ before the
2357: deformation,
2358: will now move inside the region bounded by $\gamma$ (i.e.\ the region to the left of $\gamma$),
2359: since all points on $\Gamma_p'$ lie inside the region bounded by $\Gamma_p$.
2360: This follows from the fact that the path $\Gamma_p$ is concave-up.
2361: The point $r-p$ will, however,
2362: stay to the right of the imaginary axis, since after the deformation
2363: all points $r$ on $\Gamma_p'$ lie to the right of the point $p$.
2364: This follows from the fact that the curve
2365: $\gamma$ never turns back on itself.
2366: Since $G(p')$ is analytic for all $p'$ between the imaginary axis and $\gamma$,
2367: the value of the integral \eqref{undeformed-integral} will not be affected by
2368: the deformation.
2369: [In the particular case of the functions $U(p)$ and $W(p)$
2370: considered in section \ref{BLT}, we have deliberately chosen the contours of integration
2371: so that the analyticity condition is satisfied.]
2372:
2373: After the deformation, both integrals in \eqref{undeformed-integral}
2374: follow the path $\gamma$, with the inner one starting at the point $p$.
2375: By parameterising the path $\gamma$ it is now straightforward
2376: to change the order of integration, and end up with the desired
2377: result \eqref{rtp}.
2378:
2379: The argument holds, with the obvious modifications, for paths
2380: in the second quadrant.
2381:
2382:
2383: \section{Proof that singularities do not accumulate to the imaginary axis}
2384: \label{B}
2385:
2386: The roots of eq.\,\eqref{singularities}
2387: with the top and bottom signs give singularities of
2388: $U(p)$ and $W(p)$, respectively. Our aim
2389: here
2390: is to show that the integration contours $\gamma_u$ and $\gamma_s$
2391: in section \ref{BLT} can be
2392: chosen to lie above all the complex singularities -- i.e., that
2393: singularities of
2394: $U(p)$ and $W(p)$ do not accumulate to the imaginary axis.
2395:
2396: The roots of
2397: eq.\,\eqref{singularities} are zeros of the functions $F(\pm p)$, where
2398: \begin{equation}
2399: F(p) \equiv \cosh p - 1 - i\tan k(\sinh p - p).
2400: \end{equation}
2401: The imaginary zeros of $F(p)$ are at $p = - iq_n$,
2402: where $q_n$ are the real roots of eq.\,\eqref{leading-order-disp}.
2403: We let $N$ denote
2404: the number of these imaginary zeros;
2405: for $k \neq 0$, $N$ is finite.
2406: We recall that all $q_n$ are positive; hence all $N$ imaginary
2407: zeros of $F(p)$ are
2408: on the negative imaginary axis, and all $N$ zeros of $F(-p)$ on the positive
2409: imaginary axis.
2410:
2411: Let $\kappa$ and $q$ be the real and imaginary part of $p$: $p = \kappa
2412: +iq$. We let $\mathcal{D}$ denote the rectangular region in the
2413: complex-$p$ plane bounded by the vertical lines $\kappa = \pm \varepsilon$
2414: and horizontal lines
2415: $q = \varepsilon^{1/2}$ and $q = Q$, where $\varepsilon$ is small and
2416: $Q$ is large enough for the region to contain all $N$ zeros of $F(-p)$ on the
2417: positive imaginary axis. By the argument principle, the total number
2418: of (complex) zeros of the function $F(-p)$ in the region $\mathcal{D}$ is given by
2419: $(2 \pi)^{-1}$ times
2420: the variation of its argument (i.e., its phase) along the boundary of $\mathcal{D}$.
2421: In a similar way, one may count the number of zeros of the
2422: function $F(p)$ in $\mathcal{D}$.
2423:
2424: We have
2425: \begin{equation*}
2426: \tan \arg F(\pm p) \! = \! \frac
2427: {\sinh\kappa\sin q \mp \tan k(\sinh\kappa\cos q - \kappa)}
2428: {\cosh\kappa\cos q - 1 \pm \tan k(\cosh\kappa\sin q - q)}.
2429: \end{equation*}
2430: Points on the vertical line $\kappa = \varepsilon$
2431: satisfy
2432: \begin{equation}
2433: % \tan \arg F(\pm p) \approx {-\epsilon} \left[ \frac
2434: % {\sin q \pm \tan k(1-\cos q)}
2435: % {1-\cos q \pm \tan k(q-\sin q )}
2436: % \right]. \\
2437: \tan \arg F(\pm p) \approx {-\varepsilon} \frac{d}{dq} \ln
2438: |1-\cos q \pm \tan k(q-\sin q )|.
2439: \label{B3}
2440: \end{equation}
2441: Consider, first, the case of the bottom sign in eq.\,\eqref{B3}.
2442: The expression between the bars in \eqref{B3}
2443: crosses through zero $N$ times as the line $\kappa = \varepsilon$ is
2444: traced out. Each time
2445: zero is crossed, the logarithmic derivative in \eqref{B3} jumps from $-\infty$ to $+\infty$,
2446: and the argument of $F(-p)$ increases by $\pi$ as we move from one crossing
2447: to the next one.
2448: The net increase of the argument
2449: as the line $\kappa = \varepsilon$
2450: is traversed from its bottom to the top, is $N \pi$.
2451: In the case of the top sign in \eqref{B3}, on the other hand, the expression in $|\ldots|$ never crosses
2452: through zero and hence the total increment of the argument
2453: of $F(p)$ is zero.
2454:
2455: As we move along the vertical
2456: line $\kappa = -\varepsilon$ from top to bottom, the argument of $F(-p)$
2457: increases by another $N \pi$ while the argument of $F(p)$
2458: does not acquire any increment. Since there are no zero crossings on the horizontal segments,
2459: the net change of
2460: the argument of $F(-p)$ along the boundary of $\mathcal{D}$ is
2461: $2N\pi$ while the total increment of ${\rm arg\/} \, F(p)$ is zero.
2462: Therefore $F(-p)$ has only $N$ zeros in the region $\mathcal{D}$ (and they all
2463: lie on the imaginary axis) while the function $F(p)$
2464: has no zeros in $\mathcal{D}$. This implies that complex singularities
2465: of $U(p)$ and $W(p)$ cannot accumulate to
2466: the positive imaginary axis.
2467:
2468:
2469: \begin{thebibliography}{22}
2470:
2471: \bibitem{braun}
2472: O. Braun and Yu. S. Kivshar,
2473: Phys. Rep. \textbf{306}, 1 (1998).
2474:
2475: \bibitem{hennig}
2476: D. Hennig and G. P. Tsironis
2477: Phys. Rep. \textbf{307}, 333 (1999).
2478:
2479: \bibitem{scott}
2480: A. Scott, \emph{Nonlinear Science: emergence and dynamics of
2481: coherent structures} (Oxford University Press, 1999).
2482:
2483: %-------------------------------------------
2484:
2485: \bibitem{christodoulides}
2486: D. N. Christodoulides, F. Lederer, and Y. Silberberg,
2487: Nature \textbf{424}, 817 (2003).
2488:
2489: \bibitem{campbell}
2490: D. K. Campbell, S. Flach, and Yu. S. Kivshar,
2491: Phys. Today \textbf{57}(1), 43 (2004).
2492:
2493: %--------------------------------------------
2494:
2495: \bibitem{trambettoni}
2496: A. Trombettoni and A. Smerzi,
2497: Phys. Rev. Lett. \textbf{86}, 2353 (2001).
2498:
2499: %-------------------------------------------
2500:
2501: \bibitem{Feddersen}
2502: H. Feddersen, in \emph{Nonlinear Coherent Structures in Physics
2503: and Biology. Proceedings of the 7th Interdisciplinary
2504: Workshop Held at Dijon, France, 4-6 June 1991.}
2505: Eds. M. Remoissenet and M. Peyrard, Lecture Notes
2506: in Physics vol. 393 (Springer, Berlin, 1991), pp. 159--167.
2507:
2508: \bibitem{duncan}
2509: D. B. Duncan, J. C. Eilbeck, H. Feddersen, and J. A. D. Wattis,
2510: Physica D \textbf{68}, 1 (1993).
2511:
2512: \bibitem{flach}
2513: S. Flach and C. R. Willis,
2514: Phys. Rep. \textbf{295}, 181 (1998).
2515:
2516: \bibitem{Flach_Kladko}
2517: S. Flach and K. Kladko,
2518: Physica D \textbf{127}, 61 (1999).
2519:
2520: \bibitem{FKZ}
2521: S. Flach, Y. Zolotaryuk and K. Kladko,
2522: Phys. Rev. E \textbf{59}, 6105 (1999).
2523:
2524: \bibitem{kevrekidis}
2525: P. G. Kevrekidis, K. \O{}. Rasmussen, and A. R. Bishop,
2526: Int. J. Mod. Phys. B \textbf{15}, 2833 (2001).
2527:
2528: \bibitem{ablowitz}
2529: M. J. Ablowitz, Z. H. Musslimani, and G. Biondini,
2530: Phys. Rev. E \textbf{65}, 026602 (2002).
2531:
2532: \bibitem{Kundu} K. Kundu, J. Phys. A: Math. Gen. {\bf 35}, 8109 (2002).
2533:
2534: \bibitem{eilbeck}
2535: J. C. Eilbeck and M. Johansson,
2536: in \emph{Proceedings of the Third Conference on Localization
2537: and Energy Transfer in Nonlinear Systems, San Lorenzo de El Escorial, Madrid,
2538: Spain, June 17--21, 2002} (World Scientific, Singapore, 2003), pp. 44--67.
2539:
2540: \bibitem{Malomed_collisions_1}
2541: I. E. Papacharalampous, P. G. Kevrekidis, B. A. Malomed, and D. J. Frantzeskakis,
2542: Phys. Rev. E \textbf{68}, 046604 (2003).
2543:
2544: \bibitem{Malomed_collisions_2}
2545: S. V. Dmitriev, P. G. Kevrekidis, B. A. Malomed, and D. J. Frantzeskakis,
2546: Phys. Rev. E \textbf{68}, 056603 (2003).
2547:
2548: \bibitem{pelinovsky}
2549: D. E. Pelinovsky and V. M. Rothos,
2550: Physica D \textbf{202}, 16 (2005).
2551:
2552: \bibitem{Malomed_nonlinearity_management}
2553: J. Cuevas, B. A. Malomed, and P. G. Kevrekidis, Phys. Rev. E {\bf 71}, 066614 (2005).
2554:
2555: %----------------------------------------------------
2556:
2557: \bibitem{gomez1}
2558: J. G\'omez-Garde\~nez, F. Falo, and L. M. Flor\'ia,
2559: Phys. Lett. A \textbf{332}, 213 (2004).
2560:
2561: \bibitem{gomez}
2562: J. G\'omez-Garde\~nez, L. M. Flor\'ia, M. Peyrard, and A. R. Bishop,
2563: Chaos \textbf{14}, 1130 (2004).
2564:
2565: %----------------------------------------------------
2566:
2567: \bibitem{efremidis}
2568: N. K. Efremidis, S. Sears, D. N. Christodoulides, J. W. Fleischer, and M. Segev,
2569: Phys. Rev. E \textbf{66}, 046602 (2002).
2570:
2571: \bibitem{fleischer}
2572: J. W. Fleischer, T. Carmon, M. Segev, N. K. Efremidis, and D. N. Christodoulides,
2573: Phys. Rev. Lett. \textbf{90}, 023902 (2003).
2574:
2575: \bibitem{chen}
2576: F. Chen, C. E. R\"uter, D. Runde, D. Kip, V. Shandarov, O. Manela, and M. Segev,
2577: Opt. Exp. \textbf{13}, 4314 (2005).
2578:
2579: %----------------------------------------------------
2580:
2581: \bibitem{fleischer-nature}
2582: J. W. Fleischer, M. Segev, N. K. Efremidis, and D. N. Christodoulides,
2583: Nature \textbf{422}, 147 (2003).
2584:
2585: \bibitem{hadzievski}
2586: L. Had\v{z}ievski, A. Maluckov, M. Stepi\'c, and D. Kip,
2587: Phys. Rev. Lett. \textbf{93}, 033901 (2004).
2588:
2589: \bibitem{stepic}
2590: M. Stepi\'c, D. Kip, L. Had\v{z}ievski, and A. Maluckov,
2591: Phys. Rev. E \textbf{69}, 066618 (2004).
2592:
2593: \bibitem{cuevas}
2594: J. Cuevas and J. C. Eilbeck,
2595: Phys. Lett. A \textbf{358}, 15 (2006).
2596:
2597: \bibitem{vicencio}
2598: R. A. Vicencio and M. Johansson,
2599: Phys. Rev. E \textbf{73}, 046602 (2006).
2600:
2601: %-----------------------------------------------------
2602:
2603: \bibitem{gatz1}
2604: S. Gatz and J. Herrmann,
2605: J. Opt. Soc. Am. B \textbf{8}, 2296 (1991).
2606:
2607: \bibitem{gatz2}
2608: S. Gatz and J. Herrmann,
2609: Opt. Lett. \textbf{17}, 484 (1992).
2610:
2611: %------------------------------------------------------
2612:
2613: \bibitem{tikhonenko}
2614: V. Tikhonenko, J. Christou, and B. Luther-Davies,
2615: Phys. Rev. Lett. \textbf{76}, 2698 (1996).
2616:
2617: %------------------------------------------------------
2618:
2619: \bibitem{vidal}
2620: F. Vidal and T. W. Johnston,
2621: Phys. Rev. E \textbf{55}, 3571 (1997).
2622:
2623: %------------------------------------------------------
2624:
2625: \bibitem{khare}
2626: A. Khare, K. \O{}. Rasmussen, M. R. Samuelsen, and A. Saxena,
2627: J. Phys. A: Math. Gen. \textbf{38}, 807 (2005).
2628:
2629: %------------------------------------------------------
2630:
2631: \bibitem{BOP} I. V. Barashenkov, O. F. Oxtoby, and
2632: D. E. Pelinovsky, Phys. Rev. E {\bf 72} 035602(R) (2005).
2633:
2634: %------------------------------------------------------
2635:
2636: \bibitem{OPB}
2637: O. F. Oxtoby, D. E. Pelinovsky, and I. V. Barashenkov,
2638: Nonlinearity \textbf{19}, 217 (2006).
2639:
2640: %------------------------------------------------------
2641:
2642: \bibitem{vinetskii}
2643: V. O. Vinetskii and N. V. Kukhtarev,
2644: Sov. Phys. Solid State \textbf{16}, 2414 (1975).
2645:
2646: %------------------------------------------------------
2647:
2648: \bibitem{Lindstedt} See e.g. A. H. Nayfeh and D. T. Mook, Nonlinear
2649: Oscillations (J. Wiley, New York, 1979); D. W. Jordan and P. Smith, Nonlinear
2650: Ordinary Differential Equations (Oxford University Press, 1999).
2651:
2652: %------------------------------------------------------
2653:
2654: \bibitem{AL} M.J. Ablowitz and J.F. Ladik, Stud. Appl. Math.
2655: {\bf 55} 213 (1976); J. Math. Phys. {\bf 17} 10011 (1976).
2656:
2657: %------------------------------------------------------
2658:
2659: \bibitem{Laedke_Kluth_Spatschek}
2660: E. W. Laedke, O. Kluth, and K. H. Spatschek, Phys. Rev.
2661: {\bf E 54} 4299 (1996).
2662:
2663: %------------------------------------------------------
2664:
2665: \bibitem{pelinovsky-exceptional}
2666: D. E. Pelinovsky, Nonlinearity \textbf{19}, 2695 (2006).
2667:
2668: %------------------------------------------------------
2669:
2670: \bibitem{eleonskii}
2671: V. M. Eleonksii, N. E. Kulagin, N. S. Novozhilova, and V. P. Silin,
2672: Teor. Mat. Fiz. \textbf{60}, 395 (1984).
2673:
2674: %------------------------------------------------------
2675:
2676: \bibitem{sk}
2677: H. Segur and M. D. Kruskal,
2678: Phys. Rev. Lett. \textbf{58}, 747 (1987).
2679:
2680: \bibitem{sk2}
2681: M. D. Kruskal and H. Segur,
2682: Stud. Appl. Math. \textbf{85}, 129 (1991).
2683:
2684: %------------------------------------------------------
2685:
2686: \bibitem{pomeau}
2687: Y. Pomeau, A. Ramani, and B. Grammaticos,
2688: Physica D \textbf{31}, 127 (1988).
2689:
2690: %------------------------------------------------------
2691:
2692: \bibitem{GrimshawJoshi}
2693: R. H. J. Grimshaw and N. Joshi,
2694: SIAM J. Appl. Math. \textbf{55}, 124 (1995).
2695:
2696: \bibitem{GrimshawNLS}
2697: R. H. J. Grimshaw,
2698: Stud. Appl. Math. \textbf{94}, 257 (1995).
2699:
2700: %------------------------------------------------------
2701:
2702: \bibitem{Tovbis1}
2703: A. Tovbis, M. Tsuchiya, and C. Jaff\'e,
2704: Chaos \textbf{8}, 665 (1998).
2705:
2706: \bibitem{Tovbis2}
2707: A. Tovbis,
2708: Contemporary Mathematics \textbf{255}, 199 (2000).
2709:
2710: \bibitem{Tovbis3}
2711: A. Tovbis, Stud. Appl. Math. \textbf{104}, 353 (2000).
2712:
2713: \bibitem{TovPel}
2714: A. Tovbis and D. Pelinovsky,
2715: Nonlinearity \textbf{19}, 2277 (2006).
2716:
2717: %------------------------------------------------------
2718:
2719: \bibitem{embedded_review}
2720: J. Fujioka, A. Espinosa-Cer\'on, and R. F. Rodr\'iguez,
2721: Rev. Mex. F\'is. \textbf{52}, 6 (2006).
2722:
2723: %------------------------------------------------------
2724:
2725: \bibitem{Malomed_embedded_0}
2726: S. Gonz\'alez-P\'erez-Sandi, J. Fujioka, and B. A. Malomed,
2727: Physica D \textbf{197}, 86 (2004).
2728:
2729: %------------------------------------------------------
2730:
2731: \bibitem{Malomed_embedded}
2732: B. A. Malomed, J. Fujioka, A. Espinosa-Cer\'on, R. F. Rodr\'iguez, and
2733: S. Gonz\'alez,
2734: Chaos \textbf{16}, 013112 (2006).
2735:
2736: %------------------------------------------------------
2737:
2738: \bibitem{Malomed_embedded_2}
2739: K. Yagasaki, A. R. Champneys, and B. A. Malomed,
2740: Nonlinearity {\bf 18}, 2591 (2005).
2741:
2742: %------------------------------------------------------
2743:
2744: \bibitem{melvin}
2745: T. R. O. Melvin, A. R. Champneys, P. G. Kevrekidis, and J. Cuevas,
2746: Phys. Rev. Lett. \textbf{97}, 124101 (2006).
2747:
2748:
2749: \end{thebibliography}
2750:
2751: \end{document}
2752: