math0608475/a5.tex
1: \documentclass[11pt, keyword, address]{article}
2: \usepackage{graphicx, comment}
3: \usepackage{times}
4: \usepackage{amsmath}
5: \usepackage{amssymb}
6: \usepackage{latexsym}
7: \usepackage{amsthm}
8: \newtheorem{theorem}{Theorem}[section]
9: \newtheorem{lemma}[theorem]{Lemma}
10: \newtheorem{proposition}[theorem]{Proposition}
11: \newtheorem{example}[theorem]{Example} 
12: \newtheorem{definition}[theorem]{Definition}
13: \newtheorem{corollary}[theorem]{Corollary}
14: \newtheorem{remark}[theorem]{Remark}
15: \newcommand{\As}{\mathcal{A}_{\mathrm{s}}}
16: \newcommand{\Aw}{\mathcal{A}_{\mathrm{w}}}
17: \newcommand{\A}{\mathcal{A}}
18: \newcommand{\Areg}{\mathcal{A}_{\mathrm{reg}}}
19: \newcommand{\ds}{\mathrm{d}_{\mathrm{s}}}
20: \newcommand{\dw}{\mathrm{d}_{\mathrm{w}}}
21: \newcommand{\dd}{\mathrm{d}}
22: %\newcommand{\Ec}{\mathcal{E}}
23: \newcommand{\Dc}{\mathcal{E}}
24: \newcommand{\Fc}{\mathcal{F}}
25: \newcommand{\s}{\mathrm{s}}
26: \newcommand{\w}{\mathrm{w}}
27: \newcommand{\ww}{\omega_{\mathrm{w}}}
28: \newcommand{\ws}{\omega_{\mathrm{s}}}
29: \newcommand{\Xw}{X_{\mathrm{w}}}
30: \newcommand{\Hw}{H_{\mathrm{w}}}
31: \newcommand{\Ab}{\mathcal{A}_{\bullet}}
32: \newcommand{\db}{\mathrm{d}_{\bullet}}
33: \newcommand{\I}{\mathcal{I}}
34: \newcommand{\Bb}{B_{\bullet}}
35: \newcommand{\wb}{{\omega}_{\bullet}}
36: \newcommand{\ddt}{\frac{d}{dt}}
37: \newcommand{\glim}{\widetilde{\lim} \, }
38: \newcommand{\ts}{\textstyle}
39: \def\setmarsing{
40: \oddsidemargin-0in
41: \evensidemargin-0in
42: \textwidth5.5in
43: \textheight8in
44: }
45: \setmarsing
46: 
47: \title{On global attractors of the 3D Navier-Stokes equations}
48: \author{A. Cheskidov\thanks{University of Michigan, Department of Mathematics,
49: Ann Arbor, MI 48109 (acheskid@umich.edu)} \ and
50: C. Foias\thanks{Texas A\&M University, Department of Mathematics,
51: College Station, TX 77843}}
52: \begin{document}
53: \maketitle
54: \abstract{
55: In view of the possibility that the 3D Navier-Stokes equations (NSE)  might
56: not always have regular solutions, we introduce an abstract
57: framework for studying the asymptotic behavior
58: of multi-valued dissipative evolutionary systems with respect to two
59: topologies---weak and strong. Each such system
60: possesses a global attractor in the weak topology, but not necessarily
61: in the strong. In case the latter exists and is weakly closed, it coincides
62: with the weak global attractor. We give a sufficient condition for
63: the existence of the strong global attractor, which is verified for the
64: 3D NSE when all solutions on the weak global attractor are strongly continuous.
65: We also introduce and study a two-parameter family of models for the
66: Navier-Stokes equations, with similar properties and open problems.
67: These models always possess weak global attractors, but on some of them
68: every solution blows up (in a norm stronger than the standard energy one) in finite time.
69: \\
70: \\
71: {\bf Keywords:} Navier-Stokes equations, global attractor, blow-up in finite time.
72: }
73: 
74: 
75: 
76: \section{Introduction}
77: A remarkable feature of many dissipative partial differential equations (PDEs)
78: is the existence of a global attractor to which all the solutions converge as
79: time goes to infinity \cite{SY,T2}. The global attractor $\A$
80: is the minimal closed set in
81: a phase space $H$ (i.e., the functional space, usually a Banach space,
82: in which the solutions exist) that uniformly attracts the trajectories starting
83: from any a priori given
84: bounded set in $H$. When the topology on $H$ is referred as the strong
85: (weak) topology of $H$, we will call $\A$ the strong (respectively weak)
86: global attractor.
87: 
88: It is possible that a dissipative PDE does not have a strong global attractor.
89: For instance, the 2D Navier-Stokes equations (NSE) on a bounded domain
90: $\Omega \subset \mathbb{R}$, when supplemented with appropriate boundary
91: conditions, possess a strong global attractor in $H$ (a certain subspace of
92: $L^2(\Omega)^3$) \cite{L,T2}, but it is not yet known
93: whether this holds for the 3D NSE.
94: 
95: Nevertheless, even for the 3D NSE one can prove that there exists
96: %a minimal weakly closed set in $H$ for which the attracting property holds
97: %with respect to the weak topology in $H$
98: a weak global attractor \cite{FT85}.
99: % Therefore, this attractor is referred to as the weak global attractor.
100: %For convenience, the first attractor mentioned above will be called the strong
101: %global attractor.
102: When the strong global attractor is strongly compact in $H$ (e.g., in the 2D NSE),
103: then it is also the weak global attractor. But, in any case, the weak global
104: attractor is an
105: appropriate generalization of the strong global attractor since it captures
106: the long-time behavior of the solutions. In particular, the support of any
107: time-average measure of the 3D NSE is included in the weak global attractor
108: (see \cite{FMRT}).
109: One should note that Sell \cite{S} introduced a related notion of a trajectory
110: attractor $\mathfrak{A}$ in the space of all trajectories, which was further
111: studied in \cite{CV, FS, S, SY}. The weak global attractor coincides with the
112: set of values of all trajectories in $\mathfrak{A}$ at any fixed time $t$.
113: 
114: The aim of this study is to present a general abstract framework which
115: is applicable to the 3D NSE even in the case where they do not
116: possess a strong global attractor. This framework may be also useful in
117: the study of other PDEs for which the existence of the strong global
118: attractor is in limbo. This aim forces us to consider multi-valued
119: evolutionary systems.
120: A number of papers have
121: been published concerning attractors of multi-valued semiflows. See
122: \cite{CMR} for a comparison of two canonical abstract frameworks by
123: Melnik and Valero \cite{MV} and Ball \cite{B1}. The main difference between
124: our evolutionary system $\Dc$ and Ball's generalized semiflow is that
125: we do not include
126: the hypotheses of concatenation and upper semicontinuity  with respect to
127: the initial data. This allows us to consider an evolutionary system whose
128: trajectories are all Leray-Hopf weak solutions of the 3D NSE.
129: 
130: Our definition of the evolutionary system $\Dc$ already exploits the effect of
131: dissipativity, namely the existence of an absorbing ball.
132: In fact, the space $X$ in which the trajectories of $\Dc$ live is, in applications,
133: precisely such an absorbing ball. Since in most applications the phase
134: space is a separable reflexive Banach space, both the strong and the
135: weak topologies on $X$ are metrizable. This is the motivation for us
136: to define strong and weak topologies on $X$ to be the ones induced by
137: appropriate metrics.
138: 
139:  We show
140: that every evolutionary system always possesses a weak global
141: attractor;  moreover, if the strong global attractor exists and is weakly closed,
142: then it has to coincide with the weak global attractor.
143: Note that some classical definitions (see, e.g., \cite{T2}) require a global
144: attractor to be an invariant set. We will see that under
145: a condition, which is, for example, satisfied by the Leray-Hopf weak solutions
146: of the 3D NSE, the weak global attractor is also the maximal bounded
147: invariant set. We recall that those solutions are always weakly continuous
148: in $L^2(\Omega)^3$.
149: 
150: It is known that if a weak global attractor for the 3D NSE is bounded in $V$,
151: then it is in fact strong \cite{FT85}. Moreover,
152: Ball \cite{B1} showed that if a generalized semiflow for a dissipative evolutionary
153: system is  asymptotically compact, then a strong global attractor exists.
154: This generalizes corresponding results for semiflows (see {\cite{H,HLS,L1}})
155: and implies that the strong global attractor for the 3D NSE exists
156: under the condition that all weak solutions are strongly continuous from $(0, \infty)$
157: to $L^2(\Omega)^3$ (see \cite{B1}). In this paper we show that even without the assumptions of concatenation and upper
158: semicontinuity  with respect to the initial data, the asymptotic compactness implies
159: that the weak global attractor is the minimal compact attracting set in the strong
160: metric, i.e., the weak global attractor is in fact the strong compact global attractor
161: (Theorem~\ref{t:asymptoticcompact}).
162: Applied to the 3D NSE, this result implies the existence of a strong compact global
163: attractor in the case when the solutions on the weak global attractor are
164: continuous in $L^2(\Omega)^3$ (Theorem~\ref{thm:main}).
165: 
166: The convergence of Leray-Hopf weak solutions was also studied by Rosa \cite{R}, 
167: namely, he introduced an asymptotic regularity condition
168: to insure the strong convergence of a weak solution towards its weak
169: $\omega$-limit. This condition requires the limit solutions to be strongly 
170: continuous in $L^2(\Omega)^3$, and implies that the weak global attractor is a
171: strongly compact strongly attracting set if the weak solutions on the weak global
172: attractor are strongly continuous in $L^2(\Omega)^3$.
173: Moreover, since a trajectory of the 3D NSE that is not strongly continuous
174: in  $L^2(\Omega)^3$ also, obviously, is not relatively strongly compact in $L^2(\Omega)^3$,
175: the strong continuity of weak solutions on the weak global
176: attractor $\Aw$ is also a necessary condition for $\Aw$ to be strongly compact
177: (see \cite{R}).
178: 
179: Recall that we define a global attractor as the minimal closed (uniformly)
180: attracting set in the corresponding metric, and hence, allow the possibility
181: for the solutions on a strong global attractor to be discontinuous.
182: We address this issue by
183: studying the weak convergence of weak solutions towards a weak solution strongly continuous from the right in $L^2(\Omega)^3$.  
184: We show that the weak convergence is strong under the
185: condition that the "energy jumps" of solutions converge to the "energy jumps"
186: of the limit solution (Theorem~\ref{t:deltat}). However, at this stage, no
187: necessary condition for the strong convergence is known.
188: 
189: 
190: Finally, we provide an example of a dissipative evolutionary system for
191: which all solutions on the weak global attractor blow-up in finite time.
192: We introduce a two-parameter family of simple infinite-dimensional
193: systems of differential equations.
194: These systems, called {\it tridiagonal models for the NSE (TNS models)},
195: display basic features of the NSE.
196: In particular, they are examples of dissipative systems that possess a weak
197: global attractor, but the existence of a strong global attractor is not
198: known.
199: 
200: TNS models have similar form and some similar properties to shell
201: models, specifically dyadic models studied in \cite{FP,KP}.
202: Moreover, a similar analysis of the dyadic models also results in a finite
203: time  blow-up (see \cite{C}).
204: In the TNS models though, the coefficients in the equations are chosen
205: to yield NSE-like scaling properties.
206: More precisely, we first mimic the Stokes operator in 3D via
207: the choice a positive definite operator on $l^2$ whose eigenvalues grow with the same speed as the eigenvalues of the Stokes operator. 
208: Second, we mimic the nonlinear term of the NSE via the choice of a bilinear
209: form, which scales like the Sobolev estimate for the NSE.
210: Then we obtain the following system of differential equations:
211: \begin{equation} \label{model-intro}
212: \ddt u + \nu Au + B(u,u) = g,
213: \end{equation}
214: where $u=(u_1,u_2,\dots)$,
215: \[
216: (Au)_n = n^{\alpha} u_n,
217: \]
218: and
219: \[
220: (B(u,v))_n=- n^\beta u_{n-1} v_{n-1} + (n+1)^\beta u_n v_{n+1},
221: \qquad n =1, 2, \dots,
222: \]
223: with $u_0=0$.
224: Here $\alpha$ and $\beta$ are two positive parameters. Note that
225: the orthogonality property in $l^2$ holds for $B$, which implies the existence
226: of an absorbing ball and a weak global attractor. Moreover, when
227: $\alpha = 2/3$, which corresponds to the speed with which the eigenvalues
228: of the Stokes operator grow in three-dimensional space, and $\beta=11/6$,
229: we have the following sharp estimate:
230: \[
231: | (B(u,u),Au) | \lesssim |Au |^{3/2} |A^{1/2}u|^{3/2},
232: \]
233: where
234: $
235: |v|^2=\sum v_n^2. 
236: $
237: This estimate is exactly the same as the the estimate based on the Sobolev
238: inequalities for the nonlinear term of the 3D NSE, with $|\cdot|$ being the $L^2$-norm.
239: It is an open question whether solutions of (\ref{model-intro})
240: can lose regularity in the case $(\alpha,\beta)=(2/3,11/6)$. However,
241: we show that in the nonviscous case $\nu=0$, for every $\alpha>0$, $\beta>0$,
242: $\gamma>0$,
243: and $g_n\geq 0$ for all $n\in\mathbb{N}$, the norm
244: $|A^{(\beta+\gamma-1)/(3\alpha)} u|$ of every solution
245: with $u_n(0)\geq 0$ and $u(0)\ne0$
246: blows up (Theorem~\ref{t:Eblow}).
247: 
248: When the viscosity is not zero, the model always possesses a weak global
249: attractor $\Aw$. Moreover, we prove that if the force $g$ is
250: large enough, then all solutions on $\Aw$  blow up in finite time in
251: an appropriate norm when $2\beta>3\alpha+3$ (Remark~\ref{rem:bl}).
252: The question whether $\Aw$ is a strong global attractor 
253: remains open in the case where $\beta > \alpha+1$.
254: 
255: 
256: 
257: \section{Evolutionary system and global attractors}
258: 
259: Let $(X,\ds(\cdot,\cdot))$ be a metric space endowed with
260: a metric $\ds$, which will be referred to as a strong metric.
261: Let $\dw(\cdot, \cdot)$ be another metric on $X$ satisfying
262: the following conditions:
263: \begin{enumerate}
264: \item $X$ is $\dw$-compact.
265: \item If $\ds(u_n, v_n) \to 0$ as $n \to \infty$ for some
266: $u_n, v_n \in X$, then $\dw(u_n, v_n) \to 0$ as $n \to \infty$,
267: that is, the identity map $(X, \ds) \mapsto (X,\dw)$ is uniformly continuous.
268: \end{enumerate}
269: Due to the latter property, $\dw$ will be referred to as the weak metric on $X$.
270: Note that any strongly compact ($\ds$-compact) set is weakly compact
271: ($\dw$-compact), and any weakly closed set is strongly closed.
272: Also it will be convenient to denote by $\overline{A}^{\bullet}$ the
273: closure of the set $A\subset X$ in the topology generated by $\db$; here and
274: throughout $\bullet$ stands for either $\mathrm{s}$ or $\mathrm{w}$.
275: 
276: To define an evolutionary system, first let
277: \[
278: \mathcal{T} := \{ I: \ I=[T,\infty) \mbox{ for some } T\in \mathbb{R}, \mbox{ or } 
279: I=(-\infty, \infty) \},
280: \]
281: and for each $I \subset \mathcal{T}$ let $\mathcal{F}(I)$ denote
282: the set of all $X$-valued functions on $I$.
283: Now we define an evolutionary system $\Dc$ as follows.
284: \begin{definition} \label{d:E}
285: A map $\Dc$ that associates to each $I\in \mathcal{T}$ a subset
286: $\Dc(I) \subset \mathcal{F}$ will be called an evolutionary system if
287: the following conditions are satisfied:
288: \begin{enumerate}
289: \item $\Dc([0,\infty))\ne \emptyset$.
290: \item
291: $\Dc(I+s)=\{u(\cdot): \ u(\cdot -s) \in \Dc(I) \}$ for
292: all $s \in \mathbb{R}$.
293: \item $\{u(\cdot)|_{I_2} : u(\cdot) \in \Dc(I_1)\}
294: \subset \Dc(I_2)$ for all
295: pairs of $I_1,I_2 \in \mathcal{T}$, such that $I_2 \subset I_1$.
296: \item
297: $\Dc((-\infty , \infty)) = \{u(\cdot) : \ u(\cdot)|_{[T,\infty)}
298: \in \Dc([T, \infty)) \ \forall T \in \mathbb{R} \}.$
299: \end{enumerate}
300: \end{definition}
301: We will refer to $\Dc(I)$ as the set of all trajectories (solutions)
302: on the time interval $I$. Let $P(X)$ be the set of all subsets of $X$.
303: For every $t \geq 0$ define a map
304: \begin{eqnarray*}
305: &R(t):P(X) \to P(X),&\\
306: &R(t)K := \{u(t): u(0)\in K, u(\cdot) \in \Dc([0,\infty))\}, \qquad
307: K \subset X.&
308: \end{eqnarray*}
309: Note that the assumptions on $\Dc$ imply that $R(t)$ enjoys
310: the following property:
311: \begin{equation} \label{eq:propR(T)}
312: R(t+s)K \subset R(t)R(s)K, \qquad K \subset X,\quad t,s \geq 0.
313: \end{equation}
314: 
315: Let us first point out that the trajectories are not required to be continuous
316: (even with respect to $\dw$) nor are they uniquely determined by their
317: starting points, i.e.,  it is possible to have two different
318: trajectories $u,v \in \Dc([0,\infty))$, such that $u(0)=v(0)$. Second, there is no assumption of concatenation. If $u \in \Dc([0,\infty))$ and
319: $v \in \Dc([T,\infty))$ for some $T>0$, so that
320: $u(T)=v(T)$, then the following function
321: \[
322: w(t)= \left\{
323: \begin{aligned}
324: u(t), \qquad &\mbox{if } t\in[0,T],\\
325: v(t), \qquad &\mbox{if } t\in(T,\infty).
326: \end{aligned}
327: \right.
328: \]
329: need not be in $\Dc([0,\infty))$.
330: We avoid the assumptions of the continuity,
331: uniqueness, and concatenation in order to be able to consider
332: an evolutionary system consisting of Leray-Hopf weak solutions to the
333: 3D Navier-Stokes equations.
334: 
335: Often an evolution of a dynamical system can be described by
336: a semigroup of continuous mappings acting on some metric space $H$:
337: \[
338: S(t): H \to H, \quad t \geq 0.
339: \]
340: The semigroup properties are the following:
341: \begin{equation} \label{eq:semigroup}
342: S(t+s) = S(t)S(s), \quad t,s \geq 0, \qquad S(0)= \mbox{Identity operator.}
343: \end{equation}
344: Then, if $u(t)\in H$ represents a state of the dynamical system at 
345: time $t$, we have
346: \[
347: u(t+s) = S(t)u(s), \qquad t,s \geq 0.
348: \]
349: A ball $B\subset H$ is called an absorbing ball, if for any bounded
350: set $K \subset H$ there exists $t_0$, such that
351: \[
352: S(t)K \subset B, \qquad \forall t \geq t_0.
353: \]
354: Hence, if we are interested in the long-time behavior of the dynamical system,
355: it is enough to consider a restriction of the system to an absorbing ball.
356: So, we let $X$ be a closed absorbing ball, and
357: define the map $\Dc$
358: in the following way:
359: \[
360: \Dc(I) := \{u(\cdot): u(t+s)=S(t)u(s) \mbox{ and }
361: u(s) \in X \ \forall s\in I, \ t \geq 0 \}.
362: \]
363: Note that conditions 1--4 for the evolutionary system $\Dc$ automatically follow
364: from the semigroup properties (\ref{eq:semigroup}) of $S(t)$.
365: In addition, let $T$ be such that
366: \[
367: S(t)X \subset X \qquad \forall t \geq T.
368: \]
369: Then we have
370: \[
371: R(t)K = S(t)K, \qquad \forall K \subset S(T)X, \ t \geq 0.
372: \]
373: The 3D Navier-Stokes equations will serve as an instructive illustration and 
374: application  of our consideration. 
375: As yet in our knlowledge of the 3D NSE, the time evolution of the 3D NSE cannot
376: be described by a semigroup of maps. Therefore, for the 3D NSE we will have a
377: more involved definition of $\Dc$ (see Section~\ref{3DNSE}).
378: 
379: 
380: Having defined the evolutionary system $\Dc$, we proceed to define attracting
381: sets and global attractors.
382: For $A \subset X$ and $r>0$, denote
383: $
384: B_{\bullet}(A,r) = \{u: \ \db(u, A) < r\},
385: $
386: where
387: \[
388: \db(u, A):=\inf_{x\in A}\db(u,x).
389: \]
390: 
391: 
392: \begin{definition}
393: A set $A \subset X$ is a $\mathrm{d}_{\bullet}$-attracting set
394: ($\bullet = \mathrm{s,w}$) if it uniformly
395: attracts $X$ in $\mathrm{d}_{\bullet}$-metric, i.e.,
396: for any $\epsilon>0$, there exists
397: $t_0$, such that
398: \[
399: R(t)X \subset B_{\bullet}(A, \epsilon), \qquad \forall t \geq t_0.
400: \]
401: \end{definition}
402: 
403: \begin{definition}
404: $\mathcal{A}_{\bullet}\subset X$ is a
405: $\db$-global attractor ($\bullet = \mathrm{s,w}$) if
406: $\mathcal{A}_{\bullet}$ is a minimal $\db$-closed
407: $\db$-attracting  set, i.e., $\mathcal{A}_{\bullet}$
408: is $\db$-closed $\db$-attracting
409: and every subset $A \subset \Ab$ that is also $\db$-closed and $\db$-attracting
410: satisfies $A=\Ab$.
411: \end{definition}
412: Note that the empty set is never an attracting set.
413: Note also that since $X$ is not strongly compact, the intersection of two 
414: $\ds$-closed $\ds$-attracting sets might not be $\ds$-attracting.
415: Nevertheless, the uniqueness of a global attractor is a direct consequence
416: of the following lemma.
417: 
418: 
419: \begin{lemma}
420: If $\mathcal{A}_{\bullet}$ exists and $A$ is a $\db$-closed $\db$-attracting set,
421: then $\Ab \subset A$ ($\bullet = \mathrm{s,w}$).
422: \end{lemma}
423: \begin{proof}
424: Let $A$ be an arbitrary $\db$-closed $\db$-attracting set.
425: Take any point $a \in \Ab$. Let $\epsilon>0$. If there exists $t_\epsilon>0$, such that 
426: \[
427: R(t)X \cap \Bb(a,\epsilon) = \emptyset, \qquad \forall t\geq t_\epsilon,
428: \]
429: then $\Ab \setminus \Bb(a,\epsilon/2)$ is a $\db$-closed $\db$-attracting
430: set contradicting the minimality of $\Ab$. 
431: So, there exists a sequence $t_n \to \infty$ as $n\to \infty$,
432: such that
433: \[
434: R(t_n)X \cap \Bb(a,\epsilon) \ne \emptyset, \qquad \forall n.
435: \]
436: On the other hand, since $A$ is $\db$-attracting, we infer that
437: \[
438: R(t_n)X \subset \Bb(A,\epsilon),
439: \]
440: for $n$ large enough. It follows that
441: \[
442: A \cap \Bb(a,2\epsilon) \ne \emptyset.
443: \]
444: Since $A$ is $\db$-closed, we have that $a \in A$.  Thus,
445: $\Ab \subset A$.
446: \end{proof}
447: As a direct consequence of this lemma we have the following.
448: 
449: \begin{corollary}
450: If $\mathcal{A}_{\bullet}$ exists, then it is unique
451: ($\bullet = \mathrm{s,w}$). 
452: \end{corollary}
453: Assume now that the weak global attractor $\Aw$ exists. If $\Aw$ is
454: a strongly attracting set, does it follow that the strong global attractor
455: exists? In general, this may not be true. However, if the strong global attractor
456: exists, then clearly it is a $\dw$-attracting set. Moreover, we have the following.
457: 
458: \begin{theorem} \label{thm:coinside}
459: If $\As$ exists, then $\Aw$ exists and
460: \[
461: \Aw=\overline{\As}^{\w}.
462: \]
463: \end{theorem}
464: \begin{proof}
465: If there exists
466: a $\dw$-closed $\dw$-attracting set $A\subset \overline{\As}^{\mathrm{w}}$ and
467: $A\ne \overline{\As}^{\mathrm{w}}$, then there exists
468: $u_0 \in \As$, such that
469: \[
470: d=\dw(u_0, A)>0.
471: \]
472: By the definition of an attracting set, there exists a time $t_0>0$, such that
473: \begin{equation} \label{e:111}
474: R(t)X \subset B_{\mathrm{w}}(A, d/2) \qquad \forall t\geq t_0.
475: \end{equation}
476: Note that
477: \[
478: \dw(u_0, B_{\mathrm{w}}(A, d/2)) \geq d/2.
479: \]
480: Therefore, by virtue of Property 2 in the definition of $\dw$, there
481: exists $\delta>0$, such that
482: \[
483: \ds(u_0, B_{\mathrm{w}}(A, d/2)) > \delta,
484: \]
485: whence,
486: \[
487: B_{\mathrm{s}}(u_0, \delta) \cap B_{\mathrm{w}}(A, d/2) = \emptyset.
488: \]
489: Now from \eqref{e:111} it follows that
490: \[
491: B_{\mathrm{s}}(u_0, \delta)\cap R(t)X =\emptyset \qquad \forall t \geq t_0.
492: \]
493: Consequently, $\As \setminus B_{\mathrm{s}}(u_0, \delta/2)$ is a $\ds$-closed 
494: $\ds$-attracting set strictly included in $\As$, a contradiction.
495: Hence, $\overline{\As}^{\mathrm{w}}$ is the weak global attractor.
496: \end{proof}
497: 
498: 
499: The following are two simple examples of evolutionary systems
500: that possess a weak global attractor $\Aw$, but not a strong global
501: attractor $\As$.
502: 
503: \noindent
504: \textbf{Example 1.}
505: Let
506: \[
507: X=\left\{ u \in L^2(-\infty,\infty): \int_{-\infty}^\infty u(x)^2 \, dx \leq 1,
508: u(x)=0 \mbox{ for } x> 0\right\},
509: \]
510: and define on $X$ the distances
511: \[
512: \ds(u,v):=\left( \int_{-\infty}^\infty(u(x)-v(x))^2 \, dx\right)^{1/2}, \qquad
513: \dw(u,v)=\int_{-\infty}^\infty \frac{1}{2^{|x|}}
514: \frac{|u(x)-v(x)|}{1 + |u(x)-v(x)|} \, dx.
515: \]
516: Consider the following partial differential equation:
517: \[
518: \frac{\partial u}{\partial t} = \frac{\partial u}{\partial x}.
519: \]
520: The trajectories of the evolutionary system $\Dc$
521: will be solutions of this equation, i.e.,
522: \[
523: \begin{aligned}
524: &\Dc([s,\infty)) =\{u\in \Fc([s,\infty)) : u(t) =u_0(\cdot+t-s), t\in[s,\infty), u_0 \in X\}, \qquad
525: \forall s \in \mathbb{R},\\
526: & \Dc ((-\infty,\infty))=\{0\}.
527: \end{aligned}
528: \]
529: Then it is easy to show that $\Aw=\{0\}$ (see also Theorem~\ref{c:weakA}). However, no trajectory except the trivial one $u=0$ strongly
530: converges to $0$ as $t\to \infty$.
531: 
532: \noindent
533: \textbf{Example 2.}
534: Take
535: \[
536: X=\left\{ u\in l^2: \sum_{n=1}^\infty u_n^2  \leq 1\right\}.
537: \]
538: where $u=(u_n)$, $u_n \in \mathbb{R}$ for all $n$.  For $u, v \in X$, let
539: \[
540: \ds(u,v):=\left( \sum_{n=1}^\infty(u_n-v_n)^2 \right)^{1/2}, \qquad
541: \dw(u,v)= \sum_{n=1}^\infty \frac{1}{2^n}
542: \frac{|u_n-v_n|}{1 + |u_n-v_n|}.
543: \]
544: Consider the following differential equation:
545: \[
546: \ddt u_n = -\frac{1}{n} u_n, \qquad n \in \mathbb{N}
547: \]
548: The trajectories of the evolutionary system $\Dc$
549: will be solutions of this equation, i.e.,
550: \[
551: \Dc([s,\infty)) =\{u\in\Fc([s,\infty)): (u_n(t))=(u_n^0e^{(s-t)/n}), t\in[s,\infty),
552: u^0 \in X\},
553: \]
554: for $s \in \mathbb{R}$, and
555: \[
556: \Dc((-\infty,\infty))=\{0\}.
557: \]
558: Take any $u^0 \in X$ and consider the trajectory $u \in \Dc([0,\infty))$ starting
559: at $u^0$, i.e., $u(0)=u^0$. Then we have
560: \[
561: \ds(u(t),0)^2=\sum_{n=1}^\infty u_n(t)^2 = \sum_{n=1}^\infty(u^0_n)^2 e^{-\frac{2t}{n}}
562: \to 0 \qquad \mbox{as} \qquad t \to \infty.
563: \]
564: However, the convergence is not uniform in the $\ds$-metric,
565: although it is uniform in the $\dw$-metric. So
566: again $\Aw=\{0\}$, but $\As$ does not exist.
567: 
568: Note that the nonexistence of $\As$ in the two examples is due to
569: two different behaviors of the trajectories.
570: In the first example all the nontrivial trajectories converge
571: to $\Aw$ weakly, but not strongly. In the second example all the nontrivial
572: trajectories converge to $\Aw$ strongly, but not uniformly.
573: 
574: \begin{definition} \label{d:ucomp}
575: The map $R(t)$ is uniformly $\db$-compact
576: ($\bullet= \mathrm{s, w}$) if
577: there exists $t_0\geq 0$, such that
578: \[
579: \bigcup_{t \geq t_0} R(t) X
580: \]
581: is relatively $\db$-compact.
582: \end{definition}
583: 
584: Note that since $X$ is $\dw$-compact, 
585: $R(t)$ is automatically uniformly $\dw$-compact.
586: 
587: \begin{definition} The $\wb$-limit ($\bullet= \mathrm{s, w}$)
588: of a set $K \subset X$ is
589: \[
590: \wb (K):=\bigcap_{T\geq 0} \overline{\bigcup_{t \geq T} R(t) K}^{\bullet},
591: \]
592: where the closure is taken in $\db$-metric.
593: \end{definition}
594: 
595: \begin{lemma} \label{l:wbinA}
596: Let $A$ be a $\db$-closed $\db$-attracting set. Then
597: \[
598: \wb(X) \subset A.
599: \]
600: \end{lemma}
601: \begin{proof}
602: Suppose that there exists $a \in \wb(X) \setminus A$.
603: Since $A$ is $\db$-closed, there exists $\epsilon>0$, such that
604: \[
605: A \cap \Bb(a,\epsilon) = \emptyset.
606: \]
607: By the definition of the $\wb$-limit,
608: there exist a sequence $t_n \to \infty$ as $n \to \infty$ and
609: a sequence $x_n \in R(t_n)X$, such that $\db(x_n,a) \to 0$
610: as $n \to \infty$. Hence, there exists $N>0$, such that 
611: \[
612: x_n \notin \Bb(A, \epsilon/2), \qquad \forall n\geq N.
613: \]
614: This means that $A$ is not $\db$-attracting, a contradiction.
615: \end{proof}
616: 
617: \begin{lemma} \label{l:oncompactsemigroup}
618: If the map $R(t)$ is uniformly $\db$-compact
619: ($\bullet= \mathrm{s, w}$), then 
620: $\wb(X)$ is a nonempty $\db$-compact $\db$-attracting set.
621: \end{lemma}
622: \begin{proof}
623: By Definition~\ref{d:ucomp} and the fact that $\Dc([0,\infty)) \ne \emptyset$,
624: there exists $t_0$, such that
625: \[
626: W(T):=\overline{\bigcup_{t \geq T} R(t) X}^{\bullet}
627: \]
628: is a nonempty $\db$-compact set for all $T\geq t_0$.
629: In addition, $W(s) \subset W(t)$ for all $s\geq t \geq 0$. Thus,
630: \[
631: \wb(X)= \bigcap_{T\geq t_0} W(T)
632: \]
633: is a nonempty  $\db$-compact set.
634: 
635: We will now prove that $\wb (X)$ uniformly $\db$-attracts $X$. Assume it
636: does not. Then there exists $\epsilon>0$, such that
637: \[
638: V(t):=W(t)\cap(X\setminus \Bb(\wb(X), \epsilon)) \ne \emptyset, \qquad \forall t\geq 0.
639: \]
640: Since $V(t)$ is $\db$-compact for $t\geq t_0$ and $V(s) \subset V(t)$ for
641: $s\geq t \geq 0$, we have that there exists
642: \[
643: x \in \bigcap_{t\geq t_0} V(t).
644: \]
645: Hence, $x \in \wb(X)$. However, this implies that $x\notin V(t)$, $t\geq 0$,
646: a contradiction.
647: \end{proof}
648: 
649: 
650: 
651: 
652: \begin{theorem} \label{thm:exofA}
653: If the map $R(t)$ is uniformly $\db$-compact
654: ($\bullet= \mathrm{s, w}$), then the  $\db$-global
655: attractor exists and satisfies the following additional properties:
656: \begin{enumerate}
657: \item[(a)]
658: $\Ab = \wb(X)$.
659: \item[(b)]
660: $\Ab$ is $\db$-compact.
661: \end{enumerate}
662: \end{theorem}
663: \begin{proof}
664: By Lemma~\ref{l:oncompactsemigroup},
665: $\wb(X)$ is a nonempty $\db$-compact $\db$-attracting set.
666: Moreover, $\wb(X)$ is the
667: minimal $\db$-closed $\db$-attracting set due to Lemma~\ref{l:wbinA}.
668: Therefore, $\wb(X)$ is the $\db$-global attractor.
669: \end{proof}
670: 
671: Note that since $X$ is $\dw$-compact (see the definition of $\dw$), $R(t)$ is uniformly weakly compact.
672: Hence we have the following.
673: \begin{corollary}
674: The evolutionary system $\Dc$ always possesses a weak global attractor $\Aw$.
675: \end{corollary}
676: Our next goal is to investigate whether $\Aw$ is an invariant set in the following
677: sense.
678: 
679: \begin{definition} The set $A \subset X$ is invariant, if
680: \[
681: \{u(t): u\in \Dc((-\infty,\infty)), u(0) \in A \}= A, \qquad \forall t\geq 0.
682: \]
683: \end{definition}
684: Assume that $u_0$ belongs to some invariant set. Then for all $t>0$
685: we have that $u_0 \in R(t)X$.
686: Hence, $u_0 \in \omega_{\mathrm{w}}(X)=\Aw$. Therefore, $\Aw$ contains every invariant set. Moreover, we will show that $\Aw$ is invariant under some compactness
687: property that is for instance satisfied by the family of all 
688: Leray-Hopf solutions of the 3D NSE (see Section~\ref{3DNSE}).
689: 
690: Let $C([a, b];X_\bullet)$ be the space of $\db$-continuous $X$-valued
691: functions on $[a, b]$ endowed with the metric
692: \[
693: \dd_{C([a, b]; X_\bullet)}(u,v) = \sup_{t\in[a,b]}\db(u(t),v(t)). 
694: \]
695: Let also $C([a, \infty);X_\bullet)$ be the space of $\db$-continuous
696: $X$-valued functions on $[a, \infty)$
697: endowed with the metric
698: \[
699: \dd_{C([a, \infty); X_\bullet)}(u,v) = \sum_{T\in \mathbb{N}} \frac{1}{2^T} \frac{\sup\{\db(u(t),v(t)):a\leq t\leq a+T\}}
700: {1+\sup\{\db(u(t),v(t)):a\leq t\leq a+T\}}.
701: \]
702: 
703: \begin{theorem} \label{c:weakA}
704: If $\Dc([0,\infty))$ is compact in $C([0, \infty);\Xw)$, then
705: \begin{enumerate}
706: \item[(a)]
707: $
708: \Aw = \mathcal{I}:=\{ u_0: \ u_0=u(0) \mbox{ for some }
709: u \in \Dc((-\infty, \infty))\}.
710: $
711: \item[(b)] $\Aw$ is the maximal invariant set.
712: \end{enumerate}
713: \end{theorem}
714: \begin{proof}
715: Since obviously $\I$ is the maximal invariant set, we have that $\I \subset \Aw$.
716: It remains to prove that $\Aw \subset \I$. Take any $a \in \Aw$.
717: Since $\Aw = \omega_{\mathrm{w}}(X)$,  there exist
718: $t_n \to \infty$, as $n\to \infty$ and $a_n\in R(t_n)X$, such that
719: $a_n \to a$ weakly as $n \to \infty$. 
720: Using Property 2 in Definition~\ref{d:E}, there exist
721: $u_n \in \Dc([-t_n, \infty))$, such that
722: $u_n(0) =a_n$. Also, Properties 2 and 3 in Definition~\ref{d:E}
723: of $\Dc$ imply that
724: $\Dc([-t_n,\infty))$ is compact in $C([-t_n, \infty);\Xw)$
725: and
726: \[
727: \{u|_{[-t_1,\infty)} : u \in \Dc([-t_{n},\infty))\}
728: \subset \Dc([-t_1,\infty)),
729: \]
730: for every $n$.
731: Now, passing to a subsequence and dropping a subindex, we can assume
732: that $u_n|_{[-t_1,\infty)} \to u^1\in \Dc([-t_1, \infty))$ in  $C([-t_1, \infty);\Xw)$
733: as $n\to \infty$. 
734: By a standard diagonalization process we obtain that there exist
735: $u \in  \Fc((-\infty,\infty))$ and a subsequence of $u_n$, still denoted by
736: $u_n$, such that 
737: $u_n|_{[-T,\infty)} \to u_{[-T,\infty)}$ in $C([-T, \infty);\Xw)$ for all $T>0$.
738: Thus, by the compactness we have that $u|_{[-T,\infty)} \in \Dc([-T, \infty))$
739: for all $T>0$, and hence $u \in \Dc((-\infty, \infty))$.
740: Finally, since $u(0)=a$. we have $a \in \mathcal{I}$.
741: Hence, $\Aw\subset \I$.
742: \end{proof}
743: 
744: Now that we know that the weak global attractor always exists, we can
745: weaken the condition on the existence of the strong global attractor. 
746: 
747: \begin{definition}
748: The map $R(t)$ is asymptotically $\db$-compact
749: ($\bullet= \mathrm{s, w}$) if for any $t_n \to \infty$ and
750: any $x_n \in R(t_n) X$, the  sequence
751: $\{x_n\}$ is relatively $\db$-compact.
752: \end{definition}
753: 
754: 
755: \begin{theorem} \label{t:asymptoticcompact}
756: If the map $R(t)$ is asymptotically  $\ds$-compact, then
757: $\Aw$ is $\ds$-compact strong global attractor.
758: \end{theorem}
759: \begin{proof}
760: First note that $\ws(X)\subset \ww(X)=\Aw$. On the other hand,
761: let $a \in \Aw=\ww(X)$. By
762: the definition of $\ww$-limit, there exist $t_n \to \infty$ as $n \to \infty$ and
763: $x_n \in R(t_n)X$, such that 
764: \[
765: \dw(x_n, a) \to 0 \qquad \mbox{ as } \qquad n \to \infty.
766: \]
767: Thanks to the asymptotic compactness of $R(t)$, this convergence
768: is in fact strong. Therefore, $a\in \ws(X)$. Hence, $\ws(X)= \Aw$.
769: 
770: Now let us show that $\ws(X)$ is a $\ds$-attracting set. Assume that it is not. Then there
771: exist $\epsilon>0$, $x_n \in X$, and $t_n \to \infty$ as
772: $n\to \infty$, such that
773: \[
774: x_n \in R(t_n)X \setminus B_{\s}(\ws(X), \epsilon), \qquad
775: \forall n \in \mathbb{N}.
776: \]
777: Since  $R(t)$ is asymptotically $\ds$-compact, then
778: $\{x_n\}$ is relatively $\ds$-compact.
779: Passing to a subsequence and dropping a subindex, we may assume that
780: \[
781: x_n \to x \in X \qquad \mbox{strongly},
782:  \qquad \mbox{as} \ n \to \infty.
783: \]
784: Therefore, we have that $x  \in \omega_{\mathrm{s}}(X)$,
785: a contradiction.
786: 
787: Now note that $\ws(X)$ is the
788: minimal $\ds$-closed $\ds$-attracting set due to Lemma~\ref{l:wbinA}.
789: Therefore, $\ws(X)$ is the strong global attractor $\As$.
790: Finally, let us show that $\ws(X)$
791: is strongly compact. Take any sequence $a_n \in \ws(X)$. By the definition of
792: $\ws$-limit, there exist $t_n \to \infty$ and
793: $x_n \in R(t_n)X$, such that 
794: \[
795: \ds(x_n, a_n) \to 0 \qquad \mbox{ as } \qquad n \to \infty.
796: \]
797: Note that $\{x_n\}$ is relatively $\ds$-compact due to the asymptotic compactness
798: of $R(t)$. Hence, $\{a_n\}$ is relatively $\ds$-compact, which concludes the proof.
799: 
800: \end{proof}
801: 
802: Finally, in the following example we show that the asymptotic compactness
803: is a weaker condition than the uniform strong compactness.
804: 
805: \noindent
806: \textbf{Example.}
807: Take
808: \[
809: X=\left\{ u\in l^2: \sum_{n=1}^\infty u_n^2  \leq 1\right\}.
810: \]
811: where $u=(u_n)$, $u_n \in \mathbb{R}$ for all $n$.  For $u, v \in X$, let
812: \[
813: \ds(u,v):=\left( \sum_{n=1}^\infty(u_n-v_n)^2 \right)^{1/2}, \qquad
814: \dw(u,v)= \sum_{n=1}^\infty \frac{1}{2^n}
815: \frac{|u_n-v_n|}{1 + |u_n-v_n|}.
816: \]
817: Consider the following differential equations:
818: \[
819: \ddt u_1=0,
820: \]
821: and
822: \[
823: \ddt u_n = - u_n, \qquad n \in \mathbb{N}.
824: \]
825: The trajectories of the evolutionary system $\Dc$ will be solutions of this
826: equation, i.e.,
827: \[
828: \begin{split}
829: \Dc([s,\infty)) =\{&u\in\Fc([s,\infty)): u_1(t)=u^0_1, u_n(t)=u_n^0e^{(s-t)}
830: \mbox{ for } n\geq 2,\\ & t\in[s,\infty), u^0 \in X\},
831: \end{split}
832: \]
833: for $s \in \mathbb{R}$, and
834: \[
835: \Dc((-\infty,\infty))=\{u: u_1\in [-1,1], u_n=0, n\geq 2\}.
836: \]
837: Clearly, $R(t)$ is not uniformly strongly compact, but it is
838: asymptotically compact. Hence, the strong global attractor
839: exists and
840: \[
841: \As = \{u: u_1\in [-1,1], u_n=0, n\geq 2\}.
842: \]
843: 
844: 
845: Finally,
846: Theorems~\ref{c:weakA} and \ref{t:asymptoticcompact}
847: imply the following
848: \begin{remark} 
849: If $\Dc([0,\infty))$ is compact in $C([0, \infty);\Xw)$ and $R(t)$ is asymptotically  
850: $\ds$-compact, then  the strong global attractor $\As$ exists, and is the strongly compact
851: maximal invariant set. Consequently, $\As$ is a compact global attractor in the
852: conventional sense.
853: \end{remark}
854: 
855: 
856: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
857: \section{3D Navier-Stokes equations} \label{3DNSE}
858: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
859: 
860: Here we apply the results from the previous section 
861: to the space periodic 3D Navier-Stokes equations (NSE)
862: \begin{equation} \label{NSE1}
863: \left\{
864: \begin{aligned}
865: &\ddt u - \nu \Delta u + (u \cdot \nabla)u + \nabla p = f,\\
866: &\nabla \cdot u =0,\\
867: &u, p, f \mbox{ are periodic with period } L \mbox{ in each space variable,}\\
868: &u, f \mbox{ are in } L^2_{\mathrm{loc}}(\mathbb{R}^3)^3,\\
869: &u|_{t=0}=u_0,
870: \end{aligned}
871: \right.
872: \end{equation}
873: where $u$, the velocity, and $p$, the pressure, are unknowns, $f$ is
874: a given driving force, and $\nu>0$ is the kinematic  viscosity coefficient
875: of the fluid. By a Galilean change of
876: variables, we can assume that the space average of $u$ is
877: zero, i.e.,
878: \[
879: \int_\Omega u(x,t) \, dx =0, \qquad \forall t,
880: \]
881: where $\Omega=[0,L]^3$ is a periodic box.
882: 
883: In this section we will apply our general results from the previous section
884: to the study of the assymptotical behaviour of weak solutions to \eqref{NSE1}.
885: 
886: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
887: \subsection{Functional setting} \label{funcset}
888: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
889: 
890: First, let us introduce some notations and functional setting for \eqref{NSE1}.
891: %Let $\Omega:=[0,L]^3$ be a periodic box.
892: We denote by
893: $(\cdot,\cdot)$ and $|\cdot|$ the $L^2(\Omega)^3$-inner product and the
894: corresponding $L^2(\Omega)^3$-norm.
895: Let $\mathcal{V}$ be the space of all $\mathbb{R}^3$ trigonometric polynomials of
896: period $L$ in each variable satisfying 
897: $\nabla \cdot u =0$ and $\int_\Omega u(x) \, dx =0$.
898: Let $H$ and $V$ to be the
899: closures of $\mathcal{V}$ in $L^2(\Omega)^3$ and $H^1(\Omega)^3$, respectively.
900: Also, define the distances $\db$ by
901: \[
902: \ds(u,v):=|u-v|, \qquad
903: \dw(u,v)= \sum_{\kappa \in \mathbb{Z}^3} \frac{1}{2^{|\kappa|}}
904: \frac{|u_{\kappa}-v_{\kappa}|}{1 + |u_{\kappa}-v_{\kappa}|},
905: \qquad u,v \in H,
906: \]
907: where $u_{\kappa}$ and $v_{\kappa}$ are Fourier coefficients of $u$
908: and $v$ respectively.
909: 
910: We denote by $P_{\sigma} : L^2(\Omega)^3 \to H$ the $L^2$-orthogonal
911: projection, referred to as the Leray projector, and by
912: $A=-P_{\sigma}\Delta = -\Delta$ the Stokes operator with the domain
913: $D(A)=(H^2(\Omega))^3 \cap V$. The Stokes operator is a self-adjoint
914: positive operator with a compact inverse.
915: Denote
916: \[
917: \|u\| := |A^{1/2} u| = \left(\int_\Omega \sum_{i,j=1}^3\left|\frac{\partial u_i}{\partial x_j}\right|^2 \, dx \right)^{1/2}.
918: \]
919: Note that $\|u\|$ is equivalent to the $H^1$-norm of $u$ for $u\in D(A^{1/2})$.
920: 
921: For a rigorous mathematical study of  the equation \eqref{NSE1} we need
922: a few more concepts from functional analysis. Namely, let $V'$ be the set
923: of all distributions of the form $v=\Delta u$, with $u\in V$. The $V'$-norm of
924: this $v$ is by definition $\|u\|$. Endowed with this norm, $V'$ becomes
925: the dual space of $V$, and if $v\in H$, then the value of $v$ at a point $w\in H$
926: equals to the usual scalar product $(v,w)$ in $H$. Now for $u$ and $v$ in $V$,
927: let $B(u,v):=P_{\sigma}(u \cdot \nabla v)$, which is an element of $V'$. If
928: $v \in \mathcal{D}(A)$, then $B(u,v) \in H$. Moreover, 
929: \[
930: \langle B(u,v),w \rangle =- \langle B(u,w),v \rangle, \qquad u,v,w \in V,
931: \]
932: in particular, $\langle B(u,v),v \rangle=0$ for all $u,v \in V$.
933: 
934: Equations (\ref{NSE1}) now can be condensed in the functional
935: differential equation
936: \begin{equation} \label{NSE}
937: \ddt u + \nu Au +B(u,u) = g \qquad \mbox{in} \qquad  V',
938: \end{equation}
939: where $u$ is a $V$-valued function of time and $g = P_{\sigma} f$.
940: Throughout, we will assume that  $g$ is time independent and $g\in H$.
941: 
942: \begin{definition}
943: A weak solution  of \eqref{NSE} on $[T,\infty)$ is an $H$-valued
944: function $u(t)$ defined for $t \in [T, \infty)$, such that
945: \[
946: u(\cdot) \in C([T, \infty); \Hw) \cap 
947: L_{\mathrm{loc}}^2([T, \infty); V),
948: \]
949: and
950: \begin{equation} \label{333}
951: w(\cdot):=g-\nu A u(\cdot) - B(u(\cdot),u(\cdot))\in L_{\mathrm{loc}}^1([T, \infty); V'),
952: \end{equation}
953: \begin{equation} \label{444}
954: \begin{split}
955: (u(t)-u(T),v)&= \langle u(t)-u(T),v \rangle\\
956: &=\int_T^t\langle w(s),v\rangle \, ds \qquad
957: \forall v \in V, t\geq T.
958: \end{split}
959: \end{equation}
960: \end{definition}
961: The relations \eqref{333} and \eqref{444} imply that
962: $\ddt u$ exists in $V'$ a.e. in $[T,\infty)$. Therefore, often in
963: the literature \eqref{444} and \eqref{333} are written as \eqref{NSE} and
964: \[
965: \ddt u \in L_{\mathrm{loc}}^1([T, \infty); V'),
966: \] 
967: respectively.
968: 
969: The classical fundamental result concerning \eqref{NSE} is the following.
970: 
971: \begin{theorem}[Leray, Hopf] \label{thm:Leray}
972: For every $u_0 \in H$,
973: there exists a weak solution $u(t)$ of (\ref{NSE}) on $[T,\infty)$ with $u(T)=u_0$
974: satisfying the following energy inequality: 
975: \begin{equation} \label{EI}
976: |u(t)|^2 + 2\nu \int_{t_0}^t \|u(s)\|^2 \, ds \leq
977: |u(t_0)|^2 + 2\int_{t_0}^t (g(s), u(s)) \, ds
978: \end{equation}
979: for all $t \geq t_0$, $t_0$ a.e. in $[T,\infty)$.
980: \end{theorem}
981: See \cite{K} for a hypothesis under which the energy equality holds. However,
982: in general, the existence of weak solutions satisfying the energy equality is
983: not known.  Therefore, we introduce the following definition. 
984: \begin{definition} \label{d:ex}
985: A Leray-Hopf solution of the \eqref{NSE} on the interval $[T, \infty)$
986: is a weak solution on $[T,\infty)$ satisfying the
987: energy inequality (\ref{EI}) for all $T \leq t_0 \leq t$,
988: $t_0$ a.e. in $[T,\infty)$. The set $Ex$ of measure $0$ of points $t_0$ for
989: which the energy
990: inequality does not hold will be called the exceptional set (of the solution).
991: In addition, the solution $u(t)$ will be called regular on an interval $(\alpha, \beta) \subset [T,\infty)$ if $u(t)\in V$ and $\|u(t)\|$ is continuous on $(\alpha,\beta)$.
992: \end{definition}
993: 
994: Note that the uniqueness of Leray-Hopf solutions of the Initial Value Problem is
995: not known.
996: 
997: 
998: \begin{theorem}[Leray] \label{existence}
999: For every $u_0 \in V$, there exists a strong solution
1000: $u(t)$ of \eqref{NSE} on some interval $[0,T)$, $T>0$, with $u(0)=u_0$.
1001: \end{theorem}
1002: 
1003: \begin{theorem}[Leray] \label{structure}
1004: Let $u(t)$ be a Leray-Hopf solution of \eqref{NSE} on $[T,\infty)$. Then there are
1005: at most countably many distinct open intervals $I_j$, such that
1006: \[
1007: [T,\infty) = \bigcup_j \overline{I}_j,
1008: \]
1009: $u(t)$ is regular on each $I_j$, and the measure of $[T,\infty)\setminus \cup_j I_j$
1010: is zero.
1011: \end{theorem}
1012: 
1013: \begin{theorem} \label{uniqueness}
1014: Let $u(t)$ be a regular solution on $[0,T)$. Then every Leray-Hopf solution
1015: $v(t)$ on $[0,\infty)$ with $v(0)=u(0)$ coincides with $u(t)$ in $[0,T)$.
1016: \end{theorem}
1017: 
1018: Finally, we recall several well-known supplementary facts.
1019: \begin{remark} \label{r:ex}
1020: The complement of the exceptional set $Ex$ coincides with the set of points of
1021: strong continuity from the right.
1022: \end{remark}
1023: 
1024: \begin{theorem} \label{t:doubleconv}
1025: Let $u(t)$, $u_n(t)$ be Leray-Hopf solutions on the interval $[0,\infty)$,
1026: such that 
1027: \[
1028: u_n \to u \qquad \mbox{in} \qquad C([0,T];\Hw),
1029: \]
1030: as $n\to \infty$, for some $T>0$. Let $(\alpha,\beta) \subset (0, T)$ be an interval of regularity
1031: of $u(t)$. Then for every $0 <\delta <(\beta-\alpha)/2$,
1032: \[
1033: \|u_n(t)-u(t)\|\to 0 \qquad \mbox{uniformly on} \qquad [\alpha +\delta, \beta -\delta],
1034: \]
1035:  as $n \to \infty$.
1036: \end{theorem}
1037: 
1038: \begin{remark}
1039: Let $u(t)$ be a Leray-Hopf solution. As a $V$-valued function, $u(t)$ is analytic in time on every interval of regularity.
1040: \end{remark}
1041: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1042: \subsection{The weak global attractor for the 3D NSE} \label{s:weakattractor}
1043: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1044: A ball $B\subset H$ is called an absorbing ball for the equation \eqref{NSE}
1045: if for any bounded set $K \subset H$, there exists $t_0$, such that
1046: \[
1047: u(t) \in B, \qquad \forall t \geq t_0,
1048: \]
1049: for all Leray-Hopf solutions $u(t)$ of \eqref{NSE} on $[0,\infty)$ with $u(0)\in K$.
1050: It is well known that there exists an absorbing ball in $H$ for the 3D NSE.
1051: In fact, one has the following.
1052: \begin{proposition} \label{p:aset}
1053: The 3D Navier-Stokes equations possess an absorbing ball
1054: \[
1055: B = B_{\mathrm{s}}(0, R),
1056: \]
1057: where $R$ is any number larger than
1058: $|g|\nu^{-1} L/(2\pi)$ (see, e.g., \cite{CF}).
1059: \end{proposition}
1060: 
1061: Fix $R>|g|\nu^{-1} L/(2\pi)$ and let $X$ be the closed absorbing ball
1062: \[
1063: X= \{u\in H: |u| \leq R\},
1064: \]
1065: which is, clearly, weakly compact. Then for any bounded set $K \subset H$,
1066: there exists a time $t_0$, such that
1067: \[
1068: u(t) \in X, \qquad \forall t\geq t_0,
1069: \]
1070: for every Leray-Hopf solution $u(t)$ with the initial data $u(0) \in K$.
1071: Classical NSE estimates (see \cite{CF}) imply that for any sequence of Leray--Hopf solutions $u_n(t)$ (not only for the ones guaranteed by Theorem~\ref{thm:Leray}),
1072: the following result holds.
1073: 
1074: \begin{lemma} \label{l:convergenceofLH}
1075: Let $u_n(t)$ be a sequence of Leray-Hopf solutions,
1076: such that $u_n(t) \in X$ for all $t\geq t_0$. Then 
1077: \[
1078: \begin{aligned}
1079: u_n \ \ &\mbox{is bounded in} \ \ L^2([t_0,T];V),\\
1080: \ddt u_n \ \  &\mbox{is bounded in} \ \ L^{4/3}([t_0,T];V'),
1081: \end{aligned}
1082: \]
1083: for all $T>t_0$.
1084: Moreover, there exists a subsequence $u_{n_j}$ of $u_n$, which converges
1085: in $C([t_0, T]; \Hw)$ to some Leray-Hopf solution $u(t)$, i.e.,
1086: \[
1087: (u_{n_j},v) \to (u,v) \qquad
1088: \mbox{uniformly on} \qquad  [t_0,T],
1089: \]
1090: as $n_j\to \infty$, for all $v \in H$.
1091: \end{lemma}
1092: 
1093: Consider an evolutionary system for which
1094: a family of all trajectories consists
1095: of all Leray-Hopf solutions of the 3D Navier-Stokes equations
1096: in $X$. More precisely, define
1097: \[
1098: \begin{split}
1099: \Dc([T,\infty)) := \{&u(\cdot): u(\cdot)
1100: \mbox{ is a Leray-Hopf solution on } [T,\infty)\\
1101: & \mbox{and } u(t) \in X \ \forall t \in [T,\infty)\},
1102: \qquad T \in \mathbb{R},
1103: \end{split}
1104: \]
1105: \[
1106: \begin{split}
1107: \Dc((\infty,\infty)) :=\{&u(\cdot): u(\cdot)
1108: \mbox{ is a Leray-Hopf solution on } (-\infty,\infty)\\
1109: & \mbox{and } u(t) \in X \ \forall t \in (-\infty,\infty)\}.
1110: \end{split}
1111: \]
1112: Since $X$ is weakly compact, the existence of the weak
1113: global attractor is a direct consequence of Theorem~\ref{c:weakA}.
1114: Moreover, we have the following.
1115: \begin{lemma} \label{l:compact}
1116: $\Dc([0,\infty))$ is compact in $C([0, \infty);\Hw)$.
1117: \end{lemma}
1118: \begin{proof}
1119: Take any sequence
1120: $u_n \in \Dc([0,\infty))$, $n\in \mathbb{N}$.
1121: Thanks to Lemma~\ref{l:convergenceofLH}, there exists
1122: a subsequence, still denoted by  $u_n$, that converges
1123: to some $u^{1} \in \Dc([0,\infty))$ in $C([0, 1];\Hw)$ as $n \to \infty$.
1124: Now, passing to a subsequence and dropping a subindex, we obtain that
1125: there exists $u^{2}\in \Dc([0,\infty))$, such that
1126: $u_n \to u^2$ in $C([0, 2];\Hw)$ as $n \to \infty$.
1127: Note that $u^1(t)=u^2(t)$ on $[0, 1]$.
1128: Continuing
1129: this diagonalization process, we obtain a subsequence $u_{n_j}$
1130: of $u_n$ that converges
1131: to some $u \in \Dc([0,\infty))$ in $C([0, \infty);\Hw)$ as $n_j \to \infty$, which
1132: concludes the proof.
1133: \end{proof}
1134: 
1135: Now Theorem~\ref{c:weakA} yields the following.
1136: \begin{theorem} \label{t:AweakNSE}
1137: The weak global attractor $\Aw$ for the 3D
1138: Navier-Stokes equations exists and satisfies
1139: \begin{enumerate}
1140: \item[(a)]
1141: $
1142: \Aw = \{u(0): u \in \Dc((-\infty, \infty))\}.
1143: $
1144: \item[(b)] $\Aw$ is the maximal invariant set.
1145: \end{enumerate}
1146: \end{theorem}
1147: 
1148: \begin{lemma}
1149: If $u(t)$, a Leray-Hopf solution of the 3D NSE, satisfies
1150: \[
1151: \limsup_{t \to \infty} \|u(t)\| < \infty,
1152: \]
1153: then $u(t)$ converges strongly in $H$ to the weak global
1154: attractor $\Aw$.
1155: \end{lemma}
1156: \begin{proof}
1157: Suppose that $u(t)$ does not converge strongly in $H$ to $\Aw$. 
1158: Then there exist $M>0$ and a sequence $t_n \to \infty$ as
1159: $n \to \infty$, such that
1160: \begin{equation} \label{eq:farfromAw}
1161: \ds(u(t_n), \Aw) > M, \qquad n \in \mathbb{N}.
1162: \end{equation} 
1163: Note that there exists a time $T>0$, such that
1164: $\{u(t): t\geq T\}$ is relatively compact in $H$. Therefore, passing
1165: to a subsequence, we may assume that $u(t_n)$ converges strongly
1166: (and hence weakly) in $H$ to some $a \in H$. Therefore, $a \in \Aw$, which
1167: contradicts
1168: (\ref{eq:farfromAw}).
1169: \end{proof}
1170: 
1171: 
1172: \subsection{Strong convergence of Leray-Hopf solutions}
1173: 
1174: The aim of this subsection is to give sufficient conditions for a
1175: sequence of Leray-Hopf solutions on $[T,\infty)$ to converge
1176: in $C([T,\infty);H)$, provided it converges in $C([T,\infty);\Hw)$.
1177: We start with preliminary properties, some of which may have
1178: an intrinsic interest.
1179: \begin{theorem} \label{t:lim}
1180: Let $u(t)$ be a Leray-Hopf solution of \eqref{NSE} on $[T,\infty)$.
1181: Let $Ex$ be the exceptional set for this
1182: solution (see Def.~\ref{d:ex}). Then for any time $t_0 >T$,
1183: there exist $A_-$ and $A_+$, such that
1184: \begin{enumerate}
1185: \item[(a)]
1186: For every sequence $\{t_n\} \subset [T,\infty) \setminus Ex$,
1187: such that $t_n \to t_0$, $t_n < t_0$, it follows that  $|u(t_n)| \to A_-$
1188: as $n \to \infty$.
1189: \item[(b)]
1190: For every sequence $\{t_n\} \subset [T,\infty) \setminus Ex$,
1191: such that $t_n \to t_0$, $t_n > t_0$, it follows that $|u(t_n)| \to A_+$
1192: as $n \to \infty$.
1193: \end{enumerate}
1194: For these $A_-$ and $A_+$ we will use the following notations:
1195: \[
1196: \widetilde{\lim_{t \to t_0 -}} |u(t)| := A_- \qquad \mbox{and}
1197: \qquad \widetilde{\lim_{t \to t_0 +}} |u(t)| := A_+.
1198: \]
1199: \end{theorem}
1200: \begin{proof}
1201: For $\{t_n\}$ as in (a), the energy inequality (\ref{EI}) on
1202: $[t_n, t_{n+k}]$ is 
1203: \[
1204: |u(t_{n+k})|^2 + 2\nu \int_{t_n}^{t_{n+k}}\|u(s)\|^2 \, ds \leq
1205: |u(t_n)|^2 + 2\int_{t_n}^{t_{n+k}} (g, u(s)) \, ds,
1206: \]
1207: provided $t_{n+k} \geq t_n$.
1208: Taking the upper limit as $k \to \infty$, we obtain
1209: \[
1210: \limsup_{n \to \infty} |u(t_n)|^2 + 2\nu \int_{t_n}^{t_0}\|u(s)\|^2 \, ds \leq
1211: |u(t_n)|^2 + 2\int_{t_n}^{t_0} (g, u(s)) \, ds.
1212: \]
1213: Taking the lower limit as $n \to \infty$, we arrive at
1214: \[
1215: \limsup_{n \to \infty} |u(t_n)|^2 \leq
1216: \liminf_{n \to \infty} |u(t_n)|^2,
1217: \]
1218: i.e., $\lim_{n \to \infty} |u(t_n)|$ exists. Since the limit exists for any sequence
1219: $t_n$, it does not depend on the choice of a sequence.
1220: 
1221: For $\{t_n\}$ as in (b), the energy inequality (\ref{EI}) on
1222: $[t_{n+k}, t_n]$ is
1223: \[
1224: |u(t_{n})|^2 + 2\nu \int_{t_{n+k}}^{t_n}\|u(s)\|^2 \, ds \leq
1225: |u(t_{n+k})|^2 + 2\int_{t_{n+k}}^{t_n} (g, u(s)) \, ds,
1226: \]
1227: provided $t_{n+k} \leq t_n$.
1228: Taking the lower limit as $k \to \infty$, we obtain
1229: \[
1230: |u(t_n)|^2 + 2\nu \int_{t_0}^{t_n}\|u(s)\|^2 \, ds \leq
1231: \liminf_{n \to \infty} |u(t_n)|^2 + 2\int_{t_0}^{t_n} (g, u(s)) \, ds.
1232: \]
1233: Finally, taking the upper limit as $n \to \infty$, we arrive at
1234: \[
1235: \limsup_{n \to \infty} |u(t_n)|^2 \leq
1236: \liminf_{n \to \infty} |u(t_n)|^2,
1237: \]
1238: i.e., $\lim_{n \to \infty} |u(t_n)|$ exists. Since the limit exists for any sequence
1239: $t_n$, it does not depend on the choice of a sequence.
1240: \end{proof}
1241: 
1242: 
1243: \begin{lemma} \label{l:simple}
1244: Let $u(t)$ be a Leray-Hopf solution of \eqref{NSE}
1245: on $[T,\infty)$. Then
1246: \[
1247: \begin{split}
1248: \widetilde{\lim_{t \to t_0 -}} |u(t)| =
1249: \limsup_{t \to t_0 -} |u(t)|,\\
1250: \widetilde{\lim_{t \to t_0 +}} |u(t)| =
1251: \limsup_{t \to t_0 +} |u(t)|,
1252: \end{split}
1253: \]
1254: and,
1255: \begin{equation} \label{eq:orderoflim}
1256: \widetilde{\lim_{t \to t_0 -}} |u(t)| \geq 
1257: \widetilde{\lim_{t \to t_0 +}} |u(t)| \geq |u(t_0)|.
1258: \end{equation}
1259: for all $t_0 >T$.
1260: \end{lemma}
1261: \begin{proof}
1262: Take any $t_0 > T$. Obviously, we have
1263: \[
1264: \begin{split}
1265: \widetilde{\lim_{t \to t_0 -}} |u(t)| \leq
1266: \limsup_{t \to t_0 -} |u(t)|,\\
1267: \widetilde{\lim_{t \to t_0 +}} |u(t)| \leq
1268: \limsup_{t \to t_0 +} |u(t)|.
1269: \end{split}
1270: \]
1271: To show the opposite inequalities, note that for any
1272: $t_1 \in [T,\infty) \setminus Ex$ and $t_2>t_1$,
1273: the energy inequality (\ref{EI}) on $[t_1, t_2]$ is
1274: \begin{equation*}
1275: |u(t_2)|^2 + 2\nu \int_{t_1}^{t_2}\|u(s)\|^2 \, ds \leq
1276: |u(t_1)|^2 + 2\int_{t_1}^{t_2} (g, u(s)) \, ds.
1277: \end{equation*}
1278: First, we fix $t_1<t_0$ and take the upper limit
1279: as $t_2 \to t_0-$, obtaining
1280: \[
1281: \limsup_{t \to t_0 -} |u(t)|^2 + 2\nu \int_{t_1}^{t_0}\|u(s)\|^2 \, ds \leq
1282: |u(t_1)|^2 + 2\int_{t_1}^{t_0} (g, u(s)) \, ds.
1283: \]
1284: Now we take the limit as $t_1 \to t_0 -$ avoiding the exceptional set
1285: (see Theorem~\ref{t:lim}). We get
1286: \[
1287: \limsup_{t \to t_0 -} |u(t)|^2 \leq
1288: \widetilde{\lim_{t \to t_0 -}} |u(t)|.
1289: \]
1290: 
1291: Second, we fix $t_2>t_0$ and take the limit
1292: as $t_1 \to t_0+$ avoiding the exceptional set. We arrive at
1293: \[
1294: |u(t_2)|^2 + 2\nu \int_{t_0}^{t_2}\|u(s)\|^2 \, ds \leq
1295: \widetilde{\lim_{t \to t_0+}} |u(t)|^2 + 2\int_{t_0}^{t_2} (g, u(s)) \, ds.
1296: \]
1297: Taking the upper limit as $t_2 \to t_0 +$, we get
1298: \[
1299: \limsup_{t \to t_0 +} |u(t)|^2 \leq
1300: \widetilde{\lim_{t \to t_0 +}} |u(t)|.
1301: \]
1302: 
1303: Third, we fix $t_2>t_0$ and take the limit
1304: as $t_1 \to t_0-$ avoiding the exceptional set. We obtain
1305: \[
1306: |u(t_2)|^2 + 2\nu \int_{t_0}^{t_2}\|u(s)\|^2 \, ds \leq
1307: \widetilde{\lim_{t \to t_0-}} |u(t)|^2 + 2\int_{t_0}^{t_2} (g, u(s)) \, ds.
1308: \]
1309: Taking the limit as $t_2 \to t_0 +$ avoiding the exceptional set, we get
1310: \[
1311: \widetilde{\lim_{t \to t_0 +}} |u(t)|^2 \leq
1312: \widetilde{\lim_{t \to t_0 -}} |u(t)|.
1313: \]
1314: Finally the weak continuity of $u(t)$ yields
1315: \[
1316: \widetilde{\lim_{t \to t_0 +}} |u(t)| \geq |u(t_0)|,
1317: \]
1318: which concludes the proof.
1319: \end{proof}
1320: 
1321: \begin{remark} We can now rewrite the energy
1322: inequality for a Leray-Hopf solution $u(t)$ in the following form:
1323: \begin{equation} \label{e:gen}
1324: |u(t)|^2 + 2\nu \int_{t_0}^{t}
1325: \|u(s)\|^2 \, ds \leq
1326: \widetilde{\lim_{t\to t_0+}} |u(t)|^2 + 2\int_{t_0}^{t} (g, u(s)) \, ds,
1327: \end{equation}
1328: for all $0\leq t_0 \leq t$.
1329: \end{remark}
1330: 
1331: Recall that if the energy norm $|u(t)|$ of a Leray-Hopf solution 
1332: is continuous from the right at some $t=t_0$, then $t_0$ does not belong to
1333: the exceptional set for $u(t)$, i.e. the energy inequality holds
1334: for $t_0$ (see Remark~\ref{r:ex}).
1335: 
1336: \begin{lemma} \label{l:contfromtheright}
1337: Let $u(t)$ be a Leray-Hopf solution of \eqref{NSE} on $[T,\infty)$. Then
1338: $|u(t)|$ is continuous from the right at $t=t_0 \geq T$
1339: if and only if
1340: \[
1341: \widetilde{\lim_{t \to t_0 +}} |u(t)| = |u(t_0)|.
1342: \]
1343: \end{lemma}
1344: \begin{proof}
1345: If $|u(t)|$ is continuous from the right
1346: at $t=t_0 \geq T$, then, thanks to Lemma~\ref{l:simple}, we have that
1347: \begin{equation} \label{eq:tempcontr}
1348: \widetilde{\lim_{t \to t_0 +}} |u(t)| = \limsup_{t \to t_0 +} |u(t)| = |u(t_0)|. 
1349: \end{equation}
1350: Assume now that (\ref{eq:tempcontr}) holds.
1351: Due to the weak continuity of $u(t)$, we have
1352: \[
1353: \liminf_{t \to t_0 +} |u(t)|^2 \geq |u(t_0)|.
1354: \]
1355: Hence, 
1356: \[
1357: \lim_{t \to t_0 +} |u(t)|^2 = |u(t_0)|.
1358: \]
1359: \end{proof}
1360: 
1361: 
1362: 
1363: Now we will show that the strong
1364: continuity of a Leray-Hopf solution is equivalent to the strong
1365: continuity from the left (avoiding the exceptional set).
1366: 
1367: 
1368: \begin{lemma} \label{l:simplecontin}
1369: Let $u(t)$ be a Leray-Hopf solution of \eqref{NSE} on $[T, \infty)$. Then
1370: $|u(t)|$ is continuous at $t=t_0>T$ if and only
1371: if
1372: \[
1373: \widetilde{\lim_{t \to t_0 -}} |u(t)| = |u(t_0)|.
1374: \]
1375: \end{lemma}
1376: \begin{proof}
1377: Clearly, if $|u(t)|$ is continuous at
1378: $t=t_0 >T$, then
1379: \begin{equation} \label{eq:tempcont}
1380: \widetilde{\lim_{t \to t_0 -}} |u(t)| =\limsup_{t \to t_0 -} |u(t)| =  |u(t_0)|. 
1381: \end{equation}
1382: Assume now that (\ref{eq:tempcont}) holds. 
1383: Then due to the weak continuity of $u(t)$, we have
1384: \[
1385: \lim_{t \to t_0 -} |u(t)| = |u(t_0)|.
1386: \]
1387: In addition, Lemma~\ref{l:simple} (equation \eqref{eq:orderoflim}) implies that
1388: \[
1389: \widetilde{\lim_{t \to t_0 +}} |u(t)| = |u(t_0)|.
1390: \]
1391: Finally, thanks to Lemma~\ref{l:contfromtheright}, we have
1392: \[
1393: \lim_{t \to t_0 +} |u(t)| = |u(t_0)|.
1394: \]
1395: Therefore, $|u(t)|$ is continuous at $t=t_0$.
1396: \end{proof}
1397: 
1398: 
1399: 
1400: 
1401: 
1402: 
1403: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1404: %       LEFT Lemma
1405: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1406: 
1407: We will now study a weak convergence of Leray-Hopf solutions. Our goal
1408: is to obtain sufficient conditions for a strong convergence.
1409: 
1410: \begin{lemma} \label{l:left}
1411: Let $\{u_n(t)\}$, $u(t)$ be Leray-Hopf solutions of \eqref{NSE} on $[T_1, \infty)$.
1412: If $u_n \to u$ in $C([T_1, T_2]; \Hw)$, then
1413: \[
1414: \limsup_{n \to \infty} \widetilde{\lim_{t \to t_0 -}} |u_n(t)| \leq
1415: \widetilde{\lim_{t \to t_0 -}} |u(t)|,
1416: \]
1417: for all $t_0 \in (T_1, T_2]$.
1418: \end{lemma}
1419: \begin{proof}
1420: Suppose this is not true for some $t_0 \in (T_1, T_2]$. Then passing to a subsequence and dropping the subindexes, we can assume that
1421: \[
1422: \widetilde{\lim_{t \to t_0-}} |u_n(t)| - 
1423: \widetilde{\lim_{t \to t_0-}} |u(t)| \geq  \delta > 0, \qquad \forall n
1424: \]
1425: and $|u_n(t)| \to |u(t)|$ on $[T_1, T_2]\setminus S$, where $S$ is
1426: a set of zero measure, which includes the exceptional set for $u(t)$.
1427: 
1428: Let $S':= S\cup \left( \bigcup_n Ex_n \right)$, where 
1429: $Ex_n$ is the exceptional set for $u_n(t)$.
1430: The energy inequality for $u_n(t)$ implies
1431: \[
1432: \widetilde{\lim_{\tau \to t_0 - }} |u_n(\tau)|^2 \leq |u_n(t)|^2 +
1433: 2\int_{t}^{t_0}\left(g, u_n(s)\right) \, ds,
1434: \]
1435: for all $t\in [T_1,t_0] \setminus S'$.
1436: Taking the upper limit as $n \to \infty$ and using the strong convergence
1437: of $u_n(t)$ to $u(t)$ on $[T_1,T_2] \setminus S'$, we obtain
1438: \[
1439: \limsup_{n \to \infty} \widetilde{\lim_{\tau \to t_0 - }} |u_n(\tau)|^2 \leq
1440: |u(t)|^2 + 2\int_{t}^{t_0}\left(g, u(s)\right) \, ds,
1441: \qquad t\in [T_1, t_0] \setminus S'.
1442: \]
1443: Finally, letting $t \to t_0 -$, we get
1444: \[
1445: \limsup_{n \to \infty} \widetilde{\lim_{t \to t_0 - }} |u_n(t)|^2
1446: \leq \widetilde{\lim_{t \to t_0-}} |u(t)|^2,
1447: \]
1448: which is in contradiction with the definition of $\delta$.
1449: \end{proof}
1450: 
1451: 
1452: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1453: %       RIGHT Lemma
1454: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1455: 
1456: \begin{lemma} \label{l:right}
1457: Let $\{u_n(t)\}$, $u(t)$ be Leray-Hopf solutions of \eqref{NSE} on $[T_1, \infty)$.
1458: If $u_n \to u$ in $C([T_1, T_2]; \Hw)$, then
1459: \[
1460: \liminf_{n \to \infty} \widetilde{\lim_{t \to t_0 +}} |u_n(t)| \geq
1461: \widetilde{\lim_{t \to t_0 +}} |u(t)|,
1462: \]
1463: for all $t_0 \in [T_1, T_2)$.
1464: \end{lemma}
1465: \begin{proof}
1466: Suppose this is not true for some $t_0 \in [T_1,T_2)$. Then passing to a
1467: subsequence and dropping the subindexes, we can assume that
1468: \[
1469: \widetilde{\lim_{t \to t_0+}} |u_n(t)| - 
1470: \widetilde{\lim_{t \to t_0+}} |u(t)| \leq \delta < 0,
1471: \qquad \forall n
1472: \]
1473: and
1474: $|u_n(t)| \to |u(t)|$ on $[T_1, T_2]\setminus S$, where $S$ is a zero
1475: measure set, which includes the exceptional set for $u(t)$.
1476: The energy inequality (\ref{e:gen}) for $u_n(t)$ implies
1477: \[
1478: |u_n(t)|^2  \leq \widetilde{\lim_{\tau \to t_0 + }} |u_n(\tau)|^2 +
1479: 2\int_{t_0}^{t}\left(g, u_n(s)\right) \, ds,
1480: \]
1481: for all $t\in [t_0, T_2]$.
1482: Taking the lower limit as $n \to \infty$ and using the strong convergence
1483: of $u_n(t)$ to $u(t)$ on $[T_1, T_2] \setminus S$, we obtain
1484: \[
1485: |u(t)|^2  \leq \liminf_{n \to \infty} \widetilde{\lim_{\tau \to t_0 + }}
1486: |u_n(\tau)|^2 +
1487: 2\int_{t_0}^{t}\left(g, u(s) \right) \, ds,
1488: \qquad t\in [t_0, T_2] \setminus S.
1489: \]
1490: Finally, letting $t \to t_0 +$, we get
1491: \[
1492: \widetilde{\lim_{t \to t_0+}} |u(t)|^2  \leq
1493: \liminf_{n \to \infty} \widetilde{\lim_{t \to t_0 + }} |u_n(t)|^2,
1494: \]
1495: which is in contradiction with the definition of $\delta$.
1496: \end{proof}
1497: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1498: 
1499: 
1500: 
1501: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1502: \begin{definition}
1503: For $u(t)$, a Leray-Hopf solution of \eqref{NSE}, denote
1504: \[
1505: [u(t_0)] := \widetilde{\lim_{t \to t_0 -}} |u(t)| -
1506: \widetilde{\lim_{t \to t_0 +}} |u(t)|,
1507: \]
1508: which we will call the energy norm jump (loss) at $t=t_0$. 
1509: \end{definition}
1510: 
1511: Note that due to (\ref{eq:orderoflim}), the energy norm jumps of the Leray-Hopf
1512: solutions are never negative, i.e.,
1513: \[
1514: [u(t)] \geq 0, \qquad \forall t,
1515: \]
1516: for any Leray-Hopf solution $u(t)$. Moreover, we have the following result.
1517: 
1518: 
1519: \begin{theorem} \label{thm:weak-jumps}
1520: Let $\{u_n(t)\}$, $u(t)$ be Leray-Hopf solutions of \eqref{NSE} on $[T_1, \infty)$.
1521: If $u_n \to u$ in $C([T_1, T_2], \Hw)$, then
1522: \[
1523: \limsup_{n \to \infty} [u_n(t)] \leq [u(t)],
1524: \]
1525: for all $t \in (T_1, T_2)$.
1526: \end{theorem}
1527: \begin{proof}
1528: Indeed, Lemmas \ref{l:left} and \ref{l:right}  yield
1529: \[
1530: \begin{split}
1531: \limsup_{n \to \infty} [u_n(t_0)] &= \limsup_{n \to \infty}
1532: \left(\widetilde{\lim_{t \to t_0-}}|u_n(t)| -
1533: \widetilde{\lim_{t \to t_0+}}|u_n(t)| \right) \\
1534: &\leq \limsup_{n \to \infty} \widetilde{\lim_{t \to t_0-}}|u_n(t)| -
1535: \liminf_{n \to \infty} \widetilde{\lim_{t \to t_0+}}|u_n(t)| \\
1536: &\leq \widetilde{\lim_{t \to t_0-}}|u(t)| -
1537: \widetilde{\lim_{t \to t_0+}}|u(t)|\\
1538: &= [u(t_0)].
1539: \end{split}
1540: \]
1541: \end{proof}
1542: 
1543: \begin{comment}
1544: For our strong convergence criteria we will need the following "uniform"
1545: version of Lemma~\ref{l:left}
1546: \begin{theorem} \label{t:left}
1547: Let $\{u_n\}$ be a sequence of Leray-Hopf solutions on $[T_1, \infty)$.
1548: If $u_n \to u$ in $C([T_1, T_2]; \Hw)$, then for any $\epsilon>0$
1549: there exists $N$, such that
1550: \[
1551: \widetilde{\lim_{t \to t_0 -}} |u_n(t)| \leq \widetilde{\lim_{t \to t_0 -}} |u(t)| + \epsilon
1552: \qquad \forall n\geq N, \forall t_0 \in (T_1, T_2].
1553: \]
1554: \end{theorem}
1555: \begin{proof}
1556: Suppose this is not true for some $t_0 \in (T_1, T_2]$. Then passing to a subsequence and dropping the subindexes, we can assume that
1557: \[
1558: \widetilde{\lim_{t \to t_0-}} |u_n(t)| - 
1559: \widetilde{\lim_{t \to t_0-}} |u(t)| \geq  \delta > 0, \qquad \forall n
1560: \]
1561: and $|u_n(t)| \to |u(t)|$ on $[T_1, T_2]\setminus S$, where $S$ is
1562: a set of zero measure, which includes the exceptional set.
1563: 
1564: Let $S':= S\cup \left( \bigcup_n Ex_n \right)$, where 
1565: $Ex_n$ is the exceptional set for $u_n$.
1566: The energy inequality for $u_n$ implies
1567: \[
1568: \widetilde{\lim_{\tau \to t_0 - }} |u_n(\tau)|^2 \leq |u_n(t)|^2 +
1569: 2\int_{t}^{t_0}\left(g, u_n(\tau)\right) \, d \tau.
1570: \]
1571: for all $t\in [T_1,t_0] \setminus S'$.
1572: Taking the upper limit as $n \to \infty$ and using the strong convergence
1573: of $u_n(t)$ on $[T_1,T_2] \setminus S'$, we obtain
1574: \[
1575: \limsup_{n \to \infty} \widetilde{\lim_{\tau \to t_0 - }} |u_n(\tau)|^2 \leq
1576: |u(t)|^2 + 2\int_{t}^{t_0}\left(g, u(\tau)\right) \, d \tau.
1577: \qquad t\in [T_1, t_0] \setminus S'.
1578: \]
1579: Finally, letting $t \to t_0 -$, we get
1580: \[
1581: \limsup_{n \to \infty} \widetilde{\lim_{t \to t_0 - }} |u_n(t)|^2
1582: \leq \widetilde{\lim_{t \to t_0-}} |u(t)|^2,
1583: \]
1584: which is in contradiction with the definition of $\delta$.
1585: \end{proof}
1586: \end{comment}
1587: 
1588: 
1589: 
1590: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1591: % Main Theorem!!!
1592: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1593: 
1594: Assume now that Leray-Hopf solutions converge weakly to a strongly
1595: continuous from the right in $H$ Leray-Hopf solution. We will show that the weak 
1596: convergence is strong if the energy jumps of solutions converge to the
1597: energy jumps of the limit solution.
1598: 
1599: \begin{theorem} \label{t:deltat}
1600: Let $\{u_n(t)\}$, $u(t)$ be Leray-Hopf solutions of \eqref{NSE} on $[T_1, \infty)$.
1601: If $u_n \to u$ in $C([T_1, T_2]; \Hw)$, 
1602: \[
1603: \widetilde{\lim_{t \to t_0 +}}|u(t)|=|u(t_0)|, \qquad \mbox{and} \qquad
1604: \liminf_{n \to \infty} [u_n(t_0)] \geq [u(t_0)],
1605: \]
1606: for some $t_0 \in (T_1, T_2)$, then
1607: \[
1608: \lim_{n\to \infty} |u_n(t_0)| = |u(t_0)|.
1609: \]
1610: \end{theorem}
1611: \begin{proof}
1612: Thanks to \eqref{eq:orderoflim} and Lemma~\ref{l:left}, we have
1613: \[
1614: \begin{split}
1615: \limsup_{n \to \infty} |u_n(t_0)| &\leq \limsup_{n \to \infty}
1616: \widetilde{\lim_{t\to t_0+}} |u_n(t)|\\
1617: &\leq \liminf_{n \to \infty} \widetilde{\lim_{t\to t_0-}} |u_n(t)|
1618: - \liminf_{n\to \infty}[u_n(t_0)]\\
1619: &\leq\limsup_{n \to \infty} \widetilde{\lim_{t\to t_0-}} |u_n(t)| - [u(t_0)]\\
1620: &\leq \widetilde{\lim_{t\to t_0-}} |u(t)| - [u(t_0)]\\
1621: &= \widetilde{\lim_{t\to t_0+}} |u(t)|\\
1622: &= |u(t_0)|,
1623: \end{split}
1624: \]
1625: which concludes the proof.
1626: \end{proof}
1627: 
1628: Note that if a Leray-Hopf solution $u(t)$ is continuous at $t=t_0$, then
1629: $[u(t_0)]=0$. Therefore, in particular, Theorem~\ref{t:deltat}
1630: immediately implies the following.
1631: 
1632: \begin{corollary} \label{c:gaex}
1633: Let $\{u_n\}$ be a sequence of Leray-Hopf solutions of \eqref{NSE} on
1634: $[T_1, \infty)$.
1635: If $u_n \to u$ in $C([T_1, T_2]; \Hw)$, and
1636: $|u(t)|$ is continuous at some $t=t_0 \in (T_1, T_2)$,
1637: then $u_n(t_0) \to u(t_0)$ strongly in $H$.
1638: \end{corollary}
1639: So, if a solution $u(t)$ on the weak global atractor is continuous
1640: in $H$, then it attracts its basin strongly.
1641: Note that this is exactly Rosa's asymptotic regularity condition (see \cite{R}).
1642: 
1643: \subsection{The strong global attractor for the 3D NSE}
1644: In this subsection we will see that if
1645: all solutions on the weak global attractor $\Aw$ are continuous in $H$, then
1646: $\Aw$ is the strong global attractor.
1647: First, we will show that if a solution
1648: belongs to $\Aw$ on some open time-interval $I$, then on any closed subinterval
1649: of $I$ it has to coincide  with a solution that stays on $\Aw$ for all time.
1650: 
1651: \begin{lemma} \label{l:glue1}
1652: Let $u \in \Dc([T,\infty))$ and $\tilde{u} \in \Dc([0,\infty))$, such that
1653: $u(T)=\tilde{u}(T)$ and
1654: \[
1655: \widetilde{\lim_{t \to T-}} |\tilde{u}(t)| \geq 
1656: \widetilde{\lim_{t \to T+}} |u(t)|,
1657: \]
1658: for some $T >0$. Let
1659: \[
1660: v(t)=\left\{
1661: \begin{aligned}
1662: &\tilde{u}(t), &0 \leq t \leq T, \\
1663: &u(t), &t > T.
1664: \end{aligned}
1665: \right.
1666: \]
1667: Then $v \in \Dc([0,\infty))$.
1668: \end{lemma}
1669: \begin{proof}
1670: Obviously, $v(t)$ is a weak solution of the 3D NSE. To show that it satisfies
1671: the energy inequality, take any $t\geq T$ and $t_0 \in (0,T)$,
1672: $t_0 \notin Ex$, where $Ex$ is the exceptional set for $\tilde{u}$. Then
1673: we have
1674: \[
1675: |v(t)|^2 + 2\nu \int_{T}^t \|v(s)\|^2 \, ds \leq
1676: \widetilde{\lim_{s \to T+}} |v(s)|^2 + 2\int_{T}^t (g, v(s)) \, ds,
1677: \]
1678: and
1679: \[
1680: \widetilde{\lim_{s \to T-}}|v(s)|^2 + 2
1681: \nu \int_{t_0}^{T} \|v(s)\|^2 \, ds \leq
1682: |v(t_0)|^2 + 2\int_{t_0}^{T} (g, v(s)) \, ds.
1683: \]
1684: Adding these inequalities, we obtain
1685: \[
1686: |v(t)|^2 + 2\nu \int_{t_0}^t \|v(s)\|^2 \, ds \leq
1687: |v(t_0)|^2 + 2\int_{t_0}^t (g, v(s)) \, ds,
1688: \]
1689: which concludes the proof.
1690: \end{proof}
1691: 
1692: \begin{corollary}
1693: \label{l:glue}
1694: Let $u\in \Dc([T,\infty))$ and $\tilde{u} \in \Dc((-\infty,\infty))$, such that
1695: $u(T)=\tilde{u}(T)$ and $u(t)$ is strongly continuous from the
1696: right at $t=T$. Let
1697: \[
1698: v(t)=\left\{
1699: \begin{aligned}
1700: &\tilde{u}(t), &t \leq T, \\
1701: &u(t), &t > T.
1702: \end{aligned}
1703: \right.
1704: \]
1705: Then $v \in \Dc((-\infty,\infty))$.
1706: \end{corollary}
1707: 
1708: \begin{lemma} \label{l:contona}
1709: Let $u(t)$ be a Leray-Hopf solution $u\in \Dc ([T_1,\infty))$,
1710: such that $u(t) \in \Aw$ for all
1711: $t \in (T_1, T_2)$. Then for any $T_0\in(T_1,T_2)$, there
1712: exists $v\in \Dc ((-\infty, \infty))$,
1713: such that $u(t) = v(t)$ on $[T_0,\infty)$. In particular, $u(t) \in \Aw$ for all
1714: $t>T_1$.
1715: \end{lemma}
1716: \begin{proof}
1717: First note that $(T_1, T_0)$ contains
1718: some interval of regularity of $u(t)$ and take $T$ in the interior of this
1719: interval. Since $u(T) \in \Aw$, there exists a solution
1720: $\tilde{u} \in \Dc ((-\infty, \infty))$, such that $\tilde{u}(T) =u(T)$.
1721: We will now glue them at point $t=T$, obtaining
1722: \[
1723: v(t)=\left\{
1724: \begin{aligned}
1725: &\tilde{u}(t), &t \leq T, \\
1726: &u(t), &t > T.
1727: \end{aligned}
1728: \right.
1729: \]
1730: Since $u(t)$ is strongly continuous at $t=T$,
1731: Corolary~\ref{l:glue} implies that $v \in \Dc((-\infty, \infty))$.
1732: \end{proof}
1733: 
1734: Assume now that all solutions that stay on
1735: the weak global attractor are strongly continuous in $H$. We will prove that
1736: in this case the weak global attractor is strong. 
1737: This generalizes a well-known result in \cite{FT85}.
1738: Indeed, if $\Aw$ is bounded in $V$, then all the solutions,
1739: as long as they stay on the weak global attractor, are regular, and
1740: are therefore continuous in $H$. This also weakens the condition
1741: of Ball \cite{B1} -- the continuity of all Leray-Hopf solutions from $(0,\infty)$
1742: to $H$.
1743: 
1744: \begin{lemma} \label{l:cont-comp}
1745: If  all solutions on the weak global attractor are strongly continuous
1746: in $H$, i.e., if $\Dc ((-\infty, \infty)) \subset C((-\infty, \infty); H)$,
1747: then $R(t)$ is asymptotically compact.
1748: \end{lemma} 
1749: \begin{proof}
1750: Take
1751: any $\{t_n\}$, such that $t_n \to \infty$ as $n \to \infty$, and $x_n \in R(t_n)X$.
1752: Then there exists
1753: a sequence of solutions $v_n\in \Dc ([0,\infty))$,
1754: such that $v_n(t_n) =x_n$.
1755: We will show that $\{x_n\}$ has a convergent subsequence.
1756: Without loss of generality, there exists $T>0$, such that $t_n \geq 2T$ for all $n$. Consider a sequence $u_n(t)=v_n(t+t_n -T)$, where $t\geq0$.
1757: Due to Lemma~\ref{l:compact}, $\Dc([0,\infty))$ is compact in
1758: $C([0, \infty);\Hw)$. Hence,
1759: passing to a subsequence and dropping a subindex,
1760: we can assume that $u_n$ coverges
1761: to some $u\in \Dc ([0,\infty))$ in $C([0, \infty); \Hw)$ as $n\to \infty$.
1762: By the definition of the weak global attractor, $u(t) \in \Aw$ for all $t \in [0, \infty)$.
1763: Applying Lemma~\ref{l:contona} with $(T_1,T_2)=(0,T)$, we obtain that
1764: there exists a complete trajectory
1765: $v\in \Dc((-\infty, \infty))$, such that $u(t)=v(t)$ on $[T/2, \infty)$. Therefore,
1766: \[
1767: u \in C([T/2, \infty); H).
1768: \]
1769: Hence, Corollary~\ref{c:gaex} yields that $u_n(T) \to u(T)$ strongly in $H$,
1770: i.e., $x_n  \to u(T)$  strongly in $H$.
1771: \end{proof}
1772: 
1773: We can now conclude with the two main results of this section. First, as
1774: a direct consequence of Theorem~\ref{t:asymptoticcompact} and
1775: Lemma~\ref{l:cont-comp} we obtain
1776: 
1777: \begin{theorem} \label{thm:main}
1778: If  all solutions on the weak global attractor are strongly continuous
1779: in $H$, then the
1780: strong global attractor $\As$ exists, is strongly compact, and coincides with $\Aw$.
1781: \end{theorem} 
1782: \begin{proof}
1783: Due to Lemma~\ref{l:cont-comp} we have that $R(t)$ is asymptotically compact.  Then Theorem~\ref{t:asymptoticcompact} implies that
1784: $\As$ exists, is strongly compact, and coincides with $\Aw$.
1785: \end{proof}
1786: 
1787: Second, we prove that the condition
1788: $\Dc((-\infty, \infty)) \subset C((-\infty, \infty);H)$
1789: is equivalent to a condition that all the ``energy jumps'' 
1790: uniformly converge to zero as time goes to infinity. More precisely,
1791: let
1792: \[
1793: [R(t)X]:= \sup \{[u(t)] : u \in \Dc([0,\infty))\}.
1794: \]
1795: Then we have the following.
1796: \begin{theorem}
1797: $\Dc((-\infty, \infty)) \subset C((-\infty, \infty); H)$ if and only
1798: if $[R(t)X] \to 0$ as $t \to \infty$.
1799: \end{theorem} 
1800: \begin{proof}
1801: It is obvious that if  $[R(t)X] \to 0$ as $t \to \infty$, then we have
1802: $\Dc((-\infty, \infty)) \subset C((-\infty, \infty); H)$.
1803: 
1804: Assume now that $\Dc((-\infty, \infty)) \subset C((-\infty, \infty); H)$,
1805: but $[R(t)X]$ does not converge to $0$ as $t \to \infty$.
1806: Then there exist a sequence of Leray-Hopf solutions
1807: $v_n \in \Dc([0,\infty))$
1808: and a time sequence $t_n \to \infty$ as $n\to \infty$, such that 
1809: \begin{equation} \label{a:faraway}
1810: \limsup_{n \to \infty}[v_n(t_n)] >0.
1811: \end{equation}
1812: Proceeding as in the proof of Lemma~\ref{l:cont-comp}, we can now
1813: assume that there exists some $T>0$, such that the sequence
1814: $u_n(t)=v_n(t+t_n-T)$ (where $t\geq 0$) converges in $C([0,\infty);\Hw)$
1815: to the restriction $u=v|_{[0,\infty)}$ of some complete trajectory
1816: $v \in \Dc((-\infty,\infty))$. Since $v(t)$ is continuous in $H$,
1817: by Theorem~\ref{thm:weak-jumps} we must have
1818: \[
1819: \lim_{n \to \infty} [v_n(t_n)] =\lim_{n \to \infty} [u_n(T)]=[u(T)]=[v(T)]=0, 
1820: \]
1821: a contradiction.
1822: 
1823: \end{proof}
1824: 
1825: \subsection{Regular part of the global attractor}
1826: We define the regular part of the weak global attractor, first introduced is
1827: \cite{FT85}, as follows.
1828: \[
1829: \begin{aligned}
1830: \Areg := \left\{ \right.&u_0:  \exists \tau>0, u \in \Dc((-\infty, \infty)) \mbox{ with }
1831: u(0)=u_0, \mbox{ such that }
1832: u(t) \mbox{ is regular} \\ &\mbox{on } (-\tau, \tau),
1833: \mbox{ and for each } \tilde{u} \in \Dc((-\infty, \infty))  \mbox{ with } \tilde{u}(0)=u(0) \mbox{ we have }\\
1834: & u(t)=\tilde{u}(t) \  \forall t \in(-\tau, \tau) \left. \right\}.
1835: \end{aligned}
1836: \]
1837: 
1838: The following result was proven in \cite{FT85}:
1839: \begin{theorem} The regular part of the global attractor satisfies the
1840: following properties:
1841: \begin{enumerate}
1842: \item[(a)] $\Areg$ is weakly open in $\Aw$,
1843: \item[(b)] $\Areg$ is weakly dense in $\Aw$,
1844: \item[(c)] If all solutions in  $\Aw$ are regular, then $\Aw$ is bounded in $V$
1845: (hence, $\Areg=\Aw$). 
1846: \end{enumerate}
1847: \end{theorem}
1848: Part $(c)$ was misstated in \cite{FT85} as
1849: \begin{enumerate}
1850: {\it \item[($c'$)] If $\Aw \subset V$, then $\Areg=\Aw$.}
1851: \end{enumerate}
1852: However, the proof provided in \cite{FT85} yields only $(c)$. As yet it is not
1853: known whether $(c')$ is true.
1854: 
1855: Now we will further study the case when all weak solutions on the
1856: weak global attractor of the 3D NSE are continuous in $H$. Under this
1857: assumption, Theorem~\ref{thm:main} implies that the strong compact
1858: global attractor $\As$ exists and $\As=\Aw$. Moreover,
1859: 
1860: \begin{theorem} If $\Dc((-\infty, \infty)) \subset C((-\infty, \infty); H)$,
1861: then the regular part of the global attractor $\Areg$
1862: is strongly dense in $\As$.
1863: \end{theorem}
1864: \begin{proof}
1865: Due to Theorem~\ref{thm:main} $\As$ exists and is strongly compact.
1866: Therefore, weak and strong topologies are equivalent on $\As$.
1867: Then since $\Areg$ is weakly dense in $\Aw$, it is also strongly dense
1868: in $\As=\Aw$.
1869: \end{proof}
1870: 
1871: We conclude this section with the following remark. In the case when the cubic
1872: box $\Omega=[0,L]^3$ is replaced by the cuboid $\Omega=[0,L]^2 \times [0,l]$
1873: with $0<l \ll L$, there exists a function $\alpha(g)>0$, such that if
1874: \begin{equation} \label{e:condonlL}
1875: \frac{l}{L} \leq \alpha(g),
1876: \end{equation}
1877: then $\Aw$ is the strong global attractor and is regular (see \cite{RS,TZ}). It would be interesting to
1878: show that for some values of $l/L$ larger than $\alpha(g)$, all the Leray-Hopf solutions
1879: on $\Aw$ are strongly continuous, i.e., $\Aw$ is also the strong global attractor.
1880: 
1881: \section{Tridiagonal models for the Navier-Stokes equations}
1882: In this section we introduce a two-parameter family of new simple models
1883: for the Navier-Stokes equations with a nonlinear term enjoying the
1884: same basic properties as the nonlinear term $B(u,u)$ in the NSE \eqref{NSE}.
1885: 
1886: The role of the space $H$ will be played by $l^2$ with the usual inner
1887: product and norm:
1888: \[
1889: (u,v)= \sum_{n=1}^{\infty} u_nv_n, \qquad |u|=\sqrt{(u,u)}.
1890: \]
1891: The norm $|u|$ will be called the energy norm.
1892: Let $A:D(A) \to H$ be the Laplace operator defined by
1893: \[
1894: (Au)_n = n^{\alpha} u_n, \qquad n\geq 1,
1895: \]
1896: for some $\alpha >0$. The domain $D(A)$ of this operator is
1897: \[
1898: \left\{u: \sum_{n=1}^\infty n^{2\alpha}u_n^2 <\infty \right\}.
1899: \]
1900: Clearly, $D(A)$ is dense in $H$ and 
1901: $A$ is a positive definite operator whose eigenvalues are
1902: \[
1903: 1, 2^{\alpha},  3^{\alpha}, \dots
1904: \]
1905: 
1906: Let $V=A^{-1/2}H$ endowed with the following inner product and norm:
1907: \[
1908: ((u,v))= \sum_{n=1}^{\infty} n^{\alpha}u_nv_n, \qquad \|u\|=\sqrt{((u,u))}.
1909: \]
1910: Here $\|u\|$ is an analog of $H^1$-norm of $u$ and we will call it the enstrophy
1911: norm. Let also
1912: \[
1913: \|u\|_\gamma = \left(\sum_{n=1}^{\infty} n^\gamma u_n^2 \right)^{1/2},
1914: \]
1915: which is an analog of $H^{\gamma/\alpha}$-norm of $u$.
1916: 
1917: Our models for the NSE are given by the
1918: following equations:
1919: \begin{equation} \label{model}
1920: \left\{
1921: \begin{aligned}
1922: &\ddt u_n + \nu n^{\alpha}u_n - n^\beta u_{n-1}^2 + (n+1)^\beta u_n u_{n+1}
1923: =g_n, \qquad n=1,2,3\dots\\
1924: &u_0=0.
1925: \end{aligned}
1926: \right.
1927: \end{equation}
1928: Here, $\nu>0$, $\alpha>0$, and $\beta>1$. Note that the value of $\ddt u_n$ is determined
1929: only by the values of $u_{n-1}$, $u_n$, and $u_{n+1}$. Therefore, we will
1930: refer to the equations \eqref{model} as the tridiagonal model for the Navier-Stokes
1931: equations or shortly TNS equations.
1932: For $u=(u_1,u_2,\dots)$ they can be written in a more condensed form as
1933: \begin{equation}
1934: \ddt u + \nu Au + B(u,u) = g,
1935: \end{equation}
1936: where
1937: \[
1938: (B(u,v))_n=- n^\beta u_{n-1} v_{n-1} + (n+1)^\beta u_n v_{n+1},
1939: \]
1940: and $u_0=0$. Note that the orthogonality property holds for $B$:
1941: \begin{eqnarray*}
1942: \left(B(u,v),v\right)&=&\sum_{n=1}^{\infty}\left(
1943: -n^{\beta}u_{n-1} v_{n-1} v_n + (n+1)^\beta u_n v_{n+1} v_n \right)\\
1944: &=&
1945: \sum_{n=1}^{\infty}\left(
1946: -n^{\beta}u_{n-1} v_{n-1} v_n +n^\beta u_{n-1} v_n v_{n-1} \right)\\
1947: &=&0.
1948: \end{eqnarray*}
1949: 
1950: In the case of TNS equations, 
1951: a weak solution on $[T,\infty)$ (or $(-\infty, \infty)$, if
1952: $T=-\infty$) of \eqref{model} is actually a locally bounded $H$-valued
1953: function $u(t)$ on $[T, \infty)$, such that $u_n \in C^1([T,\infty))$
1954: and $u_n(t)$ satisfies \eqref{model} for all $n$. From now on weak solutions will
1955: be called just solutions.
1956: 
1957: A solution $u(t)$ is
1958: strong (or regular) on some interval $[T_1,T_2]$, if $\|u(t)\|$ is bounded
1959: on $[T_1,T_2]$. A solution is strong on $[T_1,\infty)$, if it is strong on every
1960: interval $[T_1,T_2]$, $T_2\geq 0$.
1961: 
1962: A Leray-Hopf solution of \eqref{model} on the interval $[T, \infty)$
1963: is a solution of \eqref{model} on $[T,\infty)$ satisfying the
1964: energy inequality
1965: \[
1966: |u(t)|^2 + 2\nu \int_{t_0}^t \|u(\tau)\|^2 \, d\tau \leq
1967: |u(t_0)|^2 + 2\int_{t_0}^t (g, u(\tau)) \, d\tau,
1968: \]
1969: for all $T \leq t_0 \leq t$, $t_0$ a.e. in $[T,\infty)$.
1970: The set $Ex$ of those $t_0$ for which the energy
1971: inequality does not hold will be called the exceptional set.
1972: 
1973: 
1974: \subsection{A priori estimates and the existence of strong solutions}
1975: Taking a limit of the Galerkin approximation, the existence of
1976: Leray-Hopf solutions follows in exactly the same way as for the 3D NSE.
1977: In this paper we will show some {\it a priori} estimates, which
1978: can be obtained rigurously for the Leray-Hopf solutions.
1979: For simplicity, we assume that $g$ is independent of time, $g\in H$, and
1980: $g_n \geq 0$ for all $n$.
1981: \\
1982: 
1983: \noindent
1984: {\bf Energy estimates.} Formally taking a scalar product of (\ref{model}) with $u$,
1985: we obtain
1986: \[
1987: \begin{split}
1988: \frac{1}{2} \ddt |u|^2  &\leq  -\nu \|u\|^2 + |g||u|\\
1989: &\leq  -\nu |u |^2 + \frac{\nu}{2}|u|^2 + \frac{|g|^2}{2\nu}\\
1990: &= -\frac{\nu}{2}|u|^2 + \frac{|g|^2}{2\nu}.
1991: \end{split}
1992: \]
1993: Using Gronwall's inequality, we conclude that
1994: \begin{equation} \label{eq:defK}
1995: |u(t)|^2 \leq e^{- \nu t} |u(0)|^2 +
1996: \frac{|g|^2}{\nu^2}(1-e^{-\nu t}).
1997: \end{equation}
1998: Hence $B=\{u\in H: \ |u| \leq R\}$ is an absorbing ball for the
1999: Leray-Hopf solutions, where $R$ is any number larger that $|g|/\nu$.
2000: \\
2001: 
2002: \noindent
2003: {\bf Enstrophy estimates.} Let $v=A^{1/2}u$ and 
2004: \[
2005: c_\mathrm{b} := \left\{
2006: \begin{split}
2007: &\alpha 2^\beta, &0<\alpha \leq 1,\\
2008: &\alpha 2^{\alpha+\beta-1}, &\alpha > 1.
2009: \end{split}
2010: \right.
2011: \]
2012: Using H\"older's inequality, we obtain the following estimate
2013: for the nonlinear term:
2014: \[
2015: \begin{split}
2016: | (B(u,u),Au) | &\leq
2017: \left| \sum_{n=1}^\infty \left[(n+1)^{\alpha}-n^{\alpha} \right]
2018: (n+1)^{\beta }u_n^2 u_{n+1} \right|\\
2019: &\leq c_\mathrm{b} \sum_{n=1}^\infty n^{\beta-\alpha/2-1} |v_n|^2 |v_{n+1}|\\
2020: &\leq c_\mathrm{b} (\max_n |v_n|) \sum_{n=1}^{\infty}
2021: n^{\beta-\alpha/2-1} v_n^2 \\
2022: &\leq c_\mathrm{b} |v| |A^{1/2} v|^{2\beta/\alpha-2/\alpha-1} |v|^{-2\beta/\alpha+
2023: 2/\alpha+3}\\
2024: &= c_\mathrm{b} |A^{1/2} v |^{2\beta/\alpha-2/\alpha-1}
2025: |v|^{-2\beta/\alpha+2/\alpha+4}\\
2026: &= c_\mathrm{b} |Au |^{2\beta/\alpha-2/\alpha-1}
2027: \|u\|^{-2\beta/\alpha+2/\alpha+4},
2028: \end{split}
2029: \]
2030: whenever $\beta\in[\alpha/2+1,3\alpha/2+1]$.
2031: Choosing $u$ to have only two consecutive nonzero terms, it is easy to
2032: check that this estimate is sharp. Moreover, when $\alpha=2/3$ and
2033: $\beta=11/6$, we have
2034: \[
2035: | (B(u,u),Au) | \leq c_\mathrm{b} |Au |^{3/2}
2036: \|u\|^{3/2},
2037: \]
2038: which is exactly what Sobolev estimates give for the 3D NSE.
2039: Therefore, formally taking a scalar product of (\ref{model}) with $Au$,
2040: we obtain
2041: \[
2042: \begin{split}
2043: \frac{1}{2}\ddt \|u\|^2 &\leq -\nu|Au|^2 + c_\mathrm{b} |Au |^{3/2}
2044: \|u\|^{3/2} + (g, Au)\\
2045: &\leq -\nu|Au| + \frac{\nu}{3} |Au|^2 + \frac{3^6c_\mathrm{b}^4}{2^8\nu^3} \|u\|^6 +
2046: \frac{3}{4\nu} |g|^2 + \frac{\nu}{3} |Au|^2\\
2047: &\leq -\frac{\nu}{3}|Au|^2 + \frac{3^6c_\mathrm{b}^4}{2^8\nu^3} \|u\|^6 +
2048: \frac{3}{4\nu} |g|^2,
2049: \end{split}
2050: \]
2051: a Riccati-type equation for $\|u\|^2$. Hence, the model has the same
2052: enstrophy estimates as the 3D NSE, similar properties, and the same open question concerning the regularity of the solutions in the case
2053: $(\alpha, \beta) = (2/3, 11/6)$. In particular, we have a global existence of
2054: Leray-Hopf weak solutions (see Theorem~\ref{thm:Leray}), local existence of
2055: strong solutions (see Theorem~\ref{existence}), Leray's
2056: structure theorem (see Theorem~\ref{structure}),
2057: uniqueness of strong solutions in the class of
2058: Leray-Hopf solutions (see Theorem~\ref{uniqueness}), and existence of
2059: a weak global attractor (see Theorem~\ref{t:AweakNSE}).
2060: 
2061: In the case $(\alpha,\beta) = (1/2, 7/4)$, we have 
2062: \[
2063: | (B(u,u),Au) | \leq c_\mathrm{b} |Au |^{2}
2064: \|u\|,
2065: \]
2066: which corresponds to the 4D Navier-Stokes equations.
2067: In the case $(\alpha,\beta) = (2/5, 17/10)$, we have 
2068: \[
2069: | (B(u,u),Au) | \leq c_\mathrm{b} |Au |^{5/2} \|u\|^{1/2},
2070: \]
2071: which corresponds to the 5D Navier-Stokes equations.
2072: In general, the choice
2073: \[
2074: \alpha = \frac{2}{d}, \qquad \beta = \frac{3}{2} + \frac{1}{d}
2075: \]
2076: would correspond to the d-dimensional Navier-Stokes equations.
2077: 
2078: 
2079: The similarity of the TNS equations \eqref{model} with the NSE holds also
2080: for values of $d\ne 3$. Indeed,
2081: when $2\beta < 3\alpha +2$, the enstrophy estimate implies a local existence of strong solutions, i.e.,
2082: solutions whose enstrophy norms are continuous.
2083: More precisely, 
2084: if $2\beta < 3\alpha +2$, then for any initial data
2085: $u_0\in V$, there exists a
2086: strong solution $u(t)$ with $u(0)=u_0$ on some interval $[0,T]$. 
2087: In terms of the dimension,
2088: a sufficient condition for the local existence of strong solutions is
2089: $d<4.$
2090: In the case when $\beta \leq \alpha +1$, the enstrophy estimate implies
2091: a global regularity. In terms of the dimension,
2092: a sufficient condition for the global existence of strong solutions is
2093: $d\leq2$.
2094: 
2095: Now we will concentrate on the solutions with initial data $u_n(0)\geq0$
2096: for all $n$.
2097: 
2098: \begin{theorem} \label{t:poditofsol}
2099: Let $u(t)$ be a solution of \eqref{model} with $u_n(0)\geq 0$.
2100: Then $u_n(t) \geq 0$ for all $t>0$, and 
2101: $u(t)$ satisfies the energy inequality
2102: \begin{equation} \label{ee}
2103: |u(t)|^2 + 2\nu \int_{t_0}^t \|u(\tau)\|^2 \, d\tau \leq
2104: |u(t_0)|^2 + 2\int_{t_0}^t (g, u(\tau)) \, d\tau
2105: \end{equation}
2106: for all $0 \leq t_0 \leq t$.
2107: \end{theorem}
2108: \begin{proof}
2109: A general solution for $u_n(t)$ can be written as
2110: \begin{multline*}
2111: u_n(t)=u_n(0)\exp\left(-
2112: \int_{0}^t\left[\nu n^\alpha + (n+1)^\beta u_{n+1}(\tau)\right] \, d\tau\right)\\
2113: + \int_{0}^t(g_n+n^\beta u_{n-1}^2(s) )\exp\left(-\int_{s}^{t}
2114: \left[ \nu n^\alpha+(n+1)^\beta u_{n+1}(\tau)\right] \, d\tau \right) \, ds.
2115: \end{multline*}
2116: Since $u_n(0) \geq 0$ for all $n$, then  $u_n(t)\geq 0$ for all $n$, $t>0$.
2117: Hence, multiplying \eqref{model} by $u_n$,
2118: taking a sum from $1$ to $N$, and integrating between $t_0$ and $t$,
2119: we obtain
2120: \[
2121: \begin{split}
2122: \sum_{n=1}^N u_n(t)^2  -& \sum_{n=1}^N u_n(t_0)^2
2123:  +2\nu \int_{t_0}^t \sum_{n=1}^N n^\alpha u_n^2 \, d\tau \\
2124: &= - 2\int_{t_0}^t  (N+1)^\beta u_N^2 u_{N+1} \, d\tau
2125: +2\int_{t_0}^t\sum_{n=1}^N  g_n u_n \, d\tau\\
2126: &\leq 2\int_{t_0}^t\sum_{n=1}^N g_n u_n \, d\tau.
2127: \end{split}
2128: \]
2129: Taking the limit as $N \to \infty$, we obtain \eqref{ee}.
2130: 
2131: \end{proof}
2132: 
2133: \subsection{Blow-up in finite time} \label{sub:finitetime}
2134: Here, when $2\beta -3\alpha -3 >0$ and $g_1>0$ is large enough,
2135: we will show that every solution $u(t)$ of \eqref{model} with $u_n(0) \geq 0$
2136: blows up in finite time in an appropriate norm.
2137: First, we need the following two lemmas.
2138: 
2139: %%%%%%%%%%%% LEMMA %%%%%%%%%%%%%%
2140: \begin{lemma} \label{l:big}
2141: Let $u(t)$ be a solution to \eqref{model} on $[0,\infty)$ with $u_n(0) \geq 0$
2142: for all $n$. Assume that
2143: $\|u(t)\|_{2(\beta+\gamma-1)/3} \in L^3_{\mathrm{loc}}([0,\infty);\mathbb{R})$.
2144: Then
2145: \begin{equation} \label{e:tmp777}
2146: \int_{t_0}^t \sum_{n=1}^\infty n^{\beta+\gamma-1}u_n^2u_{n+1} \, d\tau < \infty,
2147: \qquad \int_{t_0}^t \sum_{n=1}^\infty n^{\beta+\gamma-1}u_n^3
2148: \, d\tau < \infty,
2149: \end{equation}
2150: and
2151: \begin{equation} \label{e:tmp888}
2152: \|u(t)\|_\gamma^2 - \|u(t_0)\|_\gamma^2 + 2\nu \int_{t_0}^{t} \|u\|_{\alpha+\gamma}^2 \, d\tau \geq
2153: 2\gamma \int_{t_0}^{t} \sum_{n=1}^{\infty} (n+1)^{\beta+\gamma-1}u_n^2 u_{n+1}
2154: \, d\tau 
2155: \end{equation}
2156: for all $0\leq t_0 \leq t$, $0<\gamma\leq1$.
2157: \end{lemma}
2158: \begin{proof}
2159: Thanks to Theorem~\ref{t:poditofsol}, $u_n(t) \geq 0$ for all $n, t>0$.
2160: Since $\|u(t)\|^3_{2(\beta+\gamma-1)/3}$ is integrable on $[t_0,t]$  for all
2161: $0\leq t_0\leq t$, we obtain
2162: \begin{equation*} 
2163: \begin{split}
2164: \int_{t_0}^t \sum_{n=1}^\infty n^{\beta+\gamma-1}u_n^2u_{n+1} \, d\tau &\leq
2165: 2\int_{t_0}^t \sum_{n=1}^\infty n^{\beta+\gamma-1}u_n^3 \, d\tau\\
2166: &\leq 2\int_{t_0}^t \left(\sum_{n=1}^\infty
2167: n^{\frac{2}{3}(\beta+\gamma-1)}u_n^2\right)^{3/2} \, d\tau\\
2168: &= 2 \int_{t_0}^{t} \|u\|_{2(\beta+\gamma-1)/3}^3 \, d\tau\\
2169: &< \infty.
2170: \end{split}
2171: \end{equation*}
2172: Hence, the relations in \eqref{e:tmp777} hold. In particular,
2173: \begin{equation} \label{integrq}
2174: \liminf_{n \to \infty} \int_{t_0}^t n^{\beta+\gamma} u_n^2 u_{n+1} \, d\tau =0.
2175: \end{equation}
2176: Now multiplying \eqref{model} by
2177: $n^\gamma u_n$, taking a sum from $1$ to $N$, and integrating from
2178: $t_0$ to $t$, we obtain
2179: \begin{equation*} 
2180: \begin{split}
2181: \sum_{n=1}^N & n^{\gamma}u_n(t)^2  - \sum_{n=1}^N n^{\gamma} u_n(t_0)^2
2182: +2\nu \int_{t_0}^{t} \sum_{n=1}^N n^{\alpha+\gamma} u_n^2 \, d\tau \\
2183: &= 2\int_{t_0}^{t} \sum_{n=1}^{N-1} (n+1)^\beta((n+1)^\gamma -n^\gamma)u_n^2 u_{n+1}
2184: \, d\tau\\
2185: & \quad - 2\int_{t_0}^{t} (N+1)^\beta N^\gamma u_N^2 u_{N+1} \, d\tau
2186: + 2\int_{t_0}^{t}\sum_{n=1}^N n^\gamma g_n u_n \, d\tau\\
2187: &\geq 2\gamma \int_{t_0}^{t} \sum_{n=1}^{N-1} (n+1)^{\beta+\gamma-1}u_n^2 u_{n+1}
2188: \, d\tau
2189:  - 2\int_{t_0}^{t} (N+1)^\beta N^\gamma u_N^2 u_{N+1} \, d\tau
2190: \end{split}
2191: \end{equation*}
2192: Thanks to \eqref{integrq}, taking the lower limit as $N\to \infty$, we get
2193: \eqref{e:tmp888}.
2194: 
2195: \end{proof}
2196: 
2197: 
2198: \begin{lemma} \label{l:doublelarge}
2199: For any $c>0$, there exists $g_1>0$, such that 
2200: \[
2201: \int_t^{t+1} |u|^2 \, d\tau > c, \qquad \forall t\geq0,
2202: \]
2203: for all solutions $u(t)$ with $g=(g_1,g_2,\dots)$.
2204: \end{lemma}
2205: \begin{proof}
2206: Take any $c>0$. Thanks to Theorem~\ref{t:poditofsol}, $u_n(t) \geq 0$ for all $n, t>0$.
2207: Therefore,
2208: \[
2209: u_1(t+1/2) \geq u_1(t) - \nu \int_t^{t+1/2} u_1 \, d\tau -
2210: 2^\beta \int_t^{t+1/2} u_1u_2 \, d\tau +g_1.
2211: \]
2212: Now integrating this inequality over $[t,t+1/2]$, we obtain
2213: \[
2214: \int_t^{t+1} u_1 \, d\tau + \nu \int_t^{t+1} u_1 \, d\tau +
2215: 2^{\beta-1} \int_t^{t+1} u_1^2 \, d\tau + 2^{\beta-1}
2216: \int_t^{t+1} u_2^2 \, d\tau  \geq \frac{g_1}{2}. 
2217: \]
2218: Hence, using Cauchy-Schwarz inequality, we get
2219: \[
2220: (\nu+1) \left(\int_t^{t+1} (u_1^2 +u_2^2)\, d\tau \right)^{1/2} + 2^{\beta-1}
2221: \int_t^{t+1} (u_1^2 + u_2^2) \, d\tau \geq \frac{g_1}{2}. 
2222: \]
2223: Obviously, for $g_1$ large enough, we have that
2224: \[
2225: \int_t^{t+1} |u|^2 \, d\tau \geq \int_t^{t+1} (u_1^2+u_2^2) \, d\tau > c.
2226: \]
2227: \begin{comment}
2228: Note that
2229: \begin{eqnarray*}
2230: u_n(t+1) \geq u_n(t) \exp\left(-2\nu n^{\alpha} - (n+1)^\beta
2231: \int_t^{t+1} u_{n+1}(\tau) \, d\tau  \right)\\
2232: +\int_t^{t+1} g_n \exp\left(-2\nu n^{\alpha}(t+2-s) - (n+1)^\beta
2233: \int_t^{t+1-s} u_{n+1}(\tau) \, d\tau  \right) \, ds
2234: \end{eqnarray*}
2235: \end{comment}
2236: \end{proof}
2237: 
2238: Now we proceed to our main result in this section.
2239: 
2240: \begin{theorem} \label{thm:blowup}
2241: For every solution $u(t)$ to equation (\ref{model}) with $u_n(0)\geq0$, $\nu>0$, $2\beta - 3\alpha -3>0$, and $g_1$ large enough,
2242: $\|u(t)\|^3_{2(\beta+\gamma-1)/3}$ is not locally integrable for all $\gamma>0$.
2243: \end{theorem}
2244: \begin{proof}
2245: Thanks to Theorem~\ref{t:poditofsol}, $u_n(t) \geq 0$ for all $n, t>0$.
2246: Assume that there exist 
2247: \[
2248: \gamma\in(0,\min\{2\beta -3\alpha -3, 1\})
2249: \]
2250: and a solution $u(t)$ to \eqref{model}, such that
2251: $\|u(t)\|_{2(\beta+\gamma-1)/3} \in L^3_{\mathrm{loc}}([0,\infty);\mathbb{R})$.
2252: Then Lemma~\ref{l:big} implies that
2253: \[
2254:  \int_0^T \sum_{n=1}^\infty n^{\beta+\gamma-1}u_n^2u_{n+1} \, d\tau < \infty
2255: \qquad \text{and} \qquad  \int_0^T \sum_{n=1}^\infty n^{\beta+\gamma-1}u_n^3
2256: \, d\tau < \infty,
2257: \]
2258: for all $T>0$.
2259: Moreover, since $\alpha + \gamma < 2(\beta + \gamma -1)/3$, we have that
2260: $\|u(t)\|^2_{\alpha +\gamma}$ is locally integrable on $[0,\infty)$.
2261: 
2262: Now note that if $u_{n+1} \geq 2u_n$, then $u_nu_{n+1}^2 \leq \frac{1}{2}u_{n+1}^3$.
2263: Otherwise, $u_nu_{n+1}^2 \leq 2 u_n^2 u_{n+1}$. Hence,
2264: \begin{equation} \label{eq:ineq1}
2265: u_nu_{n+1}^2 \leq {\ts \frac{1}{2}}u_{n+1}^3 + 2 u_n^2 u_{n+1}, \qquad n \in \mathbb{N}.
2266: \end{equation}
2267: This also implies
2268: \begin{equation} \label{eq:ineq2}
2269: \begin{split}
2270: u_nu_{n+1} u_{n+2} &\leq {\ts \frac{1}{2}}u_n^2u_{n+1} + {\ts \frac{1}{2}}u_{n+1}u_{n+2}^2
2271: \\
2272: &\leq {\ts \frac{1}{2}}u_n^2u_{n+1} + {\ts \frac{1}{4}}u_{n+2}^3 + u_{n+1}^2u_{n+2},
2273: \end{split}
2274: \end{equation}
2275: for all $n \in \mathbb{N}$. From \eqref{model} we have
2276: \begin{multline*}
2277: \ddt(u_nu_{n+1}) =
2278:   -\nu (n^\alpha + (n+1)^\alpha)
2279: u_nu_{n+1}
2280:  +n^\beta u_{n-1}^2u_{n+1}- (n+1)^\beta u_nu_{n+1}^2\\
2281:   +(n+1)^\beta u_n^3- (n+2)^\beta u_nu_{n+1}u_{n+2} +u_ng_{n+1}+u_{n+1}g_n.
2282: \end{multline*}
2283: From this, using inequalities (\ref{eq:ineq1}) and (\ref{eq:ineq2}), we obtain
2284: \begin{multline} \label{e:ineq11}
2285: \sum_{n=1}^\infty (n+1)^{\gamma-1}(u_nu_{n+1})(t+1)-\sum_{n=1}^\infty (n+1)^{\gamma-1}(u_nu_{n+1})(t) \\
2286: + 2\nu \int_t^{t+1} \sum_{n=1}^\infty (n+1)^{\alpha+\gamma-1}  u_nu_{n+1} \, d\tau
2287: \\
2288: + (3+(3/2)^\beta) \int_t^{t+1}\sum_{n=1}^\infty (n+1)^{\beta+\gamma-1}  u_n^2u_{n+1} \, d\tau \\
2289: \geq \frac{1}{4}\int_t^{t+1}\sum_{n=1}^\infty n^{\beta+\gamma -1} u_n^3 \, d\tau, 
2290: \end{multline}
2291: for all $t>0$.
2292: On the other hand, Lemma~\ref{l:big} yields
2293: \begin{equation} \label{e:ineq22}
2294: \|u(t+1)\|^2_\gamma - \|u(t)\|^2_\gamma + 2\nu \int_t^{t+1} \|u\|_{\alpha+\gamma}^2 \, d\tau \geq
2295: 2\gamma \int_t^{t+1} \sum_{n=1}^{\infty} (n+1)^{\beta+\gamma-1}u_n^2 u_{n+1}
2296: \, d\tau 
2297: \end{equation}
2298: for all $t>0$.
2299: Denote
2300: \[
2301: \Theta(t) = \int_t^{t+1}\|u(\tau)\|_{\gamma}^2 \, d\tau + 
2302: \frac{2\gamma}{3+(3/2)^\beta} \int_t^{t+1}
2303: \sum_{n=1}^{\infty} (n+1)^{\gamma-1}(u_nu_{n+1})(\tau) \, d\tau.
2304: \]
2305: Note that $\Theta(t)$ is absolutely continuous on $[0,\infty)$.
2306: We will show that $\Theta(t)$ is a Lyapunov function for the equation, i.e.,
2307: $\Theta(t)$ is always increasing. Indeed, multiplying the inequality
2308: (\ref{e:ineq11}) by $2\gamma/(3+(3/2)^\beta)$ and adding
2309: (\ref{e:ineq22}), we obtain
2310: \begin{multline*}
2311: \ddt \Theta(t) \geq -2\nu \int_t^{t+1} \|u(\tau)\|_{\alpha + \gamma}^2 \, d\tau
2312: - \frac{4\gamma \nu}{3+(3/2)^\beta}
2313: \int_t^{t+1}\sum_{n=1}^{\infty} (n+1)^{\alpha+\gamma-1}  u_nu_{n+1} \, d\tau \\
2314: + \frac{\gamma}{6+2(3/2)^\beta}
2315: \int_t^{t+1}\sum_{n=1}^{\infty}  n^{\beta+\gamma -1}u_n^3 \, d\tau,
2316: \end{multline*}
2317: a.e. on $(0,\infty)$.
2318: Since $\gamma$ is such that
2319: \[
2320: \epsilon:= 2\beta -3\alpha -\gamma -3 >0,
2321: \]
2322: let
2323: \[
2324: A:=\left(\sum_{n=1}^{\infty} n^{-1-\epsilon} \right)^{-1/2}.
2325: \]
2326: Now H\"older's inequality yields
2327: \[
2328: \begin{split}
2329: \int_t^{t+1} \sum_{n=1}^{\infty} n^{\alpha + \gamma} u_n^2 \, d\tau &\leq 
2330: \left(\int_t^{t+1} \sum_{n=1}^{\infty} n^{-1-\epsilon} \, d\tau\right)^{1/3}
2331: \left(\int_t^{t+1} \sum_{n=1}^{\infty} n^{\beta + \gamma -1} u_n^3 \, d\tau\right)^{2/3}
2332: \\
2333: &=
2334: A^{-2/3}\left(\int_t^{t+1} \sum_{n=1}^{\infty} n^{\beta + \gamma -1} u_n^3 \, d\tau\right)^{2/3}.
2335: \end{split}
2336: \]
2337: Hence,
2338: \[
2339: \int_t^{t+1} \sum_{n=1}^{\infty} n^{\beta + \gamma -1} u_n^3 \, d\tau \geq
2340: A \left(\int_t^{t+1} \sum_{n=1}^{\infty} n^{\alpha + \gamma} u_n^2 \, d\tau \right)^{3/2}.
2341: \]
2342: Finally, we obtain
2343: \begin{multline*}
2344: \ddt \Theta(t)
2345: \geq -2\nu\left( 1+ 2\gamma \frac{(3/2)^{\alpha+\gamma}}{3+(3/2)^\beta} \right)
2346: \int_t^{t+1} \|u(\tau)\|_{\alpha + \gamma}^2 \, d\tau\\
2347: + \frac{\gamma A}{6+2(3/2)^\beta}
2348: \left(\int_t^{t+1} \|u(\tau)\|_{\alpha + \gamma}^2 \, d\tau\right)^{3/2},
2349: \end{multline*}
2350: a.e. on $(0,\infty)$.
2351: Due to Lemma~\ref{l:doublelarge}, if $g_1$ is large enough, then there
2352: exists a positive constant $c$, such that
2353: \[
2354: \ddt \Theta(t) \geq c\Theta(t)^{3/2}, \qquad \text{a.e. on } (0,\infty).
2355: \]
2356: This is a Riccati-type equation. Hence, $\Theta(t)$ blows up in finite time,
2357: which contradicts the fact that it is continuous on $[0,\infty)$.
2358: \end{proof}
2359: 
2360: Figure~\ref{fig:regions} shows three regions, the ones where we were able
2361: to prove local regularity, global regularity, and  and blow-up in finite time.
2362: The labels $2D$, $3D$, and $4D$ show the dimensions of the 
2363: Navier-Stokes systems corresponding to the models at those points.
2364: 
2365: 
2366: \begin{figure}
2367: \center
2368: \includegraphics[width=4in]{regions.ps}
2369: \caption{Regions of local regularity, global regularity, and blow-up.} 
2370: \label{fig:regions}
2371: \end{figure}
2372: 
2373: 
2374: 
2375: \subsection{Non-regular weak global attractor}
2376: As in Subsection~\ref{s:weakattractor}, we can define an evolutionary
2377: system $\Dc$ whose trajectories are all Leray-Hopf solutions of the
2378: TNS equations.
2379: The weak global attractor for this system is
2380: \[
2381: \begin{split}
2382: \Aw = \{&u_0 \in H: \mbox{ there exists a Leray-Hopf solution } u(t) \mbox{ on }
2383: (-\infty,\infty),\\
2384: &\mbox{ such that } u(0)=u_0 \mbox{ and } |u(t)| \mbox{ is  bounded on }  (-\infty, \infty)\}.
2385: \end{split}
2386: \]
2387: 
2388: Recall that $g_n\geq 0$ for all $n\in \mathbb{N}$. Obviously, if $g=0$, then $\Aw=\{0\}$.
2389: Henceforth we will assume that $g \ne 0$.
2390: 
2391: \begin{theorem} \label{t:posA}
2392: If $g_n=0$ for all $n \geq N_g$, then
2393: every $u=(u_1,u_2,\dots) \in \Aw$ satisfies
2394: \[
2395: u_n \geq 0, \qquad n=1,2,\dots
2396: \]
2397: \end{theorem}
2398: \begin{proof}
2399: A general solution for $u_n(t)$ can be written as
2400: \begin{multline} \label{eq:gensolution}
2401: u_n(t)=u_n(t_0)\exp\left(-
2402: \int_{t_0}^t\nu n^\alpha + (n+1)^\beta u_{n+1}(\tau) \, d\tau\right)\\
2403: + \int_{t_0}^t\exp\left(-\int_{s}^{t}\nu n^\alpha+(n+1)^\beta
2404: u_{n+1}(\tau) \, d\tau\right) (g_n+n^\beta u_{n-1}^2(s) )\, ds.
2405: \end{multline}
2406: Clearly, this implies the following facts.
2407: \begin{enumerate}
2408: \item[(a)] If $u_n(t_0) \geq 0$ for some $n$ and $t_0$, then
2409: $u_n(t) \geq 0$ for all $t \geq t_0$.
2410: \item[(b)] If $|u(t)|$ is bounded for all $t\in\mathbb{R}$, then
2411: $u_n(t) \geq 0$ for all $t\in \mathbb{R}$, whenever
2412:  $u_{n+1}(t) \geq 0$ for all $t\in \mathbb{R}$.
2413: \end{enumerate}
2414: 
2415: Now assume that there exists  $u^0 \in \Aw$, such that $u^0_N < 0$
2416: for some $N \geq N_g$. Then there exists a Leray-Hopf solution $u(t)$,
2417: such that $u(0) = u^0$ and $|u(t)|$ is bounded on $(-\infty, \infty)$. For such a
2418: solution we have $u_N(t) <0$ for all $t \leq 0$.
2419: In addition, from the energy inequality for $u(t)$ we deduce that
2420: \[
2421: \begin{split}
2422: \sum_{n=N}^{\infty} u_n(t_1)^2 - \sum_{n=N}^{\infty} u_n(t_0)^2 &\leq
2423: 2\int_{t_0}^{t_1} \left[ N^\beta u_{N-1}(\tau)^2u_{N}(\tau)
2424:  -\nu \sum_{n=N}^\infty n^\alpha u_n(\tau)^2 \right] \, d\tau\\
2425:  &\leq  -2 \nu \int_{t_0}^{t_1} \sum_{n=N}^\infty u_n(\tau)^2 \, d\tau,
2426: \end{split}
2427: \]
2428: for all $t_0 \leq t_1 \leq 0$. Hence,
2429: \[
2430: \sum_{n=N}^{\infty} u_n(t_0)^2 \geq e^{-2\nu t_0}\sum_{n=N}^{\infty} u_n(0)^2,
2431: \]
2432: for all $t_0 \leq 0$. This implies that $|u(t)|^2$ is not bounded backwards
2433: in time, a contradiction.
2434: 
2435: Now take any $u^0 \in \Aw$. There exists a Leray-Hopf solution $u(t)$,
2436: such that $u(0) = u^0$ and $|u(t)|$ is bounded on $(-\infty, \infty)$.
2437: For such a solution we proved that
2438: \[
2439: u_n(t) \geq 0, \qquad \forall t\in\mathbb{R}, n \geq N_g.
2440: \]
2441: Now note that due to the remark (b) above, if $u_{n+1}(t) \geq 0$ for all
2442: $t\in \mathbb{R}$, then $u_n(t) \geq0$ for all $t\in \mathbb{R}$,
2443: which concludes the proof.
2444: \end{proof}
2445: 
2446: Now Theorem~\ref{t:posA} allows us to apply all the results in
2447: Subsection~\ref{sub:finitetime} to the solutions on the weak global
2448: attractor $\Aw$. Therefore, we have the following.
2449: \begin{remark} \label{rem:bl}
2450: Let $2\beta> 3\alpha +3$ and $g_1$ be large enough. Then
2451: $\|u(t)\|_{2(\beta+\gamma-1)/3}$ blows up in finite time
2452: for every solution $u(t)$ on $\Aw$, i.e., $\Aw$ is not bounded in
2453: $H^{2(\beta + \gamma-1)/(3\alpha)}$ for any $\gamma>0$.
2454: \end{remark}
2455: However, this does not mean that the weak global attractor cannot be strong.
2456: The question whether $\Aw$ is the strong global attractor remains open
2457: in the case $\beta > \alpha +1$.
2458: 
2459: 
2460: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2461: \subsection{Tridiagonal models for the Euler equations}
2462: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2463: In this section we consider the tridiagonal models for the Euler equations
2464: (TE), the equations \eqref{model} with $\nu=0$. First, let us show the
2465: global existence of the weak solutions to the TE equations.
2466: Take a sequence
2467: $\nu_j\to 0$ as $j\to \infty$. Given $u^0\in H$,
2468: let $u^j(t)$ be a solution of \eqref{model} with $\nu=\nu_j$ and $u^j(0)=u^0$.
2469: It is easy
2470: to show that the sequence $\{u^j\}$ is weakly equicontinuous. Therefore, thanks
2471: to Ascoli-Arzela theorem, passing to a subsequence and dropping
2472: a subindex, we obtain that there exists a function $u:[0,\infty) \to H$,
2473: such that $u^j \to u$ in $C([0,\infty);\Hw)$
2474: as $j\to \infty$. Clearly, $u(t)$ is a solution of the TE equations, in
2475: the sense that it is a locally bounded $H$-valued
2476: function on $[T, \infty)$, such that
2477: $u_n \in C^1([0,\infty))$ and $u_n(t)$ satisfies \eqref{model}
2478: for all $n$.
2479: 
2480: Now let us show that in the nonviscous case $\nu=0$, for every solution $u(t)$
2481: of \eqref{model}, the norm $\|u(t)\|_{2(\beta+\gamma-1)/3}$ blows up
2482: for any $\alpha>0$, $\beta>1$, and $\gamma>0$, reflecting the fact that there is no backward energy 
2483: transfer for this model.
2484: 
2485: \begin{theorem} \label{t:Eblow}
2486: Let $u(t)$ be a solution of \eqref{model} on $[0,\infty)$ with $\nu=0$,
2487: $g_n \geq 0$, $u_n(0) \geq 0$ for all $n$, and $u(0) \ne 0$.
2488: Then $\|u(t)\|_{2(\beta+\gamma-1)/3}$ is not bounded on $[0,\infty)$
2489: for every $\gamma >0$ .
2490: 
2491: \end{theorem}
2492: \begin{proof}
2493: Clearly, it is enough to prove the theorem in the case where
2494: $0<\gamma < \min\{1,2(\beta-1)\}$.
2495: Assume that $\|u(t)\|_{2(\beta+\gamma-1)/3}$ is bounded on $[0,\infty)$.
2496: Then Lemma~\ref{l:big} implies that
2497: \begin{equation} \label{eq:dobnorme}
2498: \|u(t)\|_\gamma^2 - \|u(t_0)\|_\gamma^2  \geq
2499: 2\gamma \int_{t_0}^{t} \sum_{n=1}^{\infty} (n+1)^{\beta+\gamma-1}u_n^2 u_{n+1}
2500: \, d\tau \geq 0, 
2501: \end{equation}
2502: for all $0\leq t_0 \leq t$. Thus, $\|u(t)\|^2_{\gamma}$ is non-decreasing. 
2503: Since $\gamma < 2(\beta+\gamma-1)/3$, $\|u(t)\|_{\gamma}$ is bounded on $[0,\infty)$. Then there
2504: exists $E_0>0$ such that
2505: \[
2506: \lim_{t \to \infty} \|u(t)\|_{\gamma}^2 = E_0.
2507: \]
2508: Then (\ref{eq:dobnorme}) implies that 
2509: \begin{equation} \label{eq:tempeuler}
2510: \lim_{t\to \infty} \int_t^\infty u_n(\tau)^2u_{n+1}(\tau) \, d\tau = 0, \qquad n\in \mathbb{N}. 
2511: \end{equation}
2512: Hence,
2513: \[
2514: \begin{split}
2515: u_n(t)^2 -u_n(0)^2&= 2n^\beta\int_0^tu_{n-1}^2u_n \, d\tau - 2(n+1)^\beta
2516: \int_0^t u_n^2u_{n+1} \, d\tau + 2\int_0^t g_n u_n \, d\tau \\
2517: &\to 2n^\beta\int_0^\infty u_{n-1}^2u_n \, d\tau - 2(n+1)^\beta
2518: \int_0^\infty u_n^2u_{n+1} \, d\tau + 2\int_0^\infty g_n u_n \, d\tau,
2519: \end{split}
2520: \]
2521: as $t\to \infty$. Hence $u_n(\infty) := \lim_{t\to \infty} u_n(t)$ exists for
2522: all $n$.
2523: Now (\ref{eq:tempeuler}) implies that $u_n(\infty) u_{n+1}(\infty)=0$ 
2524: for all $n$. Suppose that $u_k(\infty) \ne 0$ for some $k$. Then
2525: $u_{k+1}(\infty)=0$ and there exists $t_0>0$, such that
2526: \[
2527: (k+2)^{\beta}u_{k+1}(t)u_{k+2}(t) \leq  {\ts \frac{1}{3}}(k+1)^\beta u_k(\infty)^2 \qquad \mbox{and}
2528: \qquad u_k^2(t) \geq {\ts \frac{2}{3}}u_k(\infty)^2, 
2529: \]
2530: for all $t\geq t_0$. Thus,
2531: \[
2532: \begin{split}
2533: \ddt u_{k+1} &= (k+1)^\beta u_k(t)^2 - (k+2)^\beta u_{k+1}(t)u_{k+2}(t) +
2534: g_{k+1} \\
2535: &\geq {\ts \frac{1}{3}}(k+1)^\beta u_k(\infty)^2,
2536: \end{split}
2537: \]
2538: for all $n \geq t_0$. Therefore,
2539: \[
2540: \lim_{t \to \infty} u_{k+1}(t) =\infty,
2541: \]
2542: a contradiction.
2543: \end{proof}
2544: 
2545: \section*{Acknowledgment}
2546: We thank the referee for his constructive criticism and pertinent remarks.
2547: 
2548: \begin{comment}
2549: We conclude with the following result.
2550: \begin{theorem}
2551: If $2\beta > 3\alpha +3$,
2552: If $g_n=0$ for all $n \geq N_g$, and $g_1$ is large enough, then
2553: \[
2554: \sup_{t\in \mathbb{R}}u_n(t) \geq \frac{\nu n^{\beta-\alpha}}{(n+1)^{2(\beta-\alpha)}},
2555: \qquad \forall n\geq N_g,
2556: \]
2557: for each $u$, such that $u(t)\in \Aw$, $t\in\mathbb{R}$ .
2558: \end{theorem}
2559: \begin{proof}
2560: Take any weak solution with $|u(t)|$ bounded on $\mathbb{R}$. Let
2561: \[
2562: \bar{u}_n := \sup_{t\in\mathbb{R}} u_n(t)^2.
2563: \]
2564: Taking the limit as $t_0\to -\infty$ of (\ref{eq:gensolution}), we obtain
2565: \[
2566: u_n(t) \leq \frac{n^{\beta-\alpha}}{\nu}\bar{u}_{n-1}^2, \qquad n\geq N_g, \ t \in \mathbb{R}.
2567: \]
2568: Hence,
2569: \[
2570: \bar{u}_n \leq \frac{n^{\beta-\alpha}}{\nu}\bar{u}_{n-1}^2, \qquad n\geq N_g.
2571: \]
2572: Let
2573: \[
2574: v_n := \frac{(n+1)^{2(\beta-\alpha)}}{\nu n^{\beta-\alpha}}\bar{u}_n.
2575: \]
2576: Then we have
2577: \[
2578: v_n \leq v_{n-1}^2.
2579: \]
2580: Assume that $v_N <1$ for some $N \geq N_g$. Then
2581: \[
2582: v_n < v_N^{2(n-N)}, \qquad \mbox{and} \qquad \bar{u}_n \leq
2583: \frac{\nu n^{\beta-\alpha}}{(n+1)^{2(\beta-\alpha)}}v_N^{2(n-N)}.
2584: \]
2585: It is obvious that in this case $\|u(t)\|_{\theta}$  must be bounded on
2586: $\mathbb{R}$ for all $\theta>0$. On the other hand, 
2587: due to Theorem~\ref{thm:blowup}, 
2588: $\|u(t)\|_{2(\beta+\gamma-1)/3}$ blows up in finite time for any $\gamma>0$.
2589: Hence, $v_n\geq1$ for all $N \geq N_g$, i.e.,
2590: \[
2591: \bar{u}_n \geq \frac{\nu n^{\beta-\alpha}}{(n+1)^{2(\beta-\alpha)}}, \qquad n\geq N_g,
2592: \]
2593: which concludes the proof.
2594: \end{proof}
2595: \end{comment}
2596: \begin{thebibliography}{99}
2597: 
2598: 
2599: \bibitem{B1} J. M. Ball, Continuity properties and global attractors of generalized
2600: semiflows and the Navier-Stokes equations, {\it J. Nonlinear Sci.} {\bf7} (1997),
2601: 475Ð502. Erratum: {\it J. Nonlinear Sci.} {\bf 8} (1998),  233.
2602: 
2603: \bibitem{B2} J. M. Ball, Global attractors for damped semilinear wave equations,
2604: {\it Discr. Cont. Dyn. Sys.} {\bf10} (2004), 31Ð52.
2605: 
2606: \bibitem{CMR} T. Caraballo, P. Mar’n-Rubio, and J. C. Robinson,
2607: A comparison between two theories for multi-valued semiflows and their
2608: asymptotic behaviour, {\it Set-Valued Anal.} {\bf 11} (2003), 297--322.
2609: 
2610: \bibitem{CV} V. V. Chepyzhov and M. I. Vishik, {\it Attractors for Equations of
2611: Mathematical Physics}, American Mathematical Society Colloquium Publications
2612: {\bf 49}, American Mathematical Society, Providence, RI, 2002.
2613: 
2614: \bibitem{C} A. Cheskidov, Blow-up in finite time for the dyadic model of the
2615: Navier-Stokes equations, {\it  Trans. Amer. Math. Soc.}, to appear.
2616: 
2617: \bibitem{CF} P. Constantin and C. Foias, {\it Navier-Stokes Equation}, University of Chicago Press, Chicago, 1989. 
2618: 
2619: \bibitem{FS}F. Flandoli and B. Schmalfu\ss, 
2620: Weak solutions and attractors for three-dimensional Navier-Stokes equations with
2621: nonregular force, {\it J. Dynam. Differential Equations} {\bf 11} (1999),  355--398.
2622: 
2623: \bibitem{FMRT} C. Foias, O. P. Manley, R. Rosa, and R. Temam, {\it Navier-Stokes
2624: equatinon and Turbulence}, Encyclopedia of Mathematics and its Applications
2625: {\bf 83}, Cambridge University Press, Cambridge, 2001. 
2626: 
2627: \bibitem{FT85}  C. Foias and R. Temam, The connection between the Navier-Stokes
2628: equations, and turbulence theory, {\it Directions in Partial
2629: Differential Equations} (Madison, WI, 1985), Publ. Math. Res. Center
2630: Univ. Wisconsin, 55--73.
2631: 
2632: \bibitem{FP}
2633: S. Friedlander and N. Pavlovi\'c, Blowup in a three-dimensional vector
2634: model for the Euler equations, {\it Comm. Pure Appl. Math.} {\bf 57} (2004),
2635: 705--725.
2636: 
2637: \bibitem{H} J. K. Hale, {\it Asymptotic behavior of dissipative systems}, 
2638: Amer. Math. Soc., Providence, RI, 1988.
2639: 
2640: \bibitem{HLS}J. K. Hale,  J. P. LaSalle, and M. Slemrod,
2641: Theory of a general class of dissipative processes, {\it J. Math. Anal. Appl.}
2642: {\bf 39} (1972), 177--191.
2643: 
2644: 
2645: \bibitem{K} I. Kukavica, Role of the pressure for validity of the energy equality for solutions of
2646: the NavierÐStokes equation,
2647: {\it J. Dynam. Differential Equations} {\bf 18} (2006), 461--482.
2648: 
2649: \bibitem{KP}
2650: N. H. Katz and N. Pavlovi\'c, Finite time blow-up for a dyadic model of the Euler
2651: equations, {\it Trans. Amer. Math. Soc.} {\bf 357} (2005), 695--708.
2652: 
2653: \bibitem{L} O. A. Ladyzhenskaya, On the dynamical system generated by the Navier-Stokes
2654: equations, {\it J. Soviet Math.}  {\bf 3} (1975), 458--479.
2655: 
2656: \bibitem{L1} O. A. Ladyzhenskaya, {\it Attractors for Semigroups and Evolution
2657: Equations}, Cambridge Univ. Press, Cambridge, 1991.
2658: 
2659: \bibitem{MV} V. S. Melnik and J. Valero,  On attractors of multivalued semi-flows
2660: and differential inclusions, {\it Set-Valued Anal.} {\bf 6} (1998), 83--111.
2661: 
2662: \bibitem{RS} G. Raugel and G. R. Sell, 
2663: Navier-Stokes equations on thin 3D domains. I. Global attractors and global regularity of solutions.
2664: {\it J. Amer. Math. Soc.} {\bf 6} (1993), 503--568
2665: 
2666: \bibitem{R} R. M. S. Rosa,  Asymptotic regularity condition for the strong
2667: convergence towards weak limit sets and weak attractors of the 3D
2668: Navier-Stokes equations, {\it J. Diff. Eq.}, to appear.
2669: 
2670: \bibitem{S} G. R. Sell, Global attractors for the three-dimensional Navier-Stokes
2671: equations, {\it J. Dynam. Differential Equations} {\bf 8} (1996), 1--33. 
2672: 
2673: \bibitem{SY} G. R. Sell and Y. You, {\it Dynamics of Evolutionary Equations},
2674: Applied Mathematical Sciences {\bf 143}, Springer-Verlag, New York, 2002.
2675: 
2676: \bibitem{T2} R. Temam, {\it Infinite Dimensional Dynamical Systems in Mechanics and
2677: Physics}, Applied Mathematical Sciences {\bf 68}, (2nd Edition, 1997) Springer Verlag, New York, 1988.
2678: 
2679: \bibitem{TZ} R. Temam and M. Ziane, 
2680: Navier-Stokes equations in three-dimensional thin domains with various boundary conditions.
2681: {\it Adv. Differential Equations} {\bf 1} (1996), 499--546.
2682: \end{thebibliography}
2683: 
2684: 
2685: \end{document}
2686: 
2687: 
2688: