math0604077/infty
1: \documentclass[10pt]{amsart}
2: 
3: \usepackage{graphicx}
4: \usepackage{amssymb,mathrsfs}
5: 
6: \usepackage{enumerate}        %for easier lists
7: 
8: \usepackage[all,dvips]{xy}    %to get \xymatrix for diagrams
9: \UseComputerModernTips        %better looking tips for diagrams
10: %we may not need any diagrams; this is here just in case for now
11: 
12: 
13: %new commands and aliases
14: \newcommand{\im}{\operatorname{im}}
15: \newcommand{\dom}{\operatorname{dom}}
16: \newcommand {\Z}{\mathbb{Z}}
17: \newcommand{\bw}[1]{\langle #1 \rangle_{b}}     
18: \newcommand{\aw}[1]{\langle #1 \rangle_{a}}     
19: \newcommand {\cC}{\mathcal{C}}
20: \newcommand {\cF}{\mathcal{F}}
21: \newcommand{\dbpw}{\mathcal{D}_{b} \left( \mathcal{P}, w \right)}
22: \newcommand{\kbpw}{\widetilde{K}_{b} \left( \mathcal{P}, w \right)}
23: \newcommand{\kapw}{\widetilde{K}_{a} \left( \mathcal{P}, w \right)}
24: \newcommand{\kpw}{\widetilde{K} \left( \mathcal{P}, w \right)}
25: \newcommand{\dapw}{\mathcal{D}_{a} \left( \mathcal{P}, w \right)}
26: \newcommand{\dpw}{\mathcal{D} \left( \mathcal{P}, w \right)}
27: \newcommand {\st}{\hspace{-.4ex}*}
28: \newcommand {\dc}{\left( d_{C} \circ c \right)}
29: \newcommand{\di}[3]{\mathcal{D}_{#1} \left( \mathcal{#2}, #3 \right)} 
30: \newcommand{\kt}[3]{\widetilde{K}_{#1} \left( \mathcal{#2}, #3 \right)} 
31: 
32: \newtheorem{theorem}{Theorem}[section]
33: \newtheorem{lemma}[theorem]{Lemma}
34: \newtheorem{sublemma}[theorem]{Sub-Lemma}
35: \newtheorem{proposition}[theorem]{Proposition}
36: \newtheorem{corollary}[theorem]{Corollary}
37: \newtheorem{conjecture}[theorem]{Conjecture}
38: \newtheorem{question}[theorem]{Question}
39: \theoremstyle{definition}
40: \newtheorem{definition}[theorem]{Definition}
41: \newtheorem{example}[theorem]{Example}
42: \newtheorem{remark}[theorem]{Remark}
43: 
44: \sloppy 
45: 
46: \begin{document}
47: 
48: 
49: \title[Thompson's Group]{The Action of Thompson's Group on a CAT(0) Boundary}
50: \author[D.Farley]{Daniel Farley}
51:       \address{Max Planck Institute for Mathematics\\
52:                53111 Bonn}
53:       \email{farley@math.uiuc.edu}
54: 
55: \begin{abstract}  
56: For a given locally finite CAT(0) cubical complex $X$ with base vertex $\ast$, we define the 
57: \emph{profile} of a given geodesic ray $c$ issuing from $\ast$ to be the collection of all hyperplanes
58: (in the sense of \cite{Sag}) crossed by $c$.  We give necessary conditions for a collection
59: of hyperplanes to form the profile of a geodesic ray, and conjecture that these conditions
60: are also sufficient.
61: 
62: We show that profiles in diagram and picture complexes can be expressed naturally as infinite 
63: pictures (or diagrams), and use this fact to describe the fixed points at infinity of the actions  by 
64: Thompson's groups $F$, $T$, and $V$ on their respective  
65: CAT(0) cubical complexes.  
66: In particular, the actions of $T$ and $V$ have no global fixed points.
67: We obtain a partial description of the fixed set of $F$; it consists, at least, of
68: an arc $c$ of Tits length $\pi /2$, and any other fixed points of $F$ must have one particular profile,
69: which we describe.  We conjecture that all of the fixed points of $F$ lie on the arc $c$. 
70: 
71: Our results are motivated by the problem of determining whether $F$ is amenable.
72: \end{abstract}
73: 
74: \subjclass[2000]{Primary 20F65 Secondary 20F69 }
75: 
76: \keywords{amenability, CAT(0) cubical complex, Thompson's group, diagram group, space at infinity}
77: 
78: \maketitle
79: 
80: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
81: 
82: \section{Introduction}
83: 
84: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
85: 
86: Thompson's group $F$ is the group of piecewise linear 
87: homeomorphisms $h: [0,1] \rightarrow [0,1]$
88: satisfying:
89: \begin{enumerate}
90: \item the finitely many points at which $h$ is non-differentiable are all dyadic rational
91: numbers, and
92: \item if $h$ is differentiable at $x_0$, then $h'(x_0) \in \{ 2^{i} \mid i \in \mathbb{Z} \}$.
93: \end{enumerate}
94: Thompson also described two other groups, $T$ and $V$,
95: which are (respectively) the groups of  piecewise linear homeomorphisms $h$ of the circle
96: $[0,1]/(0=1)$ and the right-continuous bijections $h$ of $[0,1)$; in both cases the functions $h$ 
97: are required to satisfy (1) and (2).
98: The survey by Cannon, Floyd, and Parry \cite{CFP} is a useful introduction to all of these groups.
99: 
100: We are interested in the following question:
101: \begin{question} \label{biggie} \cite{Ross}
102: Is Thompson's group $F$ amenable?
103: \end{question}
104: To explain the original interest in \ref{biggie}, we will need a few definitions.
105: A group $G$ is \emph{elementary amenable} if it is in the smallest class of groups
106: that is closed under extensions and direct limits, and contains finite and abelian groups.
107: A group $G$ is \emph{amenable} if there is a measure $\mu : \mathcal{P}(G) \rightarrow [0,1]$
108: ($\mathcal{P}(G)$ is the power set of $G$) 
109: such that: i)  $\mu$ is finitely additive; ii) $\mu$ is left invariant; and iii) $\mu(G) = 1$.
110: We let $EG$, $AG$, and $NF$ denote the classes of elementary amenable groups, amenable groups, 
111: and groups with no free non-abelian subgroups (respectively).  Von Neumann showed that
112: $EG \subseteq AG \subseteq NF$.  The problem of determining whether these 
113: inclusions are proper was posed by Day \cite{Day}.   
114: 
115: Brin and Squier \cite{BS} showed that $F \in NF$.
116: It is proved in \cite{CFP} that $F \not \in EG$.  
117: Thus, the existence of $F$ implies that at least one of the inclusions 
118: $EG \subseteq AG \subseteq NF$ is proper for finitely presented groups:  a positive answer to
119: \ref{biggie} shows that $F \in AG - EG$, and a negative answer shows that $F \in NF - AG$.  Since at
120: least 1980, when Geoghegan posed Question \ref{biggie}, Thompson's group was expected by many to be 
121: an example of a finitely presented non-amenable group with no free subgroups.  
122: 
123: We know today that $EG \subsetneq AG \subsetneq NF$, and that the inclusions are proper for both 
124: finitely generated and finitely presented groups.
125: Grigorchuk found examples of finitely generated groups in $AG-EG$ \cite{Gri1} and, in 1998,
126: finitely presented examples as well \cite{Gri2}.  
127: In 1980, Ol'shanskii \cite{Ol1} found finitely generated
128: groups in $NF-AG$.  He and Sapir constructed finitely presented groups in $NF-AG$ in 2002 \cite{OlS}.  
129: 
130: Although the original reason to consider \ref{biggie} is thus obsolete, 
131: the problem of determining whether $F$ is amenable is still of great interest, and
132: motivates much of the current work about $F$.
133: 
134: Here we attempt to resolve \ref{biggie} negatively using CAT(0) geometry.  Two earlier results are of vital
135: importance in this.  First, Adams and Ballmann \cite{AB} showed 
136: that an amenable group $G$ which acts by isometries
137: on a locally compact CAT(0) space $X$ must either leave a finite-dimensional flat 
138: invariant or fix a point at infinity. 
139: Second, \cite{Me1, Me2} showed that Thompson's
140: groups $F$, $T$, and $V$ act properly, discretely, and by isometries on 
141: proper CAT(0) cubical complexes $X_{F}$, $X_{T}$, and $X_{V}$.  If $F$ is amenable, it must
142: therefore either leave a flat invariant or fix a point at infinity in $X_{F}$.  Elementary properties of $F$
143: (for instance, the fact that $\oplus_{n=1}^{\infty} \mathbb{Z} \subseteq F$ \cite{CFP}) 
144: imply that $F$ cannot act properly and freely
145: on a finite-dimensional flat, 
146: so we would have a proof that $F$ is non-amenable if we showed that $F$ has no global fixed points
147: in $\partial X_{F}$.  (Note that the groups $T$ and $V$ are known to be non-amenable, since
148: both are known to contain non-abelian free subgroups.) 
149: 
150: Most of the effort in this paper goes into describing the spaces at infinity of the locally finite complexes
151: $X_F$, $X_T$, and $X_V$.  (In fact, our methods apply to all diagram groups and picture groups \cite{Me2}.)
152: At this point, some background
153: on $X_F$, $X_T$, and $X_V$ is in order; we restrict our remarks to $X_F$ for the sake of simplicity. 
154: The constructions in \cite{Me1, Me2} come from the theory of diagram groups, which is
155: due to Guba and Sapir \cite{GS}.  Each vertex of $X_F$ is labelled by a semigroup diagram, which is essentially
156: a picture demonstrating how to derive an equality $w_1 = w_2$ between words $w_1$, $w_2$ over a semigroup
157: presentation.  The group $F$ is 
158: itself a diagram group, so every element in $F$ can be represented by a semigroup
159: diagram as well.  The action of $F$ on $X_F$ is given by a natural operation on diagrams:  if 
160: $x \in F$ and $v \in X_{F}^{0}$, then $x \cdot v$ is obtained by stacking the pictures $x$ and $v$, and then 
161: ``reducing dipoles".  (We refer the reader to Section \ref{diagram} for more specifics, 
162: or to \cite{GS} for a 
163: complete introduction.)  
164: 
165: Our description of $\partial X_{F}$ is of the same character.
166: We represent regions of $\partial X_{F}$ as infinite diagrams, which we call
167: \emph{profiles}.
168: The action of $F$ on profiles is determined, as before, by stacking diagrams.  As a result, we can
169: largely reduce the problem of finding fixed points in $\partial X_F$ to a much easier
170: algebra problem, which can be handled by a case analysis.  Our main theorem is as follows:
171: 
172: \begin{theorem} \label{biggie2}
173: Thompson's group $F$ fixes an arc in the boundary $\partial X_F$ of Tits length
174: $\pi /2$.  Any other fixed points on the boundary of $F$ lie in the profile $\Delta_{\infty}$.
175: Thompson's groups $T$ and $V$ act without global fixed points on their respective boundaries.
176: \end{theorem}
177: 
178: This unfortunately leaves \ref{biggie} open.  
179: 
180: The problem of finding any remaining fixed points of the action by $F$ appears to be rather delicate.
181: In Section \ref{infty}, we give some evidence for and against the existence of additional fixed points
182: in $\partial X_F$.
183: 
184: %It would be interesting to have a complete list of the fixed points for $F$, since such a list
185: %could suggest how to alter the space $X_F$ in order to produce a new locally compact 
186: %CAT(0) space $X_{F}'$ such that
187: %the action of $F$ (or a subgroup) on the boundary $\partial X_{F}'$ has no global fixed points.  
188: %One possible candidate for $X_{F}'$
189: %can be described as follows.  Let $\zeta \in \partial X_F$ 
190: %be the midpoint of the arc $c$ from Theorem \ref{biggie2}.  
191: %The group $[F,F] \rtimes \mathbb{Z}/2\mathbb{Z}$
192: %(where $\mathbb{Z}/2\mathbb{Z}$ is generated by 
193: %the linear homeomorphism $f(x) = 1-x$) acts on $X_F$ in an obvious way,
194: %leaves every horofunction centred at $\zeta$ invariant, and leaves no other point along the arc $c$ fixed.
195: %Let $X_{F}' = X_{F} - B$, where $B$ is a suitable open horoball centred at $\zeta$.  
196: %Now $[F,F] \rtimes \mathbb{Z}/2\mathbb{Z}$
197: %acts on $X_{F}'$ by isometries, and has no obvious fixed points on the boundary.  
198: %If one could show that there really are
199: %no fixed points, and that $X_{F}'$ is still CAT(0), then $[F,F]$ and, therefore, $F$ would be non-amenable.
200: 
201: %It is natural to ask more generally
202: %if Thompson's group $F$ fixes points at infinity of any locally compact CAT(0) space  
203: %$X$ on which it acts by isometries.  Our combinatorial methods do not really help to resolve this
204: %question.  
205: 
206: 
207: The paper is organized as follows.  In Section \ref{CAT0}, we collect various facts about CAT(0)
208: geometry which will be useful in later sections.  In Section \ref{diagram}, we briefly sketch the
209: definitions of diagram groups and the cubical complexes, called diagram complexes, on which they act.  
210: In Section \ref{pict},
211: we describe regions in the space at infinity of diagram complexes using infinite diagrams, and
212: describe the action of a diagram group on this space at infinity.  Section \ref{main} contains the
213: main part of the argument, where it is proved that Thompson's group $F$ fixes the profiles 
214: $\Delta_{L}$, $\Delta_{R}$, $\Delta_{L-R}$, and $\Delta_{\infty}$, each of which can be described 
215: by an infinite tree. The groups $T$ and $V$ fix only the profile
216: $\Delta_{\infty}$.  
217: In Section \ref{end},
218: we show that the profiles $\Delta_{L-R}$, $\Delta_{L}$, and $\Delta_{R}$ represent (respectively)
219: the interior of, the ``left'' endpoint of, and the ``right'' endpoint of an arc of length $\pi /2$
220: in the Tits metric.  We show moreover that all points on this arc are fixed by $F$.
221: Finally, in Section \ref{infty} we show that the region at infinity which we call 
222: $\Delta_{\infty}$  contains no fixed points of $T$ or $V$, even though it is fixed as a set.
223: As a result, one has a proof that $T$ and $V$ fix no point at infinity.  We also discuss the 
224: problem of determining whether $F$ fixes any points in $\Delta_{\infty}$.  
225: 
226: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
227: 
228: \section{Background on CAT(0) Spaces} \label{CAT0}
229: 
230: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
231: 
232: \subsection{Basic Definitions}
233: 
234: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
235: 
236: We begin by recalling several basic facts about CAT(0) spaces, all of which are taken directly from 
237: \cite{BH}.
238: 
239: A metric space $X$ is \emph{geodesic} if, for any $x_1 , x_2 \in X$, there is an
240: isometric embedding $c: [ 0, d(x_1 , x_2 ) ] \rightarrow X$, called a \emph{geodesic},
241: such that $c(0) = x_1$ and $c( d(x_1 , x_2 )) = x_2$.  We frequently confuse a geodesic
242: with its image.  A \emph{geodesic triangle} $\Delta(x,y,z)$
243: consists of three points $x,y,z \in X$ and choices of geodesics $[x,y]$, $[y,z]$, $[x,z]$
244: connecting them.  Given such a triangle, it is always possible to find points $\overline{x}$,
245: $\overline{y}$, $\overline{z}$ in two-dimensional Euclidean space $\mathbb{E}^{2}$  
246: such that
247: $d_X ( x,y ) = d_{\mathbb{E}^{2}} ( \overline{x} , \overline{y} )$,  
248: $d_X ( y,z ) = d_{\mathbb{E}^{2}} ( \overline{y} , \overline{z} )$,  
249: and $d_X ( x,z ) = d_{\mathbb{E}^{2}} ( \overline{x} , \overline{z} )$.  The triangle 
250: $\overline{\Delta}(\overline{x}, \overline{y}, \overline{z})$
251: in $\mathbb{E}^{2}$ determined by $\overline{x}$, $\overline{y}$, and $\overline{z}$ is called
252: a \emph{comparison triangle} for $\Delta$.  There is a map $h: \Delta \rightarrow \overline{\Delta}$ which
253: sends sides of $\Delta$ isometrically to the corresponding sides of $\overline{\Delta}$.  We say that the
254: triangle $\Delta$ satisfies the \emph{CAT(0) inequality} if $d_{X}(a,b) \leq d_{\mathbb{E}^{2}}( h(a), h(b))$
255: whenever $a,b \in \Delta$.  
256: A geodesic metric space $X$ is \emph{CAT(0)} if all geodesic triangles in $X$ satisfy
257: the CAT(0) inequality.  CAT(0) spaces are contractible, and \emph{uniquely geodesic}, i.e., given any two points
258: $x_1$, $x_2$ in a CAT(0) space $X$, there is a unique geodesic connecting $x_1$ to $x_2$.
259: 
260: If $X$ is an arbitrary metric space, and $c: [0,a] \rightarrow X$, $c' : [0,a'] \rightarrow X$
261: are geodesic segments satisfying $c(0) = c'(0)$, then  we define the \emph{Alexandrov angle} 
262: $\angle (c, c')$ as follows:
263: $$ \angle(c,c') := \lim_{\epsilon \rightarrow 0} 
264: \sup_{0 <  t, t' < \epsilon} \overline{\angle}_{c(0)} (c(t), c'(t')).$$
265: Here $\overline{\angle}_{c(0)} (c(t), c'(t'))$ is the angle at $\overline{c(0)}$ in the comparison
266: triangle $\overline{\Delta}$ for $\Delta \left( c(0), c(t), c'(t') \right)$.  Given
267: three points $x$, $y$, $z$ 
268: in a CAT(0) space $X$, we let $\angle_{y}(x,z)$ denote the Alexandrov angle
269: between the (unique) geodesics $[y,x]$ and $[y,z]$.
270: 
271: The CAT(0) inequality can also be expressed in terms of the Alexandrov angle.  If $\Delta$ is a 
272: geodesic triangle in the metric space
273: $X$, then $\Delta$ satisfies the CAT(0) inequality
274: if and only if each Alexandrov angle in $\Delta$ measures less than the corresponding angle in the
275: comparison triangle $\overline{\Delta}$.  We say that a geodesic metric space $X$ is CAT(0) if every 
276: geodesic triangle in $X$ satisfies
277: this version of the CAT(0) inequality.  Bridson and Haefliger \cite{BH} show that this definition 
278: of CAT(0) spaces is
279: equivalent to the earlier one.  
280: 
281: A complete CAT(0) space $X$ has a natural space at infinity $\partial X$, which we now define.
282: Two geodesic rays $c,c' : [0, \infty) \rightarrow X$ are said to be \emph{asymptotic} if there exists a constant
283: $K$ such that $d(c(t), c'(t)) \leq K$ for all $t \geq 0$.  The set $\partial X$ of \emph{boundary points}
284: of $X$ (or \emph{points at infinity}) is the set of equivalence classes of geodesic rays, where two
285: geodesic rays are equivalent if and only if they are asymptotic.  In practice, we will always use
286: a basepointed version of this construction.  Fix a point $x \in X$.  We define $\partial X$
287: to be the set of geodesic rays $c: [0, \infty) \rightarrow X$ issuing from $x$, i.e., satisfying $c(0)=x$.  These
288: two definitions of $\partial X$ are equivalent in a complete CAT(0) space 
289: by the following proposition:
290: 
291: \begin{proposition} \cite{BH} \label{parallel} 
292: If $X$ is a complete CAT(0) space and $c : [0, \infty) \rightarrow X$ is a geodesic ray issuing
293: from $x$, then for every point $x' \in X$ there is a unique geodesic ray $c'$ which issues from
294: $x'$ and is asymptotic to $c$.
295: \qed
296: \end{proposition}  
297: 
298: If a group $G$ acts by isometries on the CAT(0) space $X$, then it is clear that there is an induced
299: action on $\partial X$, if we regard the latter as the collection of equivalence classes of geodesic rays
300: in $X$.  If we use the basepointed version of the construction, then the action $\ast$ can be described as follows:
301: Let $c \in \partial X$; i.e., $c: [0, \infty) \rightarrow X$ is a geodesic ray and $c(0)=x$.  For an isometry
302: $g \in G$, $g \ast c$ is the unique geodesic ray issuing from $x$ and asymptotic to the left-translate $g \cdot c$ of $c$.
303:  
304: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
305: 
306: \subsection{Convexity in CAT(0) Spaces}
307: 
308: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
309: 
310: A subset $C$ of a CAT(0) space $X$ is \emph{convex} if, given 
311: any two points $x_{1}, x_{2} \in C$, the (unique) geodesic segment $[ x_{1}, x_{2} ]$ 
312: is contained in $C$.  A function 
313: $f: X \rightarrow \mathbb{R}$ on a geodesic metric space is \emph{convex} if, for any 
314: geodesic $c: I \rightarrow X$, the composition $f \circ c$ is convex in the ordinary 
315: sense, i.e., if, for any $t, t' \in I$ and $s \in [0,1]$, 
316: $$ (f \circ c) \left( (1-s)t + st' \right) \leq (1-s) \left( f \circ c \right)(t) 
317: + s \left( f \circ c \right) \left( t' \right).$$
318: 
319: Bridson and Haefliger \cite{BH} show that there is a natural projection $\pi_{C}: X \rightarrow C$
320: defined whenever $X$ is a complete CAT(0) space and $C$ is a closed convex subspace.  We collect some
321: basic properties of $\pi_{C}$ here.   
322: 
323: \begin{proposition} \label{projchar}
324: Let $C$ be a closed, convex subspace of a complete CAT(0) space $X$.
325: \begin{enumerate}
326: \item \cite{BH}   
327: Let 
328: $d_{C}: X \rightarrow \mathbb{R}$ be defined by the rule
329: $$d_{C}(x) = \mathrm{inf}_{y \in C} d(x,y).$$   
330: The function $d_{C}$ is convex.
331: \item \cite{BH}
332: For any $x \in X$, there is a unique point $\pi_{C} (x) \in C$ such that 
333: $d(x, \pi_{C}(x) ) = d_{C}(x)$.  The function $\pi_{C} : X \rightarrow C$
334: does not increase distances.
335: \item \cite{BH}
336: If $x \in X-C$ and $y \in C- \left\{ \pi(x) \right\}$, then $\angle_{\pi(x)}(x,y) \geq \pi /2$. 
337: \item Fix $x \in X-C$.  If $y \in C$ satisfies $\angle_{y}( x, y' ) \geq \pi /2$ for
338: all $y' \in C - \{ y \}$, then $y = \pi_{C}(x)$.
339: \end{enumerate} 
340: \end{proposition}
341: 
342: \begin{proof}
343: (4)  Assume that $y$ satisfies the above condition and $y \neq \pi(x)$.  Consider the comparison triangle 
344: $\overline{\Delta} \left( \overline{x}, \overline{y}, \overline{\pi(x)} \right)$ for 
345: the geodesic triangle $\Delta ( x, y, \pi(x) )$ in $X$.  Since the comparison triangle is
346: non-degenerate by our assumptions, at least one of the comparison angles
347: $\overline{\angle}_{\overline{y}} \left( \overline{x}, \overline{\pi(x)} \right)$,
348: $\overline{\angle}_{\overline{\pi(x)}} \left( \overline{x}, \overline{y} \right)$ measures less than
349: $\pi /2$.  Our assumptions and the CAT(0) inequality imply that 
350: $\overline{\angle}_{\overline{\pi(x)}} \left( \overline{x}, \overline{y} \right) < \pi /2$.  By the
351: CAT(0) inequality, $\angle_{\pi(x)}(x,y) < \pi /2$ as well.  This violates (3). 
352: \end{proof}
353: 
354: We say that a geodesic ray $c: [0, \infty) \rightarrow X$ \emph{crosses} a closed, convex subset $C$
355: if $(\mathrm{Im} \, c) \cap C$ is a non-empty, compact interval and $(\mathrm{Im} \, c) - C$ is disconnected.
356: 
357: \begin{lemma} \label{monotone}
358: Let $X$ be a CAT(0) space.
359: \begin{enumerate}
360: \item
361:  Suppose that $C$ is a closed convex subset of $X$,   
362: $c: [0, \infty) \rightarrow X$
363: is a geodesic ray which crosses $C$, and    
364: $(\mathrm{Im} \, c) \cap C = c \left( \left[ t_1, t_2 \right] \right)$.  The function 
365: $d_{C} \circ c$ is strictly monotonically increasing on $[ t_{2}, \infty)$, and 
366: $\left( d_{C} \circ c \right)(t) \rightarrow \infty$ as $t \rightarrow \infty$.
367: \item 
368: If $c, c' : [0, \infty) \rightarrow X$ are two asymptotic geodesic rays in $X$,
369: then the function $d( c( \underline{~ ~}), c' ( \underline{~ ~}) ): [0,\infty) \rightarrow [0, \infty)$ is
370: non-increasing.
371: \end{enumerate} 
372: \end{lemma}
373: 
374: \begin{proof}
375: Both parts are standard exercises using basic properties of convex functions.
376: Part (1) follows from the fact that $d_{C} \circ c : [0, \infty) \rightarrow [0, \infty)$ is convex, 
377: $\left( d_{C} \circ c \right) (t_{2}) = 0$,
378: and $\left( d_{C} \circ c \right) (t') >0$ for some $t' > t_{2}$.  Part (2) follows from the
379: fact that the function $d( c( \underline{~ ~}), c'( \underline{~ ~}) ): [0, \infty) \rightarrow [0, \infty)$ is convex 
380: and bounded (see \cite{BH}, page 261). 
381: \end{proof}
382: 
383: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
384: 
385: \subsection{CAT(0) Cubical Complexes}
386: 
387: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
388: 
389: We take the following definition of a cubical complex from \cite{BH}:
390: \begin{definition} (\cite{BH}, pg. 112) A cubical complex $K$ is the quotient of a disjoint union of
391: cubes $X = \coprod_{\Lambda} I^{n_{\lambda}}$ by an equivalence relation $\sim$.  The restrictions
392: $p_{\lambda} : I^{n_{\lambda}} \rightarrow K$ of the natural projection $p: X \rightarrow K = X/\sim$
393: are required to satisfy:
394: \begin{enumerate}
395: \item for every $\lambda \in \Lambda$ the map $p_{\lambda}$ is injective;
396: \item if $p_{\lambda}( I^{n_{\lambda}} ) \cap p_{\lambda'}(I^{n_{\lambda'}}) \neq \emptyset$ then there
397: is an isometry $h_{ \lambda, \lambda'}$ from a face $T_{\lambda} \subseteq I^{n_{\lambda}}$ onto a face
398: $T_{\lambda'} \subseteq I^{n_{\lambda'}}$ such that $p_{\lambda}(x) = p_{\lambda'}(x')$ if and only if
399: $x' = h_{\lambda, \lambda'}(x)$.
400: \end{enumerate}
401: \end{definition}   
402: 
403: Let $x$ and $y$ be points in $X$, and let $l(c)$
404: denote the length of a path $c$.   
405: $$ d_{\ell}(x,y) = inf \{ l(c) \mid c(0) = x; c(1) = y \}.$$
406: The function $d_{\ell}: X \times X \rightarrow [0, \infty]$ defines a metric on any 
407: cubical complex, called the \emph{length metric} \cite{BH}.  A well-known theorem 
408: due to Gromov \cite{BH} 
409: says that the length metric on a cubical complex $X$ is  a CAT(0)  
410: metric if and only if $X$ satisfies the link condition.  
411: We avoid recounting the precise statement here, but will 
412: work exclusively with CAT(0) cubical complexes from now on.    
413: 
414: Let $X$ be a complete CAT(0) cubical complex.  Following \cite{Sag}, define a relation 
415: $\sim$ on edges of $X$, such that $e_1 \sim e_2$ if and only if $e_1$ and $e_2$ are opposite sides
416: of a square ($2$-cell)  in $X$.  We will sometimes 
417: call this relation \emph{simple square equivalence}, although it
418: is not an equivalence relation.  The transitive, reflexive closure of this relation, also
419: denoted $\sim$, is called \emph{square equivalence}.  It is clear that square equivalence is an 
420: equivalence relation.
421: 
422: A \emph{combinatorial hyperplane} in $X$ is an equivalence class of edges under $\sim$.  One 
423: obtains a geometric hyperplane $H$ as follows: let $M_{e}$ be the set of all midpoints of all edges
424: square equivalent to $e$.  If $C$ is an arbitrary cube, define $H \cap C$ to be the convex hull
425: of $M_{e} \cap C$ in the cube $C$.  This description determines $H$.  
426: 
427: We now collect some basic properties of CAT(0) cubical complexes.
428: 
429: \begin{theorem} \label{Sageev} \cite{Sag} Let $X$ be a CAT(0) cubical complex. 
430: \begin{enumerate}
431: \item If $J$ is a geometric hyperplane in $X$, then $J$ does not intersect itself and partitions
432: $X$ into two convex components.
433: \item If $J_1, \ldots, J_k$ are a collection of geometric hyperplanes in $X$ such that
434: $J_m \cap J_n \neq \emptyset$ for all $m,n$, then $\bigcap J_i \neq \emptyset$.
435: \item If $x$ and $y$ are vertices in $X$ connected by a geodesic edge-path $p$ of length $n$, then
436: $p$ crosses $n$ distinct hyperplanes $J_1, \ldots, J_n$, and these are precisely the hyperplanes
437: which separate $x$ from $y$.  In particular, any other geodesic edge-path $p'$ connecting $x$ to $y$ must
438: cross precisely the same hyperplanes.   
439: \item Each geometric hyperplane $J$ is itself a CAT(0) cubical complex.
440: \qed
441: \end{enumerate}
442: \end{theorem}
443: 
444: The following lemma will be useful in Subsection \ref{profiles}.
445: 
446: \begin{lemma} \label{crossing} Let $X$ be a locally finite CAT(0) cubical complex. 
447: \begin{enumerate}
448: \item The closed 
449: $\frac{1}{2}$-neighborhood of a hyperplane $H$ in $X$ factors isometrically as $H \times [0,1]$. 
450: \item
451: Let $c: [0, \infty) \rightarrow X$ be a geodesic ray issuing from a vertex $\ast$ of $X$.   
452: If $c \left( [0, \infty) \right) \cap C \neq \emptyset$ for some open cube $C$ in $X$, then
453: $c$ crosses every hyperplane passing through $C$.  
454: \item  
455: If $H$ is a hyperplane in $X$, $d_{H}$ is non-constant on an open cell $C$ in $X$, and $H \cap C = \emptyset$, 
456: then there
457: exists some hyperplane $H_1$ passing through $C$ such that $H_1 \cap H = \emptyset$.
458: \item Let $H_1$ and $H_2$ be hyperplanes in the CAT(0) cubical complex $X$, and, for $i=1,2$, let $H_{i}^{+}$, $H_{i}^{-}$ 
459: be the two open, convex components of $X - H_i$.  If the intersections 
460: $H_{1}^{+} \cap H_{2}^{-}$, $H_{1}^{+} \cap H_{2}^{+}$, $H_{1}^{-} \cap H_{2}^{-}$,
461: and $H_{1}^{-} \cap H_{2}^{+}$ are all non-empty, then $H_{1} \cap H_{2}$ is also non-empty. 
462: \end{enumerate}
463: \end{lemma}
464: 
465: \begin{proof}
466: \begin{enumerate}
467: \item This is a consequence of the proof for Theorem 4.10 (page 611) of \cite{Sag}. 
468: \item Suppose that $c(t_1) \in C$; let $H$ be any hyperplane passing through $C$.  Thus 
469: $c(t_1) \in H \times (0,1)$; say $c(t_1) \in H \times \{ t' \}$.  Let $s_1$, $s_2$ be arbitrary numbers such
470: that $s_1 < t' < s_2$ and $\frac{1}{2} \in (s_1 , s_2)$.  The hyperplanes $H \times \{ s_i \} = H_{s_i}$
471: $(i=1,2)$ separate $X$ into three distinct connected components, one of which is $H \times (s_1 , s_2 )$.  Note
472: that $H \times [s_1 , s_2 ]$ contains no vertices of $X$, so $c$ must therefore cross either $H_{s_1}$ or
473: $H_{s_2}$.  If we assume, without loss of generality, that $c$ crosses $H_{s_1}$, then it must be that
474: $d_{H_{s_1}} (c(t)) \rightarrow \infty$ monotonically on $[ t_1 , \infty )$, for appropriate $t_1$, by 
475: Lemma \ref{monotone}.  The function
476: $d_{H_{s_1}}$ is bounded on $H \times [s_1 , s_2 ]$, so the geodesic ray $c$ eventually leaves
477: $H \times [s_1 , s_2 ]$, and it cannot cross $H_{s_1}$ a second time, due to the monotonicity of
478: $d_{H_{s_1}} \circ c$ on $[ t_1 , \infty )$.  It follows that $c$ crosses $H_{s_2}$, and thus also
479: $H_{1/2} = H$.
480: 
481: \item We prove the contrapositive.  Suppose that $H$ is a hyperplane, $C$ is an open cell such that $H \cap C = \emptyset$, and
482: every hyperplane $H'$ passing through $C$ satisfies $H' \cap H \neq \emptyset$.  We wish to show that $d_{H}$ is constant on
483: $C$.
484: 
485: Identify $C$ with $(0,1)^{n}$ and fix a factor of $(0,1)^{n}$ (the last one, without loss of generality).  There is a hyperplane
486: $H'$ such that $H' \cap C = (0,1)^{n-1} \times \left\{ \frac{1}{2} \right\}$.  Let $x_1$, $x_2 \in C$ be two points in $C$ which differ
487: only in the last coordinate, say $x_1 = ( c_1 , c_2 , \ldots , c_n )$ and $x_2 = ( c_1 , c_2 , c_3 , \ldots , \hat{c}_{n} )$.
488: We regard these as points in $H'_{c_n} := H' \times \{ c_n \}$ and $H'_{\hat{c}_n} := H' \times \{ \hat{c}_n \}$, respectively.
489: Let $x = \left( c_1 , c_2 , \ldots , c_{n-1}, \frac{1}{2} \right)$ be in $H'$, which we identify with 
490: $H' \times \left\{ \frac{1}{2} \right\}$.
491: 
492: We consider the projection $\pi_{H \cap H'} : H' \rightarrow H \cap H'$.  Let us suppose that
493: $\pi_{H \cap H'} (x) \in C'$ where $C' = (0,1)^{m}$ is an open cube 
494: (of dimension at least $2$, since $H$ and $H'$ both
495: pass through $C'$, and $H \neq H'$).  We make the identifications 
496: $H' \cap C' = \{ (d_1 , \ldots, d_m ) \in C' \mid d_m = 1/2 \}$ and
497: $H \cap C' = \{ (d_1, \ldots, d_m ) \in C' \mid d_{m-1} = 1/2 \}$.  
498: 
499: Suppose that $\pi_{H \cap H'} (x) = (d'_1 , d'_2, \ldots, d'_{m-2}, 1/2, 1/2 )$.  The requirement that
500: $\angle_{\pi_{H \cap H'}(x)}(x,y) \geq \pi/2$ for all $y \neq \pi_{H \cap H'}(x)$ 
501: in $H \cap H'$ guarantees that $[x, \pi_{H \cap H'}(x) ] \cap C' 
502: \subseteq \{ d'_1 \} \times \ldots \times \{ d'_{m-2} \} \times (0,1) \times \{ 1/2 \}$ (note: the last
503: coordinate must be $1/2$ since $[x, \pi_{H \cap H'}(x)] \subseteq H'$).  But it follows from this that
504: $\angle_{\pi_{H \cap H'}(x)}(x,y) \geq \pi/2$ for all $y \neq \pi_{H \cap H'}(x)$ in $H$.  That is:
505: $\pi_{H \cap H'} (x) = \pi_{H} (x)$
506: ,  by Proposition \ref{projchar} (4), where $\pi_{H \cap H'} : H' \rightarrow H \cap H'$ and $\pi_{H}: X \rightarrow H$
507: are the projections.   
508: 
509: Therefore the geodesic segment $[x, \pi_{H}(x)]$ ($= [x, \pi_{H \cap H'}(x)]$) is a subset
510: of $H'$.  Now we consider the geodesic segments $[x, \pi_{H}(x)] \times \{ c_n \} \subseteq H' \times \{ c_n \}$ and
511: $[x, \pi_{H}(x)] \times \{ \hat{c}_{n} \} \subseteq H' \times \{ \hat{c}_{n} \}$.  These run parallel
512: to $[x, \pi_{H}(x)]$, and meet $H$ perpendicularly for similar reasons.  It follows that
513: $[x_1 , \pi_{H}(x_1)] = [x, \pi_{H}(x)] \times \{ c_n \}$ and $[x_2, \pi_{H}(x_2)] = [x, \pi_{H}(x)] \times \{ \hat{c}_{n} \}$.
514: 
515: This implies that $d_{H}(x_1 ) = d_{H}(x) = d_{H} ( x_2 )$, which implies that
516: the value of $d_{H}$ is independent of the last coordinate.  Now we can argue
517: coordinate by coordinate to conclude that $d_H$ is constant on $C$.
518: 
519: \item
520: Assume that the four intersections in the hypothesis are all non-empty.  If we assume also 
521: that $H_1 \cap H_2 = \emptyset$, 
522: then it follows that $\left\{ H_{1}^{+} \cup H_{2}^{+}, H_{1}^{-} \cup H_{2}^{-} \right\}$ is an open cover of $X$.  Now each of the
523: half-spaces $H_{1}^{+}$, $H_{1}^{-}$, $H_{2}^{+}$, and $H_{2}^{-}$ is a convex subspace of a CAT(0) space, and therefore CAT(0) itself.
524: It follows that each is contractible.  The same reasoning also applies to the four intersections in the hypothesis: each is CAT(0), and
525: therefore contractible.  
526: 
527: It then follows that each of the sets $X^{+} = H_{1}^{+} \cup H_{2}^{+}$, $X^{-} = H_{1}^{-} \cup H_{2}^{-}$ is simply connected, since
528: each is the union of two open contractible sets which intersect in an open contractible set.   The intersection $X^{+} \cap X^{-}$ is the 
529: union of two disjoint open contractible sets:  
530: $H_{1}^{+} \cap H_{2}^{-}$ and  $H_{2}^{+} \cap H_{1}^{-}$.  Let $c$ be an arc contained in $X^{+}$ connecting $H_{1}^{+} \cap H_{2}^{-}$
531: to $H_{2}^{+} \cap H_{1}^{-}$ and meeting each in an open segment.  
532: 
533: We apply van Kampen's theorem to the pieces $X^{-} \cup c$ and $X^{+}$.  The first piece $X^{-} \cup c$ satisfies 
534: $\pi_{1} \left( X^{-} \cup c \right) \cong \mathbb{Z}$,
535: while the second is simply connected.  The intersection of these two pieces is the simply connected set 
536: $\left( H_{1}^{+} \cap H_{2}^{-} \right) \cup \left( H_{2}^{+} \cap H_{1}^{-} \right) \cup c$.  It follows that
537: $\pi_{1} \left( X^{-} \cup X^{+} \right) = \pi_{1} (X)$ is isomorphic to $\mathbb{Z}$.  The space $X$ is CAT(0), however, 
538: and therefore contractible.
539: We have a contradiction.
540: \end{enumerate} 
541: \end{proof}
542: 
543: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
544: 
545: \subsection{Profiles of Geodesic Rays in CAT(0) Cubical Complexes} \label{profiles}
546: 
547: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
548: Suppose now that $X$ is locally finite, and let $\ast$ be a vertex, which will serve as a basepoint.
549: If $H$ is any hyperplane in $X$, let $H^{+}$, the \emph{positive half-space} determined by $H$,
550: be the open complementary component of $X - H$ that doesn't contain $\ast$;  we let $H^{-}$ be the other
551: open complementary component of $X-H$. 
552: If $H_1$
553: and $H_2$ are hyperplanes in $X$, write $H_{1}^{+} \leq H_{2}^{+}$ if
554: $H_{2}^{+} \subseteq H_{1}^{+}$.  Clearly $\leq$ is a partial order on positive half-spaces.
555: We also regard $\leq$ as a partial order on hyperplanes, writing $H_1 \leq H_2$ if
556: $H_{1}^{+} \leq H_{2}^{+}$.  
557: 
558: For any geodesic ray $c$ in $X$ issuing from $\ast$, define $P(c)$, the \emph{profile of $c$},
559: to be the collection of all positive half-spaces $H^{+}$ such that $H$ is crossed by $c$.
560: 
561: \begin{proposition} \label{fund}
562: If $c: [0, \infty) \rightarrow X$ is a geodesic ray issuing from $\ast$, then $P(c)$ is non-empty
563: and satisfies:  
564: \begin{enumerate}
565: \item for any finite subset $\left\{ H_{1}^{+}, \ldots,  H_{n}^{+} \right\} \subseteq P(c)$,
566: $H_{1}^{+} \cap \ldots \cap H_{n}^{+} \neq \emptyset$;
567: \item the partially ordered set $\left( P(c), \leq \right)$ has no maximal
568: elements, and
569: \item if $H_{1}^{+} \in P(c)$ and $H_{2}^{+} \leq H_{1}^{+}$, then
570: $H_{2}^{+} \in P(c)$.
571: \end{enumerate}
572: \end{proposition}
573: 
574: \begin{proof}
575: If $H_{1}^{+}$, $H_{2}^{+}$, $\ldots$, $H_{n}^{+}$ are in $P(c)$, then there exist real numbers
576: $t_1, t_2, \ldots, t_n > 0$ such that $c( [ t_i , \infty) ) \subset H_{i}^{+}$, by Lemma \ref{monotone}.
577: If $t$ is the largest number in $\{ t_1 , \ldots, t_n \}$, then clearly 
578: $c( [t, \infty)) \subseteq H_{1}^{+} \cap H_{2}^{+} \cap \ldots \cap H_{n}^{+}$.  This proves that property (1) holds
579: for $P(c)$.  
580: 
581: It is obvious that property (3) is true of $P(c)$.
582: 
583: Suppose that $H^{+} \in P(c)$ is a maximal element, 
584: and let $c( [ t, \infty )) \subseteq H^{+}$.  Since $d_{H}(c(t')) \rightarrow  \infty$ as $t' \rightarrow \infty$ by Lemma
585: \ref{monotone},
586: we can choose $t$ so that $d_{H}(c(t)) > 1/2$.  We consider the collection $\mathcal{C}$ of all open cells $C$ such that
587: $c( [t, \infty )) \cap C \neq \emptyset$.  
588: Note that if $\widehat{C} \cap H \neq \emptyset$,
589: then any point in $c([ t, \infty)) \cap \widehat{C}$ is at most $1/2$-distant from $H$, so that every cell $C$ in $\mathcal{C}$
590: satisfies $C \cap H = \emptyset$.   
591: 
592: Let $C \in \mathcal{C}$.  If $d_{H}$ is non-constant on $C$, then, by Lemma \ref{crossing}(3), there is
593: a hyperplane $H'$ passing through $C$ such that $H' \cap H = \emptyset$.  We claim that 
594: this implies $H^{+} < (H')^{+}$.  
595: If $H^{+} \not < (H')^{+}$, i.e., 
596: if $(H')^{+} \not \subseteq H^{+}$, then $(H')^{+} \cap H^{-} \neq \emptyset$.  We also know
597: that $ \ast \in H^{-} \cap (H')^{-}$, $H^{+} \cap (H')^{-} \neq \emptyset$ 
598: (since $C \subseteq H^{+}$ and $C \cap (H')^{-} \neq \emptyset$),
599: and $H^{+} \cap (H')^{+} \neq \emptyset$ (since the geodesic ray must cross $H'$ by Lemma \ref{crossing}(2)).  Now it follows
600: from Lemma \ref{crossing}(4) that $H' \cap H \neq \emptyset$, which is a contradiction.  This proves the claim.  
601: Now $H^{+} < (H')^{+}$ and $c$ crosses $H'$ by Lemma \ref{crossing}(2), which contradicts the maximality of $H^{+}$.
602: 
603: It follows that
604: $d_{H}$ is constant on all of the cells $C$ in $\mathcal{C}$.   
605: This contradicts the fact that $d_{H} \circ c$ is strictly
606: monotonically increasing on $[t, \infty)$.  It follows that property (2) holds.
607: 
608: It is obvious that $P(c)$ is non-empty.
609: \end{proof} 
610: 
611: From now on, we call a collection of positive half-spaces $\mathcal{H}$ a \emph{profile}
612: if it is non-empty and satisfies properties (1)-(3) in Proposition \ref{fund}.  I don't know if
613: every profile $\mathcal{H}$ in this sense is 
614: realized by a geodesic ray, i.e., if there is some geodesic ray $c$ issuing
615: from $\ast$ such that $\mathrm{Im} \, c \, \cap H^{+} \neq \emptyset$ if and only if $H^{+} \in \mathcal{H}$.  
616: I make
617: the following conjecture.
618: 
619: \begin{conjecture} \label{conj} Let $X$ be a locally finite CAT(0) cubical complex with base vertex $\ast$.
620: \begin{enumerate}  
621: \item Every profile is realized by a geodesic ray issuing from $\ast$.
622: \item The collection of geodesic rays $c$ having a given, fixed profile $\mathcal{H}$
623: forms a subset of $\partial X$ of diameter less than or equal to $\pi/2$, where the distance
624: in question is the angle metric (see \cite{BH}).  
625: \end{enumerate}
626: \end{conjecture}
627: 
628: Part (2) of Conjecture \ref{conj} implies, in particular, that each profile represents
629: a contractible subset of $\partial X$, since $\partial X$ is a CAT(1) space with respect to
630: the angular metric, and sets of diameter less than $\pi$ are contractible in CAT(1) spaces
631: (Proposition 1.4(4) in \cite{BH}).  
632: If Conjecture \ref{conj} is true, then the description of $\partial X$ in terms of profiles 
633: may therefore give a useful homotopical view of the space at infinity, especially if profiles have a convenient description.  
634: We will give such a description of profiles 
635: in diagram complexes later in Section \ref{pict}.  
636: 
637: \begin{example}
638: We give a quick example to show why the most obvious approach to proving Conjecture \ref{conj} (1) fails.
639: Suppose that $X$ is a CAT(0) cubical complex with base vertex $v$; let $\mathcal{H} = \{ H_{1}^{+}, \ldots, H_{n}^{+}, 
640: \ldots \}$
641: be a profile of $X$.  For $n \in \mathbb{N}$, let $v_{n} \in H_{1}^{+} \cap \ldots \cap H_{n}^{+}$.  One might hope
642: that the sequence $( v_{n} )$ converges to a point at infinity which realizes the profile $\mathcal{H}$. 
643: 
644: Consider $\mathbb{R}^{2}$ endowed with the usual square complex structure in which integer lattice points
645: are the vertices.  We let $(0,0)$ be the base vertex.  It is not difficult to see that there are precisely
646: $8$ profiles; four of these profiles are realized by any geodesic ray issuing from $(0,0)$ and travelling
647: through one of the four open quadrants, and the other four are realized by geodesic rays travelling along the
648: coordinate axes.  The conjecture thus clearly holds in this case.  If we try to realize the profile corresponding
649: to the open quadrant $\mathbb{R}^{2,+} = \{ (x,y) \in \mathbb{R}^{2} \mid x,y > 0 \}$ by the above method, however, then we
650: find that nothing prevents us from choosing all of our points to be of the form $(x,x^{2})$.  Any such sequence would converge
651: to a point at infinity having the wrong profile.
652: \end{example}
653:  
654: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
655: 
656: \section{Diagram Groups} \label{diagram}
657: 
658: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
659: 
660: \subsection{Basic Definitions}
661: 
662: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
663: 
664: If $A$ is a set (alphabet), then the \emph{free semigroup on $A$}, denoted $A^{+}$,
665: is the collection of all positive, non-empty words in $A$, with the operation of
666: concatenation.  Let $\mathcal{P} = \langle \Sigma \mid \mathcal{R} \rangle$ be a semigroup presentation.
667: Thus, $\Sigma$ is an alphabet and $\mathcal{R} \subseteq \Sigma^{+} \times \Sigma^{+}$ is a collection
668: of equalities between elements of $\Sigma^{+}$.  We will follow the convention of \cite{Me1} and impose the additional
669: assumption that no relation of the form $(w,w)$ occurs in $\mathcal{R}$.
670: 
671: We now define pictures over $\mathcal{P}$.  Begin with a \emph{frame} $\partial \left( [0,1]^{2} \right)$, a finite, possibly empty,
672: collection $\mathcal{T}$ of \emph{transistors}, each homeomorphic to $[0,1]^{2}$, and a finite, non-empty collection $\mathcal{W}$
673: of \emph{wires}, each homeomorphic to $[0,1]$.  The frame and transistors all have well-defined top, bottom, left, and right sides,
674: which are the open sides parallel to the coordinate axes, and do not include the corners.  A wire has well-defined initial and terminal
675: points (i.e., $0$ and $1$, respectively).  A \emph{picture over $\mathcal{P}$}, denoted $\Delta$, is a quotient of
676: $$\partial \left( [0,1]^2 \right) \coprod \left( \coprod_{T \in \mathcal{T}} T \right) \coprod 
677: \left( \coprod_{w \in \mathcal{W}} w \right)$$
678: (for a choice of sets $\mathcal{T}$ and $\mathcal{W}$) by an equivalence relation $\sim$, together with a labelling function
679: $\ell : \mathcal{W} \rightarrow \Sigma$, satisfying:
680: \begin{enumerate}
681: \item The initial point of any given wire is attached either to the bottom of a transistor, or to the top of the frame.
682: The terminal point of any given wire is attached either  to the bottom of the frame, or to the top of some transistor.
683: If $w$ is a wire and $T$ is a transistor, then $w \cap T \subseteq \Delta$  is either empty or a singleton set.
684: \item Let $T_1$ and $T_2$ be transistors.  Write $T_1 < T_2$ if there is some wire $w$ such that the initial point of $w$
685: is attached to the bottom of $T_1$ and the terminal point of $w$ is attached to the top of $T_2$.  Let $<$ also denote the
686: transitive closure of the above relation.  The relation $<$ is required to be a strict partial order.
687: \item The equivalence classes of $\sim$ are either singleton sets or consist of exactly two points, exactly one of which is an endpoint
688: of a wire.  In other words, the only identifications in 
689: $$ \partial \left( [0,1]^{2} \right) \coprod \left( \coprod_{T \in \mathcal{T}} T \right) \coprod 
690: \left( \coprod_{w \in \mathcal{W}} w \right)$$
691: are generated by the attaching maps of the wires, and no two wires have points in common.  The endpoints of the wires
692: are called \emph{contacts}.
693: \item Suppose the top of the transistor $T$ meets the wires $w_{i_1}$, $w_{i_2}$, $\ldots$, $w_{i_m}$, reading from
694: the left side of $T$ to the right.  Suppose that the bottom of the transistor $T$ meets the wires $w_{j_1}$, $\ldots$,
695: $w_{j_n}$, again reading from left to right.  The \emph{top label of $T$}, denoted $L_{T}$, is
696: $$ \ell \left( w_{i_1} \right) \ell \left( w_{i_2} \right) \ldots \ell \left( w_{i_m} \right);$$
697: the \emph{bottom label of $T$}, denoted $L_{B}$, is
698: $$ \ell \left( w_{j_1} \right) \ell \left( w_{j_2} \right) \ldots \ell \left( w_{j_n} \right).$$
699: We require that $( L_{T} , L_{B} ) \in \mathcal{R}$ or $( L_{B}, L_{T}) \in \mathcal{R}$. 
700: \end{enumerate}
701: We can define the top and bottom labels of the frame just as we did for a transistor $T$.  If
702: the top label of the frame is $w_1$ and the bottom label is $w_2$, then $\Delta$ is a
703: \emph{$(w_1 , w_2)$-picture over $\mathcal{P}$}.  We say that $\Delta$ is a $(w, \ast )$-picture if
704: the top label of $\Delta$ is $w$, and the bottom label is arbitrary.
705: 
706: Two pictures $\Delta_{1}$ and $\Delta_{2}$ are \emph{isomorphic}, $\Delta_{1} \equiv \Delta_{2}$, if there is a homeomorphism
707: between them which matches labels and preserves the top-bottom- and left-right-orientations on the frame and transistors.
708: 
709: Given a $(u,v)$-picture $\Delta_1$ and a $(v,w)$-picture $\Delta_2$, one can define the \emph{concatenation}
710: $\Delta_{1} \circ \Delta_{2}$, which is the $(u,w)$-picture obtained by identifying the bottom of the frame
711: for $\Delta_1$ with the top of the frame for $\Delta_2$ by a homeomorphism which matches the endpoints of the wires, and
712: then removing the line segment corresponding to the bottom of $\Delta_1$ in the quotient, while keeping the wires passing
713: through this line segment intact.
714: 
715: \begin{figure}[!h] 
716: \begin{center}
717: \includegraphics{figure1.eps}
718: \end{center}
719: \caption{On the left, we have two pictures $\Delta_{1}$ and $\Delta_{2}$ (reading from top to bottom); on the right
720: we have the concatenation $\Delta_{1} \circ \Delta_{2}$.}
721: \label{concat}
722: \end{figure}
723: 
724: Figure \ref{concat} illustrates the operation of concatenation in a particular case.  All of the semigroup pictures
725: in the figure are pictures over the presentation $\langle x \mid x = x^{2} \rangle$.  For this reason, we leave off the labels
726: of the wires, since the label of each one is $x$.  Note that on the top left is an $(x, x^{3})$-picture, and 
727: on the bottom left is an $(x^{3}, x)$-picture.  If we denote these pictures $\Delta_{1}$ and $\Delta_{2}$, respectively, then
728: the $(x,x)$-picture on the right is $\Delta_{1} \circ \Delta_{2}$.   
729: 
730: 
731: Note that Figure \ref{concat} also illustrates our conventions for drawing pictures in the plane.  If $\Delta$ is a 
732: semigroup picture, then a function $\rho: 
733: \Delta \rightarrow \mathbb{R}^{2}$ is a \emph{projection} of $\Delta$ if:
734: \begin{enumerate}
735: \item the image of each transistor is a rectangle whose sides are parallel to the coordinate axes.  The map $\rho$
736: takes the top, left, right, and bottom of any given transistor to the corresponding sides in the image.
737: \item the image of the frame is an empty rectangle, and the map $\rho$ is again orientation-preserving, in the sense of (1).
738: The image of $\rho$ is contained inside the image of the frame. 
739: \item the image of each wire meets any given horizontal line at most once, and
740: \item $\rho$ is an embedding, except possibly at finitely many double points.  The inverse image of any double point $x$ is
741: a set of two points on distinct wires $w_1$ and $w_2$.  We assume that the images of $w_1$ and $w_2$ are transverse at $x$.
742: \end{enumerate}
743: It is rather clear that all of the defining features of a semigroup picture can be recovered from any of its 
744: suitably labelled projections.  From
745: now on, we will usually confuse a picture with any of its projections without further comment. 
746: 
747: 
748: Two transistors $T_1 < T_2$ form a \emph{dipole} if the top label of $T_1$ is identical (as a word in $\Sigma^{+}$) to the 
749: bottom label of $T_2$, and the bottom contacts of $T_1$ are paired off by wires in order with the top contacts of $T_2$.  To
750: \emph{remove a dipole}, delete the transistors $T_1$ and $T_2$ and all wires connecting them, and then glue together in order
751: the wires that formed top contacts of $T_1$ with those that formed bottom contacts of $T_2$.  The inverse operation is
752: called \emph{inserting a dipole}.  Two pictures are \emph{equal modulo dipoles}, $\Delta_1 = \Delta_2$, if one can be obtained
753: from the other by repeatedly inserting and removing dipoles.  A picture is called \emph{reduced} if it contains no dipoles.  Any
754: equivalence class modulo dipoles contains a unique reduced picture \cite{Me2, GS}.
755: 
756: In Figure \ref{dipole} we have two $(acbd, abab)$-pictures over the presentation $\mathcal{P} = \langle a, b, c, d \mid
757: ab=cd, cb=bc, ab=ba \rangle$.  In the left picture, we've circled two transistors which form a dipole.  If we remove this
758: dipole, we arrive at the picture on the right.  Notice that the two right-most transistors in the right half of the Figure do
759: not form a dipole:  the top label of the top transistor is $cd$, but the bottom label of the bottom transistor is $ba$. 
760: 
761: \begin{figure}[!h] 
762: \begin{center}
763: \includegraphics{figure2.eps}
764: \end{center}
765: \caption{The circled transistors in the left half of the figure form a dipole; on the right, we have the
766: result of removing this dipole.}
767: \label{dipole}
768: \end{figure}
769: 
770: For a fixed word $w \in \Sigma^{+}$, the set of all $(w,w)$-pictures over $\mathcal{P}$, modulo dipoles, forms a group
771: $\dbpw$ under the operation of concatenation.  We will follow \cite{GS} and call $\dbpw$
772: the \emph{braided diagram group} over $\mathcal{P}$, based at $w$.  (Warning:  the word ``braided" is rather unfortunate.  In fact, 
773: as the above definition shows, we don't care about any possible braiding of the wires, since equivalence between pictures doesn't
774: depend on any embedding into an ambient space.  Moreover, there is now a growing literature 
775: (see for example \cite{Brin})
776: on a braided version of Thompson's group $V$, which is something quite different from the older group $V$ which we consider here.
777: Nevertheless, there seems to be no better term.)  
778: A picture $\Delta$ is \emph{planar} if there is a projection 
779: $\rho : \Delta \rightarrow \mathbb{R}^{2}$ which is also an embedding.  
780: The set of all planar $(w,w)$-pictures over $\mathcal{P}$, modulo dipoles, forms a group
781: $\dpw$, which we will call the \emph{diagram group} over $\mathcal{P}$, based at $w$.  Annular pictures can be defined as follows.
782: Suppose that $\Delta$ is a picture, and let $\Delta'$ be the space obtained from $\Delta$ by removing the sides of the frame.  We say that
783: $\Delta$ is \emph{annular} if there is an orientation-preserving immersion of $\Delta'$ into
784: $A = \{ (x,y) \in \mathbb{R}^{2} \mid 1 \leq x^2 + y^2 \leq 4 \}$ such that:
785: i) the top of the frame for $\Delta'$ is wrapped around the circle $x^2 + y^2 =1$ once in the counterclockwise direction.  The initial
786: and terminal points of the top are both mapped to $(1,0) \in A$;
787: ii) the bottom of the frame for $\Delta'$ is wrapped around the circle $x^2 + y^2 =4$ once in the counterclockwise direction.  The initial
788: and terminal points of the bottom are both mapped to $(2,0) \in A$;
789: iii) the only double points of $\rho$ are $(1,0)$ and $(2,0)$; $\rho$ is an embedding otherwise.
790: The set of all annular pictures over $\mathcal{P}$ is a group $\dapw$, called the \emph{annular diagram group} over 
791: $\mathcal{P}$, based at $w$. 
792: 
793: Three groups are of special interest to us.  Let $\mathcal{P} = \langle x \mid x = x^{2} \rangle$.  
794: The groups $\di{}{\mathcal{P}}{x}$, $\di{a}{\mathcal{P}}{x}$, and $\di{b}{\mathcal{P}}{x}$ are, respectively,
795: Thompson's groups $F$, $T$, and $V$.  The original observation that $\di{}{\mathcal{P}}{x} \cong F$ was due to Victor Guba;
796: Guba and Sapir (in \cite{GS}) 
797: sketched the theory of annular and braided diagram groups expressly for the purpose of bringing their techniques
798: to bear on the study of $T$ and $V$.  Section 6 of \cite{Me2} describes an isomorphism between the groups $F$, $T$, and $V$, and
799: the corresponding diagram groups.  
800: 
801: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
802: 
803: \subsection{Diagram Complexes}
804: 
805: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
806: 
807: If $G$ is a diagram group of the standard, annular, or braided variety, then
808: a theorem of \cite{Me2} (see also \cite{Me1}) says that $G$ acts properly by isometries on a CAT(0) cubical
809: complex.  We briefly describe the construction of the cubical complex in this
810: subsection.
811: 
812: Fix a braided diagram group $\dbpw$.  We define a complex $\kbpw$, called the
813: \emph{diagram complex for $\dbpw$}, as follows.  A vertex $v \in \kbpw^{0}$ is an
814: equivalence class $\sim$ of reduced braided $(w, \ast)$-pictures, where $\Delta_{1} \sim
815: \Delta_{2}$ if and only if there is some braided permutation picture $\Psi$, such that
816: $\Delta_{1} \circ \Psi = \Delta_{2}$.  Here a \emph{permutation picture} is one 
817: with no transistors.  It is convenient to depict a vertex as a $(w,\ast)$-picture
818: in which all wires 
819: which would ordinarily be connected to the bottom of the frame have been
820: cut, as in Figure \ref{vertexpic}.  
821: 
822: \begin{figure} [!h] 
823: \begin{center}
824: \includegraphics{num1.eps}
825: \end{center}
826: \caption{This is a vertex in the cubical complex $\kt{b}{\mathcal{P}}{x}$, where
827: $\mathcal{P} = \langle x \mid x = x^{2} \rangle$.}
828: \label{vertexpic}
829: \end{figure}
830: 
831: An $n$-dimensional cube in $\kbpw$ is denoted by a reduced braided $(w,\ast)$-picture $\Delta$ in which
832: all of the bottom wires have been cut (as above), and $n$ of the maximal
833: transistors of $\Delta$ have been
834: drawn as white.  The picture Figure \ref{cubepic}a) denotes a $2$-cube, for example.
835: 
836: \begin{figure} [!h] 
837: \begin{center}
838: \includegraphics{num2.eps}
839: \end{center}
840: \caption{a) This notation describes a cube in the complex $\kt{b}{\mathcal{P}}{x}$; b) This is the labelled
841: cube in $\kt{b}{\mathcal{P}}{x}$ denoted by the picture in a).}
842: \label{cubepic}
843: \end{figure}
844: 
845: If we arbitrarily number the white transistors $1, 2, \ldots, n$, then there is a natural
846: way to label the vertices of an $n$-cube $[0,1]^{n}$, corresponding to this numbering of
847: $\Delta$.  Namely, if $(a_1 , \ldots , a_n ) \in \{ 0, 1 \}^{n}$ label $(a_1, \ldots , a_n )$
848: by the picture $\Delta_{(a_1, \ldots , a_n )}$, where the $i$th transistor is left off if $i=0$
849: and the $i$th transistor is filled in if $i=1$.  For instance, Figure \ref{cubepic}b) shows how to label the
850: corners of $[0,1]^{2}$ if $\Delta$ is as in Figure \ref{cubepic}a) and the white transistors are numbered
851: from left to right.
852: 
853: If we let $\Delta$ vary over all possible isomorphism classes of cube representatives ( where isomorphisms
854: send white transistors to white transistors), and, for each $\Delta$, choose as above a labelling
855: of $[0,1]^{n}$ for the appropriate $n$, then $\kbpw$ is the quotient of the resulting labelled cubes by the
856: equivalence relation which identifies the cubes along faces with the same labels.  It is proved in \cite{Me2} that
857: $\kbpw$ is a proper CAT(0) cubical complex if $\mathcal{P}$ is a finite presentation, and that
858: $\dbpw$ acts properly and cellularly on $\kbpw$.  The action is usually not cocompact, and, in particular,
859: isn't for any 
860: of the groups $F$, $T$, and $V$. 
861: 
862: We note that entirely similar statements are true for ordinary and annular diagram groups.  It is only
863: necessary to replace pictures with planar pictures and annular pictures (respectively) in the above
864: discussion to get the descriptions of $\kpw$ and $\kapw$, respectively.
865: 
866: Lastly, we recall a useful partial order on vertices. 
867: If $\left[ \Delta_{1} \right]$ and $\left[ \Delta_{2} \right] $ are vertices in a diagram complex, we write
868: $\left[ \Delta_{1} \right] \leq \left[ \Delta_{2} \right]$ if there exists some picture $\theta$ such that
869: $\Delta_{1} \circ \theta \equiv \Delta_{2}$.  Note that this means $\Delta_{1} \circ \theta$ and $\Delta_{2}$
870: are isomorphic before reducing dipoles.  It is not difficult to see that $\leq$ is a well-defined partial order.
871: 
872: Suppose $\mathcal{T}' \subseteq \mathcal{T}_{\Delta}$, where $\mathcal{T}_{\Delta}$ is the collection of transistors
873: in a picture $\Delta$. We say that $\mathcal{T}'$ is an \emph{initial subset} of $\mathcal{T}_{\Delta}$ if whenever
874: $T_{1} < T_{2}$ and $T_{2} \in \mathcal{T}'$, then $T_{1} \in \mathcal{T}'$ also.   
875: We reproduce a lemma from \cite{Me2}.
876: 
877: \begin{lemma} \label{bccc} \cite{Me2}
878: Let $\Delta$ be a vertex, and let $\mathcal{T}_{\Delta}$ be its set of transistors.  
879: There is a one-to-one correspondence $\psi$
880: between initial subsets of $\mathcal{T}_{\Delta}$ and vertices $\Delta_{1}$ satisfying
881: $\Delta_{1} \leq \Delta$.  The function $\psi$ is order-preserving and has an order-preserving 
882: inverse, i.e., the initial subsets $\mathcal{T}', \mathcal{T}''$ satisfy $\mathcal{T}' \subseteq \mathcal{T}''$
883: if and only if $\psi \left( \mathcal{T}' \right) \leq
884:  \psi \left( \mathcal{T}'' \right)$.
885: \qed
886: \end{lemma}
887: 
888: The map $\psi$ in the above lemma is easy to define:  if $\mathcal{T}'$ is an initial subset of transistors, then
889: $\psi \left( \mathcal{T}' \right)$ is obtained by removing all transistors in $\mathcal{T}_{\Delta} - \mathcal{T}'$,
890: along with all of their bottom wires.  The result is easily seen to be a vertex.  The argument that the map $\psi$ is 
891: injective can be extended to prove that the automorphism group of a diagram is trivial, 
892: at least combinatorially speaking.  That is,
893: if $\phi : \Delta \rightarrow \Delta$ is a isomorphism, then $\phi$ leaves the frame, each transistor, and each wire invariant,
894: and restricts to a self-homeomorphism of each of these.  It therefore follows, for instance, that in a concatenation
895: $\Delta_{1} \circ \Delta_{2}$ of pictures, one can speak of the transistors that were contributed by $\Delta_{i}$ for $i=1,2$, and
896: this is a well-defined notion even after reducing dipoles.  We shall need this observation in future sections, and use it without
897: further comment.
898: 
899: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
900: 
901: \section{Geodesic Profiles in Diagram Complexes} \label{pict}
902: 
903: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
904: 
905: We now describe profiles in diagram complexes.  Our main goals here are, first,
906: to describe a profile as an infinite picture of a certain kind, and then to describe
907: the action on profiles in terms of picture multiplication.
908: 
909: Throughout this section, we use only the complex $\kbpw$, but the discussion carries over
910: to $\kpw$ and $\kapw$ in an obvious way.
911: 
912: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
913: \subsection{Description of Profiles}
914: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
915: 
916: The first step is to describe hyperplanes in $\kbpw$.  Recall that a combinatorial hyperplane
917: is an equivalence class of $1$-cells under the relation $\sim$ of square equivalence.  The square
918: equivalence relation is generated by simple square equivalence (also denoted $\sim$), 
919: where two $1$-cells $e_1$, $e_2$ are simple square equivalent  
920:  if they are opposite faces of a $2$-cell (square).
921: 
922: \begin{figure} [!h] 
923: \begin{center}
924: \includegraphics{numa.eps}
925: \end{center}
926: \caption{a) An edge $\Delta$ in $\kt{b}{\mathcal{P}}{x}$, and b) a collection of
927: edges that are square equivalent to $\Delta$.} 
928: \label{hyperplane}
929: \end{figure}
930: 
931: We describe combinatorial hyperplanes in $\kbpw$ with help from an example.  First, fix a $1$-cell
932: in $\kbpw$, such as the one in Figure \ref{hyperplane}a), which we'll call $\Delta$.
933: 
934: 
935: We consider a small number of $1$-cells that are square equivalent to $\Delta$ (there are infinitely many such
936: $1$-cells for this $\Delta$).  These are the vertical edges in Figure \ref{hyperplane}b).  We denote these edges
937: $\Delta_1$, $\Delta_2 = \Delta$, $\Delta_3$, and $\Delta_4$, reading from left to right.  Note the interpretation
938: of simple square equivalence in terms of diagrams:  for $i=1,2,3$, $\Delta_{i} \sim \Delta_{i+1}$ since $\Delta_{i}$
939: can be obtained from $\Delta_{i+1}$ by removing a maximal shaded transistor and all of its bottom wires
940: from $\Delta_{i+1}$, or the reverse, i.e., $\Delta_{i+1}$ can be obtained in the same way from $\Delta_{i}$.
941: This observation is general, and holds true in all of the complexes $\kpw$, $\kapw$, and $\kbpw$, for all $\mathcal{P}$
942: and $w$, and indeed follows easily from the definition of the $2$-cells in a diagram complex.  We record this 
943: in a lemma.
944: 
945: \begin{lemma}
946: Let $\Delta'$, $\Delta''$ be $1$-cells in $\kbpw$.  The following statements are equivalent:
947: \begin{enumerate}
948: \item $\Delta' $ and $\Delta''$ are simple square equivalent; 
949: \item There is some maximal shaded transistor $T$ in $\Delta'$ such that $\Delta''$ is
950: the result of removing $T$ and all of its bottom wires from $\Delta'$ (or the reverse statement
951: is true, with $\Delta'$ and $\Delta''$ reversing roles).
952: \qed
953: \end{enumerate}
954: \end{lemma}
955: 
956: Fix a $1$-cell $\Delta \subseteq \kbpw$.  Let $H_{\Delta}$ denote the combinatorial hyperplane corresponding to
957: $\Delta$.  Let $\mathcal{T}_{\Delta}$ denote the collection of transistors  of $\Delta$.
958:  Let $T$ denote the (unique) white transistor in $\mathcal{T}_{\Delta}$.  Consider the collection
959: $M_{\Delta} = \{   T' \mid   T' \mathrm{~is~ a~ transistor~in~} \Delta; \, \, T' \leq T \}$
960: (the inequality sign refers to the partial order on transistors).  We can associate to this collection of transistors
961: a vertex $min \left( H_{\Delta} \right)$, called the \emph{minimal vertex of $H_{\Delta}$}.  Simply remove all transistors
962: in $\mathcal{T}_{\Delta} - M_{\Delta}$ along with their bottom wires, and then shade the white transistor.  The result is
963: necessarily a vertex by Lemma \ref{bccc}.  It is clear that $min \left( H_{\Delta} \right)$ depends only on the hyperplane
964: $H_{\Delta}$.
965: 
966: For example, if $\Delta = \Delta_1 , \Delta_2 , \Delta_3,  ~ \mathrm{or} ~ \Delta_4$ from Figure \ref{hyperplane}b), then
967: $min \left( H_{\Delta} \right)$ is the vertex at the top of $\Delta_2$.  We note one property of minimal vertices:  a vertex
968: is minimal if and only if it contains a unique maximal transistor.  If a vertex $\Delta$ 
969: has a unique maximal transistor, 
970: then the hyperplane $H_{\Delta}$ corresponding to $\Delta$ is the square equivalence class of the edge obtained by painting the
971: maximal transistor white. 
972: 
973: We are interested in $min \left( H_{\Delta} \right)$ because of the following lemma.  In all that follows, 
974: we let our basepoint
975: $\ast$ be the unique vertex in $\kbpw$ having no transistors. 
976: 
977: \begin{lemma} \label{sep}
978: Let $H$ be a hyperplane in $\kbpw$.  If $\overline{\Delta}$ is an arbitrary vertex
979: in $\kbpw$, then $\overline{\Delta}$ and the basepoint $\ast$ lie in different components of
980: $\kbpw - H$ if and only if $min(H) \leq \overline{\Delta}$.
981: \end{lemma}
982:  
983: \begin{proof}
984: Let $\overline{\Delta}$ be a vertex in $\kbpw$; let $\mathcal{T}_{\overline{\Delta}}$ denote the collection 
985: of transistors in $\overline{\Delta}$.  Choose a function $\alpha : \mathcal{T}_{\overline{\Delta}}   
986: \rightarrow \left\{ 1, \ldots , \left| \mathcal{T}_{\overline{\Delta}} \right| \right\}$ satisfying:
987: \begin{enumerate}
988: \item $\alpha$ is one-to-one;
989: \item if $T_1 < T_2$, then $\alpha \left( T_1 \right) < \alpha \left( T_2 \right)$.
990: \end{enumerate}
991: We associate a sequence of vertices
992: $\ast = \overline{\Delta}_{0}, \overline{\Delta}_{1}, \ldots, 
993: \overline{\Delta}_{\mid \mathcal{T}_{\overline{\Delta}} \mid } = \overline{\Delta}$, where
994: $\overline{\Delta}_{i}$ is the (unique) vertex determined by $\alpha^{-1} \left( \{ 1, 2, \ldots, i \} \right)$
995: under the correspondence in Lemma \ref{bccc}.  It is not difficult to see that $\overline{\Delta}_{i}$ is 
996: connected to $\overline{\Delta}_{i+1}$
997: by a unique edge for $i=0, 1, \ldots, \left| \mathcal{T}\left( \overline{\Delta} \right) \right| - 1 $.  We let
998: $p_{\alpha}$ denote the edge-path consisting of these edges.
999: 
1000: \begin{figure}[!h] 
1001: \begin{center}
1002: \includegraphics{numb.eps}
1003: \end{center}
1004: \caption{a) A labelling $\alpha$ of the 
1005: picture $\overline{\Delta}$, and b) the associated edge-path in $\kt{b}{\mathcal{P}}{x}$.} 
1006: \label{edgepath}
1007: \end{figure}
1008: 
1009: For example, Figure \ref{edgepath}a) shows a picture with a numbering $\alpha$ of its transistors, along with the corresponding 
1010: edge-path ( Figure \ref{edgepath}b) ).
1011: 
1012: 
1013: We claim that $p_{\alpha}$ is a geodesic in the $1$-skeleton $\kbpw^{1}$.
1014: Suppose that $p$ is an arbitrary edge-path connecting $\ast$ to $\overline{\Delta}$; let
1015: $\ast = \overline{\Delta}_{0}', \overline{\Delta}_{1}', \ldots, \overline{\Delta}_{m}' = \overline{\Delta}$
1016: be the vertices lying along the path $p$, listed in the order they are visited.  It is clear from the definition
1017: of edges in $\kbpw$ that $\overline{\Delta}_{i+1}'$ is obtained from $\overline{\Delta}_{i}'$ $( 0 \leq i \leq m-1 )$
1018: by either removing a maximal transistor from the bottom of $\overline{\Delta}_{i}'$, or adding a new maximal transistor
1019: to $\overline{\Delta}_{i}'$.  It immediately follows from this that 
1020: $\ell(p) \geq \left| \mathcal{T}_{\overline{\Delta}} \right|$.  This proves the claim.
1021: 
1022: Suppose that $min \left( H_{\Delta} \right) \leq \overline{\Delta}$.  Lemma \ref{bccc} implies that
1023: $min \left( H_{\Delta} \right)$ corresponds to an initial collection $\mathcal{T}$ of transistors.
1024: Suppose that $\left| \mathcal{T} \right| = n$.  It follows that we can define 
1025: $\alpha : \mathcal{T}_{\overline{\Delta}} \rightarrow \left\{ 1, \ldots, n, \ldots, 
1026: \left| \mathcal{T}_{\overline{\Delta}} \right|  \right\}$ in such a way that
1027: $\alpha_{\mid_{\mathcal{T}}} : \mathcal{T} \rightarrow \{ 1, \ldots, n \}$ is another labelling function satisfying
1028: (1) and (2) above.  In this case, $\overline{\Delta}_{n} = min\left( H_{\Delta} \right)$ and 
1029: $\overline{\Delta}_{n-1}$ is the vertex obtained by removing the (unique) maximal transistor in
1030: $\overline{\Delta}_{n}$.  It immediately follows that the edge $\left[ \overline{\Delta}_{n-1} , \overline{\Delta}_{n} \right]$
1031: is a member of the combinatorial hyperplane $H_{\Delta}$.  Since a geodesic edge-path $p_{\alpha}$ from $\ast$ to $\overline{\Delta}$
1032: crosses $H_{\Delta}$, Theorem \ref{Sageev}(3) implies 
1033: that $\ast$ and $\overline{\Delta}$ lie on opposite sides of $H_{\Delta}$, proving one direction.
1034: 
1035: Conversely, suppose that $\overline{\Delta}$ and $\ast$ are separated by the hyperplane $H_{\Delta}$.  
1036: Theorem \ref{Sageev}(3) says 
1037: that a geodesic edge-path $p$ crosses precisely the hyperplanes separating
1038: the initial vertex of $p$ from the terminal vertex of $p$.  It follows that, for some
1039: $k \in \left\{ 0, \ldots, \left| \mathcal{T}_{\overline{\Delta}} \right| - 1 \right\}$, the edge
1040: $\left[ \overline{\Delta}_{k} , \overline{\Delta}_{k+1} \right]$ represents the hyperplane $H_{\Delta}$.
1041: Under the correspondence in Lemma \ref{bccc}, $\overline{\Delta}_{k}$ corresponds to a collection
1042: $\mathcal{T}_{\overline{\Delta}_{k+1}} - \{ T \}$ of transistors in $\overline{\Delta}_{k+1}$, 
1043: where $T \in \mathcal{T}_{ \overline{\Delta}_{k+1}}$.  The edge
1044: $\left[ \overline{\Delta}_{k} , \overline{\Delta}_{k+1} \right]$ can be described in terms of pictures as follows:
1045: draw $\overline{\Delta}_{k+1}$, but leave the transistor $T$ unshaded.  According to the definition, we obtain
1046: $min \left( H_{\Delta} \right)$ by shading $T$, and then taking the picture corresponding to
1047: $\left\{ T' \in \mathcal{T}_{\overline{\Delta}_{k+1}} \mid T' \leq T \right\}$.  It is thus clear that
1048: $min \left( H_{\Delta} \right) \leq \overline{\Delta}_{k+1} \leq \overline{\Delta}$.
1049: \end{proof}  
1050: 
1051: \begin{proposition} \label{ord}
1052: Let $H_{1}^{+}$ and $H_{2}^{+}$ be two positive half-spaces in $\kbpw$.  Let $\Delta_{1}$ and
1053: $\Delta_{2}$ be their minimal vertices.
1054: \begin{enumerate}
1055: \item $H_{2}^{+} \leq H_{1}^{+}$ if and only if $\Delta_{2} \leq \Delta_{1}$.
1056: \item $H_{1}^{+} \cap H_{2}^{+} \neq \emptyset$ if and only if $\{ \Delta_{1} , \Delta_{2} \}$
1057: has an upper bound in $\kbpw^{0}$.
1058: \end{enumerate}
1059: \end{proposition}
1060: 
1061: \begin{proof}
1062: (1) $( \Rightarrow )$ Suppose $H_{2}^{+} \leq H_{1}^{+}$.  This means that $H_{1}^{+} \subseteq H_{2}^{+}$.  Thus,
1063: every vertex $\Delta \in H_{1}^{+}$ is separated from $\ast$ by $H_{2}$.  In particular, $\Delta_{1}$ is so separated
1064: from $\ast$.  By the previous lemma, $\Delta_{2} \leq \Delta_{1}$.
1065: 
1066: $( \Leftarrow )$ Suppose $\Delta_{2} \leq \Delta_{1}$.  It is sufficient to check the inclusion
1067: $H_{1}^{+} \subseteq H_{2}^{+}$ on vertices.  If $\Delta$ is a vertex in $H_{1}^{+}$, then
1068: $\Delta_{1} \leq \Delta$.  It follows that $\Delta_{2} \leq \Delta$, so $\Delta \in H_{2}^{+}$ by the previous lemma.
1069: 
1070: (2) Both directions are immediate consequences of the previous lemma.
1071: \end{proof}
1072: 
1073: We now obtain the desired characterization of profiles in terms of pictures.
1074: 
1075: \begin{theorem} \label{profile}
1076: Let the basepoint $\ast \in \kt{b}{\mathcal{P}}{w}$ be the unique vertex having no transistors. 
1077: Let $\Delta$ be an infinite $(w,\ast)$-picture over the semigroup presentation $\mathcal{P}$, 
1078: i.e.,  a picture in the sense of Section \ref{diagram}, except that the transistor and wire sets
1079: are countably infinite.  Let us suppose as well that $\Delta$ satisfies the following conditions:
1080: \begin{enumerate}
1081: \item For any transistor 
1082: $T \in \mathcal{T}_{\Delta}$, the set $( -\infty, T] = \{ T' \in \mathcal{T}_{\Delta} \mid
1083: T' \leq T \}$ is finite;
1084: \item There are no maximal elements in the set $\mathcal{T}_{\Delta}$ of transistors in $\Delta$; 
1085: \item No wire is attached to the bottom of the frame of $\Delta$.
1086: \end{enumerate}
1087: The picture $\Delta$ determines a unique profile, i.e., a 
1088: non-empty collection of positive half-spaces $\mathcal{H}_{\Delta}$ in $\kbpw$  
1089: satisfying       
1090: properties (1)-(3) in Proposition \ref{fund}.  Conversely, a profile in $\kbpw$ determines a unique
1091: infinite picture $\Delta$ satisfying properties (1)-(3) above.
1092: 
1093: The indicated correspondences are mutually inverse.
1094: \end{theorem}
1095: 
1096: \begin{proof}
1097: Suppose that $\Delta$ is an infinite $(w, \ast)$-picture over the semigroup presentation $\mathcal{P}$ satisfying
1098: the properties above.  The transistors of $\Delta$ are in one-to-one correspondence with a collection of hyperplanes in
1099: the following way.  Let $T$ be a transistor of $\Delta$; we consider the collection $(- \infty, T]$ of all transistors in
1100: $\Delta$ which are less than or equal to $T$ in the partial order on transistors.  By Lemma \ref{bccc} and 
1101: our assumption that $(- \infty, T]$ is finite, 
1102: this collection of transistors
1103: corresponds to a unique vertex, and this vertex is the minimal vertex of a unique hyperplane $H_{T}$.  Note that Lemma \ref{bccc}
1104: also implies that the correspondence between transistors and hyperplanes is one-to-one.
1105: 
1106: We consider the properties of the collection $\mathcal{H}_{\Delta} = \left\{ H_{T} \mid T \in \mathcal{T}_{\Delta} \right\}$.  First, let
1107: $H_{T_{1}}$, $\ldots$, $H_{T_{n}}$ be hyperplanes in $\mathcal{H}_{\Delta}$.  Consider the collection 
1108: $\mathcal{T}' = \left\{ T \in \mathcal{T}_{\Delta} \mid T \leq T_{i} \mathrm{~for~some~} i \in \{ 1, \ldots, n \} \right\}$.  By 
1109: Lemma \ref{bccc}, the collection $\mathcal{T}'$ corresponds to a vertex $\Delta_{\mathcal{T}'}$.  Moreover, we have
1110: that $( -\infty, T_{i}] \subseteq \mathcal{T}'$, for $i=1, \ldots, n$, from which it follows that 
1111: $min \left( H_{T_{i}} \right) \leq \Delta_{\mathcal{T}'}$, for $i=1, \ldots, n$.  This, in turn, implies that
1112: $\Delta_{\mathcal{T}'} \in H_{T_{1}}^{+} \cap \ldots \cap H_{T_{n}}^{+}$, by Lemma \ref{sep}.  Thus property (1)
1113: from Proposition \ref{fund} holds.
1114: 
1115: If $H_{T} \in \mathcal{H}_{\Delta}$, then, by the assumption that $T$ is not maximal, there is some 
1116: $T_{1} \in \mathcal{T}_{\Delta}$ such that
1117: $T < T_{1}$.  It follows from this that $(-\infty, T] \subsetneq ( -\infty, T_{1}]$; this implies that the
1118: vertices $min \left( H_{T} \right)$, $min \left( H_{T_1} \right)$ under the correspondence from Lemma \ref{bccc}
1119: satisfy $min \left( H_{T} \right) < min \left( H_{T_1} \right)$.  By Lemma \ref{ord}, $H_{T}^{+} < H_{T_1}^{+}$, 
1120: so property (2) from Proposition \ref{fund} holds. 
1121:    
1122: Checking Property (3) from Proposition \ref{fund} is an easy exercise using the properties of the correspondence in Lemma \ref{bccc}.
1123: 
1124: Conversely, suppose that $\mathcal{H}$ is a non-empty collection of hyperplanes in $\kbpw$ satisfying properties (1)-(3) of Proposition \ref{fund}.
1125: Choose a finite collection of
1126: hyperplanes $H_{T_{1}}, H_{T_{2}}, \ldots, H_{T_{n}}$.
1127: By property (1) from 
1128: Proposition \ref{fund}, $H_{T_{1}}^{+} \cap H_{T_{2}}^{+} \cap \ldots \cap H_{T_{n}}^{+} \neq \emptyset$.  
1129: This implies that there is some vertex $\hat{\Delta}$ in the latter intersection, which means, by Lemma \ref{sep}, that 
1130: $min \left( H_{T_{i}} \right) \leq \hat{\Delta}$ for $i = 1, \ldots, n$.  Since the collection
1131: $\left\{ min \left( H_{T_{i}} \right) \mid i = 1, \ldots, n \right\}$ has an upper bound, Lemma 3.2(2) of \cite{Me2}
1132: implies that it has a least upper bound.  
1133: Thus, we've shown that any finite collection of minimal vertices for hyperplanes in
1134: $\mathcal{H}$ has a least upper bound.  (This least upper bound is a ``union" of the labels for these 
1135: vertices, in an appropriate sense.
1136: Note that it won't in general be a minimal vertex itself.) 
1137: 
1138: Let $\Delta_{1}, \Delta_{2}, \ldots, \Delta_{n}, \ldots $ be the sequence consisting of all minimal vertices for hyperplanes
1139: in $\mathcal{H}$.  Since any finite collection of these hyperplanes has a least upper bound, we can identify the direct  
1140: limit of this sequence with an infinite diagram $\Delta$.  It is clear that $\Delta$ has properties (1) and (3) from the statement of the Theorem. 
1141: Property (2) follows easily from the fact that the collection $\mathcal{H}$ satisfies (2) from Proposition \ref{fund}.
1142: 
1143: We leave the final statement as an exercise.
1144: \end{proof}      
1145: 
1146: We will sometimes require a lemma which gives a necessary condition
1147: on the open cube $C$ through which a geodesic ray $c$ with profile $P(c)$
1148: can travel.  The condition involves the largest vertex $\Delta_{C}$ in the
1149: closure $\overline{C}$, which always exists, and can be obtained by
1150: shading each transistor in the picture representative for $C$
1151: (see for instance Figure \ref{cubepic}, from Section \ref{diagram}).
1152: 
1153: \begin{lemma} \label{trav}
1154: Let $c: [0, \infty) \rightarrow \kbpw$ be a geodesic ray issuing from the base vertex
1155: $\ast$.  Let $\Delta$ be the infinite picture representing the profile of $c$.  Let $C$
1156: be an open cube of $\kbpw$ satisfying $ \left( \mathrm{Im} \, c \right) \cap C \neq \emptyset$.
1157: 
1158: If $\Delta_{C}$ is the largest vertex in $\overline{C}$, then $\Delta_{C} \leq \Delta$.
1159: \end{lemma}
1160: 
1161: \begin{proof}
1162: Let $\Delta_{C}$ be the largest vertex of $C$.  Consider the collection of all maximal transistors
1163: $T_{1}, \ldots, T_{n}$ in $\Delta_{C}$; let $\Delta_{T_{i}}$ be the unique vertex determined by
1164: $( -\infty, T_{i}]$ under the correspondence from Lemma \ref{bccc}.  Note that each $\Delta_{T_{i}}$
1165: is the minimal vertex  of a hyperplane $H_{T_{i}}$, and all of the hyperplanes
1166: $H_{T_{1}}, \ldots, H_{T_{n}}$ are distinct.
1167: 
1168: Let $\widehat{\Delta}_{C}$ denote the representative for $C$, which is a picture consisting of shaded and
1169: unshaded transistors, as in Figure \ref{cubepic}.  Some of the transistors $T_i$ are shaded in $\widehat{\Delta}_{C}$;
1170: others are unshaded in $\widehat{\Delta}_{C}$.  
1171: 
1172: If $T_i$ is shaded, then $\Delta_{T_{i}} \leq \Delta_{C}$, so that $H_{T_{i}}$ separates $\Delta_{C}$ from $\ast$.
1173: Moreover, $H_{T_{i}}$ doesn't pass through $C$, since the hyperplanes $H$ satisfying $H \cap C \neq \emptyset$ are the precisely
1174: the collection of all $H_{T_{j}}$ such that $T_j$ is unshaded in $\widehat{\Delta}_{C}$.  It follows that each point $x$ in 
1175: $\left( \mathrm{Im} \, c \right) \cap C$ can be connected to $\Delta_{C}$ without crossing $H_{T_{i}}$; therefore $x \in H^{+}_{T_{i}}$,
1176: so $H^{+}_{T_{i}} \in P(c)$.
1177: 
1178: If $T_i$ is unshaded in $\widehat{\Delta}_{C}$, then $H_{T_{i}} \cap C \neq \emptyset$.  It follows from Lemma \ref{crossing}(2) that
1179: $H^{+}_{T_{i}} \in P(c)$.
1180: 
1181: The correspondence of Theorem \ref{profile} implies that $\Delta_{T_{1}}, \ldots, \Delta_{T_{n}} \leq \Delta$.  This means that the least
1182: upper bound $\widetilde{\Delta}$ of $\left\{ \Delta_{T_{1}}, \ldots, \Delta_{T_{n}} \right\}$ exists, and satisfies $\widetilde{\Delta} \leq \Delta$.
1183: But clearly $\widetilde{\Delta} = \Delta_{C}$, since the latter vertex is an upper bound of $\left\{ \Delta_{T_{1}}, \ldots, \Delta_{T_{n}} \right\}$,
1184: and any proper initial subset subset of $\mathcal{T}_{\Delta_{C}}$ would fail to contain at least one of the transistors $T_{1}, \ldots, T_{n}$.
1185: \end{proof}
1186: 
1187: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1188: \subsection{The Action on Profiles}
1189: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1190: 
1191: \begin{proposition} \label{action}
1192: There is a well-defined action of $\dbpw$ on the set of all profiles.  If $\Delta$ is
1193: a profile and $\Delta_{1} \in \dbpw$, then $\Delta_{1} \ast \Delta$ can be computed as
1194: follows.  First, form the concatenation $\Delta_{1} \circ \Delta$ and remove all dipoles.
1195: Second, remove all maximal transistors from the resulting infinite diagrem, until no maximal
1196: transistors remain.  The result is $\Delta_{1} \cdot \Delta$.
1197: \end{proposition}
1198:   
1199: \begin{proof}
1200: Let $\mathcal{H}$, $\mathcal{H}'$ be two collections of positive half-spaces in 
1201: $\kbpw$.  We write $\mathcal{H} \sim \mathcal{H}'$ and $[\mathcal{H}] = [\mathcal{H}']$
1202: if $\mathcal{H}$ and $\mathcal{H}'$ are \emph{cofinal}, i.e., if for any
1203: $H_{1}^{+} \in \mathcal{H}$, there exists
1204: $\left( H'_{1} \right)^{+} \in \mathcal{H}'$ such that $H_{1}^{+} \leq \left( H_{1}' \right)^{+}$ and
1205: for any $\left( H_{2}' \right)^{+} \in \mathcal{H}'$, there exists $H_{2}^{+} \in \mathcal{H}$
1206: such that $\left( H_{2}' \right)^{+} \leq H_{2}^{+}$.  It is fairly clear that $\sim$ is an
1207: equivalence relation on the set of all collections of positive half-spaces in $\kbpw$.
1208: 
1209: The group $\dbpw$ doesn't act in an obvious way on the set of equivalence classes, since a group
1210: element $\overline{\Delta} \in \dbpw$ doesn't necessarily map a positive half-space to a positive
1211: half-space.  Indeed, $\overline{\Delta} \cdot H^{+}$ is a negative half-space if and only if
1212: $\ast \in \overline{\Delta} \cdot H^{+}$, that is, if and only if $\overline{\Delta}^{-1} \cdot \ast \in H^{+}$.
1213: Now $\overline{\Delta}^{-1} \cdot \ast \in H^{+}$ if and only if $min \left( H^{+} \right) \leq \overline{\Delta}^{-1} \cdot \ast$.
1214: There are only finitely many vertices $\Delta'$ satisfying $\Delta' \leq \overline{\Delta}^{-1} \cdot \ast$ (all determined by
1215: the correspondence in Lemma \ref{bccc}).  It follows from this that a given $\overline{\Delta} \in \dbpw$ maps at most finitely 
1216: many positive half-spaces to negative ones.
1217: 
1218: We obtain an action on profiles in the following way.  
1219: Identify a profile $\Delta$ with the unique equivalence class $\left[ \mathcal{H} \right]$
1220: such that $\Delta \in \left[ \mathcal{H} \right]$.  For a given $\overline{\Delta} \in \dbpw$, choose a collection
1221: $\mathcal{H}' \in \left[ \mathcal{H} \right]$ such that $\overline{\Delta} \cdot H^{+}$ is a positive half-space,
1222: for each $H^{+} \in \mathcal{H'}$.  
1223: It is possible to do this since each $\mathcal{H}' \in \left[ \mathcal{H} \right]$ 
1224: is necessarily infinite.  We define
1225: $\overline{\Delta} \ast \Delta$ to be $\left[ \overline{\Delta} \cdot \mathcal{H}' \right]$.  It is not difficult to see
1226: that $\left[ \overline{\Delta} \cdot \mathcal{H}' \right]$ contains a 
1227: unique profile, and that the definition of $\ast$ doesn't depend
1228: upon the choice of $\mathcal{H}' \in \left[ \mathcal{H} \right]$.  It follows that $\ast$ is an action on profiles.
1229: 
1230: We claim two things:  first, that $\Delta \ast P(c) = P ( \Delta \ast c )$, for any geodesic ray
1231: $c: [0, \infty) \rightarrow \kbpw$ issuing from $\ast$; second, that the action $\ast$ from the previous paragraph
1232: has the description promised in the statement of the proposition.
1233: 
1234: Recall the definition of the action $\ast: \dbpw \times \partial \kbpw \rightarrow \partial \kbpw$ on the
1235: space at infinity.  If $c \in \partial \dbpw$, then, for any $\overline{\Delta} \in \dbpw$, 
1236: $\overline{\Delta} \cdot c$ is simply the translate of $c$ by the usual action of $\overline{\Delta}$ on
1237: $\kbpw$.  The ray $\overline{\Delta} \ast c$ is the unique ray issuing from $\ast$ and asymptotic to $\Delta \cdot c$,
1238: the existence of which is guaranteed by Proposition \ref{parallel}.
1239: 
1240: We now prove the first claim.  Let $c \in \partial \kbpw$ and let $\overline{\Delta} \in \dbpw$.  We choose
1241: some cofinal subset $\mathcal{H} \subseteq P(c)$ such that $\overline{\Delta} \cdot H^{+}$ is a positive half-space,
1242: for any $H^{+} \in \mathcal{H}$.  It is clear that $\overline{\Delta} \cdot c$ crosses each of the hyperplanes
1243: in $\overline{\Delta} \cdot \mathcal{H}$, since $c$ crosses each of the hyperplanes in $\mathcal{H}$.  Since
1244: $\overline{\Delta}$ maps positive half-spaces in $\mathcal{H}$ to positive half-spaces, $\overline{\Delta} \cdot c$
1245: intersects each $\overline{\Delta} \cdot H^{+} \in  \overline{\Delta} \cdot \mathcal{H}$ in an open ray, just as $c$
1246: intersects each $H^{+} \in \mathcal{H}$ in an open ray.  By Lemma \ref{monotone},
1247: $d_{\overline{\Delta} \cdot H} \left( \overline{\Delta} \cdot c(t) \right) \rightarrow \infty$ as 
1248: $t \rightarrow \infty$ for any $H$ such that $H^{+} \in \mathcal{H}$.  It follows from this, first, that
1249: $d_{\overline{\Delta} \cdot H} \left( \overline{\Delta} \ast c(t) \right)$ also goes to infinity as $t \rightarrow \infty$,
1250: and second, that $\left( \overline{\Delta} \ast c \right) (t) \in \overline{\Delta} \cdot H^{+}$ for $t$ sufficiently large.
1251: This implies that $\overline{\Delta} \cdot \mathcal{H} \subseteq P \left( \overline{\Delta} \ast c \right)$.
1252: 
1253: Next, we need to show that, for any positive half-space $H^{+} \in P \left( \overline{\Delta} \ast c \right)$, there
1254: is $\overline{H}^{+} \in \overline{\Delta} \cdot \mathcal{H}$ such that $H^{+} \leq \overline{H}^{+}$.  Choose a sequence
1255: of positive half-spaces $H^{+} = H^{+}_{0} < H^{+}_{1} < H^{+}_{2} < \ldots$ in $P \left( \overline{\Delta} \ast c \right)$.
1256: By Lemma \ref{monotone}, we know that $d_{H_{i}} \left( ( \overline{\Delta} \ast c ) (t) \right) \rightarrow \infty$ as
1257: $t \rightarrow \infty$, for any $i \in \{ 0, 1, \ldots, n, \ldots \}$.  By the definition of $P \left( \overline{\Delta} \ast c \right)$,
1258: for any $i$, $\left( \overline{\Delta} \ast c \right) (t) \in H^{+}_{i}$ for $t$ sufficiently large.  It now follows from the fact that
1259: $\overline{\Delta} \cdot c$ and $\overline{\Delta} \ast c$ are asymptotic that, for any $i$, $\left( \overline{\Delta} \cdot c \right)(t) \in H^{+}_{i}$
1260: for $t$ sufficiently large.  This implies that $\overline{\Delta} \cdot c$ crosses at least one of the $H_i$, and therefore all $H_j$ for $j$ sufficiently large,
1261: for otherwise
1262: $$\left( \overline{\Delta} \cdot c \right) \left( [0, \infty) \right) \subseteq \bigcap_{i=0}^{\infty} H^{+}_{i} = \emptyset.$$ 
1263: We choose some $H_{j}^{+}$ large enough that $\overline{\Delta}^{-1} \cdot H_{j}^{+}$ is a positive half-space.
1264: Since $c$ crosses $\overline{\Delta}^{-1} \cdot H_{j}$, it follows that $\overline{\Delta}^{-1} \cdot H_{j}^{+} \in P(c)$.
1265: Since $\mathcal{H}$ is cofinal in $P(c)$, there is $\hat{H}^{+} \in \mathcal{H}$ such that 
1266: $\overline{\Delta}^{-1} \cdot H_{j}^{+}
1267: \leq \hat{H}^{+}$.  This implies that $H_{j}^{+} \leq \overline{\Delta} \cdot \hat{H}^{+} \in \overline{\Delta} \cdot 
1268: \mathcal{H}$.  Now
1269: we've shown that $H^{+} \leq H_{j}^{+} \in \overline{\Delta} \cdot \mathcal{H}$.
1270: 
1271: It follows that $P \left( \overline{\Delta} \ast \mathcal{H} \right) = \left[ \overline{\Delta} \cdot \mathcal{H} \right]$ under the identification
1272: of $\left[ \overline{\Delta} \cdot \mathcal{H} \right]$ with a profile.  We've thus shown that $P \left( \overline{\Delta} \ast c \right) =
1273: \overline{\Delta} \ast P(c)$.  It immediately follows from this that the action of $\dbpw$ on profiles is well-defined:
1274: if $c, c'$ have the same profile, then so also do $\overline{\Delta} \ast c$ and $\overline{\Delta} \ast c'$.
1275: 
1276: Now we prove the second claim.  Let $\Delta$ be an infinite picture representing a profile and let
1277: $\overline{\Delta} \in \dbpw$.  We choose a cofinal collection of transistors $\mathcal{T}'$ in $\Delta$ (which are identified with positive half-spaces
1278: by the correspondence in Theorem \ref{profile}) such that no transistor in $\mathcal{T}'$ forms a dipole in the
1279: concatenation $\overline{\Delta} \circ \Delta$.  The above description of the action implies that
1280: $\overline{\Delta} \ast \Delta$ is the collection of all positive half-spaces such that
1281: $H^{+} \leq \overline{\Delta} \cdot H_{T'}^{+}$, for some $T' \in \mathcal{T}'$.  This collection of positive half-spaces
1282: may be identified with the collection of transistors $T$ in the reduced concatenation $\overline{\Delta} \circ \Delta$
1283: satisfying $T \leq T'$, for some $T' \in \mathcal{T}'$.  By the cofinality of $\mathcal{T}'$ in $\Delta$, these transistors $T$ are precisely those
1284: for which there exists an infinite sequence $T = T_0 < T_1 < \ldots < T_n < \ldots$ where each $T_i$ is a transistor of
1285: $\overline{\Delta} \circ \Delta$.  The second claim follows.
1286: \end{proof}
1287: 
1288: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1289: 
1290: \section{Fixed Profiles under the Actions of $F$, $T$, and $V$} \label{main}
1291: 
1292: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1293: 
1294: The results of the previous section largely
1295: reduce the problem of finding 
1296: fixed points in $\partial \kt{}{\mathcal{P}}{x}$,  $\partial \kt{a}{\mathcal{P}}{x}$, 
1297: and $\partial \kt{b}{\mathcal{P}}{x}$ ( where $\mathcal{P} = \langle x \mid x = x^{2} \rangle$) to   
1298: the algebraic problem of finding globally fixed profiles.  The latter problem is quite
1299: easy; we give a complete classification of fixed profiles for $F$, $T$, and $V$ in
1300: this section.
1301: 
1302: 
1303: %%%%%%%%%%%%%%%%%%%%%%%%
1304: \subsection{Conventions}
1305: %%%%%%%%%%%%%%%%%%%%%%%%%%
1306: 
1307: We fix some conventions for portraying profiles (and pictures) over the semigroup
1308: presentation $\langle x \mid x = x^{2} \rangle$.  
1309: 
1310: First, we draw every transistor in a
1311: picture or profile as a point and omit the frame, so that, for instance, the element $x_0 \in F$
1312: (depicted as an ordinary picture on the left) looks like the right half of Figure \ref{gen}.
1313: 
1314: \begin{figure} [!h] 
1315: \begin{center}
1316: \includegraphics{num3.eps}
1317: \end{center}
1318: \caption{A convention for drawing pictures over the semigroup presentation
1319: $\mathcal{P} = \langle x \mid x=x^{2} \rangle$.  On the left, we have a picture drawn in 
1320: the usual fashion; on the right is its equivalent. } 
1321: \label{gen}
1322: \end{figure}
1323: 
1324: We also need some conventions that will allow us to portray infinite pictures using a finite amount
1325: of space.
1326: We draw an empty dot at the end of a wire to indicate that the bottom of the wire doesn't
1327: connect to any transistor.  A solid dot at the end of a wire indicates that the wire
1328: connects to the top of some transistor.  This transistor may be either an $(x,x^{2})$-transistor
1329: or an $(x^{2}, x)$-transistor.  
1330: A wire with no dot at the end may connect to a transistor or not; we make no assumption one way or the
1331: other.
1332: 
1333: Finally, we let $T_{m}$ denote the full ordered rooted binary tree of depth $m$.  We let $\dot{T}_{m}$
1334: denote the full ordered rooted binary tree of depth $m$, where each leaf ends in a solid dot.
1335: %For instance, $\dot{T}_{2}$ is depicted in Figure \ref{exmple}. 
1336: %\begin{figure}[!h] 
1337: %\begin{center}
1338: %\includegraphics{num11.eps}
1339: %\end{center}
1340: %\caption{(label: exmple)  The picture $\dot{T}_{2}$.} 
1341: %\label{exmple}
1342: %\end{figure}
1343: Thus, $\dot{T}_m$ is the unique ordered rooted binary tree having $2^{m}$ dotted leaves, each at distance $m$ from the root.
1344: The dot on each leaf (wire) implies that each connects to the top of some transistor.  The picture $T_{m}$ is the same tree, but without
1345: the dots on the leaves.  
1346: Thus, no particular wire in $T_m$ which corresponds to a leaf necessarily leads 
1347: to the top of a transistor.  Notice, however,
1348: that in a profile 
1349: at least one of the leaves beneath a given degree $3$ vertex must attach to the top of a transistor, since there are no maximal transistors
1350: in a profile (by Theorem \ref{profile}). 
1351: 
1352: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1353: \subsection{Thompson's Group $F$}
1354: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1355: 
1356: Let $\Delta$ be a profile of $\kpw$ which is fixed by all of $F$.  Without loss of generality,
1357: we can assume that $\Delta$ has one of the forms in Figure \ref{cases}.
1358: \begin{figure} [!h] 
1359: \begin{center}
1360: \includegraphics{num4.eps}
1361: \end{center}
1362: \caption{The six cases.}
1363: \label{cases}
1364: \end{figure}
1365: (The only other possible cases are $1' - 6'$, which are obtained by reflecting $1 - 6$ across a vertical axis.  Note
1366: that the resulting cases are not mutually exclusive.)
1367: 
1368: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1369: \subsubsection{The Even-numbered Cases}
1370: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1371: 
1372: We consider even-numbered cases first.  Let $x_0$ act on any profile $\Delta$ covered by Case $2$, $4$, or $6$.  The
1373: results appear in Figure \ref{evencases}.
1374: \begin{figure} [!h] 
1375: \begin{center}
1376: \includegraphics{num5.eps}
1377: \end{center}
1378: \caption{The even-numbered cases.}
1379: \label{evencases}
1380: \end{figure}
1381: After cancelling dipoles twice, we arrive at the infinite pictures
1382: in the column at the far right of the Figure.  We claim
1383: that these infinite pictures 
1384: necessarily contain no dipoles, no matter how the wires terminating in  black dots are connected
1385: to transistors.  (It is clear also that these contain no maximal transistors.)
1386: 
1387: To prove the claim, first note that any dipole in the product $x_0 \cdot \Delta$ must be formed of one transistor
1388: in $x_0$ and another in $\Delta$.  Thus, an infinite picture 
1389: on the right side of Figure \ref{evencases} contains a dipole only if
1390: one of the two pictured vertices of degree $3$ 
1391: (both of which represent transistors from $x_0$) can form the top half of a dipole.
1392: 
1393: This is clearly impossible in Cases $2$ and $6$.  In Case $4$, it is enough to show
1394: that the lower transistor cannot form the top half of a dipole. 
1395: If we assume that it does, then the original profile $\Delta$ would have the form in Figure \ref{cont}.
1396: \begin{figure} [!h] 
1397: \begin{center}
1398: \includegraphics{num6.eps}
1399: \end{center}
1400: \caption{An impossible profile.  No matter how we attach the black-dotted wire to the top of a
1401: transistor, a dipole will be formed.} 
1402: \label{cont}
1403: \end{figure}
1404: No profile can have this form, however, since it is impossible to connect the black-dotted wire to a transistor
1405: without forming a dipole, and $\Delta$ cannot contain dipoles.  This proves the claim. 
1406: 
1407: Finally, we compare the reduced profiles $x_0 \ast \Delta$ at the right in Figure \ref{evencases} with the originals in
1408: Figure \ref{cases}. Since $\Delta$ is fixed by all of $F$, we must have that $x_0 \ast \Delta = \Delta$.  This is
1409: impossible, as we easily see.  For instance, in Case $2$, the left wire dangling from the bottom of the topmost
1410: transistor in $\Delta$ doesn't connect to a transistor, but the wire of the same description in $x_0 \ast \Delta$
1411: does.  The other even cases are left as easy exercises.
1412: 
1413: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1414: \subsubsection{Cases 3 and 5}
1415: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1416: If $\Delta$ is represented in Case $5$, then it must have the form in Figure \ref{sub5}a), where $\Delta'$ is another
1417: profile.
1418: \begin{figure} [!h] 
1419: \begin{center}
1420: \includegraphics{num7.eps}
1421: \end{center}
1422: \caption{a) The general form of a profile from Case 5.  b) The effect of letting $x_{0}^{-1}$ act.
1423: c) The profile $\Delta_{R}$.  d)  The element $x_1 \in F$.} 
1424: \label{sub5}
1425: \end{figure}
1426: If we let $x_{0}^{-1}$ act on $\Delta$, then after removing a dipole and an exposed transistor, we arrive at the
1427: profile on the far right of Figure \ref{sub5}b).  A simple induction using the fact that $\Delta = x_{0}^{-1} \ast \Delta$
1428: now shows $\Delta$ is the (unique) profile of the form depicted in Figure \ref{sub5}c).  It is easy to check that
1429: $\Delta$ is fixed by all of $F$; indeed, it is enough to show that $\Delta$ is fixed by the generators $x_0$ and $x_1$.
1430: We leave this verification as an exercise.
1431: 
1432: If $\Delta$ is represented in Case $3$, then it must have the form in Figure \ref{lr}a), where $\Delta'$ and $\Delta''$
1433: are profiles.
1434: \begin{figure} [!h] 
1435: \begin{center}
1436: \includegraphics{num8.eps}
1437: \end{center}
1438: \caption{a) The general form of a profile from Case 3.  b) The profile $\Delta_{L-R}$.}
1439: \label{lr}
1440: \end{figure}
1441: An argument similar to the one used for Case $5$ shows that $\Delta''$ has the form depicted in Figure \ref{sub5}c), and
1442: $\Delta'$ is the result of reflecting $\Delta''$ across a vertical axis.  The details are left as an exercise.  It
1443: follows  that $\Delta$ is the profile depicted in Figure \ref{lr}b), which is indeed fixed by both $x_0$ and $x_1$.
1444: 
1445: \begin{figure} [!h] 
1446: \begin{center}
1447: \includegraphics{num9.eps}
1448: \end{center}
1449: \caption{The action of $x_0$ on a profile from Case 1.} 
1450: \label{case1}
1451: \end{figure}
1452: 
1453: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1454: \subsubsection{Case 1}
1455: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1456: 
1457: We now turn to Case $1$.  Let $x_0$ act on $\Delta$.
1458: There are two subcases to consider:  either $\Delta$ has the form depicted in Figure \ref{subcases1}a) 
1459: (and thus the infinite picture 
1460: at the far right in Figure \ref{case1} contains a dipole) or the infinite picture 
1461: at the far right in Figure \ref{case1} is reduced. 
1462: 
1463:  We
1464: now rule out the first possibility using the fact that $\Delta$ is invariant under the action of $F$.  Let
1465: $x_1$ act on $\Delta$; after reducing two dipoles we arrive at the infinite picture 
1466: $\widehat{\Delta}$ on the far right of
1467: Figure \ref{subcases1}b).  The transistors enclosed by the dotted circle 
1468: were contributed by $x_1$, and any dipole in $\widehat{\Delta}$ would have to involve one of these three 
1469: transistors.  Now note that, 
1470: of these, only the transistor labelled $\ast$ could form half of a dipole; the others could not, even
1471: after we cancel any dipole involving $\ast$.  Now we compare $\Delta$ and $\widehat{\Delta}$.  Under any isomorphism
1472: between $\Delta$ and $\widehat{\Delta}$, the transistors labelled $1$ and $2$ in $\widehat{\Delta}$ must correspond
1473: (respectively) to the transistors labelled i) and ii) in $\Delta$ (as depicted in Figure \ref{subcases1}a).  This is not possible,
1474: since ii) is a $(x^{2}, x)$-transistor and $2$ is a $(x, x^{2})$-transistor.
1475: 
1476: \begin{figure} [!h] 
1477: \begin{center}
1478: \includegraphics{num10.eps}
1479: \end{center}
1480: \caption{a)  If we let $x_0$ act on this profile, the transistor labelled ii) will form half of a dipole.  b) The action
1481: of $x_1$ on the profile from a).}
1482: \label{subcases1}
1483: \end{figure}
1484: 
1485: It follows that we can assume that there are no dipoles in the 
1486: infinite picture at the far right of Figure \ref{case1}.  After comparing
1487: this profile with the profile $\Delta$, we can conclude that $\Delta$ has the form $\dot{T}_{2}$. 
1488: We now multiply $\Delta$ by $x_0$, $x_0 x_1 x_{0}^{-1}$, $x_1 x_{0}^{-1}$ and $x_{0}^{-1}$.  The results are listed in
1489: Figure \ref{arg}(a-d) (in the same order).
1490: \begin{figure} [!h] 
1491: \begin{center}
1492: \includegraphics{num12.eps}
1493: \end{center}
1494: \caption{The actions of a) $x_0$, b) $x_0 x_1 x_{0}^{-1}$, c) $x_1 x_{0}^{-1}$ and d) $x_{0}^{-1}$ on $\Delta$.}
1495: \label{arg}
1496: \end{figure}
1497: Dotted circles enclose the transistors that were contributed by the acting element.  If we knew that there were
1498: no dipoles in the infinite 
1499: pictures at the right, we could use the fact that all of them are equal to $\Delta$ in order to
1500: conclude that $\Delta$ has the form  $\dot{T}_{3}$.  
1501: \begin{figure} [!h]
1502: \begin{center}
1503: \includegraphics{num13.eps}
1504: \end{center}
1505: \caption{This profile would form dipoles in cases b) and c) from Figure \ref{arg}.}
1506: \label{lucky}
1507: \end{figure}
1508: The infinite pictures 
1509: on the far right of a) and d) in Figure \ref{arg} are necessarily reduced.  The profiles in b) and c) will
1510: be reduced unless $\Delta$ has the form in Figure \ref{lucky}b).
1511: 
1512: We now rule out the latter possibility.  Let $x_1$ act on $\Delta$.
1513: \begin{figure} [!h]
1514: \begin{center}
1515: \includegraphics{num14.eps}
1516: \end{center}
1517: \caption{The action of $x_1$ on the profile from Figure \ref{lucky}.}  
1518: \label{lucky2}
1519: \end{figure}
1520: Any isomorphism between $\widehat{\Delta}$ and $\Delta$ must match the transistors labelled 1) and 2) with
1521: the transistors labelled i) and ii), respectively.  This is impossible, since 2) is an $(x, x^{2})$-transistor and
1522: ii) is an $(x^{2}, x)$-transistor.
1523: 
1524: It now follows that $\Delta$ has the form of $\dot{T}_{3}$.  We multiply $\Delta$ by $x_0$, $x_0 x_1 x_{0}^{-1}$, 
1525: $x_1 x_{0}^{-1}$, $x_{0}^{-1}$.  The results appear in Figure \ref{deeper}.
1526: \begin{figure} [!h]
1527: \begin{center}
1528: \includegraphics{num15.eps}
1529: \end{center}
1530: \caption{The results of letting a) $x_0$, b) $x_0 x_1 x_{0}^{-1}$, c) $x_1 x_{0}^{-1}$, and d) $x_{0}^{-1}$ act on
1531: $\dot{T}_{3}$.}
1532: \label{deeper}
1533: \end{figure}
1534: Note that it is no longer possible for the transistors from the acting elements (circled) to form dipoles, so
1535: all of the profiles in Figure \ref{deeper} are reduced.  Each of these profiles is equal to $\Delta$, since $\Delta$ is invariant
1536: under the action of $F$.  It follows that $\Delta$ has the form of $\dot{T}_{4}$.
1537: 
1538: We now repeat this argument, letting the same four elements act on $\Delta$.  In this way, we conclude by induction that
1539: $\Delta$ has the form of the full infinite binary $T_{\infty}$.  We write $\Delta = \Delta_{\infty}$.
1540: 
1541: We've proved the following theorem:
1542: \begin{theorem}
1543: Thompson's group $F$ fixes exactly four profiles:
1544: $$ \Delta_{L}, \Delta_{R}, \Delta_{L-R}, ~and~ \Delta_{\infty}.$$
1545: \qed
1546: \end{theorem}
1547: These profiles come from Cases $5'$, $5$, $3$, and $1$, respectively.
1548: 
1549: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1550: \subsection{Thompson's Groups $T$ and $V$}
1551: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1552: 
1553: This subsection is devoted to a proof of the following theorem:
1554: \begin{theorem} \label{fixTV}
1555: Thompson's groups $T$ and $V$ fix only the profile $\Delta_{\infty}$.
1556: \end{theorem}
1557: \begin{proof}
1558: Let $\mathcal{P} = \langle x \mid x = x^{2} \rangle$.
1559: Suppose that the group $\di{a}{\mathcal{P}}{x} \cong T$ fixes the profile $\Delta$.  It is not difficult
1560: to see that $\Delta$ has the form in Figure \ref{TV}a), without loss of generality.
1561: \begin{figure} [!h]
1562: \begin{center}
1563: \includegraphics{num17.eps}
1564: \end{center}
1565: \caption{a) The general form of a profile in the cubical complexes for $T$ and $V$; b) the acting
1566: element $\pi_1$, which represents a half-rotation of the circle; c) the effect of letting $\pi_1$ act on
1567: the profile from a).}       
1568: \label{TV}
1569: \end{figure}
1570: We let $\pi_{1} \in \di{a}{\mathcal{P}}{x}$ act on $\Delta$; the result is portrayed in Figure \ref{TV}c), where
1571: the circled transistor is contributed by $\pi_1$.  It follows that the picture
1572: in c) is reduced.  Combining a) and c), which are both equivalent since $\pi_1 \ast \Delta = \Delta$, we have
1573: that $\Delta$ has the form $T_{2}$.
1574: 
1575: Next, we claim that at least one of the four leaves of $T_{2}$ connects to the top of an
1576: $(x, x^{2})$-transistor.  If not, then consider an $(x, x^{2})$-transistor $T$ in $\Delta$
1577: which is distinct from the three $(x, x^{2})$-transistors in $T_{2}$, and minimal among $(x, x^{2})$-transistors
1578: with this property.  The wire attached to the top of such a transistor could only lead up to
1579: the bottom of an $(x^{2}, x)$-transistor $T'$.  This implies that $T$ and $T'$ form a dipole, which
1580: contradicts the fact that $\Delta$ is reduced.
1581: 
1582: We therefore assume, without loss of generality, that $\Delta$ has the form in Figure \ref{wlog}a).
1583: \begin{figure} [!h]
1584: \begin{center}
1585: \includegraphics{num19.eps}
1586: \end{center}
1587: \caption{a) Without loss of generality, $\Delta$ has this form.  b) The element $\pi_2$, which represents
1588: a quarter-turn of the circle.}  
1589: \label{wlog}
1590: \end{figure}
1591: Now we multiply $\Delta$ by $\pi_2$, $\pi_{2}^{2}$, and $\pi_{2}^{3}$, where $\pi_{2}$ is as in Figure \ref{wlog}b)    
1592: to get the three profiles in Figure \ref{three}.
1593: \begin{figure} [!h]
1594: \begin{center}
1595: \includegraphics{num20.eps}
1596: \end{center}
1597: \caption{The results of letting the powers of $\pi_2$ act on the profile from Figure \ref{wlog}a).} 
1598: \label{three}
1599: \end{figure}
1600: Since the profiles from Figure \ref{three} and the profile from Figure \ref{wlog}a) are all equal to $\Delta$, it follows that $\Delta$ has the form
1601: $T_3$.
1602: 
1603: We then argue, as before, that at least one of the eight leaves at the bottom of $T_3$ must be attached to the top of
1604: an $(x, x^{2})$-transistor.  We then multiply $\Delta$ by $1$, $\pi_{3}$, $\pi_{3}^{2}$, $\pi_{3}^{3}$, $\ldots$,
1605: $\pi_{3}^{7}$, where $\pi_{3}$ is a picture representing a one-eighth turn of the circle. 
1606: 
1607: If we compare the eight resulting profiles, we conclude that $\Delta$ is equivalent to $T_4$.  We can continue in a similar
1608: way, and eventually conclude that $\Delta = \Delta_{\infty}$.
1609: 
1610: This proves the theorem in the case of $T$, and the proof for $V$ is the same word for word.
1611: \end{proof}
1612: 
1613: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1614: 
1615: \section{The Cases of $\Delta_{L}$, $\Delta_{R}$, and $\Delta_{L-R}$} \label{end}
1616: 
1617: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1618: 
1619: Now we consider the geodesic rays $c \in \Delta_{L} \cup \Delta_{R} \cup \Delta_{L-R}$.  Consider 
1620: first the profile $\Delta_{L-R}$.  By Lemma \ref{trav}, any geodesic ray $c$ in $\Delta_{L-R}$ is contained in
1621: the subcomplex $K$ of $X_F \, \, ( = \dpw$, where $\mathcal{P} = \langle x \mid x = x^{2} \rangle$ and $w=x$). 
1622: pictured in Figure \ref{funky}. 
1623: 
1624: %FIGURE HERE \# 26
1625: \begin{figure} [!h]  
1626: \begin{center}
1627: \includegraphics{num26.eps}
1628: \caption{A flat sector in $X_F$.  An integer lattice point $(m,n)$ ($m,n \geq 0$) corresponds to the 
1629: tree $T_{m,n}$ having a root caret, $m$ carets dangling to the left, and $n$ carets dangling to the right.} 
1630: \label{funky}
1631: \end{center}
1632: \end{figure}
1633: 
1634: Thus $K$ may be naturally identified with $\mathbb{R}^{2,+} \cup I$, where
1635: $\mathbb{R}^{2,+} = \{ (x,y) \in \mathbb{R}^{2} \mid x,y \geq 0 \}$, $I$ is the unit interval, and 
1636: $\mathbb{R}^{2,+} \cap I = \{ (0,0) \}$.
1637: 
1638: \begin{lemma} \label{Crisp} 
1639: The inclusion $i: K \rightarrow X_F$ is an isometric embedding.
1640: \end{lemma}
1641: 
1642: \begin{proof}
1643: We appeal to Theorem 1(2) of \cite{CW}, which says:
1644: If $X$ and $Y$ are finite dimensional CAT(0) cubical complexes and $\Phi: X \rightarrow Y$ is a cubical map, then  
1645: the map $\Phi$ is an isometric embedding if and only if, for every vertex $v \in X$, the simplicial map between links
1646: $\mathrm{Lk} (x,X) \rightarrow \mathrm{Lk}(\Phi(x), Y)$ induced by $\Phi$ is injective with image a full subcomplex
1647: of $\mathrm{Lk}(\Phi(x),Y)$.  (We refer the reader to \cite{BH}, page 102 for a discussion of the link; Crisp and Wiest
1648: define cubical maps on page 443 of \cite{CW}, and it is clear that the inclusion map is cubical.)
1649: 
1650: We consider the link of a vertex $T_{m,n}$, where $m,n > 1$, and leave the verifications for the other vertices as
1651: an exercise.  
1652: The link $\mathrm{Lk}(T_{m,n},K)$ is a square, i.e., the obvious one-dimensional simplicial complex consisting of $4$ vertices and $4$ edges.
1653: This link will be embedded in $\mathrm{Lk}(T_{m,n},X_F)$ as a full subcomplex if and only if (1) there is no two-dimensional cube $C$ in
1654: $X_F$ such that $T_{m,n-1}$ and $T_{m,n+1}$ are both vertices of $C$, and (2) there is no two-dimensional cube $C$ in $X_F$ such that
1655: $T_{m-1, n}$ and $T_{m+1, n}$ are both vertices of $C$.        
1656: 
1657: We now check (1); the argument for (2) is similar.  
1658: If there is such a cube $C$, then $C$ can be represented by a picture as in Figure \ref{cubepic}, consisting of $k$ shaded
1659: and $2$ unshaded transistors.  The four corners of $C$ are labelled by pictures having $k$, $k+1$, $k+1$, and $k+2$ transistors.  It follows
1660: that $k = m+n$, that $T_{m,n-1}$ is the result of leaving off both unshaded transistors, and that $T_{m,n+1}$ is the result of 
1661: shading both transistors.  From this we get a contradiction, since the unshaded transistors of the cube $C$ must both be maximal, and there is
1662: no way to remove two transistors that are both maximal in $T_{m,n+1}$ and arrive at $T_{m,n-1}$.
1663: \end{proof}
1664: 
1665: \begin{theorem}
1666: If $c$ is a geodesic ray in $X_F$ issuing from $\ast$ and $c \not \in \Delta_{\infty}$, then $c$ represents a point at infinity that is fixed by all
1667: of $F$ if and only if $c \in \Delta_{L} \cup \Delta_{R} \cup \Delta_{L-R}$.  The subspace of $\partial X_F$
1668: consisting of $\Delta_{L} \cup \Delta_{R} \cup \Delta_{L-R}$ is an arc of Tits length $\pi / 2$.
1669: \end{theorem}
1670: 
1671: \begin{proof}
1672: If $f: X \rightarrow Y$ is an isometric embedding between CAT(0) spaces, then the induced map $f_{\infty}: \partial X
1673: \rightarrow \partial Y$ is an isometry, where the boundary is endowed with the angular metric (\cite{BH}, page 280).    
1674: By the previous lemma, $K$ is isometrically embedded in $X_F$; by Lemma \ref{trav}, a geodesic ray $c$ issuing from $\ast$
1675: represents $[c] \in \Delta_{L} \cup \Delta_{R} \cup \Delta_{L-R}$ if and only if $\mathrm{Im} \, c \subseteq K$.  It follows
1676: from this that the image of $\partial K$ under the map $\partial K \rightarrow \partial X_F$ is
1677: precisely $\Delta_{L} \cup \Delta_{R} \cup \Delta_{L-R}$.  The second statement now follows from the fact that
1678: $\partial K$ is isometric to $[0, \pi/2]$.
1679: 
1680: Now suppose that $c \not \in \Delta_{\infty}$ 
1681: is a geodesic ray in $X_F$ issuing from $\ast$.  If $c$ is fixed by all of $F$ under the
1682: action $\ast$, then the argument
1683: of Section \ref{main} shows that $c \in \Delta_{L} \cup \Delta_{R} \cup \Delta_{L-R}$.
1684: 
1685: Conversely, suppose that 
1686: $c \in \Delta_{L} \cup \Delta_{R} \cup \Delta_{L-R}$.  It follows from this and Lemma \ref{trav} that
1687: $~Im~ c \subseteq K$, so every point $x \in ~Im~ c$ is within $1 + 2\sqrt{2}$ of a point in
1688: $\{ 2, 3, \ldots \} \times \{ 2, 3, \ldots \} \subseteq \mathbb{R}^{2,+}$.  
1689: We let $\widehat{T}_{m,n}$ denote
1690: the tree in Figure \ref{done}, which consists of $T_{m,n}$ and one additional caret:
1691: 
1692: %FIGURE HERE \# 27
1693: \begin{figure} [!h]
1694: \begin{center}
1695: \includegraphics{num27.eps}
1696: \caption{A picture of the tree $\widehat{T}_{m,n}$, which consists of $T_{m,n}$ and one
1697: additional caret.}   
1698: \label{done}
1699: \end{center}
1700: \end{figure}
1701: 
1702: If $m,n \geq 2$, it is routine to check that 
1703: \begin{eqnarray*}
1704: x_0 \cdot T_{m,n} & = & T_{m+1, n-1}; \\
1705: x_1 \cdot T_{m,n} & = & \widehat{T}_{m,n-1}.
1706: \end{eqnarray*}
1707: The trees $T_{m,n}$ and $T_{m+1, n-1}$ can be joined by an edge-path of length $2$ in $X_F$.  
1708: The same goes for
1709: $T_{m,n}$ and $\widehat{T}_{m,n-1}$, so $d( x_i \cdot T_{m,n} , T_{m,n} ) \leq 2$ $(i=0,1)$.
1710: 
1711: Let $t \geq 0$.  We have:
1712: \begin{eqnarray*}
1713: d( c(t) , x_i \cdot c(t) ) & \leq & d(c(t), T_{m,n}) + d(T_{m,n}, x_i \cdot T_{m,n} ) + 
1714: d( x_i \cdot T_{m,n} , x_i \cdot c(t)) \\
1715: & \leq & 4 + 4 \sqrt{2}.
1716: \end{eqnarray*}
1717: 
1718: Since this estimate doesn't depend upon $t$, it follows that $x_0$, $x_1$ both fix $c$  under the action
1719: $\ast$.  This implies that $F$ fixes $c$,
1720: since $x_0$, $x_1$ generate $F$.
1721: \end{proof}
1722: 
1723: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1724: 
1725: \section{The Case of $\Delta_{\infty}$} \label{infty}
1726: 
1727: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1728: 
1729: This section is devoted to an investigation of fixed points in
1730: $\Delta_{\infty}$.  Our main result is the following:
1731: 
1732: \begin{theorem} \label{good}
1733: The profile $\Delta_{\infty}$ contains no global fixed point of $T$ or $V$.
1734: \end{theorem}
1735: 
1736: An immediate consequence of Theorem \ref{good} and Theorem \ref{fixTV} is:
1737: 
1738: \begin{corollary}
1739: Thompson's groups $T$ and $V$ act without global fixed points on the boundaries-at-infinity of
1740: their respective picture complexes.
1741: \qed
1742: \end{corollary}
1743: 
1744: The case for Thompson's group $F$ is more complicated.  I don't know whether 
1745: $\Delta_{\infty} \subseteq \partial X_F$ contains fixed points of $F$ or not.
1746: Example \ref{clue1} and Proposition \ref{clue2} give evidence for and against the existence
1747: of fixed points in $\Delta_{\infty}$, respectively.
1748: 
1749: Our arguments will use a simple procedure for embedding any CAT(0) cubical complex into Hilbert space. 
1750: \begin{proposition}
1751: If $X$ is a CAT(0) cubical complex, then there is an embedding $\rho : X \rightarrow \ell^{2} \left( \mathcal{H} \right)$,
1752: where $\mathcal{H}$ is the collection of hyperplanes in $X$.  The map $\rho$ doesn't increase distances.
1753: \end{proposition}
1754: 
1755: \begin{proof}
1756: Let $X$ be a CAT(0) cubical complex with a distinguished base vertex $\ast$.  For any given hyperplane $H$ in $X$, we
1757: identify the closed $1/2$-neighborhood of $H$ with $H \times [0,1]$ in such a way that
1758: $d( \ast, H \times \{ 0 \} ) < d ( \ast, H \times \{ 1 \} )$, and let $H_t$ denote $H \times \{ t \}$ for
1759: $t \in [0,1]$.
1760: 
1761: Let $\rho : X \rightarrow \ell^{2} ( \mathcal{H} )$ send $x$ to
1762: $\Sigma_{H \in \mathcal{H}} f_{H}(x) H$, where
1763: $$f_{H}(x) = ~sup~ \left( \{ t \in [0,1] \mid [ \ast, x ] \cap H_t \neq \emptyset \} \cup \{ 0 \} \right).$$
1764: It is clear that $\rho$ is an embedding, that each cube of $x$ is embedded into $\ell^{2}( \mathcal{H})$ isometrically,
1765: and that the restrictions $\rho_{\mid C}$ (where $C$ is a cube) agree on overlaps.  Note also that each sum
1766: $\Sigma_{H \in \mathcal{H}} f_{H}(x) H$ is finite, since any pair of vertices in $X$ are separated by at most
1767: finitely many hyperplanes.  The first statement follows.
1768: 
1769: Let $[x,y]$ be a geodesic in $X$.  Since the restriction of $\rho$ to each cube is an isometry, $\rho$ preserves the
1770: lengths of paths.  Therefore,
1771: $$ d_{X}(x,y) = \ell \left( \rho [x,y] \right) \geq d_{\ell^{2}(\mathcal{H})} \left( \rho(x) , \rho(y) \right).$$
1772: \end{proof}
1773: 
1774: \begin{example} \label{hilbert}
1775: We consider the image of a certain $x \in X_F$ under the map $\rho$.
1776: \begin{figure} [!h]  
1777: \begin{center}
1778: \includegraphics{num23.eps}
1779: \end{center}
1780: \caption{A point $x$ in the cubical complex for Thompson's group $F$ (left) and its image under the map $\rho$ (right).}
1781: \label{embed}
1782: \end{figure}
1783: Note that each hyperplane occurring in the sum $\rho(x)$ corresponds in a straightforward way
1784: to a particular transistor $\Delta_{\infty}$.  We can use this fact to simplify our notation -- compare 
1785: Figure \ref{simple} to the right
1786: half of Figure \ref{embed}.
1787: \begin{figure} [!h]
1788: \begin{center}
1789: \includegraphics{num24.eps}
1790: \end{center}
1791: \caption{A simpler notation for $\rho(x)$.}
1792: \label{simple}
1793: \end{figure}
1794: In this way, we identify each point of $\mathrm{Im} \, \rho$ with a picture $\Delta$ such that
1795: each maximal transistor is labelled by a number $t \in (0,1]$, and every other transistor
1796: is labelled by the number $1$.  We will usually just omit the label of a transistor $T$ if $T$ is not maximal.  
1797: 
1798: With this convention, an element of $F$, $T$, or $V$ acts on a point of $\mathrm{Im} \, \rho$ by the usual picture multiplication.
1799: This action is somewhat tricky to describe if a transistor from the acting element forms a dipole with an
1800: exposed transistor labelled by a number $t \neq 1$.  In practice, however, we will always be able to avoid 
1801: considering this
1802: situation.  If there are no such dipoles, then the action is simple to describe: concatenate and reduce dipoles.
1803: \end{example}
1804: 
1805: It will be helpful to have a vocabulary for describing subtrees of a given labelled tree $T$.  If $T_{\infty}$ is the full 
1806: ordered rooted binary tree of infinite depth, we use binary strings to denote the vertices of degree three in $T_{\infty}$.
1807: We give each edge in $T_{\infty}$ a label of $0$ or $1$; the label is $0$ if the edge forms the left half of a caret, and
1808: $1$ if the edge forms the right half of a caret.  Now label each vertex $v$ in $T_{\infty}$ by the label of the unique geodesic
1809: path from the root to $v$.  For instance, if the geodesic path from the root to $v$ passes through the right half of a caret twice,
1810: and then through the left half of a caret, and then finally through the right half of a caret again, then the label of $v$ is $1101$.
1811: The root has the empty label.  
1812: 
1813: Now if $T$ is an arbitrary labelled subtree (as in Figure \ref{simple}) 
1814: of $T_{\infty}$ and $\mathrm{bin}$ is a binary string, we let $T_{\mathrm{bin}}$  
1815: denote the labelled tree having the vertex $\mathrm{bin}$ as its root.  For instance, if $T$ is the labelled tree in Figure \ref{simple},
1816: then $T_{1}$ consists of a single caret, labelled by the number $2/3$.  The tree $T_{01}$ is a single caret labelled by $3/4$.    
1817: 
1818: The following partial result will be used in the proof of Theorem \ref{good}.
1819: 
1820: \begin{proposition} \label{clue2}
1821: Let $c: [0, \infty) \rightarrow X$ be a geodesic ray, where $X= \kpw$, $\kapw$, or $\kbpw$, 
1822: $\mathcal{P} = \langle x \mid x = x^{2} \rangle$, and $w=x$.  Let $T$ be some rooted ordered binary tree in which each transistor is
1823: labelled by a $1$ (and thus $T$ corresponds to a vertex in $\kpw$, $\kapw$, or $\kbpw$, as the case may be) such that the subtrees 
1824: $T_{10}$ and $T_{11}$ each contain at least one caret.   
1825: 
1826: If $c(t) = T$ for some $t \in [0, \infty)$, then $x_{0} \ast c \neq c$.
1827: \end{proposition}
1828: 
1829: \begin{proof}
1830: We can express $\rho(c(t))$ as a tree of the form in Figure \ref{sec6arg} a), 
1831: \begin{figure} [!h]
1832: \begin{center}
1833: \includegraphics{num25.eps}
1834: \end{center}
1835: \caption{a)  The form of $\rho(c(t))$.  The trees $T_2$ and $T_3$ both contain at least one caret.  b) the effect of letting
1836: $x_0$ act on $\rho(c(t))$.}
1837: \label{sec6arg}
1838: \end{figure}
1839: where each tree $T_{i}$, $i=1,2,3$, is a labelled tree, i.e., a picture in which every transistor is an 
1840: $(x, x^{2})$-transistor, and the trees $T_2$ and $T_3$ each have at least one transistor.  
1841: The effect of the action by $x_0$
1842: is to transform the tree in Figure \ref{sec6arg}a) into the tree in Figure \ref{sec6arg}b).
1843: 
1844: Figure \ref{sec6arg}b) also serves as a definition of three different operations on labelled trees.  We now make this more explicit.
1845: If $T$ is a labelled tree having at least one transistor labelled $1$, then the act of removing the topmost transistor
1846: of $T$ leaves an ordered pair of trees $(\pi_1 T , \pi_2 T)$.  If $T_1$ and $T_2$ are labelled trees, then
1847: $T_1 \wedge T_2$ is the unique labelled tree satisfying $\pi_1 ( T_1 \wedge T_2 ) = T_1$ and
1848: $\pi_2 ( T_1 \wedge T_2 )$.
1849: 
1850: \begin{figure}
1851: \begin{center}
1852: \includegraphics{par.eps}
1853: \caption{On the right, we have the general picture describing a geodesic ray $c$ issuing from the basepoint $\ast$.  Every such
1854: geodesic ray must cross the dotted horizontal line; at the moment it does so, it is precisely $\sqrt{2}$ units distant from 
1855: its translate $x_{0} \cdot c$.}
1856: \label{contra}
1857: \end{center}
1858: \end{figure}
1859: 
1860: Note that 
1861: $$ || T - x_{0} \cdot T ||_{2}^{2} = || T_{1} - T_{1} \wedge T_{2} ||_{2}^{2} + || T_{2} - \pi_{1} T_{3} ||_{2}^{2} + ||T_{3} - \pi_{2} T_{3}||_{2}^{2}.$$
1862: Since each caret in each tree is labelled with a $1$, each of the terms on the right side of the equation is an integer which counts the number of carets 
1863: that are in one tree but not in the other.  It follows that
1864: $|| T_{1} - T_{1} \wedge T_{2} ||_{2}^{2} \geq 2$ and $||T_{3} - \pi_{2}T_{3}||_{2}^{2} \geq 1$.  This implies that
1865: $$d(c(t), x \cdot c(t)) \geq || T- x_{0} \cdot T ||_{2} \geq \sqrt{3}.$$
1866: 
1867: Now we appeal to Lemma \ref{monotone}(2), which implies that if there is $t' < t$ such that $d( c(t') , x_{0} \cdot c(t')) < \sqrt{3}$, then
1868: $c$ and $x_{0} \cdot c$ are not asymptotic, i.e., $c \neq x_{0} \ast c$.  We produce a $t' < t$ 
1869: where $d( c(t'), x_{0} \cdot c(t') ) = \sqrt{2}$
1870: in Figure \ref{contra}. 
1871: 
1872: \end{proof}
1873: 
1874: \begin{remark}  \label{rem1}
1875: \noindent (1)  It is reasonably clear from the proof that there are variations on this Proposition, in which the hypothesis that
1876: $T_{10}$ and $T_{11}$ are non-trivial is replaced by similar assumptions on different subtrees.
1877: 
1878: \noindent (2)  
1879: Let $v$ be a vertex in the cubical complex for Thompson's group $F$ such that
1880: $\rho(v)$ is an ordered labelled rooted binary tree $\widehat{T}$.  Note that 
1881: the coefficient of each caret in $\widehat{T}$ is either $1$ or $0$, since $v$ is a vertex.  
1882: The above argument shows that
1883: $$ \ast)  \quad d(v, x_{0} \cdot v) \geq \sqrt{ \left( 1 + ( \# \mathrm{~of~carets~in~} \widehat{T}_{10} ) \right) + 
1884: \left( 1 + ( \# \mathrm{~of~carets~in~} \widehat{T}_{11} \right)},$$         
1885: provided the trees $\widehat{T}_{10}$ and $\widehat{T}_{11}$ both contain at least one caret. 
1886: 
1887: This suggests a strategy for proving that $\Delta_{\infty}$ contains no fixed points of Thompson's group $F$.
1888: Suppose that $c$ is a geodesic ray having the profile $\Delta_{\infty}$.  It follows from Lemma \ref{trav}
1889: that $\rho(c(t))$ is a labelled ordered rooted binary tree $T(t)$, for any $t \geq 0$.  
1890: For any given caret $C \in T_{\infty}$, $C$ occurs in $T(t)$ with the coefficient $1$
1891: for $t$ sufficiently large, 
1892: since $c$ crosses every hyperplane
1893: in $\Delta_{\infty}$. 
1894: If $c(t)$ passed close to a vertex $v$ for a large value of $t$, then the
1895: inequality $\ast)$ shows that $d( v, x_{0} \cdot v)$ would be large, so that $d( c(t), x_{0} \cdot c(t))$ would
1896: also be large, and thus $c \neq x_{0} \ast c$.  (Indeed, the proof of Proposition \ref{clue2} shows it
1897: suffices to prove that $d( c(t), x_{0} \cdot c(t)) > \sqrt{2}$.)  Unfortunately, I know of no way to
1898: control the distance of $c(t)$ from a vertex, since $c$ will generally travel through cubes of higher and higher
1899: dimension, whose diameters go to infinity. 
1900: \end{remark}  
1901: 
1902: \noindent \emph{Proof of Theorem \ref{good}:} 
1903: Let $c \in \Delta_{\infty}$ be a geodesic ray in either $X_{T}$ or $X_{V}$
1904: which represents a fixed point at infinity. 
1905: We note first that $c(1)$ is the vertex labelled
1906: by the finite tree $T_1$. If $\pi_1 \in T \subseteq V$ is as in Figure \ref{TV}b), it is clear that
1907: $\pi_{1} \cdot T_1 = T_1$.  Now since $(\pi \cdot c)(1) = c(1)$ and 
1908: $\pi \cdot c$, $c$ are asymptotic by our assumptions, 
1909: it follows from Lemma \ref{monotone}(2) 
1910: that $(\pi \cdot c)(t) = c(t)$, for $t \geq 1$.  
1911: 
1912: Now we refer to Figure \ref{contra}.  The element $\pi_1$ flips the diamond
1913: on the right along its vertical axis, stabilizing the vertices $T_1$ and $T_2$.
1914: Since the geodesic ray travels through this diamond and $(\pi_1 \cdot c)(t) = c(t)$
1915: for $t \geq 1$, it must be that the ray $c$ travels along the straight line
1916: connecting $T_1$ to $T_2$, which is vertical in the Figure.
1917: 
1918: Next, we consider the action of $\pi_2 \in T$ (see Figure \ref{wlog}b)) on $c$.  Note
1919: first that $\pi_2$ stabilizes the vertex $T_2$.  It follows from this, Lemma \ref{monotone}(2), the 
1920: equality $c( 1 + \sqrt{2}) = T_2$, and the assumption that $\pi_{2}$ fixes $c$, 
1921: that $( \pi_{2} \cdot c )(t) = c(t)$ for $t \geq 1 + \sqrt{2}$.  
1922: Lemma \ref{trav}
1923: implies that $c( 1 + \sqrt{2} + \epsilon )$ is in the four dimensional cube $C$ having $T_2$ as its minimal
1924: vertex and $T_3$ as its maximal vertex.  If we identify the cube $C$ with $[0,1]^{4}$ in such a way that
1925: $(0,0,0,0) = T_2$ and $(1,1,1,1) = T_3$, then $\pi_2$ acts by cyclically permuting the coordinates of $C$.  It follows
1926: from this that the geodesic ray travels along the diagonal of $C$ from $T_2$ to $T_3$.  In particular,
1927: $c$ passes through $T_3$.
1928: 
1929: It now follows from Proposition \ref{clue2} that $c \neq x_0 \ast c$, which is a contradiction.
1930: \qed 
1931: 
1932: \vspace{20pt}
1933: 
1934: We conclude with an example giving some evidence that
1935: there may be fixed points of $F$ in $\Delta_{\infty}$. 
1936: 
1937: \begin{figure} [!h]
1938: \begin{center}
1939: \includegraphics{tree.eps}
1940: \caption{This labelled tree represents a point in $X_F$ which is moved only a 
1941: small distance by the generators of $F$. 
1942: It should be possible to 
1943: build a sequence of similar, arbitrarily large labelled trees which converge to a fixed point at infinity.}  
1944: \label{trxm}
1945: \end{center}  
1946: \end{figure}
1947: 
1948: \begin{example} \label{clue1}       
1949: Figure \ref{trxm} depicts a labelled tree $T$ (i.e., point in Hilbert space)
1950: such that $||T - x_{0} \cdot T||_{2} = \sqrt{7}/2$.    The check is left as an exercise.
1951: It is not difficult to see that $x_0$ acts on (most) of the trees along the bottom by a leftward shift.
1952: In particular, the tree $T_{1\ldots 10}$ (where there are $n$ ones and a single $0$) is mapped
1953: to $T_{1 \ldots 10}$ (where there are $n-1$ ones); the tree $T_{0 \ldots 01}$ ($n$ zeros) is mapped
1954: to $T_{0 \ldots 01}$ ($n+1$ zeros).  
1955: This suggests a principle for building larger trees that are moved only a small distance by $x_0$:
1956: Begin with the tree $T_{m,m}$ and attach new trees $T'$ to the leaves, making sure that the tree attached 
1957: at a given leaf is within $\epsilon$ of its neighbor to the immediate left, where $\epsilon$ will depend on
1958: $m$.  The object is to make sure that $|| T - x_{0} \cdot T ||_{2}$ is less than or equal to $\sqrt{2}$, and
1959: it is not too difficult to see that this can be done for any $m$.  (A little experimentation also shows that
1960: it is useful to label the leftmost and rightmost carets of $T_{m,m}$ with $1/2$.)  Moreover, the trees $T'$
1961: that are attached ``close" to the root (in a sense that depends on $m$)
1962: can be made arbitrarily large.  Thus, we can make a sequence of labelled
1963: trees $\overline{T}_{k}$, which are each moved less than $\sqrt{2}$ units in the Hilbert metric, and 
1964: gradually fill up the complete binary tree $T_{\infty}$.  One can then hope  that the corresponding
1965: points $z_k$ in $X_F$ are also moved only a small distance by $x_0$ (as seems likely), so that  some subsequence of $z_k$
1966: converges to a point $\zeta$ at infinity which is fixed by the action of $x_0$.  With additional care, it should also be
1967: possible to do this so that each point $z_k$ is likewise moved only a small distance by $x_1$, and therefore $\zeta$ would
1968: be fixed by all of $F$.     
1969: 
1970: It seems very likely that all of the above can be done.  This is not enough, however, because the point $\zeta$ may well
1971: fail to have the profile $\Delta_{\infty}$.  Indeed, the tree in Figure \ref{trxm} has the property that most of its norm
1972: is contributed by $T_{7,7}$.  It appears likely that any tree in the 
1973: sequence $\overline{T}_k$ will have most of its norm contributed by $T_{n,n}$ (for appropriate $n$), and this may mean
1974: that $\zeta \in \Delta_{L} \cup \Delta_{L-R} \cup \Delta_{R}$.  I conjecture the following:
1975: 
1976: \begin{conjecture}
1977: A point $\zeta \in \partial X_F$ is fixed by all of $F$ if and only if $\zeta \in \Delta_{L} \cup \Delta_{L-R} \cup \Delta_{R}$.
1978: In particular, the only fixed points for the action of $F$ on $\partial X_F$ lie on an arc of Tits length $\pi /2$. 
1979: \end{conjecture}
1980: \end{example}
1981: 
1982: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1983: 
1984: \bibliography{infty}
1985: \nocite{*}
1986: \bibliographystyle{plain}
1987: 
1988: \end{document}
1989: 
1990: 
1991: 
1992: 
1993: