math0111165/hv.tex
1: \documentclass[12pt]{amsart}
2: \usepackage{amsfonts}
3: \usepackage{epsfig,graphics}
4: \usepackage{amssymb}
5: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
6: \usepackage{amscd}
7: 
8: %TCIDATA{Created=Fri Mar 29 17:07:40 2002}
9: %TCIDATA{LastRevised=Sat Mar 30 20:12:35 2002}
10: 
11: \newtheorem{proposition}{Proposition}[section]
12: \newtheorem{theorem}{Theorem}[section]
13: \newtheorem{lemma}{Lemma}[section]
14: \newtheorem{corollary}{Corollary}[section]
15: \newtheorem{remark}{Remark}[section]
16: \newtheorem{exercise}{Notations}[section]
17: \newtheorem{definition}{Definition}[section]
18: \newtheorem{conjecture}{Conjecture}[section]
19: \newtheorem{example}{Example}[section]
20: \newtheorem{claim}{Claim}[section]
21: 
22: \addtolength{\hoffset}{-0.5cm}
23: \addtolength{\textwidth}{1cm}
24: 
25: \newcommand{\BZ}{{\mathbb{Z}}}
26: \newcommand{\BN}{{\mathbb{N}}} 
27: \newcommand{\BR}{{\mathbb{R}}}
28: \newcommand{\BC}{{\mathbb{C}}}
29: \newcommand{\BF}{{\mathbb{F}}}
30: \newcommand{\BQ}{{\mathbb{Q}}}
31: \newcommand{\BP}{{\mathbb{P}}}
32: \newcommand{\BA}{{\mathbb{A}}}
33: \newcommand{\BG}{{\mathbb{G}}}
34: \newcommand{\BT}{{\mathbb{T}}}
35: \newcommand{\BH}{{\mathbb{H}}}
36: 
37: \newcommand{\gD}{\Delta}
38: \newcommand{\gd}{\delta}
39: \newcommand{\gb}{\beta}
40: \newcommand{\gi}{\iota}
41: \newcommand{\gC}{\Gamma}
42: \newcommand{\gc}{\gamma}
43: \newcommand{\gs}{\sigma}
44: \newcommand{\gS}{\Sigma}
45: \newcommand{\gO}{\Omega}
46: \newcommand{\gep}{\epsilon}
47: \newcommand{\gl}{\lambda}
48: \newcommand{\ga}{\alpha}
49: \newcommand{\gt}{\tau}
50: \newcommand{\gz}{\zeta}
51: \newcommand{\gf}{\phi}
52: \newcommand{\gF}{\Phi}
53: \newcommand{\gth}{\theta}
54: \newcommand{\gTh}{\Theta}
55: \newcommand{\gT}{\Theta}
56: \newcommand{\fT}{{\frak{T}}}
57: \newcommand{\fH}{{\frak{H}}}
58: \newcommand{\ff}{{\frak{f}}}
59: \newcommand{\fg}{{\frak{g}}}
60: \newcommand{\fM}{{\frak{M}}}
61: \newcommand{\fO}{{\frak{O}}}
62: 
63: \def\a{\aleph}
64: \def\w{\omega}
65: \def\k{\kappa}
66: \def\De{\Delta}
67: \def\bbr{{\Bbb R}}
68: \def\vp{\varphi}
69: \def\e{\varepsilon}
70: 
71: \newcommand{\ti}[1]{\tilde{#1}}
72: \newcommand{\p}{\prod}
73: \newcommand{\ol}[1]{\overline{#1}}
74: \newcommand{\Pic}{\text{Pic}}
75: \newcommand{\jac}{\text{Jac}}
76: \newcommand{\prym}{\text{Prym}}
77: \newcommand{\gal}{\text{Gal}}
78: \newcommand{\Sym}{\text{Sym}}
79: \newcommand{\NS}{\text{NS}}
80: \newcommand{\dg}{\text{deg}}
81: \newcommand{\aut}{\text{Aut}}
82: \newcommand{\dm}{\text{dim}}
83: \newcommand{\norm}{\text{Norm}}
84: \newcommand{\odd}{\text{odd}}
85: \newcommand{\even}{\text{even}}
86: \newcommand{\ver}{\text{Ver}}
87: 
88: \newcommand{\rk}{\text{rank}}
89: \newcommand{\rank}{\text{rank}}
90: \newcommand{\vol}{\text{vol}}
91: \newcommand{\diam}{\text{diam}}
92: \newcommand{\diag}{\text{diag}}
93: 
94: \newcommand{\SL}{\text{SL}}
95: \newcommand{\GL}{\text{GL}}
96: \newcommand{\PSL}{\text{PSL}}
97: \newcommand{\PGL}{\text{PGL}}
98: \newcommand{\SO}{\text{SO}}
99: \newcommand{\PSO}{\text{PSO}}
100:    
101: \def\un{\underbar}
102: \def\uhr{\upharpoonright}
103: \def\dist{\text{dist}}
104: \def\Aut{\text{Aut}}
105: \def\N{{\cal N}}
106: 
107: \newcommand{\Cy}[1]{\mathbb{Z}/#1\mathbb{Z}}
108: \newcommand{\hra}{\hookrightarrow}
109: \newcommand{\im}[1]{\text{Im}(#1)}
110: \newcommand{\id}[1]{Id_{#1}}
111: \newcommand{\g}[1]{Gal(\ol{#1}/#1)}
112: \newcommand{\n}[2]{Norm_{#1/#2}}
113: \newcommand{\ord}[2]{\text{ord}_{(\Cy{#1})^{*}}#2}
114: \newcommand{\np}[2]{||#1||_{ #2}}
115: \newcommand{\h}[1]{H^{#1}}
116: \newcommand{\res}{\text{res}}
117: \newcommand{\df}{\stackrel{\text{def}}{=}}
118: \newcommand{\conan}{\ol{\ol{Z}}}
119: 
120: \newtheorem{prop}{Proposition}[section]
121: \newtheorem{thm}[prop]{Theorem}
122: \newtheorem{lem}[prop]{Lemma}
123: \newtheorem{cor}[prop]{Corollary}
124: \newtheorem{conj}[prop]{Conjecture}
125: \theoremstyle{definition}
126: \newtheorem{nota}[prop]{Notations}
127: \newtheorem{defn}[prop]{Definition}
128: \newtheorem{rem}[prop]{Remark}
129: \newtheorem{exam}[prop]{Example}
130: \newtheorem{prob}[prop]{Problem}
131: \newtheorem{ques}[prop]{Question}
132: \newtheorem{ob}[prop]{Observation}
133: \newtheorem{clm}[prop]{Claim}
134: \newtheorem{Ack}[prop]{Acknowledgments}
135: 
136: %\input tcilatex
137: 
138: \begin{document}
139: \author{Tsachik Gelander}
140: \address{Tsachik Gelander, Institute of Mathematics, Hebrew university, Jerusalem, Israel}
141: \email{tsachik@math.huji.ac.il}
142: %\date{\today}
143: \thanks{Research partially supported by the Clore and the Marie Curie 
144: fellowships.}
145: 
146: \title{Homotopy type and volume of locally symmetric manifolds}
147: \maketitle
148: 
149: 
150: 
151: 
152: 
153: \begin{center}
154: \epsfig{file=b.eps,height=0.5cm}
155: \end{center}
156: 
157: 
158: 
159: 
160: 
161: 
162: \begin{abstract}
163: We consider locally symmetric manifolds with a fixed universal 
164: covering, and construct for each such manifold $M$ a simplicial complex 
165: $\mathcal{R}$ whose size is proportional to the volume of $M$. 
166: When $M$ is non-compact, $\mathcal{R}$ is homotopically equivalent to $M$, 
167: while when $M$ is compact, $\mathcal{R}$ is homotopically equivalent to 
168: $M\setminus N$, where $N$ is a finite union of submanifolds of fairly 
169: smaller dimension.
170: This reflects how the volume controls the topological structure of $M$, and 
171: yields concrete bounds for various finiteness statements which previously had 
172: no quantitative proofs. 
173: For example, it gives an explicit upper bound for the 
174: possible number of locally symmetric manifolds of volume bounded by $v>0$,
175: and it yields an estimate for the size of a minimal presentation for the 
176: fundamental group of a manifold in terms of its volume.
177: It also yields a number of new finiteness results.
178: \end{abstract}
179: 
180: 
181: 
182: 
183: \setcounter{tocdepth}{1}
184: \tableofcontents
185: 
186: 
187: 
188: 
189: %NEW SECTION
190: 
191: 
192: 
193: 
194: 
195: \section{Introduction and statements of the main results}
196: 
197: In this article we study relations between the volume and the topological 
198: structure of locally symmetric manifolds. 
199: We are interested in asymptotic properties, 
200: when the volume tends to infinity. 
201: 
202: \medskip
203: 
204: We shall always fix a symmetric space, $S$, of non-compact type without 
205: Euclidean de-Rham factors, and consider the class of $S$-manifolds, 
206: by which we mean complete Riemannian manifolds locally isometric to $S$, 
207: or equivalently, manifolds of the form
208: $M=\gC\backslash S$ where $\gC$ is a discrete torsion-free group of isometries of $S$.
209: Sometimes we shall restrict our attention to arithmetic manifolds,
210: i.e. to Riemannian manifolds of the form 
211: $M=\gC\backslash S$, where $\gC\leq\textrm{Isom}(S)$ is a torsion-free arithmetic lattice
212: in the Lie group $\textrm{Isom}(S)$ of isometries of $S$.
213: A theorem of Borel and Harish-Chandra \cite{BoHC} says that if $M$ is arithmetic then 
214: $\vol (M)<\infty$. When $\rk (S)\geq 2$, Margulis' arithmeticity theorem \cite{Mar2}
215: gives the converse, i.e. $\vol (M)<\infty$ iff $M$ is arithmetic\footnote{The equivalence between finite
216: volume and arithmeticity is also known for the rank-1 cases $\text{Sp}(n,1)$ and $\text{F}_{4}^{-20}$ 
217: by \cite{GS}, \cite{cor}}. 
218: 
219: The ``complexity'' of the topology of locally symmetric manifolds is
220: controlled by the volume. This is illustrated by a theorem of 
221: Gromov (see \cite{BGS} theorem 2) which asserts  
222: that the Betti numbers are bounded by a constant times the volume, i.e. 
223: $$
224: \sum_{i=1}^{n}b_i(M)\leq c(S)\cdot\vol (M)
225: $$ 
226: for any $S$-manifold $M$.
227: Gromov's theorem applies also to non-locally symmetric manifolds, under appropriate conditions.
228: 
229: We conjecture (and prove in many cases) that, for locally symmetric manifolds,
230: the volume forces stronger topological restrictions. 
231: 
232: \begin{defn}\label{(d,v)-s.c.}
233: A {\bf $(d,v)$-simplicial complex} is a simplicial complex with 
234: at most $v$ vertices, all of them of valence $\leq d$.
235: \end{defn}
236: 
237: \begin{rem}
238: The number of $k$-simplexes in a $(d,v)$-simplicial complex $\mathcal{R}$ is 
239: $\leq\frac{v}{k+1}{n\choose k}$. Thus, the size of $\mathcal{R}$ (the number of its simplexes)
240: is at most $v\cdot\sum_{k=0}^d\frac{1}{k+1}{n\choose k}$, and this depends linearly on 
241: $v$.
242: \end{rem}          
243: 
244: \begin{conj}\label{conjA}
245: For any symmetric space of non-compact type $S$,
246: there are constants $\ga (S),d(S)$, such that any irreducible
247: $S$-manifold $M=\gC\backslash S$ (which is assumed also to be arithmetic in the case
248: $\dim (S)=3$) is homotopically equivalent to a 
249: $\big( d(S),\ga (S)\vol (M)\big)$-simplicial complex.
250: \end{conj}
251: 
252: \begin{rem}
253: The analogous statement is false for non-arithmetic manifolds in dimension $3$.
254: \end{rem}
255: 
256: In this paper we shall establish the following partial answers to Conjecture \ref{conjA}.
257: Since we failed in proving it in full generality, our results cannot be organized in a compact form, and
258: we have to split our statements and formulate strong results under appropriate conditions, and 
259: weaker results for more general cases.
260: 
261: \begin{thm}\label{thmA}
262: Let $S$ be a symmetric space of non-compact type. Then:
263: \begin{enumerate}
264: \item
265: Conjecture \ref{conjA} holds for non-compact arithmetic $S$-manifolds.
266: \item
267: If $S$ is neither isometric to $\SL_3(\BR )/\SO_3(\BR ),~\BH^2\times\BH^2$ nor to
268: $\BH^3$, then for some constants $\ga (S),d(S)$, the fundamental
269: group $\pi_1(M)$ of any 
270: $S$-manifold $M$ is isomorphic to the fundamental group of some
271: $\big( d(S),\ga (S)\vol (M)\big)$-simplicial complex.
272: \item
273: If $S$ is isometric to $\SL_3(\BR )/\SO_3(\BR ),~\BH^2\times\BH^2$ or to
274: $\BH^3$, then the fundamental group $\pi_1(M)$ of any  
275: $S$-manifold $M$ is a quotient of the fundamental group of some
276: $\big( d(S),\ga (S)\vol (M)\big)$-simplicial complex.
277: \end{enumerate}
278: \end{thm}
279: 
280: \begin{rem}
281: For compact arithmetic manifolds, Conjecture \ref{conjA} would follow, 
282: by a straightforward argument, if one could prove that the infimum of the
283: lengths of closed geodesics, taken over all compact arithmetic $S$-manifolds, 
284: is strictly positive. This conjectured phenomenon is strongly related to some 
285: properties of algebraic integers, such as the Lehmer conjecture,
286: which are still a mystery.
287: \end{rem}
288: 
289: Conjecture \ref{conjA} and Theorem \ref{thmA} yield
290: quantitative versions for some classical finiteness 
291: statements. We shall describe our main two applications 
292: in paragraphs \ref{BMP} and \ref{QVWT}. 
293: 
294: 
295: 
296: 
297: 
298: 
299: %subsection
300: 
301: \subsection{A linear bound on the size of a minimal presentation}\label{BMP}
302: 
303: \medskip
304: 
305: It is well known that the fundamental group of a locally symmetric manifold
306: with finite volume, or more generally a lattice in a connected semisimple Lie
307: group, is finitely presented. 
308: The compact case is quite standard, but the non-compact case was proved, step by step, by several authors
309: over several years.
310: Garland and Raghunathan \cite{Ga-Ra} proved it in the rank one case.
311: In the higher rank case, the finite generation was proved by Kazhdan \cite{Kazhdan} 
312: by defining and proving property-$T$ when $\rank (G)>2$,
313: and then by S.P. Wang \cite{S-P.W} when $\rank (G)=2$.
314: Using the finite generation, Margulis \cite{Mar2} proved the classical arithmeticity 
315: theorem for higher rank irreducible lattices\footnote{Margulis proved arithmeticity for higher rank 
316: non-uniform lattices \cite{Mar3}
317: few years before he proved the general case and without using his supper rigidity theorem}.
318: For arithmetic groups the finite presentability follows from the reduction 
319: theory of Borel and Harish-Chandra \cite{BoHC}.  
320: Later, using Morse theory, Gromov (\cite{BGS} theorem 2) gave a geometric proof of the finite 
321: presentability by proving that any locally 
322: symmetric manifold is diffeomorphic to the interior of a compact manifold with 
323: boundary. We remark that a slight modification of the argument of section \ref{6} below yields 
324: a completely elementary proof of this result of Gromov,
325: and hence of finite presentability (see also Remark \ref{finite-presentability}). 
326: However, in order to get a concrete estimate on the minimal possible size of such presentation 
327: in terms of the volume, we shall use arithmeticity (in the higher rank non-compact case).
328: We obtain the following quantitative version of finite presentability:
329: 
330: \begin{thm}\label{thmP}
331: Assume that $S$ is neither isometric to $\BH^3$,
332: $\SL_3(\BR )/\SO_3(\BR )$ nor to $\BH^2\times\BH^2$.
333: Then there is a constant $\eta =\eta (S)$ such that for any irreducible
334: $S$-manifold $M$, the fundamental group $\pi_1(M)$ admits a presentation
335: $$
336:  \pi_1(M)\cong\langle\Sigma :W\rangle
337: $$ 
338: with both $|\Sigma |,|W|\leq\eta\cdot\vol (M)$, in which all the relations 
339: $w\in W$ are of length $\leq 3$. 
340: \end{thm}
341: 
342: \begin{rem}
343: We believe (see also Conjecture \ref{conjA}) that the assumptions that $S$ is not
344: isometric to $\SL_3(\BR )/\SO_3(\BR )$ and to $\BH^2\times\BH^2$, are not really necessary.
345: However, our proof does not work in these cases.
346: It follows, however, from Theorem \ref{thmA}(2) that the analogous statements for non-compact 
347: $S$-manifolds hold also in these cases.
348: \end{rem}
349: 
350: \begin{rem}
351: The analogous statement for non-arithmetic hyperbolic $3$-manifolds is evidently false. 
352: However, if Conjecture \ref{conjA} is true, then the analogue of Theorem \ref{thmP} should
353: hold for arithmetic $3$-manifolds. By Theorem \ref{thmA}, it holds for non-compact arithmetic 
354: $3$-manifolds. Moreover, It was shown in \cite{cooper} that for any hyperbolic $3$-manifold $M$, 
355: the sum of the relations length, in any presentation of $\pi_1(M)$, is at 
356: least $\frac{\vol (M)}{\pi}$. This means that our upper bound is tight in 
357: this case.
358: \end{rem}
359: 
360: In the general case, Theorem \ref{thmA}(3) implies the following (weaker)
361: statement for which even the finiteness is in some sense surprising (in dimension $3$).
362: For a group $\gC$ let $d(\gC )$ denotes the minimal size of a generating set.
363: 
364: \begin{thm}\label{d(gC)}
365: For any $S$, there is a constant $\eta$ (depending on $S$), 
366: such that for any $S$-manifold $M$, $d\big(\pi_1(M)\big)\leq \eta\cdot\vol (M)$.
367: In other words, for any $v>0$, 
368: $$
369:  \sup\{d\big(\pi_1(M)\big) :\vol (M)\leq v\}\leq \eta v.
370: $$
371: Moreover, if $S$ is not isomorphic to $\SL_3(\BR )/\SO_3(\BR )$ then there is
372: a presentation
373: $$
374: \pi_1(M)\cong\langle\Sigma :W\rangle
375: $$ 
376: with $|\Sigma |,|W|\leq\eta\cdot\vol (M)$.
377: \end{thm} 
378: 
379: Note that Theorem \ref{d(gC)} does not give a bound on the length of the relations $w\in W$.
380: 
381: 
382: 
383: 
384: 
385: 
386: 
387: %subsection
388: 
389: \subsection{A quantitative version of Wang's theorem}\label{QVWT}
390: 
391: \medskip
392: 
393: We apply our results in order to estimate the number of $S$-manifolds 
394: with bounded volume (or more generally the number of conjugacy classes of
395: lattices in $G=\text{Isom}(S)$). 
396: This can be considered as a continuous analogue to
397: asymptotic group theory which studies the subgroup growth
398: of discrete groups (where ``covolume of lattices'' extends the notion ``index of subgroups''). 
399: Unlike the situation in the discrete (finitely generated) case, even the 
400: finiteness statements are not clear, and in general not true. 
401: However, a classical theorem of H.C. Wang states that if $S$ is not isometric 
402: to one of the hyperbolic spaces $\BH^2,\BH^3$, then for any $v>0$ there 
403: are only finitely many irreducible $S$-manifolds with total volume $\leq v$, 
404: up to isometries (see \cite{Wa} 8.1, and paragraph \ref{SL_2} below). 
405: We remark that Wang's result and proof do not give explicit estimates.
406: 
407: Denote by $\rho_S(v)$ the number of non-isometric irreducible $S$-manifolds
408: with volume $\leq v$.
409: 
410: By Mostow's rigidity theorem, a locally symmetric manifold of dimension 
411: $\geq 3$ is determined by its fundamental group. Applying 
412: Theorem \ref{thmA}(2), we obtain: 
413: 
414: \begin{thm}
415: If $\dim (S)\geq 4$ and $S$ is neither isometric to $\SL_3(\BR )/\SO_3(\BR )$ 
416: nor to $\BH^2\times\BH^2$ then there is a constant $c$, depending on $S$, with
417: respect to which
418: $$
419: \log\rho_S(v)\leq c\cdot v\log v
420: $$ 
421: for any $v>0$.
422: \end{thm}
423: 
424: This upper bound was first proved for hyperbolic manifolds by M. Burger, A. 
425: Lubotzky, S. Mozes and the author in \cite{BGLM}, where also a lower bound of
426: the same type was established, proving that this estimate is the true 
427: asymptotic behavior in the hyperbolic case. 
428: 
429: \begin{thm}[BGLM]
430: For $n\geq 4$, there are constants $c_n>b_n>0$ and $v_n>0$ such that
431: $$
432:  b_nv\log v\leq \log\rho_{\BH^n}(v)\leq c_nv\log v,
433: $$
434: whenever $v>v_n$.
435: \end{thm}
436: 
437: However, we suspect that in the higher rank case, where all manifolds are
438: arithmetic and conjectured to possess the congruence subgroup property, 
439: this upper bound is far from the true asymptotic behavior.  
440: It seems, in view of \cite{Lub}, that the problem of determining the real
441: asymptotic behavior is closely related to the congruence subgroup problem.
442: It is also related to the analysis of the Galois cohomology of compact 
443: extensions of $G=\text{Isom}(S)$.
444: 
445: For hyperbolic manifolds, the weaker upper bound  
446: $\rho_{\BH^n}(v)\leq v\exp (\exp (\exp (v+n)))$ was proved previously by Gromov \cite{Gr1}.
447: In this paper we establish a first concrete estimate for $\rho_S(v)$ for general $S$.
448: 
449: In dimension $2$ and $3$ the analogue of Wang's finiteness theorem is false.
450: However, as was shown by Borel \cite{Bor}, it remains true when 
451: considering only arithmetic manifolds.
452: The following estimate for the number of non-compact arithmetic $3$-manifolds
453: follows from Theorem \ref{thmA}(1):
454: 
455: \begin{prop}\label{116}
456: For some constant $c>0$, there are at most $v^{cv}$ non-isometric arithmetic 
457: non-compact hyperbolic 3-manifolds with volume $\leq v$.
458: \end{prop}
459: 
460: For compact arithmetic $3$-manifolds we conjecture that the analogous 
461: statement holds, but prove only the following weaker statement:
462: 
463: 
464: \begin{prop}\label{117}
465: Let $M$ be a compact (arithmetic) hyperbolic 3-manifold.
466: Then for some constant $c(M)$, the number of non-isometric hyperbolic
467: manifolds commensurable to $M$, with volume $\leq v$, is at most $v^{c(M)v}$.
468: \end{prop}
469: 
470: We remark that the analogue of Propositions \ref{116} and \ref{117} hold also in 
471: the cases $S=\SL_3(\BR )/\SO_3(\BR )$ and $S=\BH^2\times\BH^2$.
472: 
473: \medskip
474: 
475: In section \ref{complements}, we shall generalize some of our results concerning $S$-manifolds
476: to the larger family of $S$-orbifolds. We shall also indicate how to construct triangulations
477: for rank-1 manifolds which are not necessarily arithmetic.
478: 
479: \medskip
480: 
481: Let us now give a short and not precise explanation of the basic lines of the proofs of our main 
482: results \ref{thmA}.
483: The idea is to construct, inside each $S$-manifold $M$, a submanifold with boundary $M'$, which is
484: similar enough to $M$ and for which we can construct a triangulation of size $\leq c\cdot\vol (M)$.
485: In order to construct such a triangulation we shall require a lower bound on the injectivity radius
486: of $M'$ (independent of $M$) and some bounds on the geometry of the boundary $\partial M'$.
487: We shall construct $M'$ inside the $\gep_s$-thick part, in a way that its pre-image in the 
488: universal cover $S$ will be the complement of some ``locally finite'' union of convex sets, each
489: has a smooth boundary whose curvature is bounded uniformly from below, and the angles at the 
490: corners of $\partial M'$ (where the boundaries of two or more such convex sets meet) are bounded 
491: uniformly from below.
492: 
493: The non-compact arithmetic case is easier to deal with. The reason is that any non-uniform 
494: arithmetic lattice $\gC\leq G=\text{Isom}(S)^0$ comes from a rational structure on $G$ (rather then
495: on a compact extension of $G$ as the case might be when $\gC$ is uniform). 
496: This implies that there is 
497: some constant $\gep_s'$ such that any non-uniform arithmetic $S$-manifold contains no closed 
498: geodesics of length $\leq\gep_s'$ (see section \ref{AML}). In other words, any non-trivial closed 
499: loop which is short enough corresponds to a unipotent element in the fundamental group.
500: We define (the pre-image in $S$ of) $M'$ to be the complement of the union of appropriate
501: sub-level sets for the displacement functions $\{ d_\gc\}$ where $\gc$ runs over all unipotents
502: in $\pi_1(M)$. The injectivity radius in $M'$ is then uniformly bounded from below,
503: and using the fact that unipotents acts nicely on $S$ and on $S(\infty )$ we can 
504: both estimate the 
505: geometry of $\partial M'$, and also construct a deformation retract from $M$ to $M'$.
506: 
507: The situation is more subtle in the compact case. We do not have sufficient information
508: on the thick-thin decomposition. 
509: In this case we construct $M'$ and prove that it is diffeomorphic to $M\setminus N$ where $N$ is a 
510: finite union of submanifolds of codimension $\geq 3$. $N$ will contain the union of all 
511: closed geodesics of length smaller than some fixed constant. 
512: The idea is that at any point outside $N$
513: there is a preferred direction, such that when we move along it, the injectivity radius is
514: increasing most rapidly. This will help us to define a deformation retract from $M\setminus N$ to 
515: $M'$. As $N$ has codimension $\geq 3$, $\pi_1(M)\cong\pi_1(M\setminus N)\cong\pi_1(M ')$.
516: A main difficulty (which arises also in the compact rank one case) is how to control the geometry 
517: of the boundary of $M'$. We shall handle this difficulty in section \ref{bigangles}, where the 
518: main idea is Lemma \ref{bigA} which says that if two isometries commute then the exterior angle 
519: between their sub-level sets is $\geq\frac{\pi}{2}$.
520: 
521: \medskip
522: 
523: The present paper generalizes the work \cite{BGLM} which treated the special case of real 
524: hyperbolic spaces. Although some of the ideas from \cite{BGLM} appear again in the present paper, 
525: the situation in the general case is significantly more complicated than the hyperbolic case. 
526: The proof in \cite{BGLM} uses the explicit description of a hyperbolic compact 
527: thin component of the thick-thin decomposition as a cone over a coaxial Euclidean ellipsoids, 
528: as well as some computations in constant curvature. Hence the
529: argument in \cite{BGLM} does not apply to more general rank-1 symmetric spaces.
530: More crucially, in contrast to the situation in the rank-1 case, in the higher rank case
531: there is currently no good understanding of the structure of the thin components in the 
532: thick-thin decomposition, and hence new ingredients are required also in the skeleton of the proof.
533: 
534: Most of this paper is devoted to the treatment of the higher rank case. 
535: However, whenever it is not required, we shall not make
536: any assumption on the rank. We remark also that many of the arguments in this paper stay valid for
537: general manifolds of non positive curvature. In particular, some parts of this work could be 
538: generalized to non symmetric Hadamard spaces.
539: 
540: 
541: 
542: 
543: 
544: 
545: 
546: 
547: 
548: %NEW SECTION
549: 
550: 
551: 
552: 
553: \section{Notations, definitions and background}
554: In this section we shall fix our notations and
555: summarize some basic facts about semisimple Lie groups, symmetric spaces of 
556: non-compact type, and manifolds of non-positive curvature. For a comprehensive
557: treatment of these subjects we refer the reader to \cite{Rag}, \cite{BGS} and \cite{BH}.  
558: 
559: Let $S$ be a symmetric space of non-compact type. We shall always assume, 
560: that $S$ has no Euclidean 
561: de Rham factors. Let $G=\textrm{Isom}(S)$ be the Lie group of isometries of $S$. $G$ is 
562: center-free, semi-simple without compact factors, and with finitely many connected components. 
563: We denote by $G^0$ the identity component of $G$ with respect to the real 
564: topology. There is an algebraic group $\Bbb{G}$ defined over $\BQ$, such that
565: $G^0$ coincides with the connected component of the group of real points $\Bbb{G}(\BR )^0$.
566: In particular $G^0$ admits, apart from the real topology, a Zariski topology
567: which is defined as the trace in $G^0$ of the Zariski topology in $\Bbb{G}$.
568: $G^0$ acts transitively on $S$, and  we can identify $S$ with $G^0/K$ where $K\leq G^0$ is
569: a maximal compact subgroup. We remark that there is a bijection between
570: symmetric spaces of non-compact type and connected center free semisimple Lie 
571: groups without compact factors.
572: 
573: $S$ is a Riemannian manifold with non-positive
574: curvature such that for each point $p\in S$ there is an isometry $\sigma_p$
575: of $S$ which stabilizes $p$ and whose differential $d_p(\sigma_p)$ at $p$ is $-1$. 
576: The composition of two such isometries $\sigma_p\cdot \sigma_q$ is called a 
577: {\bf transvection}, and is belong to $G^0$.
578: The non-positivity of the sectional curvature means that the distance 
579: function $d:S\times S\to \BR ^+$ is convex and, in fact, its restriction to 
580: a geodesic line $c(t)=\big( c_1(t),c_2(t)\big)$ in $S\times S$ is strictly 
581: convex, unless $c_1$ and $c_2$ are contained in a flat plane in $S$.
582: 
583: For a real valued function $f:X\to\BR$ on a set $X$ we denote by
584: $\{ f<t\}$ the $t$-sub-level set
585: $$
586:  \{ f<t\}=\{ x\in X:f(x)<t\}.
587: $$
588: 
589: For $\gc \in G$ we denote by $d_{\gc}$ the {\bf displacement function}
590: $$
591:   d_{\gc}(x)=d(\gc \cdot x,x).
592: $$
593: This function is convex and smooth outside Fix$(\gc )$. 
594: In particular the sub-level sets $\{ d_{\gc} <t\}$ are convex with smooth boundary.
595: We denote by {\bf min}$(\gc )$ the set 
596: $$
597:  \min (\gc )=\{x\in S:d_{\gc}(x)=\inf d_{\gc}\}.
598: $$
599: 
600: For a set $A\subset S$ we denote by 
601: $$
602:  D_A(x)=d(A,x)
603: $$ 
604: the {\bf distance function}.
605: If $A$ is a convex set then the function $D_A$ is convex and smooth at any
606: point outside the boundary of $A$. We denote its $t$-sub-level set by
607: $$
608:  (A)_t=\{ D_A<t\} 
609: $$
610: and call it the $t$-{\bf neighborhood} of $A$. 
611: Note that $(A)_{t+s}=\big( (A)_t\big)_s$.
612: Similarly, we let $)A(_t$ denote the $t$-{\bf shrinking} of $A$
613: $$
614:  )A(_t=S\setminus (S\setminus A)_t.
615: $$
616: If $A$ is convex then $\big) (A)_t\big(_t= A$ but in general these are 
617: different sets. 
618: 
619: A subset $\mathcal{N}$ of a metric space $X$ is called an {\bf $\gep$-net}, if
620: $(\mathcal{N})_\gep =X$. A subset $\mathcal{N}\subset X$ is called 
621: {\bf $\gep$-discrete} is $d(x,y)>\gep$ for any $x,y\in\mathcal{N},~x\neq y$.
622: If $\mathcal{N}$ is a maximal $\gep$-discrete subset, then it is an $\gep$-net.
623: 
624: If $A$ is closed and convex then for any $x\in S$ there is a unique closest 
625: point $p_A(x)$ in $A$. The map $p_A:S\to A$ is called the {\bf projection} 
626: on $A$, and it is distance decreasing, i.e.
627: $d\big(p_A(x),p_A(y)\big)\leq d(x,y)$ for any $x,y\in S$.
628: 
629: A {\bf flat subspace} is a totally geodesic sub-manifold which is 
630: isometric to a Euclidean space. A {\bf flat} is a maximal flat subspace. 
631: Any flat subspace is contained in a flat. 
632: All flats in $S$ have the same dimension
633: $\textrm{rank}(S)=\textrm{rank}(G)$\footnote{By $\textrm{rank}(G)$ we shall 
634: always mean the real rank of $G$}. A geodesic $c\subset S$ is called 
635: {\bf regular} if it is contained in a unique flat.
636: 
637: An element $\gc \in G$ is unipotent if $\text{Ad}(\gc )-1$ is a nilpotent 
638: endomorphism. A subgroup of $G$ is called a {\bf unipotent subgroup} if all its 
639: elements are unipotent. If $H\leq G^0$ is a unipotent subgroup then its 
640: Zariski closure $\overline{H}^z\leq\Bbb{G}$ is also unipotent, and 
641: $\overline{H}^z_\BR$ is connected. Moreover, a unipotent subgroup $H\leq G^0$ 
642: is connected iff $H=\overline{H}^z_\BR$. 
643: 
644: $S(\infty )$ is the set of equivalence classes of geodesic rays, where two
645: rays are considered equivalent if their traces are of bounded Hausdorff 
646: distance from each other. 
647: There is a structure of a spherical building on $S(\infty )$ - the Tits 
648: building of $S$. The chambers of this building are sometimes called Weyl 
649: chambers. A maximal unipotent subgroup of $G^0$ is the unipotent radical of 
650: some minimal parabolic subgroup. The minimal parabolic subgroups of $G^0$ are 
651: the stabilizers (and the fixator) of the chambers of this building, and there
652: are canonical bijections between the set of minimal parabolic subgroups, the 
653: set of maximal unipotent subgroups and the set of chambers. All maximal 
654: unipotent subgroups are conjugate in $G^0$, or in other words, $G^0$ acts 
655: transitively on the set of chambers $W\subset S(\infty )$. Moreover, each 
656: parabolic subgroup acts transitively on $S$, and hence $G$
657: acts transitively on the set of couples $(W,x)$ of a chamber 
658: $W\subset S(\infty )$ and a point $x\in S$.
659: There is a canonical metric on $S(\infty )$, called the Tits metric, with 
660: respect to which the apartments of the Tits building are isometric to spheres
661: and the chambers are isometric to spherical simplices. The induced action of 
662: $G$ on $S(\infty )$ preserves the Tits metric.
663: A point $p\in S(\infty )$ is called {\bf regular} if it is an interior point
664: of a chamber. For any $p\in S(\infty )$ and $x\in S$ there is a unique geodesic
665: $c$ with $c(0)=x$ and $c(\infty )=p$, and $p$ is regular iff $c$ is regular.
666: 
667: When $\gC$ is a group of isometries of $S$, we say that a subset $A\subset S$ 
668: is $\gC$-{\bf precisely invariant} if $\gc \cdot A=A$
669: whenever $\gc \cdot A \cap A\neq \emptyset$.   
670: If $\gC$ acts freely and discretely (i.e. $\gC \subset G$ is discrete
671: and torsion free) then we denote by $\gC\backslash A$ the image of $A$ in $\gC\backslash S$.
672: If $A$ is a connected simply connected $\gC$-precisely invariant set
673: then 
674: $$
675:  \pi_1(\gC\backslash A )\cong \{ \gc \in \gC :
676:  \gc \cdot A=A\}. 
677: $$
678: 
679: For a subset $M'$ of $M=\gC\backslash S$ we let $\ti M'$ denote its pre-image in $S$, and
680: we shall usually denote by $\ti{M'}^0$ an arbitrary connected component of 
681: $\ti M'$.  
682: 
683: For a subgroup $\gC\leq G$, a constant $\gep >0$ and a point $x\in S$,  
684: we let $\gC_\gep (x)$ denote the group generated by the ``small elements'' 
685: at $x$
686: $$
687:  \gC_\gep (x)=\langle \gc\in\gC :d_\gc (x)\leq\gep\rangle.
688: $$
689: Recall the classical Margulis lemma:
690: 
691: \begin{thm}[{\bf The Margulis lemma}]
692: There are constants $\gep_s>0$ and $n_s\in\BN$, depending only on $S$,
693: such that for any discrete subgroup $\gC\leq G$, and any point $x\in S$,
694: the group $\gC_{\gep_s}(x)$ contains a subgroup 
695: of index $\leq n_s$ which is contained in a connected nilpotent subgroup of 
696: $G$.
697: \end{thm}
698: 
699: The {\bf $\gep$-thick thin} decomposition of an $S$-manifold $M=\gC\backslash S$ reads 
700: $$
701:  M=M_{\geq\gep}\cup M_{\leq\gep}.
702: $$
703: The {\bf $\gep$-thick part} $M_{\geq\gep}$ is defined as the set of all 
704: points $x\in M$ at which the injectivity radius is $\geq\frac{\gep}{2}$, 
705: and the {\bf $\gep$-thin part} $M_{\leq\gep}$
706: is the complement of the interior of the $\gep$-thick part.
707: Note that $M_{\leq\gep}$ is the set of points in $M$ through which there is a
708: non-contractible closed loop of length $\leq\gep$.
709: The pre-image of $M_{\leq\gep}$ in $S$ has a nice description as a 
710: locally finite union of convex sets
711: $$
712:  \ti M_{\leq\gep}=\cup_{\gc\in\gC\setminus\{ 1\}}\{ d_\gc\leq\gep\}.
713: $$
714: The Margulis lemma yields an important piece of information on the structure 
715: of the $\gep$-thick-thin decomposition when $\gep\leq\gep_s$.
716: 
717: If $\gC \subset G$ is a uniform lattice then any element 
718: $\gc \in \gC$ is semisimple, i.e. $\text{Ad}(\gc )$ is diagonalizable over 
719: $\BC$, or equivalently $\min (\gc )\neq \emptyset$. If, in addition, $\gC$ is 
720: torsion free then all its elements are hyperbolic.
721: A semisimple element is hyperbolic iff it has a complex eigenvalue outside the
722: unit disk.
723: A hyperbolic isometry $\gc$ has an axis, i.e. a geodesic line on which
724: $\gc$ acts by translation. Any two axes  are parallel, and 
725: $x\in \min (\gc )$ iff the geodesic line $\overline{x,\gc \cdot x}$
726: is an axis of $\gc$. 
727: We define a {\bf regular isometry} to be a hyperbolic isometry whose axes are regular 
728: geodesics (if one axis is a regular geodesic, then so are all of them, since axes are parallel). 
729: If $g\in G$ acts
730: as a regular isometry then $g$ is a regular element in $G$ in the usual sense,
731: but not necessarily vice versa.
732: If $\gc\in G$ acts as a regular isometry, and $F$ is the unique flat containing
733: an axis of $\gc$ then $\gc$ preserves $F$ and acts by translation on it, and
734: in particular $\min (\gc )=F$.
735: (This follows for example from the facts that the Weyl group 
736: $N_{G^0}(A)/A$, for $A=C_{G^0}(\gc )$, acts simply transitively on the set of Weyl chambers 
737: in $F(\infty )$, and that a
738: regular point $p\in F(\infty )$ determines its ambient Weyl chamber.)
739: 
740: If $A\subset S$ is a closed convex set, and $\ga\in G$ preserves $A$, i.e.
741: $\ga\cdot A=A$, then the projection $P_A:S\to A$ commutes with $\ga$, and 
742: since $P_A$ does not increase distances, we have
743: $d_\ga \big( P_A(x)\big)\leq d_\ga(x)$. In particular, $\ga$ is semisimple iff
744: $\min (\ga )\cap A\neq\emptyset$. Since for any isometry $i$ of a flat 
745: subspace $F$, $\min (i)\neq\emptyset$, it follows that if $\ga\in G$ preserves
746: a flat subspace then $\ga$ is semisimple. If $g\in G^0$ is semisimple, then
747: its centralizer group $C_{G^0}(g)$ is a closed reductive group.
748: If $g\in G$ is a transvection with axis $c$, then
749: $$
750:  C_{G^0}(g)=\{ h\in G^0:h\cdot c \text{ is parallel to } c\}.
751: $$
752:    
753: If $\gC \subset G$ is a non-uniform lattice, then it is almost generated by
754: unipotents. $\gc$ is unipotent iff $\text{inf}(d_\gc )=0$ while $d_\gc$ never 
755: vanishes on $S$. In particular a unipotent element stabilizes a point at infinity but not in $S$. 
756: 
757: 
758: If $f:G^0\to H$ is a surjective Lie homomorphism, and $g\in G$ is semisimple
759: (resp. unipotent) then $f(g)$ is semisimple (resp. unipotent).
760: 
761: 
762: 
763: 
764: 
765: 
766: 
767: %NEW SECTION
768: 
769: 
770: 
771: \section{Some deformation retracts}\label{3}
772: 
773: In several places in this paper we will need to produce a deformation 
774: retract from a submanifold (maybe with boundary) to a subset of it which
775: is defined by the condition that all the functions from some given family
776: are larger than a given constant. A typical example of such a situation is 
777: given by the manifold itself and the $\gep$-thick part.
778: The family of functions in this example is indexed by the fundamental group.
779: To each homotopy class of closed loops corresponds the function for which 
780: the value at any point  $x$ is the minimal length of a loop from the class, 
781: which passes through $x$. The $\gep$-thick part is exactly the subset where 
782: all these functions are $\geq\gep$. The Margulis lemma yields information 
783: about the set of functions which are $<\gep$ at $x$, which can sometimes be exploited in 
784: order to define a deformation retract to the $\gep$-thick part.
785: Another example is given by a subset $M'$
786: of the manifold $M$ and its $\gep$-shrinking $)M'(_\gep$. If the complement of $M'$ is 
787: given by the union of subsets satisfying certain properties,
788: it is sometimes convenient to use the distance functions from these subsets
789: in order to define a deformation retract to $)M'(_\gep$.
790: 
791: The idea is to construct a continuous vector
792: field which makes an acute angle at any point $x$ with the gradients of all the
793: functions from the given family which are small at $x$, and then to let points
794: flow along its integral curves.
795: This idea has been used in an elegant way in other places (c.f. \cite{BGS}). The situation concerned here, however,
796: is more subtle since we consider manifolds with boundary, and have to make sure that on the 
797: boundary, the vector field is pointing towards the interior.
798: 
799: \begin{defn}
800: For a family of real valued continuous functions 
801: $\mathcal{F} =\{\phi_i\}_{i\in I}$ on
802: a manifold $Y$ (with or without boundary) and $\gep >0$, we define the {\bf  
803: $(\mathcal{F},\gep )$-thick part} (or simply the {\bf $\mathcal{F}$-thick part) 
804: $\mathcal{F}_{\geq\gep}$} as follows
805: $$
806:  \mathcal{F}_{\geq\gep}=\cap_{\phi\in\mathcal{F}}\overline{\{\phi >\gep\}}.
807: $$
808: When $\mathcal{F}$ is given, we denote by 
809: $\Psi_{x,\gt }~(x\in Y,\gt\in\BR )$ the subset
810: $$
811:  \Psi_{x,\gt } =\{\phi\in\mathcal{F} : \phi (x)\leq\gt\}.
812: $$
813: Abusing notations, we shall sometimes write $\Psi_{x,\gt }$ for the 
814: corresponding set of indices
815: $\Psi_{x,\gt } =\{i\in I: \phi_i(x)\leq\gt\}$.
816: We say that the family $\mathcal{F}$ is {\bf locally finite} if
817: $\Psi_{x,\gt }$ is finite for any $x\in Y$ and $\gt\in\BR$. 
818: We say that $\gt\in\BR$ is a {\bf critical value} of a continuous function
819: $\phi$ if $\overline{\{\phi >\gt\}}\neq \{\phi\geq\gt\}$, i.e. if $\gt$ is 
820: a value of a local maximum of $\phi$.
821: \end{defn}
822: 
823: If $Z\subset Y$ is a submanifold with smooth boundary, and $\phi$ a real 
824: valued function on $Z$. The gradient $\nabla\phi$ is well defined on $Z$ 
825: wherever it is unique (on 
826: the boundary it is the tangent vector with length and direction equal to the 
827: value of the maximal directional 
828: derivative and the direction at which it occurs). The function $\phi$ is said 
829: to be smooth (or $C^1$) if its gradient $\nabla\phi$ is a continuous function on the 
830: whole of $Z$. When $\phi$ is smooth, its directional derivative with respect to
831: $v\in T_x(Y)$ is given by the usual formula $v\cdot\nabla\phi (x)$.
832:  
833: In later sections, we will consider families $\mathcal{F}$ which consist of functions of the 
834: following 2 types:
835: \begin{itemize}
836: \item 
837: The displacement function $d_\gc (x)=d(x,\gc\cdot x)$ associated with
838: a non-trivial isometry $\gc$ of $S$.
839: \item 
840: The distance function $D_A(x)=d(x,A)$ from a given closed convex set with
841: smooth boundary.
842: \end{itemize}
843: In both cases the function is convex, non-negative, and without positive 
844: critical values. The function $d_\gc$ is smooth on $\{ d_\gc >0\}$, and the 
845: function $D_A$ is smooth on $\overline{\{ D_A>0\}}$.
846: 
847: \medskip
848: 
849: %Let us now formulate a precise statement.
850: 
851: \begin{lem}[\bf A deformation retract which increases functions]
852: \label{deformation-retract}
853: Consider a submanifold without boundary $Y\subset S$.
854: \begin{itemize}
855: \item 
856: Let $\mathcal{F} =\{\phi_i\}_{i\in I}$ be a locally finite
857: family of non-negative continuous functions on $Y$.
858: \item
859: Assume that for each $\phi\in\mathcal{F}$ with $\{\phi =0\}\neq\emptyset$, the set 
860: $\overline{\{\phi >0\}}$ is a submanifold with smooth boundary.
861: \item
862: Assume that each $\phi\in\mathcal{F}$ is $C^1$ on $\overline{\{\phi >0\}}$.
863: \item 
864: Let $L=\mathcal{F}_{\geq 0}=\cap_{\phi\in\mathcal{F}}\overline{\{\phi >0\}}$.
865: \item
866: Let $\gb :[0,3\gep ]\to\BR^{>0}$ be a given continuous positive function 
867: (in case all $\phi\in\mathcal{F}$ are strictly positive we allow $\gb$ to
868: be defined only on $(0,3\gep ]$).
869: \item
870: Assume that $\gep$ is not a critical value for any $\phi\in\mathcal{F}$.
871: \item
872: Assume that for any $x\in Y$    
873: there is a unit tangent vector $\hat n(x)\in T_x(S)$ such that  
874: $$
875:  \hat n(x)\cdot \nabla \phi (x)\geq\gb\big( \phi (x)\big)
876: $$
877: for any $\phi\in\Psi_{x,3\gep}$. 
878: \end{itemize}
879: Then there is a deformation retract from $L$ to the $\mathcal{F}$-thick 
880: part $\mathcal{F}_{\geq\gep}$. 
881: (In particular it follows that $\mathcal{F}_{\geq\gep}\neq\emptyset$.)
882: 
883: If $\gC\leq\text{Isom}(S)$ is a discrete subgroup, and if $L$ and the family 
884: $\mathcal{F}$ are $\gC$-invariant, in the sense that $\gc\cdot L=L$, and
885: the function $\gc\cdot\phi (x):=\phi(\gc^{-1}\cdot x)$ belongs to $\mathcal{F}$ for any
886: $\phi\in\mathcal{F},~\gc\in\gC$,
887: then there exists such a deformation retract
888: which is $\gC$-invariant (in the obvious sense)
889: and hence induces a deformation retract between 
890: the corresponding subsets $\gC\backslash L$ and 
891: $\gC\backslash\mathcal{F}_{\geq\gep}$ of $M=\gC\backslash S$.
892: \end{lem}
893: 
894: \begin{proof}
895: We will define an appropriate continuous vector field on $L$. 
896: The desired deformation retract will be the flow along this vector field.
897: 
898: Let $\gd (x)$ denote 
899: $$
900:  \gd (x)=\min_{\phi\in \mathcal{F}}\phi (x).
901: $$
902: 
903: For any non-empty subset $\Psi\subset\mathcal{F}$ for which 
904: $\cap_{\phi\in\Psi}\{\phi\leq 3\gep\}\neq\emptyset$ and any point $x$ of this intersection, let 
905: $\hat f(x,\Psi )\in T_x(S)$ be a unit tangent vector which maximizes the 
906: expression
907: $$
908:  \min\{\hat f\cdot \nabla\phi (x):\phi\in\Psi\},
909: $$
910: and let 
911: $$
912:  \gb_\Psi (x)=\min_{\phi\in\Psi}\gb\big(\phi (x)\big).
913: $$
914: Then for any $\phi\in\Psi$
915: \begin{eqnarray*}
916:  \hat f(x,\Psi )\cdot \nabla \phi (x)\geq
917:  \min_{\phi\in\Psi}\hat f(x,\Psi )\cdot \nabla \phi (x)\geq
918:  \min_{\phi\in\Psi}\hat n(x)\cdot \nabla \phi (x)\geq
919:  \min_{\phi\in\Psi}\gb\big(\phi (x)\big)=
920:  \gb_\Psi (x).
921: \end{eqnarray*} 
922: Moreover, it follows from the strict convexity of the Euclidean unit disk that
923: $\hat f(x,\Psi )$ is uniquely determined, and consequently, that for a fixed
924: $\Psi$, the vector field $\hat f(x,\Psi )$ is continuous on
925: $\cap_{\phi\in\Psi}\{\phi\leq 3\gep\}$.
926: 
927: The desired vector field is defined as follows:
928: \begin{eqnarray*}
929:  &&\!\!\!\!\!\!\! \overrightarrow V(x)=
930:  \sqrt{2\big(\gep -\gd(x)\big)\lor 0}\cdot\\
931: &\phantom{\leq}&
932:  \cdot\sum_\Psi
933:  \frac{\big( 3\gep -\max_{\phi\in\Psi}\phi (x) \big)\lor 0}{\gep}\cdot
934:  \frac{\big(\min_{\phi\notin\Psi}\phi (x)-\gep \big)\lor 0}{\gep}\cdot
935:  \frac{1}{\gb_\Psi(x)}\hat f(x,\Psi ),
936: \end{eqnarray*}
937: where the sum is taken over all non-empty finite subsets 
938: $\Psi \subset\mathcal{F}$.
939: 
940: Notice that $\gb_\Psi(x)$ is strictly positive, and 
941: all the terms in each summand are continuous, and $\overrightarrow V\equiv 0$
942: on the $\mathcal{F}$-thick part
943: $$
944:  \mathcal{F}_{\geq\gep}=
945:  \cap_{\phi\in\mathcal{F}}\overline{\{\phi >\gep\}}=
946:  \cap_{\phi\in\mathcal{F}}\{\phi\geq\gep\}=
947:  \{\gd\geq\gep\}.
948: $$
949: The term $\sqrt{2\big(\gep -\gd(x)\big)\lor 0}$ takes care of the continuity
950: on the boundary $\{\gd =\gep\} =\partial\{\gd \leq\gep\}$ of the 
951: $\mathcal{F}$-thin part. 
952: The terms $\frac{\big( 3\gep -\max_{\phi\in\Psi}\phi (x) \big)\lor 0}{\gep}$ 
953: guarantee that all the non-zero summands correspond to sets which are 
954: contained in the finite set $\Psi_{x,3\gep}$. In particular the summation is 
955: finite for any $x\in L$, and $\hat f(x,\Psi )$ is well defined for any 
956: non-zero summand.
957: The terms $\frac{\big(\min_{\phi\notin\Psi}\phi (x)-\gep \big)\lor 0}{\gep}$ 
958: guarantee that all the non-zero summands correspond to $\Psi$'s which contains
959: $\Psi_{x,\gep}$.   
960: 
961: If $\gd (x)<\gep$ then $\Psi_{x,\gep}\neq\emptyset$, and, when looking only on the summand 
962: corresponding to $\Psi_{x,2\gep}$, we see that
963: $$
964:  \nabla \phi (x)\cdot\overrightarrow V(x)\geq\sqrt{2\big(\gep -\gd(x)\big)}
965: $$
966: for any $\phi\in\Psi_{x,\gep}\subset\Psi_{x,2\gep}$, and in particular for any 
967: $\phi$ with $\phi (x)=\gd (x)$.
968: 
969: It follows that if $x(t)$ is an integral curve of $\overrightarrow V$ with
970: $\gd\big( x(0)\big) <\gep$, then
971: $$
972:  \frac{d}{dt}\Big(\gd\big( x(t)\big)\Big)\geq 
973:  \sqrt{2\Big(\gep -\gd\big( x(t)\big)\Big)}.
974: $$
975: (To be more precise, since $\gd\big( x(t)\big)$ is not necessarily 
976: differentiable, we should write 
977: $\liminf_{\gt\to 0}\frac{\gd\big( x(t+\gt)\big)-\gd\big( x(t)\big)}{\gt}$ 
978: instead of $\frac{d}{dt}\Big(\gd\big( x(t)\big)\Big)$ in the last inequality.)
979: Thus, for $t=\sqrt{2\Big(\gep -\gd\big( x(0)\big)\Big)}$ we have
980: $\gd\big( x(t)\big) =\gep$. 
981: 
982: Since $x\in L$ belongs to $\partial L$ iff $\gd (x)=0$, and hence, 
983: the vector field 
984: $\overrightarrow V$ points everywhere towards the interior $\textrm{int}(L)$,
985: it follows from the Peano existence theorem of solution for ordinary 
986: differential equations, that for any $x\in L$ there is an integral curve 
987: $c_x(t)$ of $\overrightarrow V$, defined for all $t\geq 0$ with $c_x(0)=x$ and 
988: with $c_x(t)\in\textrm{int} L$ for $t>0$. Stability of the solutions implies
989: that $c_x(t)$ depends continuously on $x$ and $t\geq 0$.
990: 
991: As a conclusion, we get that the flow along 
992: $\overrightarrow V$ for
993: $\sqrt{2\gep}$ time units defines a deformation retract from $L$ to 
994: $\mathcal{F}_{\geq\gep}=\{\gd\geq\gep\}.$
995: 
996: If $L$ and $\mathcal{F}$ are
997: $\gC$-invariant, then $\overrightarrow V(x)$, as it is defined above, is also 
998: $\gC$-invariant. Hence it induces a vector field on $\gC\backslash L$, 
999: and a deformation retract from $\gC\backslash L$ to $\gC\backslash\mathcal{F}_{\geq\gep}$.
1000: \end{proof}
1001:  
1002: A second kind of deformation retracts that we shall often use is the following.
1003: Let $A\subset S$ be a closed convex set, and $B\subset S$ a set containing $A$.
1004: We say that $B$ is {\bf star-shaped} 
1005: with respect to $A$ if for any $b\in B$ the geodesic segment 
1006: connecting $b$ to its closest point (the projection) in $A$ is contained 
1007: in $B$. In that case we can define a deformation retract from $B$ to $A$ by 
1008: moving $b$ at a constant rate (equal to the initial distance) 
1009: towards its projection in $A$. We call this the {\bf star-contraction} 
1010: from $B$ to 
1011: $A$. From the convexity of the distance function we conclude that
1012: the star contraction from $B$ to $A$
1013: is distance decreasing, and hence continuous, and if 
1014: the Hausdorff distance $\textrm{Hd}(A,B)$ is finite, then the star-contraction 
1015: gives a homotopy equivalence between $(B,\partial B)$ and $(A,\partial A)$.
1016: 
1017: More generally, if $A\subset B_1\subset B_2\subset S$, and if the $B_i$'s are closed and
1018: star-shaped with respect to $A$, we can define a deformation retract from $B_2$
1019: to $B_1$ by letting any $b\in B_2\setminus B_1$ flow in the direction of
1020: its projection $P_A(b)$ in $A$ at constant speed $s$, where $s$ equals to
1021: the length of the segment $[b,P_A(b)]\cap (B_2\setminus B_1)$. 
1022: By a similar procedure one can show that there is a deformation retract from $S\setminus B_1$
1023: to $\overline{S\setminus B_2}$.
1024: In this way we obtain:
1025: 
1026: \begin{lem}[{\bf A generalized star-contraction}]\label{GSC}
1027: Assume that 
1028: \begin{itemize}
1029: \item
1030: $A$ is convex and closed,
1031: \item
1032: $A\subset B_1\subset B_2$ where $B_i$ are closed and star-shaped with respect 
1033: to $A$.
1034: \item
1035: $\textrm{Hd}(A,B_2)<\infty$.
1036: \end{itemize}
1037: Then there is a deformation retract from $(B_2,\partial B_2)$ to $(B_1,\partial B_1)$.
1038: Similarly, there is a deformation retract from $S\setminus B_1$
1039: to $\overline{S\setminus B_2}$.
1040: \end{lem}
1041: 
1042: 
1043: 
1044: 
1045: 
1046: 
1047: %NEW SECTION
1048: 
1049: 
1050: 
1051: 
1052: \section{Constructing a simplicial complex in a thick submanifold with 
1053: nice boundary}\label{4}
1054: 
1055: Throughout this section, $M=\gC\backslash S$ is a fixed $S$-manifold with finite volume,
1056: $M'\subset M$ a connected submanifold with boundary, and $\gep >0$ is fixed. 
1057: The main result of this section is Lemma \ref{simplicial-complex}. 
1058: 
1059: We wish to formulate some convenient conditions on $M'$, under which $M'$ is 
1060: homotopically equivalent to a simplicial complex $\mathcal{R}$ whose 
1061: combinatorics is bounded in terms of $\vol (M)$. 
1062: More precisely, we would like $\mathcal{R}$ to be 
1063: a $\big( d,\ga\cdot\vol (M')\big)$-simplicial complex, where $d$ and $\ga$
1064: are some constants depending only on $S$ and $\gep$ (see Definition \ref{(d,v)-s.c.}).
1065: 
1066: To construct $\mathcal{R}$ we will use a ``good covering'' argument.
1067: Recall that a cover of a topological space $T$ is called a {\bf good cover} 
1068: if any non empty intersection of sets of the cover is contractible.
1069: In this case the simplicial complex $\mathcal{R}$ which corresponds to
1070: the nerve of the cover is homotopically equivalent to $T$. By definition, the
1071: vertices of $\mathcal{R}$ correspond to the sets of the cover, and a 
1072: collection of vertices form a simplex when the intersection of the 
1073: corresponding sets is non-empty (see \cite{BT}, theorem 13.4). 
1074: 
1075: In a manifold with injectivity radius bounded uniformly 
1076: from below by $\epsilon$, such a good cover is achieved by taking 
1077: $\epsilon$-balls for which the set of centers form an
1078: $\epsilon /2$ (say) discrete net. In our case, in order to be able to use 
1079: $\gep$-balls (or more precisely $\gep /c$-balls for some constant $c$) 
1080: we shall require that $M'$ lies inside $M_{\geq\gep}$. 
1081: However, $M'$ is not a manifold but a manifold with boundary, and balls may 
1082: not be ``nice'' subsets - they may not be convex or even contractible, and an intersection of
1083: balls may not be connected.
1084: Therefore we should be more careful.
1085: Our problems arise only near the boundary (far away from 
1086: the boundary balls are nice). So we need some control on the geometry of the boundary.
1087: 
1088: \begin{lem}\label{simplicial-complex}
1089: Let $M=\gC\backslash S$ be a fixed $S$-manifold with finite volume,
1090: let $M'\subset M$ be a connected submanifold with boundary, 
1091: and let $\gep >0$ be fixed. 
1092: \begin{itemize}
1093: \item
1094: Assume that $M'$ is contained in the $\gep$-thick part $M_{\geq \gep}$.
1095: \item
1096: Write $X=M\setminus M'$, and assume that its pre-image $\tilde X\subset S$ 
1097: under the universal covering map is a locally finite union of convex open sets 
1098: with smooth boundary $\ti X=\cup_{i\in I}X_i$
1099: (by locally finite we mean that every compact set in $S$ intersects only finitely many $X_i$'s).
1100: \item
1101: Assume that $M'$ is homotopically equivalent to its $\frac{\gep}{2}$-shrinking 
1102: $)M'(_{\frac{\gep}{2}}$.
1103: \item
1104: Let $b>1$ be a constant which depends only on $S$ and on $\gep$.
1105: \item
1106: Assume that for any point $x\in S\setminus\ti X$ with $d(x,\ti X)\leq\gep$, 
1107: there is a unit tangent vector $\hat n(x)\in T_x(S)$,
1108: such that for each $i\in I$ with $d(x,X_i)=d(x,\ti X)$, 
1109: the inner product of $\hat n(x)$ with the gradient $\nabla D_{X_i}(x)$ satisfies 
1110: $$
1111:  \hat n(x)\cdot\nabla D_{X_i}(x)>\frac{1}{b}.
1112: $$
1113: \end{itemize}
1114: Then there are constants $\ga$ and $d$, depending only on $S$ and on $\gep$, 
1115: such that $M'$ is homotopically equivalent to some 
1116: $\big( d,\ga\cdot\vol (M')\big)$-simplicial complex.
1117: \end{lem}
1118: 
1119: Throughout this section we use the notation of the statement of 
1120: Lemma \ref{simplicial-complex}.\\
1121: 
1122: The proof of Lemma \ref{simplicial-complex} relies on a uniform estimate on the distance
1123: between shrinkings of $M'$.
1124: 
1125: \begin{prop}\label{closeN}
1126: For any $\gt$ and $\gd$ with $\gep\geq \gt+\gd\geq\gt>0$ we have
1127: $$
1128:  \big( )M'(_{\gt+\gd}\big)_{b\gd}\supset )M'(_\gt.
1129: $$
1130: \end{prop}
1131: 
1132: The proposition can also be stated as follows: For any such $\gt$ and $\gd$, 
1133: the Hausdorff distance between the corresponding sets satisfies
1134: $$
1135:  \textrm{Hd}\big()M'(_{\gt+\gd}~,~)M(_{\gt}\big)= \leq b\gd.
1136: $$
1137: In other words, in order to prove the proposition, we need to show that for any
1138: $x\in )M(_\gt\setminus)M(_{\gt+\gd}=(X)_{\gt+\gd}\setminus (X)_{\gt}$ there is 
1139: $y\notin(X)_{\gt+\gd}$ with 
1140: $d(x,y)\leq b\gd$.
1141: 
1142: \begin{proof}
1143: Let $x\in(X)_{\gt+\gd}\setminus (X)_{\gt}$. Choose a lifting
1144: $\ti x\in(\ti X)_{\gt+\gd}\setminus (\ti X)_{\gt}$ of $x$.
1145: As $\ti x\in(\ti X)_{\gt+\gd}\setminus (\ti X)_{\gt}$, there is $\gt_1~(\gt+\gd\geq\gt_1\geq\gt)$
1146: such that $\ti x\in\partial(\ti X)_{\gt_1}$.
1147: 
1148: Let $c(t)$ be the piecewise
1149: geodesic curve of constant speed $b$, passing through $\ti x$, which is defined 
1150: inductively (for $t\geq \gt_1$) as follows:
1151: At time $t_1=\gt_1$ set $c(t_1)=\ti x$ and define its one sided derivative by 
1152: $$
1153:  \frac{d}{dt}_+ c(t_1)=b\hat n(\ti x),
1154: $$ 
1155: and define 
1156: $$
1157:  I_c(t_1)=\{ i\in I:x\in\partial (X_i)_{t_1}\}.
1158: $$
1159: Identify $c$ with the constant speed geodesics determined by this condition, as
1160: long as $t >t_1$ and $c(t )\notin \overline{(X_i)_\gt}$ for any $i\notin 
1161: I_c(t_1)$. Let $t_2$ be the first time (if such exist) $t_1\leq t<\gt+\gd$ at 
1162: which $c(t)$ hits some $\overline{(X_i)_t}$ for $i\notin I_c(t_1)$. 
1163: As the collection $\{ X_i\}_{i\in I}$ is locally finite, $t_2$ is well defined and strictly 
1164: bigger than $t_1$. We claim that $c(t_2)\in\partial (\tilde X)_{t_2}$, and 
1165: that $c(t)\notin (\ti X)_t$ for $t_1<t<t_2$. To prove this, we
1166: need to show that $c(t )\notin (\ti X_i)_t$ for any 
1167: $t_1\leq t\leq t_2$ and $i\in I_c(t_1)$. Fix such $i$ and observe that
1168: $$
1169:  \frac{d}{dt}|_{t=t_1} D_{X_i}\big(c(t)\big)=
1170:  b\hat n(x)\cdot \nabla D_{X_i}(x)>b\frac{1}{b}=1.
1171: $$ 
1172: Since the convex function $D_{X_i}\big( c(t )\big)$ has 
1173: non-decreasing derivative we get that $\frac{d}{dt}\Big( D_{X_i}\big( c(t)\big)\Big) >1$ 
1174: for any $t_2>t>t_1$, and hence, 
1175: $D_{X_i}\big( c(t)\big)>D_{X_i}\big( c(t_1)\big)+(t-t_1)=t$ 
1176: for any such $t$. 
1177: Thus, the point $c(t)$ is outside $\big( (X_i)_{t_1}\big)_{t -t_1}=(X_i)_t$.
1178: 
1179: Then we define the second piece of $c$ by the condition that its one sided
1180: derivative $\frac{d}{dt}_+ c(t_2)$ satisfies 
1181: $\frac{d}{dt}_+ c(t_2)=b\hat n\big( c(t_2)\big)$, 
1182: and we define 
1183: $$ 
1184: I_c(t_2)=\{ i\in I:c(t_2)\in \partial (X_i)_{t_2}\}.
1185: $$
1186: 
1187: Note that since $t_2$ is smaller than $\gt+\gd <\gep$ and 
1188: $c(t_2)\in\partial (X_2)$, the  direction $\hat n\big( c(t_2)\big)$ is well 
1189: defined. In this way, we continue to define $c$ inductively.
1190: 
1191: If $t_i$ converge to some $t_{\omega_0}<\gt+\gd$, then we define 
1192: $c(t_{\omega_0})$ to be the limit of $c(t_i)$ and 
1193: $\frac{d}{dt_+}c(t_{\omega_0})=b\hat n\big( c(t_{\omega_0})\big)$
1194: and $I_c(t_{\omega_0})=\{i\in I:c(t_{\omega_0})\in\partial 
1195: (X_i)_{t_{\omega_0}}\}$,
1196: and denote the next index by $\omega_0+1$, and so on.
1197: By this way we obtain a piecewise
1198: geodesic curve with at most countably many pieces, connecting 
1199: $\ti x=c(\gt_1)$ to $\ti y=c(\gt +\gd)$, 
1200: with $\ti y=c(\gt +\gd )\notin (\tilde X)_{\gt +\gd}$. 
1201: Since the length of 
1202: $c\big( [\gt_1,\gt +\gd ]\big)$ is $b(\gt +\gd-\gt_1)\leq b\gd$ 
1203: this proves the proposition.
1204: \end{proof}
1205: 
1206: \begin{cor}\label{cover}
1207: Let $\gt$ and $\gd$ be as in Proposition \ref{closeN}. Let $\mathcal{C}$ be a
1208: collection of balls of radius $(b+1)\gd$ for which the set of centers
1209: $\mathcal{C}'$ form a maximal $\gd$ discrete subset of $)M(_{\gt+\gd}$.
1210: Then $\mathcal{C}$ is a cover of $)M(_\gt$.
1211: \end{cor}
1212: 
1213: \begin{proof}
1214: As $\mathcal{C}'$ is maximal $(\mathcal{C}')_\gd\supset )M(_{\gt+\gd}$.
1215: By Proposition \ref{closeN} we have 
1216: $$
1217:  \big( )M'(_{\gt+\gd}\big)_{b\gd}\supset )M'(_\gt.
1218: $$
1219: Thus
1220: $$ 
1221:  \cup_{C\in \mathcal{C}}C=(\mathcal{C}')_{(b+1)\gd}=
1222:  \big( (\mathcal{C}')_\gd\big)_{b\gd}\supset
1223:  \big( )M'(_{(\gt +\gd)}\big)_{b\gd}\supset )M'(_\gt.
1224: $$
1225: \end{proof}
1226: 
1227: In the sequel we will take $\gt=\frac{\gep}{2}$.
1228: The next proposition provides a uniform bound on the curvature of the smooth pieces
1229: of $\partial )M'(_{\frac{\gep}{2}}$.
1230: 
1231: \begin{prop}\label{smallC}
1232: For any $X_i$ and any point $x \in \partial (X_i)_{\frac{\epsilon}{2}}$,
1233: the $\frac{\gep}{2}$-ball, whose boundary sphere is tangent at $x$ to  
1234: $\partial (X_i)_{\frac{\epsilon}{2}}$, which lies on the same side of 
1235: $\partial (X_i)_{\frac{\gep}{2}}$ as $(X_i)_{\frac{\gep}{2}}$,
1236: is contained in $(X_i)_{\frac{\epsilon}{2}}$.
1237: \end{prop}
1238: 
1239: \begin{proof}
1240: The distance between $x$ and its closest point $p_{X_i}(x)$ in the closed 
1241: convex set $X_i$ is easily seen to be $\frac{\gep}{2}$, and the 
1242: $\frac{\gep}{2}$-ball 
1243: centered at $p_{X_i}(x)$ is the required one.
1244: \end{proof}
1245: 
1246: The following proposition follows directly from the definition of a deformation 
1247: retract.
1248: 
1249: \begin{prop}\label{B-B'}
1250: Let $B'\subset B$ be topological spaces, and let $F_t~\big( t\in [0,1]\big)$ 
1251: be a deformation retract of $B$ such that $F_t(b)\in B'$ for any 
1252: $b\in B',t\in [0,1]$. 
1253: Then $F_t|_{B'}$ is a deformation retract of $B'$.
1254: \end{prop}
1255: 
1256: Let $B_r$ denote an arbitrary ball in $S$ of radius $r$, and $B_r(x)$ the ball 
1257: of radius $r$ centered at $x$. The following proposition follows from the fact 
1258: that the volume of a ball of radius $r$ is a convex function of $r$ (because 
1259: the surface area of the $r$-sphere is an increasing function of $r$).
1260: 
1261: \begin{prop}
1262: There is a constant $m$ such that for any $\delta <1$, 
1263: $$
1264:  m\cdot \vol (B_{\delta /2})
1265:  \geq \vol (B_{(b+1.5)\delta}).
1266: $$ 
1267: Thus any $\delta$-discrete subset of 
1268: $B_{(b+1)\delta}$ consists of at most $m$ elements.
1269: \end{prop}
1270: 
1271: For a finite set
1272: $\{ y_1,...,y_t\} \subset S$ we denote by $\sigma (y_1,...,y_t)$
1273: its Chebyshev center, i.e. the unique point $x$ which minimizes the function
1274: $\max_{1\leq i\leq t}d(x,y_i)$.
1275: 
1276: We will soon take $B$ to be an intersection of balls, and $B'\subset B$ to be
1277: the intersection of $B$ with $)M'(_{\frac{\gep}{2}}$. 
1278: We intend to use \ref{B-B'} in order to show that under some certain 
1279: conditions, $B'$ is contractible.
1280: It will be natural to use the star contraction to the Chebyshev center 
1281: of the centers of the associated balls - The deformation retract which 
1282: makes any point of $B$ flow along the geodesic segment which connects it to 
1283: the required Chebyshev center, at constant velocity $s$, where $s$ 
1284: equals the initial distance. In order to do this, we need the following:
1285: 
1286: \begin{prop}[Defining the constant $\gd$] \label{positiveD}
1287: There exists $\gd~(0<\delta < \frac{\epsilon}{2(b+1)})$ 
1288: such that for any point $x\in S$, any $\frac{\gep}{2}$-ball
1289: $C$ which contains $x$ on its boundary sphere, and  
1290: any $m$ points $y_1,\ldots ,y_m\in B_{(b+1)\delta }(x)\setminus 
1291: (C)_{\delta}$, the inner product of the external normal vector of $C$ at $x$ 
1292: with the tangent at $x$ to the geodesic segment $[x,\sigma (y_1,\ldots ,y_m)]$ 
1293: is positive.
1294: \end{prop}
1295: 
1296: \begin{proof}
1297: Assume the contrary. Then there is a sequence $\delta_n \to 0$,
1298: a corresponding sequence $(C_n)$ of $\frac{\epsilon}{2}$-balls tangent to some 
1299: $x_0\in S$ (which we may assume converge to some fixed such ball), and a 
1300: corresponding sequence of
1301: $m$-tuples of points 
1302: $y_1^n,\ldots ,y_m^n\in B_{(b+1)\delta_n}(x_0)\setminus ( C_n)_{\gd_n}$, 
1303: such that the inner product of the
1304: tangent at $x_0$ to the geodesic segment $[x_0,\sigma (y_1^n,...,y_m^n)]$ with
1305: the external normal vector of the corresponding sphere $\partial C_n$ is non-positive (we may fix $x=x_0$ 
1306: since $G$ acts transitively).
1307: 
1308: Rescaling the metric each time we can
1309: assume that $\delta_n$ is fixed and equals $1$. We then get a sequence of 
1310: Riemannian metrics converging, on the ball of radius $b+1$ around $x_0$,
1311: to the Euclidean metric on the ball of radius $b+1$. 
1312: More precisely, we look at the ball of radius $b+1$ in the tangent space
1313: $T_{x_0}S$. We identify it
1314: each time with the ball of radius $(b+1)\delta_n$ around $x_0$ in $S$ via the 
1315: map $X\to \exp_{x_0}(\delta_n X)$, and we rescale the metric there to 
1316: $$
1317:  d_n(X,Y)=\frac{1}{\delta_n}\cdot d\big(\exp_{x_0}(\delta_nX),\exp_{x_0}(\delta_nY)\big).
1318: $$
1319: ( All these metrics induce the same topology.)
1320: 
1321: Now, in the rescaled metrics our tangent balls tend to a half space 
1322: (since $\epsilon /\delta_n \to \infty$), and we may as well assume that our 
1323: $m$-tuples also converge. 
1324: In the limit, we get an $m$-tuple of points
1325: in a Euclidean space at distance at least $1$ (and at most $b+1$)
1326: from a half space, for which
1327: the inner product of the external normal vector to this half space with the 
1328: vector $\overrightarrow v$, pointing from \textrm{so}me $x_0$ on the boundary 
1329: hyper-plane, towards the Chebyshev center of this $m$-tuple is non-positive. This is an absurd.
1330: \end{proof}  
1331: 
1332: Finally, we claim
1333: 
1334: \begin{prop}
1335: Let $\mathcal{C}$ be a collection of balls of radius $(b+1)\delta$, for which 
1336: the set of centers $\mathcal{C}'$ form a maximal $\delta$-discrete subset of
1337: $)M'(_{\frac{\gep}{2} +\gd}$. Then $\mathcal{C}$, i.e. the restrictions of
1338: its sets to $)M'(_{\frac{\gep}{2}}$, is a good cover of 
1339: $)M'(_{\frac{\gep}{2}}$.
1340: \end{prop}
1341: 
1342: \begin{proof}
1343: By corollary \ref{cover}, $\mathcal{C}$ is a cover of $)M'(_{\frac{\gep}{2}}$.
1344: 
1345: Let $B$ be the intersection of $m$ (not necessarily different) balls of 
1346: $\mathcal{C}$, with centers $y_1,\ldots,y_m\in\mathcal{C}$. 
1347: 
1348: Proposition \ref{smallC} implies that for any 
1349: $\tilde x\in \partial )\tilde M'(_{\frac{\gep}{2}}$ 
1350: and for any  $X_i$ with
1351: $\tilde x\in \partial  (X_i)_{\frac{\epsilon}{2}}$
1352: there is an $\frac{\epsilon}{2}$-ball,
1353: tangent to $\partial (X_i)_{\frac{\gep}{2}}$ at $\tilde x$,
1354: which is contained in $(X_i)_{\frac{\gep}{2}} \subset 
1355: (\tilde X)_{\frac{\gep}{2}}$.
1356: Thus, if in addition the image $x$ of $\tilde x$ belongs to $B$, then 
1357: Proposition \ref{positiveD} implies that the geodesic segment 
1358: $[x,\sigma (y_1,\ldots ,y_m)]$ lies inside $B'=B\cap )M'(_{\frac{\gep}{2}}$.
1359: Let us explain this point as follows. Let $c(t),~t\in [0,1]$ be a 
1360: parameterization of the geodesic segment $[x,\sigma (y_1,\ldots ,y_m)]$. 
1361: Propositions \ref{smallC} and \ref{positiveD} imply that 
1362: $c(t)\notin \overline{(X)}_{\frac{\gep}{2}}$ for all sufficiently small $t>0$.
1363: Assume that $c(t)\in\partial (X)_{\frac{\gep}{2}}$ for some $t_0$, $0<t_0<1$, 
1364: (and let $t_0$ be the first such  time). Then Propositions \ref{smallC} and 
1365: \ref{positiveD} applied to $c(t_0)$ imply that 
1366: $c(t_0-\gD t)\in(X)_{\frac{\gep}{2}}$ for small $\gD t$. But this is a 
1367: contradiction.
1368: We conclude that if $B$ is not empty 
1369: then $\sigma (y_1,\ldots ,y_m)\in B'$ and the star-contraction from $B$ to
1370: $\sigma (y_1,\ldots ,y_m)$ induces a contraction of $B'$. Hence the 
1371: set $B'$ is non-empty and contractible.
1372: This means that $\{ C':C\in \mathcal{C} \}$ is a good cover of 
1373: $)M'(_{\frac{\gep}{2}}$, where $C':=C\cap )M'(_{\frac{\gep}{2}}$.
1374: \end{proof}
1375: 
1376: We conclude that $)M'(_{\frac{\gep}{2}}$, and therefore $M'$, is homotopically 
1377: equivalent to the simplicial complex $\mathcal{R}$ which corresponds to the 
1378: nerve of $\mathcal{C}$. Since the collection of centers $\mathcal{C}'$ is 
1379: $\gd$-discrete, and therefore the $\frac{\gd}{2}$-balls with the same centers
1380: are disjoint, we get that $|\mathcal{C}'|$,  
1381: the number of the vertices of $\mathcal{R}$, is 
1382: $\leq\frac{\vol (M)}{\vol (B_{\gd /2})}$. 
1383: Since the sets of $\mathcal{C}$ are subsets of $(b+1)\gd$-balls, each of them 
1384: intersects at most 
1385: $\frac{\vol ( B_{2(b+1)\gd+\gd /2})}{\vol (B_{\gd /2})}$ of the 
1386: others. Hence, the degree of any vertex in $\mathcal{R}$ is 
1387: $\leq d:=\frac{\vol ( B_{2(b+1.25)\gd})}{\vol (B_{\gd /2})}$.
1388: This completes the proof of Lemma \ref{simplicial-complex}.
1389: 
1390: 
1391: 
1392: 
1393: 
1394: 
1395: %NEW SECTION
1396: 
1397: 
1398: 
1399: 
1400: 
1401: \section{An arithmetic variant of the Margulis lemma}\label{AML}
1402: 
1403: The classical Margulis lemma yields information on the algebraic structure
1404: of a discrete group of isometries which is generated by ``small elements''. If,
1405: in addition, this discrete group lies inside an arithmetic group then we can 
1406: say a little more. In this section we shall explain this idea in the
1407: non-uniform case. 
1408: 
1409: The Lie group of isometries $G=\textrm{Isom}(S)$
1410: is center-free, semi-simple, without compact factors and with 
1411: finitely many connected components. Its identity component $G^0$ coincides 
1412: with the connected component of the group of real points $\Bbb{G}(\BR )^0$ of some
1413: $\BQ$-algebraic group $\Bbb{G}$. 
1414: We shall identify $G^0$ with its adjoint group
1415: $\text{Ad}(G^0)\leq\text{GL}(\mathfrak{g})$ and think of it as a group of 
1416: matrices. We will denote by $\mu$ a fixed Haar measure on $G$.
1417: 
1418: \begin{lem}[\bf An arithmetic variant of the Margulis lemma]\label{arith-M-L}
1419: There are constants $\gep =\gep (S)>0$ and $m=m(S)\in\BN$, such that 
1420: if $\gC\leq G$ is a non-uniform torsion-free arithmetic lattice, then 
1421: for any $x\in S$, the group of real points of the Zariski closure
1422: $\big(\overline{\gC_\gep (x)}^z\big)_\BR$ of the group 
1423: $$
1424:  \gC_\gep (x)=\langle\gc\in\gC:d_\gc (x)\leq\gep\rangle
1425: $$
1426: has at most $m$ connected components, and its identity component is a 
1427: unipotent subgroup.
1428: \end{lem}
1429: 
1430: The following claims (\ref{Q-structure}, \ref{5.3}, \ref{5.4}) 
1431: are well known (c.f. \cite{Mar1}, chapter $IX$ section 4). 
1432: 
1433: \begin{lem}\label{Q-structure}\label{5.2}
1434: For any non-uniform arithmetic lattice $\gD\leq G^0$, there is a rational 
1435: structure on $G^0$ (coming from a rational structure on the vector space $\mathfrak {g}$)
1436: with respect to which $\gD$ is contained in $G^0(\BQ )$ and
1437: commensurable to $G^0(\BZ )$. Conjugating by an element $g\in G^0(\BQ )$ we can
1438: assume that $\gD\subset\ G^0(\BZ )$.
1439: \end{lem}
1440: 
1441: \begin{proof}[Explanation]
1442: In general, if $\gD$ is an arithmetic lattice in $G$, there is 
1443: a compact extension $H$ of $G$, a rational structure on $H$, and a subgroup
1444: $\gD'\leq H_\BQ$ commensurable to $H_\BZ$ whose projection to $G$ is $\gD$.   
1445: However, when $\gD$ is non-uniform we can always take $H=G$.
1446: This could be deduced, for example, from the fact that $\gD$ is almost 
1447: generated by unipotent elements, and that compact groups contain no non-trivial 
1448: unipotent element.
1449: 
1450: For a given rational structure, the group $G_\BZ$ is defined only up to 
1451: commensurability. A subgroup $\gD\leq G_\BQ$ which is commensurable to 
1452: $G_\BZ$ is always conjugate, by an element of $G_\BQ$, to a subgroup of $G_\BZ$.
1453: \end{proof}
1454: 
1455: If $g\in G$ is close to $1\in G$, then its characteristic polynomial (i.e. the characteristic 
1456: polynomial of the endomorphism $\text{Ad}(g)$) is close to $(\gl -1)^n$ where 
1457: $n=\dim (\mathfrak {g})$. In particular:
1458: 
1459: \begin{prop}\label{5.3}
1460: There is an identity neighborhood $\gO_1\subset G$ such that, if $g\in \gO_1$,
1461: and the characteristic polynomial of $g$ has integral coefficients, 
1462: then $g$ is unipotent.
1463: \end{prop}
1464: 
1465: \ref{5.2} and \ref{5.3} implies:
1466: 
1467: \begin{cor}[\cite{Mar1} 4.21]\label{5.4}
1468: For any non-uniform arithmetic lattice $\gD\leq G$, the intersection 
1469: $\gO_1\cap\gD$ consists of unipotent elements only.
1470: \end{cor}
1471: 
1472: Recall also the following theorem of Zassenhaus and Kazhdan-Margulis (see \cite{Rag} 
1473: theorem 8.16):   
1474: 
1475: \begin{thm}[Zassenhaus, Kazhdan-Margulis]
1476: There exists an identity neighborhood $\gO_2\subset G$, called a
1477: Zassenhaus neighborhood, so that for any discrete subgroup $\gS\leq G$,
1478: the intersection $\gS\cap\gO_2$ is contained in a connected nilpotent Lie
1479: subgroup of $G$.
1480: \end{thm}
1481: 
1482: \begin{proof}[Proof of Lemma \ref{arith-M-L}]
1483: Let $\gO\subset G$ be a relatively compact symmetric identity neighborhood 
1484: which satisfies $\gO^2\subset\gO_1\cap\gO_2$.
1485: 
1486: Fix an integer $m$,
1487: $$
1488:  m> \inf_{h\in G}\frac{\mu \big(\{g\in G:d_g(x)\leq 1\}\cdot h\gO h^{-1} \big)}
1489:  {\mu (\gO )}, 
1490: $$ 
1491: and
1492: $$
1493:  \gep =\frac{1}{m}.
1494: $$ 
1495: As $G$ is unimodular, $m$ can be chosen independently of $x$. 
1496: Replacing $\gO$ by some conjugate $h\gO h^{-1}$, if needed, we can assume that
1497: $$
1498:  m>\frac{\mu \big(\{g\in G:d_g(x)\leq 1\}\cdot\gO \big)}{\mu (\gO )}.
1499: $$ 
1500: Let
1501: $$
1502:  \gC_\gO=\langle\gC\cap\gO^2\rangle.
1503: $$
1504: Then 
1505: $$
1506:  [\gC_\gep (x): \gC_\gep (x)\cap\gC_\gO ]\leq m.
1507: $$
1508: To see this, assume for a moment that this index was $\geq m+1$. Then we 
1509: could find $m+1$ representatives $\gc_1,\gc_2,\ldots,\gc_{m+1}\in\gC_\gep (x)$
1510: for different cosets of $\gC_\gep (x)\cap\gC_\gO$ in the ball of radius $m$
1511: in $\gC_\gep (x)$ according to the word metric with respect to the generating
1512: set $\{\gc\in\gC_\gep (x): d_\gc (x)\leq\gep\}$. As they belong to different
1513: cosets, $\gc_i\gO\cap\gc_j\gO =\emptyset$ for any $1\leq i<j\leq m+1$.
1514: Since $d_{\gc_i}(x)\leq m\cdot\gep =1$ these $\gc_i$'s
1515: are all inside $\{ g\in G:d_g(x)\leq 1\}$. This contradicts the assumption 
1516: $m\cdot\mu (\gO )>\mu \big(\{ g\in G:d_g(x)\leq 1\}\cdot\gO\big)$.
1517: 
1518: It follows from the Zassenhaus-Kazhdan-Margulis theorem that $\gC_\gO$ is contained in a 
1519: connected nilpotent Lie subgroup of $G$, and therefore, by Lie's theorem,
1520: $\gC_\gO$ is triangulable over $\BC$. As $\gC_\gO$ is generated by unipotent
1521: elements, it follows that $\gC_\gO$ is a group of unipotent elements.
1522: Thus the Zariski closure $\overline{\gC_\gO}^z$ is a unipotent algebraic group,
1523: and hence, the group of its real points $(\overline{\gC_\gO}^z)_\BR$ is 
1524: connected in the real topology. Similarly, its subgroup
1525: $(\overline{\gC_\gep (x)\cap\gC_\gO}^z)_\BR$ is connected.
1526: Clearly, $(\overline{\gC_\gep (x)\cap\gC_\gO}^z)_\BR$ is the identity 
1527: connected component of $(\overline{\gC_\gep (x)}^z)_\BR$, 
1528: and its index is at most $m$.
1529: 
1530: \end{proof}
1531: 
1532: \begin{rem}
1533: Although $\gep$ could be taken to be $1/m$, we use different letters for them
1534: because they play different roles.
1535: In the sequel we will assume that the above lemma is satisfied with $\gep$
1536: replaced by $10\gep$.
1537: \end{rem}
1538: 
1539: \begin{rem}\label{5.7}
1540: It follows from Lemma \ref{arith-M-L} that if $\gc$
1541: is an element of a non-uniform arithmetic lattice of $G$ and 
1542: $\inf d_\gc < \gep$, then $\gc^j$ is unipotent for some $j\leq m$.
1543: This implies that in a non-compact arithmetic $S$-manifold there are no closed geodesic of length 
1544: $\leq\gep$, and that in the $\gep$-thick thin decomposition, the thin part has no
1545: compact connected component. For example, the $\gep$-thin part of any 
1546: non-compact arithmetic hyperbolic surface (or more generally of any rank-1
1547: manifold) consists only of cusps.
1548: \end{rem}
1549: 
1550: 
1551: 
1552: 
1553: 
1554: %new section
1555: 
1556: 
1557: 
1558: 
1559: 
1560: \section{The proof of \ref{thmA}(1)}\label{6}
1561: 
1562: In this section we shall prove:
1563: 
1564: \begin{thm}[{\bf Theorem \ref{thmA}(1) of the introduction}]\label{thm1}
1565: Let $S$ be a symmetric space of non-compact type. Then there are constants
1566: $\ga$ and $d$ (depending only on $S$) such that any non-compact arithmetic
1567: $S$-manifold $M$ is homotopically equivalent some 
1568: $\big( d,\ga\cdot\vol (M)\big)$-simplicial complex.
1569: \end{thm}
1570: 
1571: Fix $\gep' =\gep' (S),m=m(S)$, such that Lemma \ref{arith-M-L} is satisfied 
1572: with $10\gep'$, and $m$. Assume that $M=\gC\backslash S$ is a given non-compact 
1573: arithmetic $S$-manifold. Denote by $\gC^u$ the set of unipotent elements in 
1574: $\gC$,
1575: $$
1576:  \gC^u=\{ \gc\in\gC :\gc \textrm{ is unipotent}\},
1577: $$
1578: and define
1579: $$
1580:  \tilde X=\cup_{\gc\in\gC^u\setminus\{ 1\}}\{ d_\gc<\gep'\},
1581: $$
1582: and
1583: $$
1584:  \tilde M'=\tilde M \setminus \tilde X,
1585: $$
1586: and let $X\subset M$ and $M'\subset M$ be the images of $\ti X$ and 
1587: $\ti M'$ under the universal covering map.
1588: 
1589: In order to prove Theorem \ref{thm1} we shall show:
1590: 
1591: \begin{enumerate}
1592: \item\label{1}
1593: There is a deformation retract from $M$ to $M'$. In particular $M$ is 
1594: homotopically equivalent to $M'$.
1595: \item\label{2}
1596: $M'$ satisfies the conditions of Lemma \ref{simplicial-complex} with 
1597: $\gep =\gep'/m$.
1598: \end{enumerate}
1599: 
1600: Let us start with some preliminaries. Let $W\subset S(\infty )$ be a Weyl
1601: chamber of the Tits spherical building. Then $W$ is isometric to a spherical
1602: simplex.
1603: We define its center of mass $z$ to be
1604: $$
1605:  z=\frac{\int_W xd\mu (x)}{\|\int_W xd\mu (x)\|}
1606: $$
1607: where $\mu$ is the Lebesgue measure on the sphere. Since $W$ is 
1608: contained in a half 
1609: sphere (see \cite{BGS} appendix 3), we get $\|\int_W xd\mu(x)\|\neq 0$. Since $W$
1610: is convex (in the spherical metric), and since its interior is non-empty and is 
1611: exactly the set of regular points in $W$, it follows that $z$ is regular
1612: and contained in $W$.
1613: 
1614: We shall also use:
1615: 
1616: \begin{lem}\label{321}
1617: Assume that $g$ is a parabolic isometry (i.e. $\min (g)=\emptyset$), $c(t)$ is a regular geodesic, and
1618: $g$ stabilizes $c(-\infty )$. Then $\frac{d}{dt}d_g\big( c(t)\big)\neq0$
1619: for any $t\in\BR$.
1620: \end{lem}
1621: 
1622: \begin{proof}
1623: Let $c(t)$ be a regular geodesic in $S$, and let $g$ be an isometry which 
1624: stabilizes $c(-\infty )$. Assume that 
1625: $\frac{d}{dt}|_{t=t_o}d_g\big( c(t)\big) =0$, then the function 
1626: $d_g\big( c(t)\big)$, being analytic and convex 
1627: must be constant, since it doesn't tend to $\infty$ as 
1628: $t\to -\infty$. This means that $g\cdot c$ is a geodesic parallel to $c$.
1629: We conclude that $g$ preserves the unique flat which contains $c$, since this
1630: flat is exactly the set of points through which there is a geodesic parallel
1631: to $c$. But this implies that $g$ is semisimple, a contradiction.
1632: \end{proof}
1633: 
1634: Furthermore, if $g$ is unipotent, then $d_g\big( c(t)\big)\to 0$ as $t\to -\infty$
1635: (see \cite{BGS} appendix 5, section 4).
1636: 
1637: \begin{proof}[{\bf The proof of (\ref{1})}]
1638: We will show that the conditions of Lemma \ref{deformation-retract} are 
1639: satisfied with $L=Y=S$ and the $\gC$-invariant family of functions
1640: $\mathcal{F}=\{d_\gc\}_{\gc\in\gC^u\setminus\{ 1\}}$. 
1641: Then $\ti M'=\mathcal{F}_{\geq\gep'}$.
1642: 
1643: The finiteness of the sets 
1644: $$
1645:  \Psi_{x,\gt }=\{\gc\in\gC^u\setminus\{ 1\}:d_\gc (x)\leq\gt\}
1646: $$ 
1647: follows from the compactness of $\{ g\in G:d_g(x)\leq\gt\}$
1648: together with the discreteness of $\gC$. All the functions 
1649: $\{ d_\gc\}_{\gc\in\gC^u\setminus\{ 1\}}$ are convex, strictly positive, 
1650: without critical values. 
1651: We need to define the continuous function $\gb :(0,3\gep )\to\BR^{>0}$, 
1652: and the appropriate direction $\hat n(x)\in T_x(S)$ for any $x\in S$ with
1653: $\Psi_{x,3\gep}\neq\emptyset$.
1654: 
1655: Fix $x\in S$ with $\Psi_{x,3\gep}\neq\emptyset$ and let 
1656: $\Psi_{x,3\gep}=\{\gc_1,\gc_2,\ldots,\gc_k\}$.
1657: By Lemma \ref{arith-M-L} the Zariski closure $\overline\gD^z$ of
1658: the group $\gD =\langle \gc_1,\gc_2,\ldots,\gc_k \rangle$ 
1659: has a unipotent identity component, and $\leq m$ connected components. 
1660: Since $\gc_i$ is unipotent,
1661: it is contained in the Zariski closure of the cyclic group generated by any 
1662: power of it. 
1663: As $\gc_i^j$ belongs to the identity component $(\overline{\gD}^z)^0$ for some
1664: $j\leq m$, also $\gc_i\in(\overline{\gD}^z)^0$. Since this holds for any
1665: generator $\gc_i$, we get that
1666: $\overline{\gD}^z=(\overline{\gD}^z)^0$ and hence the group of its
1667: real points $(\overline{\gD}^z)_\BR$ is a connected unipotent group.
1668: Hence $\gc_1,\gc_2,\ldots,\gc_k$ are contained in a connected unipotent group.
1669: 
1670: Let $N\leq G$ be a maximal 
1671: connected unipotent subgroup which contains $\gc_1,\gc_2,\ldots ,\gc_k$.
1672: Let $W\leq S(\infty )$ be the Weyl chamber of the Tits boundary of $S$ 
1673: which corresponds to $N$. Let $c(t)=c_x(t)$ be the
1674: geodesic line with $c(0)=x$ for which $c(-\infty )$ is the center of mass of
1675: $W$.
1676: 
1677: By Lemma \ref{321}
1678: $$
1679:  \frac{d}{dt}|_{t=0}\Big( d_g\big( c(t)\big)\Big)>0
1680: $$ 
1681: for any $g\in N\setminus\{ 1\}$. In addition, the continuous function
1682: $$
1683:  h(g)=\dot c(0)\cdot\nabla d_g(x)
1684: $$ 
1685: attains a minimum on the compact set
1686: $$
1687:  \{g\in N:d_g(x)=\gt\}.
1688: $$
1689: We define $\gb (\gt )$ to be this minimum (then $\gb (\gt )$ is defined for
1690: any $\gt >0$). By definition, 
1691: $\dot c(0)\cdot\nabla d_{\gc_i}(x)\geq\gb\big(d_{\gc_i}(x)\big)$.  
1692: Moreover, it is easy to see that $\gb$ is a 
1693: continuous positive function. Since $G$ acts transitively on the
1694: set of couples $(W,x)$ of a Weyl chamber $W\subset S(\infty )$ and a point $x\in S$,
1695: $\gb$ is independent of $N$ and of $x$.
1696: The conditions of Lemma \ref{deformation-retract} are satisfied with the 
1697: tangent vector $\hat n(x)=\dot c_x(0)$.
1698: \end{proof}
1699: 
1700: \begin{rem}
1701: We used the existence and uniqueness of the center of mass
1702: of $W$ in order to define 
1703: $\gb$ in a canonical way. However, as any two Weyl chambers are isometric,
1704: we could choose arbitrarily a regular point in one Weyl chambers, and translate
1705: it to any other Weyl chambers, and to use these points when defining 
1706: $c(-\infty )$ and then $c, \hat n(x)$ and
1707: $\gb$ in our proof.
1708: \end{rem}
1709: 
1710: \begin{rem}\label{finite-presentability}
1711: A slight modification of the argument above, yields an elementary proof of the 
1712: fact that $M=\gC\backslash S$ is homotopic to a compact manifold with boundary, and 
1713: hence that $\gC$ is finitely presented, when $\gC\leq G$ is any torsion free (non-uniform)
1714: lattice. We used the arithmeticity of $\gC$ in order to get a uniform estimate for all $\gC$'s.
1715: However, for a fixed $\gC$, we could have used corollary 11.18 from \cite{Rag} instead of 
1716: corollary \ref{5.4} above, and by this to avoid the assumption that $\gC$ is arithmetic. Moreover, 
1717: we don't really have to assume that $\gC$ is torsion free (when proving just the finiteness without
1718: explicit estimate). The same argument shows that in general, any $S$-orbifold 
1719: $\gC\backslash S$ is homotopic to a compact orbifold with boundary, and it is not hard to show by the same 
1720: means that $\gC\backslash G$ is homotopic to a compact manifold with boundary. In particular, our method 
1721: gives a quite elementary proof of the finite presentability of lattices.
1722: Furthermore, changing appropriately the factor $\sqrt{\cdot}$ in the vector field which induces
1723: the deformation retract, so that it will decay more slowly, we can get a simple proof that 
1724: $\gC\backslash G$ is diffeomorphic to the interior of a compact manifold with boundary.
1725: \end{rem}
1726: 
1727: \begin{proof}[{\bf The proof of (\ref{2})}]
1728: In order to check that the conditions of Lemma \ref{simplicial-complex}
1729: are satisfied, we will show that
1730: 
1731: \begin{itemize}
1732: \item $M'\subset M_{\geq\gep}$ (where $\gep =\frac{\gep '}{m}$).
1733: \item $\ti X$ is a locally finite union of convex open sets with smooth boundary.
1734: \item $M'$ is homotopically equivalent to its $\frac{\gep'}{2m}$-shrinking 
1735: $)M'(_{\frac{\gep'}{2m}}$.
1736: \end{itemize}
1737: And that there are
1738: \begin{itemize}
1739: \item $b=b(S)>0$ and
1740: \item $\hat n(x)\in T_x(S)$ for any 
1741: $x\in (\ti X)_{\frac{\gep'}{2m}}\setminus (\ti X)$
1742: \end{itemize}
1743: such that
1744: $$
1745: \hat n(x)\cdot\nabla D_{\{\gc\leq\gep'\}}(x)>\frac{1}{b}
1746: $$
1747: for any $\gc\in\gC^u\setminus\{ 1\}$ with 
1748: $D_{\{\gc\leq\gep'\}}(x)\leq\frac{\gep'}{2m}$ (this is stronger than the condition on the
1749: inner products required in Lemma \ref{simplicial-complex}).\\
1750: 
1751: If $x\in S$ and $d_\gc (x)\leq\frac{\gep'}{m}\leq\gep'$ for some $\gc\in\gC$, 
1752: then by Lemma \ref{arith-M-L} $\gc^j$ is unipotent for some $j\leq m$,
1753: and since $d_{\gc^j} (x)\leq\frac{j\cdot\gep'}{m}\leq\gep'$, we obtain that
1754: $x\in\ti X$. This proves that $M'$ is contained in the $\frac{\gep'}{m}$-thick
1755: part $M_{\geq\frac{\gep'}{m}}$.\\
1756: 
1757: Since for $\gc\in\gC^u\setminus\{ 1\}$ the displacement function is convex and
1758: analytic and $\inf d_\gc=0$, and since $\gC$ is discrete, we have that 
1759: $\ti X=\cup_{\gc\in\gC^u\setminus\{ 1\}}\{d_\gc<\gep\}$ is a locally finite 
1760: union of convex open sets with smooth boundary.
1761: \\
1762: 
1763: We shall prove that there is a deformation retract from $M'$ to 
1764: its $\frac{\gep'}{2m}$-shrinking $)M'(_{\frac{\gep'}{2m}}$
1765: by showing that the conditions of Lemma \ref{deformation-retract}
1766: are satisfied with 
1767: ${\mathcal F}=\{ D_{\{ d_\gc\leq\gep'\}}\}_{\gc\in\gC^u\setminus\{ 1\}}$,
1768: $Y=S$, $L={\mathcal F}_{\geq 0}=\ti M'$, and 
1769: ${\mathcal F}_{\geq\frac{\gep'}{2m}}=)\ti M'(_{\geq\frac{\gep'}{2m}}$.
1770: (In this proof $\frac{\gep'}{2m}$ plays the rule of $\gep$ in Lemma
1771: \ref{deformation-retract}.)
1772: 
1773: The finiteness of the sets 
1774: $$
1775:  \Psi_{x,\gt}=\{\gc\in\gC^u\setminus\{ 1\}:
1776:  D_{\{d_\gc\leq\gep'\}}(x)\leq\gt\}
1777: $$ 
1778: follows from the discreteness of $\gC$
1779: together with the compactness of 
1780: $$
1781: \{g\in G:D_{\{d_g\leq\gep\}}(x)\leq\gt\}.
1782: $$ 
1783: 
1784: We shall define the direction $\hat n(x)\in T_x(S)$ analogously 
1785: to the way it is done in the proof of (\ref{1}), 
1786: and shall show that the conditions of Lemma \ref{deformation-retract} are 
1787: satisfied with the constant function $\frac{\gb (\gep' )}{2}$, where $\gb$ 
1788: is the function defined in the proof of (\ref{1}).
1789: 
1790: Fix $x\in L=\mathcal{F}_{\geq 0}$, with $\Psi_{x,\frac{3\gep'}{2m}}\neq\emptyset$, and denote
1791: $\Psi_{x,\frac{3\gep'}{2m}}=\{\gc_1,\gc_2,\ldots,\gc_k\}$.
1792: Since $\gep'+2\frac{3\gep'}{2m}=4\gep'$ we have
1793: $(\{ d_{\gc_i}\leq\gep'\} )_{\frac{3\gep'}{2m}}\subset\{d_{\gc_i}\leq 4\gep'\}$
1794: which implies that $\cap_1^k\{d_{\gc_i}\leq 4\gep'\}\neq\emptyset$ and hence
1795: $\gc_1,\gc_2,\ldots,\gc_k$ are contained in a connected 
1796: unipotent group. As in the proof of (\ref{1}), let $N$ be a 
1797: maximal connected unipotent group which contains $\gc_1,\gc_2,\ldots,\gc_k$, 
1798: let $W\subset S(\infty )$ be the Weyl chamber which corresponds to $N$ in the 
1799: Tits boundary. Let $c(t)=c_x(t)$ be the geodesic line with $c(0)=x$ and $c(-\infty)=$ the center 
1800: of mass of $W$. Since $\gc_i\in N$,
1801: $d\big( c(t),\gc_i\cdot c(t)\big)$ tends to $0$ as $t\to -\infty$, and since
1802: $d_{\gc_i}(x)>\gep'$ we have
1803: $d_{\gc_i}\big( c(t_0)\big)=\gep'$ for some negative $t_0$. 
1804: 
1805: The function $d_{\gc_i}\big( c(t)\big)$ is convex and hence has a non-decreasing derivative,
1806: thus, taking $\hat n(x)=\dot c(0)$, we have:
1807: 
1808: \begin{eqnarray*}
1809:   \hat n(x)\cdot\nabla D_{\{d_{\gc_i}\leq\gep'\}}
1810:  (x)
1811: %&=&\\
1812: &=& \dot c(0)\cdot\nabla D_{\{d_{\gc_i}\leq\gep'\}}(x)\\ 
1813: & =& 
1814:  \dot c(t_0)\cdot\frac{\nabla d_{\gc_i}\big( c(0)\big)}
1815:  {\|\nabla d_{\gc_i}\big( c(0)\big)\|}\\
1816: &\geq &
1817:  \dot c(0)\cdot\frac{\nabla d_{\gc_i}\big( c(0)\big)}{2}\\
1818: &=& 
1819: \frac{1}{2}\cdot\frac{d}{dt}|_{t=0}d_{\gc_i}\big( c(t)\big)\\
1820: & \geq& 
1821: \frac{1}{2}\cdot\frac{d}{dt}|_{t=t_0}d_{\gc_i}\big( c(t)\big)\\
1822: &=&
1823: \frac{1}{2}\cdot\dot c(t_0)\cdot\nabla d_{\gc_i}\big( c(t_0)\big)
1824:  \geq\frac{\gb(\gep')}{2},
1825: \end{eqnarray*}
1826: where $\gb$ is the function defined in the proof of 
1827: \ref{1}.
1828: In the above computation we made use of the facts that 
1829: $\|\nabla  D_{\{d_{\gc_i}\leq\gep'\}}\| =1$ everywhere outside 
1830: $\{d_{\gc_i}<\gep '\}$, 
1831: and that $d_{\gc_i}$ is 2-Lipschitz and hence $\|\nabla d_{\gc_i}\|\leq 2$.
1832: 
1833: We completed the verification that the conditions of Lemma 
1834: \ref{deformation-retract} are satisfied, and hence proved that $M'$ is 
1835: homotopic to its $\frac{\gep'}{2m}$-shrinking.\\
1836: 
1837: Finally, observe that we can take also the constant $b$
1838: to be $b=\frac{2}{\gb(\gep')}$, and the unit tangent vector $\hat n(x)$ from 
1839: \ref{simplicial-complex} to be the same $\hat n(x)$ used above. Then all the conditions of Lemma
1840: \ref{simplicial-complex} are satisfied, and the proof of (\ref{2}) is 
1841: completed. 
1842: \end{proof}
1843: 
1844: 
1845: 
1846: 
1847: 
1848: 
1849: 
1850: %NEW SECTION
1851: 
1852: 
1853: 
1854: 
1855: 
1856: 
1857: 
1858: \section{Estimating angles at corners of the boundary}\label{bigangles}\label{7}
1859: 
1860: The next 3 sections are devoted to the remaining cases (2), (3) of Theorem \ref{thmA}.
1861: The main result on this section is Theorem \ref{nice-boundary}.
1862: 
1863: Our method is to replace a manifold $M$ by a submanifold with boundary $M'$, 
1864: in which the injectivity radius is uniformly bounded from below, and then  
1865: to apply Lemma \ref{simplicial-complex}. This requires some
1866: information on the boundary. The pre-image $\ti X$ of the complement 
1867: $X=M\setminus M'$ in the universal covering is required to be
1868: a locally finite union of 
1869: convex open sets with smooth boundary, and what we need is a control on the 
1870: angles of the corners - where the boundaries of two or more such sets intersect.
1871: In the non-uniform case we used the presence of many unipotents, and the nice
1872: actions of unipotent groups on the boundary at infinity. In the compact case
1873: there are no unipotents, so different tools are required.
1874: 
1875: Let $A,B\subset S$ be convex bodies with smooth boundary and with a common
1876: interior point in their intersection. The angle between the boundaries at a 
1877: common point
1878: $x\in \partial A \cap \partial B$ is measured by $\pi$ minus the angle between
1879: the external normal vectors $\hat n_A(x), \hat n_B(x)$
1880: $$
1881:  \phi_x(\partial A,\partial B)=
1882:  \pi -\angle \big(\hat n_A(x),\hat n_B(x)\big).
1883: $$ 
1884: Thus, a big angle between the boundaries corresponds to a small angle between 
1885: the normals.
1886: The following lemma states that when $A$ and $B$ are sub-level sets 
1887: of commuting isometries, these angles are $\geq\frac{\pi}{2}$. A similar statement appeared in \cite{BGS}.
1888: 
1889: For an isometry $\gc$, and for $x\in S$ with 
1890: $d_{\gc}(x)=a>\min d_\gc$ 
1891: we denote by $\hat n_{\gc}(x)$ the external (with respect to 
1892: $\{ d_{\gc} \leq a \}$)
1893: normal to 
1894: $$
1895:  \{ d_{\gc} =a \}=\partial\{ d_{\gc}\leq a \}.
1896: $$
1897: 
1898: \begin{lem}\label{bigA}{\bf (Commutativity implies big angles):}
1899: If the isometries $\alpha$ and
1900: $\beta$ commute then $\hat n_{\alpha}(x)\cdot \hat n_{\beta}(x)\geq 0$.
1901: \end{lem}
1902: 
1903: \begin{proof}
1904: Let $a_\alpha =d_\alpha (x),a_\beta =d_\beta (x)$.
1905: Since $\{ d_\alpha =a_\alpha \}$ is a level set for $d_\alpha$, we see that
1906: $\hat n_{\alpha}(x)$ is the direction of the gradient
1907: $(\nabla \cdot d_\alpha) (x)$.
1908: Thus if $c(t)$ is the geodesic line through $x$ with 
1909: $\dot c(0)=\hat n_{\beta}(x)$
1910: then it is enough to show that
1911: $$
1912:  (\nabla \cdot d_\alpha )(x) \cdot \hat n_{\beta}(x)=
1913:  \frac{d}{dt}|_{t=0}\{ d_\alpha \big( c(t)\big)\} \geq 0.
1914: $$
1915: Let $p$ denote the projection on the convex set 
1916: $\{ d_{\beta} \leq a_\beta \}$. Since $\alpha$ and $\beta$ commute, the 
1917: set $\{ d_{\beta} \leq a_\beta \}$ is $\alpha$-invariant and therefore
1918: $\alpha$ commute with $p$. Since $p$ decreases distances and since the 
1919: geodesic lines $c(t),\ga\cdot c(t)$ are both orthogonal to 
1920: $\partial\{ d_\gb\leq a_\gb\}$ one sees that $d\big( c(t),\ga\cdot c(t)\big)$ 
1921: is a non-decreasing function of $t$. Thus 
1922: $$
1923:  (\nabla \cdot d_\alpha )(x)\cdot \hat n_{\beta}(x)=
1924:  \frac{d}{dt}|_{t=0}\{ d_\alpha \big( c(t)\big)\} =
1925:  \frac{d}{dt}|_{t=0}\{d\big( c(t),\alpha \cdot c(t)\big)\} \geq 0.
1926: $$
1927: \end{proof}
1928: 
1929: We shall need similar information when $A,B$ are replaced by their 
1930: $\gep$-neighborhoods $(A)_\gep ,(B)_\gep$. The following lemma explains that
1931: the situation then only improves.
1932:  
1933: Define
1934: $$
1935:  \varphi_t= \sup_{x\in \partial (A)_t \cap \partial (B)_t}
1936:  \angle\big(\hat n_{(A)_t}(x),\hat n_{(B)_t}(x)\big),
1937: $$
1938: then we have
1939: 
1940: \begin{lem}\label{monoton}{\bf (Monotonicity of angles):}
1941: $\varphi_t$ is a non-increasing function of $t$.
1942: \end{lem}
1943: 
1944: \begin{proof}
1945: We need to show that if $t_1>t_2\geq 0$ then $\varphi_{t_1}\leq 
1946: \varphi_{t_2}$. If $\varphi_t$ vanishes at some point $t=t_0$ then, as it is
1947: easy to verify, the union $A_{t_0}\cup B_{t_0}$ is convex, and $\varphi_t=0$
1948: for any $t>t_0$. We may therefore assume that $\varphi_t$ is strictly positive
1949: in our segment $[t_2,t_1]$. 
1950: 
1951: Fix a common interior point 
1952: $$
1953:  y\in\textrm{int}(A)\cap\textrm{int}(B).
1954: $$
1955: Now 
1956: $$
1957:  \varphi_t=\sup_{x\in \partial (A)_t \cap \partial (B)_t}
1958:  \angle \big(\hat n_{(A)_t}(x),\hat n_{(B)_t}(x)\big) =
1959:  \sup_{R\subset S}\{\max_{x\in R\cap\partial (A)_t \cap \partial (B)_t} 
1960:  \angle \big(\hat n_{(A)_t}(x),\hat n_{(B)_t}(x)\big)\}
1961: $$
1962: where $R$ runs over the (compact) balls centered at $y$.
1963: It is therefore enough to show that 
1964: $$
1965:  \varphi_{t_2}\geq\max_{x\in R\cap\partial (A)_{t_1} \cap \partial (B)_{t_1}} 
1966:  \angle \big(\hat n_{(A)_{t_1}}(x),\hat n_{(B)_{t_1}}(x)\big)
1967: $$
1968: for any such $R$.
1969: 
1970: Fix $R$ large enough (so that the intersection 
1971: $R\cap\partial (A)_{t_1} \cap \partial (B)_{t_1}$ is not empty)
1972: and let 
1973: $$
1974:  \tilde\varphi_t=\max_{x\in R\cap\partial (A)_t \cap 
1975:  \partial (B)_t} \angle \big(\hat n_{(A)_t}(x),\hat n_{(B)_t}(x)\big).
1976: $$
1977: Since $\tilde\varphi_t$ is obviously continuous,
1978: it is enough to show that for any $t_2 < t \leq t_1$ we have 
1979: $\tilde\varphi_{t-\gD t}\geq\tilde\varphi_t$ whenever $\gD t$ is small enough.
1980: Fix $t$, and let $\gD t$ be sufficiently small so that the argument below holds.
1981: 
1982: To simplify notation we replace $A$ (resp. $B$) by $(A)_{t-\gD t}$ 
1983: (resp. $(B)_{t-\gD t}$) and assume that $t-\gD t=0$ (this is just a matter of
1984: changing names after $t$ and $\gD t$ are fixed). 
1985: Let $x_t\in R\cap\partial (A)_t \cap\partial (B)_t$ be such that 
1986: $$
1987:  \angle \big(\hat n_{(A)_t}(x_t),\hat n_{(B)_t}(x_t)\big) =\tilde\varphi_t.
1988: $$
1989: It is enough to show that there is $x_0\in R\cap\partial A \cap \partial B$
1990: with
1991: $$
1992:  \angle \big(\hat n_A(x_0),\hat n_B(x_0)\big)\geq
1993:  \angle \big(\hat n_{(A)_t}(x_t),\hat n_{(B)_t}(x_t)\big).
1994: $$
1995: 
1996: Let $x_0\in \partial A \cap \partial B$ be a point at the minimal possible distance from $x_t$.
1997: Denote by $\hat u_0$ (resp. $\hat u_t$) the tangent to the geodesic line
1998: $\overline{x_0,x_t}$ at $x_0$ (resp. at $x_t$).
1999: 
2000: Since $\gD t$ is arbitrarily small, $\hat u_t$ is roughly in the direction of
2001: the bisector of the angle between $\hat n_{(A)_t}(x_t)$ and 
2002: $\hat n_{(B)_t}(x_t)$, and $x_0$ is closer to $y$ than $x_t$. In particular 
2003: $x_0\in R$.
2004: 
2005: By the Lagrange multipliers theorem $\hat u_0$ is a linear combination of
2006: $\hat n_A(x_0)$ and $\hat n_B(x_0)$ and since $\gD t$ is arbitrarily small, we
2007: can assume that $\hat u_0$ is in the convex cone spanned by   
2008: $\hat n_A(x_0)$ and $\hat n_B(x_0)$ (again, in the limit case $\hat u_0$
2009: is the direction of the bisector of the angle between
2010: $\hat n_A(x_0),\hat n_B(x_0)$). Thus 
2011: $$ 
2012:  \angle \big(\hat n_A(x_0),\hat n_B(x_0)\big) =
2013:  \angle \big(\hat n_A(x_0),\hat u_0\big) +
2014:  \angle \big(\hat u_0,\hat n_B(x_0)\big),
2015: $$
2016: and since (by the triangle inequality for angles) 
2017: $$
2018:  \angle \big(\hat n_{(A)_t}(x_t),\hat n_{(B)_t}(x_t)\big)\leq
2019:  \angle \big(\hat n_{(A)_t}(x_t),\hat u_t\big)+
2020:  \angle \big(\hat u_t,\hat n_{(B)_t}(x_t)\big),
2021: $$
2022: it is enough to show that 
2023: $$
2024:  \angle \big(\hat n_A(x_0),\hat u_0\big)\geq 
2025:  \angle \big(\hat n_{(A)_t}(x_t),\hat u_t\big)
2026: $$
2027: (and the analogous inequality for $B$ instead of $A$ whose proof is the same).
2028: 
2029: Let $c(s)$ be the geodesic line of unit speed 
2030: with $c(0)=x_0,\dot c(0)=\hat u_0$, and let
2031: $c(s_0)=x_t$ (i.e. $s_0=d(x_0,x_t)$). Then the above inequality follows from
2032: the following
2033: $$
2034:  \hat u_0\cdot \hat n_A(x_0)=\frac{d}{ds}|_{s=0}D_A\big( c(s)\big)\leq 
2035:  \frac{d}{ds}|_{s=s_0}D_A\big( c(s)\big)=
2036:  \hat u_t \cdot \hat n_{(A)_t}(x_t)
2037: $$
2038: and this follows from the convexity of the function $D_A$.
2039: \end{proof}
2040: 
2041: The following lemma, explains how we can use the information obtained above,
2042: in order to find a direction with respect to which the directional derivatives
2043: of all the corresponding distance functions are large.
2044: 
2045: \begin{lem}{\bf (Existence of a good direction):}\label{b(d)}
2046: There is a constant $b(d)>0$ such that for any set of unit vectors
2047: $\{ \hat n_i\}_{i\in I}\subset \BR^d$, that satisfy the condition 
2048: $\hat n_i\cdot \hat n_j \geq 0$ for any $i,j\in I$, there is a unit vector
2049: $\hat f$ such that $\hat f\cdot \hat n_i > \frac{1}{b(d)}$ for any $i\in I$.
2050: \end{lem}
2051: 
2052: \begin{proof}
2053: Let $\gD$ be a maximal $1$-discrete subset of $\{ \hat n_i\}_{i\in I}$.
2054: Then there are at most $\frac{b(d)}{2}$ elements in $\gD$, where $b(d)$ is 
2055: some constant.
2056: For any $\hat n_i \in \{ \hat n_i\}_{i\in I}$ there is 
2057: $\hat n_{i_0}\in \gD$ with $\| \hat n_i-\hat n_{i_0}\| <1$, which implies
2058: $\hat n_i \cdot \hat n_{i_0} >1/2$.
2059: Let 
2060: $$
2061:  \hat f=\frac{\Sigma_{\hat n_j\in \gD} \hat n_j}{\|\Sigma_{\hat n_j\in \gD} 
2062:  \hat n_j\|},
2063: $$
2064: then 
2065: $$
2066:  \hat n_i \cdot \hat f >\hat n_i \cdot \frac{\Sigma_{\hat n_j\in \gD} 
2067:  \hat n_j}{\frac{b(d)}{2}}\geq
2068:  \frac{2}{b(d)}\hat n_i \cdot \hat n_{i_0}\geq \frac{1}{b(d)}.
2069: $$
2070: \end{proof}
2071: 
2072: Summarizing the above discussion together with the arguments of the previous 
2073: section, we conclude the following less general but more practical version of 
2074: Lemma \ref{simplicial-complex}:
2075: 
2076: \begin{thm}\label{nice-boundary}
2077: Let $S$ be a symmetric space of non-compact type, $M=\gC\backslash S$ an $S$-manifold, and
2078: $\gep'\geq\gep >0$. Assume that: 
2079: \begin{itemize}
2080: \item 
2081: $M'\subset M_{\geq\gep}$ is a submanifold with boundary
2082: for which the complement of the pre-image in the universal covering is given
2083: by 
2084: $$
2085:  \ti X=S\setminus\ti M'=\cup_{\gc\in\gC'}\{d_\gc <\gep'\},
2086: $$
2087: where $\gC'$ is a subset of $\gC$ which is invariant under conjugation by elements of $\gC$.
2088: \item
2089: $M'$ is homotopically equivalent to its $\frac{\gep}{2}$-shrinking
2090: $)M'(_\frac{\gep}{2}$.
2091: \item 
2092: For any $x\in S$ the group 
2093: $$
2094:  \langle \gc\in\gC':d_\gc (x) \leq 3\gep'\rangle
2095: $$
2096: is either unipotent or commutative.
2097: \end{itemize} 
2098: Then there are positive constants $d,\ga$ which depend on $S,~\gep,\gep'$, 
2099: such that $M'$ is homotopically equivalent to some 
2100: $\big( d,\ga\cdot\vol (M')\big)$-simplicial complex.
2101: \end{thm}
2102: 
2103: \begin{proof}
2104: We shall show that the conditions of Lemma \ref{simplicial-complex} are 
2105: satisfied. 
2106: The sets $\{d_\gc <\gep'\}$ are convex and open with smooth boundary and 
2107: $X$ is the locally finite union of them.\\
2108: 
2109: It is given that $M'$ is contained in the $\gep$-thick part and that it is
2110: homotopic to its $\frac{\gep}{2}$-shrinking.\\
2111: 
2112: We should indicate how to define the constant $b$ and the direction 
2113: $\hat n(x)$, from the conditions of Lemma \ref{simplicial-complex}.
2114: 
2115: We define
2116: $$
2117:  b=\max \{\frac{2}{\gb (\gep')},b(d)\}
2118: $$ 
2119: where $\gb$ is the function defined in the proof of Theorem \ref{thm1} in the 
2120: previous section, and $b(d)$ is the
2121: constant defined in Lemma \ref{b(d)} for $d=\dim S$.
2122: 
2123: Next we define the directions $\hat n(x)$. Let $x\in S\setminus\ti X$ be a 
2124: point with $d(x,\ti X)\leq\gep$. If the group 
2125: $\langle \gc\in\gC':d_\gc (x) \leq 3\gep'\rangle$ is unipotent, then we define
2126: the direction $\hat n(x)$ in the same way as it is done in the previous 
2127: section.
2128: (It is shown in the previous section that in this case the conditions of 
2129: Lemma \ref{simplicial-complex} are satisfied.)
2130: Assume that the group $\langle \gc\in\gC':d_\gc (x) \leq 3\gep'\rangle$ is 
2131: abelian. Let $A(x)\subset \gC'$ be the set of elements $\gc\in \gC'$
2132: with $D_{\{d_\gc <\gep'\}}(x)=d(x,\ti X)$. Then for $\gc\in A(x)$
2133: $$
2134:  d_\gc (x)\leq 2d(x,\ti X)+\gep'\leq 3\gep',
2135: $$
2136: and hence $A(x)$ is contained in 
2137: $\langle \gc\in\gC':d_\gc (x) \leq 3\gep'\rangle$. Thus $A(x)$ is abelian.
2138: By Lemma \ref{bigA}, for any $\ga ,\gb\in A(x)$ and any 
2139: $y\in\partial \{d_\ga <\gep'\}\cap\partial \{d_\gb <\gep'\}$, the inner 
2140: product of the corresponding external normal vectors at $y$ is non-negative. 
2141: By Lemma \ref{monoton},
2142: the inner product at $T_x(S)$ of the external normal vectors to
2143: $\partial (\{d_\ga <\gep'\})_{d(x,\ti X)}$ and
2144: $\partial (\{d_\gb <\gep'\})_{d(x,\ti X)}$ at $x$, is also non-negative. 
2145: Therefore it follows from Lemma \ref{b(d)} there is a direction $\hat f\in T_x(S)$ 
2146: for which the inner product of $\hat f$ with the external normal vector at $x$ to
2147: $\partial (\{d_\gc <\gep'\})_{d(x,\ti X)}$ is $\geq\frac{1}{b(d)}$ for any 
2148: $\gc\in A(x)$.
2149: Thus, we can take $\hat n (x)=\hat f$.
2150: \end{proof}
2151: 
2152: 
2153: 
2154: 
2155: 
2156: 
2157: %NEW SECTION
2158: 
2159: 
2160: 
2161: 
2162: 
2163: \section{The proof of \ref{thmA}(2) in the rank-1 case}\label{rk-1}
2164: 
2165: Theorem \ref{thmA}(2) can be proved, independently of the rank, using the argument
2166: that we shall present in section \ref{pfBC}. 
2167: For two reasons we chose to do the rank-1
2168: case separately. The first reason is that in the remaining higher rank case all
2169: locally symmetric manifolds of finite volume are arithmetic. Since Theorem
2170: \ref{thmA} already took care of the non-compact arithmetic case, we can assume compactness when 
2171: considering higher rank manifolds. This would make the argument in section \ref{9} simpler.
2172: The second reason is that in the rank-1 case there is another proof, which is
2173: in some sense more simple. This proof, which we shall present below, is 
2174: basically the same as the one given in \cite{BGLM} for the hyperbolic case, and with the tools
2175: developed in sections 3,4,6 and 7, it could be applied to general rank-1 symmetric spaces which
2176: are not necessarily of constant curvature.
2177: 
2178: \begin{rem}
2179: For $S=\BH^2$ the statement of Conjecture \ref{conjA} follows 
2180: from the Gauss Bonnet theorem without the arithmeticity assumption.
2181: This is because the volume determines the genus and bound the possible number of
2182: cusps. For this reason we allow 
2183: ourselves to ignore this case in this and in the following section.
2184: \end{rem}
2185: 
2186: Recall what we intend to prove:
2187: 
2188: \begin{thm}[{\bf Theorem \ref{thmA}(2) for the rank-1 case}]\label{rk-1thm}
2189: If $S$ is a rank-$1$ symmetric space of dimension $\geq 4$, then for some 
2190: constants $\ga ,d$, the fundamental group $\pi_1(M)$ of any 
2191: $S$-manifold $M=\gC\backslash S$ is isomorphic to the fundamental group of some
2192: $\big( d,\ga\cdot\vol (M)\big)$-simplicial complex.
2193: \end{thm}
2194: 
2195: We shall use the ordinary thick-thin decomposition. 
2196: Let $\gep_s$ be the constant from the Margulis lemma, and let 
2197: $\gep =\frac{\gep_s}{2}$.
2198: 
2199: \begin{thm}[{\bf Thick-thin decomposition in rank-1}](see \cite{Th} section 4.5, and \cite{BGS} 
2200: section 10)\label{thick-thin} 
2201: Assume that $\rk (S)=1$. Let $M$ be an $S$-manifold of finite volume and $M_{\leq \epsilon}^0$ an 
2202: arbitrary connected component of $M_{\leq \epsilon}$. Then $M_{\leq \epsilon}^0$ belongs to either 
2203: one of the following two types:
2204: \begin{enumerate}
2205: \item 
2206: A tube, i.e. a tubular neighborhood of a short geodesic. In this case 
2207: $M_{\leq \epsilon}^0$ is topologically a ball-bundle 
2208: over the circle. Its fundamental group $\pi_1(M_{\leq \epsilon}^0)$ is 
2209: infinite cyclic, and in particular abelian.
2210: 
2211: or
2212:  
2213: \item  
2214: A cusp. In which case $M_{\leq \epsilon}^0$ is homeomorphic to 
2215: $\BR^{\geq 0}\times \partial M_{\leq \epsilon}^0$, where 
2216: $\partial M_{\leq \epsilon}^0$ is topologically a sub-manifold of 
2217: codimension $1$.
2218: A connected component of its pre-image  $\tilde  M_{\leq \epsilon}^0$ is
2219: given by $\tilde  M_{\leq \epsilon}^0=
2220: \cup_{\gc \in \gC_0}\{ d_\gc \leq \epsilon \}$ where $\gC_0\leq\gC$ is a 
2221: subgroup isomorphic to $\pi_1(M_{\leq\gep}^0)$, and it is a ``star-shaped''
2222: neighborhood of some point $z\in S(\infty )$ (i.e. each geodesic line with
2223: $c(\infty )=z$ enters once into $\tilde M_{\leq \epsilon}^0$ and stays in it).
2224: $\gC_0$ consists of unipotent elements only,
2225: and each $\gc \in \gC_0$
2226: fixes $z$ and leaves the horospheres around $z$ invariant.
2227: Moreover $\gC_0$ is metabelian.
2228: \end{enumerate}
2229: \end{thm}
2230: 
2231: There are only finitely many connected components of $M_{\leq\gep}$.
2232: Moreover, as $\dim (M)\geq 3$ the boundary of each 
2233: connected component of $M_{\leq\gep}$ is connected, and hence the thick 
2234: part  $M_{\geq\gep}$ is connected. Each cusp is homotopically equivalent to its
2235: boundary. As $\dim (M)\geq 4$ each tube is a ball bundle over the circle,
2236: with fibers of dimension $\geq 3$, and hence, its boundary is a sphere bundle
2237: for a sphere of dimension $\geq 2$. Therefore, for each connected component of
2238: the thin part, the injection of the boundary into the component induces an
2239: isomorphism between the fundamental groups (which are both $\cong\BZ$). 
2240: Thus by Van-Kampen's theorem we obtain:
2241: 
2242: \begin{cor}
2243: $$
2244:  \pi_1(M)\cong\pi_1(M_{\geq\gep}).
2245: $$
2246: \end{cor}
2247: 
2248: Take $M'=M_{\geq\gep}$. Then the following claim 
2249: finishes the proof of Theorem \ref{rk-1thm}.
2250: 
2251: \begin{clm}
2252: $M'$ satisfies the condition of Theorem \ref{nice-boundary}, with respect 
2253: to $\gep'=\gep$.
2254: \end{clm}
2255: 
2256: \begin{proof}
2257: We only need to explain why $M'$ is homotopic to its 
2258: $\frac{\gep}{2}$-shrinking, since the other conditions of Theorem
2259: \ref{nice-boundary} follow directly from the definition of $M'$ and from Theorem \ref{thick-thin}. 
2260: 
2261: Since $\gep =\frac{\gep_s}{2}$, the $\frac{\gep}{2}$-neighborhoods of different 
2262: connected components of
2263: $M_{\leq\gep}$ are still disjoint, and hence, we can prove the homotopy 
2264: equivalence by
2265: showing that there is a deformation retract from 
2266: $\overline{(M_{\leq\gep}^0)_\frac{\gep}{2}}\setminus M_{\leq\gep}^0$ to
2267: $\partial (M_{\leq\gep}^0)_\frac{\gep}{2}$, for each connected component 
2268: $M_{\leq\gep}^0$ of $M_{\leq\gep}$.
2269: 
2270: If $M_{\leq\gep}^0$ is a cusp, then one can define such a deformation retract 
2271: by letting each point flow (at constant speed = the initial distance) along the 
2272: unique geodesic line which connects it to the unique end of the cusp, i.e. along
2273: the geodesic whose lifting in the universal covering converges, when $t\to -\infty$, to the unique
2274: limit point in $S(\infty )$ of a lifting of the cusp.
2275: 
2276: If $M_{\leq\gep}^0$ is a tube, then the lifting of $M_{\leq\gep}^0$ and 
2277: its $\frac{\gep}{2}$-neighborhood in $S=\ti M$ are both star-shaped with 
2278: respect to the lifting $c$ of the short closed geodesic which lies inside $M_{\leq\gep}^0$. 
2279: The generalized star-contraction (see \ref{GSC}) from 
2280: $\overline{(\ti M_{\leq\gep}^0)_\frac{\gep}{2}}\setminus \ti M_{\leq\gep}^0$ to
2281: $\partial (\ti M_{\leq\gep}^0)_\frac{\gep}{2}$ 
2282: projects to a deformation retract of the corresponding subsets of $M$. 
2283: \end{proof}
2284: 
2285: 
2286: 
2287: 
2288: 
2289: 
2290: 
2291: %NEW SECTION
2292: 
2293: 
2294: 
2295: 
2296: \section{The proofs of \ref{thmA}(2) and \ref{thmA}(3)}\label{pfBC}
2297: 
2298: We shall use the fact that when a low dimensional submanifold is 
2299: removed from a high dimensional manifold, it doesn't change the low dimensional
2300: homotopy groups. More precisely:
2301: 
2302: \begin{lem}\label{M-N}
2303: Let $M$ be a connected manifold and $N\subset M$ a closed (not 
2304: necessarily connected) sub-manifold.
2305: \begin{itemize}
2306: \item If $\text{codim}_M(N)\geq 2$ then there is a surjective homomorphism
2307: $$
2308:  \pi_1(M\setminus N)\to\pi_1(M).
2309: $$
2310: \item If $\text{codim}_M(N)\geq 3$ then $\pi_1(M\setminus N)\cong\pi_1(M)$.
2311: \end{itemize}
2312: \end{lem}
2313: 
2314: \begin{proof}
2315: If $\text{codim}_M(N)\geq 2$ then any closed loop can be pushed to one which 
2316: dose not intersect $N$. Similarly, if $\text{codim}_M(N)\geq 3$ then any 
2317: homotopy of loops can be pushed to $M\setminus N$.
2318: \end{proof}
2319: 
2320: The submanifold we are about to remove consists
2321: of the union of all short closed geodesics of some certain type. Observe that
2322: the lifting to $S=\ti M$ of a closed geodesic in $M=\gC\backslash S$ which corresponds to
2323: $\gc\in\gC$ is an axis of $\gc$. In particular, any two closed geodesics from
2324: the same homotopy class are parallel.
2325: 
2326: \begin{lem}\label{codimN}
2327: Let $\gc\in G^0$ be a hyperbolic isometry which projects non-trivially to each 
2328: simple factor of $G^0$, and let $\ti N\subset S$ be the union of all geodesics 
2329: which are axes of $\gc$.
2330: 
2331: \begin{itemize}
2332: \item If $S$ is not isometric to $\BH^2$ then $\text{codim}_S(\ti N)\geq 2$.
2333: \item If additionally $S$ is neither isometric to $\BH^3, \BH^2\times\BH^2$ nor to
2334: $\PSL_3(\BR )/\PSO_3(\BR )$, then $\text{codim}_S(\ti N)\geq 3$.
2335: \end{itemize}
2336: \end{lem}
2337: 
2338: \begin{proof}
2339: Let $S^*$ be an irreducible factor of $S$. Denote by $\gc^*$ the projection of 
2340: $\gc$ to the corresponding simple factor $G^*$ of $G^0$.
2341: We need to estimate $\text{codim}_{S^*}\big(\min (\gc^*)\big)$. 
2342: 
2343: If $\gc^*$ fixes a point then we just remark that since the projection $\gc^*$
2344: is non-trivial, 
2345: $\text{codim}_{S^*}\big(\min (\gc^*)\big) >0$. Otherwise let $c$ be an axis
2346: of $\gc^*$ in $S^*$ and let
2347: $$
2348:  G^*(c) = \{ g\in G^*:g\cdot c \text{~is parallel to~} c\}.
2349: $$
2350: An alternative way to define $G^*(c)$ is as follows. Pick two points
2351: $p,q\in c(\BR )$ and let $g_{p,q}=\gs_p\cdot\gs_q$ be the corresponding 
2352: transvection,  then $G^*(c)$ coincides with the centralizer group 
2353: $C_{G^*}(g_{p,q})\leq G^*$ of $g_{p,q}$.
2354: $G^*(c)$ is a closed reductive Lie subgroup of $G^*$ which acts 
2355: transitively on $N^*=$ the union of all geodesics parallel to $c$ in $S^*$. 
2356: Observe that $N^*$ contains the projection of $\ti N$ to $S^*$.
2357: 
2358: Choose an Iwasawa decomposition of $G^*$ which induces an Iwasawa 
2359: decomposition of $G^*(c)^0$, i.e. write the Iwasawa decompositions of
2360: $G^*$ and $G^*(c)$ simultaneously 
2361: $$
2362:  G^*=K\cdot A\cdot U, \text{~and~} G^*(c)^0=K(c)\cdot A\cdot U(c),
2363: $$
2364: such that $K(c)=K\cap G^*(c)^0$ and $U(c)=U\cap G^*(c)^0$.
2365: (This could be done by choosing a flat $F\supset c$ and defining the torus $A$
2366: to be the subgroup of $G^*$ which acts on $F$ by translations, and then 
2367: choosing an order on the dual of $\mathfrak{a}=\text{Lie}(A)$ such that if 
2368: $c(t)$ is given by $c(t)=e^{tH}$ for $H\in\mathfrak{a}$, then $\ga (H)\geq 0$ 
2369: for any simple 
2370: root $\ga$, and then taking the Iwasawa decomposition which corresponds to 
2371: $A$ with this ordering.) 
2372: 
2373: Then $\dim (S^*)=\dim (A\cdot U)$ while $\dim (N^*)=\dim \big( A\cdot U(c)\big)$,
2374: and hence
2375: $$
2376:  \text{codim}_{S^*}(N^*)=\text{codim}_U\big( U(c)\big).
2377: $$
2378: Let $P_{c(-\infty )}$ be the parabolic subgroup of $G^*$ which corresponds to
2379: the point $c(-\infty )\in S^*(\infty )$, and let $P^-$ be the minimal parabolic
2380: opposite to $P^+=AU$. Then, as it is easy to verify, the Lie algebra of $G^*$
2381: is the direct sum
2382: $$
2383:  \text{Lie}(G^*)=\text{Lie}(U)+\text{Lie}(P^-),
2384: $$
2385: while the Lie algebra of the parabolic $P_{c(-\infty )}$ is the direct sum
2386: $$
2387:  \text{Lie}\big( P_{c(-\infty )}\big)=
2388:  \text{Lie}\big( U(c)\big)+\text{Lie}(P^-).
2389: $$
2390: Hence $\text{codim}_U\big( U(c)\big)=\text{codim}_{G^*}(P_{c(-\infty )})$.
2391: The lemma follows from the following proposition by a standard
2392: case analysis.
2393:   
2394: \begin{prop}[see \cite{BaNe} lemma 3.4 and corollary 8]
2395: Let $H$ be a connected simple Lie group and $P\leq H$ a proper closed connected 
2396: subgroup. Then $\text{codim}_H(P)\geq\rk (H)$, and the equality can hold only
2397: if $H$ is locally isomorphic to $\SL_n(\BR )$.
2398: \end{prop}
2399: \end{proof}
2400: 
2401: We shall now prove: 
2402: 
2403: \begin{thm}[{\bf Theorem \ref{thmA}(2) from the introduction
2404: in the general case}]\label{thm2}
2405: Let $S$ be a symmetric space of non-compact type, not isometric to
2406: $\BH^2,\BH^3, \BH^2\times\BH^2, \PSL_3(\BR )/\PSO_3(\BR )$.
2407: Then there are constants $\ga ,d$ such that the fundamental group of any
2408: irreducible $S$-manifold of finite volume $M$ is isomorphic to the fundamental 
2409: group of some $\big( d,\ga\cdot\vol (M)\big)$-simplicial complex.
2410: \end{thm}
2411: 
2412: \begin{proof}[Proof of Theorem \ref{thm2}] 
2413: Let $G=\text{Isom}(S)$, let $r$ be the rank of $S$, i.e. the real rank of $G$, let $n_s$ be the
2414: index from the Margulis lemma and let $n=n_s!$. Fix $\gep$ to be one third of 
2415: the $\gep_s$ from the Margulis lemma.
2416: 
2417: Since we already proved the theorem in the rank one case, we can assume that $\rk (S)\geq 2$.
2418: Then any $S$-manifold is arithmetic. Moreover, as Theorem \ref{thm1} implies Theorem \ref{thm2}
2419: for non-compact arithmetic manifold, we may assume that $M=\gC\backslash S$ is compact.
2420: In this case, we have the following strengthening of the Margulis lemma.
2421: 
2422: \begin{lem}
2423: For any $x\in S$, the group $\gC_{\gep_s} (x)$ contains an abelian subgroup of
2424: index $n_s$.
2425: \end{lem}
2426: 
2427: \begin{proof}
2428: By the Margulis lemma $\gC_{\gep_s} (x)$ contains a subgroup 
2429: $\gC_{\gep_s}^0 (x)$ of index $n_s$ which is contained in a connected nilpotent
2430: Lie subgroup of $G$. By Lie's theorem $\gC_{\gep_s}^0 (x)$ is triangulable
2431: over $\BC$. Therefore its commutator $[\gC_{\gep_s}^0 (x),\gC_{\gep_s}^0 (x)]$
2432: contains only unipotents. As $\gC$ is cocompact it has no 
2433: non-trivial unipotents. Thus $\gC_{\gep_s}^0 (x)$ is abelian.
2434: \end{proof}
2435: 
2436: It follows that when $\gc_1 ,\gc_2\in\gC$ satisfy 
2437: $\{ d_{\gc_1}\leq\gep\}\cap\{ d_{\gc_2}\leq\gep\}\neq\emptyset$ then $\gc_1^n$
2438: and $\gc_2^n$ commute.
2439: 
2440: Consider an element $\gc\in\gC$. Replacing $\gc$ by $\gc^{[G:G^0]}$, we may
2441: assume that $\gc\in G^0$. As $\gC\cap G^0$ is a uniform 
2442: lattice in $G^0$, $\gc$ is semisimple.
2443: Consider the sequence of centralizers
2444: $$
2445:  C_G(\gc)\leq C_G(\gc^n)\leq C_G(\gc^{n^2})\leq C_G(\gc^{n^3})\leq\ldots
2446:  \leq C_G(\gc^{n^r}).
2447: $$
2448: As the centralizer of a semisimple element is determined by the type of the 
2449: singularity of the element and by a torus which contains it, two consecutive
2450: terms $C_G(\gc^{n^i})$ and $C_G(\gc^{n^{i+1}})$ in this sequence 
2451: must coincide. Take $i$ to be the first time at which this happens and write 
2452: $\gc'=\gc^{n^i}$. In this way we attach $\gc'$ to each $\gc\in\gC$.
2453: Notice that if $\gc\in G^0$ then $\gc'=\gc^j$ for some $j\leq n^r$, 
2454: and in general $\gc'=\gc^j$ for some $j\leq [G:G^0]n^r$. Set $m=[G:G^0]n^r$.
2455: 
2456: The reason we prefer to work with the $\gc'$'s is the following. If 
2457: $$
2458: \{d_{\gc_1'}\leq 3\gep\}\cap\{d_{\gc_2'}\leq 3\gep\}\neq\emptyset
2459: $$ 
2460: then $\gc_1'$ and $\gc_2'$ commute. Of course, by the above lemma, the 
2461: non-empty 
2462: intersection implies that $\gc_1'^n,\gc_2'^n$ commute. Since $\gc_1'$ has the 
2463: same centralizer as $\gc_1'^n$, this implies that $\gc_1',\gc_2'^n$ commute, 
2464: and since $\gc_2'$ has the same centralizer as $\gc_2'^n$,  
2465: this implies that $\gc_1',\gc_2'$ commute.
2466: 
2467: Let $N\subset M$ be the subset which consists of the union of all
2468: closed geodesics of length $\leq\gep$ which correspond to elements of the
2469: form $\gc'$ for $\gc\neq 1$. In other words, its pre-image in $S$ is given by 
2470: $$
2471:  \ti N=\cup\{\min (\gc' ):\gc\in\gC\setminus\{ 1\} ,\min (d_{\gc'})\leq\gep\}.
2472: $$ 
2473: Then $N$ is a finite union of totally geodesic closed submanifolds. 
2474: Since $\gC$ is irreducible and $G$ is center free, any $\gc\in\gC\setminus\{ 1\}$ 
2475: projects non-trivially to each factor of $G$. So it follows from
2476: Lemma \ref{codimN} that $\text{codim}_M(N)\geq 3$, and hence, by Lemma 
2477: \ref{M-N}
2478: $$
2479:  \gC\cong\pi_1(M)\cong\pi_1(M\setminus N).
2480: $$
2481: 
2482: Define
2483: $$
2484:  \ti X=\cup_{\gc\in\gC\setminus\{ 1\}}\{ d_{\gc'}<\gep\},~~
2485:  \ti M'=S\setminus \ti X,
2486: $$
2487: and 
2488: $$
2489:  X=\gC\backslash\ti X,~~M'=\gC\backslash\ti M'=M\setminus X.
2490: $$
2491: 
2492: Clearly $N\subset X$.
2493: In order to prove the theorem we shall show:
2494: \begin{enumerate}
2495: \item 
2496: $M\setminus N$ and $M'$ are homotopically equivalent.
2497: \item 
2498: $M'$ satisfies the conditions of Theorem \ref{nice-boundary}
2499: with respect to $\gep$ and $\frac{\gep}{m}$ (corresponding to $\gep'$ and 
2500: $\gep$ respectively in \ref{nice-boundary} ).
2501: \end{enumerate}
2502: 
2503: {\bf Proof of (1):}
2504: We shall construct the desired homotopy in two steps.
2505: Since $M$ and $N$ are compact, and each connected component of $N$ is a 
2506: finite union of totally geodesic submanifolds, there is small positive number 
2507: $\eta>0$, such that the 
2508: $\eta$-neighborhoods of the components of $N$ are still disjoint and contained 
2509: in $X$. It is easy to verify that if $\eta$ is small,
2510: $M\setminus N$ and $M\setminus \overline{(N)_\eta}$ are diffeomorphic.
2511: 
2512: Next, we claim that there is a deformation retract from 
2513: $M\setminus \overline{(N)_\eta}$ to $M'$. In order to show this, we shall
2514: apply Lemma \ref{deformation-retract} to the 
2515: set of functions 
2516: $$
2517:  \mathcal{F}=\{d_{\gc'}:\gc\in\gC\setminus\{1\},\min (d_{\gc'})<\gep\}
2518: $$ 
2519: on $Y=L=S\setminus\overline{(\ti N)_\eta}$, and $M'=\mathcal{F}_{\geq\gep}$. 
2520: We have to indicate what are the directions 
2521: $\hat n(x)$, and what is the continuous function $\gb$.
2522: 
2523: Let $x\in \ti X\setminus\overline{(\ti N)_\eta}$, and let 
2524: $\{ d_{\gc'_1},\ldots ,d_{\gc'_k}\}\subset \mathcal{F}$ be the subset of
2525: functions satisfying $d_{\gc'_i}(x)\leq 3\gep$. Then the group
2526: $\langle \gc'_1,\ldots ,\gc'_k\rangle$ is abelian. Thus for each $j\leq k$
2527: the convex set $\cap_{i=1}^{j-1}\min (\gc'_i)$ is $\gc'_j$ invariant, 
2528: and hence, by induction 
2529: $$
2530: \cap_{i=1}^j\min (\gc'_i)\neq\emptyset.
2531: $$
2532: Pick arbitrarily $y\in\cap_{i=1}^k\min ({\gc'_i})$ and define 
2533: $\hat n(x)$ to be the tangent at $x$ 
2534: to the geodesic line which goes from $y$ to $x$.
2535: 
2536: For $t\leq 3\gep$ we define $\gb (t)$ to be the minimum of the directional 
2537: derivative of $d_{\gc'}$ at $x$ with respect to the tangent to the geodesics 
2538: $\overline{z,x}$, where this minimum is taken over all 
2539: $x\in\ti X\setminus (\ti N)_\eta$, all $d_{\gc'}\in \mathcal{F}$ with 
2540: $d_{\gc'}(x)=t$, and all
2541: $z\in\min (\gc' )$.
2542: Since 
2543: \begin{itemize}
2544: \item
2545: $N$ and $M\setminus (N)_\eta$ are compact, 
2546: \item
2547: up to conjugations there are only finitely many $\gc'$'s with 
2548: $\min (d_{\gc'})<\gep'$, and 
2549: \item 
2550: for any selection
2551: of $x,z,\gc'$ as above, the corresponding directional derivative is positive,
2552: \end{itemize}
2553: it follows that $\gb$ is a well defined
2554: continuous positive function.\\
2555: 
2556: {\bf Proof of (2):}
2557: We shall check that the conditions of Theorem
2558: \ref{nice-boundary} are satisfied.
2559: 
2560: Let $x\in S$ and assume $d_\gc (x)\leq\frac{\gep}{m}$
2561: for some $\gc\neq 1$ in $\gC$. As $\gc'=\gc^j$ for some $j\leq m$, it follows
2562: that 
2563: $$
2564: d_{\gc'}(x)=d_{\gc^j}(x)\leq j\cdot d_\gc (x)\leq\frac{j\cdot\gep}{m}\leq\gep.
2565: $$
2566: Thus $x\in\ti X$. This shows that $M'$ is contained in 
2567: $M_{\geq\frac{\gep}{m}}$.
2568: 
2569: If $\{ d_{\gc_1'}\leq 3\gep\}\cap\{ d_{\gc_2'}\leq 3\gep\}$ then 
2570: $\gc'_1$ commutes with $\gc'_2$. Thus, the last condition of Theorem
2571: \ref{nice-boundary} is also satisfied.
2572: 
2573: We shall show that there is a deformation retract from $M'$ to its 
2574: $\frac{\gep}{2m}$-shrinking, by an analogous way to the second step in the 
2575: proof of (1) above. 
2576: This time take
2577: $$
2578:  \mathcal{F}=\{D_{\{ d_{\gc'}\leq\gep\}}:\gc\in\gC\setminus\{1\},\min (d_{\gc'})<\gep\}
2579: $$ 
2580: while $\frac{\gep}{2m}$ plays the role of $\gep$ in Lemma 
2581: \ref{deformation-retract}.
2582: We define the directions $\hat n(x)\in T_x(S)$ 
2583: to be the tangent to the geodesic $\overline{y,x}$ for arbitrary
2584: $y\in\cap\{\min (\gc' ):d_{\gc'}\in\Psi_{x,3\gep}\}$. 
2585: Additionally, we define
2586: $\gb (t)$ to be the minimum of the directional 
2587: derivative of $D_{\{d_{\gc'}\leq\gep\}}$ at $x$ with respect to the tangent 
2588: of the geodesics $\overline{z,x}$, where the minimum is taken over all 
2589: $x\in S$, all $D_{\{d_{\gc'}\leq\gep\}}\in \mathcal{F}$ with 
2590: $D_{\{d_{\gc'}\leq\gep\}}(x)=t$, and all
2591: $z\in\min (\gc' )$. Since for $t\geq 0$ and for a relevant $\gc'$, the set
2592: $\{ D_{\{d_{\gc'}\leq\gep\}}\leq t\}$ is a convex body with smooth boundary and since
2593: any $z$ as above belongs to the interior of this set, the directional
2594: derivatives mentioned above are always positive. By compactness we get that
2595: $\gb$ is a continuous positive function defined for any $0\leq t\leq\frac{3\gep}{2m}$.
2596: This finishes the proof of Theorem \ref{thm2} 
2597: \end{proof}
2598: 
2599: Next we prove:
2600: 
2601: \begin{thm}[{\bf Theorem \ref{thmA}(3) of the introduction}]\label{thm3}
2602: For any $S$, there are constants $\ga ,d$, such that the fundamental group
2603: of any irreducible $S$-manifold $M$ is isomorphic to a quotient of the fundamental group 
2604: of some $\big( d,\ga\cdot\vol (M)\big)$-simplicial complex. 
2605: \end{thm}
2606: 
2607: \begin{proof}
2608: There are only $3$ cases left to deal with. $\PSL_3(\BR )/\PSO_3(\BR ),
2609: \BH^2\times\BH^2$ and $\BH^3$.
2610: 
2611: The proof for the first two cases goes verbatim as the proof of Theorem 
2612: \ref{thm2}, since these cases are of higher rank and we can assume compactness.
2613: The only difference is that in these cases, we have only $\text{codim}_M(N)\geq 2$ (instead of $\geq 3$) by 
2614: the first part of Lemma \ref{codimN}, 
2615: so the result follows from the first part of Lemma \ref{M-N}.
2616: 
2617: For hyperbolic $3$-manifolds, one should only throw out finitely many circles
2618: which are all the closed geodesics of length $\leq\gep$ and then prove that
2619: what is left is homotopically equivalent to the $\gep$-thick part. This is 
2620: done by deforming each cusp to its boundary as in section 
2621: \ref{rk-1}, and deforming each compact component minus a circle to its 
2622: boundary, as it is done in the proof of \ref{thm2}. 
2623: Then, again, Theorem \ref{nice-boundary} finishes the proof. 
2624: \end{proof}
2625: 
2626: 
2627: 
2628: 
2629: 
2630: 
2631: %new section
2632: 
2633: 
2634: 
2635: 
2636: 
2637: 
2638: 
2639: \section{Some remarks on Conjecture \ref{conjA}, Theorem \ref{thmA} and their relations to algebraic number 
2640: theory}\label{9} 
2641: 
2642: Let $p(x)\in\BZ [x]$ be an integral monic polynomial, and let
2643: $$
2644:  p(x)=\prod_{i=1}^k(x-\ga_i)
2645: $$ 
2646: be its factorization into linear factors over $\BC$. Denote by $m(p)$ its
2647: {\bf exponential Mahler measure}
2648: $$
2649:  m(p)=\prod_{|\ga_i|>1}|\ga_i|.
2650: $$
2651: The following is known as Lehmer's conjecture. 
2652: 
2653: \begin{conj}
2654: There exists a constant $\ell >0$ such that if $p(x)$ is an integral 
2655: monic polynomial with $m(P)\neq 1$, then $m(p)>1+\ell$.
2656: \end{conj}   
2657: 
2658: Denote by $d(p)$ the number of roots $\ga_i$ with absolute value $>1$
2659: $$
2660:  d(p)=\#\{ \ga_i:|\ga_i|>1\}.
2661: $$
2662: 
2663: The following conjecture of Margulis 
2664: is weaker than Lehmer's conjecture.
2665: 
2666: \begin{conj}[see \cite{Mar1} $IX$ 4.21]\label{MargulisConj}
2667: There is a function $\ell :\BN\to\BR^{>0}$ such that 
2668: $m(p)\geq 1+\ell \big( d(p)\big)$ for any non-cyclotomic monic
2669: polynomial $p(x)\in\BZ [x]$.
2670: \end{conj}
2671: 
2672: If Conjecture \ref{MargulisConj} is true, then for any 
2673: symmetric space of  non-compact type $S$, the minimal injectivity radius of any compact 
2674: arithmetic $S$-manifold is bounded from below by some positive constant $r=r(S)$ (see also 
2675: \cite{Mar1}, page 322 for a similar statement). 
2676: To see this, we argue as follows: 
2677: Let $G^0$ be the identity component of $G=\textrm{Isom}(S)$. As $G^0$
2678: is center-free we can identify it with its adjoint group 
2679: $\textrm{Ad}(G^0)\leq\textrm{GL}(\mathfrak{g})$.
2680: Let $\gC\leq G^0$ be a torsion-free uniform arithmetic lattice in $G^0$. 
2681: We think of $\gC$ as the intersection of the fundamental group of some compact
2682: arithmetic $S$-manifold with $G^0$.
2683: Since $\gC$ is arithmetic, there is a compact extension 
2684: $G^0\times O$ of $G^0$ and a $\BQ$-rational structure on the Lie algebra 
2685: $\mathfrak{g}\times \mathfrak{o}$ of $G^0\times O$,
2686: such that $\gC$ is the projection to $G^0$ of a lattice $\ti{\gC}$, which is 
2687: contained in $(G^0\times O)_{\BQ}$ and commensurable to the group of integral
2688: points $(G^0\times O)_{\BZ}$ with respect to some $\BQ$-base of
2689: $(\mathfrak{g}\times \mathfrak{o})_{\BQ}$. By changing this $\BQ$-base, we 
2690: can assume that $\ti{\gC}$ is in fact contained in $(G^0\times O)_{\BZ}$.
2691: This means that the characteristic polynomial $p_{\ti{\gc}}$ of any 
2692: $\ti{\gc}\in\ti{\gC}$ is a monic integral polynomial. 
2693: As $\gC$ is discrete and
2694: torsion-free, $m(p_{\ti{\gc}})>1$ for any $\ti{\gc}\in\ti{\gC}$
2695: which projects to a non-trivial element in $\gC$.
2696: Since $O$ is compact, any eigenvalue of $\tilde\gc$ with absolute value 
2697: different from $1$ is also an eigenvalue of its projection $\gc\in G^0$. 
2698: In particular
2699: $$
2700:  m(p_{\gc})= m(p_{\ti\gc})\geq 1+\min_{i\leq\dim G}\ell (i).
2701: $$ 
2702: We conclude that $\gc$ is outside the open set 
2703: $$
2704:  U=\{ g\in G^0: m(p_g)<1+\min_{i\leq\dim G}\ell (i)\}.
2705: $$ 
2706: Clearly $U$ contains any
2707: compact subgroup of $G$, and contains a subset of the form
2708: $$
2709:  \{ g\in G^0:g \textrm{~is semisimple and~} \min d_g < \ti r(S)\}
2710: $$ 
2711: for some positive constant $\ti r(S)$.
2712: Hence the minimal injectivity radius of $\gC\backslash S$ is $\geq \ti r(S)$.
2713: Finally, if $\gC\leq G$ is a torsion free lattice which is not necessarily 
2714: contained in $G^0$ then the minimal injectivity radius of $\gC\backslash S$ is at 
2715: least $\frac{\ti r(S)}{[G:G^0]}=r(S)$.
2716: This implies: 
2717: 
2718: \begin{cor}
2719: Conjecture \ref{MargulisConj} implies Conjecture \ref{conjA} (for compact 
2720: arithmetic manifolds).
2721: \end{cor}
2722: 
2723: \begin{proof}
2724: For a given $M$, choose a maximal $r$-discrete net $\mathcal{C}$, and take $\mathcal{R}$ to be the
2725: simplicial complex which corresponds to the nerve of the cover of $M$ by the $r$-balls whose centers form $\mathcal{C}$.
2726: Then $M$ is homotopic to $\mathcal{R}$ which is a $\big(\frac{\vol (B_{2.5r})}{\vol (B_{r/2})},
2727: \frac{\vol (M)}{\vol (B_r)}\big)$-simplicial complex.
2728: \end{proof}
2729: 
2730: For compact locally symmetric manifolds, the minimal injectivity radius  
2731: equals half of the length of the shortest close geodesic. 
2732: For non compact arithmetic $S$-manifolds we already know the absence of short 
2733: closed geodesics (see Remark \ref{5.7}). Therefore
2734: Conjecture \ref{MargulisConj} implies 
2735: 
2736: \begin{conj}\label{injrarCong}
2737: For any $S$, there exists a constant $l=l(S)$ such that no arithmetic 
2738: $S$-manifold contains a closed geodesic of length $\leq l$.
2739: \end{conj}
2740: 
2741: A simple argument, which uses the fact that for non-cyclotomic monic integral
2742: polynomials $F(x)$ of a fixed degree $k=\dim (G)+\dim (O)$, $m(F)$ is bounded
2743: away from $1$, shows that 
2744: for all the compact arithmetic $S$-manifolds, which
2745: arise by constructions in which the compact extending group $O$ is fixed, 
2746: the infimum on the minimal injectivity radius is positive, and the statement
2747: of Conjecture \ref{injrarCong} holds, independently of the rational structure 
2748: and of the specific choice of a manifold within a commensurability class. 
2749: More precisely:
2750: 
2751: \begin{prop}\label{fixedO}
2752: Given a compact semi-simple Lie group $O$, there is a positive constant
2753: $r=r(S,O)$, such that the minimal injectivity radius is $\geq r$,
2754: for any compact manifold $M=\gC\backslash S$ such that $\gC\cap G^0$ is commensurable to the group 
2755: $\pi_{G^0}\big( (G^0\times O)_\BZ\big)$ for some $\BZ$-structure on 
2756: $\mathfrak{g}\times \mathfrak{o}$.
2757: In particular, any such manifold is homotopically equivalent to a 
2758: $\big(\frac{\vol (B_{2.5r})}{\vol (B_{r/2})},
2759: \frac{\vol (M)}{\vol (B_r)}\big)$-simplicial complex.
2760: \end{prop}
2761:  
2762: A real algebraic integer $\gt >1$ is called a {\bf Salem number} if all its 
2763: conjugates in $\BC$ have absolute value $\leq 1$.
2764: In Conjecture \ref{MargulisConj} it is not even known whether $\ell (1)>0$, i.e.
2765: whether there is a positive gap between $1$ and the set of Salem 
2766: numbers. Sury \cite{sury} showed that the existence of such a gap is equivalent to the 
2767: existence of an identity neighborhood in $\SL_2(\BR )$ which intersects 
2768: trivially any uniform arithmetic lattice. 
2769: Therefore the existence of such a gap implies
2770: a positive infimum on the length of closed geodesics, when considering all the 
2771: arithmetic surfaces.
2772: We shall now discuss the relation between this gap and 
2773: manifolds locally isometric to $S=\textrm{SL}_3(\BR )/\textrm{SO}_3(\BR )$ or to 
2774: $\BH^2\times\BH^2$ (the two cases for which we could not prove the analog of Theorem \ref{thmA}(2)).
2775: 
2776: \begin{clm}
2777: Assume there is a positive gap between $1$ and the set of Salem numbers (i.e. $\ell (1)>0$).
2778: Then the analog of Theorem \ref{thmA}(2) holds also for the symmetric space
2779: $S=\textrm{SL}_3(\BR )/\textrm{SO}_3(\BR )$. 
2780: \end{clm}
2781: 
2782: \begin{proof}
2783: Let $\gt >1$, and assume that $\gc =\textrm{diag}(\gt ,\gt ,\gt^{-2})$
2784: is an element of a lattice $\gC\leq\textrm{SL}_3(\BR )$. One 
2785: can easily compute the eigenvalues of $\textrm{Ad}(\gc )$ : they are 
2786: $\gt^3,\gt^{-3}$ and $1$. As $\gC$ is arithmetic, $\gt$ is a Salem number, 
2787: and thus, by
2788: our assumption, is bounded away from $1$. We conclude that the displacement 
2789: functions $d_\gc$ of such elements are uniformly bounded away from $0$.
2790: Take 
2791: $$
2792:  \gep^* < \text{inf}\{\min d_\gc :\gc =\textrm{diag}(\gt ,\gt ,\gt^{-2}), \gt 
2793:  \text{~is a Salem number}\}
2794: $$
2795: which is also smaller than $\gep_s$ from the Margulis lemma.
2796: 
2797: If $g$ is a hyperbolic element of an arithmetic lattice in $\SL_3(\BR )$ with $\min (d_g )\leq\gep^*$, then
2798: $g$ has no real eigenvalues of multiplicity $2$. It is easy to see that $\min (g)$ is then a flat or a single
2799: geodesic. In particular $\dim\big(\min (g)\big)\leq 2$.
2800: As $\dim (S)=5$, the submanifold $N$ which consists of the union of all closed
2801: geodesics of length $\leq\gep$ has codimension $\geq 3$. 
2802: Therefore we can apply word by word the argument of
2803: the proof of Theorem \ref{thm2} for $S$ with $\gep^*$.
2804: \end{proof}     
2805: 
2806: If moreover $\ell (2)>0$, then the same is true also for the 
2807: second remaining case $S=\BH^2\times\BH^2$. 
2808: 
2809: \begin{clm}
2810: Assume $\ell (1)\cdot\ell (2)>0$. Then the analog of \ref{thmA}(2) holds also for $S=\BH^2\times\BH^2$.
2811: \end{clm}
2812: 
2813: \begin{proof}
2814: In contrast with the situation for $\SL_3(\BR )$, our problem here arises only for {\it regular}
2815: elements $\gc\in\SL_2(\BR )\times\SL_2(\BR )$ (for otherwise
2816: $\min (\gc )$ is a single geodesic and its codimension is $3$). 
2817: But such an element has always
2818: the form $\gc =(\gc_1,\gc_2)$ where $\gc_i\in\SL_2(\BR )$ are diagonalizable
2819: over $\BR$. If $\gc_i$ is conjugate to 
2820: $\text{diag}(\gt_i,\gt_i^{-1}),~\gt_i >1$
2821: then the only possible non-trivial conjugate of $\gt_1$ outside the unit disk
2822: is $\gt_2$.
2823: Therefore, assuming $\ell (1),\ell (2)>0$ we can choose $\gep^*$ small enough, so that all 
2824: arithmetic elements with minimal displacement $\leq\gep^*$ are singular, and each has a unique axis.
2825: Then we can apply the argument of the proof of Theorem \ref{thm2} to this case. 
2826: \end{proof}
2827: 
2828: \begin{rem}
2829: Similarly, $\ell (1)>0$ implies also the analog of  
2830: \ref{thmA}(2) for compact arithmetic $3$-manifolds.
2831: \end{rem} 
2832: 
2833: 
2834: 
2835: 
2836: %new section
2837: 
2838: 
2839: 
2840: 
2841: 
2842: \section{Estimating the size of a minimal presentation}
2843: 
2844: If Conjecture \ref{conjA} is true, 
2845: then, given a symmetric space of non-compact type $S$, and an $S$-manifold $M$, 
2846: such that either
2847: \begin{itemize}
2848: \item $S$ is not $\BH^3$, or
2849: \item $M$ is arithmetic,
2850: \end{itemize}
2851: the minimal size of a presentation for the fundamental group should be bounded
2852: linearly by the volume. 
2853: From Theorem \ref{thmA} we deduce this for most cases: 
2854: 
2855: \begin{defn}
2856: We say that a presentation of a group $\langle \gS : W\rangle$ is 
2857: {\bf standard} if the length of each $w\in W$ is $\leq 3$.
2858: \end{defn}
2859: 
2860: \begin{thm}\label{presentation-thm}
2861: Assume that either
2862: \begin{itemize}
2863:  \item $S$ is not isomorphic to $\BH^3, \BH^2\times\BH^2, \PSL_3(\BR )/\PSO_3(\BR )$, or
2864:  \item $M$ is non-compact arithmetic.
2865: \end{itemize}
2866: Then for some constant $\eta =\eta (S)$, independent of $M$,
2867: the fundamental group $\pi_1(M)$ admits a standard presentation 
2868: $$
2869:  \pi_1(M)\cong\langle \gS : W\rangle
2870: $$
2871: with $|\gS |,|W|\leq\eta\cdot\vol (M)$.
2872: 
2873: \end{thm}
2874: 
2875: \begin{proof}
2876: Let $\mathcal{R}$ be the $\big( d,\ga\cdot\vol (M)\big)$-simplicial complex which corresponds to $M$ by Theorem
2877: \ref{thmA} (1) or (2). Fix a spanning tree $T$ for 
2878: $\mathcal{R}$, and take the generating set $\gS$ for $\pi_1(\mathcal{R})\cong\pi_1(M)$ 
2879: which consists those closed loops which contain exactly one edge outside $T$.
2880: We thus obtain a generating set of size less then the number of edges of the 
2881: 1-skeleton $\mathcal{R}^1$ which is at most $\frac{\ga (S)\vol (M)d(S)}{2}$.
2882: In other words, we take for each edge of $\mathcal{R}^1\setminus T$ the element
2883: of $\pi_1(\mathcal{R}^1)$ which corresponds to the unique cycle (with 
2884: arbitrarily chosen orientation) which is obtained by adding this edge to $T$.
2885: Additionally, let the set of relations $W$ consist exactly 
2886: those words which are induced from 2-simplexes of $\mathcal{R}^2$ (we take
2887: one such relation for each 2-simplex). In this way we obtain
2888: a set of relations of size $\leq\ga (S)\vol (M) d^2(S)$ which is a bound for
2889: the number of triangles in $\mathcal{R}^1$.
2890: Thus, 
2891: $$
2892:  \eta =\ga (S)d^2(S)=\max\{ \ga (S)d^2(S),~\ga (S)d(S)/2\}
2893: $$
2894: will do.
2895: Finally, the length of each $w\in W$ is exactly the number of edges in the 
2896: corresponding 2-simplex which lie outside $T$, and thus the 
2897: presentation $\langle \gS : W\rangle$ is standard.
2898: \end{proof} 
2899: 
2900: \begin{rem}
2901: Lower bounds for the size of any presentation, are known for hyperbolic 
2902: $3$-manifolds (see \cite{cooper}). In this case, the upper bound obtained 
2903: above (for non-compact arithmetic $3$-manifolds) is tight. 
2904: \end{rem}
2905: 
2906: For non-arithmetic hyperbolic $3$-manifolds the analogous statement is 
2907: evidently false (see Remark \ref{55}). Surprisingly the following result holds:
2908: 
2909: \begin{thm}\label{Pre-Hyp}
2910: There is a constant $\eta$ such that the fundamental group of 
2911: any complete hyperbolic $3$-manifold $M$ admits a presentation
2912: $$
2913:  \pi_1(M)\cong\langle \gS : W\rangle
2914: $$
2915: for which both $|\gS |$ and $|W|$ are $\leq\eta\cdot\vol (M)$.
2916: \end{thm}
2917: 
2918: \begin{lem}
2919: Let $S$ be a rank one symmetric space, and let $\gep =\frac{\gep_s}{3}$
2920: where $\gep_s$ is the constant from the Margulis lemma. 
2921: There is a constant $c=c(S)$ such that for every $S$-manifold $M$, the number of closed geodesics of length $\leq\gep$ in $M$ is at most
2922: $c\cdot\vol (M)$.
2923: \end{lem}
2924: 
2925: \begin{proof}
2926: Write $M=\gC\backslash S$ and let $\ga, \gb\in\gC$ be elements which correspond to two
2927: different closed geodesics in $M$ of length $\leq\gep$. Then the axes of $\ga$
2928: and $\gb$ are bounded away from each other, and hence, for large enough $m$, $\langle\ga^m,\gb^m\rangle$ is a non-abelian
2929: free group. It follows that $\{ d_\ga <\gep_s\}\cap\{ d_\gb <\gep_s\}=\emptyset$.
2930: 
2931: Since $d_\gc (x)\leq \gep+2D_{\{ d_\gc <\gep\}}(x)$, we see that
2932: $$
2933:  \{ D_{\{ d_\ga <\gep\}}<\gep\}\cap\{ D_{\{ d_\gb <\gep\}}<\gep\}=\emptyset.
2934: $$
2935: This implies that if we take, for each connected component $M_{\leq\gep}^0$
2936: of $M_{\leq\gep}$, an $\gep$-ball $B_{\gep}$, whose center lies on the
2937: boundary of $M_{\leq\gep}^0$, then these balls are disjoint and injected.
2938: Thus the number of geodesics of length $\leq\gep$ (which coincides with
2939: the number of connected components of $M_{\leq\gep}$ and hence with the number 
2940: of these ${\gep}$-balls) is $\leq\vol (M)/\vol (B_{\gep})$.
2941: \end{proof}
2942: 
2943: \begin{proof}[Proof of Theorem \ref{Pre-Hyp}]
2944: Let $\gep =\frac{\gep_s}{3}$, let $M$ be a complete hyperbolic 
2945: $3$-manifold, and let $N\subset M$ be the union of all closed geodesics in $M$
2946: of length $\leq\gep$. Then $N$ consists of 
2947: $\leq\frac{\vol (M)}{\vol (B_{\gep})}$
2948: circles. As in the proof of \ref{thm3}, $\pi_1(M\setminus N)$ is
2949: isomorphic to $\pi_1(\mathcal{R})$ for some 
2950: $\big(\ga,d\cdot\vol(M)\big)$-simplicial complex $\mathcal{R}$.
2951: It follows that $\pi_1(M\setminus N)$ admits a presentation with 
2952: $\leq\frac{\ga d}{2}\vol (M)$ generators, and $\leq \ga d^2\vol (M)$ relations.
2953: Van-Kampen's theorem implies that when adding these circles one by one
2954: to $M\setminus N$, we should add one relation for each (note that each circle
2955: has a neighborhood homeomorphic to a solid torus or a solid Klein bottle) .
2956: Hence, we get a presentation of $\pi_1(M)$
2957: with the same number of generators and with at most 
2958: $\frac{\vol (M)}{\vol (B_\gep)}$ additional relations.
2959: \end{proof}
2960: 
2961: \begin{rem}\label{55}
2962: In contrast with Theorem \ref{presentation-thm}, Theorem \ref{Pre-Hyp} does 
2963: not yield bounds for the length of the relations. Since for $v$ large enough, 
2964: there are infinitely many complete hyperbolic $3$-manifolds with volume 
2965: $\leq v$, the length of the relations in the above presentations can not be 
2966: bound in terms of $\vol (M)$.
2967: \end{rem}
2968: 
2969: \begin{rem}
2970: For $S=\BH^2\times\BH^2$ the analog of Theorem \ref{Pre-Hyp} holds. 
2971: The proof is almost the same, except
2972: that in this case, each of the $\leq\frac{\vol (M)}{\vol (B_{\gep /2} )}$ 
2973: connected components of the union $N$ of all closed 
2974: geodesics of length $\leq\gep$ is either a circle or a two dimensional torus
2975: or a Klein bottle.
2976: \end{rem}
2977: 
2978: However for the symmetric space $S=\SL_3(\BR )/\SO_3(\BR )$ we have only the 
2979: following result which follows directly from Theorem \ref{thm3}.
2980: 
2981: \begin{prop}
2982: For any $S$, there is a constant $\eta (S)$ such that the fundamental group
2983: of any $S$-manifold $M$ has a generating set of size $\eta (S)\cdot\vol (M)$.
2984: \end{prop}
2985: 
2986: 
2987: 
2988: 
2989: 
2990: 
2991: 
2992: 
2993: 
2994: %NEW SECTION
2995: 
2996: 
2997: 
2998: 
2999: \section{A quantitative version of Wang's theorem}\label{QuaWang}
3000: 
3001: Denote by $\rho_S(v)$ the number of isometric classes of irreducible 
3002: $S$-manifolds of volume $\leq v$. 
3003: 
3004: If Conjecture \ref{conjA} is true then for any
3005: symmetric space, $S$, of non-compact type of dimension $\geq 4$, there is some constant $c=c(S)$,
3006: such that
3007: $$
3008:  \rho_S(v)\leq v^{c\cdot v}
3009: $$
3010: for any $v>0$.
3011: While in dimension $3$, the validity of Conjecture \ref{conjA} would yield an analogous upper bound
3012: for the growth of {\it arithmetic} $3$-manifolds.
3013: 
3014: Theorems \ref{thmA}(2) implies: 
3015: 
3016: \begin{thm}\label{QWthm}
3017: Assume that $S$ is neither isometric to 
3018: $\BH^2,\BH^3,\SL_3(\BR )/\SO_3(\BR )$ nor to $\BH^2\times\BH^2$.
3019: Then we have
3020: $$
3021:  \rho_S(v)\leq v^{c(S)\cdot v}.
3022: $$
3023: \end{thm}
3024: 
3025: And \ref{thmA}(1) implies:
3026: 
3027: \begin{prop}\label{12.2}
3028: The number of non-compact arithmetic hyperbolic $3$-manifolds of volume
3029: $\leq v$ is at most $v^{cv}$ for some constant $c$.
3030: \end{prop}
3031: 
3032: \begin{rem}\label{12.3}
3033: Similarly, for $S=\SL_3(\BR )/\SO_3(\BR )$ or $\BH^2\times\BH^2$, the
3034: number of non-compact irreducible manifolds with volume $\leq v$ is bounded
3035: by $v^{cv}$ for some $c$.
3036: \end{rem}
3037: 
3038: 
3039: \begin{proof}[Proof of \ref{QWthm}, \ref{12.2} and \ref{12.3}]
3040: Since $\dim (S)\geq 3$, it follows from Mostow's rigidity theorem that an 
3041: irreducible $S$-manifold $M$ is characterized by its fundamental 
3042: group, which, by Theorem \ref{presentation-thm}, has a presentation 
3043: $\pi_1(M)\cong\langle \gS , W\rangle$ with $|\gS |,|W|\leq\eta (S)\vol (M)$ in which all the relations has length $\leq 3$.
3044: A rough estimate of the number
3045: of groups admitting such a presentation yields \ref{QWthm}. 
3046: \end{proof} 
3047: 
3048: \begin{rem}
3049: It was shown in \cite{BGLM} that for $\BH^n$ when $n\geq 4$ this estimate
3050: is tight. However in the higher rank case it is very likely that $\rho_S(v)$ grows
3051: much slower. It was guessed in \cite{BGLM} that when $\rk (S)\geq 2$
3052: $$
3053: \log\rho_S(v)\approx {c(S)\frac{(\log V)^2}{\log\log V}}.
3054: $$
3055: \end{rem}
3056: 
3057: \begin{rem}
3058: Since Mostow rigidity does not hold for 
3059: surfaces, our method does not yield a quantitative version for Borel's 
3060: finiteness theorem for arithmetic hyperbolic surfaces of a given genus.
3061: \end{rem}
3062: 
3063: 
3064: 
3065: 
3066: 
3067: %NEW SECTION
3068: 
3069: 
3070: 
3071: 
3072: 
3073: \section{Some complements}\label{complements}
3074: 
3075: 
3076: \subsection{Extending some of the results to non-compact orbifolds} 
3077: 
3078: In the previous sections we have considered $S$-manifolds of finite volume. It is natural to try 
3079: to generalize the results obtained, to the larger family of $S$-orbifolds of finite volume. 
3080: This amounts to consider general lattices in $G$ instead of just torsion free lattices.
3081: 
3082: It turns out that some of the main statements could be generalized to  
3083: non-compact $S$-orbifolds, i.e. to general non-uniform lattices 
3084: $\gC\leq G$. We remark that we do not know how to deal with general compact orbifolds.
3085: In the non-compact case, our generalizations rely on the following effective version of Selberg's lemma: 
3086: 
3087: \begin{lem}\label{i(t.f.)}
3088: There is a constant $i=i(G)\in\BN$ such that any non-uniform arithmetic lattice
3089: $\gC\leq G$ has a torsion free normal subgroup of index $\leq i$.
3090: \end{lem}
3091: 
3092: \begin{proof}
3093: Let $\mathfrak{g}$ denote the Lie algebra of $G$ and let $n$ be its dimension.
3094: Replacing $G$ by its identity component we can assume it is connected and 
3095: center free, and therefore may be identified with its image under the adjoint 
3096: representation $\textrm{Ad}(G)\leq\textrm{GL}(\mathfrak{g})\cong GL_n(\BR )$. 
3097: As follows from the proof of Margulis' arithmeticity theorem, for any 
3098: non-uniform arithmetic lattice $\gC\leq G$ there is a base $B$ for the
3099: vector space $\mathfrak{g}\cong \BR^n$ with respect to which 
3100: $\gC\leq \textrm{GL}_n(\BQ )$ and is commensurable to $\textrm{GL}_n(\BZ )$.
3101: Replacing this base by a $\BZ$-base for the $\BZ$-span of $\gC\cdot B$ (which
3102: is easily seen to be a $\BZ$-lattice in $\BR^n$), we can assume that $\gC$ is 
3103: contained in $\textrm{GL}_n(\BZ )$.
3104: 
3105: Let $T\leq\textrm{GL}_n(\BZ )$ be a fixed torsion-free congruence subgroup  
3106: (which exists, for instance, by Selberg's lemma) and let $i=i(G)$ be its index
3107: $$
3108:  i=[\textrm{GL}_n(\BZ ):T].
3109: $$
3110: Clearly, $\gC\cap T$ is torsion free and $[\gC :\gC\cap T]\leq i$.
3111: \end{proof}
3112: 
3113: The following generalization of Theorem \ref{presentation-thm} follows 
3114: immediately:
3115: 
3116: \begin{thm}\label{Gpresentation}
3117: There is a constant $\eta (G)$ such that any non-uniform arithmetic 
3118: lattice $\gC\leq G$ has a presentation 
3119: $\gC\cong \langle \Sigma :W\rangle$
3120: with 
3121: $$
3122:  |\Sigma |,|W|\leq\eta (G)\cdot\vol (G/\gC ).
3123: $$
3124: \end{thm}
3125: 
3126: \begin{proof}
3127: Take a torsion free normal subgroup $\gC_1$ of index $\leq i$ in $\gC$.
3128: Then $\vol (G/\gC_1)=\vol (G/\gC )|\gC/\gC_1|\leq\vol (G/\gC )\cdot i$, and 
3129: $\gC_1$, being torsion free, has a presentation with $\leq\eta'\vol (G/\gC_1)$ generators and 
3130: relations by \ref{presentation-thm}. We should add at most $|\gC/\gC_1|\leq i$ generators and 
3131: $|\gC/\gC_1|^{|\gC/\gC_1|}\leq i^i$ relations to get a presentation for $\gC$.
3132: \end{proof}
3133: 
3134: However, unlike the case of \ref{presentation-thm}, we do not know how to
3135: bound the lengths of the relations in $W$.
3136: 
3137: \medskip
3138: 
3139: The following extends the results of section \ref{QuaWang}.
3140: 
3141: \begin{thm}\label{countingL}
3142: There is a constant $c=c(G)$ such that for any $v>0$, the number of conjugacy 
3143: classes of non-uniform arithmetic lattices of covolume $\leq v$ is at most
3144: $v^{c\cdot v}$.
3145: \end{thm}
3146: 
3147: \begin{proof}
3148: We already know (by section \ref{QuaWang}) that for any $v>0$ there are at most 
3149: $v^{c'\cdot v}$ 
3150: conjugacy classes of torsion free lattices of covolume $\leq v$, and therefore 
3151: at most $v^{c''\cdot v}$ conjugacy classes of torsion free lattices of covolume
3152: $\leq i\cdot v$ where $i=i(G)$ is the constant from Lemma \ref{i(t.f.)}.
3153: 
3154: Let $v_0$ be the minimal covolume of a lattice $\gC\leq G$. Let $\gC\leq G$ be 
3155: a lattice of covolume $\leq v$. By Lemma \ref{i(t.f.)}, $\gC$ contains a 
3156: torsion free normal subgroup $\gC '$ of index $\leq i$. Let $N_G(\gC ')$ be 
3157: the normalizer of $\gC '$ in $G$. Then $N_G(\gC ')$ is a lattice containing 
3158: $\gC$ whose covolume satisfies
3159: $$
3160:  v_0\leq \vol\big( G/N_G(\gC ')\big)\leq v.
3161: $$ 
3162: 
3163: It follows from Theorem \ref{Gpresentation} that $N_G(\gC ')$ has a generating 
3164: set of size $\leq [\eta\cdot v]$. Thus, for any $j$, the number of subgroups 
3165: of $N_G(\gC ')$ of index $j$ is no more than 
3166: $\big(j+[\eta\cdot v]\big) !^2$ 
3167: (which is a trivial upper bound for the number of index $j$ subgroups of the 
3168: free group of rank $[\eta\cdot v]$). The index $[N_G(\gC '):\gC ]$ is at most 
3169: $v/v_0$, so there are at most
3170: $\big( [v/v_0]+[\eta\cdot v]\big) !^2$ choices for $\gC$ as a
3171: subgroup of index
3172: $[N_G(\gC '):\gC ]$ of $N_G(\gC ')$. Summing over all indexes $\leq v/v_0$, 
3173: we get that there are at most
3174: $[v/v_0]\cdot \big( [v/v_0]+[\eta\cdot v]\big) !^2\leq v^{c'''\cdot v}$ 
3175: choices for $\gC$ as a
3176: subgroup of index $\leq [N_G(\gC '):\gC ]$ of $N_G(\gC ')$. 
3177: 
3178: Thus, the number of lattices $\gC$ which contain the same $\gC '$ as a 
3179: normal subgroup of index $\leq i$ is at most $v^{c'''\cdot v}$. 
3180: Since there are at most $v^{c''v}$ possibilities for $\gC'$,
3181: it follows that the number
3182: of conjugacy classes of lattices of covolume $\leq v$ is at most 
3183: $v^{c''\cdot v}\cdot v^{c'''\cdot v}\leq v^{c\cdot v}$.
3184: \end{proof}
3185: 
3186: 
3187: 
3188: 
3189: 
3190: %NEW subSECTION
3191: 
3192: 
3193: 
3194: \subsection{Commensurable growth}
3195: Let us now restrict our attention to a fixed commensurability class. 
3196: 
3197: \begin{defn}
3198: Two $S$-manifolds $M,N$ are called {\bf commensurable} if they have a common
3199: finite cover. I.e. $\gC_1\backslash S$ is commensurable to $\gC_2\backslash S$ iff $\gC_1$ is
3200: commensurable to some conjugate of $\gC_2$ in $G=\text{Isom}(S)$.
3201: \end{defn}
3202:    
3203: The following definition is natural.
3204: 
3205: \begin{defn}
3206: The {\bf commensurable growth} $\kappa_M(v)$ of a locally symmetric manifold 
3207: $M$, is the number of non-isometric manifolds commensurable to $M$  
3208: with volume $\leq v$.
3209: The {\bf commensurable growth} $\kappa_\gC (v)$ of a lattice $\gC\leq G$ is the
3210: number of conjugacy classes of lattices commensurable to $\gC$ with
3211: covolume $\leq v$.
3212: \end{defn}
3213: 
3214: %{\bf Some remarks and questions}
3215: 
3216: %\begin{itemize}
3217: %\item 
3218: One can define the notion of commensurable growth for arbitrary subgroup
3219: $\gC\leq G$, not necessarily a lattice, as follows: define the 
3220: generalized index between commensurable subgroups
3221: $\gC ,\gC'\leq G$ to be the rational number 
3222: $$
3223:  [\gC :\gC']=\frac{[\gC :\gC\cap\gC']}{[\gC' :\gC\cap\gC']},
3224: $$ 
3225: and use this concept instead of ``covolume'' in the above definition.
3226: %\item 
3227: 
3228: Clearly, for a locally symmetric manifold $M=\gC\backslash S$, we have
3229: $\kappa_M(v)\leq\kappa_\gC (v)$. It is natural to ask what is the relation
3230: between these functions. In particular, 
3231: do they have the same asymptotic behavior?    
3232: %\item 
3233: 
3234: Another interesting question is what is the relation between the commensurable growth  
3235: and the congruence subgroup problem.
3236: %\end{itemize}
3237: 
3238: \medskip
3239: 
3240: We shall now give upper bounds for the commensurable growth of locally 
3241: symmetric manifolds and for its fundamental group when the dimension is $>2$.
3242: 
3243: \begin{clm}
3244: Let $M=\gC\backslash S$ be an irreducible locally symmetric manifold of dimension $>2$ 
3245: with finite volume. Then there is a constant $c=c(M)$ such that 
3246: $\kappa_M(v)\leq v^{cv}$.
3247: \end{clm}
3248: 
3249: \begin{proof}
3250: If $M=\gC\backslash S$ is not arithmetic then, by Margulis' criterion for arithmeticity,
3251: the commensurability class of $\gC$ admits a unique maximal element which 
3252: contains all the others, and the result follows by considering the subgroup 
3253: growth of this maximal element. If $M$ is arithmetic, this follows from \ref{fixedO} and from 
3254: \ref{thm1}.
3255: \end{proof}
3256: 
3257: When $\gC\leq G$ is a non-arithmetic lattice, the above proof applies also to $\kappa_\gC$ and gives 
3258: $\kappa_\gC (v)\leq v^{cv}$ as well. In the arithmetic case, as in 
3259: \ref{countingL}, we obtain similar upper bounds by using the following 
3260: lemma, which can be proved in the same way as Lemma \ref{i(t.f.)}.
3261: 
3262: \begin{lem}\label{180}
3263: For any commensurability class $\mathfrak{N}$ of arithmetic lattices in $G$,
3264: there is a constant $i=i(\mathfrak{N})$ such that any $\gC\in\mathfrak{N}$ 
3265: contains a torsion free subgroup of index $\leq i$.
3266: \end{lem}
3267: 
3268: The following is immediate from \ref{fixedO} and \ref{180}.
3269: 
3270: \begin{lem}
3271: Given a commensurability class $\mathfrak{N}$ of arithmetic lattices in $G$,
3272: there is a constant $\eta =\eta (\mathfrak{N})$ such that any 
3273: $\gC\in\mathfrak{N}$
3274: has a presentation $\gC\cong \langle \Sigma :W\rangle$
3275: with $|\Sigma |,|W|\leq\eta (G)\cdot\vol (G/\gC )$.
3276: \end{lem}
3277: 
3278: As in the proof of \ref{countingL}, these two lemmas imply:
3279: 
3280: \begin{prop}
3281: For any lattice $\gC\leq G$, $\kappa_\gC (v)\leq v^{c(\gC )v}$. 
3282: \end{prop}
3283: 
3284: 
3285: 
3286: 
3287: 
3288: 
3289: %NEW subSECTION
3290: 
3291: 
3292: 
3293: 
3294: 
3295: \subsection{How to construct a simplicial complex for non-arithmetic manifolds}
3296: 
3297: We shall now explain how to attach simplicial complexes to
3298: non-arithmetic, or more generally to rank-$1$ manifolds, of dimension $\geq 4$. We are not
3299: trying to do it in the most economical way, but just to explain an idea of how this could be done.
3300: 
3301: The following proposition follows from a rough estimate for the diameter and 
3302: the minimal injectivity radius of compact connected components of the thin 
3303: part.
3304: 
3305: \begin{prop}\label{rank-1prop}
3306: There are positive constants $\ga =\ga (n)$ and $d=d(n)$ such that 
3307: any compact rank-$1$ locally symmetric manifold $M$ of dimension $n\geq 4$ is 
3308: homotopically equivalent to a 
3309: $\big(d,\ga\cdot\vol (M)^{3n^2+1}\big)$-simplicial complex.
3310: \end{prop}
3311: 
3312: \begin{proof}
3313: Fix $n$, let $\gep (n)$ be the constant of the Margulis lemma, and let 
3314: $M_{\leq\gep (n)}$ be the thin part of the ordinary $\gep (n)$ thick-thin 
3315: decomposition.
3316: 
3317: It follows from \cite{BS} (see proposition 3.2 there) that for some constant
3318: $c$, the diameter of any compact connected component of $M_{\leq\gep (n)}$ is
3319: at most $3\log \big( c\cdot\vol (M)\big)$. 
3320: Applying formula 8.5 from \cite{Gr} (page 381), which implies that the 
3321: injectivity radius decreases at most exponentially as one moves along the 
3322: manifold, we get that the injectivity radius at any point belonging to
3323: $M_{\leq\gep (n)}$, and therefore at any point of $M$, 
3324: is at least $c'\cdot\vol (M)^{-3n}$, for some positive constant $c'$. 
3325: 
3326: The proposition follows by applying a good covering argument with ordinary 
3327: balls of radius $\gep =c'\cdot\vol (M)^{-3n}$. The needed number of balls in such a
3328: cover is $\leq c''\frac{\vol(M)}{\gep^n}\leq\ga\cdot\vol (M)^{3n^2+1}$.
3329: \end{proof}     
3330: 
3331: In fact, it is easy to obtain stronger results by means of elementary 
3332: computations of volumes of neighborhoods of short closed geodesics. 
3333: We shall demonstrate this in the real hyperbolic case. We remark that in all
3334: other rank-$1$ cases, one can obtain similar estimates by using the same
3335: means, and applying Rauch's comparison theorems (see \cite{CH} 1.10). 
3336: However, also the following estimate is probably not tight, and
3337: it should be possible (and not necessarily very hard) to obtain better 
3338: estimates. 
3339: 
3340: \begin{prop}
3341: For $n\geq 4$, there are constants $\ga =\ga (n), d=d(n)$, such that any 
3342: compact hyperbolic $n$-manifold $M$ is homotopically 
3343: equivalent to some 
3344: $\big(d,\ga\cdot \vol (M)^{(1+n( [\frac{n-1}{2}]+1))/(n-2-[\frac{n-1}{2}])}
3345: \big)$-simplicial complex. 
3346: \end{prop}
3347: 
3348: \begin{proof}[Sketched proof for $n=4$]
3349: In this case any connected
3350: component of the thin part of the ordinary thick-thin decomposition (with 
3351: respect to some fixed $\gep$) is a neighborhood of a short closed geodesic 
3352: which is topologically a ball bundle over a circle.
3353: In order to understand the geometry of the thin components it is most 
3354: convenient to look at the upper half-space model for the hyperbolic space
3355: $\BH^4$ (see \cite{BePe} for details). Lift the component so that the short 
3356: closed geodesic is lifted to the line connecting $0$ to $\infty$. 
3357: Then our lifted 
3358: component is a cone, centered by this line. The intersection of this cone with
3359: a horosphere perpendicular to this line is a union of coaxial ellipsoids (with 
3360: respect to the induced $(n-1)$-Euclidean structure on the horosphere). 
3361: 
3362: Using the fact that any abelian subgroup of $\textrm{SO}_3(\BR )$ 
3363: ($\cong$ to the fixator group of this line) is contained 
3364: in a $1$-dimensional torus, one can show, by a simple pigeonhole argument on the powers of the corresponding 
3365: isometry, that if the length of our short closed
3366: geodesic is $a<<\gep$ then, for some constant $c$, the $3$ dimensional
3367: Euclidean ball of radius $c\frac{1}{a^{1/2}}$ is contained in the union of 
3368: the above ellipsoids. This implies that the volume of the component is at least
3369: $c'(1/a^{1/2})^3\cdot a$. Thus $a\geq c''\frac{1}{\vol (M)^2}$. 
3370: 
3371: Thus the injectivity radius at any point of $M$ is at least 
3372: $\rho =c''\frac{1}{\vol (M)^2}$, 
3373: and one can construct, as above, a simplicial complex with at most
3374: $c'''\frac{\vol (M)}{\rho^4}=\ga\cdot\vol (M)^9$ vertices, all of them of 
3375: degree bounded by some constant $d$.
3376: 
3377: We remark that the proof for general dimension $n$ uses the fact that any 
3378: abelian subgroup of $\textrm{SO}_{n-1}(\BR )$ is contained in some 
3379: $[\frac{n-1}{2}]$-dimensional torus.
3380: \end{proof}
3381: 
3382: \begin{rem}
3383: In order to obtain analogous estimates for non-compact rank-$1$ manifolds, one 
3384: should use the thick-thin decomposition with $\gep =\gep \big(\vol (M)\big)$
3385: as above, so that all the components of $M_{\leq\gep}$ would be cusps, and 
3386: then estimate explicitly the function $\gb (\gep)$ which is defined in the
3387: proof of Theorem \ref{thm1},
3388: and detect its influence on the determination of the constant $b=b(\gep )$
3389: in Proposition \ref{simplicial-complex} in order to finally calculate the resulting 
3390: simplicial complex.
3391: \end{rem}
3392: 
3393: 
3394: 
3395: %subsection
3396: 
3397: 
3398: 
3399: \subsection{Wang's theorem for products of $\SL_2$'s}\label{SL_2}
3400: 
3401: This paragraph is not precisely a part of the main theme of this paper, but only a part of the 
3402: same subject of mathematics. 
3403: Moreover, we are not presenting any new result here, but only clarify things which are
3404: evidently known to some peoples. The author decided to write this paragraph because it might serve
3405: as a complement to Wang's paper \cite{Wa}.
3406: 
3407: Wang's theorem states that if $G$ is a connected semisimple
3408: Lie group without compact factors, and $G$ is not locally isomorphic to 
3409: $\textrm{PSL}_2(\BR )$ or $\textrm{PSL}_2(\BC )$,
3410: then for any $v>0$, there are only 
3411: finitely many conjugacy classes of irreducible lattices in $G$ of covolume 
3412: $\leq v$.
3413: 
3414: In \cite{Wa} Wang didn't consider the case where $G$ is of higher rank and
3415: has factors locally 
3416: isomorphic to $\textrm{PSL}_2(\BR )$ or $\textrm{PSL}_2(\BC )$.
3417: Margulis' arithmeticity theorem implies that if $G$ has both a 
3418: factor which is locally isomorphic to $\textrm{PSL}_2(\BR )$ or 
3419: $\textrm{PSL}_2(\BC )$ and a factor which is not locally isomorphic
3420: $\textrm{PSL}_2(\BR )$ or $\textrm{PSL}_2(\BC )$ then $G$ contains no 
3421: irreducible lattices (see \cite{Mar1} corollary 4.5, page 315). 
3422: It was remarked by Borel (see \cite{Bor} 8.3 and 8.1)
3423: that Wang's argument implies also the finiteness
3424: of the number of conjugacy classes of irreducible lattices for groups locally
3425: isometric to $G_{a,b}=\SL_2(\BR )^a\times\SL_2(\BC )^b$ when 
3426: $(a,b)\neq (1,0),(0,1)$. 
3427: Let as now explain this remark of Borel.
3428: 
3429: It follows from Margulis' super rigidity theorem that irreducible higher rank
3430: lattices are locally rigid. Thus, the missing 
3431: ingredient in Wang's argument (\cite{Wa}, 8.1) when applied to groups 
3432: locally isometric to $G_{a,b}$ is the following statement (which was also 
3433: noted without proof in \cite{Bor}).
3434: 
3435: \begin{prop}\label{irr-lim}
3436: Let $G$ be a semi-simple Lie group with no compact factors, and let 
3437: $\Gamma_n\leq G$ be a sequence of irreducible lattices. Assume that 
3438: $(\Gamma_n)$ converges to a lattice $\gD$ in the topology of closed 
3439: subgroups (Hausdorff convergence on compact sets). Then $\gD$ is also 
3440: irreducible. 
3441: \end{prop}
3442: 
3443: \begin{proof}
3444: Assume that $\gD$ is reducible. Then we can write $G$ as an almost direct 
3445: product $G=G_1\cdot G_2$ in such a way that $\gD$ is commensurable with
3446: $\gD_1\cdot\gD_2$ where $\gD_i=\gD\cap G_i$. Fix a finite generating set for
3447: $\gD$, and for large $n$, denote by $f_n:\gD\to\gC_n$ the homomorphism induced
3448: by sending each generator to the closest element in $\gC_n$. As explained in 
3449: \cite{Wa}, since $\gD$ is finitely presented, $f_n$ is a well defined 
3450: homomorphism whenever $\gC_n$ is close enough to $\gD$.
3451: 
3452: For any $\gd\in\gD$, $f_n(\gd )\to \gd$.  
3453: We will show that $f_n(\gd )$ is central for each non-central $\gd\in\gD_1$, and for any large enough $n$. Since the center of $G$ is 
3454: discrete, this will imply the desired contradiction. 
3455:      
3456: Fix $\gd\in\gD_1$ non-central. As $\gC_n$ is irreducible, 
3457: we will show that $f_n(\gd )$ is central by showing that its projection
3458: to the second factor $\pi_2(f_n(\gd ))$ is the unit element in $G_2$. 
3459: 
3460: $\gD_2$ is a lattice in $G_2$.
3461: Let $\{\gd_{2,1},\gd_{2,2},...,\gd_{2,k}\}\in\gD_2$ be a finite set of 
3462: generators for $\gD_2$. By Borel's density theorem, 
3463: $\{\text{Ad}(\gd_{2,1}),...,\text{Ad}(\gd_{2,k})\}$ generates the algebra 
3464: $$
3465:  \langle \text{Ad} (G_2)\rangle \leq
3466:  \textrm{End}(\mathfrak{g}_2)
3467: $$ 
3468: (here $\mathfrak{g}_2$ denotes the Lie algebra 
3469: of $G_2$). Since this algebra is finite dimensional, it is generated by
3470: $\{ \text{Ad}\big(\pi_2 ( f_n (\gd_{2,i} ))\big)\}_{i=1}^k$ 
3471: whenever $n$ is 
3472: large enough. Since $G_2$ is semi-simple, the adjoint representation 
3473: $\text{Ad}:G_2\to\textrm{GL}(\mathfrak{g}_2)$ has no invariant vectors.
3474: 
3475: Let $\gep_n=\pi_2 (f_n (\gd ))$. Since $f_n (\gd )$ is close to $\gd$,
3476: $\gep_n=\pi_2 (f_n (\gd ))$ is close to the identity of $G_2$. We can therefore
3477: assume that $\gep_n$ is contained in an identity neighborhood of $G_2$ where 
3478: $\log =\exp^{-1}:G_2\to\mathfrak{g}_2$ is a well defined diffeomorphism.
3479: As $\gd$ commutes with each $\gd_{2,i}$, $\gep_n$ commutes with each 
3480: $\pi_2 (f_n (\gd_{2,i} ))$, and it follows that
3481: $$
3482:  \text{Ad}\big(\pi_2 ( f_n (\gd_{2,i} ))\big)(\log \gep_n )=
3483:  \log \gep_n,
3484: $$
3485: which in turn implies $\log \gep_n=0$, i.e. $\gep_n=1$.
3486: \end{proof}
3487: 
3488: Together with Proposition \ref{irr-lim}, the original argument from \cite{Wa} 8.1 gives:
3489: 
3490: \begin{thm}[Wang's theorem]\label{WT}
3491: Let G be a connected semi-simple Lie group without compact factors, which is 
3492: not locally isomorphic to $\SL_2(\BR )$ or $\SL_2(\BC )$. Then for any $v>0$
3493: there are only finitely many conjugacy classes of irreducible lattices in $G$
3494: with covolume $\leq v$. 
3495: \end{thm}  
3496:        
3497: \begin{rem}
3498: In \cite{BP}, Borel and Prasad established a very strong and general finiteness result, but they
3499: omitted the cases of $G=G_{a,b}$ by requiring absolute rank $\geq 2$. This requirement was used in 
3500: their proof of the stronger finiteness statement, in which the ambient group $G$ can be varied. 
3501: However, Prasad remarked to the author that, when $G$ is fixed, this requirement is unnecessary
3502: in their argument, and hence, the finiteness of the number of 
3503: conjugacy classes of arithmetic lattices of covolume $\leq v$ in $G$
3504: could be proved also by using their methods.
3505: \end{rem}
3506: 
3507: \begin{rem}
3508: More generally, for any $G$, the finiteness statement holds for lattices (not necessarily 
3509: arithmetic) which are irreducible with respect to the $\SL_2$ factors of $G$.
3510: I.e. for the set of conjugacy classes of lattices in $G$ which project densely
3511: to any factor of $G$ which is locally isomorphic to $\SL_2(\BR )$ or
3512: $\SL_2(\BC )$.
3513: \end{rem}
3514: 
3515: 
3516: 
3517: 
3518: 
3519: 
3520: 
3521: 
3522: 
3523: 
3524: 
3525: 
3526: \begin{Ack}
3527: I would like to thank 
3528: Shahar Mozes (my Ph.D. adviser) for his guidance throughout 
3529: the last few years and for insightful suggestions for this work,
3530: to Pierre Pansu for hours of helpful conversations and for 
3531: suggestions which were essential to this research,
3532: to Alex Lubotzky for many discussions, suggestions and ideas, 
3533: and to Uri Bader, Yair Glasner, and Yehuda 
3534: Shalom for many discussions and clever suggestions. Proofs of some of the 
3535: statements presented here were established during these discussions. 
3536: I would also like to thank Emmanuel Breuillard, Assaf Naor and Gopal Prasad 
3537: for some remarks concerning early versions of this paper.
3538: Finally, I would like to express my sincere gratitude to the anonymous referees 
3539: for their careful reading of the manuscript and for many 
3540: suggestions and corrections which tremendously improved the exposition of this paper.
3541: 
3542: \end{Ack}
3543: 
3544: 
3545: 
3546: 
3547: 
3548: 
3549: \begin{thebibliography}{99}
3550: 
3551: \bibitem{BaNe} U. Bader, A. Nevo, Conformal actions of simple Lie groups on compact pseudo-Riemannian manifolds, 
3552: J. Differential Geom. {\bf 60} (2002), no. 3, 355--387.
3553: 
3554: \bibitem{BGS} W. Ballmann, M. Gromov, V. Schroeder, 
3555: {\it Manifolds of Nonpositive Curvature}, Birkhauser, 1985.
3556: 
3557: \bibitem{BePe} R. Benedetti, C. Petronio, {\it Lectures on Hyperbolic 
3558: Geometry}, Springer-Verlag, 1992.
3559: 
3560: \bibitem{Bor} A. Borel, Commensurability classes and volumes of hyperbolic 
3561: 3-manifolds, Ann. Scuola Norm. Sup. Pisa, Ser. IV, {\bf 8} (1981) 1-33.
3562: 
3563: \bibitem{BoHC} A. Borel, Harish-Chandra, Arithmetic subgroups of algebraic groups. Ann. Math.
3564: (2) {\bf 75} (1962) 485-535.
3565: 
3566: \bibitem{BP} A. Borel, G. Prasad, 
3567: Finiteness theorems for discrete subgroups of bounded covolume in 
3568: semi-simple groups, Publ. Math. I.H.E.S. {\bf 69} (1989), 119-171.
3569: 
3570: \bibitem{BT} R. Bott, L.W. Tu, {\it Differential Forms in Algebraic 
3571: Topology}, Springer-Verlag, 1982.
3572: 
3573: \bibitem{BH} R. Bridson, A. Haefliger, {\it Metric Spaces of Non-Positive 
3574: Curvature}, Springer, 1999.
3575: 
3576: \bibitem{BGLM} M. Burger, T. Gelander, A. Lubotzky, S. Mozes, 
3577: Counting hyperbolic manifolds, Geom. Funct. Anal. 12 (2002), no. 6, 1161--1173. 
3578: 
3579: \bibitem{BS} M. Burger, V. Schroeder, Volume, diameter and the first 
3580: eigenvalue of locally symmetric spaces of rank one, 
3581: J. Differential Geometry {\bf 26} (1987), 273-284.
3582: 
3583: \bibitem{CH} J. Cheeger, D.G. Ebin, {\it Comparison theorems in Riemannian 
3584: geometry}, North-Holland
3585: 
3586: \bibitem{chinburg} T. Chinburg, Volume of hyperbolic manifolds, 
3587: J. Differential Geometry {\bf 18} (1983), 783-789.
3588: 
3589: \bibitem{cooper} D. Cooper, The volume of a closed hyperbolic 3-manifold
3590: is bounded by pi times the length of any presentation of its fundamental group,
3591: Proc. Amer. Math. Soc. {\bf 127} (1999), 941-942.
3592: 
3593: \bibitem{cor} k. Corlette, Archimedean superrigidity and hyperbolic geometry. 
3594: Ann. of Math. {\bf 135} (1992), no. 1, 165--182. 
3595: 
3596: \bibitem{Ga-Ra} H. Garland, M.S. Raghunathan, Fundamental domains for 
3597: lattices in (R-)rank 1 Lie groups, Ann. of Math. {\bf 92} (1970), 279-326.
3598: 
3599: \bibitem{Gr} M. Gromov, {\it Metric Structures for Riemannian and 
3600: Non-Riemannian Spaces}, Birkhauser, 1998.
3601: 
3602: \bibitem{Gr1} M. Gromov, Hyperbolic manifolds according to Thurston and Jorgensen, Semin. 
3603: Bourbaki, 32e annee, vol. 1979/80, exp. 546, Lect. Notes Math. 842, (1981), 40-53.
3604: 
3605: \bibitem{GS} M. Gromov, R. Schoen, Harmonic maps into singular spaces and $p$-adic superrigidity 
3606: for lattices in groups of rank one. Inst. Hautes Études Sci. Publ. Math. {\bf 76} (1992), 165--246.
3607:  
3608: \bibitem{Kazhdan} D.A. Kazhdan, Connection of the dual space of a group with the
3609: structure of its closed subgroups, Functional Analysis and Application {\bf 1}
3610: (1967), 63-65. 
3611: 
3612: \bibitem{Lub} A. Lubotzky, Subgroup growth and congruence subgroups, Invent.
3613: Math. Springer-Verlag {\bf 119} (1995), 267-295.
3614: 
3615: \bibitem{Mar1} G.A. Margulis, {\it Discrete Subgroups of Semisimple Lie 
3616: Groups}, Springer-Verlag, 1990.
3617:  
3618: \bibitem{Mar2} G.A. Margulis, Arithmeticity of the irreducible lattices in the semi-simple groups
3619: of rank greater then $1$, (Russian), Invent. Math. {\bf 76} (1984) 93-120.
3620: 
3621: \bibitem{Mar3} G.A. Margulis, Non-uniform lattices in semisimple algebraic groups. 
3622: Lie groups and their representations (Proc. Summer School on Group Representations of the Bolyai János Math. 
3623: Soc., Budapest, 1971), Halsted, New York, (1975) 371--553. 
3624: 
3625: \bibitem{MR} G.A. Margulis, J. Rohlfs, On the proportionality of covolumes of
3626: discrete subgroups. Math. Ann {\bf 275} (1986) 197-205 
3627: 
3628: \bibitem{PR} V. Platonov, A. Rapinchuk, {\it Algebraic Groups and Number 
3629: Theory}, Academic Press, 1994.
3630: 
3631: \bibitem{P} G. Prasad, Volume of $S$-arithmetic quotients of semi-simple 
3632: groups, Publ. Math. I.H.E.S. {\bf 69} (1989), 91-117.
3633: 
3634: \bibitem{Rag} M.S. Raghunathan, {\it Discrete Subgroups of Lie Groups}, 
3635: Springer, New York, 1972.
3636: 
3637: \bibitem{Sp} E. Springer, {\it Algebraic Topology}, 
3638: Springer-Verlag, 1966.
3639: 
3640: \bibitem{sury} B. Sury, Arithmetic groups and Salem numbers, Manuscripta Math
3641: {\bf 75} (1992), 97-102.
3642: 
3643: \bibitem{Th} W.P. Thurston, {\it Three-Dimensional Geometry and Topology},
3644: Volume 1, Princeton univ. press, 1997.
3645: 
3646: \bibitem{Wa} H.C. Wang, Topics on totally discontinuous groups,
3647: {\it Symmetric Spaces}, edited by W. Boothby and G. Weiss (1972), 
3648: M. Dekker, 460-487. 
3649: 
3650: \bibitem{S-P.W} S.P. Wang, The dual space of semisimple Lie groups, Amer. J. 
3651: Math {\bf 91} (1969), 921-937. 
3652: 
3653:          
3654: \end{thebibliography}
3655: 
3656: 
3657: 
3658: 
3659: 
3660: 
3661: \end{document}
3662: