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

The rates of growth in an acylindrically hyperbolic group

Koji Fujiwara [email protected] Department of Mathematics, Kyoto University, Kyoto, 606-8502, Japan
Abstract.

Let GG be an acylindrically hyperbolic group on a δ\delta-hyperbolic space XX. Assume there exists MM such that for any finite generating set SS of GG, the set SMS^{M} contains a hyperbolic element on XX. Suppose that GG is equationally Noetherian. Then we show the set of the growth rates of GG is well-ordered (Theorem 1.1). The conclusion was known for hyperbolic groups, and this is a generalization.

Our result applies to all lattices in simple Lie groups of rank-1 (Theorem 1.3), and more generally, some family of relatively hyperbolic groups (Theorem 1.2). It also applies to the fundamental group, of exponential growth, of a closed orientable 33-manifold except for the case that the manifold has Sol-geometry (Theorem 5.7). A potential application is a mapping class group, to which the theorem applies if it is equationally Noetherian.

The author is supported in part by Grant-in-Aid for Scientific Research (No. 15H05739, 20H00114).

1. Introduction

1.1. Definitions and results

Let GG be a finitely generated group with a finite generating set SS. We always assume that S=S1S=S^{-1}. Let Bn(G,S)B_{n}(G,S) be the set of elements in GG whose word lengths are at most nn with respect to the generating set SS. We also denote SnS^{n} instead of Bn(G,S)B_{n}(G,S). Let βn(G,S)=|Bn(G,S)|\beta_{n}(G,S)=|B_{n}(G,S)|. The exponential growth rate of (G,S)(G,S) is defined to be:

e(G,S)=limnβn(G,S)1n.e(G,S)=\lim_{n\to\infty}\beta_{n}(G,S)^{\frac{1}{n}}.

A finitely generated group GG has exponential growth if there exists a finite generating set SS such that e(G,S)>1e(G,S)>1. The group GG has uniform exponential growth if there exists c>1c>1, such that for every finite generating set SS, e(G,S)ce(G,S)\geq c.

Given a finitely generated group GG, we define:

e(G)=inf|S|<e(G,S),e(G)=\inf_{|S|<\infty}e(G,S),

where the infimum is taken over all the finite generating sets SS of GG. Since there are finitely generated groups that have exponential growth, but do not have uniform exponential growth [W], the infimum, e(G)e(G), is not always obtained by a finite generating set of a finitely generated group.

Given a finitely generated group GG, we further define the following set in \mathbb{R}:

ξ(G)={e(G,S)||S|<},\xi(G)=\{e(G,S)||S|<\infty\},

where SS runs over all the finite generating sets of GG. The set ξ(G)\xi(G) is always countable.

A non-elementary hyperbolic group contains a non-abelian free group, hence, it has exponential growth. In fact, a non-elementary hyperbolic group has uniform exponential growth [K]. Recently it is proved that ξ(G)\xi(G) of a non-elementary hyperbolic group GG is well-ordered (hence, in particular, has the minimum), [FS]. It was new even for free groups.

In this paper, we deal with larger classes of groups. We state a main result. See Definition 1.8 for the definition of acylindricity. See Definition 1.9 for the definition of equational Noetherianity.

Theorem 1.1 (Well-orderedness for acylindrical actions).

Suppose GG acts on a δ\delta-hyperbolic space XX acylindrically, and the action is non-elementary. Assume that there exists a constant MM such that for any finite generating set SS of GG, the set SMS^{M} contains a hyperbolic element on XX. Assume that GG is equationally Noetherian. Then, ξ(G)\xi(G) is a well-ordered set.

In particular, infξ(G)\inf\xi(G) is realized by some SS, ie, e(G)=e(G,S)e(G)=e(G,S).

The theorem holds under a weaker assumption, namely, we may replace the acylindricity of the action by that SMS^{M} contains a hyperbolic and WPD element (Theorem 3.1). See Definition 2.1 for the definition of WPD. Theorem 1.1 is an immediate consequence of Theorem 3.1 by Lemma 2.3. See the explanation at the beginning of the section 3.

We give some applications.

Theorem 1.2 (Theorem 5.4).

Let GG be a group that is hyperbolic relative to a collection of subgroups {P1,,Pn}\{P_{1},\cdots,P_{n}\}. Suppose GG is not virtually cyclic, and not equal to PiP_{i} for any ii. Suppose each PiP_{i} is finitely generated and equationally Noetherian. Then ξ(G)\xi(G) is well-ordered.

As examples of this theorem, we prove:

Theorem 1.3 (Rank-1 lattices, Theorem 5.5).

Let GG be one of the following groups:

  1. (1)

    A lattice in a simple Lie group of rank-1.

  2. (2)

    The fundamental group of a complete Riemannian manifold MM of finite volume such that there exist a,b>0a,b>0 with b2Ka2<0-b^{2}\leq K\leq-a^{2}<0, where KK denotes the sectional curvature.

Then ξ(G)\xi(G) is well-ordered.

Another family of examples are 33-manifold groups.

Theorem 1.4 (Theorem 5.7).

Let MM be a closed orientable 33-manifold, and G=π1(G)G=\pi_{1}(G). If MM is one of the following, then GG has exponential growth and ξ(G)\xi(G) is well-ordered.

  1. (1)

    MM is not irreducible and GG is not isomorphic to 22\mathbb{Z}_{2}*\mathbb{Z}_{2}.

  2. (2)

    MM is irreducible, not a torus bundle over a circle, and MM has a non-trivial JSJ-decomposition.

  3. (3)

    MM admits hyperbolic geometry.

  4. (4)

    MM is Seifert fibered such that the base orbifold is hyperbolic.

Potential examples of application of the main theorem are mapping class groups. We discuss this class in Section 5.3. See Example 1.12 for non-examples.

We also show some finiteness result as follows. This was known for hyperbolic groups too, [FS].

Theorem 1.5 (Theorem 7.1).

Suppose the same assumption holds for GG as in Theorem 1.1. Then for any ρξ(G)\rho\in\xi(G), up to the action of Aut(G)Aut(G), there are at most finitely many finite generating sets SS such that e(G,S)=ρe(G,S)=\rho.

As a part of the proof of the main result, we show a basic result on the growth of a group, generalizing a result known for hyperbolic group in [AL]. Given a group GG that satisfies the assumption in Theorem 1.1, where we do not need that GG is equationally Noetherian, there exists a constant A>0A>0 such that for any finite generating set SS of GG, we have

e(G,S)A|S|A.e(G,S)\geq A|S|^{A}.

The constant AA depends only on δ\delta and the acylindricity constants. See Proposition 2.10 for the statement. Examples include mapping class groups and rank-1 lattices, see Example 2.11.

We also discuss the set of growth of subgroups in a finitely generated group GG. Define

Θ(G)={e(H,S)|SG,|S|<,H=S,e(H,S)>1}.\Theta(G)=\{e(H,S)|S\subset G,|S|<\infty,H=\langle S\rangle,e(H,S)>1\}.

The set Θ(G)\Theta(G) is countable and contains ξ(G)\xi(G). If GG is a hyperbolic group, it is known by [FS, Section 5] that Θ(G)\Theta(G) is well-ordered.

Similarly, we prove:

Theorem 1.6 (Theorem 6.6).

Suppose GG is one of the groups in Theorem 1.3. Then Θ(G)\Theta(G) is a well-ordered set.

We also prove:

Theorem 1.7 (Finiteness, Theorem 7.15).

Let GG be one of the groups in Theorem 1.3. Then for each ρΘ(G)\rho\in\Theta(G), there are at most finitely many (H,S)(H,S), up to isomorphism, such that SS is a finite generating set of H<GH<G with e(H,S)=ρe(H,S)=\rho.

This kind of finiteness is known for hyperbolic groups [FS], and we generalize it (Theorem 7.1), which implies the above theorem as examples.

Some more definitions are in order in the following section.

1.2. Acylindrical actions

To generalize the properness of an action, Bowditch [Bo] introduced the following definition.

Definition 1.8 (acylindrical action).

An action of a group GG on a metric space XX is acylindrical if for any ϵ>0\epsilon>0, there exist R=R(ϵ)>0R=R(\epsilon)>0 and N=N(ϵ)>0N=N(\epsilon)>0 such that for all x,yXx,y\in X with d(x,y)Rd(x,y)\geq R, the set

{gG|d(x,g(x))ϵ,d(y,g(y))ϵ}\{g\in G|d(x,g(x))\leq\epsilon,d(y,g(y))\leq\epsilon\}

contains at most NN elements.

A group GG is called an acylindrically hyperbolic group, [O], if it acts on some δ\delta-hyperbolic space XX such that the action is acylindrical and non-elementary. Here, we say the action is elementary if the limit set of GG in the Gromov boundary X\partial X has at most two points. If the action is non-elementary, it is known that GG contains hyperbolic isometries. Non-elementary hyperbolic groups and non-virtually-abelian mapping class groups are examples of acylindrically hyperbolic groups, [Bo]. There are many other examples.

1.3. Limit groups and equational Noetherianity

Let GG be a group and Γ\Gamma a finitely generated (or countable) group. Let Hom(Γ,G)\text{Hom}\,(\Gamma,G) be the set of all homomorphisms from Γ\Gamma to GG.

A sequence of homomorphisms {fn}\{f_{n}\} from Γ\Gamma to GG is stable if for each gΓg\in\Gamma, either fn(g)=1f_{n}(g)=1 for all sufficiently large nn; or fn(g)1f_{n}(g)\not=1 for all sufficiently large nn. If the sequence is stable, then the stable kernel of the sequence, ker(fn)\underrightarrow{\ker}(f_{n}), is defined by

ker(fn)={gΓ|fn(g)=1 for all sufficiently large n}.\underrightarrow{\ker}(f_{n})=\{g\in\Gamma|f_{n}(g)=1\text{ for all sufficiently large }n\}.

We call the quotient Γ/ker(fn)\Gamma/\underrightarrow{\ker}(f_{n}) a GG-limit group, or the limit group over GG, associated to {fn}\{f_{n}\}, and the homomorphism f:ΓΓ/ker(fn)f:\Gamma\to\Gamma/\underrightarrow{\ker}(f_{n}) the limit homomorphism. We say the sequence {fn}\{f_{n}\} converges to ff.

Let GG be a group and F(x1,,x)F(x_{1},\cdots,x_{\ell}) the free group on X={x1,,x}X=\{x_{1},\cdots,x_{\ell}\}. For an element sF(x1,,x)s\in F(x_{1},\cdots,x_{\ell}) and (g1,,g)G(g_{1},\cdots,g_{\ell})\in G^{\ell}, let s(g1,,g)Gs(g_{1},\cdots,g_{\ell})\in G denote the element after we substitute every xix_{i} with gig_{i} and xi1x_{i}^{-1} by gi1g_{i}^{-1} in ss. Given a subset SF(x1,,x)S\subset F(x_{1},\cdots,x_{\ell}), define

VG(S)={(g1,,g)G|s(g1,,g)=1 for all sS}.V_{G}(S)=\{(g_{1},\cdots,g_{\ell})\in G^{\ell}|s(g_{1},\cdots,g_{\ell})=1\text{ for all }s\in S\}.

SS is called a system of equations (with XX the set of variables), and VG(S)V_{G}(S) is called the algebraic set over GG defined by SS. We sometimes suppress GG from VG(S)V_{G}(S).

Definition 1.9 (Equationally Noetherian).

A group GG is equationally Noetherian if for every 1\ell\geq 1 and every subset SS in F(x1,,x)F(x_{1},\cdots,x_{\ell}), there exists a finite subset S0SS_{0}\subset S such that VG(S0)=VG(S)V_{G}(S_{0})=V_{G}(S).

Remark 1.10.

This definition appears in for example [GrH]. There is another version of the definition that considers SGF(x1,,x)S\subset G*F(x_{1},\cdots,x_{\ell}), which is originally in [BMR] and also in [RW]. They are equivalent, see [RW, Lemma5.1].

Examples of equationally Noetherian groups include free groups, [Gu]; linear groups, [BMR]; hyperbolic groups without torsion, [S], then possibly with torsion, [RW]; and hyperbolic groups relative to equationally Noetherian subgroups, [GrH].

What is important for us is the following general principle.

Lemma 1.11 (Basic principle).

Let η:FL\eta:F\to L be the limit map of a sequence of homomorphisms, fn:FGf_{n}:F\to G. Suppose GG is equationally Noetherian. Then for sufficiently large nn, fnf_{n} factors through η\eta, namely, there exists a homomorphism hn:LGh_{n}:L\to G such that hnη=fnh_{n}\circ\eta=f_{n}.

Proof.

Let X={x1,,x}X=\{x_{1},\cdots,x_{\ell}\} and suppose F=F(X)F=F(X). Let R={ri}F(X)R=\{r_{i}\}\subset F(X) be a set of defining relations for LL. In general this is an infinite set. Each rir_{i} is a word on XX, so that we can see RR as a system of equations with XX the variable set. Since GG is equationally Noetherian, there is a finite subset R0RR_{0}\subset R such that V(R)=V(R0)V(R)=V(R_{0}), namely, every solution (an element in GG^{\ell}) for R0R_{0} is a solution for RR.

Now, since η(r1)=1\eta(r_{1})=1 in LL for a large enough nn, we have fn(r1)=1f_{n}(r_{1})=1 in GG since η\eta is the limit of {fn}\{f_{n}\}. By the same reason, since R0R_{0} is a finite set, there exists NN such that for every nNn\geq N, we have fn(ri)=1f_{n}(r_{i})=1 in GG for all riR0r_{i}\in R_{0}. In other words, if nNn\geq N, then (fn(x1),,fn(x))V(R0)(f_{n}(x_{1}),\cdots,f_{n}(x_{\ell}))\in V(R_{0}). But since V(R0)=V(R)V(R_{0})=V(R), this implies that if nNn\geq N, then (fn(x1),,fn(x))V(R)(f_{n}(x_{1}),\cdots,f_{n}(x_{\ell}))\in V(R), namely, fn(ri)=1f_{n}(r_{i})=1 in GG for all riRr_{i}\in R. Since RR is a system of defining relations for LL, it implies that each fnf_{n} with nNn\geq N factors through η:FL\eta:F\to L. ∎

Regarding Theorem 1.1, there exists a finitely generated group GG that acylindrically acts on a δ\delta-hyperbolic space, in fact a simplicial tree TT, in the non-elementary way such that for any finite generating set SS, the set S2S^{2} contains a hyperbolic isometry on TT; and that there is no finite generating set SS with e(G)=e(G,S)>1e(G)=e(G,S)>1. In particular ξ(G)\xi(G) is not well-ordered.

The following example is pointed out by Ashot Minasyan. A group GG is called Hopfian if every surjective homomorphism f:GGf:G\to G is an isomorphism.

Example 1.12.

Take a finitely generated non-Hopfian group, GG, for example BS(2,3)BS(2,3), (a Baumslag-Solitar group). Put H=GH=G*\mathbb{Z}. Then HH is non-Hopfian, ie, there exists a surjection f:HHf:H\to H that is not an isomorphism. It is a standard fact that a finitely generated equationally Noetherian group is Hopfian, therefore HH is not equationally Noetherian, so that Theorem 1.1 does not apply to HH.

But all other assumptions in the theorem are satisfied. Let TT be the Bass-Serre tree for GG*\mathbb{Z}. The tree TT is 0-hyperbolic. The action of HH on TT is acylindrical and non-elementary. Also, for any finite generating set SS of HH, it is a well-known lemma (due to Serre, cf. [BrF]) that S2S^{2} contains a hyperbolic isometry on TT.

On the other hand, it is known, [Sa], that a finitely generated group KK that is a free product is growth-tight, namely, for any surjective homomorphism h:KKh:K\to K that is not an isomorphism, and for any finite generating set SS of KK, e(K,S)>e(K,h(S))e(K,S)>e(K,h(S)). Now, it follows that e(H)e(H) is not achieved by any SS, since if it did, then take such SS. But then, the non-isomorphic, surjective homomorphism ff in the above would imply e(H,S)>e(H,f(S))e(H,S)>e(H,f(S)), a contradiction.

Acknowledgement. This work is a continuation of [FS]. The author would like to thank Zlil Sela for generously offering numerous comments and suggestions throughout the work. He is grateful to Mladen Bestvina, Emmanuel Breuillard, Thomas Delzant, Daniel Groves and Wenyuan Yang for useful comments.

He would like to thank the referee for reading the paper thoroughly and making many suggestions, by which the paper significantly improved not only in terms of the presentation, but also mathematically, for example on Proposition 7.2.

2. Lower bound of a growth rate

Although most statements in the paper are for geodesic spaces XX, we consider a graph for XX instead of a geodesic space in the arguments throughout the paper unless we indicate otherwise. The advantage is that an infimum is achieved for various notions, for example, L(S)L(S), in the arguments. But, we do not lose generality if we assume XX to be a graph since we can consider the 11-skeleton of a Rips complex of a geodesic space with a group action. Also, the various assumptions we consider (such as the hyperbolicity of the space, acylindricity of the action) remain valid, maybe with slightly difference constants.

2.1. Hyperbolic isometries and axes

Suppose a group GG acts on a geodesic space XX by isometries. Choose a base point xXx\in X and for gGg\in G, put

L(g)=infxX|xg(x)|,λ(g)=limn|xgn(x)|n.L(g)=\inf_{x\in X}|x-g(x)|,\lambda(g)=\lim_{n\to\infty}\frac{|x-g^{n}(x)|}{n}.

L(g)L(g) is called the minimal displacement. λ(g)\lambda(g) is called the translation length, and does not depend on the choice of xx, and for any n>0n>0 we have λ(gn)=nλ(g)\lambda(g^{n})=n\lambda(g). We also have

λ(g)L(g)λ(g)+7δ.\lambda(g)\leq L(g)\leq\lambda(g)+7\delta.

The first inequality is trivial and we leave the second as an exercise (for example, use [AL, Corollary 1]). For a finite set SGS\subset G and xXx\in X, define

L(S,x)=maxsS|xs(x)|,L(S,x)=\max_{s\in S}|x-s(x)|,

then define

L(S)=infxXL(S,x).L(S)=\inf_{x\in X}L(S,x).

We recall a few definitions and facts from δ\delta-hyperbolic spaces. An isometry gg of a hyperbolic space XX is called elliptic if the orbit of a point by gg is bounded, and hyperbolic if there are xXx\in X and C>0C>0 such that for any n>0n>0, we have |xgn(x)|>Cn|x-g^{n}(x)|>Cn. The element gg is hyperbolic iff λ(g)>0\lambda(g)>0.

A hyperbolic isometry gg is associated with a bi-infinite quasi-geodesic, γ\gamma, called an axis in XX. If there exists a bi-infinite geodesic γ\gamma that is invariant by gg, that would be an ideal choice for an axis, but that is not always the case.

As a remedy, if L(g)10δL(g)\geq 10\delta, take a point xXx\in X where L(g)L(g) is achieved. Then take a geodesic [x,g(x)][x,g(x)] between xx and g(x)g(x) and consider the union of its gg-orbit, which defines a gg-invariant path, (see for example [De]). If L(g)<10δL(g)<10\delta, then take n>0n>0 such that L(gn)10δL(g^{n})\geq 10\delta and apply the construction to gng^{n}, and use this path for gg, which is not gg-invariant. We denote this axis as A(g)A(g) in this paper. Also, for gng^{n} with g0g\not=0, we may also take A(g)A(g) as an axis for gng^{n}.

For gg, an axis A(g)A(g) is not unique, but uniformly (over all hyperbolic gg) quasi-geodesic, such that for any two points x,yA(g)x,y\in A(g), the Hausdorff distance between the segment between x,yx,y on A(g)A(g) and a geodesic between x,yx,y, [x,y][x,y], is at most 10δ10\delta. Also, if Hd(A(g),h(A(g))Hd(A(g),h(A(g)) for hGh\in G is finite, then it is bounded by 10δ10\delta, where HdHd is the Hausdorff distance. We sometimes call A(g)A(g) a 10δ10\delta-axes. We consider a direction on the 10δ10\delta-axis using the action of gg.

A hyperbolic isometry gg defines two limit points in X\partial X, the visual(Gromov) boundary of XX, by g=limngn(x),g=limngn(x)g^{\infty}=\lim_{n\to\infty}g^{n}(x),g^{-\infty}=\lim_{n\to-\infty}g^{n}(x), where xx is a base point. We say two hyperbolic isometries g,hg,h are independent if {g±}\{g^{\pm\infty}\} and {h±}\{h^{\pm\infty}\} are disjoint. If the Hausdorff distance between two axes is finite, then we say they are parallel.

2.2. WPD elements

We consider another version of properness of a group action that is weaker than acylindricity.

Definition 2.1 (WPD, uniformly WPD, DD-WPD).

Let GG act on a δ\delta-hyperbolic space XX. Suppose that gGg\in G is hyperbolic on XX. We say gg is WPD if there is a 10δ10\delta-axis, γ\gamma, of gg such that for any ϵ>0\epsilon>0, there exists D=D(ϵ)>0D=D(\epsilon)>0 such that for any x,yγx,y\in\gamma with |xy|Dλ(g)|x-y|\geq D\lambda(g), the number of the elements in the following set is at most DD:

(1) {hG||h(x)x|ϵ,|h(y)y|ϵ}.\{h\in G||h(x)-x|\leq\epsilon,|h(y)-y|\leq\epsilon\}.

In application we often take y=gD(x)y=g^{D}(x). If we want to make the function DD explicit, we say that gg is DD-WPD, or WPD w.r.t. DD. We say that DD is a function for WPD.

If there is a function DD such that if a set of hyperbolic elements in GG are DD-WPD, then we say they are uniformly WPD, or uniformly DD-WPD. If all hyperbolic elements in GG are uniformly DD-WPD, then we say the action is uniformly (DD-)WPD.

Some remarks are in order. The notion of WPD (weak proper discontinuity) was introduced in [BeF], where the function DD is not used, but the definitions are equivalent.

If gg is DD-WPD, then it is DD^{\prime}-WPD for any function DD^{\prime} such that D(ϵ)D(ϵ)D^{\prime}(\epsilon)\geq D(\epsilon) for all ϵ\epsilon. So, without loss of generality, we assume that D(ϵ)D(\epsilon) does not decrease when we increase ϵ\epsilon. We often use the value D(100δ)D(100\delta) in this paper, for example, see Lemma 2.2. For convenience we also assume that D(100δ)50D(100\delta)\geq 50, which we use in the proof of Lemma 7.4.

The choice of a 10δ10\delta-axis is not important in the definition. Also, one can use CC-axis for any C>0C>0. It only changes the function D(ϵ)D(\epsilon). Uniformly WPD is related to but weaker than the notion of weak acylindricity in [De]. See also [De, Example 1] for the difference between acylindricity and variations of WPD.

For an acylindrical action, it is known ([F, Lemma 2.1]) that there exists T>0T>0 such that for any hyperbolic element gg, we have λ(g)T\lambda(g)\geq T. This holds for uniformly WPD actions too, and the argument is same, but for the readers’ convenience, we prove it.

Lemma 2.2 (Lower bound on λ(g)\lambda(g)).

Suppose GG acts on a δ\delta-hyperbolic space XX. Let gGg\in G be hyperbolic with a 10δ10\delta-axis γ\gamma.

  1. (1)

    If gg is DD-WPD, then λ(g)50δD(100δ)\lambda(g)\geq\frac{50\delta}{D(100\delta)}.

  2. (2)

    If the action is acylindrical, then λ(g)50δN(150δ)\lambda(g)\geq\frac{50\delta}{N(150\delta)}.

Proof.

(1) Set 𝒟=D(100δ)\mathcal{D}=D(100\delta). Let x,yγx,y\in\gamma with |xy|𝒟λ(g)|x-y|\geq\mathcal{D}\lambda(g). Then there must be some nn with 0n𝒟0\leq n\leq\mathcal{D} such that |xgn(x)|>100δ|x-g^{n}(x)|>100\delta or |ygn(y)|>100δ|y-g^{n}(y)|>100\delta. This is because otherwise, all the elements 1,g,,g𝒟1,g,\cdots,g^{\mathcal{D}}, which are 𝒟+1\mathcal{D}+1 distinct elements, are contained in the set

{hG||h(x)x|100δ, and |h(y)y|100δ}.\{h\in G||h(x)-x|\leq 100\delta,\text{ and }|h(y)-y|\leq 100\delta\}.

This is impossible since the contains at most 𝒟\mathcal{D} elements. Now suppose, say, xx satisfies |xgn(x)|>100δ|x-g^{n}(x)|>100\delta. Then it implies that λ(gn)50δ\lambda(g^{n})\geq 50\delta. Since n𝒟n\leq\mathcal{D}, we have λ(g)50δ𝒟\lambda(g)\geq\frac{50\delta}{\mathcal{D}}.

(2) The argument is similar to (1). Take x,yγx,y\in\gamma with |xy|=R(150δ)|x-y|=R(150\delta). If |xgn(x)|100δ|x-g^{n}(x)|\leq 100\delta for some nn, then |ygn(y)|150δ|y-g^{n}(y)|\leq 150\delta. This implies, by acylindricity, there must be nn with 1nN(150δ)1\leq n\leq N(150\delta) s.t. |xgn(x)|>100δ|x-g^{n}(x)|>100\delta. It follows λ(gn)50δ\lambda(g^{n})\geq 50\delta, so that λ(g)50δN(150δ)\lambda(g)\geq\frac{50\delta}{N(150\delta)}. ∎

For a function D(ϵ)D(\epsilon), put

T=50δD(100δ),T=\frac{50\delta}{D(100\delta)},

then by Lemma 2.2 (1) we have λ(g)T\lambda(g)\geq T for a DD-WPD element gg.

The following lemma is straightforward.

Lemma 2.3 (Acylindricity implies uniform WPD).

If an action of GG on a δ\delta-hyperbolic space XX is acylindrical, then it is uniformly WPD.

Proof.

Suppose gGg\in G is hyperbolic on XX, and let γ\gamma be a 10δ10\delta-axis. Let R(ϵ),N(ϵ)R(\epsilon),N(\epsilon) be the acylindricity constants. Also, let T>0T>0 be a uniform bound for λ(g)T\lambda(g)\geq T by Lemma 2.2 (2). TT does not depend on gg nor ϵ\epsilon. Suppose ϵ\epsilon is given. Let K=K(ϵ)K=K(\epsilon) be a smallest integer with KR(ϵ)TK\geq\frac{R(\epsilon)}{T}. Then λ(gK)R(ϵ)\lambda(g^{K})\geq R(\epsilon). The constant KK does not depend on gg. Then for any xγx\in\gamma and nKn\geq K, we have |xgn(x)|R(ϵ)|x-g^{n}(x)|\geq R(\epsilon). By acylindricity, there are at most N(ϵ)N(\epsilon) elements which simultaneously move each of x,gn(x)x,g^{n}(x) by at most ϵ\epsilon. Put D(ϵ)=max{K(ϵ),N(ϵ)}D(\epsilon)=\max\{K(\epsilon),N(\epsilon)\}, then the action is uniformly DD-WPD. ∎

We state a lemma which is useful for us.

Lemma 2.4.

Suppose there are at most DD elements that satisfies the condition (1) in the definition 2.1 for ϵ=100δ\epsilon=100\delta if |xy|Dλ(g)|x-y|\geq D\lambda(g). Then gg is WPD, and moreover, there is a function DD^{\prime} with D(100δ)=DD^{\prime}(100\delta)=D that depends only on D,δD,\delta such that gg is DD^{\prime}-WPD.

Proof.

Let γ\gamma be a 10δ10\delta-axis of gg. Suppose ϵ>0\epsilon>0 is given. We may assume ϵ>100δ\epsilon>100\delta. Take x,yγx,y\in\gamma such that |xy|Dλ(g)+2ϵ+1000δ|x-y|\geq D\lambda(g)+2\epsilon+1000\delta. Take p,qγp,q\in\gamma between xx and yy with |xp|=|yq|=ϵ+50δ|x-p|=|y-q|=\epsilon+50\delta. Then |pq|Dλ(g)+800δ|p-q|\geq D\lambda(g)+800\delta.

Let JJ be the collection of elements in GG such that |xj(x)|ϵ|x-j(x)|\leq\epsilon and |yj(y)|ϵ|y-j(y)|\leq\epsilon. If jJj\in J, then |pj(p)|ϵ+30δ|p-j(p)|\leq\epsilon+30\delta and |qj(q)|ϵ+30δ|q-j(q)|\leq\epsilon+30\delta; and j(p),j(q)j(p),j(q) are in the 15δ15\delta-neighborhood of γ\gamma. It implies that JJ contains a subset J0J_{0} that contains at most (2ϵ+200δ)/(10δ)(2\epsilon+200\delta)/(10\delta) elements such that for any jJj\in J, one can find j0J0j_{0}\in J_{0} with |pj1j0(p)|70δ|p-j^{-1}j_{0}(p)|\leq 70\delta. To see it, consider points p1,p2γp_{1},p_{2}\in\gamma with |p1p2|=2ϵ+200δ|p_{1}-p_{2}|=2\epsilon+200\delta such that pp is the mid point of the segment between p1,p2p_{1},p_{2} on γ\gamma, which we denote by [p1,p2][p_{1},p_{2}]. Then the points j(p)j(p) with jJj\in J is contained in the 30δ30\delta-neighborhood of [p1,p2][p_{1},p_{2}]. One should imagine that this neighborhood is a narrow tube around [p1,p2][p_{1},p_{2}]. Now by a pigeon hole argument, one can find a desirable subset J0J_{0}. (Notice that |pj1j0(p)|=|j(p)j0(p)||p-j^{-1}j_{0}(p)|=|j(p)-j_{0}(p)|, so that one needs to find a point j0(p)j_{0}(p) near (i.e. at most 70δ70\delta) a given point j(p)j(p), which is possible.)

But |pj1j0(p)|70δ|p-j^{-1}j_{0}(p)|\leq 70\delta implies |qj1j0(q)|100δ|q-j^{-1}j_{0}(q)|\leq 100\delta. This is because |pq|800δ|p-q|\geq 800\delta and both [j(p),j(q)][j(p),j(q)] and [j0(p),j0(q)][j_{0}(p),j_{0}(q)] are contained in the 20δ20\delta-neighborhood of γ\gamma. We have shown that the element j1j0j^{-1}j_{0} moves both p,qp,q by at most 100δ100\delta. By our assumption, there are at most DD possibilities for such element. In conclusion, JJ contains at most D×(2ϵ+200δ)/(10δ)D\times(2\epsilon+200\delta)/(10\delta) elements. We proved that gg is WPD.

We compute a WPD-function for gg, which we denote by DD^{\prime}. By assumption, we may set D(100δ)=DD^{\prime}(100\delta)=D. First,

D×(2ϵ+200δ)/(10δ)=D×(ϵ/(5δ)+20).D\times(2\epsilon+200\delta)/(10\delta)=D\times(\epsilon/(5\delta)+20).

Next,

D+(2ϵ+1000δ)/λ(g)D+D(2ϵ+1000δ)/(50δ)=D(21+ϵ/(25δ))D+(2\epsilon+1000\delta)/\lambda(g)\geq D+D(2\epsilon+1000\delta)/(50\delta)=D(21+\epsilon/(25\delta))

by Lemma 2.2 (1). So, if ϵ>100δ\epsilon>100\delta, we set

D(ϵ)=Dmax{20+ϵ/(5δ),21+ϵ/(25δ)}=D×(21+ϵ/(5δ)).D^{\prime}(\epsilon)=D\max\{20+\epsilon/(5\delta),21+\epsilon/(25\delta)\}=D\times(21+\epsilon/(5\delta)).

2.3. Elementary closure

Suppose GG acts on a hyperbolic space XX and let gGg\in G be a hyperbolic isometry with an axis γ\gamma. The elementary closure of gg is defined by

E(g)={hG|Hd(γ,h(γ))<}.E(g)=\{h\in G|Hd(\gamma,h(\gamma))<\infty\}.

It turns out that E(g)E(g) is a subgroup of GG. Clearly, g<E(g)\langle g\rangle<E(g).

We denote the aa-neighborhood of a subset YXY\subset X by Na(Y)N_{a}(Y).

Lemma 2.5 (parallel axes).

Suppose GG acts on a δ\delta-hyperbolic space XX. Let gGg\in G be hyperbolic with a 10δ10\delta-axis γ\gamma. Let hGh\in G, then

  1. (1)

    If hE(g)h\in E(g), then Hd(γ,h(γ))50δHd(\gamma,h(\gamma))\leq 50\delta.

  2. (2)

    Assume that gg is DD-WPD. If hE(g)h\not\in E(g), then the diameter of h(γ)N50δ(γ)h(\gamma)\cap N_{50\delta}(\gamma) is at most

    2D(100δ)L(g)+100δ.2D(100\delta)L(g)+100\delta.

This lemma is well-known in slightly different versions, for example [F, Lemma 2.2] for acylindrical actions, so the proof will be brief.

In general if two axes have finite Hausdorff distance, then we say they are parallel.

Proof.

(1) By definition, Hd(γ,h(γ))<Hd(\gamma,h(\gamma))<\infty. Since both γ,h(γ)\gamma,h(\gamma) are 10δ10\delta-axes, we have a desired bound.

(2) Suppose not. Suppose that the direction of γ,hγ\gamma,h\gamma coincide along the parallel part. Set 𝒟=D(100δ)\mathcal{D}=D(100\delta). Take xh(γ)x\in h(\gamma) near one end of the intersection such that gn(x)g^{n}(x) for 0n2𝒟0\leq n\leq 2\mathcal{D} are in the 50δ50\delta-neighborhood of the intersection. This is possible since the intersection is long enough. Consider the points x,g𝒟(x)x,g^{\mathcal{D}}(x). Then |xg𝒟(x)|𝒟λ(g)|x-g^{\mathcal{D}}(x)|\geq\mathcal{D}\lambda(g). Letting (hgh1)ngn(hgh^{-1})^{-n}g^{n} with 0n𝒟0\leq n\leq\mathcal{D} act on x,g𝒟(x)x,g^{\mathcal{D}}(x), we have

|x(hgh1)ngn(x)|100δ,|g𝒟(x)(hgh1)ngn(g𝒟(x))|100δ|x-(hgh^{-1})^{-n}g^{n}(x)|\leq 100\delta,\,\,|g^{\mathcal{D}}(x)-(hgh^{-1})^{-n}g^{n}(g^{\mathcal{D}}(x))|\leq 100\delta

for all 0n𝒟0\leq n\leq\mathcal{D}.

But since gg is DD-WPD, there are at most 𝒟\mathcal{D} such elements, so that it must be that for some nmn\not=m, we have (hgh1)ngn=(hgh1)mgm(hgh^{-1})^{-n}g^{n}=(hgh^{-1})^{-m}g^{m}, so that gnmg^{n-m} and hh commute. It implies that γ\gamma and h(γ)h(\gamma) are parallel, a contradiction.

If the direction for γ\gamma and hγh\gamma are opposite, consider (hgh1)ngn(hgh^{-1})^{n}g^{n} instead of (hgh1)ngn(hgh^{-1})^{-n}g^{n} , and the rest is same. ∎

We quote a fact.

Proposition 2.6 (elementary closure).

Suppose GG acts on a δ\delta-hyperbolic space XX. Let gGg\in G be a hyperbolic element on XX. Assume gg is DD-WPD. Then,

  1. (1)

    E(g)E(g) is virtually \mathbb{Z}, and contains g\langle g\rangle as a finite index subgroup.

  2. (2)

    If hGh\in G is a hyperbolic element such that gg and hh are independent, then E(g)E(h)E(g)\cap E(h) is finite.

Proof.

(1) This is [DGO, Lemma 6.5].

(2) If E(g)E(h)E(g)\cap E(h) is infinite, then it contains gN\langle g^{N}\rangle for some N>0N>0. But since gg and hh are independent, for a sufficiently large mm, we have gmE(h)g^{m}\not\in E(h), impossible. ∎

Note that under the assumption of the proposition, the action of GG is elementary if and only if GG is virtually \mathbb{Z}, which is equivalent to that G=E(g)G=E(g) in this case.

In general, if a group HH is virtually \mathbb{Z}, then there exists an exact sequence

1FHC1,1\to F\to H\to C\to 1,

where CC is either \mathbb{Z} or 22\mathbb{Z}_{2}*\mathbb{Z}_{2}, and FF is finite. In the case that HH is E(g)E(g) in the above, we denote the finite group FF by F(g)F(g). If C=C=\mathbb{Z}, F(g)F(g) is the set of elements of finite order in E(g)E(g).

The axis γ\gamma defines two points in the ideal boundary of XX, which we denote {γ(),γ()}\{\gamma(\infty),\gamma(-\infty)\}. E(g)E(g) is exactly the set of elements that leaves this set invariant. If hE(g)h\in E(g) swaps those two points, we say it flips the axis since it flips the direction of the axis γ\gamma.

If the action of GG on XX is DD-uniform WPD, then for any hyperbolic gg, we have |F(g)|2D(100δ)|F(g)|\leq 2D(100\delta). Moreover, if C=C=\mathbb{Z}, then |F(g)|D(100δ)|F(g)|\leq D(100\delta).

Note that g<E(g)\langle g\rangle<E(g). If the action is non-elementary, then E(g)GE(g)\not=G, which is equivalent to that GG is not virtually \mathbb{Z}, so that GG contains two independent hyperbolic isometries.

We say a hyperbolic element gg is primitive if C=C=\mathbb{Z} and each element hE(g)h\in E(g) is written as h=fgnh=fg^{n} for some fF(g)f\in F(g) and nn\in\mathbb{N}.

2.4. About the constants

From now on, there will be many constants to make the argument concrete and precise. In the argument we consider a sequence of generators SnS_{n}. It would be a good idea to keep in mind that there are two kinds of constants.

The first kind are those that are fixed once the constants δ,D(100δ),M\delta,D(100\delta),M are given by the action:

δ,D(100δ),M;T,k,m,b.\delta,D(100\delta),M;T,k,m,b.

The constants k,mk,m will appear in Lemma 2.8 and bb in the proof of Lemma 4.2.

The second kind are those that depend on a generating set SnS_{n} of GG:

L(Sn),λ(g),L(g),λ(u),L(Sn2MD),Δn.L(S_{n}),\lambda(g),L(g),\lambda(u),L(S_{n}^{2MD}),\Delta_{n}.

If L(Sn)L(S_{n}) is bounded, then all of the constants in the second will be bounded, but if not, then, roughly they all diverge in the same order as L(Sn)L(S_{n}).

We remark that if L(Sn)L(S_{n}) diverges, then one way to argue is to rescale XX by 1L(Sn)\frac{1}{L(S_{n})}, then go to the limit, which is a tree, and use the geometry of the tree. This approach is the one taken in [FS], where only this case happens. But in our setting, the new feature is that possibly, L(Sn)L(S_{n}) is bounded. In this paper, we use a unified approach.

For convenience, we assume δ>0\delta>0 from now on.

2.5. Lower bound of a growth rate

Lemma 2.7 (hyperbolic element of large displacement).

Let XX be a δ\delta-hyperbolic space and SS a finite set of isometries of XX. Suppose that L(S)30δL(S)\geq 30\delta. Let xXx\in X be such that L(S)=L(S,x)L(S)=L(S,x). Then there is a hyperbolic element gS2g\in S^{2} such that

L(S)8δ|xg(x)|L(S)-8\delta\leq|x-g(x)|

and

|xg(x)|16δL(g).|x-g(x)|-16\delta\leq L(g).

This is exactly same as a part of [AL, Lemma 7] (in that paper, our element gg is denoted as bb in the proof), although their setting is that XX is a Cayley graph of a hyperbolic group GG and SGS\subset G.

The proof is same verbatim after a suitable translation of notions, so we omit it (cf. [BrF, Theorem 1.4] for somewhat similar result).

Note that if sSs\in S then |xs(x)|L(S)|x-s(x)|\leq L(S); and if sS2s\in S^{2} then L(s)|xs(x)|2L(S)L(s)\leq|x-s(x)|\leq 2L(S) by the definition of L(S)L(S) and the choice of xx.

We summarize the properties of gg we use later:

  • L(S)24δL(g)2L(S).L(S)-24\delta\leq L(g)\leq 2L(S).

  • The distance between xx and the 10δ10\delta-axis of gg is at most 20δ20\delta.

The second one follows from L(g)=minyX|yg(y)||xg(x)|16δL(g)=\min_{y\in X}|y-g(y)|\geq|x-g(x)|-16\delta.

Lemma 2.8 (free subgroup with primitive hyperbolic elements).

Let XX be a δ\delta-hyperbolic geodesic space and GG a group acting on XX. Assume GG is not virtually cyclic. Let D(ϵ)D(\epsilon) be a function for WPD. Set

k=60D(200δ),m=66k+4=3960D(200δ)+4.k=60D(200\delta),m=66k+4=3960D(200\delta)+4.

Let SS be a finite set that generates GG with L(S)50δL(S)\geq 50\delta, and xXx\in X with L(S)=L(S,x)L(S)=L(S,x). Suppose gS2g\in S^{2} is a hyperbolic element with a 10δ10\delta-axis γ\gamma with d(x,γ)20δd(x,\gamma)\leq 20\delta such that

L(g)|xg(x)|16δL(S)24δ.L(g)\geq|x-g(x)|-16\delta\geq L(S)-24\delta.

Assume that gg is DD-WPD.

Then there exists sSs\in S such that gk,sgks1g^{k},sg^{k}s^{-1} are independent and freely generate a rank-22 free group F<GF<G with the following property. There is a WPD-function D(ϵ)D^{\prime}(\epsilon) with

D(100δ)=D(200δ)D^{\prime}(100\delta)=D(200\delta)

such that every non-trivial element hFh\in F is hyperbolic on XX and DD^{\prime}-WPD. Also, hh satisfies:

λ(h)10(2D(200δ)L(g)+100δ)10L(S).\lambda(h)\geq 10(2D(200\delta)L(g)+100\delta)\geq 10L(S).

Moreover, there is an element uFu\in F that satisfies:

  1. (1)

    uu has a 10δ10\delta-axis α\alpha with d(x,α)50δd(x,\alpha)\leq 50\delta.

  2. (2)

    uu is primitive, namely, there exists

    1F(u)E(u)11\to F(u)\to E(u)\to\mathbb{Z}\to 1

    such that any element hE(u)h\in E(u) is written as h=fuph=fu^{p} with fF(u)f\in F(u) and pp\in\mathbb{Z}.

  3. (3)

    |F(u)|D(100δ)=D(200δ)|F(u)|\leq D^{\prime}(100\delta)=D(200\delta).

  4. (4)

    uSmu\in S^{m}.

By L(S)50δL(S)\geq 50\delta, we have L(g)26δL(g)\geq 26\delta, 2L(g)L(S)2L(g)\geq L(S), and 2λ(g)L(g)2\lambda(g)\geq L(g).

Remark 2.9.

In the proof of Proposition 2.10 we will apply Lemma 2.8 to SMD(200δ)S^{MD(200\delta)} instead of a generating set SS itself. In that case, the element sSMD(200δ)s\in S^{MD(200\delta)} that appears in the above lemma can be chosen from SS itself. We choose such ss in the beginning of the proof of the lemma and the rest of the argument is exactly same.

Proof.

Set 𝒟=D(200δ)\mathcal{D}=D(200\delta) and T=50δ𝒟T=\frac{50\delta}{\mathcal{D}}. Since gg is DD-WPD, by Lemma 2.2 (1), we have λ(g)T\lambda(g)\geq T. Remember that by our assumption, we always have D(100δ)D(200δ)D(100\delta)\leq D(200\delta), so that the estimate in the lemma holds for D(200δ)D(200\delta) as well.

Since GG is not virtually cyclic, there is an element sSs\in S with sE(g)s\not\in E(g). By Lemma 2.5(2), the diameter of the intersection s(γ)N50δ(γ)s(\gamma)\cap N_{50\delta}(\gamma) is at most

2𝒟L(g)+100δ.2\mathcal{D}L(g)+100\delta.

In view of this, set

k=10(4𝒟+100δT)=60𝒟10.k=10\left(4\mathcal{D}+\frac{100\delta}{T}\right)=60\mathcal{D}\geq 10.

Then

λ(gk)kλ(g)\displaystyle\lambda(g^{k})\geq k\lambda(g) 10(4𝒟λ(g)+100δ)\displaystyle\geq 10(4\mathcal{D}\lambda(g)+100\delta)
10(2𝒟L(g)+100δ).\displaystyle\geq 10(2\mathcal{D}L(g)+100\delta).

Here, we used 2λ(g)L(g),λ(g)T2\lambda(g)\geq L(g),\lambda(g)\geq T.

Note that sgs1sgs^{-1} is hyperbolic and DD-WPD with a 10δ10\delta-axis s(γ)s(\gamma). Since λ(gk)\lambda(g^{k}) is at least 10 times longer than the above intersection, gkg^{k} and sgks1sg^{k}s^{-1} freely generate a free group, FF. Also, its non-trivial elements hh are hyperbolic and

λ(h)λ(gk)10(2𝒟L(g)+100δ)10L(S).\lambda(h)\geq\lambda(g^{k})\geq 10(2\mathcal{D}L(g)+100\delta)\geq 10L(S).

For the last inequality, we used 2L(g)L(S)2L(g)\geq L(S).

We argue that there is a function D(ϵ)D^{\prime}(\epsilon) with D(100δ)=D(200δ)D^{\prime}(100\delta)=D(200\delta) such that every non-trivial hFh\in F is uniformly DD^{\prime}-WPD. We only need to argue for ϵ=100δ\epsilon=100\delta to check WPD by Lemma 2.4. Let γ\gamma be a 10δ10\delta-axis of hh. We will show that for any y,zγy,z\in\gamma with |yz|3λ(h)|y-z|\geq 3\lambda(h), there are at most D(200δ)D(200\delta) elements jGj\in G satisfying

|j(y)y|100δ and |j(z)z|100δ.|j(y)-y|\leq 100\delta\text{ and }|j(z)-z|\leq 100\delta.

We denote the collection of those elements jj by JJ. We will use the fact that the axis of hh is, roughly speaking, a concatenation of some translates of the segment [x,gk(x)][x,g^{k}(x)] by elements in FF (cf. a more precise description of the axis of the element uu below.) Also, we point out that without loss of generality, one may take a conjugate of hh in GG in the argument. Since hh is a word on gkg^{k} and sgks1sg^{k}s^{-1}, by taking a conjugate of hh in GG, one may assume that the segment [x,gk(x)][x,g^{k}(x)] is contained in the 30δ30\delta-neighborhood of the segment [y,z][y,z] except for some small neighborhood of xx and gk(x)g^{k}(x). Taking a further conjugate of hh if necessary, one may assume that the segment [x,gk/2(x)][x,g^{k/2}(x)] is contained in the 30δ30\delta-neighborhood of [y,z][y,z]. Here, we used |yz|3λ(h)|y-z|\geq 3\lambda(h).

That implies that one can find points p,q[x,gk/2(x)]p,q\in[x,g^{k/2}(x)] with |pq|k4λ(g)=15𝒟λ(g)|p-q|\geq\frac{k}{4}\lambda(g)=15\mathcal{D}\lambda(g) such that |pj(p)|150δ|p-j(p)|\leq 150\delta and |qj(q)|150δ|q-j(q)|\leq 150\delta if jJj\in J.

But since gg is DD-WPD where 𝒟=D(200δ)\mathcal{D}=D(200\delta), the set JJ contains at most D(200δ)D(200\delta) elements. (Strictly speaking the points pp and qq are maybe not exactly on a 10δ10\delta-axis of gg, but one can choose nearby points of pp and qq on the axis and argue.) We showed that hh is uniformly WPD.

Refer to caption
Figure 1. The thick line is α¯\bar{\alpha}. It is a broken geodesic from xx to u(x)u(x) with four short “gaps”. The first gap is [gk(x),gks(x)][g^{k}(x),g^{k}s(x)].

Now we argue for the moreover part. We define an element uu by

u=gk(sg10ks1)g20k(sgks1)gk.u=g^{k}(sg^{10k}s^{-1})g^{20k}(sg^{k}s^{-1})g^{k}.

(1) To construct an axis α\alpha of uu, consider the following (see Figure 1):

[x,gk(x)]\displaystyle[x,g^{k}(x)]\cup gks[x,g10k(x)]gk(sg10ks1)[x,g20k(x)]\displaystyle g^{k}s[x,g^{10k}(x)]\cup g^{k}(sg^{10k}s^{-1})[x,g^{20k}(x)]
\displaystyle\cup gk(sg10ks1)g20ks[x,gk(x)]gk(sg10ks1)g20k(sgks1)[x,gk(x)],\displaystyle g^{k}(sg^{10k}s^{-1})g^{20k}s[x,g^{k}(x)]\cup g^{k}(sg^{10k}s^{-1})g^{20k}(sg^{k}s^{-1})[x,g^{k}(x)],

which consists of five geodesic segments, which we call pieces. We name them as P1,,P5P_{1},\cdots,P_{5}. Their length are U,10U,20U,U,UU,10U,20U,U,U for some constant U>0U>0 maybe with some error up to 100δ100\delta. Remember that

Uλ(gk)10L(S).U\geq\lambda(g^{k})\geq 10L(S).

Note that P1P_{1} is contained in γ\gamma, which is the axis of gg. Put α1=γ\alpha_{1}=\gamma. Similarly, P2,,P5P_{2},\cdots,P_{5} are contained in the axes of the conjugates of gg by gks,gk(sg10ks1),gk(sg10ks1)g20ks,gk(sg10ks1)g20k(sgks1)g^{k}s,g^{k}(sg^{10k}s^{-1}),g^{k}(sg^{10k}s^{-1})g^{20k}s,g^{k}(sg^{10k}s^{-1})g^{20k}(sg^{k}s^{-1}), respectively. We will call those axes α2,α3,α4,α5\alpha_{2},\alpha_{3},\alpha_{4},\alpha_{5}.

There are four short (L(S)\leq L(S)) gaps between PiP_{i} and Pi+1P_{i+1}. We put (short) geodesics between the gaps and obtain a path, α¯\bar{\alpha}. Then take the union of the uu-orbit of α¯\bar{\alpha}, which we call α^\hat{\alpha}. This is an axis for uu since gg and sgs1sgs^{-1} are independent and their 10δ10\delta-axes stay close (50δ\leq 50\delta) to each other along a short (compared to UU) segment.

We call the image of a piece PiP_{i} by up(p)u^{p}(p\in\mathbb{Z}) also a piece. The axis α^\hat{\alpha} is a sequence of pieces (with short gaps in between).

By construction, xα^x\in\hat{\alpha}, but maybe α^\hat{\alpha} is not exactly a 10δ10\delta-axis for uu, so we take a 10δ10\delta-axis, α\alpha. One can check that d(x,α)20δd(x,\alpha)\leq 20\delta. (This is the reason we put gkg^{k} at the end of uu. Without gkg^{k} at the end, uu is still hyperbolic.) Also, α^\hat{\alpha} and α\alpha stay close to each other in the sense that most part of each piece in α^\hat{\alpha}, except for some short parts near the two ends, is in the 20δ20\delta-neighborhood of α\alpha.

(2) If two axes β,β\beta,\beta^{\prime} are parallel, we write ββ\beta\sim\beta^{\prime}. Since sE(g)s\not\in E(g), α1≁α2\alpha_{1}\not\sim\alpha_{2}. Also, α2≁α3\alpha_{2}\not\sim\alpha_{3}, α3≁α4\alpha_{3}\not\sim\alpha_{4}, and α4≁α5\alpha_{4}\not\sim\alpha_{5}.

We first argue that E(u)E(u) maps to =C\mathbb{Z}=C in the exact sequence after Proposition 2.6. Suppose hE(u)h\in E(u). By definition, αh(α)\alpha\sim h(\alpha). It suffices to show that hh does not flip the direction of α\alpha. For that, we examine the sequence of the lengths of the pieces on α\alpha and on h(α)h(\alpha). Here, we say that h(P)h(P) is a piece if PP is a piece in α\alpha.

The sequence for the part P1,,P5P_{1},\cdots,P_{5} is U,10U,20U,U,UU,10U,20U,U,U, so that on α\alpha, the sequence is (from the left to the right in Figure 2):

;U,10U,20U,U,U;U,10U,20U,U,U;\cdots;U,10U,20U,U,U;\,\,U,10U,20U,U,U;\cdots

Now, if the direction of h(α)h(\alpha) was opposite to α\alpha, then the sequence on h(α)h(\alpha) would be (from the left to the right in the figure):

;U,U,20U,10U,U;U,U,20U,10U,U;\cdots;U,U,20U,10U,U;\,\,U,U,20U,10U,U;\cdots
Refer to caption
Figure 2. The direction of α\alpha is to the right, and the direction of h(α)h(\alpha) is to the left. The figure indicates the three positions of PP.

From those two sequences we observe that one of the followings must hold:

  1. (1)

    there is a piece PP on h(α)h(\alpha) that intersects both N50δ(P1)N_{50\delta}(P_{1}) and N50δ(P2)N_{50\delta}(P_{2}) at least U2\frac{U}{2} in diameter, or

  2. (2)

    there is a piece PP on h(α)h(\alpha) that intersects both N50δ(P2)N_{50\delta}(P_{2}) and N50δ(P3)N_{50\delta}(P_{3}) at least U2\frac{U}{2} in diameter, or

  3. (3)

    there is a piece PP on h(α)h(\alpha) that intersects both N50δ(P3)N_{50\delta}(P_{3}) and N50δ(P4)N_{50\delta}(P_{4}) at least U2\frac{U}{2} in diameter.

In the above, PP has length (approximately) 5U5U or 10U10U.

Let β\beta be the axis that contains the piece PP. Then, since U2\frac{U}{2} is at least 5(2𝒟L(g)+100δ)5(2\mathcal{D}L(g)+100\delta), Lemma 2.5(2) would imply that either (1) α1βα2\alpha_{1}\sim\beta\sim\alpha_{2}, (2) α2βα3\alpha_{2}\sim\beta\sim\alpha_{3}, or (3) α3βα4\alpha_{3}\sim\beta\sim\alpha_{4}, respectively. In either case, we obtain a contradiction since α1≁α2\alpha_{1}\not\sim\alpha_{2}, α2≁α3\alpha_{2}\not\sim\alpha_{3} and α3≁α4\alpha_{3}\not\sim\alpha_{4}. We showed that hh does not flip α\alpha.

By the same reason, ie, the constrain from the combinatorics, h(x)h(x) must be close to up(x)u^{p}(x) for some pp\in\mathbb{Z}, namely, the distance is at most 2𝒟L(g)+100δ2\mathcal{D}L(g)+100\delta. This implies that α1(hup)(α1)\alpha_{1}\sim(hu^{-p})(\alpha_{1}) and also, α2(hup)(α2)\alpha_{2}\sim(hu^{-p})(\alpha_{2}). It then follows that

hupE(g),hupE(gnsgs1gn),hu^{-p}\in E(g),hu^{-p}\in E(g^{n}sgs^{-1}g^{-n}),

which implies gk(hup)gk(E(g)E(sgs1))g^{k}(hu^{-p})g^{-k}\in(E(g)\cap E(sgs^{-1})). But the right hand side is a finite group, so that huphu^{-p} has finite order, therefore hupF(g)hu^{-p}\in F(g). And, of course, hupF(u)hu^{-p}\in F(u) namely, there is fF(u)f\in F(u) with h=fuph=fu^{p}.

(3). This is a consequence of the DD^{\prime}-uniform WPD, and is mentioned in the paragraphs following Proposition 2.6.

(4). Since gS2g\in S^{2} and sSs\in S, we have uS14k+4u\in S^{14k+4}. We are done since m=66k+4m=66k+4. ∎

Proposition 2.10 (Lower bound of growth).

Suppose GG acts on a δ\delta-hyperbolic space XX. Let D(ϵ)D(\epsilon) be a function for WPD. Assume that there exists a constant MM such that for any finite generating set SS of GG, the set SMS^{M} contains a hyperbolic element hh that is DD-WPD.

Set

A=179220M(D(200δ))3>0.A=\frac{1}{79220M(D(200\delta))^{3}}>0.

Then for any finite generating set SS of GG, we have e(G,S)A|S|Ae(G,S)\geq A|S|^{A}.

This result is a generalization of the result on hyperbolic groups by [AL]. We adapt their argument to our setting, which is straightforward.

Proof.

Set 𝒟=D(200δ)\mathcal{D}=D(200\delta). Since the element hSMh\in S^{M} is DD-WPD, we have λ(h)50δ𝒟\lambda(h)\geq\frac{50\delta}{\mathcal{D}} by Lemma 2.2 (1). It implies that λ(h𝒟)50δ\lambda(h^{\mathcal{D}})\geq 50\delta. Since h𝒟SM𝒟h^{\mathcal{D}}\in S^{M\mathcal{D}}, we have L(SM𝒟)50δL(S^{M\mathcal{D}})\geq 50\delta.

Lemma 2.7 applies to SM𝒟S^{M\mathcal{D}} since L(SM𝒟)50δL(S^{M\mathcal{D}})\geq 50\delta. By Lemma 2.7 applied to SM𝒟S^{M\mathcal{D}} with xx such that L(SM𝒟)=L(SM𝒟,x)L(S^{M\mathcal{D}})=L(S^{M\mathcal{D}},x), there is gS2M𝒟g\in S^{2M\mathcal{D}} such that

L(SM𝒟)8δ|xg(x)|L(S^{M\mathcal{D}})-8\delta\leq|x-g(x)|

and the 10δ10\delta-axis γ\gamma of gg satisfies d(x,γ)20δd(x,\gamma)\leq 20\delta.

Then Lemma 2.8 applies to SM𝒟S^{M\mathcal{D}} and gS2M𝒟g\in S^{2M\mathcal{D}}. By Lemma 2.8 applied to SM𝒟S^{M\mathcal{D}} and gg, there exists sSM𝒟s\in S^{M\mathcal{D}} such that gk,sgks1=F\langle g^{k},sg^{k}s^{-1}\rangle=F and FF contains uu that is primitive such that

|F(u)|𝒟;uSM𝒟m;L(u)10L(SM𝒟),|F(u)|\leq\mathcal{D};\,u\in S^{M\mathcal{D}m};\,L(u)\geq 10L(S^{M\mathcal{D}}),

where k=60𝒟,m=3960𝒟+4k=60\mathcal{D},m=3960\mathcal{D}+4, and that uu is DD^{\prime}-WPD, where D(100δ)=D(200δ)D^{\prime}(100\delta)=D(200\delta).

We note that the element ss can be chosen from SS since GG is not virtually cyclic. This can be easily seen in the proof of Lemma 2.8 since we only need to choose ss such that sE(g)s\not\in E(g). See Remark 2.9.

Now, take a maximal subset WSW\subset S such that any two distinct elements w,vWw,v\in W are in different F(u)F(u)-(right) cosets. Then, |W||S|𝒟|W|\geq\frac{|S|}{\mathcal{D}}.

Let α\alpha be a 10δ10\delta-axis of uu with d(x,α)50δd(x,\alpha)\leq 50\delta.
Claim 1. For distinct v,wWv,w\in W, vα,wαv\alpha,w\alpha are not parallel. Indeed, if they were parallel, then w1vE(u)w^{-1}v\in E(u). Moreover, we will argue w1vF(u)w^{-1}v\in F(u), which is a contradiction since w,vw,v are in distinct F(u)F(u)-cosets. The reason for w1vF(u)w^{-1}v\in F(u) is that since w1vS2SM𝒟w^{-1}v\in S^{2}\subset S^{M\mathcal{D}}, we have

L(w1v)|w1v(x)x|L(SM𝒟).L(w^{-1}v)\leq|w^{-1}v(x)-x|\leq L(S^{M\mathcal{D}}).

But on the other hand, since uu is primitive, if w1vF(u)w^{-1}v\not\in F(u), then

L(w1v)L(u)100δ6L(SM𝒟).L(w^{-1}v)\geq L(u)-100\delta\geq 6L(S^{M\mathcal{D}}).

The last inequality is from L(u)10L(SM𝒟)L(u)\geq 10L(S^{M\mathcal{D}}). Those two estimates contradict. We showed the claim.

It implies that for distinct v,wWv,w\in W, the intersection of vαv\alpha and the 50δ50\delta-neighborhood of wαw\alpha is bounded by

2D(100δ)L(u)+100δ2D^{\prime}(100\delta)L(u)+100\delta

by Lemma2.5 (2) since uu is DD^{\prime}-WPD. Remember that D(100δ)=D(200δ)=𝒟D^{\prime}(100\delta)=D(200\delta)=\mathcal{D}. So, this bound is 2𝒟L(u)+100δ2\mathcal{D}L(u)+100\delta.

Set U=u20𝒟U=u^{20\mathcal{D}}. Then US20M𝒟2mU\in S^{20M\mathcal{D}^{2}m} and L(U)19𝒟L(u)L(U)\geq 19\mathcal{D}L(u) (maybe not quite 20𝒟L(u)20\mathcal{D}L(u)). α\alpha is an axis for UU as well. Set

B={wUw1|wW}.B=\{wUw^{-1}|w\in W\}.

Claim 2. We have |B|=|W||B|=|W| and BB freely generates a free group of rank |W||W|.

This is because for any wWw\in W, we have |xw(x)|L(SM𝒟)|x-w(x)|\leq L(S^{M\mathcal{D}}) since WSM𝒟W\subset S^{M\mathcal{D}}, which means the axis of wUw1wUw^{-1}, wαw\alpha, is close to xx. To be precise, close means that the distance is much smaller than L(u)L(u) since L(u)10L(SM𝒟)L(u)\geq 10L(S^{M\mathcal{D}}). Also, for the axes wαw\alpha and vαv\alpha of any distinct v,wWv,w\in W, the intersection of one with the 50δ50\delta-neighborhood of the other is 99 times shorter than L(U)L(U) since L(U)19𝒟L(u)L(U)\geq 19\mathcal{D}L(u). In this setting, the usual ping-pong argument shows the claim.

Since wSw\in S, we have B={wUw1|wW}S20M𝒟2m+2B=\{wUw^{-1}|w\in W\}\subset S^{20M\mathcal{D}^{2}m+2}. It follows that for any nn\in\mathbb{N},

|S(20M𝒟2m+2)n||Bn||B|n=|W|n|S|n𝒟n.|S^{(20M\mathcal{D}^{2}m+2)n}|\geq|B^{n}|\geq|B|^{n}=|W|^{n}\geq\frac{|S|^{n}}{\mathcal{D}^{n}}.

It implies

e(G,S)𝒟120M𝒟2m+2|S|120M𝒟2m+2.e(G,S)\geq\mathcal{D}^{-\frac{1}{20M\mathcal{D}^{2}m+2}}|S|^{\frac{1}{20M\mathcal{D}^{2}m+2}}.

Since

min{𝒟120M𝒟2m+2,120M𝒟2m+2}=120M𝒟2m+2,\min\left\{\mathcal{D}^{-\frac{1}{20M\mathcal{D}^{2}m+2}},\frac{1}{20M\mathcal{D}^{2}m+2}\right\}=\frac{1}{20M\mathcal{D}^{2}m+2},

which is at least 179220M𝒟3\frac{1}{79220M\mathcal{D}^{3}}, since m=3960𝒟+4m=3960\mathcal{D}+4. Setting

A=179220M𝒟3,A=\frac{1}{79220M\mathcal{D}^{3}},

we have e(G,S)A|S|Ae(G,S)\geq A|S|^{A}. This is a desired conclusion since 𝒟=D(200δ)\mathcal{D}=D(200\delta). ∎

Example 2.11.

Proposition 2.10 applies to the following examples:

  1. (1)

    Non-elementary hyperbolic groups (the original case in [AL]).

  2. (2)

    The mapping class groups of a compact orientable surface Σg,p\Sigma_{g,p} with 3g+p43g+p\geq 4. See Section 5.3, where the assumptions are checked.

  3. (3)

    A lattice in a simple Lie group of rank-1. See Theorem 5.5.

  4. (4)

    The fundamental group of a complete Riemannian manifold of finite volume whose sectional curvature is pinched by two negative constants. See Theorem 5.5.

3. Well-orderedness

3.1. Main theorem

We prove the following theorem. Note that Theorem 1.1 immediately follows from this theorem combined with Lemma 2.3, since the lemma says that an acylindrical action is uniformly WPD, so that every hyperbolic element is WPD.

Theorem 3.1 (Well-orderedness for uniform WPD actions).

Suppose GG acts on a δ\delta-hyperbolic space XX, and GG is not virtually cyclic. Let D(ϵ)D(\epsilon) be a function for WPD. Assume that there exists a constant MM such that for any finite generating set SS of GG, the set SMS^{M} contains a hyperbolic element on XX that is DD-WPD. Assume that GG is equationally Noetherian. Then, ξ(G)\xi(G) is a well-ordered set.

Proof.

We will prove that ξ(G)\xi(G) does not contain a strictly decreasing convergent sequence. To argue by contradiction, suppose that there exists a sequence of finite generating sets {Sn}\{S_{n}\}, such that {e(G,Sn)}\{e(G,S_{n})\} is a strictly decreasing sequence and limne(G,Sn)=d\lim_{n\to\infty}e(G,S_{n})=d, for some d>1d>1.

By Proposition 2.10, we may assume that the cardinality of the generating sets |Sn||S_{n}| from the decreasing sequence is bounded, and by possibly passing to a subsequence we may assume that the cardinality of the generating sets is fixed, |Sn|=|S_{n}|=\ell.

Let Sn={x1(n),,x(n)}S_{n}=\{x_{1}^{(n)},\cdots,x_{\ell}^{(n)}\}. Let FF be the free group of rank \ell with a free generating set: S={s1,,s}S=\{s_{1},\ldots,s_{\ell}\}. For each index nn, we define a map: fn:FGf_{n}:F\to G, by setting: fn(si)=xi(n)f_{n}(s_{i})=x_{i}^{(n)}. Since SnS_{n} are generating sets, the map fnf_{n} is an epimorphism for every nn. Note that e(G,Sn)=e(G,fn(S))e(G,S_{n})=e(G,f_{n}(S)).

Since FF is countable, the sequence {fn:FG}\{f_{n}:F\to G\} subconverges to a surjective homomorphism η:FL\eta:F\to L. LL is called a limit group over the group GG.

By assumption, GG is equationally Noetherian. By the general principle (Lemma 1.11), there exists an epimorphism hn:LGh_{n}:L\to G such that by passing to a subsequence we may assume that all the homomorphisms {fn}\{f_{n}\} factor through the limit epimorphism: η:FL\eta:F\to L, ie, fn=hnηf_{n}=h_{n}\circ\eta.

(F,S)\textstyle{(F,S)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}η\scriptstyle{\eta}fn\scriptstyle{f_{n}}(L,η(S))\textstyle{(L,\eta(S))\ignorespaces\ignorespaces\ignorespaces\ignorespaces}hn\scriptstyle{h_{n}}(G,fn(S))\textstyle{(G,f_{n}(S))}

Notice that since fn=hnηf_{n}=h_{n}\circ\eta for every index nn, we have e(G,fn(S))e(L,η(S))e(G,f_{n}(S))\leq e(L,\eta(S)).

We will show the following key result:

Proposition 3.2.

Suppose GG satisfies the assumption in Theorem 3.1. Let (L,η(S))(L,\eta(S)) be the limit group over GG of a sequence fn:(F,S)(G,fn(S))f_{n}:(F,S)\to(G,f_{n}(S)), where FF is a free group with a free generating set SS and fn(S)f_{n}(S) are generating sets of GG. Then limne(G,fn(S))=e(L,η(S))\lim_{n\to\infty}e(G,f_{n}(S))=e(L,\eta(S)).

We postpone proving this proposition until the next section and finish the proof of the theorem.

We assumed that the sequence {e(G,fn(S))}\{e(G,f_{n}(S))\} is strictly decreasing, hence, it can not converge to an upper bound of the sequence, e(L,η(S))e(L,\eta(S)). But the proposition says that it must converge to e(L,η(S))e(L,\eta(S)), a contradiction. Theorem 3.1 is proved. ∎

4. Continuity of the growth rate

We prove Proposition 3.2. We already know that e(L,η(S))e(G,fn(S))e(L,\eta(S))\geq e(G,f_{n}(S)) from the existence of the surjections hnh_{n}, which follows from that GG is equationally Noetherian.

It suffices to show that given ϵ>0\epsilon>0, for a large enough nn,

loge(L,η(S))ϵloge(G,fn(S)).\log e(L,\eta(S))-\epsilon\leq\log e(G,f_{n}(S)).

The strategy of the proof of this is same as [FS]. We note that from now on, we do not use that GG is equationally Noetherian in the proof.

Since the proof is long and complicated, we first informally describe the idea, which already appeared in [FS]. We want to show e(G,fn(S))e(G,f_{n}(S)) is almost equal to e(L,η(S))e(L,\eta(S)) for a large enough nn. First of all, if we take a large enough rr, then Br(L,η(S))B_{r}(L,\eta(S)) contains elements roughly as many as e(L,η(S))re(L,\eta(S))^{r} by the definition. Fix such rr. Then if we take nn large enough, Br(L,η(S))B_{r}(L,\eta(S)) and Br(G,fn(S))B_{r}(G,f_{n}(S)) are idential via the map hnh_{n} since LL is a limit group. But it does not mean that e(G,fn(S))e(G,f_{n}(S)) is almost equal to e(L,η(S))e(L,\eta(S)) since the growth of the balls in (G,fn(S))(G,f_{n}(S)) may decay if we take the radius larger. But it turns out that if we take rr large enough, then roughly speaking, the growth of the ball of radius rr in (G,fn(S))(G,f_{n}(S)) is almost equal to e(G,fn(S))e(G,f_{n}(S)). This is due to the well-known “local-to-global” principle in hyperbolic groups, and it is implemented by inserting “separators” in our argument. The threshold for the radius is given by mm in the proof (see Section 4.5).

We explain the idea more in detail. By the definition of the growth rate, we have for all rr,

loge(L,η(S))1rlog|Br(L,η(S))|.\log e(L,\eta(S))\leq\frac{1}{r}\log|B_{r}(L,\eta(S))|.

This is because the sequence {log|Br(L,η(S))|}\{\log|B_{r}(L,\eta(S))|\} is sub-additive.

Fix rr (we will choose rr sufficiently large in the argument we will give later). Then choose nn large enough such that hn:Br(L,η(S))Br(G,fn(S))h_{n}:B_{r}(L,\eta(S))\to B_{r}(G,f_{n}(S)) is injective. The following map is naturally induced from hnh_{n} for each qq\in\mathbb{N}:

Br(L,η(S))qBqr(G,fn(S))GB_{r}(L,\eta(S))^{q}\to B_{qr}(G,f_{n}(S))\subset G

by mapping (w1,,wq)(w_{1},\cdots,w_{q}) to hn(w1wq)h_{n}(w_{1}\cdots w_{q}). If there is an rr such that this map is injective for all qq, then an easy computation would show the desired inequality for the nn by letting qq\to\infty.

But of course this map is not injective in general. For example, the concatenation of hn(w1),hn(w2),,hn(wq)h_{n}(w_{1}),h_{n}(w_{2}),\cdots,h_{n}(w_{q}) may have lots of backtracks at the concatenation points. As a remedy, we insert elements uiu_{i} of bounded length, called separators, and define a new map sending (w1,,wq)(w_{1},\cdots,w_{q}) to hn(w1u1w2u2wquq)h_{n}(w_{1}u_{1}w_{2}u_{2}\cdots w_{q}u_{q}). This map is denoted by Φn\Phi_{n}. The separators are constructed in Lemma 4.2. We arrange that the concatenation of elements, after we insert hn(ui)h_{n}(u_{i})’s, is a uniform quasi-geodesic in GG. This follows from that GG is a hyperbolic group.

But it is still not the case that the map Φn:Br(L,η(S))qBq(r+b)(G,fn(S))\Phi_{n}:B_{r}(L,\eta(S))^{q}\to B_{q(r+b)}(G,f_{n}(S)) is injective, where bb is the bound of the length of the separators. What we actually show is that Φn\Phi_{n} is injective if we restrict it to the qq-tuples in some fixed portion of Br(L,η(S))B_{r}(L,\eta(S)), (see Lemma 4.8), which is enough for our purpose. This part is very technical. To argue that Φn\Phi_{n} is injective on the certain fixed portion, we use the action of GG on XX, ie, we map a tuple by Φn\Phi_{n} to GG, then let it act on XX. Then we analyze the orbit of a base point in XX.

4.1. Separators

We review the setting. XX is a δ\delta-hyperbolic space, and GG acts on it. We assume δ1\delta\geq 1. For each nn, SnS_{n} is a finite generating set of GG such that SnMS_{n}^{M} contains a hyperbolic element that is DD-WPD. Using the homomorphism hn:LGh_{n}:L\to G, we let LL act on XX. We first construct separators as elements in GG then pull them back to LL by hnh_{n}.

Let gSnMg\in S_{n}^{M} be a hyperbolic element that is DD-WPD. Set

𝒟=D(200δ).\mathcal{D}=D(200\delta).

Then λ(g)50δ𝒟\lambda(g)\geq\frac{50\delta}{\mathcal{D}} by Lemma 2.2(1). It implies

100δλ(g2𝒟)L(Sn2𝒟M).100\delta\leq\lambda(g^{2\mathcal{D}})\leq L(S_{n}^{2\mathcal{D}M}).

Fix nn. Let ynXy_{n}\in X be a point where L(Sn2𝒟M)L(S_{n}^{2\mathcal{D}M}) is achieved. Put

Δn=100δ+4𝒟L(Sn2𝒟M).\Delta_{n}=100\delta+4\mathcal{D}L(S_{n}^{2\mathcal{D}M}).

We define a germ w.r.t. the constant Δn\Delta_{n}. Recall that given three points x,y,zXx,y,z\in X, the Gromov product, (y,z)x(y,z)_{x}, is defined as follows:

(y,z)x=|xy|+|xz||yz|2.(y,z)_{x}=\frac{|x-y|+|x-z|-|y-z|}{2}.
Definition 4.1 (germs, equivalent and opposite germs).

Let [x,y][x,y] be a (directed) geodesic segment in XX. Suppose that |xy|10Δn|x-y|\geq 10\Delta_{n}. Then, the initial segment of [x,y][x,y] of length 10Δn10\Delta_{n} is called the germ of [x,y][x,y] at xx, denoted by germ([x,y])germ([x,y]). If |xy|<10Δn|x-y|<10\Delta_{n}, then we define the germ to be empty.

We say two non-empty germs, [x,y],[x,z][x,y],[x,z], at a common point xx, are equivalent if (y,z)x4Δn(y,z)_{x}\geq 4\Delta_{n}, and opposite if (y,z)x2Δn(y,z)_{x}\leq 2\Delta_{n}.

We sometimes call the germ of [y,x][y,x] at yy as the germ of [x,y][x,y] at yy. If γ\gamma is the germ of [x,y][x,y] at xx, then for gGg\in G, the segment g(γ)g(\gamma) is the germ of [g(x),g(y)][g(x),g(y)] at g(x)g(x). For gGg\in G, we consider the germ of [yn,g(yn)][y_{n},g(y_{n})] and call it the germ of gg at yny_{n}, and write germ (g)\text{germ\,}(g).

Recall from Lemma 2.8 that k=60𝒟,m=3964𝒟+4k=60\mathcal{D},m=3964\mathcal{D}+4.

We consider germs w.r.t. Δn\Delta_{n}.

Lemma 4.2 (The constant bb and separators. c.f. Lemma 2.4 [FS]).

There exists a constant bb with the following property, where bb depends only on δ,M\delta,M and D(ϵ)D(\epsilon). For every nn, there exist primitive, hyperbolic elements u1,u2,u3,u4Snbu_{1},u_{2},u_{3},u_{4}\in S_{n}^{b}; and mutually opposite germs γ1,γ2,γ3,γ4\gamma_{1},\gamma_{2},\gamma_{3},\gamma_{4} of some elements in Sn2𝒟MS_{n}^{2\mathcal{D}M} at yny_{n} that satisfy:

  1. (i)

    The germs are all at yny_{n} in the following.

    • The germ of [yn,u1(yn)][y_{n},u_{1}(y_{n})] is equivalent to γ1\gamma_{1} . The germ of [yn,u11(yn)][y_{n},u_{1}^{-1}(y_{n})] is equivalent to γ3\gamma_{3}.

    • The germ of [yn,u2(yn)][y_{n},u_{2}(y_{n})] is equivalent to γ1\gamma_{1}. The germ of [yn,u21(yn)][y_{n},u_{2}^{-1}(y_{n})] is equivalent to γ4\gamma_{4}.

    • The germ of [yn,u3(yn)][y_{n},u_{3}(y_{n})] is equivalent to γ2\gamma_{2}. The germ of [yn,u31(yn)][y_{n},u_{3}^{-1}(y_{n})] is equivalent to γ3\gamma_{3}.

    • The germ of [yn,u4(yn)][y_{n},u_{4}(y_{n})] is equivalent to γ2\gamma_{2}. The germ of [yn,u41(yn)][y_{n},u_{4}^{-1}(y_{n})] is equivalent to γ4\gamma_{4}.

  2. (ii)

    For all ii,

    λ(ui)100Δn,\lambda(u_{i})\geq 100\Delta_{n},

    and the distance from yny_{n} to the 10δ10\delta-axis of uiu_{i} is at most Δn\Delta_{n}.

  3. (iii)

    For every wGw\in G, and all i,ji,j (possibly i=ji=j), if the 20δ20\delta-neighborhood of [yn,ui(yn)][y_{n},u_{i}(y_{n})] intersects the segment [w(yn),wuj(yn)][w(y_{n}),wu_{j}(y_{n})], then the diameter of the intersection is bounded by: 110d(yn,ui(yn))\frac{1}{10}d(y_{n},u_{i}(y_{n})) and 110d(yn,uj(yn))\frac{1}{10}d(y_{n},u_{j}(y_{n})). If i=ji=j, we assume in addition that wF(ui)w\not\in F(u_{i}), where F(ui)F(u_{i}) is the finite normal subgroup in E(ui)E(u_{i}).

Remark 4.3.

(1) Regarding Lemma 4.2 (iii), if wF(ui)w\in F(u_{i}) then the Hausdorff-distance between [yn,ui(yn)][y_{n},u_{i}(y_{n})] and [w(yn),wui(yn)][w(y_{n}),wu_{i}(y_{n})] is at most Δn+100δ\Delta_{n}+100\delta since ww moves any point on a 10δ10\delta-axis of uiu_{i} by at most 50δ50\delta since otherwise ww would be hyperbolic.
(2) As we will see in the proof it suffices to take bb to be

(2) b=343440M(D(200δ))2+22.b=343440M(D(200\delta))^{2}+22.

The elements uiu_{i} are called separators and the property (iii) is called the small cancellation property of separators.

Note that we have L(Sn2M𝒟)L(Sn)L(S_{n}^{2M\mathcal{D}})\geq L(S_{n}).

Proof.

Since L(Sn2M𝒟)100δL(S_{n}^{2M\mathcal{D}})\geq 100\delta and yny_{n} is a point where L(Sn2M𝒟)L(S_{n}^{2M\mathcal{D}}) is achieved, by Lemma 2.7 (also see the properties at the bullets after the lemma), there is a hyperbolic element gSn4M𝒟g\in S_{n}^{4M\mathcal{D}} such that

2L(Sn2M𝒟)L(g)L(Sn2M𝒟)24δ12L(Sn2M𝒟)2L(S_{n}^{2M\mathcal{D}})\geq L(g)\geq L(S_{n}^{2M\mathcal{D}})-24\delta\geq\frac{1}{2}L(S_{n}^{2M\mathcal{D}})

and the 10δ10\delta-axis of gg, γ\gamma, is at distance at most 10δ10\delta from yny_{n}.

By Lemma 2.8 applied to Sn2M𝒟S_{n}^{2M\mathcal{D}} and gg and the 10δ10\delta-axis γ\gamma, there exists sSns\in S_{n} (see the remark 2.9) such that gkg^{k} and sgks1sg^{k}s^{-1} are independent hyperbolic elements, which freely generate a free group FF whose non-trivial element, hh, satisfies

λ(h)10(2𝒟L(g)+100δ)5Δn.\lambda(h)\geq 10(2\mathcal{D}L(g)+100\delta)\geq 5\Delta_{n}.

Recall that k=60D(200δ)=60𝒟k=60D(200\delta)=60\mathcal{D}. Note that gk,sgks1Sn4M𝒟k+2g^{k},sg^{k}s^{-1}\in S_{n}^{4M\mathcal{D}k+2}. The distance from yny_{n} to sγs\gamma is at most 40δ+L(Sn2M𝒟)40\delta+L(S_{n}^{2M\mathcal{D}}). Also, the intersection of γ\gamma and the 50δ50\delta-neighborhood of sγs\gamma is at most 2𝒟L(g)+100δ2\mathcal{D}L(g)+100\delta in length, which is Δn\leq\Delta_{n}, by Lemma 2.5 since γ\gamma and sγs\gamma are not parallel.

Consider the following four germs at yny_{n} w.r.t. the constant Δn\Delta_{n}:

γ1=germ (gk),γ2=germ (sgks1),γ3=germ (gk),γ4=germ (sgks1).\gamma_{1}=\text{germ\,}(g^{k}),\gamma_{2}=\text{germ\,}(sg^{k}s^{-1}),\gamma_{3}=\text{germ\,}(g^{-k}),\gamma_{4}=\text{germ\,}(sg^{-k}s^{-1}).

Note that any two of them are mutually opposite.

Then, there exist separators uiFu_{i}\in F such that uiSnbu_{i}\in S_{n}^{b}, where

b=14314𝒟Mk+22=343440𝒟2M+22.b=1431\cdot 4\mathcal{D}Mk+22=343440\mathcal{D}^{2}M+22.

For example, set w=gk,z=sgks1w=g^{k},z=sg^{k}s^{-1} and:

u1\displaystyle u_{1} =wzw2zw3zw19zw20,\displaystyle=wzw^{2}zw^{3}z\cdots w^{19}zw^{20},
u2\displaystyle u_{2} =w21zw22zw39zw40z,\displaystyle=w^{21}zw^{22}z\cdots w^{39}zw^{40}z,
u3\displaystyle u_{3} =zw41zw42zw43zw59zw60,\displaystyle=zw^{41}zw^{42}zw^{43}z\cdots w^{59}zw^{60},
u4\displaystyle u_{4} =zw61zw62zw63zw79zw80z.\displaystyle=zw^{61}zw^{62}zw^{63}z\cdots w^{79}zw^{80}z.

We compute that they are in SnbS_{n}^{b} since wSn4𝒟Mk,zSn4𝒟Mk+2w\in S_{n}^{4\mathcal{D}Mk},z\in S_{n}^{4\mathcal{D}Mk+2}. It will be important that this number bb does not depend on SnS_{n}. (See the proof of Proposition 3.2.)

Because of the combinatorial reason, they are primitive hyperbolic elements. The argument is similar to the one we used to showed that uu is primitive hyperbolic, using γ\gamma and sγs\gamma are not parallel, in the proof of Lemma 2.8(2). We omit it.

(i) By definition of uiu_{i},

germ [yn,u1(yn)]=germ [yn,u2(yn)]=germ (w)=γ1;\text{germ\,}[y_{n},u_{1}(y_{n})]=\text{germ\,}[y_{n},u_{2}(y_{n})]=\text{germ\,}(w)=\gamma_{1};
germ [yn,u3(yn)]=germ [yn,u4(yn)]=germ (z)=γ2;\text{germ\,}[y_{n},u_{3}(y_{n})]=\text{germ\,}[y_{n},u_{4}(y_{n})]=\text{germ\,}(z)=\gamma_{2};
u11germ [u1(yn),yn]=u31germ [u3(yn),yn]=γ3;u_{1}^{-1}\text{germ\,}[u_{1}(y_{n}),y_{n}]=u_{3}^{-1}\text{germ\,}[u_{3}(y_{n}),y_{n}]=\gamma_{3};
u21germ [u2(yn),yn]=u41germ [u4(yn),yn]=γ4.u_{2}^{-1}\text{germ\,}[u_{2}(y_{n}),y_{n}]=u_{4}^{-1}\text{germ\,}[u_{4}(y_{n}),y_{n}]=\gamma_{4}.

Here, we mean that the four germs in each line are same or equivalent to each other.

(ii). The estimate for λ(ui)\lambda(u_{i}) is straightforward from the definitions of uiu_{i}. The claim on the axes are also shown similarly to the case of uu in Lemma 2.8(1) and we omit it. We remark that the distance estimate on the axes differ since it comes from the fact that some of the separators start or end in ww, so that distance becomes larger.

(iii) follows from the definition of uiu_{i}, namely, the combinatorial structure of the words. The argument is similar to show that uu in Lemma 2.8 is primitive (Lemma 2.8(2)), so we will be brief. Suppose iji\not=j, and the intersection was longer. Then, because of the combinatorial structure of the words uiu_{i} and uju_{j}, it follows the axes γ\gamma and sγs\gamma would be parallel, by Lemma 2.5, which is impossible. We are done. Suppose i=ji=j, and the intersection was longer. Then since wF(ui)w\not\in F(u_{i}), by the same reason, γ\gamma and sγs\gamma would be parallel, a contradiction. So we are done in this case too. ∎

Remember uiGu_{i}\in G depend on the index nn of SnS_{n}, so let’s write them as ui(n)u_{i}(n). Now, since hnh_{n} is surjective, let u^i(n)L\hat{u}_{i}(n)\in L be an element with hn(u^i(n))=ui(n)h_{n}(\hat{u}_{i}(n))=u_{i}(n). Note that the word length of u^i(n)\hat{u}_{i}(n) in terms of η(S)\eta(S) is also bounded by bb. The elements u^i(n)L\hat{u}_{i}(n)\in L are also called the separators for hnh_{n}.

In the following we may just write uu (instead of u^\hat{u}) to denote a separator for hnh_{n} to simplify the notation. We note that |F(hn(u))|𝒟|F(h_{n}(u))|\leq\mathcal{D} for any separator uu for any hnh_{n} since hn(u)h_{n}(u) is primitive (see the comment after Proposition 2.6).

4.2. Forbidden elements

Given mm, choose and fix nn large enough such that the map hnh_{n} is injective on B2m(L,η(S))B_{2m}(L,\eta(S)). This is possible since this set is finite in the limit group LL. Remember

Δn=100δ+4𝒟L(Sn2𝒟M).\Delta_{n}=100\delta+4\mathcal{D}L(S_{n}^{2\mathcal{D}M}).

Given wBm(L,η(S))w\in B_{m}(L,\eta(S)) we choose one of the separators, uLu\in L, for hnh_{n} s.t. the germ of [yn,hn(w)(yn)][y_{n},h_{n}(w)(y_{n})] at hn(w)(yn)h_{n}(w)(y_{n}) and the germ of [hn(w)(yn),hn(wu)(yn)][h_{n}(w)(y_{n}),h_{n}(wu)(y_{n})] at hn(w)(yn)h_{n}(w)(y_{n}) are opposite. (Geometrically, it implies that the concatenation [yn,hn(w)(yn)][hn(w)(yn),hn(wu)(yn)][y_{n},h_{n}(w)(y_{n})]\cup[h_{n}(w)(y_{n}),h_{n}(wu)(y_{n})] is almost a geodesic.) Such a separator uu exists by the property (i) in Lemma 4.2. Here, if |ynhn(w)(yn)|<10Δn|y_{n}-h_{n}(w)(y_{n})|<10\Delta_{n}, then we choose any separator uu. We say uu is admissible for ww. Note that

|ynhn(wu)(yn)||ynhn(w)(yn)|+|hn(w)(yn)hn(wu)(yn)|4Δn.|y_{n}-h_{n}(wu)(y_{n})|\geq|y_{n}-h_{n}(w)(y_{n})|+|h_{n}(w)(y_{n})-h_{n}(wu)(y_{n})|-4\Delta_{n}.

Also, the Hausdorff distance between [yn,hn(wu)(yn)][y_{n},h_{n}(wu)(y_{n})] and [yn,hn(w)(yn)][hn(w)(yn),hn(wu)(yn)][y_{n},h_{n}(w)(y_{n})]\cup[h_{n}(w)(y_{n}),h_{n}(wu)(y_{n})] is at most 2Δn+10δ2\Delta_{n}+10\delta.

Moreover, given w,wBmw,w^{\prime}\in B_{m} then we can choose a separator uu such that uu is admissible for ww and u1u^{-1} is admissible for w1w^{\prime-1}. We say uu is admissible for w,ww,w^{\prime}. Note that

|ynhn(wuw)(yn)|\displaystyle|y_{n}-h_{n}(wuw^{\prime})(y_{n})|\geq |ynhn(w)(yn)|+|hn(w)(yn)hn(wu)(yn)|\displaystyle|y_{n}-h_{n}(w)(y_{n})|+|h_{n}(w)(y_{n})-h_{n}(wu)(y_{n})|
+|hn(wu)(yn)hn(wuw)(yn)|8Δn.\displaystyle+|h_{n}(wu)(y_{n})-h_{n}(wuw^{\prime})(y_{n})|-8\Delta_{n}.

As before, the Hausdorff distance between [yn,hn(wuw)(yn)][y_{n},h_{n}(wuw^{\prime})(y_{n})] and [yn,hn(w)(yn)][hn(w)(yn),hn(wu)(yn)][hn(wu)(yn),hn(wuw)(yn)][y_{n},h_{n}(w)(y_{n})]\cup[h_{n}(w)(y_{n}),h_{n}(wu)(y_{n})]\cup[h_{n}(wu)(y_{n}),h_{n}(wuw^{\prime})(y_{n})] is at most 2Δn+10δ2\Delta_{n}+10\delta.

For q>0q>0, we define a map

Φn:Bm(L,η(S))qBq(m+b)(G,fn(S))G\Phi_{n}:B_{m}(L,\eta(S))^{q}\to B_{q(m+b)}(G,f_{n}(S))\subset G

by sending (w1,,wq)(w_{1},\cdots,w_{q}) to hn(w1u1wquq)h_{n}(w_{1}u_{1}\cdots w_{q}u_{q}), where uiLu_{i}\in L are separators we choose that are admissible for wi,wi+1w_{i},w_{i+1} for hnh_{n}. Remember that uiBb(L,η(S))u_{i}\in B_{b}(L,\eta(S)).

It is easy to show that Φn\Phi_{n} maps an element (1,,1)\not=(1,\cdots,1) to a non-trivial element in GG using the property of the separators, but what we need is more. We will argue that on a large portion, called the set of “feasible elements”, Φn\Phi_{n} is injective by showing the image of the base point ynXy_{n}\in X by those elements are all distinct.

For the given mm, we will define forbidden elements in Bm(L,η(S))B_{m}(L,\eta(S)), which depend on nn. We are assuming that nn is large enough so that hnh_{n} is injective on B2m(L,η(S))B_{2m}(L,\eta(S)).

Definition 4.4 (forbidden elements and tails).

Given wBm(L,η(S))w\in B_{m}(L,\eta(S)), if there exist wBm(L,η(S))w^{\prime}\in B_{m}(L,\eta(S)) and a separator u(n)u(n) which is admissible for ww such that

(3) |hn(w)(yn)hn(wu(n))(yn)|15|ynhn(u(n))(yn)||h_{n}(w^{\prime})(y_{n})-h_{n}(wu(n))(y_{n})|\leq\frac{1}{5}|y_{n}-h_{n}(u(n))(y_{n})|

then we say ww is forbidden w.r.t. nn (or in terms of hn)h_{n}). We call the segment [hn(w)(yn),hn(wu(n))(yn)][h_{n}(w)(y_{n}),h_{n}(wu(n))(y_{n})] a tail of ww. The tail depends on the choice of uu, so if we want to specify it, we say the tail of the pair (w,u)(w,u).

Let w1,w2w_{1},w_{2} be two (forbidden) elements and u1,u2u_{1},u_{2} admissible separators , respectively. If u1=u2u_{1}=u_{2} and hn(w11w2)F(hn(u1))h_{n}(w_{1}^{-1}w_{2})\in F(h_{n}(u_{1})), then the Hausdorff-distance between the two tails [hn(w1)(yn),hn(w1u)(yn)][h_{n}(w_{1})(y_{n}),h_{n}(w_{1}u)(y_{n})] and [hn(w2)(yn),hn(w2u)(yn)][h_{n}(w_{2})(y_{n}),h_{n}(w_{2}u)(y_{n})] is at most Δn+100δ\Delta_{n}+100\delta. This is an immediate consequence of Remark 4.3. In this case, we say that the tails of (w1,u1)(w_{1},u_{1}) and (w2,u2)(w_{2},u_{2}) are parallel.

On the other hand, if two tails (w1,u1)(w_{1},u_{1}) and (w2,u2)(w_{2},u_{2}) are not parallel, then the intersection of one of the tails with the 20δ20\delta-neighborhood of the other tail is bounded by the 110\frac{1}{10} of the length of each tail. This is by the small cancellation property of the separators (Lemma 4.2 (iii)),

We record one immediate consequence we use later.

Lemma 4.5 (Parallel tails).

Assume that hnh_{n} is injective on B2m(L,η(S))B_{2m}(L,\eta(S)). Suppose wBm(L,η(S))w\in B_{m}(L,\eta(S)) is forbidden w.r.t. nn, and uu is a separator admissible for ww. Then there are at most 𝒟\mathcal{D} possibilities for w1Bmw_{1}\in B_{m}, including w=w1w=w_{1}, such that w1w_{1} is forbidden, uu is admissible for w1w_{1}, and the tails for (w,u)(w,u) and (w1,u)(w_{1},u) are parallel.

Proof.

Since the two tails are parallel, we have hn(w11w)F(hn(u))h_{n}(w_{1}^{-1}w)\in F(h_{n}(u)) by the definition that two tails are parallel. Recall that (see the paragraphs after Proposition 2.6.)

|F(hn(u))|𝒟|F(h_{n}(u))|\leq\mathcal{D}

for all separators uu since uu is primitive. Therefore, we have at most 𝒟\mathcal{D} possibilities for hn(w11w)h_{n}(w_{1}^{-1}w). But since hnh_{n} is injective on B2m(L,η(S))B_{2m}(L,\eta(S)), we have at most 𝒟\mathcal{D} possibilities for w11wB2m(L,η(S))w_{1}^{-1}w\in B_{2m}(L,\eta(S)), and we are done. ∎

4.3. Ratio of forbidden elements

The proof of the following lemma occupies this subsection.

Lemma 4.6.

Assume that hnh_{n} is injective on B2m(L,η(S))B_{2m}(L,\eta(S)). Consider the forbidden/non-forbidden elements in Bm(L,η(S))B_{m}(L,\eta(S)) w.r.t. nn. Then

|{the forbidden elements}|𝒟|{the non-forbidden elements}|.|\{\text{the forbidden elements}\}|\leq\mathcal{D}|\{\text{the non-forbidden elements}\}|.

We denote Bm(L,η(S))B_{m}(L,\eta(S)) as BmB_{m}. The strategy of the proof is to show that if wBmw\in B_{m} is forbidden, it will force some other elements in BmB_{m} to be non-forbidden.

Proof.

The proof has three parts.
Part I. We first construct a subset C(w)C(w) in BmB_{m} and a tree like graph T(w)T(w) in XX. The construction is inductive. In each step, a subset and a tree like graph grows, which will end in finite steps.

Suppose wBmw\in B_{m} is forbidden. We explain the inductive steps to construct C(w)C(w) and T(w)T(w).
Step 0. Set C0(w)=wC_{0}(w)=w and T0(w)=[yn,hn(w)(yn)]T_{0}(w)=[y_{n},h_{n}(w)(y_{n})].
Step 1. Since ww is forbidden, there exist wBmw^{\prime}\in B_{m} and a separator uu that is admissible for ww that satisfy (3). Let w=s1sr,siη(S),rmw^{\prime}=s_{1}\cdots s_{r},s_{i}\in\eta(S),r\leq m be a shortest representative w.r.t. the word metric by η(S)\eta(S). Then this defines a sequence of points in XX, which we call a path, γ\gamma, from yny_{n} to hn(w)(yn)h_{n}(w^{\prime})(y_{n}) as follows:

yn,hn(s1)(yn),hn(s1s2)(yn),,hn(s1sr)(yn).y_{n},h_{n}(s_{1})(y_{n}),h_{n}(s_{1}s_{2})(y_{n}),\cdots,h_{n}(s_{1}\cdots s_{r})(y_{n}).

The distance between any two adjacent points on γ\gamma is at most L(Sn2MD)L(S_{n}^{2MD}) since it is achieved at yny_{n} and hn(si)SnSn2MDh_{n}(s_{i})\in S_{n}\subset S_{n}^{2MD}.

Consider the nearest point projection, denoted by π\pi, from a point xγx\in\gamma to the tail [hn(w)(yn),hn(wu)(yn)][h_{n}(w)(y_{n}),h_{n}(wu)(y_{n})]. The nearest points may not be unique, but we choose one for π(x)\pi(x). Then the distance between the projection of any two adjacent points on γ\gamma is L(Sn2MD)+100δΔn4\leq L(S_{n}^{2MD})+100\delta\leq\frac{\Delta_{n}}{4}. Note that |π(yn)hn(w)(yn)|2Δn+150δ|\pi(y_{n})-h_{n}(w)(y_{n})|\leq 2\Delta_{n}+150\delta since uu is admissible for ww. Also, |π(hn(w)(yn))hn(wu)(yn)|15|hn(w)(yn)hn(wu)(yn)|+10δ|\pi(h_{n}(w^{\prime})(y_{n}))-h_{n}(wu)(y_{n})|\leq\frac{1}{5}|h_{n}(w)(y_{n})-h_{n}(wu)(y_{n})|+10\delta by (3).

By Lemma 4.2 (ii), the length of a tail is at least 100Δn100\Delta_{n}. Let P,QP,Q be the two points on the tail that trisect the tail into three pieces of equal length, where PP is closer to hn(w)(yn)h_{n}(w)(y_{n}) than QQ is. Each of the three pieces has lengh at least 33Δn33\Delta_{n}. Let hn(s1sp)(yn)h_{n}(s_{1}\cdots s_{p})(y_{n}) be a point on γ\gamma whose projection is closest to PP, and hn(s1sq)(yn)h_{n}(s_{1}\cdots s_{q})(y_{n}) a point whose projection is closest to QQ. Then

|π(hn(s1sp)(yn))P|L(Sn2MD)Δn4,|\pi(h_{n}(s_{1}\cdots s_{p})(y_{n}))-P|\leq L(S_{n}^{2MD})\leq\frac{\Delta_{n}}{4},
|π(hn(s1sq)(yn))Q|L(Sn2MD)Δn4.|\pi(h_{n}(s_{1}\cdots s_{q})(y_{n}))-Q|\leq L(S_{n}^{2MD})\leq\frac{\Delta_{n}}{4}.

We denote s1sp,s1sqBms_{1}\cdots s_{p},s_{1}\cdots s_{q}\in B_{m} as w0,w1w_{0},w_{1} and call them the candidates for non-forbidden elements obtained from ww. They depend on u,wu,w^{\prime} too. See Figure 3.

Refer to caption
Figure 3. If ww is forbidden, then we have two new candidates w¯,w~\bar{w},\tilde{w} for non-forbidden elements.

Consider the union of the three segements as follows:

Π(w)=[hn(w)(yn),hn(wu)(yn)]\displaystyle\Pi(w)=[h_{n}(w)(y_{n}),h_{n}(wu)(y_{n})] [hn(w0)(yn),π(hn(w0)(yn))]\displaystyle\cup[h_{n}(w_{0})(y_{n}),\pi(h_{n}(w_{0})(y_{n}))]
[hn(w1)(yn),π(hn(w1)(yn))].\displaystyle\cup[h_{n}(w_{1})(y_{n}),\pi(h_{n}(w_{1})(y_{n}))].

This depends not only ww, but also u,wu,w^{\prime}. Π(w)\Pi(w) is a tree embedded in XX. We call π(hn(w0)(yn))\pi(h_{n}(w_{0})(y_{n})) and π(hn(w1)(yn))\pi(h_{n}(w_{1})(y_{n})) the branch points of Π(w)\Pi(w), and P,QP,Q the trisecting points of (the tail of) Π(w)\Pi(w). This union is a disjoint union except for the two branch points. The distance betwee the branch points is at least 32Δn32\Delta_{n}. This in particular implies that w0w1w_{0}\not=w_{1}.

We set C1(w)={w,w0,w1}BmC_{1}(w)=\{w,w_{0},w_{1}\}\subset B_{m}, and T1(w)=[yn,hn(w)(yn)]Π(w)T_{1}(w)=[y_{n},h_{n}(w)(y_{n})]\cup\Pi(w). If both w0,w1w_{0},w_{1} are non-forbidden, this is the end of the construction, and put C(w)=C1(w),T(w)=T1(w)C(w)=C_{1}(w),T(w)=T_{1}(w). Otherwise we go to the second step.

T1(w)T_{1}(w) is a tree like graph in the sense that it is the union of two trees, [yn,hn(w)(yn)][y_{n},h_{n}(w)(y_{n})] and Π(w)\Pi(w) attached at the point hn(w)(yn)h_{n}(w)(y_{n}), and the intersection is contained in the (2Δn+100δ)(2\Delta_{n}+100\delta)-neighborhood of this point since uu is admissible for ww.

Step 2. By assumption, at least one of w0w_{0} and w1w_{1} is forbidden. For example, assume that w1w_{1} is forbidden. Then w1w_{1} has a separator u1u_{1} admissible for w1w_{1}, and an element w1Bmw_{1}^{\prime}\in B_{m} that satisfy (3). Let π1\pi_{1} denote the nearest point projection to the tail [hn(w1)(yn),hn(w1u1)(yn)][h_{n}(w_{1})(y_{n}),h_{n}(w_{1}u_{1})(y_{n})].

Then as in the first step, there is a path (a sequence of points obtained from a shortest expression of w1w_{1}^{\prime} in η(S)\eta(S)) between yny_{n} and hn(w1)(yn)h_{n}(w_{1}^{\prime})(y_{n}) in XX, from which we obtain two elements, w10,w11Bmw_{10},w_{11}\in B_{m}, using the projection π1\pi_{1} from the path to the tail [hn(w1)(yn),hn(w1u1)(yn)][h_{n}(w_{1})(y_{n}),h_{n}(w_{1}u_{1})(y_{n})]. Also, we obtain a tree embedded in XX:

Π(w1)=[hn(w1)(yn),hn(w1u1)(yn)]\displaystyle\Pi(w_{1})=[h_{n}(w_{1})(y_{n}),h_{n}(w_{1}u_{1})(y_{n})] [hn(w10)(yn),π(hn(w10)(yn))]\displaystyle\cup[h_{n}(w_{10})(y_{n}),\pi(h_{n}(w_{10})(y_{n}))]
[hn(w11)(yn),π(hn(w11)(yn))].\displaystyle\cup[h_{n}(w_{11})(y_{n}),\pi(h_{n}(w_{11})(y_{n}))].

We point out that ([yn,hn(w)(yn)]Π(w))N10δ(Π(w1))([y_{n},h_{n}(w)(y_{n})]\cup\Pi(w))\cap N_{10\delta}(\Pi(w_{1})) is contained in the (2Δn+100δ)(2\Delta_{n}+100\delta)-neighborhood of hn(w1)(yn)h_{n}(w_{1})(y_{n}). The reason is because uu is admissible to ww, and u1u_{1} is admissible to w1w_{1}, while the tails of (w,u)(w,u) and (w1,u1)(w_{1},u_{1}) are both at least 100Δn100\Delta_{n} long.

Also, the trisecting points P1,Q1P_{1},Q_{1} and the branch points on the tail [hn(w1)(yn),hn(w1u1)(yn)][h_{n}(w_{1})(y_{n}),h_{n}(w_{1}u_{1})(y_{n})] are out of the 10Δn10\Delta_{n}-neighborhood of the tail [hn(w)(yn),hn(wu)(yn)][h_{n}(w)(y_{n}),h_{n}(wu)(y_{n})].

It follows that w,w1,w10,w11w,w_{1},w_{10},w_{11} are distinct elements in BmB_{m}. If w0w_{0} is not forbidden, then we put C2(w)=C1(w){w10,w11}C_{2}(w)=C_{1}(w)\cup\{w_{10},w_{11}\} and

T2(w)=T1(w)Π(w1).T_{2}(w)=T_{1}(w)\cup\Pi(w_{1}).

If w0w_{0} is forbidden, then we do the same construction as we did to w1w_{1} and obtain two candidates: w00,w01Bmw_{00},w_{01}\in B_{m}, and a tree Π(w0)\Pi(w_{0}) which intersects the previously obtained treelike graph [yn,hn(w)(yn)]Π(w)Π(w1)[y_{n},h_{n}(w)(y_{n})]\cup\Pi(w)\cup\Pi(w_{1}) only in the (2Δn+100δ)(2\Delta_{n}+100\delta)-neighborhood of hn(w0)(yn)h_{n}(w_{0})(y_{n}). In particular, w,w0,w00,w01,w1,w10,w11w,w_{0},w_{00},w_{01},w_{1},w_{10},w_{11} (if they exist) are all distinct. See Figure 4

Refer to caption
Figure 4. This is T2(w)T_{2}(w) in the case that w,w0,w1w,w_{0},w_{1} are forbidden. Π(w)\Pi(w) is in bold lines. We have two more trees Π(w0),Π(w1)\Pi(w_{0}),\Pi(w_{1}). Those three trees stay away from each other except for the balls of radius 2Δn2\Delta_{n} at the points where they are connected.

We set C2(w)=C1(w){w00,w01,w10,w11}C_{2}(w)=C_{1}(w)\cup\{w_{00},w_{01},w_{10},w_{11}\} and

T2(w)=T1(w)Π(w0)Π(w1).T_{2}(w)=T_{1}(w)\cup\Pi(w_{0})\cup\Pi(w_{1}).

If all of the elements w00,w01w_{00},w_{01} and also w10,w11Bmw_{10},w_{11}\in B_{m} (if they exist) are non-forbidden, we stop here and set C(w)=C2(w),T(w)=T2(w)C(w)=C_{2}(w),T(w)=T_{2}(w), otherwise we go to the third step.

In the third step, we do the same construction to the forbidden elements among w00,w01,w10,w11Bmw_{00},w_{01},w_{10},w_{11}\in B_{m} For example, if w00w_{00} is forbidden, then we obtain two elements w000,w001w_{000},w_{001} and a tree Π(w00)\Pi(w_{00}). This tree intersects the tree like graph T2(w)T_{2}(w) in the (2Δn+100δ)(2\Delta_{n}+100\delta)-neighborhood of hn(w00)(yn)h_{n}(w_{00})(y_{n}). We define C3(w)C_{3}(w) from C2(w)C_{2}(w) by adding a pair of elements that are obtained from each forbidden element ww_{**} in this step. Also, we define T3(w)T_{3}(w) be adding the trees Π(w)\Pi(w_{**}) for forbidden elements ww_{**}.

We formally describe the (N+1)(N+1)-th step. In the NN-step, we have CN(w),TN(w)C_{N}(w),T_{N}(w). If all of the elements in CN(w)\CN1C_{N}(w)\backslash C_{N-1} are non-forbidden, then we stop and put C(w)=CN(w),T(w)=TN(w)C(w)=C_{N}(w),T(w)=T_{N}(w). Otherwise, we go to the (N+1)(N+1)-th step.

(N+1)(N+1)-th step. By assumption, at least one element in CN(w)\CN1C_{N}(w)\backslash C_{N-1}, wi1iNw_{i_{1}\cdots i_{N}}, is forbidden. Then we do the same construction as in the step 1, and obtain the tree Π(wi1iN)\Pi(w_{i_{1}\cdots i_{N}}); two elements, wi1iN0w_{i_{1}\cdots i_{N}0} and wi1iN1w_{i_{1}\cdots i_{N}1}. We add those two elements to CN(w)C_{N}(w), and also add Π(wi1iN)\Pi(w_{i_{1}\cdots i_{N}}) to TN(w)T_{N}(w). We do this all forbidden elements in CN(w)\CN1(w)C_{N}(w)\backslash C_{N-1}(w), and get CN+1(w)C_{N+1}(w) and TN+1(w)T_{N+1}(w). Note that the trees Π(wI)\Pi(w_{I}) we added in this step are disjoint from each other, moreover, the distance between two Π(wI),Π(wI\Pi(w_{I}),\Pi(w_{I^{\prime}} with III\not=I^{\prime} is at least 30Δn30\Delta_{n}, where I,II,I^{\prime} are multi-subscripts of length NN.

Since Cn(w)BmC_{n}(w)\subset B_{m} for all nn, and the set Cn(w)C_{n}(w) gets bigger at least by two as nn increases by one, so the process must end in finite steps since BmB_{m} is finite. At the end of the process, we have a set C(w)C(w) in BmB_{m}, and a tree like graph T(w)T(w). We remark that C(w),T(w)C(w),T(w) depend on the choice of separators in the construction, but it is not important.

No that by construction, in C(w)C(w), we have

|{forbidden elements}|+1=|{non-forbidden elements}|.|\{\text{forbidden elements}\}|+1=|\{\text{non-forbidden elements}\}|.

Also, the orbit of yny_{n} by the elements in C(w)C(w) appears as vertices of T(w)T(w).

Part II. Let w1,w2Bmw_{1},w_{2}\in B_{m} be two forbidden elements, and we analyze how the sets C(w1)C(w_{1}) and C(w2)C(w_{2}) are related. In the construction of C(w)C(w), for each forbidden element, vv, we chose a separator uu and an element vBmv^{\prime}\in B_{m}. The pair (v,u)(v,u) defines the tail [hn(v)(yn),hn(vu)(yn)][h_{n}(v)(y_{n}),h_{n}(vu)(y_{n})], which is an arc of T(w)T(w).

Claim. Let w1,w2Bmw_{1},w_{2}\in B_{m} be forbidden elements and u2u_{2} the separator we chose for w2w_{2} to construct T(w2),C(w2)T(w_{2}),C(w_{2}). Assume that the tail τ=(w2,u2)\tau=(w_{2},u_{2}) is not parallel to any of (v,u)(v,u) that appears to construct T(w1),C(w1)T(w_{1}),C(w_{1}). Then, C(w1)C(w2)C(w_{1})\cap C(w_{2}) is empty or {w2}\{w_{2}\}.

Let P,QτP,Q\in\tau be the trisecting points of Π(w2)\Pi(w_{2}). Then neither of them is in the 10Δn10\Delta_{n}-neighborhood of any trisecting point that appears in T(w1)T(w_{1}). Indeed, suppose it was, and let RR be a trisecting point of the tail, σ\sigma, in T(w1)T(w_{1}) such that |PR||P-R| or |QR||Q-R| is 10Δn\leq 10\Delta_{n}. Then, since T(w1),T(w2)T(w_{1}),T(w_{2}) have yny_{n} as the common “root” vertex that they start from, the intersection of σ\sigma and the 20δ20\delta-neighborhood of τ\tau is at least 15\frac{1}{5} of σ\sigma, which contradicts the small cancellation property.

This implies that not only the branch points on τ\tau, but also all branch points of T(w2)T(w_{2}) are outside of the 9Δn9\Delta_{n}-neighborhood of T(w1)T(w_{1}). This is because both T(w1),T(w2)T(w_{1}),T(w_{2}) are tree like graphs. It then follows that C(w1)C(w2)C(w_{1})\cap C(w_{2}) is empty or just w2w_{2} because all other points in C(w2)C(w_{2}) appear on T(w2)T(w_{2}) after the first two branch points on τ\tau, that are close to P,QP,Q. We showed the claim.

Part III. It follows from the claim that there is a finite collection of forbidden elements ww in BmB_{m} such that the C(w)C(w)’s are mutually disjoint, and that any forbidden element vBmv\in B_{m} is either contained in the union of those C(w)C(w)’s, or vv has an admissible separator uu such that the pair (tail) (v,u)(v,u) is parallel to one of the tails that appears in one of the C(w)C(w)’s.

To see that, order the forbidden elements in BmB_{m} as w1,w2,w_{1},w_{2},\cdots. Choose admissible separators u1,u2,u_{1},u_{2},\cdots. First, construct C(w1)C(w_{1}) using the separators we chose. Next, if w2w_{2} is already contained in C(w1)C(w_{1}), then disregard it. Also, if (w2,u2)(w_{2},u_{2}) is parallel to one of the tails in T(w1)T(w_{1}), then disregard w2w_{2} also. Otherwise, construct C(w2)C(w_{2}) and keep it. If w3w_{3} is contained in C(w1)C(w_{1}) or C(w2)C(w_{2}); or (w3,u3)(w_{3},u_{3}) is parallel to one of the tails of T(w1)T(w_{1}) or T(w2)T(w_{2}), then disregard w3w_{3}. Otherwise, construct C(w3)C(w_{3}) and keep it, and so on. By the claim we have shown, those C(w)C(w)’s are mutually disjoint such that for any forbidden element vBmv\in B_{m}, either vv is contained in one of the C(w)C(w)’s we have, or (v,u)(v,u) is parallel to one of the tails of the T(w)T(w)’s.

To finish the proof, by Lemma 4.5, the union of those C(w)C(w)’s contains at least 1𝒟\frac{1}{\mathcal{D}} of the forbidden elements in BmB_{m}. But since in each C(w)C(w), there are more non-forbidden elements than forbidden elements (by one), we conclude that in BmB_{m},

|{forbidden elements}|𝒟|{non-forbidden elements}|.|\{\text{forbidden elements}\}|\leq\mathcal{D}|\{\text{non-forbidden elements}\}|.

Lemma 4.6 is proved. ∎

4.4. Feasible elements

By the lemma 4.6, at least 1𝒟+1\frac{1}{\mathcal{D}+1} of Bm(L,η(S))B_{m}(L,\eta(S)) consists of non-forbidden elements.

Now, we choose a maximal subset in the set of non-forbidden elements in Bm(L,η(S))B_{m}(L,\eta(S)) such that for any two distinct elements w,ww,w^{\prime} in the set, hn(w),hn(w)h_{n}(w),h_{n}(w^{\prime}) are not in the same coset w.r.t. F(hn(ui))F(h_{n}(u_{i})) for any separator uiu_{i}. We call an element in this set an adequate element. Then, in Bm(L,η(S))B_{m}(L,\eta(S)),

|{non-forbidden elements}|𝒟4|{adequate elements}|.\frac{|\{\text{non-forbidden elements}\}|}{\mathcal{D}^{4}}\leq|\{\text{adequate elements}\}|.

This is because there are only four separators, uu, and |F(hn(u))|𝒟|F(h_{n}(u))|\leq\mathcal{D} for each uu. Combining this with Lemma 4.6 we get:

(4) |Bm(L,η(S))|𝒟4(𝒟+1)|{adequate elements in Bm(L,η(S))}|.\frac{|B_{m}(L,\eta(S))|}{\mathcal{D}^{4}(\mathcal{D}+1)}\leq|\{\text{adequate elements in }B_{m}(L,\eta(S))\}|.
Definition 4.7 (feasible elements).

Fix mm, then nn. An element of the form w1u1wquqw_{1}u_{1}\cdots w_{q}u_{q} with wiBm(L,η(S))w_{i}\in B_{m}(L,\eta(S)) are called feasible of type qq if all wiw_{i} are adequate and each uiu_{i} is admissible for wi,wi+1w_{i},w_{i+1}. We define q=0q=0 for the empty element.

For this feasible elements, we consider the following broken geodesic:

α\displaystyle\alpha =[yn,hn(w1)(yn)][hn(w1)(yn),hn(w1u1)(yn)]\displaystyle=[y_{n},h_{n}(w_{1})(y_{n})]\cup[h_{n}(w_{1})(y_{n}),h_{n}(w_{1}u_{1})(y_{n})]\cup\cdots
[hn(w1u1wq)(yn),hn(w1u1wquq)(yn)].\displaystyle\cup[h_{n}(w_{1}u_{1}\cdots w_{q})(y_{n}),h_{n}(w_{1}u_{1}\cdots w_{q}u_{q})(y_{n})].

The Hausdorff-distance between α\alpha and the geodesic [yn,hn(w1u1wquq)(yn)][y_{n},h_{n}(w_{1}u_{1}\cdots w_{q}u_{q})(y_{n})] is at most 2Δn+100δ2\Delta_{n}+100\delta.

Lemma 4.8 (cf. [FS],Lemma 2.6).

For every qq, the map Φn\Phi_{n} is injective on the set of qq-tuples of adequate elements in Bm(L,η(S))B_{m}(L,\eta(S)).

By definition, given a qq-tuples of adequate elements, we choose admissible separators (they are not unique) and form a feasible element of type qq, then map it by Φn\Phi_{n}.

Proof.

We argue by induction on the type qq. If q=0q=0 then nothing to prove since the element is empty.

Assume the conclusion holds for q0q\geq 0. Suppose (q+1)(q+1)-tuples of adequate elements w1,w2,,wq+1w_{1},w_{2},\cdots,w_{q+1} and w1,w2,,wq+1w_{1}^{\prime},w_{2}^{\prime},\cdots,w_{q+1}^{\prime} are given. Let W=w1u1w2u2wq+1uq+1W=w_{1}u_{1}w_{2}u_{2}\cdots w_{q+1}u_{q+1} and W=w1u1w2u2wq+1uq+1W^{\prime}=w_{1}^{\prime}u_{1}^{\prime}w_{2}^{\prime}u_{2}^{\prime}\cdots w_{q+1}^{\prime}u_{q+1}^{\prime} be two feasible elements of type q+1q+1. We assume that hn(W)=hn(W)h_{n}(W)=h_{n}(W^{\prime}) and want to show wi=wiw_{i}=w_{i}^{\prime} for all ii.

The elements W,WW,W^{\prime} define two broken geodesics α,α\alpha,\alpha^{\prime} between yny_{n} and hn(W)(yn)=hn(W)(yn)h_{n}(W)(y_{n})=h_{n}(W^{\prime})(y_{n}). Look at the initial parts of α,α\alpha,\alpha^{\prime}: [yn,hn(w1)(yn)][hn(w1)(yn),hn(w1u1)(yn)][y_{n},h_{n}(w_{1})(y_{n})]\cup[h_{n}(w_{1})(y_{n}),h_{n}(w_{1}u_{1})(y_{n})] and [yn,hn(w1)(yn)][hn(w1)(yn),hn(w1u1)(yn)][y_{n},h_{n}(w_{1}^{\prime})(y_{n})]\cup[h_{n}(w_{1}^{\prime})(y_{n}),h_{n}(w_{1}^{\prime}u_{1}^{\prime})(y_{n})]. We first show that u1=u1u_{1}=u_{1}^{\prime}. Suppose not. Then it would imply that either w1w_{1} or w1w_{1}^{\prime} is forbidden. We explain the reason. First, by the small cancellation property of the separators (Lemma 4.2 (iii)), either [yn,hn(w1)(yn)][hn(w1)(yn),hn(w1u1)(yn)][y_{n},h_{n}(w_{1})(y_{n})]\cup[h_{n}(w_{1})(y_{n}),h_{n}(w_{1}u_{1})(y_{n})] is contained in the 10Δn10\Delta_{n}-neighborhood of [yn,hn(w1)(yn)][y_{n},h_{n}(w_{1}^{\prime})(y_{n})], or [yn,hn(w1)(yn)][hn(w1)(yn),hn(w1u1)(yn)][y_{n},h_{n}(w_{1}^{\prime})(y_{n})]\cup[h_{n}(w_{1}^{\prime})(y_{n}),h_{n}(w_{1}^{\prime}u_{1}^{\prime})(y_{n})] is contained in the 10Δn10\Delta_{n}-neighborhood of [yn,hn(w1)(yn)][y_{n},h_{n}(w_{1})(y_{n})], since u1u_{1} is admissible for w1w_{1} and u1u_{1}^{\prime} is admissible for w1w_{1}^{\prime}. Suppose we are in the first case. Then w1w_{1} would be forbidden since there is a subword of w1w_{1}^{\prime}, denoted by w1¯\overline{w_{1}^{\prime}}, such that

|hn(w1¯)(yn)hn(w1u1)(yn)|15|hn(w1)(yn)hn(w1u1)(yn)|.|h_{n}(\overline{w_{1}^{\prime}})(y_{n})-h_{n}(w_{1}u_{1})(y_{n})|\leq\frac{1}{5}|h_{n}(w_{1})(y_{n})-h_{n}(w_{1}u_{1})(y_{n})|.

Such a subword exists because L(Sn)Δn4L(S_{n})\leq\frac{\Delta_{n}}{4} and |hn(w1)(yn)hn(w1u1)(yn)|100Δn|h_{n}(w_{1})(y_{n})-h_{n}(w_{1}u_{1})(y_{n})|\geq 100\Delta_{n}. We got a contradiction since w1w_{1} is not forbidden. Similarly, if we are in the second case, then w1w_{1}^{\prime} would be forbidden, which is a contradiction. We showed that u1=u1u_{1}=u_{1}^{\prime}.

Next we show w1=w1w_{1}=w_{1}^{\prime}. We first show that hn(w1w11)F(hn(u1))h_{n}(w_{1}w_{1}^{\prime-1})\in F(h_{n}(u_{1})). Suppose not. Then, as in the previous paragraph, the small cancellation property of the separators implies that either w1w_{1} or w1w_{1}^{\prime} is forbidden, a contradiction. We are left with the case that hn(w1w11)F(hn(u1))h_{n}(w_{1}w_{1}^{\prime-1})\in F(h_{n}(u_{1})). But, this means that hn(w1)h_{n}(w_{1}) and hn(w1)h_{n}(w_{1}^{\prime}) are in the same (right) coset w.r.t. F(hn(u1))F(h_{n}(u_{1})). Since both w1w_{1} and w1w_{1}^{\prime} are adequate, it implies w1=w1w_{1}=w_{1}^{\prime}.

Since u1=u1,w1=w1u_{1}=u_{1}^{\prime},w_{1}=w_{1}^{\prime}, it follows that W1=w2u2wq+1uq+1W_{1}=w_{2}u_{2}\cdots w_{q+1}u_{q+1} and W1=w2u2wq+1uq+1W_{1}^{\prime}=w_{2}^{\prime}u_{2}^{\prime}\cdots w_{q+1}^{\prime}u_{q+1}^{\prime} are feasible elements of type qq with hn(W1)=hn(W1)h_{n}(W_{1})=h_{n}(W_{1}^{\prime}). By the induction hypothesis, we have wi=wiw_{i}=w_{i}^{\prime} for all i2i\geq 2, and we are done. ∎

4.5. Proof of Proposition 3.2

We prove Proposition 3.2.

Proof.

Recall that D(ϵ)D(\epsilon) is the WPD function and we set 𝒟=D(100δ)\mathcal{D}=D(100\delta). Recall that for every mm, we have

1mlog(|Bm(L,η(S))|loge(L,η(S)).\frac{1}{m}\log(|B_{m}(L,\eta(S))|\geq\log e(L,\eta(S)).

Given ϵ>0\epsilon>0, choose and fix a large enough mm such that

1m+b(log|Bm(L,η(S))|log(D4(D+1))1mlog|Bm(L,η(S))|ϵ.\frac{1}{m+b}\left(\log|B_{m}(L,\eta(S))|-\log(D^{4}(D+1)\right)\geq\frac{1}{m}\log|B_{m}(L,\eta(S))|-\epsilon.

The constant bb is from Lemma 4.2. Such mm exists since limmlog|Bm(L,η(S))|m=loge(L,η(S))\lim_{m\to\infty}\frac{\log|B_{m}(L,\eta(S))|}{m}=\log e(L,\eta(S)), and bb does not depend on n,mn,m. (We will choose nn right after this.)

Choose nn large enough such that hnh_{n} is injective on B2m(L,η(S))B_{2m}(L,\eta(S)), which defines the forbidden elements in Bm(L,η(S))B_{m}(L,\eta(S)). Then for all qq:

|Bq(m+b)(G,fn(S))|(|Bm(L,η(S))|𝒟4(𝒟+1))q|B_{q(m+b)}(G,f_{n}(S))|\geq\left(\frac{|B_{m}(L,\eta(S))|}{\mathcal{D}^{4}(\mathcal{D}+1)}\right)^{q}

because the number of adequate elements in Bm(L,η(S))B_{m}(L,\eta(S)) is at least |Bm(L,η(S))|𝒟4(𝒟+1)\frac{|B_{m}(L,\eta(S))|}{\mathcal{D}^{4}(\mathcal{D}+1)} by the estimate (4), and Φn\Phi_{n} is injective on the set of feasible elements of type qq by Lemma 4.8.

Then by the above three inequalities,

log(e(G,fn(S))\displaystyle\log(e(G,f_{n}(S)) =limq1q(m+b)log|Bq(m+b)(G,fn(S))|\displaystyle=\lim_{q\to\infty}\frac{1}{q(m+b)}\log|B_{q(m+b)}(G,f_{n}(S))|
1m+blog(|Bm(L,η(S))|𝒟4(𝒟+1))1mlog(|Bm(L,η(S))|ϵ\displaystyle\geq\frac{1}{m+b}\log\left(\frac{|B_{m}(L,\eta(S))|}{\mathcal{D}^{4}(\mathcal{D}+1)}\right)\geq\frac{1}{m}\log(|B_{m}(L,\eta(S))|-\epsilon
loge(L,η(S))ϵ.\displaystyle\geq\log e(L,\eta(S))-\epsilon.

Since we have this for all large enough nn, and ϵ>0\epsilon>0 is arbitrary, we have

limnlog(e(G,fn(S))loge(L,η(S)).\lim_{n\to\infty}\log(e(G,f_{n}(S))\geq\log e(L,\eta(S)).

As we said the other direction is trivial, hence the equality holds. ∎

4.6. A family version

We state a family version of Proposition 3.2. We do not use this proposition in this paper, but it would be useful in the future.

We explain the setting. Let δ>0\delta>0 be a constant, M>0M>0 an integer and D(ϵ)D(\epsilon) a function for WPD. Suppose we have a family of finitely generated groups {Gα}\{G_{\alpha}\} such that each GαG_{\alpha} is not virtually cyclic and acts on a δ\delta-hyperbolic space XαX_{\alpha}. Assume that for any finite generating set SS of any GαG_{\alpha}, the set SMS^{M} contains a hyperbolic element on XαX_{\alpha} that is DD-WPD.

Now suppose that SnS_{n} is a finite generating set of GnG_{n} that is in the family for n>0n>0. Assume that the (infinite) sequence e(Gn,Sn)e(G_{n},S_{n}) is bounded from above. Then by Proposition 2.10, there is a constant A>0A>0 that depends only on M,δM,\delta and the function DD such that e(Gn,Sn)A|Sn|Ae(G_{n},S_{n})\geq A|S_{n}|^{A}, therefore, the sequence |Sn||S_{n}| is bounded from above.

As before, passing to a subsequence, we may assume that there exists >0\ell>0 with |Sn|=|S_{n}|=\ell for all nn. This gives a sequence of surjections fn:(F,S)(Gn,fn(S))f_{n}:(F,S)\to(G_{n},f_{n}(S)) with fn(S)=Snf_{n}(S)=S_{n}, where (F,S)(F,S) is a pair of free group FF and a free generating set SS with |S|=|S|=\ell.

Passing to a further subsequence if necessary, we may assume that the sequence fnf_{n} converges to a limit epimorphism η:(F,S)(L,η)\eta:(F,S)\to(L,\eta).

Then we have:

Proposition 4.9.

Assume that there exists an epimorphism hn:LGnh_{n}:L\to G_{n} for all nn such that all the homomorphisms {fn}\{f_{n}\} factor through the limit epimorphism: η:FL\eta:F\to L, ie, fn=hnηf_{n}=h_{n}\circ\eta. Then,

limne(Gn,fn(S))=e(L,η(S)).\lim_{n\to\infty}e(G_{n},f_{n}(S))=e(L,\eta(S)).

Before we explain the proof, we point out that in Proposition 3.2, the assumption on the existence of hnh_{n} is a consequence of that GG is equationally Noetherian. In the current setting, we state it as an assumption and avoid an assumption related to equational Noetherianity.

That said, the proof is identical to the argument for Proposition 3.2. We recall that by the existence of hnh_{n}, we immediately have limne(Gn,fn(S))e(L,η(S))\lim_{n}e(G_{n},f_{n}(S))\leq e(L,\eta(S)). The main issue was to show the other inequality limne(Gn,fn(S))e(L,η(S))\lim_{n}e(G_{n},f_{n}(S))\geq e(L,\eta(S)) by constructing separators for each action of LL on XX, via hn:LGnh_{n}:L\to G_{n}. That argument applies without change to the current setting since the constants δ,M\delta,M and the function D(ϵ)D(\epsilon) are common for all (Gn,Sn)(G_{n},S_{n}) and XnX_{n}.

5. Examples

An obvious example for Theorem 1.1 is a non-elementary hyperbolic group, GG. Let XX be a Cayley graph of GG, then it is δ\delta-hyperbolic, and the action by GG is (uniformly) proper, so that acylindrical, and non-elementary. The existence of the constant MM is known ([K]). As we said GG is equationally Noetherian ([S], [RW]). Therefore, ξ(S)\xi(S) is well-ordered. This is proved in [FS] and we adapted their argument to prove Theorem 3.1 in this paper. In this section we discuss some other examples.

5.1. Relatively hyperbolic groups

We treat a relatively hyperbolic group. For example see [Bo1] for the definition and basic properties. Suppose GG is hyperbolic relative to a collection of finitely generated subgroups {P1,,Pn}\{P_{1},\cdots,P_{n}\}. Suppose GG is not virtually cyclic and not equal to any PiP_{i}. Then it acts properly discontinuously on a proper δ\delta-hyperbolic space XX such that [Bo1, Proposition 6.13]:

  1. (1)

    There is a GG-invariant collection of points, ΠX\Pi\subset\partial X, with Π/G\Pi/G finite. For each ii, there is a point piΠp_{i}\in\Pi such that the stabilizer of pip_{i} is PiP_{i}. The union of the GG-orbits of the pip_{i}’s is Π\Pi.

  2. (2)

    For every r>0r>0, there is a GG-invariant collection of horoballs B(p)B(p) at each pΠp\in\Pi such that they are rr-separated, ie, d(B(p),B(q))rd(B(p),B(q))\geq r for every distinct p,qΠp,q\in\Pi.

  3. (3)

    The action of GG on X\pΠintB(p)X\backslash\cup_{p\in\Pi}{\rm int}B(p) is co-compact.

A subgroup H<GH<G is called parabolic if it is infinite, fixes a point in X\partial X, and contains no hyperbolic elements. The fixed point is unique and called a parabolic point. In fact, Π\Pi is the set of parabolic points ([Bo1, Proposition 6.1 and 6.13]).

For this action, we have:

Lemma 5.1 (Proposition 5.1[X]).

Let GG and XX be as above. Then there exists MM such that for any finite generating set SS of GG, the set SMS^{M} contains a hyperbolic element on XX.

Also we have the following:

Lemma 5.2.

Let GG and XX be as above. Then the action is uniformly WPD.

Proof.

Given ϵ>0\epsilon>0, take a GG-invariant collection of horoballs that are (10ϵ+100δ)(10\epsilon+100\delta)-separated in XX. Let \mathcal{B} denote the union of the interior of the horoballs in the collection. Then since the action of GG on X\X\backslash\mathcal{B} is co-compact, there exists a constant D(ϵ)D(\epsilon) such that for any yX\y\in X\backslash\mathcal{B}, the cardinality of the following set is at most D(ϵ)D(\epsilon):

{hG||yh(y)|ϵ+10δ}.\{h\in G||y-h(y)|\leq\epsilon+10\delta\}.

We argue that the action is uniformly WPD w.r.t. D=D(ϵ)D=D(\epsilon). Let gGg\in G be hyperbolic with an 10δ10\delta-axis γ\gamma. Let x,yγx,y\in\gamma with |xy|λ(g)|x-y|\geq\lambda(g). It suffices to show that the following set contains at most D(ϵ)D(\epsilon) elements:

{hG||xh(x)|ϵ,|yh(y)|ϵ}.\{h\in G||x-h(x)|\leq\epsilon,|y-h(y)|\leq\epsilon\}.

We divide the case into four:
(1) xx\not\in\mathcal{B}. Then there are at most D(ϵ)D(\epsilon) elements hGh\in G s.t. |xh(x)|ϵ|x-h(x)|\leq\epsilon, and we are done.
(2) yy\not\in\mathcal{B}. This is same as (1).
(3) x,yx,y\in\mathcal{B} such that xB(p)x\in B(p) and yB(q)y\in B(q) with pqp\not=q. Then, there must be z[x,y]z\in[x,y] with zN5ϵ+30δ()z\not\in N_{5\epsilon+30\delta}(\mathcal{B}) since the horoballs in \mathcal{B} are (10ϵ+100δ)(10\epsilon+100\delta)-separated. Then |zh(z)|ϵ+10δ|z-h(z)|\leq\epsilon+10\delta. But there are at most D(ϵ)D(\epsilon) such elements.
(4) x,yB(p)x,y\in B(p) for some pp. Then g(x)B(q)g(x)\in B(q) for some qpq\not=p, since gg is hyperbolic and does not preserve any horoball. Since horoballs are (10ϵ+100δ)(10\epsilon+100\delta)-separated, we have |xg(x)|10ϵ+100δ|x-g(x)|\geq 10\epsilon+100\delta. So, λ(g)10ϵ+50δ\lambda(g)\geq 10\epsilon+50\delta. Now, there are two possibilities: one is that g(x)γg(x)\in\gamma is between xx and yy. Then, as in (3), there must be z[x,y]z\in[x,y] with zN5ϵ+30δ()z\not\in N_{5\epsilon+30\delta}(\mathcal{B}), and we are done. The other possibility is that yy is between xx and g(x)g(x) on γ\gamma. But in this case, since |xy|λ(g)|x-y|\geq\lambda(g), we have |yg(x)|50δ|y-g(x)|\leq 50\delta. Then the distance between B(p)B(p) and B(q)B(q) is at most 50δ50\delta since yB(p)y\in B(p) and g(x)B(q)g(x)\in B(q). But it contradicts the separation of horoballs, so this case does not happen. ∎

We quote a theorem, [GrH, Theorem D].

Theorem 5.3 (Equationally Noetherian, [GrH]).

If GG is hyperbolic relative to equationally Noetherian subgroups, then GG is equationally Noetherian.

We are ready to state a theorem.

Theorem 5.4 (Well-orderedness for relatively hyperbolic groups).

Let GG be a group that is hyperbolic relative to a collection of subgroups {P1,,Pn}\{P_{1},\cdots,P_{n}\}. Suppose GG is not virtually cyclic, and not equal to PiP_{i} for any ii. Suppose each PiP_{i} is finitely generated and equationally Noetherian. Then ξ(G)\xi(G) is well-ordered.

Proof.

GG is equationally Noetherian by Theorem 5.3. Take a hyperbolic space XX with a GG action as above. The action is non-elementary since GG contains a hyperbolic isometry and is not virtually \mathbb{Z}. Then Theorem 3.1 applies by Lemma 5.1 and Lemma 5.2. ∎

5.2. Rank-1 lattices

There are many examples of relatively hyperbolic groups, but we mention one standard family that Theorem 5.4 applies to.

Let GG be a lattice in a simple Lie group of rank-1. It is always finitely generated and has exponential growth, and in fact uniform exponential growth, [EMO]. If it is a uniform lattice, then it is hyperbolic, so that ξ(G)\xi(G) is well-ordered. We prove the following as an immediate application of Theorem 5.4.

Theorem 5.5 (Rank-1 lattices).

Let GG be one of the following groups:

  1. (1)

    A lattice in a simple Lie group of rank-1.

  2. (2)

    The fundamental group of a complete Riemannian manifold MM of finite volume such that there exist a,b>0a,b>0 with b2Ka2<0-b^{2}\leq K\leq-a^{2}<0, where KK denotes the sectional curvature.

Then ξ(G)\xi(G) is well-ordered.

Proof.

We only need to argue for non-uniform lattices since otherwise, GG is a non-elementary hyperbolic group and the conclusion holds. Suppose that GG is a non-uniform lattice. Then, it is known that GG is relatively hyperbolic w.r.t. the the parabolic subgroups, called peripheral subgroups {Hi}\{H_{i}\} that are associated to the cusps, [Fa], [Bo1]. Moreover GG is not virtually cyclic, and not equal to any HiH_{i}. Also, those HiH_{i} are finitely generated virtually nilpotent groups (see for example, [Fa]). It is known that finitely generated virtually nilpotent groups are equationally Noetherian, [Br], so that HiH_{i} are equationally Noetherian. With those facts, Theorem 5.4 applies to GG and the conclusion holds. ∎

We remark that it is known that if GG is linear over a field, then it is equationally Noetherian, [BMR]. For lattices in a simple Lie group, one can apply this result as well to see that GG is equationally Noetherian (consider the adjoint representation on its Lie algebra. It’s faithful since the Lie group is simple).

5.3. Mapping class groups

We discuss mapping class groups as another possible application. Let MCG(Σ)MCG(\Sigma) be the mapping class group of a compact oriented surface Σ=Σg,p\Sigma=\Sigma_{g,p} with genus gg, punctures pp, and complexity c(Σ)=3g+pc(\Sigma)=3g+p. It is known that it is either virtually abelian or has exponential growth, and then uniform exponential growth ([Ma2]). Let 𝒞(Σ)\mathcal{C}(\Sigma) be the curve graph of Σ\Sigma. The graph 𝒞(Σ)\mathcal{C}(\Sigma) has vertex set representing the non-trivial homotopy classes of simple closed curves on Σ\Sigma, and edges joining vertices representing the homotopy classes of disjoint curves. The group MCG(Σ)MCG(\Sigma) naturally acts on it by isometries.

We assume c(Σ)>4c(\Sigma)>4. Then MCG(Σ)MCG(\Sigma) has exponential growth. We recall some facts:

  1. (1)

    The graph 𝒞(Σ)\mathcal{C}(\Sigma) is δ\delta-hyperbolic, and any pseudo-Anosov element in MCG(Σ)MCG(\Sigma) acts hyperbolically on 𝒞(Σ)\mathcal{C}(\Sigma), [MM1].

  2. (2)

    The action of MCG(Σ)MCG(\Sigma) on 𝒞(Σ)\mathcal{C}(\Sigma) is acylindrical, [Bo]. It is non-elementary.

  3. (3)

    There exists T(Σ)>0T(\Sigma)>0 such that for any pseudo-Anosov element gg, λ(g)T(Σ)>0\lambda(g)\geq T(\Sigma)>0 for the action on 𝒞(Σ)\mathcal{C}(\Sigma), [MM1]. (This follows from the acylindricity as we pointed out.)

  4. (4)

    There exists M(Σ)M(\Sigma) such that for any finite SMCGS\subset MCG such that S\langle S\rangle contains a pseudo-Anosov element, then SMS^{M} contains a pseudo-Anosov element, [Ma].

In summary, the action of MCG(Σ)MCG(\Sigma) on 𝒞(Σ)\mathcal{C}(\Sigma) is known to satisfy all the assumptions of Theorem 1.1 except we do not know if MCG(Σ)MCG(\Sigma) is equationally Noetherian or not. We expect it to hold (for example, see an announcement by Daniel Groves in [GrH]), but it does not exist in the literature yet. Once that is verified, it will imply that ξ(MCG(Σ))\xi(MCG(\Sigma)) is well-ordered if c(Σ)>4c(\Sigma)>4.

We remark that for Σ1,1,Σ1,0,Σ0,4\Sigma_{1,1},\Sigma_{1,0},\Sigma_{0,4}, the conclusion holds by [FS] since MCG(Σ)MCG(\Sigma) is hyperbolic (a well-known fact, for example, [MM1]).

Question 5.6.

Let G=MCG(Σ)G=MCG(\Sigma). Is ξ(G)\xi(G) well-ordered? If so, the infimum is attained. It would be interesting to know its value and generating sets that attain the minimum for each Σ\Sigma.

5.4. Three manifold groups

We discuss three manifold groups. Let MM be a closed, orientable 33-manifold. MM is called irreducible if π1(M)\pi_{1}(M) does not admit a non-trivial splitting over the trivial group.

Let MM be a closed, orientable, irreducible 3-manifold which is not a torus bundle over a circle. Then there is a finite collection of embedded disjoint essential tori TiT_{i} in MM such that each connected component of M\iTiM\backslash\cup_{i}T_{i} is geometric, ie, either Seifert fibred, or admitting hyperbolic or Sol-geometry. Such a collection of smallest number of tori is called the JSJ decomposition of MM. A JSJ-decomposition in this sense exists by the solution of the geometrization conjecture of 3-manifolds. The collection of tori is maybe empty. Otherwise we say MM has non-trivial JSJ-decomposition. A non-trivial JSJ-decomposition gives a graph of groups decomposition of π1(M)\pi_{1}(M) along subgroups isomorphic to 2\mathbb{Z}^{2}, and π1(M)\pi_{1}(M) has exponential growth. Its Bass-Serre tree is called the JSJ-tree, TMT_{M}. An action of a group GG on a tree TT is called kk-acylindrical if for every non-trivial element gg, the subtree of fixed points by gg is either empty or of diameter at most kk. It is proved that the action of π1(M)\pi_{1}(M) on TMT_{M} is 44-acylindrical, [WZ, Proposition 4.2]. It implies that the action is uniformly 44-WPD. It is known that if an action of GG on a tree TT is kk-acylindrical for some kk, then it is acylindrical [MO, Lemma 5.2], therefore it is uniformly WPD (Lemma 2.3). Moreover, the acylindricity constants R(ϵ),N(ϵ)R(\epsilon),N(\epsilon) and the uniformly WPD constant D(ϵ)D(\epsilon) depend only on kk and ϵ\epsilon. The moreover part easily follows from the proof of Lemma 5.2in [MO].

Theorem 5.7.

Let MM be a closed orientable 33-manifold, and G=π1(M)G=\pi_{1}(M). If MM is one of the following, then GG has exponential growth and ξ(G)\xi(G) is well-ordered.

  1. (1)

    MM is not irreducible such that GG is not isomorphic to 22\mathbb{Z}_{2}*\mathbb{Z}_{2}.

  2. (2)

    MM is irreducible such that it is not a torus bundle over a circle, and that it has a non-trivial JSJ-decomposition.

  3. (3)

    MM admits hyperbolic geometry.

  4. (4)

    MM is Seifert fibered such that the base 22-orbifold is hyperbolic.

Proof.

First, every three manifold group is equationally Noetherian, [GHL].

(1) In this case, GG is a non-trivial free product ABA*B. Since it is not 22\mathbb{Z}_{2}*\mathbb{Z}_{2}, it has exponential growth. GG acts on the Bass-Seree tree TT of this free product. Then for any finite generating set SS, the set S2S^{2} contains a hyperbolic element on TT, [Se]. The action of GG on TT is 0-acylindrical, so that the action is uniformly DD-WPD for some DD. Since TT is hyperbolic, Theorem 3.1 applies with M=2M=2.

(2) In this case, let TMT_{M} be the JSJ-tree of MM. Then the action of GG on TMT_{M} is 44-acylindrical, so that it is uniformly DD-WPD for some DD. Theorem 3.1 applies with M=2M=2.

(3) If MM is hyperbolic, then GG is a non-elementary, hyperbolic group. Then ξ(G)\xi(G) is well-ordered by [FS].

(4) In this case, we have the following exact sequence:

1GH1,1\to\mathbb{Z}\to G\to H\to 1,

where HH is the orbifold-fundamental group of the base 2-orbifold, and \mathbb{Z} is the fundamental group of the fiber circle. We denote this subgroup CC. By assumption, HH is a non-elementary, hyperbolic group.

We claim that ξ(G)=ξ(H)\xi(G)=\xi(H). To see it, let SS be a finite generating set of GG. Let S¯\bar{S} be the image of SS by the projection GHG\to H. We have |S¯n||Sn||\bar{S}^{n}|\leq|S^{n}|. But it is known that the subgroup CC is not distorted in GG in the sense that there is a constant K>0K>0 such that for any n>0n>0, we have |SnC|Kn|S^{n}\cap C|\leq Kn, [NS, Proposition 1.2 (2)]. It implies that for all nn, we have |Sn|Kn|S¯n||S^{n}|\leq Kn|\bar{S}^{n}|. It follows that e(G,S)=e(H,S¯)e(G,S)=e(H,\bar{S}), therefore ξ(G)ξ(H)\xi(G)\subset\xi(H). On the other hand, if SS is a finite generating set of HH, then there is a finite generating set S~\tilde{S} of GG which projects to SS, by lifting each element of SS to GG, then adding a generator of CC, which gives S~\tilde{S}. Then as we saw, e(G,S~)=e(H,S)e(G,\tilde{S})=e(H,S), which implies ξ(H)ξ(G)\xi(H)\subset\xi(G). We proved the claim.

But by [FS], ξ(H)\xi(H) is well-ordered, therefore so is ξ(G)\xi(G). ∎

A torus bundle over a circle either admits Sol-geometry, or it is Seifert fibered (see for example, [WZ]). Therefore the theorem covers all closed, orientable 33-manifolds with fundamental groups of exponential growth except that MM has Sol-geometry. We leave it as a question.

Question 5.8.

Let MM a closed, orientable 33-manifold that has the geometry of three dimensional solvable group, Sol. Then is ξ(π1(M))\xi(\pi_{1}(M)) well-ordered ? Also, in view of Proposition 2.10, is it true that if |S||S|\to\infty then e(π1(M),S)e(\pi_{1}(M),S)\to\infty?

6. The set of growth of subgroups

We discuss the set of growth of subgroups in a finitely generated group GG. Define

Θ(G)={e(H,S)|SG,|S|<,H=S,e(H,S)>1}.\Theta(G)=\{e(H,S)|S\subset G,|S|<\infty,H=\langle S\rangle,e(H,S)>1\}.

The set Θ(G)\Theta(G) is countable and contains ξ(G)\xi(G) as a subset. If GG is a hyperbolic group, it is known by [FS, Section 5] that Θ(G)\Theta(G) is well-ordered.

6.1. Subgroups with hyperbolic elements

As usual, suppose GG acts on a δ\delta-hyperbolic space XX. We introduce a subset in Θ(G)\Theta(G) as follows:

ΘX(G)={e(H,S)|SG,|S|<,H=S,e(H,S)>1},\Theta_{X}(G)=\{e(H,S)|S\subset G,|S|<\infty,H=\langle S\rangle,e(H,S)>1\},

where in addition we only consider SS such that S\langle S\rangle contains a hyperbolic element on XX. This set depends on the action on XX.

Theorem 6.1 (cf. Theorem 3.1).

Suppose GG acts on a δ\delta-hyperbolic space XX, and GG is not virtually cyclic. Let D(ϵ)D(\epsilon) be a function for WPD. Assume that there exists a constant MM such that if S\langle S\rangle contains a hyperbolic element on XX for a finite subset SGS\subset G, then SMS^{M} contains a hyperbolic element that is DD-WPD. Assume that GG is equationally Noetherian. Then, ΘX(G)\Theta_{X}(G) is a well-ordered set.

Proof.

The proof is nearly identical to Theorem 3.1.

Let {Sn}\{S_{n}\} be a sequence of finite generating sets of subgroups of exponential growth in GG, {Hn}\{H_{n}\}, such that each Hn=SnH_{n}=\langle S_{n}\rangle contains a hyperbolic element on XX and that {e(Hn,Sn)}\{e(H_{n},S_{n})\} is a strictly decreasing sequence with limne(Hn,Sn)=d\lim_{n\to\infty}e(H_{n},S_{n})=d, for some d1d\geq 1.

By our assumption, Proposition 2.10 applies to HnH_{n}. Therefore, |Sn||S_{n}| is uniformly bounded from above. By passing to a subsequence we may assume that |Sn||S_{n}| is constant, |Sn|=|S_{n}|=\ell, for the entire sequence.

Let Sn={x1(n),,x(n)}S_{n}=\{x_{1}^{(n)},\cdots,x_{\ell}^{(n)}\}. Let FF be the free group of rank \ell with a free generating set: S={s1,,s}S=\{s_{1},\ldots,s_{\ell}\}. For each index nn, we define a map: fn:FGf_{n}:F\to G, by setting: fn(si)=xi(n)f_{n}(s_{i})=x_{i}^{(n)}. By construction: e(Hn,Sn)=e(Hn,fn(S))e(H_{n},S_{n})=e(H_{n},f_{n}(S)).

Then, as before, the sequence {fn:FG}\{f_{n}:F\to G\} subconverges to a surjective homomorphism η:FL\eta:F\to L, where LL is a limit group over GG.

By assumption, GG is equationally Noetherian. By the general principle (Lemma 1.11), there exists an epimorphism hn:LGh_{n}:L\to G such that by passing to a subsequence we may assume that all the homomorphisms {fn}\{f_{n}\} factor through the limit epimorphism: η:FL\eta:F\to L, ie, fn=hnηf_{n}=h_{n}\circ\eta.

Then, we have

Proposition 6.2 (cf. Proposition 3.2).
limne(Hn,fn(S))=e(L,η(S)).\lim_{n\to\infty}e(H_{n},f_{n}(S))=e(L,\eta(S)).
Remark 6.3.

This proposition is a special case of Proposition 4.9, when all δ\delta-hyperbolic spaces XnX_{n} are XX.

Proof.

The proof is identical to Proposition 3.2. As in the beginning of the argument (Section 4.1), for each nn, we pick one hyperbolic isometry gSnMg\in S_{n}^{M} on XX that is DD-WPD to start with. This is possible since Hn=SnH_{n}=\langle S_{n}\rangle contains a such hyperbolic element by the definition of ΘX(G)\Theta_{X}(G). The rest is same, and we omit it. ∎

We continue the proof of the theorem. As in the proof of Theorem 3.1, Proposition 6.2 proves that there is no strictly decreasing sequence of rates of growth, {e(Hn,Sn)}\{e(H_{n},S_{n})\}, since a strictly decreasing sequence can not approach its upper bound, a contradiction. Hence, ΘX(G)\Theta_{X}(G) is well-ordered. The theorem is proved. ∎

6.2. Relatively hyperbolic groups

We apply Theorem 6.1 relatively hyperbolic groups. Let (G,{Pi})(G,\{P_{i}\}) be a relatively hyperbolic group with PiP_{i} finitely generated. Define

Θnonelem.(G)={e(H,S)|SG,|S|<,H=S,e(H,S)>1},\Theta_{{\rm non-elem.}}(G)=\{e(H,S)|S\subset G,|S|<\infty,H=\langle S\rangle,e(H,S)>1\},

where in addition we only consider HH that is not conjugate into any PiP_{i}. This is a subset in Θ(G)\Theta(G).

We first characterize the subgroups HH that appear in the definition in terms of the action on a δ\delta-hyperbolic space XX that we described in Section 5.1. Fix such XX and an action by GG.

We recall a lemma. This is straightforward from the classification of subgroups that act on a hyperbolic space, [G, Section 3.1].

Lemma 6.4.

Let H<GH<G be a subgroup. Then the following two are equivalent:

  1. (1)

    HH has an element gg that is hyperbolic on XX, and HH is not virtually \mathbb{Z}.

  2. (2)

    HH is infinite, not virtually \mathbb{Z} and not conjugate into any PiP_{i}.

The lemma implies:

Θnonelem.(G)=ΘX(G).\Theta_{{\rm non-elem.}}(G)=\Theta_{X}(G).

We prove:

Theorem 6.5.

Let GG be a group that is hyperbolic relative to a collection of subgroups {P1,,Pn}\{P_{1},\cdots,P_{n}\}. Suppose GG is not virtually cyclic, and not equal to PiP_{i} for any ii. Suppose each PiP_{i} is finitely generated and equationally Noetherian. Then Θnonelem.(G)\Theta_{{\rm non-elem.}}(G) is well-ordered.

Proof.

It suffices to argue that ΘX(G)\Theta_{X}(G) is well-ordered. For that, we apply Theorem 6.1 to the action of GG on XX. We already checked the assumptions in the proof of Theorem 5.4, except that Lemma 5.1 holds for subgroups. Namely, there is a constant MM such that for any finite set SGS\subset G such that S\langle S\rangle contains a hyperbolic isometry on XX, then SMS^{M} contains a hyperbolic isometry. But this is true and the argument is identical to the proof of [X, Proposition 5.1], and we omit it. ∎

As an example of Theorem 6.5 we prove:

Theorem 6.6.

Let GG be a group in Theorem 5.5. Then Θ(G)\Theta(G) is a well-ordered set.

Proof.

As we said in the proof of Theorem 5.5, GG is relatively hyperbolic w.r.t. the parabolic subgroups {Hi}\{H_{i}\}, which are associated to the cusps, and HiH_{i} are virtually nilpotent. By applying Theorem 6.5 to (G,{Hi})(G,\{H_{i}\}), we have that Θnonelem(G)\Theta_{{\rm non-elem}}(G) is well-ordered. But if a subgroup H=SH=\langle S\rangle is conjugate into one of HiH_{i}, then it is virtually nilpotent, so that HH has polynomial growth. It follows that Θnonelem=Θ(G)\Theta_{{\rm non-elem}}=\Theta(G) holds for GG, so that Θ(G)\Theta(G) is well-ordered. ∎

6.3. Subgroups in mapping class groups

We discuss mapping class groups. A subgroup H<MCG(Σ)H<MCG(\Sigma) is called large if it contains two independent pseudo-Anosov elements. Such HH has exponential growth.

We define:

Θlarge(MCG(Σ))={e(H,S)|SMCG(Σ),|S|<,S=H,e(H,S)>1},\Theta_{{\rm large}}(MCG(\Sigma))=\{e(H,S)|S\subset MCG(\Sigma),|S|<\infty,\langle S\rangle=H,e(H,S)>1\},

where in addition H<MCG(Σ)H<MCG(\Sigma) is large. Note that ξ(MCG(Σ))Θlarge(MCG(Σ))\xi(MCG(\Sigma))\subset\Theta_{{\rm large}}(MCG(\Sigma)).

Theorem 6.7.

Assume that G=MCG(Σ)G=MCG(\Sigma) is equationally Noetherian. Then Θlarge(MCG(Σ))\Theta_{{\rm large}}(MCG(\Sigma)) is well-ordered.

Proof.

We suppress Σ\Sigma and denote MCGMCG. Let XX the curve graph of Σ\Sigma. Then, as we said, Theorem 6.1 applies to the action of MCGMCG on XX. We conclude that ΘX(MCG)\Theta_{X}(MCG) is well-ordered. Now, we claim ΘX(MCG)=Θlarge(MCG)\Theta_{X}(MCG)=\Theta_{{\rm large}}(MCG). Indeed, given SMCGS\subset MCG, if H=SH=\langle S\rangle contains a hyperbolic element on XX, then it is a pseudo-Anosov element, and moreover, from e(H,S)>1e(H,S)>1, HH must be large. We showed ΘX(MCG)Θlarge(MCG)\Theta_{X}(MCG)\subset\Theta_{{\rm large}}(MCG). On the other hand, for SMCGS\subset MCG if H=SH=\langle S\rangle is large in MCGMCG, then HH contains hyperbolic isometries on XX, so that Θlarge(MCG)ΘX(MCG)\Theta_{{\rm large}}(MCG)\subset\Theta_{X}(MCG). ∎

It is natural to ask the following question. To deal with a non-large subgroup, considering the action on the curve graph does not seem to be enough.

Question 6.8.

Is Θ(MCG(Σ))\Theta(MCG(\Sigma)) well-ordered ?

7. Finiteness

7.1. Finiteness of equal growth generating sets

If GG is a hyperbolic group, it is known by [FS, Section 3] that for ρξ(G)\rho\in\xi(G), there are only finitely many generating sets SS of GG, up to Aut (G)\text{Aut\,}(G), such that ρ=e(G,S)\rho=e(G,S). We discuss this issue.

Theorem 7.1 (Finiteness. cf. Theorem 3.1 in [FS]).

Suppose a finitely generated group GG acts on a δ\delta-hyperbolic space XX and GG is not virtually cyclic. Let D(ϵ)D(\epsilon) be a WPD-function. Assume that there exists a constant MM such that if SS is a finite generating set of GG, then SMS^{M} contains a hyperbolic element that is DD-WPD. Assume that GG is equationally Noetherian.

Then for any ρξ(G)\rho\in\xi(G), up to the action on Aut (G)\text{Aut\,}(G), there are at most finitely many finite generating set SS such that e(G,S)=ρe(G,S)=\rho.

Proof.

We argue by contradiction. Suppose that there are infinitely many finite sets of generators {Sn}\{S_{n}\} that satisfy: e(G,Sn)=ρe(G,S_{n})=\rho, and no pair of generating sets SnS_{n} is equivalent under the action of the automorphism group Aut (G)\text{Aut\,}(G). As in the proof of Theorem 3.1, by Proposition 2.10, the cardinality of the generating sets {Sn}\{S_{n}\} is bounded, so we may pass to a subsequence that have a fixed cardinality \ell. Hence, each generating set SnS_{n} corresponds to an epimorphism, fn:FGf_{n}:F\to G, where SS is a fixed free generating set of FF, and fn(S)=Snf_{n}(S)=S_{n}.

By passing to a further subsequence, we may assume that the sequence of epimorphisms {fn}\{f_{n}\} converges to a limit group LL with η:FL\eta:F\to L the associated quotient map. As in the proof of Theorem 3.1, since GG is equationally Noetherian, by Lemma 1.11, for large nn, fn=hnηf_{n}=h_{n}\circ\eta, where hn:LGh_{n}:L\to G is an epimorphism. In particular, Sn=hn(η(S))S_{n}=h_{n}(\eta(S)). We pass to a further subsequence such that for every nn, fn=hnηf_{n}=h_{n}\circ\eta.

Since for every index nn, hnh_{n} is an epimorphism from LL onto GG that maps η(S)\eta(S) to fn(S)f_{n}(S), e(G,fn(S))e(L,η(S))e(G,f_{n}(S))\leq e(L,\eta(S)). We prove:

Proposition 7.2 (cf. Proposition 3.2 [FS]).

If ker(hn0)\ker(h_{n_{0}}) is infinite for some n0n_{0}, then e(G,fn0(S))<e(L,η(S))e(G,f_{n_{0}}(S))<e(L,\eta(S)).

We postpone the proof of the proposition until the next section, and proceed. We prove a lemma.

Lemma 7.3.

The group LL contains a finite normal subgroup N=NLN=N_{L} that contains all finite normal subgroups in LL, such that |N|2D(100δ)|N|\leq 2D(100\delta).

We recall one fact we use in the proof. If a finitely generated group GG acts on a δ\delta-hyperbolic space XX such that GG is not virtually cyclic and GG contains a hyperbolic element on XX that is DD-WPD, then GG contains a maximal finite normal subgroup N<GN<G. Moreover, |N|2D(100δ)|N|\leq 2D(100\delta). We sometimes denote NN by NGN_{G}.

The existence of such NN is known for an acylindrically hyperbolic group [DGO, Theorem 6.14], and the same proof applies to our setting, which we briefly recall. Indeed, NN is the intersection of E(g)E(g) for all hyperbolic elements gGg\in G on XX. It is obvious that NN is normal. By assumption, there must be a hyperbolic and WPD element, gg. Also, there is another element hh such that gg and hh are independent. Then, by Proposition 2.6, E(g)E(h)E(g)\cap E(h) is finite, so that NN is finite. On the other hand, if NN^{\prime} is a finite normal subgroup in GG, then for every hyperbolic element gGg\in G, there is n>0n>0 such that NN^{\prime} is contained in the centralizer of gng^{n}, so that N<E(g)N^{\prime}<E(g). It implies that N<NN^{\prime}<N. We showed that NN is maximal.

Lastly, to see |N|2D(100δ)|N|\leq 2D(100\delta), consider the exact sequence 1F(g)E(g)C11\to F(g)\to E(g)\to C\to 1 for the hyperbolic and DD-WPD element gg. Recall that |F(g)|D(100δ)|F(g)|\leq D(100\delta) if CC is cyclic. From this we have |N|2D(100δ)|N|\leq 2D(100\delta).

We prove the lemma.

Proof.

Let N<LN<L be a finite normal subgroup. Since hnh_{n} is surjective, hn(N)<Gh_{n}(N)<G is a finite normal subgroup, therefore hn(N)<NGh_{n}(N)<N_{G} for any hnh_{n}. Also, for sufficiently large nn, the surjection hn:LGh_{n}:L\to G is injective on NN. But since |NG|2D(100δ)|N_{G}|\leq 2D(100\delta), we have |N|2D(100δ)|N|\leq 2D(100\delta).

If N1,N2<LN_{1},N_{2}<L are two finite normal subgroups, then N1N2N_{1}N_{2} is a finite normal subgroup. Combined with the fact in the previous paragraph, there must be the maximal finite normal subgroup NLN_{L} in LL with |NL|2D(100δ)|N_{L}|\leq 2D(100\delta). ∎

By Proposition 3.2, limne(G,Sn)=e(L,η(S))\lim_{n\to\infty}e(G,S_{n})=e(L,\eta(S)). By our assumption, for every index nn, e(G,Sn)=ρe(G,S_{n})=\rho. Hence, e(L,η(S))=ρe(L,\eta(S))=\rho, so that for every nn, e(G,Sn)=e(L,η(S))e(G,S_{n})=e(L,\eta(S)).

It follows from Proposition 7.2 that for every nn, ker(hn)\ker(h_{n}) is finite. Since ker(hn)\ker(h_{n}) is a normal subgroup in LL, by Lemma 7.3, ker(hn)<NL\ker(h_{n})<N_{L}. Since NLN_{L} is a finite group, there are only finitely many possibilities for ker(hn)\ker(h_{n}). It follows that there must be N0<NLN_{0}<N_{L} such that ker(hn)=N0\ker(h_{n})=N_{0} for infinitely many nn.

The map hnh_{n} induces an isomorphism from L/ker(hn)L/\ker(h_{n}) to GG. Notice that this gives an isomorphism from (L/ker(hn),η(S))(L/\ker(h_{n}),\eta(S)) to (G,Sn)(G,S_{n}) since hnh_{n} gives a bijection between η(S)\eta(S) and SnS_{n}. (Here, we may assume that each SnS_{n} consists of distinct elements, so that no two elements in η(S)\eta(S) are identifies by hnh_{n}.) But this implies that (L/N0,η(S))(L/N_{0},\eta(S)) is isomorphic to (G,Sn)(G,S_{n}) for infinitely many nn by hnh_{n}, ie, those (G,Sn)(G,S_{n}) are isomorphic to each other. This is a contradiction since all of them must be non-isomorphic. Theorem 7.1 is proved. ∎

7.2. Idea of the proof of Proposition 7.2

We prove Proposition 7.2. The argument is long and complicated, but the main idea is same as the proof of [FS, Proposition 3.2], and we adapt it to our setting. Also, the proof is similar to the proof of Proposition 3.2, which also follows the counterpart in the paper [FS]. The difference between this paper and [FS] is that while they use the action of the limit group LL on a limit object, called a limit tree, while in our paper we use the actions of LL on XX induced from the maps hn:LGh_{n}:L\to G. But this approach is already taken in the proof of Proposition 3.2.

So, rather than giving a full formal proof, we first explain the strategy of the proof, then give all definitions and intermediate claims, which appear in the proof of [FS, Proposition 3.2], then explain the part where we need to make technical modifications, most of which already appeared in Section 4. One advantage of not using the action of LL on a limit object is that one does not need to deal with the degeneration of the action on the limit object. A trade-off is that we need to keep attention to the various constants related to the actions induced by hnh_{n} through the argument.

Strategy of the proof. We start with an informal description of the idea. The constant n0n_{0} is given as the assumption, which gives the homomorphism hn0:(L,η(S))(G,fn0(S))h_{n_{0}}:(L,\eta(S))\to(G,f_{n_{0}}(S)) with an infinite kernel. To show that e(G,fn0(S))<e(L,η(S))e(G,f_{n_{0}}(S))<e(L,\eta(S)), we will produce not only infinitely many (which is obvious by the assumption), but “exponentially many” elements in the preimage of gg by hn0h_{n_{0}} for each gGg\in G. Those elements are given by the map ϕn\phi_{n}. They are exponentially many in terms of the word length of gg w.r.t. fn0(S)f_{n_{0}}(S). See the estimate (6).

In the proof two constants mm and nn appear. They will be chosen and fixed around the end of the proof. The constant m>1m>1 is first chosen. It will be used to measure the gap between e(G,fn0(S))e(G,f_{n_{0}}(S)) and e(L,η(S))e(L,\eta(S)). The constant nn depends on mm, so that hnh_{n} is injective on B2m(L,η(S))B_{2m}(L,\eta(S)). Also, a positive integer qq is used to make the word length of gg, which is mqmq longer and take a limit at the end of the proof.

We now explain more concretely. Similar to Bm(G,Sn)B_{m}(G,S_{n}), the ball of radius mm in Cayley(G,Sn)Cayley(G,S_{n}) centered at the identity, let Sph m(G,Sn)\text{Sph\,}_{m}(G,S_{n}) denote the sphere of radius mm. It is an elementary fact that if e(G,Sn)>1e(G,S_{n})>1, then

e(G,Sn)=lim supm|Sph m(G,Sn)|1m.e(G,S_{n})=\limsup_{m\to\infty}|\text{Sph\,}_{m}(G,S_{n})|^{\frac{1}{m}}.

Given m>0m>0, for a large enough n>0n>0 depending on mm, we will define a “map” ϕn\phi_{n}:

ϕn:Sph mq(G,Sn0)Bq(m+2b)(L,η(S))\phi_{n}:\text{Sph\,}_{mq}(G,S_{n_{0}})\to B_{q(m+2b)}(L,\eta(S))

for all q>0q>0, where bb is a constant that does not depend on n,m,qn,m,q. The map ϕn\phi_{n} is similar to the map Φn\Phi_{n} in the section 4.2, but strictly speaking ϕn\phi_{n} is not a map, but ϕn(g)\phi_{n}(g) is a finite set of elements in Bq(m+2b)(L,η(S))B_{q(m+2b)}(L,\eta(S)) for each gg. But we abuse the notation and call them maps in the following account.

It satisfies for every gSph mq(G,Sn0)g\in\text{Sph\,}_{mq}(G,S_{n_{0}}) to have the following two properties. Set 𝒟=D(200δ)\mathcal{D}=D(200\delta).

  1. (i).
    (m1𝒟4(𝒟1))q|ϕn(g)|,\left(\frac{m-1}{\mathcal{D}^{4}(\mathcal{D}-1)}\right)^{q}\leq|\phi_{n}(g)|,

    see the estimate (5).

  2. (ii).

    hn0ϕn(g)=g,h_{n_{0}}\circ\phi_{n}(g)=g, which implies that for ghg\not=h, we have ϕn(g)ϕn(h)=\phi_{n}(g)\cap\phi_{n}(h)=\emptyset.

Once we have such a map ϕn\phi_{n}, we argue as follows: fix a (large) mm. Since ϕn(g)Bq(m+2b)(L,η(S))\phi_{n}(g)\subset B_{q(m+2b)}(L,\eta(S)), we have from (i) and (ii) that

|Sph mq(G,Sn0)|(m1𝒟4(𝒟1))q|Bq(m+2b)(L,η(S))|.|\text{Sph\,}_{mq}(G,S_{n_{0}})|\left(\frac{m-1}{\mathcal{D}^{4}(\mathcal{D}-1)}\right)^{q}\leq|B_{q(m+2b)}(L,\eta(S))|.

Taking log\log, dividing by mqmq, and letting qq\to\infty, we have as the limsup

loge(G,Sn0)+log(m1)log(𝒟4(𝒟1))mm+2bmloge(L,η(S)).\log e(G,S_{n_{0}})+\frac{\log(m-1)-\log(\mathcal{D}^{4}(\mathcal{D}-1))}{m}\leq\frac{m+2b}{m}\log e(L,\eta(S)).

Since bb does not depend on mm, choosing mm large enough, this inequality implies

loge(G,Sn0)<loge(L,η(S)).\log e(G,S_{n_{0}})<\log e(L,\eta(S)).

Roughly speaking, the construction of ϕn\phi_{n} is as follows. As in the construction of Φn\Phi_{n}, we first construct separators. To define separators in LL, we use a non-trivial element rnkerhn0r_{n}\in\ker h_{n_{0}}. Separators will be products of conjugates of rnr_{n}, so that they are also in kerhn0\ker h_{n_{0}}, which will imply the property (ii) in the above. The separators depend on nn.

For each nn, the map hn:(L,η(S))(G,Sn)h_{n}:(L,\eta(S))\to(G,S_{n}) gives a canonical bijection between η(S)\eta(S) and Sn=fn(Sn)S_{n}=f_{n}(S_{n}). This gives a bijection between the words on η(S)\eta(S) (not elements in LL) and the words on SnS_{n}.

Let m,q>0m,q>0 be integers. We will fix mm and let qq\to\infty later. Given an element gSph mq(G,Sn0)g\in\text{Sph\,}_{mq}(G,S_{n_{0}}), we choose a word w(g)w(g) of length mqmq on Sn0S_{n_{0}} that represents gg. We divide w(g)w(g) into qq subwords of length mm. As we said, each subword of length mm canonically gives a word of length mm on η(S)\eta(S) via the map hn0h_{n_{0}}. We further subdivide each of the subwords of length mm on η(S)\eta(S) into two words of length kk and mkm-k. We choose kk to satisfy 1km11\leq k\leq m-1. In this way, for each choice of a qq-tuple of such kk’s, we divided the word on η(S)\eta(S) corresponding to w(g)w(g) into 2q2q subwords. There are (m1)q(m-1)^{q} ways to subdivide it.

To each of such subdivision, we insert separators to the (2q1)(2q-1) break points and obtain an element in Bq(m+2b)(L,η(S))B_{q(m+2b)}(L,\eta(S)) since the word length of each separator is at most bb. We obtain (m1)q(m-1)^{q} such elements. Since separators are in ker(hn0)\ker(h_{n_{0}}), those elements (words) are mapped to gg by hn0h_{n_{0}}.

But we do not know if they are all distinct as elements in LL, but we will show that there are at least (m1𝒟4(𝒟1))q(\frac{m-1}{\mathcal{D}^{4}(\mathcal{D}-1)})^{q} elements that are distinct. They are called feasible elements. This collection of feasible elements is denoted by ϕn(g)\phi_{n}(g). They are significantly many, so that e(L,η(S))e(L,\eta(S)) is strictly larger than e(G,Sn0)e(G,S_{n_{0}}) as we computed in the above.

Lastly, to show that those feasible elements in LL defined for each gg are distinct, as in Section 4, we let them act on the space XX via the map hnh_{n} for a large enough nn (We choose nn such that hnh_{n} is injective on B2m(L,η(S))B_{2m}(L,\eta(S)).) We then argue that the images of the base point ynXy_{n}\in X by those elements are distinct. In the paper [FS], they use a limit tree YY on which LL acts to argue that feasible elements are distinct. Here, we use the action on XX. This difference already appeared in Section 4.

7.3. Proof

We prove Proposition 7.2.

Proof.

As in Section 4.1, set 𝒟=D(200δ)\mathcal{D}=D(200\delta), and for each nn let ynXy_{n}\in X be a point where L(Sn2𝒟M)L(S_{n}^{2\mathcal{D}M}) is achieved and put

Δn=100δ+4𝒟L(Sn2𝒟M).\Delta_{n}=100\delta+4\mathcal{D}L(S_{n}^{2\mathcal{D}M}).

We consider germs w.r.t. Δn\Delta_{n}.

As we did in Lemma 4.2, in the next lemma we construct separators uiGu_{i}\in G, which give u^1,u^2,u^3,u^4L\hat{u}_{1},\hat{u}_{2},\hat{u}_{3},\hat{u}_{4}\in L by pulling them back by hnh_{n}. In addition to the properties in Lemma 4.2 they satisfy hn0(u^i)=1h_{n_{0}}(\hat{u}_{i})=1 in GG. We state it as a lemma. The constant bb is different from the constant bb in Lemma 4.2, and it depends on ker(hn0)\ker(h_{n_{0}}) but not on nn.

Lemma 7.4 (cf Lemma 3.3 and Lemma 3.9 in [FS]).

Suppose ker(hn0)\ker(h_{n_{0}}) is infinite. Then there exists a constant bb with the following property: if nn is sufficiently large, then there are elements u1,u2,u3,u4Snbu_{1},u_{2},u_{3},u_{4}\in S_{n}^{b} that satisfy the conditions (i), (ii) and (iii) in Lemma 4.2, and in addition to that, all uiu_{i} satisfy

  • (iv)

    uihn(kerhn0)u_{i}\in h_{n}(\ker h_{n_{0}}).

Moreover, those elements are such that there are elements u^1,u^2,u^3,u^4η(S)b\hat{u}_{1},\hat{u}_{2},\hat{u}_{3},\hat{u}_{4}\in\eta(S)^{b} with hn(u^i)=uih_{n}(\hat{u}_{i})=u_{i} and uiker(hn0)u_{i}\in\ker(h_{n_{0}}) for each ii.

We remark that the moreover part is rather than an additional property, but immediate from the construction, ie, we first construct u^i\hat{u}_{i} then map them by hnh_{n}. We will call both uiu_{i} and u^i\hat{u}_{i} separators.

Proof.

In the proof there will be several constants bib_{i}, for which we do not try to give explicit values. The important property of those constants is that they do not depend on nn.

First, since ker(hn0)\ker(h_{n_{0}}) is infinite, choose distinct elements r1,,r𝒟+1Lr_{1},\cdots,r_{\mathcal{D}+1}\in L that are in the kernel. Let b1b_{1} be the maximum of the word lengths of the rir_{i} in terms of η(S)\eta(S). If nn is large enough, then the image of those 𝒟+1\mathcal{D}+1 elements by hnh_{n} are all distinct. From now on, we only consider such nn. In the following we fix each such nn and argue. We have 𝒟+1\mathcal{D}+1 distinct elements {hn(ri)}\{h_{n}(r_{i})\} and the word lengths of those w.r.t. SnS_{n} are bounded by b1b_{1}.

Secondly, choose an element gnSnMg_{n}\in S_{n}^{M} that is hyperbolic on XX such that its 10δ10\delta-axis is at distance at most 10δ10\delta from the point yny_{n}. By assumption, such an element exists. Also there is sSns\in S_{n} such that gng_{n} and sgns1sg_{n}s^{-1} are independent. Choose such ss.

Thirdly, we choose one element, rr, from the rir_{i}’s as follows: if there is rir_{i} such that
(I) The element hn(ri)h_{n}(r_{i}) is hyperbolic on XX,
then choose one of such rir_{i} and set r=rir=r_{i}. Otherwise choose rir_{i} with
(II) hn(ri)F(gn)h_{n}(r_{i})\not\in F(g_{n})
and set r=rir=r_{i}. This is clearly possible since |F(gn)|𝒟|F(g_{n})|\leq\mathcal{D}.

Note that the element rr depends on nn. From now on we suppress nn and write gng_{n} as gg.

Now we divide the case into two depending on (I) or (II) in the above. Suppose we are in the case (I). We consider the power gkg^{k} with k=60𝒟k=60\mathcal{D}, then we have (see Lemma 2.8)

gk,sgks1=gksgks1.\langle g^{k},sg^{k}s^{-1}\rangle=\langle g^{k}\rangle*\langle sg^{k}s^{-1}\rangle.

Note that we have

λ(gk)Δn100δ.\lambda(g^{k})\leq\Delta_{n}-100\delta.

This is because since k=60𝒟k=60\mathcal{D} and gSnMg\in S_{n}^{M}, we have

λ(gk)=30λ(g2𝒟)30L(Sn2𝒟M)Δn100δ.\lambda(g^{k})=30\lambda(g^{2\mathcal{D}})\leq 30L(S_{n}^{2\mathcal{D}M})\leq\Delta_{n}-100\delta.

The last inequality is by 𝒟10\mathcal{D}\geq 10 and Δn=100δ+4𝒟L(Sn2𝒟M).\Delta_{n}=100\delta+4\mathcal{D}L(S_{n}^{2\mathcal{D}M}).

Recall that the axes of gk,sgks1g^{k},sg^{k}s^{-1} are at at most 40δ+L(Sn2M𝒟)40\delta+L(S_{n}^{2M\mathcal{D}}) from yny_{n}.

In the proof of Lemma 4.2, we set w=gk,z=sgks1w=g^{k},z=sg^{k}s^{-1} and produce uiu_{i} as words on w,zw,z, but this time we take into account the germs of the element hn(r)h_{n}(r) and choose z,wgksgks1z,w\in\langle g^{k}\rangle*\langle sg^{k}s^{-1}\rangle as follows.

Notice that 6 elements gk,gk,sgks1,sgks1,gksgks1,gksgks1,g^{k},g^{-k},sg^{k}s^{-1},sg^{-k}s^{-1},g^{k}sg^{k}s^{-1},g^{k}sg^{-k}s^{-1}, define six germs at yny_{n} that are mutually opposite since λ(gk)Δn100δ\lambda(g^{k})\leq\Delta_{n}-100\delta as we noted. From the six, choose four distinct germs that are opposite to the germs for hn(r)h_{n}(r) and hn(r)1h_{n}(r)^{-1}. If those two germs are empty, then ignore this condition. Denote those four germs as γ1,γ2,γ3,γ4\gamma_{1},\gamma_{2},\gamma_{3},\gamma_{4}. Now choose w,zgk,sgks1w,z\in\langle g^{k},sg^{k}s^{-1}\rangle such that the germ for w,w1,z,z1w,w^{-1},z,z^{-1} is equivalent to γ1,γ3,γ2,γ4\gamma_{1},\gamma_{3},\gamma_{2},\gamma_{4}, respectively, such that the axes of w,zw,z are at at most 60δ+L(Sn2𝒟M)60\delta+L(S_{n}^{2\mathcal{D}M}) from yny_{n}.

We also arrange that the axes of w,zw,z are not parallel to each other, and no element of GG flips the axes of ww or zz. This is achieved by a similar technique to the one we used to prove Lemma 2.8, so we do not repeat. As in the proof of Lemma 4.2, there exists a constant b2b_{2} that depends only on δ,M\delta,M and 𝒟=D(200δ)\mathcal{D}=D(200\delta) such that the word lengths of w,zw,z w.r.t. SnS_{n} are bounded by b2b_{2}. The constant b2b_{2} does not depend on nn, nor the choice of rr.

Now, choose z^,w^L\hat{z},\hat{w}\in L with hn(z^)=z,hn(w^)=wh_{n}(\hat{z})=z,h_{n}(\hat{w})=w, whose word length w.r.t. η(S)\eta(S) are also bounded by b2b_{2}. Using z^,w^,r\hat{z},\hat{w},r we define u^iL\hat{u}_{i}\in L as follows:

u^1\displaystyle\hat{u}_{1} =w^rw^1z^rz^1w^2rw^2z^rz^1w^19rw^19z^rz^1w^20rw^20,\displaystyle=\hat{w}r\hat{w}^{-1}\cdot\hat{z}r\hat{z}^{-1}\cdot\hat{w}^{2}r\hat{w}^{-2}\cdot\hat{z}r\hat{z}^{-1}\cdots\hat{w}^{19}r\hat{w}^{-19}\cdot\hat{z}r\hat{z}^{-1}\cdot\hat{w}^{-20}r\hat{w}^{20},
u^2\displaystyle\hat{u}_{2} =w^21rw^21z^1rz^w^22rw^22z^1rz^w^40rw^40z1rz^,\displaystyle=\hat{w}^{21}r\hat{w}^{-21}\cdot\hat{z}^{-1}r\hat{z}\cdot\hat{w}^{22}r\hat{w}^{-22}\cdot\hat{z}^{-1}r\hat{z}\cdots\hat{w}^{40}r\hat{w}^{-40}\cdot z^{-1}r\hat{z},
u^3\displaystyle\hat{u}_{3} =z^rz^1w^41rw^41z^rz^1w^42rw^42z^rz^1z^rz^1w^60rw^60,\displaystyle=\hat{z}r\hat{z}^{-1}\cdot\hat{w}^{-41}r\hat{w}^{41}\cdot\hat{z}r\hat{z}^{-1}\cdot\hat{w}^{-42}r\hat{w}^{42}\cdot\hat{z}r\hat{z}^{-1}\cdots\hat{z}r\hat{z}^{-1}\cdot\hat{w}^{-60}r\hat{w}^{60},
u^4\displaystyle\hat{u}_{4} =z^rz^1w^61rw^61z^rz^1w^62rw^62z^rz^1w^80rw^80z^1rz^.\displaystyle=\hat{z}r\hat{z}^{-1}\cdot\hat{w}^{61}r\hat{w}^{-61}\cdot\hat{z}r\hat{z}^{-1}\cdot\hat{w}^{62}r\hat{w}^{-62}\cdots\hat{z}r\hat{z}^{-1}\cdot\hat{w}^{80}r\hat{w}^{-80}\cdot\hat{z}^{-1}r\hat{z}.

The word length of u^i\hat{u}_{i} in terms of η(S)\eta(S) is at most b3b_{3}, which does not depend on nn. Indeed, we may set b3=21b1+1420b2b_{3}=21b_{1}+1420b_{2}.

Clearly, u^i\hat{u}_{i} are in kerhn0\ker h_{n_{0}} since they are products of conjugates of rkerhn0r\in\ker h_{n_{0}}. Finally, define ui=hn(u^i)Snb3u_{i}=h_{n}(\hat{u}_{i})\in S_{n}^{b_{3}}. Then, they satisfy the property (iv).

We need to check those elements satisfy the other properties, (i), (ii) and (iii) in Lemma 4.2. Regarding (i), the germ γ1\gamma_{1} is the germ of ww, the germ γ3\gamma_{3} is the germ of w1w^{-1}, the germ γ2\gamma_{2} is the germ of zz, and the germ γ4\gamma_{4} is the germ of z1z^{-1}. Then u1,u2,u3,u4u_{1},u_{2},u_{3},u_{4} satisfy (i). The property (iii) is a consequence of that the axes of ww and zz are not parallel to each other. We skip details of the arguments since it is similar to Lemma 4.2. We point out that some of the argument slightly differs depending if the germs for hn(r),hn(r)1h_{n}(r),h_{n}(r)^{-1} are defined or empty. In the empty case, we use the property that the axes of w,zw,z are not flipped by any element of GG.

In conclusion, those are desired elements, so in this case we take b=b3b=b_{3} and we are done.

Suppose we are in the case (II) when we chose rr. In this case, we replace rr with another element rr^{\prime} such that hn(r)h_{n}(r^{\prime}) is hyperbolic on XX. We explain how we produce such an rr^{\prime}. First choose an element g^L\hat{g}\in L with hn(g^)=gh_{n}(\hat{g})=g, where the word length of g^\hat{g} is at most MM in terms of η(S)\eta(S). Then we consider an element of the form:

r=rg^Qrg^Q.r^{\prime}=r\hat{g}^{Q}r\hat{g}^{-Q}.

We will show that if Q40𝒟Q\geq 40\mathcal{D}, hn(r)h_{n}(r^{\prime}) is hyperbolic. See Lemma 7.6. Also, since rkerhn0r\in\ker h_{n_{0}}, we have rkerhn0r^{\prime}\in\ker h_{n_{0}}. It is a well-known method to produce a hyperbolic element as a product of two non-hyperbolic elements, and we postpone an explanation on this, and proceed.

But, then the word length of rr^{\prime} in terms of η(S)\eta(S) is at most b5b_{5}, where b5=2b1+60𝒟Mb_{5}=2b_{1}+60\mathcal{D}M, which does not depend on nn.

With the element rr^{\prime} we repeat the same argument as we did for rr in the case (I) and obtain desired uiu_{i} with a bound on the word length, uniformly over all nn, which will finish the proof of Lemma 7.4.

We now explain some details on how to produce rr^{\prime} from rr. For an element kIsom(X)k\in Isom(X), define a set

Min(k)={xX||xk(x)|100δ}.Min(k)=\{x\in X||x-k(x)|\leq 100\delta\}.

This is a kk-invariant set. We state a few standard facts on this set.

  1. (M1)

    If kk is not hyperbolic then Min(k)Min(k) is not empty, since if it is empty, then L(k)>100δL(k)>100\delta, which implies kk is hyperbolic, a contradiction.

  2. (M2)

    If kk is not hyperbolic, then for any point yXy\in X, we have

    |yk(y)|2(d(y,Min(k))400δ).|y-k(y)|\geq 2(d(y,Min(k))-400\delta).

We prove (M2). If d(y,Min(k))400δd(y,Min(k))\leq 400\delta, then nothing to show so suppose d(y,Min(k))>400δd(y,Min(k))>400\delta. Let xMin(k)x\in Min(k) be a point with d(x,y)=d(Min(k),y)d(x,y)=d(Min(k),y).

Let z[x,y]z\in[x,y] be the point with |xz|=350δ|x-z|=350\delta. We claim that zN10δ([k(x),k(y)])z\not\in N_{10\delta}([k(x),k(y)]) and k(z)N10δ([x,y])k(z)\not\in N_{10\delta}([x,y]). To prove the first claim, suppose not, ie, zN10δ([k(x),k(y)])z\in N_{10\delta}([k(x),k(y)]). Let v,w[x,y]v,w\in[x,y] be with |xv|=150δ|x-v|=150\delta and |xw|=200δ|x-w|=200\delta. Then we have k([v,w])N10δ([x,y])k([v,w])\subset N_{10\delta}([x,y]) since |xk(x)|100δ|x-k(x)|\leq 100\delta and zN10δ([k(x),k(y)])z\in N_{10\delta}([k(x),k(y)]). Then we have |vk(v)|50δ|v-k(v)|\leq 50\delta since kk is not hyperbolic. (Otherwise, vv is “pushed” along the geodesic [x,y][x,y] by at least 40δ40\delta by kk, which implies that kk is hyperbolic, impossible.) But it implies that vMin(k)v\in Min(k), which is a contradiction. We showed zN10δ([k(x),k(y)])z\not\in N_{10\delta}([k(x),k(y)]).

By the same argument, we can show k(z)N10δ([x,y])k(z)\not\in N_{10\delta}([x,y]).

Having those two claims, we have

|yk(y)||yz|+|zk(z)|+|k(z)k(y)|100δ2(|yx|350δ)100δ=2(|yx|400δ).|y-k(y)|\geq|y-z|+|z-k(z)|+|k(z)-k(y)|-100\delta\geq 2(|y-x|-350\delta)-100\delta=2(|y-x|-400\delta).

We showed (M2).

We go back to the explanation. For a hyperbolic isometry gg and its 10δ10\delta-axis Ax(g)Ax(g), we consider the nearest points projection in XX to Ax(g)Ax(g). We denote the projection by πg\pi_{g}. For every point xXx\in X, although πg(x)\pi_{g}(x) is not a point, the diameter of πg(x)\pi_{g}(x) is bounded by 100δ100\delta. The following lemma is well-known. See Figure 5.

Lemma 7.5 (Bounded projection).

If gGg\in G is hyperbolic and DD-WPD; and kG\E(g)k\in G\backslash E(g) and kk is not hyperbolic, then the image of Min(k)Min(k) in Ax(g)Ax(g) by the projection πg\pi_{g} is bounded by 2D(100δ)L(g)+200δ2D(100\delta)L(g)+200\delta in diameter.

We prove the lemma for readers’ convenience.

Proof.

Let x,yMin(k)x,y\in Min(k) and suppose pπg(x),qπg(y)p\in\pi_{g}(x),q\in\pi_{g}(y). Assume that |pq||p-q| is larger than 200δ200\delta, since otherwise there is nothing to show. Then [x,p][p,q][q,y][x,p]\cup[p,q]\cup[q,y] is a uniform quasi-geodesic and every point on it is moved by kk by at most 200δ200\delta. Moreover, each point xx on [p,q][p,q] with |xp|,|xq|50δ|x-p|,|x-q|\geq 50\delta is moved by kk by at most 20δ20\delta since kk is not hyperbolic.

Now, consider the points p,q[p,q]p^{\prime},q^{\prime}\in[p,q] with |pp|=|qq|=50δ|p-p^{\prime}|=|q-q^{\prime}|=50\delta. Then since gg is DD-WPD, we have |pq|2D(100δ)L(g)+100δ|p^{\prime}-q^{\prime}|\leq 2D(100\delta)L(g)+100\delta by Lemma 2.5 (2). This is because, otherwise, kE(g)k\in E(g), impossible. It follows that |pq|2D(100δ)L(g)+200δ|p-q|\leq 2D(100\delta)L(g)+200\delta. ∎

Refer to caption
Figure 5. The Min sets and the projection to an axis

The following lemma is also standard, and this is what we need for our purpose. See Figure 5.

Lemma 7.6 (Producing hyperbolic element).

For g,kg,k as in the lemma 7.5, the element kgQkgQkg^{Q}kg^{-Q} is hyperbolic if Q40D(100δ)Q\geq 40D(100\delta).

We also give a brief proof.

Proof.

Set 𝒟=D(100δ)\mathcal{D}=D(100\delta). (Or one can set 𝒟=D(200δ)\mathcal{D}=D(200\delta) as usual. It does not matter since D(100δ)D(200δ)D(100\delta)\leq D(200\delta).) In general, Ax(g)Ax(g) is not exactly gg-invariant, but if L(g)10δL(g)\geq 10\delta by definition. We also know that L(g)50δ/𝒟L(g)\geq 50\delta/\mathcal{D} by Lemma 2.2 (1) for our gg. So, if necessary, by replacing gg by g𝒟g^{\mathcal{D}}, we may assume that Ax(g)Ax(g) is gg-invariant. For simplicity, in the following argument, we assume that Ax(g)Ax(g) is gg-invariant.

Consider the set Min(gQkgQ)Min(g^{Q}kg^{-Q}), which is equal to the set gQ(Min(k))g^{Q}(Min(k)). We consider the projection of those two sets by πg\pi_{g}. Then since Ax(g)Ax(g) is gg-invariant, the projection πg\pi_{g} is gg-equivariant. It implies that πg(Min(gQkgQ))=gQ(πg(Min(k)))\pi_{g}(Min(g^{Q}kg^{-Q}))=g^{Q}(\pi_{g}(Min(k))).

Then by Lemma 7.5, the distance between πg(Min(gQkgQ))\pi_{g}(Min(g^{Q}kg^{-Q})) and πg(Min(k)))\pi_{g}(Min(k))) is at least

QL(g)(2𝒟L(g)+200δ)38𝒟L(g)200δ1700δQL(g)-(2\mathcal{D}L(g)+200\delta)\geq 38\mathcal{D}L(g)-200\delta\geq 1700\delta

since L(g)50δ/𝒟L(g)\geq 50\delta/\mathcal{D}. It follows that the distance between Min(gQkgQ)Min(g^{Q}kg^{-Q}) and Min(k)Min(k) is at least 1600δ1600\delta. It follows that the product of kk and gQkgQg^{Q}kg^{-Q} is hyperbolic (this is a well-known fact in δ\delta-hyperbolic geometry, ie, if the distance between Min(a)Min(a) and Min(b)Min(b) is at least 1000δ1000\delta, then abab is hyperbolic since both Min(a),Min(b)Min(a),Min(b) satisfy the property (M2)). ∎

This finishes the explanation for the part to produce rr^{\prime} from rr, and the case (ii) is done. We proved Lemma 7.4. ∎

We go back to the proof of Proposition 7.2. With Lemma 7.4, the rest is very similar to [FS]. Fix nn that is large enough to apply Lemma 7.4. We explain how we define the map ϕn\phi_{n}.

Let gSph m(G,Sn0)g\in\text{Sph\,}_{m}(G,S_{n_{0}}). Choose a shortest representative word w(g)w(g) of length mm on Sn0S_{n_{0}} for gg. By the bijection hn0h_{n_{0}} between Sn0S_{n_{0}} and η(S)\eta(S), w(g)w(g) canonically gives a word ww of length mm on η(S)\eta(S).

From this word ww, we construct a collection of elements in LL. Given a positive integer kk with 1km11\leq k\leq m-1, we divide the word ww into a prefix of length kk, and a suffix of length mkm-k. The prefix corresponds to an element in LL that we denote wpkw^{k}_{p}, and the suffix corresponds to an element in LL that we denote wskw^{k}_{s}.

Now, from the four separators we constructed in Lemma 7.4 we choose a separator u^\hat{u} for hnh_{n} such that u^\hat{u} is admissible for wpkw^{k}_{p} and u1u^{-1} is admissible for (wsk)1(w^{k}_{s})^{-1}, after we map them to GG by hnh_{n}. We are writing u^\hat{u} instead of uu to indicate that the separator is in LL.

To the pair wpk,wskw^{k}_{p},w^{k}_{s}, we associate the following element in LL: wpku^wskBm+b(L,η(S))w_{p}^{k}\hat{u}w_{s}^{k}\in B_{m+b}(L,\eta(S)).

wwpkwskwpku^wskw\leadsto w_{p}^{k}w_{s}^{k}\leadsto w_{p}^{k}\hat{u}w_{s}^{k}

Note that hn0(wpku^wsk)=hn0(w)h_{n_{0}}(w_{p}^{k}\hat{u}w_{s}^{k})=h_{n_{0}}(w) for all kk since u^ker(hn0)\hat{u}\in\ker(h_{n_{0}}).

In this way, we obtain m1m-1 “words” in LL from ww, but possibly, some of them represent the same elements in LL. To address this issue, we define a subcollection of words, called forbidden words (in somewhat similar way to what we did in Section 4).

Given mm, take nn large enough such that hnh_{n} is injective on B2m(L,η(S))B_{2m}(L,\eta(S)) from now on.

Definition 7.7 (Forbidden words. cf. Def 3.4 in [FS]).

Let ww be a word of length mm on η(S)\eta(S) in the above explanation. We say that a word wpku^wskw^{k}_{p}\hat{u}w^{k}_{s}, from the collection that is built from ww, is forbidden if there exists ff, 1fm1\leq f\leq m such that:

dX(hn(wpku^)(yn),hn(wpf)(yn))15dX(yn,hn(u^)(yn)).d_{X}(h_{n}(w^{k}_{p}\hat{u})(y_{n}),h_{n}(w^{f}_{p})(y_{n}))\leq\frac{1}{5}d_{X}(y_{n},h_{n}(\hat{u})(y_{n})).

We give a bound on the number of forbidden words.

Lemma 7.8 (cf. Lemma 3.5, [FS]).

For mm and each word ww as in Definition 7.7, there are at most 1𝒟+1m\frac{1}{\mathcal{D}+1}m forbidden words of the form: wpku^wskw^{k}_{p}\hat{u}w^{k}_{s} for k=1,,m1k=1,\ldots,m-1.

In the proof of this lemma we use the assumption that hnh_{n} is injective on B2m(L,η(S))B_{2m}(L,\eta(S)) as we did in the proof of Lemma 4.6.

Proof.

The strategy of the proof is same as the proof of Lemma 4.6. In there, we ran the argument in Bm(L,η(S))B_{m}(L,\eta(S)), but here, we do it in the set Z(w)={wpku^wsk|k=1,,m1}Z(w)=\{w^{k}_{p}\hat{u}w^{k}_{s}|k=1,\cdots,m-1\}. Namely, if wpku^wskZ(w)w^{k}_{p}\hat{u}w^{k}_{s}\in Z(w) is forbidden for some kk, then there are two other elements in Z(w)Z(w) that are candidates for non-forbidden elements. Then if at least one of them is forbidden, then there are two other elements in Z(w)Z(w) that are candidates for non-forbidden elements, and so on. We omit details. ∎

Then we have:

Lemma 7.9 (cf. Lemma 3.6 [FS]).

For mm and the word ww as above, the non-forbidden words: wpkv^i,jwskw^{k}_{p}\hat{v}_{i,j}w^{k}_{s}, for all kk, 1km11\leq k\leq m-1, are distinct elements in LL.

Proof.

The proof is nearly identical to the proof of Lemma 3.6 in [FS], and we omit it. ∎

As in Section 4, we define adequate elements in Z(w)Z(w). We choose a maximal subset in the set of non-forbidden elements in Z(w)Z(w) such that for any two distinct elements z1,z2z_{1},z_{2} in the subset, hn(z1),hn(z2)h_{n}(z_{1}),h_{n}(z_{2}) are not in the same coset w.r.t. F(hn(ui))F(h_{n}(u_{i})) for any separator uiu_{i}. We call those elements adequate elements. This notion depends on nn. Then, as before,

|{non-forbidden elements}|𝒟4|{adequate elements}|.\frac{|\{\text{non-forbidden elements}\}|}{\mathcal{D}^{4}}\leq|\{\text{adequate elements}\}|.

This is because, as we explained, there are only four separators, u^\hat{u}, and |F(hn(u^))|𝒟|F(h_{n}(\hat{u}))|\leq\mathcal{D}. In conclusion, since |Z(w)|=m1|Z(w)|=m-1,

(5) m1𝒟4(𝒟+1)|{adequate elements in Z(w)}|.\frac{m-1}{\mathcal{D}^{4}(\mathcal{D}+1)}\leq|\{\text{adequate elements in }Z(w)\}|.

Using adequate elements, we construct a collection of feasible words in LL.

Definition 7.10 (Feasible words in LL. cf. Definition 3.7 [FS]).

Let m,qm,q be positive integers. Let ww be a word of length mqmq on η(S)\eta(S) that is associated to an element gSph mq(G,Sn0)g\in\text{Sph\,}_{mq}(G,S_{n_{0}}) as in the above discussion. We present ww as a concatenation of qq subwords of length mm: w=w(1)w(q)w=w(1)\ldots w(q).

Then for any choice of integers: k1,,kqk_{1},\ldots,k_{q} with 1ktm11\leq k_{t}\leq m-1, and t=1,,qt=1,\ldots,q, for which all the elements, w(t)pktvtw(t)sktw(t)^{k_{t}}_{p}v^{t}w(t)^{k_{t}}_{s}, are adequate (here we drop the “hat” from vtv^{t} deliberately although they are in LL to avoid confusion since we want to use it right in the below) , we associate a feasible word (of type qq) on η(S)\eta(S) (in LL):

w(1)pk1v1w(1)sk1v^1w(2)pk2v2w(2)sk2v^2w(q)pkqvqw(q)skq,w(1)^{k_{1}}_{p}v^{1}w(1)^{k_{1}}_{s}\,\hat{v}^{1}\,w(2)^{k_{2}}_{p}v^{2}w(2)^{k_{2}}_{s}\,\hat{v}^{2}\,\ldots w(q)^{k_{q}}_{p}v^{q}w(q)^{k_{q}}_{s},

where for each tt, 1tq11\leq t\leq q-1, v^t\hat{v}^{t} is one of the separators from Lemma 4.2 such that v^t\hat{v}^{t} is admissible for w(t)sktw(t)^{k_{t}}_{s} and (v^t)1(\hat{v}^{t})^{-1} is admissible for (w(t+1)pkt+1)1\left(w(t+1)^{k_{t+1}}_{p}\right)^{-1}.

Finally we define the map ϕn\phi_{n}. Suppose positive integers m,qm,q are given. Then for gSph mq(G,Sn0)g\in\text{Sph\,}_{mq}(G,S_{n_{0}}), choose one shortest representative w(g)w(g) of length mqmq on Sn0S_{n_{0}}, which defines a word w~(g)\tilde{w}(g) of length mqmq on η(S)\eta(S) as in the definition 7.10. From w~(g)\tilde{w}(g) we produce feasible words on η(S)\eta(S), which define feasible elements in LL. Note that those elements are in Bq(m+2b)(L,η(S))B_{q(m+2b)}(L,\eta(S)) and mapped to gg by hn0h_{n_{0}}. We denote this collection as ϕn(g)\phi_{n}(g).

We have :

Lemma 7.11 (cf. Lemma 3.8 [FS]).

For any positive integers m,qm,q, the feasible elements in the collection ϕn(g)\phi_{n}(g) we obtain for each gSph mq(G,Sn0)g\in\text{Sph\,}_{mq}(G,S_{n_{0}}) are all distinct in LL.

Moreover, all the feasible elements obtained from all the elements gg in Sph mq(G,Sn0)\text{Sph\,}_{mq}(G,S_{n_{0}}) are all distinct in LL.

In this lemma, nn must be large enough in the sense that Lemma 7.4 applies and also, for the given mm, the map hnh_{n} is injective on B2m(L,η(S))B_{2m}(L,\eta(S)) (cf. Lemma 4.5). We will choose such nn in the proof.

Proof.

The proof of the first sentence is similar to the proof of Lemma 4.8. So we omit it. (cf. the proof of Lemma 3.8 [FS].) Then the moreover part immediately follows. ∎

We note that for each gg, we have

(6) (m1𝒟4(𝒟1))qϕn(g).\left(\frac{m-1}{\mathcal{D}^{4}(\mathcal{D}-1)}\right)^{q}\leq\phi_{n}(g).

This is from the lower bound (5) on the number of adequate elements and the lemma.

We finish the proof of the proposition 7.2. We review the setting. The constant n0n_{0} is given to start with. We want to show e(G,fn0(S))<e(L,η(S))e(G,f_{n_{0}}(S))<e(L,\eta(S)). Then the constant bb is given by Lemma 7.4, which does not depend on nn. Choose nn large enough so that we can apply Lemma 7.4.

Then choose mm such that

log(m1)>2blog(e(L,η(S))+log(𝒟4(𝒟1)).\log(m-1)>2b\log(e(L,\eta(S))+\log(\mathcal{D}^{4}(\mathcal{D}-1)).

This implies:

log(m1)>2blog(e(G,Sn0))+log(𝒟4(𝒟1)).\log(m-1)>2b\log(e(G,S_{n_{0}}))+\log(\mathcal{D}^{4}(\mathcal{D}-1)).

Then choose n>0n>0 larger if necessary such that hnh_{n} is injective on B2m(L,η(S))B_{2m}(L,\eta(S)). We need this to apply Lemma 7.9.

Now, for all q>0q>0, by combining the lower bound (6) and (the moreover part of) Lemma 7.11, we have

|Sph mq(G,Sn0)|(m1𝒟4(𝒟1))q|Bq(m+2b)(L,η(S))|.|\text{Sph\,}_{mq}(G,S_{n_{0}})|\left(\frac{m-1}{\mathcal{D}^{4}(\mathcal{D}-1)}\right)^{q}\leq|B_{q(m+2b)}(L,\eta(S))|.

From this,

loge(L,η(S))\displaystyle\log e(L,\eta(S)) limqlog(|Sph mq(G,Sn0)|(m1𝒟4(𝒟1))q)q(m+2b)\displaystyle\geq\lim_{q\to\infty}\frac{\log(|\text{Sph\,}_{mq}(G,S_{n_{0}})|(\frac{m-1}{\mathcal{D}^{4}(\mathcal{D}-1)})^{q})}{q(m+2b)}
=limqlog(|Sph mq(G,Sn0)|)q(m+2b)+q(log(m1)log(𝒟4(𝒟1)))q(m+2b)\displaystyle=\lim_{q\to\infty}\frac{\log(|\text{Sph\,}_{mq}(G,S_{n_{0}})|)}{q(m+2b)}+\frac{q(\log(m-1)-\log(\mathcal{D}^{4}(\mathcal{D}-1)))}{q(m+2b)}
=log(e(G,Sn0))mm+2b+log(m1)log(𝒟4(𝒟1))m+2b\displaystyle=\log(e(G,S_{n_{0}}))\frac{m}{m+2b}+\frac{\log(m-1)-\log(\mathcal{D}^{4}(\mathcal{D}-1))}{m+2b}
>log(e(G,Sn0)).\displaystyle>\log(e(G,S_{n_{0}})).

The last inequality is by the way we chose mm. Hence, e(L,η(S))>e(G,Sn0)e(L,\eta(S))>e(G,S_{n_{0}}). We proved Proposition 7.2. ∎

7.4. Family version

We state a family version of Proposition 7.2. Let δ,M\delta,M be constants and D(ϵ)D(\epsilon) a function for WPD. Suppose XnX_{n} is δ\delta-hyperbolic, a group GnG_{n} acts on XnX_{n} and SnS_{n} is a finite generating set of GnG_{n} such that SnMS_{n}^{M} contains a hyperbolic element on XnX_{n} that is DD-WPD.

Suppose that |Sn|=|S_{n}|=\ell for all nn, and let FF be a free group of rank \ell with a free generating set SS. Let fn:(F,S)(Gn,Sn)f_{n}:(F,S)\to(G_{n},S_{n}) be a surjection with a bijection fn(S)=Snf_{n}(S)=S_{n}. Assume that the sequence {fn}\{f_{n}\} converges to η:(F,S)(L,η(S))\eta:(F,S)\to(L,\eta(S)).

Also, assume that for all nn, there is a surjection hn:(L,η(S))(Gn,Sn)h_{n}:(L,\eta(S))\to(G_{n},S_{n}) with fn=hnηf_{n}=h_{n}\circ\eta. As in Proposition 4.9, we put this as an assumption. Then we have the following generalizing Proposition 7.2. The proof is identical and we omit it.

Proposition 7.12.

Assume that for all nn, there is a surjection hn:(L,η(S))(Gn,Sn)h_{n}:(L,\eta(S))\to(G_{n},S_{n}) with fn=hnηf_{n}=h_{n}\circ\eta. If ker(hn0)\ker(h_{n_{0}}) is infinite for some n0n_{0}, then

e(Gn0,fn0(S))<e(L,η(S)).e(G_{n_{0}},f_{n_{0}}(S))<e(L,\eta(S)).

7.5. Finiteness for ΘX(G)\Theta_{X}(G)

A finiteness result similar to Theorem 7.1 holds for subgroups. It is known for hyperbolic groups, [FS, Theorem 5.3].

Let S1,S2GS_{1},S_{2}\subset G be two finite subsets. Let Hi=Si<GH_{i}=\langle S_{i}\rangle<G be the subgroup generated by SiS_{i}. We say (H1,S1)(H_{1},S_{1}) and (H2,S2)(H_{2},S_{2}) are isomorphic if there is a bijection between S1,S2S_{1},S_{2} that induces an isomorphism between H1,H2H_{1},H_{2}.

Theorem 7.13 (Finiteness in the subgroups case).

Assume the same condition on GG as in Theorem 7.1. Moreover, we assume that if SS is a finite set of GG such that S\langle S\rangle contains a hyperbolic isometry on XX, then SMS^{M} contains a hyperbolic isometry that is DD-WPD. Let ρΘX(G)\rho\in\Theta_{X}(G), Then there are at most finitely many (H,S)(H,S), up to isomorphism, such that SGS\subset G is finite, H=SH=\langle S\rangle, HH contains a hyperbolic element on XX, and ρ=e(H,S)\rho=e(H,S).

The proof is nearly identical to the proof of Theorem 7.1 and we only need to modify the setting from the entire group GG to subgroups.

Proof.

To argue by contradiction, let ρΘX(G)\rho\in\Theta_{X}(G) and suppose that there are infinitely many distinct, up to isomorphism, (Hn,Sn)(H_{n},S_{n}) with e(Hn,Sn)=ρe(H_{n},S_{n})=\rho such that HnH_{n} contains a hyperbolic isometry on XX. Note that by assumption, SnMS_{n}^{M} contains a hyperbolic isometry that is DD-WPD.

As before, by Proposition 2.10, passing to a subsequence, one may assume that there is \ell with |Sn|=|S_{n}|=\ell for all nn. Then we obtain fn:FGf_{n}:F\to G with fn(S)=Snf_{n}(S)=S_{n}, where FF is the free group on SS with |S|=|S|=\ell. Then, passing to a subsequence again, fnf_{n} converges to a limit group (L,η(S))(L,\eta(S)) with η:FL\eta:F\to L. Since GG is equationally Noetherian, by Lemma 1.11, passing to a further subsequence, we may assume that there are hn:LHn<Gh_{n}:L\to H_{n}<G with hnη=fnh_{n}\circ\eta=f_{n} for all nn.

First, by Proposition 6.2, we have that e(Hn,Sn)=e(L,η(L))e(H_{n},S_{n})=e(L,\eta(L)) for all nn.

On the other hand, we prove a version of Proposition 7.2 for subgroups: if ker(hn0)\ker(h_{n_{0}}) is infinite for some n0n_{0}, then e(Hn0,Sn0)<e(L,η(S))e(H_{n_{0}},S_{n_{0}})<e(L,\eta(S)). The proof is same and we only outline it. As before set 𝒟=D(100δ)\mathcal{D}=D(100\delta). First, since HnH_{n} contains a hyperbolic isometry on XX that is DD-WPD, each HnH_{n} contains the maximal finite normal subgroup, which we denote by NHnN_{H_{n}} with |NHn|2𝒟|N_{H_{n}}|\leq 2\mathcal{D}, (by Lemma 7.3).

The key step is to prove a lemma similar to Lemma 7.4. The argument is same. We use that SnMS_{n}^{M} has a hyperbolic and DD-WPD element for all nn. Then, as before, for all sufficiently large nn, there exists an element r(n)ker(hn0)r(n)\in\ker(h_{n_{0}}) such that hn(r(n))h_{n}(r(n)) is hyperbolic on XX, and that the word length of r(n)r(n) in terms of η(S)\eta(S) is bounded uniformly on nn. Then then rest is same as proving the lemma. Once we have the lemma, the rest is same to show the proposition. (We point out that this is a special case of Proposition 7.12, where XnX_{n} are common. But we did not describe the details of the argument for that.)

Combining those two, we conclude that ker(hn)\ker(h_{n}) is finite for all nn.

Finally, there are only finitely many possibilities for ker(hn)\ker(h_{n}) since it is contained in NLN_{L} and |NL|2𝒟|N_{L}|\leq 2\mathcal{D}. It implies that the desired finiteness for (Hn,Sn)(H_{n},S_{n}) holds as before. ∎

7.6. Examples

We give some examples of Theorem 7.1 and Theorem 7.13. We start with relatively hyperbolic groups.

Theorem 7.14 (Finiteness for relatively hyperbolic groups).

Let GG be a group that is hyperbolic relative to a collection of subgroups {P1,,Pn}\{P_{1},\cdots,P_{n}\}. Suppose GG is not virtually cyclic, and not equal to PiP_{i} for any ii. Suppose each PiP_{i} is finitely generated and equationally Noetherian. Then for each ρξ(G)\rho\in\xi(G) there are at most finitely many finite generating sets SnS_{n} of GG, up to Aut (G)\text{Aut\,}(G), s.t. e(G,Sn)=ρe(G,S_{n})=\rho.

Moreover, for each ρΘnonelem.(G)\rho\in\Theta_{{\rm non-elem.}}(G), there are at most finitely many (Hn,Sn)(H_{n},S_{n}), up to isomorphism, s.t. e(Hn,Sn)=ρe(H_{n},S_{n})=\rho, where SnGS_{n}\subset G is finite and Hn=SnH_{n}=\langle S_{n}\rangle is not conjugate into any PiP_{i}.

Proof.

Let XX be a hyperbolic space on which GG acts as we explained in Section 5.1. We also verified that all the assumption of Theorem 7.1 for GG and the action of GG on XX. Recall that the action of GG on XX is uniformly WPD by Lemma 5.2. It implies the first part of the theorem.

For the moreover part, we apply Theorem 7.13. As Lemma 6.4 shows, Θnonelem.(G)=ΘX(G)\Theta_{{\rm non-elem.}}(G)=\Theta_{X}(G), which implies the conclusion. ∎

Theorem 7.14 immediately implies the following as Theorem 6.5 implies Theorem 6.6:

Theorem 7.15 (Finiteness for lattices).

Let GG be a group in Theorem 5.5. Then for each ρξ(G)\rho\in\xi(G) there are at most finitely many finite generating sets SnS_{n}, up to Aut (G)\text{Aut\,}(G), s.t. e(G,Sn)=ρe(G,S_{n})=\rho.

Moreover, for each ρΘ(G)\rho\in\Theta(G), there are at most finitely many (Hn,Sn)(H_{n},S_{n}), up to isomorphism, s.t. e(Hn,Sn)=ρe(H_{n},S_{n})=\rho, where SnGS_{n}\subset G is finite and Hn=SnH_{n}=\langle S_{n}\rangle.

Lastly we record the following (potential) example:

Theorem 7.16 (Finiteness for MCG).

Let MCG=MCG(Σ)MCG=MCG(\Sigma) be the mapping class group of a compact orientable surface Σ\Sigma. Assume that it is equationally Noetherian.

Then for each ρξ(MCG)\rho\in\xi(MCG) there are at most finitely many finite generating sets SnS_{n}, up to Aut (MCG)\text{Aut\,}(MCG), such that e(MCG,Sn)=ρe(MCG,S_{n})=\rho.

Moreover, for each ρΘlarge(MCG)\rho\in\Theta_{{\rm large}}(MCG), there are at most finitely many (Hn,Sn)(H_{n},S_{n}), up to isomorphism, such that e(Hn,Sn)=ρe(H_{n},S_{n})=\rho, where SnMCGS_{n}\subset MCG is finite and Hn=SnH_{n}=\langle S_{n}\rangle is a large subgroup.

Proof.

As we explained in Section 5.3, the action of MCG(Σ)MCG(\Sigma) on the curve graph X=𝒞(Σ)X=\mathcal{C}(\Sigma) satisfies the assumption of Theorem 7.1. It is uniformly WPD.

For the moreover part, the conclusion holds for ΘX(MCG)\Theta_{X}(MCG) by Theorem 7.13. But, as we said in the proof of Theorem 6.7, we have ΘX(MCG)=Θlarge(MCG)\Theta_{X}(MCG)=\Theta_{{\rm large}}(MCG), so that the conclusion holds. ∎

References

  • [AL] G. N. Arzhantseva and I. G. Lysenok, A lower bound on the growth of word hyperbolic groups, Jour. of the LMS 73 (2006), 109-125.
  • [BMR] Gilbert Baumslag, Alexei Myasnikov, Vladimir Remeslennikov, Algebraic geometry over groups. I. Algebraic sets and ideal theory. J. Algebra 219 (1999), no. 1, 16–-79.
  • [BMR2] Gilbert Baumslaga, Alexei Myasnikov, Vitaly Roman’kov. Two Theorems about Equationally Noetherian Groups. Journal of Algebra Volume 194, Issue 2, Pages 654-664
  • [BeF] Mladen Bestvina, Koji Fujiwara, Bounded cohomology of subgroups of mapping class groups. Geom. Topol. 6 (2002), 69-–89.
  • [BeF2] Mladen Bestvina, Koji Fujiwara. Handlebody subgroups in a mapping class group. In the Tradition of Ahlfors–Bers, VII. Contemporary Mathematics Volume 696, 2017, Pages 30-50. AMS.
  • [Bo1] B. H. Bowditch, Relatively hyperbolic groups. Internat. J. Algebra Comput. 22 (2012), no. 3, 66 pp.
  • [Bo] Brian H. Bowditch, Tight geodesics in the curve complex. Invent. Math. 171 (2008), no. 2, 281-–300.
  • [BrF] E. Breuillard, K. Fujiwara. On the joint spectral radius for isometries of non-positively curved spaces and uniform growth. to appear in Annales de l’institut Fourier.
  • [Br] Roger M. Bryant, The verbal topology of a group. J. Algebra 48 (1977), no. 2, 340–-346.
  • [DGO] F. Dahmani, V. Guirardel, D. Osin, Hyperbolically embedded subgroups and rotating families in groups acting on hyperbolic spaces. Mem. Amer. Math. Soc. 245 (2017), no. 1156, 152 pp.
  • [De] Thomas Delzant, Kaehler groups, \mathbb{R}-trees, and holomorphic families of Riemann surfaces. Geom. Funct. Anal. 26 (2016), no. 1, 160-–187.
  • [DrS] Cornelia Druţ̧u, Mark Sapir, Tree-graded spaces and asymptotic cones of groups. With an appendix by Denis Osin and Mark Sapir. Topology 44 (2005), no. 5, 959-–1058.
  • [EMO] Alex Eskin, Shahar Mozes, Hee Oh, On uniform exponential growth for linear groups. Invent. Math. 160 (2005), no. 1, 1-–30.
  • [Fa] B. Farb, Relatively hyperbolic groups. Geom. Funct. Anal. 8 (1998), no. 5, 810-–840.
  • [F] K. Fujiwara, Subgroups generated by two pseudo-Anosov elements in a mapping class group. I. Uniform exponential growth. Groups of diffeomorphisms, 283–-296, Adv. Stud. Pure Math., 52, Math. Soc. Japan, 2008.
  • [FS] K. Fujiwara, Z. Sela. The rates of growth in a hyperbolic group. to appear in Invent. Math.
  • [G] M. Gromov. Hyperbolic groups. Essays in group theory, 75–263, Math. Sci. Res. Inst. Publ., 8, Springer, New York, 1987.
  • [GHL] Daniel Groves, Michael Hull, Hao Liang. Homomorphisms to 3-manifold groups. arXiv:2103.06095
  • [GrH] D. Groves, M. Hull, Homomorphisms to acylindrically hyperbolic groups I: Equationally noetherian groups and families. Trans. Amer. Math. Soc. 372 (2019), no. 10, 7141-–7190.
  • [Gu] Victor S. Guba. Equivalence of infinite systems of equations in free groups and semigroups to finite subsystems. Mat. Zametki, 48(3):321-–324, 1986.
  • [K] M. Koubi, Croissance uniforme dans les groupes hyperboliques, Ann. Inst. Fourier 48 (1998), 1441–1453.
  • [Ma] Johanna Mangahas, A recipe for short-word pseudo-Anosovs. Amer. J. Math. 135 (2013), no. 4, 1087–-1116.
  • [Ma2] Johanna Mangahas, Uniform uniform exponential growth of subgroups of the mapping class group. Geom. Funct. Anal. 19 (2010), no. 5, 1468-–1480.
  • [MM1] Howard A. Masur, Yair N. Minsky, Geometry of the complex of curves. I. Hyperbolicity. Invent. Math. 138 (1999), no. 1, 103-–149.
  • [MO] Ashot Minasyan, Denis Osin. Acylindrical hyperbolicity of groups acting on trees. Mathematische Annalen 362, (2015), 1055–1105
  • [NS] Hoang Thanh Nguyen, Hongbin Sun, Subgroup distortion of 3-manifold groups. Trans. Amer. Math. Soc. 373 (2020), no. 9, 6683–-6711.
  • [O] D. Osin. Acylindrically hyperbolic groups. Trans. Amer. Math. Soc. 368 (2016), no. 2, 851–-888.
  • [RW] C. Reinfeldt, R. Weidmann, Makanin-Razborov diagrams for hyperbolic groups. Ann. Math. Blaise Pascal 26 (2019), no. 2, 119-–208.
  • [Sa] A. Sambusetti. Growth tightness of free and amalgamated products. Ann. Sci. Ecole Norm. Sup. serie 35 (2002), no. 4, 477–-488.
  • [S] Z. Sela, Diophantine geometry over groups I: Makanin-Razborov diagrams, Publication de la IHES 93 (2001), 31-105.
  • [Se] Jean-Pierre Serre, Trees. Translated from the French by John Stillwell. Springer-Verlag, Berlin-New York, 1980.
  • [Si] A. Sisto. Projections and relative hyperbolicity. Enseign. Math. (2) 59 (2013), no. 1-2, 165-–181.
  • [W] J. S. Wilson, On exponential growth and uniformly exponential growth for groups, Inventiones math. 155 (2004), 287-303.
  • [WZ] Henry Wilton, Pavel Zalesskii, Profinite detection of 3-manifold decompositions. Compos. Math. 155 (2019), no. 2, 246–259.
  • [X] Xiangdong Xie, Growth of relatively hyperbolic groups. Proc. Amer. Math. Soc. 135 (2007), no. 3, 695–-704.