nlin0611021/run.tex
1: \documentclass{conm-p-l}
2: 
3: \copyrightinfo{2006}{}
4: 
5: \setcounter{page}{1}
6: 
7: \usepackage{graphicx}
8: 
9: 
10: \newtheorem{theorem}{Theorem}[section]
11: \newtheorem{lemma}[theorem]{Lemma}
12: \newtheorem{corollary}[theorem]{Corollary}
13: 
14: \theoremstyle{definition}
15: \newtheorem{definition}[theorem]{Definition}
16: \newtheorem{example}[theorem]{Example}
17: \newtheorem{xca}[theorem]{Exercise}
18: 
19: \theoremstyle{remark}
20: \newtheorem{remark}[theorem]{Remark}
21: 
22: \numberwithin{equation}{section}
23: 
24: %    Absolute value notation
25: \newcommand{\abs}[1]{\lvert#1\rvert}
26: 
27: %    Blank box placeholder for figures (to avoid requiring any
28: %    particular graphics capabilities for printing this document).
29: \newcommand{\blankbox}[2]{%
30:   \parbox{\columnwidth}{\centering
31: %    Set fboxsep to 0 so that the actual size of the box will match the
32: %    given measurements more closely.
33:     \setlength{\fboxsep}{0pt}%
34:     \fbox{\raisebox{0pt}[#2]{\hspace{#1}}}%
35:   }%
36: }
37: 
38: \newcommand{\tla}{\tilde{\lambda}}
39: \newcommand{\tu}{\tilde{u}}
40: \newcommand{\im}{\mathop{\rm Im}\nolimits}
41: \newcommand{\hk}{\hat{k}}
42: \newcommand{\Z}{\mathbb{Z}/\{0\}}
43: \newcommand{\ZZ}{\mathbb{Z}^2/\{0\}}
44: \newcommand{\ZZZ}{\mathbb{Z}^3/\{0\}}
45: \newcommand{\E}{{\mathcal E}}
46: \newcommand{\HH}{{\mathcal H}}
47: \newcommand{\cS}{{\mathcal S}}
48: \newcommand{\q}{\vec{q}}
49: \newcommand{\vq}{\vec{q}}
50: \newcommand{\p}{\vec{p}}
51: \newcommand{\f}{\vec{f}}
52: \newcommand{\F}{{\mathcal F}}
53: \newcommand{\e}{\epsilon}
54: \newcommand{\ve}{\varepsilon}
55: \newcommand{\vth}{\vartheta}
56: \newcommand{\U}{{\mathcal U}}
57: \newcommand{\N}{{\mathcal N}}
58: \newcommand{\cq}{\tilde{q}}
59: \newcommand{\tDl}{\tilde{\Delta}}
60: \newcommand{\tdl}{\tilde{\delta}}
61: \newcommand{\tQ}{\tilde{Q}}
62: \newcommand{\vQ}{\vec{Q}}
63: \newcommand{\A}{{\mathcal A}}
64: \newcommand{\B}{{\mathcal B}}
65: \newcommand{\C}{{\mathcal C}}
66: \newcommand{\D}{{\mathcal D}}
67: \newcommand{\R}{{\mathcal R}}
68: \newcommand{\BD}{B\"acklund-Darboux transformations}
69: \renewcommand{\k}{\kappa}
70: \newcommand{\ga}{\gamma}
71: \newcommand{\Ga}{\Gamma}
72: \newcommand{\he}{\hat{e}}
73: \newcommand{\vv}{\vec{v}}
74: \newcommand{\hcS}{\hat{\cS}}
75: \newcommand{\tcS}{\tilde{\cS}}
76: \newcommand{\hS}{\hat{S}}
77: \newcommand{\tS}{\tilde{S}}
78: \newcommand{\hD}{\hat{D}}
79: \newcommand{\hJ}{\hat{J}}
80: \newcommand{\tD}{\tilde{D}}
81: \newcommand{\tC}{\tilde{C}}
82: \newcommand{\dl}{\delta}
83: \newcommand{\Dl}{\Delta}
84: \renewcommand{\th}{\theta}
85: \newcommand{\ra}{\rightarrow}
86: \newcommand{\lra}{\longrightarrow}
87: \newcommand{\al}{\alpha}
88: \newcommand{\be}{\beta}
89: \newcommand{\sg}{\sigma}
90: \newcommand{\Sg}{\Sigma}
91: \newcommand{\bM}{\bar{M}}
92: \newcommand{\pa}{\partial}
93: \newcommand{\z}{\zeta}
94: \newcommand{\hQ}{\hat{Q}}
95: \newcommand{\hq}{\hat{q}}
96: \newcommand{\hv}{\hat{v}}
97: \newcommand{\hw}{\hat{w}}
98: \newcommand{\hx}{\hat{x}}
99: \newcommand{\hy}{\hat{y}}
100: \newcommand{\hz}{\hat{z}}
101: \newcommand{\bv}{\bar{v}}
102: \newcommand{\bw}{\bar{w}}
103: \newcommand{\La}{\Lambda}
104: \newcommand{\tLa}{\tilde{\Lambda}}
105: \newcommand{\la}{\lambda}
106: \newcommand{\bq}{\bar{q}}
107: \newcommand{\bp}{\bar{p}}
108: \newcommand{\bQ}{\bar{Q}}
109: \newcommand{\bE}{\bar{E}}
110: \newcommand{\nid}{\noindent}
111: \newcommand{\cF}{{\mathcal F}}
112: \newcommand{\rc}{S_\omega}
113: \newcommand{\hrc}{\hat{S}_\omega}
114: \newcommand{\bW}{\bar{W}}
115: \newcommand{\hN}{\hat{N}}
116: \newcommand{\hF}{\hat{F}}
117: \newcommand{\tF}{\tilde{F}}
118: \newcommand{\tf}{\tilde{f}}
119: \newcommand{\om}{\omega}
120: \newcommand{\Om}{\Omega}
121: \newcommand{\na}{\nabla}
122: \newcommand{\lag}{\langle}
123: \newcommand{\rag}{\rangle}
124: \newcommand{\tx}{\tilde{x}}
125: \newcommand{\tq}{\tilde{q}}
126: \newcommand{\tom}{\tilde{\omega}}
127: \newcommand{\tE}{\tilde{E}}
128: \newcommand{\hE}{\hat{E}}
129: \newcommand{\cH}{{\mathcal H}}
130: \newcommand{\W}{{\mathcal W}}
131: \renewcommand{\O}{{\mathcal O}}
132: \newcommand{\ty}{\tilde{y}}
133: \newcommand{\tz}{\tilde{z}}
134: \newcommand{\ttau}{\tilde{\tau}}
135: \newcommand{\htau}{\hat{\tau}}
136: \newcommand{\hrho}{\hat{\rho}}
137: \newcommand{\hvth}{\hat{\vartheta}}
138: \newcommand{\vphi}{\varphi}
139: \newcommand{\tb}{\tilde{b}}
140: \newcommand{\bd}{\bar{d}}
141: \newcommand{\bb}{\bar{b}}
142: \newcommand{\ba}{\bar{a}}
143: \newcommand{\bc}{\bar{c}}
144: \newcommand{\LL}{{\mathcal L}}
145: \newcommand{\non}{\nonumber}
146: \newcommand{\tG}{\tilde{G}}
147: \newcommand{\hG}{\hat{G}}
148: \newcommand{\tT}{\tilde{T}}
149: \newcommand{\hLL}{\hat{\LL}}
150: \newcommand{\tk}{\tilde{k}}
151: 
152: %Commuting Diagram Need
153: \def\maprightu#1{\smash{
154:     \mathop{\longrightarrow}\limits^{#1}}}
155: \def\maprightd#1{\smash{
156:     \mathop{\longrightarrow}\limits_{#1}}}
157: \def\mapdownl#1{
158:     \llap{$\vcenter{\hbox{$\scriptstyle#1$}}$}\Big\downarrow}
159: \def\mapdownr#1{\Big\downarrow
160:     \rlap{$\vcenter{\hbox{$\scriptstyle#1$}}$}}
161: \def\mapupl#1{
162:     \llap{$\vcenter{\hbox{$\scriptstyle#1$}}$}\Big\uparrow}
163: \def\mapupr#1{\Big\uparrow
164:     \rlap{$\vcenter{\hbox{$\scriptstyle#1$}}$}}
165: 
166: 
167: 
168: \begin{document}
169: 
170: \title[Navier-Stokes and Euler Equations]
171: {On the Dynamics of Navier-Stokes and Euler Equations}
172: 
173: \author{Yueheng Lan}
174: \address{Department of Chemistry, University of North Carolina, 
175: Chapel Hill, NC 27599-3290 USA}
176: \curraddr{}
177: \email{ylan@email.unc.edu}
178: \thanks{}
179: 
180: \author{Y. Charles Li}
181: \address{Department of Mathematics, University of Missouri, 
182: Columbia, MO 65211, USA}
183: \curraddr{}
184: \email{cli@math.missouri.edu}
185: \thanks{}
186: 
187: 
188: \subjclass{Primary 35, 37; Secondary 34, 46}
189: \date{}
190: 
191: 
192: 
193: \dedicatory{}
194: 
195: 
196: 
197: \keywords{Heteroclinic orbit, chaos, turbulence, control, Melnikov integral, 
198: zero viscosity limit, sine-Gordon equation, Navier-Stokes equations, Euler equations}
199: 
200: 
201: \begin{abstract}
202: This is a rather comprehensive study on the dynamics of Navier-Stokes and 
203: Euler equations via a combination of analysis and numerics. We focus upon two 
204: main aspects: (a). zero viscosity limit of the spectra of linear Navier-Stokes operator,
205: (b). heteroclinics conjecture for Euler equation, its numerical verification, 
206: Melnikov integral, and simulation and control of chaos. Besides Navier-Stokes and 
207: Euler equations, we also study two models of them.   
208: \end{abstract}
209: 
210: \maketitle
211: 
212: \tableofcontents
213: 
214: %\section*{}
215: %This is an example of an unnumbered first-level heading.
216: 
217: %\specialsection*{This is a Special Section Head}
218: %This is an example of a special section head%
219: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
220: %\footnote{Here is an example of a footnote. Notice that this footnote
221: %text is running on so that it can stand as an example of how a footnote
222: %with separate paragraphs should be written.
223: %\par
224: %And here is the beginning of the second paragraph.}%
225: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
226: 
227: %\section{This is a numbered first-level section head}
228: %This is an example of a numbered first-level heading.
229: 
230: %\subsection{This is a numbered second-level section head}
231: %This is an example of a numbered second-level heading.
232: 
233: %\subsection*{This is an unnumbered second-level section head}
234: %This is an example of an unnumbered second-level heading.
235: 
236: %\subsubsection{This is a numbered third-level section head}
237: %This is an example of a numbered third-level heading.
238: 
239: %\subsubsection*{This is an unnumbered third-level section head}
240: %This is an example of an unnumbered third-level heading.
241: 
242: \section{Introduction}
243: 
244: The dynamics of Navier-Stokes and Euler equations is a challenging problem. In 
245: particular, such dynamics can be chaotic or turbulent. The main challenge 
246: comes from the large dimensionality of the phase space where the Navier-Stokes 
247: and Euler equations pose extremely intricate flows. Here the dynamics we refer to 
248: is the so-called Eulerian, in contrast to the so-called Lagrangian, dynamics of 
249: fluids. The Eulerian dynamics is a dynamics in an infinite dimensional phase space
250: (e.g. a Banach space) posed by the Cauchy problem of either the Navier-Stokes or the 
251: Euler equations as partial differential equations. The Lagrangian dynamics of 
252: fluid particles is a dynamics of a system of two or three ordinary differential 
253: equations with vector fields given by fluid velocities. The Lagrangian dynamics of 
254: fluid particles in three dimensions can be chaotic even when the Eulerian dynamics is 
255: steady (e.g. the ABC flow \cite{Dom86}). Nevertheless, as shown in Appendix B, 
256: the Lagrangian dynamics of 2D inviscid fluid particles is always integrable. 
257: 
258: Even though the global well-posedness of 3D Navier-Stokes and Euler equations 
259: is still an interesting open mathematical problem, 3D Navier-Stokes and Euler 
260: equations have local well-posedness which is often enough for a dynamical system 
261: study in the phase space. To begin such a dynamical system study, one needs to 
262: understand the spectra of the linear Navier-Stokes and/or Euler operators \cite{Li05}.
263: The spectra of the linear Navier-Stokes operators consist of eigenvalues, 
264: whereas the spectra of the linear Euler operators contain continuous spectra. 
265: Existence of invariant manifolds can be proved for Navier-Stokes equations \cite{Li05},
266: but is still open for Euler equations. The size of the invariant manifolds for 
267: Navier-Stokes equations tends to zero in the zero viscosity limit \cite{Li05}.
268: We find that 
269: the spectra of the linear Navier-Stokes and Euler operators can be classified into four 
270: categories in the zero viscosity limit: 
271: \begin{enumerate}
272: \item {\em Persistence:} These are the eigenvalues that persist and approach 
273: to the eigenvalues of the corresponding linear Euler operator when the viscosity 
274: approaches zero.
275: \item {\em Condensation:} These are the eigenvalues that approach and form 
276: a continuous spectrum for the corresponding linear Euler operator when the viscosity 
277: approaches zero.
278: \item {\em Singularity:} These are the eigenvalues that approach to a 
279: set that is not in the spectrum of the corresponding linear Euler operator when 
280: the viscosity approaches zero.
281: \item {\em Addition:} This is a subset of the spectrum of the linear Euler operator,
282: which has no overlap with the zero viscosity limit set of the spectrum of the linear 
283: NS operator.
284: \end{enumerate} 
285: We also find that as the viscosity approaches zero, the spectrum of the linear Navier-Stokes 
286: operator undergoes a fascinating deformation. Focusing upon the persistent unstable 
287: eigenvalue, we propose a heteroclinics conjecture, i.e. there should be a heteroclinic 
288: orbit (in fact heteroclinic cycles) associating to the instability for Euler equation.
289: We will present both analytical and numerical study upon this heteroclinics conjecture. 
290: Then we conduct a Melnikov integral calculation along the numerically obtained 
291: approximate heteroclinic orbit. We also compare the Melnikov prediction with the 
292: numerical simulation and control of chaos for the Navier-Stokes equations. Numerically
293: we mainly use the Liapunov exponent as a measure of chaos. In some case, we also 
294: plot the Poincar\'e return map. We realize that the size of Galerkin truncations for 
295: the full Navier-Stokes equations is limited by the computer ability. Thus we propose 
296: two simpler models of the Navier-Stokes equations. For the so-called line model, we 
297: obtain numerically exact heteroclinic orbits for any size of Galerkin truncations. 
298: We also realize that due to viscosity, the chaos in Galerkin truncations of 
299: Navier-Stokes equations is often 
300: transient chaos, i.e. the chaos has a finite life time. Infinite life time chaos can 
301: be observed in Galerkin truncations of Euler equations.
302: 
303: Chaos and turbulence have no good averages \cite{Li06i}. The matter is more fundamental 
304: than just poor understanding of averages. The very mechanism of chaos leads to the 
305: impossibility of a good average \cite{Li04}. On the other hand, chaos and turbulence are 
306: ubiquitous. In high dimensional systems, there exists tubular chaos \cite{Li04} 
307: \cite{Li03a} \cite{Li03b} \cite{Li04c} which further confirms that there is no good 
308: average. The hope is that chaos and turbulence can be controlled. Two aspects of control 
309: are practically important in applications: Taming and enhancing. When an airplane meets 
310: turbulence, it will be nice, safer and economic if we can tame the turbulence into a 
311: laminar flow or a less turbulent flow \cite{ABT01} \cite{Kim03} \cite{LB98}. In a 
312: combustor, enhancing turbulence can get the fuel and oxidant mixed and burned more 
313: efficiently \cite{LB98} \cite{Kim03}. Theoretically, one can also make use of the 
314: ergodicity of chaos to gear an orbit to a specific target \cite{OGY90}. Many other 
315: possibilities of applications of control can be designed too. An advantage of the 
316: control is that it can be done in a trial-correction manner without a detailed knowledge 
317: of turbulence.
318: 
319: Clearly, control of chaos and turbulence has great industrial value. From a mathematical 
320: point of view, the question is how much mathematis is in this control theory. So far, the 
321: mathematical merit of the theory of control of chaos is not nearly as great as proving 
322: the existence of chaos \cite{Li04}. Obviously, a lot of good numerics is in this control 
323: theory. In this article, we will address this control theory from a mathematical 
324: perspective, and try to formulate some good mathematical problems. One can add a control 
325: to any equation. But the only meaningful controls are the ones that are practical. 
326: Consider the 3D Navier-Stokes equations for example
327: \[
328: u_{i,t} +u_j u_{i,j} = -p_{,i} + \e  u_{i,jj} + f_i + \C_i \ ,
329: \]
330: defined on a spatial domain $\D$ with appropriate boundary conditions, where $\e =1 / 
331: \text{Re}$ is the inverse of the
332: Reynolds number, and $f_i=f_i(t,x)$ is the external force. Assume that without the 
333: control $\C_i$, the solutions are turbulent. The goal is to find a practical control 
334: to either tame or enhance turbulence. For instance, a practical control $\C_i = 
335: \C_i(t,x)$ should be spatially localized (perhaps near the boundary).
336: 
337: Recently, there has been quite amount of works on numerical simulations of chaos in 
338: Navier-Stokes equations \cite{FS95a} \cite{FS95b} \cite{KK01} \cite{MCH05} \cite{TOAG05} 
339: \cite{SYE06}. Here we try to combine numerics with analysis in terms of Melnikov integrals. 
340: Unlike the sine-Gordon system studied in Appendix A, analytical calculation of the 
341: Melnikov integrals is not feasible at this moment for Navier-Stokes equations. So we 
342: will resort to numerical calculations. It is an interesting open mathematical problem 
343: that whether or not 2D Euler equation is integrable as a Hamiltonian system in the 
344: Liouville sense. Since 2D Euler equation possesses infinitely many constants of motion,
345: it is tempting to conjecture that 2D Euler equation is integrable. Another support to 
346: such a conjecture is that both 2D and 3D Euler equations have Lax pairs \cite{Li01}
347: \cite{LY03} \cite{Li05L}. In fact, it is even rational to conjecture that 3D Euler equations
348: are integrable. As mentioned above, we propose the so-called heteroclinics conjecture
349: for Euler equations, i.e. there exist heteroclinic cycles for Euler equations. We numerically 
350: simulate the heteroclinic orbits and use the numerical results to conduct numerical calculations 
351: on Melnikov integrals. In these numerical simulations, it is crucial to
352: make use of known results on the spectra of linear 
353: Navier-Stokes and Euler operators \cite{Li00} \cite{LLM04} \cite{Li05}. We use the 
354: numerical Melnikov integral as a tool for both predicting and controling chaos. As a 
355: measure of chaos, we calculate the Liapunov exponents. We find that the calculated 
356: Liapunov exponents depend on the computational time interval and the precision of the 
357: computation. Nevertheless, as a measure of chaos, Liapunov exponents prove to be very robust.
358: Since the chaos is often transient, we make comparison on the base of fixed time interval
359: and fixed precision of computation.
360: 
361: Our numerics resorts to Galerkin truncations. But Galerkin truncations are somewhat 
362: singular perturbations of Euler equations. Higher single Fourier modes have more unstable 
363: eigenvalues. Therefore, it is difficult to derive dynamical pictures for Euler equations
364: from their Galerkin truncations. On the other hand, higher single Fourier modes have more
365: dissipation under Navier-Stokes flows. So Galerkin truncations perform better for 
366: Navier-Stokes equations than Euler equations. Today's computer ability still limits the 
367: size of the Galerkin truncations. With better future computer ability, Galerkin truncations
368: will paint better and better pictures of Navier-Stokes and Euler equations. It seems also 
369: important to design special models that can picture special aspects of the dynamics of 
370: Navier-Stokes and Euler equations.
371: 
372: 
373: \section{Zero Viscosity Limit of the Spectrum of 2D Linear Navier-Stokes Operator}
374: 
375: We will study the following form of 2D Navier-Stokes (NS) equation with a control,
376: \begin{equation}
377: \pa_t \Om + \{ \Psi, \Om \} = \e [\Dl \Om + f(t,x) +b\tdl (x)] \ ,
378: \label{2DNS}
379: \end{equation}
380: where $\Om$ is the vorticity which is a real scalar-valued function
381: of three variables $t$ and $x=(x_1, x_2)$, the bracket $\{\ ,\ \}$ 
382: is defined as
383: \[
384: \{ f, g\} = (\pa_{x_1} f) (\pa_{x_2}g) - (\pa_{x_2} f) (\pa_{x_1} g) \ ,
385: \]
386: where $\Psi$ is the stream function given by,
387: \[
388: u_1=- \pa_{x_2}\Psi \ ,\ \ \ u_2=\pa_{x_1} \Psi \ ,
389: \]
390: the relation between vorticity $\Om$ and stream 
391: function $\Psi$ is,
392: \[
393: \Om =\pa_{x_1} u_2 - \pa_{x_2} u_1 =\Dl \Psi \ ,
394: \]
395: and $\e = 1/\text{Re}$ is the inverse of the Reynolds number, $\Dl$ is the 
396: 2D Laplacian, $f(t,x)$ is the external force, $b\tdl (x)$ is the spatially 
397: localized control, and $b$ is the control parameter. We pose the periodic 
398: boundary condition
399: \[
400: \Om (t, x_1 +2\pi , x_2) = \Om (t, x_1 , x_2) = \Om (t, x_1, x_2 +2\pi /\al ),
401: \]
402: where $\al$ is a positive constant, i.e. the 2D NS is defined on the 2-torus $\mathbb{T}^2$. 
403: We require that $\Psi$, $f$ and $\tdl$ have mean zero
404: \[
405: \int_{\mathbb{T}^2} \Psi dx = \int_{\mathbb{T}^2} f dx = 
406: \int_{\mathbb{T}^2} \tdl dx = 0\ .
407: \]
408: Of course $\Om$ always has zero mean. In this case, $\Psi = \Dl^{-1} \Om $.
409: 
410: In both 2D and 3D, the linear NS operator obtained by linearizing NS 
411: at a fixed point has only point spectrum consisting of 
412: eigenvalues lying in a parabolic region \cite{Li05}. On the other hand, the 
413: corresponding linear Euler can have continuous spectrum besides 
414: point spectrum \cite{Li05}. The spectra of the linear NS and Euler operators can 
415: be classified into four classes in the zero viscosity limit:
416: \begin{enumerate}
417: \item {\em Persistence:} These are the eigenvalues that persist and approach 
418: to the eigenvalues of the corresponding linear Euler operator when the viscosity 
419: approaches zero.
420: \item {\em Condensation:} These are the eigenvalues that approach and form 
421: a continuous spectrum for the corresponding linear Euler operator when the viscosity 
422: approaches zero.
423: \item {\em Singularity:} These are the eigenvalues that approach to a 
424: set that is not in the spectrum of the corresponding linear Euler operator when 
425: the viscosity approaches zero.
426: \item {\em Addition:} This is a subset of the spectrum of the linear Euler operator,
427: which has no overlap with the zero viscosity limit set of the spectrum of the linear 
428: NS operator.
429: \end{enumerate} 
430: 
431: \subsection{A Shear Fixed Point}
432: 
433: For the external force $f= \Ga \cos x_1$ ($b=0$), $\Om = \Ga \cos x_1$ is a shear fixed 
434: point, where $\Ga$ is an arbitrary real nonzero constant. Choose $\al \in (0.5, 0.84)$. There is a 
435: $\e_* > 0$ such that when $\e > \e_*$, the fixed point has no eigenvalue 
436: with positive real part, and when $\e \in [0, \e_*)$, the fixed point has 
437: a unique positive eigenvalue \cite{Li05}. 
438: Notice that this unique eigenvalue persists even for linear Euler 
439: ($\e =0$). In fact, for linear Euler ($\e =0$), there is a pair of 
440: eigenvalues, and the other one is the negative of the above eigenvalue. 
441: Precise statements on such results are given in the theorem below. Later we will 
442: discover numerically that
443: some of the rest eigenvalues of the linear Navier-Stokes operator somehow form the 
444: continuous spectrum of linear Euler ($\e =0$) as $\e \ra 0$, while others
445: do not converge to the spectrum of linear Euler ($\e =0$) at all \cite{Li05} 
446: \cite{Li05a}. Using the Fourier series 
447: \[
448: \Om = \sum_{k \in \ZZ} \om_k e^{i(k_1 x_1 + \al k_2 x_2)}\ , 
449: \]
450: where $\om_{-k} = \overline{\om_k}$ (in fact, we always work in the 
451: subspace where all the $\om_k$'s are real-valued), one gets
452: the spectral equation of the linearized 2D Navier-Stokes operator at the 
453: fixed point $\Om = 2 \cos x_1$, 
454: \begin{equation}
455: A_{n-1} \om_{n-1} -\e |\hk +np|^2 \om_n - A_{n+1} \om_{n+1} = \la \om_n \ ,
456: \label{le}
457: \end{equation}
458: where $\hk \in \ZZ$, $p=(1,0)$, $\om_n = \om_{\hk +np}$, $A_n = A(p, \hk +np)$, and
459: \[
460: A(q,r) = \frac{\al}{2}\left [ \frac{1}{r_1^2+(\al r_2)^2} - 
461: \frac{1}{q_1^2+(\al q_2)^2}\right ]\left | \begin{array}{lr} 
462: q_1 & r_1 \\ q_2 & r_2 \\ \end{array} \right | \ .
463: \]
464: (In fact, the $A_n$'s should be counted twice due to switching $q$ and $r$, but the 
465: difference is only a simple scaling of $\e$ and $\la$.)
466: Thus the 2D linear NS decouples according to lines labeled by $\hk$. The following 
467: detailed theorem on the spectrum of the 2D linear NS at the fixed point 
468: $\Om = 2 \cos x_1$ was proved in \cite{Li05}.
469: \begin{theorem}[The Spectral Theorem \cite{Li05}]
470: The spectra of the 2D linear NS operator (\ref{le}) have the following 
471: properties.
472: \begin{enumerate}
473: \item $(\al \hk_2)^2+(\hk_1+n)^2 > 1$, $\forall n \in \Z$. 
474: When $\e > 0$, there is no eigenvalue of non-negative real part. 
475: When $\e = 0$, the entire spectrum is the continuous spectrum
476: \[
477: \left [ -i\al |\hk_2|, \ i\al |\hk_2| \right ]\ .
478: \]
479: \item $\hk_2 = 0$, $\hk_1 = 1$. The spectrum consists of the eigenvalues 
480: \[
481: \la = - \e n^2 \ , \quad n \in \Z \ .
482: \]
483: The eigenfunctions are the Fourier modes
484: \[
485: \tom_{np} e^{inx_1} + \ \mbox{c.c.}\ \ , \quad \forall \tom_{np} \in 
486: \C\ , \quad n \in \Z \ .
487: \]
488: As $\e \ra 0^+$, the eigenvalues are dense on the negative half of the real 
489: axis $(-\infty, 0]$. Setting $\e =0$, the only eigenvalue is $\la = 0$ of 
490: infinite multiplicity with the same eigenfunctions as above.
491: \item $\hk_2 = -1$, $\hk_1 = 0$. (a). $\e >0$. For any $\al \in (0.5, 0.95)$,
492: there is a unique $\e_*(\al)$,
493: \begin{equation}
494: \frac{\sqrt{32-3\al^6-17\al^4-16\al^2}}{2(\al^2+1)(\al^2+4)} < 
495: \e_*(\al) < \frac{1}{(\al^2+1)} \sqrt{\frac{1-\al^2}{2}}\ ,
496: \label{nuda}
497: \end{equation}
498: where the term under the square root on the left is positive for 
499: $\al \in (0.5, 0.95)$, and the left term is always less than the right term.
500: When $\e > \e_*(\al)$, there is no eigenvalue of non-negative real part. 
501: When $\e = \e_*(\al)$, $\la =0$ is an eigenvalue, and all the rest 
502: eigenvalues have negative real parts. When $\e < \e_*(\al)$, there is 
503: a unique positive eigenvalue $\la (\e )>0$, and all the rest 
504: eigenvalues have negative real parts. $\e^{-1} \la (\e )$ is a strictly 
505: monotonically decreasing function of $\e$. When $\al \in (0.5, 0.8469)$,
506: we have the estimate
507: \begin{eqnarray*}
508: & & \sqrt{\frac{\al^2(1-\al^2)}{2(\al^2+1)}-\frac{\al^4 (\al^2+3)}{4
509: (\al^2+1)(\al^2+4)}} - \e (\al^2+1) < \la (\e ) \\
510: & & < \sqrt{\frac{\al^2(1-\al^2)}{2(\al^2+1)}}- \e \al^2 \ ,
511: \end{eqnarray*}
512: where the term under the square root on the left is positive for 
513: $\al \in (0.5, 0.8469)$.
514: \[
515: \sqrt{\frac{\al^2(1-\al^2)}{2(\al^2+1)}-\frac{\al^4 (\al^2+3)}{4
516: (\al^2+1)(\al^2+4)}} \leq \lim_{\e \ra 0^+} \la (\e )  \leq 
517: \sqrt{\frac{\al^2(1-\al^2)}{2(\al^2+1)}} \ .
518: \]
519: In particular, as $\e \ra 0^+$, $\la (\e ) =\O (1)$.
520: 
521: (b). $\e =0$. When $\al \in (0.5, 0.8469)$, we have only two eigenvalues
522: $\la_0$ and $-\la_0$, where $\la_0$ is positive,
523: \[
524: \sqrt{\frac{\al^2(1-\al^2)}{2(\al^2+1)}-\frac{\al^4 (\al^2+3)}{4
525: (\al^2+1)(\al^2+4)}} < \la_0 <
526: \sqrt{\frac{\al^2(1-\al^2)}{2(\al^2+1)}} \ .
527: \]
528: The rest of the spectrum is a continuous spectrum $[-i\al , \ i\al ]$.
529: 
530: (c). For any fixed $\al \in (0.5, 0.8469)$,
531: \begin{equation}
532: \lim_{\e \ra 0^+} \la (\e ) = \la_0 \ .
533: \label{pet1}
534: \end{equation}
535: \item Finally, when $\e = 0$, the union of all the above pieces of 
536: continuous spectra is the imaginary axis $i\mathbb{R}$.
537: \end{enumerate}
538: \label{PET}
539: \end{theorem}
540: \begin{remark}
541: In the current periodic boundary condition case, viscosity does not destablize 
542: the flow in contrast to the non-slip boundary condition case \cite{Lin45}. 
543: The Orr-Sommerfeld equation and Rayleigh equation have the same periodic boundary 
544: condition in the former case, and different number of boundary conditions in the 
545: latter case.
546: \end{remark}
547: The following invariant manifold theorem of the 2D NS at the fixed point 
548: $\Om = 2 \cos x_1$ was also proved in \cite{Li05}.
549: \begin{theorem}[Invariant Manifold Theorem \cite{Li05}]
550: For any $\al \in (0.5, 0.95)$, and $\e \in (0, \e_*(\al ))$ where 
551: $\e_*(\al ) > 0$ satisfies (\ref{nuda}), in a neighborhood of $\Om = 2 \cos x_1$ 
552: in the Sobolev space $H^\ell (\mathbb{T}^2)$ ($\ell \geq 3$),
553: there are an $1$-dimensional $C^\infty$ unstable manifold and an 
554: $1$-codimensional $C^\infty$ stable manifold.
555: \end{theorem}
556: 
557: One of the goals of the work \cite{Li05} is to study the zero viscosity limit of 
558: the invariant manifolds of the 2D NS. For this study, it is crucial to understand 
559: the deformation of the linear spectra as $\e \ra 0^+$. Below we will study this 
560: numerically.
561: 
562: 
563: When $\hk_1=0$ and $\hk_2=1$, $\al =0.7$, the unique $\e_*$ in (\ref{nuda}) 
564: belongs to the interval $0.332 < \e_* < 0.339$,
565: such that when $\e < \e_*$, a positive eigenvalue appears. We test this criterion 
566: numerically and find that it is very sharp even when the truncation of 
567: the linear system (\ref{le}) is as low as $|n| \leq 100$. 
568: As $\e \ra 0^+$, we tested the truncation of 
569: the linear system (\ref{le}) up to $|n| \leq 1024$ for $\al =0.7$, the patterns are 
570: all the same. Below we present the case $|n| \leq 200$ for which the pattern is more clear. 
571: \begin{figure}
572: \includegraphics[width=4.0in,height=4.0in]{fige1.eps}
573: \caption{The eigenvalues of the linear system (\ref{le}) when $\hk_1=0$ and $\hk_2=1$, 
574: $\al =0.7$, and $\e =0.14$.}
575: \label{ge1}
576: \end{figure}
577: \begin{figure}
578: \includegraphics[width=4.0in,height=4.0in]{fige2.eps}
579: \caption{The eigenvalues of the linear system (\ref{le}) when $\hk_1=0$ and $\hk_2=1$, 
580: $\al =0.7$, and $\e =0.13$.}
581: \label{ge2}
582: \end{figure}
583: \begin{figure}
584: \includegraphics[width=4.0in,height=4.0in]{fige3.eps}
585: \caption{The eigenvalues of the linear system (\ref{le}) when $\hk_1=0$ and $\hk_2=1$, 
586: $\al =0.7$, and $\e =0.07$.}
587: \label{ge3}
588: \end{figure}
589: \begin{figure}
590: \includegraphics[width=4.0in,height=4.0in]{fige4.eps}
591: \caption{The eigenvalues of the linear system (\ref{le}) when $\hk_1=0$ and $\hk_2=1$, 
592: $\al =0.7$, and $\e =0.03$.}
593: \label{ge4}
594: \end{figure}
595: \begin{figure}
596: \includegraphics[width=4.0in,height=4.0in]{fige5.eps}
597: \caption{The eigenvalues of the linear system (\ref{le}) when $\hk_1=0$ and $\hk_2=1$, 
598: $\al =0.7$, and $\e =0.0004$.}
599: \label{ge5}
600: \end{figure}
601: \begin{figure}
602: \includegraphics[width=4.0in,height=4.0in]{fige6.eps}
603: \caption{The eigenvalues of the linear system (\ref{le}) when $\hk_1=0$ and $\hk_2=1$, 
604: $\al =0.7$, and $\e =0.00013$.}
605: \label{ge6}
606: \end{figure}
607: \begin{figure}
608: \includegraphics[width=4.0in,height=4.0in]{fige7.eps}
609: \caption{The ($\e \ra 0^+$) limiting picture of the eigenvalues of the linear system (\ref{le}) 
610: when $\hk_1=0$ and $\hk_2=1$, and $\al =0.7$.}
611: \label{ge7}
612: \end{figure}
613: \begin{figure}
614: \includegraphics[width=4.0in,height=4.0in]{fige8.eps}
615: \caption{The spectrum of the linear system (\ref{le}) when $\e = 0$, $\hk_1=0$ and $\hk_2=1$, 
616: and $\al =0.7$.}
617: \label{ge8}
618: \end{figure}
619: \begin{figure}
620: \includegraphics[width=4.0in,height=4.0in]{fige9.eps}
621: \caption{The eigenvalues of the linear system (\ref{le}) when $\hk_1=0$ and $\hk_2=2$, 
622: $\al =0.7$, and $\e =1.5$.}
623: \label{ge9}
624: \end{figure}
625: \begin{figure}
626: \includegraphics[width=4.0in,height=4.0in]{fige10.eps}
627: \caption{The eigenvalues of the linear system (\ref{le}) when $\hk_1=0$ and $\hk_2=2$, 
628: $\al =0.7$, and $\e =0.00025$.}
629: \label{ge10}
630: \end{figure}
631: \begin{figure}
632: \includegraphics[width=4.0in,height=4.0in]{fige11.eps}
633: \caption{The ($\e \ra 0^+$) limiting picture of the entire spectrum of the linear 
634: NS operator (\ref{le}) when $\al =0.7$.}
635: \label{ge11}
636: \end{figure}
637: \begin{figure}
638: \includegraphics[width=4.0in,height=4.0in]{fige12.eps}
639: \caption{The spectrum of the linear Euler operator (\ref{le}) where $\e = 0$ and $\al =0.7$.}
640: \label{ge12}
641: \end{figure}
642: 
643: \nid
644: Figure \ref{ge1} shows the case $\e =0.14$ where there is one positive eigenvalue 
645: and all the rest eigenvalues are negative. Figure \ref{ge2} shows the case $\e =0.13$ where
646: a pair of eigenvalues jumps off the real axis and becomes a complex conjugate pair.
647: Figure \ref{ge3} shows the case $\e =0.07$ where
648: another pair of eigenvalues jumps off the real axis and becomes a complex conjugate pair.
649: Figure \ref{ge4} shows the case $\e =0.03$ where
650: another pair of eigenvalues jumps off the real axis and becomes a complex conjugate pair,
651: while the former two pairs getting closer to each other. 
652: Figure \ref{ge5} shows the case $\e =0.0004$ where
653: many pairs of eigenvalues have jumped off the real axis and a bubble is formed. 
654: Figure \ref{ge6} shows the case $\e =0.00013$ where the bubble has expanded.
655: As $\e \ra 0^+$, the limiting picture is shown in Figure \ref{ge7}. Setting $\e =0$, the 
656: spectrum of the line $\hk_1=0$ and $\hk_2=1$ of the linear Euler operator is shown in 
657: Figure \ref{ge8}, where the segment on the imaginary axis is the continuous spectrum. 
658: Comparing Figures \ref{ge7} and \ref{ge8}, we see that the two eigenvalues represent 
659: ``persistence'', the vertical segment represents ``condensation'', and the two horizontal 
660: segments represent ``singularity''. Next we study one more line: $\hk_1=0$ and $\hk_2 =2$ 
661: ($\al =0.7$). In this case, there is no unstable eigenvalue. Figure \ref{ge9} shows the 
662: case $\e =1.5$ where all the eigenvalues are negative. As $\e$ is decreased, the eigenvalues 
663: go through the same process of jumping off the real axis and developing a bubble. 
664: Figure \ref{ge10} shows the case $\e =0.00025$ where the bubble has expanded. As $\e \ra 0^+$, 
665: the limiting picture is similar to Figure \ref{ge7} except that there is no persistent 
666: eigenvalue. The cases $\hk_1=0$ and $\hk_2 > 2$ ($\al =0.7$) are all the 
667: same with the case $\hk_1=0$ and $\hk_2 =2$ ($\al =0.7$). 
668: Figure \ref{ge11} shows the limiting picture of the entire spectrum of the 
669: linear NS operator as $\e \ra 0^+$. Figure \ref{ge12} shows the entire spectrum of the 
670: linear Euler operator ($\e=0$). 
671: 
672: The fascinating deformation of the spectra as $\e \ra 0^+$ and the limiting spectral picture 
673: clearly depict the nature of singular limit of the spectra as $\e \ra 0^+$. In the 
674: ``singularity'' part of the limit, there is a discrete set of values for the imaginary 
675: parts of the eigenvalues, which represent decaying oscillations with a discrete set of 
676: frequencies. Overall, the ``singularity'' part represents the temporally irreversible 
677: nature of the $\e \ra 0^+$ limit, in contrast to the reversible nature of the linear 
678: Euler equation ($\e =0$).
679: 
680: \subsection{A Cat's Eye Fixed Point}
681: 
682: In this subsection, we will study another important fixed point -- a cat's eye fixed point.
683: The periodic domain now is the square, i.e. $\al =1$ (instead of $0.7$). 
684: The cat's eye fixed point in physical variable is given by 
685: \begin{equation}
686: \Om = 2\cos mx_1 + 2 \ga \cos mx_2 \ , 
687: \label{cateye}
688: \end{equation}
689: where $m$ is a positive integer, and $\ga \in (0, 1]$.
690: In terms of Fourier modes: Let $p=(m,0)$ and $q=(0,m)$, then the cat's eye is given by
691: \[
692: \om^*_p = 1, \ \om^*_{-p} = 1, \ \om^*_q = \ga , \ \om^*_{-q} = \ga \ ,
693: \]
694: and all other $\om^*_k$'s are zero. The spectral equation for the linear 
695: 2D NS operator at the Cat's eye is then given by
696: \begin{eqnarray}
697: \la \om_k &=& A(p,k-p)\om_{k-p} - A(p,k+p)\om_{k+p} -\e |k|^2 \om_k \non \\
698: & & + \ga A(q,k-q)\om_{k-q} - \ga A(q,k+q)\om_{k+q} \ , \label{catl}
699: \end{eqnarray}
700: where 
701: \[
702: A(k,r) = \left [ \frac{1}{r_1^2+r_2^2} - 
703: \frac{1}{k_1^2+k_2^2}\right ]\left | \begin{array}{lr} 
704: k_1 & r_1 \\ k_2 & r_2 \\ \end{array} \right | \ .
705: \]
706: \begin{figure}
707: \includegraphics[width=4.0in,height=4.0in]{cat1m1.eps}
708: \caption{The spectrum of the linear NS operator (\ref{catl}) where $\e = 150$, $m=1$ and $\ga =0.5$.}
709: \label{kat1m1}
710: \end{figure}
711: \begin{figure}
712: \includegraphics[width=4.0in,height=4.0in]{cat2m1.eps}
713: \caption{The spectrum of the linear NS operator (\ref{catl}) where $\e = 120$, $m=1$ and $\ga =0.5$.}
714: \label{kat2m1}
715: \end{figure}
716: \begin{figure}
717: \includegraphics[width=4.0in,height=4.0in]{cat3m1.eps}
718: \caption{The spectrum of the linear NS operator (\ref{catl}) where $\e = 80$, $m=1$ and $\ga =0.5$.}
719: \label{kat3m1}
720: \end{figure}
721: \begin{figure}
722: \includegraphics[width=4.0in,height=4.0in]{cat4m1.eps}
723: \caption{The spectrum of the linear NS operator (\ref{catl}) where $\e = 10$, $m=1$ and $\ga =0.5$.}
724: \label{kat4m1}
725: \end{figure}
726: \begin{figure}
727: \includegraphics[width=4.0in,height=4.0in]{cat5m1.eps}
728: \caption{The spectrum of the linear NS operator (\ref{catl}) where $\e = 1$, $m=1$ and $\ga =0.5$.}
729: \label{kat5m1}
730: \end{figure}
731: \begin{figure}
732: \includegraphics[width=4.0in,height=4.0in]{cat6m1.eps}
733: \caption{The limiting picture of the spectrum of the linear NS operator (\ref{catl}) as $\e \ra 0$, 
734: where $m=1$ and $\ga =0.5$.}
735: \label{kat6m1}
736: \end{figure}
737: \begin{figure}
738: \includegraphics[width=4.0in,height=4.0in]{cat7m1.eps}
739: \caption{The spectrum of the linear Euler operator (\ref{catl}) where $\e = 0$, 
740: $m=1$ and $\ga =0.5$.}
741: \label{kat7m1}
742: \end{figure}
743: First we study the case of $m=1$ and $\ga =0.5$. Changing the value of $\ga$ does not 
744: affect the deformation patterns of the eigenvalues of the linear NS (\ref{catl}) as 
745: $\e \ra 0^+$. We truncate the system (\ref{catl}) via the Galerkin truncation 
746: $|k_1| \leq 32$ and $|k_2| \leq 32$. This is the largest Galerkin truncation that we are 
747: able to compute in a reasonable time. For smaller Galerkin truncations, the deformation 
748: patterns of the eigenvalues are the same. When $\e = 150$, all the eigenvalues of the 
749: linear NS (\ref{catl}) are negative as shown in Figure \ref{kat1m1}. When $\e$ is decreased 
750: to $\e = 120$, a pair of eigenvalues jumps off the real axis as shown in Figure \ref{kat2m1}.
751: When $\e$ is decreased to $\e = 80$, three pairs of eigenvalues jump off the real axis as 
752: shown in Figure \ref{kat3m1}. When $\e$ is decreased to $\e = 10$, many pairs of eigenvalues 
753: have jumped off the real axis and form several parabolas as shown in Figure \ref{kat4m1}.
754: When $\e$ is decreased to $\e = 1$, many parabolas are formed as shown in Figure \ref{kat5m1}.
755: The $\e \ra 0^+$ limiting picture of the eigenvalues of the linear NS (\ref{catl}) is that the 
756: eigenvalues are dense on the entire left half plane as shown in Figure \ref{kat6m1}.
757: The continuous spectrum of the linear Euler, i.e. $\e =0$ in (\ref{catl}), in any Sobolev space 
758: $H^s(\mathbb{T}^2)$ where $s$ is a non-negative integer, is a vertical band of width 
759: $2s \sg$ symmetric with respect to the imaginary axis 
760: $\{ \la \ : \ |\text{Re}(\la )| \leq s \sg \}$ as shown in Figure \ref{kat7m1}, where 
761: $\sg >0$ is the largest Liapunov exponent of the vector field given by the cat's eye (\ref{cateye})
762: \cite{SL03}. Thus the width of the vertical band is proportional to the scale $s$ of the 
763: Sobolev space $H^s(\mathbb{T}^2)$. The union of all such bands for all integers $s \geq 0$ is 
764: the entire complex plane. The eigenfunctions of the linear NS (\ref{catl}) when $\e >0$ 
765: belong to $H^s(\mathbb{T}^2)$ for all integers $s \geq 0$. All the eigenvalues of the 
766: linear NS (\ref{catl}) condense into the entire left half plane -- ``condensation''. The 
767: right half plane (or right half of the vertical band corresponding to $H^s(\mathbb{T}^2)$) 
768: represents ``addition''. Thus the possible instability hinted by the right half band of the 
769: continuous spectrum of linear Euler in $H^s(\mathbb{T}^2)$ can not be realized by real 
770: viscous fluids.
771: 
772: Next we study the case of $m=2$ and $\ga =0.5$. Changing the value of $\ga$ does not 
773: affect the deformation patterns of the eigenvalues of the linear NS (\ref{catl}) as 
774: $\e \ra 0^+$. We truncate the system (\ref{catl}) via the Galerkin truncation 
775: $|k_1| \leq 32$ and $|k_2| \leq 32$. We increase the value of $\e$ up to $2\times 10^4$, 
776: there are still eigenvalues with nonzero imaginary parts. These eigenvalues seem always 
777: complex no matter how large is $\e$. The imaginary parts of these eigenvalues are 
778: unchanged between $\e =2\times 10^4$ and $\e = 800$ as can be seen from Figures 
779: \ref{kat1m2} and \ref{kat2m2}. Decreasing $\e$, the deformation patterns are similar to 
780: those of $m=1$. When $\e = 0.2$, many eigenvalues have jumped off the real axis and 
781: form a dense parabolic region as shown in Figure \ref{kat3m2}. Decreasing $\e$ further, 
782: $6$ eigenvalues with positive real parts appear, two of which are real, and the rest 
783: four are complex. The limiting picture of the spectrum of the linear NS operator 
784: (\ref{catl}) as $\e \ra 0$ is the same with the $m=1$ case as shown in Figure \ref{kat6m1}
785: except the extra six unstable eigenvalues. The continuous spectrum of the linear Euler 
786: operator (\ref{catl}) where $\e = 0$ is the same with the $m=1$ case as shown in Figure 
787: \ref{kat7m1}. There is no analytical result on the eigenvalues of the linear Euler 
788: operator (\ref{catl}) where $\e = 0$. The numerics indicates that the $6$ eigenvalues 
789: in the right half plane and other $6$ eigenvalues in the left half plane, of the linear 
790: NS operator (\ref{catl}) persist as $\e \ra 0$, and result in $6$ eigenvalues 
791: in the right half plane and their negatives for the linear Euler operator (\ref{catl}) 
792: where $\e = 0$. 
793: 
794: \section{The Heteroclinics Conjecture for 2D Euler Equation}
795: 
796: 
797: \begin{figure}
798: \includegraphics[width=4.0in,height=4.0in]{cat1m2.eps}
799: \caption{The spectrum of the linear NS operator (\ref{catl}) where $\e = 20000$, 
800: $m=2$ and $\ga =0.5$.}
801: \label{kat1m2}
802: \end{figure}
803: \begin{figure}
804: \includegraphics[width=4.0in,height=4.0in]{cat2m2.eps}
805: \caption{The spectrum of the linear NS operator (\ref{catl}) where $\e = 800$, 
806: $m=2$ and $\ga =0.5$.}
807: \label{kat2m2}
808: \end{figure}
809: \begin{figure}
810: \includegraphics[width=4.0in,height=4.0in]{cat3m2.eps}
811: \caption{The spectrum of the linear NS operator (\ref{catl}) where $\e = 0.2$, 
812: $m=2$ and $\ga =0.5$.}
813: \label{kat3m2}
814: \end{figure}
815: 
816: Setting $\e = 0$ in the 2D Navier-Stokes equation (\ref{2DNS}), one gets the corresponding 
817: 2D Euler equation for which one has the following constants of motion:
818: \[
819: \int_{\mathbb{T}^2} |u|^2dx\ , \quad \int_{\mathbb{T}^2} F(\Om )dx 
820: \]
821: where $F$ is an arbitrary function. Consider the simple fixed point $\Om = \Ga \cos x_1$ 
822: ($\Ga \neq 0$ real constant). It has one unstable  and one stable real eigenvalues 
823: which are negative of each other. The rest of the spectrum is the entire imaginary 
824: axis which is a continuous spectrum \cite{Li00} \cite{Li05}. We will use the constant 
825: of motion 
826: \[
827: G = \int_{\mathbb{T}^2} \Om^2 dx - \int_{\mathbb{T}^2} |u|^2dx 
828: \]
829: to build a Melnikov integral for the corresponding 2D Navier-Stokes equation (\ref{2DNS}).
830: We will try to make use of the Melnikov integral as a measure of chaos and to conduct 
831: a control of chaos, around the line of fixed points $\Om = \Ga \cos x_1$ parametrized 
832: by $\Ga$. $G$ is a linear combination of the kinetic energy and the enstrophy. The 
833: gradient of $G$ in $\Om$ is given by
834: \[
835: \na_\Om G = 2(\Om +\Dl^{-1} \Om )
836: \]
837: which is zero along the line of fixed points $\Om = \Ga \cos x_1$. We define the 
838: Melnikov integral for the 2D NS (\ref{2DNS}) as 
839: \begin{eqnarray}
840: M &=& \frac{\al}{8\pi^2}\int_{-\infty}^{+\infty} \int_{\mathbb{T}^2} \na_\Om G 
841: [\Dl \Om + f(t,x) +b\tdl (x)] dxdt \label{gMel} \\
842: &=& M_0 + b M_c \ , \non 
843: \end{eqnarray}
844: where
845: \begin{eqnarray*}
846: M_0 &=& \frac{\al}{4\pi^2}\int_{-\infty}^{+\infty} \int_{\mathbb{T}^2} (\Om +\Dl^{-1} \Om )
847: [\Dl \Om + f(t,x)]dxdt \ , \\
848: M_c &=& \frac{\al}{4\pi^2}\int_{-\infty}^{+\infty} \int_{\mathbb{T}^2} (\Om +\Dl^{-1} \Om )
849: \tdl (x) dxdt \ .
850: \end{eqnarray*}
851: The question is: Where do we evaluate $M$? We propose the following conjecture.
852: \begin{itemize}
853: \item The Heteroclinics Conjecture: There is a heteroclinic orbit of the 2D Euler 
854: equation that connects $\Om = \Ga \cos x_1$ and $-\Om$. 
855: \end{itemize}
856: The rationality of this conjecture has been discussed in the Introduction.
857: If this conjecture is true, we can evaluate $M$ along the heteroclinic orbit. 
858: Also, under the perturbation of the $\e$ term, the heteroclinic orbits may 
859: break and re-connect somewhere, thereby generating the heteroclinic chaos.
860: In Appendix A, we show that one can use a Melnikov integral as a criterion to rigorously 
861: prove the existence chaos, and to conduct control of chaos. In the current case of 
862: 2D NS, the phase space is much more complicated. We do not expect the above Melnikov 
863: integral to be a rigorous measure of chaos, rather we believe that it is still relevant 
864: to chaos and control of chaos. In fact, often the chaos in 2D NS is a transient chaos.
865: In general, the zeros of a Melnikov integral do not immediately imply the existence of
866: heteroclinic or homoclinic chaos, rather they imply the intersections of the broken 
867: heteroclinic orbit with certain center-stable manifold \cite{Li04}. In fact, the 
868: Melnikov integral represents the leading order distance between the broken 
869: heteroclinic orbit and the center-stable manifold \cite{Li04}. Here for 2D NS, rigorous 
870: justification even on the above claim is an open problem due to the fact that existence 
871: of invariant manifolds for 2D Euler is an open problem.
872: 
873: The 2D Euler equation has several symmetries:
874: \begin{enumerate}
875: \item $\Om (t, x_1, x_2) \lra \Om (t, -x_1, -x_2)$,
876: \item $\Om (t, x_1, x_2) \lra -\Om (-t, x_1, x_2)$,
877: \item $\Om (t, x_1, x_2) \lra -\Om (t, -x_1, x_2)$, or $\Om (t, x_1, x_2) \lra -\Om (t, x_1, -x_2)$,
878: \item $\Om (t, x_1, x_2) \lra \Om (t, x_1+\th_1, x_2+\th_2)$, $\quad \forall \th_1, \th_2$.
879: \end{enumerate}
880: The first symmetry allows us to work in an invariant subspace in which all the $\om_k$'s are 
881: real-valued. This corresponds to the cosine transform in (\ref{FS}). In this paper, we will 
882: always work in the invariant subspace where all the $\om_k$'s are real-valued.
883: The second symmetry 
884: maps the unstable manifold of $\Ga \cos x_1$ into the stable manifold of $-\Ga \cos x_1$. The
885: third symmetry maps the unstable manifold of $\Ga \cos x_1$ into the unstable manifold of 
886: $-\Ga \cos x_1$. By choosing $\th_1 =\pi$, the fourth symmetry maps the unstable manifold of 
887: $\Ga \cos x_1$ into the unstable manifold of $-\Ga \cos x_1$. To maintain the cosine transform, 
888: the $\th_1$ and $\th_2$ in the fourth symmetry can only be $\pi$ and $\pi /\al$. 
889: 
890: If the heteroclinics conjecture is true, there is in fact a pair of heteroclinic cycles due 
891: to the above symmetries. Indeed, if there is a heteroclinic orbit asymptotic to $\Ga \cos x_1$ 
892: and $- \Ga \cos x_1$ as $t \ra -\infty$ and $+\infty$, then the third symmetry generates 
893: another heteroclinic orbit asymptotic to $-\Ga \cos x_1$ and $\Ga \cos x_1$ as 
894: $t \ra -\infty$ and $+\infty$. Together they form a heteroclinic cycle. Finally the second 
895: symmetry generates another heteroclinic cycle. That is, we have a pair of heteroclinic cycles.
896: 
897: Using the Fourier series
898: \begin{equation}
899: \Om = \sum_{k \in \ZZ} \om_k e^{i(k_1 x_1 + \al k_2 x_2)}\ , 
900: \label{FS}
901: \end{equation}
902: where $\om_{-k} = \overline{\om_k}$ and $F_{-k} = \overline{F_k}$, one gets
903: the kinetic form of the 2D Euler equation
904: \[
905: \dot{\om}_k = \sum_{k=m+n} A(m,n) \ \om_m \om_n \ ,
906: \]
907: where 
908: \[
909: A(m,n) = \frac{\al}{2}\left [ \frac{1}{n_1^2+(\al n_2)^2} - 
910: \frac{1}{m_1^2+(\al m_2)^2}\right ]\left | \begin{array}{lr} 
911: m_1 & n_1 \\ m_2 & n_2 \\ \end{array} \right | \ .
912: \]
913: Denote by $\Sg$ the hyperplane
914: \[
915: \Sg = \left \{ \om \ | \ \om_k = 0 \ , \quad \forall \text{ even } k_2 \right \} \ .
916: \]
917: Notice that the existence of invariant manifolds around the fixed point $\Om =\Ga \cos x_1$ 
918: is an open problem. We have the following theorem.
919: \begin{theorem}
920: Assume that the fixed point $\Om = \Ga \cos x_1$ has a 1-dimensional 
921: local unstable manifold $W^u$, 
922: and $W^u \cap \Sg \neq \emptyset$; then the heteroclinics conjecture is true, i.e.  
923: there is a heteroclinic orbit to the 2D Euler equation that connects 
924: $\Om = \Ga \cos x_1$ and $-\Om$.
925: \label{hcthm}
926: \end{theorem}
927: \begin{proof}
928: Let $\Om (t, x_1, x_2)$ be an orbit in $W^u$ parametrized such that 
929: \[
930: \Om (0, x_1, x_2) \in \Sg \ .
931: \]
932: Then by the definition of $\Sg$,
933: \begin{equation}
934: \Om (0, x_1, x_2) = -\Om (0, x_1, x_2+\pi /\al ) \ .
935: \label{cpt}
936: \end{equation}
937: By the second and fourth symmetries, 
938: \[
939: -\Om (-t, x_1, x_2+\pi /\al )
940: \]
941: is in the stable manifold of $-\Om$. Thus 
942: \[
943: \Om (t, x_1, x_2) \text{  and  } -\Om (-t, x_1, x_2+\pi /\al )
944: \]
945: are connected at $t=0$, and together they form a heteroclinic orbit that connects 
946: $\Om = \Ga \cos x_1$ and $-\Om$.
947: \end{proof}
948: 
949: 
950: \section{Numerical Verification of the Heteroclinics Conjecture for 2D Euler Equation}
951: 
952: Besides the symmetries mentioned in last section, we will also make use of the 
953: conserved quantities: kinetic energy $E=\sum |k|^{-2} \om_k^2$ 
954: (where $|k|^2=k_1^2+\al^2 k_2^2$) and enstrophy $S=\sum \om_k^2$, which 
955: will survive as conserved quantities for any symmetric Galerkin truncation, to help us to 
956: track the heteroclinic orbit. We will only consider the case that all the $\om_k$'s are 
957: real-valued (i.e. $\cos$-transform). 
958: 
959: We make a Galerkin truncation by keeping modes: $\{ |k_1| \leq 2, |k_2| \leq 2 \}$, which results
960: in a $12$ dimensional system. We choose $\al =0.7$. After careful consideration of the above 
961: mentioned symmetries and conserved quantities ($E=S=1$), we discover the following initial 
962: condition that best tracks the heteroclinic orbit:
963: \begin{eqnarray}
964: & & \om_{(j,0)}=\om_{(j,2)} = 0 \ , \quad \forall j\ , \non \\
965: & & \om_{(0,1)}=0.603624\ , \quad \om_{(1,1)}=- \om_{(-1,1)}=0.357832\ , \label{IC} \\
966: & & \om_{(2,1)}=\om_{(-2,1)}=0.435632\ . \non 
967: \end{eqnarray}
968: Starting from this initial condition, we calculate the solution in both forward and backward time
969: for the same duration of $T =11.8$, and we discover the approximate heteroclinic orbit 
970: asymptotic to $2\cos x_1$ and $-2\cos x_1$ as $t \ra -\infty$ and $+\infty$, as shown in Figure 
971: \ref{fi1}. Then the third symmetry generates another heteroclinic orbit asymptotic to 
972: $-2\cos x_1$ and $2\cos x_1$ as $t \ra -\infty$ and $+\infty$. Together they form a heteroclinic 
973: cycle. Finally the second symmetry generates another heteroclinic cycle. That is, we have 
974: a pair of heteroclinic cycles. Notice also that the approximate heteroclinic orbit in 
975: Figure \ref{fi1} has an extra loop before landing near $-2\cos x_1$. This is due to the 
976: $k_2=2$ modes in the Galerkin truncation. For smaller Galerkin truncations, the heteroclinic orbits 
977: can be calculated exactly by hand and have no such extra loop \cite{Li03e} \cite{Li06e}, 
978: and existence of chaos generated by the heteroclinic orbit can be rigorously proved in some case \cite{Li06e}.
979: \begin{figure}
980: \includegraphics[width=4.0in,height=3.0in]{fig1.eps}
981: \caption{The approximate heteroclinic orbit projected onto the ($\om_{(1,0)},\om_{(1,1)}$)-plane
982: in the case of the $\{ |k_1| \leq 2, |k_2| \leq 2 \}$ Galerkin truncation of the 2D Euler 
983: equation.}
984: \label{fi1}
985: \end{figure}
986: \begin{remark}
987: We have also conducted numerical experiments on Galerkin truncations by keeping more modes: 
988: $\{ |k_1| \leq 4, |k_2| \leq 4 \}$ and $\{ |k_1| \leq 8, |k_2| \leq 8 \}$. We found orbits 
989: that have similar behavior as the approximate heteroclinic orbit in Figure \ref{fi1}, but 
990: their approximations to heteroclinics are not as good as the one in Figure \ref{fi1}. 
991: \end{remark}
992: 
993: \section{Melnikov Integral and Numerical Simulation of Chaos in 2D Navier-Stokes Equation}
994: 
995: Without the control ($b=0$), using Fourier series for the 2D NS equation (\ref{2DNS}),
996: \[
997: \Om = \sum_{k \in \ZZ} \om_k e^{i(k_1 x_1 + \al k_2 x_2)}\ , \quad 
998: f = \sum_{k \in \ZZ} F_k e^{i(k_1 x_1 + \al k_2 x_2)} \ , 
999: \]
1000: where $\om_{-k} = \overline{\om_k}$ and $F_{-k} = \overline{F_k}$ (in fact, we always work in the 
1001: subspace where all the $\om_k$'s and $F_k$'s are real-valued), one gets
1002: the kinetic form of 2D NS
1003: \[
1004: \dot{\om}_k = \sum_{k=m+n} A(m,n) \ \om_m \om_n  +\e \left ( -\left [k_1^2 + (\al k_2)^2\right ]
1005: \om_k +F_k \right ) \ ,
1006: \]
1007: where 
1008: \[
1009: A(m,n) = \frac{\al}{2}\left [ \frac{1}{n_1^2+(\al n_2)^2} - 
1010: \frac{1}{m_1^2+(\al m_2)^2}\right ]\left | \begin{array}{lr} 
1011: m_1 & n_1 \\ m_2 & n_2 \\ \end{array} \right | \ .
1012: \]
1013: 
1014: For the numerical simulation of chaos, we continue the study on the Galerkin truncation:
1015: $\{ |k_1| \leq 2, |k_2| \leq 2 \}$. We will use the Melnikov integral (\ref{gMel}) to 
1016: test the existence of chaos. We always start from the initial condition (\ref{IC}). 
1017: We choose the external force
1018: \begin{equation}
1019: f = a \sin t \cos (x_1 +\al x_2)\ .
1020: \label{EF}
1021: \end{equation}
1022: Then the Melnikov integral (\ref{gMel}) has the expression
1023: \begin{equation}
1024: M_0 = M_1 + a \sqrt{M_2^2+M_3^2} \sin (t_0 +\th )\ ,
1025: \label{12mel}
1026: \end{equation}
1027: where 
1028: \begin{eqnarray*}
1029: & & \sin \th = \frac{M_3}{\sqrt{M_2^2+M_3^2}} \ , \quad \cos \th = \frac{M_2}{\sqrt{M_2^2+M_3^2}} \ , \\
1030: M_1 &=& \frac{\al}{4\pi^2}\int_{-\infty}^{+\infty} \int_0^{2\pi /\al} \int_0^{2\pi}(\Om +\Dl^{-1} \Om )
1031: \Dl \Om \ dx_1 dx_2dt \ , \\
1032: M_2 &=& \frac{\al}{4\pi^2}\int_{-\infty}^{+\infty} \int_0^{2\pi /\al} \int_0^{2\pi}(\Om +\Dl^{-1} \Om )
1033: \cos t \cos (x_1 +\al x_2)  \ dx_1 dx_2dt \ , \\
1034: M_3 &=& \frac{\al}{4\pi^2}\int_{-\infty}^{+\infty} \int_0^{2\pi /\al} \int_0^{2\pi}(\Om +\Dl^{-1} \Om )
1035: \sin t \cos (x_1 +\al x_2)  \ dx_1 dx_2dt \ ,
1036: \end{eqnarray*}
1037: where $\Om (t)$ is the approximate heteroclinic orbit in Figure \ref{fi1} with $\Om (0)$ given by 
1038: (\ref{IC}). The time integral is in fact over the interval [$-11.8 ,11.8$] rather than ($-\infty , 
1039: \infty$). Direct numerical computation gives that 
1040: \[
1041: M_1 =  -29.0977\ , \quad M_2 = -0.06754695 \ , \quad M_3 = 0 \ .
1042: \]
1043: Setting $M_0 = 0$ in (\ref{12mel}), we obtain that
1044: \[
1045: \sin (t_0 +\pi ) =  \frac{430.77741}{a} \ .
1046: \]
1047: Thus, when 
1048: \begin{equation}
1049: |a| > 430.77741\ ,
1050: \label{MC}
1051: \end{equation}
1052: there are solutions to $M_0 = 0$. Next we will test the Melnikov criterion (\ref{MC}) and see if it is 
1053: related to chaos. We define 
1054: an average Liapunov exponent $\sg$ in the following manner: For a large time interval $t \in [0, T]$,
1055: let $t_0 = T$ and 
1056: \[
1057: t_n = T + n 2\pi \ , \quad \text{ where } 0 \leq n \leq N \ , \text { and } 
1058: N = 10^3 \text{ or } 2\times 10^3\ .
1059: \]
1060: We define
1061: \[
1062: \sg_n = \frac{1}{2\pi} \ln \frac{\| \Dl \om (t_n + 2\pi )\|}{\| \Dl \om (t_n )\|} \ .
1063: \]
1064: Then the average Liapunov exponent $\sg$ is given by
1065: \[
1066: \sg = \frac{1}{N} \sum_{n=0}^{N-1} \sg_n \ .
1067: \]
1068: We introduce the Poincar\'e return map on the section given by $\om (1,0) = 0$ and we only record 
1069: one direction intersection (from $\om (1,0)$ positive to negative). For a large time interval 
1070: $t \in [0, T]$, we only record the last $1000$ intersections and we use $\bullet$ to denote the 
1071: intersections with the two heteroclinic cycles. All the numerical simulations start from the 
1072: initial condition (\ref{IC}). The average Liapunov exponent computed here depends on the time 
1073: interval, the ensemble of average, and computer accuracy. It only makes sense when it is compared in 
1074: the same setting. When $\e =0$, there is no dissipation 
1075: and no forcing. For a large time interval $t \in [0, T]$, the average Liapunov exponent $\sg$
1076: is as follows:
1077: \[
1078: \begin{array}{cccc} T= 4\times 10^4 \pi & \quad T= 8\times 10^4 \pi & \quad  T= 12 \times 10^4 \pi
1079: & \quad T= 80 \times 10^4 \pi \ , \cr 
1080: \sg = 0.042 & \sg = 0.0344 & \sg = 0.044 & \sg = 0.0848 \ .\cr 
1081: \end{array}
1082: \]
1083: Figures \ref{fi2}-\ref{fi3} are the corresponding Poincar\'e return map plots. The dynamics is chaotic.
1084: It seems that the life time of the chaos is infinite (i.e. non-transient chaos). 
1085: \begin{figure}
1086: \includegraphics[width=4.0in,height=4.0in]{fig2.eps}
1087: \caption{The Poincar\'e return map plot projected onto the plane ($\om_{(0,1)},\om_{(1,1)}$)-plane,
1088: in the case of the $\{ |k_1| \leq 2, |k_2| \leq 2 \}$ Galerkin truncation of the 2D Euler 
1089: equation (i.e. $\e =0$), where $t \in [0, T]$, $T=4\times 10^4 \pi$, only the last $1000$ 
1090: intersections are recorded.}
1091: \label{fi2}
1092: \end{figure}
1093: \begin{figure}
1094: \includegraphics[width=4.0in,height=4.0in]{fig3.eps}
1095: \caption{The same as Figure \ref{fi2}, except $T=80\times 10^4 \pi$.}
1096: \label{fi3}
1097: \end{figure}
1098: 
1099: When $\e >0$, we find that in all the cases that we tested, the dynamics is always a transient 
1100: chaos. The Melnikov criterion is only some sort of necessary condition for the existence of 
1101: heteroclinic chaos \cite{Li04}. When the Melnikov integral is zero, it gives an indication 
1102: of a re-intersection of the broken heteroclinic orbit with certain large dimensional center-stable 
1103: manifold \cite{Li04}. We believe that such a re-intersection will be reflected by the Liapunov 
1104: exponent as inducing transient chaos. 
1105: When $\e = 10^{-5}$, $a \in [0, 1208]$, and $T=4\times 10^4 \pi$, we find that 
1106: \[
1107: \sg \sim 10^{-4}\ .
1108: \]
1109: For instances,
1110: \begin{equation}
1111: \begin{array}{lllll}  a=400 &  a=430 & a=440 & a=650 & a=850 \ , \cr 
1112: \sg = 4.9 \times 10^{-4}  & \sg = 2.6 \times 10^{-4} & \sg = 5.3 \times 10^{-4} & 
1113: \sg = 5.9 \times 10^{-4} & \sg = 1.0 \times 10^{-4}  \ .\cr 
1114: \end{array}
1115: \label{noj}
1116: \end{equation}
1117: But we discover a sharp jump of $\sg$ near $a = 1208$ as shown below:
1118: \begin{equation}
1119: \begin{array}{lllll}  & T= 2\times 10^4 \pi &  T= 4 \times 10^4 \pi
1120: & T= 8 \times 10^4 \pi & T= 8\times 10^5 \pi\ , \cr 
1121: a=1208  & \sg = 3.6 \times 10^{-4} & \sg = 3.9 \times 10^{-4} & 
1122: \sg = 4.1 \times 10^{-4} & \sg = 0 \ ,\cr 
1123: a=1208.2 & \sg = 6.1 \times 10^{-2} & \sg = 6.1 \times 10^{-2} & 
1124: \sg = 2.5 \times 10^{-2}& \sg =0 \ .\cr 
1125: \end{array}
1126: \label{shar}
1127: \end{equation}
1128: For $T= 4 \times 10^4 \pi$, the corresponding Poincar\'e return map plots are shown in 
1129: Figures \ref{fi4}-\ref{fi5}. When $T= 80 \times 10^4 \pi$, the chaos for the $a=1208.2$ 
1130: case also disappears. When $a > 1208.2$, $\sg$ can still be $\sim 10^{-4}$. But we 
1131: did not observe any sharp jump of $\sg$.
1132: \begin{figure}
1133: \includegraphics[width=4.0in,height=4.0in]{fig4.eps}
1134: \caption{The same as Figure \ref{fi2}, except $\e = 10^{-5}$, $a =1208$, and $T=2\times 10^4 \pi$.}
1135: \label{fi4}
1136: \end{figure}
1137: \begin{figure}
1138: \includegraphics[width=4.0in,height=4.0in]{fig5.eps}
1139: \caption{The same as Figure \ref{fi2}, except $\e = 10^{-5}$, $a =1208.2$, and $T=2\times 10^4 \pi$.}
1140: \label{fi5}
1141: \end{figure}
1142: We believe that the sharp jump of $\sg$ near $a=1208$ is due to the 
1143: generation of new transient heteroclinic chaos. Our Melnikov integral calculation 
1144: predicts when $|a| > 430.77741$, the broken heteroclinic orbit re-intersects 
1145: with certain center-stable manifold. This does not mean that new heteroclinic 
1146: orbit is automatically generated. Perhaps near $a=1208$, new heteroclinic cycle 
1147: is generated and leads to transient heteroclinic chaos. Since our analysis cannot 
1148: access such information in the phase space, our comments here are purely speculations.
1149: 
1150: \begin{remark}
1151: We have also conducted numerical experiments on Galerkin truncations by keeping more modes: 
1152: $\{ |k_1| \leq 4, |k_2| \leq 4 \}$ and $\{ |k_1| \leq 8, |k_2| \leq 8 \}$. We found that 
1153: when $\e =0$, the strength of chaos increases as the modes are increased: For $T=8\times 10^3 \pi$,
1154: \[
1155: \begin{array}{lll} |k_1|, |k_2| \leq 2 &  |k_1|, |k_2| \leq 4 &  |k_1|, |k_2| \leq 8 \ , \cr 
1156: \sg = 4.7 \times 10^{-2} & \sg = 1.3 \times 10^{-1} & \sg = 1.7 \times 10^{-1} \ . \cr 
1157: \end{array}
1158: \]
1159: Nevertheless, this does not hint that the dynamics of 2D Euler equation is chaotic since all
1160: Galerkin truncations are perturbations of the 2D Euler equation. In fact, higher 
1161: single Fourier modes (as fixed points) have more eigenvalues with positive real parts. 
1162: 
1163: When $\e >0$, the strength of chaos decreases as the modes are increased. Higher modes 
1164: have more dissipations. Also all the chaos are transient. After enough time ($\sim 
1165: 2\times 10^4 \pi$), the $\e =0$ chaos is almost smeared away by dissipation, we believe that at this stage 
1166: the re-intersected heteroclinic orbits play a role and can enhance the transient chaos. It is this 
1167: stage where the Melnikov calculation is effective. 
1168: \end{remark}
1169: 
1170: \section{Melnikov Integral and Control of Chaos in 2D Navier-Stokes Equation} 
1171: 
1172: Now we turn on the control ($b \neq 0$). we continue the study on the Galerkin truncation:
1173: $\{ |k_1| \leq 2, |k_2| \leq 2 \}$.  
1174: We choose $\tdl (x)$ as follows
1175: \begin{equation}
1176: \tdl (x)= \sum_k e^{i(k_1 x_1 + \al k_2 x_2)}\ .
1177: \label{SC}
1178: \end{equation}
1179: Then the Melnikov integral $M$ in (\ref{gMel}) is given by 
1180: \begin{equation}
1181: M=M_0 +b M_c \ ,
1182: \label{CM}
1183: \end{equation}
1184: where $M_0$ is given by (\ref{12mel}) and 
1185: \[
1186: M_c = \frac{\al}{4\pi^2}\int_{-\infty}^{+\infty} \int_0^{2\pi /\al} \int_0^{2\pi}(\Om +\Dl^{-1} \Om )
1187: \tdl (x) \ dx_1 dx_2dt \ , 
1188: \]
1189: evaluated along the approximate heteroclinic orbit in Figure \ref{fi1}. We find that
1190: \[
1191: M_c = -18.6884 \ .
1192: \]
1193: Thus
1194: \[
1195: M= -29.0977 -18.6884 \ b + 0.06754695 \ a \sin (t_0 +\pi )\ .
1196: \]
1197: When 
1198: \[
1199: b = - 1.557 \ ,
1200: \]
1201: the Melnikov integral $M$ has roots for any $a \neq 0$. 
1202: 
1203: All the numerical simulations start from the initial condition (\ref{IC}).
1204: When $\e = 10^{-5}$, $b = - 1.557$, and $T=10^4 \pi$, we find that:
1205: \[
1206: \begin{array}{lllll}  a=1 &  a=10 & a=200 & a=400 & a=800 \ , \cr 
1207: \sg = 7.1 \times 10^{-4}  & \sg = 8.8 \times 10^{-4} & \sg = 9.3 \times 10^{-4} & 
1208: \sg = 9.0 \times 10^{-4} & \sg = 6.6 \times 10^{-4}  \ .\cr 
1209: \end{array}
1210: \]
1211: \[
1212: \begin{array}{lllll}  a=1000 &  a=1208 & a=1208.2 & a=1500 & a=3000 \ , \cr 
1213: \sg = 8.6 \times 10^{-4}  & \sg = 8.1 \times 10^{-4} & \sg = 8.4 \times 10^{-4} & 
1214: \sg = 8.6 \times 10^{-4} & \sg = 7.7 \times 10^{-4}  \ .\cr 
1215: \end{array}
1216: \]
1217: In comparison with (\ref{noj}), the values of the Liapunov exponents under the control 
1218: are doubled. Thus the control seems enhancing chaos but not dramatically. We did not observe 
1219: the sharp jump of the values of $\sg$ around $a=1208$ as in the $b=0$ case (\ref{shar}).
1220: Here the control theory is not as rigorous as that of sine-Gordon system in Appendix A. 
1221: But we believe that the Melnikov integral can play a significant role in the control of 
1222: chaos in NS equation. After all, the chaos in NS is generated by instabilities characterized 
1223: by unstable eigenvalues. And these unstable eigenvalues persist for Euler equation as 
1224: shown in a previous section. For Euler equation, these unstable eigenvalues characterize
1225: hyperbolic structures which are very likely of heteroclinics type due to infinitely many 
1226: constants of motion. Thus Melnikov integrals supported upon these hyperbolic structures 
1227: should play an important role in predicting and controling chaos. 
1228: 
1229: 
1230: \section{Zero Viscosity Limit of the Spectrum of 3D Linear Navier-Stokes Operator}
1231: 
1232: We will study the following form of 3D Navier-Stokes equation with a control,
1233: \begin{equation}
1234: \pa_t \Om + (u \cdot \na) \Om - (\Om \cdot \na) u = \e [\Dl \Om + f(t,x) +b \tdl (x)] \ ,
1235: \label{3DNS}
1236: \end{equation}
1237: where $u = (u_1, u_2, u_3)$ is the velocity, $\Om = (\Om_1, \Om_2, \Om_3)$
1238: is the vorticity, $\na = (\pa_{x_1}, \pa_{x_2}, \pa_{x_3})$, 
1239: $\Om = \na \times u$, $\na \cdot u = 0$, $\e = 1/\text{Re}$ is the inverse of 
1240: the Reynolds number, $\Dl$ is 
1241: the 3D Laplacian, and $f(t,x) = (f_1(t,x), f_2(t,x), f_3(t,x))$ is the external force,
1242: $b\tdl (x)$ is the spatially localized control, and $b$ is the control parameter.
1243: We also pose periodic boundary condition of period ($2\pi / \al , 2\pi / \be , 2\pi$), i.e.
1244: the 3D NS is defined on the 3-torus $\mathbb{T}^3$. We require that $u$, $\Om$,
1245: $f$ and $\tdl$ all have mean zero. In this case, $u$ can be uniquely 
1246: determined from $\Om$ by Fourier transform:
1247: \begin{eqnarray*}
1248: U_1(k) &=& i |k|^{-2} [ k_2 \om_3(k) - k_3 \om_2(k)]\ , \\
1249: U_2(k) &=& i |k|^{-2} [ k_3 \om_1(k) - k_1 \om_3(k)]\ , \\
1250: U_3(k) &=& i |k|^{-2} [ k_1 \om_2(k) - k_2 \om_1(k)]\ ,
1251: \end{eqnarray*}
1252: which can be rewritten in the compact form 
1253: \[
1254: U_\ell (k) = i |k|^{-2} \ve_{\ell m n} k_m \om_n(k) \ ,
1255: \]
1256: where $\ve_{\ell m n}$ is the permutation symbol ($\ell , m, n = 1,2,3$),
1257: \begin{eqnarray*}
1258: & & k = (k_1,k_2,k_3)=(\al \k_1,\be \k_2, \k_3) \ , \quad \k = (\k_1,\k_2,\k_3) \ , \\
1259: & & u_\ell (x) =\sum_{\k \in \ZZZ} U_\ell (k) e^{ik\cdot x} \ , \quad 
1260: \Om_\ell (x) =\sum_{\k \in \ZZZ} \om_\ell (k) e^{ik\cdot x} \ .
1261: \end{eqnarray*}
1262: Using these Fourier transforms together with
1263: \[
1264: f_\ell (x) =\sum_{\k \in \ZZZ} F_\ell (k) e^{ik\cdot x} \ , \quad 
1265: \tdl_\ell (x) =\sum_{\k \in \ZZZ} \Dl_\ell (k) e^{ik\cdot x} \ ,
1266: \]
1267: we can rewrite the 3D NS (\ref{3DNS}) into the kinetic form 
1268: \begin{eqnarray}
1269: & & \pa_t \om_\ell (k) + k_s \sum_{k=\tk + \hk}|\tk|^{-2}\tk_m \om_n(\tk )
1270: [\ve_{\ell m n} \om_s(\hk ) - \ve_{s m n} \om_\ell (\hk )] \non \\
1271: & & = \e [-|k|^2 \om_\ell (k) + F_\ell (k) + b \Dl_\ell (k)]\ . \label{K3D}
1272: \end{eqnarray}
1273: A popular example of fixed points of the 3D NS (\ref{3DNS}) is the so-called 
1274: ABC flow \cite{Dom86}
1275: \begin{equation}
1276: u_1 = A \sin x_3 + C \cos x_2, \
1277: u_2 = B \sin x_1 + A \cos x_3, \
1278: u_3 = C \sin x_2 + B \cos x_1, 
1279: \label{ABC}
1280: \end{equation}
1281: where $\al =\be =1$ and $\Om = u = f$ and $b=0$. The popularity comes from the fact that 
1282: the Lagrangian fluid particle flow generated 
1283: by the vector field (\ref{ABC}) can still be chaotic \cite{Dom86}. On the other hand, in 
1284: Appendix B, we prove that the Lagrangian
1285: flow generated by any solution to the 2D Euler equation is always integrable.
1286: 
1287: \subsection{A 3D Shear Fixed Point}
1288: 
1289: Below we will study the simplest fixed point -- the 3D shear flow (which is also a special 
1290: case of the ABC flow (\ref{ABC}) where $A=2$ and $B=C=0$):
1291: \begin{equation}
1292: \Om_1 = 2 \sin x_3 \ , \quad 
1293: \Om_2 = 2 \cos x_3 \ , \quad
1294: \Om_3 = 0 \ . 
1295: \label{SABC}
1296: \end{equation}
1297: Let $p=(0,0,1)$, the Fourier transform $\om^*$ of the fixed point is given by:
1298: \[
1299: \om^*_1(p) = - i , \ \om^*_2(p) = 1 , \ \om^*_1(-p) = i  , \ \om^*_2(-p) = 1 , 
1300: \ \om^*_3(p) = \om^*_3(-p) = 0 ,
1301: \]
1302: and $\om^*_\ell (k) = 0$, $\forall k \neq p \text{ or } -p$. We choose 
1303: $\al =0.7$ and $\be =1.3$ hoping that the fixed point $\om^*$ has only one unstable eigenvalue.
1304: The spectral equations of the 3D linear NS operator at the fixed point $\om^*$ are given by 
1305: \begin{eqnarray*}
1306: & & [(k_1+ik_2)-ik_2|k-p|^{-2}]\om_1(k-p) \\
1307: & & +[i|k-p|^{-2}k_1+i|k-p|^{-2}(k_1+ik_2)(k_3-1)]\om_2(k-p) \\
1308: & & +[-1-ik_2(k_1+ik_2)|k-p|^{-2}]\om_3(k-p) +\e |k|^2 \om_1(k) \\
1309: & & +[-(k_1-ik_2)-ik_2|k+p|^{-2}]\om_1(k+p)\\
1310: & & +[i|k+p|^{-2}k_1-i|k+p|^{-2}(k_1-ik_2)(k_3+1)]\om_2(k+p) \\
1311: & & +[-1+ik_2(k_1-ik_2)|k+p|^{-2}]\om_3(k+p) =-\la  \om_1(k) \ ,
1312: \end{eqnarray*}
1313: \begin{eqnarray*}
1314: & & [k_2|k-p|^{-2}-i(k_3-1)(k_1+ik_2)|k-p|^{-2}]\om_1(k-p) \\
1315: & & +[(k_1+ik_2)-k_1|k-p|^{-2}]\om_2(k-p) \\
1316: & & +[-i+ik_1(k_1+ik_2)|k-p|^{-2}]\om_3(k-p) +\e |k|^2 \om_2(k) \\
1317: & & +[-k_2|k+p|^{-2}+i(k_3+1)(k_1-ik_2)|k+p|^{-2}]\om_1(k+p) \\
1318: & & +[-(k_1-ik_2)+k_1|k+p|^{-2}]\om_2(k+p) \\
1319: & & +[i-ik_1(k_1-ik_2)|k+p|^{-2}]\om_3(k+p)  =-\la  \om_2(k) \ ,
1320: \end{eqnarray*}
1321: \begin{eqnarray*}
1322: & & [ik_2(k_1+ik_2)|k-p|^{-2}]\om_1(k-p) 
1323: +[-ik_1(k_1+ik_2)|k-p|^{-2}]\om_2(k-p) \\
1324: & & +[k_1+ik_2]\om_3(k-p) +\e |k|^2 \om_3(k) 
1325: +[-ik_2(k_1-ik_2)|k+p|^{-2}]\om_1(k+p)\\
1326: & & +[ik_1(k_1-ik_2)|k+p|^{-2}]\om_2(k+p) 
1327: +[-(k_1-ik_2)]\om_3(k+p) =-\la  \om_3(k) \ .
1328: \end{eqnarray*}
1329: Thus the 3D linear NS operator also decouples according to the lines $\hk + j p$ ($j\in 
1330: \mathbb{Z}$). Next we study the zero viscosity limit of the spectrum of this 3D linear NS operator.
1331: When $\hk = (\al ,0,0)$, we have tested the truncation of the line $\hk + j p$ up to $|j| \leq 400$.
1332: The deformation pattern stays the same. Below we present the case $|j|\leq 100$ for which the 
1333: pictures are more clear. Figure \ref{3ge1} shows the case $\e =2.0$ where all the eigenvalues 
1334: are negative. Figure \ref{3ge2} shows the case $\e =1.8$ where a pair of eigenvalues
1335: jumps off the real axis. When $\e \leq 0.66$, a unique positive eigenvalue appears.
1336: Figure \ref{3ge1} shows the case $\e =0.0007$ where a bubble has developed.
1337: \begin{figure}
1338: \includegraphics[width=4.0in,height=4.0in]{3fige1.eps}
1339: \caption{The eigenvalues of the line $\hk = (\al ,0,0)$ are all negative when $\e =2.0$.}
1340: \label{3ge1}
1341: \end{figure}
1342: \begin{figure}
1343: \includegraphics[width=4.0in,height=4.0in]{3fige2.eps}
1344: \caption{A pair of eigenvalues of the line $\hk = (\al ,0,0)$ jumps off the real axis when $\e =1.8$.}
1345: \label{3ge2}
1346: \end{figure}
1347: \begin{figure}
1348: \includegraphics[width=4.0in,height=4.0in]{3fige3.eps}
1349: \caption{A bubble of eigenvalues of the line $\hk = (\al ,0,0)$ has developed when $\e =0.0007$.}
1350: \label{3ge3}
1351: \end{figure}
1352: As $\e \ra 0^+$, the limiting picture is the same with Figure \ref{ge7}. When 
1353: $\e =0$, The spectrum picture is the same with Figure \ref{ge8}.
1354: All other decoupled systems have the same bifurcation patterns but without the 
1355: pair of persistent eigenvalues. For the entire spectrum of the 3D linear NS operator,
1356: the limiting picture is the same with Figure \ref{ge11} as $\e \ra 0^+$; and the spectrum is the 
1357: same with Figure \ref{ge12} when setting $\e =0$ \cite{Shv06}.
1358: It seems that the Unstable Disk Theorem \cite{Li00} of the 2D linear Euler case is still valid: 
1359: $|\hk +jp| < |p|$ for some $j$, implies that there is an eigenvalue of positive real part; while 
1360: $|\hk +jp| > |p|$ for all $j$, implies that there is no eigenvalue of positive real part.
1361: 
1362: \subsection{The ABC Fixed Point}
1363: 
1364: In this case, the periodic domain is the cube, i.e. $\al =\be =1$. The ABC flow
1365: is given specifically by
1366: \[
1367: \Om^*_1 = A \sin mx_3 + C \cos mx_2 , \ 
1368: \Om^*_2 = B \sin mx_1 + A \cos mx_3 , \
1369: \Om^*_3 = C \sin mx_2 + B \cos mx_1 ,  
1370: \]
1371: where $m$ is a positive integer, and ($A,B,C$) are real parameters.
1372: In terms of Fourier modes: Let $p=(m,0,0)$, $q=(0,m,0)$, and $r=(0,0,m)$, 
1373: then the ABC flow is given by
1374: \begin{eqnarray*}
1375: & & \om_1^*(q) = \frac{1}{2} C, \ \om_1^*(-q) = \frac{1}{2} C, \\
1376: & & \om_1^*(r) = \frac{1}{2i} A, \ \om_1^*(-r) = -\frac{1}{2i} A, \\
1377: & & \om_2^*(p) = \frac{1}{2i} B, \ \om_2^*(-p) = -\frac{1}{2i} B, \\
1378: & & \om_2^*(r) = \frac{1}{2} A, \ \om_2^*(-r) = \frac{1}{2} A, \\
1379: & & \om_3^*(p) = \frac{1}{2} B, \ \om_3^*(-p) = \frac{1}{2} B, \\
1380: & & \om_3^*(q) = \frac{1}{2i} C, \ \om_3^*(-q) = -\frac{1}{2i} C .
1381: \end{eqnarray*}
1382: The spectral equation for the linear 3D NS operator at the ABC flow is 
1383: then given by
1384: \begin{eqnarray*}
1385: \la \om_\ell (k) &=& -\e |k|^2 \om_\ell (k) -
1386: k_s \sum_{k=\tk + \hk}\bigg [ |\tk|^{-2}\tk_m \om^*_n(\tk )
1387: [\ve_{\ell m n} \om_s(\hk ) - \ve_{s m n} \om_\ell (\hk )] \\
1388: & & +|\tk|^{-2}\tk_m \om_n(\tk )
1389: [\ve_{\ell m n} \om^*_s(\hk ) - \ve_{s m n} \om^*_\ell (\hk )] \bigg ]. 
1390: \end{eqnarray*}
1391: Calculating the eigenvalues of the Galerkin truncations of this system becomes challenging.
1392: Beyond the size $\{ |k_n| \leq 6, \ n=1,2,3 \}$, the computing time is too long. 
1393: Below we present some pictures for the Galerkin truncation $\{ |k_n| \leq 4 , \ n=1,2,3 \}$.
1394: We choose $m=1$, $A=1.2$, $B=0.7$ and $C=0.9$. When $\e = 20000$, all the eigenvalues 
1395: are negative as shown in Figure \ref{eabc1}. As $\e$ is decreased, eigenvalues start to jump 
1396: off the real axis and form vertical lines as shown in Figure \ref{eabc2} when $\e =10$,
1397: in contrast to the parabolas in the cases of cat's eye and 3D shear. 
1398: When $\e$ is decreased to $\e =0.1$, many eigenvalues move to the right half plane, i.e. there 
1399: are many unstable eigenvalues as shown in Figure \ref{eabc3}. Notice that for the 
1400: full linear NS operator, there should be an infinite tail of negative eigenvalues to the left.
1401: When $\e =0$, the eigenvalues of the Galerkin truncation of linear NS are symmetric with respect 
1402: to the real and imaginary axes as shown in Figure \ref{eabc4}. When $\e =0$, the full linear 
1403: Euler operator at the ABC flow has a continuous spectrum similar to that at the cat's eye 
1404: \cite{Shv06}. That is, the continuous spectrum of the linear Euler at the ABC flow, in any Sobolev space 
1405: $H^s(\mathbb{T}^2)$ where $s$ is a non-negative integer, is a vertical band of width 
1406: $2s \sg$ symmetric with respect to the imaginary axis 
1407: $\{ \la \ : \ |\text{Re}(\la )| \leq s \sg \}$ as shown in Figure \ref{kat7m1}, where 
1408: $\sg >0$ is the largest Liapunov exponent of the vector field given by ABC flow. Thus the width of 
1409: the vertical band is proportional to the scale $s$ of the 
1410: Sobolev space $H^s(\mathbb{T}^2)$ \cite{Shv06}. The union of all such bands for all integers $s \geq 0$ is 
1411: the entire complex plane. The eigenfunctions of the linear NS at the ABC flow  when $\e >0$ 
1412: belong to $H^s(\mathbb{T}^2)$ for all integers $s \geq 0$. As $\e$ is decreased, the eigenvalues 
1413: move into the right half plane. The $\e \ra 0^+$ limiting picture of the eigenvalues of the linear NS 
1414: at the ABC flow is that the 
1415: eigenvalues are dense on the entire plane, in contrast to the left half plane in the case of cat's 
1416: eye as shown in Figure \ref{kat6m1}. That is, all the eigenvalues of the 
1417: linear NS at the ABC flow  condense into the entire plane -- ``condensation''. Thus the possible 
1418: instability hinted by the right half band of the 
1419: continuous spectrum of linear Euler in $H^s(\mathbb{T}^2)$ may be realized by real 
1420: viscous fluids in this case in contrast to the cat's eye case.
1421: 
1422: \begin{figure}
1423: \includegraphics[width=4.0in,height=4.0in]{abc1.eps}
1424: \caption{The eigenvalues of the (Galerkin truncation of) linear NS at the ABC flow when $\e = 20000$.}
1425: \label{eabc1}
1426: \end{figure}
1427: \begin{figure}
1428: \includegraphics[width=4.0in,height=4.0in]{abc2.eps}
1429: \caption{The eigenvalues of the (Galerkin truncation of) linear NS at the ABC flow when $\e = 10$.}
1430: \label{eabc2}
1431: \end{figure}
1432: \begin{figure}
1433: \includegraphics[width=4.0in,height=4.0in]{abc3.eps}
1434: \caption{The eigenvalues of the (Galerkin truncation of) linear NS at the ABC flow when $\e = 0.1$.}
1435: \label{eabc3}
1436: \end{figure}
1437: \begin{figure}
1438: \includegraphics[width=4.0in,height=4.0in]{abc4.eps}
1439: \caption{The eigenvalues of the (Galerkin truncation of) linear NS at the ABC flow when $\e = 0$.}
1440: \label{eabc4}
1441: \end{figure}
1442: 
1443: 
1444: 
1445: 
1446: \section{Numerical Verification of the Heteroclinics Conjecture for 3D Euler Equations}
1447: 
1448: For 3D Euler equations, one can also pose the heteroclinics conjecture.
1449: \begin{itemize}
1450: \item The Heteroclinics Conjecture: There is a heteroclinic orbit of the 3D Euler 
1451: equations that connects $\Om$ (\ref{SABC}) and $-\Om$. 
1452: \end{itemize}
1453: As discussed in the Introduction, the rationality of this conjecture comes from the fact 
1454: that 3D Euler equations have a Lax pair \cite{LY03}.
1455: Below we will verify this conjecture for the Galerkin truncation: $|\k_n| \leq 1$ ($n=1,2,3$)
1456: where $k=(\al \k_1, \be \k_2, k_3)$. Even though this is the smallest Galerkin truncation, 
1457: the dimension of the resulting system is still very large.
1458: For this Galerkin truncation, the fixed point (\ref{SABC})
1459: is still a fixed point. The linearized Galerkin truncation operator at this fixed point 
1460: can be obtained by the corresponding Galerkin truncation the 3D linear Euler operator. 
1461: In this case, the line segment labeled by $\hk = (\al ,0,0)$ ($\e=0$) has a positive eigenvalue 
1462: $\la = 0.5792$, and the corresponding eigenvector $v$ is given by:
1463: \begin{eqnarray*}
1464: & & \om_1 (1,0,-1) = 0.328919 -i \ 0.246347, \
1465: \om_2 (1,0,-1) = -0.246347 -i \ 0.328919 , \\
1466: & & \om_3 (1,0,-1) = 0.230243 -i \ 0.172443  ,
1467: \ \om_1 (1,0,0) = 0  ,  \\
1468: & & \om_2 (1,0,0) = -0.19583 -i \ 0.26147  , \\
1469: & & \om_3 (1,0,0) = 0.183029 -i \ 0.137081  ,  
1470: \ \om_1 (1,0,1) = 0.328919 -i \ 0.246347  , \\
1471: & & \om_2 (1,0,1) = 0.246347 +i \ 0.328919  , 
1472: \ \om_3 (1,0,1) = -0.230243 +i \ 0.172443 , 
1473: \end{eqnarray*}
1474: and all other $\om_\ell (k)$'s are zero. Starting from the initial condition
1475: \begin{equation}
1476: \om = \om^* + 10^{-3} v \ , \label{3IC}
1477: \end{equation}
1478: where $\om^*$ is the Fourier transform of the fixed point (\ref{SABC}), the approximate 
1479: heteroclinic orbit reaches order $\sim 10^{-3}$ neighborhood of $-\om^*$ during the 
1480: time interval [$0, 29.33$]. This approximate heteroclinic orbit is the lower branch of the 
1481: approximate heteroclinic cycle shown in Figure \ref{3gh}.
1482: \begin{figure}
1483: \includegraphics[width=4.0in,height=4.0in]{3figh.eps}
1484: \caption{An approximate heteroclinic cycle of the Galerkin truncation: $|\k_n| \leq 1$ ($n=1,2,3$) 
1485: of the 3D Euler equations.}
1486: \label{3gh}
1487: \end{figure}
1488: Notice that the approximate heteroclinic orbit here does not have the extra loop as in 
1489: Figure \ref{fi1}. When more modes are included in the Galerkin truncation, extra loops may be 
1490: generated. 
1491: 
1492: 
1493: \section{Melnikov Integral and Numerical Simulation of Chaos in 3D Navier-Stokes Equation}
1494: 
1495: 
1496: Setting $\e = 0$ in the 3D NS (\ref{3DNS}), one gets the corresponding 
1497: 3D Euler equation for which one has the following constants of motion:
1498: \[
1499: E=\int_{\mathbb{T}^3} |u|^2dx\ , \quad H=\int_{\mathbb{T}^3} u \cdot \Om dx 
1500: \]
1501: where $E$ is the kinetic energy and $H$ is the helicity. We will use the constant of motion
1502: \[
1503: G=E-H=\int_{\mathbb{T}^3} |u|^2dx - \int_{\mathbb{T}^3} u \cdot \Om dx 
1504: \]
1505: to build a Melnikov integral for the corresponding 3D Navier-Stokes equation (\ref{3DNS}).
1506: We will try to make use of the Melnikov integral as a measure of chaos and to conduct 
1507: a control of chaos, around the 3D shear flow (\ref{SABC}). The gradient of $G$ in $u$ or $\Om$ is given by
1508: \[
1509: \na_uG=2(u-\Om )\ , \quad \na_\Om G =2\text{ curl}^{-1} (u-\Om )\ ,
1510: \]
1511: where curl $= \na \times \ $. The gradient is zero at the 3D shear flow (\ref{SABC}). We define the 
1512: Melnikov function for the 3D NS (\ref{3DNS}) as
1513: \begin{eqnarray*}
1514: M &=& \frac{\al \be}{16 \pi^3}\int_{-\infty}^{+\infty} \int_{\mathbb{T}^3} \na_\Om G 
1515: [\Dl \Om + f(t,x) +b\tdl (x)] dxdt \\
1516: &=& \frac{\al \be}{16 \pi^3}\int_{-\infty}^{+\infty} \int_{\mathbb{T}^3} 2\text{ curl}^{-1} (u-\Om )
1517: [\Dl \Om + f(t,x) +b\tdl (x)] dxdt \ .
1518: \end{eqnarray*}
1519: 
1520: Next we conduct numerical simulations on the Galerkin truncation $|\k_n| \leq 1$ ($n=1,2,3$). 
1521: When $\e =0$, the Liapunov exponent $\sg =0$ for all the numerical tests 
1522: that we run. This indicates that there 
1523: is no chaos when $\e =0$. Often the smallest Galerkin truncation is an integrable system 
1524: \cite{Li03e} \cite{Li06e}. In such a circumstance, the Melnikov integral represents the 
1525: leading order term of the distance between the broken heteroclinic orbit and the 
1526: center-stable manifold of the fixed point. But the dimension of the center-stable manifold is 
1527: large. The zero of the Melnikov integral implies that the unstable manifold in which the broken 
1528: heteroclinic orbit lives, intersects with the center-stable manifold. Therefore, there is a 
1529: new heteroclinic orbit which lives in the intersection. Such a heteroclinic orbit does not 
1530: immediately imply the existence of chaos, even though it may lead to some transient chaos 
1531: characterized by finite time positive Liapunov exponent (infinite time positive Liapunov exponent 
1532: is zero). To compute the Melnikov integral, we choose the external force and control as follows
1533: \begin{eqnarray*}
1534: & & f_1 = a \sin t \cos (x_1 +\al x_2)\ , \quad f_2=f_3=0\ , \\
1535: & & \tdl_1 (x)= \sum_{\k} e^{ik\cdot x} \ , \quad \tdl_2=\tdl_3 = 0 \ ,
1536: \end{eqnarray*}
1537: where the sum is over the Galerkin truncation.
1538: Then the Melnikov integral has the expression
1539: \begin{equation}
1540: M = M_1 + a \sqrt{M_2^2+M_3^2} \sin (t_0 +\th )+bM_4 \ ,
1541: \label{3mel}
1542: \end{equation}
1543: where 
1544: \begin{eqnarray*}
1545: & & \sin \th = \frac{M_3}{\sqrt{M_2^2+M_3^2}} \ , \quad \cos \th = \frac{M_2}{\sqrt{M_2^2+M_3^2}} \ , \\
1546: M_1 &=& -\int_{-\infty}^{+\infty} \sum_k \text{Re} \left \{ i \ve_{\ell mn}k_m (i|k|^{-2} 
1547: \ve_{nsr} k_s \om_r -\om_n)\overline{\om_\ell (k)}\right \} dt \ , \\
1548: M_2 &=& \int_{-\infty}^{+\infty} \cos t \ \text{Re} \left \{ i |k|^{-2}\ve_{1 mn}k_m (i|k|^{-2} 
1549: \ve_{nsr} k_s \om_r -\om_n)\right \}_{k=(\al , 0 ,1)} dt \ , \\
1550: M_3 &=& \int_{-\infty}^{+\infty} \sin t \ \text{Re} \left \{ i |k|^{-2}\ve_{1 mn}k_m (i|k|^{-2} 
1551: \ve_{nsr} k_s \om_r -\om_n)\right \}_{k=(\al , 0 ,1)} dt \ , \\
1552: M_4 &=& \int_{-\infty}^{+\infty} \sum_k \text{Re} \left \{ i |k|^{-2}\ve_{1 mn}k_m (i|k|^{-2} 
1553: \ve_{nsr} k_s \om_r -\om_n)\right \} dt \ , 
1554: \end{eqnarray*}
1555: where the sum is over the Galerkin truncation, all the integrals are evaluated along the 
1556: lower heteroclinic orbit in Figure \ref{3gh} for the time interval [$-29.33/2, 29.33/2$],
1557: rather than ($-\infty , \infty$). Direct numerical computation gives that 
1558: \[
1559: M_1 = 645.7 \ , \quad  M_2 = 1.581\ , \quad M_3= 0 \ , \quad 
1560: M_4= 47.86 \ .
1561: \]
1562: When $b =0$ (no control), $M$ has roots when 
1563: \[
1564: |a| >  408.4 \ .
1565: \]
1566: When $\e = 10^{-5}$, $b= 0$, and $T=4\times 10^4 \pi$, we find that 
1567: \[
1568: \begin{array}{lllll}  a=300 & a=400 & a=420 & a=600 & a=800 \cr 
1569: \sg = 0.8 \times 10^{-5} & \sg = 1.3 \times 10^{-4} & \sg = 0.8 \times 10^{-4} & 
1570: \sg = 4.9 \times 10^{-4}  & \sg = 4.9 \times 10^{-4} \ .\cr 
1571: \end{array}
1572: \]
1573: Around $a=400$, $\sg$ has a jump of one order which seems to be in agreement with the 
1574: Melnikov prediction. However, when $a=100$, $\sg = 3.6 \times 10^{-4}$ which shows that 
1575: the Melnikov prediction is not very effective. We do not know the specific reason. 
1576: 
1577: 
1578: \section{Melnikov Integral and Control of Chaos in 3D Navier-Stokes Equation}
1579: 
1580: Now we turn on the control ($b \neq 0$). When 
1581: \[
1582: b = - M_1/M_4 \approx  -13.5 \ ,
1583: \]
1584: the Melnikov integral $M$ (\ref{3mel}) has roots for any $a$.
1585: 
1586: 
1587: When $\e = 10^{-5}$, $b= -13.5$, and $T=4\times 10^4 \pi$, we find that 
1588: \[
1589: \begin{array}{llll}  a=1 & a=10 & a=100 & a=200 \cr 
1590: \sg = 0.098 & \sg = 0.125 & \sg = 0.095 & 
1591: \sg = 0.083  \ .\cr 
1592: \end{array}
1593: \]
1594: Thus under the control, chaos exists even when $a=1$. Also changing the value of 
1595: $b$, $\sg$ does not change dramatically. Therefore, our control prediction above 
1596: may not be very effective.
1597: 
1598: 
1599: \section{Numerical Verification of the Heteroclinics Conjecture for a Line Model}
1600: 
1601: Returning to the 2D Navier-Stokes equation (\ref{2DNS}), numerical simulations on large 
1602: Galerkin truncations are still challenging to the current computer ability. Here we 
1603: will study a simple line model \cite{Li03e} obtained by a special Galerkin truncation
1604: \cite{Li03e}. Let $p=(1,0)$ and $\hk = (0, \al )$, the line model is given by the 
1605: Galerkin truncation:
1606: \[
1607: \{ \pm p, \ \pm (\hk + n p), \quad \forall n \in \mathbb{Z} \}
1608: \]
1609: We will work in the invariant subspace where $\om_k$'s are real-valued. The governing 
1610: equation of the line model is
1611: \begin{eqnarray}
1612: \dot{\om}_n &=& A_{n-1} \om_* \om_{n-1} - A_{n+1} \om_* \om_{n+1}\non \\
1613: & & +\e [-(n^2+\al^2)\om_n + F_n +b \Dl_n] \ , \label{LM1} \\
1614: \dot{\om}_* &=& -\sum_{n\in \mathbb{Z}}A_{n-1,n} \om_{n-1} \om_n 
1615: +\e [-\om_* + F_* +b \Dl_*]\ , \label{LM2} 
1616: \end{eqnarray}
1617: where $\om_n = \om_{\hk + np}$, $\om_*=\om_p$, similarly for $F$ and $\Dl$ as the Fourier 
1618: transforms of $f$ and $\tdl$, and
1619: \begin{eqnarray*}
1620: & & A_n = 2A(p, \hk + np) = \al \left [ \frac{1}{n^2+\al^2}-1 \right ] \ , \\
1621: & & A_{n-1,n}=2A(\hk +(n-1)p, \hk + np) = \al \left [ \frac{1}{(n-1)^2+\al^2}- 
1622: \frac{1}{n^2+\al^2}\right ] \ .
1623: \end{eqnarray*}
1624: For the line model, verification of the heteroclinics conjecture is relatively easier.
1625: First of all, for the line model ($\e=0$), it can be proved that the fixed point 
1626: $\Om = 2 \cos x_1$ 
1627: has a 1-dimensional local unstable manifold $W^u$. The basic idea of the proof is 
1628: that one can apply the Riesz projections to the spectrum of the linearized line
1629: model operator at the fixed point, and the nonlinear terms have bounded coefficients 
1630: so that they are quadratic in a Banach algebra. For the full 2D Euler equation, 
1631: the difficulty lies at the fact that the nonlinear term is not quadratic in a Banach algebra.
1632: 
1633: Denote by $\Sg$ the 1 co-dimensional hyperplane
1634: \[
1635: \Sg = \left \{ \om \ | \ \om_{(1,0)} = 0 \right \} \ .
1636: \]
1637: We have the corollary of Theorem \ref{hcthm}.
1638: \begin{corollary}
1639: Assume that $W^u \cap \Sg \neq \emptyset$; then the heteroclinics conjecture is true, i.e.  
1640: there is a heteroclinic orbit that connects $\Om = 2 \cos x_1$ and $-\Om$.
1641: \end{corollary}
1642: For any truncation ($|n| \leq N$) of the line model, we first calculate the 
1643: unstable eigenvector. Then we track the heteroclinic orbit with the 
1644: initial condition provided by the unstable eigenvector. 
1645: Numerically exact heteroclinic orbit is obtained for any $N$ ($|n| \leq N$). That is,
1646: for any $N$ ($|n| \leq N$), it can be verified numerically that 
1647: \[
1648: W^u \cap \Sg \neq \emptyset \ .
1649: \]
1650: For $|n| \leq 32$, the heteroclinic orbit is shown in Figure \ref{1L}.
1651: In comparison with the full 2D Euler equation, the hyperplane $\Sg$ here is only 
1652: $1$ co-dimensional. This is the simplest nontrivial case to study the intersection 
1653: $W^u \cap \Sg$.
1654: \begin{figure}
1655: \includegraphics[width=4.0in,height=4.0in]{1L32.eps}
1656: \caption{Numerically exact heteroclinic orbit of the line model ($\e=0$) for $|n| \leq 32$.}
1657: \label{1L}
1658: \end{figure}
1659: We also conduct calculations on the Liapunov exponents. When $\e =0$, $|n| \leq 32$, 
1660: $\sg =0$ for all the computed time intervals, which means that there is no chaos.
1661: This is true for any $N$ ($n \leq N$) and any computational time interval. This 
1662: indicates that the line model may be integrable when $\e =0$. From these facts, it is clear 
1663: that the line model is a good starting point for a rigorous analysis. For instance, it is 
1664: hopeful to make the Melnikov integral theory rigorous. 
1665: 
1666: 
1667: \section{Melnikov Integral and Numerical Simulation of Chaos in the Line Model}
1668: 
1669: For the line model, the kinetic energy and enstrophy are still invariants when $\e=0$. 
1670: Choosing the same external force (\ref{EF}) and control (\ref{SC}), we have the Melnikov 
1671: integral which is the same with that of 2D NS except that the Fourier modes summation is over 
1672: the line model,
1673: \begin{equation}
1674: M_0 = M_1 + a \sqrt{M_2^2+M_3^2} \sin (t_0 +\th )+ bM_c\ .
1675: \label{Lmel}
1676: \end{equation}
1677: For the truncation $|n| \leq 32$, we evaluate these integrals along the heteroclinic orbit
1678: in Figure \ref{1L}, and obtain that
1679: \[
1680: M_1 = -6.0705, \ M_2 = -0.10665, \ M_3 = 0, \ M_c = 11.9728
1681: \]
1682: For the case of no control ($b=0$), when 
1683: \[
1684: |a| > 56.92 
1685: \]
1686: the Melnikov integral $M$ has roots. 
1687: 
1688: We conduct some numerical simulations on 
1689: the (transient) chaos. When $\e = 10^{-3}$, and $T=2\times 10^3 \pi$, we find that 
1690: \[
1691: \begin{array}{ll}  a \leq 200 & a=400 \cr 
1692: \sg < 0 & \sg = 7.2 \times 10^{-4}  \ .\cr 
1693: \end{array}
1694: \]
1695: According to the roots of the Melnikov integral, when $|a| > 56.92$, the broken
1696: heteroclinic orbit may re-intersect with certain center-stable manifold. According 
1697: the above Liapunov exponent result, transient chaos is generated when $a=400$ 
1698: which may be due to the generation of new heteroclinic cycles.
1699: As can be expected, the Melnikov prediction performs better here for the line model.
1700: When $\e =0$, there is no chaos. When $\e > 0$, we see that transient chaos 
1701: appears when $|a|$ is greater than certain threshold which is in the range 
1702: $|a| > 56.92$ predicted by the Melnikov integral.
1703: 
1704: 
1705: \section{Melnikov Integral and Control of Chaos in the Line Model}
1706: 
1707: Now we turn on the control ($b \neq 0$). When 
1708: \[
1709: b=-M_1 /M_c = 0.50702425
1710: \]
1711: the Melnikov integral $M$ has roots for any $a$. To test the effectiveness of the 
1712: control, we set $b$ to the above value and conduct some numerical simulations on 
1713: the (transient) chaos.
1714: When $\e = 10^{-3}$, $b= 0.50702425$, and $T=2\times 10^3 \pi$, we find that 
1715: \[
1716: \begin{array}{lllll}  a =1 & a=10 & a=50 & a=200 & a=400  \cr 
1717: \sg = -12.6 \times 10^{-4}  & \sg = 2 \times 10^{-4} & \sg = 0.2 \times 10^{-4} & 
1718: \sg = 2 \times 10^{-4} & \sg = 0.87 \times 10^{-4} \ .\cr 
1719: \end{array}
1720: \]
1721: The control clearly enhanced the transient chaos. The control effectively pushed 
1722: the threshold of $a$ backward from $400$ to $10$ for the generation of transient chaos.
1723: This shows that the Melnikov integral control performs better here for the line model.
1724: 
1725: 
1726: \section{Numerical Verification of the Heteroclinics Conjecture for a Two Lines Model}
1727: 
1728: To gain an understanding of the effect of the other modes $np$ ($|n|\geq 2$) on the 
1729: line model, we introduce the two lines model which is the Galerkin truncation: 
1730: \[
1731: \{ (k_1,k_2), \quad |k_2|\leq 1 \}.
1732: \]
1733: We also work in the invariant subspace where $\om_k$'s are real-valued. 
1734: 
1735: For the two lines model, one can derive the governing equations in the 
1736: physical variables. Let 
1737: \[
1738: \Om = \om (t, x) + e^{i\al y} q(t,x) + e^{-i\al y} \bq (t,x) \ ,
1739: \]
1740: where $\om$ is real-valued, $q$ is complex-valued (the Fourier transform of $q$ is 
1741: real-valued), and 
1742: \[
1743: \int_0^{2\pi } \om (t, x) dx = 0 \ .
1744: \]
1745: Let
1746: \[
1747: f +b\tdl = \eta (t, x) + e^{i\al y}F(t,x) + e^{-i\al y} \bar{F}(t,x) \ ,
1748: \]
1749: where $\eta$ is real-valued, $F$ is complex-valued (the Fourier transform of $F$ is 
1750: real-valued), and 
1751: \[
1752: \int_0^{2\pi } \eta (t, x) dx = 0 \ .
1753: \]
1754: Substituting the above expressions into the 2D NS (\ref{2DNS}), and 
1755: ignoring the terms involving $e^{i2\al y}$ and $e^{-i2\al y}$, one gets
1756: the two lines model in the physical variables,
1757: \begin{eqnarray}
1758: & & i\pa_t q + \al \left [ (\pa_x \om ) (\pa_x^2 -\al^2)^{-1} 
1759: -(\pa_x^{-1}\om ) \right ] q = i \e \left [ (\pa_x^2 -\al^2) q + 
1760: F \right ] \ ,  \label{me1} \\
1761: & & \pa_t \om + i \al \pa_x \left [ q (\pa_x^2 -\al^2)^{-1} \bq - 
1762: \bq (\pa_x^2 -\al^2)^{-1} q \right ] = \e \left [ \pa_x^2 \om + 
1763: \eta \right ] \ . \label{me2} 
1764: \end{eqnarray}
1765: Introducing $\th = \pa_x^{-1}\om$, $\vphi = (\pa_x^2 -\al^2)^{-1} q$, 
1766: and $h = \pa_x^{-1} \eta$, one gets
1767: \begin{eqnarray}
1768: & & i\pa_t q + \al (\vphi \pa^2_x \th - \th q ) = i \e \left [ 
1769: (\pa_x^2 -\al^2) q + F \right ] \ ,  \label{eme1} \\
1770: & & \pa_t \th + i \al ( q \bar{\vphi} - \bq \vphi ) = \e \left [ 
1771: \pa_x^2 \th + h \right ] \ , \label{eme2} \\
1772: & & \quad (\pa_x^2 -\al^2) \vphi = q \ . \label{eme3}
1773: \end{eqnarray}
1774: When $\e =0$, then Kinetic energy and enstrophy
1775: \[
1776: E_0 = \int_0^{2\pi} \left [\th^2 +2 \al^2 |\vphi|^2 +2|\pa_x \vphi |^2 
1777: \right ] dx \ , \quad E_1 = \int_0^{2\pi} \left [ \om^2 + 2 |q|^2 
1778: \right ] dx \ ,
1779: \]
1780: are still invariants. 
1781: 
1782: Denote by $\Sg$ the hyperplane
1783: \[
1784: \Sg = \left \{ \om \ | \ \om_k = 0 \ , \quad \text{ whenever } k_2=0 \right \} \ .
1785: \]
1786: For the two lines model, when $\e =0$, existence of a local unstable manifold 
1787: for the fixed point $\Om = 2 \cos x_1$ is an open problem due to the fact that the 
1788: coefficients of the nonlinear terms are not bounded, i.e. the nonlinear terms are 
1789: not quadratic in a Banach algebra. Then we have the corollary of Theorem \ref{hcthm}.
1790: \begin{corollary}
1791: Assume that the fixed point $\Om = 2 \cos x_1$ has a 1-dimensional 
1792: local unstable manifold $W^u$, 
1793: and $W^u \cap \Sg \neq \emptyset$; then the heteroclinics conjecture is true, i.e.  
1794: there is a heteroclinic orbit that connects $\Om = 2 \cos x_1$ and $-\Om$.
1795: \end{corollary}
1796: For any truncation ($|k_1| \leq N$) of the two lines model, we first calculate the 
1797: unstable eigenvector. Then we track the heteroclinic orbit with the 
1798: initial condition provided by the unstable eigenvector.
1799: For $|k_1| \leq 2$, numerically exact heteroclinic orbit is obtained. That is,
1800: it can verified numerically that 
1801: \[
1802: W^u \cap \Sg \neq \emptyset \ .
1803: \]
1804: Figure \ref{2L2b} shows the numerically exact heteroclinic orbit. For $|k_1| \leq 4$,
1805: \[
1806: \text{Distance } (W^u, \Sg ) \approx 0.0086 \ .
1807: \]
1808: Figure \ref{2L4b} shows the approximate heteroclinic orbit. For $|k_1| \leq 16$,
1809: \[
1810: \text{Distance } (W^u, \Sg ) \approx 0.012 \ .
1811: \]
1812: Figure \ref{2L16b} shows the corresponding approximate heteroclinic orbit.
1813: Unlike the line model, here we do not always get numerically exact heteroclinic orbits. 
1814: This is due to the influence of the modes ($k_1,0$).
1815: \begin{figure}
1816: \includegraphics[width=4.0in,height=4.0in]{2L2.eps}
1817: \caption{Numerically exact heteroclinic orbit of the two lines model for $|k_1| \leq 2$.}
1818: \label{2L2b}
1819: \end{figure}
1820: \begin{figure}
1821: \includegraphics[width=4.0in,height=4.0in]{2L4.eps}
1822: \caption{Approximate heteroclinic orbit of the two lines model for $|k_1| \leq 4$.}
1823: \label{2L4b}
1824: \end{figure}
1825: \begin{figure}
1826: \includegraphics[width=4.0in,height=4.0in]{2L16.eps}
1827: \caption{Approximate heteroclinic orbit of the two lines model for $|k_1| \leq 16$.}
1828: \label{2L16b}
1829: \end{figure}
1830: 
1831: 
1832: \section{Melnikov Integral and Numerical Simulation of Chaos in the Two Lines Model}
1833: 
1834: When $\e =0$, there is very weak chaos for the computational interval 
1835: $t \in [0, 4\times 10^4 \pi ]$:
1836: \[
1837: \begin{array}{ccc} |n| \leq 4 & |n| \leq 8 & |n| \leq 16 \cr
1838: \sg = 9.5 \times 10^{-4} & \sg = 8.2 \times 10^{-4} & \sg = 8.5 \times 10^{-4} \ .\cr 
1839: \end{array}
1840: \]
1841: 
1842: For the two lines model, the kinetic energy and enstrophy are still invariants when $\e=0$. 
1843: Choosing the same external force (\ref{EF}) and control (\ref{SC}), we have the Melnikov 
1844: integral which is the same with that of 2D NS except that the Fourier modes summation is over 
1845: the two lines model,
1846: \begin{equation}
1847: M_0 = M_1 + a \sqrt{M_2^2+M_3^2} \sin (t_0 +\th )+ bM_c\ .
1848: \label{2Lmel}
1849: \end{equation}
1850: For the truncation $|k_1| \leq 16$, we evaluate these integrals along the heteroclinic orbit
1851: in Figure \ref{2L16b}, and obtain that
1852: \[
1853: M_1 = -4.9\ , \quad M_2 =-0.0948\ , \quad  M_3 = 0 \ , \quad  M_c = 12.2498\ .
1854: \]
1855: For the case of no control ($b=0$), when 
1856: \[
1857: |a| > 51.688 \ , 
1858: \]
1859: the Melnikov integral $M$ has roots. We conduct some numerical simulations on 
1860: the (transient) chaos. When $\e = 10^{-3}$, and $T=2\times 10^3 \pi$, we find that 
1861: \[
1862: \begin{array}{ll}  a \leq 200 & a=400 \cr 
1863: \sg < 0 & \sg = 1.27 \times 10^{-2}  \ .\cr 
1864: \end{array}
1865: \]
1866: According to the roots of the Melnikov integral, when $|a| > 51.688$, the broken
1867: heteroclinic orbit may re-intersect with certain center-stable manifold. According 
1868: the above Liapunov exponent result, strong transient chaos is generated when $a=400$ 
1869: which may be due to the generation of new heteroclinic cycles.
1870: Thus the result is almost the same as that of the line model.
1871: As can be expected, the Melnikov prediction also performs well here for the two lines model.
1872: When $\e =0$, there is weak transient chaos. 
1873: When $\e > 0$, we see that strong transient chaos 
1874: appears when $|a|$ is greater than certain threshold which is in the range 
1875: $|a| > 51.688$ predicted by the Melnikov integral. In comparison with the line model,
1876: here the transient chaos seems very strong. On the other hand, when $\e =0$, the two 
1877: lines model here still has weak chaos, in contrast to the fact of no chaos at all for 
1878: the line model. 
1879: 
1880: \section{Melnikov Integral and Control of Chaos in the Two Lines Model}
1881: 
1882: Now we turn on the control ($b \neq 0$). When 
1883: \[
1884: b=-M_1 /M_c = 0.4 \ ,  
1885: \]
1886: the Melnikov integral $M$ has roots for any $a$. To test the effectiveness of the 
1887: control, we set $b$ to the above value and conduct some numerical simulations on 
1888: the (transient) chaos. When $\e = 10^{-3}$, $b=0.4$, and $T=2\times 10^3 \pi$, 
1889: we find that 
1890: \[
1891: \begin{array}{lllll}  a \leq 400 & a=500 & a=600 & a=700 & a=800 \cr 
1892: \sg < 0 & \sg = 0.97 \times 10^{-3} & \sg = 3.0 \times 10^{-4} & \sg = 4.6 \times 10^{-2} 
1893: & \sg = 5.0 \times 10^{-2}  \ .\cr 
1894: \end{array}
1895: \]
1896: The control seems to tame the transient chaos, in contrast to the line model. The control 
1897: pushed the threshold of $a$ forward from $400$ to $700$ for the generation of strong 
1898: transient chaos. 
1899: 
1900: 
1901: 
1902: \section{Conclusion and Discussion}
1903: 
1904: Through a combination of analytical and numerical studies, we now have a better 
1905: understanding on the zero viscosity limit of the spectra of linear NS operators. 
1906: We proposed and numerically studied the so-called heteroclinics conjecture for both 
1907: 2D and 3D Euler equations. We also proposed the Melnikov integral as a tool for 
1908: predicting and controling chaos.
1909: 
1910: Our numerical verification on the heteroclinics conjecture was limited by 
1911: resonable computing time. We realized that increasing the size of the Galerkin 
1912: truncations can quickly reach the limit of our computer ability. We did not try to utilize 
1913: today's supercomputer due to the fact that Galerkin truncations are essentially 
1914: singular perturbations of Euler equations. We believe that analysis is the key to 
1915: a better understanding of the heteroclinics conjecture.
1916: 
1917: We realized through our numerical simulations that Liapunov exponent performs very 
1918: well as a measure of (even transient) chaos. Our Melnikov prediction for NS equations 
1919: is of course not as rigorous and effective as that for sine-Gordon system. In fact,
1920: it is a rough indicator for predicting and controling chaos. Nevertheless, we 
1921: believe that the Melnikov integral theory for NS equations has a lot of potential 
1922: especially in the current circumstance that there is no effective tools dealing 
1923: with chaos in NS equations.
1924: 
1925: We also believe that both the line and the two lines models have great potential
1926: in future analytical studies on modeling the dynamics of 2D NS equations.
1927: 
1928: 
1929: \section{Appendix A: Melnikov Integral and Control of Chaos in a Sine-Gordon Equation}
1930: 
1931: Consider the sine-Gordon equation \cite{Li04c}
1932: \begin{equation}
1933: u_{tt}=\frac{9}{16} u_{xx} + \sin u +\e \left [ -a u_t + \left ( 1+b\tdl (x) \right )
1934: \cos t \sin^3 u \right ]\ ,
1935: \label{PSG}
1936: \end{equation}
1937: which is subject to periodic boundary condition and odd constraint
1938: \begin{equation}
1939: u(t, x+2\pi ) = u(t, x)\ , \quad u(t, -x) = - u(t, x)\ ,
1940: \label{obc}
1941: \end{equation}
1942: where $u$ is a real-valued function of two real variables ($t,x$), $\e$ is a small 
1943: perturbation parameter, $a> 0$ is the damping coefficient, $b \tdl (x)$ is the spatially 
1944: localized control, $\tdl (x)$ is an even and $2\pi$-periodic function of $x$, and $b$ is 
1945: the control parameter. The system (\ref{PSG}) is invariant under the transform $u \ra -u$.
1946: 
1947: The natural phase space for (\ref{PSG}) is $(u,u_t) \in H^{n+1}\times H^n$ ($n \geq 0$)
1948: where $H^n$ is the Sobolev space on [$0,2\pi$]. Let $P$ be the Poincar\'e period-$2\pi$
1949: map of (\ref{PSG}) in $H^{n+1}\times H^n$. Without the control ($b=0$), we have the 
1950: following chaos theorem \cite{Li04c} \cite{Li05c}.
1951: \begin{theorem}[\cite{Li04c}]
1952: There exists a constant $a_0 >0$, when $\e$ is sufficiently small, for any $a \in \left [ 
1953: \frac{1}{100} a_0, a_0 \right ]$ there exists a symmetric pair of homoclinic orbits 
1954: $h_\pm$ ($h_-=-h_+$) asymptotic to $(u,u_t)=(0,0)$. In the neighborhood of $h_\pm$, there 
1955: exists chaos to the sine-Gordon equation (\ref{PSG}) in the following sense: There is a 
1956: Cantor set $\Xi$ of points in $H^{n+1}\times H^n$ ($n \geq 0$), which is invariant under 
1957: an iterated Poincar\'e period-$2\pi$ $P^K$ for some $K$. The action of $P^K$ on $\Xi$ 
1958: is topologically conjugate to the Bernoulli shift on two symbols $0$ and $1$. 
1959: \label{CT}
1960: \end{theorem}
1961: In the product topology, the Bernoulli shift has the property of sensitive dependence 
1962: upon initial data - the signature of chaos. 
1963: 
1964: When we turn on the control ($b \neq 0$), we hope to find values of $b$ such that the 
1965: chaos in Theorem \ref{CT} is controlled (tamed - annihilated or less chaotic, 
1966: enhanced - more chaotic). Our main tool is the Melnikov function. To build such a function,
1967: we need results from integrable theory. When $\e =0$, the fixed point ($u=0$) of the 
1968: sine-Gordon equation (\ref{PSG}) has a figure eight connecting to it \cite{Li04c}:
1969: \begin{equation}
1970: u = \pm 4 \arctan \left [ \frac{\sqrt{7}}{3} \text{ sech }\tau \sin x \right ] \ , 
1971: \label{f8}
1972: \end{equation}
1973: where $\tau = \frac{\sqrt{7}}{4} (t-t_0)$ and $t_0$ is a real parameter. Along this 
1974: figure eight, a Melnikov vector has the expression \cite{Li04c}:
1975: \begin{equation}
1976: \frac{\pa F_1}{\pa u_t} = \pm \frac{7\pi}{12\sqrt{2}} \text{ sech }\tau \tanh \tau 
1977: \sin x \left [ \frac{9}{16}+\frac{7}{16}\text{ sech}^2\ \tau \sin^2 x \right ]^{-1}\ ,
1978: \label{MV}
1979: \end{equation}
1980: where $F_1$ is a constant of motion. When $\e \neq 0$, the Melnikov function for 
1981: (\ref{PSG}) is given by \cite{Li04c}:
1982: \[
1983: M(t_0,a,b) = \int_{-\infty}^{+\infty} \int_0^{2\pi} \frac{\pa F_1}{\pa u_t} 
1984: \left [ -au_t + \left ( 1+b\tdl (x) \right )\cos t \sin^3 u \right ] dx dt \ ,
1985: \]
1986: where $u$ and $\frac{\pa F_1}{\pa u_t}$ are given in (\ref{f8}) and (\ref{MV}). 
1987: Using the odd and even property of (\ref{f8}) and (\ref{MV}) in $t$ and $x$, we 
1988: obtain
1989: \begin{equation}
1990: M(t_0,a,b) = -a M_a +\sin t_0 \ (M_0 +bM_b)\ ,
1991: \label{MF}
1992: \end{equation}
1993: where 
1994: \begin{eqnarray*}
1995: M_a &=& \int_{-\infty}^{+\infty} \int_0^{2\pi} \frac{\pa F_1}{\pa u_t} u_t \ dx dt \ , \\
1996: M_0 &=& -\int_{-\infty}^{+\infty} \int_0^{2\pi} \frac{\pa F_1}{\pa u_t}\sin 
1997: \frac{4}{\sqrt{7}}\tau \ \sin^3 u \ dx dt \ , \\
1998: M_b &=& -\int_{-\infty}^{+\infty} \int_0^{2\pi} \frac{\pa F_1}{\pa u_t}\tdl (x) \sin 
1999: \frac{4}{\sqrt{7}}\tau \ \sin^3 u \ dx dt \ .
2000: \end{eqnarray*}
2001: In the phase space $H^{n+1}\times H^n$ ($n \geq 0$), $(u,u_t)=(0,0)$ is a saddle 
2002: point under the Poincar\'e period-$2\pi$ map of (\ref{PSG}) with one-dimensional 
2003: unstable manifold $W^u$ and one-codimensional stable manifold $W^s$. The Melnikov function
2004: $\e M(t_0,a,b)$ is the leading order term of the distance between $W^u$ and $W^s$. For the 
2005: entire rigorous theory, see \cite{Li04}. When $|aM_a| < |M_0+bM_b|$, the roots of $M$ 
2006: are given by
2007: \begin{equation}
2008: \sin t_0 = \frac{a M_a}{M_0 +bM_b} \ .
2009: \label{rts}
2010: \end{equation}
2011: Near these roots, $W^u$ and $W^s$ intersect. This leads to the existence of a symmetric 
2012: pair of homoclinic orbits and chaos in Theorem \ref{CT}. When $|aM_a| > |M_0+bM_b|$, 
2013: i.e.
2014: \begin{equation}
2015: -a|M_a| - M_0 <bM_b< a|M_a|- M_0 \ ,
2016: \label{T1}
2017: \end{equation}
2018: the Melnikov function is not zero, and we have the following theorem.
2019: \begin{theorem}
2020: When the control parameter $b$ satisfies (\ref{T1}), the chaos in Theorem \ref{CT} 
2021: disappears.
2022: \label{TT}
2023: \end{theorem}
2024: \begin{proof}
2025: When the  control parameter $b$ satisfies (\ref{T1}), the Melnikov function is not zero 
2026: for any $t_0$, and $W^u$ and $W^s$ do not intersect. Thus the pair of homoclinic orbits
2027: and the corresponding chaos in Theorem \ref{CT} disappear.
2028: \end{proof}
2029: Theorem \ref{TT} only claims that the chaos in Theorem \ref{CT} disappears. This 
2030: does not mean that there is no chaos in the entire phase space $H^{n+1}\times H^n$ 
2031: ($n \geq 0$). An important point here is that by manipulating the localized control 
2032: $b \tdl (x)$, one can change the Melnikov function which leads to the disappearance of 
2033: the non-localized (in $x$) chaos. The control condition (\ref{T1}) is also interesting: 
2034: It is not true that the larger the control parameter $b$ is, the better the taming is. 
2035: In fact, when $b$ is large enough, the chaos will reappear.
2036: 
2037: When $|aM_a| < |M_0+bM_b|$, the Melnikov function (\ref{MF}) has roots (\ref{rts}), and 
2038: Theorem \ref{CT} holds. As a function of $t_0$, the Melnikov function $M$ has the maximal 
2039: absolute value (the $L^\infty$ norm),
2040: \[
2041: M_*(a,b) = a|M_a|+|M_0+bM_b|\ ,
2042: \]
2043: for $t_0 \in [0,2\pi ]$. $M_*$ is the leading order term of the maximal distance between 
2044: $W^u$ and $W^s$. Notice that $W^u$ and $W^s$ intersects near the $t_0$ given by (\ref{rts}).
2045: So the larger the $M_*$ is, the more violent the chaos is. Thus $M_*(a,b)$ serves as a 
2046: measure of the strength of the chaos. By changing the control parameter $b$, we can adjust 
2047: the strength $M_*$ of the chaos - enhancing or decreasing.
2048: 
2049: 
2050: \section{Appendix B: The Lagrange Flow Induced by a Solution to the 2D Euler Equation 
2051: Is Always Integrable}
2052: 
2053: It is well-known that 2D Euler equation is globally well-posed \cite{Kat75} \cite{Kat86}.
2054: For any solution to the 2D Euler equation, let $\Psi = \Psi (t,x_1,x_2)$ be the 
2055: corresponding stream function. Then the Lagrange flow induced by the solution is given by
2056: \begin{equation}
2057: \frac{dx_1}{dt} = - \frac{\pa \Psi}{\pa x_2} \ , \quad
2058: \frac{dx_2}{dt} =  \frac{\pa \Psi}{\pa x_1} \ .
2059: \label{LF}
2060: \end{equation}
2061: \begin{theorem}
2062: The Lagrange flow (\ref{LF}) induced by a solution to the 2D Euler equation is always 
2063: integrable.
2064: \end{theorem}
2065: \begin{proof}
2066: Assume that $\Psi (t,x_1,x_2)$ is not a steady state, i.e. 
2067: it depends upon $t$ (in this case $\Dl \Psi$ is functionally independent of 
2068: $\Psi$, otherwise, $\Psi (t,x_1,x_2)$ would be a steady state).
2069: Introducing the new Hamiltonian $H =\Psi (\th , x_1,x_2)-\psi$, 
2070: and converting (\ref{LF}) into an autonomous system
2071: \begin{equation}
2072: \frac{dx_1}{dt} = - \frac{\pa H}{\pa x_2} \ , \quad
2073: \frac{dx_2}{dt} =  \frac{\pa H}{\pa x_1} \ , \quad 
2074: \frac{d\th}{dt} = - \frac{\pa H}{\pa \psi} \ , \quad
2075: \frac{d\psi}{dt} =  \frac{\pa H}{\pa \th} \ . \quad 
2076: \label{LF1}
2077: \end{equation}
2078: Notice that the vorticity $\Om = \Dl \Psi$ is another constant of motion of (\ref{LF1}) 
2079: besides $H$:
2080: \[
2081: \frac{d}{dt} \Dl \Psi = \pa_\th \Dl \Psi -\pa_{x_1}\Dl \Psi  \pa_{x_2}\Psi 
2082: + \pa_{x_2}\Dl \Psi  \pa_{x_1}\Psi = 0\ .
2083: \] 
2084: Since $\Dl \Psi$ is independent of $\psi$, $\Dl \Psi$ and $H$ are 
2085: functionally independent. Thus (\ref{LF1}) is integrable in the Liouville sense. 
2086: In the case that $\Psi$ is independent of $t$ (i.e. a steady state), then 
2087: (\ref{LF}) is an autonomous system, thus also integrable in the Liouville sense.
2088: \end{proof}
2089: A common way to obtain steady states of 2D Euler equation is by solving 
2090: \[
2091: \Dl \Psi = f(\Psi )
2092: \]
2093: where $f(\Psi )$ is an arbitrary function of $\Psi$, i.e. $\Dl \Psi$ and $\Psi$ are 
2094: functionally dependent.
2095: 
2096: 
2097: 
2098: 
2099: 
2100: \begin{thebibliography}{99}
2101: 
2102: \bibitem{ABT01} N. Alexeeva, et al., Taming spatiotemporal chaos by 
2103: impurities in the parametrically driven nonlinear Schr{\"{o}}dinger
2104: equation, {\it J. Nonlinear Math. Phys.} {\bf 8, suppl.} (2001), 5-12.
2105: 
2106: \bibitem{Dom86} T. Dombre, et al., Chaotic streamlines in the ABC flows,
2107: {\it J. Fluid Mech.} {\bf 167} (1986), 353-391.
2108: 
2109: \bibitem{FS95a} F. Feudel, N. Seehafer, On the bifurcation phenomena in truncations 
2110: of the 2D Navier-Stokes equations, {\it Chaos Solitons Fractals} {\bf 5, no.10} 
2111: (1995), 1805-1816.
2112: 
2113: \bibitem{FS95b} F. Feudel, N. Seehafer, Bifurcations and pattern formation in a 
2114: two-dimensional Navier-Stokes fluid, {\it Phys. Rev. E} {\bf 52, no.4} 
2115: (1995), 3506-3511.
2116: 
2117: \bibitem{Kat75} T. Kato, Quasi-linear equations of evolution, with 
2118: applications to partial differential equations, {\it Lecture Notes in Mathematics}
2119: {\bf 448} (1975), 25-70.
2120: 
2121: \bibitem{Kat86} T. Kato, Remarks on the Euler and Navier-Stokes equations 
2122: in ${\mathbb{R}}^2$, {\it Proc. Symp. Pure Math., Part 2} {\bf 45} (1986), 1-7.
2123: 
2124: \bibitem{KK01} G. Kawahara, S. Kida, Periodic motion embedded in plane 
2125: Couette turbulence: regeneration cycle and burst, {\it J. Fluid Mech.} 
2126: {\bf 449} (2001), 291-300.
2127: 
2128: \bibitem{Kim03} J. Kim, Control of turbulent boundary layers, {\it 
2129: Physics of Fluids} {\bf 15, no.5} (2003), 1093-1105.
2130: 
2131: \bibitem{LLM04} Y. Latushkin, Y. Li, M. Stanislavova, The spectrum of a linearized 
2132: 2D Euler operator, {\it Studies in Appl. Math.} {\bf 112} (2004), 259-270.
2133: 
2134: \bibitem{Li00} Y. Li, On 2D Euler equations: Part I. On the energy-Casimir stabilities
2135: and the spectra for linearized 2D Euler equations,
2136: {\it J. Math. Phys.} {\bf 41, no.2} (2000), 728-758.
2137: 
2138: \bibitem{Li01} Y. Li, A Lax pair for 2D Euler equation, {\it J. Math. Phys.} {\bf 42, No.8}
2139: (2001), 3552-3553.
2140: 
2141: \bibitem{Li03e} Y. Li, On 2D Euler equations: Part II. Lax pairs and homoclinic structures,
2142: {\it Comm. on Appl. Nonlinear Analysis} {\bf 10, no.1} (2003), 1-43.
2143: 
2144: \bibitem{Li03a} Y. Li, Chaos and shadowing lemma for autonomous 
2145: systems of infinite dimensions, {\it J. Dynam. Diff. Eq.}
2146: {\bf 15} (2003), 699-730.
2147: 
2148: \bibitem{Li03b} Y. Li, Chaos and shadowing around a homoclinic tube,
2149: {\it Abstr. Appl. Anal.} {\bf 2003, no. 16} (2003), 923-931.
2150: 
2151: \bibitem{Li04c} Y. Li, Homoclinic tubes and chaos in perturbed 
2152: sine-Gordon equation, {\it Chaos Solitons Fractals} {\bf 20, no. 4}
2153: (2004), 791-798.
2154: 
2155: \bibitem{Li04} Y. Li, {\it Chaos in Partial Differential Equations}, 
2156: International Press, Somerville, MA, USA, (2004).
2157: 
2158: \bibitem{Li05} Y. Li, Invariant manifolds and their zero-viscosity limits
2159: for Navier-Stokes equations, {\it Dynamics of PDE} {\bf 2, no.2} (2005), 159-186.
2160: 
2161: \bibitem{Li05a} Y. Li, Zero dispersion and viscosity limits of invariant 
2162: manifolds for focusing nonlinear Schr{\"{o}}dinger equations, {\it 
2163: J. Math. Anal. Appl.} {\bf 315} (2006), 642-655.
2164: 
2165: \bibitem{Li05c} Y. Li, Chaos and shadowing around a heteroclinically 
2166: tubular cycle with an application to sine-Gordon equation, {\it Studies in 
2167: Appl. Math.} {\bf 116} (2006), 145-171.
2168: 
2169: \bibitem{Li05L} Y. Li, Ergodic isospectral theory of the Lax pairs of Euler
2170: equations with harmonic analysis flavor, {\it Proc. Amer. Math. Soc.} {\bf 133}
2171: (2005), 2681-2687.
2172: 
2173: \bibitem{Li06i} Y. Li, On the true nature of turbulence, {\it The Mathematical 
2174: Intelligencer, in press} (2006).
2175: 
2176: \bibitem{Li06e} Y. Li, Chaos in wave interactions, {\it Intl. J. Bifurc. Chaos,
2177: in press} (2006).
2178: 
2179: \bibitem{LY03} Y. Li, A. Yurov, Lax pairs and Darboux transformations for Euler 
2180: equations, {\it Studies in Appl. Math.} {\bf 111} (2003), 101-113.
2181: 
2182: \bibitem{Lin45} C. Lin, On the stability of two-dimensional parallel flows, 
2183: {\it Quarterly of Appl. Math.} {\bf 3} (1945-1946), 117-142, 218-234, 277-301.
2184: 
2185: \bibitem{LB98} J. Lumley, P. Blossey, Control of turbulence, {\it Annu. Rev. 
2186: Fluid Mech.} {\bf 30} (1998), 311-327.
2187: 
2188: \bibitem{MCH05} D. Molenaar, H. Clercx, G. van Heijst, Transition to chaos
2189: in a confined two-dimensional fluid flow, {\it Phys. Rev. Lett.} {\bf 95}
2190: (2005), 104503. 
2191: 
2192: \bibitem{OGY90} E. Otto, C. Grebogi, J. Yorke, Controlling chaos, {\it Phys.
2193: Rev. Lett.} {\bf 64} (1990), 1196-1199.
2194: 
2195: \bibitem{Shv06} R. Shvydkoy, The essential spectrum of advective equations, 
2196: {\it Preprint} (2006).
2197: 
2198: \bibitem{SL03} R. Shvydkoy, Y. Latushkin, The essential spectrum of the 
2199: linearized 2D Euler operator is a vertical band, {\it Contemp. Math.}
2200: {\bf 327} (2003), 299-304.
2201: 
2202: \bibitem{SYE06} J. Skufca, J. Yorke, B. Eckhardt, Edge of chaos in a parallel 
2203: shear flow, {\it Phys. Rev. Lett.} {\bf 96} (2006), 174101.
2204: 
2205: \bibitem{TOAG05} Y. Tsang, E. Ott, T. Antonsen, P. Guzdar, Intermittency in 
2206: two-dimensional turbulence with drag, {\it Phys. Rev. E.} {\bf 71} (2005), 066313.
2207: 
2208: \end{thebibliography}
2209: 
2210: 
2211: \end{document}
2212: 
2213: %-----------------------------------------------------------------------
2214: % End of article.tex
2215: %-----------------------------------------------------------------------
2216: