1: \documentclass{elsart}
2: \usepackage[english]{babel}
3: \usepackage{graphicx,subfigure}
4: \usepackage{amsfonts,amsmath}
5: \usepackage{comment}
6: \graphicspath{{./Figsnew2/}}
7: \newcommand{\im}{\mathrm{i}}
8: \newcommand{\CC}{\mathcal C}
9: \newcommand{\co}{\cos(\psi_0)}
10: \newcommand{\so}{\sin(\psi_0)}
11: \begin{document}
12: \begin{frontmatter}
13: \title{Birhythmicity, Synchronization, and Turbulence in an Oscillatory
14: System with Nonlocal Inertial Coupling}
15: \author{Vanessa Casagrande\thanksref{kuramoto}}
16: and
17: \author{Alexander S. Mikhailov\corauthref{cor}\thanksref{kuramoto}}
18: \address{Abteilung Physikalische Chemie,\\
19: Fritz-Haber-Institut der Max-Planck-Gesellschaft,\\
20: Faradayweg 4-6, 14195 Berlin, Germany}
21: \corauth[cor]{Corresponding author}
22: \thanks[kuramoto]{Dedicated to Prof. Y. Kuramoto on the occasion of
23: his retirement}
24: \ead{mikhailov@fhi-berlin.mpg.de}
25: \begin{abstract}
26: We consider a model where a population of diffusively coupled
27: limit-cycle oscillators, described by the complex Ginzburg-Landau
28: equation, interacts nonlocally via an inertial field.
29: For sufficiently high intensity of nonlocal inertial coupling, the
30: system exhibits birhythmicity with two oscillation modes at largely
31: different frequencies. Stability of uniform oscillations in the
32: birhythmic region is analyzed by means of the phase
33: dynamics approximation. Numerical simulations show that, depending on
34: its parameters, the system has irregular intermittent regimes with
35: local bursts of synchronization or desynchronization.
36: \end{abstract}
37: \end{frontmatter}
38:
39:
40: \section{Introduction}
41: Beginning with the pioneering contributions by Kuramoto \cite{kurbook}
42: and Winfree \cite{winfree}, studies of synchronization and
43: spatiotemporal chaos (turbulence) in populations of coupled active
44: oscillators have developed into a classical field of research.
45: In chemical systems, each oscillator
46: represents a reaction element and coupling between such elements is
47: usually due to diffusion of reactants in the system. This coupling
48: extends only within a short diffusion length and is therefore local.
49: The complex Ginzburg-Landau equation is the canonical model for
50: oscillatory systems with local coupling
51: near a supercritical Hopf bifurcation. In surface chemical reactions
52: \cite{ertl}, reaction elements are however additionally interacting
53: through the gas phase where instantaneous complete mixing occurs.
54: As a result, global coupling between chemical oscillators arises
55: \cite{global}.
56: Moreover, global delayed feedbacks through the gas phase could be
57: artificially introduced to control turbulence in surface reactions
58: \cite{batt96,batt97,kawamura,science,bertram1,bertram2}.
59: A special class of systems are arrays of active oscillators that do not
60: directly interact one with another, but are all coupled to a single
61: diffusing field (in biochemistry, an example of such a system would be
62: provided by arrays of allosteric enzymes interacting through a chemical
63: messenger \cite{enzymes}).
64: Adiabatically eliminating this field, models with nonlocal
65: coupling between the oscillators, described by a finite-range integral,
66: were derived \cite{kur95,kur98}.
67: Investigations of nonlocally coupled oscillator
68: arrays have revealed that such form of coupling does not always
69: synchronize neighbouring oscillators and discontinuous distributions
70: with scaleless fractal structure can develop \cite{kur97,kur98}.
71: Spatiotemporal chaos can develop in such systems even in the
72: Benjamin-Feir stable regime \cite{batt00}
73: and spiral waves with phase-randomized cores can exist here
74: \cite{shima,chapter9}.
75:
76: Furthermore, arrays with both local and nonlocal coupling between
77: oscillators are possible. This situation is characteristic, for example,
78: for surface chemical reactions. In such systems, diffusion provides local
79: coupling between neighbouring surface oscillators, whereas much faster
80: heat conduction is responsible for nonlocal coupling between them.
81:
82: Recently, Tanaka and Kuramoto \cite{tanaka} have shown how, in the
83: vicinity of a supercritical Hopf bifurcation, the description of
84: arrays of nonlocally coupled oscillators can be reduced to the complex
85: Ginzburg-Landau equation with nonlocal coupling. Because of the critical
86: slowing down near the bifurcation point, the coupling is effectively
87: instantaneous in the reduced description.
88:
89: In realistic systems, which are not too close to the Hopf bifurcation, one
90: can however also encounter the opposite situation, with a very slow
91: inertial field giving rise to nonlocal coupling. The aim of our study is
92: to analyze what new effects, primarily due to slow nonlocal coupling, are
93: possible in this class of systems. For our investigations, we have chosen
94: an abstract model of the complex Ginzburg-Landau equation interacting with
95: an additional slow field, so that the coupling is nonlocal both with
96: respect to space and time. We expect that the behaviour found in this
97: general model would be typical for a broad class of realistic systems.
98:
99: The principal effect of coupling inertiality is that birhythmicity,
100: leading to chaotic intermittency, can develop in such systems. In
101: contrast to the birhythmicity in reaction-diffusion systems near a
102: pitchfork-Hopf bifurcation (see \cite{michael01,michael02}), the two
103: oscillatory states
104: are characterized here by very different frequencies and oscillation
105: amplitudes. For the rapid oscillatory state, the additional field
106: responsible for the inertial long-range coupling is almost absent
107: and only diffusive local coupling between the oscillators is
108: important. In slow oscillations, the elements are however entrained by
109: the long-range inertial field.
110:
111: Depending on the parameters, two new kinds of intermittency are found
112: here. When rapid oscillations are modulationally unstable and give rise to
113: turbulence, spatial regions occupied by slow entrained oscillations
114: spontaneously develop and die out in the medium. They can be interpreted
115: as bursts of synchronization on a turbulent background. On the other hand,
116: relatively small, irregularly evolving islands filled with rapid
117: turbulence can persist on the background of slow almost uniform
118: oscillations. They can be therefore described as bursts of
119: desynchronization on the background of regular slow oscillations. Our
120: analysis of such phenomena is based on the phase dynamics approximation,
121: derived for the birhythmic system. It is complemented by 1D and 2D
122: numerical simulations.
123:
124: \section{The Model}
125:
126: The investigated model describes a system of diffusively coupled active
127: oscillators coupled to an additional inertial diffusive field.
128: Introducing the local complex oscillation amplitude $\eta (x,t)$ and
129: denoting as $z(x,t)$ the additional complex-valued diffusive field,
130: we have therefore a system of two equations
131: \begin{subequations}
132: \begin{align}
133: &\dot\eta&=&\quad(1+\im\omega)\eta-(1+\im\alpha)|\eta|^2\eta+
134: (1+\im\beta)\nabla^2\eta+K(z-\eta)\label{model:one}\\
135: &\tau\dot z&=&\quad\eta-z+l^2\nabla ^2 z\label{model:two}
136: \end{align}\label{model}
137: \end{subequations}
138: The additional last term in the complex Ginzburg-Landau equation
139: (\ref{model:one}) takes into account coupling of the oscillatory
140: subsystem with the field $z$ whose evolution obeys equation
141: (\ref{model:two}); $K$ is the respective coupling constant.
142: The equations are brought into a dimensionless form by choosing
143: the characteristic diffusion length in the oscillatory subsystem as the
144: length unit and taking the characteristic relaxation time scale of the
145: oscillators as the time unit. The parameters $\tau$ and $l$ determine
146: characteristic time and length scales of the additional field $z$.
147: We assume that this field is inertial ($\tau \gg 1$) and slowly
148: varying in space ($l\gg 1$).
149:
150: The linear equation (\ref{model:two}) can easily be solved as
151: \begin{equation}
152: z(x,t)=\int_{-\infty }^{\infty }\int_{0}^{t}G(x-x^{\prime },t-t^{\prime
153: })\eta (x^{\prime },t^{\prime })dx^{\prime }dt^{\prime }.
154: \end{equation}
155: where the kernel is given by
156: \begin{equation}
157: G(x,t)=\frac{1}{2\tau \sqrt{\pi Dt}}\exp \left( -\frac{x^{2}}{4Dt}-
158: \frac{t}{\tau }\right) . \label{kernel}
159: \end{equation}
160: Substituting this into equation (\ref{model:one}), we obtain an
161: equivalent integro-differential form of the considered model,
162: \begin{eqnarray}
163: \dot\eta &=&(1+\im\omega)\eta-(1+\im\alpha)|\eta|^{2}\eta
164: +(1+\im\beta)\nabla^2\eta \nonumber\\
165: &&+K\int_{-\infty }^{\infty }\int_{0}^{t}G(x-x^{\prime },t-t^{\prime
166: })\left[ \eta (x^{\prime },t^{\prime })-\eta (x,t)\right] dx^{\prime
167: }dt^{\prime } \label{integralform}
168: \end{eqnarray}
169: We see that, besides of diffusion, the model also includes an
170: additional coupling, nonlocal both with respect to space and time.
171:
172: Note that very close to a supercritical Hopf bifurcation all processes
173: become fast as compared to the relaxation time scale of individual
174: oscillators, because this time is inversely proportional to the
175: distance from the bifurcation point (critical slowing down).
176: However, it is known
177: that the complex Ginzburg-Landau equation yields qualitatively correct
178: description even relatively far from the bifurcation.
179: Equations (\ref{model}) and (\ref{integralform}) can be viewed as
180: providing a simple model of an oscillating system coupled to an
181: inertial diffusive field.
182:
183:
184: \section{Birhythmicity}
185:
186: The system described by equations (\ref{model}) is birhythmic,
187: i.e. it can have two different oscillatory uniform states.
188: Assuming that
189: $\eta (t)=\rho e^{-\im\gamma t}$ and $z(t)=re^{-\im\gamma t}$ and
190: substituting this into equations (\ref{model:one}) and
191: (\ref{model:two}), we obtain a
192: cubic equation for the oscillation frequency $\gamma $,
193: \begin{equation}
194: \tau ^{2}\gamma ^{3}+\tau ^{2}(\omega -\alpha +\alpha K)
195: \gamma ^{2}+(1+\tau
196: K)\gamma +\omega -\alpha =0. \label{3rd_order}
197: \end{equation}
198:
199: When the frequency $\gamma$ is known, the oscillation amplitude
200: $\rho$ of the field $\eta$ is given by
201: \begin{equation}
202: \rho ^{2}=1-\frac{K\tau ^{2}\gamma ^{2}}{1+\tau ^{2}\gamma ^{2}}.
203: \label{ampl}
204: \end{equation}
205: The respective oscillation amplitude of the field $z$ is
206: \begin{equation}
207: r=\frac{\rho}{\sqrt{1+\tau ^{2}\gamma ^{2}}}.
208: \end{equation}
209: Depending on its parameters, equation (\ref{3rd_order}) can have
210: either one or three real roots. The three roots correspond to three
211: possible modes of uniform oscillations with frequencies
212: $\gamma_{1,2,3}$, such that $\gamma_{1}<\gamma_{2}<\gamma_{3}$.
213: It can be checked that oscillations with the
214: middle frequency $\gamma_{2}$ are always unstable.
215: In contrast to this,
216: uniform oscillations with frequencies $\gamma_{1}$ and $\gamma_{3}$
217: are possible (but may still be unstable with respect to nonuniform
218: perturbations, see the discussion below).
219:
220: Figure \ref{ak_bif} shows the region in the parameter planes
221: ($\alpha-\omega,K$) and ($\tau,K$) where birhythmicity exists.
222: The birhythmicity is possible only for sufficiently strong coupling
223: $K$. Moreover, it develops only if the
224: characteristic time $\tau$ of the field $z$ is sufficiently large
225: (Fig. \ref{tk_bif}).
226:
227: \begin{figure}
228: \begin{center}
229: \subfigure[]{\scalebox{0.4}[0.4]
230: {\label{ak_bif}\includegraphics*{delta_border_alpha.eps}}}
231: \subfigure[]{\scalebox{0.4}[0.4]
232: {\label{tk_bif}\includegraphics*{delta_border_tau2.eps}}}
233: \caption{Birhythimicity regions (gray) for the model (\ref{model})
234: in the parameter
235: planes $(\alpha-\omega ,K)$ and $(\tau ,K).$ The fixed parameters are
236: $\tau=10$ in the first diagram and $\omega =2,\alpha =2.5$ in the
237: second diagram.
238: The boundaries of the displayed regions do not depend on the parameter
239: $\beta$ of the model.}
240: \label{bif}
241: \end{center}
242: \end{figure}
243:
244: Numerical solutions of equation (\ref{3rd_order}) are displayed in Fig.
245: \ref{bif_diag}. If the parameter $\alpha$ is kept constant and the
246: coupling strength $K$ is varied (Fig. \ref{bif_k}), birhythmicity
247: is found inside the interval $0.4<K<1$.
248: When $K\ge 1,$ the system has a stable stationary state coexisting
249: with oscillations. The dependence of the effective oscillation
250: frequency $\gamma$ on the difference $\alpha-\omega$ for $K=0.5$ is
251: shown in Fig.\ref{bif_a}.
252: Two stable limit cycles coexist within the interval $-0.1<\alpha-\omega
253: <0.8$.
254: \begin{figure}
255: \begin{center}
256: \subfigure[]{\scalebox{.45}[.45]{\label{bif_k}
257: \includegraphics*{bifurcation_k.eps}}}
258: \subfigure[]{\scalebox{.45}[.45]{\label{bif_a}
259: \includegraphics*{bifurcation_a.eps}}}
260: \subfigure[]{\scalebox{.45}[.45]{\label{bif_t}
261: \includegraphics*{bifurcation_tau.eps}}}
262: \caption{Bifurcation diagrams. Frequencies $\gamma$ of uniform
263: oscillations, together with the respective amplitudes $\rho =|\eta|$
264: and $r=|z|$ as functions of $K$, $\alpha -\omega $ and $\tau $ for
265: slow (bold line), rapid (dashed) and absolutely unstable (dotted)
266: uniform oscillation modes. The parameters are (a) $\omega =2,
267: \alpha =2.5,\tau =10$ (b) $\omega =2,K=0.5,\tau =10,$ (c)
268: $\omega =2,\alpha =2.5,K=0.5.$}
269: \label{bif_diag}
270: \end{center}
271: \end{figure}
272:
273: The difference between the two limit cycles becomes clear if we compare
274: the amplitudes of oscillations of the fields $\eta$ and $z$ which are
275: presented below in the same figures. As we have already observed,
276: at $K=0$ only one oscillating mode is present, as we are in this limit
277: reduced to the standard CGLE.
278: Thus, we can expect that the branch starting at $K=0$ (which is the
279: upper one in the bifurcation diagrams of Fig. \ref{bif_diag})
280: would have the strongest similarities to the CGLE, even though for
281: finite $K$ it would be affected by the coupling to the second field.
282: This branch is characterized by a large amplitude of the field $\eta$,
283: which never gets significantly smaller than 1.
284: The corresponding frequency is quite small, starting from $\gamma =0.5$
285: at $K=0$ and slowly decreasing as $K$ increases.
286: Below, we refer to this attractor as corresponding to the {\em slow\/}
287: limit cycle.
288:
289: In contrast to this, oscillations in the lower branch have a smaller
290: amplitude $\rho$, which decreases for higher coupling strengths $K$ and
291: eventually vanishes at $K=1$. Note that $\gamma$ is negative for this
292: mode and therefore such oscillations are counter-rotating with respect
293: to the upper branch. The frequency varies significantly for this mode.
294: Inside the birhythmicity interval, the frequency of such {\em rapid\/}
295: limit cycle is always much larger (in its magnitude) than the frequency
296: of the slow mode. A characteristic feature of the rapid limit cycle is
297: that the amplitude $r$ of the field $z$ is very small here.
298: This is because the oscillations of the
299: field $\eta$ are so rapid here that the inertial field $z$ cannot
300: respond to them. Hence, the coupling term $Kz$ in the system
301: (\ref{model}) is almost negligible for the rapid limit cycle so that
302: the nonlocal coupling is ineffective and oscillators interact only via
303: the local diffusional coupling in this mode.
304:
305: \section{The Phase Dynamics Approximation}
306:
307: Above, uniform oscillations in the model have been discussed.
308: Now, we want to consider the properties of {\em almost\/} uniform
309: oscillations, characterized by small phase gradients.
310: Because the system is invariant with
311: respect to the uniform shift of all phases, the evolution of phase
312: distributions with small phase gradients should be slow, in contrast
313: to fast local relaxation of the amplitude perturbations.
314: Therefore, a reduced description, known as the phase dynamics
315: approximation, can be constructed in such cases \cite{kurbook}.
316: When birhythmicity is present, the coefficients
317: of the phase dynamics equation are different for the two oscillation
318: modes.
319:
320: The derivation of the phase dynamics equation for this system is
321: lengthy and has been performed by using a computer program for
322: algebraic calculations ({\em Mathematica\/}).
323: Below in this section we show the scheme of derivation
324: and discuss the results. The complete derivation is presented in
325: Appendix A.
326:
327: We define the phases ($\phi$ and $\varphi$) and amplitudes
328: ($\rho$ and $r$) of complex fields $\eta$ and $z$ as
329: $\eta =\rho e^{\im\phi }$, $z=re^{\im\varphi}$.
330: It is convenient to use instead of $\phi $ and $\varphi $
331: the variables $\psi =\phi-\varphi $ and $\Theta =\phi+\varphi$.
332: Note that the phase sum $\Theta$ is the slow variable in this system.
333: The phases of the two fields are rigidly correlated in the uniform
334: state and their difference $\psi $ undergoes rapid relaxation,
335: similar to oscillation amplitudes $\rho$ and $z$.
336:
337: In terms of the new variables, the model (\ref{model}) takes the form
338: \begin{subequations}
339: \begin{eqnarray}
340: \dot\rho&=&\rho-\rho^3+K r \cos\psi-K\rho+\rho\left
341: (-\displaystyle\frac{\nabla\Theta^2}{4}+\displaystyle\frac{1}{2}
342: \beta\nabla^2\Theta\right)\\
343: \dot r&=&\displaystyle\frac{\rho}{\tau}\cos{\psi}-
344: \displaystyle\frac{r}{\tau}-\displaystyle\frac{l^2}{4\tau}r
345: \nabla\Theta^2\\
346: \dot\psi&=&-\omega+\alpha\rho^2-\left(\displaystyle\frac{k r}{\rho}
347: +\displaystyle\frac{\rho}{\tau r}\right)\sin\psi
348: +\displaystyle\frac{\beta}{4}\nabla\Theta^2
349: +\displaystyle\frac{1}{2}\left(1-
350: \displaystyle\frac{l^2}{\tau}\right)\nabla^2\Theta\\
351: \dot\Theta&=&-\omega+\alpha\rho^2-\left(\displaystyle\frac{k r}{\rho}
352: -\displaystyle\frac{\rho}{\tau r}\right)\sin\psi+\displaystyle
353: \frac{\beta}{4}\nabla\Theta^2+\displaystyle\frac{1}{2}\left(
354: 1+\displaystyle\frac{l^2}{\tau}\right)\nabla^2\Theta
355: \end{eqnarray}\label{four}
356: \end{subequations}
357: Suppose now that the birhythmic system is close to one of its two
358: oscillatory states, characterized by some amplitudes $\rho_0$ and
359: $r_0$, and the relative phase $\psi_0$, whereas the phase sum
360: $\Theta$ remains arbitrary.
361: If the phase sum $\Theta$ is slowly varying in space,
362: the variables $\rho$, $r$ and $\psi $ would only slightly deviate from
363: their equilibrium values.
364: Therefore, we can introduce small perturbations
365: of such variables, $\rho =\rho_0+\delta \rho$, $r=r_0+\delta r$,
366: $\psi=\psi_0+\delta \psi $, and linearize the equations for with
367: respect to such perturbations.
368: The linearized evolution equation for the phase $\Theta $ is
369: \begin{eqnarray}
370: \dot\Theta&=&-\omega+\alpha\rho_0^2-\left(\frac{Kr_0}{\rho_0}-
371: \frac{\rho_0}{\tau r_0}\right)\sin\psi_0\nonumber\\
372: &+&\delta\rho\left[2\alpha\rho_0+\left(\frac{Kr_0}{\rho_0^2}
373: +\frac{1}{\tau r_0}\right)\sin\psi_0\right]\nonumber\\
374: &+&\delta r\left(-\frac{K}{\rho_0}-\frac{\rho_0}{\tau r_0^2}\right)
375: \sin \psi_0+\delta\psi\left(-\frac{Kr_0}{\rho_0}
376: +\frac{\rho_0}{\tau r_0}\right) \cos\psi_0\nonumber\\
377: &+&\frac{\beta}{4}\nabla\Theta^2+\frac{1}{2}\left(1
378: +\frac{l^2}{\tau}\right)\nabla^2\Theta
379: \label{sum}
380: \end{eqnarray}
381: The linearized equations for the variables $\delta \rho ,\delta r,
382: \delta\psi$ can be solved in the adiabatic approximation, because
383: these fast variables are adjusting to slow variation of $\Theta$.
384: Substituting the result into (\ref{sum}), a closed evolution equation
385: for the phase variable $\Theta $ is derived.
386: This equation has the form:
387: \begin{equation}
388: \dot\Theta=C_0+C_1(\nabla\Theta)^2+C_2\nabla^2\Theta
389: \end{equation}
390: The analytical expressions for the coefficients $C_0,$ $C_1$
391: are given in Appendix A. Note that inside the
392: birhythmic region there are two uniform oscillation modes with
393: different $\rho_0$, $r_0$, and $\psi _0$, and thus with
394: different values of these coefficients.
395:
396: The most important coefficient is $C_2$ because its sign controls
397: stability of uniform oscillations with respect to phase modulation
398: (the Benjamin-Feir instability).
399: If $C_2>0$, uniform oscillations are stable,
400: if $C_2<0$ they are unstable. We have computed this coefficient as
401: functions of several model parameters, by using the obtained analytical
402: expressions. The computed dependences are shown in Fig. \ref{coeff2}.
403: Two different branches correspond to the two different limit cycles.
404: The upper curves in Fig. \ref{coeff2}(a,b,d) are for the slow mode,
405: whereas the lower curves are for the
406: rapid oscillations. In Fig. \ref{coeff2}(c), the upper curves
407: correspond to the slow mode for larger values of the parameter $l$.
408: \begin{figure}
409: \begin{center}
410: \subfigure[]{\label{coeff2:k}\scalebox{.27}[.27]
411: {\includegraphics*{k_stability.eps}}}
412: \hspace{0.5cm}
413: \subfigure[]{\label{coeff2:a}\scalebox{.27}[.27]
414: {\includegraphics*{alpha_stability.eps}}}\\
415: \subfigure[]{\label{coeff2:l}\scalebox{.27}[.27]
416: {\includegraphics*{l_stability.eps}}}
417: \hspace{0.5cm}
418: \subfigure[]{\label{coeff2:t}\scalebox{.27}[.27]
419: {\includegraphics*{t_stability.eps}}}
420: \end{center}
421: \caption{
422: The coefficient $C_2$ of the phase dynamics equations as functions
423: of $K,\alpha ,l$ and $\tau $ for several different values of the
424: parameter $\beta $. The upper branches in (a,b,d) and for the larger
425: values of $l$ in (c) correspond to slow oscillations, the lower
426: branches are for the rapid oscillations. The oscillations are unstable
427: if $C_2<0$. The parameters are (a) $\alpha =2.5,$ $l=10$, $\tau =10,$
428: (b) $K=0.5,l=10,\tau =10,$ (c) $\alpha =2.5,$ $K=0.5,$ $\tau =10,$
429: (d) $\alpha =2.5,$ $K=0.5,$ $l=10;$ for all curves $\omega =2.$
430: }
431: \label{coeff2}
432: \end{figure}
433:
434: Examining these plots, several observations can be made: Slow mode
435: oscillations are always stable inside the birhythmic region, even when
436: the system without coupling is Benjamin-Feir unstable
437: (i.e., $1+\alpha \beta <0$ and therefore $C_2<0$ at $K=0$).
438: In contrast to this, rapid oscillations are always unstable when the
439: condition $1+\alpha \beta <0$ is satisfied.
440: Moreover, they are unstable near the boundary of the birhythmic region,
441: at which they first appear, even when this condition is violated.
442: Increasing the diffusion length $l$ of the field $z$ favors instability
443: of rapid uniform oscillations.
444:
445: \section{Numerical Simulations}
446:
447: To investigate nonlinear spatiotemporal dynamics in the considered
448: model, simulations of equations (\ref{model:one}) and (\ref{model:two})
449: have been performed. For numerical integration of these equations,
450: the fourth-order Runge-Kutta algorithm has been used.
451: The mesh size for space discretization and the time
452: step have been chosen to optimize the computational time for each
453: parameter choice. Since the diffusion length and the diffusion constant
454: of the oscillatory field have both been chosen to be equal to unity,
455: $\Delta x$ varied between 0.3 and 0.5, while $\Delta t$ could vary
456: between $(\Delta x)^{2}/5$ and $(\Delta x)^{2}/2$.
457: Both one- and two-dimensional systems were
458: investigated. No-flux boundary conditions were employed.
459: Initial conditions varied depending on a particular simulation.
460:
461: To display the results of one-dimensional simulations, space-time
462: diagrams have been constructed.
463: In such diagrams, time always runs along the
464: horizontal axis and the vertical axis is corresponding to the spatial
465: coordinate. In two dimensions, snapshots at some time moments and
466: space-time diagrams showing evolution of the pattern along some linear
467: cross-section are presented. The patterns are shown by using a gray
468: scale where the white color encodes the smallest and the black color
469: the largest values of the displayed variable.
470:
471: First, the behavior of fronts separating spatial regions with two
472: different oscillatory states in the birhythmic system has been studied
473: (Fig. \ref{front}).
474: The front travels towards the slow oscillating region, in such a way
475: that the system ends up with uniform oscillations of the rapid mode.
476: Its propagation is characterized by periodic appearance of amplitude
477: defects which occur when the phase difference between the two
478: oscillating regions equals $2\pi $ (see Fig \ref{front:a25}).
479: The front moves at a constant velocity which depends on the
480: parameters of the system. Fig. \ref{front_vel_a}
481: shows the dependence of the front velocity
482: on the parameter $\alpha $ for several values of the coupling strength
483: $K$. In Fig. \ref{front_vel_dom}, the dependence of the front
484: propagation velocity on the frequency difference
485: $\Delta \gamma =\gamma _{3}-\gamma _{1}$ of two uniform
486: oscillatory modes is presented.
487: The front always moves faster when this difference is increased.
488: \begin{figure}
489: \begin{center}
490: \subfigure[Re$(\eta)$]{\label{front:re25}\includegraphics*
491: [width=3cm,height=5cm, angle=-90]{front_real_k05_al25.eps}}
492: \hspace{0.5cm}
493: \subfigure[$|\eta|$]{\label{front:a25}\includegraphics*
494: [width=3cm,height=5cm,angle=-90]{front_ampl_k05_al25.eps}}
495: \caption{
496: Spatiotemporal diagrams showing front propagation in the birhythmic
497: system with $\omega =2,\alpha =2.5,\beta =1,K=0.5,l=10,$ and
498: $\tau =10.$
499: The total system length is $L=120$ and the total displayed time
500: interval is $T=300$.
501: Bright regions correspond to small values of displayed variables.
502: Time runs from left to right along the horizontal axis, the spatial
503: coordinate is varied in vertical direction.}
504: \label{front}
505: \end{center}
506: \end{figure}
507:
508: \begin{figure}
509: \begin{center}
510: \subfigure[]{\label{front_vel_a}{\scalebox{0.25}[0.25]
511: {\includegraphics*{frontvelocity_alpha.eps}}}}
512: \hspace{0.5cm}
513: \subfigure[]{\label{front_vel_dom}{\scalebox{0.25}[0.25]
514: {\includegraphics*{frontvelocity_dom.eps}}}}
515: \caption{Front velocity as functions of the parameter $\alpha $ and
516: difference $\Delta \gamma $ of oscillation frequencies in the two modes
517: in the birhythmic regime. The fixed parameters are $\omega =2,\beta =1,
518: l=10,$ and $\tau =10.$}
519: \label{front_vel}
520: \end{center}
521: \end{figure}
522:
523: As follows from the stability analysis of uniform oscillatory states
524: (see, e.g., Fig. \ref{coeff2:a}),
525: rapid oscillations should be unstable with respect to phase
526: modulation near the birhythmicity boundary and complex spatiotemporal
527: regimes can be expected there. We have performed a series of numerical
528: simulations, exploring pattern formation in the system within the
529: intervals $2.55<\alpha <2.8$ and $-1<\beta <2.5$, while keeping fixed
530: the other parameters ($\omega =2,K=0.5,l=10,$ and $\tau =10$).
531: Each simulation started from independently chosen, completely random
532: initial conditions.
533:
534: Figure \ref{phasediagram} is a schematic summary of these
535: numerical investigations. The solid
536: line is the stability boundary of rapid oscillations ($C_2=0$); rapid
537: oscillations are unstable below this boundary. Slow oscillations are
538: always stable in the considered region. Simulations have been performed
539: for a set of parameter values indicated by symbols in this diagram.
540: \begin{figure}
541: \begin{center}
542: \scalebox{.4}[.4]{\includegraphics*{phasediagram_new.eps}}
543: \caption{
544: Schematic phase diagram. Numerical simulations of the
545: one-dimensional model were performed at the values of $\alpha $ and
546: $\beta $ indicated by symbols in this diagram.
547: Depending on the observed properties of patterns, the diagram is
548: divided into regions 1 to 6. The circles indicate the values of these
549: parameters used to produce a typical patterns
550: for the corresponding region, displayed in Fig. \ref{patterns}.
551: The solid line shows the stability boundary
552: of rapid uniform oscillations, given by the condition $C_2=0$. Other
553: parameters are $\omega =2,K=0.5,l=10,$ and $\tau =10$.}
554: \label{phasediagram}
555: \end{center}
556: \end{figure}
557:
558: Comparing properties of observed patterns, we can qualitatively divide
559: them into several groups occupying different regions in the parameter
560: plane. Each group is marked by its own symbol and the respective
561: regions in Fig. \ref{phasediagram} are
562: numbered. Examples of typical spatiotemporal patterns, observed in
563: one-dimensional simulations within each region, are displayed in
564: Fig. \ref{patterns}.
565: These simulations were all performed for a system of the total length
566: $L=200$ and for a total time varying between 500 to 3000 time units.
567: To clearly present the observed patterns, only parts of their entire
568: evolution are however displayed.
569: \begin{figure}
570: \begin{tabular}{cccc}
571: & Re$(\eta)$ & $|\eta|$ & $|z|$\\
572: \begin{minipage}[b]{1ex}
573: \textbf{1}\\\mbox{ }\\\mbox{ }
574: \end{minipage}
575: &{\includegraphics*[width=2.5cm,height=4cm,angle=-90,origin=rB]
576: {patt1re.eps}}
577: & {\includegraphics*[width=2.5cm,height=4cm,angle=-90,origin=rB]
578: {patt1ae.eps}}
579: & {\includegraphics*[width=2.5cm,height=4cm,angle=-90,origin=rB]
580: {patt1az.eps}}\\
581: \begin{minipage}[b]{1ex}
582: \textbf{2}\\\mbox{ }\\\mbox{ }
583: \end{minipage}
584: & {\includegraphics*[width=2.5cm,height=4cm,angle=-90,origin=rB]
585: {patt2re.eps}}
586: & {\includegraphics*[width=2.5cm,height=4cm,angle=-90,origin=rB]
587: {patt2ae.eps}}
588: & {\includegraphics*[width=2.5cm,height=4cm,angle=-90,origin=rB]
589: {patt2az.eps}}\\
590: \begin{minipage}[b]{1ex}
591: \textbf{3}\\\mbox{ }\\\mbox{ }
592: \end{minipage}
593: & {\includegraphics*[width=2.5cm,height=4cm,angle=-90,origin=rB]
594: {patt3re.eps}}
595: & {\includegraphics*[width=2.5cm,height=4cm,angle=-90,origin=rB]
596: {patt3ae.eps}}
597: & {\includegraphics*[width=2.5cm,height=4cm,angle=-90,origin=rB]
598: {patt3az.eps}}\\
599: \begin{minipage}[b]{1ex}
600: \textbf{4}\\\mbox{ }\\\mbox{ }
601: \end{minipage}
602: & {\includegraphics*[width=2.5cm,height=4cm,angle=-90,origin=rB]
603: {patt4re.eps}}
604: & {\includegraphics*[width=2.5cm,height=4cm,angle=-90,origin=rB]
605: {patt4ae.eps}}
606: & {\includegraphics*[width=2.5cm,height=4cm,angle=-90,origin=rB]
607: {patt4az.eps}}\\
608: \begin{minipage}[b]{1ex}
609: \textbf{5}\\\mbox{ }\\\mbox{ }
610: \end{minipage}
611: & {\includegraphics*[width=2.5cm,height=4cm,angle=-90,origin=rB]
612: {patt5re.eps}}
613: & {\includegraphics*[width=2.5cm,height=4cm,angle=-90,origin=rB]
614: {patt5ae.eps}}
615: & {\includegraphics*[width=2.5cm,height=4cm,angle=-90,origin=rB]
616: {patt5az.eps}} \\
617: \begin{minipage}[b]{1ex}
618: \textbf{6}\\\mbox{ }\\\mbox{ }
619: \end{minipage}
620: & {\includegraphics*[width=2.5cm,height=4cm,angle=-90,origin=rB]
621: {patt6re2.eps}}
622: & {\includegraphics*[width=2.5cm,height=4cm,angle=-90,origin=rB]
623: {patt6ae2.eps}}
624: & {\includegraphics*[width=2.5cm,height=4cm,angle=-90,origin=rB]
625: {patt6az2.eps}}\\
626: \begin{minipage}[b]{1ex}
627: \textbf{7}\\\mbox{ }\\\mbox{ }
628: \end{minipage}
629: & {\includegraphics*[width=2.5cm,height=4cm,angle=-90,origin=rB]
630: {patt7re.eps}}
631: & {\includegraphics*[width=2.5cm,height=4cm,angle=-90,origin=rB]
632: {patt7ae.eps}}
633: & {\includegraphics*[width=2.5cm,height=4cm,angle=-90,origin=rB]
634: {patt7az.eps}} \\
635: \end{tabular}
636: \caption{Spatiotemporal diagrams displaying evolution of
637: Re($\eta$), $\rho=|\eta| $ and $r=|z| $ in typical patterns
638: observed in the regions 1-6 for the one-dimensional system.
639: The respective parameter values are given in Fig. \ref{phasediagram}.
640: The displayed space and time intervals are (1)
641: $L=100,$ $T=166,$ (2) $L=100,$ $T=166,$ (3) $L=100,$ $T=250,$
642: (4) $L=200,$ $T=500,$ (5) $L=100,$ $T=250,$ (6) $L=200,$ $T=500,$
643: and (7) $L=200,$ $T=250.$
644: The contrast level is adjusted individually in each plot to ensure best
645: visualization of the pattern.}
646: \label{patterns}
647: \end{figure}
648:
649: In region 1, the slow limit cycle is Benjamin-Feir stable, while the
650: rapid limit cycle is unstable. The system is found here in the regime
651: of fully developed amplitude turbulence characterized by creation of
652: multiple amplitude defects. Increasing parameter $\beta$, we enter a
653: region 2 closer to the stability boundary where turbulence becomes
654: intermittent. We observed the emergence of larger groups of
655: synchronized oscillators which become able to reach the rapid limit
656: cycle and perform harmonical oscillations for several periods.
657: Inside the small parameter region 3 near the stability boundary,
658: the frequency and the amplitude of oscillations correspond almost
659: everywhere to those of the rapid limit cycle.
660: However, the system does not undergo
661: complete synchronization. Lines of amplitude defects travelling
662: through the system act as wave sources here.
663:
664: In regions 5 and 7, both lying above the stability boundary of rapid
665: oscillations, uniform oscillations are observed starting from random
666: initial conditions.
667: The oscillations are rapid inside region 5 and slow inside
668: region 7. These two domains are separated by regions 4 and 6 in the
669: parameter plane, where competition between two oscillation mode takes
670: place leading to complex spatiotemporal regimes.
671:
672: The patterns inside region 4 are characterized by a background of rapid
673: chaotic oscillations, with small amplitude and numerous amplitude
674: defects.
675: On this highly desynchronized background, {\em bursts of
676: synchronization\/} emerge. Such bursts consist of large groups of
677: elements which suddenly start to oscillate together with a large
678: amplitude and a small frequency corresponding to the stable slow
679: limit cycle.
680: However, these groups do not keep synchronized over a long time: after
681: less than one oscillation period the turbulence overwhelms again.
682: As already noticed in our discussion of
683: birhythmicity, the amplitude of the coupling field $z$ is very small
684: in the rapid limit cycle, while in the slow limit cycle it gets a
685: larger value comparable to the amplitude of $\eta $. Analyzing patterns
686: in region 4, it can be seen that during the synchronization bursts,
687: the amplitude $|z|$ is indeed much larger than in the turbulent
688: background. Note that such patterns
689: with synchronization bursts have also been observed for some parameter
690: values below the stability boundary of rapid uniform oscillations.
691:
692: Region 6 lies near region 7, where slow uniform oscillations are
693: winning the competition.
694: Here, the patterns can be described as exhibiting {\em bursts
695: of desynchronization\/} on the background of slow uniform oscillations.
696: Inside such bursts, the coupling field $z$ is strongly decreased in
697: amplitude and only short-range diffusive coupling between oscillators
698: is effective.
699:
700: To further illustrate this regime, phase portraits of the patterns in
701: region 6 were constructed at different time moments. Fig.
702: \ref{phaseportr_snap} shows several subsequent snapshots from the
703: recorded video.
704: Each point in a phase portrait represents a state of one of the
705: oscillators. The spatial information (i.e., spatial distances between
706: the displayed oscillators) is lost in this
707: representation, but the motions performed by the oscillator population
708: are seen more clearly. The two circles indicate the two coexisting
709: limit cycles.
710: The cycle with a smaller oscillation amplitude is rapid, whereas the
711: cycle with a larger amplitude is slow.
712: \begin{figure}
713: \begin{center}
714: \scalebox{0.5}[0.5]{\includegraphics*{phaseportr_snap_ax2.eps}}
715: \caption{Selected snapshots of the phase portraits for the pattern 6 in
716: Fig. \ref{patterns}.
717: The circles show two different limits cycles in the birhythmic system.}
718: \label{phaseportr_snap}
719: \end{center}
720: \end{figure}
721:
722: We can see from these snapshots that the system is usually divided into
723: two groups of oscillators, occupying the two coexisting attractors
724: (limit cycles). Some oscillators are found in the vicinity of the
725: origin $\eta =0$.
726: They correspond to amplitude defects generated at the border between
727: spatial regions with different oscillation frequencies (bright dots in
728: the spatiotemporal plot of $|\eta|$ for pattern 6 in Fig.
729: \ref{patterns}). Such defects are not present at all times, and the
730: respective points are not, for example, seen in the phase portraits at
731: $t=16.7$ and $t=281.7$.
732: The oscillators sitting on the rapid limit cycle with the smaller
733: amplitude belong to the desynchronization bursts.
734:
735: For selected parameter values, two-dimensional simulations of the
736: system have additionally been performed.
737: Figure \ref{patt3_2d} displays the behavior of a
738: two-dimensional system at the same parameters as for the pattern 3 in
739: Fig. \ref{patterns}.
740: The space-time diagrams in the right panels show the pattern
741: development along a horizontal cross section. Starting from random
742: initial conditions, a population of rotating spiral waves develops.
743: After a transient, a stable configuration of rotating spirals is
744: reached in two dimensions.
745:
746: Figure \ref{patt4_2d} shows two-dimensional patterns corresponding
747: to synchronization bursts (pattern 4 in Fig. \ref{patterns}).
748: Inside such a burst, occupying for example the
749: left central region in the left panels, the coupling field $z$ is
750: increased in magnitude. The space-time diagrams (right panels) reveal
751: that the synchronized regions with slow oscillations have only
752: relatively short lifetimes and are replaced by irregular rapid
753: oscillations. However, they appear again and again in the course of
754: time.
755: \begin{figure}
756: \begin{center}
757: \begin{tabular}{ccc}
758: \begin{minipage}[b]{2em}
759: Re$(\eta)$\\\mbox{ }\\\mbox{ }
760: \end{minipage}
761: &{\includegraphics[width=2.5cm]{patt3re2d.eps}}
762: & {\includegraphics[width=2.5cm,height=6cm,angle=-90,origin=rB]
763: {patt3_cut_re.eps}}\\
764: \begin{minipage}[b]{2em}
765: $|\eta|$\\\mbox{ }\\\mbox{ }
766: \end{minipage}
767: &{\includegraphics[width=2.5cm]{patt3ae2d.eps}}
768: & {\includegraphics[width=2.5cm,height=6cm,angle=-90,origin=rB]
769: {patt3_cut_ae.eps}}\\
770: \begin{minipage}[b]{2em}
771: $|z|$ \\\mbox{ }\\\mbox{ }
772: \end{minipage}
773: &{\includegraphics[width=2.5cm]{patt3az2d.eps}}
774: & {\includegraphics[width=2.5cm,height=6cm,angle=-90,origin=rB]
775: {patt3_cut_az.eps}}\\
776: \end{tabular}
777: \caption{
778: Multiple spiral waves in a two-dimensional system of size $120\times
779: 120;$ the same parameter values as for the pattern 3 in
780: Fig. \ref{patterns}. The right
781: panels are space-time plots showing evolution of the pattern along one
782: horizontal cross section; the displayed time interval is $T=480$}
783: \label{patt3_2d}
784: \end{center}
785: \end{figure}
786: \begin{figure}
787: \begin{center}
788: \begin{tabular}{ccc}
789: \begin{minipage}[b]{2em}
790: Re$(\eta)$\\\mbox{ }\\\mbox{ }
791: \end{minipage}
792: &{\includegraphics*[width=2.5cm]{patt4re2d.eps}}
793: & {\includegraphics*[width=2.5cm,height=6cm,angle=-90,origin=rB]
794: {patt4_cut_re.eps}}\\
795: \begin{minipage}[b]{2em}
796: $|\eta|$\\\mbox{ }\\\mbox{ }
797: \end{minipage}
798: &{\includegraphics*[width=2.5cm]{patt4ae2d.eps}}
799: & {\includegraphics*[width=2.5cm,height=6cm,angle=-90,origin=rB]
800: {patt4_cut_ae.eps}}\\
801: \begin{minipage}[b]{2em}
802: $|z|$ \\\mbox{ }\\\mbox{ }
803: \end{minipage}
804: &{\includegraphics*[width=2.5cm]{patt4az2d.eps}}
805: & {\includegraphics*[width=2.5cm,height=6cm,angle=-90,origin=rB]
806: {patt4_cut_az.eps}}\\
807: \end{tabular}
808: \caption{Bursts of synchronization in a two-dimensional system of size
809: $120\times 120;$ the same parameter values as for the pattern 4 in
810: Fig. \ref{patterns}.
811: The right panels are space-time plots showing evolution of the pattern
812: along one horizontal cross-section; $T=480$.}
813: \label{patt4_2d}
814: \end{center}
815: \end{figure}
816:
817: A behavior, which can be described as bursts of desynchronization, was
818: found in our two-dimensional simulations even outside of the
819: birhythmicity region, where only slow uniform oscillations are possible
820: (Fig. \ref{des_bursts_first} and \ref{des_bursts_second}).
821: We started here
822: with the initial condition, which is commonly used to generate rotating
823: spirals. A rotating spiral was indeed first formed. However, some
824: instability has then developed beginning from the its central region,
825: where the oscillation amplitude $z$ was decreased (see Fig.
826: \ref{des_bursts_first}).
827: The development of this instability has resulted
828: (Fig. \ref{des_bursts_second}) in complete destruction of the spiral
829: and the appearance of relatively small chaotic domains on the
830: background of almost uniform slow oscillations.
831: \begin{figure}
832: \begin{center}
833: \begin{tabular}{cccc}
834: \includegraphics*[width=2.5cm]{des_bursts1.eps}
835: &\includegraphics*[width=2.5cm]{des_bursts2.eps}
836: &\includegraphics*[width=2.5cm]{des_bursts3.eps}
837: &\includegraphics*[width=2.5cm]{des_bursts4.eps}\\
838: \includegraphics*[width=2.5cm]{des_bursts5.eps}
839: &\includegraphics*[width=2.5cm]{des_bursts6.eps}
840: &\includegraphics*[width=2.5cm]{des_bursts7.eps}
841: &\includegraphics*[width=2.5cm]{des_bursts8.eps}
842: \end{tabular}
843: \end{center}
844: \caption{Instability of a rotating spiral. Simulation for a
845: two-dimensional system of size $120\times 120$ with parameters
846: $\omega =2,\alpha =3,\beta =1,K=0.4,l=10,$ and $\tau =10$.
847: Subsequent snapshots of the field Re($\eta$) at times
848: $T$ = 2.8, 10.0, 15.2, 37.2, 96.8, 119.6, 130.0, and 139.2 are
849: presented.}
850: \label{des_bursts_first}
851: \end{figure}
852:
853: \begin{figure}
854: \begin{center}
855: \begin{tabular}{ccc}
856: \begin{minipage}[b]{2em}
857: Re$(\eta)$\\\mbox{ }\\\mbox{ }
858: \end{minipage}
859: &{\includegraphics*[width=2.5cm]{des_bursts_r_e.eps}}
860: & {\includegraphics*[width=2.5cm,height=6cm,angle=-90,origin=rB]
861: {des_bursts_cut_re.eps}}\\
862: \begin{minipage}[b]{2em}
863: $|\eta|$\\\mbox{ }\\\mbox{ }
864: \end{minipage}
865: &{\includegraphics*[width=2.5cm]{des_bursts_a_e.eps}}
866: & {\includegraphics*[width=2.5cm,height=6cm,angle=-90,origin=rB]
867: {des_bursts_cut_ae.eps}}\\
868: \begin{minipage}[b]{2em}
869: $|z|$ \\\mbox{ }\\\mbox{ }
870: \end{minipage}
871: &{\includegraphics*[width=2.5cm]{des_bursts_a_z.eps}}
872: & {\includegraphics*[width=2.5cm,height=6cm,angle=-90,origin=rB]
873: {des_bursts_cut_az.eps}}\\
874: \end{tabular}
875: \caption{
876: Bursts of desynchronization in a two-dimensional system of size
877: $120\times 120;$ continuation of the simulation presented in Fig.
878: \ref{des_bursts_first}.
879: The parameters are $\omega =2,\alpha =3,\beta =1,K=0.4,l=10,$ and
880: $\tau =10$; $T=720$}
881: \label{des_bursts_second}
882: \end{center}
883: \end{figure}
884:
885: Typical snapshots of spatial distributions of the variables Re$(\eta)$,
886: $|\eta|$ and $|z|$ in this pattern are displayed in the left panels
887: in Fig. \ref{des_bursts_second}.
888: Inside the domains, the coupling field $z$ is
889: reduced in magnitude and these small spatial regions are filled with
890: irregular rapid variations of the complex field $\eta$. Thus, they can
891: be classified as desynchronization bursts. The space-time diagrams
892: through one horizontal cross section of the medium (right panels in
893: Fig. \ref{des_bursts_second}) show
894: that such domains can spread or shrink, and travel through the medium.
895: The edges of these structures are marked by the appearance of amplitude
896: defects where $|\eta|$ is close to zero.
897:
898: Remarkably, if we keep fixed all other parameter values, but eliminate
899: diffusional coupling between the oscillators (the Laplacian term in
900: equation (\ref{model:one})), evolution starting from the same initial
901: conditions leads to
902: spiral waves with phase-randomized cores, similar to those which were
903: previously considered \cite{shima,chapter9}. Thus, the inclusion of
904: diffusive coupling has a strong effect on pattern formation in the
905: system.
906: Though patterns resembling spiral waves with phase-randomized core are
907: indeed initially developing in the diffusively coupled case, they are
908: unstable and, after a transient, lead to the development of
909: intermittent spatiotemporal regimes with desynchronization bursts.
910:
911: \section{Discussion}
912:
913: Our study can be viewed as an extension of a series of detailed
914: investigations of systems with nonlocal coupling between oscillators
915: performed by Y. Kuramoto with his coworkers. We have chosen the model
916: (\ref{model}), which belongs to the previously discussed class, and
917: focused our analysis on the effects of coupling inertiality and
918: diffusion in this model.
919: The principal effect of the inertiality in nonlocal coupling
920: is that the system becomes birhythmic and has uniform slow and rapid
921: oscillations as two different coexisting attractors.
922:
923: For rapid oscillations, the inertial field $z$ responsible for nonlocal
924: coupling is not excited and the system is essentially reduced to an
925: array of only diffusively coupled oscillators with amplitudes $\eta$.
926: In contrast to this, the nonlocal coupling field $z$ is involved in
927: slow oscillations and nonlocal coupling between oscillators is in
928: operation under these conditions.
929:
930: Our stability analysis based on the phase dynamics equation has shown
931: that, though slow uniform oscillations are always stable in the
932: considered model, rapid uniform oscillations may become modulationally
933: unstable near the birhythmicity boundary.
934: Traveling fronts, separating spatial regions with two different
935: oscillation modes, have been seen in our numerical simulations.
936: Note that the fronts separating oscillating and non-oscillating regions
937: have been recently observed in the vicinity of a subcritical
938: Hopf bifurcation \cite{coullet04}.
939:
940: Irregular intermittent spatiotemporal regimes of two different kinds
941: have been numerically observed. In the patterns with synchronization
942: bursts, the background is occupied by rapid chaotic oscillations.
943: On this background, short-living small domains with slow and rather
944: uniform oscillations spontaneously develop.
945: Inside such domains, the nonlocal coupling field is
946: much larger in magnitude than for the rapidly oscillating background.
947: The characteristic size of such domains was comparable to the
948: nonlocality radius in the model. This kind of intermittent turbulence
949: is to some extent similar to the behavior observed for a system near
950: the Andronov homoclinic bifurcation, exhibiting bistability between
951: a stationary and an oscillatory state \cite{coullet97,coullet98}.
952: However, the developing domains are filled here not by a stationary
953: state, but by slow oscillations.
954:
955: In the spatiotemporal patterns with desynchronization bursts, the
956: background is occupied by slow oscillations which are almost uniform.
957: On this background, a cascade of desynchronization bursts develops.
958: Each burst represents a localized turbulent spot filled with rapid
959: irregular oscillations.
960: Inside it, the nonlocal coupling field is reduced in
961: magnitude. Such patterns correspond to the regimes of intermittent
962: localized turbulence previously seen for the complex Ginzburg-Landau
963: equation with global coupling \cite{batt97}, in the experiments with
964: global delayed feedback in
965: surface chemical reactions \cite{science} and in the respective
966: realistic simulations \cite{bertram1}. In some cases, they
967: develop in the situations when rotating spiral waves with
968: phase-randomized cores have been seen \cite{shima,chapter9} in systems
969: where only nonlocal coupling was present.
970: Birhythmicity as a cause of weak turbulence has recently been
971: investigated in a biological model of glycolitic oscillations
972: \cite{batt04}.
973:
974: Stimulating discussions with Prof. Y. Kuramoto are gratefully
975: acknowledged.
976: We thank M. Stich for the discussion of questions related to
977: birhythmicity.
978:
979: \section*{Appendix A}
980: Starting from system (\ref{four}) we linearize the first three
981: equations
982: around the values $\rho_0, r_0, \psi_0$ of the uniform oscillations.
983: The system that we obtain can be written as
984: \begin{subequations}
985: \begin{eqnarray}
986: \dot{\delta\rho}&=&a_1\delta\rho+b_1\delta r+c_1\delta\psi
987: +d_1\nabla\Theta^2+e_1\nabla^2\Theta\\
988: \dot{\delta r}&=&a_2\delta\rho+b_2\delta r+c_2\delta\psi
989: +d_2\nabla\Theta^2\\
990: \dot{\delta\psi}&=&a_3\delta\rho+b_3\delta r+c_3\delta\psi
991: +d_3\nabla\Theta^2+e_3\nabla^2\Theta\\
992: \dot\Theta&=&a_{4}\delta\rho+b_{4}\delta r
993: +c_{4}\delta\psi+d_{4}\nabla\Theta^2+e_{4}\nabla^2\Theta+f_4
994: \label{linear}
995: \end{eqnarray}
996: \end{subequations}
997: where the coefficients are given in the following table
998: \begin{center}
999: \begin{tabular}{|l||l|}
1000: \hline
1001: $a_1 = 1-3 \rho_0^2-K$
1002: & $a_3 = 2\alpha\rho_0+\left[\frac{K r_0}{\rho_0^2}-r_0\frac{1}{\tau}
1003: \right]\so$\\
1004: \hline
1005: $b_1 = K\co$
1006: & $b_3 = \so\left[-\frac{K}{\rho_0}+\frac{\rho_0}{\tau r_0^2}\right]$\\
1007: \hline
1008: $c_1 = -K r_0\so$
1009: & $c_3 = \co\left[-K \frac{r_0}{\rho_0}+\frac{\rho_0}{\tau r_0}
1010: \right]$\\
1011: \hline
1012: $d_1 = -\frac{\rho_0}{4}$
1013: & $d_3 = -l^2\frac{r_0}{4\tau}$\\
1014: \hline
1015: $e_1 = \frac{\zeta}{2}\rho_0\beta$
1016: & $e_3 = \frac{1}{2}\left(1-\frac{l^2}{\tau}\right)$\\
1017: \hline\hline
1018: $a_2 = \frac{1}{\tau}\co$
1019: & $a_4 = 2\alpha\rho_0+\left[K\frac{r_0}{\rho_0^2}+\frac{1}
1020: {\tau\rho_0}\right]\so$\\
1021: \hline
1022: $b_2 = -\frac{1}{\tau}$
1023: & $b_4 = \so\left[-\frac{K}{\rho_0}-\frac{\rho_0}{\tau r_0^2}\right]$\\
1024: \hline
1025: $c_2 = -\frac{\rho_0}{\tau}\so$
1026: & $c_4 = \co\left[-K\frac{r_0}{\rho_0}+\frac{\rho_0}{\tau r_0}
1027: \right]$\\
1028: \hline
1029: $d_2 = -l^2\frac{r_0}{4\tau}$
1030: & $d_{4} = \frac{\beta}{4}$\\
1031: \hline
1032: $e_2 = 0$
1033: & $e_4 = \frac{1}{2}\left(1+\frac{l^2}{\tau}\right)$\\
1034: \hline\hline
1035: \multicolumn{2}{|c|}{$f_4 = -\omega+\alpha\rho_0^2+
1036: \left[-K\frac{r_0}{R_0}+\frac{\rho_0}{\tau r_0}\right]\so$} \\
1037: \hline
1038: \end{tabular}
1039: \end{center}
1040:
1041: Now, since we are in the approximation where $\rho, r, \psi$ adjust
1042: adiabatically to $\Theta$, we can assume
1043: $\dot{\delta\rho}=\dot{\delta r}=\dot{\delta\psi}=0$, so that we get
1044: from (\ref{linear})
1045: \begin{eqnarray}
1046: \delta \rho &=& \frac{b_3 c_2 d_1-b_2 c_3 d_1-b_3 c_1 d_2+b_1 c_3 d_2
1047: +b_2 c_1 d_3-b_1 c_2 d_3}
1048: {-a_3 b_2 c_1 + a_2 b_3 c_1 + a_3 b_1 c_2-a_1 b_3 c_2 + a_2 b_1 c_3
1049: - a_1 b_2 c_3}\nabla\Theta^2\nonumber\\
1050: & &+\frac{b_3 c_2 e_1-b_2c_3e_1+b_2c_1e_3-b_1c_2e_3}
1051: {-a_3 b_2 c_1 + a_2 b_3 c_1 + a_3 b_1 c_2-a_1 b_3 c_2 + a_2 b_1 c_3
1052: - a_1 b_2 c_3}\nabla^2\Theta\\
1053: \delta r &=& \frac{a_3c_2d_1-a_2c_3d_1-a_3c_1d_2+a_1c_3d_2+a_2c_1d_3
1054: -a_1c_2d_3}
1055: {a_3b_2c_1-a_2b_3c_1-a_3b_1c_2+a_1b_3c_2+a_2b_1c_3-a_1b_2c_3}
1056: \nabla\Theta^2\nonumber\\
1057: & &+\frac{a_3c_2e_1-a_2c_3e_1+a_2c_1e_3-a_1c_2e_3}
1058: {a_3b_2c_1-a_2b_3c_1-a_3b_1c_2+a_1b_3c_2+a_2b_1c_3-a_1b_2c_3}
1059: \nabla^2\Theta\\
1060: \delta\psi &=& \frac{a_3 b_2 d_1-a_2b_3d_1-a_3b_1d_2+a_1b_3d_2
1061: +a_2b_1d_3-a_1b_2d_3}
1062: {-a_3 b_2 c_1+a_2b_3c_1+a_3b_1c_2-a_1 b_3 c_2 -a_2 b_1 c_3
1063: + a_1 b_2 c_3}\nabla\Theta^2\nonumber\\
1064: & &+\frac{a_3 b_2 e_1-a_2 b_3 e_1+a_2 b_1 e_3-a_1 b_2 e_3}
1065: {-a_3 b_2 c_1+a_2b_3c_1+a_3b_1c_2-a_1 b_3 c_2 -a_2 b_1 c_3
1066: + a_1 b_2 c_3}\nabla^2\Theta.
1067: \end{eqnarray}
1068: These expressions can be put into the equation for $\Theta$ to get
1069: \begin{eqnarray}
1070: \dot\Theta=C_0+C_1(\nabla\Theta)^2+C_2\nabla^2\Theta
1071: \end{eqnarray}
1072: where
1073: \begin{eqnarray}
1074: C_0&=&f_4\\
1075: C_1&=&d_{4}+\left[c_4(a_3b_2d_1-a_2b_3d_1-a_3b_1d_2+a_1b_3d_2
1076: +a_2b_1d_3-a_1b_2d_3)\right.\nonumber\\
1077: & &-b_{4}(a_3c_2d_1-a_2c_3d_1-a_3c_1d_2+a_1c_3d_2+a_2c_1d_3
1078: -a_1c_2d_3)\nonumber\\
1079: & &\left.+a_{4}(b_3c_2d_1-b_2c_3d_1-b_3c_1d_2+b_1c_3d_2+b_2c_1d_3
1080: -b_1c_2d_3)\right]/\nonumber\\
1081: & &(-a_3b_2c_1+a_2b_3c_1+a_3b_1c_2-a_1b_3c_2-a_2b_1c_3+a_1b_2c_3)\\
1082: C_2&=&e_{4}+\left[c_{4}(a_3b_2e_1-a_2b_3e_1+a_2b_1e_3-a_1b_2e_3)
1083: \right.\nonumber\\
1084: & &-b_{4}(a_3c_2e_1-a_2c_3e_1+a_2c_1e_3-a_1c_2e_3)\nonumber\\
1085: & &\left.+a_{4}(b_3c_2e_1-b_2c_3e_1+b_2c_1e_3-b_1c_2e_3)\right]/
1086: \nonumber\\
1087: & &(-a_3b_2c_1+a_2b_3c_1+a_3b_1c_2-a_1b_3c_2-a_2b_1c_3+a_1b_2c_3)
1088: \end{eqnarray}
1089:
1090: % REFERENCES
1091: \begin{thebibliography}{9}
1092:
1093: \bibitem{kurbook}
1094: Y. Kuramoto,
1095: {\em Chemical Oscillations, Waves and Turbulence\/},
1096: (Springer, New York, 1984).
1097:
1098: \bibitem{winfree}
1099: A. T. Winfree,
1100: {\em J. Theor. Biol. \/} \textbf{16} (1967) 15.
1101:
1102: \bibitem{ertl}
1103: R. Imbihl and G. Ertl,
1104: {\em Chem. Rev.\/} \textbf{95} (1995) 697.
1105:
1106: \bibitem{global}
1107: G. Veser, F. Mertens, A. S. Mikhailov, and R. Imbihl,
1108: {\em Phys. Rev. Lett.\/} \textbf{71} (1993) 935-939.
1109:
1110: \bibitem{enzymes}
1111: A. S. Mikhailov and B. Hess,
1112: {\em J. Phys. Chem.\/} \textbf{100} (1996) 19059--19065.
1113:
1114: \bibitem{kur95}
1115: Y. Kuramoto,
1116: {\em Prog. of Theor. Phys.\/} \textbf{94} (1995) 321--330.
1117:
1118: \bibitem{kur96}
1119: Y. Kuramoto and H. Nakao,
1120: {\em Phys. Rev. Lett.\/} \textbf{76} (1996) 4352--4355.
1121:
1122: \bibitem{kur97}
1123: Y. Kuramoto and H. Nakao,
1124: {\em Physica D\/} \textbf{103} (1997) 294--313.
1125:
1126: \bibitem{kur98}
1127: Y. Kuramoto, D. Battogtokh and H. Nakao,
1128: {\em Phys. Rev. Lett.\/}\textbf{81} (1998) 3543--3546
1129:
1130: \bibitem{batt99}
1131: D. Battogtokh,
1132: {\em Prog. of Theor. Phys.\/} \textbf{102} (1999) 947--952.
1133:
1134: \bibitem{kur00}
1135: Y. Kuramoto, H. Nakao and D. Battogtokh,
1136: {\em Physica A\/} \textbf{288} (2000) 244--264.
1137:
1138: \bibitem{batt00}
1139: D. Battogtokh and Y. Kuramoto,
1140: {\em Phys. Rev. E\/} \textbf{61} (2000) 3227--3230.
1141:
1142: \bibitem{batt02}
1143: D. Battogtokh,
1144: {\em Phys. Lett. A\/} \textbf{299} (2002) 558--564.
1145:
1146: \bibitem{chapter9}
1147: Y. Kuramoto,
1148: in {\em Nonlinear Dynamics and Chaos: Where do we go from here?\/},
1149: edited by S.J. Hogan, A.R. Champneys, B. Krauskopf, M.d. Bernardo,
1150: R.E. Wilson, H.M. Osinga, and M.E. Homer (Institute of Physics,
1151: Bristol, 2002) Chapter: Reduction methods applied to nonlocally
1152: coupled oscillator systems 800--819.
1153:
1154: \bibitem{shima}
1155: S. Shima and Y. Kuramoto,
1156: {\em Phys. Rev. E} \textbf{69} (2004) 036213.
1157:
1158: \bibitem{tanaka}
1159: D. Tanaka and Y. Kuramoto,
1160: {\em Phys. Rev. E\/} \textbf{68} (2003) 026219.
1161:
1162: \bibitem{batt96}
1163: D. Battogtokh and A.S. Mikhailov,
1164: {\em Physica D\/} \textbf{90} (1996) 84.
1165:
1166: \bibitem{batt97}
1167: D. Battogtokh, A.S. Mikhailov, and A. Preusser
1168: {\em Physica D\/} \textbf{106} (1997) 327.
1169:
1170: \bibitem{kawamura}
1171: Y. Kawamura and Y. Kuramoto,
1172: {\em Phys. Rev. E\/} \textbf{69} (2004) 016202.
1173:
1174: \bibitem{michael01}
1175: M. Stich, M. Ipsen and A. Mikhailov,
1176: {\em Phys. Rev. Lett.} \textbf{86} (2001) 4406.
1177:
1178: \bibitem{michael02}
1179: M. Stich, M. Ipsen and A. Mikhailov,
1180: {\em Physica D} \textbf{171} (2002) 19.
1181:
1182: \bibitem{coullet97}
1183: M. Argentina and P. Coullet,
1184: {\em Phys. Rev. E\/} \textbf{56} (1997) R2359-R2362.
1185:
1186: \bibitem{coullet98}
1187: M. Argentina and P. Coullet,
1188: {\em Physica A\/} \textbf{257} (1998) 45--60.
1189:
1190: \bibitem{coullet04}
1191: P. Coullet and L. Kramer,
1192: {\em Chaos} \textbf{14} (2004) 244.
1193:
1194: \bibitem{batt04}
1195: D. Battogtokh and J.J. Tyson,
1196: {\em Phys. Rev. E\/} \textbf{70} (2004) 026212.
1197:
1198: \bibitem{science}
1199: M. Kim, M. Bertram, M. Pollmann, A. von Oertzen, A.S. Mikhailov,
1200: H.H. Rotermund, and G. Ertl,
1201: {\em Science\/} \textbf{292} (2001) 1357.
1202:
1203: \bibitem{bertram1}
1204: M. Bertram and A. S. Mikhailov,
1205: {\em Phys. Rev. E\/} \textbf{67} (2003) 036207.
1206:
1207: \bibitem{bertram2}
1208: M. Bertram, C. Beta, M. Pollmann, A. S. Mikhailov, H. H. Rotermund,
1209: and G. Ertl,
1210: {\em Phys. Rev. E\/} \textbf{67} (2003) 036208.
1211:
1212: \end{thebibliography}
1213:
1214: \end{document}
1215: