1: \documentstyle[12pt]{article}
2: \textheight 18cm
3: \textwidth 16cm
4: \begin{document}
5: \centerline{\LARGE { Solitons and wavelets: Scale analysis and bases}}
6: \vskip 1cm
7: \centerline{A. Ludu, R. F. O'Connell and J. P. Draayer}
8: \centerline{Department of Physics and Astronomy,}
9: \centerline{Louisiana State University, Baton Rouge, LA 70803-4001}
10:
11: \begin{abstract}
12: We use a one-scale similarity analysis which gives specific relations between
13: the velocity, amplitude and width of localized solutions of nonlinear differential
14: equations, whose exact solutions are generally difficult to obtain.
15: We also introduce kink-antikink compact solutions for the nonlinear-nonlinear dispersion
16: K(2,2) equation, and we construct a basis of scaling functions similar with those used in the
17: multiresolution analysis. These approaches are useful in describing
18: nonlinear structures and patterns, as well as in the derivation of the time evolution of
19: initial data for nonlinear equations with finite wavelength soliton solutions.
20: \end{abstract}
21: \vskip 0.8cm
22: PACS numbers: 42.30.Sy, 02.30.Px, 43.35.+d, 47.20.Ky, 05.45.-a,
23:
24: \vskip 3cm
25: \section{Introduction}
26:
27:
28: The nonlinear partial differential equations (NPDE) of physical interest can describe
29: a variety of patterns, particle-like traveling solutions, solitons and breather modes
30: in nuclear
31: and particle physics, nonlinear molecular and solid state physics phenomena, and
32: features found in nonlinear optics \cite{general1}. Their solutions are usually localized and
33: demonstrate stability in time and in collisions with each other. In the asymptotic domain
34: these solutions consist of isolated traveling pulses that are free of interactions and
35: have a shape related to the velocity, thus making nonlinear patterns distinct from linear
36: results. In the scattering domain, the nonlinear solutions obey some nonlinear superposition
37: principle. Nonlinear dynamics require NPDE which display very
38: strong interaction between the initial conditions and the dynamics and involve
39: multiple scales \cite{chaos,multiple}, being able to produce self-similar or fractal-like
40: patterns. Recent
41: examples show that the traditional nonlinear tools (inverse scattering, group symmetry,
42: functional transforms) are not always applicable \cite{compacton1}. From the experimental point
43: of view one knows that such patterns generally have finite space-time extension and a
44: multi-scale structure. Since the traditional solitons or the soliton-like solutions have
45: infinite extent, one needs rather appropriate compact supported basis functions to
46: investigate such structures.
47:
48: Multi-resolution analysis (MRA), \cite{wavelet1}, could be a useful method for the construction
49: of such nonlinear bases, since the linear harmonic analysis is inadequate
50: for describing nonlinear systems. The MRA uses wavelets, which are functions that
51: have a space-dependent scale which renders them an invaluable tool for analyzing multi-scale
52: phenomena. Wavelets have been used in signal processing, problems
53: involving singular potentials in quantum mechanics, in discussions concerning q-algebras,
54: and even in nuclear structure studies \cite{wavelet2}. It follows that the use of MRA
55: in the study of NPDE is natural because wavelets can analyse nonlinear features like
56: strong variations or singularities. The NPDE describe
57: a variety of patterns \cite{chaos}, features in quantum optics \cite{optics},
58: molecular and solid state physics phenomena \cite{molecular} and
59: solitons in nuclear and particle physics \cite{nuclear,particle}.
60:
61:
62: In this paper, wavelet-inspired approaches for localized solutions of NPDE are explored.
63: We propose two different formalisms for the scale analysis, and
64: the classification of soliton solutions of NPDE. A first method provides
65: relations between the characteristics of such solutions (amplitude, width and velocity)
66: without the need of solving the corresponding NPDE. The method uses the
67: multi-resolution analysis
68: \cite{wavelet1} instead of the traditional tools like the Fourier integrals or linear harmonic
69: analysis which are inadequate for describing such systems. Moreover, the
70: introduction of wavelet analysis in the study of NPDE is somehow natural because it can
71: accommodate behaviour diplaying strong variations, even singularities, to a smooth behavior.
72: This scale approach has the advantage that it does not need the explicit form of the exact
73: solutions. Hence, it is useful especially in situations when such solutions are unknown. In
74: section 2 we provide many examples, predictions and applications of this one-scale
75: approximation method (OSA) for a large class of NPDE with respect to their localized solutions.
76: The NPDE is mapped into an algebraic equation relating the amplitude, width and velocity of such
77: signals. The results are succinctly presented in tabular form. In section 2 we show an
78: example of the construction of a nonlinear basis for NPDE i.e. the construction of a
79: kink-antikink basis for the K(2,2) equation, involving nonlinear dispersion. There are many
80: physical reasons favoring wavelets in the construction of such nonlinear bases. The
81: self-similar character of the fission process of fluid drops is an example where the same type of
82: singularity occurs in any scale \cite{chaos,prl,bona}. In section 3, we use also this basis (or
83: frame) for the investigation of the time evolution of a given initial data profile for the
84: nonlinear-nonlinear dispersion K(2,2) equation. The proofs are introduced in two Appendices.
85:
86:
87: \section{Scale analysis of NPDE}
88:
89:
90: In this section we introduce a one-scale analysis (OSA) for the NPDE, in terms of their
91: localized traveling solutions. The NPDE which describe physical phenomena are usually
92: not susceptible to analytic solutions. Moreover, there are examples \cite{compacton1,fred}
93: when the mathematical tools like the inverse scattering theory or the transformation group
94: method are not applicable. When phenomena of interest have many space-time scales,
95: or the scale of the process varies in time, the numerical methods may fail, like in
96: the case of propagating discontinuities or shock waves. A simple option is the expansion of
97: solutions in a basis of appropriate chosen basis. The
98: Fourier series, which has the advantage of orthogonality, cannot discriminate the local
99: behavior of phenomena. Moreover, the analysis of the Fourier
100: coefficients of a wave $u(x,t)$ is not sufficient for drawing conclusions about the
101: scale of the localized structures.
102:
103: The so called OSA analysis, described and applied here for localized
104: traveling solutions belonging to any type of NPDE, provides algebraic connections between the
105: width $L$, amplitude $A$, and the velocity $V$ of the solution, without actually solving the
106: equation. The procedure consists in the substitution of all the terms in the NPDE, according
107: to the rules:
108: $$
109: \hbox{i.} \ \ u_t \rightarrow -Vu_x \ \ \ \hbox{substitution for traveling solutions},
110: $$
111: \begin{equation}
112: \hbox{ii.} \ \ \ u \rightarrow \pm A, \ \ \ \ \
113: u_x \rightarrow \pm A/L, \ \ \ \ \
114: u_{xx} \rightarrow \pm A/L^2 \dots ,
115: \end{equation}
116: and so forth for higher order of derivatives. This substitution in eq.(1 ii)
117: is possible only for localized (finite extended support) solutions having at least one local
118: maximum (like solitons or Gauss functions), if they exist. The advantage
119: of the substitution follows from the fact that it enables one to say something about the
120: hypothetical localized traveling solutions even if (or especially when) they are not known or
121: discovered. However, this averaging method may fail in the case of strictly monotonic solutions,
122: or solutions with singularities. This will be made more explicit in the next paragraph and
123: Appendix 1.
124:
125: Since we are interested in traveling solutions, the first substitution reduces the number of
126: variables from 2 to 1 so that we are now dealing with an ordinary differential equation
127: instead of a PDE. Then, the second substitution transforms the ordinary differential equation
128: into an equation in the parameters describing the amplitude, width and velocity. Consequently,
129: the NPDE is mapped into an algebraic equation in
130: $A,L$ and $V$.
131:
132: The proof of the method follows from the expansion of the soliton-like
133: solution $u(x,t)=u(s)$ with $s=x-Vt$, in a Gaussian family of wavelets $\Psi (s)=Ne^{Q(s)}$,
134: where $Q(s)$ is a polynomial and $N$ the normalization constant \cite{wavelet1,wavelet3}.
135: If we choose $Q= -i s-{{s^2}\over 2}$ we obtain a very particular wavelet with the support
136: mainly confined in the $(-1,1)$ interval, namely $\Psi (s)= \hbox{exp}[-i
137: s-{{s^2}\over 2}]/\pi ^{1/4}$. We have the discrete wavelet expansion of
138: $u$
139: \begin{equation}
140: u(s)=\sum_{j}\sum_{k}C_{j,k}\Psi (2^j s -k)=\sum_{j,k}C_{j,k}\Psi _{j,k}(s),
141: \end{equation}
142: in terms of integer translations ($k$) of $\Psi $, which provide the analysis of localization,
143: and in terms if dyadic dilations ($2^j$) of $\Psi $, which provide the description of different
144: scales. The idea of the proof is to choose a point where the expansion in eq.(2) can be well
145: approximated by one single scale such that $u(s)$ can be approximated with a sum of phases in
146: $s$. In order to reduce the number of scales, the range of the summation
147: should be chosen with an eye to the underlying physics. The exact proof of this approach is
148: provided in Appendix 1.
149:
150: In the following we apply this procedure to three classes of NPDE, namely:
151: dispersive, diffusive and dispersive-diffusive nonlinear equations, and we present
152: the results in tabular form.
153: Before describing different examples of NPDE, we point out that for each equation we try
154: to present one example of an exact solution (second column Tables 1,2) in comparison with the
155: results of the much simplest OSA approach (third column of first two
156: tables and second column of Table 3). We stress that we do not choose only that solution which
157: fits the similarity analysis. On the contrary, our approach fits a large number of solutions, as
158: can be checked directly.
159:
160:
161: Our first example of application of the OSA approach is given by the convective-dispersive
162: equations, for which the most celebrated example is provided by the KdV equation
163: \begin{equation}
164: u_t +uu_x + u_{xxx} =0.
165: \end{equation}
166: In general, the stable solutions of the nonlinear-dispersive
167: equations are dependent of the initial conditions, through their conservation laws.
168: Consequently, they can generate a large class of patterns, shaped by the balance between
169: nonlinear interaction and dispersion, among which the most interesting examples are solitons,
170: breathers and kinks.
171:
172: In Table 1 we present examples of pure dispersive NPDE, identified in the
173: first column by the form of the equation, and in the second column by a
174: corresponding traveling localized solution, if the analytical form is available. Such
175: exact solutions provide special relations between $L,A$ and $V$, which are given in the third
176: column of Table 1. In the last column we introduce the results of OSA, namely the relations
177: between these three parameters, provided by eqs.(1 ii). The usefulness of the approach may be
178: checked, by a quick comparison between the second and the third columns. While the results in
179: the second column are possible only when one knows the analytical solutions, the results
180: presented in the last column, obtained by the OSA, result directly from the
181: NPDE form, without actually solving it.
182:
183: The case of the KdV equation, eq(3), is described in the first row of the Table 1.
184: The OSA method gives a general expression for $L=L(A,V)$, shown in the first line, first row of
185: the last column. In the second line of the last column we show that a specification of the
186: relation between $V$ and $A$, together with the general result given in the first line, results
187: in a relation between $L$ and $A$. For $L$ to be related
188: to $A$ only, it results from the third column that the velocity $V$ must be proportional
189: to $A$. In this case we obtain the well-known relation (first row, column two) among
190: the parameters in the soliton solution $V=2A$. Moreover, OSA method allows $V$ to depend on a
191: higher power of $A$ ($V \sim A^{p}$, $p \geq 1$). If such a solution could exist, a lower bound
192: for $A$ will occur. Such solitons would have only amplitudes higher than this limit, while
193: solitons with a smaller amplitude than this limit move with velocity proportional to $A$.
194:
195: Similar results are obtained for the MKdV equation (second row), except that here $A$ needs to
196: be proportional to the square root of $V$ in order to have $L$ a function of $A$ only. This
197: prediction is again identical with that in the exact solution (second column). Moreover, the
198: same relations remain valid even for the solutions of the "compacton" type
199: \cite{electronic}
200: $$
201: u(x,t)={{\sqrt{32}k \cos [k(x-4k^2 t)]^2 }\over {3(1-{2\over 3} \cos [k(x-4k^2 t)]^2)}},
202: $$
203: where $L= \pi /6 k$, that is $L\sim 1/A$, like in the Table 1.
204:
205: Next example (third row) is provided by a generalised KdV equation, in which the dispersion
206: term is quadratic
207: \begin{equation}
208: u _t + (u ^2)_x +(u ^2)_{xxx}=0.
209: \end{equation}
210: Eq.(4), known as K(2,2) equation because of the two quadratic terms, admits compact supported
211: traveling solutions, named compactons \cite{compacton1,fred,rosenau1,rosenau2,rosenau3}. In
212: general, the compactons are obtained in the form of a power of some trigonometric function
213: defined only on its half-period, and zero otherwise, in such a way that the solution is enough
214: smooth for the NPDE in discussion. In the above example the square of the solution has to be
215: continuous up to its third derivative with respect to $x$.
216:
217:
218: Different from solitons, the compacton width is
219: independent of the amplitude and this fact provides the special connection with the wavelet
220: bases. The compactons are characterize by a unique scale, and it is this feature
221: that makes it possible to introduce a nonlinear basis starting from a unique generic
222: function. For eq.(4) the compacton solution is given by
223: \begin{eqnarray}
224: \eta _c (x-Vt)={{4V}\over {3}} cos ^2 \biggl [ {{x-Vt}\over {4}} \biggr ],
225: \end{eqnarray}
226: if $|x-Vt|<2\pi $ and zero otherwise. Here we notice that the velocity is proportional to the
227: amplitude and the width of the wave is independent of the
228: amplitude, $L=4$. As a field of application we mention that the quadratic
229: dispersion term is characteristic for the dynamics of a chain with nonlinear coupling.
230:
231: The general compacton solution for eq.(4) is actually a ''dilated" version of eq.(5).
232: That is, a combination of the first rising half of the squared cos in eq.(5), followed
233: by a flat domain of arbitrary length $\lambda $, and finally followed by the second,
234: descending part of eq.(5). Actually, this combination is just a kink compacton joined
235: smoothly with an antikink one
236: \begin{equation}
237: {\eta}_{kak} (x-Vt; \lambda)=
238: \left\{ \begin{array}{ll}
239: 0 ...\\
240: {{4V}\over 3} \ cos^{2}\biggl [ {{x-Vt}\over 4} \biggr ],
241: \ \ -2\pi \leq x-Vt \leq 0
242: \\
243: {{4V}\over 3} ,
244: \ \ 0 \leq x-Vt \leq \lambda \\
245: {{4V}\over 3} \ cos^{2}\biggl [ {{x-Vt-\lambda }\over 4} \biggr ],
246: \ \lambda \leq x -Vt \leq \lambda +2\pi
247: \\
248: 0 ... \\
249: \end{array}
250: \right.
251: \end{equation}
252: In Fig. 1 we present compactons (upper line), kink-antikink
253: pairs (KAK) described by eq.(6), both with the same amplitude and velocity (middle line).
254: Although the second derivative of this generalized compacton is discontinuous at its edges, the
255: KAK, eq.(6), is still a solution of eq.(4) because the third derivative acts on $u^2$, which is
256: a function of class $C_3$. Finally, we can construct solutions by placing a compacton on the top
257: of a KAK, as in the bottom line of Fig. 1. Such a solution exists only for a short interval
258: of time ($\lambda /V$), since the two structures have different velocities. The analytic
259: expression of the solution is given by
260: \begin{equation}
261: \eta (x,t)=\eta _{kak} (x-Vt;\lambda )+\biggl ( \eta _c (x-V't-2\pi) + {{4V}\over 3 }
262: \biggr ) \chi ({{x-V't-2\pi} \over {2\pi}} ),
263: \end{equation}
264: for $0< t < (\lambda -4\pi)/(V'-V)$ and zero in the rest. Here $\chi(x)$ is the support function,
265: equal with 1 for
266: $|x|\leq 1$ and 0 in the rest, and $V'=3\hbox{max}\{ \eta _c\}/4 +2V $.
267:
268: For the K(2,2) compacton eq.(5) fulfills some relations between the parameters:
269: $A=4V/3$ and $L=4$ \cite{rosenau1}. The relations provided by OSA in the last
270: column of the third row, predict such relations, and hence also proove the existence of the
271: compacton. That is, for a linear relation between the amplitude and the speed, the half-width
272: is constant and does not depend on $A$. Indeed, if we choose $V=\pm 3A/2$ or $V=\pm 5A/2$, we
273: obtain $L=4$, like for the compacton. The constant value of $L\equiv L_0=const.$ is a
274: typical feature of the K(2,2) compactons. Moreover, it was found numerically that for any
275: compact supported initial data, wider than $L_0$, the solution decomposes in time into a series
276: of
277: $L_0$ compactons, Fig. 2. For narrower initial data the numeric solution blows up. There
278: is no exact or analytic explanation of this effect, so far. The scale relations can give a
279: hint in this situation, too, by using the graph of the relation
280: $L=L(V,A)$ provided by this qualitative method. In Fig. 3, $L$ from the third column of the
281: third row is plotted versus $V$, for several values of $A$ (larger values of $A$ translate the
282: curves to the right). Above the value $L=4$ of the half-width of a stable compacton, wider
283: compact pulses produce an intersection for each curve (each $A$) with the axis $L=4$, providing a
284: series of compactons of different heights, like in the numerical experiments described in
285: \cite{compacton1,fred,rosenau2}. Below the $L=4$ line, all the curves $L(V)$ approach $L=0$,
286: towards infinite amplitude, explaining the instability of the narrower initial data.
287:
288:
289: Another good example of the predictive power of the method is exemplified
290: in the case of a general convection-nonlinear dispersion
291: equations, denoted by K(n,m)
292: \begin{equation}
293: \eta _t + (\eta ^n)_x +(\eta ^m)_{xxx}=0.
294: \end{equation}
295: Compacton solution for any $n\neq m$ are not known in general, except for some particular
296: cases. In this case we find a general relation among the parameters, for any $n,m$, shown in
297: the
298: fourth and fifth rows. These general relations $L(A,V)$ approach the known relations for
299: the exact solutions, in the particular cases like $n=m$ (fourth row),
300: $n=m=2$ (third row), $n=m=3$ (first reference in \cite{compacton1}) and $n=3,m=2$;
301: $n=2,m=3$ (fifth row). These results can be used to predict the behavior of
302: solutions for all values of $n,m$.
303:
304: In the following, we present another example of applications of the OSA approach,
305: related to a new type of behavior of nonlinear systems. Traditional solitons move
306: with constant speed on a rectilinear path (except for the roton \cite{prl} which has a
307: circular trajectory with constant angular velocity). The speed is usually equal to the
308: amplitude scaled with a constant. Higher solitons travel faster and there are no
309: solitons at rest (zero speed implies zero amplitude). They can travel in both
310: directions with opposite signs for the amplitude.
311: The situation is different in the case of compactons, which allow also stationary
312: solutions. When linear and nonlinear disspersion occur simultaneously, like in the so
313: called K(2,1,2) equation
314: $$
315: u_t +(u^2 )_{x} + (u)_{xxx} +\epsilon (u^2 )_{xxx}=0,
316: $$
317: where $\epsilon $ is a control parameter, the OSA yields a dependence of the
318: form
319: $$
320: L=\sqrt{(\pm A+\epsilon )/(V\pm A)},
321: $$
322: which still provides a constant width if $V=\pm A +2\epsilon $. In this case, the speed is
323: proportional to the amplitude, but can change its sign even at non-zero amplitude.
324: Solutions with larger amplitude than a critical one ($A_{crit}=\mp 2\epsilon$)
325: move to the right, solutions having the critical amplitude are at rest, and solutions
326: smaller than the critical amplitude move to the left. This behavior was
327: explored in \cite{rosenau1}, too. However, such a switching of the speed is
328: not necessarily a feature of the nonlinear dispersion.
329: A compacton of amplitude $A$ on the top of a infinite-length KAK solution of
330: amplitude $\delta $
331: \begin{equation}
332: u(x,t)= A cos ^2 \biggl ( {{x-Vt}\over{4}} \biggr ) +\delta ,
333: \end{equation}
334: is still a solution of the K(2,2) equation, with the
335: velocity given by $V={3 \over 4}\biggl (2 \delta + A \biggr )$. For $A=-2 \delta$ the
336: solution becomes an anti-compacton moving together with the KAK.
337: In the case of a slow-scale time-dependent amplitude
338: the oscillations in amplitude can transform into oscillations in the velocity.
339: The key to such a conversion of oscillations is the coupling between the
340: traditional nonlinear picture (convection-dispersion-diffusion) and the typical
341: Schr\"odinger terms.
342:
343:
344: In Table 2 we present another class of NPDE, namely the dissipative ones.
345: These equations generalize the linear wave equation (first row) where there is no
346: typical length of the traveling solutions. The wavelet analysis provides the
347: correct expression for the dispersion relation ($V=c$ $\rightarrow$ $k^2=\omega^2/c^2$) with
348: no constraint on either the amplitude $A$ or on the width $L$.
349:
350: In the second row we introduce the celebrated Burgers equation which
351: represents the simplest model for the convective-dissipative interaction. Dissipative systems
352: are to a large extent indifferent to how they were initialized, and follow their own intrinsic
353: dynamics. We provide in the second column an analytic solution of the Burgers equation.
354: For some special of values of the integration constants ($2C<V^2$, $D=0$)
355: the solution becomes a traveling kink
356: \begin{equation}
357: u(x,t)=V+\sqrt{V^2-2C} \hbox{tanh}[\sqrt{V^2 -2C} (x-Vt)] .
358: \end{equation}
359: By applying the OSA approach to the Burger equation (third column) we obtain
360: the same relation between amplitude and half-width, like in the case of the exact solution
361: eq.(10), providing the velocity is proportional with the amplitude.
362:
363: In the following we apply the OSA approach to investigate a
364: nonlinear Burgers equation
365: \begin{equation}
366: u_t +a (u^m )_x -\mu (u^k )_{xx}+cu^{\gamma}=0,
367: \end{equation}
368: called quasi-linear parabolic equation \cite{rosenau3}, and used to describe the flow of
369: fluids in porous media or the transport of thermal energy in plasma. The last term describes
370: the volumetric absorption (Bremsstrahlung for $\gamma = 1/2$, synchrotron radiation for
371: $\gamma$ in the range $1.5 - 2$, etc). The second term in eq.(11) describes the convection
372: process and the coefficient $\beta $ ranges from 0 to 1 in the case of difussion in plasma, and
373: further to higher values, for unsaturated porous medium in the presence of gravity.
374: The existence and stability of waves or patterns is strongly dependent on the coefficients $a,
375: \mu, c, m, k$ and $\gamma $, and at this point the OSA can be
376: useful again since there is no general analytic solution for eq.(11). The result of the
377: OSA approach is presented in the third row of Table 2. The typical scale of
378: patterns depends on the parameters in the equations and the amplitude of the excitations, in a
379: complicated way.
380:
381: However, in order to test OSA again, we found a simple class of exact solutions when
382: $c=0$, presented in the fourt row in Table 2, and expressed as the inverse of a degenerated
383: hypergeometric function. In this expression we have ${\cal A}= \mu k V^{\alpha -1 }
384: ((k-1)a^{\alpha })^{-1}$,
385: $z=(a/V) u^{m-1}$ and $\alpha =(k-1)/(m-1)$.
386: The asymptotic behavior of the left hand side of the solution given in fourth row, second column,
387: is described by
388: $$
389: \Gamma (\alpha +1) \biggl [
390: (-1)^{\alpha} +{1 \over {\Gamma (\alpha )}}z^{\alpha -1}e^{\alpha }
391: \biggr ]+{\cal O}(1/z).
392: $$
393: If $z$ approaches $+ \infty$ the solution increases indefinitely like an exponential.
394: For $\alpha >1$ (strong difussion effects), for even $k$ and for even $m$, the traveling wave
395: $u(x-Vt)$ has a negative singularity towards $-\infty$ at $x+x_0 =\Gamma (\alpha +1)
396: (-1)^{\alpha}<0$. For $k$ odd there is also a singularity at $x+x_0 >0$. These solutions are
397: not likely to provide viable physical results. If $k$ is even and $m$ is odd (the singularity
398: is pushed towards imaginary $x$), or if
399: $0 < \alpha <1$, the singularity is eliminated and the solution becomes semi-bounded, like in
400: the particular situations investigated in the article \cite{rosenau3}. In this case,
401: OSA provides again the correct relations, since we
402: obtaine the special behavior of the solution if the velocity is proportional to the power
403: $m-1$ of the amplitude $A$. Also, we predict the space scale of these
404: semi-compact pulses, namely the length $L={{\mu k^2 A^{k-1}}\over{V\pm amA^{m-1}}}$ .
405:
406: The OSA analysis can be applied in the case of sine-Gordon equation, fifth row of Table 2.
407: The solutions with the velocity proportional with $L^2$ are characterized
408: through the OSA approach by a transcendental equation in $A$, identical
409: with the equation fulfilled by the amplitude $A$ of the exact sine-Gordon soliton.
410:
411: In the sixth row, we present the cubic nonlinear Schr\"odinger equation (NLS3) which has
412: a soliton solution. This type of equation is applied in nonlinear optics, elementary
413: particle physics or in the polaron model in solid state physics \cite{optics,molecular,nuclear}.
414: Recently, different effects of cluster physics could be explained by using the NLS3 equation.
415: In the sixth row of Table 2 we present the NLS3 equation together with its one soliton
416: solution of amplitude $\eta_0$, obtained by the inverse scattering method. In the last column we
417: also show the relation between the parameters of a localized solution, obtained by OSA. The
418: equation for $L(A,V)$ is more general than that one fulfilled by the soliton, and hence is
419: related to more general localized solutions. By choosing the velocity proportional to the
420: amplitude, we reobtain the $L\sim 1/A =1/\eta_0 \sim 1/V$ typical relations for the soliton given
421: in the second column. If a more general solution of the NLS3 equation describes, for example, the
422: dynamics of some cluster states, or the dynamics of hard spheres in a hard core potential model
423: $$
424: -{{\hbar ^2}\over {2m}} \Psi _{xx} +(E-V)\Psi + a \Psi ^3 =0,
425: $$
426: then the $L$ parameter gives an estimation for the wavelength
427: of the wavefunctions, or for the correlation length in a Bose model
428: $$
429: L={{\hbar}\over {\sqrt{2m(E-V)+a A^2}}}\simeq {{\hbar}\over {\sqrt{2m(E-V)}}} \ \
430: \hbox{for
431: small } \ A.
432: $$
433: For the general case of a NLS equation of order
434: $n$ (seventh row), where a general analytical solution is unknown, the method predicts a
435: special $L=L(A,V)$ dependence, shown in the third column and in Fig. 4. Contrary to third order
436: NLS, where the dependence of $L$ with $A$ is monotonous for $V=\sim \pm A$ ($n=3$ in Fig. 4), at
437: higher orders than 3, the $L(A)$ function has discontinuities in the first derivative. This
438: wiggle of the function (Fig. 4, $n=4$) holds at a critical width, possibly producing
439: bifurcations in the solutions and scales. As a consequence, initial data close to this
440: width can split into doublet (or even triplet, for higher order NLS) solutions, with
441: different amplitudes. Such phenomena have been put into evidence in several numerical
442: experiments for quintic nonlinear equations \cite{rosenau2,rosenau3,kart}.
443:
444: The final example of Table 2 is provided by the Gross-Pitaevski (GP) mean field equation, which
445: is used to describe the dilute Bose condensate \cite{bose1}. The scalar field (or order
446: parameter) governed by this equation was shown to behave in a particle manner, too, since it
447: can contain topological deffects, namely dark solitons. The space scale $L$ of such solutions
448: is important, for both the theory and experiment, since is related to the trap dimensions
449: and to the scattering length. In the last row of the Table 2 we give one particular solution
450: of a simplified one-dimensional version of the GP equation \cite{bose2}
451: \begin{equation}
452: i\hbar {{\partial \Psi ({\vec x},t)} \over {\partial t}}=\biggl (
453: -{{\hbar ^2 \triangle} \over {2m}} +V_{ext}({\vec r}) +{{4\pi \hbar ^2 a} \over {m}}
454: |\Psi ({\vec x}, t)|^2
455: \biggr ) \Psi ({\vec x}, t),
456: \end{equation}
457: where $a$ is the s-wave scattering length and $V_{ext}$ is the confining potential. In the
458: solution provided in the table, the half-width of the exact nonstationary solution
459: is $L=1/\sqrt{v_c ^2 - p^2}$, where $v_c \def \sqrt {1-aV}$ is the Landau critical velocity, and
460: $p={\dot q(t)}$ is the momentum associated with the motion of this disturbance.
461: It is easy to check that the OSA provides a good match
462: with this exact solution, and also $L$ fits the correlation length $l_0 = \sqrt{m/4\pi \hbar ^2
463: a}$. We stress that such estimation of the length is also important in nuclear physics
464: where one can explain the fragmentation process as a bosonization in $\alpha $-particles, inside
465: the nucleus. Such systems are coherent if the wavelength associated with the cluster (the
466: resulting $L$ in the GP equation) is comparable with the distance between the $\alpha
467: $-clusters.
468:
469:
470: The more complex the NPDE is, the richer the conclusions of the OSA approach. A good example of
471: such an analysis is related to the convective-dissipative-dispersive NPDE, example provided by
472: the model equation
473: \begin{equation}
474: u_t + a(u^m ) _{x} + b (u^k )_{xx}+ c(u^n)_{xxx}=0
475: \end{equation}
476: Here $m, k$ and $n$ are integers and the corresponding terms are responsible
477: of the nonlinear interaction (convective term), dissipation and dispersion \cite{rosenau3}.
478: The above equation is related to weakly nonlinear phenomena, and it occurs in
479: modeling porous medium, magma, interfacial phenomena in fluids (and hence applications to drop
480: physics), etc. General solutions are difficult to obtain for such a complex equation.
481: We show in the following that most of the essential conclusions for the behavior of its solutions
482: can be obtain, in a simple way, by the OSA. This approach maps this equation
483: into
484: \begin{equation}
485: (amA^{m-1}-V)L^2 -\mu k^2 A^{k-1}L+n^3 A^{n-1}=0,
486: \end{equation}
487: Table 3, first row.
488: The most symmetric case is obtained when either $V=0$ (stationary patterns)
489: or $V\sim A^{m-1}$. In this situation the condition to have a monotonous dependence
490: of $L$ as a function of $A$ is $2k=m+n$ which yields a scale structure
491: \begin{equation}
492: L=A^{k-m}
493: \biggl (
494: {{\mu k^2 \pm \sqrt{
495: \mu ^2 k^4 -4mn^3 (a-V_0 )
496: }}\over {2m(a-V_0 )}}
497: \biggr )\sim A^{k-m},
498: \end{equation}
499: where we put $V=mV_0 A^{m-1}$.
500: The condition $2k=m+n$ is just the condition obtained in \cite{rosenau3} from a scaling
501: approach. This condition assures the universality of the corresponding patterns, and it is the
502: unique case in which $L$ depends on a power of $A$. In any other situation, this dependence is
503: more complicated and introduces singularities which broke the self-similarity.
504: In the above cited paper, the author finds out the condition for mass invariance at scaling
505: transformations as $m=n+2=k+1$. In our case we just have to request the product $AL$ (which
506: gives a measure of the mass, or volume of the pattern, like in the case of
507: one-dimensional solitons) to be a constant. This gives the condition $k-m+1=0$ which, together
508: with the general invariance condition $2k=m+n$, reproduces $m=n+2=k+1$. In this case we have
509: patterns characterized by a width
510: $$
511: L={{\mu k^2 \pm \sqrt{\mu ^2 k^4 -4mn^3 (a-V_0 )}}\over {2mA(a-V_0 )}}\rightarrow n^3 /A\mu k^2.
512: $$
513: If $a\sim V_0$ the width approaches $ n^3 /A\mu k^2$.
514: In order to make $L$ independen of $A$, like in the compacton case, we need $m=k$, which
515: together with the first invariance condition $2k=m+n$, yields $m=n=k$. This is the
516: exceptional case when the dissipative and dispersive processes have the same scaling, resulting
517: form the invariance of the eq.(13) under the group of scales.
518: Finaly, if we choose $L\sim V$ we obtain the condition $k+1=2m$ which (together with $2k=m+n$)
519: is the condition for spiral symmetry and occurence of similarity structures \cite{rosenau3}.
520:
521: The above comments are not intended to be a complete study of such a complex equation, but the
522: just a proof of how many conclusions one can obtain from the simple equation in $A,L$ and $V$,
523: eq.(14).
524:
525: The next example is provided by one of the most generalized KdV equation, which is
526: generated from the Lagrangian \cite{fred}
527: \begin{equation}
528: {\cal L}(n,l,m,p)=\int \biggl [
529: {{\phi_x \phi _t}\over {2}}+\alpha {{(\phi_x)^{p+2}}\over {(p+1)(p+2)}}
530: -\beta (\phi_x)^m (\phi_{xx})^2 +{{\gamma }\over 2} (\phi _x)^n (\phi _{xx})^l (\phi _{xxx}) ^2
531: \biggr ],
532: \end{equation}
533: where $\alpha, \beta$ and $\gamma$ are parameters adjusting the relative strentgh of the
534: interactions, and $n,l,m,p$ are integers. For example, for $\gamma =0$ one re-obtains the K(2,2)
535: equation, and for $\gamma=m=0, p=1$ one obtains the KdV equation.
536: The associated Euler-Lagrange equation in the function $\phi_x =u(x,t)\rightarrow u(x-Vt)=u(y)$,
537: reads after one integration
538: $$
539: Vu={{\alpha }\over {p+1}}u^{p+1}-\beta mu^{m-1}(u_y )^2 +2\beta (u^{m} u_y )_{y}
540: +{{\gamma n}\over {2}} u^{n-1} (u_y )^l (u_{yy})^ 2
541: $$
542: \begin{equation}
543: -{{\gamma l} \over 2} (u^{n} (u_y )^{l-1}
544: (u_{yy})^2 )_{y}+\gamma (u^n (u_y )^l u_{yy})_{yy} +C,
545: \end{equation}
546: where $C$ is the integration constant. By using the OSA we obtain
547: the following important result, expressed in the second row of Table 3: The unique case when such
548: an equation allows compact supported traveling solutions is when $m=p=n+r$, $C=0$ and $V=V_0
549: A^{m}$. This result is in full agrement with the variational calculation in \cite{fred}.
550:
551: Both eqs.(13) and (17) are rather more qualitative than capable of modeling measurable phenomena.
552: That is why we introduce now a more general model equation, in the form
553: \begin{equation}
554: u_t +f(u)_x +g(u)_{xx}+h(u)_{xxx}=0,
555: \end{equation}
556: where $f,g$ and $h$ are differentiable functions of the the function $u(x,t)$ itself.
557: The OSA approach gives the equation
558: \begin{equation}
559: -V+f'(A)+{{Ag''(A)+g'(A)}\over L}+{{A^2 h'''(A)+3Ah''(A)+h'(A)}\over {L^2}}=0.
560: \end{equation}
561: A general analysis of eq.(19) is difficult, and the best ways are numerical investigations
562: obtained for particular choices of the three functions. We confine ourselves here only to show
563: that the class of solutions which have similarity properties are those for which
564: $V=V_0 f'(A)$. In this case eq.(19) can be reduced to
565: \begin{equation}
566: L^2 f' (1-V_0 )+L(Ag'' +g' )+A^2 h''' +3A h''+h'=0,
567: \end{equation}
568: case which is presented in the third row of Table 3.
569: This last relation can be used for different purposes. For example, given a certain
570: type of dispersion and difusion ($g,h$ fixed), we can estimate for what types of nonlinearity
571: ($f$) the width $L$ will have a given dependence with $A$. Or, if we know for instance
572: $f(u)=f_0 u^{q_1}$ and $h(u)=h_0 u^{q_2}$, we can ask what type of diffusion $g$
573: we need, to have constant scale (width) of the patterns (waves), no matter of the magnitude of
574: the amplitude $A$. In other words, which is the compatible diffusion term, for given
575: nonlinearity-dispersion terms, which provides fixed scale solutions.
576: The result is obtained by integration eq.(20) with respect to $g(u)$
577: \begin{equation}
578: g(u)=-{{h_0 }\over L} \biggl ( 1+q_2 +{1 \over {q_2 -1}}\biggr ) u^{q_2} -{{Lf_0 (1-V_0
579: )}\over{q_1 -1}}u^{q_1} +C_3 \hbox{Log} u +C_4,
580: \end{equation}
581: where $C_{3,4}$ are constants of integration. In a similar way one can check the existence of
582: different other configurations by solving eq.(20), or more general, eq.(19).
583:
584:
585:
586: A last application of this method, occurs if the KdV equation has an
587: additional term depending on the square of the curvature
588: \begin{equation}
589: u_t +uu_x +u_{xxx} +\epsilon (u_{xx}^{2})_{x}=0.
590: \end{equation}
591: This is the case for extremely sharp surfaces (surface waves in solids or
592: granular materials) when the hydrodynamic surface pressure cannot be linearized
593: in curvature. Such a new term yields a new type of localized solution fulfilling the relations
594: $$
595: L=\sqrt{
596: {{4\epsilon A}\over {\pm \sqrt{1-8\epsilon A(A\pm V)}-1}}}.
597: $$
598: If we look for a constant half-width solution (compacton of $1/L =\alpha $) we need
599: a dependence of velocity of the form $V=(1+\alpha ^2 \epsilon /8)A+1/8\epsilon A +\alpha /4$.
600: There are many new effects in this situation. The non-monoton dependence of the
601: speed on $A$ introduces again bifurcations of a unique pulse in dublets and triplets.
602: Also, there is an upper bound for the amplitude at some critical values of the width. Pulses
603: narrower than this critical width drop to zero. Such bumps can exist in
604: pairs of identical amplitude at different widths. They may be related with
605: the recent observed "oscillations" in granular materials \cite{rosenau1,oscillon}.
606:
607: The examples presented in Tables 1-3 prove that the above method provides a reliable
608: criterion for finding compact suported solutions. The reason this simple
609: prescription works in so many cases follows from the advantages of wavelet
610: analysis on localized solutions. We stress that this method has little to do
611: with the traditional similarity (dimensional) analysis
612: \cite{compacton1,bona,fred,rosenau1,rosenau2,simil}. In the latter case one obtains
613: relations among powers of $A, L$ and $V$, not relations with numeric coefficients like those
614: found in our method.
615:
616:
617:
618: \section{The frame of KAK pairs}
619:
620:
621:
622: In the following we investigate the posibility of construction of a nonlinear frame
623: (an over determined or incomplete basis) by using
624: some compact solutions of the K(2,2) equation.
625: The high stability against scattering of the K(2,2) compactons, or
626: compacton generation from compact initial data, suggest they may play the role of a nonlinear
627: local basis. We know from many numerical experiments \cite{bona,fred,electronic},
628: that any positive compact initial data decomposes into a finite series of compactons and
629: anticompactons. This suggests that an intrinsic ingredient for a nonlinear basis could be the
630: multiresolution structure of the solutions, similar with the structure of scaling functions in
631: wavelet theory.
632:
633: The compactons given in eqs.(5,6) have constant half-width and hence
634: describe a unique scale, which can cover all the space by integer translations. From the point
635: of view of multi-resolution analysis, the K(2,2) equations act like a
636: $L$-band filter, allowing only a particular scale to emerge for any given set
637: of initial condition. To each scale, from zero to infinity, we can associate a K(2,2)
638: equation with different coefficients. However, the compacton solution is not the
639: unique one with this property. For a given K(2,2) equation, we can thus extend the scale from
640: $L$ to any larger scale. These more general compact supported solutions are still
641: $C_{2}({\bf R})$ and are combinations of piece-wise constant and piece-wise $\cos ^2$
642: functions. The simplest shape is given by a half-compacton prolonged with a constant level,
643: that is a kink solution. The
644: basis solution is a kink-antikink (KAK) compact supported combination, Fig. 1. Such
645: kink-antikink pairs of different length, can be associated with other compactons, or KAK pairs,
646: one on the top of the other
647: \begin{equation}
648: {\eta}_{comp+KAK} (x-Vt; \lambda)=
649: \left\{ \begin{array}{ll}
650: 0 ...\\
651: {{4V}\over 3} \ cos^{2}\biggl [ {{x-Vt}\over 4} \biggr ],
652: \ \ -2\pi \leq x-Vt \leq 0 \\
653: {{4V}\over 3} , \ \ 0 \leq x-Vt \leq \delta \\
654: {{4V}\over 3}+{{4}\over 3}(V'-2V) \ cos^{2}\biggl [ {{x-V't}\over 4} \biggr ],
655: \ \ \delta \leq x-Vt \leq \delta +4\pi \\
656: {{4V}\over 3} , \ \ \delta+4\pi \leq x-Vt \leq \lambda \\
657: {{4V}\over 3} \ cos^{2}\biggl [ {{x-Vt-\lambda }\over 4} \biggr ],
658: \ \lambda \leq x -Vt \leq \lambda +2\pi
659: \\
660: 0 ... \\
661: \end{array}
662: \right.
663: \end{equation}
664: where $\delta < \lambda$ characterizes the initial position (at $t=0$) of the top compacton,
665: with respect to the flat part of the KAK solution. The amplitude $4(V'-2V)/3$ of the compacton,
666: and the amplitude $4V/3$ of the KAK, are related to their velocities $V'$ and $V$,
667: respectively. The length of the flat part, $\lambda $, is arbitrary. A compound solution is
668: not stable in time since the different elements travel with different velocities. The total
669: height of the compacton is $4(V'-V)/3$. Since the higher the amplitude is, the faster the
670: structure travels, the top compacton moves faster than the KAK, and at a certain moment it
671: passes the KAK. Because the area of the solution is conserving, such a compound structure
672: decomposes into compactons and KAK pairs. Similar and even more complicated constructions can be
673: imagined, with indefinite number of compactons and KAK's, if one just
674: fulfills the $C_3$ continuity condition for the square of the total structure.
675: Such structures, defined at the initial moment can interpolate any function, playing a similar
676: role with wavelets or spline bases.
677: It has been also proved that the KAK solutions are stable, by using both
678: a linear stability analysis and Lyapunov stability criteria.
679:
680: For a given K(2,2) equation, the compacton solution, eq.(5) and in addition the family of KAK
681: solutions, eq.(6) can be organized as a scaling functions system. They act
682: like a low-pass filter in terms of space-time scales and
683: give the opportunity to construct frames of functions from the wavelet model
684: \cite{wavelet1,wavelet2,wavelet3}.
685:
686: With the notation from Appendix 2, and from eq.(32), we have the elements of the frame
687: $$
688: \eta _{k,j} (x)=\eta_{kak} (\pi (x-2^{j}Vt-k), 2^{j}-1)|_{t=0},
689: $$
690: where $t=0$ means that we neglect the time evolution, but the amplitude is still amplified with
691: a factor of $2^j$, in virtue of relation $\eta_{max}=4V/3$.
692: We can now expand any
693: initial data for the K(2,2) equation in this frame
694: \begin{equation}
695: u_{0} (x) = \sum_{k}\sum _{j} C_{k,j} \eta _{k,j} (x),
696: \end{equation}
697: and taking into account that
698: \begin{equation}
699: \eta _{k,j} \eta _{k',j'}
700: \left\{ \begin{array}{ll}
701: \neq 0 & k'=k\cdot 2^{j'-j}, ..., (k+1)\cdot 2^{j'-j}-1 \\
702: =0 & \mbox{otherwise,}
703: \end{array}
704: \right.
705: \end{equation}
706: we can show
707: that the square of the initial data can be linearized by
708: $$
709: u^2 (x)=\sum _{k,j} \sum _{j' \geq j} \sum _{k' \in I}C_{k,j}C_{k',j'}
710: $$
711: \begin{equation}
712: \times \biggl (
713: \sum_{i_1 =0}^{1} \sum_{i_2 =0}^{1}... \sum_{i_{j'-j}=0}^{1} \eta _{\sigma (i_1
714: ,i_2 , ..., i_{j'-j}),j'}
715: \biggr ) \eta _{k',j'},
716: \end{equation}
717: where $I$ is the range of $k'$ described in the first line of eq. (25), and
718: $$
719: \sigma (i_1 , i_2 , ..., i_{j'-j}) =
720: \sum_{l=1}^{j'-j} i_l 2^{j'-j'l+{{(j'-j)(j'-j+1)-l(l+1)}\over 2}}
721: $$
722: \begin{equation}
723: +k2^{(j'-j)j+{{(j'-j)(j'-j+1)}\over 2}}.
724: \end{equation}
725: From this relation and from eq.(25) we notice that in eq.(27) the unique nonzero terms are
726: those for which $\sigma (i_1 , i_2 ,..., i_{j'-j} )=k'$ with $k' \in I$. Hence the initial data
727: is expanded in different scales at different translations. The translations are mutually
728: orthogonal so they do not give a contribution to the square. In the multiplication of two
729: different scales in the expression of the square, we reduce the wider scale in
730: terms of linear combination of the narrower ones, by using the two-scale equation,
731: eq.(31) in Appendix 2. All the nonzero terms in this product
732: are of the order $(2^{-j}-1)/(2^{-j'}-1)\simeq 2^{j'-j}$. This number, given
733: by the number of solutions of the equations $\sigma (i_1 , i_2 ,..., i_{j'-j} )=k'$,
734: with $k' \in I$, is much smaller than the initial number of terms, where from the advantage of
735: the frame. This is the advantage of treating nonlinear problems with a basis that
736: has a scale criterion.
737:
738: Another application of the KAK basis occurs when one needs to understand the dynamics of
739: the initial data for the K(2,2) equation. Many numerical simulations show that
740: in the case when the width of the initial data is larger that $L_{compacton}$,
741: the initial shape decomposes into a finite number of compactons having the same width and
742: different amplitudes \cite{compacton1,fred,rosenau2,rosenau3}.
743: If we take such an initial pulse, Fig. 5a, and we want to see its time evolution untill the
744: break up process, we have first to expand it into the KAK frame, Fig. 5a.
745: Then we just let the system develope and the different KAK's move with their correspon
746: velocity, eq.(23), providing the new shapes, Fig. 5b and 5c, at different moments. We notice a
747: good agreement between these theoretical calculation and the above quoted numerical experiments.
748: At this stage we can not yet predict the way that the KAK breaks back into compactons. However,
749: we notice that the area of such structures in the phase space is the Poincare invariant and
750: should give a hint towards the breaking process. This invariant is nothing that the $u^3$
751: (denoted
752: $D_2$ invariant in [4]) invariant of the K(2,2) equation.
753:
754:
755:
756:
757: \section{Comments and conclusions}
758:
759: First, we make a general statement concerning the compact solutions of one-dimensional NPDE.
760: A one-dimensional dynamical model is described by a general NPDE equation
761: $\partial _t u ={\cal O}(x,{\partial }_{x} ) u$
762: where ${\cal O}$ is a nonlinear differential operator. By taking into account {\it
763: only} traveling solutions, this NPDE reduces to a nonlinear ordinary differential equation in
764: the coordinate $y =x-Vt$ for an arbitrary velocity
765: $V$. If $u(y )$ is a compact supported solution it follows that it
766: is not unique with respect to fixed initial compact data. Indeed, if we choose the initial
767: data such that the function and its derivatives (up to the
768: requested order) are zero in the neighborhood of a certain point $y _0$ of the $y $ axis,
769: these conditions can be fulfilled by any translated version of
770: a compact supported particular solution, placed everywhere
771: on the axis outside this neighborhood. Consequently, for such initial data, the
772: solution is not unique. This result shows that the compact supported property of the initial
773: data and of the solution implies its non-uniqueness.
774:
775: Since we can transform the NODE into a nonlinear differential
776: system of order one
777: \begin{equation}
778: {\vec U}_y ={\vec F}(y ,
779: {\vec U}), \ \ \ {\vec U}=(u, u_y , ...) ,
780: \end{equation}
781: we can apply the fundamental theorem of existence and uniqueness to solutions
782: of eq.(28), for given initial data ${\vec U}(y _0 )={\vec U}_0$.
783: If the function ${\vec F}$ in eq.(28) fulfills the Lipschitz condition (its relative
784: variation is bounded) than, for any initial condition, the solution is unique \cite{hure}.
785: Since any linear function is analytic and hence Lipschitz, we conclude
786: that only nonlinear functions ${\vec F}$ allow the existence of
787: compact supported solutions. Thus, a compact soliton
788: implies non-uniqueness in the underlying NPDE, which implies non-Lipschitzian structure
789: of the NPDE and hence the existence of nonlinear terms.
790:
791: In this paper we introduce new physical applications for
792: wavelets, that is the study of localized solutions of nonlinear partial differential
793: equations. The existence of compactons underlines a common feature of NPDE, discrete
794: wavelets, and also finite differences equations. We propose a new scale approach for
795: the similarity analysis and clasiffication of soliton solutions, without the need of solving
796: the corresponding NPDE. Also, we proved that starting from any unique soliton solution of a
797: NPDE, we can construct a frame of solutions organized under a multiresolution criterium.
798: This approach provides the possibility of constructing a
799: nonlinear basis for NPDE. We show that frames
800: of self-similar functions are related with
801: solitons with compact support. In addition, we notice the evidence
802: that compactons fulfil both characteristics of solitons
803: and wavelets, suggesting possible new applications.
804: Such a unifying direction between nonlinearity and self-similarity, can bring
805: new applications of wavelets in cluster formation, at any scale, from
806: supernovae through fluid dynamics to atomic and nuclear systems. The scale approach
807: can be applied with success to the physics of droplets,
808: bubbles, patterns, fragmentation, fission and
809: fusion.
810:
811: \vskip 1cm
812: Supported by the U.S. National Science Foundation through a regular grant,
813: No. 9970769, and a Cooperative Agreement, No. EPS-9720652, that includes
814: matching from the Louisiana Board of Regents Support Fund.
815:
816:
817:
818: \section{Appendices}
819:
820:
821: \subsection{Appendix 1}
822:
823:
824: We base our proof on the similar proposition in \cite{ijmpe}, excepting that here
825: we use Gaussian filtering instead of Morlet.
826: We start form the the discrete wavelet expansion of the signal $u(s)$
827: given in eq.(2)
828: $$
829: u(s)=\sum_{j}\sum_{k}C_{j,k}\Psi (2^j s -k)=\sum_{j,k}C_{j,k}\Psi _{j,k}(s),
830: $$
831: in terms of integer translations ($k$) and dyadic dilations ($2^j$) of the $\Psi$
832: wavelet. In order to reduce the number of scales needed, the range of the summation
833: should be choosen as a function of $s$.
834: We use in the following the asymptotic formula describing the pointwise
835: behavior of the Gaussian wavelet series around a point $s_0$ of interest \cite{approx}.
836: For a chosen $s_0$ and scale $j$, there is only one $k$ and $|\epsilon | \leq 1$
837: such that the support of the corresponding $\Psi _{j,k}$ contains this point,
838: $k=2^j s_0 +\epsilon$. We can express the solution and its derivatives in a neighborhood
839: of this point
840: $$
841: u(s_0) \approx \Psi (-\epsilon )\sum_{j}C_{j,2^j s_0 +\epsilon}\equiv
842: \sum_j u_j (s_0 ) ,
843: $$
844: \begin{equation}
845: u_{x}(s_0) \approx -i \Psi (-\epsilon )\sum_{j}2^j C_{j,2^j
846: s_0 +\epsilon}=-i \sum_j 2^j u_j (s_0 ) ,
847: \end{equation}
848: for $\epsilon $ being chosen enough large compared to one, and where $u_j (s_0 )=\Psi
849: (-\epsilon ) C_{j,2^j s_0 +
850: \epsilon}\approx \Psi (0) C_{j,2^j s_0 }$... Since the coefficient
851: $1/ 2^j
852: $ represents the scale for each $\Psi _{j, k}$ Gaussian wavelet, we can define it as
853: a characteristic half-width
854: $L_{j}$. Also, we finally have for the $n$-th order derivative in $s_0$
855: \begin{equation}
856: u_{xx... \ x}(s_0 )\approx \sum_{j} {{u_{j}(s_0 )}\over {L_{j}^{n}}}.
857: \end{equation}
858: Eq.(30) is the multi-scale generalization
859: of the simpler formula $ii$ in eq(1). With eq.(30) in hand we can investigate the structure of
860: hypothetic soliton solutions of NPDE, by choosing $s_0$ in the neighborhood of the maximum
861: value of the solution, $u(s_0)=A$. Around this maximum, such solutions can be described
862: very well by a unique scale $L$, and hence the solution and its derivatives can be
863: approximated with the corresponding dominant term, by the substitutions in eq.(1).
864:
865: \subsection{Appendix 2}
866:
867:
868: For the sake of simplicity we will renormalize the coefficients of the
869: K(2,2) equation such that the support of the simple compacton is one.
870: That is, we take $\eta _c (x,t)=\eta _{kak}(\pi (x-Vt),0)$ on the interval $|x-Vt|$
871: in $[-1/2,1/2]$. We construct a multiresolution approximation of $L^{2}({\bf R})$, that is an
872: increasing sequence of closed subspaces $V_j$, $j\in {\bf Z}$, of $L^{2}({\bf
873: R})$ with the following properties \cite{wavelet2,wavelet3}
874: \begin{enumerate}
875: \item
876: The $V_j$ subspaces are all disjoint and their union is dense in
877: $L^{2}({\bf R})$.
878:
879: \item
880: For any function $f \in L^{2}({\bf R})$ and for any integer $j$ we have
881: $f(x) \in V_j $ if and only if $D^{-1}f(x) \in V_{j-1}$ where $D^{-1}$ is an
882: operator that will be defined later.
883:
884: \item
885: For any function $f \in L^{2}({\bf R})$ and for any integer $k$, we have
886: $f(x) \in V_0 $ is equivalent to $f(x-k) \in V_0$.
887:
888:
889: \item
890: There is a function $g(x) \in V_0$ such that the sequence $ g(x-k)$ with
891: $k\in {\bf Z}$ is a Riesz basis of $V_0$.
892:
893: \end{enumerate}
894: In the case of compact solutions of K(2,2) of unit length, we chose for the
895: space $V_0$ that which is generated by all translation of $\eta _c$ with any
896: integer $k$. The subspaces $V_j$ for $j \geq 0$ are generated by all integer
897: translations of the compressed version of this function, namely, by
898: $\eta_{kak} (2^{j}\pi(x-Vt),0)$. The subspaces $V_j$ for $j\leq 0$
899: are generated by all integer translations of the KAK solution of length $\lambda 2^j -1$. For
900: example, $V_{-1} $ is generated by $\eta_{kak} (\pi (x-2^{j}Vt),0)$. The spaces $V_j, j\geq
901: 0$ are all solutions of K(2,2); the others are not. The function $g(x)$ is taken to be
902: $\eta_{kak}
903: (\pi (x-Vt),0)$. It is not difficult to prove that these definitions fulfill restrictions
904: one, three, and four. As for the second criterion, we define the action of the operator
905: $D^{-1} f(x) = f(2x)$ if $f(x) \in V_{j}$ with a $j$ positive integer, and
906: $D^{-1}\eta_{kak} (\pi2^{j}(x-2^{j}Vt),2^{-j}-1)=\eta_{kak} (\pi 2^{j}(x-2^{-j+1}Vt), 2^{-j+1}-1
907: )$ for negative
908: $j$. In conclusion, we construct a frame of functions made of contractions of compactons
909: and sequences of KAK solutions. We can write the corresponding two-scale
910: equation which connects the subspaces
911: \begin{equation}
912: \eta_{kak} (\pi (x-Vt), 1)=\eta_{kak} (\pi (x-Vt),0)+\eta_{kak} (\pi (x-Vt-1),0).
913: \end{equation}
914: We will denote generically by $\eta _{k,j}$ the elements of this frame, that is
915: \begin{equation}
916: \eta _{k,j} (x)=\eta_{kak} (\pi (x-2^{j}Vt-k), 2^{j}-1)|_{t=0},
917: \end{equation}
918: where $t=0$ means that we neglect the time evolution, but the amplitude is still amplified with
919: a factor of $2^j$, in virtue of relation $\eta_{max}=4V/3$.
920:
921:
922:
923:
924: \vfill
925: \eject
926:
927: \begin{thebibliography}{99}
928: %1
929: \bibitem{general1}
930: E. Alfinito, {\it et al}, eds., {\it Nonlinear Physics, Theory and
931: Experiment} (World Scientific, Singapore, 1996);
932: M. Remoissenet, {\it Waves called Solitons} (Springer-Verlag, Berlin, 1999).
933: %2
934: \bibitem{chaos}
935: L. P. Kadanoff, {\it From Order to Chaos} (World Scientific, Singapore, 1993).
936: %3
937: \bibitem{multiple}
938: J. U. Brackbill and B. I. Cohen, eds., {\it Multiple Time Scales}
939: (Academic Press, Inc., Orlando, 1985); C. Etrich, U. Peschel, F. Lederer and
940: B. A. Malomed, {\it Phys. Rev. E} {\bf 55} (1997) 6155;
941: %4
942: \bibitem{compacton1}
943: P. Rosenau and J. M. Hyman, {\it Phys. Rev. Let.} {\bf 70}
944: (1993) 564; B. Dey, {\it Phys. Rev. E} {\bf 57} (1998) 4733.
945: %5
946: \bibitem{wavelet1}
947: I. Doubechies and A. Grossmann, {\it J. Math. Phys.}{\bf \ 21} (1980) 2080;
948: A. Grossman and J. Morlet, {\it SIAM J. Math. Anal.} {\bf 15}
949: (1984) 72;
950: %6
951: \bibitem{wavelet2}
952: G. Kaiser, {\it A Friendly Guide to Wavelets} (Boston, Birkhauser, 1994);
953: A. Ludu and J. P. Draayer, in {\it Group 22. Proceedings of The XXII Int. Colloq. on Group Th.
954: Methods in Phys.}, Eds. S. P. Corney, R. Delbourgo and P. D. Jarvis
955: (Int. Press, Cambridge, Massachusetts, 1999).
956: %7
957: \bibitem{optics}
958: J. R. Taylor, ed.
959: {\it Optical Solitons - Theory and Experiment} (Cambridge University Press,
960: Cambridge, 1992).
961: %8
962: \bibitem{molecular}
963: A. S. Davydov, {\it Solitons in Molecular Systems} (Reidel, Dordrecht, 1990).
964: %9
965: \bibitem{nuclear}
966: A. Chodos, ed., {\it Solitons in Nuclear and Elementary Particle Physics} (World Scientific,
967: Singapore, 1984);
968: A. Ludu, A. Sandulescu and W. Greiner, {\it J. Phys. G: Nucl. Part.} {\bf 21} (1995) L41;
969: V. Kartavenko and W. Greiner, {\it Int. J.Mod. Phys. E}, 7 (1998) 287.
970: %10
971: \bibitem{particle}
972: R. Rajaraman, {\it Solitons and Instantons} (North-Holland, Amsterdam, 1984)
973: %11
974: \bibitem{prl}
975: A. Ludu and J. P. Draayer, {\it Phys. Rev. Lett.} {\bf 10} (1998) 2125;
976: J. M. Lina and M. Mayrand, {\it Phys. Rev. E} {\bf 48} (1994) R4160.
977: %12
978: \bibitem{bona}
979: J. L. Bona, {\it et al}, {\it Contemp. Math.} {\bf 200} (1996) 17 and
980: {\it Phil. Trans. R. Soc. Lond. A} {\bf 351} (1995) 107;
981: G. Zimmermann, Proceedings {\it Int. Conf. on Group Theoretical Methods in Physics,
982: G22} (Hobart, 10-18 July, 1998) in press and private communication.
983: %13
984: \bibitem{fred}
985: F. Cooper, J. M. Hyman and A. Khare, {\it Compacton Solutions in a Class of
986: Generalized Fifth Order KdV Equations} in press.
987: %14
988: \bibitem{wavelet3}
989: C. K. Chui and A. Cohen, in {\it Approximation Theory VII}, E. W.
990: Cheuey, C. K. Chui and L. L. Schumaker, eds. (Academic Press, Boston, 1993);
991: D. Han, Y. S. Kim and M. E. Noz, {\it Phys. Lett. A} {\bf 206} (1996) 299.
992: %15
993: \bibitem{electronic}
994: C. N. Kumar and P. K. Panigrahi, Preprint {\it solv-int/9904020}.
995: %16
996: \bibitem{rosenau1}
997: P. Rosenau, {\it Phys. Lett. A} {\bf 211} (1996) 265 and {\it Phys. Rev. Lett}
998: {\bf 73} (1994) 1737.
999: %17
1000: \bibitem{rosenau2}
1001: J. M. Hyman and P. Rosenau, {\it Physica D} {\bf 123} (1999) 502.
1002: %18
1003: \bibitem{rosenau3}
1004: P. Rosenau, {\it Physica D} {\bf 123} (1999) 525; {\bf 8D} (1983) 273.
1005: %19
1006: \bibitem{kart}
1007: V. G. Kartavenko, A. Ludu, A. Sandulescu and W. Greiner, {\it Int. J. Mod. Phys. E}
1008: {\bf 5} (1996) 329.
1009: %20
1010: \bibitem{bose1}
1011: M. H. Anderson {\it et al} {\it Science} {\bf 269} (1995) 198.
1012: %21
1013: \bibitem{bose2}
1014: Th. Busch and J. R. Anglin, {\it Phys. Rev. Lett.}, {\bf 84} (2000) 2298.
1015: %22
1016: \bibitem{oscillon}
1017: P. B. Umbanhowar and H. L. Swinney, {\it Nature} {\bf 382} (1996) 793;
1018: %23
1019: \bibitem{simil}
1020: V. I. Karpman, {\it Phys. Lett.} {\bf A 210} (1996) 77;
1021: M. I. Weinstein, {\it Comm. Math. Phys.} {\bf 87} (1983) 567;
1022: A. Khare and F. Cooper, {\it Phys. Rev. E} {\bf 48} (1993) 4853;
1023: B. Dey, C. N. Kumar and A. Khare, {\it Phys. Lett. } {\bf A 223} (1996) 449.
1024: %24
1025: \bibitem{hure}
1026: W. Hurewicz, {\it Lectures on Ordinary Differential Equations} (M.I.T. Press, Cambridge,
1027: Massachusetts, 1985) pp. 5-10.
1028: %25
1029: \bibitem{ijmpe}
1030: A. Ludu, G. Stoitcheva and J. P. Draayer, {\it Int. J. Mod. Phys. E} in print.
1031: % 26
1032: \bibitem{approx}
1033: N. N. Reyes, {\it J. Approx. Th. } {\bf 89} (1997) 89.
1034:
1035:
1036:
1037: \end{thebibliography}
1038: \vfill
1039: \eject
1040:
1041: %*****************************************************************
1042:
1043:
1044: \vskip 1cm
1045: \centerline{Figure Captions}
1046: \vskip 1cm
1047: \begin{itemize}
1048:
1049: \item
1050: Fig. 1
1051:
1052: Examples of compactons, kink-antikink pairs, and mixed solutions of the K(2,2) equation,
1053: together with their velocities.
1054:
1055:
1056:
1057: \item
1058: Fig. 2
1059:
1060: A finite series of K(2,2) compactons emerging from initial compact data of width
1061: larger than of the compacton.
1062:
1063: \item
1064: Fig. 3
1065:
1066: The half-width $L$ versus velocity $V$ for the K(2,2) equation, for different amplitudes A.
1067: Widths larger than $L_{compacton}=4$ behave different than
1068: narrower widths, with $L<4$. Amplitude increases from left to right, in the range
1069: 0.01-0.85.
1070:
1071: \item
1072: Fig. 4
1073:
1074: The half-width $L$ versus amplitude $A$, for the third ($n=3$), forth ($n=4$) and fifth
1075: ($n=5$) order NLS equation, in two $V=\pm A$ cases. We notice that the higher order ($n>3$) NLS
1076: equations have bifurcations.
1077:
1078: \item
1079: Fig. 5
1080:
1081: The dynamical evolution of the expansion in the KAK basis of a finite initial pulse of width $L$
1082: much larger than a compacton, drawn for three moments if time.
1083:
1084: \end{itemize}
1085:
1086: \vfill
1087: \eject
1088:
1089: \begin{table}[t]
1090: \caption{Traveling localized solutions for nonlinear dispersive equations.\label{tab:exp}}
1091: \vspace{0.2cm}
1092: \begin{center}
1093: \begin{tabular}{cccc}
1094: \hline \\
1095: NPDE & Analytic solution and the && OSA \\
1096: &relations among parameters & &approach\\
1097: \hline \\
1098:
1099: %2 ------------KdV ----------
1100: $u_{t}+6uu_{x}+u_{xxx}=0$ & $A~\hbox{sech} ^2 {{x-Vt} \over L};$ \ \ \ $L=\sqrt{2 / A}$, &&
1101: $L=|V\pm 6A|^{-1/2}$\\
1102: & $V=2A$ &&If $V\sim A$, $L\sim A^{-1/2}$ \\
1103: \hline \\
1104: %3 ---------MKdV-----------------------
1105: $u_{t}+u^2u_{x}+u_{xxx}=0$ & $A~\hbox{sech } {{x-Vt} \over L};$ \ \ \ $L=1/A$, & &
1106: $L=|V \pm 6A^2 |^{-1/2}$
1107: \\
1108: & $A=\sqrt{V}$ &&If $V\sim A^2 ,L\sim A^{-1}$\\
1109: \hline \\
1110: %4 -----------K(2,2)-----------------------
1111: $u_{t}+(u^2)_{x}+(u^2
1112: )_{xxx}=0$ & $A \hbox{cos} ^2 {{x-Vt}\over L}$, \ \ \hbox{if} \ \
1113: $|(x-Vt)/4|\leq \pi /2 $;
1114: & & $L=\biggl ( {{8A}\over
1115: {|V
1116: \pm 2A|}}\biggr ) ^{1/2}$ \\
1117: & L=4 && \\
1118: \hline \\
1119: %5 ----------K(n,n)-----------------------
1120: $u_t +(u^n )_x +(u^n )_{xxx}=0$ & $\biggl [ A cos^{2}\biggl ( {{x-Vt} \over {L}}
1121: \biggr ) \biggr ] ^{1 \over {n-1}}$, \ \ if $|x-Vt| \leq {{2n\pi}\over{n-1}}$ & &
1122: $L=\biggl ( {{n(n^2 +1)}\over{\alpha \pm n}}\biggr )^{1/2}$ \\
1123: &and 0 else; & & \\
1124: &$L={{4n}\over {(n-1)}},$ \ \ $A={{2Vn}\over {n+1}} $ && if $V=\alpha A^{n-1}$ \\
1125: \hline \\
1126: %6 ----------K(n,m)-----------------------
1127: $u_t +(u^n )_x +(u^m )_{xxx}=0$ & unknown
1128: & &
1129: $L=\biggl ({{{n(n^2 +1) A^{n-1}} \over {V \pm mA^{m-1}}}}\biggr )^{1/2}$
1130: \\
1131: $n\neq m$ &in general &
1132: & \\
1133: \hline \\
1134: \end{tabular}
1135: \end{center}
1136: \end{table}
1137:
1138: \vfill
1139: \eject
1140:
1141:
1142:
1143: \begin{table}[t]
1144: \caption{Traveling localized solutions for nonlinear diffusive equations.\label{tab:exp}}
1145: \vspace{0.2cm}
1146: \begin{center}
1147: \begin{tabular}{cccc}
1148: \hline \\
1149: NPDE & Analytic solution and the & OSA \\
1150: &relations among parameters & approach\\
1151: \hline \\
1152: %1 ------------linear -----------
1153: $u_{xx}-{1 \over {c^2}} u_{tt}=0$ & $\sum C_k e^{i(kx\pm \omega t )}$; &
1154: $V=c$ \\
1155: &$k^2 = \omega ^2 /c^2$ & $ A, \ L$ arbitrary \\
1156: \hline \\
1157: %1 ------------Burgers -----------
1158: $u_t + uu_x - u_{xx}=0$ & $ \sqrt{C -V^2} \hbox{tan}(\sqrt{C -V^2}{{x-Vt}\over {2}} +D)$
1159: & $L=(A \pm V)^{-1}$ \\
1160: & $+V$ & If $V\sim A$, $L\sim 1/A$\\
1161: \hline \\
1162: %1 ------------ full NL Burgers -----------
1163: $u_t +a (u^m )_x -\mu (u^k )_{xx}+cu^{\gamma}$ & only particular cases
1164: \ \ & $cA^{\gamma} L^2 + (V\pm amA^{m-1})L$ \\
1165: =0 &known & $\pm \mu k^2 A^{k-1}=0$ \\
1166: \hline \\
1167: %1 ------------ NL Burgers -----------
1168: $u_t +a (u^m )_x -\mu (u^k )_{xx}=0$ & $ -{\cal A} z^{\alpha }\ _{1}F_{1}(\alpha , \alpha +1 ,
1169: z) = x+x_0$
1170: \ \ &$L={{\mu k^2}\over{am-\alpha }}A^{k-m}$, \\
1171: & & if $ V=\alpha A^{m-1}$ \\
1172: \hline \\
1173: %7 ----------sine-G-----------------------
1174: $u_{xt}-\sin u=0$ & $A~\tan^{-1}\gamma ~e^{{x-Vt}\over {L}}$ \ \ \ &
1175: $\pm {{VA}\over {L^2 }}=sin A$ \\
1176: & & If $V=L^2, A=sinA$ \\
1177: \hline \\
1178: %8 ------------NLS(3) ----------
1179: $i\Psi_{t}+\Psi _{xx}+2|\Psi |^2 \Psi =0$ & $\eta_0 e^{i(\omega t + kx)} sech[ \eta_0
1180: (x-Vt)]$;
1181: \ \ & $L={{\pm V \pm \sqrt{|V^2 - 4 A^2 |}}\over {2A^2 }}$ \\
1182: &$ L=1/ \eta_0$& If $A\sim V, \ \ L=1/A$ & \\
1183: \hline \\
1184: %9 ------------NLS(n) ----------
1185: $i\Psi_{t}+\Psi _{xx}+|\Psi| ^{n-1}\Psi=0$ & unknown in general
1186: &$L={{\pm V \pm \sqrt{|V^2 - 4 A^n |}}\over {2A^n }}$ \\
1187: & & \\
1188: \hline
1189: %9 ------------GP ----------
1190: $i\Psi_t=-{1 \over 2}\triangle \Psi $ &
1191: $ip+\sqrt{v_{c}^{2}-p^2}\times $ &
1192: $L=(aA^2 \pm V -1)^{-1/2} $\\
1193: $+[a|\Psi|^2 +V(x) -1]\Psi$ & $\hbox{tanh}[a\sqrt{v_{c}^{2}-p^2}(x-q(t) )]$
1194: & If $V\sim \pm 1, L\sim 1/(A\sqrt{a})$ \\
1195: \hline
1196: \end{tabular}
1197: \end{center}
1198: \end{table}
1199:
1200: \vfill
1201: \eject
1202:
1203:
1204: \begin{table}[t]
1205: \caption{Traveling localized solutions for dissipative-dispersive equations.\label{tab:exp}}
1206: \vspace{0.2cm}
1207: \begin{center}
1208: \begin{tabular}{cccc}
1209: \hline \\
1210: The NPDE equation & OSA approach \\
1211: \hline \\
1212:
1213: %2 ----------------------
1214: $u_t + a(u^m ) _{x} + b (u^k )_{xx}+ c(u^n)_{xxx}=0$; &
1215: $L=A^{m-k} \cdot {{\mu k^2 \pm \sqrt{\mu ^2 k^4 -4mn^3
1216: (a-V_0 )}}\over {2m(a-V_0 )}}$ \\
1217: &if $V=mV_0A^{k-1}$ \\
1218: \hline \\
1219: %3 -------------------------------
1220: $Vu={{\alpha }\over {p+1}}u^{p+1}-\beta mu^{m-1}(u_y )^2 +2\beta (u^{m} u_y )_{y}$ &
1221: $2L^{l+4}((n+l+1)V_0 -\alpha) $\\
1222: $+{{\gamma n}\over {2}} u^{n-1} (u_y )^l (u_{yy})^ 2 -{{\gamma l} \over 2} (u^{n} (u_y
1223: )^{l-1}(u_{yy})^2 )_{y}
1224: $&$-2L^{l+2}(l+n+1)(l+n+2)\beta$ \\
1225: $+\gamma (u^n (u_y )^l u_{yy})_{yy} +C$& $-(l+n+1)(2+2n^2 +3l+l^2 +n(5+3l))\gamma$ \\
1226: &if $C=0$, $V=V_0 A^{m}$ and $m=p=n+l$ \\
1227: \hline \\
1228: %4 ---------------------------------
1229: $u_t +f(u)_x +g(u)_{xx}+h(u)_{xxx}=0$ & $L=-\biggl [ g'+Ag''\mp \biggl ( (Ag''+g')^2 $ \\
1230: & $ -4f'(1-V_0 )(A^2 h'''+3Ah''+h') \biggr ) ^{1/2} \biggr ] $ \\
1231: & $ \times (2A^2 h'''+6Ah''+2h')^{-1}$ \\
1232: & if $V=V_0 f'(A)$ \\
1233: \hline \\
1234: \end{tabular}
1235: \end{center}
1236: \end{table}
1237:
1238:
1239:
1240:
1241: \end{document}
1242:
1243: