This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Graph automaton groups

Matteo Cavaleri Matteo Cavaleri, Università degli Studi Niccolò Cusano - Via Don Carlo Gnocchi, 3 00166 Roma, Italia [email protected] Daniele D’Angeli Daniele D’Angeli, Università degli Studi Niccolò Cusano - Via Don Carlo Gnocchi, 3 00166 Roma, Italia [email protected] Alfredo Donno Alfredo Donno, Università degli Studi Niccolò Cusano - Via Don Carlo Gnocchi, 3 00166 Roma, Italia [email protected]  and  Emanuele Rodaro Emanuele Rodaro, Politecnico di Milano - Piazza Leonardo da Vinci, 32 20133 Milano, Italia [email protected]
Abstract.

In this paper we define a way to get a bounded invertible automaton starting from a finite graph. It turns out that the corresponding automaton group is regular weakly branch over its commutator subgroup, contains a free semigroup on two elements and is amenable of exponential growth. We also highlight a connection between our construction and the right-angled Artin groups. We then study the Schreier graphs associated with the self-similar action of these automaton groups on the regular rooted tree. We explicitly determine their diameter and their automorphism group in the case where the initial graph is a path. Moreover, we show that the case of cycles gives rise to Schreier graphs whose automorphism group is isomorphic to the dihedral group. It is remarkable that our construction recovers some classical examples of automaton groups like the Adding machine and the Tangled odometer.

Key words and phrases:
Graph automaton group, Schreier graph, Growth, Fractal group, Self-similar representation, Diameter, Graph automorphism.

Mathematics Subject Classification (2010): 20F65, 20F05, 20E08, 05C10, 05C25.

1. Introduction

Algebraic structures can be usually described by means of their combinatorial nature. A typical example is the relation between groups and graphs. In Geometric Group Theory, for instance, many algebraic properties of a group can be detected by investigating the combinatorial properties of the corresponding Cayley graphs. Another typical example of this interaction is achieved by the automaton group theory. Automata (or Mealy machines or transducers) are directed graphs whose transitions describe the action of the states on a finite alphabet. If, for instance, the automaton is complete and invertible, then its states generate a group that has, in many cases, very interesting properties.
As an example, in 1980 R. I. Grigorchuk described in [20] the first group of intermediate (i.e., faster than polynomial and slower than exponential) growth, that later appeared to be generated by a finite automaton. This group has a number of other interesting properties: for example, it is infinite and finitely generated, but each of its elements has finite order (Burnside group). It was also the first example of an amenable but not elementary amenable group. Over the last decades, a new exciting direction of research focusing on finitely generated automaton groups acting by automorphisms on rooted trees has been developed. It turned out that it has deep connections with the theory of profinite groups and with complex dynamics. The interested reader can refer for more details to the following list of works (and bibliography therein) [1, 2, 21, 27]. Recently a special interest has been pointed out for decision problems for automaton groups and semigroups (see [16, 18, 19, 29] and references therein).
Beside Cayley graphs the best way to encode the action of automaton groups is given by the structure of the corresponding Schreier graphs. These graphs better represent the length-preserving action of an infinite automaton group on a finite alphabet and constitute an infinite sequence of finite graphs converging to infinite graphs in the Gromov-Hausdorff topology. Finite Schreier graphs can be regarded as orbital graphs with respect to the action of the automaton group on each level of a regular rooted tree, whereas their limits are orbital graphs of the action of the group on the boundary of the tree. Finite Schreier graphs have been studied from a combinatorial point of view in several contexts (e.g., the investigation of models coming from Statistical Mechanics [13, 14]). Classification of infinite Schreier graphs have been studied in several papers (see [6, 7, 11, 12, 15, 23] for further discussions about this topic).

In this paper, we introduce a family of automaton groups arising from finite graphs, that we call graph automaton groups (Definition 3.1). More precisely, with any finite graph G=(V,E)G=(V,E), we associate an invertible automaton 𝒜G\mathcal{A}_{G} whose state set is identified with the edge set EE and that acts on words over an alphabet identified with the vertex set VV. Basically, any edge has a nontrivial action only on words starting with a letter identified with one of its endpoints. In this way, we can define an automaton that has the remarkable property of being bounded. The class of bounded automaton groups is very popular and it contains the most famous examples of automaton groups (e.g., the Grigorchuk group and the Basilica group [25]). It is a remarkable fact that all groups generated by bounded automata are amenable, and so all our graph automaton groups have this property. It is interesting that, with our construction, we recover some classical examples of automaton groups (e.g., the Adding machine and the Tangled odometer group, see Example 3.1). Among other properties, we prove that our groups are fractal and weakly regular branch over their commutator subgroup (Theorem 3.1), that they all contain (except some trivial cases) a free semigroup (Theorem 3.2) and so they have exponential growth (Corollary 3.1). We also highlight a connection between our groups and the right-angled Artin groups (Proposition 3.4).
The second part of the paper is devoted to the investigation of the Schreier graphs associated with the action of graph automaton groups. In Theorem 4.1, we give a rigidity result for the automorphism group of Schreier graphs when the initial graph is cyclic. Then we present an explicit recursive description of the Schreier graphs of graph automaton groups generated by path graphs, and we are able to determine their diameter (Theorem 4.2) and their automorphism group (Theorem 4.3).

2. Preliminaries

We recall in this preliminary section some basic definitions and properties about automaton groups and growth.

Definition 2.1.

An automaton is a quadruple 𝒜=(S,X,λ,μ)\mathcal{A}=(S,X,\lambda,\mu), where:

  1. (1)

    SS is the set of states;

  2. (2)

    X={1,2,,k}X=\{1,2,\ldots,k\} is an alphabet;

  3. (3)

    λ:S×XS\lambda:S\times X\rightarrow S is the restriction map;

  4. (4)

    μ:S×XX\mu:S\times X\rightarrow X is the output map.

The automaton 𝒜\mathcal{A} is finite if SS is finite and it is invertible if, for all sSs\in S, the transformation μ(s,):XX\mu(s,\cdot):X\rightarrow X is a permutation of XX. An automaton 𝒜\mathcal{A} can be visually represented by its Moore diagram: this is a directed labeled graph whose vertices are identified with the states of 𝒜\mathcal{A}. For every state sSs\in S and every letter xXx\in X, the diagram has an arrow from ss to λ(s,x)\lambda(s,x) labeled by x|μ(s,x)x|\mu(s,x). A sink idid in 𝒜\mathcal{A} is a state with the property that λ(id,x)=id\lambda(id,x)=id and μ(id,x)=x\mu(id,x)=x for any xXx\in X.

An important class of automata is given by the so-called bounded automata [28]. An automaton is said to be bounded if the sequence of numbers of paths of length nn avoiding the sink state (along the directed edges of the Moore diagram) is bounded.

For each n1n\geq 1, let XnX^{n} denote the set of words of length nn over the alphabet XX and put X0={}X^{0}=\{\emptyset\}, where \emptyset is the empty word. The action of 𝒜\mathcal{A} can be easily extended to the infinite set X=n=0XnX^{\ast}=\bigcup_{n=0}^{\infty}X^{n} as follows:

(1) λ(s,xw)=λ(λ(s,x),w)μ(s,xw)=μ(s,x)μ(λ(s,x),w),\displaystyle\lambda(s,xw)=\lambda(\lambda(s,x),w)\qquad\qquad\mu(s,xw)=\mu(s,x)\mu(\lambda(s,x),w),

for every wXw\in X^{\ast}. For a state sSs\in S, we denote by 𝒜s\mathcal{A}_{s} the transformation μ(s,)\mu(s,\cdot) on XX^{\ast} defined by Eqs. (1), which is a bijection if 𝒜\mathcal{A} is invertible. Given the invertible automaton 𝒜\mathcal{A}, the automaton group generated by 𝒜\mathcal{A} is by definition the group generated by the transformations 𝒜s\mathcal{A}_{s}, for sSs\in S, and it is denoted G(𝒜)G(\mathcal{A}). In the rest of the paper, we will often use the notation ss instead of 𝒜s\mathcal{A}_{s}. Moreover, the maps λ\lambda and μ\mu can be naturally extended to each element of G(𝒜)G(\mathcal{A}). Notice that the action of G(𝒜)G(\mathcal{A}) on XX^{\ast} preserves the sets XnX^{n}, for each nn.
It is a remarkable fact that an automaton group can be regarded in a very natural way as a group of automorphisms of the regular rooted tree of degree |X|=k|X|=k, i.e., the rooted tree TkT_{k} in which each vertex has kk children, via the identification of the knk^{n} vertices of the nn-th level of TkT_{k} with the set XnX^{n}.
The group G(𝒜)G(\mathcal{A}) is said to be spherically transitive if its action is transitive on XnX^{n}, for any nn. Let gG(𝒜)g\in G(\mathcal{A}). The action of gg on XX^{\ast} can be factorized by considering the action on XX and |X||X| restrictions as follows. Let Sym(k)Sym(k) be the symmetric group on kk elements. Then an element gG(𝒜)g\in G(\mathcal{A}) can be represented as

(2) (g1,,gk)σ,\displaystyle(g_{1},\ldots,g_{k})\sigma,

where gi:=λ(g,i)G(𝒜)g_{i}:=\lambda(g,i)\in G(\mathcal{A}) and σSym(k)\sigma\in Sym(k) describes the action of gg on XX. We say that Eq. (2) is the self-similar representation of gg. Notice that, given g=(g1,,gk)σg=(g_{1},\ldots,g_{k})\sigma and h=(h1,,hk)τh=(h_{1},\ldots,h_{k})\tau, the self-similar representation of ghgh is

gh=(g1hσ(1),,gkhσ(k))στ.gh=(g_{1}h_{\sigma(1)},\ldots,g_{k}h_{\sigma(k)})\sigma\tau.

In the tree interpretation of Eq. (2), the permutation σ\sigma corresponds to the action of gg on the first level of TkT_{k}, and the automorphism gig_{i} is the restriction of the action of gg to the subtree (isomorphic to TkT_{k}) rooted at the ii-th vertex of the first level.

Let us denote by StabG(𝒜)(X)={gG(𝒜):g(x)=x,xX}Stab_{G(\mathcal{A})}(X)=\{g\in G(\mathcal{A}):g(x)=x,\ \forall\ x\in X\} the stabilizer of XX, that is, the subgroup of G(𝒜)G(\mathcal{A}) consisting of all elements acting trivially on XX. In particular, if gStabG(𝒜)(X)g\in Stab_{G(\mathcal{A})}(X), then one can identify gg with the kk-tuple (g1,,gk)(g_{1},\ldots,g_{k}), where gi=λ(g,i)g_{i}=\lambda(g,i).

Lemma 2.1.

Let g=(g1,,gk)StabG(𝒜)(X)g=(g_{1},\ldots,g_{k})\in Stab_{G(\mathcal{A})}(X). If gi{g,id}g_{i}\in\{g,id\} for each i=1,,ki=1,\ldots,k, then gg is the trivial element of G(𝒜)G(\mathcal{A}).

Proof.

Let wXnw\in X^{n}. We will prove by induction on nn that g(w)=wg(w)=w. The case n=1n=1 is obvious since gStabG(𝒜)(X)g\in Stab_{G(\mathcal{A})}(X). Now let w=xww=xw^{\prime}, with xXx\in X and wXn1w^{\prime}\in X^{n-1}. Then one has g(xw)=xgx(w)g(xw^{\prime})=xg_{x}(w^{\prime}). If gx=idg_{x}=id, the claim easily follows. If gx=gg_{x}=g, the inductive hypothesis gives gx(w)=wg_{x}(w^{\prime})=w^{\prime} and so one gets g(w)=wg(w)=w. ∎

Given two groups G1G_{1} and G2G_{2}, the direct product of G1G_{1} and G2G_{2} will be denoted by G1×G2G_{1}\times G_{2}. In particular, we denote by GkG^{k} the kk-times iterated direct product of a group GG with itself. The following definitions are given in the literature for the broader class of self-similar groups (see [27]). Here we restrict our interest to automaton groups.

Definition 2.2.

Let G(𝒜)G(\mathcal{A}) be an automaton group. Then G(𝒜)G(\mathcal{A}) is said to be

  1. (1)

    fractal, if the map ψ:StabG(𝒜)(X)G(𝒜)k\psi:Stab_{G(\mathcal{A})}(X)\to G(\mathcal{A})^{k} given by g(g1,,gk)g\mapsto(g_{1},\ldots,g_{k}) is surjective on each factor.

  2. (2)

    weakly regular branch over its subgroup NN, if Nkψ(NStabG(𝒜)(X))N^{k}\subset\psi(N\cap Stab_{G(\mathcal{A})}(X)), where NN is supposed to be nontrivial.

Let 𝒢\mathcal{G} be a finitely generated group, with symmetric generating set S=S1S=S^{-1}. The length of gg with respect to SS is denoted by |g|S|g|_{S} (we will omit the subscript SS when the generating set is fixed) and it is defined as the minimal length of a word in SS^{\ast} representing the element gg. Then one puts γS(n)=#{g𝒢:|g|Sn}\gamma_{S}(n)=\#\{g\in\mathcal{G}:|g|_{S}\leq n\}. This defines the growth function

γS:.\gamma_{S}:\mathbb{N}\to\mathbb{N}.

This map clearly depends on the generating set SS. However, one can prove that changing the generating set does not affect the asymptotic properties of γS\gamma_{S}. In particular, the growth rate

λ𝒢=limnγS(n)n\lambda_{\mathcal{G}}=\lim_{n\to\infty}\sqrt[n]{\gamma_{S}(n)}

is greater than or equal to 11 independently on the particular generating set (see, for instance, [17]).

Definition 2.3.

A finitely generated group 𝒢\mathcal{G} has exponential growth (resp. sub-exponential growth) if λ𝒢>1\lambda_{\mathcal{G}}>1 (resp. λ𝒢=1\lambda_{\mathcal{G}}=1).

3. Automata groups from graphs

We present in this section the main construction of the paper, that is, we are going to associate an invertible automaton with a given finite graph.
Let G=(V,E)G=(V,E) be a finite graph. Let V={x1,,xk}V=\{x_{1},\ldots,x_{k}\} (we will often use also the notation {1,2,,k}\{1,2,\ldots,k\}) be its vertex set and let EE be its edge set.
First of all, we choose an orientation for the edges of GG. Let EE^{\prime} be the set of edges, where an orientation of each edge has been chosen. Notice that elements in EE are unordered pairs of type {xi,xj}\{x_{i},x_{j}\}, whereas elements in EE^{\prime} are ordered pairs of type (xi,xj)(x_{i},x_{j}), meaning that the edge has been oriented from the vertex xix_{i} to the vertex xjx_{j}.
We then define an automaton 𝒜G=(E{id},V,λ,μ)\mathcal{A}_{G}=(E^{\prime}\cup\{id\},V,\lambda,\mu) such that:

  • E{id}E^{\prime}\cup\{id\} is the set of states;

  • VV is the alphabet;

  • λ:E×VE\lambda:E^{\prime}\times V\to E^{\prime} is such that, for each e=(x,y)Ee=(x,y)\in E^{\prime}, one has

    λ(e,z)={eif z=xidif zx;\lambda(e,z)=\left\{\begin{array}[]{ll}e&\hbox{if }z=x\\ id&\hbox{if }z\neq x;\end{array}\right.
  • μ:E×VV\mu:E^{\prime}\times V\to V is such that, for each e=(x,y)Ee=(x,y)\in E^{\prime}, one has

    μ(e,z)={yif z=xxif z=yzif zx,y.\mu(e,z)=\left\{\begin{array}[]{ll}y&\hbox{if }z=x\\ x&\hbox{if }z=y\\ z&\hbox{if }z\neq x,y.\end{array}\right.

In other words, any directed edge e=(x,y)e=(x,y) represents a state of the automaton 𝒜G\mathcal{A}_{G} and it has just one restriction to itself (given by λ(e,x)\lambda(e,x)) and all other restrictions to the sink idid. Its action is nontrivial only on the letters xx and yy, which are switched since μ(e,x)=y\mu(e,x)=y and μ(e,y)=x\mu(e,y)=x. It is easy to check that 𝒜G\mathcal{A}_{G} is invertible for any GG and any choice of the orientation of the edges. This makes us able to define an associated automaton group.

Definition 3.1.

The graph automaton group 𝒢G\mathcal{G}_{G} is the automaton group generated by 𝒜G\mathcal{A}_{G}.

Remark 3.1.

Notice that any loop in GG gives rise to the trivial element of 𝒢G\mathcal{G}_{G}. Moreover, any multiedge produces a set of equal generators (up to consider the inverse).

The previous remark basically says that we can just consider simple graphs. We make from now on this assumption. The following proposition shows that the graph automaton group 𝒢G\mathcal{G}_{G} does not depend on the particular orientation of the edges in GG nor on the order of the vertices in VV.

Proposition 3.1.

The group 𝒢G\mathcal{G}_{G} does not depend on the choice of the edge orientation. Moreover, a rearrangement of the vertices in VV produces groups that are isomorphic.

Proof.

Without loss of generality, we can suppose that there exists an edge connecting the vertices x1x_{1} and x2x_{2}. If we choose the orientation from x1x_{1} to x2x_{2}, then the self-similar representation of the corresponding edge ee as a generator of 𝒢G\mathcal{G}_{G} is

e=(e,id,,id)(1,2).e=(e,id,\ldots,id)(1,2).

On the other hand, if the opposite orientation is chosen, the self-similar representation of the corresponding edge ff becomes

f=(id,f,,id)(1,2).f=(id,f,\ldots,id)(1,2).

A direct computation gives ef=(ef,id,,id)ef=(ef,id,\ldots,id), so that ef=idef=id by Lemma 2.1, and so f=e1f=e^{-1}. Therefore, the choice of the orientation does not affect the structure of the group 𝒢G\mathcal{G}_{G}.
In order to prove the second claim, it is enough to observe that changing the name of the vertices is equivalent to act by a permutation σ\sigma on the alphabet VV. In this case, we just get the group 𝒢Gσ\mathcal{G}_{G}^{\sigma}, which is obtained from 𝒢G\mathcal{G}_{G} via conjugation by σ\sigma. ∎

In the light of Proposition 3.1, given an oriented edge e=(xi,xj)Ee=(x_{i},x_{j})\in E^{\prime}, we can denote by e1=(xj,xi)e^{-1}=(x_{j},x_{i}) the edge with the opposite orientation, keeping in mind that the choice of the opposite orientation corresponds to take the inverse in 𝒢G\mathcal{G}_{G}.

Remark 3.2.

For any graph GG the automaton 𝒜G\mathcal{A}_{G} is bounded. In fact, it is easy to check that the Moore diagram of 𝒜G\mathcal{A}_{G}, regarded as a simple graph, is a star graph with one internal vertex corresponding to the sink idid and |E||E| leaves corresponding to the directed edges of GG. Each leaf has exactly one directed loop and all other transitions go to the sink. In particular, for any edge of GG, the Moore diagram contains exactly one cycle of fixed length avoiding the sink and disconnected from the other cycles. As a consequence, the group 𝒢G\mathcal{G}_{G} does not contain free non abelian subgroups, as follows from Sidki’s theorem [28]. Moreover, since all groups generated by bounded automata are amenable [3], then each graph automaton group 𝒢G\mathcal{G}_{G} is amenable. We do not provide a special description of Følner sets of 𝒢G\mathcal{G}_{G}. The matter certainly deserves further investigation in the future. Notice that, having 𝒢G\mathcal{G}_{G} solvable world problem, a sequence of Følner sets must be computable (in the sense of [8, 9]).

The following result shows that taking subgraphs in GG corresponds to obtain subgroups in 𝒢G\mathcal{G}_{G}. By a subgraph of GG, here we mean a subset of its edges, together with the subset of vertices of VV consisting of their endpoints.

Proposition 3.2.

Let H=(VH,EH)H=(V_{H},E_{H}) be a graph isomorphic to a subgraph of GG. Then 𝒢H𝒢G\mathcal{G}_{H}\leq\mathcal{G}_{G}.

Proof.

Let e1,,ee_{1},\ldots,e_{\ell} be the edges of G=(V,E)G=(V,E) belonging to the subgraph G~\widetilde{G} isomorphic to HH. Consider the subgroup 𝒢~𝒢G\widetilde{\mathcal{G}}\leq\mathcal{G}_{G} generated by e1,,ee_{1},\ldots,e_{\ell}. We claim that the groups 𝒢~\widetilde{\mathcal{G}} and 𝒢H\mathcal{G}_{H} are isomorphic. Let ψ\psi be the isomorphism between G~\widetilde{G} and HH. By Proposition 3.1 we can suppose, without loss of generality, that e1,,ee_{1},\ldots,e_{\ell} are edges of GG, whose endpoints are the first tt vertices of VV, and we can consider in VHV_{H} the order induced by ψ\psi. If we focus on the self-similar representation of the automorphisms generating 𝒢H\mathcal{G}_{H}, this is exactly the self-similar representation of the eie_{i}’s, regarded as generators of 𝒢G\mathcal{G}_{G}, where the last |V|t|V|-t restrictions are erased. Therefore, it is clear that a relation in 𝒢~\widetilde{\mathcal{G}} holds if and only if the same relation in 𝒢H\mathcal{G}_{H} holds. ∎

Proposition 3.3.

Let G=(V,E)G=(V,E) be the disjoint union of the graphs G1=(V1,E1)G_{1}=(V_{1},E_{1}), \ldots, Gt=(Vt,Et)G_{t}=(V_{t},E_{t}). Then 𝒢G=𝒢G1××𝒢Gt\mathcal{G}_{G}=\mathcal{G}_{G_{1}}\times\cdots\times\mathcal{G}_{G_{t}}.

Proof.

The claim easily follows by observing that the generators in EiE_{i} act trivially on the set of vertices VViV\setminus V_{i}. ∎

By virtue of Proposition 3.3, we can suppose GG to be connected. Therefore, from now on, we assume G=(V,E)G=(V,E) to be simple and connected.

Example 3.1.
  1. (1)

    If GG is the path graph P2P_{2} on 22 vertices, the associated automaton 𝒜P2\mathcal{A}_{P_{2}} is represented in Fig. 1.

    \psfrag{a}{$a$}\psfrag{1}{$1$}\psfrag{2}{$2$}\psfrag{id}{$id$}\psfrag{1|1,2|2}{$1|1,2|2$}\psfrag{1|2}{$1|2$}\psfrag{2|1}{$2|1$}\includegraphics[width=250.38562pt]{danielesegmento.eps}
    Figure 1. The path P2P_{2} and the associated automaton 𝒜P2\mathcal{A}_{P_{2}}.

    The automaton 𝒜P2\mathcal{A}_{P_{2}} is the so-called Adding machine (see, e.g., [1]) and it generates the group \mathbb{Z}. The self-similar representation of its generator is

    a=(a,id)(1,2).a=(a,id)(1,2).
  2. (2)

    If GG is the path graph P3P_{3} on 33 vertices, the associated automaton 𝒜P3\mathcal{A}_{P_{3}} is represented in Fig. 2.

    \psfrag{a}{$a$}\psfrag{b}{$b$}\psfrag{1}{$1$}\psfrag{2}{$2$}\psfrag{3}{$3$}\psfrag{id}{$id$}\psfrag{1|2}{$1|2$}\psfrag{2|3}{$2|3$}\psfrag{2|1,3|3}{$2|1,3|3$}\psfrag{1|1,3|2}{$1|1,3|2$}\psfrag{1|1,2|2,3|3}{$1|1,2|2,3|3$}\includegraphics[width=250.38562pt]{danielep3.eps}
    Figure 2. The path P3P_{3} and the associated automaton 𝒜P3\mathcal{A}_{P_{3}}.

    It corresponds to the so-called Tangled odometer (see, e.g., [22]). The two generators of the group 𝒢P3\mathcal{G}_{P_{3}} have the self-similar representation:

    a=(a,id,id)(1,2)b=(id,b,id)(2,3).a=(a,id,id)(1,2)\qquad b=(id,b,id)(2,3).
  3. (3)

    If GG is the cyclic graph C3C_{3} on 33 vertices, the associated automaton 𝒜C3\mathcal{A}_{C_{3}} is represented in Fig. 3.

    \psfrag{a}{$a$}\psfrag{b}{$b$}\psfrag{c}{$c$}\psfrag{1}{$1$}\psfrag{2}{$2$}\psfrag{3}{$3$}\psfrag{id}{$id$}\psfrag{1|2}{$1|2$}\psfrag{2|3}{$2|3$}\psfrag{2|1,3|3}{$2|1,3|3$}\psfrag{1|1,3|2}{$1|1,3|2$}\psfrag{1|1,2|2,3|3}{$1|1,2|2,3|3$}\psfrag{1|3,2|2}{$1|3,2|2$}\psfrag{3|1}{$3|1$}\includegraphics[width=273.14922pt]{danieletriangolo.eps}
    Figure 3. The cycle C3C_{3} and the associated automaton 𝒜C3\mathcal{A}_{C_{3}}.

    The automaton 𝒜C3\mathcal{A}_{C_{3}} generates a group that we call the uncle of the Hanoi Towers group on three pegs (see, e.g., [24]), and the three generators of the group 𝒢C3\mathcal{G}_{C_{3}} have the self-similar representation:

    a=(a,id,id)(1,2)b=(id,b,id)(2,3)c=(id,id,c)(3,1).a=(a,id,id)(1,2)\qquad b=(id,b,id)(2,3)\qquad c=(id,id,c)(3,1).

    Observe that the automorphisms A:=abcA:=abc, B:=bcaB:=bca, and C:=cabC:=cab generate the classical Hanoi Towers group on three pegs, that is therefore a subgroup of its uncle.

Notice that if e,fEe,f\in E^{\prime} do not share any vertex, then [e,f]=e1f1ef=id[e,f]=e^{-1}f^{-1}ef=id in 𝒢G\mathcal{G}_{G}, since their actions are nontrivial on disjoint subsets of VV. The next theorem collects some more algebraic properties shared by (almost) all graph automaton groups.
A directed path of length tt from the vertex vv to the vertex ww in a graph G=(V,E)G=(V,E) is a sequence of vertices vi0=v,vi1,vi2,,vit=wv_{i_{0}}=v,v_{i_{1}},v_{i_{2}},\ldots,v_{i_{t}}=w in which all edges e1,e2,,ete_{1},e_{2},\ldots,e_{t} are oriented in the direction from vv to ww, that is, one has e1=(v,vi1),e2=(vi1,vi2),,et=(vit1,w)e_{1}=(v,v_{i_{1}}),e_{2}=(v_{i_{1}},v_{i_{2}}),\ldots,e_{t}=(v_{i_{t-1}},w). A directed cycle of length tt is a directed path such that vi0=vitv_{i_{0}}=v_{i_{t}}.

Theorem 3.1.

Let G=(V,E)G=(V,E) be a graph such that |E|2|E|\geq 2. Then the following properties hold.

  1. (1)

    𝒢G\mathcal{G}_{G} is fractal.

  2. (2)

    𝒢G\mathcal{G}_{G} contains an element of finite order and is non-abelian.

  3. (3)

    If GG contains a cycle e1,,ete_{1},\ldots,e_{t}, then (e1ε1etεt)t1(e_{1}^{\varepsilon_{1}}\cdots e_{t}^{\varepsilon_{t}})^{t-1}, with εi{±1}\varepsilon_{i}\in\{\pm 1\}, is a relation in 𝒢G\mathcal{G}_{G} whenever e1ε1,,etεte_{1}^{\varepsilon_{1}},\ldots,e_{t}^{\varepsilon_{t}} is a directed cycle in GG.

  4. (4)

    𝒢G\mathcal{G}_{G} is weakly regular branch over its commutator subgroup 𝒢G\mathcal{G}_{G}^{\prime}.

Proof.
  1. (1)

    We have to show that the map ψ:Stab𝒢G(V)𝒢Gk\psi:Stab_{\mathcal{G}_{G}}(V)\to\mathcal{G}_{G}^{k} given by g(g1,,gk)g\mapsto(g_{1},\ldots,g_{k}), with gi=λ(g,xi)g_{i}=\lambda(g,x_{i}), is surjective on each factor. Let ee be an edge oriented from the vertex xix_{i} to the vertex xjx_{j} (suppose i<ji<j). In order to prove fractalness, it is enough to produce an element of Stab𝒢G(V)Stab_{\mathcal{G}_{G}}(V) whose hh-th restriction is ee, for each h=1,,kh=1,\ldots,k. An explicit computation gives:

    e2=(id,,id,ei-th place,id,,id,ej-th place,id,,id)Stab𝒢G(V).e^{2}=(id,\ldots,id,\underbrace{e}_{i\textrm{-th place}},id,\ldots,id,\underbrace{e}_{j\textrm{-th place}},id,\ldots,id)\in Stab_{\mathcal{G}_{G}}(V).

    Therefore, we can focus on the case where hi,jh\neq i,j. Consider a path of length tt in GG from the vertex xix_{i} to the vertex xhx_{h} through the vertices xi=xi0,xi1,,xit1,xit=xhx_{i}=x_{i_{0}},x_{i_{1}},\ldots,x_{i_{t-1}},x_{i_{t}}=x_{h}. Suppose that the path passes along the edges ei1,,eite_{i_{1}},\ldots,e_{i_{t}}, where the endpoints of eile_{i_{l}} are xil1,xilx_{i_{l-1}},x_{i_{l}}, for each l=1,,tl=1,\ldots,t. Take the group element g:=ei1ε1eitεtg:=e_{i_{1}}^{\varepsilon_{1}}\cdots e_{i_{t}}^{\varepsilon_{t}}, where εs=1\varepsilon_{s}=1 if in the path from xix_{i} to xhx_{h} the edge eise_{i_{s}} is oriented in the direction of the path, and εs=1\varepsilon_{s}=-1 otherwise. One can check that

    g=(id,,id,gi-th place,id,,id)σg=(id,\ldots,id,\underbrace{g}_{i\textrm{-th place}},id,\ldots,id)\sigma

    where σ\sigma is a cyclic permutation of length t+1t+1 such that σ(il)=il1\sigma(i_{l})=i_{l-1} for any l=1,,tl=1,\ldots,t and σ(i)=h\sigma(i)=h. A direct computation gives

    gt+1=(id,,id,gi-th place,,gi1-th place,,gh-th place,id,,id),g^{t+1}=(id,\ldots,id,\underbrace{g}_{i\textrm{-th place}},\ldots,\underbrace{g}_{i_{1}\textrm{-th place}},\ldots,\underbrace{g}_{h\textrm{-th place}},id,\ldots,id),

    i.e., the nontrivial restrictions in the self-similar representation of gt+1g^{t+1} coincide with gg and correspond to the positions i=i0,i1,,it=hi=i_{0},i_{1},\ldots,i_{t}=h. Moreover, notice that gt+1Stab𝒢G(V)g^{t+1}\in Stab_{\mathcal{G}_{G}}(V). Being Stab𝒢G(V)Stab_{\mathcal{G}_{G}}(V) a normal subgroup of 𝒢G\mathcal{G}_{G}, one has that g1e2gStab𝒢G(V)g^{-1}e^{2}g\in Stab_{\mathcal{G}_{G}}(V) and it is easy to check that

    λ(g1e2g,xit)=g1eg.\lambda(g^{-1}e^{2}g,x_{i_{t}})=g^{-1}eg.

    Then we obtain

    λ(gt+1g1e2gg(t+1),xit)=e,\lambda(g^{t+1}g^{-1}e^{2}gg^{-(t+1)},x_{i_{t}})=e,

    i.e., we have moved the generator ee to the hh-th restriction in the self-similar representation of gt+1g1e2gg(t+1)g^{t+1}g^{-1}e^{2}gg^{-(t+1)}. The same method can be applied to any generator and to any hh, and this concludes the proof.

  2. (2)

    Let e=(xi,xj)e=(x_{i},x_{j}) and f=(xj,xh)f=(x_{j},x_{h}), so that ee and ff share the vertex xjx_{j}, and their self-similar representations as generators of 𝒢G{\mathcal{G}_{G}} are

    e=(id,,id,ei-th place,id,,id)(i,j)f=(id,,id,fj-th place,id,,id)(j,h).\displaystyle\hskip 42.67912pte=(id,\ldots,id,\underbrace{e}_{i\textrm{-th place}},id,\ldots,id)(i,j)\quad f=(id,\ldots,id,\underbrace{f}_{j\textrm{-th place}},id,\ldots,id)(j,h).

    A direct computation gives

    [e,f]=(id,,id,fj-th place,id,,id,f1h-th place,id,,id)(i,j,h).[e,f]=(id,\ldots,id,\underbrace{f}_{j\textrm{-th place}},id,\ldots,id,\underbrace{f^{-1}}_{h\textrm{-th place}},id,\ldots,id)(i,j,h).

    Therefore [e,f]id[e,f]\neq id but [e,f]3=id[e,f]^{3}=id.

  3. (3)

    Up to change the orientation of some edges (see Proposition 3.1), we can assume that e1,,ete_{1},\ldots,e_{t} is a directed cycle centered at xix_{i}, i.e., a directed closed path with all edges oriented in the same direction. Suppose that the cycle contains the vertices xi=xj0,,xjt=xix_{i}=x_{j_{0}},\ldots,x_{j_{t}}=x_{i}. In particular, eie_{i} corresponds to the directed edge (xji1,xji)(x_{j_{i-1}},x_{j_{i}}). A direct computation gives

    e1et=(id,,id,e1eti-th place,id,,id)σe_{1}\cdots e_{t}=(id,\ldots,id,\underbrace{e_{1}\cdots e_{t}}_{i\textrm{-th place}},id,\ldots,id)\sigma

    where σ\sigma is a cyclic permutation of length t1t-1 such that σ(i)=i\sigma(i)=i. In particular, σt1=id\sigma^{t-1}=id and so

    (e1et)t1=(id,,id,(e1et)t1i-th place,id,,id)(e_{1}\cdots e_{t})^{t-1}=(id,\ldots,id,\underbrace{(e_{1}\cdots e_{t})^{t-1}}_{i\textrm{-th place}},id,\ldots,id)

    and so (e1et)t1=id(e_{1}\cdots e_{t})^{t-1}=id by Lemma 2.1.

  4. (4)

    We have to prove that, given any kk-tuple (g1,,gk)(g_{1},\ldots,g_{k}), with gi𝒢Gg_{i}\in\mathcal{G}_{G}^{\prime} for each ii, there exists an element g𝒢GStab𝒢G(V)g\in\mathcal{G}_{G}^{\prime}\cap Stab_{\mathcal{G}_{G}}(V) such that ψ(g)=(g1,,gk)\psi(g)=(g_{1},\ldots,g_{k}). We write 𝒢G>𝒢G××𝒢G\mathcal{G}_{G}^{\prime}>\mathcal{G}_{G}^{\prime}\times\cdots\times\mathcal{G}_{G}^{\prime} to denote this condition. First, recall that if ee and ff do not share any vertex, then they commute in 𝒢G\mathcal{G}_{G}, since they act nontrivially on sets of letters which are disjoint.

    It follows that 𝒢G\mathcal{G}_{G}^{\prime} is normally generated by the commutators [e,f][e,f] such that ee and ff share a vertex. So let e,fe,f be generators of 𝒢G\mathcal{G}_{G} such that e=(xi,xj)e=(x_{i},x_{j}) and f=(xj,xk)f=(x_{j},x_{k}). We can suppose, without loss of generality, that i=1,j=2i=1,j=2 and k=3k=3. In particular:

    e2=(e,e,id,,id)f2=(id,f,f,id,,id).e^{2}=(e,e,id,\ldots,id)\qquad f^{2}=(id,f,f,id,\ldots,id).

    A direct computation gives

    [e2,f2]=(id,[e,f],id,,id).[e^{2},f^{2}]=(id,[e,f],id,\ldots,id).

    By proceeding as in the proof of Claim (1), we get that, given an index h2h\neq 2, it is possible to construct an element gg such that

    g1[e2,f2]g=(id,,id,[e,f]h-th place,id,,id)𝒢G.g^{-1}[e^{2},f^{2}]g=(id,\ldots,id,\underbrace{[e,f]}_{h\textrm{-th place}},id,\ldots,id)\in\mathcal{G}_{G}^{\prime}.

    Then, by using that 𝒢G\mathcal{G}_{G}^{\prime} is normal and that 𝒢G\mathcal{G}_{G} is fractal, we get

    (id,,id,[e,f]𝒢Gh-th place,id,,id)𝒢G.(id,\ldots,id,\underbrace{[e,f]^{\mathcal{G}_{G}}}_{h\textrm{-th place}},id,\ldots,id)\subseteq\mathcal{G}_{G}^{\prime}.

    Being hh arbitrary and by applying the same argument to any pair of generators, we get

    𝒢G>𝒢G××𝒢G,\mathcal{G}_{G}^{\prime}>\mathcal{G}_{G}^{\prime}\times\cdots\times\mathcal{G}_{G}^{\prime},

    which is our claim.

If we consider the Case (1) in Example 3.1, where the initial graph contains only one edge, it gives rise to the Adding machine. This group is fractal but it is abelian, free and not weakly regular branch. Basically, this is the only nontrivial case for which the properties (2)-(4) of Theorem 3.1 do not hold.

Let us focus now on the semigroup structure of graph automaton groups. This will allow to know the growth of graph automaton groups (see Corollary 3.1).

Theorem 3.2.

Let G=(V,E)G=(V,E) be a graph such that |E|2|E|\geq 2. Let e,fe,f be edges that share a vertex in GG. Then the semigroup generated by ee and ff is free.

Proof.

By virtue of Proposition 3.2, it is enough to consider the group HH generated by the elements

e=(e,id,id)(1,2)f=(id,id,f)(2,3),e=(e,id,id)(1,2)\qquad f=(id,id,f)(2,3),

since it will be a subgroup of any group 𝒢G\mathcal{G}_{G} associated with a connected graph GG with more than one edge. If the semigroup 𝒮e,f\mathcal{S}_{e,f} is free, then we are done. In particular, we have to prove that if uvu\neq v in {e,f}\{e,f\}^{\ast}, then it must be uvu\neq v in 𝒮e,f\mathcal{S}_{e,f} (notice that, in the notation of Case (2) of Example 3.1, we are going to show that 𝒮a,b1\mathcal{S}_{a,b^{-1}} is free).
Observe that, by using the self-similar representations of ee and ff, we get:

e2=(e,e,id),ef=(e,id,f)(1,3,2),fe=(e,id,f)(1,2,3),f2=(id,f,f).e^{2}=(e,e,id),\quad\ ef=(e,id,f)(1,3,2),\quad\ fe=(e,id,f)(1,2,3),\quad\ f^{2}=(id,f,f).

This implies that any word in {e,f}\{e,f\}^{\ast} of length greater than 11 has restrictions of shorter length. Moreover, if u=(u1,u2,u3)σu=(u_{1},u_{2},u_{3})\sigma and v=(v1,v2,v3)τv=(v_{1},v_{2},v_{3})\tau, then a relation u=vu=v corresponds to the permutation equality σ=τ\sigma=\tau and to the restriction equality ui=viu_{i}=v_{i} in 𝒮e,f\mathcal{S}_{e,f}, for each i=1,2,3i=1,2,3. In particular, if at least one between u,vu,v has length greater than 11, this would give rise to relations ui=viu_{i}=v_{i} such that |ui|+|vi|<|u|+|v||u_{i}|+|v_{i}|<|u|+|v|, for i=1,2,3i=1,2,3.
Now let u=vu=v be a relation in the semigroup with smallest length |u|+|v||u|+|v| in {e,f}\{e,f\}^{\ast}. By the cancellativity of the semigroup, we may assume that the words uu and vv do not start and end with the same letter; in particular, we can suppose u=euu=eu^{\prime} and v=fvv=fv^{\prime}.

Since we have supposed that u=vu=v is a relation in the semigroup with smallest length |u|+|v||u|+|v|, then uu and vv must have the same restrictions (as words in {e,f}\{e,f\}^{\ast}, and not only in 𝒮e,f\mathcal{S}_{e,f}) at each position, otherwise, by considering restrictions, we would get relations of a shorter length. In what follows, we show that in all cases in which uu and vv do not coincide, there exists some restriction in which they differ, and this contradicts minimality.
If u=eu=e^{\ell}, and v=fnvv=f^{n}v^{\prime} (with ,n1\ell,n\geq 1), then by considering the restriction to the position 33 we get the equation id=fv′′id=fv^{\prime\prime}, which is a contradiction by the minimality of the relation u=vu=v. By using a symmetric argument, we may assume

u=emfkz,v=fnetyu=e^{m}f^{k}z,\quad v=f^{n}e^{t}y

for some m,k,n,t1m,k,n,t\geq 1.
Let m,n2m,n\geq 2. By considering the restriction to the position 22 we get ew=fwew=fw^{\prime}, which is a contradiction again.
Therefore, without loss of generality, we may assume m=1m=1. Suppose now that n2n\geq 2. We consider two cases: either zz is the empty word or not. In the first case, the restriction to the position 22 gives id=fwid=fw, and this is a contradiction for the same reason as above. If zz is not the empty word, we necessarily have z=ezz=ez^{\prime} and in this case, looking at the restriction to the position 22, we get ez′′=fwez^{\prime\prime}=fw, a contradiction again.
Thus, we deduce that it must be m=n=1m=n=1, i.e., u=efkzu=ef^{k}z, v=fetyv=fe^{t}y. If z,yz,y are nonempty, then necessarily we have u=efkezu=ef^{k}ez^{\prime}, v=fetfyv=fe^{t}fy^{\prime}. Now, by restricting to the position 22, we get ez′′=fy′′ez^{\prime\prime}=fy^{\prime\prime}, a contradiction. Thus, without loss of generality, we may assume z=z=\emptyset so that u=efku=ef^{k}.
Now if yy is nonempty and v=fetfyv=fe^{t}fy^{\prime}, by considering the restriction to the position 22, we get id=fy′′id=fy^{\prime\prime}, a contradiction. Therefore, we reduce to the case z=y=z=y=\emptyset, that is, we can assume u=efk,v=fetu=ef^{k},v=fe^{t}. Since we have μ(efk,2)=13=μ(fet,2)\mu(ef^{k},2)=1\neq 3=\mu(fe^{t},2), we have a contradiction and the proof is completed. ∎

Corollary 3.1.

Let G=(V,E)G=(V,E) be a graph such that |E|2|E|\geq 2. Then 𝒢G\mathcal{G}_{G} has exponential growth.

Proof.

It follows from the fact that 𝒢G\mathcal{G}_{G} contains a free semigroup of two generators a,ba,b. In fact, in this case, with any generating set containing aa and bb we can construct at least 2n2^{n} distinct group elements of length nn. ∎

Remark 3.3.

Notice that, for the notion of semigroup, the chosen orientation of the edges is important. Observe that the Adding machine, which is isomorphic to the infinite cyclic group \mathbb{Z}, has polynomial growth; on the other hand, the associated graph GG is the path P2P_{2} on two vertices (see Case (1) in Example 3.1), which does not satisfy the hypothesis of Theorem 3.2 and Corollary 3.1.

3.1. A connection to right-angled Artin groups

Let G=(V,E)G=(V,E) be a simple graph. One can construct a group associated with such a graph in the following way: the vertex set VV is the generating set and the only relations are given by the commutators of adjacent vertices. More precisely, given G=(V,E)G=(V,E), the group with presentation

W(G)=vV|vu=uv if {u,v}EW(G)=\langle v\in V|vu=uv\textrm{ if }\{u,v\}\in E\rangle

is the associated right-angled Artin group. For more details about the theory, the reader is referred to [10].

Given a graph G=(V,E)G=(V,E), one can define its dual graph GG^{\prime} to be the graph with vertex set EE, where ee and ff are adjacent in GG^{\prime} if they share a common vertex vv in GG.
Moreover, one can construct the complement G¯\overline{G} of GG having the same vertex set VV, and where two vertices are adjacent if and only if they are not adjacent in GG.

Let G=(V,E)G=(V,E) be a graph. The following proposition shows that there exists a relation between the groups 𝒢G\mathcal{G}_{G} and W(G¯)W(\overline{G^{\prime}}).

Proposition 3.4.

Let G=(V,E)G=(V,E) be a simple graph. Then there exists an epimorphism ϕ:W(G¯)𝒢G\phi:W(\overline{G^{\prime}})\to\mathcal{G}_{G}.

Proof.

First of all notice that the generating set of W(G¯)W(\overline{G^{\prime}}) is precisely EE (up to consider inverses). Let eEe\in E, we define ϕ(e)=e\phi(e)=e, where ee is supposed to be oriented in GG. The map ϕ\phi is a well defined homomorphism. Moreover, the set of adjacent vertices in G¯\overline{G^{\prime}} corresponds exactly to those edges in GG that do not share any common vertex. These edges commute by definition, when regarded as elements of 𝒢G\mathcal{G}_{G}. In this way, we have that the set of relations in W(G¯)W(\overline{G^{\prime}}) is contained in the set of relations of 𝒢G\mathcal{G}_{G}. This concludes the proof. ∎

4. Schreier graphs

In this section we recall the notion of Schreier graphs and we study some properties of them in the context of graph automaton groups.
Let 𝒢\mathcal{G} be a finitely generated group with a set SS of generators such that idSid\not\in S and S=S1S=S^{-1}, and suppose that 𝒢\mathcal{G} acts on a set MM. Then one can consider a graph Γ(𝒢,S,M)\Gamma(\mathcal{G},S,M) with vertex set MM, where two vertices m,mm,m^{\prime} are joined by an edge if there exists sSs\in S such that s(m)=ms(m)=m^{\prime}. If this is the case, we label the edge from mm to mm^{\prime} by ss, and the edge from mm^{\prime} to mm by s1s^{-1}. Equivalently, we can think that the same (undirected) edge is labeled by ss near mm and by s1s^{-1} near mm^{\prime}.
If the action of 𝒢\mathcal{G} on MM is transitive, then the graph Γ(𝒢,S,M)\Gamma(\mathcal{G},S,M) is connected and corresponds to the classical notion of Schreier graph Γ(𝒢,S,Stab𝒢(m))\Gamma(\mathcal{G},S,Stab_{\mathcal{G}}(m)) of the group 𝒢\mathcal{G} with respect to the stabilizer subgroup Stab𝒢(m)Stab_{\mathcal{G}}(m) for some (any) mMm\in M (see [27]).

Definition 4.1.

Let 𝒜=(S,X,λ,μ)\mathcal{A}=(S,X,\lambda,\mu) be an invertible automaton and let G(𝒜)G(\mathcal{A}) be the associated automaton group. The nn-th Schreier graph Γn=Γn(G(𝒜),S,Xn)\Gamma_{n}=\Gamma_{n}(G(\mathcal{A}),S,X^{n}) is the Schreier graph given by the action of G(𝒜)G(\mathcal{A}) over XnX^{n}, with respect to the generators given by SS and their inverses.

Notice that, although Schreier graphs are defined as directed and labeled graphs, for our purposes we will often consider them as undirected and unlabeled graphs.

In what follows, we denote by ΓnG\Gamma_{n}^{G} the nn-th Schreier graph of the graph automaton group 𝒢G\mathcal{G}_{G}. The vertex set of ΓnG\Gamma_{n}^{G} is identified with VnV^{n}, where G=(V,E)G=(V,E). Observe that Γ1G\Gamma_{1}^{G} coincides with GG up to remove loops and multi-edges from Γ1G\Gamma_{1}^{G}. Therefore, we can say that Γ1G\Gamma_{1}^{G} and GG coincide as simple graphs. Moreover, the Schreier graph ΓnG\Gamma_{n}^{G} is a regular graph of degree 2|E|2|E| by definition.

Example 4.1.

The Schreier graphs ΓnG\Gamma_{n}^{G}, for n=1,2,3,4n=1,2,3,4, of the Tangled odometer group introduced in Example 3.1, obtained when GG is the path P3P_{3} on 33 vertices, are shown in Fig. 4 and Fig. 5. Notice that infinite Schreier graphs of the same group, with a different system of generators, have been classified in [11].

\psfrag{0}{$1$}\psfrag{1}{$2$}\psfrag{2}{$3$}\psfrag{11}{$22$}\psfrag{00}{$11$}\psfrag{22}{$33$}\psfrag{02}{$13$}\psfrag{20}{$31$}\psfrag{111}{$222$}\psfrag{000}{$111$}\psfrag{222}{$333$}\psfrag{202}{$313$}\psfrag{020}{$131$}\includegraphics[width=364.19664pt]{livello3eti.eps}
Figure 4. The Schreier graphs Γ1P3\Gamma_{1}^{P_{3}}, Γ2P3\Gamma_{2}^{P_{3}}, Γ3P3\Gamma_{3}^{P_{3}} of the Tangled odometer group.
\psfrag{2020}{$3131$}\psfrag{0202}{$1313$}\psfrag{1111}{$2222$}\psfrag{0000}{$1111$}\psfrag{2222}{$3333$}\psfrag{2223}{$2223$}\psfrag{2113}{$2113$}\psfrag{2123}{$2123$}\psfrag{2213}{$2213$}\psfrag{2313}{$2313$}\psfrag{3113}{$3113$}\psfrag{3123}{$3123$}\includegraphics[width=409.71689pt]{livello4eti.eps}
Figure 5. The Schreier graph Γ4P3\Gamma_{4}^{P_{3}} of the Tangled odometer group.

The Schreier graph ΓnG\Gamma_{n}^{G} only depends on the initial graph G=(V,E)G=(V,E) and not on the orientation of its edges, since the generating set that we let act on VnV^{n} is supposed to be symmetric. The following lemma is well known [1].

Lemma 4.1.

Γn+1G\Gamma_{n+1}^{G} is a covering of ΓnG\Gamma_{n}^{G} of degree |V||V|.

Sketch of the Proof.

The basic idea is that, if vxvx and wywy are adjacent vertices in Γn+1G\Gamma_{n+1}^{G}, with v,wVnv,w\in V^{n} and x,yVx,y\in V, then vv and ww are adjacent in ΓnG\Gamma_{n}^{G}. ∎

Proposition 4.1.

For every n1n\geq 1, the graph ΓnG\Gamma_{n}^{G} is connected if and only if GG is. In particular, the group 𝒢G\mathcal{G}_{G} is spherically transitive if and only if GG is connected.

Proof.

Let us start by proving that, if GG is connected, then ΓnG\Gamma_{n}^{G} is connected for any n1n\geq 1. The connectedness of GG implies that 𝒢G\mathcal{G}_{G} acts transitively on VV. By Theorem 3.1, the group 𝒢G\mathcal{G}_{G} is fractal and it is a standard inductive argument to show that these properties imply the transitivity of 𝒢G\mathcal{G}_{G} on VnV^{n}.
Conversely, if x,yVx,y\in V are not in the same connected component of GG, then there is no way to connect the vertices xnx^{n} and yny^{n} in ΓnG\Gamma_{n}^{G}. ∎

We are going to show how cut-vertices in GG propagate in the Schreier graphs ΓnG\Gamma_{n}^{G}. Recall that a cut-vertex of a graph is a vertex whose deletion increases the number of connected components of the graph (see, for instance, [4]). Observe that, for our purposes, when we delete a cut-vertex from a graph we do not remove the edges which are incident to that vertex. We need some technical preparation.
Let xx be a cut-vertex of GG, whose deletion disconnects GG into cc connected components G1,,GcG_{1},\ldots,G_{c} with Gi=(Vi,Ei)G_{i}=(V_{i},E_{i}) for each i=1,,ci=1,\ldots,c. If a vertex vv of VnV^{n} has the form v=xkyvv=x^{k}yv^{\prime}, with k1k\geq 1, and yViy\in V_{i}, then a special subgraph of ΓnG\Gamma_{n}^{G} associated with the vertex vv can be constructed as follows.

  • Let EEiE\setminus E_{i} act on vv. Since only edges not belonging to EiE_{i} are acting, this action can only change the prefix xkx^{k} of vv, but the suffix yvyv^{\prime} remains unchanged. Moreover, each edge eEEie\in E\setminus E_{i} generates an Adding machine, so that its action on vv produces an orbit which is a cyclic graph whose length is a power of 22, and in particular it contains vv again. Let us denote by Xv,0X_{v,0} the orbit of vv under EEiE\setminus E_{i}. Then put Yv,0=Xv,0{v}Y_{v,0}=X_{v,0}\setminus\{v\}.

  • Let EiE_{i} act on Yv,0Y_{v,0} and get the set Xv,1X_{v,1}. Concretely, we are appending new cycles of length a power of 22 to the vertices contained in the cycles constructed at the previous step. Then put Yv,1=Xv,1{v}Y_{v,1}=X_{v,1}\setminus\{v\}.

  • Let EEiE\setminus E_{i} act Yv,1Y_{v,1} and get the set Xv,2X_{v,2}. Then put Yv,2=Xv,2{v}Y_{v,2}=X_{v,2}\setminus\{v\}.

Continue in this way by alternating the action of generators in EEiE\setminus E_{i} and EiE_{i}; in this way, we construct an increasing sequence Yv,mYv,m+1Y_{v,m}\subseteq Y_{v,m+1}. Since our alphabet is finite, after a finite number of steps, the sequence of sets Yv,mY_{v,m} stabilizes to a set YvY_{v}. Let DvD_{v} be the graph induced by YvY_{v} (in particular, DvD_{v} contains the vertex vv itself). We call the corresponding subgraph DvD_{v} of ΓnG\Gamma_{n}^{G} the decoration of vv in ΓnG\Gamma_{n}^{G}.

Remark 4.1.

If v=xkyvv=x^{k}yv^{\prime} then the decoration DvD_{v} is isomorphic to the subgraph of ΓkG\Gamma_{k}^{G} obtained by considering the alternate action of EEiE\setminus E_{i} and EiE_{i} on xkx^{k} as described above. In fact, in each vertex of DvD_{v} the suffix yvyv^{\prime} remains unchanged. In particular, the isomorphism is obtained by deleting the last nkn-k letters of each vertex of DvD_{v}. Moreover if the vertex xx is a leaf in GG, that is, it has only one adjacent vertex, then it gives rise in ΓnG\Gamma_{n}^{G} to special cut-vertices which separate a component that is a loop.

Example 4.2.

Consider the Schreier graph Γ4P3\Gamma_{4}^{P_{3}} of Fig. 5, where G=P3G=P_{3} is the path on 33 vertices (as represented in Fig. 2). Take the vertex v=2223v=2223, where x=2x=2 disconnects the path P3P_{3} into the two components (V1,E1)=({1},{a})(V_{1},E_{1})=(\{1\},\{a\}) and (V2,E2)=({3},{b})(V_{2},E_{2})=(\{3\},\{b\}). In particular, y=3V2y=3\in V_{2}. Let us construct the decoration of vv. We let the generator aa act on vv obtaining the 88-cycle on the right of vv. Now we let bb act on all vertices of this cycle different from vv: we obtain two 22-cycles attached to the vertices 21132113 and 21232123, a 44-cycle attached to 22132213, together with four loops attached to the remaining vertices of the 88-cycle. Then we let aa act again and we obtain: two loops attached to the vertices 31133113 and 31233123, two loops attached to two vertices of the 44-cycle, and the 22-cycle containing 23132313 and 13131313. We conclude by letting bb act, what produces the loop attached to the rightmost vertex 13131313.
Roughly speaking, the decoration of vv is the subgraph of Γ4P3\Gamma_{4}^{P_{3}} containing vv and all the vertices on the right of vv. Notice that such a subgraph is isomorphic to the subgraph of Γ3P3\Gamma_{3}^{P_{3}} (Fig. 4) obtained by taking the vertex 222222 and all the vertices on its left.

Notice that by construction every decoration DvD_{v} corresponds to a subgraph that can be disconnected from the Schreier graph just by removing the vertex vv. What said above can be summarized in the following result.

Proposition 4.2.

Let xx be a cut-vertex for GG. Then, for each n2n\geq 2, the vertex v=xwVnv=xw\in V^{n} is a cut-vertex in ΓnG\Gamma_{n}^{G} for every wVn1w\in V^{n-1}, and it separates the decoration DvD_{v} from the remaining part of the graph ΓnG\Gamma_{n}^{G}.

Remark 4.2.

There exists a connection between our construction and a special class of graph products, which allows to give a purely combinatorial construction of the Schreier graphs of graph automaton groups. Actually, this description was our first attempt in defining the graphs ΓnG\Gamma_{n}^{G} without observing that ΓnG\Gamma_{n}^{G} arises from the action of an automaton group. The construction was inspired by the definition of Sierpiński graphs (see, for instance, [26] and bibliography therein), although in order to generate the latter by automata we should use a partial automaton.
Let G=(V,E)G=(V,E) be a finite graph. One can define the nn-th automaton power GanG^{n}_{a} of GG as the graph with vertex set VnV^{n} and edge set consisting of pairs of vertices of type

(3) {xtyw,ytxw}{xtzw,ytzw}{xn,yn}\{x^{t}yw,y^{t}xw\}\qquad\{x^{t}zw,y^{t}zw\}\qquad\{x^{n},y^{n}\}

where {x,y}E\{x,y\}\in E and zx,yz\neq x,y, with t=0,1,,n1t=0,1,\ldots,n-1 and |w|=nt1|w|=n-t-1.
It is easy to check that the edges described above are exactly the edges of ΓnG\Gamma_{n}^{G}. In fact the connections described by (3) precisely correspond to the action of the states of the automaton 𝒜G\mathcal{A}_{G} on VnV^{n}. Therefore, the graphs ΓnG\Gamma_{n}^{G} and GanG^{n}_{a} are isomorphic.

4.1. Automorphisms of Schreier graphs of a graph automaton group

In this subsection we are going to investigate the relation between the automorphisms of the Schreier graph ΓnG\Gamma_{n}^{G} and the automorphisms of the initial graph G=(V,E)G=(V,E). Observe that Proposition 4.2 implies that any nontrivial automorphism of a decoration DvD_{v} in ΓnG\Gamma_{n}^{G} is a nontrivial automorphism of ΓnG\Gamma_{n}^{G} fixing vv. This observation will lead us to the description of the full automorphism group of the Schreier graphs associated with path graphs (Section 4.2). We start with an extension result.

Proposition 4.3.

Let ϕ\phi be an automorphism of G=(V,E)G=(V,E), with V={x1,x2,,xk}V=\{x_{1},x_{2},\ldots,x_{k}\}. Then ϕn:ΓnGΓnG\phi_{n}:\Gamma_{n}^{G}\to\Gamma_{n}^{G} defined by

ϕn(xi1xin)=ϕ(xi1)ϕ(xin)\phi_{n}(x_{i_{1}}\cdots x_{i_{n}})=\phi(x_{i_{1}})\cdots\phi(x_{i_{n}})

is an automorphism of ΓnG\Gamma_{n}^{G}.

Proof.

First of all, notice that the map ϕn\phi_{n} is a bijection by definition. We have to prove that, if vv, vVnv^{\prime}\in V^{n} are adjacent vertices in ΓnG\Gamma_{n}^{G}, then ϕn(v)\phi_{n}(v) and ϕn(v)\phi_{n}(v^{\prime}) are adjacent too.
The adjacency in the graph ΓnG\Gamma_{n}^{G} are described by the rules from Eq. (3). We have three possibilities: either v=xtywv=x^{t}yw and v=ytxwv^{\prime}=y^{t}xw, or v=xtzwv=x^{t}zw and v=ytzwv^{\prime}=y^{t}zw, or v=xnv=x^{n} and v=ynv^{\prime}=y^{n}, where {x,y}E\{x,y\}\in E and zx,yz\neq x,y, with t=0,1,,n1t=0,1,\ldots,n-1 and |w|=nt1|w|=n-t-1. Since the vertices vv and vv^{\prime} are supposed to be adjacent in ΓnG\Gamma_{n}^{G} then in all the three cases described above the vertices xx and yy are adjacent in GG. This implies that ϕ(x)\phi(x) and ϕ(y)\phi(y) are adjacent in GG, because ϕ\phi is an automorphism of G=(V,E)G=(V,E). If v=xtywv=x^{t}yw and v=ytxwv^{\prime}=y^{t}xw, then

ϕn(v)=ϕ(x)tϕ(y)ϕnt1(w),ϕn(v)=ϕ(y)tϕ(x)ϕnt1(w);\phi_{n}(v)=\phi(x)^{t}\phi(y)\phi_{n-t-1}(w),\ \ \phi_{n}(v^{\prime})=\phi(y)^{t}\phi(x)\phi_{n-t-1}(w);

it follows from Eq. (3) that ϕn(v)\phi_{n}(v) and ϕn(v)\phi_{n}(v^{\prime}) are adjacent. The other two cases can be treated analogously. ∎

When the graph GG is cyclic, we can prove that actually all automorphisms of ΓnG\Gamma_{n}^{G} are of this type. Let us denote by D2kD_{2k} the dihedral group of 2k2k elements.

Theorem 4.1.

Let CkC_{k} be the cyclic graph on kk vertices. Then

Aut(ΓnCk)Aut(Ck)D2k,for each n1.Aut(\Gamma_{n}^{C_{k}})\cong Aut(C_{k})\cong D_{2k},\qquad\mbox{for each }n\geq 1.
Proof.

Observe that for n=1n=1 the statement is obvious by definition of Γ1Ck\Gamma_{1}^{C_{k}}. Therefore, we assume n2n\geq 2. Let V={1,,k}V=\{1,\ldots,k\} be the vertex set of Ck=(V,E)C_{k}=(V,E). Observe that in ΓnCk\Gamma_{n}^{C_{k}} one has kn1k^{n-1} sets of vertices of type Vwk={1w,,kw}V_{w}^{k}=\{1w,\ldots,kw\}, with wVn1w\in V^{n-1}, each yielding a copy of CkC_{k} in ΓnCk\Gamma_{n}^{C_{k}}. This property can be obtained by taking t=0t=0 in the rule {xtyw,ytxw}\{x^{t}yw,y^{t}xw\} described in Eq. (3). Moreover, one has a copy of CkC_{k} consisting of the vertices labeled by {1n,2n,,kn}\{1^{n},2^{n},\ldots,k^{n}\}, according to the rule {xn,yn}\{x^{n},y^{n}\} in Eq. (3).
Let φ\varphi be an automorphism of ΓnCk\Gamma_{n}^{C_{k}}. First of all, we want to prove that the set X={1n,,kn}X=\{1^{n},\ldots,k^{n}\} is invariant under the action of φ\varphi. To prove that, we show that a vertex of ΓnCk\Gamma_{n}^{C_{k}} is the unique common vertex of two copies of CkC_{k} if and only if it belongs to XX.

In order to avoid heavy notation, we omit here and in the sequel to specify every time that sums and differences must be taken modulo kk. We denote by eie_{i} the generator associated with the edge (i,i+1)(i,i+1) of CkC_{k}. It follows from Eq. (3) that the neighbours of ini^{n} in ΓnCk\Gamma_{n}^{C_{k}} are exactly the vertices (i+1)in1(i+1)i^{n-1}, (i1)in1(i-1)i^{n-1} (together with ini^{n}, they belong to the copy of CkC_{k} associated with Vin1kV_{i^{n-1}}^{k}) and the vertices (i1)n(i-1)^{n}, (i+1)n(i+1)^{n} (together with ini^{n}, they belong to the copy of CkC_{k} given by XX). Therefore, for n2n\geq 2, the vertices of XX have the required property. We have to prove that no other vertex has this property.

Notice that a vertex of type xtywx^{t}yw or xtzwx^{t}zw, with t=1t=1 and {x,y}E\{x,y\}\in E, {x,z}E\{x,z\}\not\in E, has less than four distinct neighbours according to Eq. (3), and so it cannot be the unique common vertex of two copies of CkC_{k}. Hence, for n=2n=2, our characterization of XX is proved. From now, we assume n3n\geq 3 and we only consider vertices of type xtywx^{t}yw or xtzwx^{t}zw for 2tn12\leq t\leq n-1.

We want to prove that a vertex of this type cannot be the unique common vertex of two copies of CkC_{k}, although it has four distinct adjacent vertices in ΓnCk\Gamma_{n}^{C_{k}}.
Let us start by considering words of type is(i±1)wi^{s}(i\pm 1)w, with s2s\geq 2 and ww possibly empty (i.e., words starting with a block of ii’s followed by a neighbour of ii).
Let us focus our attention on the case is(i+1)wi^{s}(i+1)w (the case is(i1)wi^{s}(i-1)w is analogous). Its distinct neighbours are:

  • (i+1)is1(i+1)w(i+1)i^{s-1}(i+1)w, via the action of ei1e_{i}^{-1}, and (i1)is1(i+1)w(i-1)i^{s-1}(i+1)w, via the action of ei1e_{i-1} (together with the vertex is(i+1)wi^{s}(i+1)w, they both belong to the copy of CkC_{k} associated with Vis1(i+1)wkV_{i^{s-1}(i+1)w}^{k});

  • (i+1)siw(i+1)^{s}iw, via the action of eie_{i}, and (i1)s(i+1)w(i-1)^{s}(i+1)w, via the action of ei11e_{i-1}^{-1}.

We claim that is(i+1)wi^{s}(i+1)w, (i+1)siw(i+1)^{s}iw and (i1)s(i+1)w(i-1)^{s}(i+1)w do not belong together to a copy of CkC_{k}. In particular we claim that we cannot start from is(i+1)wi^{s}(i+1)w , pass to (i+1)siw(i+1)^{s}iw and go back in k1k-1 steps to is(i+1)wi^{s}(i+1)w passing in the last step through (i1)s(i+1)w(i-1)^{s}(i+1)w by avoiding vertex repetitions.
Notice that by applying eie_{i} to is(i+1)wi^{s}(i+1)w we get (i+1)siw(i+1)^{s}iw. Such a vertex belongs to the copy of CkC_{k} corresponding to V(i+1)s1iwkV_{(i+1)^{s-1}iw}^{k} (its neighbours in this copy of CkC_{k} are obtained by applying the generators ei+11e_{i+1}^{-1} and eie_{i}). The other cycle to which the vertex should belong must be obtained by applying the generator ei+1e_{i+1} and ei1e_{i}^{-1}. By using the latter we go back to is(i+1)wi^{s}(i+1)w, by using the former we get (i+2)siw(i+2)^{s}iw. In the same way by applying k2k-2 times the generators we get a sequence of vertices until we get (i1)siw(i-1)^{s}iw. However, this vertex is not equal to the vertex (i1)s(i+1)w(i-1)^{s}(i+1)w, which was supposed to be the last vertex in our copy of CkC_{k}. Hence we cannot go back in kk steps to is(i+1)wi^{s}(i+1)w.

Consider now vertices of type iszwi^{s}zw, with s2s\geq 2, z(i±1)z\neq(i\pm 1) and ww possibly empty (i.e., words starting with a block of ii’s followed by a letter non adjacent to ii in CkC_{k}). It can be easily seen that such vertices have four distinct neighbours: the vertices (i±1)is1zw(i\pm 1)i^{s-1}zw (together with iszwi^{s}zw, they belong to the copy of CkC_{k} associated with Vis1zwkV_{i^{s-1}zw}^{k}), and the vertices (i±1)szw(i\pm 1)^{s}zw. For the latter pair of vertices, one can use the same argument as above to show that they cannot live, together with iszwi^{s}zw, in a copy of CkC_{k}.
We have thus proved our characterization of XX for each nn, which implies that XX is invariant under the action of φ\varphi.

Now suppose that XX is fixed (not only invariant) by φ\varphi. It is possible to prove by induction on nn that in this case φ\varphi is the trivial automorphism. The key idea is that the graph ΓnCk\Gamma_{n}^{C_{k}} is obtained by gluing together, in a suitable way, kk copies of the graph Γn1Ck\Gamma_{n-1}^{C_{k}}, each consisting of the vertex set Gi={wi:wVn1}G_{i}=\{wi:w\in V^{n-1}\}, with i=1,,ki=1,\ldots,k.
We have to consider now the case where XX is not fixed by φ\varphi. Let us denote by φX\varphi_{X} the automorphism of CkC_{k} defined by putting φX(i)=j\varphi_{X}(i)=j if φ(in)=jn\varphi(i^{n})=j^{n}. Now, let ψ\psi be the automorphism of ΓnCk\Gamma_{n}^{C_{k}} induced by φX\varphi_{X} as in Proposition 4.3. By definition, the composition of φ\varphi and ψ1\psi^{-1} gives an automorphism of ΓnCk\Gamma_{n}^{C_{k}} fixing XX. By the previous discussion, we get φ=ψ\varphi=\psi, and the thesis follows. ∎

From a geometric point of view inherited from CkC_{k} via Proposition 4.3, a nontrivial automorphism of ΓnCk\Gamma_{n}^{C_{k}} is a composition of reflections around the axis of the path connecting vertices of type ini^{n} and (i+1)n(i+1)^{n}, with rotations by 2π/k2\pi/k.

Example 4.3.

In Fig. 6 the graphs Γ2C3\Gamma_{2}^{C_{3}} (on the left), Γ3C3\Gamma_{3}^{C_{3}} (in the middle) and Γ2C4\Gamma_{2}^{C_{4}} (on the right) are depicted (notice that the automaton associated with the graph C3C_{3} is represented in Case (3) of Example 3.1). It can be easily seen that Aut(Γ2C3)=Aut(Γ2C3)=D6Aut(\Gamma_{2}^{C_{3}})=Aut(\Gamma_{2}^{C_{3}})=D_{6} and Aut(Γ2C4)=D8Aut(\Gamma_{2}^{C_{4}})=D_{8}. Looking, for instance, at the copy G2G_{2} of Γ2C3\Gamma_{2}^{C_{3}} contained in Γ3C3\Gamma_{3}^{C_{3}}, we see that when passing from the level 22 to the level 33 the edge {22,11}\{22,11\} (resp. {22,33}\{22,33\}) produces the edges {221,112}\{221,112\} and {222,111}\{222,111\} (resp. {223,332}\{223,332\} and {222,333}\{222,333\}) connecting the copy G2G_{2} with the copy G1G_{1} (resp. G3G_{3}); on the other hand, the edges {222,112}\{222,112\} and {222,332}\{222,332\} do not appear in Γ3C3\Gamma_{3}^{C_{3}}.

\psfrag{00}{$11$}\psfrag{11}{$22$}\psfrag{22}{$33$}\psfrag{33}{$44$}\psfrag{000}{$111$}\psfrag{111}{$222$}\psfrag{222}{$333$}\psfrag{002}{$113$}\psfrag{220}{$331$}\psfrag{110}{$221$}\psfrag{001}{$112$}\psfrag{221}{$332$}\psfrag{112}{$223$}\psfrag{a}{$33$}\psfrag{b}{$13$}\psfrag{c}{$31$}\psfrag{d}{$11$}\psfrag{e}{$21$}\psfrag{f}{$12$}\psfrag{g}{$22$}\psfrag{h}{$32$}\psfrag{i}{$23$}\psfrag{G1}{$G_{1}$}\psfrag{G2}{$G_{2}$}\psfrag{G3}{$G_{3}$}\includegraphics[width=418.82372pt]{figurecicliche.eps}
Figure 6. The graphs Γ2C3\Gamma_{2}^{C_{3}}, Γ3C3\Gamma_{3}^{C_{3}} and Γ2C4\Gamma_{2}^{C_{4}}.
Remark 4.3.

The previous theorem can be used to show that each isomorphism class of infinite Schreier graphs associated with the action of the group 𝒢Ck\mathcal{G}_{C_{k}} contains finitely many graphs. In fact any isomorphism of infinite graphs induces an automorphism of finite graphs ΓnCk\Gamma_{n}^{C_{k}} and, since the number of such automorphisms is bounded, one gets the assert. The same phenomenon have been already noticed in [7] for the Hanoi Towers group.

4.2. The case of a path graph PkP_{k}

In this subsection, we give a precise description of the diameter and of the automorphism group of the Schreier graph ΓnG\Gamma_{n}^{G} in the case where GG is a path graph.
A path PkP_{k} on kk vertices, that we denote by {1,,k}\{1,\ldots,k\}, is a tree with two leaves and k2k-2 vertices of degree 22 (see Fig. 7). We will call extremal the edges containing the two leaves (denoted by e1e_{1} and ek1e_{k-1}) and internal the other ones. We have already remarked in Example 3.1 that the group 𝒢P2\mathcal{G}_{P_{2}} is isomorphic to \mathbb{Z} (whose nn-th Schreier graph is a cyclic graph of length 2n2^{n}) and that the group 𝒢P3\mathcal{G}_{P_{3}} is the so-called Tangled odometer group (whose first four Schreier graphs are drawn in Fig. 4 and Fig. 5).
Given a graph G=(V,E)G=(V,E), we denote by d(u,v)d(u,v) the geodesic distance between uu and vv, that is, the length of a shortest path in GG connecting uu and vv. Then the diameter of GG is defined as

diam(G)=max{d(u,v):u,vV}.diam(G)=\max\{d(u,v):u,v\in V\}.
\psfrag{G}{$P_{k}$}\psfrag{0}{$1$}\psfrag{1}{$2$}\psfrag{2}{$3$}\psfrag{k-2}{$k-1$}\psfrag{k-1}{$k$}\psfrag{e1}{$e_{1}$}\psfrag{e2}{$e_{2}$}\psfrag{ek-1}{$e_{k-1}$}\includegraphics[width=318.66946pt]{figpk.eps}
Figure 7. The path graph PkP_{k}.

First we want to understand the structure of the graphs ΓnPk\Gamma_{n}^{P_{k}}. Fix k3k\geq 3. Since PkP_{k} is a tree, all its k2k-2 vertices of degree two are cut-vertices, and so, according to Proposition 4.2 they give rise to cut-vertices in ΓnPk\Gamma_{n}^{P_{k}}. The two leaves correspond to loops (see Remark 4.1).

Example 4.4.

The Schreier graphs ΓnP4\Gamma^{P_{4}}_{n}, for n=1,2,3n=1,2,3 are shown in Fig. 8 and Fig. 9.

\psfrag{0}{$1$}\psfrag{1}{$2$}\psfrag{2}{$3$}\psfrag{3}{$4$}\psfrag{11}{$22$}\psfrag{00}{$11$}\psfrag{22}{$33$}\psfrag{33}{$44$}\psfrag{30}{$41$}\psfrag{03}{$14$}\includegraphics[width=364.19664pt]{p4inizio.eps}
Figure 8. The Schreier graphs Γ1P4\Gamma^{P_{4}}_{1} and Γ2P4\Gamma^{P_{4}}_{2}.
\psfrag{111}{$222$}\psfrag{000}{$111$}\psfrag{222}{$333$}\psfrag{333}{$444$}\psfrag{303}{$414$}\psfrag{030}{$141$}\psfrag{221}{$221$}\psfrag{334}{$334$}\includegraphics[width=364.19664pt]{p4tosto.eps}
Figure 9. The Schreier graph Γ3P4\Gamma^{P_{4}}_{3}.

Notice that in this case we can observe a “wedge”shape of the Schreier graphs, contrary to the straight shape of the graphs ΓnP3\Gamma^{P_{3}}_{n} shown in Fig. 4 and Fig. 5. This property depends on the existence of an internal edge in P4P_{4}, which does not appear in P3P_{3}.

Our aim is to describe how one can recursively construct the graph ΓnPk\Gamma_{n}^{P_{k}} starting from Γn1Pk\Gamma_{n-1}^{P_{k}}. Following [5], we can proceed as follows.

We take kk copies of Γn1Pk\Gamma_{n-1}^{P_{k}} and append to the end of the vertices of the ii-th copy the letter ii, for i=1,,ki=1,\ldots,k. Then, for each i=2,,k1i=2,\ldots,k-1, we remove the edges {in,(i1)n1i}\{i^{n},(i-1)^{n-1}i\} and {in,(i+1)n1i}\{i^{n},(i+1)^{n-1}i\}. We also remove the edges {1n,2n11}\{1^{n},2^{n-1}1\} and {kn,(k1)n1k}\{k^{n},(k-1)^{n-1}k\}. Finally, for i=1,,k1i=1,\ldots,k-1, we join the ii-th and (i+1)(i+1)-th copies by adding the edges {in,(i+1)n}\{i^{n},(i+1)^{n}\} and {(i+1)n1i,in1(i+1)}\{(i+1)^{n-1}i,i^{n-1}(i+1)\}. The last operation gives rise to new cycles of doubled length with respect to the level n1n-1.

Example 4.5.

In Fig. 10 the construction of Γ3P4\Gamma_{3}^{P_{4}} starting from 44 copies of Γ2P4\Gamma_{2}^{P_{4}} is shown. The copies are separated by dotted lines; the deleted edges are represented by dashed lines; the new edges producing cycles of length 88 are in bold lines.

\psfrag{111}{$222$}\psfrag{000}{$111$}\psfrag{222}{$333$}\psfrag{333}{$444$}\psfrag{j}{$332$}\psfrag{k}{$223$}\psfrag{u}{$221$}\psfrag{v}{$112$}\psfrag{y}{$443$}\psfrag{z}{$334$}\psfrag{m}{$141$}\psfrag{n}{$414$}\psfrag{c1}{copy $1$}\psfrag{c2}{copy $2$}\psfrag{c3}{copy $3$}\psfrag{c4}{copy $4$}\includegraphics[width=364.19664pt]{copie.eps}
Figure 10. The construction of Γ3P4\Gamma^{P_{4}}_{3} from Γ2P4\Gamma^{P_{4}}_{2}.

Recall that each generator eie_{i}, i=1,,k1i=1,\ldots,k-1 corresponds to an edge of PkP_{k} connecting two consecutive vertices ii and i+1i+1 (Fig. 7). More effectively, the action of eie_{i} on vertices of types wxvVnwxv\in V^{n} with w{i,i+1}kw\in\{i,i+1\}^{k}, xi,i+1x\neq i,i+1 and vv arbitrary (with x,vx,v possibly empty) gives rise to a cycle of length 2k2^{k}, since eie_{i} acts on {i,i+1}k\{i,i+1\}^{k} like an Adding machine. For i=1,,k2i=1,\ldots,k-2, such cycles share vertices of type (i+1)kxv(i+1)^{k}xv with cycles corresponding to the action of the generator ei+1e_{i+1}, which is the only other generator that acts nontrivially on the letter i+1i+1. In particular, maximal cycles in ΓnPk\Gamma_{n}^{P_{k}} are those containing the vertices of type ini^{n}: for i=2,,k2i=2,\ldots,k-2, such vertices belong to the two cycles of length 2n2^{n} and labeled by eie_{i} and ei+1e_{i+1}, whereas for i=1i=1 (resp. i=k1i=k-1) the vertex ini^{n} belong to the unique maximal cycle labeled by e1e_{1} (resp. ek1e_{k-1}). It follows that ΓnPk\Gamma_{n}^{P_{k}} has a cactus structure, where adjacent cycles can be labeled by generators which correspond to incident edges in PkP_{k}.
Let us analyze the behaviour of maximal cycles when constructing ΓnPk\Gamma_{n}^{P_{k}} from Γn1Pk\Gamma_{n-1}^{P_{k}}. Notice that cycles of Γn1Pk\Gamma_{n-1}^{P_{k}} labeled by eie_{i} containing vertices of type wxvVn1wxv\in V^{n-1} with w{i,i+1}kw\in\{i,i+1\}^{k} and xi,i+1x\neq i,i+1 nonempty, also appear in ΓnPk\Gamma_{n}^{P_{k}}: more precisely, they correspond to cycles containing vertices of type wxvyVnwxvy\in V^{n} with w{i,i+1}kw\in\{i,i+1\}^{k}, y=1,,ky=1,\ldots,k and xi,i+1x\neq i,i+1 nonempty. Maximal cycles in Γn1Pk\Gamma_{n-1}^{P_{k}}, which are the cycles containing the vertices in1i^{n-1}, appear in ΓnPk\Gamma_{n}^{P_{k}}. They correspond to nonmaximal cycles containing vertices in1xVni^{n-1}x\in V^{n}, for xi,i+1x\neq i,i+1 nonempty. For x=i,i+1x=i,i+1 the generator eie_{i} gives rise to a new bigger (doubled length) unique cycle of length 2n2^{n} in ΓnPk\Gamma_{n}^{P_{k}}. All these maximal cycles in ΓnPk\Gamma_{n}^{P_{k}} are connected through the path 1n,2n,,kn1^{n},2^{n},\ldots,k^{n} (see, for instance, the path 13,23,33,431^{3},2^{3},3^{3},4^{3} in Fig. 10).

As explained in Fig. 11 and Fig. 12 each new maximal cycle generated by eie_{i} has attached decorations isomorphic to the decorations appended to vertices belonging to the biggest cycle labeled by eie_{i} in Γn1Pk\Gamma_{n-1}^{P_{k}}. More precisely, the decorations DvD_{v} and DwD_{w} corresponding to the vertices v=(i+1)n1iv=(i+1)^{n-1}i and w=in1(i+1)w=i^{n-1}(i+1) of the cycle of length 2n2^{n} generated by eie_{i} in ΓnPk\Gamma_{n}^{P_{k}} are isomorphic to the decorations attached to the vertices (i+1)n1(i+1)^{n-1} and in1i^{n-1}, respectively, both belonging to the cycle of length 2n12^{n-1} generated by eie_{i} in Γn1Pk\Gamma_{n-1}^{P_{k}}.

\psfrag{G}{$\Gamma^{P_{k}}_{n-1}$}\psfrag{GG}{$\Gamma^{P_{k}}_{n}$}\psfrag{e1}{$e_{1}$}\psfrag{e2}{$e_{2}$}\psfrag{a}{$2^{n-2}1$}\psfrag{b}{$2^{n-1}$}\psfrag{c}{$\ell=2^{n-1}$}\psfrag{d}{$\ell=2^{n-1}$}\psfrag{e}{$2^{n-1}1$}\psfrag{f}{$2^{n}$}\psfrag{m}{$2^{n-2}12$}\psfrag{n}{$2^{n-2}11$}\psfrag{h}{$\ell=2^{n}$}\includegraphics[width=318.66946pt]{fig1.eps}
Figure 11. Change of cycles of extremal edges from level n1n-1 to level nn.
\psfrag{G}{$\Gamma^{P_{k}}_{n-1}$}\psfrag{GG}{$\Gamma^{P_{k}}_{n}$}\psfrag{e-}{$e_{i-1}$}\psfrag{e}{$e_{i}$}\psfrag{e+}{$e_{i+1}$}\psfrag{c}{$\ell=2^{n-1}$}\psfrag{d}{$\ell=2^{n}$}\psfrag{a}{$i^{n-1}$}\psfrag{b}{$(i+1)^{n-1}$}\psfrag{n}{$i^{n-1}(i+1)$}\psfrag{m}{$(i+1)^{n-1}i$}\psfrag{u}{$i^{n}$}\psfrag{v}{$(i+1)^{n}$}\includegraphics[width=409.71689pt]{fig2.eps}
Figure 12. Change of cycles of internal edges from level n1n-1 to level nn.

It is possible to prove by induction that the diameter of ΓnPk\Gamma_{n}^{P_{k}} is realized by the distance of the two vertices ak,na_{k,n} and bk,nb_{k,n}, where

ak,n={(1k)(n1)/21for odd n(1k)n/2for even nbk,n={(k1)(n1)/2kfor odd n(k1)n/2for even n.a_{k,n}=\left\{\begin{array}[]{ll}(1k)^{(n-1)/2}1&\hbox{for odd }n\\ (1k)^{n/2}&\hbox{for even }n\end{array}\right.\qquad b_{k,n}=\left\{\begin{array}[]{ll}(k1)^{(n-1)/2}k&\hbox{for odd }n\\ (k1)^{n/2}&\hbox{for even }n.\end{array}\right.

For instance, we have a4,3=141a_{4,3}=141 and b4,3=414b_{4,3}=414 in Fig. 9, with d(a4,3,b4,3)=19d(a_{4,3},b_{4,3})=19.
In fact, for n=1n=1 the statement clearly holds. Suppose now that nn is odd (the case nn even is similar) and that ak,n1=(1k)(n1)/2a_{k,n-1}=(1k)^{(n-1)/2} and bk,n1=(k1)(n1)/2b_{k,n-1}=(k1)^{(n-1)/2} realize the diameter in Γn1Pk\Gamma_{n-1}^{P_{k}}. Notice that a path from ak,n1a_{k,n-1} to bk,n1b_{k,n-1} must visit the vertices 2n12^{n-1} and (k1)n1(k-1)^{n-1}. In particular ak,n1a_{k,n-1} is the vertex in Γn1Pk\Gamma_{n-1}^{P_{k}} whose distance from the vertex 2n12^{n-1} is maximal. Analogously bk,n1b_{k,n-1} is the vertex in Γn1Pk\Gamma_{n-1}^{P_{k}} whose distance from the vertex (k1)n1(k-1)^{n-1} is maximal.

Therefore, passing from Γn1Pk\Gamma_{n-1}^{P_{k}} to ΓnPk\Gamma_{n}^{P_{k}} we have that ak,n=ak,n11a_{k,n}=a_{k,n-1}1 is the vertex with maximal distance from 2n112^{n-1}1 among the vertices belonging to the copy 11 of Γn1Pk\Gamma_{n-1}^{P_{k}} within ΓnPk\Gamma_{n}^{P_{k}}; similarly, bk,n=bk,n1kb_{k,n}=b_{k,n-1}k is the vertex with maximal distance from (k1)n1k(k-1)^{n-1}k among the vertices belonging to the copy kk of Γn1Pk\Gamma_{n-1}^{P_{k}} within ΓnPk\Gamma_{n}^{P_{k}}.

It follows that the diameter of ΓnPk\Gamma_{n}^{P_{k}} is realized by a path from ak,na_{k,n} to bk,nb_{k,n} given by the sequence of paths

(4) (1k)(n1)/21p12n112np(k1)n(k1)n1kp2(k1)(n1)/2k.\displaystyle(1k)^{(n-1)/2}1\to_{p_{1}}2^{n-1}1\to 2^{n}\to_{p_{\ast}}(k-1)^{n}\to(k-1)^{n-1}k\to_{p_{2}}(k1)^{(n-1)/2}k.

Observe that transition pp_{\ast} is the path 2n,3n,,(k1)n2^{n},3^{n},\ldots,(k-1)^{n} whose length is k3k-3. Moreover, inside the transitions p1p_{1} and p2p_{2} there are the transitions

(5) (k1)n11,(k2)n11,,3n11, 2n11(k-1)^{n-1}1,\ (k-2)^{n-1}1,\ldots,3^{n-1}1,\ 2^{n-1}1

and

(6) (k1)n1k,(k2)n1k,,3n1k, 2n1k,(k-1)^{n-1}k,\ (k-2)^{n-1}k,\ \ldots,3^{n-1}k,\ 2^{n-1}k,

respectively, which are the analogue at the level n1n-1 of the transition pp_{\ast} described above, so that each of them has length k3k-3: they come from the previous level and are contained in the subgraphs attached at 2n2^{n} and (k1)n(k-1)^{n}, respectively. More precisely, the projection of the path p1p_{1} on Γn1Pk\Gamma_{n-1}^{P_{k}} consists of a path p1p_{1}^{\prime} from (1k)(n1)/2(1k)^{(n-1)/2} to (k1)n1(k-1)^{n-1} followed by the path p1′′p_{1}^{\prime\prime} given by the projection of (5). Similarly, the projection of the path p2p_{2} on Γn1Pk\Gamma_{n-1}^{P_{k}} consists of the path p2′′p_{2}^{\prime\prime} given by the projection of (6) followed by the path p2p_{2}^{\prime} from 2n12^{n-1} to (k1)(n1)/2(k1)^{(n-1)/2}.

Denote by (p)\ell(p) the length of the path pp and observe that

(p1)+(p)+(p2)=diam(Γn1Pk).\ell(p_{1}^{\prime})+\ell(p_{\ast})+\ell(p_{2}^{\prime})=diam(\Gamma_{n-1}^{P_{k}}).

We are now ready to prove the following theorem.

Theorem 4.2.

Let PkP_{k} be the path graph on kk vertices, with k3k\geq 3. Then, for every n1n\geq 1:

diam(ΓnPk)=2n+1+(k1)(2n1)4n.diam(\Gamma_{n}^{P_{k}})=2^{n+1}+(k-1)(2n-1)-4n.
Proof.

Put δk(n):=diam(ΓnPk)\delta_{k}(n):=diam(\Gamma_{n}^{P_{k}}). As in the preliminary discussion preceding this theorem, we assume that nn is odd, the case of an even nn being similar. By looking at Eq. (4), we get δk(n)=i=15di\delta_{k}(n)=\sum_{i=1}^{5}d_{i}, where

d1:=d(ak,n,2n11)=(p1),d2:=d(2n11,2n),d3:=d(2n,(k1)n)=(p)d_{1}:=d(a_{k,n},2^{n-1}1)=\ell(p_{1}),\quad\quad d_{2}:=d(2^{n-1}1,2^{n}),\quad\quad d_{3}:=d(2^{n},(k-1)^{n})=\ell(p_{\ast})
d4:=d((k1)n,(k1)n1k),d5:=d((k1)n1k,bk,n)=(p2).d_{4}:=d((k-1)^{n},(k-1)^{n-1}k),\qquad\qquad d_{5}:=d((k-1)^{n-1}k,b_{k,n})=\ell(p_{2}).

From what we said above, we have:

d1+d3+d5\displaystyle d_{1}+d_{3}+d_{5} =\displaystyle= ((p1)+(p1′′))+(p)+((p2′′)+(p2))\displaystyle(\ell(p_{1}^{\prime})+\ell(p_{1}^{\prime\prime}))+\ell(p_{\ast})+(\ell(p_{2}^{\prime\prime})+\ell(p_{2}^{\prime}))
=\displaystyle= δk(n1)+2(k3).\displaystyle\delta_{k}(n-1)+2(k-3).

Since the vertices 2n112^{n-1}1 and 2n2^{n} (resp. (k1)n(k-1)^{n} and (k1)n1k(k-1)^{n-1}k) belong to the maximal cycle generated by e1e_{1} (resp. ek1e_{k-1}) and they are in opposite positions within this cycle, we have d2=d4=2n1d_{2}=d_{4}=2^{n-1}. Therefore, we obtain the following recursive description of δk(n)\delta_{k}(n):

{δk(n)=δk(n1)+2(k3)+2nδk(1)=k1.\left\{\begin{array}[]{ll}\delta_{k}(n)=\delta_{k}(n-1)+2(k-3)+2^{n}\\ \delta_{k}(1)=k-1.\end{array}\right.

A direct computation gives:

δk(n)\displaystyle\delta_{k}(n) =\displaystyle= δk(1)+2(n1)(k3)+i=2n2i\displaystyle\delta_{k}(1)+2(n-1)(k-3)+\sum_{i=2}^{n}2^{i}
=\displaystyle= 2n+1+(k1)(2n1)4n.\displaystyle 2^{n+1}+(k-1)(2n-1)-4n.

In the remaining part of the paper, we give a precise description of the automorphism group of the Schreier graph ΓnPk\Gamma_{n}^{P_{k}}.
First of all, notice that a cycle of a given size must be either invariant or moved to another cycle of the same size under the action of an automorphism. Moreover, maximal cycles (of length 2n2^{n}) generated by extremal edges are the only two maximal cycles which are connected to only one cycle of the same size. This implies that they are either invariant or swapped by an automorphism. Let us denote by AnA_{n} the set of automorphisms of ΓnPk\Gamma_{n}^{P_{k}} leaving invariant the two maximal cycles labeled with extremal edges, and by BnB_{n} the set of automorphisms of ΓnPk\Gamma_{n}^{P_{k}} swapping them.

If ϕAn\phi\in A_{n}, it is possible to prove that actually ϕ\phi fixes the vertices in every cycle labeled with internal edges. The claim easily holds for maximal cycles labeled by internal edges (see for instance the cycle of length 88 containing the vertices 232^{3} and 333^{3} in Fig. 9). In fact, they contain a pair of adjacent vertices attached to cycles of the same length and they cannot be swapped; therefore they must be fixed, being invariant cycles with two fixed vertices. As a consequence, ϕ\phi acts nontrivially only on the decorations attached to the maximal cycles generated by the action of the internal generators. Then ϕ\phi can be characterized by the automorphisms that it induces on such decorations. Since such decorations already appear in Γn1Pk\Gamma_{n-1}^{P_{k}} with the same labels and their automorphisms extend to automorphism of Γn1Pk\Gamma_{n-1}^{P_{k}} in An1A_{n-1}, we can conclude by using an inductive argument that all cycles labeled by internal edges in ΓnPk\Gamma_{n}^{P_{k}} must be fixed by any element in AnA_{n}.
On the other hand, cycles of length greater or equal to 44 generated by the action of the extremal generators e1e_{1} and ek1e_{k-1} (for instance, the cycles of length 88 in Fig. 9 containing the vertices 131^{3} and 434^{3}) give rise to blocks that have nontrivial symmetries around the vertices 2tw2^{t}w, 2t11w2^{t-1}1w and (k1)tw(k-1)^{t}w, (k1)t1kw(k-1)^{t-1}kw. Such symmetries are generated by the reflection around the axis connecting such vertices (i.e., that keep such vertices fixed). As an example, again in Fig. 9, one can consider in the leftmost cycle of length 88 the symmetry around the axis connecting the vertices 222222 and 221221. This implies that each cycle of length greater or equal to 44 generated by the action of e1e_{1} and ek1e_{k-1} gives rise to a nontrivial automorphism of order 22. Observe that such automorphisms commute, since each one acts nontrivially on a prescribed block and fixes the others. In particular, if ϕAn\phi\in A_{n} then it is a composition of these reflections.

Now let us denote by ψ\psi the automorphism of ΓnPk\Gamma_{n}^{P_{k}} induced, in the sense of Proposition 4.3, by the nontrivial automorphism of PkP_{k} switching the vertex ii with the vertex ki+1k-i+1. Clearly ψBn\psi\in B_{n} and ψ2=id\psi^{2}=id. Moreover, for any φBn\varphi\in B_{n} we have that the composition of ψ\psi with φ\varphi is in AnA_{n}. In particular, every automorphism of ΓnPk\Gamma_{n}^{P_{k}} is a composition of the aforementioned reflections in AnA_{n} and possibly ψ\psi. Moreover, ψ\psi commutes with these reflections. We are now ready to prove the following result.

Theorem 4.3.

Let PkP_{k} be the path graph on kk vertices, with k3k\geq 3. Then, for every n2n\geq 2, the group of automorphisms of ΓnPk\Gamma_{n}^{P_{k}} is 2ϕk(n)+1\mathbb{Z}_{2}^{\phi_{k}(n)+1}, where

ϕk(n)=2(kn12kn2+1)k1\phi_{k}(n)=\frac{2(k^{n-1}-2k^{n-2}+1)}{k-1}

is the number of cycles of length greater or equal to 44 in ΓnPk\Gamma_{n}^{P_{k}} generated by the action of the generators e1e_{1} and ek1e_{k-1}.

Proof.

It follows from the above discussion that we have to count the number of cycles of length greater or equal to 44 in ΓnPk\Gamma_{n}^{P_{k}} generated by the action of e1e_{1} and ek1e_{k-1}. We proceed by induction in order to prove that such a number coincides with ϕk(n)\phi_{k}(n). For n=2n=2, we just have the two cycles of length 44 with vertex sets {11,22,12,21}\{11,22,12,21\} and {(k1)2,k2,(k1)k,k(k1)}\{(k-1)^{2},k^{2},(k-1)k,k(k-1)\}, with the nontrivial automorphisms flipping 11,1211,12 and k2,k(k1)k^{2},k(k-1). Now, passing from Γn1Pk\Gamma_{n-1}^{P_{k}} to ΓnPk\Gamma_{n}^{P_{k}}, all cycles generated by e1e_{1} and ek1e_{k-1} are preserved just appending a final letter in {1,,k}\{1,\ldots,k\} to the vertices, except the maximal cycle in Γn1Pk\Gamma_{n-1}^{P_{k}} containing 2n12^{n-1}, and the maximal cycle in Γn1Pk\Gamma_{n-1}^{P_{k}} containing (k1)n1(k-1)^{n-1}, since each of them will produce a bigger maximal cycle in ΓnPk\Gamma_{n}^{P_{k}} (the one containing 2n112^{n-1}1 and 2n2^{n}, and the one containing (k1)n1k(k-1)^{n-1}k and (k1)n(k-1)^{n}), so that each of them gives rise to k1k-1 cycles to be taken into account at level nn. Therefore, we obtain the following recursive formula for ϕk(n)\phi_{k}(n):

{ϕk(n)=kϕk(n1)2ϕk(2)=2.\left\{\begin{array}[]{ll}\phi_{k}(n)=k\phi_{k}(n-1)-2\\ \phi_{k}(2)=2.\end{array}\right.

A direct computation gives:

ϕk(n)\displaystyle\phi_{k}(n) =\displaystyle= kn2ϕk(2)2i=0n3ki\displaystyle k^{n-2}\phi_{k}(2)-2\sum_{i=0}^{n-3}k^{i}
=\displaystyle= 2(kn12kn2+1)k1.\displaystyle\frac{2(k^{n-1}-2k^{n-2}+1)}{k-1}.

The claim follows by adding to the ϕk(n)\phi_{k}(n) automorphisms the one induced, in the sense of Proposition 4.3, by the nontrivial automorphism of PkP_{k} switching the vertex ii with the vertex ki+1k-i+1. ∎

Remark 4.4.

Theorem 4.2 and Theorem 4.3 do not work for k=2k=2. In this case, we get the Adding machine (see Case (1) in Example 3.1) whose Schreier graphs are cycles. More precisely, the nn-th Schreier graph ΓnP2\Gamma_{n}^{P_{2}} is the cyclic graph C2nC_{2^{n}}, whose diameter is 2n12^{n-1} and whose automorphism group is isomorphic to the dihedral group D2n+1D_{2^{n+1}}.

References

  • [1] L. Bartholdi, R. Grigorchuk, V. Nekrashevych, From fractal groups to fractal sets, in Fractals in Graz 2001, 25–118, Trends Math., Birkhäuser, Basel, 2003.
  • [2] L. Bartholdi, A. G. Henriques, V. Nekrashevych, Automata, groups, limit spaces, and tilings, J. Algebra 305 (2006), no. 2, 629–663.
  • [3] L. Bartholdi, V. Kaimanovich and V. Nekrashevych, On amenability of automata groups, Duke Math. J. 154 (2010), no. 3, 575–598.
  • [4] B. Bollobás, Modern graph theory, Graduate Texts in Mathematics, 184, Springer-Verlag, New York, 1998, xiv + 394 pp.
  • [5] I. Bondarenko, Groups generated by bounded automata and their Schreier graphs, Ph.D. Dissertation, Texas A&M University, 2007, available at core.ac.uk/reader/4273904
  • [6] I. Bondarenko, T. Ceccherini-Silberstein, A. Donno, V. Nekrashevych, On a family of Schreier graphs of intermediate growth associated with a self-similar group, European J. Combin. 33 (2012), no. 7, 1408–1421.
  • [7] I. Bondarenko, D. D’Angeli, T. Nagnibeda, Ends of Schreier graphs and cut-points of limit spaces of self-similar groups, J. Fractal Geom. 4 (2017), no. 4, 369–424.
  • [8] M. Cavaleri, Computability of Følner sets, Internat. J. Algebra Comput. 27 (2017), no. 7, 819–830.
  • [9] M. Cavaleri, Følner functions and the generic word problem for finitely generated amenable groups, J. Algebra 511 (2018), 388–404.
  • [10] R Charney, An introduction to right-angled Artin groups, Geom. Dedicata 125 (2007), 141–158.
  • [11] D. D’Angeli, Schreier graphs of an extended version of the binary adding machine, Electron. J. Combin. 21 (2014) no. 4, Paper 4.20, 14 pp.
  • [12] D. D’Angeli, A. Donno, M. Matter, T. Nagnibeda, Schreier graphs of the Basilica group, J. Mod. Dyn. 4 (2010), no. 1, 167–205.
  • [13] D. D’Angeli, A. Donno, T. Nagnibeda, Partition functions of the Ising model on some self-similar Schreier graphs, in: Random walks, boundaries and spectra, 277–304, Progr. Probab., 64, Birkhäuser/Springer Basel AG, Basel, 2011.
  • [14] D. D’Angeli, A. Donno, T. Nagnibeda, Counting dimer coverings on self-similar Schreier graphs, European J. Combin. 33 (2012), no. 7, 1484–1513.
  • [15] D. D’Angeli, D. Francoeur, E. Rodaro, J.P. Wächter, Infinite automaton semigroups and groups have infinite orbits, J. Algebra 553 (2020), 119–137.
  • [16] D. D’Angeli, E. Rodaro, Freeness of automaton groups vs boundary dynamics, J. Algebra 462 (2016), 115–136.
  • [17] P. de la Harpe, Topics in geometric group theory, Chicago Lectures in Mathematics, University of Chicago Press, Chicago, IL, 2000, vi + 310 pp.
  • [18] P. Gillibert, The finiteness problem for automaton semigroups is undecidable, Internat. J. Algebra Comput. 24 (2014), no. 1, 1–9.
  • [19] P. Gillibert, An automaton group with undecidable order and Engel problems, J. Algebra 497 (2018), 363–392.
  • [20] R. Grigorchuk, On Burnside’s problem on periodic groups, (Russian) Funktsional. Anal. i Prilozhen. 14 (1980), no. 1, 53–54.
  • [21] R. Grigorchuk, Some problems of the dynamics of group actions on rooted trees, Proc. Steklov Inst. Math. 273 (2011), no. 1, 64–175.
  • [22] R. Grigorchuk, V. Nekrashevich, Z. Šunić, From self-similar groups to self-similar sets and spectra, Fractal geometry and stochastics V, 175–207, Progr. Probab. 70, Birkhäuser/Springer, Cham, 2015.
  • [23] R. Grigorchuk, V. Nekrashevich, V. I. Sushchanskii, Automata, Dynamical Systems, and Groups, Proc. Steklov Inst. Math. 231 (2000), no. 4, 128–203.
  • [24] R. Grigorchuk, Z. Šunić, Asymptotic aspects of Schreier graphs and Hanoi Towers groups, C. R. Math. Acad. Sci. Paris 342 (2006), no. 8, 545–550.
  • [25] R. Grigorchuk, A. Żuk, On a torsion-free weakly branch group defined by a three-state automaton, International Conference on Geometric and Combinatorial Methods in Group Theory and Semigroup Theory (Lincoln, NE, 2000), International J. Algebra Comput. 12 (2002), no. 1–2, 223–246.
  • [26] A. Hinz, S. Klavžar, S. Zemljič, A survey and classification of Sierpiński-type graphs, Discrete Appl. Math. 217 (2017), part 3, 565–600.
  • [27] V. Nekrashevych, Self-similar Groups, Mathematical Surveys and Monographs, 117, American Mathematical Society, Providence, RI, 2005, xii + 231 pp.
  • [28] S. Sidki, Automorphisms of one-rooted trees: growth, circuit structure and acyclicity, J. Math. Sci. (N.Y.) 100 (2000), no. 1, 1925–1943.
  • [29] Z. Šunić, E. Ventura, The conjugacy problem in automaton groups is not solvable, J. Algebra 364 (2012) 148–154.