Avoshifts
Abstract
An avoshift is a subshift where for each set from a suitable family of subsets of the shift group, the set of all possible valid extensions of a globally valid pattern on to the identity element is determined by a bounded subpattern. This property is shared (for various families of sets ) by for example cellwise quasigroup shifts, TEP subshifts, and subshifts of finite type with a safe symbol. In this paper we concentrate on avoshifts on polycyclic groups, when the sets are what we call ``inductive intervals''. We show that then avoshifts are a recursively enumerable subset of subshifts of finite type. Furthermore, we can effectively compute lower-dimensional projective subdynamics and certain factors (avofactors), and we can decide equality and inclusion for subshifts in this class. These results were previously known for group shifts, but our class also covers many non-algebraic examples as well as many SFTs without dense periodic points. The theory also yields new proofs of decidability of inclusion for SFTs on free groups, and SFTness of subshifts with the topological strong spatial mixing property.
1 Introduction
Let be a finite set called the alphabet and a discrete group. A subshift of finite type or SFT is a set closed under the shift maps , and which is defined by removing exactly the points whose orbit hits a particular clopen set , or equivalently contain a forbidden pattern from a finite set.
It is well-known that one-dimensional SFTs (the case ) behave in a much nicer and uniform way than multidimensional SFTs (). As two examples of nice dynamical properties that hold in one dimension, topologically transitive subshifts of finite type in one dimension have a unique measure of maximal entropy and have dense periodic points (these fail for general subshifts of finite type, but in completely manageable ways). Furthermore, SFTs (and even their subshift factors called sofic shifts) can be algorithmically compared for equality and inclusion. See e.g. the standard reference [16].
In higher dimensions, even under strong irreducibility (a strong mixing/niceness assumption), subshifts may have multiple measures of maximal entropy [7], -SFTs are not known to have dense periodic points (see e.g. [9]), and it is not even possible to algorithmically tell whether a given subshift of finite type is empty [4]. For this reason, a recurring endeavor in the field of symbolic dynamics is to try to find classes of multidimensional subshifts which capture some of the interesting phenomena from the two-dimensional setting, but where at least some properties are decidable or otherwise understandable.
Group shifts – groups where the elements are points of the subshift, and group operations are continuous and shift-commuting – are the most successful such a class. They are of finite type when the shift group is polycyclic [31]. They have dense periodic points at least on abelian groups [12]. They also of course have a ``most uniform'' measure, namely the Haar measure. Many of their basic properties (inclusions, and even some dynamical properties) are decidable [3]. An extensive theory of group shifts (with emphasis on the abelian case) is developed in [31].
In [25], the author introduced the so-called -TEP subshifts. They share some properties of group shifts and one-dimensional SFTs: there is a natural measure (the measure that ``samples patterns uniformly on convex sets''). They have dense periodic points, and there are also some decidable properties, for example inclusion is easily seen to be decidable for this class.111The last results are not explicitly stated in [25], but are obvious corollaries of results proved therein; in any case, they follow from the results proved in the present paper.
Different mixing assumptions can have useful implications also in higher dimensions (even if usually not as nice as in one dimension). Strong irreducibility implies dense periodic points at least for two-dimensional SFTs [14, 15]. Two stronger mixing properties are the existence of a safe symbol, and the more general topological strong spatial mixing. Here, we say is a safe symbol for an SFT if turning a non- symbol to in a valid configuration can never introduce a forbidden pattern, for the latter property see Definition 4.8. All SFTs in these classes dense periodic points, and thus their languages and inclusion are decidable.
In the present paper, we define a new class of subshifts called avoshifts. This class is defined with respect to a family of subsets of the acting shift group , and the definition is that if a partial configuration (with in the specified class of shapes) extends to a complete configuration of the subshift, then the set of valid extensions to the identity element is determined by a finite subpattern of the configuration, which is bounded as a function of , but not necessarily .
We consider the class of avoshifts for a quite restricted222This is a good thing – the smaller the class of shapes, the larger the class of avoshifts. class of shapes called inductive intervals. With this choice, avoshifts generalize (cellwise quasi-)group shifts, -TEP subshifts (on with the standard convex geometry). It also covers all SFTs with a safe symbol, and topological strong spatial mixing in fact precisely coincides with the avo property for the class of all possible shapes. Inductive intervals make sense on any polycyclic group, and depend on the chosen subnormal series.
The following is our main ``practical'' result.
Theorem 1.1.
Let be polycyclic, and let be an avoshift for inductive intervals. Then
-
•
is of finite type,
-
•
the language of can be computed algorithmically (uniformly in ),
-
•
can be decided for any given SFT ,
-
•
forbidden patterns for the restrictions of to the groups in the subnormal series (a.k.a. projective subdynamics) can be effectively computed,
-
•
``avofactors'' can be effectively computed,
-
•
is uniformly SFT on inductive intervals, and the corresponding forbidden patterns can be computed, and
-
•
we can algorithmically semidecide that is an avoshift from its forbidden patterns.
Avofactors are factors that can be expressed as projections from a product that satisfies a similar property as that defining avoshifts. This includes all factors given by algebraic maps between group shifts.
The first five results (see below for a discussion of the sixth) were previously known for group shifts (at least for abelian groups). The first is proved in [13, 31] (in the former on , in the latter for general polycyclic groups), the second and third are from [12], and the fourth and fifth are from [3].
An important takeaway are that avoshifts on polycyclic groups satisfy the basic decidability properties, and SFTness, of group shifts. Notably, our proof of these results uses the group structure ``only once'', and the only fact used is ``equal extension counts'' (see Section 3.1 for the definition) for all finite patterns of the same shape, which is trivially true for (cellwise quasi)group shifts. For group shifts, both [12, 3] strongly use Wang's algorithm [32], which is based on the density of periodic points. Avoshifts do not have dense periodic points, and our methods are not related to Wang's algorithm in any obvious way.
The fact that avoshifts are of finite type does not seem to generalize beyond polycyclic groups [24], but the other properties are really consequences of the uniform avo property (meaning the set of valid extensions of to the identity element is determined by a subpattern which is bounded as a function of and ), which in the case of polycyclic groups and inductive intervals is automatic. We illustrate this by giving an avoshift proof of the well-known fact that SFTs on free groups have decidable languages and decidable inclusion, by showing that SFTs are characterized by the uniform avo property on the tree convex sets from [25].
The sixth property in Theorem 1.1 means that there is a finite set of forbidden patterns such that for in the family, configurations that are locally legal on are globally legal (similarly to the main property of -TEP subshifts described in the first bullet point of [25, Theorem 1.1]). This is closely tied with the uniform avo property.
We also make the following two observations that may be of interest despite not having decidability implications: The topological Markov property [1] is equivalent to being avo for the class of cofinite sets (Proposition 4.12). The property of being topologically -mixing for all can be stated equivalently as being uniformly SFT on the sets of cardinality for all .
A lengthy discussion of practicality of these methods, possible extensions of the theory, and future work is included in Section 10.
2 Definitions
We have . We denote the power set of a set by . A subset (not necessarily proper) is written . A finite subset is written . We write for the cyclic group on elements, for finite. By we denote the zero vector of for any .
Throughout, will denote a group. Our groups are for the most part finitely generated and countable. The identity element is denoted by or just . We typically think of as acting on itself by left translation. We consider the group to carry a generating set (usually called when a name is needed, but this symbol is not reserved for this purpose) and the resulting left-invariant word metric (typically called ). The ball of radius or -ball is . Specifically this is the metric closed ball of radius around the identity element, but is the metric ball around the element so we do not need special notation for it. We usually have in mind the right Cayley graph with nodes and edges for . Subsets of are sometimes called areas or shapes. For we write .
An alphabet is a finite (discrete) set . A subshift is which is topologically closed ( having the product topology) and closed under the shift or translation action of defined by formula , where in turn is the notation for indexing at . Note that is compact, and thus so are subshifts.
The subsets also carry the standard Fell topology (through identification of a set with its indicator function in , this is the same as the topology used for subshifts). Subsets of are also considered as an ordered set with inclusion as the order, and means .
In this paper, we use the terminology that a point is always an element of for a group , a configuration is any element of where for some group (possibly a point, possibly ), and a pattern is a configuration whose domain is finite. We sometimes use the term ``finite pattern'' (meaning pattern), and ``partial configuration'' (meaning configuration) for emphasis, and a point is sometimes referred to as a ``complete configuration''. A -pattern is pattern with domain . We mostly use to refer to points and configurations, and to refer to patterns.
If and , we write restriction as , as we tend to have non-trivial formulas for the sets that we prefer not to have in a subscript. If , . We say a configuration is a subconfiguration of if and . If is the set of elements of at some distance from the origin, we also call a prefix of . We say two configurations agree on if .
For a symbol and , write for the pattern with . We sometimes identify with the pattern . For two patterns write for the pattern defined by , when such a pattern exists. If and , then is the configuration defined by (equivalently , so shifting a point is a special case of this definition). The empty pattern is the unique pattern with domain .
We say occurs or appears in if is a subconfiguration of for some ; equivalently we say contains . We denote this by . We write (and use the previous three terms as well) if for some . We write for the language of . Note that the empty pattern is in the language of if and only is nonempty.
A subshift of finite type or SFT is defined by , where is clopen. A clopen set is finite union of cylinders where for (the notation means finite subset). The patterns giving the cylinders comprising are called the forbidden patterns. A window for an SFT is any such that there is a set of defining forbidden patterns whose domains are contained in . A window size is any such that is a window. Note that on a group, we can always pick to be the same for all the finitely many forbidden patterns, but it will be necessary later (when we later talk about being SFT on a subset of the group) to allow for the forbidden patterns to have different domains.
If and is a subshift, the -SFT approximation of is the SFT defined by
Equivalently, this is the set of configurations defined by forbidding precisely the patterns which do not appear in .
Suppose a group and a subshift are clear from context. Then we say a configuration is globally valid or globally legal if . In the case where is an SFT, and is a defining set of forbidden patterns, then is locally -valid or locally -legal if no appears in (with omitted if clear from context). Globally valid configurations are of course locally valid, and locally valid points (complete configurations) are globally valid.
If is a subshift and for , then the follower set is the set of configurations such that for some we have , or equivalently (and is well-defined). We refer to configurations as the extensions of to . If is a singleton (indeed even if ), we may simply write the unique element of in this place to refer to an extension.
We say a subshift is topologically mixing if for any nonempty open , for any large enough (in word norm) there is a point with . More generally topological -mixing means for any nonempty open there exists such that for any with for , there exists such that for all . In the introduction we used the term mixing/gluing property; by this we refer to notions that in some (intuitive, informal) sense generalize the idea of topological mixing.
If are subshifts, a conjugacy is a shift-commuting homeomorphism , and we say and are conjugate in this case. A factor map is a shift-commuting continuous surjection and we say covers or is a factor of in this case. A factor of an SFT is called sofic (note that in this paper, a factor is always a subshift).
If itself is a finite group, then a subshift is called a group shift if it is closed under the operations obtained from those of by applying them cellwise. Group shifts are up to conjugacy the same as subshifts on which one can define a group structure by shift-invariant continuous operations [13, 26, 28]. More generally, for any universal algebraic structure on , we can talk about subshifts with this structure by applying the operations cellwise.
3 Avoshifts and basic properties
We begin with the precise definition of an avoshift. This is given for general groups and families of subsets of groups.
Definition 3.1.
Let be a group. Let be a subshift. We say is avo for (or -avo, or a -avoshift) if there exists such that for all we have
In this case we say is determining for , or -determining. Similarly we say is determined by , or -determined. Let . We say is -avo, or a -avoshift if it is avo for all .
Note that to say is -determining is a bit of an abuse of language: -determining means that determines the possible continuations of (patterns with domain) to the identity, not that determines in any way. We use this terminology as it is concise, and hopefully not too confusing.
Example 3.2.
In the case of , one natural family we could use comes from Euclidean convexity. If is the set of intersections where is convex (for simplicity we refer to such subsets of as convex as well), then we take to be the set of all such that also . For this choice, being an avoshift on equivalently means that if we take a convex subset of , and add a new element so that stays convex, then for globally valid patterns on , the set of valid extensions to is determined by looking at only a bounded subpattern of (bounded as a function of ). This is the natural family to use for the -TEP subshifts from [25], discussed in Section 4.
Example 3.3.
The class we concentrate on in the present paper is ``inductive intervals''. In , an inductive interval is defined by including all elements where the last coordinate is in a particular interval of one of the forms
or if it is zero, then the projection to the first coordinates belongs to an inductive interval of . These are a proper subset of the convex sets defined in the previous example, so the corresponding notion of avoshifts covers a larger class of subshifts. A typical inductive interval for is illustrated in Figure 1.
We call these inductive intervals because they are defined inductively, and we prove most of their properties by induction.

Another convenient way of expressing that is -avo is through follower sets.
Lemma 3.4.
A subshift is -avo for with determining set if and only if for all .
Proof.
Suppose first is -avo, and is -determining. If then there exists such that and . Then also and , so . Conversely, if then there exists such that and . Thus, . Of course and . By the -avo property,
so , meaning .
Suppose then . We show is -avo with determining set . Suppose , let for . We need to show
Suppose first . Then , so by the assumption also , meaning there exists with . Conversely, if , then trivially . ∎
Note that does not really play a special role in the definition of the avo property, since is always a subshift, and thus if satisfies for all , then equivalently for all . Sometimes it simplifies life not to have to translate things back to the origin, and we use the preposition ``at'' as follows:
Definition 3.5.
If , then we say is determining for (or -determining) at .
Lemma 3.6.
The follower set function is decreasing in the leftmost argument with respect to the subconfiguration relation.
Proof.
Let . Let and with . Suppose for some . This means there exists such that and . Then , so , concluding the proof. ∎
By the previous lemma, for implies that for all . We obtain the following immediate consequence:
Lemma 3.7.
If is -determining, then is -determining for all such that , and every with , is -determining.
We call such intermediate sets. It is useful to observe that if is -avo, then, since the determining set can be any sufficiently large finite intermediate set, we may simply take to be the set of elements in which are at distance from the identity in the group, for any sufficient . Any such that is -determining is called an avoradius for (the term ``radius'' comes from the theory of cellular automata, see Proposition 4.10). This allows us to express the following notion, which plays a key role later in the paper.
Definition 3.8.
Let be any family of sets. We say is uniformly -avo or a uniform -avoshift if there is a common avoradius for all .
In other words, if we have a configuration with domain that is globally valid, then to know the set of its legal continuations to , it suffices to look at the subpattern on , where does not depend on .
The following characterization of the avo property is mainly used to explain the name, and to point out a connection to a well-known property of SFTs in one dimension, namely that the shift map being open characterizes one-sided SFTs [20, Theorem 1]. For , write .
Lemma 3.9.
A subshift is avo for a set if and only if the projection from to is an open map.
Here, the set is topologized by the relative topology inherited from the product topology on .
Proof.
Suppose is such that for all we have
Note that, as explained above, this property holds for any superset of (contained in ) as well.
Write for the projection. Observe that this map is surjective, since and are both just projections of to different sets of coordinates. Let be an open set. Let , say where . It suffices to show that some open set (in ) containing is contained entirely in .
Since and is open in , there exists such that whenever and , then . Note that this property holds for any superset of (contained in ) as well. Thus, we may assume by possibly replacing both with a larger set. We claim that then the open set (in ) is contained in . To see this, let be arbitrary, meaning and .
We now have where . Thus trivially, . Thus, by the defining property of , . This means there exists such that . Then and , so , concluding the proof.
Conversely, suppose the property fails for all , meaning for some we can find, for all , patterns such that and but . Let be any limit point of . Then clearly and . For the open set , , but none of the points are in , so it is not a neighborhood . Thus, the projection is not open. ∎
Remark 3.10.
The prefix ``avo'' [Avo] refers to ``open'' in Finnish. This and the previous lemma are the source of the term ``avoshift''.
The forward direction of the following result (that avoshifts are of finite type) is based on the classical proof of Kitchens for group shifts being subshifts of finite type [13].
Lemma 3.11.
If , is an avoshift for the sets if and only if it is of finite type.
Proof.
Suppose first that is of finite type. Let be the maximal diameter of a forbidden pattern. Then is determined by : if and then let be any point with . The point , i.e. the point defined by is clearly locally valid, and thus a point of , and , proving the avo property for .
Suppose then that is an avoshift for this set. By assumption, then there is a determining set for , we can take and . We claim that then is equal to its -SFT approximation. Equivalently, we claim that if every subpattern of a point is a translate of a pattern in , then . For this, suppose that indeed is such that for all .
For , consider the intervals . We show by induction that for all . For the basis of induction, our assumption directly implies . Suppose then that is globally valid, and let satisfy . We have in particular since . Since is determining at for , the set is determining at for . Thus it is also determining at for the intermediate set (recall that this means ). Since (because even ) and , we conclude that , in particular is globally valid.
By compactness, we conclude that is globally valid. Since was arbitrary, we conclude that . ∎
The direction that all SFTs are avo is specific to dimension (or rather, it generalizes naturally to free groups, see Theorem 7.7). As avoshifts have nice computational properties, no decidability-theoretically useful notion of avoshift can cover all SFTs in higher dimensions. The direction that avoshifts are SFT on the other hand generalizes to all polycyclic groups, as we will see.
Note that the previous lemma implies that even in the case of , avoshifts need not have dense periodic points, and may have any number of measures of maximal entropy (though not in the topologically mixing case).
3.1 Equal extension counts
In this section, we define subshifts with equal extension counts. This is an intermediate class that contains all group shifts and -TEP subshifts (as we show in Section 4), and is contained in the avoshifts. In the present paper, all results we prove are proved for general avoshifts (or uniform avoshifts), so equal extension counts do not play a direct role in the proofs.
Definition 3.12.
A subshift has equal extension counts for a set if the cardinality does not depend on . If is a family of subsets of , we say has equal extension counts for if it has equal extension counts for all .
The following proof again should remind the reader of Kitchens' argument that group shifts are of finite type.
Lemma 3.13.
If has equal extension counts for a set , then it is avo for this set.
Proof.
A map between compact spaces with fibers of constant finite cardinality is open. Or more concretely, can only decrease as , so this is reached by some finite , and by uniformity must be -determining. ∎
Again, we may define a notion of uniform equal extension counts by requiring that the in the definition such that can be taken for some fixed . Uniform equal extension counts for a family of sets implies uniform avo for the same set, by the previous proof.
4 Examples of avoshifts
Our first example are the group shifts. The proof should again should remind the reader of Kitchens' argument [13] that group shifts are of finite type.
Lemma 4.1.
Let be a finite quasigroup, and suppose is a subshift closed under cellwise application of operations of . Then has equal extension counts for any . Thus, such is an avoshift for .
Proof.
We extend the cellwise quasigroup operations to any patterns with the same domain in the usual way (the operations are defined cellwise, so apply them cellwise in the input patterns). Let be arbitrary. If then where . List the extensions . Let be arbitrary. Then clearly . Since multiplication by a constant in a quasigroup is injective, and the operations are applied cellwise, we obtain that if is injective, and the only possibility is that these patterns differ at . Thus, has at least as many extensions to as . ∎
Recall that group shifts are cellwiseable [13, 26] meaning up to conjugacy the operations of the group can be taken cellwise (the term is from [30]). Thus, this lemma indeed applies to all group shifts up to conjugacy. Quasigroup shifts are not cellwiseable [26], so in the case of quasigroups the assumption that the alphabet itself is a quasigroup may be crucial.
Example 4.2.
The Ledrappier subshift
is an abelian group shift with cellwise group operation addition in the two-element group . In particular, it is a quasigroup subshift with cellwise operations, thus it has equal extension counts. Thus it is an avoshift for the family of all subsets of .
Example 4.3.
Let be any f.g. infinite group. Then there is a group shift which is not uniformly -avo. For example, is not uniformly -avo.
Our next example are the -TEP subshifts. In [25] these are defined in high generality (with respect to any ``-UCP translation-invariant convexoid '' on the shift group ), but we restrict here to the Euclidean case. Let , and let be a set of patterns. A corner of is any such that there exists a linear functional such that for all .
Definition 4.4.
We say is -TEP if for all corners of , for any there are exactly patterns in which agree with on . We say is -TEP if it is defined by forbidden patterns for some -TEP set .
Let be the family of all intersections of real convex sets with , and let be the set of all such that and .
Lemma 4.5.
Let be a -TEP subshift. Then has uniform equal extension counts for , thus is uniformly -avo.
Proof.
This is the restriction of Lemma 5.2 in [23] to the Euclidean case. ∎
Definition 4.6.
Let , and let . We say is a safe symbol for if whenever , and for some we have for all , then .
Proposition 4.7.
Let , and suppose is a subshift of finite type and is a safe symbol for . Then is uniformly avo for .
Proof.
Let be arbitrary. Let be large enough so that is larger than the diameter of any forbidden pattern. Suppose and let . Define a new configuration by
Note that the first two cases intersect in , but so the definitions agree.
We claim that does not contain any forbidden patterns. This is because in every which is the support of a forbidden pattern, agrees with either the configuration or , which are legal by the assumption that is a safe symbol. ∎
Next we consider the topological strong spatial mixing property [5], which generalizes the previous proposition.
Definition 4.8.
A subshift satisfies topological strong spatial mixing or TSSM with gap , if for any disjoint sets such that , and for every and ,
Briceño defines this in the case of in [5], but the definition makes sense in any group. He does not state that are disjoint, but this does not change the definition.
Theorem 4.9.
A subshift has topological strong spatial mixing if and only if it is uniformly avo for .
Proof.
Suppose a subshift has strong spatial mixing with gap . Let be arbitrary, and let . Suppose and with . We should show . Pick large , and take , , , , , . Then the assumptions of TSSM are easy to check and we thus have . Since was arbitrary, we conclude that by compactness.
Next suppose is uniformly -avo with avoradius for all sets . Pick as the gap, and consider any disjoint sets such that , and patterns and such that . Write .
In particular , so the uniform avo property applied for the set at , gives us because meaning from the assumption.
We can proceed by induction, observing that if
then because
and , we can apply the avo property for the set at to deduce
We eventually obtain as desired. ∎
The following is related to Lemma 3.11 where one-sided intervals were used to characterize SFTs. Here is the set of inductive intervals that are restricted to being negative in one axis (as we will see when we give the general definition). The proposition shows that in the two-dimensional case, this does not lead to being an avoshift for the two-sided version of .
Here, a cellular automaton is a shift-commuting continuous function . A neighborhood is such that is a function of (which always exists by continuity). The function is called a local rule for . A cellular automaton is reversible if it is bijective (this is equivalent to any of the following: injectivity, being a homeomorphism, or having a cellular automaton inverse ).
Proposition 4.10.
Let be a surjective cellular automaton. Let be its spacetime subshift where is defined by . Let consist of sets where is an interval of one of the forms and is an interval of one of the forms . Then
-
•
is always a -avoshift,
-
•
if is reversible, then is avo for the sets and for the sets , and
-
•
if and is the cellular automaton with neighborhood and local rule given by the rules and otherwise (taken in ), then is not a -avoshift.
Proof.
For the first item, consider an inductive interval with axis intervals . If , then the local rule of can be used to determine the unique symbol used at from the contents of if has neighborhood . If , any symbol is legal since is just full shift on due to the surjectivity of .
For the second item, if is reversible we can use the same argument for .
For the third item, consider the inductive interval with axis intervals where and , i.e. the set . then we cannot determine from any finite subset of the all-zero pattern what the possible symbols at are (i.e. we cannot locally compute the set of possible symbols at the origina of -preimages of ), since for either symbol can appear, but if we change a to far to the right, then the symbol becomes forced. ∎
It is a nice exercise (but possibly not an easy one) to show that in the previous proposition (even without assuming surjective), the spacetime subshift is an avoshift for (the set of all inductive intervals) if and only if is stable (reaches its limit set in finitely many steps) and is a constant-to-one (equivalently, open) endomorphism of its limit set.
Next, we show that the avo property is also related to the topological Markov property. The reason we include this is that in [1] the authors are able to prove this property for abelian group shifts on many groups, and some interesting properties of group shifts can be deduced by only using the topological Markov property. Specifically, this property suffices to show some interesting measure-theoretic properties. It is a much weaker property than being SFT; even on , strong TMP subshifts need not be SFT, although they are sofic [8], and on , there are uncountably many subshifts with strong TMP, so they are far from even being sofic. Thus one cannot expect decidability results.
Definition 4.11.
Let be a subshift. We say has the topological Markov property if for all there exists containing such that for all with , the point defined by and is in . If there exists such that we can always take , then has the strong topological Markov property.
Proposition 4.12.
A subshift is (uniformly) avo for the family of cofinite subsets of if and only if it has the (strong) TMP.
Proof.
Suppose first that is avo for the cofinite subsets. Let be arbitrary. Write without repetition. For all the set is a cofinite subset of . Thus, for any such there exists a finite set such that and such that
for all . Writing , translating by , and substituting , we deduce
for all . Let . We claim that we can pick in the definition of TMP.
Namely, suppose with . We prove by induction (in finitely many steps) that the point defined by and is in , by showing that is globally valid. For , this is clear since . Now assume is globally valid. We have
from the above, so is a legal symbol for extending to if and only if appears in . But since , we have .
Conversely, suppose we have the topological Markov property, and let be any cofinite set, meaning for some finite . We will show that is -avo. For this, let be as in the definition of TMP for the set . We claim that is -determining. Namely, suppose . We show the non-trivial direction
For this, suppose , say satisfies .
Note that , since . Thus by TMP, the unique point with is in . But clearly , so as desired.
By retracing the proof one sees that the uniform avo property similarly corresponds to strong TMP. ∎
5 Being SFT in an area
In this section, we introduce the idea of a set of configurations in a subset of a group being SFT. Let be a subshift, and let .
We say is SFT on if there is a finite set of finite patterns such that if and only if and none of the patterns in appear as a subpattern in . When and with clear from context, we also say directly that is SFT if there is a finite set of forbidden patterns with domains , such that if and only if , and does not contain a translate of one of these patterns whose domain fits inside (even if might not extend to any subshift on ). If , then we say is uniformly SFT on the sets or simply uniformly -SFT if there is a finite set of finite patterns that define all the restrictions for . In each case a window is a set which can contain the domains of all the defining forbidden patterns, and a window size is such that is a window.
We make some simple general observations about this notion. These results are not used in the following sections. Nevertheless, the first two propositions may be useful for understanding the notion, and the reader may find the third observation of independent interest.
Proposition 5.1.
Every subshift is SFT on any (single) finite area.
Proof.
Let be a subshift, and let . Let be all -pattern that do not appear in . Then trivially defines . ∎
Proposition 5.2.
A subshift is SFT if and only if it is SFT on every (single) cofinite area.
Proof.
Since is cofinite in , a subshift SFT on all cofinite sets must itself be an SFT.
For the other direction, let be an SFT, and let . Let be a symmetric set containing such that a defining set of forbidden patterns exists. Let be the set of all patterns with domains contained in . If , then clearly does not contain an occurrence of any pattern from . Conversely, suppose does not contain any pattern from . Then in particular appears in , thus there exists a legal pattern with . Define by and , so that in fact . If , then is contained in and thus . If , then , so . Thus, does not contain any of the defining forbidden patterns, and we conclude . ∎
Proposition 5.3.
A subshift is uniformly SFT in finite sets of every fixed cardinality if and only if it is topologically -mixing for all .
To clarify the reading, let be the set of sets of cardinality . The left-hand-side of the equivalence states that for every , is uniformly SFT on the sets (but not necessarily uniformly in ).
Proof.
Suppose first that is uniformly SFT in all finite sets of a fixed cardinality . Let be any finite set of patterns that appear in . Let be the sum of cardinalities of their supports. Let be a set of forbidden patterns defining the restrictions for all . Let be the maximal norm of any element of the domain of any pattern in , and let be the maximal norm of the domain of any pattern from any , then the pattern cannot contain any pattern from as long as (since any such pattern can see at most one translate of a pattern , and all of them appear in ). Since was arbitrary, we conclude that is topologically -mixing for all .
For the converse, we show by induction on that if is topologically -mixing for all , then it is uniformly SFT in , the sets whose support is contained in a (not necessarily disjoint) union of -balls. For , this follows from Proposiiton 5.1, since contains only finite sets for any . Suppose now the claim holds up to , and all . We prove the claim or , for an arbitrary .
By -mixing, there exists such that for any many patterns with supports contained in the -ball, and which are globally legal, any union of those patterns whose separation (distance between the centers of the containing -balls) is at least is also globally legal. Thus, we can find a finite set of forbidden patterns such that the set of patterns with support are correctly defined, when can be partitioned into many -balls whose separation is at least . We need to show that we can add enough forbidden patterns so that also the patterns whose domain cannot be partitioned this way are correctly defined.
But note that for any such , we can break its domain into equivalence classes, by putting two of the -balls in the same class when their distance is less than . Each single equivalence class fits into a -ball. Thus, it suffices to apply induction to topological -mixing and this choice of . ∎
5.1 Connections between the (uniform) avo and uniform SFT properties
In this section, We show that
-
•
under minor assumptions on , uniform SFT implies uniform avo,
-
•
under very strong assumptions on , avoshift implies uniformly avo, and
-
•
under intermediate assumptions on , uniform avo implies uniform SFT.
Definition 5.4.
We say a family is good if
Lemma 5.5.
If is uniformly SFT on the sets , and suppose is good. Then is uniformly avo on .
Proof.
Let be a set of finite patterns that defines on the sets . Let be globally valid for . We need to show
for some finite . Pick such that contains the domain of every . Now if , then does not contain any pattern from . Since , neither does . Thus, neither does . Since is good, we have for some , thus defines the restriction as an SFT, thus it defines also since is shift-invariant. This is the domain of , so the locally valid configuration is globally valid in , proving the uniform avo property. ∎
Recall that a well-quasi-order or wqo is a preorder such that every infinite sequence contains a nondecreasing subsequence. We only consider partial orders (in fact, set containment). Equivalently, a wqo is a preorder that is well-founded (there are no infinite decreasing sequences) and has no infinite antichains (infinite sets of incomparable elements).
Lemma 5.6.
Let be any family of subsets of which is wqo under inclusion, and closed under increasing union. Suppose is -avo. Then is uniformly -avo.
Proof.
It suffices to show that a single avoradius works for all . Suppose not, and for each , let be any set that does not admit radius . By the wqo assumption, we may restrict to a subsequence which are increasing as sets, and do not admit radius . Then by closure under increasing union, and thus admits a determining set by the -avo assumption. For all large enough , we have . Thus, is also a determining set for the intermediate set . Taking also large enough so that , we conclude that is an avoradius for , a contradiction. ∎
If is a set where an order is understood from context, then . Note that typically with this notation one has , but not here.
Definition 5.7.
Let be a family of subsets of . A -well-ordering of is a well-ordering of such that for each , the set is in . If has a -well-ordering then we say is -constructible. A family is constructible if every admits a -well-ordering, and the group also admits one.
Lemma 5.8.
Let be a subshift that is uniformly -avo for with avoradius . Then whenever is -constructible, is SFT with window size .
Proof.
If is empty, the claim is trivial as we may forbid the empty pattern to define its restriction to any set .
Consider now any constructible set . Let be the set of all patterns that do not appear in configurations of , and whose domain is contained in the -ball. Of course if , then does not contain any of these patterns. Suppose then that does not contain any of these patterns. We will show that is the restriction of a point in .
Take a -well-ordering of , and recall the notation . Note that is isomorphic to an initial segment of the ordinals, and also isomorphic to (ordered by containment) under , with limit ordinals corresponding to increasing unions. We show by induction along that is globally valid in . This is obviously true for the minimal element of , since is the empty pattern, and is nonempty. For limit ordinals, this follows immediately from compactness of (an increasing union of globally valid patterns is globally valid).
For successor ordinals , say with predecessor for , we are dealing with a -prefix with maximal element , such that is globally valid and . By shift-invariance of , also is globally valid where . Since is a radius for , the set is determining for .
Since (this is just another way to say that is globally valid), the definition of a determining set states that
We have indeed , since is contained in , and is a pattern that appears in (since does not contain any pattern from ). Thus, .
Shifting back by , we conclude that is globally valid, concluding the induction step. ∎
6 Polycyclic groups and inductive intervals
Let be polycyclic. Fix a sequence of subgroups each normal in the next, so that quotients are cyclic (the existence of such a sequence is the definition of a polycyclic group). Pick so that generates , equivalently . Let .
Definition 6.1.
The are called a polycycle structure for .
Of course and are determined by the choices of , but it is convenient to always have refer to this data. Each by default inherits a polycycle structure from , by restricting to an initial subset of the . We always consider a polycyclic group as carrying a fixed polycycle structure. We call the size of the structure. The various are informally referred to as axes or dimensions, and a dimension is called finite or infinite depending on whether .
Note that the size is not equal to the standard Hirsch length of the group, which only counts infinite dimensions. A polycyclic group is called strongly polycyclic if all the are infinite. Every polycyclic group is virtually strongly polycyclic, and by a bit of recoding it should be possible to remove the finite dimensions from the following discussion completely, but the theory should be more easily applicable if they are allowed, and they do not add much length to the discussion.
Each element of corresponds uniquely to a tuple where for all such that we have . The tuple is given inductively for : The last coordinate is a direct projection to , then we turn it to by multiplying by a power of from the right, and extract the tuple for inductively (using its inherited polycycle structure) to get the first coordinates. Conversely we write the element corresponding to a tuple as . We are typically only interested in the value of the last nonzero element of the tuple, which is independent of the choice of the (it only depends on the ).
We say an interval (possibly infinite in one directions) grazes if it is of one of the forms . It is positive if it is one of the last two, and negative if it is one of the first two. This is the sign of the interval. The empty set has no sign zero, say. In the case of a finite cyclic group , intervals are images of intervals on under projection. Then an interval graces if it is empty, or it contains one of or (or both), but not .
Definition 6.2.
Let be a polycyclic group with a polycycle structure as above. Its inductive intervals, or IIs, are defined inductively as sets of elements whose tuple satisfies that either
-
•
comes from a particular interval gracing in , or
-
•
and comes from a certain inductive interval in .
Concretely, an II is thus determined by an -tuple of intervals where is a -gracing interval in (of finite) or (if ). The are called axis intervals. Then if for some , and for the maximal such we have .
For the group , we by default use the polycyclic structure where is the th standard generator with in the th position. Again, Figure 1 illustrates an inductive interval for the group with this choice. For example, since the last axis interval is and this axis points upward, the ground below the player is completely included in the set, all the way until ; but on the first axis (pointing right), we only fill finitely many positions.
6.1 Main properties of inductive intervals
Lemma 6.3.
Inductive intervals are a good family.
Proof.
If an inductive interval is given by axis intervals where or then is given by where or , and then is an inductive interval, where is the generator for the first axis. The same is true if . If or , then is an inductive interval. ∎
We show that the family of inductive intervals has good order-theoretic properties, and that polycyclic groups can be ordered so that all prefixes of the order are also inductive intervals up to translation. These are the properties that later allow us to conclude that II-avo subshifts are SFTs, as well as computable properties.
Lemma 6.4.
Inductive intervals form a well-quasi-order under inclusion.
Proof.
Consider a sequence of inductive intervals . If some inductive interval appears infinitely many times, we are done. If some takes on only finitely many intervals, we may restrict to the infinite subsequence where this interval is fixed. Next, we may restrict to an infinite subsequence such that each is of the same sign, i.e. grazes from the same side. By symmetry, we may assume that is empty or for all . We may assume that, for all , if does not stay fixed, it becomes a longer and longer prefix of . The sequence of IIs we end up with is then increasing under set containment.
For the latter claim, observe that -signed inductive intervals are a subfamily of the inductive intervals. ∎
Lemma 6.5.
Inductive intervals are closed under increasing union.
Proof.
Let be an increasing sequence of inductive intervals. This means just that each sequence of intervals is increasing in . The union is easily verified to be . Again the same is true automatically for -signed inductive intervals. ∎
In particular it follows from the above lemmas that the (-signed) inductive intervals are also topologically closed in .
Lemma 6.6.
The inductive intervals are topologically closed in .
Proof.
If we have a converging sequence of inductive intervals, the previous lemmas show we can find an increasing subsequence that converges to an inductive interval. Thus the sequence itself converges to an inductive interval. ∎
Lemma 6.7.
The family of inductive intervals is constructible.
Proof.
We start with the claim that inductive intervals are constructible. We now prove the statement of the lemma by induction on the dimension . For , the claim is easy, simply enumerate prefixes of the unique axis interval in the case of an inductive interval, and for the entire group , enumerate first the nonnegative numbers, and then the negative numbers (for example).
For general , first consider an inductive interval . Up to symmetry we may assume that for all nonempty , i.e. the intervals are not negative. Let be the maximal element of the last axis interval (or if and the interval is ). Pick a -well-ordering of the group by induction. Shift this well-ordering to for by picking any base point. Then order by first comparing the power of , and then in case of equality. This is a -well-ordering, since for any individual element the translated downset is of the form where is a prefix of the order. Since is an inductive interval on , is one in .
At this point we have -well-ordered the subset with axis intervals . Next, we well-order by induction, as an inductive interval on ). We add this at the end of the well-order described in the previous paragraph. The order remains a -well-order, since the new translated downsets are those of , with all of included on the last axis.
In the case of the entire group , mix the previous idea and the case of : first list the positive cosets, and then the negative ones. ∎
Definition 6.8.
If is such that admits a well-order such that is in for all , then we say is -extendable. We say a family of shapes is extendable if every is -extendable.
Lemma 6.9.
The inductive intervals are extendable.
Proof.
Consider an inductive interval with axis intervals . Begin ordering the complement by adding elements on the sides of in, say, alternating, order. It is easy to see that up to a shift, the prefixes of the order, together with , are of the form for larger and larger intervals with alternating signs, and after adding all elements of the first axis this way, we have ordered a translate of where now is with one new element. We can now order a new coset of the first axis (on either side of ). We can continue similarly up the dimensions to order (with order type at most ). ∎
7 SFTness and the avo property in concrete examples
7.1 Polycyclic groups
Theorem 7.1.
Let be a polycyclic group, and let be a subshift that is avo for the class of all inductive intervals. Then is uniformly SFT on .
Proof.
Suppose is a polycyclic and is avo for inductive intervals . By Lemma 6.4, is a wqo under inclusion, and by Lemma 6.5 it is closed under increasing union. Thus by Lemma 5.6, is uniformly -avo. By Lemma 6.7, is constructible, meaning any admits a -well-order. Then Lemma 5.8 implies that is a subshift of finite type with window size for any of these sets. In other words, is uniformly SFT on . ∎
Corollary 7.2.
Let be a polycyclic group, and let be a subshift that is avo for inductive intervals. Then is SFT.
Proof.
. ∎
For group shifts, the following is shown in [31]. We obtain a new proof.
Corollary 7.3.
Let be a polycyclic group, and let be a quasigroup shift (i.e. is a quasigroup and is closed under cellwise quasigroup operations). Then is SFT.
Proof.
We showed in Lemma 4.1 that is avo for all subsets of . In particular it is avo for iterated intervals, thus it is SFT. ∎
For shift groups beyond polycyclic, one cannot expect to obtain SFTness from an avo property. Namely, every group that has a non-f.g. subgroup admits a group shift which is not of finite type [24]. Group shifts are avo for all sets, so in particular there cannot exist any family of shapes on such a group, which would allow deducing SFTness of avoshifts for that shape. (Finitely-generated groups where all proper subgroups are finitely-generated are known to exist, see e.g. [19], but simple examples are not known.)
For the lamplighter group , we note that there is even a sofic group shift which is not SFT, so ``avo implies SFT'' is not even true for sofic shifts, for any family of sets . (The same can be proved on many other groups with a sufficiently strong simulation theory.)
Proposition 7.4.
On the lamplighter group , there is a sofic group shift (thus an avoshift for all subsets ), which is not SFT.
Proof.
Every group shift is avo for the family of all sets, so it suffices to find a sofic group shift which is not SFT. The two-symbol group shift on is of course not SFT [24]. It is easy to see that also its pullback in the sense of [2] to is not SFT but is still a group shift. By [24], its free extension to any supergroup (i.e. the subshift defined on the larger group by the same forbidden patterns) is also non-SFT but is still a group shift, in particular we get a non-SFT group shift on the lamplighter group. By the simulation theorem of [2], this specific form of subshift is automatically sofic, since the two-symbol subshift on is obviously computable. ∎
7.2 Free group SFTs are avo
We recall a definition from [25].
Definition 7.5.
Let be free on generators . A tree convex set of is such that if and lies on the geodesic between , then if , the ball is contained in . A limit tree convex set is a possibly infinite set with the same property.
Lemma 7.6.
A set is a limit tree convex set if and only if it is a limit of tree convex sets in Cantor space.
Proof.
The condition is clearly closed so limits of tree convex sets are limit tree convex. Let then be limit tree convex. Then satisfies the condition, so is tree convex for all , and is a limit of these sets. ∎
We also need the standard notion of a geodesically convex set: is geodesically convex in a free group if the unique path between any is contained in .
If is a family of sets, the corresponding extension sets are .
Theorem 7.7.
A subshift on a free group is an SFT if and only if it is uniformly avo on the extension sets of limit tree convex sets.
Proof.
Let be a free group. Suppose first that is an SFT. Consider where and are limit tree convex sets. Let be an SFT with forbidden patterns of diameter at most . Recall that is geodesically convex [25]. If there is no geodesic path from which is of length at least , we have a bound on the diameter of and we are done.
Since the radius- ball is a cutset of the group, when determining the possible legal continuations we need not look past any such ball. For any geodesic path of length at least in , the -ball around the central element is contained in since is a limit tree convex set. We conclude that it suffices to look a bounded distance away from the identity element to know the legal extensions, which implies the uniform avo property.
Suppose then that is uniformly avo on the limit tree convex sets. In particular it is uniformly avo with some avoradius on the tree convex sets. By [25], the tree convex sets form a convexoid. This implies that we can order with order type so that every prefix of it is tree convex. Then Lemma 5.8 implies that is a subshift of finite type with window size . ∎
Example 7.8.
On the free group , there is an SFT which is not avo for geodesically convex sets. Consider an SFT over alphabet with the rule that if then . Then is geodesically convex and has avoradius .
7.3 More applications of Lemma 5.8
Lemma 5.8 (which deduces uniform SFT from uniform avo) has implications on groups other than polycyclic ones. It only needs a constructible family of sets, and these are easy to find. We illustrate the result by proving the following result from [5], which applies to all groups.
Proposition 7.9.
Let be a topologically strong spatial mixing subshift on any f.g. group . Then is SFT.
Proof.
Recall that TSSM is equivalent to being uniformly avo for the class of all sets . Obviously is constructible (one can well-order any set in any way). By Lemma 5.8 is SFT with window size for any , in particular this is true for . ∎
Less trivial examples exist of constructible families, for example the convexoids (slightly generalizing the standard convex geometries) considered in [25] have this property. A convexoid is a family of finite subsets of a set which satisfies and , and which have the corner addition property
Lemma 7.10.
Let be a convexoid. Then the family of extension sets of is constructible.
Proof.
Let be the family of extension sets, i.e. all such that and . We check that each and itself are -constructible, i.e. admit -well-orders.
First, we recall that repeated application of the corner addition property for provides an anti-shelling from to , or ordering of such that any prefix of this order, together with , is convex (meaning belongs to the convexoid ). If , then one can verify that an anti-shelling from to is a -well-order.
From the second property in the definition, we can list an increasing sequence of convex sets . Starting from , and concatenating anti-shellings from to for all , one can check that the resulting order of (with order type ) is a -well-order. ∎
Thus for example the extension sets of tree convex sets defined in Section 4 are extendable. We could thus conclude from Lemma 5.8 that the uniform avo property on the free group, for extension sets of tree convex sets, implies the SFT property. Since we already characterized SFTs are the uniformly avo subshifts for the extension sets of limit tree convex sets, this is not worth stating here (in the uniform avo case, taking the limits does not matter). Convexoids for strongly polycyclic groups are built in [25], as well as on some other groups which are not discussed in the present paper.
8 Decidability results
Next, we explain how one can, starting from a finite set of forbidden patterns, prove (algorithmically or in any sufficient proof system) that an SFT is indeed an avoshift for a set of shapes and compute its language and perform comparisons between such subshifts. In the case of inductive intervals on polycyclic groups, we can also compute traces (restrictions to subgroups in the subnormal series). In the next section we explain how one can also compute some factors (avofactors), which in the case of group shifts include all algebraic factors.
A common situation in symbolic dynamics is that we are applying partial algorithms to undecidable problems, and we know that they work on some subclass of the problems, but this class itself is not obviously decidable, we usually want that our algorithm never make an incorrect claim (in addition to making correct claims on the subclass). Namely, this allows us to blindly apply our algorithms without even knowing whether they belong to the class. The algorithm is ``allowed'' to work correctly even work on some instances that are not in this class, or it may never halt, it is simply not allowed to give an incorrect answer. We make this slightly more precise in the following definition.
Definition 8.1.
Let be some countable set of (codings of) problem instances, a subset of these instances (thought of as a nice subclass), and the ``positive instances''. We say an algorithm safely partially solves if, when given , it either eventually outputs ``yes'', ``no'', or never halts. Furthermore, if it says ``yes'' then , and if it says ``no'' then . We say the algorithm safely solves on if it safely partially solves and, given , then it eventually gives an answer (thus indeed the correct answer) if .
This definition readily generalizes from decision problems to problems with more complex output (like computing the language of a subshift).
We define some computability properties for the sets . A family of subsets of is upper semicomputable if we can uniformly in compute a sequence of upper approximations to , which eventually reaches (though we cannot necessarily detect when this has happened). We say such a family is computable if it is also lower semicomputable, meaning we can actually compute the set of finite subsets of sets in (note that this does not necessarily mean we can decide for a given finite set ).
Lemma 8.2.
Let be a finitely-generated group and let be any topologically closed family of shapes. Let be an SFT defined by a finite set of forbidden patterns , such that uniformly defines on all . Then for all there exists such that whenever and is locally -valid in , then is globally valid in .
Proof.
Suppose that the conclusion fails for some fixed . Then we can find for arbitrarily large a shape , and a locally valid pattern (containing no -pattern), such that is not globally valid.
Since is closed, we can pass to a subsequence of that tends to , in which case also tends to . By compactness of Cantor space, we can find a converging subsequence of the patterns that tends to a configuration on which contains no pattern from
By the assumption that defines the restriction , we have . But for large enough , a contradiction. ∎
Lemma 8.3.
Let be a finitely-generated group with decidable word problem, and let be a topologically closed, upper semicomputable, extendable family of shapes. Then there is an algorithm that, given a finite set of patterns defining an SFT on , enumerates all finite families of forbidden patterns that uniformly define the -restrictions of .
In particular, this algorithm recognizes the SFTs that are uniformly -SFTs. Note that here we need not explicitly discuss safety; the algorithm simply will not enumerate any sets if such sets do not exist.
Proof.
Since is SFT and has decidable word problem, we can compute upper approximations to the language. Note also that the equality of SFTs is semidecidable (see e.g. [29]), so we can enumerate the complete list of sets of patterns which define .
Thus it suffices to show that once we have found a set of patterns equivalent to the original, which additionally defines uniformly on , then we can recognize this fact. For this, consider any such and let be given for obtained from the previous lemma for the family , meaning if is a locally -legal configuration, then is globally legal.
We claim that then whenever is -legal for , then has a -legal extension to . Namely, otherwise every extension is actually globally illegal. But this is a contradiction, since clearly then already cannot be a globally valid pattern.
This means that we can (for the fixed ) go through larger and larger , and for fixed we can compute better and better approximations to the set of -sized prefixes of shapes in . Since the limits of shapes from are upper semicomputable, we eventually know the correct set of their subshapes of size at most (though of course we cannot necessarily recognize that we do). At this point, by the assumption on and the discussion of the previous paragraph, no matter what -legal pattern we put on one of these shapes , it will have at least one legal continuation to that does not contain a -pattern.
At this point, the algorithm can conclude that is a set of forbidden patterns defining uniformly on . Namely, if we have a configuration on which has no -pattern, then by -extendability of , we can order so that prefixes of the order together with are (up to a shift) in , and inductively prove along this order that at any point of the process there is an extension containing no pattern from . ∎
Lemma 8.4.
Let be a finitely-generated group with decidable word problem, and let be a topologically closed, upper semicomputable and extendable family of shapes. Then there is an algorithm that, given a finite set of patterns defining an SFT on , safely solves emptiness on all which are uniformly SFT on . If is constructible, it also safely solves emptiness on uniform -avoshifts .
Proof.
If the given SFT is empty, we can recognize this fact. Otherwise, we will try to apply the previous theorem. Since its output is correct on all SFTs (when it gives an output), if it eventually outputs some , we can determine emptiness: The empty pattern is in the language of a subshift if and only if the subshift is nonempty. Since , the only possible reason why the empty pattern would be forbidden is that the empty pattern is directly in . I.e. is empty if and only if . If is uniformly SFT on , then eventually we find can can check this.
It suffices to show the last claim. Indeed, if is constructible and is a uniform -avoshift, then is uniformly SFT on by Lemma 5.8. ∎
The following theorem implies in particular the theorem of [3] that projective subdynamics of group shifts on can be computed effectively.
Theorem 8.5.
Let be a polycyclic group with polycycle structure . Then given an SFT , we can safely compute a set of forbidden patterns that defines uniformly on all inductive intervals if is an II-avoshift. The restriction is an avoshift for all . Furthermore, we can effectively compute the forbidden patterns for .
Proof.
For the first claim, by Theorem 7.1, is uniformly SFT on iterated intervals whenever is an II-avoshift. By Lemma 8.3, we can find a set of forbidden patterns defining it uniformly on these sets, since by Lemma 6.9 the II are extendable (the other properties are obvious).
The second claim (that is an avoshift) is immediate from the definition.
For the third (decidability) claim, since each of the groups is an iterated interval, the forbidden patterns in that admit a translate contained in directly define . Obviously it is decidable whether a forbidden pattern admits such a translate. ∎
Corollary 8.6.
The set of II-avoshifts on a polycyclic group is recursively enumerable.
Proof.
For each finite set of forbidden patterns defining an SFT , we try to compute forbidden patterns that define uniformly on all inductive intervals. If we find such, then is indeed uniformly SFT on the inductive intervals. Since the inductive intervals are a good family, Lemma 5.5 implies that is (uniformly) avo on inductive intervals. ∎
Next, we proceed to show that we can compute the language of an avoshift under suitable assumptions.
Lemma 8.7.
Let be a group with decidable word problem, let be a topologically closed, computable family of sets, and suppose that is -constructible. Then there is an algorithm that, given a finite set of forbidden patterns such that defines an SFT -uniformly, decides the language of the SFT .
This is a promise problem, i.e. we don't care what the algorithm does if does not have the property stated. (But we will only apply this when we actually know has this property.)
Proof.
Our algorithm computes better and better upper approximations to the set of patterns of on finite subsets of using upper semicomputability of the language.
At times, it will declare that it has found the precise set of patterns in the language on a particular domain. We will be sure to only add such deduction rules when they are actually safe (assuming the given property of ), i.e. when the algorithm declares it knows the patterns of a particular shape, this can be trusted to be actually true.
To root the process, the algorithm can check whether the empty pattern is in . If it is, then is empty and we are done. Otherwise, it concludes that it knows the exact set of globally legal patterns on the domain already (i.e. the empty pattern appears in the language). Also, if the algorithm has already deduced the globally legal patterns on a domain , then it knows the globally legal patterns on any , and on any subset of .
The crucial point is that if we know the exact patterns on for , where is larger than the diameter of any domain of a pattern in , then we also know the -patterns. This is because a pattern on is in if and only if it does not contain a -pattern, so if we know the contents of a -pattern on the annulus , then since patterns in cannot see over the annulus, the algorithm just needs to list the possible extensions that do not directly introduce a pattern from , and these will give exactly the globally legal patterns on the extended shape.
We can now prove by transfinite induction on , w.r.t. the -well-order of , that the algorithm eventually knows all -shaped patterns for finite sets whose elements are strictly below , and all their translates. (Stated like this, if the order on has a maximum , then we do not actually deduce the finite patterns containing directly, but any pattern can be shifted so as not to include the maximum.) For limit ordinals , the induction step is trivial. For successor ordinals, suppose is the predecessor and let be any finite set. It suffices to show that the algorithm eventually deduces the correct patterns on .
By induction, we eventually know the exact set of patterns on . If is larger than the diameter of any pattern in and large enough so that , then we note that by the main deduction rule, the algorithm eventually deduces the set of patterns on . After a shift, it deduces them on , and then also the subset . ∎
Lemma 8.8.
Let be a finitely-generated group with decidable word problem. Let be any algorithm that, given an SFT , safely enumerates the language of every if belongs to a class of subshifts . Then given SFTs , the inclusion is safely decidable for pairs such that .
Proof.
The inclusion is semidecidable. If , then there exists a pattern such that . The latter is semidecidable. The former is semidecidable if , because eventually the algorithm outputs the pattern . Since outputs the language of safely, if it outputs then indeed , so if the deduction is made, it is correct even if we do not necessarily have , so is indeed decided safely. ∎
While the previous lemma is not commonly stated as we have here, it is well-known and commonly applied in the case that is the class of subshifts where periodic points are dense. Namely, Wang's algorithm [32, 29] is more or less the proof of precisely this result. Note that here the idea of ``safety'' is quite useful – Wang's algorithm naturally allows use to conclude for also many where periodic points are not dense, since the algorithm that enumerates the subpatterns of periodic points indeed safely enumerates the language of a given SFT when it has dense periodic points.
Theorem 8.9.
The language of an II-avoshift on a polycyclic group is safely computable, uniformly in the SFT. In particular, given two SFTs on a polycyclic group , such that is an II-avoshift, we can safely decide the inclusion .
Proof.
We recall that even in the case , the previous theorem covers some cases Wang's algorithm does not, as it does not require dense periodic points. (Similarly, Wang's algorithm covers many cases ours does not.)
The following is an easy exercise in automata theory, but it may be of some interest that the avoshift technology covers this.
Theorem 8.10.
The language of an SFT on the free group is decidable. In particular, given two SFTs on a free group , we can decide the inclusion .
Proof.
As we showed in Section 4, any SFT on the free group is uniformly avo on the extension set of the limit tree convex sets . Lemma 5.8 says that if is a subshift that is uniformly -avo for with avoradius , then whenever is -constructible, is SFT with window size . The set is constructible since is a convexoid by Lemma LABEL:lem:ConvexoidConstructible.
Next, Lemma 8.3 says that if is a topologically closed, upper semicomputable, extendable family of shapes, then there is an algorithm that, given the finite set of patterns defining an SFT on , enumerates all finite families of forbidden patterns that uniformly define the -restrictions of . Topological closure and upper semicomputability are clear for from the corresponding properties of (which are clear from the definition). Extendability follows from the proof of constructibility in Lemma LABEL:lem:ConvexoidConstructible since we can take for some in the proof. We conclude that such a set exists, and thus the algorithm eventually finds one.
Next, Lemma 8.7 says that if be a topologically closed, computable family of sets, and is -constructible, then there is an algorithm that, from such a set , decides the language of the SFT . Since the family is guaranteed to indeed define uniformly on , the language is decided safely for all SFTs.
Lemma 8.8 then implies that inclusion of SFTs is safely decidable for all SFTs, in other words the problem is simply decidable. ∎
9 Avofactors
We slightly generalize the avoshift concept, to give an avoshift proof for the useful fact that algebraic factors of group shifts can be computed [3]. We give only the outline, as this is analogous to the theory developed above.
Let be any set, and any set of cornered subsets meaning . Then we say is -avo if for any , there is a finite subset of such that for all . In the case of subshifts, we could always take and always translate the cell to be determined to the origin, so we did not discuss corners explicitly.
Now, if we have a family of subsets on a group , then we can extend it to a family of cornered shapes on where , by taking the shapes with corner . This corresponds roughly to the inductive intervals on the group , which on the last axis graze the origin from the right, but now is not a group but an interval. Now we define an avoshift relation to be a subset which is closed, -invariant, and is avo for the cornered set described in this paragraph. The construction order one should imagine is that we build the ``cosets'' for decreasing , and the possible constructions in each coset come from .
In the case that is the set of inductive intervals on a polycyclic group, one can repeat the arguments of the previous sections to conclude that if is -invariant and a -avoshift, then we can compute a set of -invariant finite patterns on that define uniformly on all sets in .
Definition 9.1.
Let be a polycyclic group with fixed polycycle structure. Given an avoshift , we say is an avofactor of if there is a factor map such that the graph of , defined by is an avorelation.
Note that we assume the same alphabet just for notational convenience; if they are not the same, we can consider both with the union alphabet. It is important that the image is written on the second component, so that the ``image of is constructed before the preimage'' when we consider -well-orderings. The decidability result below effectively comes from the fact that we find local rules for building preimages for any image.
Now we can show the following analogously to Theorem 8.5.
Theorem 9.2.
Every avofactor of an II-avoshift on a polycyclic group is an avoshift, and we can effectively compute its forbidden patterns.
Sketch of proof.
Let be the inductive intervals of the polycyclic group . Consider the subshift , assumed avo. Note that must be an avoshift, since for all , proving the first claim.
Next, we go through the entire theory with in place of inductive intervals. We can easily show that is wqo under inclusion and closed under increasing union. Now the proof of Lemma 5.6 shows that any -avoshift is uniformly -avo.
Next, analogously to Lemma 5.8 we see that is uniformly SFT for , i.e. there is a set of forbidden patterns for it, by showing that sets in are constructible in an appropriate sense.
Since are also extendable and are computable in any reasonable sense, we see as in Section 8 that we can compute a set of forbidden patterns defining uniformly on . In particular these patterns work on , which means precisely that they give forbidden patterns for the image subshift. ∎
Corollary 9.3.
We can effectively compute the algebraic factors of any group shift on a polycyclic group.
Proof.
It is easy to verify that the graph of a shift-invariant continuous group homomorphism between group shifts is an avorelation. ∎
Example 9.4.
Not all factors of avoshifts are avo – in fact, avoshifts on inductive intervals are not closed under conjugacy even on the group (note that on they are just the SFTs, so they are closed under conjugacy). For example, take the binary full shift. Now double the alphabet and always write the -image of the first track of the row below, to the second track of the present row, where is the cellular automaton from Example 7.8. As in Example 7.8, it is easy to see that the subshift is not avo for inductive intervals.
10 Discussion and future work
The avoshift notion we have developed here already has strong implications, and already generalizes many of the known important properties of group shifts. However, the theory is still in its infancy, and we outline here some future directions.
10.1 Practical implementations
One motivation of the present paper is to provide a (first approximation to a) theoretical framework for practical decision procedures for basic operations on SFTs, at least comparing them and computing their languages. We have not attempted to make the procedures efficient in the present paper (and do not even provide pseudocode), nor have we implemented them. Our hope is that they can be implemented with SAT solvers. If they end up being practical, the plan is also to fit them in the framework of [29].
For now we do not know how practical the methods are, as we have not applied them (on the computer). A particularly interesting question is how well these methods perform compared to Wang-style algorithms for deciding properties of multi-dimensional SFTs. From our experiments with the Diddy toolbox [29] (which implements Wang's algorithm), our impression is that the basic algorithm of Wang from [32] for computing the language of a subshift using only density of periodic points is remarkably effective also for comparing SFTs when implemented with SAT solvers.
While of course multidimensional SFTs in practice tend to have periodic points and counterexamples are always carefully engineered rather than arise naturally, dense periodic points is a stringent property that fails for many natural examples, in particular it is trivial to find pairs of SFTs whose comparison cannot be made with Wang's algorithm directly. It is indeed trivial to find examples where the avoshift theory can check noninclusion of SFTs, but Wang's algorithm cannot, even in one dimension.
This is not a true weakness of Wang style algorithms: one could say that the core idea of Wang's algorithm is that one can prove the global validity of a finite pattern, by giving a complete finitary description of a configuration, in a sufficiently simple logical framework so that one can check it belongs to the subhsift, and that finding such a description is reasonable to find. Periodic points are of course an obvious example of such structures, but one may more generally consider semilinear points (see e.g. [27]) possible with restrictions on the eventual periods (which Diddy has very limited support for, at present [29]), and it is plausible that one can more generally use automatic structures [11]. These already cover a lot of ground, and in fact we are not aware of any avoshifts that do not have dense semilinear points. Thus, we are not aware of any deep theoretical reason why Wang-style methods could not be better than avoshifts.
On the other hand, there are situations where there is an easy-to-describe way to extend any configuration on a subset, but a simple configuration can nevertheless have an incredibly complicated extension with even uncomputable properties (think of spacetime diagrams of a Turing complete reversible cellular automaton). This might intuitively suggest that there are subshifts where an avoshift-style theory can make nontrivial deductions, but truly Wang-type algorithms are a logical impossibility in the sense that global configurations do not have direct finite descriptions (at least on the level of concreteness of automatic configurations). Again, we do not have concrete examples of such a phenomenon.
There are of course also other decidability methods that the present methods should be compared to. For example, in one dimension, classical automata theory seems unrivaled in its expressivity and practicality. One can also use automata theory in two dimensions by applying classical automata theory to slices of configurations. The references [10, 27] are able to prove some interesting results about concrete examples of -SFTs this way.
10.2 Extensions to be explored
In the case of , in Lemma 3.11 we only used one-sided inductive intervals on (only the infinite one, because for finite sets the avo propery is automatic). One should indeed be able to extend the present theory so that it would cover polycyclic groups and ``one-sided/signed inductive intervals''. However, constructibility fails for these sets, and one has to consider weak constructibility, meaning we do not directly well-order sets, but present them as a well-ordered union of constructible sets. We believe the theory can be easily augmented to cover this, and also in two dimensions is covers some additional examples (such as spacetime diagrams of cellular automata Proposition 4.10). The main reason for not attempting to write the results in this generality is that while we can extend the SFTness proofs to this case, we do not know how to extend any of the decidability theory without the property of extendability (Definition LABEL:def:Extendable), which fails for any notion of one-sided/signed inductive intervals. These decidability aspects are what we consider the most important facet of the theory developed here.
The theory of avoshifts could be extended to subshifts on monoids like or . However, we do not know what the optimal framework is. One might consider something like ``signed polycyclic groups'' where some of the infinite dimensions are restricted to only including the positive cosets of the normal subgroup. However, requiring intermediate dimensions to be positive can lead to complications in the non-abelian case (even in simple examples like the Heisenberg group), and we do not know what the cleanest approach is.
One possible solution is to develop the theory more generally in a setting of graph subshifts, to hopefully both cover groups like , and to make the decidability theory of avofactors in Section 9 a direct corollary of the general theory instead of needing an ad hoc discussion.
10.3 Dynamical properties
Avoshifts as defined her lack some desirable properties. First, they are not conjugacy invariant, thus being an avoshift is itself not actually a dynamical property (except on where it corresponds to being an SFT). This is the case also for SFTs with safe symbols and topological strong spatial mixing subshifts, as well as -TEP subshifts, which as we recall as some of classes generalized by avoshifts. However, group shifts (arguably the main class of motivating examples) can be defined purely abstractly as internal groups in the category of subshifts. While we have to some extent swept this under a rug, they are indeed avoshifts only up to conjugacy. The failure is drastic here, as full shifts admit a group shift structure, but there are subshifts conjugate to full shifts which are not avo (Example 9.4).
The mixing properties finite extension property [6], as well as the related recently defined map extension property [17] and contractibility [22], all imply dense periodic points on a large class of groups [22], and are conjugacy invariant. The finite extension property was actually introduced as a natural conjugacy-invariant generalization of the topological strong spatial mixing property. Of course, one can trivially extend the present theory to cover the class of subshifts that are avoshifts up to conjugacy, in the sense that SFTness is conjugacy-invariant, and for the decidability theory one can go through possible conjugacies and the avo property through a conjugating map can be recognized. However, it seems likely that there is a more natural approach.
Finally, in the present paper we have only concentrated on decidability results. Group shifts, -TEP subshifts and TSSM subshifts also have many other desirable properties such as dense periodic points, and interesting things can be said about the invariant measures in at least the first two cases (for TSSM and SFTs with a safe symbol, see [21] for some information). We do not know how much one can say, in the present generality, about the dynamical properties of avoshifts (other than the fact they are SFT).
As we have shown, avoshifts do not have dense periodic points in general. For decidability theory, as discussed in Section 10.1, this is in a sense a good thing, as we want the class to cover as much dynamical ground as possible, and dense periodic points is a restrictive property. In a future paper, our plan is to find subclasses of avoshifts with more restricted dynamics (and still covering group shifts). Specifically, for avoshifts on convex sets in , and for subshifts with equal extension counts, we claim that there are non-trivial things that can be said.
References
- [1] Sebastián Barbieri, Ricardo Gómez, Brian Marcus, and Siamak Taati. Equivalence of relative Gibbs and relative equilibrium measures for actions of countable amenable groups. Nonlinearity, 33(5):2409–2454, Mar 2020.
- [2] Laurent Bartholdi and Ville Salo. Shifts on the lamplighter group, 2024.
- [3] Pierre Béaur and Jarkko Kari. Effective projections on group shifts to decide properties of group cellular automata. International Journal of Foundations of Computer Science, 35(01n02):77–100, 2024.
- [4] Robert Berger. The undecidability of the domino problem. Mem. Amer. Math. Soc. No., 66, 1966. 72 pages.
- [5] Raimundo Briceño. The topological strong spatial mixing property and new conditions for pressure approximation. Ergodic Theory and Dynamical Systems, 38(5):1658–1696, 2018.
- [6] Raimundo Briceño, Kevin McGoff, and Ronnie Pavlov. Factoring onto subshifts with the finite extension property. Proceedings of the American Mathematical Society, 146(12):5129–5140, 2018.
- [7] Robert Burton and Jeffrey E Steif. Non-uniqueness of measures of maximal entropy for subshifts of finite type. Ergodic Theory and Dynamical Systems, 14(2):213–235, 1994.
- [8] Nishant Chandgotia, Guangyue Han, Brian Marcus, Tom Meyerovitch, and Ronnie Pavlov. One-dimensional markov random fields, markov chains and topological markov fields. Proceedings of the American Mathematical Society, 142(1):227–242, 2014.
- [9] Michael Hochman. Irreducibility and periodicity in symbolic systems, 2024.
- [10] Emmanuel Jeandel and Michael Rao. An aperiodic set of 11 Wang tiles. arXiv e-prints, page arXiv:1506.06492, Jun 2015.
- [11] Bakhadyr Khoussainov and Mia Minnes. Three lectures on automatic structures. In Proceedings of Logic Colloquium, volume 35, pages 132–176, 2007.
- [12] Bruce Kitchens and Klaus Schmidt. Periodic points, decidability and Markov subgroups. In James C. Alexander, editor, Dynamical Systems, pages 440–454, Berlin, Heidelberg, 1988. Springer Berlin Heidelberg.
- [13] Bruce P. Kitchens. Expansive dynamics on zero-dimensional groups. Ergodic Theory Dyn. Syst., 7:249–261, 1987.
- [14] Samuel J Lightwood. Morphisms from non-periodic subshifts I: constructing embeddings from homomorphisms. Ergodic Theory and Dynamical Systems, 23(2):587–609, 2003.
- [15] Samuel J Lightwood. Morphisms from non-periodic subshifts II: constructing homomorphisms to square-filling mixing shifts of finite type. Ergodic Theory and Dynamical Systems, 24(4):1227–1260, 2004.
- [16] Douglas Lind and Brian Marcus. An introduction to symbolic dynamics and coding. Cambridge university press, 2021.
- [17] Tom Meyerovitch. An embedding theorem for multidimensional subshifts. arXiv preprint arXiv:2312.05650, 2023.
- [18] Mojang Studios. Minecraft, 2024. Version 1.21.1, Video game.
- [19] A. Yu. Olshanskii. On a geometric method in the combinatorial group theory. In International Congress of Mathematicians, Proceedings of the International Congress of Muthematicians, pages 415–424, 1983.
- [20] William Parry. Symbolic dynamics and transformations of the unit interval. Transactions of the American Mathematical Society, 122(2):368–378, 1966.
- [21] Ronnie Pavlov. Shifts of finite type with nearly full entropy. Proceedings of the London Mathematical Society, 108(1):103–132, 2014.
- [22] Leo Poirier and Ville Salo. Contractible subshifts, 2024.
- [23] Ville Salo. Topologically mixing cellular automata on groups. MathOverflow. URL:https://mathoverflow.net/questions/364509 (version: 2022-03-23).
- [24] Ville Salo. When are group shifts of finite type? arXiv preprint arXiv:1807.01951, 2018.
- [25] Ville Salo. Cutting corners. Journal of Computer and System Sciences, 128:35–70, 2022.
- [26] Ville Salo and Ilkka Törmä. On shift spaces with algebraic structure. In How the World Computes: Turing Centenary Conference and 8th Conference on Computability in Europe, CiE 2012, Cambridge, UK, June 18-23, 2012. Proceedings 8, pages 636–645. Springer, 2012.
- [27] Ville Salo and Ilkka Törmä. Gardens of eden in the game of life. In Automata and Complexity: Essays Presented to Eric Goles on the Occasion of His 70th Birthday, pages 399–415. Springer, 2022.
- [28] Ville Salo and Ilkka Törmä. What can oracles teach us about the ultimate fate of life? arXiv e-prints, page arXiv:2202.07346, February 2022.
- [29] Ville Salo and Ilkka Törmä. Diddy: a Python toolbox for infinite discrete dynamical systems. In International Workshop on Cellular Automata and Discrete Complex Systems, pages 33–47. Springer, 2023.
- [30] Ville Salo and Ilkka Törmä. Recoding lie algebraic subshifts, 2020.
- [31] Klaus Schmidt. Dynamical systems of algebraic origin, volume 128 of Progress in Mathematics. Birkhäuser Verlag, Basel, 1995.
- [32] Hao Wang. Proving theorems by pattern recognition II. Bell System Technical Journal, 40:1–42, 1961.