1: \documentclass[runningheads]{llncs}
2:
3: \RequirePackage[english]{babel}
4: \usepackage{graphicx,epsfig}
5:
6: \pagestyle{headings}
7:
8: \begin{document}
9:
10: \title{Nodal two-dimensional solitons in
11: nonlinear parametric resonance}
12:
13: \author{N.V. Alexeeva\inst{1} \and
14: E.V. Zemlyanaya\inst{2} }
15:
16: \institute{Department of Mathematics, University of Cape Town,
17: South Africa; \email{nora@figaro.mrh.uct.ac.za} \and Joint
18: Institute for Nuclear Research, Dubna 141980, Russia;
19: \email{elena@jinr.ru} }
20:
21:
22: \titlerunning{Nodal two-dimensional solitons in nonlinear parametric resonance}
23: \toctitle{Nodal two-dimensional solitons in nonlinear parametric
24: resonance}
25:
26: \authorrunning{N.V. Alexeeva and E.V. Zemlyanaya}
27: \tocauthor{N.V. Alexeeva and E.V. Zemlyanaya}
28:
29: \maketitle
30:
31: \begin{abstract}
32: The parametrically driven damped nonlinear Schr\"odinger equation
33: serves as an amplitude equation for a variety of resonantly forced
34: oscillatory systems on the plane. In this note, we consider its
35: nodal soliton solutions. We show that although the nodal solitons
36: are stable against radially-symmetric perturbations for
37: sufficiently large damping coefficients, they are always unstable
38: to azimuthal perturbations. The corresponding break-up scenarios
39: are studied using direct numerical simulations. Typically, the
40: nodal solutions break into symmetric ``necklaces" of stable
41: nodeless solitons.
42:
43: \end{abstract}
44:
45: {\bf 1.} Two-dimensional localised oscillating structures,
46: commonly referred to as oscillons, have been detected in
47: experiments on vertically vibrated layers of granular material
48: \cite{Swinney}, Newtonian fluids and suspensions
49: \cite{Faraday,Astruc}. Numerical simulations established the
50: existence of stable oscillons in a variety of pattern-forming
51: systems, including the Swift-Hohenberg and Ginsburg-Landau
52: equations, period-doubling maps with continuous spatial coupling,
53: semicontinuum theories and hydrodynamic models
54: \cite{Astruc,numerical}. These simulations provided a great deal
55: of insight into the phenomenology of the oscillons; however, the
56: mechanism by which they acquire or loose their stability remained
57: poorly understood.
58:
59: In order to elucidate this mechanism, a simple model of a
60: parametrically forced oscillatory medium was proposed recently
61: \cite{us}. The model comprises a two-dimensional lattice of
62: diffusively coupled, vertically vibrated pendula. When driven at
63: the frequency close to their double natural frequency, the pendula
64: execute almost synchronous librations whose slowly varying
65: amplitude satisfies the 2D parametrically driven, damped nonlinear
66: Schr\"odinger (NLS) equation. The NLS equation was shown to
67: support radially-symmetric, bell-shaped (i.e. nodeless) solitons
68: which turned out to be stable for sufficiently large values of the
69: damping coefficient. These stationary solitons of the amplitude
70: equation correspond to the spatio-temporal envelopes of
71: the oscillons in the original lattice system. By reducing the NLS
72: to a finite-dimensional system in the vicinity of the soliton, its
73: stabilisation mechanism (and hence, the oscillon's stabilisation
74: mechanism) was clarified \cite{us}.
75:
76: In the present note we consider a more general class of
77: radially-symmetric solitons of the parametrically driven, damped
78: NLS on the plane, namely solitons with nodes. We will demonstrate
79: that these solitons are unstable against azimuthal modes, and
80: analyse the evolution of this instability.
81:
82:
83:
84: {\bf 2.} The parametrically driven, damped NLS equation has the
85: form:
86: \begin{equation}
87: i \psi_t + \nabla^2 \psi+ 2|\psi|^2 \psi - \psi = h \psi^* -i
88: \gamma \psi. \label{2Dnls}
89: \end{equation}
90: Here $\nabla^2=\partial^2 /\partial x^2 +\partial^2 /\partial
91: y^2.$ Eq.(\ref{2Dnls}) serves as an amplitude equation for a wide
92: range of nearly-conservative two-dimensional oscillatory systems
93: under parametric forcing. This equation was also used as a
94: phenomenological model of nonlinear Faraday resonance in water
95: \cite{Astruc}. The coefficient $h>0$ plays the role of the
96: driver's strength and $\gamma>0$ is the damping coefficient.
97:
98: \begin{figure}[t]
99: \begin{center}
100: %%%\psfig{file=fff.eps,width=8cm,height=5.cm}
101: \psfig{file=fig0.ps,width=8cm,height=5.cm}
102: \end{center}
103: \caption{\sf Solutions of eq.(\ref{master}):
104: ${\cal
105: R}_0(r)$ (thin continuous line), ${\cal R}_1(r)$ (thick line),
106: ${\cal R}_2(r)$ (dashed). } \label{fig0}
107: \end{figure}
108:
109: We start with the discussion of its nodeless solitons and their
110: stability. The exact (though not explicit) stationary
111: radially-symmetric solution is given by
112: \begin{equation}
113: \psi_0= {\cal A} e^{- i \theta} \, {\cal R}_0({\cal A} r),
114: \label{soliton}
115: \end{equation}
116: where $r^2=x^2+y^2$,
117: $${\cal A}^2=1 + \sqrt{h^2-\gamma^2}, \quad
118: \theta= \frac12 \arcsin\left(\frac{\gamma}{h}\right),$$ and ${\cal
119: R}_0(r)$ is the bell-shaped nodeless solution of the equation
120: \begin{equation}
121: {\cal R}_{rr} +\frac 1 r {\cal R} - {\cal R} + 2 {\cal R}^3 =0
122: \label{master}
123: \end{equation}
124: with the boundary conditions ${\cal R}_r(0)={\cal R}(\infty)=0$.
125: (Below we simply write ${\cal R}$ for ${\cal R}_0$.)
126: %Here $ \nabla_r^2=\partial_r^2+ \frac{D-1}{r} \partial_r$.
127: Solutions of eq.(\ref{master}) are well documented in literature
128: \cite{Rypdal}; see Fig.\ref{fig0}.
129:
130: \begin{figure}[t]
131: \begin{center}
132: \psfig{file=spr_f1.ps,width=8cm,height=5.cm}
133: %previous file nn.ps
134: \end{center}
135: \caption{\sf Stability diagram for two-dimensional solitons. The
136: $(\gamma, h-\gamma)$-plane is used for visual clarity. No
137: localised or periodic attractors exist for $h< \gamma$ (below the
138: horisontal axis). The region of stability of the soliton
139: $\psi_{0}$ lies to the right of the
140: solid curve marked ``$n=0$". Also shown are the regions of stability of
141: the solitons $\psi_1$ and $\psi_2$ with respect to the radially-symmetric
142: perturbations. (These lie to the right of the corresponding curves in the
143: figure.)
144: } \label{chart}
145: \end{figure}
146:
147: {\bf 3.} To examine the stability of the solution (\ref{soliton})
148: with nonzero $h$ and $\gamma$, we
149: linearise eq.(\ref{2Dnls}) in the small perturbation
150: $$
151: \delta \psi({\bf x},t)= e^{(\mu-\Gamma) {\tilde t}-i
152: \theta_{\pm}}[u({\tilde {\bf x}})+i v({\tilde {\bf x}})],
153: %\label{pert}
154: $$
155: where ${\tilde {\bf x}}={\cal A} {\bf x}$, ${\tilde t}={\cal A}^2
156: t$. This yields an eigenvalue problem
157: \begin{equation}
158: L_1 u =- (\mu+ \Gamma )v, \quad
159: (L_0 - \epsilon)v = (\mu- \Gamma )u, \label{EV_gamma}
160: \end{equation}
161: where $\Gamma = \gamma /{\cal A}^2$ and the operators
162: \begin{equation}
163: L_0 \equiv - {\tilde \nabla}^2+1 -2 {\cal R}^2({\tilde r}), \quad
164: L_1 \equiv L_0 -4 {\cal R}^2({\tilde r}), \label{op_nabla}
165: \end{equation}
166: with ${\tilde \nabla}^2= \partial^2/\partial {\tilde
167: x}^2+\partial^2/\partial {\tilde y}^2$. (We are dropping the
168: tildas below.) For further convenience, we introduce the positive
169: quantity $\epsilon= 2 \sqrt{h^2-\gamma^2}/{\cal A}^2$.
170: Fixing $\epsilon$ defines a curve
171: on the $(\gamma,h)$-plane:
172: \begin{equation}
173: h= \sqrt{ \epsilon^2/(2- \epsilon)^2 + \gamma^2}.
174: \label{h_of_gamma}
175: \end{equation}
176: Introducing
177: \begin{equation}
178: \lambda^2=\mu^2-\Gamma^2 \label{mugamma}
179: \end{equation}
180: and performing the transformation \cite{BBK} $$v({\bf x}) \to (\mu
181: +\Gamma) \lambda^{-1} v({\bf x}),$$ reduces eq.(\ref{EV_gamma})
182: to a {\it one-}parameter eigenvalue problem:
183: \begin{equation}
184: (L_0-\epsilon) v= \lambda u, \quad L_1 u= -\lambda v. \label{EV}
185: \end{equation}
186:
187: We first consider the stability with respect to radially symmetric
188: perturbations $u=u(r), \, v=v(r)$. In this case the operators
189: (\ref{op_nabla}) become
190: \begin{equation}
191: L_0 =- \frac{d^2}{dr^2} -\frac{1}{r} \frac{d}{dr}+1 -2 {\cal
192: R}^2(r), \quad L_1 = L_0 -4 {\cal R}^2(r). \label{op_radsym}
193: \end{equation}
194: In the absence of the damping and driving, all localised initial
195: conditions in the unperturbed 2D NLS equation are known to either
196: disperse or blow-up in finite time \cite{Rypdal,Malkin,collapse}.
197: It turned out, however, that the soliton $\psi_0$ stabilises as
198: the damping $\gamma$ is increased above a certain value \cite{us}.
199: The stability condition
200: is
201: $\gamma \ge \gamma_c$, where
202: \begin{equation}
203: \gamma_c=\gamma_c(\epsilon) \equiv \frac{2}{2-\epsilon} \cdot
204: \frac{{\rm Re\/} \lambda(\epsilon)
205: \, {\rm Im\/} \lambda(\epsilon)}
206: {\sqrt{({\rm Im\/} \lambda)^2- ({\rm Re\/} \lambda)^2}}.
207: \label{gce}
208: \end{equation}
209: We obtained $\lambda(\epsilon)$ by solving the eigenvalue problem
210: (\ref{EV}) directly. Expressing $\epsilon$ via $\gamma_c$ from
211: (\ref{gce}) and feeding into (\ref{h_of_gamma}), we get the
212: stability boundary on the $(\gamma,h)$-plane (Fig.\ref{chart}).
213:
214: {\bf 4.} To study the stability to asymmetric perturbations we
215: factorise, in (\ref{EV}),
216: $$u({\bf x})={\tilde u}(r) e^{i m \varphi}, \quad v({\bf x})={\tilde v}(r)
217: e^{i m \varphi},$$ where $\tan \varphi= y/x$ and $m$ is an
218: integer. The functions ${\tilde u}(r)$ and ${\tilde v}(r)$ satisfy
219: the eigenproblem (\ref{EV}) where the operators (\ref{op_radsym})
220: should be replaced by
221: \begin{equation}
222: L_0^{(m)}
223: \equiv L_0
224: +{m^2}/{r^2},
225: \quad L_1^{(m)} \equiv L_1+ {m^2}/{r^2}, \label{m2r2}
226: \end{equation}
227: respectively. This modified eigenvalue problem can be analysed in
228: a similar way to eqs.(\ref{EV}). It is not difficult to
229: demonstrate that all discrete eigenvalues of (\ref{EV}) (if any
230: exist) have to be pure imaginary in this case, and hence the
231: azimuthal perturbations do not lead to any instabilities of the
232: solution in question \cite{us}.
233:
234: \begin{figure}[t]
235: \begin{center}
236: \psfig{file=f3a.ps,width=6cm,height=4.7cm}\\
237: \psfig{file=f3b.ps,width=6cm,height=4.7cm}
238: \psfig{file=f3c.ps,width=6cm,height=4.7cm}
239: \end{center}
240: \caption{\sf The discrete eigenvalues of the linearised operator
241: (\ref{EV}) for the one-node soliton, $\psi_1$. Panels (a) and (b)
242: show the {\it complex\/} eigenvalues, $\mbox{Im} \, \lambda$ vs
243: $\mbox{Re} \, \lambda$. Arrows indicate the direction of increase
244: of $\epsilon$.
245: Panel (c) shows the {\it real\/} eigenvalues,
246: as functions of $\epsilon$.} \label{fig2}
247: \end{figure}
248:
249:
250: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
251: \begin{figure}[t]
252: \begin{center}
253: \psfig{file=f4a.ps,width=6cm,height=4.7cm}
254: \psfig{file=f4b.ps,width=6cm,height=4.7cm}\\
255: \psfig{file=f4c.ps,width=6cm,height=4.7cm}
256: \psfig{file=f4d.ps,width=6cm,height=4.7cm}\end{center}
257: \caption{\sf The discrete eigenvalues of the linearised operator
258: (\ref{EV}) for the two-node soliton, $\psi_2$. Panels (a),(b), and
259: (c) show the complex eigenvalues, $\mbox{Im} \, \lambda$ vs
260: $\mbox{Re} \, \lambda$. Panel (d) shows the real eigenvalues, as
261: functions of $\epsilon$.} \label{fig3}
262: \end{figure}
263:
264:
265: \begin{figure}[t]
266: \begin{center}
267: \psfig{file=fig5n.eps}
268: %\psfig{file=fig5a.ps,width=6cm,height=4.cm}
269: %\psfig{file=fig5b.ps,width=6cm,height=4.cm}\\
270: %\psfig{file=fig5c.ps,width=6cm,height=4.cm}
271: %\psfig{file=fig5d.ps,width=6cm,height=4.cm}
272: \end{center}
273: \caption{\sf Evolution of the azimuthal instability of the
274: one-node soliton. (a): the initial condition, soliton $\psi_1$ ;
275: (b) and (c):
276: dissociation of the ring-like ``valley'' into 4 nodeless solitons;
277: (d): divergence of the fragments.
278: Here $\gamma=3.5$ and $h=3.6$; shown is $\mbox{Re} \, \psi$.
279: Note the change of the
280: vertical scale in (d).} \label{fig4}
281: \end{figure}
282:
283: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
284:
285: \begin{figure}
286: \begin{center}
287: \psfig{file=fig6n.eps}
288: %\psfig{file=fig6a.ps,width=6cm,height=4.cm}
289: %\psfig{file=fig6b.ps,width=6cm,height=4.cm}\\
290: %\psfig{file=fig6c.ps,width=6cm,height=4.cm}
291: %\psfig{file=fig6d.ps,width=6cm,height=4.cm}\\
292: %\psfig{file=fig6e.ps,width=6cm,height=4.cm}
293: %\psfig{file=fig6f.ps,width=6cm,height=4.cm}
294: \end{center}
295: \caption{\sf The evolution of the azimuthal instability of the
296: two-node soliton $\psi_2$. (a): the initial condition; (b)-(c):
297: the rapid dissociation of the ``valley" into 4 nodeless solitons;
298: (c)-(d): a slower decay of the ``ridge" into 8 solitons $\psi_0$;
299: (e)-(f): the annihilation of the internal ring and the central
300: soliton, and the repulsion of the persisting 8 solitons. Here
301: $\gamma=4.5$ and $h=4.53$; shown is $\mbox{Re} \, \psi$. Note the
302: change of the vertical scale in (e)-(f) w.r.t. that in (a)-(d).}
303: \label{fig5}
304: \end{figure}
305:
306: {\bf 5.} Besides the nodeless solution ${\cal R}_0(r)$, the
307: ``master" equation (\ref{master}) has solutions ${\cal R}_n(r)$
308: with $n$ nodes, $n \ge 1$. (See Fig.\ref{fig0}). These give rise
309: to a sequence of nodal solutions $\psi_n$ of the damped-driven NLS
310: (\ref{2Dnls}), defined by eq.(\ref{soliton}) with ${\cal R}_0 \to
311: {\cal R}_n$. To examine the stability of the
312: $\psi_{n}$, % solution,
313: we solved the eigenvalue problem (\ref{EV}) numerically,
314: with % the
315: operators $L_{0,1}^{(m)}$ as in
316: (\ref{m2r2}). The {\it radial\/} stability properties of the nodal
317: solitons turned out to be similar to those of the nodeless soliton
318: $\psi_0$. Namely, the $\psi_n$ solutions are stable against
319: radially-symmetric perturbations for sufficiently large $\gamma$.
320: The corresponding stability regions for $\psi_1$ and $\psi_2$ are
321: depicted in Fig.(\ref{chart}). However, the {\it azimuthal\/}
322: stability properties of the nodal solitons have turned out to be
323: quite different.
324:
325: Both $\psi_1$ and $\psi_2$ solutions do have eigenvalues
326: $\lambda$ with nonzero real parts for
327: orbital numbers $m \ge 1$.
328: (See Fig.\ref{fig2} and \ref{fig3}.) Having found eigenvalues
329: $\lambda$ for each $\epsilon$, one still has to identify those
330: giving rise to the largest growth rates
331: \begin{equation}
332: \nu= \mbox{Re} \, \mu -\Gamma \label{nu}
333: \end{equation}
334: for each pair $(\epsilon, \gamma)$ [or, equivalently, for each
335: $(h,\gamma)$]. In (\ref{nu}), $\mu$ is reconstructed using
336: eq.(\ref{mugamma}). The selection of {\it real\/} eigenvalues is
337: straightforward; in this case we
338: have the following two simple rules:\\
339: $\bullet$ If, for some $\epsilon$, there are eigenvalues
340: $\lambda_1>0$, $\lambda_2>0$ such that $\lambda_1>\lambda_2$, then
341: $\nu_1>\nu_2$ for this $\epsilon$ and all $\gamma$. That is, of
342: all real eigenvalues $\lambda$ one has to
343: consider only the largest one.\\
344: $\bullet$ If, for some $\epsilon$, there is a real eigenvalue
345: $\lambda_1>0$ and a complex eigenvalue $\lambda_2$, with
346: $\mbox{Re}\, \lambda_2>0$ and $\lambda_1 > \mbox{Re} \,
347: \lambda_2$, then $\nu_1>\nu_2$ for this $\epsilon$ and all
348: $\gamma$. That is, one can ignore all complex eigenvalues with
349: real parts smaller than a real eigenvalue --- if there is one.
350:
351: The comparison of two complex eigenvalues is not so
352: straightforward. In particular, the fact that $\mbox{Re} \,
353: \lambda_1
354: > \mbox{Re} \, \lambda_2$ does not necessarily imply that
355: $\nu_1>\nu_2$. Which of the two growth rates, $\nu_1$ or $\nu_2$,
356: is larger will depend on the imaginary parts of $\lambda$, as well
357: as on $\gamma$.
358:
359: In figures \ref{fig2} and \ref{fig3}, we illustrate the real and
360: imaginary parts of the eigenvalues, arising for different $m$, for
361: the solitons $\psi_1$ and $\psi_2$. The soliton $\psi_1$ has
362: discrete eigenvalues $\lambda$ associated with orbital numbers
363: $m=0,1,...,5$ and the soliton $\psi_2$ with
364: $m=0,1,...,10$.
365:
366: \begin{table}
367: \begin{center}
368: %\begin{tabular}{cc}
369: %$\qquad\qquad\qquad n=1\qquad\qquad\qquad $ & $\qquad\qquad\qquad
370: %n=2\qquad\qquad\qquad $ \\
371: %\end{tabular}
372: \begin{tabular}{|c|c|c|c|c|c|c|c|c|} \cline{1-4} \cline{6-9}
373: m & $\nu$ & Re$\lambda$ & Im$\lambda$&$\qquad$& m & $\nu$ & Re$\lambda$ & Im$\lambda$ \\
374: \cline{1-4} \cline{6-9}
375: 0 & -0.1620 & 1.5797 & 4.2181 && 0 & -0.3361 & 2.3531 & 6.1585 \\
376: 0 & -1.4827 & 0.0609 & 1.8743 && 0 & -0.5877 & 0.1819 &
377: 1.7847 \\
378: & & & && 0 & -0.4168 & 0.3093 & 1.5572 \\
379: \cline{1-4} \cline{6-9}
380: 1 & -0.8255 & 0.2272 & 1.6198 && 1 & -0.8089 & 0.3818 & 2.1021 \\
381: 1 & 2.79e-6 & 0.0033 & 0.0000 && 1 & -0.4891 & 0.1111 & 1.6352 \\
382: & & & & & 1 & 1.07e-5 & 0.0079 & 0.0000 \\
383: \cline{1-4} \cline{6-9}
384: 2 & -0.3012 & 0.2213 & 1.0602 && 2 & -0.3497 & 0.3737 & 1.4597 \\
385: & & & & & 2 & -0.0328 & 0.2602 & 0.5128 \\
386: \cline{1-4} \cline{6-9}
387: 3 & 0.0872 & 0.5821 & 0.0000 && 3 & 0.5406 & 1.8686 & 0.0000 \\
388: \cline{1-4} \cline{6-9}
389: 4 & 0.3689 & 1.2399 & 0.0000 && 4 & 0.5286 & 1.8462 & 0.0000 \\
390: \cline{1-4} \cline{6-9}
391: 5 & 0.2057 & 0.9076 & 0.0000 && 5 & 0.0263 & 0.3958 & 0.0000 \\
392: \cline{1-4} \cline{6-9}
393: &&&&& 6 & 0.1611 & 0.9898 & 0.0000 \\
394: \cline{6-9}
395: &&&&& 7 & 0.2490 & 1.2392 & 0.0000 \\
396: \cline{6-9}
397: &&&&& 8 & 0.2783 & 1.3133 & 0.0000 \\
398: \cline{6-9}
399: &&&&& 9 & 0.2288 & 1.1861 & 0.0000 \\
400: \cline{6-9}
401: &&&&& 10 & 0.0720 & 0.6567 & 0.0000 \\
402: \cline{1-4} \cline{6-9}
403: \end{tabular}
404: \end{center}
405: \caption{\sf Eigenvalues $\lambda$
406: and corresponding growth rates
407: $\nu$ for the
408: solitons $\psi_1$ (left panel) and $\psi_2$ (right panel).
409: We included only the eigenvalues which can,
410: potentially, give rise to the largest growth rate in each
411: ``symmetry class" $m$. Some other eigenvalues have been
412: filtered out using the above selection rules.}
413: \end{table}
414:
415: In order to compare the conclusions based on the linearised
416: analysis with direct numerical simulations of the unstable
417: solitons $\psi_1$ and $\psi_2$, we fix some $h$ and $\gamma$ and
418: identify the eigenvalue with the maximum growth rate in each case.
419: In the case of the soliton $\psi_1$, we choose $\gamma=3.5$ and
420: $h=3.6$; the corresponding $\epsilon=0.9146$. The real and
421: imaginary parts of $\lambda$ for each $m$ as well as the resulting
422: growth rates $\nu$ are given in Table 1 (left panel). The
423: eigenvalue with the largest ${\rm Re} \lambda$ is associated with
424: $m=0$; however, for the given $\epsilon$ and $\gamma$ the
425: resulting $\nu <0$. (This is because we have chosen a
426: point in the ``radially stable'' part of the $(\gamma, h)$-plane,
427: to the right of the ``$n=1$'' curve in Fig.{\ref{chart}}.) On the
428: contrary, the growth rates corresponding to the real eigenvalues
429: associated with
430: $m=3,4,5$ are
431: positive for all $\gamma$. The maximum growth rate is associated
432: with $m=4$. The corresponding eigenfunctions $u(r)$ and $v(r)$
433: have a single maximum near the position of the minimum of the
434: function ${\cal R}_1(r)$; that is, the perturbation is
435: concentrated near the circular ``valley" in the relief of
436: $\psi(x,y)|^2$.
437: This
438: observation suggests that for $\gamma =3.4$ and $h=3.5$, the
439: soliton $\psi_1$ should break into a symmetric pattern of 5
440: solitons $\psi_0$: one at the origin and four around it.
441:
442: Next, in the case of the soliton $\psi_2$ we fix $\gamma=4.5$ and
443: $h=4.53$; this gives $\epsilon=0.6846$. The corresponding
444: eigenvalues, for each $m$, are presented in Table 1 (right panel).
445: Again, the eigenvalue with the largest $\mbox{Re} \lambda$ is the
446: one for $m=0$ but the resulting $\nu$ is negative. The largest
447: growth rates ($\nu_3=0.54$ and $\nu_4=0.53$, respectively) are
448: those pertaining to $m=3$ and $m=4$. The corresponding
449: eigenfunctions have their maxima near the position of the minimum
450: of the function ${\cal R}_2(r)$. Therefore, the circular ``valley"
451: of the soliton $\psi_2$ is expected to break into three or four
452: nodeless solitons $\psi_0$. (Since $\nu_3$ is so close to $\nu_4$,
453: the actual number of resulting solitons
454: --- three or four --- will be very sensitive
455: to the choice of the initial perturbation.) Next, eigenfunctions
456: pertaining to $m=5,6,...10$ have their maxima near the second,
457: lateral, maximum of the function ${\cal R}_2(r)$. The largest
458: growth rate in this group of eigenvalues arises for $m=8$. Hence,
459: the circular ``ridge" of the soliton $\psi_2$ should break into 8
460: nodeless solitons, with this process taking longer than the
461: bunching of the ``valley" into the ``internal ring" of solitons.
462:
463: The direct numerical simulations corroborate the above scenarios.
464: The $\psi_1$ soliton with $\gamma=3.5$ and $h=3.6$ splits into a
465: constellation of 5 nodeless solitons: one at the origin and four
466: solitons of opposite polarity at the vertices of the square
467: centered at the origin. The emerging nodeless solitons are stable
468: but repelling each other, see Fig.\ref{fig4}. Hence, no
469: stationary nonsymmetric configurations are possible; the
470: peripheral solitons escape to infinity. The $\psi_2$ soliton with
471: $\gamma=4.5$ and $h=4.53$ has a more complicated evolution. As
472: predicted by the linear stability analysis, it dissociates into 13
473: nodeless solitons: one at the origin, four solitons of opposite
474: polarity forming a square around it and eight more solitons (of
475: the same polarity as the one at the origin) forming an outer ring.
476: (The fact that the inner ring consists of four and not three
477: solitons, is due to the square symmetry of our domain of
478: integration which favours perturbations with $m=4$ over those with
479: $m=3$.) In the subsequent evolution the central soliton and the
480: nearest four annihilate
481: and only the eight outer solitons persist. They
482: repel each other and eventually escape to infinity, see
483: Fig.\ref{fig5}.
484:
485: In conclusion, our analysis suggests the interpretation of the
486: nodal solutions as degenerate, unstable coaxial complexes of the
487: nodeless solitons $\psi_0$.
488:
489:
490: {\bf Acknowledgements.} This project was supported by the NRF of
491: South Africa under grant 2053723. The work of E.Z. was
492: supported by an RFBR grant 03-01-00657.
493:
494:
495: \begin{thebibliography}{99}
496: \bibitem{Swinney} P.B. Umbanhowar, F. Melo, and H.L. Swinney,
497: Nature {\bf 382}, 793 (1996)
498:
499: \bibitem{Faraday} O. Lioubashevski {\it et al\/},
500: %Y. Hamiel, A. Agnon, Z. Reches, and J. Fineberg,
501: Phys. Rev. Lett. {\bf 83}, 3190 (1999);
502: H. Arbell and J. Fineberg,
503: {\it ibid.\/} {\bf 85}, 756 (2000)
504:
505: \bibitem{Astruc} D. Astruc and S. Fauve, in: Fluid Mechanics and Its
506: Applications, vol.62, p. 39-46
507: %in: IUTAM Symposium on
508: %Free Surface Flows (A.C.King, Y.D.Shikhmurzaev, eds.,
509: (Kluwer, 2001)
510:
511:
512: \bibitem{numerical} L.S. Tsimring and I.S. Aranson, Phys. Rev. Lett.
513: {\bf 79}, 213 (1997);
514: E. Cerda, F. Melo, and S. Rica, {\it ibid.\/} {\bf 79}, 4570 (1997);
515: % discrete-time map with spatial coupling
516: S.C. Venkataramani and E. Ott, {\it ibid.\/}
517: {\bf 80}, 3495 (1998); % oscillons in discrete time model
518: D. Rothman, Phys. Rev. E {\bf 57}, 1239 (1998);
519: % \semicontinuum model for granular Faraday resonance
520: J. Eggers and H. Riecke, {\it ibid.\/} {\bf 59}, 4476 (1999);
521: % hydrodynamic-type equations for granular materials
522: C. Crawford, H. Riecke, Physica D {\bf 129}, 83 (1999); % oscillons in
523: %Swift-Hohenberg
524: % including provisionally:
525: H. Sakaguchi, H.R. Brand, Europhys. Lett. {\bf 38}, 341 (1997);
526: Physica D {\bf 117}, 95 (1998)
527:
528:
529: \bibitem{us} I.V. Barashenkov, N.V. Alexeeva, and E.V. Zemlyanaya, Phys.
530: Rev. Lett. {\bf 89}, 104101 (2002)
531:
532: \bibitem{Rypdal} See e.g. K. Rypdal, J.J. Rasmussen and K. Thomsen,
533: Physica {\bf D 16}, 339 (1985) and references therein.
534: \bibitem{Malkin} E.A. Kuznetsov and S.K. Turitsyn,
535: Phys. Lett. {\bf 112A}, 273 (1985); V.M. Malkin and E.G. Shapiro,
536: Physica D {\bf 53}, 25 (1991)
537:
538: \bibitem{collapse} For a recent review and references on the blowup
539: in 2D and 3D NLS equations, see e.g. L. Berge, Phys. Rep. {\bf
540: 303}, 259 (1998);
541: G. Fibich and G. Papanicolaou,
542: SIAM J. Appl. Math. {\bf 60}, 183 (1999)
543:
544:
545: \bibitem{BBK} I.V. Barashenkov, M.M. Bogdan and V.I. Korobov,
546: Europhys. Lett. {\bf 15}, 113 (1991)
547:
548:
549: \end{thebibliography}
550:
551:
552: \end{document}
553: