0806.1286/ch.tex
1: \documentclass{amsart} %{amsart}%{article}%{elsart}
2: \usepackage{graphicx}
3: \usepackage{amssymb, amsmath}
4: \usepackage{amsfonts}
5: \usepackage{amssymb}
6: \usepackage{sidecap}
7: 
8: %\usepackage{psfig}
9: \usepackage{float}
10: 
11: \floatplacement{figure}{H}
12: 
13: \newtheorem{theorem}{Theorem}
14: \newtheorem{acknowledgement}[theorem]{Acknowledgment}
15: %\newtheorem{algorithm}[theorem]{Algorithm}
16: \newtheorem{assumption}{Assumption}
17: \newtheorem{axiom}[theorem]{Axiom}
18: \newtheorem{case}{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}{Example}
27: \newtheorem{exercise}[theorem]{Exercise}
28: \newtheorem{lemma}[theorem]{Lemma}
29: \newtheorem{notation}[theorem]{Notation}
30: \newtheorem{problem}[theorem]{Problem}
31: \newtheorem{proposition}[theorem]{Proposition}
32: \newtheorem{remark}[theorem]{Remark}
33: \newtheorem{solution}[theorem]{Solution}
34: \newtheorem{summary}[theorem]{Summary}
35: %\newenvironment{proof}[1][Proof]{\textbf{#1.} }{\ \rule{0.5em}{0.5em}}
36: %\newtheorem{figure}{Figure}
37: \newcommand{\calm}{\mathcal M}
38: \newcommand{\circd}{\mathop{D}\limits^\circ}
39: \newcommand{\circm}{\mathop{M}\limits^\circ}
40: \newcommand{\circq}{\mathop{Q}\limits^\circ}
41: \newcommand{\divv}{\text{\rm div}}
42: \newcommand{\ind}{\text{\rm ind}}
43: \newcommand{\cA}{\mathcal A}
44: \newcommand{\cB}{\mathcal B}
45: \newcommand{\cC}{\mathcal C}
46: \newcommand{\cH}{\mathcal H}
47: \newcommand{\cL}{\mathcal L}
48: \newcommand{\C}{\mathbb C}
49: \newcommand{\N}{\mathbb N}
50: \newcommand{\Z}{\mathbb Z}
51: \newcommand{\bZ}{{\bf Z}}
52: \newcommand{\Q}{\mathbb Q}
53: \newcommand{\bQ}{{\bf Q}}
54: \newcommand{\R}{\mathbb R}
55: \newcommand{\im}{\text{\rm im}}
56: \newcommand{\id}{\text{\rm id}}
57: \newcommand{\ext}{\text{\rm Ext}}
58: \newcommand{\pdim}{\text{$P$-$\dim$}}
59: \newcommand{\socle}{\text{\rm socle}}
60: \newcommand{\sgn}{\text{\rm sign}}
61: \newcommand{\dist}{\text{\rm dist}}
62: \def\T{\mathbb T}
63: \def\la{\label}
64: \hyphenation{Lem-ma}
65: %\newcommand{\tg}{t_g}
66: %%% can't define \ker it is already defined
67: %%  \qed already defined
68: 
69: 
70: \def\bt{\begin{thm}}
71: \def\et{\end{thm}}
72: 
73: \def\bl{\begin{lem}}
74: \def\el{\end{lem}}
75: 
76: \def\bd{\begin{defi}}
77: \def\ed{\end{defi}}
78: 
79: \def\bc{\begin{cor}}
80: \def\ec{\end{cor}}
81: 
82: \def\bp{\begin{proof}}
83: \def\ep{\end{proof}}
84: 
85: \def\br{\begin{rem}}
86: \def\er{\end{rem}}
87: 
88: \newtheorem{thm}{Theorem}[section]
89: %\newtheorem{lem}[thm]{Lemma}
90: \newtheorem{lem}{Lemma}[section]
91: \newtheorem{defi}{Definition}[section]
92: %\newtheorem{example}[Example]{chapter}
93: \newtheorem{ex}{Example}[section]
94: %\newtheorem{example}{Example}[chapter]
95: \newtheorem{prop}[thm]{Proposition}
96: \newtheorem{xca}[thm]{Exercise}
97: \newtheorem{rem}{Remark}[section]
98: \newtheorem{cor}{Corollary}[section]
99: %\newtheorem{assumption}[thm]{Assumption}
100: \newtheorem{method}[thm]{Method}
101: 
102: \renewcommand{\theequation}{\thesection.\arabic{equation}}
103: \numberwithin{equation}{section}
104: \numberwithin{theorem}{section}
105: %\numberwithin{lemma}{section}
106: \numberwithin{example}{section}
107: %\numberwithin{definition}{section}
108: %\numberwithin{remark}{section}
109: 
110: \numberwithin{figure}{section}
111: 
112: \begin{document}
113: 
114: \title{Cahn-Hilliard Equations  and Phase Transition Dynamics  for Binary Systems}
115: 
116: \author[Ma]{Tian Ma}
117: \address[TM]{Department of Mathematics, Sichuan University,
118: Chengdu, P. R. China}
119: 
120: \author[Wang]{Shouhong Wang}
121: \address[SW]{Department of Mathematics,
122: Indiana University, Bloomington, IN 47405}
123: \email{showang@indiana.edu}
124: 
125: \thanks{The work was supported in part by the
126: Office of Naval Research and by the National Science Foundation.}
127: 
128: \keywords{}
129: \subjclass{}
130: 
131: \begin{abstract} The process of phase separation of binary systems is described by the Cahn-Hilliard equation.  The main objective of this article is to give a classification on the dynamic phase transitions for binary systems using either the classical Cahn-Hilliard equation or the Cahn-Hilliard equation coupled with entropy, leading to some interesting physical predictions. 
132: The analysis is based on dynamic transition theory for nonlinear systems and new classification scheme for dynamic transitions, developed recently by the authors. 
133: \end{abstract}
134: \maketitle
135: \tableofcontents
136: 
137: \section{Introduction}
138: \label{sc1}
139: Cahn-Hilliard equation  describes the process of phase separation, by which the two components of a binary fluid spontaneously separate and form domains pure in each component. The main objective of this article is to provide a theoretical approach to dynamic phase transitions for binary systems.
140: 
141: Classically, phase transitions are classified by  the Ehrenfest classification scheme,  based on  the lowest derivative of the free energy that is discontinuous at the transition. 
142: In general, it is a difficult task to classify phase transitions of higher order, which appears in many equilibrium phase transition systems, such as the PVT system, the ferromagnetic system, superfluids as well as the binary systems studied in this article. 
143: 
144: For this purpose, a new  dynamic transition theory is developed recently by the authors. This new theory provides an efficient tool to analyze phase transitions of higher order. With this theory in our disposal, a new  dynamic  classification scheme is obtained, and classifies  phase transitions into three categories: Type-I, Type-II and Type-III, corresponding mathematically  to continuous,  jump, and  mixed transitions, respectively; see the Appendix as well as two recent books by the authors \cite{b-book, chinese-book} for details. 
145: 
146: There have been extensive studies in the past on the dynamics of 
147: the Cahn-Hilliard equations. However, very little is known about the higher order transitions encountered for this problem, and this article gives a complete classification of the dynamics transitions for binary systems. The results obtained lead in particular to various physical predictions. First,  the order of  phase transitions is precisely determined by the sign of  a nondimensional 
148: parameter $K$ such that if $K>0$, the transition is first-order with latent heat and if $K <0$, the transition is second-order. Second, a theoretical transition diagram is derived, leading in particular  to  a prediction that there is only second-order transition for molar fraction near $1/2$. This is different from the prediction made by the classical transition diagram. Third, a critical length scale  is derived such that no phase separation occurs at any temperature if the length of the container is smaller than the critical length scale. These physical predictions will be addressed in another article.
149: 
150: 
151: This article is organized as follows. In Section 2, both the classical Cahn-Hilliard equation and the Cahn-Hilliard equation coupled with entropy are introduced in a unified fashion using a general principle for equilibrium phase transitions outlined in Appendix B. Sections 3-6 analyze dynamic transitions for 
152: the Cahn-Hilliard equation in general domain, rectangular domain, with periodic boundary conditions, and for the Cahn-Hilliard equation coupled with entropy. 
153:  Physical conclusions are given in Section 7, and the dynamic transition theory is recalled in  Appendix A.
154: 
155: 
156: \section{Dynamic Phase Transition  Models for Binary Systems}
157: Materials compounded by two components $A$ and $B$, such as  binary
158: alloys, binary solutions and polymers, are called binary
159: systems. Sufficient cooling of a binary system may lead to phase
160: separations, i.e., at the critical temperature, the concentrations
161: of both components $A$ and $B$ with homogeneous distribution undergo  changes,
162: leading to heterogeneous distributions in space. Phase separation
163: of binary systems observed will be in one of two main ways. The first is by nucleation in which sufficiently large nuclei of the
164: second phase appear randomly and grow, and this corresponds to Type-II phase transitions. The second is  by spinodal
165: decomposition in which the systems appear to nuclear at once, and
166: periodic or semi-periodic structure is seen, and this corresponds to Type-I phase transitions. 
167: 
168: Since binary systems are conserved, the equations describing the
169: Helmholtz process and the Gibbs process are the same. Hence,
170: without distinction we use the term "free energy" to discuss this
171: problem. 
172: 
173: Let $u_A$   and $u_B$ be the concentrations of components $A$  and  $B$ respectively,  then $u_B=1-u_A$.  In a homogeneous state, $u_B=\bar{u}_B$ is a constant, and the
174: entropy density $S_0=\bar{S}_0$ is also a constant. We take $u, S$
175: the concentration and entropy density deviations:
176: $$u=u_B-\bar{u}_B,\ \ \ \ S=S_0-\bar{S}_0$$
177: By (\ref{7.28}) and (\ref{7.29}), the free energy is given
178: by
179: \begin{equation}
180: F(u,S)=F_0+\int_{\Omega} \Big[\frac{\mu_1}{2}|\nabla u|^2+\frac{\mu_2}{2}|\nabla
181: S|^2+\frac{\beta_1}{2}|S|^2+\beta_0Su\label{8.49}
182:  +\beta_2Su^2+\sum^{2p}_{k=1}\alpha_ku^k\Big]dx.
183: \end{equation}
184: Since  entropy is increasing as $u\rightarrow 0$, and by
185: $$
186: \frac{\delta}{\delta S}F(u,S)=-\mu_2\Delta
187: S+\beta_1S+\beta_0u+\beta_2u^2=0,
188: $$ 
189: we have 
190: \begin{equation}
191: \beta_0=0,\ \ \ \ \beta_1>0,\ \ \ \ \beta_2>0,\label{8.50}
192: \end{equation}
193: which implies that $S$ is a decreasing function of $|u|$.
194: 
195: According (\ref{7.38}) and (\ref{7.32}),  we derive from (\ref{8.49}) and
196: (\ref{8.50})  the  following equations governing a binary system:
197: \begin{equation}
198: \left.
199: \begin{aligned} &\frac{\partial S}{\partial t}=k_1\Delta
200: S-a_1S-a_2u^2,\\
201: &\frac{\partial u}{\partial t}=-k_2\Delta^2u+b_0\Delta (Su)+\Delta
202: f(u),\\
203: &\int_{\Omega}u(x,t)dx=0,
204: \end{aligned}
205: \right.\label{8.51}
206: \end{equation}
207: where $k_1,k_2,b_0,a_1,a_2$ are positive constants, and
208: \begin{equation}
209: f(u)=\sum^{2p-1}_{k=1}b_ku^k,\ \ \ \ b_{2p-1}>0 \ \ \ \ (p\geq
210: 2).\label{8.52}
211: \end{equation}
212: 
213: Physically sound boundary condition for (\ref{8.51}) is either the Neumann boundary condition:
214: \begin{equation}
215: \frac{\partial u}{\partial n}=\frac{\partial\Delta u}{\partial
216: n}=0,\ \ \ \ \frac{\partial S}{\partial n}=0\ \ \ \ \text{on}\
217: \partial\Omega ,\label{8.53}
218: \end{equation}
219: with  $\Omega\subset \R^n$  $(1\leq n\leq 3)$ being  a bounded domain, or 
220: the periodic boundary condition:
221: \begin{equation}
222: u(x+KL)=u(x),\ \ \ \ S(x+KL)=S(x)\label{8.54}
223: \end{equation}
224: with  $\Omega =[0,L]^n,K=(k_1,\cdots,k_n),1\leq n\leq 3.$
225: 
226: \medskip
227: 
228: For simplicity, in this section we always assume that $p=2$. Thus,
229: function (\ref{8.52}) is rewritten as
230: \begin{equation}
231: f(u)=b_1u+b_2u^2+b_3u^3,\ \ \ \ b_3>0.\label{8.55}
232: \end{equation}
233: Based on Theorem \ref{t5.1}, we have to assume that there exists a
234: temperature $T_1>0$ such that $b_1=b_1(T)$ satisfies
235: \begin{equation}
236: b_1(T)\left\{
237: \begin{aligned}
238: &  >0&&  \quad \text{  if } T>T_1,\\
239: & =0&&  \quad \text{ if } T=T_1,\\
240: & <0&&  \quad \text{ if } T<T_1.
241: \end{aligned}
242: \right.\label{8.56}
243: \end{equation}
244: 
245: %\subsection{Cahn-Hilliard equation}
246: %\la{s8.2.2}
247: 
248: If we ignore the coupled action of entropy in (\ref{8.49}), then
249: the free energy $F$ is in the following form
250: $$F(u)=F_0+\int_{\Omega}\left[\frac{\mu_1}{2}|\nabla
251: u|^2+\frac{\alpha_1}{2}u^2+\frac{\alpha_2}{3}u^3+\frac{\alpha_3}{4}u^4\right]dx,$$
252: and the equations (\ref{8.51}) are the following classical  Cahn-Hilliard
253: equation:
254: \begin{equation}
255: \left.
256: \begin{aligned} 
257: &\frac{\partial u}{\partial
258: t}=-k\Delta^2u+\Delta f(u),\\
259: &\int_{\Omega}u(x,t)dx=0,
260: \end{aligned}
261: \right.\label{8.57}
262: \end{equation}
263: where $f(u)$ is as in (\ref{8.55}).
264: 
265: \section{Phase Transition in General Domains}
266: \la{s8.2.3}
267: In this section, we shall discuss 
268: the Cahn-Hilliard equation from the mathematical point of view. We start with  the nondimensional form of equation. Let
269: \begin{align*}
270: &x=lx^{\prime}, &&  t=\frac{l^4}{k}t^{\prime}, && u=u_0u^{\prime},\\
271: &\lambda =-\frac{l^2b_1}{k},&& \gamma_2=\frac{l^2b_2u_0}{k},
272: && \gamma_3=\frac{l^2b_3u^2_0}{k},
273: \end{align*}
274: where $l$ is a given length,  $u_0=\bar{u}_B$ is the constant
275: concentration of $B$, and $\gamma_3>0$. Then the equation (\ref{8.57}) can be
276: rewritten as follows (omitting the primes)
277: \begin{equation}
278: \left.
279: \begin{aligned} 
280: &\frac{\partial u}{\partial
281: t}=-\Delta^2u-\lambda\Delta u+\Delta (\gamma_2u^2+\gamma_3u^3),\\
282: &\int_{\Omega}u(x,t)dx=0,\\
283: &u(x,0)=\varphi .
284: \end{aligned}
285: \right.\label{8.58}
286: \end{equation}
287: 
288: Let
289: $$
290:  H=\left\{u\in L^2(\Omega )\ |\ \int_{\Omega}udx=0\right\}.$$
291: For the Neumann boundary condition (\ref{8.53}) we define
292: $$
293: H_1=\left\{u\in H^4(\Omega )\cap H\ \Big|\ \frac{\partial u}{\partial
294: n}=\frac{\partial\Delta u}{\partial
295: n}=0\ \text{ on } \partial \Omega \right\},$$ 
296: and for the periodic boundary
297: condition (\ref{8.54}) we define
298: $$H_1=\left\{u\in H^4(\Omega )\cap H\ \Big|\ u(x+KL)=u(x) \ \ \forall
299: K\in \Z^n\right\}.$$
300: 
301: Then we define the operators $L_{\lambda}=-A+B_{\lambda}$ and
302: $G:H_1\rightarrow H$ by
303: \begin{equation}
304: \left.
305: \begin{aligned} 
306: & Au=\Delta^2u,\\
307: & B_{\lambda}u=-\lambda\Delta u,\\
308: & G(u)=\gamma_2\Delta u^2+\gamma_3\Delta u^3.
309: \end{aligned}
310: \right.\label{8.59}
311: \end{equation}
312: 
313: Thus, the Cahn-Hilliard equation (\ref{8.58}) is equivalent to the
314: following operator equation
315: \begin{equation}
316: \left.
317: \begin{aligned} &\frac{du}{dt}=L_{\lambda}u+G(u), \\
318: &u(0)=\varphi
319: \end{aligned}
320: \right.\label{8.60}
321: \end{equation}
322: It is known that the operators defined by (\ref{8.59}) satisfy the
323: conditions (\ref{5.2}) and (\ref{5.3}).
324: 
325: 
326: We first consider the case where $\Omega\subset \R^n(1\leq n\leq
327: 3)$ is a general bounded and smooth domain. Let $\rho_k$ and $e_k$ be the
328: eigenvalues and eigenfunctions of the following eigenvalue problem:
329: \begin{equation}
330: \left.
331: \begin{aligned} 
332:  &-\Delta e_k=\rho_ke_k,\\
333: &\frac{\partial e_k}{\partial n}|_{\partial\Omega}=0,\\
334: &\int_{\Omega}e_kdx=0.
335: \end{aligned}
336: \right.\label{8.61}
337: \end{equation}
338: The eigenvalues of (\ref{8.61}) satisfy
339: $ 0<\rho_1\leq\rho_2\leq\cdots\leq\rho_k\leq\cdots,$  and $ \lim_{k \to \infty} \rho_k= \infty$. The eigenfunctions $\{e_k\}$ of (\ref{8.61})
340: constitute an orthonormal  basis of $H$. Furthermore,  the
341: eigenfunctions of (\ref{8.61}) satisfy
342: $$\frac{\partial\Delta e_k}{\partial n}|_{\partial\Omega}=0,\ \ \
343: \ k=1,2,\cdots .$$ 
344: Hence, $\{e_k\}$ is also an orthogonal basis of
345: $H_1$ under the following equivalent norm
346: $$\|u\|_1=\left[\int_{\Omega}|\Delta^2u|^2dx\right]^{{1}/{2}}.$$
347: 
348: We are now in position to give a phase transition theorem 
349: for the problem (\ref{8.58}) with the following
350: Neumann boundary condition:
351: \begin{equation}
352: \frac{\partial u}{\partial n} =0,\ \ \ \
353: \frac{\partial\Delta u}{\partial
354: n}=0 \qquad \text{ on } \partial \Omega.\label{8.62}
355: \end{equation}
356: 
357: \bt\la{t8.3}
358: Assume that $\gamma_2=0$ and $\gamma_3>0$ in
359: (\ref{8.58}), then the following assertions hold true:
360: 
361: \begin{enumerate}
362: 
363: \item[(1)] If the first eigenvalue $\rho_1$ of (\ref{8.61}) has
364: multiplicity $m\geq 1$, then the problem (\ref{8.58}) with
365: (\ref{8.62}) bifurcates from $(u,\lambda )=(0,\rho_1)$ on $\lambda
366: >\rho_1$ to an attractor $\Sigma_{\lambda}$, homeomorphic to an
367: $(m-1)$-dimensional sphere $S^{m-1}$, and $\Sigma_{\lambda}$
368: attracts $H \setminus \Gamma$,  where $\Gamma$ is the stable manifold of $u=0$
369: with codimension $m$. 
370: 
371: \item[(2)] $\Sigma_{\lambda}$ contains at
372: least $2m$ singular points. If $m=1, \Sigma_{\lambda}$ has exactly
373: two steady states $\pm u_{\lambda}$, and if $m=2,
374: \Sigma_{\lambda}=S^1$ has at most eight singular points.
375: 
376: 
377: \item[(3)]  Each singular point $u_{\lambda}$ in
378: $\Sigma_{\lambda}$ can be expressed as
379: $$u_{\lambda}=\left(\lambda
380: -\rho_1\right)^{{1}/{2}} w +o(|\lambda
381: -\rho_1|^{{1}/{2}}),$$ 
382: where $w$ is an eigenfunction corresponding to the first eigenvalue of  (\ref{8.61}).
383: \end{enumerate}
384: \et
385: 
386: \bp
387: We proceed in several steps as follows.
388: 
389: \medskip
390: 
391: {\sc Step 1.} It is clear that the eigenfunction $\{e_k\}$ of
392: (\ref{8.61}) are also eigenvectors of the linear operator
393: $L_{\lambda}=-A+B_{\lambda}$ defined by (\ref{8.59}) and the
394: eigenvalues of $L_{\lambda}$ are given by
395: \begin{equation}
396: \beta_k(\lambda )=\rho_k(\lambda -\rho_k),\ \ \ \ k=1,2,\cdots
397: .\label{8.63}
398: \end{equation}
399: It is easy to verify the conditions (\ref{5.4}) and (\ref{5.5})
400: in our case  at $\lambda_0=\rho_1$. We shall prove this theorem
401: using the attractor bifurcation theory introduced in \cite{b-book}.
402: 
403: We need to verify that $u=0$ is a global asymptotically stable
404: singular point of (\ref{8.60}) at $\lambda =\rho_1$. By
405: $\gamma_2=0$, from the energy integration of (\ref{8.58}) we can
406: obtain
407: \begin{equation}
408: \frac{1}{2}\frac{d}{dt}\int_{\Omega}u^2dx=\int_{\Omega}\left[-|\Delta
409: u|^2+\rho_1|\nabla u|^2-3\gamma_3u^2|\nabla
410: u|^2\right]dx\label{8.64}
411: \end{equation}
412: $$\leq -C\int_{\Omega}|\Delta
413: v|^2dx-3\gamma_3\int_{\Omega}u^2|\nabla u|^2dx,$$ where $C>0$ is a
414: constant, $u=v+ w $, and $\int_{\Omega}v wdx=0$, and $w$ is a first engenfunction. It follows from
415: (\ref{8.64}) that $u=0$ is global asymptotically stable. 
416: Hence, %forbookTheorems \ref{t2.24} and \ref{t5.4}, 
417: for Assertion (1), we only have to prove that  
418: $\Sigma_{\lambda}$ is homeomorphic to $S^{m-1}$, as the rest of  this assertion follows directly from the attractor bifurcation theory introduced in \cite{b-book}.
419: 
420: \bigskip
421: 
422: {\sc Step 2.} Now we prove that the bifurcated attractor $\Sigma_{\lambda}$
423: from $(0,\rho_1)$ contains at least $2m$ singular points.
424: 
425: Let $g(u)=-\Delta u-\lambda u+\gamma_3u^3$. Then the stationary
426: equation of (\ref{8.58})  is given by 
427: \begin{align*}
428: &\Delta g(u)=0,\\
429: &\int_{\Omega}udx=0,
430: \end{align*}
431: which is equivalent, by the maximum principle,  to
432: \begin{equation}
433: \left.
434: \begin{aligned} 
435: &-\Delta u-\lambda
436: u+\gamma_3u^3=\text{constant},\\
437: &\int_{\Omega}udx=0,\\
438: &\frac{\partial u}{\partial n}|_{\partial\Omega}=0.
439: \end{aligned}
440: \right.\label{8.65}
441: \end{equation}
442: By the Lagrange multiplier theorem, (\ref{8.65}) is the Euler
443: equation of the following functional with zero average constraint:
444: \begin{align}
445: & F(u)=\int_{\Omega}\left[\frac{1}{2}|\nabla
446: u|^2-\frac{\lambda}{2}u^2+\frac{\gamma_3}{4}u^4\right]dx,\label{8.66}
447: \\
448: & u\in \left\{u\in H^1(\Omega )\cap H\ \Big|\ \
449: \frac{\partial u}{\partial n}|_{\partial\Omega}=0\quad \int_{\Omega}udx=0\right\}. \nonumber 
450: \end{align}
451:  Since $F$ is an even functional, %forbookby Theorem \ref{t3.8}, 
452:  by the classical Krasnoselskii bifurcation theorem for even functionals, 
453:  (\ref{8.66}) bifurcates from
454: $\lambda >\rho_1$ at least to $2m$ mini-maximum points, i.e., 
455: equation (\ref{8.65}) has at least $2m$ bifurcated solutions on
456: $\lambda >\rho_1$. Hence, the attractor $\Sigma_{\lambda}$
457: contains at least $2m$ singular points.
458: 
459: \bigskip
460: 
461: {\sc Step 3.} 
462: To complete the proof, we reduce the equation (\ref{8.60}) to the
463: center manifold near $\lambda =\rho_1$. 
464: By the approximation formula given in \cite{b-book},  
465: the reduced equation of (\ref{8.60}) is given by:
466: \begin{equation}
467: \frac{dx_i}{dt}=\beta_1(\lambda
468: )x_i-\gamma_3\rho_1\int_{\Omega}v^3e_{1i}dx  + o(|x|^3)\qquad \text{ for }  1\leq i\leq
469: m,\label{8.67}
470: \end{equation}
471: where $\beta_1(\lambda )=\rho_1(\lambda -\rho_1)$, 
472: $v=\sum^m_{i=1}x_ie_{1i}$, 
473: and $\{e_{11},\cdots,e_{1m}\}$ are the first eigenfunctions
474: of (\ref{8.61}). Equations (\ref{8.67}) can be rewritten as
475: \begin{equation}
476: \frac{dx_i}{dt}=\beta_1(\lambda
477: )x_i-\gamma_3\rho_1\int_{\Omega}\left(\sum^m_{j=1}x_je_{1j}\right)^3e_{1i}dx+o(|x|^3).\label{8.68}
478: \end{equation}
479: Let
480: $$ g(x)=\left(\int_{\Omega}v^3e_{11}dx,\cdots,\int_{\Omega}v^3e_{1m}dx\right),
481: \qquad  v(x)=\sum^m_{j=1}x_je_{1j}.
482: $$
483:  Then for any $x\in \R^m$,
484: \begin{equation}\label{8.69}
485: <g(x),x>=\sum^m_{i=1}x_i\int_{\Omega}v^3e_{1i}dx
486: =\int_{\Omega}v^4dx \geq C|x|^4,
487: \end{equation}
488: for some constant $C>0$.
489: 
490: Thus by the attractor bifurcation theorem \cite{b-book}, 
491: it follows from (\ref{8.68}) and (\ref{8.69}) 
492:  that the attractor $\Sigma_{\lambda}$ is homeomorphic to
493: $S^{m-1}$. Hence, Assertion (1) is proved. The other conclusions
494: in Assertions (2) and (3) can be derived from (\ref{8.68}) and
495: (\ref{8.69}). The proof is complete.
496: \qed
497: \ep
498: 
499: 
500: Physically, the coefficients $\gamma_2$ and $\gamma_3$ depend on
501: $u_0=\bar{u}_B$, the temperature $T$, and the pressure $p$:
502: $$
503: \gamma_k=\gamma_k(u_0,T,p), \qquad k=2,3.
504: $$
505: The set of points satisfying
506: $\gamma_2(u_0,T,p)=0$ has measure zero in $(u_0,T,p)\in \R^3$.
507: Hence, it is more interesting to consider the case where
508: $\gamma_2\neq 0$.
509: 
510: For this purpose, let the multiplicity of the first eigenvalue $\rho_1$ of
511: (\ref{8.61}) be $m\geq 1$, and $\{e_1,\cdots,e_m\}$ be the first
512: eigenfunctions. We introduce the following quadratic equations
513: \begin{equation}
514: \left.
515: \begin{aligned} 
516: & \sum^m_{i,j=1}a^k_{ij}x_ix_j=0\quad &&\text{ for } 1\leq
517: k\leq m,\\
518: & a^k_{ij}=\int_{\Omega}e_ie_je_kdx.
519: \end{aligned}
520: \right.\label{8.70}
521: \end{equation}
522: 
523: \bt\la{t8.4}
524: Let $\gamma_2\neq 0$,  $\gamma_3>0$, and
525: $x=0$ be an isolated singular point of (\ref{8.70}). Then the
526: phase transition of (\ref{8.58}) and (\ref{8.62}) is either
527: Type-II or Type-III. Furthermore, the problem (\ref{8.58}) with
528: (\ref{8.62}) bifurcates to at least one singular point on each side
529: of $\lambda =\rho_1$, and has a saddle-node bifurcation on
530: $\lambda <\rho_1$. In particular,  if $m=1$, then the following
531: assertions hold true:
532: 
533: \begin{enumerate}
534: 
535: \item[(1)] The phase transition is Type-III, and a neighborhood
536: $U\subset H$ of $u=0$ can be decomposed into two sectorial regions
537: $\bar{U}=\bar{D}_1(\pi )+\bar{D}_2(\pi )$ such that the phase
538: transition in $D_1(\pi )$ is the first order, and in $D_2(\pi )$
539: is the $n$-th order with $n\geq 3$. 
540: 
541: \item[(2)] The bifurcated
542: singular point $u_{\lambda}$ on $\lambda >\rho_1$ attracts
543: $D_2(\pi )$, which can be expressed as
544: \begin{equation} 
545:  u_{\lambda}=(\lambda
546: -\rho_1)e_1/\gamma_2a+o(|\lambda -\rho_1|),\label{8.71}
547: \end{equation}
548: where, by assumption,   
549: $a=\int_{\Omega}e^3_1dx  \neq 0$.
550: 
551: \item[(3)] When $|\gamma_2a|=\varepsilon$ is small, the
552: assertions in the transition perturbation theorems (Theorems \ref{t6.10}
553: and \ref{t6.11}) hold true.
554: \end{enumerate}
555: \et
556: 
557: \br\la{r8.4}
558: {\rm
559: We shall see later that when $\Omega$ is a
560: rectangular domain,  i.e., 
561: $$\Omega =\Pi^n_{k=1}(0,L_k)\subset \R^n
562: \qquad (1\leq n\leq 3),$$ 
563: then $a=\int_{\Omega} e_1^3 dx =0$. However, for almost all non-rectangular
564: domains $\Omega$,  the first eigenvalues are
565: simple and $a\neq 0$. Hence, the Type-III phase transitions for
566: general domains are generic.\qed
567: }
568: \er
569: 
570: \bp[Proof of Theorem \ref{t8.4}]
571: Assertions (1)-(3) can be directly
572: proved using Theorems \ref{t5.9}, \ref{t6.10}, and \ref{t6.11}. 
573: By assumption, $u=0$ is a
574: second order non-degenerate singular point of (\ref{8.60}) at
575: $\lambda =\rho_1$, which implies that $u=0$ is not locally
576: asymptotically stable. Hence, it follows from Theorems \ref{t5.3}  and
577: the steady state bifurcation theorem for  even-order
578: nondegenerate singular points \cite{b-book}  that the phase transition of (\ref{8.58}) with (\ref{8.62})
579: is either Type-II or Type-III, and there is at least one singular
580: point bifurcated on each side of $\lambda =\rho_1$.
581: 
582: Finally, we shall apply Theorem \ref{t6.4} to prove that there exists a
583: saddle-node bifurcation on $\lambda <\rho_1$.
584: 
585: It is known that
586: $$\text{ind}(L_{\lambda}+G,0)
587: = \left\{
588: \begin{aligned}
589: & \text{even}&& \text{ if }  \lambda =\rho_1,\\
590: & 1&& \text{ if }  \lambda <\rho_1.
591: \end{aligned}
592: \right.$$ 
593: Moreover, since $L_{\lambda}+G$ defined by (\ref{8.59})
594: is a gradient-type operator,  we can
595: derive  that
596: $$\text{ind}(L_{\rho_1}+G,0)\leq 0.$$
597: Hence there is a bifurcated branch $\Sigma_{\lambda}$ on $\lambda
598: <\rho_1$ such that
599: $$
600: \text{ind}(L_{\lambda}+G,u_{\lambda})=-1
601: \qquad \forall u_{\lambda}\in\Sigma_{\lambda},\ \ \ \ \lambda <\rho_1.
602: $$ 
603: It is
604: clear that the eigenvalues (\ref{8.63}) of $L_{\lambda}$ satisfy
605: (\ref{5.4}) and (\ref{5.5}). For any
606: $\lambda\in \R^1$ (\ref{8.60}) possesses a global attractor.
607: Therefore, for bounded $\lambda ,b<\lambda\leq\rho_1$, the
608: bifurcated branch $\Sigma_{\lambda}$ is bounded:
609: $$\|u_{\lambda}\|_H\leq C\qquad \forall
610: u_{\lambda}\in\Sigma_{\lambda},\ \ \ \ -\infty
611: <b<\lambda\leq\rho_1.$$ 
612: We need to prove that there exists
613: $\widetilde{\lambda}<\rho_1$ such that for all $\lambda
614: <\widetilde{\lambda}$ equation (\ref{8.60}) has no nonzero
615: singular points.
616: 
617: By the energy estimates of (\ref{8.60}),   
618: for any $\lambda <\widetilde{\lambda}=-{\gamma^2_2}/{2\gamma_3}$  
619: and  $u\neq 0$ in $ H$, 
620: \begin{align*}
621: &\int_{\Omega}\left[|\Delta u|^2-\lambda |\nabla
622: u|^2+2\gamma_2u|\nabla u|^2+3\gamma_3u^2|\nabla u|^2\right]dx\\
623: &\geq \int_{\Omega}|\Delta u|^2dx+\int_{\Omega}|\nabla
624: u|^2(-\lambda -2|\gamma_2u|+3\gamma_3u^2)dx\\
625: &\geq \int_{\Omega}|\nabla u|^2dx+\int_{\Omega}|\nabla
626: u|^2(-\lambda
627: +\gamma_3u^2+2\gamma_3(u-\frac{\gamma_2}{\gamma_3})^2-\frac{\gamma^2_2}{2\gamma_3})dx\\
628: &\geq \int_{\Omega}|\nabla u|^2dx+\int_{\Omega}|\nabla
629: u|^2(-\lambda -\frac{\gamma^2_2}{2\gamma_3})dx\\
630: &> 0.
631: \end{align*}
632: Therefore, when $\lambda <\widetilde{\lambda}$,
633: (\ref{8.60}) has no nontrivial singular points in $H$. Thus we
634: infer from Theorem \ref{t6.4} that there exists a saddle-node bifurcation
635: on $\lambda <\rho_1$. This proof is complete. 
636: \ep
637: 
638: \section{Phase Transition in Rectangular Domains}
639: \la{s8.2.4}
640: The dynamical properties of phase separation of a binary system in
641: a rectangular container is very different from that in a general
642: container. We see in the previous section that the phase
643: transitions in general domains are Type-III,  and  we shall show
644: in the following that the phase transitions in rectangular domains
645: are either Type-I or Type-II, which are distinguished by a critical
646: size of the domains.
647: 
648: Let $\Omega =\Pi^n_{k=1}(0,L_k)\subset \R^n$  $ (1\leq n\leq 3)$ be a
649: rectangular domain. 
650: We first consider the case where 
651: \begin{equation}
652: L=L_1>L_j\qquad  \forall 2\leq j\leq n.\label{8.72}
653: \end{equation}
654: 
655: 
656: \bt\la{t8.5}
657: Let $\Omega =\Pi^n_{k=1}(0,L_k)$ satisfy
658: (\ref{8.72}). The following assertions hold true:
659: 
660: \begin{enumerate}
661: 
662: \item[(1)]  If $$\gamma_3<\frac{2L^2}{9\pi^2}\gamma^2_2,$$ then  the
663: phase transition of (\ref{8.58}) and (\ref{8.62}) at $\lambda
664: =\lambda_0=\pi^2/L^2$ is Type-II. In particular, the problem
665: (\ref{8.58}) with (\ref{8.62}) bifurcates from $(u,\lambda
666: )=(0,\pi^2/L^2)$ on $\lambda <\pi^2/L^2$ to exactly two equilibrium
667: points which are saddles, and there are two saddle-node
668: bifurcations on $\lambda < {\pi^2}/{L^2}$ as shown in Figure
669: \ref{f8.8-1}. 
670: 
671: \item[(2)] If $$\gamma_3>\frac{2L^2}{9\pi^2}\gamma^2_2,$$  
672: then the transition is Type-I. In particular, 
673: the problem bifurcates on $\lambda >\pi^2/L^2$ to exactly two
674: attractors $u^T_1$ and $u^T_2$ which can be expressed as
675: \begin{equation}
676:  u^T_{1,2}=\pm\frac{\sqrt{2}(\lambda
677: -\frac{\pi^2}{L^2})^{{1}/{2}}}{\sigma^{{1}/{2}}}\cos\frac{\pi
678: x_1}{L}+o(|\lambda -\frac{\pi^2}{L^2}|^{{1}/{2}}),\label{8.73}
679: \end{equation}
680: where 
681: $\sigma =\frac{3\gamma_3}{2}-\frac{L^2\gamma^2_2}{3\pi^2}.
682: $
683: \end{enumerate}
684: \et
685: \begin{SCfigure}[25][t]
686:   \centering
687:   \includegraphics[width=0.5\textwidth]{8-8.pdf}
688:   \caption{Type-II transition as given by Theorem~\ref{t8.5}.}\la{f8.8-1}
689:  \end{SCfigure}
690: 
691: \bp
692: With the spatial domain as given,   the first eigenvalue and eigenfunction of
693: (\ref{8.61}) are given by
694: $$\rho_1=\pi^2/L^2,\ \ \ \ e_1=\cos\frac{\pi x_1}{L}.$$
695: The eigenvalues and eigenfunctions of $L_{\lambda}=-A+B_{\lambda}$
696: defined by (\ref{8.59}) are as follows:
697: \begin{align}
698: & \beta_K=|K|^2(\lambda -|K|^2),\label{8.74}\\
699: & e_K=\cos\frac{k_1\pi x_1}{L_1}\cdots\cos\frac{k_n\pi x_n}{L_n},\label{8.75}
700: \end{align}
701: where
702: \begin{align*}
703: &K=\left(\frac{k_1\pi}{L_1},\cdots,\frac{k_n\pi}{L_n}\right)&&  \forall 
704:   k_i\in \Z,\ \ \ \ 1 \leq i\leq n, \\
705: &|K|^2=\pi^2\sum^n_{i=1}k^2_i/L^2_i&& |K|^2\neq 0.
706: \end{align*}
707: 
708: By the approximation of the center manifold obtained in \cite{b-book},  the  reduced equation of
709: (\ref{8.60}) to the center manifold is given by
710: \begin{equation}
711: \frac{dy}{dt}= \beta_1(\lambda )y 
712:  -\frac{2\pi^2}{(L_1\cdots
713: L_n)L^2_1}
714: \int_{\Omega}\left[\gamma_2y^2e^3_1 +\gamma_3y^3 e^4_1
715: +\gamma_2 (ye_1+\Phi (y))^2e_1\right]dx,
716: \label{8.76}
717: \end{equation}
718: where $y\in \R^1$, $\Phi (y)$ is the center manifold function, and
719: \begin{equation}
720: \beta_1(\lambda )=\frac{\pi^2}{L^2}(\lambda -\frac{\pi^2}{L^2}).\label{8.77}
721: \end{equation}
722: Direct calculation implies that 
723: \begin{eqnarray}
724: &&\int_{\Omega}e^3_1dx=\int^{L_1}_0\cdots\int^{L_n}_0\cos^3\frac{\pi
725: x_1}{L}dx=0,\label{8.78}\\
726: &&\int_{\Omega}e^4_1dx=L_2\cdots L_n\int^L_0\cos^4\frac{\pi
727: x_1}{L}dx_1=\frac{3}{8}L_1\cdots L_n.\label{8.79}
728: \end{eqnarray}
729: By (\ref{8.78}) and $\Phi (y)=O(|y|^2)$ we obtain
730: \begin{equation}
731: \int_{\Omega}(ye_1+\Phi )^2e_1dx=2y\int_{\Omega}\Phi
732: (y)e^2_1dx+o(|y|^3)\label{8.80}
733: \end{equation}
734: It follows that
735: \begin{align*}
736: & 
737: \Phi (y) = \sum^{\infty}_{|K|^2>\pi^2/L^2}\phi_K(y)e_K+o(|y|^2)\\
738: & \phi_K(y)= \frac{\gamma_2y^2}{-\beta_K\|e_K\|^2_H}\int_{\Omega}\Delta
739: e^2_1\cdot e_Kdx = 
740: \frac{|K|^2\gamma_2y^2}{\beta_K\|e_K\|^2_H}\int_{\Omega}e_Ke^2_1dx.
741: \end{align*}
742: Notice that
743: $$\int_{\Omega}e_Ke^2_1dx=
744: \left\{
745: \begin{aligned}
746: &0&& \forall  K\neq\left(\frac{2\pi}{L_1},0,\cdots,0\right),\\
747: &L_1\cdots L_n/4&& \forall 
748: K=\left(\frac{2\pi}{L_1},0,\cdots,0\right).
749: \end{aligned}
750: \right.
751: $$ 
752: Then we have
753: $$
754: \Phi (y)=\frac{\gamma_2y^2}{2(\lambda
755: -4\pi^2/L^2)}\cos\frac{2\pi x_1}{L}+o(|y|^2).
756: $$ 
757: Inserting $\Phi
758: (y)$ into (\ref{8.80}) we find
759: \begin{align}
760: \int_{\Omega}(ye_1+\Phi )^2e_1dx
761: =&\frac{\gamma_2y^3}{\lambda -\frac{4\pi^2}{L^2}}\int_{\Omega}\cos\frac{2\pi
762: x_1}{L}e^2_1dx+o(|y|^3)\label{8.81}\\
763: =&\frac{L_1\cdots L_n}{4}\cdot\frac{\gamma_2}{\lambda
764: -\frac{4\pi^2}{L^2}}y^3+o(|y|^3).\nonumber
765: \end{align}
766: Finally, by (\ref{8.78}), (\ref{8.79}) and (\ref{8.81}),   we derive from (\ref{8.76})
767:  the following reduced equation of (\ref{8.60}): 
768:  \begin{equation}
769: \frac{dy}{dt}=\beta_1(\lambda
770: )y-\frac{\pi^2}{2L^2}\left(\frac{3\gamma_3}{2}
771: +\frac{\gamma^2_2}{\lambda -\frac{4\pi^2}{L^2}}\right)y^3+o(|y|^3).\label{8.82}
772: \end{equation}
773: Near the critical point $\lambda_0=\pi^2/L^2$, the coefficient
774: $$\frac{3\gamma_3}{2}+\frac{\gamma^2_2}{\lambda
775: -\frac{4\pi^2}{L^2}}=\frac{3\gamma_3}{2}-\frac{L^2\gamma^2_2}{3\pi^2}.$$
776: 
777: Thus, by Theorem \ref{t5.2} we derive from (\ref{8.82}) the assertions of
778: the theorem except the claim for the saddle-node bifurcation in
779: Assertion (1), which can be proved in the same fashion as used
780: in Theorem \ref{t8.4}. The proof is complete. 
781: \ep
782: 
783: 
784: We now consider the case where $\Omega =\Pi^n_{k=1}(0,L_k)$
785: satisfies  that 
786: \begin{equation}
787:  L=L_1=\cdots =L_m > L_j  \qquad \text{for } 2\leq m\leq 3, m < j \le n.
788:  \label{8.83}
789: \end{equation}
790: 
791: 
792: \bt\la{t8.6}
793:  Let $\Omega =\Pi^n_{k=1}(0,L_k)$ satisfy
794: (\ref{8.83}). Then the following assertions hold true:
795: 
796: \begin{enumerate}
797: \item  If 
798: $$\gamma_3>\frac{26L^2}{27\pi^2}\gamma^2_2,$$ 
799: then the phase transition of the problem (\ref{8.58}) with (\ref{8.62}) at $\lambda_0=\pi^2/L^2$ is Type-I, satisfying the following properties:
800: 
801: \begin{enumerate}
802: 
803: \item  The problem bifurcates on $\lambda >\pi^2/L^2$ to an attractor $\Sigma_{\lambda}$, containing exactly $3^m-1$ non-degenerate singular points, 
804: and $\Sigma_{\lambda}$ is homeomorphic to an $(m-1)$-dimensional 
805: sphere $S^{m-1}$.
806: 
807: \item For $m=2$, the attractor $\Sigma_{\lambda}=S^1$
808: contains 4 minimal attractors, as shown in Figure \ref{f8.9-1}. 
809: 
810: \item For
811: $m=3, \Sigma_{\lambda}=S^2$ contains  $8$ minimal attractors as shown in Figure~\ref{f8.10}(a),  if 
812: $$\gamma_3 < \frac{22 L^2}{9\pi^2} \gamma_2^2,$$
813: and contains $6$ minimal attractors as shown  in Figure \ref{f8.10}(b)
814: if 
815: $$\gamma_3 > \frac{22 L^2}{9\pi^2} \gamma_2^2.$$
816: \end{enumerate}
817: 
818: 
819: \item If 
820: $$\gamma_3<\frac{26L^2}{27\pi^2}\gamma^2_2,$$
821: then the transition is Type-II. 
822: In particular, the problem has
823: a saddle-node bifurcation on $\lambda <\lambda_0=\pi^2/L^2$, and
824: bifurcates on both side of $\lambda =\lambda_0$ to exactly $3^m-1$
825: singular points which are non-degenerate.
826: \end{enumerate}
827: \et
828: \begin{SCfigure}[25][t]
829:   \centering
830:   \includegraphics[width=0.4\textwidth]{8-9-1.pdf}
831:   \caption{For  $m=2, \Sigma_{\lambda}=S^1$ and
832: $Z_{2k}$  $(1\leq k\leq 4)$ are attractors.}\la{f8.9-1}
833:  \end{SCfigure}
834: \begin{figure}[hbt]
835:   \centering
836:   \includegraphics[width=0.38\textwidth]{8-10a.pdf}\qquad 
837:   \includegraphics[width=0.3\textwidth]{8-10b.pdf}
838:   \caption{For $m=3$, $\Sigma_{\lambda}=S^2$, (a) $\pm
839: Z_k$  $(1\leq k\leq 4)$ are attractors, (b) $\pm Y_k$  $(1\leq k\leq 3)$
840: are attractors.}\la{f8.10}
841:  \end{figure}
842: 
843: 
844: \bp 
845:  We proceed in several steps as follows.
846: 
847: \medskip
848: 
849: {\sc Step 1.} Consider the center manifold reduction. It is known
850: that the eigenvalues and eigenfunctions of
851: $L_{\lambda}=-A+B_{\lambda}$ are given by (\ref{8.74}) and
852: (\ref{8.75}) with $L_1=\cdots =L_m$.  %forbook(\ref{1.123}), 
853: As before, the reduced
854: equations of (\ref{8.60}) are  given by 
855: \begin{equation}
856: \frac{dy}{dt}=\beta_1(\lambda)y+g(y)+o(|y|^3),\label{8.84}
857: \end{equation}
858: where $y=(y_1,\cdots,y_m)\in \R^m$, $\beta_1(\lambda )$ is as in
859: (\ref{8.77}), and
860: \begin{align*}
861: &g(y)=\frac{2\pi^2}{L_1\cdots L_nL^2_1}(G_2(y)+G_3(y)+G_{23}(y)),\\
862: & G_2(y)=-\gamma_2\left(\int_{\Omega}v^2e_1dx,\cdots,\int_{\Omega}v^2e_mdx\right),\\
863: & G_3(y)=-\gamma_3\left(\int_{\Omega}v^3e_1dx,\cdots,\int_{\Omega}v^3e_mdx\right),\\
864: &G_{23}(y)=-\gamma_2\left(\int_{\Omega}u^2e_1dx,\cdots,\int_{\Omega}u^2e_mdx\right).
865: \end{align*}
866: Here 
867: $ e_i=\cos\pi x_i/L $  for   $ 1\leq i\leq m$,  $L$  is given by  (\ref{8.83}), 
868: $ v=\sum^m_{i=1}y_ie_i,$   $u=v+\Phi (y)$   and $\Phi$ is the center manifold function.
869: Direct computation shows that  
870: \begin{align}
871: & 
872: \left.\begin{aligned}
873:  \int_{\Omega}v^2e_idx
874:       =&  \int_{\Omega}\left(\sum^m_{j=1}y_j\cos\pi
875:             x_j/L\right)^2\cos\pi x_i/Ldx=0, \\
876:  \int_{\Omega}v^3e_idx=& 
877:      \int_{\Omega}\left(\sum^n_{j=1}y_k\cos\pi x_j/L\right)^3\cos\pi x_i/Ldx\\
878:  =& \frac{3}{4}L_1\cdots
879: L_n\left(\frac{1}{2}y^3_i+y_i\sum_{j\neq i}y^2_j\right),
880: \end{aligned}
881: \right.\label{8.85} \\
882: &
883: \int_{\Omega}u^2e_idx=\int_{\Omega}\left(\sum^m_{j=1}y_je_j+\Phi
884: (y)\right)^2e_idx \nonumber \\
885: & \qquad \qquad = 2\sum^m_{j=1}y_j\int_{\Omega}\Phi (y)e_je_idx
886: + \int_\Omega \Phi^2(y) e_idx. \label{8.86}
887: \end{align}
888: 
889: 
890: We need to compute the center manifold function $\Phi (y)$. 
891: As in \cite{b-book}, we have
892: \begin{equation}
893: \Phi
894: (y)=\sum^{\infty}_{|K|>\pi^2/L^2}\phi_K(y)e_K+o(|y|^2)+O(|y|^2|\beta_1|),\label{8.87}
895: \end{equation}
896: where
897: \begin{align*}
898: \phi_K(y) =&\frac{\gamma_2}{-\beta_K(\lambda)<e_K,e_K>}
899: \int_{\Omega}\Delta v^2e_Kdx\\
900: =&\frac{|K|^2\gamma_2}{\beta_K(\lambda)<e_K,e_K>}
901: \int_{\Omega}v^2e_Kdx\\
902: =&\frac{|K|^2\gamma_2}{\beta_K<e_K,e_K>}\sum^m_{i,j=1}y_iy_i\int_{\Omega}e_ie_ie_Kdx.
903: \end{align*}
904: It is clear that
905: $$\int_{\Omega}e_ie_je_Kdx=\int_{\Omega}\cos\frac{\pi
906: x_i}{L}\cos\frac{\pi x_j}{L}e_kdx=0,$$ 
907: if 
908: $$ K\neq K_i+K_j, \qquad  K_i=\left(\frac{\pi}{L_1}\delta_{1i},\cdots,\frac{\pi}{L_n}\delta_{ni}\right).
909: $$
910: By (\ref{8.74}) and (\ref{8.75}) we have
911: \begin{align*}
912: \phi_K(y) =
913: &
914: \frac{\gamma_2y_iy_j(2-\delta_{ij})}{(\lambda -|K|^2)<e_K,e_K>}\int_{\Omega}e_ie_je_Kdx\\
915: =
916: &\left\{
917: \begin{aligned} 
918: &  \frac{\gamma_2y^2_i}{2(\lambda  -\frac{4\pi^2}{L^2})}&& \text{ if }  i=j,\\
919: & \frac{2\gamma_2y_iy_j}{(\lambda -\frac{2\pi^2}{L^2})}&& \text{ if }   i\neq j,
920: \end{aligned}
921: \right.
922: \end{align*}
923: for $K=K_i+K_j$. 
924: Thus, by (\ref{8.87}), we obtain 
925: $$
926: \Phi (y)=\sum^m_{i=1}\frac{\gamma_2}{2(\lambda
927: -\frac{4\pi^2}{L^2})}y^2_i\cos\frac{2\pi x_i}{L}
928: +\sum^m_{l>r}\frac{2\gamma_2}{(\lambda
929: -\frac{2\pi^2}{L^2})}y_ly_r\cos\frac{\pi x_l}{L}\cos\frac{\pi x_r}{L}.
930: $$
931: Inserting $\Phi (y)$ into (\ref{8.86}) we derive
932: \begin{align*}
933: \int_{\Omega}u^2e_idx= &
934:  \frac{\gamma_2}{\lambda-\frac{4\pi^2}{L^2}} 
935: \int_{\Omega} \left[y^3_i  \cos^2\frac{\pi
936: x_i}{L}\cos\frac{2\pi x_i}{L}+ 4y_i\sum_{j\neq
937: i}y^2_j \cos^2\frac{\pi x_j}{L}\cos^2\frac{\pi
938: x_i}{L}\right]dx \\
939: & + o(|y|^3).
940: \end{align*}
941:  Direct computation gives that
942: \begin{equation}
943: \int_{\Omega}u^2e_idx=\frac{\gamma_2L_1\cdots
944: L_n}{4}\left[\frac{y^3_i}{\lambda -\frac{4\pi^2}{L^2}}+\frac{4y_i}{\lambda
945: -\frac{2\pi^2}{L^2}}\sum_{j\neq i}y^2_j\right] + o(|y|^3).\label{8.88}
946: \end{equation}
947: Putting (\ref{8.85}) and (\ref{8.88}) in (\ref{8.84}), we get the
948: reduced equations in the following form
949: \begin{equation}
950: \frac{dy_i}{dt}=\beta_1(\lambda
951: )y_i-\frac{\pi^2}{2L^2}\left[\sigma_1y^3_i+\sigma_2y_i\sum_{j\neq
952: i}y^2_j\right]+o(|y|^3) \quad \forall 1\leq i\leq m,\label{8.89}
953: \end{equation}
954: where
955: \begin{equation}
956: \sigma_1=\frac{3\gamma_3}{2}+\frac{\gamma^2_2}{\lambda
957: -\frac{4\pi^2}{L^2}},\qquad 
958: \sigma_2=3\gamma_3+\frac{4\gamma^2_2}{\lambda -\frac{2\pi^2}{L^2}}.
959: \label{8.90}
960: \end{equation}
961: 
962: {\sc Step 2.}
963:  It is known that the transition type of (\ref{8.60}) at
964: the critical point $\lambda_0=\pi^2/L^2$ is completely determined
965: by (\ref{8.89}), i.e., by the following equations
966: \begin{equation}
967: \frac{dy_i}{dt}=-\frac{\pi^2}{2L^2}\left[\sigma^0_1y^3_i+\sigma^0_2y_i\sum_{k\neq
968: i}y^2_k\right]\qquad \forall  1\leq i\leq m,\label{8.91}
969: \end{equation}
970: where
971: $$\sigma^0_1=\frac{3\gamma_3}{2}-\frac{L^2\gamma^2_2}{3\pi^2},\qquad \sigma^0_2=3\gamma_3-\frac{4L^2\gamma^2_2}{\pi^2}.
972: $$ 
973: It is easy to see that
974: \begin{equation}
975: \left.
976: \begin{aligned}
977: &\sigma^0_1+\sigma^0_2>0\Leftrightarrow\gamma_3>\frac{26L^2}{27\pi^2}\gamma^2_2,\\
978: &\sigma^0_1+\sigma^0_2<0\Leftrightarrow\gamma_3<\frac{26L^2}{27\pi^2}\gamma^2_2.
979: \end{aligned}
980: \right.\label{8.92}
981: \end{equation}
982: 
983: {\sc Step 3.} We consider the case where $m=2$. Thus, the transition
984: type of (\ref{8.91}) is equivalent to that of the following equations
985: \begin{equation}
986: \left.
987: \begin{aligned}
988: &\frac{dy_1}{dt}=-y_1[\sigma^0_1y^2_1+\sigma^0_2y^2_2],\\
989: &\frac{dy_2}{dt}=-y_2[\sigma^0_1y^2_2+\sigma^0_2y^2_1].
990: \end{aligned}
991: \right.\label{8.93}
992: \end{equation}
993: We can see that on the straight lines
994: \begin{equation}
995: y^2_1=y^2_2, \label{8.94}
996: \end{equation}
997: equations (\ref{8.93}) satisfy that
998: $$\frac{dy_2}{dy_1}=\frac{y_2}{y_1}\qquad \text{ for }\
999: \sigma^0_1+\sigma^0_2\neq 0,\ \ \ \ (y_1,y_2)\neq 0.$$
1000: Hence the straight lines (\ref{8.94}) are orbits of
1001: (\ref{8.93})   if  $\sigma^0_1+\sigma^0_2\neq 0$. Obviously, the
1002: straight lines
1003: \begin{equation}
1004: y_1=0 \text{  and }  y_2=0 \label{8.95}
1005: \end{equation}
1006: are also orbits of (\ref{8.93}).
1007: 
1008: There are four straight lines determined by (\ref{8.94}) and
1009: (\ref{8.95}), and each of them contains two orbits. Hence, the
1010: system (\ref{8.93}) has at least eight straight line orbits.  
1011: Hence it is not hard to see that the number of straight line orbits of
1012: (\ref{8.93}), if finite, is eight.
1013: 
1014: Since (\ref{8.60}) is a gradient-type equation, 
1015: there are no elliptic regions at $y=0$; see \cite{b-book}. Hence, when
1016: $\sigma^0_1+\sigma^0_2>0$ all the straight line orbits  on
1017: (\ref{8.94}) and (\ref{8.95}) tend to $y=0$, as shown in Figure
1018: \ref{f8.11} (a), which implies that the regions are parabolic and
1019: stable, therefore $y=0$ is asymptotically stable for (\ref{8.93}).
1020: Accordingly, by the attractor bifurcation theorem in \cite{b-book}, the transition of (\ref{8.89}) at
1021: $\lambda_0=\pi^2/L^2$ is Type-I.
1022: 
1023: When $\sigma^0_1+\sigma^0_2<0$ and $\sigma^0_1>0$, namely
1024: $$\frac{2}{9}\frac{L^2\gamma^2_2}{\pi^2}<\gamma_3<\frac{26}{27}\frac{L^2\gamma^2_2}{\pi^2},$$
1025: the four straight line orbits on (\ref{8.94}) are outward from
1026: $y=0$, and other four on (\ref{8.95}) are toward $y=0$, as shown
1027: in Figure \ref{f8.11} (b), which implies that all regions at $y=0$ are
1028: hyperbolic. Hence, by Theorem \ref{t5.3},  the transition of (\ref{8.89})
1029: at $\lambda_0=\pi^2/L^2$ is Type-II.
1030: 
1031: When $\sigma^0_1\leq 0$, then $\sigma^0_2<0$ too. In this case, no
1032: orbits of (\ref{8.93}) are toward $y=0$, as shown in Figure \ref{f8.11}
1033: (c), which implies by Theorem \ref{t5.3} that the transition is Type-II.
1034: 
1035: Thus by (\ref{8.92}), for $m=2$ we prove that the transition is
1036: Type-I if $\gamma_3>\frac{26L^2}{27\pi^2}\gamma^2_2$, and Type-II
1037: if  $\gamma_3<\frac{26 L^2}{27\pi^2}\gamma^2_2$.
1038: \begin{figure}[hbt]
1039:   \centering
1040:   \includegraphics[width=0.3\textwidth]{8-11a.pdf} 
1041:   \includegraphics[width=0.3\textwidth]{8-11b.pdf} 
1042:   \includegraphics[width=0.28\textwidth]{8-11c.pdf}
1043:   \caption{The topological structure of flows of
1044: (\ref{8.89}) at $\lambda_0=\pi^2/L^2$, (a) as
1045: $\gamma_3>\frac{26L^2}{27\pi^2}\gamma^2_2$; (b) as
1046: $\frac{2L^2}{9\pi^2}\gamma^2_2<\gamma_3<\frac{26L^2}{27\pi^2}\gamma^2_2$;
1047: and (c) $\gamma_3\leq\frac{2L^2}{9\pi^2}\gamma^2_2$.}
1048:   \la{f8.11}
1049:  \end{figure}
1050: 
1051: 
1052: {\sc Step 4.} Consider the case where $m=3$. Thus, (\ref{8.91}) are
1053: written as
1054: \begin{equation}
1055: \left.
1056: \begin{aligned}
1057: &\frac{dy_1}{dt}=-y_1[\sigma^0_1y^2_1+\sigma^0_2(y^2_2+y^2_3)],\\
1058: &\frac{dy_2}{dt}=-y_2[\sigma^0_1y^2_2+\sigma^0_2(y^2_1+y^2_3)],\\
1059: &\frac{dy_3}{dt}=-y_3[\sigma^0_1y^2_3+\sigma^0_2(y^2_1+y^2_2)].
1060: \end{aligned}
1061: \right.\label{8.96}
1062: \end{equation}
1063: It is clear that the straight lines
1064: \begin{align}
1065: & y_i=0, \quad  y_j=0\qquad  \text{ for } i\neq j,\ 1\leq i,\ j\leq
1066: 3,\label{8.97} \\
1067: &
1068: \left\{
1069: \begin{aligned} 
1070: & y^2_i=y^2_j, \quad y_k=0&& \text{ for } i\neq j,i\neq
1071: k,j\neq k,1\leq i,j,k\leq 3,\\
1072: & y^2_1=y^2_2=y^2_3,
1073: \end{aligned}
1074: \right.\label{8.98}
1075: \end{align}
1076: consist of orbits of (\ref{8.96}). There are total 13 straight
1077: lines in (\ref{8.97}) and (\ref{8.98}), each of which consists of
1078: two orbits. Thus, (\ref{8.96}) has at least 26 straight line
1079: orbits. We shall show that (\ref{8.96}) has just the straight
1080: line orbits given by (\ref{8.97}) and (\ref{8.98}). In fact, we
1081: assume that the line
1082: $$y_2=z_1y_1,\ \ \ \ y_3=z_2y_1\ \ \ \ (z_1,z_2\ \text{are\ real\
1083: numbers})$$ is a straight line orbit of (\ref{8.96}). Then
1084: $z_1,z_2$ satisfy
1085: \begin{equation}
1086: \left.
1087: \begin{aligned}
1088: &\frac{dy_2}{dy_1}=z_1=z_1\frac{\sigma^0_1z^2_1+\sigma^0_2(1+z^2_2)}{\sigma^0_1+\sigma^0_2(z^2_1+
1089: z^2_2)},\\
1090: &\frac{dy_3}{dy_1}=z_2=z_2\frac{\sigma^0_1z^2_2+\sigma^0_2(1+z^2_1)}{\sigma^0_1+\sigma^0_2(z^2_1+z^2_2)}.
1091: \end{aligned}
1092: \right.\label{8.99}
1093: \end{equation}
1094: It is easy to see that when $\sigma^0_1\neq\sigma^0_2$ the
1095: solutions $z_1$ and $z_2$ of (\ref{8.99}) take only the values
1096: $$z_1=0,\pm 1;\ \ \ \ z_2=0,\pm 1.$$
1097: In the same fashion, we can prove that the straight line orbits of
1098: (\ref{8.96}) given by
1099: $$y_1=\alpha_1y_3,\ \ \ \ y_2=\alpha_2y_3,\ \ \ \ \text{and}\
1100: y_1=\beta_1y_2,\ \ \ \ y_3=\beta_2y_2$$ have to satisfy that
1101: $$\alpha_i=0,\pm 1\ \ \ \ \text{and}\ \ \ \ \beta_i=0,\pm 1\ \ \ \
1102: (i=1,2).$$ Thus, we prove that when $\sigma^0_1\neq\sigma^0_2$,
1103: the number of straight line orbits of (\ref{8.96}) is exactly 26.
1104: 
1105: When $\sigma^0_1=\sigma^0_2$, we have that
1106: $\gamma_3=\frac{22}{9}\frac{L^2}{\pi^2}\gamma^2_2$ which implies
1107: that $\sigma^0_1=\sigma^0_2>0$. In this case, it is clear that
1108: $y=0$ is an asymptotically stable singular point of (\ref{8.96}).
1109: Hence, the transition of (\ref{8.89}) at $\lambda_0=\pi^2/L^2$ is
1110: I-type.
1111: 
1112: When $\sigma^0_1+\sigma^0_2>0$ and $\sigma^0_1\neq\sigma^0_2$, all
1113: straight line orbits of (\ref{8.96}) are toward $y=0$, which
1114: implies  that the regions at $y=0$, are stable, and
1115: $y=0$ is asymptotically stable; see \cite{b-book}. Thereby the transition of
1116: (\ref{8.89}) is Type-I.
1117: 
1118: When $\sigma^0_1+\sigma^0_2<0$ with $\sigma^0_1>0$, we can see, as
1119: in the case of $m=2$, that the regions at $y=0$ are hyperbolic,
1120: and when $\sigma^0_1+\sigma^0_2<0$ with $\sigma_1\leq 0$ the
1121: regions at $y=0$ are unstable. Hence, the transition is Type-II.
1122: 
1123: 
1124: \medskip
1125: 
1126: {\sc Step 5.} We prove Assertion (1). By Steps 3 and 4, if 
1127: $\gamma_3>\frac{26L^2}{27\pi^2}\gamma^2_2$,  
1128: the reduced equation
1129: (\ref{8.89}) bifurcates on $\lambda >\lambda_0=\pi^2/L^2$ to an
1130: attractor $\Sigma_{\lambda}$. All bifurcated equilibrium points of
1131: (\ref{8.60}) are one to one correspondence  to the bifurcated
1132: singular points of (\ref{8.89}). Therefore,  we only have to consider the stationary equations:
1133: \begin{equation}
1134: \beta_1(\lambda
1135: )y_i-\frac{\pi^2}{2L^2}[\sigma_1y^3_i+\sigma_2y_i\sum_{j\neq
1136: i}y^2_j]+o(|y|^3)=0,\ \ \ \ 1\leq i\leq m,\label{8.100}
1137: \end{equation}
1138: where $\sigma_1$ and $\sigma_2$ are as in (\ref{8.90}).
1139: 
1140: Consider the following approximative equations of (\ref{8.100})
1141: \begin{equation}
1142: \beta_1(\lambda )y_i-y_i(a_1y^2_i+a_2\sum_{j\neq i}y^2_j)=0\qquad 
1143: \text{ for }  1\leq i\leq m,\label{8.101}
1144: \end{equation}
1145: where $a_1=\pi^2\sigma_1/2L^2, a_2=\pi^2\sigma_2/2L^2$. It is
1146: clear that each regular bifurcated solution of (\ref{8.101})
1147: corresponds to a regular bifurcated solution of (\ref{8.100}).
1148: 
1149: We first prove that (\ref{8.101}) has $3^m-1$ bifurcated solutions
1150: on $\lambda >\lambda_0$. For each $k$  $(0\leq k\leq m-1)$,
1151: (\ref{8.101}) has $C^k_m\times 2^{m-k}$ solutions as follows: 
1152: \begin{equation}
1153: \left.
1154: \begin{aligned} 
1155: & y_{j_1}=0,\cdots,y_{j_k}=0 &&\text{ for } 1\leq j_l\leq
1156: m,\\
1157: & y^2_{r_1}=\cdots=y^2_{r_{m-k}}=\beta_1(a_1+(m-k-1)a_2)^{-1}
1158: &&\text{ for } r_i\neq  j_l.
1159: \end{aligned}
1160: \right.\label{8.102}
1161: \end{equation}
1162: Hence, the number of all bifurcated solutions of (\ref{8.101}) is
1163: $$\sum^{m-1}_{k=0}C^k_m\times 2^{m-k}=(2+1)^m-1=3^m-1.$$
1164: 
1165: We need to prove that all bifurcated solutions of (\ref{8.101})
1166: are regular. The Jacobian matrix of (\ref{8.101}) is given by
1167: \begin{equation}
1168: Dv=
1169: \begin{pmatrix}
1170: \beta_1-h_1(y)  &  2a_2y_1y_2      &\cdots   &    2a_2y_1y_m\\
1171: 2a_2y_2y_1      &   \beta_1-h_2(y) &\cdots   &    2a_2y_2y_m\\
1172: \vdots                &\vdots                  &             &    \vdots\\
1173: 2a_2y_my_1     & 2a_2y_my_2      &\cdots   &  \beta_1-h_m(y)
1174: \end{pmatrix},
1175: \label{8.103}
1176: \end{equation}
1177: where
1178: $$h_i(y)=3a_1y^2_i+a_2\sum_{j\neq i}y^2_j.$$
1179: For the solutions in (\ref{8.102}), without loss of generality, we
1180: take
1181: \begin{align*}
1182: &y_0=(y^0_1,\cdots,y^0_m),\\
1183: &y^0_i=0\qquad   \qquad 1\leq i\leq k, \\
1184: &y^0_{k+1}=\cdots=y^0_m=\beta^{{1}/{2}}_1(a_1+(m-k-1)a_2)^{-{1}/{2}}
1185: \end{align*}
1186: Inserting them into (\ref{8.103}) we find
1187: \begin{equation}
1188: Dv(y_0)=
1189: \left(\begin{matrix} \beta I_k&0\\
1190: 0&A_{m-k}
1191: \end{matrix}
1192: \right), \label{8.104}
1193: \end{equation}
1194: where
1195: \begin{align*}
1196: & 
1197: \beta
1198: =\beta_1\left(1-\frac{(m-k)a_2}{a_1+(m-k-1)a_2}\right)=\frac{(\sigma_1-\sigma_2)\beta_1}{\sigma_1+
1199: (m-k-1)\sigma_2}, \\
1200: & A_{m-k} =
1201: \left(\begin{matrix}
1202: \beta_1-(3a_1+(m-k-1)a_2)(y^{0}_{k+1})^2  &  \cdots & 2a_2(y^{0}_{k+1})^2\\
1203: \vdots&&\vdots\\
1204: 2a_2(y^{0}_{k+1})^2&\cdots&\beta_1-(3a_1+(m-k-1)a_2)(y^{0}_{k+1})^2
1205: \end{matrix}
1206: \right).
1207: \end{align*}
1208: Direct computation shows that 
1209: $$\text{det}A_{m-k}=\frac{\pi^{2m}\beta^m_1}{(a_1+(m-k-1)a_2)^mL^{2m}}\text{det}
1210: \left(\begin{matrix}
1211: -\sigma_1&\sigma_2&\cdots&\sigma_2\\
1212: \sigma_2&-\sigma_1&\cdots&\sigma_2\\
1213: \vdots&\vdots& & \vdots\\
1214: \sigma_2&\sigma_2&\cdots&-\sigma_1
1215: \end{matrix}
1216: \right),$$ 
1217: where $\sigma_1,\sigma_2$ are given by (\ref{8.90}).
1218: 
1219: Obviously, there are only finite number of $\lambda >\pi^2/L^2$
1220: satisfying
1221: \begin{align*}
1222: & \beta (\lambda )=\frac{(\sigma_1(\lambda )-\sigma_2(\lambda
1223: ))\beta_1(\lambda )}{\sigma_1(\lambda )+(m-k-1)\sigma_2(\lambda
1224: )}=0,\\
1225: & \text{det}A_{m-k}(\lambda )=0.
1226: \end{align*}
1227: Hence, for any $\lambda -\pi^2/L^2>0$ sufficiently small
1228: the Jacobian matrices (\ref{8.103}) at the singular points
1229: (\ref{8.102}) are non-degenerate. Thus, the bifurcated solutions of
1230: (\ref{8.101}) are regular.
1231: 
1232: Since all bifurcated singular points of (\ref{8.58}) with
1233: (\ref{8.62}) are non-degenerate, and when $\Sigma_{\lambda}$ is
1234: restricted on $x_ix_j$-plane $(1\leq i,j\leq m)$ the singular
1235: points are connected by their stable and unstable manifolds. 
1236: Hence all singular points in $\Sigma_{\lambda}$ are
1237: connected by their stable and unstable manifolds. Therefore,
1238: $\Sigma_{\lambda}$ must be homeomorphic to a sphere $S^{m-1}$.
1239: 
1240: Assertion (1) is proved.
1241: 
1242: 
1243: \medskip
1244: 
1245: {\sc Step 6. Proof of Assertions (2) and (3).} When $m=2$, by
1246: Step 5, $\Sigma_{\lambda}=S^1$ contains 8 non-degenerate singular
1247: points. By a theorem on minimal attractors in \cite{b-book},  $4$ singular points must be 
1248: attractors and the others are repellors, as shown in Figure \ref{f8.9-1}.
1249: 
1250: When $m=3$, we take the six singular points
1251: $$\pm Y_1=(\pm\beta_1a^{-1}_1,0,0),\pm
1252: Y_2=(0,\pm\beta_1a^{-1}_1,0),\pm Y_3=(0,0,\pm\beta_1a^{-1}_1)$$
1253: Then the Jacobian matrix (\ref{8.103}) at $Y_i$  $(1\leq i\leq 3)$ is
1254: $$Dv(\pm Y_i)=\left(\begin{array}{lll}
1255: \rho_1&&0\\
1256: &\rho_2\\
1257: 0&&\rho_3
1258: \end{array}
1259: \right),
1260: $$ 
1261: where
1262: $\rho_j=\beta_1\left(1-\frac{\sigma_2}{\sigma_1}\right)$ 
1263: as $j\neq i$ and $\rho_i=-2\beta_1$. Obviously, as $\sigma_2<\sigma_1$, 
1264: $0< \rho_j$ $(j\neq i)$ and $\rho_i<0$, in this case, $\pm Y_k$ $(1\leq
1265: k\leq 3)$ are repellors in $\Sigma_{\lambda}=S^2$, which implies
1266: that $\Sigma_{\lambda}$ contains $8$ attractors $\pm z_k$  $(1\leq k\leq
1267: 4)$ as shown in Figure \ref{f8.10} (a). As $\sigma_2>\sigma_1,
1268: \rho_j<0$ $(1\leq j\leq 3)$, the six singular point $\pm Y_k$  $(1\leq
1269: k\leq 3)$ are attractors, which implies that $\Sigma_{\lambda}$
1270: contains only six minimal attractors as shown in Figure \ref{f8.10} (b).
1271: Thus Assertion (2) is proved.
1272: 
1273: The claim for the saddle-node bifurcation in Assertion (3) can be
1274: proved by using the same method as in the proof of  Theorem \ref{t8.4}, and the claim
1275: for the singular point bifurcation can be proved by the same
1276: fashion as used in Step 5.
1277: 
1278: The proof of this theorem is complete.
1279: \ep
1280: 
1281: \br\la{r8.5}
1282: {\rm
1283: For the domain $\Omega =[0,L]^m\times D\subset
1284: \R^n(1\leq m<n)$ where $n\geq 2$ is arbitrary and $D\subset
1285: \R^{n-m}$ a bounded open set, Theorems \ref{t8.5} and \ref{t8.6} are also valid
1286: provided $\pi^2/L^2<\lambda_1$, where $\lambda_1$ is the first
1287: eigenvalue of the equation
1288: $$\left.
1289: \begin{aligned}
1290: &-\Delta e=\lambda e,&\ \ \ \ x\in D\subset \R^{n-m},\\
1291: &\frac{\partial e}{\partial n}|_{\partial D}=0,\\
1292: &\int_Dedx=0.
1293: \end{aligned}
1294: \right.
1295: $$ 
1296: }\er
1297: 
1298: 
1299: \br\la{r8.6}
1300: {\rm
1301: In Theorem \ref{t8.6}, the minimal
1302: attractors in the bifurcated attractor $\Sigma_{\lambda}$ can be
1303: expressed as
1304: \begin{equation}
1305: u_{\lambda}=(\lambda -\pi^2/L^2)^{{1}/{2}}e+o(|\lambda
1306: -\pi^2/L^2|^{{1}/{2}}),\label{8.105}
1307: \end{equation}
1308: where $e$ is a first eigenfunction of (\ref{8.61}). The expression
1309: (\ref{8.105}) can be derived from the reduced equations
1310: (\ref{8.89}).
1311: 
1312: We address here that the exponent $\beta ={1}/{2}$ in
1313: (\ref{8.105}), called the critical exponent in physics, is an
1314: important index in the phase transition theory in statistical
1315: physics, which arises only in the Type-I or the continuous phase
1316: transitions. It is interesting to point out that the critical
1317: exponent $\beta =1$ in (\ref{8.71}) is different from these $\beta
1318: ={1}/{2}$ appearing in (\ref{8.73}) and (\ref{8.105}). The
1319: first one occurs when  the container $\Omega\subset \R^3$ is a 
1320: non rectangular region, and the second one occurs when $\Omega$ is a
1321: rectangle or a cube. We shall continue to discuss this problem later
1322: from the physical viewpoint. \qed
1323: }\er
1324: 
1325: \section{Phase Transitions Under Periodic Boundary Conditions}
1326: \la{s8.2.5}
1327: 
1328: 
1329: When the sample or container $\Omega$ is a loop, or a torus, or 
1330: bulk in size, then the periodic boundary conditions are necessary.
1331: In this section, we shall discuss the problems in a loop domain
1332: and in the whole space $\Omega =\R^n$.
1333: 
1334: Let $\Omega =S^1\times (r_1,r_2)\subset \R^2$ be a loop domain,
1335: $0<r_1<r_2$. Then the boundary condition is given by
1336: \begin{equation}
1337: \left.
1338: \begin{aligned} 
1339: &u(\theta +2k\pi ,r)=u(\theta ,r)&& \text{ for } 
1340: 0\leq\theta\leq 2\pi ,\  r_1<r<r_2,\   k\in \Z,\\
1341: &\frac{\partial u}{\partial r}=0,\ \ \ \
1342: \frac{\partial^3u}{\partial r^3}=0&& \text{ at } r=r_1,r_2.
1343: \end{aligned}
1344: \right.\label{8.106}
1345: \end{equation}
1346: 
1347: Assume that the gap $r_2-r_1$ is small in comparison with the mean
1348: radius $r_0=(r_1+r_2)/2$. With proper scaling, we take
1349: the gap $r_2-r_1$ as $r_2-r_1=1$. Then this assumption is
1350: \begin{equation}
1351: r_0=(r_1+r_2)/2\gg 1, \qquad r_2-r_1 =1.\label{8.107}
1352: \end{equation}
1353: With this  condition, the Laplacian operator can be
1354: approximately expressed as
1355: \begin{equation}
1356: \Delta =\frac{\partial^2}{\partial
1357: r^2}+\frac{1}{r^2_0}\frac{\partial^2}{\partial\theta^2}.\label{8.108}
1358: \end{equation}
1359: 
1360: With the boundary condition (\ref{8.106}) and the operator
1361: (\ref{8.108}), the eigenvalues and eignfunctions of the linear
1362: operator $L_{\lambda}=-A+B_{\lambda}$ defined by (\ref{8.59}) are
1363: given by
1364: \begin{align*}
1365: & 
1366: \beta_K(\lambda )=K^2(\lambda -K^2), \\
1367: & e^1_K=\cos k_1\theta\cos k_2\pi (r-r_1),\\
1368: & e^2_K=\sin k_1\theta\cos k_2\pi (r-r_1), \\
1369: & K^2=\frac{k^2_1}{r^2_0}+k^2_2 &&  k_1,k_2\in \Z.
1370: \end{align*}
1371: 
1372: By (\ref{8.107}), the first eigenvalue of $L_{\lambda}$
1373: is
1374: $$\beta_1(\lambda )=\frac{1}{r^2_0}(\lambda -\frac{1}{r^2_0}),$$
1375: which has multiplicity 2, and the first eigenfunctions are
1376: $e^1_1=\cos\theta $  and $e^2_1=\sin\theta .$
1377: 
1378: 
1379: \bt\la{t8.7}
1380: Let $\Omega =S^1\times (r_1,r_2)$ 
1381: satisfy (\ref{8.107}). Then the following assertions hold true:
1382: 
1383: \begin{enumerate}
1384: 
1385: \item 
1386: If 
1387: $$\gamma_3>\frac{2r^2_0}{9}\gamma^2_2, $$
1388: then 
1389: the phase transition of (\ref{8.58}) with (\ref{8.106}) at
1390: $\lambda_0=1/ r^2_0$ is Type-I. Furthermore, the problem
1391: (\ref{8.58}) with (\ref{8.106}) bifurcates on $\lambda
1392: >\lambda_0=r^{-2}_0$ to a cycle attractor $\Sigma_{\lambda}=S^1$
1393: which consists of singular points, as shown in Figure \ref{f8.12}, and
1394: the singular points in $\Sigma_{\lambda}$ can be expressed as
1395: \begin{align*}
1396: & u_{\lambda}=\sigma^{-{1}/{2}}\left(\lambda
1397: -\frac{1}{r^2_0}\right)^{{1}/{2}}\cos (\theta
1398: +\theta_0)+o\left(|\lambda -\frac{1}{r^2_0}|^{{1}/{2}}\right),\\
1399: & \sigma =\frac{3}{4}\gamma_3-\frac{1}{6} r^2_0\gamma^2_2,
1400: \end{align*}
1401: where $\theta_0$ is the angle of $u_{\lambda}$ in
1402: $\Sigma_{\lambda}$. 
1403: 
1404: \item If
1405: $$\gamma_3<\frac{2r^2_0}{9}\gamma^2_2, $$  
1406: then the transition  is Type-II. Moreover,  the problem bifurcates on
1407: $\lambda <\lambda_0$ to  a cycle invariant set $\Gamma_{\lambda}=S^1$
1408: consisting of singular points, and there is a singularity
1409: separation at $\lambda^*<\lambda_0$,  where $\Gamma_{\lambda}$
1410: and an $S^1$ attractor $\Sigma_{\lambda}=S^1$ are generated such
1411: that the system undergoes a transition at $\lambda =\lambda_0$ from $u=0$ to
1412: $u_{\lambda}\in\Sigma_{\lambda}$, as shown in Figure \ref{f8.13}.
1413: \end{enumerate}
1414: \et
1415: 
1416: \begin{SCfigure}[25][t]
1417:   \centering
1418:   \includegraphics[width=0.4\textwidth]{8-12.pdf}
1419:   \caption{All points in $\Sigma_{\lambda}$ are singular
1420: points.}\la{f8.12}
1421:  \end{SCfigure}
1422: \begin{SCfigure}[25][t]
1423:   \centering
1424:   \includegraphics[width=0.4\textwidth]{8-13.pdf}
1425:   \caption{The point $\lambda^*$ is a singularity separation point
1426: from where two invariant sets $\Sigma_{\lambda}$ and
1427: $\Gamma_{\lambda}$ are separated with $\Sigma_{\lambda}$ being an
1428: attractor, and $\Gamma_{\lambda}\rightarrow 0$ as
1429: $\lambda\rightarrow\lambda_0$.}\la{f8.13}
1430:  \end{SCfigure}
1431: 
1432: \bp
1433: Let  $v=y\cos\theta +z\sin\theta$, 
1434: $u=v+\Phi (y,z)$,  and  $\Phi$ is the center manifold function. Then the
1435: reduced equations of (\ref{8.58}) with (\ref{8.106}) are given by
1436: \begin{equation}
1437: \left.
1438: \begin{aligned} 
1439: &\frac{dy}{dt}=\beta_1(\lambda
1440: )y+\frac{1}{\pi}\int^{r_1+1}_{r_1}\int^{2\pi}_0[\gamma_2\Delta
1441: v^2+\gamma_3\Delta v^3+\gamma_2\Delta u^2]e^1_1drd\theta ,\\
1442: &\frac{dz}{dt}=\beta_1(\lambda
1443: )z+\frac{1}{\pi}\int^{r_1+1}_{r_1}\int^{2\pi}_0[\gamma_2\Delta
1444: v^2+\gamma_3\Delta v^3+\gamma_2\Delta u^2]e^2_1drd\theta .
1445: \end{aligned}
1446: \right.\label{8.109}
1447: \end{equation}
1448: Direct computation shows that  (\ref{8.109})  can be  rewritten as
1449: \begin{equation}
1450: \left.
1451: \begin{aligned} 
1452: &\frac{dy}{dt}
1453: =\beta_1y-\frac{1}{\pi
1454: r^2_0}\left[\frac{3\pi}{4}\gamma_3y(y^2+z^2)
1455: +
1456: \int^{2\pi}_0 2\gamma_2 \Phi [y \cos^2\theta
1457:  + z \sin\theta\cos\theta] d\theta
1458: \right],\\
1459: &\frac{dz}{dt}=\beta_1z-\frac{1}{\pi
1460: r^2_0}\left[\frac{3\pi}{4}\gamma_3z(y^2+z^2)+
1461: \int^{2\pi}_02\gamma_2 \Phi [ z\sin^2\theta
1462:  + y \cos\theta\sin\theta] d\theta \right],
1463: \end{aligned}
1464: \right.\label{8.110}
1465: \end{equation}
1466: and the center manifold function $\Phi =\Phi (y,z)$ is
1467: $$
1468: \Phi =\frac{2\gamma_2}{\beta_2(\lambda )r^2_0}(y^2-z^2)\cos 2\theta
1469: +\frac{4\gamma_2}{\beta_2(\lambda )r^2_0}yz\sin 2\theta,
1470: $$ 
1471: where
1472: $$
1473: \beta_2(\lambda )=\frac{4}{r^2_0}(\lambda -\frac{4}{r^2_0}).
1474: $$
1475: Putting $\Phi$ into (\ref{8.110}), we obtain the approximate
1476: equations of (\ref{8.109}) as follows
1477: \begin{equation}
1478: \left.
1479: \begin{aligned}
1480: &\frac{dy}{dt}=\beta_1y-\frac{1}{r^2_0}\left(\frac{3}{4}\gamma_3+\frac{\gamma^2_2}{2(\lambda
1481: -4r^{-2}_0)}\right)y(y^2+z^2),\\
1482: &\frac{dz}{dt}=\beta_1z-\frac{1}{r^2_0}\left(\frac{3}{4}\gamma_3+\frac{\gamma^2_2}{2(\lambda
1483: -4r^{-2}_0)}\right)z(y^2+z^2).
1484: \end{aligned}
1485: \right.\label{8.111}
1486: \end{equation}
1487: At $\lambda =\lambda_0=r^{-2}_0$, we have
1488: $$\frac{3}{4}\gamma_3+\frac{\gamma^2_2}{2(\lambda
1489: -4r^{-2}_0)}=\frac{3}{4}\gamma_3-\frac{1}{6}r^2_0\gamma^2_2.$$
1490: 
1491: In the same fashion as used in Theorem \ref{t8.6}, we derive from (\ref{8.111}) the assertions of this theorem. Here the statement that the 
1492: bifurcated cycle $\Sigma_{\lambda}=S^1$ consists of singular
1493: points  can proved by that the equation (\ref{8.58}) with
1494: (\ref{8.106}) is invariant under the translation
1495: $$u(\theta ,r)\rightarrow u(\theta +\theta_0,r) \ \ \ \
1496: \forall\theta_0\in \R^1,$$ which ensures that the singular points
1497: of (\ref{8.58}) with (\ref{8.106}) arise as a cycle $S^1$. Thus,
1498: the theorem is proved. 
1499: \ep
1500: 
1501: \bigskip
1502: 
1503: Now, we consider the problem that the equation (\ref{8.58}) is
1504: defined in the whole space $\Omega =\R^n$  $(n\geq 2)$, with the
1505: periodic boundary condition
1506: \begin{equation}
1507: u(x+2K\pi )=u(x)\qquad \forall  K=(k_1,\cdots,k_n)\in \Z^n.\label{8.112}
1508: \end{equation}
1509: In this case, the eigenvalues and eigenfunctions of $L_{\lambda}$
1510: are given by
1511: \begin{align*}
1512: & 
1513: \beta_K(\lambda )=|K|^2(\lambda -|K|^2),&& K=(k_1,\cdots,k_n),\quad  |K|^2=k^2_1+\cdots+k^2_n, \\
1514: & e^1_K=\cos (k_1x_1+\cdots+k_nx_n),&& e^2_K=\sin (k_1x_1+\cdots+k_nx_n).
1515: \end{align*}
1516:  It is clear that the first eigenvalue $\beta_1(\lambda
1517: )=\lambda -1$ of $L_{\lambda}$ has multiplicity $2n$, and the
1518: first eigenfunctions are 
1519: $$e^1_j=\cos x_j,\ \ \ \ e^2_j=\sin x_j\qquad  \forall 1\leq j\leq n.$$
1520: 
1521: \bt\la{t8.8}
1522: \begin{enumerate}
1523: 
1524: \item If
1525: $$\gamma_3>\frac{14}{27}\gamma^2_2,$$ 
1526: then the phase transition of
1527: (\ref{8.58}) with (\ref{8.112}) at $\lambda_0=1$ is Type-I. Moreover, 
1528: 
1529: \begin{enumerate}
1530: \item the problem bifurcates
1531: from $(u,\lambda )=(0,1)$  to an attractor $\Sigma_{\lambda}$
1532: homeomorphic to a $(2n-1)$-dimensional sphere $S^{2n-1}$, and 
1533: \item 
1534: for each $k\ (0\leq
1535: k\leq n-1)$, the attractor $\Sigma_{\lambda}$ contains
1536: $C^k_n$-dimensional tori $\T^{n-k}$
1537:  consisting of singular points.
1538: \end{enumerate}
1539: 
1540: \item 
1541: If 
1542: $$\gamma_3<\frac{14}{27}\gamma^2_2,$$ 
1543:  then the transition  is Type-II. 
1544: \end{enumerate}
1545: \et
1546: 
1547: \bp
1548: We only have to prove Assertion (2), as the remaining part of the theorem is 
1549: essentially the same as  the proof for Theorem \ref{t8.6}. 
1550: 
1551: Since the space of all even functions is an invariant space of
1552: $L_{\lambda}+G$ defined by (\ref{8.59}), the problem (\ref{8.58})
1553: with (\ref{8.112}) has solutions given in (\ref{8.102}) with
1554: $m=n$ in the space of even functions.
1555: 
1556: By the translation invariance of (\ref{8.58}) and (\ref{8.112}),
1557: for each $k$  $(0\leq k\leq n-1)$ and a fixed index
1558: $(j_1,\cdots,j_k)$, the steady state solution associated with
1559: (\ref{8.102}) generates an $(n-k)$-dimensional torus  $\T^{n-k}$ which
1560: consists of steady state solutions of (\ref{8.58}) and
1561: (\ref{8.112}). For example if $(j_1,\cdots,j_k)=(1,\cdots,k)$, the
1562: $(n-k)$-dimensional singularity torus $\T^{n-k}$ is
1563: $$\T^k=\left\{u(x+\theta )=\sum^n_{j=k+1}y_j\cos (x_j+\theta_j)+o(|y|),\
1564: \ \ \ \forall (\theta_{n+1},\cdots,\theta_n)\in \R^{n-k}\right\},
1565: $$ 
1566: where
1567: $u(x)$ is the steady state solution of (\ref{8.58}) with
1568: (\ref{8.112}) associated with (\ref{8.102}) with $y_1=\cdots
1569: =y_k=0, y_{k+1}=\cdots=y_n$.
1570: 
1571: Obviously, for a fixed $(j_1,\cdots,j_k)$, the $2^{n-k}$ steady
1572: state solutions of (\ref{8.58}) and (\ref{8.112}) associated with
1573: (\ref{8.102}) are in the same singular torus $\T^{n-k}$.
1574: Furthermore, for two different index $k$-tuples $(j_1,\cdots,j_k)$
1575: and $(i_1,\cdots,i_k)$, the two associated singularity tori are
1576: different. Hence, for each $0\leq k\leq n-1$, there are exactly
1577: $C^k_n$  $(n-k)$-dimensional singularity tori in $\Sigma_{\lambda}$.
1578: Thus the proof is complete.
1579: \ep
1580: 
1581: \section{Cahn-Hilliard Equations Coupled with Entropy}
1582: \la{s8.2.6}
1583: 
1584: 
1585: When a phase separation takes place in a binary system, the entropy
1586: varies, and if the phase transition is Type-II, it will yield
1587: latent heat. Hence, it is necessary to discuss the equations
1588: (\ref{8.51}), which are called the Cahn-Hilliard equations coupled
1589: with entropy.
1590: 
1591: To make the equations (\ref{8.51}) non-dimensional, let
1592: \begin{align*}
1593: &x=lx^{\prime},&&  t=\frac{1}{k_2}l^4t^{\prime},&& 
1594: u=u_0u^{\prime},&& S=S_0S^{\prime},\\
1595: &\mu =\frac{k_1}{k_2}l^2, &&  \alpha_1=\frac{1}{k_2}l^4a_1,&& \alpha_2=\frac{l^2}{k_2S_0}u^2_0a_2,&& \lambda =-\frac{l^2}{k_2}b_1,\\
1596: &
1597: \gamma_1=\frac{l^2}{k_2}S_0b_0,&&
1598: \gamma_2=\frac{l^2}{k_2}u_0b_2,&&
1599: \gamma_3=\frac{l^2}{k_2}u^2_0b_3.
1600: \end{align*}
1601: Omitting the primes, equations (\ref{8.51}) are in the following
1602: form
1603: \begin{equation}
1604: \left.\begin{aligned} &\frac{\partial S}{\partial t}=\mu\Delta
1605: S-\alpha_1S-\alpha_2u^2,\\
1606: &\frac{\partial u}{\partial t}=-\Delta^2u-\lambda\Delta u+\Delta
1607: (\gamma_1Su+\gamma_2u^2+\gamma_3u^3),\\
1608: &\int_{\Omega}u(x,t)dx=0,\\
1609: &\frac{\partial }{\partial n} ( u,     \Delta u, S) =0 \qquad  \text{ on } \partial \Omega,\\
1610: &u(x,0)=\varphi (x).
1611: \end{aligned}
1612: \right.\label{8.113}
1613: \end{equation}
1614: 
1615: By assumptions (\ref{8.50}) and (\ref{8.52}), the coefficients
1616: satisfy
1617: $$\mu >0,\ \ \ \ \alpha_1>0,\ \ \ \ \alpha_2>0,\ \ \ \
1618: \gamma_1>0,\ \ \ \ \gamma_3>0.$$ 
1619: 
1620: 
1621: 
1622: \bt\la{t8.9}
1623: Let $\Omega
1624: =\Pi^n_{k=1}(0,L_k)\subset \R^n$ satisfy that $L=L_1=\cdots
1625: =L_m>L_j(1\leq m\leq n)$ for  any $j> m$. 
1626: 
1627: \begin{enumerate}
1628: 
1629: \item[(1)] For the case where  $m=1$, let 
1630: $$\sigma
1631: =\frac{3}{2}\gamma_3-\frac{\alpha_2\gamma_1}{\alpha_1}-\frac{\alpha_2\gamma_1L^2}{2(\alpha_1L^2+4\pi^2\mu
1632: )}-\frac{L^2\gamma^2_2}{3\pi^2}.$$
1633: 
1634: \begin{enumerate}
1635: 
1636: \item  If $\sigma <0$, then  the phase transition of
1637: (\ref{8.113}) at $\lambda =\pi^2/L^2$ is Type-II and Assertion (1)
1638: in Theorem \ref{t8.5} holds  true.
1639: 
1640: \item If $\sigma >0$ the phase transition is
1641: Type-I and Assertion (2) in Theorem \ref{t8.5} holds  true.
1642: 
1643: \end{enumerate}
1644: 
1645: \item[(2)] For the case where  $m\geq 2$, let 
1646: $$\widetilde{\sigma}=\frac{9}{2}\gamma_3-\alpha_2\gamma_1\left(\frac{2}{\alpha_1}+\frac{L^2}{2(\alpha_1L^2+
1647: 4\pi^2\mu
1648: )}+\frac{2L^2}{\alpha_1L^2+2\pi^2\mu}\right)-\frac{13L^2\gamma^2_2}{3\pi^2}.$$
1649: 
1650: \begin{enumerate}
1651: 
1652: \item If
1653: $\widetilde{\sigma}>0$,  the phase transition of (\ref{8.113}) at
1654: $\lambda =\pi^2/L^2$ is Type-I and Assertions (1) and (2) in
1655: Theorem \ref{t8.6} hold true. 
1656: 
1657: \item If $\widetilde{\sigma}<0$, then  the phase transition
1658: is Type-II and Assertion (3) in  Theorem \ref{t8.6} holds  true.
1659: \end{enumerate}
1660: 
1661: \end{enumerate}
1662: \et
1663: 
1664: 
1665: \bp
1666:  It suffices to compute the reduced equations of
1667: (\ref{8.113}) on the center manifold. Similar to (\ref{8.89}), the
1668: second order approximation of the reduced equation can be
1669: expressed as
1670: \begin{align}
1671: \frac{dy_i}{dt}=&\beta_1y_i-\frac{\pi^2}{2L^2}[\sigma_1y^3_i+\sigma_2y_i\sum_{j\neq
1672: i}y^2_j]\label{8.114}\\
1673: &-\frac{2\pi^2}{L_1\cdots
1674: L_nL^2}\gamma_1\sum^m_{j=1}y_j\int_{\Omega}\Phi_1(y)\cos\frac{\pi
1675: x_i}{L}\cos\frac{\pi x_j}{L}dx,\nonumber
1676: \end{align}
1677: where $\beta_1,\sigma_1$ and $\sigma_2$ are as in (\ref{8.89}),
1678: and the center manifold function $\Phi_1(y)$ derived
1679: from the first equation in (\ref{8.113}) can be expressed as
1680: \begin{equation}
1681: \left.
1682: \begin{aligned} 
1683: &\Phi_1(y)=\sum^{\infty}_{|K|\geq
1684: 0}\Phi_K(y)\varphi_K\\
1685: &\Phi_K(y)=\frac{-\alpha_2}{\lambda_K\|\varphi_K\|^2}\int_{\Omega}
1686: \left(\sum^m_{i=1}y_i\cos
1687: \frac{\pi x_i}{L}\right)^2\varphi_Kdx
1688: \end{aligned}
1689: \right.\label{8.115}
1690: \end{equation}
1691: where $\lambda_K$ and $\varphi_K$ are the eigenvalues and
1692: eigenfunctions of the following equation
1693: \begin{align*}
1694: &-\mu\Delta\varphi_K+\alpha_1\varphi_K=\lambda_K\varphi_K,\\
1695: &\frac{\partial\varphi_K}{\partial n}|_{\partial\Omega}=0,
1696: \end{align*} 
1697: which are given by 
1698: \begin{align*}
1699: & \lambda_K=\alpha_1+\mu K^2\pi^2,\\
1700: & \varphi_0=1,\ \ \ \ \varphi_K=\cos\frac{k_1\pi
1701: x_1}{L^2_1}\cdots\cos\frac{k_n\pi x_n}{L^2_n}, \\
1702: & K=(k_1/L_1,\cdots,k_n/L_n),\quad   |K|^2=\sum^n_{i=1}k^2_i/L^2_i&& \text{ for } k_i\in \Z,\ \ \ \ 1\leq i\leq n.
1703: \end{align*}
1704: 
1705: Let 
1706: $$K_i=(\delta_{i1}/L_1,\cdots,\delta_{in}/L_n).$$
1707: Then  we find
1708: \begin{align*}
1709: \Phi_0 =&  \frac{-\alpha_2}{\alpha_1L_1\cdots L_n}\int_{\Omega}\left(\sum^n_{i=1}y_i\cos\frac{\pi
1710: x_i}{L}\right)^2dx=\frac{-\alpha_2}{2\alpha_1}\sum^m_{j=1}y^2_j, 
1711: \\
1712: \Phi_k=&\frac{-\alpha_2}{\lambda_K\|\varphi_K\|^2}\int_{\Omega}\left(\sum^m_{i=1}y_i\cos\frac{\pi
1713: x_i}{L}\right)^2dx\\
1714: =&\left\{
1715:   \begin{aligned} 
1716:      &  0&& \text{ if }  K\neq K_l+K_r,\\
1717:      &\frac{-\alpha_2}{\lambda_K\|\varphi_K\|^2}\sum^m_{i,j=1}y_iy_j\int_{\Omega}\cos\frac{\pi
1718: x_i}{L}\cos\frac{\pi x_j}{L}\varphi_Kdx&& \text{ if }  K=K_l+K_r,
1719: \end{aligned}
1720: \right.
1721: \end{align*}
1722: for some $1\leq r, l\leq m$. Thus we derive that
1723: \begin{align*}
1724: \Phi_K=
1725: &\left\{
1726: \begin{aligned}
1727:   &\frac{-\alpha_2}{\lambda_K\|\varphi_K\|^2}y^2_j\int_{\Omega}\cos^2\frac{\pi
1728: x_j}{L}e_k&& \text{ if }  K=K_j+K_j, \ \ 1\leq j\leq m\\
1729: &\frac{-2\alpha_2}{\lambda_K\|\varphi_K\|^2}y_iy_j\int_{\Omega}\cos\frac{\pi
1730: x_i}{L}\cos\frac{\pi x_j}{L}e_K && \text{ if } K=K_i+K_j, \ \ i\neq j
1731: \end{aligned}
1732: \right.\\
1733: =&\left\{\begin{aligned} &\frac{-\alpha_2}{2(\alpha_1+4\pi^2\mu
1734: /L^2)}y^2_j  \qquad \qquad && \text{ if }K=2K_j, \ \ 1\leq j\leq m,\\
1735: &\frac{-2\alpha_2}{\alpha_1+2\pi^2\mu /L^2}y_iy_j  && \text{ if }
1736: K=K_i+K_j, \ \ i\neq j,\  \ 1\leq i,j\leq m.
1737: \end{aligned}
1738: \right.
1739: \end{align*}
1740: Putting $\Phi_0$ and $\Phi_K$ in (\ref{8.115}) we obtain
1741: \begin{align*}
1742: \Phi_1=&\frac{-\alpha_2}{2\alpha_1}\sum^n_{j=1}y^2_j-\frac{\alpha_2}{2(\alpha_1+4\pi^2\mu
1743: /L^2)}\sum^n_{j=1}y^2_j\cos\frac{2\pi x_j}{L}\\
1744: &-\frac{2\alpha_2}{\alpha_1+2\pi^2\mu
1745: /L^2}\sum_{i<j}y_iy_j\cos\frac{\pi x_i}{L}\cos\frac{\pi x_j}{L}.
1746: \end{align*}
1747: 
1748: Then, inserting $\Phi_1$ into (\ref{8.114}),  we derive the following reduced
1749: equations:
1750: \begin{align}
1751: \frac{dy_i}{dt}=
1752: &\beta_1(\lambda)y_i-\frac{\pi^2}{2L^2}\left[\left(\sigma_1-\frac{\alpha_2\gamma_1}{\alpha_1}-\frac{\alpha_2\gamma_1}
1753: {2(\alpha_1+4\pi^2\mu /L^2}\right)y^3_i\right.\label{8.116}\\
1754: &+ \left. \left(\sigma_2-\frac{\alpha_2\gamma_1}{\alpha_1}-\frac{2\alpha_2\gamma_1}{\alpha_1+2\pi^2\mu
1755: /L^2}\right)y_i\sum_{j\neq i}y^2_j \right],\ \ \ \ 1\leq i\leq
1756: m. \nonumber
1757: \end{align}
1758: Then the remaining part of the proof can be achieved in the same fashion as the proofs for Theorems \ref{t8.5} and \ref{t8.6}. The proof is complete.
1759: \ep
1760: 
1761: 
1762: \br\la{r8.7}
1763: {\rm
1764: From the phenomenological viewpoint, the
1765: coefficient $\mu >0$ in (\ref{8.113}) is small. If let $\mu =0$,
1766: then in the equilibrium state we have that
1767: $S=-\frac{\alpha_2}{\alpha_1}u^2$. In this case, (\ref{8.113}) are
1768: referred to the original Cahn-Hilliard equation (\ref{8.58}) with
1769: $\gamma^{\prime}_3=\gamma_3-\alpha_2\gamma_1/\alpha_1$ as  
1770: the coefficient of the cubic term, and the
1771: criterion  $\sigma =0$ in Assertions (1) and (2) of Theorem \ref{t8.9}
1772: are respectively equivalent to
1773: $$\gamma^{\prime}_3=\frac{2}{9}\frac{L^2\gamma^2_2}{\pi^2}\ \ \ \
1774: \text{and}\ \ \ \
1775: \gamma^{\prime}_3=\frac{26}{27}\frac{L^2\gamma^2_2}{\pi^2},$$
1776: which coincide with these in Theorems \ref{t8.5} and \ref{t8.6}. Hence, if we
1777: consider the coefficient $\mu >0$ small, then the criterion
1778: $\sigma =0$ are respectively equivalent to
1779: $$\gamma^{\prime}_3=\frac{2}{9}\frac{L^2\gamma^2_2}{\pi^2}-\frac{4}{3}\frac{\pi^2\mu\alpha_2\gamma_1}{\alpha^2_1L^2},
1780: \ \ \ \ \text{and}\ \ \ \ \
1781: \gamma^{\prime}_3=\frac{26}{27}\frac{L^2\gamma^2_2}{\pi^2}-\frac{4}{3}\frac{\pi^2\mu\alpha_2\gamma_1}{\alpha^2_1L^2}.$$
1782: Hence the item $-(4\pi^2\mu\alpha_2\gamma_1)/3\alpha^2_1L^2$ is
1783: the effect yielded by $\mu\Delta S$.
1784: \qed
1785: }\er
1786: 
1787: \section{Physical remarks}
1788: \la{s8.2.7}
1789: 
1790: We now address  the physical significance for the phase
1791: transition theorems obtained in the previous sections.
1792: 
1793: \subsection{Equation of critical parameters}For a binary system, the equation describing the control
1794: parameters $T,p,\Omega$ at the critical states is simple. 
1795: 
1796: We first
1797: consider the critical temperature $T_c$. There are two different
1798: critical temperatures $T_1$ and $T_0$ in the Cahn-Hilliard
1799: equation. $T_1$ is the one given by (\ref{8.56}),  at which the
1800: coefficient $b_1(T,p)$ or $\lambda =-l^2b_1(T,p)/k$ will change its
1801: sign, and $T_0$ satisfies that $\lambda_0<\lambda_1$, and for
1802: fixed $p$,
1803: \begin{equation}
1804: \lambda (T)\left\{
1805: \begin{aligned} 
1806: &  <\rho_1&& \text{ if } T>T_0,\\
1807: & =\rho_1  && \text{ if } T=T_0,\\
1808: & >\rho_1 && \text{ if } T<T_0,
1809: \end{aligned}
1810: \right.
1811: \label{8.117}
1812: \end{equation}
1813: where $\rho_1$ is the first eigenvalue of (\ref{8.61}), which
1814: depends on the geometrical properties of the material such as the size of the container of the sample $\Omega$. When $\Omega =(0,L)^m\times D\subset \R^n$ is a
1815: rectangular domain with $L>$ diameter of $D, \rho_1=\pi^2/L^2$.
1816: Hence, in general at the critical temperature $T_1$ a binary
1817: system does not undergo any phase transitions, but the phase transition  does occur at $T=T_0$.
1818: 
1819: At $T_1$ and $T_0$ we know that
1820: \begin{equation}
1821: \lambda (T_1)=0,\ \ \ \ \lambda (T_0)=\rho_1.\label{8.118}
1822: \end{equation}
1823: For a rectangular domain, $\rho_1=\pi^2/L^2$, therefore from
1824: (\ref{8.118}) we see that $T_1$ is a limit of the critical
1825: temperature $T_0$ of phase transition as the size of $\Omega$
1826: tends to infinite.
1827: 
1828: In fact,  for a general domain, it is easy to see that the first eigenvalue $\rho_1$ of the Laplace operator
1829: is inversely  proportional to the square of  the  maximum diameter of
1830: $\Omega$:
1831: \begin{equation}
1832: \rho_1\sim\frac{1}{L^2},\label{8.120}
1833: \end{equation}
1834: where $L$ represents the diameter scaling of $\Omega$. 
1835: 
1836: Thus the equation  of critical parameters in the
1837: Cahn-Hilliard equation, by (\ref{8.118}) and (\ref{8.120}), is
1838: given by
1839: \begin{equation}
1840: \lambda (T,p)=\frac{C}{L^2},\label{8.121}
1841: \end{equation}
1842: where $C>0$ is a constant depending on the geometry of $\Omega$.
1843: According to the Hildebrand theory (see Reichl \cite{reichl}), the function
1844: $\lambda (T,p)$ can be expressed in a explicit formula. If
1845: regardless of the term $|\nabla u|^2$, the molar Gibbs free energy
1846: takes the following form
1847: \begin{equation}
1848: g=\mu_A(1-u)+\mu_Bu+RT(1-u)\ln (1-u)+RTu\ln
1849: u+au(1-u),\label{8.122}
1850: \end{equation}
1851: where $\mu_A,\mu_B$ are the chemical potential of $A$ and $B$
1852: respectively, $R$ the molar gas constant, $a>0$ the measure of
1853: repel action between $A$ and $B$. Therefore, the coefficient $b_1$
1854: in (\ref{8.55}) with constant $p$ is
1855: $$b_1=\frac{\partial^2g}{\partial
1856: u^2}|_{u=u_0}=\frac{1}{u_0(1-u_0)}RT-a\ \ \ \ (a=a(p)),$$ where
1857: $u_0=\bar{u}_B$ is the constant concentration of $B$. Hence
1858: $$\lambda
1859: (T,p)=-\frac{l^2}{k}b_1=\frac{2al^2}{k}-\frac{l^2R}{ku_0(1-u_0)}T.$$
1860: Thus, equation (\ref{8.121}) is expressed as
1861: \begin{equation}
1862: \frac{l^2R}{ku_0(1-u_0)}T=\frac{2al^2}{K}-\frac{C}{L^2}.\label{8.123}
1863: \end{equation}
1864: Equation (\ref{8.123}) gives the critical parameter curve of a
1865: binary system with constant pressure for temperature $T$ and
1866: diameter scaling $L$ of container $\Omega$.  Because $T\geq 0$,
1867: from (\ref{8.122}) we can deduce the following physical
1868: conclusion.
1869: 
1870: \bigskip
1871: 
1872: \noindent
1873: {\bf Physical Conclusion 7.1.}  %8.3
1874: {\it 
1875: Under constant pressure, for any
1876: binary system with given geometrical shape of the container $\Omega$,
1877: there is a value $L_0>0$ such that as the diameter scaling
1878: $L<L_0$, no phase separation takes place at all temperature $T\geq
1879: 0$, and as $L>L_0$ phase separation will occur at some critical
1880: temperature $T_0>0$ satisfying (\ref{8.123}).
1881: }
1882: 
1883: \medskip
1884: 
1885: We shall see later that it is a universal property  that the dynamical
1886: properties of phase transitions depend on the geometrical shape
1887: and size of the container or sample $\Omega$.
1888: 
1889: 
1890: 
1891: \subsection{Physical explanations of phase transition theorems}
1892: 
1893: We first briefly recall the classical
1894: thermodynamic theory for a binary system. Physically, phase
1895: separation processes taking place in an unstable state are called
1896: spinodal decompositions; see Cahn and Hilliard \cite{CH57} and Onuki \cite{onuki}. 
1897: When consider the concentration $u$ as homogeneous in $\Omega$,
1898: then by (\ref{8.122}) the dynamic equation of a binary system is
1899: an ordinary differential equation:
1900: \begin{equation}
1901: \frac{du}{dt} = -\frac{dg}{du}\label{8.124}=
1902: 2au-RT\ln\frac{u}{1-u}+\mu_A-\mu_B-a.
1903: \end{equation}
1904: 
1905: Let $u_0$  $(0<u_0<1)$ be the steady state solution of (\ref{8.124}).
1906: Then, by the Taylor expansion at $u=u_0$, omitting the $n$th order
1907: terms with $n\geq 4$, (\ref{8.124}) can be rewritten as
1908: \begin{equation}
1909: \frac{dv}{dt}=\lambda v+b_2v^2+b_3v^3,\label{8.125}
1910: \end{equation}
1911: where 
1912: \begin{align*}
1913: & v=u-u_0,   &&   \lambda =2a-\frac{1}{u_0(1-u_0)}RT,\\
1914: & b_2=\frac{1-2u_0}{2u^2_0(1-u_0)^2}RT, &&b_3=-\frac{1}{3}\frac{1-u_0-2u^2_0+3u^3_0}{u^3_0(1-u_0)^4}RT.
1915: \end{align*}
1916: It is easy to see that
1917: \begin{align*}
1918: &b_2\left\{
1919:   \begin{aligned} 
1920:      & =0&& \text{ if } u_0=\frac{1}{2},\\
1921: &   \neq 0&& \text{ if } u_0\neq\frac{1}{2},
1922: \end{aligned}
1923: \right.\\
1924: &b_3<0 \qquad   \forall 0<u_0<1.
1925: \end{align*}
1926: 
1927: It is clear that the critical parameter curve $\lambda=0$  in the $T-u_0$ plane is given by 
1928: $$T_0=2a u_0 (1-u_0)/R,$$
1929: which is schematically illustrated in the classical phase diagram; see the dotted line in Figure~\ref{f8.14}. We obtain from (\ref{8.125}) the following  transition steady states: 
1930: $$
1931: v^\pm = \frac{-1}{2b_3}(b_2 \pm \sqrt{b_2^2 -4b_3 \lambda})\qquad \text{ for } b_2^2 -4b_3 \lambda>0.
1932: $$
1933: By Theorem~\ref{t8.4}, we see that there is $T^*=T^*(u_0)$   satisfying 
1934: that $b_2^2-4b_3 \lambda=0$; namely, 
1935: \begin{equation}
1936:  T^* = \frac{ 2 a u_0 (1-u_0)}{R(1-\beta(u_0))}, \qquad 
1937:  \beta(u_0)= \frac{3 (1-2u_0)^2(1-u_0)}{16(1-u_0-2u_0^2 + 3u_0^3)},
1938:  \label{8.126-1}
1939: \end{equation}
1940: such that if $T_0 < T < T^*$, 
1941: \begin{align*}
1942: &
1943: v^+ \text{ is } \left\{
1944: \begin{aligned}
1945: & \text{ locally stable (metastable) } && \text{ for } 0 < u_0 < \frac12, \\
1946: & \text{ unstable } && \text{ for } \frac12 < u_0 < 1, 
1947: \end{aligned}
1948: \right.  \\
1949: &
1950: v^- \text{ is } \left\{
1951: \begin{aligned}
1952: & \text{ locally stable (metastable) } && \text{ for }   \frac12< u_0 <1, \\
1953: & \text{ unstable } && \text{ for } 0 < u_0 <  \frac12.
1954: \end{aligned}
1955: \right.  
1956: \end{align*}
1957: Here 
1958: $$T^*(u_0) =\frac{T_0(u_0)}{1-\beta_0(u_0)} \ge T_0(u_0)$$
1959: is illustrated by the solid line in Figure~\ref{f8.14}. 
1960: This shows that the region $T_0(u_0) < T < T^*(u_0)$  is metastable, which is marked 
1961: by the shadowed region in Figure \ref{f8.14}. See, among others,  Reichl \cite{reichl}, 
1962:  Novick-Cohen and Segal \cite{NS84} ,  and Langer \cite{langer71} 
1963: for the phase transition diagram from the classical thermodynamic theory.
1964: 
1965: \begin{SCfigure}[25][t]
1966:   \centering
1967:   \includegraphics[width=0.5\textwidth]{8-14.pdf}
1968:   \caption{Typical phase diagram from classical
1969: thermodynamic theory.}
1970:   \la{f8.14}
1971:  \end{SCfigure}
1972: 
1973: In the following we shall discuss the spinodal decomposition in a unified fashion by
1974: applying the phase transition theorems presented in the previous sections.
1975: 
1976: As mentioned in the Introduction, phase separation processes
1977: of binary systems occur in two ways, one of which proceeds
1978: continuously depending on $T$, and the other one does not.
1979: Obviously, the classical theory does not explain these phenomena.
1980: In fact, the first one can be described by the Type-I phase
1981: transition, and the second one can be explained by the Type-II and
1982: Type-III phase transitions.
1983: 
1984: We first consider the case where the container $\Omega
1985: =\Pi^n_{i=1}(0,L_i)$ with $L=L_1=\cdots=L_m>L_j(j>m)$ is a
1986: rectangular  domain. Thus, by Theorems \ref{t8.5} and \ref{t8.6} 
1987: (or Theorem \ref{t8.9})
1988: there are only two phase transition types: Type-I  and Type-II, 
1989: with the type of transition  depending on $L$. We see that if
1990: $$L^2<\left\{\begin{aligned}
1991: &\frac{9}{2}\frac{\pi^2\gamma_3}{\gamma^2_2} && \text{for}\
1992: m=1,\\
1993: &\frac{27}{26}\frac{\pi^2\gamma_3}{\gamma^2_2}  && \text{for}\
1994: m\geq 2, \end{aligned} \right.
1995: $$ 
1996: then the transition  is Type-I, i.e., the phase
1997: pattern formation gradually varies as the  temperature
1998: decreases. In this case, no meta-stable states and no latent heat
1999: appear. The phase diagram is given by Figure \ref{f8.15}, 
2000: where the solid lines $u_i^T$ ($i=1,2$) represent   the transition solutions.
2001: \begin{SCfigure}[25][t]
2002:   \centering
2003:   \includegraphics[width=0.4\textwidth]{8-15.pdf}
2004:   \caption{The state $u_0=\bar{u}_B$ is
2005: stable  if $T_c=T_0 < T$, and  the state $u_0$ is unstable  if $T < T_0$, where $T_0$
2006: is as in (\ref{8.117}).}
2007:   \la{f8.15}
2008:  \end{SCfigure}
2009: 
2010: If $L$ satisfies that
2011: $$L^2>\left\{\begin{aligned}
2012: &\frac{9}{2}\frac{\pi^2\gamma_3}{\gamma^2_2} &&  \text{for}\
2013: m=1,\\
2014: &\frac{27}{26}\frac{\pi^2\gamma_3}{\gamma^2_2} && \text{for}\
2015: m\geq 2,
2016: \end{aligned}
2017: \right.
2018: $$ 
2019: then the phase transition is Type-II. Namely,  there is a
2020: leaping change in phase pattern formation at the critical temperature
2021: $T_c$. The phase diagram for Type-II  transition is given by Figure \ref{f8.16}.
2022: \begin{SCfigure}[25][t]
2023:   \centering
2024:   \includegraphics[width=0.4\textwidth]{8-16.pdf}
2025:   \caption{A Type-II phase transition.}
2026: \la{f8.16}
2027:  \end{SCfigure}
2028: 
2029: 
2030: In Figure \ref{f8.16}, 
2031: $T_0$ is the critical temperature as in (\ref{8.117}), 
2032: $T^*$  is  defined by (\ref{8.126-1}) and is the saddle-node bifurcation point of (\ref{8.125}).
2033: The constant concentration $u=u_0$ is stable in $T_c<T$,  is
2034: meta-stable in $T_0<T<T_c$, and is unstable in $T<T_0$.
2035: The two bifurcated states $U^T_1$
2036: and $U^T_2$  from $T^*$ are meta-stable in $T_0<T<T^*$, and  are stable in $T<T_0$. 
2037: Here for $i=1,2$, $U_i^T=u^T_i + u_0$, and $u_i$ are the separated solutions of 
2038: (\ref{8.58}) with (\ref{8.62}) from $T^*$.
2039: 
2040: There is a remarkable difference between Type-I and Type-II transitions.  The
2041: Type-I phase transition occurs at $T=T_0$ and Type-II does in
2042: $T_0<T<T^*$. Furthermore, latent heat is 
2043: accompanied  the Type-II phase transition. Actually, when a binary
2044: system undergoes a transition  from $u_0$ to $U^T_i$  $(i=1,2)$, 
2045: there is a gap  $|U^T_i-u_0|^2=|u_i^T|^2>\varepsilon >0$  for any $T_0<T<T^*$. 
2046: By the first
2047: equation in (\ref{8.113}) it yields a jump of entropy between
2048: $u_0$ and $U^T_i$:
2049: $$
2050: \delta S_i=\int_{\Omega}Sdx
2051: =-\frac{\alpha_2}{\alpha_1}\int_{\Omega}|u^T_i|^2dx<0,
2052: $$
2053: where $S=S_i-\bar{S}_0$ represents the entropy density deviation.
2054: Hence the latent heat is given by
2055: $$\delta H=T\delta
2056: S_i=-\frac{\alpha_2T}{\alpha_1}\int_{\Omega}|u^T_i|^2dx<0,$$
2057: which implies that the process from $u_0$ to $U^T_i$ is
2058: exothermic, and the process from $U^T_i$ to $u_0$ is endothermic.
2059: 
2060: Now, we consider the case where the container $\Omega$ is
2061: non rectangular. Thus, by Theorem \ref{t8.4} the transition is Type-III,   and its
2062: phase diagram is given by Figure \ref{f8.17}.
2063: \begin{SCfigure}[25][t]
2064:   \centering
2065:   \includegraphics[width=0.4\textwidth]{8-17.pdf}
2066:   \caption{A Type-III phase transition.}\la{f8.17}
2067:  \end{SCfigure}
2068: 
2069: 
2070: In Figure \ref{f8.17},  $T_0$  and $T^*$ are the same as those in Figure \ref{f8.16}. The state $u_0$ is stable in
2071: $T^*<T$, is metastable in $T_0<T<T^*$, and unstable in $T<T_0$.
2072: The equilibrium state $U^T_1$ separated from $T^*$ is metastable in
2073: $T_0<T<T^*$, and is stable in $T<T_0$. However, the equilibrium
2074: state $U^T_2$ separated from $T^*$ is unstable in $T_0<T<T^*$, and
2075: is metastable in $T<T_0$.
2076: % Hence, $u_0$ is supercooled in
2077: %$T_0<T<T^*$ and $u^T_1$ is superheated in $T_0<T<T^*$.
2078: 
2079: Similar to the Type-II, the Type-III phase transition has also
2080: latent heat, which occurs in $T_0<T<T^*$. But the difference
2081: between Type-II and Type-III is that Type-II has $2m $  $(m\geq 1)$
2082: stable equilibrium states separated from $T=T^*$, but Type-III has
2083: just one. The $2m$ stable states of a Type-II transition  are of some symmetry
2084: caused by $\Omega$, and we shall investigate it later. A
2085: particular aspect of Type-III is that there is a  state
2086: $U^T_2$ bifurcated from $(u,T)=(u_0,T_0)$, which is rarely
2087: observed in experiments.
2088: 
2089: \subsection{Symmetry and periodic structure}Physical experiments have shown that in pattern formation via
2090: phase separation,  periodic or semi-periodic structure appears. From
2091: Theorems \ref{t8.7} and \ref{t8.8} we  see that for the loop domains and bulk
2092: domains which can be considered as $\R^n$ or $\R^m\times D$  $(D\subset
2093: \R^{n-m})$ the steady state solutions of the Cahn-Hilliard equation
2094: are periodic, and for rectangular domains they are semi-periodic,  and the periodicity is associated with the mirror image symmetry.
2095: 
2096: Let $\Omega =(0,L)^m\times D$  $(m\geq 1)$. By Remark  \ref{r8.5},
2097: Theorem \ref{t8.5} is valid for $\Omega$. Actually, in this case the
2098: following space
2099: $$
2100: \widetilde{H}=\left\{u\in L^2(\Omega )|\
2101: u=\sum^{\infty}_{|K|=1}y_k\cos\frac{k_1\pi}{L}x_1\cdots\cos\frac{k_m\pi}{L}x_m
2102: \right\}\subset
2103: H
2104: $$ 
2105: is invariant for the Cahn-Hilliard equation (\ref{8.58}) and
2106: (\ref{8.62}). All separated equilibrium states in Theorem \ref{t8.5} are
2107: in $\widetilde{H}$. From the physical viewpoint, all equilibrium states
2108: $u(x)$ and their mirror image states
2109: $u(L-x^{\prime},x^{\prime\prime})$ are the same to describe  the pattern
2110: formation. Mathematically, under the mirror image transformation
2111: $$
2112: x\rightarrow (L-x^{\prime},x^{\prime\prime}),\ \ \ \
2113: x^{\prime}=(x_1,\cdots,x_m),\ \ \ \
2114: x^{\prime\prime}=(x_{m+1},\cdots,x_n),
2115: $$ 
2116: the Cahn-Hilliard equation (\ref{8.58}) with (\ref{8.62}) is invariant. Hence,
2117:  the steady state solutions will appear in pairs. In
2118: particular, for  Type-I phase transition, there is a remarkable mirror
2119: image symmetric. We address this problem as follows.
2120: 
2121: Let $m=1$ in $\Omega =(0,L)^m\times D$. By (\ref{8.73}) there are
2122: two bifurcated stable equilibrium states, and their projections on
2123: the first eigenspace are
2124: \begin{eqnarray*}
2125: &&u_1=y\cos\frac{\pi x_1}{L},\ \ \ \ u_2=-y\cos\frac{\pi
2126: x_1}{L},\\
2127: &&y=\sqrt{2(\lambda
2128: -\pi^2/L^2)}/\left(\frac{3}{2}\gamma_3-\frac{L^2}{3\pi^2}\gamma^2_2\right).
2129: \end{eqnarray*}
2130: It is clear that $u_2(x_1)=u_1(L-x_1)$.
2131: 
2132: Let $m=2$. By Theorem \ref{t8.6} the bifurcated attractor
2133: $\Sigma_{\lambda}$ contains 8 equilibrium states, whose projections
2134: are given by
2135: \begin{eqnarray*}
2136: &&u^{\pm}_1=\pm y_0e_1,\ \ \ \ u^{\pm}_2=\pm y_1(e_1+e_2),\\
2137: &&u^{\pm}_3=\pm y_0e_2,\ \ \ \ u^{\pm}_4=\pm y_1(-e_1+e_2),
2138: \end{eqnarray*}
2139: where $e_1=\left(\cos\frac{\pi
2140: x_1}{L},0\right)$  and $e_2=\left(0,\cos\frac{\pi x_2}{L}\right)$ form an
2141: orthogonal basis in $\R^2$, and 
2142: \begin{eqnarray*}
2143: &&y_0=\sqrt{2(\lambda -\pi^2/L^2)}/\sqrt{\sigma_1},\\
2144: &&y_1=\sqrt{2(\lambda -\pi^2/L^2)}/\sqrt{\sigma_1+\sigma_2}\ \ \ \ \text{ with }
2145: \sigma_1,\sigma_2\ \text{as\ in\ (\ref{8.90})}.
2146: \end{eqnarray*}
2147: These eight equilibrium states constitute an octagon in $\R^2$, as
2148: shown in Figure \ref{f8.18}, and they are divided into two classes:
2149: ${\mathcal{A}}_1=\left\{u^{\pm}_1,u^{\pm}_3\right\}$ and
2150: ${\mathcal{A}}_3=\left\{u^{\pm}_2,u^{\pm}_4\right\}$ by the
2151: $\frac{\pi}{2}$-rotation group $G(\frac{\pi}{2})$. 
2152: Namely,  with the action of $G(\frac{\pi}{2}), {\mathcal{A}}_i$  $(i=1,2)$ are
2153: invariant:
2154: $$Bu\in{\mathcal{A}}_i,\ \ \ \ \forall u\in{\mathcal{A}}_i\
2155: \text{and}\ B\in G(\frac{\pi}{2}),$$ 
2156: where $G(\frac{\pi}{2})$ consists of
2157: the orthogonal matrices
2158: $$B^{\pm}_1=\pm\left(\begin{matrix}
2159: 1&0\\
2160: 0&1
2161: \end{matrix}
2162: \right),\ \ \ \ 
2163: B^{\pm}_2=\pm\left(\begin{matrix} 1&0\\
2164: 0&-1
2165: \end{matrix}
2166: \right),\ \ \ \ 
2167: B^{\pm}_3=\pm\left(\begin{matrix} 0&-1\\
2168: 1&0
2169: \end{matrix}
2170: \right),\ \ \ \ 
2171: B^{\pm}_4=\pm\left(\begin{matrix} 0&1\\
2172: 1&0
2173: \end{matrix}
2174: \right).$$ 
2175: The stability of the equilibrium states $u_k$ is associated
2176: with  both classes ${\mathcal{A}}_1$ and ${\mathcal{A}}_2$,
2177: i.e., either the elements in ${\mathcal{A}}_1$ are stable, or those
2178: in ${\mathcal{A}}_2$ are stable; see Figure \ref{f8.9-1}. By (\ref{8.104})
2179: we can derive  the criterion as follows
2180: \begin{align*}
2181: &  u^{\pm}_{2k+1}\in{\mathcal{A}}_1 
2182: \text{ are stable }  &&  \Leftrightarrow \quad \frac{22}{9}\frac{L^2 \gamma^2_2}{\pi^2}>\gamma_3>\frac{26}{27}\frac{L^2\gamma^2_2}{\pi^2} && \text{ for } k=0,1, \\
2183: & u^{\pm}_{2k}\in{\mathcal{A}}_2  \text{ are stable } && \Leftrightarrow \quad \gamma_3>\frac{22}{9}\frac{L^2 \gamma^2_2}{\pi^2} && \text{ for } k=1, 2,
2184: \end{align*}
2185: \begin{SCfigure}[25][t]
2186:   \centering
2187:   \includegraphics[width=0.4\textwidth]{8-18.pdf}
2188:   \caption{The eight equilibrium states in the case where $m=2$ 
2189:   given by Theorem~\ref{t8.6}.}\la{f8.18}
2190:  \end{SCfigure}
2191: 
2192: In Figure \ref{f8.18}, we see that elements  in ${\mathcal{A}}_1$ and
2193: ${\mathcal{A}}_2$ have a $\frac{\pi}{4}$ difference in their phase
2194: angle. However, in their pattern structure,
2195: $u^{\pm}_{2K+1}\in{\mathcal{A}}_1$ and
2196: $u^{\pm}_{2k}\in{\mathcal{A}}_2$ also have a $\pi /4$ deference at
2197: the angle between the lines of $u^{\pm}_{2K+1}=0$ and
2198: $u^{\pm}_{2K}=0$. In fact, the lines that
2199: \begin{eqnarray*}
2200: &&u^+_1=-u^-_1=y_0\cos\frac{\pi x_1}{L}=0,\ \ \ \ \text{and}\\
2201: &&u^+_3=-u^-_3=y_0\cos\frac{\pi x_2}{L}=0
2202: \end{eqnarray*}
2203: are given by $x_1=L/2$ and $x_2=L/2$ respectively, as shown in
2204: Figure \ref{f8.19}(a), and the lines
2205: \begin{eqnarray*}
2206: &&u^+_2=-u^-_2=y_1\left(\cos\frac{\pi x_1}{L}+\cos\frac{\pi
2207: x_2}{L}\right)=0,\ \ \ \ \text{and}\\
2208: &&u^+_4=-u^-_4=y_1\left(-\cos\frac{\pi x_1}{L}+\cos\frac{\pi
2209: x_2}{L}\right)=0
2210: \end{eqnarray*}
2211: are given by $x_2=L-x_1$ and $x_2=x_1$ respectively as shown in
2212: Figure \ref{f8.19}(b).
2213: \begin{figure}[hbt]
2214:   \centering
2215:   \includegraphics[width=0.28\textwidth]{8-19a.pdf}\qquad
2216:   \includegraphics[width=0.28\textwidth]{8-19b.pdf}
2217:   \caption{}\la{f8.19}
2218:  \end{figure}
2219: 
2220: Let $m=3$. Then the bifurcated attractor $\Sigma_{\lambda}$
2221: contains 26 equilibrium states which can by divide into three
2222: classes by the 3-dimensional
2223: $\left(\frac{\pi}{2},\frac{\pi}{2},\frac{\pi}{2}\right)$-rotation
2224: group $G\left(\frac{\pi}{2},\frac{\pi}{2},\frac{\pi}{2}\right)$ as
2225: follows
2226: \begin{align*}
2227: &{\mathcal{A}}_1=\left\{u^{\pm}_1=\pm y_0e_1,u^{\pm}_2=\pm
2228: y_0e_2,u^{\pm}_3=\pm y_0e_3\right\}, \\
2229: &{\mathcal{A}}_2=\left\{u^{\pm}_4=\pm y_1(e_1+e_2),u^{\pm}_5=\pm
2230: y_1(e_1+e_3),u^{\pm}_6=\pm y_1(e_2+e_3),\right.\\
2231: &\ \ \ \ u^{\pm}_7=\pm y_1(-e_1+e_2),u^{\pm}_8=\pm
2232: y_1(-e_1+e_3),u^{\pm}_9=\pm y_1(-e_2+e_3)\}, \\
2233: &{\mathcal{A}}_3=\{u^{\pm}_{10}=\pm
2234: y_1(e_1+e_2+e_3),u^{\pm}_{11}=\pm y_1(-e_1+e_2+3_3),\\
2235: &\ \ \ \ u^{\pm}_{12}=\pm y_1(e_1-e_2+e_3),u^{\pm}_{13}=\pm
2236: y_1(e_1+e_2-e_3)\}.
2237: \end{align*}
2238: Only these elements in ${\mathcal{A}}_1$ or in ${\mathcal{A}}_3$
2239: are stable, and  they are determined by the
2240: following criterion
2241: \begin{align*}
2242: &\text{elements in}\ {\mathcal{A}}_1\ \text{is\
2243: stable}\Leftrightarrow\frac{22}{9}\frac{L^2\gamma^2_2}{\pi^2}>\gamma_3>\frac{26}{27}\frac{L^2\gamma^2_2}{\pi^2},\\
2244: &\text{elements\ in}\ {\mathcal{A}}_3\ \text{is\
2245: stable}\Leftrightarrow \gamma_3>\frac{22}{9}\frac{L^2\gamma^2_2}{\pi^2}.
2246: \end{align*}
2247: 
2248: \subsection{Critical exponents}
2249: From (\ref{8.73}) and (\ref{8.105}) we see that for Type-I phase
2250: transition of a binary system the critical exponent $\beta
2251: =\frac{1}{2}$. In this case, it is a second order phase transition with the Ehrenfest classification scheme,
2252: and there is a gap in heat capacity at critical temperature $T_0$.
2253: To see this, by (\ref{8.73}) and (\ref{8.105}) we have
2254: $$u^T=\left\{\begin{aligned}
2255: & 0 &&\text{if}\ T_0<T,\\
2256: &\alpha (\lambda (T)-\pi^2/L^2)^{{1}/{2}}e_1+o(|\lambda
2257: -\pi^2/L^2|^{{1}/{2}})&&\text{if}\ T<T_0,
2258: \end{aligned}
2259: \right.
2260: $$ 
2261: and the free energy for (\ref{8.58}) at $u^T$ is
2262: \begin{align*}
2263: F(u^T)=&\int_{\Omega}\left[\frac{1}{2}|\nabla
2264: u^T|^2-\frac{\lambda}{2}|u^T|^2+o(|u^T|^2)\right]dx\\
2265: =&\int_{\Omega}\frac{1}{2}[-\Delta u^T-\lambda
2266: u^T]u^T+\frac{1}{3}\gamma_2(u^T)^3+\frac{1}{4}\gamma_3(u^T)^4dx\\
2267: =&\left\{\begin{aligned} 
2268: &0 &&  \text{if}\ T_0<T,\\
2269: &-\frac{\alpha^2}{2}(\lambda
2270: -\pi^2/L^2)^2\cdot\int_{\Omega}[e^2_1+\frac{\alpha^2}{4}\gamma_3e^4_1]dx+o(|\lambda
2271: -\pi^2/L^2|^2) && \text{if} \ T > T_0.
2272: \end{aligned}
2273: \right.
2274: \end{align*}
2275: Thus, the heat capacity $C$ at $T=T_0$ satisfies
2276: $$C^+-C^-=-T_0\frac{\partial^2F(u^T)}{\partial
2277: T^2}\Big|_{T_0^+}+T_0\frac{\partial^2F(u^T)}{\partial
2278: T^2}\Big|_{T_0^-}=\alpha_1T_0\left(\frac{d\lambda}{dT}\right)^2|_{T=T_0}.$$
2279: It is known that $d\lambda /dT\neq 0$; hence the heat capacity at
2280: $T=T_0$ has a finite jump.
2281: 
2282: From (\ref{8.71}) we know that for the Type-III case, the critical
2283: exponent $\beta =1$. Thus, it is not hard to  deduce
2284: that the continuous phase transition in Type-III is of the 3rd order.
2285: 
2286: 
2287: \appendix
2288: 
2289: \section{Dynamic Transition Theory for Nonlinear Systems}
2290: In this appendix we recall some basic elements of the dynamic transition theory developed by the authors \cite{b-book, chinese-book}, which are used to carry out the dynamic transition analysis for the binary systems in this article. 
2291: 
2292: \subsection{New classification scheme}
2293: Let $X$  and $ X_1$ be two Banach spaces,   and $X_1\subset X$ a compact and
2294: dense inclusion. In this chapter, we always consider the following
2295: nonlinear evolution equations
2296: \begin{equation}
2297: \left. 
2298: \begin{aligned} 
2299: &\frac{du}{dt}=L_{\lambda}u+G(u,\lambda),\\
2300: &u(0)=\varphi ,
2301: \end{aligned}
2302: \right.\label{5.1}
2303: \end{equation}
2304: where $u:[0,\infty )\rightarrow X$ is unknown function,  and 
2305: $\lambda\in \R^1$  is the system parameter.
2306: 
2307: Assume that $L_{\lambda}:X_1\rightarrow X$ is a parameterized
2308: linear completely continuous field depending continuously on
2309: $\lambda\in \R^1$, which satisfies
2310: \begin{equation}
2311: \left. 
2312: \begin{aligned} 
2313: &L_{\lambda}=-A+B_{\lambda}   && \text{a sectorial operator},\\
2314: &A:X_1\rightarrow X   && \text{a linear homeomorphism},\\
2315: &B_{\lambda}:X_1\rightarrow X&&  \text{a linear compact  operator}.
2316: \end{aligned}
2317: \right.\label{5.2}
2318: \end{equation}
2319: In this case, we can define the fractional order spaces
2320: $X_{\sigma}$ for $\sigma\in \R^1$. Then we also assume that
2321: $G(\cdot ,\lambda ):X_{\alpha}\rightarrow X$ is $C^r(r\geq 1)$
2322: bounded mapping for some $0\leq\alpha <1$, depending continuously
2323: on $\lambda\in \R^1$, and
2324: \begin{equation}
2325: G(u,\lambda )=o(\|u\|_{X_{\alpha}}),\ \ \ \ \forall\lambda\in
2326: \R^1.\label{5.3}
2327: \end{equation}
2328: 
2329: Hereafter we always assume the conditions (\ref{5.2}) and
2330: (\ref{5.3}), which represent that the system (\ref{5.1}) has
2331: a dissipative structure.
2332: 
2333: \bd\la{d5.1}
2334: We say that the system (\ref{5.1}) has a
2335: transition of equilibrium from $(u,\lambda )=(0,\lambda_0)$ on
2336: $\lambda >\lambda_0$ (or $\lambda <\lambda_0)$ if  the following two conditions are 
2337: satisfied:
2338: 
2339: \begin{itemize}
2340: 
2341: \item[(1)] when $\lambda <\lambda_0$ (or $\lambda >\lambda_0),
2342: u=0$ is locally asymptotically stable for (\ref{5.1}); and
2343: 
2344: \item[(2)] when $\lambda >\lambda_0$ (or $\lambda <\lambda_0)$,
2345: there exists a neighborhood $U\subset X$ of $u=0$ independent of
2346: $\lambda$, such that for any $\varphi\in U \setminus \Gamma_{\lambda}$ the
2347: solution $u_{\lambda}(t,\varphi )$ of (\ref{5.1}) satisfies that
2348: $$\left. 
2349: \begin{aligned}
2350: &\limsup_{t\rightarrow\infty}\|u_{\lambda}(t,\varphi
2351: )\|_X\geq\delta (\lambda )>0,\\
2352: &\lim_{\lambda\rightarrow\lambda_0}\delta(\lambda )\geq 0,
2353: \end{aligned}
2354: \right.$$ 
2355: where $\Gamma_{\lambda}$ is the stable manifold of
2356: $u=0$, with  codim $\Gamma_{\lambda}\geq 1$ in $X$
2357: for $\lambda >\lambda_0$ (or $\lambda <\lambda_0)$.
2358: \end{itemize}
2359: \ed
2360: 
2361: Obviously, the attractor bifurcation of (\ref{5.1}) is a type of
2362: transition. However,  bifurcation and
2363: transition are two different, but related concepts. 
2364: Definition \ref{d5.1} defines the transition of (\ref{5.1}) from a stable
2365: equilibrium point to other states (not necessary equilibrium state).
2366: In general, we can define transitions from one attractor to
2367: another as follows.
2368: 
2369: 
2370: Let the eigenvalues (counting multiplicity) of $L_{\lambda}$ be given by
2371: $$\{\beta_j(\lambda )\in \C\ \   |\ \ j=1,2,\cdots\}$$
2372: Assume that
2373: \begin{align}
2374: &  \text{Re}\ \beta_i(\lambda )
2375: \left\{ 
2376:  \begin{aligned} 
2377:  &  <0 &&    \text{ if } \lambda  <\lambda_0,\\
2378: & =0 &&      \text{ if } \lambda =\lambda_0,\\
2379: & >0&&     \text{ if } \lambda >\lambda_0,
2380: \end{aligned}
2381: \right.   &&  \forall 1\leq i\leq m,  \label{5.4}\\
2382: &\text{Re}\ \beta_j(\lambda_0)<0 &&  \forall j\geq
2383: m+1.\label{5.5}
2384: \end{align}
2385: 
2386: The following theorem is a basic principle of transitions from
2387: equilibrium states, which provides sufficient conditions and a basic
2388: classification for transitions of nonlinear dissipative systems.
2389: This theorem is a direct consequence of the center manifold
2390: theorems and the stable manifold theorems; we omit the proof.
2391: 
2392: \bt\la{t5.1}
2393:  Let the conditions (\ref{5.4}) and
2394: (\ref{5.5}) hold true. Then, the system (\ref{5.1}) must have a
2395: transition from $(u,\lambda )=(0,\lambda_0)$, and there is a
2396: neighborhood $U\subset X$ of $u=0$ such that the transition is one
2397: of the following three types:
2398: 
2399: \begin{itemize}
2400: \item[(1)] {\sc Continuous Transition}: 
2401: there exists an open and dense set
2402: $\widetilde{U}_{\lambda}\subset U$ such that for any
2403: $\varphi\in\widetilde{U}_{\lambda}$,  the solution
2404: $u_{\lambda}(t,\varphi )$ of (\ref{5.1}) satisfies
2405: $$\lim\limits_{\lambda\rightarrow\lambda_0}\limsup_{t\rightarrow\infty}\|u_{\lambda}(t,\varphi
2406: )\|_X=0.$$ In particular, the attractor bifurcation of (\ref{5.1})
2407: at $(0,\lambda_0)$ is a continuous transition.
2408: 
2409: \item[(2)] {\sc Jump Transition}: 
2410: for any $\lambda_0<\lambda <\lambda_0+\varepsilon$ with some $\varepsilon >0$, there is an open
2411: and dense set $U_{\lambda}\subset U$ such that 
2412: for any $\varphi\in U_{\lambda}$, 
2413: $$\limsup_{t\rightarrow\infty}\|u_{\lambda}(t,\varphi
2414: )\|_X\geq\delta >0,$$ 
2415: where $\delta >0$ is independent of $\lambda$. 
2416: This type of transition  is also called the discontinuous 
2417: transition. 
2418: 
2419: \item[(3)] {\sc Mixed Transition}: 
2420: for any $\lambda_0<\lambda <\lambda_0+\varepsilon$  with some $\varepsilon >0$, 
2421: $U$ can be decomposed into two open sets
2422: $U^{\lambda}_1$ and $U^{\lambda}_2$  ($U^{\lambda}_i$ not necessarily
2423: connected):
2424: $$\bar{U}=\bar{U}^{\lambda}_1+\bar{U}^{\lambda}_2,\ \ \
2425: \ U^{\lambda}_1\cap U^{\lambda}_2=\emptyset ,$$ 
2426: such that
2427: \begin{align*}
2428: &\lim\limits_{\lambda\rightarrow\lambda_0}\limsup_{t\rightarrow\infty}\|u(t,\varphi
2429: )\|_X=0   &&   \forall\varphi\in U^{\lambda}_1,\\
2430: & \limsup_{t\rightarrow\infty}\|u(t,\varphi
2431: )\|_X\geq\delta >0 && \forall\varphi\in U^{\lambda}_2.
2432: \end{align*}
2433: %where  $U^{\lambda}_1$ is called the stable domain,  and  $U^{\lambda}_2$
2434: %is the unstable domain.
2435: \end{itemize}
2436: \et
2437: 
2438: 
2439: With this theorem in our disposal, we are in position to give a new dynamic classification scheme for dynamic phase transitions.
2440: 
2441: \begin{definition}[Dynamic Classification of Phase Transition]
2442: The phase transitions for  (\ref{5.1}) at $\lambda =\lambda_0$ is classified using  their  dynamic properties: continuous, jump, and mixed as given in Theorem~\ref{t5.1}, which are called Type-I, Type-II and Type-III respectively.
2443: \end{definition}
2444: 
2445: An important aspect of the  transition theory is to determine which 
2446: of the three types of transitions given by Theorem \ref{t5.1} occurs in
2447: a specific  problem. By  reduction to the center manifold of
2448: (\ref{5.1}), we know that the type of transitions for (\ref{5.1}) at
2449: $(0,\lambda_0)$ is completely dictated  by its reduction equation
2450: near $\lambda =\lambda_0$, which can be
2451: expressed as:
2452: \begin{equation}
2453: \frac{dx}{dt}=J_{\lambda}x+PG(x + \Phi (x,\lambda ),\lambda ) \ \ \
2454: \ \text{ for } x\in \R^m,\label{5.6}
2455: \end{equation}
2456: where $J_{\lambda}$ is the $m\times m$ order Jordan matrix
2457: corresponding to the eigenvalues given by (\ref{5.4}),  $\Phi
2458: (x,\lambda )$ is the center manifold function of (\ref{5.1}) near
2459: $\lambda_0$, $P:X\rightarrow E_{\lambda}$ is the canonical
2460: projection, and 
2461: $$
2462: E_{\lambda}=\cup_{1\leq i\leq m}\cup_{k\in N}\{u\in X_1|\ \
2463: (L_{\lambda}-\beta_i(\lambda ))^ku=0\}
2464: $$ 
2465: is the eigenspace of $L_{\lambda}$.
2466: 
2467: By the spectral theorem, (\ref{5.6}) can be expressed into the
2468: following explicit form
2469: \begin{equation}
2470: \frac{dx}{dt}=J_{\lambda}x+g(x,\lambda ),\label{5.7}
2471: \end{equation}
2472: where
2473: \begin{equation}
2474: \left. 
2475: \begin{aligned} 
2476: &g(x,\lambda )=(g_1(x,\lambda ),\cdots
2477: ,g_m(x,\lambda )),\\
2478: &g_j(x,\lambda )= <G(\sum^m_{i=1}x_ie_i + \Phi (x,\lambda ),\lambda
2479: ), e^*_j> \quad   \forall 1\leq j\leq m.
2480: \end{aligned}
2481: \right.\label{5.8}
2482: \end{equation}
2483: Here $e_j$ and $e^*_j$  $(1\leq j\leq m)$ are the eigenvectors of
2484: $L_{\lambda}$ and $L^*_{\lambda}$ respectively corresponding to the
2485: eigenvalues $\beta_j(\lambda )$ as in (\ref{5.4}).
2486: 
2487: In particular, if $G(u,\lambda )$ has the Taylor expansion
2488: \begin{equation}
2489: G(u,\lambda )=G_k(u,\lambda )+o(\|u\|^k_{X_{\alpha}}),\label{5.9}
2490: \end{equation}
2491: for some $k\geq 2$, where $G_k(u,\lambda )$ is a $k$-multilinear operator, then (\ref{5.7})  can be rewritten as
2492: \begin{equation}
2493: \frac{dx}{dt}=J_{\lambda}x+g_k(x,\lambda
2494: )+o(|x|^k),\label{5.10}
2495: \end{equation}
2496: where
2497: \begin{align*}
2498: &g_k(x,\lambda )=(g_{k1}(x,\lambda ),\cdots ,g_{km}(x,\lambda)),\\
2499: &g_{kj}(x,\lambda )=<G_k(\sum^n_{i=1}x_ie_i,\lambda ),e^*_j>
2500: &&    \forall 1\leq j\leq m.
2501: \end{align*}
2502: 
2503: When $x=0$ is an isolated singular point of $g_k(x,\lambda )$, in
2504: general the transition of (\ref{5.1}) is determined by the
2505: first-order approximate bifurcation equation of (\ref{5.10}) as follows:
2506: \begin{equation}
2507: \frac{dx}{dt}=J_{\lambda}x+g_k(x,\lambda ).\label{5.11}
2508: \end{equation}
2509: 
2510: The following theorem is useful to distinguish the transition
2511: types of (\ref{5.1}) at $(u,\lambda )=(0,\lambda_0)$.
2512: 
2513: 
2514: \bt\la{t5.2}
2515: Let the conditions (\ref{5.4}) and
2516: (\ref{5.5}) hold true, and $U\subset \R^m$ be a neighborhood of $x=0$.
2517: Then we have the following assertions:
2518: 
2519: \begin{itemize}
2520: 
2521: \item[(1)] If the transition of (\ref{5.1}) at $(0,\lambda_0)$ is
2522: continuous, then there is an open and dense set
2523: $\widetilde{U}\subset U$ such that for any $x_0\in\widetilde{U}$
2524: the solution $x(t,x_0)$ of (\ref{5.7}) at $\lambda =\lambda_0$
2525: with $x(0,x_0)=x_0$ satisfies that
2526: $$\lim\limits_{t\rightarrow\infty}x(t,x_0)=0.$$
2527: 
2528: \item[(2)] If there exists an open and dense set
2529: $\widetilde{U}\subset U$ such that for any $x_0\in\widetilde{U}$
2530: the solution $x(t,x_0)$ of (\ref{5.7}) at $\lambda =\lambda_0$
2531: satisfies
2532: $$\limsup_{t\rightarrow\infty}|x(t,x_0)|\neq 0,$$
2533: then the transition is a jump transition. 
2534: 
2535: \item[(3)] If the transition is
2536: mixed, then there exists an open set $\widetilde{U}\subset U$
2537: such that for any $x_0\in\widetilde{U}$ the solution $x(t,x_0)$ of
2538: (\ref{5.7}) at $\lambda =\lambda_0$ with $x(0,x_0)=x_0$ satisfies
2539: $$\lim\limits_{t\rightarrow\infty}x(t,x_0)=0.$$
2540: 
2541: 
2542: \item[(4)] If the vector field in (\ref{5.7}) at $\lambda
2543: =\lambda_0$ satisfies
2544: $$<J_{\lambda_0}x+g(x,\lambda_0),x><0\quad \forall x\in U,\ x\neq 0,$$
2545: then the transition of (\ref{5.1}) at $(0,\lambda_0)$ is an
2546: $S^{m-1}$-attractor bifurcation, and  the transition  is continuous. 
2547: 
2548: \item[(5)] If the vector field $g(x,\lambda_0)$ given by (\ref{5.8})
2549: satisfies
2550: $$<g(x,\lambda_0),x>>0 \ \ \ \ \forall x\in U,\ \ \ \ x\neq 0,$$
2551: then the transition is jump.
2552: \end{itemize}
2553: \et
2554: 
2555: In general, the conditions in Assertions (1)-(3) of Theorem \ref{t5.2}
2556: are not sufficient. They, however, do give sufficient conditions when (\ref{5.1}) has
2557: a variational structure. To see this, let (\ref{5.1}) be a gradient-type equation. Under the conditions (\ref{5.4}) and (\ref{5.5}), in a
2558: neighborhood $U\subset X$ of $u=0$, the center manifold $M^c$ in
2559: $U$ at $\lambda =\lambda_0$ consists three subsets
2560: $$M^c=W^u+W^s+D,$$
2561: where $W^s$ is the stable set, $W^u$ is the unstable set, and $D$ is  the
2562: hyperbolic set of (\ref{5.7}). Then we have the following
2563: theorem.
2564: 
2565: 
2566: \bt\la{t5.3} 
2567: Let (\ref{5.1}) be a gradient-type equation,
2568: and the conditions (\ref{5.4}) and (\ref{5.5}) hold true. If $u=0$
2569: is an isolated singular point of (\ref{5.1}) at $\lambda
2570: =\lambda_0$, then we have the following assertions:
2571: 
2572: \begin{itemize}
2573: 
2574: \item[(1)] The transition of (\ref{5.1}) at $(u,\lambda)=(0,\lambda_0)$ is continuous if and only if $u=0$ is locally asymptotically stable at $\lambda =\lambda_0$, i.e., the center
2575: manifold is stable: $M^c=W^s$. Moreover, (\ref{5.1}) 
2576: bifurcates from $(0,\lambda_0)$ to minimal attractors consisting  of
2577: singular points of (\ref{5.1}). 
2578: 
2579: \item[(2)] If the stable set $W^s$ of (\ref{5.1}) has no interior points in $M^c$, i.e.,
2580: $M^c=\bar{W}^u+\bar{D}$, then the transition is jump.
2581: \end{itemize}
2582: \et
2583: 
2584: \subsection{Transitions from simple eigenvalues}
2585: We consider the transition of (\ref{5.1}) from a simple critical
2586: eigenvalue. Let the eigenvalues $\beta_j(\lambda )$ of
2587: $L_{\lambda}$ satisfy (\ref{5.4})  and (\ref{5.5}) with $m=1$.
2588: Then the first eigenvalue $\beta_1(\lambda )$ must be a real eigenvalue. 
2589: Let $e_1(\lambda )$ and $e^*_1(\lambda )$   be  the eigenvectors of
2590: $L_{\lambda}$ and $L^*_{\lambda}$ respectively corresponding to
2591: $\beta_1(\lambda )$ with
2592: $$L_{\lambda_0}e_1=0,\ \ \ \ L^*_{\lambda_0}e^*_1=0,\ \ \ \
2593: <e_1,e^*_1>=1.$$ 
2594: Let $\Phi (x,\lambda )$    be the center manifold
2595: function of (\ref{5.1}) near $\lambda =\lambda_0$. We assume that
2596: \begin{equation}
2597: <G(xe_1+\Phi (x,\lambda_0),\lambda_0),e^*_1>=\alpha
2598: x^k+o(|x|^k),\label{5.36}
2599: \end{equation}
2600: where $k\geq 2$ an integer and $\alpha\neq 0$ a real number.
2601: \begin{figure}%[hbt]
2602:   \centering
2603:   \includegraphics[width=0.35\textwidth]{5-5a.pdf} \quad 
2604:   \includegraphics[width=0.2\textwidth]{5-5b.pdf}
2605:   \caption{Topological structure of the jump transition of
2606: (\ref{5.1}) when $k$=odd and $\alpha >0$: (a) $\lambda
2607: <\lambda_0$; (b) $\lambda\geq\lambda_0$. Here the horizontal line
2608: represents the center manifold.}\la{f5.5}
2609:  \end{figure}
2610:  \begin{figure}%[hbtp]
2611:   \centering
2612:   \includegraphics[width=0.23\textwidth]{5-6a.pdf}
2613:   \includegraphics[width=0.35\textwidth]{5-6b.pdf}
2614:   \caption{Topological structure of the continuous transition
2615: of (\ref{5.1}) when $k$=odd and $\alpha <0$: (a)
2616: $\lambda\leq\lambda_0$; (b) $\lambda >\lambda_0$.}\la{f5.6}
2617:  \end{figure}
2618:  \begin{figure}%[hbtp]
2619:   \centering
2620:   \includegraphics[width=0.32\textwidth]{5-7a.pdf}
2621:    \includegraphics[width=0.22\textwidth]{5-7b.pdf} 
2622:    \includegraphics[width=0.32\textwidth]{5-7c.pdf}
2623:   \caption{Topological structure of the mixing transition of
2624: (\ref{5.1}) when $k$=even and $\alpha\neq 0$: (a) $\lambda
2625: <\lambda_0$; (b) $\lambda =\lambda_0$; (c) $\lambda >\lambda_0$. Here
2626: $U^{\lambda}_1$ is the unstable domain, and $U^{\lambda}_2$ the
2627: stable domain.}\la{f5.7}
2628:  \end{figure}
2629: 
2630: \bt\la{t5.8}
2631: Assume  (\ref{5.4})  and (\ref{5.5}) with $m=1$, and (\ref{5.36}).  If $k$=odd and $\alpha\neq 0$ in (\ref{5.36}) then
2632: the following assertions hold true:
2633: 
2634: \begin{itemize}
2635: 
2636: \item[(1)] If $\alpha >0$,  then (\ref{5.1}) has a jump
2637: transition from $(0,\lambda_0)$, and bifurcates on $\lambda
2638: <\lambda_0$ to exactly two saddle points $v^{\lambda}_1$ and
2639: $v^{\lambda}_2$ with the Morse index one, as shown in Figure \ref{f5.5}.
2640: 
2641: \item[(2)] If $\alpha <0$,  then (\ref{5.1}) has a continuous
2642: transition from $(0,\lambda_0)$, which is an attractor bifurcation
2643:  as shown in Figure \ref{f5.6}. 
2644: 
2645: \item[(3)] The bifurcated singular points $v^{\lambda}_1$ and $v^{\lambda}_2$ 
2646: in the above cases can
2647: be expressed in the following form
2648: $$v^{\lambda}_{1,2}=\pm |\beta_1(\lambda )/\alpha
2649: |^{{1}/{k-1}}e_1(\lambda )+o(|\beta_1|^{{1}/{k-1}}).$$
2650: 
2651: \end{itemize}
2652: \et
2653: 
2654: 
2655: 
2656: \bt\la{t5.9}
2657:  Assume  (\ref{5.4})  and (\ref{5.5}) with $m=1$,  and
2658: (\ref{5.36}). If $k$=even and $\alpha\neq 0$, then we have the
2659: following assertions:
2660: 
2661: \begin{enumerate}
2662: 
2663: \item (\ref{5.1}) has a mixed transition from
2664: $(0,\lambda_0)$. More precisely, there exists a neighborhood
2665: $U\subset X$ of $u=0$ such that $U$ is separated into two disjoint
2666: open sets $U^{\lambda}_1$ and $U^{\lambda}_2$ by the stable
2667: manifold $\Gamma_{\lambda}$ of $u=0$ satisfying the following properties:
2668: 
2669: \begin{enumerate}
2670: 
2671: \item $U=U^{\lambda}_1+U^{\lambda}_2+\Gamma_{\lambda}$,
2672: 
2673: \item the transition in $U^{\lambda}_1$ is jump, and 
2674: 
2675: \item the transition in $U^{\lambda}_2$ is
2676: continuous. The local transition structure is as shown in Figure \ref{f5.7}.
2677: 
2678: \end{enumerate}
2679: 
2680: \item (\ref{5.1}) bifurcates in $U^{\lambda}_2$ to a unique
2681: singular point $v^{\lambda}$ on $\lambda >\lambda_0$, which is an
2682: attractor such that for any $\varphi\in U^{\lambda}_2$, 
2683: $$\lim\limits_{t\rightarrow\infty}\|u(t,\varphi
2684: )-v^{\lambda}\|_X=0,$$
2685: where $u(t,\varphi )$ is the solution of (\ref{5.1}). 
2686: 
2687: \item (\ref{5.1})\ bifurcates on $\lambda <\lambda_0$ to a unique saddle
2688: point $v^{\lambda}$ with the Morse index one. 
2689: 
2690: \item The bifurcated singular point $v^{\lambda}$ can be expressed as
2691: $$v^{\lambda}=-(\beta_1(\lambda )/\alpha
2692: )^{{1}/{(k-1)}}e_1+o(|\beta_1|^{{1}/{(k-1)}}).$$
2693: \end{enumerate}
2694: \et
2695: 
2696: \subsection{Singular Separation}
2697: \la{s6.2}
2698: In this section, we study  an important
2699: problem associated with the discontinuous transition of
2700: (\ref{5.1}), which we call  the singular separation.
2701: 
2702: \bd\la{d6.1}
2703: \begin{enumerate}
2704: 
2705: \item 
2706: An invariant set $\Sigma$ of (\ref{5.1}) is called a singular
2707: element if $\Sigma$ is either a singular point or a periodic
2708: orbit. 
2709: 
2710: \item Let $\Sigma_1\subset X$ be a singular
2711: element of (\ref{5.1}) and $U\subset X$ a neighborhood of
2712: $\Sigma_1$. We say that (\ref{5.1}) has a singular separation of
2713: $\Sigma$ at $\lambda =\lambda_1$ if 
2714: 
2715: \begin{enumerate}
2716: 
2717: \item (\ref{5.1}) has no singular
2718: elements in $U$ as $\lambda <\lambda_1$ (or $\lambda >\lambda_1$),
2719: and generates a singular element $\Sigma_1\subset  U$ at $\lambda
2720: =\lambda_1$,  and 
2721: 
2722: \item there are branches of singular elements
2723: $\Sigma_{\lambda}$, which are  separated from $\Sigma_1$ for $\lambda
2724: >\lambda_1$ (or $\lambda <\lambda_1$), i.e.,
2725: $$\lim\limits_{\lambda\rightarrow\lambda_1}\max_{x\in\Sigma_{\lambda}}\text{dist}(x,\Sigma_1)=0.$$
2726: \end{enumerate}
2727: 
2728: \end{enumerate}
2729: \ed
2730: 
2731: A special case of singular separation is the saddle-node
2732: bifurcation defined as follows.
2733: 
2734: \bd\la{d6.2}
2735: Let $u_1\in X$ be a singular point of
2736: (\ref{5.1}) at $\lambda =\lambda_1$ with $u_1\neq 0$. We say that
2737: (\ref{5.1}) has a saddle-node bifurcation at $(u_1,\lambda_1)$ if
2738: 
2739: \begin{enumerate}
2740: 
2741: \item the index of $L_{\lambda}+G$ at $(u_1,\lambda_1)$ is zero, i.e.,
2742: ind$(-(L_{\lambda_1}+G),u_1)=0$, 
2743: 
2744: \item  there are two branches
2745: $\Gamma_1(\lambda )$ and $\Gamma_2(\lambda )$ of singular points
2746: of (\ref{5.1}), which  are separated from $u_1$ for $\lambda >\lambda_1$
2747: (or $\lambda <\lambda_1)$,  i.e.,  for any
2748: $u_{\lambda}\in\Gamma_i(\lambda )$ $(i=1,2)$ we have
2749: $$u_{\lambda}\rightarrow u_1\ \text{in}\ X\ \ \ \ \text{as}\
2750: \lambda\rightarrow\lambda_1, $$ 
2751: and
2752: 
2753: \item  the indices of
2754: $u^i_{\lambda}\in\Gamma_i(\lambda )$ are as follows
2755: $$\text{ind}(-(L_{\lambda}+G),u_{\lambda})=
2756: \left\{
2757: \begin{aligned}
2758: &  1  &&  \text{ if } u_{\lambda}\in\Gamma_2(\lambda ),\\
2759: &  -1  &&  \text{ if } u_{\lambda}\in\Gamma_1(\lambda ).
2760: \end{aligned}
2761: \right.$$
2762: \end{enumerate}
2763: \ed
2764: 
2765: 
2766: Intuitively, the saddle-node bifurcation is
2767: schematically shown as in Figure~\ref{f6.1}, where the singular points in
2768: $\Gamma_1(\lambda )$ are saddle points and in $\Gamma_2(\lambda )$
2769: are nodes, and the singular separation of periodic orbits is as in
2770: shown Figure~\ref{f6.2}.
2771: \begin{figure}[hbt]
2772:   \centering
2773:   \includegraphics[width=0.35\textwidth]{6-1.pdf}
2774:   \caption{Saddle-node bifurcation.}\la{f6.1}
2775:  \end{figure}
2776: \begin{figure}[hbt]
2777:   \centering
2778:   \includegraphics[width=0.35\textwidth]{6-2.pdf}
2779:   \caption{Singular separation of periodic orbits.}\la{f6.2}
2780:  \end{figure}
2781: 
2782: 
2783: For the singular separation we can give a general principle as
2784: follows, which provides a basis for singular separation theory.
2785: 
2786: \bt\la{t6.4}
2787: Let the conditions (\ref{5.4}) and
2788: (\ref{5.5}) hold true. Then we have the following assertions.
2789: 
2790: \begin{enumerate}
2791: 
2792: \item[(1)] If (\ref{5.1}) bifurcates from $(u,\lambda
2793: )=(0,\lambda_0)$   to a branch $\Sigma_{\lambda}$ of singular elements on
2794: $\lambda <\lambda_0$ which is bounded in $X\times (-\infty
2795: ,\lambda_0)$ then (\ref{5.1}) has a singular separation of
2796: singular elements at some $(\Sigma_0,\lambda_1)\subset X\times
2797: (-\infty ,\lambda_0)$. 
2798: 
2799: \item[(2)] If the bifurcated branch
2800: $\Sigma_{\lambda}$ consists of singular points which has index $-1$,
2801: i.e., 
2802: $$\text{ind}(-(L_{\lambda}+G),u_{\lambda})=-1 \ \ \ \ \forall u_{\lambda}\in
2803: E_{\lambda},\ \ \ \ \lambda <\lambda_0,$$ then the singular
2804: separation is a saddle-node bifurcation from some
2805: $(u_1,\lambda_1)\in X\times (-\infty ,\lambda_0).$
2806: \end{enumerate}
2807: \et
2808: 
2809: 
2810: We consider the equation (\ref{5.1}) defined on the Hilbert spaces
2811: $X=H, X_1=H_1$. Let $L_{\lambda}=-A+\lambda B$. For $L_{\lambda}$
2812: and $G(\cdot ,\lambda ):H_1\rightarrow H$, we assume that
2813: $A:H_1\rightarrow H$ is symmetric, and
2814: \begin{eqnarray}
2815: &&<Au,u>_H\geq c\|u\|^2_{H_{{1}/{2}}},\label{6.46}\\
2816: &&<Bu,u>_H\geq c\|u\|^2_H,\label{6.47}\\
2817: &&<Gu,u>_H\leq -c_1\|u\|^p_H+c_2\|u\|^2_H,\label{6.48}
2818: \end{eqnarray}
2819: where $p>2, c, c_1, c_2>0$ are constants.
2820: 
2821: \bt\la{t6.5}
2822:  Assume the conditions (\ref{5.3}),
2823: (\ref{5.4}) and (\ref{6.46})-(\ref{6.48}), then (\ref{5.1}) has a
2824: transition at $(u,\lambda )=(0,\lambda_0)$, and the following
2825: assertions hold true:
2826: 
2827: \begin{enumerate}
2828: \item[(1)] If $u=0$ is an even-order nondegenerate singular point
2829: of $L_{\lambda}+G$ at $\lambda =\lambda_0$, then (\ref{5.1}) has a
2830: singular separation of singular points at some $(u_1,\lambda_1)\in
2831: H\times (-\infty ,\lambda_0)$. 
2832: 
2833: \item[(2)]  If $m=1$ and $G$
2834: satisfies (\ref{5.36}) with $\alpha >0$ if $k$=odd and $\alpha\neq
2835: 0$ if $k$=even, then (\ref{5.1}) has a saddle-node bifurcation at
2836: some singular point $(u_1,\lambda_1)$ with $\lambda_1<\lambda_0$.
2837: \end{enumerate}
2838: \et
2839: 
2840: \subsection{Transition and Singular Separation of Perturbed Systems}
2841: We consider the following perturbed equation of (\ref{5.1}): 
2842: \begin{equation}
2843: \frac{du}{dt}=(L_{\lambda}+S^{\varepsilon}_{\lambda})u+G(u,\lambda
2844: )+T^{\varepsilon}(u,\lambda ),\label{6.56}
2845: \end{equation}
2846: where $L_{\lambda}$ and $G_{\lambda}$ are as in (\ref{5.1}),
2847: $S^{\varepsilon}_{\lambda}:X_{\sigma}\rightarrow X$ is a linear
2848: perturbed operator,
2849: $T^{\varepsilon}_{\lambda}:X_{\sigma}\rightarrow X$ a $C^1$
2850: nonlinear perturbed operator,  and  $X_{\sigma}$ the fractional order
2851: space, $0\leq\sigma <1$. Also assume that 
2852: $G_{\lambda},T^{\varepsilon}_{\lambda}$ are
2853: $C^3$ on $u$, and
2854: \begin{equation}
2855: \left.
2856: \begin{aligned} 
2857: & \|S^{\varepsilon}_{\lambda}\|<\varepsilon,\\
2858: & \|T^{\varepsilon}_{\lambda}\|<\varepsilon ,\\
2859: & T^{\varepsilon}(u,\lambda )=o(\|u\|_{X_{\alpha}}).
2860: \end{aligned}
2861: \right.\label{6.57}
2862: \end{equation}
2863: 
2864: Let  (\ref{5.4})  and (\ref{5.5}) with $m=1$ hold true, $G(u,\lambda )=G_2(u,\lambda
2865: )+o(\|u\|^2_{X_1})$, where $G_2(\cdot ,\lambda )$ is a bilinear
2866: operator, and
2867: \begin{equation}
2868: b=<G_2(e,\lambda_0),e^*>\neq 0,\label{6.65}
2869: \end{equation}
2870: where $e\in X$ and $e^*\in X^*$ are the eigenvectors of
2871: $L_{\lambda}$ and $L^*_{\lambda}$ corresponding to
2872: $\beta_1(\lambda )$ at $\lambda =\lambda_0$ respectively.
2873: 
2874: We now consider the transition associated with the saddle-node
2875: bifurcation of the perturbed system (\ref{6.56}). Let $h(x,\lambda
2876: )$ be the center manifold function of (\ref{5.1}) near $\lambda
2877: =\lambda_0$. Assume that
2878: \begin{equation}
2879: <G(xe+h(x,\lambda_0),\lambda_0),e^*>=b_1x^3+o(|x|^3),\label{6.66}
2880: \end{equation}
2881: where $b_1\neq 0$, and $e$ and $e^*$ are as in (\ref{6.65}).
2882: 
2883: Then we have the following theorems.
2884: 
2885: \bt\la{t6.10}
2886:  Let the conditions (\ref{5.4})  and (\ref{5.5}) with $m=1$, and
2887: (\ref{6.66}) hold true, and $b_1<0$. Then there is an
2888: $\varepsilon >0$ such that if  $S^{\varepsilon}_{\lambda}$ and
2889: $T^{\varepsilon}_{\lambda}$ satisfy (\ref{6.57}), then the transition of
2890: (\ref{6.56}) is either continuous or mixed. If the transition is
2891: continuous, then Assertions (2) and (3) of Theorem \ref{t5.8} are valid
2892: for (\ref{6.56}). If the transition is mixed, then the
2893: following assertions hold true:
2894: 
2895: \begin{enumerate}
2896: 
2897: \item[(1)] (\ref{6.56}) has a saddle-node bifurcation at some
2898: point $(u_1,\lambda_1)\in X\times (-\infty
2899: ,\lambda^{\varepsilon}_0)$, and there are exactly two branches
2900: $$\Gamma^{\lambda}_i=\{(u^{\lambda}_i,\lambda )|\ \lambda_1<\lambda
2901: <\lambda^{\varepsilon}_0+\delta\} \qquad   i=1,2, $$ 
2902: separated from
2903: $(u_1,\lambda_1)$ as shown in Figure~\ref{f6.1}, which satisfy that 
2904: \begin{align*}
2905: &  \|u^{\lambda}_2\|_X\neq 0  &&\forall (u^{\lambda}_2,\lambda
2906: )\in\Gamma^{\lambda}_2,\ \ \ \ \lambda_1<\lambda
2907: <\lambda^{\varepsilon}_0+\delta ,\\
2908: & \lim_{\lambda\rightarrow\lambda^{\varepsilon}_0}\|u^{\lambda}_1\|_X=0 &&
2909: \forall  (u^{\lambda}_1,\lambda )\in\Gamma^{\lambda}_1.
2910: \end{align*}
2911: 
2912: 
2913: \item[(2)] There is a neighborhood $U\subset X$ of
2914: $u=0$, such that for each $\lambda$ with $\lambda_1<\lambda
2915: <\lambda^{\varepsilon}_0+\delta$ and
2916: $\lambda\neq\lambda^{\varepsilon}_0, U$ contains only two
2917: nontrivial singular points $u^{\lambda}_1$ and $u^{\lambda}_2$ of
2918: (\ref{6.56}). 
2919: 
2920: \item[(3)] For each $\lambda_1<\lambda
2921: <\lambda^{\varepsilon}_0+\delta , U$ can be decomposed into two
2922: open sets $\bar{U}=\bar{U}^{\lambda}_1+\bar{U}^{\lambda}_2$ with
2923: $U^{\lambda}_1\cap U^{\lambda}_2=\emptyset$, such that
2924: 
2925: \begin{enumerate}
2926: 
2927: \item if $  \lambda_1<\lambda <\lambda^{\varepsilon}_0$, 
2928: $$0\in U^{\lambda}_1,\ \ \ \ u^{\lambda}_2\in U^{\lambda}_2,\ \ \
2929: \ u^{\lambda}_1\in\partial U^{\lambda}_1\cap\partial
2930: U^{\lambda}_2,$$ with $u=0$ and $u^{\lambda}_2$ being
2931: attractors which attract $U^{\lambda}_1$ and $U^{\lambda}_2$
2932: respectively, and
2933: 
2934: \item if  $ \lambda^{\varepsilon}_0<\lambda <\lambda^{\varepsilon}_0+\delta$,
2935: $$u^{\lambda}_1\in U^{\lambda}_1,\ \ \ \ u^{\lambda}_2\in
2936: U^{\lambda}_2,\ \ \ \ 0\in\partial U^{\lambda}_1\cap\partial
2937: U^{\lambda}_2, $$ 
2938: with $u^{\lambda}_1$ and
2939: $u^{\lambda}_2$ being attractors which attract $U^{\lambda}_1$ and
2940: $U^{\lambda}_2$ respectively. 
2941: 
2942: \end{enumerate}
2943: 
2944: \item[(4)] Near $(u,\lambda)=(0,\lambda^{\varepsilon}_0),u^{\lambda}_1$ 
2945: and $u^{\lambda}_2$
2946: can be expressed as 
2947: \begin{equation} 
2948: \left.
2949: \begin{aligned}
2950: & u^{\lambda}_1=\alpha_1(\lambda ,\varepsilon )e+o(|\alpha_1|),\\
2951: & u^{\lambda}_2=\alpha_2(\lambda ,\varepsilon )e+o(|\alpha_2|),\\
2952: & \lim_{  \lambda\rightarrow\lambda^{\varepsilon}_0  }  
2953: \alpha_1(\lambda ,\varepsilon )= 0,\\
2954: & \alpha_2(\lambda^{\varepsilon}_0,\varepsilon )\neq 0,
2955: \end{aligned}
2956: \right.\label{6.67}
2957: \end{equation}
2958: where $e$ is as in (\ref{6.66}).
2959: \end{enumerate}
2960: \et
2961: 
2962: 
2963: \bt\la{t6.11} Assume the conditions (\ref{5.4})  and (\ref{5.5}) with $m=1$, and
2964: (\ref{6.66}) with $b_1>0$. Then, there is an $\varepsilon >0$ such
2965: that when $S^{\varepsilon}_{\lambda}$ and
2966: $T^{\varepsilon}_{\lambda}$ satisfy (\ref{6.57}), the transition of
2967: (\ref{6.56}) is either jump or mixed. If it is jump transition,  then
2968: Assertions (1) and  (3) of Theorem \ref{t5.8} are valid for (\ref{6.56}). 
2969: If it is mixed, then the following assertions hold true:
2970: 
2971: \begin{enumerate}
2972: 
2973: \item[(1)] (\ref{6.56}) has a saddle-node bifurcation at some
2974: point $(u_1,\lambda_1)\in X\times (\lambda^{\varepsilon}_0,+\infty
2975: )$, and there are exactly two branches
2976: $$\Gamma^{\lambda}_i=\{(u^{\lambda}_i,\lambda )|\
2977: \lambda^{\varepsilon}_0-\delta <\lambda <\lambda_1\} \qquad   (i=1,2), $$
2978: separated from $(u_1,\lambda_1)$, which satisfy
2979: \begin{align*}
2980: &  \|u^{\lambda}_2\|_X=0 &&\forall (u^{\lambda}_2,\lambda)\in\Gamma^*_2,\ \ \ \ \lambda^{\varepsilon}_0-\varepsilon
2981: <\lambda <\lambda_1,\\
2982: & \lim\limits_{\lambda\rightarrow\lambda^{\varepsilon}_0}\|u^{\lambda}_1\|_X=0
2983: && \forall 
2984: (u^{\lambda}_1,\lambda )\in\Gamma^{\lambda}_1.
2985: \end{align*}
2986: 
2987: 
2988: \item[(2)] There is a neighborhood $U\subset X$ of
2989: $u=0$, such that for each $\lambda$ with
2990: $\lambda^{\varepsilon}_0-\delta <\lambda <\lambda_1, U$ contains
2991: only two nontrivial singular points $u^{\lambda}_1$ and
2992: $u^{\lambda}_2$ of (\ref{6.56}). 
2993: 
2994: 
2995: \item[(3)] For every
2996: $\lambda^{\varepsilon}_0-\delta <\lambda <\lambda_1, U$ can be
2997: decomposed into three open sets
2998: $\bar{U}=\bar{U}_0+\bar{U}_1+\bar{U}_2$ with $U_i\cap
2999: U_j=\emptyset$ $ (i\neq j)$ such that
3000: 
3001: \begin{enumerate}
3002: 
3003: \item  if   $ \lambda^{\varepsilon}_0-\delta <\lambda <\lambda^{\varepsilon}_0$, 
3004: then 
3005: $$u=0\in U^{\lambda}_0,\ \ \ \ u^{\lambda}_i\in\partial
3006: U^{\lambda}_i\cap\partial U^{\lambda}_0(i=1,2),$$ 
3007: with $u=0$ being an attractor which
3008: attracts $U^{\lambda}_0$ and $U^{\lambda}_i(i=1,2)$ two saddle
3009: points with the Morse index one, and
3010: 
3011: \item  if   $ \lambda^{\varepsilon}_0<\lambda <\lambda_1$,  then 
3012: $$u^{\lambda}_1\in U^{\lambda}_1,\ \ \ \ u^{\lambda}_2\in\partial
3013: U^{\lambda}_2\cap\partial U^{\lambda}_1,\ \ \ \ 0\in\partial
3014: U^{\lambda}_1\cap\partial U^{\lambda}_0,$$ 
3015: with $u^{\lambda}_1$
3016: being an attractor which attracts $U^{\lambda}_1$ and
3017: $u^{\lambda}_2$ and $u=0$ being saddle points with the Morse index
3018: one. 
3019: 
3020: \end{enumerate}
3021: 
3022: \item[(4)] Near $(0,\lambda^0_{\varepsilon}),u^{\lambda}_1$
3023: and $u^{\lambda}_2$ can be expressed by (\ref{6.67}).
3024: \end{enumerate}
3025: \et
3026: 
3027: \section{Ginzburg-Landau Models}
3028: %\subsection{Thermodynamic potentials}
3029: \label{s7.2.2}
3030: In this subsection, we introduce the time-dependent Ginzburg-Landau model for equilibrium phase transitions. 
3031: 
3032: We start with thermodynamic potentials and the Ginzburg-Landau free energy. As we know, four thermodynamic potentials-- internal energy, the enthalpy, the Helmholtz free energy and the Gibbs free energy--are useful in the chemical thermodynamics of reactions and non-cyclic processes. 
3033: 
3034: Consider a thermal system, its order parameter $u$ changes in
3035: $\Omega\subset \R^n$ $(1\leq n\leq 3)$. In this situation,  the free energy of this system is of the form
3036: %\begin{widetext}
3037: \begin{equation}
3038: {\mathcal{H}}(u,\lambda
3039: )= {\mathcal{H}}_0 
3040:  +\int_{\Omega}\Big[\frac{1}{2}\sum^m_{i=1}\mu_i|\nabla
3041: u_i|^2  +g(u,\nabla u,\lambda )\Big]dx\label{7.28}
3042: \end{equation}
3043: %\end{widetext}
3044: where $N\geq 3$ is an integer,
3045: $u=(u_1,\cdots,u_m),\mu_i=\mu_i(\lambda )>0$, and $g(u,\nabla
3046: u,\lambda )$ is a $C^r(r\geq 2)$ function of $(u,\nabla u)$ with
3047: the Taylor expansion \begin{equation} g(u,\nabla u,\lambda
3048: )=\sum\alpha_{ijk}u_iD_ju_k+\sum^N_{|I|=1}\alpha_Iu^I+o(|u|^N)-fX,\label{7.29}
3049: \end{equation}
3050: where $I=(i_1,\cdots,i_m),i_k\geq 0$ are integer,
3051: $|I|=\sum^m_{k=1}i_k$, the coefficients $\alpha_{ijk}$ and
3052: $\alpha_I$ continuously depend on $\lambda$, which are determined
3053: by the concrete physical problem, $u^I=u^{i_1}_1\cdots u^{im}_m$
3054: and $fX$ the generalized work.
3055: 
3056: A thermal system is controlled by some parameter $\lambda$. When
3057: $\lambda$ is far from the critical point $\lambda_0$ the system
3058: lies on a stable equilibrium state $\Sigma_1$, and when $\lambda$
3059: reaches or exceeds $\lambda_0$ the state $\Sigma_1$ becomes unstable, and 
3060: meanwhile the system will undergo a transition  from $\Sigma_1$ to another stable
3061: state $\Sigma_2$. The basic principle is that there often exists  fluctuations in the system leading  to a deviation from the equilibrium states, and  the phase transition process is a
3062: dynamical behavior, which should be described by a time-dependent
3063: equation.
3064: 
3065: To derive a general time-dependent model, first we recall that  the classical  
3066: le Ch\^atelier  principle amounts to saying that 
3067: for a stable  equilibrium state of a system $\Sigma$, when the system deviates from
3068: $\Sigma$ by a small perturbation or fluctuation, there will be a
3069: resuming force to restore this system to return to the stable state
3070: $\Sigma$.
3071: Second, we know that  a stable equilibrium state of a thermal system must
3072: be the minimal value point of the thermodynamic potential. 
3073: 
3074: By the mathematical characterization of gradient systems and the le Ch\^atelier principle, for a system with
3075: thermodynamic potential ${\mathcal{H}}(u,\lambda )$, the governing
3076: equations are essentially determined by the functional
3077: ${\mathcal{H}}(u,\lambda )$.
3078: When the order parameters $(u_1,\cdots,u_m)$ are nonconserved
3079: variables, i.e., the integers
3080: $$\int_{\Omega}u_i(x,t)dx=a_i(t)\neq\text{constant}.$$
3081: then the time-dependent equations are given by
3082: \begin{equation}
3083: \left.
3084: \begin{aligned} 
3085: &\frac{\partial u_i}{\partial
3086: t}=-\beta_i\frac{\delta}{\delta u_i}{\mathcal{H}}(u,\lambda
3087: )+\Phi_i(u,\nabla u,\lambda ),\\
3088: &\frac{\partial u}{\partial n}|_{\partial\Omega}=0\ \ \ \
3089: (\text{or}\ u|_{\partial\Omega}=0),\\
3090: &u(x,0)=\varphi (x),
3091: \end{aligned}
3092: \right.\label{7.30}
3093: \end{equation}
3094: for any $1 \le i \le m$, 
3095: where $\delta /\delta u_i$ are the variational derivative,
3096: $\beta_i>0$ and $\Phi_i$ satisfy
3097: \begin{equation}
3098: \int_{\Omega}\sum_i\Phi_i\frac{\delta}{\delta
3099: u_i}{\mathcal{H}}(u,\lambda )dx=0.\label{7.31}
3100: \end{equation}
3101: The condition (\ref{7.31})  is  required by
3102: the Le Ch\^atelier principle. In the concrete problem, the terms
3103: $\Phi_i$ can be determined by physical laws and (\ref{7.31}).
3104: 
3105: When the order parameters are the number density and the system
3106: has no material exchange with the external, then $u_j$  $(1\leq j\leq
3107: m)$ are conserved, i.e.,
3108: \begin{equation}
3109: \int_{\Omega}u_j(x,t)dx=\text{constant}.\label{7.32}
3110: \end{equation}
3111: This conservation law requires a continuous equation
3112: \begin{equation}
3113: \frac{\partial u_j}{\partial t}=-\nabla\cdot J_j(u,\lambda
3114: ),\label{7.33}
3115: \end{equation}
3116: where $J_j(u,\lambda )$ is the flux of component $u_j$. In
3117: addition, $J_j$ satisfy
3118: \begin{equation}
3119: J_j=-k_j\nabla (\mu_j-\sum_{i\neq j}\mu_i),\label{7.34}
3120: \end{equation}
3121: where $\mu_l$ is the chemical potential of component $u_l$, 
3122: \begin{equation}
3123: \mu_j-\sum_{i\neq j}\mu_i=\frac{\delta}{\delta
3124: u_j}{\mathcal{H}}(u,\lambda )-\phi_j(u,\nabla u,\lambda
3125: ), \label{7.35}
3126: \end{equation}
3127: and  $\phi_j(u,\lambda )$ is a function depending on the other
3128: components $u_i$ $(i\neq j)$. When $m=1$, i.e., the system consists of
3129: two components $A$ and $B$, this term $\phi_j=0$. Thus, from
3130: (\ref{7.33})-(\ref{7.35}) we obtain the dynamical equations as
3131: follows
3132: \begin{equation}
3133: \left.
3134: \begin{aligned} &\frac{\partial u_j}{\partial
3135: t}=\beta_j\Delta\left[\frac{\delta}{\delta
3136: u_j}{\mathcal{H}}(u,\lambda )-\phi_j(u,\nabla u,\lambda )\right],\\
3137: &\frac{\partial u}{\partial n}|_{\partial\Omega}=0,\ \ \ \
3138: \frac{\partial\Delta u}{\partial n}|_{\partial\Omega}=0,\\
3139: &u(x,0)=\varphi (x),
3140: \end{aligned}
3141: \right.\label{7.36}
3142: \end{equation}
3143: for $1 \le j \le m$, 
3144: where $\beta_j>0$ are constants, $\phi_j$ satisfy
3145: \begin{equation}
3146: \int_{\Omega}\sum_j\Delta\phi_j\cdot\frac{\delta}{\delta
3147: u_j}{\mathcal{H}}(u,\lambda )dx=0.\label{7.37}
3148: \end{equation}
3149: 
3150: If the order parameters $(u_1,\cdots,u_k)$ are coupled to the
3151: conserved variables $(u_{k+1},\cdots,u_m)$, then the dynamical
3152: equations are
3153: \begin{equation}
3154: \left.
3155: \begin{aligned} 
3156: &\frac{\partial u_i}{\partial t}
3157:    =-\beta_i\frac{\delta}{\delta u_i}{\mathcal{H}}(u,\lambda)+\Phi_i(u,\nabla u,\lambda ),\\
3158: & \frac{\partial u_j}{\partial t}
3159:   =\beta_j\Delta\left[\frac{\delta}{\delta u_j}{\mathcal{H}}(u,\lambda )
3160:     -\phi_j(u,\nabla u,\lambda )\right],\\
3161: &\frac{\partial u_i}{\partial n}|_{\partial\Omega}=0\ \ \ \
3162: (\text{or}\ u_i|_{\partial\Omega}=0),\\
3163: &\frac{\partial u_j}{\partial n}|_{\partial\Omega}=0,\ \ \ \
3164: \frac{\partial\Delta u_j}{\partial n}|_{\partial\Omega}=0,\\
3165: &u(x,0)=\varphi (x).
3166: \end{aligned}
3167: \right.\label{7.38}
3168: \end{equation}
3169: for $1 \le i \le k$  and $k+1 \le j \le m$.
3170: 
3171: The model (\ref{7.38}) gives a general form of the governing
3172: equations to thermodynamic phase transitions.  Hence, the dynamics of
3173: equilibrium phase transition in statistic physics is based on the new Ginzburg-Landau formulation  (\ref{7.38}).
3174: 
3175: Physically, the initial value condition $u(0)=\varphi$ in
3176: (\ref{7.38}) stands for the fluctuation of system or perturbation
3177: from the external. Hence, $\varphi$ is generally small. However,
3178: we can not exclude the possibility of a bigger noise $\varphi$.
3179: 
3180: \bibliographystyle{siam}
3181: \def\cprime{$'$}
3182: \begin{thebibliography}{1}
3183: 
3184: \bibitem{CH57}
3185: {\sc J.~Cahn and J.~E. Hilliard}, {\em Free energy of a nonuniform system I.
3186:   interfacial energy}, J. Chemical Physics, 28 (1957), pp.~258--267.
3187: 
3188: \bibitem{langer71}
3189: {\sc J.~Langer}, {\em Theory of spinodal decomposition in allays}, Ann. of
3190:   Physics, 65 (1971), pp.~53--86.
3191: 
3192: \bibitem{b-book}
3193: {\sc T.~Ma and S.~Wang}, {\em Bifurcation theory and applications}, vol.~53 of
3194:   World Scientific Series on Nonlinear Science. Series A: Monographs and
3195:   Treatises, World Scientific Publishing Co. Pte. Ltd., Hackensack, NJ, 2005.
3196: 
3197: \bibitem{chinese-book}
3198: \leavevmode\vrule height 2pt depth -1.6pt width 23pt, {\em Stability and
3199:   Bifurcation of Nonlinear Evolutions Equations}, Science Press (in Chinese),
3200:   Beijing, 2007.
3201: 
3202: \bibitem{NS84}
3203: {\sc A.~Novick-Cohen and L.~A. Segel}, {\em Nonlinear aspects of the
3204:   {C}ahn-{H}illiard equation}, Phys. D, 10 (1984), pp.~277--298.
3205: 
3206: \bibitem{onuki}
3207: {\sc O.~Onuki}, {\em Phase transition dynamics}, Combridge Univ. Press.,
3208:   (2002).
3209: 
3210: \bibitem{reichl}
3211: {\sc L.~E. Reichl}, {\em A modern course in statistical physics}, A
3212:   Wiley-Interscience Publication, John Wiley \& Sons Inc., New York,
3213:   second~ed., 1998.
3214: 
3215: \end{thebibliography}
3216: 
3217: \end{document}
3218: