1: %\documentstyle[aps,multicol,prl,epsf]{revtex}
2: %\input{psfig}
3: %\renewcommand{\baselinestretch}{2.0}
4: %\draft
5:
6:
7: \documentstyle[aps,multicol,prl,epsfig]{revtex}
8:
9: \pagenumbering{arabic}
10:
11: \begin{document}
12: \begin{title}
13: {\bf
14: Feedback Loops Between Fields and Underlying Space Curvature: an Augmented
15: Lagrangian Approach}
16: \end{title}
17:
18: \author{P.G. Kevrekidis$^1$, F.L. Williams$^1$, A.R. Bishop$^2$,
19: I.G. Kevrekidis$^3$, and B.A. Malomed$^4$}
20: \address{$^1$ Department of Mathematics and Statistics,
21: University of Massachusetts, Amherst, MA 01003-4515, USA \\
22: $^2$
23: Theoretical Division, MS B210,
24: Los Alamos National Laboratory, Los Alamos, NM 87545, USA \\
25: $^3$ Department of Chemical Engineering, Princeton University,
26: 6 Olden Str. Princeton, NJ 08544 \\
27: $^4$ Department of Interdisciplinary Studies, Faculty of Engineering, Tel
28: Aviv University, Tel Aviv 69978, Israel \\
29: }
30: %\title{}
31: \maketitle
32:
33: \begin{abstract}
34: We demonstrate a systematic implementation of coupling between
35: a scalar field and the geometry of the space
36: (curve, surface, etc.) which carries the field. This naturally
37: gives rise to a feedback mechanism
38: between the field and the geometry. We develop a systematic model for the
39: feedback in a general form, inspired by a specific implementation in the
40: context of molecular dynamics (the so-called Rahman-Parrinello molecular
41: dynamics, or RP-MD). We use a generalized Lagrangian that allows for the
42: coupling of the space's metric tensor (the first fundamental form) to the
43: scalar field, and add terms motivated by RP-MD. We present two
44: implementations of the scheme: one in which the metric is only
45: time-dependent [which gives rise to ordinary differential equation (ODE) for
46: its temporal evolution], and one with spatio-temporal dependence [wherein
47: the metric's evolution is governed by a partial differential equation (PDE)].
48: Numerical results are reported for the (1+1)-dimensional model with a
49: nonlinearity of the sine-Gordon type.
50: \end{abstract}
51:
52: %\date{\today}
53:
54: \vspace{5mm}
55:
56: %\section{Introduction}
57:
58: \begin{multicols}{2}
59:
60: Recently, much attention has been focused on soft-condensed-matter objects,
61: such as vesicles, microtubules, and membranes \cite{av1,av2,gaid1,gaid2}.
62: Many nanoscale physical systems, including nanotubes and electronic and
63: photonic waveguide structures \cite{gaid3,gaid4}, have nontrivial geometry
64: and are influenced by substrate effects. These classes of systems, many of
65: which are inherently nonlinear, raise the question of the interplay between
66: nonlinearity and a substrate with variable curvature. Of particular interest
67: is a possibility of developing curvature in the substrate due to forces
68: generated by the nonlinear field. The resulting curvature can in turn affect
69: the field.
70:
71: There is an increasing body of literature dealing with the interplay of
72: nonlinearity and a curved substrate. Usually, however, the substrate
73: geometry is assumed to be {\it fixed}, see, e.g., \cite{curv}.
74: Nevertheless, for many applications, ranging from condensed matter to optics
75: to biophysics, it is relevant to introduce models that admit
76: a flexible substrate, which is affected by the field(s) that it carries, as
77: well as feeding back into the field dynamics. In this situation, equations for
78: the fields in a nonlinear system abutting on the flexible substrate should
79: include both the field dynamics proper and the feedback coupling to the
80: substrate. Equations for the evolution of the substrate should in turn be
81: affected by the evolution of the field. A prototypical physical example of
82: this type is Euler buckling \cite{antm}, where the evolution of a thermal
83: profile causes the underlying surface to buckle (and hence locally modify
84: its curvature).
85:
86: In a discrete setting, a model of this type has recently been presented in
87: \cite{kmb}. However, it was limited to a system of masses coupled by
88: nonlinear springs. Some studies have also been performed in a special case
89: of the continuum limit of classical spin systems (such as the Heisenberg
90: chain) coupled to the curvature; geometric frustration was found to arise in
91: such settings \cite{dand}.
92:
93: About twenty years ago, a problem similar to the theme of our study was
94: examined in the context of molecular dynamics studies of
95: structural transitions in crystals.
96: In particular, in a series of papers \cite{RP}, Rahman and
97: Parrinello introduced a new idea for studying such transitions by means of
98: an augmented Lagrangian that would account for the degrees of freedom of the
99: ``box'' (the cell) in which the MD particles lie. In
100: studying the time evolution of the box dynamics (naturally obtained through
101: the Euler-Lagrange equations of the augmented Lagrangian for the box degrees
102: of freedom), they were able to identify structural transitions (under
103: external shear) from square to hexagonal patterns, fcc to bcc etc. We will
104: hereafter refer to this technology as the RP-MD method. The relevant
105: Lagrangian for the particles and the box in this case reads:
106: \begin{equation}
107: L=\frac{1}{2}\sum_{i}m_{i}{\bf \dot{s}}_{i}G{\bf \dot{s}}_{i}-
108: \sum_{i,j>i}V(r_{ij})+\frac{1}{2}W{\rm Tr}(\dot{f}^{{\rm T}}\dot{f}),
109: \label{rpeq1}
110: \end{equation}
111: where $m_{i}$ is the mass of the $i$-th particle, ${\bf \dot{s}}_{i}$ is its
112: vectorial velocity, the spatial part $G$ of the spatiotemporal metric tensor
113: may be represented in terms of another matrix $f$ as $G=f^{{\rm T}}f$ ($G$
114: is positive definite). $^{{\rm T}}$ and ${\rm Tr}$ denote the transposition
115: and trace respectively,
116: $r_{ij}$ and $V$ are distances between the particles and the
117: potential of interaction between them, and $W$ is an effective mass of the
118: box.
119:
120: Our purpose in this work is to extend the RP-MD methodology
121: to the case of a continuum scalar field,
122: coupled to either a spatially averaged geometric
123: characteristic (``average curvature''), which will give rise to an ODE, or
124: to a spatiotemporal curvature field, that will generate a PDE.
125: The continuum field
126: may represent, e.g., a chemical concentration propagating over a membrane,
127: or a salt solution, causing the swelling of a polymer gel \cite{yy},
128: or an envelope
129: wave of the electric
130: field in nanosystems.
131: The
132: spatiotemporal metric is assumed to have the simple form,
133: \begin{equation}
134: g=\left(
135: \begin{array}{cc}
136: -1 & 0 \\
137: 0 & G
138: \end{array}
139: \right) . \label{rpeq2}
140: \end{equation}
141: We will first consider the general case, where $G$ is a $d\times d$ matrix,
142: $d$ being the space dimension.
143:
144: One can define a field-type generalization of the RP-MD model, with a scalar
145: field $\phi $, as follows:
146: \begin{equation}
147: L=\int d^{d}x\left[ -g^{ij}\frac{\partial \phi }{\partial x^{i}}\frac{
148: \partial \phi }{\partial x^{j}}-V(\phi )\right] +\frac{1}{2}W{\rm Tr}(\dot{f}
149: ^{{\rm T}}\dot{f}), \label{rpeq3}
150: \end{equation}
151: where $V(\phi )$ is a potential governing the nonlinear evolution of the field
152: $\phi $ and $f=f(t)$ (only).
153: If $f$ (and hence $G$) is a function of both spatial coordinates and
154: time, an elastic-energy term \cite{elastic} should be added to the Lagrangian
155: (\ref{rpeq3}), so that it becomes
156: \[
157: L=\int d^{d}x\left[ -g^{ij}\frac{\partial \phi }{\partial x^{i}}\frac{
158: \partial \phi }{\partial x^{j}}-V(\phi )\right] +
159: \]
160: \begin{equation}
161: \int d^{d}x\left\{ \frac{1}{2}W\left[ {\rm Tr}(\dot{f}^{{\rm T}}\dot{f})-
162: \frac{1}{2}{\rm Tr}(\frac{\partial f^{{\rm T}}}{\partial x_{i}}\frac{
163: \partial f}{\partial x^{i}})\right] \right\} . \label{rpeq4}
164: \end{equation}
165:
166: %This general formulation and can be adapted to the cases $d=1$, $2$ or $3$.
167: %Here we aim to put forward the simplest one-dimensional (1d) model, with the
168: %single component of the metric tensor, $G\equiv f^{2}$, where the scalar $f$
169: %can be a function of time $t$ only or of $x$ and $t$.
170:
171: An objective of this brief report is to propose two models that include the
172: feedback to the curvature, and, simultaneously, admit as particular
173: solutions the (unperturbed) solutions for the flat
174: (original) metric. In particular, we choose $G=1+f^2$
175: (so that the metric is positive definite and for $f=0$ has as a special
176: case the original Minkowskian metric). Notice a slight deviation in
177: our choice from the RP case of $G=f^2$ \cite{choice}. Although the
178: formulation of Eqs. (\ref{rpeq3})-(\ref{rpeq4}) is very general, we
179: hereafter focus on the $d=1$ case that we will examine
180: in more detail.
181:
182: Assuming initially that $f=f(t)$ only
183: (e.g., including only an effect of the ``mean
184: curvature'' on the scalar-field dynamics), the Lagrangian (\ref{rpeq3})
185: becomes
186: \begin{equation}
187: L=\frac{1}{2}W\dot{f}^{2}+\int dx\left[ \frac{1}{2}\left( \frac{\partial
188: \phi }{\partial t}\right) ^{2}-\frac{1 + f^{2}}{2}\left( \frac{\partial \phi }
189: {\partial x}\right) ^{2}-V(\phi )\right] . \label{rpeq6}
190: \end{equation}
191: %The equations of motion following from it are
192: %\begin{eqnarray}
193: %\phi _{tt} &=&f^{2}\phi _{xx}-\frac{\partial V}{\partial \phi }
194: %\label{rpeq7} \\
195: %Wf_{tt} &=&-f\int \phi _{x}^{2}dx, \label{rpeq8}
196: %\end{eqnarray}
197: %where the subscript stands for the corresponding derivative.
198: Then, the resulting equations of motion (to which we will hereafter refer
199: as model A) are
200: %we will refer to the following variation of Eqs. (\ref{rpeq7})-
201: %(\ref{rpeq8}) as a model A,
202: \begin{eqnarray}
203: \phi _{tt} &=&(1+f^{2})\phi _{xx}-\frac{\partial V}{\partial \phi }
204: \label{rpeq13} \\
205: Wf_{tt} &=&-f\int \phi _{x}^{2}dx, \label{rpeq14}
206: \end{eqnarray}
207: where the subscripts stand for the corresponding partial derivatives.
208:
209: Notice that the function $f$ is directly related to the
210: scalar curvature of the 1-d space. In particular, the Ricci scalar,
211: which is $R=2R_{1212}/det(g)$ \cite{weinberg}
212: in the general case, in the 1-d case
213: is $R=-2 \tilde{f}_{tt}/\tilde{f}$, where $\tilde{f}=\sqrt{1+f^2}$.
214: %where the
215: %relevant entry of the curvature tensor is given by:
216: %$$
217: %R_{1212}=\frac{1}{2}
218: %\left( \frac{\partial^2 g_{22}}
219: %{\partial t^2} - \frac{\partial^2 g_{22}}
220: %{\partial x^2} \right)-\frac{1}{4 g_{22}}
221: %\left[ \left(\frac{\partial g_{22}}{\partial t} \right)^2
222: %+\frac{\partial g_{11}}{\partial x}
223: %\frac{\partial g_{22}}{\partial x} \right]
224: %$$
225: %\begin{eqnarray}
226: %-\frac{1}{4 g_{11}}
227: % \left[ \left(\frac{\partial g_{11}}{\partial x} \right)^2
228: %+\frac{\partial g_{11}}{\partial t}
229: %\frac{\partial g_{22}}{\partial t} \right]=f f_{tt}
230: %\label{rpeq12}
231: %\end{eqnarray}
232: %and hence the scalar curvature is directly
233: %related to $f$ through $R=-2 f_{tt}/f$.
234:
235:
236: On the other hand, for a metric with both spatial and temporal
237: dependence (e.g., for $f=f(x,t)$),
238: one arrives at the following Lagrangian:
239: \[
240: L=\int dx\left[ \frac{1}{2}\left( \frac{\partial \phi }{\partial t}\right)
241: ^{2}-\frac{1 + f^{2}}{2}\left( \frac{\partial \phi }{\partial x}\right)
242: ^{2}-V(\phi )\right] +
243: \]
244: %\[
245: \begin{equation}
246: \int dx\left[ \frac{W}{2}\left( \frac{\partial f}{\partial t}\right) ^{2}-
247: \frac{W}{2}\left( \frac{\partial f}{\partial x}\right) ^{2}\right] .
248: \label{add1}
249: \end{equation}
250: %\]
251: The ensuing coupled equations for the scalar field and the curvature
252: (to which we will refer as model B) are
253: \begin{eqnarray}
254: \phi _{tt} &=&\left( (1+f^{2})\phi _{x}\right) _{x}-\frac{\partial V}
255: {\partial \phi } \label{rpeq15} \\
256: Wf_{tt} &=&Wf_{xx}-f\phi _{x}^{2}. \label{rpeq16}
257: \end{eqnarray}
258:
259: %\begin{eqnarray}
260: %\phi _{tt} &=&(f^{2}\phi _{x})_{x}-\frac{\partial V}{\partial \phi }
261: %\label{rpeq10} \\
262: %W\left( f_{xx}-f_{tt}\right) &=&f\phi _{x}^{2}. \label{rpeq11}
263: %\end{eqnarray}
264:
265: %It is then particularly interesting to examine the effects of the
266: %curvature for a specific nonlinear potential.
267: As a particular application of models A and B,
268: we examine the physically ubiquitous
269: sine-Gordon (sG) potential, $V(\phi )=1-\cos (\phi )$ \cite{eilbeck}. It is
270: clear that Eqs. (\ref{rpeq13})-(\ref{rpeq14}) and
271: (\ref{rpeq15})-(\ref{rpeq16}) have particular
272: solutions with $f=0$, for which the latter equation of each pair is satisfied
273: trivially, while the former reduces exactly to the sG equation. Basic
274: solitary-wave solutions of the sG equation are the topological soliton
275: (kink),
276: \begin{equation}
277: \phi _{{\rm k}}(x,t)=4\tan ^{-1}\left[ \exp \left( \gamma
278: (x-x_{0}-vt)\right) \right] , \label{kink}
279: \end{equation}
280: where $v$ is its velocity, $\gamma =(1-v^{2})^{-1/2}$ is the Lorentz factor,
281: and $x_{0}$ is the initial position of the kink's center, and the breather,
282: $\phi _{{\rm br}}(x,t)=4\tan^{-1} \{ \sqrt{(1-\omega ^{2})/\omega ^{2}}
283: \sin \left[ \omega \gamma (t-v(x-x_{0}))\right]
284: \times \\
285: {\rm sech}\left[
286: \gamma \sqrt{1-\omega ^{2}}(x-x_{0}-vt)\right] \} $, where $\omega $
287: is the frequency of its internal oscillations ($0<\omega <1$).
288:
289: %However, Eqs. (\ref{rpeq7}) and (\ref{rpeq8}), as well as (\ref{rpeq10}) and
290: %(\ref{rpeq11}), do not offer an appropriate model to study the coupling of
291: %the field dynamics to curvature -- first of all, because the flat-metric
292: %state (the one with $f={\rm const}$) containing a soliton is not a solution
293: %in this case. In fact, the kink and breather are bound to be destroyed in
294: %such a model. This occurs (in all the cases that we have examined, for both
295: %the kinks and breathers) through the formation of sharp gradients and the
296: %trapping of the solitons, while the geometrical factor $f$ develops
297: %oscillations around $f=0$ (see an example in Fig. \ref{rfig1}).
298:
299: The results of the interaction of the kink with the curvature in model A
300: are shown in Fig. \ref{rfig2} \cite{ww}.
301: The curvature variable $f$, initialized with
302: a small random value, performs smooth oscillations with a frequency of
303: $\omega \approx 2.866$. Notice that this is natural in this case, since the
304: kink has an approximately fixed ``mass'', $M_{k}=\int_{-\infty }^{+\infty
305: }u_{x}^{2}dx$, which can be found to be $M_{k}=8.247$ for the velocity $%
306: v\approx 0.25$). Then, Eq. (\ref{rpeq14}) predicts the frequency of these
307: oscillations $\omega \approx \sqrt{M_{k}}=2.871$, which is very close to the
308: above-mentioned numerically exact value. Fast small-amplitude oscillations
309: of the kink's velocity, observed in Fig. \ref{rfig2} both with and without
310: the curvature, are due to ``hopping'' over sites of a lattice (with spacing
311: $h=0.1$) employed in the numerical scheme which solves Eq. (\ref{rpeq14}).
312: Notice, however, that in the top panel the mean velocity is $\approx 0.2503$,
313: while in the bottom panel it is $\approx 0.2499$, hence the curvature
314: oscillations increase the kink's velocity. This may be anticipated due to
315: the presence of the positive definite factor $1+f^{2}$ in front of $\phi
316: _{xx}$ in Eq. (\ref{rpeq13}), which is expected to renormalize $v^{2}$.
317:
318: The curvature-breather interaction in model A is shown in Fig. \ref
319: {rfig3}. The frequency of the breather does not change significantly (it
320: fluctuates between $0.87$ and $0.93$), but its amplitude decreases
321: substantially (by more than $50\%$), resulting in its becoming much more
322: mobile (the velocity increases to $\approx 0.425$ from the initial value
323: $0.25$).
324:
325: In model B, we examine collision of a kink with a localized pulse of the
326: substrate field $f$. For the breather, we have obtained results which are
327: qualitatively similar to those presented below for the kink. We create the
328: curvature pulse to the left or to the right of the kink. As $\phi _{x}$
329: vanishes far from the kink, the equation (\ref{rpeq16}) for $f$ becomes a
330: linear wave equation. Hence, we observe splitting of the pulse into left-
331: and one right-traveling ones. In the case where the kink is initially to the
332: left of the pulse, it collides with the left-propagating fragment of
333: the (split) pulse. In the opposite case,
334: the kink is eventually caught by the co-propagating right fragment of
335: the (split) pulse. Numerical
336: simulations shown in Fig. \ref{rfig4} demonstrate that the collision with
337: the counter-propagating pulse reduces the kink's velocity, while the
338: interaction with the co-propagating pulse gives rise to an increase of the
339: velocity. In particular, in the former case, the mean speed of the kink
340: after the collision is $\approx 0.2476$, while in the latter one, it
341: increases to $\approx 0.2527$. Notice that in both cases a small fragment of
342: the pulse that collides with the kink passes through it, while a larger
343: fraction of the pulse is reflected by it. The latter feature may be
344: explained by a momentum-balance analysis.
345:
346: In conclusion, we have extended field theory in the spirit of
347: the Rahman-Parrinello
348: Molecular Dynamics technique. The resulting
349: equations couple the spatio-temporal evolution of the field to that of the
350: underlying curvature of the space which carries the field. Coupled equations
351: for the temporal or spatiotemporal evolution of the metric are obtained in a
352: general setting, and, as an example,
353: are solved together with the sine-Gordon field
354: equation. The purely temporal evolution of the metric has been found to
355: increase the velocity of the field solitons, while the model allowing
356: spatio-temporal evolution of the metric can induce both increase and
357: decrease of the velocity.
358:
359: It would be particularly interesting to extend the models A and B to higher
360: dimensions. It is also worth studying how the local evolution of the curvature
361: affects kinematics and dynamics of the solitons, and to correlate such
362: observations with the behavior of reactant chemical concentrations in
363: chemical or biological environments with non-trivial geometry \cite{yannis}.
364:
365: {\it Acknowledgements}: The authors acknowledge D. Maroudas for
366: stimulating discussions. The support of NSF and AFOSR (IGK) and
367: NSF (DMS-0204585), UMass and the Clay Institute (PGK) is gratefully
368: acknowledged. Work at Los Alamos is supported by the US DoE.
369:
370: \begin{references}
371: \bibitem{av1} E. Sackmann, in R. Lipowsky and E. Sackmann (Eds.), The
372: Structure and Dynamics of Membranes: Handbook of Biological Physics, vol. 1
373: (Elsevier, Amsterdam, 1995).
374:
375: \bibitem{av2} U. Seifert, \newblock Adv. in Phys. {\bf 46}, 13 (1997).
376:
377: \bibitem{gaid1} W. Saenger.
378: \newblock{\it Principles of Nucleic Acid
379: Structure} (Springer-Verlag: Berlin, 1984).
380:
381: \bibitem{gaid2} J. Meunier, D.
382: Langevin, and D. Boccara (Eds.),
383: {\it Physics of Amphiphilic Layers}, Springer-Verlag (Berlin, 1997).
384:
385: \bibitem{gaid3} Y. Imry. {\it Introduction to Mesoscopic Physics} (Oxford
386: University Press: New York, 1997).
387:
388: \bibitem{gaid4} C.M. Soukoulis
389: (Ed.), {\it Photonic Band Gaps and Localization}, Plenum Press
390: (New York, 1993).
391:
392: \bibitem{curv} See e.g., Yu. B. Gaididei, S. F. Mingaleev, and P. L.
393: Christiansen, Phys. Rev. E {\bf 62}, R53 (2000); M. Ibanes, J. M. Sancho,
394: and G. P. Tsironis, Phys. Rev. E {\bf 65}, 041902 (2002);
395:
396: \bibitem{antm} S.P. Timoshenko and S. Woinowsky-Krieger,
397: {\it Theory of plates and shells}, McGraw-Hill (New York, 1959)
398: S.S. Antman,
399: \newblock {\it Nonlinear problems of elasticity}, Springer-Verlag
400: (New York 1995).
401:
402: \bibitem{kmb} P.G. Kevrekidis, B.A. Malomed and A.R. Bishop, \newblock
403: Phys. Rev. E {\bf 66}, 046621 (2002).
404:
405: \bibitem{dand} S. Villain-Guillot, R. Dandoloff, A. Saxena and A.R. Bishop,
406: \newblock Phys. Rev. B {\bf 52}, 6712 (1995); R. Dandoloff, S.
407: Villain-Guillot, A. Saxena and A.R. Bishop, \newblock Phys. Rev. Lett. {\bf
408: 74}, 813 (1995).
409:
410:
411: \bibitem{RP} see e.g., M. Parrinello and A. Rahman, \newblock Phys. Rev.
412: Lett. {\bf 45}, 1196 (1980). M. Parrinello and A. Rahman, \newblock J. Appl.
413: Phys. {\bf 52}, 7182 (1981).
414:
415: \bibitem{yy} E. C. Achilleos, R. K. Prud'homme, I. G. Kevrekidis, K. N. Christodoulou and
416: K.R. Gee,
417: { AIChE Journal}, {\bf 46}, 2128 (2000).
418:
419: \bibitem{elastic} Elastic dynamics is a natural first approximation,
420: even though it is possible for the substrate to have more
421: complicated dynamical evolution laws (which can also be incorporated
422: in this setting).
423:
424: \bibitem{choice} We also ran detailed simulations of the case with
425: $G=f^2$. Here, the non-existence of the original steady state
426: of the field and its competition with the oscillatory behavior of $f$
427: around $f=0$ yields rather unphysical behavior involving large gradients
428: of the field. For these reasons, this case is not presented in detail here.
429:
430: \bibitem{eilbeck} R. K. Dodd, J. C. Eilbeck, J. D. Gibbon, and H. C.
431: Morris, Solitons and Nonlinear Wave Equations (Academic Press, London, 1982).
432:
433: \bibitem{weinberg} B. O'Neill, Semi-Riemannian Geometry with Applications
434: to Relativity (Academic Press, London, 1983).
435:
436: \bibitem{ww} In the numerical computations $W$ has been fixed to $W=1$.
437: By varying $W$, we have found, similarly to the original works of RP,
438: that its increase slows down the rate at which the presented
439: phenomenology occurs.
440:
441: \bibitem{yannis} J. Wolff, A.G. Papathanasiou, I.G. Kevrekidis, H.H.
442: Rotermund, and G. Ertl, Science {\bf 294}, 134 (2001).
443: \end{references}
444:
445: \end{multicols}
446:
447: %\centerline{{\bf Figure Captions}}
448:
449: %\begin{figure}[tbp]
450: %\centering
451: %{\epsfig{file=cn1.ps, width=6.4cm,angle=0, clip=}}
452: %\caption{Results of simulations of Eqs. (\ref{rpeq7}) and (\ref{rpeq8}) with
453: %the initial configuration in the form of the sine-Gordon kink (\ref{kink})
454: %with the velocity $v=0.25$ in the $\protect\phi$) field. The metric variable
455: %$f$ is initialized at a small random value. The top left and right panel
456: %show, respectively, the $\protect\phi$-field profile at $t=28.5$, and the
457: %actual position of the kink's center (solid line) vs. the position that it
458: %would have, moving at the constant velocity equal to the initial value.
459: %(dashed line). The bottom panel shows the time evolution of $f$.}
460: %\label{rfig1}
461: %\end{figure}
462:
463: \begin{figure}[tbp]
464: \centering
465: {\epsfig{file=rpfig1a.ps, width=6.4cm,angle=0, clip=}}
466: \centering
467: {\epsfig{file=rpfig1b.ps, width=6.4cm,angle=0, clip=}}
468: \caption{The top left panel shows the kink's spatial profile at $t=60$ in
469: model A. The initial kink profile (centered around the new center
470: position) is shown by the dashed line, and is practically indistinguishable
471: from the solution at $t=60$, indicating that the kink maintains its shape.
472: The bottom left panel shows smooth oscillations of $f(t)$. The top and
473: bottom right panels show, respectively, the kink's velocity $v(t)$ vs. $t$,
474: and the same quantity but for $f=0$.}
475: \label{rfig2}
476: \end{figure}
477:
478: \begin{figure}[tbp]
479: \centering
480: {\epsfig{file=rpfig2.ps, width=6.4cm,angle=0, clip=}}
481: \caption{The top left and right panels show, respectively, the breather in
482: model A at the end of the simulation period, $t=120$, and its position
483: vs. time (solid), as compared to that which it would have moving at the
484: initial velocity, $v=0.25$ (dashed). The middle left panel shows the time
485: evolution of $f(t)$, while the middle right and bottom left panels present
486: exchange between the curvature-mode's energy, $E_f= f_t^2/2 + (f^2/4) \int
487: \protect\phi_x^2 dx$, and the rest of the energy, $E_{\protect\phi}=E-E_f$.
488: Finally, the bottom right panel shows the time evolution of the field
489: $\protect\phi$ at the center of the breather.}
490: \label{rfig3}
491: \end{figure}
492:
493: \begin{figure}[tbp]
494: \centering
495: {\epsfig{file=rpfig3a.ps, width=6.4cm,angle=0, clip=}}
496: \centering
497: {\epsfig{file=rpfig3b.ps, width=6.4cm,angle=0, clip=}}
498: \centering
499: {\epsfig{file=rpfig3c.ps, width=6.4cm,angle=0, clip=}}
500: \centering
501: {\epsfig{file=rpfig3d.ps, width=6.4cm,angle=0, clip=}}
502: \caption{The left part of the figure shows spatial profiles of the fields
503: $\protect\phi$ and $f$ at $t=60$ in model B. It also shows the time
504: evolution of the energies $E_{\protect\phi}=E-E_f$ and $E_f=\int (1/2)
505: (f_t^2+f_x^2+ (f \protect\phi_x)^2/2 ) dx$, as well as the kink's velocity
506: as a function of time. The kink's center is initially at $x=90$, while the
507: center of the curvature pulse is at $x=110$. The right subplots show
508: counterparts of those in the left part, but for the case of the kink and
509: curvature pulse initially centered at $x=110$ and $x=90$, respectively.}
510: \label{rfig4}
511: \end{figure}
512:
513: \end{document}
514: