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

Lipschitz metric isometries between Outer Spaces of virtually free groups

Rylee Alanza Lyman
Abstract

Dowdall and Taylor observed that given a finite-index subgroup of a free group, taking covers induces an embedding from the Outer Space of the free group to the Outer Space of the subgroup, that this embedding is an isometry with respect to the (asymmetric) Lipschitz metric, and that the embedding sends folding paths to folding paths. The purpose of this note is to extend this result to virtually free groups. We further extend a result Francaviglia and Martino, proving the existence of “candidates” for the Lipschitz distance between points in the Outer Space of the virtually free group. Additionally we identify a deformation retraction of the spine of the Outer Space for the virtually free group with the space considered by Krstić and Vogtmann.

Let FF be a finitely generated virtually free group. Throughout this paper we will always assume FF is virtually non-abelian free, but will just write virtually free. Consider the deformation space 𝒯=𝒯(F)\mathscr{T}=\mathscr{T}(F) in the sense of Forester and Guirardel–Levitt [GL07a, For02] whose points are actions of FF on simplicial \mathbb{R}-trees with finite stabilizers. This deformation space is canonical in the sense that all of Out(F)\operatorname{Out}(F) acts on it, and we call it the unprojectivized Outer Space of FF. Projectivizing (by for example requiring the volume of the quotient graph of groups to be 11) yields the Outer Space 𝒫𝒯=𝒫𝒯(F)\mathscr{PT}=\mathscr{PT}(F).

When F=FnF=F_{n}, the space 𝒫𝒯(Fn)\mathscr{PT}(F_{n}) is the Culler–Vogtmann Outer Space CVn\operatorname{CV}_{n}. When FF is a free product of finite and cyclic groups, the space 𝒫𝒯\mathscr{PT} is not quite Guirardel–Levitt’s Outer Space 𝒪(F)\mathscr{O}(F) for FF [GL07b]: the difference is that trees in 𝒪(F)\mathscr{O}(F) have trivial edge stabilizers, while there is no such requirement for trees in 𝒫𝒯(F)\mathscr{PT}(F).

In [DT17, Section 5.1], Dowdall–Taylor show that if HH is a finite-index subgroup of FnF_{n}, then thinking of each FnF_{n}-tree T𝒫𝒯(Fn)T\in\mathscr{PT}(F_{n}) as an HH-tree yields an embedding i:𝒫𝒯(Fn)𝒫𝒯(H)i\colon\mathscr{PT}(F_{n})\to\mathscr{PT}(H). They show that ii is an isometry with respect to the (asymmetric) Lipschitz metric on 𝒫𝒯(Fn)\mathscr{PT}(F_{n}) and 𝒫𝒯(H)\mathscr{PT}(H), and that ii sends a particular class of (directed) geodesics in 𝒫𝒯(Fn)\mathscr{PT}(F_{n}) called folding paths to folding paths in 𝒫𝒯(H)\mathscr{PT}(H). The purpose of this note is to extend these results to virtually free groups.

Theorem A.

Let HFH\leq F be a finite-index subgroup of the virtually free group FF. The induced map i:𝒫𝒯(F)𝒫𝒯(H)i\colon\mathscr{PT}(F)\to\mathscr{PT}(H) is an isometry with respect to the Lipschitz metric. Moreover, ii maps folding paths in 𝒫𝒯(F)\mathscr{PT}(F) to folding paths in 𝒫𝒯(H)\mathscr{PT}(H).

The Lipschitz metric for 𝒫𝒯(F)\mathscr{PT}(F) when FF is not free was first defined and studied by Meinert [Mei15]. He identifies 𝒫𝒯\mathscr{PT} with the subset of 𝒯\mathscr{T} consisting of those trees Γ\Gamma such that the sum of the lengths of edges of the quotient graph F\ΓF\backslash\Gamma is 11. This notion of volume is not so well-behaved for general 𝒫𝒯\mathscr{PT}: the Lipschitz metric is only a pseudometric, and taking covers is not multiplicative with respect to volume. We introduce a new notion of volume which rectifies both of these issues: we divide the length of an edge by the order of its stabilizer and then add.

Meinert proves [Mei15, Theorem 4.14] that there are witnesses for the Lipschitz pseudometric distance between any two trees in 𝒯\mathscr{T}: hyperbolic group elements gFg\in F such that

d(Γ,Γ)=loggΓgΓ,d(\Gamma,\Gamma^{\prime})=\log\frac{\|g\|_{\Gamma^{\prime}}}{\|g\|_{\Gamma}},

where gΓ\|g\|_{\Gamma} denotes the hyperbolic translation length of gg on Γ\Gamma. He further shows that these witnesses may be taken to satisfy a weak finiteness condition. By contrast, Francaviglia–Martino [FM11, FM15] show that in the case of FF a free product, the projection of gg to the quotient graph of groups has certain nice properties. To reduce the computational complexity of computing the Lipschitz distance between trees in 𝒫𝒯\mathscr{PT}, we extend Francaviglia–Martino’s result to FF virtually free.

Theorem B (cf. Theorem 9.10 of [FM15]).

Let Γ\Gamma and Γ\Gamma^{\prime} be trees in 𝒯\mathscr{T}. There is a witness gFg\in F, the projection of whose axis into the quotient graph of groups F\\ΓF\backslash\mkern-5.0mu\backslash\Gamma has one of the following forms.

  1. 1.

    An embedded simple loop.

  2. 2.

    An embedded figure-eight (a wedge of two circles).

  3. 3.

    An embedded barbell (two simple loops joined by a segment).

  4. 4.

    An embedded singly degenerate barbell (a non-free vertex and a simple loop joined by a segment).

  5. 5.

    An embedded doubly degenerate barbell (two non-free vertices joined by a segment).

There are finitely many graph-of-groups edge paths of the given forms, which we call candidates, and the set of candidates is independent of the choice of Γ\Gamma^{\prime}.

In [BFH20] and [KV93], given a finite-rank non-abelian free group FnF_{n} and a homomorphism α:GOut(Fn)\alpha\colon G\to\operatorname{Out}(F_{n}) with finite domain, a certain subcomplex of the spine of Outer Space they term LαL_{\alpha} is considered. In those papers, vertices of LαL_{\alpha} are marked graphs equipped with a GG-action via α\alpha, and the contractibility of these simplicial complexes is used to prove that centralizers of finite subgroups of Out(Fn)\operatorname{Out}(F_{n}) are virtually of type F.

Let GOut(Fn)G\leq\operatorname{Out}(F_{n}) be a finite subgroup, and let EE be the full preimage of GG in Aut(Fn)\operatorname{Aut}(F_{n}). Then EE is a virtually free group. Let Aut0(E)\operatorname{Aut}^{0}(E) be the subgroup of Aut(E)\operatorname{Aut}(E) that leaves FnInn(Fn)F_{n}\cong\operatorname{Inn}(F_{n}) invariant and let Out0(E)\operatorname{Out}^{0}(E) be its image in Out(E)\operatorname{Out}(E). We define a deformation retraction L=L(E)L=L(E) of the spine of the outer space of EE. Our final result is the following theorem; see 3.1 for a precise statement.

Theorem C.

Given GOut(Fn)G\leq\operatorname{Out}(F_{n}) and EE as above, the isometry i:𝒫𝒯(E)𝒫𝒯(Fn)i\colon\mathscr{PT}(E)\to\mathscr{PT}(F_{n}) induces a simplicial isomorphism L(E)LαL(E)\to L_{\alpha}, where α:GOut(Fn)\alpha\colon G\to\operatorname{Out}(F_{n}) is the natural inclusion. The normalizer N(G)N(G) of GG in Out(Fn)\operatorname{Out}(F_{n}) leaves LαL_{\alpha} invariant. We have the following short exact sequence

1{1}G{G}N(G){N(G)}Out0(E){\operatorname{Out}^{0}(E)}1.{1.}

In particular, if FnF_{n} is characteristic in EE, then Out0(E)=Out(E)\operatorname{Out}^{0}(E)=\operatorname{Out}(E).

Here is the organization of this note. We begin by recalling work of Meinert in Section 1 culminating in the proof of A. Section 2 is dedicated to the proof of B, and C is proved in Section 3.

1 Outer Space and the Lipschitz metric

Unprojectivized Outer Space.

A point of 𝒯=𝒯(F)\mathscr{T}=\mathscr{T}(F) for a finitely generated virtually free group FF is the equivalence class of a minimal isometric action of FF on a simplicial \mathbb{R}-tree Γ\Gamma such that all stabilizers are finite. The equivalence relation is FF-equivariant isometry of trees. We will, as is standard, always speak of actual trees, rather than isometry classes of trees.

The group Aut(F)\operatorname{Aut}(F) acts on 𝒯\mathscr{T} on the right by twisting the action, i.e. the tree Γ.Φ\Gamma.\Phi is the tree Γ\Gamma equipped with the FF-action defined by first applying Φ\Phi and then acting. Inner automorphisms of FF correspond to FF-equivariant isometries, so we get an action of Out(F)\operatorname{Out}(F) on 𝒯\mathscr{T}.

Quotient graph of groups.

Associated to such a tree Γ\Gamma, there is a quotient graph of groups 𝒢=F\\Γ\mathcal{G}=F\backslash\mkern-5.0mu\backslash\Gamma, which we now describe, following [Bas93, 3.2]. By subdividing at midpoints of edges of Γ\Gamma which are inverted by the FF-action, we may assume that the action of FF on Γ\Gamma is without inversions in edges, so the quotient graph G=F\ΓG=F\backslash\Gamma naturally inherits a graph structure. Write p:ΓGp\colon\Gamma\to G for the quotient map. Choose connected subgraphs TSΓT\subset S\subset\Gamma such that p|S:SGp|_{S}\colon S\to G is a bijection on edges (so SS is a fundamental domain for the action) and p|T:TGp|_{T}\colon T\to G is a bijection on vertices (so TT projects to a spanning tree). For ee an edge and vv a vertex of GG, write e~\tilde{e} and v~\tilde{v} for the unique preimage of ee and vv in SS and TT respectively. For an oriented edge ee of a graph, write τ(e)\tau(e) for its terminal vertex. For each oriented edge ee of GG, let geFg_{e}\in F be an element such that ge.τ(e~)=τ(e)~g_{e}.\tau(\tilde{e})=\widetilde{\tau(e)}. If τ(e~)T\tau(\tilde{e})\in T, choose ge=1g_{e}=1. The groups 𝒢v\mathcal{G}_{v} and 𝒢e\mathcal{G}_{e} for vv a vertex and ee an edge of GG are the stabilizers in FF of v~\tilde{v} and e~\tilde{e} respectively. For an oriented edge ee of GG, the injective homomorphism 𝒢e𝒢τ(e)\mathcal{G}_{e}\to\mathcal{G}_{\tau(e)} is the restriction of the map xgexge1x\mapsto g_{e}xg_{e}^{-1} to 𝒢e\mathcal{G}_{e}.

The graph of groups 𝒢=F\\Γ\mathcal{G}=F\backslash\mkern-5.0mu\backslash\Gamma is equipped with a metric defined so that p:Γ𝒢p\colon\Gamma\to\mathcal{G} restricts to an isometry on each edge and a marking, an isomorphism Fπ1(𝒢)F\cong\pi_{1}(\mathcal{G}) well-defined up to inner automorphism.

Volume, projectivized Outer Space.

Given a tree Γ𝒯\Gamma\in\mathscr{T}, we define the volume of Γ\Gamma to be

vol(Γ)=vol(F\\Γ)=e(e)|𝒢e|,\operatorname{vol}(\Gamma)=\operatorname{vol}(F\backslash\mkern-5.0mu\backslash\Gamma)=\sum_{e}\frac{\ell(e)}{|\mathcal{G}_{e}|},

where (e)\ell(e) denotes the length of the edge ee, and where the sum is taken over the edges ee of F\\ΓF\backslash\mkern-5.0mu\backslash\Gamma (equivalently the edges of some and hence any fundamental domain SS). It is clear that if Γ\Gamma^{\prime} has hyperbolic length function Γ\|\cdot\|_{\Gamma^{\prime}} (see e.g. [CM87]) equal to λΓ\lambda\|\cdot\|_{\Gamma} for some λ>0\lambda>0, then vol(Γ)=λvol(Γ)\operatorname{vol}(\Gamma^{\prime})=\lambda\operatorname{vol}(\Gamma), so we may and will identify 𝒫𝒯(F)\mathscr{PT}(F) with the subset of 𝒯(F)\mathscr{T}(F) comprising the trees with volume 11. Let us remark for experts that because 𝒯(F)\mathcal{T}(F) is locally compact (and its spine is locally finite), this identification is as innocuous as it is for CVn\operatorname{CV}_{n}.

Stretch factors.

We begin by reporting on work of Meinert [Mei15]. Given two trees Γ\Gamma and Γ\Gamma^{\prime} in 𝒯\mathscr{T}, there is an FF-equivariant map f:ΓΓf\colon\Gamma\to\Gamma^{\prime}, which we may take to be Lipschitz continuous, and we write Lip(f)\operatorname{Lip}(f) for its Lipschitz constant. As in [Mei15, p. 998], it suffices to consider only those FF-equivariant maps which are piecewise linear, which is to say for each edge e~\tilde{e} of Γ\Gamma, the map ff is either constant on e~\tilde{e} or linear with constant slope on e~\tilde{e}. We define

Lip(Γ,Γ)=inf{Lip(f):f:ΓΓ is F-equviariant}.\operatorname{Lip}(\Gamma,\Gamma^{\prime})=\inf\{\operatorname{Lip}(f):f\colon\Gamma\to\Gamma^{\prime}\text{ is $F$-equviariant}\}.

Meinert proves [Mei15, Theorem 4.6] using nonprincipal ultrafilters that there exists an FF-equivariant piecewise linear Lipschitz map f:ΓΓf\colon\Gamma\to\Gamma^{\prime} such that Lip(f)=Lip(Γ,Γ)\operatorname{Lip}(f)=\operatorname{Lip}(\Gamma,\Gamma^{\prime}). We remark that since our trees are locally finite and the quotient graphs of groups are compact, it should be possible to give a proof using the Arzela–Ascoli theorem.

Meinert further proves [Mei15, Theorem 4.14] that

Lip(Γ,Γ)=supggΓgΓ,\operatorname{Lip}(\Gamma,\Gamma^{\prime})=\sup_{g}\frac{\|g\|_{\Gamma^{\prime}}}{\|g\|_{\Gamma}},

where the supremum is taken over all hyperbolic elements gFg\in F, and that the supremum is realized.

Train track structures.

Given a tree Γ𝒯\Gamma\in\mathscr{T} and a point x~Γ\tilde{x}\in\Gamma, a direction at x~\tilde{x} is a germ of geodesics beginning at x~\tilde{x}. Equivalently, a direction is a component of the complement Γ{x~}\Gamma\setminus\{\tilde{x}\}. The set of directions at a point x~\tilde{x} is denoted Dx~ΓD_{\tilde{x}}\Gamma. The action of FF on Γ\Gamma induces a permutation of the set of direction in Γ\Gamma. A train track structure on Γ\Gamma is an FF-equivariant collection of equivalence relations on the set of directions of Γ\Gamma with the property that there are at least two equivalence classes at each point. That is to say, if δ1δ2\delta_{1}\sim\delta_{2} as directions in Dx~ΓD_{\tilde{x}}\Gamma, then for all gFg\in F, we have g.δ1g.δ2g.\delta_{1}\sim g.\delta_{2} as directions in Dg.x~ΓD_{g.\tilde{x}}\Gamma. A train track structure partitions the directions at Dx~ΓD_{\tilde{x}}\Gamma into gates, and an FF-equivariant collection of equivalence relations on the set of directions of Γ\Gamma is a train track structure if and only if there are at least two gates at every point. A turn is a pair of directions based at a common point. It is illegal if the directions belong to the same gate, and legal otherwise. A geodesic γ~\tilde{\gamma} in Γ\Gamma through x~\tilde{x} is said to cross the turn at x~\tilde{x} determined by the components of the complement Γ{x~}\Gamma\setminus\{\tilde{x}\} containing γ~\tilde{\gamma}. A geodesic γ~\tilde{\gamma} is legal if it makes only legal turns and illegal otherwise.

Tension forest.

Let f:ΓΓf\colon\Gamma\to\Gamma^{\prime} be a piecewise linear FF-equivariant map between trees in 𝒯\mathscr{T}. The union of all closed edges of Γ\Gamma on which ff attains its maximal slope (which must be nonzero) is Δ(f)\Delta(f), the tension forest of ff. It is FF-invariant. The map ff sends directions pointing into Δ=Δ(f)\Delta=\Delta(f) to directions of Γ\Gamma^{\prime}, and thus defines an equivalence relation on turns in Δ\Delta by saying two directions of Δ\Delta are equivalent if they are mapped to the same direction by ff.

Optimal maps.

A piecewise linear FF-equivariant map f:ΓΓf\colon\Gamma\to\Gamma^{\prime} is an optimal map if Lip(f)=Lip(Γ,Γ)\operatorname{Lip}(f)=\operatorname{Lip}(\Gamma,\Gamma^{\prime}) and its tension forest Δ(f)\Delta(f) has an induced train track structure. Meinert proves [Mei15, Proposition 4.12] that every FF-equivariant piecewise linear map realizing Lip(Γ,Γ)\operatorname{Lip}(\Gamma,\Gamma^{\prime}) is homotopic to an optimal map f:ΓΓf^{\prime}\colon\Gamma\to\Gamma^{\prime} with Δ(f)Δ(f)\Delta(f^{\prime})\subset\Delta(f).

Meinert proves [Mei15, Lemma 4.15] that for an optimal map f:ΓΓf\colon\Gamma\to\Gamma^{\prime}, Lip(f)=gΓgΓ\operatorname{Lip}(f)=\frac{\|g\|_{\Gamma^{\prime}}}{\|g\|_{\Gamma}} if and only if the axis of gg in Γ\Gamma is contained in the tension forest Δ(f)\Delta(f) and is legal with respect to the train track structure determined by ff.

Lipschitz metric.

We define the Lipschitz metric on 𝒫𝒯\mathscr{PT} as

d(Γ,Γ)=logLip(Γ,Γ).d(\Gamma,\Gamma^{\prime})=\log\operatorname{Lip}(\Gamma,\Gamma^{\prime}).
Proposition 1.1.

For all Γ\Gamma, Γ\Gamma^{\prime} and Γ′′𝒫𝒯\Gamma^{\prime\prime}\in\mathscr{PT}, we have

  1. 1.

    d(Γ,Γ)0d(\Gamma,\Gamma^{\prime})\geq 0 and d(Γ,Γ)=0d(\Gamma,\Gamma^{\prime})=0 implies Γ\Gamma and Γ\Gamma^{\prime} are FF-equivariantly isometric.

  2. 2.

    d(Γ,Γ′′)d(Γ,Γ)+d(Γ,Γ′′)d(\Gamma,\Gamma^{\prime\prime})\leq d(\Gamma,\Gamma^{\prime})+d(\Gamma^{\prime},\Gamma^{\prime\prime}).

Proof.

Meinert proves all of these statements as [Mei15, Proposition/Definition 4.2] with the exception of the assertion that d(Γ,Γ)=0d(\Gamma,\Gamma^{\prime})=0 implies Γ\Gamma and Γ\Gamma^{\prime} are FF-equivariantly isometric. Of course, his volume is defined slightly differently, but the arguments apply with our new notion of volume.

So suppose d(Γ,Γ)=0d(\Gamma,\Gamma^{\prime})=0 but that Γ\Gamma and Γ\Gamma^{\prime} are not equivariantly isometric, and let f:ΓΓf\colon\Gamma\to\Gamma^{\prime} be an optimal map. Because ff is FF-equivariant, there is an induced map F\\f:F\\ΓF\\ΓF\backslash\mkern-5.0mu\backslash f\colon F\backslash\mkern-5.0mu\backslash\Gamma\to F\backslash\mkern-5.0mu\backslash\Gamma^{\prime} (it is in fact a homotopy equivalence of graphs of groups as defined in [Lym22, Section 1]). Since Γ\Gamma and Γ\Gamma^{\prime} are minimal, ff is surjective, so F\\fF\backslash\mkern-5.0mu\backslash f is as well, and the argument in the proof of [FM11, Lemma 4.16] [Mei15, Proposition 4.16] shows that F\\fF\backslash\mkern-5.0mu\backslash f is an isometry. Because ff is not an FF-equivariant isometry, it must fold a pair of edges, which must be in the same edge orbit, so some 𝒢F\\f(e)\mathcal{G}^{\prime}_{F\backslash\mkern-5.0mu\backslash f(e)} is larger than 𝒢e\mathcal{G}_{e}. But this implies that vol(Γ)<1\operatorname{vol}(\Gamma^{\prime})<1, a contradiction. ∎

Folding paths.

An optimal map f:ΓΓf\colon\Gamma\to\Gamma^{\prime} between trees in 𝒯\mathscr{T} is a folding map if the tension forest satisfies Δ(f)=Γ\Delta(f)=\Gamma. Associated to a folding map, one can use a technique of Skora known as “folding all illegal turns at speed 11” to obtain a 11-parameter family of trees Γt\Gamma_{t} in 𝒯\mathscr{T} for t0t\geq 0 such that Γ0=Γ\Gamma_{0}=\Gamma and Γt=Γ\Gamma_{t}=\Gamma^{\prime} for tt large and folding maps fst:ΓsΓtf_{st}\colon\Gamma_{s}\to\Gamma_{t} for sts\leq t satisfying

f0T=f for all T large, fss=IdΓs, and frt=fstfrs for rst,f_{0T}=f\text{ for all $T$ large, }f_{ss}=\operatorname{Id}_{\Gamma_{s}},\text{ and }f_{rt}=f_{st}f_{rs}\text{ for $r\leq s\leq t$},

with frsf_{rs} and frtf_{rt} inducing the same train track structure on Γr\Gamma_{r}. See [BF14, FM11] for more details, in particular the argument of [BF14] that shows that the folding paths that we are about to define are unique, depending only on the folding map f:ΓΓf\colon\Gamma\to\Gamma^{\prime}. If we rescale each Γt\Gamma_{t} so that it belongs to 𝒫𝒯\mathscr{PT}, Meinert shows [Mei15, Theorem 4.23] that the map tΓtt\mapsto\Gamma_{t} defines a directed geodesic in the sense that for rstr\leq s\leq t, we have

d(Γr,Γs)+d(Γs,Γt)=d(Γr,Γt).d(\Gamma_{r},\Gamma_{s})+d(\Gamma_{s},\Gamma_{t})=d(\Gamma_{r},\Gamma_{t}).

With our result 1.1 in hand, we may reparametrize our directed geodesics by arc length, i.e. as a path [0,d(Γ,Γ)]𝒫𝒯[0,d(\Gamma,\Gamma^{\prime})]\to\mathscr{PT} such that Γ=Γ0\Gamma=\Gamma_{0}, Γd(Γ,Γ)=Γ\Gamma_{d(\Gamma,\Gamma^{\prime})}=\Gamma^{\prime} and for sts\leq t we have

d(Γs,Γt)=ts.d(\Gamma_{s},\Gamma_{t})=t-s.

This is a folding path.

The Outer Space of a subgroup

Fix HFH\leq F a subgroup of finite index nn. Each tree Γ𝒯(F)\Gamma\in\mathscr{T}(F) is, by restricting the action, a minimal HH-tree such that all stabilizers are finite, so we may view Γ\Gamma as a tree in 𝒯(H)\mathscr{T}(H). Since an FF-equivariant isometry is in particular HH-equivariant, we see that this map i:𝒯(F)𝒯(H)i\colon\mathscr{T}(F)\to\mathscr{T}(H) is well-defined. The natural map H\\ΓF\\ΓH\backslash\mkern-5.0mu\backslash\Gamma\to F\backslash\mkern-5.0mu\backslash\Gamma is a covering map in the sense of Bass [Bas93, 2.6]. The equation in [Bas93, Proposition 2.7] implies that

vol(H\\Γ)=nvol(F\\Γ).\operatorname{vol}(H\backslash\mkern-5.0mu\backslash\Gamma)=n\operatorname{vol}(F\backslash\mkern-5.0mu\backslash\Gamma).

Thus if Γ𝒫𝒯(F)\Gamma\in\mathscr{PT}(F), scaling edge lengths by 1n\frac{1}{n} yields a point 1nΓ𝒫𝒯(H)\frac{1}{n}\Gamma\in\mathscr{PT}(H). We are ready to prove A, which we restate here.

Theorem 1.2.

Suppose HFH\leq F has finite index nn. The map i:𝒫𝒯(F)𝒫𝒯(H)i\colon\mathscr{PT}(F)\to\mathscr{PT}(H) given by Γ1nΓ\Gamma\mapsto\frac{1}{n}\Gamma is an isometry with respect to the Lipschitz metric. Moreover, ii maps folding paths in 𝒫𝒯(F)\mathscr{PT}(F) to folding paths in 𝒫𝒯(H)\mathscr{PT}(H).

Proof.

Fix Γ\Gamma and Γ\Gamma^{\prime} in 𝒫𝒯(F)\mathscr{PT}(F). If f:ΓΓf\colon\Gamma\to\Gamma^{\prime} is an optimal map, then 1nf:1nΓ1nΓ\frac{1}{n}f\colon\frac{1}{n}\Gamma\to\frac{1}{n}\Gamma^{\prime} has Lipschitz constant Lip(f)\operatorname{Lip}(f), so

d𝒫𝒯(H)(1nΓ,1nΓ)d𝒫𝒯(F)(Γ,Γ).d_{\mathscr{PT}(H)}\left(\frac{1}{n}\Gamma,\frac{1}{n}\Gamma^{\prime}\right)\leq d_{\mathscr{PT}(F)}(\Gamma,\Gamma^{\prime}).

We have seen that

d𝒫𝒯(H)(1nΓ,1nΓ)=logsuphh1nΓh1nΓ=logsuphhΓhΓd_{\mathscr{PT}(H)}\left(\frac{1}{n}\Gamma,\frac{1}{n}\Gamma^{\prime}\right)=\log\sup_{h}\frac{\|h\|_{\frac{1}{n}\Gamma^{\prime}}}{\|h\|_{\frac{1}{n}\Gamma}}=\log\sup_{h}\frac{\|h\|_{\Gamma^{\prime}}}{\|h\|_{\Gamma}}

where the supremum is taken over the hyperbolic elements of HH. We have seen that there exists a hyperbolic g0Fg_{0}\in F realizing the supremum

supggΓgΓ,\sup_{g}\frac{\|g\|_{\Gamma^{\prime}}}{\|g\|_{\Gamma}},

where now the supremum is taken over the hyperbolic elements of FF. If we take k1k\geq 1 such that g0kHg_{0}^{k}\in H, we see that

d𝒫𝒯(H)(1nΓ,1nΓ)d𝒫𝒯(F)(Γ,Γ),d_{\mathscr{PT}(H)}\left(\frac{1}{n}\Gamma,\frac{1}{n}\Gamma^{\prime}\right)\geq d_{\mathscr{PT}(F)}(\Gamma,\Gamma^{\prime}),

so ii is an isometry. It is clear that if f:ΓΓf\colon\Gamma\to\Gamma^{\prime} is a folding map, then 1nf:1nΓ1nΓ\frac{1}{n}f\colon\frac{1}{n}\Gamma\to\frac{1}{n}\Gamma^{\prime} is a folding map, and thus it follows that the folding path induced by 1nf\frac{1}{n}f agrees with that induced by ff and then scaling by 1n\frac{1}{n}. ∎

2 Candidates

The purpose of this section is to prove B. Given trees Γ\Gamma and Γ\Gamma^{\prime} in 𝒯\mathscr{T}, let f:ΓΓf\colon\Gamma\to\Gamma^{\prime} be an optimal map. It may not be the case that ff maps vertices to vertices. We may remedy this by (FF-equivariantly) declaring the images of vertices to be vertices. This done, [Lym22, Proposition 1.2] allows us to consider the induced map of quotient graphs of groups F\\f:F\\ΓF\\ΓF\backslash\mkern-5.0mu\backslash f\colon F\backslash\mkern-5.0mu\backslash\Gamma\to F\backslash\mkern-5.0mu\backslash\Gamma^{\prime}. The tension forest Δ(f)\Delta(f) projects to a tension subgraph of groups Δ(F\\f)\Delta(F\backslash\mkern-5.0mu\backslash f) in F\\ΓF\backslash\mkern-5.0mu\backslash\Gamma. Note that FF-equivariance of ff implies that it is a subgraph of groups and not a subgraph of subgroups in the sense of Bass [Bas93]. (Put another way, being optimally stretched is a property of the edge orbit.) Since ff is optimal, Δ(F\\f)\Delta(F\backslash\mkern-5.0mu\backslash f) is a core subgraph of groups in the sense that each valence-one vertex of Δ(F\\f)\Delta(F\backslash\mkern-5.0mu\backslash f) satisfies that the edge-to-vertex group inclusion of the incident edge is not surjective.

If gg is hyperbolic in Γ\Gamma, the projection of the axis of gg to F\\ΓF\backslash\mkern-5.0mu\backslash\Gamma, after subdividing at preimages of vertices and modding out by the action of gg yields an immersion (in the sense of Bass [Bas93, 1.2]) of a subdivided circle C=CgC=C_{g} into F\\ΓF\backslash\mkern-5.0mu\backslash\Gamma. The image of CC in F\\ΓF\backslash\mkern-5.0mu\backslash\Gamma is a cyclically reduced graph-of-groups edge path

γ=g0e1g1ekgk\gamma=g_{0}e_{1}g_{1}\ldots e_{k}g_{k}

in F\\ΓF\backslash\mkern-5.0mu\backslash\Gamma. (See [Ser03, Bas93, Lym22] for more details on graphs of groups.) In what follows we will confuse gg, CgC_{g} and γ\gamma. We think of the edges of CC as being labeled by the edges eie_{i} of F\\ΓF\backslash\mkern-5.0mu\backslash\Gamma crossed by this edge path and vertices of CC as being labeled by the corresponding vertex group elements. (Note that up to cyclically re-ordering, we may assume that g0g_{0} is trivial; or put another way, one vertex is labeled gkg0g_{k}g_{0}.) If gg is a witness, in the sense that

Lip(Γ,Γ)=gΓgΓ,\operatorname{Lip}(\Gamma,\Gamma^{\prime})=\frac{\|g\|_{\Gamma^{\prime}}}{\|g\|_{\Gamma}},

then in fact, the image of CgC_{g} lies in Δ(F\\f)\Delta(F\backslash\mkern-5.0mu\backslash f), and the map CgΔ(F\\f)F\\ΓC_{g}\to\Delta(F\backslash\mkern-5.0mu\backslash f)\to F\backslash\mkern-5.0mu\backslash\Gamma^{\prime} becomes an immersion in the sense of Bass [Bas93, 1.2] after subdividing at the preimages of vertices. Thus it will sometimes instead be convenient to perform that subdivision and think of CgC_{g} as being labeled by edges and vertex group elements in F\\ΓF\backslash\mkern-5.0mu\backslash\Gamma^{\prime} instead.

This done, the main step of the proof follows nearly exactly as in Francaviglia and Martino [FM11].

Lemma 2.1 (“Sausages Lemma” cf. Lemma 3.14 of [FM11]).

Let f:ΓΓf\colon\Gamma\to\Gamma^{\prime} be an optimal map. There is a witness gg such that the immersed circle CgC_{g} is a “sausage”: the image γ\gamma of CgC_{g} is a cyclically reduced path in F\ΓF\backslash\Gamma of the form γ=γ1γ¯2\gamma=\gamma_{1}\bar{\gamma}_{2}, where each γi\gamma_{i} is a path in F\\ΓF\backslash\mkern-5.0mu\backslash\Gamma that can be parametrized by [0,1][0,1] in such a way that γ1\gamma_{1} and γ2\gamma_{2} are embedded, and there exist a finite family of disjoint closed intervals Ij[0,1]I_{j}\subset[0,1], each one possibly consisting of a single point, such that γ1(t)=γ2(s)\gamma_{1}(t)=\gamma_{2}(s) if and only if t=st=s and tt belongs to some IjI_{j}.

Proof.

We follow the proof of [FM11, Lemma 3.14]. The first claim is that we can choose gg such that the image of CgC_{g} in F\\ΓF\backslash\mkern-5.0mu\backslash\Gamma has no triple points. That is, if we can write γ=δ1δ2δ3\gamma=\delta_{1}\delta_{2}\delta_{3}, where the endpoints of each δi\delta_{i} map to the same point in F\\ΓF\backslash\mkern-5.0mu\backslash\Gamma, then we may replace γ\gamma (and thus gg) with a shorter loop. If some δi\delta_{i} maps to a cyclically reduced loop in F\\ΓF\backslash\mkern-5.0mu\backslash\Gamma^{\prime}, choose that δi\delta_{i}. If all δi\delta_{i} map to reduced but not cyclically reduced paths in F\\ΓF\backslash\mkern-5.0mu\backslash\Gamma^{\prime}, then (after shuffling vertex group elements around,) the labellings of δi\delta_{i} by their images in F\\ΓF\backslash\mkern-5.0mu\backslash\Gamma^{\prime} may be written as

δ1\displaystyle\delta_{1} =e1e¯1\displaystyle=e_{1}\ldots\bar{e}_{1}
δ2\displaystyle\delta_{2} =e2e¯2\displaystyle=e_{2}\ldots\bar{e}_{2}
δ3\displaystyle\delta_{3} =e3e¯3,\displaystyle=e_{3}\ldots\bar{e}_{3},

where, since γ\gamma immerses into F\\ΓF\backslash\mkern-5.0mu\backslash\Gamma^{\prime}, we must have e1e2e_{1}\neq e_{2}, so the loop δ1δ2\delta_{1}\delta_{2} is immersed in F\\ΓF\backslash\mkern-5.0mu\backslash\Gamma^{\prime}, and we choose it instead. This process shortens γ\gamma in F\\ΓF\backslash\mkern-5.0mu\backslash\Gamma, so eventually we obtain gg such that CgC_{g} has no triple points.

The second claim is that we can choose gg such that the image of CgC_{g} in F\\ΓF\backslash\mkern-5.0mu\backslash\Gamma has no double points which cross in the sense that we have γ=δ1δ2δ3δ4\gamma=\delta_{1}\delta_{2}\delta_{3}\delta_{4}, where the initial endpoints of δ1\delta_{1} and δ3\delta_{3} are equal and the initial endpoints of δ2\delta_{2} and δ4\delta_{4} are equal. So suppose we do have such a configuration. Now if some δiδi+1\delta_{i}\delta_{i+1} (with subscripts taken mod 44) maps to a cyclically reduced path in F\\ΓF\backslash\mkern-5.0mu\backslash\Gamma^{\prime}, then we can choose that path as our new gg. Otherwise, as before we have that (after shuffling vertex group elements around) the labellings of δi\delta_{i} by their images in F\\ΓF\backslash\mkern-5.0mu\backslash\Gamma^{\prime} may be written as

δ1δ2\displaystyle\delta_{1}\delta_{2} =e1e¯1\displaystyle=e_{1}\ldots\bar{e}_{1}
δ2δ3\displaystyle\delta_{2}\delta_{3} =e2e¯2\displaystyle=e_{2}\ldots\bar{e}_{2}
δ3δ4\displaystyle\delta_{3}\delta_{4} =e3e¯3\displaystyle=e_{3}\ldots\bar{e}_{3}
δ4δ1\displaystyle\delta_{4}\delta_{1} =e4e¯4.\displaystyle=e_{4}\ldots\bar{e}_{4}.

Therefore we conclude that

δi=eie¯i+3,\delta_{i}=e_{i}\ldots\bar{e}_{i+3},

with subscripts taken mod 44. Since γ\gamma immerses into F\\ΓF\backslash\mkern-5.0mu\backslash\Gamma^{\prime}, we have e1e3e_{1}\neq e_{3} and e2e4e_{2}\neq e_{4}, so the loop δ1δ¯3\delta_{1}\bar{\delta}_{3} is immersed in F\\ΓF\backslash\mkern-5.0mu\backslash\Gamma^{\prime}, and we choose that as our new gg. Again this process shortens γ\gamma in F\\ΓF\backslash\mkern-5.0mu\backslash\Gamma, so eventually we obtain gg such that CgC_{g} has no crossing double points.

Finally, we claim that we can avoid bad triangles: i.e. a configuration γ=δ1δ2δ3δ4δ5δ6\gamma=\delta_{1}\delta_{2}\delta_{3}\delta_{4}\delta_{5}\delta_{6}, where δ1\delta_{1}, δ3\delta_{3} and δ5\delta_{5} are closed paths and thus the other three form a “triangle,” which need not be embedded. If any of the paths δ1\delta_{1}, δ2\delta_{2} or δ3\delta_{3} is cyclically reduced in F\\ΓF\backslash\mkern-5.0mu\backslash\Gamma^{\prime}, then we may take that loop as our new gg. If none of these subpaths are cyclically reduced in F\\ΓF\backslash\mkern-5.0mu\backslash\Gamma^{\prime}, then arguing as before, we have that δ1δ2δ3δ¯2\delta_{1}\delta_{2}\delta_{3}\bar{\delta}_{2} is an immersed closed path in F\\ΓF\backslash\mkern-5.0mu\backslash\Gamma whose image in F\\ΓF\backslash\mkern-5.0mu\backslash\Gamma^{\prime} is cyclically reduced. The only worry is that this path may be longer than γ\gamma because δ2\delta_{2} is longer than δ4δ5δ6\delta_{4}\delta_{5}\delta_{6}. This would imply that the path δ3δ4δ5δ¯4\delta_{3}\delta_{4}\delta_{5}\bar{\delta}_{4}, to which the same argument applies, is shorter than γ\gamma, so we choose it. Again this process shortens γ\gamma in F\\ΓF\backslash\mkern-5.0mu\backslash\Gamma, so eventually we obtain gg such that CgC_{g} has no bad triangles.

We now think of CgC_{g} as labeled by the images of edges and vertices (with vertex group elements) in F\\ΓF\backslash\mkern-5.0mu\backslash\Gamma. Identify vertices with the same vertex label and (oriented) edges with the same edge label. We claim that the resulting graph is a sausage. If there are no identifications or only a single pair of vertex identifications, we are done. By assumption there are no triple points, so each vertex of the resulting graph is at most 44-valent. Since double points do not cross, if we have two pairs of double points in CgC_{g}, they “nest”.

If there is only a single pair of double points, we are done, as γ\gamma is a figure-eight. If we suppose there are at least two pairs of double points, then there are two “innermost” pairs which map to vertices in F\\ΓF\backslash\mkern-5.0mu\backslash\Gamma. Unlike in the original [FM11], if vv, ww is an innermost pair, the path between vv and ww may be of the form ege¯eg\bar{e}, so is not embedded. In this case, choose the middle vertex of ege¯eg\bar{e}, while in other cases choose a point on the embedded path from vv to ww arbitrarily. Since there are two innermost pairs this yields two points on CgC_{g} and therefore two subpaths γ1\gamma_{1} and γ2\gamma_{2} from the first point to the second,, so that γ=γ1γ¯2\gamma=\gamma_{1}\bar{\gamma}_{2}. Since we have no bad triangles and no crossing double points, both γ1\gamma_{1} and γ2\gamma_{2} are embedded in F\\ΓF\backslash\mkern-5.0mu\backslash\Gamma. This shows that the graph is an embedded sausage: the intervals IjI_{j} correspond to subpaths (which may be single vertices) which have more than one preimage in γ\gamma. Since double points do not cross, these intervals appear in the same order in γ1\gamma_{1} and γ2\gamma_{2}. ∎

With the Sausages Lemma in hand, we can prove B.

Theorem 2.2.

Let Γ\Gamma and Γ\Gamma^{\prime} be trees in 𝒯\mathscr{T}. There is a witness gFg\in F which is a candidate in the sense of B.

Proof.

Again the proof is nearly identical to [FM11, Proposition 3.15]. Let f:ΓΓf\colon\Gamma\to\Gamma^{\prime} be an optimal map. We begin with the loop γ=γ1γ¯2\gamma=\gamma_{1}\bar{\gamma}_{2} supplied by 2.1. If the family of intervals IjI_{j} is empty or contains at most a single interval II, then γ\gamma is a candidate and we are done.

So suppose that the family of intervals IjI_{j} contains at least two intervals. We will replace γ\gamma with a candidate. Let [a,b][a,b] and [c,d][c,d] be the two extremal intervals of IjI_{j}, i.e. 0ab<cd10\leq a\leq b<c\leq d\leq 1, and if 0<a0<a or d<1d<1, there is no IjI_{j} in the interval (0,a)(0,a) or (d,1)(d,1) respectively. We replace the path γ2\gamma_{2} with the path that follows γ2\gamma_{2} if t<bt<b or t>ct>c and follows γ1\gamma_{1} on the interval [b,c][b,c]. Note that δ2\delta_{2} embeds in F\\ΓF\backslash\mkern-5.0mu\backslash\Gamma because γ1(t)=γ2(s)\gamma_{1}(t)=\gamma_{2}(s) if and only if t=st=s. Note as well that the F\\fF\backslash\mkern-5.0mu\backslash f-image of δ2\delta_{2} in F\\ΓF\backslash\mkern-5.0mu\backslash\Gamma^{\prime} is reduced for the same reason and because the F\\fF\backslash\mkern-5.0mu\backslash f-images of γ1\gamma_{1} and γ2\gamma_{2} are reduced. The loop γ=γδ¯2\gamma^{\prime}=\gamma\bar{\delta}_{2} is a candidate. ∎

3 Induced maps on the spine and reduced spine

Suppose Γ\Gamma is a tree in 𝒯\mathscr{T}. In this section we will forget the metric on Γ\Gamma, thinking of it merely as a simplicial tree. By FF-equivariantly collapsing an orbit of edges, we obtain a new tree Γ\Gamma^{\prime}. If Γ𝒯\Gamma^{\prime}\in\mathscr{T}, then we say that Γ\Gamma^{\prime} is obtained from Γ\Gamma by forest collapse. There is an induced collapse map in the sense of the author in [Lym22, Section 1] F\\ΓF\\ΓF\backslash\mkern-5.0mu\backslash\Gamma\to F\backslash\mkern-5.0mu\backslash\Gamma^{\prime}, and in the sense of that paper the edge orbit determines an edge which is a collapsible forest. The set of FF-equivariant homeomorphism classes of trees in 𝒯\mathscr{T} is partially ordered under the relation of forest collapse, and the spine of Outer Space for FF, denoted K=K(F)K=K(F), is the geometric realization of this poset. That is, it is an “ordered” simplicial complex whose vertices are elements of the poset, whose edges witness the partial order, and whose higher-dimensional simplices span vertices representing totally ordered subsets of the poset. A minimal element of this poset is a reduced tree, i.e. one for which collapsing any orbit of edges yields a tree not in 𝒯\mathscr{T}.

The reduced spine of Outer Space L=L(F)L=L(F) comprises those trees Γ\Gamma all of whose edges e~\tilde{e} are surviving, in the sense that there is a reduced collapse Γ\Gamma^{\prime} of Γ\Gamma in which e~\tilde{e} is not collapsed. For example, a tree Γ\Gamma is in L(Fn)L(F_{n}) if and only if the quotient graph Fn\\ΓF_{n}\backslash\mkern-5.0mu\backslash\Gamma has no separating edges. In the case where FF is a free product of finite groups, L(F)L(F) is all of the spine of Guirardel–Levitt’s Outer Space P𝒪(F)P\mathscr{O}(F) [GL07b]. Let us remark that there is an easy characterization of those edges that are surviving by considering “shelters” in the quotient graph of groups F\\ΓF\backslash\mkern-5.0mu\backslash\Gamma in the sense of Clay and Guirardel–Levitt [GL07a, Cla09]. We have no further need of shelters, which are somewhat complicated to describe, so we refer the interested reader to Guirardel–Levitt [GL07a, before Lemma 7.4] or to Clay [Cla09, Section 1.3] for details.

In parallel, given a homomorphism α:GOut(Fn)\alpha\colon G\to\operatorname{Out}(F_{n}) with finite domain, Krstić–Vogtmann and Bestvina–Feighn–Handel consider a subcomplex of the fixed-point subcomplex Kα=Kα(Fn)K_{\alpha}=K_{\alpha}(F_{n}) of K(Fn)K(F_{n}) under the action of α(G)Out(Fn)\alpha(G)\leq\operatorname{Out}(F_{n}). By results of Culler [Cul84] and Zimmermann [Zim81], the subcomplex Kα(Fn)K_{\alpha}(F_{n}) is always nonempty, and a tree Γ\Gamma representing a vertex of K(Fn)K(F_{n}) belongs to Kα(Fn)K_{\alpha}(F_{n}) when, thinking in the quotient graph, the quotient graph Fn\\ΓF_{n}\backslash\mkern-5.0mu\backslash\Gamma is equipped with a GG-action via α\alpha.

We, like Krstić–Vogtmann and Bestvina–Feighn–Handel, are interested in a subcomplex of Kα(Fn)K_{\alpha}(F_{n}) which we term Lα=Lα(Fn)L_{\alpha}=L_{\alpha}(F_{n}). A tree Γ\Gamma representing a vertex of Kα(Fn)K_{\alpha}(F_{n}) belongs to Lα(Fn)L_{\alpha}(F_{n}) when, thinking in the quotient graph, every edge ee of Fn\\ΓF_{n}\backslash\mkern-5.0mu\backslash\Gamma is essential, in the sense that it is not contained in every maximal GG-invariant forest in Fn\\ΓF_{n}\backslash\mkern-5.0mu\backslash\Gamma.

Given a homomorphism α:GOut(Fn)\alpha\colon G\to\operatorname{Out}(F_{n}) with finite domain, let EE be the full preimage of α(G)\alpha(G) in Aut(Fn)\operatorname{Aut}(F_{n}). We can form the fiber product

E×α(G)G={(Φ,g)E×G:[Φ]=α(g)},E\times_{\alpha(G)}G=\{(\Phi,g)\in E\times G:[\Phi]=\alpha(g)\},

where [Φ][\Phi] denotes the outer class of Φ\Phi in Out(Fn)\operatorname{Out}(F_{n}). Then F=E×α(G)GF=E\times_{\alpha(G)}G is a virtually free group; in fact we have the following commutative diagram of short exact sequences

1{1}Fn{F_{n}}F{F}G{G}1{1}1{1}Fn{F_{n}}E{E}α(G){\alpha(G)}1.{1.}

Conversely, if FF sits in a short exact sequence as in the top row above, the conjugation action of FF on itself induces a map GOut(Fn)G\to\operatorname{Out}(F_{n}) which we suppose to be α\alpha, and FAut(Fn)F\to\operatorname{Aut}(F_{n}) whose image is EE, so the diagram above commutes. Furthermore, Goursat’s lemma coupled with the fact that the kernels of the maps FGF\to G and FEF\to E meet trivially implies that FF is isomorphic to the fiber product E×α(G)GE\times_{\alpha(G)}G. We are ready to state and prove C.

Theorem 3.1.

Given α:GOut(Fn)\alpha\colon G\to\operatorname{Out}(F_{n}) a homomorphism with finite domain, we may consider the groups FF and EE above. The isometry i:𝒫𝒯(F)𝒫𝒯(Fn)i\colon\mathscr{PT}(F)\to\mathscr{PT}(F_{n}) induces a simplicial isomorphism L(F)LαL(F)\to L_{\alpha}. The normalizer N(α(G))N(\alpha(G)) of α(G)\alpha(G) in Out(Fn)\operatorname{Out}(F_{n}) leaves LαL_{\alpha} invariant. Let Aut0(E)\operatorname{Aut}^{0}(E) be the subgroup of Aut(E)\operatorname{Aut}(E) consisting of those automorphisms that leave FnF_{n} invariant, and let Out0(E)\operatorname{Out}^{0}(E) be its image in Out(E)\operatorname{Out}(E). We have the following short exact sequence

1{1}α(G){\alpha(G)}N(α(G)){N(\alpha(G))}Out0(E){\operatorname{Out}^{0}(E)}1.{1.}
Proof.

As in Culler–Vogtmann [CV86], we may view vertices of the spine as corresponding to trees in 𝒫𝒯(F)\mathscr{PT}(F), all of whose natural edges have the same length. (That is, if a vertex of Γ\Gamma has valence two, so that the incident edges are in the same orbit, we set the lengths of these edges to half the length of edges that are not incident to vertices of valence two).

Under the isometric embedding of A, if HFH\leq F is a finite-index subgroup, vertices of the spine K(F)K(F), thought of as trees in 𝒫𝒯(F)\mathscr{PT}(F) as above, are sent to vertices of the spine K(H)K(H). It is clear that this induces a poset map from the vertex set of K(F)K(F) to K(H)K(H).

In the situation of the statement, the action of FF on Γ\Gamma descends to an action of GG on Fn\\ΓF_{n}\backslash\mkern-5.0mu\backslash\Gamma such that the quotient graph of groups in the sense of Bass [Bas93, 3.2] is F\\ΓF\backslash\mkern-5.0mu\backslash\Gamma, and conversely every FnF_{n}-tree Γ\Gamma in K(Fn)K(F_{n}) such that the quotient Fn\\ΓF_{n}\backslash\mkern-5.0mu\backslash\Gamma is equipped with a GG-action via α\alpha is actually an FF-tree in K(F)K(F). We will show that a non-surviving edge in a tree Γ\Gamma corresponding to a vertex of K(F)K(F) corresponds to an inessential GG-invariant forest in Fn\\ΓF_{n}\backslash\mkern-5.0mu\backslash\Gamma. Indeed, the image in F\\ΓF\backslash\mkern-5.0mu\backslash\Gamma of a maximal collapsible forest in Γ\Gamma is a (collapsible) forest in the sense of the author in [Lym22, Section 2], and thus a GG-invariant forest in Fn\\ΓF_{n}\backslash\mkern-5.0mu\backslash\Gamma. An edge is non-surviving if it is contained in every forest in F\\ΓF\backslash\mkern-5.0mu\backslash\Gamma, and thus in every GG-invariant forest in Fn\\ΓF_{n}\backslash\mkern-5.0mu\backslash\Gamma and conversely. Thus the poset map above sends L(F)L(F) isomorphically to LαL_{\alpha}.

If φOut(Fn)\varphi\in\operatorname{Out}(F_{n}) belongs to the normalizer of α(G)\alpha(G), then for all gGg\in G we have φα(g)=α(g)φ\varphi\alpha(g)=\alpha(g^{\prime})\varphi for some gGg^{\prime}\in G and for ΓLα\Gamma\in L_{\alpha}, we have

(Γ.φ).α(g)=(Γ.α(g)).φ=Γ.φ,(\Gamma.\varphi).\alpha(g)=(\Gamma.\alpha(g^{\prime})).\varphi=\Gamma.\varphi,

so Γ.φ\Gamma.\varphi belongs to the fixed-point set of α(G)\alpha(G), and it is clear that all of its edges remain essential, so we conclude φ\varphi leaves LαL_{\alpha} invariant.

Let Aut0(E)\operatorname{Aut}^{0}(E) be the subgroup of automorphisms leaving FnF_{n} invariant and let Out0(E)\operatorname{Out}^{0}(E) be its image in Out(E)\operatorname{Out}(E). Similarly, write NN for the normalizer of α(G)\alpha(G) in Out(Fn)\operatorname{Out}(F_{n}) and write MM for its full preimage in Aut(Fn)\operatorname{Aut}(F_{n}). Following Krstić [Krs92], we will show that MM is isomorphic to Aut0(E)\operatorname{Aut}^{0}(E). The final claim in the theorem will then follow from the snake lemma (which applies because the image of each subgroup in the top row is normal in the bottom row) applied to the following commutative diagram

1{1}Fn{F_{n}}E{E}α(G){\alpha(G)}1{1}1{1}M{M}Aut0(E){\operatorname{Aut}^{0}(E)}1{1}1,{1,}\scriptstyle{\cong}

which yields the following exact sequence with indicated connecting homomorphism

1{1}1{1}α(G){\alpha(G)}N{N}Out0(E){\operatorname{Out}^{0}(E)}1.{1.}δ\scriptstyle{\delta}

By definition, MM consists of automorphisms Φ:FnFn\Phi\colon F_{n}\to F_{n} such that the outer class φ\varphi satisfies φα(G)φ1=α(G)\varphi\alpha(G)\varphi^{-1}=\alpha(G). Since EE is the preimage of α(G)Out(Fn)\alpha(G)\leq\operatorname{Out}(F_{n}) in Aut(Fn)\operatorname{Aut}(F_{n}), for eEe\in E, we have that the automorphism of FnF_{n} defined by xΦ(eΦ1(x)e1)x\mapsto\Phi(e\Phi^{-1}(x)e^{-1}) is in EE, say it is equal to xeΦxeΦ1x\mapsto e_{\Phi}xe_{\Phi}^{-1}. We claim that the map ΨΦ\Psi_{\Phi} defined as eeΦe\mapsto e_{\Phi} is an automorphism of EE. For fFnf\in F_{n}, we have fΦf_{\Phi} is the automorphism

xΦ(fΦ1(x)f1)=Φ(f)xΦ(f)1,x\mapsto\Phi(f\Phi^{-1}(x)f^{-1})=\Phi(f)x\Phi{(f)}^{-1},

so the restriction of ψΦ\psi_{\Phi} to FnF_{n} is Φ\Phi. An easy calculation expressing Φ(eeΦ1(x)(ee)1)\Phi(ee^{\prime}\Phi^{-1}(x){(ee^{\prime})}^{-1}) two ways shows that ψΦ\psi_{\Phi} is a homomorphism. Since the restriction of ψΦ\psi_{\Phi} to FnF_{n} is an automorphism, its kernel is a finite normal subgroup of EE, hence trivial. Its image contains a finite-index subgroup so has finite index. But EE has a well-defined, negative Euler characteristic, from which we conclude that ψΦ\psi_{\Phi} is an automorphism of EE. Similarly we have ψΦψΦ=ψΦΦ\psi_{\Phi^{\prime}}\psi_{\Phi}=\psi_{\Phi^{\prime}\Phi}, so the rule ΦψΦ\Phi\mapsto\psi_{\Phi} defines a homomorphism MAut0(E)M\to\operatorname{Aut}^{0}(E).

Conversely, given ΦAut0(E)\Phi\in\operatorname{Aut}^{0}(E), restriction to FnF_{n} yields a homomorphism Aut0(E)Aut(Fn)\operatorname{Aut}^{0}(E)\to\operatorname{Aut}(F_{n}), which we want to show is in MM. Indeed, for all eEe\in E we have

xΦ|Fn(eΦ|Fn1(x)e1)=Φ(e)xΦ(e)1,x\mapsto\Phi|_{F_{n}}(e\Phi|_{F_{n}}^{-1}(x)e^{-1})=\Phi(e)x\Phi{(e)}^{-1},

which is clearly in EE. By definition, we have that ψΦ|Fn(e)=Φ(e)\psi_{\Phi|_{F_{n}}}(e)=\Phi(e), and we saw already that ψΦ|Fn=Φ\psi_{\Phi}|_{F_{n}}=\Phi, so the operations defined above are inverse homomorphisms and we conclude that MAut0(E)M\cong\operatorname{Aut}^{0}(E). ∎

References

  • [Bas93] Hyman Bass. Covering theory for graphs of groups. J. Pure Appl. Algebra, 89(1-2):3–47, 1993.
  • [BF14] Mladen Bestvina and Mark Feighn. Hyperbolicity of the complex of free factors. Adv. Math., 256:104–155, 2014.
  • [BFH20] Mladen Bestvina, Mark Feighn, and Michael Handel. A McCool Whitehead type theorem for finitely generated subgroups of 𝖮𝗎𝗍(Fn)\mathsf{Out}(F_{n}). Available at arXiv:2009.10052 [math.GR], September 2020.
  • [Cla09] Matt Clay. Deformation spaces of GG-trees and automorphisms of Baumslag-Solitar groups. Groups Geom. Dyn., 3(1):39–69, 2009.
  • [CM87] Marc Culler and John W. Morgan. Group actions on 𝐑{\bf R}-trees. Proc. London Math. Soc. (3), 55(3):571–604, 1987.
  • [Cul84] Marc Culler. Finite groups of outer automorphisms of a free group. In Contributions to group theory, volume 33 of Contemp. Math., pages 197–207. Amer. Math. Soc., Providence, RI, 1984.
  • [CV86] Marc Culler and Karen Vogtmann. Moduli of graphs and automorphisms of free groups. Invent. Math., 84(1):91–119, 1986.
  • [DT17] Spencer Dowdall and Samuel J. Taylor. The co-surface graph and the geometry of hyperbolic free group extensions. J. Topol., 10(2):447–482, 2017.
  • [FM11] Stefano Francaviglia and Armando Martino. Metric properties of outer space. Publ. Mat., 55(2):433–473, 2011.
  • [FM15] Stefano Francaviglia and Armando Martino. Stretching factors, metrics and train tracks for free products. Illinois J. Math., 59(4):859–899, 2015.
  • [For02] Max Forester. Deformation and rigidity of simplicial group actions on trees. Geom. Topol., 6:219–267, 2002.
  • [GL07a] Vincent Guirardel and Gilbert Levitt. Deformation spaces of trees. Groups Geom. Dyn., 1(2):135–181, 2007.
  • [GL07b] Vincent Guirardel and Gilbert Levitt. The outer space of a free product. Proc. Lond. Math. Soc. (3), 94(3):695–714, 2007.
  • [Krs92] Sava Krstić. Finitely generated virtually free groups have finitely presented automorphism group. Proc. London Math. Soc. (3), 64(1):49–69, 1992.
  • [KV93] Sava Krstić and Karen Vogtmann. Equivariant outer space and automorphisms of free-by-finite groups. Comment. Math. Helv., 68(2):216–262, 1993.
  • [Lym22] Rylee Alanza Lyman. Train track maps on graphs of groups. Groups Geom. Dyn., 16(4):1389–1422, 2022.
  • [Mei15] Sebastian Meinert. The Lipschitz metric on deformation spaces of GG-trees. Algebr. Geom. Topol., 15(2):987–1029, 2015.
  • [Ser03] Jean-Pierre Serre. Trees. Springer Monographs in Mathematics. Springer-Verlag, Berlin, 2003. Translated from the French original by John Stillwell, Corrected 2nd printing of the 1980 English translation.
  • [Zim81] Bruno Zimmermann. Über Homöomorphismen nn-dimensionaler Henkelkörper und endliche Erweiterungen von Schottky-Gruppen. Comment. Math. Helv., 56(3):474–486, 1981.