1: \documentclass{ejm}%
2: \usepackage{amsmath}
3: \usepackage{amsfonts}
4: \usepackage{amssymb}
5: \usepackage{graphicx}%
6: \setcounter{MaxMatrixCols}{30}
7: %TCIDATA{OutputFilter=latex2.dll}
8: %TCIDATA{Version=5.00.0.2606}
9: %TCIDATA{CSTFile=40 LaTeX article.cst}
10: %TCIDATA{Created=Monday, December 06, 2004 15:59:33}
11: %TCIDATA{LastRevised=Thursday, March 17, 2005 15:31:33}
12: %TCIDATA{<META NAME="GraphicsSave" CONTENT="32">}
13: %TCIDATA{<META NAME="SaveForMode" CONTENT="1">}
14: %TCIDATA{BibliographyScheme=Manual}
15: %TCIDATA{<META NAME="DocumentShell" CONTENT="Standard LaTeX\Blank - Standard LaTeX Article">}
16: \newtheorem{theorem}{Theorem}
17: \newtheorem{axiom}[theorem]{Axiom}
18: \newtheorem{case}[theorem]{Case}
19: \newtheorem{claim}[theorem]{Claim}
20: \newtheorem{conclusion}[theorem]{Conclusion}
21: \newtheorem{condition}[theorem]{Condition}
22: \newtheorem{conjecture}[theorem]{Conjecture}
23: \newtheorem{corollary}[theorem]{Corollary}
24: \newtheorem{criterion}[theorem]{Criterion}
25: \newtheorem{definition}[theorem]{Definition}
26: \newtheorem{example}[theorem]{Example}
27: \newtheorem{exercise}[theorem]{Exercise}
28: \newtheorem{lemma}[theorem]{Lemma}
29: \newtheorem{notation}[theorem]{Notation}
30: \newtheorem{problem}[theorem]{Problem}
31: \newtheorem{oproblem}[theorem]{Open problem}
32: \newtheorem{proposition}[theorem]{Proposition}
33: \newtheorem{remark}[theorem]{Remark}
34: \newtheorem{solution}[theorem]{Solution}
35: \newtheorem{summary}[theorem]{Summary}
36: \def\eps{\varepsilon}
37: \def\goesto{\rightarrow}
38: \def\dirac{{\mbox{\boldmath$\delta$}}}
39:
40: %%%%%%%%%%%%%%%%%%%%%%
41:
42: \title[Localized patterns in the Brusselator]
43: {Mesa-type patterns in the one-dimensional Brusselator and their stability}
44:
45: \author[Theodore Kolokolnikov, Thomas Erneux and Juncheng Wei]{%
46: T.\ns K\ls O\ls L\ls O\ls K\ls O\ls L\ls N\ls I\ls K\ls O\ls V$~^\dag$, \ns
47: T.\ns E\ls R\ls N\ls E\ls U\ls X$~^\dag$ \ns
48: \and J.\ns W\ls E\ls I$~^\ddag$
49: }
50:
51: \affiliation{$~^\dag$\ Universit\'e Libre de Bruxelles, Optique Nonlin\'eaire
52: Th\'eorique, Campus Plaine, C.P. 231, 1050 Brussels, Belgium\\
53: tkolokol\symbol{64}gmail.com, terneux\symbol{64}ulb.ac.be
54:
55: $~^\ddag$\ Department of Mathematics,
56: The Chinese University of Hong Kong,
57: Shatin, Hong Kong \\
58: wei\symbol{64}math.cuhk.edu.hk}
59: \pubyear{2005}
60: \volume{000}
61: \pagerange{\pageref{firstpage}--\pageref{lastpage}}
62:
63:
64:
65: \begin{document}
66:
67: \maketitle
68:
69: \begin{abstract}
70: The Brusselator is a generic reaction-diffusion model for a tri-molecular
71: chemical reaction. We consider the case when the input and output reactions
72: are slow. In this limit, we show the existence of $K$-periodic, spatially
73: bi-stable structures, \emph{mesas}, and study their stability. Using singular
74: perturbation techniques, we find a threshold for the stability of $K$ mesas.
75: This threshold occurs in the regime where the exponentially small tails of the
76: localized structures start to interact. By comparing our results with Turing
77: analysis, we show that in the generic case, a Turing instability is followed
78: by a slow coarsening process whereby logarithmically many mesas are
79: annihilated before the system reaches a steady equilibrium state. We also
80: study a ``breather''-type instability of a mesa, which occurs due to a Hopf
81: bifurcation. Full numerical simulations are shown to confirm the analytical results.
82:
83: \end{abstract}
84:
85: \label{firstpage}
86:
87: \section{Introduction}
88:
89: In 1952, Turing proposed that the formation of spatial patterns during
90: morphogenesis could be explained in terms of the instability of a homogeneous
91: steady-state solution of a reaction-diffusion network describing the evolution
92: of a set of morphogens \cite{Turing}. Turing himself illustrated his ideas on
93: two chemical models. Turing's original work is primarily concerned with the
94: stability analysis of the homogeneous steady-state solution of the rate
95: equations for the interacting morphogens \cite{NP}. The main point of
96: biological interest, however, is whether stable spatial structures may be
97: generated beyond the instability, i.e., whether the rate equations admit
98: stable (and positive) inhomogeneous solutions exhibiting the most
99: characteristic features of morphogenetic patterns.
100:
101: This point has been taken up seriously in the early seventies. More systematic
102: numerical studies of Turing's model have been performed showing irregular
103: spatial structures \cite{bard}. This led to serious reservations about the
104: relevance of Turing's theory in developmental biology, particularly its
105: ability to generate regular patterns. But these criticisms originate from a
106: somewhat unfortunate choice of Turing's example and they do not touch the
107: essential points of Turing's theory. More specifically, the obvious
108: requirement that the rate equations must admit positive and bounded solutions
109: is not satisfied in Turing's example \cite{erneux2}.\
110:
111: Because of this misunderstanding on Turing's fundamental contribution, other
112: two-variable models satisfying the law of mass action have been explored. In
113: 1968, Prigogine and Lefever \cite{PL} introduced a two variable system that
114: exhibits an autocatalytic reaction (called the ``Brusselator''). The
115: simplicity of the rate equations motivated analytical and numerical studies
116: which showed the existence of stable structures \cite{auchmuty}. The
117: Brusselator is based on the following intermediate reactions for the two
118: chemical intermediates $X$ and $Y$:
119: \begin{equation}
120: A \rightarrow X, ~~~ C + X \rightarrow Y + F, ~~~ 2X + Y \rightarrow3 X, ~~~ X
121: \rightarrow E.
122: \end{equation}
123:
124:
125: \begin{figure}[tb]
126: \begin{center}
127: \setlength{\unitlength}{1in} \begin{picture}(6.5,3.0)(0.8,0)
128: %\put(0.1,0.0){\includegraphics[width=5in, height=3in]{turing.eps}}
129: \put(0.16,0.25){\includegraphics[width=3.5in]{turing.eps}}
130: \put(3.333,0.25){\includegraphics[width=3.5in]{fig4.eps}}
131: \put(1.7,0.1){(a)}
132: \put(5.0,0.1){(b)}
133: \end{picture}
134: \end{center}
135: \caption{(a) Turing instability and mesa-type localized structures. The
136: parameters are $A=1,B=8,~\varepsilon=10^{-4},~D=10,~\tau=10.$ Initial
137: condition was set to $u=A, v=A/B$, perturbed by a very small random noise.
138: Note the logarithmic scale for time. Initial Turing instability triggers a
139: $k=7$ mode at time $t \sim50$. Thereafter a coarsening process takes place
140: until there are only two mesas left. (b) Slow-time oscillatory instability of
141: a single spike solution to (\ref{6dec4:04}). The parameter values are
142: $A=1,~B=8,~D=10,~\varepsilon=0.00025,~\tau=0.999.$}%
143: \label{fig:turing}%
144: \end{figure}
145:
146: The global reaction is $A+C\rightarrow F+E$ and corresponds to the
147: transformation of inputs products $A$ and $C$ into output products $F$ and
148: $E.$ We assume (without loss of generality) that the rate constants for the
149: first and last step are equal to $r$ whereas the intermediate rate constants
150: are one. The rates equations then become
151: \begin{align}
152: X_{t} & =D_{x}X_{xx}+r A-C X+X^{2}Y-rX\\
153: Y_{t} & =D_{y}Y_{xx}+CX-X^{2}Y.
154: \end{align}
155:
156:
157: In this paper we concentrate on the case where the input and the output reaction
158: steps are slow in comparison to the intermediate steps, so that $r$ is small. In the
159: absence of diffusion terms, this implies that the total concentration $X+Y$
160: is a slowly changing quantity. Linear stability reveals the presence of a periodic
161: orbit which arise through a Hopf bifurcation of the steady state $X=A, Y=C/A$. In
162: the limit $r \rightarrow0$ such a periodic orbit consists of a slow and a fast
163: phase, so that a pulse train can be observed.
164:
165: There is a large body of literature on the space-independent Brusselator
166: ($D_{x}=0=D_{y}$) and its extensions. Under different assumptions on the
167: parameters, the equilibrium state is only marginally stable and an addition of
168: small random perturbations (or small periodic forcing) can lead regular to
169: large-amplitude pulses \cite{OP}, \cite{MVE}. Other authors have looked at
170: coupled brusselator systems which can exhibit chaos and syncrhonization
171: \cite{Tyson}, \cite{YuGumel}, \cite{ChakravartiMarekRay}.
172:
173: Since the discovery of spatial patterns in 1970's, various Turing patterns in
174: the Brusselator were studied both numerically and analytically in one, two and
175: three space dimensions. These include spots, stripes, labyrinths and and
176: hexagonal patterns \cite{Erneux}, \cite{NP}, \cite{deWit}, \cite{Pena},
177: \cite{deWit3D}, oscillatory instabilities and spatio-temporal chaos
178: \cite{deWitChaos}, \cite{YZE}. While Turing analysis and its weakly nonlinear
179: extensions have been successful at detecting and classifying possible pattern
180: types, its range of applicability is limited. Indeed Turing patterns are
181: assumed to be small sinusoidal perturbations of a homogeneous state. In practice
182: however, many patterns are localized and contain sharp transitions such as spikes
183: and kinks. As we will see, this is particularly the case under the assumption of
184: small $r$. In this paper we study \emph{mesa}-type patterns.\footnote{Mesa means
185: table in Spanish; it is also a name given to square boulders which are found in the
186: Colorado desert. The use of this term was suggested by Fife \cite{Fife}.} These are
187: box-like patterns that join two flat regions of space with sharp transition layers,
188: such as shown in Figure \ref{fig:1}. Such patterns are not amenable to Turing
189: analysis since they are far away from the homogeneous state.
190:
191: Previous mathematical efforts on the formation of localized patterns concentrated on
192: two variable reaction-diffusion models where the dynamics is controlled by the
193: interaction between a slow and a fast variable. For the Brusselator without
194: diffusion terms and with $r$ small, the total concentration $X+Y$ plays the role of
195: a slow variable. As a result, localized patterns patterns are expected when
196: diffusion is permitted.
197:
198: For our analysis it will be convenient to use the following scaling,
199: \begin{equation}\label{6dec4:04}
200: \begin{aligned} \tau u_{t} & =D \varepsilon u_{xx}+\varepsilon A-Bu+u^{2}v-\varepsilon u\\ v_{t} & =D \varepsilon v_{xx}+Bu-u^{2}v, \end{aligned}
201: \end{equation}
202: where
203: \begin{equation}
204: \tau=\frac{D_{y}}{D_{x}},~~~\varepsilon=r\frac{D_{y}}{D_{x}},~~~B=C\frac
205: {D_{y}}{D_{x}},~~~D=\frac{D_{x}}{r},
206: \end{equation}
207: \begin{equation}
208: u=X,~~~~v=\tau Y
209: \end{equation}
210: and the spatial domain is $x\in\lbrack0,1]$ with the zero flux boundary
211: conditions
212: \begin{equation}
213: u_{x}=0=v_{x}~~\mbox{at}~x=0~\mbox{and}~1.
214: \end{equation}
215:
216:
217: In this work we assume the following conditions,
218: \begin{align}
219: r \ll D_{x} \le O(D_{y}) \ll1,~~~~ A=O(1), ~~ C = O\left( \frac{D_{x}}{D_{y}%
220: }\right) .
221: \end{align}
222: In terms of our scaling we have
223: \begin{equation}
224: \label{epsD}\varepsilon D\ll1,~~~D\gg1 ; ~~~~~O(A)=1=O(B),
225: \end{equation}
226: \begin{equation}
227: O(\tau) \ge1.
228: \end{equation}
229:
230:
231: As a motivation, let us present two numerical examples. On Figure
232: \ref{fig:turing}.a we show a typical time evolution over a very long time
233: interval. The homogeneous steady state is unstable because of Turing
234: instability but the expected Turing sinusoidal pattern only appears after a
235: delay ($t > 50$). Turing's structure then gradually deteriorates and
236: relatively quickly moves into a new pattern formed by several localized
237: mesa-type structures. They then undergo a coarsening process over a
238: logarthmically long time-scale and, eventually, only two mesas remain.
239: Turing's analysis can be used to predict the first pattern ($t \sim50$) but it
240: cannot anticipate the coarsening process or the number of final mesas. As we
241: shall demonstrate, in the case $B > A^{2}$, the Turing pattern is
242: characterized by a wave number proportional to $k=O\left( \frac{1}%
243: {\sqrt{D\varepsilon}}\right) $ while the number of stable long-time mesas has
244: the order $K=O\left( \frac{1}{\sqrt{D\varepsilon} \ln\frac{1}{\varepsilon}%
245: }\right) $, so that $k$ is logarithmically bigger than $K$. This is the
246: underlying cause of the coarsening process observed in Figure \ref{fig:turing}.a.
247:
248: In the second experiment shown in Figure \ref{fig:turing}.b, a single mesa
249: undergoes a ``breather''-type oscillatory instability which eventually leads
250: to its extinction. Both coarsening and the breather instability occur at a
251: slow timescale. The main goal of this work is to describe these instabilities analytically.
252:
253: Our results are related to the study of bistable systems, see for example
254: \cite{RW}, \cite{GMP}, \cite{N0} \cite{N1}, \cite{N2}, \cite{N3}, \cite{N4},
255: \cite{DI}. Mesa patterns also appear in the FitzHugh-Nagumo model \cite{GMP},
256: certain phase separation models such as Cahn-Hilliard, Allen-Cahn, \cite{ACF},
257: \cite{AFS}, \cite{C} and block-copolymers \cite{RW}. For these systems the
258: resulting spectral problem has small eigenvalues, also called critical
259: spectra, that tend to zero with the thickness of the interfaces. Typically $k$
260: such layers are stable \cite{N3}. However as we show in this paper, if the
261: number of layers is excessively large, instabilities can occur. This happens
262: when the exponentially small interaction outside the interface locations
263: cannot be ignored. Our main new contribution is to study this interaction, and
264: to show that it has a destabilizing effect.
265:
266: \subsection{Summary of main results}
267:
268: We now summarize our main results. We first describe the shape of the
269: equilibrium $K$ mesa solutions. In Section \ref{sec:2} we show the following.
270:
271: \begin{figure}[tb]
272: \begin{center}
273: \setlength{\unitlength}{1in} \begin{picture}(5.0,3.1)(0,0)
274: \put(2.5, 0){$x$}
275: \put(0, 1.6){$v$}
276: \put(0.2,0.2){\includegraphics[width=4.5in]{fig1.eps}}
277: \end{picture}
278: \end{center}
279: \caption{An example of a three-mesa equilibrium solution for $v$. Here, $K=3,~
280: A=2,~B=18,~\varepsilon D=0.02^{2},$ $w_{0}=9$, $l=0.11.$ }%
281: \label{fig:1}%
282: \end{figure}
283:
284: \begin{proposition}
285: \label{prop:1}Consider the equilibrium-state problem,%
286: \begin{gather}
287: 0=\varepsilon Dv_{xx}+Bu-u^{2}v,\ \ \ \ \ \ 0=\varepsilon Du_{xx}+\varepsilon
288: A+u^{2}v-\left( B+\varepsilon\right) u,\ \ \ \ x\in\left[ 0,1\right] ,\\
289: u^{\prime}=0=\ v^{\prime}\text{ \ \ \ at }\ \ \ x=0\text{ and }x=1.
290: \end{gather}
291: in the limit (\ref{epsD}) and suppose that
292: \begin{align}
293: \label{15apr17:20}A^{2}<2B.
294: \end{align}
295: Then there exists a K-mesa symmetric solution to $(v,u)$ of the following form.
296:
297: Let%
298: \begin{align}
299: \label{wld}w_{0}=3\sqrt{B/2},\ \ \ \ l=\frac{A}{K\sqrt{2B}},~~~~ d = \frac
300: {1}{K}-l,
301: \end{align}
302: \begin{equation}
303: \label{14apr15:50}x_{li}\equiv x_{i}-\frac{l}{2},\ \ \ x_{ri}\equiv
304: x_{i}+\frac{l}{2},\ \ \ x_{i}\equiv\frac{\left( \frac{1}{2}+i\right) }%
305: {K}\ \ \text{for} \ \ i=1\ldots K.
306: \end{equation}
307:
308:
309: For $x$ away from $x_{ri},$ $x_{li}$, we have:%
310: \begin{equation}
311: v\sim\left\{
312: \begin{array}
313: [c]{l}%
314: w_{0},\ \ \ \ x\in\lbrack0,1]\backslash\cup\left[ x_{li},x_{ri}\right] \\
315: \frac{w_{0}}{3},\ \ \ \ x\in\cup\left( x_{li},x_{ri}\right)
316: \end{array}
317: \right.
318: \end{equation}
319: For $x$ near the interfaces $x_{ri},$ $x_{li}$ we have,%
320: \begin{equation}
321: v\sim\left\{
322: \begin{array}
323: [c]{l}%
324: v_{l}(x-x_{li}),\ \ \ \ x-x_{li}\leq O(\sqrt{\frac{\varepsilon D}{B}})\\
325: v_{r}(x-x_{ri}),\ \ \ \ x-x_{ri}\leq O(\sqrt{\frac{\varepsilon D}{B}})
326: \end{array}
327: \right.
328: \end{equation}
329: where
330: \begin{equation}
331: v_{r}=w_{0}\frac{2}{3}+w_{0}\frac{1}{3}\tanh\left( \frac{w_{0}}{3}\frac
332: {x}{\sqrt{2\varepsilon D}}\right) ,\ \ \ \ \ \ \ v_{l}=w_{0}\frac{2}{3}%
333: -w_{0}\frac{1}{3}\tanh\left( \frac{w_{0}}{3}\frac{x}{\sqrt{2\varepsilon D}%
334: }\right) .
335: \end{equation}
336: Finally,
337: \begin{equation}
338: u\sim w_{0}-v.
339: \end{equation}
340:
341: \end{proposition}
342:
343: A typical such solution with $K=3$ is illustrated on Figure \ref{fig:1}.
344:
345: We next analyse the stability of such equilibrium. There are two distinguished
346: limits of interest, either $DK^{2}\ll O\left( \frac{1}{\varepsilon{\ln
347: ^{2}{\varepsilon}}}\right) $ or $DK^{2}=O\left( \frac{1}{\varepsilon{\ln
348: ^{2}{\varepsilon}}}\right) .$ The former is studied in Sections \ref{sec:3}
349: and \ref{sec:4} while the latter in Section \ref{sec:5}. In Section
350: \ref{sec:3} we derive rather precise results for eigenvalues, summarized in
351: the following theorem.
352:
353: \begin{theorem}
354: \label{thm:largeD}Consider a $K$ mesa solution of Proposition \ref{prop:1}.
355: Suppose in addition that
356: \begin{equation}
357: 1\ll DK^{2}\ll O\left( \frac{1}{\varepsilon{\ln^{2}{\varepsilon}}}\right)
358: \text{ \ \ \ \ and \ \ }O\left( \tau-1\right) \gg0.
359: \end{equation}
360: Such solution is stable when $\tau-1\gg0$ and unstable when $\tau-1\ll0$.
361: \ There are $2K$ small eigenvalues of order $O\left( \varepsilon\right) ;$
362: all other eigenvalues are negative and have order $\geq O\left(
363: D\varepsilon\right) .$ The smallest $2K$ eigenvalues are given by%
364: \begin{align}
365: \lambda_{j\pm} & \sim\frac{-1\pm\sqrt{1-2K^{2}dl\left[ 1-\cos\left(
366: \frac{\pi j}{K}\right) \right] }}{2\left( \tau-1\right) }\varepsilon
367: \ \ \text{for}\ \ j=1\ldots K-1;\\
368: \lambda_{-} & \sim\frac{-Kl}{\tau-1}\varepsilon,\ \ \ \ \ \ \lambda
369: _{+}=\frac{-1}{\tau-1}\varepsilon.
370: \end{align}
371: and are all negative when $\tau>1,$ and positive when $\tau<1.$ The transition
372: from stability to instability occurs via a Hopf bifurcation as $\tau$ is
373: decreased past $\tau_{h}$ where to leading order, $\tau_{h}\sim1$.
374: \end{theorem}
375:
376: Note that near the interfaces, the gradient changes on the order $\delta
377: =\sqrt{\varepsilon D}.$ In terms of $\delta,$ the scaling $DK^{2}=O(\frac
378: {1}{\varepsilon\ln^{2}{\varepsilon}})$ can be written as
379: \begin{equation}
380: D=O\left( \delta^{2}\exp\left( \frac{1}{K\delta}\right) \right) .
381: \end{equation}
382: Thus Theorem \ref{thm:largeD} confirms the stability of $K$ mesas when
383: $\tau>1$ as long as $D$ is not exponentially large in $\delta$. In the
384: contrary case, we derive the following result in Section \ref{sec:5}.
385:
386: \begin{theorem}
387: \label{thm:hugeD}Suppose that
388: \begin{equation}
389: \tau>1
390: \end{equation}
391: and let
392: \begin{equation}
393: D_{K}=\frac{1}{K^{2}}D_{1}\text{ \ where \ }D_{1}\sim\left\{
394: \begin{array}
395: [c]{c}%
396: \frac{A^{2}}{2\varepsilon\ln^{2}\left( \frac{12\sqrt{2}AB^{3/2}}%
397: {\varepsilon\left( \sqrt{2B}-A\right) ^{2}}\right) },\ \ \ \ 2A^{2}<B\ \\
398: \frac{\left( \sqrt{2B}-A\right) ^{2}}{2\varepsilon\ln^{2}\left(
399: \frac{12\sqrt{2}}{\varepsilon A}B^{3/2}\right) },\ \ \ \ 2A^{2}>B
400: \end{array}
401: +l.s.t.\right.
402: \end{equation}
403: Here, l.s.t. denotes logarithmically small terms. Then a $K$ mesa symmetric
404: equilibrium with $K\geq2$ is stable if $D<D_{K}$ and is unstable otherwise.
405: Moreover, a single-mesa equilibria $K=1$ is always stable. A more precise
406: value for $D_{1}$ is given in Proposition \ref{prop:expstab}.
407: \end{theorem}
408:
409: Theorem \ref{thm:hugeD} states that the instability threshold occurs when $D$
410: is exponentially large. In this case the exponentially fast decay outside the
411: interface locations must be taken into account. Their exponentially weak
412: interaction is responsible for an eventual loss of stability. Figure
413: \ref{fig:3} illustrates this proposition. Indeed using the parameters used in
414: that simulation we deduce from Proposition \ref{prop:expstab} that
415: $D_{1}=20.96;$ so that $D_{2}=5.28.$ Since $D=10>D_{2},$ the two-mesa
416: equilibrium state is unstable.
417:
418: \begin{figure}[ptbh]
419: \begin{center}
420: \setlength{\unitlength}{1in} \begin{picture}(5.2,3.1)(0,0)
421: \put(2.5, 0){$x$}
422: \put(2.5, 3.0){$v(x,t)$}
423: \put(0.1, 1.6){$t$}
424: \put(0.1,0.0){\includegraphics[width=5in, height=3in]{fig3.eps}}
425: \end{picture}
426: \end{center}
427: \caption{ Slow-time competition instability of a two-mesa solution to
428: (\ref{6dec4:04}). The parameter values are $A=1,~B=8,~D=10,~\varepsilon
429: =0.00025,~\tau=1.3$}%
430: \label{fig:3}%
431: \end{figure}
432:
433: Our next result is about the presence of a Hopf bifurcation when $\tau$ is
434: near 1.
435:
436: \begin{theorem}
437: \label{thm:hopf} Suppose that
438: \begin{equation}
439: \text{ \ \ \ }\sqrt{\frac{B}{\varepsilon D}}\ll DK^{2}\ll O\left( \frac
440: {1}{\varepsilon\ln^{2}{\varepsilon}}\right) . \label{9mar14:03}%
441: \end{equation}
442: Let
443: \begin{equation}
444: \tau_{h_{+}}=1+\frac{1}{4D}\left( ld-\frac{K}{3}\left( d^{3}+l^{3}\right)
445: \right) \label{9mar14:14}%
446: \end{equation}
447: where $K,d,l$ are as in Proposition \ref{prop:1}. Then a K-mesa solution
448: undergoes a Hopf bifurcation when $\tau=\tau_{h_{+}}.$ It is stable when
449: $\tau>\tau_{h_{+}}$ and unstable otherwise. When $\tau=\tau_{h_{+}},$ the
450: corresponding eigenvalue has value%
451: \begin{equation}
452: \lambda_{+}\sim i\sqrt{8K}\left( \varepsilon^{3}DB\right) ^{1/4}.
453: \end{equation}
454:
455: \end{theorem}
456:
457: Figure \ref{fig:turing}.b illustrates the type of oscillations that occur when
458: $\tau< \tau_{h+}$. Theorem \ref{thm:hopf} is able to predict the onset of such
459: oscillations even though it says nothing about whether this bifurcation is
460: supercritical or subcritical.
461:
462: Finally we perform a Turing stability analysis of the Brusselator in Section
463: \ref{sec:6}. We find that in the generic case, the modes $k$ within the Turing
464: instability band have the order $O(\frac{1}{\delta})$ where $\delta=
465: \sqrt{\varepsilon D}$ is the width of the interface. On the other hand the
466: mesa instability threshold occurs when $K=O\left( \frac1{\delta\ln\frac
467: {1}{\varepsilon}}\right) $. It is then clear that $k \gg K$ by a
468: logarithmically large amount. This is the underlying reason for the coarsening
469: process observed in Figure \ref{fig:turing}.a.
470:
471: One of the easy consequences of Turing's analysis is the existence of the
472: regime for which mesa-type structures are stable at the same time as the
473: homogeneous steady state. A more interesting question is the following.
474:
475: \begin{oproblem}
476: Does there exist a parameter regime for which the mesa-type solution is
477: unstable, and in addition the homogeneous steady state $u=A,\ v=\frac{B}{A}$
478: is unstable with respect to the Turing instability?
479: \end{oproblem}
480:
481: An existence of such a regime would imply spatio-temporal chaos in the
482: Brusselator. We answer this question in the negative for the case when
483: \begin{equation}
484: \frac{1}{\delta}\ll D\ll\delta\exp\left( \frac{1}{\delta}\right)
485: ,\ \ \ \ \ \ \delta=\sqrt{\varepsilon D}.
486: \end{equation}
487:
488:
489: \section{Steady state}
490:
491: \label{sec:2}
492:
493: In this section we derive the asymptotics of the steady-state solution to the
494: Brusselator (\ref{6dec4:04}). Let
495: \begin{equation}
496: w=v+u.
497: \end{equation}
498: and rewrite (\ref{6dec4:04}) as
499: \begin{align}
500: v_{t} & =\delta^{2}v_{xx}+B\left( w-v\right) -\left( w-v\right) ^{2}v,\\
501: \frac{1}{\varepsilon}\left( v_{t}+\tau\left( w_{t}-v_{t}\right) \right)
502: & =Dw_{xx}-w+v+A,
503: \end{align}
504: where%
505: \begin{equation}
506: \delta^{2}=\varepsilon D.
507: \end{equation}
508: The steady-state equations then become
509: \begin{gather}\label{15dec2:23}
510: \left\{
511: \begin{array}
512: [c]{c}%
513: 0=\delta^{2}v_{xx}+B\left( w-v\right) -\left( w-v\right) ^{2}v,\\
514: 0=Dw_{xx}-w+v+A
515: \end{array}
516: \right. x\in\lbrack0,L]\\
517: v_{x}(0)=0=v_{x}(L),\ \ \ \ \ w_{x}(0)=0=w_{x}(L).
518: \end{gather}
519: Next we expand (\ref{15dec2:23}) in $\frac{1}{D},$
520: \begin{align}
521: v & =v_{0}+\frac{1}{D}v_{1}+\ldots,\\
522: w & =w_{0}+\frac{1}{D}w_{1}+\ldots
523: \end{align}
524: We obtain $w_{0}^{\prime\prime}=0$ so that $w_{0}$ is a constant to be
525: determined. For $v_{0}$ we obtain
526: \begin{equation}
527: \delta^{2}v_{0xx}=F(v_{0},w_{0}),
528: \end{equation}
529: where
530: \begin{equation}
531: F(v,w)\equiv-B\left( v-w\right) +\left( v-w\right) ^{2}v.
532: \end{equation}
533:
534:
535: We seek solutions for $w_{0}$ such that $F(v_{0},w_{0})$ is a cubic in $v_{0}$
536: with equidistant roots. This is the so-called Maxwell line condition, and
537: implies that the integral of $F$ between its first two roots is the negative
538: of the integral between the last two. When this is the case, the solution for
539: $v_{0}$ will be a kink-like solution in the form of a tanh. An example of such
540: solution is shown on Figure \ref{fig:1}. The condition of equidistant roots is
541: equivalent to solving two equations
542: \begin{equation}
543: F_{vv}(v,w)=0=\ F(v,w)
544: \end{equation}
545: for unknowns $B$ and $w.$ A simple computation shows that this is equivalent
546: to%
547: \begin{equation}
548: B=\frac{2}{9}w_{0}^{2}.\label{Bw}%
549: \end{equation}
550: Substituting for $B$ we obtain
551: \begin{equation}
552: F(v_{0,}w_{0})=\left( v_{0}-w_{0}\right) ^{2}v_{0}-\frac{2}{9}w_{0}%
553: ^{2}\left( v_{0}-w_{0}\right) =\left( v_{0}-\frac{1}{3}w_{0}\right)
554: \left( v_{0}-\frac{2}{3}w_{0}\right) \left( v_{0}-w_{0}\right)
555: .\label{Feq}%
556: \end{equation}
557: On the entire space, the ODE $\delta^{2}v_{0}^{\prime\prime}=F(v_{0},w_{0})$
558: with $F$ as in (\ref{Feq}) admits the following two solutions,
559: \begin{align}
560: v_{r} & =\frac{2}{3}w_{0}+w_{0}\frac{1}{3}\tanh\left( \frac{w_{0}}{3}%
561: \frac{x}{\sqrt{2}\delta}\right) ,\label{vr}\\
562: v_{l} & =\frac{2}{3}w_{0}-w_{0}\frac{1}{3}\tanh\left( \frac{w_{0}}{3}%
563: \frac{x}{\sqrt{2}\delta}\right) .\label{vl}%
564: \end{align}
565: We are interested in mesa-type solutions. A single, symmetric mesa-type
566: solution on an interval $[0,L]$ has the form,%
567: \begin{equation}
568: v_{0}\sim\left\{
569: \begin{array}
570: [c]{c}%
571: v_{l}\left( x-x_{l}\right) ,\ \ \ x<\frac{L}{2}\\
572: v_{r}\left( x-x_{r}\right) ,\ \ x>\frac{L}{2}%
573: \end{array}
574: .\right.
575: \end{equation}
576: Here, we choose%
577: \begin{equation}
578: x_{l}=\frac{L-l}{2},\ \ \ \ x_{r}=\frac{L+l}{2}.
579: \end{equation}
580: where $l$ is the width of the mesa to be determined. A K-spike symmetric
581: solution on the interval $\left[ 0,1\right] $ is then obtained by glueing
582: together $K$ solutions on the interval $L=\frac{1}{K}.$ Such a solution has
583: $2K$ interfaces whose locations are given by $x_{li},x_{ri}$ defined in
584: (\ref{14apr15:50}). To find $l$ we need the second order equations. We have,
585: \begin{equation}
586: \delta^{2}v_{1xx}=F_{v}(v_{0,}w_{0})v_{1}+F_{w}(v_{0,}w_{0})w_{1},\label{v1}%
587: \end{equation}%
588: \begin{equation}
589: w_{1xx}=w_{0}-v_{0}-A,\label{w1}%
590: \end{equation}
591: where
592: \begin{equation}
593: F_{v}\left( v,w\right) =B+\left( v-w\right) \left( 3v-w\right)
594: ,\ \ \ F_{w}\left( v,w\right) =-B+2\left( w-v\right) v.
595: \end{equation}
596: Integrating (\ref{w1}) and
597: using the boundary condition $w_{1}^{\prime}\left( \pm L\right) =0$ we
598: obtain
599: \begin{equation}
600: \int_{0}^{L}\left( w_{0}-v_{0}-A\right) =0.
601: \end{equation}
602: We evaluate
603: \begin{equation}
604: \int_{0}^{L}v_{0}\sim l\frac{w_{0}}{3}+w_{0}\left( L-l\right)
605: \end{equation}
606: from where
607: \begin{equation}
608: l\sim\frac{3}{2}\frac{LA}{w_{0}}\sim\frac{LA}{\sqrt{2B}}.\label{lwB}%
609: \end{equation}
610: Substituting $L=\frac{1}{K}$ then yields Proposition \ref{prop:1}.
611:
612: We remark that the equation (\ref{w1}) for $w_{1}$ with Neumann boundary
613: condition is solvable up to a constant. See Section \ref{sec:5}, Lemma
614: \ref{lemma:9} for the determination of this constant.
615:
616: \section{Stability in the regime $DK^{2}\ll O\left( \frac{1}{\varepsilon}%
617: {\ln^{2}{\varepsilon}}\right) $}
618:
619: \label{sec:3}
620:
621: In this section we derive Theorem \ref{thm:largeD}, valid when $D \ll O\left(
622: \frac{1}{\varepsilon\ln^{2} \varepsilon}\right) .$ Before showing this
623: result, we derive a more general formula for eigenvalues which is valid for
624: all $\tau.$ We show the following.
625:
626: \begin{lemma}
627: \label{lemma:4} Consider a K-mesa symmetric equilibrium solution as given in
628: Proposition \ref{prop:1}. Moreover suppose that
629: \begin{equation}
630: 1\ll DK^{2}\ll O\left( \frac{1}{\varepsilon}{\ln^{2}{\varepsilon}}\right) .
631: \end{equation}
632: The eigenvalues of such equilibrium state are asymptotically given implicitly
633: by%
634: \begin{equation}
635: \lambda\sim2\sqrt{B\frac{\varepsilon}{D}}\left( ldK-\frac{2}{\sigma}%
636: \frac{\left( \tau-1\right) \lambda+\varepsilon}{\varepsilon}\right)
637: \label{12dec1:52}%
638: \end{equation}
639: where $\sigma$ may take one of the following $2K$ values%
640: \begin{gather}
641: \sigma_{j\pm}=\left( c\pm\sqrt{a^{2}+b^{2}+2ab\cos\left( \theta\right)
642: }\right) ,\ \ \theta=\frac{\pi j}{K}\ \ \text{for}\ \ j=1\ldots K-1,\\
643: \sigma_{\pm}=c+a\pm b,
644: \end{gather}
645: where%
646: \begin{equation}
647: a=\frac{-\mu_{d}}{\sinh\left( \mu_{d}d\right) },\ \ \ \ \ \ \ b=\frac
648: {-\mu_{l}}{\sinh\left( \mu_{l}l\right) },\ \ \ \ \ \ \ c=\mu_{d}\coth\left(
649: \mu_{d}d\right) +\mu_{l}\coth\left( \mu_{l}l\right) ,
650: \end{equation}%
651: \begin{equation}
652: \mu_{l}\equiv\frac{\sqrt{2\varepsilon+\lambda\left( 2\tau-1\right) }}%
653: {\delta},\ \ \ \ \ \mu_{d}\equiv\frac{\sqrt{\lambda}}{\delta}.
654: \end{equation}
655:
656: \end{lemma}
657:
658: \textbf{Proof}. We start by linearizing around the equilibrium solution
659: $(v,w).$ We write,
660: \begin{equation}
661: v(x,t)=v\left( x\right) +e^{\lambda t}\phi\left( x\right)
662: ,\ \ \ w(x,t)=w\left( x\right) +e^{\lambda t}\psi\left( x\right)
663: \end{equation}
664: where we assume that $\psi$ and $\phi$ are small. We obtain
665: \begin{subequations}
666: \begin{align}
667: \lambda\phi & =\delta^{2}\phi_{xx}-F_{v}(v,w)\phi-F_{w}(v,w)\psi
668: ,\label{eqphi}\\
669: \frac{1}{\varepsilon}\left( \phi+\tau\left( \psi-\phi\right) \right)
670: \lambda & =D\psi_{xx}-\psi+\phi, \label{eqpsi}%
671: \end{align}
672: where $\delta=\sqrt{\varepsilon D}$. Using (\ref{Bw}) we obtain
673: \end{subequations}
674: \begin{align}
675: F_{v}(v_{0},w_{0}) & =3v_{0}^{2}-4w_{0}v_{0}+\left( w_{0}^{2}+B\right)
676: =3v_{0}^{2}-4w_{0}v_{0}+\frac{11}{9}w_{0}^{2},\\
677: F_{w}(v_{0},w_{0}) & =-2v_{0}^{2}+2w_{0}v_{0}-B=-2v_{0}^{2}+2w_{0}%
678: v_{0}-\frac{2}{9}w_{0}^{2}%
679: \end{align}
680: so that
681: \begin{align}
682: F_{v}(w_{0},w_{0}) & =B,\ \ \ \ F_{w}(w_{0},w_{0})=-B,\\
683: F_{v}(\frac{w_{0}}{3},w_{0}) & =B,\ \ \ \ F_{w}(\frac{w_{0}}{3},w_{0})=B.
684: \end{align}
685:
686:
687: Note that away from kink locations $x_{li},x_{ri},$ the diffusion term
688: $\varepsilon D\phi^{\prime\prime}$ may be neglected and we have $\phi
689: \sim-\frac{F_{w}(v_{0},w_{0})}{F_{v}(v_{0},w_{0})}\psi.$ On the other hand,
690: near the kink locations we locally estimate the eigenfunction by the
691: derivative of the profile. This suggests the following asymptotic form:%
692: \begin{equation}
693: \phi\sim\left\{
694: \begin{array}
695: [c]{l}%
696: c_{li}v_{li}^{\prime},\ \ \ x\sim x_{li}\\
697: c_{ri}v_{ri}^{\prime},\ \ \ x\sim x_{ri}\\
698: \psi,\ \ \ x\notin\left( x_{li},x_{ri}\right) ,\ \ i=1\ldots K\\
699: -\psi,\ \ \ x\in\left( x_{li},x_{ri}\right) ,\ \ i=1\ldots K
700: \end{array}
701: \right.
702: \end{equation}
703: where the constants $c_{li}$ and $c_{ri}$ are to be found. We multiply
704: (\ref{eqphi}) by $v_{li}^{\prime}$ and integrate. Because of exponential
705: decay, we obtain that $\int v_{li}^{\prime}\phi\sim c_{li}\int v_{li}%
706: ^{\prime2}.$ Therefore%
707: \begin{equation}
708: \lambda\int v_{li}^{\prime2}+\int v_{li}^{\prime}F_{w}(v_{li},w_{0})\psi
709: \sim\int\left[ \delta^{2}v_{li}^{\prime}\phi_{xx}-v_{li}^{\prime}\phi
710: F_{v}\left( v_{li},w_{0}\right) -v_{li}^{\prime}\phi\frac{1}{D}\left(
711: F_{vv}\left( v_{li},w_{0}\right) v_{1}+F_{vw}\left( v_{li},w_{0}\right)
712: w_{1}\right) \right] .
713: \end{equation}
714: Using integration by parts we obtain,%
715: \begin{equation}
716: \int\left[ \varepsilon Dv_{li}^{\prime}\phi_{xx}-v_{li}^{\prime}\phi
717: F_{v}\left( v_{li},w_{0}\right) \right] \sim0.
718: \end{equation}
719: Next we evaluate
720: \begin{equation}
721: I=\int v_{li}^{\prime}\phi\left( F_{vv}\left( v_{li},w_{0}\right)
722: v_{1}+F_{vw}\left( v_{li},w_{0}\right) w_{1}\right) .
723: \end{equation}
724: Differentiating (\ref{v1}) we have%
725: \begin{equation}
726: \delta^{2}v_{1}^{\prime\prime}-v_{l}^{\prime}\left( F_{vv}\left(
727: v_{li},w_{0}\right) v_{1}+F_{vw}\left( v_{li},w_{0}\right) w_{1}\right)
728: -F_{v}\left( v_{li},w_{0}\right) v_{1}^{\prime}-F_{w}\left( v_{li}%
729: ,w_{0}\right) w_{1}^{\prime}=0
730: \end{equation}
731: so that
732: \begin{align}
733: I & =\int\phi\left( \delta^{2}v_{1}^{\prime\prime}-F_{v}\left(
734: v_{li},w_{0}\right) v_{1}^{\prime}-F_{w}\left( v_{li},w_{0}\right)
735: w_{1}^{\prime}\right) \\
736: & \sim-\int c_{li}v_{l}^{\prime}F_{w}\left( v_{li},w_{0}\right)
737: w_{1}^{\prime}.
738: \end{align}
739: Therefore we obtain%
740: \begin{equation}
741: c_{li}\lambda\int_{x_{li}^{-}}^{x_{li}^{+}}v_{li}^{\prime2}+\psi\left(
742: x_{li}\right) \int_{x_{li}^{-}}^{x_{li}^{+}}v_{li}^{\prime}F_{w}(v_{0}%
743: ,w_{0})\sim c_{li}w_{1}^{\prime}(x_{li})\int_{x_{li}^{-}}^{x_{li}^{+}}%
744: v_{li}^{\prime}F_{w}\left( v_{0},w_{0}\right) .
745: \end{equation}
746: Here and below, the symbol $\int_{x_{li}^{-}}^{x_{li}^{+}}$ denotes
747: integration over the interface located at $x_{li}$. Since $v_{0}^{\prime}$
748: decays exponentially outside the interface, this symbol is unambigious; that
749: is $\int_{x_{li}^{-}}^{x_{li}^{+}}=\int_{0}^{1}+e.s.t.$ Next we show that
750: \begin{equation}\label{15dec2:31}
751: \int_{x_{li}^{-}}^{x_{li}^{+}}v_{li}^{\prime}F_{w}\left( v_{0},w_{0}\right)
752: \sim-\frac{8}{81}w_{0}^{3},\ \ \ \ \int_{x_{li}^{-}}^{x_{li}^{+}}%
753: v_{li}^{\prime2}\sim\frac{2\sqrt{2}w_{0}^{3}}{81\delta}.
754: \end{equation}
755: We have $\int_{x_{l}^{-}}^{x_{l}^{+}}v_{li}^{\prime}F_{w}\left( v_{0}%
756: ,w_{0}\right) =G(v(x_{li}^{+}))-G(v(x_{li}^{-}))$ where $G(v)\equiv\int
757: F_{w}(v,w_{0})dv=-\frac{2}{3}v^{3}+w_{0}v^{2}-\frac{2}{9}w_{0}^{2}v.$ We have
758: $v(x_{li}^{+})=w_{0}/3,\ v(x_{li}^{-})=w_{0}$ so that $G(v(x_{li}%
759: ^{+}))-G(v(x_{li}^{-}))=\left[ \frac{1}{81}-\frac{1}{9}\right] w_{0}%
760: ^{3}=-\frac{8}{81}w_{0}^{3}.$
761:
762: To evaluate the second integral in (\ref{15dec2:31}),
763: use the explicit formula (\ref{vl}) for
764: $v_{l}$ and the fact that $\int_{-\infty}^{+\infty}\operatorname{sech}%
765: ^{4}\left( y\right) dy=\frac{4}{3}.$
766:
767: Using (\ref{w1}) and (\ref{lwB}) yields%
768: \begin{equation}
769: w_{1}^{\prime}\left( x_{li}\right) \sim-\frac{w_{0}ldK}{3D}.\
770: \end{equation}
771:
772:
773: Therefore we obtain%
774: \begin{equation}
775: \label{23feb16:02}c_{li}\lambda\frac{\sqrt{2}}{\delta}=4\left( c_{li}\left(
776: \frac{1}{3} \frac{ldK}{D}w_{0}\right) +\psi\left( x_{li}\right) \right) ,
777: \end{equation}
778: An analogous computation yields%
779: \begin{equation}
780: \label{23feb16:03}-c_{ri}\lambda\frac{\sqrt{2}}{\delta}=4\left(
781: -c_{ri}\left( \frac{1}{3}\frac{ldK}{D}w_{0}\right) +\psi\left(
782: x_{ri}\right) \right) ,
783: \end{equation}
784:
785:
786: It remains to compute $\ \psi\left( x_{li}\right) .$ Inside the intervals
787: $\left( x_{li},x_{ri}\right) $ we have $\phi\sim-\psi$ and outside those
788: intervals we have $\phi\sim\psi.$ In addition we assume that near kinks,
789: $\psi$ changes much slower than $\phi.$ In this case we may replace
790: \begin{equation}
791: v_{li}^{\prime}\sim-\frac{2}{3}w_{0}{\mbox{\boldmath$\delta$}}\left(
792: x-x_{li}\right) ,\ \ v_{ri}^{\prime}\sim\frac{2}{3}w_{0}%
793: {\mbox{\boldmath$\delta$}}\left( x-x_{ri}\right)
794: \end{equation}
795: inside the equation (\ref{eqpsi}). Here and below, ${\mbox{\boldmath$\delta$}}%
796: $ denotes the Dirac delta function. Therefore we obtain,%
797: \begin{equation}
798: \psi_{xx}-\mu^{2}\psi\sim-\alpha\sum_{i=1}^{K}c_{li}{\mbox{\boldmath$\delta$}}%
799: \left( x-x_{li}\right) -c_{ri}{\mbox{\boldmath$\delta$}}\left(
800: x-x_{ri}\right)
801: \end{equation}
802: where
803: \begin{align}\label{mul}
804: \mu & \sim\mu_{l}\equiv\frac{\sqrt{2\varepsilon+\lambda\left( 2\tau
805: -1\right) }}{\delta},\ \ \ x\in\left( x_{li},x_{ri}\right) ,\\
806: \label{mud}
807: \mu & \sim\mu_{d}\equiv\frac{\sqrt{\lambda}}{\delta},\ \ \ x\notin\left(
808: x_{li},x_{ri}\right) ,
809: \end{align}
810: and%
811: \begin{equation}
812: \alpha=\frac{2}{3}w_{0}\frac{\left( 1-\tau\right) \lambda-\varepsilon
813: }{\delta^{2}}.
814: \end{equation}
815: Next we apply the following lemma.
816:
817: \begin{lemma}
818: \label{lemma:1} Suppose that
819: \begin{align}
820: u^{\prime\prime}-\mu^{2}u & =-\left( \sum b_{li}{\mbox{\boldmath$\delta$}}%
821: \left( x-x_{li}\right) +b_{ri}{\mbox{\boldmath$\delta$}}\left(
822: x-x_{ri}\right) \right) \label{15dec1:49}\\
823: \text{with \ }u^{\prime}\left( 0\right) & =0=u^{\prime}\left( 1\right) ,
824: \end{align}
825: where $x_{li},x_{ri}$ are given in (\ref{14apr15:50}) and
826: \begin{equation}
827: \mu=\left\{
828: \begin{array}
829: [c]{cc}%
830: \mu_{l}, & x\in\left( x_{li},x_{ri}\right) \\
831: \mu_{d}, & x\notin\left( x_{li},x_{ri}\right)
832: \end{array}
833: \right. .
834: \end{equation}
835: Let
836: \begin{equation}
837: \vec{u}\equiv\left[
838: \begin{array}
839: [c]{c}%
840: u\left( x_{l1}\right) \\
841: u\left( x_{r1}\right) \\
842: \vdots\\
843: u\left( x_{lK}\right) \\
844: u\left( x_{rK}\right)
845: \end{array}
846: \right] ,\ \ \ \ \ \ \vec{b}\equiv\left[
847: \begin{array}
848: [c]{c}%
849: b_{l1}\\
850: b_{r1}\\
851: \vdots\\
852: b_{lK}\\
853: b_{rK}%
854: \end{array}
855: \right] .\ \ \ \ \ \
856: \end{equation}
857: \ \ \ \ \ \ \ \ \ \ Then
858: \begin{equation}
859: \vec{u}=\mathbf{M}\vec{b},
860: \end{equation}
861: where
862: \begin{equation}
863: M^{-1}=\left[
864: \begin{array}
865: [c]{ccccccc}%
866: a+c & b & & & & & \\
867: b & c & a & & & & \\
868: & a & c & b & & & \\
869: & & & \cdots & & & \\
870: & & & & a & c & b\\
871: & & & & & b & c+a
872: \end{array}
873: \right]
874: \end{equation}
875: with
876: \begin{equation}
877: a=\frac{-\mu_{d}}{\sinh\left( \mu_{d}d\right) },\ \ \ \ \ \ \ b=\frac
878: {-\mu_{l}}{\sinh\left( \mu_{l}l\right) },\ \ \ \ \ \ \ c=\mu_{d}\coth\left(
879: \mu_{d}d\right) +\mu_{l}\coth\left( \mu_{l}l\right) .
880: \end{equation}
881: The eigenvalues $\sigma$ of $\mathbf{M}^{-1}$ are given as follows%
882: \begin{gather}
883: \sigma_{j\pm}=\left( c\pm\sqrt{a^{2}+b^{2}+2ab\cos\left( \theta\right)
884: }\right) ,\ \ \theta=\frac{\pi j}{K}\ \ for\ \ j=1\ldots K-1\\
885: \sigma_{\pm}=c+a\pm b.
886: \end{gather}
887: The two eigenvalues $\sigma_{\pm}$ may be written explicitly as%
888: \begin{align}
889: \sigma_{+} & =\mu_{d}\tanh\left( \mu_{d}d/2\right) +\mu_{l}\tanh\left(
890: \mu_{l}l/2\right) ,\\
891: \sigma_{-} & =\mu_{d}\tanh\left( \mu_{d}d/2\right) +\mu_{l}\coth\left(
892: \mu_{l}l/2\right) .
893: \end{align}
894:
895: \end{lemma}
896:
897: This lemma was derived in \cite{RW}; for convenience of the reader we give its
898: proof in Appendix A.
899:
900: Define
901: \begin{equation}
902: \vec{\psi}=\left[
903: \begin{array}
904: [c]{c}%
905: \psi\left( x_{l1}\right) \\
906: \psi\left( x_{r1}\right) \\
907: \vdots\\
908: \psi\left( x_{lK}\right) \\
909: \psi\left( x_{rK}\right)
910: \end{array}
911: \right] ,\ \ \ \ \ \ \ \vec{d}=\left[
912: \begin{array}
913: [c]{c}%
914: c_{l1}\\
915: -c_{r1}\\
916: \vdots\\
917: c_{lK}\\
918: -c_{rK}%
919: \end{array}
920: \right] .
921: \end{equation}
922: Using Lemma \ref{lemma:1} we have,%
923: \begin{equation}
924: \vec{\psi}=\alpha\mathbf{M}\vec{d}%
925: \end{equation}
926: and we write (\ref{23feb16:02}, \ref{23feb16:03}) as%
927:
928: \begin{equation}
929: \vec{d}\lambda\sqrt{\frac{2}{\varepsilon D}.}=4\left( \vec{d}\left( \frac
930: {1}{3}\frac{ldK}{D}w\right) +\alpha\mathbf{M}\vec{d}\right) .
931: \end{equation}
932: Therefore we have%
933: \begin{equation}
934: \lambda_{j}\sqrt{\frac{2}{\varepsilon D}}=4\left( \frac{1}{3}\frac{ldK}%
935: {D}w+\alpha\mathbf{\sigma}^{-1}\right) \ \ \text{for}\ \ j=1\ldots2K
936: \end{equation}
937: where $\sigma$ are $2K$ eigenvalues of $\mathbf{M}^{-1}\mathbf{.}$
938:
939: This completes the proof of Lemma \ref{lemma:4}. $\blacksquare$
940:
941: We now come back to the proof of Theorem \ref{thm:largeD}. Assuming
942: $O(\tau-1)\gg0,$ the dimensional analysis shows that two scalings are
943: possible, either $\lambda=O(\varepsilon)$ or $\lambda=O(D\varepsilon).$
944:
945: We first consider the case
946: \begin{equation}
947: \lambda=\varepsilon\lambda_{0}.
948: \end{equation}
949: Using the notation of Lemma \ref{lemma:4} we then obtain from (\ref{mul}) and
950: (\ref{mud})
951: \begin{equation}
952: \mu_{d}^{2}=\frac{\lambda_{0}}{D}\ll1,\ \ \ \ \mu_{l}^{2}=\frac{2+\lambda
953: _{0}\left( 2\tau-1\right) }{D}\ll1.
954: \end{equation}
955: and
956: \begin{equation}
957: a\sim-\frac{1}{d},\ \ b\sim-\frac{1}{l},\ \ \ \ c\sim\frac{1}{d}+\frac{1}%
958: {l}=\frac{1}{Kdl},
959: \end{equation}%
960: \begin{align}
961: \sigma_{j\pm} & =c\pm\sqrt{\left( a+b\right) ^{2}-2abt}%
962: ,\ \ \ \ \ \ \ \ t=1-\cos\left( \frac{\pi j}{K}\right) \in\left( 0,2\right)
963: \\
964: & =\frac{1}{Kdl}\pm\sqrt{\left( \frac{1}{Kdl}\right) ^{2}-2\frac{1}{dl}t}\\
965: & =\frac{1}{Kdl}\left( 1\pm\sqrt{1-2K^{2}dlt}\right) .
966: \end{align}
967: Therefore we obtain%
968: \begin{equation}
969: \lambda_{0}\sim2ldK\sqrt{\frac{B}{\varepsilon D}}\left( 1-2\frac{\left(
970: \tau-1\right) \lambda_{0}+1}{1\pm\sqrt{1-2K^{2}dlt}}\right) .
971: \end{equation}
972: When $\lambda_{0}=O\left( 1\right) ,$ the right hand side dominates and we
973: therefore obtain%
974: \begin{equation}
975: \lambda_{0}=\frac{-1\pm\sqrt{1-2K^{2}dlt}}{2\left( \tau-1\right) }.
976: \label{8mar18:01}%
977: \end{equation}
978: Moreover, since $d=\frac{1}{K}-l,\ \ l\in\left( 0,\frac{1}{K}\right) ,$ and
979: since $t\in(0,2)$ we see that $2K^{2}dlt\leq1.$ This shows that in the case
980: $\tau>1,$ the eigenvalues corresponding to the nodes $\sigma_{j\pm}$ are all
981: real negative. The node $\sigma_{+}$ is,%
982:
983: \begin{align}
984: \sigma_{+} & =\mu_{d}\tanh\left( \mu_{d}\frac{d}{2}\right) +\mu_{l}%
985: \tanh\left( \mu_{l}\frac{l}{2}\right) \\
986: & \sim\mu_{d}^{2}\frac{d}{2}+\mu_{l}^{2}\frac{l}{2}\\
987: & \sim\frac{1}{2D}\left( \lambda_{0}d+\left( 2+\lambda_{0}\left(
988: 2\tau-1\right) \right) l\right) ,\\
989: \lambda_{0} & \sim2\sqrt{\frac{B}{\varepsilon D}}\left( ldK-\frac{4D\left[
990: \left( \tau-1\right) \lambda_{0}+1\right] }{\left( \lambda_{0}d+\left(
991: 2+\lambda_{0}\left( 2\tau-1\right) \right) l\right) }\right) .
992: \end{align}
993: Since the third term is asymptotically bigger, the assumption $\lambda
994: _{0}=O(1)$ leads to%
995: \begin{equation}
996: \lambda_{0}=\frac{-1}{\tau-1}.
997: \end{equation}
998: Note that this is a special case of the formula (\ref{8mar18:01})\ with
999: $\pm=-$ and $t=0.$ For the mode $\sigma_{-}$ we have%
1000: \begin{equation}
1001: \sigma_{-}=\mu_{d}\tanh\left( \mu_{d}\frac{d}{2}\right) +\mu_{l}\coth\left(
1002: \mu_{l}\frac{l}{2}\right) \sim\frac{2}{l};
1003: \end{equation}%
1004: \begin{align}
1005: \lambda_{0} & \sim2l\sqrt{B\frac{1}{\varepsilon D}}\left( dK-\left[
1006: \left( \tau-1\right) \lambda_{0}+1\right] \right) ,\\
1007: \lambda_{0} & \sim\frac{-1+Kd}{\tau-1}\sim\frac{-Kl}{\tau-1}.
1008: \end{align}
1009: Note that this is a special case of the formula (\ref{8mar18:01})\ with
1010: $\pm=-$ and $t=2.$
1011:
1012: Next we consider large eigenvalues, $\lambda\gg O(\varepsilon).$ Then we have
1013: \begin{equation}
1014: \mu_{l}\sim\frac{\sqrt{\left( 2\tau-1\right) \lambda}}{\sqrt{D\varepsilon}%
1015: },\ \ \ \mu_{d}\sim\frac{\sqrt{\lambda}}{\sqrt{D\varepsilon}},
1016: \end{equation}
1017: and we write,
1018: \begin{equation}
1019: \lambda\sim2\sqrt{B}\left( \frac{1}{D}\sqrt{D\varepsilon}ldK-\frac{2}{\left(
1020: \sigma/\frac{\sqrt{\lambda}}{\sqrt{D\varepsilon}}\right) }\left(
1021: \tau-1\right) \sqrt{\lambda}\right) .
1022: \end{equation}
1023: Dimensional analysis shows that the only way to achieve this balance is when
1024: $\sigma$ is large. But this is only possible when%
1025: \begin{equation}
1026: \sinh\left( \mu_{l}l\right) =0\text{ \ \ or \ \ \ }\sinh\left( \mu
1027: _{d}d\right) =0.
1028: \end{equation}
1029: Thus either $\mu_{l}l=i\pi m$ or $\mu_{d}d=i\pi m$ where $m$ is some integer.
1030: This yields the following eigenvalues,%
1031: \begin{equation}
1032: \lambda\sim-D\varepsilon\frac{m^{2}\pi^{2}}{l^{2}\left( 2\tau-1\right)
1033: }\text{ \ \ \ \ or \ \ }\lambda\sim-D\varepsilon\frac{m^{2}\pi^{2}}{d^{2}}.
1034: \end{equation}
1035:
1036:
1037: Finally, we show that a Hopf bifurcation occurs in the regime $O\left(
1038: \tau-1\right) \ll1.$ Since the small eigenvalues are negative when $\tau
1039: -1\gg0$ and positive when $\tau-1\ll0$, the real part changes sign precisely
1040: when $O\left( \tau-1\right) \ll1.$ \ To show that this occurs via a\ Hopf
1041: bifurcation, it suffices to show that $\lambda$ can never be zero. Suppose
1042: not. Then using some algebra we arrive at the following for the modes
1043: $\sigma_{j\pm}$%
1044: \begin{equation}
1045: -1\pm\sqrt{1-2K^{2}dlt}=0,\text{ \ }\ t=1-\cos\left( \frac{\pi j}{K}\right)
1046: \in\left( 0,2\right) .
1047: \end{equation}
1048: But this is clearly impossible since $K^{2}dlt<\frac{1}{4}$ as mentioned
1049: above. The modes $\sigma_{\pm}$ are handled similarly.
1050:
1051: This completes the proof of Theorem \ref{thm:largeD}. $\blacksquare$
1052:
1053: \section{\bigskip Hopf bifurcation, $DK^{2}\ll O\left( \frac{1}{\varepsilon
1054: }{\ln^{2}{\varepsilon}}\right) $}
1055:
1056: \label{sec:4}
1057:
1058: As Theorem \ref{thm:largeD} shows, a Hopf bifurcation occurs when $\tau$ is
1059: near 1. In this section we study this regime in more detail, culminating in
1060: the proof of Theorem \ref{thm:hopf}.
1061:
1062: We start by analysing the $\sigma_{+}$ node. We make the assumption that%
1063: \begin{equation}
1064: \lambda\ll O\left( \varepsilon D\right) .
1065: \end{equation}
1066: This assumption will be verified later on to be consistent with the condition
1067: (\ref{9mar14:03}), $D\gg\sqrt{\frac{B}{\varepsilon D}}$. We have,
1068: \begin{equation}
1069: \mu_{l}^{2}\sim\frac{\lambda+2\varepsilon}{\varepsilon D}\ll1,\ \ \ \mu
1070: _{d}^{2}\sim\frac{\lambda}{\varepsilon D}\ll1,
1071: \end{equation}
1072: and we expand $\sigma_{+}$ up to second order,%
1073: \begin{align}
1074: \sigma_{+} & =\mu_{d}\tanh\left( \mu_{d}\frac{d}{2}\right) +\mu_{l}%
1075: \tanh\left( \mu_{l}\frac{l}{2}\right) \\
1076: & \sim\frac{\lambda}{\varepsilon D}\frac{d}{2}-\frac{1}{24}\left(
1077: \frac{\lambda}{\varepsilon D}\right) ^{2}d^{3}+\frac{\lambda+2\varepsilon
1078: }{\varepsilon D}\frac{l}{2}-\frac{1}{24}\left( \frac{\lambda+2\varepsilon
1079: }{\varepsilon D}\right) ^{2}l^{3}\\
1080: & \sim-\left( \frac{\lambda}{\varepsilon D}\right) ^{2}\left( \frac
1081: {d^{3}+l^{3}}{24}\right) +\frac{\lambda}{\varepsilon D}\left( \frac{1}%
1082: {2K}\right) +\frac{l}{D}.
1083: \end{align}
1084: We write the equation for $\lambda$ as%
1085: \begin{equation}
1086: \lambda\sigma\sim2\sqrt{B\frac{\varepsilon}{D}}\left( ldK\sigma
1087: -2\frac{\left( \tau-1\right) \lambda+\varepsilon}{\varepsilon}\right)
1088: \end{equation}
1089: or%
1090: \begin{equation}
1091: a\lambda^{3}+b\lambda^{2}+c\lambda+e=0,
1092: \end{equation}
1093: where%
1094: \begin{align}
1095: a & =\left( \frac{1}{\varepsilon D}\right) ^{2}\left( \frac{d^{3}+l^{3}%
1096: }{24}\right) ,\\
1097: b & =-\frac{1}{\varepsilon D}\left( \frac{1}{2K}\right) -2\sqrt
1098: {B\frac{\varepsilon}{D}}ldK\left( \frac{1}{\varepsilon D}\right) ^{2}\left(
1099: \frac{d^{3}+l^{3}}{24}\right) \\
1100: & \sim-\frac{1}{\varepsilon D}\left( \frac{1}{2K}\right) \ \ \text{using
1101: (\ref{9mar14:03});}\\
1102: c & =-\frac{l}{D}+2\sqrt{B\frac{\varepsilon}{D}}\left( \frac{ld}%
1103: {2\varepsilon D}-\frac{2}{\varepsilon}\left( \tau-1\right) \right) \\
1104: & \sim2\sqrt{B\frac{\varepsilon}{D}}\left( \frac{ld}{2\varepsilon D}%
1105: -\frac{2}{\varepsilon}\left( \tau-1\right) \right) ;\\
1106: e & =2\sqrt{B\frac{\varepsilon}{D}}\left( ldK\left\{ \frac{l}{D}\right\}
1107: -2\right) \\
1108: & \sim-4\sqrt{B\frac{\varepsilon}{D}}.
1109: \end{align}
1110: Substituting $\lambda=i\lambda_{i}$ and separating the real and imaginary
1111: part, we find that%
1112: \begin{equation}
1113: \lambda_{i}=\sqrt{\frac{e}{b}},\ \ \ c=\frac{ae}{b}.
1114: \end{equation}
1115: This yields,%
1116: \begin{equation}
1117: \lambda_{i}=\sqrt{8K}\left( \varepsilon^{3}DB\right) ^{1/4}%
1118: \end{equation}
1119: and, keeping also all lower order terms for reference,%
1120: \begin{equation}
1121: \tau-1\sim\frac{ld}{4D}-\frac{\left( \frac{d^{3}+l^{3}}{12}\right)
1122: K}{D+\sqrt{B\frac{1}{\varepsilon D}}ldK^{2}\left( \frac{d^{3}+l^{3}}%
1123: {6}\right) }-\frac{l\varepsilon}{4\sqrt{B\varepsilon D}}. \label{9mar14:15}%
1124: \end{equation}
1125: The formula (\ref{9mar14:14}) is obtained by dropping the last term of the
1126: right hand side of (\ref{9mar14:15}) (which is of smaller order than the first
1127: term on the right hand side), as well as dropping the second term in the
1128: denominator of the second term.
1129:
1130: Finally we must also show that the $\sigma_{+}$ mode undergoes a Hopf
1131: bifurcation before all other modes.\ Let us now consider the $\sigma_{-}$
1132: mode. \ We assume here that $\mu_{l}\sim\frac{\sqrt{\lambda}}{\sqrt
1133: {\varepsilon D}}\sim\mu_{d}.$ and we rewrite (\ref{12dec1:52}) as
1134: \begin{align}\label{15dec2:43}
1135: -4\frac{1}{D}\sqrt{\frac{B}{\varepsilon D}}\left( \tau_{0}\lambda
1136: _{0}D+1\right) & =F(\lambda_{0})\equiv\lambda_{0}\sqrt{\lambda_{0}}\left(
1137: \tanh\sqrt{\lambda_{0}}\frac{d}{2}+\coth\sqrt{\lambda_{0}}\frac{l}{2}\right)
1138: \\
1139: \text{where }\tau & =1+\tau_{0},\text{ \ }\lambda=\varepsilon D\lambda_{0}.
1140: \end{align}
1141: Near the origin and for real $\lambda_{0}$, the curve $F\left( \lambda
1142: _{0}\right) \sim\frac{2}{l}\lambda_{0}+\left( \frac{d}{2}+\frac{l}%
1143: {6}\right) \lambda_{0}^{2}$ is convex and increasing. The left hand side
1144: of (\ref{15dec2:43}) is a
1145: line in $\lambda_{0}$ and it intersects the y axis at $-\frac{4}{D}\sqrt
1146: {\frac{B}{\varepsilon D}}$ which is a very small value by assumption
1147: (\ref{9mar14:03}). Therefore this line will intersect the curve $F\left(
1148: \lambda_{0}\right) $ for some small value of $\lambda_{0}$ unless its slope
1149: is precisely the slope of $F(\lambda_{0})$ at the origin, i.e. $-4\sqrt
1150: {\frac{B}{\varepsilon D}}\tau_{0}\sim\frac{2}{l}.$ This is precisely the
1151: scaling on which the Hopf bifurcation occurs. Subsituting $l=\frac{A}%
1152: {\sqrt{2B}}$ we obtain,%
1153: \begin{equation}
1154: \tau_{h_{-}}\sim1-\frac{\sqrt{\varepsilon D}}{\sqrt{2}A}. \label{10mar18:11}%
1155: \end{equation}
1156: Performing a similar study for the modes $\sigma_{j\pm}$ we obtain
1157: \begin{equation}
1158: \tau_{h_{j\pm}}\sim1-\frac{\sqrt{\varepsilon D}}{\sqrt{B}4Kdl}\left(
1159: 1\pm\sqrt{1-2K^{2}dlt}\right) ,\ \ \ \ t=\ 1-\cos\left( \frac{\pi j}%
1160: {K}\right) \in\left( 0,2\right) . \label{10mar18:12}%
1161: \end{equation}
1162: But clearly, $\tau_{h_{j\pm}},\tau_{h_{-}}<\tau_{h_{+}}$ since $\frac
1163: {\sqrt{\varepsilon D}}{A}\gg O\left( \frac{1}{D}\right) $ by the assumption
1164: (\ref{9mar14:03})$.$ This shows that the eigenvalue $\lambda_{+}$
1165: corresponding to $\sigma_{+}$ undergoes the Hopf bifurcation before any of the
1166: other eigenvalues, as $\tau$ decreases past $\tau_{h_{+}}$. $\blacksquare$
1167:
1168: In figure \ref{fig:tauhopf} we show the Hopf bifurcation values $\tau_{h_{+}}$
1169: and $\tau_{h_{-}}$ computed numerically as well as the asymptotic results
1170: (\ref{9mar14:14}), (\ref{10mar18:11}), for various values of $D$ while fixing
1171: $\delta=\sqrt{D\varepsilon}=0.01.$ This figure shows a very good agreement
1172: when $D\gg\delta.$
1173:
1174: \begin{figure}[ptb]
1175: \begin{center}
1176: \setlength{\unitlength}{1.75cm} \begin{picture}(5.0,3.1)(0,0)
1177: \put(2.5, 0){$\log_{10}(D \delta)$}
1178: \put(-0.05, 1.2){\rotatebox{90}{$\log_{10}(1-\tau)$}}
1179: \put(0.2,0.2){\includegraphics{tauhopf.eps}}
1180: \end{picture}
1181: \end{center}
1182: \caption{The value of $\tau_{h_{+}}$ and $\tau_{h_{-}}$ as a function of $D$,
1183: while $\delta=\sqrt{\varepsilon D}=0.01$ is held fixed. The dots represent the
1184: numerical solution obtained by substituting $\lambda=i \lambda_{i}$ into
1185: (\ref{12dec1:52}) and solving for $\tau$ and $\lambda_{i}$ using Newton's
1186: method. The solid lines are represent formulas (\ref{9mar14:14}) and
1187: (\ref{10mar18:11}). Here, $A=1$ and $B=15.$ }%
1188: \label{fig:tauhopf}%
1189: \end{figure}
1190:
1191: \section{\bigskip Asymmetric K-mesa solutions and instability with $DK^{2}\sim
1192: O\left( \frac{1}{\varepsilon}{\ln^{2}{\varepsilon}}\right) $}
1193:
1194: \label{sec:5}
1195:
1196: In Section \ref{sec:3} we have shown that $K$ mesas are always stable provided
1197: that $\tau> 1$. In our analysis there, we have ignored the effect of the
1198: exponentially decaying tail of $v$. However as $D K^{2}$ increases, this
1199: effect eventually must be taken into account. As we will see in this section,
1200: this occurs when $D\geq O\left( \frac{1}{\varepsilon\ln^{2}\varepsilon
1201: }\right) .$ The main result of this section is the following.
1202:
1203: \begin{proposition}
1204: \label{prop:expstab} Let
1205: \begin{equation}
1206: f(D)=3\sqrt{B/2}+\frac{A}{16BD}\left( \sqrt{2B}-A\right) ^{2}+3\sqrt
1207: {2B}\left( \exp\left\{ -\frac{A}{\sqrt{2\varepsilon D}}\right\}
1208: +\exp\left\{ -\frac{1}{\sqrt{2\varepsilon D}}\left( \sqrt{2B}-A\right)
1209: \right\} \right) ,
1210: \end{equation}
1211: and let $D_{1}$ be the minimum of $f(D)$. Let%
1212: \begin{equation}
1213: D_{K}=\frac{1}{K^{2}}D_{1}.
1214: \end{equation}
1215: Suppose that $\tau>1$ and $K\geq2.$ Then $K-mesa$ solution of Proposition
1216: \ref{prop:1} is stable when $D<D_{K}$ and unstable when $D>D_{K}.$ The minimum
1217: $D_{1}$ satisfies the following transcendental equation,%
1218: \begin{equation}
1219: \ \ \ \ \ \frac{A}{\sqrt{D_{1}}8B}\left( \sqrt{2B}-A\right) ^{2}%
1220: =\frac{3\sqrt{2B}}{\sqrt{2\varepsilon}}\left( A\exp\left\{ -\frac{A}%
1221: {\sqrt{D_{1}2\varepsilon}}\right\} +\left( \sqrt{2B}-A\right) \exp\left\{
1222: -\frac{1}{\sqrt{D_{1}2\varepsilon}}\left( \sqrt{2B}-A\right) \right\}
1223: \right) . \label{D1}%
1224: \end{equation}
1225:
1226:
1227: Suppose that $2A^{2}<B.$ Then%
1228: \begin{equation}
1229: D_{1}\sim\frac{A^{2}}{2\varepsilon\ln^{2}\left( \frac{12\sqrt{2}AB^{3/2}%
1230: }{\varepsilon\left( \sqrt{2B}-A\right) ^{2}}\right) }.
1231: \end{equation}
1232: Suppose that $2A^{2}>B.$ Then%
1233: \begin{equation}
1234: D_{1}\sim\frac{\left( \sqrt{2B}-A\right) ^{2}}{2\varepsilon\ln^{2}\left(
1235: \frac{12\sqrt{2}}{\varepsilon A}B^{3/2}\right) }.
1236: \end{equation}
1237:
1238: \end{proposition}
1239:
1240: Before providing a proof, consider a numerical example. Take
1241: \begin{equation}
1242: \varepsilon=0.001,\ A=2,\ B=18.
1243: \end{equation}
1244: Then solving (\ref{D1}) we obtain $D_{1}=21.16$ so that
1245: \begin{equation}
1246: D_{2}=5.3,\ \ \ D_{3}=2.35.
1247: \end{equation}
1248: To verify Proposition \ref{prop:expstab}, we ran the full numerical simulation
1249: of (\ref{6dec4:04}) for various values of $D$. We took $\tau=3,$ the initial
1250: condition to be as given in Proposition \ref{prop:1} with $K=2,$ and we took
1251: $D$ from 5 to 6 with 0.1 increments every 2500 time units. For each value of
1252: $D$, we then plotted the difference of the height of the two mesas versus
1253: time. See Figure \ref{fig:K2diff}.a. From this computation, we see that a
1254: change of stability occurs when $D\approx5.5$: if $D<5.5$ then the difference
1255: in height is decreasing but is increasing if $D>5.5$. This agrees well with
1256: the theoretical prediction $D=5.3.$ We then took $K=3$, and $D$ from 1.9 to 3
1257: with 0.1 increments every 2500 time units. Figure \ref{fig:K2diff}.b shows the
1258: the difference in heights of the first and second mesa. From this figure we
1259: conclude that the change in stability occurs when $D\sim2.45.$ Again, this
1260: agrees well with the theoretical prediction of $D_{3}=2.35$.
1261:
1262: \begin{figure}[ptbh]
1263: \begin{center}
1264: \setlength{\unitlength}{1in} \begin{picture}(6.5,2.4)(0.4,0)
1265: %\put(2.5, 0){$x$}
1266: %\put(2.5, 3.0){$v(x,t)$}
1267: %\put(0.1, 1.6){$t$}
1268: \put(0.16,0.25){\includegraphics{K2diff.eps}}
1269: \put(3.333,0.25){\includegraphics{K3diff.eps}}
1270: \put(1.5,0.1){(a)}
1271: \put(4.8,0.1){(b)}
1272: \end{picture}
1273: \end{center}
1274: \caption{ (a) The difference in height of a two-mesa solution for various
1275: values of $D$. Here, $D=5+0.1\text{floor}(t/2500)$. Note the change of
1276: stability when $D \sim5.5$. (b) The difference in height of the first two
1277: mesas of a three-mesa solution. Here, $D=1.9+0.1\text{floor}(t/2500)$. Note
1278: the change of stability when $D \sim2.45$. In both figures, $\varepsilon
1279: =0.001,\ A=2,\ B=18, \tau=3.$ }%
1280: \label{fig:K2diff}%
1281: \end{figure}
1282:
1283: Proposition \ref{prop:expstab} follows from the existence of asymmetric
1284: patterns. Indeed a similar phenomena was studied for the spike solutions of
1285: the Gierer-Meinhardt model \cite{IWW} and in the Gray-Scott model \cite{KWW}.
1286: In both of these models, an asymmetric spike pattern was found to bifurcate
1287: from a symmetric $K$ spike solution when $K>1$. Moreover a change of stability
1288: of a $K$-spike pattern occured precisely at that point.
1289:
1290: To show existence of asymmetric patterns, it suffices to compute $w(L)$ as a
1291: function of the domain length $L$, and to show an existence of a minimum of
1292: this curve. Below we show that such minimum occurs precisely when the the
1293: interaction in the $u$ component is balanced by the interaction of $v$ in the
1294: exponential tail. The main result is the following.
1295:
1296: \begin{lemma}
1297: \label{lemma:9} Consider a symmetric mesa-type solution on domain $[0,L].$
1298: Then we have,%
1299: \begin{equation}
1300: w\left( L\right) \sim3\sqrt{B/2}+\frac{1}{D}\frac{A}{16B}L^{2}\left(
1301: \sqrt{2B}-A\right) ^{2}+3\sqrt{2B}\left( \exp\left\{ -\frac{LA}%
1302: {\sqrt{2\varepsilon D}}\right\} +\exp\left\{ -\frac{L}{\sqrt{2\varepsilon
1303: D}}\left( \sqrt{2B}-A\right) \right\} \right) .\label{w(L)}%
1304: \end{equation}
1305:
1306: \end{lemma}
1307:
1308: \begin{figure}[ptbh]
1309: \begin{center}
1310: \setlength{\unitlength}{1in} \begin{picture}(5.2,3.1)(0,0)
1311: %\put(2.5, 0){$x$}
1312: %\put(2.5, 3.0){$v(x,t)$}
1313: %\put(0.1, 1.6){$t$}
1314: \put(0.1,0.0){\includegraphics[width=5in, height=3in]{wL.eps}}
1315: \end{picture}
1316: \end{center}
1317: \caption{The value of $w(L)$ as a function of $L$. Here,
1318: $A=2,\ B=18,\ \varepsilon=0.001$ and $D=10.$ The solid curve represents an
1319: exact numerical value computed using the boundary value problem solver; the
1320: dashed curve represents the asymptotic formula (\ref{w(L)}). Note that both
1321: curves give almost the same minimum value of $L \approx0.68$}%
1322: \label{fig:5}%
1323: \end{figure}
1324:
1325: \textbf{Proof}. We recall that upon expanding the solution in $\frac{1}{D}$ as
1326: $w=w_{0}+\frac{1}{D}w_{1}+\ldots$, $v=v_{0}+\frac{1}{D}v_{1}+\ldots$,
1327: the\ equation for $w_{1}$ is%
1328: \begin{equation}
1329: \delta^{2}v_{1xx}=F_{v}(v_{0,}w_{0})v_{1}+F_{w}(v_{0,}w_{0})w_{1}%
1330: ,\label{14dec8:32}%
1331: \end{equation}%
1332: \begin{equation}
1333: w_{1xx}=w_{0}-v_{0}-A.
1334: \end{equation}
1335:
1336:
1337: \bigskip
1338:
1339: \bigskip Note that we have%
1340: \begin{equation}
1341: \delta^{2}v_{0xx}^{\prime}=F_{v}(v_{0},w_{0})v_{0}^{\prime}.
1342: \end{equation}
1343: Therefore, upon multiplying (\ref{14dec8:32}) by $v_{0}^{\prime}$ and
1344: integrating we obtain,%
1345: \begin{equation}
1346: \int_{0}^{L/2}v_{0}^{\prime}F_{w}(v_{0}w_{0})w_{1}=\delta^{2}\left(
1347: v_{1x}v_{0x}-v_{1}v_{0xx}\right) _{x=0}^{x=L/2}. \label{21feb15:44}%
1348: \end{equation}
1349: To evaluate the right hand side, we write
1350: \begin{equation}
1351: \int_{0}^{L/2}v_{0}^{\prime}F_{w}(v_{0}w_{0})w_{1}\sim w_{1}\left(
1352: x_{l}\right) \int_{x_{l}^{-}}^{x_{l}^{+}}\frac{d}{dx}G\left( v_{0}%
1353: ,w_{0}\right) ,
1354: \end{equation}
1355: where
1356: \begin{gather}
1357: G(v_{0})=\int F_{w}dv_{0}=-Bv_{0}+v_{0}^{2}w_{0}-\frac{2}{3}v_{0}^{3},\\
1358: G(w_{0})=\frac{1}{9}w_{0}^{3},\ \ \ \ G(\frac{w_{0}}{3})=\frac{1}{81}w_{0}%
1359: ^{3}.
1360: \end{gather}
1361: It follows that
1362: \begin{equation}
1363: \int_{0}^{L/2}v_{0}^{\prime}F_{w}(v_{0}w_{0})w_{1}\sim-w_{1}\left(
1364: x_{l}\right) \frac{8}{81}w_{0}^{3}.
1365: \end{equation}
1366: To evaluate the right hand side of (\ref{21feb15:44}) we expand the solution
1367: near the boundary.
1368:
1369: At $x=0,$ we assumed that $v\sim w_{0};$ writing
1370: \begin{equation}
1371: v=w_{0}+V(x)
1372: \end{equation}
1373: we then obtain to leading order,%
1374: \begin{align}
1375: \delta^{2}V^{\prime\prime} & =F_{v}\left( w_{0},w_{0}\right) V+O\left(
1376: \frac{1}{D}\right) \\
1377: & \sim BV.
1378: \end{align}
1379: Imposing the boundary condition $V^{\prime}\left( 0\right) =0$ we then
1380: obtain%
1381: \begin{equation}
1382: V\sim K\left( \exp\left\{ -\frac{\sqrt{B}}{\delta}x\right\} +\exp\left\{
1383: +\frac{\sqrt{B}}{\delta}x\right\} \right) +O\left( \frac{1}{D}\right)
1384: \end{equation}
1385: for some constant $K.$ To determine $K$, we impose the matching condition
1386: $w_{0}+V\sim v_{0}+\frac{1}{D}v_{1}$ in the region $\frac{\delta}{\sqrt{B}}\ll
1387: x\ll x_{l}.$ In this region we obtain%
1388: \begin{align}
1389: v_{0} & =\frac{2}{3}w_{0}-\frac{w_{0}}{3}\tanh\frac{x-x_{l}}{2}\frac
1390: {\sqrt{B}}{\delta}\\
1391: & \sim w_{0}+\frac{2}{3}w_{0}\exp\left\{ -\frac{\sqrt{B}}{\delta}%
1392: x_{l}\right\} \exp\left\{ \frac{\sqrt{B}}{\delta}x\right\} .
1393: \end{align}
1394: Thus we obtain%
1395: \begin{equation}
1396: K=\frac{2}{3}w_{0}\exp\left\{ -\frac{\sqrt{B}}{\delta}x_{l}\right\} .
1397: \end{equation}
1398: Moreover, for $x\ll x_{l}$ we have%
1399: \begin{align}
1400: \delta^{2}v_{1xx} & \sim Bv_{1}-Bw_{1}\\
1401: & \sim Bv_{1}-Bw_{1}(0)
1402: \end{align}
1403: so that
1404: \begin{equation}
1405: v_{1}\sim w_{1}\left( 0\right) +DK\exp\left\{ -\frac{\sqrt{B}}{\delta
1406: }x\right\} .
1407: \end{equation}
1408: It follows that
1409: \begin{align}
1410: v_{1x}\left( 0\right) v_{0x}\left( 0\right) & \sim\frac{-B}{\delta^{2}%
1411: }K^{2}D,\ \ \ \ \ v_{1x}\left( 0\right) v_{0xx}\left( 0\right) \ \sim
1412: \frac{B}{\delta^{2}}\frac{K^{2}}{D}\ ,\\
1413: \delta^{2}\left( v_{1x}v_{0x}-v_{1}v_{0xx}\right) _{x=0} & =-4B^{2}%
1414: D\exp\left\{ -\frac{\sqrt{B}}{\delta}\left( L-l\right) \right\} .
1415: \end{align}
1416: Similarly, near $x=\frac{L}{2}$ we have%
1417: \begin{equation}
1418: v_{0}\sim\frac{w_{0}}{3}+\frac{2}{3}w_{0}\exp\left\{ \frac{-\sqrt{B}}{\delta
1419: }\frac{l}{2}\right\} \exp\left\{ \frac{-\sqrt{B}}{\delta}\left( x-\frac
1420: {L}{2}\right) \right\}
1421: \end{equation}
1422: from where we deduce
1423: \begin{equation}
1424: v_{1}\sim-w_{1}(0)+D\frac{2}{3}w_{0}\exp\left\{ \frac{-\sqrt{B}}{\delta}%
1425: \frac{l}{2}\right\} \exp\left( \frac{\sqrt{B}}{\delta}\left( x-\frac{L}%
1426: {2}\right) \right) ,
1427: \end{equation}
1428: %
1429:
1430: \begin{equation}
1431: \delta^{2}\left( v_{1x}v_{0x}-v_{1}v_{0xx}\right) _{x=L/2}=-\frac{4B^{2}}%
1432: {D}\exp\left\{ -\frac{\sqrt{B}}{\delta}l\right\} .
1433: \end{equation}
1434: Therefore we obtain%
1435: \begin{align}
1436: w_{1}\left( x_{l}\right) \frac{8}{81}w_{0}^{3} & \sim4B^{2}D\left(
1437: \exp\left\{ -\frac{\sqrt{B}}{\delta}l\right\} +\exp\left\{ -\frac{\sqrt{B}%
1438: }{\delta}\left( L-l\right) \right\} \right) \\
1439: w_{1}\left( x_{l}\right) & \sim3\sqrt{2B}D\left( \exp\left\{ -\frac
1440: {LA}{\sqrt{2}\delta}\right\} +\exp\left\{ -\frac{L}{\sqrt{2}\delta}\left(
1441: \sqrt{2B}-A\right) \right\} \right) .
1442: \end{align}
1443: In the region $0<x<x_{l}$, we have $v_{0}\sim w_{0}$ so from (\ref{w1}), \ we
1444: obtain
1445: \begin{equation}
1446: w_{1}^{\prime\prime}\sim-A,\text{ \ \ \ \ \ }0<x<x_{l}.
1447: \end{equation}
1448: It follows that%
1449: \begin{equation}
1450: w_{1}\left( x_{l}\right) -w_{1}\left( 0\right) \sim-\frac{A}{2}x_{l}%
1451: ^{2}=-\frac{A}{2}\left( \frac{L-l}{2}\right) ^{2}.
1452: \end{equation}
1453: Using $w_{0}=3\sqrt{B/2}$ and\ $l=LA/\sqrt{2B},$ we then obtain (\ref{w(L)}).
1454: $\blacksquare$
1455:
1456: In Figure \ref{fig:5} we compare the asymptotic formula (\ref{w(L)}) with the
1457: numerically computed value for $A=2,\ B=18,\ \varepsilon=0.001$ and $D=10.$
1458: Note that the function $L\rightarrow w(L)$ has a minimum at $L\approx0.7.$
1459: This shows the existence of a fold point. Now suppose that $D$ is chosen such
1460: that this minimum occurs precisely at $L=\frac{1}{K},$ with $K>1$ an integer.
1461: Then the corresponding $K$-mesa equilibrium solution will have a zero
1462: eigenvalue. Since the exponential terms in (\ref{w(L)}) quickly die out as $L$
1463: increases, the solution becomes stable to the right of $L=\frac{1}{K},$ and
1464: therefore unstable to the left of it. Proposition \ref{prop:expstab} is
1465: precisely this statement; it is obtained simply by scaling the $L$ out and
1466: stating the existence of the fold point in terms of $D$ instead.
1467:
1468: \section{Turing analysis}
1469:
1470: \label{sec:6}
1471:
1472: In this section we perform a Turing analysis of the homogenous steady state
1473: $u=A,\ v=\frac{B}{A}.$ In particular, we are interested in examining any
1474: possible connections between the Turing instability regime (which leads to
1475: cosinusoidal-like patterns $\cos\left( 2k\pi x\right) $ of mode $k$) and the
1476: localized mesa-like structures. We start by linearizing (\ref{6dec4:04}%
1477: ) around the steady state as follows,%
1478: \begin{equation}
1479: u=A+\xi e^{\lambda t}\cos\left( 2k\pi x\right) ,\ \ v=\frac{B}{A}+\eta
1480: e^{\lambda t}\cos\left( 2k\pi x\right) ,\ \ \ \ \ \xi,\eta\ll1;\ \ \ 2k\in%
1481: %TCIMACRO{\U{2115} }%
1482: %BeginExpansion
1483: \mathbb{N}
1484: %EndExpansion
1485: .
1486: \end{equation}
1487: This yields a 2x2 eigenvalue problem for $\lambda.$ Its solution is given by%
1488: \begin{equation}
1489: \lambda^{2}-T\lambda+\Delta=0,
1490: \end{equation}
1491: where%
1492: \begin{equation}
1493: T=B-\tau A^{2}-n\left( 1+\tau\right) -\varepsilon,\ \ \ \ \Delta=\tau\left[
1494: n\left( n-B+A^{2}\right) +\varepsilon\left( A^{2}+n\right) \right]
1495: ,\ \ \ \ \ n=4k^{2}\pi^{2}\varepsilon D.
1496: \end{equation}
1497:
1498:
1499: Note that $\Delta_{n=0}>0$ so that the zero mode is unstable if $T_{n=0}>0$ or
1500: $B-\tau A^{2}>0.$ Numerically, we observe that when the zero mode is unstable,
1501: it dominates and the system moves away from the equilibrium and quickly
1502: approaches a very long relaxation cycle, before any of the non-zero modes are
1503: activated. Therefore no spatial instability is observed. This leads to the
1504: following \emph{necessary condition }for the Turing instability to appear:%
1505: \begin{equation}
1506: B-\tau A^{2}<0. \label{11mar11:03}%
1507: \end{equation}
1508: Provided this condition is satisfied, we have $T<0$ for all $n.$ Therefore
1509: Turing instability will occur iff $\Delta<0.$ In particular the second
1510: necessary condition is that
1511: \begin{equation}
1512: B-A^{2}>0. \label{11mar11:04}%
1513: \end{equation}
1514: In this case, the most unstable mode is of the same order as the minimum of
1515: $\Delta$,
1516: \begin{align}
1517: \label{15apr17:15}k_{\ast}^{2}\sim\frac{B-A^{2}-\varepsilon}{\pi
1518: ^{2}2\varepsilon D}.
1519: \end{align}
1520: Shortly after the Turing instability is triggered, localized mesa-type
1521: structures appear due to the presence of steep gradients. In this regime,
1522: Turing instability cannot predict the final number $K$ of mesas. Indeed, we
1523: have $K\leq K_{\ast}$ where%
1524: \begin{equation}
1525: K_{\ast}=\max\left( 1,\sqrt{\frac{D_{1}}{D}}\right)
1526: \end{equation}
1527: with $D_{1}=O\left( \frac{1}{\varepsilon\ln^{2}\varepsilon}\right) $ as
1528: obtained in Theorem \ref{thm:hugeD}. It follows that as long as $B-A^{2}\gg0$
1529: and $\tau-1\gg0$, we have
1530: \begin{equation}
1531: K_{\ast} = O\left( \frac{1}{\delta\ln\frac{1}{\varepsilon}}\right) ,~~
1532: k_{\ast} = O\left( \frac{1}{\delta} \right) , ~~~\delta= \sqrt{\varepsilon
1533: D}
1534: \end{equation}
1535: so that $K_{\ast} \ll k_{\ast}.$ Therefore we expect that shortly after the
1536: patterns appear, a coarsening process takes place whereby some of the
1537: resulting mesas dissapear until there are at most $K_{\ast}$ of them left.
1538:
1539: Consider an example shown on Figure \ref{fig:turing}.a. Take $A=1,\ B=8,$
1540: $\varepsilon=10^{-4},\ D=10.$ We take $\tau=10$ to satisfy (\ref{11mar11:03}).
1541: From (\ref{D1})\ we obtain $D_{1}\sim44.5$ so that $K_{\ast}=2$ and from
1542: (\ref{15apr17:15}) we obtain $k_{\ast}=9.$ In a numerical simulation, we
1543: started from the homogenous steady state $u=A,v=\frac{B}{A}$, perturbed by a
1544: very small random noise. A Turing instability corresponding to the mode $k=7$
1545: first develops. At time $t\sim50$ only six modes remain from which six mesas
1546: develop. One by one, these mesas are annihilated until only two remain,
1547: confirming our theory. We integrated the system until $t=\text{one million}$,
1548: but we do not expect any more mesa refinement since $K_{\ast}=2$.
1549:
1550: Formulas (\ref{15apr17:15}) together with (\ref{15apr17:20}) show that it is
1551: possible for the mesa patterns to be stable even as the homogenous steady
1552: state is stable. This occurs when $A^{2}/2<B<A^{2}$ (i.e. $l>\frac{1}{2}$)
1553: with $\tau>\frac{B}{A^{2}}.$ A more difficult question is whether one can find
1554: a regime in which both are \emph{unstable}, and the system iterates between
1555: the two. Here we consider the case where $D$ satisfies (\ref{9mar14:03}). In
1556: this regime the instability of a single mesa solution can only occur when
1557: $\tau$ is near 1. Then (\ref{11mar11:03}) and (\ref{11mar11:04}) together
1558: imply that $A\sim B^{2}.$ So we set:%
1559: \begin{equation}
1560: B=A^{2}+\alpha,\ \ \ \alpha\ll1.
1561: \end{equation}
1562: From the condition $T_{n=0}\leq0$ we obtain
1563: \begin{equation}
1564: \tau\geq1+\frac{\alpha-\varepsilon}{A^{2}}%
1565: \end{equation}
1566: and we suppose that a single mode $k=1\implies n=n_{\ast}=4\pi^{2}\varepsilon
1567: D$ is unstable. The condition $\Delta_{n=n_{\ast}}<0$ then leads, to leading
1568: order:%
1569: \begin{equation}
1570: 4\pi^{2}\varepsilon D\left( \varepsilon D4\pi^{2}-\alpha\right) +\varepsilon
1571: A^{2}<0
1572: \end{equation}
1573: so that to leading order,%
1574: \begin{equation}
1575: \alpha\geq\frac{A^{2}}{D4\pi^{2}}+4\pi^{2}\varepsilon D
1576: \end{equation}
1577: from where%
1578: \begin{equation}
1579: \tau\geq1+\frac{1}{D4\pi^{2}}\sim1+0.025\frac{1}{D}.
1580: \end{equation}
1581: On the other hand, we have $B\sim A^{2}\implies l=\frac{1}{\sqrt{2}};$ and we
1582: have
1583: \begin{equation}
1584: ld-\frac{l^{3}+d^{3}}{3}=2l-2l^{2}-\frac{1}{3}=\sqrt{2}-1-\frac{1}{3}%
1585: \end{equation}
1586: so that from Theorem \ref{thm:hopf} we obtain
1587: \begin{equation}
1588: \tau_{h}\sim1+0.020\frac{1}{D}.
1589: \end{equation}
1590: Therefore the instability of a mesa cannot follow the Turing instability since
1591: $\tau>\tau_{h}.$
1592:
1593: \section{Discussion}
1594:
1595: In Section \ref{sec:5} we were able to determine the instability thresholds
1596: without actually computing the eigenvalues; but simply by showing the
1597: existence of an asymmetric pattern bifurcating from a fold point. It is an
1598: open problem to find the full expression for the eigenvalues near this
1599: threshold. This would give a theoretical timescale for each of the step in the
1600: coarsening process. We expect that the unstable eigenvalue will decrease
1601: exponentially in the distance between the mesas. This would explain the
1602: exponential time increase between the successive coarsening events, as
1603: observed in Figure \ref{fig:turing}.a
1604:
1605: In Theorem \ref{thm:hopf} under condition (\ref{9mar14:03}) we have shown that
1606: as $\tau$ is decreased near 1, the first eigenvalue to cross the imaginary
1607: axis is $\lambda_{+}$, whose eigenfunction is even. This corresponds to a
1608: ``breather''-type instability shown on Figure \ref{fig:turing}.b. An open
1609: question is whether there exists a regime for which other eigenvalues undergo
1610: a Hopf bifurcation before the $\lambda_{+}$ eigenvalue. For the Gray-Scott
1611: model it is known that a single spike can undergo a Hopf bifurcation due to a
1612: slow translational instability -- which corresponds to an odd eigenfunction --
1613: whereby the center of the spike oscillates periodically \cite{Doelman},
1614: \cite{KWW}, \cite{Muratov}. The analogy of this phenomenon for a single mesa
1615: of the Brusselator would be the Hopf bifurcation of the $\lambda_{-}$
1616: eigenvalue. One can also imagine spike-type solutions for the Brusselator
1617: simply by taking the limit $l \rightarrow0$ or equivalently, $B \rightarrow
1618: \infty.$ It is an open problem to study this regime.
1619:
1620: It would be interesting to study the slow dynamics of the mesas, of which
1621: there are several types, corresponding to different eigenvalues. The first
1622: type is the slow translational motion of the mesa such as seen in Figure
1623: \ref{fig:3} after time $t\sim2200$. Similar motion has been analysed for the
1624: FitzHugh-Nagumo model on an infinite line \cite{GMP}. In contrast to the
1625: Brusselator however, the FitzHugh-Nagumo model does not have a mass
1626: conservation constraint $lK\sim\frac{A}{\sqrt{2B}}$ derived in Proposition
1627: \ref{prop:1}, and does not undergo a coarsening process. In this sense the
1628: Brusselator resembles more the Cahn-Hilliard model or the Allen-Cahn model
1629: with mass constraint \cite{WardCH1}, \cite{WardCH2}. However unlike the
1630: Cahn-Hilliard model, the Brusselator does not have a variational formulation,
1631: and the mass conservation is only asymptotically valid. We remark that a
1632: similar phenomenon was also studied for Gray Scott and Gierer-Meinhardt models
1633: in the context of spike solutions \cite{IWW}, \cite{Doelman-speed},
1634: \cite{Muratov-speed}.
1635:
1636: A second type of slow instability is the mass exchange that occurs prior to
1637: mesa annhiliation as seen in Figure \ref{fig:3} at time $t \sim2000$. This
1638: phenomenon also occurs in some flame-propagation problems \cite{flame1},
1639: \cite{flame2} and in the Keller-Segel model \cite{KKW}, where an exchange of
1640: mass takes place between two boundary spikes, and eventually leads to an
1641: annihilation of one of them.
1642:
1643: The coarsening process in the brusselator terminates when there are
1644: $K=O\left( \frac{1}{\delta\ln\varepsilon^{-1}}\right) $ mesas left, where
1645: $\delta$ is the characteristic width of the interface. This is in contrast to
1646: the the Cahn-Hilliard model, where the coarsening proceeds until all but one
1647: interface remains \cite{WardCH1}.
1648:
1649: Localized structures far from the Turing regime are commonplace in
1650: reaction-diffusion systems such as the Brusselator, and provide an alternative
1651: pattern-formation mechanism to Turing instability. These structures appear
1652: whenever the Turing instability band is very large or when the diffusivity
1653: ratio of the activator and inhibitor is large. As we demonstrate in this work,
1654: Turing analysis cannot explain the diverse phenomena that can occur in this
1655: regime, such as coarsening and the ``breather''-type instabilities. However
1656: singular perturbation tools can be successfully applied to asnwer many of
1657: these questions.
1658:
1659: \section{Appendix A: proof of Lemma \ref{lemma:1}}
1660:
1661: \textbf{Proof}. Note that (\ref{15dec1:49})\ is equivalent to solving%
1662: \begin{align}
1663: u^{\prime\prime}-\mu_{l}^{2}u & =0,\ \ \ x\in\cup\left( x_{li}%
1664: ,x_{ri}\right) \\
1665: u^{\prime\prime}-\mu_{d}^{2}u & =0,\ \ \ x\notin\cup\left( x_{li}%
1666: ,x_{ri}\right) \\
1667: u^{\prime}\left( x_{li}^{+}\right) -u^{\prime}\left( x_{li}^{-}\right) &
1668: =-b_{li,}\ \ \ \ \ u^{\prime}\left( x_{ri}^{+}\right) -u^{\prime}\left(
1669: x_{ri}^{-}\right) =-b_{ri,}\\
1670: u^{\prime}\left( 0\right) & =0=u^{\prime}\left( 1\right) .
1671: \end{align}
1672:
1673:
1674: When $x\in\left[ x_{li,}x_{ir}\right] $ we have%
1675: \begin{equation}
1676: u=u_{li}\cosh\left( \mu_{2}\left( x-x_{li}\right) \right) +B_{li}%
1677: \sinh\left( \mu_{2}\left( x-x_{li}\right) \right) ,\ \ \ x\in\left[
1678: x_{li,}x_{ir}\right] \ \ \text{for}\ \ i=1\ldots K,
1679: \end{equation}
1680: where $u_{li}=u\left( x_{li}\right) $ and $B_{li}$ is to be found. We
1681: similarly have%
1682: \begin{equation}
1683: u=u_{ir}\cosh\left( \mu_{1}\left( x-x_{ri}\right) \right) +B_{di}%
1684: \sinh\left( \mu_{1}\left( x-x_{ri}\right) \right) ,\ \ \ x\in\left[
1685: x_{ri,}x_{l\left( i+1\right) }\right] \ \ \text{for}\ \ i=1\ldots K-1.
1686: \end{equation}
1687: We define%
1688: \begin{equation}
1689: d\equiv x_{r\left( i+1\right) }-x_{li}=\frac{1-l}{K},
1690: \end{equation}%
1691: \begin{align}
1692: c_{1} & \equiv\cosh\left( \mu_{l}l\right) ,\ \ s_{1}\equiv\cosh\left(
1693: \mu_{l}l\right) ,\\
1694: c_{2} & \equiv\cosh\left( \mu_{d}d\right) ,\ \ s_{2}\equiv\cosh\left(
1695: \mu_{d}d\right) .
1696: \end{align}
1697: We have $u_{ri}=u_{li}c_{l}+B_{li}s_{l},$ from where%
1698: \begin{equation}
1699: B_{li}=\frac{u_{ri}-u_{li}c_{l}}{s_{l}}~~\text{for}\ \ i=1\ldots K
1700: \end{equation}
1701: and similarly $u_{l\left( i+1\right) }=u_{ir}c_{d}+B_{ri}s_{d}$ so that%
1702: \begin{equation}
1703: B_{di}=\frac{u_{l\left( i+1\right) }-u_{ri}c_{d}}{s_{d}}~~\text{for}%
1704: \ \ i=1\ldots K-1.
1705: \end{equation}
1706: We also have
1707: \begin{align}
1708: b_{li} & =u^{\prime}\left( x_{li}^{-}\right) -u^{\prime}\left( x_{li}%
1709: ^{+}\right) =\mu_{d}\left( u_{r\left( i-1\right) }s_{d}+B_{d\left(
1710: i-1\right) }c_{d}\right) -\mu_{l}B_{li}\\
1711: & =\mu_{d}\left( u_{r\left( i-1\right) }s_{d}+\frac{u_{li}-u_{r\left(
1712: i-1\right) }c_{d}}{s_{d}}c_{d}\right) -\mu_{l}\frac{u_{ri}-u_{li}c_{l}%
1713: }{s_{l}}\\
1714: & =-\frac{1}{s_{d}}\mu_{d}u_{r\left( i-1\right) }+\left( \frac{c_{d}%
1715: }{s_{d}}\mu_{d}+\frac{c_{l}}{s_{l}}\mu_{l}\right) u_{li}-\frac{1}{s_{l}}%
1716: \mu_{l}u_{ri}\ \ \text{for}\ \ i=2\ldots K
1717: \end{align}
1718: and similarly%
1719: \begin{equation}
1720: b_{ri}=-\frac{1}{s_{l}}\mu_{l}u_{li}+\left( \frac{c_{d}}{s_{d}}\mu_{d}%
1721: +\frac{c_{l}}{s_{l}}\mu_{l}\right) u_{ri}-\frac{1}{s_{d}}\mu_{d}u_{r\left(
1722: i+1\right) }\ \ \text{for}\ \ i=1\ldots K-1.
1723: \end{equation}
1724: Next note that
1725: \begin{equation}
1726: u=A\cosh\left( \mu x\right) ,\text{ \ }x\in\lbrack0,x_{l1}]
1727: \end{equation}
1728: for some constant $A.$ Matching $u\left( x_{l1}^{-}\right) =u\left(
1729: x_{l1}^{+}\right) $ we then obtain
1730: \begin{align}
1731: u & =\frac{u\left( x_{l1}\right) }{c_{d/2}}\cosh\left( \mu x\right) ,\\
1732: u^{\prime}\left( x_{l1}^{-}\right) & =u\left( x_{l1}\right) \mu
1733: \frac{s_{d/2}}{c_{d/2}}.
1734: \end{align}
1735: where $s_{d/2}\equiv\sinh\left( \mu_{d}d/2\right) ,$\ $c_{d/2}\equiv
1736: \cosh\left( \mu_{d}d/2\right) .$ \ Next we use the following identity,%
1737: \begin{equation}
1738: \frac{\sinh\left( x/2\right) }{\cosh(x/2)}=\frac{\cosh\left( x\right)
1739: -1}{\sinh(x)}=\frac{\sinh(x)}{1+\cosh(x)}%
1740: \end{equation}
1741: to obtain
1742: \begin{equation}
1743: u^{\prime}\left( x_{l1}^{-}\right) =\mu\frac{c_{d}-1}{s_{d}}u\left(
1744: x_{l1}\right) .
1745: \end{equation}
1746: Therefore we obtain%
1747: \begin{align}
1748: b_{l1} & =\mu\frac{c_{d}-1}{s_{d}}u_{l1}-u^{\prime}\left( x_{li}^{+}\right)
1749: \\
1750: & =\mu\frac{c_{d}-1}{s_{d}}u_{l1}-\mu\frac{u_{r1}-u_{l1}c_{l}}{s_{l}}\\
1751: & =\mu\left( \frac{-1}{s_{d}}\mu_{d}+\frac{c_{d}}{s_{d}}\mu_{d}+\frac{c_{l}%
1752: }{s_{l}}\mu_{l}\right) u_{l1}-\frac{1}{s_{l}}\mu u_{r1}%
1753: \end{align}
1754: and similarly%
1755: \begin{equation}
1756: b_{rK}=-\frac{1}{s_{l}}\mu_{l}u_{lK}+\left( \frac{-1}{s_{l}}\mu_{l}%
1757: +\frac{c_{d}}{s_{d}}\mu_{d}+\frac{c_{l}}{s_{l}}\mu_{l}\right) u_{lK}.
1758: \end{equation}
1759: This yields the matrix $\mathbf{M}:$%
1760: \begin{equation}
1761: \mathbf{M}=\left[
1762: \begin{array}
1763: [c]{ccccccc}%
1764: a+c & b & & & & & \\
1765: b & c & a & & & & \\
1766: & a & c & b & & & \\
1767: & & & \cdots & & & \\
1768: & & & & a & c & b\\
1769: & & & & & b & c+a
1770: \end{array}
1771: \right]
1772: \end{equation}
1773: where
1774: \begin{equation}
1775: a=\frac{-\mu_{d}}{s_{d}},\ \ \ \ \ \ \ b=\frac{-\mu_{l}}{s_{l}}%
1776: ,\ \ \ \ \ \ \ c=\frac{c_{d}}{s_{d}}\mu_{d}+\frac{c_{l}}{s_{l}}\mu_{l}.
1777: \end{equation}
1778: Consider the matrix%
1779: \begin{equation}
1780: Q=\left[
1781: \begin{array}
1782: [c]{cccccccc}%
1783: a & b & & & & & & \\
1784: b & 0 & a & & & & & \\
1785: & a & 0 & b & & & & \\
1786: & & & \cdots & & & & \\
1787: & & & & b & 0 & a & \\
1788: & & & & & a & 0 & b\\
1789: & & & & & & b & a
1790: \end{array}
1791: \right] .
1792: \end{equation}
1793: The eigenvalues of this matrix were computed in \cite{RW} (see Appendix B). It
1794: was found that $Q$ has the following eigenvalues,%
1795: \begin{gather}
1796: \pm\sqrt{a^{2}+b^{2}+2ab\cos\left( \theta\right) },\ \ \theta=\frac{\pi
1797: j}{K}\ \ \text{for}\ \ j=1\ldots K-1,\\
1798: a+b,\ \ \ \ a-b.
1799: \end{gather}
1800: But w\bigskip e have $M=\left( Q+c\right) .\ $Therefore the eigenvalues of
1801: $M$ are given by
1802: \begin{gather}
1803: \left( c\pm\sqrt{a^{2}+b^{2}+2ab\cos\left( \theta\right) }\right)
1804: ,\ \ \theta=\frac{\pi j}{K}\ \ \text{for}\ \ j=1\ldots K-1,\\
1805: c+a+b,\ \ c+a-b.
1806: \end{gather}
1807:
1808: \section*{Acknowledgments}
1809: We would like to thank an anonymous referee and Y. Nishiura for suggestions that
1810: helped to improve the manuscript. This work was supported by the Fonds National de
1811: la Recherche Scientifique (Belgium) and the InterUniversity Attraction Pole program
1812: of the Belgian government. The first author is grateful for the financial support
1813: of NSERC PDF grant and Bourse de Post doctorat de l'ULB.
1814:
1815:
1816: \begin{thebibliography}{99} %
1817:
1818:
1819: \bibitem {ACF}N. D. Alikakos, X. Chen, G. Fusco, \emph{Motion of a Droplet by
1820: Surface Tension Along the Boundary}, Calc. Var. Partial Differential
1821: Equations, 11(3)(2000), 233-305.
1822:
1823: \bibitem {AFS}N. D. Alikakos, G. Fusco, C. Stephanopolous, \emph{Critical
1824: Spectrum and Stability of Interfaces for a Class of Reaction-Diffusion
1825: Equations}, J. Differential Equations, 126(1)(1996), 106-167.
1826:
1827: \bibitem {auchmuty}J.F.G.\ Auchmuty and G.\ Nicolis, \emph{Bifurcation
1828: analysis of nonlinear reaction-diffusion equations - I. Evolution equations
1829: and the steady state solutions}, Bull.\ Math.\ Biol. 37(1975), 323-365.
1830:
1831: \bibitem {bard}J.\ Bard and I.\ Lauder, \emph{How well does Turing's theory of
1832: morphogensis work?}, J.\ Theor. Biol.\ 45(1974), 501-531.
1833:
1834: \bibitem {flame1}H.\ Berestycki, S.\ Kamin and G.\ Sivashinsky,
1835: \emph{Nonlinear dynamics and metastability in a Burgers type equation}, Compte
1836: Rendus Acad. Sci., Paris 321(1995), 185-190.
1837:
1838: \bibitem {ChakravartiMarekRay}S. Chakravarti, M. Marek and W.H. Ray,
1839: \emph{Reaction-diffusion system with Brusselator kinetics: Control of a
1840: quasiperiodic route to chaos}, Phys. Rev. E, 52(3)(1995), 2407-2423.
1841:
1842: \bibitem {C}X. Chen, \emph{Spectrum for the Allen-Cahn, Cahn-Hilliard, and
1843: Phase-Field Equations for Generic Interfaces}, Comm. Partial Differential
1844: Equations, 19(7-8)(1994), 1371-1395.
1845:
1846: \bibitem {DI}A. Doelman and D. Iron, \emph{Destabilization of fronts in a
1847: class of bistable systems}, Physica D, 35(6), 1420-1450.
1848:
1849: \bibitem {Doelman-speed}A. Doelman, W. Eckhaus and T.J. Kaper, \emph{Slowly
1850: Modulated Two-Pulse Solutions in the Gray-Scott Model I: Asymptotic
1851: Construction and Stability}, SIAM J. Appl. Math., 61(3)(2000), 1080-1102.
1852:
1853: \bibitem {Doelman}A. Doelman, R. A. Gardner and T. J. Kaper, \emph{Stability
1854: Analysis of Singular Patterns in the 1D Gray-Scott Model: A Matched
1855: Asymptotics Approach}, Physica D, 122(1-4)(1998), 1-36.
1856:
1857: \bibitem {erneux2}T.\ Erneux, J.\ Hiernaux and G.\ Nicolis, \emph{Turing's
1858: theory in morphogenesis}, Bull.\ Math.\ Biol., 40(1978), 771-789.
1859:
1860: \bibitem {Erneux}T. Erneux and E. Reiss, \emph{Brusselator isolas}, SAIM J.
1861: Appl. Math. 43(1983), 1240-1246.
1862:
1863: \bibitem {Fife}Paul C. Fife, personal communications.
1864:
1865: \bibitem {GMP}R.E. Goldstein, D.J. Muraki and D.M. Petrich, \emph{Interface
1866: proliferation and the growth of labyrinths in a reaction-diffusion system},
1867: Phys. Rev. E, 53(4)(1996), 3933-3957.
1868:
1869: \bibitem {IWW}D. Iron, M. J. Ward and J. Wei, \emph{The Stability of Spike
1870: Solutions to the One-Dimensional Gierer-Meinhardt face Model}, Physica D,
1871: 150(1-2)(2001), 25-62.
1872:
1873: \bibitem {KKW}K. Kang, T. Koloklnikov and M.J. Ward, \emph{Spike stability and
1874: dynamics in the one dimensional Keller-Segel model}, submitted, IMA J. Appl.
1875: Math., 26 pages.
1876:
1877: \bibitem {KWW}T. Kolokolnikov, M. Ward and J. Wei, \emph{Slow Translational
1878: Instabilities of Spike Patterns in the One-Dimensional Gray-Scott Model},
1879: submitted, Interfaces and Free Boundaries, 39 pages.
1880:
1881: \bibitem {Muratov}C. Muratov and V. V. Osipov, \emph{Stability of the Static
1882: Spike Autosolitons in the Gray-Scott Model,} SIAM 1-205. J. Appl. Math.,
1883: 62(5)(2002), 1463-1487.
1884:
1885: \bibitem {Muratov-speed}C.B. Muratov, and V.V. Osipov, \emph{Static spike
1886: autosolitons in the Gray-Scott model}, J. Phys. A, 33(2000), 8893-8916.
1887:
1888: \bibitem {MVE}C.B. Muratov, E. Vanden-Eijnded and W. E, \emph{Self-induced
1889: stochastic resonance in excitable systems}, Physica D, 210(2005), 227-240.
1890:
1891: \bibitem {NP}G.\ Nicolis and I.\ Prigogine, \emph{Self-Organization in
1892: Nonequilibrium Systems}, Wiley, New York, 1977.
1893:
1894: \bibitem {N0}Y.Nishiura, \emph{Coexistence of infinitely many stable solutions
1895: to reaction-diffusion system in the singular limit}, in Dynamics Reported:
1896: Expositions in Dynamical Systems, Edited by C.R.K.T. Jones, U. Kirchgraber and
1897: H.O. Walther, volume 3. Springer-Verlag, New York, 1994.
1898:
1899: \bibitem {N1}Y. Nishiura and H. Fujii, \emph{Stability of Singularly Perturbed
1900: Solutions to Systems of Reaction-Diffusion Equations}, SIAM J. Math. Anal.,
1901: 18(1987), 1726-1770.
1902:
1903: \bibitem {N2}Y. Nishiura, and H. Fujii, \emph{SLEP Method to the Stability of
1904: Singularly Perturbed Solutions with Multiple Internal Transition Layers in
1905: Reaction-Diffusion Systems}, in Dynamics of Infinite-Dimensional Systems
1906: (Lisbon, 1986), pp. 211-230, NATO Adv. Sci. Inst. Ser. F Comput. Systems Sci,
1907: 37(1987), Springer, Berlin.
1908:
1909: \bibitem {N3}Y. Nishiura, M. Mimura, H. Ikeda and H. Fujii, \emph{Singular
1910: Limit Analysis of Stability of Traveling Wave Solutions in Bistable
1911: Reaction-Diffusion Systems}, SIAM J. Math. Anal., 21(1990), 85-122.
1912:
1913: \bibitem {N4}Y. Nishiura and M. Taniguchi, \emph{Stability and characteristic
1914: wavelength of planar interfaces in the large diffusion limit of the
1915: inhibitor}, Proc. Roy. Soc. Edinburgh Sect. A, 126(1)(1996), 117-145.
1916:
1917: \bibitem {OP}V.V. Osipov and E.V. Ponizovskaya, \emph{Stochastic resonance in
1918: the Brusselator model}, Phys. Rev. E, 61(4)(2000), 4603-4605.
1919:
1920: \bibitem {PL}I.\ Prigogine and R.\ Lefever, \emph{Symmetry-breaking
1921: instabilities in dissipative systems}, J.\ Chem. Phys. 48(1968), 1965-1700.
1922:
1923: \bibitem {RW}X.\ Ren and J.\ Wei, \emph{On the spectra of 3-D lamellar
1924: solutions of the Diblock Copolymer problem,} SIAM J. Math. Anal. 35(1)(2003), 1-32.
1925:
1926: \bibitem {Pena}B. Pe\~na and C. P\'erez-Garc\'ia, \emph{Stability of Turing
1927: patterns in the Brusselator model}, Phys. Rev. E., 64(5), 2001.
1928:
1929: \bibitem {Turing}A.M.\ Turing, \emph{The chemical basis of morphogenesis},
1930: Phil.\ Trans.\ R.\ Soc., London B237(1952) 37-72.
1931:
1932: \bibitem {Tyson}J. Tyson, \emph{Some further studies of non-linear
1933: oscillations in chemical systems}, J. Chem. Phys., 58(1972), 3919-3930.
1934:
1935: \bibitem {WardCH1}L. Reyna, and M.J. Ward, \emph{Metastable Internal Layer
1936: Dynamics for the Viscous Cahn-Hilliard Equation}, Methods and Applications of
1937: Analysis, 2(1995), 285-306.
1938:
1939: \bibitem {flame2}X.\ Sun and M.J.\ Ward, \emph{ Metastability for a
1940: Generalized Burgers Equation with Applications to Propagating Flame--Fronts},
1941: European J. Appl. Math., 10(1999), 27-53.
1942:
1943: \bibitem {WardCH2}M.J. Ward, \emph{Metastable Bubble Solutions for the
1944: Allen-Cahn Equation with Mass Conservation} SIAM J. Appl. Math., 56(5)(1996), 1247-1279.
1945:
1946: \bibitem {deWit}A.\ De Wit, \emph{Spatial Patterns and Spatiotemporal Dynamics
1947: in Chemical Systems}, Advances in Chemical Physics, 1999, I. Prigogine and
1948: S.A. Rice, editors.
1949:
1950: \bibitem {deWit3D}A. De Wit, P.\ Borckmans and G.\ Dewel \emph{Twist grain
1951: boundaries in three-dimensional lamellar Turing structures}, Proc. Natl. Acad.
1952: Sci. USA, 94(1997), 12765-12768.
1953:
1954: \bibitem {deWitChaos}A.\ De Wit, D.\ Lima, G.\ Dewel and P.\ Borckmans,
1955: \emph{Spatiotemporal dynamics near a codimension-two point}, Phys. Rev. E.,
1956: 54(1)(1996), 261--271.
1957:
1958: \bibitem {YZE}L. Yang, A.M. Zhabotinsky and I.R. Epstein \emph{Stable Squares
1959: and Other Oscillatory Turing Patterns in a Reaction-Diffusion Model}, Phys.
1960: Rev. Lett., 92(19)(2004), 198303-1 - 198303-4.
1961:
1962: \bibitem {YuGumel}P. Yu and A.B. Gumel, \emph{Bifurcation and stability
1963: analyses for a coupled Brusselator model}, Journal of Sound and Vibration,
1964: 244(5)(2001), 795-820.
1965: \end{thebibliography}
1966:
1967:
1968: \end{document}
1969: