1: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2: %% %%
3: %% Title: %%
4: %% Non-ergodic actions, cocycles and superrigidity %%
5: %% %%
6: %% Authors: %%
7: %% David Fisher, Dave Witte Morris, and Kevin Whyte %%
8: %% %%
9: %% Date: %%
10: %% 29 July 2004 %%
11: %% %%
12: %% Latex2e file, 21 pages %%
13: %% requires NYJ.CLS style file, available from %%
14: %% http://nyjm.albany.edu:8000/nyj.cls %%
15: %% %%
16: %% to appear in %%
17: %% New York Journal of Mathematics %%
18: %% %%
19: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
20: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
21:
22:
23: \documentclass[reqno, hyp]{nyj}
24: \title{Non-ergodic actions, cocycles and superrigidity}
25:
26: \author{David Fisher}
27: \address{Department of Mathematics and Computer Science,
28: Lehman College -- CUNY,
29: 250 Bedford Park Boulevard W.,
30: Bronx, NY 10468}
31: \email{dfisher@lehman.cuny.edu,
32: http://comet.lehman.cuny.edu/fisher/}
33:
34: \author{Dave Witte Morris}
35: \address{Department of Mathematics and Computer Science,
36: University of Lethbridge,
37: 4401 University Drive,
38: Lethbridge, AB, T1K~3M4, Canada}
39: \email{Dave.Morris@uleth.ca,
40: http://people.uleth.ca/$\sim$dave.morris/}
41:
42: \author{Kevin Whyte}
43: \address{Department of Mathematics,
44: University of Illinois at Chicago,
45: 851 S.~Morgan Street,
46: Chicago, IL 60607--7045}
47: \email{kwhyte@math.uic.edu,
48: http://www2.math.uic.edu/$\sim$kwhyte/}
49:
50:
51:
52: \thanks{This research was partially supported by the National Science
53: Foundation (grants DMS--0226121, DMS--0100438, and DMS--0204576).
54: The first author was also partially supported by a PSC-CUNY
55: grant.}
56:
57: \keywords{Borel action, non-ergodic, ergodic component, Borel cocycle,
58: superrigidity, von Neumann Selection Theorem}
59:
60: \subjclass{28D15}
61:
62:
63: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
64: %% define math symbols: %%
65: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
66:
67: \newcommand{\symmdiff}{\mathbin{\Delta}}
68: \DeclareMathOperator{\Id}{Id}
69: \newcommand{\real}{\mathbb{R}}
70: \renewcommand{\natural}{\mathbb{N}}
71: \newcommand{\func}{\mathcal{F}}
72: \newcommand{\F}{\mathbf{F}}
73: \newcommand{\erg}{\Omega}
74: \newcommand{\cobound}{\delta}
75: \DeclareMathOperator{\Aut}{Aut}
76: \newcommand{\A}{\Aut_{[\mu]}(S)}
77: \newcommand{\B}{\mathcal{B}}
78: \DeclareMathOperator{\cpct}{Cpct}
79: \DeclareMathOperator{\dist}{dist}
80: \DeclareMathOperator{\diam}{diam}
81: \newcommand{\triv}{\mathord{\bf 1}}
82: \newcommand{\trivact}{\mathbb{I}}
83: \DeclareMathOperator{\Const}{Const}
84: \DeclareMathOperator{\EssRg}{EssRg}
85: \DeclareMathOperator{\Stab}{Stab}
86: \newcommand{\field}{\mathbb{F}}
87: \newcommand{\proj}[1]{\mathbb{P}(#1)}
88: \newcommand{\Unitary}{\mathord{\mathbf{U}\bigl( L^2(S) \bigr)}}
89: \DeclareMathOperator{\Prob}{Prob}
90:
91: \newcommand{\G}{\Gamma}
92: \newcommand{\g}{\gamma}
93: \newcommand{\Ga}{\mathbb G}
94: \newcommand{\Ks}{\mathcal K}
95:
96: % this makes set braces "{}" tall enough
97: % to fit what is in between
98: \newcommand{\bigset}[2]{\left\{\, #1
99: \mathrel{\left| \vphantom {\left\{ #1 \mid #2 \right\} }
100: \right.} #2 \,\right\} }
101:
102:
103: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
104: %% For cross-references within the paper: %%
105: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
106:
107: \renewcommand{\see}[1]{{\upshape(}see~\ref{#1}{\upshape)}}
108: \newcommand{\seeand}[2]{{\upshape(}see~\ref{#1} and~\ref{#2}{\upshape)}}
109: \newcommand{\cf}[1]{(cf.~\ref{#1})}
110: \newcommand{\pref}[1]{{\upshape(}\ref{#1}{\upshape)}}
111: \newcommand{\fullref}[2]{\ref{#1}\pref{#1-#2}}
112: \newcommand{\fullsee}[2]{{\upshape(}see~\ref{#1}\pref{#1-#2}{\upshape)}}
113:
114:
115: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
116: %% Set up theorem environments and numbering style: %%
117: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
118:
119: \renewcommand{\thesubsection}{\thesection\Alph{subsection}}
120:
121: \numberwithin{equation}{section}
122:
123: \newtheorem{thm}[equation]{Theorem}
124: \newtheorem{lem}[equation]{Lemma}
125: \newtheorem{prop}[equation]{Proposition}
126: \newtheorem{cor}[equation]{Corollary}
127:
128: \theoremstyle{definition}
129: \newtheorem{defn}[equation]{Definition}
130: \newtheorem{rem}[equation]{Remark}
131: \newtheorem{notn}[equation]{Notation}
132:
133: \newtheorem{ack}{Acknowledgments\ignorespaces}
134: \renewcommand{\theack}{}
135:
136: \renewcommand{\labelitemii}{$\circ$}
137:
138:
139: \newcounter{steppf}
140:
141: \newenvironment{steppf}[1][\unskip]{\refstepcounter{steppf}
142: \em
143: \medskip \noindent Step \thestep\ #1.\ }{\unskip\upshape}
144: \newcommand{\thestep}{\arabic{steppf}}
145:
146: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
147: %% END OF MACRO DEFINITIONS %%
148: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
149:
150:
151:
152: \begin{document}
153:
154: \begin{abstract}
155: This paper proves various results concerning non-ergodic actions
156: of locally compact groups and particularly Borel cocycles defined
157: over such actions. The general philosophy is to reduce the study
158: of the cocycle to the study of its restriction to each ergodic
159: component of the action, while being careful to show that all
160: objects arising in the analysis depend measurably on the ergodic
161: component. This allows us to prove a version of the superrigidity
162: theorems for cocycles defined over non-ergodic actions.
163: \end{abstract}
164:
165: \maketitle
166:
167: \tableofcontents
168:
169:
170: \section{Introduction}
171:
172: It is often the case that one has extensive information about each
173: ergodic component of an action, and would like to piece this local
174: information together (measurably) in order to obtain a global
175: conclusion about the entire action. This note addresses a number of
176: problems of this type, mostly dealing with cocycles. For example:
177: \begin{enumerate}
178: \item If $\alpha$ and $\beta$ are Borel cocycles, and the restriction
179: of~$\alpha$ to almost every ergodic component is cohomologous to the
180: restriction of~$\beta$, then $\alpha$ is cohomologous to~$\beta$
181: \see{CohoErg}.
182: \item If $\alpha$ is a Borel cocycle, and the restriction of~$\alpha$
183: to almost every ergodic component is cohomologous to a homomorphism
184: cocycle, then $\alpha$ is cohomologous to a homomorphism cocycle
185: \see{CohoHomo}.
186: \item If almost every ergodic component of a $G$-action has a certain
187: standard Borel $G$-space~$X$ as a measurable quotient, then $X$~is a
188: measurable quotient of the entire action \see{QuotErgComps}.
189: \end{enumerate}
190: (We consider only Borel actions of second countable, locally compact
191: groups on standard Borel spaces with a quasi-invariant probability
192: measure.)
193:
194: We also prove a superrigidity theorem for cocycles that applies to
195: non-ergodic actions \see{theorem:Gsuperrigidityne}. In fact our
196: work was motivated by the discovery, during the writing of
197: \cite{FisherMargulis}, that no proof of any version of
198: superrigidity for cocycles concerning cocycles over non-ergodic
199: actions, exists in the literature. For many applications to
200: non-ergodic actions, including those in \cite{FisherMargulis},
201: other information concerning the action and cocycle allows one to,
202: with some additional work make do with superrigidity theorems for
203: cocycles which are defined over ergodic actions. However, the
204: results in section \ref{section:superrigid} allow some
205: simplification of the arguments in section $5$ of
206: \cite{FisherMargulis}, and should have other applications as well.
207:
208: \begin{ack}
209: Most of this research was accomplished during a visit to the Institute
210: for Mathematical Research (FIM) of the Swiss Federal Institute of
211: Technology (ETH) in Zurich. We are pleased to thank FIM for its
212: hospitality, and for the financial support that made the visit possible.
213: We would also like to thank Scot Adams, A.~Kechris, and G.~A.~Margulis
214: for helpful comments during the course of the investigation.
215: \end{ack}
216:
217:
218:
219: \section{Some lemmas on measurability} \label{LemmaSect}
220:
221: This section records basic definitions and notation, and also proves that
222: various natural constructions of sets, functions, and actions yield
223: results that are measurable. Some of the conclusions are known (or even
224: well known), but others may be of independent interest.
225:
226: \subsection{Properties of standard Borel spaces}
227:
228: We assume the basic theory of Polish spaces, standard Borel spaces and
229: analytic Borel spaces, which can be found in a number of textbooks, such
230: as \cite[Chap.~3]{Arveson}. We recall the definition of a standard space
231: and an analytic set.
232:
233: \begin{defn} \
234: \begin{itemize}
235: \item A topological space is \emph{Polish} if it is homeomorphic to
236: a complete, separable metric space.
237: \item A Borel space is \emph{standard} if it is Borel isomorphic to
238: a Polish topological space.
239: \item The pair $(S,\mu)$ is a \emph{standard Borel probability space} if
240: $S$~is a standard Borel space and $\mu$ is a probability measure on~$S$.
241: \item A subset~$A$ of a standard Borel space $S$ is \emph{analytic} if
242: there exist
243: \begin{itemize}
244: \item a standard Borel space~$X$
245: and
246: \item a Borel map $\psi \colon X \to S$,
247: \end{itemize}
248: such that $A = \psi(X)$ is the image of~$\psi$.
249: \end{itemize}
250: \end{defn}
251:
252: \begin{rem}[{\cite[Thm.~3.2.4, p.~67]{Arveson}}] \label{Anal->Meas}
253: Any analytic subset~$A$ of any standard Borel space~$S$ is
254: \emph{absolutely measurable}; that is, for any probability measure~$\mu$
255: on~$S$, there exist Borel subsets $B_1$ and~$B_2$ of~$S$ with $B_1
256: \subset A \subset B_2$ and $\mu(B_2 \smallsetminus B_1) = 0$.
257: \end{rem}
258:
259: \begin{thm}[{von Neumann Selection Theorem, cf.\ \cite[Thm.~3.4.3,
260: p.~77]{Arveson}}] \label{vonNeumann}
261: Let
262: \begin{itemize}
263: \item $(\erg,\nu)$ be a standard Borel probability space,
264: \item $L$ be a standard Borel space,
265: \item $\func$ be an analytic subset of $\erg \times L$,
266: and
267: \item $\erg_{\func}$ be the projection of~$\func$ to~$\erg$.
268: \end{itemize}
269: Then there are
270: \begin{itemize}
271: \item a conull, Borel subset~$\erg_0$ of~$\erg_{\func}$,
272: and
273: \item a Borel function $\Phi \colon \erg_0 \to L$,
274: \end{itemize}
275: such that $\bigl( \omega, \Phi(\omega) \bigr) \in \func$, for all
276: $\omega \in \erg_0$.
277: \end{thm}
278:
279: \begin{notn}
280: Suppose $(S,\mu)$ is a standard Borel probability space, and
281: $(X,d)$~is a separable metric space.
282: \begin{enumerate}
283: \item We use $\F(S,X)$ to denote the space of
284: measurable functions $f \colon S \to X$, where two functions are
285: identified if they are equal almost everywhere. If $X$~is complete, then
286: $\F(S,X)$ is a Polish space, under the topology of convergence in
287: measure (cf.\ \cite[\S4.4]{WheedenZygmund}). A metric can be given by
288: $$ d_{\F}(f,g) = \min \bigset{ \vphantom{\Bigl|} \epsilon \ge 0}{
289: \mu \bigl\{\, s \in S \mid d \bigl( f(s), g(s) \bigr) > \epsilon
290: \,\bigr\} \le \epsilon} $$
291:
292: \item We use $\B(S)$ to denote the Boolean algebra of measurable subsets
293: of~$S$, where two subsets are identified if their symmetric difference
294: has measure~$0$. It is well known that this is a complete separable metric
295: space, with metric
296: $$ d_{\B}(A,B) = \mu(A \symmdiff B) .$$
297: \end{enumerate}
298: \end{notn}
299:
300: \begin{rem}
301: The $\sigma$-algebra of Borel subsets of $\F(S,X)$ is generated by the
302: sets of the form
303: $$ \Delta_{S_0,X_0,\epsilon} = \bigset{f }{ \mu \bigl( S_0 \cap
304: f^{-1}(X_0) \bigr) < \epsilon } ,$$
305: where $S_0$ and~$X_0$ are Borel subsets of $S$ and~$X$,
306: respectively, and $\epsilon > 0$. This implies that if $(X',d')$ is Borel
307: isomorphic to $(X,d)$, then $\F(S,X')$ is Borel isomorphic to $\F(S,X)$.
308: \end{rem}
309:
310: The following is well known:
311:
312: \begin{lem}[{cf.\ \cite[Lem.~7.1.3, p.~215]{MargulisBook}}]
313: \label{FiberFunc}
314: Let
315: \begin{itemize}
316: \item $L$ and~$S$ be standard Borel spaces,
317: \item $\mu$ be a probability measure on~$S$,
318: \item $X$ be a separable metric space,
319: and
320: \item $f \colon L \times S \to X$ be Borel.
321: \end{itemize}
322: Then
323: \begin{enumerate}
324: \item \label{FiberFunc-fiber}
325: for each $\ell \in L$, the function $f_\ell \colon S \to X$, defined by
326: $f_\ell(s) = f(\ell,s)$, is Borel,
327: and
328: \item \label{FiberFunc-Borel}
329: the induced function $\check f \colon L \to \F(S,X)$, defined by $\check
330: f(\ell) = f_\ell$, is Borel.
331: \end{enumerate}
332: \end{lem}
333:
334: In short, any Borel function from $L \times S$ to~$X$ yields a Borel
335: function from~$L$ to $\F(S,X)$. The converse is true:
336:
337: \begin{lem} \label{DefdFiberwise}
338: Let
339: \begin{itemize}
340: \item $L$ and $S$ be standard Borel spaces,
341: \item $\mu$ be a probability measure on~$S$,
342: \item $(X,d)$ be a separable metric space,
343: and
344: \item $\phi \colon L \to \F(S,X)$ be a Borel function.
345: \end{itemize}
346: Then there is a Borel function $\hat\phi \colon L \times S \to X$, such
347: that, for each $\ell \in L$, we have
348: \begin{equation} \label{phi(l,s)=phi(l)(s)}
349: \mbox{$\hat\phi(\ell,s) = \phi(\ell)(s)$ for a.e.\ $s \in
350: S$.}
351: \end{equation}
352: \end{lem}
353:
354: \begin{proof}
355: For each $n \in \natural$, let $\{D_n^i\}_{i=1}^\infty$ be a partition
356: of $\F(S,X)$ into countably many (nonempty) Borel sets of diameter
357: less than $2^{-n}$, and choose $\phi_n^i \in D_n^i$. Define
358: $\phi_n \colon L \times S \to X$ by
359: $$ \mbox{$\phi_n(\ell,s) = \phi_n^i(s)$ if $\phi(\ell) \in D_n^i$.}
360: $$
361: Then each $\phi_n$ is Borel.
362:
363: By replacing each $\{D_n^i\}_{i=1}^\infty$ by the join of $\{D_1^i\}_{i=1}^\infty,
364: \ldots, \{D_n^i\}_{i=1}^\infty$ we may assume that $\{D_{n+1}^i\}_{i=1}^\infty$ is
365: a refinement of $\{D_n^i\}_{i=1}^\infty$. For each $m,n \in \natural$ and each $\ell \in L$
366: $$ \mu \bigl\{\, s \in S \mid d\bigl( \phi_m(\ell,s), \phi_n(\ell,s)
367: \bigr) > 2^{-\min(m,n)} \,\bigr\} < 2^{-\min(m,n)} .$$
368: Thus, on each fiber $\{\ell\} \times S$, the sequence $\{\phi_n\}$ not only
369: converges in measure, but converges quickly. There is no harm in
370: assuming that $X$~is complete; then $\{\phi_n\}$ converges
371: pointwise a.e. Let $\hat\phi$ be the pointwise limit of
372: $\{\phi_n\}$. (Define $\hat\phi$ to be constant on the set where
373: $\{\phi_n\}$ does not converge.) Then $\hat\phi$ is Borel, and
374: satisfies \pref{phi(l,s)=phi(l)(s)}.
375:
376: \end{proof}
377:
378: \begin{cor} \label{F(F)}
379: Let
380: \begin{itemize}
381: \item $L$ and $S$ be standard Borel probability spaces,
382: and
383: \item $(X,d)$ be a separable metric space.
384: \end{itemize}
385: Then $\F( L \times S, X)$ is naturally homeomorphic to $\F \bigl( L, \F(
386: S, X) \bigr)$.
387: \end{cor}
388:
389: \begin{proof}
390: From the preceding two lemmas, we know that there is a natural bijection
391: between the two spaces. Fubini's Theorem implies that a sequence
392: converges in one of the spaces if and only if the corresponding sequence
393: converges in the other space.
394: \end{proof}
395:
396: \begin{defn} Suppose $(S,\mu)$ is a standard Borel probability space.
397: \begin{itemize}
398: \item Let $\A$ be the group of all equivalence classes of measure-class-preserving Borel
399: automorphisms of $(S,\mu)$, where two automorphisms are equivalent if they are
400: equal almost everywhere.
401: \item Let $\Unitary$ be the group of unitary operators on the Hilbert
402: space $L^2(S)$, with the strong operator topology (that is, $T_n \to T$
403: if $\|T_n(f) - T(f)\| \to 0$, for every $f \in L^2(S)$, equivalently, the topology on
404: $\Unitary$ has a subbasis of open sets $\mathcal{U}(f,g,\epsilon)=\{T : \|Tf-g\|<\epsilon\}$). Note that $\Unitary$ is a Polish space.
405: \item There is a well known embedding of $\A$ in $\Unitary$, given by
406: $$ T_\phi(f)(s) = D_\phi(s)^{1/2} \, f\bigl( \phi^{-1}(s) \bigr) ,$$
407: for $\phi \in \A$ and $f \in L^2(S)$, where $D_\phi$~is the Radon-Nikodym
408: derivative of~$\phi$. This provides $\A$ with the topology
409: of a separable metric space, and thereby makes $\A$ into a topological
410: group.
411: \end{itemize}
412: \end{defn}
413:
414: \begin{rem}
415: Note that $\A$ is \emph{not} locally compact.
416: On the other hand, $\A$ is a closed subset of $\Unitary$ (because it
417: consists of the operators that map nonnegative functions to nonnegative
418: functions \cite[\S3]{GoodrichEtAl}), so it is a Polish space.
419: \end{rem}
420:
421: \begin{prop}[{Ramsay, cf.\ \cite[Cor.~3.4]{Ramsay}}]
422: If $(S,\mu)$ is a standard Borel probability space, then $\A$ acts
423: continuously on the Borel algebra $\B(S)$.
424: \end{prop}
425:
426: \begin{proof}
427: We wish to show that if $\phi$ is close to $\phi_0$ in $\A$, and $A$ is close
428: to~$A_0$ in $\B(S)$, then $\mu \bigl( \phi(A) \smallsetminus \phi_0(A_0) \bigr)$
429: and $\mu \bigl( \phi_0(A_0) \smallsetminus \phi(A) \bigr)$ are close to~$0$.
430: Thus, letting $\psi$ be $\chi_A$ and $\chi_{A_0}$, it suffices to show that if
431: $\psi \in L^2(S)$, with $\|\psi\| \le 1$, then
432: $\bigl| \int_{\phi(A)} \psi^2 \,d\mu - \int_{\phi_0(A_0)} \psi^2 \,d\mu \bigr| $
433: is close to~$0$. To simplify the notation, we replace $\phi$ and~$\phi_0$
434: by their inverses in the following calculation.
435:
436: We have
437: \begin{align*}
438: &\left| \int_{\phi^{-1}(A)} \psi^2 \,d\mu - \int_{\phi_0^{-1}(A_0)}
439: \psi^2 \,d\mu \right| \\
440: &\kern2em = \left| \int_A T_\phi(\psi)^2 \,d\mu - \int_{A_0}
441: T_{\phi_0}(\psi)^2 \,d\mu \right| \\
442: &\kern2em \le \left| \int_A T_\phi(\psi)^2 \,d\mu - \int_{A}
443: T_{\phi_0}(\psi)^2 \,d\mu \right| \\
444: &\kern4em + \left| \int_A T_{\phi_0}(\psi)^2 \,d\mu -
445: \int_{A_0} T_{\phi_0}(\psi)^2 \,d\mu \right| \\
446: &\kern2em = \left| \int_A \bigl( T_\phi(\psi) + T_{\phi_0}(\psi)
447: \bigr)
448: \bigl( T_\phi(\psi) - T_{\phi_0}(\psi) \bigr) \,d\mu
449: \right| \\
450: &\kern4em + \left| \int_A T_{\phi_0}(\psi)^2 \,d\mu -
451: \int_{A_0} T_{\phi_0}(\psi)^2 \,d\mu \right| \\
452: &\kern2em \le 2
453: \sqrt{\int_A \bigl( T_\phi(\psi) - T_{\phi_0}(\psi) \bigr)^2 \,d\mu}
454: && \begin{matrix}
455: \mbox{(H\"older's Inequality} \\
456: \mbox{and $\|\psi\| \le
457: 1$)}
458: \end{matrix}
459: \\
460: &\kern4em + \int_{A \symmdiff A_0} T_{\phi_0}(\psi)^2
461: \,d\mu
462: && \begin{matrix}
463: \mbox{(integrand is 0} \\
464: \mbox{on $A \cap A_0$)}
465: . \end{matrix}
466: \end{align*}
467: By definition of the topology on $\Unitary$, the first term in the final
468: expression is small whenever $\phi$ is close to~$\phi_0$ (since $\psi$ is
469: fixed). Because $T_{\phi_0}(\psi)^2$ is a fixed $L^1$~function, the second
470: term is small whenever $A \symmdiff A_0$ has sufficiently small measure
471: \cite[Exer.~6.10(a)]{Rudin-RCAnal}; that is, whenever $A$ is sufficiently
472: close to~$A_0$ in $\B(S)$.
473: Thus,
474: $\left| \int_{\phi^{-1}(A)} \psi^2 \,d\mu - \int_{\phi_0^{-1}(A_0)}
475: \psi^2 \,d\mu \right| $
476: is close to~$0$, as desired.
477: \end{proof}
478:
479: The fact that the action on $\B(S)$ is continuous
480: at the empty set~$\emptyset$ can be restated as follows:
481:
482: \begin{cor} \label{TendToZero}
483: If $\phi_n \to \phi$ in $\A$, and $A_n$ is a sequence of Borel subsets
484: of~$S$, such that $\mu(A_n) \to 0$, then $\mu \bigl( \phi_n(A_n) \bigr)
485: \to 0$.
486: \end{cor}
487:
488: In the following result, we assume that $S$ is a separable metric space,
489: so that we can speak of convergence in measure. This is a very mild
490: assumption, because any standard Borel space is, by definition, Borel
491: isomorphic to such a space.
492:
493: \begin{cor} \label{ConvInMeas}
494: Assume that $S$ is a separable metric space.
495: If $\phi_n \to \phi$ in $\A$, then $\phi_n \to \phi$ in measure.
496: \end{cor}
497:
498: \begin{proof}
499: By considering $\phi_n \circ \phi^{-1}$ (and using the fact that the
500: action of~$\phi$ on $\B(S)$ is continuous at~$\emptyset$), we may assume
501: that $\phi = \Id$. Let $\{A^i\}_{i=1}^\infty$ be a partition of~$S$ into
502: countably many Borel sets, such that $\diam(A^i) < \epsilon$ for each~$i$.
503: Let
504: $$ \Sigma_n^i = \{\, s \in A^i \mid d \bigl( \phi_n(s), s \bigr) >
505: \epsilon \,\} .$$ Then, for each fixed~$i$, we have
506: $$ \mu \bigl( \phi_n (\Sigma^n_i) \bigr)
507: \le \mu \bigl( \phi_n(A^i) \smallsetminus A^i \bigr)
508: \le d \bigl( \phi_n(A^i), A^i \bigr)
509: \to 0
510: \mbox{ as $n \to \infty$}.
511: $$
512: Thus, $\phi_n (\Sigma_n^i) \to \emptyset$ in $\B(S)$. Because $\A$ acts
513: continuously on $\B(S)$, this implies that
514: $$ \mu(\Sigma_n^i) \to 0 \mbox{ as $n \to \infty$}.$$
515: Therefore
516: \begin{align*}
517: &\lim_{n \to \infty} \mu \{\, s \in S \mid d \bigl( \phi_n(s), s
518: \bigr) > \epsilon \,\} \\
519: &\kern3em = \lim_{n \to \infty} \mu \bigl( \cup_{i=1}^\infty \Sigma_n^i
520: \bigr) \\
521: &\kern3em \le \inf_{k \in \natural} \lim_{n \to \infty} \mu \Bigl(
522: \bigl( \cup_{i=1}^k \Sigma_n^i \bigr) \cup \bigl( S
523: \smallsetminus (A^1 \cup \cdots \cup A^k) \bigr) \Bigr) \\
524: &\kern3em = \lim_{k \to \infty} \mu \bigl( S
525: \smallsetminus (A^1 \cup \cdots \cup A^k) \bigr) \\
526: &\kern3em= 0
527: . \end{align*}
528: So $\phi_n \to \Id$ in measure.
529: \end{proof}
530:
531: \begin{defn}
532: There is a natural action of $\A$ on $\F(S,X)$, defined by $\phi(f) = f
533: \circ \phi^{-1}$, for $\phi \in \A$ and $f \in \F(S,X)$.
534: \end{defn}
535:
536: \begin{prop} \label{Autmu(S)action}
537: If $(S,\mu)$ is a standard Borel probability space, and $(X,d)$ is a
538: separable metric space, then the natural action of $\A$ on $\F(S,X)$ is
539: continuous.
540: \end{prop}
541:
542: \begin{proof}
543: Suppose $\phi_n \to \phi$ in $\A$, and $f_n \to f$ in $\F(S,X)$. Note
544: that
545: $$ d ( f_n \phi_n, f \phi )
546: \le d ( f_n\phi_n, f\phi_n ) + d ( f \phi_n, f \phi ) .$$
547: We have $d ( f_n\phi_n, f\phi_n ) \to 0$, because
548: \begin{align*}
549: \mu \bigset{ s }{ d \bigl( f_n\phi_n(s), f \phi_n(s) \bigr)
550: \bigr)> \epsilon }
551: &= \mu \Bigl( \phi_n^{-1} \bigl\{\, s \mid d \bigl( f_n(s), f(s)
552: \bigr) > \epsilon \,\bigr\} \Bigr) \\
553: & \to 0,
554: \end{align*}
555: using \pref{TendToZero} and the fact that $f_n \to f$ in measure to get
556: the final limit.
557:
558: Because $(S,\mu)$ is standard, there is no harm in assuming that $S$ is a
559: complete, separable metric space. Then, by Lusin's Theorem, there is a
560: (large) subset~$K$ of~$S$, such that $f$ is uniformly continuous on~$K$.
561: Let
562: $$ A_n = \phi^{-1}(S \smallsetminus K) \cup \phi_n^{-1}(S \smallsetminus
563: K) .$$
564: From \pref{TendToZero}, we see that, by requiring $\mu(K)$ to be
565: sufficiently large, we may ensure that
566: $$ \mu (A_n) < \epsilon .$$
567: Choose $\delta > 0$, such that $d \bigl( f(s), f(t) \bigr) < \epsilon$,
568: for all $s,t \in K$ with $d_S(s,t) < \delta$.
569: We have
570: \begin{align*}
571: \limsup d ( f \phi_n, f \phi )
572: &= \limsup \mu \bigl\{\, s \mid d \bigl( f\phi_n(s), f \phi(s) \bigr)
573: > \epsilon\, \bigr\} \\
574: &\le \limsup \mu(A_n) + \lim \mu \bigl\{\, s \mid d \bigl( \phi_n(s),
575: \phi(s) \bigr) > \delta \,\bigr\}
576: \\
577: &\le \epsilon + 0
578: , \end{align*}
579: using \pref{ConvInMeas} to obtain the term``$0$" in the final expression.
580:
581: Since $\epsilon > 0$ is arbitrary, we conclude that $d(f_n\phi_n, f\phi)
582: \to 0$.
583: \end{proof}
584:
585:
586: We will use the following easy observation.
587:
588: \begin{lem} \label{IntIsCont}
589: If $A$ is any Borel subset of any standard Borel probability space
590: $(S,\mu)$, then the integration functional
591: $ I \colon \F(S,\real^{\ge0}) \to \real \cup \{\infty\}$,
592: defined by $I(f) = \int_A f\, d\mu$, is Borel.
593: \end{lem}
594:
595: \begin{proof}
596: Let $\chi_A$ be the characteristic function of~$A$. It is easy to see that
597: the map $f \mapsto \chi_A f$ is a continuous function on
598: $F(S,\real^{\ge0})$, so we may (and will) assume $A = S$.
599:
600: For each $n \in \natural$, choose a continuous function $\xi_n \colon
601: \real^{\ge0} \to [0,n]$, such that
602: $$ \xi_n(x) =
603: \begin{cases}
604: x & \mbox{if $x \le n$}, \\
605: 0 & \mbox{if $x \ge n+1$}.
606: \end{cases}
607: $$
608: Note, for each $f \in \F(S,\real^{\ge0})$, that the composition $\xi_n
609: \circ f$ is bounded by~$n$, and $\xi_n \circ f \uparrow f$ pointwise.
610:
611: Because $\xi_n$ is uniformly continuous (being a continuous function with
612: compact support), it is easy to see that the map $\F(S,\real^{\ge0}) \to
613: \F(S,[0,n])$, defined by $f \mapsto \xi_n \circ f$, is continuous.
614: Furthermore, the restriction of~$I$ to $\F(S,[0,n])$ is continuous. Thus,
615: the function
616: $$ \mbox{$ I_n \colon \F(S,\real^{\ge0}) \to \real^{\ge0}$,
617: defined by $I_n(f) = I(\xi_n \circ f)$,}$$
618: is continuous.
619: The Monotone Convergence Theorem implies that
620: $$I(f) = \lim_{n \to \infty} I_n(f) ,$$ so $I$~is a pointwise limit of
621: continuous functions. Therefore, $I$ is Borel.
622: \end{proof}
623:
624:
625: \subsection{Ergodic decomposition and near actions}
626: A proof of the following folklore theorem has been provided by
627: G.~Greschonig and K.~Scmidt \cite{Greschonig-Schmidt}. The statement here
628: is slightly stronger than \cite[Thm.~1.1]{Greschonig-Schmidt} (because
629: Thm.~\fullref{ErgDecompThm}{atom} is more precise than
630: \cite[Thm.~1.1(3)]{Greschonig-Schmidt}), but it follows immediately from
631: \cite[Thm.~5.2]{Greschonig-Schmidt}, by letting
632: $\erg$ be a $p_*\mu$-conull Borel subset of $\Prob(X)$,
633: $X' = p^{-1}(\erg)$,
634: $\psi(x) = p(x)$,
635: and
636: $\xi(\omega) = \omega$.
637:
638: \begin{notn}
639: For any standard Borel space~$X$, we use $\Prob(X)$ to denote the space
640: of all probability measures on~$X$. It is well known that $\Prob(X)$ is a
641: standard Borel space, under an appropriate weak$^*$ topology.
642: \end{notn}
643:
644: \begin{thm}[{Ergodic Decomposition \cite{Greschonig-Schmidt}}]
645: \label{ErgDecompThm}
646: Let
647: \begin{itemize}
648: \item $G$ be a locally compact second countable group,
649: \item $(X,\mu)$ be a standard Borel probability space,
650: and
651: \item $\rho \colon G \times X \to X$ be a Borel action, such that
652: $\mu$~is quasi-invariant.
653: \end{itemize}
654: Then there exist:
655: \begin{itemize}
656: \item a standard Borel probability space $(\erg,\nu)$,
657: \item a conull $G$-invariant Borel subset $X'$ of~$X$,
658: \item a $G$-invariant Borel map $\psi \colon X' \to \erg$,
659: and
660: \item a Borel map $\xi \colon \erg \to \Prob(X')$,
661: \end{itemize}
662: such that
663: \begin{enumerate}
664: \item \label{ErgDecompThm-atom}
665: $\xi(\omega) \bigl( \psi^{-1}(\omega) \bigr) = 1$ for each $\omega \in
666: \erg$,
667: \item $\mu = \int_{\erg} \xi(\omega) \, d\nu(\omega)$,
668: and
669: \item For each $\omega \in \erg$, $\xi(\omega)$ is quasi-invariant and ergodic.
670: \end{enumerate}
671: \end{thm}
672:
673: \begin{rem} \label{UnionOfProds}
674: To simplify the notation in the conclusion of Thm.~\ref{ErgDecompThm}, we
675: will often assume that the space~$X$ can be written as a Cartesian product
676: $X = \erg \times S$, such that
677: \begin{itemize}
678: \item $\mu$ is the product measure on $X = \erg \times S$,
679: and
680: \item $\psi(\omega, s) = \omega$, for a.e.\ $(\omega,s) \in X$.
681: \end{itemize}
682: (For example, we make this assumption in the statement of
683: Prop.~\ref{BorelMaps}.)
684: The following decomposition theorem of V.~A.~Rohlin \cite{Rohlin}
685: asserts that (up to isomorphism) the general case is a countable union of
686: examples of this type.
687: \end{rem}
688:
689: \begin{prop}[Rohlin] \label{RohlinDecomp}
690: Assume the notation of Theorem~\ref{ErgDecompThm}. There is a partition
691: of~$\Omega$ into countably many Borel sets $\Omega_1,\Omega_2,\ldots$
692: {\rm(}some of these sets may be empty{\rm)}, such that, for each~$k$,
693: there exist
694: \begin{enumerate}
695: \item a conull subset $X'_k$ of $\psi^{-1}(\Omega_k)$,
696: \item a standard Borel probability space $(S_k,\mu_k)$,
697: \item a $(\psi_*\mu \times \mu_k)$-conull subset $(\erg_k \times
698: S_k)'$ of $\erg_k \times S_k$,
699: and
700: \item a measure-class-preserving Borel isomorphism $\theta \colon X'_k
701: \to (\erg_k \times S_k)'$,
702: \end{enumerate}
703: such that $\psi(x) = \pi_1 \bigl( \theta(x) \bigr)$
704: for each $x \in X'_k$, where $\pi_1(\omega,s) = \omega$.
705: \end{prop}
706:
707: \begin{proof}
708: Say that two standard Borel probability spaces $(S_1,\mu_1)$ and
709: $(S_2,\mu_2)$ are \emph{of the same type} if there exists a
710: measure-class-preserving Borel isomorphism from a conull subset of~$S_1$
711: onto a conull subset of~$S_2$ \cite[pp.~10--11]{Rohlin}. This is obviously
712: an equivalence relation. It has only countably many equivalence classes
713: \cite[p.~18]{Rohlin}. Thus, there is a decomposition $\erg = \erg_1 \cup
714: \erg_2 \cup \cdots$, such that if $\omega$ and~$\omega'$ belong to the
715: same~$\Omega_k$, then the fibers $\bigl( \psi^{-1}(\omega), \xi(\omega)
716: \bigr)$ and $\bigl( \psi^{-1}(\omega'), \xi(\omega') \bigr)$ are of the
717: same type \cite[(I), p.~41]{Rohlin}. This implies that, modulo sets of
718: measure~$0$, $\psi^{-1}(\erg_k)$ is Borel isomorphic to the Cartesian
719: product $\erg_k \times \psi^{-1}(\omega)$, for any $\omega \in \erg_k$
720: \cite[p.~42]{Rohlin}.
721: \end{proof}
722:
723: \begin{prop} \label{BorelMaps}
724: Let
725: \begin{itemize}
726: \item $G$ be a second countable, locally compact group,
727: \item $(\erg,\nu)$ and $(S,\mu)$ be standard Borel probability
728: spaces,
729: \item $\erg'$ be a conull Borel subset of~$\erg$,
730: and
731: \item $\rho \colon G \times (\erg \times S) \to \erg \times S$ be a
732: Borel action of~$G$ on $\erg \times S$, such that, for each $\omega \in
733: \erg'$, the probability measure $\mu_\omega$ on $\{\omega\} \times S$
734: {\rm(}induced by the natural isomorphism with~$S${\rm)} is
735: quasi-invariant.
736: \end{itemize}
737: Then:
738: \begin{enumerate}
739: \item \label{BorelMaps-intoA}
740: There are Borel maps $\rho_S \colon G \times
741: \erg' \times S \to S$ and $\rho_{\Aut} \colon G \times \erg' \to \A$,
742: defined by
743: \begin{equation} \label{BorelMaps-RhoDefn}
744: \mbox{$\rho(g,\omega,s) = \bigl( \omega, \rho_S(g,\omega,s) \bigr)
745: = \bigl( \omega, \rho_{\Aut}(g,\omega)(s) \bigr)$ for a.e.\ $s \in S$}
746: \end{equation}
747: for each $(g,\omega) \in G \times \erg'$.
748: \item \label{BorelMaps-RN}
749: There is a Borel function $D \colon G \times \erg \times S \to \real^{\ge0}$, such that
750: $$\int_S D(g, \omega,s) f(s) \, d\mu(s) = \int_S f \bigl(
751: \rho_S(g,\omega,s) \bigr)\, d\mu(s) $$
752: for every $g \in G$, $\omega \in \erg$, and $f \in \F(S,
753: \real^{\ge0})$.
754: \end{enumerate}
755: \end{prop}
756:
757: \begin{rem}
758: The function $D$ in conclusion $2$ above is a \emph{fiberwise}
759: Radon-Nikodym derivative. In particular, in the case when
760: $\Omega$ is a one point set, conclusion $2$ implies that the Radon-Nikodym
761: derivative is a Borel function on $G{\times}S$. \end{rem}
762:
763: \begin{proof}
764: Equation~\pref{BorelMaps-RhoDefn} determines a well-defined function
765: $\rho_{\Aut} \colon G \times \erg' \to \F(S,S)$. Because $G$ is a group,
766: we know that $\rho_{\Aut}(g,\omega)$ is (essentially) a Borel
767: automorphism. Because $\mu_\omega$ is quasi-invariant, we see that
768: $\rho_{\Aut}(g,\omega)$ is measure-class preserving. Thus,
769: $\rho_{\Aut}(g,\omega) \in \A$, so $\rho_{\Aut}$ is actually a map
770: into $\A$. Thus, in order to complete the proof of
771: \pref{BorelMaps-intoA}, it only remains to show that $\rho_{\Aut}$ is
772: Borel. For this, we will use the conclusion of \pref{BorelMaps-RN}, so
773: let us establish the latter.
774:
775: \pref{BorelMaps-RN}
776: Let $\{A_n\}$ be a countable, dense subset of the Borel algebra $\B(S)$,
777: and define
778: \begin{align*}
779: D^\Delta &= \bigcap_{n \in \natural} \bigset{
780: (g, \omega,f)
781: }{
782: \mu \bigl( \rho_{\Aut}(g, \omega)(A_n) \bigr)
783: = \int_{A_n} f \, d\mu} \\
784: &\subset G \times \erg' \times \F(S,\real^{\ge0})
785: . \end{align*}
786: For each $g \in G$ and $\omega \in \erg'$, the fiber
787: $\{\, s \in S \mid (g,\omega,s) \in D^\Delta \,\}$
788: of~$D^\Delta$ consists precisely of the Radon-Nikodym derivative of the
789: transformation $\rho_{\Aut}(g, \omega)$. Therefore, $D^\Delta$ is the graph of
790: a function $\check D \colon G \times \erg' \to \F(S,\real^{\ge0})$, and
791: $\check D(g, \omega)$ is the Radon-Nikodym derivative of $\rho_{\Aut}(g,
792: \omega)$.
793:
794: Note:
795: \begin{itemize}
796: \item Because the Borel map $(g,\omega,s) \mapsto \bigl( g,\omega,
797: \rho_S(g,\omega,s) \bigr)$ is injective, we know that it maps $G \times
798: \erg' \times A_n$ (or any other Borel set) to a Borel subset of $G \times
799: \erg' \times S$. So Fubini's Theorem implies that $\mu \Bigl(
800: \rho_S \bigl( (g, \omega) \times A_n \bigr) \Bigr)$ is a Borel function of
801: $(g,\omega)$.
802: \item Lemma~\ref{IntIsCont} implies that $\int_{A_n} f \, d\mu$ is a
803: Borel function of~$f$.
804: \end{itemize}
805: Therefore $D^\Delta$ is a Borel set, so the corresponding function
806: $\check D$ is a Borel function. Then the desired Borel function $D \colon
807: (G \times \erg') \times S \to \real^{\ge0}$ is obtained by applying
808: Lem.~\ref{DefdFiberwise}.
809:
810: \pref{BorelMaps-intoA}
811: Given $L^2$ functions $f,g \colon S \to \real$, and any $\epsilon > 0$,
812: define
813: $$ \Theta \colon G \times \erg' \times S \to \real^{\ge0}$$
814: by
815: \begin{align*}
816: \Theta(g,\omega,s)
817: &= \bigl| T_{\rho_{\Aut}(g,\omega)} f(s) - g(s) \bigr|^2 \\
818: &= \bigl| D(g,\omega,s)^{1/2} \, f\bigl( \rho_{\Aut}(g,\omega)^{-1}(s)
819: \bigr) - g(s) \bigr|^2 ,
820: \end{align*}
821: where $D(g,\omega,s)$ is given by \fullref{BorelMaps}{RN}.
822: Then $\Theta$ is a Borel function on $G \times \erg' \times S$, so
823: Fubini's Theorem implies that
824: $$ \bigset{ (g, \omega) \in G \times \erg' }{ \int_S \Theta(g,\omega,s) <
825: \epsilon^2} $$
826: is Borel. In other words, if $\mathcal{U}(f,g,\epsilon)$ is any basic
827: open set in $\Unitary$, then $\rho_{\Aut}^{-1} \bigl(
828: \mathcal{U}(f,g,\epsilon) \bigr)$ is Borel. So $\rho_{\Aut}$ is Borel.
829: \end{proof}
830:
831: \begin{cor} \label{BorelIntoF}
832: Let
833: \begin{itemize}
834: \item $G$, $\erg$, $S$, $\erg'$, and $\rho$ be as in
835: Prop.~\ref{BorelMaps},
836: and
837: \item $X$ be a complete, separable metric space.
838: \end{itemize}
839: Then the action~$\rho$ induces a Borel map $\rho_{\F} \colon G \times
840: \erg' \times \F(S,X) \to \F(S,X)$, defined by
841: \begin{equation} \label{BorelIntoF-eq}
842: \mbox{$\rho_{\F}(g,\omega,f)(s) = f \bigl( \rho_S(g,\omega,s) \bigr)$
843: \quad for a.e.~$s$} .
844: \end{equation}
845: \end{cor}
846:
847: \begin{proof}
848: The existence of an abstract function~$\rho_{\F}$ satisfying
849: \pref{BorelIntoF-eq} is not an issue. Because $\rho_{\Aut}$ is Borel
850: \fullsee{BorelMaps}{intoA}, and the natural action of $\A$ on $\F(S,X)$ is
851: continuous \see{Autmu(S)action}, we know that $\rho_{\F}$ is Borel.
852: \end{proof}
853:
854:
855: \subsection{Borel cocycles}
856:
857: We assume the basic theory of Borel cocycles, as in
858: \cite[\S7.2]{MargulisBook} or \cite[\S4.2]{ZimmerBook}.
859:
860: \begin{defn}[{\cite[Defns.~4.2.1 and 4.2.2]{ZimmerBook}}] \label{CocDefn}
861: Suppose $H$~is a topological group, and $\rho \colon G \times X \to X$
862: is a Borel action of a locally compact group~$G$ on a standard Borel
863: probability space~$X$ with quasi-invariant measure~$\mu$.
864: \begin{enumerate}
865: \item A Borel function $\alpha \colon G \times X \to H$ is a Borel
866: \emph{cocycle} (for the action~$\rho$) if, for all $g_1,g_2 \in G$, we
867: have
868: \begin{equation} \label{CocDefn-CocEqn}
869: \mbox{$\alpha(g_1 g_2, x) = \alpha \bigl( g_1, \rho(g_2 , x) \bigr) \,
870: \alpha(g_2, x)$
871: \quad for a.e.\ $x \in X$}.
872: \end{equation}
873: \item The cocycle $\alpha$ is \emph{strict} if the equality in
874: \pref{CocDefn-CocEqn} holds for \emph{all} $x \in X$, not merely almost
875: all.
876: \item Two Borel cocycles $\alpha,\beta \colon G \times X \to H$ are
877: \emph{cohomologous} if there is a Borel function $\phi \colon X \to H$,
878: such that, for each $g \in G$, we have
879: $$ \mbox{$\beta(g,x) = \phi \bigl( \rho(g,x) \bigr) \, \alpha(g,x) \,
880: \phi(x)^{-1}$
881: \quad for a.e.\ $x \in X$} .$$
882: This is an equivalence relation.
883: \end{enumerate}
884: \end{defn}
885:
886: \begin{rem} \label{CanBeStrict}
887: There is usually no harm in assuming that a Borel cocycle is strict,
888: because any Borel cocycle is
889: equal a.e.\ to a strict Borel cocycle \cite[Thm.~B.9, p.~200]{ZimmerBook}.
890: More precisely, if $\alpha$ is a Borel cocycle, then there is a strict
891: Borel cocycle $\alpha'$, such that, for each $g \in G$, $\alpha(g,x) =
892: \alpha'(g,x)$ for a.e.\ $x \in X$.
893: \end{rem}
894:
895: \begin{lem} \label{PointwiseOps}
896: Let
897: \begin{itemize}
898: \item $(S,\mu)$ be a standard Borel probability space.
899: \item $X$, $Y$, and~$Z$ be complete, separable, locally compact
900: metric spaces,
901: \item $\tau \colon X \times Y \to Z$ be a continuous function.
902: \end{itemize}
903: Then the induced map $\tau^{\F} \colon \F(S,X) \times \F(S,Y) \to
904: \F(S,Z)$, defined by
905: $$ \tau^{\F}(\phi,\psi)(s) = \tau \bigl( \phi(s), \psi(s) \bigr), $$
906: is continuous.
907: \end{lem}
908:
909: \begin{proof}
910: Given sequences $\phi_n \to \phi \in \F(S,X)$, $\psi_n \to \psi \in
911: \F(S,Y)$, and $\epsilon > 0$, Lusin's Theorem gives us a compact
912: subset~$K$ of~$S$, such that $\phi$ and~$\psi$ are continuous on~$K$, and
913: $\mu(K) > 1 - \epsilon$. Because $X$ and~$Y$ are locally compact, we may
914: let $K_X$ and $K_Y$ be compact neighborhoods of $\phi(K)$ and $\psi(K)$
915: in $X$ and~$Y$, respectively.
916: Because $\tau$ is uniformly continuous on $K_X \times K_Y$, there is
917: some $\delta > 0$, such that
918: \begin{align*}
919: &\limsup_{n \to \infty} \mu \bigset{ s \in S }{
920: d \Bigl( \tau \bigl( \phi_n(s), \psi_n(s) \bigr), \tau \bigl( \phi(s),
921: \psi(s) \bigr) \Bigr) > \epsilon } \\
922: &\kern3em\le
923: \limsup_{n \to \infty} \mu \{\, s \in S \mid \mbox{$\phi_n(s) \notin
924: K_X$ or $\psi_n(s) \notin K_Y$} \,\} \\
925: &\kern6em + \limsup_{n \to \infty} \mu \{\, s \in S \mid d
926: \bigl( \phi_n(s), \phi(s) \bigr) > \delta \,\} \\
927: &\kern6em + \limsup_{n \to \infty} \mu \{\, s \in S \mid d
928: \bigl( \psi_n(s), \psi(s) \bigr) > \delta \,\} \\
929: &\kern3em \le 2 \epsilon + 0 + 0
930: . \end{align*}
931: Because $\epsilon > 0$ is arbitrary, we conclude that
932: $\tau^{\F}(\phi_n,\psi_n) \to \tau^{\F}(\phi,\psi)$.
933: \end{proof}
934:
935: \begin{cor}[{cf.\ \cite[Rmk.\ after Lem.~7.2.1, p.~217]{MargulisBook}}]
936: \label{TwistisBorel}
937: Let
938: \begin{itemize}
939: \item $G$, $\erg$, $S$, $\erg'$, and $\rho$ be as in
940: Prop.~\ref{BorelMaps},
941: \item $H$ be a locally compact, second countable group,
942: \item $X$ be a complete, separable metric space,
943: \item $\tau \colon H \times X \to X$ be a continuous action of~$H$
944: on~$X$,
945: and
946: \item $\alpha \colon G \times (\erg \times S) \to H$ be a Borel cocycle.
947: \end{itemize}
948: Then the function $\rho^{\alpha,\tau} \colon G \times \erg' \times
949: \F(S,X) \to \F(S,X)$, defined by
950: $$ \rho^{\alpha,\tau}(g,\omega,\phi)(s) =
951: \tau \Bigl( \alpha(g,\omega, s)^{-1}, \phi \bigl( \rho_S(g,\omega,s)
952: \bigr) \Bigr) ,$$
953: is Borel.
954: \end{cor}
955:
956: \begin{proof}
957: Because the map $\rho_{\F}$ of Cor.~\ref{BorelIntoF} is Borel, the
958: action~$\rho$ does not affect the measurability of~$\rho^{\alpha,\tau}$,
959: so it may be ignored. Furthermore, by replacing $\erg$ with $G \times
960: \erg'$, we may assume that $G$~is trivial and $\erg' = \erg$; in
961: particular, $G$ may be ignored. Thus,
962: \begin{itemize}
963: \item $\alpha$ is a Borel map from $\erg \times S$ to $H$,
964: and
965: \item $\rho^{\alpha,\tau} \colon \erg \times \F(S,X) \to \F(S,X)$ is
966: defined by
967: $$ \rho^{\alpha,\tau}(\omega,\phi)(s) =
968: \tau \bigl( \alpha(\omega,s)^{-1}, \phi(s) \bigr) .$$
969: \end{itemize}
970: Then, because $\check\alpha \colon \erg \to \F(S,X)$ and $\tau^{\F}$
971: are Borel (see \fullref{FiberFunc}{Borel} and~\ref{PointwiseOps}), we
972: conclude that $\rho^{\alpha,\tau}$ is a composition of Borel functions.
973: Therefore, it is Borel.
974: \end{proof}
975:
976: Although we do not need the following result in this paper, we include
977: the proof to provide a convenient reference.
978:
979: \begin{cor}[{\cite[Rmk.\ after Lem.~7.2.1, p.~217]{MargulisBook}}]
980: \label{TwistContinuous}
981: Let
982: \begin{itemize}
983: \item $G$ and $H$ be second countable, locally compact groups,
984: \item $(S,\mu)$ be a standard Borel probability space,
985: \item $\rho \colon G \times S \to S$ be a Borel action of~$G$ on~$S$,
986: such that $\mu$ is quasi-invariant,
987: \item $\alpha \colon G \times S \to H$ be a strict Borel cocycle,
988: and
989: \item $\tau \colon H \times X \to X$ be a continuous action of~$H$
990: on~$X$.
991: \end{itemize}
992: Then
993: \begin{enumerate}
994: \item the function $\check\alpha \colon G \to \F(S,H)$, induced
995: by~$\alpha$ \see{FiberFunc}, is continuous,
996: and
997: \item the $\alpha$-twisted action $\zeta_{\rho,\tau,\alpha}$ of~$G$ on
998: $\F(S,H)$ is continuous {\rm(}see Defn.~\fullref{TwistDefn}{action}{\rm)}.
999: \end{enumerate}
1000: \end{cor}
1001:
1002: \begin{proof}
1003: Let $\triv \colon G \times S \to H$ be the trivial cocycle, defined
1004: by $\triv(g,s) = e$ (the identity element of~$H$). For convenience, let
1005: $\F = \F(S,H)$.
1006:
1007: \setcounter{steppf}{0}
1008:
1009: \begin{steppf} \label{TwistContinuous-translate}
1010: The action $\rho^{\triv,\alpha}$ is continuous.
1011: \end{steppf}
1012: From Prop.~\fullref{BorelMaps}{intoA} (with $\erg$ consisting of a single
1013: point), we know that $\rho$ induces a Borel function $\rho_{\Aut} \colon
1014: G \to \A$. This function is a homomorphism, and any measurable
1015: homomorphism into a second countable topological group is continuous
1016: \cite[Thm.~B.3, p.~198]{ZimmerBook}, so we conclude that $\rho_{\Aut}$ is
1017: continuous. Because the action of $\A$ on $\F$ is continuous
1018: \see{Autmu(S)action}, we conclude that $\rho^{\triv,\alpha}$ is
1019: continuous.
1020:
1021: \begin{steppf} \label{TwistContinuous-cocycle}
1022: $\check\alpha$ is continuous.
1023: \end{steppf}
1024: Note that $\F$ is a (second countable) topological group under
1025: pointwise multiplication \cf{PointwiseOps}. Furthermore,
1026: $\rho^{\triv,\alpha}$ is a continuous action of~$G$ on~$\F$ by
1027: automorphisms, so we may form the semidirect product $G \ltimes \F$. The
1028: cocycle identity implies that the Borel function $\alpha^{G \ltimes \F}
1029: \colon G \to G \ltimes \F$, defined by $\alpha^{G \ltimes \F}(g,f)
1030: = \bigl( g, \check\alpha(g) \bigr)$ is a homomorphism. Because (as noted
1031: above) measurable homomorphisms are continuous, we conclude that
1032: $\check\alpha$ is continuous.
1033:
1034: \begin{steppf}
1035: $\rho^{\tau,\alpha}$ is continuous.
1036: \end{steppf}
1037: Because $\rho^{\tau,\alpha}(g,f) = \tau^{\F} \bigl( \check\alpha(g),
1038: \rho^{\triv,\alpha}(g,f) \bigr)$, this conclusion can be obtained by
1039: combining Steps~\ref{TwistContinuous-translate}
1040: and~\ref{TwistContinuous-cocycle} with Lem.~\ref{PointwiseOps}.
1041: \end{proof}
1042:
1043:
1044: \section{Restricting cocycles to ergodic components}
1045: \label{RestrictCocycleSect}
1046:
1047: This section uses the von Neumann Selection Theorem \pref{vonNeumann} to
1048: obtain information about a cocycle from its restrictions to ergodic
1049: components. The main result is \pref{AnalErgComp}; the others are
1050: corollaries.
1051:
1052: \begin{notn} \label{ResCocNotn}
1053: Let
1054: \begin{itemize}
1055: \item $G$, $\erg$, $S$, $\erg'$ and~$\rho$ be as in
1056: Prop.~\ref{BorelMaps} (or, equivalently, as in Thm.~\ref{AnalErgComp}
1057: below),
1058: \item $H$ be a locally compact, second countable group,
1059: and
1060: \item $\alpha \colon G \times (\erg \times S) \to H$ be a strict Borel
1061: cocycle.
1062: \end{itemize}
1063: For each $\omega \in \erg'$, define
1064: $\rho_\omega \colon G \times S \to S$
1065: and
1066: $\alpha_\omega \colon G \times S \to H$
1067: by
1068: $$ \mbox{$\rho_\omega(g,s) = \rho_S(g,\omega,s)$
1069: \quad and \quad
1070: $\alpha_\omega(g,s) = \alpha(g,\omega,s)$}
1071: ,$$
1072: where $\pi_S(*,s) = s$.
1073: \end{notn}
1074:
1075:
1076:
1077: \begin{defn}[{\cite[Lem.~3.1]{Ramsay}, \cite[Defn.~3.1]{ZimmerExtn}}]
1078: Suppose $G$ is a locally compact second countable group, and $(S,\mu)$ is
1079: a standard Borel probability space.
1080: A Borel map $\rho \colon G \times S \to S$ is a \emph{near action}
1081: of~$G$ on~$S$ if
1082: \begin{itemize}
1083: \item for all $g_1,g_2 \in G$, we have
1084: $\rho(g_1 g_2 , s) = \rho \bigl( g_1, \rho(g_2, s) \bigr)$
1085: for a.e.\ $s \in S$,
1086: \item $\rho(e,s) = s$ for a.e.\ $s \in S$,
1087: and
1088: \item each $g \in G$ preserves the measure class of~$\mu$.
1089: \end{itemize}
1090: Note that the definition of a cocycle \pref{CocDefn} can be applied to
1091: near actions, not only actions.
1092: \end{defn}
1093:
1094: Let us record the following elementary observation.
1095: Part~\pref{ResCocLem-Invt} follows easily from the assumption that
1096: $\mu_\omega$ is quasi-invariant. The other two parts are consequences of
1097: the first.
1098:
1099: \begin{lem} \label{ResCocLem}
1100: Assume the setting of Notation~\ref{ResCocNotn}, and let $\omega \in
1101: \erg'$. Then:
1102: \begin{enumerate}
1103: \item \label{ResCocLem-Invt}
1104: $\rho(\omega,s) \in \{\omega\} \times S$, for a.e.\ $s \in S$,
1105: \item $\rho_\omega$ is a near action of~$G$ on~$S$,
1106: and
1107: \item \label{ResCocLem-coc}
1108: $\alpha_\omega$ is a Borel cocycle for~$\rho_\omega$.
1109: \end{enumerate}
1110: \end{lem}
1111:
1112: \begin{thm} \label{AnalErgComp}
1113: Let
1114: \begin{itemize}
1115: \item $G$ and~$H$ be second countable, locally compact groups,
1116: \item $(\erg,\nu)$ and $(S,\mu)$ be standard Borel probability
1117: spaces,
1118: \item $\erg'$ be a conull subset of~$\erg$,
1119: \item $\rho \colon G \times (\erg \times S) \to \erg \times S$ be a
1120: Borel action of~$G$ on $\erg \times S$, such that, for each $\omega' \in
1121: \erg$, the probability measure $\mu_\omega$ on $\{\omega\} \times S$
1122: {\rm(}induced by the natural isomorphism with~$S${\rm)} is
1123: quasi-invariant,
1124: \item $\alpha \colon G \times (\erg \times S) \to H$ be a strict Borel
1125: cocycle,
1126: \item $\func$ be an analytic subset of $\erg \times \F(G \times
1127: S,H)$,
1128: and
1129: \item $\erg_\func = \{\, \omega \in \erg' \mid \mbox{$\alpha_\omega$
1130: is cohomologous to a cocycle in $\func_\omega$} \,\}$,
1131: where
1132: $$\func_\omega = \{\, f \in \F(G \times
1133: S,H) \mid (\omega, f) \in \func\,\} .$$
1134: \end{itemize}
1135: Then:
1136: \begin{enumerate}
1137: \item \label{AnalErgComp-anal}
1138: $\erg_\func$ is analytic,
1139: and
1140: \item \label{AnalErgComp-coho}
1141: $\alpha$ is cohomologous to a Borel cocyle
1142: $\beta \colon G \times (\erg \times S) \to H$, such that $\beta_\omega \in
1143: \func_\omega$, for a.e.\ $\omega \in \erg_{\func}$.
1144: \end{enumerate}
1145: \end{thm}
1146:
1147: \begin{proof}
1148: The cocycle~$\alpha$ defines a Borel map $\check\alpha \colon \erg \to
1149: \F(G \times S,H)$ \fullsee{FiberFunc}{Borel}. Define
1150: $$ \cobound \colon \erg' \times \F(G \times S,H) \times \F(S,H) \to \F(G
1151: \times S,H)$$
1152: by
1153: \begin{equation} \label{AnalErgComp-delta}
1154: \cobound(\omega, \phi, f)(g,s) = f \bigl( \rho_\omega(g,s) \bigr)
1155: \, \phi(g,s) \, f(s)^{-1} .
1156: \end{equation}
1157:
1158: We claim that $\cobound$ is Borel.
1159: \begin{itemize}
1160: \item We know that the map $\rho_{\F}$ (defined in Cor.~\ref{BorelIntoF})
1161: is Borel. This induces a Borel map $\check\rho_{\F} \colon \erg' \times
1162: \F(S,H) \to \F \bigl( G, \F(S,H) \bigr)$ \see{FiberFunc}. From
1163: Cor.~\ref{F(F)}, we see that we may think of this as a map into $\F(G
1164: \times S, H)$. Thus, the first factor on the right-hand side of
1165: \pref{AnalErgComp-delta} represents a Borel function from $\erg' \times
1166: \F(S,H)$ into $\F(G \times S, H)$.
1167: \item The second factor on the right-hand side of
1168: \pref{AnalErgComp-delta} represents the identity function on $\F(G \times
1169: S, H)$, and the term $f(s)$ represents the inclusion of $\F(S,H)$ into
1170: $\F(G \times S, H)$. These are obviously Borel maps into $\F(G \times S,
1171: H)$.
1172: \end{itemize}
1173: Because pointwise multiplication and pointwise inversion are continuous
1174: operations on $\F(G \times S, H)$ \see{PointwiseOps}, we conclude that
1175: $\cobound$ is Borel, as claimed.
1176:
1177: Therefore, the function
1178: $$\sigma \colon \erg' \times \F(S,H) \to \erg' \times \F(G \times S,H)
1179: ,$$
1180: defined by
1181: $$ \sigma(\omega, f) = \Bigl( \omega, \cobound \bigl( \omega,
1182: \check\alpha(\omega), f \bigr) \Bigr) ,$$
1183: is Borel, so $\sigma^{-1}(\func)$ is analytic. Then, because $\erg_\func$
1184: is the projection of~$\sigma^{-1}(\func)$ to~$\erg'$, we conclude that
1185: $\erg_\func$ is analytic. This establishes \pref{AnalErgComp-anal}.
1186:
1187: The von Neumann Selection Theorem \pref{vonNeumann} implies that there is
1188: a Borel function $\hat\Phi \colon \erg' \to \F(S,H)$, such that
1189: $\sigma\bigl( \omega, \hat\Phi(\omega) \bigr) \in \func$, for a.e.\
1190: $\omega \in \erg_\func$. Corresponding to~$\hat\Phi$, there is a Borel
1191: function $\Phi \colon \erg' \times S \to H$ \see{DefdFiberwise}. Letting
1192: $$ \beta(g,\omega,s)
1193: = \cobound \bigl( \omega, \check\alpha(\omega), \hat\Phi(\omega) \bigr)
1194: (g,s)
1195: = \Phi \bigl( \rho(g,\omega,s) \bigr) \, \alpha(g,\omega,s) \,
1196: \Phi(\omega, s)^{-1} ,$$
1197: we obtain \pref{AnalErgComp-coho}.
1198: \end{proof}
1199:
1200: \begin{cor} \label{CohoErg}
1201: Let
1202: \begin{itemize}
1203: \item $G$ and $H$ be locally compact, second countable groups,
1204: \item $(X,\mu)$ be a standard Borel probability space,
1205: \item $\rho \colon G \times X \to X$ be a Borel action, such that $\mu$
1206: is quasi-invariant,
1207: \item $\psi \colon X' \to \erg$ be the corresponding ergodic
1208: decomposition \see{ErgDecompThm},
1209: and
1210: \item $\alpha,\beta \colon G \times X \to H$ be Borel cocycles.
1211: \end{itemize}
1212: For each $\omega \in \erg$, let $\alpha_\omega$ and $\beta_\omega$ be
1213: the restrictions of~$\alpha$ and~$\beta$ to $G \times \psi^{-1}(\omega)$.
1214: Then:
1215: \begin{enumerate}
1216: \item \label{CohoErg-coc}
1217: There is a conull Borel subset $\erg'$ of~$\erg$, such that, for each
1218: $\omega \in \erg'$, the maps $\alpha_\omega$ and $\beta_\omega$ are Borel
1219: cocycles.
1220: \item \label{CohoErg-anal}
1221: $\{\, \omega \in \erg' \mid \mbox{$\alpha_\omega$ is cohomologous
1222: to~$\beta_\omega$}\,\}$ is an analytic subset of~$\erg$.
1223: \item \label{CohoErg-trivial}
1224: If $\alpha_\omega$ is cohomologous to~$\beta_\omega$, for a.e.\ $\omega
1225: \in \erg'$, then $\alpha$ is cohomologous to~$\beta$.
1226: \end{enumerate}
1227: \end{cor}
1228:
1229: \begin{proof}
1230: From Prop.~\ref{RohlinDecomp} (and Rem.~\ref{UnionOfProds}), we may
1231: assume the notation of Thm.~\ref{AnalErgComp}. By changing $\alpha$
1232: and~$\beta$ on a null set, we may assume these cocycles are strict
1233: \see{CanBeStrict}.
1234:
1235: Conclusion \pref{CohoErg-coc} is immediate from
1236: Lem.~\fullref{ResCocLem}{coc}.
1237:
1238: Recall that $\beta$ induces a Borel function $\check\beta \colon \erg'
1239: \to \F(G \times S, H)$, defined by $\check\beta(\omega) = \beta_\omega$
1240: \fullsee{FiberFunc}{Borel}.
1241: Let
1242: $$\func = \{\, (\omega, \check\beta(\omega) \mid \omega \in \erg' \,\}
1243: \subset \erg' \times \F(G \times S, H) .$$
1244: Because $\func$ is an analytic subset (in fact, it is closed),
1245: \pref{CohoErg-anal} is immediate from \fullref{AnalErgComp}{anal}.
1246:
1247: Assume, now, that $\alpha_\omega$ is cohomologous to~$\beta_\omega$, for a.e.\ $\omega
1248: \in \erg'$. From \fullref{AnalErgComp}{coho} and Fubini's
1249: Theorem, we conclude that $\alpha$ is cohomologous to a
1250: cocycle~$\tilde\alpha$, such that for a.e.\ $g \in G$,
1251: \begin{equation} \label{CohoErg-AlmostBeta}
1252: \mbox{for a.e.~$x \in \erg' \times S$, we have $\tilde\alpha(g,x) =
1253: \beta(g,x)$.}
1254: \end{equation}
1255: From the cocycle identity, one easily concludes that
1256: \pref{CohoErg-AlmostBeta} must hold for every $g \in G$, not merely
1257: almost every~$g$. Therefore $\tilde\alpha$ is (obviously) cohomologous
1258: to~$\beta$. By transitivity, then $\alpha$ is also cohomologous
1259: to~$\beta$; this establishes~\pref{CohoErg-trivial}.
1260: \end{proof}
1261:
1262: \begin{defn}
1263: Suppose $\rho \colon G \times X \to X$ is a Borel action with
1264: quasi-invariant measure, and $H$~is a locally compact second countable
1265: group.
1266: \begin{enumerate}
1267: \item The \emph{trivial cocycle} $\triv_{G \times X} \colon G \times X
1268: \to H$ is defined by $\triv_{G\times X}(g,x) = e$.
1269: \item A Borel cocycle $\alpha \colon G \times X \to H$ is a
1270: \emph{coboundary} if it is cohomologous to the trivial cocycle.
1271: \end{enumerate}
1272: \end{defn}
1273:
1274: \begin{cor} \label{TrivialCocycle}
1275: Let $G$, $H$, $X$, $\rho$, $\psi$, $\Omega$, $\alpha$, $\alpha_\omega$,
1276: and~$\erg'$ be as in Cor.~\ref{CohoErg}. Then:
1277: \begin{enumerate}
1278: \item \label{TrivialCocycle-anal}
1279: $\{\, \omega \in \erg' \mid \mbox{$\alpha_\omega$ is a coboundary}\,\}$
1280: is an analytic subset of~$\erg$.
1281: \item \label{TrivialCocycle-trivial}
1282: If $\alpha_\omega$ is a coboundary, for a.e.\ $\omega \in \erg'$, then
1283: $\alpha$ is a coboundary.
1284: \end{enumerate}
1285: \end{cor}
1286:
1287: \begin{proof}
1288: Let $\beta$ be the trivial cocycle, and apply Cor.~\ref{CohoErg}.
1289: \end{proof}
1290:
1291: \begin{defn} \label{HomoCocDefn}
1292: Recall that a Borel cocycle $\alpha \colon G \times S \to H$ is a
1293: \emph{constant} (or \emph{homomorphism}) cocycle if $\alpha(g,s)$ is
1294: essentially independent of~$s$, for each $g \in G$.
1295: \end{defn}
1296:
1297: \begin{cor} \label{CohoHomo}
1298: Let $G$, $H$, $X$, $\rho$, $\psi$, $\Omega$, $\alpha$, $\alpha_\omega$,
1299: and~$\erg'$ be as in Cor.~\ref{CohoErg}.
1300: If $\alpha_\omega$ is cohomologous to a constant cocycle, for a.e.\
1301: $\omega \in \erg'$, then $\alpha$ is cohomologous to a constant cocycle.
1302: \end{cor}
1303:
1304: \begin{proof}
1305: Similar to the proof of Cor.~\fullref{CohoErg}{trivial}, but with
1306: $\func = \erg' \times \Const$, where
1307: $$ \Const = \{\, f \colon G \times S \to H \mid \mbox{$f(g,s)$ is
1308: essentially independent of~$s$} \,\} .$$
1309: \end{proof}
1310:
1311: \begin{notn}
1312: For any locally compact, second countable group~$H$,
1313: we use $\cpct(H)$ to denote the set of compact subgroups of~$H$. It is
1314: well known that this is a complete, separable metric space, under the
1315: Hausdorff metric
1316: $$ d(K_1,K_2) = \max_{k_1 \in K_1} \dist(k_1,K_2) + \max_{k_2 \in K_2}
1317: \dist(K_1,k_2) ,$$
1318: where $\dist$ is any metric on~$H$.
1319: \end{notn}
1320:
1321: \begin{cor} \label{CohoCpctImg}
1322: Let $G$, $H$, $X$, $\rho$, $\psi$, $\Omega$, $\alpha$, $\alpha_\omega$,
1323: and~$\erg'$ be as in Cor.~\ref{CohoErg}.
1324:
1325: If, for a.e.\ $\omega \in \erg'$, the cocycle $\alpha_\omega$ is
1326: cohomologous to a cocycle whose essential range is contained in a compact
1327: subgroup of~$H$, then there are
1328: \begin{itemize}
1329: \item a Borel function $\kappa \colon \erg' \to \cpct(H)$,
1330: and
1331: \item a Borel cocycle $\beta$ that is cohomologous to~$\alpha$,
1332: \end{itemize}
1333: such that the essential range of~$\beta_\omega$ is
1334: contained in~$\kappa(\omega)$, for a.e.\ $\omega \in \erg'$.
1335: \end{cor}
1336:
1337: \begin{proof}
1338: As in the proof of Cor.~\ref{CohoErg}, we assume the notation of
1339: Thm.~\ref{AnalErgComp}, and we assume the cocycle~$\alpha$ is strict.
1340: Let
1341: $$ \func^+ = \{\, (f,K) \in \F(G \times S, H) \times \cpct(H) \mid
1342: \EssRg(f) \subset K \,\} ,$$
1343: and let $\func$ be the projection of~$\func^+$ to $\F(G \times S, H)$.
1344: Then $\func^+$ is closed, so $\func$ is analytic. Applying
1345: Thm.~\fullref{AnalErgComp}{coho} yields a Borel cocycle~$\beta$,
1346: cohomologous to~$\alpha$, such that $\beta_\omega \in F$, for a.e.\
1347: $\omega \in \erg'$. Now the Borel function $\kappa$ is obtained from
1348: the von Neumann Selection Theorem \pref{vonNeumann}.
1349: \end{proof}
1350:
1351:
1352: \begin{cor} \label{HomoModCpct}
1353: Let $G$, $H$, $X$, $\rho$, $\psi$,
1354: $\Omega$, $\alpha$, $\alpha_\omega$, and~$\erg'$ be as in
1355: Cor.~\ref{CohoErg}.
1356:
1357: If, for a.e.\ $\omega \in \erg'$, there are a compact subgroup~$K_\omega$
1358: of~$H$ and a Borel cocycle~$\beta_\omega$, cohomologous
1359: to~$\alpha_\omega$, such that
1360: \begin{enumerate} \renewcommand{\theenumi}{\alph{enumi}}
1361: \item \label{HomoModCpct-range}
1362: the essential range of~$\beta_\omega$ is contained in the normalizer
1363: $N_H(K_\omega)$,
1364: and
1365: \item \label{HomoModCpct-homo}
1366: the induced cocycle $\overline{\beta_\omega} \colon G \times S \to
1367: N_H(K_\omega)/K_\omega$ is a homomorphism cocycle,
1368: \end{enumerate}
1369: then there are
1370: \begin{enumerate}
1371: \item a Borel cocycle~$\beta$, cohomologous to $\alpha$,
1372: and
1373: \item a Borel function $\kappa \colon \erg \to \cpct(H)$,
1374: \end{enumerate}
1375: such that \pref{HomoModCpct-range} and~\pref{HomoModCpct-homo} hold with
1376: $\beta_\omega = \check\beta(\omega)$ and $K_\omega = \kappa(\omega)$, for
1377: a.e.\ $\omega \in \erg'$.
1378: \end{cor}
1379:
1380: \begin{proof}
1381: Let
1382: $$ \func^+ = \bigset{ (f,K) \in \F \bigl( G, \F(S, H) \bigr) \times
1383: \cpct(H) }{\
1384: \begin{matrix}
1385: \mbox{for a.e.\ $g \in G$, $ \exists h \in N_H(K)$, } \\
1386: \mbox{such that $\EssRg\bigl( f(g) \bigr) \subset h \, K$}
1387: \end{matrix}
1388: } .$$
1389: Then $\func^+$ is closed, so the proof is similar to that of
1390: Cor.~\ref{CohoCpctImg}. (Recall that $\F \bigl( G, \F(S, H) \bigr)$ is
1391: naturally homeomorphic to $\F(G \times S, H)$ \see{F(F)}.)
1392: \end{proof}
1393:
1394:
1395:
1396: \section{Superrigidity for non-ergodic actions}
1397: \label{section:superrigid}
1398:
1399: One main application of the results in this paper is to prove
1400: general versions of superrigidity for cocycles for non-ergodic
1401: actions of certain groups. We now define this class of groups. Let $I$ be a
1402: finite index set and for each $i{\in}I$, we let $k_i$ be a local
1403: field of characteristic zero and ${\mathbb G}_i$ be a connected
1404: simply connected semisimple algebraic $k_i$-group. We first define
1405: groups $G_i$, and then let $G=\prod_{i{\in}I}G_i$. If $k_i$ is
1406: non-Archimedean,
1407: $G_i={\mathbb G}_i(k_i)$ the
1408: $k_i$-points of ${\mathbb G}_i$. If $k_i$ is Archimedean, then
1409: $G_i$ is either ${\mathbb G}_i(k_i)$ or its topological universal
1410: cover. (This makes sense, since when $\Ga_i$ is simply connected
1411: and $k_i$ is Archimedean, ${\mathbb G}_i(k_i)$
1412: is topologically connected.) We assume that the $k_i$-rank of any
1413: simple factor of any ${\mathbb G}_i$ is at least two.
1414:
1415: We will need one assumption on the cocycles we consider.
1416:
1417: \begin{defn}
1418: \label{defn:quasi-integrable} Let $D$ be a locally compact group,
1419: $(S,\mu)$ a standard probability measure space on which $D$ acts
1420: preserving $\mu$ and $H$ be a normed topological group. We call a
1421: cocycle $\alpha\colon D{\times}S{\rightarrow}H$ over the $D$
1422: action \emph{$D$-integrable} if for any compact subset
1423: $M\subset{D}$, the function
1424: $Q_{M,\alpha}(x)=\sup_{m{\in}M}\ln^+\|\alpha(m,x)\|$ is in
1425: $L^1(S)$ (recall that $\ln^+ x = \max{(\ln x, 0)}$).
1426: \end{defn}
1427:
1428: Any continuous cocycle over a continuous action on a compact
1429: topological space is automatically $D$-integrable. We remark that
1430: a cocycle over a cyclic group action is $D$-integrable if and only
1431: if $\ln^+\|(\alpha(\pm{1},x)\|$ is in $L^1(S)$.
1432:
1433: We first recall the superrigidity theorems from \cite{FisherMargulis} for ergodic actions.
1434:
1435: \begin{thm}
1436: \label{theorem:Gsuperrigidity} Let $G$ be as above, let $(S,\mu)$
1437: be a standard probability measure space and let $H$ be the $k$
1438: points of a $k$-algebraic group where $k$ is a local field of
1439: characteristic $0$. Assume $G$ acts ergodically on $S$ preserving
1440: $\mu$. Let $\alpha\colon G{\times}S{\rightarrow}H$ be a
1441: $G$-integrable Borel cocycle. Then $\alpha$ is cohomologous to a
1442: cocycle $\beta$ where $\beta(g,x)=\pi(g){c(g,x)}$. Here
1443: ${\pi\colon G{\rightarrow}H}$ is a continuous homomorphism and
1444: $c\colon G{\times}S{\rightarrow}C$ is a cocycle taking values in a
1445: compact group centralizing $\pi(G)$.
1446: \end{thm}
1447:
1448: \begin{thm}
1449: \label{theorem:Gammasuperrigidity} Let $G,S,H$ and $\mu$ be as
1450: Theorem \ref{theorem:Gsuperrigidity} and let $\G<G$ be a lattice.
1451: Assume $\Gamma$ acts ergodically on $S$ preserving $\mu$. Assume
1452: $\alpha\colon \Gamma{\times}S{\rightarrow}H$ is a
1453: $\Gamma$-integrable, Borel cocycle. Then $\alpha$ is cohomologous
1454: to a cocycle $\beta$ where
1455: $\beta(\gamma,x)=\pi(\gamma){c(\gamma,x)}$. Here ${\pi\colon
1456: G{\rightarrow}H}$ is a continuous homomorphism of $G$ and $c\colon
1457: \Gamma{\times}X{\rightarrow}C$ is a cocycle taking values in a
1458: compact group centralizing $\pi(G)$.
1459: \end{thm}
1460:
1461: To state non-ergodic versions of the above theorems, we will need
1462: a Borel structure on the space of homomorphisms for $G$ to $H$.
1463: Given $G$ as above and $H$ as in Theorem
1464: \ref{theorem:Gsuperrigidity}, it is well-known that there are only
1465: finitely many conjugacy classes of homomorphisms $\pi\colon
1466: G{\rightarrow}H$. We choose a set $\Pi=\{\pi_i\}$ of
1467: representatives and endow it with the discrete topology, so as to
1468: be able to consider measurable maps to $\Pi$.
1469:
1470: Given a group $D$ acting on a standard probability measure space
1471: $(X,\mu)$, we denote by $\erg$ the space of ergodic components of
1472: the action and let $p\colon S{\rightarrow}\erg$ be the natural
1473: projection. We now state the general versions of the
1474: superrigidity theorems above.
1475:
1476: \begin{thm}
1477: \label{theorem:Gsuperrigidityne} Let $G$ be as above, let
1478: $(X,\mu)$ be a standard probability measure space and let $H$ be
1479: the $k$ points of a $k$-algebraic group where $k$ is a local field
1480: of characteristic $0$. Assume $G$ acts on $S$ preserving $\mu$.
1481: Let $\alpha\colon G{\times}S{\rightarrow}H$ be a $G$-integrable
1482: Borel cocycle. Then there exist measurable maps
1483: $\pi\colon\erg{\rightarrow}\Pi,
1484: \kappa\colon\erg{\rightarrow}\cpct(H)$ and $\phi\colon
1485: S{\rightarrow}H$ with $\kappa(p(x)){\subset}Z_H(\pi(p(x))$ almost
1486: everywhere such that
1487: $$\alpha(g,x)=\phi(gx)^{-1}\beta(g,x)\phi(x)$$
1488: where $\beta(g,x)=\pi(p(x))(g){c(g,x)}$. Here $c\colon
1489: G{\times}S{\rightarrow}H$ is a measurable cocycle with
1490: $c(g,x){\in}\kappa(p(x))$ almost everywhere.
1491: \end{thm}
1492:
1493: \begin{thm}
1494: \label{theorem:Gammasuperrigidityne} Let $G,S,H$ and $\mu$ be as
1495: Theorem \ref{theorem:Gsuperrigidityne} and let $\G<G$ be a
1496: lattice. Assume $\Gamma$ acts on $X$ preserving $\mu$. Assume
1497: $\alpha\colon\Gamma{\times}S{\rightarrow}H$ is a
1498: $\Gamma$-integrable, Borel cocycle. Then there exist measurable
1499: maps $\pi\colon\erg{\rightarrow}\Pi,
1500: \kappa\colon\erg{\rightarrow}\cpct(H)$ and $\phi\colon
1501: S{\rightarrow}H$ with $\kappa(p(x)){\subset}Z_H(\pi(p(x)))$ such
1502: that $\alpha(\g,x)=\phi({\g}x)^{-1}\beta(\g,x)\phi(x)$ where
1503: $\beta(\g,x)=\pi(p(x))(\g){c(\g,x)}$. Here
1504: $c\colon\G{\times}S{\rightarrow}H$ is a measurable cocycle with
1505: $c(\g,x){\in}\kappa(p(x))$ almost everywhere. \end{thm}
1506:
1507: \begin{proof}[Proof of Theorems \ref{theorem:Gsuperrigidityne} and
1508: \ref{theorem:Gammasuperrigidityne}] These are an immediate
1509: consequence of Theorems \ref{theorem:Gsuperrigidity} and
1510: \ref{theorem:Gammasuperrigidity}, Corollary \ref{HomoModCpct}, and
1511: Proposition \ref{RohlinDecomp}. Moreover, one can also prove
1512: these results by using the proof of Theorems
1513: \ref{theorem:Gsuperrigidity} and \ref{theorem:Gammasuperrigidity}
1514: from \cite{FisherMargulis}, Proposition \ref{RohlinDecomp} and
1515: Corollary \ref{AlgHull} below.
1516: \end{proof}
1517:
1518: There are also versions of Theorems \ref{theorem:Gsuperrigidityne}
1519: and \ref{theorem:Gammasuperrigidityne} which do not require that
1520: we assume the cocycle is $G$-integrable and versions, in that
1521: context, where the class of $G$ considered can be somewhat
1522: broader, i.e. $G$ of rank at least $2$, with some/all simple
1523: factors of rank $1$. To remove the $G$-integrability assumption
1524: requires assumptions on the algebraic hull of the cocycle, while
1525: weakening the rank assumption requires both assumptions on the
1526: algebraic hull and the assumption that the $G$ action on each
1527: ergodic component of $(X,\mu)$ is weakly irreducible. These
1528: assumptions are less natural in the non-ergodic setting, so we
1529: leave it to the interested reader to formulate and prove such
1530: results, using Theorems $3.6$ and $3.7$ of \cite{FisherMargulis}
1531: in place of Theorems \ref{theorem:Gsuperrigidity} and
1532: \ref{theorem:Gammasuperrigidity} above.
1533:
1534:
1535: \section{Equivariant maps on ergodic components}
1536:
1537: This section uses von Neumann Selection Theorem \pref{vonNeumann} to
1538: prove that if almost every ergodic component of a $G$-action has a
1539: fixed standard Borel $G$-space~$X$ as a measurable quotient, then
1540: $X$~is a measurable quotient of the entire action. Actually, the
1541: conclusion is proved in a more general setting that includes twisting by
1542: cocycles \see{QuotErgComps}. This yields a corollary \pref{AlgHull} that
1543: obtains information about a cocycle from the algebraic hulls of its
1544: restrictions to ergodic components.
1545:
1546:
1547: \begin{defn}[{\cite[\S3.2.0, pp.~216--217]{MargulisBook}}]
1548: \label{TwistDefn}
1549: Suppose
1550: \begin{itemize}
1551: \item $G$ and $H$ are locally compact, second countable groups,
1552: \item $S$ and $X$ are Borel spaces,
1553: \item $\rho \colon G \times S \to S$ and $\tau \colon H \times X \to X$
1554: are Borel actions,
1555: \item $\mu$ is a probability measure on~$S$,
1556: and
1557: \item $\alpha \colon G \times S \to H$ is a Borel cocycle.
1558: \end{itemize}
1559: Then:
1560: \begin{enumerate}
1561: \item \label{TwistDefn-action}
1562: We define an action $\zeta_{\rho,\tau,\alpha} \colon G \times \F(S,X)
1563: \to \F(S,X)$ by
1564: $$ \zeta_{\rho,\tau,\alpha}(g,\phi)(s) = \tau \Bigl( \alpha(g,s), \phi
1565: \bigl( \rho(g^{-1},s) \bigr) \Bigr) .$$
1566: We may refer to $\zeta_{\rho,\tau,\alpha}$ as the
1567: \emph{$\alpha$-twisted action} of~$G$ on $\F(S,X)$. It is Borel
1568: \see{TwistisBorel}.
1569: \item \label{TwistDefn-equi}
1570: A function $\phi \colon S \to X$ is \emph{essentially $(\rho,\tau,
1571: \alpha)$-equivariant} if, for each $g \in G$, we have
1572: $$\mbox{$\phi \bigl( \rho(g,s) \bigr) = \tau \bigl( \alpha(g,s),
1573: \phi(s) \bigr)$ for a.e.~$s \in S$}. $$
1574: In other words, a Borel function $\phi \colon S \to X$ is essentially
1575: $(\rho,\tau, \alpha)$-equivariant if and only if it represents a fixed
1576: point of the $\alpha$-twisted action of~$G$ on $\F(S,X)$.
1577: \end{enumerate}
1578: \end{defn}
1579:
1580: \begin{prop} \label{AlphaEquiComps}
1581: Let
1582: \begin{itemize}
1583: \item $G$, $H$, $X$, $\rho$, $\psi$, $\alpha$, and~$\alpha_\omega$ be as
1584: in Cor.~\ref{CohoErg},
1585: \item $\rho_\omega$ be the restriction of~$\rho$ to $G
1586: \times \psi^{-1}(\omega)$, for each $\omega \in \erg$,
1587: \item $(Y,d)$ be a complete, separable metric space,
1588: \item $\tau \colon H \times Y \to Y$ be a continuous action of~$H$
1589: on~$Y$,
1590: \item $\erg'$ be a conull Borel subset of~$\erg$, such that
1591: $\alpha_\omega$ is a Borel cocycle, for each $\omega \in \erg'$,
1592: and
1593: \item $\erg^X = \bigset{ \omega \in \erg' }{
1594: \begin{matrix}
1595: \mbox{there is a Borel map $\phi_\omega \colon \psi^{-1}(\omega) \to Y$}
1596: \\
1597: \mbox{that is essentially $(\rho_\omega, \tau,
1598: \alpha_\omega)$-equivariant}
1599: \end{matrix}} $.
1600: \end{itemize}
1601: Then:
1602: \begin{enumerate}
1603: \item \label{AlphaEquiComps-anal}
1604: $\erg^X$ is analytic,
1605: and
1606: \item \label{AlphaEquiComps-equi}
1607: there is an essentially $(\rho,\tau,\alpha)$-equivariant Borel map
1608: $\phi \colon X \to Y$ if and only if $\erg^X$ is conull
1609: in~$\erg$.
1610: \end{enumerate}
1611: \end{prop}
1612:
1613: \begin{proof}
1614: From Prop.~\ref{RohlinDecomp} (and Rem.~\ref{UnionOfProds}), we may
1615: assume the notation of Thm.~\ref{AnalErgComp}.
1616: Recall that $\rho$, $\tau$, and~$\alpha$ induce Borel maps
1617: \begin{itemize}
1618: \item $\rho_{\F} \colon G \times \erg \times \F(S,X) \to \F(S,X)$
1619: \see{BorelIntoF},
1620: \item $\tau^{\F} \colon \F(S,H) \times \F(S,X) \to \F(S,X)$
1621: \see{PointwiseOps}, and
1622: \item $\check\alpha \colon G \times \erg \to \F(S,H)$ \see{FiberFunc}.
1623: \end{itemize}
1624: Let $G_0$ be a countable, dense subset of~$G$, and define
1625: $$ \func = \bigl\{\, (\omega,f ) \in \erg \times \F(S,X)
1626: \mid \mbox{$\rho_{\F}(g,\omega,f) = \tau^{\F} \bigl(
1627: \check\alpha(g,\omega),f \bigr)$ for all $g \in G_0$} \,\} .$$
1628: Then $\func$ is an analytic set (in fact, it is Borel, because
1629: $\rho_{\func}$, $\tau^{\func}$, and~$\check\alpha$ are Borel and $G_0$ is
1630: countable).
1631:
1632: \pref{AlphaEquiComps-anal} $\erg^X$ is the projection of the analytic
1633: set~$\func$ to~$\erg$.
1634:
1635: (\ref{AlphaEquiComps-equi}~$\Rightarrow$) For a.e.\ $\omega \in \erg$,
1636: the map $\phi_\omega$, defined by $\phi_\omega(s) = \phi(\omega,s)$, is
1637: essentially $(\rho_\omega, \tau, \alpha_\omega)$-equivariant.
1638:
1639: (\ref{AlphaEquiComps-equi}~$\Leftarrow$)
1640: Because $\func$ is analytic, we may apply the von Neumann Selection
1641: Theorem \pref{vonNeumann}. By assumption, $\erg_{\func}$ is conull
1642: in~$\erg$, so we conclude that there is a Borel function $\phi \colon
1643: \erg \to \F(S,X)$, such that $\bigl( \omega, \phi(\omega) \bigr) \in
1644: \func$, for a.e.\ $\omega \in \erg$. Lemma~\ref{DefdFiberwise} provides
1645: us with a corresponding Borel function $\hat\phi \colon \erg \times S \to
1646: X$. By applying Fubini's Theorem, we see, for each $g \in G_0$, that
1647: \begin{equation} \label{QuotOfComps-equi}
1648: \mbox{$\hat\phi \bigl( \rho(g,\omega, s) \bigr) = \tau \bigl(
1649: \alpha(g,\omega,s), \hat\phi(\omega, s) \bigr)$ for a.e.~$(\omega,s) \in
1650: \erg \times S$}.
1651: \end{equation}
1652: Let
1653: $$ H = \{\, g \in G \mid \mbox{\pref{QuotOfComps-equi} holds} \,\} .$$
1654: Then $H$ is the stabilizer of~$\hat\phi$ under the $\alpha$-twisted
1655: action of~$G$ on $\F(\erg \times S, X)$ (see
1656: Defn.~\fullref{TwistDefn}{action}). Because the action is Borel, we know
1657: that $H$ is a closed subgroup of~$G$. On the other hand, $H$ is dense,
1658: because it contains~$G_0$. Therefore, $H = G$, which means that
1659: $\hat\phi$~is essentially $(\rho,\tau,\alpha)$-equivariant.
1660: \end{proof}
1661:
1662: Any continuous homomorphism $\pi \colon G \to H$ determines a
1663: ``constant" cocycle $\pi^\times \colon G \times X \to H$, defined by
1664: $\pi^\times(g,x) = \pi(g)$ \cf{HomoCocDefn}. In the statement of the
1665: following corollary, we ignore the distinction between $\pi$
1666: and~$\pi^\times$.
1667:
1668: \begin{cor} \label{QuotErgComps}
1669: Let
1670: \begin{itemize}
1671: \item $G$, $H$, $X$, $\rho$, $\psi$, $\rho_\omega$, $Y$, and $\tau$ be
1672: as in Prop.~\ref{AlphaEquiComps},
1673: and
1674: \item $\pi \colon G \to H$ be a continuous homomorphism.
1675: \end{itemize}
1676: There exists an essentially $(\rho,\tau,\pi)$-equivariant Borel map $\phi
1677: \colon X \to Y$ if and only if there exists an essentially
1678: $(\rho_\omega, \tau, \pi)$-equivariant Borel map $\phi_\omega \colon
1679: \psi^{-1}(\omega) \to Y$ for a.e.\ $\omega \in \erg$.
1680: \end{cor}
1681:
1682: \begin{defn}[{\cite[Defn.~9.2.2]{ZimmerBook}}]
1683: If
1684: \begin{itemize}
1685: \item $G$ acts ergodically on $X$,
1686: \item $\alpha \colon G \times X \to H$ is a Borel cocycle,
1687: \item $\field$ is a local field,
1688: and
1689: \item $H$ is the $\field$-points of an algebraic group over~$\field$,
1690: \end{itemize}
1691: then there exists a Zariski-closed subgroup~$L$ of~$H$, such that
1692: \begin{enumerate}
1693: \item $\alpha$ is cohomologous to a cocycle taking values in~$L$,
1694: and
1695: \item $\alpha$ is \emph{not} cohomologous to a cocycle taking values in
1696: any proper Zariski-closed subgroup of~$H$.
1697: \end{enumerate}
1698: The subgroup~$L$ is unique up to conjugacy. It is called
1699: the \emph{algebraic hull} of~$\alpha$.
1700: \end{defn}
1701:
1702: \begin{cor} \label{AlgHull}
1703: Let
1704: \begin{itemize}
1705: \item $(X,\mu)$ be a standard Borel probability space,
1706: \item $\rho \colon G \times X \to X$ be a Borel action, such that $\mu$
1707: is quasi-invariant,
1708: \item $\psi \colon X' \to \erg$ be the corresponding ergodic
1709: decomposition \see{ErgDecompThm},
1710: \item $\field$ be a local field,
1711: \item $H$ be the $\field$-points of an algebraic group over~$\field$,
1712: and
1713: \item $\alpha \colon G \times X \to H$ be a Borel cocycle.
1714: \end{itemize}
1715: Then $\alpha$ is cohomologous to a Borel cocyle $\beta \colon G \times
1716: X \to H$, such that, for a.e.\ $\omega \in \erg$, the Zariski closure of
1717: the range of~$\beta_\omega$ is equal to the algebraic hull
1718: of~$\beta_\omega$ {\rm(}where $\beta_\omega$ is the restriction
1719: of~$\beta$ to $G \times \psi^{-1}(\omega)${\rm)}.
1720: \end{cor}
1721:
1722: \begin{proof}
1723: Choose $\erg'$ as in \fullref{CohoErg}{coc} (with $\beta = \alpha$).
1724: Chevalley's Theorem \cite[Thm.~5.1, p.~89]{Borel-LinAlgGrps} implies
1725: there is a countable collection $\{(\tau_i,V_i)\}_{i=0}^\infty$ of
1726: rational representations of~$H$, such that every Zariski-closed subgroup
1727: of~$H$ is the stabilizer of some point in some projective space
1728: $\proj{V_i}$. For $c,d, i\in \natural$, define
1729: \begin{itemize}
1730: \item $\erg_{c,d} = \bigset{ \omega \in \erg' }{
1731: \begin{matrix}
1732: \mbox{the algebraic hull of $\alpha_\omega$ is $d$-dimensional} \\
1733: \mbox{and has exactly $c$ connected components}
1734: \end{matrix} }$,
1735: \item $ Y_{c,d}^i = \bigset{ v \in \proj{V_i} }{
1736: \begin{matrix}
1737: \mbox{$\dim \Stab_H(v) = d$, and} \\
1738: \mbox{$\Stab_H(v)$ has exactly $c$ connected components} \\
1739: \end{matrix}
1740: }$,
1741: \item $Y_{c,d}
1742: = \coprod_{i=0}^\infty Y_{c,d}^i$ (disjoint union),
1743: and
1744: \item $\tau \colon H \times Y_{c,d} \to Y_{c,d}$ by $\tau(h,x) = \tau_i(h)
1745: x$ if $x \in \proj{V_i}$.
1746: \end{itemize}
1747: Each $Y_{c,d}^i$ is Borel (see \ref{DimBorel} below), so $Y_{c,d}$ is a
1748: standard Borel space. Therefore, combining the Cocycle Reduction Lemma
1749: \cite[Lem.~5.2.11, p.~108]{ZimmerBook} with
1750: Prop.~\fullref{AlphaEquiComps}{anal} and Rem.~\ref{Anal->Meas} implies
1751: that $\erg_{c,d}$ is absolutely measurable. Thus, we may let
1752: $\erg'_{c,d}$ be a conull, Borel subset of~$\erg_{c,d}$, for each~$d$.
1753:
1754: There is no harm in assuming $\erg = \erg'_{c,d}$, for some~$c$ and~$d$.
1755: Then there is an essentially $(\rho_\omega, \tau,
1756: \alpha_\omega)$-equivariant Borel map from $\psi^{-1}(\omega)$ to
1757: $Y_{c,d}$, for a.e.\ $\omega \in \erg$, so
1758: Cor.~\fullref{AlphaEquiComps}{equi} implies that there is an essentially
1759: $(\rho, \tau, \alpha)$-equivariant Borel map from $X$ to~$Y_{c,d}$. This
1760: means that $\alpha$ is cohomologous to a cocycle~$\beta$, such that the
1761: essential range of~$\beta_\omega$ is contained in an algebraic group
1762: whose dimension and number of connected components are no more than those
1763: of the algebraic hull of~$\beta_\omega$. By changing $\beta$ on a set of
1764: measure~$0$, we may assume that the entire range of~$\beta_\omega$ is
1765: contained in this subgroup. So the desired conclusion follows from the
1766: minimality (and uniqueness) of the algebraic hull of~$\beta_\omega$.
1767: \end{proof}
1768:
1769: The following observation, used in the proof of Cor.~\ref{AlgHull} above,
1770: must be well known, but the authors do not know of a reference.
1771:
1772: \begin{lem} \label{DimBorel}
1773: Suppose
1774: \begin{itemize}
1775: \item $G$ is a real Lie group,
1776: \item $M$ is a Polish space,
1777: and
1778: \item $\rho \colon G \times M \to M$ is a continuous action, such that
1779: the stabilizer $\Stab_G(m)$ has only finitely many connected components,
1780: for each $m \in M$.
1781: \end{itemize}
1782: Then:
1783: \begin{enumerate}
1784: \item \label{DimBorel-dim}
1785: $\dim \Stab_G(m)$ is a Borel function of $m \in M$.
1786: \item \label{DimBorel-cpct}
1787: For each compact subgroup~$K$ of~$G$,
1788: $$ \{\, m \in M \mid \mbox{$K$ is conjugate to a maximal compact
1789: subgroup of $\Stab_G(m)$} \,\}$$
1790: is Borel.
1791: \item \label{DimBorel-dimcpct}
1792: The dimension of the maximal compact subgroup of $\Stab_G(m)$ is a
1793: Borel function of~$m$.
1794: \item \label{DimBorel-comp}
1795: The number of connected components of $\Stab_G(m)$ is a Borel
1796: function of~$m$.
1797: \item \label{DimBorel-dimcomp}
1798: For each $d$ and~$c$,
1799: $$ \bigset{ m \in M }{
1800: \begin{matrix}
1801: \mbox{$\dim \Stab_G(m) = d$, and} \\
1802: \mbox{$\Stab_G(m)$ has exactly~$c$ connected components}
1803: \end{matrix} }$$
1804: is Borel.
1805: \end{enumerate}
1806: \end{lem}
1807:
1808: \begin{proof}
1809: \pref{DimBorel-dim}
1810: It is well known (and easy to see) that $\dim \Stab_G(m)$ is an upper
1811: semicontinuous function of $m \in M$.
1812:
1813: \pref{DimBorel-cpct} The fixed-point set $M^K$ is closed, and $G$ is
1814: $\sigma$-compact, so $\rho(G, M^K)$ is a countable union of closed sets.
1815: Therefore $\rho(G, M^K)$ is Borel. Now
1816: $$ \rho(G, M^K) = \{\, m \in M \mid \mbox{$\Stab_G(m)$
1817: contains a conjugate of~$K$}\,\} ,$$
1818: so
1819: $$\bigset{ m \in M }{ \begin{matrix}
1820: \mbox{$K$ is conjugate to a maximal} \\
1821: \mbox{compact subgroup of $\Stab_G(m)$}
1822: \end{matrix} }
1823: = \rho(G, M^K) \smallsetminus \bigcup_{K' \supset K} \rho(G, M^{K'}) .$$
1824: Any real Lie group has only countably many conjugacy classes of
1825: compact subgroups \cite[Prop.~10.12]{Adams}, so the union is countable.
1826: Therefore, this is a Borel set.
1827:
1828: (\ref{DimBorel-dimcpct}, \ref{DimBorel-comp}) Immediate from
1829: \pref{DimBorel-cpct}. (Recall that the number of components of
1830: $\Stab_G(m)$ is the same as the number of components of any of its
1831: maximal compact subgroups.)
1832:
1833: \pref{DimBorel-dimcomp} Combine \pref{DimBorel-dim} and
1834: \pref{DimBorel-comp}.
1835: \end{proof}
1836:
1837:
1838: \begin{thebibliography}{GGM}
1839:
1840: \bibitem[Ad]{Adams}
1841: S.~Adams,
1842: \emph{Reduction of cocycles with hyperbolic targets,}
1843: Ergodic Theory Dynam. Systems {\bf 16} (1996), no. 6, 1111--1145,
1844: MR1424391 (98i:58135),
1845: Zbl 0869.58031.
1846:
1847: \bibitem[Ar]{Arveson}
1848: W.~Arveson,
1849: \emph{An invitation to $C\sp*$-algebras,}
1850: Springer-Verlag, New York, 1976,
1851: MR0512360 (58 \#23621),
1852: Zbl 0344.46123.
1853:
1854: \bibitem[Bo]{Borel-LinAlgGrps}
1855: A.~Borel,
1856: \emph{Linear Algebraic Groups,} 2nd ed.,
1857: Springer, New York, 1991,
1858: MR1102012 (92d:20001),
1859: Zbl 0726.20030.
1860:
1861: \bibitem[FM]{FisherMargulis}
1862: D.~Fisher and G.~A.~Margulis,
1863: \emph{Local rigidity for cocycles,}
1864: in S.T.Yau, ed.,
1865: \emph{Surveys in Differential Geometry,}
1866: Vol. 8,
1867: International Press, Boston, MA,
1868: (to appear).
1869:
1870: \bibitem[GGM]{GoodrichEtAl}
1871: K.~Goodrich, K.~Gustafson, and B.~Misra,
1872: \emph{On converse to Koopman's lemma,}
1873: Phys. A {\bf 102} (1980), no. 2, 379--388,
1874: MR0582372 (82i:82007).
1875:
1876: \bibitem[GS]{Greschonig-Schmidt}
1877: G.~Greschonig and K.~Schmidt,
1878: \emph{Ergodic decomposition of
1879: quasi-invariant probability measures,}
1880: Colloq. Math. {\bf 84/85} (2000), part 2, 495--514,
1881: MR1784210 (2001i:28021),
1882: Zbl 0972.37003.
1883:
1884: \bibitem[Mc]{Mackey-Ind}
1885: G.~Mackey,
1886: \emph{Induced representation of locally compact groups~I,}
1887: Ann. Math. (2) {\bf 55} (1952) 101--139,
1888: MR0044536 (13,434a),
1889: Zbl 0046.11601.
1890:
1891: \bibitem[Mr]{MargulisBook}
1892: G.~A.~Margulis,
1893: \emph{Discrete subgroups of semisimple Lie groups,}
1894: Springer-Verlag, Berlin, 1991,
1895: MR1090825 (92h:22021),
1896: Zbl 0732.22008.
1897:
1898: \bibitem[Ra]{Ramsay}
1899: A.~Ramsay,
1900: \emph{Virtual groups and group actions,}
1901: Adv. Math. {\bf 6} (1971) 253--322,
1902: MR0281876 (43 \#7590),
1903: Zbl 0216.14902.
1904:
1905: \bibitem[Ro]{Rohlin}
1906: V.~A.~Rohlin,
1907: \emph{On the fundamental ideas of measure theory,}
1908: Amer. Math. Soc. Translation {\bf 10} (1962) 1--54,
1909: MR0047744 (13,924e).
1910:
1911: \bibitem[Ru]{Rudin-RCAnal}
1912: W.~Rudin,
1913: \emph{Real and Complex Analysis,}
1914: 2nd ed.,
1915: McGraw Hill, New York, 1974,
1916: MR0344043 (49 \#8783),
1917: Zbl 0278.26001.
1918:
1919: \bibitem[WZ]{WheedenZygmund}
1920: R.~Wheeden and A.~Zygmund,
1921: \emph{Measure and integral,}
1922: Marcel Dekker, New York, 1977,
1923: MR0492146 (58 \#11295),
1924: Zbl 0362.26004.
1925:
1926: \bibitem[Z1]{ZimmerExtn}
1927: R.~J.~Zimmer,
1928: \emph{Extensions of ergodic group actions,}
1929: Illinois J. Math. {\bf 20} (1976) no. 3, 373--409,
1930: MR0409770 (53 \#13522),
1931: Zbl 0334.28015.
1932:
1933: \bibitem[Z2]{ZimmerBook}
1934: R.~J.~Zimmer,
1935: \emph{Ergodic theory and semisimple groups,}
1936: Birkh\"auser, Basel, 1984,
1937: MR0776417 (86j:22014),
1938: Zbl 0571.58015.
1939:
1940:
1941: \end{thebibliography}
1942:
1943: \end{document}
1944:
1945: