1: %\renewcommand{\theequation}{\thesection.\arabic{equation}}
2: %\usepackage{amsmath}
3: %\setcounter{MaxMatrixCols}{10}
4:
5:
6: \documentclass[pra,final,titlepage,showpacs]{revtex4}
7: \usepackage{amsfonts}
8: \usepackage{graphicx}
9: \usepackage{amssymb}
10: %%\usepackage{latexsym}
11: %%\usepackage{cite}
12:
13: \newcommand{\eq}[1]{(\ref{#1})}
14: \newcommand{\eps}{\varepsilon}
15: \newcommand{\beq}{\begin{equation}}
16: \newcommand{\eeq}{\end{equation}}
17: \newcommand{\bea}{\begin{eqnarray}}
18: \newcommand{\eea}{\end{eqnarray}}
19: \newcommand{\bean}{\begin{eqnarray}}
20: \newcommand{\eean}{\end{eqnarray}}
21: %%\eqnobysec
22: \textheight 23cm
23: \textwidth 17cm
24: \topmargin=0pt
25: \headheight=0cm
26: \headsep=12pt
27: \footskip=1.5cm
28: \oddsidemargin=-0.2cm
29: \evensidemargin=-0.2cm
30: \pagestyle{plain}
31: \def\d{{\mathrm{d}}}
32: \def\Rset{\mathbb{R}}
33: \def\Zset{\mathbb{Z}}
34: \def\fix{\mathrm{Fix}}
35: \def\sgn{{\mathrm{sgn}}}
36: \def\eg{{e.g.}}
37: \def\ie{{i.e.}}
38: \renewcommand{\topfraction}{1}
39: \renewcommand{\bottomfraction}{1}
40: \renewcommand{\textfraction}{0}
41:
42: \begin{document}
43:
44: \title{Discrete Embedded Solitons}
45: \author{Kazuyuki~Yagasaki}
46: \affiliation{Department of Mechanical and Systems Engineering,
47: Gifu University, Gifu 501-1193, Japan}
48: \author{Alan R.~Champneys}
49: \affiliation{Department of Engineering Mathematics, University of
50: Bristol, Bristol BS8 1TR, UK}
51: \author{Boris A.~Malomed}
52: \affiliation{
53: Department of Interdisciplinary Studies,
54: School of Electrical Engineering,
55: Faculty of Engineering, Tel Aviv University, Tel Aviv 69978, Israel}
56:
57:
58: \begin{abstract}
59: We address the existence and properties of discrete
60: \textit{embedded solitons} (ESs), that is localized waves existing
61: inside the phonon band in a nonlinear dynamical-lattice model. The
62: model describes a one-dimensional array of optical waveguides with
63: both $\chi ^{(2)}$ (second-harmonic generation) and $\chi ^{(3)}$
64: (Kerr) nonlinearities, for which a rich family of ESs are known to
65: occur in the continuum limit. First, a simple motivating problem
66: is considered, in which the $\chi ^{(3)}$ nonlinearity acts in a
67: single waveguide. An explicit solution is constructed
68: asymptotically in the large-wavenumber limit. The general problem
69: is then shown to be equivalent to the existence of a homoclinic
70: orbit in a four-dimensional reversible map. From properties of
71: such maps, it is shown that (unlike ordinary gap solitons),
72: discrete ESs have the same codimension as their continuum
73: counterparts. A specific numerical method is developed to compute
74: homoclinic solutions of the map, that are symmetric under a
75: specific reversing transformation. Existence is then studied in
76: the full parameter space of the problem. Numerical results agree
77: with the asymptotic results in the appropriate limit and suggest
78: that the discrete ESs may be semi-stable as in the continuous
79: case.
80: \end{abstract}
81:
82: \pacs{05.45.Yv, 05.45.-a, 42.65.Tg, 42.65.Ky}
83:
84: \maketitle
85: \section{Introduction}
86:
87: An \emph{embedded soliton} (ES) is an isolated solitary wave in a
88: non-integrable system that resides insides the continuous spectrum
89: of linear waves. Unlike regular \emph{gap solitons}, the existence
90: of ESs in continuum models is generally of codimension-one in the
91: model's parameter space. That is, they are embedded into wider
92: families of \emph{generalized} solitary waves which have
93: nonvanishing periodic tails attached to them, such that the ES
94: occurs at isolated values of the frequency, or other intrinsic
95: parameter of the solution family at which the tail's amplitude
96: vanishes. The concept of ESs was introduced in \cite{YaMaKa:99}
97: (see also \cite{ChMaYaKa:01,YaMaKaCh:01} for more details),
98: although particular examples of waves of this kind were actually
99: known earlier (but not under this name)
100: \cite{ChGr:97,FuEs:97,ChMaFr:98}. In the last few years, ESs have
101: been shown to be of relevance in a diverse range of nonlinear-wave
102: models, see, e.g., ~\cite{KoChBuSa:02,AtMa:01,EsFuGo:03,Ya:03} and
103: references therein. One of their characteristic properties is that
104: they are typically, at best, \emph{semi-stable}. That is, the
105: linearization around them gives rise to no unstable eigenvalues,
106: but a non-trivial zero-eigenvalue mode can be found. The latter
107: leads to an algebraic-in-time instability for energy-decreasing
108: perturbations, and a similar algebraic-in-time relaxation back to
109: the original pulse for energy-increasing perturbations
110: \cite{PeYa:02}. Only in special cases of systems which have more
111: than one dynamical invariant, or for which the travelling-wave
112: system is reversible but not Hamiltonian, have examples been found
113: of ESs that are truly asymptotically stable \cite{Ya:03}. Also
114: known are physically relevant models with special symmetries that
115: support continuous (rather than discrete) families of ESs, and a
116: large part thereof may be stable \cite{Mak,Merhasin}.
117:
118: All the above-cited examples pertain to continuum models. A
119: fundamental issue is whether solitons of the embedded type may
120: also occur in discrete nonlinear equations modelling dynamical
121: lattices. For example, in \cite{ChKi:99} it was argued that moving
122: topological defects in Frenkel-Kontorova (discrete sine-Gordon)
123: lattices can meaningfully be described as \emph{embedded kinks}.
124: Some of specific examples of moving discrete solitons constructed
125: by means of the \textit{inverse method} (a model is sought for to
126: which a given moving soliton is an exact solution \cite{Flach})
127: may also have the embedded character. Such waves, which connect a
128: zero state to itself, translated through a multiple of the period
129: of the lattice potential (the `topological charge'), were first
130: (indirectly) identified by Peyrard and Kruskal \cite{PeKr:84} and
131: have later received much attention in a wide variety of
132: applications \cite{KiBr:98}. In a co-moving frame, these waves
133: should be thought of as localized solutions to
134: infinite-dimensional \emph{advance and delay} (nonlocal)
135: equations, see works \cite{SaZoEi:00,AiChRo:03} where branches of
136: embedded kinks with the topological charge 1 and 2 were computed.
137: Another recent study \cite{ESdiscrete} reported the existence of
138: ESs in a discrete lattice of the Ablowitz-Ladik type, but with a
139: quintic nonlinearity added to the usual cubic terms, and also with
140: an additional next-nearest-neighbor linear coupling. By tuning the
141: nonlinear terms, an explicit analytical solution for a discrete ES
142: was found in that model (which somehow resembles the
143: above-mentioned inverse method). Apart from these, we are aware of
144: no other work that systematically considers the issue of ESs in
145: discrete models.
146:
147: A general objective of this paper is to understand the existence
148: and multiplicity properties of discrete solitons in lattice
149: systems for which the corresponding continuum model is known to
150: support ESs. The first fundamental issue is whether the
151: codimension of their existence is preserved under discretization.
152: In dissipative systems, solitary pulses are supported due to the
153: balance between forcing and damping, hence the corresponding
154: homoclinic orbits are always of codimension one. Thus, formally
155: similar to ESs, dissipative solitons typically exist as solutions
156: to the corresponding stationary ordinary differential equation at
157: discrete values of a relevant parameter (temporal or spatial
158: frequency, depending on the physical realization). Then, it will
159: generically happen that, under the discretization of that
160: equation, the codimension of the solitons \emph{decreases}. This
161: is possible because discretization of homoclinic orbits to
162: hyperbolic equilibria leads to transverse homoclinic connections
163: that exist for a range of parameter values \cite{FiSc}. However,
164: for ESs we shall show that, in contrast, discrete ESs typically
165: keep \emph{the same} codimension as their continuum counterparts,
166: i.e., they also exist at discrete values of the corresponding
167: physical parameter.
168:
169: In this work, we restrict ourselves to a discretization of a
170: single continuum model, for which the discrete counterpart finds a
171: direct physical interpretation. In fact, our results and
172: methodology will carry over to other discretized models that
173: support ESs (this expectation is supported by the general fact
174: that equations of the discrete nonlinear-Schr\"{o}dinger type,
175: that we consider below, may be derived as an asymptotic limit from
176: a much larger class of discrete models \cite{Morgante}).
177: Specifically, we shall start with the continuum model in which ESs
178: were identified \cite{YaMaKa:99}. It describes an optical medium
179: with competing quadratic ($\chi ^{(2)}$) and cubic ($\chi ^{(3)}$)
180: nonlinearities (see \cite{Bang1,Bang2,Bang3} and references
181: therein for physical applications). In this model, the evolution
182: variable $z$ is the propagation distance in the nonlinear optical
183: waveguide, and the variable $t$ is either the reduced time (in the
184: case of temporal solitons) or, which is more physically relevant,
185: a transverse spatial coordinate in a planar waveguide. In the
186: temporal case, such a model can be written in dimensionless form
187: as (see detailed explanations in reviews
188: \cite{Review,DiscrChi2soliton})
189: \begin{eqnarray}
190: &&i\frac{\partial u}{\partial z}+\frac{1}{2}\frac{\partial ^{2}u}{\partial
191: t^{2}}+u^{\ast }v+\gamma _{1}(|u|^{2}+2|v|^{2})u=0, \nonumber \\[-1ex]
192: && \label{eqn:PDE} \\[-1ex]
193: &&i\frac{\partial v}{\partial z}+\frac{\delta }{2}\frac{\partial
194: ^{2}v}{\partial t^{2}}+qv+\frac{1}{2}u^{2}+2\gamma
195: _{2}(|v|^{2}+2|u|^{2})v=0, \nonumber
196: \end{eqnarray}where $u$ and $v$ are the local amplitudes of the fundamental-frequency (FF)
197: and second-harmonic (SH) fields, an asterisk denotes complex
198: conjugation, $\delta $ is the relative dispersion of the SH, $q$
199: is the wavenumber mismatch, and $\gamma _{1,2}$ measure the
200: relative strength of the cubic (Kerr) nonlinearity compared to its
201: quadratic (SH-generating) counterpart, whose strength is
202: normalized to be $1$. In the physical model, $\gamma _{1,2} $ must
203: have the same sign (typically, they are positive, but may, in
204: principle, be negative too). The spatial variant of the system
205: (\ref{eqn:PDE}) takes the same form, with a partly different
206: interpretation of the dimensionless parameters, and is easier to
207: implement experimentally \cite{Bang1,Bang2,Bang3}. In the spatial
208: domain, one always has $\delta \equiv 1$.
209:
210: The existence of ESs in this model was established in
211: \cite{YaMaKa:99,YaMaKaCh:01} for both cases, $\gamma _{1,2}>0$ and
212: $\gamma _{1,2}<0$. In the case of relevance to spatial waveguides
213: ($\delta =1$), fundamental ESs are absent in the model, while
214: higher-order ones can be found.
215:
216: In models such as (\ref{eqn:PDE}) which includes $\chi ^{(2)}$
217: interaction, ESs can be embedded only in the continuous spectrum
218: of the SH component, while the FF wavenumber can never be located
219: inside the continuous spectrum. This feature is stipulated by the
220: asymmetry between the two equations in system (\ref{eqn:PDE}). If
221: the SH supports linear waves, while the FF has the possibility of
222: exponential localization like $e^{-\lambda |t|}$, then the $u^{2}$
223: term, which drives the SH field in the second equation
224: (\ref{eqn:PDE}), allows it to have tails that decay exponentially
225: at a rate $e^{-2\lambda |t|}$. Such a solitary wave was said in
226: reference \cite{YaMaKa:99} to be \emph{tail-locked} and,
227: accordingly, the SH equation to be \emph{nonlinearisable} (because
228: the quadratic term can never be neglected in this equation). The
229: tail locking does not take place if the SH field supports its own
230: exponentially decaying tails that asymptotically (for
231: $|t|\rightarrow \infty $) dominate over the tails induced by the
232: term $u^{2}$.
233:
234: In this paper we study the existence of discrete ESs in a model
235: arising from discretization of the (actually, spatial)
236: $t$-coordinate in (\ref{eqn:PDE}). One motivation for doing this
237: is to understand the effect of numerical approximation (which also
238: implies discretization, once a finite-difference scheme is
239: employed) on the computation of ESs in model systems. But this is
240: not our primary motivation. Lattice models play an increasingly
241: important role in describing physical phenomena in a number of
242: newly developed experimental settings. In particular, the discrete
243: version of the $\chi ^{(2)}:\chi ^{(3)}$ model describes an array
244: of the corresponding linearly-coupled waveguides. The creation of
245: discrete spatial solitons in a system of channel waveguides with
246: the $\chi ^{(2)}$ nonlinearity was recently reported
247: \cite{DiscrChi2soliton}. Our assumption concerning relevant
248: solutions is that, once the continuum version of the model readily
249: gives rise to ESs, then there is a chance to find them in the
250: corresponding lattice model too. In this paper, we show that this
251: is the case indeed and investigate how we may then pass to the
252: continuum limit. We report the corresponding discrete ES that can
253: be found, under special conditions, in an approximate analytical
254: form, and, in the general case, numerically.
255:
256: The paper is organized as follows. In Section~2, we present the
257: lattice model to be considered in this work and discuss its
258: physical applicability. In Section~3, an asymptotic analysis is
259: undertaken for a reduced model where the cubic nonlinearity is
260: present only at a single site. Section~4 then goes on to discuss
261: the meaning of discrete ESs, as a solution to a system of
262: stationary finite-difference equations, in terms of a homoclinic
263: orbit to a fixed point of a discrete-time reversible map. This
264: makes it possible to understand that the solutions we seek must
265: have codimension one, similar to ESs in continuum models. In terms
266: of the same approach, in Section~5 we describe the numerical
267: procedure developed to search for discrete ES solutions. Note that
268: this requires special adaptation or other methods for finding
269: homoclinic orbits of maps, owing to the non-hyperbolic nature of
270: the fixed point. Our approach heavily relies on the reversibility
271: although it can be readily modified to treat non-reversible ESs.
272: Numerical results are reported in Section~6, in the form of the
273: corresponding codimension-one solution branches in the parameter
274: space. An array of actual shapes of the ESs is displayed too. Our
275: numerical results also show that the asymptotic analysis of
276: Section~3 can well predict the codimension-one set of ESs in the
277: parameter space in the model when the wavenumber is large. The
278: paper is concluded by Section~7 along with a preliminary stability
279: result for the fundamental ES.
280:
281: \section{The models}
282:
283: We now consider Eqs.~(\ref{eqn:PDE}) in the spatial domain. Direct
284: discretization of the spatial second derivative produces a lattice
285: model
286: \begin{eqnarray}
287: &&i\frac{du_{n}}{dz}+\frac{1}{2}D(u_{n+1}+u_{n-1}-2u_{n})+u_{n}^{\ast
288: }v_{n}+\gamma _{1}(|u_{n}|^{2}+2|v_{n}|^{2})u_{n}=0, \nonumber \\[-1ex]
289: && \label{eqn:lattice} \\[-1ex]
290: &&i\frac{dv_{n}}{dz}+\frac{1}{2}\delta
291: D(v_{n+1}+v_{n-1}-2v_{n})+qv_{n}+\frac{1}{2}u_{n}^{2}+2\gamma
292: _{2}(|v_{n}|^{2}+2|u_{n}|^{2})v_{n}=0, \nonumber
293: \end{eqnarray}where $n$ is the discrete transverse coordinate which assumes integer
294: values, and $D=1/h^{2}$ with $h$ the stepsize of the
295: discretization. Effective coefficients of the lattice diffraction
296: for the FF and SH waves are $D_{1}\equiv D$ and $D_{2}=\delta D$;
297: %Note that straightforward discretization of the spatial-domain model,
298: with $\delta =1$, this yields $D_{2}=D_{1}$ in
299: Eqs.~(\ref{eqn:lattice}).
300: %However,
301: Unequal values of these coefficients, $D_{1}\neq D_{2}$ (i.e.,
302: $\delta \neq 1 $), including those with \emph{opposite signs}, can
303: in principle be realized experimentally (the latter possibility
304: was recently employed to predict the existence of discrete gap
305: solitons \cite{PanosZiad}). To do this one would use a
306: \textit{diffraction-management} technique, which is based on
307: oblique propagation of the beam across the waveguide array which
308: is modeled by the discrete system
309: \cite{DiffrManagement1,DiffrManagement2}. In that case, the
310: effective lattice-diffraction coefficients are
311: \begin{equation}
312: D_{m}=D_{m}^{(0)}\cos (mQ),\quad m=1,2, \label{management}
313: \end{equation}where $Q\equiv k\sin \theta $ is the transverse wavenumber, $k$ is the
314: beam's wavenumber (normalization is such that the array spacing is
315: unity), $\theta $ is the angle between the Poynting vector and the
316: coordinate $z$ running along the waveguides, and $D_{1,2}^{(0)}$
317: are the original lattice-dispersion coefficients, corresponding to
318: $Q=0$. As seen in Eq.~(\ref{management}), one can efficiently
319: control the size and signs of $D_{1,2}$ by choosing an appropriate
320: angle $\theta $. In particular, the ratio $D_{2}/D_{1}$ can be
321: made large by choosing $Q=\pi /2+\varepsilon $, or $Q=3\pi
322: /2-\varepsilon $, for a small positive $\varepsilon $.
323:
324: To cast the model in a normalized form, we note that $D_{1}$ can
325: be made positive, if it was originally negative, by means of the
326: complex conjugation (i.e., Eqs.~(\ref{eqn:lattice}) are replaced
327: by their counterparts for $u_{n}^{\ast }$ and $v_{n}^{\ast }$),
328: combined with the changes $v_{n}\rightarrow -v_{n}$ and $\gamma
329: _{1,2}\rightarrow -\gamma_{1,2}$. The size of $D_{1}$ may be set
330: equal to $1$ by means of the rescaling: $zD_{1}\rightarrow z$,
331: $\left( u_{n},v_{n}\right)/D_{1} \rightarrow \left(
332: u_{n},v_{n}\right)$, which leads to the final normalized form of
333: the discrete model,
334: \begin{eqnarray}
335: && i\frac{du_{n}}{dz}+\frac{1}{2}(u_{n+1}+u_{n-1}-2u_{n})+u_{n}^{\ast}v_{n}
336: +\tilde{\gamma}_{1}(|u_{n}|^{2}+2|v_{n}|^{2})u_{n}=0, \nonumber \\[-1ex]
337: && \label{normalised} \\[-1ex]
338: && i\frac{dv_{n}}{dz}+\frac{1}{2}\delta(v_{n+1}+v_{n-1}-2v_{n})
339: +\tilde{q}v_{n}+\frac{1}{2}u_{n}^{2}
340: +2\tilde{\gamma}_{2}(|v_{n}|^{2}+2|u_{n}|^{2})v_{n}=0, \nonumber
341: \end{eqnarray}
342: where
343: \begin{equation}
344: \tilde{q}=q/D_{1},\quad \tilde{\gamma}_{1,2}=D_{1}\gamma _{1,2}.
345: \label{delta}
346: \end{equation}
347: The dimensionless constants (\ref{delta}), including $\delta$, may, in
348: general, be positive, zero, or negative.
349:
350: To obtain some analytical results, we shall also introduce a simplified
351: model where ESs can be found in approximate form. In the simplified model,
352: only the central site (the one corresponding to $n=0$) carries the Kerr
353: nonlinearity. Such a model is physically feasible too, if the central
354: waveguide in the array is doped to enhance its Kerr nonlinearity. Thus, the
355: simplified model is based on the equations
356: \begin{eqnarray}
357: && i\frac{d u_{n}}{d z} +\frac{1}{2}\left(u_{n+1}+u_{n-1}-2u_{n}\right)+u_{n}^{\ast }v_{n}
358: +\tilde{\gamma}_{1}\epsilon_{n}\left(|u_{0}|^{2}+2|v_{0}|^{2}\right)u_{0}=0, \nonumber \\[-1ex]
359: && \label{eqn:mod} \\[-1ex]
360: && i\frac{d v_{n}}{d z} +\frac{1}{2}\delta
361: \left(v_{n+1}+v_{n-1}-2v_{n}\right)
362: +\tilde{q}v_{n}+\frac{1}{2}u_{n}^{2}
363: +2\tilde{\gamma}_{2}\epsilon_{n}\left(|v_{0}|^{2}+2|u_{0}|^{2}\right)v_{0}=0,
364: \nonumber
365: \end{eqnarray}
366: where $\epsilon _{n}=0$ for $n\neq 0$, and $\epsilon _{n}=1$ for $n=0$.
367:
368: In either model, (\ref{eqn:lattice}) or (\ref{eqn:mod}), the ES
369: corresponds to a solution with FF component exponentially
370: localized:
371: \begin{equation}
372: u_{n}\sim A\exp (i\tilde{k}z-\lambda |n|)\qquad
373: \mbox{as
374: $|n|\rightarrow\infty$}, \label{uasympt}
375: \end{equation}where $A$ is a real amplitude and $\lambda $ is positive and real.
376: Simultaneously, the propagation constant $\tilde{k}$ must belong to the
377: \textit{phonon band} of the SH equation. That is, the SH component of the
378: solution generically will have a nonvanishing tail, of the form
379: \begin{equation}
380: v_{n}\sim C\exp (2i\tilde{k}z-ip|n|)\qquad \mbox{as $|n|\rightarrow \infty$},
381: \label{vasympt}
382: \end{equation}where $p$ is real. The existence of an ES corresponds to the vanishing of
383: the constant $C$ in Eq.~(\ref{vasympt}). Linearization of
384: Eqs.~(\ref{eqn:lattice}) and (\ref{eqn:mod}) and subsequent
385: substitution of expressions (\ref{uasympt}) and (\ref{vasympt})
386: yield the following relations:
387: \begin{equation}
388: \tilde{k}=2\sinh ^{2}(\lambda /2), \label{k}
389: \end{equation}\begin{equation}
390: \tilde{q}-2\tilde{k}=2\delta \sin ^{2}(p/2). \label{q}
391: \end{equation}From Eqs.~(\ref{k}) and (\ref{q}) we see that a discrete ES may exist in the
392: case of $\tilde{k}>0$, with $\tilde{q}$ belonging to the region
393: \begin{equation}
394: 0<\tilde{q}-2\tilde{k}<2\delta \quad \mathrm{or}\quad 2\delta
395: <\tilde{q}-2\tilde{k}<0, \label{area}
396: \end{equation}depending on whether $\delta $ is positive or not. Note that we can only
397: choose $\tilde{k}$, $\tilde{q}$ and $\delta $ independently in the
398: set of $(\tilde{k},\tilde{q},\delta ,\lambda ,p)$ since $\lambda $
399: and $p$ are determined by the other three parameters via Eqs.
400: (\ref{k}) and (\ref{q}).
401:
402: \section{Asymptotic analysis for the simplified model}
403:
404: We first consider the simplified model (\ref{eqn:mod}), with the
405: $\chi ^{(3)} $ nonlinearity present only at the central site. An
406: ES solution is sought asymptotically under the assumption that
407: $|v_{n}|\ll |u_{n}|=O(1)$, the validity of which will be checked
408: \textit{a posteriori}. Accordingly, the quadratic term in the
409: first equation of the system (\ref{eqn:mod}) may be dropped to
410: first approximation, which makes the expressions (\ref{uasympt})
411: and (\ref{k}) an \emph{exact solution} for $n\neq 0$. At the point
412: $n=0$, the cross-phase-modulation term, $|v_{0}|^{2}u_{0}$, in the
413: first equation (\ref{eqn:mod}) may be dropped too to leading
414: order. Then, the equation at $n=0$ yields a final result for the
415: FF component of the soliton, $A^{2}=\tilde{\gamma}_{1}^{-1}\sinh
416: \lambda $, which implies that the discrete soliton in the FF
417: component is supported by itself, without coupling to the SH
418: component. Further, using Eq.~(\ref{k}) one can express $A^{2}$ in
419: terms of $\tilde{k}$,
420: \begin{equation}
421: A^{2}=\tilde{\gamma}_{1}^{-1}\sqrt{\tilde{k}(\tilde{k}+2)}, \label{Ak}
422: \end{equation}
423: which means that the solution exists only in the case of
424: $\tilde{\gamma}_{1}>0$.
425:
426: Now, we tackle the second equation of (\ref{eqn:mod}) and
427: substitute the FF field in the form of Eqs.~(\ref{uasympt}) and
428: (\ref{k}). Obviously, at $n\neq 0$, an exact solution can be
429: found, which precisely corresponds to an ES, as it does not
430: contain the nonvanishing tail (\ref{vasympt}) and is `tail-locked'
431: to the square of the FF field, as explained in the Introduction.
432: We find explicitly that
433: \begin{equation}
434: v_{n}=B\exp (2i\tilde{k}z-2\lambda |n|), \label{vsolution}
435: \end{equation}where
436: \begin{equation}
437: B=-\frac{A^{2}}{2[2\delta \sinh ^{2}\lambda
438: +(\tilde{q}-2\tilde{k})]}\equiv
439: -\frac{1}{2\tilde{\gamma}_{1}}\frac{\sqrt{\tilde{k}(\tilde{k}+2)}}{2\delta
440: \cdot \tilde{k}(\tilde{k}+2)+(\tilde{q}-2\tilde{k})}. \label{B}
441: \end{equation}Substituting the expressions (\ref{uasympt}) and (\ref{vsolution}) into the
442: second equation of (\ref{eqn:mod}) at $n=0$, we have
443: \begin{equation}
444: \cosh \lambda =\frac{2\tilde{\gamma}_{2}}{\delta \tilde{\gamma}_{1}}.
445: \label{eqn:lambda}
446: \end{equation}Here, following the assumption $|v_{n}|\ll |u_{n}|$, we neglected the
447: self-phase-modulation term $|v_{0}|^{2}v_{0}$ at this point.
448: Equation~(\ref{eqn:lambda}) also implies that
449: \begin{equation}
450: \frac{2\tilde{\gamma}_{2}}{\delta \tilde{\gamma}_{1}}>1, \label{eqn:12}
451: \end{equation}and, especially, that $\delta $ has the same sign as $\tilde{\gamma}_{2}$
452: since $\lambda ,\,\tilde{\gamma}_{1}>0$. Finally, using the
453: relations (\ref{k}) and (\ref{eqn:lambda}), we obtain
454: \begin{equation}
455: \tilde{k}=\frac{2}{\delta }\frac{\tilde{\gamma}_{1}}{\tilde{\gamma}_{2}}-1,
456: \label{final}
457: \end{equation}which is positive by virtue of (\ref{eqn:12}).
458: This result selects the \emph{single value} of $\tilde{k}$ at which the simplified model admits the
459: existence of an ES, with tail-locked SH component. It follows from
460: Eq.~(\ref{final}) that such a solution exists if the relative
461: lattice-diffraction coefficient (see Eq.~(\ref{delta})) belongs to
462: the interval
463: \begin{equation}
464: 0<|\delta |<2\tilde{\gamma}_{1}/|\tilde{\gamma}_{2}|,\quad
465: \tilde{\gamma}_{1}>0, \label{12+}
466: \end{equation}where $\delta $ and $\tilde{\gamma}_{2}$ must be of the same sign.
467:
468: It is now necessary to check compatibility of the solution with
469: the underlying assumption, $|v_{n}|\ll |u_{n}|$, which implies
470: $A^{2}\ll B^{2}$ (see Eqs.~(\ref{Ak}) and (\ref{B})).
471: Straightforward consideration shows that this condition amounts to
472: $|\delta|\ll|\tilde{\gamma}_{1,2}|$. We also note that, if the FF
473: component of the soliton is strongly localized, so that
474: $e^{-\lambda }\ll 1$ (see Eq.~(\ref{uasympt})), the simplified
475: model is actually equivalent to the full one, as the nonlinearity
476: in the first equation of (\ref{eqn:lattice}) is then negligible at
477: $n\neq 0$. From (\ref{k}) we see that this condition implies that
478: $\tilde{k}$ must be large and hence, from (\ref{final}), that
479: $|\delta|\ll 1$.
480:
481: A noteworthy feature of Eqs.~(\ref{final}) and (\ref{12+}) is that
482: they do not involve the mismatch parameter $\tilde{q}$. However,
483: the condition that the soliton found above is embedded implies
484: that $\tilde{k}$ given by Eq.~(\ref{final}) must belong to the
485: interval (\ref{area}), which relates $\tilde{k}$ to $\tilde{q}$,
486: as $|\delta|\ll 1$. This approximate analysis, valid too for the
487: original model in the limit of small $|\delta|$, means that, for
488: an isolated selected $\tilde{k}$-value given by (\ref{final}),
489: there exists a curve of single-humped ESs approximately
490: parametrized by $\tilde{q}$ in the narrow interval
491: \begin{equation}
492: \tilde{q}\in (2\tilde{k},2\tilde{k}+2\delta) \quad\mbox{for
493: $\delta>0$}\quad\mbox{or}\quad (2\tilde{k}+2\delta,2\tilde{k})
494: \quad\mbox{for $\delta<0$} \label{qinterval}
495: \end{equation}
496: in the $(\tilde{q},\tilde{k})$-plane for $\delta$ fixed.
497:
498: \section{Reversible maps}
499:
500: Let us now consider the model (\ref{eqn:lattice}) in the general
501: case, and understand, from a dynamical systems point of view, why
502: discrete ESs should exist. In this context, one of essential
503: issues is the transition to the continuum limit. The scaling
504: leading to Eq. (\ref{normalised}), in which $D_{1}=1$, does not
505: allow this. Hence, we undo this scaling and set $D_{1}=D$,
506: $D_{2}=\delta D$. As above, we look for stationary solutions in
507: the form
508: \begin{equation}
509: u_{n}=U_{n}e^{ikz},\quad v_{n}=V_{n}e^{2ikz}, \label{eqn:UV}
510: \end{equation}where $U_{n}$ and $V_{n}$ are real. Equations~(\ref{eqn:lattice}) thus
511: reduce to
512: \begin{eqnarray}
513: &&\frac{1}{2}D(U_{n+1}+U_{n-1}-2U_{n})-kU_{n}+U_{n}V_{n}+\gamma
514: _{1}(U_{n}^{2}+2V_{n}^{2})U_{n}=0, \nonumber \\[-1ex]
515: && \label{eqn:disc} \\[-1ex]
516: &&\frac{1}{2}\delta
517: D(V_{n+1}+V_{n-1}-2V_{n})+(q-2k)V_{n}+\frac{1}{2}U_{n}^{2}+2\gamma
518: _{2}(V_{n}^{2}+2U_{n}^{2})V_{n}=0, \nonumber
519: \end{eqnarray}where parameters without the tildes are related to those with tildes by
520: \begin{equation}
521: q=\tilde{q}D,\quad k=\tilde{k}D,\quad \gamma _{1,2}=\tilde{\gamma}_{1,2}/D
522: \label{rescaledpars}
523: \end{equation}Note that the possible existence regions of ESs, given by expression (\ref{area}), now becomes
524: \begin{equation}
525: 0<q-2k<2\delta D\quad \mathrm{or}\quad 2\delta D<q-2k<0. \label{area_scale}
526: \end{equation}
527:
528: First and foremost, we want to establish the codimension of ES
529: solutions to (\ref{eqn:disc}). In order to do that, it is useful
530: to recast the system as a four-dimensional map. Specifically, upon
531: scaling the variables as $\xi _{n}\equiv \sqrt{2|\gamma _{1}|/D}\
532: U_{n}$ and $\eta _{n}\equiv \sqrt{2|\gamma _{1}|/D}\ V_{n}$, we
533: can indeed view Eqs.~(\ref{eqn:disc}) as defining a\emph{\
534: four-dimensional map},
535: \begin{equation}
536: \zeta _{n+1}=F_{\epsilon }(\zeta _{n}), \label{eqn:F}
537: \end{equation}
538: where $\zeta _{n}\equiv (\chi _{n},\xi _{n},\mu _{n},\eta _{n})$ and
539: \[
540: F_{\epsilon }(\zeta _{n})=\left(
541: \begin{array}{c}
542: \xi _{n} \\
543: f_{\epsilon }^{(1)}(\zeta _{n}) \\
544: \eta _{n} \\
545: f_{\epsilon }^{(2)}(\zeta _{n})\end{array}\right) .
546: \]
547: Here we define
548: \begin{eqnarray}
549: f_{\epsilon }^{(1)}(\zeta _{n}) &=&-\chi _{n}+2\nu _{1}\xi _{n}-(\epsilon
550: \xi _{n}\eta _{n}+\kappa _{1}(\xi _{n}^{2}+2\eta _{n}^{2})\xi _{n}), \\
551: f_{\epsilon }^{(2)}(\zeta _{n}) &=&-\mu _{n}+2\nu _{2}\eta _{n}-\delta
552: ^{-1}\left( \frac{1}{2}\epsilon \xi _{n}^{2}+2\kappa _{2}(\eta _{n}^{2}+2\xi
553: _{n}^{2})\eta _{n}\right) ,
554: \end{eqnarray}
555: where
556: \begin{eqnarray}
557: \nu _{1} &=&1+\frac{k}{D},\quad \nu _{2}\equiv 1+\frac{2k-q}{\delta D},\quad
558: \epsilon \equiv \sqrt{\frac{2}{D|\gamma _{1}|}}, \nonumber \\[-1.5ex]
559: && \label{eqn:para} \\[-1.5ex]
560: \kappa _{1} &\equiv &{\mathrm{sgn}}(\gamma _{1}),\quad \kappa _{2}\equiv
561: \gamma _{2}/\left\vert \gamma _{1}\right\vert . \nonumber
562: \end{eqnarray}
563:
564: The map $F_{\epsilon }$ is \textit{reversible} \cite{De:76b,LaRo:98}, in the
565: sense that $F_{\epsilon }^{-1}(R(\zeta _{n}))=R(F_{\epsilon }(\zeta _{n}))$
566: holds, where $R:\mathbb{R}^{4}\rightarrow \mathbb{R}^{4}$ is the (linear)
567: involution given by
568: \begin{equation}
569: R:(\chi _{n},\xi _{n},\mu _{n},\eta _{n})\mapsto (\xi _{n},\chi _{n},\eta
570: _{n},\mu _{n}). \label{R1}
571: \end{equation}
572: Note that the map $F_{\epsilon }$ is also reversible, under the second
573: reversibility
574: \begin{equation}
575: \hat{R}:(\chi _{n},\xi _{n},\mu _{n},\eta _{n})\mapsto (f_{\epsilon
576: }^{(1)}(\zeta _{n}),\xi _{n},f_{\epsilon }^{(2)}(\zeta _{n}),\eta _{n}).
577: \label{R2}
578: \end{equation}
579: In what follows, we consider orbits of the map that are reversible
580: with respect to $R$. These will correspond to ESs that have their
581: central peaks on a lattice site, which are the kind of lattice
582: solitons that are most frequently observed in physical
583: applications. In contrast, ESs that are symmetric under $\hat{R}$
584: would have their central peak between two lattice sites. Both $R$
585: and $\hat{R}$ arise from the fact that the ordinary differential
586: equations for steady solutions of the continuum model
587: (\ref{eqn:PDE}),
588: \begin{eqnarray}
589: &&\frac{1}{2}\frac{d^{2}U}{dt^{2}}-kU+UV+\gamma _{1}(U^{2}+2V^{2})U=0,
590: \nonumber \\[-1ex]
591: && \label{eqn:cont} \\[-1ex]
592: &&\frac{\delta }{2}\frac{d^{2}V}{dt^{2}}+(q-2k)V+\frac{1}{2}U^{2}+2\gamma
593: _{2}(V^{2}+2U^{2})V=0, \nonumber
594: \end{eqnarray}
595: are reversible too, under an involution
596: $\tilde{R}:(U,dU/dt,V,dV/dt)\mapsto(U,-dU/dt,V,-dV/dt)$.
597:
598: A fundamental characteristic of reversible maps is that if
599: $\{\zeta_{n}\}_{n=-\infty }^{\infty }$ is an orbit then
600: $\{R(\zeta_{-n})\}_{n=-\infty }^{\infty }$ is also an orbit. We
601: say that an orbit $\{\zeta _{n}\}_{n=-\infty}^{\infty}$ is
602: \textit{symmetric} (with respect to the reversibility) if
603: $\zeta_{n+1}=R(\zeta _{-n})$. Denote $\mathrm{Fix}(R)=\{\zeta \in
604: \mathbb{R}^{4}\,|\,F_{\epsilon }(\zeta )=R(\zeta )\}$. We easily
605: see that $\zeta =(\chi ,\xi ,\mu ,\eta )\in \mathrm{Fix}(R)$ if
606: and only if $\chi =f_{\epsilon }^{(1)}(\zeta )$ and $\mu
607: =f_{\epsilon }^{(2)}(\zeta )$. Thus, $\mathrm{Fix}(R)$ depends on
608: the particular form of the map $F_{\epsilon }$. An orbit $\{\zeta
609: _{n}\}_{n=-\infty }^{\infty }$ is symmetric if and only if $\zeta
610: _{0}\in \mathrm{Fix}(R)$, i.e., $\xi _{-1}=\xi _{1}$ and $\eta
611: _{-1}=\eta _{1}$.
612:
613: The map $F_{\epsilon }$ has a fixed point at the origin $O$, whose Jacobian
614: matrix is
615: \[
616: J=\left(
617: \begin{array}{cccc}
618: 0 & 1 & 0 & 0 \\
619: -1 & 2\nu _{1} & 0 & 0 \\
620: 0 & 0 & 0 & 1 \\
621: 0 & 0 & -1 & 2\nu _{2}\end{array}\right) .
622: \]It follows from Eq.~(\ref{area_scale}) that $\nu _{1}>1$ and $|\nu _{2}|<1$,
623: hence the origin is a fixed point of $F_{\epsilon }$ of
624: saddle-center type. The saddle-center has one-dimensional stable
625: and unstable manifolds, $W^{s}(O)$ and $W^{u}(O)$, which are
626: tangent to the stable and unstable subspaces spanned by the
627: vectors $(1,\nu _{1}-\sqrt{\nu _{1}^{2}-1},0,0)$ and $(1,\nu
628: _{1}+\sqrt{\nu _{1}^{2}-1},0,0)$, respectively, and a
629: two-dimensional center manifold, $W^{c}(O)$, which is tangent to
630: the center subspace spanned by a set of two vectors, $(0,0,1,0)$
631: and $(0,0,0,1)$. By the fundamental property of reversible maps,
632: $W^{s}(O)=R(W^{u}(O))$ and $W^{u}(O)=R(W^{s}(O))$. Thus, if there
633: exists an orbit $\{\zeta _{n}\}_{n=-\infty }^{\infty }$ on
634: $W^{u}(O)$ such that $\zeta _{0}\in \mathrm{Fix}(R)$, then it is
635: also contained in $W^{s}(O)$ and is a symmetric homoclinic orbit
636: to $O$.
637:
638: If $F_{\epsilon }$ were not reversible, then such intersections
639: between the one-dimensional stable and unstable manifolds in the
640: four-dimensional phase space would be of codimension two. However,
641: symmetric homoclinic orbits are of codimension one, since, for
642: their existence, we require an intersection between the
643: one-dimensional unstable manifold $W^{u}(O)$ and the
644: two-dimensional manifold $\mathrm{Fix}(R)$. Thus, since a
645: homoclinic orbit to $O$ for $F_{\epsilon }$ represents precisely
646: an ES in the lattice system (\ref{eqn:disc}), we have established
647: that:
648:
649: \noindent \emph{Embedded solitons of the lattice model are of
650: \textbf{codimension-one} in the parameter space, provided they are
651: symmetric under a reversibility equivalent to $R$ (or $\hat{R}$)}.
652:
653: \noindent Note that this is identical to the multiplicity result known in
654: the continuum version of the model \cite{ChMaYaKa:01}.
655:
656: Finally, in what follows we shall also treat the case of pure
657: $\chi ^{(2)}$ nonlinearity, $\gamma _{1}=\gamma _{2}=0$. This case
658: is physically important, as experiments could be quite conceivably
659: be set up in a medium with negligible Kerr nonlinearity. However,
660: without the $\chi^{(3)}$ terms, no ES exists in the continuum
661: model \cite{ChMaYaKa:01}. It will therefore be important to find
662: out whether the same is true in the discrete model too.
663:
664: With $\gamma _{1}=\gamma _{2}=0$, the scaling leading to $F_{\epsilon}$
665: becomes invalid, so we shall consider instead a family of four-dimensional
666: maps
667: \begin{equation}
668: \zeta _{n+1}=G_{\epsilon }(\zeta _{n}), \label{eqn:G}
669: \end{equation}
670: where we define
671: \[
672: G_{\epsilon }(\zeta _{n})=\left(
673: \begin{array}{c}
674: \xi _{n} \\
675: g_{\epsilon }^{(1)}(\zeta _{n}) \\
676: \eta _{n}, \\
677: g_{\epsilon }^{(2)}(\zeta _{n})\end{array}\right) ,
678: \]
679: with
680: \begin{eqnarray}
681: g_{\epsilon }^{(1)}(\zeta _{n}) &=&-\chi _{n}+2\nu _{1}\xi _{n}-(\epsilon
682: \xi _{n}\eta _{n}+(1-\epsilon )(\xi _{n}^{2}+2\eta _{n}^{2})\xi _{n}), \\
683: g_{\epsilon }^{(2)}(\zeta _{n}) &=&-\mu _{n}+2\nu _{2}\eta _{n}-\delta
684: ^{-1}\left( \frac{1}{2}\epsilon \xi _{n}^{2}+2(1-\epsilon )(\eta
685: _{n}^{2}+2\xi _{n}^{2})\eta _{n}\right),
686: \end{eqnarray}
687: with $\nu _{1}$ and $\nu _{2}$ given by (\ref{eqn:para}). The map
688: $G_{\epsilon }$ is reversible under the same involutions $R$ and
689: $\hat{R}$. The map $G_{1}$ is equivalent to (\ref{eqn:disc}) with
690: $\gamma _{1}=\gamma _{2}=0$ if one sets $\xi _{n}=U_{n}$ and $\eta
691: _{n}=V_{n}$, and simultaneously $G_{0}$ coincides with $F_{0}$
692: with $\kappa_{1}=\kappa _{2}=1 $. The origin $O$ is also a
693: saddle-center of $G_{\epsilon }$ and has the same stable, unstable
694: and center subspaces as those of $F_{\epsilon }$. So we can apply
695: all the arguments presented above for $F_{\epsilon }$ to
696: $G_{\epsilon }$ if we replace $F_{\epsilon }$ with $G_{\epsilon }$
697: in the definition of $\mathrm{Fix}(R)$. In particular, a
698: homoclinic orbit to $O$ for $G_{\epsilon }$ represents an ES in
699: the lattice system (\ref{eqn:lattice}) with $\gamma _{1}=\gamma
700: _{2}=0$.
701:
702: \section{Numerical procedure}
703:
704: Several numerical procedures exist for finding homoclinic orbits to fixed
705: points in maps, see, e.g., \cite{K81,BK97a,BK97b,Y98a,BBV00}. Here we shall
706: use an adaptation of these methods that takes into regard both the
707: reversibility and the non-hyperbolic nature of the fixed point.
708:
709: To compute symmetric homoclinic orbits for $F_{\epsilon }$, we consider the
710: three-dimensional algebraic problem
711: \begin{eqnarray}
712: &&\chi _{-N}-\left( \nu _{1}-\sqrt{\nu _{1}^{2}-1}\right) \xi _{-N}=0,
713: \label{eqn:con1} \\
714: &&\xi _{1}=\xi _{-1},\quad \eta _{1}=\eta _{-1}, \label{eqn:con2}
715: \end{eqnarray}
716: for $N>0$ sufficiently large, where $\zeta_{\pm 1}
717: \equiv(\chi_{\pm 1},\xi_{\pm 1},\mu_{\pm 1},\eta_{\pm 1})
718: =F_{\epsilon }^{N\pm 1}(\chi_{-N},\xi_{-N},0,0)$. The condition
719: (\ref{eqn:con1}) means that the point $(\chi _{-N},\xi _{-N},0,0)$
720: lies in the one-dimensional unstable subspace of the fixed point
721: at the origin, and condition (\ref{eqn:con2}) means that $\zeta
722: _{0}=(\chi _{0},\xi _{0},\mu_{0},\eta _{0})\in \mathrm{Fix}(R)$,
723: i.e., the finite orbit $\{\zeta_{n}\}_{n=-N}^{0}$ intersects
724: $\mathrm{Fix}(R)$ at $n=0$. Thus, adding a parameter as an extra
725: unknown variable, Eqs.~(\ref{eqn:con1}) and (\ref{eqn:con2})
726: represent a formally well-posed system of three equations for two
727: unknowns, $\chi _{-N}$ and $\xi _{-N}$, and the free parameter,
728: which can be chosen to be any of $\epsilon$, $\delta$, $k$, $D$,
729: $q$, $\kappa_{1}$ or $\kappa_{2}$. A non-trivial solution for $N$
730: sufficiently large gives an approximate homoclinic orbit $\{\zeta
731: _{n}\}_{n=-N}^{N+1}$ of $F_{\epsilon }$, that is symmetric under
732: $R$, which in turn corresponds to a discrete ES solution of the
733: lattice equation (\ref{eqn:lattice}) for $\gamma _{1,2}\neq 0$. In
734: the case $\gamma _{1,2}=0$, the same treatment can be used to find
735: symmetric homoclinic orbits for $G_{\epsilon }$. Note that one
736: test of the validity of this approximation is to measure the
737: distance of the point $(\chi _{-N},\xi _{-N},0,0)$ from the
738: origin. By virtue of the stable manifold theorem for maps
739: \cite{GH83}, we know that the error will be proportional to this
740: distance squared.
741:
742: Branches of solutions to Eqs.~(\ref{eqn:con1}) and
743: (\ref{eqn:con2}) can be continued in a second parameter using
744: pseudo-arclength continuation. To this end, we use the code
745: \texttt{AUTO} \cite{Auto}. The problem now is finding a good
746: initial point along a branch of discrete ESs. Here we can use a
747: regular perturbation approach by first finding solutions to the
748: map $F_{0}$, making use of the fact that when $\epsilon =0$, the
749: $(\chi _{n},\xi _{n})$-plane is invariant under $F_{\epsilon }$.
750: The restriction of $F_{0}$ onto the invariant plane is given by
751: \begin{equation}
752: (\chi _{n+1},\xi _{n+1})=f(\chi _{n},\xi _{n}), \label{eqn:2d}
753: \end{equation}
754: where
755: \[
756: f(\chi _{n},\xi _{n})\equiv (\xi _{n},-\chi _{n}+2\nu _{1}\xi _{n}-\kappa
757: _{1}\xi _{n}^{3}).
758: \]The two-dimensional map $f$ also has a hyperbolic saddle at the origin, and
759: is reversible under an involution,
760: \[
761: \bar{R}:(\chi _{n},\xi _{n})\mapsto (\xi _{n},\chi _{n}).
762: \]The stable (resp. unstable) manifold of the saddle, $\bar{W}^{s}(O)$
763: (resp. $\bar{W}^{u}(O)$), is tangent to its stable (resp. unstable) subspace spanned
764: by a vector $(1,\nu _{1}-\sqrt{\nu _{1}^{2}-1})$ (resp.
765: $(1,\nu_{1}+\sqrt{\nu _{1}^{2}-1}$)). By the reversibility,
766: $\bar{W}^{s}(O)=\bar{R}(\bar{W}^{u}(O))$ and vice versa. When $\nu
767: _{1}>0$ and $\kappa _{1}>0$, the stable and unstable manifolds
768: intersect transversely and form homoclinic tangles, as shown in
769: Fig.~\ref{fig:3a}.
770:
771: \begin{figure}[t]
772: \begin{center}
773: \includegraphics[scale=1.2]{fig1.eps}
774: \end{center}
775: \caption{Homoclinic tangles for the two-dimensional map $f$ with
776: $\protect\nu_1=1.25$ and $\protect\kappa _{1}=1$, drawn using the
777: software \texttt{Dynamics} \protect\cite{NY97}. The circle
778: ``$\bullet$'' represents the saddle at the origin.} \label{fig:3a}
779: \end{figure}
780:
781: Let $N>0$ be a sufficiently large integer, and let
782: $(\bar{\chi}_{-N},\bar{\xi}_{-N})$ be a point on the unstable
783: subspace such that
784: $(\bar{\xi}_{N+1},\bar{\chi}_{N+1})=(\bar{\chi}_{-N},\bar{\xi}_{-N})$,
785: where
786: $(\bar{\chi}_{N+1},\bar{\xi}_{N+1})=f^{2N+1}(\bar{\chi}_{-N},\bar{\xi}_{-N})$.
787: By the reversibility of $f$, the point
788: $(\bar{\chi}_{N+1},\bar{\xi}_{N+1})$ must be contained in the
789: stable subspace. The two points $(\bar{\chi}_{-N},\bar{\xi}_{-N})$
790: and $(\bar{\chi}_{N+1},\bar{\xi}_{N+1})$ are also close to the
791: saddle when $N>0$ is large. Hence, the orbit leaving at
792: $(\bar{\chi}_{-N},\bar{\xi}_{-N})$ on the unstable subspace and
793: arriving at $(\bar{\chi}_{N+1},\bar{\xi}_{N+1})$ on the stable
794: subspace is a good approximation to a symmetric homoclinic orbit
795: of $f$. Using an adaptation of the driver \texttt{HomMap}
796: \cite{Y98a,Y98b} to \texttt{AUTO97}, we can easily find such
797: approximate homoclinic points.
798:
799: Now, in order to find non-trivial symmetric homoclinic solutions
800: of $F_{\epsilon }$ for $\epsilon >0$, we take $\epsilon $ as the
801: additional free parameter and choose $(\chi _{-N},\xi
802: _{-N},\epsilon )=(\bar{\chi}_{-N},\bar{\xi}_{-N},0)$ as the
803: starting solution to (\ref{eqn:con1}) and (\ref{eqn:con2}), where
804: $(\bar{\chi}_{-N},\bar{\xi}_{-N})$ denotes the homoclinic point on
805: the unstable subspace for $f$, obtained using the above procedure.
806: Fixing $\kappa _{1}=1$, we then performed continuation of these
807: algebraic equations in \texttt{AUTO}, using $\delta$ as the
808: continuation parameter. To obtain symmetric homoclinic orbits for
809: $\kappa _{1}=-1$, we also take $\delta $ and $\kappa _{1}$ as the
810: free and continuation parameters, respectively, and continue the
811: solution obtained above for $\kappa _{1}=1$ and $\epsilon
812: =\sqrt{2/(D|\gamma _{1}|)}$. The results are presented in
813: Figs.~\ref{fig:3b}-\ref{fig:3d}. As shown in Figs.~\ref{fig:3b}
814: and \ref{fig:3c}, new branches bifurcate from the one with
815: $\epsilon=0$ at discrete values of $\delta$ and can be continued
816: to $\epsilon=\sqrt{2/(D|\gamma_{1}|)} $ by varying $\delta $ and
817: $\epsilon $. As shown in Fig.~\ref{fig:3d}, the branches were also
818: continued from $\kappa _{1}=1$ to $\kappa _{1}=-1$ by varying
819: $\delta $ and $\kappa_{1}$. For $\kappa _{1}=-1$ and $\kappa
820: _{2}=1$, we could not find such symmetric homoclinic orbits of
821: $F_{\epsilon }$ with sufficient precision. In these computations,
822: we also chose the value of $N$ such that the distance between the
823: approximate homoclinic point and the origin was $1.5\times
824: 10^{-3}$ at most, and checked that the results did not change
825: significantly under increase of $N$. Our computations also
826: suggested that a possibly infinite number of branches of symmetric
827: homoclinic orbits could be obtained when $k$ or $q$ is large or
828: $\delta $ is small. These branches show oscillations in the
829: parameter plane, which are sensitive to the value of $N$ and do
830: not appear to converge as $N\rightarrow \infty$. These, probably
831: spurious branches, result from a large ratio between the imaginary
832: eigenvalue of the linearization and the real eigenvalue, which is
833: well known to be a singular limit for equations that bear ES
834: \cite{Ch:01,Lo:01}, and needs to be treated with great caution. In
835: the results that follow, we have stopped computation at points
836: where such oscillations first become evident, and have checked all
837: results for consistency in the limit of $N\rightarrow \infty $.
838:
839: \begin{figure}[tbp]
840: \begin{center}
841: \includegraphics[scale=0.7]{fig2a.eps}\quad \includegraphics[scale=0.7]{fig2b.eps}
842: \end{center}
843: \caption{Numerical continuation of symmetric homoclinic orbits for
844: $F_{\protect\epsilon }$ when $D=100$ and $N=85$: (a) $k=0.3$,
845: $q=5$ and $\protect\kappa _{1}=\protect\kappa _{2}=1$; (b)
846: $k=0.5$, $q=1$, $\protect\kappa _{1}=1$ and $\protect\kappa
847: _{2}=-1$. Here $\protect\epsilon $ and $\protect\delta $ are
848: varied. The dotted line represents $\protect\epsilon
849: =\protect\sqrt{2/(D|\protect\gamma _{1}|)}$ with $\protect\gamma
850: _{1}=0.05$. } \label{fig:3b}
851: \end{figure}
852:
853: \begin{figure}[tbp]
854: \begin{center}
855: \includegraphics[scale=0.7]{fig3.eps}
856: \end{center}
857: \caption{Numerical continuation of symmetric homoclinic orbits for
858: $G_{\protect\epsilon }$ when $D=10$, $k=3$, $q=1$ and $N=15$. Here
859: $\protect\epsilon $ and $\protect\delta $ are varied. The dotted
860: line represents $\protect\epsilon =1$. } \label{fig:3c}
861: \end{figure}
862:
863: \begin{figure}[tbp]
864: \begin{center}
865: \includegraphics[scale=0.7]{fig4.eps}
866: \end{center}
867: \caption{Numerical continuation of symmetric homoclinic orbits for
868: $F_{\protect\epsilon }$ when $D=100$, $k=0.5$, $q=1$,
869: $\protect\kappa _{2}=-1 $, $N=85$ and $\protect\epsilon
870: =0.6324555\approx \protect\sqrt{2/\left( D|\protect\gamma
871: _{1}|\right) }$ with $\protect\gamma _{1}=0.05$. Here
872: $\protect\delta $ and $\protect\kappa _{1}$ are varied. The dotted
873: line represents $\protect\kappa_{1}=-1$. } \label{fig:3d}
874: \end{figure}
875:
876: \section{Numerical results}
877:
878: We now present continuation results obtained with \texttt{AUTO}
879: for symmetric homoclinic orbits of $F_{\epsilon }$ (or
880: $G_{\epsilon }$) under simultaneous variation of two relevant
881: parameters within the saddle-center parameter regions. By varying
882: the discreteness coefficient $D$ up to large values, we are also
883: able to compare the results to those of the continuum limit,
884: $D\rightarrow \infty $. We shall treat the cases $\delta >0$ and
885: $\delta <0$ separately and also discuss the possibility of
886: discrete ESs in the absence of cubic nonlinearity.
887:
888: \subsection{The case of\/ $\protect\delta >0$}
889:
890: Figure~\ref{fig:4a} depicts branches of discrete ESs in the
891: presence of $\chi ^{(3)}$ terms, with $\gamma
892: _{1}=\gamma_{2}=0.05$ and $k=0.3$. Two distinct branches of ESs
893: are displayed. Along each branch, the profile changes
894: continuously. Figure~\ref{fig:4b} displays an array of profiles of
895: these ESs for $D=100$ and $D=5$. Note that the second
896: (higher-$\delta$) branch can be considered to be a family of
897: \emph{fundamental}, i.e.\ single-humped, solitons throughout the
898: range of existence, whereas the double-humped structure of the SH
899: component of the first branch becomes evident for small $q$. This
900: property, and the location of the branches in the $(\delta ,q)$
901: plane, are fully consistent with results obtained in the continuum
902: limit \cite[Fig.~3]{ChMaYaKa:01}. Our numerical results also
903: suggest that there may exist further branches for lower $\delta
904: $-values, each subsequent one containing more oscillations in core
905: of the SH component. However, computation becomes unreliable
906: beyond the first two branches for the reasons given at the end of
907: the last section, and so we do not display those results here.
908: Also experience from the continuum model suggests that
909: non-fundamental ESs never have a chance to be stable.
910:
911: \begin{figure}[tbp]
912: \begin{center}
913: \includegraphics[scale=0.7]{fig5a.eps}\quad \includegraphics[scale=0.7]{fig5b.eps}
914: \end{center}
915: \caption{Branches of the discrete ES solutions for $k=0.3$ and
916: $\protect\gamma _{1}=\protect\gamma _{2}=0.05$. (a) Solutions in
917: the $(\protect\delta ,q)$-plane for $(D,N)=(100,86)$ (solid line)
918: $(10,31)$ (dashed line) and $(5,23)$ (broken line). According to
919: (\protect\ref{area_scale}) ESs also exist only above the dotted
920: line $q=0.6$ for this $k$-value. Labeled points correspond to the
921: relevant subpanels of Fig.~\protect\ref{fig:4b} where solutions
922: profiles are displayed. (b) Solutions in the $(D,\protect\delta
923: )$-plane for $q=5$ and $N=22$. Note that, according to
924: (\protect\ref{area_scale}), we must have $\protect\delta D>2.2$
925: (the boundary is shown as a dotted line). } \label{fig:4a}
926: \end{figure}
927:
928: \begin{figure}[tbp]
929: \begin{center}
930: \includegraphics[scale=0.55]{fig6a.eps}\qquad \includegraphics[scale=0.55]{fig6b.eps}\\[2ex]
931: \includegraphics[scale=0.55]{fig6c.eps}\qquad \includegraphics[scale=0.55]{fig6d.eps}\\[2ex]
932: \includegraphics[scale=0.55]{fig6e.eps}\qquad \includegraphics[scale=0.55]{fig6f.eps}\\[2ex]
933: \includegraphics[scale=0.55]{fig6g.eps}\qquad \includegraphics[scale=0.55]{fig6h.eps}
934: \end{center}
935: \caption{Profiles of the discrete ESs corresponding to the labeled
936: points in Fig.~\protect\ref{fig:4a}(a): (a) $\protect\delta
937: =0.58637 $ and $q=5$; (b) $\protect\delta =2.9894$ and $q=5$; (c)
938: $\protect\delta =0.65467$ and $q=0.6$; (d) $\protect\delta
939: =3.3693$ and $q=0.6$; (e) $\protect\delta =1.3496 $ and $q=5$; (f)
940: $\protect\delta =3.1432$ and $q=5$; (g) $\protect\delta =0.80213$
941: and $q=0.6$; (h) $\protect\delta =3.3824$ and $q=0.6$. In
942: panels~(a)-(d), $D=100$ and $N=85$, and in panels (e)-(h), $D=5$
943: and $N=22$. In this and all subsequent plots, the FF
944: ($u$-component) is interpolated by a solid line and the SH
945: ($v$-component) by a broken line. } \label{fig:4b}
946: \end{figure}
947:
948: Figure \ref{fig:4a}(b) depicts continuation in $D$ of the solutions on these
949: two branches for $q=5$. In this case, we find that there is a lower limit on
950: $D$ beyond which no ESs exist. This corresponds to the right-hand inequality
951: in the first condition in (\ref{area_scale}).
952: %This is symptomatic of the fact that in the anticontinuum limit
953: %($D\rightarrow 0$) only a narrow band of $q$ values is allowed, as
954: %shown by the asymptotic analysis (\ref{qinterval}).
955:
956: We will now check the validity of the analytical approximation
957: developed in Section~3 when $\tilde{k}$ and hence $k$ are large.
958: In Fig.~\ref{fig:4b1}, we display the two numerically computed ES
959: branches with the same values of $\gamma _{1}$ and $\gamma _{2}$
960: as those in Fig.~\ref{fig:4a} for $D=10$ when $k$ is rather large,
961: and compare them with the analytical prediction (\ref{final}),
962: taking into regard the relations (\ref{rescaledpars}). In
963: Fig.~\ref{fig:4b1}(a), the parameters $\delta $ and $k$ are varied
964: with $q=2k$, and in Fig.~\ref{fig:4b1}(b) the parameters $\delta $
965: and $q$ are varied for $k=50$. The analytical predictions are
966: plotted as dashed lines. We find good agreement between the
967: numerical and analytical results, especially for large $k$.
968: Figure~\ref{fig:4b2} displays the profiles of these ESs. The
969: fundamental soliton in Fig.~\ref{fig:4b2}(a) has a steep peak at
970: $n=0$, as assumed in the approximate analysis of Section~3.
971:
972: \begin{figure}[tbp]
973: \begin{center}
974: \includegraphics[scale=0.7]{fig7a.eps}\quad \includegraphics[scale=1.23]{fig7b.eps}
975: \end{center}
976: \caption{Branches of discrete ESs for $\protect\gamma
977: _{1}=\protect\gamma_{2}=0.05$ and $D=10$; $N=7$ is chosen for
978: larger $\protect\delta$, and $N=5$ for smaller $\protect\delta $.
979: The analytical prediction given by Eqs.~(\protect\ref{final}) and
980: (\protect\ref{rescaledpars}) is plotted as a dashed line. (a)
981: Solutions in the $(\protect\delta,k)$-plane for $q=2k$. (b)
982: Solutions in the $(\protect\delta,q)$-plane for $k=50$. The ESs
983: exist in the region $100<q<100+10\protect\delta $, whose boundary
984: is shown as a dotted line.} \label{fig:4b1}
985: \end{figure}
986:
987: \begin{figure}[tbp]
988: \begin{center}
989: \includegraphics[scale=0.55]{fig8a.eps}\qquad \includegraphics[scale=0.55]{fig8b.eps}
990: \end{center}
991: \caption{Examples of the ESs on solution branches in
992: Fig.~\protect\ref{fig:4b1}(b): (a) $\protect\delta =0.34718$,
993: $q=102$ and $N=7$; (b) $\protect\delta =0.0029014$, $q=100.02$ and
994: $N=5$. } \label{fig:4b2}
995: \end{figure}
996:
997: It is well known that in the continuum case multi-pulse homoclinic
998: orbits exist (under a mild \emph{Birkhoff signature} condition
999: \cite{MiHoOR:92,KoChBuSa:02}) along families of curves in the
1000: parameter plane that accumulate on the curves of the fundamental
1001: ES solutions. Unlike the branches computed above, they do not
1002: feature a single-humped shaped in either component, but rather
1003: look like bound-states of two spatially separated fundamental
1004: solitons. We have found exactly the same solution families in the
1005: discrete model too. An example is presented in
1006: Fig.~\ref{bound-state}.
1007:
1008: \begin{figure}[tbp]
1009: \begin{center}
1010: \includegraphics[scale=0.55]{fig9.eps}
1011: \end{center}
1012: \caption{A typical example of a two-pulse bound state ES in the
1013: lattice system (\protect\ref{eqn:disc}) with
1014: $\protect\delta=1.1261$, $\protect\gamma _{1}=\protect\gamma
1015: _{2}=0.05$, $D=5$, $k=0.3$, $q=7$ and $N=51$.} \label{bound-state}
1016: \end{figure}
1017:
1018: \subsection{The case of\/ $\protect\delta <0$}
1019:
1020: Figure~\ref{fig:4c} shows ES branches in the case of
1021: self-defocusing $\chi ^{(3)}$ terms, with $\gamma _{1}=\gamma
1022: _{2}=-0.05$ and $\delta <0$. Figures~\ref{fig:4d} and \ref{fig:4e}
1023: display the profiles of these ESs for $D=100$, $D=10$ and $D=6$.
1024: Note that these curves and the solutions on them for $D=100 $ are
1025: identical, to the accuracy depicted, to the corresponding ones
1026: found in the continuum counterpart of the model in Ref.
1027: \cite[Figs.~2,3]{YaMaKaCh:01}.
1028: \begin{figure}[tbp]
1029: \begin{center}
1030: \includegraphics[scale=0.7]{fig10a.eps}\quad \includegraphics[scale=0.7]{fig10b.eps}
1031: \end{center}
1032: \caption{Branches of ESs in the discrete system
1033: (\protect\ref{eqn:lattice}) with $q=1$ and $\protect\gamma
1034: _{1}=\protect\gamma _{2}=-0.05$: (a) In the $(\protect\delta
1035: ,k)$-plane for $D=100$ and $10$; (b) in the $(D,k)$-plane for
1036: $\protect\delta =-0.5$ and $N=30$. In panel~(a), the solid and
1037: dashed curves represent the results for $D=100$ and $10$,
1038: respectively. When $D=100$ (resp. $D=10$), $N=85$ (resp. $N=30$)
1039: were used for the most part but $N=93$ or $N=109$ (resp. $N=35$ or
1040: $N=40$) were used for the two outer curves with large
1041: $|\protect\delta |$. According to condition
1042: (\protect\ref{area_scale}), ESs exist only above $k=0.5$ and below
1043: $k=0.5-\protect\delta D$, when $q=1 $. } \label{fig:4c}
1044: \end{figure}
1045: Notice the variety of multi-humped shapes of the solitons
1046: belonging to these families. Like the continuum model, only the
1047: inner-most of these branches represents a fundamental soliton.
1048: Continuation towards the anti-continuum limit, $D\rightarrow 0$,
1049: becomes numerically problematic for these solitons. It was found
1050: difficult to compute the solutions with repeatable accuracy while
1051: varying $N$ below $D\approx 6$. Clearly, these branches become
1052: increasingly spiky as $D$ is decreased (see Fig.~\ref{fig:4e}),
1053: and it may happen that branches of ES solutions actually terminate
1054: before they reach the minimum value of $D$ at which they remain
1055: embedded, which would be $D_{\min}=2k-1$, for the values of $q$
1056: and $\delta $ used in Fig.~\ref{fig:4c}(b).
1057:
1058: \begin{figure}[tbp]
1059: \begin{center}
1060: \includegraphics[scale=0.55]{fig11a.eps}\qquad \includegraphics[scale=0.55]{fig11b.eps}\\[2ex]
1061: \includegraphics[scale=0.55]{fig11c.eps}\qquad \includegraphics[scale=0.55]{fig11d.eps}\\[2ex]
1062: \includegraphics[scale=0.55]{fig11e.eps}\qquad \includegraphics[scale=0.55]{fig11f.eps}\\[2ex]
1063: \includegraphics[scale=0.55]{fig11g.eps}\qquad \includegraphics[scale=0.55]{fig11h.eps}
1064: \end{center}
1065: \caption{ESs on the branches in Fig.~\protect\ref{fig:4c}(a) for
1066: $\protect\delta =-1$: (a) $k=0.69564$; (b) $k=0.71312$; (c)
1067: $k=0.71383$; (d) $k=0.71387$; (e) $k=0.6895$; (f) $k=0.70811$; (g)
1068: $k=0.70898$; (h) $k=0.70902$. In panels (a)-(d), $D=100$ and
1069: $N=85$, and in panels~(e)-(h), $D=10$ and $N=30$. } \label{fig:4d}
1070: \end{figure}
1071:
1072: \begin{figure}[tbp]
1073: \begin{center}
1074: \includegraphics[scale=0.55]{fig12a.eps}\quad \includegraphics[scale=0.55]{fig12b.eps}
1075: \end{center}
1076: \caption{ESs at the end of the above two branches in
1077: Fig.~\protect\ref{fig:4c}(b) for $D=6$: (a) $k=0.78427$; (b)
1078: $k=0.78426$. } \label{fig:4e}
1079: \end{figure}
1080:
1081: Figure~\ref{fig:4f} shows the ES branches in a still more exotic
1082: case of opposite signs in front of the FF and SH $\chi ^{(3)}$ SPM
1083: terms, $\gamma _{1}=0.05$ and $\gamma _{2}=-0.05$. This case may,
1084: in principle, also be physically relevant -- not to ordinary
1085: optical materials, but rather to photonic crystals (see, e.g.,
1086: \cite{PhotCryst} and references therein). Figure~\ref{fig:4g}
1087: displays the profiles of these ESs for $D=100$ and $D=10$. Note
1088: similarity with the branches in Fig.~\ref{fig:4c} for small
1089: $\delta $. However, in this case it is found that the solution
1090: branches still exist for large $k$, rather than undergoing turning
1091: back with the increase of $k$. Also the internal oscillations on
1092: the non-fundamental branches become far less pronounced.
1093:
1094: \begin{figure}[tbp]
1095: \begin{center}
1096: \includegraphics[scale=0.7]{fig13a.eps}\quad \includegraphics[scale=0.7]{fig13b.eps}
1097: \end{center}
1098: \caption{Branches of ESs with $q=1$, $\protect\gamma _{1}=0.05$
1099: and $\protect\gamma _{2}=-0.05$: (a) In the $(\protect\delta
1100: ,k)$-plane for $D=100$ and $10$; (b) in the $(D,k)$-plane for
1101: $k=1$ and $N=24$. In panel~(a), the solid and dashed curves
1102: represent the results for $D=100$ and $10$, respectively. When
1103: $D=100$ (resp. $D=10$), $N=85$ (resp. $N=30$) was used. In
1104: panel~(b), the forth solid line from the bottom still exists
1105: although it is very short and almost invisible. The ESs exist only
1106: in the region $0.5<k<0.5-\protect\delta D$ when $q=1$, whose
1107: boundary is indicated by a dotted line. } \label{fig:4f}
1108: \end{figure}
1109:
1110: \begin{figure}[tbp]
1111: \begin{center}
1112: \includegraphics[scale=0.55]{fig14a.eps}\quad \includegraphics[scale=0.55]{fig14b.eps}\\[2ex]
1113: \includegraphics[scale=0.55]{fig14c.eps}\quad \includegraphics[scale=0.55]{fig14d.eps}\\[2ex]
1114: \includegraphics[scale=0.55]{fig14e.eps}\quad \includegraphics[scale=0.55]{fig14f.eps}\\[2ex]
1115: \includegraphics[scale=0.55]{fig14g.eps}\quad \includegraphics[scale=0.55]{fig14h.eps}
1116: \end{center}
1117: \caption{Examples of ESs on the branches in
1118: Fig.~\protect\ref{fig:4f}(a) for $k=1$: (a) $\protect\delta
1119: =-0.23601$; (b) $\protect\delta =-0.099245$; (c) $\protect\delta
1120: =-0.053237$; (d) $\protect\delta =-0.033776$; (e) $\protect\delta
1121: =-0.24838$; (f) $\protect\delta =-0.13304$; (g) $\protect\delta
1122: =-0.096003$; (h) $\protect\delta =-0.05109$. In panels (a)-(d),
1123: $D=100$ and $N=85$; in panels (e)-(h), $D=10$ and $N=30$. }
1124: \label{fig:4g}
1125: \end{figure}
1126:
1127: The approximate analysis of Section~3 is also valid for $\gamma
1128: _{1}>0$ and $\gamma _{2},\,\delta <0$ when $k$ is large.
1129: Figure~\ref{fig:4h} shows the fundamental ES branch (the left one
1130: in Fig.~\ref{fig:4f}(a)) with the same values of $\gamma _{1}$ and
1131: $\gamma _{2}$ as those in Fig.~\ref{fig:4f} for $D=10$ when $k$ is
1132: rather large. In Fig.~\ref{fig:4h}(a), the parameters $\delta $
1133: and $k$ are varied for $q=2k$, and in Fig.~\ref{fig:4h}(b) the
1134: parameters $\delta $ and $q$ are varied for $k=50$, as in
1135: Figs.~\ref{fig:4b1}(a) and (b). The predictions based on Eqs.
1136: (\ref{final}) with (\ref{rescaledpars}) are plotted as dashed
1137: lines. A profile of the ES is also displayed in
1138: Fig.~\ref{fig:4h}(a). We see that Eq.~(\ref{final}) again
1139: approximates well the numerical result for the fundamental
1140: solitons when $k$ is large. The ES in Fig.~\ref{fig:4h}(a)
1141: features a steep peak at $n=0$, as assumed in the analytical
1142: approximation.
1143:
1144: \begin{figure}[t]
1145: \begin{center}
1146: \includegraphics[scale=0.7]{fig15a.eps}\quad \includegraphics[scale=1.23]{fig15b.eps}
1147: \end{center}
1148: \caption{Branches of discrete ESs for $\protect\gamma _{1}=0.05$,
1149: $\protect\gamma _{2}=-0.05$, $D=10$ and $N=7$. The analytical
1150: prediction, given by Eqs.~(\protect\ref{final}) and
1151: (\protect\ref{rescaledpars}) is plotted by dashed lines. (a)
1152: Solutions in the $(\protect\delta ,k)$-plane for $q=2k$. (b)
1153: Solutions in the $(\protect\delta ,q)$-plane for $k=50$. The ESs
1154: exist in the region $100+10\protect\delta<q<100$, whose boundary
1155: is shown by the dotted line. An example of the ES, for
1156: $\protect\delta =-0.31873$ and $q=98$, is plotted in panel~(b). }
1157: \label{fig:4h}
1158: \end{figure}
1159:
1160: Finally, we briefly discuss the case in which $\gamma _{1}=\gamma
1161: _{2}=0$, i.e., the $\chi ^{(3)}$ terms are absent. As shown in
1162: Fig.~\ref{fig:4i}, we could follow solutions of the
1163: three-dimensional algebraic problem for $G_{\epsilon}$. However,
1164: these solutions do not represent ESs since the origin is not a
1165: saddle-center but a hyperbolic saddle. Although the region
1166: $q/2<k<q/2-\delta D$ in which the origin is a saddle-center
1167: becomes wider when $D$ is larger, for the solution $k$ diverges to
1168: $\infty$ as $D\rightarrow \infty$. Thus, there seems to be no hope
1169: that an ES exists in this case. As mentioned above, this situation
1170: is very similar to that in the continuum model (\ref{eqn:PDE}).
1171:
1172: \begin{figure}[t]
1173: \begin{center}
1174: \includegraphics[scale=1.23]{fig16a.eps}\quad \includegraphics[scale=0.7]{fig16b.eps}
1175: \end{center}
1176: \caption{Branches of solutions of the three-dimensional algebraic problem
1177: for $G_{\protect\epsilon }$ in the absence of the $\protect\chi ^{(3)}$
1178: nonlinearity, with $\protect\gamma _{1}=\protect\gamma _{2}=0$, and $q=1$.
1179: (a) In the $(\protect\delta ,k)$-plane for $D=10$; (b) in the $(D,k)$-plane
1180: for $\protect\delta =-0.1$ and $N=15$. In panel~(a), $N=15$ and $25$ were
1181: used for $k>1$ and $k<1$, respectively. The ESs would exist only in the
1182: region $0.5<k<0.5-\protect\delta D$, the boundaries of which are plotted as
1183: dotted lines, when $q=1$. An example of a \emph{non-embedded} (gap) soliton,
1184: found in this case for $\protect\delta =-0.1$ and $k=1.9027$, is plotted in
1185: panel~(a).}
1186: \label{fig:4i}
1187: \end{figure}
1188:
1189: \section{Conclusions}
1190:
1191: In this paper, we have established the existence of embedded
1192: solitons (ESs) in the discrete model of the second-harmonic
1193: generation in the presence of cubic nonlinearity. The model can be
1194: naturally realized as an array of channel waveguides in the
1195: spatial domain, therefore our results suggest possibilities for
1196: new experiments with discrete spatial solitons in nonlinear
1197: optics. We have also introduced a simplified model, with the $\chi
1198: ^{(3)}$ nonlinearity present solely at the central site, in which
1199: the existence of an ES was demonstrated in an asymptotic
1200: analytical form. In the general case, the existence region for the
1201: ESs in the discrete model was found numerically. Moreover, we have
1202: checked that the asymptotic analysis of the simplified model in
1203: the limit of large wavenumbers gives a good approximation of the
1204: codimension-one set in parameter space, on which the ESs exist.
1205:
1206: More generally, we have established, that unlike the
1207: discretizations of other (dissipative) continuum models bearing
1208: localized solutions, the codimension of ESs does not change when
1209: one passes to a discrete version. Whereas continuum ESs may be
1210: regarded as homoclinic orbits to saddle-center equilibria in
1211: reversible flows (ordinary differential equations), discrete ESs
1212: should be thought of as homoclinic orbits to saddle-center fixed
1213: points of reversible maps. Both have codimension one in the
1214: parameter space. Understanding this property led us to the
1215: derivation of a numerical method for computing homoclinic orbits
1216: to nonhyperbolic equilibria of reversible maps.
1217:
1218: Accurate investigation of stability of ESs in the lattice model is
1219: beyond the scope of the present investigation. Systematic results
1220: for the stability will be presented elsewhere; however, some
1221: preliminary results suggest that the fundamental discrete ESs
1222: inherit the \textit{semi-stability} property found in the
1223: corresponding continuous model \cite{YaMaKaCh:01,PeYa:02}. We
1224: expect that general arguments in favor of the semi-stable
1225: character of these solitons for the continuous case should apply
1226: in the discrete setting as well.
1227:
1228: Figure~\ref{fig:7a} shows a preliminary computational result for
1229: $D=10$, $\delta =-1$, $k=0.6895$, $q=1$ and $\gamma _{1}=\gamma
1230: _{2}=-0.05$, corresponding to the fundamental ES of
1231: Fig.~\ref{fig:4d}(e). Here Eq.~(\ref{eqn:lattice}) was integrated
1232: numerically by the fourth-order Runge-Kutta method under the
1233: boundary condition
1234: \begin{equation}
1235: u_{-\bar{N}-1}(z)=u_{\bar{N}+1}(z)=v_{-\bar{N}-1}(z)=v_{\bar{N}+1}(z)=0
1236: \end{equation}for all $z\geq 0$ with $\bar{N}=93$ and the initial condition
1237: \begin{equation}
1238: u_{n}(0)=(1+c_{1})U_{n},\quad v_{n}(0)=(1+c_{2})V_{n} \label{eqn:init}
1239: \end{equation}where $(U_{n},V_{n})$ represents an approximate ES given by the data of
1240: Fig.~\ref{fig:4d}(e) for $|n|\leq N=30$ and by $U_{n}=V_{n}=0$ for $|n|>N$, and $c_{1,2}$
1241: are small amplitudes of the initial disturbance, which were chosen
1242: to be $c_{1}=c_{2}=0.01$. The positive values of $c_{1,2}$ mean that the
1243: norm
1244: \begin{equation}
1245: E=\sum_{n=-\bar{N}}^{\bar{N}}(|u_{n}|^{2}+2|v_{n}|^{2})
1246: \end{equation}of the perturbed state (in the temporal-domain optical model, it plays the
1247: role of energy) is greater than that of the unperturbed ES. In
1248: Fig.~\ref{fig:7a} we see the characteristic hallmark of
1249: semi-stability for the ES: The shapes of the perturbed wave are
1250: almost unchanged for a long period of $t $ in
1251: Figs.~\ref{fig:7a}(a) and(b), and the amplitudes $|u_{0}(t)|$ and
1252: $|v_{0}(t)|$ exhibit small oscillations near the unperturbed
1253: values in Figs.~\ref{fig:7a}(c) and(d). However, the perturbed ES
1254: was destroyed in the simulations when different signs of $c_{1}$
1255: and $c_{2}$ were chosen. Further investigation of stability and
1256: bifurcation will be the subject of future work.
1257:
1258: \begin{figure}[t]
1259: \begin{center}
1260: \includegraphics[scale=0.8]{fig17a.eps}\quad \includegraphics[scale=0.8]{fig17b.eps}\\[3ex]
1261: \includegraphics[scale=0.6]{fig17c.eps}\quad \includegraphics[scale=0.6]{fig17d.eps}\\[0pt]
1262: \end{center}
1263: \caption{Evolution of the fundamental ES (corresponding to
1264: Fig.~\protect\ref{fig:4d}(e)) initiated by the nondestructive
1265: perturbation with $c_{1}=c_{2}=0.01$ in the initial conditions
1266: (\protect\ref{eqn:init}). The dashed lines in panels~(c) and (d)
1267: are the amplitudes of the $u_{0}$- and $v_{0}$-components of the
1268: unperturbed soliton.} \label{fig:7a}
1269: \end{figure}
1270:
1271: Finally, our results so far pertain only to those discrete ESs
1272: that are symmetric with respect to the involution $R$ defined in
1273: Eq.~(\ref{R1}); solitons with this symmetry have an on-site
1274: central peak. It is also possible to apply techniques developed in
1275: this work to solutions symmetric with respect to involution
1276: $\hat{R}$, see Eq.~(\ref{R2}), that should feature an inter-site
1277: central peak. In other physical context, such waves are less
1278: likely to be stable than waves that are centered on a lattice site
1279: \cite{Review}. However, that understanding typically applies to
1280: regular (gap) discrete solitons and need not necessarily apply to
1281: ESs. We shall address this issue in future work.
1282:
1283: \section*{Acknowledgements}
1284:
1285: K.Y. acknowledges support from the Japanese Society for
1286: the Promotion of Science, which enabled him to stay in Bristol and
1287: to perform this work. B.A.M. appreciates support of EPSRC
1288: ``critical mass" grant and mathematics programme and hospitality
1289: of the Department of Engineering Mathematics at the University of
1290: Bristol.
1291:
1292: \section*{References}
1293:
1294: %\bibliography{embed}
1295:
1296: \begin{thebibliography}{99}
1297: \bibitem{YaMaKa:99} Yang J, Malomed B A and Kaup D J 1999 Embedded solitons
1298: in second-harmonic-generating systems \textit{Phys. Rev. Lett.} \textbf{83}
1299: 1958--1961
1300:
1301: \bibitem{ChMaYaKa:01} Champneys A R, Malomed B A, Yang J and Kaup D J 2001
1302: Embedded solitons: Solitary waves in resonance with the linear spectrum
1303: \textit{Physica D} \textbf{152-153} 340--354
1304:
1305: \bibitem{YaMaKaCh:01} Yang J, Malomed B A, Kaup D J and Champneys A R 2001
1306: Embedded solitons: A new type of solitary wave \textit{Math. Comp. Sim}
1307: \textbf{56} 585--600
1308:
1309: \bibitem{ChGr:97} Champneys A R and Groves M D 1997 A global investigation
1310: of solitary waves solutions to a two-parameter model for water
1311: waves \textit{J. Fluid. Mech.} \textbf{342} 299--229
1312:
1313: \bibitem{FuEs:97} Fujioka J and Espinosa A 1997 Soliton-like solution of an
1314: extended NLS equation existing in resonance with linear dispersive waves
1315: \textit{J. Phys. Soc. Jpn.} \textbf{66} 2601--2607
1316:
1317: \bibitem{ChMaFr:98} Champneys A R, Malomed B A and M. J.~Friedman 1998
1318: Thirring solitons in the presence of dispersion \textit{Phys. Rev. Lett.}
1319: \textbf{80} 4168--4171
1320:
1321: \bibitem{AtMa:01} Atai J and Malomed B A 2001 Solitary waves in systems with
1322: separated Bragg grating and nonlinearity \textit{Phys. Rev. E} \textbf{64}
1323: 066617
1324:
1325: \bibitem{KoChBuSa:02} Kolossovski K, Champneys A R, Buryak A and R. A.
1326: Sammut 2002 Multi-pulse embedded solitons as bound states of quasi-solitons
1327: \textit{Physica D} \textbf{171} 153--177
1328:
1329: \bibitem{EsFuGo:03} Espinosa-Ceron A, Fujioka J and Gomez-Rodriguez A 2003
1330: Embedded solitons: Four-frequency radiation, front propagation and radiation
1331: inhibition \textit{Physica Scripta} \textbf{67} 314--324
1332:
1333: \bibitem{Ya:03} Yang J K 2003 Stable embedded solitons \textit{Phys. Rev.
1334: Lett.} \textbf{91} 143903
1335:
1336: \bibitem{Mak} Mak W C K, Malomed B A and Chu P L 2004 Symmetric and
1337: asymmetric solitons in linearly coupled Bragg gratings \textit{Phys. Rev.} E
1338: 69 066610
1339:
1340: \bibitem{Merhasin} Merhasin I M and Malomed B A 2004 Four-wave solitons in
1341: Bragg cross-gratings \textit{J. Optics B: Quant. Semiclass. Opt.} 6 S323-S332
1342:
1343: \bibitem{PeYa:02} Pelinovsky D E and Yang J 2002 A normal form for nonlinear
1344: resonance of embedded solitons, \textit{Proc. Roy. Soc. Lond. A} \textbf{458}
1345: 1469--1497
1346:
1347: \bibitem{ChKi:99} Champneys A R and Kivshar Yu S 2000 Origin of multikinks
1348: in dispersive nonlinear systems \textit{Phys. Rev. E} \textbf{61} 2551--2554
1349:
1350: \bibitem{Flach} Flach S, Zolotaryuk Y and Kladko K 1999 Moving lattice kinks
1351: and pulses: An inverse method \textit{Phys. Rev E} \textbf{59} 6105-6115
1352:
1353: \bibitem{Morgante} Morgante A M, Johansson M, Kopidakis G and Aubry S 2002
1354: Standing wave instabilities in a chain of nonlinear coupled oscillators
1355: \textit{Physica D} \textbf{162 }53-94
1356:
1357: \bibitem{PeKr:84} Peyrard M and Kruskal M D 1984 Kink dynamics in the highly
1358: discrete {sine-Gordon} system \textit{Physica D} \textbf{14} 88--102
1359:
1360: \bibitem{KiBr:98} Braun O M and Kivshar Yu S 1998 Nonlinear dynamics of the
1361: {Frenkel-Kontorova} model \textit{Phys. Rep.} \textbf{306} 1--108
1362:
1363: \bibitem{SaZoEi:00} Savin A V. Zolotaryuk Y and Eilbeck J C 2000 Moving
1364: kinks and nanopterons in the nonlinear {Klein-Gordon} lattice
1365: \textit{Physica D} \textbf{138} 267--281
1366:
1367: \bibitem{AiChRo:03} Aigner A A, Champneys A R and Rothos V R 2003 A new
1368: barrier to the existence of moving kinks in Frenkel-Kontorova lattices
1369: \textit{Physica D} \textbf{186} 148--170
1370:
1371: \bibitem{ESdiscrete} Gonzalez-Perez-Sandi S, Fujioka J and Malomed B A 2004
1372: Embedded solitons in dynamical lattices \textit{Physica D} \textbf{197}
1373: 86--100
1374:
1375: %%\bibitem{Scott} Scott, A 2003 \textit{Nonlinear Science: Emergence and
1376: %%Dynamics of Coherent Structures} 2nd ed (Oxford: Oxford University Press)
1377:
1378: \bibitem{FiSc} Fiedler B and Scheurle J 1996 \textit{Discretization of
1379: Homoclinic Orbits, Rapid Forcing and ``Invisible'' Chaos} Memoirs of Amer.
1380: Math. Soc. 119 (Providence RI: American Mathematical Society)
1381:
1382: \bibitem{Bang1} Bang O, Clausen C B, Christiansen P L and Torner L 1999
1383: Engineering competing nonlinearities \textit{Opt. Lett.} \textbf{24}
1384: 1413--1415
1385:
1386: \bibitem{Bang2} Corney J F and Bang O 2001 Solitons in quadratic nonlinear
1387: photonic crystals \textit{Phys. Rev. E} \textbf{64} 047601
1388:
1389: \bibitem{Bang3} Johansen S K, Carrasco S, Torner L and Bang O 2002
1390: Engineering of spatial solitons in two-period QPM structures \textit{Opt.
1391: Commun.} \textbf{203} 393--402
1392:
1393: \bibitem{Review} Etrich C, Lederer F, Malomed B A, Peschel T, and Peschel U
1394: 2000 Optical solitons in media with a quadratic nonlinearity \textit{Progr.
1395: Opt.} \textbf{41} 483-568
1396:
1397: \bibitem{DiscrChi2soliton} Iwanow R, Schiek R, Stegeman G I, Pertsch T,
1398: Lederer F, Min Y and Sohler W 2004 Observation of discrete quadratic
1399: solitons \textit{Phys. Rev. Lett.} \textbf{93} 113902
1400:
1401: \bibitem{PanosZiad} Kevrekidis P G Malomed B A, and Musslimani Z 2003
1402: Discrete gap solitons in a diffraction-managed waveguide array \textit{Eur.
1403: Phys. J. D} \textbf{67} 013605
1404:
1405: \bibitem{DiffrManagement1} Eisenberg H S, Silberberg Y, Morandotti R, Boyd A
1406: and Aitchison J S 2000 Diffraction management \textit{Phys. Rev. Lett.}
1407: \textbf{85} 1863--1866
1408:
1409: \bibitem{DiffrManagement2} Pertsch T, Zentgraf T, Peschel U and Lederer F
1410: 2002 Anomalous refraction and diffraction in discrete optical systems
1411: \textit{Phys. Rev. Lett.} \textbf{88} 093901
1412:
1413: \bibitem{De:76b} Devaney R L 1976 Reversible diffeomorphisms and flows
1414: \textit{Trans. Amer. Math. Soc.} \textbf{218} 89--113
1415:
1416: \bibitem{LaRo:98} Lamb J S W and Roberts J A G 1998 Time-reversal symmetry
1417: in dynamical systems: A survey \textit{Physica D} \textbf{112} 1--39
1418:
1419: \bibitem{GH83} Guckenheimer J and Holmes P 1983 \textit{Nonlinear
1420: Oscillations, Dynamical Systems, and Bifurcations of Vector Fields} (New
1421: York: Springer)
1422:
1423: \bibitem{Auto} Doedel E, Champneys A R, Fairgrieve T F, Kuznetsov Y A,
1424: Sandstede B and Wang X 1997 \textit{AUTO97: Continuation and
1425: Bifurcation Software for Ordinary Differential Equations (with
1426: HomCont)} Available by anonymous ftp from
1427: \texttt{ftp.cs.concordia.ca}, directory \texttt{pub/doedel/auto}
1428:
1429: \bibitem{NY97} Nusse E H and Yorke J A 1997 \textit{Dynamics: Numerical
1430: Explorations} 2nd ed. (New York: Springer)
1431:
1432: \bibitem{K81} Kawakami H 1981 Algorithme num\'{e}rique d\'{e}finissant la
1433: bifurcation d'un point homocline \textit{C. R. Acad. Sc. Paris, S\'{e}rie I}
1434: \textbf{293} 401--403
1435:
1436: \bibitem{BK97a} Beyn W-J and Kleinkauf J N 1997 The numerical computation of
1437: homoclinic orbits for maps \textit{SIAM J. Num. Anal.} \textbf{34} 1207--1236
1438:
1439: \bibitem{BK97b} Beyn W-J and Kleinkauf J N 1997 Numerical approximation of
1440: homoclinic chaos \textit{Numer. Algorithms} \textbf{14} 25--53
1441:
1442: \bibitem{Y98a} Yagasaki K 1998 Numerical detection and continuation of
1443: homoclinic points and their bifurcations for maps and periodically forced
1444: systems \textit{Int. J. Bifurcation Chaos} \textbf{7} 1617--1627
1445:
1446: \bibitem{BBV00} Bergamin J M, Bountis T and Vrahatis M N 2002 Homoclinic
1447: orbits of invertible maps \textit{Nonlinearity} \textbf{15} 1603--1619
1448:
1449: \bibitem{Y98b} Yagasaki K 1998 \textit{HomMap: An Auto driver for homoclinic
1450: bifurcation analysis of maps and periodically forced systems} (Gifu Japan:
1451: Gifu University)
1452:
1453: \bibitem{Lo:01} Lombardi E 2000 \textit{Oscillatory Integrals and Phenomena
1454: beyond All Algebraic Orders: With Applications to Homoclinic Orbits in
1455: Reversible Systems} (Berlin: Springer)
1456:
1457: \bibitem{Ch:01} Champneys A R 2001 Codimension-one persistence beyond all
1458: orders of homoclinic orbits to singular saddle centers in
1459: reversible systems \textit{Nonlinearity} \textbf{14} 87--112
1460:
1461: \bibitem{MiHoOR:92} Mielke A, Holmes P and O'Reilly O 1992 Cascades of
1462: homoclinic orbits to, and chaos near, a Hamiltonian saddle-center \textit{J.
1463: Dyn. Diff. Eqns.} \textbf{4} 95--126
1464:
1465: \bibitem{PhotCryst} Andreani L C, Agio M, Bajoni D, Belotti M, Galli M,
1466: Guizzetti G, Malvezzi A M, Marabelli F, Patrini M and Vecchi G 2003
1467: Morphology and optical properties of bare and polydiacetylenes-infiltrated
1468: opals \textit{Synthetic Metals} \textbf{139} 695--700
1469: \end{thebibliography}
1470:
1471: \end{document}
1472: