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

The dynamics and geometry of free group endomorphisms

Jean Pierre Mutanguha Email:[email protected], Web address:https://mutanguha.com
Department of Mathematical Sciences, University of Arkansas, Fayetteville, AR
Abstract

We prove that ascending HNN extensions of free groups are word-hyperbolic if and only if they have no Baumslag-Solitar subgroups. This extends the theorem of Brinkmann that free-by-cyclic groups are word-hyperbolic if and only if they have no free abelian subgroups of rank 2. The paper is split into two independent parts:

1) We study the dynamics of injective nonsurjective endomorphisms of free groups. We prove a canonical structure theorem that initializes the development of improved relative train tracks for endomorphisms; this structure theorem is of independent interest since it makes many open questions about injective endomorphisms tractable.

2) As an application of the structure theorem, we are able to (relatively) combine Brinkmann’s theorem with our previous work and obtain the main result stated above. In the final section, we further extend the result to HNN extensions of free groups over free factors.

Overview

Word-hyperbolic groups are groups that act geometrically (properly and cocompactly) on proper δ\delta-hyperbolic spaces and these spaces are the coarse equivalents of negatively curved geometric spaces. Word-hyperbolic groups are an important class of groups introduced by Misha Gromov [14] and one of the fundamental problems in geometric group theory is determining necessary and sufficient algebraic conditions for a group to be word-hyperbolic.

Question (Coarse hyperbolization [1, Question 1.1]).

Let GG be a group of finite type. If GG contains no Baumslag-Solitar subgroups, then must it be word-hyperbolic?

Finite type can be thought of as a strengthening of finite presentation for torsion-free groups. Baumslag-Solitar groups are two-generator one-relator groups with the presentation BS(m,n)=a,t|t1amt=anBS(m,n)=\langle a,t~{}|~{}t^{-1}a^{m}t=a^{n}\rangle for m,n0m,n\neq 0. It is now a classical fact that a word-hyperbolic group cannot have subgroups isomorphic to Baumslag-Solitar groups. The question, historically attributed to Gromov, asks if Baumslag-Solitar subgroups are the only essential obstruction to word-hyperbolicity for finite type groups. It has been answered in the affirmative for certain classes of groups: the fundamental groups of closed 33-manifolds (Thurston [30, 31], Perelman) and free-by-cyclic groups (Brinkmann [8]). The main result of this paper extends the latter to a larger class of HNN extensions.

Main Theorem.

Let AFA\leq F be a free factor of a finite rank free group FF and ϕ:AF\phi:A\to F be an injective homomorphism. The HNN extension FAF*_{A} of FF over AA and ϕ\phi is word-hyperbolic if and only if it contains no BS(1,n)BS(1,n) subgroups for n1n\geq 1.

The HNN extension has the presentation FA=F,t|t1at=ϕ(a),aAF*_{A}=\langle F,t~{}|~{}t^{-1}at=\phi(a),~{}\forall a\in A\rangle. If A=F=ϕ(A)A=F=\phi(A), then FA=FϕF*_{A}=F\rtimes_{\phi}\mathbb{Z} is a free-by-cyclic group and this case was proven by Peter Brinkmann. Our proof of the main theorem above uses Brinkmann’s result. If A=Fϕ(A)A=F\neq\phi(A), then FA=FϕF*_{A}=F*_{\phi} is a strictly ascending HNN extension and coarse hyperbolization for this class of groups was our original motivation. Finally, when AFA\neq F, we more or less reduce this to the case A=FA=F.

The case A=FA=F of the main theorem (Theorem 6.7) proved to be difficult since it required an understanding of the dynamics of injective nonsurjective endomorphisms of free groups and no such complete study had been carried out yet. Patrick Reynolds studied the dynamics of irreducible nonsurjective endomorphisms of free groups [27] and we previously used this work to prove an instance of the main theorem where ϕ:FF\phi:F\to F is irreducible but not surjective [23, 24]. In this case, we showed that the strictly ascending HNN extension FϕF*_{\phi} is always word-hyperbolic. Unfortunately, it remained unclear how Reynolds’ work could be generalized to all injective nonsurjective endomorphisms and no further progress on the problem had been made.

The first part of this paper (Sections 14) is a self-contained systematic study of all injective nonsurjective endomorphisms of free groups and is the main novel contribution of the paper. We hereby present a summary of the important results from these sections. Note that the statements given here are not the complete statements of the cited results.

Summary.

If ϕ:FF\phi:F\to F is injective but not surjective, then there is:

  1. 1.

    a unique maximal proper [ϕ][\phi]-fixed free factor system 𝒜\mathcal{A}; (Proposition 2.3)

  2. 2.

    a unique minimal proper [ϕ][\phi]-invariant free factor system 𝒜\mathcal{A}^{*} that carries 𝒜\mathcal{A} and is fixed under backward iteration, i.e., ϕ1𝒜=𝒜\phi^{-1}\cdot\mathcal{A}^{*}=\mathcal{A}^{*}; (Proposition 2.4)

  3. 3.

    a unique expanding 𝒜\mathcal{A}^{*}-relative immersion for ϕ\phi. (Corollary 4.7)

The free factor systems 𝒜\mathcal{A} and 𝒜\mathcal{A}^{*} could be trivial.

This means any injective nonsurjective endomorphism ϕ:FF\phi:F\to F is induced by a graph map f:ΓΓf:\Gamma\to\Gamma such that: 1) some possibly empty or disconnected proper subgraph ΓΓ\Gamma^{\prime}\subset\Gamma is ff-invariant and the restriction of ff to Γ\Gamma^{\prime} is a homotopy equivalence; 2) some ff-invariant proper subgraph Γ′′Γ\Gamma^{\prime\prime}\supset\Gamma^{\prime} has an fnf^{n}-image contained in Γ\Gamma^{\prime} for some n1n\geq 1; and 3) collapsing the lifts of the subgraph Γ′′\Gamma^{\prime\prime} in the universal cover Γ~\tilde{\Gamma} induces an expanding immersion f¯:TT\bar{f}:T\to T on a simplicial FF-tree. In particular, if Γ\Gamma^{\prime} is empty, then ff is an expanding graph immersion.

In our previous work [23], we extended a result of Ilya Kapovich [17] and proved the special case of the main theorem with A=FA=F and ϕ:FF\phi:F\to F induced by an expanding graph immersion. The expanding immersion was crucial to the proof and so we developed expanding relative immersions with the intent of, in a sense, relativizing our previous proof and extending it to all injective nonsurjective endomorphisms. This application of the structure theorem is carried out in the second part of the paper (Sections 57).

Expanding relative immersions are interesting in their own right as they can potentially be applied to many currently open problems about nonsurjective endomorphisms. To name a few: relative immersions may be needed to extend Hagen-Wise’s cubulation of word-hyperbolic free-by-cyclic groups [16, 15] to the groups FAF*_{A}; we suspect that they could be used to extend the main theorem to a characterization of the possible Dehn/isoperimetric functions of the groups FAF*_{A} following work by Bridson-Groves [7] (See also Problem 2); they provide a framework for generalizing Bestvina-Feighn-Handel’s construction of improved relative train tracks for automorphisms [5], which were used in Brinkmann’s, Hagen-Wise’s, and Bridson-Groves’ results; for a different direction of generalization, the main theorem may be extended to HNN extensions GAG*_{A} of torsion-free word-hyperbolic groups GG over free factors AGA\leq G; and finally, the virtual fibering question is a particularly interesting problem that was first posed to us by Dawid Kielak and our structure theorem along with the cubulation problem could be significant steps towards its resolution:

Question (Virtual fibering).

Suppose ϕ:FF\phi:F\to F is injective but not surjective. If FϕF*_{\phi} is word-hyperbolic, then does it have a free-by-cyclic finite index subgroup? What if it has a quadratic Dehn function?

Besides Section 7, the results in this paper were the author’s doctoral thesis. With future applications of the structure theorem in mind, we have chosen to emphasize in this paper the independence of the “theory” (first part) and “application” (second part). The two parts can be read independently. If you choose to skip the first part, then we encourage you to read the prologue and interlude; these optional sections contain several examples that capture the key points. The main results of the first part are summarized again in the interlude with a bit more detail since we are assuming you will, by then, be familiar with the terms defined in the definitions and conventions section. The summary is all you will need to use expanding relative immersions in your own work. The epilogue is a brief discussion of some questions about endomorphisms in the context of expanding relative immersions.

The main tools used in the first part are Bass-Serre theory, Stallings graphs, bounded cancellation, and train track theory. The second part uses some Bass-Serre theory and the Bestvina-Feighn combination theorem.

Acknowledgments: I am greatly indebted to my advisor Matt Clay. I encountered all the tools used in this paper in our first directed readings, before I even knew what geometric group theory was.

Prologue

We will start with a few motivating examples to illustrate the main construction of the first part of this paper. Let F=F(a,b)F=F(a,b) be the free group on two generators and ϕ,φ,ψ:FF\phi,\varphi,\psi:F\to F be injective nonsurjective endomorphisms given by

ϕ:(a,b)(ab,ba),φ:(a,b)(a,bab1),andψ:(a,b)(a,abab).\phi:(a,b)\mapsto(ab,ba),\quad~{}\varphi:(a,b)\mapsto(a,bab^{-1}),\quad\text{and}\quad\psi:(a,b)\mapsto(a,abab).

The standard rose is a rose RR with two petals and an identification F=π1(R)F=\pi_{1}(R) such that the basis {a,b}\{a,b\} corresponds to the petals; it will be graphically represented by a rose with two oriented petals labelled by aa and bb respectively.

Refer to caption
Figure 1: The standard rose.

For each integer k1k\geq 1, let R^k\hat{R}_{k} be the cover of RR corresponding to the subgroup ϕk(F)\phi^{k}(F), i.e., it is the quotient of the universal cover R~\tilde{R} by the action of the subgroup ϕk(F)\phi^{k}(F). The Stallings graph S(ϕk(F))S(\phi^{k}(F)) is the core of R^k\hat{R}_{k}, i.e., the smallest connected subgraph of R^k\hat{R}_{k} with rank 2. Alternatively, we can define R~(ϕk(F))R~\tilde{R}(\phi^{k}(F))\subset\tilde{R} to be the smallest subtree invariant under the ϕk(F)\phi^{k}(F)-action and S(ϕk(F))S(\phi^{k}(F)) to be the quotient ϕk(F)\R~(ϕk(F))\phi^{k}(F)\backslash\tilde{R}(\phi^{k}(F)). Evidently, there is a natural isomorphism ϕk(F)π1(S(ϕk(F)))\phi^{k}(F)\cong\pi_{1}(S(\phi^{k}(F))). A Stallings graph SS is graphically represented as an RR-digraph, i.e., SS will be an oriented graph whose oriented edges are labelled by aa or bb as shown in Figure 2.

Refer to caption
Figure 2: Stallings graphs S(ϕk(F))S(\phi^{k}(F)) with respect to the standard rose for k=1,2,3k=1,2,3.

The Stallings graphs S(ϕk(F))S(\phi^{k}(F)) are roses for all k1k\geq 1 and each petal doubles in size with each iteration. Since ϕ\phi is injective, it is an isomorphism onto its image. In particular, we may use ϕ\phi to get an isomorphism Fϕk(F)π1(S(ϕk(F)))F\cong\phi^{k}(F)\cong\pi_{1}(S(\phi^{k}(F))) for k1k\geq 1. Under this isomorphism, the basis {a,b}\{a,b\} corresponds to the petals of S(ϕk(F))S(\phi^{k}(F)) and we recover the standard rose for all k1k\geq 1. This is equivalent to the existence of an immersion (locally injective map) on the rose f:RRf:R\to R such that π1(f)=ϕ\pi_{1}(f)=\phi [24, Lemma 3.2]. Indeed, the obvious map f:RRf:R\to R that maps the aa-petal to the path abab and the bb-petal to the path baba is an immersion. The petals of S(ϕk(F))S(\phi^{k}(F)) doubling in size with each iteration implies ff is in fact an expanding immersion, i.e., all edges expand under ff-iteration.

Refer to caption
Figure 3: Stallings graphs S(φk(F))S(\varphi^{k}(F)) for k=1,2,3k=1,2,3.

The Stallings graphs S(φk(F))S(\varphi^{k}(F)) are barbells for all k1k\geq 1, the middle bars roughly double in size with each iteration, and the plates are labelled by aa (See Figure 3). Under the isomorphism Fφk(F)π1(S(φk(F)))F\cong\varphi^{k}(F)\cong\pi_{1}(S(\varphi^{k}(F))), the conjugacy classes {[a],[b]}\{[a],[b]\} correspond to the two plates of the barbell S(φk(F))S(\varphi^{k}(F)) and we recover the so called standard barbell for all k1k\geq 1. Just as in the first example, this is equivalent to the existence of an immersion on the barbell g:BBg:B\to B such that π1(g)=φ\pi_{1}(g)=\varphi under some identification F=π1(B)F=\pi_{1}(B). However, the immersion is not expanding since the plates are not growing in size with each iteration.

What we are now about to do will be the crux of the first part of the paper. Since the (nongrowing) plates of S(φk(F))S(\varphi^{k}(F)) are all labelled aa, we deduce that a\langle a\rangle is a free factor fixed by φ\varphi and b\langle b\rangle is mapped to a conjugate of a\langle a\rangle. Thus, {a,b}\{\langle a\rangle,\langle b\rangle\} forms an invariant free factor system that contains a fixed free factor. Now consider g~:B~B~\tilde{g}:\tilde{B}\to\tilde{B} to be the lift of gg to the universal cover B~\tilde{B}. Collapsing all translates of axes of aa and bb in B~\tilde{B}, i.e., collapsing all edges of B~\tilde{B} labelled by the plates of BB, will produce a so-called (F,{a,b})(F,\{\langle a\rangle,\langle b\rangle\})-tree TT with a nonfree FF-action and g~\tilde{g} induces an expanding immersion g¯:TT\bar{g}:T\to T where each edge doubles in size under g¯\bar{g}-iteration. The map g¯\bar{g} will be referred to as a relative immersion.

Refer to caption
Figure 4: Stallings graphs S(ψk(F))S(\psi^{k}(F)) for k=1,2k=1,2.

For the final example, the Stallings graphs S(ψk(F))S(\psi^{k}(F)) are roses again where one petal roughly doubles in size with each iteration and another is labelled by aa (See Figure 4). Unlike the first example, the roses with the isomorphisms Fψk(F)π1(S(ψk(F)))F\cong\psi^{k}(F)\cong\pi_{1}(S(\psi^{k}(F))) are not standard roses since bb does not correspond to a petal for k1k\geq 1. In fact, applying ψ\psi to the labels of S(ψk(F))S(\psi^{k}(F)) forces a full fold with one petal to get S(ψk+1(F))S(\psi^{k+1}(F)) which means the Stallings graphs are all distinct marked roses. In particular, ψ\psi cannot be induced by a graph immersion! But the observation that the nongrowing petals of S(ψk(F))S(\psi^{k}(F)) are all labelled by aa implies a\langle a\rangle is a fixed free factor. Despite ψ\psi not being induced by an immersion, the obvious map on the standard rose h:RRh:R\to R will induce an expanding immersion h¯:YY\bar{h}:Y\to Y on an (F,{a})(F,\{\langle a\rangle\})-tree YY where every edge doubles in size.

The main result of the first part of this paper is that this construction always produces expanding relative immersions, Theorem 4.5: any injective nonsurjective endomorphism of a free group is induced by a graph map f:ΓΓf:\Gamma\to\Gamma such that collapsing the translates of axes in Γ~\tilde{\Gamma} of some canonical invariant free factor system containing a fixed free factor system will induce an expanding relative immersion f¯:TT\bar{f}:T\to T. This can be summarized into two steps: first establishing the existence of a unique (possibly trivial) maximal fixed free factor system (Section 2); then showing that collapsing this free factor system and its preimages in the universal cover of an appropriately chosen graph Γ\Gamma will produce a tree TT on which we can define an expanding relative immersion (Section 4). The two guiding principles will be bounded cancellation (Lemmas 1.4 & 3.1) and the fact that Stallings graphs for iterated injective nonsurjective endomorphisms have unbounded size (Lemma 1.1 & Proposition 3.9).

Definitions and conventions

FF will always be a free group with finite rank at least 22. A nontrivial subgroup system of FF is a nonempty finite collection of nontrivial subgroups 𝒜={A1,,Al}\mathcal{A}=\{A_{1},\ldots,A_{l}\} of FF. A nontrivial free factor system of FF is a nonempty collection of nontrivial free factors 𝒜={A1,,Al}\mathcal{A}=\{A_{1},\ldots,A_{l}\} of FF such that AiAjA_{i}\cap A_{j} is trivial if iji\neq j and A1,,Al\langle A_{1},\ldots,A_{l}\rangle is free factor of FF. We define the trivial system to be the collection consisting of the trivial subgroup; this will allow us to treat absolute and relative cases simultaneously in our proofs. For any subgroup system 𝒜\mathcal{A}, the subgroups Ai𝒜A_{i}\in\mathcal{A} are called the components of 𝒜\boldsymbol{\mathcal{A}} or 𝓐\boldsymbol{\mathcal{A}}-components. A subgroup system 𝒜\mathcal{A} is finitely generated if its components are finitely generated. A free factor system 𝒜\mathcal{A} is proper if some component is proper, i.e., 𝒜{F}\mathcal{A}\neq\{F\}. Let 𝒜\mathcal{A} and \mathcal{B} be subgroup systems of FF. We shall say 𝓐\boldsymbol{\mathcal{A}} carries \boldsymbol{\mathcal{B}} if for every component BB\in\mathcal{B}, there is a component A𝒜A\in\mathcal{A} and element xFx\in F such that BxAx1B\leq xAx^{-1} and, when 𝒜\mathcal{A} and \mathcal{B} are free factor systems, we shall denoted it by 𝒜\mathcal{B}\preceq\mathcal{A}. For a conjugacy class of elements [g][g] in FF, we shall also say 𝓐\boldsymbol{\mathcal{A}} carries [g]\boldsymbol{[g]} if there is a component A𝒜A\in\mathcal{A} and element xFx\in F such that gxAx1g\in xAx^{-1}. If 𝒜={A}\mathcal{A}=\{A\} (or ={B}\mathcal{B}=\{B\}) is a singleton, then we will write, “AA carries [g]/[g]\,/\,\mathcal{B} (or BB).” One can easily verify that the \preceq-relation on free factor systems is a preorder, i.e., it is reflexive and transitive. So it determines an equivalence relation on the set of free factor systems and a partial order on the set of equivalence classes. Free factor systems will always be considered up to this equivalence relation. In particular, we can replace components in a system with conjugates whenever convenient.

Remark.

Suppose F=a,bF=\langle a,b\rangle be the free group of rank 2. With our definition, {a,b}\{\langle a\rangle,\langle b\rangle\} and {bab1,(ba)b(ba)1}\{\langle\,bab^{-1}\,\rangle,\langle\,(ba)b(ba)^{-1}\,\rangle\} are equivalent free factor systems but {a,(ba)b(ba)1}\{\langle a\rangle,\langle\,(ba)b(ba)^{-1}\,\rangle\} is not a free factor system. Thus, when we replace components in a free factor system with conjugates, we need to ensure the resulting subgroup system is still a free factor system.

Two group homomorphisms h1,h2:ABh_{1},h_{2}:A\to B are equivalent if there is an inner automorphism ib:BBi_{b}:B\to B such that h2=ibh1h_{2}=i_{b}\circ h_{1} and this is denoted by [h1]=[h2][h_{1}]=[h_{2}]. Outer endomorphisms of FF are the equivalence classes on the set of endomorphisms of FF. For instance, properties such as being irreducible are not just properties of an endomorphism ϕ\phi but also its outer class [ϕ][\phi].

Let ϕ:FF\phi:F\to F be an endomorphism. We say a subgroup system 𝒜={A1,,Al}\mathcal{A}=\{A_{1},\ldots,A_{l}\} is [ϕ]\boldsymbol{[\phi]}-invariant if 𝒜\mathcal{A} carries the subgroup system ϕ(𝒜)={ϕ(A1),,ϕ(Al)}\phi(\mathcal{A})=\{\phi(A_{1}),\ldots,\phi(A_{l})\}, i.e, there exists a set of elements {x1,,xl}F\{x_{1},\ldots,x_{l}\}\subset F and a function σ:{1,,l}{1,,l}\sigma:\{1,\ldots,l\}\to\{1,\ldots,l\} such that ϕ(Ai)xiAσ(i)xi1\phi(A_{i})\leq x_{i}A_{\sigma(i)}x_{i}^{-1} for all ii. A [ϕ][\phi]-invariant subgroup system 𝒜\mathcal{A} is [ϕ]\boldsymbol{[\phi]}-fixed if its subgroups are permuted up to conjugacy, i.e., σ\sigma is a permutation and [ϕ(Ai)]=[Aσ(i)][\phi(A_{i})]=[A_{\sigma(i)}] for all ii. When ϕ\phi is an automorphism, all [ϕ][\phi]-invariant free factor systems are [ϕ][\phi]-fixed. When ϕ\phi is injective, then, for any k1k\geq 1 and free factor system 𝒜\mathcal{A} of FF, ϕk(𝒜)\phi^{k}(\mathcal{A}) is a free factor system of ϕk(F)\phi^{k}(F); conversely, any free factor system of ϕk(F)\phi^{k}(F) gives us a free factor system of FF via the isomorphism ϕk:Fϕk(F)\phi^{k}:F\to\phi^{k}(F). Finally, note that when 𝒜\mathcal{A} is a [ϕ][\phi]-fixed free factor system but ϕ\phi is not surjective, ϕk(𝒜)\phi^{k}(\mathcal{A}) will typically not be a free factor system of FF (See the remark above). A subgroup AFA\leq F is eventually [ϕ]\boldsymbol{[\phi]}-periodic if [ϕm(A)]=[ϕn(A)][\phi^{m}(A)]=[\phi^{n}(A)] for some m>n1m>n\geq 1, and it is [ϕ]\boldsymbol{[\phi]}-periodic if [ϕm(A)]=[A][\phi^{m}(A)]=[A] for some m1m\geq 1.

The endomorphism ϕ\phi is reducible if it has a nontrivial proper invariant free factor system and irreducible otherwise. One immediate consequence of Stallings folds [29] is the injectivity of irreducible endomorphisms (Observation below). So we can drop the injectivity hypothesis when specializing results to the irreducible case.

For the topological perspective, graphs are 1-dimensional CW-complexes and a graph map f:XXf:X\to X^{\prime} will be a continuous map of graphs that sends vertices to vertices and any edge to a vertex or immersed path. An edge ee of XX is pretrivial if f(e)f(e) is a vertex. For a graph map f:ΓΓf:\Gamma\to\Gamma^{\prime} of finite graphs, let KK be the maximum of the (combinatorial) length of the edge-path f(e)f(e) as ee varies over all the edges of Γ\Gamma. Then ff is KK-Lipschitz, a fact that will be used throughout the paper. Generally, K(f)K(f) will denote some convenient Lipschitz constant for ff rather than the infimum. For the moment, XX is used for arbitrary graphs but, for most of the paper, Γ\Gamma will be used for finite connected noncontractible graphs and TT for infinite simply-connected graphs (simplicial trees). A core graph is a graph with no proper deformation retract.

A direction at a vertex vXv\in X is a half-edge attached to the vertex. A vertex is bivalent if it has exactly two directions. The set of directions at vertex vv is denoted by TvXT_{v}X. branch points are vertices with at least three directions and natural edges are maximal edge-paths whose interior vertices are bivalent. We will say a graph map is natural if it maps branch points to branch points and any natural edge to a branch point or immersed path. If the graph map f:XXf:X\to X^{\prime} has no pretrivial edges, then the restriction to initial segments induces the derivative map at vv, dfv:TvXTf(v)Xdf_{v}:T_{v}X\to T_{f(v)}X^{\prime}. The graph map ff is an immersion if it is locally injective, i.e., it has no pretrivial edges and the derivative maps dfvdf_{v} are injective for all vertices vv; note that immersions are natural. An immersion ff is expanding if the length of fn(e)f^{n}(e) is unbounded as nn\to\infty for every edge ee of XX.

With immersions defined, we preface the observation with a summary of Stallings’ folding theorem. Let RR be a rose whose edges are indexed by a basis {a1,,an}\{a_{1},\ldots,a_{n}\} of FF and let f:RRf:R\to R be the map where f(ai)f(a_{i}) is the immersed edge-path in RR labelled by ϕ(ai)\phi(a_{i}). Stallings [29] showed that ff factors as ιflf1\iota\circ f_{l}\circ\cdots\circ f_{1} where each fif_{i} is a fold and ι\iota is an immersion. We will use this factorization again to prove bounded cancellation in Lemma 1.4 and Proposition 3.1.

Observation.

If ϕ:FF\phi:F\to F is irreducible, then it is injective.

Proof.

If ϕ\phi is not injective, then at least one of the folds in Stallings’ factorization of ϕ\phi collapses a subgraph of the domain. In particular, the kernel of ϕ\phi contains a proper free factor AFA\leq F; therefore, ϕ(A)={1}A\phi(A)=\{1\}\leq A and [ϕ][\phi] is reducible. ∎

Let 𝒜={A1,,Al}\mathcal{A}=\{A_{1},\ldots,A_{l}\} be a nontrivial subgroup system of FF, then an 𝓐\boldsymbol{\mathcal{A}}-marked graph (Γ,α)(\Gamma_{*},\alpha_{*}) is a collection of graphs Γ={Γ1,,Γl}\Gamma_{*}=\{\Gamma_{1},\ldots,\Gamma_{l}\} where the finite connected core graphs Γi\Gamma_{i} are indexed by markings, i.e., isomorphisms αi:Aiπ1(Γi)\alpha_{i}:A_{i}\to\pi_{1}(\Gamma_{i}). When 𝒜={A}\mathcal{A}=\{A\} is a singleton, we may write AA-marked graph in place of 𝒜\mathcal{A}-marked graph. A marked graph (Γ,α)(\Gamma,\alpha) is an FF-marked graph. For any marked graph (Γ,α)(\Gamma,\alpha) and any conjugacy class of a nontrivial element gFg\in F, denoted by [g][g], we define its length with respect to α\boldsymbol{\alpha}, gα\lVert g\rVert_{\alpha}, to be the (combinatorial) length of the immersed loop in Γ\Gamma representing [g][g].

More generally, we want to consider pairs 𝒜\mathcal{A}\preceq\mathcal{B} of free factor systems of FF. For each component BiB_{i}\in\mathcal{B}, let 𝒜i\mathcal{A}_{i} be either the nonempty maximal subset of 𝒜\mathcal{A} carried by BiB_{i} or the trivial system if no such subset exists. Typically, we shall replace components of 𝒜\mathcal{A} with conjugates so that 𝒜\mathcal{A} is also a free factor system of \mathcal{B}, i.e., each 𝒜i\mathcal{A}_{i} is a free factor system of BiB_{i} (next remark below). A (𝓑,𝓐)\boldsymbol{(\mathcal{B},\mathcal{A})}-forest TT_{*} is a simplicial forest of (Bi,𝒜i)(B_{i},\mathcal{A}_{i})-trees TiT_{i}, i.e., a collection of simplicial trees T={T1,,Tk}T_{*}=\{T_{1},\ldots,T_{k}\} where each tree TiT_{i} is equipped with a minimal simplicial BiB_{i}-action whose edge stabilizers are trivial and point stabilizers are trivial or conjugates (in BiB_{i}) of 𝒜i\mathcal{A}_{i}-components. We note that a (B,{B})(B,\{B\})-tree is a point also known as a degenerate tree.

When 𝒜\mathcal{A} is the trivial system, an (F,𝒜)(F,\mathcal{A})-tree is a free minimal FF-tree TT. In that case, the quotient F\TF\backslash T is a marked graph. When 𝒜\mathcal{A} is a nontrivial proper free factor system, the quotient of an (F,𝒜)(F,\mathcal{A})-tree is a graph of groups decomposition of FF with trivial edge groups and 𝒜\mathcal{A} as the nontrivial vertex groups [28]; such decompositions are sometimes known as free splittings of F\boldsymbol{F} and they will be our relative analogues for marked graphs. Any given (F,𝒜)(F,\mathcal{A})-tree endowed with the combinatorial metric has an associated length function lT:Fl_{T}:F\to\mathbb{R}. Precisely, any isometry of a simplicial tree is either elliptic (fixes a point) or loxodromic (preserves an axis of least displacement). If gFg\in F is elliptic, then lT(g)=0l_{T}(g)=0; otherwise, lT(g)>0l_{T}(g)>0 is the translation distance of gg acting on its axis. When 𝒜\mathcal{A} is trivial and F\TF\backslash T is the marked graph (Γ,α)(\Gamma,\alpha), then lT()=αl_{T}(\cdot)=\|\cdot\|_{\alpha} as functions FF\to\mathbb{R}. The minimal subtree T(H)TT(H)\subset T of a nontrivial subgroup HFH\leq F is the smallest subtree with a minimal induced HH-action, i.e., the union of all fixed points and axes of nontrivial elements in HH.

Remark.

Let 𝒜\mathcal{A}\preceq\mathcal{B} be free factor systems of FF, BiB_{i}\in\mathcal{B} be a free factor with nontrivial 𝒜i\mathcal{A}_{i}, and TT be an (F,𝒜)(F,\mathcal{A})-tree. Choose a connected fundamental domain of BiB_{i} acting on T(Bi)T(B_{i}) and the nontrivial stabilizers in BiB_{i} of vertices in the domain form a free factor system 𝒜i\mathcal{A}_{i}^{\prime} of BiB_{i}. Then 𝒜i\mathcal{A}_{i}^{\prime} and 𝒜i\mathcal{A}_{i} are equivalent as free factor systems of FF since 𝒜i{Bi}\mathcal{A}_{i}\preceq\{B_{i}\}.

Suppose ψ:FF\psi:F\to F^{\prime} is an injective homomorphism. Then it determines a contravariant preimage function ψ1\psi^{-1} from the poset of free factor systems of FF^{\prime} to the poset of free factor systems of FF. One way to define this function is through trees. Let 𝒜\mathcal{A}^{\prime} be a free factor system of FF^{\prime}, TT be any (F,𝒜)(F^{\prime},\mathcal{A}^{\prime})-tree, and T(ψ(F))TT(\psi(F))\subset T be the minimal subtree of ψ(F)F\psi(F)\leq F^{\prime}. In particular, the quotient ψ(F)\T(ψ(F))\psi(F)\backslash T(\psi(F)) is a free splitting of ψ(F)\psi(F) whose nontrivial vertex groups form a nontrivial free factor system 𝒜′′\mathcal{A}^{\prime\prime} of ψ(F)\psi(F). If T(ψ(F))T(\psi(F)) has a free ψ(F)\psi(F)-action, then this condition is independent of the choice of (F,𝒜)(F^{\prime},\mathcal{A}^{\prime})-tree TT and we set ψ1𝒜\psi^{-1}\cdot\mathcal{A}^{\prime} to be the trivial system. Otherwise, we have a nontrivial free factor system 𝒜′′\mathcal{A}^{\prime\prime} of ψ(F)\psi(F) and a corresponding nontrivial free factor system of FF via the isomorphism ψ:Fψ(F)\psi:F\to\psi(F); the equivalence class of the latter free factor system is independent of the (F,𝒜)(F^{\prime},\mathcal{A}^{\prime})-tree TT and we will denote it by ψ1𝒜\psi^{-1}\cdot\mathcal{A}^{\prime}.

Let TT and TT^{\prime} be (F,𝒜)(F,\mathcal{A})- and (F,𝒜)(F^{\prime},\mathcal{A}^{\prime})-trees respectively and ψ:FF\psi:F\to F^{\prime} be an injective homomorphism such that 𝒜\mathcal{A}^{\prime} carries ψ(𝒜)\psi(\mathcal{A}) or, equivalently, ψ1𝒜𝒜\psi^{-1}\cdot\mathcal{A}^{\prime}\succeq\mathcal{A}. This carrying condition ensures that elliptic elements in FF (with respect to TT) have elliptic ψ\psi-images (with respect to TT^{\prime}). A tree map f:TTf:T\to T^{\prime} is 𝝍\boldsymbol{\psi}-equivariant if f(gx)=ψ(g)f(x)f(g\cdot x)=\psi(g)\cdot f(x) for all gFg\in F and xTx\in T. If we require ψ1𝒜=𝒜\psi^{-1}\cdot\mathcal{A}^{\prime}=\mathcal{A}, then loxodromic elements in FF have loxodromic ψ\psi-images; this is the relative analogue of a π1\pi_{1}-injective graph map.

Suppose 𝒜\mathcal{A} is a nontrivial [ϕ][\phi]-invariant free factor system of FF for some injective endomorphism ϕ:FF\phi:F\to F; so there is a function σ:{1,,l}{1,,l}\sigma:\{1,\ldots,l\}\to\{1,\ldots,l\} and inner automorphisms ixi:FFi_{x_{i}}:F\to F such that the restrictions ϕi=(ixiϕ)|Ai\phi_{i}=\left.\left(i_{x_{i}}\circ\phi\right)\right|_{A_{i}} are homomorphisms ϕi:AiAσ(i)\phi_{i}:A_{i}\to A_{\sigma(i)}. The collection {ϕi}\{\phi_{i}\} is a restriction of ϕ\boldsymbol{\phi} to 𝒜\boldsymbol{\mathcal{A}}, which we denote by ϕ|𝒜:𝒜𝒜\left.\phi\right|_{\mathcal{A}}:\mathcal{A}\to\mathcal{A}. A weak representative for [ϕi]\boldsymbol{[\phi_{i}]} is a graph map fi:ΓiΓσ(i)f_{i}:\Gamma_{i}\to\Gamma_{\sigma(i)} from an AiA_{i}\,-marked graph (Γi,αi)(\Gamma_{i},\alpha_{i}) to an Aσ(i)A_{\sigma(i)}\,-marked graph (Γσ(i),ασ(i))(\Gamma_{\sigma(i)},\alpha_{\sigma(i)}) such that [π1(fi)αi]=[ασ(i)ϕi][\pi_{1}(f_{i})\circ\alpha_{i}]=[\alpha_{\sigma(i)}\circ\phi_{i}] for all i{1,,l}i\in\{1,\ldots,l\}. A weak representative for [ϕ|𝒜]\boldsymbol{\left[\left.\phi\right|_{\mathcal{A}}\,\right]} is a graph map f:ΓΓf_{*}:\Gamma_{*}\to\Gamma_{*} of an 𝒜\mathcal{A}-marked graph (Γ,α)(\Gamma_{*},\alpha_{*}) that is a disjoint union of weak representatives for [ϕi][\phi_{i}]. A natural representative is a weak representative that is also natural. A topological representative is a weak representative that has no pretrivial edges and whose underlying graph Γ\Gamma_{*} has no bivalent vertices.

Recall 𝒜\mathcal{A}\preceq\mathcal{B}. Suppose ϕ1𝒜=𝒜\phi^{-1}\cdot\mathcal{A}=\mathcal{A}, \mathcal{B} is [ϕ][\phi]-invariant, and let σ:{1,k}{1,k}\sigma^{\prime}:\{1,\ldots k\}\to\{1,\ldots k\} be the function corresponding to the [ϕ][\phi]-invariance of \mathcal{B} and ϕ|={ϕi:BiBσ(i)}\left.\phi\right|_{\mathcal{B}}=\{\phi_{i}:B_{i}\to B_{\sigma^{\prime}(i)}\} be a restriction of ϕ\phi. An 𝓐\boldsymbol{\mathcal{A}}-relative weak representative for ϕ|\boldsymbol{\left.\phi\right|_{\mathcal{B}}} is a forest map f:TTf_{*}:T_{*}\to T_{*} of a (,𝒜)(\mathcal{B},\mathcal{A})-forest TT_{*} that is a disjoint union of ϕi\phi_{i}-equivariant tree maps fi:TiTσ(i)f_{i}:T_{i}\to T_{\sigma^{\prime}(i)}. An 𝓐\boldsymbol{\mathcal{A}}-relative natural representative is an 𝒜{\mathcal{A}}-relative weak representative that is also natural. An 𝓐\boldsymbol{\mathcal{A}}-relative representative is an 𝒜{\mathcal{A}}-relative weak representative with no pretrivial edges and whose underlying forest has no bivalent vertices. A relative representative is minimal if it has no orbit-closed invariant subforests whose components are bounded. An (expanding resp.) 𝒜\boldsymbol{\mathcal{A}}-relative immersion for ϕ|\boldsymbol{\left.\phi\right|_{\mathcal{B}}} is an 𝒜\mathcal{A}-relative representative for ϕ|\left.\phi\right|_{\mathcal{B}} that is also an (expanding resp.) immersion.

In the first part of the paper, we show that any injective endomorphism of FF has a canonical invariant free factor system 𝒜\mathcal{A} and an expanding 𝒜\mathcal{A}-relative immersion ff such that 𝒜\mathcal{A} is eventually mapped into a canonical fixed free factor system. In the second part of the paper, we use the canonical systems and the expanding relative immersion to prove the main theorem.

Dynamics of free group endomorphisms

1 Stallings graphs and bounded cancellation

Let (Γ,α)(\Gamma,\alpha) be a marked graph with no bivalent vertices. For any nontrivial subgroup system \mathcal{H} of FF, the Stallings (subgroup) graph for \mathcal{H} with respect to a marked graph (Γ,α)(\Gamma,\alpha) with is the smallest \mathcal{H}-marked graph (S(),β)(S(\mathcal{H}),\beta_{*}) along with an immersion ι:S()Γ\iota_{*}:S(\mathcal{H})\to\Gamma such that [π1(ιi)βi]=[α|Hi][\pi_{1}(\iota_{i})\circ\beta_{i}]=\left[\left.\alpha\right|_{H_{i}}\right] for every HiH_{i}\in\mathcal{H}. Alternatively, S()S(\mathcal{H}) is the collection of cores S(Hi)S(H_{i}) of the covers Γ^Hi\hat{\Gamma}_{H_{i}} of Γ\Gamma corresponding to α(Hi)\alpha(H_{i}), i.e., the smallest deformation retract of Γ^Hi\hat{\Gamma}_{H_{i}}, and ιi\iota_{i} is the restriction to S(Hi)S(H_{i}) of the covering map Γ^HiΓ\hat{\Gamma}_{H_{i}}\to\Gamma. We may sometimes refer to the marked graph (Γ,α)(\Gamma,\alpha) as the ambient graph. In the notation, the marking and immersion for Stallings graphs will usually be omitted. If HH and HH^{\prime} are in the same conjugacy class, [H][H], then there is a homeomorphism h:S(H)S(H)h:S(H)\to S(H^{\prime}) such that ι=ιh\iota=\iota^{\prime}\circ h. The converse holds as well. So the Stallings graph S[H]S[H] is uniquely determined by the conjugacy class [H][H]. Furthermore, the Stallings graph S[]S[\mathcal{H}] is a finite graph if and only if \mathcal{H} is finitely generated. Suppose ϕ:FF\phi:F\to F is injective, 𝒜\mathcal{A} is a nontrivial free factor system of FF. We will be studying the (iterated) Stallings graphs (S[ϕk(𝒜)],β,k)(S[\phi^{k}(\mathcal{A})],\beta_{*,k}) for k1k\geq 1.

Remark.

For any free factor system \mathcal{F} of FF and nontrivial subgroup system \mathcal{H} carried by \mathcal{F}, we can similarly define the Stallings graph for \mathcal{H} with respect to \mathcal{F}-marked graphs (Γ,α)(\Gamma_{*},\alpha_{*}). If ϕ:\phi_{*}:\mathcal{F}\to\mathcal{F} is an injective endomorphism, i.e., a collection of injective endomorphism {ϕi:FiFσ(i)}\{\phi_{i}:F_{i}\to F_{\sigma(i)}\} that need not be a restriction of an endomorphism of FF, then we can still consider the Stallings graphs S[ϕk(𝒜)]S[\phi^{k}_{*}(\mathcal{A})] for k1k\geq 1 and 𝒜\mathcal{A}\preceq\mathcal{F}. The point of this remark is that the results of this paper hold when ϕ:FF\phi:F\to F is replaced with ϕ:\phi_{*}:\mathcal{F}\to\mathcal{F}. In other words, the ambient graph Γ\Gamma need not be connected for our results.

Unlike the ambient graph Γ\Gamma, we allow Stallings graphs S=S[]S=S[\mathcal{H}] to have bivalent vertices. More precisely, we subdivide SS so that the immersion ι:SΓ\iota_{*}:S\to\Gamma is simplicial, i.e., maps edges to edges. With this subdivision, we get a combinatorial metric on (S,β)(S,\beta_{*}) that is compatible with (Γ,α)(\Gamma,\alpha), i.e., for any nontrivial element gg in HiH_{i}, gα=gβi\lVert g\rVert_{\alpha}=\lVert g\rVert_{\beta_{i}}.

Lemma 1.1.

Let ϕ:FF\phi:F\to F be injective and HH be a finitely generated nontrivial subgroup of FF. If HH is not eventually [ϕ][\phi]-periodic, then the length of the longest natural edge in S[ϕk(H)]S[\phi^{k}(H)] is unbounded as kk\to\infty.

Proof.

Suppose the length of the longest natural edge in S[ϕk(H)]S[\phi^{k}(H)] with respect to some marked graph (Γ,α)(\Gamma,\alpha) was uniformly bounded as kk\to\infty. We want to show that HH is eventually [ϕ][\phi]-periodic. The number of natural edges in S[ϕk(H)]S[\phi^{k}(H)] is bounded above by 3rank(H)33\cdot\mathrm{rank}(H)-3. Our assumption implies there is a bound on the volume of (number of edges in) the graphs S[ϕk(H)]S[\phi^{k}(H)] as kk\to\infty. So the sequence S[ϕk(H)]S[\phi^{k}(H)] is eventually periodic, i.e., there are integers m>n1m>n\geq 1 and an isometry h:S[ϕm(H)]S[ϕn(H)]h:S[\phi^{m}(H)]\to S[\phi^{n}(H)] such that ιm=ιnh\iota_{m}=\iota_{n}\circ h. Since a Stallings graph determines the conjugacy class of its defining subgroup, we have [ϕm(H)]=[ϕn(H)][\phi^{m}(H)]=[\phi^{n}(H)], i.e., HH is eventually [ϕ][\phi]-periodic. ∎

This lemma will be invoked on invariant free factor systems containing a component that is not eventually [ϕ][\phi]-periodic. Conversely, the next lemma handles the case when an invariant free factor system consists entirely of eventually periodic free factors.

Lemma 1.2.

Let ϕ:FF\phi:F\to F be injective and 𝒜\mathcal{A} be a nontrivial [ϕ][\phi]-invariant free factor system. If all components in 𝒜\mathcal{A} are eventually [ϕ][\phi]-periodic, then some nonempty subset 𝒜\mathcal{B}\subset\mathcal{A} is a [ϕ][\phi]-fixed free factor system and ϕk(𝒜)\phi^{k}(\mathcal{A}) is carried by \mathcal{B} for some k0k\geq 0.

E.g., if ϕ:FF\phi:F\to F is injective and FF is eventually [ϕ][\phi]-periodic, then ϕ\phi is an automorphism.

Proof.

Let σ:{1,,l}{1,,l}\sigma:\{1,\ldots,l\}\to\{1,\ldots,l\} be the function used to define the [ϕ][\phi]-invariance of 𝒜={A1,,Al}\mathcal{A}=\{A_{1},\ldots,A_{l}\}. Then there is a nonempty subset J{1,,l}J\subset\{1,\ldots,l\} on which σ\sigma acts as a bijection and σl({1,,l})=J\sigma^{l}(\{1,\ldots,l\})=J. Let 𝒜\mathcal{B}\preceq\mathcal{A} by the nontrivial [ϕ][\phi]-invariant free factor system corresponding to JJ. Then ϕl(𝒜)\phi^{l}(\mathcal{A}) is carried by \mathcal{B} since the image of σl{\sigma\,}^{l} is JJ. It remains to show that \mathcal{B} is [ϕ][\phi]-fixed. Set jj to be the order of σ|J\left.\sigma\right|_{J}, fix BB\in\mathcal{B}, and let ix:FFi_{x}:F\to F be the inner automorphism such that ixϕj(B)Bi_{x}\circ\phi^{j}(B)\leq B. Define ψ=ixϕj\psi=i_{x}\circ\phi^{j}. As BB is eventually [ϕ][\phi]-periodic and hence eventually ϕj\phi^{j}-periodic, there are integers m>n1m>n\geq 1 such that [ψm(B)]=[ϕjm(B)]=[ϕjn(B)]=[ψn(B)].\displaystyle[\psi^{m}(B)]=[\phi^{jm}(B)]=[\phi^{jn}(B)]=[\psi^{n}(B)]. Therefore, there is an element yFy\in F such that yψn(B)y1=ψm(B)ψn(B)y\psi^{n}(B)y^{-1}=\psi^{m}(B)\leq\psi^{n}(B). But no finitely generated subgroup of FF is conjugate to a proper subgroup of itself (Lemma 1.3 below). So ψm(B)=ψn(B)\psi^{m}(B)=\psi^{n}(B) and, by injectivity of ϕ\phi, ψ(B)=B\psi(B)=B. Since BB\in\mathcal{B} was arbitrary, ϕj\phi^{j} fixes the free factors of \mathcal{B} up to conjugation; as \mathcal{B} is [ϕ][\phi]-invariant, it must be [ϕ][\phi]-fixed. ∎

The following fact will be used again in the proof of Proposition 2.4.

Lemma 1.3.

No finitely generated subgroup of FF is conjugate to a proper subgroup of itself.

Proof.

By Marshall Hall’s theorem, free groups are subgroup separable/locally extended residual finiteness (LERF), i.e., for any finitely generated subgroup HFH\leq F and element gFHg\in F\setminus H, there is a finite group GG and a surjective homomorphism φ:FG\varphi:F\to G such that φ(g)φ(H)\varphi(g)\notin\varphi(H). See [29] for a proof due to Stallings.

For a contradiction, suppose there is an element yFy\in F such that yHy1HyHy^{-1}\leq H and gHyHy1g\in H\setminus yHy^{-1}. By subgroup separability, there is a finite group GG and homomorphism φ:FG\varphi:F\to G such that φ(g)φ(yHy1)\varphi(g)\notin\varphi(yHy^{-1}). But gHg\in H implies φ(g)φ(H)\varphi(g)\in\varphi(H) and, by finiteness of GG, yHy1HyHy^{-1}\leq H implies φ(yHy1)=φ(H)\varphi(yHy^{-1})=\varphi(H) — a contradiction. ∎

The next lemma, also known as the Bounded Cancellation Lemma, will be used extensively in this paper. At the risk of overloading notation, for an edge-path pp in a graph Γ\Gamma, [p][p] denotes the immersed edge-path that is homotopic to pp rel. endpoints; for a loop ρ\rho in Γ\Gamma, [ρ][\rho] will be the immersed loop that is freely homotopic to ρ\rho.

Lemma 1.4 (Bounded Cancellation).

Let g:ΓΓg:\Gamma\to\Gamma^{\prime} be a π1\pi_{1}-injective graph map. Then there is a constant C(g)C(g) such that, for any natural edge-path decomposition p1p2p_{1}\cdot p_{2} of an immersed path in the universal cover Γ~\tilde{\Gamma}, the edge-path [g~(p1)][g~(p2)][\tilde{g}(p_{1})]\cdot[\tilde{g}(p_{2})] is contained in the C(g)C(g)-neighborhood of [g~(p1)g~(p2)][\tilde{g}(p_{1})\cdot\tilde{g}(p_{2})].

The following proof is due to Bestvina-Feighn-Handel [4, Lemma 3.1].

Proof.

Any graph map g:ΓΓg:\Gamma\to\Gamma^{\prime} factors as a pretrivial edge collapse and edge subdivision g0g_{0}, a composition of r0r\geq 0 folds grg1g_{r}\circ\cdots\circ g_{1}, and an simplicial immersion gr+1g_{r+1}. The collapse, subdivision, and immersion have cancellation constants 0 while each fold has cancellation constant 11 by π1\pi_{1}-injectivity. Thus we may choose C(g)=rC(g)=r. ∎

Let f:ΓΓf:\Gamma\to\Gamma be a topological representative for an injective endomorphism ϕ:FF\phi:F\to F, 𝒜\mathcal{A} be a nontrivial [ϕ][\phi]-invariant free factor system of FF, and Γ^k\hat{\Gamma}_{k} be the disjoint union of covers of Γ\Gamma corresponding to ϕk(𝒜)\phi^{k}(\mathcal{A}) for some k1k\geq 1. Then ff lifts to a map f^k:Γ^kΓ^k\hat{f}_{k}:\hat{\Gamma}_{k}\to\hat{\Gamma}_{k} and the deformation retraction Γ^kS[ϕk(𝒜)]\hat{\Gamma}_{k}\to S[\phi^{k}(\mathcal{A})] induces a map f¯k:S[ϕk(𝒜)]S[ϕk(𝒜)]\bar{f}_{k}:S[\phi^{k}(\mathcal{A})]\to S[\phi^{k}(\mathcal{A})] with K(f¯k)=K(f)K(\bar{f}_{k})=K(f) and C(f¯k)=C(f)C(\bar{f}_{k})=C(f). We shall call f¯k\bar{f}_{k} the (k\boldsymbol{k}-th) homotopy lift of f\boldsymbol{f}.

Stallings graphs S[ϕk(𝒜)]S[\phi^{k}(\mathcal{A})] are ϕk(𝒜)\phi^{k}(\mathcal{A})-marked graphs by definition and the maps f¯k\bar{f}_{k} are weak representatives for [ϕ|ϕk(𝒜)][\left.\phi\right|_{\phi^{k}(\mathcal{A})}\,] that might map branch points to bivalent vertices. Note that injectivity of ϕ\phi allows us to also consider the graphs as 𝒜\mathcal{A}-marked graphs and f¯k\bar{f}_{k} as weak representatives for [ϕ|𝒜][\left.\phi\right|_{\mathcal{A}}\,]. We hope to replace the weak representatives f¯k\bar{f}_{k} with homotopic natural representatives while maintaining uniform control on the Lipschitz and cancellation constants. The next lemma allows us to measure how close the homotopy lift f¯k\bar{f}_{k} is to mapping branch points to branch points.

Lemma 1.5.

Suppose g:ΓΓg:\Gamma\to\Gamma^{\prime} be a π1\pi_{1}-injective graph map with cancellation constant C=C(g)C=C(g). Then gg maps branch points to the CC-neighborhood of branch points.

Proof.

Set C=C(g)C=C(g). If Γ\Gamma^{\prime} is the CC-neighborhood of its branch points, then there is nothing to prove. Suppose νΓ\nu\in\Gamma^{\prime} is a bivalent vertex whose distance to the nearest branch point is greater than CC. We need to show that ν\nu is not the gg-image of any branch point. Set ϵ1\epsilon_{1} and ϵ2\epsilon_{2} to be the distinct oriented directions originating from ν\nu and ϵ¯1,ϵ¯2\bar{\epsilon}_{1},\bar{\epsilon}_{2} are the same directions with opposite orientation.

Let vΓv\in\Gamma be a branch point and g(v)=νg(v)=\nu. As vv is a branch point, there are at least three distinct oriented directions originating from vv: e1e_{1}, e2e_{2}, and e3e_{3}. Let p12p_{12} be an immersed path that starts and ends with e1e_{1} and e¯2\bar{e}_{2} respectively and define p23p_{23} similarly. Set p13=[p12p23]p_{13}=[p_{12}\cdot p_{23}] and p13=p12p¯23p^{\prime}_{13}=p_{12}\cdot\bar{p}_{23}, where p¯23\bar{p}_{23} is the reversal of the path p23p_{23}. See Figure 5 for an illustration. Although the paths are loops, we still treat them as paths, i.e., tightening is done rel. the endpoints. Without loss of generality, assume [g(p12)][g(p_{12})] starts with ϵ1\epsilon_{1}.

Refer to caption
Figure 5: Schematic for paths p12,p23,p13,p_{12},p_{23},p_{13}, and p13p_{13}^{\prime}. The path p13p_{13} starts with e1e_{1} follows the dashed path and ends with e¯3\bar{e}_{3}. The path p13p_{13}^{\prime} is the “figure 8” path traced by p12p_{12} then p¯23\bar{p}_{23}.

If [g(p12)][g(p_{12})] ends with ϵ¯1\bar{\epsilon}_{1}, then [g(p12)]=μ1ρμ¯1[g(p_{12})]=\mu_{1}\cdot\rho\cdot\bar{\mu}_{1} , where μ1\mu_{1} is an extension of ϵ1\epsilon_{1} to an embedded path from ν\nu to a branch point and ρ\rho is an immersed nontrivial loop. By hypothesis, μ1\mu_{1} is longer than CC. Since p12p_{12} starts and ends with e1e_{1} and e¯2\bar{e}_{2} respectively, the concatenation p12p12p_{12}\cdot p_{12} is an immersed path such that [g(p12)][g(p12)][g(p_{12})]\cdot[g(p_{12})] has μ¯1μ1\bar{\mu}_{1}\cdot\mu_{1} as a subpath, violating bounded cancellation. So we may assume [g(p12)][g(p_{12})] starts and ends with ϵ1\epsilon_{1} and ϵ¯2\bar{\epsilon}_{2}.

If [g(p23)][g(p_{23})] starts and ends with ϵ2\epsilon_{2} and ϵ¯1\bar{\epsilon}_{1}, then [g(p13)]=[g(p12)g(p23)][g(p_{13})]=[g(p_{12})\cdot g(p_{23})] starts and ends with ϵ1\epsilon_{1} and ϵ¯1\bar{\epsilon}_{1} respectively, which violates bounded cancellation for the same reason given in the previous paragraph. Similarly, if [g(p23)][g(p_{23})] starts and ends with ϵ1\epsilon_{1} and ϵ¯2\bar{\epsilon}_{2}, we rule out this possibility by considering [g(p13)][g(p^{\prime}_{13})]. We have ruled out all cases, and therefore, no branch point vv of Γ\Gamma is mapped to ν\nu. ∎

Corollary 1.6.

Let g:ΓΓg:\Gamma\to\Gamma^{\prime} be a KK-Lipschitz π1\pi_{1}-injective graph map with cancellation constant CC. Then gg is homotopic to a (K+C)(K+C)-Lipschitz natural graph map with cancellation constant 2C2C.

Proof.

By Lemma 1.5, gg maps branch points to the CC-neighborhood of branch points. So we can find a graph map gg^{\prime} homotopic to gg that maps branch points to branch points. Since the homotopy is moving images of branch points a distance at most CC, we can use K(g)=K+CK(g^{\prime})=K+C and C(g)=2CC(g^{\prime})=2C.

The bounded cancellation lemma only considers natural edge-paths, so homotopies that are supported in the interior of natural edges will not affect the cancellation constant. Using a tightening homotopy supported in the interior of natural edges in Γ\Gamma, we may assume gg^{\prime} is a natural graph map. The homotopy will not affect the Lipschitz and cancellation constants. ∎

So we can replace the homotopy lift f¯k\bar{f}_{k} with a homotopic natural representative that has Lipschitz and cancellation constants K(f)+C(f)K(f)+C(f) and 2C(f)2C(f) respectively. One would usually collapse the pretrivial edges and forget bivalent vertices to get a topological representative but we will not since we want to preserve compatibility: β\lVert\cdot\rVert_{\beta_{*}} is the restriction of α\lVert\cdot\rVert_{\alpha} to ϕk(𝒜)\phi^{k}(\mathcal{A}). To summarize the properties of homotopy lifts that will be used in the next sections:

Proposition 1.7.

Suppose ϕ:FF\phi:F\to F is injective, f:ΓΓf:\Gamma\to\Gamma is a topological representative for [ϕ][\phi], and 𝒜\mathcal{A} is a nontrivial [ϕ][\phi]-invariant free factor system. For any k1k\geq 1, there is a natural representative f¯k:S[ϕk(𝒜)]S[ϕk(𝒜)]\bar{f}_{k}:S[\phi^{k}(\mathcal{A})]\to S[\phi^{k}(\mathcal{A})] for [ϕ|ϕk(𝒜)][\left.\phi\right|_{\phi^{k}(\mathcal{A})}\,] with Lipschitz and cancellation constants K(f¯k)=K(f)+C(f)K(\bar{f}_{k})=K(f)+C(f) and C(f¯k)=2C(f)C(\bar{f}_{k})=2C(f) respectively.

The crucial point is that the Lipschitz and cancellation constants are independent of kk.

2 Canonical and elliptic free factor systems

In this section, we will construct a canonical invariant free factor system for any given injective endomorphism of FF. This free factor system, called the elliptic free factor system, is crucial for the construction of (expanding) relative immersions later in the paper.

Suppose 𝒜\mathcal{A} is a free factor system and ϕ:FF\phi:F\to F is an injective endomorphism. We shall say that a conjugacy class [g][g] in FF has an infinite [ϕ]\boldsymbol{[\phi]}-tail if for every n1n\geq 1, there is a conjugacy class [gn][g_{n}] in FF such that [ϕn(gn)]=[g][\phi^{n}(g_{n})]=[g]. The system 𝒜\mathcal{A} carries an infinite [ϕ]\boldsymbol{[\phi]}-tail of a conjugacy class [g][g] in FF if for every n1n\geq 1, there is a conjugacy class [gn][g_{n}] carried by 𝒜\mathcal{A} such that [ϕn(gn)]=[g][\phi^{n}(g_{n})]=[g]. We now state and prove the main technical result of this section.

Theorem 2.1.

If ϕ:FF\phi:F\to F is injective, then [ϕ][\phi] has a nontrivial fixed free factor system if and only if some nontrivial conjugacy class has an infinite [ϕ][\phi]-tail.

Remark.

Reynolds defined expansive endomorphisms [27, Definition 3.8] and [ϕ][\phi] is expansive in this sense exactly when only the trivial conjugacy class has an infinite [ϕ][\phi]-tail. Under this equivalence, Reynolds’ Remark 3.12 in [27] is a weaker form of this theorem.

Proof.

Let ϕ:FF\phi:F\to F be an injective endomorphism. The forward direction is obvious: if [ϕ][\phi] has a nontrivial fixed free factor system 𝒜\mathcal{A}, then any nontrivial conjugacy class [g][g] carried by 𝒜\mathcal{A} has an infinite [ϕ][\phi]-tail. The main content of the theorem is in the reverse direction. Fix a nontrivial conjugacy class [g][g] in FF with an infinite [ϕ][\phi]-tail. We proceed by descending down the poset of free factor systems. The following claim is the key idea:

Claim (Descent).

Let 𝒟\mathcal{D} be a [ϕ][\phi]-invariant free factor system that carries an infinite [ϕ][\phi]-tail of [g][g]. If 𝒟\mathcal{D} contains a free factor that is not eventually [ϕ][\phi]-periodic, then some [ϕ][\phi]-invariant free factor system 𝒟𝒟\mathcal{D}^{\prime}\prec\mathcal{D} carries an infinite [ϕ][\phi]-tail of [g][g].

Since there are no infinite chains in the poset of free factor systems, the descent (proven below) starts with the free factor system {F}\{F\} and then finds a [ϕ][\phi]-invariant free factor sytem 𝒜\mathcal{A} that carries an infinite [ϕ][\phi]-tail of [g][g] and whose free factors are eventually [ϕ][\phi]-periodic; such a free factor system contains a nontrivial [ϕ][\phi]-fixed free factor system by Lemma 1.2 and we are done.∎

Note that the assumption “𝒟\mathcal{D} contains a free factor that is not eventually [ϕ][\phi]-periodic” in the descent claim is more general than we really need for our conclusion. We could have worked with a more natural assumption “𝒟\mathcal{D} has no periodic free factors” and reached the same conclusion. However, we give the general argument instead since we will invoke a variation of it in the next proposition.

Proof of descent.

Let (Γ,α)(\Gamma,\alpha) be a marked graph, f:ΓΓf:\Gamma\to\Gamma be a topological representative for [ϕ][\phi], and set K=K(f)+C(f)K=K(f)+C(f) and C=2C(f)C=2C(f). Suppose 𝒟\mathcal{D} is a [ϕ][\phi]-invariant free factor system that carries an infinite [ϕ][\phi]-tail [gn]n1[g_{n}]_{n\geq 1} of [g][g] and contains a free factor that is not eventually [ϕ][\phi]-periodic. Then, for all k1k\geq 1 and nkn\geq k, there is an immersed loop ρk(gn)\rho_{k}(g_{n}) in Δk=S[ϕk(𝒟)]\Delta_{k}=S[\phi^{k}(\mathcal{D})] corresponding to [ϕk(gn)][\phi^{k}(g_{n})], where Δk=S[ϕk(𝒟)]\Delta_{k}=S[\phi^{k}(\mathcal{D})] is the Stallings graph with respect to (Γ,α)(\Gamma,\alpha). Set L=max{gα,C}L=\max\{\lVert g\rVert_{\alpha},C\}.

For all k1k\geq 1, let f¯k:ΔkΔk\bar{f}_{k}:\Delta_{k}\to\Delta_{k} be the natural representatives for [ϕ|ϕk(𝒟)][\left.\phi\right|_{\phi^{k}(\mathcal{D})}\,] given by Proposition 1.7. In particular, these representatives have Lipschitz and cancellation constants KK and CC respectively. As [gn]n1[g_{n}]_{n\geq 1} is an infinite [ϕ][\phi]-tail of [g][g] carried by 𝒟\mathcal{D}, we get, for any fixed k1k\geq 1, an infinite sequence of immersed loops (ρk(gn))nk(\rho_{k}(g_{n}))_{n\geq k} in Δk\Delta_{k} such that the free homotopy classes of f¯knk(ρk(gn))\bar{f}_{k}^{n-k}(\rho_{k}(g_{n})) have length gα\lVert g\rVert_{\alpha} for all nkn\geq k.

Form a directed graph 𝔾k\mathbb{G}_{k} whose vertices are the natural edges of Δk\Delta_{k} and there is a directed edge EiEjE_{i}\to E_{j} if f¯k\bar{f}_{k} maps natural edge EiE_{i} over EjE_{j}. Note that the number of natural edges of Δk\Delta_{k} is at most N=3rank(F)3N=3\cdot\mathrm{rank}(F)-3 and so 𝔾k\mathbb{G}_{k} has at most NN vertices.

Since 𝒟\mathcal{D} contains a free factor that is not eventually [ϕ][\phi]-periodic, the length of natural edges in Δk\Delta_{k} is unbounded as kk\to\infty by Lemma 1.1. Fix k0k\gg 0 such that the longest natural edge in Δk\Delta_{k} is longer than LKN1L\cdot K^{N-1}. Let 0\mathcal{L}_{0} be the natural edges of Δk\Delta_{k} longer than LKN1L\cdot K^{N-1} and \mathcal{L} be the union of 0\mathcal{L}_{0} and all natural edges on a directed path to 0\mathcal{L}_{0} in 𝔾k\mathbb{G}_{k}. Since f¯k\bar{f}_{k} is KK-Lipschitz and the shortest directed path in 𝔾k\mathbb{G}_{k} from a natural edge in \mathcal{L} to 0\mathcal{L}_{0} has at most NN natural edges on it, every natural edge in \mathcal{L} is longer than LL. The natural edges in \mathcal{L} will be referred to as the long natural edges and the remaining natural edges as 𝑠ℎ𝑜𝑟𝑡{\it short}.

Set ΔΔk\Delta^{\prime}\subset\Delta_{k} to be the union of short natural edges, which will be a proper subgraph since long natural edges exist. The subgraph is automatically f¯k\bar{f}_{k}-invariant by the construction of \mathcal{L}. Since ρk(gk)\rho_{k}(g_{k}) is an immersed loop in Δk\Delta_{k} with length gαL\lVert g\rVert_{\alpha}\leq L, its natural edges are short and Δ\Delta^{\prime} is a nonempty, noncontractible proper subgraph of Δk\Delta_{k}. Therefore, Δ\Delta^{\prime} determines a nontrivial [ϕ][\phi]-invariant proper free factor system 𝒟𝒟\mathcal{D}^{\prime}\prec\mathcal{D}. Technically, it determines an invariant free factor system of ϕk(F)\phi^{k}(F) but, as ϕ\phi is injective, this corresponds to an invariant free factor system of FF. It remains to show that 𝒟\mathcal{D}^{\prime} carries an infinite [ϕ][\phi]-tail of [g][g].

Let 𝕃𝔾k\mathbb{L}\subset\mathbb{G}_{k} be the full subgraph generated by the long natural edges \mathcal{L}. If there are no directed cycles in 𝕃\mathbb{L}, then f¯kN(Δk)Δ\bar{f}_{k}^{N}(\Delta_{k})\subset\Delta^{\prime}; in this case, the sequence of nontrivial loops (f¯kN(ρk(gn)))nk+N(\bar{f}_{k}^{N}(\rho_{k}(g_{n})))_{n\geq k+N} in Δ\Delta^{\prime} determines an infinite [ϕ][\phi]-tail of [g][g] carried by 𝒟\mathcal{D}^{\prime} and we are done. Now suppose there are directed cycles in 𝕃\mathbb{L} and let ρ\rho be an immersed loop in Δk\Delta_{k} that contains a long natural edge in such a cycle. Then, by bounded cancellation and the fact long natural edges are longer than LCL\geq C, [f¯km(ρ)][\bar{f}_{k}^{m}(\rho)] contains a long natural edge in the same directed cycle in 𝕃\mathbb{L} for all m1m\geq 1. Consequently, none of the immersed loops ρk(gn)\rho_{k}(g_{n}) in Δk\Delta_{k} contain a long natural edge that is in a directed cycle of 𝕃\mathbb{L}. Therefore, as far as the sequence of loops (ρk(gn))nk(\rho_{k}(g_{n}))_{n\geq k} is concerned, we may assume there are no directed cycles in 𝕃\mathbb{L} and, as before, the sequence (f¯kN(ρk(gn)))nk+N(\bar{f}_{k}^{N}(\rho_{k}(g_{n})))_{n\geq k+N} determines an infinite [ϕ][\phi]-tail of [g][g] carried by 𝒟\mathcal{D}^{\prime}. ∎

The following dichotomy is (equivalent to) a result in Reynolds’ thesis.

Corollary 2.2 ([27, Proposition 3.11]).

If ϕ:FF\phi:F\to F is irreducible, then either ϕ\phi is an automorphism or only the trivial conjugacy class has an infinite [ϕ][\phi]-tail.

Proof.

Suppose ϕ\phi is irreducible and there is a nontrivial conjugacy class with an infinite [ϕ][\phi]-tail. By Theorem 2.1, there is a nontrivial [ϕ][\phi]-fixed free factor system 𝒜\mathcal{A}. Since ϕ\phi is irreducible, 𝒜={F}\mathcal{A}=\{F\} and ϕ\phi is an automorphism. ∎

The fixed free factor system given by Theorem 2.1 may depend on the chosen conjugacy class [g][g] with an infinite tail or the marked graphs (Γ,α)(\Gamma,\alpha) chosen in the descent. The next proposition constructs a canonical fixed free factor system for [ϕ][\phi]; this system carries all conjugacy classes with an infinite tail as well as all finitely generated fixed subgroup system. The proof will use both descent and ascent (like a losing game of Tetris) in the poset of free factor systems!

Proposition 2.3.

If ϕ:FF\phi:F\to F is injective, then there is a unique maximal [ϕ][\phi]-fixed free factor system 𝒜\mathcal{A}. Precisely, 𝒜\mathcal{A} carries every conjugacy class with an infinite [ϕ][\phi]-tail and every finitely generated [ϕ][\phi]-fixed subgroup system.

Proof.

Let ϕ:FF\phi:F\to F be an injective endomorphism. If the trivial conjugacy class is the only conjugacy class with an infinite [ϕ][\phi]-tail, then the trivial system is the only [ϕ][\phi]-fixed subgroup system. In this case, the trivial system is the unique maximal [ϕ][\phi]-fixed free factor system and we are done as it vacuously carries all conjugacy classes with infinite [ϕ][\phi]-tails and all [ϕ][\phi]-fixed subgroup systems. We can now assume some nontrivial conjugacy class has an infinite [ϕ][\phi]-tail. By Theorem 2.1, ϕ\phi has a nontrivial fixed free factor system 𝒟0\mathcal{D}_{0}. We proceed by ascending up the poset of free factor systems:

Claim (Ascent).

Let \mathcal{B} be a finitely generated [ϕ][\phi]-fixed subgroup system, [g][g] a conjugacy class with an infinite [ϕ][\phi]-tail, and 𝒟\mathcal{D} a nontrivial [ϕ][\phi]-fixed free factor system. If 𝒟\mathcal{D} does not carry both \mathcal{B} and [g][g], then some [ϕ][\phi]-fixed free factor system 𝒟𝒟\mathcal{D}^{\prime}\succ\mathcal{D} carries both \mathcal{B} and [g][g].

Once again, as there are no infinite chains in the poset of free factor systems, the ascent (proven below) starts with the nontrivial [ϕ][\phi]-fixed proper free factor 𝒟0\mathcal{D}_{0} and stops at a necessarily unique maximal [ϕ][\phi]-fixed free factor system 𝒜\mathcal{A} that carries all finitely generated [ϕ][\phi]-fixed subgroup systems and all conjugacy classes with infinite [ϕ][\phi]-tails.∎

Proof of ascent.

Let \mathcal{B} be a finitely generated nontrivial [ϕ][\phi]-fixed subgroup system of FF, [g][g] be a nontrivial conjugacy class in FF with an infinite [ϕ][\phi]-tail, and 𝒟\mathcal{D} be a nontrivial [ϕ][\phi]-fixed free factor system of FF that does not carry both \mathcal{B} and [g][g]. We now describe the descent:

Claim (Descent).

Let 𝒟′′\mathcal{D}^{\prime\prime} be a [ϕ][\phi]-invariant free factor system that carries 𝒟\mathcal{D}, \mathcal{B}, and an infinite [ϕ][\phi]-tail of [g][g]. If 𝒟′′\mathcal{D}^{\prime\prime} contains a free factor that is not eventually [ϕ][\phi]-periodic, then some [ϕ][\phi]-invariant free factor system 𝒟′′′𝒟′′\mathcal{D}^{\prime\prime\prime}\prec\mathcal{D}^{\prime\prime} carries 𝒟\mathcal{D}, \mathcal{B}, and an infinite [ϕ][\phi]-tail of [g][g].

Starting with {F}\{F\}, the descent (proven below) will find a nontrivial [ϕ][\phi]-invariant free factor system 𝒟\mathcal{D}^{*} that carries 𝒟\mathcal{D}, \mathcal{B}, and an infinite [ϕ][\phi]-tail of [g][g] and whose free factors are eventually [ϕ][\phi]-periodic. By Lemma 1.2, 𝒟\mathcal{D}^{*} contains a [ϕ][\phi]-fixed free factor subsystem 𝒟𝒟\mathcal{D}^{\prime}\subset\mathcal{D}^{*} such that ϕk(𝒟)\phi^{k}(\mathcal{D}^{*}) is carried by 𝒟\mathcal{D}^{\prime} for some k0k\geq 0. As 𝒟\mathcal{D} and \mathcal{B} are [ϕ][\phi]-fixed, they are carried by 𝒟\mathcal{D}^{\prime}. Similarly, 𝒟\mathcal{D}^{\prime} carries [g][g] since 𝒟\mathcal{D}^{*} carries an infinite [ϕ][\phi]-tail of [g][g]. So 𝒟\mathcal{D}^{\prime} is a [ϕ][\phi]-fixed free factor system that carries 𝒟\mathcal{D}, \mathcal{B}, and [g][g] as needed for ascent.∎

Proof of descent.

Let (Γ,α)(\Gamma,\alpha) be a marked graph, f:ΓΓf:\Gamma\to\Gamma be a topological representative for [ϕ][\phi], and set K=K(f)+C(f)K=K(f)+C(f) and C=2C(f)C=2C(f). Suppose 𝒟′′\mathcal{D}^{\prime\prime} is a [ϕ][\phi]-invariant free factor system that carries 𝒟\mathcal{D}, \mathcal{B}, and an infinite [ϕ][\phi]-tail of [g][g]. Let S[ϕk(𝒟)]S[\phi^{k}(\mathcal{D})] and S[ϕk()]S[\phi^{k}(\mathcal{B})] be the Stallings graphs with respect to (Γ,α)(\Gamma,\alpha). Since 𝒟\mathcal{D} and \mathcal{B} are finitely generated and [ϕ][\phi]-fixed, the length of the longest immersed loop in S[ϕk(𝒟)]S[\phi^{k}(\mathcal{D})] or S[ϕk()]S[\phi^{k}(\mathcal{B})] that covers any edge at most twice is uniformly bounded by some L0L_{0} for all k1k\geq 1. Set L=max{L0,gα,C}L=\max\{L_{0},\lVert g\rVert_{\alpha},C\}. The proof now mimics that of the descent in Theorem 2.1 and we only give a sketch.

For all k1k\geq 1, let Δk′′=S[ϕk(𝒟′′)]\Delta_{k}^{\prime\prime}=S[\phi^{k}(\mathcal{D}^{\prime\prime})] and, by Proposition 1.7, there is a KK-Lipschitz natural representative for [ϕ|𝒟′′]\left[\left.\phi\right|_{\mathcal{D}^{\prime\prime}}\,\right], f¯k:Δk′′Δk′′\bar{f}_{k}:\Delta_{k}^{\prime\prime}\to\Delta_{k}^{\prime\prime}, that has cancellation constant CC. As some free factor in 𝒟′′\mathcal{D}^{\prime\prime} is not eventually [ϕ][\phi]-periodic, we can fix k0k\gg 0 so that the longest natural edge in Δk′′\Delta_{k}^{\prime\prime} is longer than LKN1L\cdot K^{N-1} by Lemma 1.1. Define the long and short natural edges as before and deduce long natural edges are longer than LL. Set Δ′′′Δk′′\Delta^{\prime\prime\prime}\subset\Delta_{k}^{\prime\prime} to be the union of the short natural edges, which is necessarily proper and f¯k\bar{f}_{k}-invariant. Recall that immersed loops of S[ϕk(𝒟)]S[\phi^{k}(\mathcal{D})] and S[ϕk()]S[\phi^{k}(\mathcal{B})] that cover any edge at most twice have length bounded by L0LL_{0}\leq L and these Stallings graphs have simplicial immersions into Δk′′\Delta_{k}^{\prime\prime}. Hence the images of these immersions lie in the subgraph of short natural edges Δ′′′\Delta^{\prime\prime\prime}. So Δ′′′\Delta^{\prime\prime\prime} is neither empty nor contractible. The subgraph Δ′′′\Delta^{\prime\prime\prime} determines a [ϕ][\phi]-invariant proper free factor system 𝒟′′′𝒟′′\mathcal{D}^{\prime\prime\prime}\prec\mathcal{D}^{\prime\prime} that carries both 𝒟\mathcal{D} and \mathcal{B} since both S[ϕk(𝒟)]S[\phi^{k}(\mathcal{D})] and S[ϕk()]S[\phi^{k}(\mathcal{B})] have immersions into Δ′′′\Delta^{\prime\prime\prime}. From the proof of descent in Theorem 2.1, LCL\geq C implies 𝒟′′′\mathcal{D}^{\prime\prime\prime} carries an infinite [ϕ][\phi]-tail of [g][g]. ∎

Although this proposition produces a canonical fixed free factor system for an injective endomorphism, we shall enlarge the system again to get a better [ϕ][\phi]-invariant free factor system that gives us some control of the relative dynamics of [ϕ][\phi]. We do this by taking iterated preimages of the maximal fixed free factor system. We then show that the resulting invariant free factor system is a disjoint union of the maximal fixed free factor system with a free factor system that eventually gets mapped into the fixed system.

Proposition 2.4.

If ϕ:FF\phi:F\to F is injective and 𝒜\mathcal{A} is the maximal [ϕ][\phi]-fixed free factor system, then there is a unique maximal [ϕ][\phi]-invariant free factor system 𝒜𝒜\mathcal{A}^{*}\succeq\mathcal{A} such that ϕk(𝒜)\phi^{k}(\mathcal{A}^{*}) is carried by 𝒜\mathcal{A} for some k0k\geq 0. After replacing the free factors of 𝒜\mathcal{A} with conjugates if necessary, we can assume 𝒜𝒜\mathcal{A}\subset\mathcal{A}^{*}.

We shall call the free factor system given by this proposition the [ϕ]\boldsymbol{[\phi]}-elliptic free factor system. We call it elliptic since it will be carried by the vertex groups in free splittings of FF in the subsequent sections.

Proof.

Let ϕ:FF\phi:F\to F be an injective endomorphism and 𝒜\mathcal{A} be the maximal [ϕ][\phi]-fixed free factor system of FF given by Proposition 2.3. If ϕ\phi is surjective or has no nontrivial fixed free factor systems, then 𝒜=𝒜\mathcal{A}^{*}=\mathcal{A} is {F}\{F\} or trivial respectively and we are done. Thus, we assume that 𝒜\mathcal{A} is a nontrivial proper free factor system. For all k1k\geq 1, define 𝒜k=ϕk𝒜\mathcal{A}_{k}=\phi^{-k}\cdot\mathcal{A}.

Since 𝒜\mathcal{A} is [ϕ][\phi]-invariant, we get that 𝒜𝒜k𝒜k+1\mathcal{A}\preceq\mathcal{A}_{k}\preceq\mathcal{A}_{k+1} for all k1k\geq 1 and, consequently, all 𝒜k\mathcal{A}_{k} are [ϕ][\phi]-invariant. As there are no infinite chains in the poset of free factor systems of FF, the chain of [ϕ][\phi]-invariant free factor systems 𝒜k(k1)\mathcal{A}_{k}~{}(k\geq 1) stabilizes and we can set 𝒜\mathcal{A}^{*} to be the maximal free factor system in the chain. By construction, 𝒜\mathcal{A}^{*} carries a subgroup system \mathcal{B} if and only if ϕk()\phi^{k}(\mathcal{B}) is carried by 𝒜\mathcal{A} for some k0k\geq 0 and this implies the uniqueness of 𝒜\mathcal{A}^{*} amongst [ϕ][\phi]-invariant free factor system that have ϕ\phi-iterates carried by 𝒜\mathcal{A}. It remains to show that 𝒜𝒜\mathcal{A}\subset\mathcal{A}^{*} after replacing the free factors of 𝒜\mathcal{A} with conjugates if necessary.

As in the proof of Lemma 1.2, suppose σ:{1,l}{1,,l}\sigma:\{1,\ldots l\}\to\{1,\ldots,l\} is the function associated to the [ϕ][\phi]-invariance of 𝒜={A1,Al}\mathcal{A}^{*}=\{A_{1},\ldots A_{l}\}. Then there is a maximal nonempty subset J{1,,l}J\subset\{1,\ldots,l\} on which σ\sigma is a bijection. Let 𝒜J={Aj𝒜:jJ}\mathcal{A}_{J}=\{\,A_{j}\in\mathcal{A}^{*}\,:\,j\in J\,\}. Since 𝒜\mathcal{A}^{*} carries the maximal [ϕ][\phi]-fixed free factor system 𝒜\mathcal{A}, it follows that 𝒜𝒜J\mathcal{A}\preceq\mathcal{A}_{J}. Replace components of 𝒜\mathcal{A} with conjugates if necessary and assume 𝒜\mathcal{A} is a free factor system of 𝒜J\mathcal{A}_{J}; in particular, each A𝒜A\in\mathcal{A} is a subgroup of some Aj𝒜JA_{j}\in\mathcal{A}_{J}. We want to show that 𝒜𝒜J\mathcal{A}\subset\mathcal{A}_{J}. Choose a component A𝒜A\in\mathcal{A} and let Aj𝒜JA_{j}\in\mathcal{A}_{J} be the component such that AAjA\leq A_{j}. Furthermore, fix an inner automorphism ix:FFi_{x}:F\to F such that ixϕs(Aj)Aji_{x}\circ\phi^{s}(A_{j})\leq A_{j} for some s0s\geq 0. Set ψ=ixϕs\psi=i_{x}\circ\phi^{s}.

By construction, Aj𝒜=ϕks𝒜A_{j}\in\mathcal{A}^{*}=\phi^{-ks}\cdot\mathcal{A} implies ψk(Aj)\psi^{k}(A_{j}) is conjugate to a subgroup of a [ϕ][\phi]-periodic, and hence [ψ][\psi]-periodic, free factor A𝒜A^{\prime}\in\mathcal{A} for some k0k\geq 0. We must have AAjA^{\prime}\leq A_{j} since AAj𝒜JA^{\prime}\leq A_{j\,^{\prime}}\in\mathcal{A}_{J}, {A}\{A^{\prime}\} carries {ψk(Aj)}\{\psi^{k}(A_{j})\}, ψ(Aj)Aj\psi(A_{j})\leq A_{j}, and σ|J\left.\sigma\right|_{J} is a bijection. So iyψk(Aj)AAji_{y}\circ\psi^{k}(A_{j})\leq A^{\prime}\leq A_{j} for some inner automorphism iy:FFi_{y}:F\to F. The [ψ][\psi]-periodicity of AA^{\prime} implies (iyψk)m(A)A(i_{y}\circ\psi^{k})^{m}(A^{\prime})\leq A^{\prime} is conjugate to AA^{\prime} for some m1m\geq 1. But Lemma 1.3 says no finitely generated subgroup of FF is conjugate to a proper subgroup of itself. Therefore, (iyψk)m(A)=A(i_{y}\circ\psi^{k})^{m}(A^{\prime})=A^{\prime} and, by injectivity of ψ\psi, iyψk(Aj)=A=Aji_{y}\circ\psi^{k}(A_{j})=A^{\prime}=A_{j}. In particular, A=A=AjA=A^{\prime}=A_{j}. As this holds for arbitrary free factors A𝒜A\in\mathcal{A}, we get 𝒜𝒜J\mathcal{A}\subset\mathcal{A}_{J}. ∎

It is obvious that the maximal [ϕ][\phi]-fixed free factor system is proper exactly when ϕ\phi is not surjective. The same holds for the [ϕ][\phi]-elliptic free factor system:

Observation.

Let ϕ:FF\phi:F\to F be an injective endomorphism. The [ϕ][\phi]-elliptic free factor system is

  1. 1.

    proper exactly when ϕ\phi is not surjective.

  2. 2.

    trivial exactly when only the trivial conjugacy class has an infinite [ϕ][\phi]-tail;

3 Relative representatives

We will now use relative (weak) representatives as the basis for inductively studying dynamical properties of free group endomorphisms. For the whole section, we suppose ϕ:FF\phi:F\to F is an injective endomorphism and ϕ1𝒜=𝒜\phi^{-1}\cdot\mathcal{A}=\mathcal{A}, i.e, 𝒜\mathcal{A} is a [ϕ][\phi]-invariant proper free factor system such that there is no free factor system 𝒜𝒜\mathcal{A}^{\prime}\succ\mathcal{A} such that 𝒜\mathcal{A} carries ϕ(𝒜)\phi(\mathcal{A}^{\prime}); e.g., consider an injective nonsurjective ϕ\phi and its elliptic free factor system.

The first step is to establish the relative version of the bounded cancellation.

Lemma 3.1 (Bounded Cancellation).

Let TT and TT^{\prime} be (F,𝒜)(F,\mathcal{A})- and (F,𝒜)(F^{\prime},\mathcal{A}^{\prime})-trees respectively and ψ:FF\psi:F\to F^{\prime} be an injective homomorphism such that ψ1𝒜=𝒜\psi^{-1}\cdot\mathcal{A}^{\prime}=\mathcal{A}. If g:TTg:T\to T^{\prime} is a ψ\psi-equivariant tree map, then there is a constant C(g)C(g) such that for every natural edge-path decomposition p1p2p_{1}\cdot p_{2} of an immersed path in TT, the edge-path [g(p1)][g(p2)][g(p_{1})]\cdot[g(p_{2})] is contained in the C(g)C(g)-neighborhood of [g(p1)g(p2)][g(p_{1})\cdot g(p_{2})].

Proof.

The proof is the same as before. Since the trees are simplicial, the map gg factors as an equivariant pretrivial edge collapse and subdivision, a composition of r0r\geq 0 equivariant folds, and an equivariant simplicial embedding. As ψ\psi is injective and ψ1𝒜=𝒜\psi^{-1}\cdot\mathcal{A}^{\prime}=\mathcal{A}, no fold identifies vertices in the same orbit and, hence, each fold has cancellation constant 1. We may choose C(g)=rC(g)=r. ∎

Let 𝒜\mathcal{A}\preceq\mathcal{B} be a chain of [ϕ][\phi]-invariant free factor systems, TT_{*} be a (,𝒜)(\mathcal{B},\mathcal{A})-forest, and f:TTf_{*}:T_{*}\to T_{*} be a 𝒜\mathcal{A}-relative representative for ϕ|\left.\phi\right|_{\mathcal{B}}. For all k1k\geq 1, set T(ϕk())TT_{*}(\phi^{k}(\mathcal{B}))\subset T_{*} to be the minimal subforest for ϕk()\phi^{k}(\mathcal{B}); minimal subforests are the relative analogues of iterated Stallings graphs. We will assume the minimal subforests T(ϕk())T_{*}(\phi^{k}(\mathcal{B})) inherit their simplicial structure from the ambient forest TT_{*} and so they might have bivalent vertices unlike TT_{*}. For a graph of groups decomposition with bivalent vertices, branch points are vertices that are images of branch points of the Bass-Serre tree and natural edges are images of natural edges of the tree.

For any k1k\geq 1, let f,k:T(ϕk())Tf_{*,k}:T_{*}(\phi^{k}(\mathcal{B}))\to T_{*} be the restriction of ff_{*} to T,kT_{*,k} and then replace it with an equivariantly homotopic map T(ϕk())T(ϕk())T_{*}(\phi^{k}(\mathcal{B}))\to T_{*}(\phi^{k}(\mathcal{B})) that is induced by the deformation retraction of f,k(T(ϕk()))f_{*,k}(T_{*}(\phi^{k}(\mathcal{B}))) to T(ϕk())T_{*}(\phi^{k}(\mathcal{B})), which we call the (k\boldsymbol{k}-th) homotopy restriction of f\boldsymbol{f_{*}}. Note that if XT(ϕk())X\subset T_{*}(\phi^{k}(\mathcal{B})) is an axis such that f|X\left.f_{*}\right|_{X} is an immersion, then f,k|X\left.f_{*,k}\right|_{X} is still an immersion.

These homotopy restrictions are the relative analogues of homotopy lifts. The proof of the following lemma is also almost the same as that of Lemma 1.5.

Lemma 3.2.

Let TT and TT^{\prime} be (F,𝒜)(F,\mathcal{A})- and (F,𝒜)(F^{\prime},\mathcal{A}^{\prime})-trees respectively and ψ:FF\psi:F\to F^{\prime} be an injective homomorphism such that ψ1𝒜=𝒜\psi^{-1}\cdot\mathcal{A}^{\prime}=\mathcal{A}. If g:TTg:T\to T^{\prime} is a ψ\psi-equivariant tree map with cancellation constant C=C(g)C=C(g), then gg maps branch points to the CC-neighborhood of branch points.

Proof.

Set C=C(g)C=C(g) and let ν\nu be a bivalent vertex in TT^{\prime} whose distance to the nearest branch point is greater than CC. In particular, ν\nu has a trivial stabilizer. We denote by ϵ1,ϵ2\epsilon_{1},\epsilon_{2} the 2 distinct directions at ν\nu oriented away from the vertex. Suppose vv is a branch point of TT such that g(v)=νg(v)=\nu. As ψ\psi is injective, vv has a trivial stabilizer under the action of FF. Choose 3 distinct directions at vv: e1,e2,e_{1},e_{2}, and e3e_{3}. Let p12p_{12} be an embedded path in TT that starts with e1e_{1} and ends with a translate of e¯2\bar{e}_{2}. Since vv has a trivial stabilizer, the path determines a unique loxodromic element x12x_{12} in FF with axis a12a_{12} such that p12p_{12} is a fundamental domain of the axis under the translation action of x12x_{12}. Without loss of generality, [g(p12)][g(p_{12})] starts with ϵ1\epsilon_{1}.

If [g(p12)][g(p_{12})] ends with the translate ϕ(x12)ϵ¯1\phi(x_{12})\bar{\epsilon}_{1}, then [g(p12)]=μρ(ϕ(x12)μ¯)[g(p_{12})]=\mu\cdot\rho\cdot(\phi(x_{12})\bar{\mu}), where μ\mu is an extension of ϵ1\epsilon_{1} to an embedded path from ν\nu to the axis of ϕ(x12)\phi(x_{12}) and ρ\rho is a fundamental domain of the axis of loxodromic element ϕ(x12)\phi(x_{12}). By assumption, μ\mu is longer than CC. Decompose the axis a12=aa+a_{12}=a_{-}\cdot a_{+} at vv, then [g(a)][g(a+)][g(a_{-})]\cdot[g(a_{+})] has μ¯μ\bar{\mu}\cdot\mu as a subpath, violating bounded cancellation. The remaining cases are handled similarly. Upon ruling out all cases, we conclude that no branch point vv of TT is mapped to ν\nu. ∎

As before, we get a corollary whose proof is essentially the same as that of Corollary 1.6.

Corollary 3.3.

Let TT and TT^{\prime} be (F,𝒜)(F,\mathcal{A})- and (F,𝒜)(F^{\prime},\mathcal{A}^{\prime})-trees respectively and ψ:FF\psi:F\to F^{\prime} be an injective homomorphism such that ψ1𝒜=𝒜\psi^{-1}\cdot\mathcal{A}^{\prime}=\mathcal{A}. If g:TTg:T\to T^{\prime} is a ψ\psi-equivariant KK-Lipschitz tree map with cancellation constant CC, then gg is equivariantly homotopic to a ψ\psi-equivariant (K+C)(K+C)-Lipschitz natural tree map with cancellation constant 2C2C.

The corollary allows us to replace f,kf_{*,k} with an equivariantly homotopic ϕk(𝒜)\phi^{k}(\mathcal{A})-relative natural representative that has Lipschitz and cancellation constants K(f,k)=K(f)+C(f)K(f_{*,k})=K(f_{*})+C(f_{*}) and C(f,k)=2C(f)C(f_{*,k})=2C(f_{*}) respectively. If XT(ϕk())X\subset T_{*}(\phi^{k}(\mathcal{B})) is an axis and f,k|X\left.f_{*,k}\right|_{X} is an immersion before the homotopy, then f,k(X)f_{*,k}(X) is an immersed path after the homotopy; however, the restriction f,k|X\left.f_{*,k}\right|_{X} may fail to be an immersion due to pretrivial edges. The following summary is a relative analogue of Proposition 1.7.

Proposition 3.4.

Let f:TTf_{*}:T_{*}\to T_{*} be an 𝒜\mathcal{A}-relative representative for ϕ|\left.\phi\right|_{\mathcal{B}}. For any k1k\geq 1, there is an ϕk(𝒜)\phi^{k}(\mathcal{A})-relative natural representative f,k:T(ϕk())T(ϕk())f_{*,k}:T_{*}(\phi^{k}(\mathcal{B}))\to T_{*}(\phi^{k}(\mathcal{B})) for ϕ|ϕk()\left.\phi\right|_{\phi^{k}(\mathcal{B})} such that:

  1. 1.

    K(f,k)=K(f)+C(f)K(f_{*,k})=K(f_{*})+C(f_{*}) and C(f,k)=2C(f)C(f_{*,k})=2C(f_{*}).

  2. 2.

    If ff_{*} restricted to the axes of some conjugacy class [b][b] in ϕk()\phi^{k}(\mathcal{B}) is an immersion, then f,kf_{*,k} restricted to the axes of [b][b] is also an immersion modulo pretrivial edges.

Collapsing a maximal (orbit-closed) f,kf_{*,k}-invariant subforest of T(ϕk())T_{*}(\phi^{k}(\mathcal{B})) with bounded components and forgetting the bivalent vertices induces a minimal ϕk(𝒜)\phi^{k}(\mathcal{A})-relative representative g,k:Y,kY,kg_{*,k}:Y_{*,k}\to Y_{*,k} for ϕ|ϕk()\left.\phi\right|_{\phi^{k}(\mathcal{B})} defined on a (ϕk(),ϕk(𝒜))(\phi^{k}(\mathcal{B}),\phi^{k}(\mathcal{A}))-forest Y,kY_{*,k}. Note that the collapsed maximal subforest contains the pretrivial edges as ϕ1𝒜=𝒜\phi^{-1}\cdot\mathcal{A}=\mathcal{A}. Since g,kg_{*,k} is induced by equivariantly collapsing a forest and forgetting bivalent vertices, we have K(g,k)=K(f)+C(f)K(g_{*,k})=K(f_{*})+C(f_{*}), C(g,k)=2C(f)C(g_{*,k})=2C(f_{*}), and lY,klT|ϕk()l_{Y_{*,k}}\leq\left.l_{T_{*}}\right|_{\phi^{k}(\mathcal{B})}. If XX is an axis of bb in T(ϕk())T_{*}(\phi^{k}(\mathcal{B})) and f,k|X\left.f_{*,k}\right|_{X} is an immersion modulo pretrivial edges, then g,k|X\left.g_{*,k}\right|_{X^{\prime}} is an immersion, where XX^{\prime} is the axis of bb in Y,kY_{*,k}.

Proposition 3.5.

Let f:TTf_{*}:T_{*}\to T_{*} be an 𝒜\mathcal{A}-relative representative for ϕ|\left.\phi\right|_{\mathcal{B}}. For any k1k\geq 1, there is a (ϕk(),ϕk(𝒜))(\phi^{k}(\mathcal{B}),\phi^{k}(\mathcal{A}))-forest Y,kY_{*,k} and a minimal ϕk(𝒜)\phi^{k}(\mathcal{A})-relative representative g,k:Y,kY,kg_{*,k}:Y_{*,k}\to Y_{*,k} for ϕ|ϕk()\left.\phi\right|_{\phi^{k}(\mathcal{B})} such that:

  1. 1.

    K(g,k)=K(f)+C(f)K(g_{*,k})=K(f_{*})+C(f_{*}) and C(g,k)=2C(f)C(g_{*,k})=2C(f_{*}).

  2. 2.

    lY,k:ϕk()l_{Y_{*,k}}:\phi^{k}(\mathcal{B})\to\mathbb{R} is dominated by (\leq) the restriction lT|ϕk()=lT(ϕk())\left.l_{T_{*}}\right|_{\phi^{k}(\mathcal{B})}=l_{T_{*}(\phi^{k}(\mathcal{B}))};

  3. 3.

    If ff_{*} restricted to the axes of some conjugacy class [b][b] in ϕk()\phi^{k}(\mathcal{B}) is an immersion, then g,kg_{*,k} restricted to the axes of [b][b] is also an immersion.

For an 𝒜\mathcal{A}-relative weak representative ff_{*} for ϕ|\left.\phi\right|_{\mathcal{B}}, we define the transition matrix A(f)A(f_{*}). Let A(f)A(f_{*}) be a nonnegative integer square matrix whose rows and columns are indexed by the orbits of edges in TT_{*}; the entry of A(f)A(f_{*}) in row-ii and column-jj, A(f)(i,j)A(f_{*})(i,j), is given by the number of translates of eie_{i} that are contained in the immersed edge-path f(ej)f_{*}(e_{j}), where eie_{i} is a orbit representative for the ii-th orbit of edges. An 𝒜\mathcal{A}-relative weak representative ff_{*} is irreducible if the matrix A(f)A(f_{*}) is irreducible, i.e., for any pair (i,j)(i,j), there is a positive integer nijn_{ij} such that A(f)nij(i,j)>0A(f_{*})^{n_{ij}}(i,j)>0. In this case, the stretch factor of ff_{*}, λ(f)1\lambda(f_{*})\geq 1, is the Perron-Frobenius eigenvalue of A(f)A(f_{*}). An irreducible 𝒜\mathcal{A}-relative weak representative is expanding if λ(f)>1\lambda(f_{*})>1. Note that irreducible 𝒜\mathcal{A}-relative representatives are minimal.

We say ϕ|\left.\phi\right|_{\mathcal{B}} is irreducible relative to 𝒜\boldsymbol{\mathcal{A}} if there is no [ϕ][\phi]-invariant free factor system 𝒞\mathcal{C} such that 𝒜𝒞\mathcal{A}\prec\mathcal{C}\prec\mathcal{B}. If 𝒜\mathcal{A} is trivial and ={F}\mathcal{B}=\{F\}, then we recover the definition of ϕ\phi’s irreducibility. The next lemma is the most useful property of an irreducible restriction for our purposes.

Lemma 3.6.

If ϕ|\left.\phi\right|_{\mathcal{B}} is irreducible relative to 𝒜\mathcal{A}, then every minimal 𝒜\mathcal{A}-relative representative for ϕ|\left.\phi\right|_{\mathcal{B}} is irreducible.

Proof.

Suppose some minimal 𝒜\mathcal{A}-relative representative for ϕ|\left.\phi\right|_{\mathcal{B}} has a reducible transition matrix; in particular, it has an invariant \mathcal{B}-equivariant proper subforest (with unbounded components) that determines a [ϕ][\phi]-invariant free factor system 𝒞\mathcal{C} such that 𝒜𝒞\mathcal{A}\prec\mathcal{C}\prec\mathcal{B}. So ϕ|\left.\phi\right|_{\mathcal{B}} is not irreducible relative to 𝒜\mathcal{A}. ∎

Remark.

Bestvina-Handel give the absolute version of this property as the definition of irreducibility and then prove that it is equivalent to the definition of irreducibility given in this paper [6, Lemma 1.16]. The relative version of this equivalence holds as well but we will not prove it as it is not needed.

Bestvina-Handel used the next proposition to construct train tracks [6, Theorem 1.7].

Proposition 3.7.

If ϕ|\left.\phi\right|_{\mathcal{B}} is irreducible relative to 𝒜\mathcal{A}, then there is an irreducible 𝒜\mathcal{A}-relative representative f:TTf_{*}:T_{*}\to T_{*} for ϕ|\left.\phi\right|_{\mathcal{B}} with the minimal stretch factor, i.e., if f:TTf_{*}^{\prime}:T_{*}^{\prime}\to T_{*}^{\prime} is an irreducible 𝒜\mathcal{A}-relative representative for ϕ|\left.\phi\right|_{\mathcal{B}}, then λ(f)λ(f)\lambda(f_{*}^{\prime})\geq\lambda(f_{*}).

The minimal stretch factor will be denoted by λ([ϕ],,𝒜)\lambda([\phi],\mathcal{B},\mathcal{A}).

Proof.

Let g:YYg_{*}:Y_{*}\to Y_{*} be a minimal 𝒜\mathcal{A}-relative representative for ϕ|\left.\phi\right|_{\mathcal{B}} and suppose ϕ|\left.\phi\right|_{\mathcal{B}} is irreducible relative to 𝒜\mathcal{A}. Then gg_{*} is an irreducible 𝒜\mathcal{A}-relative representative by Lemma 3.6 with stretch factor λ(g)\lambda(g_{*}). By the lack on bivalent vertices, any irreducible 𝒜\mathcal{A}-relative representative has a transition matrix of size N=3rank(F)3\leq N=3\cdot\mathrm{rank}(F)-3. Suppose BB is an irreducible integer square matrix with Perron-Frobenius eigenvalue λ(B)λ(g)\lambda(B)\leq\lambda(g_{*}). Then BB has a positive right eigenvector v\vec{v} associated with λ(B)\lambda(B). So for all k1k\geq 1, BkB^{k} has right eigenvector v\vec{v} associated with eigenvalue λ(B)k\lambda(B)^{k}. Assuming the smallest entry of v\vec{v} is 11 (rescale if necessary), we get that the minimum row-sum of BkB^{k} is at most λ(B)k\lambda(B)^{k} for any k1k\geq 1. If BB has no more than NN rows, then the largest entry of BB is at most the minimum row-sum of BN!B^{N!}, which we know is at most λ(B)N!λ(g)N!\lambda(B)^{N!}\leq\lambda(g_{*})^{N!}. So there are finitely many irreducible integer square matrices with size N\leq N and Perron-Frobenius eigenvalue λ(g)\leq\lambda(g_{*}). Thus, there is a finite set of stretch factors λ(g)\leq\lambda(g_{*}) for irreducible 𝒜\mathcal{A}-relative representatives for ϕ|\left.\phi\right|_{\mathcal{B}}. In particular, there is an irreducible 𝒜\mathcal{A}-relative representative f:TTf_{*}:T_{*}\to T_{*} for ϕ|\left.\phi\right|_{\mathcal{B}} with the minimal stretch factor. ∎

Bestvina-Handel’s work [6] can be adapted to show that an irreducible 𝒜\mathcal{A}-relative representative for ϕ|\left.\phi\right|_{\mathcal{B}} with the minimal stretch factor is an 𝒜\mathcal{A}-relative train track (Appendix A) and, conversely, bounded cancellation implies all irreducible 𝒜\mathcal{A}-relative train tracks for ϕ|\left.\phi\right|_{\mathcal{B}} have the minimal stretch factor. We do not prove this converse as it is not needed. The next lemma is an application of train track theory that will be invoked once, in the second half of the proof of Proposition 3.9.

Lemma 3.8 (Train Track Theory).

If ϕ|\left.\phi\right|_{\mathcal{B}} is irreducible relative to 𝒜\mathcal{A} and f:TTf_{*}:T_{*}\to T_{*} is an irreducible 𝒜\mathcal{A}-relative representative for ϕ|\left.\phi\right|_{\mathcal{B}} with the minimal stretch factor, then there is an element gg in \mathcal{B} with an axis aga_{g} such that the restriction of fkf_{*}^{k} to aga_{g} is an immersion for all k1k\geq 1.

Such an axis will be known as an 𝒇\boldsymbol{f_{*}}-legal axis.

Proof.

If λ(f)=1\lambda(f_{*})=1, then ff_{*} is a simplicial embedding and we are done. So we may assume λ(f)>1\lambda(f_{*})>1. By minimality of its stretch factor, ff_{*} is an expanding irreducible 𝒜\mathcal{A}-relative train track for ϕ|\left.\phi\right|_{\mathcal{B}} (Theorem A.1), i.e., for any edge ee in TT_{*}, fk(e)f_{*}^{k}(e) is an expanding immersed path for all k1k\geq 1. A 2-edge path e1e2e_{1}\cdot e_{2} is ff_{*}-legal if it is a translate of a subpath of fk(e)f_{*}^{k}(e) for some edge ee and integer k1k\geq 1. By irreducibility of ff_{*}, every edge ee is contained in a 3-edge path eee+e_{-}\cdot e\cdot e_{+} whose 2-edge subpaths are both ff_{*}-legal. This means we can form an axis aga_{g} whose 2-edge subpaths are all ff_{*}-legal. By the train track property, the restriction of fkf_{*}^{k} to aga_{g} is an immersion for all k1k\geq 1. ∎

The main tools from Section 1 that were used in the previous section were Lemma 1.1, bounded cancellation, and Proposition 1.7. The relative analogues of the latter two have already been established in this section. We now state the main technical result of this section, an analogue of Lemma 1.1 — analogous in the sense that both give sufficient conditions for iterated subgroup graphs to have arbitrarily long natural edges.

Proposition 3.9.

Let 𝒜\mathcal{A}\prec\mathcal{B} be a chain of [ϕ][\phi]-invariant free factor systems with ϕ|\left.\phi\right|_{\mathcal{B}} irreducible relative to 𝒜\mathcal{A} and λ([ϕ],,𝒜)>1\lambda([\phi],\mathcal{B},\mathcal{A})>1. If 𝒜\mathcal{A} carries the maximal [ϕ][\phi]-fixed free factor system, then the length of the longest natural edge in T(ϕk())T_{*}(\phi^{k}(\mathcal{B})) is unbounded as kk\to\infty.

Before starting the proof, we will first describe the (absolute) vertex blow-up construction. Let 𝒜\mathcal{A}\prec\mathcal{B} be a chain of free factor systems and TT_{*} be some (,𝒜)(\mathcal{B},\mathcal{A})-forest. Recall that we assume 𝒜i𝒜\mathcal{A}_{i}\subset\mathcal{A} is also a free factor system of BiB_{i}\in\mathcal{B}. Fix some 𝒜\mathcal{A}-marked roses (R𝒜,α𝒜)(R_{\mathcal{A}},\alpha_{\mathcal{A}}) . Define Γ\Gamma_{\mathcal{B}} to be the graph formed by identifying the appropriate vertices of the graph of groups \T\mathcal{B}\backslash T_{*} with the basepoints of roses (R𝒜,α𝒜)(R_{\mathcal{A}},\alpha_{\mathcal{A}}). If c:R𝒜Γc:R_{\mathcal{A}}\to\Gamma_{\mathcal{B}} is the inclusion map, then Bass-Serre theory gives markings α={αi:Biπ1(Γi)}\alpha_{\mathcal{B}}=\{\alpha_{i}:B_{i}\to\pi_{1}(\Gamma_{i})\} such that [π1(c)α𝒜]=[(α)|𝒜][\,\pi_{1}(c)\circ\alpha_{\mathcal{A}}\,]=[\,\left.(\alpha_{\mathcal{B}})\right|_{\mathcal{A}}\,]. Thus, (Γ,α)(\Gamma_{\mathcal{B}},\alpha_{\mathcal{B}}) is a \mathcal{B}-marked graph. This construction and, in general any pair of graphs Γ𝒜Γ\Gamma_{\mathcal{A}}^{\prime}\subset\Gamma_{\mathcal{B}}^{\prime} with collections of markings α𝒜,α\alpha_{\mathcal{A}}^{\prime},\alpha_{\mathcal{B}}^{\prime} such that π1(c)α𝒜=(α)|𝒜\pi_{1}(c^{\prime})\circ\alpha_{\mathcal{A}}^{\prime}=\left.(\alpha_{\mathcal{B}}^{\prime})\right|_{\mathcal{A}} will be referred to as vertex blow-up.

We note that the Stallings graph S[ϕk()]S[\phi^{k}(\mathcal{B})] with respect to (Γ,α)(\Gamma_{\mathcal{B}},\alpha_{\mathcal{B}}), as a ϕk()\phi^{k}(\mathcal{B})-marked graph, is a vertex blow-up of ϕk()\T(ϕk())\phi^{k}(\mathcal{B})\backslash T_{*}(\phi^{k}(\mathcal{B})): let ι:S[ϕk()]Γ\iota:S[\phi^{k}(\mathcal{B})]\to\Gamma_{\mathcal{B}} be the Stallings graph’s immersion and S𝒜S[ϕk()]S_{\mathcal{A}}\subset S[\phi^{k}(\mathcal{B})] be the core of the subgraph ι1(R𝒜)\iota_{\mathcal{B}}^{-1}(R_{\mathcal{A}}). Since ϕ1𝒜=𝒜\phi^{-1}\cdot\mathcal{A}=\mathcal{A}, S𝒜=S[ϕk(𝒜)]S_{\mathcal{A}}=S[\phi^{k}(\mathcal{A})] is marked by an isomorphism α𝒜:ϕk(𝒜)π1(S𝒜)\alpha_{\mathcal{A}}^{\prime}:\phi^{k}(\mathcal{A})\to\pi_{1}(S_{\mathcal{A}}) and α𝒜\alpha_{\mathcal{A}}^{\prime} is the restriction of the marking α:ϕk()π1(S[ϕk()])\alpha_{\mathcal{B}}^{\prime}:\phi^{k}(\mathcal{B})\to\pi_{1}(S[\phi^{k}(\mathcal{B})]) to ϕk(𝒜)\phi^{k}(\mathcal{A}) with respect to the inclusion S𝒜S[ϕk()]S_{\mathcal{A}}\subset S[\phi^{k}(\mathcal{B})]. Therefore, S[ϕk()]S[\phi^{k}(\mathcal{B})] is also a vertex blow-up of ϕk()\Y,k\phi^{k}(\mathcal{B})\backslash Y_{*,k}. The noncontractible components of the subgraph ι1(R𝒜)\iota_{\mathcal{B}}^{-1}(R_{\mathcal{A}}) will be known as the lower stratum and the rest of the graph as the top stratum.

Now suppose 𝒜\mathcal{A}\prec\mathcal{B} are also [ϕ][\phi]-invariant and let f:TTf_{*}:T_{*}\to T_{*} be a 𝒜\mathcal{A}-relative representative for ϕ|\left.\phi\right|_{\mathcal{B}} defined on some (,𝒜)(\mathcal{B},\mathcal{A})-forest TT_{*} and f𝒜:R𝒜R𝒜f_{\mathcal{A}}:R_{\mathcal{A}}\to R_{\mathcal{A}} be a topological representative for ϕ|𝒜\left.\phi\right|_{\mathcal{A}}. Construct a topological representative f:ΓΓf_{\mathcal{B}}:\Gamma_{\mathcal{B}}\to\Gamma_{\mathcal{B}} for [ϕ|][\left.\phi\right|_{\mathcal{B}}\,] that agrees with f𝒜f_{\mathcal{A}} on the 𝒜\mathcal{A}-marked roses R𝒜R_{\mathcal{A}} and induces ff_{*} on the Bass-Serre forest TT_{*} upon collapsing the roses R𝒜R_{\mathcal{A}}.

For any k1k\geq 1, we let g,k:Y,kY,kg_{*,k}:Y_{*,k}\to Y_{*,k} be the minimal ϕk(𝒜)\phi^{k}(\mathcal{A})-relative representative for ϕ|ϕk()\left.\phi\right|_{\phi^{k}(\mathcal{B})} given by Proposition 3.5 using ff_{*} and f¯k:S[ϕk()]S[ϕk()]\bar{f}_{k}:S[\phi^{k}(\mathcal{B})]\to S[\phi^{k}(\mathcal{B})] be the natural representative for [ϕ|ϕk()][\left.\phi\right|_{\phi^{k}(\mathcal{B})}\,] given by Proposition 1.7 using ff_{\mathcal{B}}. By Proposition 3.5(3), if an element bb in \mathcal{B} has an ff_{*}-legal axis, then ϕk(b)\phi^{k}(b) has a g,kg_{*,k}-legal axis. It can be arranged for S[ϕk(𝒜)]S[ϕk()]S[\phi^{k}(\mathcal{A})]\subset S[\phi^{k}(\mathcal{B})] to be f¯k\bar{f}_{k}-invariant and f¯k\bar{f}_{k} to induce g,kg_{*,k} on the (ϕk(),ϕk(𝒜))(\phi^{k}(\mathcal{B}),\phi^{k}(\mathcal{A}))-forest Y,kY_{*,k} upon collapsing a maximal invariant proper subgraph of S[ϕk()]S[\phi^{k}(\mathcal{B})] containing S[ϕk(𝒜)]S[\phi^{k}(\mathcal{A})] and forgetting bivalent vertices.

Here is the idea behind the proof. By irreducibility of the restriction ϕ|\left.\phi\right|_{\mathcal{B}}, we may assume the map g,kg_{*,k} is an expanding irreducible representative for ϕ|ϕk()\left.\phi\right|_{\phi^{k}(\mathcal{B})}. For the contrapositive, suppose the forests T(ϕk())T_{*}(\phi^{k}(\mathcal{B})) had uniformly bounded natural edges. There is a sequence of loxodromic elements bkb_{k} in ϕk()\phi^{k}(\mathcal{B}) with uniformly bounded translation lengths lT(gk)l_{T_{*}}(g_{k}). Now suppose that the vertex blow-up S[ϕk()]S[\phi^{k}(\mathcal{B})] had natural edges with aribtrarily long top stratum subpaths. Bounded cancellation, the fact f¯k\bar{f}_{k} induces g,kg_{*,k}, and the irreducibility of g,kg_{*,k} imply g,kg_{*,k} is an expanding irreducible immersion. However, this contradicts the first assumption since lY,klT(ϕk())l_{Y_{*,k}}\leq l_{T_{*}(\phi^{k}(\mathcal{B}))}. So the second supposition is false and the natural edges of S[ϕk()]S[\phi^{k}(\mathcal{B})] have top stratum subpaths with uniformly bounded length. Using the Lipschitz property, expanding irreducibility of g,kg_{*,k}, and existence of a g,kg_{*,k}-legal axis (train track theory), we find uniformly bounded lower stratum paths in S[ϕk()]S[\phi^{k}(\mathcal{B})] connecting the origin of any oriented top stratum subpath of a natural edge to another top stratum subpath of a natural edge. Consequently, we are able to build uniformly bounded immersed loops in S[ϕk()]S[\phi^{k}(\mathcal{B})] that contain top stratum subpaths. This implies some loxodromic conjugacy class in \mathcal{B} has an infinite [ϕ][\phi]-tail. By Proposition 2.3, any conjugacy class in FF with an infinite [ϕ][\phi]-tail is carried by the maximal [ϕ][\phi]-fixed free factor system. Thus the maximal [ϕ][\phi]-fixed free factor system cannot be carried by 𝒜\mathcal{A} as it carries a loxodromic conjugacy class.

Proof of Proposition 3.9.

Suppose 𝒜\mathcal{A}\prec\mathcal{B} are [ϕ][\phi]-invariant free factor systems, ϕ|\left.\phi\right|_{\mathcal{B}} is irreducible relative to 𝒜\mathcal{A}, f𝒜:R𝒜R𝒜f_{\mathcal{A}}:R_{\mathcal{A}}\to R_{\mathcal{A}} is a topological representative for ϕ|𝒜\left.\phi\right|_{\mathcal{A}} defined on 𝒜\mathcal{A}-marked roses (R𝒜,α𝒜)(R_{\mathcal{A}},\alpha_{\mathcal{A}}), and f:TTf_{*}:T_{*}\to T_{*} is an expanding irreducible 𝒜\mathcal{A}-relative representative for ϕ|\left.\phi\right|_{\mathcal{B}} with the minimal stretch factor λ(f)>1\lambda(f_{*})>1 (Proposition 3.7). By Lemma 3.8, there is an element bb in \mathcal{B} with an ff_{*}-legal axis. Set (Γ,α)(\Gamma_{\mathcal{B}},\alpha_{\mathcal{B}}) to be the vertex blow-up of \T\mathcal{B}\backslash T_{*} with respect to the 𝒜\mathcal{A}-marked roses (R𝒜,α𝒜)(R_{\mathcal{A}},\alpha_{\mathcal{A}}). The discussion preceding the proof gives minimal ϕk(𝒜)\phi^{k}(\mathcal{A})-relative representatives g,k:Y,kY,kg_{*,k}:Y_{*,k}\to Y_{*,k} for ϕ|ϕk()\left.\phi\right|_{\phi^{k}(\mathcal{B})} and natural representatives f¯k:S[ϕk()]S[ϕk()]\bar{f}_{k}:S[\phi^{k}(\mathcal{B})]\to S[\phi^{k}(\mathcal{B})] for [ϕ|ϕk()][\left.\phi\right|_{\phi^{k}(\mathcal{B})}\,] that have these properties: for all k1k\geq 1,

  1. 1.

    f¯k\bar{f}_{k} induces g,kg_{*,k} on Y,kY_{*,k} upon collapsing the ff_{\mathcal{B}}-invariant subgraph R𝒜ΓR_{\mathcal{A}}\subset\Gamma_{\mathcal{B}};

  2. 2.

    K=K(f¯k)=K(f)+C(f)K=K(\bar{f}_{k})=K(f_{\mathcal{B}})+C(f_{\mathcal{B}}) and C=C(f¯k)=2C(f)C=C(\bar{f}_{k})=2C(f_{\mathcal{B}});

  3. 3.

    lY,k:ϕk()l_{Y_{*,k}}:\phi^{k}(\mathcal{B})\to\mathbb{R} is dominated by the restrictions lT|ϕk()=lT(ϕk())\left.l_{T_{*}}\right|_{\phi^{k}(\mathcal{B})}=l_{T_{*}(\phi^{k}(\mathcal{B}))}; and

  4. 4.

    ϕk(b)\phi^{k}(b) has a g,kg_{*,k}-legal axis.

The collection ϕ|ϕk()\left.\phi\right|_{\phi^{k}(\mathcal{B})} is conjugate to ϕ|\left.\phi\right|_{\mathcal{B}} by injectivity of ϕ\phi. So ϕ|ϕk()\left.\phi\right|_{\phi^{k}(\mathcal{B})} is irreducible relative to ϕk(𝒜)\phi^{k}(\mathcal{A}) and λ(f)\lambda(f_{*}) is the minimal stretch factor for ϕ|ϕk()\left.\phi\right|_{\phi^{k}(\mathcal{B})} relative to ϕk(𝒜)\phi^{k}(\mathcal{A}). Furthermore, the minimal ϕk(𝒜)\phi^{k}(\mathcal{A})-relative representatives g,kg_{*,k} are irreducible (Lemma 3.6) and λ(g,k)λ(f)>1\lambda(g_{*,k})\geq\lambda(f_{*})>1 by the minimality of λ(f)\lambda(f_{*}).

Suppose for the contrapositive that there is a bound L1L\geq 1 such that all natural edges in T(ϕk())T_{*}(\phi^{k}(\mathcal{B})) are shorter than LL for all k1k\geq 1. Then, for all k1k\geq 1, there is a loxodromic element bkb_{k} in ϕk()\phi^{k}(\mathcal{B}) such that lT(bk)(3N3)Ll_{T_{*}}(b_{k})\leq(3N-3)L, where N=3rank(F)3N=3\cdot\mathrm{rank}(F)-3. Every edge EE in Γ,k=ϕk()\Y,k\Gamma_{*,k}=\phi^{k}(\mathcal{B})\backslash Y_{*,k} lifts to a ϕk()\phi^{k}(\mathcal{B})-orbit of a natural edge EE^{\prime} in ϕk()\T(ϕk())\phi^{k}(\mathcal{B})\backslash T_{*}(\phi^{k}(\mathcal{B})), which corresponds to a top stratum subpath E¯\bar{E} of a natural edge in S[ϕk()]S[\phi^{k}(\mathcal{B})].

Claim.

The subpath E¯\bar{E} in S[ϕk()]S[\phi^{k}(\mathcal{B})] has length CKN1\leq C{\cdot}K^{N-1} for all edges EE in Γ,k\Gamma_{*,k} and k1k\geq 1.

Suppose, the graph Γ,k\Gamma_{*,k} has an edge E0E_{0} whose corresponding subpath E¯0\bar{E}_{0} in S[ϕk()]S[\phi^{k}(\mathcal{B})] is longer than CKN1\displaystyle C\cdot K^{N-1} for some k1k\geq 1. As we did in the proof of Theorem 2.1, we construct the set of long edges \mathcal{E} by looking at all the edges of Γ,k\Gamma_{*,k} that are eventually mapped over E0E_{0}. Here, an edge E1E_{1} in Γ,k\Gamma_{*,k} mapped over E0E_{0} if there are lifts E1E_{1}^{\prime} and E0E_{0}^{\prime} in Y,kY_{*,k} such that g,kg_{*,k} maps E1E_{1}^{\prime} over E0E_{0}^{\prime}. Since f¯k:S[ϕk()]S[ϕk()]\bar{f}_{k}:S[\phi^{k}(\mathcal{B})]\to S[\phi^{k}(\mathcal{B})] is KK-Lipschitz and it induces g,kg_{*,k} on Y,kY_{*,k}, each long edge in Γ,k\Gamma_{*,k} corresponds to a top statrum subpath in S[ϕk()]S[\phi^{k}(\mathcal{B})] longer than CC. Since g,kg_{*,k} is an irreducible ϕk(𝒜)\phi^{k}(\mathcal{A})-relative representative, all edges eventually map over E0E_{0} and hence are long. The long natural edges of S[ϕk()]S[\phi^{k}(\mathcal{B})] will be the natural edges in S[ϕk()]S[\phi^{k}(\mathcal{B})] containing top stratum subpaths.

Suppose an edge EE of Γ,k\Gamma_{*,k} had a lift EE^{\prime} in Y,kY_{*,k} that is the initial segment of the g,kg_{*,k}-image of two edges that share an initial vertex. Then the top stratum subpath E¯\bar{E} is in a long natural edge of S[ϕk()]S[\phi^{k}(\mathcal{B})] that is the initial segment of f¯k\bar{f}_{k}-images of natural edges that share an initial vertex; this violates bounded cancellation since long natural edges of S[ϕk()]S[\phi^{k}(\mathcal{B})] longer than C=C(f¯k)C=C(\bar{f}_{k}). Hence, there is no folding in g,kg_{*,k}, i.e., g,kg_{*,k} is an expanding irreducible ϕk(𝒜)\phi^{k}(\mathcal{A})-relative immersion. We may now find an m1m\geq 1 such that all loxodromic elements bb in ϕk()\phi^{k}(\mathcal{B}) have lY,k(ϕm(b))>(3N3)Ll_{Y_{*,k}}(\phi^{m}(b))>(3N-3)L. Since lY,kl_{Y_{*,k}} is dominated by lT|ϕk()\left.l_{T_{*}}\right|_{\phi^{k}(\mathcal{B})}, we get that lT(b)>(3N3)Ll_{T_{*}}(b^{\prime})>(3N-3)L for all loxodromic elements bb^{\prime} in ϕk+m()\phi^{k+m}(\mathcal{B}). This contradicts the assumption that lT(bk+m)(3N3)Ll_{T_{*}}(b_{k+m})\leq(3N-3)L for some loxodromic bk+mb_{k+m} in ϕk+m()\phi^{k+m}(\mathcal{B}). So the top stratum subpath E¯\bar{E} in S[ϕk()]S[\phi^{k}(\mathcal{B})] has length CKN1\leq C\cdot K^{N-1} for all natural edges EE of Γ,k\Gamma_{*,k} and k1k\geq 1. This ends the proof of the claim.

Next, we prove the existence of paths in the lower stratum of S[ϕk()]S[\phi^{k}(\mathcal{B})] with uniformly bounded lengths connecting top stratum paths. Suppose E0,E1,E_{0},E_{1}, and E2E_{2} are edges of Γ,k\Gamma_{*,k} with lifts E0,E1,E_{0}^{\prime},E_{1}^{\prime}, and E2E_{2}^{\prime} in Y,kY_{*,k} such that E1E2E_{1}^{\prime}\cdot E_{2}^{\prime} is a subpath of the immersed path g,k(E0)g_{*,k}(E_{0}^{\prime}). Then E¯1P12E¯2\bar{E}_{1}\cdot P_{12}\cdot\bar{E}_{2} is a subpath of immersed path f¯k(E¯0)\bar{f}_{k}(\bar{E}_{0}) for some lower stratum path P12P_{12} in S[ϕk()]S[\phi^{k}(\mathcal{B})]. Since E¯0\bar{E}_{0} has length bounded by CKN1C\cdot K^{N-1} and f¯k\bar{f}_{k} is KK-Lipschitz, the path P12P_{12} has length bounded by CKNC\cdot K^{N}. We say the 2-edge path E1E2E_{1}\cdot E_{2} in Γ,k\Gamma_{*,k} has a nondegenerate turn bounded by CKNC\cdot K^{N}.

As g,kg_{*,k} is an expanding irreducible relative representative that has a legal axis (this is where the argument invokes train track theory), every edge EE^{\prime} in Y,kY_{*,k} can be extended to an immersed 3-edge path EEE+E_{-}^{\prime}\cdot E^{\prime}\cdot E_{+}^{\prime} that is a translate of a subpath of g,kn(E)g_{*,k}^{n}(E^{\prime}) and n2N!n\leq 2{\cdot}N!. In particular, any edge in Γ,k\Gamma_{*,k} can be extended to a 3-edge path whose 2-edge subpaths both have nondegenerate turns bounded by CKN1K2N!C\cdot K^{N-1}\cdot K^{2{\cdot}N!}, i.e., every top stratum subpath E¯\bar{E} can be extended to an immersed path E¯PE¯P+E¯+\bar{E}_{-}{\cdot}P_{-}{\cdot}\bar{E}{\cdot}P_{+}{\cdot}\bar{E}_{+} with top stratum subpaths E¯\bar{E}_{-}, E¯+\bar{E}_{+} and lower stratum paths PP_{-}, P+P_{+} with length bounded by CKN1K2N!C\cdot K^{N-1}\cdot K^{2{\cdot}N!}.

Using this bound on lower stratum paths and the bound on top stratum subpaths given by the claim, we can now form an immersed loop ρk\rho_{k} in Γ,k\Gamma_{*,k} with the properties:

  1. 1.

    ρk\rho_{k} lifts to an axis in Y,kY_{*,k} for some loxodromic conjugacy class [bk][b_{k}^{\prime}];

  2. 2.

    ρk\rho_{k} passes any edge of Γ,k\Gamma_{*,k} at most twice and only takes short turns (including the turn at the endpoint), which implies it has at most 2N2N edges and (short) turns; and

  3. 3.

    ρk\rho_{k} represents a loop in S[ϕk()]S[\phi^{k}(\mathcal{B})] with length bounded by 2NC(1+K2N!)KN12N{\cdot}C(1+K^{2{\cdot}N!})K^{N-1}.

In summary, for each k1k\geq 1, we construct a loxodromic conjugacy class [bk][b_{k}^{\prime}] in ϕk()\phi^{k}(\mathcal{B}) whose α\alpha_{\mathcal{B}}-length is bounded by a constant independent of kk. As there are finitely many conjugacy classes with α\alpha_{\mathcal{B}}-length bounded by any given constant, the sequence of conjugacy classes [bk]k=1[b_{k}^{\prime}]_{k=1}^{\infty} has a constant infinite subsequence. Thus, some loxodromic conjugacy class [b][b^{\prime}] has an infinite [ϕ][\phi]-tail carried by \mathcal{B}. Recall that the maximal [ϕ][\phi]-fixed free factor system carries all conjugacy classes with an infinite [ϕ][\phi]-tail (Proposition 2.3); on the other hand, 𝒜\mathcal{A} does not carry loxodromic conjugacy classes in \mathcal{B} by definition. Therefore, 𝒜\mathcal{A} cannot carry the maximal [ϕ][\phi]-fixed free factor system. ∎

4 Canonical expanding relative immersions

The main result of this section is the existence of expanding immersions for nonsurjective endomorphisms relative to their elliptic free factor systems.

Let ϕ:FF\phi:F\to F be an injective endomorphism, 𝒜\mathcal{A}\prec\mathcal{B} be a pair of [ϕ][\phi]-invariant free factor systems such that ϕ1𝒜=𝒜\phi^{-1}\cdot\mathcal{A}=\mathcal{A}, and ϕ|:\left.\phi\right|_{\mathcal{B}}:\mathcal{B}\to\mathcal{B} be a restriction of ϕ\phi to \mathcal{B}. Recall that a relative representative is minimal if it has no orbit-closed invariant subforests with bounded components and an expanding 𝒜{\mathcal{A}}-relative immersion for ϕ|𝓑\boldsymbol{\left.\phi\right|_{\mathcal{B}}} is an 𝒜\mathcal{A}-relative representative f:TTf_{*}:T_{*}\to T_{*} for ϕ|\left.\phi\right|_{\mathcal{B}} that is a minimal immersion whose edges expand under ff_{*}-iteration.

There will be two possible ways of obtaining a relative immersion from a relatively irreducible restriction with a minimal stretch factor λ\lambda. If λ=1\lambda=1, then an irreducible representative with stretch factor λ\lambda is automatically a simplicial immersion. The next proposition shows how to construct an immersion when λ>1\lambda>1. This construction is unique to nonsurjective endomorphisms because we require that the restriction be irreducible relative to a free factor system that carries the [ϕ][\phi]-elliptic free factor system — when ϕ\phi is an automorphism, the [ϕ][\phi]-elliptic free factor system is {F}\{F\} and no such restriction exists.

Proposition 4.1.

Let ϕ:FF\phi:F\to F be injective and 𝒜\mathcal{A}\prec\mathcal{B} be a chain of [ϕ][\phi]-invariant free factor systems that carry the [ϕ][\phi]-elliptic free factor system. If ϕ|\left.\phi\right|_{\mathcal{B}} is irreducible relative to 𝒜\mathcal{A} and λ([ϕ],,𝒜)>1\lambda([\phi],\mathcal{B},\mathcal{A})>1, then there is an expanding irreducible 𝒜\mathcal{A}-relative immersion for ϕ|\left.\phi\right|_{\mathcal{B}}.

Proof.

Suppose ϕ:FF\phi:F\to F is injective, 𝒜\mathcal{A}\prec\mathcal{B} are [ϕ][\phi]-invariant free factor systems that carry the [ϕ][\phi]-elliptic free factor system, ϕ|\left.\phi\right|_{\mathcal{B}} is irreducible relative to 𝒜\mathcal{A}, and f:TTf_{*}:T_{*}\to T_{*} is an expanding irreducible (,𝒜)(\mathcal{B},\mathcal{A})-relative representative for ϕ|\left.\phi\right|_{\mathcal{B}} with minimal stretch factor λ(f)>1\lambda(f_{*})>1. Set K=K(f)+C(f)K=K(f_{*})+C(f_{*}) and C=2C(f)C=2C(f_{*}). By Proposition 3.4, there is a natural representative f,k:T(ϕk())T(ϕk())f_{*,k}:T_{*}(\phi^{k}(\mathcal{B}))\to T_{*}(\phi^{k}(\mathcal{B})) for ϕ|ϕk(B)\left.\phi\right|_{\phi^{k}(B)} with Lipschitz and cancellation constants K(f,k)=KK(f_{*,k})=K and C(f,k)=CC(f_{*,k})=C respectively for all k1k\geq 1.

The first part of the proof proceeds as a relativized version of the proof of Theorem 2.1. By Proposition 3.9, we may fix k0k\gg 0 such that the set of natural edges 0\mathcal{L}_{0} in T(ϕk())T_{*}(\phi^{k}(\mathcal{B})) longer than CKN1C\cdot K^{N-1} is not empty, where N=3rank(F)3N=3\cdot\mathrm{rank}(F)-3. Choose \mathcal{L} to be the set of all natural edges that eventually get mapped over those in 0\mathcal{L}_{0} by f,kf_{*,k} and call \mathcal{L} the long natural edges. As f,kf_{*,k} is KK-Lipschitz and there are at most NN orbits of natural edges in T(ϕk())T_{*}(\phi^{k}(\mathcal{B})), the long natural edges are longer that CC. Injectivity of ϕ\phi implies ϕ|ϕk()\left.\phi\right|_{\phi^{k}(\mathcal{B})} is conjugate to ϕ|\left.\phi\right|_{\mathcal{B}}. So ϕ|ϕk()\left.\phi\right|_{\phi^{k}(\mathcal{B})} is irreducible relative to ϕk(𝒜)\phi^{k}(\mathcal{A}), λ(f)\lambda(f_{*}) is the minimal stretch factor for ϕ|ϕk()\left.\phi\right|_{\phi^{k}(\mathcal{B})} relative to ϕk(𝒜)\phi^{k}(\mathcal{A}), and the short natural edges of T(ϕk())T_{*}(\phi^{k}(\mathcal{B})) form an orbit-closed f,kf_{*,k}-invariant subforest with bounded components.

Collapse a maximal f,kf_{*,k}-invariant subforest of T(ϕk())T_{*}(\phi^{k}(\mathcal{B})) that has bounded components and contains the short natural edges then forget the bivalent vertices; this induces a minimal ϕk(𝒜)\phi^{k}(\mathcal{A})-relative representative g,k:Y,kY,kg_{*,k}:Y_{*,k}\to Y_{*,k} for ϕ|ϕk()\left.\phi\right|_{\phi^{k}(\mathcal{B})}. The map g,kg_{*,k} is an irreducible ϕk(𝒜)\phi^{k}(\mathcal{A})-relative representative for ϕ|ϕk()\left.\phi\right|_{\phi^{k}(\mathcal{B})} (Lemma 3.6) and λ(g,k)λ(f)\lambda(g_{*,k})\geq\lambda(f_{*}) by the minimality of λ(f)\lambda(f_{*}). So g,kg_{*,k} is an expanding irreducible ϕk(𝒜)\phi^{k}(\mathcal{A})-relative representative.

Since the lifts in T(ϕk())T_{*}(\phi^{k}(\mathcal{B})) of all edges in Y,kY_{*,k} are longer than the cancellation constant CC, there is no folding in g,kg_{*,k} — otherwise, there would be folding in f,kf_{*,k} identifying paths longer than its cancellation constant, absurd. Thus, g,kg_{*,k} is an immersion. By injectivity of ϕ\phi, we can view Y,kY_{*,k} as a (,𝒜)(\mathcal{B},\mathcal{A})-forest and g,kg_{*,k} as an expanding irreducible 𝒜\mathcal{A}-relative immersion for ϕ|\left.\phi\right|_{\mathcal{B}}. ∎

We are now ready to state and prove our base case for the construction. In light of the previous proposition, the point is that a restriction ϕ|\left.\phi\right|_{\mathcal{B}} that is irreducible relative to the [ϕ][\phi]-elliptic free factor system 𝒜\mathcal{A} will satisfy λ([ϕ],,𝒜)>1\lambda([\phi],\mathcal{B},\mathcal{A})>1.

Proposition 4.2.

Let ϕ:FF\phi:F\to F be injective and 𝒜\mathcal{A}\prec\mathcal{B} be a chain of [ϕ][\phi]-invariant free factor systems where 𝒜\mathcal{A} is the [ϕ][\phi]-elliptic free factor system. If ϕ|\left.\phi\right|_{\mathcal{B}} is irreducible relative to 𝒜\mathcal{A}, then there is an expanding irreducible 𝒜\mathcal{A}-relative immersion for ϕ|\left.\phi\right|_{\mathcal{B}}.

Proof.

Let ϕ:FF\phi:F\to F be injective and ϕ|\left.\phi\right|_{\mathcal{B}} be irreducible relative to the [ϕ][\phi]-elliptic free factor system 𝒜\mathcal{A}. Then there is an irreducible 𝒜\mathcal{A}-relative representative f:TTf_{*}:T_{*}\to T_{*} for ϕ|\left.\phi\right|_{\mathcal{B}} with stretch factor λ(f)=λ([ϕ],,𝒜)1\lambda(f_{*})=\lambda([\phi],\mathcal{B},\mathcal{A})\geq 1 (Proposition 3.7). We say BiB_{i}\in\mathcal{B} is loxodromic if TiTT_{i}\in T_{*} is not a point, i.e., BiB_{i} contains a loxodromic element; similarly, the component Bi\TiB_{i}\backslash T_{i} of the graph of groups \T\mathcal{B}\backslash T_{*} is loxodromic if BiB_{i} is loxodromic. If λ(f)=1\lambda(f_{*})=1, then the induced map on the loxodromic components of \T\mathcal{B}\backslash T_{*} is a graph isomorphism. So for some k1k\geq 1, if A𝒜A\in\mathcal{A} is carried by a loxodromic BB\in\mathcal{B}, then AA is ϕk\phi^{k}-invariant. By Proposition 2.4, the subset of all A𝒜A\in\mathcal{A} carried by the loxodromic component of \mathcal{B} form a [ϕ][\phi]-fixed free factor subsystem. As ff_{*} induces a graph isomorphism on the loxodromic components of \T\mathcal{B}\backslash T_{*} and these components’ vertex groups form a [ϕ][\phi]-fixed free factor system, we get that ff_{*} is surjective when restricted to the unbounded components of the forest TT_{*} and the loxodromic components of \mathcal{B} form a [ϕ][\phi]-fixed free factor system. This is a contradiction since [ϕ][\phi]-periodic free factors are elliptic (Propositions 2.3 and 2.4). Therefore, λ(f)>1\lambda(f_{*})>1 and the result follows from Proposition 4.1. ∎

Specializing this proposition to the case where ϕ\phi is irreducible and nonsurjective yields an alternate proof to a result due to Reynolds.

Corollary 4.3 ([27, Corollary 3.23]).

If ϕ:FF\phi:F\to F is irreducible but not surjective, then ϕ\phi is induced by an expanding irreducible graph immersion.

Remark.

This proof of Reynolds’ result is a variation of our previous proof [24, Theorem 4.5] with two crucial differences: 1) it makes no use of limit trees in the compactification of outer space; 2) the specialization of Proposition 4.1 to irreducible nonsurjective endomorphisms need not invoke train track theory since we can use Lemma 1.1 in place of Proposition 3.9.

The next proposition is the induction step for our construction.

Proposition 4.4.

Let ϕ:FF\phi:F\to F be injective, 𝒜\mathcal{A} be the [ϕ][\phi]-elliptic free factor system, and 𝒜𝒞\mathcal{A}\prec\mathcal{B}\prec\mathcal{C} be a chain of [ϕ][\phi]-invariant free factor systems. If there is an expanding 𝒜\mathcal{A}-relative immersion for ϕ|\left.\phi\right|_{\mathcal{B}} and a \mathcal{B}-relative immersion for ϕ|𝒞\left.\phi\right|_{\mathcal{C}}, then there is an expanding 𝒜\mathcal{A}-relative immersion for ϕ|𝒞\left.\phi\right|_{\mathcal{C}}.

Although the proof gets a bit technical, the idea is rather simple: a \mathcal{B}-relative immersion for ϕ|𝒞\left.\phi\right|_{\mathcal{C}} (top stratum) and an expanding 𝒜\mathcal{A}-relative immersion for ϕ|\left.\phi\right|_{\mathcal{B}} (lower stratum) can be patched together via a (relative) vertex blow-up to get a minimal 𝒜\mathcal{A}-relative representative g:YYg_{*}:Y_{*}\to Y_{*} whose only possible folds would have to happen between a top and lower stratum edge of YY_{*}. As the restriction of gg_{*} to the lower stratum is an expanding immersion, we may assume the edges in the lower stratum are longer than the cancellation constant. This means no lower stratum edge is identified by a fold and so no folding in gg_{*} is possible. Thus, gg_{*} is a minimal 𝒜\mathcal{A}-relative immersion for ϕ|𝒞\left.\phi\right|_{\mathcal{C}}, which will be expanding if 𝒜\mathcal{A} is the [ϕ][\phi]-elliptic free factor system.

Let T𝒞T_{\mathcal{C}} be a (𝒞,)(\mathcal{C},\mathcal{B})-forest and TT_{\mathcal{B}} be a (,𝒜)(\mathcal{B},\mathcal{A})-forest. For any free factor Ci𝒞C_{i}\in\mathcal{C}, let i\mathcal{B}_{i} the maximal subset of \mathcal{B} that is carried by CiC_{i}. Replace the free factors of \mathcal{B} with conjugates if necessary and assume \mathcal{B} is also a free factor system of 𝒞\mathcal{C}. In particular, the free factors BiB\in\mathcal{B}_{i} are subgroups of the free factors Ci𝒞C_{i}\in\mathcal{C}. Identifying the appropriate vertices of the graph of groups 𝒞\T𝒞\mathcal{C}\backslash T_{\mathcal{C}} with basepoints on the graph of groups \T\mathcal{B}\backslash T_{\mathcal{B}} results in a graph of groups decomposition for 𝒞\mathcal{C} whose Bass-Serre forest TT_{*} is a (𝒞,𝒜)(\mathcal{C},\mathcal{A})-forest that contains TT_{\mathcal{B}}. We call TT_{*} the vertex blow-up of T𝒞T_{\mathcal{C}} with respect to TT_{\mathcal{B}}.

Proof of Proposition 4.4.

Let ϕ:FF\phi:F\to F be injective, 𝒞\mathcal{B}\prec\mathcal{C} be a chain of [ϕ][\phi]-invariant free factor systems that carry the [ϕ][\phi]-elliptic free factor system 𝒜\mathcal{A}. Suppose f:TTf_{\mathcal{B}}:T_{\mathcal{B}}\to T_{\mathcal{B}} is an expanding 𝒜\mathcal{A}-relative immersion for ϕ|\left.\phi\right|_{\mathcal{B}} and f𝒞:T𝒞T𝒞f_{\mathcal{C}}:T_{\mathcal{C}}\to T_{\mathcal{C}} is a \mathcal{B}-relative immersion for ϕ|𝒞\left.\phi\right|_{\mathcal{C}} then define TT_{*} to be the vertex blow-up of T𝒞T_{\mathcal{C}} with respect to TT_{\mathcal{B}}. The edges of TT_{*} are of two types: the lower stratum, which are edges that are contained in the 𝒞\mathcal{C}-orbit of TT_{\mathcal{B}}, and the top stratum, which are the remaining edges.

Let f:TTf_{*}:T_{*}\to T_{*} be a minimal 𝒜\mathcal{A}-relative representative for ϕ|𝒞\left.\phi\right|_{\mathcal{C}} such that TT_{\mathcal{B}} is an ff_{*}-invariant subforest, the restriction of ff_{*} to TT_{\mathcal{B}} agrees with ff_{\mathcal{B}}, and ff_{*} induces f𝒞f_{\mathcal{C}} upon collapsing the lower stratum. For all k1k\geq 1, set T(ϕk(𝒞))T_{*}(\phi^{k}(\mathcal{C})) and T(ϕk())T_{*}(\phi^{k}(\mathcal{B})) to be the minimal subforests of TT_{*} for ϕk(𝒞)\phi^{k}(\mathcal{C}) and ϕk()\phi^{k}(\mathcal{B}) respectively. Similarly, define the minimal subforest T(ϕk())TT_{\mathcal{B}}(\phi^{k}(\mathcal{B}))\subset T_{\mathcal{B}}. By the inclusion of TT_{\mathcal{B}} in TT_{*}, we get a simplicial identification of T(ϕk())T_{*}(\phi^{k}(\mathcal{B})) with T(ϕk())T_{\mathcal{B}}(\phi^{k}(\mathcal{B})). However, we want to consider these two forests differently with respect to their branch points and natural edges. In particular, there may be branch points of T(ϕk(𝒞))T_{*}(\phi^{k}(\mathcal{C})) that are bivalent when considered as points on the subforest T(ϕk())T_{*}(\phi^{k}(\mathcal{B})). So by “natural edges of T(ϕk())T_{*}(\phi^{k}(\mathcal{B})), we mean those inherited from the parent forest T(ϕk(𝒞))T_{*}(\phi^{k}(\mathcal{C})); on the other hand, by “natural edges of T(ϕk())T_{\mathcal{B}}(\phi^{k}(\mathcal{B})), we do mean exactly that. Under the identification of the two forests, the natural edges of T(ϕk())T_{*}(\phi^{k}(\mathcal{B})) partition any natural edge of T(ϕk())T_{\mathcal{B}}(\phi^{k}(\mathcal{B})) into at most 2𝒳2\mathcal{X} segments, where 𝒳=rank(F)1\mathcal{X}=\mathrm{rank}(F)-1.

Since f𝒞:T𝒞T𝒞f_{\mathcal{C}}:T_{\mathcal{C}}\to T_{\mathcal{C}} is a \mathcal{B}-relative immersion for ϕ|𝒞\left.\phi\right|_{\mathcal{C}}, the restrictions of f𝒞f_{\mathcal{C}} to T𝒞(ϕk(𝒞))T_{\mathcal{C}}(\phi^{k}(\mathcal{C})) are ϕk()\phi^{k}(\mathcal{B})-relative immersions f𝒞,k:T𝒞(ϕk(𝒞))T𝒞(ϕk(𝒞))f_{\mathcal{C},k}:T_{\mathcal{C}}(\phi^{k}(\mathcal{C}))\to T_{\mathcal{C}}(\phi^{k}(\mathcal{C})) for ϕ|ϕk(𝒞)\left.\phi\right|_{\phi^{k}(\mathcal{C})} that are conjugate to f𝒞f_{\mathcal{C}}. As ff_{*} induces f𝒞f_{\mathcal{C}} upon collapsing the lower stratum, any edges in f(T(ϕk(𝒞)))f_{*}(T_{*}(\phi^{k}(\mathcal{C}))) but not T(ϕk(𝒞))T_{*}(\phi^{k}(\mathcal{C})) must be in the lower stratum and the restriction of ff_{*} to T(ϕk(𝒞))T_{*}(\phi^{k}(\mathcal{C})) induces f𝒞,kf_{\mathcal{C},k} upon collapsing the lower stratum, i.e., the ϕk(𝒞)\phi^{k}(\mathcal{C})-orbit of T(ϕk())T_{*}(\phi^{k}(\mathcal{B})). By Proposition 3.4, there is an ϕk(𝒜)\phi^{k}(\mathcal{A})-relative natural representative f,k:T(ϕk(𝒞))T(ϕk(𝒞))f_{*,k}:T_{*}(\phi^{k}(\mathcal{C}))\to T_{*}(\phi^{k}(\mathcal{C})) for ϕ|ϕk(𝒞)\left.\phi\right|_{\phi^{k}(\mathcal{C})} with Lipschitz and cancellation constants K=K(f)+C(f)K=K(f_{*})+C(f_{*}) and C=2C(f)C=2C(f_{*}) respectively. Furthermore, T(ϕk())T_{*}(\phi^{k}(\mathcal{B})) is an f,kf_{*,k}-invariant subforest and f,kf_{*,k} still induces f𝒞,kf_{\mathcal{C},k} upon collapsing the lower stratum. As ff_{*} agrees with the immersion ff_{\mathcal{B}} on TT_{\mathcal{B}}, f,kf_{*,k} differs from ff_{\mathcal{B}} on T(ϕk())T_{*}(\phi^{k}(\mathcal{B})) by a homotopy supported in the natural edges of T(ϕk())T_{\mathcal{B}}(\phi^{k}(\mathcal{B})).

Since f:TTf_{\mathcal{B}}:T_{\mathcal{B}}\to T_{\mathcal{B}} is an expanding 𝒜\mathcal{A}-relative immersion for ϕ|\left.\phi\right|_{\mathcal{B}}, the minimal subforest T(ϕk())T_{\mathcal{B}}(\phi^{k}(\mathcal{B})) has natural edges whose lengths are all exponential in kk. Fix k0k\gg 0 such that all natural edges of T(ϕk())T_{\mathcal{B}}(\phi^{k}(\mathcal{B})) are longer than 2𝒳CK3𝒳12\mathcal{X}\cdot C\cdot K^{3\mathcal{X}-1}. By the pigeonhole principle, each natural edge of T(ϕk())T_{\mathcal{B}}(\phi^{k}(\mathcal{B})) contains a natural edge of T(ϕk())T_{*}(\phi^{k}(\mathcal{B})) longer than CK3𝒳C\cdot K^{3\mathcal{X}}. Let 𝔾k\mathbb{G}_{k} be the directed graph of natural edges of the f,kf_{*,k}-invariant subforest T(ϕk())T_{*}(\phi^{k}(\mathcal{B})) where a directed edge EiEjE_{i}\to E_{j} corresponds to f,kf_{*,k} mapping EiE_{i} over EjE_{j}. Set 𝒮0\mathcal{S}_{0} to be those natural edges with length at most CC and 𝒮\mathcal{S} to be those natural edges with directed path from 𝒮0\mathcal{S}_{0} in 𝔾k\mathbb{G}_{k} and their ϕk(𝒞)\phi^{k}(\mathcal{C})-translates; these lower stratum natural edges will be the short natural edges of T(ϕk(𝒞))T_{*}(\phi^{k}(\mathcal{C})). Since f,kf_{*,k} is KK-Lipschitz and the shortest path between any two natural edges in 𝔾k\mathbb{G}_{k} has 3𝒳3\mathcal{X} natural edges, all the short natural edges have length at most CK3𝒳1C\cdot K^{3\mathcal{X}-1}. So the short natural edges 𝒮\mathcal{S} form an orbit-closed f,kf_{*,k}-invariant lower stratum subforest of T(ϕk(𝒞))T_{*}(\phi^{k}(\mathcal{C})) with bounded components as 𝒮\mathcal{S} does not cover any natural edge of T(ϕk())T_{\mathcal{B}}(\phi^{k}(\mathcal{B})).

Collapsing the short natural edges of T(ϕk())T_{*}(\phi^{k}(\mathcal{B})) induces a map g,k:Y,kY,kg_{*,k}^{\prime}:Y_{*,k}^{\prime}\to Y_{*,k}^{\prime} with the same cancellation constant C(g,k)=CC(g_{*,k}^{\prime})=C. Now iteratively collapse pretrivial edges until the induced map g,k:Y,kY,kg_{*,k}:Y_{*,k}\to Y_{*,k} has none. As the collapses are supported in the lower stratum, the new map g,kg_{*,k} still induces the immersion f𝒞,kf_{\mathcal{C},k} upon collapsing the rest of the lower stratum and, as a result, folding in g,kg_{*,k} may only occur between initial segments of natural edges of Y,kY_{*,k} whose g,kg_{*,k}-images are a lower stratum natural edge. However, all natural edges in the lower stratum of Y,kY_{*,k} are longer than CC by construction and so no folding in g,kg_{*,k} is possible by bounded cancellation, i.e., g,kg_{*,k} is an immersion.

Collapsing a maximal invariant subforest with bounded components and forgetting bivalent vertices if necessary, we may assume g,k:Y,kY,kg_{*,k}:Y_{*,k}\to Y_{*,k} is a minimal ϕk(𝒜)\phi^{k}(\mathcal{A})-relative immersion for ϕ|ϕk(𝒞)\left.\phi\right|_{\phi^{k}(\mathcal{C})}. By the injectivity of ϕ\phi, we can view g,kg_{*,k} as a minimal 𝒜\mathcal{A}-relative immersion for ϕ|𝒞\left.\phi\right|_{\mathcal{C}}. It remains to show that every edge of Y,kY_{*,k} expands under g,kg_{*,k}-iteration.

For a contradiction, suppose there is an edge of Y,kY_{*,k} whose g,kg_{*,k}-iterates have uniformly bounded length. Since g,kg_{*,k} is minimal, the non-expanding edges in the graph of groups 𝒞\Y,k\mathcal{C}\backslash Y_{*,k} contain a setwise fixed subgraph \mathcal{F} that carries a loxodromic element. The subgraph \mathcal{F} is a free splitting of a [ϕ][\phi]-invariant free factor system. Recall that the [ϕ][\phi]-elliptic free factor system 𝒜\mathcal{A} decomposes as a union of the maximal [ϕ][\phi]-fixed free factor system and free factors that eventually get mapped into this fixed system (Proposition 2.4). By construction, the point stabilizers of Y,kY_{*,k} are conjugates of 𝒜\mathcal{A}. So any vertex of the fixed graph \mathcal{F} is labelled by either the trivial group or a free factor of the maximal [ϕ][\phi]-fixed free factor system. Thus \mathcal{F} is a free splitting of a [ϕ][\phi]-fixed free factor system that carries some loxodromic element. However, Propositions 2.3 and 2.4 imply that all [ϕ][\phi]-fixed free factors systems are elliptic — a contradiction. ∎

We are now ready to inductively construct expanding relative immersions.

Theorem 4.5.

If ϕ:FF\phi:F\to F is injective but not surjective, then there is an expanding 𝒜\mathcal{A}-relative immersion for ϕ\phi, where 𝒜\mathcal{A} is the [ϕ][\phi]-elliptic free factor system.

Proof.

Suppose ϕ:FF\phi:F\to F is injective but not surjective. By Proposition 2.4, the [ϕ][\phi]-elliptic free factor system 𝒜\mathcal{A} is proper. The naive approach is to assume there exists a chain 𝒜=0n={F}\mathcal{A}=\mathcal{B}_{0}\prec\cdots\prec\mathcal{B}_{n}=\{F\} in the poset of [ϕ][\phi]-invariant free factor system such that the restrictions ϕ|m+1\left.\phi\right|_{\mathcal{B}_{m+1}} are irreducible relative to m\mathcal{B}_{m} for all m1m\geq 1. This assumption is typical when working with automorphisms. For each restriction ϕ|m+1\left.\phi\right|_{\mathcal{B}_{m+1}}, if the minimal stretch factor λ([ϕ],m+1,m)=1\lambda([\phi],\mathcal{B}_{m+1},\mathcal{B}_{m})=1, then there is automatically a m\mathcal{B}_{m}-relative immersion for ϕ|m+1\left.\phi\right|_{\mathcal{B}_{m+1}}; and if λ([ϕ],m+1,m)>1\lambda([\phi],\mathcal{B}_{m+1},\mathcal{B}_{m})>1, then there is an expanding m\mathcal{B}_{m}-relative immersion for ϕ|m+1\left.\phi\right|_{\mathcal{B}_{m+1}} by Proposition 4.1. In either case, there is a m\mathcal{B}_{m}-relative immersion for ϕ|m+1\left.\phi\right|_{\mathcal{B}_{m+1}}. By Proposition 4.2, λ([ϕ],1,0)>1\lambda([\phi],\mathcal{B}_{1},\mathcal{B}_{0})>1 and there is an expanding 0\mathcal{B}_{0}-relative immersion for ϕ|1\left.\phi\right|_{\mathcal{B}_{1}}. By inductively patching these immersions together using Proposition 4.4, we get an expanding 0\mathcal{B}_{0}-relative immersion for ϕ\phi and we are done. Unfortunately, since ϕ\phi is not surjective, it could be that no chain 𝒜=0n={F}\mathcal{A}=\mathcal{B}_{0}\prec\cdots\prec\mathcal{B}_{n}=\{F\} satisfies the naive assumption we made at the start. Recall that our definition of ϕ|m+1\left.\phi\right|_{\mathcal{B}_{m+1}} being irreducible relative to m\mathcal{B}_{m} presupposed ϕ1m=m\phi^{-1}\cdot\mathcal{B}_{m}=\mathcal{B}_{m}. Fortunately, this is a minor complication that can be easily addressed. The proof follows the approach described above closely but uses a chain with slightly weaker conditions on it.

We first construct a chain 𝒜=0n={F}\mathcal{A}=\mathcal{B}_{0}\prec\cdots\prec\mathcal{B}_{n}=\{F\} in the poset of [ϕ][\phi]-invariant free factor system that we will induct on. Let 𝒜1\mathcal{A}\prec\mathcal{B}_{1} be a chain of [ϕ][\phi]-invariant free factor systems such that ϕ|1\left.\phi\right|_{\mathcal{B}_{1}} is irreducible relative to 𝒜\mathcal{A}. Suppose m\mathcal{B}_{m} has been constructed for some m1m\geq 1 and let 𝒞m\mathcal{C}\succeq\mathcal{B}_{m} be the maximal free factor system in the chain mϕ1mϕ2m\mathcal{B}_{m}\preceq\phi^{-1}\cdot\mathcal{B}_{m}\preceq\phi^{-2}\cdot\mathcal{B}_{m}\preceq\cdots of [ϕ][\phi]-invariant free factor systems. If ϕ1m=m=𝒞\phi^{-1}\cdot\mathcal{B}_{m}=\mathcal{B}_{m}=\mathcal{C}, then let mm+1\mathcal{B}_{m}\prec\mathcal{B}_{m+1} be a chain of [ϕ][\phi]-invariant free factor systems such that ϕ|m+1\left.\phi\right|_{\mathcal{B}_{m+1}} is irreducible relative to m\mathcal{B}_{m}. If m𝒞\mathcal{B}_{m}\prec\mathcal{C}, then let mm+k=𝒞\mathcal{B}_{m}\prec\cdots\prec\mathcal{B}_{m+k}=\mathcal{C} be the chain of [ϕ][\phi]-invariant free factor systems such that m+i=ϕim\mathcal{B}_{m+i}=\phi^{-i}\cdot\mathcal{B}_{m} for 1ik1\leq i\leq k.

We proceed by inducting on the resulting chain between 𝒜\mathcal{A} and {F}\{F\}. For the base case, ϕ|1\left.\phi\right|_{\mathcal{B}_{1}} is irreducible relative to 𝒜\mathcal{A}; therefore, there is an expanding 𝒜\mathcal{A}-relative immersion for ϕ|1\left.\phi\right|_{\mathcal{B}_{1}} by Proposition 4.2. For our induction hypothesis, suppose that there is an expanding 𝒜\mathcal{A}-relative immersion fm:TmTmf_{\mathcal{B}_{m}}:T_{\mathcal{B}_{m}}\to T_{\mathcal{B}_{m}} for ϕ|m\left.\phi\right|_{\mathcal{B}_{m}} for some m1m\geq 1. By our construction of the chain, either ϕ|m+1\left.\phi\right|_{\mathcal{B}_{m+1}} is irreducible relative to \mathcal{B} or ϕ(m+1)\phi(\mathcal{B}_{m+1}) is carried by m\mathcal{B}_{m}. We deal with these two cases separately.

Case 1. Suppose ϕ|m+1\left.\phi\right|_{\mathcal{B}_{m+1}} is irreducible relative to m\mathcal{B}_{m}. By Proposition 3.7, there is an irreducible m\mathcal{B}_{m}-relative representative f:TTf_{*}:T_{*}\to T_{*} for ϕ|m+1\left.\phi\right|_{\mathcal{B}_{m+1}} with minimal stretch factor. If λ(f)=1\lambda(f_{*})=1, then fm+1=ff_{\mathcal{B}_{m+1}}=f_{*} is a m\mathcal{B}_{m}-relative simplicial immersion for ϕ|m+1\left.\phi\right|_{\mathcal{B}_{m+1}}. If λ(f)>1\lambda(f_{*})>1, then there is an expanding \mathcal{B}-relative immersion fm+1f_{\mathcal{B}_{m+1}} for ϕ|m+1\left.\phi\right|_{\mathcal{B}_{m+1}} by Proposition 4.1. In either case, we get a m\mathcal{B}_{m}-relative immersion fm+1:Tm+1Tm+1f_{\mathcal{B}_{m+1}}:T_{\mathcal{B}_{m+1}}\to T_{\mathcal{B}_{m+1}} for ϕ|m+1\left.\phi\right|_{\mathcal{B}_{m+1}} defined on a (m+1,m)(\mathcal{B}_{m+1},\mathcal{B}_{m})-forest Tm+1T_{\mathcal{B}_{m+1}}. Thus, there is an expanding 𝒜\mathcal{A}-relative immersion for ϕ|m+1\left.\phi\right|_{\mathcal{B}_{m+1}} by the induction hypothesis and Proposition 4.4.

Case 2. Now suppose ϕ(m+1)\phi(\mathcal{B}_{m+1}) is carried by m\mathcal{B}_{m}. Let Tm(ϕ(m+1))TmT_{\mathcal{B}_{m}}(\phi(\mathcal{B}_{m+1}))\subset T_{\mathcal{B}_{m}} be the minimal subforest of ϕ(m+1)\phi(\mathcal{B}_{m+1}) and Tm+1=Tm(ϕ(m+1))T_{\mathcal{B}_{m+1}}=T_{\mathcal{B}_{m}}(\phi(\mathcal{B}_{m+1})) be the same forest after forgetting bivalent vertices. By injectivity of ϕ\phi, we may consider Tm+1T_{\mathcal{B}_{m+1}} as a (m+1,𝒜)(\mathcal{B}_{m+1},\mathcal{A})-forest that comes with a natural ϕ|m+1\left.\phi\right|_{\mathcal{B}_{m+1}}-equivariant immersion g:Tm+1Tmg:T_{\mathcal{B}_{m+1}}\to T_{\mathcal{B}_{m}}. Since fm:TmTmf_{\mathcal{B}_{m}}:T_{\mathcal{B}_{m}}\to T_{\mathcal{B}_{m}} is an immersion, we can identify a subdivision of TmT_{\mathcal{B}_{m}} with the minimal subforest Tm+1(m)Tm+1T_{\mathcal{B}_{m+1}}(\mathcal{B}_{m})\subset T_{\mathcal{B}_{m+1}} of m\mathcal{B}_{m}. Composing gg with the subdivision and inclusion Tm+1(m)Tm+1T_{\mathcal{B}_{m+1}}(\mathcal{B}_{m})\subset T_{\mathcal{B}_{m+1}} gives an 𝒜\mathcal{A}-relative immersion fm+1:Tm+1Tm+1f_{\mathcal{B}_{m+1}}:T_{\mathcal{B}_{m+1}}\to T_{\mathcal{B}_{m+1}} for ϕ|m+1\left.\phi\right|_{\mathcal{B}_{m+1}}, which is expanding since its image lies in Tm+1(m)T_{\mathcal{B}_{m+1}}(\mathcal{B}_{m}) and its restriction to Tm+1(m)T_{\mathcal{B}_{m+1}}(\mathcal{B}_{m}) is the expanding 𝒜\mathcal{A}-relative immersion fmf_{\mathcal{B}_{m}} after forgetting bivalent vertices. ∎

It will follow from bounded cancellation that expanding relative immersion from the theorem is in fact canonical.

Proposition 4.6.

Suppose ϕ:FF\phi:F\to F is injective and there is an expanding 𝒜\mathcal{A}-relative immersion for ϕ\phi, where 𝒜\mathcal{A} is a [ϕ][\phi]-invariant free factor system. Then there is a unique expanding 𝒜\mathcal{A}-relative immersion for ϕ\phi.

Proof.

Let ϕ:FF\phi:F\to F be injective, 𝒜\mathcal{A} a [ϕ][\phi]-invariant free factor system. Suppose f:TTf:T\to T and f:TTf^{\prime}:T^{\prime}\to T^{\prime} are expanding 𝒜\mathcal{A}-relative immersions for ϕ\phi. The goal is to show that TT and TT^{\prime} are equivariantly homeomorphic. Let g:TTg:T\to T^{\prime} be an equivariant tree map, i.e., ψ\psi-equivariant tree map with ψ:FF\psi:F\to F being the identity automorphism; such a map always exists between (F,𝒜)(F,\mathcal{A})-trees. By taking restrictions to the ϕk(F)\phi^{k}(F)-minimal subtrees T(ϕk(F))T(\phi^{k}(F)) and T(ϕk(F))T^{\prime}(\phi^{k}(F)) and applying deformation retractions, we get a ϕk(F)\phi^{k}(F)-equivariant tree map gk:T(ϕk(F))T(ϕk(F))g_{k}:T(\phi^{k}(F))\to T^{\prime}(\phi^{k}(F)) with cancellation constant C(g)C(g). By Corollary 3.3, we assume gkg_{k} is an equivariant natural tree map with cancellation constant 2C(g)2C(g).

Recall that ff^{\prime} is an expanding 𝒜\mathcal{A}-relative immersion for ϕ\phi, so all the natural edges of T(ϕk(F))T^{\prime}(\phi^{k}(F)) are longer than 2C(g)2C(g) for large enough kk. Fix k0k\gg 0, then no folding occurs in gkg_{k} by bounded cancellation. So gkg_{k} is a forest collapse and T(ϕk(F))T^{\prime}(\phi^{k}(F)) has at most the same number of orbits of branch points as T(ϕk(F))T(\phi^{k}(F)). By the same argument, T(ϕk(F))T(\phi^{k}(F)) is an equivariant forest collapse of T(ϕk(F))T^{\prime}(\phi^{k}(F)) and has at most the same number of orbits of branch points as T(ϕk(F))T^{\prime}(\phi^{k}(F)). This implies the two minimal subtrees have the same number of orbits of branch points and gkg_{k} is an equivariant homeomorphism. Since ff and ff^{\prime} are immersions, TT and TT^{\prime} are ϕk\phi^{k}-equivariantly homeomorphic to T(ϕk(F))T(\phi^{k}(F)) and T(ϕk(F))T^{\prime}(\phi^{k}(F)) respectively. Therefore, TT and TT^{\prime} are equivariantly homeomorphic by injectivity of ϕ\phi. So ff and ff^{\prime} are homotopic immersions and hence agree up to isotopy. ∎

Corollary 4.7.

If ϕ:FF\phi:F\to F is injective but not surjective, then there is a unique expanding 𝒜\mathcal{A}-relative immersion for ϕ\phi, where 𝒜\mathcal{A} is the [ϕ][\phi]-elliptic free factor system.

This gives us a complete characterization of when an injective endomorphism is induced by a unique expanding graph immersion.

Corollary 4.8.

Let ϕ:FF\phi:F\to F be an injective endomorphism. Then the following conditions are equivalent:

  1. 1.

    [ϕ][\phi] is induced by a unique expanding graph immersion;

  2. 2.

    [ϕ][\phi] is induced by an expanding graph immersion;

  3. 3.

    the trivial conjugacy class is the only conjugacy class with an infinite [ϕ][\phi]-tail;

  4. 4.

    the trivial system is the only [ϕ][\phi]-fixed free factor system.

Proof.

We leave the implication (2)(3)(2)\implies(3) as an exercise. By Theorem 2.1, we have (3)(4)(3)\iff(4). It remains to show (3)(1)(3)\implies(1). Suppose the trivial conjugacy class is the only conjugacy class with an infinite [ϕ][\phi]-tail. So ϕ\phi is not surjective and the [ϕ][\phi]-elliptic free factor system is trivial. By Corollary 4.7, there is a unique expanding ϕ\phi-equivariant immersion f:TTf:T\to T defined on a free FF-tree TT, i.e., [ϕ][\phi] is induced by a unique expanding immersion on the marked graph F\TF\backslash T. ∎

Interlude

Let us summarize the main results of the first part of the paper.

Summary.

If ϕ:FF\phi:F\to F is injective but not surjective, then there is:

  1. 1.

    a unique maximal proper [ϕ][\phi]-fixed free factor system 𝒜\mathcal{A}; (Proposition 2.3)

  2. 2.

    a unique maximal proper [ϕ][\phi]-invariant free factor system 𝒜𝒜\mathcal{A}^{*}\succeq\mathcal{A} such that ϕk(𝒜)\phi^{k}(\mathcal{A}^{*}) is carried by 𝒜\mathcal{A} for some k0k\geq 0; after replacing the free factors of 𝒜\mathcal{A} with conjugates if necessary, we can assume 𝒜𝒜\mathcal{A}\subset\mathcal{A}^{*}; (Proposition 2.4)

  3. 3.

    a unique expanding 𝒜\mathcal{A}^{*}-relative immersion for ϕ\phi. (Corollary 4.7)

We call the free factor system 𝒜\mathcal{A}^{*} the [ϕ][\phi]-elliptic free factor system. By expanding immersion, we mean every edge of the tree grows under iteration.

As a corollary of this summary, we have the following characterization: [ϕ][\phi] is induced by an expanding graph immersion if and only if the trivial system is the only [ϕ][\phi]-fixed free factor system (Corollary 4.8). We will now contextualize these results and, especially, this corollary. In our previous work [23], we determined exactly when the mapping torus of an expanding graph immersion has a word-hyperbolic fundamental group.

Theorem ([23, Theorem 6.3]).

Let ϕ:FF\phi:F\to F be induced by an expanding graph immersion. FϕF*_{\phi} is word-hyperbolic if and only if it has no BS(1,d)BS(1,d) subgroups for d2d\geq 2.

By Reynolds’ result (Corollary 4.3), we knew this theorem applied to all irreducible nonsurjective endomorphisms but, at the time, there was no known algebraic characterization of the general class of endomorphisms induced by expanding graph immersions. It is clear that nontrivial [ϕ][\phi]-periodic subgroups are an obstruction to ϕ\phi being induced by an expanding graph immersion. Corollary 4.8 means nontrivial fixed free factor systems are the only (essential) obstruction. So the above theorem can be restated as:

Theorem.

Let ϕ:FF\phi:F\to F be an injective endomorphism with no nontrivial fixed free factor system. FϕF*_{\phi} is word-hyperbolic if and only if it has no BS(1,d)BS(1,d) subgroups for d2d\geq 2.

The interesting thing about the restatement is that it is purely algebraic, i.e., there is no mention of topological maps. Since fixed free factor systems correspond to free-by-cyclic subgroups FF^{\prime}\rtimes\mathbb{Z} of FϕF*_{\phi}, the restatement also suggests how to extend our previous work to all injective nonsurjective endomorphisms.

In the next sections, we use expanding relative immersions to relativize our previous work in [23]. For instance, the sequence of lemmas/propositions in the next section is essentially identical to the sequence in [23, Section 3]. In Section 6, we extend our previous theorem to all injective endomorphisms. However, this will not constitute an alternate proof of Brinkmann’s theorem [8]: FF\rtimes\mathbb{Z} is word-hyperbolic if and only if it has no 2\mathbb{Z}^{2} subgroups. We assume Brinkmann’s theorem as the base case of our generalization.

Before diving into the details, we give an illustration of the consideration in the second part of this paper. To prove that BS(1,d)BS(1,d) subgroups are the only obstruction to word-hyperbolicity, the main tool we use is the Bestvina-Feighn combination theorem [3]: as long as the group in question satisfies the annuli flaring condition, it will be word-hyperbolic. See Section 7 for the exact statement. For non-examples, we now explain why/how BS(1,d)BS(1,d) groups fail the annuli flaring condition.

Consider two Baumslag-Solitar groups BS(1,d)BS(1,d) with d=1,2d=1,2. Note that BS(1,1)=2BS(1,1)=\mathbb{Z}^{2}. Fix the presentation BS(1,d)=a,t|t1at=adBS(1,d)=\langle a,t~{}|~{}t^{-1}at=a^{d}\rangle and notice that the group is a mapping torus of an endomorphism ψ\psi of the cyclic group a\langle a\rangle. So BS(1,d)BS(1,d) acts on a simplicial tree 𝒯\mathcal{T} with point and edge stabilizers conjugate to a\langle a\rangle and the quotient graph BS(1,d)\𝒯BS(1,d)\backslash\mathcal{T} has one vertex and edge. The edge has a natural orientation inherited from the nontrivial loop corresponding to tt and we can lift the orientation to an equivariant orientation of 𝒯\mathcal{T}. Under this orientation, 𝒯\mathcal{T} is like a rooted tree: every vertex has exactly dd incoming edges and one outgoing edge. Choose a vertex \star on the axis of tt whose stabilizer by a\langle a\rangle and define 𝒯a\mathcal{T}_{\langle a\rangle} to be the family tree of \star-descendants, i.e., union of directed edge-paths that terminate on \star.

Refer to caption
Figure 6: Portion of the trees 𝒯a𝒯\mathcal{T}_{\langle a\rangle}\subset\mathcal{T} when d=2d=2. The line on the left is the axis of tt.

A class of annuli of length nn in BS(1,d)BS(1,d) is an ordered pair (g,pg)(g,p_{g}) where gBS(1,d)g\in BS(1,d) is nontrivial and pgp_{g} is a geodesic of length nn in 𝒯\mathcal{T} fixed by gg. Since gg is elliptic, then (up to conjugacy) we may assume gag\in\langle a\rangle, pg\star\in p_{g}, and pg𝒯ap_{g}\subset\mathcal{T}_{\langle a\rangle}. The orientation on 𝒯\mathcal{T} implies there are two types of annuli: 1) unidirectional — those whose geodesics are directed paths terminating on \star; 2) bidirectional — those whose geodesics are a wedge of two directed paths terminating on \star. When d=1d=1, 𝒯\mathcal{T} is a line and all annuli are unidirectional. Loosely speaking, flaring means expanding exponentially towards some end. It should be intuitive that annuli will not flare because the associated endomorphism ψ\psi is the identity map. When d=2d=2, 𝒯\mathcal{T} is a regular trivalent tree and annuli can be unidirectional or bidirectional. Since the associated endomorphism ψ\psi is represented by a degree 2 map of the circle, the unidirectional annuli will flare. However, due to the branching in 𝒯\mathcal{T}, it is possible to wedge two flaring unidirectional annuli along their expanding ends to get a bidirectional annulus that shrinks exponentially towards both its ends.

The objective is to rule out both of these behaviors in the setting of a mapping torus FϕF*_{\phi} that has no BS(1,d)BS(1,d) subgroups with d1d\geq 1 and thus conclude that the mapping torus satisfies the annuli flaring condition. Here is an outline of the second part of the paper:

  1. 1.

    In Section 5, we prove that if FϕF*_{\phi} has no BS(1,d)BS(1,d) subgroups with d2d\geq 2, then there is a uniform bound on one of the two directed pieces of any bidirectional annulus. Thus long bidirectional annuli are unidirectional up to bounded error. However, since annuli are not defined yet in this section, everything is done in terms of pullbacks, which are special classes of bidirectional annuli.

  2. 2.

    In Section 6, we define annuli and give the correspondence with pullbacks. We then use expanding relative immersions and Brinkmann’s theorem to prove the section’s main result: if FϕF*_{\phi} has no 2\mathbb{Z}^{2} subgroups, then long unidirectional annuli flare. Thus, together with the previous section, FϕF*_{\phi} satisfies the annuli flaring condition if it has no BS(1,d)BS(1,d) subgroups with d1d\geq 1.

  3. 3.

    Finally, in Section 7, we extend the result to HNN extensions of free groups over free factors FAF*_{A} by showing that there is a canonical finite collection of mapping tori FϕFAF^{\prime}*_{\phi^{\prime}}\leq F*_{A} that carry all long annuli of FAF*_{A}. Intuitively speaking, the action of FAF*_{A} on its Bass-Serre tree is acylindrical relative to {Fiϕi}\{F_{i}^{\prime}*_{\phi_{i}^{\prime}}\}. So FAF*_{A} satisfies the annuli flaring condition if and only if the subgroups {Fiϕi}\{F_{i}^{\prime}*_{\phi_{i}^{\prime}}\} satisfy the condition.

Geometry of ascending HNN extensions

5 Iterated pullbacks

In our previous work [23, Sections 3], topological pullbacks for a graph immersion f:ΓΓf:\Gamma\to\Gamma were used to give sufficient conditions for π1(f)=ϕ:FF\pi_{1}(f)=\phi:F\to F to have an invariant nonfixed cyclic subgroup system. The goal of this section is to drop the immersion hypothesis and give sufficient conditions that apply to all injective endomorphisms of FF.

Suppose immersions f1:Γ1Γf_{1}:\Gamma_{1}\to\Gamma and f2:Γ2Γf_{2}:\Gamma_{2}\to\Gamma induce inclusions of free groups H1,H2FH_{1},H_{2}\leq F respectively. Then components of the core (topological) pullback of (f1,f2)(f_{1},f_{2}), also known as fibered products are in one-to-one correspondence with nontrivial intersection H1gH2g1H_{1}\cap gH_{2}g^{-1} as gg ranges over (H1,H2)(H_{1},H_{2})-double coset representatives of H1\F/H2H_{1}\backslash F/H_{2}. For a graph immersion f:ΓΓf:\Gamma\to\Gamma that induces an endomorphism ϕ:FF\phi:F\to F, we get a one-to-one correspondence between components of the core pullback of (fk,fk)(f^{k},f^{k}) and nontrivial conjugacy classes [ϕk(F)gϕk(F)g1][\phi^{k}(F)\cap g\phi^{k}(F)g^{-1}] as [[g]][[g]] ranges over ϕk(F)\phi^{k}(F)-double cosets for all k1k\geq 1. We will not define pullbacks topological maps since we will be working (semi-)algebraically in this section.

Given subgroups H1,H2FH_{1},H_{2}\leq F, we define the (algebraic) pullback of (H1,H2)(H_{1},H_{2}), denoted by H1H2H_{1}\wedge H_{2}, to be the set of all nontrivial components [H1gH2g1][H_{1}\cap gH_{2}g^{-1}] as [[g]][[g]] ranges over the (H1,H2)(H_{1},H_{2})-double cosets in FF. Note that pullbacks are not subgroup systems since they are sets of conjugacy classes of subgroups. When H1H_{1} and H2H_{2} are finitely generated, then their pullback is a finite set. For an injective endomorphism ϕ:FF\phi:F\to F and k1k\geq 1, define the iterated (algebraic) pullbacks of ϕ\boldsymbol{\phi} to be Λk(ϕ)=ϕk(F)ϕk(F)\Lambda_{k}(\phi)=\phi^{k}(F)\wedge\phi^{k}(F). Furthermore, the iterated pullback depends only on the outer class [ϕ][\phi] and will be denoted later by Λk[ϕ]\Lambda_{k}[\phi]. In this section, we will simply write Λk\Lambda_{k} for brevity.

The (algebraic) mapping torus of an injective endomorphism ϕ\phi is the ascending HNN extension of FF with the presentation Fϕ=F,t|t1xt=ϕ(x),xF.F*_{\phi}=\langle\,F,t~{}|~{}t^{-1}xt=\phi(x),\forall x\in F\,\rangle. We now give a third description of Λk\Lambda_{k} that will actually be the working description. The mapping torus FϕF*_{\phi} is the fundamental group of a circle of groups with one vertex group FF and edge group FF. The edge monomorphisms for this circle of groups are the identity map idF:FFid_{F}:F\to F and endomorphism ϕ:FF\phi:F\to F and the corresponding Bass-Serre tree 𝒯\mathcal{T} has one orbit of edges and vertices. The tree also comes with a natural orientation where each vertex has exactly one outgoing oriented edge and the stable letter tFϕt\in F*_{\phi} acts on its directed axis with positive translation.

By construction, there is a vertex \star of 𝒯\mathcal{T} whose stabilizer in FϕF*_{\phi} is exactly FF. Let 𝒯F\mathcal{T}_{F} be the full subtree of 𝒯\mathcal{T} rooted at \star. Colloquially, 𝒯F\mathcal{T}_{F} is the family tree of \star-descendants. We will use the action of FϕF*_{\phi} on 𝒯\mathcal{T} to index the vertices of 𝒯F\mathcal{T}_{F} by cosets in FF. The root \star is indexed gFgF, where gF=FgF=F is the coset of FF in FF. The kthk^{\text{th}}-generation vertices gtkgt^{-k}\cdot\star are indexed by the cosets gϕk(F)g\phi^{k}(F) in FF. The intersection g1ϕk(F)g11g2ϕk(F)g21g_{1}\phi^{k}(F)g_{1}^{-1}\cap g_{2}\phi^{k}(F)g_{2}^{-1} corresponds to the simultaneous stabilizer of the ordered pair of vertices indexed by g1ϕk(F)g_{1}\phi^{k}(F) and g2ϕk(F)g_{2}\phi^{k}(F); the intersection is conjugate in FF to the stabilizer of the vertices indexed by ϕk(F)\phi^{k}(F) and gϕk(F)g\phi^{k}(F), where g=g11g2g=g_{1}^{-1}g_{2}. The intersection is also conjugate (in FF) to the stabilizers of the vertices indexed by ϕk(F)\phi^{k}(F) and hgϕk(F)hg\phi^{k}(F), where hϕk(F)h\in\phi^{k}(F). Thus, FF-orbits of kthk^{\text{th}}-generation ordered pairs of vertices are indexed by ϕk(F)\phi^{k}(F)-double cosets in FF and conjugacy classes of their nontrivial stabilizers are the components of Λk\Lambda_{k}.

There is a function ϕ|Λk:ΛkΛk+1\left.\phi\right|_{\Lambda_{k}}:\Lambda_{k}\to\Lambda_{k+1} induced by ϕ\phi, given algebraically by

[ϕk(F)gϕk(F)g1][ϕk+1(F)ϕ(g)ϕk+1(F)ϕ(g)1].[\phi^{k}(F)\cap g\phi^{k}(F)g^{-1}]\quad\mapsto\quad[\phi^{k+1}(F)\cap\phi(g)\phi^{k+1}(F)\phi(g)^{-1}].

Graphically, the function maps (the conjugacy class of) the stabilizer of (g1tk,g2tk)(g_{1}t^{-k},g_{2}t^{-k})\cdot\star to (the conjugacy class of) the stabilizer of t1(g1tk,g2tk)t^{-1}\cdot(g_{1}t^{-k},g_{2}t^{-k})\cdot\star.

Lemma 5.1.

ϕ|Λk:ΛkΛk+1\left.\phi\right|_{\Lambda_{k}}:\Lambda_{k}\to\Lambda_{k+1} as a function on the set of components is injective.

Graphical proof.

If t1(g1tk,g2tk)t^{-1}\cdot(g_{1}t^{-k},g_{2}t^{-k})\cdot\star and t1(g1tk,g2tk)t^{-1}\cdot(g_{1}^{\prime}t^{-k},g_{2}^{\prime}t^{-k})\cdot\star are in the same FF-orbit, then they are in fact in the same ϕ(F)\phi(F)-orbit (descendants of t1t^{-1}\cdot\star). So (g1tk,g2tk)(g_{1}t^{-k},g_{2}t^{-k})\cdot\star and (g1tk,g2tk)(g_{1}^{\prime}t^{-k},g_{2}^{\prime}t^{-k})\cdot\star are in the same FF-orbit by the action of tt. ∎

One can give an algebraic proof of the lemma that replaces the action of tt with the injectivity of ϕ\phi. The lemma implies there is a chain of injections Λ0Λ1Λ2\Lambda_{0}\to\Lambda_{1}\to\Lambda_{2}\to\cdots. Furthermore, the restriction to each component is an isomorphism since it is a conjugation by tt in FϕF*_{\phi}. We will be mainly interested in the set compliment Λ^k=Λkϕ(Λk1)\hat{\Lambda}_{k}=\Lambda_{k}\setminus\phi(\Lambda_{k-1}). Equivalently, Λ^k={[ϕk(F)gϕk(F)g1]Λk:gϕ(F)}.\hat{\Lambda}_{k}=\{\,[\phi^{k}(F)\cap g\phi^{k}(F)g^{-1}]\in\Lambda_{k}~{}:~{}g\notin\phi(F)\,\}. Graphically, Λ^k\hat{\Lambda}_{k} is the set of components that stabilizes FF-orbits of kthk^{\text{th}}-generation pairs of vertices in 𝒯F\mathcal{T}_{F} whose only common ancestor is \star. We might say iterated pullbacks stabilize if Λ^k=\hat{\Lambda}_{k}=\emptyset for some kk. The image ϕ(F)\phi(F) is malnormal in FF if and only if Λ^1=\hat{\Lambda}_{1}=\emptyset. Iterated pullback stability is a sort of generalization of malnormality for ϕk(F)\phi^{k}(F) with respect to ϕ\phi.

Lemma 5.2.

Suppose ϕ:FF\phi:F\to F is injective and k1k\geq 1. If Λ^k\hat{\Lambda}_{k} is empty, then so is Λ^k+1\hat{\Lambda}_{k+1}; and if Λ^k\hat{\Lambda}_{k} has only cyclic components, then Λ^k+1\hat{\Lambda}_{k+1} is empty or has only cyclic components.

Proof.

There is an “inclusion” of components, Λ^k+1Λ^k\hat{\Lambda}_{k+1}\preccurlyeq\hat{\Lambda}_{k}, since a stabilizer of a (k+1)th(k{+}1)^{\text{th}}-generation pair of vertices is contained in a stabilizer of a kthk^{\text{th}}-generation pair of vertices. So Λ^k=\hat{\Lambda}_{k}=\emptyset implies Λ^k+1=\hat{\Lambda}_{k+1}=\emptyset. Suppose Λ^k\hat{\Lambda}_{k} has cyclic components, then Λ^k+1\hat{\Lambda}_{k+1} is empty or has cyclic components as the subgroups of a cyclic group are trivial or cyclic. ∎

The reduced rank of a nontrivial finite rank free group HH is rr(H)=rank(H)1\mathrm{rr}(H)=\mathrm{rank}(H)-1 and the reduced rank of a pullback H1H2H_{1}\wedge H_{2}, where H1,H2H_{1},H_{2} are finitely generated free groups, is the sum of the reduced ranks of components in H1H2H_{1}\wedge H_{2}. The latter is denoted by rr(H1H2)\mathrm{rr}(H_{1}\wedge H_{2}). Since the restriction ϕ|Λk\left.\phi\right|_{\Lambda_{k}} gives natural isomorphisms of the components, the chain of injections produces a nondecreasing sequence of positive integers

rank(F)1=rr(F)rr(Λ1)rr(Λ2).\mathrm{rank}(F)-1=\mathrm{rr}(F)\leq\mathrm{rr}(\Lambda_{1})\leq\mathrm{rr}(\Lambda_{2})\leq\cdots.

Observe that rr(Λi)=rr(Λi+1)\mathrm{rr}(\Lambda_{i})=\mathrm{rr}(\Lambda_{i+1}) if and only if Λ^i+1\hat{\Lambda}_{i+1} is empty or has cyclic components only. By Lemma 5.2, the sequence becomes constant once two consecutive entries are equal. Walter Neumann used topological pullbacks of Stallings graphs to bound the reduced ranks of algebraic pullbacks and improve Hanna Neumann’s bound [25].

Theorem 5.3 ([26, Proposition 2.1]).

If H1,H2FH_{1},H_{2}\leq F are nontrivial finitely generated subgroups, then rr(H1H2)2rr(H1)rr(H2)\mathrm{rr}(H_{1}\wedge H_{2})\leq 2\,\mathrm{rr}(H_{1})\,\mathrm{rr}(H_{2}).

Remark.

Although this bound is weaker than the Strengthened Hanna Neumann Conjecture, now Friedman-Mineyev’s Theorem [11, 19], it is sufficient for our purposes. W. Neumann’s proof is an elementary Euler characteristic computation for the topological pullback.

The uniform bound given by W. Neumann’s theorem allows us to conclude that the sequence of reduced ranks given above is eventually constant, and hence, past a certain point we are adding only cyclic components or nothing at all to the iterated pullbacks.

Lemma 5.4.

If ϕ:FF\phi:F\to F is injective, then either Λ^k\hat{\Lambda}_{k} has cyclic components for all k2rr(F)2k\geq 2\,\mathrm{rr}(F)^{2} or Λ^k=\hat{\Lambda}_{k}=\emptyset for some kk.

Proof.

Theorem 5.3 gives us a uniform bound on the reduced ranks of the iterated pullbacks rr(Λk)2rr(ϕk(F))2=2rr(F)2\mathrm{rr}(\Lambda_{k})\leq 2\,\mathrm{rr}(\phi^{k}(F))^{2}=2\,\mathrm{rr}(F)^{2} for all k1k\geq 1. By Lemma 5.2 and the uniform bound on the nondecreasing sequence of reduced ranks, rr(Λk)\mathrm{rr}(\Lambda_{k}) are all equal for kk0=2rr(F)2k\geq k_{0}=2\,\mathrm{rr}(F)^{2}. Therefore, Λ^k0\hat{\Lambda}_{k_{0}} is empty or has only cyclic components. The lemma follows by inductively applying Lemma 5.2. ∎

We say ϕ:FF\phi:F\to F has an invariant cyclic subgroup system with index d𝟏\boldsymbol{d\geq 1} if there is an integer k1k\geq 1, element xFx\in F, and nontrivial cyclic subgroup cF\langle c\rangle\leq F such that ϕk(c)xcx1\phi^{k}(\langle c\rangle)\leq x\langle c\rangle x^{-1} and has index dd. Note that this is actually a property of the outer endomorphism [ϕ][\phi]. We can now give the main result of this section:

Proposition 5.5.

Let ϕ:FF\phi:F\to F be injective. If Λ^k\hat{\Lambda}_{k} is not empty for all k1k\geq 1, then [ϕ][\phi] has an invariant cyclic subgroup system with index d2d\geq 2.

Proof.

Let k0=2rr(F)2k_{0}=2\,\mathrm{rr}(F)^{2} and ϕ:FF\phi:F\to F be an injective endomorphism such that Λ^k\hat{\Lambda}_{k} is not empty for all k1k\geq 1. By Lemma 5.4, Λ^k\hat{\Lambda}_{k} has cyclic components for kk0k\geq k_{0}.

So far nothing in the section has used relative immersions but our main motivation for constructing them was this proposition. Note that Λ^1\hat{\Lambda}_{1}\neq\emptyset automatically implies ϕ\phi is not surjective. So ϕ\phi is injective but not surjective. By Theorem 4.5, there is an expanding 𝒜\mathcal{A}^{*}-relative immersion f:TTf:T\to T for ϕ\phi, where 𝒜\mathcal{A}^{*} is the [ϕ][\phi]-elliptic free factor system and TT is an (F,𝒜)(F,\mathcal{A}^{*})-tree. Recall that the equality ϕ1𝒜=𝒜\phi^{-1}\cdot\mathcal{A}^{*}=\mathcal{A}^{*} is part of our definition of 𝒜\mathcal{A}^{*}-relative representatives. An element of FF is elliptic if it fixes a point on TT, i.e., its conjugacy class is carried by 𝒜\mathcal{A}^{*}. By construction of the [ϕ][\phi]-elliptic free factor system (Proposition 2.4), there is an m1m\geq 1 such that [ϕk(g)][\phi^{k}(g)] is carried by the maximal [ϕ][\phi]-fixed free factor system 𝒜\mathcal{A} for all kmk\geq m and elliptic gFg\in F. As 𝒜\mathcal{A} is a [ϕ][\phi]-fixed free factor system, the components of ϕk(𝒜)\phi^{k}(\mathcal{A}) are free factors of FF for all kk.

Suppose [ϕk(F)gϕk(F)g1]Λ^k[\phi^{k}(F)\cap g\phi^{k}(F)g^{-1}]\in\hat{\Lambda}_{k} for some kk1=max(k0,m+1)k\geq k_{1}=\max(k_{0},m+1). As kk0k\geq k_{0}, this component is cyclic and we may assume it has a representative ϕk(F)gϕk(F)g1\phi^{k}(F)\cap g\phi^{k}(F)g^{-1} generated by a nontrivial element ϕk(x)ϕk(F)\phi^{k}(x)\in\phi^{k}(F). In particular, there exists ϕk(y)ϕk(F)\phi^{k}(y)\in\phi^{k}(F) and gFϕ(F)g\in F\setminus\phi(F) such that ϕk(x)=gϕk(y)g1\phi^{k}(x)=g\phi^{k}(y)g^{-1}. We first show that xx is loxodromic, i.e., not elliptic. Suppose xFx\in F is elliptic, i.e., 𝒜\mathcal{A}^{*} carries [x][x]. Then some free factor ϕk(A)ϕk(𝒜)\phi^{k}(A)\in\phi^{k}(\mathcal{A}) carries [ϕk(x)]=[ϕk(y)][\phi^{k}(x)]=[\phi^{k}(y)] as km+1k\geq m+1. So yy must be elliptic too since ϕ1𝒜=𝒜\phi^{-1}\cdot\mathcal{A}^{*}=\mathcal{A}^{*} and, as noted in the previous paragraph, ϕm(A)\phi^{m}(A) carries ϕm(x)\phi^{m}(x) and ϕm(y)\phi^{m}(y). Thus, up to conjugation in FF, we may assume ϕm(x),ϕm(y)ϕm(A)\phi^{m}(x),\phi^{m}(y)\in\phi^{m}(A) and, up to conjugation in ϕkm(F)\phi^{k-m}(F), we may assume ϕk(x),ϕk(y)ϕk(A)\phi^{k}(x),\phi^{k}(y)\in\phi^{k}(A). Malnormality of the free factor ϕk(A)\phi^{k}(A) implies gϕkm(F)g\in\phi^{k-m}(F). But km1k-m\geq 1 leads to the contradiction gϕ(F)g\in\phi(F). Therefore, xx is loxodromic. The integer kk1k\geq k_{1} and component [ϕk(x)]Λ^k[\,\langle\phi^{k}(x)\rangle\,]\in\hat{\Lambda}_{k} were arbitrary, so all components of Λ^k\hat{\Lambda}_{k} are loxodromic for kk1k\geq k_{1}. Recall from the proof of Lemma 5.2 that there is an infinite descending chain:

Λ^1Λ^2Λ^3\hat{\Lambda}_{1}\succcurlyeq\hat{\Lambda}_{2}\succcurlyeq\hat{\Lambda}_{3}\succcurlyeq\cdots

Since Λ^k1\hat{\Lambda}_{k_{1}} has finitely many components and the components are all cyclic, there is a cyclic component in Λ^k1\hat{\Lambda}_{k_{1}} which “carries” some component of Λ^k\hat{\Lambda}_{k} for all kk1k\geq k_{1}. Suppose this component has a representative generated by c=ϕk1(x)ϕk1(F)c=\phi^{k_{1}}(x)\in\phi^{k_{1}}(F). Then for all kk1k\geq k_{1}, there is a cyclic component of Λ^k\hat{\Lambda}_{k} with a representative generated by ϕk(xk)ϕk(F)\phi^{k}(x_{k})\in\phi^{k}(F) such that c\langle c\rangle carries ϕk(xk)\langle\phi^{k}(x_{k})\rangle. If we let αT\alpha\subset T be the axis for element cc, then the previous sentence implies there are sequences of element (xk)kk1(x_{k})_{k\geq k_{1}} and (tk)kk1(t_{k})_{k\geq k_{1}} such that the (unoriented) axes of ϕk(xk)\phi^{k}(x_{k}) are all translates tkαt_{k}\cdot\alpha of the (unoriented) axis α\alpha.

For any kk1k\geq k_{1}, replace xkx_{k} with its inverse if necessary so that the action of ϕk(xk)\phi^{k}(x_{k}) on its axis is coherent (respects orientation) with the action of cc on α\alpha. By passing to a strictly increasing subsequence (ki)i1(k_{i})_{i\geq 1}, we may assume there is an oriented edge ee of TT such that the axes αki\alpha_{k_{i}} of (xki)i1(x_{k_{i}})_{i\geq 1} all contain a translate of ee. We now pass to the graph of groups Γ=F\T\Gamma=F\backslash T in order to avoid mentions of translates and orbits. The axis α\alpha will project to an immersed loop α¯\bar{\alpha} in Γ\Gamma representing cc and axes αki\alpha_{k_{i}} project to immersed loops α¯ki\bar{\alpha}_{k_{i}} that represent xkix_{k_{i}} and whose f¯ki\bar{f}^{k_{i}}-image is a power α¯di\bar{\alpha}^{d_{i}} up to rotation/cyclic homotopy, where di1d_{i}\geq 1 and f¯:ΓΓ\bar{f}:\Gamma\to\Gamma is the immersion induced by f:TTf:T\to T. The edge ee projects to an edge e¯\bar{e} that is contained in all the loops α¯ki\bar{\alpha}_{k_{i}} for i1i\geq 1

The proof now mimics the proof of [23, Proposition 3.11]. Since ff is an immersion, it maps axes in TT onto axes and f¯\bar{f} maps immersed loops in Γ\Gamma to immersed loops. So f¯ki(e¯)\bar{f}^{k_{i}}(\bar{e}) is a subpath of the immersed loop f¯ki(α¯ki)α¯di\bar{f}^{k_{i}}(\bar{\alpha}_{k_{i}})\simeq\bar{\alpha}^{d_{i}} for all ii, and since ff is expanding, f¯ki(e¯)\bar{f}^{k_{i}}(\bar{e}) contains arbitrarily long powers of α¯\bar{\alpha} as ii\to\infty. Set NN to be the number of subpaths of α¯\bar{\alpha} (up to rotation) that are also loops. Choose n0n\gg 0 such that f¯kn(e¯)\bar{f}^{k_{n}}(\bar{e}) contains the loop α¯N+1\bar{\alpha}^{N+1} as a subpath. Then f¯kn+1(e¯)\bar{f}^{k_{n+1}}(\bar{e}) is a subpath of α¯dn+1\bar{\alpha}^{d_{n+1}} that contains the loop f¯kn+1kn(α¯N+1)\bar{f}^{k_{n+1}-k_{n}}(\bar{\alpha}^{N+1}) as a subpath. In fact, for all positive integers jN+1j\leq N+1, the loop f¯kn+1kn(α¯j)\bar{f}^{k_{n+1}-k_{n}}(\bar{\alpha}^{j}) is a subpath of α¯dn+1\bar{\alpha}^{d_{n+1}}. Thus, there is a sequence of loops (ϵj)j=1N+1(\epsilon_{j})_{j=1}^{N+1} that are subpaths of α¯\bar{\alpha} and strictly increasing positive integers (sj)j=1N+1(s_{j})_{j=1}^{N+1} such that f¯kn+1kn(α¯j)ϵj\bar{f}^{k_{n+1}-k_{n}}(\bar{\alpha}^{j})\cdot\epsilon_{j} is α¯sj\bar{\alpha}^{s_{j}} up to rotation. By definition of NN and pigeonhole principle, some ϵt=ϵt=ϵ\epsilon_{t}=\epsilon_{t^{\prime}}=\epsilon for some t<tt<t^{\prime} and f¯kn+1kn(α¯tt)\bar{f}^{k_{n+1}-k_{n}}(\bar{\alpha}^{t^{\prime}-t}) is α¯stst\bar{\alpha}^{s_{t^{\prime}}-s_{t}} up to rotation. Lifting back to the (F,𝒜)(F,\mathcal{A})-tree TT, we find that fkn+1knf^{k_{n+1}-k_{n}} maps the axis α\alpha to a translate of itself. So ϕkn+1kn(c)\phi^{k_{n+1}-k_{n}}(c) is conjugate to a nontrivial power cdc^{d} and d2d\geq 2 since ff is expanding. ∎

Remark.

A careful examination of the proof reveals that it can be made “more effective” with the pigeonhole principle. For any expanding relative immersion ff, we can construct a specific number k=k(f)k=k(f) for which Λ^k\hat{\Lambda}_{k}\neq\emptyset implies ϕ\phi has an invariant cyclic subgroup system with index d2d\geq 2. Thus, one would not have to check infinitely many iterated pullbacks of ϕ\phi to know that an invariant cyclic subgroup system with index d2d\geq 2 exists. Of course, for this to be useful, we need to know whether finding the [ϕ][\phi]-elliptic free factor system and expanding relative immersion for ϕ\phi can be made effective. Finally, we remark that the converse of Proposition 5.5 holds but we omit the proof as it is not needed for the rest of the paper.

In the next section, we will actually be using the contrapositive of the proposition: if ϕ\phi is injective and has no invariant cyclic subgroup system with index d2d\geq 2, then iterated pullbacks of ϕ\phi stabilize. In this case, we get control on the types of annuli in the mapping torus FϕF*_{\phi}, which allows us to prove the main theorem: FϕF*_{\phi} is word-hyperbolic if ϕ\phi additionally has no fixed cyclic subgroup system, i.e., invariant cyclic subgroup system with index d=1d=1.

6 Hyperbolic endomorphisms

We are finally ready to put all the major pieces together. The first piece involves understanding the relationship between annuli in the mapping torus FϕF*_{\phi} and iterated pullbacks of ϕ\phi. The second piece involves building on Brinkmann’s theorem to show that atoroidal injective endomorphisms are hyperbolic. In our previous work [23], we used these two pieces to give sufficient conditions for the mapping torus to be word-hyperbolic.

Theorem 6.1 ([23, Theorem 6.4]).

If f:ΓΓf:\Gamma\to\Gamma is a based-hyperbolic graph map and all strictly bidirectional annuli in its mapping torus MfM_{f} are shorter than some integer, then π1(Mf)\pi_{1}(M_{f}) is word-hyperbolic.

We will define the new terms in the theorem as we go. Suppose ϕ:FF\phi:F\to F is an injective endomorphism and f:ΓΓf:\Gamma\to\Gamma is its topological representative. Recall that the (topological) mapping torus of ff is the quotient space Mf=(Γ×[0,1])/fM_{f}=\left(\Gamma\times[0,1]\right)/{\sim_{f}} with the identification (x,1)f(f(x),0)(x,1)\sim_{f}(f(x),0) for all xΓx\in\Gamma and the algebraic mapping torus FϕF*_{\phi} is isomorphic to the fundamental group π1(Mf)\pi_{1}(M_{f}). The edge-space of MfM_{f} will be the cross-section in MfM_{f} represented by Γ×{12}\Gamma\times\{\frac{1}{2}\}.

Strictly bidirectional annuli of the mapping torus MfM_{f} with length 2L2L can be thought of as the iterated pullbacks Λ^L\hat{\Lambda}_{L} of ϕ\phi but it does take a bit of work to give the correspondence. Fix a basepoint νS1\nu\in S^{1}. For integers L1<L2L_{1}<L_{2}, an (topological) annulus in MfM_{f} of length L=L2L1L=L_{2}-L_{1} is a homotopy of loops h:S1×[L1,L2]Mfh:S^{1}\times[L_{1},L_{2}]\to M_{f} satisfying the following conditions:

  1. 1.

    it is transverse to the edge-space of MfM_{f};

  2. 2.

    the hh-preimage of the edge-space is S1×([L1,L2])S^{1}\times([-L_{1},L_{2}]\cap\mathbb{Z});

  3. 3.

    for integers i[L1,L2]i\in[L_{1},L_{2}], the rings of the annulus hi=h(,i):S1Mfh_{i}=h(\cdot,i):S^{1}\to M_{f} are locally injective every where except possibly at the basepoint ν\nu;

  4. 4.

    and for the trace of the basepoint hν=h(ν,):[L1,L2]Mfh^{\nu}=h(\nu,\cdot):[L_{1},L_{2}]\to M_{f}, no subpath between consecutive integer coordinates [i,i+1][i,i+1] is homotopic rel. endpoints into the edge-space.

This is the definition used in [23]. In light of our last description of Λ^k\hat{\Lambda}_{k} in the previous section, we give alternative definition. Let 𝒯\mathcal{T} be the Bass-Serre tree for FϕF*_{\phi} and \star be the point whose stabilizer is FF. An (algebraic) annulus ([α],pα)([\alpha],p_{\alpha}) in FϕF*_{\phi} of length L2L\geq 2 is a choice a nontrivial conjugacy class [α][\alpha] in FϕF*_{\phi} and an orbit of a geodesic path in pα𝒯p_{\alpha}\subset\mathcal{T} of length LL fixed by α\alpha. Since elements of [α][\alpha] act on 𝒯\mathcal{T} with fixed points, we can always choose a representative αF\alpha\in F. Technically, we have defined a conjugacy class of algebraic annuli but the distinction will not be relevant for us.

Lemma 6.2.

Let ϕ:FF\phi:F\to F be injective and f:ΓΓf:\Gamma\to\Gamma be a topological representative for [ϕ][\phi]. There is a one-to-one correspondence between (homotopy classes of) annuli in MfM_{f} of length L1L\geq 1 and (conjugacy classes of) annuli in FϕF*_{\phi} of length L+12L+1\geq 2.

Proof.

Given a topological annulus hh in MfM_{f} of length L1L\geq 1, then the generator of the image π1(h):π1(Mf)\pi_{1}(h):\mathbb{Z}\to\pi_{1}(M_{f}) determines a conjugacy class [α][\alpha] in π1(Mf)Fϕ\pi_{1}(M_{f})\cong F*_{\phi}. Condition (3) ensures π1(h)\pi_{1}(h) is injective and α\alpha is nontrivial. Let h~:×[L1,L2]M~f\tilde{h}:\mathbb{R}\times[L_{1},L_{2}]\to\tilde{M}_{f} be the lift of the annulus to the universal cover of MfM_{f}. Collapsing the Γ~\tilde{\Gamma}-direction of M~f\tilde{M}_{f} produces the Bass-Serre tree 𝒯\mathcal{T} and Condition (2) ensures the induced map h¯:×([L1,L2])𝒯\bar{h}:\mathbb{R}\times([L_{1},L_{2}]\cap\mathbb{Z})\to\mathcal{T} is constant on the first factor and its image is a collection of edge-midpoint; each ring hih_{i} determines a conjugacy class in the stabilizer of the corresponding edge-midpoint. By Conditions (1)(1) and (4)(4), the midpoints extend to a geodesic edge-path pαp_{\alpha} in 𝒯\mathcal{T} of length L+1L+1 fixed by α\alpha.

The other direction works in a similar fashion. For any conjugacy class [α][\alpha] in FϕF*_{\phi} and two consecutive edge-midpoints in 𝒯F\mathcal{T}_{F} fixed by αF\alpha\in F, we can construct an annulus of length 11 as follows. Fix a basepoint in the edge-space and assume \star is the vertex between the midpoints. If the midpoints are increasing/decreasing, then α=ϕ(x)ϕ(F)\alpha=\phi(x)\in\phi(F) without loss of generality. Let σ,ρ\sigma,\rho be based loop in the edge-space representing x,ϕ(x)Fx,\phi(x)\in F and τ\tau a based loop in MfM_{f} representing tFϕt\in F*_{\phi}. Then the based path στρ¯τ¯\sigma\cdot\tau\cdot\bar{\rho}\cdot\bar{\tau} is null-homotopic and can be extended to an annulus with ends σ,ρ\sigma,\rho and trace τ\tau. If the midpoints are at the same height, then α=ϕ(x)ϕ(F)\alpha=\phi(x)\in\phi(F) and α=gϕ(y)g1gϕ(F)g1\alpha=g\phi(y)g^{-1}\in g\phi(F)g^{-1} for some gϕ(F)g\notin\phi(F). Let σ,ρ,γ\sigma,\rho,\gamma be based loops in the edge-space representing x,y,gFx,y,g\in F respectively and τ\tau a based loop in MfM_{f} representing tFϕt\in F*_{\phi}. So the based path στγτ¯ρ¯τγ¯τ¯\sigma\cdot\tau\gamma\bar{\tau}\cdot\bar{\rho}\cdot\tau\bar{\gamma}\bar{\tau} is null-homotopic and can be extended to an homotopy between σ,ρ\sigma,\rho satisfying Conditions (1)-(3) and having trace τγτ¯\tau\gamma\bar{\tau}. This trace satisfies Condition (4) because tgt1Ftgt^{-1}\notin F and thus the homotopy is a topological annulus with ends σ,ρ\sigma,\rho. Given a geodesic path in 𝒯\mathcal{T} of length L2L\geq 2 fixed by α\alpha, we can replace the path with a translate in 𝒯F\mathcal{T}_{F} without affecting the class [α][\alpha] and then construct a topological annulus in MfM_{f} of length L1L-1 by concatenating the length 1 annuli from the preceding discussion. This concludes the correspondence between topological and algebraic annuli. ∎

The natural orientation on 𝒯\mathcal{T} gives a dichotomy for annuli ([α],pα)([\alpha],p_{\alpha}) in FϕF*_{\phi}:

  1. 1.

    all edges of pαp_{\alpha} have the same orientation— we say α\alpha is unidirectional.

  2. 2.

    pαp_{\alpha} switches from increasing to decreasing exactly once — we say α\alpha is bidirectional.

Each vertex of 𝒯\mathcal{T} has exactly one outgoing edge and hence the geodesic pαp_{\alpha} cannot switch from decreasing to increasing because [F:F]=1[F:F]=1. For similar reasons, bidirectional annuli do not exist if and only if ϕ(F)\phi(F) is malnormal in FF. The next proposition generalizes this equivalence of bidirectional annuli (or lack thereof) and malnormality.

An annulus ([α],pα)([\alpha],p_{\alpha}) in FϕF*_{\phi} is strictly bidirectional if the switch from increasing to decreasing occurs at the midpoint of pαp_{\alpha}.

Lemma 6.3.

Let ϕ:FF\phi:F\to F be injective. For any integer L1L\geq 1, the mapping torus FϕF*_{\phi} has a strictly bidirectional annulus of length 2L2L if and only if Λ^L[ϕ]\hat{\Lambda}_{L}[\phi] is not empty.

Proof.

If there is a strictly bidirectional annulus ([α],pα)([\alpha],p_{\alpha}) in FϕF*_{\phi} of length 2L2L, then we may assume the midpoint of pαp_{\alpha} is \star after replacing pαp_{\alpha} with a translate. Then the stabilizer of pαp_{\alpha} contains α\alpha and so it is nontrivial. Since the stabilizer of pαp_{\alpha} is the stabilizer of its (ordered) endpoints, the conjugacy class of the stabilizer is a component of Λ^L[ϕ]\hat{\Lambda}_{L}[\phi].

If Λ^L[ϕ]\hat{\Lambda}_{L}[\phi] is not empty, then some path in 𝒯F\mathcal{T}_{F} of length 2L2L with midpoint at \star has a nontrivial stabilizer. Choose a nontrivial element α\alpha in this stabilizer and ([α],pα)([\alpha],p_{\alpha}) is a strictly bidirectional annulus in FϕF*_{\phi} of length 2L2L. ∎

Let ϕ:FF\phi:F\to F be injective and f:ΓΓf:\Gamma\to\Gamma be a topological representative for ϕ\phi. If [ϕ][\phi] has no invariant cyclic subgroup system with index d2d\geq 2, then there is an integer L1L\geq 1 for which Λ^L[ϕ]\hat{\Lambda}_{L}[\phi] is empty (Proposition 5.5) and all strictly bidirectional annuli in MfM_{f} are shorter than 2L2L (Lemma 6.3). This sets up the second hypothesis of Theorem 6.1.

As for the first hypothesis, we begin by defining (based-) hyperbolicity. For a real number λ>1\lambda>1 and integer n1n\geq 1, we say a graph map f:ΓΓf:\Gamma\to\Gamma is (based-) (λ,n)\boldsymbol{(\lambda,n)}-hyperbolic if all (based) loops σ:S1Γ\sigma:S^{1}\to\Gamma (with the basepoint mapped to a vertex) satisfy the inequality

λ|fn(σ)|max(|f2n(σ)|,|σ|)\lambda\,|f^{n}(\sigma)|\leq\max(\,|f^{2n}(\sigma)|,|\sigma|\,)~{}

where |||\cdot| is the combinatorial length after tightening; whether tightening respects a basepoint (based homotopy) or not (free homotopy) will be apparent from the context.

When a graph map is (based-) (λ,n)(\lambda^{\prime},n)-hyperbolic for some λ>1,n1\lambda^{\prime}>1,n\geq 1, then it is (based-) (λk,nk)(\lambda^{k},nk)-hyperbolic for all k1k\geq 1 and λ(1,λ]\lambda\in(1,\lambda^{\prime}]. So the constants can be omitted and when we do need them, we can assume λ>1\lambda>1 is any preferred integer. As defined, hyperbolicity is a property of the homotopy class [f][f]. Meanwhile, based-hyperbolicity is a property of the map ff.

A graph map f:ΓΓf:\Gamma\to\Gamma is atoroidal if it is π1\pi_{1}-injective and there is no nontrivial loop σ\sigma in Γ\Gamma and integer k1k\geq 1 such that fk(σ)σf^{k}(\sigma)\simeq\sigma. This again is a property of the homotopy class [f][f]. Bestvina-Feighn-Handel showed, as a step in [4, Theorem 5.1], that hyperbolic atoroidal homotopy equivalences are based-hyperbolic. Their argument is reproduced here, modified to drop the π1\pi_{1}-surjectivity assumption. This allows us to consider the growth rate of loops without basepoints for the rest of the section

Lemma 6.4.

If the graph map f:ΓΓf:\Gamma\to\Gamma is atoroidal and (3,n)(3,n)-hyperbolic, then it is based-(2,n)(2,n^{\prime})-hyperbolic.

To avoid context-ambiguity in the proof, we use \lVert\cdot\rVert for lengths of free homotopy classes of loops and |||\cdot| for lengths of loops rel. basepoints. However, the distinction is not needed after the proof as all loops afterwards will be considered up to free homotopy.

Proof.

Suppose f:ΓΓf:\Gamma\to\Gamma is atoroidal and (3,n)(3,n)-hyperbolic for some integer n1n\geq 1. Set MM to be the maximum length of fk(s)f^{k}(s) rel. basepoint over all embedded based loops ss in Γ\Gamma for k{0,n,2n}k\in\{0,n,2n\}.

Suppose |fn(σ)|4M|f^{n}(\sigma)|\geq 4M for some immersed based loop σ\sigma and pick an embedded based loop ss with same basepoint as σ\sigma so that the concatenation sσs\cdot\sigma is an immersed loop, i.e., sσ=|s|+|σ|\lVert s\cdot\sigma\rVert=|s|+|\sigma|. As the graph map ff is (3,n)(3,n)-hyperbolic, we get

3fn(sσ)max(f2n(sσ),sσ).3\,\lVert f^{n}(s\cdot\sigma)\rVert\leq\max(\,\lVert f^{2n}(s\cdot\sigma)\rVert,\lVert s\cdot\sigma\rVert\,).

For a concatenation of based loops ρ1ρ2\rho_{1}\cdot\rho_{2}, we get ||ρ1||ρ2||ρ1ρ2|ρ1|+|ρ2|\left|\,|\rho_{1}|-|\rho_{2}|\,\right|\leq\lVert\rho_{1}\cdot\rho_{2}\rVert\leq|\rho_{1}|+|\rho_{2}|.

Case 1. If f2n(sσ)3fn(sσ)\lVert f^{2n}(s\cdot\sigma)\rVert\geq 3\,\lVert f^{n}(s\cdot\sigma)\rVert, then

|f2n(σ)|f2n(sσ)|f2n(s)|\displaystyle|f^{2n}(\sigma)|\geq\lVert f^{2n}(s\cdot\sigma)\rVert-|f^{2n}(s)| 3fn(sσ)M\displaystyle\geq 3\,\lVert f^{n}(s\cdot\sigma)\rVert-M
3|fn(σ)|3|fn(s)|M\displaystyle\geq 3\,|f^{n}(\sigma)|-3\,|f^{n}(s)|-M
3|fn(σ)|4M\displaystyle\geq 3\,|f^{n}(\sigma)|-4M
2|fn(σ)|.\displaystyle\geq 2\,|f^{n}(\sigma)|.

Case 2. If sσ3fn(sσ)\lVert s\cdot\sigma\rVert\geq 3\,\lVert f^{n}(s\cdot\sigma)\rVert, then

|σ|=sσ|s|3fn(sσ)M3|fn(σ)|4M2|fn(σ)||\sigma|=\lVert s\cdot\sigma\rVert-|s|\geq 3\,\lVert f^{n}(s\cdot\sigma)\rVert-M\geq 3\,|f^{n}(\sigma)|-4M\geq 2\,|f^{n}(\sigma)|

Combining both cases: 2|fn(σ)|max(|f2n(σ)|,|σ|)2\,|f^{n}(\sigma)|\leq\max(\,|f^{2n}(\sigma)|,|\sigma|\,). If |fnk(σ)|4M|f^{nk}(\sigma)|\geq 4M for an immersed based loop σ\sigma in Γ\Gamma and k1k\geq 1, then by induction

2k|fnk(σ)|max(|f2nk(σ)|,|σ|).2^{k}\,|f^{nk}(\sigma)|\leq\max(\,|f^{2nk}(\sigma)|,|\sigma|\,).

For any bound BB, there are only finitely many immersed based loops σ\sigma^{\prime} in Γ\Gamma with |σ|B|\sigma^{\prime}|\leq B. Since ff is atoroidal, there is an integer k0k\gg 0 such that |fnk(σ)|8M|f^{nk}(\sigma^{\prime})|\geq 8M for every based loop σ\sigma^{\prime} with |σ|4M|\sigma^{\prime}|\leq 4M and we conclude that ff is based-(2,nk)(2,nk)-hyperbolic . ∎

When pp is a subpath of an immersed loop σ\sigma and n1n\geq 1, then [fn(p)]σ[f^{n}(p)]_{\sigma} is the subpath of [fn(p)][f^{n}(p)] that survives in [fn(σ)][f^{n}(\sigma)] and |fn(p)|σ|f^{n}(p)|_{\sigma} is the length of [fn(p)]σ[f^{n}(p)]_{\sigma}. Bounded cancellation implies |fn(p)||fn(p)|σ+2C(fn)|f^{n}(p)|\leq|f^{n}(p)|_{\sigma}+2C(f^{n}). The next lemma is based on Brinkmann’s Lemma 4.2 in [8] with a few changes made to account for graph maps that are not π1\pi_{1}-surjective.

Lemma 6.5.

Let f:ΓΓf:\Gamma\to\Gamma be a graph map and RΓR_{*}\subset\Gamma be an ff-invariant union of roses such that the restriction f|R:RR\left.f\right|_{R_{*}}:R_{*}\to R_{*} is (4,n)(4,n)-hyperbolic. For some constant LcL_{c}, if pRp\subset R_{*} is a subpath (edge-path) of some immersed loop σ\sigma in Γ\Gamma and |fn(p)|σLc|f^{n}(p)|_{\sigma}\geq L_{c}, then

3|fn(p)|σmax(|f2n(p)|σ,|p|).3\,|f^{n}(p)|_{\sigma}\leq\max\left(\,|f^{2n}(p)|_{\sigma},|p|\,\right).

The number LcL_{c} is the critical length of the triple (f,Γ,R)(f,\Gamma,R_{*}).

Proof.

Let f:ΓΓf:\Gamma\to\Gamma be a graph map and RΓR_{*}\subset\Gamma be an ff-invariant union of roses such that the restriction f|R\left.f\right|_{R_{*}} is (4,n)(4,n)-hyperbolic. Set MM to be the maximum length of fk(s)f^{k}(s) rel. basepoint over all petals ss in RR_{*} and k{n,2n}k\in\{n,2n\}. Choose Lc=2C(f2n)+5ML_{c}=2C(f^{2n})+5M where C(f2n)C(f^{2n}) is the cancellation constant for f2nf^{2n}. Recall the triangle inequality: ||p1||p2|||p1p2||p1|+|p2|\left|\,|p_{1}|-|p_{2}|\,\right|\leq|p_{1}\cdot p_{2}|\leq|p_{1}|+|p_{2}| for any path decomposition of a loop p1p2p_{1}\cdot p_{2}. A remark on the context: paths [pi][p_{i}] are tightened rel. endpoints but the loop [p1p2][p_{1}\cdot p_{2}] is tightened by free homotopy.

Given a subpath pRp\subset R_{*} of some immersed loop σ\sigma in Γ\Gamma, pick a petal ss in RR_{*} such that sps\cdot p is an immersed loop in RR_{*}. As the restriction to RR_{*} is (4,n)(4,n)-hyperbolic, we get

4|fn(sp)|max(|f2n(sp)|,|sp|).4\,|f^{n}(s\cdot p)|\leq\max(\,|f^{2n}(s\cdot p)|,|s\cdot p|\,).

If 4|fn(sp)||f2n(sp)|4\,|f^{n}(s\cdot p)|\leq|f^{2n}(s\cdot p)|, then

4|fn(p)|σ4|fn(p)|\displaystyle 4\,|f^{n}(p)|_{\sigma}\leq 4\,|f^{n}(p)| 4|fn(sp)|+4|fn(s)|\displaystyle\leq 4\,|f^{n}(s\cdot p)|+4\,|f^{n}(s)|
|f2n(sp)|+4M\displaystyle\leq|f^{2n}(s\cdot p)|+4M
|f2n(p)|+5M\displaystyle\leq|f^{2n}(p)|+5M
|f2n(p)|σ+2C(f2n)+5M(bounded cancellation)\displaystyle\leq|f^{2n}(p)|_{\sigma}+2C(f^{2n})+5M\quad\text{(bounded cancellation)}

Similarly, if 4|fn(sp)||sp|4\,|f^{n}(s\cdot p)|\leq|s\cdot p|, then

4|fn(p)|σ\displaystyle 4\,|f^{n}(p)|_{\sigma} |p|+1+4M\displaystyle\leq|p|+1+4M
|p|+5M+2C(f2n)(since M1)\displaystyle\leq|p|+5M+2C(f^{2n})\quad(\text{since }M\geq 1)

Since Lc=2C(f2n)+5ML_{c}=2C(f^{2n})+5M, we have the desired implication:

|fn(p)|σLc3|fn(p)|σmax(|f2n(p)|σ,|p|).|f^{n}(p)|_{\sigma}\geq L_{c}\implies 3\,|f^{n}(p)|_{\sigma}\leq\max\left(\,|f^{2n}(p)|_{\sigma},|p|\,\right).\qed

An (outer class of an) injective endomorphism ϕ:FF\phi:F\to F is atoroidal if it has no invariant cyclic subgroup system with index d=1d=1. If f:ΓΓf:\Gamma\to\Gamma is a topological representative for ϕ\phi, then [f][f] is atoroidal if and only if [ϕ][\phi] is atoroidal. The following proposition is an extension of Brinkmann’s result [8, Proposition 7.1] and is the main technical result of the section.

Proposition 6.6.

If ϕ:FF\phi:F\to F is injective and atoroidal, then [ϕ][\phi] has a (2,n)(2,n)-hyperbolic topological representative for some integer n1n\geq 1.

Proof.

Suppose ϕ:FF\phi:F\to F is an injective and atoroidal endomorphism. If ϕ\phi is surjective, then the proposition is precisely Brinkmann’s result. So we assume ϕ\phi is not surjective. By Theorem 4.5, there is an expanding 𝒜\mathcal{A}^{*}-relative immersion g:TTg:T\to T for ϕ\phi, where 𝒜\mathcal{A}^{*} is the [ϕ][\phi]-elliptic free factor system. Fix some 𝒜\mathcal{A}^{*}-marked roses (R𝒜,α𝒜)(R_{\mathcal{A}^{*}},\alpha_{\mathcal{A}^{*}}) and set (Γ,α)(\Gamma,\alpha) to be the (R𝒜,α𝒜)(R_{\mathcal{A}^{*}},\alpha_{\mathcal{A}^{*}})-vertex blow-up of the graph of groups F\TF\backslash T, i.e., (Γ,α)(\Gamma,\alpha) is a marked graph with subgraphs R𝒜R_{\mathcal{A}^{*}} corresponding to the free factor system α(𝒜)\alpha(\mathcal{A}^{*}). The roses R𝒜R_{\mathcal{A}^{*}} form the lower stratum of Γ\Gamma and the remaining edges the top stratum.

We outline the proof which follows closely Brinkmann’s strategy. Patch together a homotopy equivalence of the lower stratum with the expanding relative immersion to get some topological representative ff of [ϕ][\phi]. By Brinkmann’s theorem, the restriction of ff to the lower stratum is hyperbolic. The expanding relative immersion on the top stratum means loops that are mostly top stratum will have uniform exponential growth under forward iteration. Lemma 6.5 implies loops that are mostly lower stratum will have uniform exponential growth under forward and/or backward iteration. The heart of the proof lies in quantifying what being mostly top or lower stratum means and showing that all loops are one or the other. Of course, there are a few minor technicalities that need addressing; for instance, the restriction to the lower stratum is almost but not exactly a homotopy equivalence. This concludes the outline.

Recall that the maximal [ϕ][\phi]-fixed free factor system 𝒜\mathcal{A} is a subset of 𝒜\mathcal{A}^{*} and there is an integer k00k_{0}\geq 0 such that ϕk0(𝒜)\phi^{k_{0}}(\mathcal{A}^{*}) is carried by 𝒜\mathcal{A} (Proposition 2.4). So we may find a topological representative f𝒜:R𝒜R𝒜f_{\mathcal{A}^{*}}:R_{\mathcal{A}^{*}}\to R_{\mathcal{A}^{*}} for [ϕ|𝒜][\left.\phi\right|_{\mathcal{A}^{*}}\,] whose restriction to the periodic roses R𝒜R_{\mathcal{A}}, denoted by f𝒜f_{\mathcal{A}}, is a homotopy equivalence. As ϕ\phi is atoroidal, the restriction f𝒜f_{\mathcal{A}} is (4,n0)(4,n_{0})-hyperbolic for some n01n_{0}\geq 1 (Brinkmann’s theorem).

If σ\sigma is an immersed loop in R𝒜R_{\mathcal{A}^{*}}, then [fk(σ)][f^{k}(\sigma)] is a loop in the periodic roses R𝒜R_{\mathcal{A}} for all kk0k\geq k_{0}. Since the restriction f𝒜f_{\mathcal{A}} is (4,n0)(4,n_{0})-hyperbolic and f𝒜n0k(σ)f_{\mathcal{A}^{*}}^{n_{0}k}(\sigma) is a loop in R𝒜R_{\mathcal{A}} for any loop σ\sigma in R𝒜R_{\mathcal{A}^{*}} and kk0k\geq k_{0}, we get the inequality

4|f𝒜n0(f𝒜n0k(σ))|max(|f𝒜2n0(f𝒜n0k(σ))|,|f𝒜n0k(σ)|) for all loops σ in R𝒜 and kk0.\displaystyle 4\cdot|f_{\mathcal{A}}^{n_{0}}(f_{\mathcal{A}^{*}}^{n_{0}k}(\sigma))|\leq\max(\,|f_{\mathcal{A}}^{2n_{0}}(f_{\mathcal{A}^{*}}^{n_{0}k}(\sigma))|,|f_{\mathcal{A}^{*}}^{n_{0}k}(\sigma)|\,)\text{ for all loops }\sigma\text{ in }R_{\mathcal{A}^{*}}\text{ and }k\geq k_{0}.

Choose an integer k11k_{1}\geq 1 so that 4k14Kn0k04^{k_{1}}\geq 4K^{n_{0}k_{0}}, where K=K(f𝒜)K=K(f_{\mathcal{A}^{*}}) is the Lipschitz constant for f𝒜f_{\mathcal{A}^{*}}. Suppose σ\sigma is a loop in R𝒜R_{\mathcal{A}^{*}}.

If 4|f𝒜n0(f𝒜n0(k0+k11)(σ))||f𝒜2n0(f𝒜n0(k0+k11)(σ))|4\cdot|f_{\mathcal{A}}^{n_{0}}(f_{\mathcal{A}^{*}}^{n_{0}(k_{0}+k_{1}-1)}(\sigma))|\leq|f_{\mathcal{A}}^{2n_{0}}(f_{\mathcal{A}^{*}}^{n_{0}(k_{0}+k_{1}-1)}(\sigma))|, then by induction

4k0+k1|f𝒜n0(k0+k1)(σ)||f𝒜2n0(k0+k1)(σ)|.4^{k_{0}+k_{1}}\cdot|f_{\mathcal{A}^{*}}^{n_{0}(k_{0}+k_{1})}(\sigma)|\leq|f_{\mathcal{A}^{*}}^{2n_{0}(k_{0}+k_{1})}(\sigma)|.

If 4|f𝒜n0(f𝒜n0(k0+k11)(σ))||f𝒜n0(k0+k11)(σ)|4\cdot|f_{\mathcal{A}}^{n_{0}}(f_{\mathcal{A}^{*}}^{n_{0}(k_{0}+k_{1}-1)}(\sigma))|\leq|f_{\mathcal{A}^{*}}^{n_{0}(k_{0}+k_{1}-1)}(\sigma)|, then by induction and Lipschitz property

4k1|f𝒜n0(k0+k1)(σ)||f𝒜n0k0(σ)|Kn0k0|σ|4^{k_{1}}\cdot|f_{\mathcal{A}^{*}}^{n_{0}(k_{0}+k_{1})}(\sigma)|\leq|f_{\mathcal{A}^{*}}^{n_{0}k_{0}}(\sigma)|\leq K^{n_{0}k_{0}}\cdot|\sigma|

and 4|f𝒜n0(k0+k1)(σ)||σ| by choice of k1.4\cdot|f_{\mathcal{A}^{*}}^{n_{0}(k_{0}+k_{1})}(\sigma)|\leq|\sigma|\text{ by choice of }k_{1}.

Therefore, the lower stratum map f𝒜f_{\mathcal{A}^{*}} is (4,n1)(4,n_{1})-hyperbolic with n1=n0(k0+k1)n_{1}=n_{0}(k_{0}+k_{1}).

Let f:ΓΓf:\Gamma\to\Gamma be a topological representative for [ϕ][\phi] that extends f𝒜f_{\mathcal{A}^{*}} to the top stratum and induces the expanding 𝒜\mathcal{A}^{*}-relative immersion g:TTg:T\to T upon collapsing the lower stratum in the universal cover Γ~\tilde{\Gamma}. For an arbitrary immersed loop σ\sigma in Γ\Gamma, define σtop\sigma_{top} (σlow\sigma_{low} resp.) to be the collection of maximal subpaths of σ\sigma in the top (lower resp.) stratum. For all n1n\geq 1, define [fn(σtop)]σ[f^{n}(\sigma_{top})]_{\sigma} ([fn(σlow)]σ[f^{n}(\sigma_{low})]_{\sigma} resp.) to be the collection of paths [fn(p)]σ[f^{n}(p)]_{\sigma} where pp is some path in σtop\sigma_{top} (σlow\sigma_{low} resp.). That ff induces an immersion gg upon collapsing the lower stratum implies that the top stratum is persistent: if σ\sigma is an immersed loop in Γ\Gamma, then f(σ)topf(\sigma)_{top} survives in [f(σ)][f(\sigma)].

As the relative immersion g:TTg:T\to T is expanding, there is an integer k21k_{2}\geq 1, such that gk2(e)g^{k_{2}}(e) has length 2\geq 2 for all edges ee in TT; and as ff induces gg, for any immersed loop σ\sigma in Γ\Gamma and path pp in σtop\sigma_{top}, we get 2|p|σ|fk2(p)|σ2|p|_{\sigma}\leq|f^{k_{2}}(p)|_{\sigma}. We may replace n1n_{1} and k2k_{2} with a common multiple and assume n1=k2n_{1}=k_{2}. A similar inequality holds in the lower stratum. By the (4,n1)(4,n_{1})-hyperbolicity of f|R𝒜\left.f\right|_{R_{\mathcal{A}^{*}}} and Lemma 6.5, there is a critical length Lc=Lc(f,Γ,R𝒜)L_{c}=L_{c}(f,\Gamma,R_{\mathcal{A}^{*}}) such that for any immersed loop σ\sigma in Γ\Gamma and path pp in σlow\sigma_{low},

|fn1(p)|σLc3|fn1(p)|σmax(|f2n1(p)|σ,|p|).|f^{n_{1}}(p)|_{\sigma}\geq L_{c}\implies 3|f^{n_{1}}(p)|_{\sigma}\leq\max(\,|f^{2n_{1}}(p)|_{\sigma},|p|\,).

Set MM to be the maximal length amongst all paths in fn1(e)lowf^{n_{1}}(e)_{low} for all top stratum edges ee of Γ\Gamma. For any integer k1k\geq 1, we distinguish two cases:

Case 1. If |fn1k(σ)low|(Lc+6M)|fn1k(σ)top||f^{n_{1}k}(\sigma)_{low}|\leq(L_{c}+6M)\,|f^{n_{1}k}(\sigma)_{top}|, then

|fn1k(σ)|=|fn1k(σ)low|+|fn1k(σ)top|\displaystyle|f^{n_{1}k}(\sigma)|=|f^{n_{1}k}(\sigma)_{low}|+|f^{n_{1}k}(\sigma)_{top}| (Lc+6M+1)|fn1k(σ)top| and\displaystyle\leq(L_{c}+6M+1)\,|f^{n_{1}k}(\sigma)_{top}|\text{ and }
2k|fn1k(σ)top||fn1k(fn1k(σ)top)|σ\displaystyle 2^{k}\,|f^{n_{1}k}(\sigma)_{top}|\leq|f^{n_{1}k}(f^{n_{1}k}(\sigma)_{top})|_{\sigma} |f2n1k(σ)|.\displaystyle\leq|f^{2n_{1}k}(\sigma)|.

Additionally, if 2k2(Lc+6M+1)2^{k}\geq 2(L_{c}+6M+1), then 2|fn1k(σ)||f2n1k(σ)|2\,|f^{n_{1}k}(\sigma)|\leq|f^{2n_{1}k}(\sigma)|.

Case 2. Suppose |fn1k(σ)low|(Lc+6M)|fn1k(σ)top||f^{n_{1}k}(\sigma)_{low}|\geq(L_{c}+6M)\,|f^{n_{1}k}(\sigma)_{top}|. Set mm to be the number of paths in fn1k(σ)lowf^{n_{1}k}(\sigma)_{low}. Then mm is also the number of paths in fn1k(σ)topf^{n_{1}k}(\sigma)_{top} and m|fn1k(σ)top|m\leq|f^{n_{1}k}(\sigma)_{top}|. By the pigeonhole principle, some path ρ\rho in fn1k(σ)lowf^{n_{1}k}(\sigma)_{low} satisfies |ρ|Lc+6M|\rho|\geq L_{c}+6M. As |ρ|6M|\rho|\geq 6M, we have 3(|ρ|2M)2|ρ|3(|\rho|-2M)\geq 2|\rho|. Set σ=fn1(k1)(σ)\sigma^{\prime}=f^{n_{1}(k-1)}(\sigma). By definition of MM and persistence of fn1(σ)topf^{n_{1}}(\sigma^{\prime})_{top}, there must be a path pp^{\prime} in σlow\sigma^{\prime}_{low} such that [fn1(p)]σ[f^{n_{1}}(p^{\prime})]_{\sigma^{\prime}} is a subpath of ρ\rho, |fn1(p)|σ|ρ|2MLc|f^{n_{1}}(p^{\prime})|_{\sigma^{\prime}}\geq|\rho|-2M\geq L_{c}, and 3|fn1(p)|σmax(|f2n1(p)|σ,|p|).3\,|f^{n_{1}}(p^{\prime})|_{\sigma^{\prime}}\leq\max(\,|f^{2n_{1}}(p^{\prime})|_{\sigma^{\prime}},|p^{\prime}|\,).

If |f2n1(p)|σ3|fn1(p)|σ|f^{2n_{1}}(p^{\prime})|_{\sigma^{\prime}}\geq 3|f^{n_{1}}(p^{\prime})|_{\sigma^{\prime}}, then |f2n1(p)|σ2|ρ||f^{2n_{1}}(p^{\prime})|_{\sigma^{\prime}}\geq 2|\rho| and |fn1(k+1)(p)|σ3k12|ρ||f^{n_{1}(k+1)}(p^{\prime})|_{\sigma^{\prime}}\geq 3^{k-1}\cdot 2|\rho|.

If |p|3|fn1(p)|σ|p^{\prime}|\geq 3|f^{n_{1}}(p^{\prime})|_{\sigma^{\prime}}, then |p|2|ρ||p^{\prime}|\geq 2|\rho|. By inducting on the same argument used at the start of the case, there must be a path pp in σlow\sigma_{low} such that |fn1k(p)|σ|f^{n_{1}k}(p)|_{\sigma} is a subpath of ρ\rho and |p|2k|ρ||p|\geq 2^{k}\,|\rho|. In either case, we get 2k|ρ|max(|fn1(k+1)(p)|σ,|p|)2^{k}|\rho|\leq\max(\,|f^{n_{1}(k+1)}(p^{\prime})|_{\sigma^{\prime}},|p|\,). Define fn1k(σ)critf^{n_{1}k}(\sigma)_{crit} to be the set of paths ρ\rho in fn1k(σ)lowf^{n_{1}k}(\sigma)_{low} with |ρ|Lc+6M|\rho|\geq L_{c}+6M. Altogether, we have shown:

2k|fn1k(σ)crit|max(|f2n1k(σ)|,|σ|).2^{k}|f^{n_{1}k}(\sigma)_{crit}|\leq\max(\,|f^{2n_{1}k}(\sigma)|,|\sigma|\,).

The following computation is lifted from Brinkmann [8, Proof of Proposition 7.1]. Set A=|fn1k(σ)crit|A=|f^{n_{1}k}(\sigma)_{crit}| to be the total length of paths in fn1k(σ)critf^{n_{1}k}(\sigma)_{crit}, B=|fn1k(σ)low|AB=|f^{n_{1}k}(\sigma)_{low}|-A to be the total length of the remaining paths in fn1k(σ)lowf^{n_{1}k}(\sigma)_{low}, and C=|fn1k(σ)top|C=|f^{n_{1}k}(\sigma)_{top}|. We now find a positive lower bound of AA+B+C\frac{A}{A+B+C} that is independent of σ\sigma and kk. We assumed A+B(Lc+6M)CA+B\geq(L_{c}+6M)\,C and so AA+B+CA(Lc+6M)(A+B)(Lc+6M+1)\frac{A}{A+B+C}\geq\frac{A(L_{c}+6M)}{(A+B)(L_{c}+6M+1)} and we can focus on the factor AA+B=1BA+B\frac{A}{A+B}=1-\frac{B}{A+B}. Recall mCm\leq C, so A+B(Lc+6M)mA+B\geq(L_{c}+6M)\,m. Since each path pp in fn1k(σ)lowf^{n_{1}k}(\sigma)_{low} but not in fn1k(σ)critf^{n_{1}k}(\sigma)_{crit} satisfies |p|<Lc+6M|p|<L_{c}+6M and there are at most mm of them, Bm(Lc+6M1)B\leq m\,(L_{c}+6M-1). Combining the last two inequalities gives the bound

1BA+B1m(Lc+6M1)m(Lc+6M)1Lc+6M.1-\frac{B}{A+B}\geq 1-\frac{m\,(L_{c}+6M-1)}{m\,(L_{c}+6M)}\geq\frac{1}{L_{c}+6M}.

Altogether, AA+B+C1Lc+6M+1\frac{A}{A+B+C}\geq\frac{1}{L_{c}+6M+1}.

Additionally, if 2k2(Lc+6M+1)2^{k}\geq 2(L_{c}+6M+1), then 2|fn1k(σ)|max(|f2n1k(σ)|,|σ|)2|f^{n_{1}k}(\sigma)|\leq\max(\,|f^{2n_{1}k}(\sigma)|,|\sigma|\,).

Choose k1k\geq 1 so that 2k2(Lc+6M+1)2^{k}\geq 2(L_{c}+6M+1); the two exhaustive cases above imply ff is (2,n1k)(2,n_{1}k)-hyperbolic. ∎

All the heavy lifting is done and we can prove our main theorem

Theorem 6.7.

Let ϕ:FF\phi:F\to F be an injective endomorphism. Then the following statements are equivalent:

  1. 1.

    FϕF*_{\phi} is word-hyperbolic;

  2. 2.

    FϕF*_{\phi} contains no BS(1,d)BS(1,d) subgroups with d1d\geq 1;

  3. 3.

    [ϕ][\phi] has no invariant cyclic subgroup system with index d1d\geq 1;

  4. 4.

    all/some topological representatives of [ϕ][\phi] are based-hyperbolic and all strictly bidirectional annuli in their mapping tori are shorter than some integer.

Proof.

This proof is mostly a matter of bookkeeping.

(1)(2)(1){\implies}(2): BS(1,d)BS(1,d) subgroups are well-known obstructions to word-hyperbolicity.

(2)(3)(2){\implies}(3): If [ϕ][\phi] has an invariant cyclic subgroup system with index d1d\geq 1, then there is a subgroup of FϕF*_{\phi} isomorphic to a quotient of BS(1,d)BS(1,d); use normal forms to show the subgroup is in fact isomorphic to BS(1,d)BS(1,d) (See [17, Lemma 2.3] for details).

(3)(4)(3){\implies}(4): Suppose [ϕ][\phi] has no invariant cyclic subgroup system with index d1d\geq 1. Then, by Proposition 6.6, [ϕ][\phi] has a (2,n)(2,n)-hyperbolic topological representative f:ΓΓf:\Gamma\to\Gamma defined on a marked graph (Γ,α)(\Gamma,\alpha). That is, for (any conjugacy class) of xFx\in F, we have

2ϕn(x)αmax(ϕ2n(x)α,xα).2\,\|\phi^{n}(x)\|_{\alpha}\leq\max(\|\phi^{2n}(x)\|_{\alpha},\|x\|_{\alpha}).

For any marked graph (Γ,β)(\Gamma^{\prime},\beta), choose difference of markings g:ΓΓg:\Gamma\to\Gamma^{\prime} and h:ΓΓh:\Gamma^{\prime}\to\Gamma, i.e., graph maps such that [π1(g)α]=[β][\pi_{1}(g)\circ\alpha]=[\beta] and [π1(h)β]=[α][\pi_{1}(h)\circ\beta]=[\alpha]. Then K=max(K(g),K(h))K=\max(K(g),K(h)) satisfies 1KxβxαKxβ\frac{1}{K}\|x\|_{\beta}\leq\|x\|_{\alpha}\leq K\|x\|_{\beta} for all xFx\in F, where K(g),K(h)K(g),K(h) are the respective Lipschitz constants. Fix an integer m1m\geq 1 such that 2m>K22^{m}>K^{2}, then for every xFx\in F,

2mKϕmn(x)βmax(Kϕ2mn(x)β,Kxβ)\frac{2^{m}}{K}\|\phi^{mn}(x)\|_{\beta}\leq\max(K\,\|\phi^{2mn}(x)\|_{\beta},K\,\|x\|_{\beta})

and (Γ,β,[ϕ])(\Gamma^{\prime},\beta,[\phi]) is (2mK2,mn)(2^{m}K^{-2},mn)-hyperbolic, i.e., all topological representatives of [ϕ][\phi] on (Γ,β)(\Gamma^{\prime},\beta) are hyperbolic. Since [ϕ][\phi] is atoroidal, all topological representatives of [ϕ][\phi] on (Γ,β)(\Gamma^{\prime},\beta) are based-hyperbolic (Lemma 6.4). As (Γ,β)(\Gamma^{\prime},\beta) was arbitrary, all topological representatives of [ϕ][\phi] are based-hyperbolic.

Let f:ΓΓf^{\prime}:\Gamma^{\prime}\to\Gamma^{\prime} be any topological representative for [ϕ][\phi]. Since [ϕ][\phi] has no invariant cyclic subgroup with index d2d\geq 2, all strictly bidirectional annuli in MfM_{f^{\prime}} are shorter than some integer (Proposition 5.5 with Lemmas 6.3 and 6.2).

(4)(1)(4){\implies}(1) — Theorem 6.1. See Theorem 7.7 below for a sketch of the proof. ∎

7 HNN extensions over free factors

In the last section, we extend the previous characterization to HNN extensions of free groups defined over free factors. Precisely, let AFA\leq F be a (nontrivial) free factor and ϕ:AF\phi:A\to F be an injective homomorphism. The HNN extension of FF over AA is given by:

FA=F,t|t1xt=ϕ(x),xA.F*_{A}=\langle F,t~{}|~{}t^{-1}xt=\phi(x),\forall x\in A\rangle.

HNN extensions, just like mapping tori, have associated Bass-Serre trees which we can use to define algebraic annuli as in the previous section. Unlike the Bass-Serre tree of a mapping torus, that of an HNN extension need not behave like a rooted tree: generally, the number of outgoing and incoming edges at a vertex will match the indices of AA and ϕ(A)\phi(A) in FF respectively. However, since free factors are malnormal, these HNN extensions share with mapping tori the dichotomy of annuli: annuli are either unidirectional or bidirectional.

We will say a conjugacy class of elements [g][g] in FF has one forward iterate if it is contains an element aAa\in A. By malnormality of AA, the conjugacy class of [ϕ(a)][\phi(a)] depends only on the class [g][g]. We set the forward iterate of such a class [g][g] to be ϕ[g]=[ϕ(a)]\phi\cdot[g]=[\phi(a)]. Inductively, [g][g] has n+1n+1 forward iterates if it has nn forward iterates and the iterate ϕn[g]\phi^{n}\cdot[g] has one forward iterate; in this case, set ϕn+1[g]=ϕ(ϕn[g])\phi^{n+1}\cdot[g]=\phi\cdot(\phi^{n}\cdot[g]).

Similar to our use of infinite tails (backward iteration) to construct a canonical fixed free factor system in Section 2, we shall use forward iteration to construct a canonical invariant free factor system of AA. The proof will use a simpler variation of descent.

Proposition 7.1.

If AFA\leq F is a free factor and ϕ:AF\phi:A\to F is injective, then there is a unique maximal [ϕ][\phi]-invariant free factor system of AA. Precisely, \mathcal{F} carries every conjugacy class with 2rank(A)2\cdot\mathrm{rank}(A) forward iterates and every [ϕ][\phi]-invariant subgroup system.

Proof.

Let AFA\leq F be a free factor, ϕ:AF\phi:A\to F be injective, and L(A)1=2rank(A)1L(A)-1=2\cdot\mathrm{rank}(A)-1 be the length of the longest chain of free factor systems in AA. Any conjugacy class in a [ϕ][\phi]-invariant free factor system of AA has infinitely many forward iterates. So if only the trivial conjugacy class has L(A)L(A) forward iterates, then the trivial system is the unique [ϕ][\phi]-invariant free factor system. Thus, we may assume some fixed nontrivial conjugacy class [g][g] has L(A)L(A) forward iterates. Since AFA\leq F is a free factor and ϕ:AF\phi:A\to F is injective, we can iterate ϕ1\phi^{-1} on the poset of free factor systems of AA!

Claim (Descent).

For some n0n\geq 0, let 𝒜n𝒜0\mathcal{A}_{n}\prec\cdots\prec\mathcal{A}_{0} be the chain of nontrivial free factor systems of AA such that 𝒜i=ϕi{A}\mathcal{A}_{i}=\phi^{-i}\cdot\{A\} and 𝒜n\mathcal{A}_{n} carries all [ϕ][\phi]-invariant subgroup systems of AA. If 𝒜n\mathcal{A}_{n} is not [ϕ][\phi]-invariant, then 𝒜n+1=ϕ1𝒜n𝒜n\mathcal{A}_{n+1}=\phi^{-1}\cdot\mathcal{A}_{n}\prec\mathcal{A}_{n} is a nontrivial free factor system of AA that carries all [ϕ][\phi]-invariant subgroup systems.

Starting with {A}\{A\}, the descent will eventually stop at some n<L(A)1n<L(A)-1 with a nontrivial free factor system 𝒜n=ϕn{A}\mathcal{A}_{n}=\phi^{-n}\cdot\{A\} that carries all conjugacy classes with n+1n+1 forward iterates and all [ϕ][\phi]-invariant subgroup systems of AA. Since descent stopped, 𝒜n\mathcal{A}_{n} must be [ϕ][\phi]-invariant and hence the unique maximal system amongst all [ϕ][\phi]-invariant free factor systems of AA. ∎

Proof of descent.

Let n0n\geq 0 and 𝒜nmissingA0={A}\mathcal{A}_{n}\prec\cdots\prec\mathcal{\mathcal{missing}}A_{0}=\{A\} be the chain of nontrivial free factor systems such that 𝒜i=ϕi{A}\mathcal{A}_{i}=\phi^{-i}\cdot\{A\} and 𝒜n\mathcal{A}_{n} carries all [ϕ][\phi]-invariant subgroup systems of AA. Nontriviality of 𝒜n\mathcal{A}_{n} implies n<L(A)1n<L(A)-1. Now suppose that 𝒜n\mathcal{A}_{n} is not [ϕ][\phi]-invariant.

The existence of a nontrivial conjugacy class [g][g] with L(A)n+2L(A)\geq n+2 forward iterates implies 𝒜n+1=ϕ1𝒜n\mathcal{A}_{n+1}=\phi^{-1}\cdot\mathcal{A}_{n} is a nontrivial free factor system of AA. If n=0n=0, then clearly 𝒜n+1𝒜n\mathcal{A}_{n+1}\preceq\mathcal{A}_{n}. If n>1n>1, then 𝒜n𝒜n1\mathcal{A}_{n}\prec\mathcal{A}_{n-1} by assumption; hence, ϕ1𝒜nϕ1𝒜n1\phi^{-1}\cdot\mathcal{A}_{n}\preceq\phi^{-1}\cdot\mathcal{A}_{n-1} and, equivalently, 𝒜n+1𝒜n\mathcal{A}_{n+1}\preceq\mathcal{A}_{n}. If 𝒜n=𝒜n+1\mathcal{A}_{n}=\mathcal{A}_{n+1}, then ϕ(𝒜n)\phi(\mathcal{A}_{n}) is carried by 𝒜n\mathcal{A}_{n}. But 𝒜n\mathcal{A}_{n} is not [ϕ][\phi]-invariant, thus 𝒜n+1𝒜n\mathcal{A}_{n+1}\prec\mathcal{A}_{n}. Let \mathcal{B} be any [ϕ][\phi]-invariant subgroup system in AA. By assumption, 𝒜n\mathcal{A}_{n} carries \mathcal{B} and ϕ()\phi(\mathcal{B}); therefore, 𝒜n+1=ϕ1𝒜n\mathcal{A}_{n+1}=\phi^{-1}\cdot\mathcal{A}_{n} carries \mathcal{B}. ∎

The unique [ϕ][\phi]-invariant free factor system of AA given by Proposition 7.1, denoted by \mathcal{F}, is the canonical [ϕ][\phi]-invariant free factor system and it captures the long-term dynamics of [ϕ][\phi]. The system \mathcal{F} is proper (in AA) if and only if AA is not [ϕ][\phi]-invariant; it is trivial exactly when only the trivial conjugacy class has L(A)=2rank(A)L(A)=2\cdot\mathrm{rank}(A) forward iterates. The canonical invariant free factor system allows us to naturally extend definitions and results that required iteration of an injective endomorphism. For instance, we can now say ϕ:AF\phi:A\to F has an invariant cyclic subgroup system with index d1\boldsymbol{d}\geq 1 if a restriction ϕ|:\left.\phi\right|_{\mathcal{F}}:\mathcal{F}\to\mathcal{F} has an invariant cyclic subgroup system with index d1d\geq 1. In particular, [ϕ][\phi] is atoroidal if [ϕ|]\left[\left.\phi\right|_{\mathcal{F}}\right] is atoroidal.

For kL(A)k\geq L(A), we define the iterated pullbacks 𝚲k[ϕ]\boldsymbol{\Lambda_{k}[\phi]} of [ϕ]\boldsymbol{[\phi]} to be the iterated pullbacks Λk[ϕ|]\Lambda_{k}[\left.\phi\right|_{\mathcal{F}}] of the restriction [ϕ|][\left.\phi\right|_{\mathcal{F}}]. Similarly define Λ^k\hat{\Lambda}_{k} for kL(A)k\geq L(A). Since connectedness did not play any role in the proof of Proposition 5.5, we immediately get the following extension when we replace endomorphisms of FF with endomorphisms of \mathcal{F}.

Proposition 7.2.

If AFA\leq F is a free factor, ϕ:AF\phi:A\to F is injective, and Λ^k[ϕ]\hat{\Lambda}_{k}[\phi] is not empty for all kL(A)k\geq L(A), then ϕ\phi has an invariant cyclic subgroup system with index d2d\geq 2.

Remark.

By the same token, the results in this section do not need AA to be “connected.” So there is a natural generalization of these results that easily follows if we replace the free factor AFA\leq F with a free factor system 𝒜\mathcal{A} of FF.

For the rest of the section, we will extend the results of Section 6 to the injective homomorphism ϕ:AF\phi:A\to F. Fix a marked graph (Γ,α)(\Gamma,\alpha) such that the free factor systems \mathcal{F} and AA correspond to nested core subgraphs ΓΓAΓ\Gamma_{\mathcal{F}}\subset\Gamma_{A}\subset\Gamma respectively, i.e., Γ\Gamma_{\mathcal{F}} and ΓA\Gamma_{A} have markings α:π1(Γ)\alpha_{\mathcal{F}}:\mathcal{F}\to\pi_{1}(\Gamma_{\mathcal{F}}) and αA:Aπ1(ΓA)\alpha_{A}:A\to\pi_{1}(\Gamma_{A}) respectively such that the inclusion maps c:ΓΓAc_{\mathcal{F}}:\Gamma_{\mathcal{F}}\to\Gamma_{A} and cA:ΓAΓc_{A}:\Gamma_{A}\to\Gamma satisfy [π1(c)α]=[αA|][\pi_{1}(c_{\mathcal{F}})\circ\alpha_{\mathcal{F}}]=[\left.\alpha_{A}\right|_{\mathcal{F}}] and [π1(cA)αA]=[α|A][\pi_{1}(c_{A})\circ\alpha_{A}]=[\left.\alpha\right|_{A}]. Assume \mathcal{F} is not trivial so that Γ\Gamma_{\mathcal{F}} is not degenerate (finite set of points). A topological representative for [ϕ]\boldsymbol{[\phi]} will be a graph map f:(ΓA,Γ)(Γ,Γ)f:(\Gamma_{A},\Gamma_{\mathcal{F}})\to(\Gamma,\Gamma_{\mathcal{F}}) with no pretrivial edges such that [π1(f)αA]=[αϕ][\pi_{1}(f)\circ\alpha_{A}]=[\alpha\circ\phi]. Thus, the invariant restriction f|:ΓΓ\left.f\right|_{\mathcal{F}}:\Gamma_{\mathcal{F}}\to\Gamma_{\mathcal{F}} is a topological representative for the restriction [ϕ|][\left.\phi\right|_{\mathcal{F}}].

The following is the analogue of the mapping torus. Let ff be a topological representative for ϕ:AF\phi:A\to F. We define the classifying space to be Mf=(Γ(ΓA×[0,1]))/M_{f}=\left(\Gamma\sqcup(\Gamma_{A}\times[0,1])\right)/\sim where we identify (x,0)x(x,0)\sim x and (x,1)f(x)(x,1)\sim f(x) for all xΓAΓx\in\Gamma_{A}\subset\Gamma. The edge-space of MfM_{f} will be the cross-section represented by ΓA×{12}\Gamma_{A}\times\{\frac{1}{2}\}. By construction, π1(Mf)FA\pi_{1}(M_{f})\cong F*_{A}. Topological annuli in MfM_{f} and algebraic annuli in FAF*_{A} are defined exactly as before and the correspondence between them is the same. Similarly, the correspondence between strictly bidirectional annuli and the iterated pullbacks remains. We then get this natural extension of Lemma 6.3.

Lemma 7.3.

Let AFA\leq F be a free factor and ϕ:AF\phi:A\to F be injective. For any integer LL(A)L\geq L(A), the HNN extension FAF*_{A} has a strictly bidirectional annulus of length 2L2L if and only if Λ^L[ϕ]\hat{\Lambda}_{L}[\phi] is not empty.

The definition of (based-) hyperbolicity extends naturally using the invariant restriction. For λ>1\lambda>1 and nL(A)n\geq L(A), a topological representative f:(ΓA,Γ)(Γ,Γ)f:(\Gamma_{A},\Gamma_{\mathcal{F}})\to(\Gamma,\Gamma_{\mathcal{F}}) for [ϕ][\phi] is (based-) (λ,n)\boldsymbol{(\lambda,n)}-hyperbolic if f|:ΓΓ\left.f\right|_{\mathcal{F}}:\Gamma_{\mathcal{F}}\to\Gamma_{\mathcal{F}} is (based-) (λ,n)(\lambda,n)-hyperbolic and [f][f] is atoroidal if [f|][\left.f\right|_{\mathcal{F}}] is atoroidal. When \mathcal{F} is trivial, any topological representative of ϕ\phi is vacuously (based-) hyperbolic. Again, since connectedness played no role in the proofs of Lemma 6.5 and Proposition 6.6, we get this extension:

Proposition 7.4.

Let AFA\leq F be a free factor. If ϕ:AF\phi:A\to F is injective and atoroidal, then [ϕ][\phi] has a (2,n)(2,n)-hyperbolic topological representative for some integer nL(A)n\geq L(A).

We are now ready to state and almost prove the main result of the section.

Theorem 7.5.

Let AFA\leq F be a free factor and ϕ:AF\phi:A\to F be an injective homomorphism. Then the following are equivalent:

  1. 1.

    FAF*_{A} is word-hyperbolic;

  2. 2.

    FAF*_{A} contains no BS(1,d)BS(1,d) subgroups with d1d\geq 1;

  3. 3.

    [ϕ][\phi] has no invariant cyclic subgroup system with index d1d\geq 1;

  4. 4.

    all/some topological representative of [ϕ][\phi] are based-hyperbolic and all strictly bidirectional annuli in their classifying spaces are shorter than some integer.

The proof follows the same steps as Theorem 6.7 and the only missing ingredient is an extension of Theorem 6.1 (Theorem 7.7 below) that proves the implication (4)(1)(4)\implies(1).

Remark.

Every HNN extension of the form FAF*_{A} where FF is a finite rank free group and AFA\leq F is a free factor can be written as a mapping torus 𝔽ψ\mathbb{F}*_{\psi} of an injective endomorphism ψ:𝔽𝔽\psi:\mathbb{F}\to\mathbb{F} where 𝔽\mathbb{F} is a free group with possibly infinite rank. Conversely, every finitely generated subgroup of a mapping torus 𝔽ψ\mathbb{F}*_{\psi} where 𝔽\mathbb{F} has possibly infinite rank can be written as an HNN extension FAF*_{A} where FF has finite rank and AFA\leq F is a free factor [9]. This argument is Feighn-Handel’s result that FAF*_{A} and 𝔽ψ\mathbb{F}*_{\psi} are coherent. As a corollary, we get the following amusing statement:

Corollary 7.6.

Let ψ:𝔽𝔽\psi:\mathbb{F}\to\mathbb{F} be an injective endomorphism of an infinite rank free group. 𝔽ψ\mathbb{F}*_{\psi} is locally word-hyperbolic if and only if it contains no BS(1,d)BS(1,d) subgroups with d1d\geq 1.

By locally word-hyperbolic, we mean every finitely generated subgroup is word-hyperbolic.

The following theorem will complete our proof of Theorem 7.5.

Theorem 7.7.

Let AFA\leq F be a free factor and ϕ:AF\phi:A\to F be injective. If a topological representative ff of [ϕ][\phi] is based-hyperbolic and all strictly bidirectional annuli in its classifying space MfM_{f} are shorter than some integer, then FAF*_{A} is word-hyperbolic.

Roughly speaking, the extension follows from the fact that annuli of MfM_{f} longer than L(A)L(A) are annuli of the mapping torus of f|\left.f\right|_{\mathcal{F}} up to some controlled error. Due to the close similarity with the proof of Theorem 6.1 given in [23], we will only sketch a proof of the extension.

We start by defining annuli flaring; let h:S1×[M,M]Mfh:S^{1}\times[-M,M]\to M_{f} be a topological annulus in MfM_{f} of length 2M2L(A)2M\geq 2L(A). The girth of h\boldsymbol{h} is the combinatorial length |h0||h_{0}| of the middle ring h0=h(,0):S1Mfh_{0}=h(\cdot,0):S^{1}\to M_{f} in the edge-space. Let λ>1\lambda>1 be a real number. The annulus hh is 𝝀\boldsymbol{\lambda}-hyperbolic if

λ|h0|max(|hL|,|hL|).\lambda|h_{0}|\leq\max(|h_{-L}|,|h_{L}|).

For integers ii between M-M and M1M-1 (inclusive), define τi:(i,i+1)Γ\tau_{i}:(i,i+1)\to\Gamma by projecting the trace hν|(i,i+1)\left.h^{\nu}\right|_{(i,i+1)} to the vertex-space Γ\Gamma as follows:

τi(t)={xif hν(t)=(x,s)ΓA×[0,1] and s<12f(x)if hν(t)=(x,s)ΓA×[0,1] and s>12xif hν(t)=xΓ\tau_{i}(t)=\begin{cases}x&\text{if }h^{\nu}(t)=(x,s)\in\Gamma_{A}\times[0,1]\text{ and }s<\frac{1}{2}\\ f(x)&\text{if }h^{\nu}(t)=(x,s)\in\Gamma_{A}\times[0,1]\text{ and }s>\frac{1}{2}\\ x&\text{if }h^{\nu}(t)=x\in\Gamma\\ \end{cases}

We say hh is 𝝆\boldsymbol{\rho}-thin if |τi|+1ρ|\tau_{i}|+1\leq\rho for all ii, where |||\cdot| is the length after tightening the path rel. endpoints. We say MfM_{f} satisfies the annuli flaring condition if there are λ>1,M2L(A)\lambda>1,M\geq 2L(A), and a function H:H:\mathbb{R}\to\mathbb{R} such that any ρ\rho-thin annulus of length 2M2M with girth at least H(ρ)H(\rho) is λ\lambda-hyperbolic. Bestvina-Feighn’s combination theorem [3] states that π1(Mf)\pi_{1}(M_{f}) is word-hyperbolic if MfM_{f} satisfies the annuli flaring condition.

Sketch proof of Theorem 7.7.

Suppose AFA\leq F is a free factor, ϕ:AF\phi:A\to F is injective, and \mathcal{F} is the canonical [ϕ][\phi]-invariant free factor system of AA. Let f:(ΓA,Γ)(Γ,Γ)f:(\Gamma_{A},\Gamma_{\mathcal{F}})\to(\Gamma,\Gamma_{\mathcal{F}}) be a based-hyperbolic topological representative for [ϕ][\phi] and K=K(f)>1K=K(f)>1 be some Lipschitz constant for ff. If \mathcal{F} is trivial, then all unidirectional annuli in MfM_{f} are shorter than L(A)L(A). So MfM_{f} vacuously satisfies the annuli flaring condition for some M=2L(A)M=2L(A) by the annulus dichotomy (malnormality of AA). Thus, we may assume \mathcal{F} is not trivial. The assumptions and tool introduced in the next two paragraphs are needed in case AA is not [ϕ][\phi]-invariant.

Let 𝒜n𝒜0𝒜1\mathcal{A}_{n}\prec\cdots\prec\mathcal{A}_{0}\prec\mathcal{A}_{-1} be the chain where 𝒜1={F},𝒜0={A},𝒜n=,\mathcal{A}_{-1}=\{F\},\mathcal{A}_{0}=\{A\},\mathcal{A}_{n}=\mathcal{F}, and generally 𝒜i=ϕi{A}\mathcal{A}_{i}=\phi^{-i}\cdot\{A\} for 0in0\leq i\leq n. By invariance of 𝒜n\mathcal{A}_{n}, we can extend the chain by setting 𝒜i=𝒜n\mathcal{A}_{i}=\mathcal{A}_{n} for i>ni>n. For 0i<n0\leq i<n, 𝒜i1\mathcal{A}_{i-1} carries ϕ(𝒜i)\phi(\mathcal{A}_{i}) by definition. We assume Γ\Gamma has a filtration of nonempty core subgraphs Γ1Γ0Γ1\cdots\subset\Gamma_{1}\subset\Gamma_{0}\subset\Gamma_{-1} where Γ1=ΓR\Gamma_{-1}=\Gamma_{\mathcal{F}}\vee R is a wedge between Γ\Gamma_{\mathcal{F}} and a rose RR and each Γi\Gamma_{i} is a 𝒜i\mathcal{A}_{i}-marked graph. If n>0n>0, we will also assume the topological representative f:(Γ0,Γn)(Γ1,Γn)f:(\Gamma_{0},\Gamma_{n})\to(\Gamma_{-1},\Gamma_{n}) is of the form (Γ0,,Γn1,Γn)(Γ1,,Γn2,Γn)(\Gamma_{0},\ldots,\Gamma_{n-1},\Gamma_{n})\to(\Gamma_{-1},\ldots,\Gamma_{n-2},\Gamma_{n}). Note that since Γn=Γ\Gamma_{n}=\Gamma_{\mathcal{F}} was ff-invariant to begin with, changes can made to the given representative ff to allow for these additional assumptions without affecting the invariant restriction f|\left.f\right|_{\mathcal{F}}. So ff is still based-hyperbolic after the changes.

Using this filtration, we introduce a tool needed to iterate based loops not in Γ\Gamma_{\mathcal{F}}: for any integer i0i\geq 0 and any immersed based loop s:S1Γ0s:S^{1}\to\Gamma_{0} that is freely homotopic into Γ\Gamma_{\mathcal{F}}, there is a “closest point” projection (with respect to Hausdorff distance of lifts to the universal cover) of ss to an immersed based loop in Γ0\Gamma_{0} denoted by si:S1Γ0\lfloor s\rfloor_{i}:S^{1}\to\Gamma_{0} freely homotopic to ss such that fi(si)f^{i}(\lfloor s\rfloor_{i}) is a based loop in Γ0\Gamma_{0}. Recall that based loops must send basepoints to vertices. Since we assumed the “complement” of Γn\Gamma_{n} are roses, the projections si\lfloor s\rfloor_{i} are immersed based loops in Γi\Gamma_{i}. So si=sn\lfloor s\rfloor_{i}=\lfloor s\rfloor_{n} for ini\geq n. Implicit in this is the fact that, for all i0i\geq 0, any conjugacy class carried by 𝒜i\mathcal{A}_{i} but not 𝒜i+1\mathcal{A}_{i+1} is mapped by ϕ\phi to a conjugacy class carried by 𝒜i1\mathcal{A}_{i-1} but not 𝒜i\mathcal{A}_{i}. We are ready to start proving the theorem.

Suppose ff is based-(2,m)(2,m)-hyperbolic for some m1m\geq 1 and strictly bidirectional annuli in MfM_{f} are shorter than 2L2L(A)2L\geq 2L(A). Fix r1r\geq 1 such that 2L+r3KLm2^{L+r}\geq 3\cdot K^{Lm}. Now set M=(2L+r)mM=(2L+r)m and

H(ρ)=2M4ρK2M1K1.H(\rho)=2^{M}\cdot 4\rho\cdot\frac{K^{2M}-1}{K-1}.

Suppose hh is an arbitrary ρ\rho-thin annulus of length 2M2M with girth |h0|H(ρ)|h_{0}|\geq H(\rho). We need to show that hh is 22-hyperbolic. Without loss of generality, assume the truncation h|[M,0]\left.h\right|_{[-M,0]} is unidirectional and increasing.

For Mk0-M\leq k\leq 0, consider the rings of hk:S1Mfh_{k}:S^{1}\to M_{f} in the edge-space ΓA×{12}\Gamma_{A}\times\{\frac{1}{2}\}. This can be viewed as an immersed based loop hk:S1Γ0h_{k}:S^{1}\to\Gamma_{0} by forgetting the {12}\{\frac{1}{2}\}-factor. Since ML(A)M\geq L(A) and hMMfh0h_{-M}\simeq_{M_{f}}h_{0} (free homotopy in MfM_{f}) is not trivial, the ring hMh_{-M} is freely homotopic (in Γ\Gamma) into Γn=Γ\Gamma_{n}=\Gamma_{\mathcal{F}}. Since Γn\Gamma_{n} is ff-invariant and hMMfhkh_{-M}\simeq_{M_{f}}h_{k} for all k0k\leq 0, the rings hkh_{k} are freely homotopic into Γn\Gamma_{n}. In particular, these rings all have projections to subgraphs Γi\Gamma_{i} for all ii.

For the first step, project h1h_{-1} to Γ1\Gamma_{1} so that the image f(h11):S1Γ0f(\lfloor h_{-1}\rfloor_{1}):S^{1}\to\Gamma_{0} is a based loop freely homotopic to h0h_{0} in Γ\Gamma. The ρ\rho-thin assumption (and “closest point” projection) means: |f(h11)||h0|2ρ|f(\lfloor h_{-1}\rfloor_{1})|\geq|h_{0}|-2\rho. Similarly, we project h2h_{-2} to Γ2\Gamma_{2} so that f(h22):S1Γ1f(\lfloor h_{-2}\rfloor_{2}):S^{1}\to\Gamma_{1} is a based loop freely homotopic to h11\lfloor h_{1}\rfloor_{1} in Γ\Gamma. Then |f(h22)||h11|2ρ|f(\lfloor h_{-2}\rfloor_{2})|\geq|\lfloor h_{1}\rfloor_{1}|-2\rho by ρ\rho-thinness and, by the KK-Lipschitz property of ff,

|f2(h22)||f(h11)|2ρK|h0|2ρ(1+K).|f^{2}(\lfloor h_{-2}\rfloor_{2})|\geq|f(\lfloor h_{1}\rfloor_{1})|-2\rho\cdot K\geq|h_{0}|-2\rho\cdot(1+K).

By induction, the ff-invariance of Γn\Gamma_{n}, and the fact M>L(A)>nM>L(A)>n, we get this inequality:

|fM(hMn)||h0|2ρ(1++KM1)=|h0|2ρKM1K1.|f^{M}(\lfloor h_{-M}\rfloor_{n})|\geq|h_{0}|-2\rho\cdot(1+\cdots+K^{M-1})=|h_{0}|-2\rho\cdot\frac{K^{M}-1}{K-1}.

Since ff is based-(2,m)(2,m)-hyperbolic, we know that

2|fM(hMn)|max(|fM+m(hMn)|,|fMm(hMn)|).2|f^{M}(\lfloor h_{-M}\rfloor_{n})|\leq\max(|f^{M+m}(\lfloor h_{-M}\rfloor_{n})|,|f^{M-m}(\lfloor h_{-M}\rfloor_{n})|).

There are three cases to consider of which we will only prove one.

Case 1: Suppose 2|fM(hMn)||fMm(hMn)|2|f^{M}(\lfloor h_{-M}\rfloor_{n})|\leq|f^{M-m}(\lfloor h_{-M}\rfloor_{n})|.

Then by induction on based-(2,m)(2,m)-hyperbolicity, we get:

|hM||hMn|\displaystyle|h_{-M}|\geq|\lfloor h_{-M}\rfloor_{n}| 22L+r|fM(hMn)|\displaystyle\geq 2^{2L+r}|f^{M}(\lfloor h_{-M}\rfloor_{n})|
22L+r|h0|22L+r2ρKM1K1\displaystyle\geq 2^{2L+r}|h_{0}|-2^{2L+r}\cdot 2\rho\cdot\frac{K^{M}-1}{K-1}
3|h0|H(ρ)\displaystyle\geq 3|h_{0}|-H(\rho)
2|h0|\displaystyle\geq 2|h_{0}| as |h0|H(ρ).\displaystyle\text{as }|h_{0}|\geq H(\rho).

So hh is 22-hyperbolic in this case. The proof so far has been nearly identical to the proof of Theorem 6.1 in [23]. But here, we need to be careful when applying iterates of ff to the rings of hh since AA may not be [ϕ][\phi]-invariant. This is why we introduced the projections i\lfloor\cdot\rfloor_{i}.

There are two more cases remaining.

Case 2: Suppose 2|fM(hMn)||fM+m(hMn)|2|f^{M}(\lfloor h_{-M}\rfloor_{n})|\leq|f^{M+m}(\lfloor h_{-M}\rfloor_{n})| and hh is unidirectional.

Case 3: Suppose 2|fM(hMn)||fM+m(hMn)|2|f^{M}(\lfloor h_{-M}\rfloor_{n})|\leq|f^{M+m}(\lfloor h_{-M}\rfloor_{n})| and hh is bidirectional.

Case 3 is where the bound on strictly bidirectional annuli is needed. We leave the details of these cases to the reader. Alternatively, one could read the proof of Theorem 6.1, compare how Case 1 was handled in the two proofs, and adjust the proofs of the remaining cases accordingly. We have covered all the cases and MfM_{f} satisfies the annuli flaring condition. By the combination theorem, FAπ1(Mf)F*_{A}\cong\pi_{1}(M_{f}) is word-hyperbolic. ∎

Epilogue

We would like to conclude this paper with a discussion of a few questions: the first question is a natural generalization of Section 7; the rest are problems from Ilya Kapovich’s paper [17, Section 6] that could be answered using expanding relative immersions.

Problem 1.

Suppose AFA\leq F is a vertex group of a cyclic splitting of FF and ϕ:AF\phi:A\to F is injective. Is FAF*_{A} word-hyperbolic if it contains no BS(1,d)BS(1,d) subgroups for d1d\geq 1? Can it have BS(m,d)BS(m,d) subgroups with m,d>1m,d>1?

A cyclic splitting of FF is an edge of groups decomposition of FF with a nontrivial cyclic edge group. Vertex groups of cyclic splittings are generalizations of free factors and it would be interesting to see if the ideas in Section 7 can be adapted to this case. The first obstacle is finding the appropriate generalization of Proposition 7.1 since ϕ1\phi^{-1}-iteration need not be as well-behaved. Furthermore, the vertex groups AA are not always malnormal which means annuli in FAF*_{A} could exhibit more complicated behavior.

For the remaining problems, assume ϕ:FF\phi:F\to F is injective.

Problem 2 ([17, Problem 6.4]).

What kind of isoperimetric functions can FϕF*_{\phi} have?

The automorphism case of this problem was answered by Bridson-Groves [7]: they showed that free-by-cyclic groups have quadratic isoperimetric functions. Implicit in the second part of the paper is the idea that FϕF*_{\phi} is hyperbolic relative to a canonical finite collection of free-by-cyclic groups when it has no BS(1,d)BS(1,d) subgroups for d2d\geq 2. Furthermore, when AFA\leq F is a free factor and ψ:AF\psi:A\to F is injective, then FAF*_{A} is hyperbolic relative to a canonical finite collection of ascending HNN extensions. This would imply that FAF*_{A} has a quadratic isoperimetric function when it has no BS(1,d)BS(1,d) subgroups for d2d\geq 2. We intend to complete this direction in future work by employing a combination theorem for relatively hyperbolic groups [12, 22].

When FϕF*_{\phi} is word-hyperbolic, Mahan Mj proved there is a continuous extension to the Gromov boundary of the inclusion map FFϕF\leq F*_{\phi} [21]; this map is known as the Cannon-Thurston map.

Problem 3 ([17, Problem 6.7]).

If FϕF*_{\phi} is word-hyperbolic, is the Cannon-Thurston map (uniformly) finite-to-one? Is there a corresponding ending laminations theorem?

For the first of these questions, it seems that expanding relative immersions reduce the problem to the automorphism case, which has been answered [13, 20]. As for the ending laminations, what is missing is the appropriate formulation of the theorem that replaces the usual short exact sequence with a graph of groups decomposition.

Problem 4 ([17, Problem 6.8]).

Let ϕ:FF\partial\phi:\partial F\to\partial F be the extension of ϕ\phi to the boundary. Can we classify injective endomorphisms φ\varphi by the dynamics of boundary extensions φ\partial\varphi?

Levitt-Lustig covered the surjective case when they proved that most automorphisms of non-elementary word-hyperbolic groups have boundary extensions with north-south dynamics [18]: two fixed points — a repellor and an attractor. Expanding relative immersions should imply that most injective nonsurjective endomorphisms of FF have sink dynamics: a single fixed point and it is an attractor.

There are of course more unresolved questions about injective endomorphisms and their expanding relative immersions. These are left for the reader to ask and answer.

Appendix A Relative train tracks

The objective in this appendix is to sketch the proof that irreducible relative representatives with minimal stretch factor are train tracks. Bestvina-Handel’s construction of train tracks for irreducible automorphisms [6] translates verbatim to the non-free forest setting.

Let ϕ:FF\phi:F\to F be an injective endomorphism, 𝒜\mathcal{A}\prec\mathcal{B} be [ϕ][\phi]-invariant free factor systems, and TT_{*} be a (,𝒜)(\mathcal{B},\mathcal{A})-forest. We allow forests to have bivalent vertices. Recall, an 𝒜\mathcal{A}-relative weak representative for the restriction ϕ|\left.\phi\right|_{\mathcal{B}} is a ϕ|\left.\phi\right|_{\mathcal{B}}-equivariant graph map f:TTf_{*}:T_{*}\to T_{*}. An 𝒜\mathcal{A}-relative representative is an 𝒜\mathcal{A}-relative weak representative ff_{*} with no pretrivial edges and whose underlying forest TT_{*} has no bivalent vertices. Additionally, we say the relative representative is minimal if it has no orbit-closed invariant subforests with bounded components.

For any 𝒜\mathcal{A}-relative weak representative ff_{*}, we get the transition matrix A(f)A(f_{*}). An 𝒜\mathcal{A}-relative representative ff_{*} is irreducible if the matrix A(f)A(f_{*}) is irreducible and, in this case, the stretch factor of ff_{*}, denoted by λ(f)\lambda(f_{*}), is the Perron-Frobenius eigenvalue of A(f)A(f_{*}). An 𝓐\boldsymbol{\mathcal{A}}-relative train track for ϕ|\left.\phi\right|_{\mathcal{B}} is an 𝒜\mathcal{A}-relative representative ff_{*} for ϕ|\left.\phi\right|_{\mathcal{B}} that additionally satisfies the property: the edge-paths fn(e)f_{*}^{n}(e) are immersed for all edges ee in TT_{*} and integers n1n\geq 1. We have set the stage for the Bestvina-Handel’s result.

Theorem A.1 ([6, Theorem 1.7]).

Let ϕ|\left.\phi\right|_{\mathcal{B}} be irreducible relative to 𝒜\mathcal{A} and f:TTf_{*}:T_{*}\to T_{*} be an irreducible 𝒜\mathcal{A}-relative representative for ϕ|\left.\phi\right|_{\mathcal{B}}. If ff_{*} has minimal stretch factor, then it is an irreducible 𝒜\mathcal{A}-relative train track.

Minimality is understood to be amongst irreducible representatives rather than weak representatives. The argument relies on understanding how various moves on an irreducible 𝒜\mathcal{A}-relative representative ff_{*} affect λ(f)\lambda(f_{*}) and invoking minimality of λ(f)\lambda(f_{*}) to conclude that no such moves are possible. Note that although the moves are described locally, they must be performed equivariantly if we want the resultant forests to be (,𝒜)(\mathcal{B},\mathcal{A})-forests. The proofs of these moves/lemmas are omitted since they are the same as the proofs in [6].

Remark.

Recently, Bestvina [2] and Francaviglia-Martino [10] gave an alternative approach to proving this theorem using the Lipschitz metric on relative outer space.

The first move is subdivision, which occurs at an interior point of an edge that is in the preimage of vertices under the representative.

Lemma A.2 ([6, Lemma 1.10]).

If f:TTf_{*}:T_{*}\to T_{*} is an irreducible 𝒜\mathcal{A}-relative weak representative for ϕ|\left.\phi\right|_{\mathcal{B}} and f:TTf_{*}^{\prime}:T_{*}^{\prime}\to T_{*}^{\prime} is induced by a subdivision, then ff_{*}^{\prime} is an irreducible 𝒜\mathcal{A}-relative weak representative and λ(f)=λ(f)\lambda(f_{*}^{\prime})=\lambda(f_{*}).

The next move is bivalent homotopy, which occurs at a bivalent vertex and decreases the number of edges.

Lemma A.3 ([6, Lemma 1.13]).

If f:TTf_{*}:T_{*}\to T_{*} is an irreducible 𝒜\mathcal{A}-relative weak representative for ϕ|\left.\phi\right|_{\mathcal{B}} and f′′:T′′T′′f_{*}^{\prime\prime}:T_{*}^{\prime\prime}\to T_{*}^{\prime\prime} is an irreducible 𝒜\mathcal{A}-relative weak representative induced by a bivalent homotopy followed by collapse of a maximal invariant subforest with bounded components, then λ(f′′)λ(f)\lambda(f_{*}^{\prime\prime})\leq\lambda(f_{*}).

The last move we need is folding, which occurs between a pair of oriented edges originating from the same vertex that have the same image under the representative.

Lemma A.4 ([6, Lemma 1.15]).

Suppose f:TTf_{*}:T_{*}\to T_{*} is an irreducible 𝒜\mathcal{A}-relative weak representative for ϕ|\left.\phi\right|_{\mathcal{B}} and f:TTf_{*}^{\prime}:T_{*}^{\prime}\to T_{*}^{\prime} is induced by a fold. If ff_{*}^{\prime} is an 𝒜\mathcal{A}-relative weak representative, then it is irreducible and λ(f)=λ(f)\lambda(f_{*}^{\prime})=\lambda(f_{*}). Otherwise, if f′′:T′′T′′f_{*}^{\prime\prime}:T_{*}^{\prime\prime}\to T_{*}^{\prime\prime} is an irreducible 𝒜\mathcal{A}-relative weak representative induced by a homotopy of ff_{*}^{\prime} that makes the final map locally injective on the interior of edges, followed by collapse of a maximal invariant subforest with bounded components, then λ(f′′)<λ(f)\lambda(f_{*}^{\prime\prime})<\lambda(f_{*}).

Sketch proof of Theorem A.1.

Let ϕ|\left.\phi\right|_{\mathcal{B}} be irreducible relative to 𝒜\mathcal{A} and f:TTf_{*}:T_{*}\to T_{*} be an irreducible 𝒜\mathcal{A}-relative representative. If λ(f)=1\lambda(f_{*})=1, then ff_{*} is a simplicial embedding with minimal stretch factor and we are done. So we may assume λ(f)>1\lambda(f_{*})>1.

Suppose for the contrapositive that ff_{*} is not an 𝒜\mathcal{A}-relative train track, then the edge-path fn(e)f_{*}^{n}(e) is not immersed for some edge ee in TT_{*} and integer n1n\geq 1. Let nn be the smallest such integer and assume \star is an interior point of an edge ee at which fnf_{*}^{n} fails to be locally injective. We appropriately subdivide TT_{*} so that a neighborhood UU of \star and its iterates fk(U)f_{*}^{k}(U) (1kn)(1\leq k\leq n) satisfy nice properties: 1) UU is an interval whose boundary consists of distinct vertices; 2) fkf_{*}^{k} is locally injective on UU for 1k<n1\leq k<n; 3) fnf_{*}^{n} folds UU at \star to an edge; and 4) fk(U)\star\notin f_{*}^{k}(U) for 1kn1\leq k\leq n. We can then iteratively fold fn1(U)f_{*}^{n-1}(U), …, f2(U)f_{*}^{2}(U), and f(U)f_{*}(U). By minimality of nn, all the folds except the last one induce an irreducible 𝒜\mathcal{A}-relative weak representative. By the first case of Lemma A.4, this irreducible 𝒜\mathcal{A}-relative weak representative has the same stretch factor as ff_{*}. By construction, the last fold induces a map ff_{*}^{\prime} that fails to be an 𝒜\mathcal{A}-relative weak representative as it fails to be locally injective at \star. We can apply a tightening homotopy on ff_{*}^{\prime} to make it locally injective at \star, then collapse a maximal invariant subforest with bounded components to get f′′:T′′T′′f_{*}^{\prime\prime}:T_{*}^{\prime\prime}\to T_{*}^{\prime\prime}, a minimal 𝒜\mathcal{A}-relative weak representative for ϕ|\left.\phi\right|_{\mathcal{B}}. By Lemma 3.6, the map f′′f_{*}^{\prime\prime} is irreducible. By the second case of Lemma A.4, the stretch factor is strictly smaller: λ(f′′)<λ(f)\lambda(f_{*}^{\prime\prime})<\lambda(f_{*}). We then sequentially apply bivalent homotopies and collapse maximal invariant subforests with bounded components until we get an irreducible 𝒜\mathcal{A}-relative representative f′′′f_{*}^{\prime\prime\prime}. The stretch factor satisfies λ(f′′′)λ(f′′)<λ(f)\lambda(f_{*}^{\prime\prime\prime})\leq\lambda(f_{*}^{\prime\prime})<\lambda(f_{*}) by Lemma A.3. So ff_{*} did not have minimal stretch factor. ∎

References

  • [1] Mladen Bestvina. Questions in geometric group theory. preprint, 2004.
  • [2] Mladen Bestvina. A Bers-like proof of the existence of train tracks for free group automorphisms. Fund. Math., 214(1):1–12, 2011.
  • [3] Mladen Bestvina and Mark Feighn. A combination theorem for negatively curved groups. J. Differential Geom., 35(1):85–101, 1992.
  • [4] Mladen Bestvina, Mark Feighn, and Michael Handel. Laminations, trees, and irreducible automorphisms of free groups. Geom. Funct. Anal., 7(2):215–244, 1997.
  • [5] Mladen Bestvina, Mark Feighn, and Michael Handel. The Tits alternative for Out(Fn){\rm Out}(F_{n}). I. Dynamics of exponentially-growing automorphisms. Ann. of Math. (2), 151(2):517–623, 2000.
  • [6] Mladen Bestvina and Michael Handel. Train tracks and automorphisms of free groups. Ann. of Math. (2), 135(1):1–51, 1992.
  • [7] Martin R. Bridson and Daniel Groves. The quadratic isoperimetric inequality for mapping tori of free group automorphisms. Mem. Amer. Math. Soc., 203(955):xii+152, 2010.
  • [8] Peter Brinkmann. Hyperbolic automorphisms of free groups. Geom. Funct. Anal., 10(5):1071–1089, 2000.
  • [9] Mark Feighn and Michael Handel. Mapping tori of free group automorphisms are coherent. Ann. of Math. (2), 149(3):1061–1077, 1999.
  • [10] Stefano Francaviglia and Armando Martino. Stretching factors, metrics and train tracks for free products. Illinois J. Math., 59(4):859–899, 2015.
  • [11] Joel Friedman. Sheaves on graphs, their homological invariants, and a proof of the Hanna Neumann conjecture: with an appendix by Warren Dicks. Mem. Amer. Math. Soc., 233(1100):xii+106, 2015.
  • [12] François Gautero. Geodesics in trees of hyperbolic and relatively hyperbolic spaces. Proc. Edinb. Math. Soc. (2), 59(3):701–740, 2016.
  • [13] Pritam Ghosh. Limits of conjugacy classes under iterates of hyperbolic elements of Out(𝔽\mathbb{F}). Groups Geom. Dyn., 14(1):177–211, 2020.
  • [14] Misha Gromov. Hyperbolic groups. In Essays in group theory, volume 8 of Math. Sci. Res. Inst. Publ., pages 75–263. Springer, New York, 1987.
  • [15] Mark F. Hagen and Daniel T. Wise. Cubulating hyperbolic free-by-cyclic groups: the general case. Geom. Funct. Anal., 25(1):134–179, 2015.
  • [16] Mark F. Hagen and Daniel T. Wise. Cubulating hyperbolic free-by-cyclic groups: the irreducible case. Duke Math. J., 165(9):1753–1813, 2016.
  • [17] Ilya Kapovich. Mapping tori of endomorphisms of free groups. Comm. Algebra, 28(6):2895–2917, 2000.
  • [18] Gilbert Levitt and Martin Lustig. Most automorphisms of a hyperbolic group have very simple dynamics. Ann. Sci. École Norm. Sup. (4), 33(4):507–517, 2000.
  • [19] Igor Mineyev. Submultiplicativity and the Hanna Neumann conjecture. Ann. of Math. (2), 175(1):393–414, 2012.
  • [20] Mahan Mitra. Ending laminations for hyperbolic group extensions. Geom. Funct. Anal., 7(2):379–402, 1997.
  • [21] Mahan Mitra. Cannon-Thurston maps for trees of hyperbolic metric spaces. J. Differential Geom., 48(1):135–164, 1998.
  • [22] Mahan Mj and Lawrence Reeves. A combination theorem for strong relative hyperbolicity. Geom. Topol., 12(3):1777–1798, 2008.
  • [23] Jean Pierre Mutanguha. Hyperbolic immersions of free groups. Groups Geom. Dyn., 14(4):1253–1275, 2020.
  • [24] Jean Pierre Mutanguha. Irreducible nonsurjective endomorphisms of FnF_{n} are hyperbolic. Bull. Lond. Math. Soc., 52(5):960–976, 2020.
  • [25] Hanna Neumann. On the intersection of finitely generated free groups. Addendum. Publ. Math. Debrecen, 5:128, 1957.
  • [26] Walter D. Neumann. On intersections of finitely generated subgroups of free groups. In Groups—Canberra 1989, volume 1456 of Lecture Notes in Math., pages 161–170. Springer, Berlin, 1990.
  • [27] Patrick R. Reynolds. Dynamics of irreducible endomorphisms of FNF_{N}. PhD thesis, University of Illinois at Urbana-Champaign, 2011.
  • [28] Jean-Pierre Serre. Arbres, amalgames, SL2{\rm SL}_{2}. Société Mathématique de France, Paris, 1977. Avec un sommaire anglais, Rédigé avec la collaboration de Hyman Bass, Astérisque, No. 46.
  • [29] John R. Stallings. Topology of finite graphs. Invent. Math., 71(3):551–565, 1983.
  • [30] William P. Thurston. Three-dimensional manifolds, Kleinian groups and hyperbolic geometry. Bull. Amer. Math. Soc. (N.S.), 6(3):357–381, 1982.
  • [31] William P. Thurston. Hyperbolic structures on 33-manifolds. I. Deformation of acylindrical manifolds. Ann. of Math. (2), 124(2):203–246, 1986.