math0204324/dyn.tex
1: \def\versiondate{23 Jan.\ 2003}
2: \input math.macros
3: %\input Ref.macros
4: \input lrlEPSfig.macros
5: 
6: %\proofmodetrue
7: %\leftsectionheadtrue
8: \checkdefinedreferencetrue
9: %\continuousnumberingtrue
10: \continuousfigurenumberingtrue
11: \theoremcountingtrue
12: \sectionnumberstrue
13: %\figuresectionnumberstrue
14: \forwardreferencetrue
15: %\lefteqnumberstrue
16: \citationgenerationtrue
17: \nobracketcittrue
18: \tocgenerationtrue
19: \hyperstrue
20: \initialeqmacro
21: 
22: \bibsty{myapalike}
23: %\input\jobname.lbl 
24: 
25: \def\T{{\Bbb T}}
26: \def\Td{{\T^d}}
27: \def\dbar{{\overline{d}}}
28: \def\Pfp{{(\P^f)^+}}
29: \def\Pfm{{(\P^f)^-}}
30: \def\calA{{\cal A}}
31: \def\calT{{\cal T}}
32: \def\F{{\cal F}}
33: \def\U{{\cal U}}
34: \def\C{{\Bbb C}}
35: \def\hE{\ell^2(E)}
36: \def\Hs{{\ell^2(E)}}
37: \def\hD{\ell^2({\Z}^d)}
38: \def\id{{\rm id}}  %% The identity operator.
39: \def\ev#1{{\cal #1}}
40: \def\A{{\ev A}}
41: \def\ip#1{(\changecomma #1)}
42: \def\bigip#1{\bigl(\bigchangecomma #1\bigr)}
43: \def\biggip#1{\biggl(\bigchangecomma #1\biggr)}
44: \def\leftip#1{\left(\leftchangecomma #1\right)}  
45: \def\changecomma#1,{#1,\,}
46: \def\bigchangecomma#1,{#1,\;}
47: \def\leftchangecomma#1,{#1,\ }
48: \def\Var{{\rm Var}}
49: \def\bfz{{\bf 0}}
50: \def\bfo{{\bf 1}}
51: \def\GM{{\ss GM}}  %% geometric mean
52: \def\mult#1{M_{\times #1}} %% multiplication of the variable
53: \def\strle{\preccurlyeq_{\rm s}}   %% strongly dominated by
54: %\def\frac#1#2{{#1 \over #2}}  % to work with Mathematica TeXForm
55: %\def\imag{i}  % for TeXForm
56: \def\fav{\widehat f(0)}  % the average of $f$
57: \def\past{{\ss Past}}
58: \def\cat{{\rm G}}  % Catalan's constant
59: \def\others{{\cal Z}}  % the $\sigma$-field generated by non-zero sites
60: \def\HM{{\ss HM}}  % harmonic mean
61: \def\fullle{\preccurlyeq_{\rm f}}   %% fully dominated by
62: \def\Zd{{\Z^d}}
63: \def\ee{{\bf e}}  % exponential
64: \def\frac#1#2{{#1 \over #2}}
65: \def\nonfrac#1#2{{#1 / #2}}
66: \def\hl{{\ss HL}}  % Helson-Lowdenslager space
67: \def\supp{{\rm supp}\,}
68: \def\Cov{{\rm Cov}}
69: \def\ord{\Pi}  % order
70: 
71: \def\BLPSusf{Benjamini, Lyons, Peres, and Schramm (2000)%
72: \def\BLPSusf{BLPS (2000)}}
73: 
74: 
75: 
76: \def\firstheader{\eightpoint\ss\underbar{\raise2pt\line 
77:     {
78:     To appear in {\it Duke Math. J.}
79:     \hfil Version of \versiondate}}}
80: 
81: \beginniceheadline
82: 
83: %\ifproofmode \relax \else\head{} {Version of \versiondate}\fi 
84: \vglue20pt
85: 
86: \title{Stationary Determinantal Processes:}
87: \title{Phase Multiplicity, Bernoullicity,}
88: \title{Entropy, and Domination}
89: 
90: \author{Russell Lyons and Jeffrey E. Steif}
91: 
92: 
93: 
94: 
95: 
96: \abstract{%
97: We study a class of stationary processes indexed by $\Z^d$ that
98: are defined via minors of $d$-dimensional (multilevel) Toeplitz matrices.
99: We obtain necessary and sufficient conditions for
100: phase multiplicity (the existence of a phase transition)
101: analogous to that which occurs in 
102: statistical mechanics.
103: Phase uniqueness is equivalent to the presence of a strong
104: $K$ property, a particular strengthening of the usual $K$ 
105: (Kolmogorov) property. 
106: We show that all of these processes are
107: Bernoulli shifts (isomorphic to i.i.d.\ processes in the sense of ergodic
108: theory).
109: We obtain estimates of their entropies and
110: we relate these processes via stochastic domination to product measures.
111: }
112: 
113: 
114: \bottomII{Primary 
115: 82B26, %phase transitions (general)
116: 28D05, %measure-preserving transformations
117: 60G10. %Stationary processes
118: Secondary
119: 82B20, %lattice systems (Ising, dimer, Potts, etc.) and systems on graphs
120: 37A05, %measure-preserving transformations
121: 37A25, %ergodicity, mixing, rates of mixing
122: 37A60, %dynamical systems in statistical mechanics
123: 60G25, %Prediction theory
124: 60G60, %Random fields
125: 60B15. %Probability measures on groups, Fourier transforms, factorization
126: }
127: {Random field, Kolmogorov mixing, Toeplitz determinants, negative
128: association, fermionic lattice model, stochastic domination, entropy,
129: Szeg\H{o} infimum, Fourier coefficients, insertion tolerance, deletion
130: tolerance, geometric mean.}
131: {Research partially supported by 
132: NSF grants DMS-9802663, DMS-0103897 (Lyons), and
133: DMS-0103841 (Steif), the Swedish Natural Science Research Council (Steif)
134: and the Erwin Schr\"odinger Institute for Mathematical
135: Physics, Vienna (Lyons and Steif).}
136: 
137: 
138: \articletoc
139: 
140: 
141: 
142: 
143: \bsection {Introduction}{s.intro}
144: 
145: Determinantal probability measures and point processes arise in numerous
146: settings, such as
147: mathematical physics (where they are called fermionic point processes),
148: random matrix theory, representation theory, and certain other areas
149: of probability theory. 
150: See \ref b.Sosh/ for a survey and \ref b.L:det/ for additional developments
151: in the discrete case.
152: We present here a detailed analysis of the discrete
153: stationary case.
154: After this paper was first written, we learned of independent but slightly
155: prior work of \ref b.ShiTak:II/, announced in \ref b.ShiTak:ann/.
156: We discuss the (small) overlap between their work and ours in the appropriate
157: places below.
158: See also \ref b.ShiTak:I/ and \ref b.ShiYoo:glauber/ for related
159: contemporaneous work.
160: 
161: Stationary determinantal processes are interesting from several viewpoints.
162: First, they have interesting relations with the 
163: theory of Toeplitz determinants.
164: As in that theory, the {\bf geometric mean} of a nonnegative function $f$,
165: defined as 
166: $$
167: \GM(f) := \exp \int \log f
168: \,,
169: $$
170: will play an important role in some of our results.
171: (In fact, the arithmetic mean and the harmonic mean of $f$ will also
172: characterize certain properties of our processes.)
173: Second, such processes arise in certain combinatorial models, such as uniform
174: spanning trees and dimer models.
175: Third, these systems have a rich infinite-dimensional parameter space,
176: consisting, in the case of a $\Z^d$ action, of all measurable functions $f$
177: from the $d$-dimensional torus to $[0, 1]$.
178: We illustrate the variety of possible behaviors with examples throughout
179: the paper.
180: Fourth, they have the unusual property of negative association.
181: Though unusual, negative association occurs in various places and a fair
182: amount is known about it
183: (see \ref b.Joag/, \ref b.Pemantle:negass/, 
184: \ref b.Newman:asympt/, \ref b.ShaoSu:LIL/, \ref b.Shao:comparison/,
185: \ref b.ZhangWen:weak/, \ref b.Zhang:Strassen/, and the references therein,
186: for example). We offer a whole new class of examples of negatively
187: associated stationary processes.
188: As such, determinantal processes provide an easy way to construct examples of
189: many kinds of behavior that might otherwise be difficult to construct, such as
190: negatively associated (stationary) processes with slow decay of correlations,
191: or with the even sites independent of the odd sites (in one dimension, say),
192: or with the property of being finitely dependent.
193: Fifth, all our processes are Bernoulli shifts, i.e., isomorphic to i.i.d.\
194: processes. This may be surprising in view of the fact that only measurability,
195: rather than smoothness, of the parameter $f$ is required.
196: Sixth, in one dimension,
197: some of the processes are strong $K$, while others are not.
198: Namely, strong $K$ is equivalent to $f(\bfo-f)$ having a positive
199: geometric mean.
200: Similarly, we characterize exactly, in all dimensions, which $f$ give strong
201: {\it full\/} $K$ systems. It turns out that not only the rate at which $f$
202: approaches 0 or 1 matters, but also where. For example, in two dimensions,
203: if $f$ is real analytic, then the system is strong full $K$ iff the
204: (possibly empty) sets $f^{-1}(0)$ and $f^{-1}(1)$ belong to nontrivial
205: algebraic varieties. The strong full $K$ property is 
206: analogous to phase uniqueness
207: in statistical physics, as
208: we explain in \ref s.sk/.
209: 
210: We now state our results somewhat more precisely and present several
211: examples.
212: Let $f : \Td \to [0, 1]$ be a Lebesgue-measurable function on the
213: $d$-dimensional torus $\Td := \R^d/\Z^d$.
214: Define a $\Z^d$-invariant probability measure $\P^f$ on the Borel sets of $ \{
215: 0, 1 \}^{\Z^d}$ by defining the probability of the cylinder sets 
216: \begineqalno
217: \P^f[\eta(e_1)=1,\dots, \eta(e_k)=1] 
218: &:= 
219: \P^f[\{ \eta \in \{ 0, 1 \}^{\Z^d} \st \eta(e_1)=1,\dots, \eta(e_k)=1\}]
220: \cr&:= 
221: \det [\widehat{f}(e_j-e_i)]_{1\le i, j \le k}
222: \cr
223: \endeqalno
224: for all $e_1,\ldots,e_k\in \Z^d$, where $\widehat f$ denotes the Fourier
225: coefficients of $f$.
226: We shall prove in \ref s.back/ that this does indeed define a probability
227: measure.
228: Note that when $d=1$ and when $e_1,\ldots,e_k$
229: are chosen to be $k$ consecutive integers, the
230: right-hand side above is the usual $k \times k$ Toeplitz determinant of $f$
231: denoted $D_{k-1}(f)$.
232: In particular, we have $\P^f\big[\eta(e) = 1\big] = \widehat f(\bfz)
233: = \int_{\Td} f$ for every $e \in \Z^d$.
234: 
235: \procl x.Bernoulli
236: As a simple example, if $f \equiv p$, then $\P^f$ is i.i.d.\ Bernoulli($p$)
237: measure.
238: \endprocl
239: 
240: 
241: \procl x.ustH
242: For a more interesting example, consider
243: $$
244: f(x, y) := {\sin^2 \pi x \over \sin^2 \pi x + \sin^2 \pi y}
245: \,.
246: \label e.ustH
247: $$
248: A portion of a sample of a configuration from $\P^f$ is shown in \ref
249: f.window-ustH/, where a square with upper left corner $(i, j)$ is colored
250: black iff $\eta(i, j) = 1$.
251: These correspond to the horizontal edges of a uniform spanning tree in the
252: square lattice. That is, if $T$ is a spanning tree of $\Z^2$, then $\eta(i,
253: j)$ is the indicator that the edge from $(i, j)$ to $(i+1, j)$ belongs to
254: $T$. The portion of the spanning tree from which \ref f.window-ustH/ was
255: constructed is shown in \ref f.window-ust/.
256: When $T$ is chosen ``uniformly" (see \ref b.Pemantle:ust/, \ref
257: b.Lyons:bird/, or \ref b.BLPSust/ for definitions and information on
258: this), then $\eta$ has the law $\P^f$. This follows from the Transfer
259: Current Theorem of \ref b.BurPem/ and the representation of the Green
260: function as an integral; see \ref b.Lyons:book/ for more details.
261: Similarly, the edges of the uniform spanning forest (it is a tree only for $d
262: \le 4$, as shown by \ref b.Pemantle:ust/) parallel to the $x_1$-axis in $d$
263: dimensions correspond to the
264: function 
265: $$
266: f(x_1, x_2, \ldots, x_d) := {\sin^2 \pi x_1 \over \sum_{j=1}^d \sin^2 \pi x_j}
267: \,.
268: \label e.ustH-d
269: $$
270: We remark that
271: the uniform spanning tree is the so-called random-cluster model when 
272: one takes the limit $q \downarrow 0$,
273: then $p \downarrow 0$, and finally the thermodynamic limit, the latter
274: shown to exist by \ref b.Pemantle:ust/.
275: \endprocl
276: 
277: \twoefiginslabel window-ustH {A sample from $\P^f$ of \ref e.ustH/.} x3 
278: window-ust {A uniform spanning tree.} y3
279: 
280: \procl x.ustHx
281: Let 
282: $$
283: g(x) :=
284: {\sin \pi x \over \sqrt{1+\sin^2 \pi x}}
285: \,.
286: $$
287: Then the edges of the uniform spanning tree in the plane that
288: lie on the $x$-axis have the law $\P^g$, as shown in \ref x.ustH_x-dom/ below.
289: \endprocl
290: 
291: 
292: \procl x.USTzz
293: Let 
294: $$
295: f(x) := {1 \over 2} + {|\sin 2\pi x| - 1 \over 2 \cos 2\pi x}
296: \,.
297: $$
298: An elementary calculation shows that 
299: $$
300: \widehat f(k) = \cases{1/2  &if $k=0$,\cr
301:                        0    &if $k \ne 0$ is even,\cr
302: \displaystyle
303: (-1)^{(k-1)/2}\left(-{1 \over 2} + {2 \over \pi} 
304:       \sum_{j=0}^{(k-1)/2} {(-1)^j \over 2j+1}\right) - {1 \over \pi k}
305:                             &if $k$ is odd.\cr}
306: $$
307: We shall show in \ref x.USTzz-dom/ that the measure $\P^f$ arises as follows.
308: Given a spanning tree $T$ of the square lattice, let $\eta(n)$ be the
309: indicator that $e_n \in T$, where $e_n$ is the edge 
310: $$
311: e_n := \cases{[(n/2, n/2), (n/2 + 1, n/2)]  &if $n$ is even,\cr
312:               [((n+1)/2, (n-1)/2), ((n+1)/2, (n+1)/2)]  &if $n$ is odd.\cr}
313: $$
314: The collection of edges $\{e_n \st n \in \Z \}$ is a zig-zag path in the
315: plane. The law of $\eta$ is $\P^f$ when $T$ is chosen as a uniform spanning
316: tree.
317: (Although it is not hard to see that the law of $\eta$ is $\Z$-invariant,
318: by using planar duality, and although the law of $\eta$ must be a
319: determinantal probability measure because the law of $T$ is,
320: it is not apparent {\it a priori\/} that the law of $\eta$ has the
321: form $\P^f$ for some $f$.)
322: \endprocl
323: 
324: \procl x.lozenge
325: Fix a horizontal edge of the hexagonal lattice (also known as the
326: honeycomb lattice) and index all its vertical translates by $\Z$.
327: If one considers the standard measure of maximal entropy on perfect
328: matchings of the hexagonal lattice, also called the dimer model
329: and equivalent to lozenge tilings of the plane, and
330: looks only at the edges indexed as above by $\Z$, then the law is $\P^f$
331: for $f := \I{[1/3, 2/3]}$, as shown by \ref b.Kenyon:local/.
332: \endprocl
333: 
334: \procl x.domino It is interesting that the function
335: $f := \I{[0, 1/2]}$ for $d = 1$ also arises from a combinatorial model.
336: In this case,
337: $$
338: \widehat f(n) = \cases{1/2 &if $n=0$,\cr
339:                   0 &if $n \ne 0$ is even,\cr
340:                   1/(\pi i n) &if $n$ is odd.\cr}
341: $$
342: The measure $\P^f$ is the zig-zag process of \ref b.Joh:noninter/
343: derived from uniform domino tilings in the plane.
344: For the definition of ``uniform" in this case, see \ref b.BurPem/.
345: A picture of a portion of such a tiling is shown in \ref f.window-tiling/.
346: Consider the squares on a diagonal from upper left to lower right.
347: The domino covering any such square also covers a second square.
348: If this second square is above the diagonal, then we color the original
349: square black, as shown in \ref f.window-diagprocess/.
350: \ref b.Joh:noninter/ showed that the law of this process is $\P^f$ when the
351: diagonal squares are indexed by $\Z$ in the natural way.
352: More generally, the processes $\P^f$ for $f$ the indicator of any arc of $\T$
353: are used by \ref b.BOO:plancherel/, Theorem 3, to describe the typical shape
354: of Young diagrams.
355: \endprocl
356: 
357: \twoefiginslabel window-tiling {A uniform domino tiling.} x3 
358: window-diagprocess {A sample from $\P^f$ of \ref x.domino/.} y3
359: 
360: 
361: \procl x.renewal Let $0 < a < 1$ and $d = 1$.
362: If $f(x) := (1 - a)^2/|e^{2\pi i x} - a|^2$, then $\P^f$ is a renewal
363: process (\ref B.Sosh/).
364: The number of 0s between successive 1s has the same distribution as
365: the number of tails until 2 heads appear for a coin that has probability $a$
366: of coming up tails.
367: More explicitly, for $n\ge 1$,
368: $$
369: \P^f[\eta(1) = \cdots = \eta(n-1) = 0, \eta(n) = 1 \mid \eta(0) = 1]
370: =
371: n (1-a)^2 a^{n-1}
372: \,.
373: \label e.renewal-distribution
374: $$
375: Since 
376: $$
377: f(x) = {1-a \over 1+a} \left( {a e^{2\pi i x} \over 1 - a e^{2\pi i x}} + {1
378: \over 1 - a e^{-2\pi i x}} \right)
379: \,,
380: $$
381: expansion in a geometric series shows that
382: $$
383: \widehat f(k) = {1-a \over 1+a} a^{|k|}
384: \,.
385: $$
386: We prove that $\P^f$ is indeed this explicit renewal process after we prove
387: \ref p.regenerate/, in which we extend this example to other regenerative
388: processes.
389: \endprocl
390: 
391: 
392: \procl x.perconf
393: If $0 < p < 1$ and $f : \Td \to [0, 1]$ is measurable, then a fair sample of
394: $\P^{pf}$ can be obtained from a fair sample of $\P^f$ simply by independently
395: changing each 1 to a 0 with probability $1-p$.
396: \endprocl
397: 
398: 
399: For general systems, note that the covariance of
400: $\eta(\bfz)$ and $\eta(k)$ for $k \in \Z^d$ is $-|\widehat f(k)|^2$. This is
401: summable in $k$ since $f \in L^2(\Td)$, but that is essentially the most one
402: can say for its rate of decay.
403: That is, given any $\Seq{a_k} \in \ell^1(\Z^d)$, there is some
404: (even continuous) $f : \Td \to [0, 1]$ and some constant $c > 0$ such that
405: $|\widehat f(k)|^2 \ge c |a_k|$ for all $k \in \Z^d$, as shown by \ref
406: b.LKK/.
407: Observe also that as is the case for Gaussian processes, the processes
408: studied here have the property that if the random variables are uncorrelated, then
409: they are mutually independent.
410: 
411: It is shown (in a much more general context) by \ref b.L:det/ that these
412: measures, as well as these measures conditioned on the values of $\eta$
413: restricted to any finite subset of $\Z^d$,
414: have the following negative association property:
415: If $A$ and $B$ are increasing events that are measurable with respect to the
416: values of $\eta$ on disjoint subsets of $\Z^d$, then $A$ and $B$ are
417: negatively correlated.
418: 
419: In the order presented in this paper, our principal results are the following.
420: \beginbullets
421: 
422: For all $f$, the process $\P^f$ is a Bernoulli shift. % (\ref t.Bern/).
423: This was shown by \ref b.ShiTak:II/ for those $f$ such that $\sum_{n \ge 1}
424: n |\widehat f(n)|^2 <\infty$ by showing that those $\P^f$ are weak Bernoulli.
425: 
426: For all $f$, the process $\P^f$ stochastically dominates product measure
427: $\P^{\GM(f)}$ and is stochastically dominated by product measure
428: $\P^{1-\GM(\bfo-f)}$, and these bounds are optimal.
429: This is rather unexpected for the process of \ref x.ustH/ related to the
430: uniform spanning tree; explicit calculations are given below in \ref
431: x.ustH-dom/ for this particular process.
432: We give similar optimal bounds for full domination (uniform insertion and
433: deletion tolerance) in terms of harmonic means.
434: 
435: We present methods to estimate the entropy of $\P^f$.
436: For example, we show that for the function $g$ in \ref x.ustHx/,
437: the entropy of the system $\P^g$ lies in the interval $[0.69005, 0.69013]$;
438: see \ref x.ustH-x-ent/.
439: 
440: The process $\P^f$ is strong full $K$ iff there is a nonzero
441: trigonometric polynomial $T$ such that 
442: ${|T|^2 \over f (\bfo-f)} \in L^1(\Td)$.
443: This is equivalent to phase uniqueness in the sense that
444: no conditioning at infinity can change the measure.
445: 
446: In one dimension, $\P^f$ is strong $K$ iff $f(\bfo-f)$ has a positive
447: geometric mean.
448: This is equivalent to phase uniqueness when conditioning on one
449: side only. Higher-dimensional versions of this will also be obtained.
450: \endbullets
451: 
452: We shall give full definitions as they become needed. 
453: Some general background on determinantal probability measures is presented
454: in \ref s.back/, where we also exhibit a key representation of certain
455: conditional probabilities as Szeg\H{o} infima.
456: The property of being a Bernoulli shift 
457: is proved in \ref s.bern/, while the auxiliary result
458: that all $\P^f$ have full support (except in two degenerate cases) is shown in
459: \ref s.fs/.
460: Properties concerning stochastic domination are proved in \ref
461: s.dom/. These are used to estimate entropy in \ref s.ent/.
462: More sophisticated methods
463: of estimating entropy are also developed and illustrated in \ref s.ent/.
464: The main result about phase multiplicity is proved in \ref s.sk/, while the
465: one-sided case for $d=1$ and its higher-dimensional generalizations are
466: treated in \ref s.1side/.
467: Finally, we end with some open questions in \ref s.open/.
468: 
469: The above definitions can be generalized to any countable abelian group with
470: the discrete topology. 
471: For example, if we use $\Z \times \Z_2$, 
472: we obtain nontrivial joinings of the above systems with themselves. That
473: is, suppose $f : \T \to [0, 1]$ and $h : \T \to [0, 1]$ are such that $f \pm
474: h : \T \to [0, 1]$. Then the function $f_h : \T \times \{ -1, 1 \}
475: \to [0, 1]$ defined by $(x, \epsilon) \mapsto f(x) + \epsilon h(x)$
476: gives a system that, when restricted to each copy of $\Z$, is just $\P^f$, but
477: has correlations between the two copies that are given by $h$
478: (the case $h = \bfz$ gives the independent joining). 
479: We can obtain slow decay of correlations between the two copies and
480: negative associations in the joining itself;
481: this is perhaps something that is not easy to construct directly.
482: 
483: 
484: \bsection {Background}{s.back}
485: 
486: We first quickly review the probability measures studied in 
487: \ref b.L:det/;
488: see that paper for complete details.
489: 
490: Let $E$ be a finite or countable set and consider the complex
491: Hilbert space $\hE$. 
492: Given any closed subspace $H \subseteq \ell^2(E)$,
493: let $P_H$ denote the orthogonal projection onto $H$.
494: There is a unique probability measure $\P^H$ on $2^E := \{ 0, 1 \}^E$ defined
495: by 
496: $$
497: \P^H[\eta(e_1) = 1, \ldots, \eta(e_k) = 1]
498: =
499: \det [(P_H e_i, e_j)]_{1\le i,j\le k}
500: \label e.(3)
501: $$
502: for all $k \ge 1$ and any set of distinct $e_1,\ldots,e_k\in E$; see, e.g.,
503: \ref b.L:det/ or \ref b.DVJ/, Exercises 5.4.7--5.4.8.
504: On the right-hand side, we are identifying each $e \in E$ with the element
505: of $\hE$ that is 1 in coordinate $e$ and 0 elsewhere. 
506: In case $H$ is finite-dimensional, then $\P^H$ is concentrated on subsets
507: of $E$ of cardinality equal to the dimension of $H$.
508: 
509: 
510: More generally,
511: let $Q$ be a positive contraction, meaning that $Q$ is a self-adjoint
512: operator on $\Hs$ such that for all $u \in \Hs$, we have $0 \le (Q u, u) \le
513: (u, u)$.
514: There is a unique probability
515: measure $\P^Q$ such that 
516: $$
517: \P^Q[\eta(e_1) = 1, \ldots, \eta(e_k) = 1]
518: = \det [(Q {e_i}, {e_j})]_{i, j \le
519: k}
520: \label e.Qcorrel
521: $$
522: for all $k \ge 1$ and distinct $e_1, \ldots, e_k \in E$.
523: When $Q$ is the orthogonal projection onto a closed
524: subspace $H$, then $\P^Q = \P^H$.
525: In fact, properties of $\P^Q$ can be deduced from the special case of
526: orthogonal projections. Since this will be useful for our analysis, we
527: review this reduction procedure.
528: 
529: Note first that uniqueness follows from the fact that \ref e.Qcorrel/
530: determines all finite-dimen\-sion\-al marginals via the inclusion-exclusion
531: theorem. Indeed, we have the following formula for any disjoint pair of
532: finite sets $A, B \subseteq E$ (see, e.g., \ref b.L:det/):
533: $$
534: \P^Q[\eta \restrict A \equiv 1,\, \eta \restrict B \equiv 0]
535: =
536: \det \left[\big(\I B(e)e + (-1)^{\I B(e)} Q e, e'\big)\right]_{e, e' \in A
537: \cup B}
538: \,.
539: \label e.incexc
540: $$
541: To show existence, let $P_H$ be any orthogonal projection that is
542: a dilation of $Q$, i.e., $H$ is
543: a closed subspace of $\ell^2(E')$ for some $E' \supseteq E$ and for all $u
544: \in \Hs$, we have $Qu = P_{\Hs} P_H u$, where we regard $\ell^2(E') = \Hs
545: \oplus \ell^2(E' \setminus E)$. (In this case, $Q$ is also called the
546: compression of $P_H$ to $\Hs$.)
547: The existence of a dilation is standard and is easily
548: constructed; see, e.g., \ref b.L:det/.
549: Having chosen a dilation, we simply define $\P^Q$ as the law of $\eta$
550: restricted to $E$ when $\eta$ has the law $\P^H$.
551: Then \ref e.Qcorrel/ is a special case of \ref e.(3)/.
552: 
553: A probability measure $\P$ on $2^E$ is said to have {\bf negative
554: associations} if for all pairs
555: $A$ and $B$ of increasing events that are measurable with respect to the
556: values of $\eta$ on disjoint subsets of $E$, we have that $A$ and $B$ are
557: negatively correlated with respect to $\P$.
558: The following {\bf conditional negative association} (CNA)
559: property is proved in \ref b.L:det/ and a consequence of this
560: (see \ref p.JNRD/) will be used frequently here.
561: 
562: \procl t.negass
563: If $Q$ is any positive contraction on $\ell^2(E)$, $A$ is a
564: finite subset of $E$, and $\eta_0 \in 2^A$,
565: then $\P^Q[\;\cdot \mid \eta \restrict A = \eta_0]$ has
566: negative associations.
567: \endprocl
568: 
569: We now assume that $E=\Z^d$. Then the group structure of $\Z^d$ allows
570: $\Z^d$ to act naturally on $\ell^2(\Z^d)$ and on $2^{\Z^d}$.
571: The proof of the following lemma is straightforward and 
572: therefore skipped.
573: 
574: \procl l.pm
575: If $Q$ is a $\Z^d$-invariant positive contraction on
576: $\ell^2(\Z^d)$, then $\P^Q$ is also $\Z^d$-invariant.
577: \endprocl
578: 
579: As is well known, there exists a complex Hilbert-space isomorphism
580: between $L^2(\T^d, \lambda_d)$ and $\ell^2(\Z^d)$ where
581: $\T^d$ is the $d$-dimensional torus $\R^d/\Z^d$ and
582: $\lambda_d$ is unit Lebesgue measure on $\T^d$. 
583: This isomorphism is given by the Fourier transform
584: $f\mapsto \widehat f$, where for $f \in L^2(\T^d, \lambda_d)$, we have
585: $\widehat{f}(k)=\int_{\T^d} f(x) e^{-2\pi i k\cdot x} \, d\lambda_d(x)$ 
586: for $k\in\Z^d$.
587: If $\ee_k$ denotes the function $x \mapsto e^{2\pi i k \cdot x}$, then the
588: isomorphism takes the set $ \{ \ee_k \st k \in \Zd \}$ to the standard basis
589: for $\ell^2(\Zd)$.
590: From now on, we shall abbreviate $L^2(\T^d, \lambda_d)$ by $L^2(\T^d)$.
591: 
592: The following is well known.
593: 
594: \procl t.rudin 
595: \beginitems
596: \itemrm{(i)} Let $A\subseteq\Td$ be measurable and consider the operator 
597: $T_A:L^2(\T^d)\to L^2(\T^d)$ given by
598: $$
599: T_A(g)=g \I A\,.
600: $$
601: Then these projections (as $A$ varies over the measurable subsets of $\Td$) 
602: correspond (via the Fourier
603: isomorphism) to the $\Z^d$-invariant projections on $\ell^2(\Z^d)$.
604: \itemrm{(ii)} More generally, let $f:\Td\to [0,1]$ be measurable and
605: consider the operator $M_f:L^2(\T^d)\to L^2(\T^d)$ given by
606: $$
607: M_f(g)=fg\,.
608: $$
609: Then these positive contractions (as $f$ varies
610: over the measurable functions from $\T^d$ to $[0,1]$) correspond (via the
611: Fourier
612: isomorphism) to the $\Z^d$-invariant positive contractions on $\ell^2(\Z^d)$.
613: More specifically, $M_f$ corresponds to convolution with $\widehat{f}$.
614: \enditems
615: \endprocl
616: 
617: As in the above theorem, an $f:\Td\to [0,1]$ yields a
618: $\Z^d$-invariant positive contraction $Q_f$ on
619: $\ell^2(\Z^d)$, which in turn yields a translation-invariant
620: probability measure $\P^{Q_f}$ on $2^{\Z^d}$ that we denote more simply
621: by $\P^f$.
622: 
623: \procl l.fourier
624: Given $f:\Td\to [0,1]$ measurable and $e_1,\ldots,e_k\in \Z^d$,
625: $$
626: \P^f[\eta(e_1)=1,\dots, \eta(e_k)=1] = 
627: \det [\widehat{f}(e_j-e_i)]_{1\le i, j \le k}
628: \,.
629: $$
630: \endprocl
631: 
632: \proof 
633: By definition, the left-hand side is
634: $\det [(Q_f e_i,e_j)]_{1\le i, j \le k}$. 
635: By \ref t.rudin/(ii), 
636: $$
637: \ip{Q_f e_i,e_j}= \bigip{M_f \ee_{e_i}, \ee_{e_j}}
638: = \widehat{f}(e_j-e_i)\,.
639: \Qed
640: $$
641: 
642: 
643: \procl r.toeplitz
644: \ref l.fourier/ says that for $d=1$,
645: the probability of having 1s on some finite collection of elements of
646: $\Z$ is a particular minor of the Toeplitz matrix associated to $f$.
647: \endprocl
648: 
649: Equation \ref e.incexc/ shows a symmetry of $\P^f$ and
650: $\P^{\bfo-f}$, namely, if $\eta$ has the distribution $\P^f$, then $\bfo - \eta$
651: has the distribution $\P^{\bfo-f}$.
652: \ref b.ShiTak:II/ prove the existence of $\P^f$ by a different method and
653: also note this symmetry.
654: 
655: Although the last lemma gives us a formula for $\P^f$ directly in terms of
656: $f$ without reference to any projections, it is still useful 
657: to know a specific projection of which $M_f$ is a compression. 
658: Let 
659: $f:\Td\to [0,1]$ be measurable. 
660: Identifying $\T^{d+1}$ with $\Td \times \CO{0, 1}$,
661: we let $A_f\subseteq \T^{d+1}$ be the set
662: $\{(x,y)\in \T^{d+1} \st y\le f(x)\}$. 
663: Consider the projection $T_{A_f}$ of $L^2(\T^{d+1})$ given 
664: in \ref t.rudin/(i).
665: We view $L^2(\T^d)$ as a subspace of $L^2(\T^{d+1})$ by identifying
666: $g\in L^2(\T^d)$ with $g\otimes \bfo \in L^2(\T^{d+1})$, where
667: $(g\otimes \bfo) (x,y):= g(x)$ for $x\in \Td$, $y\in \T$.
668: The orthogonal projection $P$ of $L^2(\T^{d+1})$ onto $L^2(\T^d)$ is then
669: given by $g \mapsto (x \mapsto \int_\T g(x,y) \, dy)$. A very simple
670: calculation, 
671: left to the reader, shows that $M_f$ is a compression of $T_{A_f}$;
672: i.e., 
673: $$
674: M_f=P T_{A_f}
675: \label e.compression
676: $$ 
677: on $L^2(\T^d)$ viewed as a subspace of $L^2(\T^{d+1})$. For later use, let 
678: $$
679: H_f:=\{g \in L^2(\T^{d+1}) \st g = 0 \hbox{ a.e.\ on } (A_f)^c \}
680: $$
681: be the image of $T_{A_f}$.
682: 
683: We next remind the reader of the notion of stochastic domination between 
684: two probability measures
685: on $2^E$. First, if $\eta,\delta$ are elements of $2^E$, we write 
686: $\eta \preccurlyeq\delta$ if $\eta(e) \le \delta(e)$ for all $e\in E$. A 
687: subset $A$ of
688: $2^E$ is called {\bf increasing} if $\eta\in A$ and $\eta
689: \preccurlyeq\delta$ imply that $\delta\in A$.
690: If $\nu$ and $\mu$ are two probability measures on $2^E$, we write $\nu
691: \preccurlyeq \mu$ if $\nu(A) \le \mu(A)$ for all increasing sets $A$.
692: A theorem of \ref b.Strassen/ says that this is equivalent to the 
693: existence of a probability
694: measure $m$ on $2^E \times 2^E$ that has $\nu$ and $\mu$ as its first and
695: second marginals (i.e., $m$ is a {\bf coupling} of $\nu$ and $\mu$)
696: and such that $m$ is concentrated on the set
697: $\{(\eta,\delta):\eta \preccurlyeq\delta\}$ (i.e., $m$ is {\bf monotone}).
698: 
699: Throughout the paper, we shall use 
700: the following consequence of conditional negative association (\ref
701: t.negass/), sometimes called
702: {\bf joint negative regression dependence}
703: (see \ref b.Pemantle:negass/).
704: 
705: \procl p.JNRD
706: Assume that $\{X_i\}_{i\in I}$ has conditional negative association,
707: $I$ is the disjoint union of $I_1$ and $I_2$, and 
708: $a,b\in \{0,1\}^{I_1}$ with $a_i\le b_i$ for each $i\in I_1$.
709: Then 
710: $$
711: [\{X_i\}_{i\in I_2}\mid X_i=b_i, i\in I_1]
712: \preccurlyeq [\{X_i\}_{i\in I_2}\mid X_i=a_i, i\in I_1]\,,
713: $$
714: where $[ Y \mid A]$ stands for the law of $Y$ conditional on $A$.
715: \endprocl
716: 
717: \procl l.mon
718: Let $f_1,f_2:\Td\to [0,1]$ with $f_1\le f_2$ a.e. Then
719: $\P^{f_1} \preccurlyeq  \P^{f_2}$.
720: \endprocl 
721: 
722: \proof
723: This follows from a more general result (see \ref b.L:det/) that 
724: says that if $Q_1$ and $Q_2$ are two commuting positive contractions 
725: on $\ell^2(E)$ such that $Q_1\le Q_2$ in the sense that $Q_2-Q_1$ 
726: is positive, then $\P^{Q_1} \preccurlyeq  \P^{Q_2}$. 
727: However, here is a more concrete proof in our case. Since
728: $f_1\le f_2$ a.e., it follows that $H_{f_1}\subseteq H_{f_2}$, which implies
729: by Theorem 6.2 in \ref b.L:det/ that the projection measures 
730: $\P^{H_{f_1}}$ and $\P^{H_{f_2}}$ on $\Z^{d+1}$ satisfy 
731: $\P^{H_{f_1}}  \preccurlyeq  \P^{H_{f_2}}$ and hence their restrictions to $\Z^d$ satisfy
732: the same relationship; i.e., $\P^{f_1} \preccurlyeq\P^{f_2}$.
733: \Qed
734: 
735: 
736: We close this section with a key
737: representation of certain conditional probabilities and an application.
738: The minimum in \ref e.closest/ below is often referred to as a Szeg\H{o}
739: infimum.
740: For an infinite set $B \subseteq \Z^d \setminus \{ \bfz \}$, write
741: $$
742: \P^f[\eta(\bfz) = 1 \mid \eta \restrict B \equiv 1]
743: :=
744: \lim_{n \to\infty}
745: \P^f[\eta(\bfz) = 1 \mid \eta \restrict B_n \equiv 1]
746: \,,
747: $$
748: where $B_n$ is any increasing sequence of finite subsets of $B$ whose union is
749: $B$. This is a decreasing limit by virtue of \ref p.JNRD/.
750: Let $\ip{\cdot, \cdot}_f$ denote the usual inner product in the complex
751: Hilbert space $L^2(f)$.
752: For any set $B \subset \Z^d$, write $[B]_f$ for the closure in $L^2(f)$
753: of the linear span of the complex exponentials $ \{ \ee_k \st k
754: \in B \} $.
755: 
756: \procl t.closest
757: Let $f : \Td \to [0, 1]$ be measurable and $B\subset\Z^d$ with
758: $\bfz \notin B$.
759: Then 
760: $$
761: \P^f[\eta(\bfz) = 1 \mid \eta \restrict B \equiv 1]
762: =
763: \min \left\{ \int |\bfo - u|^2 f \,d\lambda_d \st u \in [B]_f \right\}
764: \,.
765: \label e.closest
766: $$
767: \endprocl
768: 
769: 
770: \proof
771: It suffices to prove the theorem for $B$ finite since the infinite case then
772: follows by a simple limiting argument.
773: So assume that $B$ is finite.
774: Note that $\widehat f(k-j) = \ip{\ee_j, \ee_k}_f$, so that 
775: $$
776: \P^f[\eta \restrict B \equiv 1] = 
777: \det [ \ip{\ee_j, \ee_k}_f]_{j, k \in B}
778: \,,
779: $$
780: and similarly for $\P^f[\eta \restrict (B \cup \{ \bfz \}) \equiv 1]$.
781: Thus, the left-hand side of \ref e.closest/ is a quotient of
782: determinants. The fact that such a quotient has the form of the right-hand
783: side is sometimes called Gram's formula. We include the proof
784: for the convenience of the reader.
785: Since $\ee_\bfz = \bfo$, it follows by row operations on the matrix
786: $[\ip{\ee_j, \ee_k}]_{j, k \in B \cup \{ \bfz \}}$ that
787: $$
788: \P^f[\eta(\bfz) = 1 \mid \eta \restrict B \equiv 1]
789: =
790: \|P_{[B]_f}^\perp \bfo\|_f^2
791: \,,
792: \label e.byproj
793: $$
794: where $P_{[B]_f}^\perp$ denotes orthogonal projection onto the orthogonal
795: complement of $[B]_f$ in $L^2(f)$.
796: Since this is the squared distance from $\bfo$ to $[B]_f$,
797: the equation \ref e.closest/ now follows.% from the next lemma.
798: \Qed
799: 
800: An extension of the above reasoning, given in \ref b.L:det/, provides
801: the entire conditional probability measure:
802: 
803: \procl t.condonB
804: Let $f : \Td \to [0, 1]$ be measurable and $B\subset\Z^d$.
805: Then the law of $\eta \restrict (\Zd \setminus B)$ conditioned on
806: $\eta \restrict B \equiv 1$ is the determinantal probability measure
807: corresponding to the positive contraction on $\ell^2(\Zd \setminus B)$
808: whose $(j, k)$-matrix entry is 
809: $$
810: \bigip{P_{[B]_f}^\perp \ee_j, P_{[B]_f}^\perp \ee_k}_f
811: \,.
812: $$
813: for $j, k \notin B$.
814: \endprocl
815: 
816: 
817: 
818: The Szeg\H{o} infimum that appears in \ref t.closest/ involves
819: trigonometric approximation, a classical area that
820: has strong connections to the topics of prediction and interpolation
821: for wide-sense stationary processes. We briefly discuss these topics
822: now. In later sections, we describe more explicit
823: connections to our results.
824: Recall that a mean-0 wide-sense stationary process is a
825: (not necessarily stationary) process
826: $\Seq{Y_n}_{n\in \Zd}$ for which all the variables have finite
827: variance and mean 0, and such that for each $k \in \Zd$, the covariance
828: $\Cov(Y_{n+k}, Y_n) = \E[Y_{n+k} \overline{Y_n}]$ does not depend on $n$.
829: There is then a positive measure $G$ (called the spectral measure) 
830: on $\T^d$ satisfying
831: $$
832: \widehat{G}(k)=\Cov(Y_{n+k}, Y_n)
833: $$
834: for $n,k\in \Zd$.
835: It turns out that for a one-dimensional wide-sense stationary process,
836: if $G$ is absolutely continuous with density $g$, then
837: $\GM(g)=0$ iff 
838: perfect linear prediction is possible, which means that $Y_0$ is in the closed
839: linear span of $\{Y_n\}_{n\le -1}$ in $L^2(\Omega)$,
840: where $\Omega$ is the underlying probability
841: space. This was proved in various versions by
842: Szeg\H{o}, Kolmogorov and Kre{\u\i}n.
843: Since the covariance with respect to $\P^f$ of 
844: $\eta(\bfz)$ and $\eta(k)$ for $k \in \Z^d$ is also
845: given by a Fourier coefficient (namely $-|\widehat f(k)|^2$ for $k \ne
846: \bfz$)
847: and since, as we shall see, the geometric mean of $f$ will play an 
848: important role in classifying the behavior of $\P^f$ as well,
849: one might wonder about the relationship between our
850: determinantal processes and wide-sense stationary processes.
851: It is not hard to show that $\P^f$ 
852: (viewed as a wide-sense stationary process)
853: has a spectral measure $G$ that is absolutely continuous with
854: density $g$ given by the formula
855: $$
856: g:= \widehat{f}(\bfz) \bfo -f*\widetilde{f}
857: \,,
858: $$
859: where $*$ denotes convolution and
860: $\widetilde{f}(t):=f(1-t)$.
861: Other than the trivial cases $f = \bfz$ and $f = \bfo$,
862: it is easy to check that $g$ is bounded away from 0 
863: and so, in particular, its geometric mean is always
864: strictly positive.
865: This suggests that our results are perhaps not so connected to
866: prediction and interpolation. However, it turns out that
867: some of the questions that
868: we deal with here (such as phase multiplicity and domination)
869: concerning $\P^f$ do have interpretations in terms of prediction and
870: interpolation for wide-sense stationary processes whose spectral
871: density is $f$ (which does not include $\P^f$).
872: More specifics will be given in the relevant sections.
873: 
874: Special attention is often devoted to stationary Gaussian processes, one
875: reason being that their distribution is determined entirely by their
876: spectral measure. It is known that
877: a stationary Gaussian process with no deterministic component
878: is a multistep Markov chain iff its spectral density is the reciprocal of a
879: trigonometric polynomial; see \ref b.Doob:elem/. 
880: The analogous property for determinantal processes is regeneration:
881: 
882: \procl p.regenerate
883: If $d=1$, then $f$ is the reciprocal of a trigonometric polynomial of
884: degree at most $n$ iff $\P^f$ is a regenerative process that regenerates
885: after $n$ successive $1$s appear.
886: \endprocl
887: 
888: This last property means that for any $k$, given that
889: $\eta \restrict [k+1, k+n] \equiv 1$, the future, $\eta \restrict
890: \CO{k+n+1,\infty}$, is conditionally independent of the past, $\eta
891: \restrict \OC{-\infty, k}$.
892: 
893: \proof
894: Note first that because of \ref t.condonB/, this
895: regenerative property holds for $\P^f$ iff for all $j \ge 0$ and all $C
896: \subset \OC{-\infty, -n-1}$, we have $P_{[B]_f}^\perp \ee_j = P_{[B \cup
897: C]_f}^\perp \ee_j$, where $B := [-n, -1]$.
898: (Here, we are relying on the fact that $\|P_{[B]_f}^\perp \ee_j\|_f > \|P_{[B
899: \cup C]_f}^\perp \ee_j\|_f$ if the vectors are not equal.)
900: This is the same as $P_{[B]_f}^\perp \ee_j \perp [C]_f$ for all $j \ge 0$, or,
901: in other words, there exists some $u_j \in [B]_f$ with $\ee_j - u_j \perp [B
902: \cup C]_f$. As this would have to hold for all $C$
903: (and $u_j$ is independent of $C$), 
904: it is the same as the existence of
905: some $u_j \in [B]_f$ such that $T_j := (\ee_j - u_j) f$ is analytic, i.e.,
906: $\widehat {T_j}(k) = 0$ for all $k < 0$. 
907: Now this implies that $\overline{T_0} = (\bfo - \overline{u_0}) f$, whence 
908: $T_0 (\bfo - \overline{u_0}) = \overline{T_0} (\bfo - u_0)$.
909: %%(Note that we may divide by analytic and conjugate analytic functions since
910: %%they vanish only on a set of measure 0; see \ref b.Rudin:RCA/, Theorem 17.18,
911: %%p.~345.)
912: The left side of this last 
913: equation is an analytic function, while the right side
914: is the conjugate of an analytic function (i.e., its Fourier coefficients
915: vanish on $\Z^+$).
916: Therefore, both equal some constant, $c$.
917: Hence 
918: $$
919: f = {T_0 \over \bfo - u_0} = {c \over (\bfo - u_0) (\bfo - \overline{u_0})}
920: = {c \over |\bfo - u_0|^2}
921: \,,
922: $$
923: which is indeed the reciprocal of a trigonometric polynomial of degree at most
924: $n$.
925: 
926: Conversely, if $1/f$ is the reciprocal of a trigonometric polynomial of degree
927: at most $n$, then since $f \ge 0$, the theorem of Fej\'er and Riesz
928: (see \ref b.GrenanderSzego/, p.~20)
929: allows us to write $f = c/|\bfo - u_0|^2$ for some conjugate-analytic
930: polynomial $u_0 \in [B]_f$ such that the analytic extension of
931: $\bfo - \overline{u_0}$ to the unit disc has no zeroes.
932: We may rewrite this as $(\bfo - u_0) f = T_0$ for $T_0 := c/(\bfo -
933: \overline{u_0})$. 
934: Since $T_0$ has an extension to the unit disc as the reciprocal of an
935: analytic polynomial with no zeroes, it follows that $T_0$ is also analytic.
936: Multiplying both sides of this equation by $\ee_1$ and rewriting $\ee_1 u_0 f =
937: c_1 (\bfo - u'_1) f = c_1 (u_0 - u'_1) f + c_1 T_0$ for some constant $c_1$ and
938: some $u'_1 \in [B]_f$, we see that $(\ee_1 - u_1) f = T_1$ for $u_1 := c_1 (u_0
939: - u'_1) \in [B]_f$ and $T_1 := \ee_1 T_0 + c_1 T_0$, an analytic function.
940: Similarly, we may establish by induction that for each $j \ge 0$,
941: there is some $u_j \in [B]_f$ such that $T_j := (\ee_j - u_j) f$ is 
942: analytic.
943: This proves the equivalence desired.
944: %(One can also see that
945: %$T_{k+1} = \ee_1 T_k + c_{k+1} T_0$ for some constant $c_{k+1}$.)
946: %\msnote{There is a more explicit form available, so I don't think this adds
947: %much.}
948: \Qed
949: 
950: We now use this proof to establish the explicit probabilistic form of the
951: renewal process in \ref x.renewal/.
952: In this case, $B = \{-1\}$ and one easily verifies that $u_j = a^{j+1}
953: \ee_{-1}$, as is standard in the theory of linear prediction. 
954: Therefore, 
955: $$
956: \P^f[\eta(j) = 1 \mid \eta(-1) = 1]
957: =
958: \|P_{[B]_f}^\perp \ee_j\|_f^2
959: =
960: \|\ee_j - u_j\|_f^2
961: =
962: \|\ee_j\|_f^2 - \|u_j\|_f^2
963: =
964: {1-a \over 1+a} (1 - a^{2j+2})
965: \,.
966: $$
967: It now suffices to verify that this is also true for the explicit renewal
968: process described in \ref x.renewal/.
969: First, it is well known from basic renewal theory that for a renewal 
970: process, the probabilities
971: $\P^f[\eta(j) = 1 \mid \eta(-1) = 1]$ determine the distribution of the 
972: number of 0s between two 1s. Hence to verify the above statement,
973: one simply needs to check that these latter probabilities are related to
974: the interrenewal distribution via the appropriate convolution-type equation.
975: In this case, it comes down to verifying that for all $j\ge 0$,
976: $$
977: {1-a \over 1+a}(1-a^{2j+2})
978: =
979: (j+1)(1-a)^2a^j
980: +{1-a \over 1+a}\sum_{k=1}^j k (1-a)^2a^{k-1}(1-a^{2(j-k)+2})\,.
981: $$
982: This identity is easy to check.
983: 
984: 
985: 
986: \bsection {The Bernoulli Shift Property}{s.bern}
987: 
988: In this section,
989: we prove that the stationary determinantal processes studied here
990: are Bernoulli shifts. We 
991: assume the reader is familiar with the
992: basic notions of a Bernoulli shift (see, e.g., \ref b.Ornstein:book/).
993: 
994: \procl t.Bern
995: Let $f : \T^d \to [0, 1]$ be measurable.
996: Then $\P^f$ is a Bernoulli shift; i.e., it is isomorphic 
997: (in the sense of ergodic theory) to an i.i.d.\ process.
998: \endprocl
999: 
1000: Before beginning the proof, we present a few preliminaries.
1001: We first recall the definition of the $\dbar$-metric.
1002: 
1003: \procl d.overd
1004: If $\mu$ and $\nu$ are 
1005: $\Z^d$-invariant probability measures
1006: on $2^{\Z^d}$, then
1007: $$
1008: \dbar(\mu,\nu) := \inf_m m \Big[ \big \{ (\eta,\delta) \in 2^{\Z^d} \times
1009: 2^{\Z^d} \st \eta(\bfz)\ne \delta(\bfz) \big \}\Big]\, ,
1010: $$
1011: where the infimum is taken over all couplings $m$ of $\mu$ and 
1012: $\nu$ that are $\Z^d$-invariant.
1013: \endprocl 
1014: 
1015: The following is a slight generalization of a well-known result
1016: (see, for example, page 75 in \ref b.Liggett/). The 
1017: proof is an immediate consequence of the existence of a monotone coupling
1018: and is left to the reader.
1019: 
1020: 
1021: \procl l.weak
1022: Suppose that $\sigma_1$, $\sigma_2$, $\nu$ and $\mu$ are
1023: $\Z^d$-invariant probability measures on $2^{\Z^d}$ such that $\nu
1024: \preccurlyeq \sigma_i \preccurlyeq  \mu$ for both $i=1, 2$. 
1025: Then 
1026: $$
1027: \dbar(\sigma_1, \sigma_2) \le \mu[\eta(\bfz) = 1] - \nu[\eta(\bfz)= 1]
1028: \,.
1029: $$
1030: \endprocl 
1031: 
1032: \procl p.L1givesdbar
1033: Let $f, g : \T^d \to [0, 1]$ be measurable.
1034: Then 
1035: $$
1036: \dbar(\P^f, \P^g) \le \int_{\T^d} |f-g| \,d\lambda_d
1037: \,.
1038: $$
1039: \endprocl 
1040: 
1041: \proof
1042: Because of \ref l.mon/, we may apply \ref l.weak/ to $\sigma_1 := \P^f$,
1043: $\sigma_2 := \P^g$, $\nu := \P^{f \wedge g}$, and $\mu := \P^{f \vee g}$.
1044: We obtain that 
1045: $$
1046: \dbar(\P^f, \P^g)
1047: \le
1048: \int_{\T^d} f \vee g \,d\lambda_d - \int_{\T^d} f \wedge g \,d\lambda_d
1049: =
1050: \int_{\T^d} |f-g| \,d\lambda_d
1051: \,.
1052: \Qed
1053: $$
1054: 
1055: \proofof t.Bern
1056: The first step is to approximate $f$ by 
1057: trigonometric polynomials. Let $K_r$ be the $r$th Fej\'er kernel for $\T$, 
1058: $$
1059: K_r := \sum_{|j| \le r} \Big(1-{|j|\over r+1}\Big) \ee_j 
1060: \,,
1061: $$
1062: and define $K^d_r(x_1, \ldots, x_d) := \prod_{i=1}^d K_r(x_i)$.
1063: It is well known 
1064: that $K^d_r$ is a positive summability kernel, so that
1065: if we define $g_r$ by
1066: $$
1067: g_r:= f * K^d_r\,,
1068: $$
1069: then $0\le g_r \le 1$ and $\lim_{r\to\infty} g_r =f$ a.e.\ and in $L^1(\Td)$. 
1070: 
1071: Next, since each $K^d_r$ is a trigonometric polynomial, so is each $g_r$.
1072: {}From this and \ref l.fourier/, it is easy to see that there is a constant 
1073: $C$ such that $\P^{g_r}(A\cap B) = \P^{g_r}(A)\P^{g_r}(B)$ 
1074: if $A$ is of the form $\eta\equiv 1 \hbox{ on } S_1$ and 
1075: $B$ is of the form $\eta\equiv 1 \hbox{ on } S_2$ 
1076: with $S_1$ and $S_2$ being finite sets 
1077: and having distance at least $C$ between them.
1078: {}From this, a standard argument in probability (a $\pi$-$\lambda$
1079: argument) shows that if $A$ is any event depending only on
1080: locations $S_1$ and if $B$ is any event depending only on locations $S_2$
1081: with $S_1$ and $S_2$ possibly infinite sets
1082: having distance at least $C$ between them, then
1083: $\P^{g_r}(A\cap B) = \P^{g_r}(A)\P^{g_r}(B)$. A process 
1084: with this property is called a {\bf finitely dependent} process 
1085: and this property implies it is a Bernoulli shift 
1086: (e.g., the so-called very weak Bernoulli condition 
1087: is immediately verified; see \ref b.Ornstein:book/ for this definition for
1088: $d=1$ and, e.g., \ref b.Steif/ for general $d$).
1089: 
1090: Since the processes that are Bernoulli shifts 
1091: are closed in the $\dbar$ metric (see
1092: \ref b.Ornstein:book/) and \ref p.L1givesdbar/ tells us that
1093: $\lim_{r\to\infty} \dbar(\P^{g_r},\P^f) = 0$, we 
1094: conclude that $\P^f$ is a Bernoulli shift.
1095: \Qed
1096: 
1097: \procl r.WB
1098: An important property of 1-dimensional 
1099: stationary processes is the weak Bernoulli
1100: (WB) property. This is also referred to as ``$\beta$-mixing" and ``absolute
1101: regularity" in the literature. Despite its name, it is known that WB is
1102: strictly stronger than Bernoullicity. It is easy to check that ``aperiodic''
1103: regenerative processes and finitely dependent processes are WB. Hence,
1104: by our earlier results, if $f$ is a trigonometric polynomial or the inverse
1105: of a trigonometric polynomial (or if $\bfo-f$ is the inverse
1106: of a trigonometric polynomial), then $\P^f$ is WB. This is subsumed by the
1107: independent
1108: work of \ref b.ShiTak:II/, who showed that $\P^f$ is WB whenever
1109: $\sum_{n \ge 1} n |\widehat f(n)|^2 <\infty$.
1110: The precise class of $f$ for which $\P^f$ is WB is not known.
1111: We also note that it follows from 
1112: \ref b.Smor/ that if
1113: $f$ is a trigonometric polynomial, then $\P^f$ is finitarily isomorphic
1114: to an i.i.d.\ process.
1115: \endprocl
1116: 
1117: \bsection {Support of the Measures}{s.fs}
1118: 
1119: In this section, we show that all of the probability measures $\P^f$ have
1120: full support except in two degenerate cases. We begin with the following
1121: lemma.
1122: 
1123: \procl l.lindep
1124: Let $H$ be a closed subspace of $\hE$ and let $A$ and $B$ be finite disjoint
1125: subsets of $E$. Then
1126: $$
1127: \P^H[\eta(e)=1 \hbox{ for } e\in A, \,\, \eta(e)=0 \hbox{ for } e\in B] >0
1128: $$
1129: if and only if 
1130: $\{P_H(e)\}_{e\in A}\cup\{P_{H^\perp}(e)\}_{e\in B}$ is linearly independent.
1131: \endprocl
1132: 
1133: \proof
1134: In the special case that $A$ or $B$ is empty, the result follows from
1135: \ref b.L:det/.
1136: In general,
1137: $\{P_H(e)\}_{e\in A}\cup\{P_{H^\perp}(e)\}_{e\in B}$ is linearly independent
1138: if and only if
1139: $\{P_H(e)\}_{e\in A}$ and $\{P_{H^\perp}(e)\}_{e\in B}$ are each linearly
1140: independent sets. Also,
1141: $\P^H[\eta \restrict A \equiv 1,\, \eta \restrict B \equiv 0] >0$
1142: if and only if
1143: $\P^H[\eta \restrict A \equiv 1] >0$ and
1144: $\P^H[\eta \restrict B \equiv 0] >0$ since 
1145: $$
1146: \P^H[\eta \restrict A \equiv 1, \,\, \eta \restrict B \equiv 0] \ge  
1147: \P^H[\eta \restrict A \equiv 1] \P^H[\eta \restrict B \equiv 0] 
1148: $$ 
1149: by the negative association property, \ref t.negass/.
1150: Thus, the general case follows from the special case.
1151: \Qed
1152: 
1153: \procl t.support
1154: $\P^f$ has full support
1155: for any function $f:\Td\to [0,1]$ other than $\bfz$ or $\bfo$.
1156: \endprocl
1157: 
1158: \proof
1159: Since marginals of probability measures with full support clearly
1160: have full support, it suffices to prove the result for
1161: $\P^H$ when $H$ is a closed $\Z^d$-invariant
1162: subspace of $\hD$ other than
1163: $\hD$ or 0. If we translate over to $L^2(\T^d)$, we see, by \ref l.lindep/,
1164: that it suffices to show that
1165: if $A\subseteq \Td$ with $0< \lambda_d(A) < 1$
1166: and $n_1,\ldots,n_k, m_1,\ldots,m_\ell$ are
1167: all distinct elements of $\Z^d$, then
1168: $$
1169: \{\ee_{n_j} \I A\}_{1\le j \le k} \cup
1170: \{\ee_{m_r} \I {A^c}\}_{1\le r \le \ell} 
1171: $$
1172: is linearly independent.
1173: Suppose that
1174: $c_1,\ldots,c_k, d_1,\ldots,d_\ell$ are complex numbers such that
1175: $$
1176: \sum_{j=1}^k c_j \ee_{n_j} \I A +
1177: \sum_{r=1}^\ell d_r \ee_{m_r}  \I {A^c} = \bfz 
1178: \,.
1179: $$
1180: From this it follows that
1181: $\sum_{j=1}^k c_j \ee_{n_j}=0$ a.e.\ on $A$. 
1182: Since $\lambda_d(A) >0$,
1183: we have that $\sum_{j=1}^k c_j \ee_{n_j}$ is 0 
1184: on a set of positive measure.
1185: It is well known that this implies that $c_1,\ldots,c_k$ vanish (the proof
1186: uses induction on $d$ and Fubini's theorem). Similarly, $d_1,\ldots,d_\ell$
1187: vanish.
1188: \Qed
1189: 
1190: 
1191: \bsection {Domination Properties}{s.dom}
1192: 
1193: In this section, we study the question of
1194: which product measures are stochastically dominated by $\P^f$ and
1195: which product measures stochastically dominate $\P^f$. 
1196: For simplicity, we give our first results for $d=1$, and only afterwards
1197: describe how these results extend to higher
1198: dimensions. We also treat a different 
1199: notion of ``full domination" at the end of the
1200: section.
1201: Recall that for $f : \Td \to [0, 1]$, we define 
1202: $$
1203: \GM(f) := \exp \int_\Td \log f \,d\lambda_d
1204: \,.
1205: $$
1206: 
1207: We introduce an auxiliary stronger
1208: notion of domination than $\preccurlyeq$, but only in the case where
1209: one of the measures is a product measure. Let $\mu_p$ denote product
1210: measure with density $p$. 
1211: 
1212: \procl d.sdom
1213: A stationary process $\{\eta_n\}_{n\in\Z}$ with distribution $\nu$
1214: {\bf strongly dominates} $\mu_p$, written 
1215: $\mu_p\strle\nu$,
1216: if for any $n$ and any $a_1,\ldots,a_n \in \{0,1\}$,
1217: $$
1218: \P[\eta_0=1 \mid \eta_i=a_i, i=1,\ldots, n]\ge p
1219: \,.
1220: $$
1221: Similarly, we define $\nu\strle\mu_p$ if the above inequality holds
1222: when $\ge$ is replaced by $\le$.
1223: \endprocl
1224: 
1225: The following lemma is easy and well known; it is sometimes referred to as 
1226: Holley's lemma.
1227: 
1228: \procl l.holley
1229: If $\mu_p\strle \nu$, then $\mu_p\preccurlyeq\nu$.
1230: \endprocl 
1231: 
1232: The converse is not true, as we shall see later (\ref r.support/).
1233: In light of \ref l.mon/, we have that
1234: if $p \le f \le q$, then 
1235: $$
1236: \mu_p \preccurlyeq \P^f \preccurlyeq \mu_q
1237: \,.
1238: $$
1239: The optimal improvement of these stochastic bounds is as follows.
1240: 
1241: \procl t.dom
1242: For any $f : \T \to [0, 1]$, we have
1243: $\mu_p\preccurlyeq\P^f$ iff $p\le \GM(f)$ iff $\mu_p\strle\P^f$. Similarly,
1244: $\P^f\preccurlyeq\mu_q$ iff $q\ge 1-\GM(\bfo-f)$ iff $\P^f\strle\mu_q$.
1245: In addition, for any stationary process $\mu$ that 
1246: has conditional negative association, we have
1247: $\mu_p\preccurlyeq\mu$ iff $\mu_p\strle\mu$.
1248: \endprocl 
1249: 
1250: \proof
1251: Let $d_n := D_{n-1}(f)$ be the probability of having $n$ 1s in a row.
1252: According to Szeg\H{o}'s theorem (\ref b.GrenanderSzego/, pp.~44, 66),
1253: $d_{n+1}/d_n$ is decreasing in $n$
1254: and
1255: $$
1256: \lim_{n\to\infty}d_{n+1}/d_n= \GM(f)
1257: = \lim_{n \to\infty} d_n^{1/n}
1258: \,.
1259: $$
1260: In particular, $d_{n+1}/d_n\ge \GM(f)$ for all $n$.
1261: 
1262: \ref p.JNRD/ implies that
1263: for any fixed $n$,
1264: $$
1265: \P^f[\eta_0=1 \mid \eta_i=a_i, i=1,\ldots, n]
1266: $$
1267: is minimized among all $a_1,\ldots,a_n \in \{0,1\}$
1268: when $a_1 = a_2 = \cdots = a_n = 1$.
1269: In this case, the value is $d_{n+1}/d_n$.
1270: Since this is at least $\GM(f)$, 
1271: we deduce that if $p\le \GM(f)$, then
1272: $\mu_p\strle \P^f$ and hence that $\mu_p\preccurlyeq\P^f$. 
1273: 
1274: Conversely, if $\mu_p \preccurlyeq \P^f$, then certainly
1275: $p^n\le d_n$ for all $n$. 
1276: Hence $p \le \GM(f)$. The second to last statement can be proved in 
1277: the same way or can be concluded by symmetry. 
1278: Finally, the last statement can be proved in a similar fashion.
1279: \Qed
1280: 
1281: % The proof also shows the equivalence with $d_n \ge p^n$ for all $n$.
1282: 
1283: \procl r.examples
1284: The above theorem gives us two interesting examples. 
1285: First, if we take $f:=\I A$ where $A$ has Lebesgue measure $1-\epsilon$, then
1286: $\dbar(\P^f, \delta_{\bfo}) \le \epsilon$ by 
1287: \ref p.L1givesdbar/, but nonetheless, $\P^f$ does not dominate
1288: any nontrivial product measure since $\GM(f) = 0$.
1289: Second, if we take
1290: $f:={\bf 1}_{[0,1/2]}+ .4\cdot {\bf 1}_{[1/2,1]}$, 
1291: then $f < 1/2$ on a set of positive
1292: measure, but nonetheless $\P^f$ dominates $\P^{1/2}=\mu_{1/2}$
1293: since $\GM(f) = (.4)^{1/2} > 1/2$.
1294: \endprocl
1295: 
1296: 
1297: \procl x.ustH_x-dom
1298: Let $f$ be as in \ref x.ustH/, so that $\P^f$ is the law of the horizontal
1299: edges of the uniform spanning tree in the plane.
1300: In order to examine a 1-dimensional process, let us consider only
1301: the edges lying on the $x$-axis.
1302: If we let 
1303: $$
1304: g(x) := \int_\T f(x, y) \,d\lambda_1(y)
1305: \,,
1306: $$
1307: then the edges lying on the $x$-axis have the law $\P^g$
1308: since $\widehat g(k) = \widehat f(k, 0)$ for all $k \in \Z$.
1309: Since an antiderivative of $1/(1 + a\sin^2 \pi y)$ is 
1310: $$
1311: {\arctan \left( \sqrt{1+a}\tan \pi y\right) \over \pi \sqrt{1+a}}
1312: \,,
1313: $$
1314: we have 
1315: $$
1316: g(x) =
1317: {\sin \pi x \over \sqrt{1+\sin^2 \pi x}}
1318: \,,
1319: $$
1320: as given in \ref x.ustHx/.
1321: We claim that $\P^g$ strongly dominates $\mu_p$ for $p := \sqrt 2 - 1$, and
1322: this is optimal.
1323: In order to show this, we calculate $\GM(g)$.
1324: Write $g_1(x) := \sin^2 \pi x$. 
1325: Then $(\GM(g))^2 = \GM(g_1)/\GM(\bfo + g_1)$.
1326: Let $G_1$ be the harmonic extension of $\log g_1$ from the circle
1327: to the unit disc.
1328: Then $\int \log g_1 \,d\lambda_1 = G_1(0)$ by the mean value property of
1329: harmonic functions.
1330: Since $g_1 = |(1-\ee_1)/2|^2$, we see that $G_1$ is the
1331: real part of the analytic function $z \mapsto 2 \log [(1-z)/2]$
1332: in the unit disc, from which we conclude that $G_1(0) = \log (1/4)$.
1333: Therefore, $\GM(g_1) = 1/4$. 
1334: Similarly, $\bfo+g_1 = |[\sqrt 2 + 1 - (\sqrt 2 - 1)\ee_1]/2|^2$, whence
1335: $\GM(\bfo+g_1) = [(\sqrt 2 + 1)/2]^2$.
1336: Therefore $\GM(g) = \sqrt 2 - 1 = 0.4142^+$, as desired.
1337: It turns out that $q := 1 - \GM(\bfo - g) = 1 - 2(\sqrt 2 -1)e^{-2\cat/\pi} =
1338: 0.5376^+$, where 
1339: $$
1340: \cat := \sum_{k=0}^\infty {(-1)^k \over (2k+1)^2} = 0.9160^-
1341: \label e.defcat
1342: $$
1343: is {\bf Catalan's constant}.
1344: Thus, $\P^g \strle \mu_q$.
1345: It is interesting how close $p$ and $q$ are.
1346: \endprocl
1347: 
1348: We now introduce some new mixing conditions.
1349: Given a positive integer $r$, we may restrict $\P^f$ to $2^{r\Z}$.
1350: If we identify $r\Z$ with $\Z$, then we obtain the process $\P^{f_r}$, where 
1351: $$
1352: f_r(t) := {1 \over r} \sum_{x \in r^{-1} t} f(x)
1353: \,,
1354: $$
1355: where $r^{-1} t := \{ x \in \T \st r x = t \}$.
1356: The reason that this restriction is equal to $\P^{f_r}$ is that
1357: for all $k \in \Z$, we have $\widehat{f_r}(k) = \widehat f(rk)$, as
1358: is easy to check.
1359: Because of this relation, we have that $\int |f_r(t) - \widehat
1360: f(0)|^2\,d\lambda_1(t) \to 0$ as $r \to\infty$, so that by \ref l.weak/, it
1361: follows that $\dbar(\P^{f_r}, \P^{\widehat f(0)}) \to 0$ as $r \to\infty$.
1362: (It is not hard to show that a similar property holds for any Kolmogorov
1363: automorphism (see \ref d.K/), while the example in \ref r.support/ shows
1364: that this property can occur in other cases as well.)
1365: In fact, we often have a stronger convergence for our determinantal processes,
1366: as we show next.
1367: 
1368: \procl t.dom-mix
1369: Let $f : \T \to [0, 1]$ be measurable.
1370: If $\GM(f) > 0$ or $f$ is bounded away from 0 on an interval of positive length,
1371: then there exist constants $p_r \to \widehat f(0)$ such that $\P^{f_r}
1372: \succcurlyeq \P^{p_r}$ for all $r$.
1373: Therefore, if $\GM\big(f(\bfo - f)\big) > 0$ or if $f$ is continuous and not
1374: equal to $\bfz$ nor $\bfo$, then there exist constants $p_r, q_r \to \widehat
1375: f(0)$ such that $\P^{p_r} \preccurlyeq \P^{f_r} \preccurlyeq \P^{q_r}$ for all
1376: $r$.
1377: \endprocl
1378: 
1379: Of course, in view of \ref t.dom/, we use $p_r := \GM(f_r)$ and $q_r := 1 -
1380: \GM(\bfo - f_r)$.
1381: Thus, the theorem is an immediate consequence of the following lemma.
1382: 
1383: \procl l.dom-mix
1384: Let $f : \T \to [0, 1]$ be measurable.
1385: If $\GM(f) > 0$ or $f$ is bounded away from 0 on an interval of positive length,
1386: then $\GM(f_r) \to \widehat f(0)$ as $r \to\infty$.
1387: \endprocl
1388: 
1389: \proof
1390: We have seen that $f_r \to \fav \cdot \bfo$ in $L^2$, whence also in measure.
1391: Therefore $\log f_r \to (\log \fav) \cdot \bfo$ in measure.
1392: Thus, it remains to show that $\{ \log f_r \}$ is uniformly integrable (at
1393: least, for all large $r$).
1394: Suppose first that $\GM(f) > 0$.
1395: 
1396: Given $h \in L^1(\T)$, write $h^{(r)}(t) := h(r t)$, so that
1397: $\widehat{h^{(r)}}(k)$ is 0 when $k$ is not a multiple of $r$ and is
1398: $\widehat h(k/r)$ when $k$ is a multiple of $r$.
1399: For any $g, h \in L^2(\T)$, we have
1400: \begineqalno
1401: \int_\T g_r(t) \overline{h(t)} \,d\lambda_1(t)
1402: &=
1403: \sum_{k \in \Z} \widehat{g_r}(k) \overline{\widehat h(k)}
1404: =
1405: \sum_{k \in \Z} \widehat{g}(r k) \overline{\widehat h(k)}
1406: =
1407: \sum_{k \in \Z} \widehat{g}(k) \overline{\widehat{h^{(r)}}(k)}
1408: \cr&=
1409: \int_\T g(t) \overline{h^{(r)}(t)} \,d\lambda_1(t)
1410: =
1411: \int_\T g(t) \overline{h(r t)} \,d\lambda_1(t)
1412: \,.
1413: \label e.turn
1414: \cr
1415: \endeqalno
1416: Therefore, for any set $A \subseteq \T$, we have 
1417: $$
1418: \int_{A} \log (f_r) \,d\lambda_1
1419: \ge
1420: \int_{A} (\log f)_r \,d\lambda_1
1421: =
1422: \int_{r^{-1} A} \log f\,d\lambda_1
1423: \,;
1424: $$
1425: the inequality follows by concavity of log, while the equality follows from
1426: \ref e.turn/ applied to $g := \log f$ and $h := \I A$.
1427: Given any $\epsilon \in \big(0, \fav\big)$, let $A^r_\epsilon := \{ t \st
1428: f_r(t) < \epsilon \}$.
1429: Note that for any measurable $A \subset \T$, we have
1430: $\lambda_1(r^{-1} A) = \lambda_1(A)$.
1431: Since $f_r \to \fav \cdot \bfo$ in measure, it follows that
1432: $$
1433: \lim_{r \to\infty} \lambda_1(r^{-1} A^r_\epsilon) =
1434: \lim_{r \to\infty} \lambda_1(A^r_\epsilon)
1435: = 0
1436: \,.
1437: $$
1438: Since $\log f$ is integrable, it follows that 
1439: $$
1440: \lim_{r \to\infty} \int_{r^{-1} A^r_\epsilon} \log f \,d\lambda_1 =
1441: 0
1442: \,.
1443: $$
1444: This establishes the uniform integrability in the first case.
1445: 
1446: In the second case where $f \ge c > 0$ on an interval of length
1447: $\epsilon > 0$, we have that $f_r \ge c \epsilon$ on all of $\T$ for
1448: all $r > 1/\epsilon$.
1449: It is then obvious that $\{\log f_r\}$ is uniformly integrable for all 
1450: $r > 1/\epsilon$.
1451: \Qed
1452: 
1453: \procl r.nasty
1454: There are functions $f$ with $f > 0$ a.e., yet $\GM(f_r) = 0$ for all $r$.
1455: For example, enumerate the rationals $\{ x_j \st j \ge 1\}$ in $(0, 1)$ and
1456: choose $\epsilon_j > 0$ such that $x_j + \epsilon_j < 1$ and $\sum_j
1457: \epsilon_j < 1$.
1458: Define $A_n := [0, 1] \setminus \bigcup_{j > n} [x_j, x_j + \epsilon_j]$ and
1459: $f_0(x) := e^{-1/x}$.
1460: Then one can show that 
1461: $$
1462: f(x) := f_0(x) \I{A_0}(x) +
1463: \sum_{n=1}^\infty f_0(x-x_n) \I{[x_n, x_n+\epsilon_n] \cap A_n}(x) 
1464: $$
1465: is such an example.
1466: \endprocl
1467: 
1468: 
1469: 
1470: 
1471: We turn next to the extension of the prior results to higher dimensions.
1472: This is basically straightforward, but has interesting applications, as we
1473: shall see.
1474: First, we recall the usual notion of the {\bf past $\sigma$-field} $\past(k)$
1475: at a point $k \in \Z^d$.
1476: We use lexicographic ordering on $\Z^d$, i.e., write $(k_1, k_2,
1477: \ldots, k_d) \prec (l_1, l_2, \ldots, l_d)$ if $k_i < l_i$ when $i$ is the
1478: smallest index such that $k_i \ne l_i$.
1479: Then define $\past(k)$ to be the $\sigma$-field generated by $\eta(j)$ for all
1480: $j \prec k$.
1481: More generally, the past $\sigma$-field could be defined with respect to 
1482: any {\bf ordering} of $\Z^d$, which means the selection of a set $\ord \subset
1483: \Z^d$ that has the properties $\ord \cup (-\ord) = \Zd \setminus \{ \bfz \}$, 
1484: $\ord \cap (-\ord) = \emptyset$, and $\ord+\ord \subset \ord$. The associated ordering 
1485: is that where $k \prec l$ iff $ l-k \in \ord$.
1486: $\past_\ord(u)$ will denote the past of $u$ with respect to the ordering
1487: $\ord$,
1488: i.e., $\{v \st v \prec u\}$.
1489: Thus, $\past_\ord(\bfz)$ is just $-\ord$. 
1490: (For a characterization of
1491: all orders, see \ref b.Teh/, \ref b.Zaiceva/, or \ref b.Trevisan/.)
1492: As before, let $\mu_p$ denote product measure with density $p$. 
1493: 
1494: 
1495: \procl d.sdom-d
1496: Given an ordering $\ord$, 
1497: a stationary process $\{\eta_n\}_{n\in\Z^d}$ with distribution $\nu$
1498: {\bf strongly dominates} $\mu_p$, written 
1499: $\mu_p\strle\nu$,
1500: if 
1501: $$
1502: \P[\eta_\bfz=1 \mid \past_\ord(\bfz)]\ge p \quad\nu\hbox{-a.s.}
1503: $$
1504: Similarly, we define $\nu\strle\mu_p$ if the above inequality holds
1505: when $\ge$ is replaced by $\le$. (Note that the ordering $\ord$ here is 
1506: suppressed in the notation.)
1507: \endprocl
1508: 
1509: Although we have phrased it differently, in one dimension, this is equivalent
1510: to \ref d.sdom/ when $\ord=\{-1,-2,\ldots\}$.
1511: Again, we have a version of Holley's lemma:
1512: 
1513: \procl l.holley-d
1514: Given any ordering $\ord$, 
1515: if $\mu_p\strle \nu$, then $\mu_p\preccurlyeq\nu$.
1516: \endprocl 
1517: 
1518: (Note that to produce a monotone coupling for a general ordering with
1519: respect to a set $\ord$, it is enough to do so for the measures
1520: restricted to any finite $B \subset \Zd$. Given such a $B$, order $B$ by the
1521: restriction of $\prec$ to $B$ and couple the measures by adding sites from
1522: $B$ in this order.)
1523: 
1524: We now prove
1525: 
1526: \procl t.dom-d
1527: Fix any ordering $\ord$.
1528: For any measurable $f : \Td \to [0, 1]$, we have
1529: $\mu_p\preccurlyeq\P^f$ iff $p\le \GM(f)$ iff $\mu_p\strle\P^f$. Similarly,
1530: $\P^f\preccurlyeq\mu_q$ iff $q\ge 1-\GM(\bfo-f)$ iff $\P^f\strle\mu_q$.
1531: In addition, for any stationary process $\mu$ that
1532: has conditional negative association, we have
1533: $\mu_p\preccurlyeq\mu$ iff $\mu_p\strle\mu$.
1534: \endprocl 
1535: 
1536: \proof
1537: \ref p.JNRD/ and \ref t.closest/ imply that 
1538: \begineqalno
1539: \essinf \P^f\Big[\eta(\bfz) = 1 \mid \past_\ord(\bfz)\Big] 
1540: &=
1541: \inf \Big\{ \P^f[\eta(\bfz) = 1 \mid \eta \restrict A \equiv 1] \st
1542: A \subset -\ord \hbox{ is finite} \Big\}
1543: \cr&=
1544: \P^f[\eta(\bfz) = 1 \mid \eta \restrict (-\ord) \equiv 1]
1545: \cr&=
1546: \min \left\{ \int |\bfo - T|^2 f \,d\lambda_d \st T \in [-\ord]_f \right\}
1547: \,.
1548: \endeqalno
1549: By \ref b.HelLow:I/, the latter quantity equals $\GM(f)$.
1550: Therefore, $p\le \GM(f)$ iff $\mu_p\strle\P^f$. 
1551: On the other hand, if $\mu_p\preccurlyeq\P^f$, then let $A(r)$ be the
1552: finite subset of $-\ord$ consisting of all points within some large radius $r$
1553: about $\bfz$.
1554: Let $a_1 \prec a_2 \prec \cdots \prec a_n$ be the elements of $A(r)$ in
1555: order, where $n := |A(r)|$, and write $A_j := A(r) \cap (-\ord+a_j)$.
1556: Then 
1557: $$
1558: p^n \le \P^f[\eta \restrict A(r) \equiv 1]
1559: =
1560: \prod_{j=1}^n \P^f[\eta(a_j) = 1 \mid \eta \restrict A_j \equiv 1]
1561: \,.
1562: $$
1563: Most terms in this product are quite close to $\GM(f)$, while all lie in
1564: $[\GM(f), 1]$.
1565: It follows that
1566: $$
1567: \lim_{r \to\infty} \P^f[\eta \restrict A(r) \equiv 1]^{1/n}
1568: =
1569: \GM(f)
1570: \,,
1571: $$
1572: so that $p \le \GM(f)$.
1573: Finally, the remaining statements can be proved in a similar fashion.
1574: \Qed
1575: 
1576: \procl r.past
1577: It follows from \ref t.dom-d/ that the choice of ordering $\ord$ does not
1578: determine whether $\mu_p \strle \P^f$.
1579: This is not true for general stationary processes, even in one dimension. 
1580: We are grateful to Olle H\"aggstr\"om for the following simple example.
1581: Let $\mu$ be the distribution on $\{0, 1 \}^2$ given by 
1582: $$
1583: \mu = (1/7) (\delta_{(0,0)} +\delta_{(0,1)}) + (2/7)\delta_{(1,0)} + (3/7)
1584: \delta_{(1,1)}
1585: \,.
1586: $$
1587: Let $\eta \in \{ 0, 1 \}^\Z$ be such that with probability 1/2, $(\eta_{2n},
1588: \eta_{2n+1})$ are chosen independently each with distribution $\mu$ and
1589: otherwise $(\eta_{2n-1}, \eta_{2n})$ are chosen independently each with
1590: distribution $\mu$.
1591: Let $\nu$ be the law of $\eta$.
1592: If $\ord := \Z^-$, then $\mu_p \strle \nu$ iff $p \le 1/2$, while
1593: if $\ord := \Z^+$, then $\mu_p \strle \nu$ iff $p \le 4/7$.
1594: \endprocl
1595: 
1596: \procl x.ustH-dom
1597: Let $f$ be as in \ref x.ustH/, so that $\P^f$ is the law of the horizontal
1598: edges of a uniform spanning tree in the square lattice $\Z^2$.
1599: By \ref b.Kas:dim/ or \ref b.Montroll/, we have
1600: $$
1601: \int_{\T^2} \log 4(\sin^2 \pi x + \sin^2 \pi y) \,d\lambda_2(x, y)
1602: =
1603: {4\cat \over \pi}
1604: =
1605: 1.1662^+
1606: \,,
1607: \label e.catalan
1608: $$
1609: where $\cat$ is again Catalan's constant \ref e.defcat/.
1610: As we have shown in \ref x.ustH_x-dom/, $\GM(\sin^2 \pi x)=1/4$,
1611: whence by \ref e.catalan/,
1612: $\GM(f) = e^{-4\cat/\pi} =\GM(\bfo - f)$, 
1613: where we are using the observation that $1-f(x, y) = f(y,x)$. 
1614: (This last identity has a combinatorial reason arising from planar dual
1615: trees.) 
1616: Therefore $\mu_p \preccurlyeq \P^f\preccurlyeq \mu_{1-p}$ for $p :=
1617: e^{-4\cat/\pi} = 0.3115^+$ and this $p$ is optimal (on each side).
1618: This result in itself is rather surprising and it would be fascinating to see
1619: an explicit monotone coupling.
1620: \endprocl
1621: 
1622: For $f:\Td\to [0,1]$ measurable and
1623: $r = (r_1, r_2, \ldots, r_d) \in \Z^d$ with all $r_j > 0$, define
1624: $$
1625: f_r(t) := {1 \over \prod_{j=1}^d r_j} \sum_{x \in r^{-1} t} f(x)
1626: \,,
1627: $$
1628: where $r^{-1} (t_1, t_2, \ldots, t_d) := \{ (x_1, x_2, \ldots, x_d) \in \Td
1629: \st \all j\ r_j x_j = t_j \}$.
1630: Write $r \to\infty$ to mean that $\min r_j \to\infty$.
1631: The following is a straightforward extension of \ref t.dom-mix/.
1632: We leave its proof to the reader.
1633: 
1634: \procl t.dom-mix-d
1635: Let $f : \Td \to [0, 1]$ be measurable.
1636: If $\GM(f) > 0$ or $f$ is positive on a non-empty open set,
1637: then there exist constants $p_r \to \widehat f(\bfz)$ as $r \to\infty$
1638: such that $\P^{f_r}
1639: \succcurlyeq \P^{p_r}$ for all $r$.
1640: Therefore, if $\GM\big(f(\bfo - f)\big) > 0$ or if $f$ is continuous and not
1641: equal to $\bfz$ nor $\bfo$, then there exist constants $p_r, q_r \to \widehat
1642: f(\bfz)$ such that $\P^{p_r} \preccurlyeq \P^{f_r} \preccurlyeq \P^{q_r}$ for
1643: all $r$.
1644: \endprocl
1645: 
1646: Consider now a domination property even stronger than our previously defined
1647: strong domination. 
1648: Let $\others$ be the $\sigma$-field generated by $\eta(k)$ for $k \ne \bfz$.
1649: 
1650: \procl d.fulldom
1651: A stationary process $\{\eta_n\}_{n\in\Z^d}$ with distribution $\nu$
1652: {\bf fully dominates} $\mu_p$, written 
1653: $\mu_p\fullle\nu$,
1654: if 
1655: $$
1656: \P[\eta_\bfz=1 \mid \others]\ge p \quad\nu\hbox{-a.s.}
1657: $$
1658: In this situation, we also say that $\nu$ is {\bf uniformly insertion tolerant
1659: at level} $p$.
1660: Similarly, we define $\nu\fullle\mu_p$ if the above inequality holds
1661: when $\ge$ is replaced by $\le$, and
1662: say that $\nu$ is {\bf uniformly deletion tolerant at level} $1-p$.
1663: We say that $\nu$ is {\bf uniformly insertion tolerant} if
1664: $\mu_p\fullle\nu$ for some $p > 0$ and
1665: that $\nu$ is {\bf uniformly deletion tolerant}
1666: if $\nu\fullle\mu_p$ for some $p < 1$.
1667: \endprocl
1668: 
1669: We show that the optimal level of uniform insertion tolerance of a
1670: determinantal process $\P^f$
1671: is the {\bf harmonic mean} of $f$, defined as 
1672: $$
1673: \HM(f) := \left(\int_\Td {d\lambda_d  \over f} \right)^{-1}
1674: \,.
1675: $$
1676: Note that $\HM(f) = 0$ iff $1/f$ is not integrable.
1677: 
1678: \procl t.fulldom
1679: For any measurable $f : \Td \to [0, 1]$, we have
1680: $\mu_p\fullle\P^f$ iff $p\le \HM(f)$. Similarly,
1681: $\P^f\fullle\mu_q$ iff $q\ge 1-\HM(\bfo-f)$.
1682: \endprocl
1683: 
1684: \proof
1685: By \ref p.JNRD/ and \ref t.closest/, we have
1686: \begineqalno
1687: \essinf \P^f\Big[\eta(\bfz) = 1 \mid \others\Big] 
1688: &=
1689: \inf \Big\{ \P^f[\eta(\bfz) = 1 \mid \eta \restrict A \equiv 1] \st
1690: A \subset \Zd \setminus \{ \bfz \} \hbox{ is finite } \Big\}
1691: \cr&=
1692: \P^f[\eta(\bfz) = 1 \mid \eta \restrict B \equiv 1]
1693: =
1694: \min \left\{ \int |\bfo - T|^2 f \,d\lambda_d \st T \in [B]_f \right\}
1695: \, ,
1696: \endeqalno
1697: where $B := \Zd \setminus \{ \bfz \}$.
1698: As shown by \refbmulti{Kolmog:sshs,Kolmog:iesrs} for $d=1$, the latter equals
1699: $\HM(f)$.
1700: The proof extends immediately to general $d$.
1701: This proves the first assertion. The second follows by symmetry.
1702: \Qed
1703: 
1704: \procl r.HM
1705: Since a proof of Kolmogorov's theorem that 
1706: $$
1707: \min \left\{ \int |\bfo - T|^2 f \,d\lambda_d \st T \in [B]_f \right\}
1708: =
1709: \HM(f)
1710: \,,
1711: $$
1712: where $B := \Zd \setminus \{ \bfz \}$,
1713: is difficult to find in readily accessible sources, we
1714: provide one here.
1715: We have
1716: $$
1717: \min \left\{ \int |\bfo - T|^2 f \,d\lambda_d \st T \in [B]_f \right\}
1718: =
1719: \|u\|^2_f\,,
1720: $$
1721: where
1722: $$
1723: u :=
1724: \|P^\perp_{[B]_f} \bfo\|^2_f
1725: \,.
1726: $$
1727: Now $g \perp [B]_f$ iff $g \in L^2(f)$ and $\widehat{gf}(k) = 0$ for all $k
1728: \in B$.
1729: The latter condition holds iff $gf$ is a constant.
1730: If $g \ne \bfz$, then we deduce that $1/f \in L^2(f)$, i.e., $\HM(f) > 0$.
1731: Therefore,
1732: $$
1733: [B]^\perp_f = \cases{0 &if $\HM(f) = 0$,\cr
1734:                \C /f &if $\HM(f) > 0$.\cr
1735:                }
1736: $$
1737: Hence $u = \bfz$ if $\HM(f) = 0$, while otherwise, 
1738: $$
1739: \|u\|^2_f = |\ip{\bfo, \sqrt{\HM(f)}/f}_f|^2
1740: =
1741: \HM(f)
1742: \,,
1743: $$
1744: as desired.
1745: \endprocl
1746: 
1747: 
1748: \procl r.strongnotfull
1749: It is easy to find $f$ such that $\GM(f)> 0$ and $\HM(f)= 0$.
1750: Indeed, a natural such example is the function $g$ of \ref x.ustHx/.
1751: For any such $f$, the corresponding process
1752: $\P^f$ strongly dominates a
1753: nontrivial product measure, but does not
1754: fully dominate any nontrivial product measure. In addition, for any function
1755: $f$ such that $\HM(f) > 0$
1756: and $f$ is not constant a.e., $\GM(f)> \HM(f)$ (as a consequence of
1757: Jensen's inequality), so that $\P^f$ will strongly dominate strictly more
1758: product measures than it will fully dominate.
1759: \endprocl
1760: 
1761: \procl x.ustH-delete
1762: Let $f$ be as in \ref e.ustH-d/, so that $\P^f$ is the law of the edges of the
1763: uniform spanning forest in $\Z^d$ that lie parallel to the $x_1$-axis.
1764: Then $\P^f$ is uniformly deletion tolerant 
1765: iff $d \ge 4$.
1766: This is because $1/(\bfo - f)$ is integrable iff $d \ge 4$.
1767: For example, when $d=4$, we obtain full domination by $\mu_p$ with $p :=
1768: 0.66425^-$, where we have calculated $\HM(\bfo-f)$ as follows. First, we have
1769: $$
1770: 1/(\bfo - f) = 
1771: 1 + {\sin^2 \pi x_1 \over \sum_{j=2}^4 \sin^2 \pi x_j}
1772: \,,
1773: $$
1774: whence 
1775: \begineqalno
1776: \int_{\T^4} 1/(\bfo - f) \,d\lambda_4
1777: &=
1778: 1 + {1 \over 2} \int_{\T^3} {1 \over \sum_{j=2}^4 \sin^2 \pi x_j}
1779: \,d\lambda_3(x_2, x_3, x_4)
1780: \cr&=
1781: 1 + {1 \over 2} \int_{\T^2} {1 \over \sqrt{\left(\sum_{j=2}^3 \sin^2 \pi x_j
1782: \right) \left( 1 + \sum_{j=2}^3 \sin^2 \pi x_j\right)}} \,d\lambda_2(x_2, x_3)
1783: \,.
1784: \cr
1785: \endeqalno
1786: This last integral has no simple form and is calculated numerically.
1787: This gives the value reported for $1 - \HM(\bfo - f)$.
1788: For large $d$, we have $1 - \HM(\bfo - f) \sim 1/d$. Indeed, write
1789: $\lambda_{d-1}^*$ for Lebesgue measure on $\R^{d-1}$. Letting
1790: $$
1791: A_d:=\int_{\T^{d-1}} {d - 1 \over \sum_{j=2}^{d} 2 \sin^2 \pi x_j}
1792: \,d\lambda_{d-1}(x_2, \ldots, x_d)\,,
1793: $$
1794: we have that 
1795: $$
1796: (d-1) (1 - \HM(\bfo - f))= {A_d\over 1+A_d/(d-1)}
1797: \,.
1798: $$
1799: Hence it suffices to show that $\lim_{d\to\infty}A_d=1$.
1800: To show this, note that
1801: \begineqalno
1802: A_d
1803: &=
1804: \int_0^\infty \lambda_{d-1}\Big[ \sum_{j=2}^{d} 2 \sin^2 \pi x_j <
1805: (d-1)/t\Big]
1806: \,dt
1807: \cr&\le
1808: \int_0^2 \lambda_{d-1}\Big[ \sum_{j=2}^{d} 2 \sin^2 \pi x_j < (d-1)/t\Big] \,dt
1809: \cr&\quad+ \int_2^{10(d-1)} 
1810: \lambda_{d-1}\Big[ \sum_{j=2}^{d} 2 \sin^2 \pi x_j < (d-1)/2\Big] \,dt
1811: \cr&\quad+ \int_{10(d-1)}^\infty
1812: \lambda_{d-1}^*\Big[ \all j |x_j| < 1/2 \hbox{ and }
1813: \sum_{j=2}^{d} 8 x_j^2 < (d-1)/t\Big] \,dt
1814: \cr&\to
1815: \int_0^2 \I{[0, 1]}(t) \,dt
1816: =
1817: 1
1818: \endeqalno
1819: as $d \to \infty$, where we have used the weak law of large
1820: numbers and the bounded convergence theorem for the first piece,
1821: a standard large-deviation result for the second piece, and an easy 
1822: estimate on the third piece. The reverse inequality obtains by using only
1823: the first piece. Thus, $A_d \to 1$, as desired.
1824: This value of $1 - \HM(\bfo - f)$,
1825: which gives a full domination upper bound on $\P^f$, should be
1826: compared to $\widehat f(\bfz) = 1/d$ (by symmetry), which is the
1827: probability of a 1 at a site.
1828: On the other hand, for no $d$ is the process uniformly insertion tolerant
1829: since $1/f$ is never integrable.
1830: We remark that for the full uniform spanning forest measure on $\Zd$
1831: (considering edges in all directions), we have change intolerance, meaning
1832: that $\P[\eta(\bfz) = 1 \mid \others] \in \{ 0, 1 \}$ a.s. This follows from a
1833: result of \BLPSusf, as explained by \ref b.HeicklenLyons/.
1834: \endprocl
1835: 
1836: \procl x.ust-delete
1837: Let $g(x)$ be as in \ref x.ustHx/, so that
1838: the edges of the uniform spanning tree in the plane that
1839: lie on the $x$-axis have the law $\P^g$.
1840: We have $\P^g \fullle \mu_p$ for $p := (1+\pi)/(1+2\pi) = 0.56865^+$, and this
1841: is optimal.
1842: This is because 
1843: \begineqalno
1844: \int_\T 1/(1-g) \,d\lambda 
1845: &=
1846: \int_\T \left(1 + \sin^2 \pi x  + \sin \pi x \sqrt{1 + \sin^2 \pi x }\right)
1847: \,d\lambda(x)
1848: \cr&=
1849: 3/2 + \int_\T \sin \pi x \sqrt{1 + \sin^2 \pi x } \,d\lambda(x)
1850: \,.
1851: \cr
1852: \endeqalno
1853: An antiderivative of this integrand is 
1854: $$
1855: -{1 \over \pi} \arctan\left({\cos\pi x \over \sqrt{1 + \sin^2\pi x}}\right)
1856: -{1 \over 2\pi} \cos\pi x\sqrt{1 + \sin^2\pi x}
1857: \,,
1858: $$
1859: whence the remaining integral is $1/2+1/\pi$. Therefore $\int_\T 1/(1-g)
1860: \,d\lambda = 2+1/\pi$ and $1 - \HM(\bfo-g) = (1+\pi)/(1+2\pi)$, as desired.
1861: Observe that $\int 1/g =\infty$ and so the process is not uniformly
1862: insertion tolerant.
1863: Similarly, for the edges lying on the $x$-axis of the uniform spanning tree in
1864: 3 dimensions, we have full domination by $\mu_p$ with $p := 0.37732^+$.
1865: Here, the law of these edges is $\P^h$, where 
1866: $$
1867: h(x) :=
1868: \int_{\T^2} f(x, y, z) \,d\lambda_2(y, z)
1869: =
1870: \int_\T {\sin^2 \pi x \over \sqrt{\left(\sin^2 \pi x + \sin^2 \pi y\right)
1871: \left(1 + \sin^2 \pi x + \sin^2 \pi y\right)}} \,d\lambda_1(y)
1872: \,.
1873: $$
1874: This integral has no simpler form, so to compute $p := 1 - \HM(\bfo - h)$, we
1875: calculated $h$ numerically and used the result to calculate the harmonic mean
1876: numerically. Since $\int 1/h =\infty$, as is easily checked,
1877: this process is not uniformly insertion tolerant.
1878: Similarly, one can check that the process of edges lying on the $x$-axis of the 
1879: uniform spanning tree in $d\ge 4$ dimensions
1880: is not uniformly insertion tolerant.
1881: \endprocl
1882: 
1883: \procl x.USTzz-dom
1884: Let $f$ be as in \ref x.USTzz/.
1885: It turns out that $\GM(f) = e^{-2\cat/\pi}/\sqrt2 = 1 - \GM(\bfo-f) =   
1886: 0.39467^+$. This gives strong domination inequalities.
1887: It is easy to see that $\HM(f) = \HM(\bfo - f) = 0$, so there are no
1888: nontrivial full domination inequalities.
1889: The interest of this function is that it describes the process of edges
1890: along a zig-zag path in the plane for the uniform spanning tree, as claimed
1891: in \ref x.USTzz/. 
1892: We now sketch how to prove this.
1893: The Transfer Current Theorem of \ref b.BurPem/ allows one to calculate, via
1894: determinants, the law for any set of possible edges of the uniform spanning
1895: tree.
1896: Thus, it is enough to verify that for the edges belonging to the zig-zag
1897: path, the matrix entries are those of the Toeplitz matrix associated to
1898: the Fourier coefficients given in \ref x.USTzz/.
1899: For even $k$, the values of $\widehat f(k)$ are given in
1900: \ref b.Lyons:book/.
1901: (These values imply the astonishing
1902: fact that the edges in the plane that lie along a diagonal, e.g., the horizontal
1903: edges with left endpoints $(n, n)$ ($n \in \Z$), are independent, i.e., have
1904: law $\mu_{1/2}$.)
1905: Thus, it remains to treat the case of odd $k$.
1906: A straightforward application of the Transfer Current Theorem gives that
1907: the matrix entry corresponding to the edges $e_0$ and $e_{2k-1}$ is 
1908: \begineqalno
1909: &\int {e(s) + e(t) - e(s+t) - 1 \over 4 - e(s) - e(-s) - e(t) - e(-t)}
1910: e(k s + k t) \,d\lambda_2(s, t)
1911: \cr&\qquad\qquad=
1912: \int {e(s) + e(x-s) - e(x) - 1 \over 4 - e(s) - e(-s) - e(x-s) - e(s-x)}
1913: e(k x) \,d\lambda_2(s, x)
1914: \,,
1915: \endeqalno
1916: where $e(x) := e^{2 \pi i x}$.
1917: Evaluate the integral in $s$ for fixed $x$ by a contour integral 
1918: $$
1919: {1 \over 2 \pi i} \oint {z + e(x) z^{-1} - e(x) - 1 \over 4 - z - z^{-1} -
1920: e(x) z^{-1} - e(-x) z} {d z \over z}
1921: $$
1922: over the contour $|z| = 1$.
1923: The integrand has poles inside the unit disc at $z = 0$ and 
1924: $$
1925: z = {2-\sqrt{4-|1+e(x)|^2} \over 1+e(-x)}
1926: \,.
1927: $$
1928: After use of the residue theorem and integrating in $x$, one obtains
1929: $\widehat f(2k-1)$, as desired.
1930: \endprocl
1931: 
1932: 
1933: 
1934: 
1935: We close this section by
1936: describing how our domination results can be interpreted in terms of
1937: prediction and interpolation questions for wide-sense stationary 
1938: processes. In view of \ref t.closest/ and the well-known correspondence
1939: between prediction and Szeg\H{o} infima, for $d=1$,
1940: $\P^f[\eta(\bfz) = 1 \mid \eta \restrict \{-1,-2,\dots\} \equiv 1]$
1941: is exactly the 
1942: mean squared error for the 
1943: best linear predictor of $Y_0$ given $\Seq{Y_n}_{n\le -1}$,
1944: where 
1945: $\Seq{Y_n}$ is a wide-sense stationary process with spectral density $f$.
1946: Similarly,
1947: $\P^f[\eta(\bfz) = 0 \mid \eta \restrict \{-1,-2,\dots\} \equiv 0]$
1948: is exactly the mean squared error for the best linear predictor
1949: of $Y_0$ given $\Seq{Y_n}_{n\le -1}$, where 
1950: $\Seq{Y_n}$ is now a wide-sense stationary process with spectral density 
1951: $\bfo-f$. This gives us a correspondence between strong domination and 
1952: prediction. An analogous correspondence holds between
1953: full domination and interpolation, where one instead looks at the 
1954: mean squared error for the 
1955: best linear predictor of $Y_0$ given $\Seq{Y_n}_{n\neq 0}$.  
1956: 
1957: \bsection{Entropy}{s.ent}
1958: 
1959: We assume the reader is familiar with the definition of the entropy
1960: $H(\mu)$ of a process $\mu$, as well as basic results concerning entropy (see
1961: \ref b.Walters/ and \ref b.KatzWeiss/).  
1962: Because of Ornstein's theorem (and its generalizations, see \ref
1963: b.KatzWeiss/, \ref b.Conze/, \ref b.Thouvenot/, and \ref b.OrnWeiss/)
1964: that entropy characterizes
1965: Bernoulli shifts up to isomorphism, the following question is particularly
1966: interesting:
1967: 
1968: \procl q.entropy
1969: What is $H(\P^f)$?
1970: \endprocl
1971: 
1972: We know the answer only in the trivial case where $f$ is a constant and in
1973: case $f$ or $\bfo - f$ is the reciprocal of a trigonometric polynomial of
1974: degree 1. In principle, as we shall see, one can also determine the entropy
1975: when $f$ or $\bfo - f$ is the reciprocal of any trigonometric polynomial,
1976: but the formula would be rather unwieldy.
1977: 
1978: In general, then, we shall discuss how to estimate the entropy of $\P^f$.
1979: The definition of entropy always provides upper bounds, due to
1980: subadditivity, so the harder bound is the lower bound.
1981: As we shall see, reciprocals of trigonometric polynomials can be used to
1982: get arbitrarily close lower (and upper) bounds.
1983: Unfortunately, that method is not practical for precise computation.
1984: Nevertheless, in many cases, we have
1985: another method that appears to work quite well.
1986: Indeed, our method seems to provide
1987: upper bounds that converge more quickly than does use of the definition.
1988: 
1989: \ref b.ShiTak:II/ proved that $H(\P^f) > 0$ for all $f \ne \bfz, \bfo$.
1990: Of course, this also follows immediately from our \ref t.Bern/.
1991: 
1992: Let
1993: $$
1994: H[p]:= H(\mu_p)= -p\log p-(1-p)\log(1-p)
1995: \,.
1996: $$
1997: \ref t.dom-d/ yields easy  
1998: lower bounds on the entropy of those processes $\P^f$ such that $f(\bfo-f)$
1999: has a strictly positive geometric mean. We shall obtain more refined lower
2000: bounds later in the one-dimensional case. It is easy to see that
2001: $$
2002: \mu_p \strle \mu \strle \mu_{1-p}
2003: \implies
2004: H(\mu)\ge H[p]
2005: \,.
2006: \label e.doment
2007: $$
2008: By \ref t.dom-d/ and \ref e.doment/, we deduce the
2009: following lower bound on entropy:
2010: 
2011: \procl p.entropy-d 
2012: For any measurable $f : \Td \to [0, 1]$, we have
2013: $$
2014: H(\P^f)\ge \min\Big\{H\big[\GM(f)\big],H\big[\GM(\bfo-f)\big]\Big\}
2015: \,.
2016: $$
2017: \endprocl 
2018: 
2019: \procl r.AMGM
2020: This can be compared to the trivial upper bound $H(\P^f) \le H\big[\widehat
2021: f(0)\big]$. Note that $\widehat f(0)$ is the arithmetic mean of $f$.
2022: \ref b.ShiTak:II/ also note this upper bound and provide two lower bounds:
2023: $H(\P^f) \ge H\big[\widehat f(0)\big]/2$ and
2024: $H(\P^f) \ge \int_{\Td} H[f(x)] \,d\lambda_d(x)$.
2025: \endprocl
2026: 
2027: \procl r.support
2028: It is interesting that \ref e.doment/
2029: is not true if $\strle$ is replaced by $\preccurlyeq$.
2030: For example, let $(X_{2i},X_{2i+1})$ be, independently for 
2031: different $i$, $(1,1)$ or $(0,0)$, each with probability $1/2$.
2032: If $\nu$ is the distribution of this process, then $\nu$ is not
2033: stationary, but $\mu:=\big(\nu +T(\nu)\big)/2$ is stationary, where $T$ is the
2034: shift.
2035: Next, $H(\mu)=\log 2/2 < H[1/\sqrt{2}]$, even though 
2036: $\mu_{1-1/\sqrt{2}} \preccurlyeq \mu \preccurlyeq \mu_{1/\sqrt{2}}$,
2037: as is easily verified.
2038: We also observe that this measure does not even have full support. 
2039: \endprocl
2040: 
2041: \procl x.ustH-ent
2042: Let $f$ be as in \ref x.ustH/, so that $\P^f$ is the law of the horizontal
2043: edges of a uniform spanning tree in the square lattice $\Z^2$.
2044: It is known that the entropy of the entire uniform spanning tree measure (both
2045: horizontal and vertical edges included) is \ref e.catalan/;
2046: see \ref b.BurPem/. 
2047: This can be compared to the bound on $H(\P^f)$ that
2048: \ref p.entropy-d/ provides, together with the calculations of
2049: \ref x.ustH-dom/, namely,
2050: $
2051: H(\P^f) \ge H[e^{-4\cat/\pi}] = 0.6203^+
2052: $.
2053: (The results are much worse if one uses the bounds of \ref b.ShiTak:II/
2054: reported in \ref r.AMGM/ above.)
2055: Direct calculation using cylinder events corresponding to a 4-by-4 block
2056: gives an upper bound for the entropy $H(\P^f) \le 0.68864$.
2057: Of course, the vertical edges of the uniform spanning tree measure have the
2058: same entropy as the horizontal edges.
2059: \endprocl
2060: 
2061: \procl x.ust_H-x
2062: Let $g(x)$ be as in \ref x.ustHx/, so that
2063: the edges of the uniform spanning tree in the plane that
2064: lie on the $x$-axis have the law $\P^g$.
2065: In view of \ref x.ustH_x-dom/,
2066: \ref p.entropy-d/ implies a lower
2067: bound of $H(\P^g) \ge H[\sqrt 2 - 1] \ge 0.67835$.
2068: Also, direct calculation using cylinder events of length 16 gives an upper
2069: bound $H(\P^g) \le 0.69034$.
2070: In \ref x.ustH-x-ent/, we shall obtain more refined bounds on the entropy.
2071: It is not hard to see that the horizontal edges associated to the $y$-axis
2072: have law $\P^{\bfo - g}$.
2073: \endprocl
2074: 
2075: Although not relevant for our determinantal probability measures
2076: (where we have seen that $\strle$ and $\preccurlyeq$ are equivalent),
2077: it is interesting to ask whether 
2078: $\mu_p \preccurlyeq \mu \preccurlyeq \mu_{1-p}$ with $p > 0$
2079: yields any lower bound on the entropy $H(\mu)$ for general $\mu$. 
2080: We first ask the question whether $\mu_p \preccurlyeq \mu$
2081: with $p > 0$ implies that
2082: $H(\mu)> 0$ provided $\mu\neq \delta_\bfo$. The answer to
2083: this question is affirmative
2084: since it is known (see \ref b.furst/ for $d=1$ and \ref b.GTW/ for $d > 1$)
2085: that zero-entropy processes are {\bf disjoint} from i.i.d.\
2086: processes (meaning that there are no stationary couplings of them other than
2087: independent couplings).
2088: The next result provides an explicit lower bound on the entropy
2089: for processes trapped between two i.i.d.\ processes.
2090: The proof was obtained jointly with Chris Hoffman.
2091: 
2092: \procl p.hoffman
2093: For any $d\ge 1$, if
2094: $\mu_p \preccurlyeq \mu \preccurlyeq \mu_{1-p}$ with $p >0$, then
2095: $$
2096: H(\mu)\ge \max{\{a_p,b_p\}} >0
2097: \,,
2098: $$
2099: where
2100: $$
2101: a_p:=
2102: (1-p)\log\Big({1\over  1-p}\Big)-{1-2p \over 2}\log\Big({1\over
2103: 1-2p}\Big)
2104: $$
2105: and
2106: $$
2107: b_p:=
2108: 2(1-p)\log\Big({1\over  1-p}\Big)-(1-2p)\log\Big({1\over
2109: 1-2p}\Big)-(1-2p)\log 2.
2110: $$
2111: \endprocl
2112: 
2113: \procl r.log2
2114: Observe that $b_p$ approaches $\log 2$ as $p\to 1/2$.
2115: \endprocl
2116: 
2117: \proof
2118: Let $X$, $Y$ and $Z$ denote processes with respective distributions 
2119: $\mu_p,\mu$ and $\mu_{1-p}$. We construct a joining (stationary coupling)
2120: of all three processes as follows. Consider any joining $m_1$ of $X$ and $Y$
2121: with $X_i\le Y_i$ a.s.\ 
2122: and any joining $m_2$ of $Y$ and $Z$ with $Y_i\le Z_i$ a.s. We now pick a 
2123: realization for $Y$ according to $\mu$ and then choose $X$ and $Z$ 
2124: (conditionally) independently using $m_1$ and $m_2$ respectively (this
2125: is called the fibered product of $m_1$ and $m_2$ over $Y$). This gives
2126: us a joining $(X',Y',Z')$ of $X$, $Y$ and $Z$.
2127: We may now assume that $(X, Y, Z) = (X',Y', Z')$.
2128: 
2129: We now use standard facts about entropy.
2130: First, $H(X,Z)\ge H(X,Y,Z)- H(Y)$. We next note, using
2131: the conditional independence of $X$ and $Z$ given $Y$, that
2132: \begineqalno
2133: H(X,Y,Z)
2134: &= H(Y) + H(X,Z\mid Y)=
2135: H(Y) + H(X\mid Y)+H(Z\mid Y)
2136: \cr&\ge  H(X)+H(Z)-H(Y)\,. 
2137: \cr
2138: \endeqalno
2139: Hence
2140: $$
2141: H(X,Z)\ge H(X)+H(Z)-2H(Y)=2H[p]-2H(Y)\,.
2142: $$
2143: Since $X\le Z$, the one-dimensional marginal of $(X,Z)$ necessarily
2144: has 3 atoms of weights $p,1-2p$ and $p$. It follows that
2145: $$
2146: H(X,Z)\le 2p\log\Big({1\over p}\Big)+(1-2p)\log\Big({1\over 1-2p}\Big)\,.
2147: $$
2148: Combining this with the previous inequality shows that 
2149: $H(\mu)\ge a_p$. 
2150: 
2151: We modify the above proof to show that $b_p$ is also
2152: a lower bound. Rather than using $H(X,Z)\ge H(X,Y,Z)- H(Y)$, we
2153: use $H(X,Z)= H(X,Y,Z)- H(Y\mid X,Z)$. The earlier argument then gives
2154: $$
2155: H(X,Z)\ge 2H[p]-H(Y)- H(Y\mid X,Z)\,.
2156: $$
2157: Using the same upper bound on $H(X,Z)$ as before and the trivial upper bound
2158: for $H(Y\mid X,Z)$ of $(1-2p)\log 2$ (obtained by noting that $Y_i$ is
2159: determined by $X_i$ and $Z_i$ when $X_i=Z_i$) yields the lower bound of $b_p$.
2160: 
2161: The easiest way to check the strict positivity of $a_p$ is to
2162: observe that
2163: $$
2164: 2H[p] > 2p\log\Big({1\over p}\Big)+(1-2p)\log\Big({1\over 1-2p}\Big)
2165: \,.
2166: $$
2167: This, in turn, follows from the fact that the left-hand side is the entropy of
2168: the independent joining $\mu_p \times \mu_{1-p}$, while the right-hand side is
2169: the entropy of the monotonic coordinatewise-independent joining of $\mu_p$
2170: and $\mu_{1-p}$, whose existence follows from the inequality 
2171: $p \le 1-p$.
2172: \Qed
2173: 
2174: There are few cases where we know the entropy $H(\P^f)$ exactly.
2175: Besides the case when $f$ is constant, we can calculate the entropy when $f$
2176: is the reciprocal of a trigonometric polynomial of degree 1.
2177: We shall illustrate this in the simplest case, \ref x.renewal/.
2178: Thus, let $0 < a < 1$, $d = 1$, and $f(x) := (1 - a)^2/|e^{2\pi i x} - a|^2$.
2179: Recall that that for any invariant measure $\mu$ in one dimension, we have
2180: $$
2181: H(\mu) =
2182: \int H\big[\mu[\eta_0 = 1 \mid \eta_{-1}, \eta_{-2}, \ldots]\big]
2183: \,d\mu(\eta_{-1}, \eta_{-2}, \ldots)
2184: \,.
2185: \label e.ent-past
2186: $$
2187: Now by \ref e.renewal-distribution/, it is easy to check that
2188: $$
2189: P^f[\eta_0 = 1 \mid \eta_{-1}, \eta_{-2}, \ldots] =
2190: {(1 - a)^2 N \over N - (N - 1) a}
2191: \,,
2192: $$
2193: where $N := \min \{ k\ge 1 \st \eta_{-k} = 1 \} $.
2194: In addition, we may easily show that for $n\ge 1$,
2195: $$
2196: \P^f[N = n] = {(1-a)(n - (n-1)a) a^{n-1} \over 1+a}
2197: \,.
2198: $$
2199: Therefore 
2200: \begineqalno
2201: H(\P^f) 
2202: &=
2203: \sum_{n=1}^\infty 
2204: {(1-a)(n - (n-1)a) a^{n-1} \over 1+a}
2205: \Big(
2206: - {(1 - a)^2 n \over n - (n - 1) a} \log {(1 - a)^2 n \over n - (n - 1) a}
2207: \cr&\hskip1in
2208: - {n - (n - 1) a - (1 - a)^2 n \over n - (n - 1) a} 
2209: \log {n - (n - 1) a - (1 - a)^2 n \over n - (n - 1) a}
2210: \Big)
2211: \cr&=
2212: {1-a \over 1+a}
2213: \sum_{n=1}^\infty a^{n-1}
2214: \Big(
2215: - (1 - a)^2 n \log [(1 - a)^2 n]
2216: \cr&\hskip1.3in
2217: - [a(1 - a) n + a] \log [a(1 - a) n + a]
2218: \cr&\hskip1.3in
2219: + [(1-a) n + a] \log [(1-a) n + a]
2220: \Big)
2221: \,.
2222: \cr
2223: \endeqalno
2224: 
2225: In principle, one can also write an infinite series of positive terms for the
2226: entropy of $\P^f$ when $f$ is equal to the reciprocal of any trigonometric
2227: polynomial, since, by \ref p.regenerate/, the process $\P^f$ is then
2228: a regenerative process.
2229: Of course, the answer will be much more unwieldy than the above formula.
2230: However, it can be used to get arbitrarily good lower bounds on the entropy of
2231: any process $\P^f$, in theory.
2232: To see this, 
2233: we use the following lemma, which is more or less Lemma 5.10 in 
2234: \ref b.Rudolph/. As our proof is somewhat shorter and gives a more precise
2235: bound, we include it. See also \ref b.BurPem/, Lemma 6.2, for a similar
2236: proof.
2237: 
2238: \procl l.rudolph
2239: For any stationary processes $\mu$ and $\nu$, we have
2240: $|H(\mu)-H(\nu)| \le H[\dbar(\mu,\nu)]$.
2241: \endprocl
2242: 
2243: \proof
2244: Consider a
2245: stationary process $(X,Y)=\Seq{(X_i,Y_i)}_{i\in\Z^d}$,
2246: where $X=\Seq{X_i}_{i\in\Z^d}$ has distribution $\mu$,
2247: $Y=\Seq{Y_i}_{i\in\Z^d}$ has distribution $\nu$, and
2248: $P(X_0\neq Y_0)=\dbar(\mu,\nu)$. 
2249: Let $Z_i := X_i + Y_i$ mod 2. Then 
2250: $$
2251: H(\mu) \le H(X, Z) = H(Y, Z) \le H(\nu) + H(Z)        
2252: \le H(\nu) + H[\dbar(\mu, \nu)]
2253: \,.
2254: $$
2255: Since the same holds with $\mu$ and $\nu$ switched, the result follows.
2256: \Qed
2257: 
2258: 
2259: 
2260: Thus, given any measurable $f : \T \to [0, 1]$ and any $\epsilon > 0$, define
2261: $\delta$ to be the smallest positive number 
2262: so that $H[\delta] = \epsilon$.
2263: Then choose a trigonometric polynomial $T \ge 1$ such that $\int_\T |T^{-1} -
2264: \max(f,\delta/2)| \,d\lambda_1 <\delta/2$.
2265: We can calculate $H(\P^{1/T})$ as closely as desired.
2266: Since $\dbar(\P^{1/T}, \P^f) < \delta$ by \ref p.L1givesdbar/,
2267: it follows by \ref l.rudolph/ that $|H(\P^f) - H(\P^{1/T})| < \epsilon$.
2268: 
2269: Unfortunately,
2270: this method of calculation is hopeless in practice since when $T$ has a high
2271: degree, it will take a very long time to see a renewal, which means that
2272: cylinder events of great length will be needed for the estimation.
2273: When a cylinder event of length $n$ is used, one must calculate $2^n$
2274: probabilities. Thus, this is completely impractical.
2275: 
2276: However, we now exhibit an alternative method that works extremely well in
2277: practice, although we cannot prove that it works well {\it a priori}.
2278: We begin with two simple examples,
2279: $$
2280: f(x) := \I{[0, 1/2]}(x)
2281: \label e.deff
2282: $$
2283: (which also occurred in \ref x.domino/)
2284: and 
2285: $$
2286: g(x) := \sin^2 \pi x = (1 - \cos 2\pi x)/2
2287: \,.
2288: \label e.defg
2289: $$
2290: Since $\widehat f(k) = \widehat g(k) = 0$ for all even $k \ne 0$,
2291: both $\P^{f}$ and $\P^{g}$ have the property that looking at only the even
2292: coordinates, we see independent fair coin flips, i.e., $\mu_{1/2}$.
2293: Therefore the entropy of both processes is at least $(1/2) \log 2$.
2294: In either case, we know of no direct method to prove that strict inequality
2295: holds, but it does.
2296: We first show this for $\P^g$. 
2297: In \ref x.ustH_x-dom/ we showed that $\GM(g) = 1/4$.
2298: By symmetry, $\GM(1-g) = 1/4$.
2299: Since $H[1/4] = 0.56^+ > (1/2) \log 2 = 0.35^-$, we obtain by
2300: \ref p.entropy-d/ that $H(\P^g) > (1/2) \log 2$.
2301: 
2302: In order to show that $H(\P^f) > (1/2) \log 2$ for $f$ as in \ref
2303: e.deff/,
2304: we need a method to
2305: obtain more refined entropy bounds. We illustrate such a method
2306: beginning with this simple function $g$.
2307: While we shall explain afterwards a method to obtain results for more general
2308: functions for the case $d=1$, this first proof contains the essential
2309: idea of this method while at the same time relying on a more elementary 
2310: calculation, and therefore, we feel, is worth including.
2311: 
2312: \procl p.entropyg
2313: With $g$ as in \ref e.defg/, we have 
2314: $$
2315: 0.63^- = {3 \over 8} H[1/4] + {5 \over 8} H[11/28]
2316: \le H(\P^g) \le
2317: {3 \over 8} H[7/20] + {5 \over 8} H[5/12] = 0.67^-
2318: \,.
2319: $$
2320: \endprocl
2321: 
2322: \procl l.GMpert
2323: Let $h : \T \to [0, 1]$ be a trigonometric polynomial of degree at most 1,
2324: i.e., $\widehat h(k) = 0$ for $|k| \ge 2$.
2325: Fix $n > 0$ and $A \subseteq \{ 1, 2, \ldots, n \}$.
2326: For $C \subseteq \{ 1, 2, \ldots, n \}$, let $\lambda(C)$ denote the sequence
2327: of lengths of consecutive intervals in $\{ 1, 2, \ldots, n \} \setminus C$.
2328: Then 
2329: $$
2330: \P^h[\eta \restrict A \equiv 0,\, \eta \restrict (\{ 1, 2, \ldots, n \}
2331: \setminus A) \equiv 1]
2332: =
2333: \sum_{C \subseteq A} (-1)^{|A \setminus C|} \prod_{i \in\lambda(C)} D_{i-1}(h)
2334: \,.
2335: $$
2336: \endprocl
2337: 
2338: \proof
2339: Whenever two sets $F_1, F_2 \subset \Z$ neither overlap nor come within
2340: distance 1 of each other, the configurations on $F_1$ and $F_2$ are
2341: $\P^h$-independent because $\widehat h(k) = 0$ for $|k| \ge 2$.
2342: Therefore $\P^h[\eta \restrict (\{ 1, 2, \ldots, n \} \setminus C) \equiv 1] =
2343: \prod_{i \in\lambda(C)} D_{i-1}(h)$. Now the desired formula follows from the
2344: inclusion-exclusion principle.
2345: \Qed
2346: 
2347: \proofof p.entropyg
2348: Write $d_n := D_{n-1}(g)$.
2349: Recall that 
2350: $$
2351: H(\P^g) =
2352: \int H\big[\P^g[\eta_0 = 1 \mid \eta_{-1}, \eta_{-2}, \ldots]\big]
2353: \,d\P^g(\eta_{-1}, \eta_{-2}, \ldots)
2354: \,.
2355: \label e.ent-past
2356: $$
2357: 
2358: By the negative association property of $\P^g[\;\cdot \mid \eta_{-1} =
2359: \eta_{-2} = 1]$, we have that on the event $ \{ \eta_{-1} = \eta_{-2} = 1 \}$,
2360: \begineqalno
2361: \P^g[\eta_0 = 1 \mid \eta_{-1}, \eta_{-2}, \ldots]
2362: &\ge
2363: \lim_{n \to\infty} \P^g[\eta_0 = 1 \mid \eta_{-1} = \eta_{-2}= \cdots = \eta_{-n} = 1]
2364: \cr&=
2365: \lim_{n \to\infty} {\P^g[\eta_0 = \eta_{-1} = \eta_{-2}= \cdots = \eta_{-n} = 1] \over
2366: \P^g[\eta_{-1} = \eta_{-2}= \cdots = \eta_{-n} = 1]}
2367: \cr&=
2368: \lim_{n \to\infty} d_{n+1}/d_n
2369: =
2370: \GM(g)
2371: =
2372: 1/4
2373: \,.
2374: \endeqalno
2375: Likewise, on the same event, we have 
2376: \begineqalno
2377: \P^g[\eta_0 = 1 \mid \eta_{-1}, \eta_{-2}, \ldots]
2378: &\le
2379: \lim_{n \to\infty} \P^g[\eta_0 = 1 \mid \eta_{-1} = \eta_{-2}= 1,\, \eta_{-3} = 
2380: \cdots = \eta_{-n} = 0]
2381: \cr&=
2382: \lim_{n \to\infty} \P^g[\eta_0 = 0 \mid \eta_{-1} = \eta_{-2}= 0,\, \eta_{-3} = 
2383: \cdots = \eta_{-n} = 1]
2384: \cr&\hskip1in\hbox{[by symmetry]}
2385: \cr&=
2386: 1 - \lim_{n \to\infty} \P^g[\eta_0 = 1 \mid \eta_{-1} = \eta_{-2}= 0,\, \eta_{-3} = 
2387: \cdots = \eta_{-n} = 1]
2388: \cr&=
2389: 1 - \lim_{n \to\infty} {\P^g[\eta_{-1} = \eta_{-2}= 0,\, \eta_0 = \eta_{-3} = 
2390: \cdots = \eta_{-n} = 1] \over
2391: \P^g[\eta_{-1} = \eta_{-2}= 0,\, \eta_{-3} = \cdots = \eta_{-n} = 1]}
2392: \cr&=
2393: 1 - \lim_{n \to\infty} 
2394: {d_1 d_{n-2} - d_2 d_{n-2} - d_1 d_{n-1} + d_{n+1}
2395: \over
2396: d_{n-2} - d_1 d_{n-2} - d_{n-1} + d_{n}}
2397: \cr&\hskip1in\hbox{[by \ref l.GMpert/]}
2398: \cr&=
2399: 1 - 
2400: {d_1 - d_2 - d_1 \GM(g) + \GM(g)^3
2401: \over
2402: 1 - d_1 - \GM(g) + \GM(g)^2}
2403: \cr&=
2404: 7/20
2405: \endeqalno
2406: since $d_1 = 1/2$ and $d_2 = 3/16$.
2407: We may conclude that on the event $ \{ \eta_{-1} = \eta_{-2} = 1 \}$, we have 
2408: $$
2409: H\big[\P^g[\eta_0 = 1 \mid \eta_{-1}, \eta_{-2}, \ldots]\big]
2410:  \in \big[H[1/4], H[7/20]\big]
2411: \,.
2412: $$
2413: By symmetry, the same holds on the event $ \{ \eta_{-1} = \eta_{-2} = 0 \}$.
2414: 
2415: Similarly, we have that on the event $ \{ \eta_{-1} = 1,\, \eta_{-2} = 0 \}$,
2416: \begineqalno
2417: \P^g[\eta_0 = 1 \mid \eta_{-1}, \eta_{-2}, \ldots]
2418: &\ge
2419: \lim_{n \to\infty} \P^g[\eta_0 = 1 \mid \eta_{-2} = 0,\,
2420: \eta_{-1} = \eta_{-3}= \cdots = \eta_{-n} = 1]
2421: \cr&=
2422: \lim_{n \to\infty} {\P^g[\eta_{-2} = 0,\,
2423: \eta_0 = \eta_{-1} = \eta_{-3}= \cdots = \eta_{-n} = 1] \over
2424: \P^g[\eta_{-2} = 0,\, \eta_{-1} = \eta_{-3}= \cdots = \eta_{-n} = 1]}
2425: \cr&=
2426: \lim_{n \to\infty}
2427: {d_2 d_{n-2} - d_{n+1}
2428: \over
2429: d_1 d_{n-2} - d_n}
2430: \cr&=
2431: {d_2 - \GM(g)^3
2432: \over
2433: d_1 - \GM(g)^2}
2434: \cr&=
2435: 11/28
2436: \,.
2437: \endeqalno
2438: Likewise, on the same event, we have 
2439: \begineqalno
2440: \P^g[\eta_0 = 1 \mid \eta_{-1}, \eta_{-2}, \ldots]
2441: &\le
2442: \lim_{n \to\infty} \P^g[\eta_0 = 1 \mid \eta_{-1} = 1,\, \eta_{-2}= \eta_{-3} = 
2443: \cdots = \eta_{-n} = 0]
2444: \cr&=
2445: \lim_{n \to\infty} \P^g[\eta_0 = 0 \mid \eta_{-1} = 0,\, \eta_{-2}= \eta_{-3} = 
2446: \cdots = \eta_{-n} = 1]
2447: \cr&=
2448: 1 - \lim_{n \to\infty} \P^g[\eta_0 = 1 \mid \eta_{-1} = 0,\, \eta_{-2}= \eta_{-3} = 
2449: \cdots = \eta_{-n} = 1]
2450: \cr&=
2451: 1 - \lim_{n \to\infty} {\P^g[\eta_{-1} = 0,\, \eta_0 = \eta_{-2} = 
2452: \cdots = \eta_{-n} = 1] \over
2453: \P^g[\eta_{-1} = 0,\, \eta_{-2} = \cdots = \eta_{-n} = 1]}
2454: \cr&=
2455: 1 - \lim_{n \to\infty} {d_1 d_{n-1} - d_{n+1} \over d_{n-1} - d_n}
2456: \cr&=
2457: 1 - {d_1 - \GM(g)^2 \over 1 - \GM(g)}
2458: \cr&=
2459: 5/12
2460: \,.
2461: \endeqalno
2462: We may conclude that on the event $ \{ \eta_{-1} = 1,\, \eta_{-2} = 0 \}$, we have 
2463: $$
2464: H\big[\P^g[\eta_0 = 1 \mid \eta_{-1}, \eta_{-2}, \ldots]\big]
2465:  \in \big[H[11/28], H[5/12]\big]
2466: \,.
2467: $$
2468: By symmetry, the same holds on the event $ \{ \eta_{-1} = 0,\, \eta_{-2} =
2469: 1 \}$.
2470: 
2471: Putting all these bounds into \ref e.ent-past/ gives the claimed bounds on the
2472: entropy of $\P^g$.
2473: \Qed
2474: 
2475: 
2476: For more general functions, we can get lower bounds on the entropy by using
2477: the asymptotics of \ref b.BumpDiaconis/ or of \ref b.TW:like/,
2478: which serve as replacements for our use of the special \ref l.GMpert/.
2479: In fact, we use an extension of the formula of \ref b.TW:like/ due to
2480: \ref b.Lyons:szego/.
2481: Denote by $\nu_{f}$ the measure on $2^\N$ obtained by the
2482: limiting condition $\lim_{n \to\infty} \P^f[\;\cdot \mid \eta_{-1} = \eta_{-2}
2483: = \cdots = \eta_{-n} = 1]$, which exists by \ref p.JNRD/.
2484: If $\GM(f) > 0$, define
2485: $$
2486: \Phi_f(z) := \exp {1 \over 2} \int_{\T} {e^{2\pi i t} + z
2487: \over e^{2\pi i t} - z} \log f(t) \,d\lambda_1(t)
2488: \label e.defPhi_f
2489: $$
2490: for $|z| < 1$.
2491: The {\bf outer} function
2492: $$
2493: \varphi_f(t) := \lim_{r \uparrow 1} \Phi_f(r e^{2 \pi i t})
2494: \label e.defvarphi_f
2495: $$
2496: exists for $\lambda_1$-a.e.\ $t \in \T$ and satisfies $|\varphi_f|^2 = f$
2497: $\lambda_1$-a.e. 
2498: This limit also holds in $L^1(\T)$.
2499: See \ref b.Rudin:RCA/, Theorem 17.11, p.~340, and Theorem 17.16, p.~343.
2500: As is easily verified,
2501: $$
2502: \log \Phi_f(z) = \widehat F(0) + 2 \sum_{k=1}^\infty \widehat F(k) z^k
2503: \label e.preexp
2504: $$
2505: when $F := (1/2) \log f$.
2506: The mean value theorem for analytic functions and the $L^1(\T)$
2507: convergence above show that
2508: $$
2509: \widehat {\varphi_f}(0) = \Phi_f(0) = \sqrt{\GM(f)}
2510: \,.
2511: $$
2512: Define $\Phi_f := \varphi_f := \bfz$ if $\GM(f) = 0$.
2513: An analytic trigonometric polynomial is (a constant times) an outer
2514: function iff its extension to the unit disc has no zeroes in the open disc.
2515: [On the one hand, the extension of 
2516: no outer function has any zeroes in the open disc since it
2517: is an exponential.  On the other hand, by factoring a polynomial, one
2518: has to check merely that $z \mapsto z-\zeta$ is a constant times
2519: an outer function for
2520: $|\zeta| \ge 1$. Indeed, $\Phi_f(z) = (-|\zeta|/\zeta)(z -\zeta)$
2521: for $f(t) := |e^{2\pi i t} -\zeta|^2$, as can be seen by the Poisson integral
2522: formula for $|\zeta|>1$ (\ref b.Rudin:RCA/, Theorem 11.9, p.~235) and by a
2523: limiting procedure for $|\zeta| = 1$.]
2524: The results of \ref b.Lyons:szego/ show that \ref t.condonB/ reduces to
2525: the following formula:
2526: 
2527: 
2528: \procl t.szego
2529: For any measurable $f : \T \to [0, 1]$, the measure $\nu_{f}$ is equal to
2530: the determinantal probability measure corresponding to the positive
2531: contraction on $\ell^2(\N)$ whose $(j, k)$-matrix entry is 
2532: $$
2533: \sum_{l=0}^{j \wedge k} \overline{\widehat{\varphi_f}(j-l)}
2534: \widehat{\varphi_f}(k-l)
2535: \,.
2536: $$
2537: \endprocl
2538: 
2539: This can be substituted in the appropriate places in the proof of \ref
2540: p.entropyg/.
2541: For example, on the event $ \{ \eta_{-1} = \eta_{-2} = 0 \}$, one has that 
2542: \begineqalno
2543: \P^f[\eta_0 = 1 \mid \eta_{-1}, \eta_{-2}, \ldots]
2544: &\ge
2545: \lim_{n \to\infty}
2546: \P^f[\eta_0 = 1 \mid \eta_{-1} = \eta_{-2}= 0,\, \eta_{-3} = \eta_{-4} =
2547: \cdots = \eta_{-n} = 1]
2548: \cr&=
2549: \nu_{f}[\eta_2=1, \eta_1=\eta_0=0]/ \nu_{f}[\eta_1=\eta_0=0]
2550: \,.
2551: \endeqalno
2552: One can use $\nu_{\bfo-f}$ in a similar fashion for estimating such
2553: probabilities above by $\eta$ that terminate in repeating 0s, rather than in
2554: repeating 1s.
2555: For example, if 
2556: $$
2557: f := |1+\ee_{1} + \ee_{2}|^2/9
2558: \,,
2559: $$
2560: then $\varphi_f = (1+\ee_{1} + \ee_{2})/3$ since the polynomial
2561: $(1+z+z^2)/3$ has no zeroes in the open unit disc.
2562: Similarly, one can show that 
2563: $$
2564: \varphi_{\bfo-f} = {\sqrt 6 + \sqrt 2 \over 6} - {\sqrt 2 \over 3} \ee_{1} 
2565: + {\sqrt 2 - \sqrt 6 \over 6} \ee_{2}
2566: \,.
2567: $$
2568: With these and \ref t.szego/, one can show by following the method of proof of
2569: \ref p.entropyg/ that 
2570: $$
2571: 0.53 \le H(\P^f) \le 0.61
2572: \,.
2573: $$
2574: This method of estimation is relatively easy to program on a computer;
2575: decomposing by cylinder events of length 8 instead of the length 2 used in the
2576: proof of \ref p.entropyg/ and using Mathematica, we find that 
2577: $$
2578: 0.601992  \le H(\P^f) \le 0.602433
2579: \,.
2580: $$
2581: % We also find for $g$ above the range $[0.659062, 0.659087]$.
2582: Similarly, using cylinder events of length 15, we obtain that 
2583: $$
2584: 0.65907716
2585: \le H(\P^g) \le
2586:   0.65907733
2587: \label e.entropy-g
2588: $$
2589: for the function $g$ in \ref e.defg/.
2590: 
2591: We now return to the problem of showing that
2592: $H(\P^f) > (1/2) \log 2$ for $f$ as in \ref e.deff/.
2593: 
2594: \procl p.entrindic
2595: If $f := \I{[0, 1/2]}$, then $H(\P^f) > (1/2) \log 2$.
2596: \endprocl
2597: 
2598: \proof
2599: Let $\widetilde{f}:= .99\I{[0, 1/2]} +.01\I{(1/2,1)}$.
2600: \ref p.L1givesdbar/ tells us that
2601: $\dbar(\P^f, \P^{\widetilde{f}}) \le .01$, and hence from
2602: \ref l.rudolph/,
2603: $H(\P^f) \ge H(\P^{\widetilde{f}})- H[.01]$. 
2604: Now $H[.01] < 0.0561$.
2605: Write $\delta := .01$ and
2606: $L := \log (\delta ^{-1} - 1) = \log 99$.
2607: A simple integration shows that 
2608: $$
2609: \skew6\widehat {\widetilde f}(k) =
2610: \cases{(1/2) &if $k = 0$,\cr
2611: \displaystyle{(-1)^k - 1 \over 2k\pi} i (1 - 2\delta)
2612: &if $k \ne 0$\cr}
2613: $$
2614: (where $i := \sqrt{-1}$).
2615: To find $\varphi_{\widetilde f}$, we proceed as follows.
2616: Let $F := (1/2) \log \widetilde f$.
2617: Then
2618: $$
2619: \widehat F(k) =
2620: \cases{(1/4) \log \delta(1-\delta)  &if $k = 0$,\cr
2621: \displaystyle{(-1)^k - 1 \over 4k\pi} i L
2622: &if $k \ne 0$.\cr}
2623: $$
2624: By using \ref e.preexp/ and exponentiating, we obtain
2625: $$
2626: \Phi_{\widetilde f}(z)
2627: =
2628: [\delta(1-\delta)]^{1/4}
2629: \left(
2630: 1
2631: -
2632: {i \over \pi} L z
2633: -
2634: {1 \over 2\pi^2} L^2 z^2
2635: -
2636: {i \over 6\pi^3} (2\pi^2 - L^2) L z^3
2637: +
2638: \cdots
2639: \right)
2640: \,.
2641: $$
2642: Since the coefficients of this power series are the Fourier coefficients of
2643: $\varphi_{\widetilde f}$,
2644: we have a procedure to compute the Fourier coefficients
2645: $\varphi_{\widetilde f}$.
2646: Since $\widetilde f(1-x) = 1-\widetilde f(x)$, we have
2647: $\overline{\Phi_{\widetilde f}(\overline z)} =
2648: \Phi_{\bfo-\widetilde f}(z)$. 
2649: Thus, the $k$th Fourier coefficient of $\varphi_{\bfo - \widetilde f}$ equals
2650: the complex conjugate of the $k$th Fourier coefficient of $\varphi_{\widetilde
2651: f}$.
2652: Using the method of \ref p.entropyg/ with cylinder events of length 3,
2653: which requires knowledge of Fourier coefficients only for $|k| \le 2$, we
2654: obtain that $H(\P^{\widetilde f}) \ge 0.4105$.
2655: Therefore $H(\P^f) > 0.4105 - 0.0561 = 0.3544 > (1/2) \log 2$.
2656: (If we use cylinder events of length 12, then we obtain the bound $H(\P^f) >
2657: 0.4442$. However, we believe that the true entropy is significantly larger
2658: still.)
2659: \Qed
2660: 
2661: 
2662: \procl x.depends
2663: We can also use these bounds to prove that $H(\P^f)$ does {\it not\/}
2664: depend {\it only\/} on the distribution of $f$.
2665: For example, consider the function 
2666: $$
2667: f(t) := \sin^2(\pi t/2)
2668: $$
2669: on $[0, 1]$, which has the same distribution as the function $g$ of \ref
2670: e.defg/. 
2671: An elementary integration shows that 
2672: $$
2673: \widehat f(k) = \cases{1/2 &if $k=0$,\cr
2674: \displaystyle {2k i \over (2k-1)(2k+1) \pi} &if $k \ne 0$\cr}
2675: $$
2676: (where $i := \sqrt{-1}$).
2677: To find $\varphi_f$, we proceed as follows.
2678: Write $F := (1/2) \log f$.
2679: The real parts of the integrals giving $\widehat F(k)$ can be found in
2680: standard tables, while the imaginary parts are derived in \ref b.LPR/,
2681: giving
2682: $$
2683: \widehat F(k) =
2684: \cases{-\log 2  &if $k=0$,\cr
2685:        \displaystyle -{1 \over 4k} + {i \over k \pi} \sum_{j=1}^k {1 \over 2j-1}
2686:                 &if $k > 0$.\cr}
2687: $$
2688: %The list of the Fourier coefficients $\widehat F(k)$ for $k = 0, 1,
2689: %\ldots, 10$ is
2690: %\begineqalno
2691: %-\log 2,\,
2692: %&-1/4 + i/\pi,\, -1/8 + 2 i/3\pi,\, -1/12 + 23 i/45\pi,\, 
2693: %\cr& -1/16 + 44 i/105\pi,\, -1/20 + 563 i/1575\pi,\, 
2694: %\cr& -1/24 + 3254 i/10395\pi,\, -1/28 + 88069 i/315315\pi,\, 
2695: %\cr& -1/32 + 11384 i/45045\pi,\, -1/36 + 1593269 i/6891885\pi,\, 
2696: %\cr& -1/40 + 15518938 i/72747675\pi
2697: %\,.
2698: %\cr
2699: %\endeqalno
2700: As in the proof of \ref p.entrindic/, we compute
2701: \begineqalno
2702: \Phi_f(z) =
2703: 1/2 &+( -1/4 + i/\pi) z +(-1/16 + i/6\pi - 1/\pi^2) z^2
2704: \cr&  -(1/32 - 19 i/360 \pi + 5/6 \pi^2 + 2 i/3\pi^3) z^3
2705: \cr&  +(-5/256 + 89 i/5040\pi - 27/40 \pi^2 - i/\pi^3 + 1/3 \pi^4) z^4
2706:   + \cdots
2707: \,,
2708: \endeqalno
2709: whose coefficients are the Fourier coefficients of $\varphi_f$.
2710: Since $1-\sin^2(\pi t/2) = \sin^2(\pi (1-t)/2)$, it follows from the
2711: definition \ref e.defPhi_f/ that $\overline{\Phi_{f}(\overline z)} =
2712: \Phi_{\bfo-f}(z)$. Since the power series coefficients of $\Phi_{\bfo - f}$
2713: are the Fourier coefficients of $\varphi_{\bfo - f}$, we obtain that the $k$th
2714: Fourier coefficient of $\varphi_{\bfo - f}$ equals the complex conjugate of
2715: the $k$th Fourier coefficient of $\varphi_f$.
2716: Using the method of \ref p.entropyg/ with cylinder events of length 8,
2717: which requires knowledge of Fourier coefficients only for $|k| \le 7$, we
2718: obtain that 
2719: $$
2720: 0.659648
2721: \le H(\P^f) \le
2722: 0.684021
2723: \,.
2724: $$
2725: Comparing with \ref e.entropy-g/, we see that $H(\P^f) > H(\P^g)$.
2726: \endprocl
2727: 
2728: Note that even when neither $\widehat f$ nor $\widehat{\varphi_f}$ can be
2729: found explicitly, they can always be found by numerical integration and
2730: exponentiation, and then one can follow the procedure we have used in 
2731: this last example. Indeed, that will be done for part of our final example.
2732: 
2733: \procl x.ustH-x-ent
2734: As in \ref x.ust_H-x/, let 
2735: $$
2736: g(x) := 
2737: {\sin \pi x \over \sqrt{1+\sin^2 \pi x}}
2738: \,.
2739: $$
2740: Then the edges of the uniform spanning tree measure in the plane that lie on
2741: the $x$-axis have the law $\P^g$.
2742: Write $g_1(x) := \sin^2 \pi x$.
2743: The calculations in \ref x.ustHx/ show that $\Phi_{g_1}(z) = (1-z)/2$ and
2744: $\Phi_{\bfo+g_1}(z) = [\sqrt 2 + 1 - (\sqrt 2 - 1)z]/2$, whence
2745: $$
2746: \Phi_g(z)
2747: =
2748: \sqrt{\Phi_{g^2}(z)}
2749: =
2750: \sqrt{{\Phi_{g_1} \over \Phi_{\bfo+g_1}}}
2751: =
2752: \sqrt{1-z \over \sqrt 2 + 1 - (\sqrt 2 - 1)z}
2753: \,.
2754: $$
2755: Expansion of $\Phi_g(z)$ in a Maclaurin series gives 
2756: $$
2757: \frac{1}{({1 + {\sqrt{2}}})^{1/2}} - \frac{z}{( 1 + {\sqrt{2}} )
2758: ^{\nonfrac{3}{2}}} - \frac{(-1+2\sqrt{2})\,z^2}{2\,(1 +\sqrt{2})^{5/2}} -
2759: \frac{( 5 - 2\,{\sqrt{2}} ) \,z^3}{2\,( 1 + {\sqrt{2}}
2760: ) ^{\nonfrac{7}{2}}} - \frac{( -27 + 28\,{\sqrt{2}} )
2761: \,z^4}{8\,( 1 + {\sqrt{2}} ) ^{\nonfrac{9}{2}}} +
2762: \cdots
2763: \,,
2764: $$
2765: which tells us $\widehat{\varphi_g}$.
2766: The transfer currents, which can be calculated by the method in \ref
2767: b.BurPem/, tell us $\widehat g$; for example, $\widehat g(k)$ for $k=0, 1, 2,
2768: 3, 4$ is 
2769: $$
2770: \frac{1}{2}\,,\qquad \frac{1}{2} - \frac{2}{\pi
2771: }\,,\qquad\frac{5}{2} - \frac{8}{\pi }\,,\qquad\frac{25}{2} - \frac{118}{3\,\pi
2772: }\,,\qquad
2773: \frac{129}{2} - \frac{608}{3\,\pi }
2774: \,.
2775: $$
2776: We use numerical integration to find $\widehat{\varphi_{\bfo-g}}$.
2777: Then if we use cylinder events of length 8, we find that 
2778: $$
2779: 0.69005 \le H(\P^g) \le 0.69013
2780: \,.
2781: $$
2782: It is interesting how close this is to $\log 2 = 0.69315^-$.
2783: \endprocl
2784: 
2785: 
2786: We close our treatment of entropy with some elementary observations.
2787: If $f_n \to 1/2$ weak${}^*$ (meaning that we have weak${}^*$
2788: convergence of the measures having these densities),
2789: must $H(\P^{f_n}) \to \log 2$?
2790: The answer is ``no", as we now demonstrate.
2791: Given any $f : \T \to [0, 1]$ and any integer $n$, let $\mult{n} f$ denote the
2792: function 
2793: $$
2794: \mult{n} f(x) := f(nx)
2795: \,.
2796: $$
2797: Let $\Seq{\eta^{(n)}_k \st k \in \Z}$ have the distribution $\P^{\mult{n}f}$.
2798: Since $f(x) = \sum_{k \in \Z} \widehat{f}(k) e^{2\pi i k x}$ in $L^2(\T)$, the
2799: Fourier expansion of $\mult{n}f$ is $\mult{n}f(x) = \sum_{k \in \Z}
2800: \widehat{f}(k) e^{2\pi i k n x}$ for $n \ne 0$.
2801: In particular, for any $n \ne 0$ and any $r$, the processes
2802: $\Seq{\eta^{(n)}_{nk+r} \st k \in \Z}$ 
2803: each have distribution $P^f$ and are independent of each other 
2804: as $r$ ranges from 0 to $n-1$.
2805: Therefore $H(\P^{\mult{n}f}) = H(\P^f)$.
2806: Note that unless $f$ is constant, 
2807: $H(\P^{\mult{n}f}) = H(\P^f) < H\big[\widehat f(0)\big]$, even
2808: though $\mult{n}f$ tends weak$^*$ to the constant function $\widehat f(0)
2809: \cdot \bfo$.
2810: One can show that a similar phenomenon holds in higher dimensions, that is,
2811: if $f : \Td \to [0, 1]$ and $A: \Td \to \Td$ is a group epimorphism, then
2812: $H(\P^f) = H(\P^{f \circ A})$.
2813: 
2814: 
2815: \bsection {Phase Multiplicity}{s.sk}
2816: 
2817: 
2818: In this section, we classify exactly the set of functions $f$ for 
2819: which $\P^f$ satisfies a
2820: strong full $K$ property. This property, which we now describe, is an 
2821: essential strengthening of the 
2822: usual Kolmogorov or $K$ property. One of the reasons this property 
2823: is interesting is that it is closely connected to the notion
2824: of multiplicity for Gibbs states in 
2825: statistical mechanics; in particular, it corresponds to 
2826: {\it uniqueness}. In the next section, we
2827: classify exactly the set of functions $f$ for which $\P^f$ satisfies a
2828: (1-sided) strong $K$ property.
2829: 
2830: {\bf In this section and the next, we always assume that $f$ is not}
2831: {\bf identically  0 nor identically 1.}
2832: 
2833: 
2834: To begin, we first recall the $K$ property for stationary processes indexed 
2835: by $\Z$. 
2836: There are many equivalent formulations, of which we choose an appropriate one.
2837: 
2838: For $S \subseteq \Z^d$, write $\F(S)$ for the $\sigma$-field on $2^{\Z^d}$
2839: generated by $\eta(e)$ for $e \in S$.
2840: 
2841: \procl d.K
2842: A translation-invariant probability measure $\mu$ on
2843: $2^{\Z}$ is {\bf $K$\/} (or a {\bf Kolmogorov automorphism})
2844: if for any finite cylinder event $E$ and 
2845: for all $\epsilon>0$ there exists an $N$ such that 
2846: $$
2847: \mu\Big[ \big|\mu\big(E \mid \F(\OC{-\infty,-N}\cap \Z)\big)-\mu(E)\big| \ge \epsilon \Big] \le
2848: \epsilon
2849: \,.
2850: $$
2851: \endprocl
2852: 
2853: It is well known that the above definition of $K$ is equivalent 
2854: to having a trivial (1-sided) tail $\sigma$-algebra in the sense that
2855: the $\sigma$-algebra $\bigcap_{m \ge 1} \F\big(\OC{- \infty,
2856: -m}\big)$ is trivial
2857: (see page 120 of \ref b.Georgii/ for a version of this).
2858: The $K$ property is known to be an isomorphism invariant and is
2859: also known to be equivalent to the
2860: property that all nontrivial factors have strictly positive entropy.
2861: The latter implies that a process is $K$ iff its time reversal is $K$,
2862: something which is not immediate from the definition.
2863: 
2864: In order to give a complete discussion of the points that we wish to make,
2865: we need to introduce three further properties, which are respectively
2866: called the full $K$ property, the strong full $K$ property, and the
2867: (1-sided) strong $K$ property, the last property given only for $d=1$.
2868: 
2869: While the notion of the $K$ property generalizes 
2870: to $\Z^d$ (see \ref b.Conze/),
2871: there is a slight strengthening of the definition that 
2872: has a more aesthetic extension to
2873: $\Z^d$. To give this, define $B^d_n:=[-n,n]^d\cap\Z^d$.
2874: 
2875: \procl d.2K
2876: A translation-invariant probability measure $\mu$ on
2877: $2^{\Z^d}$ is {\bf full $K$\/} if for any finite cylinder event $E$ and 
2878: for all $\epsilon>0$ there exists an $N$ such that 
2879: $$
2880: \mu\Big[ \big|\mu\big(E \mid \F((B^d_N)^c)\big)-\mu(E)\big| \ge \epsilon \Big] \le
2881: \epsilon
2882: \,.
2883: $$
2884: \endprocl
2885: 
2886: Analogously to an earlier statement, the
2887: full $K$ property is equivalent (see again page 120 of \ref b.Georgii/)
2888: to having a 
2889: {\bf trivial full tail}, which means that % $\F_{\pm \infty}$
2890: $\bigcap_{m\ge 1} \F\big((B^d_m)^c\big)$ is trivial.
2891: It is well known that $K$ does not imply 
2892: full $K$ even for $d=1$ (see, for example, the bilaterally deterministic
2893: Bernoulli shift processes constructed in \ref b.OW:bilat/).
2894: However, for Markov random fields, the two notions are equivalent
2895: (see \refbmulti{HS,HS2}).
2896: \ref b.L:det/ proved that all (not necessarily $\Z^d$-invariant) 
2897: determinantal probability measures satisfy
2898: the full $K$ property. 
2899: For Gibbs states arising in statistical mechanics, this property 
2900: is equivalent
2901: to an ``extremality'' property of the Gibbs state 
2902: (see page 118 of \ref b.Georgii/ for precisely 
2903: what this means and this equivalence). 
2904: 
2905: The following definitions strengthen the $K$ property in an even more 
2906: essential way. We begin with the full version, which seems to us more 
2907: natural.
2908: 
2909: \procl d.2sK
2910: A translation-invariant probability measure $\mu$ on
2911: $2^{\Z^d}$
2912: is {\bf strong full $K$\/} if for any finite cylinder event $E$ and 
2913: for all $\epsilon>0$ there exists an $N$ such that 
2914: $$
2915: \big|\mu\big(E \mid \F((B^d_N)^c)\big)-\mu(E)\big| <\epsilon
2916: \quad \mu\hbox{-a.e.}
2917: $$
2918: \endprocl
2919: 
2920: For the full $K$ property, ``$\mu$-most'' conditionings far away have
2921: little effect on a ``local event'', while in the strong full $K$ case,
2922: {\it all\/} conditionings far away have little effect on a ``local event''.
2923: This is a substantial difference. An example that illustrates this 
2924: difference is the Ising model in $\Z^2$. 
2925: The plus state for the Ising model at high temperatures
2926: is strong full $K$, while at low temperatures,
2927: it is full $K$ (because it is extremal) but not strong full $K$.
2928: 
2929: Finally, if $d=1$, we define
2930: strong $K$ in the following way (which the reader can presumably
2931: anticipate). One extension of this definition 
2932: to $\Z^d$ (among various possible) will be given in \ref s.1side/.
2933: 
2934: \procl d.1sK
2935: A translation-invariant probability measure $\mu$ on $2^{\Z}$ 
2936: is {\bf strong $K$\/} if for any finite cylinder event $E$ and 
2937: for all $\epsilon>0$ there exists an $N$ such that 
2938: $$
2939: \big|\mu\big(E \mid \F(\OC{-\infty,-N}\cap \Z)\big)-\mu(E)\big| <\epsilon
2940: \quad\mu\hbox{-a.e.}
2941: $$
2942: \endprocl
2943: 
2944: We note that this definition is closely related to, but weaker than,
2945: the $\psi$-mixing property (see \ref b.Bradley/). The difference is that the
2946: event $E$ here is specified in advance, rather than having a uniformity 
2947: for all events $E$, provided only that they depend 
2948: on the random variables
2949: with positive index. If we extend the notion of strong $K$ 
2950: in the obvious way to the case when $\{0,1\}$ is replaced by a countable set,
2951: then the example in \ref b.Bradley/ of a process that is 
2952: $\psi$-mixing, but whose time reversal is not $\psi$-mixing, yields an example
2953: of a strong $K$ process whose time reversal is not.
2954: This cannot occur for our measures $\P^f$ since they are all time reversible.
2955: 
2956: The following development has an analogy with the ``plus and minus states''
2957: in the Ising model from statistical mechanics; however, such knowledge 
2958: is not needed by the reader.
2959: 
2960: Consider a function $f:\Td\to [0,1]$ and consider the corresponding probability measure
2961: $\P^f$ on $2^{\Z^d}$. We shall define a probability measure
2962: $(\P^f)^+$ on $2^{\Z^d}$ which will be ``$\P^f$ conditioned 
2963: on all 1s at $\infty$'';
2964: more specifically (but still not precisely), we want to define $(\P^f)^+$ by
2965: $$
2966: \Pfp := \lim_{n\to \infty} \P^f[\;\cdot \mid  \eta\equiv 1 \hbox{ on }
2967: (B^d_n)^c]\,.
2968: $$
2969: To make sure this is well defined, we proceed in stages. 
2970: Let
2971: $$
2972: (\Pfp)_n := \lim_{k\to\infty} \P^f[\;\cdot \mid  \eta\equiv 1 \hbox{ on
2973: }B^d_{n+k}\backslash B^d_n]\,.
2974: $$
2975: This limit is taken in the weak$^{*}$-topology. \ref p.JNRD/ implies
2976: that, when restricted to $B^d_n$, 
2977: this sequence is stochastically decreasing and hence necessarily converges. 
2978: $(\Pfp)_n$ is therefore well defined. One next defines
2979: $$
2980: \Pfp := \lim_{n\to \infty} (\Pfp)_n
2981: \,,
2982: $$
2983: where again the limit is taken in the weak$^{*}$-topology. 
2984: \ref p.JNRD/ again implies
2985: that for fixed $k$,
2986: $(\Pfp)_n$ restricted to $B^d_k$, is, for $n>k$,
2987: stochastically increasing in $n$ and hence converges.
2988: This implies that its limit $\Pfp$ is well
2989: defined and completes the definition of $\Pfp$.
2990: The stochastic monotonicity results also imply that 
2991: $\Pfp \preccurlyeq  \P^f$ and that
2992: $\Pfp$ is translation invariant.
2993: 
2994: In exactly the same way (using 0 instead of 1 boundary conditions), one defines
2995: $\Pfm$, which satisfies $\P^f  \preccurlyeq \Pfm$. Analogy with the
2996: ferromagnetic Ising model in statistical mechanics leads us to the
2997: following definition.
2998: 
2999: \procl d.pt
3000: The probability measure $\P^f$ has {\bf phase multiplicity} if $\Pfm\neq \Pfp$, and otherwise has {\bf phase uniqueness}.
3001: \endprocl
3002: 
3003: \procl l.pt
3004: $\P^f$ has phase uniqueness if and only if it is strong full $K$.
3005: \endprocl
3006: 
3007: The reason this is true is that \ref p.JNRD/ implies
3008: that the most
3009: extreme boundary conditions are when all 1s or all 0s are used and the
3010: measures corresponding to any other boundary conditions are
3011: ``stochastically trapped'' between these two special cases. 
3012: A detailed proof follows straightforwardly from the stochastic 
3013: monotonicity arguments above and
3014: is left to the reader.
3015: 
3016: We shall now
3017: obtain necessary and sufficient conditions for the equality of $\P^f$ and
3018: $\Pfp$. From these, a necessary and sufficient condition for the strong full $K$ property will easily emerge. 
3019: 
3020: 
3021: Let $\calT$ denote the set of trigonometric polynomials on $\T^d$.
3022: Let $L^2(1/f)$ denote the set 
3023: $$
3024: \Big\{h:\Td\to \C \st \int_\Td {|h|^2 \over f} \, d\lambda_d <\infty\Big\}\,. 
3025: $$
3026: Here we use the convention that $0/0:=0$. Note also that
3027: $h$ needs to vanish where $f$ does.
3028: 
3029: Our main result on phase multiplicity is the following.
3030: 
3031: \procl t.sn
3032: Assume that $f:\Td\to [0,1]$ is measurable. %positive a.e. 
3033: The following are equivalent. 
3034: \beginitems
3035: \itemrm{(i)} $\Pfp=\P^f$; 
3036: \itemrm{(ii)} $f$ is in the closure in $L^2(1/f)$ of $\calT\cap L^2(1/f)$;
3037: \itemrm{(iii)} There exists a nonzero trigonometric polynomial $T$ such that 
3038: ${|T|^2 \over f}\in L^1(\Td)$; i.e., $\calT\cap L^2(1/f)\neq {0}$.
3039: \enditems
3040: Moreover, if $\Pfp\neq \P^f$, then $\Pfp= \delta_\bfz$.
3041: \endprocl
3042: 
3043: 
3044: \proof
3045: We shall first show that (i) and (ii) are equivalent and then
3046: that (ii) and (iii) are equivalent.
3047: 
3048: {\bf (i) implies (ii):} Let $u_n$ be the element in $[B]_f$ achieving the
3049: minimum in \ref e.closest/ for $B := (B^d_n)^c$.
3050: Then 
3051: $$
3052: \|\bfo\|^2_f = \|\bfo - u_n\|^2_f + \|u_n\|^2_f
3053: \,.
3054: $$
3055: By (i) and \ref e.closest/, we have $\|\bfo - u_n\|_f \to \|\bfo\|_f$ as $n
3056: \to\infty$, or in other words, $\|u_n\|_f \to 0$ as $n \to\infty$.
3057: Furthermore, $\bfo - u_n \perp [(B^d_n)^c]_f$ in $L^2(f)$, which is the same
3058: as $(\bfo - u_n) f$ being a trigonometric polynomial $T_n$ with $\widehat
3059: {T_n}$ supported in $B^d_n$.
3060: We have 
3061: $$
3062: \|f - T_n\|^2_{(1/f)}
3063: =
3064: \int |f - T_n|^2 {1 \over f} \,d\lambda_d
3065: =
3066: \int \left|\bfo - {T_n \over f}\right|^2 f \,d\lambda_d
3067: =
3068: \int |u_n|^2 f \,d\lambda_d
3069: =
3070: \|u_n\|^2_f
3071: \label e.twoways
3072: $$
3073: tends to 0 as $n \to\infty$, which proves (ii).
3074: 
3075: {\bf (ii) implies (i):}
3076: Given $\epsilon > 0$, let
3077: $T$ be a trigonometric polynomial with $\|f - T\|_{(1/f)} < \epsilon$.
3078: Let $n$ be such that $\widehat T$ is supported in $B^d_n$.
3079: Define $v := T/f$, which is in $L^2(f)$.
3080: Then $v \perp [(B^d_n)^c]_f$ in $L^2(f)$ and $\|\bfo - v\|_f = \|f -
3081: T\|_{(1/f)} < \epsilon$, so 
3082: $\|P_{[(B^d_n)^c]_f}^\perp \bfo\|_f^2 > {\|\bfo\|^2_f -\epsilon^2}$.
3083: Combining this with \ref e.byproj/, we see that
3084: $\Pfp[\eta(\bfz) = 1] > \P^f[\eta(\bfz) = 1] - \epsilon^2$. Since
3085: this holds for any $\epsilon > 0$, we obtain that $\Pfp[\eta(\bfz) = 1] \ge
3086: \P^f[\eta(\bfz) = 1]$. However, 
3087: since $\Pfp \preccurlyeq \P^f$ and both are translation invariant, 
3088: the two measures are actually equal by \ref l.weak/, which proves (i).
3089: 
3090: {\bf (ii) implies (iii):} 
3091: This is immediate.
3092: 
3093: {\bf (iii) implies (ii):}
3094: Assume there exists a nonzero
3095: trigonometric polynomial $T \in L^2(1/f)$. 
3096: Let $\epsilon > 0$. 
3097: Choose $B\subseteq \Td$ such that
3098: $$
3099: \int_B \left({|T|^2 \over f} + 2|T| +f\right) < \epsilon 
3100: $$
3101: and $B$ is an open set containing the zero set of $T$ (which is a set of
3102: measure 0, as remarked in the proof of \ref t.support/). Let 
3103: $g := T  \I B + f  \I {B^c}$. Then $g/T \in L^\infty(\Td)$ since $T$ 
3104: is bounded away from 0 on $B^c$ and $0\le f \le 1$. We can now choose 
3105: trigonometric 
3106: polynomials $p_n$ on $\Td$ such that $p_n \to g/T$ a.e.\ 
3107: on $\Td$ with $\|p_n\|_\infty \le \|g/T\|_\infty$ for all $n$ (just use
3108: $p_n:=(g/T) * K^d_n$, where $K^d_n$ is, as before, the Fej\'er kernel for
3109: $\Td$).  
3110: Since $|T|^2/f \in L^1(\Td)$, it follows from the 
3111: Lebesgue dominated convergence theorem that
3112: $$
3113: \lim_{n\to\infty}\int_\Td {|T|^2 \over f}  \left|{g \over T}-p_n\right|^2 =0\,.
3114: $$
3115: Hence there exists a trigonometric polynomial $h$ such that
3116: $$
3117: \int_\Td {|T|^2 \over f}  \left|{g \over T}-h\right|^2 < \epsilon\,.
3118: \label e.1
3119: $$
3120: Minkowski's inequality (applied to $L^2(|T|^2/f)$) yields
3121: $$
3122: \left(\int_\Td {|T|^2 \over f}  \left|{f \over T}-h\right|^2\right)^{{1\over
3123: 2}} \le
3124: \left(\int_\Td {|T|^2 \over f}  \left|{f \over T}-{g \over
3125: T}\right|^2\right)^{{1\over 2}} +
3126: \left(\int_\Td {|T|^2 \over f}  \left|{g \over T}-h\right|^2\right)^{{1\over
3127: 2}}\,.
3128: $$
3129: The second summand is at most $\epsilon^{{1/ 2}}$ by \ref e.1/. 
3130: The first summand is 
3131: $$
3132: \left(\int_B {|T|^2 \over f}  \left|{f \over T}-1\right|^2\right)^{{1\over 2}}
3133: \le 
3134: \left(\int_B f+ 2|T| + {|T|^2 \over f}\right)^{{1\over 2}} <\epsilon^{{1\over
3135: 2}}
3136: $$
3137: by choice of $B$.
3138: Hence 
3139: $\int_\Td |f-Th|^2 {1 \over f} \, dx < 4\epsilon$. Since $Th$ is a 
3140: trigonometric polynomial, (ii) is proved.
3141: 
3142: Finally, assume that $\Pfp\neq \P^f$. Since (iii) fails,
3143: $[(B^d_n)^c]_f^\perp = \bfz$ for all $n$, which means by \ref e.byproj/
3144: that $\Pfp[\eta(\bfz) = 1] = 0$, or, in other words, $\Pfp = \delta_\bfz$.
3145: \Qed
3146: 
3147: \procl x.continuous-finite
3148: If $f:\T\to [0,1]$ is continuous and has a finite number of 0s with $f$
3149: approaching each of these 0s at most polynomially quickly, then $\Pfp=\P^f$
3150: since it is easy to construct a trigonometric polynomial $T$ with
3151: the same 0s as $f$ and approaching 0 at least as quickly as $f$
3152: (for example, $T$ could be of the form 
3153: $\prod_{i=1}^k \sin^n 2\pi(x-a_i)$).
3154: In particular, if $f$ is real analytic and not $\bfz$, 
3155: then $\Pfp = \P^f$. 
3156: \endprocl
3157: 
3158: \procl x.trivial
3159: If $f$ vanishes on a set of positive measure, then $\Pfp=\delta_\bfz$.
3160: \endprocl
3161: 
3162: \procl x.cont
3163: If $f:\T\to [0,1]$ is a continuous function with a single zero at $x_0\in \T$
3164: and $f(x) = e^{-1/|x-x_0|}$ in some neighborhood of $x_0$,
3165: then $\Pfp=\delta_\bfz$. Indeed, there is no nonzero 
3166: trigonometric polynomial $T$ with
3167: $|T|^2/f\in L^1(\T)$ since the rate at which a trigonometric polynomial 
3168: approaches 0 is at most polynomially quickly.  
3169: \endprocl
3170: 
3171: \procl x.depends-distribution
3172: We give an example to show that the property $\Pfp = \P^f$ does {\it not\/}
3173: depend only on the distribution of $f$, even among real-analytic functions
3174: $f$.
3175: In 2 dimensions, the function 
3176: $$
3177: f(x, y) := \sin^2 (2\pi y - \cos 2\pi x)
3178: $$
3179: generates a system for which $\Pfp \ne \P^f$.
3180: This is because $f$ vanishes (even to second order) on a curve 
3181: $$
3182: y = {1 \over 2 \pi} \cos 2 \pi x + \Z
3183: \label e.curve
3184: $$
3185: that is not in the zero set of any trigonometric polynomial in two
3186: variables except the zero polynomial.
3187: However, $f$ has the same distribution as the trigonometric polynomial
3188: $g(x, y) := \sin^2 2 \pi y$ and
3189: $(\P^g)^+ = \P^g$.
3190: To see that the curve \ref e.curve/ does not lie in the zero set of any
3191: nonzero trigonometric polynomial, suppose that $T(x, y)$ is a trigonometric
3192: polynomial that vanishes there. Write $w := e^{2\pi i x}$ and $z := e^{2\pi
3193: i y}$. By multiplying by a suitable complex
3194: exponential, we may assume that $P(w, z) = T(x, y)$ is an analytic
3195: polynomial. 
3196: Our assumption is that 
3197: $$
3198: h(w) := P(w, e^{i(w+w^{-1})/2})
3199: $$
3200: satisfies $h(w) = 0$ for all $|w|=1$.
3201: Since $h$ is an analytic function for $w \ne 0$, it follows that $h(w) = 0$
3202: for all $w \ne 0$.
3203: Now for each $z \ne 0$, there are an infinity of $w$ such that
3204: $e^{i(w+w^{-1})/2} = z$. Therefore for each $z \ne 0$,
3205: there are an infinity of $w$ such that $P(w, z) = 0$.
3206: Since a polynomial has only a finite number of zeroes if it is not
3207: identically zero, this means that for each $z \ne 0$, $P(w, z) = 0$ for all
3208: $w \in \C$.
3209: Hence $P \equiv 0$, as desired.
3210: \endprocl
3211: 
3212: \procl x.algvar
3213: For $d = 2$, if $f$ is real analytic on a neighborhood of its 
3214: zero set, then $\Pfp = \P^f$ iff its zero set is contained in a 
3215: (nontrivial) algebraic variety, where we view $\T^2$ as $ \{ (z_1, 
3216: z_2) \in \C^2 \st |z_1| = |z_2| = 1 \}$.
3217: This is because the slowest $f$ can vanish at a point is of order $x^2+y^2$
3218: (since the constant and linear terms of $f$ must vanish)
3219: and $1/(x^2+y^2)$ is not integrable. Therefore, all the zeroes of $f$ must be
3220: cancelled by those of $T$.
3221: Conversely, if the zero set of $f$ is contained in the zero set of a
3222: trigonometric polynomial $T$, then by {\L}owasiewicz's inequality (see,
3223: e.g., \ref b.BierMil/, Theorem 6.4), for a sufficiently large $n$, we have
3224: $|T|^n/f$ is bounded, whence integrable.
3225: \endprocl
3226: 
3227: 
3228: \procl x.depends-distribution1
3229: Here, we give a 1-dimensional example showing that the property $\Pfp =
3230: \P^f$ does {\it not\/} depend only on the distribution of $f$, even among
3231: continuous $f$.
3232: Let $f(x) := \sqrt{x} |\sin (\pi/x)|$ on $[0, 1]$.
3233: Then $\Pfp \ne \P^f$ since there are an infinite number of first-order zeroes.
3234: However, let $g$ be the increasing rearrangement of $f$ on $[0, 1]$ and define
3235: $$
3236: h(x) := \cases{g(2x) &if $x \le 1/2$,\cr
3237:                g(2-2x) &if $x \ge 1/2$.\cr}
3238: $$
3239: Then $h$ is continuous and has the same distribution as $f$, yet
3240: $(\P^h)^+ = \P^h$ by \ref x.continuous-finite/ together with an easy
3241: computation.
3242: Of course, there is no such example for real-analytic $f$ in one dimension
3243: by \ref x.continuous-finite/.
3244: \endprocl
3245: 
3246: We finally state and prove our necessary and sufficient condition for
3247: the strong full $K$ property.
3248: 
3249: \procl c.sn
3250: Consider $f:\Td\to [0,1]$ and the corresponding probability measure
3251: $\P^f$. Then
3252: $\P^f$ is strong full $K$ if and only if there is a nonzero
3253: trigonometric polynomial $T$ such that 
3254: ${|T|^2 \over f (\bfo-f)} \in L^1(\Td)$.
3255: \endprocl
3256: 
3257: \proof
3258: According to \ref t.sn/, 
3259: $\Pfp=\P^f$ iff there exists a nonzero
3260: trigonometric polynomial $T_1$ such that
3261: ${|T_1|^2 \over f} \in L^1(\Td)$. The same argument applied to $\bfo - f$
3262: tells us that $\Pfm=\P^f$ iff there exists a nonzero
3263: trigonometric polynomial $T_2$ such that
3264: ${|T_2|^2 \over \bfo-f} \in L^1(\Td)$. \ref l.pt/ now completes the proof.
3265: (Take $T := T_1 T_2$.)
3266: \Qed
3267: 
3268: \procl r.analytic
3269: Observe that by trivial scaling, we could have an $f$ whose Fourier
3270: coefficients go to zero slowly but which is bounded away from 0 and 1.
3271: Hence slow decay of Fourier coefficients does not imply phase multiplicity.
3272: On the other hand, if the coefficients decay exponentially, then 
3273: $\sum_{n \in \Z} \widehat{f}(n) z^n$ is complex-analytic in an annulus around
3274: the unit circle, which implies that its restriction to the unit circle is
3275: real-analytic and hence there is phase uniqueness.
3276: \endprocl
3277: 
3278: \procl r.folner
3279: We mention that if $\Pfp=\P^f$, then this measure is
3280: F{\o}lner independent in the sense of \ref b.Adams/ and
3281: if in addition $\Pfm=\P^f$, then this measure is even strong
3282: F{\o}lner independent in the sense that any conditioning outside
3283: a box yields a measure which is $\dbar$ close to the unconditioned
3284: process. The arguments for proving these facts are analogous to those
3285: of \ref b.OW:ising/ (see \ref b.Adams/ for a published version), where it is
3286: proved that the plus state for the Ising model is a Bernoulli shift.
3287: (The concept of F{\o}lner independence has been used also in \ref b.HS/ and
3288: \ref b.hoffman:MRF/.)
3289: \endprocl  
3290: 
3291: We close this section with an interpretation of
3292: our phase multiplicity results in terms 
3293: of interpolation questions for wide-sense stationary 
3294: processes. The following comments can be proved from \ref t.closest/ 
3295: together with the well-known correspondence 
3296: between interpolation and Szeg\H{o} infima.
3297: First, $\Pfp= \delta_\bfz$ is equivalent to being able to 
3298: interpolate perfectly from information far away for 
3299: a wide-sense stationary process $\Seq{Y_n}$  with spectral density $f$,
3300: which means that for every $n$, $Y_0$ is in the closed linear
3301: subspace spanned by  $\{Y_k\}_{|k|\ge  n}$. For $d=1$, it
3302: is stated on p.~102 of \ref b.Roz/ that this latter condition is 
3303: equivalent to the negation of condition (iii) in \ref t.sn/.
3304: A similar statement holds for $\Pfm$ with $f$ replaced by $\bfo-f$.
3305: The proof of \ref t.sn/ implies (via this whole correspondence) that, 
3306: for spectral measures that are
3307: absolutely continuous with a bounded nonnegative density,
3308: if perfect linear interpolation fails, then
3309: our ability to interpolate as $n$ gets large goes to 0; i.e.,
3310: the length of the projection of $Y_0$ onto the closed linear
3311: subspace spanned by  $\{Y_k\}_{|k|\ge  n}$ goes to 0 as $n\to\infty$.
3312: 
3313: \bsection{1-Sided Phase Multiplicity}{s.1side}
3314: 
3315: We first study the notion of strong $K$ for $d=1$. 
3316: This is natural since the
3317: structure of ``one-sided'' behavior can be very different from
3318: ``two-sided'' behavior in various situations; for example, the existence
3319: of bilaterally deterministic Bernoulli shift processes demonstrates this.
3320: In contrast with \ref x.depends-distribution1/, we shall see that whether
3321: $\P^f$ is strong $K$ depends only on the distribution of $f$.
3322: 
3323: Consider a function $f:\T\to [0,1]$ and the 
3324: corresponding probability measure
3325: $\P^f$ on $2^{\Z}$. We define a probability measure
3326: $(\P^f)^{+,1}$ on $2^{\Z}$ thought of as ``$\P^f$ conditioned 
3327: on all 1s at $-\infty$'' (the superscript ``1'' refers to the fact that we are
3328: doing this on 1 side); namely,
3329: $$
3330: (\P^f)^{+,1}:= \lim_{n\to \infty} \lim_{k \to\infty} 
3331: \P^f\big[\;\cdot \mid \eta\equiv 1 \hbox{ on }
3332: [-n-k,-n]\cap \Z\big]
3333: \,.
3334: $$
3335: The existence of the limit and its translation invariance follow from
3336: stochastic monotonicity, as did that of $(\P^f)^+$.
3337: As before, we also have $(\P^f)^{+,1} \preccurlyeq  \P^f$. 
3338: In the analogous way (using 0 instead of 1 boundary conditions), one defines
3339: $(\P^f)^{-,1}$, which then satisfies $\P^f  \preccurlyeq (\P^f)^{-,1}$. 
3340: 
3341: \procl d.pt1
3342: The probability measure $\P^f$ has a {\bf 1-sided phase multiplicity} if 
3343: $(\P^f)^{-,1}\neq(\P^f)^{+,1}$, and otherwise has {\bf 1-sided phase
3344: uniqueness}.
3345: \endprocl
3346: 
3347: Note the following analogue (whose proof is left to the reader) 
3348: of \ref l.pt/ for the 1-sided case.
3349: 
3350: \procl l.pt1
3351: $\P^f$ has 1-sided phase uniqueness if and only if it 
3352: is strong $K$.
3353: \endprocl
3354: 
3355: Our main result for 1-sided phase multiplicity is the following. 
3356: 
3357: \procl t.sn1GM
3358: If $f:\T\to [0,1]$, then $(\P^f)^{+,1}=\P^f$ iff $\GM(f) > 0$. 
3359: Moreover, if $(\P^f)^{+,1}\neq \P^f$, then $(\P^f)^{+,1}= \delta_\bfz$.
3360: \endprocl
3361: 
3362: \proof
3363: By \ref t.szego/, we have
3364: $$
3365: \lim_{k \to\infty} \P^f[\eta(0) = 1 \mid \eta_{-n} = \eta_{-n-1} =
3366: \cdots = \eta_{-n-k} = 1]
3367: =
3368: \sum_{l=0}^{n-1} |\widehat{\varphi_f}(l)|^2
3369: \,.
3370: $$
3371: (In fact, via \ref t.closest/, this special case of \ref t.szego/ is due to
3372: Kolmogorov and Wiener; see \ref b.GrenanderSzego/, Section 10.9.)
3373: Taking $n \to\infty$, we find that
3374: $$
3375: (\P^f)^{+,1}\big[\eta(0) = 1\big] = \|\widehat{\varphi_f}\|^2_2 = \|\varphi_f\|^2_2
3376: \,.
3377: $$
3378: If $\GM(f) > 0$, then we obtain
3379: $$
3380: (\P^f)^{+,1}\big[\eta(0) = 1\big] = \|f\|_1 = \widehat
3381: f(0) = \P^f\big[\eta(0) = 1\big]
3382: \,.
3383: $$
3384: Since $(\P^f)^{+,1} \preccurlyeq \P^f$ and both probability measures are
3385: $\Z$-invariant, it follows by \ref l.weak/ that $(\P^f)^{+,1} = \P^f$.
3386: On the other hand, if $\GM(f) = 0$, then 
3387: $$
3388: (\P^f)^{+,1}\big[\eta(0) = 1\big] = 0
3389: \,,
3390: $$
3391: so
3392: $(\P^f)^{+,1} = \delta_\bfz$.
3393: \Qed
3394: 
3395: Our necessary and sufficient condition for the strong $K$ property 
3396: now follows immediately.
3397: 
3398: \procl c.sn1
3399: For $f:\T\to [0,1]$, the corresponding probability measure $\P^f$
3400: is strong $K$ if and only if $\GM(f) \GM(\bfo-f) > 0$.
3401: \endprocl
3402: 
3403: 
3404: 
3405: We now construct a process $\P^f$ that is strong $K$ but not strong full $K$.
3406: 
3407: \procl t.diff
3408: There exists an $f:\T\to [0,1]$ such that $\P^f$ is strong $K$ but not
3409: strong full $K$.
3410: \endprocl
3411: 
3412: \proof
3413: Choose $f$ such that $f$ is continuous, bounded away from 1, vanishes at a
3414: single point $x_0$, and such that $f(x) = 
3415: e^{-1/(|x-x_0|)^{1/2}}$ in some neighborhood of $x_0$.
3416: Since a trigonometric polynomial vanishes at its zeroes at most
3417: polynomially quickly, there cannot exist a nonzero
3418: trigonometric polynomial $T$
3419: such that $|T|^2/f \in L^1(\T)$. Hence by \ref c.sn/,
3420: $\P^f$ is not strong full $K$. 
3421: On the other hand, it is clear that $\GM(f) \GM(\bfo-f) > 0$, 
3422: so by \ref c.sn1/, we know that $\P^f$ is strong $K$.
3423: \Qed
3424: 
3425: 
3426: \procl r.plus
3427: By combining Theorems 
3428: \briefref t.dom/ and \briefref t.sn1GM/, we see that
3429: $\P^f \succcurlyeq \P^g$ implies 
3430: $(\P^f)^{+,1} \succcurlyeq (\P^g)^{+,1}$.
3431: However, 
3432: it is not necessarily the case that $(\P^f)^+ \succcurlyeq (\P^g)^+$.
3433: For example, let $g \equiv 1/2$ and $f$ be a function as described in the last
3434: result, but which has a geometric mean larger than $1/2$
3435: (recall \ref t.dom/). 
3436: This example also shows that if $\P^g \preccurlyeq \P^f$,
3437: there do not necessarily exist $g'\le f'$ 
3438: such that $\P^{g'} = \P^g$ and $\P^{f'} = \P^f$.  To see this,
3439: let $f$ and $g$ be as above. Then $\P^{g'} = \P^g$ easily implies that
3440: $g' \equiv 1/2$, which in turn (using $g'\le f'$) implies that
3441: $(\P^{f'})^+ =  \P^{f'}$. Since
3442: $\Pfp \neq  \P^f$, this contradicts the fact that $\P^f = \P^{f'}$.
3443: \endprocl
3444: 
3445: 
3446: It may be of interest to see another proof (which we only sketch)
3447: of \ref t.sn1GM/ that does not
3448: depend on the asymptotics of \ref t.szego/, but rather follows
3449: the lines of the proof of \ref t.sn/.
3450: The appropriate replacement of the trigonometric polynomials
3451: for this question is the set 
3452: $$
3453: \calA :=
3454: \{h\in L^2(\T) \st \hbox{ there exists } \ell  \hbox{ such that }
3455: \widehat{h}(k) =0 \hbox{ for } k< \ell\}\,.
3456: $$
3457: Note that when $\ell$ is taken to be 0, we get the {\bf Hardy space}
3458: of analytic functions, denoted $H^2(\T)$.
3459: 
3460: 
3461: \procl t.sn1
3462: Assume that $f:\T\to [0,1]$ and let $\calA$ be as defined as above.
3463: The following are equivalent. 
3464: \beginitems
3465: \itemrm{(i)} $(\P^f)^{+,1}=\P^f$; 
3466: \itemrm{(ii)} $f$ is in the closure in $L^2(1/f)$ of $\calA\cap L^2(1/f)$;
3467: \itemrm{(iii)} There exists a nonzero element $T\in \calA$ such that 
3468: ${|T|^2 \over f}\in L^1(\T)$; i.e., $\calA\cap L^2(1/f)\neq {0}$;
3469: \itemrm{(iv)} $\GM(f) > 0$. 
3470: \enditems
3471: Moreover, if $(\P^f)^{+,1}\neq \P^f$, then $(\P^f)^{+,1}= \delta_\bfz$.
3472: \endprocl
3473: 
3474: \proof
3475: {\bf (i) iff  (ii)} and {\bf (ii) implies (iii):} These are proved
3476: exactly as in \ref t.sn/.
3477: 
3478: {\bf (iii) implies (ii):}
3479: Fix a nonzero $T \in \calA$ with ${|T|^2 \over f}\in L^1(\T)$.
3480: Given any $\epsilon$, there exists $\delta>0$ such that if $B$ is any
3481: subset of $\T$ with measure less than $\delta$, then
3482: $$
3483: \int_B \left({|T|^2 \over f} + 2|T| +f\right) < \epsilon\,.
3484: $$
3485: Since $T$ vanishes only on a set of measure 0
3486: (see \ref b.Rudin:RCA/, Theorem 17.18, p.~345),  
3487: there is a $\gamma>0$ such that
3488: $\{x\in \T \st |T(x)| < \gamma\}$ has measure less than $\delta$.
3489: If we take $B$ to be this latter set, then we can proceed exactly 
3490: as in \ref t.sn/ since now $T$ is bounded away from 0 on $B^c$. 
3491: 
3492: {\bf (iii) implies (iv):}
3493: Any function in $\calA$ is a complex exponential times a function in the
3494: Hardy space $H^2(\T)$. Hence there is 
3495: a nonzero function $T \in H^2(\T)$ with $|T|^2/f \in
3496: L^1(\T)$. Now
3497: $\GM(|T|) > 0$  (see \ref b.Rudin:RCA/, Theorem 17.17, p.~344). 
3498: Letting $g := |T|^2/f$, we have that
3499: $g \in L^1(\T)$ and hence $\GM(g) \le \int g < \infty$.
3500: Therefore $\GM(f) = \GM(|T|^2/g) = \GM(|T|)^2/\GM(g) > 0$, as desired.
3501: 
3502: {\bf (iv) implies (iii):}
3503: If $\GM(f) > 0$, then take $T := \varphi_f$, defined in
3504: \ref s.dom/.
3505: For this $T$, we have $|T|^2/f = \bfo$, which is trivially integrable.
3506: 
3507: The final statement is proved as in \ref t.sn/.
3508: \Qed
3509: 
3510: 
3511: 
3512: 
3513: We now extend \ref t.sn1GM/ and \ref c.sn1/ to higher dimensions.
3514: Fix a choice of ordering of $\Zd$ given by a set $\ord$ 
3515: and a sequence $\Seq{k_n}\in \Z^d$ such that the distance from
3516: $\bfz$ to $-(\ord+k_n)$ goes to $\infty$.
3517: 
3518: \procl d.1sK
3519: A translation-invariant probability measure $\mu$ on $2^{\Z^d}$ 
3520: {\bf has phase uniqueness 
3521: relative to the ordering induced by $\ord$ and the sequence $\Seq{k_n}$ \/} 
3522: as above if for any finite cylinder event $E$ and 
3523: for all $\epsilon>0$ there exists an $N$ such that for all $n\ge N$
3524: $$
3525: \big|\mu\big(E \mid \F(-(\ord+k_n)\big)-\mu(E)\big| <\epsilon
3526: \quad\mu\hbox{-a.e.}
3527: $$
3528: \endprocl
3529: 
3530: It is easy to check using \ref p.JNRD/
3531: that this definition {\it does not depend\/} on the choice
3532: of the sequence $\Seq{k_n}$. However,
3533: it turns out that the choice of ordering, $\ord$, of $\Zd$ can make a 
3534: difference. In particular, if the ordering is archimedean
3535: (meaning that for any two positive elements $a$ and $b$, 
3536: there exists an integer $n$ such that $na > b$), 
3537: such as\ftnote{${}^*$}{In fact, all archimedean orders arise in this way; 
3538: see \ref b.Teh/, \ref b.Zaiceva/, or \ref b.Trevisan/.} when the ordering is
3539: induced by $\ord := \{ k \in \Z^d \st k \cdot x > 0 \}$, 
3540: where $x \in \R^d$ is a fixed vector having two coordinates with an
3541: irrational ratio, then we have a complete characterization
3542: of 1-sided phase multiplicity in terms of the geometric mean.  For the standard
3543: lexicographic ordering, however, 1-sided phase multiplicity cannot be
3544: characterized by the geometric mean alone, as shown by the
3545: example $f :=   \I{[0,1]\times[0, 1/2]}$ in two dimensions.
3546: Here, both $f$ and $\bfo - f$ have 0 geometric mean, but the columns of a
3547: configuration are independent under $\P^f$, so that conditioning on the remote
3548: past has no effect on the present.
3549: On the other hand, positivity of the geometric means will still {\it suffice\/}
3550: for phase uniqueness 
3551: relative to the ordering induced by $\ord$,
3552: as we now prove. Analogously to the
3553: one-dimensional case, we let, for our given ordering $\ord$,
3554: $$
3555: (\P^f)^{+,1}:= \lim_{n\to \infty}
3556: \P^f\big[\;\cdot \mid \eta\equiv 1 \hbox{ on } -(\ord+k_n)\big]
3557: \,.
3558: $$
3559: By \ref p.JNRD/, this limit exists and is independent of $\Seq{k_n}$.
3560: As before, $(\P^f)^{-,1}$ is defined analogously using boundary
3561: conditions of 0s, one-sided phase multiplicity is defined
3562: by $(\P^f)^{-,1}\neq(\P^f)^{+,1}$, and the property
3563: in \ref d.1sK/
3564: is equivalent to one-sided phase uniqueness.
3565: 
3566: \procl t.sn1GM-d
3567: Let $f:\Td\to [0,1]$. If $\GM(f) > 0$, then $(\P^f)^{+,1}=\P^f$.
3568: If $\GM(f) = 0$ and the ordering is archimedean, then
3569: $(\P^f)^{+,1}= \delta_\bfz$. In particular, if the 
3570: ordering is archimedean, then $\P^f$ has phase 
3571: uniqueness (relative to this ordering) if and only if 
3572: $\GM(f) \GM(\bfo-f) > 0$.
3573: \endprocl
3574: 
3575: We begin with the background results on which we rely for the proof.
3576: For an ordering induced by $\ord$, define the 
3577: corresponding {\bf Helson-Lowdenslager space}
3578: $$
3579: \hl^2 := \hl^2(\T^d, \ord) := \Big \{ f \in L^2(\T^d) \st \supp \widehat f
3580: \subset \ord \cup \{ \bfz \} \Big \}
3581: \,.
3582: $$
3583: For $0 \le f \in L^1(\lambda)$, let $[\ord \cup \{ \bfz \}]$ be the linear span
3584: of $ \{ \ee_k \st k \in \ord \cup \{ \bfz \} \}$ and $[\ord \cup \{ \bfz
3585: \}]_f$ be its closure in $L^2(f)$.
3586: The replacement for outer functions is the class of {\bf spectral factors},
3587: which are the functions $\varphi \in \hl^2$ with the additional properties 
3588: $$
3589: \widehat\varphi(\bfz) > 0
3590: \label e.at0
3591: $$
3592: and
3593: $$
3594: 1/\varphi \in [\ord \cup \{ \bfz \}]_{|\varphi|^2}
3595: \,.
3596: \label e.inverse
3597: $$
3598: \ref b.HelLow:I/ show that for any $\ord$ and for any
3599: $0 \le f \in L^1(\T^d)$, the condition $\GM(f) >
3600: 0$ is equivalent to the existence of a spectral factor $\varphi_f$ such that
3601: $|\varphi_f|^2 = f$ a.e. [More precisely, they prove $\GM(f) > 0$ iff $\exists
3602: \varphi \in \hl^2$ satisfying $|\varphi|^2 = f$ a.e.\ and %\ref e.support/ and 
3603: \ref e.at0/. Their proof
3604: shows that in this case, $\varphi$ can be chosen so that also \ref e.inverse/
3605: holds.] Furthermore, \ref b.Lyons:szego/ shows that $\varphi_f$ is then unique.
3606: 
3607: Denote by $\nu_{f}$ the measure on $2^{\ord \cup \{ \bfz \}}$ given
3608: by $\P^f[\;\cdot \mid \eta \restrict (-\ord) \equiv 1]$. This is, as usual,
3609: defined by conditioning on more and more 1s in $-\ord$ and using
3610: \ref p.JNRD/. The results of \ref b.Lyons:szego/ imply the following.
3611: 
3612: \procl t.szego-d 
3613: Fix an ordering induced by a set $\ord$.
3614: Let $f : \Td \to [0, 1]$ be measurable. If $\GM(f) > 0$, define
3615: $\varphi_f$ to be its spectral factor (with respect to $\ord$).
3616: Otherwise, define $\varphi_f := \bfz$.
3617: If $\GM(f) > 0$ or the ordering of $\Zd$ given by $\ord$ is archimedean, then
3618: the measure $\nu_{f}$ is equal to
3619: the determinantal probability measure corresponding to the positive
3620: contraction on $\ell^2(\ord \cup \{ \bfz \})$ whose $(j, k)$-matrix entry is 
3621: $$
3622: \sum_{l\in \ord \cup \{ \bfz \},\; l \preccurlyeq j,k} 
3623: \overline{\widehat{\varphi_f}(j-l)}\widehat{\varphi_f}(k-l)
3624: \,.
3625: $$
3626: \endprocl
3627: 
3628: \proofof t.sn1GM-d
3629: Suppose first that $\GM(f) > 0$ or the ordering given by $\ord$ is archimedean.
3630: By \ref t.szego-d/, we have
3631: $$
3632: \P^f[\eta(\bfz) = 1 \mid \eta \restrict -(\ord+k_n) \equiv 1]
3633: =
3634: \P^f[\eta(k_n) = 1 \mid \eta \restrict (-\ord) \equiv 1]
3635: =
3636: \sum_{l \in \ord \cup \{ \bfz \},\; l \preccurlyeq k_n} 
3637: |\widehat{\varphi_f}(k_n - l)|^2
3638: \,.
3639: $$
3640: Taking $n \to\infty$, we find that
3641: $$
3642: (\P^f)^{+,1}\big[\eta(\bfz) = 1\big] = \|\widehat{\varphi_f}\|^2_2 =
3643: \|\varphi_f\|^2_2
3644: \,.
3645: $$
3646: If $\GM(f) > 0$, then we obtain
3647: $$
3648: (\P^f)^{+,1}\big[\eta(\bfz) = 1\big] = \|f\|_1 = \widehat
3649: f(\bfz) = \P^f\big[\eta(\bfz) = 1\big]
3650: \,.
3651: $$
3652: Since $(\P^f)^{+,1} \preccurlyeq \P^f$ and both probability measures are
3653: $\Zd$-invariant, it follows that $(\P^f)^{+,1} = \P^f$.
3654: However, if $\GM(f) = 0$ and the ordering is archimedean, then 
3655: $$
3656: (\P^f)^{+,1}\big[\eta(\bfz) = 1\big] = 0
3657: \,,
3658: $$
3659: so
3660: $(\P^f)^{+,1} = \delta_\bfz$.
3661: The last statement of the theorem follows as before.
3662: \Qed
3663: 
3664: 
3665: Analogously to \ref s.sk/, our one-sided phase multiplicity results 
3666: for $d=1$ can be translated to known results for prediction (rather than
3667: interpolation) questions for wide-sense stationary 
3668: processes. We leave these observations to the reader. 
3669: 
3670: \bsection{Open Questions}{s.open}
3671: 
3672: \procl q.1
3673: Calculate $H(\P^f)$.
3674: \endprocl
3675: 
3676: We conjecture that entropy is concave:
3677: 
3678: \procl g.concave
3679: For any $f$ and $g$, we have $H\big(\P^{(f+g)/2}\big) \ge \big(H(\P^f) +
3680: H(\P^g)\big)/2$.
3681: \endprocl
3682: 
3683: 
3684: \procl q.6
3685: Letting $A$ vary over all sets of measure 1/2, how do we get the largest
3686: and the smallest entropies for $\P^{\I A}$?
3687: More generally, given $f$, which $g$ with the same distribution as $f$
3688: maximize or minimize $H(\P^g)$?
3689: \endprocl
3690: 
3691: 
3692: \procl q.bounds 
3693: If $d=1$ and $\GM(f) \GM(\bfo - f) > 0$, do we get arbitrarily close lower
3694: bounds (in principle) on $H(\P^f)$ by the method in \ref s.dom/?
3695: How can one get close lower bounds on the entropy for higher-dimensional
3696: processes?
3697: \endprocl
3698: 
3699: \procl q.block
3700: Suppose $f : \T \to [0, 1]$ is a trigonometric polynomial $f$ of
3701: degree $m$.
3702: Then $\P^f$ is $m$-dependent, as are all $(m+1)$-block factors of
3703: independent processes (as defined by \ref b.Hed/).
3704: Is it the case that $\P^{f}$ is an $(m+1)$-block factor of an i.i.d.\ process?
3705: Erik Broman (personal communication) has shown this when $m = 1$.
3706: If one can find such block factors sufficiently explicitly for
3707: trigonometric polynomials, then one could find explicit factors of i.i.d.\
3708: processes that give any process $\P^f$. This would enable one to use more
3709: standard probabilistic techniques to study $\P^f$.
3710: More generally, in higher dimensions,
3711: if $f : \Td \to [0, 1]$ is a trigonometric polynomial, is $\P^f$ a block
3712: factor?
3713: \endprocl
3714: 
3715: 
3716: \procl q.otherPT
3717: Given an ordering $\ord$ on $\Zd$ and an increasing sequence of finite sets
3718: $\ord_n \subset \ord$ whose union is all of $\ord$, when is there phase
3719: multiplicity in the sense that 
3720: $$
3721: \lim_{n\to\infty}
3722: \P^f[\eta(\bfz) = 1 \mid \eta \restrict (\ord \setminus \ord_n) \equiv 1]
3723: \ne
3724: \lim_{n\to\infty}
3725: \P^f[\eta(\bfz) = 1 \mid \eta \restrict (\ord \setminus \ord_n) \equiv 0]
3726: \,?
3727: $$
3728: This clearly does not depend on the choice of $\ord_n$.
3729: In one dimension, this is the same as 1-sided phase multiplicity, where we know
3730: the answer.
3731: \endprocl
3732: 
3733: 
3734: \procl q.3
3735: Suppose that $d = 1$.
3736: Note that translation and flip of $f$ yield the same measure $\P^{f}$.
3737: Does $\P^{f}$ determine $f$ up to translation and
3738: flip? 
3739: \endprocl
3740: 
3741: We now ask about some properties that are important for models
3742: in statistical physics.
3743: Let $\others$ be as in \ref d.fulldom/.
3744: A stationary process $\nu$ is called {\bf quasi-local} if $\nu[\eta(\bfz) =
3745: 1 \mid \others]$ has a continuous version, where the product topology is
3746: used on $2^{\Zd \setminus \{ \bfz \}}$.
3747: It is called {\bf almost surely quasi-local} if $\nu[\eta(\bfz) =
3748: 1 \mid \others]$ has an almost surely continuous version.
3749: 
3750: \procl q.q-local
3751: For which $f$ is $\P^f$ quasi-local or almost surely quasi-local?
3752: When there is phase multiplicity, then
3753: $\P^f$ is as far as possible from this since then each version of
3754: $\P^f[\eta(\bfz) = 1 \mid \others]$ is nowhere continuous.
3755: In one dimension,
3756: there is a natural definition of one-sided quasi-locality, which is the same
3757: as what is sometimes referred to as the uniform martingale property
3758: (see \ref b.kalikow/). It is
3759: not hard to see that this property, or even its almost sure version, would
3760: imply an affirmative answer to \ref q.bounds/.
3761: \endprocl
3762: 
3763: We say that a stationary process $\nu$ on $2^\Zd$ is {\bf insertion
3764: tolerant} if $\nu[\eta(\bfz) = 1 \mid \others] > 0$ $\nu$-a.s.
3765: Similarly, $\nu$ is {\bf deletion tolerant} if $\nu[\eta(\bfz) = 0 \mid
3766: \others] > 0$ $\nu$-a.s.
3767: 
3768: \procl g.tolerance
3769: Every $\P^f$ is insertion and deletion tolerant, other than the trivial cases
3770: $f = \bfz$ and $f = \bfo$.
3771: \endprocl
3772: 
3773: \ref t.fulldom/ implies a positive solution to \ref g.tolerance/ when
3774: $1/[f(\bfo-f)]$ is integrable. Indeed, we see by \ref t.fulldom/ that this is
3775: the exact criterion for uniform insertion tolerance and uniform
3776: deletion tolerance.
3777: 
3778: \medbreak
3779: \noindent {\bf Acknowledgements.}\enspace
3780: We thank Chris Hoffman for his help with \ref p.hoffman/.
3781: We thank Eric Bedford and Jeff Geronimo for important references and Jean-Paul
3782: Thouvenot for discussions.
3783: We are grateful to David Wilson for permission to include Figures \briefref
3784: f.window-ust/ and \briefref
3785: f.window-tiling/ and to Peter Duren and J.M.~Landsberg for helpful remarks.
3786: \ref f.window-tiling/ was created using the linear algebraic techniques of
3787: \ref b.wilson:planar/ for generating domino tilings; the needed
3788: matrix inversion was
3789: accomplished using the formulas of \ref b.Kenyon:local/. The resulting tiling
3790: gives dual spanning trees by the bijection of Temperley (see \ref
3791: b.KPW:tr-mat/). One of the trees is \ref f.window-ust/.  
3792: 
3793: 
3794: \bibfile{\jobname}
3795: \def\noop#1{\relax}
3796: \input \jobname.bbl
3797: 
3798: 
3799: \filbreak
3800: \begingroup
3801: \eightpoint\sc
3802: \parindent=0pt\baselineskip=10pt
3803: 
3804: Department of Mathematics,
3805: Indiana University,
3806: Bloomington, IN 47405-5701
3807: \emailwww{rdlyons@indiana.edu}
3808: {http://php.indiana.edu/\string~rdlyons/}
3809: 
3810: and
3811: 
3812: School of Mathematics,
3813: Georgia Institute of Technology,
3814: Atlanta, GA 30332-0160
3815: \email{rdlyons@math.gatech.edu}
3816: 
3817: Department of Mathematics,
3818: Chalmers University of Technology,
3819: S-41296 Gothenburg, Sweden 
3820: \email{steif@math.chalmers.se}
3821: 
3822: and
3823: 
3824: 
3825: School of Mathematics,
3826: Georgia Institute of Technology,
3827: Atlanta, GA 30332-0160
3828: \email{steif@math.gatech.edu}
3829: 
3830: \endgroup
3831: 
3832: 
3833: \bye
3834: