math0603062/urn.tex
1: \def\versiondate{5 Jan. 2009}
2: \input math.macros
3: \input Ref.macros
4: 
5: %\proofmodetrue
6: %\leftsectionheadtrue
7: \checkdefinedreferencetrue
8: %\continuousnumberingtrue
9: \continuousfigurenumberingtrue
10: \theoremcountingtrue
11: \sectionnumberstrue
12: %\figuresectionnumberstrue
13: \forwardreferencetrue
14: %\lefteqnumberstrue
15: \tocgenerationtrue
16: \citationgenerationtrue
17: \nobracketcittrue
18: \hyperstrue
19: \initialeqmacro
20: 
21: \input\jobname.key
22: \bibsty{myapalike}
23: 
24: \def\cp{\tau}   % complexity
25: \def\gh{G}  % graph
26: \def\gp{\Gamma}  % group
27: \def\gpe{\gamma}  % group element
28: \def\verts{{\ss V}}
29: \def\VV{\verts}
30: \def\edges{{\ss E}}
31: \def\EE{\edges}
32: \def\tr{{\rm Tr}}  % trace
33: \def\bfo{{\bf 1}}
34: \def\bfz{{\bf 0}}
35: \def\detp{{\rm det}'}
36: \def\fo{{\frak F}}  %% forest
37: \def\tree{{\frak T}}  %% tree
38: \def\wsf{{\ss WSF}}
39: \def\fsf{{\ss FSF}}
40: \def\UST{{\ss UST}}
41: \def\ust{{\ss UST}}
42: \def\bdv{\partial_{\verts}}
43: \def\cpe{\cp_{\rm e}}
44: \def\Aut{{\rm Aut}} % automorphism group
45: \def\grn{{\cal G}}  % Green gen function = return series
46: \def\rad{{\bf r}}   % radius of cnv
47: \def\cat{{\rm G}}  % Catalan's constant
48: \def\asym{{\bf h}}   % asymptotics
49: \def\hloc{h_{\rm loc}}  % local entropy
50: \def\ip#1{(\changecomma #1)}
51: \def\bigip#1{\bigl(\changecomma #1\bigr)}
52: \def\Bigip#1{\Bigl(\changecomma #1\Bigr)}
53: \def\biggip#1{\biggl(\bigchangecomma #1\biggr)}
54: \def\leftip#1{\left(\leftchangecomma #1\right)}
55: \def\changecomma#1,{#1,\,}
56: \def\bigchangecomma#1,{#1,\;}
57: \def\leftchangecomma#1,{#1,\ }
58: \def\G{{\cal G}}
59: \def\GG{{\cal G}}
60: \def\PGW{{\ss PGW}}
61: \def\AGW{{\ss AGW}}
62: \def\UGW{{\ss UGW}}
63: \def\sgn{{\rm sgn}}
64: \def\rtd{\mu}  % the measure on rooted graphs
65: \def\cd{\Rightarrow}  % convergence in distribution
66: \def\sfrac#1#2{{#1 \over #2}}
67: \def\frac#1#2{{#1 \over #2}}
68: \def\nroot{\bp}  % network root; \root is already in TeX
69: \def\eps{\epsilon}
70: \def\mathrm#1{{\rm #1}}
71: \def\ed{\buildrel \cal D \over =}  %% equal in distribution
72: \def\A{{\cal A}}
73: \def\Ah{\widehat{\cal A}}
74: \def\marks{{\Xi}}  % space of marks
75: \def\mk{\xi}  % a mark
76: \def\mkmp{\psi}  % mark map
77: \def\etail#1{#1^-}   % tail of an edge
78: \def\ehead#1{#1^+}   % head of an edge
79: \def\bbar#1{{\overline #1}}
80: \def\fdom{{\ss FC}}  % finite components
81: \def\Ph{{\widehat{\rtd}}}
82: \def\Php{{\widehat{\rtd'}}}
83: \def\ev#1{{\cal #1}} % Events
84: \def\traj{\theta}    % Markov chain trajectory measure
85: \def\invar{{\cal I}}  %% invariant events
86: \def\SS{{\cal S}}   %% shift
87: \def\rln{{\cal R}}   %% relation
88: \def\rlnmap{\phi}
89: \def\assumptions{Let\/ $\P$ be a percolation on a
90: unimodular random network}
91: \def\genassumptions{Let\/ $\P$ be an insertion-tolerant percolation on a
92: unimodular random network}
93: \def\ct{n}  % count edges leaving
94: \def\paths{{\cal P}_*}   % paths in a network
95: \def\invarp{{\invar_\infty}}  % invariant $\sigma$-field for paths
96: \def\fin{{\ss FG_2}}  % finite graphs with 2 distinguished vertices
97: \def\gtwo{{{\cal G}_{**}}}  % networks with 2 distinguished vertices
98: \def\uni{{\cal U}}  % unimodular probability measures on $\GG_*$
99: \def\projection{{\ss prj}}  % projection on the first coordinate
100: \def\dmn{\GG_\#}  % distinct-mark networks
101: \def\itimes{\boxtimes}  % independent product
102: \def\dist{{\rm dist}}  % distance
103: \def\bn{{\rm B}}  % Bernoulli
104: \def\alg{{\ss Alg}}  % algebra
105: \def\affalg{\overline{\ss Alg}}  % affiliated algebra
106: \def\two{\{0, 1\}}
107: \def\expdeg{\overline{\rm deg}}  % expected degree
108: 
109: \def\ins{\Pi}   %% Insertion
110: \def\isoe{\iota_\edge}
111: \def\bde{\bd_\edge}
112: \def\bd{\partial}
113: \def\pc{p_{\rm c}}
114: \def\pcb{p_{\rm c}^{\rm bond}}
115: \def\pu{p_{\rm u}}
116: \def\USF{{\ss USF}}
117: \def\UST{{\ss UST}}
118: \def\MST{{\ss MST}}
119: \def\WUSF{{\ss WUSF}}
120: \def\WMSF{{\ss WMSF}}
121: \def\FUSF{{\ss FUSF}}
122: \def\FMSF{{\ss FMSF}}
123: \def\verts{{\ss V}}
124: \def\vertex{{\ss V}}
125: \def\vertices{{\ss V}}
126: \def\edges{{\ss E}}
127: \def\edge{{\ss E}}
128: \def\Stab{{\ss Stab}}
129: \def\bp{o}
130: \def\isol{\heartsuit}  % isolated vertex (type)
131: \def\lift{\lambda}  % lift probability measure from class to network
132: \def\dual#1{#1^\dagger}
133: \def\collec{{\cal C}}  % collection
134: \def\K{{\cal K}}  % compact set
135: \def\gve{{\cal G}_{*-}}  % network with vertex and edge
136: \def\type{\tau}  % type of a vertex
137: \def\HH{{\cal H}}   % a Hilbert space
138: \def\DD{{\scr D}}   % domain
139: \def\dmn{{\scr D}}   % domain
140: \def\spm{\sigma}  % trace of spectral measure
141: 
142: \def\BLPSusf{\ref b.BLPS:usf/, hereinafter referred to as %
143: \def\BLPSusf{BLPS (\refbyear {BLPS:usf})}\BLPSusf}
144: 
145: \def\BLPSgip{\ref b.BLPS:gip/, hereinafter referred to as %
146: \def\BLPSgip{BLPS (\refbyear {BLPS:gip})}\BLPSgip}
147: 
148: \def\LPSmsf{\ref b.LPS:msf/}
149: 
150: \ifproofmode \relax \else\head{{\it Electron. J. Probab.}, {\bf 12}, Paper
151: 54 (2007), 1454--1508.}
152: {Version of \versiondate}\fi
153: \vglue20pt
154: 
155: \def\firstheader{\eightpoint\ss\underbar{\raise2pt\line 
156: {{\it Electron.\ J.\ Probab.}, {\bf 12}, Paper 54 (2007), 1454--1508. 
157:      \hfil Version of \versiondate}}}
158: 
159: \beginniceheadline
160: 
161: \title{Processes on Unimodular Random Networks}
162: 
163: \author{David Aldous and Russell Lyons}
164: 
165: 
166: %\centerline{\tt NOT FOR DISTRIBUTION}
167: 
168: 
169: 
170: \abstract{
171: We investigate unimodular random networks.
172: Our motivations include 
173: their characterization via reversibility of an associated random walk
174: and their similarities to unimodular quasi-transitive graphs.
175: We extend various theorems concerning random walks, percolation, spanning
176: forests, and amenability from the known context of unimodular
177: quasi-transitive graphs to the more general context of unimodular random
178: networks.
179: We give properties of a trace associated to unimodular random networks
180: with applications to stochastic comparison of
181: continuous-time random walk.
182: }
183: 
184: \bottomII{Primary
185: 60C05. % Combinatorial probability
186: Secondary
187: 60K99,  % special processes
188:  05C80. %   Random graphs
189: }
190: {Amenability, equivalence relations, infinite graphs, percolation,
191: quasi-transitive, random walks, transitivity, weak convergence,
192: reversibility, trace, stochastic comparison, spanning forests,
193: sofic groups.}
194: {Research partially supported by NSF grants DMS-0203062,
195: DMS-0231224, DMS-0406017, and DMS-0705518.}
196: 
197: \articletoc
198: 
199: \bsection{Introduction}{s.intro}
200: 
201: In the setting of infinite discrete graphs,
202: the property of being a Cayley graph of a group is a strong form of 
203: ``spatial homogeneity":
204: many results not true for arbitrary graphs are true under this strong property.
205: As we shall soon explain, weaker regularity properties sufficient for many
206: results have been studied.  
207: In this paper, we turn to random graphs, investigating
208: a notion of
209: ``statistical homogeneity" or ``spatial stationarity"
210: that we call
211: a {\it unimodular random rooted network}.
212: The root is merely a distinguished vertex of the network and the
213: probability measure is on a certain space of rooted networks.
214: In a precise sense, the root is ``equally likely" to be any vertex of the
215: network, even though we consider infinite networks.
216: We shall show that many results known for
217: deterministic graphs under previously-studied regularity conditions 
218: do indeed extend to
219: unimodular random rooted networks.
220: 
221: Thus, a probabilistic motivation for our investigations is the study of
222: stochastic processes under unimodularity.
223: A second motivation is combinatorial: One often asks for asymptotics of
224: enumeration or optimization problems on finite networks as the size of the
225: networks tend to infinity.
226: One can sometimes answer such questions with the aid of a suitable limiting
227: infinite object. A survey of this approach is given by \ref b.AS:obj/.
228: We call ``random weak limit" the type of limit one considers; it is the
229: limiting ``view" from a uniformly chosen vertex of the finite networks.
230: What limiting objects can arise this way?
231: It has been observed before that
232: the probabilistic objects of interest,
233: unimodular random rooted networks, contain all
234: the combinatorial objects of interest, random
235: weak limits of finite networks.
236: One open question is whether these two classes in fact coincide.
237: An affirmative answer would have many powerful consequences, as we shall
238: explain.
239: 
240: To motivate this by
241: analogy, recall a simple fact about stationary sequences
242: $\Seq{Y_i}_{i \in \Z}$
243: of random variables.
244: For each $n \geq 1$, let 
245: $\Seq{Y_{n,i}}_{1 \leq i \leq n}$
246: be arbitrary. Center it at a uniform index 
247: $U_n \in \{1,2,\ldots,n\}$ 
248: to get a bi-infinite sequence
249: $\Seq{Y_{n,U_n+i}}_{i \in \Z}$,
250: interpreted arbitrarily outside its natural range.  
251: If there is a weak limit 
252: $\Seq{Y_i}_{i \in \Z}$ as $n \to \infty$ 
253: of these randomly centered sequences, then the limit is
254: stationary, and conversely any stationary sequence can be obtained
255: trivially as such a limit.
256: 
257: %Though the intrinsic definition
258: %(\ref s.notation/)
259: %is somewhat abstract, its motivation is easy to describe.
260: By analogy, then,
261: given a finite graph, take a uniform random vertex as root.
262: Such a randomly rooted graph automatically has a certain property
263: %(briefly, an associated measure on graphs with a distinguished oriented edge
264: %is invariant under edge-reversal)
265: (in short, if mass is redistributed in the graph, then the expected mass
266: that leaves the root is equal to the expected mass the arrives at the root)
267: and in \ref s.notation/,
268: we abstract this property as unimodularity.
269: It is then immediate that any infinite random rooted graph that is a limit
270: (in an appropriate sense that we call ``random weak limit")
271: of uniformly randomly rooted finite graphs will be unimodular, whereas
272: the above question asks
273: conversely whether any unimodular random rooted graph arises as a random
274: weak limit of some sequence of randomly rooted finite graphs.
275: 
276: Additional motivation for the definition arises from random walk considerations.
277: Given any random rooted graph, simple random walk induces a Markov chain on
278: rooted graphs.
279: Unimodularity of a probability measure $\rtd$ on rooted graphs
280: is equivalent to the property that a reversible stationary
281: distribution for this chain is given by the root-degree biasing of $\rtd$,
282: just as on finite graphs,
283: a stationary distribution for simple random walk is proportional to the
284: vertex degrees; see \ref s.extreme/.
285: 
286: Let us return now to the case of deterministic graphs.
287: An apparently minor relaxation of the Cayley graph property
288: is the ``transitive" property 
289: (that there is an automorphism taking any vertex to any other vertex).
290: By analogy with the shift-invariant interpretation of stationary sequences, one might expect every
291: transitive graph to fit into our set-up.
292: But this is false.
293: Substantial research over the last ten years has shown that
294: the most useful regularity condition is that of a
295: {\it unimodular transitive graph} (or, more generally, quasi-transitive).
296: %It turns out that this lets us make precise in a powerful way 
297: %the intuition that 
298: Intuitively, this is
299: an unrooted transitive graph that can be given 
300: a random root in such a way that each vertex is equally
301: likely to be the root.
302: This notion is, of course, precise in itself for a finite graph.
303: To understand how this is extended to infinite graphs, and then to
304: unimodular random rooted graphs, consider a finite graph $\gh$ and a
305: function $f(x, y)$ of ordered pairs of vertices of $\gh$.
306: Think of $f(x, y)$ as an amount of mass that is sent from $x$ to $y$.
307: Then the total mass on the graph $\gh$ before transport equals the total
308: after, since mass is merely redistributed on the graph.
309: We shall view this alternatively as saying that for a randomly
310: uniformly chosen vertex, the expected mass it receives is equal to the
311: expected mass it sends out.
312: This, of course, depends crucially on choosing the vertex uniformly and,
313: indeed, characterizes the uniform measure among all probability measures on
314: the vertices.
315: 
316: Consider now an infinite transitive graph, $\gh$.
317: Since all vertices ``look the same", we could just fix one, $\bp$,
318: rather than try to choose one uniformly.
319: However, a mass transport function $f$ will not conserve the mass at $\bp$
320: without some assumption on $f$ to make up for the fact that $\bp$ is fixed.
321: Although it seems special at first, it turns out that a very useful
322: assumption is that $f$ is invariant under the diagonal action of
323: the automorphism group of $\gh$.
324: (For a finite graph that happened to have no automorphisms other than the
325: identity, this would be no restriction at all.)
326: This is still not enough to guarantee ``conservation of mass", i.e., that
327: $$
328: \sum_x f(\bp, x) = \sum_x f(x, \bp)
329: \,,
330: \label e.mtp1
331: $$
332: but it turns out that \ref e.mtp1/ does hold when the
333: automorphism group of $\gh$ is unimodular. Here, ``unimodular" is used in
334: its original sense that the group
335: admits a non-trivial Borel measure that is invariant under both left and
336: right multiplication by group elements.
337: We call $\gh$ itself unimodular in that case; see Sections
338: \briefref s.notation/ and \briefref s.related/ for more
339: on this concept.
340: The statement that \ref e.mtp1/ holds under these assumptions is called the
341: Mass-Transport Principle for $\gh$.
342: If $\gh$ is quasi-transitive, rather than transitive, we still have a
343: version of \ref e.mtp1/, but we can no longer consider only one fixed
344: vertex $\bp$.
345: Instead, each orbit of the automorphism group must have a representative
346: vertex.
347: Furthermore, it must be weighted ``proportionally to its frequency" among
348: vertices; see \ref t.fixedgraph/. 
349: This principle was introduced to the study of percolation by \ref
350: b.Hag:deptree/, then developed and exploited heavily by \BLPSgip.
351: Another way of stating it is that \ref e.mtp1/ holds in expectation 
352: when $\bp$ is chosen randomly by an appropriate probability measure.
353: If we think of $\bp$ as the root, then we arrive at the notion of
354: random rooted graphs, and the corresponding statement that \ref e.mtp1/
355: holds in expectation is a general form of the Mass-Transport Principle.
356: This general form was called the ``Intrinsic Mass-Transport Principle" by
357: \ref b.BS:rdl/.
358: %It is, in fact, equivalent to the corresponding probability measure being
359: %unimodular by our definition.
360: We shall call a probability measure on rooted graphs unimodular precisely
361: when this general form of the Mass-Transport Principle holds.
362: We develop this in \ref s.notation/.
363: 
364: Thus, we can extend many results known
365: for unimodular quasi-transitive graphs to our new setting of
366: unimodular random rooted graphs, as noted by \ref b.BS:rdl/.
367: As a bonus, our set-up allows the treatment of
368: quasi-transitive graphs to be precisely parallel to that of
369: transitive graphs, with no additional notation or thought needed, which had
370: not always been the case previously.
371: 
372: 
373: 
374: To state results in their natural generality, as well as for technical
375: convenience, we shall work in the setting of networks,
376: which are just graphs with ``marks" (labels) on edges and vertices.
377: Mainly, this paper is organized to progress from the most general to the
378: most specific models.
379: An exception is made in \ref s.related/, where we discuss random networks
380: on fixed underlying graphs.
381: This will not only help to understand and motivate the general setting, but
382: also will be useful in deriving consequences of our general results.
383: 
384: \ref s.extreme/ elaborates the comment above about reversible stationary
385: distributions for random walk,  discussing extremality and invariant
386: $\sigma$-fields, speed of random walk, and continuous-time random walk and
387: their explosions.
388: \ref s.trace/ discusses a trace associated to unimodular random networks
389: and comparison of return probabilities of different continuous-time random
390: walks, which partially answers a question of Fontes and Mathieu.
391: We then write out the extensions to unimodular random rooted graphs of results
392: known for fixed graphs in the context of
393: percolation (\ref s.extend/),
394: spanning forests (\ref s.span/)
395: and amenability (\ref s.amen/).
396: These extensions are in most (though not all) cases straightforward.
397: Nevertheless, we think it is useful to list these extensions in the order
398: they need to be proved so that others need not check the entire (sometimes
399: long) proofs or chains of theorems from a variety of papers.
400: Furthermore, we were required to find several essentially new results along
401: the way.
402: %Moreover, occasionally our approach leads to a simpler and shorter proof
403: %than the original, albeit one based on our fundamental finite-approximation
404: %result, \ref t.TII/.
405: 
406: In order to appreciate the scope of our results, we list many examples of
407: unimodular probability measures in \ref s.ex/.
408: In particular, there is a significant and important overlap between our
409: theory and the theory of graphings of measure-preserving equivalence
410: relations. This overlap
411: is well known among a few specialists, but
412: deserves to be made more explicit. We do that here in \ref x.equivalence/.
413: 
414: Among several open problems, we spotlight a special case of
415: \ref q.coupling/: Suppose we
416: are given a partial order on the mark space and two unimodular probability
417: measures, one stochastically dominating the other. That is, there is
418: a monotone coupling of the two unimodular distributions that puts the
419: networks on the same graphs, but has higher marks for the second network
420: than for the first. Does this
421: imply the existence of a {\it unimodular\/} monotone coupling?
422: A positive answer would be of great benefit in a variety of ways.
423: 
424: Another especially important open question is \ref q.TII/, whether every
425: unimodular probability measure is a limit of uniformly rooted finite
426: networks.
427: For example, in the case that the random rooted infinite network is just a
428: Cayley graph (rooted, say, at the identity) with the edges marked by the
429: generators, a positive answer to this question on finite approximation would
430: answer a question of \ref b.Weiss:sofic/, by showing that all finitely
431: generated groups are ``sofic", although this is contrary to the belief
432: expressed by \ref b.Weiss:sofic/.
433: (Sofic groups were introduced, with a different definition, by \ref
434: b.Gromov:endo/; see \ref b.Elek:sofic/ for a proof that the definitions are
435: equivalent.)
436: This would establish several conjectures, since they are known to hold for
437: sofic groups: the direct finiteness
438: conjecture of \ref b.Kaplansky/ on group algebras
439: (see \ref b.Elek:sofic/), a
440: conjecture of \ref b.Gottschalk/ on ``surjunctive" groups in topological
441: dynamics
442: (see \ref b.Gromov:endo/), the Determinant
443: Conjecture on Fuglede-Kadison determinants
444: (see \ref b.Elek:hyper/), and Connes's (\refbyear{Connes})
445: Embedding Conjecture
446: for group von Neumann algebras (see \ref b.Elek:hyper/).
447: %The latter conjecture is referred to by saying that all countable groups are
448: %``hyperlinear".
449: %Since our work combined with that of \ref b.Elek:hyper/ shows that this is
450: %true, we ``refute the famous `theorem' of Gromov claiming that a
451: %proposition which holds for all countable discrete groups is either trivial
452: %or false", in the words of \ref b.Ozawa/.
453: The Determinant Conjecture in turn implies the Approximation Conjecture of
454: \ref b.Schick/ and the Conjecture of Homotopy Invariance of $L^2$-Torsion
455: due to \ref b.Luck:3man/; see Chap.~13 of \ref b.Luck:book/ for these
456: implications and more information.
457: \ref b.Weiss:sofic/ gave another proof
458: of Gottschalk's conjecture for sofic groups.
459: One may easily extend that proof to show a form of Gottschalk's conjecture 
460: for all quasi-transitive unimodular
461: graphs that are limits of finite graphs, but there are easy
462: counterexamples for general transitive graphs.
463: 
464: Further discussion of the
465: question on approximation by finite networks is given in
466: \ref s.finite/.
467: A positive answer would
468: provide solid support for the intuition that the root of a
469: unimodular random rooted network is equally likely to be any vertex.
470: \ref s.finite/ also contains some variations that would result from a
471: positive answer and some
472: additional consequences for deterministic graphs.
473: 
474: The notion of weak convergence of rooted locally finite graphs or networks  
475: (needed to make sense of convergence of randomly rooted
476: finite graphs to a limit infinite graph)
477: has arisen before in several different contexts.
478: Of course, the special case where the limit network is a Cayley diagram
479: was introduced by \ref b.Gromov:endo/ and \ref b.Weiss:sofic/.
480: In the other cases, the limits provide examples of unimodular
481: random rooted graphs.
482: \ref b.Aldous:fringe/
483: gives many examples of models of random finite trees
484: which have an infinite-tree limit 
485: (and one such example, the limit of uniform random labeled trees
486: being what is now called the 
487: Poisson-Galton-Watson tree,
488: ${\PGW}^\infty(1)$, goes back to 
489: \ref b.Grimmett:ltrees/).
490: The idea that random weak limits of finite planar graphs of
491: uniformly bounded degree provide an interesting class of infinite planar
492: graphs was developed by \ref b.BS:rdl/, who showed that random walk on
493: almost any such limit graph is recurrent.
494: (Thus, such graphs do not include regular trees or hyperbolic graphs,
495: other than trivial examples like $\Z$.)
496: A specialization to random weak limits of
497: plane triangulations was studied in more detail in
498: interesting recent work of \ref b.AngelSchramm/ and \ref b.Angel/.
499: %Another type of approximation is due to \ref b.VershikGordon/.
500: 
501: \ref x.PWIT/ describes an infinite-degree tree, arising as a limit
502: of weighted finite complete graphs. 
503: This example provides an interface between our setting and 
504: related ideas of 
505: ``local weak convergence"
506: and ``the objective method in the probabilistic analysis of algorithms".
507: A prototype is that the distribution of $n$ random points in a square of area $n$
508: converges in a natural sense as $n \to \infty$ to the distribution
509: of a Poisson point process on the plane of unit intensity.
510: One can ask whether solutions of combinatorial optimization
511: problems over the $n$ random points 
512: (minimum spanning tree, minimum matching, traveling salesman problem)
513: converge to limits that are the solutions of analogous
514: optimization problems over the Poisson point process in the whole plane.
515: \ref x.PWIT/ can be regarded as a mean-field analogue of random
516: points in the plane, and $n \to \infty$ limits of solutions of combinatorial optimization problems
517: within this model have been studied using the non-rigorous
518: cavity method from statistical physics.  
519: \ref b.APercus:SU/ illustrate what can be done by non-rigorous means,
520: while \ref b.AS:obj/ survey introductory rigorous theory. 
521: 
522: 
523: The reader may find it helpful to keep in mind one additional example,
524: a unimodular version of family trees of Galton-Watson branching
525: processes; see also \ref x.fdd/.
526: 
527: \procl x.AGW \procname{Unimodular Galton-Watson}
528: Let $\Seq{p_k \st k \ge 0}$ be a probability distribution on $\N$.
529: Take two independent Galton-Watson trees with offspring distribution
530: $\Seq{p_k}$, each starting with one particle, the root,
531: and join them by a new edge
532: whose endpoints are their roots. Root the new tree
533: at the root of the first Galton-Watson
534: tree. This is {\bf augmented Galton-Watson}
535: measure, \AGW. (If $p_0 \ne 0$, then we have the
536: additional options to condition on either
537: non-extinction or extinction of the joined trees.) Now bias by the
538: reciprocal of the degree of the root to get {\bf unimodular Galton-Watson}
539: measure, $\UGW$.
540: In different language, \ref b.LPP:GW/ proved that this measure,
541: $\UGW$, is unimodular.
542: Note that
543: the mean degree of the root is
544: $$
545: \expdeg(\UGW) = \left(\sum_{k \ge 0} {p_k \over k+1} \right)^{-1}
546: \,.
547: \label e.degAGW
548: $$
549: %We shall revisit this example in the contexts of limits of finite graphs
550: %and of random walks.
551: \endprocl
552: 
553: 
554: 
555: 
556: 
557: \bsection{Definitions and Basics}{s.notation}
558: 
559: We denote a (multi-)graph $\gh$ with vertex set $\verts$ and undirected
560: edge set $\edges$ by $\gh = (\verts, \edges)$.
561: When there is more than one graph under discussion, we write $\verts(\gh)$ or
562: $\edges(\gh)$ to avoid ambiguity.
563: We denote the degree of a vertex $x$ in a graph $\gh$ by $\deg_\gh(x)$.
564: {\bf Simple random walk} on $\gh$ is the Markov chain whose state space is
565: $\verts$ and whose transition probability from $x$ to $y$ equals the number of
566: edges joining $x$ to $y$ divided by $\deg_\gh(x)$.
567: 
568: A {\bf network} is a (multi-)graph $\gh = (\vertex, \edge)$ together with a
569: complete separable metric space $\marks$ called the {\bf mark space} and
570: maps from $\vertex$ and $\edge$ to $\marks$. Images in $\marks$ are called
571: {\bf marks}.
572: Each edge is given two marks, one associated to (``at") each of its endpoints.
573: The only assumption on degrees is that they are finite.
574: We shall usually assume that $\marks$ is Baire space $\N^\N$, since every
575: uncountable complete separable metric space is Borel isomorphic to Baire
576: space by Kuratowski's theorem (Theorem 15.10 of \ref b.Royden/).
577: We generally omit mention of the mark maps from our notation for
578: networks when we do not need them.
579: For convenience, we consider graphs as special cases of networks in which
580: all marks are equal to some fixed mark.
581: 
582: We now define ends in graphs.
583: In the special case of a tree, an infinite path that starts at any vertex
584: and does not backtrack is called a {\bf ray}. Two rays are {\bf equivalent}
585: if they have
586: infinitely many vertices in common. An equivalence class of rays is
587: called an {\bf end}.
588: In a general infinite graph, $\gh$, an {\bf end} of $\gh$ is an equivalence
589: class of infinite simple paths in $\gh$, where two paths
590: are equivalent if for every finite $K\subset \vertex(\gh)$,
591: there is a connected component of $\gh\setminus K$ that intersects
592: both paths.
593: 
594: Let $\gh$ be a graph.  For a subgraph $H$, let its {\bf (internal) vertex
595: boundary} $\bdv H$ be the set of vertices of $H$ that are adjacent to some
596: vertex not in $H$.
597: We say that $\gh$ is {\bf (vertex) amenable}
598: if there exists a sequence of subsets $H_n\subset \vertex(\gh)$
599: with
600: $$
601: \lim_{n \to\infty} {|\bdv H_n| \over |\verts(H_n)|} = 0\,,
602: $$
603: where $|\cbuldot|$ denotes cardinality.
604: Such a sequence is called a {\bf F\o{}lner sequence}.
605: A finitely generated group is {\bf amenable} if its Cayley graph is amenable.
606: For example, every finitely generated abelian group is amenable.
607: For more on amenability of graphs and groups, see \BLPSgip.
608: 
609: A {\bf homomorphism} $\varphi: \gh_1 \to \gh_2$ from one
610: graph $\gh_1=(\verts_1,\edges_1)$ to another $\gh_2=(\verts_2,\edges_2)$
611: is a pair of maps $\varphi_\verts:\verts_1\to\verts_2$ and
612: $\varphi_\edges:\edges_1\to\edges_2$
613: such that $\varphi_\verts$ maps the endpoints of $e$ to the endpoints
614: of $\varphi_\edges(e)$ for every edge $e \in \edges_1$.
615: When both maps $\varphi_\verts:\verts_1\to\verts_2$ and
616: $\varphi_\edges:\edges_1\to\edges_2$ are bijections, then $\varphi$ is
617: called an {\bf isomorphism}.
618: When $\gh_1 = \gh_2$, an isomorphism is called
619: an {\bf automorphism}.
620: The set of all automorphisms of $\gh$ forms a group under composition,
621: denoted by $\Aut(\gh)$.
622: The action of a group $\gp$ on a graph $\gh$ by automorphisms is said to be {\bf
623: transitive} if there is only one $\gp$-orbit in $\verts(\gh)$ and to be {\bf
624: quasi-transitive} if there are only finitely many orbits in $\verts(\gh)$.
625: A graph $\gh$ is {\bf transitive} or {\bf quasi-transitive} according as
626: whether the corresponding action of $\Aut(\gh)$ is.
627: %if for every $x,y\in\verts$, there is an
628: %automorphism of $\gh$ taking $x$ to $y$.
629: %In other words, the orbit
630: %space $\verts/\Aut(\gh)$ of $\verts$ under the group $\Aut(\gh)$ of
631: %automorphisms of $\gh$ is a singleton.
632: For example, every Cayley graph is transitive.
633: %A graph is {\bf quasi-transitive} if
634: %the orbit space $\verts/\Aut(\gh)$ is finite.
635: All the same terms are applied to networks when the maps in question preserve
636: the marks on vertices and edges.
637: 
638: A locally compact group is called {\bf unimodular} if its
639: left Haar measure is also right invariant.
640: In particular, every discrete countable group is unimodular.
641: We call a graph $\gh$ {\bf unimodular} if $\Aut(\gh)$ is unimodular, where
642: $\Aut(\gh)$ is given the weak topology generated by its action on $\gh$.
643: Every Cayley graph and, as \ref b.SoardiWoess/ and \ref b.Salvatori/ proved,
644: every quasi-transitive amenable graph is unimodular.
645: See \ref s.related/ and \BLPSgip\ for more details on unimodular graphs.
646: 
647: A {\bf rooted network} $(\gh, \bp)$ is a network $\gh$ with a distinguished
648: vertex $\bp$ of $\gh$, called the {\bf root}.
649: A {\bf rooted isomorphism} of rooted networks is an isomorphism of the
650: underlying networks that takes the root of one to the root of the other.
651: We generally do not distinguish between a rooted network and its
652: isomorphism class.
653: When needed, however, we use the following notation to make these distinctions:
654: $\gh$ will denote a graph, $\overline \gh$ will denote a network with underlying
655: graph $\gh$, and $[\overline \gh, \bp]$ will denote the class of rooted networks
656: that are rooted-isomorphic to $(\overline \gh, \bp)$.
657: We shall use the following notion introduced (in slightly different
658: generalities) by \ref b.BS:rdl/ and \ref b.AS:obj/.
659: %Given a positive integer $R$, a finite rooted network $H$, and a probability
660: %distribution $\rtd$ on rooted networks with a fixed mark space, let $p(R, H,
661: %\rtd)$ denote the probability that $H$ is rooted isomorphic to the ball of
662: %radius $R$ about the root of a network chosen with distribution $\rtd$.
663: %A sequence $\rtd_n$ is said to {\bf converge weakly} to $\rtd$, written $\rtd_n
664: %\cd \rtd$, if $\lim_{n \to\infty} p(R, H, \rtd_n) = p(R, H, \rtd)$
665: %for any positive integer $R$ and any finite network $H$.
666: %
667: %This agrees with the usual sense of weak convergence of probability
668: %measures on a metric space as follows.
669: Let $\GG_*$ denote the set of rooted isomorphism classes of rooted
670: {\it connected\/} locally finite networks.
671: Define a metric on $\GG_*$ by letting the distance between $(G_1, o_1)$ and
672: $(G_2, o_2)$ be $1/(1+\alpha)$, where $\alpha$ is
673: the supremum of those $r > 0$ such that
674: %(if we take
675: %$(G_i, o_i)$ to be representatives of their isomorphism classes) there is a
676: %network $(G, o)$ whose underlying graph is
677: %rooted isomorphic to subgraphs of the graphs
678: %underlying both $(G_i, o_i)$ with the properties
679: %\smallskip
680: %\item{(i)} every vertex of finite degree within distance $1/\epsilon$ of
681: %$o_i$ belongs to the image of $G$;
682: %\item{(ii)} every vertex of infinite degree within distance $1/\epsilon$ of
683: %$o_i$ that belongs to the image of $G$ has degree at least $1/\epsilon$
684: %in $G$; and
685: %\item{(iii)} every mark in $G$ is within distance $\epsilon$ of the
686: %corresponding marks in $G_i$.
687: %\smallskip
688: %\noindent
689: there is some rooted isomorphism of
690: the balls of (graph-distance) radius $\flr{r}$ around the roots of $G_i$ 
691: such that each pair of corresponding marks has distance less than $1/r$.
692: It is clear that $\GG_*$ is separable and complete in this metric.
693: For probability measures $\rtd$, $\rtd_n$ on $\GG_*$, we write $\rtd_n \cd
694: \rtd$ when $\rtd_n$ converges weakly with respect to this metric.
695: 
696: For a probability measure $\rtd$ on rooted networks, write $\expdeg(\rtd)$
697: for the {\it expectation\/} of the degree of the root with respect to
698: $\rtd$.
699: In the theory of measured equivalence relations (\ref x.equivalence/),
700: this is twice the {\bf
701: cost} of the graphing associated to $\rtd$.
702: Also, by the {\bf degree} of $\rtd$ we mean the {\it distribution\/} of
703: the degree of the root under $\rtd$.
704: 
705: For a locally finite connected rooted network, there is a canonical choice
706: of a rooted network in its rooted-isomorphism class.
707: More specifically, there is a continuous map $f$ from $\GG_*$ to the space of
708: networks on $\N$ rooted at 0 such that $f\big([\gh, \bp]\big) \in [\gh,
709: \bp]$ for all $[\gh, \bp] \in \GG_*$.
710: To specify this, consider the following total ordering on rooted networks
711: with vertex set $\N$ and root $0$.
712: First, total order $\N \times \N$ by the lexicographic order:
713: $(i_1, j_1) \prec (i_2, j_2)$ if
714: either $i_1 < i_2$ or $i_1 = i_2$ and $j_1 < j_2$.
715: Second, the lexicographic order on Baire space $\marks$ is also a total order.
716: We consider networks on $\N$ rooted at $0$.
717: Define a total order on such networks as follows.
718: Regard the edges as oriented for purposes of identifying the edges with $\N
719: \times \N$; the mark at $i$ of an edge between $i$ and $j$ will be considered
720: as the mark of the oriented edge $(i, j)$.
721: Suppose we are given a pair of networks on $\N$ rooted at $0$.
722: If they do not have the same edge sets, then the network that contains the
723: smallest edge in their symmetric difference is deemed to be the smaller
724: network.
725: If they do have the same edge sets, but not all the vertex marks are the same,
726: then the network that contains the vertex with the smaller mark on the
727: least vertex where they differ is deemed the smaller network.
728: If the networks have the same edge sets and the same vertex marks, but not all
729: the edge marks are the same, then
730: the network that contains the oriented edge with the smaller mark on the
731: least oriented edge where they differ is deemed the smaller network.
732: Otherwise, the networks are identical.
733: 
734: We claim that the rooted-isomorphism class of each locally finite connected
735: network contains a unique smallest rooted network on $\N$ in the above
736: ordering.
737: This is its {\bf canonical} representative.
738: To prove our claim, consider only the networks in the class that have
739: vertex set $\N$ and are rooted at $0$.
740: It is fairly easy to see that there is a smallest graph that supports a
741: network in the class:
742: the neighbors of 0 are $[1, n_1]$ for some $n_1 \ge 1$,
743: the neighbors of 1, besides 0, are $[n_1+ 1, n_2]$ for some $n_2 \ge n_1$, the
744: neighbors of 2, other than possibly 0 and 1, are $[n_2+1, n_3]$ for some
745: $n_3\ge n_2$, etc.
746: In general, the neighbors of $k \ge 1$ that are larger than $k$ are $[n_k+1,
747: n_{k+1}]$ for some $n_{k+1} \ge n_k$.
748: Write $n_0 := 0$.
749: Let $A_{-1}$ be the set of networks in the class that are on this smallest
750: graph.
751: All have the same mark at $0$.
752: Write $\mkmp(k)$ for the mark at a vertex $k$ and $\mkmp(j, k)$ for the
753: mark at the oriented edge $(j, k)$.
754: Define $A_k$ recursively as follows.
755: Given $A_{k-1}$, let $A_k$ be the subset of networks in $A_{k-1}$ such that
756: the marks $\mkmp(j)$ are increasing for $j \in [n_k+1, n_{k+1}]$.
757: Then let $B_{-1} := \bigcap_k A_k$.
758: %If the graph has no other vertices, then $A_0$ is a singleton and the
759: %canonical representative.
760: %Suppose there are other vertices, so that $n_1$ defined above is at least 1.
761: %Let $A_1$ be the subset of $A_0$ where the marks $\mkmp(k)$ are increasing
762: %in $k \in [1, n_1]$.
763: This is non-empty.
764: Now define $B_k$ recursively by letting $B_k$ be the subset of $B_{k-1}$
765: such that $j \mapsto \mkmp(k, j)$ is increasing on $[n_k+1, n_{k+1}]$.
766: The set $\bigcap_k B_k$ is then a singleton, the desired canonical
767: representative.
768: 
769: 
770: For a (possibly disconnected)
771: network $\gh$ and a vertex $x \in \verts(\gh)$, write $\gh_x$ for the
772: connected component of $x$ in $\gh$.
773: If $\gh$ is a network with probability distribution $\rtd$ on its
774: vertices, then $\rtd$ induces naturally a distribution on $\GG_*$,
775: which we also denote by $\rtd$; namely, the probability of $(\gh_x, x)$ is
776: $\rtd(x)$.
777: More precisely, $\rtd\big([\gh_x, x]\big) := \sum \big\{ \rtd(y) \st y \in
778: \verts(\gh), \ (\gh_y, y) \in [\gh_x, x] \big\}$.
779: For a finite network $\gh$, let $U(\gh)$ denote the distribution on $\GG_*$
780: obtained this way by choosing a uniform random vertex of $\gh$ as root.
781: Suppose that $\gh_n$ are finite networks and that $\rtd$ is a
782: probability measure on $\GG_*$.
783: We say the {\bf random weak limit} of $\gh_n$ is $\rtd$ if $U(\gh_n) \cd \rtd$.
784: %Note that only the component of the root matters for convergence to $\rtd$.
785: If $\rtd\Big(\big\{[\gh, \bp]\big\}\Big) = 1$ for a fixed transitive
786: network $\gh$ (and (any) $\bp \in \vertex(\gh)$), then we say that the {\bf
787: random weak limit} of $\gh_n$ is $\gh$.
788: 
789: As usual, call a collection $\collec$
790: of probability measures on $\GG_*$
791: {\bf tight} if for each $\epsilon > 0$, there is a
792: compact set $\K \subset \GG_*$ such that $\rtd(\K) > 1 - \epsilon$ for all
793: $\rtd \in \collec$.
794: Because $\GG_*$ is complete, any tight collection
795: has a subsequence that possesses a weak limit.
796: %If $\rtd_1$ and $\rtd_2$ are probability measures on $\verts(\gh)$ such that for
797: %any positive integer $R$ and any finite graph $H$, we have $p(R, H, \rtd_1) =
798: %p(R, H, \rtd_2)$, then we write $\rtd_1 \equiv \rtd_2$.
799: %If $\gh$ is a regular graph, write $d_\gh$ for its degree.
800: 
801: 
802: 
803: The class of probability measures
804: $\rtd$ that arise as random weak limits of finite networks is contained in
805: the class of unimodular $\rtd$, which we now define.
806: Similarly to the space $\GG_*$, we define the space $\gtwo$ of isomorphism
807: classes of locally
808: finite connected networks with an ordered pair of distinguished vertices
809: and the natural topology thereon.
810: We shall write a function $f$ on $\gtwo$ as $f(\gh, x, y)$.
811: 
812: \procl d.unimodular 
813: Let $\rtd$ be a probability measure on $\GG_*$.
814: We call $\rtd$ {\bf unimodular} if it obeys the {\bf Mass-Transport
815: Principle}:
816: For all Borel
817: %universally measurable\ftnote{*}{This means ``measurable with
818: %respect to every Borel measure".}
819: $f : \gtwo \to [0, \infty]$,
820: %and invariant in the sense that for any (non-rooted)
821: %network isomorphism $\gpe$ of $\gh$ and any $x, y \in \vertex(\gh)$, we have
822: %$f(\gpe \gh, \gpe x, \gpe y) = f(\gh, x, y)$.
823: we have
824: $$
825: \int \sum_{x \in \vertex(\gh)} f(\gh, \bp, x) \,d\rtd\big([\gh, \bp]\big)
826: =
827: \int \sum_{x \in \vertex(\gh)} f(\gh, x, \bp) \,d\rtd\big([\gh, \bp]\big)
828: \,.
829: \label e.mtpgen
830: $$
831: Let $\uni$ denote the set of unimodular Borel probability measures on $\GG_*$.
832: \endprocl
833: 
834: Note that to define the sums that occur here, we choose a specific network
835: from its rooted-isomorphism class, but which one we choose makes no
836: difference when the sums are computed.
837: We sometimes call $f(\gh, x, y)$ the amount of ``mass" sent from $x$ to $y$.
838: The motivation 
839: for the name ``unimodular" is two fold: One is the extension of the concept
840: of unimodular automorphism groups of networks. The second is
841: that the Mass-Transport
842: Principle expresses the equality of two measures on $\gtwo$
843: associated to $\rtd$, the ``left" measure $\rtd_{\rm L}$ defined by
844: $$
845: \int_\gtwo f \,d\rtd_{\rm L}
846: :=
847: \int_{\GG_*} \sum_{x \in \vertex(\gh)} f(\gh, \bp, x) \,d\rtd\big([\gh,
848: \bp]\big)
849: $$ 
850: and the ``right" measure $\rtd_{\rm R}$ defined by
851: $$
852: \int_\gtwo f \,d\rtd_{\rm R}
853: :=
854: \int_{\GG_*} \sum_{x \in \vertex(\gh)} f(\gh, x, \bp) \,d\rtd\big([\gh,
855: \bp]\big)
856: \,.
857: $$
858: Thus, $\rtd$ is unimodular iff $\rtd_{\rm L} = \rtd_{\rm R}$, which can
859: also be expressed by saying that the left measure is absolutely continuous
860: with respect to the right measure and has Radon-Nikod\'ym derivative 1.
861: 
862: It is easy to see that any $\rtd$ that is a random weak limit of finite
863: networks is unimodular, as observed by \ref b.BS:rdl/, who introduced this
864: general form of the Mass-Transport Principle under the name 
865: ``intrinsic Mass-Transport Principle".
866: The converse is open.
867: 
868: A special form of the Mass-Transport Principle was considered, in different
869: language, by \ref b.AS:obj/.
870: Namely, they defined $\rtd$ to be {\bf involution invariant} if \ref
871: e.mtpgen/ holds for those $f$ supported on $(\gh, x, y)$ with $x \sim y$.
872: In fact, the Mass-Transport Principle holds for general $f$ if it holds for
873: these special $f$:
874: 
875: \procl p.invinv
876: A measure is involution invariant iff it is unimodular.
877: \endprocl
878: 
879: \proof
880: Let $\rtd$ be involution invariant.
881: The idea is to send the mass from $x$ to $y$ by single steps, equally
882: spread among the shortest paths from $x$ to $y$.
883: For the proof, we may assume that $f(\gh, x, y) = 0$ unless $x$ and $y$ are
884: at a fixed distance, say $k$, from each other, since any $f$ is a sum of
885: such $f$.
886: Now write $L(\gh, x, y)$ for the set of paths of length $k$ from $x$ to $y$.
887: Let $n_j(\gh, x, y; z, w)$ be the number of paths in $L(\gh, x, y)$ such
888: that the $j$th edge goes from $z$ to $w$.
889: Define $f_j(\gh, z, w)$ for $1 \le j \le k$ and $z, w \in \verts(\gh)$
890: by
891: $$
892: f_j(\gh, z, w) :=
893: \sum_{x, y \in \verts(\gh)} {f(\gh, x, y) n_j(\gh, x, y; z, w) \over
894: |L(\gh, x, y)|}
895: \,.
896: $$
897: Then $f_j(\gh, z, w) = 0$ unless $z \sim w$. Furthermore,
898: $f_j(\gh, z, w) := f_j(\gh', z', w')$ if
899: $(\gh, z, w)$ is isomorphic to $(\gh', z', w')$.
900: Thus, $f_j$ is well defined and Borel on $\gtwo$,
901: whence involution invariance gives us
902: $$
903: \int \sum_{x \in \vertex(\gh)} f_j(\gh, \bp, x) \,d\rtd(\gh, \bp)
904: =
905: \int \sum_{x \in \vertex(\gh)} f_j(\gh, x, \bp) \,d\rtd(\gh, \bp)
906: \,.
907: $$
908: On the other hand,
909: $$
910: \sum_{x \in \vertex(\gh)} f(\gh, \bp, x)
911: =
912: \sum_{x \in \vertex(\gh)} f_1(\gh, \bp, x)
913: \,,
914: $$
915: $$
916: \sum_{x \in \vertex(\gh)} f(\gh, x, \bp)
917: =
918: \sum_{x \in \vertex(\gh)} f_k(\gh, x, \bp)
919: \,,
920: $$
921: and for $1 \le j < k$, we have
922: $$
923: \sum_{x \in \vertex(\gh)} f_j(\gh, x, \bp)
924: =
925: \sum_{x \in \vertex(\gh)} f_{j+1}(\gh, \bp, x)
926: \,.
927: $$
928: Combining this string of equalities yields the desired equation for
929: $f$.
930: \Qed
931: 
932: 
933: \comment{
934: If $\gh$ is a network, $x \in \verts(\gh)$, and $e \in \edges(\gh)$, then
935: the {\bf isomorphism class} of $(\gh, x, e)$ is the class $\{(\gpe \gh, \gpe
936: x, \gpe e) \st \gpe \hbox{ is an isomorphism of }\gh\}$.
937: Let $\gve$ be the set of such isomorphism classes with the natural
938: topology.
939: Given a rooted network $(\gh, x)$ and an edge $e$ incident to $x$, define
940: the involution $\iota(\gh, x, e) := (\gh, y, e)$, where $y$ is the other
941: endpoint of $e$.
942: Given a probability measure $\rtd$ on rooted networks, define the probability
943: measure $\widehat\rtd$ to be the law of the isomorphism class of
944: $(\gh, x, e)$, where $(\gh, x)$ is chosen
945: according to $\rtd$ and $e$ is then chosen uniformly among the edges
946: incident to $x$.
947: Thus,
948: $$
949: \widehat\rtd(\A) :=
950: \int_{\GG_*} {1 \over \deg_\gh x} |\{e \sim x \st (\gh, x, e) \in \A\}|
951: \,d\rtd\big([\gh, x]\big)
952: $$
953: for Borel sets $\A \subseteq \gve$.
954: (Note that we actually choose $[\gh, x]$ according to $\rtd$, then choose
955: $(\gh, x) \in [\gh, x]$ arbitrarily; this latter choice has no effect on
956: the number of $e \sim x$ such that $(\gh, x, e) \in \A$.)
957: Also, define $\widetilde \rtd$ to be the (non-probability) measure that is the
958: result of biasing $\widehat\rtd$ by the degree of the root; that is, the
959: Radon-Nikod\'ym derivative of $\widetilde \rtd$ with respect to $\widehat\rtd$
960: at the isomorphism class of $(\gh, x, e)$ is $\deg_\gh(x)$:
961: $$
962: \widetilde\rtd(\A) :=
963: \int_{\GG_*} |\{e \sim x \st (\gh, x, e) \in \A\}|
964: \,d\rtd\big([\gh, x]\big)
965: $$
966: for Borel sets $\A \subseteq \gve$.
967: (If the expected degree of the root is finite, one could obtain a
968: probability measure from $\widetilde \rtd$ by dividing by the expected degree;
969: but this is not always the case.)
970: The involution $\iota$ induces a pushforward map $\widetilde\rtd \mapsto
971: \widetilde\rtd \circ \iota^{-1}$.
972: We say that $\rtd$ is {\bf unimodular} or {\bf involution invariant} if
973: $\widetilde\rtd \circ \iota^{-1}= \widetilde\rtd$.
974: \comment{This definition works only for locally finite networks. We must resort
975: to the definition by \ref b.AS:obj/ in the general case. The latter seems much
976: harder for probabilists to comprehend.}
977: }%
978: 
979: %It turns out that the proof of our finite-approximation result, \ref
980: %t.TII/, will use only involution invariance.
981: %Thus, an alternative proof of \ref p.invinv/ can be given from the fact
982: %that unimodularity holds for random weak limits of finite networks, in
983: %combination with \ref t.TII/.
984: 
985: Occasionally one uses the Mass-Transport Principle for functions $f$
986: that are not nonnegative. It is easy to see that this use is justified
987: when
988: $$
989: \int \sum_{x \in \vertex(\gh)} |f(\gh, \bp, x)| \,d\rtd(\gh, \bp)
990: < \infty
991: \,.
992: $$
993: 
994: As noted by Oded Schramm (personal communication, 2004), unimodularity can
995: be defined for probability measures on other structures, such as
996: hypergraphs, while involution invariance is limited to graphs (or networks
997: on graphs).
998: 
999: \comment{
1000: Suppose that $\gh$ is an infinite quasi-transitive amenable connected network.
1001: Choose one element $o_i$ from each vertex orbit.
1002: It is shown in \BLPSgip, Proposition 3.6, that there is a probability
1003: measure $\rho$ on the set $ \{ o_i \}$ such that for any F\o{}lner sequence
1004: $H_n$, the relative frequency of vertices in $H_n$ that are in the same orbit
1005: as $o_i$ converges to $\rho(o_i)$. We call this measure $\rho$ the {\bf
1006: natural frequency distribution} of $\gh$.
1007: }%
1008: 
1009: We shall sometimes use the following property of marks.
1010: Intuitively, it says that each vertex has positive probability to be the
1011: root.
1012: 
1013: \procl l.unmark \procname{Everything Shows at the Root}
1014: Suppose that $\rtd$ is a unimodular probability measure on $\GG_*$.
1015: Let $\mk_0$ be a fixed mark and $\marks_0$ be a fixed 
1016: Borel set of marks.
1017: If the mark of the root is a.s.\ $\mk_0$, then the mark of every vertex is
1018: a.s.\ $\mk_0$.
1019: If every edge incident to the root a.s.\
1020: has its edge mark at the root in $\marks_0$,
1021: then all edge marks a.s.\ belong to $\marks_0$.
1022: \endprocl
1023: 
1024: \proof
1025: In the first case, each vertex
1026: sends unit mass to each vertex with a mark different from $\mk_0$.
1027: The expected mass received at the root is
1028: zero. Hence the expected mass sent is 0.
1029: The second case is a consequence of the first, where we put the mark $\mk_0$
1030: at a vertex when all the edge marks at that vertex lie in $\marks_0$.
1031: \Qed
1032: 
1033: When we discuss percolation in \ref s.extend/, we shall find it crucial
1034: that we have a unimodular coupling of the various measures (given by the
1035: standard coupling of Bernoulli percolation in this case).
1036: It would also be very useful to have unimodular couplings in more general
1037: settings.
1038: We now discuss what we mean.
1039: 
1040: Suppose that $\rln \subseteq \marks \times \marks$ is a closed set, which we
1041: think of as a binary relation such as the lexicographic order on Baire space.
1042: Given two measures $\rtd_1, \rtd_2 \in \uni$, say that
1043: $\rtd_1$ is {\bf $\rln$-related} to $\rtd_2$ if there is a probability measure
1044: $\nu$, called an {\bf $\rln$-coupling of $\rtd_1$ to $\rtd_2$},
1045: on rooted networks with mark space
1046: $\marks \times \marks$ such that $\nu$ is concentrated on networks all of
1047: whose marks lie in $\rln$ and whose marginal given by taking the $i$th
1048: coordinate of each mark is $\rtd_i$ for $i = 1, 2$.
1049: In particular, $\rtd_1$ and $\rtd_2$ can be coupled to have the same
1050: underlying rooted graphs.
1051: \comment{
1052: Call $\rln$ {\bf extra separable} if there is a sequence $\Seq{\mk_l}$ and two
1053: Borel maps $\rlnmap_i : \marks \to 2^{\Seq{\mk_l}}$ ($i = 1, 2$)
1054: such that for all $(\mk, \mk') \in \rln$, for all $\zeta_1 \in
1055: \rlnmap_1(\mk)$, and for all $\zeta_2 \in \rlnmap_2(\mk')$, we have $(\zeta_1,
1056: \zeta_2) \in \rln$.
1057: For example, this is the case with the relation $ \le $ on $\R$, where, e.g.,
1058: we may take $\Seq{\mk_l} := \Q$, $\rlnmap_1(\mk) := (-\infty, \mk) \cap \Q$,
1059: and $\rlnmap_2(\mk) := (\mk, \infty) \cap \Q$.
1060: }%
1061: 
1062: It would be very useful to have a positive answer to the following question.
1063: Some uses are apparent in \ref s.trace/ and in \ref s.finite/,
1064: while others appear in \ref b.Lyons:est/ and are hinted at elsewhere.
1065: 
1066: \procl q.coupling \procname{Unimodular Coupling}
1067: Let $\rln \subseteq \marks \times \marks$ be a closed set.
1068: %that is extra separable.
1069: If $\rtd_1, \rtd_2 \in \uni$ and $\rtd_1$ is
1070: $\rln$-related to $\rtd_2$, is there then a unimodular $\rln$-coupling of
1071: $\rtd_1$ to $\rtd_2$?
1072: \endprocl
1073: 
1074: The case where $\rtd_i$ are amenable is established affirmatively in \ref
1075: p.amen-average/.
1076: However, the case where $\rtd_i$ are supported by a fixed underlying
1077: non-amenable Cayley graph is open even when the marks take only two values.
1078: Here is a family of examples to illustrate what we do not know:
1079: 
1080: \procl q.end-couple
1081: Let $T$ be the Cayley graph of $\Z_2 * \Z_2 * \Z_2$ with respect to the
1082: generators $a$, $b$, $c$, which are all involutions.
1083: We label the edges with the generators.
1084: Fix three Borel symmetric
1085: functions $f_a$, $f_b$, $f_c$ from $[0, 1]^2$ to $[0, 1]$.
1086: Also, fix an end $\xi$ of $T$.
1087: Let $U(e)$ be i.i.d.\ Uniform$[0, 1]$ random variables indexed by the edges
1088: $e$ of $T$.
1089: For each edge $e$, let $I_e$ be the two edges adjacent to $e$ that lead
1090: farther from $\xi$ and let $J_e$ be the two other edges that are adjacent to
1091: $e$.
1092: Let $L(e)$ denote the Cayley label of $e$, i.e., $a$, $b$, or $c$.
1093: For an edge $e$ and a pair of edges $\{e_1, e_2\}$, write $f\big(e, \{e_1,
1094: e_2\}\big) := f_{L(e)}\big(U(e_1), U(e_2)\big)$.
1095: Define $X(e) := f(e, I_e)$ and $Y(e) := \max \big\{f(e, I_e), f(e, J_e)\big\}$.
1096: Let $\nu$ be the law of $(X, Y)$.
1097: Let $\rtd_1$ be the law of $X$ and $\rtd_2$ be the law of $Y$.
1098: We use the same notation for the measures in $\uni$ given by rooting $T$ at
1099: the vertex corresponding to the identity of the group.
1100: Let $\rln$ be $\le$ on $[0, 1] \times [0, 1]$.
1101: Since $X(e) \le Y(e)$ for all $e$, $\nu$ is an $\rln$-coupling of $\rtd_1$ to
1102: $\rtd_2$.
1103: In addition, $\rtd_2$ is clearly $\Aut(T)$-invariant (recall that the edges
1104: are labeled), while
1105: the same holds for $\rtd_1$ since it is an i.i.d.\ measure.
1106: Thus, $\rtd_i$ are both unimodular for $i = 1, 2$.
1107: On the other hand, $\nu$ is not $\Aut(T)$-invariant.
1108: Is there an invariant $\rln$-coupling of $\rtd_1$ to $\rtd_2$?
1109: In other words, is there a unimodular $\rln$-coupling of $\rtd_1$
1110: to $\rtd_2$?
1111: %Even much more elementary forms of this question have been open for several
1112: %years.
1113: \endprocl
1114: 
1115: Another example concerns monotone coupling of the wired and free uniform
1116: spanning forests (whose definitions are given below in \ref s.span/).
1117: This question was raised in \BLPSusf; a partial answer was given by \ref
1118: b.Bowen:coupling/.
1119: This is not the only interesting situation involving graph inclusion.
1120: To be more precise about this relation,
1121: for a map $\psi : \marks \to \marks$ and a network $G$, let $\psi(G)$
1122: denote the network obtained from $G$ by replacing each mark with its image
1123: under $\psi$.
1124: Given a Borel subset $\marks_0 \subseteq \marks$ and a network $\gh$,
1125: call the subnetwork consisting of those edges both of whose edge marks lie in
1126: $\marks_0$ the {\bf $\marks_0$-open subnetwork} of $\gh$.
1127: If $\rtd$ and $\rtd'$ are two probability measures on rooted networks, let
1128: us say that $\rtd$ is {\bf edge dominated} by $\rtd'$ if there exists a
1129: measure $\nu$ on $\GG_*$, a Borel subset $\marks_0 \subseteq \marks$, and
1130: Borel functions $\psi, \psi': \marks \to \marks$ such that if $(\gh', \bp)$
1131: denotes a network with law $\nu$ and $(\gh, \bp)$ the component of $\bp$ in
1132: the $\marks_0$-open subnetwork,
1133: then $\big(\psi(\gh), \bp\big)$ has law $\rtd$ and
1134: $\big(\psi'(\gh'), \bp\big)$ has law $\rtd'$.
1135: If the measure $\nu$ can be chosen to be unimodular, then we say that
1136: $\rtd$ is {\bf unimodularly edge dominated} by $\rtd'$.
1137: As a special case of \ref q.coupling/, we do not know whether the existence
1138: of such a measure $\nu$ that is not unimodular implies the existence of
1139: $\nu$ that is unimodular when $\rtd$ and $\rtd'$ are both unimodular
1140: themselves.
1141: 
1142: 
1143: \bsection{Fixed Underlying Graphs}{s.related}
1144: 
1145: Before we study general unimodular probability measures, it is useful to
1146: examine the relationship between unimodularity in the classical sense for
1147: graphs and unimodularity in the sense investigated here for random rooted
1148: network classes.
1149: 
1150: Given a graph $\gh$ and a vertex $x \in \verts(\gh)$, write $\Stab(x) := \{
1151: \gpe \in \Aut(\gh) \st \gpe x = x\}$ for
1152: the {\bf stabilizer subgroup} of $x$.
1153: Also, write $[x] := \Aut(\gh) x$ for the orbit of $x$.
1154: Recall the following principle from \BLPSgip:
1155: 
1156: \proclaim Mass-Transport Principle. If $\gh= (\vertex, \edge)$ is any graph,
1157: %$\gp \subseteq \Aut(\gh)$ is closed,
1158: $f : \vertex \times \vertex \to [0,
1159: \infty]$ is invariant under the diagonal action of $\Aut(\gh)$, and $o, o' \in
1160: \vertex$, then
1161: $$
1162: \sum_{z \in [o']} f(o, z) |\Stab(o')| = \sum_{y \in [o]} f(y, o') |\Stab(y)| \,.
1163: $$
1164: 
1165: Here, $|\cbuldot|$ denotes Haar measure on $\Aut(\gh)$, although we
1166: continue to use this notation for cardinality as well.
1167: Since $\Stab(x)$ is compact and open, $0 < |\Stab(x)| < \infty$.
1168: As shown in \ref b.Schlichting/ and \ref b.Trofimov/,
1169: $$
1170: |\Stab(x)y| / |\Stab(y)x| = |\Stab(x)| / |\Stab(y)|\,;
1171: \label e.schl-trof
1172: \,.
1173: $$
1174: It follows easily that $\gh$ is unimodular iff
1175: $$
1176: |\Stab(x)y| = |\Stab(y)x|
1177: \label e.unimod-schl-trof
1178: $$
1179: whenever $x$ and $y$ are in the same orbit.
1180: 
1181: \procl t.fixedgraph \procname{Unimodular Fixed Graphs}
1182: Let $G$ be a fixed connected graph. Then
1183: $G$ has a random root that gives a unimodular
1184: measure iff $G$ is a unimodular graph with
1185: $$
1186: c := \sum_i |\Stab(o_i)|^{-1} <\infty
1187: \,,
1188: \label e.alm-trn
1189: $$
1190: where $\{ o_i\} $ is a complete orbit
1191: section.
1192: In this case, there is only one such measure $\rtd$ on random rooted graphs
1193: from $G$ and it satisfies
1194: $$
1195: \rtd([G, x]) = c^{-1} |\Stab(x)|^{-1}
1196: \label e.mustab
1197: $$
1198: for every $x \in \verts(G)$.
1199: \endprocl
1200: 
1201: Of course, a similar statement holds for fixed networks.
1202: An example of a graph satisfying \ref e.alm-trn/, but that is not
1203: quasi-transitive, is obtained from the random weak limit of balls in a
1204: 3-regular tree.
1205: That is, let $\vertex := \N \times \N$.
1206: Join $(m, n)$ by edges to each of $(2m, n-1)$ and $(2m+1, n-1)$ for $n \ge
1207: 1$.
1208: The result is a tree with only one end and $\big|\Stab\big((m, n)\big)\big|
1209: = 2^n$.
1210: 
1211: 
1212: \proof
1213: Suppose first that $G$ is unimodular and that $c <\infty$.
1214: Define $\rtd$ by
1215: $$
1216: \all i \rtd([G, o_i]) := c^{-1} |\Stab(o_i)|^{-1}
1217: \,.
1218: $$
1219: To show that $\rtd$ is unimodular, let
1220: $f : \gtwo \to [0, \infty]$ be Borel.
1221: Since we are concerned only with the graph $\gh$, we shall write $f$
1222: instead as a function
1223: $f : \vertex \times \vertex \to [0,
1224: \infty]$ that is $\Aut(\gh)$-invariant.
1225: Then
1226: $$\eqaln{
1227: \int \sum_x f(o, x) \,d\rtd(G, o)
1228: &=
1229: c^{-1} \sum_i \sum_x f(o_i, x) |\Stab(o_i)|^{-1}
1230: \cr&=
1231: c^{-1} \sum_i |\Stab(o_i)|^{-1}
1232: \sum_j |\Stab(o_j)|^{-1} \sum_{x \in [o_j]} f(o_i, x) |\Stab(o_j)|
1233: \cr&=
1234: c^{-1} \sum_i |\Stab(o_i)|^{-1}
1235: \sum_j |\Stab(o_j)|^{-1} \sum_{y \in [o_i]} f(y, o_j) |\Stab(y)|
1236: \cr&\hskip1in\hbox{[by the Mass-Transport Principle for $\gh$]}
1237: \cr&=
1238: c^{-1} \sum_i |\Stab(o_i)|^{-1}
1239: \sum_j |\Stab(o_j)|^{-1} \sum_{y \in [o_i]} f(y, o_j) |\Stab(o_i)|
1240: \cr&\hskip1in\hbox{[by unimodularity of $\gh$]}
1241: \cr&=
1242: c^{-1} \sum_j \sum_y f(y, o_j) |\Stab(o_j)|^{-1}
1243: \cr&=
1244: \int \sum_y f(y, o) \,d\rtd(G, o)
1245: \,.
1246: }$$
1247: Since $\rtd$ satisfies the Mass-Transport Principle, it is unimodular.
1248: 
1249: Conversely, suppose that $\rtd$ is a unimodular probability measure on rooted
1250: versions of $\gh$. To see that $\gh$ is unimodular, consider any $u, v$ in the
1251: same orbit.
1252: Define
1253: $$
1254: \rtd\big([x]\big) :=
1255: \rtd\big([\gh, x]\big)
1256: \,.
1257: $$
1258: We first show that $\rtd\big([u]\big) > 0$.
1259: Every graph isomorphic to $\gh$ has a well-defined notion of vertices of
1260: type $[u]$.
1261: Let each vertex $x$ send mass 1 to each vertex of type $[u]$ that is
1262: nearest to $x$.
1263: This is a Borel function on $\gtwo$ if we transport no mass on graphs
1264: that are not isomorphic to $\gh$.
1265: The expected mass sent is positive, whence so is the expected mass
1266: received.
1267: Since only vertices of type $[u]$ receive mass, it follows that
1268: $\rtd\big([u]\big) > 0$, as desired.
1269: 
1270: Let $f(x, y) := \I{\gp_{u, x} v}(y)$, where $\gp_{u, x} := \{ \gpe
1271: \in\Aut(\gh) \st \gpe u = x \}$.
1272: Note that $y \in \gp_{u, x} v$ iff $x \in \gp_{v, y} u$.
1273: It is straightforward to check that $f$ is diagonally invariant under
1274: $\Aut(\gh)$.
1275: Note that
1276: $$
1277: |\Stab(x) y| \I{[x]}(o)
1278: =
1279: |\gp_{x, \bp} y|
1280: $$
1281: for all $x, y, \bp \in \verts(\gh)$.
1282: Therefore, we have
1283: $$\eqaln{
1284: |\Stab(u) v| \rtd([u])
1285: &=
1286: \int |\gp_{u, o} v| \,d\rtd(\gh, o)
1287: =
1288: \int \sum_x \I{\gp_{u, o} v}(x) \,d\rtd(\gh, o)
1289: \cr&=
1290: \int \sum_x f(o, x) \,d\rtd(\gh, o)
1291: =
1292: \int \sum_x f(x, o) \,d\rtd(\gh, o)
1293: \cr&\hskip1in\hbox{[by the Mass-Transport Principle for $\rtd$]}
1294: \cr&=
1295: \int \sum_x \I{\gp_{u, x} v}(o) \,d\rtd(\gh, o)
1296: =
1297: \int \sum_x \I{\gp_{v, o} u}(x) \,d\rtd(\gh, o)
1298: \cr&=
1299: \int |\gp_{v, o} u| \,d\rtd(\gh, o)
1300: =
1301: |\Stab(v) u| \rtd([v])
1302: \,.
1303: }$$
1304: That is,
1305: $$
1306: |\Stab(u) v| \rtd([u])
1307: =
1308: |\Stab(v) u| \rtd([v])
1309: \,.
1310: \label e.rtdstab
1311: $$
1312: Since $u$ and $v$ are in the same orbit, we have $[u] = [v]$, so $\rtd([u]) =
1313: \rtd([v])$. Since $\rtd([u]) > 0$, we obtain \ref e.unimod-schl-trof/.
1314: That is, $\gh$ is unimodular.
1315: Comparison of \ref e.rtdstab/ with \ref e.schl-trof/ shows \ref e.mustab/.
1316: \Qed
1317: 
1318: Automorphism invariance for random unrooted networks on fixed underlying
1319: graphs is also closely tied to unimodularity of random rooted networks.
1320: Here, we shall need to distinguish between graphs, networks, and
1321: isomorphism classes of rooted networks.
1322: Recall that
1323: $\overline \gh$ denotes a network whose
1324: underlying graph is $\gh$ and $[\overline \gh, \bp]$ denotes an
1325: equivalence class of networks $\overline \gh$ on $\gh$ with root $\bp$.
1326: %Note that $[\overline \gh, \bp]$ carries a natural probability measure
1327: %$\lift_{[\overline \gh, \bp]}$
1328: %induced by the Haar probability measure on $\Stab(\bp)$.
1329: 
1330: Let $\gh$ be a fixed connected unimodular graph satisfying \ref e.alm-trn/.
1331: Fix a complete orbit section $\{\bp_i\}$ of $\vertex(\gh)$.
1332: For a graph $\gh'$ and $x \in \vertex(\gh')$, $z \in
1333: \vertex(\gh)$, let $\Phi(x, z)$ be the set of rooted isomorphisms, if any,
1334: from $(\gh', x)$ to $(\gh, z)$.
1335: When non-empty, this set carries a natural probability measure,
1336: $\lift'_{(\gh', x; z)}$ arising from the Haar probability measure on
1337: $\Stab(z)$.
1338: When $\Phi(x, z) = \emptyset$, let $\lift'_{(\gh', x, z)} := 0$.
1339: Define
1340: $$
1341: \lift_{(\gh', x)} := \sum_{z \in \vertex(\gh)} \lift'_{(\gh', x; z)}
1342: \,.
1343: $$
1344: This is the analogue for isomorphisms from $\gh'$ to $\gh$
1345: of Haar measure on $\Aut(\gh)$.
1346: In particular, any $\gpe \in \Aut(\gh)$ pushes forward $\lift'_{(\gh', x,
1347: z)}$ to $\lift'_{(\gh', x, \gpe z)}$.
1348: 
1349: For a graph $\gh'$ isomorphic to $\gh$ and $x \in \vertex(\gh')$, let
1350: $\type(\gh', x) := \bp_i$ for the unique $\bp_i$ for which
1351: $\Phi(x, \bp_i) \ne \emptyset$.
1352: Note that $\lift_{(\gh', x)} = \lift_{(\gh', y)}$ when $\type(\gh', x) =
1353: \type(\gh', y)$.
1354: 
1355: Every probability measure $\rtd$ on $\GG_*$ that is
1356: concentrated on network classes whose underlying graph is $\gh$ induces a
1357: probability measure $\lift_\rtd$ on unrooted networks on $\gh$:
1358: $$
1359: \lift_\rtd(\A)
1360: :=
1361: \int \int_{\textstyle \Phi\big(\bp, \type(\gh', \bp)\big)}
1362: \I{\A}(\phi \overline \gh')  \,d\lift_{(\gh', \bp)}(\phi)
1363: \,d\rtd([\overline \gh', \bp])
1364: $$
1365: for Borel sets $\A$ of networks on $\gh$.
1366: It is easy to see that this is well defined (the choice of $(\overline
1367: \gh', \bp)$ in its equivalence class not mattering).
1368: 
1369: \procl t.uni-vs-invar \procname{Invariance and Unimodularity}
1370: Let $\gh$ be a fixed connected unimodular graph satisfying \ref e.alm-trn/.
1371: Let $\nu$ be an $\Aut(\gh)$-invariant probability measure on unrooted
1372: networks whose underlying graph is $\gh$.
1373: Then randomly rooting the network as in \ref e.mustab/ gives a measure $\rtd
1374: \in \uni$.
1375: Conversely, let $\rtd \in \uni$ be supported on networks whose underlying
1376: graph is $\gh$.
1377: %Fix $z \in \vertex(\gh)$.
1378: %Let $\rtd_z$ be $\rtd$ conditioned on the root being in the orbit of $z$.
1379: %Define $\nu$ to be the measure induced by $\rtd_z$
1380: %on networks on $\gh$ by identifying $z$
1381: %with $\bp$ under a random element of $\Aut(\gh)$ that takes $z$ to the root.
1382: %\msnote{Phrase better. The point is that $\rtd$ is defined on
1383: %rooted-isomorphism classes.} Then $\nu$ is $\Aut(\gh)$-invariant.
1384: Then $\lift_\rtd$ is $\Aut(\gh)$-invariant.
1385: \endprocl
1386: 
1387: \proof
1388: The first part of the theorem is proved just as is the first part of \ref
1389: t.fixedgraph/,
1390: so we turn to the second part.
1391: Let $\gpe_0 \in \Aut(\gh)$
1392: and $F$ be a bounded Borel-measurable function of networks on $\gh$.
1393: Invariance of $\lift_\rtd$ means that $\int F(\overline
1394: \gh)\,d\lift_\rtd(\overline \gh) = \int F(\gpe_0 \overline
1395: \gh)\,d\lift_\rtd(\overline \gh)$.
1396: To prove that this holds, let
1397: $$
1398: f(\overline \gh', x, y) :=
1399: \int_{\textstyle \Phi\big(x, \type(\gh', y)\big)
1400: \atop \textstyle \cap \Phi\big(y, \gpe_0 \type(\gh, y)\big)}
1401: F(\phi \overline \gh')\,d\lift_{(\gh', y)}(\phi)
1402: \,.
1403: $$
1404: It is straightforward to check that
1405: $f$ is well defined and Borel on $\gtwo$.
1406: %if we define $f$ to be 0
1407: %for networks that are not defined on graphs isomorphic to $\gh$.
1408: Therefore, unimodularity of $\rtd$ gives
1409: $$\eqaln{
1410: \int F(\gpe_0 \overline \gh)\,d\lift_\rtd(\overline\gh)
1411: &=
1412: \int \int_{\textstyle \Phi\big(\bp, \type(\gh', \bp)\big)}
1413: F(\gpe_0 \phi \overline \gh') \,d\lift_{(\gh', \bp)}(\phi)
1414: \,d\rtd([\overline \gh', \bp])
1415: \cr&=
1416: \int \int_{\textstyle \Phi\big(\bp, \gpe_0 \type(\gh', \bp)\big)}
1417: F(\phi \overline \gh') \,d\lift_{(\gh', \bp)}(\phi)
1418: \,d\rtd([\overline \gh', \bp])
1419: \cr&=
1420: \int \sum_{x \in \vertex(\gh')}
1421: \int_{\textstyle \Phi\big(x, \type(\gh', \bp)\big)
1422: \atop \textstyle \cap \Phi\big(\bp, \gpe_0 \type(\gh', \bp)\big)}
1423: F(\phi \overline \gh') \,d\lift_{(\gh', \bp)}(\phi)
1424: \,d\rtd([\overline \gh', \bp])
1425: \cr&=
1426: \int \sum_{x \in \vertex(\gh')}
1427: f(\overline \gh', x, \bp)
1428: \,d\rtd([\overline \gh', \bp])
1429: \cr&=
1430: \int \sum_{x \in \vertex(\gh')}
1431: f(\overline \gh', \bp, x)
1432: \,d\rtd([\overline \gh', \bp])
1433: \cr&=
1434: \int \sum_{x \in \vertex(\gh')}
1435: \int_{\textstyle \Phi\big(\bp, \type(\gh', x)\big)
1436: \atop \textstyle \cap \Phi\big(x, \gpe_0 \type(\gh', x)\big)}
1437: F(\phi \overline \gh') \,d\lift_{(\gh', x)}(\phi)
1438: \,d\rtd([\overline \gh', \bp])
1439: \cr&=
1440: \int \sum_{x \st \type(\gh', x) = \type(\gh', \bp)}
1441: \int_{\textstyle \Phi\big(\bp, \type(\gh', x)\big)
1442: \atop \textstyle \cap \Phi\big(x, \gpe_0 \type(\gh', x)\big)}
1443: F(\phi \overline \gh') \,d\lift_{(\gh', x)}(\phi)
1444: \,d\rtd([\overline \gh', \bp])
1445: \cr&=
1446: \int \sum_{x \st \type(\gh', x) = \type(\gh', \bp)}
1447: \int_{\textstyle \Phi\big(\bp, \type(\gh', \bp)\big)
1448: \atop \textstyle \cap \Phi\big(x, \gpe_0 \type(\gh', \bp)\big)}
1449: F(\phi \overline \gh') \,d\lift_{(\gh', \bp)}(\phi)
1450: \,d\rtd([\overline \gh', \bp])
1451: \cr&=
1452: \int \sum_{x \in \vertex(\gh')}
1453: \int_{\textstyle \Phi\big(\bp, \type(\gh', \bp)\big)
1454: \atop \textstyle \cap \Phi\big(x, \gpe_0 \type(\gh', \bp)\big)}
1455: F(\phi \overline \gh') \,d\lift_{(\gh', \bp)}(\phi)
1456: \,d\rtd([\overline \gh', \bp])
1457: \cr&=
1458: \int
1459: \int_{\textstyle \Phi\big(\bp, \type(\gh', \bp)\big)}
1460: F(\phi \overline \gh') \,d\lift_{(\gh', \bp)}(\phi)
1461: \,d\rtd([\overline \gh', \bp])
1462: \cr&=
1463: \int F(\overline \gh)\,d\lift_\rtd(\gh)
1464: \,.
1465: \Qed
1466: }$$
1467: 
1468: 
1469: \procl r.anygroup
1470: As this section shows, unimodular quasi-transitive graphs are special cases
1471: of unimodular rooted networks.
1472: However, sometimes one is interested in random networks on a graph $\gh$
1473: that are not necessarily invariant under the full group $\Aut(\gh)$, but
1474: only under some subgroup, $\gp \subset \Aut(\gh)$.
1475: This is common when $\gh$ is a Cayley graph of $\gp$.
1476: In this case, we could mark the edges by the generators they represent;
1477: that is, if $x, y \in \gp$ and $y = x a$ with $a$ one of the generators
1478: used to form $\gh$, then we can mark the edge $[x, y]$ at $x$ by $a$.
1479: This makes the full automorphism group of the network
1480: $\overline \gh$ equal to $\gp$, rather than to $\Aut(\gh)$.
1481: The theory here then goes through with only a complication of notation.
1482: However, given any graph $\gh$ and any closed subgroup $\gp \subset
1483: \Aut(\gh)$ that acts quasi-transitively on $\gh$, we do not know whether it
1484: is possible to mark the edges and vertices of $\gh$ to get a network whose
1485: automorphism group is equal to $\gp$.
1486: Yet, the theory for quasi-transitive subgroups is the same; see \BLPSgip.
1487: \endprocl
1488: 
1489: 
1490: 
1491: 
1492: 
1493: 
1494: \bsection{Random Walks and Extremality}{s.extreme}
1495: 
1496: Random walks on networks,
1497: besides being of intrinsic interest,
1498: form an important tool for studying networks.
1499: A random walk is most useful when it has a stationary measure, in other
1500: words, when the distribution of $(\gh, w_0)$ is the same as the
1501: distribution of $(\gh, w_1)$, where $w_0$ is the initial location of the
1502: random walk and $w_1$ is the next location of the random walk.
1503: 
1504: Consider simple random walk on a random graph chosen by a unimodular
1505: probability measure $\rtd$ on rooted graphs, where we start the random walk
1506: at the root.
1507: Just as for finite graphs, we do not expect $\rtd$ to be stationary for the
1508: random walk; rather, we get a stationary measure by biasing $\rtd$ by the
1509: degree of the root.
1510: The fact that this measure is stationary follows from the definition of
1511: involution invariance;
1512: in fact, the definition is precisely that simple random walk is
1513: reversible, i.e., that the distribution of $\big((\gh, w_0), (\gh,
1514: w_1)\big)$ is the same as the distribution of $\big((\gh, w_1), (\gh,
1515: w_0)\big)$, where $(\gh, w_0)$ has distribution $\rtd$ biased by the degree
1516: of the root and $w_1$ is a uniform random neighbor of the 
1517: root.\ftnote{*}{Note that the degree times counting measure is
1518: reversible on every graph, regardless of
1519: unimodularity of the measure on rooted graphs.}
1520: If $\expdeg(\rtd) < \infty$, then we can normalize the biased measure
1521: to obtain a probability measure.
1522: 
1523: In particular, recall from
1524: \ref x.AGW/ the definition of the augmented Galton-Watson measure \AGW.
1525: In \ref b.LPP:GW/, it was remarked in reference to the stationarity of $\AGW$
1526: for simple random walk that ``unlike the situation for finite
1527: graphs, there is no biasing in favor of vertices of large degree".
1528: However, we now see that contrary to this remark,
1529: the situations of finite graphs and $\AGW$ are, in fact, parallel.
1530: That is because the biasing by the degree has already been made part of the
1531: probability measure $\AGW$. The correct comparison of the uniform measure
1532: on vertices of finite graphs is to the
1533: unimodular Galton-Watson probability measure on trees, $\UGW$, because it
1534: is for this measure that ``all vertices are equally likely to be the root".
1535: 
1536: More generally, we can consider
1537: stationarity of random walk in a random environment with random scenery.
1538: Here, if the graph underlies a network, the marks are not restricted to play a
1539: passive role, but may, in fact, determine the transition probabilities (as
1540: in \ref s.trace/) and provide a scenery for the random walk.
1541: %That is, let $\mkmp(x)$ be the vertex mark at $x$.
1542: That is, a Borel function $p : \gtwo \to [0, 1]$, written as
1543: $p : (\gh, x, y) \mapsto p_\gh(x, y)$, such that $\sum_{y \in \verts}
1544: p_\gh(x, y) = 1$ for all vertices $x$
1545: %and such that
1546: %$$
1547: %p_{(\phi(\gh), \phi(\bp))}(\phi(x), \phi(y))
1548: %= p_{(\gh, \bp)}(x, y)
1549: %$$
1550: %for every non-rooted isomorphism $\phi$
1551: is called an {\bf environment}.
1552: A Borel map $\nu : \GG_* \to (0, \infty)$, written
1553: $\nu : (\gh, x) \mapsto \nu_{\gh}(x)$,
1554: %such that
1555: %$$
1556: %\nu_{(\phi(\gh), \phi(\bp))}(\phi(x)) = \nu_{(\gh, \bp)}(x)
1557: %$$
1558: %for every non-rooted isomorphism $\phi$
1559: is called an {\bf initial bias}.
1560: It is called {\bf $p$-stationary} if for all $\gh$, the measure
1561: $\nu_\gh$ is stationary for
1562: the random walk on $\gh$ given by the environment $p_\gh$.
1563: Write $\paths$ for the set of (equivalence classes of)
1564: pairs $\big((\gh, w_0), \Seq{w_n \st n \ge
1565: 0}\big)$ with $(\gh, w_0) \in \GG_*$ and $w_n \in \vertex(\gh)$.
1566: Let $\Ph$ denote the distribution on $\paths$ of
1567: the trajectory of the Markov chain
1568: determined by the environment starting at $\bp$ with initial distribution
1569: equal to $\rtd$ biased by $\nu_{\gh}(\bp)$.
1570: That is, if $\traj_{(\gh, \bp)}$ denotes the probability measure on $\paths$
1571: determined by the environment on $\gh$ with initial vertex
1572: $w_0 = \bp$, then for all events $\ev B$, we have
1573: $$
1574: \Ph(\ev B)
1575: :=
1576: \int_{\GG_*} \traj_{(\gh, \bp)}(\ev B) \nu_{\gh}(\bp)
1577: \,d\rtd(\gh, \bp)
1578: \,.
1579: $$
1580: Let $\invar$ denote the $\sigma$-field of events (in the Borel
1581: $\sigma$-field of $\GG_*$) that are invariant under non-rooted
1582: isomorphisms.
1583: To avoid possible later confusion,
1584: note that this does not depend on the measure $\rtd$, so that even if there
1585: are no non-trivial non-rooted isomorphisms $\rtd$-a.s., the $\sigma$-field
1586: $\invar$ is still not equal (mod 0) to the $\sigma$-field of
1587: $\rtd$-measurable sets.
1588: It is easy to see that
1589: for any $\rtd \in \uni$ and $\ev A \in \invar$ with $\rtd(\ev A) > 0$, the
1590: probability measure $\rtd(\;\cbuldot \mid \ev A)$ is also unimodular.
1591: %Let $\invarp$ be the collection of Borel subsets of $\paths$ that are
1592: %invariant under non-rooted isomorphisms of the underlying network with the
1593: %diagonal action, i.e., under maps $\big((\gh, \bp), \Seq{w_n}\big) \mapsto
1594: %\big((\phi(\gh), \phi(\bp)), \Seq{\phi(w_n)}\big)$ for isomorphisms $\phi:\gh
1595: %\to\gh$.
1596: Define the {\bf shift} $\SS : \paths \to \paths$ by
1597: $$
1598: \SS \big((\gh, w_0), \Seq{w_n}\big) := \big((\gh, w_1), \Seq{w_{n+1}}\big)
1599: \,.
1600: $$
1601: %and $\SS \big((\gh, \bp), \Seq{w_n}\big) := \big((\gh, \bp),
1602: %\Seq{w_n}\big)$ when $\bp \ne w_0$.
1603: The following extends Theorem 3.1 of \ref b.LS:rwrers/;
1604: the proof is essentially the same.
1605: 
1606: %\comment{Here and later, we require local finiteness if this turns out not to
1607: %be in our definition of $\GG_*$.}
1608: 
1609: \procl t.rwrers \procname{Random Walk in a Random Environment and Random
1610: Scenery}
1611: Let $\rtd$ be a unimodular probability measure on $\GG_*$.
1612: Let $p_{\cbuldot}(\cbuldot)$ be an environment and
1613: $\nu_{\cbuldot}(\cbuldot)$ be an initial bias that is $p$-stationary.
1614: Let $\Ph$ be the corresponding measure on trajectories.
1615: Then $\Ph$ is stationary for the shift.
1616: If $p$ is also reversible with respect to $\nu_{\cbuldot}(\cbuldot)$, then
1617: $\Ph$ is reversible, in other words, for all events $\ev A$, $\ev B$, we
1618: have 
1619: $$
1620: \Ph[(\gh, w_0) \in \ev A, (\gh, w_1) \in \ev B]
1621: =
1622: \Ph[(\gh, w_1) \in \ev A, (\gh, w_0) \in \ev B]
1623: \,.
1624: $$
1625: If
1626: $$
1627: \int \nu_{\gh}(\bp) \,d\rtd(\gh, \bp) = 1
1628: \,,
1629: \label e.proba
1630: $$
1631: then $\Ph$ is a probability measure.
1632: \endprocl
1633: 
1634: \proof
1635: The reversibility was not mentioned in prior work, so we give that proof
1636: here.
1637: Assuming that $p$ is $\nu$-reversible, we have 
1638: $$\eqaln{
1639: \Ph[(\gh, w_0) \in \ev A, (\gh, w_1) \in \ev B]
1640: &=
1641: \Ebig{\sum_{x \in \vertex(\gh)} \I{\ev A}(\gh, \bp) \nu_\gh(\bp) p_\gh(\bp,
1642: x) \I{\ev B}(\gh, x)}
1643: \cr&=
1644: \Ebig{\sum_{x \in \vertex(\gh)} \I{\ev A}(\gh, \bp) \nu_\gh(x) p_\gh(x,
1645: \bp) \I{\ev B}(\gh, x)}
1646: \,.
1647: \cr
1648: }$$
1649: The Mass-Transport Principle now gives that this 
1650: $$
1651: =
1652: \Ebig{\sum_{x \in \vertex(\gh)} \I{\ev A}(\gh, x) \nu_\gh(\bp) p_\gh(\bp,
1653: x) \I{\ev B}(\gh, \bp)}
1654: =
1655: \Ph[(\gh, w_1) \in \ev A, (\gh, w_0) \in \ev B]
1656: \,. \Qed
1657: $$
1658: 
1659: \procl r.choice
1660: This theorem is made more useful by noticing that for any $\rtd \in \uni$,
1661: there is a choice of $p_{\cbuldot}(\cbuldot)$ and
1662: $\nu_\cbuldot(\cbuldot)$ that satisfies all the hypotheses, including
1663: \ref e.proba/.
1664: For example, if $F_\gh(x)$
1665: denotes $\sum_{y \sim x} 1/\deg_\gh(y)$,
1666: then let $p_\gh(x, y) := 1/[F_\gh(x) \deg_\gh(y)]$ and
1667: $\nu_\gh(x) := Z^{-1} F_\gh(x)/\deg_\gh(x)$,
1668: where 
1669: $$
1670: Z := \int F_\gh(\bp)/\deg_\gh(\bp) \,d\rtd(\gh, \bp)
1671: \,.
1672: $$
1673: It is clear that $p$ is an environment.
1674: Since $F_\gh(\bp) \le \sum_{y \sim x} 1 = \deg_\gh(\bp)$, we also have that
1675: $Z < \infty$, so that $\nu$ is a $p$-stationary initial bias and $p$ is
1676: $\nu$-reversible.
1677: \endprocl
1678: 
1679: Given a network with positive edge
1680: weights and a time $t > 0$, form the {\bf transition operator} $P_t$ for
1681: continuous-time random walk whose rates are the edge weights; in the case
1682: of unbounded weights (or degrees), we take the minimal process, which dies
1683: after an explosion.
1684: That is, if the entries of a matrix $A$ indexed by the vertices are equal
1685: off the diagonal to the negative of the edge weights and the diagonal
1686: entries are chosen to make the row sums zero, then $P_t := e^{-A t}$; in
1687: the case of unbounded weights, we take the self-adjoint extension of $A$
1688: corresponding to the minimal process.
1689: The matrix $A$ is called the {\bf infinitesimal generator} or the {\bf
1690: Laplacian} of the network.
1691: %the infinitesimal generator is an extension of the Laplacian and it is true
1692: %that the transition probabilities are given by exponentiation (Theorem
1693: %13.37 of Rudin's F.A., 1st ed.). This should be in some probability text,
1694: %too.
1695: 
1696: 
1697: \procl c.continuous-time-stationary
1698: Suppose that $\rtd \in \uni$ is carried by networks with non-negative edge
1699: weights such that the corresponding continuous-time Markov chain has no
1700: explosions a.s.
1701: Then $\rtd$ is stationary and reversible.
1702: \endprocl
1703: 
1704: \proof
1705: Fix $t > 0$ and let $p_\gh(x, y) := P_t(x, y)$.
1706: It is well known that $p$ is reversible with respect to
1707: the uniform measure $\nu_\gh \equiv 1$.
1708: Thus, \ref t.rwrers/ applies.
1709: \Qed
1710: 
1711: We can also obtain a sufficient condition for lack of explosions:
1712: 
1713: \procl c.no-explode
1714: Suppose that $\rtd \in \uni$ is carried by networks with non-negative edge
1715: weights $c_\gh(e)$ such that $Z := \E[\sum_{x \sim \bp} c_\gh(\bp, x)] <
1716: \infty$.
1717: Then the corresponding continuous-time Markov chain has no explosions.
1718: \endprocl
1719: 
1720: \proof
1721: In this case, consider the discrete-time Markov chain corresponding to
1722: these weights.
1723: It has a stationary probability measure arising from the choice
1724: $\nu_\gh(x) := \sum_{y \sim \bp} c_\gh(x, y)/Z$.
1725: It is well known that explosions occur iff 
1726: $$
1727: \sum_{n \ge 0} \nu_\gh(w_n)^{-1} < \infty
1728: $$
1729: with positive probability.
1730: However, stationarity guarantees that this sum is infinite a.s.\
1731: (by the Poincar\'e recurrence theorem).
1732: \Qed
1733: 
1734: \procl r.poss-explode
1735: It is possible for explosions to occur:
1736: For example, consider the uniform spanning tree $T$ in $\Z^2$ (see
1737: \BLPSusf). The only fact we use about $T$ is that it has one end a.s.\
1738: and has an invariant distribution.
1739: Let $c_\gh(e) := 0$ for $e \notin T$ and
1740: $c_\gh(e) := 2^{f(e)}$ when $e \in T$
1741: and $f(e)$ is the number of vertices in the finite component of $T
1742: \setminus e$.
1743: Then it is easy to verify that the corresponding continuous-time Markov
1744: chain explodes a.s.
1745: 
1746: Furthermore, explosions may occur
1747: on a fixed transitive graph that is not unimodular, 
1748: even if the condition in \ref c.no-explode/ is satisfied.
1749: To see this, let
1750: $\xi$ be a fixed end of a regular tree $T$ of degree 3. 
1751: Thus, for every vertex $x$ in $T$, there is a unique ray $x_\xi := 
1752: \Seq{x_0=x, x_1, x_2, \ldots}$ starting at $x$ such that $x_\xi$ and
1753: $y_\xi$ differ by only finitely many vertices for any pair $x$, $y$.
1754: Call $x_1$ the {\bf $\xi$-parent} of $x$, call $x$ a {\bf $\xi$-child} of
1755: $x_1$, and call
1756: $x_2$ the {\bf $\xi$-grandparent} of $x$. Let $\gh$ be the graph obtained
1757: from $T$ by adding the edges $(x, x_2)$ between each $x$ and its
1758: $\xi$-grandparent.
1759: Then $\gh$ is a transitive graph, first mentioned by \ref b.Trofimov/.
1760: In fact, every automorphism of $\gh$ fixes $\xi$.
1761: Now consider the following random weights on $\gh$.
1762: Put weight 0 on every edge in $\gh$ that is not in $T$.
1763: For each vertex of $\gh$, declare open the edge to precisely
1764: one of its two $\xi$-children, chosen uniformly and
1765: independently for different vertices.
1766: The open components are rays.
1767: Let the weight of every edge that is not open also be 0.
1768: If an edge $(x, y)$ between a vertex $x$ and its $\xi$-parent $y$ is open and
1769: $y$ is at distance $n$ from the beginning of the open ray containing $(x, y)$,
1770: then let the weight of the edge be $(3/2)^n$.
1771: Since this event has probability $1/2^{n+1}$, the condition of \ref
1772: c.no-explode/ is clearly satisfied.
1773: It is also clear that the Markov chain explodes a.s.
1774: \endprocl
1775: 
1776: The class $\uni$ of unimodular probability measures on $\GG_*$ is clearly
1777: convex.
1778: An element of $\uni$ is called {\bf extremal} if it cannot be written as a
1779: convex combination of other elements of $\uni$.
1780: We shall show that the extremal measures are those for which $\invar$
1781: contains only sets of measure 0 or 1.
1782: Intuitively,
1783: they are the extremal measures for unrooted networks since
1784: the distribution of the root is forced given the distribution of the
1785: unrooted network.
1786: For example, one may show that $\UGW$ is extremal when conditioned on
1787: non-extinction.
1788: %To do this, use ergodicity of simple random walk, proved in \ref b.LPP/,
1789: %and then appeal to \ref t.erg/, in particular, \ref e.shiftev/, which
1790: %shows that if all such $\ev B$ are trivial, then so are all $\ev A$,
1791: %there, i.e., all of $\invar$.
1792: %Is there a more direct way?
1793: First, we show the following ergodicity property, analogous to Theorem 5.1
1794: of \ref b.LS:indist/.
1795: Recall that a $\sigma$-field is called {\bf $\rtd$-trivial} if all its
1796: elements have measure 0 or 1 with respect to $\rtd$.
1797: 
1798: \procl t.erg \procname{Ergodicity}
1799: Let $\rtd$ be a unimodular probability measure on $\GG_*$.
1800: Let $p_{\cbuldot}(\cbuldot)$ be an environment that satisfies
1801: $$
1802: \all{\gh} \all{x, y \in \vertex(\gh)} \quad x \sim y \Longrightarrow
1803: p_\gh(x, y) > 0
1804: \label e.getsall
1805: $$
1806: and
1807: $\nu_{\cbuldot}(\cbuldot)$ be an initial bias that is $p$-stationary and
1808: satisfies \ref e.proba/.
1809: Let $\Ph$ be the corresponding probability measure on trajectories.
1810: If $\invar$ is $\rtd$-trivial, then
1811: every event that is shift invariant is $\Ph$-trivial.
1812: More generally, the events $\ev B$ in the $\Ph$-completion of the
1813: shift-invariant $\sigma$-field are those of the form 
1814: $$
1815: \ev B
1816: =
1817: \big\{ \big((\gh, \bp), w\big) \in \paths \st (\gh, \bp) \in \ev A \big\}
1818: \xor \ev C
1819: \label e.shiftev
1820: $$
1821: for some $\ev A \in \invar$ and some event $\ev C$ with $\Ph(\ev C) = 0$.
1822: \endprocl
1823: 
1824: \proof
1825: Let $\ev B$ be a shift-invariant event.
1826: As in the proof of Theorem 5.1 of \ref b.LS:indist/, we have $\traj_{(\gh,
1827: \bp)}(\ev B) \in \{ 0, 1 \}$ $\rtd$-a.s.
1828: The set $\ev A$ of $(\gh, \bp)$ where this probability equals 1 is in
1829: $\invar$ by \ref e.getsall/, and a little thought reveals that
1830: \ref e.shiftev/ holds for some $\ev C$ with $\Ph(\ev C) = 0$.
1831: If $\invar$ is $\rtd$-trivial, then
1832: $\rtd(\ev A) \in \{0, 1\}$, whence $\Ph(\ev B) \in \{0, 1\}$ as desired.
1833: Conversely, every event $\ev B$ of the form \ref e.shiftev/ is clearly 
1834: in the $\Ph$-completion of the shift-invariant $\sigma$-field.
1835: \Qed
1836: 
1837: We may regard the space $\paths$ as the space of sequences of rooted
1838: networks, where all roots belong to the same network.
1839: Thus, $\paths$ is the natural trajectory space for the Markov chain with
1840: the transition probability from $(\gh, x)$ to $(\gh, y)$ given by $p_\gh(x,
1841: y)$.
1842: With this interpretation, \ref t.erg/ says that this Markov chain is
1843: ergodic when $\invar$ is $\rtd$-trivial.
1844: The next theorem says that this latter condition is, in turn, equivalent to
1845: extremality of $\rtd$.
1846: 
1847: \procl t.findroot \procname{Extremality}
1848: %Let $(\gh, \bp, \bp')$ be a random connected network with two distinguished
1849: %vertices such that the laws $\rtd$ of $(\gh, \bp)$ and $\rtd'$ of $(\gh,
1850: %\bp')$ are both unimodular.
1851: %Then $\rtd = \rtd'$.
1852: %Equivalently, a unimodular probability measure $\rtd$ on $\GG_*$ is
1853: A unimodular probability measure $\rtd$ on $\GG_*$ is
1854: extremal iff $\invar$ is $\rtd$-trivial.
1855: \endprocl
1856: 
1857: \comment{
1858: For the entire proof, choose an environment and stationary initial bias
1859: that satisfy \ref e.proba/ and \ref e.getsall/.
1860: Also, let $\kappa$ be the law of $(\gh, \bp, \bp')$ on $\gtwo$ that couples
1861: $\rtd$ and $\rtd'$.
1862: 
1863: Let $\A$ be an event of $\GG_*$.
1864: Let $\alpha$ be the function on $\paths$ that gives
1865: the asymptotic frequency of visits to $\A$:
1866: $$
1867: \alpha\big(((\gh, w_0), \Seq{w_n})\big) :=
1868: \liminf_{N \to\infty} {1 \over N} |\{n \le N \st (\gh, w_n) \in \A\}|
1869: \,.
1870: $$
1871: Then $\alpha$ is a shift-invariant function.
1872: %that is measurable with respect to $\invarp$.
1873: For any $t \in [0, 1]$, let $\ev B_t := \{(\gh, \bp) \st \traj_{(\gh,
1874: \bp)}[\alpha \le t] \in \{0, 1\}\}$.
1875: Then $\rtd(\ev B_t), \rtd'(\ev B_t) \in \{ 0, 1 \}$,
1876: as we saw in the proof of \ref t.erg/.
1877: Therefore, there exists a Borel function $f : \GG_* \to [0, 1]$ such
1878: that $\rtd\big[\traj_{(\gh, \bp)}[\alpha = f(\gh, \bp)] = 1 \big] = 1$.
1879: Since a random walk has positive probability of going from $\bp$ to $\bp'$,
1880: it follows that
1881: $$
1882: \kappa[\big[\traj_{(\gh, \bp)}[\alpha = f(\gh, \bp) = f(\gh, \bp')] = 1 \big]
1883: = 1
1884: \,.
1885: \label e.same
1886: $$
1887: Let $\nu \rtd$ stand for the measure $d(\nu \rtd)(\gh, \bp) = \nu_\gh(\bp)
1888: d\rtd(\gh, \bp)$, and likewise for $\nu\rtd'$.
1889: By \ref t.rwrers/, these are stationary measures for simple random walk,
1890: whence the ergodic theorem gives
1891: $$\eqaln{
1892: \int \I\A \nu_\gh(\bp) \,d\rtd(\gh, \bp)
1893: &=
1894: \int f(\gh, \bp) \nu_\gh(\bp) \,d\rtd(\gh, \bp)
1895: }$$
1896: }%
1897: 
1898: \proof
1899: Let $\A \in \invar$. If $\A$ is
1900: not $\rtd$-trivial, then we may write $\rtd$ as a convex combination of $\rtd$
1901: conditioned on $\A$ and $\rtd$ conditioned on the complement of $\A$. 
1902: Each of these two new probability measures is unimodular,
1903: yet distinct, so $\rtd$ is not extremal.
1904: 
1905: Conversely, suppose that $\invar$ is $\rtd$-trivial.
1906: Choose an environment and stationary initial bias
1907: that satisfy \ref e.proba/ and \ref e.getsall/, as in \ref r.choice/.
1908: Let $\A$ be an event of $\GG_*$.
1909: Let $\alpha$ be the function on $\paths$ that gives
1910: the frequency of visits to $\A$:
1911: $$
1912: \alpha\big((\gh, w_0), \Seq{w_n})\big) :=
1913: \liminf_{N \to\infty} {1 \over N} |\{n \le N \st (\gh, w_n) \in \A\}|
1914: \,.
1915: $$
1916: \ref t.rwrers/ allows us to apply the ergodic theorem to deduce that
1917: $\int \alpha \,d\Ph = (\nu\rtd)(\A)$,
1918: where $\nu \rtd$ stands for the measure $d(\nu \rtd)(\gh, \bp) = \nu_\gh(\bp)
1919: d\rtd(\gh, \bp)$.
1920: On the other hand, $\alpha$ is a shift-invariant function, which,
1921: according to \ref t.erg/, means that $\alpha$ is a constant $\Ph$-a.s.
1922: Thus, we conclude that
1923: $\alpha = (\nu\rtd)(\A)$ $\Ph$-a.s.
1924: %\ref t.rwrers/ allows us to apply the ergodic theorem to deduce that
1925: %$\alpha = \int \alpha \,d\Ph = (\nu\rtd)(\A)$ $\Ph$-a.s.,
1926: %where $\nu \rtd$ stands for the measure $d(\nu \rtd)(\gh, \bp) = \nu_\gh(\bp)
1927: %d\rtd(\gh, \bp)$.
1928: Consider any non-trivial
1929: convex combination of two unimodular probability
1930: measures, $\rtd_1$ and $\rtd_2$, that gives $\rtd$.
1931: Then $\Ph$ is a (possibly different) convex combination of $\Ph_1$ and $\Ph_2$.
1932: The above applies to each of $\Ph_i$ ($i=1, 2$) and the associated
1933: probability measures $a_i \nu \rtd_i$, where
1934: $a_i := \left(\int \nu_\gh(\bp) \,d\rtd_i(\gh, \bp)\right)^{-1}$.
1935: Therefore, we obtain that $a_1(\nu\rtd_1)(\A) = (\nu\rtd)(\A) =
1936: a_2(\nu\rtd_2)(\A)$.
1937: Since this holds for all $\A$, we obtain $a_1 (\nu\rtd_1) = a_2
1938: (\nu\rtd_2)$.
1939: Since $\rtd_1$ and $\rtd_2$ are probability measures, this is the same as
1940: $\rtd_1 = \rtd_2$, whence $\rtd$ is extremal.
1941: \Qed
1942: 
1943: We define the {\bf speed} of a path $\Seq{w_n}$ in a graph $\gh$ to be
1944: $\lim_{n \to\infty} \dist_\gh(w_0, w_n)/n$ when this limit exists, where
1945: $\dist_\gh$ indicates the distance in the graph $\gh$.
1946: 
1947: The following extends Lemma 4.2 of \ref b.BLS:pert/.
1948: 
1949: \procl p.speedexists \procname{Speed Exists}
1950: Let $\rtd$ be a unimodular probability measure on $\GG_*$ with an
1951: environment and stationary initial distribution $\nu_{\cbuldot}(\cbuldot)$ with
1952: $\int \nu_{\gh}(\bp) \,d\rtd(\gh, \bp) = 1$, so that the
1953: associated random walk distribution $\Ph$ is a probability measure.
1954: Then the speed of random walk exists $\Ph$-a.s.\ and is equal $\Ph$-a.s.\
1955: to an $\invar$-measurable function.
1956: The same holds for simple random walk when $\expdeg( \rtd) < \infty$.
1957: \endprocl
1958: 
1959: \proof
1960: Let $f_n\big((\gh, \bp), w\big) := \dist_\gh\big(w(0), w(n)\big)$. Clearly
1961: $$
1962: f_{n+m}\big((\gh, \bp), w\big)
1963: \le f_n\big((\gh, \bp), w\big) + f_m\big(\SS^n \big((\gh, \bp), w\big)\big)
1964: \,,
1965: $$
1966: so that the Subadditive Ergodic Theorem ensures
1967: that the speed $\lim_{n \to\infty} f_n\big( (\gh, \bp), w\big)/n$ exists
1968: $\Ph$-a.s.
1969: Since the speed is shift invariant, \ref t.erg/ shows that the speed is
1970: equal $\Ph$-a.s.\ to an $\invar$-measurable function.
1971: The same holds for simple random walk since it has an equivalent stationary
1972: probability measure (degree biasing) when $\expdeg(\rtd) < \infty$.
1973: \Qed
1974: 
1975: In the case of simple random walk on trees, we can actually calculate the
1976: speed\ftnote{*}{The publisher inadvertently changed the following proposition
1977: to a theorem in the published version. Also, the published version had an
1978: incorrect formula for \ref e.genspeed/.}:
1979: 
1980: \procl p.treespeed \procname{Speed on Trees}
1981: Let $\rtd \in \uni$ be concentrated on infinite trees.
1982: If $\rtd$ is extremal and $\expdeg(\rtd) < \infty$, then the speed of
1983: simple random walk is $\Ph$-a.s.\ 
1984: $$
1985: 1 - {2 \over \overline{\deg}(\mu)}
1986: \,.
1987: \label e.genspeed
1988: $$
1989: \endprocl
1990: 
1991: \proof Given a rooted tree $(\gh, \bp)$ and $x \in \vertex(\gh)$, write
1992: $|x|$ for the distance in $\gh$ between $\bp$ and $x$.
1993: The speed of a path $\Seq{w_n}$ is the limit
1994: $$
1995: \lim_{n \to \infty} {1 \over n} |w_n|
1996: = \lim_{n \to \infty} {1 \over n} \sum_{k = 0}^{n-1} \bigl(|w_{k+1}|
1997:        - |w_{k}|\bigr)\,.
1998: $$
1999: Now the strong law of large numbers for martingale differences (\ref
2000: b.Feller:book/, p.~243) gives
2001: $$
2002: \lim_{n \to \infty} {1 \over n} \sum_{k = 0}^{n-1} \bigl(|w_{k+1}|
2003:        - |w_{k}|\bigr)
2004: = \lim_{n \to \infty} {1 \over n} \sum_{k = 0}^{n-1} \E\bigl[|w_{k+1}|
2005:        - |w_{k}| \bigm| \Seq{w_i \st i \le k}\bigr] \quad\hbox{a.s.} %\label e.1
2006: $$
2007: Provided $w_k \ne o$, the $k$th term on the right equals
2008: $$
2009: {\deg_\gh w_k - 2 \over \deg_\gh w_k }\,. %\label e.2
2010: $$
2011: Since $\gh$ is a.s.\ infinite,
2012: $w_k = \bp$ for only a set of $k$ of density 0 a.s., whence the speed equals
2013: $$
2014: \lim_{n \to \infty} {1 \over n} \sum_{k = 0}^{n-1} {\deg_\gh w_k - 2 \over
2015: \deg_\gh w_k }\,. %\label e.sumdrift
2016: $$
2017: Since this is the limit of averages of an ergodic stationary sequence for
2018: the measure $d\sigma(\gh, \bp) = \deg_\gh (\bp)\, d\rtd(\gh,
2019: \bp)/\overline{\deg}(\rtd)$,
2020: the ergodic theorem
2021: tells us that it converges a.s.\ to the $\sigma$-mean of an element of the
2022: sequence,
2023: $$
2024: \int {\deg_\gh(\bp) - 2 \over \deg_\gh(\bp)} d\sigma(\gh, \bp)
2025: \,,
2026: $$
2027: which is the same as \ref e.genspeed/.
2028: \Qed
2029: %Also, see \ref t.deg-infinite/.
2030: 
2031: When we study percolation, the following consequence will be useful.
2032: 
2033: \procl p.transtrees \procname{Comparison of Transience on Trees}
2034: Suppose $\rtd \in \uni$ is concentrated on networks whose underlying graphs
2035: are trees that are transient for simple random walk.
2036: Suppose that the mark space is $(0, \infty)$, that marks $\mkmp(\cbuldot,
2037: \cbuldot)$ on edges are the same at both endpoints, that the environment is
2038: $p_\gh(x, y) := \mkmp(x, y)/\nu_\gh(x)$, where $\nu_\gh(x) := \sum_{y \sim
2039: x} \mkmp(x, y)$, and that $\int \nu_{\gh}(\bp) \,d\rtd(\gh, \bp) = 1$, so
2040: that the associated random walk distribution $\Ph$ is a probability
2041: measure.
2042: Then random walk is also transient with respect to the environment
2043: $p_\cbuldot(\cbuldot)$ $\rtd$-a.s.
2044: \endprocl
2045: 
2046: \proof
2047: Let $\ev A$ be the set of $p_\cbuldot(\cbuldot)$-recurrent networks.
2048: Suppose that $\rtd(\ev A) > 0$.
2049: By conditioning on $\ev A$, we may assume without loss of
2050: generality that $\rtd(\ev A) = 1$.
2051: By \ref t.deg2/ and the recurrence of simple random walk on trees with at
2052: most two ends, we have $\expdeg(\rtd) > 2$.
2053: Let $\epsilon > 0$ be sufficiently small that 
2054: $$
2055: \int |\{x \sim \bp \st \mkmp(\bp, x) \ge \epsilon\}| \,d\rtd(\gh, \bp) > 2
2056: \,.
2057: $$
2058: Since finite trees have average degree strictly less than 2, is follows
2059: that the subnetwork $(\gh_\epsilon, \bp)$, defined to be the connected
2060: component of $\bp$ formed by the edges with marks at least $\epsilon$,
2061: is infinite with positive probability.
2062: Let $\rtd'$ be the law of $(\gh_\epsilon, \bp)$ when $(\gh, \bp)$ has the
2063: law $\rtd$, and conditioned on the event $\ev B$ that 
2064: $(\gh_\epsilon, \bp)$ is infinite.
2065: Then $\rtd' \in \uni$ and $\expdeg(\rtd') > 2$.
2066: %Let $\Ph'$ be the associated measure on random walks.
2067: By \ref p.treespeed/, simple random walk has positive speed $\rtd'$-a.s.,
2068: so, in particular, is transient a.s.
2069: Now simple random walk is the walk corresponding to all edge weights in
2070: $\gh_\epsilon$ being, say, $1/\epsilon$.
2071: Rayleigh's monotonicity principle (\ref b.DoyleSnell/ or \ref b.LP:book/)
2072: now implies that random walk is transient $\Ph$-a.s.\ on $\ev B$.
2073: Thus, $\ev A \cap \ev B = \emptyset$.
2074: This contradicts our initial assumption that $\rtd(\ev A) = 1$.
2075: \Qed
2076: 
2077: The converse of \ref p.transtrees/ is not true, as there are transient
2078: reversible random walks on 1-ended trees (see \ref x.halfplane/ for an
2079: example of such graphs; weights can be defined appropriately).
2080: Also, \ref p.transtrees/ does not extend to arbitrary networks, as one may
2081: construct an invariant network on $\Z^3$ that gives a recurrent random
2082: walk.
2083: %Put very small conductances on very large periodically spaced cube
2084: %boundaries to effectively cut off the origin.
2085: 
2086: 
2087: 
2088: Given two probability measures $\rtd$ and $\rtd'$
2089: on rooted networks and one of the standard
2090: notions of product networks, one can define the {\bf independent product}
2091: $\rtd \itimes \rtd'$ of
2092: the two measures by choosing a network from each measure independently and
2093: taking their product, rooted at the ordered pair of the original roots.
2094: 
2095: \procl p.product \procname{Product Networks}
2096: Let $\rtd$ and $\rtd'$ be two unimodular probability measures on $\GG_*$.
2097: Then their independent product $\rtd \itimes \rtd'$ is also unimodular.
2098: If $\rtd$ and $\rtd'$ are both extremal, then so is $\rtd \itimes \rtd'$.
2099: \endprocl
2100: 
2101: \proof
2102: Let $\gh_n$ and $\gh'_n$ be finite connected networks whose random weak
2103: limits are $\rtd$ and $\rtd'$, respectively.
2104: Then $\gh_n \times \gh'_n$
2105: clearly has random weak limit $\rtd \itimes \rtd'$, whence the product
2106: is unimodular.
2107: Now suppose that both $\rtd$ and $\rtd'$ are extremal.
2108: Let $\ev A \in \invar$.
2109: Then $\ev A_{(\gh', \bp')} :=
2110: \{(\gh, \bp) \st (\gh, \bp) \times (\gh', \bp') \in \ev A \} \in
2111: \invar$ since $\Aut(\gh) \times \Aut(\gh') \subseteq \Aut(\gh \times
2112: \gh')$.
2113: Therefore, $\rtd \ev A_{(\gh', \bp')} \in \{0, 1\}$.
2114: On the other hand, $\ev A_{(\gh', \bp')} = \ev A_{(\gh', \bp'')}$ for all
2115: $\bp'' \in \verts(\gh')$ because $\ev A \in \invar$.
2116: Therefore,
2117: $\ev B := \{(\gh', \bp') \st \rtd \ev A_{(\gh', \bp')}  = 1\} \in \invar$,
2118: whence $\rtd' \ev B \in \{0, 1\}$.
2119: Hence Fubini's theorem tells us that $(\rtd \itimes \rtd')(\ev A) \in
2120: \{0, 1\}$, as desired.
2121: \Qed
2122: 
2123: 
2124: \procl r.weakmixing
2125: Another type of product that can sometimes be defined does not always
2126: produce an extremal measure from two extremal measures.
2127: That is, suppose that
2128: $\rtd$ and $\rtd'$ are two extremal unimodular probability measures on
2129: $\GG_*$ that, for simplicity, we assume are concentrated on networks with a
2130: fixed underlying transitive graph, $\gh$.
2131: Let $\rtd''$ be the measure on networks given by taking a fixed root,
2132: $\bp$, and choosing the marks as $(\mkmp, \mkmp')$, where $\mkmp$ gives a
2133: network with law $\rtd$, $\mkmp'$ gives a network with law $\rtd'$, and
2134: $\mkmp$, $\mkmp'$ are independent.
2135: Then it may be that $\rtd''$ is not extremal.
2136: For an example, consider the following.
2137: Fix an irrational number, $\alpha$.
2138: Given $x \in [0, 1]$, form the network $\gh_x$ on the
2139: integer lattice graph by marking each integer $n$ with the indicator
2140: that the fractional part of $n\alpha + x$ lies in $[0, 1/2]$.
2141: Let $\rtd$ be the law of $(\gh_x, 0)$ when $x$ is chosen uniformly.
2142: Then $\rtd$ is unimodular and extremal (by ergodicity of Lebesgue measure
2143: with respect to rotation by $\alpha$), but if $\rtd' = \rtd$ and $\rtd''$
2144: is the associated measure above, then $\rtd''$ is not
2145: extremal since when the marks come from $x, y \in [0, 1]$,
2146: the fractional part of $x - y$ is $\invar$-measurable.
2147: \endprocl
2148: 
2149: It may be useful to keep in mind the vast difference between stationarity
2150: and reversibility in this context.
2151: For example, let $T$ be a 3-regular tree and $\zeta$ be an end of $T$.
2152: Mark each edge by two independent random variables, one that is uniform on
2153: $[0, 1]$ and the other uniform on $[1, 2]$, with the latter one
2154: at its endpoint closer to $\zeta$ and with all these random variables
2155: mutually independent for different edges.
2156: Then simple random walk is stationary in this scenery, but not
2157: reversible, even though $T$ is a Cayley graph.
2158: 
2159: 
2160: 
2161: \bsection{Trace and Stochastic Comparison}{s.trace}
2162: 
2163: There is a natural trace associated to every measure in $\uni$.
2164: This trace is useful for making various comparisons.
2165: We illustrate this by extending results of \ref b.PittSC/ and
2166: \ref b.FontesMathieu/ on return
2167: probabilities of continuous-time random walks.
2168: %We also give some results that are used by \ref b.Lyons:est/ for
2169: %comparative enumeration.
2170: 
2171: Suppose that $\rtd$ is a unimodular probability measure on $\GG_*$.
2172: %Assume that the vertex marks are a.s.\ distinct.
2173: Consider the Hilbert space $H := \int^\oplus \ell^2\big(\vertex(\gh)\big)
2174: \,d\rtd(\gh, \bp)$, a direct integral (see, e.g., \ref b.Nielsen/ or \ref
2175: b.KR/, Chapter 14).
2176: Here, we always choose the canonical representative for each network,
2177: which, recall, is a network on the vertex set $\N$.
2178: The space $H$ is defined as the set
2179: of ($\rtd$-equivalence classes of) $\rtd$-measurable functions $f$ defined on
2180: canonical rooted networks $(\gh, \bp)$ that satisfy $f(\gh, \bp) \in
2181: \ell^2(\N)$ and $\int \|f(\gh, \bp)\|^2 \,d\rtd(\gh, \bp) < \infty$.
2182: We write $f = \int^\oplus f(\gh, \bp) \,d\rtd(\gh, \bp)$.
2183: The inner product is given by $(f, g) := \int \big(f(\gh, \bp), g(\gh,
2184: \bp)\big) \,d\rtd(\gh, \bp)$.
2185: Let $T : (\gh, \bp) \mapsto T_{\gh, \bp}$ be a measurable assignment of
2186: bounded linear operators on $\ell^2\big(\vertex(\gh)\big) = \ell^2(\N)$
2187: with finite
2188: supremum of the norms $\|T_{\gh, \bp}\|$.
2189: Then $T$ induces a bounded linear operator $T := T^\rtd := \int^\oplus T_{\gh,
2190: \bp} \,d\rtd(\gh, \bp)$ on $H$ via
2191: $$
2192: T^\rtd : \int^\oplus f(\gh, \bp)
2193: \,d\rtd(\gh, \bp) \mapsto \int^\oplus T_{\gh, \bp} f(\gh, \bp)
2194: \,d\rtd(\gh, \bp)
2195: \,.
2196: $$
2197: The norm $\|T^\rtd\|$ of $T^\rtd$ is the $\rtd$-essential supremum of
2198: $\|T_{\gh, \bp}\|$.
2199: Identify each $x \in \vertex(\gh)$ with the vector $\II x \in
2200: \ell^2\big(\vertex(\gh)\big)$.
2201: Let $\alg = \alg(\rtd)$ be the von Neumann algebra of ($\mu$-equivalence
2202: classes of) such maps $T$ that are equivariant in
2203: the sense that for all network isomorphisms $\phi : \gh_1 \to \gh_2$, all
2204: $\bp_1, x, y \in \verts(\gh_1)$ and all $\bp_2 \in \verts(\gh_2)$,
2205: we have $(T_{\gh_1, \bp_1} x, y) = (T_{\gh_2, \bp_2} \phi x, \phi y)$.
2206: For $T \in \alg$, we have in particular that $T_{\gh, \bp}$ depends on
2207: $\gh$ but not on the root $\bp$, so we shall simplify our notation and
2208: write $T_\gh$ in place of $T_{\gh, \bp}$.
2209: For simplicity, we shall even write $T$ for $T_\gh$ when no confusion can
2210: arise.
2211: Recall that if $S$ and $T$ are self-adjoint operators on a Hilbert space
2212: $H$, we write $S \le T$ if $\ip{S u, u} \le \ip{T u, u}$ for all $u \in H$.
2213: We claim that
2214: $$
2215: \tr(T) := \tr_\rtd(T)
2216: := \Ebig{(T_\gh \bp, \bp)} := \int (T_\gh \bp, \bp)
2217: \,d\rtd(\gh, \bp)
2218: $$
2219: is a {\bf trace} on $\alg$, i.e., $\tr(\cbuldot)$ is linear,
2220: $\tr(T) \ge 0$ for $T \ge 0$, and $\tr(S T) = \tr(T
2221: S)$ for $S, T \in \alg$.
2222: Linearity of $\tr$ is obvious.
2223: Also, the second property is obvious since the integrand is nonnegative for $T
2224: \ge 0$. The third property follows from the Mass-Transport Principle:
2225: We have
2226: $$\eqaln{
2227: \Ebig{(S T \bp, \bp)}
2228: &=
2229: \Ebig{(T \bp, S^* \bp)}
2230: =
2231: \Ebigg{\sum_{x \in \vertex(\gh)} (T \bp, x) (x, S^* \bp)}
2232: \cr&=
2233: \Ebigg{\sum_{x \in \vertex(\gh)} (T \bp, x) (S x, \bp)}
2234: =
2235: \Ebigg{\sum_{x \in \vertex(\gh)} (T x, \bp) (S \bp, x)}
2236: \cr&=
2237: \Ebig{(T S \bp, \bp)}
2238: \,.
2239: }$$
2240: In order to justify this use of the Mass-Transport Principle, we check
2241: absolute integrability:
2242: $$\eqaln{
2243: \Ebigg{\sum_{x \in \vertex(\gh)} |(T \bp, x) (x, S^* \bp)|}
2244: &\le
2245: \bigg(\Ebigg{\sum_{x \in \vertex(\gh)} |(T \bp, x)|^2}
2246: \Ebigg{\sum_{x \in \vertex(\gh)} |(x, S^* \bp)|^2} \bigg)^{1/2}
2247: \cr&=
2248: \bigg(\Ebig{\|T \bp\|^2} \Ebig{\|S^* \bp\|^2} \bigg)^{1/2}
2249: %\le
2250: %\|T_{\gh, \bp}\| \cdot \|S_{\gh, \bp}\|
2251: %\,.
2252: %Furthermore, we have
2253: %\Ebig{ \|T_{\gh, \bp}\| \cdot \|S_{\gh, \bp}\|}
2254: \le
2255: %\bigg(\Ebig{ \|T_{\gh, \bp}\|^2} \Ebig{\|S_{\gh, \bp}\|^2} \bigg)^{1/2}
2256: %=
2257: \|T\| \cdot \|S\|
2258: <\infty
2259: \,.
2260: }$$
2261: 
2262: A general property of traces that are finite and normal, as ours is,
2263: is that if $S \le T$,
2264: then $\tr f(S) \le \tr f(T)$ for any increasing function $f : \R \to \R$.
2265: %$\tr(e^S) \le
2266: %\tr(e^T)$. (Proof: for a positive integer $n$, we have $\tr(T^n) - \tr(S^n) =
2267: %\tr(T^n - S^n) = \tr((T-S)U) = \tr((T-S)^{1/2} U (T-S)^{1/2}) \ge 0$ for a
2268: %positive operator $U$.)
2269: One proof is as follows.
2270: First, if $f \ge 0$ and $T$ is self-adjoint, then $f(T) \ge 0$.
2271: Second, if $S, T \ge 0$, then $\tr (S T) = \tr \big(S^{1/2} T S^{1/2}\big)
2272: \ge 0$ since $S^{1/2} T S^{1/2} = (T^{1/2} S^{1/2})^* (T^{1/2}S^{1/2})  \ge 0$.
2273: Third, if $f$ is an increasing polynomial and $S \le T$, then
2274: $$
2275: {d \over d z} \Big(\tr f\big(S + z(T-S)\big)\Big)
2276: =
2277: \tr \Big(f'\big(S + z(T-S)\big)(T-S)\Big)
2278: \ge 0
2279: $$
2280: for $z \ge 0$
2281: since $f' \ge 0$, $S + z(T-S)$ is self-adjoint, and $0 \le T-S$.
2282: This shows that with these restrictive hypotheses, $\tr f(S) \le \tr f(T)$.
2283: Fourth, any monotone increasing function can be approximated by 
2284: an increasing polynomial.
2285: This shows the result in general.
2286: See \ref b.BrownKosaki/, pp.~6--7 for another proof. (They stated the result
2287: only for continuous $f$ with $f(0) = 0$ because they dealt with more
2288: general traces and operators, but such restrictions are not needed in our
2289: situation.)
2290: %They also assume the trace is faithful, which ours is, but that is not used
2291: %in their proof.
2292: 
2293: Recall from \ref s.extreme/ that given
2294: a network with positive edge
2295: weights and a time $t > 0$, we form the transition operator $P_t$ for
2296: continuous-time random walk whose rates are the edge weights; in the case
2297: of unbounded weights (or degrees), we take the minimal process, which dies
2298: after an explosion.
2299: If $A$ is the Laplacian of the network,
2300: then $P_t := e^{-A t}$.
2301: 
2302: \procl t.rtnpr \procname{Return Probabilities}
2303: Let the mark space be $\R^+$ and let $\rln$ be $\le$.
2304: Let $\rtd_i \in \uni$ have edge weights that are the same at both ends of each
2305: edge ($i = 1, 2$).
2306: Suppose that there is a unimodular $\rln$-coupling $\nu$ of
2307: $\rtd_1$ to $\rtd_2$.
2308: Let $P^{(i)}_t$ be the transition operators corresponding to the edge
2309: weights ($i = 1, 2$).
2310: Then 
2311: $$
2312: \int P^{(1)}_t(\bp, \bp) \,d\rtd_1(\gh, \bp) \ge \int P^{(2)}_t(\bp,
2313: \bp) \,d\rtd_2(\gh, \bp)
2314: $$
2315: for all $t > 0$.
2316: \endprocl
2317: 
2318: \proof
2319: Suppose first that the weights are bounded a.s.
2320: The assumptions tell us that the infinitesimal generators $A^{(i)} \in
2321: \alg$ satisfy $A^{(1)} \le A^{(2)}$, so that for all $t > 0$, we have
2322: $-A^{(1)} t \ge -A^{(2)} t$.
2323: Therefore
2324: $\int P^{(1)}_t(\bp, \bp) \,d\rtd_1(\gh, \bp) = \tr_{\nu}(e^{-A^{(1)} t})
2325: \ge \tr_{\nu}(e^{-A^{(2)} t}) = \int P^{(2)}_t(\bp, \bp)
2326: \,d\rtd_2(\gh, \bp)$.
2327: 
2328: In the case of unbounded weights, note that $P^{(i)}_t(\bp, \bp) = \lim_{n
2329: \to\infty} P^{(i, n)}_t(\bp, \bp)$ for $i = 1, 2$, where the superscript $n$
2330: refers to truncating all weights larger than $n$. (This is because we are
2331: looking at the minimal process, which is reversible.)
2332: %Given t >0, there is some distance R
2333: %such that the probability of having gone more than R while returning to o
2334: %at time t is less than epsilon. Thus, truncation will eventually give the
2335: %same rates within distance R of o, so we get the correct limiting return
2336: %probability at time t. Now use the BCT (we are taking expectation over (G,
2337: %o).)
2338: Thus, the result follows from the bounded case, which we have proved.
2339: \Qed
2340: 
2341: \ref t.rtnpr/ extends a result of \ref b.FontesMathieu/, who proved it in the
2342: case of $\Z^d$ for processes without explosions.
2343: \ref b.PittSC/, Lemma 3.1, prove an analogous
2344: comparison result for Cayley graphs, but with different assumptions on the
2345: pairs of rates (which are deterministic for them).
2346: The case of \ref t.rtnpr/ specialized to
2347: finite networks was proved earlier by Benjamini and Schramm;
2348: see Theorem 3.1 of \ref b.HeicklenHoffman/.
2349: 
2350: This theorem also gives a partial answer to a
2351: question of Fontes and Mathieu, who asked whether the same holds when
2352: $\rtd_i$ are supported on a single Cayley graph, are invariant under the
2353: group action, and are $\rln$-related.
2354: For example, the theorem shows that
2355: it holds when the networks (which are the environments for the
2356: random walks) are given by i.i.d.\ edge marks, since in such a case, it is
2357: trivial that being $\rln$-related implies the existence of a unimodular
2358: $\rln$-coupling.
2359: Also, in the amenable case, the existence of a unimodular $\rln$-coupling
2360: follows from the existence of an $\rln$-coupling, as is proved in \ref
2361: p.amen-average/; in the case of fixed amenable transitive graphs, this is a
2362: well-known averaging principle.
2363: 
2364: \procl q.rtnpr
2365: Does \ref t.rtnpr/ hold without the assumption of a unimodular
2366: $\rln$-coupling, but just an $\rln$-coupling?
2367: \endprocl
2368: 
2369: This question asks whether we can compare the traces from two different von
2370: Neumann algebras.
2371: One situation where we can do this is as follows.
2372: 
2373: 
2374: 
2375: If $\rtd_1$ and $\rtd_2$ are probability measures on $\GG_*$, then a
2376: probability measure $\nu$
2377: on $\GG_* \times \GG_*$ whose coordinate
2378: marginals are $\rtd_1$ and $\rtd_2$ is called a {\bf monotone graph
2379: coupling} of $\rtd_1$ and $\rtd_2$ if
2380: $\nu$ is concentrated on pairs of rooted
2381: networks $\big( (\gh_1, \bp), (\gh_2, \bp) \big)$
2382: that share the same roots and satisfy $\vertex(\gh_1) \subseteq
2383: \vertex(\gh_2)$.
2384: %In this case, decompose $\ell^2\big(\vertex(\gh_2)\big) =
2385: %\ell^2\big(\vertex(\gh_1)\big) \oplus \ell^2\big(\vertex(\gh_2) \setminus
2386: %\vertex(\gh_1)\big)$ and let $I = I_1 \oplus I_2$ be the corresponding
2387: %decomposition of the identity operator.
2388: In this instance,
2389: let $V_1$ be the inclusion of $\ell^2\big(\vertex(\gh_1)\big)$ in
2390: $\ell^2\big(\vertex(\gh_2)\big)$.
2391: %and $\bfz_2$ denote the 0 operator on
2392: %$\ell^2\big(\vertex(\gh_2) \setminus \vertex(\gh_1)\big)$.
2393: %Thus, if $T$ is an operator on $\ell^2\big(\vertex(\gh_1)\big)$, then $T
2394: %\oplus \bfz_2$ is its natural extension to
2395: %an operator on $\ell^2\big(\vertex(\gh_2)\big)$.
2396: %Conversely, if $T$ is an operator on $\ell^2\big(\vertex(\gh_2)\big)$, then
2397: %$V_1^* T V_1$ is its restriction to $\ell^2\big(\vertex(\gh_1)\big)$.
2398: %These give two natural ways to compare an operator on
2399: %$\ell^2\big(\vertex(\gh_1)\big)$ to an operator on
2400: %$\ell^2\big(\vertex(\gh_2)\big)$, which accounts for the two parts of our
2401: %next proposition.
2402: 
2403: When there is a unimodular coupling (as in \ref t.rtnpr/),
2404: the following result is easy.
2405: The fact that it holds more generally is useful.
2406: 
2407: 
2408: \comment{
2409: \procl p.equality \procname{Trace Comparison}
2410: Let $\nu$ be a monotone graph coupling of two unimodular probability measures
2411: $\rtd_1$ and $\rtd_2$.
2412: Let $T^{(i)} \in \alg$  ($i = 1, 2$).
2413: Suppose that 
2414: $$
2415: T^{(1)}_{\gh_1} \oplus \bfz_2 \le T^{(2)}_{\gh_2}
2416: \label e.cond1
2417: $$
2418: for $\nu$-almost all pairs $\big( (\gh_1, \bp), (\gh_2, \bp) \big)$.
2419: Then $\tr_{\rtd_1}(T^{(1)}) \le \tr_{\rtd_2}(T^{(2)})$ with equality iff
2420: $$
2421: T^{(1)}_{\gh_1} \oplus \bfz_2 = T^{(2)}_{\gh_2} \quad \nu\hbox{-a.s.}
2422: \label e.defeq1
2423: $$
2424: Similarly, if
2425: $$
2426: T^{(1)}_{\gh_1} \le V_1^* T^{(2)}_{\gh_2} V_1
2427: \label e.cond2
2428: $$
2429: for $\nu$-almost all pairs $\big( (\gh_1, \bp), (\gh_2, \bp) \big)$,
2430: then $\tr_{\rtd_1}(T^{(1)}) \le \tr_{\rtd_2}(T^{(2)})$ with equality iff
2431: $$
2432: T^{(1)}_{\gh_1} = V_1^* T^{(2)}_{\gh_2} V_1 \quad \nu\hbox{-a.s.}
2433: \label e.defeq2
2434: $$
2435: \endprocl
2436: }
2437: 
2438: \procl p.equality \procname{Trace Comparison}
2439: Let $\nu$ be a monotone graph coupling of two unimodular probability measures
2440: $\rtd_1$ and $\rtd_2$.
2441: Let $T^{(i)} \in \alg(\rtd_i)$ be self-adjoint with
2442: $$
2443: T^{(1)}_{\gh_1} \le V_1^* T^{(2)}_{\gh_2} V_1
2444: \label e.cond1
2445: $$
2446: for $\nu$-almost all pairs $\big( (\gh_1, \bp), (\gh_2, \bp) \big)$.
2447: Then $\tr_{\rtd_1}(T^{(1)}) \le \tr_{\rtd_2}(T^{(2)})$.
2448: If in addition
2449: for $\nu$-almost all pairs $\big( (\gh_1, \bp), (\gh_2, \bp) \big)$ and for
2450: all $x \in \vertex(\gh_1)$ we have
2451: $$
2452: \deg_{\gh_1}(x) < \deg_{\gh_2}(x)
2453: \implies
2454: \bigip{T^{(1)}_{\gh_1} x, x} < \bigip{T^{(2)}_{\gh_2} x, x}
2455: \,,
2456: \label e.cond2
2457: $$
2458: then either
2459: $$
2460: \tr_{\rtd_1}(T^{(1)}) < \tr_{\rtd_2}(T^{(2)})
2461: \label e.concl1
2462: $$
2463: or 
2464: $$
2465: \vertex(\gh_1) = \vertex(\gh_2)\,,\ \edges(\gh_1) = \edges(\gh_2)\,,
2466: \ \hbox{ and } \ T^{(1)}_{\gh_1} = T^{(2)}_{\gh_2} \quad \nu\hbox{-a.s.}
2467: \label e.concl2
2468: $$
2469: \endprocl
2470: 
2471: \proof
2472: Suppose that \ref e.cond1/ holds.
2473: The fact that $\tr_{\rtd_1}(T^{(1)}) \le \tr_{\rtd_2}(T^{(2)})$ is an
2474: immediate consequence of the definition of trace and of the hypothesis:
2475: $$\eqaln{
2476: \tr_{\rtd_1}(T^{(1)}) 
2477: &=
2478: \int \bigip{T^{(1)}_{\gh} \bp, \bp} \,d\rtd_1(\gh, \bp)
2479: =
2480: \int \bigip{T^{(1)}_{\gh_1} \bp, \bp} \,d\nu\big( (\gh_1, \bp), (\gh_2,
2481: \bp) \big)
2482: \cr&\le
2483: \int
2484: \bigip{V_1^* T^{(2)}_{\gh_2} V_1 \bp, \bp} \,d\nu\big( (\gh_1, \bp), (\gh_2,
2485: \bp) \big)
2486: \cr&=
2487: \int \bigip{T^{(2)}_{\gh_2} \bp, \bp} \,d\nu\big( (\gh_1, \bp), (\gh_2,
2488: \bp) \big)
2489: =
2490: \tr_{\rtd_2}(T^{(2)})
2491: \,.
2492: }$$
2493: 
2494: Suppose that equality holds in this inequality, i.e., \ref e.concl1/ fails.
2495: Then
2496: $$
2497: \bigip{T^{(1)}_{\gh_1} \bp, \bp}
2498: =
2499: \bigip{T^{(2)}_{\gh_2} \bp, \bp}
2500: \quad \nu\hbox{-a.s.}
2501: $$
2502: \comment{
2503: Let $B_n(\bp)$ be the vertex ball of radius $n$ about $\bp$.
2504: By the Mass-Transport Principle and the preceding equation, we have
2505: $$\eqaln{
2506: \int \sum_{x \in B_n(\bp)} \bigip{T^{(1)}_\gh x, x} \,d\rtd_1(\gh, \bp)
2507: &=
2508: \int |B_n(\bp)| \bigip{T^{(1)}_\gh \bp, \bp} \,d\rtd_1(\gh, \bp)
2509: \cr&=
2510: \int |B_n(\bp)| \bigip{T^{(2)}_\gh \bp, \bp} \,d\rtd_2(\gh, \bp)
2511: \cr&=
2512: \int \sum_{x \in B_n(\bp)} \bigip{T^{(2)}_\gh x, x} \,d\rtd_2(\gh, \bp)
2513: \,.
2514: }$$
2515: Comparing the first and last quantities in this chain of equalities and using
2516: the hypothesis again yields that
2517: }
2518: %
2519: Of course, we also have by hypothesis that $\nu$-a.s.,
2520: $$
2521: \bigip{T^{(1)}_{\gh_1} x, x}
2522: \le
2523: \bigip{T^{(2)}_{\gh_2} x, x}
2524: \label e.ineqatx
2525: $$
2526: for all $x \in \vertex(\gh_1)$.
2527: Assume now that \ref e.cond2/ holds.
2528: We shall prove that \ref e.concl2/ holds.
2529: First,
2530: we claim that $\nu$-a.s.,
2531: $$
2532: \vertex(\gh_1) = \vertex(\gh_2),\ 
2533: \edges(\gh_1) = \edges(\gh_2)\,,
2534: \label e.eqgraphs
2535: $$
2536: and
2537: $$
2538: \bigip{T^{(1)}_{\gh_1} x, x}
2539: =
2540: \bigip{T^{(2)}_{\gh_2} x, x}
2541: %\quad \hbox{for all } x \in B_n(\bp)\;\ \nu\hbox{-a.s.}
2542: \label e.atx
2543: $$
2544: %Hence the same holds for all $x \in \vertex(\gh_1)$.
2545: for all $x \in \vertex(\gh_2)$.
2546: If not, let $k$ be the smallest integer such that with positive
2547: $\nu$-probability, there is a vertex $x$ at $\gh_1$-distance $k$ from $\bp$
2548: where \ref e.atx/ does not hold.
2549: Such a $k$ exists by virtue of \ref e.cond2/.
2550: Consider $\bigip{T^{(i)}_{\gh_i} x, x}$ as part of the mark at $x$.
2551: According to \ref t.rwrers/, the random walk on $\gh_i$ given in \ref
2552: r.choice/ yields a shift-stationary measure $\Ph_i$ on trajectories
2553: $\big((\gh_i, w_0), \Seq{w_n \st n \ge 0}\big)$ for each $i$.
2554: In particular, the distribution of $\bigip{T^{(i)}_{\gh_i} w_k, w_k}$ is
2555: the same as that of $\bigip{T^{(i)}_{\gh_i} w_0, w_0}$.
2556: Now the latter is the same for $i=1$ as for $i=2$ (since $w_0 = \bp$).
2557: Note that for all $x$ at distance less than $k$ from $\bp$, we have
2558: $\deg_{\gh_1}(x) = \deg_{\gh_2}(x)$ by \ref e.cond2/.
2559: Thus, the walks may be coupled together up to time $k$, whence
2560: the distribution of $\bigip{T^{(i)}_{\gh_i} w_k, w_k}$ is
2561: not the same for $i=1$ as for $i=2$ in light of \ref
2562: e.ineqatx/ and choice of $k$.
2563: This is a contradiction.
2564: It follows that $\deg_{\gh_1}(x) = \deg_{\gh_2}(x)$ for all $x \in
2565: \vertex(\gh_1)$, whence \ref e.eqgraphs/ and \ref e.atx/ hold.
2566: 
2567: Now $T := T^{(2)}_{\gh_2} - T^{(1)}_{\gh_1} \ge 0$ is self-adjoint.
2568: It follows that
2569: for any $x, y \in \vertex(\gh_2)$
2570: and any complex number $\alpha$ of modulus 1,
2571: $$
2572: 0
2573: \le
2574: \bigip{T (\alpha x + y), \alpha x + y}
2575: =
2576: 2 \Re \big\{ \alpha \ip{T x, y} \big\}
2577: \,,
2578: $$
2579: whence $\bigip{T x, y} = 0$.
2580: %$$
2581: %\bigip{T^{(1)}_{\gh_1} x, y}
2582: %=
2583: %\bigip{T^{(2)}_{\gh_2} x, y}
2584: %\quad \hbox{for all } x , y \in \vertex(\gh_2)\;\ \nu\hbox{-a.s.}
2585: %$$
2586: That is, $T = 0$ and \ref e.concl2/ holds.
2587: \Qed
2588: 
2589: \comment{
2590: Since we also want to handle unbounded operators, sometimes arising as
2591: unbounded functions of bounded operators, we recall that
2592: a closed densely defined operator is {\bf affiliated}
2593: with $\alg$ if it commutes with all unitary operators
2594: that commute with $\alg$; see, e.g., \ref b.KadRing1/, p.~342.
2595: Write $\affalg$ for the set of closed densely defined operators affiliated
2596: with $\alg$.
2597: Also note that our trace on $\alg$ is obviously finite.
2598: Let $T \in \affalg$ be a self-adjoint operator 
2599: with spectral resolution $E_T$. We define the Borel measure $\spm_{\rtd, T}$
2600: by
2601: $$
2602: \spm_{\rtd, T}(B) :=
2603: \tr_\rtd\big(E_T(B)\big)
2604: \label e.specmsr
2605: $$
2606: for Borel subsets $B \subseteq \R$.
2607: We extend the trace by defining 
2608: $$
2609: \tr_\rtd(T) := 
2610: \int_0^\infty \lambda \,d\spm_{\rtd, T}(\lambda) 
2611: $$
2612: for positive operators $T \in \affalg$ and then by linearity to all of
2613: $\affalg$ when it makes sense.
2614: If $T \ge 0$ and $\tr(T) < \infty$, then $\bp \in \dmn(\sqrt T)$ since
2615: $$\eqaln{
2616: \lim_{n \to\infty} \|E_T[0, n] \sqrt T E_T[0, n] \bp \|^2 
2617: &=
2618: \lim_{n \to\infty} \bigip{E_T[0, n] T E_T[0, n] \bp, \bp}
2619: \cr&=
2620: \lim_{n \to\infty} \int_0^n \lambda \,d\spm_{\rtd, T}(\lambda) 
2621: =
2622: \tr_\rtd(T) < \infty
2623: \,.
2624: }$$
2625: In particular, 
2626: $$
2627: \|\sqrt T \bp\|^2 = \tr_\rtd(T)
2628: \,.
2629: $$
2630: Similarly, if $T$ is self-adjoint with polar decomposition $T = U |T|$ and
2631: $\tr_\rtd\big(|T|\big) < \infty$,
2632: then 
2633: $$
2634: \bigip{U \sqrt{|T|} \bp, \sqrt{|T|} \bp} = \tr_\rtd(T)
2635: \,.
2636: $$
2637: 
2638: % $S \le T$ means that the closure of $T - S$ is positive and self-adjoint.
2639: 
2640: % $V_1^* T^{(2)} V_1$ is affiliated with $\alg(\rtd_1)$ and its traces on
2641: % both algebras agree. We have the inequality of the $\rtd_1$-trace,
2642: % therefore.
2643: 
2644: We now extend \ref p.equality/.
2645: %For the definition of the measure topology, see \ref b.FackKos/, \S 1.5.
2646: 
2647: \procl p.equalityunbdd \procname{Trace Comparison of Unbounded Operators}
2648: Let $\nu$ be a monotone graph coupling of two unimodular probability measures
2649: $\rtd_1$ and $\rtd_2$.
2650: Let $T^{(i)} \in \affalg(\rtd_i)$ be self-adjoint with
2651: $\tr_{\rtd_i} (|T^{(i)}|) < \infty$ ($i = 1, 2$).
2652: %Let $e$ be the spectral resolution of $T^{(2)}$ and $e_n := e(-n, n)$ for
2653: %$n \ge 1$.
2654: %Suppose that $T^{(2, n)} \in \alg(\rtd_2)$ satisfy 
2655: %$$
2656: %|T^{(2, n)}| \le |T^{(2)}|
2657: %\label e.approxs
2658: %$$
2659: %and 
2660: %Then $V_1^* e_n T^{(2)} e_n V_1$ converges in the measure topology to an
2661: %operator $S \in \affalg(\rtd_1)$.
2662: %Suppose that the closure of $S - T^{(1)}$ is a positive operator.
2663: If $\|\sqrt{|T^{(1)}_{\gh_1}|} \bp\| \le \|\sqrt{|T^{(2)}_{\gh_2}|} \bp\|$
2664: for $\nu$-almost all pairs $\big( (\gh_1, \bp), (\gh_2, \bp) \big)$,
2665: then $x \in \dmn(\sqrt{|T^{(i)}|})$ for all $x \in \vertex(\gh_i)$ ($i = 1,
2666: 2$) and $\tr_{\rtd_1}(T^{(1)}) \le \tr_{\rtd_2}(T^{(2)})$.
2667: If in addition
2668: for $\nu$-almost all pairs $\big( (\gh_1, \bp), (\gh_2, \bp) \big)$ and for
2669: all $x \in \vertex(\gh_1)$ we have
2670: $$
2671: \deg_{\gh_1}(x) < \deg_{\gh_2}(x)
2672: \implies
2673: \|\sqrt{|T^{(1)}_{\gh_1}|} x\| < \|\sqrt{|T^{(2)}_{\gh_2}|} x\|
2674: \,,
2675: \label e.cond3
2676: $$
2677: then either
2678: \ref e.concl1/ holds or
2679: \ref e.concl2/ holds.
2680: \endprocl
2681: 
2682: \proof
2683: Note first that $e_n |T^{(2)}| e_n$ is a bounded operator, which is easily
2684: seen to be in $\alg(\rtd_1)$.
2685: Furthermore, this is an increasing sequence of operators that converges
2686: since $L^1(\rtd_1)$ is complete by Theorem 13 of \ref b.Segal/.
2687: \msnote{Or we might want to not use $\le$ in the hypo and use instead
2688: sqrts. Maybe then we don't even need the limits.}
2689: }
2690: 
2691: 
2692: \comment{
2693: \procl r.eq-ext
2694: An extension of the previous proposition is proved similarly, where we
2695: consider marks in $\marks^2$ and assume that $\nu$ is concentrated on pairs
2696: of rooted networks such that when the second coordinate is forgotten, we
2697: obtain identical rooted networks.
2698: \msnote{Explain what this means}
2699: \endprocl
2700: }
2701: 
2702: \comment{
2703: The following extension is useful as well (see \ref b.Lyons:est/).
2704: A continuous function $f : (0, \infty) \to \R$ is called {\bf operator
2705: monotone on $(0, \infty)$} if for any bounded self-adjoint operators $A, B$
2706: with spectrum in $(0, \infty)$ and $A \le B$, we have $f(A) \le f(B)$.
2707: According to L\"owner's theorem,
2708: such $f$ are precisely those that can be represented as
2709: $$
2710: f(t) =
2711: \alpha + \beta t + \int \left[{s \over 1+s^2} - {1 \over t +s}\right]
2712: d\kappa(s)
2713: \label e.pick
2714: $$
2715: for some $\alpha \in \R$, $\beta \ge 0$, and
2716: some positive Borel measure $\kappa$ on $\CO{0, \infty}$ with $\int
2717: d\kappa(s)/(1+s^2) < \infty$; see \ref b.Bhatia/, pp.~138--144 and \ref
2718: b.Donoghue/.
2719: Given $f$ as above, $0 \le T \in \alg$ and $\rtd \in \uni$, say that the
2720: pair $(f, T)$ is {\bf mild} with respect to $\rtd$ if 
2721: $$
2722: \int 
2723: \left[
2724: \int_0^1 \bigip{(T+s)^{-1} \bp, \bp}
2725: \,d\kappa(s)
2726: +
2727: \int_1^\infty \Big( {1 \over s} - \bigip{(T+s)^{-1} \bp, \bp} \Big)
2728: \,d\kappa(s)
2729: \right]
2730: \,d\rtd(\gh, \bp)
2731: < \infty
2732: \,.
2733: $$
2734: We are interested in extending the equality condition of \ref p.equality/
2735: to unbounded functions of unbounded operators.
2736: For simplicity, we extend $\alg$ to the set of equivariant measurable maps
2737: $T$ such that $T_{\gh}$ is a linear operator 
2738: defined on the linear span of the vectors $x \in \vertex(\gh)$.
2739: %Note that if the range of $T_{\gh} \ge I$
2740: %is dense in $\ell^2\big(\vertex(\gh)\big)$,
2741: %then the operator $T_\gh^{-1}$ has norm bounded by
2742: %$1$ on the range of $T_\gh$,
2743: %and therefore it extends to a bounded operator on all of
2744: %$\ell^2(\vertex(\gh)$.
2745: 
2746: 
2747: \procl l.incrlimit
2748: Let $T_n$ be bounded self-adjoint operators on a Hilbert space $\HH$ and
2749: $T$ an unbounded self-adjoint operator on $\HH$ with domain $\DD(T)$.
2750: Suppose that $I \le T_n \le T_{n+1}$ for all $n$ and that for all $x \in
2751: \DD(T)$, we have $\bigip{T_n x, x} \to \bigip{T x, x}$ as $n \to\infty$.
2752: Then $\bigip{T_n^{-1} x, x} \to \bigip{T^{-1} x, x}$ as $n \to\infty$ for
2753: all $x \in \HH$.
2754: \endprocl
2755: 
2756: \proof
2757: Note that $T$ is invertible by
2758: Theorems 13.13 and 13.31 of \ref b.Rudin:FA/.
2759: Since the function $t \mapsto -1/t$ is operator monotone on $(0, \infty)$,
2760: we have that the pointwise limit $A$ of $T_n^{-1}$ exists and satisfies
2761: $T_n^{-1} \ge A \ge T^{-1}$.
2762: Therefore
2763: $T_n \le A^{-1} \le T$.
2764: (The proof of operator monotonicity on p.~114 of \ref b.Bhatia/ applies as
2765: well to unbounded self-adjoint operators.)
2766: On the other hand, the pointwise limit of $T_n$ is $T$ on $\DD(T)$, whence
2767: $A = T$ on $\DD(T)$.
2768: Since $A$ is also self-adjoint, it follows that its domain is the same as
2769: $\DD(T)$ and so $A = T$.
2770: \Qed
2771: 
2772: \msnote{Perhaps we want to do the previous theorem for unbounded operators
2773: instead.}
2774: 
2775: \procl c.unbdd \procname{Traces of Unbounded Operators}
2776: Let $\nu$ be a monotone graph coupling of two unimodular probability measures
2777: $\rtd_1$ and $\rtd_2$.
2778: Let $T^{(i, M)} \in \alg$ for $i = 1, 2$ and $M \in \N$ be such that
2779: $$
2780: T^{(1, M)}_{\gh_1} \oplus \bfz_2 \le T^{(2, M)}_{\gh_2} 
2781: %\,.
2782: \label e.mono
2783: $$
2784: and
2785: $$
2786: 0 \le T^{(i, M)}_{\gh_i} \le T^{(i, M+1)}_{\gh_i}
2787: %\qquad \nu\hbox{-a.s.}
2788: \label e.increasing
2789: $$
2790: for $i = 1, 2$, all $M \in \N$ and for
2791: $\nu$-almost all pairs $\big( (\gh_1, \bp), (\gh_2, \bp) \big)$, where
2792: $\bfz_2$ denotes the 0 operator on
2793: $\ell^2\big(\vertex(\gh_2) \setminus \vertex(\gh_1)\big)$.
2794: Assume that for all $\gh$ and all $x, y \in \vertex(\gh)$, we have
2795: $\lim_{M \to\infty} \bigip{T^{(i, M)}_{\gh} x, y} = \bigip{T^{(i)}_{\gh} x,
2796: y}$ for some $T^{(i)} \in \alg$ that has a self-adjoint
2797: extension ($i = 1, 2$).
2798: %Suppose that for all $s > 0$ and $i = 1, 2$, the range of $T^{(i)}_{\gh_i} +
2799: %s$ is dense in $\ell^2\big(\vertex(\gh_i)\big)$.
2800: Let $f$ be an operator monotone function on $(0, \infty)$ with
2801: $$
2802: \lim_{t \downarrow 0} t f(t) = 0
2803: \label e.nomass
2804: $$
2805: and such that $(f, T^{(i, 1)})$ is mild with respect to $\rtd_i$ for $i = 1,
2806: 2$.
2807: If
2808: $$
2809: \lim_{M \to\infty} \lim_{\epsilon \downarrow 0}
2810: \left[
2811: \tr_{\rtd_2}\big(f(T^{(2, M)} + \epsilon)\big) -
2812: \tr_{\rtd_1}\big(f(T^{(1, M)} + \epsilon)\big)
2813: \right]
2814: =
2815: 0
2816: \label e.limit
2817: $$
2818: and
2819: for $\nu$-almost all pairs $\big( (\gh_1, \bp), (\gh_2, \bp) \big)$ and for
2820: all $x \in \vertex(\gh_1)$ we have
2821: $$
2822: \deg_{\gh_1}(x) < \deg_{\gh_2}(x)
2823: \implies
2824: \forall s > 0 \quad
2825: \Bigip{\big(T^{(1)}_{\gh_1} + s\big)^{-1} x, x} <
2826: \Bigip{\big(T^{(2)}_{\gh_2} + s\big)^{-1} x, x}
2827: \,,
2828: \label e.invs
2829: $$
2830: then $T^{(1)}_{\gh_1} = T^{(2)}_{\gh_2}$ $\nu$-a.s.
2831: \endprocl
2832: 
2833: \proof
2834: Write $f$ as in \ref e.pick/.
2835: If $\kappa = 0$, then the result is trivial, so assume $\kappa \ne 0$.
2836: The assumption \ref e.nomass/ ensures that $\kappa\big(\{0\}\big) = 0$.
2837: For $\epsilon > 0$, $M \in \N$ and 
2838: $\nu$-almost each pair $\big( (\gh_1, \bp), (\gh_2, \bp) \big)$, 
2839: we have
2840: $$\eqaln{
2841: \bigip{f(T^{(2, M)}_{\gh_2} + \epsilon) \bp, \bp}
2842: &-
2843: \bigip{f(T^{(1, M)}_{\gh_1} + \epsilon) \bp, \bp}
2844: \cr&\hskip-1in=
2845: \beta \bigg[\bigip{T^{(2, M)}_{\gh_2} \bp, \bp}
2846: -
2847: \bigip{T^{(1, M)}_{\gh_1} \bp, \bp} \bigg]
2848: \cr&\hskip-1in\qquad+
2849: \int
2850: \left[
2851: \Bigip{ \big(T^{(1, M)}_{\gh_1} + \epsilon +s \big)^{-1} \bp, \bp} -
2852: \Bigip{ \big(T^{(2, M)}_{\gh_2} + \epsilon +s \big)^{-1} \bp, \bp}
2853: \right]
2854: d\kappa(s)
2855: \cr&\hskip-1in\ge
2856: \int
2857: \left[
2858: \Bigip{ \big(T^{(1, M)}_{\gh_1} + \epsilon +s \big)^{-1} \bp, \bp} -
2859: \Bigip{ \big(T^{(2, M)}_{\gh_2} + \epsilon +s \big)^{-1} \bp, \bp}
2860: \right]
2861: d\kappa(s)
2862: %\cr&\hskip-1in\ge
2863: %\int
2864: %\left[
2865: %\Bigip{ \big(T^{(1)}_{\gh_1} + \epsilon +s \big)^{-1} \bp, \bp} -
2866: %\Bigip{ \big(T^{(2)}_{\gh_2} + \epsilon +s \big)^{-1} \bp, \bp}
2867: %\right]
2868: %d\kappa(s)
2869: \,,
2870: \label e.upb
2871: }
2872: $$
2873: where the inequality is from
2874: \ref e.mono/ and the fact that $\beta \ge 0$.
2875: %, while the second inequality
2876: %is from \ref e.increasing/.
2877: By \ref e.increasing/,
2878: %and the fact that
2879: %the function $t \mapsto -1/t$ is operator monotone on $(0, \infty)$,
2880: we have for every $M \ge 1$, $\epsilon > 0$, and $s > 0$,
2881: $$
2882: \bigip{(T^{(i, M)}_\gh+\epsilon + s)^{-1} \bp, \bp}
2883: \le
2884: \bigip{(T^{(i, 1)}_\gh+s)^{-1} \bp, \bp}
2885: \,,
2886: $$
2887: and so for $\epsilon < 1 < s$,
2888: $$
2889: {1 \over s} - \bigip{(T^{(i, M)}_\gh+\epsilon + s)^{-1} \bp, \bp} 
2890: \le
2891: {1 \over 1+s^2} + {1 \over 1+s} - \bigip{(T^{(i, 1)}_\gh+1+s)^{-1} \bp, \bp} 
2892: \,.
2893: $$
2894: Because $(f, T^{(i, 1)})$ is mild with respect to $\rtd_i$, we may
2895: therefore apply the Lebesgue
2896: dominated convergence theorem and \ref l.incrlimit/ to obtain that
2897: $$\eqaln{
2898: \lim_{M \to\infty}
2899: \lim_{\epsilon \downarrow 0}
2900: &\int
2901: \int_0^\infty
2902: \left[
2903: {s \over 1 + s^2} -
2904: \Bigip{ \big(T^{(i, M)}_{\gh} + \epsilon +s \big)^{-1} \bp, \bp}
2905: \right]
2906: d\kappa(s)
2907: \,d\rtd_i(\gh, \bp)
2908: \cr&=
2909: \int
2910: \int_0^\infty
2911: \left[
2912: {s \over 1 + s^2} -
2913: \Bigip{ \big(T^{(i)}_{\gh} + \epsilon +s \big)^{-1} \bp, \bp}
2914: \right]
2915: d\kappa(s)
2916: \,d\rtd_i(\gh, \bp)
2917: \,.
2918: }$$
2919: Combining this with \ref e.limit/ and \ref e.upb/ yields
2920: $$
2921: \int
2922: \int_0^\infty
2923: \left[
2924: \Bigip{ \big(T^{(1)}_{\gh_1} +s \big)^{-1} \bp, \bp} -
2925: \Bigip{ \big(T^{(2)}_{\gh_2} +s \big)^{-1} \bp, \bp}
2926: \right]
2927: d\kappa(s)
2928: \,d\nu\big( (\gh_1, \bp), (\gh_2, \bp) \big)
2929: =
2930: 0
2931: \,.
2932: $$
2933: Now the integrand is non-negative, whence
2934: for $\kappa$-almost all $s > 0$ and
2935: for $\nu$-almost all pairs $\big( (\gh_1, \bp), (\gh_2, \bp) \big)$,
2936: $$
2937: \Bigip{ \big(T^{(1)}_{\gh_1} +s \big)^{-1} \bp, \bp}
2938: =
2939: \Bigip{ \big(T^{(2)}_{\gh_2} +s \big)^{-1} \bp, \bp}
2940: \,.
2941: $$
2942: Note that the operators $\big(T^{(i)}_{\gh_i} +s \big)^{-1}$ are
2943: bounded with norm at most $s^{-1}$.
2944: Therefore, we may conclude from \ref p.equality/ that these operators are
2945: equal $\nu$-a.s.
2946: Now Theorems 13.13 and 13.31 of \ref b.Rudin:FA/ tell us that
2947: $T^{(i)}_{\gh} + s$ is a bijection of its domain with
2948: $\ell^2\big(\vertex(\gh)\big)$.
2949: Hence $T^{(1)}_{\gh_1} + s = T^{(2)}_{\gh_2} + s$ $\nu$-a.s.,
2950: which gives us the desired result.
2951: \Qed
2952: }
2953: 
2954: 
2955: \bsection{Percolation}{s.extend}
2956: 
2957: We now begin our collection of
2958: extensions of results that are known for unimodular fixed graphs.
2959: For most of the remainder of the paper, we consider graphs without marks, or,
2960: equivalently, with constant marks, except that marks are used as explained
2961: below to perform percolation on the given graphs.
2962: We begin this section on percolation with some preliminary results on
2963: expected degree.
2964: 
2965: \procl t.deg-infinite \procname{Minimal Expected Degree}
2966: If $\rtd$ is a unimodular probability measure on $\GG_*$ concentrated on
2967: infinite graphs, then $\expdeg(\rtd) \ge 2$.
2968: \endprocl
2969: 
2970: This is proved exactly like Theorem 6.1 of \BLPSgip\ is proved.
2971: In the context of equivalence relations, this is well known and was perhaps
2972: first proved by \ref b.Levitt/.
2973: 
2974: \procl t.deg2 \procname{Degree Two}
2975: If $\rtd$ is a unimodular probability measure on $\GG_*$ concentrated on
2976: infinite graphs,
2977: then $\expdeg (\rtd) = 2$ iff $\rtd$-a.s.\ $\gh$ is a tree with at most 2 ends.
2978: \endprocl
2979: 
2980: The proof is like that of Theorem 7.2 of \BLPSgip.
2981: 
2982: 
2983: \procl p.limittree \procname{Limits of Trees}
2984: If $G_n$ are finite trees with random weak limit $\rtd$, then
2985: $\expdeg(\rtd) \le 2$ and $\rtd$ is concentrated on trees with at most 2
2986: ends.
2987: \endprocl
2988: 
2989: \proof
2990: Since $\expdeg\big(U(G_n)\big) < 2$, we have $\expdeg(\rtd) \le 2$.
2991: The remainder follows from \ref t.deg2/.
2992: \Qed
2993: 
2994: We now discuss what we mean by percolation on a random rooted network.
2995: Given a probability measure $\rtd \in \uni$, we may wish to randomly
2996: designate some of the edges of the random network ``open".
2997: For example, in Bernoulli($p$) bond percolation, each edge is independently
2998: open with probability $p$.
2999: %Formally, fix a homeomorphism $\phi_2 : \marks \times [0, 1] \to \marks$.
3000: Recall that in \ref s.finite/ 
3001: we have added a second independent uniform $[0, 1]$ coordinate to
3002: the edge marks to form $\rtd^\bn \in \uni$, the standard coupling of
3003: Bernoulli percolation on $\rtd$.
3004: For $p \in [0, 1]$, one can define {\bf Bernoulli($p$) bond
3005: percolation on} $\rtd$
3006: as the measure $\rtd^\bn_p$ that replaces the second coordinate of each edge
3007: mark by ``open" if it is at most $p$ and by ``closed" otherwise.
3008: 
3009: Note that if $\phi : \marks \times [0, 1] \to \marks$ is the projection
3010: onto the first coordinate, then $\rtd = \rtd^\bn_p \circ \phi^{-1}$.
3011: We have changed the mark space, but a fixed homeomorphism would bring it back
3012: to $\marks$.
3013: Thus, more generally, if $\psi : \marks \to \marks$ is Borel, then we call
3014: $\rtd$ a {\bf percolation} on $\rtd \circ \psi^{-1}$.
3015: 
3016: 
3017: \procl d.ins
3018: Let $G=\big(\verts(G),\edges(G)\big)$ be a graph.
3019: Given a configuration $A\in \two^{\edges(G)}$ and an edge
3020: $e\in\edges(G)$, denote $\ins_e A$ the element of $\two^{\edges(G)}$ that
3021: agrees with $A$ off of $e$ and is $1$ on $e$.
3022: For $\A\subset \two^{\edges(G)}$, we write $\ins_e\ev A := \{\ins_e A \st A \in
3023: \A\}$.
3024: For bond percolation, call an edge ``closed" if it is marked ``0" and ``open"
3025: if it is marked ``1".
3026: A bond percolation process $\P$ on $G$ is {\bf insertion tolerant} if
3027: $\P(\ins_e\ev A)>0$
3028: for every $e\in\edges(G)$ and every Borel $\A\subset \two^{\edges(G)}$
3029: satisfying $\P(\A)>0$.
3030: The primary subtlety in extending this notion to percolation on unimodular
3031: random networks is that it may not be possible to pick an edge measurably from
3032: a rooted-automorphism-invariant set of edges.
3033: Thus, we shall make an extra assumption of distinguishability with marks.
3034: That is, a percolation process $\P$ on a unimodular probability measure on
3035: $\GG_*$ is {\bf insertion tolerant} if $\P$-a.s.\ there is no nontrivial
3036: rooted isomorphism of the marked network and
3037: for any event $\A \subseteq \{ (A,\gh)
3038: \st A \subseteq \edges(\gh),\; \gh \in \GG_* \}$ with $\P(\A) > 0$ and any
3039: Borel function $e : \gh \mapsto e(\gh) \in \edges(\gh)$ defined on
3040: $\GG_*$, we have $\P(\ins_e \A) > 0$.
3041: \endprocl
3042: 
3043: For example, Bernoulli($p$) bond percolation
3044: is insertion tolerant when $p\in \OC{0, 1}$.
3045: 
3046: We call a connected component of open edges (and their endpoints) a {\bf
3047: cluster}.
3048: Given a rooted graph $(\gh, \bp)$, define
3049: $$
3050: \pc(\gh, \bp) := \sup \{ p \st \hbox{Bernoulli}(p)\hbox{ percolation on }
3051: \gh \hbox{ has no infinite clusters a.s.}\}
3052: \,.
3053: $$
3054: Clearly $\pc$ is $\invar$-measurable, so if $\rtd \in \uni$ is extremal,
3055: then there is a constant $\pc(\rtd)$ such that $\pc(\gh, \bp) = \pc(\rtd)$
3056: for $\rtd$-a.e.\ $(\gh, \bp)$.
3057: 
3058: 
3059: \procl x.pclower
3060: Even if $\rtd \in \uni$ satisfies $\expdeg(\rtd) < \infty$, it does
3061: not necessarily follow that $\pc(\gh) > 0$ for $\rtd$-a.e.\ $(\gh, \bp)$.
3062: For example, let $p_k := 1/[k(k+1)]$ for $k \ge 1$ and $p_0 := 0$.
3063: Let $\UGW$ be the corresponding unimodular Galton-Watson measure (see
3064: \ref x.AGW/).
3065: Then $\expdeg(\UGW) = 6/(12-\pi^2)$ by \ref e.degAGW/,
3066: but since $\sum k p_k = \infty$,
3067: we have $\pc(\gh) = 0$ a.s.\ by \ref b.Lyons:rwpt/.
3068: \endprocl
3069: 
3070: A more elaborate example
3071: shows that no stochastic bound on the
3072: degree of the root,
3073: other than uniform boundedness,
3074: implies $\pc(\rtd) > 0$:
3075: 
3076: \procl x.pclower2
3077: Given $a_n > 0$, we shall construct $\rtd \in \uni$ such that
3078: $\rtd[\deg_\gh(\bp) \ge n] < a_n$ for all large $n$
3079: and $\pc(\gh) = 0$ for $\rtd$-a.e.\ $(\gh, \bp)$.
3080: We may assume that $\sum a_n < \infty$.
3081: Consider the infinite tree $T$ of degree $3$ and $\bp \in \vertex(T)$.
3082: Let $j \geq 2$ and set $p_j := \frac{1}{2} + \frac{1}{j}$. 
3083: Consider supercritical Bernoulli($p_j$)
3084: bond percolation on $T$,
3085: so that
3086: $$
3087: \theta_j := \P[\bp
3088: \hbox{ belongs to an infinite component}] > 0
3089: \,.
3090: $$
3091: Define $q_j < 1$ by
3092: $p_j q_j = \frac{1}{2} + \frac{1}{2j}$,
3093: so that the fragmentation of an infinite $p_j$-component by an independent
3094: $q_j$ percolation process will still contain infinite components.
3095: Let $N_j$ be the smallest integer with
3096: $(1 - \frac{1}{j^2})^{N_j} < 1 - q_j$.
3097: Finally, choose some sequence $1 > r_j \downarrow 0$ sufficiently fast.
3098: We label the edges of $T$ by the following
3099: operations, performed independently for each $j \geq 2$.
3100: 
3101: Take the infinite components of Bernoulli($p_j$) bond percolation on $T$.
3102: ``Thin" by retaining each component independently with probability $r_j$ and
3103: deleting other components.
3104: For each edge $e$ in the remaining components, let $L_j(e) := N_j$, while
3105: $L_j(e) := 1$ for deleted edges $e$.
3106: 
3107: Now let $L(e) := \sup_j L_j(e)$.
3108: Since $r_j \to 0$ fast, $L(e) < \infty$ for all $e$ a.s.
3109: Consider the graph $G$ obtained by replacing each edge $e$ of $T$ by $L(e)$
3110: parallel chains of length 2.
3111: To estimate $\deg_G(\bp)$, note that the chance that $\bp$ is incident
3112: to an edge $e$ with $L_j(e) = N_j$ (thus contributing at most $3N_j$ to the
3113: degree) equals $r_j\theta_j$.  
3114: Thus, by choice of $\Seq{r_j}$, we can make the root-degree distribution
3115: have tail probabilities eventually less than $a_n$.
3116: Now consider Bernoulli($1/j$) bond percolation on $G$.
3117: For an edge $e$ of $T$ which is replaced by $N_j$ chains of $G$,
3118: the chance of percolating across at least one of these $N_j$
3119: chains is (by definition of $N_j$) larger than $q_j$.
3120: Thus the percolation clusters on $G$ dominate the $q_j$-percolation
3121: clusters on the retained components of the original infinite
3122: $p_j$-percolation clusters on $T$, and as observed above must therefore
3123: contain infinite components.
3124: 
3125: To see how to make this into a probability measure in $\uni$, note that
3126: $L$ is an invariant random network on $T$.
3127: Therefore, we obtain a probability measure in
3128: $\uni$ by \ref t.uni-vs-invar/.
3129: We may now use the edge labels to replace an edge labeled $n$ by $n$ parallel
3130: chains of length 2, followed by
3131: a suitable re-rooting
3132: as in \ref x.stretched/ (below).
3133: This gives a new probability measure in $\uni$ that has
3134: the property desired, since 
3135: the expected degree of the root is finite by the hypothesis $\sum_n a_n <
3136: \infty$.
3137: Also, the re-rooting stochastically decreases the degree since it
3138: introduces roots of degree 2.
3139: \endprocl
3140: 
3141: 
3142: The following extends a well-known result of \ref b.HP:uniq/.
3143: The proof is the same.
3144: 
3145: \procl t.uniq \procname{Uniqueness Monotonicity and Merging Clusters}
3146: Let $\rtd$ be a unimodular probability measure on $\GG_*$.
3147: Let $p_1 < p_2$ and $\P_i$ ($i=1,2$) be the corresponding Bernoulli($p_i$) bond
3148: percolation processes on $\rtd$.
3149: If there is a unique infinite cluster $\P_1$-a.s., then
3150: there is a unique infinite cluster $\P_2$-a.s.  Furthermore, in the
3151: standard coupling of Bernoulli percolation processes, if $\rtd$ is extremal,
3152: then $\rtd$-a.s.\ for all $p_1, p_2$ satisfying $\pc(\rtd) < p_1 < p_2 \le 1$,
3153: every infinite $p_2$-cluster contains an infinite $p_1$-cluster.
3154: \endprocl
3155: 
3156: As a consequence, for extremal $\rtd \in \uni$,
3157: there is a constant $\pu(\rtd)$ such that for any
3158: $p > \pu(\rtd)$, we have $\P_{p}$-a.s., there is a
3159: unique infinite cluster, while for any $p < \pu(\rtd)$,
3160: we have $\P_{p}$-a.s., there is not a unique infinite cluster.
3161: 
3162: Every unimodular probability measure on $\GG_*$ can be written as a Choquet
3163: integral of extremal measures.
3164: In the following, we refer to these extremal measures as ``extremal
3165: components".
3166: 
3167: \procl l.ergcomp
3168: If\/ $\P$ is an insertion-tolerant percolation on a
3169: unimodular random network that is concentrated on infinite graphs, then
3170: almost every extremal component of\/ $\P$ is insertion tolerant.
3171: \endprocl
3172: 
3173: \proof
3174: The proof can be done precisely as that of Lemma 1 of \ref b.GKN/, but
3175: by using the present
3176: Theorems \briefref t.rwrers/, \briefref t.erg/, and \briefref
3177: t.findroot/, as well as \ref r.choice/ to replace the use of a
3178: measure-preserving transformation in \ref b.GKN/ by the shift on
3179: trajectories of a stationary Markov chain.
3180: The fact that $\P$ is concentrated on infinite graphs is used to deduce
3181: that the corresponding Markov chain is not positive recurrent, whence the
3182: asymptotic frequency of visits to any given neighborhood of the root is 0.
3183: \Qed
3184: 
3185: \procl c.ninf \procname{Number of Infinite Clusters}
3186: If\/ $\P$ is an insertion-tolerant percolation on a
3187: unimodular random network, then
3188: $\P$-almost surely, the number of infinite clusters is $0,1$ or $\infty$.
3189: \endprocl
3190: 
3191: The proof is standard for the extremal components; cf.\ \ref b.NS:number/.
3192: 
3193: The following extends \ref b.HP:uniq/ and Proposition 3.9 of \ref
3194: b.LS:indist/.
3195: The proof is parallel to that of the latter.
3196: 
3197: \procl p.niso
3198: \assumptions.
3199: Then $\P$-a.s.\ each infinite cluster that has at least $3$ ends has
3200: no isolated ends.
3201: \endprocl
3202: 
3203: The following corollary is proved just like Proposition 3.10 of \ref
3204: b.LS:indist/.
3205: There is some overlap with Theorem 3.1 of \ref b.Paulin/.
3206: 
3207: \procl c.isolated \procname{Many Ends}
3208: \genassumptions.
3209: If there are infinitely many infinite clusters $\P$-a.s., then $\P$-a.s.\ every
3210: infinite cluster has continuum many ends and no isolated end.
3211: \endprocl
3212: 
3213: The following extends Lemma 7.4 and Remark 7.3
3214: of \BLPSgip\ and is proved similarly.
3215: 
3216: \procl l.trim \procname{Subforests}
3217: \assumptions.
3218: If\/ $\P$-a.s.\ there is a component of the open subgraph
3219: $\omega$ with at least three ends, then
3220: there is a percolation $\fo$ on $\omega$ whose components are trees
3221: such that
3222: a.s.\ whenever a component $K$ of $\omega$ has at least three ends,
3223: there is a component of $K \cap \fo$ that has infinitely many ends and has
3224: $\pc < 1$.
3225: \endprocl
3226: 
3227: The following extends Proposition 3.11 of \ref b.LS:indist/ and is proved
3228: similarly (using the preceding \ref l.trim/).
3229: 
3230: \procl p.trans \procname{Transient Subtrees}
3231: \genassumptions.
3232: If there are $\P$-almost surely infinitely many
3233: infinite clusters,  then $\P$-a.s.\ each infinite cluster is transient and,
3234: in fact, contains a transient tree.
3235: \endprocl
3236: 
3237: In order to use this, we shall use the comparison of simple to network
3238: random walks given in \ref p.transtrees/.
3239: 
3240: \procl d.inds
3241: A percolation process $\P$ on a unimodular probability measure on
3242: $\GG_*$ has {\bf indistinguishable infinite clusters} if
3243: for any event $\A \subseteq \{ \big(A,(\gh, \bp)\big)
3244: \st A \in \two^{\vertices(\gh)} \times \two^{\edges(\gh)},\;
3245: (\gh, \bp) \in \GG_* \}$ that is invariant under non-rooted isomorphisms,
3246: almost surely, for all infinite clusters $C$ of the open subgraph
3247: $\omega$, we have $(C,\omega)\in\ev A$, or for all
3248: infinite clusters $C$, we have $(C,\omega)\notin\ev A$.
3249: \endprocl
3250: 
3251: 
3252: The following extends Theorem 3.3 of \ref b.LS:indist/ and is proved
3253: similarly using the preceding results:
3254: For example, instead of the use of delayed simple random walk by \ref
3255: b.LS:indist/, we use the network random walk in \ref r.choice/. 
3256: This is a reversible random walk corresponding to edge weights $(x, y)
3257: \mapsto 1/[(\deg x)(\deg y)]$.
3258: It is transient by Propositions \briefref p.trans/ and \briefref
3259: p.transtrees/, combined with Rayleigh's monotonicity principle.
3260: 
3261: \procl t.cerg \procname{Indistinguishable Clusters}
3262: If\/ $\P$ is an insertion-tolerant percolation on a
3263: unimodular random network, then
3264: $\P$ has indistinguishable infinite clusters.
3265: \endprocl
3266: 
3267: Among the several consequences of this result is the following extension of
3268: Theorem 4.1 of \ref b.LS:indist/, proved similarly.
3269: 
3270: \procl t.tau \procname{Uniqueness and Long-Range Order}
3271: \genassumptions, $\rtd$.
3272: If\/ $\P$ is extremal and there is more than one infinite cluster $\P$-a.s.,
3273: then $\rtd$-a.s.,
3274: $$
3275: \inf \big \{ \P[\hbox{there is an open path from $x$ to } y \mid \gh]
3276: \st x, y \in \vertex(\gh) \big \} = 0
3277: \,.
3278: $$
3279: \endprocl
3280: 
3281: The following extends Theorem 6.12 of \ref b.LS:indist/ and is proved
3282: similarly.
3283: 
3284: \procl t.product \procname{Uniqueness in Products}
3285: Suppose that $\rtd$, $\rtd_1$, and $\rtd_2$ are extremal unimodular probability
3286: measures on $\GG_*$, with $\rtd$ supported on infinite graphs and $\rtd_1$
3287: a percolation on $\rtd_2$. Then $\pu(\rtd \itimes \rtd_1) \ge
3288: \pu(\rtd \itimes \rtd_2)$. In particular, $\pu(\rtd) \ge
3289: \pu(\rtd \itimes \rtd_2)$.
3290: \endprocl
3291: 
3292: More results on percolation will be presented in \ref s.amen/.
3293: 
3294: 
3295: 
3296: \bsection{Spanning Forests}{s.span}
3297: 
3298: An interesting type of percolation other than Bernoulli is given by certain
3299: random forests.
3300: There are two classes of such random forests that have been widely studied,
3301: the uniform ones and the minimal ones.
3302: 
3303: We first discuss the uniform case.
3304: Given a finite connected graph, $\gh$, let $\ust(\gh)$ denote the uniform
3305: measure on spanning trees on $\gh$.
3306: \ref b.Pemantle:ust/
3307: proved a conjecture of Lyons, namely,
3308: that if an infinite connected graph $\gh$ is exhausted by a sequence of finite
3309: connected subgraphs $\gh_n$, then the weak limit of
3310: $\Seq{\ust(\gh_n)}$ exists.
3311: However, it may happen that the limit measure is not supported on trees,
3312: but on forests.
3313: This limit measure is now called the {\bf free (uniform) spanning forest} on
3314: $\gh$, denoted $\fsf$ or $\FUSF$.
3315: If $\gh$ is itself a tree, then this measure is
3316: trivial, namely, it is concentrated on $\{\gh\}$. Therefore, \ref
3317: b.Hag:umesft/
3318: introduced another limit that had been considered on $\Z^d$
3319: more implicitly by \ref b.Pemantle:ust/ and explicitly by
3320: \ref b.Hag:rcust/, namely, the weak limit of the uniform
3321: spanning tree measures on $\gh_n^*$, where $\gh_n^*$ is the graph $\gh_n$ with
3322: its boundary identified (``wired")
3323: to a single vertex. As \ref b.Pemantle:ust/ showed,
3324: this limit also always exists
3325: on any graph and is now called the {\bf wired (uniform) spanning forest},
3326: denoted $\wsf$ or $\WUSF$.
3327: It is clear that both $\fsf$ and $\wsf$ are concentrated on the set of
3328: spanning forests\ftnote{*}{By a
3329: ``spanning forest'', we mean a subgraph without cycles that contains every
3330: vertex.} of $\gh$ that are {\bf essential}, meaning that all their
3331: trees are infinite.
3332: %Furthermore, since the limits exist regardless of the exhaustion chosen, both
3333: %$\fsf_\gh$ and $\wsf_\gh$ are invariant under any automorphisms that $\gh$ may
3334: %have.
3335: Both $\fsf$ and $\wsf$ are important in their own right; see \ref
3336: b.Lyons:bird/ for a survey and \BLPSusf\ for a comprehensive
3337: treatment.
3338: 
3339: In all the above, one may work more generally with a weighted graph, where
3340: the graph has positive weights on its edges.
3341: In that case, $\ust$ stands for the measure such that the probability of a
3342: spanning tree is proportional to the product of the weights of its edges.
3343: The above theorems continue to hold and we use the same notation for the
3344: limiting measures.
3345: 
3346: 
3347: Most results known about the uniform spanning forest measures hold for
3348: general graphs. Some, however, require extra hypotheses such as
3349: transitivity and unimodularity.
3350: We extend some of these latter results here.
3351: 
3352: Given $\rtd$, taking the wired uniform spanning forest on each graph
3353: gives a percolation that we denote $\WUSF(\rtd)$.
3354: Our first result shows, among other things,
3355: that the kind of limit considered in this paper,
3356: i.e., random weak convergence, gives another natural way to define $\wsf$.
3357: It might be quite useful to have an explicit description of measures on
3358: finite graphs whose random weak limit is the {\it free\/} spanning forest.
3359: 
3360: \procl p.ust-limit \procname{UST Limits}
3361: If $\rtd$ is a unimodular probability measure on infinite networks in
3362: $\GG_*$, then $\expdeg\big(\WUSF(\rtd)\big) = 2$.
3363: If $\gh_n$ are finite connected networks whose random weak limit is $\rtd$,
3364: then $\UST(\gh_n) \cd \WUSF(\rtd)$.
3365: More generally, if $\rtd_n$ are unimodular probability measures on $\GG_*$
3366: with $\rtd_n \cd \rtd$, then $\WUSF(\rtd_n) \cd \WUSF(\rtd)$.
3367: \endprocl
3368: 
3369: \proof
3370: Suppose first that $\rtd$ is concentrated on recurrent networks.
3371: Then Wilson's algorithm gives both $\UST(\gh_n)$ and $\WUSF(\rtd)$.
3372: Hence it is clear that $\UST(\gh_n) \cd \WUSF(\rtd)$.
3373: Since $\expdeg(\UST(\gh_n)) = 2 - 2/|\vertex(\gh_n)| \le 2$,
3374: this limit relation in combination with
3375: Fatou's lemma shows that $\expdeg(\WUSF(\rtd)) \le 2$,
3376: whence equality results from \ref t.deg-infinite/.
3377: %We could also have used, say, \ref p.treespeed/.
3378: 
3379: Suppose next that $\rtd$ is concentrated on transient networks.
3380: Then the proof of Theorem 6.5 of \BLPSusf\ gives the same result.
3381: 
3382: Finally, if $\rtd$ is concentrated on neither recurrent nor transient
3383: networks, then we may write $\rtd$ as a mixture of two unimodular measures
3384: that are concentrated on recurrent or on transient networks and apply the
3385: preceding.
3386: 
3387: This proves the first sentence. The second sentence is a special case of
3388: the third, so we prove the third.
3389: 
3390: \comment{
3391: Suppose first that
3392: $\gh_n$ are finite connected networks whose random weak limit is $\rtd$.
3393: Given a positive integer $R$, let $\UST_R(\rtd)$ be the uniform spanning
3394: tree on the wired ball of radius $R$ about the root, and likewise for
3395: $\UST_R(\gh_n)$.
3396: By definition, we have $\UST_R(\rtd) \cd \WUSF(\rtd)$ as $R \to\infty$.
3397: Clearly, $\UST_R(\gh_n) \cd \UST_R(\rtd)$ as $n \to\infty$.
3398: Furthermore, the intersection of $\UST(\gh_n)$ with the ball of radius $R$
3399: stochastically dominates $\UST_R(\gh_n)$ by the theorem of \ref b.FedMih/.
3400: Therefore, every weak limit point of $\Seq{\UST(\gh_n)}$ stochastically
3401: dominates $\UST_R(\rtd)$ and therefore also $\WUSF(\rtd)$.
3402: On the other hand, from \ref p.limittree/,
3403: every weak limit point of $\Seq{\UST(\gh_n)}$
3404: has expected degree at most 2, while by \ref t.deg-infinite/, we have
3405: $\expdeg\big(\WUSF(\rtd)\big) \ge 2$.
3406: It follows that $\UST(\gh_n) \cd \WUSF(\rtd)$ as $n \to\infty$ and that
3407: $\expdeg\big(\WUSF(\rtd)\big) = 2$.
3408: 
3409: By \ref t.TII/, it follows that
3410: $\expdeg\big(\WUSF(\rtd)\big) = 2$ for all $\rtd \in \uni$.
3411: }%
3412: %
3413: %We may now establish the last assertion of the proposition.
3414: If $\rtd_n \cd \rtd$, then
3415: $\UST_R(\rtd_n) \cd \UST_R(\rtd)$ as $n \to\infty$ and the
3416: intersection of $\WUSF(\rtd_n)$ with the ball of radius $R$
3417: stochastically dominates $\UST_R(\rtd_n)$, so that every weak limit point of
3418: $\Seq{\WUSF(\rtd_n)}$ stochastically dominates $\WUSF(\rtd)$.
3419: However, since they all have expected degree 2 by what we have just shown,
3420: they are equal.
3421: \Qed
3422: 
3423: We next show that the trees of the $\wsf$ have only one end a.s.
3424: The first theorem of this type was proved by \ref b.Pem:ust/.
3425: His result was completed and extended in
3426: Theorem 10.1 of \BLPSusf, which
3427: dealt with the transitive unimodular case.
3428: The minor modifications needed for the quasi-transitive unimodular case
3429: were explained by \ref b.Lyons:est/.
3430: Another extension is given by \ref b.LMS:wsf/, who showed that for graphs with
3431: a ``reasonable" isoperimetric profile, each tree has only one end
3432: $\wsf$-a.s.
3433: 
3434: \procl t.1endUSF \procname{One End}
3435: If $\rtd$ is a unimodular probability measure on $\GG_*$ that is
3436: concentrated on transient networks with bounded degree,
3437: then $\WUSF(\rtd)$-a.s., each tree has exactly one end.
3438: \endprocl
3439: 
3440: \proof
3441: The proof is essentially the same as in \BLPSusf, with the following
3442: modifications.
3443: In the proof of Theorem 10.3 of \BLPSusf, which is the case where there is
3444: only one tree a.s., we replace $x$ and $y$ there by $\bp$ and $X_n$, where
3445: $\Seq{X_n}$ is the simple random walk starting from the root; to use
3446: stationarity, we bias the underlying network by the degree of the root.
3447: This gives a measure equivalent to $\rtd$, so that almost sure conclusions for
3448: it hold for $\rtd$ as well.
3449: The stationarity and reversibility give that the probability that the
3450: random walk from $\bp$ ever visits $X_n$ is equal to the probability that a
3451: random walk from $X_n$ ever visits $\bp$.
3452: By transience, this tends to 0 as $n \to\infty$, which allows the proof of
3453: \BLPSusf\ to go through.
3454: [Here, we needed finite
3455: expected degree to talk about probability since we used the equivalent
3456: probability measure of biasing by the degree.]
3457: 
3458: In the proof of Theorem 10.4 of \BLPSusf, the case when there is more than
3459: one tree a.s.,
3460: we need the degrees to be bounded for the displayed equality on p.~36 of
3461: \BLPSusf\ to hold up to a constant factor.
3462: \Qed
3463: 
3464: For our proof, we had to assume transience;
3465: there is presumably an extension to the recurrent case, which would say
3466: that the number of ends $\WUSF(\rtd)$-a.s.\ is the same as the number of
3467: ends $\rtd$-a.s.\ when $\rtd$ is concentrated on recurrent networks.
3468: Also, presumably the assumption that the degrees are bounded is not needed.
3469: In any case, our result here goes beyond what has been done before and
3470: gives a partial answer to Question 15.4 of \BLPSusf; removing the
3471: assumption of bounded degrees would completely answer that question.
3472: It also applies, e.g., to transient clusters of Bernoulli percolation; see
3473: \ref b.GKZ/ and \ref b.BLS:pert/ for sufficient conditions for transience.
3474: 
3475: We now prove analogous results for the other model of spanning trees, the
3476: minimal ones.
3477: Given a finite connected graph, $\gh$, and independent uniform $[0, 1]$
3478: random variables on its edges, the spanning tree that minimizes the sum of
3479: the labels of its edges has a distribution denoted $\MST(\gh)$, the {\bf
3480: minimal spanning tree} measure on $\gh$.
3481: If $\gh$ is infinite, there are two analogous measures, as in the uniform
3482: case. They can be defined by weak limits, but also directly (and by
3483: pointwise limits).
3484: Namely, given independent uniform $[0, 1]$ edge labels, remove all edges
3485: whose label is the largest in some cycle containing that edge.
3486: The remaining edges form the {\bf free minimal spanning forest},
3487: $\FMSF$.
3488: If one also removes the edges $e$ both of whose endpoints belong to infinite
3489: paths of edges that are all labeled smaller than $e$ is, then the resulting
3490: forest is called the {\bf wired minimal spanning forest}, $\WMSF$.
3491: 
3492: The following is analogous to \ref p.ust-limit/ above and is proved
3493: similarly to it and part of Theorem 3.12 of \LPSmsf, using \ref t.death/
3494: below. Parts of it were also proved by \ref b.AS:obj/.
3495: %It extends results of \LPSmsf, as well.
3496: %As noted in \ref b.AS:obj/, it holds without the usual assumption that the
3497: %edge labels are independent uniform random variables, but with the
3498: %assumption merely that the labels are distinct.
3499: 
3500: \procl p.mst-limit \procname{MST Limits}
3501: If $\rtd$ is a unimodular probability measure on infinite networks in
3502: $\GG_*$, then $\expdeg\big(\WMSF(\rtd)\big) = 2$.
3503: If $\gh_n$ are finite connected networks whose random weak limit is $\rtd$,
3504: then $\MST(\gh_n) \cd \WMSF(\rtd)$.
3505: More generally, if $\rtd_n$ are unimodular probability measures on $\GG_*$
3506: with $\rtd_n \cd \rtd$, then $\WMSF(\rtd_n) \cd \WMSF(\rtd)$.
3507: \endprocl
3508: 
3509: The following extends a result of \LPSmsf, which in turn extends a result
3510: of \ref b.Alexander:MSF/, who proved this in fixed Euclidean lattices.
3511: Our proof follows slightly different lines.
3512: For information on when the hypothesis is satisfied, see \ref t.death/
3513: below.
3514: 
3515: \procl t.msf1end \procname{One End}
3516: If $\rtd$ is an extremal
3517: unimodular probability measure on infinite networks in
3518: $\GG_*$ and there is $\P_{\pc(\rtd)}$-a.s.\ no infinite cluster,
3519: then $\WMSF(\rtd)$-a.s., each tree has exactly one end.
3520: \endprocl
3521: 
3522: \proof
3523: By the first part of \ref p.mst-limit/ and by \ref t.deg2/, each tree has at
3524: most 2 ends $\WMSF(\rtd)$-a.s.
3525: Suppose that some tree has 2 ends with positive probability.
3526: A tree with precisely two ends has a {\bf trunk}, the unique bi-infinite
3527: simple path in the tree.
3528: By the definition of $\WMSF$, the labels on a trunk cannot have a maximum.
3529: By the Mass-Transport Principle, the limsup in one direction must equal the
3530: limsup in the other direction, since otherwise we could identify the one
3531: edge that has label larger than the average of the two limsups and is the
3532: last such edge in the direction from the larger limsup to the smaller
3533: limsup.
3534: Let $p$ be this common limsup.
3535: By the preceding, all the labels on the trunk are strictly less than $p$.
3536: The root belongs to the trunk with positive probability.
3537: Assume this happens.
3538: Then the root belongs to an infinite $p$-cluster (the one containing the
3539: trunk).
3540: Now invasion from the root will fill (the vertices of)
3541: this $p$-cluster and is part of the
3542: tree containing the root (see \LPSmsf), whence the tree contains (the
3543: vertices of) the entire $p$-cluster of the root.
3544: By \ref t.uniq/, the tree therefore also contains an infinite $p'$-cluster
3545: for every $p' \in (\pc(\rtd), p)$ if $p > \pc(\rtd)$.
3546: Let $x$ be a vertex in the tree that is in an infinite $p'$-cluster $C$ for
3547: $p' := (\pc(\rtd) + p)/2$.
3548: Now invasion from $x$ has a finite symmetric difference with
3549: invasion from $o$ (see \LPSmsf),
3550: invasion from $x$ does not leave $C$, and invasion from $\bp$ fills the
3551: trunk.
3552: It follows from the definition of $p$ that $p' \ge p$.
3553: That is, $p = \pc(\rtd)$.
3554: Therefore, there is $\P_{\pc(\rtd)}$-a.s.\ an infinite cluster.
3555: \Qed
3556: 
3557: 
3558: 
3559: \bsection{Amenability and Nonamenability}{s.amen}
3560: 
3561: Recall that a graph $\gh$ is (vertex) amenable
3562: iff there is a sequence of subsets $H_n\subset \vertex(\gh)$
3563: with
3564: $$
3565: \lim_{n \to\infty} {|\bdv H_n| \over |\verts(H_n)|} = 0\,,
3566: $$
3567: where $|\cbuldot|$ denotes cardinality.
3568: 
3569: Amenability, originally defined for groups, now appears in several areas of
3570: mathematics, including probability theory and ergodic theory.
3571: Its presence provides many tools one is used to from $\Z$ actions, yet its
3572: absence also provides a powerful threshold principle.
3573: There are many equivalent definitions of amenability.
3574: We choose one that is not standard, but is useful for our probabilistic
3575: purposes.
3576: We show that it is equivalent to other definitions.
3577: Then we shall illustrate its uses.
3578: 
3579: \procl d.amen
3580: Let $\projection : \marks \to \marks$ be the composition of a homeomorphism of
3581: $\marks$ with $\marks^2$ followed by the projection onto the first coordinate.
3582: If a rooted network $(\gh, \bp)$ is understood,
3583: then for a subset $\marks_0 \subseteq \marks$ and vertex $x$, the {\bf
3584: $\marks_0$-component of $x$} is the set of vertices that can be reached
3585: from $x$ by edges both of whose marks lie in $\marks_0$.
3586: Write $K(\marks_0)$ for the $\marks_0$-component of the root.
3587: For a probability measure $\rtd$ on rooted graphs, denote by $\fdom(\rtd)$ the
3588: class of percolations on $\rtd$ that have only finite components.
3589: That is, $\fdom(\rtd)$ consists of pairs $(\nu, \marks_0)$ such that $\nu$ is
3590: a unimodular probability measure on $\GG_*$, $\marks_0 \subseteq \marks$ is
3591: Borel, $\rtd = \nu \circ \projection^{-1}$, and
3592: $K(\marks_0)$ is finite $\nu$-a.s.
3593: (By \ref l.unmark/, all $\marks_0$-components are then finite $\nu$-a.s.)
3594: For $\marks_0 \subseteq \marks$ and $x \in \vertex(\gh)$, write
3595: $$
3596: \ct(x, \marks_0) := | \{y \in \vertex(\gh) \st (x, y) \in \edges(\gh),
3597: \hbox{ some edge mark of } (x, y) \hbox{ is }\notin \marks_0 \} |
3598: \,.
3599: $$
3600: For $K \subset \vertex(\gh)$, define
3601: $$
3602: \ct(K, \marks_0) :=
3603: \sum_{x \in K} \ct(x, \marks_0)
3604: \,.
3605: $$
3606: Define
3607: $$
3608: \isoe(\rtd)
3609: :=
3610: \inf \left\{ \int {\ct\big(K(\marks_0), \marks_0\big) \over |K(\marks_0)|}
3611: d\rtd'(\gh, \bp)
3612: \st (\rtd', \marks_0) \in \fdom(\rtd) \right\}
3613: \,.
3614: $$
3615: Call $\rtd$ {\bf amenable} if $\isoe(\rtd) = 0$.
3616: Define 
3617: $$
3618: \expdeg(\rtd', \marks_0)
3619: :=
3620: %\int |\{x \in \vertex(\gh) \st (\bp, x) \in \edges(\gh) \hbox{ and both
3621: %edge marks of $(\bp, x)$ lie in } \marks_0\}| \,d\rtd'(\gh, \bp)
3622: \int \big[\deg_\gh(\bp) - n(\bp, \marks_0)\big] \,d\rtd'(\gh, \bp)
3623: \,,
3624: $$
3625: the expected degree in the $\marks_0$-component of the root,
3626: and
3627: $$
3628: \alpha(\rtd)
3629: :=
3630: \sup \big\{ \expdeg(\rtd', \marks_0) \st (\rtd', \marks_0) \in \fdom(\rtd)
3631: \big\}
3632: \,.
3633: $$
3634: Of course, neither $\isoe(\rtd)$ nor $\alpha(\rtd)$ depends on the choice
3635: of homeomorphism in $\projection$.
3636: Furthermore, these quantities depend only on the probability measure on
3637: the underlying graphs of the networks, not on the marks.
3638: \endprocl
3639: 
3640: This definition of amenability is justified in three ways: It agrees with
3641: the usual definition of amenability for fixed unimodular quasi-transitive
3642: graphs by Theorems 5.1 and 5.3 of \BLPSgip; it agrees with the usual notion
3643: of amenability for equivalence relations by \ref t.many/ below; and it
3644: allows us to extend to non-amenable unimodular random rooted graphs many
3645: theorems that are known for non-amenable unimodular fixed graphs, as we
3646: shall see.
3647: 
3648: We say that a graph $\gh$ is {\bf anchored amenable}
3649: if there is a sequence of subsets $H_n\subset \vertex(\gh)$
3650: such that $\bigcap_n H_n \ne \emptyset$, each $H_n$ induces a connected
3651: subgraph of $\gh$, and
3652: $$
3653: \lim_{n \to\infty} {|\bdv H_n| \over |\verts(H_n)|} = 0\,.
3654: $$
3655: The relationship between amenability of $\rtd$ and amenability or anchored
3656: amenability of $\rtd$-a.e.\ graph is as follows.
3657: The first clearly implies the third, which implies the second, but the third
3658: does not imply the first.
3659: Indeed, take a 3-regular tree and
3660: randomly subdivide its edges by a number of vertices whose
3661: distribution does not have a finite
3662: exponential tail. \ref b.ChenPeres/ show that the result is anchored
3663: amenable a.s.
3664: However, there is an appropriate unimodular version if the subdividing
3665: distribution has finite mean (see Example 2.4.4 of \ref
3666: b.Kaim:harmonic/ or \ref x.stretched/ below),
3667: and it is non-amenable by \ref c.amentrees/ below.
3668: 
3669: In order to work with this definition, we shall need some easy facts.
3670: 
3671: \procl l.bdryo
3672: If $\rtd$ is a unimodular probability measure on $\GG_*$ and $(\rtd',
3673: \marks_0) \in \fdom(\rtd)$, then
3674: $$
3675: \int {\ct\big(K(\marks_0), \marks_0\big) \over |K(\marks_0)|} d\rtd'
3676: =
3677: \int \ct(\bp, \marks_0) \, d\rtd'
3678: \,.
3679: $$
3680: \endprocl
3681: 
3682: \proof
3683: Let $K_x$ be the $\marks_0$-component of $x$.
3684: Let each vertex $x$ send mass $\ct(y, \marks_0)/|K_x|$ to each $y \in K_x$.
3685: Then the left-hand side is the expected mass sent from the root and the
3686: right-hand side is the expected mass received by the root.
3687: \Qed
3688: 
3689: 
3690: %\procl l.finiteavg
3691: %If $\rtd$ is a unimodular probability measure on $\GG_*$ and $\rtd' \in
3692: %\fdom(\rtd)$, then
3693: %$$
3694: %\expdeg(\rtd)
3695: %=
3696: %\int {\sum_{x \in K(\bp)} \deg(x) \over |K(\bp)|} d\rtd'(\gh, \bp)
3697: %\,,
3698: %\label e.alpha-alt
3699: %$$
3700: %where $K(\bp)$ is the component of the root in $(\gh, \bp)$ and degree is
3701: %measured in $\rtd$. \msnote{Fix wording.}
3702: %\endprocl
3703: %
3704: %\proof
3705: %This is a consequence of the Mass-Transport Principle:
3706: %each vertex $x$ distributes mass $\deg(x)$ equally among the vertices in its
3707: %$\rtd'$-component $K$.
3708: %The expected total mass sent out from the root is $\expdeg(\rtd)$, while the
3709: %expected total mass received at the root is the right-hand side of
3710: %\ref e.alpha-alt/.
3711: %\Qed
3712: 
3713: 
3714: \procl p.alpha
3715: If $\rtd$ is a unimodular probability measure on $\GG_*$, then
3716: $$
3717: \isoe(\rtd) + \alpha(\rtd) = \expdeg(\rtd)
3718: \,.
3719: \label e.alpha
3720: $$
3721: Therefore, if\/ $\expdeg(\rtd) < \infty$, then $\rtd$ is amenable iff
3722: $\alpha(\rtd) = \expdeg(\rtd)$.
3723: \endprocl
3724: 
3725: \proof
3726: %In the definition of $\alpha(\rtd)$, replace $\expdeg(\rtd')$ with its equivalent
3727: %expression from \ref e.alpha-alt/ (applied to $\rtd'$).
3728: %Adding this to $\isoe(\rtd)$,
3729: %the result follows by another application of
3730: %\ref e.alpha-alt/ (applied to $\rtd$).
3731: This is obvious from \ref l.bdryo/.
3732: \Qed
3733: 
3734: 
3735: \procl l.disint
3736: Let $\rtd, \nu \in \uni$ with $\nu$ a percolation on $\rtd$, that is, there
3737: is some Borel $\psi : \marks \to \marks$ such that $\rtd = \nu \circ
3738: \psi^{-1}$.
3739: Let $\kappa$ be a regular conditional probability measure of $\rtd$
3740: with respect to the $\sigma$-field generated by $\psi$, i.e., a
3741: disintegration of $\nu$ with respect to $\psi$, with $\kappa_{(\gh, \bp)}$
3742: being the probability measure on the fiber over $(\gh, \bp)$.
3743: Let $h : \gtwo \to [0, 1]$ be Borel and symmetric: $h(G, x, y) = h(G, y,
3744: x)$.
3745: Define $k(G, x, y) := \int h(G, x, y) \,d\kappa_{(\gh, x)}$.
3746: Then there is a symmetric Borel $\lambda$ such that for $\rtd$-a.e.\ $(\gh,
3747: \bp)$ and all $x \in \vertex(\gh)$, we have $k(\gh, \bp, x) = \lambda(\gh,
3748: \bp, x)$.
3749: \endprocl
3750: 
3751: %This could also be proved by showing it on finite networks and using \ref
3752: %t.TII/, as suggested by B\'alint Virag.
3753: %It is false without unimodularity.
3754: 
3755: \proof
3756: It suffices to show that for all $f : \gtwo \to [0, 1]$, we have 
3757: $$
3758: \int \sum_{x \in \vertex(\gh)} k(\gh, \bp, x) f(\gh, \bp, x) \,d\rtd(\gh,
3759: \bp)
3760: =
3761: \int \sum_{x \in \vertex(\gh)} k(\gh, x, \bp) f(\gh, \bp, x) \,d\rtd(\gh,
3762: \bp)
3763: \,,
3764: $$
3765: since this shows symmetry of $k$ a.e.\ with respect to the left measure
3766: $\rtd_{\rm L}$.
3767: To see that this equation holds, observe that 
3768: $$\eqaln{
3769: \int \sum_{x \in \vertex(\gh)} k(\gh, \bp, x) f(\gh, \bp, x) &\,d\rtd(\gh,
3770: \bp)
3771: =
3772: \int \int \sum_{x \in \vertex(\gh)} h(\gh, \bp, x) f(\gh, \bp, x)
3773: \,d\kappa_{(\gh, \bp)} \,d\rtd(\gh, \bp)
3774: \cr&=
3775: \int \int \sum_{x \in \vertex(\gh)} h(\gh, x, \bp) f(\gh, x, \bp)
3776: \,d\kappa_{(\gh, \bp)} \,d\rtd(\gh, \bp)
3777: \cr&\hskip1in\hbox{[by the Mass-Transport Principle for $\nu$]}
3778: \cr&=
3779: \int \int \sum_{x \in \vertex(\gh)} h(\gh, \bp, x) f(\gh, \bp, x)
3780: \,d\kappa_{(\gh, x)} \,d\rtd(\gh, \bp)
3781: \cr&\hskip1in\hbox{[by the Mass-Transport Principle for $\rtd$]}
3782: \cr&=
3783: \int \int \sum_{x \in \vertex(\gh)} h(\gh, x, \bp) f(\gh, \bp, x)
3784: \,d\kappa_{(\gh, x)} \,d\rtd(\gh, \bp)
3785: \cr&\hskip1in\hbox{[by symmetry of $h$]}
3786: \cr&=
3787: \int \sum_{x \in \vertex(\gh)} k(\gh, x, \bp) f(\gh, \bp, x) \,d\rtd(\gh,
3788: \bp)
3789: \,.
3790: \Qed
3791: }$$
3792: 
3793: 
3794: We now prove some properties that are equivalent to amenability.
3795: Most of these are standard in the context of equivalence relations.
3796: With appropriate modifications, these equivalences hold with a weakening of
3797: the assumption of unimodularity.
3798: They are essentially due to \ref b.CFW/ and \ref b.Kaim:amen/, although
3799: (ii) seems to be new.
3800: 
3801: \procl t.many \procname{Amenability Criteria}
3802: Let $\rtd$ be a unimodular probability measure on $\GG_*$ with $\expdeg(\rtd)
3803: < \infty$.
3804: The following are equivalent:
3805: \beginitems
3806: \itemrm{(i)} $\rtd$ is amenable;
3807: \itemrm{(ii)} there is
3808: a sequence of Borel functions $\lambda_n : \gtwo \to [0, 1]$
3809: such that for all $(\gh, x, y) \in \gtwo$ and all $n$,
3810: we have
3811: $$
3812: \lambda_n(\gh, x, y) =
3813: \lambda_n(\gh, y, x)
3814: \label e.symm
3815: $$
3816: and for $\rtd$-a.e.\ $(\gh, \bp)$, we have
3817: $$
3818: \sum_{x \in \vertex(\gh)} \lambda_n(\gh, \bp, x) = 1
3819: \label e.prmsrae
3820: $$
3821: and
3822: $$
3823: \lim_{n \to\infty}
3824: \sum_{x \in \vertex(\gh)}\sum_{y \sim x}
3825: \left|\lambda_n(\gh, \bp, x) -
3826: \lambda_n(\gh, \bp, y)\right|
3827: = 0
3828: \,;
3829: \label e.asymsame
3830: $$
3831: \itemrm{(iii)} there is
3832: a sequence of Borel functions $\lambda_n : \gtwo \to [0, 1]$
3833: such that for $\rtd$-a.e.\ $(\gh, \bp)$,
3834: $$
3835: \sum_{x \in \vertex(\gh)} \lambda_n(\gh, \bp, x) = 1
3836: $$
3837: and
3838: $$
3839: \lim_{n \to\infty}
3840: \sum_{y \sim \bp} \sum_{x \in \vertex(\gh)}
3841: \left|\lambda_n(\gh, \bp, x) -
3842: \lambda_n(\gh, y, x)\right|
3843: = 0
3844: \,;
3845: $$
3846: \itemrm{(iv)}
3847: $\rtd$ is hyperfinite, meaning that there
3848: is a unimodular measure $\nu$ on $\GG_*$, an increasing sequence of 
3849: Borel subsets $\marks_n \subseteq \marks$, and a Borel
3850: function $\psi: \marks \to \marks$ such that if $G$ denotes a
3851: network with law $\nu$ and $G_n$ the subnetwork consisting of those edges
3852: both of whose edge marks lie in $\marks_n$, then $\psi(G)$ has law $\rtd$,
3853: all components of $G_n$ are finite, and $\bigcup_n \marks_n = \marks$.
3854: \enditems
3855: \endprocl
3856: 
3857: \proof
3858: Assume from now on that $\rtd$ is carried by networks with distinct marks.
3859: We shall use the following construction.
3860: Suppose that $(\nu, \marks_0) \in \fdom(\rtd)$.
3861: Let $\kappa$ be a regular conditional probability measure of
3862: $\rtd$ with respect to the $\sigma$-field generated by $\projection$.
3863: By \ref l.disint/, there is a Borel symmetric 
3864: $\lambda : \gtwo \to [0, 1]$ such that
3865: for $\rtd$-a.e.\ $(\gh, \bp)$ and $x \in \vertex(\gh)$, we have
3866: $$
3867: \lambda(\gh, \bp, x)
3868: =
3869: \int \II{x \in K(\marks_0)}/|K(\marks_0)|
3870: \,d\kappa_{(\gh, \bp)}
3871: \,.
3872: \label e.connect
3873: $$
3874: Clearly,
3875: $$
3876: \sum_{x \in \vertex(\gh)} \lambda(\gh, \bp, x) = 1
3877: \label e.prmsr
3878: $$
3879: for $\rtd$-a.e.\ $(\gh, \bp)$.
3880: For $\rtd$-a.e.\ $(\gh, \bp)$, we have
3881: $$\eqaln{
3882: \sum_{x \in \vertex(\gh)} \sum_{y \sim x}
3883: |\lambda(\gh, \bp, x) &- \lambda(\gh, \bp, y)|
3884: \cr&\le
3885: \sum_{x \in \vertex(\gh)} \sum_{y \sim x}
3886: \int |\II{x \in K(\marks_0)} -
3887: \II{y \in K(\marks_0)}|/|K(\marks_0)|
3888: \,d\kappa_{(\gh, \bp)}
3889: \cr&\le
3890: \int {2\ct\big(K(\marks_0), \marks_0\big) \over |K(\marks_0)|}
3891: \,d\kappa_{(\gh, \bp)}
3892: \,.
3893: \label e.isobd
3894: }$$
3895: 
3896: Now if (i) holds, then we may choose $(\rtd_n, \marks_n) \in \fdom(\rtd)$
3897: such that
3898: $$
3899: \int
3900: \sum_{n}
3901: {\ct\big(K(\marks_n), \marks_n\big) \over |K(\marks_n)|}
3902: \,d\rtd_{n} < \infty
3903: \,.
3904: $$
3905: Let $\kappa^{(n)}$ and
3906: $\lambda_n$ be as above (but for $\rtd_n$).
3907: Then by \ref e.isobd/, we have
3908: $$\eqaln{
3909: \sum_{n} \int
3910: \sum_{x \in \vertex(\gh)}\sum_{y \sim x}
3911: \big|\lambda_n(\gh, \bp, x) &-
3912: \lambda_n(\gh, \bp, y)\big|
3913: \,d\rtd(\gh, \bp)
3914: \cr&\le
3915: \sum_{n} \int
3916: \int {2\ct\big(K(\marks_n), \marks_n\big) \over |K(\marks_n)|}
3917: \,d\kappa^{(n)}_{(\gh, \bp)}
3918: \,d\rtd(\gh, \bp)
3919: \cr&=
3920: \sum_{n}
3921: \int {2\ct\big(K(\marks_n), \marks_n\big) \over |K(\marks_n)|}
3922: \,d\rtd_{n} < \infty
3923: \,,
3924: }$$
3925: which shows that (ii) holds.
3926: 
3927: Next, suppose that (ii) holds.
3928: Then by the bounded convergence theorem, we have
3929: $$
3930: \lim_{n \to\infty} \int
3931: \sum_{x \in \vertex(\gh)}\sum_{y \sim x}
3932: \left|\lambda_n(\gh, \bp, x) -
3933: \lambda_n(\gh, \bp, y)\right|
3934: \,d\rtd(\gh, \bp)
3935: = 0
3936: \,.
3937: $$
3938: On the other hand, the Mass-Transport Principle and \ref e.symm/
3939: show that this integral is the same as
3940: $$
3941: \int
3942: \sum_{y \sim \bp} \sum_{x \in \vertex(\gh)}
3943: \left|\lambda_n(\gh, \bp, x) -
3944: \lambda_n(\gh, y, x)\right|
3945: \,d\rtd(\gh, \bp)
3946: \,.
3947: $$
3948: Therefore, by taking a subsequence if necessary, we achieve (iii).
3949: 
3950: Next, suppose that (iii) holds.
3951: Then we may define a sequence similar to $\lambda_n$ on
3952: the corresponding equivalence relation (see \ref x.equivalence/),
3953: which implies that the equivalence
3954: relation is amenable by \ref b.Kaim:amen/, and hence hyperfinite by
3955: a theorem of \ref b.CFW/. (Another proof of the latter theorem was sketched
3956: by \ref b.Kaim:amen/, with more details
3957: given by \ref b.KechrisMiller/).
3958: Translating the definition of hyperfinite equivalence relation to rooted
3959: networks gives (iv).
3960: 
3961: Finally, that (iv) implies (i) is an immediate consequence of Lebesgue's
3962: Dominated Convergence Theorem, our assumption that $\expdeg(\rtd) <
3963: \infty$, and \ref l.bdryo/.
3964: \Qed
3965: 
3966: Now we show how to produce unimodular networks from non-unimodular ones on
3967: amenable measures, just as we can produce invariant measures from
3968: non-invariant ones on amenable groups.
3969: We illustrate in the context of couplings.
3970: 
3971: \procl p.amen-average \procname{Coupling From Amenability}
3972: Let $\rln \subseteq \marks \times \marks$ be a closed set.
3973: If $\rtd_1, \rtd_2 \in \uni$ are amenable and $\rtd_1$ is
3974: $\rln$-related to $\rtd_2$, then there is a unimodular $\rln$-coupling of
3975: $\rtd_1$ to $\rtd_2$.
3976: \endprocl
3977: 
3978: \proof
3979: Let $\nu$ be an $\rln$-coupling of $\rtd_1$ to $\rtd_2$.
3980: Let $\lambda_n$ be as in \ref t.many/(ii) for $(\rtd_1 + \rtd_2)/2$.
3981: Define the measures $\nu_n$ by
3982: $$
3983: \nu_n(\ev B)
3984: :=
3985: \int
3986: \sum_{x \in \vertex(\gh)} \lambda_n(\gh, \bp, x) \II{(\gh, x) \in \ev B}
3987: \,d\nu(\gh, \bp)
3988: $$
3989: for Borel $\ev B \subseteq \GG_*$ (with mark space $\marks \times \marks$).
3990: Then $\nu_n$ is a probability measure by \ref e.prmsrae/.
3991: Since $\nu$ is carried by networks all of whose marks are in $\rln$, so is
3992: $\nu_n$.
3993: If $\ev B$ is an event that
3994: specifies only the first coordinates of the marks, i.e., $\ev B = \ev
3995: B_1 \times 2^\marks$ for some $\ev B_1$, then
3996: $$\eqaln{
3997: \nu_n(\ev B)
3998: &=
3999: \int
4000: \sum_{x \in \vertex(\gh)} \lambda_n(\gh, \bp, x) \II{(\gh, x) \in \ev B}
4001: \,d\nu(\gh, \bp)
4002: \cr&=
4003: \int
4004: \sum_{x \in \vertex(\gh)} \lambda_n(\gh, \bp, x) \II{(\gh, x) \in \ev B_1}
4005: \,d\mu_1(\gh, \bp)
4006: \cr&=
4007: \int
4008: \sum_{x \in \vertex(\gh)} \lambda_n(\gh, x, \bp) \II{(\gh, \bp) \in \ev B_1}
4009: \,d\mu_1(\gh, \bp)
4010: \cr&=
4011: \int
4012: \II{(\gh, \bp) \in \ev B_1}
4013: \,d\mu_1(\gh, \bp)
4014: \cr&=
4015: \mu_1(\ev B)
4016: }$$
4017: by the Mass-Transport Principle, \ref e.symm/, and \ref e.prmsrae/.
4018: Likewise, if $\ev B$ is an event that
4019: specifies only the second coordinates of the marks, then $\nu_n(\ev B) =
4020: \mu_2(\ev B)$.
4021: Thus, $\nu_n$ is an
4022: $\rln$-coupling of $\rtd_1$ to $\rtd_2$.
4023: 
4024: Now by definition of $\nu_n$, we have
4025: $$
4026: \int f(\gh, \bp)\,d\nu_n(\gh, \bp)
4027: =
4028: \int
4029: \sum_{x \in \vertex(\gh)} \lambda_n(\gh, \bp, x) f(\gh, x)
4030: \,d\nu(\gh, \bp)
4031: $$
4032: for every Borel $f : \GG_* \to [0, \infty]$.
4033: Therefore, for every Borel $h : \gtwo \to [0, 1]$ with $h(\gh, x, y) = 0$
4034: unless $x \sim y$, we have
4035: $$\eqaln{
4036: \int \sum_{y \in \vertex(\gh)} h(\gh, \bp, y) \,d\nu_n(\gh, \bp)
4037: &=
4038: \int \sum_{x \in \vertex(\gh)} \lambda_n(\gh, \bp, x)
4039: \sum_{y \in \vertex(\gh)} h(\gh, x, y) \,d\nu_n(\gh, \bp)
4040: \cr&=
4041: \int \sum_{y \in \vertex(\gh)}
4042: \sum_{x \in \vertex(\gh)} 
4043: h(\gh, x, y) 
4044: \lambda_n(\gh, \bp, x)
4045: \,d\nu(\gh, \bp)
4046: }$$
4047: and
4048: $$\eqaln{
4049: \int \sum_{y \in \vertex(\gh)} h(\gh, y, \bp) \,d\nu_n(\gh, \bp)
4050: &=
4051: \int \sum_{x \in \vertex(\gh)} \lambda_n(\gh, \bp, x)
4052: \sum_{y \in \vertex(\gh)} h(\gh, y, x) \,d\nu_n(\gh, \bp)
4053: \cr&=
4054: \int \sum_{y \in \vertex(\gh)}
4055: \sum_{x \in \vertex(\gh)} 
4056: h(\gh, x, y) 
4057: \lambda_n(\gh, \bp, y)
4058: \,d\nu(\gh, \bp)
4059: \,,
4060: }$$
4061: where, in the last step, we have interchanged $x$ and $y$.
4062: Therefore,
4063: $$\eqaln{
4064: \bigg| 
4065: \int \sum_{y \in \vertex(\gh)} h(\gh, \bp, y) \,d\nu_n(\gh, \bp)
4066: &- 
4067: \int \sum_{y \in \vertex(\gh)} h(\gh, y, \bp) \,d\nu_n(\gh, \bp)
4068: \bigg|
4069: \cr&\le
4070: \int \sum_{y \in \vertex(\gh)}
4071: \sum_{x \sim y}
4072: \left|
4073: \lambda_n(\gh, \bp, x)
4074: -
4075: \lambda_n(\gh, \bp, y)
4076: \right|
4077: \,d\nu(\gh, \bp)
4078: \cr&=
4079: \int \sum_{y \in \vertex(\gh)}
4080: \sum_{x \sim y}
4081: \left|
4082: \lambda_n(\gh, \bp, x)
4083: -
4084: \lambda_n(\gh, \bp, y)
4085: \right|
4086: \,d\rtd_1(\gh, \bp)
4087: \,,
4088: }$$
4089: which tends to 0 by \ref e.asymsame/.
4090: Thus, any limit point of $\nu_n$ is involution invariant and, since $\rln$ is
4091: closed, is an $\rln$-coupling of $\rtd_1$ to $\rtd_2$.
4092: \Qed
4093: 
4094: \comment{
4095: \proof
4096: Let $\nu$ be an $\rln$-coupling of $\rtd_1$ to $\rtd_2$.
4097: Let $\lambda_n$ be as in \ref t.many/ for, say, $(\mu_1+\mu_2)/2$.
4098: Write
4099: $$
4100: k(\gh, \bp) :=
4101: \sum_{x \in \vertex(\gh)} \lambda_n(\gh, x, \bp)
4102: \,.
4103: $$
4104: Define the measures $\nu_n$ by
4105: $$
4106: \nu_n(\ev B)
4107: :=
4108: \int
4109: \sum_{x \in \vertex(\gh)} \lambda_n(\gh, \bp, x) \II{(\gh, x) \in \ev B}
4110: k(\gh, x)^{-1} \,d\nu(\gh, \bp)
4111: $$
4112: for Borel $\ev B \subseteq \GG_*$ (with mark space $\marks \times \marks$).
4113: Then $\nu_n$ is a probability measure since
4114: $$\eqaln{
4115: \int
4116: \sum_{x \in \vertex(\gh)} \lambda_n(\gh, \bp, x)
4117: k(\gh, x)^{-1} \,d\nu(\gh, \bp)
4118: &=
4119: \int
4120: \sum_{x \in \vertex(\gh)} \lambda_n(\gh, x, \bp)
4121: k(\gh, \bp)^{-1} \,d\nu(\gh, \bp)
4122: \cr&=
4123: \int d\nu(\gh, \bp)
4124: = 1
4125: }$$
4126: by the Mass-Transport Principle.
4127: %Note that
4128: %$$
4129: %\int
4130: %\sum_{x \in \vertex(\gh)} \lambda_n(\gh, x, \bp) \,d\mu_1(\gh, \bp)
4131: %=
4132: %\int
4133: %\sum_{x \in \vertex(\gh)} \lambda_n(\gh, \bp, x) \,d\mu_1(\gh, \bp)
4134: %=
4135: %1
4136: %$$
4137: %by the Mass-Transport Principle.
4138: Since $\nu$ is carried by networks all of whose marks are in $\rln$, so is
4139: $\nu_n$.
4140: If $\ev B$ is an event that
4141: specifies only the first coordinates of the marks, i.e., $\ev B = \ev
4142: B_1 \times 2^\marks$ for some $\ev B_1$, then
4143: $$\eqaln{
4144: \nu_n(\ev B)
4145: &=
4146: \int
4147: \sum_{x \in \vertex(\gh)} \lambda_n(\gh, \bp, x) \II{(\gh, x) \in \ev B}
4148: k(\gh, x)^{-1} \,d\kappa(\gh, \bp)
4149: \cr&=
4150: \int
4151: \sum_{x \in \vertex(\gh)} \lambda_n(\gh, \bp, x) \II{(\gh, x) \in \ev B_1}
4152: k(\gh, x)^{-1} \,d\mu_1(\gh, \bp)
4153: \cr&=
4154: \int
4155: \sum_{x \in \vertex(\gh)} \lambda_n(\gh, x, \bp) \II{(\gh, \bp) \in \ev B_1}
4156: k(\gh, \bp)^{-1} \,d\mu_1(\gh, \bp)
4157: \cr&=
4158: \int
4159: \II{(\gh, \bp) \in \ev B_1}
4160: \,d\mu_1(\gh, \bp)
4161: \cr&=
4162: \mu_1(\ev B)
4163: }$$
4164: by the Mass-Transport Principle.
4165: Likewise, if $\ev B$ is an event that
4166: specifies only the second coordinates of the marks, then $\nu_n(\ev B) =
4167: \mu_2(\ev B)$.
4168: Thus, $\nu_n$ is an
4169: $\rln$-coupling of $\rtd_1$ to $\rtd_2$.
4170: Now by definition of $\nu_n$, we have
4171: $$
4172: \int f(\gh, \bp)\,d\nu_n(\gh, \bp)
4173: =
4174: \int
4175: \sum_{x \in \vertex(\gh)} \lambda_n(\gh, \bp, x) f(\gh, x)
4176: k(\gh, x)^{-1} \,d\nu(\gh, \bp)
4177: $$
4178: for every Borel $f : \GG_* \to [0, \infty]$.
4179: Therefore,
4180: \msnote{The below does not calculate in sufficient generality.}
4181: $$\eqaln{
4182: \bigg| \int \sum_{y \sim \bp} \II{(\gh, \bp) \in \ev B} \,d\nu_n(\gh, \bp)
4183: &- \int \sum_{y \sim \bp} \II{(\gh, y) \in \ev B} \,d\nu_n(\gh, \bp)\bigg|
4184: \cr&\quad=
4185: \bigg|
4186: \int \sum_{x \in \vertex(\gh)} \lambda_n(\gh, \bp, x)
4187: \sum_{y \sim x} \II{(\gh, x) \in \ev B} k(\gh, x)^{-1} \,d\nu(\gh, \bp)
4188: \cr&\qquad-
4189: \sum_{x \in \vertex(\gh)} \lambda_n(\gh, \bp, x)
4190: \sum_{y \sim x} \II{(\gh, y) \in \ev B} k(\gh, x)^{-1} \,d\nu(\gh, \bp)
4191: \bigg|
4192: \cr&\quad\le
4193: ...
4194: }$$
4195: \msnote{Now apply CBS and use that the square of a number in
4196: $[0, 1]$ is at most that number.}
4197: Thus, any limit point of $\nu_n$ is unimodular and, since $\rln$ is
4198: closed, is an $\rln$-coupling of $\rtd_1$ to $\rtd_2$.
4199: \Qed
4200: }%
4201: 
4202: 
4203: 
4204: \procl p.recurrent \procname{Recurrence Implies Amenability}
4205: If $\rtd \in \uni$ and simple random walk is $\rtd$-a.s.\ recurrent,
4206: then $\rtd$ is amenable.
4207: \endprocl
4208: 
4209: \proof
4210: Consider the ``lazy" simple random walk that moves nowhere with probability
4211: 1/2 and otherwise moves to a random neighbor, like simple random walk.
4212: It is recurrent by hypothesis and aperiodic by construction.
4213: Let $\lambda_n(\gh, \bp, x)$ be the probability that lazy simple random
4214: walk on $\gh$ starting from $\bp$ will be at $x$ at time $n$.
4215: By \ref b.Orey/ and recurrence, the functions $\lambda_n$
4216: satisfy property (iii) of \ref t.many/, whence $\rtd$ is amenable.
4217: \Qed
4218: 
4219: 
4220: The following is proved similarly to Remark 6.2 of \BLPSusf.
4221: 
4222: \procl p.amenforest \procname{Forests in Amenable Networks}
4223: If $\rtd \in \uni$ is amenable and $\marks_0 \subseteq \marks$ is such that
4224: the $\marks_0$-open subgraph $\fo$ of $(\gh, \bp)$ is a forest $\rtd$-a.s.,
4225: then the expected degree of $\bp$ in $\fo$ is at most 2.
4226: \endprocl
4227: 
4228: The following extends Theorem 5.3 of \BLPSgip.
4229: In the following, we say that $\P$ is a {\bf percolation} on $\rtd$ that
4230: {\bf gives} subgraphs $\A$ a.s.\ if there is a Borel
4231: function $\psi: \marks \to \marks$ such that $\rtd = \P \circ \psi^{-1}$
4232: and there is a Borel subset $\marks_0 \subseteq \marks$
4233: such that if $G(\marks_0)$ denotes the $\marks_0$-open subnetwork of $G$,
4234: then $\psi\big(G(\marks_0)\big) \in \A$ for $\P$-a.e.\ $G$.
4235: 
4236: \procl t.treechar
4237: Let $\rtd \in \uni$ with $\expdeg(\rtd) < \infty$. The following are equivalent:
4238: \beginitems
4239: \itemrm{(i)} $\rtd$ is amenable;
4240: \itemrm{(ii)} there is a percolation $\P$ on $\rtd$ that gives
4241: spanning trees with at most 2 ends a.s.;
4242: \itemrm{(iii)} there is a percolation $\P$ on $\rtd$ that gives
4243: non-empty connected subgraphs $\omega$ that satisfy $\pc(\omega) = 1$ a.s.
4244: \enditems
4245: \endprocl
4246: 
4247: \proof
4248: The proof that (i) implies (ii) is done as for Theorem 5.3 of \BLPSgip, but
4249: uses Propositions \briefref p.amenforest/ and \briefref p.limittree/.
4250: That (ii) implies (iii) is obvious.
4251: The proof that (iii) implies (i) follows the first part of the proof of
4252: Theorem 1.1 in \ref b.BLPS:crit/.
4253: \Qed
4254: 
4255: \procl c.amentrees \procname{Amenable Trees}
4256: A unimodular probability measure $\rtd$ on infinite
4257: rooted trees is amenable iff $\expdeg(\rtd) = 2$ iff $\rtd$-a.s.\ $\gh$ has
4258: 1 or 2 ends.
4259: \endprocl
4260: 
4261: \proof
4262: Combine \ref t.treechar/ with \ref t.deg2/.
4263: \comment{
4264: Since the average degree in any finite tree is less than 2, we have
4265: $\expdeg(\rtd') < 2$ for all $\rtd' \in \fdom(\rtd)$, and so $\alpha(\rtd) \le 2$.
4266: Therefore, $\expdeg(\rtd) > 2$ implies non-amenability by \ref p.alpha/.
4267: On the other hand, if $\expdeg(\rtd) = 2$, then by \ref t.deg2/, $\rtd$ is
4268: concentrated on trees with at most 2 ends, whence on trees with $\pc = 1$.
4269: Therefore, Bernoulli($p$) percolation on $\rtd$ produces only finite
4270: components and for $p$ close to 1, produces a $\rtd' \in \fdom(\rtd)$ with
4271: $\expdeg(\rtd')$ close to 2.
4272: Thus, $\alpha(\rtd) = 2$ and we have amenability by \ref p.alpha/.
4273: The last equivalence is from \ref t.deg2/.
4274: }%
4275: \Qed
4276: 
4277: 
4278: 
4279: The next result was proved for non-amenable unimodular transitive graphs in
4280: \BLPSgip\ with a more direct proof in \ref b.BLPS:crit/. This extension is
4281: proved similarly.
4282: Presumably, the hypothesis that $\rtd$ is non-amenable can be replaced by the
4283: assumption that $\pc(\rtd) < 1$. (This is a major open conjecture for
4284: quasi-transitive graphs.)
4285: 
4286: \procl t.death \procname{Critical Percolation}
4287: Let $\rtd$ be an extremal unimodular non-amen\-a\-ble
4288: probability measure on $\GG_*$
4289: with $\expdeg(\rtd) < \infty$.
4290: There is $\P_{\pc(\rtd)}$-a.s.\ no infinite cluster.
4291: \endprocl
4292: 
4293: %The finiteness condition is needed for ruling out 1 component by knowing
4294: %that P(edge open) \to 1 implies degree of percolation \to degree of graph.
4295: % I initially required $\rtd$ to be concentrated on non-amenable graphs,
4296: % but I don't see why this restriction is needed.
4297: 
4298: This can be interpreted for finite graphs as follows.
4299: Suppose that $\gh_n$ are finite connected graphs with bounded average
4300: degree whose random weak limit is
4301: extremal and non-amenable with critical value $\pc$.
4302: Consider Bernoulli($\pc$) percolation on $\gh_n$.
4303: Let $\alpha_n(\ell)$ be the random variable giving
4304: the proportion of vertices of $\gh_n$ that belong to simple open paths of
4305: length at least $\ell$.
4306: Then $\lim_{\ell \to\infty} \lim_{n \to\infty} \alpha_n(\ell) = 0$ in
4307: probability.
4308: 
4309: 
4310: The following theorem is proved similarly to \ref b.BS:hp/, as extended by
4311: \ref b.LP:book/, and by using \ref x.dual/.
4312: 
4313: \procl t.planarperc \procname{Planar Percolation}
4314: Let $\rtd \in \uni$ be extremal, non-amenable,
4315: and carried by plane graphs with one end and bounded degree.
4316: Then $0 < \pc(\rtd) < \pu(\rtd) < 1$ and Bernoulli($\pu(\rtd)$) percolation
4317: on $\rtd$ has a unique infinite cluster a.s.
4318: \endprocl
4319: 
4320: 
4321: 
4322: The following extends Theorems 3.1, 3.2, 3.5, 3.6, and 3.10 of \ref
4323: b.BLS:pert/ and is proved similarly.
4324: If $\P$ is a percolation on $\rtd$ with $\P \circ \psi^{-1} = \rtd$ and
4325: with $\marks_0 \subseteq \marks$
4326: defining the open subgraphs, we call $\P'$ a {\bf subpercolation} on $\P$
4327: that {\bf gives} subgraphs in $\A'$ {\bf with positive probability} if there
4328: is a Borel
4329: function $\psi': \marks \to \marks$ such that $\P = \P' \circ \psi'^{-1}$
4330: and there is a Borel subset $\marks_1 \subseteq \marks$ such that if
4331: $G(\marks_1)$ denotes the $\marks_1$-open subnetwork of $G$, then
4332: $\P'\left[\psi\Big(\psi'\big(G(\marks_1)\big)\big(\marks_0\big)\Big) \in
4333: \A'\right] > 0$.
4334: 
4335: \procl t.sub-iso \procname{Non-Amenable Subgraphs}
4336: Let $\rtd$ be a unimodular probability measure on $\GG_*$ with finite
4337: expected degree and $\P$ be a
4338: percolation on $\rtd$ with open subgraph $\omega$.
4339: \beginitems
4340: \itemrm{(i)} If $h > 0$ and $\E[\deg_\omega \bp \mid \bp \in \omega] \ge
4341: \alpha(\rtd) + 2h$, then there is a subpercolation $\P'$ on $\P$ that gives
4342: a non-empty subgraph $\omega'$ with $\isoe(\omega') \ge h$ with positive
4343: $\P'$-probability.
4344: \itemrm{(ii)} If $\omega$ is a forest a.s., $h > 0$ and $\E[\deg_\omega \bp
4345: \mid \bp \in \omega] \ge 2 + 2h$, then there is a subpercolation $\P'$ on
4346: $\P$ that gives a non-empty subgraph with $\isoe \ge h$ with positive
4347: probability.
4348: \itemrm{(iii)} If $\rtd$ is non-amenable and $\omega$ has exactly one
4349: infinite cluster $\P$-a.s., then there is a subpercolation $\P'$ on $\P$
4350: that gives a non-empty subgraph $\omega'$ with $\isoe(\omega') > 0$ $\P'$-a.s.
4351: \itemrm{(iv)} If $\omega$ has components with at least three ends
4352: $\P$-a.s., then there is a subpercolation $\P'$ on $\P$ that gives a
4353: non-empty forest $\fo$ with $\isoe(\fo) > 0$ $\P'$-a.s.
4354: \itemrm{(v)} If $\rtd$ is concentrated on subgraphs with spectral radius
4355: less than 1 and $\omega$ has exactly one infinite cluster $\P$-a.s., then
4356: there is a subpercolation $\P'$ on $\P$ that gives a non-empty forest $\fo$
4357: with $\isoe(\fo) > 0$ $\P'$-a.s.
4358: \enditems
4359: \endprocl
4360: 
4361: The following extends a result of \ref b.Hag:rcust/ and is proved
4362: similarly to Corollary 6.3 of \BLPSusf.
4363: 
4364: \procl p.amen-for \procname{Amenability and Boundary Conditions}
4365: Let $\rtd$ be an amenable unimodular probability measure on $\GG_*$.
4366: Then $\FUSF(\rtd) = \WUSF(\rtd)$ and $\FMSF(\rtd) = \WMSF(\rtd)$.
4367: \endprocl
4368: 
4369: 
4370: We may now strengthen \ref p.recurrent/, despite the fact that not every
4371: graph is necessarily non-amenable $\rtd$-a.s.
4372: It extends Theorem 4.3 of \ref b.BLS:pert/ and is proved similarly, using
4373: \ref t.sub-iso/(iii) with $\P := \rtd$.
4374: 
4375: 
4376: \procl t.posspeed \procname{Positive Speed on Non-Amenable Graphs}
4377: If $\rtd \in \uni$ is non-amenable and concentrated on graphs with bounded
4378: degree,
4379: then the speed of simple random walk is positive $\rtd$-a.s.
4380: \endprocl
4381: 
4382: 
4383: 
4384: \bsection{Examples}{s.ex}
4385: 
4386: 
4387: We present here a variety of interesting examples of unimodular
4388: measures.
4389: 
4390: \procl x.renew \procname{Renewal Processes}
4391: Given a stationary (delayed) renewal process on $\Z$, let $\rtd$ be the law
4392: of $(\Z, 0)$ with the graph $\Z$, some fixed mark at renewals, and some
4393: other fixed mark elsewhere. Then $\rtd \in \uni$.
4394: \endprocl
4395: 
4396: \procl x.halfplane \procname{Half-Plane}
4397: Fix $d \ge 3$ and let $T$ be the $d$-regular tree.
4398: Let $\rtd_1$ be the random weak limit of balls of growing radii in $T$.
4399: Note that $\rtd_1$ is carried by trees with only one end.
4400: Let $\rtd_2$ be concentrated on the fixed graph $(\Z, 0)$.
4401: Now let $\rtd := \rtd_1 \itimes \rtd_2$.
4402: This is a unimodular version of the half-plane $\N \times \Z$.
4403: \endprocl
4404: 
4405: \procl x.cover
4406: For a rooted network $(\gh, \bp)$, its {\bf universal cover} is the rooted
4407: tree $(T, \bp) = T(\gh, \bp)$ formed as follows.
4408: The vertices of $T$ are the finite paths in $\gh$ that start at $\bp$ and
4409: do not backtrack.
4410: Two such vertices are joined by an edge in $T$ if one is an extension of
4411: the other by exactly one edge in $\gh$.
4412: The path with no edges consisting of just the vertex $\bp$ in $\gh$ is the
4413: root $\bp$ of $T$.
4414: There is a natural rooted graph homomorphism $\pi : (T, \bp) \to (\gh,
4415: \bp)$ (the cover map) that maps paths to their last point.
4416: Marks on $T$ are defined by lifting the marks on $\gh$ via $\pi$.
4417: It is clear that
4418: if $\rtd \in \uni$ and $\nu$ is the law of $T(\gh, \bp)$ when $(\gh, \bp)$
4419: has the law $\rtd$, then $\nu \in \uni$
4420: and $\expdeg(\rtd) = \expdeg(\nu)$.
4421: \endprocl
4422: 
4423: 
4424: \procl x.cluster
4425: Let $\P$ be a unimodular percolation on $\rtd$ that labels edges either
4426: open or closed.
4427: Let $\nu$ be the law of the open cluster of the root when the network is
4428: chosen according to $\P$ conditional on the root belonging to an infinite
4429: open cluster.
4430: Then $\nu$ is unimodular, as a direct verification of the definition shows.
4431: When $\rtd$ is concentrated on a fixed unimodular graph, this fact has been
4432: widely used in the study of percolation.
4433: \endprocl
4434: 
4435: 
4436: \procl x.Voronoi \procname{Tilings}
4437: Let $X$ be a Euclidean space or hyperbolic space (of constant curvature).
4438: Write $\gp$ for its isometry group.
4439: There is a Mass-Transport Principle for $X$ that says the following; see
4440: \ref b.BS:hp/ for a proof.
4441: Let $\rho$ be a positive Borel measure on $X \times X$ that is invariant
4442: under the diagonal action of $\gp$.
4443: Then there is a constant $c$ such that for all Borel $A \subset X$ of
4444: volume $|A| > 0$, we have $\rho(A \times X) = \rho(X \times A) = c |A|$.
4445: Suppose that $P$ is a (countable) point process in $X$ whose law is
4446: $\gp$-invariant.
4447: For example, Poisson point processes are $\gp$-invariant.
4448: One often considers graphs $\gh$ that are functions $\gh = \beta(P)$, where
4449: $\beta$ commutes with the action of $\gp$.
4450: For a few recent examples, see
4451: \ref b.BS:hp/, \ref b.HolPer:PP/, or \ref b.Timar:PP/.
4452: For instance, the 1-skeleton of the Voronoi tessellation corresponding to
4453: $P$ is such a graph.
4454: In general, we call such measures on graphs {\bf $\gp$-equivariant factors}
4455: of $P$.
4456: They are necessarily $\gp$-invariant.
4457: \par
4458: Another way that invariant measures on graphs embedded in $X$ occur is
4459: through (aperiodic) tilings of $X$.
4460: Again, one can take the 1-skeleton.
4461: An important tool for studying aperiodic tilings is a limit measure
4462: obtained from translates of a given tiling (in the Euclidean case), or,
4463: more generally, invariant measures on tilings with special properties; see,
4464: e.g., \ref b.Rob/, \ref b.Radin:survey/, \ref b.Solomyak/,
4465: \ref b.Radin:book/, or
4466: \ref b.BowRad:densest/ for some examples.
4467: \par
4468: Let $\nu$ be any $\gp$-invariant probability measure on graphs embedded in
4469: $X$.
4470: Fix a Borel set $A \subset X$ of positive finite volume.
4471: If $v(A) := \int |\vertex(\gh) \cap A| \,d\nu(\gh) < \infty$,
4472: then define $\rtd$ as follows.
4473: Choose $\gh$ with the law $\nu$ biased by $|\vertex(\gh) \cap A|$.
4474: Then choose the root $\bp$ of $\gh$
4475: uniformly among all vertices that belong to $A$.
4476: The law of the resulting graph $(\gh, \bp)$ is $\rtd$.
4477: We claim that $\rtd$ is unimodular and does not depend on $A$.
4478: In fact, $\rtd$ is the same as the Palm measure of $(\gh, \bp)$, except
4479: that $\rtd$ is a measure on isomorphism classes of graphs that
4480: does not involve any geometric embedding.
4481: To prove our claims, we first write $\rtd$ in symbols:
4482: $$
4483: \rtd(\A) :=
4484: v(A)^{-1} \int \sum_{\bp \in A} \II{(\gh, \bp) \in \A} \,d\nu(\gh)
4485: $$
4486: for Borel $\A \subset \GG_*$.
4487: Let $f : \gtwo \to [0, \infty]$ be Borel.
4488: Define
4489: $$
4490: \rho(B \times C) :=
4491: \int \sum_{x \in \vertex(\gh) \cap B} \sum_{y \in \vertex(\gh) \cap C}
4492: f(\gh, x, y)\,d\nu(\gh)
4493: $$
4494: for Borel $B, C \subseteq X$.  Since $\nu$ is invariant, $\rho$ is
4495: diagonally invariant.
4496: Therefore,
4497: $$\eqaln{
4498: \int \sum_{x \in \vertex(\gh)} f(\gh, \bp, x) \,d\rtd(\gh, \bp)
4499: &=
4500: v(A)^{-1} \rho(A \times X)
4501: =
4502: v(A)^{-1} \rho(X \times A)
4503: \cr&=
4504: \int \sum_{x \in \vertex(\gh)} f(\gh, x, \bp) \,d\rtd(\gh, \bp)
4505: \,,
4506: }$$
4507: which means that $\rtd$
4508: satisfies the Mass-Transport Principle, i.e., is unimodular.
4509: Furthermore, if we take $f(\gh, x, y) := \II{x = y}$, then we see that
4510: $\rho(A \times X) = v(A)$, so that
4511: there is a constant $c$ such that $v(A) = c |A|$.
4512: Likewise, if $f(\gh, x, y) := \II{x = y, (\gh, x) \in \ev A}$, then we see
4513: that for each $\A$, there is another constant $c_\A$ such that $v(A)
4514: \rtd(\A) = c_\A |A|$.
4515: It follows that $\rtd$ does not depend on $A$.
4516: \endprocl
4517: 
4518: 
4519: %\procl x.aperiodic
4520: %Similarly, given any amenable graph $\gh$ and any network on it, if we take
4521: %a limit point of the sequence of probability measures gotten by taking a
4522: %uniform root in a F\o lner sequence, we get a unimodular measure.
4523: %\endprocl
4524: 
4525: 
4526: 
4527: \procl x.dual \procname{Planar Duals}
4528: Let $\rtd$ be a unimodular probability measure on plane graphs all of
4529: whose faces have finitely many sides.
4530: We are assuming that to each graph, there is a measurably associated
4531: plane embedding.
4532: Thus, each graph $\gh$ has a plane dual $\dual \gh$ with respect to its
4533: embedding.
4534: In fact, to be technically accurate in what follows, we replace the
4535: embedding by an assignment (possibly random) of marks to the edges that
4536: indicate the cyclic order in which they appear around a vertex
4537: in a fixed orientation of the plane.
4538: (For example, if a vertex $x$ has $d$ edges incident to it, then one can
4539: let the $d$ edge marks associated to $x$ be $\{1, 2, \ldots, d\}$ in cyclic
4540: order, with the one marked 1 chosen at random, independently of marks
4541: elsewhere.)
4542: Then the plane dual graph is defined entirely with respect to the
4543: resulting network in an automorphism-equivariant way, needing no reference
4544: to the plane.
4545: \par
4546: Provided a certain finiteness condition is satisfied,
4547: there is a natural unimodular probability measure on the dual graphs,
4548: constructed as follows.
4549: For a face $f$, let $\deg f$ denote the
4550: number of sides of $f$. For a vertex $x$, let $F(x) := \sum_{f \sim x}
4551: 1/\deg f$.
4552: Assume that $Z := \int F(\bp) \,d\rtd(\gh, \bp) < \infty$.
4553: To create a unimodular probability measure $\dual \rtd$ on the duals,
4554: first choose $(\gh, \bp)$ with law $\rtd$ biased by $F(\bp)/Z$.
4555: Then choose a face $f_0$ incident to $\bp$ with probability proportional to
4556: $1/\deg f_0$.
4557: The law of the resulting rooted graph $(\dual \gh, f_0)$ is $\dual \rtd$:
4558: $$
4559: \dual \rtd(\A) :=
4560: Z^{-1} \int \sum_{f_0 \sim \bp} {1 \over \deg f_0} \II{(\dual \gh, f_0) \in
4561: \A}
4562: \,d\rtd(\gh, \bp)
4563: $$
4564: for Borel $\A \subseteq \GG_*$.
4565: To prove that $\dual \rtd$ is indeed unimodular, let $k : \gtwo \to [0, \infty]$
4566: be Borel. Then
4567: $$\eqaln{
4568: Z \int \sum_{f \in \vertex(\dual \gh)} k(\dual \gh, f_0, f)
4569: \,d\dual \rtd(\dual \gh, f_0)
4570: &=
4571: \int
4572: \sum_{f_0 \sim \bp} {1 \over \deg f_0}
4573: \sum_{f \in \vertex(\dual \gh)} k(\dual \gh, f_0, f)
4574: \,d\rtd(\gh, \bp)
4575: \cr&=
4576: \int
4577: \sum_{f_0 \sim \bp}
4578: \sum_{f \in \vertex(\dual \gh)}
4579: {1 \over \deg f_0}
4580: k(\dual \gh, f_0, f)
4581: \sum_{x \sim f} {1 \over \deg f}
4582: \,d\rtd(\gh, \bp)
4583: \cr&=
4584: \int
4585: \sum_{x \in \vertex(\gh)}
4586: \sum_{f_0 \sim \bp}
4587: \sum_{f \sim x}
4588: {1 \over \deg f_0}
4589: {1 \over \deg f}
4590: k(\dual \gh, f_0, f)
4591: \,d\rtd(\gh, \bp)
4592: \cr&=
4593: \int
4594: \sum_{x \in \vertex(\gh)}
4595: \sum_{f_0 \sim x}
4596: \sum_{f \sim \bp}
4597: {1 \over \deg f_0}
4598: {1 \over \deg f}
4599: k(\dual \gh, f_0, f)
4600: \,d\rtd(\gh, \bp)
4601: \cr&\hskip1in\hbox{[by the Mass-Transport Principle for $\rtd$]}
4602: \cr&=
4603: \int
4604: \sum_{x \in \vertex(\gh)}
4605: \sum_{f \sim x}
4606: \sum_{f_0 \sim \bp}
4607: {1 \over \deg f_0}
4608: {1 \over \deg f}
4609: k(\dual \gh, f, f_0)
4610: \,d\rtd(\gh, \bp)
4611: \cr&=
4612: Z \int \sum_{f \in \vertex(\dual \gh)} k(\dual \gh, f, f_0)
4613: \,d\dual \rtd(\dual \gh, f_0)
4614: \,.
4615: }$$
4616: Thus, $\dual \rtd$ satisfies the Mass-Transport Principle, so is unimodular.
4617: A similar argument shows that $\dual{(\dual \rtd)} = \rtd$.
4618: \par
4619: Another important construction comes from combining the primal and dual
4620: graphs into a new plane graph by adding a vertex where each edge crosses
4621: its dual.
4622: That is, if $\gh$ is a plane graph and $\dual \gh$ its dual, then every
4623: edge $e \in \edge(\gh)$ intersects $\dual e \in \edge(\dual \gh)$ in one
4624: point, $v_e$.
4625: (These are the only intersections of $\gh$ and $\dual \gh$.)
4626: For $e \in \edge(\gh)$, write $\widehat e$ for the pair of edges that
4627: result from the subdivision of $e$ by $v_e$, and likewise for
4628: $\widehat{\dual e}$.
4629: This defines a new graph $\widehat \gh$, whose vertices are
4630: $\verts(\gh)\cup\verts(\dual \gh)\cup\Big\{v_e\st e\in\edges(\gh)\Big\}$
4631: and whose edges are $\bigcup_{e \in\edges(\gh)} (\widehat e \cup
4632: \widehat{\dual e})$.
4633: If $\expdeg(\rtd) < \infty$, then we may define a unimodular probability
4634: measure $\widehat \rtd$ on the graphs $\widehat \gh$ from $\rtd$ as
4635: follows.
4636: Let $\widehat Z := (3/2) \expdeg(\rtd) + Z \le (5/2) \expdeg(\rtd) <
4637: \infty$.
4638: For $x \in \vertex(\gh)$, let $\widehat N(x)$ be the set consisting of $x$
4639: itself plus the vertices of $\widehat \gh$ that correspond to edges or
4640: faces of $\gh$ that are incident to $x$.
4641: For $w \in \vertex(\widehat \gh)$, define
4642: $$
4643: \delta(w) :=
4644: |\{x \in \vertex(\gh) \st w \in \widehat N(x)\}|^{-1}
4645: =
4646: \cases{
4647: 1 &if $w \in \vertex(\gh)$,\cr
4648: 1/2 &if $w = v_e$ for some $e \in \edge(\gh)$,\cr
4649: 1/\deg w &if $w \in \vertex(\dual \gh)$.\cr
4650: }
4651: $$
4652: Define
4653: $$
4654: \widehat \rtd(\A) :=
4655: \widehat Z^{-1} \int \sum_{w_0 \in \widehat N(\bp)} \delta(w_0)
4656: \II{(\widehat \gh, w_0) \in \A} \,d\rtd(\gh, \bp)
4657: \,.
4658: $$
4659: Note that $\widehat\rtd$ is a probability measure.
4660: To prove that $\widehat\rtd$ is unimodular, let $k : \gtwo \to [0, \infty]$
4661: be Borel. Then
4662: $$\eqaln{
4663: \widehat Z \int \sum_{w \in \vertex(\widehat \gh)} k(\widehat \gh, w_0, w)
4664: &\,d\widehat\rtd(\widehat \gh, w_0)
4665: =
4666: \int
4667: \sum_{w_0 \in \widehat N(\bp)} \delta(w_0)
4668: \sum_{w \in \vertex(\widehat \gh)} k(\widehat \gh, w_0, w)
4669: \,d\rtd(\gh, \bp)
4670: \cr&=
4671: \int
4672: \sum_{w_0 \in \widehat N(\bp)}
4673: \sum_{w \in \vertex(\widehat \gh)}
4674: \delta(w_0)
4675: k(\widehat \gh, w_0, w)
4676: \sum_{x \st w \in \widehat N(x)} \delta(w)
4677: \,d\rtd(\gh, \bp)
4678: \cr&=
4679: \int
4680: \sum_{x \in \vertex(\gh)}
4681: \sum_{w_0 \in \widehat N(\bp)}
4682: \sum_{w \in \widehat N(x)}
4683: \delta(w_0)
4684: \delta(w)
4685: k(\widehat \gh, w_0, w)
4686: \,d\rtd(\gh, \bp)
4687: \cr&=
4688: \int
4689: \sum_{x \in \vertex(\gh)}
4690: \sum_{w_0 \in \widehat N(x)}
4691: \sum_{w \in \widehat N(\bp)}
4692: \delta(w_0)
4693: \delta(w)
4694: k(\widehat \gh, w_0, w)
4695: \,d\rtd(\gh, \bp)
4696: \cr&\hskip1in\hbox{[by the Mass-Transport Principle for $\rtd$]}
4697: \cr&=
4698: \int
4699: \sum_{x \in \vertex(\gh)}
4700: \sum_{w \in \widehat N(x)}
4701: \sum_{w_0 \in \widehat N(\bp)}
4702: \delta(w_0)
4703: \delta(w)
4704: k(\widehat \gh, w, w_0)
4705: \,d\rtd(\gh, \bp)
4706: \cr&=
4707: \widehat Z \int \sum_{w \in \vertex(\widehat \gh)} k(\widehat \gh, w, w_0)
4708: \,d\widehat \rtd(\widehat \gh, w_0)
4709: \,.
4710: }$$
4711: Thus, $\widehat\rtd$ satisfies the Mass-Transport Principle, so is unimodular.
4712: 
4713: \endprocl
4714: 
4715: 
4716: 
4717: \procl x.PWIT \procname{Poisson Weighted Infinite Tree}
4718: Our definition
4719: (\ref s.notation/)
4720: of the metric on the space $\GG_*$ of rooted graphs
4721: refers to ``balls of radius $r$" in which distance is
4722: graph distance, i.e., edges implicitly have length $1$.
4723: \ref b.AS:obj/ work in the setting of graphs whose edges
4724: have positive real lengths, so that distance becomes minimum
4725: path length.  This setting permits one to consider 
4726: graphs which may have infinite degree, but which are still
4727: ``locally finite" in the sense that only finitely vertices fall within
4728: any finite radius ball.
4729: Of course, edge lengths are a special (symmetric) case of edge marks.
4730: %We have restricted ourselves to locally finite graphs. 
4731: %However, if we allowed infinite degree with an $\N$-ordering of the edges
4732: %incident to any vertex, then our theory would extend.
4733: An important example is the following.
4734: Consider a regular rooted tree $T$ of infinite degree.  % marked as follows.
4735: Fix a continuous increasing function $\Lambda$ on $\CO{0, \infty}$ with
4736: $\Lambda(0) = 0$ and $\lim_{t \to\infty} \Lambda(t) = \infty$.
4737: Order the children of each vertex of $T$ via a bijection with $\Z^+$.
4738: For each vertex $x$, consider an independent Poisson process on $\R^+$ with
4739: mean function $\Lambda$.
4740: Define the length of the edge joining $x$ to its $n$th child to be the
4741: $n$th point of the Poisson process associated to $x$.
4742: This is a unimodular random network in the extended sense of \ref b.AS:obj/.
4743: It can be derived by taking the random weak limit of the complete graph on $n$
4744: vertices whose edge lengths are independent with cdf 
4745: %vertices with each edge marked independently by a random variable having cdf
4746: $t \mapsto 1 - e^{-\Lambda(t)/n}$ ($t \ge 0$)
4747: and then deleting all edges in the limit whose length
4748: is $\infty$.
4749: (We are working here with the mark space $[0, \infty]$.)
4750: %Triangles will not disappear in the weak limit. This is why
4751: %we need to delete infinitely long edges.
4752: See \ref b.Aldous:random-assign/.
4753: \endprocl
4754: 
4755: 
4756: %\procl x.planar
4757: %Many examples of random weak limits of finite random trees are described
4758: %in \ref b.Aldous:fringe/.
4759: %Interesting recent work of
4760: %\ref b.AngelSchramm/ and \ref b.Angel/ describes random weak limits of uniform plane
4761: %triangulations.
4762: %\endprocl
4763: 
4764: \procl x.stretched \procname{Edge Replacement}
4765: Here is a general way to create unimodular random rooted graphs from existing
4766: unimodular fixed graphs.
4767: This is an extension of the random subdivision (or stretching) introduced
4768: by \ref b.AdamsLyons/ and studied further in
4769: Example 2.4.4 of \ref b.Kaim:harmonic/ and \ref b.ChenPeres/.
4770: Let $\fin$ be
4771: the set of isomorphism classes of finite graphs with
4772: an ordered pair of distinct distinguished vertices.
4773: For our construction, we may start with a fixed
4774: unimodular quasi-transitive connected graph, $\gh$, or, more generally,
4775: with any unimodular probability measure $\rtd$ on $\GG_*$.
4776: In the former case, fix an orientation of the edges of $\gh$ and let $L$ be
4777: a random field on the oriented edges of $\gh$ that is invariant under the
4778: automorphism group of $\gh$ and takes values in $\fin$ and such that
4779: $\big|\vertex\big(L(e)\big)\big|$ has finite mean for each edge $e$.
4780: Replace each edge $e$ with the graph $L(e)$, where the first of the
4781: distinguished vertices of $L(e)$ is identified with the tail of $e$ and the
4782: second of its distinguished vertices is identified with the head of $e$.
4783: Call the resulting random graph $H$.
4784: There is a unimodular probability measure that is equivalent to this
4785: measure on random graphs.
4786: Namely, let $\mu_i$ be the law of $(H, o_i)$, where $\{o_i\}$ is a complete
4787: section of the vertex orbits of $\gh$.
4788: Given $H$, write
4789: $$
4790: A(x) := 2 + \sum_{e \sim x} \Big(\big|\vertex\big(L(e)\big)\big|-2\Big)
4791: $$
4792: and
4793: $$
4794: c := \sum_i \E\big[A(o_i)\big] \big|\Stab(o_i)\big|^{-1}
4795: \,.
4796: $$
4797: Choose $o_i$ with probability $c^{-1} \E\big[A(o_i)\big]
4798: \big|\Stab(o_i)\big|^{-1}$.
4799: Given $o_i$, choose $(H, o_i)$ with distribution $\mu_i$.
4800: Given this, list the non-distinguished vertices of all $L(e)$ for $e$
4801: incident to $o_i$ as $z_1, z_2, \ldots, z_{A(o_i) - 2}$ and set $z_{A(o_i) - 1}
4802: := z_{A(o_i)} := o_i$.
4803: Let $U$ be a uniform integer in $\big[1, A(o_i)\big]$.
4804: Then $(H, z_U)$ is unimodular and, clearly, has law with respect to which
4805: $\sum_i \mu_i$ is absolutely continuous.
4806: 
4807: Indeed, we state and prove this more generally.
4808: Suppose that $\rtd$ is a unimodular probability measure on $\GG_*$.
4809: Orient the edges of the rooted networks arbitrarily.
4810: Let $\mkmp(e)$ denote the ordered pair of the
4811: marks of the edge $e$ (ordered by the orientation of $e$).
4812: Suppose $L : \marks^2 \to \fin$ is Borel with the property
4813: that whenever $L(\mk_1, \mk_2) = (G, x, y)$, we also have $L(\mk_2, \mk_1) =
4814: (G, y, x)$.
4815: (This will ensure that the orientation of the edges will not affect the
4816: result.)
4817: If
4818: $$
4819: \int \sum_{e \sim \bp} \bigg[\Big|\vertex\Big(L\big(\mkmp(e)\big)\Big)\Big|
4820: - 2\bigg]
4821: \,d\rtd(\gh, \bp)
4822: < \infty
4823: \,,
4824: $$
4825: then let $\rtd'$ be the following measure.
4826: Define
4827: $$
4828: A(\gh, \bp) := 2 + \sum_{e \sim \bp}
4829: \bigg[\Big|\vertex\Big(L\big(\mkmp(e)\big)\Big)\Big| - 2\bigg]
4830: \,.
4831: $$
4832: Choose $(\gh, \bp)$ with probability distribution $\rtd$ biased by $A(\gh,
4833: \bp)$ and replace each edge $e$ by the graph $L\big(\mkmp(e)\big)$, where
4834: the tail and head of $e$ are identified with the first and second
4835: distinguished vertices of $L\big(\mkmp(e)\big)$, respectively; call the
4836: resulting graph $H$.
4837: Write $A := A(\gh, \bp)$ and
4838: list the non-distinguished vertices of all $L\big(\mkmp(e)\big)$ for $e$
4839: incident to $\bp$ as $z_1, z_2, \ldots, z_{A - 2}$ and set $z_{A - 1}
4840: := z_{A} := \bp$.
4841: Let $U$ be a uniform integer in $[1, A]$.
4842: Finally, let $\rtd'$ be the distribution of $(H, z_U)$.
4843: 
4844: This is unimodular by the following calculation.
4845: Write $H(\gh)$ for the graph $H$ formed as above from the network $\gh$.
4846: Let $\vertex_0(\mk_1, \mk_2)$ be the set of non-distinguished vertices of
4847: the graph $L(\mk_1, \mk_2)$.
4848: Write $z_i(\gh, \bp)$ ($1 \le i \le A(\gh, \bp)-2$) for the vertices of the
4849: neighborhood $B(\gh, \bp) := \bigcup_{e \sim \bp} \vertex_0\big(\mkmp(e)\big)$.
4850: Write $z_i(\gh, \bp) := \bp$ for $i = A(\gh, \bp)-1, A(\gh, \bp)$.
4851: Put $c := \int A(\gh, \bp) \,d\rtd(\gh, \bp)$.
4852: In order to show that $\rtd'$ is unimodular, let $f : \gtwo \to [0,
4853: \infty]$ be Borel.
4854: Define $\overline f : \gtwo \to [0, \infty]$ by
4855: $$\eqaln{
4856: \overline f(\gh, x, y) 
4857: &:=
4858: {1 \over c}
4859: \sum_{z \in B(\gh, x)}
4860: \sum_{z' \in B(\gh, y)}
4861: f\big(H(\gh), z, z'\big)
4862: +
4863: {2 \over c}
4864: \sum_{z \in B(\gh, x)}
4865: f\big(H(\gh), z, y\big)
4866: \cr&\qquad+
4867: {2 \over c}
4868: \sum_{z' \in B(\gh, y)}
4869: f\big(H(\gh), x, z'\big)
4870: +
4871: {2 \over c}
4872: f\big(H(\gh), x, y\big)
4873: \,.
4874: \cr
4875: }$$
4876: Then 
4877: $$\eqaln{
4878: \int \sum_{z \in \vertex(H)} &f(H, \bp, z) \,d\rtd'(H, \bp)
4879: \cr&=
4880: {1 \over c} \int {1 \over A(\gh, \bp)} \sum_{i=1}^{A(\gh, \bp)}
4881: \sum_{z \in \vertex(H(\gh))} f\big(H(\gh), z_i(\gh, \bp), z\big)
4882: A(\gh, \bp)\,d\rtd(\gh, \bp)
4883: \cr&=
4884: \int \sum_{x \in \vertex(\gh)} \overline f(\gh, \bp, x) \,d\rtd(\gh, \bp)
4885: \cr&=
4886: \int \sum_{x \in \vertex(\gh)} \overline f(\gh, x, \bp) \,d\rtd(\gh, \bp)
4887: \cr&=
4888: \int \sum_{z \in \vertex(H)} f(H, z, \bp) \,d\rtd'(H, \bp)
4889: \,.
4890: }$$
4891: %If $\rtd = U(G)$ for some finite graph $G$, then it is easy to see that
4892: %$\rtd' = U(H)$ for the graph $H$ constructed from $G$.
4893: %In the general case, we use \ref t.TII/ to see that $\rtd'$ is a random
4894: %weak limit of finite networks, whence unimodular.
4895: %Furthermore, the law of $H$ rooted at the root of $\gh$
4896: %equals $\rtd'$ conditioned on the root being one of the
4897: %distinguished vertices in the replacement graphs.
4898: \endprocl
4899: 
4900: Our final example details the correspondence between random rooted graphs
4901: and graphings of equivalence relations.
4902: 
4903: 
4904: \procl x.equivalence
4905: Let $\mu$ be a Borel probability measure on a topological space
4906: $X$ and $R$ be a Borel subset of
4907: $X^2$ that is an equivalence relation with finite or countable equivalence
4908: classes. We call the triple $(X, \mu, R)$ a {\bf measured equivalence
4909: relation}.
4910: For $x \in X$, denote its $R$-equivalence class by $[x]$.
4911: We call $R$ {\bf measure preserving} if
4912: $$
4913: \int_{x \in X} \sum_{y \in [x]} f(x, y) \,d\mu(x)
4914: =
4915: \int_{x \in X} \sum_{y \in [x]} f(y, x) \,d\mu(x)
4916: $$
4917: for all Borel $f : X^2 \to [0, \infty]$.
4918: A {\bf graphing} $\Phi$ of $R$ is a Borel subset of $X^2$ such that the
4919: smallest equivalence relation containing $\Phi$ is $R$.
4920: A graphing $\Phi$ induces the structure of a graph on the vertex set $X$
4921: by defining an edge between $x$ and $y$ if $(x, y) \in \Phi$ or $(y, x)
4922: \in \Phi$.
4923: Denote the subgraph induced on $[x]$ and rooted at $x$ by $\Phi(x)$.
4924: Given Borel maps $\psi : X \to \marks$ and $\phi : X^2 \to \marks$,
4925: we regard $\psi(x)$ as the mark at $x$ and $\phi(x, y)$ as the mark at $x$
4926: of the edge from $x$ to $y$.
4927: Thus, $\Phi(x)$ is a random rooted network. Its law (or, rather, the law of
4928: its rooted isomorphism class) is unimodular iff $R$ is measure preserving.
4929: 
4930: Conversely, suppose that $\rtd$ is a probability measure on $\GG_*$.
4931: Add independent uniform marks as second coordinates to the existing marks
4932: and call the resulting measure $\nu$.
4933: Write $\dmn \subset \GG_*$ for the set of (isomorphism classes of) rooted
4934: networks with distinct marks.
4935: Thus, $\nu$ is concentrated on $\dmn$.
4936: Define $R \subset \dmn^2$ to be the set of pairs of (isomorphism classes
4937: of) rooted networks that are
4938: non-rooted isomorphic.
4939: Define $\Phi \subset R$ to be the set of pairs of isomorphic rooted
4940: networks whose roots are neighbors in the unique (non-rooted) isomorphism.
4941: Then $(\dmn, \nu, R)$ is a measured equivalence relation with graphing
4942: $\Phi$.
4943: We have that $R$ is measure preserving iff $\rtd$ is unimodular.
4944: If we define the mark map $\projection$ that forgets the second coordinate,
4945: then $\nu$ pushes forward to $\rtd$, i.e., $\rtd = \nu \circ \projection^{-1}$.
4946: 
4947: Thus, the theory of unimodular random rooted networks has substantial
4948: overlap with the
4949: theory of graphed measure-preserving equivalence relations.
4950: The largest difference between the two theories lies in the foci of
4951: attention: We focus on probabilistic aspects of the graphing, while the
4952: other theory focuses on ergodic aspects of the equivalence relation (and,
4953: thus, considers all graphings of a given equivalence relation).
4954: The origins of our work lie in two distinct areas: one is group-invariant
4955: percolation on graphs, while the other is asymptotic analysis of finite
4956: graphs. The origin of the study of measured equivalence relations lies in
4957: the ergodic theory of group actions.
4958: Some references for the latter work, showing relations to von Neumann
4959: algebras and logic, among other things, are
4960: \refbmulti{FM1,FM2}, \ref b.FHM/,
4961: \ref b.CFW/,
4962: \ref b.Zimmer:book/,
4963: \ref b.KechrisMiller/,
4964: \ref b.BeckerKech:book/, \ref b.Kechris:turing/,
4965: \ref b.AdamsLyons/,
4966: \refbmulti {Kaim:amen,Kaim:harmonic},
4967: \ref b.Paulin/,
4968: \refbmulti{Gaboriau:cost,Gaboriau:betti},
4969: and \refbmulti{Furman1,Furman2}.
4970: \endprocl
4971: 
4972: 
4973: 
4974: 
4975: 
4976: \bsection{Finite Approximation}{s.finite}
4977: 
4978: Although we do not present any theorems in this section, because of its
4979: potential importance, we have devoted the whole section to the question of
4980: whether
4981: finite networks are weakly dense in
4982: $\uni$.
4983: Let us call random weak limits of finite networks {\bf sofic}.
4984: 
4985: \procl q.TII \procname{Finite Approximation}
4986: Is every probability measure in $\uni$ sofic? In other words, if
4987: $\rtd$ is a unimodular probability measure on $\GG_*$,
4988: do there exist finite networks $G_n$ such that
4989: $U(G_n) \cd \rtd$?
4990: \endprocl
4991: 
4992: To appreciate why the answer is not obvious, consider the
4993: special case of the (non-random) graph consisting of the
4994: infinite rooted 3-regular tree.
4995: It is true that there exist finite graphs $G_n$ that
4996: approximate the infinite 3-regular tree in the sense of random
4997: weak convergence; a moment's thought shows these {\it cannot}
4998: be finite trees.
4999: This special case is of course known
5000: (one can use finite quotient groups of $\Z_2 * \Z_2 * \Z_2$,
5001: random $3$-regular graphs, or expanders (\ref B.lub94/)),
5002: but the constructions in this special case do not readily extend to
5003: the general case.
5004: 
5005: Another known case of sofic measures is more difficult to establish.
5006: Namely, \ref b.Bowen:periodic/ showed
5007: that all unimodular networks on regular trees are sofic. (To deduce this from
5008: his result, one must use the fact, easily established, that networks with
5009: marks from a finite set are dense in $\GG_*$.)
5010: 
5011: \procl x.fdd
5012: The general unimodular Galton-Watson
5013: measure $\UGW$ (\ref x.AGW/) is also sofic.
5014: To see this, consider the following random networks,
5015: sometimes called ``fixed-degree
5016: distribution networks" and first studied by \ref b.MolloyReed/.
5017: Given
5018: $\Seq{r_k}$ and $n$ vertices, give each vertex $k$ balls with probability
5019: $r_k$, independently. Then pair the balls at random and place an edge for
5020: each pair between the corresponding vertices. There may be one ball left
5021: over; if so, ignore it. Let $m_0 := \sum k r_k$, which we assume is
5022: finite.  In the limit, we get a tree where the root has degree $k$ with
5023: probability $r_k$, each neighbor of the root, if any, has degree $k$ with
5024: probability $k r_k/m_0$, etc. In fact, we get $\UGW$ for the
5025: offspring distribution $k \mapsto (k+1) r_{k+1}/m_0$.  Thus, if we want
5026: the offspring distribution $\Seq{p_k}$, we need merely start with $r_k :=
5027: c^{-1} p_{k-1}/k$ for $k \ge 1$ and $r_0 := 0$, where $c := \sum_{k \ge 0}
5028: p_k/(k+1)$.
5029: \endprocl
5030: 
5031: Let us compare the
5032: intuitions behind amenability and unimodularity.
5033: One can define an average of any bounded function on the vertices of, say,
5034: an amenable Cayley graph; the average will be the same for any translate of
5035: the function.
5036: This can also be regarded as an average of the function with respect to a
5037: probability measure that chooses a group element uniformly at random;
5038: however, the precise justification of this requires a measure that, though
5039: group invariant, is only finitely additive.
5040: Nevertheless, this invariant measure is approximated by uniform measures on
5041: finite sets, namely, F\o{}lner sets.
5042: By contrast, the justification that a unimodular random rooted graph
5043: provides a uniform distribution on the vertices is via the Mass-Transport
5044: Principle. The measure itself is, of course, countably additive; if it is
5045: sofic, then it,
5046: too, is approximated by uniform measures on finite sets.
5047: The two intuitions concerning uniform measures that come from amenability
5048: and from unimodularity agree insofar as every amenable quasi-transitive graph
5049: is unimodular, as shown by \ref b.SoardiWoess/ and \ref b.Salvatori/.
5050: 
5051: One might think that if a sequence $\Seq{\gh_n}$ of finite graphs has a
5052: fixed transitive graph $\gh$ as its random weak limit, then any unimodular
5053: probability measure on networks supported by $\gh$ would be a random weak
5054: limit of some choice of networks on the same sequence $\Seq{\gh_n}$.
5055: This is false, however; e.g., if $\gh$ is a 3-regular tree, then almost any
5056: choice of a sequence of growing 3-regular graphs has $\gh$ as its random weak
5057: limit. However, there is a random independent set\ftnote{*}{This means that
5058: no two vertices of the set are adjacent.} of density 1/2 on $\gh$ whose
5059: law is automorphism invariant,
5060: while the density of independent sets in random 3-regular graphs is bounded
5061: away from 1/2 (see \ref b.FS/).
5062: Nevertheless, if \ref q.TII/ has a positive answer, then
5063: it is not hard to show that
5064: there is some sequence of finite graphs that has this
5065: property of carrying arbitrary networks.
5066: 
5067: Recall that the {\bf Cayley diagram} of a group $\gp$ generated by a finite
5068: subset $S$ is the network $(\gh, \bp)$ with vertex set $\gp$, edge set
5069: $\big\{(x, x s) \st x \in \gp,\ s \in S\big\}$, root $\bp$ the identity
5070: element of $\gp$, and edge marks $s$ at the endpoint $x$ of $(x, x s)$ and
5071: $s^{-1}$ at the endpoint $x s$ of $(x, x s)$, as in \ref r.anygroup/.
5072: We do not mark the vertices (or mark them all the same).
5073: \ref b.Weiss:sofic/ defined $\gp$ to be {\bf sofic} if its Cayley diagram 
5074: is a random weak limit of finite networks with edge marks from
5075: $S \cup S^{-1}$.
5076: It is easy to check that this property does not depend on the generating
5077: set $S$ chosen.
5078: By embedding $S \cup S^{-1}$ into $\marks$, we can use a positive answer to
5079: \ref q.TII/ to give every
5080: Cayley diagram as a random weak limit of {\it some\/} finite networks.
5081: Changing the marks on the finite networks to their nearest points in $S
5082: \cup S^{-1}$ gives the kind of approximating networks desired. That is, we
5083: would have that
5084: every finitely generated group is sofic.
5085: As we mentioned in the introduction, this would have plenty of
5086: consequences.
5087: 
5088: To illustrate additional consequences of
5089: a positive answer to
5090: \ref q.TII/, we establish
5091: the existence of various probability measures on sofic networks.
5092: The idea is that if a class of probability measures is specified by a
5093: sequence of local ``closed" conditions in such a way that there is a measure in
5094: that class on any finite graph, then there is an automorphism-invariant
5095: measure in that class on any sofic quasi-transitive graph.
5096: Rather than state a general theorem to that effect, we shall
5097: illustrate the principle by two examples.
5098: 
5099: 
5100: \procl x.gibbs \procname{Invariant Markov Random Fields}
5101: Consider networks with vertex marks $\pm 1$ and no (or constant) edge marks.
5102: %Let $H$ be a local potential function, i.e., for any finite network
5103: %$\gh$, $H(\gh) \in \OC{-\infty, + \infty}$.
5104: %For example, suppose $H$ is $\infty$ unless all marks are $\pm 1$.
5105: %Let us ignore edge marks.
5106: Given a finite graph $\gh$, $h \in \R$, and $\beta > 0$,
5107: the probability measure $\nu_\gh$ on mark maps $\mkmp : \verts(\gh) \to \{-1,
5108: +1\}$ given by
5109: $$
5110: \nu_\gh(\mkmp) := Z^{-1} \exp\Big\{\sum_{x \in \verts(\gh)} \beta h \mkmp(x) -
5111: \sum_{x \sim y} \beta \mkmp(x) \mkmp(y)\Big\}
5112: \,,
5113: $$
5114: where $Z$ is the normalizing constant required to make these probabilities
5115: add to 1, is known as the
5116: anti-ferromagnetic Ising model at inverse
5117: temperature $\beta$ and with external field $h$ on $\gh$.
5118: %is given by $H(\gh) =
5119: %- \beta h \epsilon$ when $\gh$ is an isolated vertex with mark $\epsilon$,
5120: %$H(\gh) = + \beta \epsilon_1 \epsilon_2$ when $\gh$ is a single edge whose
5121: %endpoints are marked $\epsilon_1$ and $\epsilon_2$, and $H(\gh) = 0$
5122: %otherwise.
5123: Let $\gh$ now be an infinite sofic transitive graph.
5124: Let $\Seq{\gh_n}$ be an approximating sequence of finite graphs
5125: and let $\nu_n := \nu_{\gh_n}$ be the corresponding probability measures.
5126: Write $\overline \gh_n$ for the corresponding random network on $\gh_n$
5127: and, as usual, $U(\overline \gh_n)$ for the uniformly rooted network.
5128: Since $U(\overline \gh_n)$ is unimodular, so is any weak limit point,
5129: $\rtd$.
5130: By tightness, there is such a weak limit point, and it is concentrated on
5131: networks with underlying graph $\gh$.
5132: By \ref t.uni-vs-invar/, we may lift $\rtd$ to an
5133: automorphism-invariant
5134: measure $\nu = \lift_\rtd$ on networks on $\gh$.
5135: This measure $\nu$ is a Markov random field with the required Gibbs
5136: specification, meaning that for any finite subgraph $H$ of $\gh$, the
5137: conditional $\nu$-distribution of $\mkmp \restrict \vertex(H)$ given $\mkmp
5138: \restrict \bdv H$ is equal to the conditional $\nu_H$-distribution of the
5139: same thing.
5140: One is often interested in invariant random fields, not just any random
5141: fields with the given Gibbs specification.
5142: One can easily get a Markov random field with the required Gibbs
5143: specification by taking a limit over subgraphs of $\gh$, but this will not
5144: necessarily produce an invariant measure. In case $\gh$ is amenable, one
5145: could take a limit of averages of the resulting measure to obtain an
5146: invariant measure, but this will not work in the non-amenable case.
5147: That is the point of the present construction.
5148: A variation on this is spin glasses:
5149: Here, for finite graphs $\gh$, the measure $\nu_\gh$ is
5150: $$
5151: \nu_\gh(\mkmp) := Z^{-1} \exp\Big\{\sum_{x \in \verts(\gh)} \beta h \mkmp(x) +
5152: \sum_{x \sim y} \beta J_{x, y} \mkmp(x) \mkmp(y)\Big\}
5153: \,,
5154: $$
5155: where $J_{x, y}$ are, say, independent $\pm 1$-valued random variables.
5156: Again, one can find an invariant spin glass (coupled to the independent
5157: interactions $J_{x, y}$) with the same parameters $h$ and $\beta$ on any
5158: transitive sofic graph by the above method.
5159: \endprocl
5160: 
5161: 
5162: \procl x.sandpile \procname{Invariant Sandpiles}
5163: Consider now networks with vertex marks in $\N$ and no edge marks.
5164: Given a finite rooted 
5165: graph $(\gh, \bp)$, a mark map $\mkmp : \vertex(\gh)
5166: \to \N$ is called {\bf critical} (or {\bf stable and recurrent}) if for all
5167: $x \in \vertex(\gh)$, we have $\mkmp(x) < \deg(x)$ and for all
5168: subgraphs $W$ of $\gh \setminus \{\bp\}$, there is some $x \in \vertex(W)$
5169: such that $\mkmp(x) \ge \deg_W(x)$; we may take $\mkmp(\bp) \equiv 0$.
5170: In this context, one usually calls the root ``the sink".
5171: It turns out that the set of such mark
5172: maps form a very interesting group, called the {\bf sandpile
5173: group} or {\bf chip-firing group} of $\gh$; see \ref b.btw/, \ref
5174: b.Dhar:survey/, \ref b.Biggs/, and \ref b.MRZ/.
5175: Let $\nu_{(\gh, \bp)}$ be the uniform measure on critical mark maps.
5176: Given a transitive sofic graph $\gh$ and an approximating sequence
5177: $\Seq{\gh_n}$ of finite graphs, let $\nu_n :=
5178: \nu_{(\gh_n, \bp_n)}$ be the corresponding probability measures for any
5179: fixed choice of roots $\bp_n$.
5180: Write $(\overline \gh_n, \bp_n)$ for the corresponding random network on
5181: $(\gh_n, \bp_n)$ and, as usual, $U(\overline \gh_n)$ for the uniformly rooted
5182: network. (This root is unrelated to $\bp_n$.)
5183: Since $U(\overline \gh_n)$ is unimodular, so is any weak limit point,
5184: $\rtd$.
5185: By tightness, there is such a weak limit point, and it is concentrated on
5186: networks with underlying graph $\gh$.
5187: (Probably the entire sequence $\Seq{U(\overline \gh_n)}$ in fact converges
5188: to $\rtd$.)
5189: By \ref t.uni-vs-invar/, we may lift $\rtd$ to an
5190: automorphism-invariant
5191: measure $\nu$ on networks on $\gh$.
5192: This measure $\nu$ is supported by networks with only critical mark maps in
5193: the sense that for all $x \in \vertex(\gh)$, we have $\mkmp(x) < \deg(x)$
5194: and for all subgraphs $W$ of $\gh$, there is some $x \in \vertex(W)$ such
5195: that $\mkmp(x) \ge \deg_W(x)$.
5196: \endprocl
5197: 
5198: However, we do not know how to answer the following question.
5199: 
5200: \procl q.color \procname{Invariant Coloring}
5201: Given a quasi-transitive infinite graph $\gh$ and a number $c$
5202: that is 
5203: at least the chromatic number of $\gh$, is there an $\Aut(\gh)$-invariant
5204: probability measure on proper $c$-colorings of the vertices of $\gh$?
5205: %The following does not work: Let $U(\gh_n) \cd \gh$ (say in the
5206: %transitive case). Consider a coloring of $\gh_n$ that maximizes the number
5207: %of vertices colored $\{0, 1, \ldots, c-1\}$. Will the resulting networks
5208: %have a limit that uses only these colors? In other words, if every large
5209: %neighborhood of a finite graph is $c$-colorable, then can a small
5210: %proportion of vertices be deleted from the graph to give a graph that is
5211: %$c$-colorable? NO; we need to be more careful. Can we be?
5212: Let $D := \max_{x \in \vertex(\gh)} \deg_\gh x $.
5213: A positive answer for $c \ge D + 1$ is due to Schramm (personal
5214: communication, 1997).
5215: If $\gh$ is sofic, then we can also obtain such a measure for $c = D$ by
5216: using a well-known result of Brooks (see, e.g., \ref b.Bollobas/, p.~148,
5217: Theorem V.3.)
5218: The question is particularly interesting when $\gh$ is planar and $c = 4$.
5219: In fact, it is then also of great interest to know whether there is a
5220: quasi-transitive proper 4-coloring of $\gh$.
5221: \endprocl
5222: 
5223: 
5224: 
5225: 
5226: 
5227: \medbreak
5228: \noindent {\bf Acknowledgements.}\enspace
5229: We are grateful to 
5230: %Hari Bercovici for helpful discussions related to \ref c.unbdd/ and to
5231: Oded Schramm for discussions of Theorems \briefref
5232: t.uni-vs-invar/ and \briefref t.msf1end/.
5233: We also thank Oded Schramm and \'Ad\'am Tim\'ar for various suggestions.
5234: We are indebted to
5235: G\'abor Elek for telling us about the connection of our work to
5236: various conjectures about groups.
5237: 
5238: 
5239: 
5240: \def\noop#1{\relax}
5241: \input \jobname.bbl
5242: 
5243: \filbreak
5244: \begingroup
5245: \eightpoint\sc
5246: \parindent=0pt\baselineskip=10pt
5247: 
5248: Department of Statistics,
5249: University of California Berkeley,
5250: Berkeley, CA, 94720-3860
5251: \emailwww{aldous@stat.berkeley.edu}
5252: {http://www.stat.berkeley.edu/users/aldous}
5253: 
5254: Department of Mathematics,
5255: Indiana University,
5256: Bloomington, IN 47405-5701
5257: \emailwww{rdlyons@indiana.edu}
5258: {http://mypage.iu.edu/\string~rdlyons/}
5259: 
5260: \endgroup
5261: 
5262: 
5263: \bye
5264: 
5265: 
5266: