The dynamics and geometry of free group endomorphisms
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 -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 be a group of finite type. If 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 for . 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 -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 be a free factor of a finite rank free group and be an injective homomorphism. The HNN extension of over and is word-hyperbolic if and only if it contains no subgroups for .
The HNN extension has the presentation . If , then 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 , then is a strictly ascending HNN extension and coarse hyperbolization for this class of groups was our original motivation. Finally, when , we more or less reduce this to the case .
The case 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 is irreducible but not surjective [23, 24]. In this case, we showed that the strictly ascending HNN extension 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 1–4) 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 is injective but not surjective, then there is:
-
1.
a unique maximal proper -fixed free factor system ; (Proposition 2.3)
-
2.
a unique minimal proper -invariant free factor system that carries and is fixed under backward iteration, i.e., ; (Proposition 2.4)
-
3.
a unique expanding -relative immersion for . (Corollary 4.7)
The free factor systems and could be trivial.
This means any injective nonsurjective endomorphism is induced by a graph map such that: 1) some possibly empty or disconnected proper subgraph is -invariant and the restriction of to is a homotopy equivalence; 2) some -invariant proper subgraph has an -image contained in for some ; and 3) collapsing the lifts of the subgraph in the universal cover induces an expanding immersion on a simplicial -tree. In particular, if is empty, then 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 and 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 5–7).
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 ; we suspect that they could be used to extend the main theorem to a characterization of the possible Dehn/isoperimetric functions of the groups 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 of torsion-free word-hyperbolic groups over free factors ; 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 is injective but not surjective. If 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 be the free group on two generators and be injective nonsurjective endomorphisms given by
The standard rose is a rose with two petals and an identification such that the basis corresponds to the petals; it will be graphically represented by a rose with two oriented petals labelled by and respectively.

For each integer , let be the cover of corresponding to the subgroup , i.e., it is the quotient of the universal cover by the action of the subgroup . The Stallings graph is the core of , i.e., the smallest connected subgraph of with rank 2. Alternatively, we can define to be the smallest subtree invariant under the -action and to be the quotient . Evidently, there is a natural isomorphism . A Stallings graph is graphically represented as an -digraph, i.e., will be an oriented graph whose oriented edges are labelled by or as shown in Figure 2.

The Stallings graphs are roses for all and each petal doubles in size with each iteration. Since is injective, it is an isomorphism onto its image. In particular, we may use to get an isomorphism for . Under this isomorphism, the basis corresponds to the petals of and we recover the standard rose for all . This is equivalent to the existence of an immersion (locally injective map) on the rose such that [24, Lemma 3.2]. Indeed, the obvious map that maps the -petal to the path and the -petal to the path is an immersion. The petals of doubling in size with each iteration implies is in fact an expanding immersion, i.e., all edges expand under -iteration.

The Stallings graphs are barbells for all , the middle bars roughly double in size with each iteration, and the plates are labelled by (See Figure 3). Under the isomorphism , the conjugacy classes correspond to the two plates of the barbell and we recover the so called standard barbell for all . Just as in the first example, this is equivalent to the existence of an immersion on the barbell such that under some identification . 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 are all labelled , we deduce that is a free factor fixed by and is mapped to a conjugate of . Thus, forms an invariant free factor system that contains a fixed free factor. Now consider to be the lift of to the universal cover . Collapsing all translates of axes of and in , i.e., collapsing all edges of labelled by the plates of , will produce a so-called -tree with a nonfree -action and induces an expanding immersion where each edge doubles in size under -iteration. The map will be referred to as a relative immersion.

For the final example, the Stallings graphs are roses again where one petal roughly doubles in size with each iteration and another is labelled by (See Figure 4). Unlike the first example, the roses with the isomorphisms are not standard roses since does not correspond to a petal for . In fact, applying to the labels of forces a full fold with one petal to get which means the Stallings graphs are all distinct marked roses. In particular, cannot be induced by a graph immersion! But the observation that the nongrowing petals of are all labelled by implies is a fixed free factor. Despite not being induced by an immersion, the obvious map on the standard rose will induce an expanding immersion on an -tree 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 such that collapsing the translates of axes in of some canonical invariant free factor system containing a fixed free factor system will induce an expanding relative immersion . 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 will produce a tree 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
will always be a free group with finite rank at least . A nontrivial subgroup system of is a nonempty finite collection of nontrivial subgroups of . A nontrivial free factor system of is a nonempty collection of nontrivial free factors of such that is trivial if and is free factor of . 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 , the subgroups are called the components of or -components. A subgroup system is finitely generated if its components are finitely generated. A free factor system is proper if some component is proper, i.e., . Let and be subgroup systems of . We shall say carries if for every component , there is a component and element such that and, when and are free factor systems, we shall denoted it by . For a conjugacy class of elements in , we shall also say carries if there is a component and element such that . If (or ) is a singleton, then we will write, “ carries (or ).” One can easily verify that the -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 be the free group of rank 2. With our definition, and are equivalent free factor systems but 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 are equivalent if there is an inner automorphism such that and this is denoted by . Outer endomorphisms of are the equivalence classes on the set of endomorphisms of . For instance, properties such as being irreducible are not just properties of an endomorphism but also its outer class .
Let be an endomorphism. We say a subgroup system is -invariant if carries the subgroup system , i.e, there exists a set of elements and a function such that for all . A -invariant subgroup system is -fixed if its subgroups are permuted up to conjugacy, i.e., is a permutation and for all . When is an automorphism, all -invariant free factor systems are -fixed. When is injective, then, for any and free factor system of , is a free factor system of ; conversely, any free factor system of gives us a free factor system of via the isomorphism . Finally, note that when is a -fixed free factor system but is not surjective, will typically not be a free factor system of (See the remark above). A subgroup is eventually -periodic if for some , and it is -periodic if for some .
The endomorphism 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 will be a continuous map of graphs that sends vertices to vertices and any edge to a vertex or immersed path. An edge of is pretrivial if is a vertex. For a graph map of finite graphs, let be the maximum of the (combinatorial) length of the edge-path as varies over all the edges of . Then is -Lipschitz, a fact that will be used throughout the paper. Generally, will denote some convenient Lipschitz constant for rather than the infimum. For the moment, is used for arbitrary graphs but, for most of the paper, will be used for finite connected noncontractible graphs and for infinite simply-connected graphs (simplicial trees). A core graph is a graph with no proper deformation retract.
A direction at a vertex is a half-edge attached to the vertex. A vertex is bivalent if it has exactly two directions. The set of directions at vertex is denoted by . 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 has no pretrivial edges, then the restriction to initial segments induces the derivative map at , . The graph map is an immersion if it is locally injective, i.e., it has no pretrivial edges and the derivative maps are injective for all vertices ; note that immersions are natural. An immersion is expanding if the length of is unbounded as for every edge of .
With immersions defined, we preface the observation with a summary of Stallings’ folding theorem. Let be a rose whose edges are indexed by a basis of and let be the map where is the immersed edge-path in labelled by . Stallings [29] showed that factors as where each is a fold and is an immersion. We will use this factorization again to prove bounded cancellation in Lemma 1.4 and Proposition 3.1.
Observation.
If is irreducible, then it is injective.
Proof.
If is not injective, then at least one of the folds in Stallings’ factorization of collapses a subgraph of the domain. In particular, the kernel of contains a proper free factor ; therefore, and is reducible. ∎
Let be a nontrivial subgroup system of , then an -marked graph is a collection of graphs where the finite connected core graphs are indexed by markings, i.e., isomorphisms . When is a singleton, we may write -marked graph in place of -marked graph. A marked graph is an -marked graph. For any marked graph and any conjugacy class of a nontrivial element , denoted by , we define its length with respect to , , to be the (combinatorial) length of the immersed loop in representing .
More generally, we want to consider pairs of free factor systems of . For each component , let be either the nonempty maximal subset of carried by or the trivial system if no such subset exists. Typically, we shall replace components of with conjugates so that is also a free factor system of , i.e., each is a free factor system of (next remark below). A -forest is a simplicial forest of -trees , i.e., a collection of simplicial trees where each tree is equipped with a minimal simplicial -action whose edge stabilizers are trivial and point stabilizers are trivial or conjugates (in ) of -components. We note that a -tree is a point also known as a degenerate tree.
When is the trivial system, an -tree is a free minimal -tree . In that case, the quotient is a marked graph. When is a nontrivial proper free factor system, the quotient of an -tree is a graph of groups decomposition of with trivial edge groups and as the nontrivial vertex groups [28]; such decompositions are sometimes known as free splittings of and they will be our relative analogues for marked graphs. Any given -tree endowed with the combinatorial metric has an associated length function . Precisely, any isometry of a simplicial tree is either elliptic (fixes a point) or loxodromic (preserves an axis of least displacement). If is elliptic, then ; otherwise, is the translation distance of acting on its axis. When is trivial and is the marked graph , then as functions . The minimal subtree of a nontrivial subgroup is the smallest subtree with a minimal induced -action, i.e., the union of all fixed points and axes of nontrivial elements in .
Remark.
Let be free factor systems of , be a free factor with nontrivial , and be an -tree. Choose a connected fundamental domain of acting on and the nontrivial stabilizers in of vertices in the domain form a free factor system of . Then and are equivalent as free factor systems of since .
Suppose is an injective homomorphism. Then it determines a contravariant preimage function from the poset of free factor systems of to the poset of free factor systems of . One way to define this function is through trees. Let be a free factor system of , be any -tree, and be the minimal subtree of . In particular, the quotient is a free splitting of whose nontrivial vertex groups form a nontrivial free factor system of . If has a free -action, then this condition is independent of the choice of -tree and we set to be the trivial system. Otherwise, we have a nontrivial free factor system of and a corresponding nontrivial free factor system of via the isomorphism ; the equivalence class of the latter free factor system is independent of the -tree and we will denote it by .
Let and be - and -trees respectively and be an injective homomorphism such that carries or, equivalently, . This carrying condition ensures that elliptic elements in (with respect to ) have elliptic -images (with respect to ). A tree map is -equivariant if for all and . If we require , then loxodromic elements in have loxodromic -images; this is the relative analogue of a -injective graph map.
Suppose is a nontrivial -invariant free factor system of for some injective endomorphism ; so there is a function and inner automorphisms such that the restrictions are homomorphisms . The collection is a restriction of to , which we denote by . A weak representative for is a graph map from an -marked graph to an -marked graph such that for all . A weak representative for is a graph map of an -marked graph that is a disjoint union of weak representatives for . 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 has no bivalent vertices.
Recall . Suppose , is -invariant, and let be the function corresponding to the -invariance of and be a restriction of . An -relative weak representative for is a forest map of a -forest that is a disjoint union of -equivariant tree maps . An -relative natural representative is an -relative weak representative that is also natural. An -relative representative is an -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.) -relative immersion for is an -relative representative for that is also an (expanding resp.) immersion.
In the first part of the paper, we show that any injective endomorphism of has a canonical invariant free factor system and an expanding -relative immersion such that 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 be a marked graph with no bivalent vertices. For any nontrivial subgroup system of , the Stallings (subgroup) graph for with respect to a marked graph with is the smallest -marked graph along with an immersion such that for every . Alternatively, is the collection of cores of the covers of corresponding to , i.e., the smallest deformation retract of , and is the restriction to of the covering map . We may sometimes refer to the marked graph as the ambient graph. In the notation, the marking and immersion for Stallings graphs will usually be omitted. If and are in the same conjugacy class, , then there is a homeomorphism such that . The converse holds as well. So the Stallings graph is uniquely determined by the conjugacy class . Furthermore, the Stallings graph is a finite graph if and only if is finitely generated. Suppose is injective, is a nontrivial free factor system of . We will be studying the (iterated) Stallings graphs for .
Remark.
For any free factor system of and nontrivial subgroup system carried by , we can similarly define the Stallings graph for with respect to -marked graphs . If is an injective endomorphism, i.e., a collection of injective endomorphism that need not be a restriction of an endomorphism of , then we can still consider the Stallings graphs for and . The point of this remark is that the results of this paper hold when is replaced with . In other words, the ambient graph need not be connected for our results.
Unlike the ambient graph , we allow Stallings graphs to have bivalent vertices. More precisely, we subdivide so that the immersion is simplicial, i.e., maps edges to edges. With this subdivision, we get a combinatorial metric on that is compatible with , i.e., for any nontrivial element in , .
Lemma 1.1.
Let be injective and be a finitely generated nontrivial subgroup of . If is not eventually -periodic, then the length of the longest natural edge in is unbounded as .
Proof.
Suppose the length of the longest natural edge in with respect to some marked graph was uniformly bounded as . We want to show that is eventually -periodic. The number of natural edges in is bounded above by . Our assumption implies there is a bound on the volume of (number of edges in) the graphs as . So the sequence is eventually periodic, i.e., there are integers and an isometry such that . Since a Stallings graph determines the conjugacy class of its defining subgroup, we have , i.e., is eventually -periodic. ∎
This lemma will be invoked on invariant free factor systems containing a component that is not eventually -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 be injective and be a nontrivial -invariant free factor system. If all components in are eventually -periodic, then some nonempty subset is a -fixed free factor system and is carried by for some .
E.g., if is injective and is eventually -periodic, then is an automorphism.
Proof.
Let be the function used to define the -invariance of . Then there is a nonempty subset on which acts as a bijection and . Let by the nontrivial -invariant free factor system corresponding to . Then is carried by since the image of is . It remains to show that is -fixed. Set to be the order of , fix , and let be the inner automorphism such that . Define . As is eventually -periodic and hence eventually -periodic, there are integers such that Therefore, there is an element such that . But no finitely generated subgroup of is conjugate to a proper subgroup of itself (Lemma 1.3 below). So and, by injectivity of , . Since was arbitrary, fixes the free factors of up to conjugation; as is -invariant, it must be -fixed. ∎
The following fact will be used again in the proof of Proposition 2.4.
Lemma 1.3.
No finitely generated subgroup of 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 and element , there is a finite group and a surjective homomorphism such that . See [29] for a proof due to Stallings.
For a contradiction, suppose there is an element such that and . By subgroup separability, there is a finite group and homomorphism such that . But implies and, by finiteness of , implies — 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 in a graph , denotes the immersed edge-path that is homotopic to rel. endpoints; for a loop in , will be the immersed loop that is freely homotopic to .
Lemma 1.4 (Bounded Cancellation).
Let be a -injective graph map. Then there is a constant such that, for any natural edge-path decomposition of an immersed path in the universal cover , the edge-path is contained in the -neighborhood of .
The following proof is due to Bestvina-Feighn-Handel [4, Lemma 3.1].
Proof.
Any graph map factors as a pretrivial edge collapse and edge subdivision , a composition of folds , and an simplicial immersion . The collapse, subdivision, and immersion have cancellation constants while each fold has cancellation constant by -injectivity. Thus we may choose . ∎
Let be a topological representative for an injective endomorphism , be a nontrivial -invariant free factor system of , and be the disjoint union of covers of corresponding to for some . Then lifts to a map and the deformation retraction induces a map with and . We shall call the (-th) homotopy lift of .
Stallings graphs are -marked graphs by definition and the maps are weak representatives for that might map branch points to bivalent vertices. Note that injectivity of allows us to also consider the graphs as -marked graphs and as weak representatives for . We hope to replace the weak representatives 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 is to mapping branch points to branch points.
Lemma 1.5.
Suppose be a -injective graph map with cancellation constant . Then maps branch points to the -neighborhood of branch points.
Proof.
Set . If is the -neighborhood of its branch points, then there is nothing to prove. Suppose is a bivalent vertex whose distance to the nearest branch point is greater than . We need to show that is not the -image of any branch point. Set and to be the distinct oriented directions originating from and are the same directions with opposite orientation.
Let be a branch point and . As is a branch point, there are at least three distinct oriented directions originating from : , , and . Let be an immersed path that starts and ends with and respectively and define similarly. Set and , where is the reversal of the path . 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 starts with .

If ends with , then , where is an extension of to an embedded path from to a branch point and is an immersed nontrivial loop. By hypothesis, is longer than . Since starts and ends with and respectively, the concatenation is an immersed path such that has as a subpath, violating bounded cancellation. So we may assume starts and ends with and .
If starts and ends with and , then starts and ends with and respectively, which violates bounded cancellation for the same reason given in the previous paragraph. Similarly, if starts and ends with and , we rule out this possibility by considering . We have ruled out all cases, and therefore, no branch point of is mapped to . ∎
Corollary 1.6.
Let be a -Lipschitz -injective graph map with cancellation constant . Then is homotopic to a -Lipschitz natural graph map with cancellation constant .
Proof.
By Lemma 1.5, maps branch points to the -neighborhood of branch points. So we can find a graph map homotopic to that maps branch points to branch points. Since the homotopy is moving images of branch points a distance at most , we can use and .
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 , we may assume is a natural graph map. The homotopy will not affect the Lipschitz and cancellation constants. ∎
So we can replace the homotopy lift with a homotopic natural representative that has Lipschitz and cancellation constants and 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: is the restriction of to . To summarize the properties of homotopy lifts that will be used in the next sections:
Proposition 1.7.
Suppose is injective, is a topological representative for , and is a nontrivial -invariant free factor system. For any , there is a natural representative for with Lipschitz and cancellation constants and respectively.
The crucial point is that the Lipschitz and cancellation constants are independent of .
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 . This free factor system, called the elliptic free factor system, is crucial for the construction of (expanding) relative immersions later in the paper.
Suppose is a free factor system and is an injective endomorphism. We shall say that a conjugacy class in has an infinite -tail if for every , there is a conjugacy class in such that . The system carries an infinite -tail of a conjugacy class in if for every , there is a conjugacy class carried by such that . We now state and prove the main technical result of this section.
Theorem 2.1.
If is injective, then has a nontrivial fixed free factor system if and only if some nontrivial conjugacy class has an infinite -tail.
Remark.
Proof.
Let be an injective endomorphism. The forward direction is obvious: if has a nontrivial fixed free factor system , then any nontrivial conjugacy class carried by has an infinite -tail. The main content of the theorem is in the reverse direction. Fix a nontrivial conjugacy class in with an infinite -tail. We proceed by descending down the poset of free factor systems. The following claim is the key idea:
Claim (Descent).
Let be a -invariant free factor system that carries an infinite -tail of . If contains a free factor that is not eventually -periodic, then some -invariant free factor system carries an infinite -tail of .
Since there are no infinite chains in the poset of free factor systems, the descent (proven below) starts with the free factor system and then finds a -invariant free factor sytem that carries an infinite -tail of and whose free factors are eventually -periodic; such a free factor system contains a nontrivial -fixed free factor system by Lemma 1.2 and we are done.∎
Note that the assumption “ contains a free factor that is not eventually -periodic” in the descent claim is more general than we really need for our conclusion. We could have worked with a more natural assumption “ 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 be a marked graph, be a topological representative for , and set and . Suppose is a -invariant free factor system that carries an infinite -tail of and contains a free factor that is not eventually -periodic. Then, for all and , there is an immersed loop in corresponding to , where is the Stallings graph with respect to . Set .
For all , let be the natural representatives for given by Proposition 1.7. In particular, these representatives have Lipschitz and cancellation constants and respectively. As is an infinite -tail of carried by , we get, for any fixed , an infinite sequence of immersed loops in such that the free homotopy classes of have length for all .
Form a directed graph whose vertices are the natural edges of and there is a directed edge if maps natural edge over . Note that the number of natural edges of is at most and so has at most vertices.
Since contains a free factor that is not eventually -periodic, the length of natural edges in is unbounded as by Lemma 1.1. Fix such that the longest natural edge in is longer than . Let be the natural edges of longer than and be the union of and all natural edges on a directed path to in . Since is -Lipschitz and the shortest directed path in from a natural edge in to has at most natural edges on it, every natural edge in is longer than . The natural edges in will be referred to as the long natural edges and the remaining natural edges as .
Set to be the union of short natural edges, which will be a proper subgraph since long natural edges exist. The subgraph is automatically -invariant by the construction of . Since is an immersed loop in with length , its natural edges are short and is a nonempty, noncontractible proper subgraph of . Therefore, determines a nontrivial -invariant proper free factor system . Technically, it determines an invariant free factor system of but, as is injective, this corresponds to an invariant free factor system of . It remains to show that carries an infinite -tail of .
Let be the full subgraph generated by the long natural edges . If there are no directed cycles in , then ; in this case, the sequence of nontrivial loops in determines an infinite -tail of carried by and we are done. Now suppose there are directed cycles in and let be an immersed loop in that contains a long natural edge in such a cycle. Then, by bounded cancellation and the fact long natural edges are longer than , contains a long natural edge in the same directed cycle in for all . Consequently, none of the immersed loops in contain a long natural edge that is in a directed cycle of . Therefore, as far as the sequence of loops is concerned, we may assume there are no directed cycles in and, as before, the sequence determines an infinite -tail of carried by . ∎
The following dichotomy is (equivalent to) a result in Reynolds’ thesis.
Corollary 2.2 ([27, Proposition 3.11]).
If is irreducible, then either is an automorphism or only the trivial conjugacy class has an infinite -tail.
Proof.
Suppose is irreducible and there is a nontrivial conjugacy class with an infinite -tail. By Theorem 2.1, there is a nontrivial -fixed free factor system . Since is irreducible, and is an automorphism. ∎
The fixed free factor system given by Theorem 2.1 may depend on the chosen conjugacy class with an infinite tail or the marked graphs chosen in the descent. The next proposition constructs a canonical fixed free factor system for ; 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 is injective, then there is a unique maximal -fixed free factor system . Precisely, carries every conjugacy class with an infinite -tail and every finitely generated -fixed subgroup system.
Proof.
Let be an injective endomorphism. If the trivial conjugacy class is the only conjugacy class with an infinite -tail, then the trivial system is the only -fixed subgroup system. In this case, the trivial system is the unique maximal -fixed free factor system and we are done as it vacuously carries all conjugacy classes with infinite -tails and all -fixed subgroup systems. We can now assume some nontrivial conjugacy class has an infinite -tail. By Theorem 2.1, has a nontrivial fixed free factor system . We proceed by ascending up the poset of free factor systems:
Claim (Ascent).
Let be a finitely generated -fixed subgroup system, a conjugacy class with an infinite -tail, and a nontrivial -fixed free factor system. If does not carry both and , then some -fixed free factor system carries both and .
Once again, as there are no infinite chains in the poset of free factor systems, the ascent (proven below) starts with the nontrivial -fixed proper free factor and stops at a necessarily unique maximal -fixed free factor system that carries all finitely generated -fixed subgroup systems and all conjugacy classes with infinite -tails.∎
Proof of ascent.
Let be a finitely generated nontrivial -fixed subgroup system of , be a nontrivial conjugacy class in with an infinite -tail, and be a nontrivial -fixed free factor system of that does not carry both and . We now describe the descent:
Claim (Descent).
Let be a -invariant free factor system that carries , , and an infinite -tail of . If contains a free factor that is not eventually -periodic, then some -invariant free factor system carries , , and an infinite -tail of .
Starting with , the descent (proven below) will find a nontrivial -invariant free factor system that carries , , and an infinite -tail of and whose free factors are eventually -periodic. By Lemma 1.2, contains a -fixed free factor subsystem such that is carried by for some . As and are -fixed, they are carried by . Similarly, carries since carries an infinite -tail of . So is a -fixed free factor system that carries , , and as needed for ascent.∎
Proof of descent.
Let be a marked graph, be a topological representative for , and set and . Suppose is a -invariant free factor system that carries , , and an infinite -tail of . Let and be the Stallings graphs with respect to . Since and are finitely generated and -fixed, the length of the longest immersed loop in or that covers any edge at most twice is uniformly bounded by some for all . Set . The proof now mimics that of the descent in Theorem 2.1 and we only give a sketch.
For all , let and, by Proposition 1.7, there is a -Lipschitz natural representative for , , that has cancellation constant . As some free factor in is not eventually -periodic, we can fix so that the longest natural edge in is longer than by Lemma 1.1. Define the long and short natural edges as before and deduce long natural edges are longer than . Set to be the union of the short natural edges, which is necessarily proper and -invariant. Recall that immersed loops of and that cover any edge at most twice have length bounded by and these Stallings graphs have simplicial immersions into . Hence the images of these immersions lie in the subgraph of short natural edges . So is neither empty nor contractible. The subgraph determines a -invariant proper free factor system that carries both and since both and have immersions into . From the proof of descent in Theorem 2.1, implies carries an infinite -tail of . ∎
Although this proposition produces a canonical fixed free factor system for an injective endomorphism, we shall enlarge the system again to get a better -invariant free factor system that gives us some control of the relative dynamics of . 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 is injective and is the maximal -fixed free factor system, then there is a unique maximal -invariant free factor system such that is carried by for some . After replacing the free factors of with conjugates if necessary, we can assume .
We shall call the free factor system given by this proposition the -elliptic free factor system. We call it elliptic since it will be carried by the vertex groups in free splittings of in the subsequent sections.
Proof.
Let be an injective endomorphism and be the maximal -fixed free factor system of given by Proposition 2.3. If is surjective or has no nontrivial fixed free factor systems, then is or trivial respectively and we are done. Thus, we assume that is a nontrivial proper free factor system. For all , define .
Since is -invariant, we get that for all and, consequently, all are -invariant. As there are no infinite chains in the poset of free factor systems of , the chain of -invariant free factor systems stabilizes and we can set to be the maximal free factor system in the chain. By construction, carries a subgroup system if and only if is carried by for some and this implies the uniqueness of amongst -invariant free factor system that have -iterates carried by . It remains to show that after replacing the free factors of with conjugates if necessary.
As in the proof of Lemma 1.2, suppose is the function associated to the -invariance of . Then there is a maximal nonempty subset on which is a bijection. Let . Since carries the maximal -fixed free factor system , it follows that . Replace components of with conjugates if necessary and assume is a free factor system of ; in particular, each is a subgroup of some . We want to show that . Choose a component and let be the component such that . Furthermore, fix an inner automorphism such that for some . Set .
By construction, implies is conjugate to a subgroup of a -periodic, and hence -periodic, free factor for some . We must have since , carries , , and is a bijection. So for some inner automorphism . The -periodicity of implies is conjugate to for some . But Lemma 1.3 says no finitely generated subgroup of is conjugate to a proper subgroup of itself. Therefore, and, by injectivity of , . In particular, . As this holds for arbitrary free factors , we get . ∎
It is obvious that the maximal -fixed free factor system is proper exactly when is not surjective. The same holds for the -elliptic free factor system:
Observation.
Let be an injective endomorphism. The -elliptic free factor system is
-
1.
proper exactly when is not surjective.
-
2.
trivial exactly when only the trivial conjugacy class has an infinite -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 is an injective endomorphism and , i.e, is a -invariant proper free factor system such that there is no free factor system such that carries ; e.g., consider an injective nonsurjective 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 and be - and -trees respectively and be an injective homomorphism such that . If is a -equivariant tree map, then there is a constant such that for every natural edge-path decomposition of an immersed path in , the edge-path is contained in the -neighborhood of .
Proof.
The proof is the same as before. Since the trees are simplicial, the map factors as an equivariant pretrivial edge collapse and subdivision, a composition of equivariant folds, and an equivariant simplicial embedding. As is injective and , no fold identifies vertices in the same orbit and, hence, each fold has cancellation constant 1. We may choose . ∎
Let be a chain of -invariant free factor systems, be a -forest, and be a -relative representative for . For all , set to be the minimal subforest for ; minimal subforests are the relative analogues of iterated Stallings graphs. We will assume the minimal subforests inherit their simplicial structure from the ambient forest and so they might have bivalent vertices unlike . 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 , let be the restriction of to and then replace it with an equivariantly homotopic map that is induced by the deformation retraction of to , which we call the (-th) homotopy restriction of . Note that if is an axis such that is an immersion, then 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 and be - and -trees respectively and be an injective homomorphism such that . If is a -equivariant tree map with cancellation constant , then maps branch points to the -neighborhood of branch points.
Proof.
Set and let be a bivalent vertex in whose distance to the nearest branch point is greater than . In particular, has a trivial stabilizer. We denote by the 2 distinct directions at oriented away from the vertex. Suppose is a branch point of such that . As is injective, has a trivial stabilizer under the action of . Choose 3 distinct directions at : and . Let be an embedded path in that starts with and ends with a translate of . Since has a trivial stabilizer, the path determines a unique loxodromic element in with axis such that is a fundamental domain of the axis under the translation action of . Without loss of generality, starts with .
If ends with the translate , then , where is an extension of to an embedded path from to the axis of and is a fundamental domain of the axis of loxodromic element . By assumption, is longer than . Decompose the axis at , then has as a subpath, violating bounded cancellation. The remaining cases are handled similarly. Upon ruling out all cases, we conclude that no branch point of is mapped to . ∎
As before, we get a corollary whose proof is essentially the same as that of Corollary 1.6.
Corollary 3.3.
Let and be - and -trees respectively and be an injective homomorphism such that . If is a -equivariant -Lipschitz tree map with cancellation constant , then is equivariantly homotopic to a -equivariant -Lipschitz natural tree map with cancellation constant .
The corollary allows us to replace with an equivariantly homotopic -relative natural representative that has Lipschitz and cancellation constants and respectively. If is an axis and is an immersion before the homotopy, then is an immersed path after the homotopy; however, the restriction 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 be an -relative representative for . For any , there is an -relative natural representative for such that:
-
1.
and .
-
2.
If restricted to the axes of some conjugacy class in is an immersion, then restricted to the axes of is also an immersion modulo pretrivial edges.
Collapsing a maximal (orbit-closed) -invariant subforest of with bounded components and forgetting the bivalent vertices induces a minimal -relative representative for defined on a -forest . Note that the collapsed maximal subforest contains the pretrivial edges as . Since is induced by equivariantly collapsing a forest and forgetting bivalent vertices, we have , , and . If is an axis of in and is an immersion modulo pretrivial edges, then is an immersion, where is the axis of in .
Proposition 3.5.
Let be an -relative representative for . For any , there is a -forest and a minimal -relative representative for such that:
-
1.
and .
-
2.
is dominated by () the restriction ;
-
3.
If restricted to the axes of some conjugacy class in is an immersion, then restricted to the axes of is also an immersion.
For an -relative weak representative for , we define the transition matrix . Let be a nonnegative integer square matrix whose rows and columns are indexed by the orbits of edges in ; the entry of in row- and column-, , is given by the number of translates of that are contained in the immersed edge-path , where is a orbit representative for the -th orbit of edges. An -relative weak representative is irreducible if the matrix is irreducible, i.e., for any pair , there is a positive integer such that . In this case, the stretch factor of , , is the Perron-Frobenius eigenvalue of . An irreducible -relative weak representative is expanding if . Note that irreducible -relative representatives are minimal.
We say is irreducible relative to if there is no -invariant free factor system such that . If is trivial and , then we recover the definition of ’s irreducibility. The next lemma is the most useful property of an irreducible restriction for our purposes.
Lemma 3.6.
If is irreducible relative to , then every minimal -relative representative for is irreducible.
Proof.
Suppose some minimal -relative representative for has a reducible transition matrix; in particular, it has an invariant -equivariant proper subforest (with unbounded components) that determines a -invariant free factor system such that . So is not irreducible relative to . ∎
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 is irreducible relative to , then there is an irreducible -relative representative for with the minimal stretch factor, i.e., if is an irreducible -relative representative for , then .
The minimal stretch factor will be denoted by .
Proof.
Let be a minimal -relative representative for and suppose is irreducible relative to . Then is an irreducible -relative representative by Lemma 3.6 with stretch factor . By the lack on bivalent vertices, any irreducible -relative representative has a transition matrix of size . Suppose is an irreducible integer square matrix with Perron-Frobenius eigenvalue . Then has a positive right eigenvector associated with . So for all , has right eigenvector associated with eigenvalue . Assuming the smallest entry of is (rescale if necessary), we get that the minimum row-sum of is at most for any . If has no more than rows, then the largest entry of is at most the minimum row-sum of , which we know is at most . So there are finitely many irreducible integer square matrices with size and Perron-Frobenius eigenvalue . Thus, there is a finite set of stretch factors for irreducible -relative representatives for . In particular, there is an irreducible -relative representative for with the minimal stretch factor. ∎
Bestvina-Handel’s work [6] can be adapted to show that an irreducible -relative representative for with the minimal stretch factor is an -relative train track (Appendix A) and, conversely, bounded cancellation implies all irreducible -relative train tracks for 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 is irreducible relative to and is an irreducible -relative representative for with the minimal stretch factor, then there is an element in with an axis such that the restriction of to is an immersion for all .
Such an axis will be known as an -legal axis.
Proof.
If , then is a simplicial embedding and we are done. So we may assume . By minimality of its stretch factor, is an expanding irreducible -relative train track for (Theorem A.1), i.e., for any edge in , is an expanding immersed path for all . A 2-edge path is -legal if it is a translate of a subpath of for some edge and integer . By irreducibility of , every edge is contained in a 3-edge path whose 2-edge subpaths are both -legal. This means we can form an axis whose 2-edge subpaths are all -legal. By the train track property, the restriction of to is an immersion for all . ∎
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 be a chain of -invariant free factor systems with irreducible relative to and . If carries the maximal -fixed free factor system, then the length of the longest natural edge in is unbounded as .
Before starting the proof, we will first describe the (absolute) vertex blow-up construction. Let be a chain of free factor systems and be some -forest. Recall that we assume is also a free factor system of . Fix some -marked roses . Define to be the graph formed by identifying the appropriate vertices of the graph of groups with the basepoints of roses . If is the inclusion map, then Bass-Serre theory gives markings such that . Thus, is a -marked graph. This construction and, in general any pair of graphs with collections of markings such that will be referred to as vertex blow-up.
We note that the Stallings graph with respect to , as a -marked graph, is a vertex blow-up of : let be the Stallings graph’s immersion and be the core of the subgraph . Since , is marked by an isomorphism and is the restriction of the marking to with respect to the inclusion . Therefore, is also a vertex blow-up of . The noncontractible components of the subgraph will be known as the lower stratum and the rest of the graph as the top stratum.
Now suppose are also -invariant and let be a -relative representative for defined on some -forest and be a topological representative for . Construct a topological representative for that agrees with on the -marked roses and induces on the Bass-Serre forest upon collapsing the roses .
For any , we let be the minimal -relative representative for given by Proposition 3.5 using and be the natural representative for given by Proposition 1.7 using . By Proposition 3.5(3), if an element in has an -legal axis, then has a -legal axis. It can be arranged for to be -invariant and to induce on the -forest upon collapsing a maximal invariant proper subgraph of containing and forgetting bivalent vertices.
Here is the idea behind the proof. By irreducibility of the restriction , we may assume the map is an expanding irreducible representative for . For the contrapositive, suppose the forests had uniformly bounded natural edges. There is a sequence of loxodromic elements in with uniformly bounded translation lengths . Now suppose that the vertex blow-up had natural edges with aribtrarily long top stratum subpaths. Bounded cancellation, the fact induces , and the irreducibility of imply is an expanding irreducible immersion. However, this contradicts the first assumption since . So the second supposition is false and the natural edges of have top stratum subpaths with uniformly bounded length. Using the Lipschitz property, expanding irreducibility of , and existence of a -legal axis (train track theory), we find uniformly bounded lower stratum paths in 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 that contain top stratum subpaths. This implies some loxodromic conjugacy class in has an infinite -tail. By Proposition 2.3, any conjugacy class in with an infinite -tail is carried by the maximal -fixed free factor system. Thus the maximal -fixed free factor system cannot be carried by as it carries a loxodromic conjugacy class.
Proof of Proposition 3.9.
Suppose are -invariant free factor systems, is irreducible relative to , is a topological representative for defined on -marked roses , and is an expanding irreducible -relative representative for with the minimal stretch factor (Proposition 3.7). By Lemma 3.8, there is an element in with an -legal axis. Set to be the vertex blow-up of with respect to the -marked roses . The discussion preceding the proof gives minimal -relative representatives for and natural representatives for that have these properties: for all ,
-
1.
induces on upon collapsing the -invariant subgraph ;
-
2.
and ;
-
3.
is dominated by the restrictions ; and
-
4.
has a -legal axis.
The collection is conjugate to by injectivity of . So is irreducible relative to and is the minimal stretch factor for relative to . Furthermore, the minimal -relative representatives are irreducible (Lemma 3.6) and by the minimality of .
Suppose for the contrapositive that there is a bound such that all natural edges in are shorter than for all . Then, for all , there is a loxodromic element in such that , where . Every edge in lifts to a -orbit of a natural edge in , which corresponds to a top stratum subpath of a natural edge in .
Claim.
The subpath in has length for all edges in and .
Suppose, the graph has an edge whose corresponding subpath in is longer than for some . As we did in the proof of Theorem 2.1, we construct the set of long edges by looking at all the edges of that are eventually mapped over . Here, an edge in mapped over if there are lifts and in such that maps over . Since is -Lipschitz and it induces on , each long edge in corresponds to a top statrum subpath in longer than . Since is an irreducible -relative representative, all edges eventually map over and hence are long. The long natural edges of will be the natural edges in containing top stratum subpaths.
Suppose an edge of had a lift in that is the initial segment of the -image of two edges that share an initial vertex. Then the top stratum subpath is in a long natural edge of that is the initial segment of -images of natural edges that share an initial vertex; this violates bounded cancellation since long natural edges of longer than . Hence, there is no folding in , i.e., is an expanding irreducible -relative immersion. We may now find an such that all loxodromic elements in have . Since is dominated by , we get that for all loxodromic elements in . This contradicts the assumption that for some loxodromic in . So the top stratum subpath in has length for all natural edges of and . This ends the proof of the claim.
Next, we prove the existence of paths in the lower stratum of with uniformly bounded lengths connecting top stratum paths. Suppose and are edges of with lifts and in such that is a subpath of the immersed path . Then is a subpath of immersed path for some lower stratum path in . Since has length bounded by and is -Lipschitz, the path has length bounded by . We say the 2-edge path in has a nondegenerate turn bounded by .
As is an expanding irreducible relative representative that has a legal axis (this is where the argument invokes train track theory), every edge in can be extended to an immersed 3-edge path that is a translate of a subpath of and . In particular, any edge in can be extended to a 3-edge path whose 2-edge subpaths both have nondegenerate turns bounded by , i.e., every top stratum subpath can be extended to an immersed path with top stratum subpaths , and lower stratum paths , with length bounded by .
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 in with the properties:
-
1.
lifts to an axis in for some loxodromic conjugacy class ;
-
2.
passes any edge of at most twice and only takes short turns (including the turn at the endpoint), which implies it has at most edges and (short) turns; and
-
3.
represents a loop in with length bounded by .
In summary, for each , we construct a loxodromic conjugacy class in whose -length is bounded by a constant independent of . As there are finitely many conjugacy classes with -length bounded by any given constant, the sequence of conjugacy classes has a constant infinite subsequence. Thus, some loxodromic conjugacy class has an infinite -tail carried by . Recall that the maximal -fixed free factor system carries all conjugacy classes with an infinite -tail (Proposition 2.3); on the other hand, does not carry loxodromic conjugacy classes in by definition. Therefore, cannot carry the maximal -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 be an injective endomorphism, be a pair of -invariant free factor systems such that , and be a restriction of to . Recall that a relative representative is minimal if it has no orbit-closed invariant subforests with bounded components and an expanding -relative immersion for is an -relative representative for that is a minimal immersion whose edges expand under -iteration.
There will be two possible ways of obtaining a relative immersion from a relatively irreducible restriction with a minimal stretch factor . If , then an irreducible representative with stretch factor is automatically a simplicial immersion. The next proposition shows how to construct an immersion when . This construction is unique to nonsurjective endomorphisms because we require that the restriction be irreducible relative to a free factor system that carries the -elliptic free factor system — when is an automorphism, the -elliptic free factor system is and no such restriction exists.
Proposition 4.1.
Let be injective and be a chain of -invariant free factor systems that carry the -elliptic free factor system. If is irreducible relative to and , then there is an expanding irreducible -relative immersion for .
Proof.
Suppose is injective, are -invariant free factor systems that carry the -elliptic free factor system, is irreducible relative to , and is an expanding irreducible -relative representative for with minimal stretch factor . Set and . By Proposition 3.4, there is a natural representative for with Lipschitz and cancellation constants and respectively for all .
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 such that the set of natural edges in longer than is not empty, where . Choose to be the set of all natural edges that eventually get mapped over those in by and call the long natural edges. As is -Lipschitz and there are at most orbits of natural edges in , the long natural edges are longer that . Injectivity of implies is conjugate to . So is irreducible relative to , is the minimal stretch factor for relative to , and the short natural edges of form an orbit-closed -invariant subforest with bounded components.
Collapse a maximal -invariant subforest of that has bounded components and contains the short natural edges then forget the bivalent vertices; this induces a minimal -relative representative for . The map is an irreducible -relative representative for (Lemma 3.6) and by the minimality of . So is an expanding irreducible -relative representative.
Since the lifts in of all edges in are longer than the cancellation constant , there is no folding in — otherwise, there would be folding in identifying paths longer than its cancellation constant, absurd. Thus, is an immersion. By injectivity of , we can view as a -forest and as an expanding irreducible -relative immersion for . ∎
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 that is irreducible relative to the -elliptic free factor system will satisfy .
Proposition 4.2.
Let be injective and be a chain of -invariant free factor systems where is the -elliptic free factor system. If is irreducible relative to , then there is an expanding irreducible -relative immersion for .
Proof.
Let be injective and be irreducible relative to the -elliptic free factor system . Then there is an irreducible -relative representative for with stretch factor (Proposition 3.7). We say is loxodromic if is not a point, i.e., contains a loxodromic element; similarly, the component of the graph of groups is loxodromic if is loxodromic. If , then the induced map on the loxodromic components of is a graph isomorphism. So for some , if is carried by a loxodromic , then is -invariant. By Proposition 2.4, the subset of all carried by the loxodromic component of form a -fixed free factor subsystem. As induces a graph isomorphism on the loxodromic components of and these components’ vertex groups form a -fixed free factor system, we get that is surjective when restricted to the unbounded components of the forest and the loxodromic components of form a -fixed free factor system. This is a contradiction since -periodic free factors are elliptic (Propositions 2.3 and 2.4). Therefore, and the result follows from Proposition 4.1. ∎
Specializing this proposition to the case where is irreducible and nonsurjective yields an alternate proof to a result due to Reynolds.
Corollary 4.3 ([27, Corollary 3.23]).
If is irreducible but not surjective, then 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 be injective, be the -elliptic free factor system, and be a chain of -invariant free factor systems. If there is an expanding -relative immersion for and a -relative immersion for , then there is an expanding -relative immersion for .
Although the proof gets a bit technical, the idea is rather simple: a -relative immersion for (top stratum) and an expanding -relative immersion for (lower stratum) can be patched together via a (relative) vertex blow-up to get a minimal -relative representative whose only possible folds would have to happen between a top and lower stratum edge of . As the restriction of 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 is possible. Thus, is a minimal -relative immersion for , which will be expanding if is the -elliptic free factor system.
Let be a -forest and be a -forest. For any free factor , let the maximal subset of that is carried by . Replace the free factors of with conjugates if necessary and assume is also a free factor system of . In particular, the free factors are subgroups of the free factors . Identifying the appropriate vertices of the graph of groups with basepoints on the graph of groups results in a graph of groups decomposition for whose Bass-Serre forest is a -forest that contains . We call the vertex blow-up of with respect to .
Proof of Proposition 4.4.
Let be injective, be a chain of -invariant free factor systems that carry the -elliptic free factor system . Suppose is an expanding -relative immersion for and is a -relative immersion for then define to be the vertex blow-up of with respect to . The edges of are of two types: the lower stratum, which are edges that are contained in the -orbit of , and the top stratum, which are the remaining edges.
Let be a minimal -relative representative for such that is an -invariant subforest, the restriction of to agrees with , and induces upon collapsing the lower stratum. For all , set and to be the minimal subforests of for and respectively. Similarly, define the minimal subforest . By the inclusion of in , we get a simplicial identification of with . 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 that are bivalent when considered as points on the subforest . So by “natural edges of ”, we mean those inherited from the parent forest ; on the other hand, by “natural edges of ”, we do mean exactly that. Under the identification of the two forests, the natural edges of partition any natural edge of into at most segments, where .
Since is a -relative immersion for , the restrictions of to are -relative immersions for that are conjugate to . As induces upon collapsing the lower stratum, any edges in but not must be in the lower stratum and the restriction of to induces upon collapsing the lower stratum, i.e., the -orbit of . By Proposition 3.4, there is an -relative natural representative for with Lipschitz and cancellation constants and respectively. Furthermore, is an -invariant subforest and still induces upon collapsing the lower stratum. As agrees with the immersion on , differs from on by a homotopy supported in the natural edges of .
Since is an expanding -relative immersion for , the minimal subforest has natural edges whose lengths are all exponential in . Fix such that all natural edges of are longer than . By the pigeonhole principle, each natural edge of contains a natural edge of longer than . Let be the directed graph of natural edges of the -invariant subforest where a directed edge corresponds to mapping over . Set to be those natural edges with length at most and to be those natural edges with directed path from in and their -translates; these lower stratum natural edges will be the short natural edges of . Since is -Lipschitz and the shortest path between any two natural edges in has natural edges, all the short natural edges have length at most . So the short natural edges form an orbit-closed -invariant lower stratum subforest of with bounded components as does not cover any natural edge of .
Collapsing the short natural edges of induces a map with the same cancellation constant . Now iteratively collapse pretrivial edges until the induced map has none. As the collapses are supported in the lower stratum, the new map still induces the immersion upon collapsing the rest of the lower stratum and, as a result, folding in may only occur between initial segments of natural edges of whose -images are a lower stratum natural edge. However, all natural edges in the lower stratum of are longer than by construction and so no folding in is possible by bounded cancellation, i.e., is an immersion.
Collapsing a maximal invariant subforest with bounded components and forgetting bivalent vertices if necessary, we may assume is a minimal -relative immersion for . By the injectivity of , we can view as a minimal -relative immersion for . It remains to show that every edge of expands under -iteration.
For a contradiction, suppose there is an edge of whose -iterates have uniformly bounded length. Since is minimal, the non-expanding edges in the graph of groups contain a setwise fixed subgraph that carries a loxodromic element. The subgraph is a free splitting of a -invariant free factor system. Recall that the -elliptic free factor system decomposes as a union of the maximal -fixed free factor system and free factors that eventually get mapped into this fixed system (Proposition 2.4). By construction, the point stabilizers of are conjugates of . So any vertex of the fixed graph is labelled by either the trivial group or a free factor of the maximal -fixed free factor system. Thus is a free splitting of a -fixed free factor system that carries some loxodromic element. However, Propositions 2.3 and 2.4 imply that all -fixed free factors systems are elliptic — a contradiction. ∎
We are now ready to inductively construct expanding relative immersions.
Theorem 4.5.
If is injective but not surjective, then there is an expanding -relative immersion for , where is the -elliptic free factor system.
Proof.
Suppose is injective but not surjective. By Proposition 2.4, the -elliptic free factor system is proper. The naive approach is to assume there exists a chain in the poset of -invariant free factor system such that the restrictions are irreducible relative to for all . This assumption is typical when working with automorphisms. For each restriction , if the minimal stretch factor , then there is automatically a -relative immersion for ; and if , then there is an expanding -relative immersion for by Proposition 4.1. In either case, there is a -relative immersion for . By Proposition 4.2, and there is an expanding -relative immersion for . By inductively patching these immersions together using Proposition 4.4, we get an expanding -relative immersion for and we are done. Unfortunately, since is not surjective, it could be that no chain satisfies the naive assumption we made at the start. Recall that our definition of being irreducible relative to presupposed . 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 in the poset of -invariant free factor system that we will induct on. Let be a chain of -invariant free factor systems such that is irreducible relative to . Suppose has been constructed for some and let be the maximal free factor system in the chain of -invariant free factor systems. If , then let be a chain of -invariant free factor systems such that is irreducible relative to . If , then let be the chain of -invariant free factor systems such that for .
We proceed by inducting on the resulting chain between and . For the base case, is irreducible relative to ; therefore, there is an expanding -relative immersion for by Proposition 4.2. For our induction hypothesis, suppose that there is an expanding -relative immersion for for some . By our construction of the chain, either is irreducible relative to or is carried by . We deal with these two cases separately.
Case 1. Suppose is irreducible relative to . By Proposition 3.7, there is an irreducible -relative representative for with minimal stretch factor. If , then is a -relative simplicial immersion for . If , then there is an expanding -relative immersion for by Proposition 4.1. In either case, we get a -relative immersion for defined on a -forest . Thus, there is an expanding -relative immersion for by the induction hypothesis and Proposition 4.4.
Case 2. Now suppose is carried by . Let be the minimal subforest of and be the same forest after forgetting bivalent vertices. By injectivity of , we may consider as a -forest that comes with a natural -equivariant immersion . Since is an immersion, we can identify a subdivision of with the minimal subforest of . Composing with the subdivision and inclusion gives an -relative immersion for , which is expanding since its image lies in and its restriction to is the expanding -relative immersion 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 is injective and there is an expanding -relative immersion for , where is a -invariant free factor system. Then there is a unique expanding -relative immersion for .
Proof.
Let be injective, a -invariant free factor system. Suppose and are expanding -relative immersions for . The goal is to show that and are equivariantly homeomorphic. Let be an equivariant tree map, i.e., -equivariant tree map with being the identity automorphism; such a map always exists between -trees. By taking restrictions to the -minimal subtrees and and applying deformation retractions, we get a -equivariant tree map with cancellation constant . By Corollary 3.3, we assume is an equivariant natural tree map with cancellation constant .
Recall that is an expanding -relative immersion for , so all the natural edges of are longer than for large enough . Fix , then no folding occurs in by bounded cancellation. So is a forest collapse and has at most the same number of orbits of branch points as . By the same argument, is an equivariant forest collapse of and has at most the same number of orbits of branch points as . This implies the two minimal subtrees have the same number of orbits of branch points and is an equivariant homeomorphism. Since and are immersions, and are -equivariantly homeomorphic to and respectively. Therefore, and are equivariantly homeomorphic by injectivity of . So and are homotopic immersions and hence agree up to isotopy. ∎
Corollary 4.7.
If is injective but not surjective, then there is a unique expanding -relative immersion for , where is the -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 be an injective endomorphism. Then the following conditions are equivalent:
-
1.
is induced by a unique expanding graph immersion;
-
2.
is induced by an expanding graph immersion;
-
3.
the trivial conjugacy class is the only conjugacy class with an infinite -tail;
-
4.
the trivial system is the only -fixed free factor system.
Proof.
We leave the implication as an exercise. By Theorem 2.1, we have . It remains to show . Suppose the trivial conjugacy class is the only conjugacy class with an infinite -tail. So is not surjective and the -elliptic free factor system is trivial. By Corollary 4.7, there is a unique expanding -equivariant immersion defined on a free -tree , i.e., is induced by a unique expanding immersion on the marked graph . ∎
Interlude
Let us summarize the main results of the first part of the paper.
Summary.
If is injective but not surjective, then there is:
-
1.
a unique maximal proper -fixed free factor system ; (Proposition 2.3)
-
2.
a unique maximal proper -invariant free factor system such that is carried by for some ; after replacing the free factors of with conjugates if necessary, we can assume ; (Proposition 2.4)
-
3.
a unique expanding -relative immersion for . (Corollary 4.7)
We call the free factor system the -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: is induced by an expanding graph immersion if and only if the trivial system is the only -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 be induced by an expanding graph immersion. is word-hyperbolic if and only if it has no subgroups for .
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 -periodic subgroups are an obstruction to 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 be an injective endomorphism with no nontrivial fixed free factor system. is word-hyperbolic if and only if it has no subgroups for .
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 of , 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]: is word-hyperbolic if and only if it has no 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 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 groups fail the annuli flaring condition.
Consider two Baumslag-Solitar groups with . Note that . Fix the presentation and notice that the group is a mapping torus of an endomorphism of the cyclic group . So acts on a simplicial tree with point and edge stabilizers conjugate to and the quotient graph has one vertex and edge. The edge has a natural orientation inherited from the nontrivial loop corresponding to and we can lift the orientation to an equivariant orientation of . Under this orientation, is like a rooted tree: every vertex has exactly incoming edges and one outgoing edge. Choose a vertex on the axis of whose stabilizer by and define to be the family tree of -descendants, i.e., union of directed edge-paths that terminate on .

A class of annuli of length in is an ordered pair where is nontrivial and is a geodesic of length in fixed by . Since is elliptic, then (up to conjugacy) we may assume , , and . The orientation on implies there are two types of annuli: 1) unidirectional — those whose geodesics are directed paths terminating on ; 2) bidirectional — those whose geodesics are a wedge of two directed paths terminating on . When , 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 is the identity map. When , is a regular trivalent tree and annuli can be unidirectional or bidirectional. Since the associated endomorphism is represented by a degree 2 map of the circle, the unidirectional annuli will flare. However, due to the branching in , 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 that has no subgroups with and thus conclude that the mapping torus satisfies the annuli flaring condition. Here is an outline of the second part of the paper:
-
1.
In Section 5, we prove that if has no subgroups with , 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.
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 has no subgroups, then long unidirectional annuli flare. Thus, together with the previous section, satisfies the annuli flaring condition if it has no subgroups with .
-
3.
Finally, in Section 7, we extend the result to HNN extensions of free groups over free factors by showing that there is a canonical finite collection of mapping tori that carry all long annuli of . Intuitively speaking, the action of on its Bass-Serre tree is acylindrical relative to . So satisfies the annuli flaring condition if and only if the subgroups satisfy the condition.
Geometry of ascending HNN extensions
5 Iterated pullbacks
In our previous work [23, Sections 3], topological pullbacks for a graph immersion were used to give sufficient conditions for 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 .
Suppose immersions and induce inclusions of free groups respectively. Then components of the core (topological) pullback of , also known as fibered products are in one-to-one correspondence with nontrivial intersection as ranges over -double coset representatives of . For a graph immersion that induces an endomorphism , we get a one-to-one correspondence between components of the core pullback of and nontrivial conjugacy classes as ranges over -double cosets for all . We will not define pullbacks topological maps since we will be working (semi-)algebraically in this section.
Given subgroups , we define the (algebraic) pullback of , denoted by , to be the set of all nontrivial components as ranges over the -double cosets in . Note that pullbacks are not subgroup systems since they are sets of conjugacy classes of subgroups. When and are finitely generated, then their pullback is a finite set. For an injective endomorphism and , define the iterated (algebraic) pullbacks of to be . Furthermore, the iterated pullback depends only on the outer class and will be denoted later by . In this section, we will simply write for brevity.
The (algebraic) mapping torus of an injective endomorphism is the ascending HNN extension of with the presentation We now give a third description of that will actually be the working description. The mapping torus is the fundamental group of a circle of groups with one vertex group and edge group . The edge monomorphisms for this circle of groups are the identity map and endomorphism and the corresponding Bass-Serre tree 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 acts on its directed axis with positive translation.
By construction, there is a vertex of whose stabilizer in is exactly . Let be the full subtree of rooted at . Colloquially, is the family tree of -descendants. We will use the action of on to index the vertices of by cosets in . The root is indexed , where is the coset of in . The -generation vertices are indexed by the cosets in . The intersection corresponds to the simultaneous stabilizer of the ordered pair of vertices indexed by and ; the intersection is conjugate in to the stabilizer of the vertices indexed by and , where . The intersection is also conjugate (in ) to the stabilizers of the vertices indexed by and , where . Thus, -orbits of -generation ordered pairs of vertices are indexed by -double cosets in and conjugacy classes of their nontrivial stabilizers are the components of .
There is a function induced by , given algebraically by
Graphically, the function maps (the conjugacy class of) the stabilizer of to (the conjugacy class of) the stabilizer of .
Lemma 5.1.
as a function on the set of components is injective.
Graphical proof.
If and are in the same -orbit, then they are in fact in the same -orbit (descendants of ). So and are in the same -orbit by the action of . ∎
One can give an algebraic proof of the lemma that replaces the action of with the injectivity of . The lemma implies there is a chain of injections . Furthermore, the restriction to each component is an isomorphism since it is a conjugation by in . We will be mainly interested in the set compliment . Equivalently, Graphically, is the set of components that stabilizes -orbits of -generation pairs of vertices in whose only common ancestor is . We might say iterated pullbacks stabilize if for some . The image is malnormal in if and only if . Iterated pullback stability is a sort of generalization of malnormality for with respect to .
Lemma 5.2.
Suppose is injective and . If is empty, then so is ; and if has only cyclic components, then is empty or has only cyclic components.
Proof.
There is an “inclusion” of components, , since a stabilizer of a -generation pair of vertices is contained in a stabilizer of a -generation pair of vertices. So implies . Suppose has cyclic components, then 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 is and the reduced rank of a pullback , where are finitely generated free groups, is the sum of the reduced ranks of components in . The latter is denoted by . Since the restriction gives natural isomorphisms of the components, the chain of injections produces a nondecreasing sequence of positive integers
Observe that if and only if 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 are nontrivial finitely generated subgroups, then .
Remark.
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 is injective, then either has cyclic components for all or for some .
Proof.
Theorem 5.3 gives us a uniform bound on the reduced ranks of the iterated pullbacks for all . By Lemma 5.2 and the uniform bound on the nondecreasing sequence of reduced ranks, are all equal for . Therefore, is empty or has only cyclic components. The lemma follows by inductively applying Lemma 5.2. ∎
We say has an invariant cyclic subgroup system with index if there is an integer , element , and nontrivial cyclic subgroup such that and has index . Note that this is actually a property of the outer endomorphism . We can now give the main result of this section:
Proposition 5.5.
Let be injective. If is not empty for all , then has an invariant cyclic subgroup system with index .
Proof.
Let and be an injective endomorphism such that is not empty for all . By Lemma 5.4, has cyclic components for .
So far nothing in the section has used relative immersions but our main motivation for constructing them was this proposition. Note that automatically implies is not surjective. So is injective but not surjective. By Theorem 4.5, there is an expanding -relative immersion for , where is the -elliptic free factor system and is an -tree. Recall that the equality is part of our definition of -relative representatives. An element of is elliptic if it fixes a point on , i.e., its conjugacy class is carried by . By construction of the -elliptic free factor system (Proposition 2.4), there is an such that is carried by the maximal -fixed free factor system for all and elliptic . As is a -fixed free factor system, the components of are free factors of for all .
Suppose for some . As , this component is cyclic and we may assume it has a representative generated by a nontrivial element . In particular, there exists and such that . We first show that is loxodromic, i.e., not elliptic. Suppose is elliptic, i.e., carries . Then some free factor carries as . So must be elliptic too since and, as noted in the previous paragraph, carries and . Thus, up to conjugation in , we may assume and, up to conjugation in , we may assume . Malnormality of the free factor implies . But leads to the contradiction . Therefore, is loxodromic. The integer and component were arbitrary, so all components of are loxodromic for . Recall from the proof of Lemma 5.2 that there is an infinite descending chain:
Since has finitely many components and the components are all cyclic, there is a cyclic component in which “carries” some component of for all . Suppose this component has a representative generated by . Then for all , there is a cyclic component of with a representative generated by such that carries . If we let be the axis for element , then the previous sentence implies there are sequences of element and such that the (unoriented) axes of are all translates of the (unoriented) axis .
For any , replace with its inverse if necessary so that the action of on its axis is coherent (respects orientation) with the action of on . By passing to a strictly increasing subsequence , we may assume there is an oriented edge of such that the axes of all contain a translate of . We now pass to the graph of groups in order to avoid mentions of translates and orbits. The axis will project to an immersed loop in representing and axes project to immersed loops that represent and whose -image is a power up to rotation/cyclic homotopy, where and is the immersion induced by . The edge projects to an edge that is contained in all the loops for
The proof now mimics the proof of [23, Proposition 3.11]. Since is an immersion, it maps axes in onto axes and maps immersed loops in to immersed loops. So is a subpath of the immersed loop for all , and since is expanding, contains arbitrarily long powers of as . Set to be the number of subpaths of (up to rotation) that are also loops. Choose such that contains the loop as a subpath. Then is a subpath of that contains the loop as a subpath. In fact, for all positive integers , the loop is a subpath of . Thus, there is a sequence of loops that are subpaths of and strictly increasing positive integers such that is up to rotation. By definition of and pigeonhole principle, some for some and is up to rotation. Lifting back to the -tree , we find that maps the axis to a translate of itself. So is conjugate to a nontrivial power and since 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 , we can construct a specific number for which implies has an invariant cyclic subgroup system with index . Thus, one would not have to check infinitely many iterated pullbacks of to know that an invariant cyclic subgroup system with index exists. Of course, for this to be useful, we need to know whether finding the -elliptic free factor system and expanding relative immersion for 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 is injective and has no invariant cyclic subgroup system with index , then iterated pullbacks of stabilize. In this case, we get control on the types of annuli in the mapping torus , which allows us to prove the main theorem: is word-hyperbolic if additionally has no fixed cyclic subgroup system, i.e., invariant cyclic subgroup system with index .
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 and iterated pullbacks of . 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 is a based-hyperbolic graph map and all strictly bidirectional annuli in its mapping torus are shorter than some integer, then is word-hyperbolic.
We will define the new terms in the theorem as we go. Suppose is an injective endomorphism and is its topological representative. Recall that the (topological) mapping torus of is the quotient space with the identification for all and the algebraic mapping torus is isomorphic to the fundamental group . The edge-space of will be the cross-section in represented by .
Strictly bidirectional annuli of the mapping torus with length can be thought of as the iterated pullbacks of but it does take a bit of work to give the correspondence. Fix a basepoint . For integers , an (topological) annulus in of length is a homotopy of loops satisfying the following conditions:
-
1.
it is transverse to the edge-space of ;
-
2.
the -preimage of the edge-space is ;
-
3.
for integers , the rings of the annulus are locally injective every where except possibly at the basepoint ;
-
4.
and for the trace of the basepoint , no subpath between consecutive integer coordinates is homotopic rel. endpoints into the edge-space.
This is the definition used in [23]. In light of our last description of in the previous section, we give alternative definition. Let be the Bass-Serre tree for and be the point whose stabilizer is . An (algebraic) annulus in of length is a choice a nontrivial conjugacy class in and an orbit of a geodesic path in of length fixed by . Since elements of act on with fixed points, we can always choose a representative . Technically, we have defined a conjugacy class of algebraic annuli but the distinction will not be relevant for us.
Lemma 6.2.
Let be injective and be a topological representative for . There is a one-to-one correspondence between (homotopy classes of) annuli in of length and (conjugacy classes of) annuli in of length .
Proof.
Given a topological annulus in of length , then the generator of the image determines a conjugacy class in . Condition (3) ensures is injective and is nontrivial. Let be the lift of the annulus to the universal cover of . Collapsing the -direction of produces the Bass-Serre tree and Condition (2) ensures the induced map is constant on the first factor and its image is a collection of edge-midpoint; each ring determines a conjugacy class in the stabilizer of the corresponding edge-midpoint. By Conditions and , the midpoints extend to a geodesic edge-path in of length fixed by .
The other direction works in a similar fashion. For any conjugacy class in and two consecutive edge-midpoints in fixed by , we can construct an annulus of length as follows. Fix a basepoint in the edge-space and assume is the vertex between the midpoints. If the midpoints are increasing/decreasing, then without loss of generality. Let be based loop in the edge-space representing and a based loop in representing . Then the based path is null-homotopic and can be extended to an annulus with ends and trace . If the midpoints are at the same height, then and for some . Let be based loops in the edge-space representing respectively and a based loop in representing . So the based path is null-homotopic and can be extended to an homotopy between satisfying Conditions (1)-(3) and having trace . This trace satisfies Condition (4) because and thus the homotopy is a topological annulus with ends . Given a geodesic path in of length fixed by , we can replace the path with a translate in without affecting the class and then construct a topological annulus in of length by concatenating the length 1 annuli from the preceding discussion. This concludes the correspondence between topological and algebraic annuli. ∎
The natural orientation on gives a dichotomy for annuli in :
-
1.
all edges of have the same orientation— we say is unidirectional.
-
2.
switches from increasing to decreasing exactly once — we say is bidirectional.
Each vertex of has exactly one outgoing edge and hence the geodesic cannot switch from decreasing to increasing because . For similar reasons, bidirectional annuli do not exist if and only if is malnormal in . The next proposition generalizes this equivalence of bidirectional annuli (or lack thereof) and malnormality.
An annulus in is strictly bidirectional if the switch from increasing to decreasing occurs at the midpoint of .
Lemma 6.3.
Let be injective. For any integer , the mapping torus has a strictly bidirectional annulus of length if and only if is not empty.
Proof.
If there is a strictly bidirectional annulus in of length , then we may assume the midpoint of is after replacing with a translate. Then the stabilizer of contains and so it is nontrivial. Since the stabilizer of is the stabilizer of its (ordered) endpoints, the conjugacy class of the stabilizer is a component of .
If is not empty, then some path in of length with midpoint at has a nontrivial stabilizer. Choose a nontrivial element in this stabilizer and is a strictly bidirectional annulus in of length . ∎
Let be injective and be a topological representative for . If has no invariant cyclic subgroup system with index , then there is an integer for which is empty (Proposition 5.5) and all strictly bidirectional annuli in are shorter than (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 and integer , we say a graph map is (based-) -hyperbolic if all (based) loops (with the basepoint mapped to a vertex) satisfy the inequality
where 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-) -hyperbolic for some , then it is (based-) -hyperbolic for all and . So the constants can be omitted and when we do need them, we can assume is any preferred integer. As defined, hyperbolicity is a property of the homotopy class . Meanwhile, based-hyperbolicity is a property of the map .
A graph map is atoroidal if it is -injective and there is no nontrivial loop in and integer such that . This again is a property of the homotopy class . 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 -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 is atoroidal and -hyperbolic, then it is based--hyperbolic.
To avoid context-ambiguity in the proof, we use for lengths of free homotopy classes of loops and 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 is atoroidal and -hyperbolic for some integer . Set to be the maximum length of rel. basepoint over all embedded based loops in for .
Suppose for some immersed based loop and pick an embedded based loop with same basepoint as so that the concatenation is an immersed loop, i.e., . As the graph map is -hyperbolic, we get
For a concatenation of based loops , we get .
Case 1. If , then
Case 2. If , then
Combining both cases: . If for an immersed based loop in and , then by induction
For any bound , there are only finitely many immersed based loops in with . Since is atoroidal, there is an integer such that for every based loop with and we conclude that is based--hyperbolic . ∎
When is a subpath of an immersed loop and , then is the subpath of that survives in and is the length of . Bounded cancellation implies . 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 -surjective.
Lemma 6.5.
Let be a graph map and be an -invariant union of roses such that the restriction is -hyperbolic. For some constant , if is a subpath (edge-path) of some immersed loop in and , then
The number is the critical length of the triple .
Proof.
Let be a graph map and be an -invariant union of roses such that the restriction is -hyperbolic. Set to be the maximum length of rel. basepoint over all petals in and . Choose where is the cancellation constant for . Recall the triangle inequality: for any path decomposition of a loop . A remark on the context: paths are tightened rel. endpoints but the loop is tightened by free homotopy.
Given a subpath of some immersed loop in , pick a petal in such that is an immersed loop in . As the restriction to is -hyperbolic, we get
If , then
Similarly, if , then
Since , we have the desired implication:
An (outer class of an) injective endomorphism is atoroidal if it has no invariant cyclic subgroup system with index . If is a topological representative for , then is atoroidal if and only if 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 is injective and atoroidal, then has a -hyperbolic topological representative for some integer .
Proof.
Suppose is an injective and atoroidal endomorphism. If is surjective, then the proposition is precisely Brinkmann’s result. So we assume is not surjective. By Theorem 4.5, there is an expanding -relative immersion for , where is the -elliptic free factor system. Fix some -marked roses and set to be the -vertex blow-up of the graph of groups , i.e., is a marked graph with subgraphs corresponding to the free factor system . The roses form the lower stratum of 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 of . By Brinkmann’s theorem, the restriction of 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 -fixed free factor system is a subset of and there is an integer such that is carried by (Proposition 2.4). So we may find a topological representative for whose restriction to the periodic roses , denoted by , is a homotopy equivalence. As is atoroidal, the restriction is -hyperbolic for some (Brinkmann’s theorem).
If is an immersed loop in , then is a loop in the periodic roses for all . Since the restriction is -hyperbolic and is a loop in for any loop in and , we get the inequality
Choose an integer so that , where is the Lipschitz constant for . Suppose is a loop in .
If , then by induction
If , then by induction and Lipschitz property
and
Therefore, the lower stratum map is -hyperbolic with .
Let be a topological representative for that extends to the top stratum and induces the expanding -relative immersion upon collapsing the lower stratum in the universal cover . For an arbitrary immersed loop in , define ( resp.) to be the collection of maximal subpaths of in the top (lower resp.) stratum. For all , define ( resp.) to be the collection of paths where is some path in ( resp.). That induces an immersion upon collapsing the lower stratum implies that the top stratum is persistent: if is an immersed loop in , then survives in .
As the relative immersion is expanding, there is an integer , such that has length for all edges in ; and as induces , for any immersed loop in and path in , we get . We may replace and with a common multiple and assume . A similar inequality holds in the lower stratum. By the -hyperbolicity of and Lemma 6.5, there is a critical length such that for any immersed loop in and path in ,
Set to be the maximal length amongst all paths in for all top stratum edges of . For any integer , we distinguish two cases:
Case 1. If , then
Additionally, if , then .
Case 2. Suppose . Set to be the number of paths in . Then is also the number of paths in and . By the pigeonhole principle, some path in satisfies . As , we have . Set . By definition of and persistence of , there must be a path in such that is a subpath of , , and
If , then and .
If , then . By inducting on the same argument used at the start of the case, there must be a path in such that is a subpath of and . In either case, we get . Define to be the set of paths in with . Altogether, we have shown:
The following computation is lifted from Brinkmann [8, Proof of Proposition 7.1]. Set to be the total length of paths in , to be the total length of the remaining paths in , and . We now find a positive lower bound of that is independent of and . We assumed and so and we can focus on the factor . Recall , so . Since each path in but not in satisfies and there are at most of them, . Combining the last two inequalities gives the bound
Altogether, .
Additionally, if , then .
Choose so that ; the two exhaustive cases above imply is -hyperbolic. ∎
All the heavy lifting is done and we can prove our main theorem
Theorem 6.7.
Let be an injective endomorphism. Then the following statements are equivalent:
-
1.
is word-hyperbolic;
-
2.
contains no subgroups with ;
-
3.
has no invariant cyclic subgroup system with index ;
-
4.
all/some topological representatives of 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.
: subgroups are well-known obstructions to word-hyperbolicity.
: If has an invariant cyclic subgroup system with index , then there is a subgroup of isomorphic to a quotient of ; use normal forms to show the subgroup is in fact isomorphic to (See [17, Lemma 2.3] for details).
: Suppose has no invariant cyclic subgroup system with index . Then, by Proposition 6.6, has a -hyperbolic topological representative defined on a marked graph . That is, for (any conjugacy class) of , we have
For any marked graph , choose difference of markings and , i.e., graph maps such that and . Then satisfies for all , where are the respective Lipschitz constants. Fix an integer such that , then for every ,
and is -hyperbolic, i.e., all topological representatives of on are hyperbolic. Since is atoroidal, all topological representatives of on are based-hyperbolic (Lemma 6.4). As was arbitrary, all topological representatives of are based-hyperbolic.
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 be a (nontrivial) free factor and be an injective homomorphism. The HNN extension of over is given by:
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 and in 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 in has one forward iterate if it is contains an element . By malnormality of , the conjugacy class of depends only on the class . We set the forward iterate of such a class to be . Inductively, has forward iterates if it has forward iterates and the iterate has one forward iterate; in this case, set .
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 . The proof will use a simpler variation of descent.
Proposition 7.1.
If is a free factor and is injective, then there is a unique maximal -invariant free factor system of . Precisely, carries every conjugacy class with forward iterates and every -invariant subgroup system.
Proof.
Let be a free factor, be injective, and be the length of the longest chain of free factor systems in . Any conjugacy class in a -invariant free factor system of has infinitely many forward iterates. So if only the trivial conjugacy class has forward iterates, then the trivial system is the unique -invariant free factor system. Thus, we may assume some fixed nontrivial conjugacy class has forward iterates. Since is a free factor and is injective, we can iterate on the poset of free factor systems of !
Claim (Descent).
For some , let be the chain of nontrivial free factor systems of such that and carries all -invariant subgroup systems of . If is not -invariant, then is a nontrivial free factor system of that carries all -invariant subgroup systems.
Starting with , the descent will eventually stop at some with a nontrivial free factor system that carries all conjugacy classes with forward iterates and all -invariant subgroup systems of . Since descent stopped, must be -invariant and hence the unique maximal system amongst all -invariant free factor systems of . ∎
Proof of descent.
Let and be the chain of nontrivial free factor systems such that and carries all -invariant subgroup systems of . Nontriviality of implies . Now suppose that is not -invariant.
The existence of a nontrivial conjugacy class with forward iterates implies is a nontrivial free factor system of . If , then clearly . If , then by assumption; hence, and, equivalently, . If , then is carried by . But is not -invariant, thus . Let be any -invariant subgroup system in . By assumption, carries and ; therefore, carries . ∎
The unique -invariant free factor system of given by Proposition 7.1, denoted by , is the canonical -invariant free factor system and it captures the long-term dynamics of . The system is proper (in ) if and only if is not -invariant; it is trivial exactly when only the trivial conjugacy class has 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 has an invariant cyclic subgroup system with index if a restriction has an invariant cyclic subgroup system with index . In particular, is atoroidal if is atoroidal.
For , we define the iterated pullbacks of to be the iterated pullbacks of the restriction . Similarly define for . 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 with endomorphisms of .
Proposition 7.2.
If is a free factor, is injective, and is not empty for all , then has an invariant cyclic subgroup system with index .
Remark.
By the same token, the results in this section do not need to be “connected.” So there is a natural generalization of these results that easily follows if we replace the free factor with a free factor system of .
For the rest of the section, we will extend the results of Section 6 to the injective homomorphism . Fix a marked graph such that the free factor systems and correspond to nested core subgraphs respectively, i.e., and have markings and respectively such that the inclusion maps and satisfy and . Assume is not trivial so that is not degenerate (finite set of points). A topological representative for will be a graph map with no pretrivial edges such that . Thus, the invariant restriction is a topological representative for the restriction .
The following is the analogue of the mapping torus. Let be a topological representative for . We define the classifying space to be where we identify and for all . The edge-space of will be the cross-section represented by . By construction, . Topological annuli in and algebraic annuli in 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 be a free factor and be injective. For any integer , the HNN extension has a strictly bidirectional annulus of length if and only if is not empty.
The definition of (based-) hyperbolicity extends naturally using the invariant restriction. For and , a topological representative for is (based-) -hyperbolic if is (based-) -hyperbolic and is atoroidal if is atoroidal. When is trivial, any topological representative of 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 be a free factor. If is injective and atoroidal, then has a -hyperbolic topological representative for some integer .
We are now ready to state and almost prove the main result of the section.
Theorem 7.5.
Let be a free factor and be an injective homomorphism. Then the following are equivalent:
-
1.
is word-hyperbolic;
-
2.
contains no subgroups with ;
-
3.
has no invariant cyclic subgroup system with index ;
-
4.
all/some topological representative of 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 .
Remark.
Every HNN extension of the form where is a finite rank free group and is a free factor can be written as a mapping torus of an injective endomorphism where is a free group with possibly infinite rank. Conversely, every finitely generated subgroup of a mapping torus where has possibly infinite rank can be written as an HNN extension where has finite rank and is a free factor [9]. This argument is Feighn-Handel’s result that and are coherent. As a corollary, we get the following amusing statement:
Corollary 7.6.
Let be an injective endomorphism of an infinite rank free group. is locally word-hyperbolic if and only if it contains no subgroups with .
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 be a free factor and be injective. If a topological representative of is based-hyperbolic and all strictly bidirectional annuli in its classifying space are shorter than some integer, then is word-hyperbolic.
Roughly speaking, the extension follows from the fact that annuli of longer than are annuli of the mapping torus of 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 be a topological annulus in of length . The girth of is the combinatorial length of the middle ring in the edge-space. Let be a real number. The annulus is -hyperbolic if
For integers between and (inclusive), define by projecting the trace to the vertex-space as follows:
We say is -thin if for all , where is the length after tightening the path rel. endpoints. We say satisfies the annuli flaring condition if there are , and a function such that any -thin annulus of length with girth at least is -hyperbolic. Bestvina-Feighn’s combination theorem [3] states that is word-hyperbolic if satisfies the annuli flaring condition.
Sketch proof of Theorem 7.7.
Suppose is a free factor, is injective, and is the canonical -invariant free factor system of . Let be a based-hyperbolic topological representative for and be some Lipschitz constant for . If is trivial, then all unidirectional annuli in are shorter than . So vacuously satisfies the annuli flaring condition for some by the annulus dichotomy (malnormality of ). Thus, we may assume is not trivial. The assumptions and tool introduced in the next two paragraphs are needed in case is not -invariant.
Let be the chain where and generally for . By invariance of , we can extend the chain by setting for . For , carries by definition. We assume has a filtration of nonempty core subgraphs where is a wedge between and a rose and each is a -marked graph. If , we will also assume the topological representative is of the form . Note that since was -invariant to begin with, changes can made to the given representative to allow for these additional assumptions without affecting the invariant restriction . So is still based-hyperbolic after the changes.
Using this filtration, we introduce a tool needed to iterate based loops not in : for any integer and any immersed based loop that is freely homotopic into , there is a “closest point” projection (with respect to Hausdorff distance of lifts to the universal cover) of to an immersed based loop in denoted by freely homotopic to such that is a based loop in . Recall that based loops must send basepoints to vertices. Since we assumed the “complement” of are roses, the projections are immersed based loops in . So for . Implicit in this is the fact that, for all , any conjugacy class carried by but not is mapped by to a conjugacy class carried by but not . We are ready to start proving the theorem.
Suppose is based--hyperbolic for some and strictly bidirectional annuli in are shorter than . Fix such that . Now set and
Suppose is an arbitrary -thin annulus of length with girth . We need to show that is -hyperbolic. Without loss of generality, assume the truncation is unidirectional and increasing.
For , consider the rings of in the edge-space . This can be viewed as an immersed based loop by forgetting the -factor. Since and (free homotopy in ) is not trivial, the ring is freely homotopic (in ) into . Since is -invariant and for all , the rings are freely homotopic into . In particular, these rings all have projections to subgraphs for all .
For the first step, project to so that the image is a based loop freely homotopic to in . The -thin assumption (and “closest point” projection) means: . Similarly, we project to so that is a based loop freely homotopic to in . Then by -thinness and, by the -Lipschitz property of ,
By induction, the -invariance of , and the fact , we get this inequality:
Since is based--hyperbolic, we know that
There are three cases to consider of which we will only prove one.
Case 1: Suppose .
Then by induction on based--hyperbolicity, we get:
So is -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 to the rings of since may not be -invariant. This is why we introduced the projections .
There are two more cases remaining.
Case 2: Suppose and is unidirectional.
Case 3: Suppose and 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 satisfies the annuli flaring condition. By the combination theorem, 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 is a vertex group of a cyclic splitting of and is injective. Is word-hyperbolic if it contains no subgroups for ? Can it have subgroups with ?
A cyclic splitting of is an edge of groups decomposition of 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 -iteration need not be as well-behaved. Furthermore, the vertex groups are not always malnormal which means annuli in could exhibit more complicated behavior.
For the remaining problems, assume is injective.
Problem 2 ([17, Problem 6.4]).
What kind of isoperimetric functions can 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 is hyperbolic relative to a canonical finite collection of free-by-cyclic groups when it has no subgroups for . Furthermore, when is a free factor and is injective, then is hyperbolic relative to a canonical finite collection of ascending HNN extensions. This would imply that has a quadratic isoperimetric function when it has no subgroups for . We intend to complete this direction in future work by employing a combination theorem for relatively hyperbolic groups [12, 22].
When is word-hyperbolic, Mahan Mj proved there is a continuous extension to the Gromov boundary of the inclusion map [21]; this map is known as the Cannon-Thurston map.
Problem 3 ([17, Problem 6.7]).
If 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 be the extension of to the boundary. Can we classify injective endomorphisms by the dynamics of boundary extensions ?
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 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 be an injective endomorphism, be -invariant free factor systems, and be a -forest. We allow forests to have bivalent vertices. Recall, an -relative weak representative for the restriction is a -equivariant graph map . An -relative representative is an -relative weak representative with no pretrivial edges and whose underlying forest 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 -relative weak representative , we get the transition matrix . An -relative representative is irreducible if the matrix is irreducible and, in this case, the stretch factor of , denoted by , is the Perron-Frobenius eigenvalue of . An -relative train track for is an -relative representative for that additionally satisfies the property: the edge-paths are immersed for all edges in and integers . We have set the stage for the Bestvina-Handel’s result.
Theorem A.1 ([6, Theorem 1.7]).
Let be irreducible relative to and be an irreducible -relative representative for . If has minimal stretch factor, then it is an irreducible -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 -relative representative affect and invoking minimality of 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 -forests. The proofs of these moves/lemmas are omitted since they are the same as the proofs in [6].
Remark.
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 is an irreducible -relative weak representative for and is induced by a subdivision, then is an irreducible -relative weak representative and .
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 is an irreducible -relative weak representative for and is an irreducible -relative weak representative induced by a bivalent homotopy followed by collapse of a maximal invariant subforest with bounded components, then .
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 is an irreducible -relative weak representative for and is induced by a fold. If is an -relative weak representative, then it is irreducible and . Otherwise, if is an irreducible -relative weak representative induced by a homotopy of that makes the final map locally injective on the interior of edges, followed by collapse of a maximal invariant subforest with bounded components, then .
Sketch proof of Theorem A.1.
Let be irreducible relative to and be an irreducible -relative representative. If , then is a simplicial embedding with minimal stretch factor and we are done. So we may assume .
Suppose for the contrapositive that is not an -relative train track, then the edge-path is not immersed for some edge in and integer . Let be the smallest such integer and assume is an interior point of an edge at which fails to be locally injective. We appropriately subdivide so that a neighborhood of and its iterates satisfy nice properties: 1) is an interval whose boundary consists of distinct vertices; 2) is locally injective on for ; 3) folds at to an edge; and 4) for . We can then iteratively fold , …, , and . By minimality of , all the folds except the last one induce an irreducible -relative weak representative. By the first case of Lemma A.4, this irreducible -relative weak representative has the same stretch factor as . By construction, the last fold induces a map that fails to be an -relative weak representative as it fails to be locally injective at . We can apply a tightening homotopy on to make it locally injective at , then collapse a maximal invariant subforest with bounded components to get , a minimal -relative weak representative for . By Lemma 3.6, the map is irreducible. By the second case of Lemma A.4, the stretch factor is strictly smaller: . We then sequentially apply bivalent homotopies and collapse maximal invariant subforests with bounded components until we get an irreducible -relative representative . The stretch factor satisfies by Lemma A.3. So 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 . 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(). 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 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 . PhD thesis, University of Illinois at Urbana-Champaign, 2011.
- [28] Jean-Pierre Serre. Arbres, amalgames, . 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 -manifolds. I. Deformation of acylindrical manifolds. Ann. of Math. (2), 124(2):203–246, 1986.