1:
2: \documentclass[twocolumn,showpacs,preprintnumbers,amsmath,amssymb]{revtex4}
3:
4: \usepackage{graphicx}% Include figure files
5: \usepackage{dcolumn}% Align table columns on decimal point
6: \usepackage{bm}% bold math
7: \renewcommand{\theequation}{\arabic{section}.\arabic{equation}}
8:
9:
10: \begin{document}
11:
12: \title{Universal fractal structures in the weak interaction of solitary waves in generalized nonlinear Schr\"{o}dinger equations}% Force line breaks with \\
13:
14: \author{Yi Zhu}
15: %\altaffiliation[Also at ]{Zhou Pei-Yuan Center for Applied Mathematics, Tsinghua University.}%Lines break automatically or can be forced with \\
16: \email{zhuyi03@mails.tsinghua.edu.cn}
17: \affiliation{%
18: Zhou Pei-Yuan Center for Applied Mathematics, Tsinghua University,
19: Beijing 100084, P.R.China\\
20: }%
21:
22: \author{Jianke Yang}
23: %\homepage{http://www.Second.institution.edu/~Charlie.Author}
24: \email{jyang@cems.uvm.edu} \affiliation{
25: Department of Mathematics and Statistics, University of Vermont, 16 Colchester Avenue, Burlington, VT 05401, USA\\% with \\
26: }%
27:
28: %\date{\today}% It is always \today, today,
29: % but any date may be explicitly specified
30: %\date{ }
31:
32: \begin{abstract}
33: Weak interactions of solitary waves in the generalized nonlinear
34: Schr\"{o}dinger equations are studied. It is first shown that these
35: interactions exhibit similar fractal dependence on initial
36: conditions for different nonlinearities. Then by using the
37: Karpman-Solov'ev method, a universal system of dynamical equations
38: is derived for the velocities, amplitudes, positions and phases of
39: interacting solitary waves. These dynamical equations contain a
40: single parameter, which accounts for the different forms of
41: nonlinearity. When this parameter is zero, these dynamical equations
42: are integrable, and the exact analytical solutions are derived. When
43: this parameter is non-zero, the dynamical equations exhibit fractal
44: structures which match those in the original wave equations both
45: qualitatively and quantitatively. Thus the universal nature of
46: fractal structures in the weak interaction of solitary waves is
47: analytically established. The origin of these fractal structures is
48: also explored. It is shown that these structures bifurcate from the
49: initial conditions where the solutions of the integrable dynamical
50: equations develop finite-time singularities. Based on this
51: observation, an analytical criterion for the existence and locations
52: of fractal structures is obtained. Lastly, these analytical results
53: are applied to the generalized nonlinear Schr\"{o}dinger equations
54: with various nonlinearities such as the saturable nonlinearity, and
55: predictions on their weak interactions of solitary waves are made.
56: \end{abstract}
57:
58: \pacs{42.65.Tg, 05.45.Yv, 42.81.Dp}
59: %\pacs{Valid PACS appear here}% PACS, the Physics and Astronomy
60: % Classification Scheme.
61: %\keywords{Suggested keywords}%Use showkeys class option if keyword
62: %display desired
63: \maketitle
64:
65:
66: \section{\label{sec:level1}Introduction}
67: \quad
68:
69: Solitary wave interactions are a fascinating and important
70: phenomenon for both physical and mathematical reasons. Physically,
71: such interactions have arisen in a wide array of disciplines such as
72: water waves \cite{Ablowitz_Segur}, optics
73: \cite{Hasegawa_Kodama,Krolikowski,Anastassiou,Chen_Kivshar,
74: Kivshar_Agrawal, Aceves_stud}, and Josephson junctions \cite{Yukon}.
75: For instance, in soliton-based fiber communication systems, optical
76: pulses traveling in different frequency channels pass through each
77: other, giving rise to collisions (strong interactions) of solitary
78: waves. In the same frequency channel, neighboring optical pulses
79: interfere with each other through overlapping tails, giving rise to
80: weak interactions of solitary waves. Motivated by these physical
81: applications, solitary wave interactions has been studied
82: extensively in both the mathematical and physical communities. If
83: the system is integrable, collisions of solitons are elastic
84: \cite{Ablowitz_Segur}, and weak interactions of solitons exhibit
85: interesting yet simple behaviors
86: \cite{Hasegawa_Kodama,Karpman_Solovev,Gerdjikov_PRL,Gerdjikov_Yang,Yang_Manakov}.
87: However, in non-integrable systems, solitary wave interactions can
88: be far more complex. The first sign of this complexity was reported
89: by Ablowitz, et al. \cite{Ablowitz_Kruskal} for kink and anti-kink
90: collisions in the $\phi^4$ model where, inside the trapping
91: interval, a reflection window was found. Later extensive numerical
92: studies on this model by Campbell, et al. \cite{Campbell1,
93: Campbell2, Campbell3, Campbell4} revealed that in fact, sequences of
94: two- and more-bounce reflection windows exist, and the physical
95: mechanism for these refection windows is a resonant energy transfer
96: between the translational motion and internal modes of
97: kinks/antikinks. Anninos, et al. \cite{anninos} pointed out further
98: that there is a fractal structure in kink-antikink collisions. Using
99: a collective-coordinate (i.e., variational) approach, they derived a
100: set of fourth-order ordinary differential equations (ODEs) for these
101: collisions, and these ODEs exhibit qualitatively similar fractal
102: structures as in the $\phi^4$ model (a comprehensive review on
103: kink-antikink collisions in $\phi^4$-type equations can be found in
104: \cite{Kudryavtsev}). These complex dynamics turn out to be not
105: restricted to kink-antikink collisions. Indeed, similar phenomena
106: have been reported on kink-defect collisions in the sine-Gordon and
107: $\phi^4$ models \cite{Kivshar1, Kivshar2, Kivshar3}, as well as
108: vector-soliton collisions in the coupled nonlinear Schr\"{o}dinger
109: (NLS) equations \cite{YangTan,YangTan2,TanYang}. Furthermore,
110: fractal scattering has also been reported on weak interactions of
111: breathers in a weakly discrete sine-Gordon equation
112: \cite{Dmitriev_Kivshar} and weak interactions of solitary waves in a
113: weakly discrete NLS equation \cite{Dmitriev}. Recently, Goodman and
114: Haberman \cite{Goodman_Haberman1, Goodman_Haberman2,
115: Goodman_Haberman3} provided a deep analysis on the
116: collective-coordinate models (ODEs) for kink-antikink collisions in
117: the $\phi^4$ model \cite{anninos}, kink-defect collisions in the
118: sine-Gordon model \cite{Kivshar1}, and vector-soliton collisions in
119: the coupled NLS equations \cite{Ueda_Kath, TanYang}, using
120: sophisticated dynamical system techniques. They derived analytical
121: formulas for the locations of reflection-window sequences, which
122: agree qualitatively with numerical results on the original partial
123: differential equations (PDEs). Their results shed much light on the
124: origins of these window sequences and fractal structures, especially
125: from a mathematical point of view.
126:
127: Despite the above progress on solitary wave interactions, our
128: understanding on these phenomena is far from satisfactory. On the
129: collision of solitary waves, the analysis done so far were all based
130: on approximate collective-coordinate approaches, hence the reduced
131: ODE models can only provide qualitative results at best. Many
132: features reported in the ODE models can not be seen in the PDE
133: simulations, thus it is not possible to make a reliable prediction
134: on the collision dynamics based on those ODE models and their
135: analysis. In addition, the ODE models obtained from the
136: collective-coordinate approaches not only are complicated, but also
137: differ significantly from one PDE system to another. This forced
138: previous researchers to analyze each PDE and its reduced ODE systems
139: on an individual basis, which prevents an overall understanding on
140: collision processes of solitary waves. On the weak interaction of
141: solitary waves, the situation is even less satisfactory. The fractal
142: nature of this weak interaction was reported only for systems which
143: are weakly perturbed integrable systems (sine-Gordon and NLS
144: equations, to be more specific) \cite{Dmitriev_Kivshar, Dmitriev}.
145: It is not known yet whether similar phenomena arise in strongly
146: non-integrable equations. More seriously, the previous work on this
147: subject is largely numerical. No analysis has been attempted yet
148: (not even the approximate collective-coordinate studies). Thus an
149: analytical understanding on weak interactions of solitary waves is a
150: completely open question.
151:
152: In this paper, we study weak interactions of solitary waves in a
153: whole class of generalized NLS equations (with arbitrary
154: nonlinearities) both analytically and numerically. These generalized
155: NLS equations are not weak perturbations of the NLS equation in
156: general. First we show by direct PDE simulations that these weak
157: interactions for different nonlinearities exhibit similar fractal
158: structures on initial parameters of solitary waves. This establishes
159: that fractal scattering is a common feature of weak interactions in
160: this class of generalized NLS equations. Next, we rigorously derive
161: a universal system of dynamical equations (ODEs) for the velocities,
162: amplitudes, positions and phases of interacting solitary waves in
163: this class of PDEs by the Karpman-Solov'ev method. This universal
164: ODE system is remarkably simple, and it contains only a single
165: parameter which depends on the individual PDEs (after variable
166: rescalings). When this parameter is zero, these dynamical equations
167: are integrable, and their exact analytical solutions are derived.
168: When this parameter is non-zero, the dynamical equations are found
169: to exhibit fractal structures for a wide range of initial
170: conditions. These fractal structures match those in the original
171: PDEs both qualitatively and quantitatively, thus the universal
172: nature of fractal scattering in the weak interaction of solitary
173: waves is analytically established. We further explore the origin of
174: these fractal structures. Our numerical studies on the ODE system
175: show that these fractal structures bifurcate from the initial
176: conditions where the solutions of the integrable dynamical equations
177: develop finite-time singularities. Based on this observation, we
178: present an analytical criterion for the existence and locations of
179: fractal structures. One corollary from this criterion is that when
180: the initial separation velocity is above a certain threshold value,
181: fractal structures should disappear --- a prediction which agrees
182: with our PDE numerics as well as previous numerics on the weakly
183: discrete NLS equation (see Fig. 6 in Ref. \cite{Dmitriev}). Lastly,
184: we apply these analytical results to the generalized NLS equations
185: with various nonlinearities such as the cubic-quintic, exponential
186: and saturable nonlinearities, and make detailed predictions on the
187: dynamics of their weak interactions.
188:
189: This paper is structured as follows. In Sec. 2, we describe
190: individual solitary waves in the generalized NLS equations. In Sec.
191: 3, we present direct PDE simulation results on weak interactions of
192: solitary waves in the generalized NLS equations with two different
193: nonlinearities, and reveal the common (universal) fractal structures
194: in this class of PDEs. In Sec. 4, we analytically derive a universal
195: system of dynamical equations (ODEs) for parameters of interacting
196: solitary waves using asymptotic methods, and show that these ODEs
197: accurately describe the weak interactions in the PDEs. In Sec. 5, we
198: solve this ODE system analytically when the single parameter in this
199: system is equal to zero (which is the integrable case). In addition,
200: we derive explicit conditions for the solutions of the integrable
201: ODE system to develop finite-time singularities. In Sec. 6, we show
202: that fractal structures appear in this ODE system when its parameter
203: is non-zero, and explore the origin of these fractal structures. In
204: Sec. 7, we apply the analytical results to the generalized NLS
205: equations with various nonlinearities. In Sec. 8, we summarize the
206: results of the paper and make some further remarks.
207:
208:
209: \section{Preliminaries}
210: The generalized NLS equation is
211: \begin{equation}
212: iU_t+U_{xx}+F(|U|^2)U=0, \label{eqn:1}
213: \end{equation}
214: where $F(\cdot)$ is a real-valued algebraic function with $F(0)=0$.
215: Equation (\ref{eqn:1}) supports solitary waves of the form
216: \begin{equation}
217: U=\Phi(x-Vt-x_0;\beta)e^{\frac{1}{2}iV(x-x_0)-\frac{1}{4}iV^2t+i\beta
218: t-i\sigma_0}, \label{eqn:solution}
219: \end{equation}
220: where $\Phi(\theta)$ is a positive function which satisfies the
221: following equation
222: \begin{eqnarray}
223: \Phi_{\theta\theta}+F(|\Phi|^2)\Phi-\beta\Phi=0, \nonumber\\
224: \Phi\rightarrow 0, |\theta|\rightarrow \infty, \label{eqn:2}
225: \end{eqnarray}
226: and $\beta \: (>0), V, x_0, \sigma_0$ are real constants. For
227: convenience, we introduce the notations
228: \begin{eqnarray}
229: \xi=Vt+x_0, \quad \theta=x-\xi,\nonumber\\
230: \sigma=(\beta+\frac{1}{4}V^2)t-\sigma_0, \quad
231: \phi=\frac{1}{2}V\theta+\sigma. \label{def_phi}
232: \end{eqnarray}
233: Physically, $\beta$ is the propagation constant which is related to
234: the solitary-wave amplitude (henceforth, we call $\beta$ an
235: amplitude parameter), $\phi$ is the phase of the solitary wave,
236: $\sigma_0$ is its initial phase, $\xi$ is its center position, $V$
237: is its velocity, and $x_0$ is its initial position. The solitary
238: wave is characterized uniquely by its four parameters:
239: $V,\beta,\sigma_0$ and $x_0$. The asymptotic behavior of this
240: solution at infinity is
241: \begin{equation}
242: \Phi(\theta)\rightarrow c\: e^{-\sqrt{\beta}|\theta|}, \quad
243: |\theta|\rightarrow\infty, \label{eqn:asy}
244: \end{equation}
245: where $c$ is the tail coefficient which is determined by the
246: nonlinear function $F$ and propagation constant $\beta$. We define
247: the power of the solitary wave as
248: \begin{equation}
249: P(\beta)=\int^{\infty}_{-\infty}\Phi^2(\theta;\beta)d\theta,
250: \end{equation}
251: which plays an important role in the linear stability of the
252: solitary wave. For general functions $F$, the analytical formulas
253: for $\Phi, \; P$ and $c$ are not available. But for some special
254: nonlinearities, one can find the analytical solutions. For instance,
255: for the cubic-quintic nonlinearity
256: \begin{equation}
257: F(|U|^2)=\alpha|U|^2+\gamma|U|^4, \label{cqNL}
258: \end{equation}
259: the analytical formulas for $\Phi, \; P$ and $c$ are
260: \cite{Kivshar_Agrawal,Kovalev}
261: \begin{eqnarray}
262: && \Phi(\theta;\beta)=\sqrt{\frac{4B\beta/\alpha}{B+\cosh{2\sqrt{\beta}\theta}}},\label{cqNLs}\\
263: &&P=\frac{4B\sqrt{\beta}(\pi/2-\arctan{\frac{B}{\sqrt{1-B^2}}})}{\alpha\sqrt{1-B^2}},
264: \label{Pformula}
265: \\&&c=\sqrt{8B\beta/\alpha}, \label{cformula}
266: \end{eqnarray}
267: where
268: \begin{eqnarray}
269: B=\textrm{sgn}(\alpha)(1+\frac{16\beta\gamma}{3\alpha^2})^{-1/2}.
270: \end{eqnarray}
271: For special values of $\alpha=1,~\gamma=0$, Eq.(\ref{eqn:1}) becomes
272: the integrable NLS equation, and then
273: \begin{eqnarray}
274: &&B=1,\quad \Phi(\theta;\beta)=\sqrt{2\beta}\textrm{sech}(\sqrt{\beta}\theta),\nonumber\\
275: &&P=4\sqrt{\beta}, \quad c=\sqrt{8\beta}.
276: \end{eqnarray}
277:
278: \section{\label{sec:PDE}Universal fractal structures in weak
279: interactions of solitary-waves \setcounter{equation}{0}}
280:
281: When two solitary waves are placed adjacent to each other, they
282: would interfere through tail overlapping. In this case, the initial
283: condition is
284: \begin{eqnarray}
285: U(x,0)=U_{1}(x,0)+U_{2}(x,0), \nonumber\\
286: U_{k}(x, 0)=\Phi(x-x_{0,k};\beta_{0,k})e^{i\phi_{0,k}},
287: \nonumber \\
288: \phi_{0,k}=\frac{1}{2}V_{0,k}(x-x_{0,k})-\sigma_{0,k},
289: \label{pdeinitial}
290: \end{eqnarray}
291: where $\Phi$ satisfies Eq.(\ref{eqn:2}). Here "0" in the subscript
292: represents the initial value of the underlying parameter. For
293: convenience, we assign the left solitary wave with index $k=1$, and
294: the right solitary wave with index $k=2$. To study the weak
295: interaction between these two solitary waves, we require that the
296: two solitary waves are both stable, well separated, and having
297: almost the same velocities and amplitudes. Introducing notations
298: \begin{eqnarray}
299: \beta=\frac{1}{2}(\beta_1+\beta_2),\;V=\frac{1}{2}(V_1+V_2),\;\xi=\frac{1}{2}(\xi_1+\xi_2),
300: \end{eqnarray}
301: and
302: \begin{eqnarray}
303: \Delta\beta=\beta_2-\beta_1,\;\;\Delta V=V_2-V_1,\;\;\Delta
304: \xi=\xi_2-\xi_1,
305: \end{eqnarray}
306: the above requirements then amount to
307: \begin{eqnarray}
308: P_\beta>0,|\Delta\beta|\ll\beta,|\Delta V|\ll
309: 1,\beta\Delta\xi\gg1\gg|\Delta\beta\Delta\xi|.\label{eqn:assu}
310: \end{eqnarray}
311: Here, $P_\beta \equiv dP/d\beta>0$ corresponds to the
312: Vakhitov-Kolokolov criterion for the linear stability of solitary
313: waves in Eq. (\ref{eqn:1}) \cite{Kivshar_Agrawal, VK}.
314:
315: Below, we numerically study the weak interaction of solitary waves
316: in Eq. (\ref{eqn:1}). This equation is numerically integrated by the
317: pseudo-spectral method coupled with the fourth-order Runge-Kutta
318: integration along the time direction. Since each solitary wave has
319: four parameters, we have eight parameters in the initial conditions.
320: Due to the phase, translation and Galilean invariances of Eq.
321: (\ref{eqn:1}), we can fix $\sigma_{0,1}=0$, $x_{0,1}+x_{0,2}=0$ and
322: $V_0\equiv (V_{0,1}+V_{0,2})/2=0$ without any loss of generality.
323: Also, for simplicity, we take $\Delta V_0=V_{0,2}-V_{0,1}=0$ in all
324: our simulations of this section, i.e. the two solitary waves are
325: initially at rest. This leaves four free parameters in the initial
326: conditions (\ref{pdeinitial}): $\Delta x_{0} \equiv x_{0,2}-x_{0,1},
327: \Delta\phi_0\equiv \phi_{0,2}-\phi_{0,1}=-\sigma_{0,2},
328: \beta_{0}\equiv (\beta_{0,1}+\beta_{0,2})/2$, and $\Delta \beta_0
329: \equiv \beta_{0,2}-\beta_{0,1}$.
330: %
331: We define the exit velocity $\Delta V_\infty \equiv
332: {\displaystyle\lim_{t\rightarrow +\infty}} \Delta V$. We also define
333: the collision time $\tilde{t}$ as the time when the two solitary
334: waves are the closest (i.e. the separation distance between peaks of
335: the two solitary waves the smallest) during interactions. The life
336: time of interaction is defined as the time length from the beginning
337: $(t=0$) to the collision time $\tilde{t}$, which is equal to the
338: collision time in value. Thus $\tilde{t}$ will be used to denote the
339: life time as well. Of the four parameters in initial conditions, we
340: will fix $\beta_0, \Delta \beta_0$ and $\Delta x_0$, and use
341: $\Delta\phi_0$ as the control parameter and vary it continuously
342: between 0 and $2\pi$. At each $\Delta\phi_0$ value, we simulate the
343: evolution of the two solitary waves and record the exit velocity and
344: the life time. Numerically, the exit velocity is determined as
345: follows. We let the solitary waves propagate for a long time. If
346: they still do not separate, we assign the exit velocity as zero. If
347: they do separate, we wait till they have separated far apart and
348: their velocities stabilized. Then we locate the positions of maximum
349: solitary wave amplitudes at serval different time values. The
350: average separation velocity of the two solitary waves in these time
351: intervals is assigned as the exit velocity. The numerical life time
352: is simply the time when the two solitary waves are the closest in
353: the simulations. Below, we carry out numerical studies of weak
354: interactions as described above on two different nonlinearities: the
355: cubic-quintic and exponential nonlinearities.
356:
357: \subsection{Weak interactions for the cubic-quintic nonlinearity}
358: Our first example of nonlinearity is the cubic-quintic nonlinearity
359: (\ref{cqNL}), which arises in a wide array of physical systems such
360: as optics \cite{Kivshar_Agrawal} and boson condensates
361: \cite{Kovalev,Barashenkov}. In this nonlinearity, we set
362: \begin{eqnarray} \label{alphagamma}
363: \alpha=1, \quad \gamma=0.04.
364: \end{eqnarray}
365: It is easy to verify that all solitary waves (\ref{cqNLs}) in this
366: case are linearly stable using the Vakhitov-Kolokolov criterion. In
367: our simulations, we set $\Delta x_0=10$ and $\beta_0=1$. The $x$
368: interval is 70 units wide, discretized by 512 grid points; and the
369: time step size is 0.004.
370: \begin{figure}
371: %\includegraphics[width=70mm,height=65mm]{cq-nonequal.eps}
372: \includegraphics[width=60mm,height=55mm]{fig1.eps}
373: \caption{\label{cqnonequal}~ The graph of exit velocity $\Delta
374: V_\infty$ versus the initial phase difference $\Delta\phi_0$ in the
375: non-equal initial amplitude case of the cubic-quintic NLS equation
376: (\ref{eqn:1}), (\ref{cqNL}). The cubic and quintic nonlinearity
377: coefficients are given in Eq. (\ref{alphagamma}), and the other
378: (fixed) initial parameters are $\Delta x_0=10$, $\beta_0=1$,
379: $\Delta\beta_0=-0.065$, and $\Delta V_0=0$. }
380: \end{figure}
381: \begin{figure*}
382: %\includegraphics[width=80mm,height=70mm]{level1vslife.eps}
383: \includegraphics[width=80mm,height=70mm]{fig2a.eps} \hspace{0.4cm}
384: \includegraphics[width=80mm,height=70mm]{fig2b.eps}
385: %\includegraphics[width=80mm,height=70mm]{life.eps} \hspace{0.4cm}
386: %\includegraphics[width=80mm,height=70mm]{level1dy.eps}
387: %\includegraphics[width=80mm,height=60mm]{lifefit.eps}
388: %\includegraphics[width=80mm,height=60mm]{peakfit.eps}
389: \caption{\label{pdelevel1}~ (a) The exit velocity versus initial
390: phase difference graph of Fig. 1 re-plotted near the accumulation
391: point of the primary hill sequence; (b) the life time versus initial
392: phase difference graph; (1)-(3): separation versus time diagrams of
393: solitary wave interactions at three values of $\Delta\phi_0$ marked
394: by circles in (a): (1) 0.1759; (2) $-0.0057$; (3) $-0.1053$.}
395: \end{figure*}
396:
397: \begin{figure*}
398: \includegraphics[width=63mm,height=47mm]{fig3a.eps}
399: \includegraphics[width=53mm,height=47mm]{fig3b.eps}
400: \includegraphics[width=53mm,height=45mm]{fig3c.eps}\\
401: \includegraphics[width=63mm,height=40mm]{fig3d.eps}
402: \includegraphics[width=53mm,height=40mm]{fig3e.eps}
403: \includegraphics[width=53mm,height=40mm]{fig3f.eps}
404: %\includegraphics[width=63mm,height=47mm]{cqlevel1.eps}
405: %\includegraphics[width=53mm,height=47mm]{cqlevel2.eps}
406: %\includegraphics[width=53mm,height=45mm]{cqlevel3.eps}\\
407: %\includegraphics[width=63mm,height=40mm]{1511.eps}
408: %\includegraphics[width=53mm,height=40mm]{095248.eps}
409: %\includegraphics[width=53mm,height=40mm]{09669.eps}
410: \caption{\label{fig:PDEzoom}~ Top: the exit velocity versus initial
411: phase difference graph of Fig. 1 and its two zoomed-in structures;
412: bottom: soliton-positions versus time diagrams at three values of
413: $\Delta\phi_0 $ marked by circles in the top panel: (1) 1.511; (2)
414: 0.95248; (3) 0.9669. }
415: \end{figure*}
416:
417: We first study the nonequal-amplitude case, and take
418: $\Delta\beta_0=-0.065$. The $\Delta V_\infty$ versus $\Delta\phi_0$
419: diagram is shown in Fig.\ref{cqnonequal}. The prominent structures
420: in this graph can be split into two regions: one region is
421: $-0.34<\Delta\phi_0<2.5$, and the other region is
422: $2.9<\Delta\phi_0<3.3$. The structures in these two regions turn out
423: to be quite similar (except a horizontal reflection with respect to
424: a vertical axis), thus we focus on the larger region
425: $-0.34<\Delta\phi_0<2.5$ below. The main structure in this region
426: forms a sequence of hills; their widths get smaller from the right
427: to the left, and their heights are about the same. These hills will
428: be called the primary hills. This primary-hill sequence converges to
429: the accumulation point $\Delta\phi_{0c}=-0.339$. In order to see
430: this hill sequence near the accumulation point $\Delta\phi_{0c}$
431: more clearly, we zoom in the region $[-0.35,0.4]$, and the zoomed-in
432: diagram is shown in Fig.\ref{pdelevel1}. In this figure, the
433: cascading sequence can be seen very clearly (see
434: Fig.\ref{pdelevel1}(a)). In Fig.\ref{pdelevel1}(b), the
435: corresponding life-time diagram is displayed. We can see that on the
436: same hill, interactions have roughly the same life time. On
437: different hills, life times are different: hills closer to the
438: accumulation point $\Delta\phi_{0c}$ have longer life times. Between
439: hills, even longer life times can be seen, suggesting more complex
440: dynamics there. To explore differences in interaction dynamics on
441: different hills, we select three points
442: $\Delta\phi_0=0.1759,-0.0057,-0.1053$ (marked in
443: Fig.\ref{pdelevel1}(a) by circles) on three adjacent primary hills.
444: These points are at the same relative positions (roughly halfway
445: between the peak and bottom) of the respective hills. At these
446: points, the interaction dynamics is plotted in
447: Fig.\ref{pdelevel1}(1-3). Here only the separation distance
448: $\Delta\xi$ versus time $t$ graphs are shown. We find that these
449: three dynamical processes are similar, except that the oscillation
450: times before final separation differ by one from one hill to the
451: next. The life times $\tilde{t}_n$ of interactions on this primary
452: hill sequence are found to be an almost perfect linear function of
453: the hill index $n$ as
454: \begin{equation}
455: \omega \tilde{t}_n=2n\pi+\delta, \label{time_PDE}
456: \end{equation}
457: where the least-square linear fit gives
458: \begin{equation} \label{PDEvalue}
459: \omega=0.08605, \quad \delta=2.8897.
460: \end{equation}
461: Here the life time of each primary hill is measured numerically at
462: the relative location of that hill shown in Fig.\ref{pdelevel1}(a)
463: by circles. This life-time formula has the same form as those for
464: all window sequences reported before
465: \cite{Campbell2,Campbell3,Kivshar1,YangTan,Goodman_Haberman2}.
466: %
467: %(see Fig.\ref{pdelevel1}(I)). The locations of the hill peaks
468: %$\Delta\phi_{0,n}$ are also examined, and we find that
469: %$((\Delta\phi_{0,n})^2-(\Delta\phi_{0c})^2)^{-1/2}$ is roughly a
470: %linear function of the window index $n$ from $n=5$ up as
471: %\begin{equation}
472: %((\Delta\phi_{0,n})^2-(\Delta\phi_{0c})^2)^{-1/2}=\mu n+ \nu,
473: %\end{equation}
474: %where the least-square linear fit gives $\mu=0.35776,\nu=1.13839$.
475:
476: In addition to the primary hill sequence as described above,
477: Fig.\ref{cqnonequal} also possesses higher-order structures between
478: primary hills. To demonstrate, we first isolate the long interval
479: $[-0.35, 2.5]$ in Fig. \ref{cqnonequal} and re-plot that part of the
480: graph in Fig.\ref{fig:PDEzoom}(a). Then we zoom into its
481: sub-interval $[0.91, \;0.995]$, which is between the two largest
482: primary hills in Fig.\ref{fig:PDEzoom}(a). The zoomed-in graph is
483: shown in Fig.\ref{fig:PDEzoom}(b). We see that the zoomed-in graph
484: is similar to Fig.\ref{fig:PDEzoom}(a), but the cascading direction
485: has reversed. This behavior is analogous to that reported in
486: \cite{anninos, YangTan2} for the $\phi^4$ model and the coupled NLS
487: equations. The main structure in this zoomed-in window is again two
488: sequences of hills, accumulating to the left and right respectively.
489: We call them secondary hills. Between secondary hills, we see even
490: higher-order structures. To see these structures more clearly, we
491: zoom into the sub-interval $[0.9657, \;0.96785]$, which is between
492: the two largest secondary hills in Fig.\ref{fig:PDEzoom}(b). The
493: zoomed-in graph is shown in Fig.\ref{fig:PDEzoom}(c). We see that it
494: is again similar to Fig. \ref{fig:PDEzoom}(b) but with a reversed
495: cascading direction. One can zoom into the regions between these
496: tertiary hills in Fig.\ref{fig:PDEzoom}(c) further, and will get
497: even higher order structures which are similar to the ones shown in
498: Fig.\ref{fig:PDEzoom}. Thus Fig.\ref{fig:PDEzoom}(a) is a fractal
499: structure! We have also explored the interaction dynamics on this
500: fractal. To demonstrate, we pick three $\Delta \phi_0$ values
501: ~$1.511,\;0.95248,\;0.9669$ which are at the same relative positions
502: of the fractal (roughly halfway between the peak and bottom of the
503: widest hills) in Fig.\ref{fig:PDEzoom}(a)-(c) (marked by circles).
504: The interaction dynamics at these three points are displayed in
505: Fig.\ref{fig:PDEzoom}(1-3) respectively. Here the positions of
506: maximum amplitudes of the interacting waves are plotted against
507: time. We see that these dynamical patterns are clearly similar,
508: except that the numbers of oscillations before final separation are
509: different.
510:
511: %Even though the three graphs in Fig. \ref{fig:PDEzoom}(a, b, c) look
512: %similar, we should point out some important differences between Fig.
513: %\ref{fig:PDEzoom}(a) and its zoomed-in graphs (b) and (c). Notice
514: %that Fig. \ref{fig:PDEzoom}(b) and (c) are indeed extremely similar
515: %not only qualitatively, but also quantitatively. If we reverse
516: %Fig.\ref{fig:PDEzoom}(c), both the locations and widths of its hills
517: %closely match those in Fig.\ref{fig:PDEzoom}(b). However, this is
518: %not so between Fig.\ref{fig:PDEzoom}(a) and (b). More importantly,
519: %if we change initial conditions (such as taking non-zero initial
520: %velocities), the number of primary hills in Fig.\ref{fig:PDEzoom}(a)
521: %can reduce greatly in number (down to two or even one, see Fig.
522: %\ref{evolution}), but if one zooms into such graphs between primary
523: %hills, the zoomed-in graph still contains an infinite sequence of
524: %secondary hills. Further zooms will see infinite tertiary and
525: %higher-order hill sequences as well. This phenomenon will be
526: %illustrated in great detail on the reduced ODE model in Sec. **.
527: %These differences indicate that the primary structure in Fig.
528: %\ref{fig:PDEzoom}(a) and its higher-order structures in Fig.
529: %\ref{fig:PDEzoom}(b,c) are of different characters, and have
530: %different origins. This fact has important implications for the
531: %in-depth analysis of these fractal structures and their underlying
532: %dynamics.
533:
534: In the above numerical simulations, the two solitary waves have
535: different initial amplitudes ($\Delta\beta_0=-0.065$). We have also
536: studied interactions of equal-amplitude solitary waves, i.e. with
537: $\Delta\beta_0=0$, while keeping the other parameters the same. In
538: this case, the graph of exit velocity $\Delta V_\infty$ versus
539: initial phase difference $\Delta \phi_0$ is shown in
540: Fig.\ref{fig:cqequal}. This graph is symmetric with respect to
541: $\Delta \phi_0$ for obvious reasons. Examination of this graph shows
542: that it is also a fractal. Thus fractal dependence arises in weak
543: interactions of both equal and non-equal amplitude solitary waves.
544:
545:
546: \begin{figure}
547: %\includegraphics[width=70mm,height=60mm]{cqequal.eps}
548: \includegraphics[width=70mm,height=60mm]{fig4.eps}
549: \caption{\label{fig:cqequal}~ The exit velocity versus initial phase
550: difference graph in the equal initial amplitude case of the
551: cubic-quintic NLS equation \ref{eqn:1}), (\ref{cqNL}). The cubic and
552: quintic nonlinearity coefficients as well as the initial conditions
553: are the same as in Fig. 1, except that $\Delta\beta_0=0$ now.}
554: \end{figure}
555:
556: \subsection{Weak interactions with exponential nonlinearity}
557: \begin{figure}
558: \includegraphics[width=70mm,height=60mm]{fig5a.eps}
559: \includegraphics[width=70mm,height=60mm]{fig5b.eps}
560: %
561: %\includegraphics[width=70mm,height=60mm]{exnonequal.eps}
562: %\includegraphics[width=70mm,height=60mm]{exequal.eps}
563: \caption{\label{fig:exPDE} The exit velocity versus initial phase
564: difference graphs for the exponential nonlinearity (\ref{exp_non}):
565: (a) the non-equal initial amplitude case with
566: $\Delta\beta_0=-0.045$; (b) the equal initial amplitude case with
567: $\Delta\beta_0=0$. The other (fixed) initial parameters are
568: $\beta_0=2.3$, $\Delta x_0=8$, and $\Delta V_0=0$. }
569: \end{figure}
570: To explore whether the above fractal structures for weak
571: interactions persist or not with other types of nonlinearities, we
572: consider in this subsection a different type of nonlinearity --- the
573: exponential nonlinearity, with
574: \begin{equation} \label{exp_non}
575: F(|U|^2)=e^{|U|}-1.
576: \end{equation}
577: Here, $-1$ is introduced into this function to meet the condition
578: $F(0)=0$. Note that this nonlinearity does not have any parameters.
579: Throughout this subsection, we set the initial separation $\Delta
580: x_0=8$, and average propagation constant $\beta_0=2.3$. We study two
581: cases, one for non-equal amplitudes with $\Delta\beta_0=-0.045$, and
582: the other for equal amplitudes with $\Delta\beta_0=0$. For both
583: cases, the control parameter is $\Delta \phi_0$ as before. In our
584: simulations, the $x$ interval was 70 units wide, discretized by 512
585: grid points. The time step size was 0.002. The $\Delta V_\infty$
586: verse $\Delta\phi_0$ graphs for both cases are plotted in Fig.
587: \ref{fig:exPDE}. We have verified that both graphs in this figure
588: are fractals. Comparing these fractals with those in Figs.
589: \ref{cqnonequal} and \ref{fig:cqequal} of the cubic-quintic
590: nonlinearity, we see that the fractal structures for these two
591: different nonlinearities are very similar. The only major difference
592: between them is in the non-equal amplitude case, where there is only
593: one primary hill sequence (accumulating toward the left) for the
594: exponential nonlinearity, while there are two primary hill sequences
595: for the cubic-quintic nonlinearity. It is remarkable that two very
596: different nonlinearities exhibit quite similar fractal dependence on
597: initial conditions. Thus fractal scattering appears to be a
598: universal feature in weak interactions of Eq. (\ref{eqn:1}) rather
599: than an accident. This leads us to the following questions: how can
600: we analytically establish the universal nature of fractal
601: scatterings for Eq. (\ref{eqn:1}) with general nonlinearities? how
602: can we analytically explain the major differences of fractals for
603: different nonlinearities? These questions will be answered in the
604: following sections.
605:
606:
607: \section{\label{sec:level3}Dynamical Equations\setcounter{equation}{0}}
608: \quad To study weak interactions analytically, we use the
609: Karpman-Solov'ev method \cite{Karpman_Solovev} by treating the
610: interference as a small perturbation to each solitary wave (see also
611: \cite{Gorshkov}). This method has been successfully used before on
612: the NLS equation \cite{Hasegawa_Kodama, Karpman_Solovev, Gorshkov,
613: Gerdjikov_PRL}, the modified NLS equation \cite{Gerdjikov_Yang}, the
614: Manakov equations \cite{Yang_Manakov}, as well as the
615: (non-integrable) coupled NLS equations \cite{YangPRE01}. To proceed,
616: we first need to consider the evolution of a single solitary wave in
617: the perturbed generalized NLS equation
618: \begin{equation}
619: iU_t+U_{xx}+F(|U|^2)U=\epsilon G, \label{eqn:3}
620: \end{equation}
621: where function $G$ is a perturbation term, and $\epsilon$ is a small
622: parameter. Without perturbations ($\epsilon=0$), the solitary wave
623: (\ref{eqn:solution}) is an exact solution of Eq.(\ref{eqn:3}), and
624: its internal parameters $V,\beta,\sigma_0,x_0$ are time-independent.
625: When the perturbation is turned on, these internal parameters of the
626: solitary wave will evolve slowly on the time scale $T=\epsilon t$.
627: The multiple-scale perturbation theory for this slow evolution is
628: well known \cite{YangKaup, YangPRE01}. We write the perturbed
629: solution as
630: \begin{equation}
631: U=\hat{\Phi}(\theta,t,T)e^{iV\theta/2+i\sigma}, \label{eqn:4}
632: \end{equation}
633: where
634: \begin{equation}
635: \theta=x-\int^t_0Vdt-x_0,\quad\sigma=\int^t_0(\beta+V^2/4)dt-\sigma_0.
636: \end{equation}
637: Here $V(T),\beta(T),\sigma_0(T),x_0(T)$ are all functions of slow
638: time $T$. Next, we will derive the dynamical equations (ODEs) for
639: the slow-time evolution of these parameters. Substituting
640: (\ref{eqn:4}) into (\ref{eqn:3}), we get the equation for
641: $\hat{\Phi}$ as
642: \begin{eqnarray}
643: i\hat{\Phi}_t+\hat{\Phi}_{\theta\theta}-\beta\hat{\Phi}+F(\hat{\Phi}^2)\hat{\Phi}=\nonumber
644: \\\epsilon
645: Ge^{-i\phi}
646: -\epsilon\left[i\hat{\Phi}_{\beta}\beta_T-i\hat{\Phi}_{\theta}x_{0T}\right]\nonumber\\
647: -\epsilon \left(Vx_{0T}/2-V_T\theta/2+\sigma_{0T} \right)\hat{\Phi},
648: \end{eqnarray}
649: where $\phi$ is defined in Eq. (\ref{def_phi}). We expand the
650: amplitude function $\hat{\Phi}$ into a perturbation series
651: \begin{equation}
652: \hat{\Phi}=\Phi(\theta;\beta)+\epsilon\tilde{\Phi}+O(\epsilon^2).
653: \end{equation}
654: The equation at order $\epsilon^0$ is satisfied automatically since
655: $\Phi$ satisfies Eq.(\ref{eqn:2}). At order $\epsilon$, the equation
656: for $\tilde{\Phi}$ can be written as
657: \begin{equation}
658: i\Psi_t+\mathcal{L}\Psi=H, \label{eqn:5}
659: \end{equation}
660: where
661: \begin{eqnarray}
662: \Psi=\left(\begin{array}{c}\tilde{\Phi}+\tilde{\Phi}^*\\\tilde{\Phi}^*-\tilde{\Phi}\end{array}\right),
663: \quad
664: \mathcal{L}=\left(\begin{array}{c}0\quad\mathcal{L}_0\\\mathcal{L}_1\quad0\end{array}\right),
665: \end{eqnarray}
666: \begin{eqnarray}
667: &&\mathcal{L}_0=-\partial_{\theta\theta}+\beta-F(\Phi^2),\nonumber\\
668: &&\mathcal{L}_1=-\partial_{\theta\theta}+\beta-F(\Phi^2)-2\Phi^2F'(\Phi^2),
669: \end{eqnarray}
670: and
671: \begin{eqnarray}
672: H=\left[\begin{array}{l}-G^*e^{i\phi}+Ge^{-i\phi}-2i\Phi_{\beta}\beta_T+2i\Phi_{\theta}x_{0T}\\
673: -G^*e^{i\phi}- Ge^{-i\phi}+(Vx_{0T}-\theta
674: V_T+2\sigma_{0T})\Phi\end{array}\right].
675: \end{eqnarray}
676: Here the superscript "*" represents complex conjugation. Operator
677: $\mathcal{L}$ has two eigenfunctions and two generalized
678: eigenfunctions associated with the zero eigenvalue,
679: \begin{eqnarray}
680: \Psi_1=\left(\begin{array}{c}\Phi_{\theta}\\0\end{array}\right),\;\;\quad\quad\Psi_2=\left(\begin{array}{c}0\\\Phi\end{array}\right),\nonumber\\
681: \tilde{\Psi}_1=\left(\begin{array}{c}0\\-\theta\Phi/2\end{array}\right),\quad\tilde{\Psi}_2=\left(\begin{array}{c}-\Phi_{\beta}\\0\end{array}\right),
682: \end{eqnarray}
683: with the relations
684: \begin{eqnarray}
685: \mathcal{L}\Psi_k=0, \hspace{0.5cm}
686: \mathcal{L}\tilde{\Psi}_k=\Psi_k, \quad k=1,2.
687: \end{eqnarray}
688: In order for the inhomogeneous solution $\Psi$ of the first-order
689: equation (\ref{eqn:5}) to be non-secular at large time, the
690: inhomogeneous term in Eq.(\ref{eqn:5}) must be orthogonal to the
691: above eigenfunctions and generalized eigenfunctions of the zero
692: eigenvalue, i.e.,
693: \begin{eqnarray}
694: \left<H,\Psi_k\right>=\left<H,\tilde{\Psi}_k\right>=0,\quad
695: k=1,2,\label{eqn:6}
696: \end{eqnarray}
697: under the inner product defined as
698: \begin{eqnarray}
699: \left<F_1,F_2\right>=\int^{\infty}_{-\infty}F_1^{\dag}\left(\begin{array}{c}
700: 0 \quad 1\\1\quad0\end{array}\right)F_2 d\theta.
701: \end{eqnarray}
702: Here $F_k^{\dag}$ is the Hermitian of $F_k$. Evaluating the four
703: integrals in Eq.(\ref{eqn:6}), the slow-time evolution equations for
704: parameters $V(T),\beta(T),\sigma_0(T),x_0(T)$ will be obtained.
705: These evolution equations can be written as
706: \begin{eqnarray}
707: P\frac{dV}{dT}=2\int^{\infty}_{-\infty}\Phi_\theta(G^*e^{i\phi}+
708: Ge^{-i\phi})d\theta, \label{eqn:dy1}
709: \end{eqnarray}
710: \begin{eqnarray}
711: P_\beta\frac{d\beta}{dT}=\frac{1}{i}\int^{\infty}_{-\infty}\Phi(Ge^{-i\phi}-
712: G^*e^{i\phi})d\theta, \label{eqn:dy2}
713: \end{eqnarray}
714: \begin{eqnarray}
715: P\frac{dx_0}{dT}=\frac{1}{i}\int^{\infty}_{-\infty}\Phi\theta(Ge^{-i\phi}-
716: G^*e^{i\phi})d\theta, \label{eqn:dy3}
717: \end{eqnarray}
718: \begin{eqnarray}
719: P_\beta(\frac{V}{2}\frac{dx_0}{dT}+\frac{d\sigma_0}{dT})=\int^{\infty}_{-\infty}\Phi_\beta(G^*e^{i\phi}+
720: Ge^{-i\phi})d\theta. \label{eqn:dy4}
721: \end{eqnarray}
722: These equations will be critical for the development of weak
723: interaction theory of solitary waves below.
724:
725: Now, we consider the weak interaction of two solitary waves. Here
726: the tail overlapping can be considered as a small perturbation which
727: causes the internal parameters of each solitary wave to evolve on a
728: slow time scale $\epsilon t$. Here $\epsilon$ is the magnitude of
729: tail overlapping which is exponentially small with solitary wave
730: spacing $\Delta \xi$. We will not introduce $\epsilon$ explicitly in
731: the next analysis. To the leading order, the interacting solution is
732: simply a superposition of two solitary waves,
733: \begin{eqnarray}
734: U=U_{1}+U_{2}, \nonumber\\
735: U_{k}=\Phi_{k}e^{i\phi_{k}}, \quad k=1,2,
736: \end{eqnarray}
737: where all parameters slowly vary over time. Picking up the dominant
738: interference terms, each solitary wave is governed by the following
739: perturbed generalized NLS equations:
740: \begin{eqnarray}
741: iU_{k,t}+iU_{k,\theta\theta}+F(|U_k|^2)U_k=H_k, \label{eqn:7}
742: \end{eqnarray}
743: where
744: \begin{eqnarray}
745: H_k=&-&(F(|U_k|^2)+F'(|U_k|^2)|U_k|^2)U_{3-k}\nonumber\\&-&F'(|U_k|^2)U_k^2U_{3-k}^*.
746: \end{eqnarray}
747: In this paper, we only study the weak interaction, so conditions
748: (\ref{eqn:assu}) are assumed. Since $|\Delta V|\ll 1$, the phase
749: difference
750: \begin{eqnarray}
751: \Delta\phi=\phi_2-\phi_1\approx-V\Delta\xi/2+\Delta\sigma,
752: \end{eqnarray}
753: which is independent of $\theta$.
754:
755: Now We apply the above solitary wave perturbation theory to
756: Eq.(\ref{eqn:7}). In this problem,
757: \begin{eqnarray}
758: \epsilon
759: Ge^{-i\phi}=&-&(F(\Phi_k^2)\Phi_{3-k}+F'(\Phi_k^2)\Phi_k^2\Phi_{3-k})e^{(-1)^{k+1}i\Delta\phi}\nonumber\\
760: &-&F'(\Phi_k^2)\Phi_k^2\Phi_{3-k}e^{(-1)^ki\Delta\phi}.
761: \label{eqn:8}
762: \end{eqnarray}
763: Substituting (\ref{eqn:8}) into Eqs.(\ref{eqn:dy1})-(\ref{eqn:dy4}),
764: we obtain the following dynamical equations
765: \begin{widetext}
766: \begin{eqnarray}
767: P_k\frac{dV_k}{dt}=-4\int^{\infty}_{-\infty}\Phi_{k,\theta}(F(\Phi_k^2)\Phi_{3-k}+2F'(\Phi_k^2)\Phi_k^2\Phi_{3-k})d\theta\cos(\Delta\phi),
768: \label{eqn:dy1-1}
769: \end{eqnarray}
770: \begin{eqnarray}
771: P_{k,\beta_k}\frac{d\beta_k}{dt}=(-1)^k2\int^{\infty}_{-\infty}\Phi_kF(\Phi_k^2)\Phi_{3-k}d\theta\sin(\Delta\phi),\label{eqn:dy2-1}
772: \end{eqnarray}
773: \begin{eqnarray}
774: P_k\frac{dx_{k,0}}{dt}=(-1)^k2\int^{\infty}_{-\infty}\Phi_k\theta
775: F(\Phi_k^2)\Phi_{3-k}d\theta\sin(\Delta\phi), \label{eqn:dy3-1}
776: \end{eqnarray}
777: \begin{eqnarray}
778: P_{k,\beta_k}(\frac{V_k}{2}\frac{dx_{k,0}}{dt}+\frac{d\sigma_{k,0}}{dt})=-2\int^{\infty}_{-\infty}\Phi_{k,\beta_k}
779: (F(\Phi_k^2)\Phi_{3-k}+2F'(\Phi_k^2)\Phi_k^2\Phi_{3-k})d\theta\cos(\Delta\phi),
780: \label{eqn:dy4-1}
781: \end{eqnarray}
782: \end{widetext}
783: where $P_k, k=1, 2$ are powers of the two individual solitary
784: waves. These equations can be simplified greatly. Due to
785: assumptions (\ref{eqn:assu}), and noticing that $\Phi(\theta)$ and
786: $\Phi_\beta(\theta)$ are even functions of $\theta$, the
787: leading-order terms of the above integrals can be explicitly
788: calculated. For instance,
789: \begin{eqnarray}
790: &&\int^{\infty}_{-\infty}\Phi_k
791: F(\Phi_k^2)\Phi_{3-k}d\theta\nonumber\\&=&\int^{\infty}_{-\infty}\Phi
792: F(\Phi^2)ce^{(-1)^{k+1}\sqrt{\beta}\theta}d\theta e^{-\sqrt{\beta}\Delta\xi} \nonumber\\
793: &=&\int^{\infty}_{-\infty}(\beta\Phi-\Phi_{\theta\theta})ce^{\sqrt{\beta}\theta}d\theta
794: e^{-\sqrt{\beta}\Delta\xi}\nonumber\\&=&2\sqrt{\beta}c^2e^{-\sqrt{\beta}\Delta\xi}.
795: \end{eqnarray}
796: Similarly,
797: \begin{eqnarray}
798: &&\int^{\infty}_{-\infty}\Phi_{k,\theta}(F(\Phi_k^2)\Phi_{3-k}+2F'(\Phi_k^2)\Phi_k^2\Phi_{3-k})d\theta\nonumber\\
799: &=&\int^{\infty}_{-\infty}\Phi_\theta(F(\Phi^2)+2F'(\Phi^2)\Phi^2)ce^{(-1)^{k+1}\sqrt{\beta}\theta}d\theta e^{-\sqrt{\beta}\Delta\xi}\nonumber\\
800: \nonumber\\&=&(-1)^k\sqrt{\beta}\int^{\infty}_{-\infty}\Phi
801: F(\Phi^2)ce^{\sqrt{\beta}\theta}d\theta e^{-\sqrt{\beta}\Delta\xi}
802: \nonumber\\&=&(-1)^k2\beta c^2e^{-\sqrt{\beta}\Delta\xi},
803: \end{eqnarray}
804: \begin{eqnarray}
805: &&\int^{\infty}_{-\infty}\Phi_k\theta F(\Phi_k^2)\Phi_{3-k}d\theta
806: \nonumber\\&=&\int^{\infty}_{-\infty}\Phi\theta
807: F(\Phi^2)ce^{(-1)^{k+1}\sqrt{\beta}\theta}d\theta e^{-\sqrt{\beta}\Delta\xi}\nonumber\\
808: &=&(-1)^{k+1}\int^{\infty}_{-\infty}\Phi\theta
809: F(\Phi^2)ce^{\sqrt{\beta}\theta}d\theta
810: e^{-\sqrt{\beta}\Delta\xi}\nonumber\\
811: & \frac{\underline{\textit{\footnotesize def}}}
812: {\;}&(-1)^{k+1}D_1e^{-\sqrt{\beta}\Delta\xi}, \label{eqD1}
813: \end{eqnarray}
814: \begin{eqnarray}
815: &&\int^{\infty}_{-\infty}\Phi_{k,\beta_k}(F(\Phi_k^2)\Phi_{3-k}+2F'(\Phi_k^2)\Phi_k^2\Phi_{3-k})d\theta\nonumber\\
816: &=&\int^{\infty}_{-\infty}\Phi_\beta(F(\Phi^2)+2F'(\Phi^2)\Phi^2)ce^{(-1)^{k+1}\sqrt{\beta}\theta}d\theta e^{-\sqrt{\beta}\Delta\xi}\nonumber\\
817: &=&\int^{\infty}_{-\infty}\Phi_\beta(F(\Phi^2)+2F'(\Phi^2)\Phi^2)ce^{\sqrt{\beta}\theta}d\theta e^{-\sqrt{\beta}\Delta\xi}\nonumber\\
818: & \frac{\underline{\textit{\footnotesize
819: def}}}{\;}&D_2e^{-\sqrt{\beta}\Delta\xi}. \label{eqD2}
820: \end{eqnarray}
821: With the above simplifications, the dynamical equations reduce to
822: \begin{eqnarray}
823: P\frac{dV_k}{dt}=(-1)^{k+1}8\beta
824: c^2\cos(\Delta\phi)e^{-\sqrt{\beta}\Delta\xi}, \label{eqn:simpD1}
825: \end{eqnarray}
826: \begin{eqnarray}
827: P_\beta\frac{d\beta_k}{dt}=(-1)^k4\sqrt{\beta}c^2\sin(\Delta\phi)e^{-\sqrt{\beta}\Delta\xi},\label{eqn:simpD2}
828: \end{eqnarray}
829: \begin{eqnarray}
830: P\frac{dx_{k,0}}{dt}=-2D_1\sin(\Delta\phi)e^{-\sqrt{\beta}\Delta\xi},
831: \label{eqn:simpD3}
832: \end{eqnarray}
833: \begin{eqnarray}
834: P_\beta(\frac{V}{2}\frac{dx_{k,0}}{dt}+\frac{d\sigma_{k,0}}{dt})=-2D_2\cos(\Delta\phi)e^{-\sqrt{\beta}\Delta\xi}
835: ,\label{eqn:simpD4}
836: \end{eqnarray}
837: where $P$ is the power of the solitary wave with propagation
838: constant $\beta$, and $D_1,D_2$ are defined in
839: (\ref{eqD1}),(\ref{eqD2}). From the above equations, we find that
840: \begin{eqnarray}
841: \beta_t=V_t=0, \label{Eq1}
842: \end{eqnarray}
843: \begin{eqnarray}
844: \Delta\xi_t=\Delta V,\label{Eq2}
845: \end{eqnarray}
846: \begin{eqnarray}
847: \Delta\phi_t=\Delta\beta,\label{Eq3}
848: \end{eqnarray}
849: \begin{eqnarray}
850: \Delta V_t=-\frac{16\beta
851: c^2}{P}\cos(\Delta\phi)e^{-\sqrt{\beta}\Delta\xi},\label{Eq4}
852: \end{eqnarray}
853: \begin{eqnarray}
854: \Delta\beta_t=\frac{8\sqrt{\beta}c^2}{P_\beta}\sin(\Delta\phi)
855: e^{-\sqrt{\beta}\Delta\xi}\label{Eq5}.
856: \end{eqnarray}
857: Equations (\ref{Eq1})-(\ref{Eq5}) are the key results in the weak
858: interaction theory of solitary waves.
859: %Thus, we can get the dynamical equations for integrable NLS equation
860: %\begin{eqnarray}
861: %\Delta
862: %V_t=-32\beta^\frac{3}{2}\cos(\Delta\phi)e^{-\sqrt{\beta}\Delta\xi},\label{Dy-inte1}
863: %\end{eqnarray}
864: %\begin{eqnarray}
865: %\Delta\beta_t=32\beta^2{P_\beta}\sin(\Delta\phi)e^{-\sqrt{\beta}\Delta\xi}\label{Dy-inte2}.
866: %\end{eqnarray}
867: %\begin{eqnarray}
868: %\Delta\xi_t=\Delta V, \Delta\phi_t=\Delta\beta,\label{Dy-inte3}
869: %\end{eqnarray}
870: These equations can be further simplified by variable rescalings.
871: Introducing notations
872: \begin{eqnarray} \label{scaling1}
873: \psi=\Delta\phi,\zeta=-\sqrt{\beta}\Delta\xi,f=\frac{16\beta^{3/2}
874: c^2}{P},g=\frac{8\sqrt{\beta}c^2}{P_\beta},\quad
875: \end{eqnarray}
876: and
877: \begin{equation} \label{time_scale}
878: \tau=\sqrt{f}\: t, \quad\varepsilon=\frac{g}{f}-1=\frac{P}{2\beta
879: P_\beta}-1,
880: \end{equation}
881: then the dynamical equations (\ref{Eq2})-(\ref{Eq5}) reduce to
882: \begin{eqnarray}\{\begin{array}{l}
883: \zeta_{\tau\tau}=\cos{\psi}e^{\zeta}\\
884: %\psi_{\tau\tau}=h\sin{\psi}e^{\zeta}=(1+\varepsilon)\sin{\psi}e^{\zeta}
885: \psi_{\tau\tau}=(1+\varepsilon)\sin{\psi}e^{\zeta}
886: \end{array}\quad. \label{Dyreduce}
887: \end{eqnarray}
888: Eq. (\ref{Dyreduce}) is the final dynamical system we obtained for
889: the analytical treatment of weak interactions in the generalized NLS
890: equations (\ref{eqn:1}). It is important to remark that Eq.
891: (\ref{Dyreduce}) is universal for the generalized NLS equations with
892: arbitrary nonlinearities. It contains only a single parameter
893: $\varepsilon$, which depends on the specific form of nonlinearity.
894:
895: Eq. (\ref{Dyreduce}) has the following general properties. First, it
896: is Hamiltonian with the conserved Hamiltonian (energy) as
897: \begin{equation} \label{Edef}
898: E=\frac{1}{2}(\dot{\zeta}^2-\dot{\psi}^2)-e^\zeta \cos\psi
899: %+\frac{1}{2}(1-\frac{1}{1+\varepsilon}) \dot{\psi}^2.
900: +\frac{\varepsilon}{2(1+\varepsilon)} \dot{\psi}^2.
901: \end{equation}
902: %\begin{eqnarray}
903: %E&=&\frac{1}{2}(\dot{\zeta}^2-h\dot{\psi}^2)-\cos\psi
904: %e^\zeta\nonumber\\&=&\frac{1}{2}(\dot{\zeta}^2-\dot{\psi}^2)-\cos\psi
905: %e^\zeta+(1-h)\dot{\psi}^2.
906: %\end{eqnarray}
907: Here $\dot{(\;)}\equiv d/d\tau$. Second, it has some symmetry
908: properties. One is that it is time-reversible, i.e., if
909: $[\zeta(\tau), \psi(\tau)]$ is a solution with initial conditions
910: $(\zeta_0, \dot{\zeta}_0, \psi_0, \dot{\psi}_0)$, then
911: $[\zeta(-\tau), \psi(-\tau)]$ is a solution with initial conditions
912: $(\zeta_0, -\dot{\zeta}_0, \psi_0, -\dot{\psi}_0)$. Another symmetry
913: is on phase flipping, i.e., if $[\zeta(\tau), \psi(\tau)]$ is a
914: solution with initial conditions $(\zeta_0, \dot{\zeta}_0, \psi_0,
915: \dot{\psi}_0)$, then $[\zeta(\tau), -\psi(\tau)]$ is a solution with
916: initial conditions $(\zeta_0, \dot{\zeta}_0, -\psi_0,
917: -\dot{\psi}_0)$. Physically, this latter symmetry corresponds to the
918: interchange of the left and right solitary waves in the PDE
919: evolutions, which of course does not change the interaction outcome.
920: Eq. (\ref{Dyreduce}) also has the property that if $\psi(\tau)$ is a
921: solution, so is $\psi(\tau)+2n\pi$ for any integer of $n$. This
922: reflects the fact that in the PDE system, solution evolution remains
923: the same if the phase difference between the solitary waves changes
924: by a multiple of $2\pi$.
925:
926: The dynamical equations (\ref{Dyreduce}) are asymptotically accurate
927: in describing weak interactions in the PDE system (to the leading
928: order) when the spacing $\Delta \xi$ is large. Surprisingly, even
929: when the two solitary waves come close to each other, Eq.
930: (\ref{Dyreduce}) can still describe the interaction process very
931: well. This is analogous to weak interactions in the (integrable) NLS
932: equation \cite{Hasegawa_Kodama, Karpman_Solovev}. Below, we make
933: detailed comparisons between the ODE solutions of
934: Eq.(\ref{Dyreduce}) and the PDE solutions in Sec. 3(A) for the
935: cubic-quintic nonlinearity. Inserting the parameter values
936: $\alpha=1, \gamma=0.04$ and $\beta_0=1$ of PDE simulations into Eqs.
937: (\ref{Pformula}) and (\ref{cformula}), we get $P=3.74720$,
938: $P_{\beta}=1.64835$, and $c=2.69495$, thus $f=31.01080$,
939: $g=35.24845$, and $\varepsilon=0.13665$. Corresponding to the
940: initial conditions for PDE simulations in Sec. 3(A), the initial
941: conditions for the ODE system (\ref{Dyreduce}) with nonequal and
942: equal initial amplitudes are
943: \begin{equation} \label{ic_nonequal}
944: \zeta_0=-10, \; \dot{\zeta}_0=0, \; \dot{\psi}_0=-0.01167,
945: \end{equation}
946: and
947: \begin{equation} \label{ic_equal}
948: \zeta_0=-10, \; \dot{\zeta}_0=0, \; \dot{\psi}_0=0,
949: \end{equation}
950: respectively. In both cases, the initial phase difference $\psi_0$
951: is the control parameter as in PDE simulations. The ODE system
952: (\ref{Dyreduce}) is numerically solved by the fourth-order
953: Runge-Kutta method, with the time step set as $0.01$. The simulation
954: results on the exit velocity $-\dot{\zeta}_\infty$ versus $\psi_0$
955: graph are shown in Fig.\ref{fig:ODEglobal}.
956: \begin{figure*}
957: \includegraphics[width=80mm,height=65mm]{fig6a.eps}
958: \includegraphics[width=80mm,height=65mm]{fig6b.eps}
959: %\includegraphics[width=80mm,height=65mm]{odenonequal.eps}
960: %\includegraphics[width=80mm,height=65mm]{odeequal.eps}
961: \caption{\label{fig:ODEglobal} The exit velocity
962: ($-\dot{\zeta}_\infty$) versus the initial phase difference
963: ($\psi_0$) graphs from the ODE model (\ref{Dyreduce}). Here the
964: initial conditions for (a) and (b) are chosen corresponding to the
965: PDE simulation results in Figs. 1 and 4 respectively (see text). }
966: \end{figure*}
967: Clearly, these graphs are very similar to Figs. \ref{cqnonequal} and
968: \ref{fig:cqequal}) from PDE simulations. We have also investigated
969: the detailed structures of Fig. \ref{fig:ODEglobal} in ways
970: analogous to what we did for Figs. 2 and 3. Specifically, we have
971: examined the primary hill sequence in Fig. \ref{fig:ODEglobal}(a),
972: and zoomed into regions between primary hills. The results are shown
973: in Figs.\ref{ODElevel1} and \ref{fig:ODE-zoom} respectively. Both
974: figures closely resemble Figs. 2 and 3 from the PDE simulations.
975:
976: The agreement between the ODE model and the PDE simulations is not
977: only qualitative, but also quantitative. To demonstrate, we compare
978: the locations and life times of primary hill sequences in Figs. 2
979: and \ref{ODElevel1}. The comparison results are summarized in Table
980: 1. Very good quantitative agreement between them can be seen. In the
981: ODE model, the life time is also an almost perfect linear function
982: of the hill index $n$ in the form (\ref{time_PDE}). When the time
983: rescaling (\ref{time_scale}) is recovered, the ODE model gives
984: \begin{equation} \label{ODEvalue}
985: \omega|_{ODE}=0.08570,\quad \delta|_{ODE}=2.9655,
986: \end{equation}
987: closely resembling the corresponding values (\ref{PDEvalue}) from
988: the PDE simulations.
989:
990:
991: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
992: \begin{table}
993: \caption{\label{tab:table1} Comparison on locations and life times
994: of primary hills in Figs. 2(a) and 7(a) from the PDE and ODE
995: simulations. }
996: \begin{ruledtabular}
997: \begin{tabular}{lcccc}
998: &\quad location (PDE) &location (ODE) & life (PDE) & life (ODE)\\
999: \hline
1000: $n$ & $\Delta\phi_{0,n}$ & $\psi_{0,n}$ & $t_n$ & $\sqrt{f}\tau_n$\\
1001: \hline
1002: 1&1.7735 & 1.7794 & 65 &68 \\
1003: 2&0.6985 & 0.7015 &117 &119\\
1004: 3&0.2430 & 0.2468 &183 &185\\
1005: 4&0.0280 & 0.0359 &253 &255\\
1006: 5&-0.0850 &-0.0763 &325 &327\\
1007: 6&-0.1530 &-0.1431 &398 &400\\
1008: 7&-0.1963 &-0.1863 &470 &473\\
1009: 8&-0.2258 &-0.2157 &544 &547\\
1010: 9&-0.2475 &-0.2367 &617 &620\\
1011: 10&-0.2630&-0.2523 &691 &693\\
1012: \hline
1013: $\infty$&-0.3392&-0.3280&$\infty$&$\infty$
1014: \end{tabular}
1015: \end{ruledtabular}
1016: \end{table}
1017: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1018:
1019: %Next, we compare the life-times of the primary hill sequence in
1020: %Figs. 2 and \ref{ODElevel1}. These life times from the ODE model are
1021: %found to be an almost perfect linear function of the window index
1022: %$n$, just like those in the PDE simulations. In addition, after the
1023: %time scales in the ODE model are adjusted back to the PDE case, the
1024: %life-time formula from the ODE model is
1025: %\begin{equation}
1026: %\omega t_n=2n\pi+\delta,
1027: %\end{equation}
1028: %with $\omega=0.01539, ???? \delta=-3.31771$. This formula agrees
1029: %well with the PDE formula (\ref{time_PDE}) quantitatively ?????.
1030:
1031: \begin{figure*}
1032: \includegraphics[width=80mm,height=70mm]{fig7a.eps}
1033: \hspace{0.4cm}
1034: \includegraphics[width=80mm,height=70mm]{fig7b.eps}
1035: %
1036: %\includegraphics[width=80mm,height=70mm]{life-ode.eps}
1037: %\hspace{0.4cm}
1038: %\includegraphics[width=80mm,height=70mm]{dyna.eps}
1039: %\includegraphics[width=80mm,height=60mm]{lifefit-ode.eps}
1040: %\includegraphics[width=80mm,height=60mm]{peakfit-ode.eps}
1041: \caption{\label{ODElevel1}~ (a) The exit velocity versus initial
1042: phase difference graph of Fig. \ref{fig:ODEglobal}(a) re-plotted
1043: near the accumulation point of the primary hill sequence; (b) the
1044: life time versus initial phase difference graph; (1)-(3): separation
1045: ($-\zeta$) versus time ($\tau$) diagrams at three values of $\psi_0$
1046: marked by circles in (a): (1) 0.187; (2) 0.0056; (3) $-0.0939$. All
1047: these graphs are obtained from the ODE model (\ref{Dyreduce}), and
1048: they should be compared to the corresponding PDE graphs in Fig.
1049: \ref{pdelevel1}. }
1050: \end{figure*}
1051:
1052:
1053: \begin{figure*}
1054: \includegraphics[width=63mm,height=47mm]{fig8a.eps}
1055: \includegraphics[width=53mm,height=47mm]{fig8b.eps}
1056: \includegraphics[width=53mm,height=45mm]{fig8c.eps}\\
1057: \includegraphics[width=60mm,height=30mm]{fig8d.eps}
1058: \includegraphics[width=53mm,height=30mm]{fig8e.eps}
1059: \includegraphics[width=53mm,height=30mm]{fig8f.eps}
1060: %
1061: %\includegraphics[width=63mm,height=47mm]{zoom1_ode.eps}
1062: %\includegraphics[width=53mm,height=47mm]{zoom2_ode.eps}
1063: %\includegraphics[width=53mm,height=45mm]{zoom3_ode.eps}\\
1064: %\includegraphics[width=60mm,height=30mm]{p1532.eps}
1065: %\includegraphics[width=53mm,height=30mm]{p095071.eps}
1066: %\includegraphics[width=53mm,height=30mm]{p09660278.eps}
1067: \caption{\label{fig:ODE-zoom} Top: the exit velocity versus
1068: initial phase difference graph of Fig. \ref{fig:ODEglobal} and
1069: its two zoomed-in structures; bottom: separation versus time
1070: diagrams at three values of $\psi_0 $ marked by circles in the top
1071: panel: (1) 1.532; (2) 0.95071; (3) 0.96603. These graphs from the
1072: ODE model should be compared to the corresponding PDE graphs in
1073: Fig. \ref{fig:PDEzoom}. }
1074: \end{figure*}
1075:
1076: Above we have established that the reduced ODE system
1077: (\ref{Dyreduce}) accurately describes weak interactions of the PDE
1078: system. Since the ODE system (\ref{Dyreduce}) is universal for Eq.
1079: (\ref{eqn:1}) regardless of details of its nonlinearities, we see
1080: that the hill sequences and fractal structures in Eq.
1081: (\ref{Dyreduce}) are universal for weak interactions of solitary
1082: waves in the PDE system (\ref{eqn:1}), as Figs. \ref{cqnonequal},
1083: \ref{fig:cqequal}, \ref{fig:exPDE}, and \ref{fig:ODEglobal} clearly
1084: indicate.
1085:
1086:
1087: Next we will turn our attention to the ODE system (\ref{Dyreduce}),
1088: and analyze its solution dynamics in more detail. In particular, we
1089: would like to understand why fractal structures arise in this
1090: system, and how to analytically predict their locations and other
1091: main features.
1092:
1093: \section{Solutions of the integrable dynamical equations and their
1094: singularity conditions
1095: \setcounter{equation}{0}}
1096:
1097: Eq.(\ref{Dyreduce}) conserves energy (\ref{Edef}) for all values of
1098: $\varepsilon$. When $\varepsilon=0$, it has another conserved
1099: quantity,
1100: \begin{eqnarray}
1101: M=\dot{\zeta}\dot{\psi}-e^\zeta \sin\psi,
1102: \end{eqnarray}
1103: which can be called the momentum of this system. In this particular
1104: case, system (\ref{Dyreduce}) is an integrable Hamiltonian system
1105: and can be analytically solved. Let $Y=\zeta+i\psi$,
1106: Eq.(\ref{Dyreduce}) becomes
1107: \begin{eqnarray}
1108: Y_{\tau\tau}=e^Y.
1109: \end{eqnarray}
1110: The general solution of this equation is
1111: \begin{eqnarray}
1112: %Y(\tau)=\ln\left(-2C^2_1\textrm{sech}^2\left(C_1\left(\tau+C_2\right)\right)\right),\label{GSolu}
1113: Y(\tau)=\ln\left[-2C^2_1\textrm{sech}^2\left(C_1\tau+C_2\right)\right],\label{GSolu}
1114: \end{eqnarray}
1115: where
1116: \begin{equation} \label{C1}
1117: C_1=\frac{1}{2}\sqrt{{\dot{Y}_0}^2-2e^{Y_0}}=\frac{1}{\sqrt{2}}\sqrt{E+i
1118: M},
1119: \end{equation}
1120: and
1121: \begin{equation} \label{C2}
1122: C_2=-\textrm{arctanh}\left(\frac{\dot{Y}_0}{2C_1}\right).
1123: \end{equation}
1124: Here the branch of the square root function in (\ref{C1}) is chosen
1125: such that $\rm{Re}(C_1)\ge 0$. It is noted that solutions $Y$ which
1126: differ by a multiple of $2\pi i$ correspond to the same physical
1127: solution, thus it does not matter which Riemann surface one takes
1128: for the logarithmic function in Eq. (\ref{GSolu}). If $C_1=0$, i.e.,
1129: $\dot{Y}_0=\pm \sqrt{2}\hspace{0.02cm} e^{Y_0/2}$, the solution
1130: (\ref{GSolu}) degenerates to the form
1131: \begin{eqnarray}
1132: Y(\tau)=-2\ln\left(e^{-\frac{1}{2}Y_0}\mp
1133: \frac{1}{\sqrt{2}}\;\tau\right).
1134: \end{eqnarray}
1135: The asymptotic behaviors of these solutions as
1136: $\tau\rightarrow\infty$ can be easily determined. Let
1137: \begin{equation} \label{C12abcd}
1138: C_1=a+b \hspace{0.05cm}i, \quad \frac{C_2}{C1}=c+d \hspace{0.05cm}
1139: i,
1140: \end{equation}
1141: where $a,b,c,d$ are real constants, then the following leading-order
1142: asymptotic expressions for the solution can be obtained when
1143: $\tau\rightarrow\infty$:
1144: \begin{eqnarray}
1145: %&&(1)\; a>0: \quad Y(\tau) \rightarrow -2a\tau-i2b\tau,
1146: %\\&&(2)\; a<0: \quad
1147: %Y(\tau) \rightarrow
1148: %2a\tau+i2b\tau, \\
1149: && (1)\; a \neq 0: \;\; Y(\tau) \rightarrow
1150: -2|a|\tau-\mbox{sgn}(a) \hspace{0.02cm} 2b\tau \hspace{0.007cm} i; \label{ane0} \\
1151: && (2)\; a=0, \; b\neq0: \nonumber \\
1152: && (2a)\; d=0: \; \nonumber\\
1153: && \quad Y(\tau) =\ln 2b^2-\ln\cos^2b(\tau+c); \label{asym_dzero}\\
1154: &&(2b)\;d\neq 0: \nonumber \\
1155: && \quad Y(\tau) = \ln
1156: 4b^2-\ln\left[\cosh 2bd+\cos 2b(\tau+c)\right] \nonumber \\
1157: &&\quad\quad\quad\quad +
1158: 2i\arctan\left[\textrm{tanh} (bd)\hspace{0.01cm} \tan b(\tau+c)\right]; \\
1159: &&(3)\; a=0,b=0: \;\; Y(\tau) \rightarrow -2\ln({\mp\tau}).
1160: \end{eqnarray}
1161:
1162: From these asymptotic expressions, we see that when $a\ne 0$, the
1163: two solitary waves eventually move away from each other with exit
1164: velocity $2|a|$; when $a=0$ but $b\ne 0$, the solution is
1165: time-periodic for both $d=0$ and $d\ne 0$, the difference being that
1166: in the former case, the periodic solution exhibits finite-time
1167: singularities (where $\zeta=\mbox{Re}(Y) \to \infty$), while in the
1168: latter case, the solution has no singularities; when $a=b=0$, the
1169: two solitary waves eventually separate logarithmically, and the exit
1170: velocity is zero. As an example, we take the initial conditions
1171: (\ref{ic_nonequal}). In this case, the graph of exit velocity
1172: $-\dot{\zeta}_\infty (=2|a|)$ versus $\psi_0$ is plotted in Fig.
1173: \ref{figS} (bottom panel). This graph is smooth everywhere, except
1174: at $\psi_0=0,\pm\pi$ where it has a cusp (due to the absolute-value
1175: function in $|a|$). The squares and diamonds on this graph will be
1176: explained later. Clearly, this graph has no fractal structure
1177: anywhere. Thus, fractal dependence is a signature of the dynamical
1178: system (\ref{Dyreduce}) when it is non-integrable (with $\varepsilon
1179: \ne 0$), not when it is integrable (with $\varepsilon=0$).
1180:
1181: The above asymptotic states do not tell the full story about the
1182: solution dynamics in the integrable system. For instance, for the
1183: case of $a\ne 0$, even though the solution has a benign-looking
1184: asymptotics (\ref{ane0}) as $\tau \to \infty$, the solution can
1185: still develop a singularity (where the separation $\zeta \to
1186: \infty$) at a finite time. These solutions with finite-time
1187: singularities turn out to be critical for the appearance of fractal
1188: structures in the non-integrable system, as our numerics in the next
1189: section will indicate. Thus we analyze these singularity solutions
1190: in more details below. The necessary and sufficient conditions for
1191: singularities in solution (\ref{GSolu}) are that
1192: \begin{eqnarray} \label{sing_cond}
1193: %e^{C_1\tilde{\tau}+C_2}+e^{-C_1\tilde{\tau}-C_2}=0,
1194: \cosh(C_1\tilde{\tau}+C_2)=0,
1195: \end{eqnarray}
1196: and $\tilde{\tau}>0$, where $\tilde{\tau}$ is the time of
1197: singularity. If $\tilde{\tau}<0$, i.e., singularities in the
1198: solution occur at a negative time, such singularities are irrelevant
1199: for the time evolution of Eq. (\ref{Dyreduce}) and need not be
1200: considered. The solutions of Eq. (\ref{sing_cond}) are
1201: \begin{eqnarray} \label{tildetau_form}
1202: %C_1(\tilde{\tau}+C_2)=\frac{1}{2}(2n+1)\pi i, \nonumber \\
1203: %n=\cdots,-2,-1,0,1,2,\cdots.
1204: C_1\tilde{\tau}+C_2=\frac{1}{2}(2n+1)\pi i, \quad n=0, \pm 1, \pm 2,
1205: \cdots.
1206: \end{eqnarray}
1207: This is a complex-valued relation, which gives two real relations on
1208: $\tilde{\tau}$, $C_1$ and $C_2$. When $a\ne 0$, i.e., $C_1$ is not
1209: purely imaginary, we find by separating the real and imaginary parts
1210: of Eq. (\ref{tildetau_form}) that the solution (\ref{GSolu}) has a
1211: single finite-time singularity of the type $\ln(\tau-\tilde{\tau})$
1212: if the following conditions are satisfied:
1213: %\begin{eqnarray}
1214: %&&S\equiv \frac{\mbox{Im}(C_2)}{\mbox{Re}(\frac{1}{C_1})}
1215: %=\frac{1}{2}\left(2n+1\right)\pi, \nonumber\\ &&\qquad \qquad n=0,
1216: %\pm 1, \pm 2, \cdots, \label{condition2}
1217: %\end{eqnarray}
1218: \begin{equation}
1219: S\equiv \frac{\mbox{Im}(C_1^*C_2)}{\mbox{Re}(C_1)}
1220: =\frac{1}{2}\left(2n+1\right)\pi, \quad n=0, \pm 1, \pm 2, \cdots,
1221: \label{condition2}
1222: \end{equation}
1223: \begin{equation}
1224: \tilde{\tau}= -\frac{\rm{Re}(C_2)}{\rm{Re}(C_1)}> 0.
1225: \label{condition1}
1226: \end{equation}
1227: Here Re$(\cdot)$ and Im$(\cdot)$ represent the real and imaginary
1228: parts of a complex number. When $a=0 \; (b\ne 0)$, singularity
1229: solutions exist if $d=0$. These solutions have an infinite number of
1230: finite-time singularities of the type $\ln(\tau-\tilde{\tau})$, as
1231: the formula (\ref{asym_dzero}) indicates. Physically, at the time of
1232: singularity $\tilde{\tau}$, the two solitary waves strongly collide,
1233: thus $\tilde{\tau}$ is the collision time. Whether conditions
1234: (\ref{condition2}) and (\ref{condition1}) can be satisfied depends
1235: on the initial conditions (which determine the $C_1$ and $C_2$
1236: values, see (\ref{C1}), (\ref{C2})). In the text below, we will call
1237: initial conditions $(\zeta_0, \dot{\zeta}_0, \psi_0, \dot{\psi}_0)$
1238: which satisfy Eqs. (\ref{condition2}) and (\ref{condition1}) as
1239: singularity points. At singularity points, solutions of the
1240: integrable dynamical system (\ref{Dyreduce}) develop finite-time
1241: singularities.
1242:
1243: %\begin{equation} \label{C11}
1244: %C_1=\frac{1}{2}i \sqrt{0.01167^2+2e^{-10+i\psi_0}},
1245: %\end{equation}
1246: %\begin{eqnarray} \label{C12}
1247: %&& C_2=-\frac{2i}{\sqrt{0.01167^2+2e^{-10+i\psi_0}}} \times
1248: %\nonumber \\
1249: %&& arctanh \frac{0.01167}{\sqrt{0.01167^2+2e^{-10+i\psi_0}}}.
1250: %\end{eqnarray}
1251: %Define
1252: %\begin{equation} \label{S}
1253: %S(\psi_0) \equiv
1254: %\mbox{Im}\left[(C_2-C_2^*)/(\frac{1}{C_1}+\frac{1}{C_1^*})\right]=
1255: %\frac{\mbox{Im}(C_2)}{\mbox{Re}(\frac{1}{C_1})},
1256: %\end{equation}
1257: %where Re$(\cdot)$ and Im$(\cdot)$ represent the real and imaginary
1258: %parts of a complex number,
1259: To demonstrate how to determine singularity points in the
1260: initial-condition space, we take initial conditions
1261: (\ref{ic_nonequal}) of Fig. \ref{fig:ODEglobal}(a) as an example.
1262: Here $\psi_0$ is a control parameter. With these initial conditions,
1263: the graph of function $S(\psi_0)$ is plotted in Fig. \ref{figS} (top
1264: panel). This graph has a maximum 0.96. As $\psi_0 \to 0^+$ or
1265: $\pi^-$, $S(\psi_0) \to -\infty$.
1266: %The singularity
1267: %condition (\ref{condition2}) now becomes
1268: %\begin{equation} \label{cond3}
1269: %S(\psi_0)=\frac{1}{2}\left(2n+1\right)\pi, \quad n=0, \pm 1, \pm 2,
1270: %\cdots.
1271: %\end{equation}
1272: As we can see from this graph, for any value of $n\le -1$, Eq.
1273: (\ref{condition2}) has two roots, $\psi_{0, n}^{(1)}$ and $\psi_{0,
1274: n}^{(2)}$. We have checked that these roots satisfy the other
1275: singularity condition (\ref{condition1}), thus these $\psi_{0,
1276: n}^{(1)}$ and $\psi_{0, n}^{(2)}$ values are singularity points. It
1277: is noted that the graph of function $S(\psi_0)$ also has another
1278: piece in the interval $\pi < \psi_0 < 2\pi$, which is the mirror
1279: image of that shown in Fig. \ref{figS} around the point
1280: $\psi_0=\pi$. But in that interval, $\tilde{\tau}<0$, not satisfying
1281: the second singularity condition (\ref{condition1}), thus we did not
1282: plot that piece of the graph in Fig. \ref{figS}.
1283:
1284: Next, we examine these singularity points $\psi_{0, n}^{(1)}$ and
1285: $\psi_{0, n}^{(2)}$ in more detail. These points form two infinite
1286: sequences with $n=-1, -2, \dots$, which accumulate at $\psi_0=0^+$
1287: and $\pi^-$ respectively. In Fig. \ref{figS} (bottom panel), these
1288: two sequences are marked by squares and diamonds on the exit
1289: velocity versus $\psi_0$ graph. Calculating the asymptotics of $C_1$
1290: from Eq. (\ref{condition2}) and substituting it into Eq.
1291: (\ref{condition1}), we find that the collision times
1292: $\tilde{\tau}_n$ of both sequences have the following asymptotic
1293: expressions:
1294: \begin{equation} \label{collide_time}
1295: \omega \tilde{\tau}_n= 2|n|\pi + \pi, \quad n \to -\infty,
1296: \end{equation}
1297: where $\omega=2\mbox{Im}(C_1|_{\psi_0=0})$ and
1298: $2\mbox{Im}(C_1|_{\psi_0=\pi})$ for the left and right sequences
1299: respectively. The asymptotic formulas for the locations of these two
1300: singularity sequences \{$\psi_{0, n}^{(1)}$\} and \{$\psi_{0,
1301: n}^{(2)}$\} can also be calculated. We find from (\ref{condition2})
1302: that
1303: \begin{equation} \label{location_sing}
1304: \psi_{0, n}^{(1)} \to \frac{A_1}{(2n+1)\pi}, \quad n \to -\infty,
1305: \end{equation}
1306: and
1307: \begin{equation} \label{location_sing2}
1308: \pi-\psi_{0, n}^{(2)} \to \frac{A_2}{(2n+1)\pi}, \quad n \to
1309: -\infty,
1310: \end{equation}
1311: where
1312: \[A_1=\left.8e^{-\zeta_0}\rm{Re}(C_2)\rm{Im}^2(C_1)\right|_{\psi_0=0},
1313: \]
1314: and
1315: \[A_2=\left.8e^{-\zeta_0}\rm{Re}(C_2)\rm{Im}^2(C_1)\right|_{\psi_0=\pi}.
1316: \]
1317: %\begin{figure*}
1318: \begin{figure}
1319: %\includegraphics[width=63mm,height=47mm]{fig_S.eps}
1320: \includegraphics[width=80mm,height=60mm]{fig9a.eps}
1321: %\includegraphics[width=80mm,height=60mm]{fig_S3.eps}
1322:
1323: \vspace{0.3cm}
1324: \includegraphics[width=84mm,height=30mm]{fig9b.eps}
1325: %\includegraphics[width=84mm,height=30mm]{fig_spsi2b.eps}
1326: %\includegraphics[width=63mm,height=40mm]{fig_S_yi2.eps}
1327: \caption{Top: graph of the function $S(\psi_0)$ defined in Eq.
1328: (\ref{condition2}) for the initial conditions (\ref{ic_nonequal}).
1329: Intersections of the graph with horizontal lines are singularity
1330: points. Bottom: exit velocity versus $\psi_0$ graph in the
1331: integrable system (\ref{Dyreduce}). Both squares and diamonds are
1332: singularity points. \label{figS}}
1333: \end{figure}
1334: %\end{figure*}
1335:
1336: The above detailed analysis on singularity points was performed for
1337: the particular initial conditions (\ref{ic_nonequal}) where the two
1338: solitary waves are initially stationary ($\dot{\zeta}_0=0$). What
1339: will happen if $\dot{\zeta}_0 \neq 0$? To answer this question, we
1340: fix $\zeta_0$ and $\dot{\psi}_0$ as in Eq. (\ref{ic_nonequal}), vary
1341: $\dot{\zeta}_0$, and examine how singularity points move in the
1342: ($\psi_0$, $\dot{\zeta}_0$) plane. The results are shown in
1343: Fig.\ref{singularpoints}. The top curve corresponds to $n=-1$, the
1344: next curve corresponding to $n=-2$, and so on. All curves are
1345: bounded from both above and below except the top one (with $n=-1$).
1346: When $n\rightarrow -\infty$, these curves approach the accumulation
1347: curve plotted by the dashed line in Fig.\ref{singularpoints}. Below
1348: this accumulation curve, there are no singularity points. The
1349: analytical formula for this accumulation curve can be easily
1350: derived. On this accumulation curve, $C_1$ must be pure imaginary,
1351: thus
1352: \begin{eqnarray}
1353: &M=\dot{\zeta}_{0c}\dot{\psi}_0-e^{\zeta_0}\sin{\psi_{0}}=0,\label{accumulation1}\\
1354: &E=\frac{1}{2}(\dot{\zeta}_{0c}^2-\dot{\psi}_0^2)-e^{\zeta_0}\cos{\psi_{0}}<0.
1355: \label{accumulation2}
1356: \end{eqnarray}
1357: Here $(\dot{\zeta}_{0c}, \psi_{0})$ is an accumulation point. From
1358: Eq. (\ref{accumulation1}), we see that the function of the
1359: accumulation curve is
1360: \begin{equation} \label{sine_curve}
1361: \dot{\zeta}_{0c}=\frac{e^{\zeta_0}\sin{\psi_{0}}}{\dot{\psi}_0}.
1362: \end{equation}
1363: The maximum and minimum of this curve are
1364: \begin{equation} \label{threshold}
1365: \dot{\zeta}_{0c, \rm{min}}=-\dot{\zeta}_{0c, \rm{max}}=
1366: -\left|\frac{e^{\zeta_0}}{\dot{\psi}_0}\right|.
1367: \end{equation}
1368: For the $\zeta_0$ and $\dot{\psi}_0$ values of Fig.
1369: \ref{singularpoints}, we get $\dot{\zeta}_{0c, \rm{min}}=-0.00389$.
1370: If $\dot{\zeta}_0 < \dot{\zeta}_{0c, \rm{min}}$, there are no
1371: singularity solutions for any value of $\psi_0$. When
1372: $\dot{\zeta}_{0c, \rm{min}}< \dot{\zeta}_0 < \dot{\zeta}_{0c,
1373: \rm{max}}$, two infinite sequences of singularity points can be
1374: found. When $\dot{\zeta}_0>\dot{\zeta}_{0c, \rm{max}}$, however, the
1375: number of singularity points becomes finite; this number gradually
1376: decreases (down to one) as $\dot{\zeta}_0$ increases.
1377:
1378: \begin{figure}
1379: %\begin{minipage}[c]{.6\textwidth}
1380: \includegraphics[width=80mm,height=65mm]{fig10.eps}
1381: %\includegraphics[width=80mm,height=65mm]{zeros.eps}
1382: %\end{minipage}
1383: %\begin{minipage}[c]{.28\textwidth}
1384: \centering \caption{\label{singularpoints}Singularity points
1385: satisfying conditions (\ref{condition2})-(\ref{condition1}) in the
1386: ($\psi_0$, $\dot{\zeta}_0$) plane. The dashed curve is the
1387: accumulation curve. Here $\zeta_0=-10, \dot{\psi}_0=-0.01167$. }
1388: %\end{minipage}
1389: \end{figure}
1390:
1391: The above calculations of singularity points and their accumulation
1392: curves were made for special choices of initial conditions
1393: $\zeta_0=-10$ and $\dot{\psi}_0=-0.01167$ (see Fig.
1394: \ref{singularpoints}). In view of the importance of singularity
1395: points for fractal structures which we will reveal in the next
1396: section, we would like to discuss singularity points and their
1397: accumulation curves further for {\it general} initial conditions
1398: below.
1399:
1400: First, we examine the accumulation curve in the ($\psi_0$,
1401: $\dot{\zeta}_0$) plane for general initial conditions $\zeta_0$ and
1402: $\dot{\psi}_0$. In this general case, the accumulation curve (if it
1403: exists) is necessarily given by Eq. (\ref{sine_curve}). But the
1404: curve (\ref{sine_curve}) (or portions of it) may not satisfy
1405: condition (\ref{accumulation2}), thus may not actually be the
1406: accumulation curve. Below we determine what portions of the curve
1407: (\ref{sine_curve}) are the accumulation curve. Before we do so, let
1408: us first point out that conditions (\ref{accumulation1}) and
1409: (\ref{accumulation2}) are not only the necessary, but also
1410: sufficient conditions for the accumulation curve. In addition, the
1411: accumulation of singularity points toward the accumulation curve is
1412: always from the upper side, not lower side. To show these, we only
1413: need to prove that condition (\ref{condition1}) holds only on the
1414: upper edge of the curve (\ref{sine_curve}), but not the lower edge
1415: of it. On the upper edge of (\ref{sine_curve}), $C_1$ is purely
1416: imaginary, and $\rm{sgn}(M)=\rm{sgn}(\dot{\psi}_{0})$. Thus
1417: $\rm{sgn}(\rm{Im}(C_1))=\rm{sgn}(M)=\rm{sgn}(\dot{\psi}_{0})$.
1418: Consequently, $\rm{Re}(\dot{Y}_0/2C_1)>0$. Notice that for any
1419: complex number $z$, $\mbox{Re}[\mbox{tanh}(z)]$ and $\mbox{Re}(z)$
1420: have the same sign, hence $\rm{Re}(C_2)<0$. Then due to
1421: $\rm{Re}(C_1)>0$, condition (\ref{condition1}) thus holds. By
1422: similar reasoning, we can show that on the lower edge of the curve
1423: (\ref{sine_curve}), condition (\ref{condition1}) does not hold. Thus
1424: singularity points accumulate toward (\ref{sine_curve}) only from
1425: above, not below.
1426:
1427: Now we turn to equations (\ref{accumulation1}) and
1428: (\ref{accumulation2}), and use them to determine the accumulation
1429: curve for the general case. Substituting Eq. (\ref{sine_curve})
1430: into inequality (\ref{accumulation2}) and simplifying, this
1431: inequality becomes
1432: \begin{equation}
1433: -\frac{1}{\dot{\psi}_0^2}\left(\dot{\psi}_0^2+e^{\zeta_0}(1+\cos\psi_{0})\right)
1434: \left(\dot{\psi}_0^2-e^{\zeta_0}(1-\cos\psi_{0})\right)<0,
1435: \end{equation}
1436: which is equivalent to
1437: \begin{equation} \label{cospsi2}
1438: \cos\psi_{0} > 1-e^{-\zeta_0}\dot{\psi}_0^2.
1439: \end{equation}
1440: Thus the accumulation curve is the parts of curve (\ref{sine_curve})
1441: where $\psi_{0}$ satisfies the constraint (\ref{cospsi2}). If
1442: $\dot{\psi}_0^2>2e^{\zeta_0}$, condition (\ref{cospsi2}) is
1443: satisfied for all values of $\psi_0$, hence the entire curve
1444: (\ref{sine_curve}) is the accumulation curve. If $0<\dot{\psi}_0^2
1445: \le 2e^{\zeta_0}$, portions of the curve (\ref{sine_curve}) centered
1446: at $\psi_0=\pi$ do not satisfy condition (\ref{cospsi2}), thus do
1447: not belong to the accumulation curve. The rest of the curve
1448: (\ref{sine_curve}) does satisfy condition (\ref{cospsi2}), thus is
1449: the accumulation curve. If $\dot{\psi}_0=0$ (equal initial amplitude
1450: case), no value of $\psi_0$ satisfies condition (\ref{cospsi2}),
1451: thus accumulation points do not exist.
1452:
1453: Next, we derive two general properties about singularity points in
1454: the $(\psi_0, \dot{\zeta}_0)$ plane for general initial conditions
1455: $\zeta_0$ and $\dot{\psi}_0$. One property is that, if
1456: $\dot{\zeta}_{0c}$ is on the accumulation curve, then for any
1457: $\dot{\zeta}_0 <\dot{\zeta}_{0c}$, singularity points can not exist.
1458: We will prove this by showing that $\tilde{\tau}<0$ for
1459: $\dot{\zeta}_0 <\dot{\zeta}_{0c}$. To show $\tilde{\tau}<0$, we only
1460: need to show $\rm{Re}(\dot{Y}_0/2C_1)<0$ (see above). Without loss
1461: of generality, we only show this for $\dot{\psi}_0<0$; the proof for
1462: $\dot{\psi}_0>0$ is similar (in fact, as has been pointed out
1463: before, flipping the sign of $\dot{\psi}_0$ physically amounts to
1464: interchanging the positions of the left and right solitary waves and
1465: thus does not affect the interaction outcome). For $\dot{\psi}_0<0$
1466: and $\dot{\zeta}_0 <\dot{\zeta}_{0c}$, $C_1$ is in the first
1467: quadrant (as $M>0$). If $\dot{\zeta}_0<0$, then $\dot{Y}_0$ is in
1468: the third quadrant, thus $\rm{Re}(\dot{Y}_0/2C_1)<0$ holds. Now we
1469: consider $0<\dot{\zeta}_0<\dot{\zeta}_{0c}$. In this case,
1470: $\dot{Y}_0$ is in the fourth quadrant, hence $i\dot{Y}_0$ lies in
1471: the first quadrant (like $C_1$). To show
1472: $\rm{Re}(\dot{Y}_0/2C_1)<0$, we only need to show
1473: $\rm{arg}(i\dot{Y}_0)<\rm{arg}(2C_1)$. Since both $i\dot{Y}_0$ and
1474: $C_1$ are in the first quadrant, we only need to show
1475: $\rm{arg}(-\dot{Y}_0^2)<\rm{arg}(4C_1^2)$. Notice that
1476: \begin{equation}
1477: -\dot{Y}_0^2+4C_1^2=-2e^{Y_0},
1478: \end{equation}
1479: which is independent of $\dot{\zeta}_0$. In addition, the angle of
1480: $-2e^{Y_0}$ falls in between those of $-\dot{Y}_0^2$ and $4C_1^2$.
1481: Thus to show $\rm{arg}(-\dot{Y}_0^2)<\rm{arg}(4C_1^2)$, we only need
1482: to show $\rm{arg}(-\dot{Y}_0^2)<\rm{arg}(-2e^{Y_0})$. Note that
1483: \begin{equation}
1484: -\dot{Y}_0^2=\dot{\psi}_0^2-\dot{\zeta}_0^2-2i\dot{\psi}_0\dot{\zeta}_0,
1485: \end{equation}
1486: whose angle is an increasing function of $\dot{\zeta}_0$ when
1487: $\dot{\psi}_0<0$, thus for $\dot{\zeta}_0 <\dot{\zeta}_{0c}$,
1488: \begin{equation} \label{argY}
1489: \rm{arg}(-\dot{Y}_0^2)<\rm{arg}(\dot{\psi}_0^2-\dot{\zeta}_{0c}^2-2i\dot{\psi}_0\dot{\zeta}_{0c}).
1490: \end{equation}
1491: Now recall that $\dot{\zeta}_{0c}$ lies on the accumulation curve,
1492: thus it satisfies the conditions (\ref{accumulation1}) and
1493: (\ref{accumulation2}). Substituting these conditions into
1494: (\ref{argY}), and recalling our assumptions of $\dot{\psi}_0<0$ and
1495: $0<\dot{\zeta}_0 <\dot{\zeta}_{0c}$, we find that the right hand
1496: side of (\ref{argY}) is less than $\rm{arg}(-2e^{Y_0})$, thus
1497: inequality $\rm{arg}(-\dot{Y}_0^2)<\rm{arg}(-2e^{Y_0})$ is proved.
1498: Summarizing the above arguments, we conclude that for any
1499: $\dot{\zeta}_0$ below the accumulation curve, singularity points do
1500: not exist in the $(\psi_0, \dot{\zeta}_0)$ plane.
1501:
1502: %
1503: %
1504: Another general property about singularity points is that, at
1505: sufficiently large values of $\dot{\zeta}_0$, there is a unique
1506: singularity point in the $\psi_0$ interval. The proof is as follows.
1507: It is easy to check that when $\dot{\zeta}_0^2+\dot{\psi}_0^2 \gg
1508: e^{\zeta_0}$ and $|\dot{\zeta}_0|\ _{\sim }^{>}\ |\dot{\psi}_0|$,
1509: functions $S$ and $\tilde{\tau}$ have the following leading-order
1510: asymptotic expressions,
1511: \begin{equation} \label{tauasym}
1512: \tilde{\tau} \to \frac{1}{\dot{\zeta}_0}\mbox{ln}
1513: \frac{2(\dot{\zeta}_0^2+\dot{\psi}_0^2)}{e^{\zeta_0}},
1514: \end{equation}
1515: \begin{equation} \label{Sasym}
1516: S \to \frac{1}{2}\mbox{sgn}(\dot{\zeta}_0)\psi_0+S_A,
1517: \end{equation}
1518: where
1519: \begin{equation}
1520: S_A=\frac{\dot{\psi}_0}{2|\dot{\zeta}_0|}\ln
1521: \frac{2(\dot{\zeta}_0^2+\dot{\psi}_0^2)}{e^{\zeta_0}}-
1522: \mbox{arctan}\frac{\dot{\psi}_0}{|\dot{\zeta}_0|}-\frac{1}{2}\pi,
1523: \end{equation}
1524: and the relative errors are
1525: $O\left(e^{\zeta_0}/(\dot{\zeta}_0^2+\dot{\psi}_0^2)\right)$. From
1526: Eq. (\ref{Sasym}), we see that the rise of $S$ value over the
1527: interval $0\le \psi_0 \le 2\pi$ is $\pi$, which guarantees that Eq.
1528: (\ref{condition2}) has a single solution in the $\psi_0$ interval
1529: for a single value of $n$. From Eq. (\ref{tauasym}), we see that
1530: when $\dot{\zeta}_0>0$ is sufficiently large, $\tilde{\tau}>0$ over
1531: the entire $\psi_0$ interval. Thus singularity conditions
1532: (\ref{condition2}) and (\ref{condition1}) admit a unique singularity
1533: point. We note by passing that when $\dot{\zeta}_0$ is sufficiently
1534: large negative, $\tilde{\tau}<0$ over the entire $\psi_0$ interval,
1535: thus there can not be any singularity points. This is consistent
1536: with the previous general property proved above.
1537:
1538: To summarize the above results on singularity points and
1539: accumulation points and slightly extend them, we present the
1540: following classifications on singularity solutions in the integrable
1541: system (\ref{Dyreduce}):
1542: \begin{enumerate}
1543: %\item $\dot{\zeta}_0=\dot{\psi}_0=0$ (zero initial velocities and
1544: %equal initial amplitudes): a singularity solution exists at the
1545: %single value $\psi_0=0$ (see case (2a) in (\ref{asym_dzero}));
1546: \item $\dot{\psi}_0=0$ (equal initial amplitudes):
1547: \begin{itemize}
1548: \item If $\dot{\zeta}_0 > -\sqrt{2}e^{\zeta_0/2}$, a singularity solution exists
1549: at the single singularity point $\psi_0=0$. Here $M=0$, and $E>0$
1550: for $\dot{\zeta}_0 > \sqrt{2}e^{\zeta_0/2}$ and $E<0$ otherwise;
1551: \item If $\dot{\zeta}_0 < -\sqrt{2}e^{\zeta_0/2}$, there are no
1552: singularity solutions for any ${\psi_0}$;
1553: \end{itemize}
1554:
1555: \item $\dot{\psi}_0\ne 0$, $\dot{\zeta}_0=0$
1556: (non-equal initial amplitudes, zero initial velocities):
1557: \begin{itemize}
1558: \item If $\dot{\psi}_0^2 > 2e^{\zeta_0}$, singularity solutions exist
1559: at two infinite sequences of $\psi_0$ values, accumulating at
1560: $\psi_0=\{0^+, \pi^-\}$, or $\{\pi^+, 2\pi^-\}$, for
1561: $\dot{\psi}_0<0$ and $\dot{\psi}_0>0$ respectively;
1562: \item If $0<\dot{\psi}_0^2 < 2e^{\zeta_0}$: singularity solutions exist at
1563: one infinite sequence of $\psi_0$ values, accumulating at
1564: $\psi_0=0^+$ or $2\pi^-$ for $\dot{\psi}_0<0$ and $\dot{\psi}_0>0$
1565: respectively;
1566: \end{itemize}
1567: On these sequences of singularity points, $M\ne 0$ and $E\ne 0$
1568: generically (at the accumulation points, $M=0$, and $E<0$);
1569:
1570: \item $\dot{\psi}_0\ne 0$ (the general non-equal initial amplitude
1571: case):
1572:
1573: In this case, the accumulation curve is the parts of curve
1574: (\ref{sine_curve}) where $\psi_0$ satisfies the constraint
1575: (\ref{cospsi2}). When $\dot{\psi}_0^2>2e^{\zeta_0}$, the entire
1576: curve (\ref{sine_curve}) is the accumulation curve. When
1577: $0<\dot{\psi}_0^2 \le 2e^{\zeta_0}$, the accumulation curve is
1578: (\ref{sine_curve}) except portions of it which are centered at
1579: $\psi_0=\pi$.
1580:
1581: For $\dot{\zeta}_0$ below the accumulation curve, there are no
1582: singularity points; at sufficiently large $\dot{\zeta}_0$ values,
1583: there is a single singularity point.
1584:
1585: At all these singularity values, $E$ and $M$ are non-zero
1586: generically (except the accumulation points where $M=0$).
1587: \end{enumerate}
1588: It is noted that in the above classifications, case (2) is just a
1589: special case of case (3), and can be readily deduced from (3). Case
1590: (1) can be deduced from (3) as well under the limit $\dot{\psi}_0
1591: \to 0$. But cases (1) and (2) are important special cases, hence we
1592: listed them out separately.
1593:
1594:
1595: \section{Origins of fractal structures in the non-integrable dynamical equations
1596: \setcounter{equation}{0}}
1597:
1598: We have known from Figs. \ref{fig:ODEglobal}, \ref{ODElevel1} and
1599: \ref{fig:ODE-zoom} that the non-integrable ODE system
1600: (\ref{Dyreduce}) exhibits hill sequences and fractal structures
1601: which coincide with those in the PDE simulations, but such
1602: structures do not exist when this ODE system becomes integrable. The
1603: natural question then is: where do the fractal structures in the
1604: non-integrable system (\ref{Dyreduce}) come from? In this section,
1605: we will establish through careful numerics that these fractal
1606: structures bifurcate from singularity points of the integrable
1607: system.
1608: \begin{figure}
1609: %\begin{minipage}[c]{.6\textwidth}
1610: \includegraphics[width=80mm,height=65mm]{fig11a.eps}\\
1611: \includegraphics[width=80mm,height=40mm]{fig11b.eps}
1612: %\includegraphics[width=80mm,height=65mm]{ga2zero.eps}\\
1613: %\includegraphics[width=80mm,height=40mm]{ga2zero2.eps}
1614: %\end{minipage}
1615: %\begin{minipage}[c]{.28\textwidth}
1616: \centering \caption{\label{ga2zero}~ The exit velocity versus
1617: initial phase difference graphs in the ODE model (\ref{Dyreduce}) at
1618: various values of $\varepsilon$: (1) 0.13665; (2) 0.036; (3) 0.0036;
1619: (4) 0; (5) $-0.0036$; (6) $-0.036$. The initial conditions are given
1620: in (\ref{ic_nonequal}). The squares and diamonds in (4) are
1621: singularity points of the integrable system (see Fig. 9, bottom). }
1622: %\end{minipage}
1623: \end{figure}
1624:
1625: \begin{figure}
1626: %\begin{minipage}[c]{.6\textwidth}
1627: \includegraphics[width=80mm,height=50mm]{fig12a.eps}\\
1628: \includegraphics[width=80mm,height=50mm]{fig12b.eps}
1629: %\includegraphics[width=80mm,height=50mm]{evolution1.eps}\\
1630: %\includegraphics[width=80mm,height=40mm]{evolution2.eps}
1631: %\end{minipage}
1632: %\begin{minipage}[c]{.28\textwidth}
1633: \centering \caption{\label{evolution} The exit velocity versus
1634: initial phase difference graphs in the ODE model (\ref{Dyreduce})
1635: at various values of $\dot{\zeta}_0$: (1) 0.00707; (2) 0.00548;
1636: (3) 0.00495; (4) 0.00350; (5) 0; (6) $-0.00350$; (7) $-0.00424$.
1637: Here $\zeta_0=-10, \dot{\psi}_0=-0.01167$, and
1638: $\varepsilon=0.0036$. }
1639: %\end{minipage}
1640: \end{figure}
1641:
1642: To determine the origin of these fractals, we take the same
1643: initial conditions (\ref{ic_nonequal}) as in Fig.
1644: \ref{fig:ODEglobal}(a), but gradually decrease the value of
1645: $\varepsilon$ from 0.13665 of Fig. \ref{fig:ODEglobal}(a) down to
1646: zero (the integrable case), then down further to negative values.
1647: In this process, we closely monitor how the fractal structure of
1648: Fig. \ref{fig:ODEglobal}(a) changes as $\varepsilon$ decreases.
1649: The result is shown in Fig. \ref{ga2zero}. Here, the
1650: $-\dot{\zeta}_\infty$ verse $\psi_0$ graphs are plotted at six
1651: decreasing $\varepsilon$ values: $\varepsilon=0.13665, 0.036,
1652: %
1653: %I changed 0.035 to 0.036 for presentation purpose 0.036,
1654: %
1655: 0.0036, 0, -0.0036$ and $-0.036$. We see that as $\varepsilon$
1656: decreases from 0.13665 but above zero, primary hill sequences and
1657: the fractal regions between them persist and are clearly visible
1658: in Fig. \ref{ga2zero}(1, 2, 3). Indeed, we have zoomed into the
1659: sensitive regions between primary hills in each of Figs.
1660: \ref{ga2zero}(1, 2, 3), and obtained higher order structures which
1661: look very similar to those shown in Fig. \ref{fig:ODE-zoom}. As
1662: $\varepsilon \to 0^+$, our key observation is that, the peaks of
1663: individual primary hills as well as the nearby fractal regions
1664: collapse to sequences of points on the smooth
1665: $-\dot{\zeta}_\infty$ curve of the integrable system (see Figs.
1666: \ref{ga2zero}(3, 4)). Closer examination tells us that, these
1667: sequences of points in Fig. \ref{ga2zero}(4) are nothing but the
1668: two sequences of singularity points of the integrable system which
1669: we plotted in Fig. \ref{figS}! In other words, hill sequences and
1670: fractal structures in the non-integrable system bifurcate from the
1671: singularity points of the integrable system! However, this
1672: bifurcation is one-sided: as $\varepsilon$ decreases below zero,
1673: no fractal regions appear, see Fig. \ref{ga2zero}(5). A finite
1674: number of primary hills, reminiscent of primary hill sequences for
1675: positive $\varepsilon$ values, do exist. But the whole graph is
1676: smooth, and it has no fractal structures inside (even the
1677: spike-looking parts of the graph in Fig. \ref{ga2zero}(5) turn out
1678: to be smooth upon closer examination). Furthermore, as
1679: $\varepsilon$ decreases further below zero, the number of primary
1680: hills keeps decreasing, and the graph becomes more smooth, see
1681: Fig. \ref{ga2zero}(6). Thus, fractal structures are a signature of
1682: the non-integrable system (\ref{Dyreduce}) only for positive
1683: values of $\varepsilon$, not negative values of $\varepsilon$.
1684:
1685: To further substantiate our claim on fractal structures of the
1686: non-integrable system bifurcating from singularity points of the
1687: integrable system, we tune initial conditions so that singularity
1688: points in the integrable system gradually disappear in the
1689: $\psi_0$ interval, and check if fractal structures in the
1690: non-integrable system disappear as well (for small $\varepsilon$).
1691: Specifically, we fix the $\zeta_0$ and $\dot{\psi}_0$ values in
1692: Eq. (\ref{ic_nonequal}) and tune the $\dot{\zeta}_0$ value, as we
1693: did in Fig. \ref{singularpoints}. The $\varepsilon$ value in Eq.
1694: (\ref{Dyreduce}) is taken as $\varepsilon=0.0036$, which is very
1695: small. Thus, the non-integrable system is weakly perturbed from
1696: the integrable one. For the above initial conditions, singularity
1697: points of the integrable system have been displayed in Fig.
1698: \ref{singularpoints} in the $(\psi_0, \dot{\zeta}_0)$ plane. We
1699: gradually decrease the $\dot{\zeta}_0$ value. For each
1700: $\dot{\zeta}_0$, we numerically compute the exit velocity versus
1701: $\psi_0$ graph in the perturbed (non-integrable) system
1702: (\ref{Dyreduce}), and compare how this graph relates to
1703: singularity points of the integrable system in Fig.
1704: \ref{singularpoints}. To illustrate, we pick seven representative
1705: $\dot{\zeta}_0$ values, which are 0.00707, 0.00548, 0.00495,
1706: 0.00350, 0, $-0.00350$ and $-0.00424$ in decreasing order. These
1707: seven $\dot{\zeta}_0$ values are marked by horizontal dashed lines
1708: in Fig. \ref{singularpoints}. As we can see from that figure, at
1709: these seven $\dot{\zeta}_0$ values, the numbers of singularity
1710: points in the $\psi_0$ interval are 1, 3, 5, $\infty$, $\infty$,
1711: $\infty$, and 0 respectively. For each of these seven
1712: $\dot{\zeta}_0$ values, the corresponding exit velocity versus
1713: $\psi_0$ graph in the perturbed system (\ref{Dyreduce}) is shown
1714: in Fig. \ref{evolution}. We notice from this figure that the
1715: numbers of primary hills and fractal regions near these hills at
1716: these $\dot{\zeta}_0$ values are equal to 1, 3, 5, $\infty$,
1717: $\infty$, $\infty$, and 0 respectively
1718: --- exactly like singularity points in the integrable system! In
1719: particular, when singularity points in the integrable system
1720: disappear, so do primary hills and fractal structures in the weakly
1721: perturbed non-integrable system. Furthermore, the locations of
1722: primary hills and fractal regions closely follow those of
1723: singularity points of the integrable system. Thus, the connections
1724: between them are unmistakable. Fig. \ref{evolution}, together with
1725: Fig. \ref{ga2zero}, establishes beyond doubt that primary hills and
1726: fractal structures in the non-integrable system (\ref{Dyreduce})
1727: indeed bifurcate from singularity points of the integrable system.
1728:
1729:
1730: The bifurcation of fractal structures from singularity points of the
1731: integrable system indicates that near such points, the solutions of
1732: the perturbed system (\ref{Dyreduce}) are very sensitive to initial
1733: conditions. To shed light on why this sensitivity occurs, we present
1734: some numerical results below. First, we look at the integrable
1735: system (with $\varepsilon=0$). Taking the initial conditions as
1736: (\ref{ic_nonequal}), evolutions of $\psi$ versus $\tau$ at the
1737: singularity point $\psi_0=0.98325$ (marked in Fig. \ref{figS},
1738: bottom panel) and its left and right near neighbors $\psi_0=0.92$
1739: and 1.05 are plotted in Fig. \ref{senstive}(a) using the solution
1740: formula (\ref{GSolu}). An interesting feature about these evolutions
1741: is that for initial $\psi_0$ values at the two sides of the
1742: singularity point, the phase functions $\psi(\tau)$ have drastically
1743: different trajectories as they go through the time $\tau\approx 700$
1744: where the two solitary waves interact strongly (this time is the
1745: singularity time of the singular solution at $\psi_0=0.98325$). For
1746: $\psi_0$ below the singularity point, the phase sharply (but
1747: continuously) decreases by $2\pi$, while for $\psi_0$ above the
1748: singularity point, the phase sharply (but continuously) increases by
1749: $2\pi$. Recall that $\dot{\psi}\propto \Delta \beta$ and it
1750: determines the relative energy (amplitude) distributions among the
1751: two solitary waves [see Eqs. (\ref{Eq3}), (\ref{scaling1}) and
1752: (\ref{time_scale})], we know that on the two sides of the
1753: singularity point, the energies have opposite distributions among
1754: the two solitary waves during their strong interactions. However,
1755: after the interaction is completed, the asymptotic slopes of the
1756: three $\psi(\tau)$ trajectories in Fig. \ref{senstive}(a) are almost
1757: the same, signaling that the interaction outcome is actually
1758: insensitive to the $\psi_0$ values (the roughly $2\pi$ difference
1759: between these phase trajectories does not affect the physical
1760: solutions). This is why outcomes of weak interactions in the
1761: integrable system (\ref{Dyreduce}) do not exhibit sensitive
1762: dependence on initial conditions (see Fig. \ref{figS} bottom panel).
1763: However, when system (\ref{Dyreduce}) is positively perturbed, the
1764: results are completely different. To demonstrate, we take
1765: $\varepsilon=0.0036$ now, while the initial conditions
1766: (\ref{ic_nonequal}) remain the same. In this slightly perturbed
1767: system, the solution develops finite-time singularity at the
1768: singularity point $\psi_0=0.9695$, which is the counterpart of the
1769: singularity point mentioned above in the integrable system. The
1770: phase function at this singularity point is plotted in Fig.
1771: \ref{senstive}(b) (solid line). (It is noted that in this perturbed
1772: case, we do not have exact solution formulas, hence this solution
1773: was obtained by numerically integrating Eq. (\ref{Dyreduce}). Due to
1774: the finite-time singularity in the solution, our numerical
1775: integration can not go beyond the singularity time $\tau \approx
1776: 700$. The solution beyond the singularity time, shown in Fig.
1777: \ref{senstive}(b) as dotted lines, was inferred from our numerics at
1778: nearby $\psi_0$ values.) On the two sides of the singularity point,
1779: we select two nearby values $\psi_0=0.92$ and 1.01. The phase
1780: trajectories at these $\phi_0$ values are also plotted in Fig.
1781: \ref{senstive}(b). We see that as these trajectories go through the
1782: time $\tau \approx 700$, one sharply decreases by $2\pi$, while the
1783: other sharply increases by $2\pi$, similar to what happens in the
1784: integrable case (see Fig. \ref{senstive}(a)). However, after these
1785: sharp decreases/increases, the trajectories turn around and start to
1786: move in the opposite direction. Eventually, these trajectories
1787: approach drastically different asymptotic slopes (one positive and
1788: the other one negative in fact), indicating that the interaction
1789: outcomes are very different for these slight changes in the $\psi_0$
1790: values. This is the phenomenon of sensitive dependence on initial
1791: conditions which occurs in the perturbed system (\ref{Dyreduce})
1792: (with $\varepsilon>0$), but not the integrable system (with
1793: $\varepsilon=0$).
1794:
1795: It is also enlightening to look at this sensitive dependence on
1796: initial conditions from the viewpoint of PDE evolutions. To
1797: illustrate, we take the cubic-quintic nonlinearity (\ref{cqNL}) in
1798: Eq. (\ref{eqn:1}), and take $\alpha=1$, $\beta_0=1$. Then for the
1799: $\varepsilon$ values and initial conditions used in the ODE
1800: simulations of Fig. \ref{senstive}, and in view of the variable
1801: rescalings (\ref{scaling1}) and (\ref{time_scale}), the
1802: corresponding PDE parameters for Fig. \ref{senstive}(a) (the
1803: integrable case) are $\gamma=0$, $\Delta \beta_0=-0.066016$, $\Delta
1804: V_0=0$, $\Delta x_0=10$, and the corresponding PDE parameters for
1805: Fig. \ref{senstive}(b) (the perturbed case) are $\gamma=0.0010$,
1806: $\Delta \beta_0=-0.066$, $\Delta V_0=0$, $\Delta x_0=10$. For these
1807: PDE parameters, the PDE evolution results (in the form of contour
1808: plots) at three $\Delta \phi_0$ values corresponding to those in the
1809: ODE simulations of Fig. \ref{senstive} are displayed in Fig.
1810: \ref{contour}. In the integrable (NLS) case (top row of Fig.
1811: \ref{contour}), we take the three $\Delta \phi_0$ values exactly the
1812: same as those in Fig. \ref{senstive}(a), i.e. $\Delta \phi_0=0.92,
1813: 0.98325, 1.05$. In this case, at the lower $\Delta \phi_0$ value,
1814: the left solitary wave retains its higher energy at the collision
1815: time; at the singularity point of $\Delta \phi_0$, the two waves
1816: completely coalesce at the collision time, signaling the singularity
1817: formation in the ODE system; at the higher $\Delta \phi_0$ value,
1818: the right solitary wave gets higher energy at the collision time.
1819: However, after interaction, the two waves always separate, and the
1820: right wave always gets higher energy, in all three cases. Recall
1821: that before interaction, the left wave has higher energy, thus we
1822: can call these interaction outcomes transmission. In these
1823: interactions, even though the intermediate process (especially the
1824: collision segment) rather strongly depends on the initial phase
1825: difference $\Delta \phi_0$, the interaction outcome is insensitive
1826: to it. These PDE evolution results completely resemble the ODE
1827: simulations in Fig. \ref{senstive}(a). In the perturbed
1828: (non-integrable) case, the PDE simulation results are quite
1829: different from the integrable ones (as in the ODE simulations). In
1830: the perturbed case, we take the three $\Delta \phi_0$ values to be
1831: $0.92, 0.972$ and $1.01$. Notice that the first and third of these
1832: $\Delta \phi_0$ values are exactly the same as those in the ODE
1833: simulations of Fig. \ref{senstive}(b), while the middle $\Delta
1834: \phi_0$ value of $0.972$ is slightly different from the
1835: corresponding ODE value of $0.9695$. This slight difference in the
1836: middle $\Delta \phi_0$ value is necessary in order for the
1837: corresponding ODE and PDE simulations to exhibit the same behaviors,
1838: and this difference is due to the modeling error of the PDE
1839: evolutions by the ODE system (\ref{Dyreduce}). At $\Delta
1840: \phi_0=0.92$ (see Fig. \ref{senstive}(1)), the interaction outcome
1841: is similar to the integrable ones (top row of Fig. \ref{senstive})
1842: in that it is also transmission. But at $\Delta \phi_0=0.972$ (Fig.
1843: \ref{senstive}(2)), the two waves strongly coalesce, then form an
1844: oscillating bound state. At $\Delta \phi_0=1.01$, on the other hand,
1845: the two exiting waves have opposite energy distributions from Fig.
1846: \ref{senstive}(1); this interaction outcome can be called
1847: reflection. Thus, in the perturbed case, the interaction outcome is
1848: sensitive to initial conditions, which distinctively contrasts the
1849: integrable case. Again, these PDE evolution results for the
1850: perturbed case completely resemble the ODE simulations in Fig.
1851: \ref{senstive}(b).
1852:
1853: The above ODE and PDE simulations corroborate the fact that the
1854: source of this sensitive dependence on initial conditions in the
1855: perturbed system lies in the finite-time singularities of solutions
1856: in the integrable dynamical system (\ref{Dyreduce}). From the PDE
1857: point of view, the origin of this sensitive dependence can be traced
1858: to the coalescing of the two solitary waves in the integrable PDE
1859: system. At the moment, our understanding on this sensitive
1860: dependence and fractal structures in the perturbed system is still
1861: very limited. For instance, we can not yet explain any quantitative
1862: details inside these fractal structures, nor can we explain why this
1863: sensitive dependence occurs only for one-sided perturbations of the
1864: integrable system (with $\varepsilon>0$). These are non-trivial
1865: questions which merit further analysis, but they are beyond the
1866: scope of the present article.
1867:
1868: %Intuitively, this may be understood as follows. On these singularity
1869: %points, the solutions of the integrable system develop finite-time
1870: %singularities. When the system is weakly perturbed, any slight
1871: %change to initial conditions near these singularity points will
1872: %cause significant changes to the solutions due to the proximity of
1873: %these solutions to the singularity solutions. In other words,
1874: %finite-time singularities in solutions of the integrable system are
1875: %the reason for sensitive dependence on initial conditions in the
1876: %perturbed system.
1877: %From the PDE point of view, when the ODE solutions
1878: %develop finite-time singularities, $\zeta \to \infty$, thus the two
1879: %solitary waves strongly collide. The collision outcome then becomes
1880: %sensitive to initial conditions.
1881:
1882: \begin{figure}
1883: \includegraphics[width=70mm,height=55mm]{fig13a.eps}
1884: \includegraphics[width=70mm,height=55mm]{fig13b.eps}
1885: %\includegraphics[width=70mm,height=55mm]{senstive_1.eps}
1886: %\includegraphics[width=70mm,height=55mm]{senstive_2.eps}
1887: \caption{\label{senstive} ~Evolutions of $\psi$ versus $\tau$ at a
1888: singularity point $\psi_0$ and its left and right neighbor points in
1889: the dynamical system (\ref{Dyreduce}) with initial conditions
1890: (\ref{ic_nonequal}). (a) $\varepsilon=0$ (the integrable case); here
1891: $\psi_0=0.98325$ is the singularity point which is marked in Fig.
1892: \ref{figS} (bottom panel); the left and right neighbor points are
1893: taken as $\psi_0=0.92$ and 1.05; (b) $\varepsilon=0.0036$ (the
1894: positively perturbed case); here $\psi_0=0.9695$ is the singularity
1895: point; its left and right neighbor points are taken as $\psi_0=0.92$
1896: and 1.01. }
1897: \end{figure}
1898:
1899: \begin{figure}
1900: \includegraphics[width=85mm,height=40mm]{fig14a.eps}
1901:
1902: \includegraphics[width=85mm,height=40mm]{fig14b.eps}
1903: %\includegraphics[width=85mm,height=40mm]{contour_1.eps}
1904: %\includegraphics[width=85mm,height=40mm]{contour_2.eps}
1905: \caption{Evolutions of $|U(x, t)|$ in the PDE (\ref{eqn:1}) with
1906: cubic-quintic nonlinearity (\ref{cqNL}), corresponding to the ODE
1907: simulations of Fig. \ref{senstive}. Top row: the integrable (NLS)
1908: equation ($\gamma=0$); (a) $\Delta\phi_0=0.92$, (b)
1909: $\Delta\phi_0=0.98325$, (c) $\Delta\phi_0=1.05$. Bottom row: the
1910: perturbed case (with $\gamma=0.0010$); (1) $\Delta\phi_0=0.92$; (2)
1911: $\Delta\phi_0= 0.972$; (3) $\Delta\phi_0=1.01$. The values of other
1912: initial parameters are given in the text. \label{contour}}
1913: \end{figure}
1914:
1915:
1916:
1917: The fact of primary hills and fractal structures in the
1918: non-integrable system (\ref{Dyreduce}) bifurcating from singularity
1919: points of the integrable system has far reaching consequences. One
1920: important consequence is that, the main features of primary hill
1921: sequences shown in Figs. \ref{pdelevel1}(a) and \ref{ODElevel1}(a)
1922: for PDEs and ODEs can now be analytically explained. For instance,
1923: the life-time formula (\ref{time_PDE}) for primary hill sequences in
1924: the weakly perturbed system (\ref{Dyreduce}) is nothing but the
1925: analogous collision-time (singularity time) formula
1926: (\ref{collide_time}) for sequences of singularity points in the
1927: integrable system. To make a quantitative comparison between these
1928: formulas, we take the initial conditions (\ref{ic_nonequal}) which
1929: was used in the PDE and ODE simulations of Figs. \ref{pdelevel1}(a)
1930: and \ref{ODElevel1}(a). When the time rescaling (\ref{time_scale})
1931: is recovered, the collision-time formula (\ref{collide_time}) of the
1932: integrable system becomes
1933: \begin{equation} \label{time_anal}
1934: 0.0839\tilde{t}_n=2n\pi +\pi,
1935: \end{equation}
1936: which compares very favorably with the life-time formulae
1937: (\ref{time_PDE}), (\ref{PDEvalue}) and (\ref{ODEvalue}) in direct
1938: PDE and ODE simulations. The small differences in the $\omega$ and
1939: $\delta$ values between the analytical formula (\ref{collide_time})
1940: and the PDE/ODE ones (\ref{time_PDE}) are caused by the not-so-small
1941: value of $\varepsilon=0.13665$. As $\varepsilon \to 0$, these
1942: quantitative differences will vanish. Regarding the locations of
1943: individual hills in the primary-hill sequence, they are described by
1944: the formula (\ref{location_sing}) for singularity-point locations of
1945: the integrable system when $\varepsilon \ll 1$. Note that the form
1946: of this formula is different from all previous ones on window
1947: sequences in solitary-wave collisions
1948: \cite{Campbell2,Campbell3,Kivshar1,TanYang,Goodman_Haberman2}.
1949:
1950: For $\varepsilon>0$, each primary hill is paired with a sensitive
1951: (fractal) region at its foot (see Figs. \ref{ga2zero} and
1952: \ref{evolution}). Similar to primary hill sequences, the locations
1953: of these fractal regions are described by the same formula
1954: (\ref{location_sing}) in the limit $\varepsilon \to 0^+$.
1955:
1956: The fact of primary hills and fractal structures bifurcating from
1957: singularity points of the integrable system also explains major
1958: features of interaction results in Fig. \ref{fig:exPDE}(a) for the
1959: exponential nonlinearity (\ref{exp_non}). We have noticed that,
1960: unlike Fig. \ref{cqnonequal}, this graph has only one infinite
1961: sequence of primary hills accumulating toward the left (the right
1962: sequence of Fig. \ref{cqnonequal} is absent). This phenomenon is due
1963: to the fact that for the choices of initial conditions for Fig.
1964: \ref{fig:exPDE}(a), there is only one infinite sequence of
1965: singularity points in the integrable system. To see it, we first
1966: calculate the $f, g$ and $\varepsilon$ values for Fig.
1967: \ref{fig:exPDE}(a), which are found to be
1968: \begin{equation}
1969: f=228.8211, \quad g=231.91770, \quad \varepsilon=0.01353.
1970: \end{equation}
1971: Thus in the scaled dynamical equation (\ref{Dyreduce}), the initial
1972: conditions corresponding to those for Fig. \ref{fig:exPDE}(a) are
1973: \begin{equation}
1974: \zeta_0=-12.13260, \quad \dot{\zeta}_0=0, \quad
1975: \dot{\psi}_0=-0.00297.
1976: \end{equation}
1977: Notice that $\dot{\psi}_0^2 < 2e^{\zeta_0}$, thus according to the
1978: classifications of singularity points in the end of the previous
1979: section (case 2), the integrable equation (\ref{Dyreduce}) with the
1980: above initial conditions has only one sequence of singularity points
1981: in the $\psi_0$ interval, accumulating to the left toward
1982: $\psi_0=0^+$. This is in perfect agreement with the primary-hill
1983: sequence of Fig. \ref{fig:exPDE}(a) from direct PDE simulations.
1984:
1985: In many of the interaction results presented in this paper, the exit
1986: velocity versus $\psi_0$ graphs have infinite sequences of primary
1987: hills (see Figs. \ref{cqnonequal} and \ref{fig:exPDE} for instance);
1988: when zooming into the sensitive regions between primary hills, one
1989: gets infinite sequences of secondary hills. It is important to
1990: understand that these two infinite sequences are pure coincidence,
1991: and are totally un-related. Each primary hill corresponds to a
1992: particular singularity point of the integrable system, thus the
1993: number of primary hills is equal to the number of singularity points
1994: in the integrable system. This number can be either infinite or
1995: finite, depending on the choices of initial conditions. For
1996: instance, Figs. \ref{evolution}(1, 2, 3) have 1, 3 and 5 primary
1997: hills, corresponding to the same numbers of singularity points on
1998: the top three dashed lines of Fig. \ref{singularpoints}. On the
1999: other hand, at the foot of each primary hill, there is always an
2000: {\it infinite} sequence of secondary hills (when $\varepsilon>0$).
2001: In other words, secondary hills always exist as an infinite, not
2002: finite, sequence. For example, if one zooms into each of the three
2003: sensitive regions at the foot of the three primary hills in Fig.
2004: \ref{evolution}(2), one always gets an infinite sequence of
2005: secondary hills. Thus secondary-hill structures are un-related to
2006: primary-hill structures. If we zoom into the sensitive regions
2007: between secondary hills, we always get infinite sequences of
2008: tertiary hills which are very similar to the sequences of secondary
2009: hills both qualitatively and quantitatively (see Figs.
2010: \ref{fig:PDEzoom}(b, c)). This process can continue indefinitely.
2011: Thus, our conclusion is that sensitive regions between primary hills
2012: are fractal structures (in the sense that portions of these
2013: structures, when amplified, are the same as the strcutures
2014: themselves). But the whole graph with primary hills is not a
2015: fractal.
2016:
2017:
2018: \section{Applications to the generalized NLS equations with various nonlinearities \setcounter{equation}{0}}
2019: In previous sections, we have shown that for the cubic-quintic and
2020: exponential nonlinearities at selected parameters ($\alpha=1,
2021: \gamma=0.04, \beta_0=1$ for the former, and $\beta_0=2.3$ for the
2022: latter), weak interactions of solitary waves exhibit hill sequences
2023: and fractal structures for a wide range of initial conditions, and
2024: the reduced ODE model (\ref{Dyreduce}) accurately captures these
2025: interaction dynamics both qualitatively and quantitatively. In this
2026: section, we consider a larger question: for a given form of
2027: nonlinearity in the PDE (\ref{eqn:1}), can it exhibit fractal
2028: structures? For example, with the cubic-quintic nonlinearity
2029: (\ref{cqNL}), for what parameters $\alpha$ and $\gamma$ can one
2030: possibly find fractal structures? This question can be answered by
2031: applying our previous results on the ODE model (\ref{Dyreduce}). For
2032: demonstration purpose, we will do so for three forms of
2033: nonlinearity: cubic-quintic, exponential, and saturable
2034: nonlinearities.
2035:
2036: From the analysis of the ODE system (\ref{Dyreduce}) in the previous
2037: section, we have found that fractal structures in weak interactions
2038: can only occur for $\varepsilon>0$, not for $\varepsilon<0$. Thus,
2039: once we have obtained the functional dependence of $\varepsilon$ on
2040: system parameters, it will quickly become clear when fractal
2041: structures can arise. The analytical expression for $\varepsilon$ is
2042: given in Eq. (\ref{time_scale}). Notice that due to the
2043: Vakhitov-Kolokolov stability criterion \cite{Kivshar_Agrawal,VK},
2044: the solitary wave is linearly stable only when $P_\beta >0$, i.e.
2045: $\varepsilon >-1$. Below, we will use the $\varepsilon$ formula in
2046: (\ref{time_scale}) to calculate $\varepsilon$ for general system
2047: parameters in the three nonlinearities mentioned above.
2048:
2049: First we consider the cubic-quintic nonlinearity (\ref{cqNL}). When
2050: $\alpha<0$, we found that $P_\beta$ is always negative, i.e. the
2051: solitary wave is always linearly unstable. Thus we only consider the
2052: $\alpha>0$ case below. In this case, it is easy to see from Eqs.
2053: (\ref{eqn:2}) and \ref{cqNL}) that by a rescaling of variables, we
2054: can make $\alpha=\beta_0=1$. Thus the only remaining parameter for
2055: this nonlinearity is $\gamma$. Using the analytical formula
2056: (\ref{Pformula}) for $P$, we can obtain the dependence of
2057: $\varepsilon$ on $\gamma$, which is plotted in Fig.
2058: \ref{varepsilon}(1). From this graph, we see that $\varepsilon>0$
2059: when $\gamma>0$, and $\varepsilon<0$ when $\gamma < 0$. Thus
2060: \emph{fractal structures in this cubic-quintic model can appear only
2061: when $\gamma>0$, not when $\gamma<0$}. If $\gamma=0$, this
2062: cubic-quintic model reduces to the integrable NLS equation, and the
2063: dynamical equations (\ref{Dyreduce}) reduce to the integrable case
2064: (with $\varepsilon=0$) studied in Sec. 5. In this integrable case,
2065: there is of course no fractal dependence in solitary wave
2066: interactions.
2067:
2068:
2069: \begin{figure*}
2070: \includegraphics[width=55mm,height=45mm]{fig15a.eps}
2071: \includegraphics[width=55mm,height=45mm]{fig15b.eps}
2072: \includegraphics[width=55mm,height=45mm]{fig15c.eps}
2073: %\includegraphics[width=55mm,height=45mm]{epsivsgamma-cq-2.eps}
2074: %\includegraphics[width=55mm,height=45mm]{betavsepsi-exp-2.eps}
2075: %\includegraphics[width=55mm,height=45mm]{betavsepsi-satu-2.eps}
2076: \caption{\label{varepsilon}~ Graphs of $\varepsilon$ versus system
2077: parameters for three different nonlinearities: (1) $\varepsilon$
2078: verse $\gamma$ for the cubic-quintic nonlinearity (\ref{cqNL}); (2)
2079: $\varepsilon$ verse $\beta$ for the exponential nonlinearity
2080: (\ref{exp_non}); (3) $\varepsilon$ verse $\beta$ for the saturable
2081: nonlinearity (\ref{sat_non}).}
2082: \end{figure*}
2083:
2084: Next, we consider the exponential nonlinearity (\ref{exp_non}). In
2085: this case, the solitary wave depends only on the propagation
2086: constant $\beta$, thus $\varepsilon$ depends only on $\beta$ as
2087: well. The analytical expression for function $\varepsilon(\beta)$ is
2088: not available, but this function can be easily determined by
2089: numerical methods, and its graph is plotted in
2090: Fig.\ref{varepsilon}(b). It is seen that this graph has two critical
2091: propagation constants, $\bar{\beta}_a=2.2457$ where $\varepsilon=0$,
2092: and $\bar{\beta}_b=14.0051$ where $\varepsilon=+\infty$. When
2093: $\beta< \bar{\beta}_a$, $\varepsilon < 0$, thus fractal structures
2094: do not exist; when $\bar{\beta}_a < \beta <\bar{\beta}_b$, thus
2095: fractal structures can appear (see previous sections); when
2096: $\beta>\bar{\beta}_b$, $\varepsilon < -1$, thus the solitary wave is
2097: linearly unstable.
2098:
2099: Next, we consider the saturable nonlinearity,
2100: \begin{equation} \label{sat_non}
2101: F(|U|^2)=1-\frac{1}{1+|U|^2},
2102: \end{equation}
2103: which is common in optics (for instance, in photorefractive crystals
2104: \cite{Segev_PRE}). Here one is added in the above formula to make
2105: $F(0)=0$ (this does not affect the solitary waves and their
2106: interaction dynamics). In this case, $\varepsilon$ also depends only
2107: on the propagation constant $\beta$. This dependence is computed
2108: numerically and plotted in Fig.\ref{varepsilon}(c). We find that
2109: $\varepsilon$ is negative for all values of $\beta$, thus fractal
2110: structures can not exist in weak interactions of solitary waves for
2111: this saturable nonlinearity. This conclusion is consistent with our
2112: earlier results for the cubic-quintic nonlinearity, as the saturable
2113: nonlinearity (\ref{sat_non}) resembles the cubic-quintic
2114: nonlinearity (\ref{cqNL}) with $\alpha >0$ and $\gamma<0$. It is
2115: noted, however, that for the saturable nonlinearity, weak
2116: interactions of solitary waves can still exhibit some interesting
2117: structures as shown in Figs. \ref{ga2zero}(5, 6), but these
2118: structures are not fractal structures.
2119:
2120: From the above three examples (as well as the previous section), we
2121: see that the reduced ODE model (\ref{Dyreduce}) enables us to
2122: accurately predict when and where fractal structures and hill
2123: sequences appear in the space of initial parameters of solitary
2124: waves. Based on this reduced model, a global and universal
2125: understanding on weak interactions of solitary waves has been
2126: achieved for the generalized NLS equations (\ref{eqn:1}) with
2127: arbitrary forms of nonlinearity.
2128:
2129: \section{conclusion and discussion}
2130:
2131: In this paper, we have analyzed weak interactions of solitary waves
2132: in the generalized nonlinear Schr\"{o}dinger equations with general
2133: forms of nonlinearity. We have shown that these interactions exhibit
2134: similar fractal dependence on initial conditions for different
2135: nonlinearities. To analytically explain these universal fractal
2136: structures, we derived a set of fourth-order dynamical equations for
2137: the solitary-wave parameters using asymptotic methods. A remarkable
2138: feature of these dynamical equations is that they contain only one
2139: parameter, which is dependent on the specific form of nonlinearity.
2140: When this parameter is zero, these dynamical equations are
2141: integrable, and the exact analytical solutions are derived. When
2142: this parameter is non-zero, the dynamical equations exhibit fractal
2143: structures which match those in the original PDEs both qualitatively
2144: and quantitatively. We have also investigated the origin of these
2145: fractal structures, and found that they bifurcate from the
2146: singularity points (i.e. initial conditions for singularity
2147: solutions) in the integrable system. Based on this observation, an
2148: analytical criterion for the existence and locations of fractal
2149: structures is obtained. Lastly, we applied these analytical results
2150: to the generalized nonlinear Schr\"{o}dinger equations with various
2151: nonlinearities such as the cubic-quintic, exponential and saturable
2152: nonlinearities, and predictions on their weak interactions of
2153: solitary waves are presented.
2154:
2155: Regarding the bifurcation of fractal structures from the integrable
2156: dynamical equations, even though we have established that this
2157: bifurcation occurs at the singularity points of the integrable
2158: system, more challenging questions are to comprehensively analyze
2159: how this bifurcation takes place, and to quantitatively predict the
2160: detailed geometric structures inside these fractals. This has not
2161: been done yet. Recently, Goodman and Haberman analyzed the
2162: approximate ODE models for the collisions of solitary waves in three
2163: physical systems where window sequences and fractal structures have
2164: been reported \cite{Goodman_Haberman1, Goodman_Haberman2,
2165: Goodman_Haberman3}. They found that the origin of window sequences
2166: and fractal structures in these systems lies in the crossing of the
2167: separatrix (homoclinic orbit). Analytical predictions on the
2168: locations of window sequences in the ODE models were derived as
2169: well. It is not clear at the moment whether similar analysis can be
2170: performed for our system (\ref{Dyreduce}). This question is beyond
2171: the scope of the present article, and will be left for future
2172: studies.
2173:
2174:
2175:
2176: \begin{acknowledgments}
2177: The authors thank Drs. Richard Haberman, Roy Goodman and Meirong
2178: Zhang for valuable discussions. This work was supported in part by
2179: the Air Force Office of Scientific Research under grant USAF
2180: 9550-05-1-0379.
2181: \end{acknowledgments}
2182: %\bibliography{apssamp}% Produces the bibliography via BibTeX.
2183:
2184: \begin{thebibliography}{200}
2185:
2186: \bibitem{Ablowitz_Segur}
2187: M. J. Ablowitz and H. Segur, {\it Solitons and the Inverse
2188: Scattering Transform}, SIAM, Philadelphia, 1981.
2189:
2190: \bibitem{Hasegawa_Kodama} A. Hasegawa and Y. Kodama, \emph{Solitons in Optical
2191: Communications}, Clarendon, Oxford, 1995.
2192:
2193: \bibitem{Krolikowski} W. Krolikowski, B. Luther-Davies, C. Denz, and T.
2194: Tschudi, "Annihilation of photorefractive solitons", Opt. Lett.
2195: 23, 97, 1998.
2196:
2197: \bibitem{Anastassiou} C. Anastassiou, M. Segev, K. Steiglitz, J. A.
2198: Giordmaine, M. Mitchell, M. Shih, S. Lan, and J. Martin,
2199: "Energy-Exchange Interactions between Colliding Vector Solitons",
2200: Phys. Rev. Lett. 83, 2332, 1999.
2201:
2202: \bibitem{Chen_Kivshar} Z. Chen, M. Acks, E.A. Ostrovskaya and Y.S. Kivshar,
2203: "Observation of bound states of interacting vector solitons", Opt.
2204: Lett. 25, 417 (2000).
2205:
2206: \bibitem{Kivshar_Agrawal} Y.S. Kivshar and G. P. Agrawal, \emph{Optical Solitons:
2207: From Fibers to Photonic Crystals}, Academic Press, San Diego, 2003.
2208:
2209: \bibitem{Aceves_stud} T. Dohnal and A.B. Aceves,
2210: "Optical soliton bullets in (2+1)D nonlinear Bragg resonant periodic
2211: geometries", Stud. Appl. Math. 115, 209-232 (2005).
2212:
2213: \bibitem{Yukon} S.P. Yukon and N.C.H. Lin, "Coupled long-Josephson junctions
2214: and the N sine-Gordon equations", IEEE Trans. Magn. 27, 2736 (1991).
2215:
2216: \bibitem{Karpman_Solovev} V. I. Karpman and V. V. Solov'ev,
2217: "A perturbation theory for soliton systems", Physica D 3, 142-164
2218: (1981).
2219:
2220: \bibitem{Gerdjikov_PRL} V.S. Gerdjikov, D.J. Kaup, I.M. Uzunov, and E.G. Evstatiev,
2221: "Asymptotic Behavior of N-Soliton Trains of the Nonlinear
2222: Schr\"{o}inger Equation", Phys. Rev. Lett. 77, 3943 (1996).
2223:
2224: \bibitem{Gerdjikov_Yang} V. S. Gerdjikov, E. V. Doktorov and J. Yang, "Adiabatic
2225: Interaction of N-Ultrashort Solitons: Universality of the Complex
2226: Toda Chain Model." Phys. Rev. E. 64, 056617 (2001).
2227:
2228: \bibitem{Yang_Manakov} J. Yang, "Suppression of Manakov-Soliton Interference in Optical Fibers."
2229: Phys. Rev. E. 65, 036606 (2002).
2230:
2231: \bibitem{Ablowitz_Kruskal} M.J. Ablowitz, M.D. Kruskal and J.F.
2232: Ladik, "Solitary Wave Collisions", SIAM J. Appl. Math. 36, 428
2233: (1979).
2234:
2235: \bibitem{Campbell1} D. K. Campbell, J. S. Schonfeld, and C. A.
2236: Wingate, "Resonant structure in kink-antikink interactions in
2237: $\phi^4$ theory", Physica D 9, 1-32 (1983).
2238:
2239: \bibitem{Campbell2} M. Peyrard and D. K. Campbell,
2240: "Kink-antikink interactions in a modified sine-Gordon model",
2241: Physica D 9, 33-51 (1983).
2242:
2243: \bibitem{Campbell3} D.K. Campbell and M. Peyrard,
2244: "Solitary wave collisions revisited", Physica D 18, 47-53 (1986).
2245:
2246: \bibitem{Campbell4}
2247: D.K. Campbell, M. Peyrard and P. Sodano, "Kink-antikink
2248: interactions in the double sine-Gordon equation", Physica D 19,
2249: 165 - 205 (1986).
2250:
2251: \bibitem{anninos} P. Anninos, S. Oliveira, and R. A. Matzner,
2252: "Fractal structure in the scalar $\lambda(\phi^2-1)^2$ theory",
2253: Phys. Rev. D 44, 1147-1160, 1991.
2254:
2255: \bibitem{Kudryavtsev} T. I. Belova, A. E. Kudryavtsev,
2256: "Solitons and their interactions in classical field theory",
2257: Physics-Uspekhi, 40, 359-386 (1997).
2258:
2259:
2260: \bibitem{Kivshar1} Y. S. Kivshar, Z. Fei, and L. V\'{a}zquez,
2261: "Resonant soliton-impurity interactions", Phys. Rev. Lett. 67,
2262: 1177-1180, 1991.
2263:
2264: \bibitem{Kivshar2} Z. Fei, Y. S. Kivshar, and L. V\'{a}quez,
2265: "Resonant kink-impurity interactions in the sine-Gordon model",
2266: Phys. Rev. A 45, 6019-6030, 1992.
2267:
2268: \bibitem{Kivshar3} Z. Fei, Y. S. Kivshar, and L. V\'{a}quez,
2269: "Resonant kink-impurity interactions in the $\phi^4$ model", Phys.
2270: Rev. A 46, 5214-5220, 1992.
2271:
2272: \bibitem{YangTan} J. Yang and Y. Tan,
2273: "Fractal structure in the collision of vector solitons", Phys. Rev.
2274: Lett. 85, 3624-3627, 2000.
2275:
2276: \bibitem{YangTan2} J. Yang and Y. Tan,
2277: ``Fractal dependence of vector-soliton collisions in birefringent fibers.''
2278: Phys. Lett. A. 280, 129 (2001).
2279:
2280: \bibitem{TanYang} Y. Tan and J. Yang, "Complexity and regularity of
2281: vector-soliton collisions." Phys. Rev. E. 64, 056616 (2001).
2282:
2283: \bibitem{Dmitriev_Kivshar} S.V. Dmitriev, Yu.S. Kivshar, and T. Shigenari,
2284: "Fractal structures and multiparticle effects in soliton
2285: scattering", Phys. Rev. E 64, 056613, 2001.
2286:
2287: \bibitem{Dmitriev} S.V. Dmitriev and T. Shigenari,
2288: "Short-lived two-soliton bound states in weakly perturbed nonlinear
2289: Schr\"odinger equation", Chaos 12, 324 (2002).
2290:
2291: \bibitem{Goodman_Haberman1}
2292: R. H. Goodman and R. Haberman, "Interaction of sine-Gordon kinks
2293: with defects: The two-bounce resonance", Phys. D. 195, pp. 303--323
2294: (2004).
2295:
2296: \bibitem{Goodman_Haberman2}
2297: R. H. Goodman and R. Haberman, "Vector soliton interactions in
2298: birefringent optical fibers", Phys. Rev. E 71, pp. 056605 (2005).
2299:
2300: \bibitem{Goodman_Haberman3}
2301: R. H. Goodman and R. Haberman, "Kink-antikink collisions in the
2302: phi-four equation: The n-bounce resonance and the separatrix map",
2303: SIAM J. Appl. Dyn. Sys. 4, pp. 1195-1228 (2005).
2304:
2305: \bibitem{Ueda_Kath} T. Ueda and W. L. Kath,
2306: "Dynamics of coupled solitons in nonlinear optical fibers", Phys.
2307: Rev. A 42, 563-571, 1990.
2308:
2309: \bibitem{Kovalev} A.S. Kovalev and A.M. Kosevich,
2310: Fiz. Nizk. Temp. 2, 913 (1976).
2311:
2312: \bibitem{VK} N. G. Vakhitov and A. A. Kolokolov, "Stationary solutions of the
2313: wave equation in the medium with nonlinearity saturation," Izv.
2314: Vyssh. Uchebn. Zaved. Radiofiz. {\bf 16}, 1020 (1973) [Radiophys.
2315: Quantum Electron. {\bf 16}, 783 (1973)].
2316:
2317: \bibitem{Barashenkov} I.V. Barashenkov, V.G. Makhankov,
2318: "Soliton-like bubbles in a system of interacting bosons", Phys.
2319: Lett. A 128, 52-56 (1988).
2320:
2321: \bibitem{Gorshkov} K.A. Gorshkov and L.A. Ostrovsky,
2322: "Interactions of solitons in nonintegrable systems: direct
2323: perturbation method and applications", Physica D 3, 428-438 (1981).
2324:
2325: \bibitem{YangPRE01} J. Yang, "Interactions of vector solitons." Phys. Rev. E 64,
2326: 026607 (2001)
2327:
2328: \bibitem{YangKaup} J. Yang, and D.J. Kaup,
2329: "Stability and evolution of solitary waves in perturbed generalized
2330: nonlinear Schroedinger equations." SIAM J. Appl. Math. 60, 967-989
2331: (2000).
2332:
2333: \bibitem{Segev_PRE}
2334: N. K. Efremidis, S. Sears, D. N. Christodoulides, J. W. Fleischer,
2335: and M. Segev, "Discrete solitons in photorefractive optically
2336: induced photonic lattices", Phys. Rev. E 66, 046602 (2002).
2337:
2338: \end{thebibliography}
2339:
2340: \end{document}
2341: