1: \documentclass[aps,prl,twocolumn,groupedaddress,showpacs,amsmath,
2: amssymb,amsfonts]{revtex4}
3: \usepackage{amsmath}
4: \usepackage{amsfonts}
5: \usepackage{amssymb}
6: \usepackage{bm}
7: \usepackage{graphicx}%
8: \def\eps{\varepsilon}
9: \def\goesto{\rightarrow}
10: \def\dirac{{\mbox{\boldmath$\delta$}}}
11: \begin{document}
12: \title{Spot deformation and replication in the two-dimensional
13: Belousov-Zhabotinski reaction in water-in-oil microemulsion}
14: \author{Theodore Kolokolnikov}
15: \author{Mustapha Tlidi}
16: \email[]{tkolokol@mathstat.dal.ca}
17: \email[]{mtlidi@ulb.ac.be}
18: \affiliation{Department of Mathematics and Statistics, Dalhousie
19: University, Halifax, Canada}
20: \affiliation{Optique Nonlin\'eaire Th\'eorique, Universit\'e Libre de Bruxelles, Campus Plaine CP 231,
21: 1050 Bruxelles, Belgium}
22:
23: \date{\today}
24: \begin{abstract}
25: In the limit of large diffusivity ratio,
26: spot-like solutions in the two-dimensional Belousov-Zhabotinski
27: reaction in water-in-oil microemulsion
28: are studied. It is
29: shown analytically that
30: such spots undergo an instability as the diffusivity ratio is
31: decreased. An instability threshold is derived. For spots of
32: small radius, it is shown that this instability leads to a spot splitting
33: into precisely two spots. For larger spots, it leads to deformation,
34: fingering patterns and space-filling curves.
35: Numerical simulations are shown to be in close agreement with the
36: analytical predictions.
37: \end{abstract}
38:
39: \pacs{45.70.Qj, 82.40.Ck, 02.30.Jr, 02.60.Lj}
40:
41: \maketitle
42: Localized patterns such as spots belong to the class of dissipative
43: structures found far from equilibrium \cite{DStruc}. In recent years,
44: considerable progress has been made in the understanding of these
45: systems.
46: The question of stability of such patterns is central and the source of
47: instabilities must be carefully scrutinized. In particular, the
48: occurrence of instability can lead to the deformation of spots followed
49: by spot multiplication (also called self-replication) or fingering. This
50: intriguing phenomenon has been the subject of research since the
51: pioneering work of Pearson \cite{pearson}. Shortly after, thanks to the
52: development of open spatial chemical reactors, self-replication was
53: observed in various experiments such as ferrocyanide-iodate-sulphite
54: reaction \cite{lmps}, the Belousov-Zhabotinsky reaction \cite{mpm, kve,
55: kve2},
56: and chloride dioxide-malonic-acid reaction \cite{DBDK}. By now this
57: phenomenon is believed to be universal \cite{mo, hayase1}. It is not
58: restricted to chemical reactions and occurs in many
59: systems in biology \cite{meron}, material science \cite{RW, ns} and
60: nonlinear optics \cite{laser}.
61:
62: Analytically, spot replication is relatively well understood in
63: one dimensional setting. Nishiura and Ueyama \cite{N5} proposed that
64: self-replication of spikes occurs when the spike solution dissapears due
65: to the presence of a fold point. Similar explanation has been reported
66: for the box-like patterns \cite{OK}. In two dimensions,
67: a mechanism for spot instability has been
68: proposed in \cite{mo} for general reaction-diffusion systems; related
69: analysis was performed earlier in the framework of a piecewise-linear
70: approximation in \cite{omk}, \cite{km} and more recently in the context
71: of diblock copolymer systems \cite{RW}, \cite{ns}.
72: See also \cite{Goldstein}, \cite{Mmotion} and \cite{gcom} and for a related
73: approach from the point of view of interface motion.
74:
75: In this letter we perform an analytical and numerical investigation of
76: the two-dimensional localized spots that were recently reported for the
77: Belousov-Zhabotinski (BZ) reaction in water-in-oil microemulsion
78: \cite{kve}, \cite{kve2}.
79: In the classical BZ reaction, spiral waves are observed \cite{KT} but no
80: localized spot solutions are possible. Indeed, localized structures
81: develop only when the ratio of diffusion coefficients is sufficiently
82: large, which occurs in the microemulsion system but not in the
83: classical BZ reaction.
84:
85: We consider the
86: water-in-oil microemulsion model of the BZ reaction as described in
87: \cite{kve} and \cite{kve2}:
88: \begin{subequations}
89: \label{bz}
90: \begin{align}
91: &\begin{aligned}
92: \varepsilon_{0}v_{t} =
93: \varepsilon_{0}D_{v}\Delta v+\left[ f_{0}
94: z+i_{0}\left( 1-mz\right) \right] \frac{v-q_{0}}{v+q_{0}}\\
95: +\left[
96: \frac{1-mz}{1-mz+\varepsilon_{1}}\right] v-v^{2}
97: \end{aligned}\\
98: &z_{t} =D_{z}\Delta z-z+v\left[ \frac{1-mz}{1-mz+\varepsilon_{1}}\right]
99: \end{align}
100: \end{subequations}
101: where $v, z$ are dimensionless concentrations of activator HBrO$_2$ and
102: oxidized catalyst $[Ru(bpy)_3]^{3+}$ respectively; $D_v$ and $D_z$ are
103: dimensionless diffusion coefficients of activator and catalyst; $f, \eps$
104: and $q$ are parameters of the standard Keener-Tyson model \cite{KT}; $i_0$
105: represents the photoinduced production of inhibitor, and $m$ represents
106: the strength of oxidized state of the catalyst with $0 <mz < 1$.
107: This reaction was shown experimentally and
108: numerically to admit localized spot patterns that persist for long time
109: \cite{kve}, \cite{kve2}.
110:
111: We rescale the variables as
112: $
113: z=1/m-m^{-3/2}w\varepsilon_{1},~~v=m^{-1/2}\hat{v},~~ t=\varepsilon
114: _{0}m^{1/2}\hat{t}.
115: $
116: In the new variables, after dropping the hats, we obtain
117: \begin{subequations}
118: \begin{align}
119: v_{t} =\varepsilon^{2}\Delta v+f(v,w);\ \ \tau w_{t}=D\Delta
120: w+g(v,w) \label{vw}
121: \end{align}
122: where
123: \begin{align}
124: f(v,z) &=-\left[ f_{0}+f_{1}w\right] \frac{v-q}{v+q}+\left[
125: \frac{w}{1+\alpha w}\right] v-v^{2};\\
126: g(v,w)&=1-\left[ \frac
127: {w}{1+\alpha w}\right] v
128: \end{align}
129: \end{subequations}
130: and with the nondimensional constants given by
131: \begin{align}
132: \begin{aligned}
133: &\alpha =m^{-1/2},~~ f_{1}=\varepsilon_{1}m^{1/2}\left( i_{0}%
134: -\frac{f_{0}}{m}\right),~~
135: q =q_{0}m^{1/2},\\
136: &\varepsilon^2=\varepsilon_{0}D_{v}m^{1/2}
137: ,~~ D=D_{z}\varepsilon_{1}m^{-1/2},~~ \tau=\frac{1}{m}\frac{\varepsilon_{1}
138: }{\varepsilon_{0}}.
139: \end{aligned}
140: \end{align}
141: In the limit $m\rightarrow\infty,\ f_{1}\rightarrow0$ and $\tau\rightarrow
142: 0\ \ $we obtain the reduced system,%
143: \begin{equation}
144: \left\{
145: \begin{aligned}
146: v_{t}&=\varepsilon^{2}\Delta v-f_{0}\frac{v-q}{v+q}+wv-v^{2}\\
147: 0&=D\Delta w+1-vw.
148: \end{aligned}
149: \right. \label{simple}%
150: \end{equation}
151: In particular, parameter values used in Fig.~14 of \cite{kve2} are
152: $f_0=2.18, i_0=0,
153: m=10, \eps_1 = 0.01, \eps_0=0.1$ and $D_z/D_v=100$ which gives $\alpha=0.3,
154: f_1 = -0.007, \tau=0.01$,
155: so that the simplification (\ref{simple}) is appropriate.
156:
157:
158: { \begin{figure}[b]
159: { \begin{minipage}[t]{0.15\textwidth}
160: \begin{center}{\small
161: \setlength{\unitlength}{1\textwidth} \begin{picture}(1,0.42)(0,0)
162: \put(-0.2,0.0){\includegraphics[width=1.2\textwidth,
163: height=1.1\textwidth]{steady3d.eps}}
164: \end{picture}
165: }(a)\end{center}
166: \end{minipage}
167: \begin{minipage}[t]{0.30\textwidth}
168: \begin{center}{
169: \setlength{\unitlength}{1\textwidth} \begin{picture}(1.0,0.42)(0,0)
170: \put(-0.05,0.03){\includegraphics[height=0.42\textwidth,
171: ]{ss1.eps}}
172: \put(+0.21,0.28){\includegraphics[height=0.12\textwidth,
173: ]{ss1b.eps}}
174: \put(0.5,0.03){\includegraphics[height=0.42\textwidth,
175: ]{ss2.eps}}
176: \end{picture}
177: }
178: (b) \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ (c)
179: \end{center}
180: \end{minipage}
181: }\caption{(a) Radially symmetric, stable spot solution of the reduced BZ
182: system (\ref{simple})
183: in the BZ system on a unit disk.
184: (b) Radial profile of $v$
185: (solid curve) and its asymptotic approximation (dashed
186: curves) given by (\ref{ss}). Insert: a blowup showing the profile of the
187: interface.
188: (c) Radial profile of $w$ (solid curve) and its two-term
189: asympotic approximations
190: (dashed curves) given by (\ref{w}).
191: Parameter values are $D = 20, f_0 = 2.0, \eps=0.02, q=0.005.$
192: }
193: \label{fig:steady}
194: \end{figure}}
195:
196: Experimental and numerical evidence in \cite{kve} and \cite{kve2}
197: suggest that (\ref{bz}) admits localized spot soltuions,
198: such as shown in Fig.~\ref{fig:steady}.
199: Such spots occur in the regime where $\eps \ll 1.$
200: Our goal is to describe analytically the radius and profile of such a spot
201: and then study its
202: stability using singular perturbation techniques
203: similar to those described in \cite{mo}.
204: As we will demonstrate, the instability thresholds appear in
205: the regime where $D \gg 1$. For sufficiently large values of $D$ a
206: spot pattern is stable. However as $D$ is decreased, instabilities of the
207: form $\exp(\lambda t) \cos(m\theta) \phi(r)$ may develop, where $\theta$
208: and $r$ are the angular and radial coordinates, respectively.
209: We provide the analytic description of the profile of the spot
210: (\ref{ss}, \ref{w}) and the dispersion relation (\ref{main-result})
211: between $m$ and $\lambda$. This relation leads directly to the estimate
212: for the instability threshold.
213:
214: We begin by constructing a stationary (time-independent)\ spot-type solution
215: on a two dimensional unit disk $\left\{ x:\left\vert x\right\vert <1\right\}
216: $. We assume that $D\gg1$. Then to leading order, $w\sim w_{0}$
217: is a constant to be determined, and the solution for $v$ consists of an
218: interface located at $\left\vert x\right\vert \sim l$ which connects two
219: nearly spatially homogeneous layers. To find the profile of such an interface
220: and its location $l$, let us rescale near $l$ as $v\left( x\right) =V\left(
221: y\right) $ with $y=\left( r-l\right) /\varepsilon$ and $r=\left\vert
222: x\right\vert .$ For the steady state we then obtain, to leading order,
223: $V^{\prime\prime}\left( y\right) +f\left( V;w_{0}\right) =0.$ The
224: interface solution corresponds to a heteroclinic orbit of this ODE. The
225: existence of such an orbit is only possible whenever the equations
226: \begin{equation}
227: f\left( V_{-};z_{0}\right) =0=f\left( V_{+};w_{0}\right) ,\ \ \ \ \int
228: _{V_{-}}^{V_{+}}f\left( s;w_{0}\right) ds=0\label{Vpm}%
229: \end{equation}
230: are simultaneously satisfied for some values $V_{-}$ $<V_{+}.$ Since $q\ll1$,
231: we have $V_{-}\sim 0$ and (\ref{Vpm}) can be written
232: as
233: \begin{equation}
234: f_{0}+f_{1}w_{0}\sim\frac{3}{16}\left[ \frac{w_{0}}{1+\alpha w_{0}}\right]
235: ^{2};\ \ \ V_{+}\sim\frac{3}{4}\left[ \frac{w_{0}}{1+\alpha w_{0}}\right]
236: .\label{Vpw0}%
237: \end{equation}
238: Next, we ignore the $O(q)$ terms and integrate $V''(y)+f(V,w_0)=0$
239: to obtain
240: \begin{equation}
241: \begin{aligned}
242: v&\sim\left\{
243: \begin{aligned}
244: &V_{+}\tanh^{2}\left( \sqrt{\frac{V_{+}}{6}}\left( \dfrac{r-l}{\varepsilon
245: }\right) \right) ,~~& r<l\\
246: &0,& r>0
247: \end{aligned}
248: \right.
249: \end{aligned}
250: \label{ss}%
251: \end{equation}
252: with $V_{+}$ and $w_{0}$ given by (\ref{Vpw0}) and where $r=\left\vert
253: x\right\vert $. This formula describes the profile of the interface in
254: the limit $q \to 0$. Its
255: thickness is of $O(\eps V_+^{-1/2}) .$ To determine its location
256: $l$, we integrate the second equation in (\ref{vw}). Zero-flux conditions then
257: yield $\int g=0$ so that
258: \begin{equation}
259: g\left( V_{+},w_{0}\right) l^{2}+\left( 1-l^{2}\right) g\left(
260: 0,w_{0}\right) \sim0\label{gio}%
261: \end{equation}
262: In the limit of the reduced system (\ref{simple}),\ we obtain%
263: \begin{align}
264: \label{ss2}
265: l\sim\frac{1}{2\sqrt{f_{0}}},\ \ w_{0}\sim\frac{4}{\sqrt{3}}\sqrt{f_{0}%
266: },\ \ \ V_{+}\sim\sqrt{3f_{0}}.
267: \end{align}
268: To determine the correction to $w,$ we write $w=w_{0}+D^{-1}%
269: w_{1}+O\left( D^{-2}\right) $. to obtain $\Delta w_{1}+g\left( w_{0}%
270: ,v_{0}\right) =0.$ Imposing continuity at the interface $r=l$, the
271: solvability condition $w_{1}(l)=0$, and using (\ref{gio})\ and $g\left(
272: 0,w_{0}\right) =1,$ we then obtain
273: \begin{align}
274: w\sim w_{0}+\frac{1}{4D}\left\{
275: \begin{aligned}
276: &-\dfrac{(1-l^2)(l^2-r^2)}{l^2}, &r<l\\
277: &
278: 2\ln(\dfrac{r}{l})+l^2-r^2, &r>l
279: \end{aligned}
280: \right. .
281: \label{w}
282: \end{align}
283: An example of a localized spot and its radial profile are shown in Fig
284: \ref{fig:steady}.
285:
286: When decreasing the diffusion coefficient $D$ of the recovery variable,
287: numerical simulations show that the spot becomes unstable. To compute the
288: threshold associated with this instability, we linearlize around the
289: localized spot solution (\ref{ss}) as
290: \begin{align*}
291: v(x,t) & =v(x)+\exp\left( \lambda t\right) \cos\left( m\theta\right)
292: \phi\left( r\right) \\
293: w(x,t) & =w(x)+\exp\left( \lambda t\right) \cos\left( m\theta\right)
294: \psi\left( r\right)
295: \end{align*}
296: where $v$ and $w$ are given to leading order in (\ref{ss} and \ref{ss2}) spot-type solution, $m$ is an integer, $\phi,\psi\ll1$
297: and $(r,\theta)$ are the polar coordinates of $x.$ Substituting into
298: (\ref{bz}) we then obtain
299: {
300: \begin{subequations}
301: \begin{align}\label{phi}
302: \lambda\phi & =\varepsilon^{2}\left( \phi_{rr}+\frac{1}{r}\phi_{r}%
303: -\frac{m^{2}}{r^{2}}\phi\right) +f_{v} \phi+f_{w} \psi\\
304: \label{psi}
305: \tau\lambda\psi & =D\left( \psi_{rr}+\frac{1}{r}\psi_{r}-\frac{m^{2}}{r^{2}%
306: }\psi\right) +g_{v} \phi+g_{w} \psi.
307: \end{align}
308: \label{phipsi}
309: \end{subequations} }
310: Note that $v_r$ satsifies
311: that $\varepsilon^{2}(v_{rrr}+1/rv_{rr}-1/r^2v_r)+f_{v}v_{r}+w_{r}f_{w}=0.$
312: Therefore multiplying (\ref{psi}) by $v_r$ and integrating by parts, we obtain
313: \begin{align}\label{2:56}
314: \left( \lambda+\frac{(m^{2}-1)\varepsilon^{2}}{l^{2}}\right) \int v_{r}^{2}%
315: \sim\int v_{r}f_{w}\left( \psi-w_{r}\right) .
316: \end{align}
317: where we have assumed that to leading order,
318: $\phi\sim v_{r}, \lambda \ll 1, \psi \ll \phi$.
319: Since $v_{r}$ is exponentially small outside the interface, we simplify $\int
320: v_{r}f_{w}\left( \psi-w_{r}\right) \sim-\left( \psi\left( l\right)
321: -w_{r}\left( l\right) \right) \int_{0}^{V_{+\ }}f_{w}\left( v,w_{0}%
322: \right) dv$. We estimate
323: $w_r(l) \sim -1/(2D)g\left( V_{+},w_{0}\right)$ and
324: to determine $\psi\left( l\right),$ we integrate (\ref{psi})
325: over the interface $l$ to obtain
326: $
327: D\psi_{r}|_{l^{-}}^{l^{+}}
328: \sim-g\left( 0,w_{0}\right) /l^{2}
329: $
330: where we have used $\phi \sim v_r$ and (\ref{gio}). Keeping only
331: leading order terms in $D$ we then obtain the following problem for
332: $\psi$:
333: \begin{subequations}
334: \label{reducedpsi}
335: \begin{align}\label{djump}
336: &0=
337: \psi_{rr}+\frac{1}{r}\psi_{r}-\frac{m^{2}}{r^{2}}\psi=0,
338: ~~~~ r\ne l
339: \\
340: & \psi^{\prime}( 0)=0=\psi^{\prime}( 1)
341: ,~~
342: \psi^{\prime}\left( l^{+}\right) -\psi^{\prime}\left( l^{-}\right)
343: =-\frac{1}{Dl^{2}}g\left( 0,w_{0}\right) .
344: \end{align}
345: \end{subequations}
346: Using the continuity of $\psi$ at $r=l$ we get
347: \[
348: \psi\left( l\right) \sim\frac{1}{2Dml}g\left( 0,w_{0}\right) \left( 1+l^{2m}\right)
349: \]
350: and substituting into (\ref{2:56}) we obtain
351:
352: \begin{align}
353: \lambda\frac{l^{2}}{\varepsilon^{2}}\sim 1-m^{2}+A\left[ 1-l^{2}-\frac{1}%
354: {m}\left( 1+l^{2m}\right) \right]
355: \label{main-result}
356: \end{align}
357: where
358: \begin{align}
359: A=\frac
360: {l}{2\varepsilon D}\frac{g\left( 0,w_{0}\right) \left( \int_{0}^{V_{+\ }%
361: }f_{w}\left( v,w_{0}\right) dv\right) }{\varepsilon\int v_{r}^{2}}.
362: \end{align}
363: For the reduced model (\ref{simple}) we obtain an explicit result
364: \begin{align}
365: \label{simple-A}
366: A \sim \frac{3^{5/4}2^{1/2}5}{64}\frac{1}{f_0^{3/4}\eps D}
367: =
368: 0.43622\frac{1}{f_{0}^{3/4}\varepsilon D}.
369: \end{align}
370: In Fig.~\ref{fig:evals}, the analytical prediction given by
371: (\ref{main-result}) is found to be in good agreement with the numerical
372: computations of problem (\ref{phipsi}). Full numerical simulations of
373: (\ref{simple}) also agree with this prediction. For example, when taking
374: $\eps=0.03, f_0=1.3$ and $q=0.01$, slight spot deformation corresponding
375: to mode $m=2$ is observed when $D = 1.0$
376: but not when $D=1.3.$ This agrees well with $D\sim1.1$,
377: the threshold predicted by (\ref{main-result}).
378: \begin{figure}[tb]
379: \setlength{\unitlength}{0.5\textwidth} \begin{picture}(1,0.30)(0,0)
380: \put(0.0,0.17){\footnotesize $\dfrac{\lambda}{\eps^2}$}
381: \put(0.04,0.09){\includegraphics[width=0.12\textwidth,
382: height=0.08\textheight]{eval1.eps}}
383: \put(0.18,0.056){\footnotesize $m$}
384: \put(0.18,0.00){(a)}
385: \put(0.38,0.09){\includegraphics[width=0.12\textwidth,
386: height=0.08\textheight]{eval2.eps}}
387: \put(0.33,0.17){\footnotesize $\dfrac{\lambda}{\eps^2}$}
388: \put(0.51,0.056){\footnotesize $D$}
389: \put(0.51,0.00){(b)}
390: \put(0.71,0.09){\includegraphics[width=0.12\textwidth,
391: height=0.08\textheight]{Dl.eps}}
392: \put(0.66,0.17){\footnotesize $D$}
393: \put(0.79,0.063){\footnotesize $l=\frac{1}{4f_0^2}$}
394: \put(0.84,0.00){(c)}
395: \end{picture}
396: \caption{(a), (b) Comparison of numerical computations of $\lambda$ given by
397: (\ref{phipsi}) (diamonds) with
398: the analytical result (\ref{main-result}, \ref{simple-A}) (dashed line)
399: for the reduced model
400: (\ref{simple}) on a unit disk. Parameter values are $q=0.01$, $\eps=0.01$ and
401: (a) $D=3, f_0=1.3$; (b) $m=2, f_0=1.3$.
402: (c) Simultaneous solution of $\lambda=0=d\lambda/dm$, showing the first
403: value of $D$ for which instability occurs. The system is stable above the
404: curve and unstable below it.
405: To compute $\lambda$ numerically,
406: (\ref{phipsi}) was reformulated as a boundary value problem by adjoining
407: the equation $d\lambda/dr=0$ along with fixing $\psi(1).$
408: Maple's numerical
409: boundary value problem solver was then used with initial guesses
410: $\phi=v_r, \lambda=0$ and $\psi =$ the solution of
411: (\ref{reducedpsi}).
412: All computations are
413: correct to four significant digits. }
414: \label{fig:evals}
415: \end{figure}
416:
417: From (\ref{main-result}) and (\ref{simple-A}), it is clear that
418: the mode $m=1$ is always stable and that
419: for large enough $D$, all modes $m\geq1$ are stable. As $D$ is decreased,
420: instability sets in when $D=O(\eps^{-1})$.
421: The threshold
422: value is found by simultaneously solving $\lambda=d\lambda/dm=0$.
423: The resulting graph is shown on Fig.~\ref{fig:evals}c.
424: In particular, note that the system is stable if $D\eps > 0.038$, independent
425: of the value of $f_0.$
426: In the limit
427: of small radius $l\rightarrow0$, the first unstable integer mode is $m=2$
428: corresponding to $A=6,$ so that
429: the spot of small radius becomes unstable whenever
430: $\eps D f^{3/4} \le 0.0727.$
431: More generally, by eliminating $A$
432: we find
433: that there exist constants $l_{1}<l_{2}<\ldots<1$ such the first unstable mode
434: is $m$ provided that $l_{m-1}<l<l_{m},$ where $l_1=0$, $l_2=0.491$, $l_3=0.667$,
435: $l_4=0.753$ and
436: $l_{m}\sim 1-0.937/m$ as $m\rightarrow\infty;$ where 0.937 is the root of
437: $e^{-2z}\left( 3+2z\right) +3-4z=0.$
438:
439:
440: \begin{figure*}
441: \begin{center}
442: \includegraphics[width=0.45\textwidth]{bif.eps}
443: \includegraphics[width=0.45\textwidth]{bif-ex.eps}\\
444: \hfill (a)\hfill$~$\hfill (b) \hfill$~$\\
445: \end{center}
446: \caption{ (a) Bifurcation diagram for (\ref{simple}) in $D$ and $f_0$ with
447: $\eps=0.05, q=0.01.$ Solid dots represent deformations of a spot without
448: topological change. Points marked by "2" represent spot-replication into
449: two spots. An empty circle represents spot-to-ring bifurcation
450: and empty circle
451: with a number inside represents spot-to-ring-to-spots bifurcation.
452: A solid line represents the boundary of spot-to-ring replication which
453: occurs at the fold point of the radially symmetric steady state.
454: The bifurcation diagram was obtained by solving the full two-dimensional
455: system (\ref{simple})
456: using the finite element package {\tt FlexPDE} \cite{FlexPDE} with zero-flux
457: boundary conditions and 800 elements on a quarter-disk. For initial
458: conditions, (\ref{ss}, \ref{ss2}, \ref{w}) was used, but with $r$
459: replaced by $r(1+0.05 \cos2\theta)$.
460: The solid
461: line was obtained by solving for the fold point of the radially symmetric
462: steady state using Maple's boundary value problem solver. (b) Full
463: numerical simulations starting from a spot-state (\ref{ss}, \ref{w}) as
464: initial conditions.
465: First row: spot deformation, $f_0=1.3 , D=0.35$.
466: Note that the final
467: stedy state of the system is a deformed blob shown here at $t=4000$.
468: Second row: self-replication, $f_0=2.2, D=0.24$.
469: Third row: spot-to-ring
470: bifurcation, $f_0=1.3, D=0.1$. Fourth row: spot-to-ring-to-spots
471: bifurcation, $f_0=2.6, D=0.1$.}
472: \label{fig:bif}
473: \end{figure*}
474:
475:
476: Numerical computations indicate that self-replication is more prevalent
477: for spots of small radius (see Fig.~\ref{fig:bif}).
478: For larger spots, a deformation usually leads to
479: to the so-called "finger growth'' and space filling curves.
480: Others studies have shown the occurrence of
481: fingering instabilities leading to labyrinthine patterns \cite{Fingers},
482: \cite{Goldstein}, \cite{mo}. The
483: question of whether the self-replication or fingering instability
484: occurs first is still open.
485:
486:
487: For smaller, $O(1)$ values of $D$, there is also a different instability
488: mechanism that can lead to splitting of a spot into a ring as illustrated
489: in Fig.~\ref{fig:bif}. Unlike spot multiplication, this instability
490: is radially symmetric, and is caused by the {\em dissapearence} of the
491: steady state solution -- whereby the steady state solution ceases to
492: exist due to the presence of a saddle-node bifurcation --
493: rather than by its lateral instability. Numerical simulations suggest
494: that that spot replication occurs only for spots of smaller radius,
495: whereas spot-to-ring instability is dominant for larger spots.
496:
497: To conclude,
498: we have estimated analytically for which
499: value of diffusion $D$ spot deformation first occurs.
500: As is $D$ is decreased further, self-replication and/or spot-to-ring
501: instability is observed. As evidenced by numerical simulations, we
502: conjecture that spot deformation is the precursor to this phenomenon.
503:
504:
505: \begin{thebibliography}{}
506: The authors are grateful to Vladimir Vanag and Irving Epstein
507: for introducing us to this subject and for illuminating discussions.
508: We are also grateful to the anonymous referees whose
509: suggestions have improved the paper significantly.
510: T.K.~is supported by NSERC Canada, and the
511: Dalhousie University startup grant. M.T.~is supported by the
512: FNRS Belgium.
513:
514:
515:
516: \bibitem{DStruc}P.~Glansdorff and I.~Prigogine, {\it {Thermodynamic Theory of Structures, Stability and Fluctuations}} (Wiley, New York, 1971); E. Meron, Phys. Rep. {\bf{218}}, 1 (1992);
517: J.~Chanu and R.~Lefever, {\it Inhomogeneous phases and pattern
518: formation}, Physica A {\bf 213}, 1-276 (1995);
519: M.C.~Cross and P.C.~Hohenberg, Mod. Phys. Rep. {\bf{65}}, 851 (1993);
520: K.~Kapral and K.~Showalter, {\it{Chemical Waves and Patterns}}
521: (Kluwer Academic press, Dordercht, 1995).
522:
523: \bibitem{pearson}J.E.~Pearson,
524: Science, {\bf {261}}, 189 (1993).
525:
526: \bibitem{lmps}K.~Lee, W.D.~McCormick, J.E.~Pearson, and H.L.~Swinney,
527: Nature, {\bf{396}}, 215 (1994).
528:
529: \bibitem {mpm}A.P. Mu\~nuzuri, V. P\'erez-Villar, and M.~Markus,
530: Phys. Rev. Lett. {\bf{79}}, 1940 (1997).
531:
532: \bibitem{kve}A.~Kaminaga, V.K.~Vanag, and I.R.~Epstein,
533: Angewandte Chemie, {\bf{45}}, 3087 (2006).
534:
535: \bibitem{kve2}A.~Kaminaga, V.K.~Vanag, and I.R.~Epstein,
536: J. Chem. Phys. {\bf 122}, 174706 (2005).
537:
538: \bibitem{DBDK}P.W.~Davis, P.~Blanchedeau, E.~Dulos, and P.~De~Kepper,
539: J.~Phys.~Chem.~A {\bf{102}} 8236 (1998).
540:
541: \bibitem{mo} C.~Muratov and V.V.~Osipov, Phys. Rev. E {\bf 53}, 3101
542: (1996); Phys. Rev. E {\bf 54}, 4860 (1996);
543: C.~Muratov, Phys. Rev. E {\bf 66}, 066108 (2002).
544:
545:
546: \bibitem{hayase1}Y. Hayase and T. Ohta, Phys. Rev. Lett. {\bf {81}}, 1726 (1998);
547: Y. Hayase, Phys. Rev. E {\bf {62}}, 5998 (2000).
548:
549: \bibitem{meron}
550: E.~Meron, E.~Gilad J.~von~Hardenberg, and M.~Shachak,
551: Chaos, Solitons and Fractals, {\bf{19}}, 367 (2004).
552:
553:
554: \bibitem{RW}X.\ Ren and J.\ Wei, SIAM J. Math. Anal. {\bf {35}}, 1 (2003).
555:
556: \bibitem{ns} Y.Nishiura and H.Suzuki, SIAM Math.Anal. {\bf 36}(3), 916
557: (2004).
558:
559: \bibitem{laser} M. Tlidi A.G. Vladimirov, and P. Mandel, Phys. Rev.
560: Lett.
561: {\bf {89}}, 233901 (2002).
562:
563:
564: \bibitem{N5}Y. Nishiura and D. Ueyama, Physica D {\bf {130}}, 73 (1999).
565:
566: \bibitem{OK}B.S. Kerner and V.V. Osipov, {\it{Autosolitons: A new Approach to Problems of Self-Organization and Turbulence}} (Kluwer, Dordrecht, 1994).
567:
568:
569: \bibitem{omk}T.~Ohta, M.~Mimura, and R.~Kobayashi, Physica D {\bf {34}}, 115 (1989).
570:
571: \bibitem{km} S.Kawaguchi and M.Mimura, SIAM J.Appl,Math {\bf 59}, 920(1999)
572:
573: \bibitem{KT}J.J.~Tyson and P. Fife, J. Chem. Phys. {\bf {73}}, 2224 (1980);
574: J.P.~Keener and J.J.~Tyson, Physica D {\bf {21}}, 307 (1986).
575:
576: \bibitem{Goldstein}
577: R.E.~Golstein, D.J. Muraki, and D.M. Petrich, Phys. Rev. E {\bf {53}}, 3939 (1996).
578:
579: \bibitem{Mmotion}
580: C.B.~Muratov, Phys. Rev. E {\bf 54}, 3369 (1996)
581:
582:
583: \bibitem{gcom}
584: D.~Gomila, P.~Colet, G.L.~Oppo and M.~San Miguel,
585: Phys.~Rev.~Lett.~{\bf 87}, 194101 (2001).
586:
587:
588: \bibitem{Fingers}
589: K. Lee and H. Swinney, Phys. Rev. E {\bf {51}}, 1899 (1995).
590:
591: \bibitem{FlexPDE} See FlexPDE website, {\tt www.pdesolutions.com}.
592:
593: \end{thebibliography}
594: \end{document}
595: