1: \documentclass{amsart}
2:
3: \usepackage{amsmath,amsfonts,amsthm,amssymb}
4: \usepackage{graphicx}
5:
6: \newcommand{\mapdown}[1]%
7: {\Big\downarrow\rlap{$\vcenter{\hbox{$\scriptstyle#1$}}$}}
8:
9:
10: \newcommand{\A}{\mathcal{A}}
11: \newcommand{\arr}{\longrightarrow}
12: \newcommand{\spec}[1]{\mathop{Sp}\left(#1\right)}
13: \newcommand{\symm}[1][\alb]{\mathop{\mathrm{Symm}}\left(#1\right)}
14: \newcommand{\supp}{\mathop{\mathrm{supp}}}
15: \newcommand{\C}{\mathbb{C}}
16: \newcommand{\R}{\mathbb{R}}
17: \newcommand{\st}{\mathop{\mathrm{St}}}
18: \newcommand{\wt}{\widetilde}
19: \newcommand{\alb}{\mathsf{X}}
20: \newcommand{\xo}{\alb^\omega}
21: \newcommand{\xmo}{\alb^{-\omega}}
22: \newcommand{\xs}{\alb^*}
23: \newcommand{\xz}{\alb^{\mathbb{Z}}}
24: \newcommand{\maxalg}{\mathcal{A}_{\textit{max}}}
25: \newcommand{\minalg}{\mathcal{A}_{\textit{min}}}
26: \newcommand{\nuke}{\mathcal{N}}
27: \newcommand{\cpg}{O_G}
28: \newcommand{\solen}[1][G]{\mathcal{S}_{#1}}
29: \newcommand{\si}{\widehat{\mathsf{s}}}
30: \newcommand{\sis}{\mathsf{s}}
31: \newcommand{\lims}[1][G]{\mathcal{J}_{#1}}
32: \newcommand{\Z}{\mathbb{Z}}
33: \newcommand{\img}[1]{\mathop{\mathrm{IMG}}\left(#1\right)}
34: \newcommand{\til}{\mathcal{T}}
35: \newcommand{\M}{\mathcal{M}}
36: \newcommand{\gi}{\M_{d^\infty}(G)}
37:
38: \newtheorem{theorem}{Theorem}[section]
39: \newtheorem{proposition}[theorem]{Proposition}
40: \newtheorem{corollary}[theorem]{Corollary}
41: \newtheorem{lemma}[theorem]{Lemma}
42:
43:
44: \theoremstyle{definition}
45:
46: \newtheorem{definition}{Definition}[section]
47: \newtheorem*{remark}{Remark}
48: \newtheorem{example}{Example}
49: \newtheorem{examples}[example]{Examples}
50: \newtheorem{question}{Problem}
51:
52: \numberwithin{equation}{section}
53:
54: \title{Self-similar groups, operator algebras and Schur complements}
55: \author{Rostislav Grigorchuk}
56: \address{Texas A\&M University, College Station, USA}
57: \email{grigorch@math.tamu.edu}
58: \thanks{Supported by NSF grants DMS-0456185 and DMS-0600975}
59:
60: \author{Volodymyr Nekrashevych}
61: \email{nekrash@math.tamu.edu}
62: \thanks{Supported by NSF grant DMS-0605019}
63: \subjclass[2000]{37F10, 47A10, 20E08}
64: \dedicatory{Dedicated to the 70th birthday of D.V.~Anosov}
65:
66: \begin{document}
67: \maketitle
68:
69: \begin{abstract}
70: In the first part of the article we introduce $C^*$-algebras
71: associated to self-similar groups and study their properties and
72: relations to known algebras. The algebras are constructed as
73: sub-algebras of the Cuntz-Pimsner algebra (and its homomorphic
74: images) associated with the self-similarity of the group. We study
75: such properties as nuclearity, simplicity and Morita equivalence
76: with algebras related to solenoids.
77:
78: The second part deals with the Schur complement transformations of
79: elements of self-similar algebras. We study properties of such
80: transformations and apply them to the spectral problem for Markov
81: type elements in self-similar $C^*-$algebras. This is related to
82: the spectral problem of the discrete Laplace operator on groups
83: and graphs. Application of the Schur complement method in many
84: situations reduces the spectral problem to study of invariant sets
85: (very often of the type of a ``strange attractor'') of a
86: multidimensional rational transformation. A number of illustrating
87: examples is provided. Finally we observe a relation between the
88: Schur complement transformations and Bartholdi-Kaimanovich-Virag
89: transformations of random walks on self-similar groups.
90: \end{abstract}
91:
92:
93: \section{Introduction}
94:
95: Self-similar groups is a class of groups which attracts more and
96: more attention of researchers from different areas of mathematics,
97: and first of all from group theory.
98:
99: Self-similar groups (whose study was initiated by the first named
100: author, N.~Gupta, S.~Sidki, A.~Brunner, P.M.~Neumann, and others)
101: posses many nice and unusual properties which allow to solve
102: difficult problems of group theory and related areas, even
103: including problems in Riemannian geometry and holomorphic
104: dynamics. They are closely related to groups generated by finite
105: automata (studied by J.~Ho{\v r}ej{\v s}, V.M.~Glushkov, S.
106: V.~Aleshin, V.I.~Sushchanski and others) and many of them belong
107: to another interesting class of groups---the class of branch
108: groups. We recommend the following sources for the basic
109: definitions and properties
110: \cite{nek:book,handbook:branch,grigorchuk:branch,sidki_monogr}.
111:
112: One of the main features that makes the class of self-similar
113: groups important is that it makes possible treatment of the renorm
114: group in a noncommutative setting. This passage from cyclic to a
115: non-commutative renormalization can be compared with the passage
116: from classical to non-commutative geometry, i.e., passage from
117: commutative $C^*$-algebras of continuous functions to
118: non-commutative $C^*$-algebras, see~\cite{connes:noncomg}.
119:
120: Recent research shows that operator algebras also play an
121: important role in the theory of self-similar groups and show
122: interesting connections with other areas (for instance with
123: hyperbolic dynamics). The first appearance of $C^*-$algebras
124: related to self-similar groups was in~\cite{bgr:spec} and was
125: related to the problem of computation of the spectra of Markov
126: type operators on the Schreier graphs related to self-similar
127: groups. In~\cite{nek:bim} the methods of~\cite{bgr:spec} were
128: interpreted in terms of the Cuntz-Pimsner algebras of the Hilbert
129: bimodules associated with self-similar groups. It was proved in~\cite{nek:bim}
130: that there exists a smallest self-similar norm (i.e., norm which
131: agrees with the bimodule) and that the Cuntz-Pimsner
132: algebra constructed using the completion of the group algebra by
133: the smallest norm is simple and purely infinite.
134:
135: This article is a survey of results and ideas on the interplay of
136: self-similar groups, renormalization, self-similar operator
137: algebras, spectra of Markov operators and random walks.
138:
139: We continue the study of self-similar completions of the group
140: algebra started in~\cite{bgr:spec} and~\cite{nek:bim}. We prove
141: existence of the largest self-similar norm on the group algebra
142: and define the maximal Cuntz-Pimsner algebra and maximal
143: self-similar completion of the group algebra
144: (Propositions~\ref{pr:cpimsner} and~\ref{pr:algmaximal}).
145:
146: Our starting point is the observation that self-similarities of a
147: Hilbert space $H$ (i.e., isomorphisms between $H$ and its $d-$th,
148: $d\ge 2$ power $H^d$) are in a natural bijection with the
149: representations of the Cuntz algebra $O_d$. Then we consider
150: self-similar unitary representations of a self-similar group and
151: following~\cite{nek:bim} define the associated universal
152: Cuntz-Pimsner algebra $\cpg$. The algebra $\cpg$ is universal in
153: the sense that any unitary self-similar representation of the
154: group can be extended to a representation of $\cpg$. Homomorphic
155: images of $\cpg$ can be constructed using different self-similar
156: representations of $G$. Among important self-similar
157: representations we study the natural unitary representation of $G$
158: on $L^2(\xo, \nu)$, where $\xo$ is the boundary of the rooted tree
159: on which $G$ acts and $\nu$ is the uniform Bernoulli measure on
160: it. Another important class of self-similar representations are
161: permutational representations of $G$ on countable $G$-invariant
162: subsets of $\xo$.
163:
164: The sub-$C^*$-algebra of $\cpg$ generated by $G$ is denoted
165: $\maxalg$. For every homomorphic image of the Cuntz-Pimsner
166: algebra $\cpg$ we get the respective image of $\maxalg$. The
167: algebra $\minalg$, for instance, is defined as the image of
168: $\maxalg$ under a permutational representation of $G$ on a generic
169: self-similar $G$-invariant countable set of $\xo$ (which can be
170: extended to a permutational representation of $\cpg$ in a natural
171: way). The algebra $\minalg$ was studied in~\cite{bgr:spec}
172: and~\cite{nek:bim}. Similarly, the algebra $\A_{mes}$ generated by
173: the natural representation of $G$ (and of $\cpg$) on $L^2(\xo,
174: \nu)$, is considered. This algebra is particularly convenient for
175: spectral computations (see Subsection~\ref{ss:scssalg}).
176:
177: It is convenient in many cases to pass to a bigger sub-algebra of
178: $\cpg$ than $\maxalg$. It is the algebra generated by $G$ and the
179: union of the matrix algebras $M_{d\times d}(\C)$ naturally
180: constructed inside $O_d\subset\cpg$. It is proved
181: in~\cite{nek:cpalg} that this algebra is denoted
182: $\M_{d^\infty}(G)$ and it is the universal algebra of the groupoid
183: of germs of the action of $G$ on the boundary $\xo$ of the rooted
184: tree. This makes it possible to apply the well-developed theory of
185: $C^*$-algebras associated to groupoids to the study of $\cpg$ and
186: $\maxalg$.
187:
188: For every homomorphic image of $\cpg$ (i.e., for every
189: self-similar representation of $G$) we get the corresponding image
190: of $\M_{d^\infty}(G)$.
191:
192: In Section~\ref{s:contracting} we show a relation of self-similar
193: groups and algebras to hyperbolic dynamics. If a self-similar
194: group $G$ is \emph{contracting}, then there is a natural Smale
195: space associated to it (the \emph{limit solenoid}). It is a
196: dynamical system $(\solen, \si)$ with hyperbolic behavior: the
197: space $\solen$ has a local structure of a direct product such that
198: the homeomorphism $\si$ is contracting on one factor and expanding
199: on the other. We prove in Theorem~\ref{th:moritaequiv} that the
200: algebra $\M_{d^\infty}(G)$ is Morita equivalent in this case to
201: the convolution algebra of the unstable equivalence relation on
202: the limit solenoid. Such convolution algebras were studied by
203: J.~Kaminker, I.~Putnam and J.~Spielberg
204: in~\cite{putnam,put_kam_sp,putnam:structure}.
205:
206: The algebras $\maxalg$, $\M_{d^\infty}(G)$ and their images under
207: self-similar representations of $G$ have a nice self-similarity
208: structure described by matrix recursions, which encode the
209: structure of the Moor diagrams of the Mealy type automata defining
210: the underlying group. This self-similarity is used in the second
211: part of the article in the study of spectra of elements of the
212: involved algebra.
213:
214: We recall at first the classical Schur complement transformation,
215: which is used in linear algebra for solving systems of linear
216: equations, in statistics for finding conditional variance of
217: multivariate Gaussian random variables and in Bruhat normal form
218: (see~\cite{cottle:schursurvey,cohn}). We show in our article that
219: Schur complement is also useful in the study of spectral problems
220: in self-similar algebras and that it can be nicely expressed in
221: terms of the Cuntz algebra. We establish some simple properties of
222: the Schur complement transformations and introduce a semigroup of
223: such transformations.
224:
225: The method that we use to treat the spectral problem could be a
226: first step in generalization of the method of Malozemov and
227: Teplyaev that they developed for the study of spectra of
228: self-similar graphs related to classical fractals (like the
229: Sierpinskii gasket)~\cite{teplyaev:sierpinski,malozemov_teplyaev}.
230: In fact they also use (in an implicit form) the Schur complement.
231: Their technique is developed for the case when only one complex
232: parameter is involved. Our technique (which is a development of
233: the technique used in~\cite{bgr:spec,gr_zu:lamp,grisunik:hanoi}
234: involves several parameters and therefore necessarily lead to
235: multidimensional rational mappings and their dynamics.
236:
237: The Schur complements in our situations are renormalization
238: transformations for the spectral problem and related problems.
239: Considered together they generate a noncommutative semigroup which
240: can be called the ``Schur renorm group'' (observe that in
241: classical situation the renorm group very often is a cyclic
242: semigroup).
243:
244: We illustrate our method by several examples the most
245: sophisticated among which is the example related to the
246: 3-generated torsion 2-group of intermedate growth constructed
247: in~\cite{grigorchuk:80_en}.
248:
249: The transformations that arise in this case are
250: \[\wt S_1:\left(\begin{array}{c} x\\ y\\ z\\ u\\
251: v\end{array} \right)\mapsto\left(\begin{array}{c} z+y\\
252: \frac{x^2(2yzv-u(y^2+z^2-u^2+v^2))}{(y+z+u+v)(y+z-u-v)(y-z+u-v)(-y+z+u-v)}\\
253: \frac{x^2(2zuv-y(-y^2+z^2+u^2+v^2))}{(y+z+u+v)(y+z-u-v)(y-z+u-v)(-y+z+u-v)}\\
254: \frac{x^2(2yuv-z(y^2-z^2+u^2+v^2))}{(y+z+u+v)(y+z-u-v)(y-z+u-v)(-y+z+u-v)}\\
255: u+v+\frac{x^2(2yzu-v(y^2+z^2+u^2-v^2))}{(y+z+u+v)(y+z-u-v)(y-z+u-v)(-y+z+u-v)}\end{array}\right)\]
256: and
257: \[
258: \wt S_2:\left(\begin{array}{c} x\\ y\\ z\\ u\\
259: v\end{array}\right)\mapsto\left(\begin{array}{c}\frac{x^2(y+z)}{(u+v+y+z)(u+v-y-z)}\\
260: u\\ y\\ z\\
261: v-\frac{x^2(u+v)}{(u+v+y+z)(u+v-y-z)}\end{array}\right)
262: \]
263:
264: The dynamical properties of these transformations are not well
265: understood and this is one of intriguing problems.
266:
267: The examples that appear demonstrate few cases (taken from
268: \cite{bgr:spec,grisunik:hanoi}) leading to easily treatable
269: transformations (when they ``integrable'' in the sense that there
270: is a nontrivial semi-conjugacy to a one-dimensional map), a couple
271: of examples (taken from
272: \cite{zukgrigorchuk:3st,grisavchuksunic:img}) when there is no
273: information about topological nature of the invariant subsets
274: (which according to computer experiments look like ``strange
275: attractors''), and a couple of examples when our method does not
276: work (but the corresponding spectral problem is important because
277: of its relation to the problem of finding new constructions of
278: expanders).
279:
280: In principle, the area of application of our method is much
281: broader but then it requires use of Schur complements in
282: infinite-dimensional spaces of matrices. At the moment we do not
283: have examples of successful applications of the method in the
284: infinite-dimensional case.
285:
286: One of important properties we are after in the study of
287: self-similar groups and related objects is amenability.
288: Self-similar groups provide a number of examples of amenable but
289: not elementary amenable groups
290: \cite{grigorchuk:growth_en,grigorchuk:notEG}. The fundamental idea
291: how to treat amenability of self-similar groups belongs to
292: Bartholdi and Virag~\cite{barthvirag}. Roughly speaking it
293: converts self-similarity of a group into self-similarity of a
294: random walk on it. This idea was further developed by
295: V.Kaimanovich (in what he calls the ``M\"unchhausen trick'') using
296: the notion of entropy of a random
297: walk~\cite{kaimanovich:munchhausen}. Kaimanovich introduced
298: transformations of the random walk under the renorm
299: transformations of the group and successfully used them to show
300: that some self-similar groups are amenable. Our observation is
301: that these transformations again can be interpreted as the Schur
302: complement transformations of the measures (or corresponding
303: elements of the group algebra) determining the random walk. More
304: precisely, we show that the transformation considered
305: in~\cite{kaimanovich:munchhausen} is the Schur complement
306: conjugated by the map $A\mapsto A+I$, where $I$ is the identity
307: matrix.
308:
309: We believe that the introduced ideas of self-similar algebras and
310: Schur complement transformations on them will be useful for the
311: study of different aspects of the theory of self-similar groups
312: and its applications.
313:
314: \medskip\noindent\textbf{Acknowledgements.} The authors are
315: grateful to Peter Kuchment for pointing out that the
316: transformation we are considering is called the Schur complement
317: and to Yaroslav Vorobets for useful remarks.
318:
319: \section{Self-similar groups}
320: \subsection{Definition}
321: Let $\alb$ be a finite alphabet and let $\xs$ denote the set of
322: finite words over the alphabet $\alb$. In other terms, $\xs$ is
323: the free monoid generated by $\alb$. We consider $\xs$ to be the
324: set of vertices of a rooted regular tree with the root coinciding
325: with the empty word $\varnothing$ and in which a word $v$ is
326: connected to every word of the form $vx$ for $x\in\alb$.
327:
328: A rooted tree is a standard model of a self-similar structure. If
329: we remove the empty word from it, then the tree $\xs$ will split
330: into $|\alb|$ subtrees $x\xs$, where $x\in\alb$. Each of the
331: subtrees $x\xs$ is isomorphic to the whole tree $\xs$ under the
332: isomorphism $xv\mapsto v$, see Figure~\ref{fig:tree}.
333:
334: \begin{figure}
335: \includegraphics{tree.eps}\\
336: \caption{Rooted tree}\label{fig:tree}
337: \end{figure}
338:
339:
340: If we consider the \emph{boundary} of the tree $\xs$, then the
341: self-similarity is even more evident. The boundary of a rooted
342: tree is the set of all infinite paths starting in the root. In our
343: case the boundary of $\xs$ is naturally identified with the set
344: $\xo$ of infinite words $x_1x_2\ldots$. The set $\xo$ is a
345: disjoint union of the \emph{cylindrical sets}
346: $x\xo=\{xx_2x_3\ldots\;:\;x_i\in\alb\}$, and again, the shift
347: $xw\mapsto w$ is a bijection (and a homeomorphism, if we endow
348: $\xo$ with the natural topology of a direct product of discrete
349: sets $\alb$).
350:
351: A group acting on the rooted tree $\xs$ is called self-similar, if
352: the action agrees with the described self-similarity structure on
353: the tree $\xs$. Namely, we adopt the following definition.
354:
355: \begin{definition}
356: A \emph{self-similar group} $(G, \alb)$ is a group $G$ acting
357: faithfully on the rooted tree $\xs$ such that for every $g\in G$
358: and every $x\in\alb$ there exist $h\in G$ and $y\in\alb$ such that
359: \[g(xw)=yh(w)\]
360: for all $w\in\xs$.
361: \end{definition}
362:
363: It follows from the definition that if $(G, \alb)$ is a
364: self-similar group, then for every $v\in\xs$ there exists $h\in G$
365: uniquely defined by the condition
366: \[g(vw)=g(v)h(w)\]
367: for all $w\in\xs$. The element $h$ is called \emph{restriction} of
368: $g$ in $v$ and is denoted $h=g|_v$.
369:
370: We have the following obvious properties of restriction
371: \[g|_{v_1v_2}=\left(g|_{v_1}\right)|_{v_2},\qquad
372: (gh)|_v=g|_{h(v)}h|_v.\]
373:
374: It is convenient to identify a letter $x$ with the \emph{creation
375: operator} on $\xs$ (or on $\xo$) given by appending the letter $x$
376: to the beginning of the word:
377: \[x\cdot v=xv.\]
378:
379: Similarly, we can identify every word $u\in\xs$ with the creation
380: operator
381: \[u\cdot v=uv.\]
382:
383: In this case the condition that $g(vw)=uh(w)$ for all $w\in\xs$
384: can be written as equality of compositions of transformations of
385: $\xs$:
386: \[g\cdot v=u\cdot h.\]
387:
388: We have the following straightforward corollary of the
389: definitions.
390:
391: \begin{proposition}
392: The set $\alb\cdot G$ of transformations of $\xs$ of the form
393: $x\cdot g:w\mapsto xg(w)$ is closed under pre- and
394: post-compositions with action of the elements of $G$. The obtained
395: left and right actions of $G$ on $\alb\cdot G$ commute and are
396: defined by
397: \[h\cdot (x\cdot g)=h(x)\cdot (h|_xg),\qquad (x\cdot g)\cdot
398: h=x\cdot (gh),\] where dot denotes composition of transformations.
399: \end{proposition}
400:
401: We call the set $\alb\cdot G$ the \emph{(permutational)
402: $G$-bimodule associated with the self-similar group $(G, \alb)$}.
403:
404: \begin{definition}
405: \label{def:replicating} A self-similar group $(G, \alb)$ is called
406: \emph{self-replicating} (\emph{recurrent} in~\cite{nek:book}) if
407: it is transitive on the first level $\alb^1$ of the tree $\xs$ and
408: for any (and thus for every) $x\in\alb$ the map $g\mapsto g|_x$
409: from the stabilizer of $x$ in $G$ to $G$ is onto. Equivalently,
410: the self-similar action is self-replicating if the left action of
411: $G$ on the bimodule $\alb\cdot G$ is transitive.
412: \end{definition}
413:
414: It is not hard to prove by induction that if an action is
415: self-replicating, then it is transitive on every level $\alb^n$ of
416: the tree $\xs$ (is \emph{level-transitive}).
417:
418: Every self-similar action naturally induces an action on the
419: boundary $\xo$ of the tree $\xs$. This is an action by
420: measure-preserving homeomorphisms.
421:
422: \begin{example} \textbf{Grigorchuk
423: group.}\label{sss:grigorchuk} Consider the binary alphabet
424: $\alb=\{0, 1\}$, the tree $\xs$ and let $\mathfrak{G}$ be the
425: group generated by the automorphisms $a, b, c, d$ of $\xs$ defined
426: recursively by the relations
427: \begin{alignat*}{2}
428: a(0w) &= 1w, &\quad a(1w) &=0w\\
429: b(0w) &= 0a(w), &\quad b(1w) &= 1c(w)\\
430: c(0w) &= 0a(w), &\quad c(1w) &= 1d(w)\\
431: d(0w) &= 0w, &\quad d(1w) &= 1b(w).
432: \end{alignat*}
433:
434: \begin{figure}
435: \includegraphics{grig.eps}\\
436: \caption{Automaton generating the Grigorchuk group}\label{fig:grig}
437: \end{figure}
438:
439:
440: This group was defined for the first time
441: in~\cite{grigorchuk:80_en}. This group is a particularly easy
442: example of a Burnside group (an infinite finitely generated
443: torsion group) and it is the first example of a group of
444: intermediate growth, which answers a question of J.~Milnor.
445: Figure~\ref{fig:grig} shows the Moore diagram of the automaton
446: generating the Grigorchuk group (see a definition of Moore
447: diagrams below).
448: \end{example}
449:
450: \begin{example}\textbf{Free group.} \label{sss:free} A convenient way
451: to define self-similar groups are Moore diagrams of automata,
452: generating them. Consider, for instance, the Moore diagram shown
453: on Figure~\ref{fig:alfree}. The vertices of the diagram correspond
454: to states of the automaton (to generators of the group), the
455: arrows describe transitions between the states and the labels
456: correspond to the output of the automaton. If there is an arrow
457: from a state $g$ to a state $h$ labeled by $(x, y)$, then this
458: means that $g(xw)=yh(w)$ for all $w\in\xs$. Thus, the automaton
459: shown on Figure~\ref{fig:alfree} describes the group generated by
460: the elements $a, b, c$ such that
461: \begin{alignat*}{2}
462: a(0w) &= 0b(w), &\quad a(1w)=1b(w),\\
463: b(0w) &= 1a(w), &\quad b(1w)=0c(w),\\
464: c(0w) &= 1c(w), &\quad c(1w)=0a(w),
465: \end{alignat*}
466: for all $w\in\xs$.
467:
468: \begin{figure}
469: \includegraphics{alfree.eps}\\
470: \caption{Automaton generating a free group}\label{fig:alfree}
471: \end{figure}
472:
473:
474: S.~Sidki conjectured in~\cite{sid:cycl} that the three states of
475: the automaton generate a free group of rank 3. This claim was
476: later proved by Y.~Vorobets and M.~Vorobets
477: in~\cite{vorobets:alfree}.
478: \end{example}
479:
480: \begin{example}\textbf{Basilica.} \label{sss:basilica} The group
481: generated by the automaton shown on Figure~\ref{fig:basilica} is
482: called the \emph{Basilica group}, since it is the iterated
483: monodromy group of the polynomial $z^2-1$ (see below for the
484: definition).
485:
486: \begin{figure}
487: \includegraphics{basilica.eps}\\
488: \caption{Automaton generating $\img{z^2-1}$}\label{fig:basilica}
489: \end{figure}
490:
491:
492: It is the first example of an amenable group, which can not be
493: constructed from the groups of sub-exponential growth using the
494: operations preserving amenability (taking quotients, extensions,
495: direct limits and passing to a subgroup).
496: \end{example}
497:
498: \subsection{Iterated monodromy groups} Let $\M$ be a path
499: connected and locally path connected topological space and let
500: $\M_1$ be its open subset. A \emph{$d$-fold partial self-covering}
501: is a covering map $f:\M_1\arr\M$, i.e., a continuous map such that
502: for every $x\in\M$ there exists a neighborhood $U\ni x$ whose
503: total preimage $f^{-1}(U)$ is a disjoint union of $d$ open subsets
504: which are mapped homeomorphically onto $U$ by $f$.
505:
506: For every $n\ge 1$ the iteration $f^n:\M_n\arr\M$ is a $d^n$-fold
507: partial self-covering (in general with a smaller domain $\M_n$).
508:
509: Choose a basepoint $t\in\M$. Then the disjoint union
510: $T=\bigsqcup_{n\ge 0}f^{-n}(t)$ has a natural structure of a
511: $d$-regular tree with the root $t\in\{t\}=f^{-0}(t)$ where a
512: vertex $z\in f^{-n}(t)$ is connected to the vertex $f(z)\in
513: f^{-(n-1)}(t)$.
514:
515: The fundamental group $\pi_1(\M, t)$ acts naturally on every level
516: $f^{-n}(t)$ of the tree $T$. The image of a vertex $z$ under the
517: action of a loop $\gamma\in\pi_1(\M, t)$ is the end of the unique
518: preimage of $\gamma$ under the covering $f^n$, which starts at
519: $z$. It is easy to check that these actions define an action of
520: $\pi_1(\M, t)$ by automorphisms of the rooted tree $T$. This
521: action is called the \emph{iterated monodromy action}.
522:
523: \begin{definition}
524: The \emph{iterated monodromy group} $\img{f}$ of a partial
525: self-covering $f:\M_1\arr\M$ is the quotient of the fundamental
526: group $\pi_1(\M, t)$ by the kernel of the iterated monodromy
527: action.
528: \end{definition}
529:
530: The iterated monodromy group is self-similar, if we identify the
531: tree of preimages $T$ with $\xs$ in a correct way
532: (see~\cite{nek:book,bgn}. The obtained self-similar action (called
533: the \emph{standard action}) is described in the following way.
534: Choose a bijection $\Lambda:\alb\arr f^{-1}(t)$ between the
535: alphabet $\alb$ of size $d$ and the set of preimages of the
536: basepoint. Choose also paths $\ell_x$ from the basepoint $t$ to
537: its preimage $\Lambda(x)$ for every $x$. Then the standard action
538: of $\img{f}$ on $\xs$ (which is conjugate to its action on $T$) is
539: given by the formula
540: \begin{equation}
541: \label{eq:imgstandard} \gamma(xw)=y(\ell_x\gamma_x\ell_y^{-1})(w),
542: \end{equation}
543: where $\gamma_x$ is the $f$-preimage of $\gamma$ starting in
544: $\Lambda(x)$, $y\in\alb$ is such that $\Lambda(y)$ is the end of
545: $\gamma_x$ and $w\in\xs$ is any word. We multiply here the paths
546: in the natural order (see Figure~\ref{fig:recur}).
547:
548: \begin{figure}
549: \includegraphics{recur.eps}\\
550: \caption{Self-similarity of iterated monodromy groups}\label{fig:recur}
551: \end{figure}
552:
553:
554: \section{Self-similar algebras}
555:
556: \subsection{Self-similarities of Hilbert spaces}
557:
558: A ($d$-\emph{fold}) \emph{similarity} of an infinite-dimensional
559: Hilbert space $H$ is an isomorphism
560: \[\psi:H\arr H^d=\underbrace{H\oplus\cdots\oplus H}_{d}.\]
561:
562: \begin{example}
563: Let $\alb$ be an alphabet with $d$ letters and let $\xs$ and $\xo$
564: be the rooted tree and its boundary, respectively. Let $\nu$ be
565: the uniform Bernoulli measure on $\xo$, i.e., the direct product
566: of uniform probability measures on $\alb$. Then the Hilbert space
567: $H=L^2(\xo, \nu)$ is decomposed into the direct sum
568: $\bigoplus_{x\in\alb}L^2(x\xo)$, where $L^2(x\xo)$ is the subspace
569: of functions with support a subset of the cylindrical set $x\xo$.
570: But the spaces $L^2(x\xo)$ are naturally isomorphic to $L^2(\xo,
571: \nu)$, where the isomorphism is the map $U_x:L^2(x\xo)\arr
572: L^2(\xo, \nu)$ given by
573: \[U_x(f)(w)=\frac 1{\sqrt{d}} f(xw).\] We view $U_x$ as
574: partial isometries of $L^2(\xo, \nu)$. Hence we get the natural
575: similarity $\sum_{x\in\alb} U_x$ of the space $L^2(\xo, \nu)$.
576: \end{example}
577:
578: \begin{example}
579: Let $W\subset\xo$ be a \emph{self-similar} subset of the boundary
580: of the tree $\xs$, i.e., such a set that $W=\bigcup_{x\in\alb}xW$.
581: Then the space $\ell^2(W)$ is naturally self-similar, since it can
582: be decomposed into a direct sum
583: \[\ell^2(W)=\bigoplus_{x\in\alb}\ell^2(xW),\]
584: where the spaces $\ell^2(xW)$ are naturally isomorphic to
585: $\ell^2(W)$, with the isomorphism $U_x:\ell^2(xW)\arr\ell^2(W)$
586: given by
587: \[U_x(f)(w)=f(xw).\]
588: \end{example}
589:
590: The above two examples will be the main types of self-similarities
591: on Hilbert spaces used in this paper.
592:
593: \subsection{Representations of the Cuntz algebra $O_d$}
594:
595: Recall that the \emph{Cuntz algebra} $O_d$, $d\ge 2$, is the
596: $C^*$-algebra given by the presentation
597: \[\langle a_1, a_2, \ldots, a_d\;:\;a_1a_1^*+a_2a_2^*+\cdots+a_da_d^*=1,\;a_k^*a_k=1,
598: k=1,\ldots, d\rangle.\] Note that as a corollary of the defining
599: relations we get $a_k^*a_l=0$ for $k\ne l$. The defining relations
600: can be also written as the following two matrix equalities
601: \[(a_1, a_2, \ldots, a_d)(a_1, a_2, \ldots, a_d)^*=1,\quad
602: (a_1, a_2, \ldots, a_d)^*(a_1, a_2, \ldots, a_d)=I,\] where $I$ is
603: the $d\times d$ unit matrix.
604:
605: The Cuntz algebra $O_d$ is simple (see~\cite{cuntz}) and hence any
606: $d$ isometries $a_1, \ldots, a_d$ such that $a_1a_1^*+\cdots
607: +a_da_d^*=1$ generate a $C^*$-algebra isomorphic to the Cuntz
608: algebra $O_d$ and determine its representation.
609:
610: Representations of the Cuntz algebra can be identified with
611: self-similarities of a Hilbert space as the following proposition
612: shows.
613:
614: \begin{proposition}
615: \label{pr:cuntzrepr} The relation putting into correspondence to a
616: $*$-representation $\rho:O_d\arr B(H)$ the map
617: \[\tau_\rho=(\rho(a_1^*), \rho(a_2^*),
618: \ldots, \rho(a_d^*)):H\arr H^d\] is a bijection between the set of
619: representations of $O_d$ on $H$ and the set of $d$-fold
620: self-similarities on $H$.
621:
622: The inverse of this bijection puts into correspondence to a
623: $d$-similarity $\psi:H\arr H^d$ the representation of $O_d$ given
624: by $\rho(a_k)=T_k$, for
625: \[T_k(\xi)=\psi^{-1}(0, \ldots, 0, \xi, 0, \ldots, 0),\]
626: where $\xi$ in the right-hand side is at the $k$th coordinate of
627: $H^d$.
628: \end{proposition}
629:
630: \begin{proof}
631: If $\rho$ is a representation of $O_d$ on $H$, then $\rho(a_k)$
632: are isometries of $H$ with the subspaces $H_k=\rho(a_k)(H)$ and
633: $H=H_1\oplus \cdots \oplus H_d$. The $k$th components of a vector
634: $\xi\in H$ with respect to this decomposition is its image under
635: the projection $\rho(a_ka_k^*)$. Hence we get an isomorphism
636: \[H\arr H^d:\xi\mapsto (\rho(a_1)^*(\xi), \rho(a_2)^*(\xi), \ldots,
637: \rho(a_d)^*(\xi))\] equal to the composition of the identical map
638: $H\arr H_1\oplus\cdots\oplus H_d$ with the isomorphism
639: $\rho(a_1)^*\oplus\cdots\oplus\rho(a_d)^*:H_1\oplus\cdots\oplus
640: H_d\arr H^d$.
641:
642: Conversely, suppose that
643: \[\psi:H\arr H^d\]
644: is a $d$-similarity. The map
645: \[\xi\mapsto(0, \ldots, 0, \xi, 0, \ldots 0)\]
646: is an isometry of $H$ with the $k$th direct summand of $H^d$.
647: Composing it with the isomorphism $\psi^{-1}$ we get an isometry
648: $T_k$ of $H$ with the direct summand $H_k$ of a decomposition
649: $H=H_1\oplus\cdots\oplus H_d$. But tuples of such isometries are
650: precisely the representations of the Cuntz algebra $O_d$. Note
651: that if $\psi(\xi)=(\xi_1, \ldots, \xi_d)$, then $\xi=\sum_k
652: T_k(\xi_k)$, hence $\xi_k=T_k^*(\xi)$. Consequently, the two
653: bijections are mutually inverse.
654: \end{proof}
655:
656: \begin{example}\label{ex:murepr}
657: The representation of $O_d$ associated with the natural
658: $d$-similarity of $L^2(\xo, \nu)$ is generated by the isometries
659: $\pi(a_x)=T_x$ on $L^2(\xo, \nu)$, given by
660: \[T_x(f)(w)=\left\{\begin{array}{ll}\sqrt{d}f(w') &\text{if $w=xw'$}\\
661: 0 & \text{otherwise.}\end{array}\right.\]
662: \end{example}
663:
664: \begin{example}
665: Let $W$ be a self-similar subset of $\xo$. The representation
666: associated to the natural $d$-similarity on $\ell^2(W)$ is
667: generated by the isometries
668: \[T_x(f)(w)=\left\{\begin{array}{ll}f(w') &\text{if $w=xw'$}\\
669: 0 & \text{otherwise.}\end{array}\right.\] Such representations of
670: $O_d$ are called \emph{permutational}. Permutational
671: representations of the Cuntz algebra related to self-affine
672: (digit) tilings of the Euclidean space are studied
673: in~\cite{cuntz_rep}.
674: \end{example}
675:
676: \subsection{Self-similar groups and their representations}
677: \label{ss:wrrecurs} If $(G, \alb)$ is a self-similar group, then
678: the associated \emph{wreath recursion} is the embedding
679: $\phi:G\arr\symm\wr_\alb G=\symm\ltimes G^\alb$ given by
680: \[\phi(g)=\sigma(g|_x)_{x\in\alb},\]
681: where $\sigma\in\symm$ is the action $x\mapsto g(x)$ of $g$ on the
682: first level $\alb$ of the tree $\xs$ and the components $g|_x$ of
683: $G^\alb$ are the restrictions of $g$ onto the subtrees $x\xs$,
684: i.e., are given by the condition
685: \[g(xw)=g(x)g|_x(w)\]
686: for all $w\in\xs$.
687:
688: Suppose that we have a self-similar group $(G, \alb)$ and let
689: $d=|\alb|$. Let $H$ be a Hilbert space together with a
690: $d$-similarity
691: \[\psi:H\arr H^\alb.\]
692: We will denote the summand of $H^\alb$ corresponding to a letter
693: $x\in\alb$ by $H_x$. Let $\rho:O_d\arr B(H)$ be the associated
694: representation of the Cuntz algebra. It is generated by isometries
695: $T_x=\rho(a_x)$, $x\in\alb$, such that $H_x=T_x(H)$.
696:
697: \begin{definition}
698: A unitary representation $\rho$ of $G$ on $H$ is said to be
699: \emph{self-similar} (with respect to the $d$-similarity $\psi$) if
700: \[\rho(g)T_x=T_y\rho(h)\]
701: whenever $g(xw)=yh(w)$ for all $w\in\xs$, i.e., whenever $g\cdot
702: x=y\cdot h$ in the associated bimodule $\alb\cdot G$.
703: \end{definition}
704:
705: \begin{example}
706: A self-similar group $G$ acts on the tree $\xs$ by automorphisms
707: and the induced action on the boundary $\xo$ is preserving the
708: Bernoulli measure $\nu$. We get hence a unitary representation
709: $\pi$ of $G$ on $L^2(\xo, \nu)$. It is easy to see that this
710: representation is self-similar with respect to the natural
711: $d$-similarity on $L^2(\xo, \nu)$.
712: \end{example}
713:
714: \begin{example}
715: Let $W$ be a self-similar $G$-invariant subset of $\xo$. Then the
716: permutational representation of $G$ on $W$ is self-similar.
717: \end{example}
718:
719: For the general notion of a Cuntz-Pimsner algebra
720: see~\cite{pimsner}.
721:
722: \begin{definition}
723: Let $G$ be a self-similar group acing on $\xs$. The
724: \emph{associated (universal) Cuntz-Pimsner algebra} $\cpg$ is the
725: universal $C^*$-algebra generated by $G$ and $a_x$, $x\in\alb$,
726: satisfying the following relations
727: \begin{itemize}
728: \item all relations of $G$;
729: \item Cuntz relations for $a_x$: $a_x^*a_x=1$ for all $x\in\alb$,
730: $\sum_{x\in\alb}a_xa_x^*=1$;
731: \item $ga_x=a_yh$ for $g, h\in G$ and $x, y\in\alb$,
732: if $g(xw)=yh(w)$ for all $w\in\xs$, i.e., if $g(x)=y$ and
733: $h=g|_x$.
734: \end{itemize}
735:
736: The algebra generated by $G$ in $\cpg$ is denoted $\maxalg$.
737: \end{definition}
738:
739: Note that as a corollary of the defining relations we get the
740: relations
741: \begin{equation}\label{eq:splitting}g=g\sum_{x\in\alb}a_xa_x^*=
742: \sum_{x\in\alb}a_{g(x)}g|_xa_x^*,\end{equation} for
743: every $g\in G$, where $g|_x$ is, as usual, the section of $g$ at
744: $x$, i.e., such an element of $G$ that $g(xw)=g(x)g|_x(w)$ for all
745: $w\in\xs$.
746:
747:
748: The next proposition follows directly from the definitions.
749:
750: \begin{proposition}
751: \label{pr:cpimsner} Let $\pi:O_d=\langle a_x\rangle_{x\in\alb}\arr
752: B(H)$ be a representation of the Cuntz algebra associated with a
753: $d$-similarity of a Hilbert space $H$.
754:
755: A unitary representation $\rho$ of $G$ on $H$ is self-similar if
756: and only if $\rho$ and $\pi$ generate a representation of the
757: Cuntz-Pimsner algebra $\cpg$.
758:
759: Consequently, self-similar representations of $G$ are precisely
760: restrictions onto $G$ of representations of the Cuntz-Pimsner
761: algebra $\cpg$.
762: \end{proposition}
763:
764:
765:
766:
767: \subsection{Matrix recursions}
768: A \emph{matrix recursion} on an algebra $A$ is a homomorphism
769: \[\phi:A\arr M_{d\times d}(A)\]
770: of $A$ into the algebra of matrices over $A$.
771:
772: \begin{example}
773: Let $\psi:H\arr H^d=H^\alb$ be a $d$-similarity on a Hilbert space
774: $H$ and let $\rho:G\arr B(H)$ be a self-similar unitary
775: representation of a group $G$. Then every operator $\rho(g)$ for
776: $g\in G$ can be written, with respect to the decomposition
777: $\psi(H)=H^\alb$, as a $d\times d$ matrix
778: \[\rho(g)=\left(A_{yx}\right)_{x, y\in\alb},\]
779: where \[A_{yx}=\left\{\begin{array}{ll}\rho(g|_x) & \text{if $g(x)=y$,}\\
780: 0 & \text{otherwise.}\end{array}\right.\]
781: \end{example}
782:
783: For every self-similar group $G$ we have the associated matrix
784: recursion on the group algebra $\C[G]$, which is the linear
785: extension of the recursion:
786: \begin{equation}\label{eq:matrrecursion}\phi(g)=\left(A_{yx}\right)_{x,y\in\alb},\quad A_{yx}=\left\{\begin{array}{ll}g|_x & \text{if $g(x)=y$,}\\
787: 0 & \text{otherwise,}\end{array}\right.\end{equation} which can be
788: interpreted as the \emph{wreath recursion} $\phi:G\arr
789: \symm\wr_\alb G$, as it was defined in Section~\ref{ss:wrrecurs}.
790:
791: In terms of the associated representation $\rho$ of the Cuntz
792: algebra, we have
793: \[g|_x=T_{g(x)}^*gT_x,\]
794: where $T_x=\rho(a_x)$, which follows from~\eqref{eq:splitting} (or
795: directly from the defining relations of $\cpg$).
796:
797: Note that the homomorphism $\phi$ usually is not injective (even
798: if the wreath recursion is and thus the map $\phi:G\arr
799: M_d\left(\C[G]\right)$ is injective as well).
800:
801: \begin{example}
802: Consider the group $\mathfrak{G}=\langle a, b, c, d\rangle$ from
803: Example~\ref{sss:grigorchuk}. Then
804: \[\phi(a)=\left(\begin{array}{cc} 0 & 1 \\ 1 & 0\end{array}\right),
805: \phi(b)=\left(\begin{array}{cc} a & 0 \\ 0 & c\end{array}\right),
806: \phi(c)=\left(\begin{array}{cc} a & 0 \\ 0
807: & d\end{array}\right), \phi(d)=\left(\begin{array}{cc} 1 & 0 \\
808: 0 & b\end{array}\right).\]
809:
810: Denote $\alpha=(b+c+d-1)/2$. Then $\phi(\alpha)=\left(\begin{array}{cc} a & 0\\
811: 0 & \alpha\end{array}\right)$,
812: $\phi(\alpha^2-1)=\left(\begin{array}{cc} 0 & 0\\ 0 &
813: \alpha^2-1\end{array}\right)$,
814: $\phi(a(\alpha^2-1)a)=\left(\begin{array}{cc} \alpha^2-1 & 0\\
815: 0 & 0\end{array}\right)$, and
816: \[\phi\left((\alpha^2-1)a(\alpha^2-1)a\right)=0.\]
817:
818: But
819: \[(\alpha^2-1)a(\alpha^2-1)a=\frac{(b+c+d-4)a(b+c+d-4)a}{16}\ne 0\]
820: in $\C[G]$.
821: \end{example}
822:
823: One can find however an ideal $I\subset \C[G]$ such that $\phi$
824: induces an injective homomorphism $\C[G]/I\arr M_{d\times
825: d}(\C[G]/I)$, see~\cite{sid:ring,nek:iterat}. The ideal $I$ is the
826: ascending union of the kernels of the matrix recursions $\C[G]\arr
827: M_{d^n\times d^n}(\C[G])$ describing the action of the group $G$
828: on the $n$th level of the tree $\xs$.
829:
830:
831: \subsection{Self-similar completions of the group algebra}
832: Let $\rho:G\arr B(H)$ be a self-similar representation (with
833: respect to a $d$-similarity $\psi:H\arr H^d$) and let
834: $\mathcal{A}_\rho$ be the completion of $\C[G]$ with respect to
835: the norm given by $\rho$. Then the matrix recursion $\phi$ extends
836: to a homomorphism
837: \[\mathcal{A}_\rho\arr M_{d\times d}(\A_\rho),\]
838: also denoted by $\phi$ (or $\phi_\rho$), which is injective, since
839: it implements the equivalence of the representation $\rho$ with
840: the representation $\psi\circ\rho\circ\psi^{-1}$.
841:
842: The following description of the completions $\mathcal{A}_\rho$
843: follows directly from Proposition~\ref{pr:cpimsner}.
844:
845: \begin{definition}
846: A completion of $\C[G]$ is called \emph{self-similar} if it is the
847: completion with respect to a self-similar representation.
848: \end{definition}
849:
850: The following proposition shows that there is a unique maximal
851: completion. We denote by $\maxalg$ the $C^*$-algebra generated by
852: $G$ in $\cpg$.
853:
854: \begin{proposition}
855: \label{pr:algmaximal} A completion $\mathcal{A}$ of $\C[G]$ is
856: self-similar if and only if it is the closure of $\C[G]$ in a
857: homomorphic image of the Cuntz-Pimsner algebra $\cpg$. In
858: particular, every such completion is a homomorphic image of the
859: algebra $\maxalg$.
860: \end{proposition}
861:
862: There also exists the smallest self-similar completion.
863:
864: \begin{definition}\label{def:Ggeneric}
865: Let $G$ be a countable group of automorphisms of $\xs$. A point
866: $w\in\xo$ of the boundary is called \emph{$G$-generic} if for
867: every $g\in G$ either $g(w)\ne w$ or $w$ is fixed by $g$ together
868: with all points of a neighborhood of $w$.
869: \end{definition}
870:
871: It is not hard to prove that the set of $G$-generic points is
872: co-meager (i.e., is an intersection of a countable collection of
873: open dense sets), see~\cite{grineksu_en} and~\cite{nek:bim}. Let
874: $W\subset\xo$ be a non-empty countable $G$-invariant set of
875: $G$-generic points (one can also just take the $G$-orbit of a
876: $G$-generic point). Denote by $\rho_W$ the permutational
877: representation of $G$ on $\ell^2(W)$. The following theorem is
878: proved in~\cite{nek:bim}.
879:
880: \begin{theorem}\label{th:minimal} Let $G$ be a group with a self-similar action on
881: $\xs$. Denote by $\|\cdot\|_{min}$ the norm on $\C[G]$ defined by
882: the representation $\rho_W$. This norm does not depend on the
883: choice of the set $W$ and $\|a\|_{min}\le\|a\|_{\rho}$ for any
884: self-similar representation $\rho$ of $G$ and any $a\in\C[G]$.
885:
886: The completion of $\C[G]$ with respect to the norm
887: $\|\cdot\|_{min}$ is the algebra $\minalg$ generated by $G$ in a
888: unique simple unital quotient of $\cpg$.
889: \end{theorem}
890:
891: Hence, if $\mathcal{A}$ is a completion of $\C[G]$ with respect to
892: a self-similar representation of $G$, then the identical map on
893: $G$ induces surjective homomorphisms
894: \[\maxalg\arr\mathcal{A}\arr\minalg.\]
895:
896: Let $\mathfrak{O}_G$ be the groupoid generated by the germs of the
897: local homeomorphisms of $\xo$ of the form
898: \begin{equation}\label{eq:tvg}T_vg:w\mapsto vg(w),\quad v\in\xs, g\in G.\end{equation}
899: Recall that a germ of a local homeomorphism $h:U\arr V$ is a pair
900: $(h, x)$, where $x$ belongs to the domain of $h$, where two pairs
901: $(h_1, x_1)$ and $(h_2, x_2)$ are identified if $x_1=x_2$ and
902: restrictions of $h_i$ on a neighborhood of $x_1$ coincide. The
903: germs are composed in a natural way and inverse of a germ $(g, x)$
904: is defined to be equal to $(g^{-1}, g(x))$. The set of germs of a
905: pseudogroup of local homeomorphisms has a natural \emph{germ
906: topology} defined by the basis consisting of the sets of the form
907: $\mathcal{U}_{h, U}=\{(h, x)\;:\;x\in U\}$ where $h$ is an element
908: of the pseudogroup and $U$ is an open subset of the domain of $h$.
909:
910: For more details and for the definition of the operator algebras
911: associated with topological groupoids,
912: see~\cite{renault:groupoids,bridhaefl,paterson:gr}.
913:
914: It is not hard to prove that the universal Cuntz-Pimsner algebra
915: $\cpg$ coincides with the universal algebra of the groupoid
916: $\mathfrak{O}_G$ (see~\cite{nek:cpalg}). One can consider also the
917: \emph{reduced $C^*$-algebra} of $\mathfrak{O}_G$ (see the
918: definitions in~\cite{renault:groupoids,khoshkamskand}). Let
919: $\A_{red}$ be the subalgebra of the reduced $C^*$-algebra of
920: $\mathfrak{O}_G$ generated by $G$. Then $\A_{red}$ is also a
921: completion of $G$ with respect to a self-similar representation
922: (since it comes from a representation of $\cpg$, see
923: Proposition~\ref{pr:cpimsner}).
924:
925: A groupoid of germs of an action of a group $G$ on a topological
926: space $\mathcal{X}$ is non-Hausdorff if and only if there exists
927: $x\in\mathcal{X}$ and an element $g\in G$ such that the germ of
928: $g$ at $x$ can not be separated from the germ of the identity at
929: $x$. The latter is equivalent to the condition that for every
930: neighborhood $U$ of $x$ there exists $y\in U$ such that the germ
931: of $g$ at $y$ is trivial (in particular $g(x)=x$) and there exists
932: $z\in U$ such that $g(z)\ne z$. Consequently, the groupoid of
933: germs is Hausdorff if and only if for every $g\in G$ the interior
934: of the set of fixed points of $g$ is closed. See an example of a
935: self-similar group with non-Hausdorff groupoid of germs of the
936: action on $\xo$ in Example~\ref{ex:nonHausd}.
937:
938: The following theorem follows from classical results of theory of
939: amenable groupoids
940: (see~\cite{renault:groupoids,delaroche_renault}). The last
941: paragraph of the theorem follows directly from the definitions of
942: the corresponding algebras.
943:
944: \begin{theorem}
945: \label{th:amenablegr} If the groupoid of germs of the action of
946: $G$ on $\xo$ is (measurewise) amenable (in particular, if the
947: orbits of the action of $G$ on $\xo$ have polynomial growth), then
948: the universal and reduced algebras of the groupoid
949: $\mathfrak{O}_G$ coincide. In particular, their sub-algebras
950: $\maxalg$ and $\A_{red}$ coincide.
951:
952: If the groupoid of germs of the action of $G$ on $\xo$ is
953: Hausdorff, then the algebras $\A_{red}$ and $\minalg$ coincide.
954: \end{theorem}
955:
956: In particular, if the action of $G$ on $\xo$ is free, then the
957: groupoid of germs is Hausdorff and moreover is \emph{principal}
958: (i.e., is an equivalence relation) and $\A_{red}=\minalg$.
959:
960: Another frequently used self-similar completion is the completion
961: $\A_{mes}$ defined by the natural unitary representation $\pi$ of
962: $G$ on $L^2(\xo, \nu)$.
963:
964: \begin{question} Find conditions when $\A_{mes}=\minalg$.\end{question}
965:
966: \subsection{Gauge-invariant subalgebra of $\cpg$}
967: Let $\M_k$ be the closed linear span in $\cpg$ of the elements
968: $a_vga_u^*$ for $g\in G$ and $v, u\in\alb^k$. Here we use the
969: multi-index notation $a_{x_1x_2\ldots x_n}=a_{x_1}a_{x_2}\cdots
970: a_{x_n}$. In particular, $\M_0=\maxalg$ is the algebra generated
971: by $G$ in $\cpg$.
972:
973: It is easy to see that for $v_1, v_2, u_1, u_2\in\alb^k$ and $g_1,
974: g_2\in G$
975: \[a_{v_1}g_1a_{u_1}^*a_{v_2}g_2a_{u_2}^*=\left\{\begin{array}{cr}a_{v_1}g_1g_2a_{u_2}^*
976: & \text{if $u_1=v_2$}\\
977: 0 & \text{otherwise.}\end{array}\right.\] Observe that $a_va_u^*$
978: are multiplied in the same way as the matrix units, hence $\M_k$
979: is isomorphic to the algebra $M_{d^k\times d^k}(\maxalg)$ of
980: $d^k\times d^k$-matrices over $\maxalg$.
981:
982: Recall that by~\eqref{eq:splitting} every element $g\in G$ can be
983: written in $\cpg$ as a sum
984: \[
985: g=\sum_{x\in\alb}a_{g(x)}g|_x a_x^*.
986: \]
987: If we apply this formula to the element $g$ in $a_vga_u^*\in\M_k$,
988: we get an element of $\M_{k+1}$. Hence $\M_{k+1}\supset\M_k$.
989:
990: The algebra $\gi$ is defined as the closure in $\cpg$ of the
991: ascending union $\bigcup_{k\ge 0}\M_k$.
992:
993: Every algebra $\M_k$ contains the sub-algebra equal to the linear
994: span of the elements $a_va_u^*$, which is isomorphic to
995: $M_{d^k\times d^k}(\C)$. Their union (i.e., the closed linear span
996: of all elements $a_va_u^*$ for $|v|=|u|$) is isomorphic to the
997: Glimm's uniformly hyperfinite algebra $M_{d^\infty}(\C)$
998: (see~\cite{glimm,davidson}), since the expansion
999: rule~\eqref{eq:splitting} for $g=1$ defines the diagonal
1000: embeddings $M_{d^k\times d^k}\hookrightarrow M_{d^{k+1}\times
1001: d^{k+1}}$. Its diagonal subalgebra, generated by the projections
1002: $a_va_v^*$ is isomorphic to the algebra $C(\xo)$ of continuous
1003: functions on $\xo$. The isomorphism identifies a projection
1004: $a_va_v^*$ with the characteristic function of the cylindrical set
1005: $v\xo$. It is easy to see that the action of $G\subset\gi$ on the
1006: diagonal algebra by conjugation coincides with the action by
1007: conjugation of $G$ on $C(\xo)$.
1008:
1009: It is proved in~\cite{nek:bim} that if the group $(G, \alb)$ is
1010: self-replicating, then the subalgebra $\gi$ of $\cpg$ is generated
1011: by $G$ and the subalgebra $C(\xo)$, i.e., that $\gi$ is the
1012: cross-product of $\maxalg$ and the algebra of continuous functions
1013: on $\xo$ induced by the usual action of $G$ on $\xo$.
1014:
1015: Similarly to the Cuntz algebra (see~\cite{cuntz,davidson}), we
1016: have a natural strongly continuous (\emph{gauge}) action $\Gamma$
1017: of the circle $\mathbb{T}=\{z\in\C: |z|=1\}$ on $\cpg$ by
1018: \begin{eqnarray*}
1019: \Gamma_z(g) & = & g \\
1020: \Gamma_z(a_x) & = & za_x.
1021: \end{eqnarray*}
1022: for $g\in G, x\in\alb$ and $z\in\mathbb{T}$.
1023:
1024: Then $\Gamma_z(a_vga_u^*)=z^{|v|-|u|}a_vga_u^*$ for $u, v\in\xs$, $g\in G$,
1025: thus the integral $\int\Gamma_z(a_vga_u^*)dz$ is equal to zero for
1026: $|v|\neq |u|$ and to $a_vga_u^*$ for $|v|=|u|$, where $dz$ is the
1027: normalized Lebesgue measure on the circle. Consequently the map
1028: \begin{equation}
1029: \label{eq:exp} \mathbb{M}_0(a)=\int\Gamma_z(a)dz
1030: \end{equation}
1031: for $a\in\cpg$ is a conditional expectation from $\cpg$ onto the
1032: subalgebra $\gi$.
1033:
1034: \begin{example}
1035: If $G$ is the cyclic group generated by the adding machine
1036: $a=\sigma(1, a)$, then the algebra $\maxalg=\minalg$ is isomorphic
1037: to the algebra $C(\mathbb{T})$ with the linear recursion
1038: $C(\mathbb{T})\arr M_2(C(\mathbb{T}))$ coming from the double
1039: self-covering of the circle (if we identify $C(\mathbb{T})$ with
1040: $C^*(\Z)$ via Fourier series, then this linear recursion is given
1041: by $z\mapsto\left(\begin{array}{cc} 0 & z\\ 1 &
1042: 0\end{array}\right)$, where $z=e^{2\pi i t}$, $t\in\R$, is the
1043: variable of the Fourier series). Consequently, the algebra $\gi$
1044: of the adding machine action is the Bunce-Deddence algebra. It is
1045: also isomorphic to the cross-product algebra of the odometer
1046: action on the Cantor space $\xo$ (see~\cite{davidson}).
1047: \end{example}
1048:
1049: \subsection{Overview of algebras associated with self-similar groups}
1050:
1051: We studied above the universal Cuntz-Pimsner algebra $\cpg$ of a
1052: self-similar group $G$. Representations of this algebras
1053: correspond to representations of $G$ on a Hilbert space $H$ which
1054: are self-similar with respect to some similarity of $H$ (see
1055: Proposition~\ref{pr:cpimsner}).
1056:
1057: The subalgebra of $\cpg$ generated by $G$ was denoted $\maxalg$.
1058: Any self-similar representation of $G$ extends to a representation
1059: of $\maxalg$, hence $\maxalg$ is the completion of the group
1060: algebra with respect to the maximal self-similar norm.
1061:
1062: Third algebra is the natural inductive limit of the algebras
1063: $M_{d^n\times d^n}(\maxalg)$, which we denoted by
1064: $\M_{d^\infty}(G)$. It is a subalgebra of $\cpg$ in a natural way.
1065:
1066: Hence we get the following tower of algebras associated to a
1067: self-similar group
1068: \[\maxalg\subset\M_{d^\infty}(G)\subset\cpg.\]
1069:
1070: There exits also the smallest self-similar norm on the group
1071: algebra, which corresponds to a unique simple homomorphic image
1072: ${\cpg}_{min}$ of the Cuntz-Pimsner algebra $\cpg$ (see
1073: Theorem~\ref{th:minimal}). The subalgebra of ${\cpg}_{min}$
1074: generated by $G$ is denoted $\minalg$ and it is the completion of
1075: the group algebra with respect to the smallest self-similar norm.
1076: The image $\M_{d^\infty}(G)_{min}$ of $\M_{d^\infty}(G)$ is also
1077: simple.
1078:
1079: For any other self-similar representation $\rho$ of $G$ we get the
1080: corresponding homomorphic image of $\cpg$, $\M_{d^\infty}(G)$ and
1081: $\maxalg$. We get hence the following commutative diagram of
1082: algebras
1083: \[\begin{array}{ccccc}
1084: \maxalg & \hookrightarrow & \M_{d^\infty}(G) & \hookrightarrow & \cpg\\
1085: \mapdown{} & & \mapdown{} & & \mapdown{} \\
1086: \A_\rho & \hookrightarrow & \M_{d^\infty}(G)_\rho & \hookrightarrow & {\cpg}_\rho\\
1087: \mapdown{} & & \mapdown{} & & \mapdown{} \\
1088: \minalg & \hookrightarrow & \M_{d^\infty}(G)_{min} &
1089: \hookrightarrow & {\cpg}_{min}
1090: \end{array}\]
1091: where all vertical arrows are surjective and all horizontal are
1092: embeddings.
1093:
1094: An important example of a self-similar representation is the
1095: natural unitary representation $\pi$ of $G$ on $L^2(\xo, \nu)$,
1096: where $\nu$ is the uniform Bernoulli measure on the boundary $\xo$
1097: of the rooted tree. Together with the natural self-similarity of
1098: the space $L^2(\xo, \nu)$ this gives a representation of $\cpg$.
1099: We denote the respective homomorphic images by
1100: $\A_{mes}\hookrightarrow
1101: \M_{d^\infty}(G)_{mes}\hookrightarrow{\cpg}_{mes}$.
1102:
1103: The algebra $\A_{mes}$ is residually finite-dimensional, since the
1104: representation $\pi$ of $G$ is a direct sum of finite-dimensional
1105: representations (coming from the action of $G$ on the levels of
1106: the tree). These finite-dimensional representations give an
1107: additional tool to the study of the algebra $\A_{mes}$.
1108:
1109: Therefore the next questions about the epimorphisms
1110: $\maxalg\arr\A_{mes}\arr\minalg$ are natural.
1111:
1112: \begin{question}
1113: When are the epimorphisms $\A_{mes}\arr\minalg$ and
1114: $\maxalg\arr\A_{mes}$ isomorphisms?
1115: \end{question}
1116:
1117: \begin{question}
1118: Under which conditions the algebra $\A_{mes}$ isomorphic to the
1119: reduced $C^*$-algebra of the group $G$? Is it so when $G$ is the
1120: free group generated by the automaton shown on
1121: Figure~\ref{fig:alfree} in Example~\ref{sss:free}?
1122: \end{question}
1123:
1124: \section{Contracting groups and limit solenoids}
1125: \label{s:contracting}
1126: \subsection{Matrix recursion for contracting groups}
1127:
1128: \begin{definition}
1129: A self-similar group $(G, \alb)$ is said to be \emph{contracting}
1130: if there exists a finite set $\nuke\subset G$ such that for every
1131: $g\in G$ there exists $n_0$ such that $g|_v\in\nuke$ for all words
1132: $v\in\xs$ of length greater than $n_0$. The smallest set $\nuke$
1133: with this property is called the \emph{nucleus} of the
1134: self-similar group.
1135: \end{definition}
1136:
1137: \begin{example}
1138: The adding machine action of $\Z$ is contracting with the nucleus
1139: $\nuke=\{1, a, a^{-1}\}$, since $a^n|_0=a^{\lfloor n/2\rfloor}$
1140: and $a^n|_1=a^{\lceil n/2 \rceil}$.
1141: \end{example}
1142:
1143: \begin{example}
1144: It is also not hard to prove that the torsion group from
1145: Example~\ref{sss:grigorchuk} is contracting with the nucleus $\{1,
1146: a, b, c, d\}$.
1147: \end{example}
1148:
1149: \begin{example}
1150: The free group considered in Example~\ref{sss:free} is not
1151: contracting. In this example any section $g|_v$ has the same
1152: length as $g$.
1153: \end{example}
1154:
1155: The nucleus $\nuke$ of a contracting group can be interpreted as
1156: an \emph{automaton}, i.e., for every $g\in\nuke$ and every letter
1157: $x\in\alb$ the restriction $g|_x$ belongs to $\nuke$.
1158: Consequently, if we denote by $N$ the linear span of $\nuke$ in
1159: $\C[G]$ and $\phi:\C[G]\arr M_{d\times d}(\C[G])$ is the
1160: associated matrix recursion, then $\phi(N)$ is a subspace of the
1161: space $M_{d\times d}(N)$ of matrices with entries in $N$.
1162: Moreover, the following is true.
1163:
1164: \begin{theorem}[\cite{nek:cpalg}]
1165: If the group $(G, \alb)$ is contracting and $\nuke$ is its
1166: nucleus, then the algebra $\cpg$ is generated by
1167: $\{a_x\}_{x\in\alb}\cup\nuke$ and is defined by the following
1168: finite set of relations
1169: \begin{enumerate}
1170: \item Cuntz relations \[a_x^*a_x=1\]
1171: \item decompositions
1172: \[g=\sum_{x\in\alb}a_{g(x)}g|_xa_x^*\]
1173: for every $g\in\nuke$ (this includes the remaining Cuntz algebra
1174: relation \[\sum_{x\in\alb}a_xa_x^*=1\] in the case $g=1$),
1175: \item all relations $g_1g_2g_3=1$ of length at most three which are true for the elements
1176: of the nucleus $\nuke$ in the group $G$ and relations
1177: $gg^*=g^*g=1$ for $g\in\nuke$.
1178: \end{enumerate}
1179: \end{theorem}
1180:
1181: Moreover, the groupoid of germs of the action of $G$ on $\xo$ is
1182: amenable in the case of a contracting group $G$. This follows from
1183: the fact that the orbits of the of the action on $\xo$ have
1184: polynomial growth (see~\cite{delaroche_renault}
1185: and~\cite{nek:book}). This implies the following corollary of
1186: Theorem~\ref{th:amenablegr}.
1187:
1188: \begin{proposition}
1189: If the group $G$ is contracting and level-transitive, then the
1190: algebras $\maxalg$ and $\A_{red}$ coincide. If, additionally, for
1191: every element $g$ of the nucleus of $G$ the interior of the set of
1192: fixed points of $g$ is closed, then all self-similar completions
1193: of $G$ are isomorphic to $\maxalg=\minalg=\A_{red}$.
1194: \end{proposition}
1195:
1196: \begin{examples}
1197: The adding machine action, the Basilica group~\ref{sss:basilica}
1198: satisfy all the conditions of the corollary, hence they have
1199: unique self-similar completions.
1200: \end{examples}
1201:
1202: \begin{example}
1203: \label{ex:nonHausd} Consider the group generated by the
1204: transformations $a, b, c$ given by
1205: \begin{alignat*}{2}
1206: a(0w) &= 1w, &\quad a(1w) &= 0w\\
1207: b(0w) &= 0a(w), &\quad b(1w) &=1c(w)\\
1208: c(0w) &= 0w, &\quad c(1w) &= 1b(c).
1209: \end{alignat*}
1210: This is one of the Grigorchuk groups $G_w$ studied
1211: in~\cite{grigorchuk:growth_en} (for $w=111\ldots$). Its growth was
1212: studied by A.~Erschler in~\cite{ershler:growth}.
1213:
1214: This group is contracting with the nucleus $\{1, a, b, c,
1215: bc=cb\}$, but its groupoid of germs is not Hausdorff. Namely, the
1216: set of fixed points of $c$ is equal to
1217: \[\{111\ldots\}\cup\bigsqcup_{k=0, 1, \ldots}\underbrace{11\ldots
1218: 1}_{2k}0\xo,\] its interior is $\bigsqcup_{k=0, 1,
1219: \ldots}\underbrace{11\ldots 1}_{2k}0\xo$, which is not closed.
1220:
1221: Let us show that in this case the element $b+c-bc-1$ belongs to
1222: the kernel of the epimorphism $\maxalg=\A_{red}\arr\minalg$.
1223:
1224: We have \begin{multline*}\phi(b+c-bc-1)=\left(\begin{array}{cc} a & 0\\
1225: 0 & c\end{array}\right)+\left(\begin{array}{cc} 1 & 0\\ 0 &
1226: b\end{array}\right)-\left(\begin{array}{cc} a & 0\\ 0 &
1227: bc\end{array}\right)-\left(\begin{array}{cc} 1 & 0\\ 0 &
1228: 1\end{array}\right)=\\ \left(\begin{array}{cc} 0 & 0\\ 0 &
1229: b+c-bc-1\end{array}\right),\end{multline*} which implies that
1230: $\pi(b+c-bc-1)=0$ for the permutation representation $\pi$ on any
1231: orbit of $G$ on $\xo$. Hence $b+c-bc-1$ is equal to zero in
1232: $\minalg$.
1233:
1234: On the other hand, it follows from contraction that the isotropy
1235: group of the point $111\ldots\in\xo$ in the groupoid of germs of
1236: the action of $G$ on $\xo$ contains 4 elements (the germs of $1,
1237: b, c$ and $bc$). Consequently, for any germ $\gamma$ with range
1238: $111\ldots$ the germs $\gamma$, $b\cdot\gamma$, $c\cdot\gamma$ and
1239: $bc\cdot\gamma$ are pairwise different (here the \emph{range} of a
1240: germ $(h, x)$ is the point $h(x)$). This implies that the element
1241: $b+c-bc-1$ is non-zero in the regular representation of the
1242: groupoid, hence it is non-zero in $\A_{red}$.
1243:
1244: Consequently, the homomorphism $\A_{red}\arr\minalg$ is not an
1245: isomorphism in general. In particular, $\A_{red}$ is not simple,
1246: even though the action of the group $G$ is minimal. (If the action
1247: is minimal and the groupoid of germs is Hausdorff, then the
1248: reduced algebra of the groupoid of germs is simple,
1249: see~\cite{renault:groupoids} Proposition~4.6.)
1250: \end{example}
1251:
1252: \begin{question} Describe the kernel of the epimorphism
1253: $\A_{red}\arr\minalg$. Is it true that for a contracting group $G$
1254: it is generated (in $\cpg$) by linear combinations of the elements
1255: of the nucleus?
1256: \end{question}
1257:
1258:
1259: \subsection{Limit solenoid}
1260: Let us fix some contracting self-similar group $(G, \alb)$ with
1261: the nucleus $\nuke$. Consider the space $\xz$ of bi-infinite
1262: sequences of the form
1263: \[\ldots x_{-2}x_{-1}\;.\;x_0x_1\ldots\]
1264: of letters $x_i\in\alb$. Here the dot marks the place between the
1265: coordinates number 0 and number $-1$. We consider $\xz$ to be a
1266: topological space with the direct product topology of discrete
1267: sets $\alb$.
1268:
1269: \begin{definition}
1270: Two sequences $\ldots x_{-2}x_{-1}\;.\;x_0x_1\ldots$ and $\ldots
1271: y_{-2}y_{-1}\;.\;y_0y_1\ldots$ are said to be \emph{asymptotically
1272: equivalent} (with respect to the action of $G$) if there exists a
1273: finite set $N\subset G$ and a sequence $g_i\in N$ such that
1274: \[g_i(x_ix_{i+1}x_{i+2}\ldots)=y_iy_{i+1}y_{i+2}\ldots\]
1275: for all $i\in\mathbb{Z}$.
1276: \end{definition}
1277:
1278: It is proved in~\cite{nek:book} that we can take $N$ equal to the
1279: nucleus of $G$ and that the asymptotic equivalence relation can be
1280: described in the following way.
1281:
1282: \begin{proposition}
1283: \label{pr:aseq} The sequences $\ldots
1284: x_{-2}x_{-1}\;.\;x_0x_1\ldots$ and $\ldots
1285: y_{-2}y_{-1}\;.\;y_0y_1\ldots$ are asymptotically equivalent if
1286: and only if there exists a sequence $g_i\in\nuke$ of elements of
1287: the nucleus such that $g_i\cdot x_i=y_i\cdot g_{i-1}$, i.e., such
1288: that $y_i=g_i(x_i)$ and $g_{i-1}=g_i|_{x_i}$.
1289: \end{proposition}
1290:
1291: \begin{definition}
1292: The \emph{limit solenoid} $\solen$ of the self-similar group $(G,
1293: \alb)$ is the quotient of the topological space $\xz$ by the
1294: asymptotic equivalence relation.
1295: \end{definition}
1296:
1297:
1298: The next proposition follows from the definition and
1299: Proposition~\ref{pr:aseq} (see details in~\cite{nek:book}).
1300:
1301: \begin{proposition}
1302: The limit solenoid $\solen$ is a compact metrizable
1303: finite-dimensional space. If the action of $G$ on $\xs$ is
1304: level-transitive, then $\solen$ is connected. The shift
1305: \[\ldots x_{-2}x_{-1}\;.\;x_0x_1\ldots\mapsto
1306: \ldots x_{-3}x_{-2}\;.\;x_{-1}x_0\ldots\] induces a homeomorphism
1307: $\si:\solen\arr\solen$.
1308: \end{proposition}
1309:
1310: \begin{definition}
1311: Let $\xmo$ be the space of sequences $\ldots x_2x_1$, over the
1312: alphabet $\alb$ with the direct product topology. Two sequences
1313: $\ldots x_2x_1, \ldots y_2y_1\in\xmo$ are \emph{asymptotically
1314: equivalent} if there exists a finite set $N\subset G$ and a
1315: sequences $g_k\in N$ such that $g_k(x_k\ldots x_1)=y_k\ldots y_1$
1316: for all $k\ge 1$. The quotient of $\xmo$ by the asymptotic
1317: equivalence relation is called the \emph{limit space} of the group
1318: $(G, \alb)$ and is denoted $\lims$.
1319: \end{definition}
1320:
1321: Here also we can take $N$ to be equal to the nucleus
1322: (see~\cite{nek:book}).
1323:
1324: We have a natural continuous projection $\solen\arr\lims$ induced
1325: by the map \[\ldots x_{-2}x_{-1}\;.\;x_0x_1\ldots\mapsto \ldots
1326: x_{-2}x_{-1}.\] This projection semiconjugates the map
1327: $\si:\solen\arr\solen$ with the map $\sis:\lims\arr\lims$ induced
1328: by the one-sided shift $\ldots x_2x_1\mapsto \ldots x_3x_2$. We
1329: call $(\lims, \sis)$ the \emph{limit dynamical system}.
1330:
1331: The following theorem is proved in~\cite{nek:book}, where a more
1332: general formulation can be found.
1333:
1334: \begin{theorem}
1335: If $f:\M_1\arr\M$ is an expanding partial covering (where $\M_1$
1336: is an open subset of $\M$ with the induced Riemann metric), then
1337: the iterated monodromy group $\img{f}$ is contracting and the
1338: limit dynamical system $\sis:\lims[\img{f}]\arr\lims[\img{f}]$ is
1339: topologically conjugate to the action of $f$ on its Julia set.
1340: \end{theorem}
1341:
1342: The limit solenoid $\solen$ can be reconstructed from the limit
1343: dynamical system as the inverse limit of the sequence
1344: \[\lims\stackrel{\sis}{\longleftarrow}\lims\stackrel{\sis}{\longleftarrow}\ldots.\]
1345: The map $\sis$ induces a homeomorphism of the inverse limit, which
1346: is conjugate with $\si:\solen\arr\solen$.
1347:
1348: \begin{example}[Lyubich-Minsky laminations]
1349: Let $f\in\C(z)$ be a post-critically finite rational function
1350: (i.e., the orbits of its critical points are finite). Consider the
1351: inverse limit $\widehat{\mathcal{S}_f}$ of the backward iteration
1352: \[\widehat\C\stackrel{f}{\longleftarrow}\widehat\C\stackrel{f}{\longleftarrow}\ldots,\]
1353: where $\widehat\C$ is the Riemann sphere. The shift along the
1354: inverse sequence (i.e., along the action of $f$) defines a
1355: homeomorphism $\widehat
1356: f:\widehat{\mathcal{S}_f}\arr\widehat{\mathcal{S}_f}$, called the
1357: \emph{natural extension} of $f$. We have the natural projection
1358: $P$ of $\widehat{\mathcal{S}_f}$ onto $\widehat\C$ (onto the first
1359: term of the inverse sequence).
1360:
1361: Let $\mathcal{S}_f$ be the preimage of the Julia set in
1362: $\widehat{\mathcal{S}_f}$ under the projection $P$. Then the space
1363: $\mathcal{S}_f$ is homeomorphic to the limit solenoid of $\img{f}$
1364: and the action of $\widehat f$ on $\mathcal{S}_f$ is topologically
1365: conjugate to the action of the shift on the limit solenoid.
1366:
1367: The space $\widehat{\mathcal{S}_f}$ and the homeomorphism
1368: $\widehat f$ were studied in~\cite{lyubichminsk}.
1369: \end{example}
1370:
1371: \subsection{The solenoid as a hyperbolic dynamical system}
1372: Let us have a look at the dynamical system $\si:\solen\arr\solen$
1373: more carefully. To avoid technicalities, we will assume that the
1374: self-similar group $(G, \alb)$ is finitely generated, contracting,
1375: self-replicating and that it is \emph{regular}, which means that
1376: for every $g\in G$ and $w\in\xo$ either $g(w)\ne w$, or $w$ is
1377: fixed by $g$ together will all points of a neighborhood of $w$
1378: (i.e., for some beginning $v$ of $w$ the restriction $g|_v$ is
1379: trivial). In other words, a group $G$ is said to be regular if
1380: every point of $\xo$ is $G$-regular in the sense of
1381: Definition~\ref{def:Ggeneric}. It is sufficient to check the
1382: regularity condition for the elements $g$ belonging to the nucleus
1383: of $G$.
1384:
1385: If $f:\mathcal{M}\arr\mathcal{M}$ is an expanding self-covering of
1386: a complete compact geodesic space, then the iterated monodromy
1387: group $G=\img{f}$ is contracting and the dynamical system
1388: $\sis:\lims[\img{f}]\arr\lims[\img{f}]$ is topologically conjugate
1389: to $(\mathcal{M}, f)$. Moreover, $\img{f}$ satisfies the
1390: regularity condition in this case (this easily follows from the
1391: description of the orbispace structure on $\lims[\img{f}]$ given
1392: in~\cite{nek:book}).
1393:
1394: We say that two points $\xi, \zeta\in\solen$ are \emph{stably
1395: (unstably) equivalent} if for every neighborhood of the diagonal
1396: $U\subset\solen\times\solen$ there exists $n_U\ge 0$ such that
1397: \[(\si^n(\xi), \si^n(\zeta))\in U\] for all $n\ge n_U$ ($n\le -n_U$,
1398: respectively).
1399:
1400: \begin{proposition}
1401: \label{pr:reghyperb} Let the points $\xi, \zeta\in\solen$ be
1402: represented by sequences $(x_n)_{n\in\Z}$ and $(y_n)_{n\in\Z}$,
1403: respectively.
1404:
1405: The points $\xi, \zeta$ are stably equivalent if and only if there
1406: exists $n\in\Z$ such that the sequences $\ldots x_{n-1}x_n$ and
1407: $\ldots y_{n-1}y_n\in\xmo$ are asymptotically equivalent, i.e.,
1408: represent the same point of $\lims$.
1409:
1410: The points $\xi, \zeta$ are unstably equivalent if and only if
1411: there exists $n\in\Z$ such that
1412: \[g(x_nx_{n+1}\ldots)=y_ny_{n+1}\ldots\] for some element $g$ of the nucleus.
1413: \end{proposition}
1414:
1415: \begin{proof}
1416: It is easy to see that stable and unstable equivalence follows
1417: from the conditions in the proposition.
1418:
1419: Let us show that the converse implications hold. Let $k$ be such
1420: that for every two elements $g, h\in\nuke$ of the nucleus and
1421: every word $v\in\alb^n$ either $g(v)\ne h(v)$ or $g(v)=h(v)$ and
1422: $g|_v=h|_v$. Such $k$ exists by the regularity condition.
1423:
1424:
1425: Denote by $\overline U_k$ the set of pairs $\left((x_n)_{n\in\Z},
1426: (y_n)_{n\in\Z}\right)$ such that $g(x_{-k}\ldots x_k)=y_{-k}\ldots
1427: y_k$ for some element $g\in\nuke$ of the nucleus. Let $U_k$ be the
1428: image of $\overline U_k$ in $\solen\times\solen$.
1429:
1430: Every set $U_k$ is a neighborhood of the diagonal and for every
1431: neighborhood of the diagonal $U\subset\solen\times\solen$ there
1432: exists $k$ such that $U_k\subset U$.
1433:
1434: Let $\xi$ and $\zeta$ be the points of $\solen$ represented by the
1435: sequences $(x_i)_{i\in\Z}$ and $(y_i)_{i\in\Z}$.
1436:
1437: Suppose that $\xi$ and $\zeta$ are stably equivalent. Then for any
1438: $k$ there exists $n_k$ such that the pair
1439: $\left(\si^n\left((x_i)_{i\in\Z}\right),
1440: \si^n\left((y_i)_{i\in\Z}\right)\right)$ belongs to $\overline
1441: U_k$ for all $n\ge n_k$. Then for every $n>n_k+k$ there exists
1442: $g_n\in\nuke$ such that $g_n(x_{-n}x_{-n+1}\ldots
1443: x_{-n+2k})=y_{-n}y_{-n+1}\ldots y_{-n+2k}$. Consequently, for
1444: $h_n=g_n|_{x_{-n}\ldots x_{-n+k-1}}$ we have \[h_n(x_{-n+k}\ldots
1445: x_{-n+2k})=y_{-n+k}\ldots y_{-n+2k}.\] Note that by the choice of
1446: $k$ the element $h_n$ is determined uniquely by the pair of words
1447: $(x_{-n}x_{-n+1}\ldots x_{-n+2k}, y_{-n}y_{-n+1}\ldots y_{-n+2k})$
1448: (i.e., does not depend on the choice of $g_n$). The uniqueness of
1449: $g_n$ implies that $h_n\cdot x_{-n+k}=y_{-n+k}\cdot h_{n-1}$,
1450: which finishes the proof.
1451:
1452: The proof of the statement about the unstable equivalence is
1453: analogous.
1454: \end{proof}
1455:
1456: Hence we get the following.
1457:
1458: \begin{corollary}
1459: In conditions of Proposition~\ref{pr:reghyperb} the points $\xi$
1460: and $\zeta$ are unstably equivalent if and only if $x_0x_1\ldots$
1461: and $y_0y_1\ldots$ belong to one $G$-orbit.
1462:
1463: In other words, the unstable equivalence relation
1464: $\mathcal{U}\subset\solen\times\solen$ is a union of the sets
1465: $\mathcal{U}_g$ equal to the images in $\solen\times\solen$ of the
1466: sets
1467: \[\{(\ldots x_{-2}x_{-1}\;.\;x_0x_1\ldots; \ldots
1468: y_{-2}y_{-1}\;.\;g(x_0x_1\ldots))\;:\;x_i, y_i\in\alb\}.\]
1469: \end{corollary}
1470:
1471: Let us introduce a topology on $\mathcal{U}$ equal to the (direct
1472: limit) topology of the union $\mathcal{U}=\bigcup_{g\in
1473: G}\mathcal{U}_g$, where $\mathcal{U}_g$ are taken with the induced
1474: topology of the subset $\mathcal{U}_g\subset\solen\times\solen$
1475: and the direct limit is taken with respect to the identical maps
1476: between the sets $\bigcup_{g\in A}\mathcal{U_g}$, where $A$ runs
1477: through the finite subsets of $G$.
1478:
1479: This topology converts the unstable equivalence relation into a
1480: Hausdorff topological groupoid, which we call the \emph{unstable
1481: groupoid}. A natural Haar system on it comes from the push forward
1482: of the uniform Bernoulli measure $\nu$ on $\xmo$.
1483:
1484: The shift $\si:\solen\arr\solen$ defines automorphisms of the
1485: stable and unstable groupoids.
1486:
1487: The next theorem can be seen as a formulation of the fact that $G$
1488: is the holonomy group of the unstable foliation of $\solen$ in
1489: terms of $C^*$-algebras.
1490:
1491: \begin{theorem}
1492: \label{th:moritaequiv} The universal convolution algebra of the
1493: unstable groupoid is strongly Morita equivalent to $\gi$.
1494: \end{theorem}
1495:
1496: \begin{proof}
1497: Let us construct an equivalence bimodule (in the sense
1498: of~\cite{muhlyrenault:equiv}) between the groupoid $\mathcal{U}$
1499: of the unstable equivalence relation and the groupoid
1500: $\mathcal{G}$ of germs of the semigroup of transformations
1501: $T_vgT_u^*$ for $u, v\in\xs, |u|=|v|$ and $g\in G$. The universal
1502: convolution algebra of $\mathcal{G}$ is $\gi$
1503: (see~\cite{nek:cpalg}). Note also that regularity of $(G, \alb)$
1504: implies that the groupoid $\mathcal{G}$ is principal (i.e., is an
1505: equivalence relation), since the isotropy groups of $\mathcal{G}$
1506: are trivial.
1507:
1508: Let us consider the bimodule $Z$ consisting of the set of pairs
1509: $(\zeta, w)$, where $\zeta\in\solen$ and $w\in\xo$ are such that
1510: $\ldots x_{-2}x_{-1}\;.\;w$ represents a point unstably equivalent
1511: to $\zeta$ (for some and hence for any $\ldots x_{-2}x_{-1}$). Let
1512: $Z_g\subset Z$ for $g\in G$ be the set of pairs $(\zeta, w)$ such
1513: that $\zeta$ is represented by a sequence $\ldots
1514: x_{-2}x_{-1}\;.\;g(w)$. Endow $Z_g$ with the induced topology from
1515: $\solen\times\xo$. The set $Z$ is a union of $Z_g$ for all $g\in
1516: G$. Consider the direct limit topology on $Z$ coming from this
1517: union.
1518:
1519: We have a natural right action of $\mathcal{U}$ on $Z$:
1520: \[(\zeta_2, w)\cdot (\zeta_1, \zeta_2)=(\zeta_1, w)\]
1521: and a left action of $\mathcal{G}$ on $Z$:
1522: \[(w_2, w_1)\cdot (\zeta, w_1)=(\zeta, w_2),\]
1523: which satisfy the conditions of the equivalence bimodule (see
1524: Definition~2.1 of~\cite{muhlyrenault:equiv}).
1525:
1526: One can find another proof of this theorem in~\cite{nek:cpalg}.
1527: \end{proof}
1528:
1529: The dynamical system $\si:\solen\arr\solen$ is an example of a
1530: \emph{Smale space} (for a definition see~\cite{put_kam_sp}). The
1531: $C^*$-algebras associated with hyperbolic dynamical systems were
1532: studied
1533: in~\cite{putnam,put_kam_sp,putnam:structure,ruelle:therm,ruelle:grpd}.
1534:
1535: \section{Schur complements and rational multidimensional dynamics}
1536: \subsection{Schur complements of operators}
1537: Schur complements are widely used in linear algebra (usually
1538: without knowledge of the name). The term ``Schur complement'' was
1539: apparently introduced for the first time by
1540: E.~V.~Haynsworth~\cite{haynsworth:inertia}. See a
1541: survey~\cite{cottle:schursurvey} of the use of Schur complement in
1542: algebra and statistics. It is also related to \emph{Bruhat normal
1543: form} of matrices over a skew field (see~\cite{cohn}), which is
1544: used to define the \emph{Dieudonn\'e determinant}.
1545:
1546: Schur complements are quite often used in different situations
1547: where renormalization principle can be used. Here we describe one
1548: such situation, which is related to computation of spectra of
1549: Hecke type operators and discrete Laplace operators attached to
1550: self-similar groups and their Schreier graphs. The material
1551: written here summarizes the ideas and results of the
1552: articles~\cite{bgr:spec,grineksu_en,gr_zu:lamp,gri:solvedunsolved,grisavchuksunic:img}.
1553:
1554: Let $H$ be a Hilbert space decomposed into a direct sum $H=H_1\oplus H_2$ of two non-zero
1555: subspaces. Let $M\in B(H)$ be a bounded operator and let
1556: \[M=\left(\begin{array}{cc}A & B\\ C & D\end{array}\right)\]
1557: be an operator matrix representing $M$ according to this decomposition.
1558:
1559: \begin{definition}
1560: (i) Assume that $D\in B(H_2)$ is invertible. Then the \emph{first Schur complement}
1561: is the operator
1562: \[S_1(M)=A-BD^{-1}C\]
1563:
1564: (ii) Assume that $A\in B(H_1)$ is invertible. Then the \emph{second Schur complement}
1565: is the operator
1566: \[S_2(M)=D-CA^{-1}B.\]
1567: \end{definition}
1568:
1569: The reason why Schur complements are useful for the spectral
1570: problem is the following well known statement.
1571:
1572: \begin{theorem}
1573: \label{th:schurcomplement} Suppose that $D$ is invertible. Then
1574: $M$ is invertible if and only if $S_1(M)$ is invertible.
1575: Similarly, if $A$ is invertible, then $M$ is invertible if and
1576: only if $S_2(M)$ is invertible.
1577:
1578: The inverse is computed then by the formula
1579: \begin{equation}\label{eq:frobenius} M^{-1}=\left(\begin{array}{cc}S_1^{-1} & -S_1^{-1}BD^{-1}\\
1580: -D^{-1}CS_1^{-1} &
1581: D^{-1}CS_1^{-1}BD^{-1}+D^{-1}\end{array}\right),\end{equation}
1582: where $S_1=S_1(M)$.
1583: \end{theorem}
1584:
1585: \begin{proof}
1586: Consider \[L=\left(\begin{array}{rr}1 & 0\\ -D^{-1}C & D^{-1}\end{array}\right),\]
1587: where $0$ and $1$ represent the zero and identity linear maps between the corresponding subspaces of
1588: $H$.
1589:
1590: Then
1591: \[L^{-1}=\left(\begin{array}{rr}1 & 0\\ C & D\end{array}\right),\quad ML=
1592: \left(\begin{array}{rr} A-BD^{-1}C & BD^{-1}\\ 0 & 1\end{array}\right).\]
1593: A triangular operator matrix \[R=\left(\begin{array}{rr} x & y\\ 0 & 1\end{array}\right)
1594: \]
1595: represents a (right) invertible operator if and only if $x$ is
1596: (right) invertible, hence invertibility of $ML$ is equivalent to
1597: invertibility of $S_1(M)$ and the result follows. The second part
1598: of the theorem is proved similarly.
1599:
1600: Direct computations show that~\eqref{eq:frobenius} takes place.
1601: \end{proof}
1602:
1603: Note that if $A$, $D$ and $M$ are invertible, then
1604: \[M^{-1}=\left(\begin{array}{cc}S_1^{-1} & B'\\
1605: C' & S_2^{-1}\end{array}\right),\] where $S_1=S_1(M)$,
1606: $S_2=S_2(M)$ and
1607: \begin{gather*}
1608: B'=-S_1^{-1}BD^{-1}=-A^{-1}BS_2^{-1},\\
1609: C'=-D^{-1}CS_1^{-1}=-S_2^{-1}CA^{-1}.
1610: \end{gather*}
1611: The last two equalities follow from
1612: \[S_1A^{-1}B=B-BD^{-1}CA^{-1}B=BD^{-1}S_2\] and
1613: \[CA^{-1}S_1=C-CA^{-1}BD^{-1}C=S_2D^{-1}C.\]
1614:
1615: Formula~\eqref{eq:frobenius} is called sometimes \emph{Frobenius
1616: formula}, see, for instance~\cite{gantmacher}.
1617:
1618: We see that taking Schur complement of $M$ is equivalent to
1619: inverting the left top corner of the matrix $M^{-1}$.
1620:
1621: The following corollary is known as \emph{Schur
1622: formula}~\cite{schur:formula} and easily follows from the proof of
1623: Theorem~\ref{th:schurcomplement}.
1624:
1625: \begin{corollary}
1626: Let $H$ be finite dimensional and suppose that the determinant $|D|$ is not equal to zero.
1627: Then \[|M|=|S_1(M)|\cdot |D|.\]
1628: \end{corollary}
1629:
1630: There is nothing special in decomposition of $H$ into a direct sum
1631: of \emph{two} subspaces. If $H=H_1\oplus H_2\oplus\cdots\oplus
1632: H_d$ and
1633: \begin{equation}\label{eq:m}
1634: M=\left(\begin{array}{ccc} M_{11} & \ldots & M_{1d}\\ \vdots &
1635: \ddots & \vdots \\ M_{d1} & \ldots & M_{dd}\end{array}\right),
1636: \end{equation}
1637: where $M_{ij}:H_j\arr H_i$, then we can write $H$ as $H=H_1\oplus
1638: H_1^{\perp}$, where $H_1^\perp=H_2\oplus\cdots\oplus H_d$ and get
1639: \[M=\left(\begin{array}{cc}M_{11} & B\\ C & D\end{array}\right),\]
1640: where $B, C$ and $D$ are operators represented by $1\times (d-1),
1641: (d-1)\times 1$ and $(d-1)\times (d-1)$-sized operator matrices
1642: coming from~\eqref{eq:m}. Then $S_1(M)$ is defined as before under
1643: the condition that $D$ is invertible.
1644:
1645: Changing the order of the summands (putting $H_i$ on the first place, say using a cyclic permutation)
1646: we define the $i$th Schur complement $S_i(M)$.
1647:
1648: We will use these definitions in situations when $H_i$ for $i=1,
1649: \ldots, m$ are isomorphic. There are two cases.
1650:
1651: (I) Finite dimensional, when $H=\underbrace{H'\oplus\cdots\oplus
1652: H'}_{\text{$d$ times}}$ for $\dim H'<\infty$.
1653:
1654: We will apply the Schur complements to the sequence of operators
1655: $M_n\in B(H_n)$, where $\dim H_n=d^n$, and
1656: \[H_{n+1}=\underbrace{H_n\oplus\cdots\oplus H_n}_{\text{$d$ times}},\]
1657: for all $n=0, 1, \ldots$ and $\dim H_0=1$.
1658:
1659: (II) Infinite dimensional case. Suppose that we have an infinite
1660: dimensional (separable) Hilbert space $H$ and a fixed
1661: $d$-similarity
1662: \[\psi:H\arr H^d.\]
1663:
1664: Let \[\widetilde{S}_i=\psi^{-1}S_i\psi:B(H)\arr B(H),\quad
1665: i=1,\ldots, d\] be partially defined transformations, where $S_i$,
1666: as before, is the $i$th Schur complement.
1667:
1668: In the sequel we will write just $S_i$ instead of
1669: $\widetilde{S}_i$, as this will not lead to a confusion.
1670:
1671: \begin{proposition}
1672: \label{pr:schurcuntz} Let $\rho$ be a representation of the Cuntz
1673: algebra $O_d$ associated with the $d$-similarity $\psi:H\arr H^d$
1674: and let $T_i=\rho(a_i)$ be the images of the generators of the
1675: Cuntz algebra. Then the Schur complements are given by
1676: \[S_i(M)=\left(T_i^*M^{-1}T_i\right)^{-1}.\]
1677: \end{proposition}
1678:
1679: \begin{proof}
1680: Follows directly from the definition of $T_i$ and inversion
1681: formula~\eqref{eq:frobenius} from
1682: Theorem~\ref{th:schurcomplement}. Recall (see
1683: Proposition~\ref{pr:cuntzrepr}) that $T_i$ are given by
1684: \[T_i(\xi)=\psi^{-1}(0, \ldots, 0, \xi, 0, \ldots, 0),\] where
1685: $\xi$ is on the $i$th coordinate.
1686: \end{proof}
1687:
1688: \begin{corollary}
1689: We have the following formula for the composition of Schur
1690: complements
1691: \[S_{i_1}\circ\cdots\circ S_{i_k}(M)=\left((T_{i_k}\cdots
1692: T_{i_1})^*\cdot M^{-1}\cdot(T_{i_k}\cdots T_{i_1})\right)^{-1}.\]
1693: \end{corollary}
1694:
1695: We see that a composition of $k$ Schur complements associated with
1696: the $d$-similarity $\psi:H\arr H^d$ is a Schur complement
1697: associated with the corresponding $d^k$-similarity of $H$ obtained
1698: by iteration of $\psi$.
1699:
1700: Consider the semigroup $\mathcal{S}=\langle S_1, \ldots,
1701: S_m\rangle$ of partial transformations of $B(H)$ generated by the
1702: transformations $S_i$. We will call it the \emph{Schur
1703: renorm-semigroup}. In the examples that will follow, we will
1704: restrict $\mathcal{S}$ (or its particular elements
1705: $s\in\mathcal{S}$) onto some $\mathcal{S}$-invariant (or
1706: $s$-invariant) subspaces $B'\subset B(H)$. In the examples that we
1707: consider $B'$ will be finite dimensional.
1708:
1709: We will see in Example~\ref{ex:resolvent} how nicely Schur
1710: complement behaves on the space of resolvents.
1711:
1712: The Schur transformations $S_i$ are homogeneous:
1713: \[S_i(tM)=tS_i(M)\]
1714: for all $t\in\mathbb{C}$. This allows to define the corresponding
1715: transformations $\widehat S_i$ on the ``projective space''
1716: $PB(H)=B(H)/\C^{\times}$ of $B(H)$ (here $\C^\times$ is the
1717: multiplicative group of $\C$). Restricting to invariant
1718: finite-dimensional subspaces $B'\subset B(H)$ (real or complex) we
1719: get transformations on the corresponding projective spaces
1720: $\mathbb{R}P^n$ or $\mathbb{C}P^n$, where $n+1=\dim B'$.
1721:
1722: \subsection{Schur complements in self-similar
1723: algebras}\label{ss:scssalg} If $B$ is a unital Banach algebra
1724: together with a unital embedding $B\arr M_{d\times d}(B)$ (we call
1725: such algebras \emph{self-similar}), then we can compute the Schur
1726: complements $S_i(a)$ of the elements of $B$.
1727:
1728: Let $(G, \alb)$ be a self-similar group, let
1729: $\phi:G\arr\symm\wr_\alb G$ be the corresponding wreath recursion
1730: and let $\cpg$ be the universal Cuntz-Pimsner algebra of $G$. Then
1731: the Schur complements $S_i$ are defined on $\maxalg\subset\cpg$
1732: and can be computed using Proposition~\ref{pr:schurcuntz} or
1733: directly using the matrix recursion $\phi:\C[G]\arr M_{d\times
1734: d}(\C[G])$.
1735:
1736: If $\psi:H\arr H^d$ is a fixed $d$-similarity on $H$ (i.e., a
1737: representation of the Cuntz algebra $O_d$, then, as it was
1738: observed in Proposition~\ref{pr:cpimsner}, every self-similar
1739: representation $\rho$ of $G$ extends to a representation
1740: $\rho:\cpg\arr B(H)$. The corresponding Schur complements on the
1741: closure $\A_\rho$ of $\C[G]$ are also computed by the formula in
1742: Proposition~\ref{pr:schurcuntz}.
1743:
1744: Let us consider, for instance, the space $L^2(\xo, \nu)$, the
1745: natural unitary representation $\pi$ of a self-similar group $(G,
1746: \alb)$ on it and the natural $d$-similarity (and the associated
1747: representation of the Cuntz algebra) on $L^2(\xo, \nu)$, see
1748: Example~\ref{ex:murepr}. Consider the respective self-similar
1749: completion $\A_{mes}$ of $\C[G]$.
1750:
1751: The partition $\xi_n$ of the boundary $\partial\xs=\xo$ into a
1752: disjoint union of $d^n$ cylinder subsets $v\xo$ ($v\in\alb^n$),
1753: corresponding to the vertices of the $n$th level of the tree is
1754: $G$-invariant (since it is invariant under the action of the whole
1755: automorphism group of the tree).
1756:
1757: Let $H_n=\{X_{A_i^{(n)}},\; 1\le i\le d^n\}$ be the finite
1758: dimensional subspace spanned by the characteristic functions of
1759: the atoms $A_i^{(n)}$ of the partition $\xi_n$. Then $H_n$ is a
1760: $\pi(G)$-invariant subspace and $\{H_n\}_{n=1}^\infty$ is a nested
1761: sequence with canonical embeddings $H_n\hookrightarrow H_{n+1}$
1762: and
1763: \begin{equation}\label{eq:Hn} H=\overline{\bigcup_{n=1}^\infty
1764: H_n}.\end{equation}
1765:
1766: Note that
1767: \[H_{n+1}=\bigoplus_{1\le i\le d}T_i(H_n),\]
1768: where $T_i=\pi(a_i)$ are the generators of the Cuntz algebra
1769: associated to the natural $d$-similarity of $L^2(\xo, \nu)$. The
1770: space $T_i(H_n)$ is the linear span of the characteristic
1771: functions of cylindrical sets of the level $n$ with the first
1772: letter $i$.
1773:
1774: Let $M=\sum_{g\in G}\lambda_g g\in\ell^1(G)$ be an arbitrary
1775: element of the Banach group algebra and let $\pi(M)$ be its image
1776: in $\A_{mes}$. Denote $M_n=\pi(M)|_{H_n}$. Relation~\eqref{eq:Hn}
1777: and $G$-invariance of the spaces $H_n$ implies that
1778: \begin{equation}\label{eq:spec}
1779: \spec{\pi(M)}=\overline{\bigcup_{n=1}^\infty\spec{M_n}}.\end{equation}
1780:
1781: The formula~\eqref{eq:spec} reduces the problem of computation of
1782: the spectrum of $M$ to finite dimensional problems of finding
1783: $\spec{M_n}$.
1784:
1785: The latter problem requires computation of the $d^n\times
1786: d^n$-size matrices $g^{(n)}$ given by the recursive relations:
1787: \[g^{(0)}=1,\quad g^{(n+1)}=\left(A_{yx}\right)_{x,y\in\alb},\quad
1788: A_{yx}=\left\{\begin{array}{ll}g|_x^{(n)} & \text{if $g(x)=y$,}\\
1789: 0 & \text{otherwise,}\end{array}\right.\] repeating the matrix
1790: recursion~\eqref{eq:matrrecursion} and finding spectra of their
1791: linear combinations. There is no general tool for solving this
1792: problem. Nevertheless, we will show how the problem can be solved
1793: using the Schur map in some particular cases.
1794:
1795: We will start with some examples which have a complete solution of
1796: the spectral problem, then show a few examples when the spectral
1797: problem is reduced to a problem in Dynamical Systems of finding
1798: invariant sets for a multidimensional rational mapping (it looks
1799: that these invariant sets are indeed ``strange attractors'' for
1800: the corresponding maps, at least it is so in many cases). Finally
1801: we will show some examples for which the spectral problem is
1802: extremely interesting and has links with other topics, but it
1803: looks like the method of Schur complements doesn't work for these
1804: examples.
1805:
1806: We perform some computations in homogeneous coordinates of the
1807: space of parameters, since the spectrum of a pencil (i.e., the set
1808: of values of parameters corresponding to degenerate matrices) is
1809: invariant under multiplication by a non-zero number and the Schur
1810: complement transformations are homogeneous. But the obtained
1811: transformations are written in non-homogeneous coordinates as
1812: transformations in the Euclidean space.
1813:
1814: In all examples that follow we use parametric family of elements
1815: of a group algebra (usually the sum with coefficients-parameters
1816: of generators and identity) or a matrix with entries of this type
1817: and consider the corresponding operator in $L^2(\xo, \nu)$ given
1818: by the representation $\pi$ (so that the operators belong to the
1819: algebra $\mathcal{A}_{mes}$).
1820:
1821: \begin{example} Let $\mathfrak{G}=\langle a, b, c, d\rangle$ be the group from
1822: Example~\ref{sss:grigorchuk}. Since the generators are of order 2,
1823: the operator $M=\pi(a+b+c+d)$ is a self-adjoint operator in
1824: $B(L^2(\partial T, \nu))$, being also a self-adjoint element of
1825: $\A_{mes}$ with the same spectrum. In what follows we omit $\pi$.
1826: We have $a=\left(\begin{array}{cc} 0 & 1\\ 1 &
1827: 0\end{array}\right)$, $b=\left(\begin{array}{cc} a & 0\\ 0 &
1828: c\end{array}\right)$, $c=\left(\begin{array}{cc} a & 0\\ 0 &
1829: d\end{array}\right)$, $d=\left(\begin{array}{cc} 1 & 0\\ 0 &
1830: b\end{array}\right)$, and
1831: \[M=a+b+c+d=\left(\begin{array}{cc} 2a+1 & 1\\ 1 &
1832: b+c+d\end{array}\right).\] Also $a_{n+1}=\left(\begin{array}{cc} 0 & 1\\
1833: 1 & 0\end{array}\right)$, $b_{n+1}=\left(\begin{array}{cc} a_n & 0\\
1834: 0 & c_n\end{array}\right)$, $c_{n+1}=\left(\begin{array}{cc} a_n & 0\\
1835: 0 & d_n\end{array}\right)$, $d_{n+1}=\left(\begin{array}{cc} 1 & 0\\
1836: 0 & b_n\end{array}\right)$ are recursive relations determining
1837: matrices of generators viewed as operators in the spaces $H_{n+1}$
1838: and $H_n$ respectively (0 and 1 represent here the zero operators
1839: and the identity operator of the corresponding size). Instead of
1840: computing spectrum of $M_n=a_n+b_n+c_n+d_n$ directly, one can try
1841: to find spectrum of the whole pencil $M_n(x, y, z,
1842: u)=xa_n+yb_n+zc_n+ud_n$ of matrices, where $x, y, z, u\in\C$.
1843:
1844: If we find spectrum of each of the pencils $M_n(x, y, z, u)$, we
1845: will solve the problem of finding spectrum of the pencil $M(x, y,
1846: z, u)=xa+yb+zc+ud$ of infinite dimensional operators, since
1847: spectrum $M(x, y, z, u)$ is the closure of the unions of spectra
1848: of $M_n(x, y, z, u)$.
1849:
1850: The 5-dimensional space of operators $M(x, y, z, u,
1851: v)=xa+yb+zc+ud+v\cdot 1$ is invariant with respect to the Schur
1852: complements which in this case are given by the following
1853: formulae.
1854: \[\widehat S_1:\left(\begin{array}{c} x\\ y\\ z\\ u\\
1855: v\end{array} \right)\mapsto\left(\begin{array}{c} z+y\\
1856: \frac{x^2(2yzv-u(y^2+z^2-u^2+v^2))}{(y+z+u+v)(y+z-u-v)(y-z+u-v)(-y+z+u-v)}\\
1857: \frac{x^2(2zuv-y(-y^2+z^2+u^2+v^2))}{(y+z+u+v)(y+z-u-v)(y-z+u-v)(-y+z+u-v)}\\
1858: \frac{x^2(2yuv-z(y^2-z^2+u^2+v^2))}{(y+z+u+v)(y+z-u-v)(y-z+u-v)(-y+z+u-v)}\\
1859: u+v+\frac{x^2(2yzu-v(y^2+z^2+u^2-v^2))}{(y+z+u+v)(y+z-u-v)(y-z+u-v)(-y+z+u-v)}\end{array}\right)\]
1860: and
1861: \[
1862: \widehat S_2:\left(\begin{array}{c} x\\ y\\ z\\ u\\
1863: v\end{array}\right)\mapsto\left(\begin{array}{c}\frac{x^2(y+z)}{(u+v+y+z)(u+v-y-z)}\\
1864: u\\ y\\ z\\
1865: v-\frac{x^2(u+v)}{(u+v+y+z)(u+v-y-z)}\end{array}\right)
1866: \]
1867:
1868: The complement $\widehat S_2$ is not defined when
1869: $(y+z)a+(u+v)\cdot 1$ is not invertible, i.e., if
1870: \begin{equation}\label{eq:s2not1}
1871: y+z+u+v = 0
1872: \end{equation}
1873: or
1874: \begin{equation}\label{eq:s2not2}
1875: y+z-u-v = 0,
1876: \end{equation}
1877: as $\spec{a}=\{\pm 1\}$.
1878:
1879: The complement $\widehat S_1$ is not defined for the same
1880: conditions~\eqref{eq:s2not1} and~\eqref{eq:s2not2} and also the
1881: conditions obtained from them by application of the cyclic
1882: permutation $(u, y, z)\mapsto (z, u, y)\mapsto (y, z, u)$, as
1883: \begin{multline*}yc+zd+ub+v\cdot 1\\ =y\left(\begin{array}{cc}a & 0\\ 0 &
1884: d\end{array}\right)+z\left(\begin{array}{cc}1 & 0\\ 0 &
1885: b\end{array}\right)+u\left(\begin{array}{cc}a & 0\\ 0 &
1886: c\end{array}\right)+v\left(\begin{array}{cc}1 & 0\\ 0 &
1887: 1\end{array}\right)\\
1888: =\left(\begin{array}{cc}(y+u)a+(z+v)1 & 0\\ 0 &
1889: yd+zb+uc+v1\end{array}\right).
1890: \end{multline*}
1891:
1892: Note that in both cases the complement $\widehat S_i$ is not
1893: defined if and only if the denominator in the corresponding
1894: formulae is equal to zero.
1895:
1896: Note that the third iteration of the map $\widehat S_2$ fixes the
1897: three middle coordinates, hence we actually get a family of maps
1898: $\C^2\arr\C^2$ depending on three parameters.
1899:
1900: It is not known at the moment if the spectrum of the pencil $M(x,
1901: y, z, u, v)$ is invariant with respect to these maps and what are
1902: its topological properties. Further investigation in this
1903: direction would be interesting.
1904:
1905: But let us simplify the problem and consider the pencil
1906: $M(x)=xa+b+c+d$. The spectral problem for $M(x)$ consists in
1907: finding those values of the parameters $x$ and $\lambda$ for which
1908: the operator $M(x)-\lambda\cdot 1$ is not invertible. We will call
1909: this set the \emph{spectrum} of the pencil. This terminology will
1910: be used also for further examples. To simplify transformations let
1911: us change the system of coordinates and consider the pencil
1912: \[R(\lambda, \mu)=-\lambda a+b+c+d-(\mu+1)1\]
1913: and its finite dimensional approximations
1914: \[R_n(\lambda, \mu)=-\lambda a_n+b_n+c_n+d_n-(\mu+1)1_n\]
1915: represented by a 2-parametric family of $2^n\times 2^n$-matrices.
1916:
1917: Let \[\Sigma=\{(\lambda, \mu)\;:\;\text{$R(\lambda, \mu)$ is not
1918: invertible}\}\] and
1919: \[\Sigma_n=\{(\lambda, \mu)\;:\;\text{$R_n(\lambda, \mu)$ is not
1920: invertible}\}.\] Define $F, G:\R^2\arr\R^2$ be two rational maps
1921: \begin{eqnarray*}
1922: F:\left(\begin{array}{c}\lambda\\
1923: \mu\end{array}\right) &\mapsto&
1924: \left(\begin{array}{c}\frac{2(4-\mu^2)}{\lambda^2}\\
1925: -\mu-\frac{\mu(4-\mu^2)}{\lambda^2}\end{array}\right),\\
1926: G:\left(\begin{array}{c}\lambda\\
1927: \mu\end{array}\right) &\mapsto&
1928: \left(\begin{array}{c}\frac{2\lambda^2}{4-\mu^2}\\
1929: \mu+\frac{\mu\lambda^2}{4-\mu^2}\end{array}\right) \end{eqnarray*}
1930: and observe, following Y.~Vorobets, that $H\circ F=G$, $H\circ
1931: G=F$, where
1932: \[H:\left(\begin{array}{c}\lambda\\
1933: \mu\end{array}\right)\mapsto\left(\begin{array}{c}4/\lambda\\
1934: -2\mu/\lambda\end{array}\right)\] is an involutive map
1935: $\R^2\arr\R^2$.\end{example}
1936:
1937: \begin{theorem}
1938: I. The maps $F$ and $G$ seen as maps on the projective space are
1939: respectively the first and the second Schur maps restricted on the
1940: pencil $R(\lambda, \mu)$.
1941:
1942: II. The set $\Sigma$ is invariant with respect to $F$ and $G$,
1943: i.e., $F^{-1}\Sigma=\Sigma$, $G^{-1}\Sigma=\Sigma$ and relations
1944: \[\Sigma_{n+1}=F^{-1}\Sigma_n=G^{-1}\Sigma_n\]
1945: hold.
1946:
1947: III. The set $\Sigma$ is the set shown on Figure~\ref{fig:sigma}.
1948:
1949: IV. The set $\Sigma_n$ is the union of the line $\lambda+\mu-2=0$
1950: and hyperbolas $H_\theta=0$,
1951: $\theta\in\bigcup_{i=0}\alpha^{-i}(0)$, where
1952: $H_\theta=4-\mu^2+\lambda^2+4\lambda\theta$. See
1953: Figure~\ref{fig:sigma5} for the spectrum $\Sigma_5$.
1954: \end{theorem}
1955:
1956: \begin{figure}
1957: \includegraphics{sigma.eps}\\
1958: \caption{The spectrum $\Sigma$}\label{fig:sigma}
1959: \end{figure}
1960:
1961: \begin{figure}
1962: \includegraphics{sigma5.eps}\label{fig:sigma5}
1963: \caption{The spectrum $\Sigma_5$}
1964: \end{figure}
1965:
1966: \begin{proof}
1967: We restrict ourselves here only to computation of the Schur maps.
1968: The rest can be found in~\cite{bgr:spec}.
1969:
1970: We have
1971: \[R(\lambda, \mu)=\left(\begin{array}{cc}2a-\mu & -\lambda \\ -\lambda &
1972: b+c+d-\mu-1\end{array}\right).\]
1973:
1974: For getting the first Schur map observe that $t=(b+c+d-1)/2$ is an
1975: idempotent. Hence
1976: \[(b+c+d-\mu-1)^{-1}=(2t-\mu)^{-1}=\frac{2t+\mu}{4-\mu^2}\]
1977: and \[\widehat S_1(R(\lambda,
1978: \mu))=2a-\mu-\frac{\lambda^2(b+c+d+\mu-1)}{4-\mu^2},\] which is
1979: proportional to $-\lambda'a+b+c+d-(\mu'+1)$ for $(\lambda',
1980: \mu')=F(\lambda, \mu)$.
1981:
1982: Similarly
1983: \begin{multline*}
1984: \widehat S_2(R(\lambda, \mu))=b+c+d-\mu-1-\lambda^2(2a-\mu)^{-1}\\
1985: =b+c+d-\mu-1-\frac{\lambda^2(2a+\mu)}{4-\mu^2}\\
1986: =-\frac{2\lambda^2}{4-\mu^2}a+b+c+d-\left(\mu+\frac{\mu\lambda^2}{4-\mu^2}+1\right)\\
1987: =-\lambda''a+b+c+d-(\mu''+1),
1988: \end{multline*}
1989: where $(\lambda'', \mu'')=G(\lambda, \mu)$.
1990: \end{proof}
1991:
1992: \begin{remark}
1993: The equations of hyperbolas $H_\theta$ are related by the rule
1994: \[H_\theta(F(\lambda, \mu))=\frac{1-\mu^2}{\lambda^4}H_{\frac
1995: 12\sqrt{2-2\theta}}(\lambda, \mu) H_{-\frac
1996: 12\sqrt{2-2\theta}}(\lambda, \mu).\]
1997:
1998:
1999: The maps $F$ and $G$ are semi-conjugate to the map $x\mapsto
2000: 2x^2-1$ via the maps $\psi_F(x, y)=\frac{4-y^2+x^2}{4x}$ and
2001: $\psi_G(x, y)=\frac{4-x^2+y^2}{4y}$, respectively. This is the
2002: crucial point in computation of the spectrum of $R_n(\lambda,
2003: \mu)$.
2004: \end{remark}
2005:
2006: \begin{example}
2007: Let $H=H^{(3)}$ be the Hanoi Towers group (on three pegs) studied
2008: in~\cite{grisunik:hanoi} and~\cite{grisunik:hanoioberwolfach}. It
2009: is a self-similar group acting on $\xs$ for $\alb=\{0, 1, 2\}$
2010: with generators $a, b, c$ satisfying the following matrix
2011: recursions
2012: \begin{align*} a &=\left(\begin{array}{ccc}0 & 1 & 0\\ 1 & 0 & 0\\ 0 & 0 &
2013: a\end{array}\right),\\
2014: b &=\left(\begin{array}{ccc}0 & 0 & 1\\ 0 & b & 0\\ 1 & 0 &
2015: 0\end{array}\right),\\
2016: c &=\left(\begin{array}{ccc}c & 0 & 0\\ 0 & 0 & 1\\ 0 & 1 &
2017: 0\end{array}\right).
2018: \end{align*}
2019: Consider the two-parametric family of elements of the
2020: $C^*$-algebra $\mathcal{A}_{mes}$.
2021: \begin{multline*}
2022: \Delta(x, y)=\left(\begin{array}{ccc}c-x & y & y\\
2023: y & b-x & y\\ y & y & a-x\end{array}\right)=\\
2024: \left(\begin{array}{ccc|ccc|ccc}c-x & 0 & 0 & y & 0 & 0 & y & 0 &
2025: 0\\
2026: 0 & -x & 1 & 0 & y & 0 & 0 & y &
2027: 0\\
2028: 0 & 1 & -x & 0 & 0 & y & 0 & 0 & y\\ \hline
2029: y & 0 & 0 & -x & 0 & 1 & y & 0 & 0\\
2030: 0 & y & 0 & 0 & b-x & 0 & 0 & y & 0\\
2031: 0 & 0 & y & 1 & 0 & -x & 0 & 0 & y\\ \hline
2032: y & 0 & 0 & y & 0 & 0 & -x & 1 & 0\\
2033: 0 & y & 0 & 0 & y & 0 & 1 & -x & 0\\
2034: 0 & 0 & y & 0 & 0 & y & 0 & 0 & a-x
2035: \end{array}\right).
2036: \end{multline*}
2037: Permuting rows and columns and dividing them into blocks we get
2038: the matrix
2039: \[\left(\begin{array}{ccc|cccccc}
2040: c-x & 0 & 0 & y & 0 & 0 & y & 0 & 0 \\
2041: 0 & b-x & 0 & 0 & y & 0 & 0 & y & 0 \\
2042: 0 & 0 & a-x & 0 & 0 & y & 0 & 0 & y \\ \hline
2043: y & 0 & 0 & -x & 0 & 1 & y & 0 & 0 \\
2044: 0 & y & 0 & 0 & -x & 0 & 0 & y & 1 \\
2045: 0 & 0 & y & 1 & 0 & -x & 0 & 0 & y \\
2046: y & 0 & 0 & y & 0 & 0 & -x & 1 & 0 \\
2047: 0 & y & 0 & 0 & y & 0 & 1 & -x & 0 \\
2048: 0 & 0 & y & 0 & 1 & y & 0 & 0 & -x
2049: \end{array}\right).\]
2050: Computation of Schur complement with respect to the given
2051: partition of the matrix yields
2052: \[S_1(\Delta(x, y))=\Delta(x', y'),\]
2053: where
2054: \[x'=x-\frac{2(x^2-x-y^2)y^2}{(x-y-1)(x^2-1+y-y^2)}\]
2055: and
2056: \[y'=\frac{(x+y-1)y^2}{(x-y-1)(x^2-1+y-y^2)}.\]
2057: These rational functions were calculated in~\cite{grisunik:hanoi}
2058: using a different base. As it was observed
2059: in~\cite{grisunik:hanoi}, the map $F:(x, y)\mapsto (x', y')$ is
2060: semiconjugate to the map $f:\R\arr\R:x\mapsto x^2-x-3$. The
2061: spectrum of $\Delta(x, y)$ is in this case the union
2062: $\bigcup_{\theta\in\bigcup f^{-n}(S)}\mathcal{H}_\theta\cup
2063: L_0\cup L_1\cup L_2$, where $S=\{-2, 0\}$, and
2064: $\mathcal{H}_\theta$ is the hyperbola $x^2-xy-2y^2-\theta y=1$ and
2065: $L_0, L_1, L_2$ are the lines given by the equations
2066: \begin{align*}
2067: x-1-2y &= 0,\\
2068: x+1+y &= 0,\\
2069: x-1+y &= 0.
2070: \end{align*}
2071: A part of the spectrum is drawn on Figure~\ref{fig:hanoi}.
2072:
2073: \begin{figure}[h]
2074: \centering\includegraphics{hanoi.eps} \caption{Spectrum of
2075: $H_3$}\label{fig:hanoi}
2076: \end{figure}
2077:
2078: \end{example}
2079:
2080: \begin{example}
2081: Let $B=\img{z^2-1}$ be the Basilica group, studied
2082: in~\cite{zukgrigorchuk:3st,zukgrigorchuk:3stsp,barthvirag,bgn,nek:book}.
2083: It is realized as a self-similar group acting on the binary tree
2084: and generated by $a, b$, which are given by the matrix recursions
2085: \[a=\left(\begin{array}{cc}1 & 0\\ 0 & b\end{array}\right),\quad
2086: b=\left(\begin{array}{cc}0 & 1\\ a & 0\end{array}\right).\]
2087: Consider the pencil
2088: \[R(\lambda, \mu)=a+a^{-1}+\lambda(b+b^{-1})-\mu=
2089: \left(\begin{array}{cc}2-\mu & \lambda(1+a^{-1})\\
2090: \lambda(1+a) & b+b^{-1}-\mu\end{array}\right).\]
2091:
2092: We have
2093: \begin{multline*}
2094: \widehat S_2(R(\lambda,
2095: \mu))=b+b^{-1}-\mu-\frac{\lambda^2(2+a+a^{-1})}{2-\mu}\\
2096: =\frac{\lambda^2}{\mu-2}(a+a^{-1})+b+b^{-1}-\mu+\frac{2\lambda^2}{\mu-2}\\
2097: \sim
2098: a+a^{-1}+\frac{\mu-2}{\lambda^2}(b+b^{-1})-\left(\frac{\mu(\mu-2)}{\lambda^2}-2\right),
2099: \end{multline*}
2100: where ``$\sim$'' means ``proportional''. We get hence the rational
2101: map
2102: \begin{equation}\label{eq:basilica}
2103: \left\{\begin{array}{rcl}\lambda & \mapsto &
2104: \frac{\mu-2}{\lambda^2},\\
2105: \mu &\mapsto & -2+\frac{\mu(\mu-2)}{\lambda^2}.\end{array}\right.
2106: \end{equation}
2107:
2108: This rational map is quite complicated and the structure of the
2109: spectrum of the pencil $R(\lambda, \mu)$ is unclear. Computer
2110: experiments suggest that it has to have a structure of a ``strange
2111: attractor''. Most probably the map~\eqref{eq:basilica} is not
2112: semiconjugate to any one-dimensional map.
2113:
2114: The pencil $R(\lambda, \mu)$ is not invariant with respect to the
2115: first Schur complement (because the inverse of $b+b^{-1}-\mu$ is
2116: not a finite sum).
2117: \end{example}
2118:
2119: \begin{example}
2120: Let $G=\img{z^2+i}$ be the group studied
2121: in~\cite{bgn,nek:book,grisavchuksunic:img}. It is generated by $a,
2122: b, c$ given by the matrix recursions
2123: \[a=\left(\begin{array}{cc}0 & 1\\ 1 & 0\end{array}\right),\quad
2124: b=\left(\begin{array}{cc}a & 0\\ 0& c\end{array}\right),\quad
2125: c=\left(\begin{array}{cc}b & 0\\ 0 & 1\end{array}\right).\] It is
2126: a branch group of intermediate growth. For the notion of branch
2127: groups see~\cite{handbook:branch} and~\cite{grigorchuk:branch};
2128: the proof of intermediate growth of $\img{z^2+i}$ is given
2129: in~\cite{buxperez:imgi}. Consider the pencil
2130: \[M(y, z, \lambda)=a+yb+zc-\lambda=M(y, z, \lambda)=\left(\begin{array}{cc}
2131: ya+zb-\lambda & 1\\ 1 & yc+z-\lambda\end{array}\right),\] then
2132: \[\widehat S_1(M(y, z, \lambda)=M(y', z', \lambda'),\] where $\Phi:(y, z,
2133: \lambda)\mapsto (y', z', \lambda')$ is the map
2134: \[\Phi:\left\{\begin{array}{rcl} y & \mapsto & \frac{z}{y},\\
2135: z & \mapsto &
2136: \frac{1}{(z-\lambda-y)(z-\lambda+y)},\\
2137: \lambda & \mapsto & \frac{-\lambda
2138: y^2+\lambda(z-\lambda)^2+z-\lambda}{y(z-\lambda-y)(z-\lambda+y)}.\end{array}\right.\]
2139: It is not known if $\Phi$ is semiconjugate to a map of a smaller
2140: dimension and therefore there is no information about the
2141: structure of the spectrum of the pencil $M(y, z, \lambda)$.
2142:
2143: In this example, as also in the previous one, we have difficulty
2144: to apply $S_2$, since in these cases we do not get finite
2145: combinations of the group elements.
2146: \end{example}
2147:
2148: \begin{example}
2149: Consider now the free group from Example~\ref{sss:free}. The
2150: corresponding matrix recursion is
2151: \[a_{n+1}=\left(\begin{array}{cc} b_n & 0\\ 0 &
2152: b_n\end{array}\right),\quad b_{n+1}=\left(\begin{array}{cc} 0 & c_n\\
2153: a_n & 0\end{array}\right),\quad c_{n+1}=\left(\begin{array}{cc} 0 & a_n\\
2154: c_n & 0\end{array}\right).\] For the inverses we have
2155: \[a_{n+1}^{-1}=\left(\begin{array}{cc} b_n^{-1} & 0\\ 0 &
2156: b_n^{-1}\end{array}\right),\quad b^{-1}_{n+1}=\left(\begin{array}{cc} 0 & a_n^{-1}\\
2157: c_n^{-1} & 0\end{array}\right),\quad c^{-1}_{n+1}=\left(\begin{array}{cc} 0 & c_n^{-1}\\
2158: a_n^{-1} & 0\end{array}\right).\] The problem is to find the
2159: spectrum of the matrix
2160: \begin{multline*}M_n=a_n+a^{-1}_n+b_n+b_n^{-1}+c_n+c_n^{-1}=\\ \left(\begin{array}{cc}
2161: b_{n-1}+
2162: b^{-1}_{n-1} & a_{n-1}+a^{-1}_{n-1}+c_{n-1}+c^{-1}_{n-1}\\
2163: a_{n-1}+a^{-1}_{n-1}+c_{n-1}+c^{-1}_{n-1} &
2164: b_{n-1}+b^{-1}_{n-1}\end{array}\right),\end{multline*} but
2165: introduction of new parameters and computation of Schur
2166: complements unfortunately does not lead us to a success.
2167: \end{example}
2168:
2169: \begin{question} What is the spectrum of the matrix $M_n$? Does there exist
2170: $\epsilon>0$ such that for every $n$ the spectrum of $M_n$ belongs
2171: to the set $[-6+\epsilon, 6-\epsilon]\cup\{6, -6\}$? If the answer
2172: is ``yes'', then the sequence of the Schreier graphs of the action
2173: of $\langle a, b, c\rangle$ on the levels of the tree $\xs$ is a
2174: sequence of expanders. (Here a \emph{Schreier graph} of a group
2175: generated by a finite set $S$ and acting on a set $M$ is the graph
2176: with the set of vertices $M$ in which two vertices $x, y\in M$ are
2177: connected by an edge if one is the image of the other under the
2178: action of an element of $S$.)
2179: \end{question}
2180:
2181:
2182: \begin{example}
2183: Consider the group generated by the transformations $a, b, c$ of
2184: $\xs$ for $\alb=\{0, 1\}$ given by the recurrent relations
2185: \begin{alignat*}{2}
2186: a(0w) &= 1b(w), &\quad a(1w) &= 0b(w),\\
2187: b(0w) &= 0a(w), &\quad b(1w) &= 1c(w),\\
2188: c(0w) &= 0c(w), &\quad c(1w) &= 1a(w),
2189: \end{alignat*}
2190: i.e., the group generated by the automaton shown on
2191: Figure~\ref{fig:freepr}.
2192:
2193: \begin{figure}
2194: \includegraphics{freepr.eps}\\
2195: \caption{An automaton generating $C_2*C_2*C_2$}\label{fig:freepr}
2196: \end{figure}
2197:
2198: This group is isomorphic (by a result of Y.~Muntyan and
2199: D.~Savchuk) to the free product $C_2*C_2*C_2$ of three groups of
2200: order 2, see~\cite{nek:book} Theorem~1.10.2. The elements $a, b,
2201: c$ are of order 2.
2202:
2203: The corresponding matrix recursions are \[
2204: a_{n+1}=\left(\begin{array}{cc} 0 & b_n \\ b_n &
2205: 0\end{array}\right),\quad b_{n+1}=\left(\begin{array}{cc} a_n & 0
2206: \\ 0 & c_n\end{array}\right),\quad c_{n+1}=\left(\begin{array}{cc} c_n & 0
2207: \\ 0 & a_n\end{array}\right).\]
2208: \end{example}
2209:
2210: \begin{question}
2211: What is the spectrum of the matrices
2212: \[M_n=a_n+b_n+c_n=\left(\begin{array}{cc} a_{n-1}+c_{n-1} & b_n \\ b_n &
2213: a_{n-1}+c_{n-1}\end{array}\right)?\] In particular, is the
2214: spectrum of $M$ a subset of $(-2\sqrt{2}, 2\sqrt{2})\cup\{3\}$? If
2215: it is, then the Schreier graphs of the action of the group on the
2216: levels of the tree $\xs$ are Ramanujan.
2217: \end{question}
2218:
2219: \section{Analytic families}
2220:
2221: Let $\psi:H\arr H_1\oplus H_2$ be an isomorphism and let
2222: \[M(z)=\left(\begin{array}{cc}A(z) & B(z)\\ C(z) &
2223: D(z)\end{array}\right)\] be an analytic on some domain $\Omega$
2224: operator valued function. Assume that $D(z)$ s invertible on
2225: $\Omega$. Then \[S_1(M(z))=A(z)-B(z)D^{-1}(z)C(z)\] is analytic on
2226: $\Omega$. We get therefore a partially defined Schur map between
2227: the spaces of analytic operator valued functions.
2228:
2229: If $H$ is a Hilbert space with a $d$-similarity $\psi:H\arr H^d$,
2230: then we can define in a similar way partially defined Schur
2231: transformations $\wt S_i$ on the space of analytic operator valued
2232: functions on $H$.
2233:
2234: \begin{example}\label{ex:resolvent} Let $M=\left(\begin{array}{cc}A & B\\ C &
2235: D\end{array}\right)$ and consider the function
2236: \[M-z=\left(\begin{array}{cc}A-z & B\\ C & D-z\end{array}\right).\]
2237: It is mapped by the Schur map to $S_1=A-z-B(D-z)^{-1}C$ and we
2238: have
2239: \[(M-z)^{-1}=\left(\begin{array}{cc}S_1^{-1} & -S_1^{-1}B(D-z)^{-1}\\
2240: -(D-z)^{-1}CS_1^{-1} &
2241: -(D-z)^{-1}CS_1^{-1}B(D-z)^{-1}+(D-z)^{-1}\end{array}\right).\] It
2242: follows from~\eqref{eq:frobenius} in
2243: Theorem~\ref{th:schurcomplement} that $S_1((M-z)^{-1})=(A-z)^{-1}$
2244: and, similarly, $S_2((M-z)^{-1})=(D-z)^{-1}$.
2245: \end{example}
2246:
2247: Consider now an operator-valued holomorphic function of $n$
2248: complex variables
2249: \[M:\C^n\arr B(H)\]
2250: and let again $\psi:H\arr H^d$ be a $d$-similarity. We will denote
2251: the respective Schur transformations $\wt S_i$ just by $S_i$.
2252:
2253: \begin{definition}
2254: The function $M(z)$ is \emph{self-similar} (with respect to $\wt
2255: S_i$) if there is a function $F:\C^n\arr\C^n$ such that
2256: \[S_i(M(z))=M(F(z)).\]
2257: \end{definition}
2258:
2259: We can consider functions $M$ from the projective space $\C
2260: P^{n-1}$ to $B(H)$. Self-similarity of such functions is defined
2261: analogically.
2262:
2263: \begin{example}
2264: Let $G$ be a self-similar group generated by $\{a_1, \ldots,
2265: a_k\}$ and acting on a rooted $d$-regular tree $\xs$. Consider the
2266: pencil of Hecke type operators
2267: \[M(z)=M(z_1, \ldots, z_k)=\sum_{i=1}^kz_i\pi(a_i+a_i^{-1}),\]
2268: where $\pi$ is the self-similar representation of $G$ on $L^2(\xo,
2269: \nu)$. The operators $M(z)$ belong to the algebra $\A_{mes}$.
2270:
2271: Assume that $M(z)$ is a self-similar function with respect to
2272: $S_i$. This means that
2273: \[S_i(M(z_1, \ldots, z_k))=\sum_{i=1}^kz_i'\pi(a_i+a_i^{-1})\]
2274: where $z_j'=Z_j(z_1, \ldots, z_k)$ for $j=1, \ldots, k$ are some
2275: functions. We get hence a map $F=(Z_1, \ldots, Z_k):\C^k\arr\C^k$
2276: and the projectivized map $\wt F:\C P^{k-1}\arr\C P^{k-1}$.
2277: \end{example}
2278:
2279: \begin{proposition}
2280: \label{pr:Linvariant} The hyperspace
2281: $L=\left\{\sum_{i=1}^kz_i=0\right\}$ is forward $S_i$-invariant
2282: for all $i$.
2283: \end{proposition}
2284:
2285: Here forward $S_i$-invariance is the condition
2286: $S_i(L\cap\mathop{\mathrm{Dom}}(S_i))\subset L$.
2287:
2288: \begin{proof}
2289: \begin{lemma}
2290: If $\left(\begin{array}{c}x_1\\ x_2\end{array}\right)$ belongs to
2291: the kernel of $M$, then $x_1\in\ker S_1(M)$. Conversly, if
2292: $x_1\in\ker S_1(M)$ then there is $x_2$ such that
2293: $\left(\begin{array}{c}x_1\\ x_2\end{array}\right)\in\ker M$.
2294: \end{lemma}
2295:
2296: \begin{proof}
2297: We have \[\left(\begin{array}{cc}A & B \\ C &
2298: D\end{array}\right)\left(\begin{array}{c} x_1\\
2299: x_2\end{array}\right)=\left(\begin{array}{c}Ax_1+Bx_2 \\
2300: Cx_1+Dx_2\end{array}\right)=\left(\begin{array}{c}0\\
2301: 0\end{array}\right).\] We get $x_2=-D^{-1}Cx_1$ if $D$ has right
2302: inverse. Consequently, $(A-BD^{-1}C)x_1=0$.
2303:
2304: Conversly, if $x_1\in\ker S_1(M)$, then
2305: $\left(\begin{array}{c}x_1\\ x_2\end{array}\right)$, where
2306: $x_2=-D^{-1}Cx_1$.
2307: \end{proof}
2308:
2309: Since $\sum_{i=1}^kz_i=0$ the constant function $c\ne 0$ on
2310: $\partial T$ belongs to $\ker M$. Its restriction onto the
2311: cylindrical set of words starting with $i$ (where $i$ is the same
2312: as in $S_i$) is also a constant function and by the above lemma,
2313: we have $S_i(M)c=0$, hence $\sum z_i'=0$. That means that
2314: $S_i(L\cap\mathop{\mathrm{Dom}}(S_i))\subset L$.
2315: \end{proof}
2316:
2317: Let \[\Sigma M(z)=\{z\;:\;M(z)\text{\ is not invertible}\}\] be
2318: the spectrum (critical set) of the pencil $M(z)$. We have then the
2319: following corollary of Theorem~\ref{th:schurcomplement}
2320:
2321: \begin{corollary}
2322: Let $M(z)$ and $A(z), B(z), C(z), D(z)$ be as before. Then
2323: \[\Sigma M(z)\backslash\Sigma D(z)=\Sigma S_1(M(z))\setminus \Sigma D(z).\]
2324: \end{corollary}
2325:
2326: If $M(z)$ is self-similar, i.e., if $S_1(M(z))=M(F(z))$, then
2327: \[\Sigma S_1(M(z))=\Sigma M(F(z))=F^{-1}(\Sigma(M(z))),\]
2328: and we get
2329:
2330: \begin{corollary}
2331: The spectrum $\Sigma M(z)$ is backward-invariant under $F$, i.e.,
2332: \[F^{-1}(\Sigma M(z))=\Sigma M(z)\]
2333: if and only if \[\Sigma D(z)\cap \Sigma M(z)=\Sigma D(z)\cap
2334: F^{-1}(\Sigma M(z)).\]
2335: \end{corollary}
2336:
2337: The condition of the corollary is not easy to check if
2338: $\Sigma=\Sigma M(z)$ is unknown. In all examples that were
2339: treated, though, we have $F^{-1}\Sigma=\Sigma$.
2340:
2341: \begin{question} Under what natural and easy-to-check
2342: conditions the equality $F^{-1}\Sigma=\Sigma$ is true?
2343: Under what conditions we have
2344: $\Sigma=\overline{\bigcup_{n=0}^\infty F^{-n}L}$?
2345: \end{question}
2346:
2347: In all treated examples the map $F$ is rational. Therefore the
2348: problem of finding the spectrum of a pencil in a self-similar
2349: group is related to the problem of description of invariant
2350: subsets of multidimensional rational mappings. This subject is of
2351: independent interest in dynamical systems, multidimensional
2352: complex analysis, etc (see~\cite{sibony}). Spectra of Hecke type
2353: elements in $C^*$-algebras related to self-similar groups is a big
2354: source of interesting examples of dynamical systems on the complex
2355: projective space.
2356:
2357: \section{Schur maps and random walks}
2358: Consider the map $J:B(H)\arr B(H):A\mapsto A+I$ where $I$ is the
2359: identity operator and suppose that we have fixed a $d$-similarity
2360: $\psi:H\arr H^d$. Consider the conjugates $k_i=JS_iJ^{-1}$ of the
2361: Schur maps. We call $k_i$ the \emph{probabilistic Schur maps} (the
2362: reason will be clarified later). If $M=\left(\begin{array}{cc}A &
2363: B\\ C & D\end{array}\right)$, then
2364: \begin{equation}\label{eq:k_1}k_1(M)=A+B(I-D)^{-1}C,\end{equation}
2365: where $I$ is the identity operator (or a matrix of the same size
2366: as $D$).
2367:
2368: Following Bartholdi, Virag~\cite{barthvirag} and
2369: Kaimanovich~\cite{kaimanovich:munchhausen}, we are going to apply
2370: the probabilistic Schur maps to study random walks on self-similar
2371: groups.
2372:
2373: Let $(G, \alb)$ be a self-similar group acting on a $d$-regular
2374: tree and let $\mu$ be a probability distribution on $G$. Thus,
2375: every element $g$ of $G$ has a mass $\mu_g$, $0\le \mu_g\le 1$ and
2376: $\sum_{g\in G}\mu_g=1$. The set $\supp\mu=\{g\;:\;\mu_g\ne 0\}$ is
2377: called the \emph{support} of $\mu$. The measure $\mu$ is
2378: \emph{non-degenerate} if $\supp\mu$ generates $G$. We will
2379: identify $\mu$ with the element
2380: \[\mu=\sum_{g\in G}\mu_g g\] of $\ell^1(G)$. Moreover, $\mu$
2381: belongs to the simplex
2382: \[\ell_+^1(G)=\{\mu\in\ell^1(G)\;:\;0\le\mu_g\le 1, \sum_{g\in G}\mu_g\}.\]
2383: The left random walk on $G$ generated by $\mu$ starts at some
2384: element of $G$ (usually at the identity) and at each step makes
2385: the move $g\mapsto hg$ with probability $\mu(h)$.
2386:
2387: When started at $e$ (the identity element of $G$) the distribution
2388: at the moment $n$ is given by the $n$th convolution
2389: $\mu^{(n)}=\underbrace{\mu*\cdots *\mu}_n$ of $\mu$ and the
2390: probability of return $P_{e, e}^{(n)}$ is equal to $\mu^{(n)}(e)$.
2391: Left random walk on $G$ is invariant with respect to the right
2392: action of $G$ on itself.
2393:
2394: The main topics that interest specialist in random walks are:
2395: \begin{enumerate}
2396: \item norm of the Markov operator \[M=\sum_{g\in G}\mu_gL_g\]
2397: in $\ell^2(G)$, where $L_g$ are the operators of the left regular
2398: representation. We identify here the elements of $\ell^1(G)$ with
2399: the left convolution operators on $\ell^2(G)$;
2400:
2401: \item spectrum of $M$;
2402:
2403: \item asymptotic behavior of $P_{e, e}^{(n)}$ when $n\to \infty$;
2404:
2405: \item Liouville property (i.e., when all harmonic functions are
2406: constant);
2407: \end{enumerate}
2408: and other asymptotic characteristics of the random walks.
2409:
2410: Let $u$ be one of the vertices of the first level of the tree and
2411: let $H=\st_G(u)$ be its stabilizer in $G$. As $H$ has finite index
2412: in $G$, the random walk hits $H$ with probability 1. Let $\mu_H$
2413: be the distribution on $H$ given by the probability of the first
2414: hit, i.e.,
2415: \[\mu_H(h)=\sum_{n=0}^\infty f_{e, h}^{(n)},\]
2416: where $f_{e, h}^{n}$ is the probability to hit $H$ at the element
2417: $h$ for the first time at step $n$. As $H<G$ is a recurrent set as
2418: a subgroup of finite index in $G$, we have $\sum_{h\in
2419: H}\mu_H(h)=1$. Let $p_i:H\arr G$ be the $i$th projection map
2420: $h\mapsto h|_i$ for $1\le i\le d$, and let $\mu_i$ be the image of
2421: $\mu_H$ under $p_i$.
2422:
2423: The next theorem and its proof are analogous to Theorem 2.3 of
2424: V.~Kaimanovich~\cite{kaimanovich:munchhausen}, but they are
2425: formulated a bit differently.
2426:
2427: \begin{theorem}
2428: In the above conditions
2429: \[\mu_i=k_i(\mu).\]
2430: \end{theorem}
2431:
2432: \begin{proof}
2433: Let $M=(m_{ij})_{1\le i, j\le d}$ be the matrix representation of
2434: the Markov operator $M$ coming from the wreath recursion
2435: $\phi:G\hookrightarrow \symm\wr_\alb G$. Let us extend it to the
2436: map $\ell_+^1(G)\arr M_d(\ell_+^1(G))$.
2437:
2438: Then the measures $\mu_i$ can be expressed as
2439: \begin{equation}\label{eq:mii}\mu_i=m_{ii}+M_{i\overline i}\left(I+M_{\overline i\overline
2440: i}+ M_{\overline i\overline i}^2+\cdots\right)M_{\overline i
2441: i}=\mu_{ii}+M_{i\overline i}\left(I-M_{\overline i\overline
2442: i}\right)^{-1}M_{\overline ii},\end{equation} where $M_{i\overline
2443: i}$ (respectively, $M_{\overline i i}$) denotes the row
2444: $(m_{ij})_{j\ne i}$ (respectively, the column $(m_{ji})_{j\ne i}$)
2445: of the matrix $M$ with deleted element $m_{ii}$, and $M_{\overline
2446: i\overline i}$ is the $(d-1)\times (d-1)$-matrix obtained from $M$
2447: by removing its $i$th row and $i$th column.
2448:
2449: The first term in~\eqref{eq:mii} corresponds to staying at the
2450: point $i$ (in the random walk on $X$ induced by the random walk on
2451: $G$), while the first factor of the second term corresponds to
2452: moving from $i$ to $\alb\setminus\{i\}$, the second its factor
2453: corresponds to staying in $\alb\setminus\{i\}$ and the third
2454: factor of the second term corresponds to moving back from
2455: $\alb\setminus\{i\}$ to $i$.
2456:
2457: Comparing~\eqref{eq:mii} with~\eqref{eq:k_1} we get the statement
2458: of the theorem.
2459: \end{proof}
2460:
2461: In case $\st_G(i)=\st_G(1)$ for all $i\in\alb$ (for instance, if
2462: $\xs$ is a binary tree, or if $G$ acts on the first level by
2463: powers of a transitive cycle $\sigma$) we can interpret the above
2464: fact as that we have a sequence of stopping times $\tau(n)$ such
2465: that if
2466: \[\phi(Y^{(n)})=\sigma^{(n)}\left(Y^{(n)}_1, Y^{(n)}_2,\ldots,
2467: Y^{(n)}_d\right)\] is image of the $n$th step $Y^{(n)}$ of the
2468: random walk under the wreath recursion $\phi$, then
2469: \[\phi(Y^{\tau(n)})=\left(Y^{(\tau(n))}_1, Y^{(\tau(n))}_2, \ldots Y^{(\tau(n))}_d\right),\]
2470: i.e., the random element at time $\tau(n)$ belongs to the
2471: stabilizer of the first level, and $Z_i^{(n)}=Y_i^{(\tau(n))}$ is
2472: the random walk on $G$ determined by the measure $\mu_i$. Thus we
2473: can treat asymptotic characteristics of $\mu$-random walk on $G$
2474: via $\mu_i$-random walks on $G$. Of course for complete
2475: reconstruction of $(G, \mu)$ we also need to know the joint
2476: distributions of the processes $(G, \mu_i)$ (as they are usually
2477: not independent). Nevertheless, some information about the random
2478: walk can be obtained without the independence.
2479:
2480: The maps $k_i:\mathcal{M}\arr\mathcal{M}:\mu\mapsto\mu_i$ on the
2481: simplex of measures on $G$ are continuous and their fixed points
2482: are of a special interest for us. The fixed points always exist
2483: (for example the unit mass concentrated on the identity), but we
2484: are interested only in non-degenerate fixed points (i.e., with
2485: support generating the group).
2486:
2487: \begin{remark}
2488: Invariance of the hyperplane $L=\{x\in\ell^1(G)\;:\;\sum_{g\in
2489: G}x_g=0\}$ under the Schur map (the proof of which is analogous to
2490: the proof of Proposition~\ref{pr:Linvariant}) implies
2491: $k_i$-invariance of the hyperplane $L=\{\lambda\;:\;\sum_{g\in
2492: G}\lambda_g=1\}$. The probabilistic meaning of the maps $k_i$
2493: shows that the simplex of probability measures $\mathcal{M}\subset
2494: L$ is also $k_i$-invariant.
2495: \end{remark}
2496:
2497: Now we are going to describe briefly V.~Kaimanovich's approach to
2498: testing amenability of $G$ by entropy method (which develops the
2499: ideas previously expressed in~\cite{barthvirag}).
2500:
2501: Having a left random walk $g_{n+1}=h_{n+1}g_n$, given by $(G,
2502: \mu)$, where $(h_n)_{n\ge 0}$ is a sequence of independent
2503: $\mu$-distributed random variables, we consider the induced Markov
2504: chain, denoted $(\alb\cdot G, \mu)$, on the $G$-bimodule
2505: $\alb\cdot G$ seen as a left $G$-space:
2506: \[x_{n+1}\cdot g_{n+1}=h_{n+1}\cdot(x_n\cdot
2507: g_n)=h_{n+1}(x_n)\cdot h_{n+1}|_{x_n}g_n.\]
2508:
2509: If we start the Markov chain $(\alb\cdot G, \mu)$ at the point
2510: $x=x\cdot 1\in\alb\cdot G$, we get a projection $\Pi_x:g_n\mapsto
2511: g_n\cdot x=g_n(x)\cdot g_n|_x$ of the random walk on $G$ onto the
2512: Markov chain on $\alb\cdot G$. If the action $(G, \alb)$ is
2513: self-replicating (Definition~\ref{def:replicating}), then the map
2514: $\Pi_x:G\arr G\cdot\alb:g\mapsto g\cdot x$ is onto.
2515:
2516: The tuple $(\Pi_1(g), \ldots, \Pi_d(g))=(g(1)\cdot g|_1, \ldots,
2517: g(d)\cdot g|_d)$ determines $g$ uniquely, since it means that the
2518: image of $g$ in $\symm\wr_\alb G$ is $\pi(g|_1, \ldots, g|_d)$,
2519: where $\pi$ is the permutation $i\mapsto g(i)$, $i\in\alb$.
2520:
2521: The right action of $G$ on the bimodule $\alb\cdot G$ commutes
2522: with the left action, hence the Markov chain $(\alb\cdot G, \mu)$
2523: is invariant under the right action of $G$. The quotient of
2524: $\alb\cdot G$ by the right $G$-action is naturally identified with
2525: $\alb$ (the orbit of $x\cdot g$ is labeled by $x$). Consequently,
2526: we get a Markov chain on $\alb$ equal to the quotient of the chain
2527: $(\alb\cdot G, \mu)$ by the right $G$-action. It is easy to see
2528: that this chain is given by the action of $G$ on $\alb$:
2529: \[x\mapsto y,\qquad\text{with probability
2530: $\sum_{g(x)=y}\mu(g)$.}\] This chain is irreducible when $G$ acts
2531: transitively on the first level $\alb$ and $\mu$ is
2532: non-degenerate.
2533:
2534: Consider now the trace of the Markov chain $(\alb\cdot G, \mu)$ on
2535: the right $G$-orbit $x\cdot G=\{x\cdot g\;:\;g\in G\}$ for a fixed
2536: letter $x\in\alb$, which is a recurrent set (due to irreducibility
2537: of the quotient chain), i.e., consider the Markov chain with the
2538: first return transitions. We can identify the points of the orbit
2539: $x\cdot G$ with the group $G$ using the map $x\cdot g\mapsto g$
2540: and get in this way a Markov chain on $G$. It is not hard to check
2541: that this Markov chain is the left random walk defined by the
2542: measure $\mu_x=k_x(\mu)$, constructed above
2543: (see~\cite{kaimanovich:munchhausen}).
2544:
2545: \begin{example}
2546: Let $B=\langle a, b\rangle$ be the Basilica group generated by the
2547: states of the automaton shown on Figure~\ref{fig:basilica}. We
2548: have $\phi(a)=\left(\begin{array}{cc} 1 & 0\\ 0 &
2549: b\end{array}\right)$ and $\phi(b)=\left(\begin{array}{cc}0 & a\\
2550: 1 & 0\end{array}\right)$. Then $\phi(a^{-1})=\left(\begin{array}{cc} 1 & 0\\
2551: 0 & b^{-1}\end{array}\right)$ and
2552: $\phi(b^{-1})=\left(\begin{array}{cc} 0 & 1\\ a^{-1} &
2553: 0\end{array}\right)$. Take
2554: $\mu=p(a+a^{-1})+q(b+b^{-1})\in\ell_+^1(B)$, where $2(p+q)=1$.
2555:
2556: The transition moves for the Markov chain on $\{0, 1\}\times B$
2557: are
2558: \begin{align*}
2559: 0\cdot g & \stackrel{a}{\arr} a(0)\cdot a|_0g=0\cdot g,\\
2560: 0\cdot g & \stackrel{a^{-1}}{\arr} a^{-1}(0)\cdot a^{-1}|_0g=0\cdot g,\\
2561: 1\cdot g & \stackrel{a}{\arr} a(1)\cdot a|_1g=1\cdot bg,\\
2562: 1\cdot g & \stackrel{a^{-1}}{\arr} a^{-1}(1)\cdot
2563: a^{-1}|_1g=1\cdot b^{-1}g,
2564: \end{align*}
2565: and
2566: \begin{align*}
2567: 0\cdot g & \stackrel{b}{\arr} b(0)\cdot b|_0g=1\cdot g,\\
2568: 0\cdot g & \stackrel{b^{-1}}{\arr} b^{-1}(0)=1\cdot a^{-1}g,\\
2569: 1\cdot g & \stackrel{b}{\arr} b(1)\cdot b|_1g=0\cdot ag,\\
2570: 1\cdot g & \stackrel{b^{-1}}{\arr} b^{-1}(1)\cdot
2571: b^{-1}|_1g=0\cdot g.
2572: \end{align*}
2573: \end{example}
2574:
2575: \begin{figure}
2576: \includegraphics{graph.eps}\\
2577: \caption{The random walk on $\{0, 1\}\cdot B$}\label{fig:graph}
2578: \end{figure}
2579:
2580: See a graphical description of the random walk on $\{0, 1\}\cdot
2581: B$ on Figure~\ref{fig:graph}. Figure~\ref{fig:smallgraph} shows
2582: the graph of the induced random walk on $\{0, 1\}=\alb$ with
2583: transition probabilities $1/2$.
2584:
2585: \begin{figure}
2586: \includegraphics{smallgraph.eps}\\
2587: \caption{The induced random walk on $\alb$}\label{fig:smallgraph}
2588: \end{figure}
2589:
2590: The main characteristics of a random walk $(G, \mu)$ on groups
2591: are:
2592: \begin{itemize}
2593: \item[(i)] the spectral radius of the random walk
2594: \[r=\limsup_{n\to\infty}\sqrt[n]{\mu^{(n)}(e)};\]
2595: \item[(ii)] the drift
2596: \[\lambda=\lim_{n\to\infty}\frac{|g_n|}{n},\]
2597: where $\{g_n\}_{n\ge 1}$ is a random trajectory and $|g_n|$ is the
2598: length of an element;
2599: \item[(iii)] entropy \[h=\lim_{n\to\infty}\frac 1n H(\mu^{(n)}),\]
2600: where $H$ is the entropy of a probability distribution.
2601: \end{itemize}
2602:
2603: Vanishing of $\lambda$ implies vanishing of the entropy, which
2604: implies amenability of $G$ (in case of a non-degenerate measure
2605: $\mu$~\cite{kaimvershik}), while $r=1$ for a symmetric measure
2606: $\mu$ is equivalent to amenability (see~\cite{kesten:amen}).
2607:
2608: Kaimanovich's approach to amenability (called in his paper
2609: ``M\"unchhausen trick'') is based on the inequalities
2610: (see~\cite{kaimanovich:munchhausen} Theorem~3.1)
2611: \[h(G, \mu)\le h(G, \mu_i)\le |\alb|h(G, \mu)\]
2612: which holds for the projections $\mu_i$, $i=1, \ldots, d$,
2613: described above. His main observation based on these inequalities
2614: and properties of the entropy is that if $\mu$ is non-degenerate
2615: and self-affine, i.e., if there is $\alpha>0$ such that
2616: \[\mu_i=(1-\alpha)\delta_e+\alpha\mu\]
2617: for some $i$, then the inequalities imply $h(\mu)=0$ and hence
2618: that $G$ is amenable (and moreover, has Liouville property).
2619:
2620: Kaimanovich calls the measure $\mu$ satisfying the self-affinity
2621: condition \emph{self-similar}. Self-similar measures are fixed
2622: points of the maps
2623: \[\alpha_i:\mu\mapsto\frac{k_i(\mu)-k_i(\mu)(e)}{1-k_i(\mu)(e)},\]
2624: which are continuous maps of the simplex of probability measures
2625: on $G$. The simplex is compact with respect to the weak topology.
2626: For all $i$ there is an $\alpha_i$-invariant measure (indeed, the
2627: measure $\delta_e$ concentrated in identity is such a measure) but
2628: the problem is to find a non-degenerate fixed point if such exist.
2629:
2630: The next few examples show when this is indeed the case. Here we
2631: follow~\cite{kaimanovich:munchhausen}.
2632:
2633: \begin{example}
2634: Take $\mathfrak{G}=\langle a, b, c, d\rangle$ --- the group of
2635: Example~\ref{sss:grigorchuk}. Consider a one-dimensional family of
2636: measures
2637: \[\mu=2\alpha m_1+4\beta m_2=\alpha a+\beta b+\beta c+\beta d+\alpha+\beta,\]
2638: where $m_1=(1+a)/2$, $m_2=(1+b+c+d)/4$ and $2\alpha+4\beta=1$ and
2639: $0\le \alpha, \beta\le 1$.
2640:
2641: Then
2642: \begin{align*}
2643: \mu_1 &= \frac\alpha{1-\alpha}+4\beta
2644: m_1+\frac{4\alpha\beta}{1-\alpha}m_2\\
2645: \mu_2 &=
2646: \frac\alpha{1-\alpha}+\frac{4\alpha\beta}{1-\alpha}m_1+4\beta m_2.
2647: \end{align*}
2648:
2649: In terms of the parameter $\alpha\in (0, 1/2)$ the corresponding
2650: transformations $\phi_i$ take the form
2651: \begin{align*}
2652: \phi_1:\alpha &\mapsto
2653: \frac{2\beta}{1-\frac\alpha{1-\alpha}}=\frac{1-\alpha}2\\
2654: \phi_2:\alpha &\mapsto
2655: \left.\frac{2\alpha\beta}{1-\alpha}\right/\left(1-\frac\alpha{1-\alpha}\right)=
2656: \frac{2\alpha\beta}{1-2\alpha}=\frac\alpha 2.
2657: \end{align*}
2658: The only fixed point for $\phi_2$ corresponds to $\alpha=0$ and
2659: $\mu$ is degenerate in this case, while for $\phi_1$ the value
2660: $\alpha=1/3$ gives a fixed point corresponding to a non-degenerate
2661: self-affine measure
2662: \[\mu=\frac 23 m_1+\frac 13 m_2=\frac 5{12}+\frac 13 a+\frac
2663: 1{12}(b+c+d).\] Removing the atom at the identity we obtain a
2664: self-affine measure concentrated on the generating set $\{a, b, c,
2665: d\}$
2666: \[\wt\mu=\frac 47+\frac 17(b+c+d)\] with self-similarity
2667: coefficient $1/2$
2668: \[\wt\mu_1=\frac 12+\frac 12\wt\mu\]
2669: \end{example}
2670:
2671: \begin{example}
2672: Basilica group, $B=\langle a, b\rangle$ is defined by
2673: \begin{alignat*}{2} a(0w) &= 0w, &\quad a(1w) &= 1b(w)\\
2674: b(0w) &= 1w, &\quad b(1w) &= 0a(w),\end{alignat*} see
2675: Example~\ref{sss:basilica}. Consider the one-parameter family of
2676: measures
2677: \[\mu=\frac{a+a^{-1}+rb+rb^{-1}}{2(r+1)},\quad r\ge 0.\]
2678: Then
2679: \begin{multline*}k_2(\mu)=\frac{b+b^{-1}}{2(r+1)}+\frac{r^2(2+a+a^{-1})}{4(1+r)^2}\left(1-\frac
2680: 1{r+1}\right)^{-1}\\
2681: =\frac r{4(r+1)}(a+a^{-1})+\frac{b+b^{-1}}{2(r+1)}+\frac
2682: r{2(r+1)}\\
2683: =\frac{r+2}{2(r+1)}\left(\frac r{2(r+2)}(a+a^{-1})+\frac
2684: 1{r+2}(b+b^{-1})\right)+\frac r{2(r+1)}\\
2685: =p(r)\phi_2(\mu)+q(r),
2686: \end{multline*}
2687: where $p(r)=\frac{r+2}{2(r+1)}$, $q(r)=\frac r{2(r+1)}$ and
2688: \[\phi_2\left(\frac 1{2(r+1)}: \frac
2689: r{2(r+1)}\right)=\left(\frac r{2(r+2)}:\frac 1{r+2}\right)\] is
2690: the projectivization of the corresponding map, which represents
2691: the map $\Lambda:\frac 1{2r}\mapsto \frac{r}2$ of $\R$ or
2692: $z\mapsto\frac 1{4z}$ if $z=\frac 1{2r}$. This map has no fixed
2693: points, but is periodic of period $2$; $\Lambda^2=id$. This fact
2694: is used in~\cite{barthvirag} to prove amenability of $B$.
2695:
2696: At the same time, as it is shown
2697: in~\cite{kaimanovich:munchhausen}, the map $\phi_1$ has a fixed
2698: point, which represents a non-degenerate measure and this gives
2699: another way to prove amenability of $B$.
2700: \end{example}
2701:
2702: \begin{question}
2703: Using the M\"unchhausen trick construct new interesting examples
2704: of self-similar amenable but not elementary amenable groups.
2705: \end{question}
2706:
2707:
2708: \bibliographystyle{plain}
2709: \bibliography{nekrash,mymath}
2710:
2711: \end{document}
2712: