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

Centralizers in the R. Thompson group VnV_{n}

Collin Bleak    Hannah Bowman    Alison Gordon    Garrett Graham    Jacob Hughes    Francesco Matucci    Eugenia Sapir
Abstract

Let n2n\geq 2 and let αVn\alpha\in V_{n} be an element in the Higman-Thompson group VnV_{n}. We study the structure of the centralizer of αVn\alpha\in V_{n} through a careful analysis of the action of α\langle\alpha\rangle on the Cantor set \mathfrak{C}. We make use of revealing tree pairs as developed by Brin and Salazar from which we derive discrete train tracks and flow graphs to assist us in our analysis. A consequence of our structure theorem is that element centralizers are finitely generated. Along the way we give a short argument using revealing tree pairs which shows that cyclic groups are undistorted in VnV_{n}.

1 Introduction

11footnotetext: Research supported by the National Science Foundation through the Research Experiences for Undergraduates grant at Cornell

In this paper, we produce a description of the structure of element centralizers within the Higman–Thompson groups VnV_{n}. As a corollary to our structure theorem we see that element centralizers in the VnV_{n} are finitely generated. Finally, we give a separate short argument which shows that all infinite-cyclic subgroups within VnV_{n} are embedded without distortion. (The groups VnV_{n} are first introduced by Higman in [14], where Vn=Gn,1V_{n}=G_{n,1} in his notation. The R. Thompson group denoted VV in [9] is V2V_{2} in our notation.)

For a given integer n2n\geq 2, our primary view of the group VnV_{n} is as a group of homeomorphisms acting on the Cantor set n\mathfrak{C}_{n} (seen as the boundary of the rooted regular infinite nn-ary tree 𝒯n\mathscr{T}_{n}). This point of view informs most of our work, where we use a close study of the dynamics of subgroup actions on n\mathfrak{C}_{n} to derive our main result.

Centralizers in Thompson’s group F=F2F=F_{2} are first classified by Guba and Sapir in [13] as a consequence of their classification of element centralizers for diagram groups. In related but separate work, Brin and Squier in [6] describe roots and centralizers in PLo(I)\textrm{PL}_{o}(I), the group of orientation-preserving, piecewise linear homeomorphisms of the unit interval (from which it is also easy to describe the element centralizers of FnF_{n}, although Brin and Squier never formally do so). Guba and Sapir also show in [12] that element centralizers in FF are embedded without distortion in FF. In Chapter 8 of the thesis [17], Bleak, Kassabov, and Matucci classify centralizers in T2T_{2} up to finite index. This paper can be seen as a continuation of the line of research leading to these results.

Higman’s work in [14] also contains information about the structure of centralizers of elements in VnV_{n} (see Theorem 9.9 of [14]). If one reads Higman’s proof of Theorem 9.9 carefully, one can derive with reasonable effort some of the information about the \mathbb{Z} factors contained within the right hand direct product in 1.1. However, our own result contains significantly more detail about the overall structure of element centralizers than is contained within [14].

Separately, Tuna Altinel and Alexey Muranov (see [1]) use model theory to analyze aspects of the groups Fn<Tn<VnF_{n}<T_{n}<V_{n}. In their work they compile some information with regard to element centralizers in these groups. Their results with regards to element centralizers are very similar to what is known from the work of Brin and Squier in [6] and of Kassabov and Matucci in [15], and appears to be contained within the results of Higman from [14].

Martínez-Pérez and Nucinkis [16] recently studied generalizations of the groups of Higman, Thompson, Stein and Brin and classified centralizers of finite subgroups in order to study finiteness properties of those groups via cohomology. Their result generalizes the one in [17] by Bleak, Kassabov, and Matucci and agrees with the one of the current paper when restricted to torsion elements.

The work in this paper uses in broad outline the approach of Bleak, Kassabov and Matucci to centralizers in Chapter 8 of [17], but we work in the more complex groups VnV_{n}, and thus we need to employ a more complex set of tools in our analysis. We chose to use Brin’s revealing pair technology (see [5]) for our supporting calculations (from amongst a fairly long list of tools that provide similar data), and we developed discrete train tracks and our flow graph objects to further support our intuitive understanding of how elements of VnV_{n} act on n\mathfrak{C}_{n}.

One advantage of considering a revealing pair (A,B,σ)(A,B,\sigma) representing an element αVn\alpha\in V_{n} over any random representative tree pair is that it is easier to understand the orbit structure of points in the Cantor set n\mathfrak{C}_{n} under the action of the group α\langle\alpha\rangle. The paper [21] studies revealing pairs in depth, and includes a solution of the conjugacy problem for VnV_{n} using revealing pairs (Salazar gives all arguments in the context of V2V_{2}, but it is easy to see that her methods extend to nn-ary trees).

We briefly describe some alternative technologies. As alluded above, one could use the seminormal and quasinormal forms from [14] to gather much of the information we obtained from revealing pairs. In fact, in response to early drafts of this article and conversations with the first author of this article, Nathan Barker has replicated and extended many of the results herein using Higman’s technology and he has gone on to work on the simultaneous conjugacy problem in VnV_{n} [2]. Another technology is the strand diagrams of Belk and Matucci (see [3]), which themselves are refinements of Pride’s pictures in [19, 20]. In turn, Pride’s pictures are essentially dual objects to the Dehn diagrams from geometric group and semigroup theory (for instance, in this context, one can study the related analysis of conjugacy in FF and other diagram groups by Guba and Sapir in [13]). In the end, these tools all provide access to similar content. We chose revealing pairs as we were comfortable with calculations using them, and because it was particularly easy to define our chief combinatorial objects, discrete train tracks and flow graphs, from a revealing pair.

The dynamical information described by discrete train tracks and their corresponding flow graphs, forms a key ingredient in the proof by Bleak and Salazar [4] of the perhaps surprising result that 2\mathbb{Z}^{2}*\mathbb{Z} does not embed in VV. In particular, those authors make significant use of our flow graph technology in their analysis.

Sections 2 and 3, and subsection 4.1, serve as a mostly expository introduction to calculations in the generalized R. Thompson groups FnF_{n}, TnT_{n}, and VnV_{n}. An informed reader in the area can likely skip ahead to section 4.2, looking back to these sections on the rare occasions in which a new term appears.


Acknowledgments. The authors would like to thank Claas Röver and Martin Kassabov for interesting and helpful conversations relating to this work. We also owe a debt of thanks to Mark Sapir for suggesting we look into the distortion of cyclic subgroups in VnV_{n}. We thank Nathan Barker for interesting conversations with respect to our flow graph object; it was by these conversations that the authors were finally convinced of the need to formally define flow-graphs, instead of just using them as informal guides to intuition. Finally, the sixth author would like to thank Conchita Martínez-Pérez for pleasant conversations about her work with Nucinkis [16] where they discussed our independent approaches to the question, and the mutual confirmation of each other’s results.

The authors also thank Cornell University for hosting, Robert Strichartz for organizing, and the NSF for funding the REU program at which most of this work took place. We also thank the Centre de Recerca Matemàtica (CRM) in Barcelona, the University of Virginia, and the University of Nebraska – Lincoln for supporting some of the authors and providing great conditions to finish this work.

1.1 A formal statement of results

If αVn\alpha\in V_{n}, we set

CVn(α):={βVnαβ=βα}.C_{V_{n}}(\alpha)\mathrel{\mathop{:}}=\left\{\beta\in V_{n}\mid\alpha\beta=\beta\alpha\right\}.

We call the group CVn(α)C_{V_{n}}(\alpha) the centralizer in VnV_{n} of α\alpha, as is standard.

Our primary theorem is the following.

Theorem 1.1

Let n>1n>1 be a positive integer and suppose αVn\alpha\in V_{n}. Then, there are non-negative integers ss, tt, mim_{i}, and rir_{i} and groups KmiK_{m_{i}}, Gn,riG_{n,r_{i}}, AjA_{j}, and PtP_{t}, for i{1,,s}i\in\{1,\ldots,s\} and j{1,,t}j\in\{1,\ldots,t\}, so that

CVn(α)(i=1sKmiGn,ri)×(j=1t((Aj)Pqj))C_{V_{n}}(\alpha)\cong\left(\prod_{i=1}^{s}K_{m_{i}}\rtimes G_{n,r_{i}}\right)\times\left(\prod_{j=1}^{t}\left(\left(A_{j}\rtimes\mathbb{Z}\right)\wr P_{q_{j}}\right)\right)

We now explain the statement of this theorem in a bit more detail.

The group α\langle\alpha\rangle acts on a subset of the nodes of the infinite nn-ary tree. The number ss represents the number of distinct lengths of finite cyclic orbits of nodes under this action. The value of ss is easy to compute from any given revealing pair representing α\alpha.

For the action mentioned above, each rir_{i} is determined as a minimal number of nodes carrying a fundamental domain (for the action of a conjugate version of α\langle\alpha\rangle) in the set of nodes supporting the cycles of length mim_{i}.

For each mim_{i}, we have Kmi=(Maps(n,mi))riK_{m_{i}}=(Maps(\mathfrak{C}_{n},\mathbb{Z}_{m_{i}}))^{r_{i}}, where Maps(n,mi)Maps(\mathfrak{C}_{n},\mathbb{Z}_{m_{i}}) is the group of continuous maps from n\mathfrak{C}_{n} to mi\mathbb{Z}_{m_{i}} under point-wise multiplication, and where mi\mathbb{Z}_{m_{i}} is the cyclic group /(mi)\mathbb{Z}/(m_{i}\mathbb{Z}) under the discrete topology. We note that KmiK_{m_{i}} is not finitely generated for mi>1m_{i}>1.

The groups Gn,riG_{n,r_{i}} are the Higman-Thompson groups from [14].

Given any element βVn\beta\in V_{n}, one can associate an infinite collection of finite labeled graphs (which we call flow graphs). Flow graphs are labeled, directed, finite graphs and which describe structural meta-data pertaining to the dynamics of certain subsets of n\mathfrak{C}_{n} under the action of α\langle\alpha\rangle. Flow graphs are themselves “quotient objects” coming from discrete train tracks, which are objects we introduce here to model dynamics in the Cantor set much as regular train tracks model dynamics in surface homeomorphism theory [22, 23, 18]

Components of a flow graph associated with α\alpha fall into equivalence classes {ICCi}\{ICC_{i}\} which model similar dynamics. The number tt is the number of equivalence classes of components carrying infinite orbits under the action of α\langle\alpha\rangle. This number happens to be independent of the representative flow graph chosen.

The right factor of the main direct product represents the restriction of the centralizer of α\alpha to elements which are the identity away from the closure of the region where α\langle\alpha\rangle acts with non-finite orbits.

For each jj, the supports of the elements of ICCjICC_{j} represent regions in the action of α\langle\alpha\rangle where the centralizers of α\alpha, with restricted actions to these regions, are isomorphic. For any such support of an element of ICCjICC_{j}, the finite order elements of the restricted centralizer form a group. We take AjA_{j} to be a representative group from set of these isomorphic finite groups.

The group PqjP_{q_{j}} is the full symmetric group on the qjq_{j} isomorphic flow graph components in ICCjICC_{j}.

Recall now that by Higman in [14], the groups Gk,rG_{k,r} are all finitely presented. This fact, together with a short analysis of the nature of the actions in the left-hand semi-direct products, shows that each of the groups KmiGn,riK_{m_{i}}\rtimes G_{n,r_{i}} are finitely generated (see Corollary 6.2 below). As the groups on the right-hand side of the central direct product are manifestly finitely generated, we obtain the following corollary to Theorem 1.1.

Corollary 1.2

Let n>1n>1 be an integer and αVn\alpha\in V_{n}. The group CVn(α)C_{V_{n}}(\alpha) is finitely generated.

One could try to improve this last corollary to obtain a statement of finite-presentation, which may be feasible. Such a proof might be accomplished through a careful study of presentations of the Higman groups Gn,rG_{n,r}. In this direction, we ask the following.

Question 1

Must the group CVn(α)C_{V_{n}}(\alpha) be finitely presented for every αVn\alpha\in V_{n}?

We have also found some evidence supporting the possibility that the answer to the following question is “Yes.”

Question 2

Is it true that for each index jj, the group AjA_{j} is abelian?

It is not completely trivial to find an example element αV=V2\alpha\in V=V_{2} where any AiA_{i} is not cyclic. In section 7 we give such an example where t=1t=1 and A12×2A_{1}\cong\mathbb{Z}_{2}\times\mathbb{Z}_{2}.

A generator of each \mathbb{Z} in the right hand terms AiA_{i}\rtimes\mathbb{Z} is given by a root of a restricted version of α\alpha which is restricted to act only on the support of an element in ICCjICC_{j}. We know that AiA_{i} commutes with the restricted version of α\alpha by definition, but it is not clear that AiA_{i} will commute with any valid choice of a generator for the \mathbb{Z} term.

Question 3

Is it possible to replace the right hand terms AiA_{i}\rtimes\mathbb{Z} with Ai×A_{i}\times\mathbb{Z}?

In the final section of the paper we prove the following theorem.

Theorem 1.3

Suppose αVn\alpha\in V_{n} so that α\langle\alpha\rangle\cong\mathbb{Z}. The group α\langle\alpha\rangle is undistorted as a subgroup of VnV_{n}.

1.2 Our general approach

We give a description of our approach to centralizers in the broadest terms.

Let αVn\alpha\in V_{n}. Let H=αH=\langle\alpha\rangle. Define the fundamental domain of the action of HH to be the space n/H\mathfrak{C}_{n}/H. Note that this space is generally not very friendly, i.e., it is typically not Hausdorff, but that fact will have almost no bearing on our work.

If an element commutes with α\alpha, it will induce an action on n/H\mathfrak{C}_{n}/H.

Thus, we get a short exact sequence:

1𝒦CVn(α)𝒬11\to\mathcal{K}\to C_{V_{n}}(\alpha)\to\mathcal{Q}\to 1

Here, 𝒦\mathcal{K} represent the elements in CVn(α)C_{V_{n}}(\alpha) which act on n\mathfrak{C}_{n} in such a way that their induced action on n/H\mathfrak{C}_{n}/H is trivial. The group 𝒬\mathcal{Q} is the natural quotient of CVn(α)C_{V_{n}}(\alpha) by the image of the inclusion map 𝒦CVn(α)\mathcal{K}\to C_{V_{n}}(\alpha). Loosely speaking, elements of 𝒬\mathcal{Q} are represented by elements of VnV_{n} which act in the “same” way on each “copy” of the fundamental domain in n\mathfrak{C}_{n} (this of course is imprecise; there may be no embedding of a fundamental domain in n\mathfrak{C}_{n}).

The first author is indebted to Martin Kassabov for pointing out this general structural approach to analyzing centralizers in groups of homeomorphisms.

2 Basic definitions

Throughout this section, let us fix an integer n>1n>1 for our discussion. We will also establish other conventions later that will hold throughout the section, and not just in a particular subsection.

We assume the reader is familiar with [9], and with the definition of a tree pair representative of an element of V=V2V=V_{2}. Nonetheless, we give an abbreviated tour through those definitions for the non-experts (extending them to include the groups FnTnVnF_{n}\leq T_{n}\leq V_{n}) and we state some essential lemmas which either occur in that source or which are easily derived by the reader with full understanding of these definitions. We give some examples demonstrating most of that language, and we add some new language to the lexicon mostly in support of our own later definition of a flow graph. In general the reader experienced with R. Thompson group literature will find little new material in this section and is encouraged to skip ahead, returning only if he or she runs into an unfamiliar term in the later parts of the paper.

2.1 Trees and Cantor sets

The only material in this section that may be unfamiliar to readers conversant with R. Thompson group literature is some of the language describing the Cantor set underlying a node of the tree 𝒯n\mathscr{T}_{n} and related concepts.

Our primary perspective will be to consider VnV_{n} as a group of homeomorphisms of the Cantor set \mathfrak{C}. In particular, VnV_{n} should be thought to act as a group of homeomorphisms of the Cantor set n\mathfrak{C}_{n}\cong\mathfrak{C}. That is, the version n\mathfrak{C}_{n} of the Cantor set that is realized as the boundary of the standard infinite, rooted nn-ary tree 𝒯n\mathscr{T}_{n}. While we assume the reader understands that realization of the Cantor set, and also the terms “node” or “vertex”, “child”, “parent”, “ancestor”, “descendant”, “leaf”, and “nn-caret” (or simply “caret”), and similar related language when referring to aspects of rooted nn-ary trees, we give a short discussion below to establish some of our standard usage.

The left-to-right ordering of the children of a vertex in 𝒯n\mathscr{T}_{n} allows us to give a name to each vertex in 𝒯n\mathscr{T}_{n}. Given a vertex pp in 𝒯n\mathscr{T}_{n}, there is a unique order preserving bijection Ordp:Children(p){0,1,2,,n1}\textrm{Ord}_{p}:Children(p)\to\left\{0,1,2,\ldots,n-1\right\}. Let vv be a vertex of 𝒯n\mathscr{T}_{n} and let (vi)i=1m(v_{i})_{i=1}^{m} be the unique descending path in 𝒯n\mathscr{T}_{n} starting at the root r=v1r=v_{1} and ending at v=vmv=v_{m}, then we name the vertex vv with the sequence (Ordvi(vi+1))i=1m1(\textrm{Ord}_{v_{i}}(v_{i+1}))_{i=1}^{m-1}. We will think of this sequence as a string (ordered from left-to-right).

Given names of two vertices, we may concatenate these strings to produce the name of a third vertex which will be a descendant of the first vertex

Below, we diagram an example of 𝒯2\mathscr{T}_{2}, with a finite tree TT highlighted within it. The vertex cc is a leaf of TT, and in both TT and 𝒯2\mathscr{T}_{2}, cc is a child of bb which is a child of aa. The vertex aa is an ancestor of cc and cc is a descendant of aa. The name of the vertex labeled by cc is 010010.

\psfrag{1}[c]{$1$}\psfrag{0}[c]{$0$}\psfrag{a}[c]{$a$}\psfrag{b}[c]{$b$}\psfrag{c}[c]{$c$}\psfrag{T}[c]{$T$}\includegraphics[height=200.0pt,width=350.0pt]{UTreeSubtree.eps}

We can view a finite rooted nn-ary tree TT as an instruction on how to partition the Cantor set n\mathfrak{C}_{n}. Consider the natural embedding of TT into the tree 𝒯n\mathscr{T}_{n}, where we send the root of TT to the root of 𝒯n\mathscr{T}_{n}, and we preserve orders of children. For instance, as in the diagram above. Now, consider the set PnP_{n} of all infinite descending paths in 𝒯n\mathscr{T}_{n} which start at the root of 𝒯n\mathscr{T}_{n}. If we consider each ordered nn-caret in 𝒯n\mathscr{T}_{n} as an instruction to pass through another inductive subdivision of the unit interval in the formation process of n\mathfrak{C}_{n}, then each element in PnP_{n} can be thought of as limiting on an element of n\mathfrak{C}_{n}. We thus identify PnP_{n} with n\mathfrak{C}_{n}. The set PnP_{n} will now be considered to be topologized using the induced topology from the metric space topology of the unit interval. Now if we consider a vertex cc of TT, we can associate cc with the subset of n\mathfrak{C}_{n} corresponding to the paths in PnP_{n} which pass through cc. We will call this the Cantor set under cc, and we will call any such subset of n\mathfrak{C}_{n} an interval of n\mathfrak{C}_{n}. It is immediate that any interval in n\mathfrak{C}_{n} is actually homeomorphic with n\mathfrak{C}_{n}. Given a node cc, the Cantor set under cc is also commonly called a cone neighborhood in n\mathfrak{C}_{n}, and by definitions these sets form the standard basis for the product topology on Pn={0,1,,n1}ωnP_{n}=\{0,1,\ldots,n-1\}^{\omega}\cong\mathfrak{C}_{n}. Thus, the leaves of TT partition the set PnP_{n}, and they also partition n\mathfrak{C}_{n}, into a set of open basis sets. We will call this the partition of n\mathfrak{C}_{n} associated with the tree TT.

Extending the language of the previous paragraph, given a set XnX\subset\mathfrak{C}_{n}, we will call any node cc of the universal tree 𝒯n\mathscr{T}_{n} which has its underlying set contained in XX, a node of XX.

Of course, for all integers m,n>1m,n>1, we have that mn2=\mathfrak{C}_{m}\cong\mathfrak{C}_{n}\cong\mathfrak{C}_{2}=\mathfrak{C}.

Using the example above, the interval of 2\mathfrak{C}_{2} under cc is 2[2/9,7/27]\mathfrak{C}_{2}\cap[2/9,7/27]. In discussion, we will generally not distinguish between a vertex of 𝒯n\mathscr{T}_{n} and the interval under it.

Remark 2.1

Any finite union of disjoint intervals in n\mathfrak{C}_{n} is homeomorphic with n\mathfrak{C}_{n}.

2.2 Elements of VnV_{n}, TnT_{n}, and FnF_{n}

Some of the language in this subsection is unusual, although the general content will be familiar to all readers with knowledge of the R. Thompson groups.

An element of Homeo(n)\textrm{Homeo}(\mathfrak{C}_{n}) is allowable if it can be represented by an allowable triple (A,B,σ)(A,B,\sigma). The triple (A,B,σ)(A,B,\sigma) is allowable if there is a positive integer mm so that AA and BB are rooted, finite, nn-ary trees with the same number mm of leaves, and σΣm\sigma\in\Sigma_{m}, the permutation group on the set {1,2,3,,m}\left\{1,2,3,\ldots,m\right\}. We explain below how to build the homeomorphism α\alpha which is represented by such an allowable triple (A,B,σ)(A,B,\sigma). We then call (A,B,σ)(A,B,\sigma) a representative tree pair for α\alpha. The group VnV_{n} consists of the set of all allowable homeomorphisms of n\mathfrak{C}_{n} under the operation of composition.

We are now ready to explain how an allowable triple (A,B,σ)(A,B,\sigma) defines an allowable homeomorphism α\alpha of n\mathfrak{C}_{n}. Suppose AA and BB both have mm leaves, for some integer m>1m>1. We consider AA to represent the domain of α\alpha, and BB to represent the range of α\alpha. We take the leaves of AA and BB and number them in their natural left-to-right ordering from 11 to mm. For each index ii with 1im1\leq i\leq m, we map the interval under leaf ii of AA to the interval under the leaf iσi\sigma of BB using an orientation-preserving affine homeomorphism (that is, the homeomorphism of the two Cantor sets underlying these leaves defined by a restriction and co-restriction of an affine map with positive slope from the real numbers \mathbb{R} to \mathbb{R}). Note for n\mathfrak{C}_{n} the slopes of such maps will be integral powers of 2n12n-1. We will say that the leaf ii of AA is mapped to the leaf iσi\sigma of BB by α\alpha.

The next remark follows immediately from the discussion above:

Remark 2.2

Suppose vVnv\in V_{n}, and vv is represented by a tree pair (A,B,σ)(A,B,\sigma). If η\eta is a node in 𝒯n\mathscr{T}_{n} so that η\eta is either a leaf of AA or a descendant node in 𝒯n\mathscr{T}_{n} of a leaf of AA, then vv will carry the Cantor set underlying η\eta affinely and bijectively in an order-preserving fashion to the Cantor set underlying some node τ\tau of 𝒯n\mathscr{T}_{n}, where τ\tau is either a leaf of BB or a descendant of a leaf of BB.

The diagram below illustrates an example for V2=VV_{2}=V. The tree AA is on the left, and the tree BB is on the right. Note that we have re-decorated the leaves of BB with the numbering from σ\sigma. The diagram beneath the tree pair is intended to indicate where intervals of 2\mathfrak{C}_{2} are getting mapped, where the intervals DD represent the domain and the intervals RR represent the range. The lower diagram is superfluous when defining a general element of VV.

\psfrag{1}[c]{$1$}\psfrag{2}[c]{$\,\,2$}\psfrag{3}[c]{$3$}\psfrag{4}[c]{$4$}\psfrag{5}[c]{$5$}\psfrag{6}[c]{$6$}\psfrag{D}[c]{$D$}\psfrag{R}[c]{$R$}\includegraphics[height=200.0pt,width=300.0pt]{treePair.exp5.eps}

In general, as we can decorate our tree leaves with numbers to indicate the bijection, we will now re-define the phrase “tree pair” throughout the remainder of the paper to mean an allowable triple. We will still discuss the permutation of a tree pair as needed.

As mentioned in the introduction, there are groups FnTnVnF_{n}\leq T_{n}\leq V_{n}. If we only allow cyclic permutations, we get the group TnT_{n}. If our permutation is trivial, we get FnF_{n}. Thus, we can think of FnF_{n} as a group of piecewise-linear homeomorphisms of the interval [0,1][0,1], while TnT_{n} can be thought of as a group of piecewise-linear homeomorphisms of the circle S1S^{1}.

Here is an example element θT=T2\theta\in T=T_{2}, which we will be considering again later.

\psfrag{1}[c]{$1$}\psfrag{2}[c]{$2$}\psfrag{3}[c]{$3$}\psfrag{4}[c]{$4$}\psfrag{5}[c]{$5$}\psfrag{6}[c]{$6$}\psfrag{7}[c]{$7$}\psfrag{D}[c]{$D$}\psfrag{R}[c]{$R$}\includegraphics[height=200.0pt,width=300.0pt]{elmt.T.hid.eps}

Below is an example element from F=F2F=F_{2}.

\psfrag{1}[c]{$1$}\psfrag{2}[c]{$2$}\psfrag{3}[c]{$3$}\psfrag{4}[c]{$4$}\psfrag{5}[c]{$5$}\psfrag{6}[c]{$6$}\psfrag{7}[c]{$7$}\psfrag{D}[c]{$D$}\psfrag{R}[c]{$R$}\includegraphics[height=200.0pt,width=300.0pt]{elmt.F.eps}

2.3 Multiplication in VnV_{n}

There is nothing new in this subsection; readers familiar with the R. Thompson groups should skip ahead.

Naturally, the group operation of VnV_{n} is given by composition of functions. Building compositions using tree pairs is not hard. The process is enabled by the fact that there are many representative tree pairs for an element of VnV_{n}.

We give an example of what we mean by the last sentence. Given a representative tree pair (A,B,σ)(A,B,\sigma), one can use a leaf ii of AA to be a root of an extra nn-caret, creating AA^{\prime}, and one can build a tree BB^{\prime} from BB by replacing the leaf iσi\sigma with an nn-caret. Label the leaves of AA^{\prime} in increasing order, and the leaves of BB^{\prime} using the induced labeling from the permutation σ\sigma on all leaves of BB^{\prime} that are also leaves of BB. For the other leaves of BB^{\prime}, use the labeling, in order, of the nn-caret in AA^{\prime} that is not a nn-caret of AA. Let us call the permutation we have built from the leaves of AA^{\prime} to the leaves of BB^{\prime} by σ\sigma^{\prime}. The process of replacing the tree pair (A,B,σ)(A,B,\sigma) by (A,B,σ)(A^{\prime},B^{\prime},\sigma^{\prime}) is called a simple augmentation.

If the reverse process can be carried out (that is, deleting an nn-caret from both the domain and range trees and re-labeling the permutation so that our initial two trees appear as a simple augmentation of our resulting tree pair) then we call this process a simple reduction. If we carry out either a simple augmentation or a simple reduction, we may instead say we have done a simple modification to our initial tree pair.

The following lemma is a straightforward consequence of the standard fact that any element in VnV_{n} has a unique tree pair representation that will not admit any simple reductions.

Lemma 2.3

Any two representative tree pairs of a particular element of VnV_{n} are connected by a finite sequence of simple modifications.

We are now ready to carry out multiplication of tree pairs. The essence of the idea is to augment the range tree of the first element and the domain tree of the second element until they are the same tree (and carry out the necessary augmentations throughout both tree pairs), and then re-label the permutations, using the labeling of the range tree of the first element to seed the re-labeling of the permutation in the second element’s tree pair. At this junction, the two inner trees are completely identical, and can be removed. The diagram below demonstrates this process.

\psfrag{1}[c]{$1$}\psfrag{2}[c]{$2$}\psfrag{3}[c]{$3$}\psfrag{4}[c]{$4$}\psfrag{5}[c]{$5$}\psfrag{6}[c]{$6$}\includegraphics[height=350.0pt,width=250.0pt]{multiplication.eps}

Thus, we now know what VnV_{n} is, and we can represent and multiply its elements.

3 Conjugation, roots, and centralizers

Recall that we will use right action notation to describe how VnV_{n} acts on =n\mathfrak{C}=\mathfrak{C}_{n}, as below.

The following discussion is completely basic, and just carries out some straightforward points from permutation group theory.

Let cnc\in\mathfrak{C}_{n} and α\alpha, βVn\beta\in V_{n}. In particular, α:nn\alpha:\mathfrak{C}_{n}\to\mathfrak{C}_{n}.

  • We denote the image of cc under α\alpha by cαc\alpha.

  • For conjugation, we denote the conjugate of α\alpha by β\beta by αβ\alpha^{\beta}, and note therefore that cαβ=cβ1αβc\alpha^{\beta}=c\beta^{-1}\alpha\beta.

We will often need to discuss what is moving under an action, so we need a definition as well. We define the support of an element αVn\alpha\in V_{n}, denoted by Suppα\operatorname{Supp}{\alpha}, as below.

Supp(α)={cncαc}.\operatorname{Supp}(\alpha)=\left\{c\in\mathfrak{C}_{n}\mid c\alpha\neq c\right\}.

Note that this is distinct from the standard analysis version of support, which would extend the definition to include the closure of the set of points which are moving.

The following lemma is now standard from the theory of permutation groups.

Lemma 3.1
Supp(αβ)=Supp(α)β\operatorname{Supp}(\alpha^{\beta})=\operatorname{Supp}(\alpha)\beta

Recall that given αVn\alpha\in V_{n} we have

CVn(α)={βVnαβ=βα}.C_{V_{n}}(\alpha)=\left\{\beta\in V_{n}\mid\alpha\beta=\beta\alpha\right\}.

We point out the following obvious facts.

Remark 3.2

Suppose α\alpha, β\beta, and γVn\gamma\in V_{n}.

  1. 1.

    If αβ=βα\alpha\beta=\beta\alpha, then αβ=β1αβ=α\alpha^{\beta}=\beta^{-1}\alpha\beta=\alpha.

  2. 2.

    If βk=α\beta^{k}=\alpha for some integer kk, then βCVn(α)\beta\in C_{V_{n}}(\alpha).

  3. 3.

    We have CVn(α)(CVn(α))γ=CVn(αγ)C_{V_{n}}(\alpha)\cong(C_{V_{n}}(\alpha))^{\gamma}=C_{V_{n}}(\alpha^{\gamma}).

We will use the third point above repeatedly to replace an element whose centralizer we are studying, by a conjugate element which admits a simpler tree-pair representative (simplifying our analysis without affecting centralizer structure).

4 Revealing pairs and related objects

Subsection 4.1 should be considered as mostly expository; it will contain definitions and lemmas from within [5] and [21]. We give very detailed examples of all of the concepts therein which are of use in our context. The following subsections on discrete train tracks, laminations and flow graphs, on the other hand, are entirely new.

Many tree pairs exist to represent a single element. Some tree pairs are more useful than others when it comes to discerning aspects of the dynamics of the element’s action on the Cantor set. Consider the element θT\theta\in T which was defined earlier as the second tree pair diagram in Subsection 2.2.

Thought of as a homeomorphism of the circle, θ\theta has a rotation number, which, roughly stated, measures the average rotation of the circle under the action of θ\theta. (Rotation numbers are a beautiful idea of Poincaré, and they are extremely useful in the analysis of circle maps). It is known, initially by work of Ghys and Sergiescu (see [11]), that the rotation number of any element of TT is rational. Thus, θ\theta has a rational rotation number p/qp/q (in lowest terms). A lemma of Poincaré now shows that some point on the circle will have a periodic orbit, with period qq. If the reader examines the tree pair representative for θ\theta, doubtless he or she will discover a point that travels on a finite orbit of length five. In fact, all points of the circle travel on their own periodic orbit of length five, as tt is torsion with order five. Under any reasonable definition of rotation number, θ\theta must have rotation number 2/52/5, since the reader can observe that after five iterations of θ\theta, the circle will have rotated a total of two whole times around.

The previous example is intended to point to the fact that some tree pairs somehow hide the dynamics of an element’s action.

In this section, we will describe revealing pairs, which are tree pairs that easily yield up all information about the orbit dynamics in n\mathfrak{C}_{n} under the action of a cyclic subgroup of VnV_{n} (or in the circle or the interval, in the cases of cyclic subgroups of TnT_{n} or FnF_{n}, respectively).

4.1 Revealing pair definitions

Throughout this section, we will work with some nontrivial αVn\alpha\in V_{n}. We will assume that the tree pair (A,B,σ)(A,B,\sigma) represents α\alpha.

Consider AA and BB as finite rooted nn-ary sub-trees (with roots at the root of 𝒯n\mathscr{T}_{n}) of 𝒯n\mathscr{T}_{n}.

We will call the set of vertices of 𝒯n\mathscr{T}_{n} which are leaves of both AA and BB the neutral leaves of (A,B,σ)(A,B,\sigma). We will simply call these the neutral leaves, if the tree pair is understood.

As both AA and BB have a root and neither are empty, we can immediately form the tree C=ABC=A\cap B. It is immediate that the neutral leaves are leaves of CC, but if ABA\neq B, then CC will have other leaves as well.

We can make the sets X=ABX=A-B and Y=BAY=B-A. The closures X¯\overline{X} and Y¯\overline{Y} in 𝒯n\mathscr{T}_{n} are both finite disjoint unions rooted nn-ary trees (their roots are not sitting at the root of 𝒯n\mathscr{T}_{n}), where the number of carets in X¯\overline{X} is the same as the number of carets of Y¯\overline{Y} (this number could be zero if A=BA=B). We call X¯\overline{X} and Y¯\overline{Y} a difference of carets for AA and BB.

In the remainder, when we write DED-E, where DD and EE are trees, we actually want to take the difference of carets, so that our result will be a collection of rooted trees.

While α\langle\alpha\rangle acts on the Cantor set, it also induces a “partial action” on an infinite subset of the vertices of 𝒯n\mathscr{T}_{n}, as we explain in this paragraph. Since the interval under a leaf λ\lambda of AA is taken affinely to the interval under a leaf of BB (we will denote this leaf by λα\lambda\alpha), we see that α\alpha induces a map from the vertices of 𝒯n\mathscr{T}_{n} under λ\lambda to the vertices of 𝒯n\mathscr{T}_{n} under λα\lambda\alpha. In particular, the full sub-tree in 𝒯n\mathscr{T}_{n} with root λ\lambda is taken to the full sub-tree in 𝒯n\mathscr{T}_{n} with root λα\lambda\alpha in order preserving fashion. We note in passing that we cannot extend this to a true action on the vertices of 𝒯n\mathscr{T}_{n}; if we consider a vertex η\eta in 𝒯n\mathscr{T}_{n} which is above a leaf of AA, the map α\alpha may take the interval underlying η\eta and map it in a non-affine fashion across multiple intervals in n\mathfrak{C}_{n}.

As an example of the behavior mentioned above, consider the element θ\theta again. The parent vertex of the domain leaves labeled 3 and 4 is mapped across multiple vertices of the range tree, in a non-affine fashion.

If η\eta is a vertex of 𝒯n\mathscr{T}_{n} we can now define a forward and backward orbit OηO_{\eta} of η\eta in 𝒯n\mathscr{T}_{n} under the action of α\langle\alpha\rangle, to the extent that we restrict ourselves to powers of α\alpha that take the interval under η\eta affinely bijectively to an interval under a vertex of 𝒯n\mathscr{T}_{n}. In more general circumstances, given any integer kk and a vertex η\eta of 𝒯n\mathscr{T}_{n}, we will use the notation ηαk\eta\alpha^{k} to denote the subset of n\mathfrak{C}_{n} which is the image of the interval under η\eta under the map αk\alpha^{k}, and if that set happens to be the interval under a vertex τ\tau in 𝒯n\mathscr{T}_{n} (so that the restriction of αk\alpha^{k} to the interval under η\eta takes the interval underlying η\eta affinely to the interval underlying τ\tau), then we may denote the vertex τ\tau by ηαk\eta\alpha^{k} as well.

Let L(A,B,σ)L_{(A,B,\sigma)} denote the set of vertices of 𝒯n\mathscr{T}_{n} which are either leaves of AA or leaves of BB, and let λL(A,B,σ)\lambda\in L_{(A,B,\sigma)}.

It is possible that λ\lambda is a neutral leaf whose vertex in 𝒯n\mathscr{T}_{n} has orbit OλO_{\lambda} entirely contained in the neutral leaves of AA and BB. By Remark 2.2, and the fact that the neutral leaves are finite in number, we see that in this case the orbit of λ\lambda is periodic in 𝒯n\mathscr{T}_{n}. In this case, we call λ\lambda a periodic leaf.

Now suppose λ\lambda is not a periodic leaf of AA. This implies that if we consider the forward and backward orbits of λ\lambda under the action of α\langle\alpha\rangle, then in both directions, the orbit will exit the set of neutral leaves. In particular, there is a minimal integer r0r\leq 0 and a maximal integer s0s\geq 0 so that for all integers ii with risr\leq i\leq s we have that λαiL(A,B,σ)\lambda\alpha^{i}\in L_{(A,B,\sigma)}. It is also immediate that λαr\lambda\alpha^{r} is a leaf of ABA-B and λαs\lambda\alpha^{s} is a leaf of BAB-A, while for all values of ii with r<i<sr<i<s we see that λαi\lambda\alpha^{i} is a neutral leaf.

Thus, we have the following lemma.

Lemma 4.1

Suppose αV\alpha\in V is non-trivial and α\alpha is represented by a tree pair (A,B,σ)(A,B,\sigma). If λL(A,B,σ)\lambda\in L_{(A,B,\sigma)}, then either λ\lambda is

  1. 1.

    a periodic neutral leaf, in which case there is a maximal integer s0s\geq 0 so that the iterated augmentation chain defined by IAC(λ):=(λαi)i=0sIAC(\lambda):=(\lambda\alpha^{i})_{i=0}^{s} is a sequence of neutral leaves so that s+1s+1 is the smallest positive power so that λαs+1=λ\lambda\alpha^{s+1}=\lambda,

  2. 2.

    a leaf of ABA-B, in which case there is a maximal integer s>0s>0 so that IAC(λ):=(λαi)i=0sIAC(\lambda):=(\lambda\alpha^{i})_{i=0}^{s} is a sequence of leaves of AA or BB, and furthermore, we then have λαs\lambda\alpha^{s} is a non-neutral leaf of BB while λαi\lambda\alpha^{i} is a neutral leaf in L(A,B,σ)L_{(A,B,\sigma)} for all indices ii with 0<i<s0<i<s,

  3. 3.

    a leaf of BAB-A, in which case there is a minimal integer r<0r<0 so that IAC(λ):=(λαi)i=r0IAC(\lambda):=(\lambda\alpha^{i})_{i=r}^{0} is a sequence of leaves of AA or BB, and furthermore, we then have λαr\lambda\alpha^{r} is a non-neutral leaf of AA while λαi\lambda\alpha^{i} is a neutral leaf in L(A,B,σ)L_{(A,B,\sigma)} for all indices ii with r<i<0r<i<0, or

  4. 4.

    a neutral, non-periodic leaf, in which case, λ\lambda is a neutral leaf in a sequence IAC(η)IAC(\eta) for some vertex η\eta which is a leaf of ABA-B, as discussed in point (2)(2). In this case we set IAC(λ):=IAC(η)IAC(\lambda):=IAC(\eta).

We now define and comment on some language from [5] and from [21]. Suppose we have the hypotheses of Lemma 4.1. The definition of iterated augmentation chain IAC(λ)IAC(\lambda) in Lemma 4.1 reflects the fact that we can augment the trees AA and BB at each vertex along an iterated augmentation chain, and end up with a new representative tree pair for α\alpha (the first vertex in an augmentation chain of type (2)(2) or (3)(3) can only be augmented in the domain tree AA, while the last such vertex can only be augmented in the range tree BB).

If λ\lambda is of type (2)(2), with IAC(λ)=(λαi)i=0sIAC(\lambda)=(\lambda\alpha^{i})_{i=0}^{s} for some positive integer ss, where λαs\lambda\alpha^{s} is an ancestor of λ\lambda, then we say that λ\lambda is a repeller. If λ\lambda is of type (3), with IAC(λ)=(λαi)i=r0IAC(\lambda)=(\lambda\alpha^{i})_{i=r}^{0} for some negative integer rr, and where λαr\lambda\alpha^{r} is an ancestor of λ\lambda, then we say that λ\lambda is an attractor.

We are now ready to define what it means for a tree pair to be a revealing pair. Suppose αV\alpha\in V and the tree pair (A,B,σ)(A,B,\sigma) represents α\alpha. If every component of ABA-B contains a repeller, and every component of BAB-A contains an attractor, then we say that (A,B,σ)(A,B,\sigma) is a revealing pair representing α\alpha.

The discussion beginning section 10.7 in [5] proves that every element of VnV_{n} admits a revealing pair. It is not hard to generate an algorithm which will transform any representative tree pair for an element of VnV_{n} into a revealing pair.

Given αVn\alpha\in V_{n}, we will denote by RαR_{\alpha} the set of all revealing pairs for α\alpha. We will use the symbol \sim as a relation in the fashion (A,B,σ)α(A,B,\sigma)\sim\alpha denoting the fact that (A,B,σ)Rα(A,B,\sigma)\in R_{\alpha}. In this case, we can further name leaves of ABA-B and BAB-A. If λ\lambda is a leaf of ABA-B and λ\lambda is not a repeller, then we say λ\lambda is a source. If λ\lambda is a leaf of BAB-A and λ\lambda is not an attractor, then we say λ\lambda is a sink.

There are finitely many process types called “Rollings” introduced by Salazar in [21]. Rollings are methods by which one can carry out a finite collection of simple expansions to a revealing tree pair (A,B,σ)(A,B,\sigma) to produce a new revealing pair (A,B,σ)(A^{\prime},B^{\prime},\sigma^{\prime}).

Below, we give the definitions and some discussion for rollings of type II. We give definitions for the other types of rollings in sub-section 4.2.

The tree pair (A,B,σ)(A^{\prime},B^{\prime},\sigma^{\prime}) is a single rolling of type II from (A,B,σ)(A,B,\sigma) if it is obtained from (A,B,σ)(A,B,\sigma) by adding a copy of a component UU of ABA-B to AA at the last leaf in the orbit of the repeller in UU and to BB at its image; or, by adding a copy of a component WW of BAB-A to AA at the first leaf in the orbit of the attractor (the leaf of AA corresponding to the root node of WW in BB) and to BB at its image.

The tree pair (A,B,σ)(A^{\prime},B^{\prime},\sigma^{\prime}) is a rolling of type II from (A,B,σ)(A,B,\sigma) if it is obtained from (A,B,σ)(A,B,\sigma) by a finite collection of single rollings of type II applied to the initial tree pair (A,B,σ)(A,B,\sigma) in some order.

We now state three lemmas about properties of revealing pairs. All of these properties are fairly straightforward to verify. In the cases of Lemma 4.2 and Lemma 4.3, the curious reader may also refer to the discussion in Sections 3.3 and 3.4 of [21] for alternative proofs.

Below, we slightly abuse the notion of a vertex name, by associating an infinite descending path with an infinite “name” string. This then represents a point in the Cantor set which is the boundary of the infinite tree.

Our first lemma discusses repellers, sources, and fixed points.

Lemma 4.2

Suppose αVn\alpha\in V_{n} and (A,B,σ)α(A,B,\sigma)\sim\alpha, and that λ\lambda is a leaf of a component CC of ABA-B, so that IAC(λ)=(λαi)i=0sIAC(\lambda)=(\lambda\alpha^{i})_{i=0}^{s}.

  1. 1.

    If λ\lambda is a repeller (so λαs\lambda\alpha^{s} is an ancestor of λ\lambda), set γi\gamma_{i} to be the name of the node λαi\lambda\alpha^{i} for each 0is0\leq i\leq s. Set Γ\Gamma to be the string which is the suffix one would append to the name of the node γs\gamma_{s} to obtain the name of the node γ0\gamma_{0} i.e. the path from the ancestor to the repeller. We notate this by the expression γsΓ=γ0\gamma_{s}\Gamma=\gamma_{0}. Thus the infinite descending path corresponding to the infinite string γi(Γ)\gamma_{i}(\Gamma)^{\infty} represents a unique repelling fixed point pγip_{\gamma_{i}} in the interval XiX_{i} of n\mathfrak{C}_{n} underneath γi\gamma_{i}, under the action of αs\langle\alpha^{s}\rangle, for each index ii with 0i<s0\leq i<s.

  2. 2.

    if λ\lambda is a source, then λαs\lambda\alpha^{s} is a sink.

  3. 3.

    α\langle\alpha\rangle\cong\mathbb{Z}.

Proof:

The inverse αs\alpha^{-s} of αs\alpha^{s} maps the Cantor set XsX_{s} under γs\gamma_{s} bijectively to the Cantor set X0X_{0} under γ0\gamma_{0}, where X0XsX_{0}\subset X_{s}, in affine fashion. The only infinite descending path which is fixed by this map is the path terminating in Γ\Gamma^{\infty}, thus pγ0=γ0Γ=γsΓΓp_{\gamma_{0}}=\gamma_{0}\Gamma^{\infty}=\gamma_{s}\Gamma\Gamma^{\infty} is the unique attracting fixed point of αs\alpha^{-s} in XsX_{s}, and is thus the unique repelling fixed point of αs\alpha^{s} within X0X_{0} (and even within all of XsX_{s}) under the action of αs\alpha^{s}.

We now show below why each of the points pγip_{\gamma_{i}} are also repelling fixed points of αs\alpha^{s}.

We first obtain a new revealing tree pair (A,B,σ)α(A^{\prime},B^{\prime},\sigma^{\prime})\sim\alpha which is a single rolling of type II from (A,B,σ)(A,B,\sigma), constructed as follows.

Glue a copy of CC at λs\lambda_{s} in BB to produce BB^{\prime}, and a further copy of CC at the leaf γs1\gamma_{s-1} of AA to produce AA^{\prime}. The resulting tree pair (A,B,σ)(A^{\prime},B^{\prime},\sigma^{\prime}) so obtained has all of the leaves of CC of ABA-B as neutral leaves (except in the case where s=1s=1, in which case AA^{\prime} is AA with a copy of CC attached at γ0\gamma_{0}, a leaf of the original CC).

In any case, the original copy of CC, within AA^{\prime}, is now contained in ABA^{\prime}\cap B^{\prime}. The new copy of CC in AA^{\prime} is a complementary component of ABA^{\prime}-B^{\prime} and contains the leaf γs1Γ\gamma_{s-1}\Gamma as a repeller. There are no other new complementary components for the tree pair (A,B,σ)(A^{\prime},B^{\prime},\sigma^{\prime}), which therefore must represent a revealing pair for α\alpha.

Now, by a minor adjustment to the argument in first paragraph, the point pγs1p_{\gamma_{s-1}} is a fixed repelling point of αs\alpha^{s}.

We can now continue inductively in this fashion to show that each point pγip_{\gamma_{i}} is the unique repelling fixed point of XiX_{i} under the action of αs\alpha^{s} by building a revealing pair (A′′,B′′,σ′′)(A^{\prime\prime},B^{\prime\prime},\sigma^{\prime\prime}) for α\alpha with the point pγip_{\gamma_{i}} as a point in a repelling leaf of a complementary component of (A′′,B′′,σ′′)(A^{\prime\prime},B^{\prime\prime},\sigma^{\prime\prime}) with shape CC and spine Γ\Gamma rooted at γi\gamma_{i}.

We leave the second point of the lemma to the reader, while the third point is immediate from the first since some power of α\alpha has a repelling fixed point.

\diamond

We call each subset XiX_{i} a basin of repulsion for α\alpha, since all points in XiX_{i} eventually flow out of XiX_{i} under repeated iteration of αs\alpha^{s}, never to return, except for pγip_{\gamma_{i}}, for each 0is0\leq i\leq s. We call the string Γ\Gamma above the spine of the repeller γ0\gamma_{0} or the spine of CC, as it describes the shape of the path in CC from the root of CC to the repeller γ0\gamma_{0}.

We call each point pγip_{\gamma_{i}} a periodic repelling point of α\alpha for each index ii with 0i<s0\leq i<s. We may also call these points fixed repelling points of αs\alpha^{s}, noting in passing that not all periodic repelling points of α\alpha are necessarily fixed by αs\alpha^{s}. In similar fashion we call the sequence of points (pγi)i=0s1(p_{\gamma_{i}})_{i=0}^{s-1} a periodic orbit of periodic repelling points for α\alpha. We denote by α\mathcal{R}_{\alpha} the set of periodic repelling points of α\alpha, noting that it is a finite set.

We now give a similar lemma discussing attractors and sinks.

Lemma 4.3

Suppose αVn\alpha\in V_{n} and (A,B,σ)α(A,B,\sigma)\sim\alpha and λ\lambda is a leaf of a component CC of BAB-A, so that IAC(λ)=(λαi)i=r0IAC(\lambda)=(\lambda\alpha^{i})_{i=r}^{0}, for some negative integer rr.

  1. 1.

    If λ\lambda is an attractor (so λαr\lambda\alpha^{r} is an ancestor of λ\lambda), set γi\gamma_{i} to be the name of the node λαi\lambda\alpha^{i} for each ri0r\leq i\leq 0. In this case the string Γ\Gamma which has γrΓ=γ0\gamma_{r}\Gamma=\gamma_{0} has the property that the infinite descending path corresponding to the name γi(Γ)\gamma_{i}(\Gamma)^{\infty} represents a unique attracting fixed point pγip_{\gamma_{i}} in the interval XiX_{i} of n\mathfrak{C}_{n} underneath γi\gamma_{i}, under the action of αr\langle\alpha^{r}\rangle, for each index ii with ri<0r\leq i<0.

  2. 2.

    If λ\lambda is a sink, then λαr\lambda\alpha^{r} is a source.

  3. 3.

    α\langle\alpha\rangle\cong\mathbb{Z}.

Proof:

This proof is similar to the proof of the previous lemma, where here α1\alpha^{-1} has a tree pair with λ\lambda as a repeller, and each pγip_{\gamma_{i}} is a periodic repelling point of α1\alpha^{-1}.

\diamond

We now extend the notation from Lemma 4.2 to apply to the sets named in Lemma 4.3 as below.

We call each subset XiX_{i} a basin of attraction for α\alpha as indicated by (A,B,σ)(A,B,\sigma), since all points in XiX_{i} eventually limit to pγip_{\gamma_{i}} under repeated iteration of αr\alpha^{-r}, for each ri0r\leq i\leq 0. We call the string Γ\Gamma above the spine of the attractor γ0\gamma_{0} or the spine of CC, as it describes the shape of the path in CC from the root of CC to the attractor γ0\gamma_{0}.

We call each point pγip_{\gamma_{i}} a periodic attracting point of α\alpha for each index ii with ri0r\leq i\leq 0. We may also call these points fixed attracting points of αr\alpha^{-r}. In similar fashion we call the sequence of points (pγi)i=r1(p_{\gamma_{i}})_{i=r}^{-1} a periodic orbit of periodic attracting points for α\alpha. We denote by 𝒜α\mathcal{A}_{\alpha} the set of periodic attracting points of α\alpha, noting that it is a finite set.

In the previous two lemmas, if λ\lambda is a source or a sink (case two in each lemma), we refer to IAC(λ)IAC(\lambda) as a source-sink chain.

The next lemma follows directly from the two above, and the classification of the orbits of the leaves of AA and BB. This lemma is a version of one result proved by Burillo, Cleary, Stein and Taback in their joint work [8] as Proposition 6.1. We give a new proof here, as the situation is greatly simplified through the use of revealing pairs.

Lemma 4.4

Suppose αVn\alpha\in V_{n}. There is an integer nn so that α\alpha has order nn if and only if there is a tree pair (A,B,σ)(A,B,\sigma) representing α\alpha with A=BA=B.

Proof:

Suppose α\alpha does not have infinite order, and let (A,B,σ)α(A,B,\sigma)\sim\alpha be a revealing pair representing α\alpha. We must have that (A,B,σ)(A,B,\sigma) admits no repellers or attractors, thus both ABA-B and BAB-A are empty, and so A=BA=B.

Suppose instead (A,B,σ)(A,B,\sigma) is a tree pair representing α\alpha with A=BA=B. Then it is immediate from the definition of multiplication for tree pairs that the order of α\alpha is the order of the permutation σ\sigma.   \diamond

We denote by 𝒫α\mathcal{P}_{\alpha} the points of n\mathfrak{C}_{n} which underlie the periodic neutral leaves of L(A,B,σ)L_{(A,B,\sigma)}. Note that the set 𝒫α\mathcal{P}_{\alpha} is independent of the choice of revealing pair used to represent α\alpha. We further denote by Per(α)Per(\alpha) the set of all periodic points of α\alpha, that is

Per(α)=α𝒜α𝒫α.Per(\alpha)=\mathcal{R}_{\alpha}\sqcup\mathcal{A}_{\alpha}\sqcup\mathcal{P}_{\alpha}.

The following tree pairs represent the previously defined element θT\theta\in T (we apply an augmentation to our first tree pair to produce a revealing pair representing θ\theta).

\psfrag{1}[c]{$1$}\psfrag{2}[c]{$2$}\psfrag{3}[c]{$3$}\psfrag{4}[c]{$4$}\psfrag{5}[c]{$5$}\psfrag{6}[c]{$6$}\psfrag{7}[c]{$7$}\psfrag{D}[c]{$D$}\psfrag{R}[c]{$R$}\includegraphics[height=165.0pt,width=240.0pt]{elmt.T.compare.eps}

A quick examination of the second tree pair above should convince the reader that θ\theta has order five and rotation number 2/52/5.

We now give a series of diagrams for a revealing pair (A,B,σ)(A,B,\sigma) representing a particular non-torsion element αV\alpha\in V. This element has a revealing pair which contains many of the structures we have been discussing. In each diagram below we illustrate some of the particular aspects we have discussed above.

Below is an example of our revealing pair (A,B,σ)(A,B,\sigma), with the neutral leaves underlined. Recall that the left tree is AA, and the right tree is BB.

\psfrag{1}[c]{$\,\,1$}\psfrag{2}[c]{$\,\,2$}\psfrag{3}[c]{$\,\,3$}\psfrag{4}[c]{$\,\,4$}\psfrag{5}[c]{$\,\,5$}\psfrag{6}[c]{$\,\,6$}\psfrag{7}[c]{$\,\,7$}\psfrag{8}[c]{$\,\,8$}\psfrag{9}[c]{$\,\,9$}\psfrag{10}[c]{$\,\,10$}\psfrag{D}[c]{$D$}\psfrag{R}[c]{$R$}\includegraphics[height=150.0pt,width=300.0pt]{revPair.1.5.eps}

In the next diagram, we point out a periodic orbit of neutral leaves of length two.

\psfrag{1}[c]{$\,\,1$}\psfrag{2}[c]{$\,\,2$}\psfrag{3}[c]{$\,\,3$}\psfrag{4}[c]{$\,\,4$}\psfrag{5}[c]{$\,\,5$}\psfrag{6}[c]{$\,\,6$}\psfrag{7}[c]{$\,\,7$}\psfrag{8}[c]{$\,\,8$}\psfrag{9}[c]{$\,\,9$}\psfrag{10}[c]{$\,\,10$}\psfrag{D}[c]{$D$}\psfrag{R}[c]{$R$}\includegraphics[height=150.0pt,width=300.0pt]{revPair.2.eps}

Below, the vertex 1111111111 in BB is an attractor; its iterated augmentation chain is (111,11111)(111,11111), a sequence of length two. The fixed point of the attractor corresponds to the value 11 in the unit interval. The word Γ\Gamma for this attractor is 1111.

\psfrag{1}[c]{$\,\,1$}\psfrag{2}[c]{$\,\,2$}\psfrag{3}[c]{$\,\,3$}\psfrag{4}[c]{$\,\,4$}\psfrag{5}[c]{$\,\,5$}\psfrag{6}[c]{$\,\,6$}\psfrag{7}[c]{$\,\,7$}\psfrag{8}[c]{$\,\,8$}\psfrag{9}[c]{$\,\,9$}\psfrag{10}[c]{$\,\,10$}\psfrag{D}[c]{$D$}\psfrag{R}[c]{$R$}\includegraphics[height=150.0pt,width=300.0pt]{revPair.3.eps}

In the diagram to follow, the vertices 00000000 and 011011 represent repellers in AA. The dotted paths track the orbits of the repellers along their iterated augmentation chains.

\psfrag{1}[c]{$\,\,1$}\psfrag{2}[c]{$\,\,2$}\psfrag{3}[c]{$\,\,3$}\psfrag{4}[c]{$\,\,4$}\psfrag{5}[c]{$\,\,5$}\psfrag{6}[c]{$\,\,6$}\psfrag{7}[c]{$\,\,7$}\psfrag{8}[c]{$\,\,8$}\psfrag{9}[c]{$\,\,9$}\psfrag{10}[c]{$\,\,10$}\psfrag{D}[c]{$D$}\psfrag{R}[c]{$R$}\includegraphics[height=150.0pt,width=300.0pt]{revPair.5.eps}

Finally, we highlight the fact that sources flow to sinks. Note how the lengths of the paths from sources to sinks are not uniform. In particular, the source 01000100 first hops to 110110 before next landing in the basin of attraction under the vertex 000000.

\psfrag{1}[c]{$\,\,1$}\psfrag{2}[c]{$\,\,2$}\psfrag{3}[c]{$\,\,3$}\psfrag{4}[c]{$\,\,4$}\psfrag{5}[c]{$\,\,5$}\psfrag{6}[c]{$\,\,6$}\psfrag{7}[c]{$\,\,7$}\psfrag{8}[c]{$\,\,8$}\psfrag{9}[c]{$\,\,9$}\psfrag{10}[c]{$\,\,10$}\psfrag{D}[c]{$D$}\psfrag{R}[c]{$R$}\includegraphics[height=150.0pt,width=300.0pt]{revPair.5.5.eps}

4.2 Discrete train tracks and laminations

In this subsection we are concerned with modelling the dynamics in n\mathfrak{C}_{n} under the action of the group generated by an element vv of VnV_{n}. Naturally, the relevant information is contained in any particular revealing pair ρ=(C,D,θ)\rho=(C,D,\theta) for the element, but we have found that tools from surface topology afford us another way to visualize these dynamics. In particular, these dynamics can be usefully described by a combinatorial object introduced by Thurston (see [22, 23]) to aid in the study of surfaces, namely, a train track. Given a revealing pair, it is easy to draw a combinatorial train track which in turn “carries” a lamination in some compact surface with boundary. It is this lamination which models, in some sense, the movement of points in the Cantor set under iteration of the map vv. As in the theory of laminations carried by a train track, one quickly realizes that most of the relevant visual information is actually contained in the train-track object (see for instance, Penner and Harer’s book [18]). We call the train track object developed here a discrete train track, even though it is continuous in nature; the name is meant to emphasize that our train tracks model dynamics in a totally disconnected set under iterations of a fixed map.

Below, we describe in detail our method for generating a discrete train track TTρTT_{\rho} from the revealing pair ρ\rho representing the general element vVnv\in V_{n} mentioned above, and briefly describe how to model the lamination it carries. We build an example discrete train track and lamination using the element α\alpha and the tree pair (A,B,σ)(A,B,\sigma) from the previous subsection. Finally, we discuss some of the utility of TTρTT_{\rho} and some facts about how TTρTT_{\rho} would change under basic operations applied to vv (conjugation, then a choice of representative revealing pair) or ρ\rho (Salazar’s rollings). In the next subsection, we describe a derived object, the flow graph, which carries less information than the train track (although flow graphs in general still contain enough information to answer many dynamics questions). In our current practice, we find discrete train tracks and flow graphs to be helpful for understanding dynamics, while the generating revealing pairs tend to be helpful for any involved computations and for specifying elements with appropriate desired dynamics.

Note that any such discrete train track TTρTT_{\rho} is a representative of an equivalence class of similar train tracks, where the equivalence class is determined by all the discrete train tracks for revealing pairs equivalent to ρ\rho up to Salazar’s rollings, choice of location for drawing the complementary trees along the orbits of repelling and attracting fixed points, and conjugacy in VnV_{n}. Thus we are picking a representative object which is in some sense less well chosen than some “minimal” train track in this class. The result of applying Belk and Matucci’s geometric conjugacy invariant, the closed strand diagram, is very close to what a “minimal” discrete train track for an element of VnV_{n} should be (leaving out the demarcations representing iterating the element, and of course, including some simplifications which result from the effects of conjugation (see [3])).

Here is how one draws a discrete train track from a revealing pair.

  1. 1.

    List all of the iterated augmentation chains for the revealing pair.

  2. 2.

    For each chain representing a non-trivial orbit of a periodic neutral leaf:

    1. (a)

      Draw a circle.

    2. (b)

      If the chain represents an orbit of length rr, demarcate the circle into rr equal subintervals (typically we end these intervals with dots), and label these with the names of the nodes carrying the neutral leaves, in a clockwise order.

  3. 3.

    For each chain of a repeller:

    1. (a)

      Draw a circle.

    2. (b)

      If the repelling periodic point travels an orbit of length rr, then demarcate the circle into rr equal subintervals as above, and label these with the names of the nodes carrying the orbit of the repeller in a counter-clockwise order.

    3. (c)

      For the segment corresponding to the interval of the repelling periodic point of the complementary tree, instead of the one label mentioned in the last point, we label the two ends of the segment with the top and bottom of the spine of the repeller (top (root node of complementary component) before bottom in counter-clockwise order).

    4. (d)

      Lay the complementary component of the repeller along the sub-arc of the circle with the spine labels, gluing the spine to the circle.

      1. i.

        Scale the complementary component so that the spine is the length of the appropriate sub-arc of the circle.

      2. ii.

        Smooth out the tree (and the spine in particular) and bend the spine so that the spine has the same shape as the sub-arc of the circle with the labels from the spine nodes (preserving the current length of the spine).

      3. iii.

        Rigidly rotate the scaled, smoothed, and bent tree in the plane, and translate it so that the spine can be identified with the appropriate sub-arc of the circle.

  4. 4.

    For each chain for an attractor:

    1. (a)

      Draw a circle.

    2. (b)

      If the attracting periodic point travels an orbit of length rr, then demarcate the circle into rr equal subintervals, and label these with the names of the nodes carrying the orbit of the attractor in a counter-clockwise order.

    3. (c)

      For the segment corresponding to the interval of the attracting periodic point of the complementary tree, instead of the one label mentioned in the last point, we label the two ends of the segment with the top and bottom of the spine of the attractor (bottom (attracting leaf node of complementary component) before top in counter-clockwise order).

    4. (d)

      Lay the complementary component of the attractor along the sub-arc of the circle with the spine labels, gluing the spine to the circle.

      1. i.

        reflect the complementary component of the attractor across a vertical axis. Thus, the root of the component is now drawn at the bottom, while the left and right hand sides are preserved, respectively, as left and right hand sides.

      2. ii.

        Scale the complementary component so that the spine is the length of the appropriate sub-arc of the circle.

      3. iii.

        Smooth out the tree (and the spine in particular) and bend the spine so that the spine has the same shape as the sub-arc of the circle with the labels from the spine nodes (preserving the current length of the spine).

      4. iv.

        Rigidly rotate the reflected, scaled, smoothed, and bent tree in the plane, and translate it so that the spine can be identified with the appropriate sub-arc of the circle.

  5. 5.

    For each source-sink chain, drawn a line connecting the appropriate sources and sinks (lines may have to cross each other in the case of VnV_{n} which is why the lamination carried by the train track can only be embedded in an surface with boundary; strips can pass “under” each other). Demarcate each line with dots representing the length of the source-sink chain, and label sub-arcs with appropriate node labels as above for other sorts of iterated augmentation chains.

  6. 6.

    Add parenthetical labels for splittings of trees where the support of the whole tree will be mapped away by one application of the element. Add parenthetical labels anywhere else as desired to improve clarity. (This step is not strictly necessary, but we find it to be helpful.)

If we follow the process above for the element α\alpha, the diagram below is an example of what we may obtain (we include some drawn under-crossings following the methods from drawing knots from knot theory).

[Uncaptioned image]

To visualize the lamination carried by a discrete train track, for each sub-arc in the demarcations of the discrete train tracks, one can draw a transverse copy of an II-fiber with an embedded copy of the Cantor set n\mathfrak{C}_{n}. If we stabilize this by building a product with the interval II (which lines we draw parallel to any train track sub-arc), then one has a picture of the lamination carried by the train track. (Glue the ends of the II-fibers of the Cantor sets using the rules provided by the map vv, as a local picture of a mapping cone on the II-fiber running transverse to any demarcation dot, allowing contraction and expansion in the transverse II-fibers near to any tree-splitting.

Note that by reflecting the attracting complementary components as directed, we are able to draw the resulting carried lamination on a compact orientable surface with boundary. (The “left” and “right” portions of a Cantor set underlying a node are correctly associated without any twisting.)

Separately, one can observe that the group homomorphism 𝒮\mathcal{S} of the penultimate subsection of this article is connected with holonomy measurements for the carried lamination along repelling cycles when applied to the element we are centralizing.

A local diagram of the lamination carried by this train track local to the repelling cycle in the upper left corner is included in the diagram below. One can see the rescaling near the periodic repeller and portions of the Cantor set moving away along flow lines.

[Uncaptioned image]

Let TT1TT_{1} be a discrete train track drawn in accordance with the method above. We define the support of the discrete train track TT1TT_{1} to be given as the union of the Cantor sets underlying the node labels in the iterated augmentation chains of the (non-trivial) periodic orbits, the source-sink chains, and the orbits of the repellers and attractors used to create the drawing of TT1TT_{1}. Seeing a discrete train track as a (possibly disconnected) graph, we define the components of TT1TT_{1} to be the connected components of TT1TT_{1} as a graph. Given any such component XX, we describe the support of the component XX to be the union of the Cantor sets underlying the node labels of XX. We call a component a torsion component if the component is a circle generated from a periodic orbit of neutral leaves, otherwise we call a component a non-torsion component. Similar language will be defined for flow graphs, below.

We now mention some technical facts to do with the relationships amongst discrete train tracks drawn from distinct revealing pair representatives for an element of VnV_{n}, and also relationships which may arise as a consequence of conjugacy.

We give the basic definitions of Salazar’s rollings below, with the intention of understanding of the variance of discrete train tracks across the set of all representative revealing pairs for an element of VnV_{n}. The reader is encouraged to read Salazar’s Section 3.5 in [21], where she defines rollings of various types and traces the various impacts of rollings on revealing tree pairs.

The tree pair (A,B,σ)(A^{\prime},B^{\prime},\sigma^{\prime}) is a single rolling of type E from (A,B,σ)(A,B,\sigma) if it is obtained from (A,B,σ)(A,B,\sigma) by adding an nn-caret to AA and BB along each of the leaves of an iterated augmentation chain corresponding to one of two types of iterated augmentation chains; either to all of the leaves in AA and BB of a periodic orbit of neutral leaves, or to the initial source leaf in AA for a source-sink chain, and then to all of the neutral leaves in AA and BB of that chain, and then to the leaf of BB corresponding to the sink of that chain. (These are called elementary rollings.)

The tree pair (A,B,σ)(A^{\prime},B^{\prime},\sigma^{\prime}) is a single rolling of type I from (A,B,σ)(A,B,\sigma) if it is obtained from (A,B,σ)(A,B,\sigma) by adding a cancelling tree along all of the leaves of AA along the orbit of a repeller, and at BB at the image of these leaves under the map, or by adding a cancelling tree at all the leaves of AA which appear in the reverse orbit of an attractor, and at the leaves of BB to which these leaves of AA are mapped. If WW is the complementary component of ABA-B or BAB-A corresponding to the repeller or attractor in this discussion, and WW is rooted at node Σ\Sigma, then a tree CC is a cancelling tree for WW if it is obtained from WW by first choosing a proper, non-empty prefix Δ\Delta of the spine Γ\Gamma of WW (so that Γ=ΔΘ\Gamma=\Delta\Theta for some suffix Θ\Theta), and then taking CC to be the maximal sub-tree of WW which has root Σ\Sigma and containing the node ΣΔ\Sigma\Delta as a leaf. (Note that if one carries out this process, the corresponding complementary component WW^{\prime} created in ABA^{\prime}-B^{\prime} or BAB^{\prime}-A^{\prime} for the tree pair (A,B,σ)(A^{\prime},B^{\prime},\sigma^{\prime}) will now be rooted at ΣΔ\Sigma\Delta and will have spine ΘΔ\Theta\Delta).

Finally, recall from sub-section 4.1 that a tree pair (A,B,σ)(A^{\prime},B^{\prime},\sigma^{\prime}) is a single rolling of type II from (A,B,σ)(A,B,\sigma) if it is obtained from (A,B,σ)(A,B,\sigma) by adding a copy of a component UU of ABA-B to AA at the last leaf in the orbit of the repeller in UU and to BB at its image; or, by adding a copy of a component WW of BAB-A to AA at the first leaf in the orbit of the attractor (the leaf of AA corresponding to the root node of WW in BB) and to BB at its image.

The following lemma lists some basic properties of discrete train tracks. The reader will not be required to use this lemma later in the paper, although it gives a separate view of some arguments.

Lemma 4.5

Suppose τ1P1=(D1,R1,σ1)\tau_{1}\sim P_{1}=(D_{1},R_{1},\sigma_{1}) and τ2P2=(D2,R2,σ2)\tau_{2}\sim P_{2}=(D_{2},R_{2},\sigma_{2}) are elements of VnV_{n}, and that TT1TT_{1} and TT2TT_{2} are the corresponding train tracks derived from the revealing pair representatives P1P_{1} and P2P_{2} of these elements. Then we have the following:

  1. 1.

    Suppose fVnf\in V_{n} so that τ1f=τ2\tau_{1}^{f}=\tau_{2}. Then,

    1. (a)

      ff induces a 111-1 correspondence between the components of TT1TT_{1} which describe dynamics around repelling/attracting orbits and the components of TT2TT_{2} which describe dynamics around repelling/attracting orbits.

    2. (b)

      This correspondence also guarantees that the individual cycles in these components are also carried to cycles of the same length and type (repellers must move to repellers and attractors to attractors).

    3. (c)

      This correspondence preserves the labelings of the spine of the complementary components for corresponding cycles.

  2. 2.

    Suppose τ1=τ2\tau_{1}=\tau_{2} and (D1,R1,σ1)(D_{1},R_{1},\sigma_{1}) differs from (D2,R2,σ2)(D_{2},R_{2},\sigma_{2}) by an application of a rolling of type XX. Then

    1. (a)

      if XX is a rolling of Type E, then a train track connecting a source node to a sink node will be split into nn such parallel tracks along its length, or a periodic circle representing a periodic orbit of neutral leaves will be split into nn copies of “parallel” periodic circles.

    2. (b)

      if XX is a single rolling of Type I, then there is a suffix Δ\Delta of the spine so that the labels of all the sub-arcs in the relevant repelling or attracting circle will be modified by the addition of Δ\Delta as a suffix to all of the labels on the circle. Also, some source-sink chains which are incident on the affected circle will be lengthened by the length of the orbit of the repeller or the attractor.

    3. (c)

      if XX is a single rolling of Type II, then the source-sink chains with sources or sinks beginning or ending in the relevant complementary component will all increase their lengths by one (each of these new arcs will need appropriate labels added), and the sub-arc of the relevant repelling (attracting) circle to which the spine of the complementary tree is glued will move one location backward (forward) in the cyclic ordering of the arcs in that circle (respectively). Finally, the labels of the affected arcs on the circle will change (the arc corresponding to the old spine will now be labelled by the old leaf label for the attractor or the repeller of that spine, while the arc corresponding to the new spine will have as root label its old label, and as leaf label, its old label concatenated with the word corresponding to the spine of the repeller of attractor).

Proof: The latter points about rollings follow directly from the definitions of rollings, as given in subsection 4.1.

The first two sub-points of 1 are a result of the fact that VnV_{n} is a group of homeomorphisms, and topological conjugacy preserves the properties mentioned. The third sub-point of 1 follows from the fact that elements of VnV_{n} do not change infinite suffixes; so, the points in the finite orbit of a repelling periodic point or of an attracting periodic point all have the same infinite repeating suffix (that is, as described in Lemmas 4.3 and 4.2). Now the conjugating element ff again cannot change this infinite suffix class, so the resulting orbit will consist of points with this same infinite suffix. (Note first that that Belk and Matucci [3] also derive the infinite suffix of a finite repelling or attracting orbit as a conjugacy invariant for elements of F<T<VF<T<V, using a method very similar to our discrete train tracks, although Belk and Matucci’s definition is mildly different, and second that one can choose a representative revealing pair for the conjugate version τ2\tau_{2} of τ1\tau_{1} so that the spine is cyclically rotated, but by applying a rolling of type IIII, this spine can be rotated back to the original spine, thus producing a discrete train track with the same spine labelling on the appropriate sub-arc of the circle (although, the prefixes of all nodes on the circle will be lengthened).)   \diamond

4.3 Flow graphs

The flow graph of any revealing pair p=(A,B,σ)αVnp=(A,B,\sigma)\sim\alpha\in V_{n} is a labeled directed graph which is, in some sense, a quotient object from a discrete train track (we loose the visual aspects of the branching of source sink lines as they leave the orbit of a repeller or join the orbit of an attractor, although even this information can be recovered from labels). We now describe how to build a flow graph from a revealing pair.

For each repeller of AA, we draw a vertex. For each attractor of BB, we draw a vertex. For each neutral leaf of L(A,B,σ)L_{(A,B,\sigma)} that is part of a periodic orbit of neutral leaves for pp, we draw a vertex. We draw a directed edge from a repeller to an attractor for each source in the basin of repulsion of the repeller whose iterated augmentation chain terminates in a sink in the basin of attraction of the attractor (we call these source-sink flow lines or by similar language; they are in one-one correspondence with the set of source-sink chains for the pair pp). We draw a directed edge from each repelling and attracting vertex to itself whenever the period of the corresponding repeller (or attractor) is greater than one (we call this a repelling (or attracting) periodic orbit). We draw an edge connecting two vertices representing periodic neutral leaves if a single iteration of α\alpha will take the first leaf to the second, whenever these leaves are not the same. We label all source-sink flow lines with the appropriate iterated augmentation chain. We label all repelling and attracting periodic orbits (even of length one) with the finite periodic orbit of the actual points in n\mathfrak{C}_{n}, each such point labeled by its infinite descending path in PnP_{n}. (Note that one can detect the names of repelling and attracting basins by deleting the infinite “Γ\Gamma^{\infty}” portion of these labels.)

The diagram below is an example of a flow graph for the tree (A,B,σ)(A,B,\sigma) we have been examining. Strings of form Γ\Gamma^{\infty} are indicated by overlines in the labels of the diagram.

[Uncaptioned image]

The components of the flow graph for a revealing pair naturally decompose into two sets; components representing the flow (under the action of α\langle\alpha\rangle along the underlying sets of a periodic orbit of neutral leaves for the revealing pair, called torsion components, and components representing flows from basins of repulsion to basins of attraction, and characterizing the orbits of repelling and attracting periodic points, called non-torsion components.

Any flow line in the graph from a repeller to an attractor can be thought of as representing the complete bi-infinite forward and backward orbit in n\mathfrak{C}_{n} of the interval underlying the source, under the action of α\langle\alpha\rangle. The set so determined is called the underlying set for the flow line, and it limits on some of the periodic repelling points of α\alpha (on one end), and on some of the periodic attracting points of α\alpha (on the other end).

Given a component CC of a flow graph for α\alpha, we can discuss the underlying set for the component CC in n\mathfrak{C}_{n}. We define this set as the closure of the underlying sets of the flow lines of the component if the component has a flow line (thus, we will capture the limiting repelling and attracting points in their finite orbits), and otherwise as the underlying set of the finite periodic orbit of an appropriate neutral leaf.

We observe in passing the following remark, which the reader can verify as a test of their understanding. The proof of this remark follows directly from the definition of flow graph or can be obtained as a consequence of Lemma 4.5.

Remark 4.6

Given any revealing pair p=(A,B,σ)p=(A,B,\sigma) representing αVn\alpha\in V_{n}, and p=(A,B,σ)p^{\prime}=(A^{\prime},B^{\prime},\sigma^{\prime}) the result of a single rolling of type II from (A,B,σ)(A,B,\sigma), the flow graph α,p\mathscr{F}_{\alpha,p^{\prime}} for α\alpha generated by pp^{\prime} is identical to the flow graph α,p\mathscr{F}_{\alpha,p} for α\alpha generated by pp, except for one pair (U,U)(U,U^{\prime}) of corresponding components. The components UU and UU^{\prime} are both non-torsion components which have the same underlying set and are isomorphic as graphs. However, the flow lines of UU^{\prime} will bear different labels from the flow lines of UU, having had some of its iterated augmentation chains lengthened by one hop, and, one of the periodic orbits in the labeling of UU^{\prime}, of either a repelling periodic point or of an attracting periodic point, may be cyclically permuted from the corresponding orbit as labeled in UU.

A mildly more difficult statement is that in the following lemma.

Lemma 4.7

Suppose αVn\alpha\in V_{n} is represented by a revealing pair p=(A,B,σ)αp=(A,B,\sigma)\sim\alpha. If z1z_{1} and z2z_{2} underlie distinct non-torsion components of p,α\mathscr{F}_{p,\alpha}, then they cannot underlie the same component of a flow graph representing αk\alpha^{k} for any integer kk.

Proof:

Powers of α\alpha may split orbits of repelling (or attracting) periodic points under the action of α\langle\alpha\rangle, but they can never move a point from the underlying set of one component of the flow graph to the underlying set of other components; these sets are themselves created as the unions of the images of the forward and backward orbits of the points underlying the interval portions of the fundamental domain of α\alpha which are carried in affine fashion by α\alpha.

\diamond

The next lemma is a direct consequence of the work of Salazar in [21].

Lemma 4.8

Suppose nn is a positive integer and αVn\alpha\in V_{n} is represented by the revealing pair (A,B,σ)(A,B,\sigma). For all integers k0k\geq 0, set mkm_{k} to be the number of finite periodic orbits of neutral leaves for the tree pair (A,B,σ)(A,B,\sigma) with period kk and set rkr_{k} to be mk(modn1)m_{k}\pmod{n-1}.

There is a conjugate u=αfu=\alpha^{f} for some fVnf\in V_{n}, where uu has a representative tree pair (E,F,χ)(E,F,\chi), so that the following hold.

  1. 1.

    The element uu has rkr_{k} periodic orbits of neutral leaves with period kk.

  2. 2.

    For each k0k\geq 0, the number rkr_{k} is a conjugacy invariant of α\alpha.

  3. 3.

    The number of components of FEF-E is the same as the number of components of BAB-A, and these components have the same shape, and the same spine.

  4. 4.

    Likewise, the number of components of EFE-F is the same as the number of components of ABA-B, and these components have the same shape, and the same spine.

Sketch of proof:

The statements in the lemma above are for the most part stated in Corollary 1 in Section 4.4 of [21]. We explain our statements relating to the modular arithmetic involved in the reduction of the number of finite periodic orbits of the same length, as that is not given Salazar’s corollary. The fact that spines are preserved as well is not explicitly stated by Salazar, but it is an immediate consequence of her techniques.

One can describe any conjugation in VnV_{n} in terms of tree pairs by a process which we now sketch. Suppose u=αfu=\alpha^{f} and let ff be represented by a tree pair (C,D,τ)(C,D,\tau). There is an expansion (E,F,χ)(E^{\prime},F^{\prime},\chi^{\prime}) of (A,B,σ)(A,B,\sigma) so that C(EF)C\subset(E^{\prime}\cap F^{\prime}). One replaces the sub-tree CC inside of (EF)(E^{\prime}\cap F^{\prime}) by DD, producing a revealing tree pair (E,F,χ)(E,F,\chi) for uu (one has to remember to percolate out the effects of τ\tau in both trees) with the same number of neutral leaves as for the (E,F,χ)(E^{\prime},F^{\prime},\chi^{\prime}) tree pair.

This process of conjugation produces a revealing pair for αf\alpha^{f} whose periodic neutral leaves are generated by expansions along the iterated augmentation chains of the original neutral cycles (together with some permuting of locations in the universal tree). Thus, the number of neutral periodic leaves of the result is in the same congruence class modulo (n1)(n-1) as for the tree pair for α\alpha.

One can now find a conjugate of α\alpha which has all finite periodic orbits of neutral leaves of the same period adjacent in the universal tree 𝒯n\mathscr{T}_{n}. If there are more than nn such orbits, one can find a conjugate such that nn of the orbits travel in parallel, all carried in the kk-orbit of an nn-caret. One can also choose a conjugation so that one can build a new revealing pair for the resulting element with the same number of periodic orbit neutral leaves as in (A,B,σ)(A,B,\sigma). By simple reductions along the full orbit of the caret, one can reduce the current representative tree pair to another revealing pair with k(n1)k\cdot(n-1) fewer periodic neutral leaves, and can repeat this overall process until there are fewer than nn orbits of neutral leaves for any period kk.

\diamond

5 A partition of n\mathfrak{C}_{n}

Given positive integer nn and some αVn\alpha\in V_{n} with representative revealing pair (A,B,σ)α(A,B,\sigma)\sim\alpha, we can decompose n\mathfrak{C}_{n} as TαZαT_{\alpha}\cup Z_{\alpha}. Here, we are using the following notation:

  • TαT_{\alpha}: the subset of n\mathfrak{C}_{n} lying under the neutral leaves in AA which are on cyclic orbits (possibly of length one).

  • ZαZ_{\alpha}: the subset of n\mathfrak{C}_{n} underlying the root node of the complementary components of ABA-B and BAB-A, as well as any neutral leaves which are part of source-sink chains.

The reader can easily verify the following.

Lemma 5.1

If αVn\alpha\in V_{n} and βCVn(α)\beta\in C_{V_{n}}(\alpha), then

  1. 1.

    Tαβ=TαT_{\alpha}\beta=T_{\alpha}, and

  2. 2.

    Zαβ=ZαZ_{\alpha}\beta=Z_{\alpha}.

This lemma allows us to work to comprehend centralizers over each set, without regard to the behavior of these centralizers in the other regions.

From here to near the end of Section 7, we will assume that nn and α\alpha are fixed, and α\alpha is the element whose centralizer in VnV_{n} we are analyzing, and that (A,B,σ)(A,B,\sigma) is a revealing pair representing α\alpha. We further assume through the use of the third point of Remark 3.2 and of Lemma 4.8 that for each periodic cycle length mim_{i} of neutral leaves of AA under the action of α\langle\alpha\rangle, that there are precisely ri<nr_{i}<n such cycles of periodic neutral leaves. These values mim_{i} and rir_{i} are the numbers which appear in the semi-direct product terms in the left-hand direct product in the statement of Theorem 1.1.

Suppose GVnG\leq V_{n} and XnX\subset\mathfrak{C}_{n}. We define

GX={vGv|n\X=Id|n\X}.G_{X}=\left\{v\in G\mid v|_{\mathfrak{C}_{n}\backslash X}=Id|_{\mathfrak{C}_{n}\backslash X}\right\}.

so that GXG_{X} is the subgroup of elements of GG which act as the identity except on the set XX. Lemma 5.1, assures us that there are two commuting elements αTVnTα\alpha_{T}\in{V_{n}}_{T_{\alpha}} and αZVnZα\alpha_{Z}\in{V_{n}}_{Z_{\alpha}} of VnV_{n}, so that α|Tα=αT|Tα\alpha|_{T_{\alpha}}=\alpha_{T}|_{T_{\alpha}}, and α|Zα=αZ|Zα\alpha|_{Z_{\alpha}}=\alpha_{Z}|_{Z_{\alpha}}. Thus, we see immediately that α=αTαZ\alpha=\alpha_{T}\alpha_{Z}.

We will therefore restrict our attention to finding the centralizers CVnTα(αT)C_{\,\,{V_{n}}_{T_{\alpha}}}(\alpha_{T}), and CVnZα(αZ)C_{\,\,{V_{n}}_{Z_{\alpha}}}(\alpha_{Z}). In fact, we have the following corollary to Lemma 5.1.

Corollary 5.2

We have

CVn(α)CVnTα(αT)×CVnZα(αZ)C_{V_{n}}(\alpha)\cong C_{\,\,{V_{n}}_{T_{\alpha}}}(\alpha_{T})\times C_{\,\,{V_{n}}_{Z_{\alpha}}}(\alpha_{Z})

This explains the central direct product in our statement of Theorem 1.1.

6 Centralizers over the set TαT_{\alpha}

Taking advantage of the decomposition given by Corollary 5.2, we analyze the centralizer of α\alpha by restricting our attention to the set TαT_{\alpha}, over which α\alpha acts as an element of torsion.

Below, when we refer to a cycle of α\alpha, we mean a cycle of neutral leaves for (A,B,σ)(A,B,\sigma).

Lemma 6.1

The leaves of AA over TαT_{\alpha} can be partitioned into disjoint sets according to cycle lengths, denoted by SmiS_{m_{i}}, where each SmiS_{m_{i}} consists of all the leaves of AA in a periodic orbit of length mim_{i}. If βCVn(α)\beta\in C_{V_{n}}(\alpha), then β\beta preserves the subsets of n\mathfrak{C}_{n} underlying the leaves in any particular set SmiS_{m_{i}}.

Proof: We first suppose TαT_{\alpha} is not empty, and we further suppose L(A,B,σ)L_{(A,B,\sigma)} admits distinct neutral leaf cycles of length kk and mm under the action of α\alpha, where we chose our labels so that k<mk<m. Finally, suppose βCVn(α)\beta\in C_{V_{n}}(\alpha) acts by mapping a point pp^{\prime} underlying a cycle of length mm to a point p=pβp=p^{\prime}\beta underlying a cycle of α\alpha of length kk. (If there is a γCVn(α)\gamma\in C_{V_{n}}(\alpha) which maps a point underlying a cycle of length kk to a point underlying a cycle of length mm, then γ1\gamma^{-1} will match our requirements.) We now have the following computation:

p=pαk=p(αβ)k=pαkβ=qβ.p=p\alpha^{k}=p(\alpha^{\beta})^{k}=p^{\prime}\alpha^{k}\beta=q\beta.

where q=pαkq=p^{\prime}\alpha^{k} is not pp^{\prime} since the orbit length for pp^{\prime} under α\alpha is mm. However, we have just shown that qβ=pq\beta=p and by assumption pβ=pp^{\prime}\beta=p, so we have a contradiction.   \diamond

We suppose throughout the remainder that there are ss distinct neutral leaf cycle lengths under the induced action of α\langle\alpha\rangle on the periodic neutral leaves in L(A,B,σ)L_{(A,B,\sigma)}, namely {m1,m2,,ms}\{m_{1},m_{2},\ldots,m_{s}\}.

We thus can focus on how a particular βCVk(α)\beta\in C_{V_{k}}(\alpha) can commute with α\alpha over the underlying set of the leaves in any particular set SmiS_{m_{i}}. This is the reason for the left-hand direct product with ss terms in our statement of Theorem 1.1. (Here, we are following the same logic as used in the beginning of this section which allowed us to focus our attention on TαT_{\alpha} based on the dynamical cause of the direct product decomposition of Corollary 5.2.)

For each leaf of SmiS_{m_{i}}, we can consider its orbit in SmiS_{m_{i}} under the induced action of α\langle\alpha\rangle. Build a set FmiSmiF_{m_{i}}\subset S_{m_{i}} by taking one leaf from each such orbit. Thus, FmiF_{m_{i}} is a collection of rir_{i} leaves. Let XmiX_{m_{i}} denote the subset of n\mathfrak{C}_{n} underlying SmiS_{m_{i}} and let EmiE_{m_{i}} denote the subset of n\mathfrak{C}_{n} underlying the set FmiF_{m_{i}}. By construction we see that a subset of the fundamental domain of α\alpha is Dmi:=Xmi/αEmiD_{m_{i}}:=X_{m_{i}}/\langle\alpha\rangle\cong E_{m_{i}}.

We now analyze the groups Gmi=CVnXmi(α)=CVn(α)VnXmiG_{m_{i}}=C_{\,\,{V_{n}}_{X_{m_{i}}}}(\alpha)=C_{V_{n}}(\alpha)\cap{V_{n}}_{X_{m_{i}}}, which are individually isomorphic to the terms in the left-hand direct product of Theorem 1.1.

As described in Section 1.2, GmiG_{m_{i}} is an extension of its subgroup KmiK_{m_{i}} consisting of the elements in GmiG_{m_{i}} which have trivial induced action on DmiD_{m_{i}} and which act trivially outside of XmiX_{m_{i}}.

We now study KmiK_{m_{i}}. Let β\beta be an element of KmiK_{m_{i}}, and fix β\beta until we have finished our classification of KmiK_{m_{i}}.

By the previous two paragraphs we see that β\beta must carry each set underlying a leaf cycle in SmiS_{m_{i}} (under the action of α\langle\alpha\rangle) to itself. Consider a leaf λFmi\lambda\in F_{m_{i}}, and how β\beta moves points from the set underlying λ\lambda (fix this choice of leaf λ\lambda for the remainder of the discussion leading to the classification of KmiK_{m_{i}}). First, set λr=λαr\lambda_{r}=\lambda\alpha^{r}, the rr’th leaf in the orbit of λ\lambda under the action of α\langle\alpha\rangle on SmiS_{m_{i}}, for r{0,1,,mi1}r\in\{0,1,\ldots,m_{i}-1\}, and denote by τλ,r\tau_{\lambda,r} the set underlying λr\lambda_{r} in n\mathfrak{C}_{n}. If we take a point p0τλ,0p_{0}\in\tau_{\lambda,0}, its orbit under the action of α\langle\alpha\rangle is p0p_{0}, p1=p0αp_{1}=p_{0}\alpha, p2=p0α2p_{2}=p_{0}\alpha^{2}, and etc., so that prτλ,rp_{r}\in\tau_{\lambda,r}. In order for β\langle\beta\rangle to have no induced action on DmiD_{m_{i}}, we see that p0β=pcp_{0}\beta=p_{c} for some index cc. Now, in order to commute with the action of α\langle\alpha\rangle, we must have prβ=p((r+c)modmi)p_{r}\beta=p_{((r+c)\mod m_{i})} for any index 0rmi0\leq r\leq m_{i}. In particular, β\beta must push the full orbit of p0p_{0} under the action of α\langle\alpha\rangle forward by some constant index less than mim_{i}. A consequence of this is that any point pEmip\in E_{m_{i}} is itself in an orbit of length less than or equal to mim_{i} under the action of β\langle\beta\rangle. Extending this discussion as possible via recalling our choices of FmiF_{m_{i}} and λ\lambda, we see that β(mi!)\beta^{(m_{i}!)} must act trivially over the whole set XmiX_{m_{i}}, so β\beta must be torsion.

This now leads to the conclusion of our classification of KmiK_{m_{i}}. Suppose that (C,D,χ)(C,D,\chi) is a revealing pair for β\beta. Since β\beta is torsion, we see that C=DC=D. We assume (by taking a larger revealing pair to represent β\beta if necessary), that λ\lambda is a node of CC. Since α\alpha takes the set τλ,r\tau_{\lambda,r} to the set τλ,((r+1)modmi)\tau_{\lambda,((r+1)\mod{m_{i}})} in affine fashion, and β\beta commutes with α\alpha in such a way as to have no induced action on DmiD_{m_{i}}, it is straightforward to verify that the sub-tree Tλ,rT_{\lambda,r} in CC rooted at node λr\lambda_{r} is identical in shape to the sub-tree T((r+1)modmi)T_{((r+1)\mod{m_{i}})} in DD rooted at λ((r+1)modmi)\lambda_{((r+1)\mod{m_{i}})}, for all indices rr. Further, by the last sentence of the previous paragraph, for each leaf γ0\gamma_{0} of Tλ,0T_{\lambda,0} there is an integer tγ00t_{\gamma_{0}}\geq 0 such that β\beta will send the corresponding leaf γr\gamma_{r} of Tλ,rT_{\lambda,r} to the corresponding leaf γ((r+tγ0)modmi)\gamma_{((r+t_{\gamma_{0}})\mod m_{i})} in Tλ,((r+tγ0)modmi)T_{\lambda,((r+t_{\gamma_{0}})\mod m_{i})}. Any map nn\mathfrak{C}_{n}\to\mathfrak{C}_{n} which is the identity outside of XmiX_{m_{i}} and which satisfies these properties can be found in VnXmi{V_{n}}_{X_{m_{i}}}, and a straightforward topological argument (using the compactness of n\mathfrak{C}_{n}, and the basis cones of the topology on n\mathfrak{C}_{n}. See Subsection 2.1 for the definition) shows that these maps form a subgroup which is isomorphic to the group Maps(n,mi)Maps(\mathfrak{C}_{n},\mathbb{Z}_{m_{i}}). (That is, any continuous map from n\mathfrak{C}_{n} to mi\mathbb{Z}_{m_{i}} can be described as a rooted, finite, labeled nn-ary tree, where each label indicates where in mi\mathbb{Z}_{m_{i}} to send the set underlying the labeled leaf and this represents the offset in the orbit under α\langle\alpha\rangle for the element β\beta on the corresponding leaf.) Since the choice of map to mi\mathbb{Z}_{m_{i}} on the set τλ,0\tau_{\lambda,0} has no bearing (for the definition of β\beta) on the choice of map to mi\mathbb{Z}_{m_{i}} for the sets underlying the other leaves in FmiF_{m_{i}}, we see that

Kmi(Maps(n,mi))ri.K_{m_{i}}\cong(Maps(\mathfrak{C}_{n},\mathbb{Z}_{m_{i}}))^{r_{i}}.

We now need to consider the structure of Gmi/Kmi=QmiG_{m_{i}}/K_{m_{i}}=Q_{m_{i}}. Thus, we are modding out the subgroup of elements of VnXmi{V_{n}}_{X_{m_{i}}} which commute with the action of α\langle\alpha\rangle by the subgroup KmiK_{m_{i}}. In particular, we are looking at the elements of VnXmi{V_{n}}_{X_{m_{i}}} which carry, for each index 0rmi10\leq r\leq m_{i}-1, the sets underlying the rr’th copy of the fundamental domain DmiD_{m_{i}} to itself, where the map on the rr’th copy of the fundamental domain is precisely the conjugate version (under the action of αr\alpha^{r}) of the map on the 0’th copy EmiE_{m_{i}}. Therefore, the group QmiQ_{m_{i}} is isomorphically represented by the restriction of the action of VnV_{n} to the domain EmiE_{m_{i}}. In particular, QmiVnEmiQ_{m_{i}}\cong{V_{n}}_{E_{m_{i}}}. Since EmiE_{m_{i}} is given as the disjoint union of rir_{i} distinct copies of n\mathfrak{C}_{n}, we see that this group is precisely the finitely presented group Gn,riG_{n,r_{i}} of Higman in [14]! That is, we have the following.

QmiGn,riQ_{m_{i}}\cong G_{n,r_{i}}

Note that we can realize an isomorphic copy of QmiQ_{m_{i}} in GmiG_{m_{i}} as follows. Let β\beta^{\prime} be an element of VnEmi{V_{n}}_{E_{m_{i}}}, represented by the tree pair (C,D,χ)(C^{\prime},D^{\prime},\chi^{\prime}). The set SmiS_{m_{i}} decomposes as mim_{i} copies of the nodes of FmiF_{m_{i}} in the universal tree 𝒯n\mathscr{T}_{n} (take copy CrC_{r} as the nodes in the set FmiαrF_{m_{i}}\alpha^{r}, using the induced action of α\langle\alpha\rangle on the set of subsets of the nodes in SmiS_{m_{i}}, for each index 0r<mi0\leq r<m_{i}, and fix this definition of the sets CrC_{r} for the remainder of this subsection). There is a revealing pair (C′′,D′′,χ′′)β(C^{\prime\prime},D^{\prime\prime},\chi^{\prime\prime})\sim\beta^{\prime} expansion of (C,D,χ)(C^{\prime},D^{\prime},\chi^{\prime}) which has all of the nodes in the set SmiS_{m_{i}} as leaves of C′′C^{\prime\prime} and D′′D^{\prime\prime} (excepting the nodes in FmiF_{m_{i}}, which themselves are roots of a forest pair f=(d,r,θ)f=(\mathscr{F}_{d},\mathscr{F}_{r},\theta) representing the element of Gn,rkG_{n,r_{k}} corresponding to β\beta^{\prime}). By simply gluing a copy of ff to each CrC_{r} for r>0r>0 (the nodes in these CrC_{r} are leaves of C′′C^{\prime\prime} and D′′D^{\prime\prime}), we can build a new revealing tree pair (C,D,χ)(C,D,\chi) representing an element βVnXmi\beta\in{V_{n}}_{X_{m_{i}}} which acts on the set Γr\Gamma_{r} underlying CrC_{r} as β\beta^{\prime} acts on Γ0\Gamma_{0} under C0C_{0}, for each index rr (fix this definition of the sets Γr\Gamma_{r} for the remainder of the section as well). It is immediate by construction that the group Q^mi\widehat{Q}_{m_{i}} of elements β\beta so constructed is isomorphic with QmiQ_{m_{i}} and is a subgroup of GmiG_{m_{i}} which splits the short exact sequence KmiGmiQmiK_{m_{i}}\hookrightarrow G_{m_{i}}\twoheadrightarrow Q_{m_{i}}.

Thus, we have the following.

GmiKmiQmiKmiGn,riG_{m_{i}}\cong K_{m_{i}}\rtimes Q_{m_{i}}\cong K_{m_{i}}\rtimes G_{n,r_{i}}

We can complete our analysis of the centralizer of α\alpha over TαT_{\alpha} by showing the following lemma, which concludes the proof of Corollary 1.2 from the introduction.

Lemma 6.2

The group GmiG_{m_{i}} is finitely generated.

Proof:

First, we recall that for all positive integer values n>1n>1 and rr, Higman’s group Gn,rG_{n,r} is finitely presented (Theorem 4.6 of [14]). Let us denote by An,r|Rn,r\langle A_{n,r}\,|\,R_{n,r}\rangle a finite presentation of Gn,rG_{n,r} for any such nn and rr.

We first describe our set of generators for GmiG_{m_{i}}. For each generator gAn,rig^{\prime}\in A_{n,r_{i}}, we will take as a generator of GmiG_{m_{i}} the element gQ^mig\in\widehat{Q}_{m_{i}} which duplicates the effect of gg^{\prime} on Emi=Γ0E_{m_{i}}=\Gamma_{0} over each of the sets Γr\Gamma_{r} underlying the copies CrC_{r} of the leaves C0C_{0} over Γ0\Gamma_{0}, for each valid index rr.

At this stage, our collection of generators generates the group Q^mi\widehat{Q}_{m_{i}}, which is finitely presented still by carrying over the relations of Gn,riG_{n,r_{i}} as well in corresponding fashion. We need add only one further generator to generate the remainder of GmiG_{m_{i}}. Let g1Kmig_{1}\in K_{m_{i}} be the element represented by the revealing pair (S,T,θ)(S,T,\theta). We define (S,T,θ)(S,T,\theta) as follows. Let λ\lambda be a node of FmiF_{m_{i}} (fix this choice and dependent derived notation for the remainder of this subsection). Let S=TS=T be the minimal nn-ary tree so that λr=λαr\lambda_{r}=\lambda\alpha^{r} is a node which is a parent of nn-leaves of SS, for all index values 0r<mi0\leq r<m_{i}. Let θ\theta be the permutation from the leaves of SS to the leaves of TT which takes the first child of λr\lambda_{r} and sends it to the first child of λ((r+1)modmi)\lambda_{((r+1)\mod m_{i})}, for all indices 0r<mi0\leq r<m_{i}, and otherwise acts as the identity. The element g1g_{1} so constructed is our last generator.

We now show that g1g_{1}, together with our other generators, is sufficient to generate KmiK_{m_{i}}. Each element of KmiK_{m_{i}} decomposes as a finite product of sub-node translations along the mim_{i} orbit of FmiF_{m_{i}} in SmiS_{m_{i}} (under the action of α\langle\alpha\rangle). That is, we choose a descendant node pp of FmiF_{m_{i}} in the universal tree 𝒯n\mathscr{T}_{n}, and a translation constant 0tmi0\leq t\leq m_{i}. Then we translate the full orbit of pp under the action of α\langle\alpha\rangle forward cyclically by the constant tt, while acting as the identity elsewhere. We denote this translation as ptp_{t}. By choosing a specific nn-ary forest rooted at FmiF_{m_{i}} and translating each leaf of the forest in such a fashion, we can obtain any element of KmiK_{m_{i}}, as described above. Now, recall that Gn,riG_{n,r_{i}} acts transitively on the set of nodes in the infinite nn-ary forest descending from FmiF_{m_{i}} which do not happen to represent the full domain of Gn,riG_{n_{,}r_{i}} (if ri=1r_{i}=1, no element of Vn=Gn,1V_{n}=G_{n,1} can take a proper sub-node to the root node). Thus, given any particular descendant node pp from a node of FmiF_{m_{i}} and a translation distance tt, we can find an element ρ\rho of Q^mi\widehat{Q}_{m_{i}} taking the first descendant of λ0\lambda_{0} to pp. It is now immediate by construction that p1=g1ρp_{1}=g_{1}^{\rho}, and pt=p1tp_{t}=p_{1}^{t}. In particular, the set consisting of g1g_{1} and the generators of Q^mi\widehat{Q}_{m_{i}} together, is sufficient to generate GmiG_{m_{i}}.   \diamond

7 Centralizers over the set ZαZ_{\alpha}

Let us fix a revealing pair 𝔭=(A,B,σ)α\mathfrak{p}=(A,B,\sigma)\sim\alpha. Let {Γ1,Γ2,,Γe}\{\Gamma_{1},\Gamma_{2},\ldots,\Gamma_{e}\} represent the set of non-torsion flow graph components of the flow graph 𝔭,α\mathscr{F}_{\mathfrak{p},\alpha}, where for each index ii, we denote by XiX_{i} the component support of Γi\Gamma_{i}. Let αi\alpha_{i} represent the element in VnXi{V_{n}}_{X_{i}} such that αi|Xi=α|Xi\alpha_{i}|_{X_{i}}=\alpha|_{X_{i}}, and suppose further that 𝔭i=(Ai,Bi,σi)αi\mathfrak{p}_{i}=(A_{i},B_{i},\sigma_{i})\sim\alpha_{i} is a revealing tree pair that is identical to 𝔭\mathfrak{p} over the support XiX_{i} of γi\gamma_{i}, so that the flow graph 𝔭i,αi\mathscr{F}_{\mathfrak{p}_{i},\alpha_{i}} is identical to Γi\Gamma_{i}. Recall that by definition Zα=iXiZ_{\alpha}=\cup_{i}X_{i}. For any βVn\beta\in V_{n} set

𝒜β:=β𝒜β.\mathcal{RA}_{\beta}:=\mathcal{R}_{\beta}\sqcup\mathcal{A}_{\beta}.

Note that for any such β\beta, we have that 𝒜β\mathcal{RA}_{\beta} is a finite discrete set.

Lemma 7.1

Let gCVn(α)g\in C_{V_{n}}(\alpha).

  1. 1.

    The group g\langle g\rangle acts on the set 𝒜α\mathcal{RA}_{\alpha}.

  2. 2.

    Given rαr\in\mathcal{R}_{\alpha}, there is a basin of repulsion UrU_{r} of α\alpha containing rr so that UrgU_{r}g is contained in a basin of repulsion UsU_{s} for some repelling periodic point ss of α\alpha.

  3. 3.

    Given r𝒜αr\in\mathcal{A}_{\alpha}, there is a basin of attraction UrU_{r} of α\alpha containing rr so that UrgU_{r}g is contained in a basin of attraction UsU_{s} for some attracting periodic point ss of α\alpha.

  4. 4.

    The group g\langle g\rangle acts bijectively on each of the sets α\mathcal{R}_{\alpha} and 𝒜α\mathcal{A}_{\alpha} of repelling and attracting periodic points of α\alpha.

Proof: Let p𝒜αp\in\mathcal{RA}_{\alpha} have an orbit of size kk under the action of α\langle\alpha\rangle. We now have

pg=pαkg=pgαkpg=p\alpha^{k}g=pg\alpha^{k}

Thus, pgpg is fixed by αk\alpha^{k}, so that g\langle g\rangle bijectively preserves 𝒜α\mathcal{RA}_{\alpha}.

The second point follows from the continuity of gg and the following computation. Let rαr\in\mathcal{R}_{\alpha} so that rgrrg\neq r (if such a repelling periodic point fails to exist, we automatically have the second part of our lemma for the repelling periodic points). Choose U~rn\widetilde{U}_{r}\subseteq\mathfrak{C}_{n} an interval neighborhood of rr small enough so that gg is affine on U~r\widetilde{U}_{r} and so that U~r\widetilde{U}_{r} is contained in a basin of repulsion for α\alpha. Expand the revealing tree pair (A,B,σ)(A,B,\sigma) representing α\alpha by a rolling of type II to create a new revealing tree pair (A,B,σ)(A^{\prime},B^{\prime},\sigma^{\prime}) representing α\alpha with a complementary component CC rooted in some node whose underlying set is contained in U~r\widetilde{U}_{r}. The root node UrU_{r} of CC represents a basin of repulsion for α\alpha which is an interval neighborhood of rr carried affinely by gg to another interval of n\mathfrak{C}_{n}. Assume that rr is in a periodic orbit of length kk under the action of α\langle\alpha\rangle. Our result for repelling periodic points follows easily from the following limit:

(Ur)gα(nk)=(Ur)α(nk)grg (as n).(U_{r})g\alpha^{-(nk)}=(U_{r})\alpha^{-(nk)}g\to rg\textrm{\qquad(as }n\to\infty).

Hence gg takes a basin of repulsion neighborhood of rr into a neighborhood NN of some periodic repelling point s=rgs=rg of α\alpha with orbit length dividing kk where NN limits on ss under powers (nk)-(nk) of α\alpha.

A similar argument shows the third point of the lemma, and the final point of the lemma is an immediate consequence of the previous three points.

\diamond

The following corollary is immediate from the first point of the lemma above, together with the fact that α\alpha only admits finitely many periodic points.

Corollary 7.2

Let gCVnZα(α)g\in C_{{V_{n}}_{Z_{\alpha}}}(\alpha), then 𝒜α𝒜g\mathcal{RA}_{\alpha}\subset\mathcal{RA}_{g}.

The following corollary now follows from Lemma 7.1, using the idea behind the proof from Section 6 that a centralizer of an element of torsion must carry the set of all finite orbits of length kk (under the action of the torsion element) to itself.

Corollary 7.3

Let gCVn(α)g\in C_{V_{n}}(\alpha), and let rαr\in\mathcal{R}_{\alpha} or r𝒜αr\in\mathcal{A}_{\alpha}, with periodic orbit (ri=rαi)i=0k1(r_{i}=r\alpha^{i})_{i=0}^{k-1} in α\mathcal{R}_{\alpha} or 𝒜α\mathcal{A}_{\alpha} respectively, then there is a periodic orbit (si)i=0k1(s_{i})_{i=0}^{k-1} in α\mathcal{R}_{\alpha} or 𝒜α\mathcal{A}_{\alpha} respectively such that si=rigs_{i}=r_{i}g.

The next corollary depends on the proof of the second and third points of Lemma 7.1.

Corollary 7.4

Suppose gCVn(α)g\in C_{V_{n}}(\alpha) sends some point zXiz\in X_{i} to a point zgXjzg\in X_{j} for some indices ii and jj, then gg will send XiX_{i} bijectively to XjX_{j}.

Proof:

We first show that gg will send all of the periodic repelling and attracting points α\alpha within XiX_{i} into XjX_{j}.

Set Pi:=Xi𝒜αP_{i}:=X_{i}\cap\mathcal{RA}_{\alpha}.

If gg takes two points in PiP_{i} to the supports of distinct components of the flow graph of α\alpha, then XiX_{i} must admit a flow line which has some point r1r_{1} in its periodic repelling orbit vertex label sent to a periodic repelling point r1gr_{1}g of one component Γk\Gamma_{k} while having another point a1a_{1} in its periodic attracting orbit vertex label sent to a periodic attracting point for α\alpha in a label of a distinct component γm\gamma_{m}. But now, there are basins of repulsion U1U_{1} and attraction W1W_{1} around r1r_{1} and a1a_{1} respectively which are carried by gg into XkX_{k} and XmX_{m} respectively. This last is a contradiction, as follows.

Recall that some non-zero power kk of α\alpha fixes all of the repelling and attracting periodic points of α\alpha. Take p1U1\{r1}p_{1}\in U_{1}\backslash\{r_{1}\}. It must be that for all integers zz,

p1gαkz=p1αkzg,p_{1}g\alpha^{kz}=p_{1}\alpha^{kz}g,

however, for zz large and negative, p1αkzp_{1}\alpha^{kz} is near r1r_{1} while for zz large and positive, p1αkzp_{1}\alpha^{kz} is near a1a_{1}. In particular αk\alpha^{k} has a flow line connecting r1gr_{1}g to a1ga_{1}g, which is not possible by Lemma 4.7.

If p1p_{1} is in the support of a flow line, then by considering powers of α\alpha using similar arguments as above we can show the whole flow line is sent by gg to a single flow graph component of α\alpha.

\diamond

We now define the function

𝒮(g)=log2n1(rαrg).\mathcal{S}(g)=\log_{2n-1}\left(\prod_{r\in\mathcal{R}_{\alpha}}rg^{\prime}\right).

where rgrg^{\prime} denotes the slope of gg at the the repeller rr. This map is well-defined by using the recognition that gg is affine in small neighborhoods of points in Per(α)Per(\alpha). It is not too hard to see that the function 𝒮:Vn\mathcal{S}:V_{n}\to\mathbb{Z} is not a homomorphism in general.

Lemma 7.5

The map 𝒮:CVn(α)\mathcal{S}:C_{V_{n}}(\alpha)\to\mathbb{Z} is a group homomorphism.

Proof: Given g1,g2CVn(α)g_{1},g_{2}\in C_{V_{n}}(\alpha) we compute 𝒮(g1g2)\mathcal{S}(g_{1}g_{2}) directly from the definition. We note that in small interval neighborhoods of the points in α\mathcal{R}_{\alpha}, the maps gig_{i} are differentiable, and so we can apply the chain rule. Since gig_{i} acts bijectively on the set α\mathcal{R}_{\alpha} (by Corollary 7.3) each of the two terms of the product appears exactly once and so 𝒮(g1g2)=𝒮(g1)+𝒮(g2)\mathcal{S}(g_{1}g_{2})=\mathcal{S}(g_{1})+\mathcal{S}(g_{2})   \diamond


We will now shift attention to the local behavior of CVnXiC_{{V_{n}}_{X_{i}}} over the region XiX_{i} for any particular index i{1,2,,t}i\in\{1,2,\ldots,t\}. For each such index ii, set

𝒮i(g)=log2n1(rαirg).\mathcal{S}_{i}(g)=\log_{2n-1}\left(\prod_{r\in\mathcal{R}_{\alpha_{i}}}rg^{\prime}\right).

By the previous lemma, 𝒮i\mathcal{S}_{i} is a group homomorphism when we restrict its domain to either CVnXi(αi)C_{{V_{n}}_{X_{i}}}(\alpha_{i}) or even to CVn(α)C_{V_{n}}(\alpha), and further, if βiCVnXi(αi)\beta_{i}\in C_{{V_{n}}_{X_{i}}}(\alpha_{i}) while βCVn(α)\beta\in C_{V_{n}}(\alpha), with βi|Xi=β|Xi\beta_{i}|_{X_{i}}=\beta|_{X_{i}}, then Si(βi)=Si(β)S_{i}(\beta_{i})=S_{i}(\beta). For our immediate purposes below, we will use Si:CVnXi(αi)S_{i}:C_{{V_{n}}_{X_{i}}}(\alpha_{i})\to\mathbb{Z}.

The previous lemma now immediately implies the existence of the following exact sequence:

0ker(𝒮i)CVnXi(αi)im(𝒮i)=00\to\ker(\mathcal{S}_{i})\hookrightarrow C_{\,\,{V_{n}}_{X_{i}}}(\alpha_{i})\twoheadrightarrow\mathrm{im}(\mathcal{S}_{i})=\mathbb{Z}\to 0 (1)
Lemma 7.6

(Stair Algorithm) Let g1,g2CVnXi(αi)g_{1},g_{2}\in C_{{V_{n}}_{X_{i}}}(\alpha_{i}) and r,sr,s be periodic repelling points of αi\alpha_{i}, for some index ii. If

rg1=rg2=srg_{1}=rg_{2}=s

and the slope of g1g_{1} at rr is equal to the slope of g2g_{2} at rr , then g1=g2g_{1}=g_{2}.

Proof:

There is a basin of repulsion UrU_{r} of αi\alpha_{i} containing rr so that g1=g2g_{1}=g_{2} on UrU_{r}. Let xUr\{r}x\in U_{r}\backslash\{r\}. Since

xg1αin=xg2αin=xαing1=xαing2xg_{1}\alpha_{i}^{n}=xg_{2}\alpha_{i}^{n}=x\alpha_{i}^{n}g_{1}=x\alpha_{i}^{n}g_{2}

for all integers nn, we see that on the underlying support of any flow line LL of αi\alpha_{i} limiting on rr we have g1|L=g2|Lg_{1}|_{L}=g_{2}|_{L}. But this now means that g1g_{1} agrees with g2g_{2} on all of the underlying support of the edges of the graph Γi\Gamma_{i}. In particular, g1=g2g_{1}=g_{2} on sets limiting to each of the attracting orbits on the other ends of the flows lines of Γi\Gamma_{i} leading away from the repelling orbit of rr. Since g1g_{1} and g2g_{2} are always affine in small neighborhoods of the repelling and attracting periodic points for α\alpha, we then see that g1g_{1} and g2g_{2} actually agree on small neighborhoods of the attracting periodic orbits which appear on the terminal ends of those flow lines leaving the orbit containing rr.

We now repeat this argument moving away from the attracting orbits to new repelling orbits along new flow lines, where we again have that g1=g2g_{1}=g_{2} along these flow lines. Now by the connectivity of Γi\Gamma_{i}, g1=g2g_{1}=g_{2} over XiX_{i}.

\diamond

Lemma 7.7

There is an exact sequence

0CVnXi(αi)𝑞Q00\to\mathbb{Z}{\to}C_{\,\,{V_{n}}_{X_{i}}}(\alpha_{i})\overset{q}{\to}Q\to 0 (2)

where QSym(αi)Q\leq\mathrm{Sym}(\mathcal{R}_{\alpha_{i}}). In particular, CVnXi(αi)C_{\,\,{V_{n}}_{X_{i}}}(\alpha_{i}) is virtually infinite cyclic.

Proof: Define M:={βCVnXi(αi)β(r)=r,rαi}M:=\{\beta\in C_{\,\,{V_{n}}_{X_{i}}}(\alpha_{i})\mid\beta(r)=r,\,\,\forall r\in\mathcal{R}_{\alpha_{i}}\}. Fix r1αir_{1}\in\mathcal{R}_{\alpha_{i}} and define the following map

φ:Mβlog2n1(r1β).\begin{array}[]{cccc}\varphi:&M&\longrightarrow&\mathbb{Z}\\ &\beta&\longmapsto&\log_{2n-1}(r_{1}\beta^{\prime}).\end{array}

By Lemma 7.6, the map φ\varphi is injective and so MM\cong\mathbb{Z}. Now we observe that MM is the kernel of the action of CVnXi(αi)C_{\,\,{V_{n}}_{X_{i}}}(\alpha_{i}) on αi\mathcal{R}_{\alpha_{i}} and so we get a natural map q:CVnXi(αi)Qq:C_{\,\,{V_{n}}_{X_{i}}}(\alpha_{i})\to Q where Q:=CVnXi(αi)/MSym(αi)Q:=C_{\,\,{V_{n}}_{X_{i}}}(\alpha_{i})/M\leq\mathrm{Sym}(\mathcal{R}_{\alpha_{i}}).   \diamond

We recall a few elementary facts from group theory. The proofs are immediate, but we provide them for completeness.

Lemma 7.8

Let G,HG,H be groups such that the following sequence

0HG𝜔0.0\to H\to G\overset{\omega}{\to}\mathbb{Z}\to 0.

is exact. Then GHG\cong H\rtimes\mathbb{Z}.

Proof: Let zω1(1)z\in\omega^{-1}(1) then, by construction, zG\mathbb{Z}\cong\langle z\rangle\leq G and Hz={1G}H\cap\langle z\rangle=\{1_{G}\}. If gGg\in G and ω(g)=k\omega(g)=k, then zkg1Hz^{k}g^{-1}\in H, and so G=HzG=H\langle z\rangle.   \diamond


By Lemmas 7.5 and 7.8 applied on the exact sequence (1) we obtain that CVnXi(αi)ker(𝒮i)C_{\,\,{V_{n}}_{X_{i}}}(\alpha_{i})\cong\ker(\mathcal{S}_{i})\rtimes\mathbb{Z}. We will now show that ker(𝒮i)\ker(\mathcal{S}_{i}) is a finite group and that it coincides with the set of torsion elements of CVnXi(αi)C_{\,\,{V_{n}}_{X_{i}}}(\alpha_{i}).

Lemma 7.9

Let G,Q,K,CG,Q,K,C be groups, where C=C=\mathbb{Z} and QQ is a finite group of order mm. Assume the following two sequences

0C=𝜑G𝜓Q00\to C=\mathbb{Z}\overset{\varphi}{\to}G\overset{\psi}{\to}Q\to 0

and

0K𝜏G𝜎00\to K\overset{\tau}{\to}G\overset{\sigma}{\to}\mathbb{Z}\to 0

are exact. Then KK is a finite group and GKG\cong K\rtimes\mathbb{Z}.

Proof:

In this proof we write MfNM\leq_{f}N to denote that MM is a finite index subgroup of a group NN. Let hσ1(1)h\in\sigma^{-1}(1) and I:=hGI:=\langle h\rangle\leq G. By Lemma 7.8 we have GKIKG\cong K\rtimes I\cong K\rtimes\mathbb{Z}. We need to show that KK is a finite group. By assumption, CfGC\leq_{f}G and so we observe that

ICIICCGCQ,\frac{I}{C\cap I}\cong\frac{IC}{C}\leq\frac{G}{C}\cong Q,

therefore implying that CIfIC\cap I\leq_{f}I. In particular, CIC\cap I is a non-trivial group, hence CIfCC\cap I\leq_{f}C is too. By definition, for every gGg\in G, we have gmCg^{m}\in C. Since CIfCC\cap I\leq_{f}C, there is an integer kk such that gkmCIg^{km}\in C\cap I, for every gGg\in G.

If gKg\in K, we have that gkmKCIKI=0g^{km}\in K\cap C\cap I\leq K\cap I=0. Therefore, KK is a torsion subgroup of finite exponent, hence KC=0K\cap C=0 and so the first exact sequence implies that Kψ(K)QK\cong\psi(K)\leq Q and therefore it is finite.

\diamond


Applying Lemma 7.9 on the two exact sequences (1) and (2) we deduce the following result:

Corollary 7.10

Let i{1,2,,e}i\in\{1,2,\ldots,e\}. The centralizer CVnXi(αi)C_{\,\,{V_{n}}_{X_{i}}}(\alpha_{i}) is isomorphic to ker(𝒮i)\ker(\mathcal{S}_{i})\rtimes\mathbb{Z} and ker(𝒮i)\ker(\mathcal{S}_{i}) is a finite group.

The following is a fairly technical statement about roots which does not particularly assist us in our exploration of the centralizer of α\alpha, but which deserves to be stated. It represents a viewpoint on the underlying reason for the corollary on page 68 of [14]. Of course, a separate argument can also be given simply by noting how the lengths of spines of complementary components for representative revealing pairs change when one takes an element to powers.

Corollary 7.11

Suppose βVn\beta\in V_{n} is a non-torsion element with precisely one flow graph component Υ\Upsilon which has component support YY. Then the set of roots of β\beta in VnY{V_{n}}_{Y} is finite.

Proof: Consider the exact sequence from Lemma 7.8, applied to CVnY(β)C_{{V_{n}}_{Y}}(\beta). If βω1(k)\beta\in\omega^{-1}(k) for some positive integer kk, then all of the roots of β\beta in VnY{V_{n}}_{Y} occur in the sets ω1(j)\omega^{-1}(j) for integers jj which divide kk. Thus the roots all occur in a finite collection of finite sets.   \diamond

We note in passing that it may be the case that for all γω1(1)\gamma\in\omega^{-1}(1) we have that γk=τγβ\gamma^{k}=\tau_{\gamma}\cdot\beta, where τγ\tau_{\gamma} is a non-trivial torsion element for each such γ\gamma.

We now analyze the kernel of our 𝒮i\mathcal{S}_{i} homomorphism a bit further.

Lemma 7.12

Let i{1,2,,e}i\in\{1,2,\ldots,e\}. The kernel ker(𝒮i)={gCVnXi(αi)gk=id,for somek}\ker(\mathcal{S}_{i})=\{g\in C_{{V_{n}}_{X_{i}}}(\alpha_{i})\mid g^{k}=id,\;\;\mbox{for some}\;\;k\in\mathbb{Z}\}, that is, the kernel of 𝒮i\mathcal{S}_{i} is the set of torsion elements in CVnXi(αi)C_{{V_{n}}_{X_{i}}}(\alpha_{i}).

Proof: Let TT be the set of torsion elements in CVnXi(αi)C_{{V_{n}}_{X_{i}}}(\alpha_{i}). A priori, this may not be a subgroup.

We observe that Tker(𝒮i)T\subseteq\ker(\mathcal{S}_{i}), because the slope of gTg\in T multiplies to one across the full cycle of every periodic orbit of neutral leaves of gg, and the orbits of the repelling periodic orbits of αi\alpha_{i} are carried to each other by the action of any g\langle g\rangle for any gCVnXi(αi)g\in C_{{V_{n}}_{X_{i}}}(\alpha_{i}). So, for gTg\in T, we have by the definition of 𝒮i\mathcal{S}_{i} that 𝒮i(g)=0\mathcal{S}_{i}(g)=0.

By the previous Corollary, ker(𝒮)\ker(\mathcal{S}) is a finite group, hence ker(𝒮)T\ker(\mathcal{S})\subseteq T.   \diamond


At this juncture, we have pushed our analysis of the centralizer of an element of VnV_{n} with a discrete train track (or flow graph) with one component, which represents a non-torsion component, as far as necessary for us to be able to support the structure described in the right hand product of our Theorem 1.1.

Proof of the right hand product structure of Theorem 1.1:

We now partition the non-torsion flow graph components {Γi}\{\Gamma_{i}\} by the rule ΓiΓj\Gamma_{i}\sim\Gamma_{j} if there exists fCVn(α)f\in C_{V_{n}}(\alpha) such that (Xi)f=Xj(X_{i})f=X_{j}. In this case we note that αif=αj\alpha_{i}^{f}=\alpha_{j} and therefore CVXi(αi)CVXj(αj)C_{V_{X_{i}}}(\alpha_{i})\cong C_{V_{X_{j}}}(\alpha_{j}). In particular, we have that ker(𝒮i)ker(𝒮j)\ker(\mathcal{S}_{i})\cong\ker(\mathcal{S}_{j}).

Let us call such supports (XX_{*} which can be carried to each other by an fCVn(α)f\in C_{V_{n}}(\alpha)) supports of isomorphic connected components. Let {ICC1,ICC2,,ICCt}\{ICC_{1},ICC_{2},\ldots,ICC_{t}\} be the set of \sim-equivalence classes of isomorphic (flow graph) connected components (each of which is denoted by ICCjICC_{j}, for some jj), where qjq_{j} is the cardinality of the set ICCjICC_{j}. For each jj, order the members of ICCjICC_{j} and let Γi,j\Gamma_{i,j} be the ii-th member of ICCjICC_{j}. We re-index the Xi,αi,𝒮iX_{i},\,\alpha_{i},\,\mathcal{S}_{i} using the new double-index notation in corresponding fashion. Let us fix conjugators γ1,k,j\gamma_{1,k,j} in CVn(α)C_{V_{n}}(\alpha) carrying X1,jX_{1,j} to Xk,jX_{k,j} (acting as the identity elsewhere), and use these to generate a full permutation group PqjP_{q_{j}} which acts on the underlying sets X,jX_{*,j} of the elements in ICCjICC_{j}. Set Aj:=ker(𝒮1,j)A_{j}:=\ker(\mathcal{S}_{1,j}) and note that ker(𝒮i,j)=Ajγ1,i,j\ker(\mathcal{S}_{i,j})=A_{j}^{\gamma_{1,i,j}}.

For fixed jj, we then have the centralizer of α\alpha over the set iXi,j\cup_{i}X_{i,j} consists of any self centralization on each individual Xi,jX_{i,j} (this group will be congruent to Bi,j:=ker(𝒮i,j)B_{i,j}:=\ker(\mathcal{S}_{i,j})\rtimes\mathbb{Z}), together with a product by any permutation of these components. Thus, this group is

Gj(i=1qjBi,j)Pqj=(Aj)Pqj.G_{j}\cong\left(\prod_{i=1}^{q_{j}}B_{i,j}\right)\rtimes P_{q_{j}}=\left(A_{j}\rtimes\mathbb{Z}\right)\wr P_{q_{j}}.

(note that the product i=1qjBi,j\prod_{i=1}^{q_{j}}B_{i,j} is normal in GjG_{j}). Now as non-isomorphic non-torsion components cannot be mapped onto each other by the action of ff, we see that the action of ff on the non-torsion components of α\alpha is describable as an element from the direct product of the individual groups GjG_{j}. Since tt is the number of isomorphism classes {ICCj}\{ICC_{j}\}, we obtain our statement of Theorem 1.1.

\diamond

Below is a revealing pair (A,B,σ)(A,B,\sigma) representing element α\alpha of V2V_{2}, which has centralizer congruent to (2×2)×(\mathbb{Z}_{2}\times\mathbb{Z}_{2})\times\mathbb{Z}; the group corresponding to A1A_{1} is the Klein 44-group. This example is included to answer a question of Nathan Barker which arose in conversation.

\psfrag{1}[c]{$\,\,1$}\psfrag{2}[c]{$\,\,2$}\psfrag{3}[c]{$\,\,3$}\psfrag{4}[c]{$\,\,4$}\psfrag{5}[c]{$\,\,5$}\psfrag{6}[c]{$\,\,6$}\psfrag{7}[c]{$\,\,7$}\psfrag{8}[c]{$\,\,8$}\psfrag{9}[c]{$\,\,9$}\psfrag{10}[c]{$\,\,10$}\psfrag{11}[c]{$\,\,11$}\psfrag{12}[c]{$\,\,12$}\psfrag{13}[c]{$\,\,13$}\psfrag{14}[c]{$\,\,14$}\psfrag{15}[c]{$\,\,15$}\psfrag{16}[c]{$\,\,16$}\psfrag{17}[c]{$\,\,17$}\psfrag{18}[c]{$\,\,18$}\psfrag{19}[c]{$\,\,19$}\psfrag{20}[c]{$\,\,20$}\psfrag{\alpha}[c]{$\alpha$}\includegraphics[height=150.0pt,width=300.0pt]{KleinCentral.eps}

Below, we sketch the flow graph associated with (A,B,σ)(A,B,\sigma). We do not label the flow lines with their iterated augmentation chains, as that will only serve to clutter the essential aspects of this graph. The three different types of flow lines for this revealing pair are encoded by the different methods used in drawing the directed edges (solid, versus two distinct flavors of dashing).

\psfrag{0}[c]{$Y$}\psfrag{1}[c]{$X$}\includegraphics[height=150.0pt,width=300.0pt]{Klein_Flow.eps}

In this graph, the repelling fixed points label the four vertices to the left and right, while the attracting fixed points are the four vertices in the middle.

Graph connectivity and the constraint on preserving flow-line types provide enough information to guarantee that if τ\tau is a non-trivial torsion element commuting with α\alpha, then τ\tau is completely determined by which repelling fixed point one chooses to send 0001¯000\overline{1} to. All such elements are order two, and the elements β\beta and γ\gamma below generate the Klein 44-group KK so determined by the orbit dynamics on the repelling fixed points. Thus, A1KA_{1}\leq K.

\psfrag{1}[c]{$\,\,1$}\psfrag{2}[c]{$\,\,2$}\psfrag{3}[c]{$\,\,3$}\psfrag{4}[c]{$\,\,4$}\psfrag{\beta}[c]{$\beta$}\includegraphics[height=100.0pt,width=200.0pt]{beta.eps}\psfrag{1}[c]{$\,\,1$}\psfrag{2}[c]{$\,\,2$}\psfrag{3}[c]{$\,\,3$}\psfrag{4}[c]{$\,\,4$}\psfrag{\beta}[c]{$\beta$}\psfrag{1}[c]{$\,\,1$}\psfrag{2}[c]{$\,\,2$}\psfrag{3}[c]{$\,\,3$}\psfrag{4}[c]{$\,\,4$}\psfrag{5}[c]{$\,\,5$}\psfrag{6}[c]{$\,\,6$}\psfrag{7}[c]{$\,\,7$}\psfrag{8}[c]{$\,\,8$}\includegraphics[height=100.0pt,width=200.0pt]{Gamma.eps}

Since β\beta and γ\gamma can be realized as a group of permutations of the leaves of the common tree ABA\cap B which are in the repeller and attractor iterated augmentation chains, while preserving the orbit dynamics of α\langle\alpha\rangle over 𝒜α\mathcal{RA}_{\alpha}, we see that every element in KK commutes with α\alpha. Note that as α\alpha has no roots, the generator of the \mathbb{Z} factor is α\alpha, so A1A1×A_{1}\rtimes\mathbb{Z}\cong A_{1}\times\mathbb{Z}, in this example.

8 Cyclic subgroup (non)distortion

We assume the reader is familiar with distortion of subgroups in a group. We use definitions consistent with their usage in [10].

The calculation of the (non)distortion of the cyclic subgroups in VnV_{n} is very close is spirit to the calculations in [7, 24, 8]. However, by using properties of revealing pairs, we remove many of the technical obstructions usually associated with counting carets and this shortens the arguments in those papers.

We believe that technology such as the revealing pair technology should generally simplify proofs of non-distortion of cyclic subgroups in the various families of generalized Thompson’s groups following arguments similar to that which is given below. It would be interesting to see a general such tool developed for these families of groups.

We are indebted to M. Kassabov for a discussion of the spirit of this type of argument.

Proof of Theorem 1.3:

Suppose αVn\alpha\in V_{n} and that α\langle\alpha\rangle\cong\mathbb{Z}. Suppose further that X={x1,x2,,xq}X=\{x_{1},x_{2},\ldots,x_{q}\} is a finite generating set of VnV_{n} which is closed under inverses.

There is a minimal positive integer PXP_{X} so that for all pnp\in\mathfrak{C}_{n}, the slope of each xjx_{j} in small neighborhoods of pp is (2n1)sp(2n-1)^{s_{p}} where |sp|<PX|s_{p}|<P_{X}.

Suppose (A,B,σ)α(A,B,\sigma)\sim\alpha, and let rr be a repeller for this tree pair. Suppose further the iterated augmentation chain of rr is r=r0{r}={r}_{0}, ri=r0αi{r}_{i}={r_{0}}{\alpha}^{i} for 0iu0\leq i\leq u (so that ru{r}_{u} is the root of the complementary component Cr{C_{r}} of AB{A}-{B} containing the repeller r{r}). Let Γr{\Gamma_{r}} be the spine of Cr{C_{r}}, and suppose the length of Γr{\Gamma_{r}} is Lr{L_{r}}.

For each index 0<i<u0<i<u, there is a jump J(r,i)J_{(r,i)} in depth in the infinite binary tree, from the depth of ri{r}_{i} to ri+1{r}_{i+1}. Further, set J(r,0)J_{(r,0)} to be the jump from the depth of ru{r}_{u} to the depth of r1{r}_{1}. (E.g., if ru{r}_{u} has depth 44, and r1{r}_{1} has depth 33, then we set J(r,0)J_{(r,0)} to be 11.) Given a positive integer zz, set

S(r,z)=Σe=1(zmodu)J(r,(e1)),S_{(r,z)}=\Sigma_{e=1}^{(z\!\!\!\!\mod u)}J_{(r,(e-1))},

the partial sum of the first zmoduz\!\!\mod u jumps. (This sum may be negative.) Note that the sum of all uu jumps is zero, so that

Σe=1zJ(r,(e1)modu)=S(r,z).\Sigma_{e=1}^{z}J_{(r,\,(e-1)\!\!\!\!\mod u)}=S_{(r,z)}.

Now fix a particular positive integer zz. Set w=z/uw=\lfloor z/u\rfloor, the largest positive integer less than or equal to z/uz/u. Now, if yi{y}_{i} is the repelling periodic point under the leaf ri{r}_{i} for 0i<u0\leq i<u, direct calculation shows that the slope of αz\alpha^{z} at y0{y}_{0} is

((2n1)Lr)w+1(2n1)S(r,z).((2n-1)^{{L_{r}}})^{w+1}\cdot(2n-1)^{S_{(r,z)}}.

By the chain rule, we require at least ((Lrw+Lr+S(r,z))/PX\lceil((L_{r}\cdot w+L_{r}+S_{(r,z)})/P_{X}\rceil generators from XX to create the element αz{\alpha}^{z} (where \lceil\cdot\rceil denotes the largest positive integer greater than or equal to its argument). This last function can be interpreted as a function in zz that is bounded below by an affine function g:g:\mathbb{N}\to\mathbb{Q} where g(z)=(1/W)z+Og(z)=(1/W)\cdot z+O with positive slope 1/W<(Lr/(uPX))1/W<(L_{r}/(u\cdot P_{X})) for some integer WW and vertical offset OO\in\mathbb{Z} (for technical reasons, choose OO so that g(0)<0g(0)<0). Now the function fpos=g1f_{pos}=g^{-1} (restricted and co-restricted to \mathbb{N}) is an affine distortion function for the positive powers of α\alpha in α\langle\alpha\rangle within VnV_{n} (αz\alpha^{z} has minimal length zz when expressed as an element in α\langle\alpha\rangle using the generating set {α,α1}\{\alpha,\alpha^{-1}\}, and by construction, z<fpos(m)z<f_{pos}(m), where mm is the minimal word length of αz\alpha^{z} as expressed in XX).

A similar argument produces an affine distortion function fneg:f_{neg}:\mathbb{N}\to\mathbb{N} for the negative powers of α\alpha. Thus, f:=fpos+fnegf:=f_{pos}+f_{neg} will be an affine distortion function for the whole of α\langle\alpha\rangle in VnV_{n}. In particular, α\langle\alpha\rangle is undistorted in VnV_{n}.

\diamond

References

  • [1] Tuna Altinel and Alexey Muranov, Interprétation de l’arithmétique dans certains groupes de permutations affines par morceaux d’un intervalle, J. Inst. Math. Jussieu 8 (2009), no. 4, 623–652.
  • [2] N. Barker, The simultaneous conjugacy problem in Thompson’s group VV, Ph.D. thesis, University of Newcastle Upon Tyne, 2011.
  • [3] J.M. Belk and F. Matucci, Conjugacy and Dynamics in Thompson’s groups, submitted, arXiv:math.GR/0708.4250v1.
  • [4] C. Bleak and O. P. Salazar-Díaz, Free products in R. Thompson’s group VV, submitted.
  • [5] Matthew G. Brin, Higher dimensional Thompson groups, Geom. Dedicata 108 (2004), 163–192.
  • [6] Matthew G. Brin and Craig C. Squier, Presentations, conjugacy, roots, and centralizers in groups of piecewise linear homeomorphisms of the real line, Comm. Algebra 29 (2001), no. 10, 4557–4596.
  • [7] José Burillo, Quasi-isometrically embedded subgroups of Thompson’s group FF, J. Algebra 212 (1999), no. 1, 65–78. MR MR1670622 (1999m:20051)
  • [8] José Burillo, Sean Cleary, Melanie Stein, and Jennifer Taback, Combinatorial and metric properties of Thompson’s group TT, Trans. Amer. Math. Soc. 361 (2009), no. 2, 631–652.
  • [9] J. W. Cannon, W. J. Floyd, and W. R. Parry, Introductory notes on Richard Thompson’s groups, Enseign. Math. (2) 42 (1996), no. 3-4, 215–256.
  • [10] Benson Farb, The extrinsic geometry of subgroups and the generalized word problem, Proc. London Math. Soc. s3–68 (1994), no. 3, 577–593.
  • [11] É Ghys and V. Sergiescu, Sur un groupe remarquable de difféomorphismes du cercle, Comment. Math. Helv. 62 (1987), 185–239.
  • [12] V. S. Guba and M. V. Sapir, On subgroups of the R. Thompson group FF and other diagram groups, Mat. Sb. 190 (1999), no. 8, 3–60. MR MR1725439 (2001m:20045)
  • [13] Victor Guba and Mark Sapir, Diagram groups, Mem. Amer. Math. Soc. 130 (1997), no. 620, viii+117. MR MR1396957 (98f:20013)
  • [14] Graham Higman, Finitely presented infinite simple groups, Department of Pure Mathematics, Department of Mathematics, I.A.S. Australian National University, Canberra, 1974, Notes on Pure Mathematics, No. 8 (1974).
  • [15] Martin Kassabov and Francesco Matucci, The Simultaneous Conjugacy Problem in Groups of Piecewise-Linear Functions, to appear in Groups, Geometry and Dynamics, arXiv:math.GR/0607167v3
  • [16] C. Martínez-Pérez and B. E. A. Nucinkis, Bredon cohomological finiteness conditions for generalisations of Thompson’s groups, submitted, arXiv:math.GR/1105.0189v3
  • [17] F. Matucci, Algorithms and classification in groups of piecewise-linear homeomorphisms, Ph.D. thesis, Cornell University, 2008, arXiv:math.GR/0807.2871v1.
  • [18] R. Penner and J. Harer, Combinatorics of train tracks, Annals of Mathematics Studies, vol. 125, Princeton University Press, 31 William Street, Princeton, New Jersey 08540, 1992.
  • [19] Stephen J. Pride, Geometric methods in combinatorial semigroup theory, Semigroups, formal languages and groups (York, 1993), NATO Adv. Sci. Inst. Ser. C Math. Phys. Sci., vol. 466, Kluwer Acad. Publ., Dordrecht, 1995, pp. 215–232. MR MR1630622 (99d:20095)
  • [20]   , Low-dimensional homotopy theory for monoids, Internat. J. Algebra Comput. 5 (1995), no. 6, 631–649. MR MR1365195 (96m:57006)
  • [21] Olga Salazar-Díaz, Thompson’s group VV from a dynamical viewpoint, Internat. J. Algebra Comput. 20 (2010), no. 1, 39–70.
  • [22] W. P. Thurston, On the geometry and dynamics of diffeomorphisms of surfaces, Bull. Amer. Math. Soc. 19 (1988), no. 2, 417–431.
  • [23]   , Three-dimensional geometry and topology, Princeton Mathematical Series, vol. 35, Princeton University Press, 31 William Street, Princeton, New Jersey 08540, 1997.
  • [24] C. Wladis, Cyclic subgroups are quasi-isometrically embedded in the Thompson-Stein groups, submitted.