0806.0669/RV2.txt
1: \documentclass[leqno,oneside]{article}
2: \usepackage{amsmath, amssymb}
3: \usepackage{amsfonts}
4: \newcommand{\field}[1]{\mathbb{#1}}
5: \newcommand{\bn}{$\boldmath$ \nabla $\unboldmath$}
6: \newcommand{\g}{\tilde{g}}
7: \newcommand{\Ti}{\tilde{T}}
8: \newcommand{\Kt}{\tilde{\hat{K}}}
9: \newcommand{\K}{\tilde{K}}
10: \newcommand{\N}{\tilde{N}}
11: \newcommand{\Rt}{\tilde{\hat{Ric}}}
12: \newcommand{\n}{\noindent}
13: \newcommand{\be}{\begin{equation}}
14: \newcommand{\ee}{\end{equation}}
15: \newcommand{\ben}{\begin{displaymath}}
16: \newcommand{\een}{\end{displaymath}}
17: \newcommand{\theone}{1\ }
18: \newcommand{\thetwo}{2\ }
19: \newcommand{\ep}{\hspace{\stretch{1}}$\Box$}
20: \usepackage{epsfig}
21: \usepackage{latexsym}
22: \usepackage{mathrsfs}
23: \newtheorem{R}{Remark}
24: \newtheorem{Def}{Definition}
25: \newtheorem{Prop}{Proposition}
26: \newtheorem{T}{Theorem}
27: \newtheorem{Con}{Conjecture}
28: \newtheorem{Lem}{Lemma}
29: \newtheorem{Cor}{\bf Corollary}
30: %\newtheorem*{Theorem1}{Theorem 1}
31: %\newtheorem*{Theorem2}{Theorem 2}
32: \addtolength{\hoffset}{-1.3cm}
33: \addtolength{\textwidth}{.5cm}
34: \addtolength{\textheight}{2.0cm}
35: \addtolength{\oddsidemargin}{1.5cm}
36: \addtolength{\voffset}{-1cm}
37: \linespread{1.2}
38: \begin{document}
39: 
40: \begin{center}
41: {\Large\bf On the asymptotic spectrum of the reduced volume in cosmological solutions of the Einstein equations}
42: 
43: \vspace{0.4cm}
44: 
45: {\large Martin Reiris}\footnote{e-mail: reiris@math.mit.edu.}\\
46: 
47: \vspace{0.1cm}
48: 
49: \textsc{Math. Dep. Massachusetts Institute of Technology}\\
50: 
51: \end{center}
52: 
53: \vspace{0.3cm}
54: \begin{center}
55: \begin{minipage}[c]{11cm}
56: \linespread{1.1}%
57: \selectfont
58: {\small Say $\Sigma$ is a compact three-manifold with non-positive Yamabe invariant. We prove that in any long time 
59: constant mean curvature Einstein flow over $\Sigma$, having bounded $C^{\alpha}$ space-time curvature at the cosmological scale, the reduced volume 
60: ${\mathcal{V}}=(\frac{-k}{3})^{3}Vol_{g(k)}(\Sigma)$ 
61: ($g(k)$ is the evolving spatial three-metric and $k$ the mean curvature) decays monotonically towards the volume 
62: value of the 
63: geometrization in which the cosmologically normalized flow decays. In more basic terms, under the given assumptions, 
64: there is volume collapse in the regions where the injectivity radius collapses (i.e. 
65: tends to zero) in the long time. We conjecture that under the curvature assumption above 
66: the Thurston geometrization is the unique global attractor.  We validate it in some special cases.}
67:  
68: \end{minipage}
69: \end{center}
70: 
71: \vspace{0.4cm}
72: \begin{center}
73: \section{Introduction}
74: \end{center}
75: \n A long-standing problem in General Relativity is to understand the long time evolution of cosmological 
76: solutions (solutions with compact space-like sections) of the Einstein equations at the cosmological scale or, in 
77: other words, to understand the large-scale shape of general cosmological solutions. Put in full mathematical 
78: generality the problem is outstandingly difficult and at present out of reach\footnote{The area of
79: cosmology, for which understand the large-scale shape is of central interest, overcame 
80: the mathematical difficulty by assuming large-scale homogeneity and isotropy and that only the averaged properties 
81: of matter contribute to the dynamic at the large-scales. The assumption reduces the mathematical complexity to the 
82: study of three well known models: the ${\mathcal{K}}=-1,0,1$ 
83: Friedman-Lema\^itre cosmologies. It is worth to remark that the justification of such assumption has now become a 
84: problem itself, the so called {\it averaging problem in cosmology}\cite{B2} which is the center of a large debate 
85: these days. The ${\mathcal{K}}=0,-1$
86: Friedman-Lema\^itre models have non-compact spatial sections, isometric to the flat $\field{R}^{3}$ for the ${\mathcal{K}}=0$ 
87: FL-model or the 
88: hyperbolic three-space
89: (with dynamical sectional curvature) still diffeomorphic to $\field{R}^{3}$ for the ${\mathcal{K}}=-1$ case. 
90: Both models however can be compactified
91: to obtain cosmological solutions (with compact slices). Perturbations 
92: around those models have also been largely studied by cosmologists.}. In this article we will present some progress 
93: in this problem for solutions satisfying suitable assumptions. More in particular we will 
94: investigate cosmological solutions of the Einstein constant mean curvature (CMC) flow equations over three-manifolds $\Sigma$ with 
95: non-positive Yamabe invariant (see later) and
96: having a uniform (in time) bound in the $C^{\alpha}$ space-time curvature (see later) at the cosmological scale
97: \footnote{The curvature assumption explicitly prohibits the formation of
98: singularities. In this sense the present work is about the evolution of cosmological solutions which
99: do not develop singularities at the cosmological scale. From a topological perspective, it deals with solutions whose
100: large-scale shape is driven by the topology of the three-dimensional space-like sections.}. 
101: 
102: Since stating the results with precision needs some technical elaboration, we will start by giving below a first glance of 
103: the ideas but in an informal manner. That may give a first flavor of the contents. After that we will comment on related developments 
104: and immediately thereafter we shall be introducing some primary terminology (as not all of it is standard in the field) and 
105: use it to give a detailed description of the contents in the rest of the article.
106: 
107: Consider a cosmological solution of the Einstein equations admitting a Cauchy hypersurface $\sim \Sigma$ of 
108: constant mean curvature (CMC) different from zero. Assume the Yamabe invariant\footnote{The Yamabe invariant of a 
109: compact three-manifold
110: is defined as the supremum of the scalar curvatures of unit volume Yamabe metric. Yamabe metrics are metrics
111: minimizing the Yamabe functional $(\int_{\Sigma}R_{g}dv_{g}/V_{g}^{\frac{1}{3}})$ over a fixed conformal class $[g]$. 
112: The Yamabe invariant is also known as {\it sigma constant} (see for instance \cite{FM2}).} $Y(\Sigma)$ of $\Sigma$ is non-positive. 
113: We will look at the flow $(g,K)$ along the (unique) 
114: CMC foliation where $g$ is the three-metric inherited from the space-time metric ${\bf g}$ and $K$ the second fundamental 
115: form at every CMC slice. The main object of study will be the {\it cosmologically normalized (CMC) flow}, namely the
116: flow $(\tilde{g},\tilde{K})=((\frac{k}{3})^{2}g,\frac{-k}{3}K)$ (see later). In elementary terms the main result will 
117: be to show that if the space-time curvature at the cosmological scale has uniformly (in time) bounded $C^{\alpha}$ norm 
118: with respect to every slice in the CMC foliation then as the mean curvature $k$ tends to zero\footnote{Note that when $k\rightarrow 0$ the cosmological time $t=-1/k$ diverges. We will use
119: the terminology ``{\it in the long time}'' to mean ``when $k\rightarrow 0$".}
120:  the flow $(\tilde{g},\tilde{K})$ separates the manifold $\Sigma$
121: persistently into a (possibly empty) $H$ (-hyperbolic) sector and a (possibly empty) $G$ (-graph) sector 
122: with particular properties that we describe next. The $H$ sector consists of a 
123: finite
124: set of manifolds admitting a complete hyperbolic metric $g_{H}$ of finite volume. Over each one of the $H$ pieces
125: the flow $(\tilde{g},\tilde{K})$ converges to $(g_{H},-g_{H})$ in the long time. The $G$ sector is instead a graph 
126: manifold\footnote{A graph manifold is a manifold obtained as a sum
127: along two-tori of $U(1)$ bundles over two-surfaces.} and over it
128: the injectivity radius collapses (i.e. tends to zero) at every point and in the long time. Moreover the 
129: volume of the $G$ sector
130: relative to the metric $\tilde{g}$ collapses to zero. This shows that in the long time, the volume of 
131: $\Sigma$ relative to the metric $\tilde{g}$, converges to the sum of the volumes of the hyperbolic pieces in the $H$ 
132: sector. The separation into the $H$ and $G$ sectors is called a {\it geometrization}. As we will explain later the results 
133: presented above
134: point towards a much deeper picture of the long time evolution of CMC solutions at the cosmological scale and under curvature 
135: bounds, namely
136: that the {\it Thurston geometrization} (see later) is the only global attractor.     
137: 
138: This article has its roots in the
139: works \cite{R}, \cite{FM}, \cite{A1}, \cite{FM2}. In \cite{FM2} Fischer and Moncrief studied for the first time 
140: the notion of volume collapse at the 
141: cosmological scale and its relations with the Yamabe invariant\footnote{The volume at the cosmological scale is the volume of 
142: $\Sigma$ relative to the metric $\tilde{g}$, namely ${\mathcal{V}}=(\frac{-k}{3})^{3}Vol_{g}$. We will call it either the {\it 
143: volume at the cosmological scale} or the {\it reduced volume} (see later). Note that Fischer and Moncrief use different 
144: terminology. They call {\it sigma constant} to what we call {\it Yamabe invariant} and 
145: {\it reduced Hamiltonian} to what we call {\it reduced volume}. We won't be following it here.}. In particular they 
146: investigated the reduced volume 
147: on a list of natural examples showing at least on those cases a connection between the asymptotic value of the reduced volume and the topology of the
148: Cauchy hypersurfaces. Their analysis validates the results of this article. A related 
149: investigation was carried out by Anderson in the seminal work \cite{A1}, where it is proved (also using the CMC gauge) that 
150: under 
151: pointwise curvature bounds (see the article for a precise statement) there is a sequence of CMC slices with $k\rightarrow 0$ 
152: on which the Einstein flow (suitable scaled) geometrizes the three-manifold. Similar
153: results but exploiting the reduced volume were obtained in \cite{R}. Finally, the notion of cosmological normalized flow 
154: that we use here was elaborated in \cite{R1} following \cite{AM}.
155: 
156: \vspace{2mm}
157: \n \begin{R} {\rm In the context of flows on manifolds with non positive Yamabe invariant, there are strong relations between the Einstein and the Ricci flow. In \cite{Ham} Hamilton
158: has been able to prove that under curvature bounds the Ricci flow geometrizes the manifold in much the same way as it has been proved here the Einstein flow does. 
159: He proves however that the tori separating the $H$ and $G$ pieces are incompressible and therefore the long time geometrization is the Thurston geometrization. It 
160: may be interesting to apply the results on volume collapse carried out in this paper to the Ricci flow under curvature bounds.}
161: \end{R}
162: 
163: We give next a more detailed description of the contents. In technical terms we will be dealing with space-times $(\bf{M},\bf{g})$ where $\bf{M}$ is a four-dimensional manifold
164: and $\bf{g}$ a $C^{\infty}$\footnote{We will assume all through that the Lorentzian space-time metric is of class $C^{\infty}$.} Lorentzian $(3,1)$ metric satisfying the Einstein equations in vacuum ${\bf Ric}=0$. 
165: Assume that there is a space-like slice of non-zero constant mean 
166: curvature (k) diffeomorphic to a three-dimensional manifold $\Sigma$. As is well known (see \cite{Re} and references therein) 
167: there is a unique region 
168: $\Omega_{CMC}$ inside 
169: ${\bf M}$ and diffeomorphic to $\field{R}\times \Sigma$ where the mean curvature $k$ (which serves as a coordinate for 
170: the first factor) varies monotonically. Assume that $\Sigma$ is of non-positive Yamabe invariant $Y(\Sigma)$ (if 
171: $Y(\Sigma)>0$ it is conjectured \cite{Re}
172: that the flow becomes extinct in finite proper time in any of the two time-directions from any CMC slice and therefore the flow
173: would not be a long time flow). In this situation it is easy to check from the energy constraint that
174: $k$ never becomes zero. The existence of a CMC slice of non zero mean curvature defines two different 
175: time directions in $\Omega_{CMC}$: the direction in which the CMC slices increase volume that we will call ``the future" and the
176: direction in which they decrease volume that we will call ``the past"\footnote{By the Hawking singularity theorem all past directed time-like geodesics starting at a common CMC slice terminate before a uniform time lapse.}. We are interested in the dynamics in the future
177: direction. The CMC foliation induces a 3+1 splitting which allows us to write the metric $\bf{g}$ as
178: \begin{equation}\label{first}
179: {\bf g}=-(N^{2}-|X|^{2})dk^{2}+X^{*}\otimes dk+dk\otimes X^{*}+g,
180: \end{equation}
181: 
182: \n where $N$ is the {\it lapse function}, $X$ the {\it shift vector} and $g$ is a three-Riemannian metric on $\Sigma$ (depending on $k$). 
183: Thus the space-time metric $\bf{g}$ is described by a flow $(N,X,g)(k)$ that we will call the {\it Einstein (CMC) flow}. 
184: Let $T$ be the normal vector field to the CMC foliation and pointing in the future direction. The second fundamental form
185: $K$ of the CMC slices is $K=-\frac{1}{2}{\mathcal{L}}_{T}g$\footnote{${\mathcal{L}}_{X}$ denotes the Lie derivative along the vector field $X$}  and therefore $k=tr_{g}K$. The Einstein equations ${\bf Ric}=0$ 
186: in the CMC 3+1 splitting are  
187: \begin{equation}\label{constraints1}
188: R=|K|^{2}-k^{2},
189: \end{equation}
190: \begin{equation}\label{constraints2}
191: \nabla .K=0,
192: \end{equation}
193: \begin{equation}\label{h-j1}
194: \dot{g}=-2NK+{\mathcal{L}}_{X}g,
195: \end{equation}
196: \begin{equation}\label{h-j2}
197: \dot{K}=-\nabla\nabla N+N(Ric+kK-2K\circ K)+{\mathcal{L}}_{X}K,
198: \end{equation}
199: \begin{equation}\label{lapse}
200: -\Delta N+|K|^{2}N=1.
201: \end{equation}
202: 
203: \noindent Equations (\ref{constraints1}),(\ref{constraints2}) are the {\it constraint equations}, equations (\ref{h-j1}),(\ref{h-j2}) are the 
204: {\it Hamilton-Jacobi} 
205: equations of motion and (\ref{lapse}) is the (fundamental) 
206: {\it lapse equation} which is obtained after contraction of (\ref{h-j2}). Thus $N$ gets uniquely determined from
207: $(g,K)$ after solving (\ref{lapse}). Different choices of the shift vector give different flows $(X,N,g)$ over $\Sigma$ but 
208: the space-time solutions ${\bf g}$ they represent via equation (\ref{first}) are isometric. Thus up to space-time
209: diffeomorphism the Einstein flow is uniquely determined from the (abstract) flow $(g,K)(k)$. We will use the choice of $X=0$ 
210: all through the article. 
211: 
212: In cosmological terms the mean 
213: curvature $k$ is a measure of the universe expansion and can be identified \cite{R1} with $-3{\mathcal{H}}$ where ${\mathcal{H}}$ is the Hubble parameter 
214: (constant over each slice of the CMC foliation). At a slice $\{k_{0}\}\times \Sigma$ the Hubble parameter is 
215: ${\mathcal{H}}_{0}=-\frac{k_{0}}{3}$ and if we scale
216: the space-time metric ${\bf g}$ as $\tilde{\bf g}={\mathcal{H}}_{0}^{2}{\bf g}$ we get a new space-time metric 
217: which is a new solution
218: of the Einstein equations in vacuum with three-metric $\tilde{g}={\mathcal{H}}_{0}^{2}g$ 
219: and second fundamental
220: form $\tilde{K}_{0}={\mathcal{H}}_{0}K$ at the same slice, thus having Hubble parameter equal to one (only in that 
221: slice). If  we perform such scaling at every
222: slice in the CMC foliation we obtain a flow $(\tilde{g},\tilde{K})({\mathcal{H}})=({\mathcal{H}}^{2}g,{\mathcal{H}}K)({\mathcal{H}})$ 
223: which we
224: will call the {\it cosmologically normalized Einstein flow} or {\it the Einstein flow at the cosmological scale}. The 
225: cosmologically normalized flow is the subject of the present article. Cosmologically normalized tensors will be denoted 
226: with a tilde either above or next to them. 
227: For example the space-time Riemannian tensor ${\bf Rm}_{\alpha\beta\gamma}^{\ \ \ \delta}$ is scale invariant, therefore
228: the cosmologically normalized Riemann curvature tensor is itself. The normal unit vector field $T$ scale as $T/{\mathcal{H}}$ 
229: and the combination $E={\bf Rm}_{\alpha\beta\gamma\delta}T^{\alpha}T^{\gamma}$ (the electric component of ${\bf Rm}$) is 
230: scale invariant. We will study the cosmologically normalized flow under the following {\it curvature assumption}.
231: 
232: \vspace{0.2cm}
233: \n {\bf Curvature assumption}\footnote{It is fundamental that we assume pointwise bounds of the curvature, i.e. 
234: bounds in the $C^{\alpha}$ norm of ${\bf Rm}$. $L^{2}$ bounds instead seem too weak to control the geometry in the 
235: thin parts.}: 
236: {\it there is a constant $\Lambda>0$ such that, at any time ${\mathcal{H}}$, the $C^{\alpha}_{\tilde{g}}({\mathcal{H}})$ norm of the cosmological normalized Riemann tensor $\tilde{{\bf Rm}}_{\alpha\beta\gamma}^{\ \ \ \delta}(={\bf Rm}_{\alpha\beta\gamma}^{\ \ \ \delta})$
237:  is bounded above by $\Lambda$}. 
238: \n \begin{R} ({\it on the $C^{\alpha}$ norm of the Riemann tensor}). {\rm Given a slice $\{{\mathcal{H}}\}\times \Sigma$ we decompose
239: the space-time Riemann tensor ${\bf Rm}$ into its electric $E_{\alpha\gamma}={\bf Rm}_{\alpha\beta\gamma\delta}T^{\beta}T^{\delta}$
240: and magnetic component $B_{\alpha\gamma}={\bf Rm}^{*}_{\alpha\beta\gamma\delta}T^{\beta}T^{\delta}$, where $^{*}$ means Hodge dual (see \cite{CK}). Now $E$ and $B$ are
241: two $(2,0)$, T-null tensors, which are symmetric and traceless. The $C^{\alpha}_{\tilde{g}}$ norm of ${\bf Rm}$ in the slice 
242: $\{{\mathcal{H}}\}\times \Sigma$ 
243: is defined as the 
244: $C^{\alpha}_{\g}$ norms of $E$ and $B$ as tensors in the Riemannian manifold $(\Sigma,\tilde{g}({\mathcal{H}}))$ 
245: (see the background section for a definition of the $C^{\alpha}_{\g}$ norm of a tensor). These $C^{\alpha}_{\g}$ 
246: norms are assumed to be uniformly bounded by $\Lambda$ for all ${\mathcal{H}}$ along the evolution.}  
247: \end{R}
248: 
249: \vspace{-0.3cm}
250: \n \begin{R} {\rm There is an example due to Ringstr\"om\footnote{The investigation is in connection with 
251: the a priori curvature condition given in \cite{A1}. It is easy to show that the example doesn't satisfy
252: our curvature assumption continuously in time.} \cite{Ring} (Prop 2) of a homogeneous Bianchi VIII model, 
253: showing that  while there are no singularities being formed the curvature assumption above is only satisfied over a 
254: divergent sequence of times, but not for all. The existence of such
255: a sequence is enough to apply many of the results of this article and to conclude in particular 
256: volume collapse.}
257: \end{R}  
258: 
259: The first main result will be the following.  
260: 
261: \begin{T}\label{theo1} Say $Y(\Sigma)\leq 0$ and say $(\tilde{g},\tilde{K})$ is a cosmologically normalized Einstein flow (in vacuum) satisfying the
262: curvature assumption. Then the range of ${\mathcal{H}}$ is of the form $(0,a)$ (it is a long time flow) and as 
263: ${\mathcal{H}}\rightarrow 0$ (i.e. in the long time) the flow (weakly or strongly) persistently geometrizes the manifold $\Sigma$.
264: \end{T}
265: 
266: \n Let us explain what a weak or strong geometrization is. Recall first the {\it thick-thin decomposition} of a Riemannian manifold\footnote{The Thick-thin decomposition
267: is a well know and standard separation of a Riemannian manifold.}.
268: Denote by $\Sigma^{\epsilon}$ the set of points in $(\Sigma,\tilde{g})$ where the injectivity radius is bounded below
269: by $\epsilon$ and $\Sigma_{\epsilon}$ the set of points where the injectivity radius is bounded above by $\epsilon$.
270: $\Sigma^{\epsilon}$ and $\Sigma_{\epsilon}$ are called the $\epsilon$-thick and $\epsilon$-thin parts of $(\Sigma,\tilde{g})$
271: and such decomposition is called the $\epsilon$-{\it thick-thin decomposition}. 
272: A flow $(\tilde{g},\tilde{K})$ in $\Sigma$ geometries $\Sigma$ iff there is (a continuous) $\epsilon({\mathcal{H}})$ with $\epsilon({\mathcal{H}})\rightarrow 0$
273: as ${\mathcal{H}}\rightarrow 0$ such that after a sufficiently
274: long time (i.e. after ${\mathcal{H}}$ gets sufficiently small) $\Sigma_{\epsilon}$ is persistently diffeomorphic
275: to a graph manifold to be denoted by $G$ and $\Sigma^{\epsilon}$ is persistently diffeomorphic to a finite set of
276: manifolds ($H_{i}$), to be denoted as $H$, admitting a complete hyperbolic metric of finite volume ($\g_{H,i}$) and with
277: $(\Sigma^{\epsilon},(\tilde{g},\tilde{K}))$ converging to $\cup_{i=1}^{i=n}(H_{i},(\g_{H,i},-\g_{H,i}))$ in $C^{2,\beta}\times C^{1,\beta}$ (see the background section for a precise description of the convergence). The manifolds separating the $G$ and $H$ sectors are
278: two tori. If all the tori are incompressible (their fundamental groups inject into the fundamental group of $\Sigma$) 
279: the geometrization is said to be {\it strong} and well known to be unique (see for instance \cite{R} Theorem 9), (actually equivalent to the Thurston decomposition
280: of the manifold). If one of the tori is not incompressible, the geometrization is said to be {\it weak}. A schematic picture 
281: of a geometrization is given in Figure \ref{fig1}. Let us exemplify the geometrization 
282: phenomenon with some simple but illustrative cases. 
283: 
284: \begin{enumerate}
285: \item $Y(\Sigma)<0$. The {\it flat cone or Robertson-Walker ${\mathcal{K}}=-1$} solution is ${\bf g}=-dt^{2}+t^{2}g_{H}$ where
286: $g_{H}$ is a hyperbolic metric on a hyperbolic manifold $\Sigma_{H}$. The mean
287: curvature is $k=\frac{-3}{t}$ and the normalized flow converges (it is actually
288: steady) to
289: $(g_{H},-g_{H})$ on the three dimensional manifold $\Sigma_{H}$. The solution
290: is flat.
291: 
292: \item $Y(\Sigma)=0$. 
293: 
294: \begin{enumerate} \item Consider now the solution ${\bf g}=-dt^{2}+\frac{t^{2}}{4}\sigma
295: +d\theta^{2}$ on $\Sigma=S_{gen}\times U(1)$, where $S_{gen}$ is a compact
296: surface of genus $gen>1$, $\sigma$ is a metric of constant scalar curvature equal to $-1$ 
297: on $S_{gen}$ and $d\theta^{2}$ is the standard element of length on $U(1)$. The mean curvature
298: is $k=\frac{-2}{t}$ and the normalized flow collapses to
299: a state $(\frac{\sigma}{9},-\frac{2\sigma}{3})$ on the two dimensional manifold
300: $S_{gen}$. The solution is flat.
301: 
302: \item The {\it Kasner} $(1,0,0)$ (with unit coefficients) is defined as ${\bf
303: g}=-dt^{2}+t^{2}d\theta_{1}^{2}+d\theta_{2}^{2}+d\theta_{3}^{2}$ on
304: $\Sigma=T^{3}$. The mean curvature $k=\frac{-1}{t}$ and the normalized flow
305: collapses to a state $(\frac{1}{9}d\theta^{2}_{1}$,
306: $\frac{-1}{3}d\theta_{1}^{2})$ on the one dimensional manifold $U(1)$. The
307: solution is flat.
308: 
309: \item The {\it Kasner} $(\frac{2}{3},\frac{2}{3},\frac{-1}{3})$ with unit coefficients is defined
310: as ${\bf g}=-dt^{2} +
311: t^{\frac{4}{3}}d\theta_{1}^{2}+t^{\frac{4}{3}}d\theta_{2}^{2}+t^{\frac{-2}{3}}d\theta_{3}^{2}$
312: on $\Sigma=T^{3}$. The mean curvature is $k=\frac{-1}{t}$ and the normalized
313: flow collapses with bounded curvature to a point, i.e. to the zero dimensional
314: space.  
315: 
316: \end{enumerate}
317: \end{enumerate}    
318: 
319: \begin{figure}[h]
320: \centering
321: \includegraphics[width=120mm,height=70mm]{ev.JPG}
322: \caption[U]{Large-scale picture of a cosmological solution. Observe that the reduced volume ${\mathcal{V}}$ is represented as
323: decreasing.}\label{fig1}
324: \end{figure}
325: 
326: \n A crucial quantity used in the proof of Theorem \theone is the {\it reduced volume} ${\mathcal{V}}={\mathcal{H}}^{3}V_{g({\mathcal{H}})}$ which is the volume of the 
327: cosmologically normalized metric $\tilde{g}$\footnote{To our knowledge Fischer and Moncrief were the first to consider the reduced volume in the context of long time
328: evolution in the CMC gauge.}. As it turns out \cite{FM} the reduced volume (which is scale invariant) is 
329: either monotonically decreasing or
330: steady in which case the solution is a flat cone. Equally important, the infimum of the reduced volume when it is thought 
331: as a function on CMC states $(g,K)$ (i.e. pairs $(g,K)$ satisfying the constraint equations) is given by ${\mathcal{V}}_{inf}
332: =(\frac{-Y(\Sigma)}{-6})^{\frac{3}{2}}$ \cite{FM}.
333: The natural question is whether it is always the case that ${\mathcal{V}}\downarrow {\mathcal{V}}_{inf}$ at least 
334: under the curvature assumption above.
335: If it does so, it is known (\cite{R} Theorem 9) that the geometrization is strong and (therefore) unique. We will call it  the 
336: {\it Thurston geometrization}. We conjecture that such is always the case for solutions satisfying the curvature assumption.
337: 
338: \begin{Con}\label{con1} Say $Y(\Sigma)\leq 0$ and say $(\tilde{g},\tilde{K})$ is a cosmologically normalized Einstein flow satisfying the 
339: assumption. Then ${\mathcal{V}}\downarrow {\mathcal{V}}_{inf}=(\frac{-Y(\Sigma)}{6})^{\frac{3}{2}}$.
340: \end{Con}
341: 
342: \n Another way to express the conjecture is that the Thurston geometrization is a global attractor for cosmologically normalized 
343: flows satisfying the curvature assumption on manifolds with non-positive Yamabe invariant. If valid, the conjecture implies 
344: that the (scale invariant) Yamabe functional 
345: \ben
346: Y(g)=\frac{\int_{\Sigma}R_{g}dv_{g}}{V_{g}^{\frac{1}{3}}},
347: \een
348: 
349: \n converges to the Yamabe invariant $Y(\Sigma)$ along the flow. This can be sketchily seen as follows. Under the curvature assumption the scalar curvature
350: $R_{\tilde{g}}$ is known to be bounded (above and below, see Prop \ref{P1} later). On the other hand it is known that for manifolds with
351: $Y(\Sigma)\leq 0$ it is \cite{A3} $(\frac{-Y(\Sigma)}{6})^{\frac{3}{2}}=\sum_{i=1}^{i=n} Vol_{\g_{H,i}}(H_{i})$
352: where $H_{i}$ are the hyperbolic pieces in the Thurston decomposition of $\Sigma$. If ${\mathcal{V}}\downarrow {\mathcal{V}}_{inf}$
353: the volume of the $G$ sector collapses to zero and therefore $\frac{\int_{G}R_{\tilde{g}}dv_{\tilde{g}}}{V_{\tilde{g}}^{\frac{1}{3}}}\rightarrow 0$.
354: As $R_{\tilde{g}}\rightarrow -6$ on the $H$ sector we have $Y(\tilde{g})=Y(g)\rightarrow Y(\Sigma)$ as desired.
355: 
356: The second main result will be to show that always the $G$ sector collapses in (reduced) volume
357: \footnote{We note that this is a non-trivial statement. Consider the two-manifold $[a,b]\times S^{1}$ with the time dependent 
358: metric $g=t^{2}dx^{2}+\frac{1}{t^{2}}d\theta^{2}$. The volume is the same for all $t$ but $inj_{g}\rightarrow 0$.}. 
359: 
360: \begin{T} Say $Y({\Sigma})\leq 0$ and say $(\tilde{g},\tilde{K})$ is a cosmologically normalized flow 
361: satisfying the curvature assumption. Then the reduced volume of the 
362: total space converges towards the volume value of the long time geometrization.
363: \end{T}
364: 
365: \n The {\it volume value} of the geometrization (see also the background section) is 
366: 
367: 
368: \n $\sum_{i=1}^{i=n} V_{g_{H,i}}(H_{i})$. This result is a first step to prove the conjecture above. In fact it 
369: validates the conjecture in some particular cases described in the Corollaries \ref{cor1}-\ref{cor4} which will be proved after the proof 
370: of Theorem \thetwo. 
371: 
372: \begin{Cor}\label{cor1} Say $Y(\Sigma)\leq 0$. Given $\Lambda$ there is $\epsilon$ such that for any cosmologically normalized
373: flow satisfying the curvature assumption (with the same $\Lambda$) and having 
374: ${\mathcal{V}}-{\mathcal{V}}_{inf}\leq \epsilon$ at an initial time, it is ${\mathcal{V}}\downarrow {\mathcal{V}}_{inf}$
375: in the long time.
376: \end{Cor}
377: 
378: \n In basic terms, what Corollary \ref{cor1} says is that if we restrict to the set of solutions satisfying the
379: curvature assumption with a fixed $\Lambda$ then the Thurston geometrization is stable (in the class).
380: 
381: \begin{Cor}\label{cor2} Say $Y(\Sigma)\leq 0$ and say $(\tilde{g},\tilde{K})$ is a cosmologically normalized flow satisfying the
382: curvature assumption that is locally collapsing at every point, i.e. there is no $H$ sector in the long time
383: geometrization. Then $Y(\Sigma)=0$ and ${\mathcal{V}}\downarrow {\mathcal{V}}_{inf}=0$.
384: \end{Cor} 
385:        
386: Let ${\mathcal{V}}_{0}$ be the infimum of the volumes of all complete hyperbolic manifolds 
387: (with sectional curvature normalized to one), with or without cusps (this number is known to be positive \cite{Th}). Then we
388: have
389: 
390: \begin{Cor}\label{cor3} Say $Y(\Sigma)\leq 0$ and say $(\tilde{g},\tilde{K})$ is a cosmologically normalized flow satisfying the
391: curvature assumption. If at an initial time it is ${\mathcal{V}}<{\mathcal{V}}_{0}$ then 
392: ${\mathcal{V}}\downarrow {\mathcal{V}}_{inf}=0$ in the long time.
393: \end{Cor}
394:  
395: \n Corollary \ref{cor3} says that if we restrict to the class of solutions satisfying the curvature assumption (with variable $\Lambda$), 
396: there is a threshold ${\mathcal{V}}_{0}$ for ${\mathcal{V}}$ at the initial time, below which the long time
397: gemetrization has only a $G$ sector and the reduced volume collapses to zero in the long time.
398: 
399: \begin{Cor}\label{cor4} Say $Y(\Sigma)\leq 0$ and say $(\tilde{g},\tilde{K})$ is a cosmologically normalized flow satisfying the
400: curvature assumption above. Then ${\mathcal{V}}\downarrow {\mathcal{V}}_{inf}$ iff the tori separating the
401: $H$ and $G$ sectors in the long time geometrization of the flow are incompressible in $\Sigma$.
402: \end{Cor}
403: 
404: \n Corollary \ref{cor4} shows that to prove Conjecture \ref{con1} is sufficient to prove the topological fact that
405: the two-tori separating the $H$ and $G$ sectors are incompressible.
406: 
407: Finally as an outcome of the proof of Theorems \theone and \thetwo we will be able to prove
408: 
409: \begin{Cor}\label{cor5} Say $Y(\Sigma)\leq 0$ and say $(\g,\K)$ is a cosmologically normalized flow satisfying the curvature 
410: assumption. Then the Bel-Robinson energy $Q_{0}({\bf Rm})$ is an $o({\mathcal{H}})$ (i.e. $\lim_{{\mathcal{H}}\rightarrow 0}Q_{0}/{\mathcal{H}}=0$). 
411: \end{Cor}
412: 
413: \begin{center}
414: \section{Background and terminology}
415: \end{center}
416: 
417: {\center \subsection{Convergence and collapse of Riemannian manifolds}}
418: 
419: \vspace{0.1cm}
420: We will make use the Cheeger-Gromov theory of convergence and collapse of Riemannian three-manifolds under curvature bounds. 
421: In particular we will use extensively the following result (\cite{A1} Prop 4 and 5. See also \cite{Pet})
422: 
423: \begin{Prop}\label{CH-G} Let $(\Sigma_{i},g_{i})$ be a sequence of Riemannian three-manifolds with or without boundary, with
424: uniformly bounded $C^{\alpha}_{g_{i}}$ Ricci curvature, i.e. $\|Ric_{g_{i}}\|_{C^{\alpha}_{g_{i}}}\leq \Lambda$. 
425: We have
426: 
427: \begin{enumerate}
428: \item\label{1} say $\{x_{i}\}$ is a sequence of points such that $dist(x_{i},\partial \Sigma_{i})\rightarrow \infty$ and 
429: $inj_{g_{i}}(x_{i})\geq inj_{0}$. Then one can extract a subsequence $(\Sigma_{i_{j}},x_{i_{j}},g_{i_{j}})$ converging
430: in $C^{2,\beta}$ to $(\Sigma_{\infty},x_{\infty},g_{\infty})$ ($\beta<\alpha$),
431: \item\label{2} suppose non of the $(\Sigma_{i}$ is diffeomorphic to a closed space form (a quotient of $S^{3}$). Then 
432: for any $\epsilon>0$ and $inj_{0}$ there is $r(\epsilon,\Lambda,inj_{0})$ with $r\rightarrow \infty$ as $\epsilon\rightarrow 0$, 
433: such that if $dist(x_{i},\partial \Sigma_{i})\geq r$ there is a finite cover of the ball $B(x_{i},r)$ with 
434: $inj_{g_{i}}(x_{i})\sim inj_{0}$.
435: \end{enumerate}
436: \end{Prop}
437: 
438: \n A number of remarks are in order.
439: 
440: \vspace{0.1cm}
441: i) If $U$ is any tensor field on a Riemannian manifold $(\Sigma,g)$ then
442: the $C^{k,\alpha}_{g}$ norm of $U$ is defined as 
443: \ben
444: \begin{split}
445: \|U\|_{C^{k,\alpha}_{g}}&=sup_{x\in\Sigma}\{|U|_{g}(x)+|\nabla U|_{g}(x)+ \dots +|\nabla^{k} U|_{g}(x)+\\
446: &\quad sup_{y\in \Sigma} \{\frac{|\nabla^{k}U(x)-\nabla^{k} U(y)|}{dist(x,y)^{\alpha}}\}\}.
447: \end{split}
448: \een
449: 
450: \n The difference $\nabla^{k} U(x)-\nabla^{k} U(y)$ is by parallel transport along any shortest geodesic joining $x$ with $y$.
451: 
452: ii) The convergence in item {\it 1} in the Proposition above is in the following sense: there is a sequence of submanifolds $\tilde{\Sigma}_{i}\subset \Sigma_{i}$ with 
453: $x_{i}\in \tilde{\Sigma_{i}}$ and $dist(x_{i},\partial \tilde{\Sigma_{i}})\rightarrow \infty$ 
454: and a sequence of diffeomorphisms (onto the image) $\varphi_{i}:\tilde{\Sigma}_{i}\rightarrow \Sigma_{i}$ such that
455: $\|\varphi_{i*}(g_{i})-g_{\infty}\|_{C^{2,\beta}_{g_{\infty}}}\rightarrow 0$.
456: 
457: c) A sequence of tensors $U_{i}$ in $(\Sigma_{i},x_{i},g_{i})$ converge in $C^{k,\beta}$ to $U_{\infty}$ in $(\Sigma_{\infty},x_{\infty},g_{\infty})$
458: if $\|\varphi_{i*}U_{i}-U_{\infty}\|_{C^{k,\beta}_{g_{\infty}}}\rightarrow 0$. 
459: 
460: d) One practical consequence of Proposition \ref{CH-G} is that when it comes to find interior elliptic estimates of certain 
461: elliptic operators on collapsed regions, we may well assume (because we can unwrap) that at the given point $x$ where one 
462: wants to extract the estimate there is a chart $(z_{1},z_{2},z_{3})$ of harmonic coordinates covering the ball $B(x,inj_{0})$
463: with $g_{ij}$ (the components of $g$ in the chart $\{z\}$) bounded in $C^{2,\alpha}_{\{z\}}$ (now the norms are standard 
464: H\"older norms on
465: the chart $\{z\}$) by a constant $C(inj_{0},\Lambda)$. We will be using this fact repeatedly all through the article.     
466: 
467: {\center \subsection{Electric-magnetic decomposition of the space-time curvature and some related formulae}}
468: 
469: In this section we introduce the electric/magnetic decomposition of the space-time curvature and useful formulae. Non of the
470: properties presented is given with a proof. The reader can consult the references \cite{CK},\cite{AM} for a detailed account 
471: on Weyl fields.  
472: 
473: Let $T$ be the normal unit vector field in the future direction and say ${\bf g}$ is a vacuum solution of the
474: Einstein equations. Then the electric and magnetic fields of ${\bf Rm}$ are defined by
475: \ben
476: E_{ab}={\bf Rm}_{acbd}T^{c}T^{d},\ \ B_{ab}=^{*}{\bf Rm}_{acbd}T^{c}T^{d}.
477: \een
478: 
479: \n The electric and magnetic fields are traceless and $T$-null $(2,0)$ vectors. In terms of $g$ and $K$ they have the
480: expressions
481: \ben
482: E=Ric+kK-K\circ K,\ \ B=-Curl K,
483: \een
484: 
485: \n where $Curl$ is the operator on symmetric $(2,0)$ tensors defined as 
486: $(Curl A)_{ab}=\frac{1}{2}(\epsilon_{a}^{\ cd}\nabla_{d}A_{cb}+\epsilon_{b}^{\ cd}\nabla_{d} A_{ca})$ ($\epsilon_{abc}$ is the volume form).
487: We also have
488: \ben
489: Div E=K\wedge B,\ \ Div B=-K\wedge B,
490: \een
491: 
492: \n where $(Div A)_{a}=\nabla^{b}A_{ba}$ is the divergence and $\wedge$ is the operation $(A\wedge C)_{a}=\epsilon_{a}^{\ bc}A_{b}^{\ d}C_{dc}$.
493: Dynamically (under zero shift) we have
494: \ben
495: \dot{E}=N Curl B-\nabla N\wedge B-\frac{5}{2}N (E\times K)-\frac{2}{3} N <E,K>g-\frac{1}{2}NkE,
496: \een
497: \ben
498: \dot{B}=-N Curl E+\nabla N\wedge E-\frac{5}{2}N (B\times K)-\frac{2}{3}N <B,K>k-\frac{1}{2}NkB.
499: \een
500:  
501: \n the dot meaning derivative with respect to $k$, $<.,.>$ is the inner product and $\times$ is the operation
502: \ben
503: (A\times C)_{ab}=\epsilon_{a}^{\ cd}\epsilon_{b}^{\ ef}A_{ce}C_{df}+\frac{1}{3}(A.C)g_{ab}-\frac{1}{3}(tr A)(tr C)g_{ab}.
504: \een 
505: 
506: The Bel-Robinson tensor is the totally symmetric, traceless, $(4,0)$ tensor $Q_{\alpha\beta\gamma\delta}$, defined
507: as
508: \ben
509: Q_{\alpha\beta\gamma\delta}={\bf Rm}_{\alpha\mu\gamma\nu}{\bf Rm}_{\beta\ \delta}^{\ \mu\ \nu}+{\bf Rm}_{\alpha\mu\gamma\nu}^{*}{\bf Rm}^{*\ \mu\ \nu}_
510: {\beta\ \delta}.
511: \een
512: 
513: \n We have $Q_{\alpha\beta\gamma\delta}({\bf Rm})T^{\alpha}T^{\beta}T^{\gamma}T^{\delta}=|E|^{2}_{g}+|B|^{2}_{g}$ and 
514: $\nabla^{\alpha}Q_{\alpha\beta\gamma\delta}({\bf Rm})=0$. Denote by $Q$ the integral in $\Sigma$ of $Q_{TTTT}$. Taking
515: the divergence of $Q_{\alpha TTT}$ and integrating we get the {\it Gauss equation}
516: \ben\label{GG}
517: \dot{Q}=-3\int_{\Sigma}NQ_{\alpha\beta TT}{\bf \Pi}^{\alpha\beta}dv_{g}.
518: \een
519: 
520: \n where ${\bf \Pi}$ is the {\it deformation tensor} ${\bf \Pi}_{\alpha\beta}=\bn_{\alpha}T_{\beta}$. Restricted to
521: any CMC slice we have ${\bf \Pi}=-K$, ${\bf \Pi}_{Ti}=\frac{1}{N}\nabla_{i}N$, ${\bf \Pi}_{Ti}=0$ and ${\bf \Pi}_{TT}=0$. 
522: Again dot means derivative with respect to $k$. All through the article we will use the formulae above but for the cosmologically 
523: normalized tensors. We will indicate that
524: by using a tilde above the referred tensor or by including a subindex $\g$ next to tensor operation, for instance $\wedge_{\g}$
525: or $| (.)|_{\g}$. The cosmological normalized versions of the equations above is straightforward to get and won't be deduced when needed.
526: 
527: \vspace{0.2cm}
528: {\center \subsection{The Newtonian potential, the reduced volume element and the logarithmic time}}
529: 
530: When working with cosmologically normalized quantities, it is convenient to use the {\it logarithmic time} $\sigma=-\ln -k$ as the time variable. 
531: Derivatives with respect to $\sigma$ of a cosmologically normalized quantity gives rise to a quantity which is also cosmologically normalized. 
532: {\it For the rest of the article derivatives with respect to $\sigma$ will be denoted with a dot.}
533: 
534: To illustrate how to work with cosmologically normalized quantities let us introduce the {\it Newtonian potential} 
535: $\phi=3\N-1=3N{\mathcal{H}}^{2}-1$ and
536: the {\it reduced volume element} $d\nu={\mathcal{H}}^{3}dv_{g}$ (both are scale invariant) and let us deduce the following 
537: pair of equations which will be fundamental
538: \be\label{locredvol}
539: \frac{\ln (d\nu)^{\frac{1}{3}}}{d\sigma}=\phi,
540: \ee
541: \be\label{Poisson}
542: \Delta_{\g}\phi-|\K|^{2}_{\g}\phi=|\Kt|^{2}_{\g}.
543: \ee
544: 
545: \n We have used a hat $\hat{}$ above $\K$ to mean the traceless part of $\K$ (with respect to $\g$)\footnote{We will use the same notation for $Ric$.}. 
546: Under zero shift we have 
547: \ben
548: \frac{d dv_{g}}{d k}=\frac{1}{2}tr_{g}\frac{d g}{d k}dv_{g}=-Nkdv_{g}.
549: \een
550: 
551: \n Now $d/d\sigma=(d/dk)(dk/d\sigma)=-kd/dk$ and so
552: \be
553: \frac{d \ln (d\nu)^{\frac{1}{3}}}{d\sigma}=\frac{d \ln {\mathcal{H}}}{d \sigma}+{\mathcal{H}}(N{\mathcal{H}}/3)=
554: 3N{\mathcal{H}}^{2}-1,
555: \ee 
556: 
557: \n as desired. To get the Poisson-like equation for the Newtonian potential $\phi$ observe that the lapse equation (\ref{lapse}) is
558: scale invariant and therefore 
559: \ben
560: -\Delta_{\g}\N+|\K|^{2}_{\g}\N=1.
561: \een
562: 
563: \n Making $\N=\frac{1}{3}(\phi+1)$ we get
564: \ben
565: -\Delta_{\g}\phi+|\K|^{2}_{\g}\phi=3-|\K|^{2}_{\g}=-|\Kt|^{2}_{\g}.
566: \een
567: 
568: A final remark. The Maximum principle applied to (\ref{Poisson}) gives $-1\leq \phi\leq 0$. This important fact implies by
569: equation (\ref{locredvol}) that the local reduced volume element $d\nu$ is non increasing (under zero shift) and in 
570: particular that the reduced volume ${\mathcal{V}}$ is monotonically decreasing unless is steady in which case the solution 
571: is a flat cone. 
572: 
573: \vspace{0.2cm}
574: {\center\subsection{Geometric states and persistent geometric states}}
575: 
576: \n In this section we introduce some definitions, that although not strictly needed, puts the main concepts 
577: used in a broad geometric context.    
578: 
579: \begin{Def} Given a compact three-manifold $\Sigma$ define the geometric spectrum to be the set of all its
580: partitions (geometric states) of the form $\Sigma=\{H_{1}\ldots, H_{i},G_{1},$
581: 
582: \n $\ldots,G_{j}\}$ where the $H$ pieces 
583: are three manifolds
584: possibly with boundary admitting a complete hyperbolic metric of finite volume, and the $G$ are graph 
585: three-manifolds possibly with boundary. If any, the boundaries in all pieces are two-tori and a torus in the 
586: boundary of a $H$ piece is always a torus in the boundary of a $G$ piece. Two geometric states are said to
587: be equivalent if there is an isotopy in $\Sigma$ carrying the $H$ and the $G$ sectors of one into the $H$
588: and $G$ sectors of the other. A geometric state is said to be pure if there is only a $H$ or a $G$ piece.  
589: \end{Def}
590: 
591: \begin{Def} Given a geometric state $\{H_{1},\ldots,H_{i},G_{1},\ldots,G_{j}\}$, its volume value $V$ is defined
592: as the sum of the volumes of the complete hyperbolic metrics of finite volume of the $H$ pieces.
593: The volumetric spectrum is defined as the set of all the volume values for all the states in the geometric 
594: spectrum. 
595: \end{Def}
596: 
597: \n In this terminology, Theorem \theone says in particular that if $\epsilon$ is chosen sufficiently small
598: then after sufficiently long time the $\epsilon$-thick-thin decomposition is a (persistent) geometric state of the 
599: manifold. We will prove in Theorem \thetwo that ${\mathcal{V}}$ decreases to the 
600: volume value of the geometric state in which the flow is decaying. 
601: 
602: We make precise now the notion of persistence of a geometrization introduced in section $1$. These notions will be used in the
603: proofs of the main results. We say that a long time, cosmologically normalized flow $(\g,\K)$ implements a persistent geometrization iff either 
604: 
605: \begin{enumerate}
606: 
607: \item $inj_{\g(\sigma)}(\Sigma)\rightarrow 0$ as $\sigma$ goes to infinity (in which case there is only one persistent $G$ 
608: piece) or
609: 
610: \item $inj_{\g{\sigma}}(\Sigma)\geq inj_{0}>0$ as $\sigma$ goes to infinity (in which case there is only one 
611: persistent $H$ piece) and there is
612: a continuous function $\varphi:(-\ln -a,\infty)\times H\rightarrow \Sigma$, differentiable in the second factor, 
613: such that $\|\varphi ^{*}\g(\sigma)-\g_{H}\|_{C^{2,\beta}_{\g_{H}}}\rightarrow 0$ as $\sigma$ goes to infinity, or
614: 
615: \item the injectivity radius collapses in some regions and remains bounded below in some others (in which case 
616: there are a set of $G$ pieces $G_{1},\ldots,G_{j}$ and a set of $H$ pieces $H_{1},\ldots,H_{k}$) and for any 
617: $\epsilon>0$ and for
618: any $H$ piece $(H_{i},\g_{Hi})$ there is a continuous function $\varphi_{i}:(-\ln -a,\infty)\times 
619: H^{\epsilon}_{i}\rightarrow \Sigma$, differentiable in the second factor such that $\|\varphi_{i}^{*}\g(\sigma)-\g_{Hi}\|_{C^{2,\beta}_{\g_{Hi}}}
620: \rightarrow 0$ as $\sigma$ goes to infinity.
621: 
622: \end{enumerate}
623: 
624: {\center\subsection{Some useful terminology}}
625:  
626: Any sequence $\{\sigma_{i}\}$ of logarithmic times ($\sigma_{i}=-\ln -k_{i}$) which is diverging i.e. $\lim_{i\rightarrow \infty}\sigma_{i}=\infty$ 
627: will be called a {\it diverging sequence of logarithmic times} and abbreviated DSLT. Given a DSLT, $\{\sigma_{i}\}$, we say 
628: that a sequence of sets 
629: $\Omega(\sigma_{i})$ has {\it asymptotically total reduced volume} (ATRV) if 
630: ${\mathcal{V}}(\Omega(\sigma_{i}))\rightarrow {\mathcal{V}}_{\infty}$
631: as $\sigma_{i}\rightarrow \infty$ where ${\mathcal{V}}_{\infty}$ is the limit of the reduced volume in the long time. 
632: Similarly we can define a set having {\it asymptotically non-zero} (ANZRV) or {\it asymptotically zero reduced volume} (AZRV). 
633: We say that a quantity $f$ controls a quantity $h$ if $|f|<M$ implies $|h|<C(M)$, and $f$ controls $h$ at zero if $M\rightarrow 0$ implies
634: $C(M)\rightarrow 0$.
635: 
636: \vspace{0.2cm}
637: {\center\section{Proof of the main results and corollaries}}
638: 
639: This section is organized as follows. We prove first three propositions (Propositions \ref{P1},\ref{P2},\ref{P3}) that would frame
640: the proofs of Theorem \theone and \thetwo. We prove then Theorems \theone and \thetwo and next the 
641: Corollaries \ref{cor1}-\ref{cor5}. 
642: 
643: Let us give a heuristic behind the proofs of Theorems \theone and \thetwo. The key ingredient is to look at
644: the reduced volume. As it is monotonic, it must settle in some limit value ${\mathcal{V}}_{\infty}$ as ${\mathcal{H}}\rightarrow 0$.
645: Therefore in the regions where the injectivity radius is bounded below (in $(\Sigma,\g(\sigma)))$ it must be $\phi\sim 0$
646: otherwise by equation (\ref{locredvol}) the volume would keep decreasing and eventually become below ${\mathcal{V}}_{\infty}$. 
647: Using equation (\ref{Poisson}) this implies $|\Kt|^{2}_{\g}\sim 0$ which after using the Einstein equations implies
648: $|\tilde{\hat{Ric}}|^{2}_{\g}\sim 0$. In other words the regions where the injectivity radius remains bounded below become 
649: hyperbolic. This argument gives in essence Theorem \theone. Theorem \thetwo is more involved because it deals with the regions
650: where the injectivity radius collapses. We are able to show however that if the $G$
651: regions (where the injectivity radius collapses) carry a non zero reduced volume (call it ${\mathcal{V}}_{0}$), 
652: then the regions inside $G$ whose unwrapped
653: geometry becomes hyperbolic carry asymptotically all the volume ${\mathcal{V}}_{0}$. This fact will imply an 
654: isoperimetric inequality showing that the regions lying at a 
655: distance between 1/2 and 1 from the collapsed regions and whose unwrapped geometry is becoming hyperbolic, carry also 
656: asymptotically a non zero reduced volume (if ${\mathcal{V}}_{0}\neq 0$). As these two regions are disjoint, the limit of the volume of the $G$ regions must be above ${\mathcal{V}}_{0}$
657: which is a contradiction.
658:     
659: \vspace{0.2cm}
660: \begin{Prop}\label{P1} Say $Y(\Sigma)\leq 0$ and say $(\tilde{g},\tilde{K})$ is a cosmologically normalized flow satisfying the
661: curvature assumption. At any logarithmic time $\sigma$ we have the following properties.
662: 
663: \begin{enumerate}
664: 
665: \item \label{i1} $\|\Kt\|_{C^{1,\alpha}_{\g}}$, $\|\Rt\|_{C^{\alpha}_{\g}}$ and $\|\phi\|_{C^{2,\alpha}_{\g}}$
666: are controlled by $\Lambda$.
667: 
668: \item For any $\epsilon>0$ there is $\delta(\epsilon,\Lambda)>0$ such that at any point $p$ if $|\phi(p)|\leq \delta$ then 
669: $|\Kt|_{\g}(p)\leq \epsilon$. In other words $-\phi(p)$ controls $|\Kt|_{\g}(p)$ at zero.
670: 
671: \end{enumerate}
672: \end{Prop}
673: 
674: \n {\bf Proof:} 
675: 
676: \n {\it 1}. As has been proved in \cite{A1} (Prop 2.2), $\|\K\|_{L^{\infty}_{\g}}$ and $\|Ric\|_{L^{\infty}_{\g}}$ are
677: controlled by $\Lambda$. Consider the elliptic system
678: \ben\label{eeo}
679: Div \Kt=0,
680: \een
681: \ben\label{eet}
682: Curl_{\g} \Kt=-B,
683: \een
684: 
685: \n Pick a point $x\in \Sigma$ and unwrap\footnote{So far we have $L^{\infty}_{\g}$ control of $\tilde{Ric}$. Still
686: proposition \ref{CH-G} holds, and one can unwrap to have $inj_{\g}\sim inj_{0}$ but this time the unwrapped geometry is controlled in
687: $C^{1,\beta}$. This is enough however to get elliptic estimates from equations \ref{eeo} and \ref{eet}. This is the only time we will need
688: an extension of proposition \ref{CH-G}.} if necessary to have $inj_{\g}(x)\geq inj_{0}$. Then interior Schauder
689: estimates \cite{Elliptic} show that $\|\Kt\|_{C^{1,\alpha}_{\g}(B(x,inj_{0}/2))}$ is controlled by $\Lambda$. Therefore $\|\Kt\|_{C^{1,\alpha}_{\g}}$ is
690: controlled by $\Lambda$. From $E=\tilde{Ric}-3\K+\K\circ\K$
691: we get that $\|\tilde{Ric}\|_{C^{\alpha}_{\g}}$ is controlled by $\Lambda$. Schauder estimates applied to
692: \ben
693: \Delta_{\g} \phi -|\Kt|^{2}_{\g}\phi=|\K|^{2}_{\g},
694: \een
695: 
696: \n show that $\|\phi\|_{C^{2,\alpha}_{\g}}$ is controlled by $\Lambda$.
697: 
698: \vspace{0.2cm}
699: \n {\it 2}. Suppose there is a sequence of logarithmic times $\{\sigma_{i}\}$ and a sequence of points $\{x_{i}\}$
700: such that $\phi(x_{i},\sigma_{i})\rightarrow 0$ but $|\Kt|_{\g(\sigma_{i})}(x_{i},\sigma_{i})\geq M$ with $M>0$. Unwrapping if 
701: necessary to have $inj_{\g}(x_{i})\geq inj_{0}$ we can extract a subsequence of $\{\sigma_{i}\}$ such that
702: on the balls $B(x_{i},inj_{0}/2)$, $\g(\sigma_{i})$ converges in $C^{2,\beta}$ to a limit metric $\g_{\infty}$, $\phi$ 
703: converges in $C^{2,\beta}$ to a limit
704: $\phi_{\infty}\leq 0$ with $\phi_{\infty}(x_{\infty})=0$ and $\Kt$ converges in $C^{1,\beta}$ to a limit $\Kt_{\infty}$ with $|\Kt|_{\g_{\infty}}
705: (x_{\infty})\geq M$, all satisfying the equation
706: \ben
707: \Delta_{\g_{\infty}}\phi_{\infty}-|\K_{\infty}|^{2}_{\g_{\infty}}\phi_{\infty}=|\Kt|^{2}_{\g_{\infty}}.
708: \een
709: 
710: \n However at $x_{\infty}$ it is $0\geq (\Delta_{\g_{\infty}}\phi_{\infty})(x_{\infty})=|\Kt|^{2}_{\g_{\infty}}(x_{\infty})>M>0$ which is
711: absurd. \hspace{\stretch{1}}$\Box$
712: 
713: \begin{Prop}\label{P2} Say $\Sigma$ is a compact three-manifold with 
714: bounded $C^{\alpha}_{g}$ norm of the curvature and bounded volume, i.e. $\|Ric\|_{C^{\alpha}_{g}}+Vol_{g}(\Sigma)\leq \Lambda$, 
715: and say $r<r'$. Then there is $C(\Lambda,r,r')$ such that for any measurable subset $\Omega$ it is $Vol_{g}(B(\Omega,r))\geq C(\Lambda,r,r') Vol_{g}(B(\Omega,r'))$
716: where $B(\Omega,s)$ is the ball of $\Omega$ with radius $s$. 
717: \end{Prop}
718: 
719: \n {\bf Proof:} 
720: 
721: Let ${\mathcal{K}}(\Lambda)<0$ be a lower bound for the sectional curvatures of any Riemannian three-manifold with 
722: $\|Ric\|_{L^{\infty}_{g}}\leq C(\Lambda)$.
723: Let $\{x_{i},i=1,\ldots,m\}\subset \Sigma$ be any set of $m$ points. By the Bishop-Gromov
724: volume comparison the function 
725: 
726: \n $Vol_{g}(\cup_{i=1}^{i=m}B(x_{i},r))/Vol_{g_{{\mathcal{K}}}}(o,r)$ is monotonically decreasing as $r$ increases, where 
727: $g_{\mathcal{K}}$ is a metric of constant sectional curvature ${\mathcal{K}}$ in $\field{R}^{3}$. We have therefore
728: \begin{equation}
729: Vol_{g}(\cup_{i=1}^{i=m}B(x_{i},r'))\leq C(\Lambda,r,r') Vol_{g}(\cup_{i=1}^{i=m}B(x_{i},r)),
730: \end{equation}
731: 
732: \n for any $r<\bar{r}$. Now consider a measurable set $\Omega$. There is $\{x_{i},i=1,\ldots,\infty\}\subset \Omega$ such that
733: \ben
734: \cup_{i=1}^{i=m}B(x_{i},r)\uparrow_{m} B(\Omega,r),
735: \een
736: 
737: \n and
738: \ben
739: \cup_{i=1}^{i=m}B(x_{i},r')\uparrow B(\Omega,r').
740: \een
741: 
742: \n Then taking volumes we have
743: \ben
744: \begin{split}
745: Vol_{g}(B(\Omega,r'))&=\lim_{m\rightarrow \infty} Vol_{g}(\cup_{i=1}^{i=m}B(x_{i},r'))\leq C(\Lambda,r,r')\lim_{m\rightarrow \infty} 
746: Vol_{g}(\cup_{i=1}^{i=m}B(x_{i},r))\\
747: &= C(\Lambda,r,r')Vol_{g}(B(\Omega,r)),
748: \end{split}
749: \een
750: 
751: \n which finishes the proof. \hspace{\stretch{1}}$\Box$\\
752: 
753: \begin{Prop}\label{P3} Say $Y(\Sigma)\leq 0$ and say $(\g,\K)$ is a cosmologically normalized flow satisfying the curvature assumption. We have the following
754: properties.
755: 
756: \begin{enumerate}
757: \item Given $\Gamma\geq 0$ and any DSLT, $\{\sigma_{i}\}$, the sequence of sets $\Omega_{\phi,\Gamma}(\sigma_{i})=\{
758: x\in \Sigma/ -\phi(x,\sigma_{i})\geq \Gamma\}$ has AZRV.
759: 
760: \item Given $\Gamma\geq 0$ and any DSLT, $\{\sigma_{i}\}$, the sequence of sets $\Omega_{\Kt,\Gamma}(\sigma_{i})=\{x\in\Sigma/|\Kt(x,\sigma_{i})|_{\g(\sigma_{i})}\geq \Gamma\}$
761: has AZRV.
762: 
763: \item Given $\Gamma\geq 0$ and any DSLT, $\{\sigma_{i}\}$, the sequence of sets $\Omega_{\nabla \Kt,\Gamma}=\{x\in \Sigma/|\nabla \Kt(x,\sigma_{i})|_{\g(\sigma_{i})}\geq \Gamma\}$
764: has AZRV.
765: 
766: \item For any pair of DSLT, $\{\sigma_{i}\}$ and $\{\sigma'_{i}\}$ with $\delta'\geq \sigma_{i}-\sigma'_{i}\geq \delta$ ($\delta'>\delta>0$ and fixed)
767: we have
768: \ben
769: \int_{\sigma'_{i}}^{\sigma_{i}}\|E\|^{2}_{L^{2}_{\g(\sigma)}}d\sigma \rightarrow 0.
770: \een
771: 
772: \item For any DSLT $\{\sigma_{i}\}$ we have
773: \ben
774: \tilde{Q}_{0}(\sigma_{i})=(\|E\|^{2}_{L^{2}_{\g}}+\|B\|^{2}_{L^{2}_{\g}})(\sigma_{i})\rightarrow 0.
775: \een
776: 
777: \n and therefore the sets 
778: \ben
779: \Omega_{B,\Gamma}(\sigma_{i})=\{x\in\Sigma/|B(x,\sigma_{i})|_{\g(\sigma_{i})}\geq \Gamma\},
780: \een
781: \ben
782: \Omega_{\tilde{\hat{Ric}},\Gamma}(\sigma_{i})=\{x\in \Sigma/|\tilde{\hat{Ric}}(x_{i},\sigma_{i})|_{\g(\sigma_{i})}\geq\Gamma\},
783: \een
784: 
785: \n have AZRV.
786: 
787: \end{enumerate}
788: \end{Prop} 
789: 
790: \n {\bf Proof:} 
791: 
792: \vspace{0.2cm}
793: \n {\it 1}. Differentiating 
794: \ben
795: \Delta_{\g}\phi-|\K|^{2}_{\g}\phi=|\Kt|^{2}_{\g},
796: \een
797: 
798: \n with respect to logarithmic time we get
799: \begin{equation}\label{Dder}
800: \Delta_{\g}\dot{\phi}-|\K|^{2}_{\g}\dot{\phi}=-({\Delta}_{\g})\dot{}\phi+(|\K|^{2}_{\g})\dot{}\phi+(|\Kt|^{2}_{\g})\dot{}.
801: \end{equation}
802: 
803: \n Appealing to
804: \ben
805: \dot{\g}=2\phi \g-6\tilde{N}\Kt,
806: \een
807: \ben
808: \dot{\Kt}=-\Kt-\phi\g-\nabla^{2}\phi+\phi E+E-\N(\Kt\circ\Kt-2\Kt),
809: \een
810: \ben
811: (\Delta_{\g})\dot{}(\phi)=<\nabla^{2}\phi,\dot{\g}>_{\g}-<\nabla\phi,Div \dot{\g}+\frac{1}{2}dtr_{\g}\dot{\g}>_{\g},
812: \een
813: 
814: \n we get by Proposition \ref{P1} {\it 1} that the right hand side of equation (\ref{Dder}) has $C^{\alpha}$ norm
815: controlled by $\Lambda$. By the maximum principle on equation (\ref{Dder}) $\|\dot{\phi}\|_{L^{\infty}}$ is controlled by $\Lambda$. Therefore
816: writing $\phi(x,\sigma)-\phi(x,\sigma_{i})=\int_{\sigma_{i}}^{\sigma}\dot{\phi}(x,\sigma)d\sigma$ we see that
817: if $-\phi(x,\sigma_{i})\geq \Gamma$ there is $T(\Lambda,\Gamma)$ such that $-\phi(x,\sigma)\geq \Gamma/2$ for every
818: $\sigma \in [\sigma_{i},\sigma_{i}+T(\Lambda,\Gamma)]$. Now suppose there is a subsequence of $\{\sigma_{i}\}$ denoted
819: by $\{\sigma_{i_{j}}\}$ such that ${\mathcal{V}}(\Omega_{\phi,\Gamma}(\sigma_{i_{j}}))\geq M$ for some $M>0$, then
820: $\dot{\mathcal{V}}(\sigma)=\int_{\Sigma}(d\nu)\dot{}=\int_{\Sigma}3\phi d\nu\leq -3MT(\Lambda)/2$ for any $\sigma\in
821: [\sigma_{i_{j}},\sigma_{i_{j}}+T(\Lambda,\Gamma)]$. Therefore as ${\mathcal{V}}$ is monotonic, 
822: it must be ${\mathcal{V}}(\sigma)\downarrow -\infty$ as $\sigma\rightarrow \infty$ which is absurd.
823: 
824: \vspace{0.2cm}
825: \n {\it 2}. This is direct from {\it 1} above and Proposition \ref{P1} {\it 2}.
826: 
827: \vspace{0.2cm}
828: \n {\it 3}. We prove first the claim that if $|\nabla \Kt(x,\sigma)|_{\g}\geq \Gamma$ then there is $r(\Lambda,\Gamma)$ and $x'\in B(x,r)$
829: such that $|\Kt(x',\sigma)|_{\g}\geq M$ for some $M(\Lambda,\Gamma)>0$. This shows that $\Omega_{\nabla\Kt,\Gamma}(\sigma)\subset B(\Omega_{\Kt,M}(\sigma),r)$.
830: Once this is proved, by Propositions \ref{P2} and \ref{P3} {\it 2} $B(\Omega_{\Kt,M}(\sigma_{i}),r)$ and therefore $\Omega_{\nabla\Kt,\Gamma}(\sigma_{i})$ have
831: AZRV which would finish this item. Now let us prove the claim. From now on if at $x$, $inj_{\g(\sigma)}(x)$ is small
832: we unwrap to have $inj_{\g(\sigma)}(x)\geq inj_{0}>0$. By Proposition \ref{P1} {\it 1} we have that $\nabla\Kt$ is controlled 
833: in $C^{\alpha}$, therefore
834: $|\nabla \Kt(x,\sigma)-\nabla \Kt(y,\sigma)|_{\g}\leq C(\Lambda) d(x,y)^{\alpha}$. Pick a unit vector $v(x)$ at $x$ such that
835: $|\nabla_{v}\Kt(x,\sigma)|_{\g}\geq \Gamma/3$. Pick a harmonic chart $\{x^{i}\}$ covering the ball $B(x,inj_{0})$
836: such that the Christoffel symbols $\Gamma_{ij}^{k}$ are zero at $x$ (this is always possible) and write in the $\{x^{i}\}$
837: coordinates
838: \ben
839: (\nabla_{v}\Kt)_{jk} (y)=v^{i}\partial_{x^{i}}\Kt_{jk}(y)-\Gamma_{jk'}^{l}(y)\Kt_{lk}(y)v^{k'}(y)-\Gamma_{kk'}^{l}(y)\Kt_{jl}(y)v^{k'}(y).
840: \een
841: 
842: \n Pick $r(\Lambda,\Gamma)\leq inj_{0}$ with $r^{\alpha}C(\Lambda)\leq \Gamma/20$ such that 
843: $|\Gamma_{ij}^{l}\Kt_{lk}|\leq \Gamma/40$ on $B(x,r)$. Then on $B(x,r)$ we have $|\partial_{v}\Kt(x)-\partial_{v}\Kt(y)|\leq \Gamma/10$
844: and therefore $|\Kt(p)-\Kt(q)|\geq \Gamma 2r/4$ where $p$ and $q$ are the intercepts of the line (in the coordinate
845: system $\{x^{i}\}$) $x+\lambda v$ and the boundary of the ball $B(x,r)$. So either $|\Kt(p)|$ or $|\Kt(q)|$ must be
846: greater or equal to $\Gamma r/4$. This finishes the proof of the claim.
847: 
848: \vspace{0.2cm} 
849: \n {\it 4}. We start by noting the following. Say $f$ and $h$ are tensorial quantities such that, given $\Gamma>0$ and any DSLT 
850: , $\{\sigma_{i}\}$, $\|h\|_{L^{\infty}_{\g}}(\sigma_{i})$ and $\|f\|_{L^{\infty}_{\tilde{g}}}$ are controlled by $\Lambda$ and 
851: the sequence of sets 
852: $\Omega_{f,\Gamma}(\sigma_{i})=\{x\in\Sigma/|f(x,\sigma_{i})|_{\g}\geq \Gamma\}$ has AZRV, then: 
853: a) for any DSLT, $\{\sigma_{i}\}$, it is $(\int_{\Sigma}|h*f|_{\g}^{2}dv_{\g})(\sigma_{i})\rightarrow 0$ ($*$ is some tensorial composition) 
854: and b) for any pair of DSLT, $\{\sigma_{i}\}$ and $\{\sigma'_{i}\}$ as
855: in the statement of this item (Prop \ref{P3}, {\it 4}), it is $\int_{\sigma'_{i}}^{\sigma_{i}}(\int_{\Sigma}|h*f|^{2}dv_{\g})d\sigma\rightarrow 0$ as 
856: $\sigma_{i}\rightarrow \infty$. The claim
857: a) is obvious by writing $|h*f|_{\g}\leq c|h|_{\g}|f|_{\g}$ for some numeric $c$. For the claim b) observe that if the claim holds
858: for $h=1$ it holds for any $h$. Now if it is false for $h=1$ we can extract a sequence of logarithmic times $\{\bar{\sigma}_{i}\}$
859: with $\sigma_{i}\geq \bar{\sigma}_{i}\geq \sigma'_{i}$ and $(\int_{\Sigma}|f|_{\g}^{2}dv_{\g})(\bar{\sigma_{i}})\nrightarrow 0$ which contradicts a).
860: 
861: Now note that by {\it 2} above, the claim a) and 
862: \ben
863: Curl_{\g}\K=-B,
864: \een
865: 
866: \n we have that for any DSLT $\{\bar{\sigma}_{i}\}$, $\|B\|_{L^{2}_{\g}}(\sigma_{i})\rightarrow 0$. 
867:      
868: We will prove this item by studying the quantity
869: \begin{equation}\label{form}
870: \int_{\Sigma}<E,\Kt>_{\g}dv_{\g},
871: \end{equation}
872: 
873: \n and its derivative with respect to logarithmic time. Differentiating it with respect to $\sigma$ we get
874: \begin{equation}\label{EK}
875: \begin{split}
876: (\int_{\Sigma}<E,\Kt>_{\g}dv_{\g})\dot{}&=\int_{\sigma} <\dot{E},\Kt>_{\g}+<E,\dot{\Kt}>_{\g}-<E\circ\Kt_{\g},\dot{\g}>+\\
877: &\quad +3<E,\Kt>_{\g}\phi dv_{\g}.
878: \end{split}
879: \end{equation}
880: 
881: \n To estimate the terms on the right hand side of the last equation we appeal to the equations
882: \begin{equation}\label{gdot}
883: \dot{\g}=2\phi \g-6\tilde{N}\Kt,
884: \end{equation}
885: \begin{equation}\label{Edot}
886: \dot{E}=\N Curl_{\g} B -\frac{\nabla \N}{\N}\wedge_{\g} B-\frac{5}{2}E\times_{\g} \K-\frac{2}{3}<E,\K>_{\g}\g-\frac{3}{2}E,
887: \end{equation}
888: \begin{equation}\label{Kdot}
889: \dot{\Kt}=-\Kt-\phi\g-\nabla^{2}\phi+\phi E+E-\N(\Kt\circ\Kt-2\Kt).
890: \end{equation}
891: 
892: \n Recall $3\N=\phi+1$. Integrate both sides of equation \ref{EK} in the interval $[\sigma'_{i},\sigma_{i}]$ and plug in
893: the equations (\ref{Edot}), (\ref{Kdot}), (\ref{gdot}). By the claim a) above the left hand side (of the integrated equation) 
894: converges to zero as $i\rightarrow \infty$. Using the items {\it 1}, {\it 2} and the claim b) above we get that
895: the only terms on the right hand side (of the integrated equation) that may not converge to zero as $i\rightarrow \infty$ are
896: \be\label{Ephi}
897: \int_{\sigma'_{i}}^{\sigma_{i}}\int_{\Sigma}<E,\nabla^{2}\phi>_{\g}dv_{\g}d\sigma,
898: \ee
899: \begin{equation}\label{BKt}
900: \int_{\sigma'_{i}}^{\sigma_{i}}\int_{\Sigma}\N <Curl_{\g} B,\Kt>_{g}dv_{\g}d\sigma,
901: \end{equation}
902: 
903: \n and
904: \begin{equation}\label{E^2}
905: \int_{\sigma'_{i}}^{\sigma_{i}}\int_{\Sigma}|E|^{2}_{\g}dv_{\g}d\sigma.
906: \end{equation}
907: 
908: \n To show that the equation (\ref{Ephi}) and (\ref{BKt}) converge to zero as $i\rightarrow \infty$ we integrate by parts.
909: In equation (\ref{Ephi}) integration by parts gives
910: \ben
911: \int_{\sigma'_{i}}^{\sigma_{i}}\int_{\Sigma}-<Div E,\nabla \phi>_{g}dv_{\g}d\sigma,
912: \een
913: 
914: \n which by the formula $Div_{\g} E=\K\wedge_{\g} B$ and the claim b) above is guaranteed to converge to zero as
915: $i\rightarrow \infty$. To integrate by parts on equation \ref{BKt} invoke the formula
916: \ben
917: Div(U\wedge U')=-<Curl U,U'>+<U,Curl U'>,
918: \een
919: 
920: \n holding for any $U$ and $U'$ traceless symmetric tensors. This gives 
921: \begin{equation}\label{intparts}
922: \int_{\sigma'_{i}}^{\sigma_{i}}\int_{\Sigma}<B,\nabla \N * \Kt+\N Curl\Kt>dv_{\g}d\sigma,
923: \end{equation}
924: 
925: \n ($*$ is some tensor operation). After using the formula $Curl_{\g}\Kt=-B$ in equation (\ref{intparts}), claim b)
926: above shows that all the expression goes to zero as $i\rightarrow \infty$. We are thus lead to conclude that the
927: expression (\ref{E^2}) goes to zero as $i\rightarrow \infty$ as desired.
928: 
929: \vspace{0.2cm}
930: \n {\it 5}. We have shown in ${\it 4}$ that $\|B\|^{2}_{L^{2}_{\g(\sigma_{i})}}$ goes to zero as $i\rightarrow \infty$
931: for any DSLT $\{\sigma_{i}\}$. To show that the same happens for $E$ we make use of the Bel-Robinson energy. Observe
932: that $\frac{1}{\mathcal{H}}\int_{\Sigma}(|E|_{g}^{2}+|B|^{2}_{g})dv_{g}
933: =\frac{1}{\mathcal{H}}\int_{\Sigma}Q_{TTTT}({\bf Rm})dv_{g}=\int_{\Sigma}Q_{\Ti\Ti\Ti\Ti}(\tilde{\bf Rm})dv_{\g}=
934: \int_{\Sigma}(|E|^{2}_{\g}+|B|^{2}_{\g})dv_{\g}$. Using the Gauss equation we get 
935: \begin{equation}\label{Gaussint}
936: \dot{\tilde{Q}}=\tilde{Q}-9\int_{\Sigma}\N \tilde{Q}_{\alpha\beta \Ti\Ti}\tilde{\bf \Pi}^{\alpha\beta}dv_{\g}.
937: \end{equation}
938: 
939: \n Now suppose there is a DSLT, $\{\sigma_{i}\}$, such that $\int_{\Sigma}|E|^{2}_{\g}dv_{g}>M$ for some $M>0$. As the
940: right hand side of equation (\ref{Gaussint}) is controlled by $\Lambda$ we can find $T(\Lambda, M)$ such that the 
941: integral in $\sigma$ of the right hand side of equation (\ref{Gaussint}) on the interval $[\sigma_{i},\sigma_{i}+T]$
942: is, in absolute value, less than $M/2$ and therefore integrating equation (\ref{Gaussint}) on the interval $[\sigma_{i},\sigma]$
943: with $\sigma\in [\sigma_{i},\sigma_{i}+T]$ we get $\tilde{Q}(\sigma)\geq M/2$. Therefore $\int_{\sigma_{i}}^{\sigma_{i}+T}\tilde{Q}d\sigma\geq TM/2$.
944: However by ${\it 4}$ we know $\int_{\sigma_{i}}^{\sigma_{i}+T}\int_{\Sigma}|E|^{2}_{\g}+|B|^{2}_{\g}dv_{\g}d\sigma\rightarrow
945: 0$ which is a contradiction.     
946: 
947: \hspace{\stretch{1}}$\Box$\\
948: 
949: We are now ready to prove Theorems $1$ and $2$ (stated conveniently).
950: 
951: \begin{T} Say $Y(\Sigma)\leq 0$ and say $(\g,\K)$ a cosmologically normalized flow satisfying the curvature assumption. 
952: Then $(\g,\K)$ induces a unique persistent geometrization on $\Sigma$.
953: \end{T}
954: 
955: \n {\bf Proof:} 
956: 
957: We prove first there is a DSLT, $\{\sigma_{i}\}$ with
958: $(\Sigma^{\frac{1}{i}},(\g,\K)(\sigma_{i}))$ converging to 
959: 
960: \n $\cup_{i=1}^{i=n}(H_{i},(\g_{H,i},-\g_{H,i}))$ (in $C^{2,\beta}$).
961: Introduce a new variable $j=1,2,3,\ldots$. For $j=1$ find a sequence $\{\sigma_{1,i}\}$ with 
962: $(\Sigma^{1},\g(\sigma_{1,i}))$ convergent in $C^{2,\beta}$. For $j=2$ find a subsequence $\{\sigma_{2,i}\}$ 
963: of $\{\sigma_{1,i}\}$ with $(\Sigma^{1/2},\g(\sigma_{2,i}))$ convergent in $C^{2,\beta}$. Proceed similarly
964: for all $j$ to have a double sequence $\{\sigma_{j,i}\}$. Now, the diagonal sequence $\{\sigma_{i,i}\}$, $(\Sigma^{1/i},\g(\sigma_{i,i}))$
965: converges into a union of complete manifolds of finite volume, denoted as $\cup_{\nu} (M_{\nu},\g_{\infty,\nu})$. 
966: By Proposition \ref{P3} {\it 2}, $\K(\sigma_{i,i})$ converges to $-\g_{\infty,\nu}$ in $C^{1,\beta}$. By proposition \ref{P3} {\it 5} and
967: the formula $E=\tilde{Ric}-3\K+\K\circ\K$ we get that each metric $\g_{\infty,\nu}$ is hyperbolic. Therefore, as there
968: is a lower bound for the volume of complete hyperbolic manifolds of finite volume and the total volume of the limit 
969: space is bounded above, there must be a finite number of components, and we can write $\cup_{\nu} (M_{\nu},\g_{\infty,\nu})=\cup_{i=1}^{i=n}(H_{i},\g_{H,i})$.
970:  
971: We prove next that each component $(H_{j},\g_{H,j})$ is persistent. For simplicity assume there is only one component
972: and therefore $(\Sigma^{1/i},\g(\sigma_{i,i}))$ converges in $C^{2,\beta}$ to $(H,\g_{H})$. There are two possibilities according to whether
973: the component is compact ot not, we discuss them separately.
974: 
975: 1.{\it (The compact case)} Assume $(H,\g_{H})$ is compact. Consider the space of metrics ${\mathcal{M}}_{H}$ in $H$. For every metric $g$ consider
976: the orbit of $g$ under the diffeomorphism group. Denote such orbit by $o(g)$. Around $\g_{H}$
977: consider a small (smooth) section ${\mathcal{S}}$ of ${\mathcal{M}}_{H}$ of $C^{2,\beta}_{\g_{H}}$ metrics transverse to the 
978: orbits generated by the action on ${\mathcal{M}}_{H}$ of the diffeomorphism group
979: \footnote{Which particular section is taken is unimportant. One can use for instance ${\mathcal{S}}=\{g/id:(H,g)\rightarrow (H,\g_{H})\}$
980: is harmonic (see \cite{AM}, \cite{Ham}). The same comment applies in the non-compact case.}. For an illustration see Figure \ref{fig2}. 
981: 
982: \begin{figure}[h]
983: \centering
984: \includegraphics[width=110mm,height=60mm]{Dif.JPG}
985: \caption[U]{Representation of the space of metrics on a neighborhood of $\g_{H}$.}\label{fig2}
986: \end{figure}
987: 
988: If $\epsilon_{0}$ is sufficiently small 
989: every ($C^{\infty}$) metric $g$ in ${\mathcal{M}}_{H}$ with $\|g-\g_{H}\|_{C^{2,\beta}_{\g_{H}}}\leq \epsilon_{0}$ can be uniquely projected 
990: into ${\mathcal{S}}$
991: by a diffeomorphism, or in other words we can consider the projection $P(g)=o(g)\cap {\mathcal{S}}$. Note that one can project
992: every $(C^{\infty}$) path of ($C^{\infty}$) metrics $g(t)$ starting close to $\g_{H}$, to a $(C^{\infty})$ path $P(g(t))$, 
993: until at least the first time when $\|P(g(t))-\g_{H}\|_{C^{2,\beta}_{\g_{H}}}= \epsilon_{0}$
994: or in other words until at least when the projection touches the boundary of the ball of center $\g_{H}$ and radius $\epsilon_{0}$ 
995: in $C^{2,\beta}_{\g_{H}}$ (denote such ball as
996: $B(\g_{H},\epsilon_{0})$)\footnote{We consider $C^{\infty}$ paths of $C^{\infty}$ metrics because we have assumed the solution ${\bf g}$ and
997: therefore the zero shift flow to be $C^{\infty}$. It is not difficult to show that independent of the section ${\mathcal{S}}$ 
998: considered, a path $g(t)$ as above (leaving or not the ball $B(\g_{H},\epsilon)$) can be projected into ${\mathcal{S}}$ until at least a first time
999: when the projection touches the boundary of the ball. Note that if $\phi^{*}_{t}(g(t))=P(g(t))\in B(g_{H},\epsilon_{0})\cap {\mathcal{S}}$ for $t<t_{*}$ 
1000: then for every $t_{1}<t_{*}$ $\phi_{t_{1}}^{*}(g(t))$ is a path in $B(g_{H},\epsilon_{0})$ for $t$ in a neighborhood of $t_{1}$ and which therefore can
1001: be projected into ${\mathcal{S}}$.} 
1002: 
1003: Recall Mostow rigidity\footnote{Mostow rigidity says that any two hyperbolic metrics on a compact manifold are necesarily isometric. What we 
1004: state as {\it Mostow rigidity} here is an obvious consequence of this fact.}
1005: \vspace{0.1cm}
1006: \n {\it Mostow rigidity (the compact case)}. There is $\epsilon_{1}$ such that if $P(g'_{H})\in B(\g_{H},\epsilon_{1})$, where
1007: $g'_{H}$ is a hyperbolic metric in $H$ then $P(g'_{H})=\g_{H}$. 
1008: 
1009: \vspace{0.1cm}
1010: Fix $\epsilon_{2}=min\{\epsilon_{0},\epsilon_{1}\}$. Observe that as $\g_{\sigma_{i,i}}\rightarrow \g_{H}$ in $C^{2,\beta}$ there is 
1011: a sequence of diffeomorphisms $\phi_{i}$ such that $\phi^{*}_{i}(g(\sigma_{i,i}))$ converges to $\g_{H}$ in $C^{2,\beta}_{\g_{H}}$.
1012: Now, if the geometrization is not persistent there is $\epsilon\leq \epsilon_{2}$ and $i_{2}$ such that if $i\geq i_{2}$ then 
1013: $P(\phi^{*}_{i}(\g(\sigma)))$ is well defined for $\sigma\geq \sigma_{i,i}$ until a first time $\sigma_{i,i}+T_{i}$ when 
1014: $P(\phi^{*}_{i}(\g(\sigma_{i,i}+T_{i})))$ is in $\partial B(\g_{H},\epsilon_{2})$. But we know the sequence of Riemannian
1015: manifolds $(H,P(\phi^{*}_{i}(\g(\sigma_{i,i}+T_{i}))))$ converge in $C^{2,\beta}$ to $\g_{H}$, and that means by the definition of $C^{2,\beta}$ convergence and 
1016: Mostow rigidity 
1017: that there is a sequence of diffeomorphisms
1018: $\varphi_{i}$ such that $P(\varphi^{*}_{i}(P(\phi^{*}_{i}(\g(\sigma_{i,i}+T_{i}))))$ converge to $\g_{H}$ in 
1019: $C^{2,\beta}_{\g_{H}}$. This contradict the fact that $P(\phi^{*}_{i}(\g(\sigma_{i,i}+T_{i})))$ is in $\partial B(\g_{H},\epsilon_{2})$.
1020: 
1021: 2. {\it (The non-compact case)}. The proof 
1022: of this case proceeds along the same lines as the compact case but special care must be taken at the cusps. Let us assume for 
1023: simplicity that there is only one cusp in the piece $(H,\g_{H})$. Given $A$
1024: sufficiently small there is a unique torus transverse to the cusp, to be denoted by $T^{2}_{A}$, of constant mean curvature 
1025: and area $A$. Denote by $H_{A}$ the ``bulk" side of the torus $T^{2}_{A}$ in $H$. Consider the set of metrics ${\mathcal{M}}_{H_{A}}$
1026: on $H_{A}$ such that for any of them $T^{2}_{A}$ has constant mean curvature and area $A$. Consider the action of the diffeomorphism
1027: group on ${\mathcal{M}}_{H_{A}}$ leaving the torus $T^{2}_{A}$ invariant. Again the orbit of $g$ will be denoted by $o(g)$.
1028: Consider a small (smooth) section of ${\mathcal{S}}$ of $C^{2,\beta}$ metrics around $\g_{H}$ and transverse to the orbits of 
1029: the action by the diffeomorphism group mentioned above. Finally consider the projection $P(g)=o(g)\cap {\mathcal{S}}$ which is
1030: well defined on a ball $B(\g_{H},\epsilon_{0})$ for $\epsilon_{0}$ small enough. Observe again that a path $g(t)$ of
1031: metrics in ${\mathcal{M}}_{H_{A}}$ can be projected into ${\mathcal{S}}$ until at least the first time 
1032: when $P(g(t))$ is in $\partial B(\g_{H},\epsilon_{0})$. Slightly abusing the notation (as we would need a pointed sequence)
1033: consider the sequence $(\Sigma,\g(\sigma_{i,i}))$ converging in $C^{2,\beta}$ to $\g_{H}$. If $i\geq i_{0}$ we can identify on $\Sigma$
1034: a torus $T^{2}_{A,\g(\sigma_{i,i})}$ of constant mean curvature and area $A$ which converges as $i\rightarrow \infty$ (and after the application of a diffeomorphism) 
1035: to $T^{2}_{A}$ in $H$\footnote{For a proof of this fact see the footnote in page 328 in \cite{Ham}.}. 
1036: More in particular there is a sequence of diffeomorphisms (onto the image) 
1037: $\phi_{i}:H_{A}\rightarrow \Sigma$ such that 
1038: $\phi_{i}(T^{2}_{A})=T^{2}_{A,\g(\sigma_{i,i})}$ and $\|\phi^{*}_{i}(\g(\sigma_{i,i}))-\g_{H}\|_{C^{2,\beta}_{\g_{H}}}$
1039: converging to zero. We note the following crucial facts (justified below).
1040: 
1041: i) The diffeomorphisms (onto the image) $\phi_{\sigma}:H_{A}\rightarrow \Sigma$ with $\phi_{\sigma}(T^{2}_{A})=T^{2}_{A,\g(\sigma)}$ can
1042: be defined (varying differentiably) as long as the tori $T^{2}_{A,\g(\sigma)}$ are well defined (varying differentiably).
1043: 
1044: ii) There are $\sigma_{0}$ and $\epsilon_{1}$ such that if $\sigma_{1}\geq \sigma_{0}$ and $\phi_{\sigma_{1}}:H_{A}\rightarrow \Sigma$ is well 
1045: defined and $\|\phi^{*}_{\sigma_{1}}(\g(\sigma_{1}))-\g_{H}\|_{C^{2,\beta}_{\g_{H}}}\leq \epsilon_{1}$ then the tori $T^{2}_{A,\g(\sigma)}$ are
1046: well defined, varying differentiably for $\sigma$ on a neighborhood of $\sigma_{1}$.
1047: 
1048: iii) If for $\sigma_{1}\geq \sigma_{0}$ $\phi_{\sigma_{1}}:H_{A}\rightarrow \Sigma$ is well defined and satisfies 
1049: $\|\phi^{*}_{\sigma_{1}}(\g(\sigma_{1}))-\g_{H}\|_{C^{2,\beta}_{\g_{H}}}\leq \epsilon_{1}$ then $\phi_{\sigma}$ is well defined 
1050: until at least the first time $\sigma_{2}\geq \sigma_{1}$ when $\|P(\phi^{*}_{\sigma_{2}}(\g(\sigma_{2})))-\g_{H}\|_{C^{2,\beta}_{\g_{H}}}
1051: =\epsilon_{1}$.
1052:     
1053: The fact i) is self evident. The fact ii) is the most important to consider and can be justified as follows. It is well known
1054: that under curvature bounds the injectivity radius cannot collapse in finite distance from a region that is non collapsed. In
1055: particular if $\|\phi^{*}_{\sigma_{1}}(\g(\sigma_{1}))-\g_{H}\|_{C^{2,\beta}_{\g_{H}}}\leq \epsilon_{1}$ for $\epsilon_{1}$ sufficiently small
1056: the ``bulk" side of $T^{2}_{A,\g(\sigma_{1})}$ is non collapsed and therefore the `` cusp" side of $T^{2}_{A,\g(\sigma_{1}))}$
1057: is non collapsed in finite distances from the ``bulk" side. Now, if $\sigma_{1}$ is big enough the $C^{\beta}$ norm of $\hat{Ric}$ around 
1058: $T^{2}_{A,g(\sigma_{1})}$ must be small otherwise one may find a DSLT for which the pointed sequence 
1059: $(\Sigma,p_{i},\g(\sigma_{i}))$ with $p_{i}\in T^{2}_{A,\g(\sigma_{i})}$
1060: is not converging into a complete hyperbolic metric of finite volume. This shows in particular that if $\sigma_{0}$ is big 
1061: enough the geometry nearby $T^{2}_{A,\g(\sigma_{1})}$ is close (in $C^{2,\beta}$) to the geometry nearby $T^{2}_{A}$ in $H$. 
1062: By the continuity of the flow the tori $T^{2}_{A,\g(\sigma)}$ are well defined for $\sigma$ in a neighborhood of $\sigma_{1}$.
1063: The fact iii) follows directly from ii).
1064: 
1065: Recall Mostow Rigidity\footnote{For a proof of this fact as well as for realted discussions the reader can consult \cite{Ham} (footnote on page 323).}
1066: 
1067: \vspace{0.1cm}
1068: \n {\it Mostow rigidity (the non-compact case)}. There is $A_{0}$ such that for any $A\leq A_{0}$ there is $\epsilon_{0}$ such
1069: that if $(\Sigma',g'_{H})$ is a complete hyperbolic manifold of finite volume and $\phi:H_{A}\rightarrow \Sigma'$ is a
1070: diffeomorphism onto the image satisfying $\|\phi^{*}(g_{H}')-\g_{H}\|_{C^{2,\beta}_{\g_{H}}}\leq \epsilon_{0}$ then $(\Sigma',g'_{H})$
1071: is isometric to $(H,\g_{H})$. In particular $P(\phi^{*}(g'_{H}))=\g_{H}$.
1072: 
1073: \vspace{0.1cm}
1074: 
1075: Given $A\leq A_{0}$ but so far arbitrary, fix $\epsilon_{2}=min\{\epsilon_{0},\epsilon_{1}\}$. Due to the facts i), ii) and iii) we have that if the 
1076: geometrization is not persistent there is $\epsilon\leq \epsilon_{2}$ and $i_{2}$ such that if $i\geq i_{2}$ then 
1077: $P(\phi^{*}_{i}(\g(\sigma)))$ is well defined for $\sigma\geq \sigma_{i,i}$ until a first time $\sigma_{i,i}+T_{i}$ when 
1078: $P(\phi^{*}_{i}(\g(\sigma_{i,i}+T_{i})))$ is in $\partial B(\g_{H},\epsilon_{2})$. Now the sequence $P(\phi^{*}_{i}(\g(\sigma_{i,i}+T_{i})))$ 
1079: has a subsequence converging in $C^{2,\beta}$ to a complete hyperbolic metric in finite volume. Again as in the compact case, by Mostow rigidity it must be
1080: converging in $C^{2,\beta}_{\g_{H}}$ to $\g_{H}$ contradicting the fact that $P(\phi^{*}_{i}(\g(\sigma_{i,i}+T_{i})))$ is in $\partial B(\g_{H},\epsilon_{2})$.
1081:  
1082: To finish the proof of the persistence of the geometrization one still needs to show that the compliment of the persistent 
1083: pieces $(H_{i},\g_{H,i})$ is the $G$ sector or in other words that for any $\epsilon>0$, $(\Sigma^{\epsilon}(\sigma),\g(\sigma))$ converges to 
1084: the $\epsilon$-thick part of the persistent pieces 
1085: $(H_{i},\g_{H,i})$. The proof of this fact follows by contradiction. If this is not the case one can extract a DSLT containing 
1086: an $H$ piece different  from the pieces $(H_{i},\g_{H,i})$. One can prove again that this new piece is persistent leading into 
1087: a contradiction for if persistent, the piece must be one of the pieces $(H_{i},\g_{H,i})$  by the way these pieces 
1088: are defined.  
1089: \ep
1090: 
1091: \begin{T} Say $Y(\Sigma)\leq 0$ and say $(\g,\K)$ is cosmologically normalized flow satisfying the curvature
1092: assumption. Then $\lim_{\epsilon\rightarrow 0}(\lim_{\sigma} {\mathcal{V}}(\Sigma_{\epsilon}))=0$.
1093: \end{T}
1094: 
1095: \n {\bf Proof:}
1096: 
1097: First observe that as the geometrization is persistent and the reduced volume is monotonic, each one of the limits, in
1098: $\sigma$ first and in $\epsilon$ later exists. 
1099:      
1100: Consider the sets $\Omega_{H,\Gamma}(\sigma)=\{x\in \Sigma / |\tilde{\hat{Ric}}(x,\sigma)|_{\g}\leq \Gamma\}$. These sets have 
1101: ATRV when $\sigma\rightarrow \infty$ for any fixed $\Gamma$ and $r$. 
1102: Now consider the set $\Omega_{H,\Gamma,r}(\sigma)=\{x\in\Omega_{H,\Gamma}/|\Rt(x',\sigma)|_{\g}\leq 2\Gamma,\ 
1103: {\rm for\ all}\ x'\in B(x,r)\}$. 
1104: This set has ATRV because the complement is contained in the set $B(\Omega_{\tilde{\hat{Ric}},\Gamma}(\sigma),r)$ which we know must have AZRV
1105: by Propositions \ref{P2} and \ref{P3}. We will need an isoperimetric inequality for the balls of radius one of the regions $\Sigma_{\epsilon}\cap \Omega_{H,\Gamma, r}$ for
1106: a suitable value of $r$ and $\Gamma$. These values of $r$ and $\Gamma$ will come out later using the proposition below. 
1107: Recall Margulis lemma (\cite{Th} corollary 5.10.2)
1108: 
1109: \begin{Lem} \label{Mar} (Margulis) There is $\epsilon_{0}$ such that for any complete hyperbolic three-manifold $\Sigma$, $\Sigma_{\epsilon_{0}}$ is
1110: (the $\Sigma_{\epsilon_{0}}$ part of) one of the following models:
1111: 
1112: \begin{enumerate}
1113: 
1114: \item A horoball modulo $\field{Z}$ or $\field{Z}\times \field{Z}$ (where the action on the half-space model is by horizontal Euclidean translations) or, 
1115: 
1116: \item a ball around a geodesic $\gamma$ of some radius $R$ modulo $\field{Z}$ (where the action is by translations along the geodesic $\gamma$).
1117: 
1118: \end{enumerate}
1119: 
1120: \end{Lem}
1121: 
1122: We use such $\epsilon_{0}$ in the proposition below.
1123: 
1124: \begin{Prop}\label{PM} For any $\delta>0$, there is $\epsilon,\sigma_{0},r>1$ and $\Gamma$ such that at any $\sigma\geq \sigma_{0}$, and
1125: for any ball $B(x,r)$ with $x\in\Sigma_{\epsilon}(\sigma)\cap\Omega_{H,\Gamma,r}(\sigma)$ we can unwrap the ball to have
1126: $inj(x)\sim \epsilon_{0}$ and such that on it the ball of radius one is $\delta$-close in $C^{2,\beta}$ to a ball of
1127: radius one in one of the Margulis models 
1128: above.
1129: \end{Prop}
1130: 
1131: \n{\bf Proof} (of Proposition \ref{PM}): 
1132: 
1133: Suppose by contrary there is a $\delta>0$ such that the proposition doesn't hold. Then for
1134: every $i$ the conclusion is false for the set of parameters $\Gamma=1/i$, $r=i$, $\sigma_{0}=i$ and $\epsilon(i)$ chosen
1135: in such a way that (at any logarithmic time) for any ball $B(x,r=i)$ with 
1136: $x\in \Sigma_{\epsilon(i)}\cap \Omega_{H,\Gamma=1/i,r=i}$ we can unwrap the ball to have $inj_{\g}(x)\sim \epsilon_{0}$. As we are assuming 
1137: the conclusion is false for any $i$, we can find for any $i$, $\sigma_{i}>\sigma_{0}=i$ and $x_{i}$ in $\Sigma_{\epsilon_{i}}(\sigma_{i})\cap 
1138: \Omega_{H,\Gamma=1/i,r=i}(\sigma)$ such the unwrapped ball is $\delta$-far from one of the Margulis models above. Now the
1139: sequence (in $i$) of such unwrapped balls converges in $C^{2,\beta}$ to a complete Riemannian manifold and because $x_{i}$ is 
1140: in $\Omega_{H,\Gamma=1/i,r=i}(\sigma_{i})$ for every $i$ it must be hyperbolic and therefore one of the Margulis models by Lemma \ref{Mar}. This
1141: is a contradiction. \hspace{\stretch{1}}$\Box$\\
1142:  
1143: Now observe that at each point in any of the Margulis models there is one and only one direction where the size of 
1144: the (collapsed) fibers expands most.
1145: In the first two examples the directions are determined by the vertical geodesic congruence and in the third by the 
1146: congruence of geodesics coming
1147: out perpendicularly from the geodesic $\gamma$. Observe that the direction is invariant under wrappings or 
1148: unwrapping. Define in each model a field $X$ having norm one and 
1149: in the direction of maximal fiber expansion. For example if the model is a cusp, i.e. a horoball 
1150: modulo $\field{Z}\times\field{Z}$ then (writing the metric as $g_{H}=dx^{2}+e^{2x}h$ where $h$ is the flat metric
1151: in the two-torus induced by the action of $\field{Z}\times\field{Z}$ in $\field{R}^{2}$) the vector field $X$ is $\partial_{x}$. 
1152: It is a straight forward 
1153: calculation that the divergence of $X$ in any one of the models is positive and bounded below and above say by 
1154: $C_{1}$ and $C_{2}$ ($0<C_{1}<C_{2}$). For example in the cusp case the divergence of $X$ is computed as $\nabla. X=\frac{1}{\sqrt{|g_{H}}|}\partial_{x}(\sqrt{|g_{H}|})=2$
1155: with $|g_{H}|=A^{2}e^{4x}$ the determinant of the metric $g_{H}$ in the coordinates $(x,\theta_{1},\theta_{2})$ ($(\theta_{1}$ and $\theta_{2}$ are the natural
1156: coordinates on $T^{2}=S^{1}\times S^{1}$ and A is the area under $h$). Suppose now that
1157: we have a manifold $U$ made out of a finite (but arbitrary) set of balls of radius one taken from any one of the 
1158: three Margulis models. The balls may touch each other in an arbitrary fashion, and therefore the boundary may not be 
1159: entirely smooth although it is in a set of total measure. Then
1160: \ben\label{II}
1161: C_{1}Vol(U)\leq \int_{U} \nabla.X dv = \int_{\partial(U)} <X,n>dS \leq Vol(\partial U),
1162: \een
1163: 
1164: \noindent (where we have used the fact that $X$ is of norm one and therefore $<X,n>\leq 1$). 
1165: Observe also that for any $s<1$ 
1166: \begin{equation}\label{II2}
1167: Vol(B(\partial U,s))\geq C(s) Vol(U).
1168: \end{equation}
1169: 
1170: \n To see that observe that every point in the 
1171: smooth part of the boundary
1172: joins with one and only one closest center. Then using the isoperimetric inequality \ref{II} above, the set formed 
1173: by all segments of length $s$ 
1174: starting at the points in the smooth parts of the boundary and in the direction of their unique center must have a 
1175: definite part of 
1176: the total volume. We have 
1177: similar inequalities if the balls are made out of balls $\delta$-close in
1178: $C^{2,\beta}$ to one of the Margulis models for $\delta$ sufficiently small. From now on take such $\delta$ in 
1179: Proposition \ref{PM}, 
1180: to get the parameters $\Gamma$, $r$ and $\epsilon$.  
1181: 
1182: Now lets go back to finish the proof of Theorem \thetwo. Assume by the contrary there is a sequence 
1183: $\{\sigma_{i}\}$ such that $\lim_{\bar{\epsilon}\rightarrow 0}(\lim_{\sigma_{i}\rightarrow \infty} 
1184: {\mathcal{V}}(\Sigma_{\bar{\epsilon}}))={\mathcal{V}}_{0}>0$. We then have $\lim_{\bar{\epsilon}\rightarrow 0}
1185: (\lim_{\sigma_{i}
1186: \rightarrow \infty}{\mathcal{V}}(\Sigma_{\bar{\epsilon}}\cap \Omega_{H,\Gamma,r}))={\mathcal{V}}_{0}$. Fix
1187: $\bar{\epsilon}\leq \epsilon$. We can use the isoperimetric inequality (\ref{II2}) with $s=1/2$, to conclude that 
1188: $\lim_{\sigma_{i}\rightarrow \infty}{\mathcal{V}}(B(\partial B(\Omega_{H,\Gamma,r}(\sigma_{i})\cap \Sigma_{\bar{\epsilon}}(\sigma_{i}),1),s))$ 
1189: is bounded below by a nonzero constant independent of $\bar{\epsilon}$. Note that the set $B(\partial B(\Omega_{H,\Gamma,r}(\sigma_{i})\cap \Sigma_{\bar{\epsilon}}(\sigma_{i}),1),s))$ 
1190: is disjoint from the set $\Omega_{H,\Gamma,r}(\sigma_{i})\cap \Sigma_{\bar{\epsilon}}(\sigma_{i})$ and that the $\sup\{inj(x)_{\g}/ x\in B(\partial B(\Omega_{H,\Gamma,r}(\sigma_{i})\cap \Sigma_{\bar{\epsilon}}(\sigma_{i}),1),s))\}$
1191: converges to zero\footnote{Under curvature bounds the injectivity radius propagates over finite distances, in particular if $inj_{\g}(x)\rightarrow 0$ and $d(x,y)<C$ then $inj_{\g}(y)\rightarrow 0$.} as $i\rightarrow \infty$, therefore $\lim_{\bar{\epsilon}}(\lim_{\sigma\rightarrow \infty}{\mathcal{V}}(\Sigma_{\bar{\epsilon}}))>{\mathcal{V}}_{0}$
1192: which is a contradiction \hspace{\stretch{1}}$\Box$\\
1193: 
1194: We prove next Corollaries \ref{cor1}-\ref{cor5}. Corollary $2$ is direct from Theorem \thetwo. For corollary $3$ observe that
1195: since ${\mathcal{V}}$ is monotonic there cannot be any $H$ piece emerging and so we are in the situation of Corollary $2$. Corollary $5$
1196: is the content of Proposition \ref{P3} {\it 5}. To prove Corollary $1$ observe that as proved in \cite{R} (Theorem 9) given $\Lambda$
1197: there is $\epsilon$ such that if ${\mathcal{V}}(\g,\K)-{\mathcal{V}}_{inf}<\epsilon$ then the thick-thin decomposition
1198: implements the Thurston geometrization, therefore if the flow starts in a state $(\g,\K)(\sigma_{0})$ with ${\mathcal{V}}(\g,\K)(\sigma_{0})-
1199: {\mathcal{V}}_{inf}<\epsilon$ as ${\mathcal{V}}$ is monotonic the difference ${\mathcal{V}}(\g,\K)(\sigma)-{\mathcal{V}}_{inf}$ 
1200: is kept along the evolution, so the Thurston geometrization is the persistent geometrization. By Theorem \thetwo
1201: it must be ${\mathcal{V}}\downarrow {\mathcal{V}}_{inf}$. For Corollary \ref{cor4} observe again that by what has been proved
1202: in \cite{R} and Theorems $1$ and $2$, ${\mathcal{V}}\downarrow {\mathcal{V}}_{inf}$ iff the persistent long time geometrization is the Thurston geometrization
1203: iff the tori separating the $H$ and $G$ sectors are incompressible.
1204: 
1205: \addcontentsline{toc}{section}{\bf Bibliography} 
1206: \begin{thebibliography}{99}  
1207: 
1208: \bibitem[1]{R} Reiris, Martin. Large-scale (CMC) evolution of cosmological solutions of the Einstein equations with a priori bounded space-time curvature. arXiv:0705.3070
1209: 
1210: \bibitem[2]{R1} Reiris, Martin. General ${\cal K}=-1$ Friedman–Lemaître models and the averaging problem in 
1211: cosmology. {\it Classical and quantum Gravity, {\bf 25} (2008)}.
1212: 
1213: \bibitem[3]{FM} Fischer, Arthur E.; Moncrief, Vincent The reduced hamiltonian of general relativity and the 
1214: $\sigma$-constant of conformal geometry. 
1215: {\em Mathematical and quantum aspects of relativity and cosmology (Pythagoreon, 1998),
1216:  70--101, Lecture Notes in Phys., 537, Springer, Berlin, 2000.} 
1217: 
1218: \bibitem[4]{A1} Anderson, Michael T. On long-time evolution in general relativity and geometrization of 3-manifolds. 
1219: {\em Comm. Math. Phys. 222 (2001), no. 3, 533--567}.
1220: 
1221: \bibitem[5]{A3} Anderson, Michael T. Canonical metrics on 3-manifolds and 4-manifolds.  {\it 
1222: Asian J. Math.  10  (2006),  no. 1, 127--163.}
1223: 
1224: \bibitem[6]{A4} Anderson, Michael T. Scalar curvature and the existence of geometric structures on 3-manifolds. I. 
1225: {\em J. Reine Angew. Math. 553 (2002), 125--182}.
1226: 
1227: \bibitem[7]{CK} Christodoulou, Demetrios; Klainerman, Sergiu The global nonlinear stability of the Minkowski space. 
1228: {\em Princeton Mathematical Series, 41. Princeton University Press, Princeton, NJ, 1993}.
1229: 
1230: \bibitem[8]{AM} Andersson, Lars; Moncrief, Vincent. 
1231: Future complete vacuum space times. 
1232: {\em The Einstein equations and the large-scale behavior of gravitational fields, 299--330, 
1233: Birkhäuser, Basel, 2004}.
1234: 
1235: \bibitem[9]{FM2} Fischer, Arthur E.; Moncrief, Vincent The reduced Einstein equations and the conformal volume collapse of 3-manifolds.  
1236: {\it Classical and Quantum Gravity  18  (2001),  no. 21, 4493--4515.} 
1237: 
1238: \bibitem[10]{Th} Thurston, William. The geometry and topology of three-manifolds (online).
1239: 
1240: \bibitem[11]{Ring} Ringström, Hans On curvature decay in expanding cosmological models.  
1241: {\it Comm. Math. Phys.  264  (2006),  no. 3, 613--630.}
1242: 
1243: \bibitem[12]{Ham} Hamilton, Richard S. Non-singular solutions of the Ricci flow on three-manifolds.  
1244: {\it Comm. Anal. Geom.  7  (1999),  no. 4, 695--729.} in the book: Collected papers on Ricci flow.
1245: Edited by H. D. Cao, B. Chow, S. C. Chu and S. T. Yau. Series in Geometry and Topology, 37. International 
1246: Press, Somerville, MA, 2003.
1247: 
1248: \bibitem[13]{Pet} Petersen, Peter Riemannian geometry. Second edition. 
1249: {\it Graduate Texts in Mathematics, 171. Springer, New York, 2006.}
1250: 
1251: 
1252: \bibitem[14]{Elliptic} Gilbarg, David; Trudinger, Neil S. Elliptic partial differential equations of second order. 
1253: {\it Reprint of the 1998 edition. Classics in Mathematics. Springer-Verlag, Berlin, 2001.}
1254: 
1255: \bibitem[15]{B2}  Buchert, Thomas. Backreaction Issues in Relativistic Cosmology and the Dark Energy Debate. 
1256: {\it Preprint: arXiv:gr-qc/0612166v2.}
1257: 
1258: \bibitem[16]{Re}Rendall, Alan D. Constant mean curvature foliations in cosmological 
1259: space-times. {\it Journées Relativistes 96, Part II (Ascona, 1996). Helv. Phys. Acta 69 (1996), 
1260: no. 4, 490--500.}
1261: 
1262: \end{thebibliography}
1263: 
1264: \end{document}
1265: