1:
2: %
3: %
4: %
5: \documentclass{amsart}
6: %
7: %
8: %
9: %\agtart
10: %
11: %
12: \usepackage{amssymb,amsthm,color}
13: \usepackage{epsfig}
14: \usepackage[all]{xy}
15:
16: %
17: %
18: %
19:
20: \title{Actions of certain arithmetic groups on Gromov hyperbolic spaces}
21:
22: %
23: \author[J. F. Manning]{Jason Fox Manning}
24: %\givenname{Jason Fox}
25: %\surname{Manning}
26: \address{University at Buffalo, SUNY}
27: \email{j399m@buffalo.edu}
28: \urladdr{http://www.math.buffalo.edu/~j399m}
29:
30: %
31: %
32: %
33: %
34: %
35:
36: %\keyword{arithmetic group}
37: %\keyword{group action}
38: %\keyword{Gromov hyperbolic space}
39: %\keyword{rigidity}
40: %\subject{primary}{msc2000}{20F65}
41: %\subject{secondary}{msc2000}{22E40}
42:
43: %
44:
45:
46: %
47: %
48: %
49: %
50: %
51: %
52: %
53: \newtheorem{theorem}{Theorem}[section]
54: \newtheorem{lemma}[theorem]{Lemma}
55: \newtheorem{corollary}[theorem]{Corollary}
56: \newtheorem{proposition}[theorem]{Proposition}
57: \newtheorem{question}[theorem]{Question}
58: \newtheorem{conjecture}[theorem]{Conjecture}
59: \newtheorem{qconjecture}[theorem]{``Conjecture''}
60: \newtheorem{case}{Case}
61: \newtheorem{subcase}{Case}[case]
62: \newtheorem{subsubcase}{Case}[subcase]
63: \newtheorem{claim}[theorem]{Claim}
64: \newtheorem{subclaim}[theorem]{Subclaim}
65: \newtheorem{step}{Step}
66:
67: \theoremstyle{definition}
68: \newtheorem{remark}[theorem]{Remark}
69: \newtheorem{definition}[theorem]{Definition}
70: \newtheorem{example}[theorem]{Example}
71: \newtheorem{observation}[theorem]{Observation}
72:
73: %
74: %
75: \newcommand{\cd}[1]{\begin{equation*}{\xymatrix{#1}}\end{equation*}}
76: \newcommand{\cdlabel}[2]{\begin{equation}\label{#1}{\xymatrix{#2}}\end{equation}}
77: \newcommand{\FA}{\ensuremath{(\mathrm{FA})}}
78: \newcommand{\QFA}{\ensuremath{(\mathrm{QFA})}}
79: \newcommand{\isom}[1]{\mathrm{Isom}(#1)}
80: \newcommand{\Mtwo}[4]{\ensuremath{\begin{pmatrix} {#1} & {#2} \\ {#3} &
81: {#4}\end{pmatrix} }}
82: \newcommand{\textMtwo}[4]{\ensuremath{\left(\begin{smallmatrix} {#1} & {#2} \\ {#3} &
83: {#4}\end{smallmatrix}\right) }}
84: \newcommand{\mc}[1]{\mathcal{#1}}
85: \newcommand{\uu}[1]{\ensuremath{\underline{#1}}}
86: \newcommand{\gp}[3]{\ensuremath{({#1}\cdot{#2})_{#3}}}
87:
88: \def\co{\colon\thinspace}
89: \def\R{\mathbb{R}}
90: \def\Z{\mathbb{Z}}
91: \def\C{\mathbb{C}}
92: \def\H{\mathbb{H}}
93: \def\N{\mathbb{N}}
94: \def\Q{\mathbb{Q}}
95:
96:
97:
98:
99: \begin{document}
100:
101: \begin{abstract}
102: We study the variety of actions of a fixed (Chevalley) group on
103: arbitrary geodesic, Gromov hyperbolic spaces. In high rank we obtain
104: a complete classification. In rank one, we obtain some partial results
105: and give a conjectural picture.
106: \end{abstract}
107:
108: \maketitle
109:
110: \tableofcontents
111:
112:
113:
114:
115:
116: \section{Introduction}
117: Given a group $G$ one may ask the question:
118: \begin{question}
119: In what ways can $G$ act non-trivially on a Gromov hyperbolic metric space?
120: \end{question}
121: Many interesting groups can be fruitfully studied via some natural action on
122: a Gromov hyperbolic space. Examples include the action of an
123: amalgam or HNN extension on its Bass-Serre tree, the action of a
124: Kleinian group on $\H^n$, and the action of the mapping class group of
125: a surface on the curve complex of that same surface. Alternatively, one can
126: study the space of actions of a fixed group on some (fixed or
127: varying) Gromov hyperbolic metric space. For example, the $PSL(2,\C)$--character variety of a group
128: parameterizes the space of actions of a fixed
129: group on the hyperbolic space $\H^3$. Analysis of the structure of
130: this variety has led to many remarkable theorems about $3$--manifolds
131: and their fundamental
132: groups (for an introduction, see \cite{shalen:handbook}).
133: A larger ``variety'' (in scare quotes because there
134: is unlikely to be any algebraic structure) would describe all
135: non-trivial actions on Gromov hyperbolic spaces, up to some
136: appropriate equivalence relation. We give
137: suggestions for how to define this equivalence relation and topologize the
138: variety in Section \ref{s:actions}. Briefly, the equivalence relation
139: is that generated by coarsely equivariant quasi-isometric embeddings.
140: This equivalence is coarser than that given by quasi-conjugacy (as in
141: for instance \cite{msw:quasiactI}), but finer than that given by
142: equivariant homeomorphism of limit sets.
143: In this paper we concentrate on cases in which the set of equivalence
144: classes is particularly simple.
145:
146: Certain equivalence classes of actions on hyperbolic spaces cannot be
147: ruled out, or
148: even really analyzed using the tools of negative curvature. These are
149: the \emph{actions with an invariant horoball} (see Section
150: \ref{ss:combhoro} for the definition, and Theorem \ref{t:horeq} for
151: some characterizations). Actions with an invariant
152: horoball (in the sense used in this paper) are always elementary; they
153: include the trivial action
154: on a point and actions which preserve some horofunction.
155: A cobounded
156: action on an unbounded space never has an invariant horoball.
157:
158: The variety of actions of an irreducible higher rank lattice is
159: expected to be very simple.
160: \begin{conjecture}\label{c:main}
161: If $\Gamma$ is an irreducible lattice in a higher rank Lie group (or
162: in a
163: nontrivial product of locally compact groups) $G$,
164: there there are finitely many Gromov hyperbolic $G$--spaces (up to
165: coarse equivalence) without
166: invariant horoballs.
167: \end{conjecture}
168: In the case where $G$ is a simple Lie group of rank at least $2$, the
169: only expected actions are those with an invariant horoball. If $G$
170: has more than one direct factor, then $\Gamma$ projects densely to
171: each factor. If the factors have rank one, then there will clearly be
172: non-elementary isometric actions of $\Gamma$ on rank one symmetric
173: spaces. We discuss this case in Section \ref{section:rankone}.
174:
175: For actions by lattices in nontrivial products of simple Lie
176: groups, the conjecture follows from rigidity results of Monod
177: and Monod-Shalom
178: \cite{monod:superrigid,monodshalom:jdg}
179: if one restricts attention
180: to Gromov hyperbolic spaces which are also either CAT$(0)$ spaces,
181: proper and cocompact spaces, or bounded valence graphs.
182: (Cf.
183: \cite{gelanderkarlssonmargulis} for CAT$(0)$ spaces.)
184: For an example
185: of an action of a lattice on a Gromov hyperbolic space which is
186: inequivalent to any action on a Gromov hyperbolic CAT$(0)$ space,
187: consider a lattice as in the Appendix to \cite{manning:qfa}, which has
188: Property (T), but admits a non-trivial pseudocharacter (or homogeneous
189: quasi-morphism). This
190: pseudocharacter gives rise to a cobounded action on a space
191: quasi-isometric to $\R$, fixing both ends. On the other hand, an
192: action on a CAT$(0)$ space preserving a point at infinity would give
193: (via the Busemann function) a homomorphism to $\R$, and a cobounded such
194: action would give an unbounded homomorphism to $\R$.
195: Such a homomorphism is ruled out by Property (T).
196:
197: \begin{theorem}\label{thm:cdr2}
198: Suppose that $G$ is a simple Chevalley-Demazure
199: group scheme of rank at least $2$, and let $\mathcal{O}$ be the ring
200: of integers of any number field. Then any isometric action of
201: $G(\mathcal{O})$ on a Gromov hyperbolic geodesic metric space has
202: an invariant horoball.
203: \end{theorem}
204: Some remarks:
205: \begin{enumerate}
206: \item Some closely related results are proved by Karlsson and Noskov
207: \cite[Sections 8.2 and 8.3]{karlssonnoskov}; part of our strategy is
208: similar to theirs, and to that of Fukunaga in \cite{fukunaga:FA}.
209: \item The rank $\geq 2$ assumption is necessary, as the
210: action of $SL(2,\mathcal{O})$ on $\H^3$ obtained from the inclusion
211: $SL(2,\mathcal{O})\longrightarrow SL(2,\C)$ never has an
212: invariant horoball.
213: \end{enumerate}
214:
215: The rank one case is discussed in Section
216: \ref{section:rankone}, where we apply a result of Carter, Keller and
217: Paige to show a weaker theorem for these groups.
218: \begin{theorem}\label{thm:cdr1}
219: Let $\mathcal{O}$ be the ring of integers of a number field, and
220: suppose that $\mathcal{O}$ has infinitely many units. Suppose $X$ is
221: quasi-isometric to a tree. Every action of $SL(2,\mathcal{O})$ on $X$
222: has a bounded orbit.
223: \end{theorem}
224:
225:
226: \subsection{Relative hyperbolicity and bounded generation}
227: If $G$ is a finitely generated
228: group, $S$ is a generating set for $G$, and
229: $\mc{P}=\{P_1,\ldots,P_n\}$ is a collection of
230: subgroups of $G$, then one can form the \emph{coned space}
231: $C(G,\mc{P},S)$ as follows: Let $\Gamma(G,S)$ be the Cayley graph of
232: $G$ with respect to $S$.
233: The coned space $C(G,\mc{P},S)$ is obtained
234: from $\Gamma(G,S)$ by coning each left coset of an element of $\mc{P}$
235: to a point. (Here the $0$--skeleton of $\Gamma(G,S)$ is implicitly
236: identified with $G$.)
237:
238: Recall that a graph is \emph{fine} if every edge is contained in only
239: finitely many circuits (i.e. embedded cycles) of any bounded length.
240: \begin{definition}
241: If $G$ is a group with finite generating set $S$, and a collection of
242: subgroups $\mc{P}=\{P_1,\ldots,P_n\}$, then
243: $G$ is \emph{weakly hyperbolic relative to} $\mc{P}$ if
244: $C(G,\mc{P},S)$ is $\delta$--hyperbolic for some $\delta$.
245:
246: If, moreover, $C(G,\mc{P},S)$ is \emph{fine}, then $G$ is
247: \emph{(strongly) hyperbolic relative to} $\mc{P}$.
248: \end{definition}
249:
250: A special case of weak relative hyperbolicity is bounded generation.
251: The following definition is easily seen to be equivalent to the
252: standard one:
253: \begin{definition}
254: A group $G$ is \emph{boundedly generated} by a collection of
255: subgroups $\mc{P}=\{P_1,\ldots,P_n\}$ if the Cayley graph of $G$ with
256: respect to $\cup\mc{P}$ has finite diameter.
257:
258: If each $P_i$ is cyclic, generated by $g_i$, we say that $G$ is
259: \emph{boundedly generated by} $\{g_1,\ldots,g_n\}$.
260: \end{definition}
261:
262: A cobounded action on an unbounded Gromov hyperbolic space does not
263: have an invariant horoball.
264: It is thus a corollary of Theorem \ref{thm:cdr2} that these higher rank
265: $G(\mathcal{O})$ are not strongly relatively hyperbolic
266: with respect to any system of proper subgroups.
267: \begin{corollary}\label{corollary:notrelhyp}
268: If $L=G(\mc{O})$ is as in Theorem \ref{thm:cdr2}, and $L$ is weakly
269: hyperbolic relative to a system of subgroups
270: $\mc{P}$, then the coned space $C(L,\mc{P},S)$ has finite
271: diameter for any finite generating set $S$.
272:
273: In particular, $L$ is not strongly hyperbolic relative to any system
274: of proper subgroups.
275: \end{corollary}
276: By possibly altering $\mc{P}$ to add some cyclic subgroups, the first
277: part of the corollary can be restated: If $L$ is weakly
278: hyperbolic relative to a system of subgroups $\mc{P}$ which generate
279: $L$, then $L$ is boundedly generated by $\mc{P}$.
280: \begin{remark}
281: Corollary \ref{corollary:notrelhyp} also can be easily deduced from
282: \cite{karlssonnoskov} in some special cases, including $L=SL(n,\Z)$
283: for $n>2$. The second part of Corollary \ref{corollary:notrelhyp}
284: (about strong relative hyperbolicity) can be deduced from known
285: theorems in a number of ways, perhaps most straightforwardly by
286: combining a theorem of Fujiwara \cite{fujiwara:gromovhyperbolic} with
287: one of Burger and Monod \cite{burgermonod:gafa}.
288: \end{remark}
289:
290:
291: \subsection{Outline}
292: In Section \ref{s:preliminaries} we recall some definitions and basic
293: results, first from the theory of Gromov hyperbolic spaces, and second
294: from Chevalley groups over number rings. In Section \ref{s:actions}
295: an equivalence relation amongst hyperbolic $G$--spaces is proposed, and
296: generalized combinatorial horoballs are introduced. In Section
297: \ref{s:elementary}, we improve on the statement and proof of
298: Proposition 3.9 of \cite{manning:qfa}, and use the improved version to
299: characterize hyperbolic $G$--spaces with invariant horoballs.
300: In Section \ref{s:bigrank} we prove Theorem \ref{thm:cdr2}, and in
301: Section \ref{section:rankone} we prove Theorem \ref{thm:cdr1}.
302:
303: \subsection{Acknowledgments}
304: The author thanks Benson Farb and Hee Oh for useful conversations, and
305: Nicolas Monod for helpful comments on an earlier version of this paper.
306: This work was partly supported by an NSF Postdoctoral Research
307: Fellowship (DMS-0301954).
308:
309: \section{Preliminaries}\label{s:preliminaries}
310: \subsection{Coarse geometry}
311: \begin{definition}
312: If $X$ and $Y$ are metric spaces, $K\geq 1$ and $C\geq 0$,
313: a \emph{$(K,C)$--quasi-isometric embedding} of $X$ into $Y$ is a function
314: $q\co X\to Y$ so that
315: For all $x_1$, $x_2\in X$
316: \[\frac{1}{K}d(x_1,x_2)-C\leq d(q(x_1),q(x_2))\leq Kd(x_1,x_2)+C\]
317:
318: If in addition the map $q$ is \emph{$C$--coarsely onto} -- \emph{i.e.},
319: every $y\in Y$ is
320: distance at most $C$ from some point in $q(X)$ -- then $q$ is called a
321: \emph{$(K,C)$--quasi-isometry}.
322: The two metric spaces $X$ and $Y$ are then
323: said to be \emph{quasi-isometric} to one
324: another. This is a symmetric condition.
325: \end{definition}
326: \begin{definition}
327: A \emph{$(K,C)$--quasi-geodesic} in $X$
328: is a $(K,C)$--quasi-isometric embedding $\gamma\co\R\to X$.
329: We will occasionally abuse
330: notation by referring to the image of $\gamma$ as a quasi-geodesic.
331: \end{definition}
332: \subsection{Gromov hyperbolic spaces}
333: For more details on Gromov hyperbolic metric spaces, see
334: \cite{bridhaef:book} and \cite{gromov:wordhyperbolic}.
335: A number of equivalent definitions of Gromov hyperbolicity are known.
336: For geodesic spaces, we will use the one based on the existence of
337: \emph{comparison tripods}.
338: Given a geodesic
339: triangle $\Delta(x,y,z)$ in
340: any metric space, there is a unique \emph{comparison tripod}, $T_\Delta$,
341: a metric tree so
342: that the distances between the three extremal points of the tree,
343: $\overline{x}$, $\overline{y}$ and $\overline{z}$ , are the same as
344: the distances between $x$, $y$ and $z$.
345: There is an obvious map $\pi\co \Delta(x,y,z)\to T_\Delta$ which takes $x$ to
346: $\overline{x}$, $y$ to $\overline{y}$ and $z$ to $\overline{z}$, and
347: which is an isometry on each side of $\Delta(x,y,z)$.
348: \begin{definition}\label{d:geodgromov}
349: A geodesic space $X$ is \emph{$\delta$--hyperbolic} if for any geodesic triangle
350: $\Delta(x,y,z)$ and any point $p$ in the comparison tripod $T_\Delta$,
351: the diameter of $\pi^{-1}(p)$ is less than $\delta$. If $\delta$ is
352: unimportant we may simply say that $X$ is \emph{Gromov hyperbolic}.
353: \end{definition}
354: Gromov hyperbolicity (of geodesic spaces) is a quasi-isometry invariant.
355:
356: The following proposition about stability of quasi-geodesics in Gromov
357: hyperbolic spaces
358: is well-known (see, e.g. \cite[III.H.1.7]{bridhaef:book}).
359: \begin{proposition}\label{p:quasistable}
360: Let $K\geq 1$, $C\geq 0$, $\delta\geq 0$. Then there is some
361: $B=B(K,C,\delta)$, so that whenever $\gamma$ and $\gamma'$ are two
362: $(K,C)$--quasi-geodesics with the same endpoints in a
363: $\delta$--hyperbolic geodesic metric space $X$, then the Hausdorff
364: distance between $\gamma$ and $\gamma'$ is at most $B$.
365: \end{proposition}
366:
367:
368: We will occasionally need to deal with spaces which are not geodesic.
369: If $X$ is $\delta$--hyperbolic in the sense of Definition
370: \ref{d:geodgromov}, then it
371: satisfies the four point condition: For all $p_1$, $p_2$, $p_3$,
372: $p_4\in X$,
373: \begin{equation}\label{fourpoint}
374: d(p_1,p_4)+d(p_2,p_3)\leq\max\{d(p_1,p_2)+d(p_3,p_4)\mbox{, }d(p_1,p_3)+d(p_2,p_4)+2\delta\}.
375: \end{equation}
376: Conversely, if a geodesic space satisfies \eqref{fourpoint}, then it
377: is $6\delta$--hyperbolic in the sense of Definition \ref{d:geodgromov}.
378: (For both these facts, and the below definition, see
379: \cite[III.H.1.22]{bridhaef:book} or \cite{gromov:wordhyperbolic}.)
380: \begin{definition}\label{d:fourpoint}
381: A (not necessarily geodesic) metric space $X$ is \emph{$(\delta)$--hyperbolic}
382: if it satisfies the condition \eqref{fourpoint} above. If $\delta$ is
383: unimportant, we simply say that $X$ is \emph{Gromov hyperbolic}.
384: \end{definition}
385:
386: In order to describe the boundary of a Gromov hyperbolic space, we
387: introduce the \emph{Gromov product} notation.
388: \begin{definition}
389: If $x,y$ and $z$ are points in a metric space with a metric
390: $d(\cdot,\cdot)$, then
391: \[ \gp{x}{y}{z}:= \frac{1}{2}(d(x,z)+d(y,z)-d(x,y)).\]
392: \end{definition}
393: We should remark that in the context of Definition \ref{d:geodgromov},
394: $\gp{x}{y}{z}$ is the distance from
395: $\overline{z}$ to the central vertex of the comparison tripod for a
396: geodesic triangle with vertices $x$, $y$ and $z$.
397: \begin{definition}
398: Let $p\in X$ where $X$ is a Gromov hyperbolic space.
399: A sequence of points $\{x_i\}$ in a Gromov hyperbolic space
400: \emph{tends to infinity} if
401: $\lim_{i,j\to\infty}\gp{x_i}{x_j}{p}=\infty$.
402: Two such sequences are \emph{equivalent}, written $\{x_i\}\sim\{y_i\}$, if
403: $\lim_{i,j\to\infty}\gp{x_i}{y_j}{p}=\infty$.
404: The boundary $\partial X$ is the set of equivalence classes of
405: sequences which tend to infinity.
406: \end{definition}
407: The Gromov product extends (by taking a lim sup) to sequences which
408: tend to infinity, and this allows convergence in $\partial X$ to be
409: defined, giving $\partial X$ a natural topology.
410:
411: \subsection{Isometries of Gromov hyperbolic spaces}
412: Isometries of geodesic hyperbolic spaces can be classified into three
413: types.
414: \begin{definition}
415: Let $f\co X\to X$ be an isometry. If $x\in X$, we let
416: $O_x=\{f^n(x)\mid n\in \Z\}$.
417: We say that $f$ is \emph{elliptic} if $O_x$ is bounded. We say that
418: $f$ is \emph{hyperbolic} if
419: $n\mapsto f^{n}(x)$ is a quasi-isometric embedding of $\Z$ into $X$.
420: We say that $f$ is \emph{parabolic} if $O_x$ has a unique limit point
421: in $\partial G$.
422: \end{definition}
423: The following was observed by Gromov \cite[8.1.B]{gromov:wordhyperbolic}.
424: Although it is often stated with an extra hypothesis of
425: properness, this hypothesis is unnecessary (See, for example the
426: proof in \cite[Chapitre 9]{cdp}, where the extra hypothesis is given,
427: but not used).
428: \begin{proposition}\label{p:isometryclass}
429: Every isometry of a geodesic Gromov hyperbolic space is elliptic,
430: parabolic, or hyperbolic.
431: \end{proposition}
432:
433:
434:
435: \begin{lemma}\label{l:commute}
436: Suppose that $G$ acts on the geodesic Gromov hyperbolic space $X$, and that
437: $p\in G$ acts parabolically, fixing $e\in \partial X$. If $g\in G$
438: commutes with $p$, then $g$ also fixes $e$.
439: \end{lemma}
440: \begin{proof}
441: Let $e'=g(e)$. We have $e'=gp(e)=pg(e)=p(e')$, so $p$ fixes $e'$.
442: Since $p$ is parabolic, it fixes a unique point in $\partial X$, and
443: so $e'=e$.
444: \end{proof}
445:
446: \begin{definition}
447: Let $G$ be a finitely generated group, and let $g\in G$. If $n\mapsto
448: g^n$ is a quasi-isometric embedding, we say that $g$ is
449: \emph{undistorted}. Otherwise, $g$ is \emph{distorted}.
450: \end{definition}
451: The proof of the following observation is left to the reader.
452: \begin{lemma}\label{lemma:distortion}
453: Let the finitely generated group $G$ act by isometries on a Gromov
454: hyperbolic space $X$. If $g\in
455: G$ acts hyperbolically on $X$, then $g$ is undistorted.
456: \end{lemma}
457:
458: \subsection{Chevalley groups}\label{defs:chevalley}
459: In this section we recall the definition of a Chevalley group over a
460: commutative ring.
461: (All rings are assumed to have $1 \neq 0$.)
462: The simplest example of a Chevalley group is
463: $SL(n,\Z)$. If one thinks of $SL(n,\Z)$ as being the ``$\Z$--points
464: of $SL(n,\C)$,'' then the Chevalley-Demazure group scheme
465: identifies what the ``$R$--points of $G$'' are, where
466: $R$ is now allowed to be an arbitrary ring, and $G$ an arbitrary
467: complex semisimple Lie group. It turns out that this idea is not entirely
468: well-formed, until one fixes an embedding of $G$ into $GL(n,\C)$ for
469: some $n$. The following exposition is largely
470: adapted from \cite{abe:chevalley} and \cite{tavgen:bg}.
471:
472: Let $\rho\co G\to GL(V)$ be a representation of a connected complex semisimple
473: Lie group into the general linear group of a complex vector space $V$
474: of dimension $n$. We will assume that
475: $d\rho\co\mathfrak{g}\to\mathfrak{gl}(V)$ is faithful (Here
476: $\mathfrak{g}$ is the Lie algebra of $G$, $\mathfrak{gl}(V)$ the Lie
477: algebra of endomorphisms of $V$.
478: Let $\mathfrak{h}$ be a Cartan subalgebra of $\mathfrak{g}$ and let
479: $\Phi$ be the (reduced) root system relative to $\mathfrak{h}$.
480: Let $\Delta$ be a
481: choice of simple roots. Then there is a \emph{Chevalley basis}
482: for $\mathfrak{g}$ (see \cite{steinberg:notes} or \cite{chevalley:tohoku}) of
483: the form $\mathcal{B}=\{X_\alpha\ |\ \alpha\in\Phi\}\cup\{H_\alpha\ |\
484: \alpha\in\Delta\}$, so that the $H_\alpha$ generate $\mathfrak{h}$ (and
485: thus commute) and the structure constants are all integral.
486: In other words, the $\Z$--span of $\mathcal{B}$ is actually a Lie algebra over
487: $\Z$.
488: In particular, if $\{\beta - r_{\beta,\alpha} \alpha, \ldots \beta,
489: \ldots \beta + q_{\beta,\alpha} \alpha\}$ are all the roots on the
490: line $\{\beta + t\alpha\ |\ t\in\Z\}$, then:
491: \begin{enumerate}
492: \item $[X_\alpha, X_{-\alpha}] = H_\alpha$.
493: \item $[H_\alpha, X_\beta] = A(\alpha,\beta) X_\beta$, where
494: $A(\alpha,\beta)$ is an integer determined by $\alpha$
495: and $\beta$.
496: \item $[X_\alpha,X_\beta] = 0$ if $\alpha + \beta$ is not in $\Phi$;
497: otherwise $[X_\alpha,X_\beta] = \pm (r_{\beta,\alpha}+1)X_{\alpha+\beta}$.
498: \end{enumerate}
499: Notice that the $X_\alpha$ are all ad-nilpotent, and that the
500: $H_\alpha$ are ad-semisimple.
501:
502: It can be shown (see, e.g. \cite[\S27]{humphreys:introduction})
503: that $V$ contains
504: an ``admissible lattice'' $V_\Z$: This is a free $\Z$--module in $V$
505: which is invariant under
506: $(d\rho(X_\alpha))^m / m!$ for any $\alpha$ and any $m$. Let
507: $\{v_1,\ldots,v_n\}$ be a basis for $V_\Z$. In terms of this basis,
508: $\rho(g)$ can be written as an $n$ by $n$ matrix with complex
509: entries for any $g$. Let $x_{ij}\co G\to \C$ be the function which
510: simply reads off the $ij$th entry of this matrix. The $x_{ij}$
511: generate an affine complex algebra $\C(G)$. Let $\Z(G)$ denote the
512: $\Z$--algebra with the same generators.
513:
514: We can endow $\Z(G)$ with a Hopf algebra structure by defining
515: a comultiplication $\mu^*$ by
516: $\mu^*(x_{ij}) = \sum_k x_{ik}\otimes x_{kj}$, a counit $\epsilon$
517: by $\epsilon(x_{ij}) = \delta_{ij}$, and an antipode $s$ by
518: $s(f)(g) = f(g^{-1})$.
519: (That $s$ maps $\Z(G)$ into itself uses the fact
520: that $\det(\rho(g))=\pm 1$ for every $g\in G$ -- this
521: follows from semisimplicity.)
522: Since $\Z(G)$ is a Hopf algebra over $\Z$, it defines a functor from
523: rings to groups as follows:
524: For any ring $R$, let
525: $G(R) = \mathrm{Hom}_\Z(\Z(G), R)$. (Note that the elements of
526: $G(R)$ are $\Z$--algebra homomorphisms, so they send $1$ to $1$.)
527: We define a group operation
528: $\bullet$ on $G(R)$ by $\rho\bullet\sigma = (\rho\otimes\sigma)\circ
529: \mu^*$.
530: (In other words, $\rho\bullet\sigma(x_{ij}) = \sum_k
531: \rho(x_{ik})\sigma(x_{kj})$.)
532:
533: Some observations:
534: \begin{enumerate}
535: \item $G(\C)$ can be naturally identified with the image of $\rho\co
536: G\to GL(V)$ -- if $\rho$ is assumed to be faithful, then clearly
537: $G(\C)\cong G$ doesn't depend on $\rho$. On the other hand, for $R$
538: arbitrary, $G(R)$ does depend on $\rho$.
539: \item The assignment $R\mapsto G(R)$ is a covariant functor from
540: commutative rings to groups. This functor is often called a
541: ``Chevalley-Demazure group scheme.''
542: \item\label{zoft} Any Hopf algebra gives such a functor. An
543: important example is
544: $\Z[t]$, with a Hopf ($\Z$--)algebra structure so that:
545: \begin{enumerate}
546: \item The comultiplication satisfies $\mu^*(t)=t\otimes 1 + 1\otimes
547: t$.
548: \item The counit satisfies $\epsilon(t)=0$.
549: \item The antipode satisfies $S(t)=-t$
550: \end{enumerate}
551: In this case the functor $R\mapsto \mathrm{Hom}(Z[t],R)$ is the
552: ``forgetful'' functor, which takes a ring to its underlying Abelian
553: group.
554: \item\label{contrafunctor} Since $G(R)$ is $Hom_\Z(\Z(G),R)$, any morphism of Hopf algebras
555: $\Z(G)\to A$ where $A$ is
556: some other Hopf algebra, will give rise to a homomorphism of groups
557: $Hom_\Z(A,R)\to G(R)$.
558: \end{enumerate}
559: \begin{definition}
560: Let $\alpha$ be a root. Then since $d\rho(X_\alpha)$ is
561: nilpotent, the formal sum $\exp ( t X_\alpha ) = \sum_{m=1}^\infty
562: t^m \frac{d\rho(X_\alpha)^m}{m!} $ is a matrix with entries which are
563: polynomials in $t$.
564: Because $\frac{d\rho(X_\alpha)^m}{m!}$ preserves
565: $V_\Z$, each $x_{ij}(\frac{d\rho(X_\alpha)^m}{m!})$ is an integer.
566: Thus we get a map
567: \[\mathrm{ev}_\alpha\co \Z(G) \to \Z[t],\]
568: which sends $x_{ij}$ to the (integral) polynomial in $t$ which appears
569: in the $ij$'th place of the matrix $\exp( t X_\alpha)$. The map
570: $\mathrm{ev}_\alpha$ is a morphism of Hopf algebras, where $\Z[t]$ is given a
571: Hopf algebra structure as in observation \eqref{zoft} above.
572: This morphism gives rise via observation \eqref{contrafunctor} to a
573: homomorphism of the additive group underlying $R$ into the group
574: $G(R)$
575: \[x_\alpha\co R\to G(R).\]
576: The image of this map is the \emph{root subgroup}
577: of $G(R)$ corresponding to $\alpha$.
578: \end{definition}
579:
580: \begin{definition}
581: The subgroup of $G(R)$ generated by the root subgroups
582: is denoted $E(R)$.
583: \end{definition}
584:
585: In this paper, we focus mainly on the special case that $R$ is the
586: ring of integers of a number field $k$ and $\Phi$ has rank at least
587: two. In this case, $G(R)=E(R)$ by
588: a result of Matsumoto \cite{matsumoto:csp}. By a result
589: of Carter and Keller \cite{carterkeller:slno} in case $\Phi=A_n$,
590: and Tavgen$'$ \cite{tavgen:bg} in general,
591: $G(R)$ is boundedly generated by the root subgroups.
592:
593: It should also be noted that in this special case, $G(R)$ is actually
594: an irreducible lattice in a semisimple Lie group. Indeed, if $s$ and
595: $t$ are the number of real
596: and complex places of $k$, respectively, then $G(R)$ is a lattice in the
597: Lie group
598: \[ G(\R)^s\times G(\C)^t.\]
599:
600: \section{Equivalence of actions}\label{s:actions}
601: We wish to study the variety of actions of a group $G$ on
602: Gromov hyperbolic spaces, up to some kind of coarse equivalence. By a
603: {\em (Gromov) hyperbolic $G$--space}, we will always mean a Gromov
604: hyperbolic geodesic metric space, equipped with an isometric
605: $G$--action.
606:
607:
608: \begin{definition}
609: Let $X$ and $Y$ be Gromov hyperbolic $G$--spaces. We say that $X$ and
610: $Y$ are {\em equivalent} if they lie in the same
611: equivalence class, under the equivalence relation generated by
612: coarsely equivariant quasi-isometric embeddings.
613: \end{definition}
614: The following proposition should serve to clarify this equivalence
615: relation.
616: \begin{proposition}If $X_1$ and $X_2$ are equivalent Gromov hyperbolic
617: $G$--spaces, then there is
618: a third Gromov hyperbolic $G$--space $V$ which coarsely equivariantly
619: quasi-isometrically embeds in both $X_1$ and $X_2$.
620: \end{proposition}
621: \begin{proof}
622: The key claim is the following:
623: \begin{claim}\label{c:diamond}
624: If $V$ and $W$ are hyperbolic $G$--spaces which coarsely equivariantly
625: quasi-isometrically embed in a third hyperbolic $G$--space $X$, then
626: there is a fourth hyperbolic $G$--space $A$ which coarsely
627: equivariantly quasi-isometrically embeds into $V$ and $W$.
628: \end{claim}
629: Before proving the claim, let us see how it implies the proposition.
630: If $X_1$ and $X_2$ are equivalent hyperbolic $G$--spaces, they
631: must be joined by a sequence of coarsely
632: equivariant quasi-isometric embeddings:
633: \cdlabel{long}{ & V_1\ar[dl]\ar[dr] & & \cdots\ar[dl]\ar[dr] & &V_n\ar[dl]\ar[dr] & &
634: V_{n+1}\ar[dl]\ar[dr] & \\
635: X_1 & & Z_1 &\cdots & Z_{n-1} & & Z_n & & X_2}
636: By applying the claim with $V=V_n$, $W=V_{n+1}$, and $X=Z_n$, we
637: obtain a hyperbolic $G$--space $A$ which coarsely equivariantly
638: quasi-isometrically embeds into both $V_n$ and $V_{n+1}$, and hence
639: into both $Z_{n-1}$ and $X_2$. We can thus shorten the sequence
640: \eqref{long}, unless it is a shortest possible such sequence,
641: \cd{ & V_1\ar[dl]\ar[dr] & \\
642: X_1 & & X_2,}
643: in which case the proposition is verified.
644:
645: \begin{proof} (of Claim \ref{c:diamond})
646: We first construct a $G$--space $A_1$ which coarsely equivariantly
647: quasi-isometrically embeds in both $V$ and $W$ and is Gromov
648: hyperbolic but not geodesic. We then show that $A_1$ is
649: quasi-isometric to a geodesic $G$--space $A$.
650:
651: Choose $\delta>0$ so that all of $V$, $W$, and $X$ are
652: $\delta$--hyperbolic spaces.
653: By hypothesis, we may choose $K\geq 1$ and $C\geq 0$ so that there are maps
654: $\phi\co V\to X$ and $\psi\co W\to Y$ which are $C$--coarsely equivariant
655: $(K,C)$--quasi-isometric embeddings. We choose constants $J_0<J_1<J_2$:
656: Let
657: $J_0 = 2B(K,C,\delta)+2\delta$,
658: where $B(K,C,\delta)$ is the constant of quasi-geodesic stability from
659: Proposition \ref{p:quasistable}, let $J_1=J_0+2C$, and let $J_2=4 J_1$.
660:
661: Let $A_0$ be the subset of $V\times
662: W$ given by $A_0=\{(v,w)\mid d(\phi(v),\psi(w))\leq J_0\}$, and let $A_1$
663: be the smallest $G$--equivariant subset of $V\times W$ containing
664: $A_0$. We endow $A_1$ with the (pseudo)metric
665: \[
666: d((v_1,w_1),(v_2,w_2))=d(\phi(v_1),\phi(v_2))+d(\psi(v_2),\psi(w_2)).
667: \]
668: \begin{subclaim}
669: $A_1$ is Gromov hyperbolic.
670: \end{subclaim}
671: \begin{proof}
672: Since $A_1$ is not a geodesic space, we must work with the four-point
673: definition \ref{d:fourpoint}. We will show that
674: $A_1$ is $(2\delta+4J_1)$--hyperbolic in the sense of Definition
675: \ref{d:fourpoint}.
676: Let $\{p_i=(v_i,w_i)\mid i=1,\ldots,4\}$ be four points in $A_1$.
677: For each $i$, we write $\uu{v}_i$ for $\phi(v_i)$ and $\uu{w}_i$ for
678: $\psi(w_i)$. By
679: reordering the points if necessary, we can assume that
680: \[d(\uu{v}_1,\uu{v}_3)+d(\uu{v}_2,\uu{v}_4)\geq
681: d(\uu{v}_1,\uu{v}_2)+d(\uu{v}_3,\uu{v}_4),\]
682: as in Figure \ref{f:palimpsest}.
683: \begin{figure}[htbp]
684: \begin{center}
685: \begin{picture}(0,0)%
686: \epsfig{file=palimpsest.pstex}%
687: \end{picture}%
688: \setlength{\unitlength}{3947sp}%
689: %
690: \begingroup\makeatletter\ifx\SetFigFont\undefined%
691: \gdef\SetFigFont#1#2#3#4#5{%
692: \reset@font\fontsize{#1}{#2pt}%
693: \fontfamily{#3}\fontseries{#4}\fontshape{#5}%
694: \selectfont}%
695: \fi\endgroup%
696: \begin{picture}(3300,1939)(2551,-2744)
697: \put(5551,-2686){\makebox(0,0)[lb]{\smash{{\SetFigFont{12}{14.4}{\familydefault}{\mddefault}{\updefault}{\color[rgb]{0,0,0}$\phi(v_3)$}%
698: }}}}
699: \put(2776,-961){\makebox(0,0)[lb]{\smash{{\SetFigFont{12}{14.4}{\familydefault}{\mddefault}{\updefault}{\color[rgb]{0,0,0}$\phi(v_2)$}%
700: }}}}
701: \put(2626,-1186){\makebox(0,0)[lb]{\smash{{\SetFigFont{12}{14.4}{\familydefault}{\mddefault}{\updefault}{\color[rgb]{0,0,0}$\psi(w_2)$}%
702: }}}}
703: \put(2551,-2011){\makebox(0,0)[lb]{\smash{{\SetFigFont{12}{14.4}{\familydefault}{\mddefault}{\updefault}{\color[rgb]{0,0,0}$\phi(v_1)$}%
704: }}}}
705: \put(2626,-2236){\makebox(0,0)[lb]{\smash{{\SetFigFont{12}{14.4}{\familydefault}{\mddefault}{\updefault}{\color[rgb]{0,0,0}$\psi(w_1)$}%
706: }}}}
707: \put(5851,-1411){\makebox(0,0)[lb]{\smash{{\SetFigFont{12}{14.4}{\familydefault}{\mddefault}{\updefault}{\color[rgb]{0,0,0}$\phi(v_4)$}%
708: }}}}
709: \put(5776,-1186){\makebox(0,0)[lb]{\smash{{\SetFigFont{12}{14.4}{\familydefault}{\mddefault}{\updefault}{\color[rgb]{0,0,0}$\psi(w_4)$}%
710: }}}}
711: \put(5776,-2461){\makebox(0,0)[lb]{\smash{{\SetFigFont{12}{14.4}{\familydefault}{\mddefault}{\updefault}{\color[rgb]{0,0,0}$\psi(w_3)$}%
712: }}}}
713: \end{picture}%
714: \caption{Points in $X$.}
715: \label{f:palimpsest}
716: \end{center}
717: \end{figure}
718: There are then two cases.
719:
720: In case
721: \[ d(\uu{w}_1,\uu{w}_3)+d(\uu{w}_2,\uu{w}_4)\geq
722: d(\uu{w}_1,\uu{w}_2)+d(\uu{w}_3,\uu{w}_4),\]
723: then
724: $\max\{d(p_1,p_3)+d(p_2,p_4),d(p_1,p_2)+d(p_3,p_4)\}=d(p_1,p_2)+d(p_2,p_4)$.
725: Applying $\delta$--hyperbolicity in $X$, we obtain
726: \begin{eqnarray*}
727: d(p_1,p_4)+d(p_3,p_2) & = &
728: d(\uu{v}_1,\uu{v}_4)+d(\uu{v}_2,\uu{v}_3)+d(\uu{w}_1,\uu{w}_4)+d(\uu{w}_2,\uu{w}_3)
729: \\
730: & \leq & d(\uu{v}_1,\uu{v}_3)+d(\uu{v}_2,\uu{v}_4)+ 2\delta
731: +d(\uu{w}_1,\uu{w}_3)+d(\uu{w}_2,\uu{w}_4)+ 2\delta \\
732: & = & d(p_1,p_3)+d(p_2,p_4)+ 4\delta \\
733: & = & \max\{d(p_1,p_3)+d(p_2,p_4),d(p_1,p_2)+d(p_3,p_4)\} + 2
734: (2\delta) \\
735: & \leq & \max\{d(p_1,p_3)+d(p_2,p_4),d(p_1,p_2)+d(p_3,p_4)\} \\
736: & & + 2(2\delta+4 J_1)
737: \end{eqnarray*}
738: as required.
739:
740: If on the other hand
741: \[ d(\uu{w}_1,\uu{w}_3)+d(\uu{w}_2,\uu{w}_4) <
742: d(\uu{w}_1,\uu{w}_2)+d(\uu{w}_3,\uu{w}_4),\]
743: then it is still true (since $d(\uu{w_i},\uu{v_i})\leq J_1)$) that
744: \begin{eqnarray*}
745: d(\uu{w}_1,\uu{w}_3)+d(\uu{w}_2,\uu{w}_4) & \geq &
746: d(\uu{v}_1,\uu{v}_3)+d(\uu{v}_2,\uu{v}_4) - 4J_1 \\
747: & \geq & d(\uu{v}_1,\uu{v}_2)+d(\uu{v}_3,\uu{v}_4) - 4J_1 \\
748: & \geq & d(\uu{w}_1,\uu{w}_2)+d(\uu{w}_3,\uu{w}_4) - 8J_1.
749: \end{eqnarray*}
750: We therefore obtain
751: \begin{eqnarray*}
752: d(p_1,p_4)+d(p_3,p_2) & = &
753: d(\uu{v}_1,\uu{v}_4)+d(\uu{v}_2,\uu{v}_3)+d(\uu{w}_1,\uu{w}_4)+d(\uu{w}_2,\uu{w}_3)
754: \\
755: & \leq & d(\uu{v}_1,\uu{v}_3)+d(\uu{v}_2,\uu{v}_4)+ 2\delta
756: +d(\uu{w}_1,\uu{w}_2)+d(\uu{w}_3,\uu{w}_4)+ 2\delta \\
757: & \leq & d(p_1,p_3)+d(p_2,p_4) + 2(2\delta+4 J_1) \\
758: & \leq & \max\{d(p_1,p_3)+d(p_2,p_4),d(p_1,p_2)+d(p_3,p_4)\}\\& & + 2
759: (2\delta+4 J_1),
760: \end{eqnarray*}
761: which finishes the proof of the Subclaim.
762: \end{proof}
763: Though the space $A_1$ is hyperbolic, and $G$ quasi-acts on $A_1$
764: via the diagonal action on $V\times W$, the space $A_1$ is not geodesic,
765: and $G$ does not act by isometries. Both of these issues can be fixed
766: at once, by replacing $A_1$ by an appropriate graph.
767: Specifically, we let $A$ be the graph with vertex set
768: $V(A)=A_1$, and with an edge between every pair of vertices $p$ and
769: $q$ so that there exists some $g$ with $d(gp,gq)\leq J_2:=4J_1$. Clearly
770: this graph is a geodesic $G$--space.
771:
772: Let $\iota\co A_1\to A$ be the map which takes a point to the
773: corresponding vertex. We claim that $\iota$ is a quasi-isometry, and
774: so $A$ is a Gromov hyperbolic space.
775:
776: Let $a=(v_a,w_a)$ and $b=(v_b,w_b)$ be points in $A_1$ so that
777: $d(\iota(a),\iota(b))=1$. There is some $g$ with
778: \[d(ga,gb)=d(\phi(gv_a),\phi(gv_b))+d(\psi(gv_a),\psi(gv_b))\leq J_2.\]
779: Since $\phi$ and $\psi$ are both $C$--coarsely equivariant,
780: $d(\phi(gv_a),\phi(gv_b))$ differs by
781: at most $2C$ from $d(\phi(v_a),\phi(v_b))$; similarly, $d(\psi(gw_a),\psi(gw_b))$ differs from $d(\psi(w_a),\psi(w_b))$ by
782: at most $2C$. It follows that $d(ga,gb)$ differs by at most $4C$ from
783: $d(a,b)$, and so $d(a,b)\leq J_2+4C<\frac{3}{2}J_2$. As this is true for every pair
784: of points connected by an edge, we deduce
785: \begin{equation}\label{lower}
786: d(\iota(p),\iota(q)) \geq \frac{2}{3J_2} d(p,q)
787: \end{equation}
788: for any pair of points $p$, $q\in A_1$.
789:
790: We now obtain the complementary bound to \eqref{lower}. Suppose $p=(v_1,w_1)$
791: and $q=(v_2,w_2)$ are any two points in $A_1$. We write $\uu{v_1}$
792: for $\phi(v_1)$ and so on as before.
793: We will show that
794: \begin{equation}\label{upper}
795: d(\iota(p),\iota(q))\leq \frac{1}{J_1}d(p,q)+2
796: \end{equation}
797: by constructing a path joining $\iota(p)$ to $\iota(q)$ in $A$. The
798: vertices of this path will be points in $A_0\subset V\times W$ so that the first
799: coordinate lies on a geodesic between $v_1$ and $v_2$, while the
800: second lies on a geodesic between $w_1$ and $w_2$.
801:
802: If $d(p,q)\leq 2J_2 = 8J_1$, then \eqref{upper} is automatically
803: satisfied. We may therefore assume that $d(p,q)\geq 2J_2$. Because
804: $d(\uu{v}_1,\uu{w}_1)$ and $d(\uu{w}_1,\uu{w}_2)$ are at most $J_0$,
805: we have
806: $\min\{d(\uu{v}_1,\uu{v}_2),d(\uu{w}_1,\uu{w}_2)\}\geq J_2-J_0 >
807: 3J_0$. It follows that
808: $\gp{\uu{v}_2}{\uu{w}_2}{\uu{v}_1}>\gp{\uu{w}_1}{\uu{w}_2}{\uu{v}_1}$, as in
809: Figure \ref{f:bar}.
810: \begin{figure}[htbp]
811: \begin{center}
812: \input{bar.pstex_t}
813: \caption{A pair of pairs of points in $X$, together with comparison tripods.
814: The dashed line on the left is the image of a geodesic in $V$; the one
815: on the right is the image of a geodesic in $W$.}
816: \label{f:bar}
817: \end{center}
818: \end{figure}
819: In fact,
820: $\gp{\uu{v}_2}{\uu{w}_2}{\uu{v}_1}-\gp{\uu{w}_1}{\uu{w}_2}{\uu{v}_1}>J_0\geq
821: J_1/2$ (since clearly $B(K,C,\delta)\geq C$).
822: We may thus choose real numbers
823: \[ \gp{\uu{w}_1}{\uu{w}_2}{\uu{v}_1}=t_0<t_1<\cdots<t_k =
824: \gp{\uu{v}_2}{\uu{w}_2}{\uu{v}_1}\]
825: satisfying $\frac{J_1}{2}<|t_{i+1}-t_i|\leq J_1$ for each $i$.
826: Note that $k \leq \frac{1}{J_1}d(p,q)$ for such a choice.
827:
828: Let
829: $\gamma$ be a unit speed geodesic from $\uu{v}_1$ to $\uu{w}_2$. For
830: each $i$ between $0$ and $k$, there are points $x_i'$ on a geodesic
831: between $\uu{v}_1$ and $\uu{w}_1$ and $y_i'$ on a geodesic between
832: $\uu{w}_1$ and $\uu{w}_2$ which are distance at most $\delta$ from
833: $\gamma(t_i)$. Since $V$ and $W$ are geodesic spaces, there are
834: geodesics $[v_1,v_2]$ in $V$ and $[w_1,w_2]$ in $W$. The maps $\phi$
835: and $\psi$ send these geodesics
836: to $(K,C)$--quasi-geodesics in $X$. Applying quasi-geodesic stability
837: (Proposition \ref{p:quasistable}), there are points
838: $x_i\in [v_1,v_2]$, $y_i\in [w_1,w_2]$ so that $\phi(x_i)$ and
839: $\psi(y_i)$ lie within $B+\delta$ of $\gamma(t_i)$. Since
840: $d(\phi(x_i),\psi(y_i))\leq 2B+2\delta=J_0$, the point $p_i=(x_i,y_i)$
841: lies in $A_0\subseteq A_1$. Moreover, $d(p_i,p_{i+1})\leq
842: 2(2(B+\delta)+J_1)<J_2$, and so $d(\iota(p_i),\iota(p_{i+1}))\leq 1$.
843:
844: Finally, one notes that
845: \[
846: d(\uu{v}_1,\phi(x_0))+d(\uu{w}_1,\psi(y_0)) \leq J_0+2\delta+2 B=2J_0<J_2
847: \]
848: and likewise for $d(\uu{v}_1,\phi(x_0))+d(\uu{w}_1,\psi(y_0))$, from
849: which it follows that $d(\iota(p),\iota(p_0))$ and
850: $d(\iota(q),\iota(p_k))$ are at most one. Thus
851: \[ d(\iota(p),\iota(q))\leq k+2 \leq \frac{1}{J_1} d(p,q)+2.\]
852: It is obvious that $\iota$ is $1$--almost onto, and so $\iota$ is a
853: quasi-isometry.
854:
855: It follows that $A$ is a Gromov hyperbolic geodesic $G$--space
856: which coarsely equivariantly, quasi-isometrically embeds into $V$ and
857: $W$.
858:
859: \end{proof}
860: \end{proof}
861:
862: \begin{remark}
863: We record some observations about this equivalence:
864: \begin{enumerate}
865: \item Let $\Gamma$ be a group acting isometrically on $\H^2$. This
866: action extends in an obvious way to either $\H^3$ or $\C\H^2$.
867: Although there is no quasi-isometric embedding either from $\H^3$ to
868: $\C\H^2$ or vice versa, these actions are equivalent under the
869: equivalence relation.
870: \item A Gromov hyperbolic $G$--space has a bounded orbit if and only if
871: it is equivalent to the trivial $G$--space consisting of a single
872: point.
873: \item If $X$ and $Y$ are equivalent Gromov hyperbolic $G$--spaces,
874: $x\in X$ and $y\in Y$, then the
875: limit sets $\Lambda(X)=\{e\in \partial X\cap \overline{Gx}\}$ and
876: $\Lambda(Y)=\{e\in \partial Y\cap \overline{Gy}\}$ are equivariantly
877: homeomorphic.
878: \item The equivalence is perfectly well-defined in the more general
879: setting of {\em quasi-actions} on geodesic Gromov hyperbolic
880: spaces. Call a geodesic Gromov hyperbolic space with a
881: $G$--quasi-action a {\em hyperbolic quasi-$G$--space}. Every
882: hyperbolic quasi-$G$--space is equivalent to some hyperbolic
883: $G$--space.
884: \end{enumerate}
885: \end{remark}
886:
887: Given a group acting on a hyperbolic $G$--space $X$ and some basepoint
888: $x_0\in X$, one obtains a metric on $G$ given by the formula
889: \[d(g,h) = d(x_0,g^{-1}hx_0).\]
890: This metric is obviously determined by its values on $\{1\}\times G$.
891: The compact-open topology on real-valued functions on $G$ thus induces
892: a topology on the set of pointed hyperbolic $G$--spaces.
893: The quotient topology on the space of equivalence classes
894: of Gromov hyperbolic $G$--spaces is not Hausdorff. For example, if
895: $G$ is a surface
896: group, then all of Teichm\"uller space is identified to a single
897: point, whose closure contains many inequivalent actions of $G$ on
898: $\R$--trees.
899:
900:
901: \subsection{Combinatorial horoballs}\label{ss:combhoro}
902: Combinatorial horoballs of the simplest possible type were defined in
903: \cite{rhds}, and used as building blocks for complexes naturally
904: associated to relatively hyperbolic groups. The point there as here
905: is that these spaces can be used to ``hide'' an action on a
906: non-hyperbolic space in an action on a hyperbolic space. There is
907: some flexibility as to how this can
908: happen which is deliberately ignored in \cite{rhds}; here we give a
909: more general construction.
910:
911: \begin{definition}\label{d:admissible}
912: Let $X$ be a graph, acted on by $G$, and suppose that
913: \[\{B_i\co X^{(0)}\to 2^{X^{(0)}}\}_{i\in \N}\]
914: is a collection of functions. We will say that $B_*$ is
915: \emph{admissible} if it
916: satisfies the following
917: four axioms:
918: \begin{enumerate}
919: \item Connectedness: If $v\in X^{(0)}$, then
920: $B_1(v)=\{w\in X^{(0)}\mid d_X(v,w)\leq 1\}$.
921: \item\label{ax:exp} Exponential growth: Let $v$, $w\in X^{(0)}$ and $n\in \N$. If
922: $v\in B_n(w)$, then $B_n(v)\subset B_{n+1}(w)$.
923: \item Symmetry: Let $v$, $w\in X^{(0)}$ and $n\in \N$.
924: If $v\in B_n(w)$, then $w\in B_n(v)$.
925: \item $G$--equivariance: If $w\in X^{(0)}$, $n\in \N$, and $g\in G$,
926: then $g B_n(w)=B_n(gw)$.
927: \end{enumerate}
928: \end{definition}
929:
930: \begin{definition}\label{d:choroball}
931: Let $B_*$ be a sequence of functions $\{B_i\co X^{(0)}\to
932: 2^{X^{(0)}}\}_{i\in \N}$. The \emph{combinatorial horoball based on
933: $X$ and $B_*$}, or $\mc{H}(X,B_*)$, is the graph defined as follows:
934: \begin{enumerate}
935: \item $\mc{H}(X,B_*)^{(0)} = X^{(0)}\times \N$.
936: \item If $n\in \N$ and $v\in X^{(0)}$, then $(v,n)$ is connected to
937: $(v,n+1)$ by an edge (called a \emph{vertical edge}).
938: \item If $n\in \N$ and $v\in B_n(w)$, then $(v,n)$ is connected to
939: $(w,n)$ by an edge (called a \emph{horizontal edge}).
940: \end{enumerate}
941: \end{definition}
942: We leave the following Lemma as an exercise.
943: \begin{lemma}\label{l:horohyp}
944: Let $G$ act on the graph $X$. If $B_*$ is admissible, then
945: $\mc{H}(X,B_*)$ is Gromov hyperbolic, and the action of $G$ on $X$
946: induces an action of $G$ on $\mc{H}(X,B_*)$.
947: \end{lemma}
948:
949: Because of the flexibility of this construction, a group
950: typically admits many inequivalent actions on horoballs.
951:
952: \begin{definition}\label{d:ihoroball}
953: A Gromov hyperbolic $G$--space {\em has an invariant horoball} if it is
954: equivalent to a $G$--space of the form $\mc{H}(X,B_*)$, for some graph
955: $X$ with a $G$--action, and some family $B_*$ which is admissible.
956: \end{definition}
957: In the next section we give some other characterizations of $G$--spaces
958: with invariant horoballs (Theorem \ref{t:horeq}).
959:
960: \section{Elementary actions and pseudocharacters}\label{s:elementary}
961: In this section we show how an elementary action by $G$ on a hyperbolic space
962: gives rise to a pseudocharacter on $G$ which ``picks out'' the
963: elements which act hyperbolically. We then see that if no element
964: acts hyperbolically, then the action has an invariant horoball.
965:
966: \subsection{The pseudocharacter coming from an elementary action}
967: Recall the following definitions.
968: \begin{definition}
969: An action of $G$ on a hyperbolic space $X$ is \emph{elementary} if it
970: is either equivalent to the trivial action on a point, or if the
971: induced action on $\partial X$ has a global fixed point.\footnote{This definition is slightly more restrictive than the
972: usual one, which allows for a pair of points in $\partial X$ to be
973: preserved.}
974: \end{definition}
975: \begin{definition}
976: A \emph{quasicharacter} (or quasi-morphism) on a group $G$ is a
977: real valued function $q$ on $G$ satisfying
978: \begin{equation}\label{defect}
979: |q(gh)-q(g)-q(h)|<C\mbox{, for all } g,h\in G.
980: \end{equation}
981: The \emph{defect} of a quasicharacter is the
982: smallest $C$ so that \eqref{defect} is satisfied. A
983: \emph{pseudocharacter} (or homogeneous quasi-morphism) is a
984: quasicharacter $p$ which satisfies the additional condition
985: \begin{equation*}
986: p(g^n)=n p(g)\mbox{, for all }g\in G, n\in \Z
987: \end{equation*}
988: \end{definition}
989: In this subsection we show how an elementary action on a hyperbolic
990: space gives rise to a pseudocharacter which is nonzero precisely on
991: the elements which act hyperbolically.
992: We begin by studying ``quasi-horofunctions'' on the space
993: $X$, corresponding to a fixed point at infinity. A
994: quasi-horofunction restricted to an arbitrary orbit will give
995: a quasicharacter, which can then be homogenized to give the desired
996: pseudocharacter.
997:
998: \begin{definition}
999: (cf. \cite[7.5.D]{gromov:wordhyperbolic})
1000: Let $\mathbf{x}=\{x_i\}$ be a sequence tending to infinity in the geodesic
1001: hyperbolic space $X$. The \emph{quasi-horofunction coming from}
1002: $\mathbf{x}$ is the function $\eta_{\mathbf{x}}\co X\to \R$ given by
1003: \[ \eta_{\mathbf{x}}(a) = \limsup_{n\to\infty}(d(a,x_n)-d(x_0,x_n)). \]
1004: \end{definition}
1005:
1006: We use the following observation repeatedly:
1007: \begin{observation}\label{o:fourdelta}
1008: Let $A$, $B$, $C$ and $D$ be four points in the $\delta$--hyperbolic
1009: space $X$. If $\gp{C}{D}{A}$ and $\gp{C}{D}{B}$ are both larger than $d(A,B)$,
1010: then
1011: \[ | (d(B,C) - d(A,C)) - (d(B,D) - d(A,D)) | \leq 4\delta. \]
1012: \end{observation}
1013: The observation \ref{o:fourdelta} implies in particular:
1014: \begin{lemma}\label{l:approxbyterm}
1015: If $a\in X$, and $\mathbf{x}=\{x_i\}$ tends to infinity in $X$, then
1016: for all $n$ sufficiently large,
1017: \[| \eta_{\mathbf{x}}(a)-(d(a,x_n)-d(x_0,x_n))| \leq 4\delta.\]
1018: \end{lemma}
1019:
1020: We now can describe the dependence of $\eta_{\mathbf{x}}$ on the sequence
1021: $\mathbf{x}$.
1022: \begin{lemma}\label{l:key}
1023: Let $\mathbf{x}=\{x_i\}$ and $\mathbf{y}=\{y_i\}$ be two sequences of
1024: points in the geodesic $\delta$--hyperbolic space $X$
1025: which tend to the same point in $\partial X$. For any point $a\in X$,
1026: we have
1027: \[ | \eta_{\mathbf{x}}(a)-\eta_{\mathbf{y}}(a)-\eta_{\mathbf{x}}(y_0)|
1028: \leq 16\delta. \]
1029: \end{lemma}
1030: \begin{proof}
1031: Since $\mathbf{x}$ and $\mathbf{y}$ tend to the same point at
1032: infinity, we may choose $N$ so that
1033: $(z,z')_\alpha> 2\thinspace \mathrm{diam}\{a,x_0,y_0\}$ for every $z$, $z'$ in
1034: $\{x_i \mid i\geq N\}\cup \{y_i \mid i\geq N\}$. Using Lemma
1035: \ref{l:approxbyterm} three times, the quantity
1036: \[|\eta_{\mathbf{x}}(a)-\eta_{\mathbf{y}}(a)-\eta_{\mathbf{x}}(y_0)|\]
1037: differs by at most $12\delta$ from
1038: \begin{equation}\label{easy}
1039: | (d(a,x_N) - d(a,y_N)) - (d(y_0,x_N)-d(y_0,y_N))|.
1040: \end{equation}
1041: By Observation \ref{o:fourdelta}, the quantity \eqref{easy} is at most
1042: $4\delta$. The Lemma follows.
1043: \end{proof}
1044:
1045: Using Lemma \ref{l:key}, we deduce that an isometry of $X$ changes
1046: $\eta_{\mathbf{x}}(a)$ by approximately the same amount, independent of
1047: the $a\in X$ chosen:
1048: \begin{proposition}\label{p:isom}
1049: Let $X$ be a geodesic $\delta$--hyperbolic space, and suppose that
1050: $\mathbf{x}=\{x_i\}$ tends to $e\in \partial X$. Let $a$ be any point
1051: in $X$.
1052: If $g\in \mathrm{Isom}(X)$ fixes $e$, then
1053: $\eta_{\mathbf{x}}(ga)$ differs from
1054: $\eta_{\mathbf{x}}(a)+\eta_{\mathbf{x}}(gx_0)$ by at most $16\delta$.
1055: \end{proposition}\label{p:trans}
1056: \begin{proof}
1057: First note that if $g\mathbf{x}$ is the sequence $\{gx_i\}$, then
1058: \[\eta_{\mathbf{x}}(a) = \eta_{g\mathbf{x}}(ga).\]
1059: But by Lemma \ref{l:key},
1060: \[ | \eta_{\mathbf{x}}(ga)-\eta_{g\mathbf{x}}(ga)-\eta_{\mathbf{x}}(g x_0)|
1061: \leq 16\delta. \]
1062: \end{proof}
1063:
1064: \begin{corollary}\label{c:quasimorphism}
1065: Suppose $X$ is a $\delta$--hyperbolic space, and that $G$ acts on $X$
1066: fixing $e\in \partial X$. Let $\mathbf{x}=\{x_i\}$ be any sequence tending to
1067: $e$. The function $q_{\mathbf{x}}\co G\to \R$ defined by
1068: $q_{\mathbf{x}}(g)=\eta_{\mathbf{x}}(g x_0)$ is a quasicharacter of
1069: defect at most $16\delta$.
1070: \end{corollary}
1071: \begin{proof}
1072: Let $g$, $h\in G$. Using Proposition \ref{p:isom},
1073: \begin{eqnarray*}
1074: |q_{\mathbf{x}}(gh)-q_{\mathbf{x}}(g)-q_{\mathbf{x}}(h)| & = &
1075: |\eta_{\mathbf{x}}(ghx_0)-\eta_{\mathbf{x}}(gx_0)-\eta_{\mathbf{x}}(hx_0)| \\
1076: & \leq & |
1077: \eta_{\mathbf{x}}(hx_0)+\eta_{\mathbf{x}}(gx_0)-\eta_{\mathbf{x}}(gx_0)-\eta_{\mathbf{x}}(hx_0)|+16\delta = 16\delta.
1078: \end{eqnarray*}
1079: \end{proof}
1080:
1081: \begin{proposition}\label{p:pseudo}
1082: Let $X$, $e$, $\mathbf{x}$, and $q_{\mathbf{x}}$ be as in the
1083: statement of Corollary \ref{c:quasimorphism}, and let the
1084: pseudocharacter $p_{\mathbf{x}}\co G\to \R$
1085: be given by
1086: \[p_{\mathbf{x}}(g) = \lim_{n\to\infty}\frac{q_{\mathbf{x}}(g^n)}{n}.\]
1087: Then $p_{\mathbf{x}}(g)\neq 0$ if and only if $g$ acts hyperbolically on $X$.
1088: \end{proposition}
1089:
1090: \begin{proof}
1091: First, we suppose that $p_{\mathbf{x}}(g)\neq 0$. Without loss of generality we
1092: assume that $p_{\mathbf{x}}(g)>0$. Since
1093: \[p_{\mathbf{x}}(g)=\lim_{n\to\infty}\frac{\eta_{\mathbf{x}}(g^n x_0)}{n}>0,\]
1094: there exists some $N$ so that
1095: $\eta_{\mathbf{x}}(g^n x_0)>\frac{1}{2}p_{\mathbf{x}}(g)n$ for all
1096: $n\geq N$.
1097:
1098: Let $a$ and $b$
1099: be integers. Choosing some sufficiently large $M$ and applying the triangle
1100: inequality and Lemma \ref{l:approxbyterm}, we obtain a lower bound for
1101: $d(g^a x_0,g^b x_0)$:
1102: \begin{eqnarray*}
1103: d(g^a x_0,g^b x_0) & = & d(g^N x_0,g^{N+|b-a|}x_0)\\
1104: & \geq & d(g^{N+|b-a|} x_0, x_M)-d(g^N x_0,x_M)\\
1105: & \geq & \eta_{\mathbf{x}}(g^{N+|b-a|})-\eta_{\mathbf{x}}(g^N) -
1106: 8\delta\\
1107: & \geq & \frac{1}{2}p_{\mathbf{x}}(g)(N+|b-a|)-\eta_{\mathbf{x}}(g^N)-8\delta\\
1108: & = & \frac{1}{2}p_{\mathbf{x}}(g) |b-a| - (\eta_{\mathbf{x}}g^N+8\delta-\frac{1}{2}p_{\mathbf{x}}(g)N).
1109: \end{eqnarray*}
1110: On the other hand, $d(g^a x_0,g^b x_0)\leq |b-a| d(x_0,gx_0)$, so the map
1111: $n\mapsto g^n(x_0)$ is a quasi-isometric embedding, and $g$ acts
1112: hyperbolically.
1113:
1114: Conversely, suppose that $g$ acts hyperbolically.
1115: It
1116: follows that there is some $\epsilon>0$ so that
1117: \[d(g^n x_0,x_0)>\epsilon n\] for all $n$.
1118: By
1119: replacing $g$ with $g^{-1}$, we may suppose that $\{g^i x_0\}$ tends
1120: to a point in $\partial X\smallsetminus\{e\}$ as $i\to\infty$.
1121: Thus there is some $R$ so that
1122: \[\gp{g^n x_0}{x_i}{{x_0}}<R\] for all positive $n$ and $i$.
1123: Lemma \ref{l:approxbyterm} implies that for sufficiently
1124: large $m$,
1125: \begin{eqnarray*}
1126: \eta_{\mathbf{x}}(g^n x_0) & \geq & d(g^n x_0,x_m)-d(x_0,x_m)-4\delta\\
1127: & = & d(g^n x_0,x_0)-2 \gp{g^nx_0}{x_n}{{x_0}} - 4\delta \\
1128: & \geq & \epsilon n - (2 R +4\delta).
1129: \end{eqnarray*}
1130: Since $q_{\mathbf{x}}(g^n)\geq \epsilon n - (2R+4\delta)$ for all $n>0$, we must
1131: have $p(g)\geq \epsilon>0$.
1132: \end{proof}
1133:
1134: \begin{remark}
1135: Proposition \ref{p:pseudo} was proved in \cite[Proposition 3.9]{manning:qfa} for the case of
1136: quasi-trees. The proof here is somewhat more efficient even in this
1137: case.
1138: \end{remark}
1139:
1140: \subsection{Characterization of $G$--spaces with invariant horoballs}
1141: \begin{theorem}\label{t:horeq}
1142: Let $X$ be a Gromov hyperbolic $G$--space. The following are
1143: equivalent:
1144: \begin{enumerate}
1145: \item\label{ih} $X$ has an invariant horoball.
1146: \item\label{enh} $X$ is elementary, and no element acts
1147: hyperbolically.
1148: \item\label{onepoint} $X$ is equivalent to a hyperbolic $G$--space $Y$, so that
1149: $\#(\partial Y)\leq 1$.
1150: \end{enumerate}
1151: \end{theorem}
1152: \begin{proof}
1153: That \eqref{ih} implies \eqref{onepoint} is trivial.
1154:
1155: We next assume \eqref{onepoint} and show \eqref{enh}. If \eqref{enh}
1156: holds for $Y$, it holds for $X$, so we may suppose that $X=Y$. If
1157: $\#(\partial X) = 1$ or $G$ has a bounded orbit in $X$, then clearly
1158: $X$ is elementary. The only case remaining is that $\partial X$ is
1159: empty, but $Gx$ is unbounded for some $x\in X$. We show that this
1160: case does not occur. Chose a sequence
1161: $\{g_i\}$ in $G$ so that $\lim_{i\to\infty}d(g_i x,x) = \infty$.
1162: Since $\partial X$ is empty,
1163: $\liminf_{i,j\to\infty}(g_ix,g_jx)_x<\infty$. It follows that there
1164: are elements $g_m$, $g_n$ so that $d(g_mx,x)$ and $d(g_nx,x)$ are much
1165: larger than $(g_mx,h_nx)_x$. It can then be shown (see, for
1166: example, \cite[Chapitre 9, Lemme 2.3]{cdp}) that $g_mg_n$ is hyperbolic.
1167: It follows that $\partial X$ contains at least two points (the fixed
1168: points of $g_mg_n$), contrary to assumption.
1169:
1170: It remains to show that \eqref{enh} implies \eqref{ih}.
1171: Let $X$ be a hyperbolic $G$--space so that the action of $G$ is
1172: elementary. If $X$ is equivalent to a point (i.e. if $Gx$ is bounded
1173: for $x\in X$), then $X$ is also
1174: equivalent to a ray, which is the combinatorial horoball based on a
1175: point. We therefore may assume that $Gx$ is unbounded for any $x\in
1176: X$. We will construct a combinatorial horoball which coarsely
1177: equivariantly quasi-isometrically embeds in $X$.
1178: Let $\delta>0$ be some number so
1179: that $X$ is $\delta$--hyperbolic.
1180: Let $e$, $\mathbf{x}$,
1181: $q_{\mathbf{x}}$ and $p_{\mathbf{x}}$ be as in the statements of
1182: Corollary \ref{c:quasimorphism} and Proposition \ref{p:pseudo}.
1183:
1184: To build the combinatorial horoball, we first must start with a graph $Y$
1185: on which $G$ acts. Choose a finite generating set $S$ for $G$, and
1186: let $C_0 = \mathrm{diam}(S x_0)$.
1187: Let $V(Y)=G$, and connect $g$ to $h$ in $Y$ if $d(g x_0, h x_0)\leq C_0$. It
1188: is clear that $G$ acts on $Y$; in fact, $Y$ is a certain Cayley graph
1189: for $G$. We next define the functions $B_n\co V(Y)\to 2^{V(Y)}$.
1190: Let $C_1 = 2C_0+20\delta$,
1191: and let
1192: \begin{equation}\label{bstar}
1193: B_n(g) = \{ h\in G\mid d(hx_0,gx_0)\leq (2n+1)C_1\}.
1194: \end{equation}
1195: \begin{claim}
1196: The sequence of functions $B_*$ in \eqref{bstar} is admissible in the sense
1197: of Definition \ref{d:admissible}.
1198: \end{claim}
1199: \begin{proof}
1200: The only axiom which is not obvious is \eqref{ax:exp}. We must show
1201: that if $a$ and $b$ are in $B_n(v)$, then $a\in B_{n+1}(b)$ (or
1202: equivalently $b\in B_{n+1}(a)$).
1203: Put another way,
1204: we must show that if $d(ax_0,vx_0)$ and $d(bx_0,vx_0)$
1205: are bounded above by $(2n+1)C_1$ then $d(a x_0,b x_0)\leq (2n+3)C_1$.
1206:
1207: Because no element of $G$ acts hyperbolically, the pseudocharacter
1208: $p_{\mathbf{x}}$ is identically zero. An easy argument shows that
1209: $|q_{\mathbf{x}}(g)|\leq 16\delta$ for all $g\in G$.
1210: Using Lemma \ref{l:approxbyterm}, we can choose some large $n$ so that
1211: \[| \eta_{\mathbf{x}}(z)-(d(z,x_n)-d(x_0,x_n))| \leq 4\delta\]
1212: for $z\in \{ax_0,bx_0,vx_0\}$. It follows that
1213: \begin{equation}\label{diff}
1214: |d(z_1,x_m)-d(z_2,x_m)|\leq 16\delta+4\delta\leq C_1
1215: \end{equation} for $z_1$,
1216: $z_2\in\{ax_0,bx_0,vx_0\}$. The assertion to be proved is symmetric
1217: in $a$ and $b$, so we may assume
1218: that $\gp{x_m}{b}{v}\leq \gp{x_m}{a}{v}$.
1219: We deduce:
1220: \begin{eqnarray*}
1221: d(a,b) & \leq & \gp{v}{x_m}{a}+[\gp{a}{x_m}{v}-\gp{b}{x_m}{v}]+\gp{v}{x_m}{b}+2\delta\\
1222: & = & d(v,a) +(d(b,x_m)-d(v,x_m))+2\delta \\
1223: & \leq & (2n+1)C_1 + C_1 + 2\delta \leq (2n+3)C_1.
1224: \end{eqnarray*}
1225: The first line follows from examining the comparison tripods
1226: for the triangles $\Delta(a,v,x_m)$ and $\Delta(b,v,x_m)$;
1227: the last follows from \eqref{diff}.
1228: \end{proof}
1229:
1230: Since $B_*$ is admissible, the combinatorial horoball
1231: $H=\mathcal{H}_{B_*}(Y)$ is a hyperbolic $G$--space. It remains
1232: to construct a coarsely equivariant quasi-isometric embedding from $H$
1233: to $X$. It
1234: suffices to define this map on the vertices of $H$.
1235: For each $g\in G = V(Y)$ and each $n\in\N$, choose some
1236: $i(g,n)$ so that $\gp{x_k}{x_l}{{gx_0}}\geq 2nC_1$ for all $k$, $l\geq
1237: i(g,n)$. Choose also some unit speed geodesic $\gamma_{g,n}$ starting
1238: at $g x_0$ and ending at $x_{i(g,n)}$. Any vertex of $H$ is a pair
1239: $(g,n)$ where $n\in \N$ and $g\in G$. We define a map $\phi\co
1240: V(H)\to X$ by
1241: \begin{equation*}
1242: \phi(g,n) = \gamma_{g,n}(nC_1).
1243: \end{equation*}
1244: A number of choices were made in the definition of $\phi$ (namely, the
1245: sequence $\mathbf{x}$, the numbers $i(g,n)$, and the geodesics $\gamma_{g,n}$).
1246: However, so long as $x_0$ is unchanged, different
1247: choices lead to a function which differs by at most $\delta$
1248: from $\phi$. In particular, we could replace $\mathbf{x}$ by
1249: $\mathbf{x}'=\{x_i'\}$, where $x_0'=x_0$ and $x_i'=hx_i$ for some
1250: fixed $h$ and for all $i\geq 1$.
1251: It follows that the distance between $\phi(hg,n)$ and $h\phi(g,n)$ is
1252: at most $\delta$ for any $h$, $g\in G$ and $n\in \N$, and so the map
1253: $\phi$ is coarsely equivariant.
1254:
1255:
1256: It remains to show that $\phi$ is a quasi-isometric
1257: embedding.
1258:
1259: Note that if $v$ and $w$ are two vertices in $H$ connected by a
1260: vertical path, then
1261: $C_1d_H(v,w)-\delta\leq d(\phi(v),\phi(w))\leq C_1d_H(v,w)+\delta$,
1262: where $d_H$ is the distance in $H$.
1263:
1264: We therefore assume that $v=(a,n)$ and $w=(b,m)$, where $a\neq b$.
1265: There is a unique $k$ so that $(2k-1)C_1<d(a,b)\leq (2k+1)C_1$. If
1266: $\max\{m,n\}\geq k$, then $d(v,w)=|m-n|+1$; otherwise
1267: $d(v,w)=2k-(m+n)+\frac{1}{2}\pm\frac{1}{2}$. Let
1268: $I \geq \max\{i(a,n),i(b,m)\}$, and observe that the points $\phi(v)$
1269: and $\phi(w)$ lie within $\delta$ of geodesics joining $ax_0$ to $x_I$
1270: and $bx_0$ to $x_I$, respectively (see Figure \ref{f:deep}).
1271: \begin{figure}[htbp]
1272: \begin{center}
1273: \input{deep.pstex_t}
1274: \caption{A pair of points in the image of $\phi$.}
1275: \label{f:deep}
1276: \end{center}
1277: \end{figure}
1278: Note that $\gp{x_I}{ax_0}{{bx_0}}$
1279: and $\gp{x_I}{bx_0}{{ax_0}}$ differ from $\frac{1}{2}d(ax_0,bx_0)$ by at
1280: most $\frac{C_1}{2}$.
1281: There are a couple of cases to consider.
1282:
1283: First, assume that one or both of $n$ and $m$ is at least $k$.
1284: Without loss of generality assume that $n\geq k$. Since
1285: $d(ax_0,bx_0)\leq (2k+1)C_1$, we have $(x_I,bx_0)_a\leq kC_1+C_1$.
1286: Since $d(\phi(v),ax_0)\geq k C_1$,
1287: it follows that $\phi(v)$ is at most $C_1+\delta$ from the
1288: geodesic joining $bx_0$ to $x_I$. Accounting for the possible
1289: difference between $\gp{x_I}{ax_0}{{bx_0}}$ and $\gp{x_I}{bx_0}{{ax_0}}$, we deduce that the distance between $\phi(v)$ and
1290: $\phi(w)$ differs from $|n-m|C_1=(d(v,w)+\frac{1}{2}\pm\frac{1}{2})C_1$ by at most $2C_1+\delta$. It
1291: follows that in case one of $n$ or $m$ is at least $k$, we have
1292: \[ C_1 d(v,w)-2C_1+\delta \leq d(\phi(v),\phi(w))\leq
1293: C_1d(v,w)+3C_1+\delta.\]
1294:
1295: In case both $n$ and $m$ are strictly less than $k$, we may argue as
1296: follows. Since $\gp{ax_0}{x_I}{{bx_0}}$ and $\gp{bx_0}{x_I}{{ax_0}}$ are both
1297: at least $\frac{1}{2}(d(ax_0,bx_0)-C_1)\geq (k-1)C_1\geq \max\{m,n\}
1298: C_1$, it follows that both $\phi(v)$ and $\phi(w)$ are within $2\delta$
1299: of the geodesic joining $ax_0$ to $bx_0$. From this it follows that
1300: $d(\phi(v),\phi(w))$ differs from $d(ax_0,bx_0)-(n+m)C_1$ by at most
1301: $4\delta$. But since $d(ax_0,bx_0)$ differs by at most $C_1$ from $k
1302: C_1$, we deduce that
1303: \[(2k-(m+n))C_1 -(C_1+4\delta) \leq d(\phi(v),\phi(w))\leq
1304: (2k-(m+n))C_1 + (C_1+4\delta),\]
1305: from which it immediately follows that
1306: \[C_1 d(v,w) - (2C_1+4\delta)\leq d(\phi(v),\phi(w))\leq C_1 d(v,w)
1307: +(2C_1+4\delta).\]
1308:
1309: In particular, $\phi$ is a $(C_1,2C_1+4\delta)$--quasi-isometric
1310: embedding from the combinatorial horoball $H$ into $X$, and the
1311: theorem is established.
1312: \end{proof}
1313:
1314:
1315:
1316: \section{Rigidity in rank $\geq 2$}\label{s:bigrank}
1317: The purpose of
1318: this section is to establish Theorem \ref{thm:cdr2}, but we will begin
1319: with some lemmas which hold in a slightly broader context.
1320: We suppose that $G$ is a simple Chevalley-Demazure
1321: group scheme of rank at least $2$, that $\Phi$ is a root system for
1322: $G$, and that $R$ is some commutative unital ring containing $\Z$.
1323:
1324: \begin{lemma}\label{l:phiprime}
1325: If $\alpha$ and $\beta\in \Phi$,
1326: then there is a $\Phi' = \mathrm{Span}(\Phi')\cap\Phi\subseteq \Phi$
1327: containing $\alpha$ and $\beta$ so that
1328: $\Phi'$ is
1329: isomorphic to $A_1\times A_1$, $A_2$, $B_2$ or $G_2$. If $\Phi'\cong
1330: A_1\times A_1$, then $\alpha\neq -\beta$.
1331: \end{lemma}
1332: \begin{proof}
1333: If $\alpha$ and $\beta$ are linearly independent, then
1334: $\Phi'=\mathrm{Span}(\{\alpha,\beta\})\cap\Phi$ is a root system of
1335: rank two, and we simply recall that such a root system is always
1336: isomorphic to one of those listed.
1337:
1338: If $\beta=-\alpha$, we assert that there must be some $\gamma\in \Phi$
1339: so that
1340: $\mathrm{Span}(\{\alpha,\gamma\})\cap\Phi$ is not equal to $A_1\times
1341: A_1$.
1342: Suppose that there is no such $\gamma$. Then either
1343: $\Phi=\overline{\Phi}\times\langle\alpha\rangle$, or
1344: $\Phi=\langle\alpha\rangle$.
1345: Because $G$ is simple $\Phi$ cannot split as a product; because $G$
1346: has rank at least two, $\Phi\neq A_1$.
1347: \end{proof}
1348: For the following two lemmas, we refer to \cite{carter:book}. Although
1349: the proofs there are done assuming that $R$ is a field, this
1350: assumption is unnecessary; see also \cite{stein71} or
1351: \cite{steinberg:notes}.
1352: \begin{lemma}\label{l:steinberg}{\em (Steinberg commutator
1353: relations)}\cite[Theorem 5.2.2]{carter:book}
1354: If $\alpha$, $\beta\in \Phi$ and $t$, $u\in R$, then
1355: \[[x_\alpha(t),x_\beta(u)]=\prod_{i,j>0\mbox{, }i\alpha+j\beta\in \Phi }
1356: x_{i\alpha+j\beta}(N_{\alpha,\beta,i,j}t_iu_j), \]
1357: where the $N_{\alpha,\beta,i,j}\in \Z$ are integers which depend only
1358: on the order in which the product is taken.
1359: \end{lemma}
1360: \begin{lemma}\label{l:weylconj}\cite[Lemma 7.2.1]{carter:book}
1361: If $\alpha\in \Phi$, $w$ is an element of the Weyl group of $\Phi$,
1362: and $t\in R$,
1363: then
1364: $x_{w(\alpha)}(t)$ is conjugate either to $x_\alpha(t)$
1365: or $x_\alpha(-t)$.
1366: \end{lemma}
1367:
1368: \begin{lemma}\label{l:distortion}
1369: Let $X$ be a hyperbolic $G(R)$--space.
1370: If $g = x_\alpha(t)$ for some $\alpha\in \Phi$ and $t\in R$, then $g$
1371: does not act hyperbolically on $X$.
1372: \end{lemma}
1373: \begin{proof}
1374: By Lemma \ref{l:phiprime}, there is a subset $\Phi'$ of $\Phi$
1375: containing $\alpha$ which is either isomorphic to $A_2$, $B_2$ or
1376: $G_2$. In each case, we may apply Lemma \ref{l:steinberg} some number
1377: of times to show that $g$ is distorted in $G$; the details of this are
1378: left to the reader. By Lemma
1379: \ref{lemma:distortion}, $g$ cannot act hyperbolically on $X$.
1380: \end{proof}
1381:
1382: \begin{proposition}\label{p:allfixone}
1383: Let $X$ be a hyperbolic $G(R)$--space,
1384: let $r$, $s\in \Phi$, and let $\rho_1$, $\rho_2\in R$.
1385: Suppose that $p=x_r(\rho_1)$ acts parabolically, fixing some $e\in
1386: \partial X$. Then $ge=e$ for any other root element $g=x_s(\rho_2)$.
1387: \end{proposition}
1388: \begin{proof}
1389: If $r=s$ or if
1390: $\langle r,s\rangle = A_1\times A_1$, then
1391: $p$ and $g$ commute. By Lemma \ref{l:commute}, $g(e)=e$, and we are
1392: done.
1393:
1394: Otherwise $r$ and $s$ are contained in a two-dimensional root system
1395: $\Phi'\subset \Phi$ which is isomorphic to
1396: $A_2$, $B_2$ or $G_2$, by Lemma \ref{l:phiprime}.
1397:
1398: Each case requires a separate argument.
1399:
1400: \begin{case}\label{c:a2}
1401: $\Phi'\cong A_2$.
1402: \end{case}
1403: $A_2 = \{\lambda_i \mid i\in \Z_6\}$ contains six roots, arranged
1404: hexagonally; the angle between $\lambda_i$ and
1405: $\lambda_j$ is $\frac{|i-j|}{3}\pi$. Suppose $r=\lambda_i$ and $s=\lambda_j$.
1406: By Lemma \ref{l:weylconj}, $h = x_s(\rho_1)$ is conjugate to either
1407: $p$ or $p^{-1}$, so $h$ is a parabolic, fixing some point
1408: $f\in\partial X$.
1409: In case $|i-j|=1$, then Lemma \ref{l:steinberg} implies that
1410: $h$ and $p$ commute, and so $f=e$ by Lemma \ref{l:commute}.
1411: In case $|i-j|>1$, one argues by induction on $|i-j|$ to the same
1412: conclusion: $f=e$.
1413:
1414: Since $g$ commutes with $h$, we must have $ge=e$, again by Lemma
1415: \ref{l:commute}.
1416:
1417: \begin{case}\label{c:b2}
1418: $\Phi'\cong B_2$.
1419: \end{case}
1420:
1421: Let $\alpha$ be a short root, and $\beta$ a long root, so that
1422: $\alpha$ and $\beta$ span $\Phi'\subset \Phi$, as in
1423: Figure \ref{f:b_2}.
1424: \begin{figure}[htbp]
1425: \begin{center}
1426: \begin{picture}(0,0)%
1427: \includegraphics{b_2.pstex}%
1428: \end{picture}%
1429: \setlength{\unitlength}{4144sp}%
1430: %
1431: \begingroup\makeatletter\ifx\SetFigFont\undefined%
1432: \gdef\SetFigFont#1#2#3#4#5{%
1433: \reset@font\fontsize{#1}{#2pt}%
1434: \fontfamily{#3}\fontseries{#4}\fontshape{#5}%
1435: \selectfont}%
1436: \fi\endgroup%
1437: \begin{picture}(4352,2006)(867,-3021)
1438: \put(4972,-2047){\makebox(0,0)[lb]{\smash{{\SetFigFont{6}{7.2}{\familydefault}{\mddefault}{\updefault}{\color[rgb]{0,0,0}$\alpha$}%
1439: }}}}
1440: \put(3455,-1448){\makebox(0,0)[lb]{\smash{{\SetFigFont{6}{7.2}{\familydefault}{\mddefault}{\updefault}{\color[rgb]{0,0,0}$\beta$}%
1441: }}}}
1442: \put(2546,-2026){\makebox(0,0)[lb]{\smash{{\SetFigFont{6}{7.2}{\familydefault}{\mddefault}{\updefault}{\color[rgb]{0,0,0}$\alpha$}%
1443: }}}}
1444: \put(882,-1238){\makebox(0,0)[lb]{\smash{{\SetFigFont{6}{7.2}{\familydefault}{\mddefault}{\updefault}{\color[rgb]{0,0,0}$\beta$}%
1445: }}}}
1446: \end{picture}%
1447: \caption{$B_2$ and $G_2$.}
1448: \label{f:b_2}
1449: \end{center}
1450: \end{figure}
1451:
1452: There are two subcases, depending on whether $r$ is a short or long
1453: root.
1454: \begin{subcase}
1455: The parabolic $p=x_r(\rho_1)$, where $r$ is a short root of $\Phi'$.
1456: \end{subcase}
1457: Without loss of generality,
1458: we may assume that $r=\alpha$. If $s\in
1459: \{\alpha,2\alpha+\beta,-\beta\}$, then $g=x_s(\rho_2)$ commutes with
1460: $p$, by Lemma \ref{l:steinberg}.
1461: Thus by Lemma \ref{l:commute} $ge=e$.
1462:
1463: Suppose next that $s = \pm (\alpha+\beta)$. Then Lemma
1464: \ref{l:steinberg} implies
1465: \begin{equation}\label{c1}
1466: p g p^{-1} g^{-1} = x_{r+s}(N \rho_1\rho_2),
1467: \end{equation}
1468: where $N=N_{r,s,1,1}$ is an integer. Since $r+s\in
1469: \{2\alpha+\beta,-\beta\}$, we already know that
1470: $h:=x_{r+s}(N\rho_1\rho_2)$ fixes $e$. We rearrange \eqref{c1} to give
1471: \begin{equation}
1472: h^{-1} p = gpg^{-1}.
1473: \end{equation}
1474: Since $p$ and $h$ both fix $e$, so must $gpg^{-1}$. The element
1475: $gpg^{-1}$ also fixes $g(e)$, since $p$ fixes $e$.
1476: Since $p$ is
1477: parabolic, it can only fix one point in $\partial X$, and so $g(e)=e$.
1478:
1479: If $s\in \{\beta, -\alpha\}$, let $r' = \alpha+\beta$; if
1480: $s=-2\alpha-\beta$ let $r' = -\alpha-\beta$. In any case there is an
1481: element of the Weyl group of $\Phi$ taking $r$ to $r'$; by Lemma
1482: \ref{l:weylconj}, there is a
1483: $p'= x_{r'}(\pm \rho_1)$ which is
1484: conjugate to $p$ in $E(\Phi,R)$. Since $p'$ is conjugate to $p$, it
1485: is parabolic; by the argument of the previous paragraph, $p'$ has the
1486: same fixed point as $p$. If $s\in \{\beta, -2\alpha-\beta\}$, then
1487: $g$ commutes with $p'$, and so $g(e)=e$ by Lemma \ref{l:commute}.
1488: Finally, if
1489: $s=-\alpha$, then we may apply the argument of the previous paragraph
1490: again (with $p'$ and $r'$ in place of $p$ and $r$), to deduce that $g(e)=e$.
1491:
1492: \begin{subcase}
1493: The parabolic $p=x_r(\rho_1)$, where $r$ is a long root of $\Phi'$.
1494: \end{subcase}
1495: In this case, we may assume for instance that $r=2\alpha+\beta$. If
1496: $s\in\{2\alpha+\beta, \alpha+\beta, \alpha, \beta, -\beta\}$, then
1497: Lemma \ref{l:steinberg} implies that $g=x_s(\rho_2)$ commutes with $p$,
1498: and so $g(e)=e$ by Lemma \ref{l:commute}.
1499:
1500: Suppose then that $s\in \{-\alpha,-2\alpha-\beta,-\alpha-\beta\}$.
1501: If $s\in\{-\alpha,-2\alpha-\beta\}$, let $r'=\beta$; if $s=
1502: -\alpha-\beta$, then let $r'=-\beta$. In either case, there is an
1503: element of the Weyl group taking $r$ to $r'$, and so there is a
1504: parabolic $p' = x_{r'}(\pm \rho_1)$ conjugate to $p$ by Lemma
1505: \ref{l:weylconj}. By the previous paragraph, $p'$ has the same fixed
1506: point as $p$ does. Applying the previous paragraph with $p'$ and $r'$
1507: in the place of $p$ and $r$ implies that $g(e)=e$ for $g =
1508: x_s(\rho_2)$. This completes the proof of Case \ref{c:b2}.
1509:
1510: \begin{case}\label{c:g2}
1511: $\Phi'\cong G_2$.
1512: \end{case}
1513: Let $\alpha$ be a short root, and $\beta$ a long root, so that
1514: $\alpha$ and $\beta$ span $\Phi'\subset \Phi$, as in
1515: the right half of Figure \ref{f:b_2}.
1516:
1517:
1518: Again there are two subcases, depending on whether $r$ is a short or long
1519: root.
1520:
1521: \begin{subcase}
1522: The parabolic $p=x_r(\rho_1)$, where $r$ is a long root of $\Phi'$.
1523: \end{subcase}
1524: If $s\in \{-\beta, 3\alpha+\beta, \pm(3\alpha+2\beta)\}$, then Lemma
1525: \ref{l:steinberg} implies that $g=x_s(\rho_2)$ commutes with $p$, and
1526: so $g(e)=e$ by Lemma \ref{l:commute}.
1527:
1528: Suppose that $s\in \{2\alpha+\beta,-\alpha-\beta\}$. Lemma
1529: \ref{l:steinberg} implies that
1530: \begin{equation}
1531: p g p^{-1} g^{-1} = x_{r+s}(N\rho_1\rho_2)=: h,
1532: \end{equation}
1533: for some integer $N$. Exactly as in Case \ref{c:b2}, $h$ commutes
1534: with $p$, and so $h(e)=e$. Thus $e= h^{-1} p (e) = gpg^{-1}(e)$ and
1535: the parabolic $gpg^{-1}$ fixes $e$. Again since $gpg^{-1}$ also fixes
1536: $g(e)$, we must have $g(e)=e$.
1537:
1538: Using Lemma \ref{l:weylconj} repeatedly we discover that for every
1539: short root $r'$ there is a parabolic element $p'=x_{r'}(\pm \rho_1)$
1540: with $p'(e)=e$. Since $g$ must commute with some such element,
1541: $g(e)=e$ as well.
1542: \begin{subcase}
1543: The parabolic $p=x_r(\rho_1)$, where $r$ is a long root of $\Phi'$.
1544: \end{subcase}
1545: Without loss of generality, we may assume that $r = 3\alpha+2\beta$.
1546:
1547: If the inner product of $s$ with $r$
1548: is nonnegative, then $s\in\{ \pm \alpha, \beta, \alpha+\beta,
1549: \alpha+2\beta,\alpha+3\beta, 3\alpha+2\beta\}$, and $g$ commutes with
1550: $p$ by Lemma \ref{l:steinberg}, and so $g(e)=e$.
1551:
1552: Otherwise, a (possibly
1553: repeated) application of Lemma \ref{l:weylconj} implies that $g$
1554: commutes with a parabolic $p' = x_{r'}(\pm \rho_1)$ for some long root
1555: $r$ of $\Phi'$, and with $p'(e)=e$. This again implies via Lemma
1556: \ref{l:commute} that $g(e)=e$. This completes the proof in Case \ref{c:g2}.
1557: \end{proof}
1558:
1559: We are now ready to give the proof of Theorem \ref{thm:cdr2}.
1560: \begin{proof}
1561: A result of Tavgen$'$ \cite{tavgen:bg} shows that $G(\mathcal{O})$ is
1562: boundedly generated by its root subgroups.
1563: The ring of integers $\mathcal{O}$ is finitely generated as an Abelian
1564: group; choose generators $\mu_1,\ldots,\mu_k$. It follows from
1565: Tavgen$'$'s result that the set
1566: \[ S=\{x_\alpha(\mu_i) \mid \alpha \in \Phi\mbox{, }1\leq i\leq k\}\]
1567: boundedly generates $G(\mathcal{O})$.
1568:
1569: Each of these generators acts hyperbolically, elliptically or
1570: parabolically on $X$.
1571: By Lemma \ref{l:distortion}, none can act hyperbolically.
1572: If all of the root elements act elliptically, then it follows from bounded
1573: generation that the orbit of a point under the action of $G$ must be
1574: bounded.
1575:
1576: We therefore may assume that some $x_\alpha(\mu_i)$ acts parabolically on
1577: $X$, fixing a single point $e\in \partial X$. It follows from
1578: Proposition \ref{p:allfixone} that all the root subgroups will fix
1579: this point $e$, and so $G(\mathcal{O})$ fixes $e$.
1580:
1581: By Proposition \ref{p:pseudo}, the pseudocharacter
1582: $p_\mathbf{x}\co G(\mathcal{O})\to \R$ determined by a sequence
1583: $\mathbf{x}=\{x_i\}$ tending to $e$
1584: is nonzero exactly on the hyperbolic elements. Thus $p_\mathbf{x}(g)=0$
1585: whenever $g$ lies in a root subgroup.
1586:
1587: An elementary argument (see for example \cite[Proposition
1588: 5]{kotschick:quasi}) shows that a pseudocharacter $p$ on a boundedly
1589: generated group
1590: is determined by its values on
1591: the bounded generators; thus $p\equiv 0$ on $G(\mathcal{O})$. It
1592: follows that no element of $G(\mathcal{O})$ acts hyperbolically on
1593: $X$.
1594: By Theorem
1595: \ref{t:horeq}, the $G(\mathcal{O})$--space $X$ has an invariant horoball.
1596: \end{proof}
1597:
1598:
1599: \section{Remarks on rank one}\label{section:rankone}
1600: One can also ask what actions rank one Chevalley groups have on
1601: hyperbolic spaces. If $\mc{O}$ is a number ring with finitely many
1602: units, then $SL(2,\mc{O})$ is a lattice either in $SL(2,\R)$ or
1603: $SL(2,\C)$. In particular, it has a proper non-elementary action on
1604: $\H^2$ or $\H^3$. Moreover, such a group admits uncountably many
1605: distinct pseudocharacters (AKA homogeneous quasi-(homo)morphisms) up
1606: to scale \cite{fujiwara:gromovhyperbolic,bestvinafujiwara:mcg}. Each
1607: such ``projective pseudocharacter'' gives rise to a
1608: quasi-action on $\R$; no two such are equivalent. Moreover, these
1609: often give rise to quasi-actions on more complicated trees
1610: \cite{manning:cocycles}.
1611: The groups $SL(2,\mc{O})$ where $\mc{O}$ has infinitely many units are
1612: more rigid.
1613: In this section we
1614: apply the main result of \cite{manning:qfa} to the special case of
1615: actions on quasi-trees (defined below), and
1616: speculate on the general situation.
1617:
1618: Recall that a group $G$ is said to have
1619: Property \FA\ if every action by $G$ on a simplicial tree $T$ has a
1620: fixed point.
1621: \begin{definition}
1622: A \emph{quasi-tree} is a graph which is quasi-isometric to a tree.
1623: \end{definition}
1624: \begin{definition}
1625: A group $G$ has property \QFA\ if every action by $G$ on a quasi-tree
1626: $X$ has a bounded orbit.
1627: \end{definition}
1628: \begin{remark}
1629: This is differently worded than the definition in \cite{manning:qfa},
1630: but easily seen to be equivalent. Note that \QFA\ implies \FA, but
1631: not vice versa.
1632: \end{remark}
1633: As quasi-trees are in particular Gromov hyperbolic spaces which admit
1634: no parabolic isometries (see Section 3.2 of \cite{manning:qfa}),
1635: Theorem \ref{thm:cdr2} implies that higher rank Chevalley
1636: groups over number rings have property \QFA.
1637:
1638: We recall a definition and a theorem from \cite[Section 4]{manning:qfa}.
1639:
1640: \begin{definition}
1641: Let $G$ be a group, and let $g$ be an element of $G$. We will say $g$
1642: is a \emph{stubborn element of $G$} if for all
1643: $H< G$ with $[G:H]\leq 2$, there exists some integer $k_H>0$ so that
1644: $g^{k_H}\in [H,H]$.
1645: \end{definition}
1646:
1647:
1648: \begin{theorem}\label{thm:qfa}\cite[Theorem 4.4]{manning:qfa}
1649: Let $G$ be a group which is boundedly generated by elements
1650: $g_1,\ldots,g_n$, so
1651: that for each $i$, $g_i$ is a stubborn element of $B_i$ for some
1652: amenable $B_i < G$. Then $G$ has Property \QFA.
1653: \end{theorem}
1654: Note that the above theorem was misstated slightly in
1655: \cite{manning:qfa}; the
1656: word ``amenable'' was inadvertently omitted.
1657:
1658: Here's an easy lemma:
1659: \begin{lemma}\label{lemma:finiteindex}
1660: Let $H<G$ be a subgroup of finite index. If $H$ has Property
1661: \QFA, then so does $G$.
1662: \end{lemma}
1663:
1664: Our methods in the higher rank case use heavily the bounded generation of
1665: $G(\mathcal{O})$ established by Tavgen$'$ in \cite{tavgen:bg}
1666: for Chevalley groups over rings of integers of algebraic
1667: number fields. There is an analogous result of Carter, Keller, and Paige
1668: in rank $1$, at least for $SL(2,\cdot)$ and certain number
1669: fields:
1670: \begin{theorem}\label{thm:ckp2}
1671: \cite{wittemorris:ckp}
1672: For any integer $d>1$ there is an $r=r(d)$ so that the following is true.
1673: Let $K$ be a number field of degree $d$ over $\Q$, and let
1674: $\mathcal{O}$ be the ring of integers of $K$. If the $\mc{O}$ has
1675: infinitely many units, then:
1676: \begin{enumerate}
1677: \item every element of $E(2,\mathcal{O})$ is a product of at most $r$
1678: elementary matrices, and
1679: \item the index of $E(2,\mathcal{O})$ in $SL(2,\mathcal{O})$ is at
1680: most $r$.
1681: \end{enumerate}
1682: \end{theorem}
1683: In the above statement, $E(2,\mathcal{O})$ is the subgroup of
1684: $SL(2,\mathcal{O})$ generated by the root subgroups (the strictly
1685: upper triangular and strictly lower triangular matrices).
1686: The following statement implies Theorem \ref{thm:cdr1}:
1687: \begin{theorem}\label{thm:rankone}
1688: If $\mathcal{O}$ is the ring of integers of an algebraic number field
1689: and $\mathcal{O}$ has infinitely many units, then $SL(2,\mathcal{O})$
1690: has Property \QFA.
1691: \end{theorem}
1692: \begin{proof}
1693: By Lemma \ref{lemma:finiteindex} it suffices to show that
1694: $E(2,\mathcal{O})$ has property \QFA.
1695: If $\Lambda= \{\lambda_1,\ldots,\lambda_n\}$ is an integral basis
1696: for the number field $\mathcal{O}$, then
1697: Theorem \ref{thm:ckp2} implies that $E(2,\mathcal{O})$ is boundedly
1698: generated by the $2n$ elements $\textMtwo{1}{\lambda_i}{0}{1}$ and
1699: $\textMtwo{1}{0}{\lambda_i}{1}$.
1700: \begin{claim}
1701: For each $i\in\{1,\ldots,n\}$,
1702: $\textMtwo{1}{\lambda_i}{0}{1}$ is a stubborn element of
1703: $B=\textMtwo{*}{*}{0}{*}\cap E(2,\mathcal{O})$.
1704: \end{claim}
1705: \begin{proof}
1706: On page 189 of \cite{carter:book}, Carter observes that
1707: $\textMtwo{t}{0}{0}{t^{-1}}$ can be written as a product of elementary
1708: matrices, for any invertible $t\in \mc{O}$, as follows.
1709: If
1710: $\lambda$ is any invertible element, we may write
1711: \[\Mtwo{0}{\lambda}{-\lambda^{-1}}{0} =
1712: \Mtwo{1}{\lambda}{0}{1}\Mtwo{1}{0}{-\lambda^{-1}}{1}\Mtwo{1}{\lambda}{0}{1},\]
1713: and then note that
1714: \begin{equation}\label{computation}
1715: \Mtwo{t}{0}{0}{t^{-1}}=\Mtwo{0}{t}{-t^{-1}}{0}\Mtwo{0}{-1}{1}{0}.
1716: \end{equation}
1717: Carter's observation shows that $\textMtwo{\omega}{0}{0}{\omega^{-1}}$ is in
1718: $E(2,\mc{O})$ for any unit $\omega$ in $\mc{O}$.
1719: Computing the commutator of \textMtwo{\omega}{0}{0}{\omega^{-1}} and
1720: \textMtwo{1}{\lambda}{0}{1} for $\omega$ a unit of $R$ and $\lambda\in R$
1721: yields:
1722: \begin{equation}\label{computation2}
1723: \left[\Mtwo{\omega}{0}{0}{\omega^{-1}},\Mtwo{1}{\lambda}{0}{1}\right]
1724: = \Mtwo{1}{(1-\omega^2)\lambda}{0}{1}.
1725: \end{equation}
1726:
1727:
1728: By assumption, the group of units of $\mc{O}$ is infinite.
1729: Dirichlet's units theorem (see, e.g. \cite[Appendix B]{stewarttall})
1730: implies that we may
1731: choose $\omega_0\in \mc{O}^*$ a unit of infinite order.
1732: Let $H<B$ be a subgroup of index at most two. Then $H$ must contain
1733: $\textMtwo{\omega_0^2}{0}{0}{\omega_0^{-2}}$ and
1734: $\textMtwo{1}{2 r}{0}{1}$ for all $r\in \mc{O}$. It follows from the
1735: computations \eqref{computation} and \eqref{computation2}
1736: that $\textMtwo{1}{i}{0}{1}\in [H,H]$ for all
1737: $i$ in the ideal $I$ generated by $2(1-\omega_0^2)$. Let $N$ be the
1738: order of $R/I$. (The number $N$ is also called the \emph{norm} of $I$;
1739: that it is finite when $I\neq (0)$ is an elementary fact of algebraic
1740: number theory; see, e.g., \cite[Chapter 5]{stewarttall}.)
1741: For any of the $\lambda_i$, we have
1742: $\textMtwo{1}{\lambda_i}{0}{1}^N=\textMtwo{1}{N\lambda_i}{0}{1}\in [H,H]$, and
1743: so $\textMtwo{1}{\lambda_i}{0}{1}$ is stubborn.
1744: \end{proof}
1745: It remains to observe that $B<E(2,\mathcal{O})$ is solvable, and hence
1746: amenable.
1747: We may now apply Theorem \ref{thm:qfa} to conclude that
1748: $E(2,\mathcal{O})$ has Property \QFA.
1749: \end{proof}
1750: \begin{remark}
1751: It was already known \cite[p. 68]{serre:trees} that the groups covered
1752: by Theorem \ref{thm:rankone} possessed property \FA.
1753: \end{remark}
1754: \begin{remark}
1755: Another proof of \ref{thm:rankone} may be given as follows: First
1756: show that every unipotent is distorted. It follows that the bounded
1757: generators cannot (quasi)-act hyperbolically. It is shown in
1758: \cite[Corollary 3.6]{manning:qfa} that there are no parabolic isometries of
1759: quasi-trees, and so each of the bounded generators (quasi)-acts
1760: elliptically. It then follows from bounded generation that any orbit
1761: is bounded.
1762: \end{remark}
1763:
1764: Finally, we speculate on the variety of hyperbolic $\Gamma$-spaces, for
1765: $\Gamma=SL(2,\mathcal{O})$, where $\mathcal{O}$ is the
1766: ring of integers of a number field $k$.
1767: We have already remarked that $\Gamma$ is a lattice in
1768: \[\prod_{i=1}^s SL(2,\R)\times \prod_{i=1}^t SL(2,\C),\]
1769: where $s$ and $t$ are the number of real and complex places
1770: respectively. Projection to some factor gives an isometric action
1771: either on $\H^2$ or $\H^3$. Call a hyperbolic $\Gamma$--space
1772: \emph{standard} if it is equivalent to $\H^2$ or $\H^3$ with one of
1773: these actions.
1774:
1775: \begin{conjecture}\label{conjecture:realquadratic}
1776: Every quasi-action by $\Gamma$ on a Gromov hyperbolic metric space
1777: either has an invariant horoball or is standard.
1778: \end{conjecture}
1779:
1780: %
1781: %
1782: %
1783: %\bibliographystyle{gtart}
1784: %
1785:
1786: %\bibliographystyle{abbrv}
1787: %\small
1788: %\bibliography{../../mybib.bib}
1789: \def\cprime{$'$} \providecommand\url[1]{\texttt{#1}}
1790: \begin{thebibliography}{10}
1791:
1792: \bibitem{abe:chevalley}
1793: E.~Abe.
1794: \newblock Chevalley groups over local rings.
1795: \newblock {\em T\^ohoku Math. J. (2)}, 21:474--494, 1969.
1796:
1797: \bibitem{bestvinafujiwara:mcg}
1798: M.~Bestvina and K.~Fujiwara.
1799: \newblock Bounded cohomology of subgroups of mapping class groups.
1800: \newblock {\em Geometry and Topology}, 6:69--89, 2002.
1801:
1802: \bibitem{bridhaef:book}
1803: M.~R. Bridson and A.~Haefliger.
1804: \newblock {\em Metric Spaces of Non--Positive Curvature}, volume 319 of {\em
1805: {G}rundlehren der mathematischen {W}issenschaften}.
1806: \newblock Springer--Verlag, Berlin, 1999.
1807:
1808: \bibitem{burgermonod:gafa}
1809: M.~Burger and N.~Monod.
1810: \newblock Continuous bounded cohomology and applications to rigidity theory.
1811: \newblock {\em Geom. Funct. Anal.}, 12(2):219--280, 2002.
1812:
1813: \bibitem{carterkeller:slno}
1814: D.~Carter and G.~Keller.
1815: \newblock Bounded elementary generation of {${\rm SL}\sb{n}({\mathcal O})$}.
1816: \newblock {\em Amer. J. Math.}, 105(3):673--687, 1983.
1817:
1818: \bibitem{carter:book}
1819: R.~W. Carter.
1820: \newblock {\em Simple groups of {L}ie type}.
1821: \newblock John Wiley \& Sons, London-New York-Sydney, 1972.
1822: \newblock Pure and Applied Mathematics, Vol. 28.
1823:
1824: \bibitem{chevalley:tohoku}
1825: C.~Chevalley.
1826: \newblock Sur certains groupes simples.
1827: \newblock {\em T\^ohoku Math. J. (2)}, 7:14--66, 1955.
1828:
1829: \bibitem{cdp}
1830: M.~Coornaert, T.~Delzant, and A.~Papadopoulos.
1831: \newblock {\em G\'eom\'etrie et th\'eorie des groupes: Les groupes
1832: hyperboliques de Gromov}, volume 1441 of {\em Lecture Notes in Mathematics}.
1833: \newblock Springer-Verlag, Berlin, 1990.
1834:
1835: \bibitem{fujiwara:gromovhyperbolic}
1836: K.~Fujiwara.
1837: \newblock The second bounded cohomology of a group acting on a
1838: {G}romov-hyperbolic space.
1839: \newblock {\em Proc. London Math. Soc. (3)}, 76(1):70--94, 1998.
1840:
1841: \bibitem{fukunaga:FA}
1842: M.~Fukunaga.
1843: \newblock Fixed points of elementary subgroups of {C}hevalley groups acting on
1844: trees.
1845: \newblock {\em Tsukuba J. Math.}, 3(2):7--16, 1979.
1846:
1847: \bibitem{gelanderkarlssonmargulis}
1848: T.~Gelander, A.~Karlsson, and G.~A. Margulis.
1849: \newblock Superrigidity, generalized harmonic maps and uniformly convex spaces.
1850: \newblock {\em Geom. Funct. Anal.}, 17(5):1524--1550, 2008.
1851:
1852: \bibitem{gromov:wordhyperbolic}
1853: M.~Gromov.
1854: \newblock Word hyperbolic groups.
1855: \newblock In S.~M. Gersten, editor, {\em Essays in Group Theory}, volume~8 of
1856: {\em Mathematical Sciences Research Institute Publications}, pages 75--264.
1857: Springer--Verlag, New York, 1987.
1858:
1859: \bibitem{rhds}
1860: D.~Groves and J.~F. Manning.
1861: \newblock {Dehn filling in relatively hyperbolic groups}.
1862: \newblock {\em Israel Journal of Mathematics}.
1863: \newblock to appear, preprint at \url{arXiv:math/0601311v3}.
1864:
1865: \bibitem{humphreys:introduction}
1866: J.~E. Humphreys.
1867: \newblock {\em Introduction to {L}ie algebras and representation theory},
1868: volume~9 of {\em Graduate Texts in Mathematics}.
1869: \newblock Springer-Verlag, New York, 1978.
1870: \newblock Second printing, revised.
1871:
1872: \bibitem{karlssonnoskov}
1873: A.~Karlsson and G.~A. Noskov.
1874: \newblock Some groups having only elementary actions on metric spaces with
1875: hyperbolic boundaries.
1876: \newblock {\em Geom. Dedicata}, 104:119--137, 2004.
1877:
1878: \bibitem{kotschick:quasi}
1879: D.~Kotschick.
1880: \newblock Quasi-homomorphisms and stable lengths in mapping class groups.
1881: \newblock {\em Proc. Amer. Math. Soc.}, 132(11):3167--3175 (electronic), 2004.
1882:
1883: \bibitem{manning:cocycles}
1884: J.~F. Manning.
1885: \newblock {Geometry of pseudocharacters}.
1886: \newblock {\em Geometry and Topology}, 9:1147--1185, 2005.
1887: \newblock \url{arXiv:math.GR/0303380}.
1888:
1889: \bibitem{manning:qfa}
1890: J.~F. Manning.
1891: \newblock {Quasi-actions on trees and Property QFA}.
1892: \newblock {\em Journal of the London Mathematical Society}, 73(1):84--108,
1893: 2006.
1894: \newblock With an appendix by Nicolas Monod and Bertrand R{\'e}my.
1895:
1896: \bibitem{matsumoto:csp}
1897: H.~Matsumoto.
1898: \newblock Sur les sous-groupes arithm\'etiques des groupes semi-simples
1899: d\'eploy\'es.
1900: \newblock {\em Ann. Sci. \'Ecole Norm. Sup. (4)}, 2:1--62, 1969.
1901:
1902: \bibitem{monod:superrigid}
1903: N.~Monod.
1904: \newblock Superrigidity for irreducible lattices and geometric splitting.
1905: \newblock {\em J. Amer. Math. Soc.}, 19(4):781--814 (electronic), 2006.
1906:
1907: \bibitem{monodshalom:jdg}
1908: N.~Monod and Y.~Shalom.
1909: \newblock Cocycle superrigidity and bounded cohomology for negatively curved
1910: spaces.
1911: \newblock {\em J. Differential Geom.}, 67(3):395--455, 2004.
1912:
1913: \bibitem{wittemorris:ckp}
1914: D.~W. Morris.
1915: \newblock Bounded generation of {${\rm SL}(n,A)$} (after {D}. {C}arter, {G}.
1916: {K}eller, and {E}. {P}aige).
1917: \newblock {\em New York J. Math.}, 13:383--421 (electronic), 2007.
1918:
1919: \bibitem{msw:quasiactI}
1920: L.~Mosher, M.~Sageev, and K.~Whyte.
1921: \newblock Quasi-actions on trees. {I}. {B}ounded valence.
1922: \newblock {\em Ann. of Math. (2)}, 158(1):115--164, 2003.
1923:
1924: \bibitem{serre:trees}
1925: J.-P. Serre.
1926: \newblock {\em Trees}.
1927: \newblock Springer-Verlag, Berlin, 1980.
1928: \newblock Translated from the French by John Stillwell.
1929:
1930: \bibitem{shalen:handbook}
1931: P.~B. Shalen.
1932: \newblock Representations of 3-manifold groups.
1933: \newblock In {\em Handbook of geometric topology}, pages 955--1044.
1934: North-Holland, Amsterdam, 2002.
1935:
1936: \bibitem{stein71}
1937: M.~R. Stein.
1938: \newblock Generators, relations and coverings of {C}hevalley groups over
1939: commutative rings.
1940: \newblock {\em Amer. J. Math.}, 93:965--1004, 1971.
1941:
1942: \bibitem{steinberg:notes}
1943: R.~Steinberg.
1944: \newblock {\em Lectures on {C}hevalley groups}.
1945: \newblock Yale University, New Haven, Conn., 1968.
1946: \newblock Notes prepared by John Faulkner and Robert Wilson.
1947:
1948: \bibitem{stewarttall}
1949: I.~Stewart and D.~Tall.
1950: \newblock {\em Algebraic number theory and {F}ermat's last theorem}.
1951: \newblock A K Peters Ltd., Natick, MA, third edition, 2002.
1952:
1953: \bibitem{tavgen:bg}
1954: O.~I. Tavgen{\cprime}.
1955: \newblock Bounded generability of {C}hevalley groups over rings of
1956: {$S$}-integer algebraic numbers.
1957: \newblock {\em Izv. Akad. Nauk SSSR Ser. Mat.}, 54(1):97--122, 221--222, 1990.
1958:
1959: \end{thebibliography}
1960:
1961:
1962:
1963: \end{document}
1964: