ABSTRACT Title of dissertation: TWO EQUIVALENCE RELATIONS IN SYMBOLIC DYNAMICS Andrew Dykstra Doctor of Philosophy, 2007 Dissertation directed by: Professor Mike Boyle Department of Mathematics A G-shift of finite type (G-SFT) is a shift of finite type which commutes with the continuous action of a finite group G. We classify irreducible G-SFTs up to right closing almost conjugacy, answering a question of Bill Parry. Then, we derive a computable set of necessary and sufficient conditions for the existence of a homomorphism from one shift of finite type to another. We consider an equivalence relation on subshifts, called weak equivalence, which was introduced and studied by Beal and Perrin. We classify arbitrary shifts of finite type up to weak equivalence. TWO EQUIVALENCE RELATIONS IN SYMBOLIC DYNAMICS by Andrew Dykstra Dissertation submitted to the Faculty of the Graduate School of the University of Maryland, College Park in partial fulfillment of the requirements for the degree of Doctor of Philosophy 2007 Advisory Committee: Professor Mike Boyle, Chair/Advisor Professor Michael Brin Professor Brian Hunt Professor Michael Jakobson Professor Jim Purtilo Dedication To my parents Tim and Nancy, and to my partner John. ii Acknowledgements I wish to thank a number of people whose support was essential to the writing of this dissertation. It was my great fortune to have Dr. Mike Boyle as my thesis advisor. Mike always had good suggestions for problems to work on, as well as possible approaches to solving those problems. Often I would come to him feeling discouraged, having made only feeble progress on a particular problem; he would instantly suggest a way to move forward with my ideas, or perhaps a new train of thought altogether. For a couple of years we worked on some truly challenging problems, but were unable to solve any of them completely. In this time, Mike taught me never to lose heart, and to learn from my failures?failure is part of being a mathematician. Leading by example, he taught me that success comes with optimism and perseverance. Thank you, Mike, for believing in me more than I did. My work at Maryland was also shaped by my colleagues in the Research Inter- action Team (RIT) in symbolic dynamics. Led by Mike Boyle, the RIT consisted of a small group of students and postdocs. Each week, we would meet to discuss our research and significant papers from our field. I would especially like to thank RIT participants Angela Desai, Todd Fisher, and Nick Long, who provided a sounding board for my ideas and taught me about the problems they were working on. Angela was especially helpful when I first joined the RIT. She had already mastered the fundamentals of symbolic dynamics and was willing to bring me up to speed. She initiated regular meetings with me outside of the RIT, in which we read and dis- cussed the book, An Introduction to Symbolic Dynamics and Coding, by Lind and iii Marcus [33]. Todd, a postdoc, brought a tremendous amount of enthusiasm and expertise to the RIT. Todd demonstrated an ability to explain difficult concepts in ways that even I could understand. Also, Todd offered me a great deal of profes- sional and career advice. Nick opened my eyes to connections between symbolic dynamics and K-theory and was an excellent listener when I presented my ideas. I wish to thank my former housemate and friend, Andy Kebo. Though his re- search is in harmonic analysis and quantum computation, Andy showed tremendous interest in my work. He would ask compelling questions, and let me bounce ideas off of him. I have fond memories of the long walks we would take, where we would discuss our daily challenges and victories. Andy is one of the most thoughtful and caring people I have ever met. Lastly, I would like to thank my former housemates and friends Aaron Lott and Yue Xiao for their consideration and emotional support. I am especially grateful to Aaron, who taught me to typeset my work in LATEX and spent several hours helping me with certain technical aspects of my thesis defense. iv Table of Contents 1 Introduction 1 2 Right closing almost conjugacy for G-shifts of finite type 10 2.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10 2.2 Background and definitions . . . . . . . . . . . . . . . . . . . . . . . 11 2.2.1 Shifts of finite type . . . . . . . . . . . . . . . . . . . . . . . . 11 2.2.2 G-shifts of finite type . . . . . . . . . . . . . . . . . . . . . . . 13 2.2.3 Skew products and matrices over Z+G . . . . . . . . . . . . . 13 2.3 Some useful results . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15 2.4 Right closing almost conjugacy for mixing free G-SFTs . . . . . . . . 19 2.5 General mixing G-SFTs . . . . . . . . . . . . . . . . . . . . . . . . . 24 2.6 The irreducible but periodic case . . . . . . . . . . . . . . . . . . . . 26 2.7 Regular isomorphism of G-Markov chains . . . . . . . . . . . . . . . . 29 3 Weak equivalence for shifts of finite type 31 3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31 3.2 Preliminaries . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32 3.2.1 Shifts of finite type . . . . . . . . . . . . . . . . . . . . . . . . 32 3.2.2 Edge shifts . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33 3.2.3 Irreducible shifts of finite type . . . . . . . . . . . . . . . . . . 34 3.2.4 Reducible shifts of finite type . . . . . . . . . . . . . . . . . . 35 3.3 Homomorphisms between irreducible SFTs . . . . . . . . . . . . . . . 35 3.4 Extending homomorphisms defined on a nonwandering set . . . . . . 37 3.5 Homomorphisms between reducible SFTs . . . . . . . . . . . . . . . . 44 3.6 Weak equivalence of SFTs . . . . . . . . . . . . . . . . . . . . . . . . 45 Bibliography 49 v Chapter 1 Introduction We work in symbolic dynamics, focusing on the classification theory of one- dimensional shifts of finite type. This classification theory is built on invariants, which enable us to describe the similarities and differences between systems. We begin with a discussion of Bernoulli shifts. For a natural number n, assign to each element i of the set {0,...,n?1} a number pi, such that p0+???+pn?1 = 1. Let X be the set of sequences x = (xi)i?Z where each xi ? {0,...,n?1}. Give to X the sigma algebra A which is the product of the algebra of subsets of {0,...,n?1}. Define a probability measure ? on X, given by ? = {p0,...,pn?1}Z. The triple (X,A,?) is a probability space. Define the shift map ? : X ? X by ?(x)i = xi+1 for each i ? Z. Then ? is a measure-preserving transformation of (X,A,?), and the dynamical system (X,A,?,?) is a Bernoulli shift. As an example, the Bernoulli shift with n = 2 and p0 = p1 = 12 corresponds to tossing a fair coin infinitely often (indexed by Z). In the 1950?s, Kolmogorov and Sinai introduced measure-theoretic entropy as an isomorphism invariant for measure-preserving transformations on a probability space. They showed that the measure-theoretic entropy of a Bernoulli shift (as above) is ? n?1summationdisplay i=0 pilogpi. It follows that not all Bernoulli shifts are measurably iso- morphic, as Bernoulli shifts defined by different probabilities pi typically do not 1 have the same measure-theoretic entropy. Then in 1970, Ornstein [36] proved that Bernoulli shifts of equal measure-theoretic entropy are measurably isomorphic. In other words, not only is measure-theoretic entropy an isomorphism invariant for Bernoulli shifts?it is a complete isomorphism invariant. Often we think of the elements of {0,...,n?1} as states, and each coordinate xk of a point x ? X represents one unit of time spent at state xk. For k > 0, we think of the state xk as occurring in the future, x?k as having occurred in the past, and x0 as the present state of x. A Bernoulli shift X has the property that, given a state i ? {0,...,n? 1}, the probability that a point x ? X has i as its present state is independent of the past or future states of x (the probability that x0 = i is simply pi). This motivates us to consider objects, more general than Bernoulli shifts, for which we tolerate a certain amount of dependence on the past. Let A be an n?n stochastic matrix. Then A is the adjacency matrix for the directed labeled graph G which has n vertices {v0,...,vn?1} and, for each A(i,j) negationslash= 0, there is a directed edge in G beginning at vi and ending at vj, labeled A(i,j). We let X be the set of sequences x = (xi)i?Z such that each xi is a vertex in G, and each A(xi,xi+1) negationslash= 0. Then, with the appropriate sigma algebra A and measure ?, (X,A,?,?) is a measure-preserving transformation called a Markov chain. Example 1.0.1. Let A = ? ?? ? .3 .7 1 0 ? ?? ?. Then A is the adjacency matrix for the following directed graph. 2 d55d54d53d52d48d49d50d51v0.3 d54d54 .7 d34d34 d55d54d53d52d48d49d50d51v1 1 d98d98 Notice here that the probability of a particular present state depends on the past. For example, if x?1 = v0, then the probability of x0 = v0 is .3 whereas, if x?1 = v1, then the probability of x0 = v0 is 1. The basic building blocks of Markov chains are the irreducible Markov chains. Generalizing [36], Friedman and Ornstein [22] classify irreducible Markov chains up to measurable isomorphism in terms of measure-theoretic entropy and one additional invariant?period. But not only are we interested in Markov chains as measurable objects. Also we may think of Markov chains topologically, as follows. For a Markov chain X and associated directed graph G having vertex set {v0,...,vn?1}, give to X ? {v0,...,vn?1}Z the topology T which is the relative of the product of the discrete topology on {v0,...,vn?1}. Then (X,T ) is compact and metrizable, and the shift map? : (X,T ) ? (X,T ) is a homeomorphism. (X,T,?) is a topological dynamical system called a shift of finite type (SFT). The topological analogue of measurable isomorphism is topological conjugacy. The problem of classifying SFTs up to topological conjugacy has turned out to be more difficult than the problem of classifying Markov chains up to measurable isomorphism. In fact, a complete set of invariants for topological conjugacy of irreducible SFTs has yet to be found. However there do exist a number of weaker equivalence relations up to which SFTs have been classified. For an irreducible SFT X, there exists a Markov measure on X with respect 3 to which the measure-theoretic entropy of X is maximized. This maximal measure- theoretic entropy of X is also the topological entropy of X (topological entropy having been defined in 1965 by Adler, Konheim and McAndrew [3] as the topolog- ical analogue of measure-theoretic entropy). In light of Friedman and Ornstein?s work [22], it is natural to ask what, if anything, can be said topologically about irreducible SFTs of equal period which are measurably isomorphic with respect to their maximal-entropy measures. This question is answered in the seminal paper of Adler and Marcus [4]. Here they introduce an equivalence relation on SFTs, called almost conjugacy (AC), for which topological entropy and period are complete invariants. SFTs X and Y are almost conjugate if they admit a common SFT extensionZ, where the mapsZ ?X and Z ? Y giving the extension are surjective and one-to-one almost everywhere. Given irreducible SFTsX andY of equal period, with measures of maximal entropy ?X and ?Y respectively, X and Y are almost conjugate if and only if there exists a measurable isomorphism (X,?X) ? (Y,?Y). To this point, all Markov chains we have considered have been two-sided, in the sense that they arise as Z-products. But also it is natural to consider N- products?the one-sided Markov chains. In contrast to the two-sided setting where the shift map is an isomorphism, the shift map on a one-sided Markov chain is merely an endomorphism (it is typically non-invertible). Surprisingly, whereas two- sided Bernoulli shifts are classified in terms of measure-theoretic entropy, it follows from [40], [42] and [43] that distinct one-sided Bernoulli shifts of equal measure- theoretic entropy are never measurably isomorphic. What invariants are necessary 4 and sufficient to classify one-sided Markov chains up to measurable isomorphism? The answer to this question involves a class of maps between SFTs called closing maps. The simplest examples of closing maps are resolving maps, which arise from certain natural matrix equations, and which play a key role in the construction of Adler and Marcus [4]. The (more general) closing maps are important to symbolic dynamics because of their connections to algebraic invariants, and also because all known general constructions of maps between SFTs involve closing maps. A map between SFTs is right closing if it is injective on each unstable set of the domain, and left closing if it is injective on each stable set of the domain. A map is closing if it is right or left closing. If X and Y are irreducible SFTs with the same topological entropy and period, then maps ?X : Z ? X and ?Y : Z ? Y giving an almost conjugacy between X and Y may be chosen to be closing maps. However, in the construction [4], it is not possible to construct ?X and ?Y that are either both right closing or both left closing. (If one of ?X or ?Y is right closing, then the other will be left closing.) The importance of closing maps is emphasized by connections to regular iso- morphism. Bill Parry introduces (and studies) regular ismomorphisms of two-sided Markov chains, first in [20] and also in [37]. Roughly speaking, a measurable isomor- phism of two-sided Markov chains ? : (X,?) ? (Y,?) is regular if knowing the past and a bounded look into the future of a point x?X determines the present state of the point ?(x) ?Y. Recall that almost conjugacies of SFTs X and Y correspond to measurable isomorphisms of Markov chains (X,?X) and (Y,?Y), where ?X and ?Y are the measures of maximal measure-theoretic entropy for X and Y respectively. 5 Perhaps there is some stronger form of almost conjugacy, corresponding to regular isomorphism? Yes, there is, and it is called right closing almost conjugacy (RCAC): SFTs X and Y are right closing almost conjugate if they admit a common SFT extension Z, where the maps Z ?X and Z ?Y giving the extension are surjective, one-to-one almost everywhere, and right closing. In [15], irreducible SFTs are classified up to RCAC in terms of topological entropy, period, and an algebraic invariant called ideal class. Regular isomorphisms are exactly the measurable isomorphisms that right closing almost conjugacies induce. In [8], Ashley, Marcus and Tuncel find computable invariants with which to classify one-sided Markov chains up to measurable isomorphism. As Boyle and Tun- cel point out in [18], these include in particular all invariants of regular isomorphism of two-sided Markov chains. In chapter 2, we investigate RCAC, but in a more general setting than was studied in the past. The objects we consider are called G-shifts of finite type (G- SFTs), and the maps between G-SFTs are called G-maps. For a finite group G, a G-SFT is an SFT equipped with a continuous and shift commuting G-action. A G-map X ?Y between G-SFTs is a map of SFTs, equivariant with respect to the G-actions on both X and Y. So a G-SFT is in particular an SFT, and may be thought of as an SFT by simply ignoring the G-action. Similarly a G-map is in particular a map of SFTs. The equivalence relations AC and RCAC are defined in our new context by simply replacing ?SFT? with ?G-SFT? and ?map? with ?G-map? in the respective definitions. 6 Our initial motivation to classify RCAC for irreducible G-SFTs was to answer a question of Bill Parry [38]. Additionally, a classification of AC for irreducible G-SFTs had appeared in [1] (and also in [37]), so the problem of classifying RCAC for irreducible G-SFTs seemed natural. Also, there is a well-defined class of G- SFTs (the ones where the G-action is free) which arise as skew products. The study of skew products dates all the way back to von Neumann, and is of independent interest. In section 2.3, we generalize a number of results from the SFT context to the G-SFT context, most noticeably with our Z+G Masking Lemma and our ZG Replacement Theorem. Then in section 2.4 we show, somewhat surprisingly, that for mixing G-SFTs arising as skew products, there are no new invariants of RCAC that weren?t already invariants in the SFT context. This is not the case in general, however, and in sections 2.5 and 2.6 we introduce some new invariants, and use them to classify arbitrary irreducible G-SFTs up to RCAC. We conclude chapter 2 with a discussion of regular isomorphism in the G-SFT context. In chapter 3, my co-author Joseph Barth and I study another equivalence re- lation, called weak equivalence. As its name suggests, weak equivalence is a weak relation?significantly weaker than AC or RCAC. Subshifts S and T are weak equiv- alent if there exist full shifts ?n and ?m containing S and T, respectively, and maps f : ?n ? ?m and g : ?m ? ?n such that f?1(T) = S and g?1(S) = T. Weak equivalence was introduced and studied by Beal and Perrin in [9]. Their main result is a classification of a very special class of irreducible SFTs (the so- called flower shifts) up to weak equivalence. This begs the question of what would 7 be needed to classify arbitrary SFTs up to weak equivalence. For an answer, first we investigate the structure of the most general class of SFTs?reducible SFTs. Reducible SFTs are classified up to flow equivalence by Huang in [23] and [24]; with Boyle, an alternative description appears in [11] and [14]. Kim and Roush show in [26] that a solution to the problem of classifying arbitrary SFTs up to topological conjugacy would follow from a solution in the irreducible case. We contribute to the theory of reducible SFTs in section 3.4 by introducing the notion of a phase matrix for an SFT. We show that the set of phase matrices for an SFT is a computable conjugacy invariant and, with Theorem 3.4.5, we give necessary and sufficient conditions in terms of phase matrices for there to exist an extension of a given map NW(S) ?NW(T) between nonwandering sets of SFTs S and T to a map S ?T. This enables us in section 3.5 to classify when there exists a map from one SFT to another. Then in section 3.6 we classify weak equivalence for SFTs, by showing that SFTs S and T are weak equivalent if and only if there exist maps S ?T and T ?S. The centerpiece of our work is an extension result (Theorem 3.4.5). As out- lined in section 3.1, extension results are significant to symbolic dynamics. In section 3.6 we indicate that our work may be relevant to a difficult open problem on ex- tensions (Problem 3.6.8). This problem is derived from the work of Maass [34], and is formulated in terms of steady maps. We mention a certain special class of SFT pairs (S,T) such that there exists at least one map S ?T which fails to be steady. AnaturalquestionrelatedtoTheorem3.4.5is, givenasurjectivemapNW(S) ? 8 NW(T) from the nonwandering set of and SFT S onto the nonwandering set of an SFT T, what are necessary and sufficient conditions for there to exist an extension of the given map to a surjective map S ? T? We hope to answer this question in the near future but, as of now, all we can say is that the answer to this question will certainly be nontrivial. Also difficult and open is the question of when there exists a surjective map from one SFT onto another. An answer to the question posed in the preceding paragraph would provide a reduction of this question to the irreducible case. How- ever the irreducible case has shown itself to be extraordinarily challenging, and will certainly involve deep algebraic invariants. Chapter 2 was published as a paper in the journal Colloquium Mathematicum. Chapter 3 is a paper in preprint, soon to be submitted for publication. 9 Chapter 2 Right closing almost conjugacy for G-shifts of finite type 2.1 Introduction For a finite group G, a G-shift of finite type (G-SFT) is a shift of finite type (X,?) together with a continuous G action on X which commutes with the shift ?. For irreducible shifts of finite type, right closing almost conjugacy is classified in terms of entropy, period, and an algebraic invariant called ideal class [15]. Bill Parry [38] posed the following question: what additional invariants are necessary to classify right closing almost conjugacy for irreducible G-SFTs? With Theorem 2.4.1 we show that for mixing G-SFTs where the G action is free, there are no additional invariants. In section 2.5 we generalize Theorem 2.4.1 to mixing G-SFTs where the G action is no longer assumed to be free. In section 2.6 we generalize further to irreducible but periodic G-SFTs. As a corollary to our results we classify regular isomorphism for G-Markov chains with respect to measures of maximal entropy. Without the right closing assumption, almost conjugacy for irreducible G- SFTs was classified by Roy Adler, Bruce Kitchens and Brian Marcus [1]. They were working in a more general setting, but by modifying the proofs given here we can arrive at the same classification of almost conjugacy for irreducible G-SFTs (as was also done in [37]). I thank Mike Boyle for many helpful discussions about this problem. 10 2.2 Background and definitions We assume some familiarity with shifts of finite type; [29] and [33] provide more complete backgrounds. All of the free G-SFTs we consider arise out of skew products, as in [16]. The study of skew products dates back to von Neumann, in the context of ergodic measure preserving transformations on a probability space. For an example of more recent work with skew products in ergodic theory, see [21]. Also see [39] and [41] (and their references) for recent results with skew products in Livsic theory. 2.2.1 Shifts of finite type Let A be an n?n matrix over the nonnegative integers Z+. Then A is the adjacency matrix for a directed graph, GA, which has vertices {v1,v2,...,vn}, and the number of edges from vI to vJ is AIJ. Let EA = {edges in GA}, and put ?A = {x = (xi)i?Z ? (EA)Z|each xixi+1 is a path in GA}. With the appropriate topology (the relative of the product of the discrete topology on EA), ?A is a compact metric space. The shift on ?A is the homeomorphism ? : ?A ? ?A given by (?x)i = xi+1. The pair (?A,?) is the edge shift of finite type (SFT) defined by A. Where ? is understood, we write just ?A to denote (?A,?). A map between SFTs pi: ?A ? ?B is a continuous function such that pi ? ?(x) = ??pi(x) for all x? ?A. The map pi is one block if it is induced by a function which sends each edge of GA to an edge of GB. A factor map is a surjective map. An injective factor map is a conjugacy. 11 The matrixAis irreducible if for each entryAIJ ofAthere is a natural number N such that (AN)IJ > 0. If A is irreducible we say also that the graph GA and the edge SFT ?A are irreducible. The matrix A is primitive if there is a natural number N such that for each entry AIJ of A, (AN)IJ > 0. If A is primitive we say also that the graph GA is primitive; in this case the edge SFT ?A is mixing. If ? is the Perron eigenvalue of A, then the entropy of ?A is log?. If v = [v1,v2,...,vn]T is a right Perron eigenvector with entries in the ring Z[1/?], then the ideal class of ?A is the class of the Z[1/?]-ideal which is generated by the components v1,...,vn of v. A point x ? ?A is periodic if there exists a natural number p such that ?p(x) = x. In this case p is a period of x; the smallest period of x is called the least period of x. We define the period of the edge SFT ?A to be the greatest common divisor of the set of periods of periodic points in ?A. The period of the graph GA is the period of ?A. Any SFT (X,?) is conjugate to some edge SFT (?A,?). Then, the terms irreducible, mixing, entropy, ideal class and period apply to X exactly as they apply to ?A. A point x?X is doubly transitive if both sets {?n(x): n? 0} and {?n: n? 0} are dense in X. Two pointsx = (xn)n?Z andy = (yn)n?Z inX are left asymptotic if there is an integer n such that xk = yk for all k ? n. A map between SFTs pi: X ? Y is 1-1 a.e. if it is injective on the set of doubly transitive points in X. The map pi is right closing if, for each pair x,y ? X of distinct left asymptotic points, pi(x) negationslash= pi(y). We say the SFTs X and Y are right closing almost conjugate as SFTs if there is a third SFT Z which factors onto both X and Y by factor maps which are 1-1 a.e. and right closing. 12 2.2.2 G-shifts of finite type LetGbe a finite group. AG-SFT is an SFT (X,?) together with a continuous right G action on X such that ?(x?g) = ?(x) ?g for all x ? X and g ? G. We say the G-SFT X (or the G action on X) is free if, for each non-identity element g of G, x?g negationslash= x for all x ? X. We say X (or the G action on X) is faithful if, for each non-identity element g of G, there exists some x ? X such that x?g negationslash= x. If Y is another G-SFT, then a G-map pi: X ? Y is a map between SFTs such that pi(x?g) = pi(x)?g for all x ? X and g ? G. A G-factor map is a surjective G-map and a G-conjugacy is an injective G-factor map. Two G-SFTs X and Y are right closing almost conjugate as G-SFTs if there is a third G-SFT Z which factors onto both X and Y by 1-1 a.e. and right closing G-factor maps. We point out that right closing almost conjugate G-SFTs are in particular right closing almost conjugate SFTs. The terms we define above for SFTs, such as irreducible, mixing, entropy, ideal class and period, apply to a G-SFT X as they apply to X as an SFT. 2.2.3 Skew products and matrices over Z+G By ZG we mean the integral group ring of G. We write an element x of ZG as x = summationtextg?Gngg, where each ng ? Z. Then for each g in G we define pig(x) = ng. If pig(x) > 0, then g is a summand of x. If pig(x) > 0 for each g in G, then we say x is very positive and write x greatermuch 0. The augmentation of x is |x| = summationtextg?Gpig(x). If A is a matrix over ZG, then A greatermuch 0 if AIJ greatermuch 0 for each entry AIJ of A. The augmentation |A| is the matrix given by |A|IJ = |AIJ| for each entry AIJ of A. We 13 let Z+G = {x?ZG: pig(x) ? 0 for each g ?G}. If G is a directed graph and l is a labeling of the edges of G by elements of G, then we say (G,l) is a G-labeled graph. If A is a square matrix over Z+G, then |A| is a square matrix over Z+ which, as before, is the adjacency matrix for a directed graph G|A|. The matrix A corresponds to a G-labeled graph (G|A|,lA), where lA is defined as follows: for each pair I, J of vertices in G|A|, AIJ = summationtextngg if and only if for each g ? G exactly ng of the edges from I to J are lA-labeled g. The edge labeling lA determines a function ?A: ?|A| ? G by ?A(x) = lA(x0) for each x = (xn)n?Z in ?|A|. The function ?A is locally constant: for each x ? ?|A|, ?A is constant on a neighborhood of x (here ?A is constant on {y ? ?|A||y0 = x0}). The function ?A is the skewing function defined by A. Given two locally constant functions ?1,?2: ?|A| ?G, we say ?1 is cohomologous to ?2 if there is another locally constant h: ?|A| ?G such that ?1(x) = [h(?x)]?1 ??2(x)?h(x) for each x? ?|A|. The Z+G matrix A determines an automorphism SA: ?|A|?G? ?|A|?G by SA(x,g) = (?(x),?A(x)?g), where ?A is the skewing function defined by A. We say the dynamical system (?|A| ?G,SA) is the skew product defined by A. There is a free right G action on (?|A| ?G,SA) which commutes with the automorphism SA, given by g: (x,h) mapsto? (x,h?g). Often we write just SA as an abbreviation for the skew product (?|A| ?G,SA) We can present the skew product SA as a free G-SFT (which we also denote by SA) as follows. As an edge SFT SA has graph G, where the vertex set of G is the product of the vertex set of G|A| with G, and for each edge e from I to J in G|A|, for each g in G, there is an edge from (I,g) to (J,lA(e)?g) in G. For each pair of 14 vertices v, vprime of G we choose an ordering of the edges from v to vprime, and let g in G act by the one block map given by the unique automorphism of G which acts on the vertex set of G by (J,h) mapsto? (J,h?g), and which is order preserving. In this way any skew product is a free G-SFT. Conversely, any free G-SFT is G-conjugate to a skew product SA for some Z+G matrix A. We say a matrix A over Z+G is very primitive if there exists a natural number N such that AN greatermuch 0. One easily checks that A is very primitive if and only if the G-SFT SA is mixing. Square matricesAandB over Z+Gare strong shift equivalent (SSE) over Z+G if they are connected by a string of elementary moves of the following sort: there are R and S over Z+G such that A = RS and B = SR. Parry has shown that A and B are SSE over Z+G if and only if the skew products SA and SB are G-conjugate [16, Prop. 2.7.1]. 2.3 Some useful results In this section we collect some results to be used later. We begin with the known classification of right closing almost conjugacy for irreducible SFTs, which is a corollary of [15, Theorem 7.1]. Theorem 2.3.1. Irreducible SFTs are right closing almost conjugate as SFTs if and only if they have the same ideal class, entropy and period. Lemma 2.3.2 (Z+G Masking Lemma). Let A and C be matrices over Z+G such that the skew product SA is G-conjugate to a subsystem of the skew product SC. Then there is a matrix B over Z+G such that A is a principal submatrix of B, and 15 SB and SC are G-conjugate skew products. Proof. If SA is G-conjugate to a subsystem of SC, then A is SSE over Z+G to a principal submatrix of C [16, Prop. 2.7.1]. Nasu?s original Masking Lemma for matrices over Z [35, Lemma 3.18] also holds for matrices over an arbitrary semiring containing 0 and 1 [13, Appendix 1]; in particular it holds for matrices over Z+G. This means there is a matrix B over Z+G such that A is a principal submatrix of B, and B is SSE over Z+G to C; SB and SC are G-conjugate skew products by [16, Prop. 2.7.1]. Lemma 2.3.3. Let A and B be matrices over Z+G. A G-factor map pi: SA ?SB induces a factor map pi: ?|A| ? ?|B| such that the skewing function ?A is coho- mologous to ?B ?pi. Conversely, if pi: ?|A| ? ?|B| is a factor map such that ?A is cohomologous to ?B ?pi, then pi induces a G-factor map pi: SA ? SB. The G-map pi is 1-1 a.e. and right closing if and only if the map pi is 1-1 a.e. and right closing. Proof. Letpi: SA ?SB be aG-factor map. Writepi = pi1?pi2, so that for an element (x,g) ? ?|A| ?G, pi(x,g) = (pi1(x,g),pi2(x,g)). Let e denote the identity element of G. Then pi: (x,g) mapsto? (pi1(x,e),pi2(x,e) ?g), since pi intertwines G actions. For x ? ?|A|, set pi(x) = pi1(x,e) and h(x) = pi2(x,e), so that pi(x,g) = (pi(x),h(x)?g). Look componentwise at the equality pi?SA = SB ?pi. The first component shows that pi: ?|A| ? ?|B| is a well-defined factor map. The second component shows that ?A(x) = [h(?x)]?1 ?(?B ?pi)(x)?h(x) for each x ? ?|A|. Hence ?A is cohomologous to ?B ?pi. 16 Conversely, suppose pi: ?|A| ? ?|B| is a factor map such that ?A is cohomolo- gous to ?B?pi. Then there is a locally constant map h: ?|A| ?G such that for each x ? ?|A|, ?A(x) = [h(?x)]?1 ?(?B ?pi)(x)?h(x). Define pi: ?|A| ?G ? ?|B| ?G by pi(x,g) = (pi(x),h(x)?g). Observe that pi is a G-factor map. For the last statement of the lemma, consider the following commutative dia- gram, where the mapsqA: SA ? ?A andqB: SB ? ?B are each given by (x,g) mapsto?x. SA pi d47d47 qA d15d15 SB qB d15d15 ?|A| pi d47d47 ?|B| Both maps qA and qB are |G|-to-1 everywhere. Therefore pi is 1-1 a.e. if and only if pi is 1-1 a.e. For the closing condition, note that if ? and ? are maps between irreducible SFTs, then ??? is right closing if and only if both ? and ? are right closing [15, Props. 4.10 and 4.11]. Because the constant-to-one maps qA and qB are in particular right closing [29, Prop. 4.3.4], it follows that pi is right closing if and only if pi is right closing. If (G,l) is a G-labeled graph, then for a cycle s = s1s2...sp in G we define the weight of s by l(s) = l(s1)l(s2)???l(sp). The ratio group ?l is the subgroup of G given by ?l = {l(s)?l(sprime)?1: s,sprime are cycles in G of the same length} Theorem 2.3.4 (ZGReplacement Theorem). Let (G,l) and (Gprime,lprime) be irreducible 17 G-labeled graphs of the same period which define edge SFTs ? and ?prime (respectively) and skewing functions ?: ? ?G and ?prime: ?prime ?G given by ?(x) = l(x0) and ?prime(x) = lprime(x0). Let pi: ? ? ?prime be a factor map such that ? is cohomologous to ?prime ?pi. If ?l = ?lprime, then there is a 1-1 a.e. factor map pi: ? ? ?prime such that ? is cohomologous to ?prime ?pi. Moreover, if pi is right closing, then pi can be taken to be right closing as well. In [5, Theorem 6.1], Ashley proves a version of his (Z) Replacement Theorem for maps between irreducible Markov chains, which can be interpreted as follows. Let R+ denote the multiplicative group of positive real numbers. Let (G,l) and (Gprime,lprime) be irreducible R+-labeled graphs of the same period, which define irreducible SFTs ? and ?prime and locally constant functions ?: ? ? R+ and ?prime: ?prime ? R+ where ?(x) = l(x0) and ?prime(x) = lprime(x0). If the ratio groups ?l and ?lprime are equal (as multiplicative subgroups of R+), and pi: ? ? ?prime is a factor map such that ? is cohomologous to ?prime ?pi, then there is a 1-1 a.e. factor map pi: ? ? ?prime such that ? is cohomologous to ?prime ?pi. Moreover, if pi is right closing, then pi can be taken to be right closing as well. If instead of R+-labeled graphs we consider G-labeled graphs, then we have the statement of Theorem 2.3.4. To prove Theorem 2.3.4, one can easily check that Ashley?s proof for R+-labeled graphs goes through for G-labeled graphs as well. Theorem 2.3.5. Let X and Y be mixing free G-SFTs. Let pi: X ?Y be a G-factor map which is right closing. Then there is a G-factor map piprime: X ? Y which is 1-1 a.e. and right closing. 18 Proof. Since X and Y are mixing free G-SFTs, assume without loss of generality that X = SA and Y = SB for very primitive matrices A and B over Z+G. By Lemma 2.3.3 the G-factor map pi induces a map pi: ?|A| ? ?|B| such that ?A is cohomologous to ?B ?pi. Since A and B are very primitive the periods of G|A| and G|B| are both 1, and furthermore ?lA = ?lB = G. So assume (by Theorem 2.3.4) that the map pi is 1-1 a.e. and right closing. Again apply Lemma 2.3.3 to obtain a G-factor map piprime: SA ?SB which is 1-1 a.e. and right closing. 2.4 Right closing almost conjugacy for mixing free G-SFTs For mixing SFTs, entropy and ideal class are a complete set of invariants of right closing almost conjugacy (Theorem 2.3.1). We show that there are no additional invariants of right closing almost conjugacy for mixing free G-SFTs. Theorem 2.4.1. Let X and Y be mixing free G-SFTs. Then the following are equivalent. 1. X and Y are right closing almost conjugate as G-SFTs. 2. X and Y are right closing almost conjugate as SFTs. 3. X and Y have the same entropy and ideal class. Moreover, assuming (2) or (3), the common extension of X and Y in (1) can be taken to be a free G-SFT. Proof. (2) ? (3) follows from Theorem 2.3.1. Right closing almost conjugate G- SFTs are in particular right closing almost conjugate as SFTs, so (1) ? (2). It 19 remains to show (2) ? (1). Let X and Y be mixing free G-SFTs which are right closing almost conjugate as SFTs. Without loss of generality, assume thatX andY are skew productsSA and SB for very primitive matrices A and B over Z+G. Let lA, lB, ?A and ?B denote the edge labelings and skewing functions defined by A and B, respectively (see section 2.2). Since SA and SB are right closing almost conjugate as SFTs, they have the same entropy and ideal class (Theorem 2.3.1). The factor maps qA: SA ? ?|A| and qB: SB ? ?|B| given by (x,g) mapsto? x are |G|-to-1 everywhere. In particular they preserve entropy and ideal class, so ?|A| and ?|B| have the same entropy and ideal class. Hence ?|A| and ?|B| are right closing almost conjugate as SFTs (Theorem 2.3.1). Let ?|C| be a common extension of ?|A| and ?|B| by 1-1 a.e. right closing factor maps pi1: ?|C| ? ?|A| and pi2: ?|C| ? ?|A|. SA qA d33d33d66d66 d66d66d66 d66d66d66 ?|C| pi1 d124d124d122d122d122d122 d122d122d122d122 pi2 d34d34d68d68 d68d68d68d68 d68d68 SBq B d125d125d123d123d123 d123d123d123 d123d123 ?|A| ?|B| Without loss of generality, assume the factor maps pi1 and pi2 are one block. Define edge labelings l1 and l2 on G|C| by l1 = lA ? pi1 and l2 = lB ? pi2. The labelings l1 and l2 correspond to matrices C1 and C2 (respectively) over Z+G such that |C1| = |C2| = |C|. Define skewing functions ?1: ?|C| ? G and ?2: ?|C| ? G by ?1(x) = l1(x0) and ?2(x) = l2(x0). Define G-factor maps pi1: SC1 ? SA and pi2: SC2 ?SB by pi1(x,g) = (pi1(x),g) and pi2(x,g) = (pi2(x),g). Let q1: SC1 ? ?|C| and q2: SC2 ? ?|C| be the factor maps (x,g) mapsto? x. Then the following diagram commutes. 20 SC1 pi1 d125d125d123d123d123 d123d123d123 d123d123 q1 d34d34d69d69 d69d69d69d69 d69d69 SC2q 2 d124d124d121d121d121d121 d121d121d121d121 pi2 d33d33d67d67 d67d67d67 d67d67d67 SA qA d33d33d66d66 d66d66d66 d66d66d66 ?|C| pi1 d124d124d122d122d122d122 d122d122d122d122 pi2 d34d34d68d68 d68d68d68d68 d68d68 SBq B d125d125d123d123d123 d123d123d123 d123d123 ?|A| ?|B| The factor maps pi1 and pi2 are 1-1 a.e. and right closing, so the factor maps pi1 and pi2 are as well (Lemma 2.3.3). In particular SC1 and SC2 are mixing free G-SFTs, so C1 and C2 are very primitive. Let l be the (G?G)-labeling l = l1 ?l2. Then l corresponds to a Z+(G?G) matrix whose augmentation is |C|. Call this matrix C. Let ?: ?|C| ?G?G denote the skewing function given by ?(x) = l(x0). Claim 2.4.2. There is a vertex I in G|C| and a natural number N such that there is a collection U of paths of length N from I to I with the following properties: 1. for each g in G there are at least |G| paths u? U with weights l1(u) = g. 2. for each g in G there are at least |G| paths u? U with weights l2(u) = g. 3. for each u = u1u2???uN ? U the point xu ? ?|C|, defined by xui = uj if i?j mod N, has least period N. 4. if u and v are distinct paths in U, then xu and xv are in different orbits under the shift. To prove the claim, let ? be the element of Z+G given by ? = summationtextg?Gg. Fix a vertex I in G|C|. Let ? be the number of cycles of length 1 in G|C|, and choose a positive integer k large enough so that k ?? ? |G|. Since C1 and C2 are very primitive matrices there is a positive integer M = M(k) such that, for i = 1,2 and 21 for all m?M, k?? is a summand of (Cmi )II. Let N ?M be a prime number. Let V be the set of all N-paths from I to I. Each v = v1v2???vN ? V defines a point xv ? ?|C| by xvi = vj if i ? j mod N. Since N is prime, each such xv has least period either N or 1. Let V1 = {v ?V : xv has least period 1} and U = V ?V1. It remains to verify that U satisfies the properties of the claim. Note that, for i = 1,2, each monomial summand g of (CNi )II corresponds to a path v ? V with weight li(v) = g. Also, N was chosen so that k?? is a summand of each (CNi )II. So for i = 1,2 and for each g ? G, there are at least k paths v ? V with weight li(v) = g. There are only ? cycles of length 1 in G|C|, so in particular |V1| ??. But k?? ? |G|. Hence, for i = 1,2 and for each g ?G, there are at least k paths u? U with weight li(u) = g, which verifies properties (1) and (2). Properties (3) and (4) are true by construction of U. This proves the claim. Now consider all points xu ? ?|C| such that u ? U. Let ?|C| denote the smallest closed ?-invariant subset of ?|C| containing all points of this form. Then ?|C|?G is a closed SC-invariant subset of ?|C|?G, so it is a subsystem of the skew product SC. Let SC denote this subsystem of SC. Construct a (G?G)-labeled graph (H,lH) as follows. The vertex set of H consists of N vertices, I1,I2,...,IN. For j = 1,2,...,N ? 1, draw exactly one edge starting at Ij and ending at Ij+1, and give this edge the lH-label (e,e), where e is the identity element of G. From IN to I1 draw exactly |U| edges, call them s1,s2,...,s|U|. Let S = {s1,s2,...,s|U|}, and fix a set bijection ?: S ? U. For si ? S, put lH(si) = l(?(si)) = (l1(?(si)),l2(?(si))). LetDbetheZ+(G?G)adjacencymatrixforthe(G?G)-labeledgraph(H,lH). 22 Observe that the set bijection ?: S ? U induces a (G?G)-conjugacy between SD and SC. Assume without loss of generality that D is a principal submatrix of C (Lemma 2.3.2), so that (H,lH) is an induced sub-labeled graph of (G|C|,l). For each g ? G, at least |G| of the edges si ? S have l-labels of the form (g,?), and at least |G| of the si ? S have l-labels of the form (?,g) (by definition). Therefore there is a way to permute the second coordinates of the l-labelings of edges in S such that each (g,h) ? G?G labels at least one si ? S. Equivalently, there exists a graph isomorphism P of G|C| which fixes all edges except those in S, and permutes the set S such that for any (g,h) ? G?G, there is at least one edge si ? S with (l1(si),l2 ?P(si)) = (g,h). Fix a graph isomorphism P with this property and set lprime to be the (G?G)-labeling of G|C| given by lprime = l1 ? (l2 ?P). Let P denote the automorphism of ?|C| induced by P. Let Cprime2 be the Z+G matrix defined by the edge labeling l2 ?P of G|C|. Note that the map ?: SCprime2 ? SC2 given by (x,g) mapsto? (P(x),g) is a G-conjugacy. Let Cprime be the Z+(G?G) matrix defined by the edge labeling lprime of G|C|, and let ?prime: ?|C| ? G?G be the skewing function given by ?prime(x) = lprime(x0). Then SCprime is the skew product (?|C| ? G ? G,SCprime), where SCprime(x,g,h) = (?(x),?prime(g,h)) = (?(x),?1(x)?g,(?2?P)(x)?h), and G?G acts by (k,l): (x,g,h) mapsto? (x,gk,hl). Note that Cprime is very primitive. (This is because, with I = I1 and N as above, (CprimeN)II has as a summand every element ofG?G.) ThereforeSCprime is a mixing free (G?G)-SFT. From now on, regard SCprime as a mixing free G-SFT by restricting the (G?G)- action to the diagonal: let an element g ? G act by (x,h,k) mapsto? (x,hg,kg). Let p1: SCprime ?SC1 be the |G|-to-one factor map (x,g,h) mapsto? (x,g), and letp2: SCprime ?SCprime2 23 be the |G|-to-one factor map (x,g,h) mapsto? (x,h). Note that p1 and p2 are G-factor maps; they are right closing because they are constant-to-one [29, Prop 4.3.4]. This gives a diagram of right closing G-factor maps: SCprime p1 d125d125d123d123d123 d123d123d123 d123d123 p2 d33d33d67d67 d67d67d67d67 d67d67 SC1 pi1 d125d125d124d124d124 d124d124d124 d124d124 SCprime2 d33d33d66d66 d66d66 d33d33d66d66 d66d66 ? pi2 d33d33 SA SB SCprime is a mixing free G-SFT, so by Theorem 2.3.5, the right closing G-factor maps pi1 ?p1 and pi2 ???p2 can be replaced by 1-1 a.e. and right closing G-factor maps. This proves the theorem. 2.5 General mixing G-SFTs In this section we classify right closing almost conjugacy for mixing G-SFTs where the G action is no longer assumed to be free. We will need this generalization to classify the irreducible but periodic case in section 2.6. We begin with a result for faithful G-SFTs, which were defined in section 2.2. Lemma 2.5.1. Any irreducible faithful G-SFT is a 1-1 a.e. right closing G-factor of an irreducible free G-SFT. Lemma 2.5.1 is a corollary of [1, Theorem 3]. If X is a G-SFT, we let HX denote the normal subgroup of G which acts by the identity map. Then X is a faithful (G/HX)-SFT where, for all g ?G and x?X, x?(gHX) = x?g. 24 Theorem 2.5.2. Let X and Y be mixing G-SFTs. Then the following are equiva- lent. 1. X and Y are right closing almost conjugate as G-SFTs 2. X and Y are right closing almost conjugate as SFTs, and HX = HY. 3. X and Y have the same entropy and ideal class, and HX = HY. Proof. (2) ? (3) follows from Theorem 2.3.1. If X and Y are right closing almost conjugate as G-SFTs, then in particular they are right closing almost conjugate as SFTs. Moreover, if Z is a common 1-1 a.e. right closing G-extension of X and Y, then HX = HZ and HY = HZ, because 1-1 a.e. G-factor maps preserve the subgroup HZ. This proves (1) ? (2). Conversely, suppose X and Y are right closing almost conjugate as SFTs, and H = HX = HY. Then X and Y are faithful (G/H)-SFTs, where for all x ? X, y ?Y and g ?G, x?(gH) = x?g and y?(gH) = y?g. Hence there are free (G/H)- SFTs hatwideX and hatwideY, and 1-1 a.e. right closing (G/H)-factor maps ?X: hatwideX ? X and ?Y : hatwideY ? Y (Lemma 2.5.1). Since X and Y are right closing almost conjugate as SFTs, they have the same entropy and ideal class. Since ?X and ?Y are right closing factor maps between irreducible SFTs, they preserve entropy and ideal classes. So hatwideX and hatwideY have the same entropy and ideal class, and are therefore right closing almost conjugate as SFTs. Thus hatwideX and hatwideY are right closing almost conjugate as (G/H)- SFTs, and the common extension can be taken to be a free (G/H)-SFT (Theorem 2.4.1). Let Z be a free (G/H)-SFT with 1-1 a.e. right closing (G/H)-factor maps piX: Z ? hatwideX and piY : Z ? hatwideY. 25 Z piX d127d127 d127d127 d127d127d127 d127 pi Y d31d31d62d62 d62d62d62 d62d62 hatwideX ?X d15d15 hatwideY ?Y d15d15 X Y For all hatwidex ? hatwideX, hatwidey ? hatwideY and g ? G, put hatwidex?g = hatwidex? (gH) and hatwidey?g = hatwidey? (gH). With these G actions, hatwideX and hatwideY are G-SFTs, and ?X and ?Y are now G-maps. For all z ? Z and g ? G, put g ?z = z ? (gH). This G action makes Z a G-SFT as well, and piX and piY are now G-maps. Thus Z together with the maps ?X ?piX and ?Y ?piY give a right closing almost conjugacy between X and Y as G-SFTs. 2.6 The irreducible but periodic case Here we classify right closing almost conjugacy for irreducible but periodic G- SFTs. If (X,?) is an irreducible G-SFT of period p, then we let X0,X1,...,Xp?1 denote the cyclically moving subsets ofX under?. Then for 0 ?n?p?1, (Xn,?p) is a mixing SFT. The (Xn,?p) are pairwise conjugate SFTs and the action of G on (X,?) permutes the (Xn,?p). If the entropy of (X,?) is log(?), then the entropy of each (Xn,?p) is log(?p). The ideal class (in Z[1/?p]) of (Xn,?p) is determined by the ideal class (in Z[1/?]) of (X,?). We let X = X0 and ? = ?p|X. Then as SFTs, X is conjugate to X ?{0,...,p?1}, where the shift for the latter is given by 26 ?(x,n) = ? ???? ??? ? (x,n+ 1) if 0 ?n?p?2, (?(x),0) if n = p?1. (2.6.1) We give toX?{0,...,p?1} theG action which is the image under conjugacy of the G action on X, so that X is G-conjugate to X?{0,...,p?1}. Without loss of generality, we assume from now on that irreducible but periodic G-SFTs are of the form (X,?) = (X ?{0,...,p?1},?), where the shift ? is given by 2.6.1. By Zp we mean the group of integers {0,1,...,p? 1} with addition mod p. The G action on X determines a homomorphism ?X: G?Zp, given by ?X(g) = k if and only if g: (X,0) mapsto? (X,k). We refer to ?X as the action homomorphism for the G-SFT (X,?). Note that for 0 ? n ? p? 1 and for each g ? G, g: (X,n) mapsto? (X,n+?X(g)(mod p)), where the action on the first coordinate is given by some automorphism Ug of (X,?). The first coordinate automorphisms {Ug}g?G define a G action on (X,?), given by g: x mapsto? Ug(x). This G action on X is not necessarily free, even if the G action on X is free. We refer to the G-SFT X as the base G-SFT forX. We point out that baseG-SFTs are mixing, so right closing almost conjugacy of base G-SFTs is classified by Theorem 2.5.2. Theorem 2.6.2. Let X and Y be irreducible G-SFTs. Then the following are equivalent. 1. X and Y are right closing almost conjugate as G-SFTs. 2. The base G-SFTs X and Y for X and Y are right closing almost conjugate as G-SFTs, and the action homomorphisms ?X and ?Y are the same. 27 Proof. Suppose (X,?) and (Y,?) are right closing almost conjugate as G-SFTs. Then there is a G-SFT (Z,?) and 1-1 a.e. right closing G-factor maps piX: Z ?X and piY : Z ? Y. The maps piX and piY preserve period, so Z must have period p, where p is the period of both X and Y. Furthermore Z must be irreducible because X and Y are irreducible. Without loss of generality, assume that Z = Z ? {0,...,p? 1} where Z is the base G-SFT for Z. Further assume (X,0) = piX(Z,0) and (Y,0) = piY(Z,0), where X and Y are the base G-SFTs for X and Y respectively. Observe that for 0 ?n?p?1, piX(Z,n) = piX ??n(Z,0) = ?n ?piX(Z,0) = ?n(X,0) = (X,n). In particular ?X = ?Z (since piX intertwines G actions). Similarly ?Y = ?Z. Let PZ: Z ? Z be the G-factor map (z,n) mapsto? z and let PX: X ? X be the G-factor map (x,n) mapsto? x. Since piX(Z,n) = (X,n) for 0 ? n ? p? 1, there is a G-factor map piX: Z ?X which makes the following diagram commute. Z piX d47d47 PZ d15d15 X PX d15d15 Z piX d47d47X piX is 1-1 a.e. and right closing because piX is. Similarly construct a 1-1 a.e. right closing G-factor map piY : Z ? Y. Then X and Y are right closing almost conjugate as G-SFTs. 28 Conversely, supposethebaseG-SFTs(X,?)and(Y,?)arerightclosingalmost conjugate as G-SFTs, and ? = ?X = ?Y. In particular, X and Y have the same period p. Let (Z,?) be a G-SFT with 1-1 a.e. right closing G-factor maps piX: Z ? X and piY : Z ? Y. Let Z = Z ? {0,...,p ? 1} with the shift defined as in 2.6.1. Define a G action on Z by g: (z,n) mapsto? (z?g,n+?(g)(mod p)). Define maps piX: Z ? X and piY : Z ? Y by piX(z,n) = (piX(z),n) and piY(z,n) = (piY(z),n). Then piX and piY are G-factor maps. They are 1-1 a.e. and right closing because piX and piY are. 2.7 Regular isomorphism of G-Markov chains Let (X,?) and (Y,?) be irreducible Markov chains with Markov measures ? and ?. Let ? and ? be the time zero partitions of X and Y, respectively. Consider the past?-algebras?? =logicalortext?n=0?n? and?? =logicalortext?n=0?n?. Then (X,?) and (Y,?) are regularly isomorphic if there is a measurable isomorphism ?: (X,?) ? (Y,?) such that ??1(??) ???N?? = ?? ???1????????N?, ?(??) ???N?? = ?? ???1????????N? for some non-negative integer N. The idea of regular isomorphism was introduced and studied by Parry, first in [20] and also in [37]. For a regular isomorphism ? (in contrast to an arbitrary measurable isomorphism), to code the present (?x)0, it 29 suffices to know the past and a bounded look into the future x(??,N]. Boyle and Tuncel [17] show this measurable coding relation has a more finite and continuous formulation, as follows. Theorem 2.7.1. Irreducible Markov chains (X,?) and (Y,?) are regularly isomor- phic if and only if there exists an irreducible Markov chain (Z,?) and 1-1 a.e. right closing factor maps piX: (Z,?) ? (X,?) and piY : (Z,?) ? (Y,?). AG-Markov chain is a Markov chain (X,?) such thatX is aG-SFT and?is a G-invariant Markov measure onX. Say that irreducibleG-Markov chains (X,?) and (Y,?) areG-regularly isomorphic if there is a regular isomorphism?: (X,?) ? (Y,?) such that ? is G-equivariant. By Theorems 2.4.1 and 2.7.1 we have the following. Corollary 2.7.2. Mixing free G-Markov chains (X,?X) and (Y,?Y), with unique measures of maximal entropy ?X and ?Y, are G-regularly isomorphic if and only if (X,?X) and (Y,?Y) are regularly isomorphic as Markov chains. In the general irreducible case, G-regular isomorphism with respect to mea- sures of maximal entropy can be classified in terms of the invariants of Theorem 2.6.2. 30 Chapter 3 Weak equivalence for shifts of finite type 3.1 Introduction In [9], Beal and Perrin introduce an equivalence relation on subshifts called weak equivalence, and classify a special collection of irreducible shifts of finite type (SFTs) up to weak equivalence. We classify arbitrary SFTs up to weak equivalence. For this, first we prove an extension theorem: Given a homomorphism from the nonwandering set of an SFT S to the nonwandering set of an SFT T, we give necessary and sufficient conditions for the existence of an extension of the given homomorphism to a homomorphism from S to T. We use this extension result to give necessary and sufficient conditions for the existence of a homomorphism from one SFT to another. Extension results are significant in symbolic dynamics. Boyle?s extension re- sult [10, Extension Lemma 2.4] is the key to characterizing when there exists an epimorphism from one mixing SFT onto another of lower entropy, and has had other applications as well. (For example, a refinement of that result, [17, Theorem 5.3], establishes that the Markovian property of a homomorphism is in a certain sense not local.) The extension theorem for inert automorphisms of Kim and Roush [25] is a central tool in the study of automorphisms of an SFT. (See [12], [27] and [28].) The study [7] provides a variety of surjective extension results. It follows from 31 the work of Maass [34] that an open problem on extensions, Problem 3.6.8 below, is very closely related to the problem of characterizing the limit sets of stable cellular automaton maps. As a consequence of Lightwood?s machinery for extending certain classes of homomorphisms of Z2 SFTs, we know that for a large class of examples the key issue for existence of an embedding between Z2 SFTs is the existence of a homomorphism between them [31, 32]. This provides some additional motivation for our existence result in the more tractable d = 1 case. We thank Danrun Huang for calling our attention to the paper of Beal and Perrin, and we thank Mike Boyle for many helpful discussions about this problem. 3.2 Preliminaries 3.2.1 Shifts of finite type We assume some familiarity with shifts of finite type. See [29] and [33] for more complete background. Given n > 0, let ?n = {0,...,n? 1}Z denote the set of doubly-infinite se- quences (xi)i?Z where each xi ? {0,...,n ? 1}. Give to ?n the product of the discrete topology on {0,...,n?1}. Then ?n is compact and metrizable. The shift is the homeomorphism ? : ?n ? ?n given by ?(x)i = xi+1. The pair (?n,?) (or just ?n for short) is the full n-shift. A full shift is a full n-shift for some n. If S is a closed, ?-invariant subset of a full shift, then the pair (S,?) (or just S for short) is a subshift. 32 A word on {0,...,n?1} is a finite concatenation w1...wk, where each wi ? {0,...,n?1}. Given a subshift S, a point x ? S, and coordinates i 0. If there exists a uniform L> 0 such that AL(i,j) > 0 for all 1 ?i,j ?n, then A and G(A) are primitive . If A is primitive, then ?A is mixing. A point x ? ?A is periodic if there exists p > 0 such that ?p(x) = x. In this case p is a period of x. When A is irreducible, we define the period of A, G(A) and ?A to be the greatest common divisor of the set of periods of the periodic points in ?A. Note that for primitive A, the period of A is 1. Given a path w = w0???wl?1 of edges in G(A), the length of w is |w| = l. Denote by in(w) and ter(w) the initial and terminal vertices of w, respectively. If A is irreducible, say that vertices vi and vj in V(A) are period equivalent if there is a path w in G(A) with in(w) = vi and ter(w) = vj such that |w| is divisible by p = the period of A. This defines an equivalence relation on V(A), and induces a partition of V(A) into a disjoint union of p sets V(A) = V0(A)unionsq???unionsqVp?1(A). The Vi(A) are called the cyclically moving vertex sets of V(A). This partition of V(A) induces a partition of ?A into p sets ?A = ?0A unionsq???unionsq?p?1A , 34 where ?iA = {x ? ?A : in(x0) ? Vi(A)}. The ?iA are called the cyclically moving subsets of ?A. By re-labelling if necessary, we may assume that?(?iA) = ?(i+1) mod pA for 0 ?i?p?1. Each ?iA is both invariant and mixing under ?p. 3.2.4 Reducible shifts of finite type Let A be as in section 3.2.2. The nonwandering set of ?A, denoted NW(A), is the closure of the set of periodic points in ?A. The nonwandering set NW(A) is an SFT, and is a disjoint union NW(A) = C(A)1 unionsq???unionsqC(A)K, where each C(A)i is an irreducible SFT. The C(A)i are called the irreducible com- ponents of ?A. If K > 1, then ?A is reducible. For 1 ? i ? K, let pi denote the period of C(A)i. Then C(A)i is partitioned into cyclically moving subsets C(A)0i,...,C(A)pi?1i , each of which is both invariant and mixing under ?pi, as in Section 3.2.3. Therefore we can find li ? 1 such that if q and qprime are any two vertices in the same cyclically moving subset of the graph G(C(A)i), then there exists a path w in G(C(A)i) of length |w| = pi ?li such that in(w) = q and ter(w) = qprime. Such li is called a cyclic transition length for C(A)i. 3.3 Homomorphisms between irreducible SFTs Let S and T be subshifts. Write S PER?? T if, for each periodic point x ? S of period p, there is a periodic point y ?T of period q, such that q divides p. 35 Theorem 3.3.1. Given irreducible SFTs S and T, there exists a homomorphism f : S ? T if and only if S PER?? T. Moreover, assuming S PER?? T, then given any pair of cyclically moving subsets S0 and T0 in S and T respectively, f : S ?T may be chosen so that f(S0) ?T0. Proof. If f : S ?T is a block code and x?S is a periodic point of period p, then f(x) ?T must be a periodic point of some period dividing p, by the ?-equivariance of f. Conversely, first suppose S PER?? T and T is mixing. In this case, let S be the empty subshift, and let f : S ?T be the empty homomorphism. By [10, Extension Lemma 2.4], f extends to a homomorphism f : S ?T. Now supposeS PER??T whereT is no longer assumed to be mixing. Decompose S and T into disjoint unions of cyclically moving subsets, each of which is invariant and mixing under the pth power of the shift, where p is the period of S (the period of T divides p by assumption). Fix cyclically moving subsets S0 of S and T0 of T, and let X = (S0,?p) and Y = (T0,?p). Then X PER?? Y and Y is mixing so, by the previous paragraph, there exists a homomorphism f0 : X ? Y. Every other cyclically moving subset of S is equal to ?i(S0) for some unique 1 ? i < p. So define a homomorphism fi on ?i(S0) by fi = ?i ?f0 ???i, and define f : S ?T by f(x) = fi(x) where x??i(S0). 36 3.4 Extending homomorphisms defined on a nonwandering set Let ?A be a reducible edge shift. Write the nonwandering set NW(A) = C(A)1unionsq???unionsqC(A)K as a disjoint union of irreducible components, and let pi denote the period of C(A)i. Fix in each C(A)i one cyclically moving subset C(A)0i, and enumerate the remaining cyclically moving subsets inC(A)i byC(A)1i,...,C(A)pi?1i , where C(A)ki = ?k(C(A)0i). Given this enumeration of the cyclically moving subsets, for each i,j, define the set of connection paths from C(A)i to C(A)j, denoted CPA(i,j), to be the set of paths in G(A) of the form x0???xt?1 which have initial vertex in some C(A)si, and terminal vertex in some C(A)rj. The number s+t?r, taken mod gcd(pi,pj), is the phase change of the path x0???xt?1. Now define a K?K matrix PA as follows. The entries of PA are sets. Specif- ically, PA(i,j) ? Zgcd(pi,pj) is the set of phase changes of paths in CP(Ai,Aj). Let M(A) denote the set of phase matrices for A. Note that the finitely many possible enumerations of the cyclically moving subsets determine the finitely many matrices in M(A). By a phase matrix for A we mean a matrix PA determined by some enumeration of cyclically moving subsets. Given x and xprime in ?A, say that x and xprime are backwardly asymptotic if there exists k ?Z such that xi = xprimei for all i?k, and forwardly asymptotic if there exists k ?Z such that xi = xprimei for all i?k. Remark 3.4.1. Let PA be a phase matrix for A. Then c ? PA(i,j) if and only if there exist x, y and z in ?A such that 37 ? z is backwardly asymptotic to x and forwardly asymptotic to y, ? x?C(A)si and y ?C(A)rj, and ? s?r ?c mod gcd(pi,pj). It follows that, if ? : ?A ? ?C is an isomorphism which respects the numberings of the irreducible components and cyclically moving subsets, then the associated phase matrices PA and PC are equal. In particular, the set M(A) of possible phase matrices for A is an isomorphism invariant. Example 3.4.2. Let A be the adjacency matrix for the following graph. v11 d11d11 v22 d11d11 v21 d55d55v01 d107d107 d47d47 d47d47 ? d47d47v02 d55d55v12 d107d107 Let C(A)1 correspond to the cycle on the left and let C(A)2 correspond to the cycle on the right. Choose the cyclically moving subset C(A)01 to be those points x ? C(A)1 such that the initial vertex of x0 is v01, and C(A)02 to be those points x ? C(A)2 such that the initial vertex of x0 is v02. Then the phase matrix for this choice is PA = ? ?? ? {0} {2} ? {0} ? ?? ?. Remark 3.4.3. Let D(A) denote the set ofK?K diagonal matricesD, where each D(i,i) is an element of the group Zpi. Note that, if M and Mprime are any two phase 38 matrices in M(A), then there is a matrix D ? D(A) such that Mprime = DMD?1, by which we mean that, for each i,j, Mprime(i,j) = {D(i,i) +s?D(j,j) : s?M(i,j)}. (3.4.4) Conversely, given M ? M(A) and D ? D(A), the matrix DMD?1 is in M(A). Thus the set M(A) is computable via the following finite process: 1. Arbitrarily choose an enumeration of the cyclically moving subsets for A. 2. Construct the phase matrix PA for this choice, as defined above, and include this PA in M(A). 3. For each D ? D(A), add to M(A) the matrix DPAD?1. Now let ?B be another reducible edge shift with nonwandering set NW(B) = C(B)1 unionsq ??? unionsqC(B)L, and let qi denote the period of the irreducible component C(B)i. Suppose there exists a homomorphism f0 : NW(A) ? NW(B). Then f0 induces a set function g : {1,...,K} ? {1,...,L} by the following rule: C(B)g(i) is the irreducible component of ?B which contains f0(C(A)i). In each irreducible component C(B)i, arbitrarily choose one cyclically moving subset C(B)0i, and enumerate the remaining cyclically moving subsets in C(B)i by C(B)ki = ?k(C(B)0i). Let PB be the phase matrix determined by this choice. Say that a phase matrix PA for A is compatible with (PB,f0) if PA is the phase matrix determined by a choice of the cyclically moving subsets C(A)0i in C(A)i such that each f0(C(A)0i) ? C(B)0g(i). For such PA let PA be the matrix such that each PA(i,j) is the set of elements in PA(i,j) taken mod gcd(qg(i),qg(j)). 39 Theorem 3.4.5. Let ?A and ?B be reducible edge shifts, as above, and let f0 : NW(A) ? NW(B) be a homomorphism. Let g : {1,...,K} ? {1,...,L} be the set function induced by f0. Choose a phase matrix PB for B, and let PA be a phase matrix for A compatible with (PB,f0). Then f0 extends to a homomorphism f : ?A ? ?B if and only if each PA(i,j) is contained in PB(g(i),g(j)). Proof. Suppose f0 extends to f : ?A ? ?B. Let c ? PA(i,j) and let c ? PA(i,j) satisfy c ? c mod gcd(qg(i),qg(j)). By Remark 3.4.1 there exist x, y and z in ?A such that ? z is backwardly asymptotic to x and forwardly asymptotic to y, ? x?C(A)si and y ?C(A)rj, and ? s?r ?c mod gcd(pi,pj). Let s (r, respectively) denote s (r, resp.) taken modulo qg(i) (qg(j), resp.). Then ? f(z) is backwardly asymptotic to f(x) and forwardly asymptotic to f(y), ? f(x) ?C(B)sg(i) and f(y) ?C(B)rg(j), and ? s?r ?c mod gcd(qg(i),qg(j)). Hence c?PB(g(i),g(j)). For the converse first note that we may assume, without loss of generality, that f0 is one-block. For, if not, then we can choose an SFT ?C and an isomorphism ? : ?C ? ?A such thatf0??|NW(C) is one-block. The isomorphism?can be taken to 40 respect the numberings of the irreducible components and cyclically moving subsets, which implies PC = PA, by Remark 3.4.1. It follows that f0 extends to all of ?A if and only if f0 ??|NW(C) extends to all of ?C. For the converse, we will use the following. Claim 3.4.6. Suppose each PA(i,j) ? PB(g(i),g(j)). Then there exists M > 0 such that, for each i,j and for each path Q ? CPA(i,j) of length |Q| ? M, there exists a path Qprime ? CPB(g(i),g(j)) such that |Qprime| = |Q|, in(Qprime) = f0(in(Q)), and ter(Qprime) = f0(ter(Q)). To prove Claim 3.4.6, first recall that each C(A)i has a cyclic transition length li. WLOG assume that li is a cyclic transition length for C(B)g(i) as well. (If not, just make li larger.) Let l = max i {li}, and let p = productdisplay i pi. Note that if v and vprime are any two vertices in the same cyclically moving subset of some C(B)g(i), then there is a path in C(B)g(i) of length lp from v to vprime. Next, pick T large enough so that, for each i,j and for each c?PA(i,j), there exists a path Q?CPB(g(i),g(j)) of length at most T with phase change c. Then, for each i,j, pick Ri,j large enough so that all integers r greater than Ri,j and divisible by gcd(qg(i),qg(j)) are contained in the N-ideal N := {mqg(i) +nqg(j) : m?N, n?N}. 41 Let R = maxi,jRi,j, and set M = T +R+ 2lp+ 2max i qg(i). To see thatM satisfies the requirements of Claim 3.4.6, letQ?CPA(i,j) have length|Q| ?M and phase changec. Assume that in(Q) is in someC(A)si and ter(Q) is in some C(A)rj, so that c?s+|Q|?r mod gcd(pi,pj). Let Q1 ?CPB(g(i),g(j)) have phase change c?c mod gcd(qg(i),qg(j)) and |Q1| ?T. Extend Q1 to the left in G(C(B)g(i)) by at most qg(i) and to the right in G(C(B)g(j)) by at most qg(j) to a path Q2 ?CPB(g(i),g(j)) with in(Q2) ?C(B)sg(i) and ter(Q2) ?C(B)rg(j), where s?s mod qg(i) and r ?r mod qg(j). Note that Q2 has phase change c. Also, |Q|?2lp?|Q2| is divisible by gcd(qg(i),qg(j)) and greater than R, so |Q|?2lp?|Q2| ? N . Hence |Q|?2lp = |Q2|+aqg(i) +bqg(j) for some a,b ? 0. So we may extend Q2 to the left in G(C(B)g(i)) by aqg(i) and to the right in G(C(B)g(j)) by bqg(j) to a path Q3 ?CPB(g(i),g(j)) with ? in(Q3) ?C(B)sg(i), ? ter(Q3) ?C(B)rg(j), and ? |Q| = |Q3|+ 2lp. Finally, by choice of l, we may extend Q3 to the left in G(C(B)g(i)) by lp and to the right in G(C(B)g(j)) by lp to a path Qprime ?CPB(g(i),g(j)) with 42 ? in(Qprime) = f0(in(Q)), ? ter(Qprime) = f0(ter(Q)), and ? |Q| = |Qprime|. This completes the proof of Claim 3.4.6 Now define f : ?A ? ?B as follows. Let M > 0 be as in Claim 3.4.6, and let Q ? Qprime be a chosen associated map on connection paths Q of length at least M. Consider the set of paths VWX in G(A) which satisfy 1. V is in a component C(A)i and |V| = M, 2. X is in a component C(A)j and |X| = M, 3. the only C(A)i-vertex of W is its initial vertex, and the only C(A)j-vertex of W is its terminal vertex, and 4. W does not contain any path of length M from NW(A). Let x? ?A. For each k ?Z such that x[k,k+|VWX|?1] is a path satisfying (1)?(4) above, set f(x)[k,k+|VW|?1] = (VW)prime. Elsewhere in x, define f(x)i = f0(x)i. As there is an upper bound on the lengths of paths VWX which satisfy (1)?(4) above, f is a homomorphism. By construction, f0 is the restriction of f to NW(A). 43 3.5 Homomorphisms between reducible SFTs Our main theorem below characterizes when there exists a homomorphism between arbitrary SFTs. Theorem 3.5.1. Let ?A and ?B be reducible edge shifts with nonwandering sets NW(A) = C(A)1 unionsq???unionsqC(A)K and NW(B) = C(B)1 unionsq???unionsqC(B)L. Let PB be a phase matrix for B. Then there exists a homomorphism f : ?A ? ?B if and only if we may choose a set function g : {1,...,K} ? {1,...,L} and a phase matrix PA for A such that, for 1 ?i,j ?K, 1. C(A)i PER??C(B)g(i), and 2. PA(i,j) ?PB(g(i),g(j)). Proof. Given a homomorphism f : ?A ? ?B, let C(B)g(i) be the irreducible com- ponent of ?B containing f(C(A)i). This defines a set function g : {1,...,K} ? {1,...,L}. Then, for 1 ? i ? K, f|C(A)i : C(A)i ? C(B)g(i) is a homomorphism, so C(A)i PER??C(B)g(i), by Theorem 3.3.1. For 1 ?j ?Landi?g?1(j), chooseC(A)0i to be a cyclically moving subset of C(A)i such that f(C(A)0i) ?C(B)0j. Let PA be the phase matrix for A induced by this choice of the cyclically moving subsets C(A)0i in C(A)i. Then PA is compatible with (PB,f0), wheref0 = f|NW(A). By Theorem 3.4.5 eachPA(i,j) ?PB(g(i),g(j)). Conversely suppose we may choose a set function g : {1,...,K} ? {1,...,L} and a phase matrix PA for A such that conditions (1) and (2) of Theorem 3.5.1 are satisfied for 1 ? i,j ? K. By condition (1) and Theorem 3.3.1, there exists a 44 homomorphism f0 : NW(A) ? NW(B) such that C(B)g(i) is the irreducible com- ponent of ?B containing f0(C(A)i). Moreover, by Theorem 3.3.1, f0 may be chosen so that each C(B)0g(i) contains f0(C(A)0i) (i.e. PA is compatible with (PB,f0)). It then follows from Theorem 3.4.5 that f0 extends to a homomorphism f : ?A ? ?B. 3.6 Weak equivalence of SFTs Definition 3.6.1. Let S and T be subshifts. If ? : ?n ? ?m is a homomor- phism between full shifts ?n and ?m which contain S and T respectively, such that ??1(T) = S, then write ? : S ??T. If such a ? exists, write S ??T Definition 3.6.2 (Beal and Perrin). Subshifts S and T are weak equivalent if S ??T and T ??S. A flower shift is an SFT presented by a directed graph made up of loops that all begin and end at a single vertex. If the lengths of the loops of a flower shift S are s := (s1,...,sn), then we define < s >N to be the N-ideal s1N+???+snN. Given flower shifts S and T, which define ideals < s >N and < t >N, Beal and Perrin [9] prove that S and T are weak equivalent if and only if N=N. Proposition 3.6.3. The following are equivalent for subshifts S and T. 1. S ??T. 2. There exists an SFT Sprime containing S, a subshift Tprime containing T, and a ho- momorphism f : Sprime ?Tprime such that f?1(T) = S. 45 Proof. (1) ? (2): Let (f,Sprime,Tprime) be (?,?n,?m) from Definition 3.6.1. (2) ? (1): Choose r ? 2 so that f is r-block. We may assume WLOG that Sprime is a 1-step SFT, which means that a point x is in Sprime if and only if xixi+1 is an allowed word of length 2 in Sprime for each i ? Z. Let ?n?1 and ?m?1 be full shifts containing Sprime and Tprime respectively. Define an r-block homomorphism ? : ?n ? ?m by: ?(x0???xr?1) = ? ???? ??? ? f(x0???xr?1) if x0???xr?1 is allowed in Sprime m?1 if x0???xr?1 is not allowed in Sprime Then ? : ?n ? ?m is well-defined, and ??1(T) = S. Theorem 3.6.4. Let S be an SFT and let T be a subshift. Then S ??T if and only if there exists a homomorphism f : S ?T. Proof. If ? : S ?? T, then f = ?|S is a homomorphism. Conversely, if f : S ? T is a homomorphism, then letting Sprime = S and Tprime = T, condition (2) of Proposition 3.6.3 is satisfied. Hence S ??T. The requirement that S be an SFT can not be removed from Theorem 3.6.4. To see this note that, if S is a non-SFT subshift and T is an SFT, then it can not be the case that S ?? T. However there are plenty of examples of homomorphisms from non-SFT subshifts to SFT subshifts. Corollary 3.6.5. Let S and T be SFTs. Then S and T are weak equivalent if and only if there exist homomorphisms f : S ?T and g : T ?S. 46 Corollary 3.6.5 follows directly from Theorem 3.6.4. Corollary 3.6.5 combined with Theorem 3.5.1 provide a classification of weak equivalence for SFTs. The next proposition is an easy observation which we record to further empha- size the local nature of the weak equivalence relation. For a subshift S, let Wk(S) denote the set of words of length k occurring in points in S. Given n ? N, let Sn denote the largest SFT such that Wn(Sn) = Wn(S). If f : S ? T is an R-block homomorphism between subshifts then, for r ?R, the R-block rule defining f also defines a homomorphism fr : Sr ?Tr?R. Proposition 3.6.6. Suppose ? : S ?? T, and let f = ?|S. Then there exists a positive integer R such that, for all r ?R, fr is well-defined and f?1r (T) = S. Proof. We have ? : ?n ? ?m for some n and m. Choose R so that ? is R-block. Observe that S ? SR ? ?n because SR ? S1 and S1 is the smallest full shift containing S. Also ?|SR = fR because ? and f are defined by the same R-block rule. Therefore ?|Sr = fr for all r ? R. We know ??1(T) = S, so (?|Sr)?1(T) = S and hence f?1r (T) = S for all r ?R. Say that a homomorphism f defined on a subshift S is steady if there is an SFT Sprime containing S such that f is well-defined on Sprime, and f applied to Sprime has the same image as does f. A subshift S is sofic if there exists an SFT R and an epimorphism R?S. Remark 3.6.7. If S ?? T where S is a non-SFT subshift, then there exists a homomorphism f : S ? T which fails to be steady. This is true in particular if S is strictly sofic. To see this, suppose ? : S ?? T and let f = ?|S. If f were steady, 47 then there would exist an SFT Sprime containing S such that f is well-defined on Sprime and f(Sprime) = T. Then, for all r sufficiently large, fr is well-defined and Sr ? Sprime, so f?1r (T) = Sr. But Sr negationslash= S because S is non-SFT, contradicting Proposition 3.6.6. Remark 3.6.7 points to a seeming resemblance between Proposition 3.6.6 and the following difficult open problem. Problem 3.6.8. Give necessary and sufficient conditions for a homomorphism de- fined on a sofic shift to be steady. One motivation for Problem 3.6.8 is to characterize the limit sets of stable cellular automaton maps. These limit sets are the subshifts T for which there exists a cellular automaton map f and a positive integer r such that T = Image(fr) = Image(fr+1). In [34], Maass studies such limit sets, and shows that they are precisely the mixing sofic shifts S which contain a receptive fixed point and which admit a steady epimorphism f : S ?S. 48 Bibliography [1] R. Adler, B. Kitchens and B. Marcus, Finite group actions on shifts of finite type, Ergodic Theory & Dynamical Systems 5 (1985), 1-25. [2] R. Adler, B. Kitchens and B. Marcus, Almost topological classification of finite-to-one factor maps between shifts of finite type, Ergodic Theory & Dynamical Systems 5 (1985), 485-500. [3] R. L. Adler, A. G. Konheim and M. H. McAndrew, Topological entropy, Trans. Amer. Math. Soc. 114 (1965), 309-319. [4] R. Adler and B. Marcus, Topological entropy and equivalence of dynamical systems, Mem. Amer. Math. Soc., 20, No. 219 (1979). [5] J. Ashley, Bounded-to-1 factors of an aperiodic shift of finite type are 1-to- 1 almost everywhere factors also, Ergodic Theory & Dynamical Systems 10 (1990), 615-625. [6] J. Ashley, An extension theorem for closing maps of shifts of finite type, Trans. Amer. Math. Soc. 336 (1993), no. 1, 389-420. [7] J. Ashley, B. Marcus, D. Perrin and S. Tuncel, Surjective extensions of sliding-block codes, SIAM J. Discrete Math. 6 (1993), no. 4, 582-611. [8] J. Ashley, B. Marcus and S. Tuncel, The classification of one-sided Markov chains, Ergodic Theory & Dynamical Systems 17 (1997), no. 2, 269-295. [9] M.-P. Beal and D. Perrin, A weak equivalence between shifts of finite type, Adv. in Appl. Math. 29 (2002), 162-171. [10] M. Boyle, Lower entropy factors of sofic systems, Ergodic Theory & Dy- namical Systems 3 (1983), no. 4, 541-557. [11] M. Boyle, Flow equivalence of shifts of finite type via positive factorizations, Pacific J. Math. 204 (2002), no. 2, 273-317. [12] M. Boyle and U. Fiebig, The action of inert finite-order automorphisms on finite subsystems of the shift, Ergodic Theory & Dynamical Systems 11 (1991), no 3, 413-425. [13] M. Boyle and D. Handelman, The spectra of nonnegative matrices via sym- bolic dynamics, Annals of Math. 133 (1991), 249-316. [14] M. Boyle and D. Huang, Poset block equivalence of integral matrices, Trans. Amer. Math. Soc. 335 (2003), no. 10, 3861-3886. 49 [15] M. Boyle, B. Marcus and P. Trow, Resolving maps and the dimension group for shifts of finite type, Mem. Amer. Math. Soc. 70 (1987). [16] M. Boyle and M. Sullivan, Equivariant flow equivalence for shifts of finite type, by matrix equivalence over group rings, Proc. London Math. Soc. 91 (2005), 184-214. [17] M. Boyle and S. Tuncel, Infinite-to-one codes and Markov measures, Trans. Amer. Math. Soc. 285 (1984), no. 2, 657-684. [18] M. Boyle and S. Tuncel, Regular isomorphism of Markov chains is almost topological, Ergodic Theory & Dynamical Systems 10 (1990), 89-100. [19] M. Boyle and J. B. Wagoner, Positive algebraic K-theory and shifts of finite type, Modern Dynamical Systems and Applications (2004) 45-66. [20] R. Fellgett and W. Parry, Endomorphisms of Lebesgue space II, Bull. Lon- don Math. Soc. 7 (1975), 151-158. [21] M. Field and M. Nicol, Ergodic theory of equivariant diffeomorphisms: Markov partitions and stable ergodicity, Mem. Amer. Math. Soc. 169, No. 803 (2004). [22] N. Friedman and D. Ornstein, On the isomorphism of weak Bernoulli trans- formations, Adv. Math. 5 (1970), 365-394. [23] D. Huang, Flow equivalence of reducible shifts of finite type, Ergodic Theory & Dynamical Systems 14 (1994), no. 4, 695-720. [24] D. Huang, Flow equivalence of reducible shifts of finite type and Cuntz- Krieger algebras, J. Reine Angew. Math. 462 (1995), 185-217. [25] K. H. Kim and F. W. Roush, On the structure of inert automorphisms of subshifts, Pure Math. Appl. Ser. B 2 (1991), no. 1, 3-22. [26] K. H. Kim and F. W. Roush, Topological classification of reducible subshifts Pure Math. Appl. Ser. B 3 (1992), no. 2-4, 87-102. [27] K. H. Kim, F. W. Roush and J. B. Wagoner, Characterization of inert actions on periodic points. I., Forum Math. 12 (2000), no. 5, 565-602. [28] K. H. Kim, F. W. Roush and J. B. Wagoner, Characterization of inert actions on periodic points. II., Forum Math. 12 (2000), no. 6, 671-712. [29] B. Kitchens, Symbolic dynamics: one-sided, two-sided and countable state Markov shifts, Springer (1991). [30] W. Krieger, On the subsystems of topological Markov chains, Ergodic The- ory & Dynamical Systems 2 (1982), 195-202. 50 [31] S. Lightwood, Morphisms from non-periodic Z2 subshifts. I. Constructing embeddings from homomorphisms, Ergodic Theory & Dynamical Systems 23 (2003), no. 2, 587-609. [32] S. Lightwood, Morphisms from non-periodic Z2 subshifts. II. Construct- ing homomorphisms to square-filling mixing shifts of finite type, Ergodic Theory & Dynamical Systems 24 (2004), no. 4, 1227-1260. [33] D. Lind and B. Marcus, An introduction to symbolic dynamics and coding, Cambridge Univ. Press (1995). [34] A. Maass, On the sofic limit sets of cellular automata, Ergodic Theory & Dynamical Systems 15 (1995), no. 4, 663-684. [35] M. Nasu, Topological conjugacy for sofic systems and extensions of auto- morphisms of finite subsystems of topological Markov shifts, Springer Lec. Notes in Math 1342 (1988), 564-607. [36] D. S. Ornstein, Bernoulli shifts with the same entropy are isomorphic, Adv. Math. 4 (1970), 337-352. [37] W. Parry, Notes on coding problems for finite state processes, Bull. London Math. Soc. 23 No. 1 (1991), 1-33. [38] W. Parry, Unpublished communication, 2003. [39] W. Parry, The Liv?sic periodic point theorem for non-abelian cocycles, Er- godic Theory & Dynamical Systems 19 (1999), 687-701. [40] W. Parry and P. Walters, Endomorphisms of a Lebesgue space, Bull. Amer. Math. Soc. 78 (1972), 272-276. [41] K. Schmidt, Remarks on Liv?sic? theory for nonabelian cocycles, Ergodic Theory & Dynamical Systems 19 (1999), 703-721. [42] S. Tuncel, Conditional pressure and coding, Israel J. Math. 39 (1981), 101- 112. [43] P. Walters, Some results on the classification of non-invertible measure- preserving transformations, Lecture Notes in Mathematics 318. Springer, Berlin, Heidelberg and New York (1973), 266-276. 51