1: \documentclass[11pt,a4paper]{article}
2: \usepackage{epsfig}
3: \usepackage{amssymb}
4: \usepackage{epic}
5: \usepackage{eepic}
6: \usepackage[latin1]{inputenc}
7: \usepackage[T1]{fontenc}
8: \usepackage{psfrag}
9: % \usepackage[xdvi]{graphicx}
10: \def\P{{\mathbb P}}
11: \def\C{{\cal C}}
12: \def\Cox{\hfill \Box}
13: \def\GM{{\rm GM}}
14: \def\supp{{\rm supp}}
15: \newtheorem{theorem}{Theorem}[section]
16: \newtheorem{lemma}[theorem]{Lemma}
17: \newtheorem{corollary}[theorem]{Corollary}
18: \newtheorem{proposition}[theorem]{Proposition}
19: \newtheorem{definition}[theorem]{Definition}
20: \newtheorem{example}[theorem]{Example}
21: \newtheorem{remark}[theorem]{Remark}
22:
23: \author{Erik I. Broman\footnote{Dept.\ of Mathematics,
24: Chalmers University of Technology, S-412 96 G\"oteborg, Sweden,
25: \texttt{http://www.math.chalmers.se/\~{ }broman/}, research
26: supported by the Swedish Research Council.},
27: Olle H\"aggstr\"om\footnote{Dept.\ of Mathematics,
28: Chalmers University of Technology, S-412 96 G\"oteborg, Sweden,
29: \texttt{http://www.math.chalmers.se/\~{ }olleh/}, research
30: supported by the Swedish Research Council.}, and
31: Jeffrey E. Steif\footnote{Dept.\ of Mathematics,
32: Chalmers University of Technology, S-412 96 G\"oteborg, Sweden,
33: \texttt{http://www.math.chalmers.se/\~{ }steif/}, research
34: supported by the Swedish Research Council
35: and the G\"{o}ran Gustafsson Foundation (KVA).}}
36: \date{\today}
37:
38: \title{Refinements of stochastic domination}
39:
40: %\fleqn
41:
42: %\secnumdepth{1}
43: \begin{document}
44: \maketitle
45:
46: \begin{abstract}
47: In a recent paper by two of the authors,
48: the concepts of upwards and downwards
49: $\epsilon$-movability were introduced, mainly as a
50: technical tool for studying dynamical percolation of
51: interacting particle systems.
52: In this paper, we further explore these concepts which can be seen as
53: refinements or quantifications of stochastic domination, and we
54: relate them to previously studied concepts such as uniform
55: insertion tolerance and extractability.
56:
57:
58: \medskip\noindent
59: {\bf AMS subject classification:} 60G99.
60:
61: \medskip\noindent
62: {\bf Keywords and phrases:} finite energy, stochastic domination,
63: extractibility, rigidity
64:
65:
66: % \medskip\noindent
67: % {\bf Short title:} Determinants and two-block-factors
68:
69: \end{abstract}
70:
71:
72:
73: \section{Introduction}\label{intro}
74:
75: In \cite{BS}, Broman and Steif introduced certain refinements
76: of stochastic domination, which we call upwards and
77: downwards
78: $\epsilon-$movability; see Definition \ref{def_eps-stab} below.
79: These concepts were introduced mainly as a technical tool in the analysis
80: of dynamical percolation for interacting particle systems, but
81: they turn out to be interesting in their own right.
82: In the present paper, we
83: explore these concepts further and relate them to various previously
84: studied concepts.
85:
86: Let $S$ be a countable set. For $p \in [0,1]$,
87: let every $s\in S$, independently of all other elements in $S$,
88: take value $1$ with probability $p$ and take value $0$
89: with probability $1-p$. We let $\pi_{p}$
90: denote the corresponding product measure on $\{0,1\}^{S}$.
91: When talking about product measures on $\{0,1\}^S$,
92: we will always mean these uniform ones (with
93: the same $p$ for every $s \in S$).
94:
95: Let $\mu$ be an arbitrary
96: probability measure on $\{0,1\}^{S}$. For $\epsilon \in (0,1)$,
97: we will let $\mu^{(+,\epsilon)}$ denote the distribution of the process
98: obtained by first choosing an element of $\{0,1\}^{S}$ according to
99: $\mu$ and then independently changing each 0 to a 1 with probability
100: $\epsilon$. Similarly, we will let $\mu^{(-,\epsilon)}$ denote the
101: distribution of the process obtained by first choosing an element of
102: $\{0,1\}^{S}$ according to $\mu$ and then independently changing each 1 to
103: a 0 with probability $\epsilon$. Finally, for $\delta \in (0,1)$,
104: we let $\mu^{(-,\epsilon,+,\delta)}$
105: denote the distribution of the process obtained by first choosing an element
106: of $\{0,1\}^{S}$ according to $\mu$ and then independently changing each 0 to
107: a 1 with probability $\delta$ and each 1 to a 0 with probability $\epsilon$.
108:
109: It turns out that for any $\epsilon\in [0,1)$,
110: $\mu_1^{(+,\epsilon)}=\mu_2^{(+,\epsilon)}$ implies
111: that $\mu_1=\mu_2$. To
112: see this, it suffices to check that $\mu_1(A)=\mu_2(A)$ for events $A$ of
113: the type ``all $s \in S'$ take value $0$'' where $S' \subseteq S$
114: (this is easy),
115: and then use inclusion-exclusion. Similarly, of course,
116: $\mu_1^{(-,\epsilon)}=\mu_2^{(-,\epsilon)}$
117: implies $\mu_1=\mu_2$.
118:
119: For $\sigma,\sigma^{\prime}\in\{0,1\}^{S}$ we write
120: $\sigma \preceq \sigma^{\prime}$
121: if $\sigma(s)\leq \sigma^{\prime}(s)$ for every $s\in S.$ A
122: function $f:\{0,1\}^{S} \rightarrow {\mathbb R}$
123: is increasing if $f(\sigma)\leq f(\sigma^{\prime})$ whenever
124: $\sigma \preceq \sigma^{\prime}.$
125: For two probability measures $\mu,\mu^{\prime}$ on $\{0,1\}^{S}$, we
126: say that $\mu$ {\bf is stochastically dominated by} $\mu'$, and
127: write $\mu \preceq \mu^{\prime}$, if for every continuous
128: increasing function $f$ we have that $\mu(f)\leq \mu^{\prime}(f)$.
129: ($\mu(f)$ is shorthand for $\int f d\mu$.) By Strassens theorem
130: (see \cite[p.\ 72]{IPS}),
131: this is equivalent to the existence of random variables
132: $X,X^{\prime}\in\{0,1\}^{S}$ such that
133: $X$ has distribution $\mu$, $X^{\prime}$
134: has distribution $\mu^{\prime}$, and $X \preceq X^{\prime}$ a.s.
135: This is also equivalent to $\mu(A)\leq \mu^{\prime}(A)$ for all
136: up-sets $A$ where an up-set is a set whose indicaor function
137: $I_A$ is increasing. From now on ``$\sim$'' will mean ``has distribution''.
138:
139: \begin{definition} \label{def_eps-stab}
140: Let $(\mu_{1},\mu_{2})$ be a pair of
141: probability measures on $\{0,1\}^{S},$ where $S$ is a countable set.
142: Assume that $\mu_{1} \preceq \mu_{2}$.
143: If, given $\epsilon>0$, we have
144: \[
145: \mu_{1} \preceq \mu_{2}^{(-,\epsilon)},
146: \]
147: then we say that the pair $(\mu_{1},\mu_{2})$ is
148: {\bf downwards $\epsilon$-movable}.
149: $(\mu_{1},\mu_{2})$ is said to be {\bf downwards movable}
150: if it is downwards
151: $\epsilon$-movable for some $\epsilon >0$.
152: Analogously, if, given $\epsilon>0$, we have
153: \[
154: \mu_{1}^{(+,\epsilon)} \preceq \mu_{2},
155: \]
156: then we say that the pair ($\mu_{1},\mu_{2}$) is
157: {\bf upwards $\epsilon$-movable},
158: and we say that ($\mu_{1},\mu_{2}$)
159: is {\bf upwards movable} if the pair is upwards $\epsilon$-movable
160: for some $\epsilon >0$.
161: \end{definition}
162:
163: Note that if we restrict to the case where both $\mu_1$ and $\mu_2$
164: are product measures, then these concepts become trivial.
165:
166: In \cite{BS} a considerable amount of effort was spent on trying to show
167: downwards movability when the pair
168: considered was two stationary
169: distributions, corresponding to two different parameter values,
170: for some specific interacting particle system.
171: In particular, the so called contact process (see Liggett \cite{SIS}
172: for definitions and a survey) was investigated.
173: Considering $(\mu_1,\mu_2)$,
174: where $\mu_i$ is the upper invariant measure for the contact process
175: with infection rate $\lambda_i$,
176: it was shown in \cite{BS}
177: that if $\lambda_1<\lambda_2$, then the pair is downwards movable.
178:
179: Another result from \cite{BS}
180: is that if $\mu_1 \preceq \mu_2$,
181: $\mu_2$ satisfies the FKG lattice condition (see \cite[p.\ 78]{IPS})
182: and
183: \[
184: \displaystyle \inf_{\tilde{S} \subset S \atop |\tilde{S}|<\infty}
185: \inf_{s\in \tilde{S}\atop \xi \in \{0,1\}^{\tilde{S}\setminus s}}
186: [\mu_{2}(\sigma(s)=1|\sigma(\tilde{S}\setminus s)\equiv \xi)-
187: \mu_{1}(\sigma(s)=1|\sigma(\tilde{S}\setminus s)\equiv \xi)]
188: >0
189: \]
190: then $(\mu_1,\mu_2)$ is downwards movable.
191: This however is not sufficient to
192: get the result for the contact process mentioned above since
193: by \cite{L1}, the upper invariant measure for the contact process on
194: ${\mathbb Z}$ does not satisfy the FKG lattice condition when
195: $\lambda < 2$.
196:
197: In the present paper, we will concentrate on the case
198: where $\mu_{1}$ is a product measure but $\mu_{2}$ is not.
199: We now proceed with some further explanations and definitions needed
200: to state our main results, Theorems \ref{thm1} and \ref{thm:finite}
201: below.
202: In Sections \ref{examples}--\ref{secFKG},
203: we will establish a number of examples and auxiliary results,
204: while Section \ref{types} will tie things together giving proofs
205: of Theorems \ref{thm1} and \ref{thm:finite}.
206:
207:
208: For a probability measure $\mu$ on $\{0,1\}^{S}$,
209: define $p_{\sup,\mu}$ by
210: \[
211: p_{\sup,\mu}:=\sup \{p \in [0,1]: \pi_{p} \preceq \mu \}.
212: \]
213: Since the relation $\preceq$ is preserved under weak limits we see that
214: \[
215: \pi_{p_{\sup,\mu}} \preceq\mu
216: \]
217: and so the supremum is achieved. Therefore we also denote this by
218: $p_{\max,\mu}$.
219:
220:
221: If $p_{\max,\mu}=0$, then trivially $(\pi_{p_{\max,\mu}},\mu)$
222: is downwards movable but not upwards movable. Assume next
223: that $\mu$ is a probability measure with $p_{\max,\mu}>0$.
224: If $p \in [0,p_{\max,\mu})$,
225: then the pair $(\pi_{p},\mu)$ is trivially upwards
226: movable. It is also easy to see that it is downwards movable
227: by arguing as follows.
228: By Strassen's theorem, we may choose
229: $X \sim \mu$ and $Y \sim \pi_{p_{\max,\mu}}$ such that $X \geq Y$ a.s.
230: Then choose $\epsilon>0$ such that
231: $(1-\epsilon)p_{\max,\mu}>p$, and let $Z \sim \pi_{1-\epsilon}$
232: be independent of both $X$ and $Y$.
233: We obtain $\min(X,Z) \geq \min(Y,Z)$ a.s., and
234: since $\min(Y,Z) \sim \pi_{p_{\max,\mu}(1-\epsilon)}$ we conclude that
235: \[
236: \pi_{p} \preceq \mu^{(-,\epsilon)},
237: \]
238: as desired.
239:
240: The final case we are left with (when one of the measures is a uniform
241: product measure) is $(\pi_{p_{\max,\mu}},\mu)$ with $p_{\max,\mu}>0$.
242: This pair is by definition not upwards movable,
243: but we believe it is an interesting
244: question to ask if it is downwards movable and
245: this question motivates the following definition.
246: \begin{definition}
247: We say that $\mu$ is {\bf nonrigid} if the pair
248: $(\pi_{p_{\max,\mu}},\mu)$ is downwards movable and
249: otherwise we will say that $\mu$ is {\bf rigid}.
250: \end{definition}
251:
252: All uniform product measures other than $\delta_0$ are trivially rigid
253: while all $\mu$ such that $p_{\max,\mu}=0$ are trivially nonrigid.
254: Heuristically, it is natural to expect that as long as
255: $p_{\max,\mu}>0$, then typically $\mu$ should be rigid.
256: This issue turns out to be quite intricate, however; see
257: Proposition \ref{prop:rigid_when_finite} and Theorem
258: \ref{extype2_2} below.
259:
260: Other well known concepts which have arisen in a number of
261: problems and which we feel belong to this same circle of
262: ideas are those of finite energy (Newman and Schulman \cite{NS})
263: and insertion and deletion tolerance (Lyons and Schramm \cite{LSch}).
264:
265: \begin{definition}
266: We say that $\mu$ is {\bf $\epsilon$-insertion tolerant} if
267: for any $s\in S$, we have that
268: \begin{equation} \label{eqnUIT}
269: \mu(\sigma(s)=1|\sigma(S \setminus s))\geq
270: \epsilon \textrm{ a.s. }
271: \end{equation}
272: We say that $\mu$ is {\bf uniformly insertion tolerant} if it is
273: $\epsilon$-insertion tolerant for some $\epsilon>0$. The analogous notions of
274: {\bf $\epsilon$-deletion tolerant} and {\bf uniformly deletion tolerant} are
275: defined similarly (the ``$1$'' is replaced by ``$0$'').
276: Finally, we say $\mu$ has {\bf finite $\epsilon$-energy} if it is both
277: $\epsilon$-insertion tolerant and $\epsilon$-deletion tolerant, and that it has
278: {\bf uniform finite energy}
279: if it has finite $\epsilon$-energy for some $\epsilon>0$.
280: \end{definition}
281: Also closely related are the following notions of extractability;
282: we discuss some background on
283: this concept at the end of the introduction.
284: \begin{definition}
285: We call
286: $\mu$ {\bf $\epsilon$-upwards extractable} if
287: there exists a probability measure $\nu$
288: such that $\mu=\nu^{(+,\epsilon)}$.
289: We call $\mu$ {\bf uniformly upwards extractable} if
290: it is $\epsilon$-upwards extractable for some $\epsilon>0$.
291: The notions of {\bf $\epsilon$-downwards extractable} and
292: {\bf uniformly downwards extractable} are defined analogously
293: (the ``$+$'' is replaced by ``$-$''). Finally,
294: $\mu$ is called {\bf $\epsilon$-extractable} if
295: there exists a probability measure $\nu$
296: such that $\mu=\nu^{(-,\epsilon,+,\epsilon)}$,
297: and it is called {\bf uniformly extractable} if it is
298: $\epsilon$-extractable for some $\epsilon>0$.
299: \end{definition}
300: We are now equipped with all the definitions needed to state our main
301: theorem. We refer to Figure 1 for a comprehensive diagram over the
302: implications and non-implications that the theorem asserts.
303:
304: \begin{theorem} \label{thm1}
305: Let $S$ be a countable set and consider the following
306: properties of a probability measure $\mu$ on
307: $\{0,1\}^S$:
308: \begin{description}
309: \item{\rm (I) } $\mu$ is uniformly upwards extractable.
310: \item{\rm (II) } $\mu$ is uniformly insertion tolerant.
311: \item{\rm (III) } $\mu$ is rigid.
312: \item{\rm (IV) } There exists a $p>0$ such that $\pi_p \preceq \mu.$
313: \end{description}
314: We then have that {\rm (I)} $\Rightarrow$ {\rm (II)} $\Rightarrow$ {\rm (IV)}
315: and that {\rm (I)} $\Rightarrow$ {\rm (III)} $\Rightarrow$ {\rm (IV)} while
316: none of the four corresponding reverse implications hold. Also,
317: {\rm (III)} does not imply {\rm (II)}.
318: Moreover, with $S={\mathbb Z}$, there exist translation invariant
319: examples for all of the asserted nonimplications.
320: % \[
321: % {\rm (I)}
322: % \left. \begin{array}{c} \Rightarrow \end{array} \right.
323: % {\rm (II)}
324: % \left. \begin{array}{c} \Rightarrow \\ \not \Leftarrow \end{array} \right.
325: % {\rm (IV)}
326: % \]
327: % \[
328: %{\rm (I)}
329: % \left. \begin{array}{c} \Rightarrow \\ \not \Leftarrow \end{array} \right.
330: % {\rm (III)}
331: % \left. \begin{array}{c} \Rightarrow \\ \not \Leftarrow \end{array} \right.
332: % {\rm (IV)}
333: % \]
334: % \[
335: %{\rm (II)}
336: % \left. \begin{array}{c} \not \Rightarrow \\ \not \Leftarrow \end{array} \right.
337: %{\rm (III)}
338: % \]
339: % \[
340: %{\rm (IV)}
341: % \left. \begin{array}{c} \not \Rightarrow \end{array} \right.
342: %{\rm (II)} \mbox{ or } {\rm (III)}
343: % \]
344: %
345: \end{theorem}
346: In addition, it turns out that {\rm (IV)} does not even imply
347: ``{\rm (II)} or {\rm (III)}''; see Remark \ref{rem:neither(II)or(III)}.
348: Note that we have not managed to work out whether or not (II) implies (III).
349:
350:
351: \begin{figure}[hbt]
352: \begin{center}
353: \psfrag{1}{I}
354: \psfrag{2}{\hspace{-1mm}II}
355: \psfrag{3}{\hspace{-1mm}III}
356: \psfrag{4}{\hspace{-1mm}IV}
357: \includegraphics[width=0.4\textwidth]{b.eps}
358: \end{center}
359: \end{figure}
360:
361: \begin{quote}
362: {\small Figure 1.
363: Hasse diagram of the implications between properties (I), (II),
364: (III) and (IV) in Theorem \ref{thm1}: we have proved that
365: one property implies another iff there
366: is a downwards path in the diagram from the former to the latter. We do not
367: know whether the dashed line between (II) and (III) should be there or not,
368: i.e., whether or not uniform insertion tolerance implies rigidity.
369: As will be seen in Theorem \ref{thmFKG}, the
370: desired implication (II) $\Rightarrow$ (III) holds under an additional
371: FKG-like assumption. If we restrict to finite $S$, then
372: some of the implications will turn into equivalences; see Theorem
373: \ref{thm:finite}.}
374: \end{quote}
375:
376:
377: Some of the asserted implications are easy: (I) trivially implies (II).
378: The implication (III) $\Rightarrow $ (IV) is also trivial as we saw.
379: It is a direct application of Holley's inequality (see, e.g.,
380: \cite[Theorem 4.8]{GHM}) to see that
381: $\epsilon$-insertion tolerance implies that $\pi_\epsilon \preceq \mu$,
382: whence (II) implies (IV).
383: Thus, apart from the implication (I) $\Rightarrow$ (III)
384: (which is in fact not so hard either), we see all the implications claimed
385: in the theorem. Therefore
386: our interest in Theorem \ref{thm1} is more in the counterexamples showing
387: the distinction between some of these properties rather than in the
388: implications.
389:
390: As mentioned above, we do not know in general whether (II) implies (III).
391: However Theorem \ref{thmFKG} provides us
392: with a partial answer, telling us that this is true under
393: the extra assumption of $\mu$ being downwards FKG,
394: a property weaker than satisfying the FKG lattice condition and
395: defined as follows.
396: \begin{definition} \label{downFKG}
397: A measure $\mu$ on $\{0,1\}^S$ is downwards FKG if for any finite
398: $S^{\prime}\subset S$ and any increasing subsets $A,B$
399: \[
400: \mu(A\cap B| \sigma(S^{\prime})\equiv 0)
401: \geq\mu(A | \sigma(S^{\prime})\equiv 0) \mu(B | \sigma(S^{\prime})\equiv 0).
402: \]
403: \end{definition}
404: The concept of downwards FKG was made explicit in \cite{LigS} where
405: it was shown,
406: among other things, that for such translation invariant measures, stochastic
407: domination of a product measure has a very simple characterization.
408: As mentioned before, the upper invariant measure for the contact process
409: in one dimension and with $\lambda<2$ is known to not satisfy the FKG lattice condition.
410: In addition this is believed to be true for any value of $\lambda$ and any dimension.
411: However, it was proven in \cite{BHK} that it is downwards FKG for any dimension
412: and for all values of $\lambda.$
413: This result was then
414: exploited in \cite{LigS} to show that the upper invariant measure for the
415: contact process dominates product measures despite the fact that
416: the measure is not uniformly insertion tolerant.
417:
418: \begin{theorem} \label{thmFKG}
419: Let $\mu$ be a translation invariant downwards FKG
420: measure on $\{0,1\}^{{\mathbb Z}^{d}}.$ Then {\rm (II)} implies {\rm (III)}.
421: \end{theorem}
422: In Section \ref{secFKG} we prove an easy technical lemma that together
423: with some results of \cite{LS} will give us the following theorem (see
424: Section \ref{secFKG} for the definition of conditional negative associativity).
425:
426: \begin{proposition}\label{prop:neg}
427: Let $\mu$ be a translation invariant,
428: conditionally negatively associated measure on $\{0,1\}^{\mathbb Z}$.
429: Then {\rm (IV)} implies {\rm (III)}.
430: \end{proposition}
431: If we restrict to finite $S$, then further implications between
432: the various properties are available. By the support of a
433: measure $\mu$ on $\{0,1\}^S$, denoted by $\supp(\mu)$, we mean
434: $\{\xi \in \{0,1\}^S:\mu(\sigma(S)\equiv \xi)>0\}$.
435: \begin{theorem} \label{thm:finite}
436: Let $S$ be finite, and consider properties {\rm (I)}--{\rm (IV)} of
437: probability measures on $\{0,1\}^S$. We then have
438: \begin{equation} \label{eqnSfin1}
439: {\rm (I)} \, \Leftrightarrow \, {\rm (II)} \Leftrightarrow \, \supp(\mu) \textrm{ is an up-set,}
440: \end{equation}
441: and
442: \begin{equation} \label{eqnSfin2}
443: {\rm (III)} \, \Leftrightarrow \, {\rm (IV)} \Leftrightarrow \, \mu(\sigma(S)\equiv 1)>0.
444: \end{equation}
445: Consequently, the properties in {\rm (\ref{eqnSfin1})} imply those in
446: {\rm (\ref{eqnSfin2})}
447: but not vice versa. Note in particular that if
448: we are in the full support case,
449: then {\rm (I)--(IV)} all hold.
450: \end{theorem}
451: Although the term {\em extractability} is our own, the concept does
452: have a history; in particular, there has been interest in finding
453: lower bounds
454: on $\epsilon$ for which $\epsilon$-extractability holds.
455: The question of uniform extractability has been
456: studied for the Ising model as well as other Markov random fields in
457: \cite{BD,HB,N}. Earlier, in \cite{G1,G2,G3},
458: a similar question was studied for Markov chains and
459: autoregressive processes.
460: Of related interest is the result in \cite{HB} that for Markov random fields,
461: uniform finite energy
462: implies uniform extractability.
463:
464: \section{Basic examples} \label{examples}
465:
466: Our first example in this section is a pair of measures which
467: is downwards but not upwards movable. Note first that if
468: $\nu^{(+,\epsilon)} \preceq \mu$,
469: then we must have
470: \[
471: \nu \preceq \mu
472: \]
473: as well as
474: \[
475: \pi_{\epsilon} \preceq \mu \, .
476: \]
477:
478: \begin{example} \label{ex-epsstab1}
479: {\rm
480: Take $\nu=\frac{1}{2} \pi_{q}+\frac{1}{2}\delta_{0}$ and $\mu=\frac{1}{2} \pi_{p}+\frac{1}{2} \delta_{0}$ where
481: $q<p$ and where $\delta_{0}$ is the measure
482: which puts probability 1 on the configuration of all zeros. Trivially
483: \[
484: \nu \preceq \mu.
485: \]
486: If $S$ is infinite, then obviously
487: $\mu$ cannot dominate a product measure with positive density.
488: Therefore there does not exist any $\epsilon>0$ such that
489: \[
490: \nu^{(+,\epsilon)} \preceq \mu.
491: \]
492: However,
493: \[
494: \mu^{(-,\epsilon)}=
495: {\textstyle
496: \frac{1}{2} \pi^{(-,\epsilon)}_{p}+\frac{1}{2} \delta^{(-,\epsilon)}_{0}
497: =\frac{1}{2} \pi_{p(1-\epsilon)}+\frac{1}{2} \delta_{0}} \, ,
498: \]
499: so if we take $\epsilon>0$ such that $p(1-\epsilon)>q$, we get that
500: \[
501: \nu \preceq \mu^{(-,\epsilon)}.
502: \]
503: Hence $(\nu,\mu)$ is downwards but not upwards movable.
504: } $\Cox$
505: \end{example}
506: Before presenting the next three examples, we recall a family of
507: stationary processes known as {\bf determinantal} processes,
508: introduced in Lyons and Steif \cite{LS}. These are
509: probability measures ${\mathbf P}^{f}$ on the Borel sets
510: of $\{0,1\}^{{\mathbb Z}}$ where $f:[0,1] \rightarrow [0,1]$
511: is a Lebesgue-measurable function (see \cite{LS}). For such
512: an $f,$ define
513: \begin{eqnarray} \label{defPf}
514: \lefteqn{{\bf P}^{f}[\sigma(e_{1})=\cdots=\sigma(e_{k})=1]}\\
515: & & :={\bf P}^{f}[\{\sigma \in \{0,1\}^{\mathbb Z}: \sigma(e_{1})=\cdots=\sigma(e_{k})=1\}]\nonumber \\
516: & & :=\textrm{det}[\hat{f}(e_{j}-e_{i})]_{1\leq,i,j\leq k}, \nonumber
517: \end{eqnarray}
518: where $e_{1},\ldots,e_{k}$ are distinct elements in ${\mathbb Z}$
519: and $k\geq 1.$ Here $\hat{f}$ denotes the
520: Fourier coefficients of $f,$ defined by
521: \[
522: \hat{f}(k):=\int_{0}^{1}f(x)e^{-i2\pi kx} dx.
523: \]
524: In \cite{LS} it is proven that ${\bf P}^{f}$ is
525: indeed a probability measure.
526: (The fact that a probability measure is determined by the values it gives
527: to cylinder sets of this type follows immediately from
528: inclusion-exclusion.) In fact they showed this for the
529: more general case of $f:{\mathbb T}^{d} \rightarrow [0,1]$ where
530: ${\mathbb T}^{d}:={\mathbb R}^{d}/{\mathbb Z}^{d}$; in this case
531: the resulting process is indexed by $ {\mathbb Z}^{d}$.
532: This result rests very strongly on the
533: results in \cite{Lyons2002}. We will also need the following definition,
534: where GM stands for geometric mean:
535: \[
536: \GM(f):=\exp{\int_{0}^{1} \log f(x) dx}.
537: \]
538:
539: \begin{example} \label{ex-epsstab2}
540: {\rm
541: Let $f$ be a function from $[0,1]$ to itself.
542: By \cite[Theorem 5.3]{LS},
543: $\pi_{p} \preceq {\mathbf P}^{f}$ iff $p\leq \GM(f)$.
544: It is easy to see from (\ref{defPf})
545: that
546: $({\mathbf P}^{f})^{(-,\epsilon)}
547: = {\mathbf P}^{(1-\epsilon)f}$. Since
548: $\GM((1-\epsilon)f)=(1-\epsilon)\GM(f)$,
549: we see that when $p>0$ and $\pi_{p} \preceq {\mathbf P}^{f},$
550: $(\pi_{p},{\mathbf P}^{f})$ is downwards movable iff
551: it is upwards movable iff $p < \GM(f)$. This implies in particular that
552: ${\mathbf P}^{f}$ is rigid iff $\GM(f)>0$.
553: } $\Cox$
554: \end{example}
555: The following example is a variant of the one in \cite[Remark 5.4]{LS}.
556: \begin{example} \label{ex-epsstab3}
557: {\rm
558: By \cite[Lemma 2.7]{LS},
559: ${\mathbf P}^{f} \preceq {\mathbf P}^{g}$ if $f\leq g.$
560: Let $I_{A}$ denote the indicator
561: function of some set $A \subseteq [0,1]$ which has
562: Lebesgue-measure $1-\delta$ for some $\delta>0.$
563: Let $f=1/2 I_{A}$ and $g=3/4 I_{A}.$ There
564: exists $\epsilon>0$ such that $f\leq g(1-\epsilon) \leq g,$
565: and so we get that
566: ${\mathbf P}^{f} \preceq {\mathbf P}^{g(1-\epsilon)} \preceq {\mathbf P}^{g}.$
567: Hence $({\mathbf P}^{f},{\mathbf P}^{g})$
568: is downwards movable. However $\GM(g)=0$ which implies
569: that ${\mathbf P}^{g}$ does not dominate any product
570: measure with positive density.
571: Therefore $({\mathbf P}^{f},{\mathbf P}^{g})$ is not
572: upwards movable.
573: } $\Cox$
574: \end{example}
575:
576: \noindent
577: For our next example we need the definition of harmonic mean (HM):
578: \[
579: HM(f):=\left(\int_0^1 \frac{dx}{f(x)} \right)^{-1}.
580: \]
581: \begin{example} \label{ex-epsstab5}
582: {\rm Let $f(x)=x.$ It is easy to see that $HM(f)=0$ and $GM(f)=1/e>0.$
583: Since $GM(f)>0$, ${\mathbf P}^{f}$ is rigid. On the other hand,
584: since $HM(f)=0,$ Theorem 5.16 of \cite{LS} shows that ${\mathbf P}^{f}$ is not uniformly
585: insertion tolerant.
586: } $\Cox$
587: \end{example}
588:
589: \section{Uniform insertion tolerance and upwards \\ extractability}
590: \label{sect:(II)vs(I)}
591:
592: In this section we focus on uniform
593: upwards extractability (property (I)) and
594: uniform insertion tolerance (property (II)). Proposition
595: \ref{prop:(II)vs(I)_finite} provides an equivalence between these
596: properties when $S$ is finite, while Theorem \ref{th:Hajek} exhibits
597: a contrasting scenario for $S$ countable.
598: %,
599: %and show that for the case of finite $S$,
600: %(I), (II) and $\supp(\mu)$ is an up-set are equivalent,
601: %while (II) $\not\Rightarrow$ (I) in the more general countable case,
602: %including obtaining a translation invariant such example.
603: \begin{proposition} \label{prop:(II)vs(I)_finite}
604: If $S$ is finite and $\mu$ is a probability measure
605: on $\{0,1\}^S$, then the following are equivalent:
606: \begin{description}
607: \item{\rm (i)} uniform insertion tolerance,
608: \item{\rm (ii)} uniform
609: upwards extractability, and
610: \item{\rm (iii)} $\supp(\mu)$ is an up-set.
611: \end{description}
612: \end{proposition}
613: \begin{theorem} \label{th:Hajek}
614: For $S$ countably infinite, there exists a probability measure $\mu$ on
615: $\{0,1\}^S$ that is uniformly insertion tolerant but not uniformly upwards extractable.
616: Moreover, we can take $\mu$ to be a translation invariant measure on $\{0,1\}^{\mathbb Z}$.
617: \end{theorem}
618: {\bf Proof of Proposition \ref{prop:(II)vs(I)_finite}.}
619: If $\mu$ is uniformly insertion tolerant, then it is immediate that
620: $\supp(\mu)$ is an up-set. Furthermore uniform upwards
621: extractability trivially implies
622: uniform insertion tolerance as we have said previously.
623: We are therefore only left with having to
624: show that if $\supp(\mu)$ is an up-set,
625: then $\mu$ is uniformly upwards extractable.
626:
627: In what follows, given a configuration $\sigma\in\{0,1\}^S$, $|\sigma|$ will be the number of 1's in $\sigma$.
628: If there is to exist a $\nu$ such that $\mu=\nu^{(+,\epsilon)}$ with $\epsilon\in[0,1)$, it is not hard
629: to see that we must have
630: \begin{equation} \label{eqnnuformel}
631: \nu(\sigma)=\sum_{\tilde{\sigma} \preceq \sigma} (-\epsilon)^{|\sigma|-|\tilde{\sigma}|}
632: (1-\epsilon)^{|\tilde{\sigma}|-|S|} \mu(\tilde{\sigma}) \ \forall \sigma \in \{0,1\}^S.
633: \end{equation}
634: This can be verified through a direct calculation, but it is easier to calculate
635: $\nu^{(+,\epsilon)}(\sigma)$ and check that it is indeed
636: equal to $\mu(\sigma)$, as follows.
637: \begin{eqnarray*}
638: \lefteqn{\nu^{(+,\epsilon)}(\sigma)} \\
639: & & = \sum_{\sigma_1 \preceq \sigma} \epsilon^{|\sigma|-|\sigma_1|}(1-\epsilon)^{|S|-|\sigma|}
640: \nu(\sigma_1) \\
641: & & =\sum_{\sigma_1 \preceq \sigma} \epsilon^{|\sigma|-|\sigma_1|}(1-\epsilon)^{|S|-|\sigma|}
642: \sum_{\sigma_2 \preceq \sigma_1}(-\epsilon)^{|\sigma_1|-|\sigma_2|}
643: (1-\epsilon)^{|\sigma_2|-|S|} \mu(\sigma_2) \\
644: & & =\sum_{\sigma_1 \preceq \sigma} \sum_{\sigma_2 \preceq \sigma_1} \epsilon^{|\sigma|-|\sigma_1|}
645: (-\epsilon)^{|\sigma_1|-|\sigma_2|}(1-\epsilon)^{|\sigma_2|-|\sigma|} \mu(\sigma_2) \\
646: & & = \sum_{\sigma_2}(1-\epsilon)^{|\sigma_2|-|\sigma|} \mu(\sigma_2)
647: \sum_{\sigma_1:\sigma_2 \preceq \sigma_1 \preceq \sigma}\epsilon^{|\sigma|-|\sigma_1|}(-\epsilon)^{|\sigma_1|-|\sigma_2|}.
648: \end{eqnarray*}
649:
650: If we fix $\sigma_2$, then the binomial theorem gives that the last summation is equal to 0
651: unless $\sigma_2=\sigma$ in which case it is equal to 1.
652: We therefore easily obtain that $\nu^{(+,\epsilon)}(\sigma)=\mu(\sigma)$ for every $\sigma.$
653:
654: What remains is to check that $\nu(\sigma)\geq 0$ for all $\sigma.$
655: From (\ref{eqnnuformel}) it is immediate that $\nu(\sigma)=0$ for
656: every $\sigma \not \in \supp(\mu)$ since $\supp(\mu)$ is an up-set. For $\sigma\in\supp(\mu)$ on the
657: other hand, it is easy to see that if we do this construction for
658: different $\epsilon$'s, then we get
659: \[
660: \lim_{\epsilon \rightarrow 0} \nu(\sigma) \, = \, \mu(\sigma) \, .
661: \]
662: Since $\mu(\sigma)>0$ for all $\sigma \in \supp(\mu)$ and $|S| <\infty$, for $\epsilon>0$
663: small enough, we get that $\nu(\sigma)>0$ for all $\sigma \in \supp(\mu).$ This
664: shows that $\mu$ is $\epsilon$-upwards extractable for all such $\epsilon$.
665: $\Cox$
666:
667: \medskip\noindent
668: {\bf Proof of Theorem \ref{th:Hajek}.}
669: Let $S= \cup_{k=2}^\infty S_k$, where
670: \[
671: S_k= ((k,1), (k,2), \ldots, (k,k)) \, .
672: \]
673: We will take the probability measure
674: $\mu$ on $\{0,1\}^S$ to be the product measure
675: \begin{equation} \label{eq:product_measure}
676: \mu \, = \, \mu_2 \times \mu_3 \times \cdots
677: \end{equation}
678: where each $\mu_k$ is a probability measure on $\{0,1\}^{S_k}$. The
679: $\mu_k$'s are constructed as follows, drawing heavily on an example
680: of Hajek and Berger \cite{HB}. For $\sigma\in\{0,1\}^{S_k}$, let
681: \begin{equation} \label{eq:Hajek's_example}
682: \mu_k(\sigma) = \left\{
683: \begin{array}{ll}
684: \frac{4}{3} 2^{-k} & \mbox{if the number of $1$'s in $\sigma$ is even} \\
685: \frac{2}{3} 2^{-k} & \mbox{if the number of $1$'s in $\sigma$ is odd.}
686: \end{array} \right.
687: \end{equation}
688: We may think of $\mu_k$ as the distribution of a $\{0,1\}^{S_k}$-valued random
689: variable $X_k$ obtained by first tossing a biased coin with
690: heads-probability $\frac{2}{3}$, and if heads pick the components of
691: $X_k$ i.i.d.\ $(\frac{1}{2}, \frac{1}{2})$ conditioned on an even number
692: of $1$'s, while if tails pick the components i.i.d.\
693: $(\frac{1}{2}, \frac{1}{2})$ conditioned on an odd number of $1$'s.
694: One can also check that this distribution is the same as choosing all but
695: (an arbitrary) one of the variables according to $\pi_{1/2}$ and then taking the last
696: variable to be 1 with probability $1/3$ ($2/3$) if there are an even (odd) number of
697: 1's in the other bits. This last description immediately implies that
698: $\mu_k$ is $\frac{1}{3}$-insertion tolerant. Because
699: of the product structure in (\ref{eq:product_measure}), this property is
700: inherited by $\mu$, which therefore is uniformly
701: insertion tolerant.
702:
703: It remains to show that $\mu$ is not uniformly upwards extractable. To this end,
704: let $X$ be a $\{0,1\}^S$-valued random variable with distribution $\mu$,
705: and for $k=2,3,\ldots$ let $Y_k$ denote the number of $1$'s
706: in $X(S_k)$. It is immediate from (\ref{eq:Hajek's_example}) that
707: \begin{equation} \label{eq:enhanced_prob_of_even}
708: \P(Y_k \mbox{ is even}) = {\textstyle \frac{2}{3}}
709: \end{equation}
710: for each $k$. Using our last description of $\mu_k$, the weak law of large numbers
711: implies that
712: \[
713: \frac{Y_k}{k} \rightarrow {\textstyle \frac{1}{2}} \mbox{ in probability as $k\to\infty$.}
714: \]
715: Hence, in particular,
716: \begin{equation} \label{eq:substantial_no_of_0s}
717: \lim_{k \rightarrow \infty} \P(Y_k \leq k - m) \, = \, 1
718: \end{equation}
719: for any fixed $m$.
720:
721: Now assume (for contradiction) that $\mu = \nu^{(+, \epsilon)}$ for some
722: fixed $\epsilon >0$;
723: since $\mu$ being $\epsilon_2$-upwards extractable implies it is
724: $\epsilon_1$-upwards extractable for $\epsilon_1 <\epsilon_2$, we may
725: without loss of generality assume that $\epsilon \leq 1/3.$
726: Pick $X'$ according to $\nu$; we may then suppose that
727: $X$ has been obtained from $X'$ by randomly switching $0$'s to $1$'s
728: independently with probability $\epsilon$. The intuition behind the argument leading up
729: to a contradiction is that the process of independently flipping $0$'s to $1$'s will cancel
730: all preferences of ending up with an even number of $1$'s.
731:
732:
733: If $X'(S_k)$ contains precisely $l$ $0$'s, then the conditional probability
734: (given $X'$) that an even number of these switch to
735: $1$'s when going from $X'$ to $X$ is easily seen to equal
736: \[
737: {\textstyle
738: \frac{1}{2} + \frac{1}{2}(1- 2 \epsilon)^l \, .
739: }
740: \]
741: The easiest way to see this is using an equivalent random mechanism where each 0
742: independently ``updates'' with probability $2\epsilon$ and then all the sites
743: which have updated then independently actually switch to a 1 with
744: probability $1/2$.
745: It follows that the conditional probability (again given $X'$) that $Y_k$ is odd
746: is at least
747: \[
748: {\textstyle
749: \min \{ \frac{1}{2} + \frac{1}{2}(1- 2 \epsilon)^l,
750: \frac{1}{2} - \frac{1}{2}(1- 2 \epsilon)^l \} \, = \,
751: \frac{1}{2} - \frac{1}{2}(1- 2 \epsilon)^l.
752: }
753: \]
754: Now pick $m$ large enough so that
755: $\frac{1}{2} - \frac{1}{2}(1- 2 \epsilon)^m > \frac{5}{12}$. Since
756: $X' \preceq X$ a.s., we get from (\ref{eq:substantial_no_of_0s}) that
757: \[
758: \lim_{k \rightarrow \infty} \P(A_k) \, = \, 1
759: \]
760: where $A_k$ is the event that there are at least $m$ $0$'s in
761: $X'(S_k)$. This gives
762: \begin{eqnarray*}
763: \lim_{k \rightarrow \infty} \P(Y_k \mbox{ is odd})
764: & \geq & \lim_{k \rightarrow \infty} \P(Y_k \mbox{ is odd} \, | \,
765: A_k) \P(A_k) \\
766: & \geq & ({\textstyle \frac{1}{2} - \frac{1}{2}(1- 2 \epsilon)^m})
767: \lim_{k \rightarrow \infty} \P(A_k) \\
768: & > & {\textstyle \frac{5}{12}} \, .
769: \end{eqnarray*}
770: This clearly contradicts (\ref{eq:enhanced_prob_of_even}).
771:
772: We now translate this example into the setting of translation
773: invariant distributions on $\{0,1\}^{\mathbb Z}$.
774:
775: Begin with randomly designating either all even integers or all odd integers
776: (each with probability $\frac{1}{2}$) in the index set ${\mathbb Z}$
777: to represent copies of $S_2$. Assume that we happened to choose the
778: even integers (the other case is handled analogously). Then we
779: toss another fair coin to decide whether to put i.i.d.\ copies of
780: $X(S_2)$ on the pairs $\{\ldots, (-4, -2), (0,2), (4,6), \ldots\}$
781: in ${\mathbb Z}$,
782: or on $\{\ldots (-2,0), (2,4), (6,8), \ldots\}$. Then use one more fair coin
783: to decide whether $\{\ldots -3, 1, 5, 9, \ldots\}$ or
784: $\{\ldots, -1, 3, 7, 11, \ldots\}$ should be designated for i.i.d.\ copies
785: of $X(S_3)$, and once this is decided toss a fair three-sided coin to
786: choose one of the three possible placements of the length-$3$ blocks in
787: this subsequence to put these copies. And so on.
788:
789: This makes the resulting process $X^*$ translation invariant. Also, since
790: the property of $\epsilon$-insertion tolerance is obviously closed under
791: convex combinations, we easily obtain that $X^*$ is $\frac{1}{3}$-insertion
792: tolerant
793: and therefore uniformly insertion tolerant.
794:
795: Furthermore, for any $k\geq 2$,
796: we may apply (\ref{eq:enhanced_prob_of_even}) to the i.i.d.\ copies
797: of $X(S_k)$ to deduce that with probability $1$ there will
798: exist $i \in \{0,1, \ldots, k2^{k-1}-1\}$ such that
799: \begin{equation} \label{eq:2/3_limit}
800: \lim_{n \rightarrow \infty} \frac{1}{n}
801: \sum_{j=1}^{n} {\bf 1}_{B_{i,j,k}} \, = \,
802: {\textstyle \frac{2}{3}}
803: \end{equation}
804: where $B_{i,j,k}$ denotes the event that the number of $1$'s in
805: \[
806: \{i+ jk2^{k-1}, i+ jk2^{k-1} + 2^{k-1}, i+ jk2^{k-1} + 2 \cdot 2^{k-1},
807: \ldots, i+ jk2^{k-1} + (k-1)2^{k-1}\}
808: \]
809: is even. The right way to think of $i$ is that it is the first place
810: to the right of
811: the origin where a copy of $X(S_k)$ starts. The summation variable $j$
812: on the other hand,
813: makes us jump to the starting
814: points of all the other copies of $X(S_k)$ to the right of the origin.
815: Furthermore,
816: by arguing as in for the non-translation invariant construction,
817: we have that if $X^*$ is uniformly upwards extractable,
818: then for large $k$ the limit in (\ref{eq:2/3_limit}) will be less than
819: $1 - \frac{5}{12} = \frac{7}{12}$ for all $i \in \{0,1, \ldots, k2^{k-1}-1\}$.
820: But this contradicts (\ref{eq:2/3_limit}), so we can conclude that
821: $X^*$ is not uniformly upwards extractable.
822: $\Cox$
823:
824: \medskip\noindent
825: Note, finally, that the examples in the above proof also show that uniform
826: finite energy does not imply uniform extractability.
827:
828: \section{Rigidity} \label{sect:slipperiness}
829:
830: We now proceed to discuss the issue of when a measure is rigid.
831: As mentioned in the introduction, any measure
832: which does not dominate a nontrivial product measure is trivially
833: nonrigid and so it would be more interesting to
834: have a nonrigid measure which dominates
835: a nontrivial product measure; such a measure is provided in
836: Theorem \ref{extype2_2} below.
837:
838: \begin{proposition} \label{prop:rigid_when_finite}
839: If $S$ is finite and $\mu$ is a probability measure
840: on $\{0,1\}^S$, then the following are equivalent.
841: \begin{description}
842: \item{\rm (i)} $\mu$ dominates $\pi_p$ for some $p>0$,
843: \item{ \rm (ii)}
844: $\mu$ is rigid, and
845: \item{\rm (iii)} $\mu(\sigma(S)\equiv1)>0$.
846: \end{description}
847: \end{proposition}
848: This does not extend to infinite $S$, as shown in the following result.
849:
850: \begin{theorem} \label{extype2_2}
851: For $S$ countably infinite,
852: there exists a $\mu$ which dominates a nontrivial product measure $\pi_p$
853: but is nevertheless nonrigid.
854: Moreover, we can take $\mu$ to be a translation invariant measure
855: on $\{0,1\}^{\mathbb Z}$.
856: \end{theorem}
857: {\bf Proof of Proposition \ref{prop:rigid_when_finite}.}
858: It is easy to see that the condition that $\mu$ dominates
859: $\pi_p$ for some $p>0$
860: is equivalent to the condition that $\mu(\sigma(S)\equiv 1)>0.$
861: Also, recall that if $\mu$ is rigid it must dominate a
862: non-trivial product measure.
863:
864: To make the proof complete, it only remains to show that (i) and (iii)
865: of the statement imply that $\mu$ is rigid.
866: We have $\pi_{p_{\max, \mu}} \preceq \mu$, so that
867: \begin{equation} \label{eq:good_ol_dom}
868: \pi_{p_{\max, \mu}}(A) \, \leq \, \mu(A)
869: \end{equation}
870: for all increasing events $A \subseteq \{0,1\}^S$. We next claim that
871: \begin{equation} \label{eq:ineq_in_fact_an_equality}
872: \exists A\neq\emptyset,\{0,1\}^S \textrm{ such that $A$ is increasing and } \pi_{p_{\max, \mu}}(A)= \mu(A).
873: \end{equation}
874: To see this, note that if we had strict inequality
875: in (\ref{eq:good_ol_dom}) for all such nontrivial increasing events $A$, then
876: we could find a sufficiently small $\delta>0$ so that
877: \[
878: \pi_{p_{\max, \mu}+ \delta}(A) \, < \, \mu(A)
879: \]
880: for all such $A$ (this uses the finiteness of $S$), contradicting
881: the definition of $p_{\max, \mu}$. Now, for such an $A$ we have that
882: $\mu(A)\geq\mu(\sigma(S)\equiv 1)>0$ and hence for any $\epsilon>0$
883: \[
884: \mu^{(-, \epsilon)}(A) \, < \, \mu(A)
885: \]
886: (again because $S$ is finite), which in combination with
887: (\ref{eq:ineq_in_fact_an_equality}) yields
888: \[
889: \pi_{p_{\max, \mu}} \, \not\preceq \, \mu^{(-, \epsilon)} \, .
890: \]
891: Since $\epsilon>0$ was arbitrary, $\mu$ is rigid.
892: $\Cox$
893:
894: \medskip\noindent
895: For the proof of Theorem \ref{extype2_2}, the following elementary
896: lemma (which is presumably known) is convenient
897: to have.
898: \begin{lemma} \label{lem:conditioned_binomial}
899: For $k \geq 1$, $p \in (0,1)$ and $m \in \{0,1, \ldots, k\}$, write
900: $\rho_{k,p,m}$ for the distribution of a Binomial$(k,p)$ random variable
901: conditioned on taking value at least $m$. For $p_1 \leq p_2$, we have
902: \[
903: \rho_{k,p_1,m} \, \preceq \, \rho_{k,p_2,m} \, .
904: \]
905: \end{lemma}
906: {\bf Proof.}
907: For $i=1,2$, let $Y_i$ be a Bin$(k, p_i)$ random variable, and let
908: $X_i$ be a random variable with distribution $\rho_{k, p_i, m}$.
909: What we need to show is that for any $n\in\{m+1, \ldots, k\}$ we have
910: \[
911: \frac{\P(X_1 \geq n)}{\P(X_1 < n)} \, \leq \,
912: \frac{\P(X_2 \geq n)}{\P(X_2 < n)}
913: \]
914: which is the same as showing that
915: \begin{equation} \label{eq:need_to_show}
916: \frac{\P(X_2 \geq n)}{\P(X_1 \geq n)} \cdot \frac{\P(X_1 < n)}{\P(X_2 < n)}
917: \, \geq \, 1 \, .
918: \end{equation}
919: Writing $Z_1$ and $Z_2$ for the probabilities that $Y_1 \geq m$ and
920: $Y_2 \geq m$, respectively, the left-hand side of (\ref{eq:need_to_show})
921: becomes
922: \begin{equation} \label{eq:first_rewrite}
923: \frac{\frac{1}{Z_2} \sum_{j=n}^k {k \choose j}
924: p_2^j(1-p_2)^{k-j}}{\frac{1}{Z_1}
925: \sum_{j=n}^k {k \choose j} p_1^j(1-p_1)^{k-j}} \cdot
926: \frac{\frac{1}{Z_1}
927: \sum_{j=m}^{n-1} {k \choose j} p_1^j(1-p_1)^{k-j}}{\frac{1}{Z_2}
928: \sum_{j=m}^{n-1} {k \choose j}
929: p_2^j(1-p_2)^{k-j}} \, .
930: \end{equation}
931: Cancelling the $Z_i$'s and introducing the notation
932: $\phi_i=\frac{p_i}{1-p_i}$ for $i=1,2$, the expression in
933: (\ref{eq:first_rewrite}) may further be rewritten as
934: \begin{eqnarray} \nonumber
935: \lefteqn{ \mbox{ } \hspace{-20mm}
936: \frac{p_2^n(1-p_2)^{k-n} \sum_{j=n}^k {k \choose j}
937: \phi_2^{j-n}}{p_1^n(1-p_1)^{k-n} \sum_{j=n}^k {k \choose j} \phi_1^{j-n}}
938: \cdot
939: \frac{p_1^n(1-p_1)^{k-n} \sum_{j=m}^{n-1} {k \choose j}
940: \phi_1^{j-n}}{p_2^n(1-p_2)^{k-n} \sum_{j=m}^{n-1} {k \choose j}
941: \phi_2^{j-n}} = } \\
942: & = &
943: \frac{ \sum_{j=n}^k {k \choose j}
944: \phi_2^{j-n}}{ \sum_{j=n}^k {k \choose j} \phi_1^{j-n}}
945: \cdot
946: \frac{ \sum_{j=m}^{n-1} {k \choose j}
947: \phi_1^{j-n}}{ \sum_{j=m}^{n-1} {k \choose j}
948: \phi_2^{j-n}} \, .
949: \label{eq:second_rewrite}
950: \end{eqnarray}
951: Note now that $\phi_1 \leq \phi_2$, so that
952: \[
953: \sum_{j=n}^k {k \choose j}
954: \phi_2^{j-n} \, \geq \, \sum_{j=n}^k {k \choose j} \phi_1^{j-n}
955: \]
956: and
957: \[
958: \sum_{j=m}^{n-1} {k \choose j}
959: \phi_1^{j-n} \, \geq \, \sum_{j=m}^{n-1} {k \choose j}
960: \phi_2^{j-n} \, .
961: \]
962: Hence, the expression in (\ref{eq:second_rewrite}) is greater than or equal
963: to $1$, so (\ref{eq:need_to_show}) is verified and the lemma is
964: established. $\Cox$
965:
966: \medskip\noindent
967: {\bf Proof of Theorem \ref{extype2_2}.}
968: As in the proof of Theorem \ref{th:Hajek}, we take
969: $S= \cup_{k=2}^\infty S_k$ where
970: $S_k= ((k,1), (k,2), \ldots, (k,k))$, and
971: the probability measure
972: $\mu$ on $\{0,1\}^S$ to be the product measure
973: \[
974: \mu \, = \, \mu_2 \times \mu_3 \times \cdots
975: \]
976: where each $\mu_k$ is a probability measure on $\{0,1\}^{S_k}$. This
977: time, we take the $\mu_k$'s to be as follows. For $\sigma \in \{0,1\}^{S_k}$,
978: set
979: \begin{equation} \label{eq_def_of_blocks}
980: \mu_k(\sigma) = \left\{
981: \begin{array}{ll}
982: k^{-1}2^{-k} & \mbox{if the number of $1$'s in $\sigma$ is exactly $1$}, \\
983: 1 - 2^{-k} & \mbox{if } \sigma=(1,1,1,\ldots, 1), \\
984: 0 & \mbox{otherwise}.
985: \end{array} \right.
986: \end{equation}
987: We now make three claims about the $\mu_k$ measures:
988: \begin{description}
989: \item{\sc Claim 1.} $p_{\max, \mu_k} \geq \frac{1}{2}$ for all $k$.
990: \item{\sc Claim 2.} $\lim_{k \rightarrow\infty}p_{\max, \mu_k}= \frac{1}{2}$.
991: \item{\sc Claim 3.} For any fixed
992: $\epsilon < \frac{1}{2}$, we have for all $k$ sufficiently large that
993: \[
994: \mu_k^{(-,\epsilon)} \succeq \pi_{\frac{1}{2}}
995: \]
996: where $\pi_{\frac{1}{2}}$ is product measure with $p=\frac{1}{2}$ on
997: $\{0,1\}^{S_k}$.
998: \end{description}
999: We slightly postpone proving the claims, and first show how they
1000: imply the existence of a nonrigid measure that dominates $\pi_{\frac{1}{2}}$.
1001:
1002: Let us modify $S$ and $\mu$ slightly by setting, for $m \geq 2$,
1003: \[
1004: \tilde{S}_m \, = \, \cup_{k=m}^\infty S_k
1005: \]
1006: and
1007: \begin{equation} \label{eq:projected_product_structure}
1008: \tilde{\mu}_m \, = \, \mu_m \times \mu_{m+1} \times \cdots
1009: \end{equation}
1010: so that in other words $\tilde{\mu}_m$ is the probability measure on
1011: $\{0,1\}^{\tilde{S}_m}$ which arises by projecting $\mu$ on
1012: $\{0,1\}^{\tilde{S}_m}$.
1013:
1014: Using the product structure (\ref{eq:projected_product_structure}), we
1015: get from {\sc Claim 1} that $p_{\max, \tilde{\mu}_m} \geq \frac{1}{2}$
1016: (for any $m$), and from {\sc Claim 2} that
1017: $p_{\max, \tilde{\mu}_m} \leq \frac{1}{2}$ (for any $m$). Hence
1018: \[
1019: p_{\max, \tilde{\mu}_m} \, = \, {\textstyle \frac{1}{2} }
1020: \]
1021: for any $m$. Fixing $\epsilon \in (0,1/2)$, we can also deduce from
1022: (\ref{eq:projected_product_structure}) and {\sc Claim 3} that
1023: \begin{equation} \label{eq:punchline_giving_nonrigidity}
1024: \tilde{\mu}_m^{(-, \epsilon)} \, \succeq \, \pi_{\frac{1}{2}}
1025: \, = \, \pi_{p_{\max, \tilde{\mu}_m}}
1026: \end{equation}
1027: for $m$ sufficiently large. For such $m$ we thus have that $\tilde{\mu}_m$
1028: is nonrigid.
1029:
1030: It remains to prove {\sc Claim 1}, {\sc Claim 2} and {\sc Claim 3}.
1031:
1032: {\sc Claim 1} is the same as saying that $\mu_k \succeq
1033: \pi_{\frac{1}{2}}$. This
1034: is immediate to verify, but the best way to think about it is as follows.
1035: Suppose that we pick $X_k \in \{0,1\}^{S_k}$ according to
1036: $\pi_{\frac{1}{2}}$,
1037: and if $X_k= (0,0, \ldots, 0)$ then we switch one of the $0$'s (chosen
1038: uniformly at random) to a $1$, while otherwise we switch {\em all} $0$'s
1039: to $1$'s. The resulting random element of $\{0,1\}^{S_k}$ then
1040: has distribution $\mu_k$.
1041:
1042: To prove {\sc Claim 2}, it suffices (in view of {\sc Claim 1}) to prove that
1043: \[
1044: \limsup_{k \rightarrow\infty}p_{\max, \mu_k}\le \frac{1}{2}
1045: \]
1046: and to this end it is enough to show for any $\delta>0$ that
1047: \begin{equation} \label{eq:doesnt_dom_1/2+delta}
1048: \mu_k \, \not\succeq \pi_{\frac{1}{2}+\delta}
1049: \end{equation}
1050: for all sufficiently large $k$. Let $A_k$ denote the event of seeing
1051: at most one $1$ in $\{0,1\}^{S_k}$; then $A_k$ is a decreasing event and
1052: its complement $\neg A_k$ is increasing. Now simply note that
1053: \begin{equation} \label{eq:first_ratio_of_probs}
1054: \frac{\mu_k(A_k)}{\pi_{\frac{1}{2}+\delta}(A_k)} \, = \,
1055: \frac{(\frac{1}{2})^k}{(\frac{1}{2}- \delta)^k +
1056: k (\frac{1}{2}+ \delta)(\frac{1}{2}- \delta)^{k-1}}
1057: \end{equation}
1058: which tends to $\infty$ as $k \rightarrow \infty$. Hence, taking $k$
1059: large enough gives $\mu_k(A_k) > \pi_{\frac{1}{2}+\delta}(A_k)$, so that
1060: $\mu_k(\neg A_k) < \pi_{\frac{1}{2}+\delta}(\neg A_k)$ and
1061: (\ref{eq:doesnt_dom_1/2+delta}) is established, proving {\sc Claim 2}.
1062:
1063: To prove {\sc Claim 3}, note first that both $\pi_{\frac{1}{2}}$ and
1064: $\mu_k^{(-, \epsilon)}$ are invariant under permutations of $S_k$,
1065: so that it suffices to show for $k$ large that
1066: \begin{equation} \label{eq:consider_only_Bn}
1067: \mu_k^{(-, \epsilon)} (B_n) \, \leq \, \pi_{\frac{1}{2}} (B_n)
1068: \end{equation}
1069: for $n=0,1, \ldots, k-1$, where $B_n$ is the event of seeing at most
1070: $n$ $1$'s in $S_k$. For $n=0$ we get
1071: \begin{equation} \label{eq:second_ratio_of_probs}
1072: \frac{\mu_k^{(-, \epsilon)} (B_0)}{\pi_{\frac{1}{2}} (B_0)} \, = \,
1073: \frac{(\frac{1}{2})^k \epsilon + (1- (\frac{1}{2})^k )
1074: \epsilon^k}{(\frac{1}{2})^k}
1075: \end{equation}
1076: while for $n=1$
1077: \begin{equation} \label{eq:third_ratio_of_probs}
1078: \frac{\mu_k^{(-, \epsilon)} (B_1)}{\pi_{\frac{1}{2}} (B_1)} \, = \,
1079: \frac{ (\frac{1}{2})^k + (1- (\frac{1}{2})^k )
1080: (\epsilon^k + k \epsilon^{k-1}(1-\epsilon))}{(k+1)(\frac{1}{2})^k } \, .
1081: \end{equation}
1082: The right-hand sides of (\ref{eq:second_ratio_of_probs}) and
1083: (\ref{eq:third_ratio_of_probs})
1084: tend to $\epsilon$ and $0$, respectively,
1085: as $k \rightarrow \infty$, so (\ref{eq:consider_only_Bn})
1086: is verified for $n=0$ and $1$ (and $k$ large enough).
1087: To verify (\ref{eq:consider_only_Bn}) for $n\geq 2$ (and all such $k$),
1088: define two random variables $Y$ and $Y'$ as the number of $1$'s in
1089: two random elements of $\{0,1\}^{S_k}$ with respective distributions
1090: $\mu_k^{(-, \epsilon)}$ and $\pi_{\frac{1}{2}}$. Note that $Y$ conditioned
1091: on taking value at least $2$ has the same distribution as a
1092: Bin $(k, 1- \epsilon)$ random variable conditional on taking value
1093: at least $2$, while the conditional distribution of $Y'$ given that it
1094: is at least $2$, is that of a Bin $(k, \frac{1}{2})$ variable conditioned
1095: on being at least $2$. Defining $\rho_{k, (1-\epsilon), 2}$ and
1096: $\rho_{k, \frac{1}{2}, 2}$ as in Lemma \ref{lem:conditioned_binomial},
1097: we thus have for $n \in \{2, \ldots, k-1\}$ that
1098: \begin{equation} \label{eq:first_compl_prob}
1099: \mu_k^{(-, \epsilon)} (B_n) \, = \, 1 - (1-\mu_k^{(-, \epsilon)} (B_1))
1100: (1- \rho_{k, (1-\epsilon), 2}(B_n))
1101: \end{equation}
1102: and
1103: \begin{equation} \label{eq:second_compl_prob}
1104: \pi_{\frac{1}{2}} (B_n) \, = \, 1 - (1-\pi_{\frac{1}{2}} (B_1))
1105: (1- \rho_{k, \frac{1}{2}, 2}(B_n)) \, .
1106: \end{equation}
1107: But we have already seen that
1108: $\mu_k^{(-, \epsilon)} (B_1) \leq \pi_{\frac{1}{2}} (B_1)$, and Lemma
1109: \ref{lem:conditioned_binomial} tells us that
1110: $\rho_{k, (1-\epsilon), 2}(B_n) \leq \rho_{k, \frac{1}{2}, 2}(B_n)$, so
1111: (\ref{eq:first_compl_prob}) and (\ref{eq:second_compl_prob}) yield
1112: \[
1113: \mu_k^{(-, \epsilon)} (B_n) \, \leq \, \pi_{\frac{1}{2}} (B_n) \, ,
1114: \]
1115: and {\sc Claim 3} is established.
1116:
1117: Finally, we translate this example into the setting of translation
1118: invariant distributions on $\{0,1\}^{\mathbb Z}$.
1119: The measure $\tilde{\mu}_m$
1120: can be turned into a translation invariant measure $\tilde{\mu}^*_m$
1121: on $\{0,1\}^{\mathbb Z}$
1122: by the same independent-copy-and-paste procedure as in
1123: Theorem \ref{th:Hajek}. The property
1124: \[
1125: \pi_{\frac{1}{2}} \, \preceq \, (\tilde{\mu}^*_m)^{(-, \epsilon)}
1126: \]
1127: is obviously inherited from (\ref{eq:punchline_giving_nonrigidity}). Thus,
1128: in order to show that $\tilde{\mu}^*_m$ is nonrigid, it only remains to
1129: show that it does not stochastically dominate $\pi_{\frac{1}{2}+ \delta}$
1130: for any $\delta >0$. This follows using (\ref{eq:first_ratio_of_probs})
1131: by an argument analogous to (\ref{eq:2/3_limit}) in
1132: Theorem \ref{th:Hajek}: If we pick $k$ depending on $\delta$ as in
1133: the justification of {\sc Claim 2}, then, under $\tilde{\mu}^*_m$,
1134: certain infinite arithmetic progressions will have subsequences
1135: of length $k$ which contain at most one $1$ often
1136: enough (under spatial averaging) that the corresponding event
1137: has $\pi_{\frac{1}{2}+ \delta}$-measure $0$. We omit the details.
1138: $\Cox$
1139:
1140: \begin{remark}
1141: \label{rem:neither(II)or(III)}
1142: {\rm
1143: The measure $\tilde{\mu}_m$ is obviously not uniformly insertion
1144: tolerant, and we have thus demonstrated the existence of a measure for which
1145: property (IV) holds while neither (II) nor (III) does.
1146: }
1147: $\Cox$
1148: \end{remark}
1149: \begin{remark}
1150: {\rm For any $p \in (0,1)$, the construction above can be modified
1151: by replacing $2^{-k}$ by $p^k$ in (\ref{eq_def_of_blocks}). Proceeding
1152: as in the rest of the proof yields the result that for any
1153: $p, \epsilon\in (0,1)$ such that $p+\epsilon<1$, there exists a measure
1154: $\mu$ on $\{0,1\}^S$ where $S$ is countably infinite, with
1155: the property that
1156: $p_{\max, \mu} = p$ and
1157: \[
1158: \pi_{p_{\max, \mu}} \preceq \mu^{(-, \epsilon)} \, .
1159: \]
1160: This is obviously sharp.
1161: $\Cox$
1162: }
1163: \end{remark}
1164:
1165: \section{Further results on rigidity} \label{secFKG}
1166:
1167: In this section, we continue the study of rigidity, and prove
1168: Theorem \ref{thmFKG} and Proposition \ref{prop:neg}.
1169:
1170:
1171:
1172:
1173:
1174: \medskip\noindent
1175: The proof of Theorem \ref{thmFKG} will make use of the following
1176: technical lemma.
1177: \begin{lemma} \label{lemmaFKG}
1178: Let $\mu$ be a measure on $\{0,1\}^{{\mathbb Z}^{d}}.$ Assume that it is
1179: $\delta$-insertion tolerant for some $\delta >0$. If for some $p\in (0,1)$ and $\epsilon >0$
1180: \begin{equation} \label{eqnassump}
1181: \mu^{(-,\epsilon)}(\sigma(\{1,\ldots,n\}^d)\equiv 0)\leq (1-p)^{n^d} \textrm{ for all } n\geq 0,
1182: \end{equation}
1183: then there exists $p'>p$ such that
1184: \[
1185: \mu(\sigma(\{1,\ldots,n\}^d)\equiv 0)\leq (1-p')^{n^d} \textrm{ for all } n\geq 0.
1186: \]
1187: \end{lemma}
1188: {\bf Proof.}
1189: Let $X\sim \mu$ and $Z\sim \pi_{1-\epsilon}$ be independent and let $X^{(-,\epsilon)}=
1190: \min(X,Z).$ It is easy to see using the $\delta$-insertion tolerance that for any $s \in \{1,\ldots,n\}^d,$
1191: and any $\zeta \in \{0,1\}^{\{1,\ldots,n\}^d \setminus s}$
1192: \begin{eqnarray*}
1193: \lefteqn{{\mathbb P}(X(s)=1 \cap X(\{1,\ldots,n\}^d \setminus s)\equiv \zeta)} \\
1194: & & \geq \frac{\delta}{1-\delta}{\mathbb P}(X(s)=0 \cap X(\{1,\ldots,n\}^d \setminus s)\equiv \zeta).
1195: \end{eqnarray*}
1196: Iterating this, we get that for any $\xi \in \{0,1\}^{\{1,\ldots,n\}^d}$
1197: \[
1198: {\mathbb P}(X(\{1,\ldots,n\}^d)\equiv \xi) \geq \left ( \frac{\delta}{1-\delta} \right ) ^{|\xi|}
1199: {\mathbb P}(X(\{1,\ldots,n\}^d)\equiv 0).
1200: \]
1201: Here $|\xi|$ denotes the cardinality of the set $\{s\in \{1,\ldots,n\}^d:\xi(s)=1\}.$
1202: We get that
1203: \begin{eqnarray*}
1204: \lefteqn{{\mathbb P}(X^{(-,\epsilon)}(\{1,\ldots,n\}^d)\equiv 0)} \\
1205: &=& \sum_{\xi \in \{0,1\}^{\{1,\ldots,n\}^d}}{\mathbb P}
1206: (X^{(-,\epsilon)}(\{1,\ldots,n\}^d)\equiv 0|X(\{1,\ldots,n\}^d)\equiv \xi)\\
1207: & & \times {\mathbb P}(X(\{1,\ldots,n\}^d)\equiv \xi) \\
1208: &\geq & \sum_{\xi \in \{0,1\}^{\{1,\ldots,n\}^d}}{\mathbb P}
1209: (X^{(-,\epsilon)}(\{1,\ldots,n\}^d)\equiv 0|X(\{1,\ldots,n\}^d)\equiv \xi) \\
1210: & & \times \left ( \frac{\delta}{1-\delta} \right ) ^{|\xi|}{\mathbb P}
1211: (X(\{1,\ldots,n\}^d)\equiv 0) \\
1212: &=& \sum_{\xi \in \{0,1\}^{\{1,\ldots,n\}^d}}\epsilon^{|\xi|}
1213: \left ( \frac{\delta}{1-\delta} \right ) ^{|\xi|}{\mathbb P}
1214: (X(\{1,\ldots,n\}^d)\equiv 0) \\
1215: &=& \left (1+ \frac{\epsilon \delta}{1-\delta} \right )^{n^d}{\mathbb P}
1216: (X(\{1,\ldots,n\}^d)\equiv 0).
1217: \end{eqnarray*}
1218: Therefore if (\ref{eqnassump}) holds we can conclude that
1219: \[
1220: \mu(\sigma(\{1,\ldots,n\}^d)\equiv 0)\leq \left (\frac{1-\delta}{1-\delta+\epsilon \delta} \right)^{n^d}
1221: (1-p)^{n^d} \, ,
1222: \]
1223: and we are done.
1224: $\Cox$
1225:
1226: \medskip \noindent
1227: {\bf Proof of Theorem \ref{thmFKG}.}
1228: The case $p_{\max,\mu}=1$ is trivial and we therefore assume that $p_{\max,\mu}\in(0,1).$
1229: In \cite{LigS}, it is shown that if $\mu$ is downwards FKG and if
1230: \begin{equation} \label{liggett}
1231: \mu(\sigma(\{1,\ldots,n\}^d)\equiv 0) \leq (1-p)^{n^d} \textrm{ for all } n\geq 0,
1232: \end{equation}
1233: then $\pi_p \preceq \mu.$ Therefore if $\pi_{p_{\max,\mu}} \preceq \mu^{(-,\epsilon)}$ for some
1234: $\epsilon >0$, then (\ref{eqnassump}) trivially holds (with $p=p_{\max,\mu}$)
1235: and so we can conclude from Lemma~\ref{lemmaFKG} and the above result in \cite{LigS}
1236: that $\pi_{p'} \preceq \mu$ for some $p'> p_{\max,\mu}$, a contradiction.
1237: $\Cox$
1238:
1239:
1240: \medskip \noindent
1241: We now define conditional negative association.
1242:
1243: \begin{definition}
1244: A probability measure $\mu$ on $\{0,1\}^{\mathbb Z}$ is said to have
1245: {\bf conditional negative association} if for any finite $S \subset {\mathbb Z}$
1246: and any two increasing functions $f,g$ that are measurable on disjoint
1247: subsets of ${\mathbb Z}\setminus S$,
1248: \[
1249: \mu(fg|\sigma(S)) \leq \mu(f|\sigma(S))\mu(g|\sigma(S)).
1250: \]
1251: \end{definition}
1252: We will use the fact (see \cite{LS}) that for conditionally
1253: negatively associated measures $\mu$, we have
1254: $\pi_{\rho} \preceq \mu$ iff
1255: \begin{equation} \label{eqncNA}
1256: \mu(\sigma(\{1,\ldots,n\})\equiv 1)\geq {\rho}^n \ \forall \ n\geq 1.
1257: \end{equation}
1258:
1259: \noindent
1260: {\bf Proof of Proposition \ref{prop:neg}.}
1261: We note that the case $p_{\max,\mu}=1$ is trivial and we
1262: therefore assume that $p_{\max,\mu}\in(0,1).$ Assume that
1263: $\pi_{p_{\max,\mu}} \preceq \mu^{(-,\epsilon)}$ for some $\epsilon > 0$.
1264: We then get that
1265: \[
1266: p_{\max,\mu}^n \leq \mu^{(-,\epsilon)}(\sigma(\{1,\ldots,n\})\equiv 1)=(1-\epsilon)^n\mu(\sigma(\{1,\ldots,n\})\equiv 1).
1267: \]
1268: Hence
1269: \[
1270: \mu(\sigma(\{1,\ldots,n\})\equiv 1) \geq
1271: \left( \frac{p_{\max,\mu}}{1-\epsilon} \right)^n.
1272: \]
1273: Therefore $\pi_{\frac{p_{\max,\mu}}{1-\epsilon}} \preceq \mu$ by the
1274: result quoted in connection with (\ref{eqncNA}).
1275: This is a contradiction since $p_{\max,\mu} < \frac{p_{\max,\mu}}{1-\epsilon}$.
1276: $\Cox$
1277:
1278:
1279:
1280:
1281:
1282:
1283:
1284:
1285:
1286:
1287:
1288:
1289:
1290:
1291:
1292:
1293:
1294:
1295:
1296:
1297:
1298:
1299:
1300:
1301:
1302:
1303:
1304:
1305:
1306:
1307:
1308:
1309:
1310:
1311: \section{Proof of main result} \label{types}
1312:
1313: \begin{lemma} \label{lemmapos}
1314: If $\mu$ is uniform upwards extractable, then
1315: for any $\epsilon>0$ there exists a
1316: $\delta>0$ such that $(\mu^{(-,\epsilon)})^{(+,\delta)} \preceq \mu.$
1317: \end{lemma}
1318: {\bf Proof.} Let $\nu$ and $\alpha>0$ be such that $\mu=\nu^{(+,\alpha)}.$
1319: One can easily compute that for any $\alpha$,
1320: $\epsilon$, and $\delta$, we have that
1321: \[
1322: ((\mu^{(+,\alpha)})^{(-,\epsilon)})^{(+,\delta)}
1323: =\mu^{(-,\epsilon(1-\delta),+,\alpha(1-\epsilon)+\alpha \epsilon \delta+(1-\alpha)\delta)}.
1324: \]
1325: Now, given $\epsilon>0,$ choose $\delta>0$ such that
1326: $\alpha(1-\epsilon)+\alpha \epsilon \delta+(1-\alpha)\delta<\alpha.$
1327: We therefore get that
1328: \[
1329: (\mu^{(-,\epsilon)})^{(+,\delta)}=((\nu^{(+,\alpha)})^{(-,\epsilon)})^{(+,\delta)}
1330: \preceq \nu^{(-,\epsilon(1-\delta),+,\alpha)} \preceq \nu^{(+,\alpha)}=\mu.
1331: \]
1332: $\Cox$
1333:
1334: \begin{lemma} \label{lemmatype1}
1335: Given a probability measure $\mu$ on $\{0,1\}^{S},$ assume that for
1336: every $\epsilon>0$, there exists a $\delta>0$ such that
1337: $(\mu^{(-,\epsilon)})^{(+,\delta)} \preceq \mu$. Then $\mu$ is rigid.
1338: \end{lemma}
1339: {\bf Proof.}
1340: The case $p_{\max,\mu}=1$ is trivial, and we will therefore assume
1341: that $p_{\max,\mu}\in (0,1)$. Assume for contradiction that $\mu$ is
1342: nonrigid. Then there exists an $\epsilon>0$ such that
1343: $\pi_{p_{\max,\mu}} \preceq \mu^{(-,\epsilon)}$. By assumption there
1344: exists a $\delta>0$ such that
1345: $(\mu^{(-,\epsilon)})^{(+,\delta)} \preceq \mu$.
1346: Hence $(\pi_{p_{\max,\mu}})^{(+,\delta)} \preceq
1347: (\mu^{(-,\epsilon)})^{(+,\delta)} \preceq \mu$.
1348: Since $p_{\max,\mu}<1,$ $(\pi_{p_{\max,\mu}})^{(+,\delta)}$ is
1349: a product measure with density strictly larger than $p_{\max,\mu}.$
1350: This is a contradiction.
1351: $\Cox$
1352:
1353: \medskip\noindent
1354: We remark that we do not know whether the reverse statement of
1355: Lemma \ref{lemmatype1} is true. It would also be interesting to know if
1356: the sufficient condition in this lemma follows from uniform insertion
1357: tolerance.
1358:
1359: Example \ref{ex-epsstab5} provides us with an example of a $\mu$ which is
1360: on one hand rigid but on the other hand not uniformly
1361: insertion tolerant. However, since it relies heavily on results not presented
1362: in this paper, we give here another more ``hands on'' example.
1363: It is a variant of \cite[Remark 6.4]{LS} and
1364: shows that the reverse statement of Lemma \ref{lemmapos} is false.
1365:
1366: \begin{example} \label{ex-epsstab4}
1367: {\rm
1368: Let $\{X_{i}\}_{i \in {\mathbb N}}$ be defined in the following way.
1369: For every even $i \geq 0$, let independently $(X_{i},X_{i+1})$ be $(1,1)$
1370: or $(0,0)$ with probability $1/2$ each. Let $\mu_e$ denote the
1371: distribution of this process. For $\epsilon, \delta>0$ let
1372: $\{X_{i}^{(-,\epsilon(1-\delta),+,\delta)}\}_{i \in {\mathbb N}}$
1373: be a sequence of random variables with distribution
1374: $\mu_e^{(-,\epsilon(1-\delta),+,\delta)}=(\mu_e^{(-,\epsilon)})^{(+,\delta)}$.
1375: By noting that for any $\epsilon>0$ there exists a $\delta>0$ such that for even $i$
1376: \[
1377: {\mathbb P}(\max(X_{i}^{(-,\epsilon(1-\delta),+,\delta)},
1378: X_{i+1}^{(-,\epsilon(1-\delta),+,\delta)})=1)<{\textstyle \frac{1}{2}},
1379: \]
1380: we see that for the same choice of $\epsilon,\delta$ we get that
1381: $(\mu_e^{(-,\epsilon)})^{(+,\delta)} \preceq \mu_e$. Lemma \ref{lemmatype1}
1382: gives us that $\mu_e$ is rigid.
1383: However, it is easy to see that $\mu_e$ is not uniform insertion tolerant.
1384:
1385: The only drawback with this construction is that it is not translation
1386: invariant. However this is easily fixed. Let $\mu_o$ be the distribution
1387: of $\{X_{i+1}\}_{i \in {\mathbb N}},$ i.e. it is $\mu_e$ shifted over by 1.
1388: Define the measure $\mu$ by
1389: \[
1390: {\textstyle \mu=\frac{1}{2}\mu_e+\frac{1}{2}\mu_o. }
1391: \]
1392: It is easy to check that
1393: \[
1394: {\textstyle (\mu^{(-,\epsilon)})^{(+,\delta)}=\frac{1}{2}(\mu_e^{(-,\epsilon)})^{(+,\delta)}+\frac{1}{2}(\mu_o^{(-,\epsilon)})^{(+,\delta)}
1395: \preceq \frac{1}{2}\mu_e+\frac{1}{2}\mu_o=\mu. }
1396: \]
1397: By Lemma \ref{lemmatype1}, it follows that $\mu$ is rigid.
1398: On the other hand, clearly
1399: $$
1400: \mu(\sigma(0)=1|\sigma(1)=0,\sigma(2)=\sigma(3)=1)=0
1401: $$
1402: and hence $\mu$ is not uniformly insertion tolerant.
1403: } $\Cox$
1404: \end{example}
1405:
1406:
1407: \noindent
1408: {\bf Proof of Theorem \ref{thm1}.}
1409: Lemma \ref{lemmapos} together with Lemma \ref{lemmatype1}
1410: shows that property (I) implies property (III) and all the other
1411: implications were indicated in the introduction. As far as all of the
1412: reversed implications claimed not to hold, we continue as follows.
1413: Example \ref{ex-epsstab4} together with Lemma \ref{lemmatype1}
1414: (or example \ref{ex-epsstab5})
1415: shows that (III) does not imply (II) (and hence that
1416: (III) does not imply (I) and that (IV) does not imply (II)).
1417: Theorem \ref{extype2_2} implies that (IV) does not imply (III).
1418: Finally, Theorem \ref{th:Hajek} shows that (II) does not imply (I).
1419: Also, all of these examples were translation invariant measures on
1420: $\{0,1\}^{{\mathbb Z}}.$
1421: $\Cox$
1422:
1423: \medskip
1424: \noindent
1425: {\bf Proof of Theorem \ref{thm:finite}.}
1426: This follows immediately from Propositions \ref{prop:(II)vs(I)_finite}
1427: and \ref{prop:rigid_when_finite}.
1428: $\Cox$
1429:
1430: \medskip
1431: \noindent
1432: We feel, finally, that it is worth mentioning the following result,
1433: which is an easy consequence of Lemma \ref{lemmapos}.
1434:
1435: \begin{corollary}
1436: Assume that $(\mu_1,\mu_2)$ is downwards movable and that $\mu_2$
1437: is uniformly upwards extractable. Then
1438: $(\mu_1,\mu_2)$ is also upwards movable.
1439: \end{corollary}
1440:
1441: \begin{thebibliography}{99}
1442:
1443: \bibitem{BD} Bassalygo, L.A. and Dobrushin, R.L. (1987)
1444: Epsilon-entropy of a Gibbs field, (Russian)
1445: {\em Problemy Peredachi Informatsii} {\bf 23}, 3--15.
1446:
1447: \bibitem{BHK}
1448: van den Berg, J., H\"aggstr\"om O. and Kahn, J. (2004)
1449: Some conditional correlation inequalities for percolation
1450: and related processes, {\em Random Structures Algorithms}, to appear.
1451:
1452: \bibitem{BS} Broman, E.I. and Steif, J.E. (2004)
1453: Dynamical stability of percolation for some interacting
1454: particle systems and $\epsilon$-movability, {\em Ann. Probab.}, to appear.
1455:
1456: \bibitem{GHM} Georgii, H.-O., H\"aggstr\"om, O. and Maes, C. (2001)
1457: The random geometry of equilibrium phases,
1458: {\em Phase Transitions and Critical Phenomena, Volume 18}
1459: (C. Domb and J.L. Lebowitz, eds), pp 1--142, Academic Press, London.
1460:
1461: \bibitem{G1} Gray, R.M. (1970)
1462: Informations rates of autoregressive processes,
1463: {\em IEEE Trans. Information Theory}
1464: {\bf IT-16}, 412--421.
1465:
1466: \bibitem{G2} Gray, R.M. (1971)
1467: Rate distortion functions for finite-state finite-alphabet Markov sources,
1468: {\em IEEE Trans. Information Theory} {\bf IT-17}, 127--134.
1469:
1470: \bibitem{G3} Gray, R.M. (1973)
1471: Correction to ``Information rates of stationary-ergodic
1472: finite-alphabet sources'',
1473: % (IEEE Trans. Information Theory IT-17 (1971), 516--523).
1474: {\em IEEE Trans. Information Theory}
1475: {\bf IT-19}, 573.
1476:
1477: \bibitem{HB} Hajek, B. and Berger, T. (1987) A decomposition theorem
1478: for binary Markov random fields, {\em Ann. Probab.} {\bf 15},
1479: 1112--1125.
1480:
1481: \bibitem{IPS} Liggett, T.M. (1985)
1482: {\it Interacting Particle Systems}
1483: Springer, New York.
1484:
1485: \bibitem{SIS} Liggett, T.M. (1999)
1486: {\em Stochastic Interacting Systems: Contact, Voter and Exclusion Processes},
1487: Springer, New York.
1488:
1489: \bibitem{L1} Liggett, T.M. (1994)
1490: Survival and coexistence in interacting particle systems,
1491: {\em Probability and Phase Transition} (ed. G. Grimmett),
1492: Kluwer, Dodrecht, pp 209--226.
1493:
1494: \bibitem{LigS} Liggett, T.M. and Steif, J.E. (2004)
1495: Stochastic domination: the contact process, Ising models,
1496: FKG measures and exchangeable processes, preprint.
1497:
1498: \bibitem{Lyons2002} Lyons, R. (2003)
1499: Determinantal probability measures,
1500: {\em Publ. Math. Inst. Hautes Etudes Sci.}, {\bf 98}, 167--212.
1501:
1502: \bibitem{LSch} Lyons, R. and Schramm, O. (1999)
1503: Indistinguishability of percolation clusters,
1504: {\em Ann. Probab.} {\bf 27}, 1809--1836.
1505:
1506: \bibitem{LS} Lyons, R. and Steif, J.E. (2003)
1507: Stationary determinantal processes: phase multiplicity, Bernoullicity,
1508: entropy, and domination,
1509: {\em Duke Math. Journal}, {\bf 120}, 515--575.
1510:
1511: \bibitem{N} Newman, C.M. (1987)
1512: Decomposition of binary random fields and zeros of partition functions,
1513: {\em Ann. Probab.} {\bf 15}, 1126--1130.
1514:
1515: \bibitem{NS} Newman, C.M. and Schulman, L.S. (1981)
1516: Infinite clusters in percolation models,
1517: {\em J. Stat. Phys.} {\bf 26}, 613--628.
1518:
1519: \end{thebibliography}
1520:
1521:
1522:
1523:
1524:
1525: \end{document}
1526:
1527:
1528:
1529:
1530:
1531: