The rates of growth in an acylindrically hyperbolic group
Abstract.
Let be an acylindrically hyperbolic group on a -hyperbolic space . Assume there exists such that for any finite generating set of , the set contains a hyperbolic element on . Suppose that is equationally Noetherian. Then we show the set of the growth rates of 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 -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.
1. Introduction
1.1. Definitions and results
Let be a finitely generated group with a finite generating set . We always assume that . Let be the set of elements in whose word lengths are at most with respect to the generating set . We also denote instead of . Let . The exponential growth rate of is defined to be:
A finitely generated group has exponential growth if there exists a finite generating set such that . The group has uniform exponential growth if there exists , such that for every finite generating set , .
Given a finitely generated group , we define:
where the infimum is taken over all the finite generating sets of . Since there are finitely generated groups that have exponential growth, but do not have uniform exponential growth [W], the infimum, , is not always obtained by a finite generating set of a finitely generated group.
Given a finitely generated group , we further define the following set in :
where runs over all the finite generating sets of . The set 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 of a non-elementary hyperbolic group 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 acts on a -hyperbolic space acylindrically, and the action is non-elementary. Assume that there exists a constant such that for any finite generating set of , the set contains a hyperbolic element on . Assume that is equationally Noetherian. Then, is a well-ordered set.
In particular, is realized by some , ie, .
The theorem holds under a weaker assumption, namely, we may replace the acylindricity of the action by that 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 be a group that is hyperbolic relative to a collection of subgroups . Suppose is not virtually cyclic, and not equal to for any . Suppose each is finitely generated and equationally Noetherian. Then is well-ordered.
As examples of this theorem, we prove:
Theorem 1.3 (Rank-1 lattices, Theorem 5.5).
Let be one of the following groups:
-
(1)
A lattice in a simple Lie group of rank-1.
-
(2)
The fundamental group of a complete Riemannian manifold of finite volume such that there exist with , where denotes the sectional curvature.
Then is well-ordered.
Another family of examples are -manifold groups.
Theorem 1.4 (Theorem 5.7).
Let be a closed orientable -manifold, and . If is one of the following, then has exponential growth and is well-ordered.
-
(1)
is not irreducible and is not isomorphic to .
-
(2)
is irreducible, not a torus bundle over a circle, and has a non-trivial JSJ-decomposition.
-
(3)
admits hyperbolic geometry.
-
(4)
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 as in Theorem 1.1. Then for any , up to the action of , there are at most finitely many finite generating sets such that .
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 that satisfies the assumption in Theorem 1.1, where we do not need that is equationally Noetherian, there exists a constant such that for any finite generating set of , we have
The constant depends only on 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 . Define
The set is countable and contains . If is a hyperbolic group, it is known by [FS, Section 5] that is well-ordered.
Similarly, we prove:
We also prove:
Theorem 1.7 (Finiteness, Theorem 7.15).
Let be one of the groups in Theorem 1.3. Then for each , there are at most finitely many , up to isomorphism, such that is a finite generating set of with .
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 on a metric space is acylindrical if for any , there exist and such that for all with , the set
contains at most elements.
A group is called an acylindrically hyperbolic group, [O], if it acts on some -hyperbolic space such that the action is acylindrical and non-elementary. Here, we say the action is elementary if the limit set of in the Gromov boundary has at most two points. If the action is non-elementary, it is known that 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 be a group and a finitely generated (or countable) group. Let be the set of all homomorphisms from to .
A sequence of homomorphisms from to is stable if for each , either for all sufficiently large ; or for all sufficiently large . If the sequence is stable, then the stable kernel of the sequence, , is defined by
We call the quotient a -limit group, or the limit group over , associated to , and the homomorphism the limit homomorphism. We say the sequence converges to .
Let be a group and the free group on . For an element and , let denote the element after we substitute every with and by in . Given a subset , define
is called a system of equations (with the set of variables), and is called the algebraic set over defined by . We sometimes suppress from .
Definition 1.9 (Equationally Noetherian).
A group is equationally Noetherian if for every and every subset in , there exists a finite subset such that .
Remark 1.10.
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 be the limit map of a sequence of homomorphisms, . Suppose is equationally Noetherian. Then for sufficiently large , factors through , namely, there exists a homomorphism such that .
Proof.
Let and suppose . Let be a set of defining relations for . In general this is an infinite set. Each is a word on , so that we can see as a system of equations with the variable set. Since is equationally Noetherian, there is a finite subset such that , namely, every solution (an element in ) for is a solution for .
Now, since in for a large enough , we have in since is the limit of . By the same reason, since is a finite set, there exists such that for every , we have in for all . In other words, if , then . But since , this implies that if , then , namely, in for all . Since is a system of defining relations for , it implies that each with factors through . ∎
Regarding Theorem 1.1, there exists a finitely generated group that acylindrically acts on a -hyperbolic space, in fact a simplicial tree , in the non-elementary way such that for any finite generating set , the set contains a hyperbolic isometry on ; and that there is no finite generating set with . In particular is not well-ordered.
The following example is pointed out by Ashot Minasyan. A group is called Hopfian if every surjective homomorphism is an isomorphism.
Example 1.12.
Take a finitely generated non-Hopfian group, , for example , (a Baumslag-Solitar group). Put . Then is non-Hopfian, ie, there exists a surjection that is not an isomorphism. It is a standard fact that a finitely generated equationally Noetherian group is Hopfian, therefore is not equationally Noetherian, so that Theorem 1.1 does not apply to .
But all other assumptions in the theorem are satisfied. Let be the Bass-Serre tree for . The tree is 0-hyperbolic. The action of on is acylindrical and non-elementary. Also, for any finite generating set of , it is a well-known lemma (due to Serre, cf. [BrF]) that contains a hyperbolic isometry on .
On the other hand, it is known, [Sa], that a finitely generated group that is a free product is growth-tight, namely, for any surjective homomorphism that is not an isomorphism, and for any finite generating set of , . Now, it follows that is not achieved by any , since if it did, then take such . But then, the non-isomorphic, surjective homomorphism in the above would imply , 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 , we consider a graph for 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, , in the arguments. But, we do not lose generality if we assume to be a graph since we can consider the -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 acts on a geodesic space by isometries. Choose a base point and for , put
is called the minimal displacement. is called the translation length, and does not depend on the choice of , and for any we have . We also have
The first inequality is trivial and we leave the second as an exercise (for example, use [AL, Corollary 1]). For a finite set and , define
then define
We recall a few definitions and facts from -hyperbolic spaces. An isometry of a hyperbolic space is called elliptic if the orbit of a point by is bounded, and hyperbolic if there are and such that for any , we have . The element is hyperbolic iff .
A hyperbolic isometry is associated with a bi-infinite quasi-geodesic, , called an axis in . If there exists a bi-infinite geodesic that is invariant by , that would be an ideal choice for an axis, but that is not always the case.
As a remedy, if , take a point where is achieved. Then take a geodesic between and and consider the union of its -orbit, which defines a -invariant path, (see for example [De]). If , then take such that and apply the construction to , and use this path for , which is not -invariant. We denote this axis as in this paper. Also, for with , we may also take as an axis for .
For , an axis is not unique, but uniformly (over all hyperbolic ) quasi-geodesic, such that for any two points , the Hausdorff distance between the segment between on and a geodesic between , , is at most . Also, if for is finite, then it is bounded by , where is the Hausdorff distance. We sometimes call a -axes. We consider a direction on the -axis using the action of .
A hyperbolic isometry defines two limit points in , the visual(Gromov) boundary of , by , where is a base point. We say two hyperbolic isometries are independent if and 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, -WPD).
Let act on a -hyperbolic space . Suppose that is hyperbolic on . We say is WPD if there is a -axis, , of such that for any , there exists such that for any with , the number of the elements in the following set is at most :
(1) |
In application we often take . If we want to make the function explicit, we say that is -WPD, or WPD w.r.t. . We say that is a function for WPD.
If there is a function such that if a set of hyperbolic elements in are -WPD, then we say they are uniformly WPD, or uniformly -WPD. If all hyperbolic elements in are uniformly -WPD, then we say the action is uniformly (-)WPD.
Some remarks are in order. The notion of WPD (weak proper discontinuity) was introduced in [BeF], where the function is not used, but the definitions are equivalent.
If is -WPD, then it is -WPD for any function such that for all . So, without loss of generality, we assume that does not decrease when we increase . We often use the value in this paper, for example, see Lemma 2.2. For convenience we also assume that , which we use in the proof of Lemma 7.4.
The choice of a -axis is not important in the definition. Also, one can use -axis for any . It only changes the function . 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 such that for any hyperbolic element , we have . 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 ).
Suppose acts on a -hyperbolic space . Let be hyperbolic with a -axis .
-
(1)
If is -WPD, then .
-
(2)
If the action is acylindrical, then .
Proof.
(1) Set . Let with . Then there must be some with such that or . This is because otherwise, all the elements , which are distinct elements, are contained in the set
This is impossible since the contains at most elements. Now suppose, say, satisfies . Then it implies that . Since , we have .
(2) The argument is similar to (1). Take with . If for some , then . This implies, by acylindricity, there must be with s.t. . It follows , so that . ∎
The following lemma is straightforward.
Lemma 2.3 (Acylindricity implies uniform WPD).
If an action of on a -hyperbolic space is acylindrical, then it is uniformly WPD.
Proof.
Suppose is hyperbolic on , and let be a -axis. Let be the acylindricity constants. Also, let be a uniform bound for by Lemma 2.2 (2). does not depend on nor . Suppose is given. Let be a smallest integer with . Then . The constant does not depend on . Then for any and , we have . By acylindricity, there are at most elements which simultaneously move each of by at most . Put , then the action is uniformly -WPD. ∎
We state a lemma which is useful for us.
Lemma 2.4.
Proof.
Let be a -axis of . Suppose is given. We may assume . Take such that . Take between and with . Then .
Let be the collection of elements in such that and . If , then and ; and are in the -neighborhood of . It implies that contains a subset that contains at most elements such that for any , one can find with . To see it, consider points with such that is the mid point of the segment between on , which we denote by . Then the points with is contained in the -neighborhood of . One should imagine that this neighborhood is a narrow tube around . Now by a pigeon hole argument, one can find a desirable subset . (Notice that , so that one needs to find a point near (i.e. at most ) a given point , which is possible.)
But implies . This is because and both and are contained in the -neighborhood of . We have shown that the element moves both by at most . By our assumption, there are at most possibilities for such element. In conclusion, contains at most elements. We proved that is WPD.
We compute a WPD-function for , which we denote by . By assumption, we may set . First,
Next,
by Lemma 2.2 (1). So, if , we set
∎
2.3. Elementary closure
Suppose acts on a hyperbolic space and let be a hyperbolic isometry with an axis . The elementary closure of is defined by
It turns out that is a subgroup of . Clearly, .
We denote the -neighborhood of a subset by .
Lemma 2.5 (parallel axes).
Suppose acts on a -hyperbolic space . Let be hyperbolic with a -axis . Let , then
-
(1)
If , then .
-
(2)
Assume that is -WPD. If , then the diameter of is at most
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, . Since both are -axes, we have a desired bound.
(2) Suppose not. Suppose that the direction of coincide along the parallel part. Set . Take near one end of the intersection such that for are in the -neighborhood of the intersection. This is possible since the intersection is long enough. Consider the points . Then . Letting with act on , we have
for all .
But since is -WPD, there are at most such elements, so that it must be that for some , we have , so that and commute. It implies that and are parallel, a contradiction.
If the direction for and are opposite, consider instead of , and the rest is same. ∎
We quote a fact.
Proposition 2.6 (elementary closure).
Suppose acts on a -hyperbolic space . Let be a hyperbolic element on . Assume is -WPD. Then,
-
(1)
is virtually , and contains as a finite index subgroup.
-
(2)
If is a hyperbolic element such that and are independent, then is finite.
Proof.
(1) This is [DGO, Lemma 6.5].
(2) If is infinite, then it contains for some . But since and are independent, for a sufficiently large , we have , impossible. ∎
Note that under the assumption of the proposition, the action of is elementary if and only if is virtually , which is equivalent to that in this case.
In general, if a group is virtually , then there exists an exact sequence
where is either or , and is finite. In the case that is in the above, we denote the finite group by . If , is the set of elements of finite order in .
The axis defines two points in the ideal boundary of , which we denote . is exactly the set of elements that leaves this set invariant. If swaps those two points, we say it flips the axis since it flips the direction of the axis .
If the action of on is -uniform WPD, then for any hyperbolic , we have . Moreover, if , then .
Note that . If the action is non-elementary, then , which is equivalent to that is not virtually , so that contains two independent hyperbolic isometries.
We say a hyperbolic element is primitive if and each element is written as for some and .
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 . 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 are given by the action:
The constants will appear in Lemma 2.8 and in the proof of Lemma 4.2.
The second kind are those that depend on a generating set of :
If 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 .
We remark that if diverges, then one way to argue is to rescale by , 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, is bounded. In this paper, we use a unified approach.
For convenience, we assume from now on.
2.5. Lower bound of a growth rate
Lemma 2.7 (hyperbolic element of large displacement).
Let be a -hyperbolic space and a finite set of isometries of . Suppose that . Let be such that . Then there is a hyperbolic element such that
and
This is exactly same as a part of [AL, Lemma 7] (in that paper, our element is denoted as in the proof), although their setting is that is a Cayley graph of a hyperbolic group and .
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 then ; and if then by the definition of and the choice of .
We summarize the properties of we use later:
-
•
-
•
The distance between and the -axis of is at most .
The second one follows from .
Lemma 2.8 (free subgroup with primitive hyperbolic elements).
Let be a -hyperbolic geodesic space and a group acting on . Assume is not virtually cyclic. Let be a function for WPD. Set
Let be a finite set that generates with , and with . Suppose is a hyperbolic element with a -axis with such that
Assume that is -WPD.
Then there exists such that are independent and freely generate a rank- free group with the following property. There is a WPD-function with
such that every non-trivial element is hyperbolic on and -WPD. Also, satisfies:
Moreover, there is an element that satisfies:
-
(1)
has a -axis with .
-
(2)
is primitive, namely, there exists
such that any element is written as with and .
-
(3)
.
-
(4)
.
By , we have , , and .
Remark 2.9.
Proof.
Set and . Since is -WPD, by Lemma 2.2 (1), we have . Remember that by our assumption, we always have , so that the estimate in the lemma holds for as well.
Since is not virtually cyclic, there is an element with . By Lemma 2.5(2), the diameter of the intersection is at most
In view of this, set
Then
Here, we used .
Note that is hyperbolic and -WPD with a -axis . Since is at least 10 times longer than the above intersection, and freely generate a free group, . Also, its non-trivial elements are hyperbolic and
For the last inequality, we used .
We argue that there is a function with such that every non-trivial is uniformly -WPD. We only need to argue for to check WPD by Lemma 2.4. Let be a -axis of . We will show that for any with , there are at most elements satisfying
We denote the collection of those elements by . We will use the fact that the axis of is, roughly speaking, a concatenation of some translates of the segment by elements in (cf. a more precise description of the axis of the element below.) Also, we point out that without loss of generality, one may take a conjugate of in in the argument. Since is a word on and , by taking a conjugate of in , one may assume that the segment is contained in the -neighborhood of the segment except for some small neighborhood of and . Taking a further conjugate of if necessary, one may assume that the segment is contained in the -neighborhood of . Here, we used .
That implies that one can find points with such that and if .
But since is -WPD where , the set contains at most elements. (Strictly speaking the points and are maybe not exactly on a -axis of , but one can choose nearby points of and on the axis and argue.) We showed that is uniformly WPD.

Now we argue for the moreover part. We define an element by
(1) To construct an axis of , consider the following (see Figure 1):
which consists of five geodesic segments, which we call pieces. We name them as . Their length are for some constant maybe with some error up to . Remember that
Note that is contained in , which is the axis of . Put . Similarly, are contained in the axes of the conjugates of by , respectively. We will call those axes .
There are four short () gaps between and . We put (short) geodesics between the gaps and obtain a path, . Then take the union of the -orbit of , which we call . This is an axis for since and are independent and their -axes stay close () to each other along a short (compared to ) segment.
We call the image of a piece by also a piece. The axis is a sequence of pieces (with short gaps in between).
By construction, , but maybe is not exactly a -axis for , so we take a -axis, . One can check that . (This is the reason we put at the end of . Without at the end, is still hyperbolic.) Also, and stay close to each other in the sense that most part of each piece in , except for some short parts near the two ends, is in the -neighborhood of .
(2) If two axes are parallel, we write . Since , . Also, , , and .
We first argue that maps to in the exact sequence after Proposition 2.6. Suppose . By definition, . It suffices to show that does not flip the direction of . For that, we examine the sequence of the lengths of the pieces on and on . Here, we say that is a piece if is a piece in .
The sequence for the part is , so that on , the sequence is (from the left to the right in Figure 2):
Now, if the direction of was opposite to , then the sequence on would be (from the left to the right in the figure):

From those two sequences we observe that one of the followings must hold:
-
(1)
there is a piece on that intersects both and at least in diameter, or
-
(2)
there is a piece on that intersects both and at least in diameter, or
-
(3)
there is a piece on that intersects both and at least in diameter.
In the above, has length (approximately) or .
Let be the axis that contains the piece . Then, since is at least , Lemma 2.5(2) would imply that either (1) , (2) , or (3) , respectively. In either case, we obtain a contradiction since , and . We showed that does not flip .
By the same reason, ie, the constrain from the combinatorics, must be close to for some , namely, the distance is at most . This implies that and also, . It then follows that
which implies . But the right hand side is a finite group, so that has finite order, therefore . And, of course, namely, there is with .
(3). This is a consequence of the -uniform WPD, and is mentioned in the paragraphs following Proposition 2.6.
(4). Since and , we have . We are done since . ∎
Proposition 2.10 (Lower bound of growth).
Suppose acts on a -hyperbolic space . Let be a function for WPD. Assume that there exists a constant such that for any finite generating set of , the set contains a hyperbolic element that is -WPD.
Set
Then for any finite generating set of , we have .
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 . Since the element is -WPD, we have by Lemma 2.2 (1). It implies that . Since , we have .
Lemma 2.7 applies to since . By Lemma 2.7 applied to with such that , there is such that
and the -axis of satisfies .
Then Lemma 2.8 applies to and . By Lemma 2.8 applied to and , there exists such that and contains that is primitive such that
where , and that is -WPD, where .
We note that the element can be chosen from since is not virtually cyclic. This can be easily seen in the proof of Lemma 2.8 since we only need to choose such that . See Remark 2.9.
Now, take a maximal subset such that any two distinct elements are in different -(right) cosets. Then, .
Let be a -axis of
with .
Claim 1.
For distinct ,
are not parallel. Indeed, if they were parallel, then .
Moreover, we will argue , which is a contradiction since are in distinct -cosets.
The reason for is that
since ,
we have
But on the other hand, since is primitive, if , then
The last inequality is from . Those two estimates contradict. We showed the claim.
It implies that for distinct , the intersection of and the -neighborhood of is bounded by
by Lemma2.5 (2) since is -WPD. Remember that . So, this bound is .
Set . Then and (maybe not quite ). is an axis for as well. Set
Claim 2. We have and freely generates a free group of rank .
This is because for any , we have since , which means the axis of , , is close to . To be precise, close means that the distance is much smaller than since . Also, for the axes and of any distinct , the intersection of one with the -neighborhood of the other is times shorter than since . In this setting, the usual ping-pong argument shows the claim.
Since , we have . It follows that for any ,
It implies
Since
which is at least , since . Setting
we have . This is a desired conclusion since . ∎
Example 2.11.
Proposition 2.10 applies to the following examples:
-
(1)
Non-elementary hyperbolic groups (the original case in [AL]).
-
(2)
The mapping class groups of a compact orientable surface with . See Section 5.3, where the assumptions are checked.
-
(3)
A lattice in a simple Lie group of rank-1. See Theorem 5.5.
-
(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 acts on a -hyperbolic space , and is not virtually cyclic. Let be a function for WPD. Assume that there exists a constant such that for any finite generating set of , the set contains a hyperbolic element on that is -WPD. Assume that is equationally Noetherian. Then, is a well-ordered set.
Proof.
We will prove that does not contain a strictly decreasing convergent sequence. To argue by contradiction, suppose that there exists a sequence of finite generating sets , such that is a strictly decreasing sequence and , for some .
By Proposition 2.10, we may assume that the cardinality of the generating sets 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, .
Let . Let be the free group of rank with a free generating set: . For each index , we define a map: , by setting: . Since are generating sets, the map is an epimorphism for every . Note that .
Since is countable, the sequence subconverges to a surjective homomorphism . is called a limit group over the group .
By assumption, is equationally Noetherian. By the general principle (Lemma 1.11), there exists an epimorphism such that by passing to a subsequence we may assume that all the homomorphisms factor through the limit epimorphism: , ie, .
Notice that since for every index , we have .
We will show the following key result:
Proposition 3.2.
Suppose satisfies the assumption in Theorem 3.1. Let be the limit group over of a sequence , where is a free group with a free generating set and are generating sets of . Then .
We postpone proving this proposition until the next section and finish the proof of the theorem.
We assumed that the sequence is strictly decreasing, hence, it can not converge to an upper bound of the sequence, . But the proposition says that it must converge to , a contradiction. Theorem 3.1 is proved. ∎
4. Continuity of the growth rate
We prove Proposition 3.2. We already know that from the existence of the surjections , which follows from that is equationally Noetherian.
It suffices to show that given , for a large enough ,
The strategy of the proof of this is same as [FS]. We note that from now on, we do not use that 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 is almost equal to for a large enough . First of all, if we take a large enough , then contains elements roughly as many as by the definition. Fix such . Then if we take large enough, and are idential via the map since is a limit group. But it does not mean that is almost equal to since the growth of the balls in may decay if we take the radius larger. But it turns out that if we take large enough, then roughly speaking, the growth of the ball of radius in is almost equal to . 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 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 ,
This is because the sequence is sub-additive.
Fix (we will choose sufficiently large in the argument we will give later). Then choose large enough such that is injective. The following map is naturally induced from for each :
by mapping to . If there is an such that this map is injective for all , then an easy computation would show the desired inequality for the by letting .
But of course this map is not injective in general. For example, the concatenation of may have lots of backtracks at the concatenation points. As a remedy, we insert elements of bounded length, called separators, and define a new map sending to . This map is denoted by . The separators are constructed in Lemma 4.2. We arrange that the concatenation of elements, after we insert ’s, is a uniform quasi-geodesic in . This follows from that is a hyperbolic group.
But it is still not the case that the map is injective, where is the bound of the length of the separators. What we actually show is that is injective if we restrict it to the -tuples in some fixed portion of , (see Lemma 4.8), which is enough for our purpose. This part is very technical. To argue that is injective on the certain fixed portion, we use the action of on , ie, we map a tuple by to , then let it act on . Then we analyze the orbit of a base point in .
4.1. Separators
We review the setting. is a -hyperbolic space, and acts on it. We assume . For each , is a finite generating set of such that contains a hyperbolic element that is -WPD. Using the homomorphism , we let act on . We first construct separators as elements in then pull them back to by .
Fix . Let be a point where is achieved. Put
We define a germ w.r.t. the constant . Recall that given three points , the Gromov product, , is defined as follows:
Definition 4.1 (germs, equivalent and opposite germs).
Let be a (directed) geodesic segment in . Suppose that . Then, the initial segment of of length is called the germ of at , denoted by . If , then we define the germ to be empty.
We say two non-empty germs, , at a common point , are equivalent if , and opposite if .
We sometimes call the germ of at as the germ of at . If is the germ of at , then for , the segment is the germ of at . For , we consider the germ of and call it the germ of at , and write .
Recall from Lemma 2.8 that .
We consider germs w.r.t. .
Lemma 4.2 (The constant and separators. c.f. Lemma 2.4 [FS]).
There exists a constant with the following property, where depends only on and . For every , there exist primitive, hyperbolic elements ; and mutually opposite germs of some elements in at that satisfy:
-
(i)
The germs are all at in the following.
-
•
The germ of is equivalent to . The germ of is equivalent to .
-
•
The germ of is equivalent to . The germ of is equivalent to .
-
•
The germ of is equivalent to . The germ of is equivalent to .
-
•
The germ of is equivalent to . The germ of is equivalent to .
-
•
-
(ii)
For all ,
and the distance from to the -axis of is at most .
-
(iii)
For every , and all (possibly ), if the -neighborhood of intersects the segment , then the diameter of the intersection is bounded by: and . If , we assume in addition that , where is the finite normal subgroup in .
Remark 4.3.
(1)
Regarding Lemma 4.2 (iii), if then
the Hausdorff-distance between
and is
at most
since moves any point on a -axis
of by at most since otherwise
would be hyperbolic.
(2) As we will see in the proof it suffices to take to be
(2) |
The elements are called separators and the property (iii) is called the small cancellation property of separators.
Note that we have .
Proof.
Since and is a point where is achieved, by Lemma 2.7 (also see the properties at the bullets after the lemma), there is a hyperbolic element such that
and the -axis of , , is at distance at most from .
By Lemma 2.8 applied to and and the -axis , there exists (see the remark 2.9) such that and are independent hyperbolic elements, which freely generate a free group whose non-trivial element, , satisfies
Recall that . Note that . The distance from to is at most . Also, the intersection of and the -neighborhood of is at most in length, which is , by Lemma 2.5 since and are not parallel.
Consider the following four germs at w.r.t. the constant :
Note that any two of them are mutually opposite.
Then, there exist separators such that , where
For example, set and:
We compute that they are in since . It will be important that this number does not depend on . (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 is primitive hyperbolic, using and are not parallel, in the proof of Lemma 2.8(2). We omit it.
(i) By definition of ,
Here, we mean that the four germs in each line are same or equivalent to each other.
(ii). The estimate for is straightforward from the definitions of . The claim on the axes are also shown similarly to the case of 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 , so that distance becomes larger.
(iii) follows from the definition of , namely, the combinatorial structure of the words. The argument is similar to show that in Lemma 2.8 is primitive (Lemma 2.8(2)), so we will be brief. Suppose , and the intersection was longer. Then, because of the combinatorial structure of the words and , it follows the axes and would be parallel, by Lemma 2.5, which is impossible. We are done. Suppose , and the intersection was longer. Then since , by the same reason, and would be parallel, a contradiction. So we are done in this case too. ∎
Remember depend on the index of , so let’s write them as . Now, since is surjective, let be an element with . Note that the word length of in terms of is also bounded by . The elements are also called the separators for .
In the following we may just write (instead of ) to denote a separator for to simplify the notation. We note that for any separator for any since is primitive (see the comment after Proposition 2.6).
4.2. Forbidden elements
Given , choose and fix large enough such that the map is injective on . This is possible since this set is finite in the limit group . Remember
Given we choose one of the separators, , for s.t. the germ of at and the germ of at are opposite. (Geometrically, it implies that the concatenation is almost a geodesic.) Such a separator exists by the property (i) in Lemma 4.2. Here, if , then we choose any separator . We say is admissible for . Note that
Also, the Hausdorff distance between and is at most .
Moreover, given then we can choose a separator such that is admissible for and is admissible for . We say is admissible for . Note that
As before, the Hausdorff distance between and is at most .
For , we define a map
by sending to , where are separators we choose that are admissible for for . Remember that .
It is easy to show that maps an element to a non-trivial element in 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”, is injective by showing the image of the base point by those elements are all distinct.
For the given , we will define forbidden elements in , which depend on . We are assuming that is large enough so that is injective on .
Definition 4.4 (forbidden elements and tails).
Given , if there exist and a separator which is admissible for such that
(3) |
then we say is forbidden w.r.t. (or in terms of . We call the segment a tail of . The tail depends on the choice of , so if we want to specify it, we say the tail of the pair .
Let be two (forbidden) elements and admissible separators , respectively. If and , then the Hausdorff-distance between the two tails and is at most . This is an immediate consequence of Remark 4.3. In this case, we say that the tails of and are parallel.
On the other hand, if two tails and are not parallel, then the intersection of one of the tails with the -neighborhood of the other tail is bounded by the 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 is injective on . Suppose is forbidden w.r.t. , and is a separator admissible for . Then there are at most possibilities for , including , such that is forbidden, is admissible for , and the tails for and are parallel.
Proof.
Since the two tails are parallel, we have by the definition that two tails are parallel. Recall that (see the paragraphs after Proposition 2.6.)
for all separators since is primitive. Therefore, we have at most possibilities for . But since is injective on , we have at most possibilities for , and we are done. ∎
4.3. Ratio of forbidden elements
The proof of the following lemma occupies this subsection.
Lemma 4.6.
Assume that is injective on . Consider the forbidden/non-forbidden elements in w.r.t. . Then
We denote as . The strategy of the proof is to show that if is forbidden, it will force some other elements in to be non-forbidden.
Proof.
The proof has three parts.
Part I. We first construct a subset in and a tree like graph in .
The construction is inductive. In each step, a subset and a tree like graph grows, which will end in finite steps.
Suppose is forbidden. We explain the inductive steps
to construct and .
Step 0. Set and .
Step 1.
Since is forbidden, there exist
and a separator that is admissible for that satisfy (3).
Let be a shortest representative
w.r.t. the word metric by .
Then this defines a sequence of points in , which we call
a path, , from to
as follows:
The distance between any two adjacent points on is at most since it is achieved at and .
Consider the nearest point projection, denoted by , from a point to the tail . The nearest points may not be unique, but we choose one for . Then the distance between the projection of any two adjacent points on is . Note that since is admissible for . Also, by (3).
By Lemma 4.2 (ii), the length of a tail is at least . Let be the two points on the tail that trisect the tail into three pieces of equal length, where is closer to than is. Each of the three pieces has lengh at least . Let be a point on whose projection is closest to , and a point whose projection is closest to . Then
We denote as and call them the candidates for non-forbidden elements obtained from . They depend on too. See Figure 3.

Consider the union of the three segements as follows:
This depends not only , but also . is a tree embedded in . We call and the branch points of , and the trisecting points of (the tail of) . This union is a disjoint union except for the two branch points. The distance betwee the branch points is at least . This in particular implies that .
We set , and . If both are non-forbidden, this is the end of the construction, and put . Otherwise we go to the second step.
is a tree like graph in the sense that it is the union of two trees, and attached at the point , and the intersection is contained in the -neighborhood of this point since is admissible for .
Step 2. By assumption, at least one of and is forbidden. For example, assume that is forbidden. Then has a separator admissible for , and an element that satisfy (3). Let denote the nearest point projection to the tail .
Then as in the first step, there is a path (a sequence of points obtained from a shortest expression of in ) between and in , from which we obtain two elements, , using the projection from the path to the tail . Also, we obtain a tree embedded in :
We point out that is contained in the -neighborhood of . The reason is because is admissible to , and is admissible to , while the tails of and are both at least long.
Also, the trisecting points and the branch points on the tail are out of the -neighborhood of the tail .
It follows that are distinct elements in . If is not forbidden, then we put and
If is forbidden, then we do the same construction as we did to and obtain two candidates: , and a tree which intersects the previously obtained treelike graph only in the -neighborhood of . In particular, (if they exist) are all distinct. See Figure 4

We set and
If all of the elements and also (if they exist) are non-forbidden, we stop here and set , otherwise we go to the third step.
In the third step, we do the same construction to the forbidden elements among For example, if is forbidden, then we obtain two elements and a tree . This tree intersects the tree like graph in the -neighborhood of . We define from by adding a pair of elements that are obtained from each forbidden element in this step. Also, we define be adding the trees for forbidden elements .
We formally describe the -th step. In the -step, we have . If all of the elements in are non-forbidden, then we stop and put . Otherwise, we go to the -th step.
-th step. By assumption, at least one element in , , is forbidden. Then we do the same construction as in the step 1, and obtain the tree ; two elements, and . We add those two elements to , and also add to . We do this all forbidden elements in , and get and . Note that the trees we added in this step are disjoint from each other, moreover, the distance between two with is at least , where are multi-subscripts of length .
Since for all , and the set gets bigger at least by two as increases by one, so the process must end in finite steps since is finite. At the end of the process, we have a set in , and a tree like graph . We remark that depend on the choice of separators in the construction, but it is not important.
No that by construction, in , we have
Also, the orbit of by the elements in appears as vertices of .
Part II. Let be two forbidden elements, and we analyze how the sets and are related. In the construction of , for each forbidden element, , we chose a separator and an element . The pair defines the tail , which is an arc of .
Claim. Let be forbidden elements and the separator we chose for to construct . Assume that the tail is not parallel to any of that appears to construct . Then, is empty or .
Let be the trisecting points of . Then neither of them is in the -neighborhood of any trisecting point that appears in . Indeed, suppose it was, and let be a trisecting point of the tail, , in such that or is . Then, since have as the common “root” vertex that they start from, the intersection of and the -neighborhood of is at least of , which contradicts the small cancellation property.
This implies that not only the branch points on , but also all branch points of are outside of the -neighborhood of . This is because both are tree like graphs. It then follows that is empty or just because all other points in appear on after the first two branch points on , that are close to . We showed the claim.
Part III. It follows from the claim that there is a finite collection of forbidden elements in such that the ’s are mutually disjoint, and that any forbidden element is either contained in the union of those ’s, or has an admissible separator such that the pair (tail) is parallel to one of the tails that appears in one of the ’s.
To see that, order the forbidden elements in as . Choose admissible separators . First, construct using the separators we chose. Next, if is already contained in , then disregard it. Also, if is parallel to one of the tails in , then disregard also. Otherwise, construct and keep it. If is contained in or ; or is parallel to one of the tails of or , then disregard . Otherwise, construct and keep it, and so on. By the claim we have shown, those ’s are mutually disjoint such that for any forbidden element , either is contained in one of the ’s we have, or is parallel to one of the tails of the ’s.
4.4. Feasible elements
By the lemma 4.6, at least of consists of non-forbidden elements.
Now, we choose a maximal subset in the set of non-forbidden elements in such that for any two distinct elements in the set, are not in the same coset w.r.t. for any separator . We call an element in this set an adequate element. Then, in ,
This is because there are only four separators, , and for each . Combining this with Lemma 4.6 we get:
(4) |
Definition 4.7 (feasible elements).
Fix , then . An element of the form with are called feasible of type if all are adequate and each is admissible for . We define for the empty element.
For this feasible elements, we consider the following broken geodesic:
The Hausdorff-distance between and the geodesic is at most .
Lemma 4.8 (cf. [FS],Lemma 2.6).
For every , the map is injective on the set of -tuples of adequate elements in .
By definition, given a -tuples of adequate elements, we choose admissible separators (they are not unique) and form a feasible element of type , then map it by .
Proof.
We argue by induction on the type . If then nothing to prove since the element is empty.
Assume the conclusion holds for . Suppose -tuples of adequate elements and are given. Let and be two feasible elements of type . We assume that and want to show for all .
The elements define two broken geodesics between and . Look at the initial parts of : and . We first show that . Suppose not. Then it would imply that either or is forbidden. We explain the reason. First, by the small cancellation property of the separators (Lemma 4.2 (iii)), either is contained in the -neighborhood of , or is contained in the -neighborhood of , since is admissible for and is admissible for . Suppose we are in the first case. Then would be forbidden since there is a subword of , denoted by , such that
Such a subword exists because and . We got a contradiction since is not forbidden. Similarly, if we are in the second case, then would be forbidden, which is a contradiction. We showed that .
Next we show . We first show that . Suppose not. Then, as in the previous paragraph, the small cancellation property of the separators implies that either or is forbidden, a contradiction. We are left with the case that . But, this means that and are in the same (right) coset w.r.t. . Since both and are adequate, it implies .
Since , it follows that and are feasible elements of type with . By the induction hypothesis, we have for all , and we are done. ∎
4.5. Proof of Proposition 3.2
We prove Proposition 3.2.
Proof.
Recall that is the WPD function and we set . Recall that for every , we have
Given , choose and fix a large enough such that
The constant is from Lemma 4.2. Such exists since , and does not depend on . (We will choose right after this.)
Choose large enough such that is injective on , which defines the forbidden elements in . Then for all :
because the number of adequate elements in is at least by the estimate (4), and is injective on the set of feasible elements of type by Lemma 4.8.
Then by the above three inequalities,
Since we have this for all large enough , and is arbitrary, we have
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 be a constant, an integer and a function for WPD. Suppose we have a family of finitely generated groups such that each is not virtually cyclic and acts on a -hyperbolic space . Assume that for any finite generating set of any , the set contains a hyperbolic element on that is -WPD.
Now suppose that is a finite generating set of that is in the family for . Assume that the (infinite) sequence is bounded from above. Then by Proposition 2.10, there is a constant that depends only on and the function such that , therefore, the sequence is bounded from above.
As before, passing to a subsequence, we may assume that there exists with for all . This gives a sequence of surjections with , where is a pair of free group and a free generating set with .
Passing to a further subsequence if necessary, we may assume that the sequence converges to a limit epimorphism .
Then we have:
Proposition 4.9.
Assume that there exists an epimorphism for all such that all the homomorphisms factor through the limit epimorphism: , ie, . Then,
Before we explain the proof, we point out that in Proposition 3.2, the assumption on the existence of is a consequence of that 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 , we immediately have . The main issue was to show the other inequality by constructing separators for each action of on , via . That argument applies without change to the current setting since the constants and the function are common for all and .
5. Examples
An obvious example for Theorem 1.1 is a non-elementary hyperbolic group, . Let be a Cayley graph of , then it is -hyperbolic, and the action by is (uniformly) proper, so that acylindrical, and non-elementary. The existence of the constant is known ([K]). As we said is equationally Noetherian ([S], [RW]). Therefore, 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 is hyperbolic relative to a collection of finitely generated subgroups . Suppose is not virtually cyclic and not equal to any . Then it acts properly discontinuously on a proper -hyperbolic space such that [Bo1, Proposition 6.13]:
-
(1)
There is a -invariant collection of points, , with finite. For each , there is a point such that the stabilizer of is . The union of the -orbits of the ’s is .
-
(2)
For every , there is a -invariant collection of horoballs at each such that they are -separated, ie, for every distinct .
-
(3)
The action of on is co-compact.
A subgroup is called parabolic if it is infinite, fixes a point in , and contains no hyperbolic elements. The fixed point is unique and called a parabolic point. In fact, 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 and be as above. Then there exists such that for any finite generating set of , the set contains a hyperbolic element on .
Also we have the following:
Lemma 5.2.
Let and be as above. Then the action is uniformly WPD.
Proof.
Given , take a -invariant collection of horoballs that are -separated in . Let denote the union of the interior of the horoballs in the collection. Then since the action of on is co-compact, there exists a constant such that for any , the cardinality of the following set is at most :
We argue that the action is uniformly WPD w.r.t. . Let be hyperbolic with an -axis . Let with . It suffices to show that the following set contains at most elements:
We divide the case into four:
(1)
.
Then there are at most elements
s.t. , and we are done.
(2)
.
This is same as (1).
(3) such that
and with .
Then, there must be with
since the horoballs in
are -separated.
Then .
But there are at most such elements.
(4) for some .
Then for some ,
since is hyperbolic and does not preserve
any horoball. Since horoballs are -separated, we have .
So, .
Now, there are two possibilities:
one is that is between and . Then, as in (3), there must be with ,
and we are done.
The other possibility is that is between
and on .
But in this case, since ,
we have .
Then the distance between and
is at most since
and . 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 is hyperbolic relative to equationally Noetherian subgroups, then is equationally Noetherian.
We are ready to state a theorem.
Theorem 5.4 (Well-orderedness for relatively hyperbolic groups).
Let be a group that is hyperbolic relative to a collection of subgroups . Suppose is not virtually cyclic, and not equal to for any . Suppose each is finitely generated and equationally Noetherian. Then is well-ordered.
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 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 is well-ordered. We prove the following as an immediate application of Theorem 5.4.
Theorem 5.5 (Rank-1 lattices).
Let be one of the following groups:
-
(1)
A lattice in a simple Lie group of rank-1.
-
(2)
The fundamental group of a complete Riemannian manifold of finite volume such that there exist with , where denotes the sectional curvature.
Then is well-ordered.
Proof.
We only need to argue for non-uniform lattices since otherwise, is a non-elementary hyperbolic group and the conclusion holds. Suppose that is a non-uniform lattice. Then, it is known that is relatively hyperbolic w.r.t. the the parabolic subgroups, called peripheral subgroups that are associated to the cusps, [Fa], [Bo1]. Moreover is not virtually cyclic, and not equal to any . Also, those 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 are equationally Noetherian. With those facts, Theorem 5.4 applies to and the conclusion holds. ∎
We remark that it is known that if 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 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 be the mapping class group of a compact oriented surface with genus , punctures , and complexity . It is known that it is either virtually abelian or has exponential growth, and then uniform exponential growth ([Ma2]). Let be the curve graph of . The graph has vertex set representing the non-trivial homotopy classes of simple closed curves on , and edges joining vertices representing the homotopy classes of disjoint curves. The group naturally acts on it by isometries.
We assume . Then has exponential growth. We recall some facts:
-
(1)
The graph is -hyperbolic, and any pseudo-Anosov element in acts hyperbolically on , [MM1].
-
(2)
The action of on is acylindrical, [Bo]. It is non-elementary.
-
(3)
There exists such that for any pseudo-Anosov element , for the action on , [MM1]. (This follows from the acylindricity as we pointed out.)
-
(4)
There exists such that for any finite such that contains a pseudo-Anosov element, then contains a pseudo-Anosov element, [Ma].
In summary, the action of on is known to satisfy all the assumptions of Theorem 1.1 except we do not know if 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 is well-ordered if .
We remark that for , the conclusion holds by [FS] since is hyperbolic (a well-known fact, for example, [MM1]).
Question 5.6.
Let . Is 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 .
5.4. Three manifold groups
We discuss three manifold groups. Let be a closed, orientable -manifold. is called irreducible if does not admit a non-trivial splitting over the trivial group.
Let 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 in such that each connected component of 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 . 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 has non-trivial JSJ-decomposition. A non-trivial JSJ-decomposition gives a graph of groups decomposition of along subgroups isomorphic to , and has exponential growth. Its Bass-Serre tree is called the JSJ-tree, . An action of a group on a tree is called -acylindrical if for every non-trivial element , the subtree of fixed points by is either empty or of diameter at most . It is proved that the action of on is -acylindrical, [WZ, Proposition 4.2]. It implies that the action is uniformly -WPD. It is known that if an action of on a tree is -acylindrical for some , then it is acylindrical [MO, Lemma 5.2], therefore it is uniformly WPD (Lemma 2.3). Moreover, the acylindricity constants and the uniformly WPD constant depend only on and . The moreover part easily follows from the proof of Lemma 5.2in [MO].
Theorem 5.7.
Let be a closed orientable -manifold, and . If is one of the following, then has exponential growth and is well-ordered.
-
(1)
is not irreducible such that is not isomorphic to .
-
(2)
is irreducible such that it is not a torus bundle over a circle, and that it has a non-trivial JSJ-decomposition.
-
(3)
admits hyperbolic geometry.
-
(4)
is Seifert fibered such that the base -orbifold is hyperbolic.
Proof.
First, every three manifold group is equationally Noetherian, [GHL].
(1) In this case, is a non-trivial free product . Since it is not , it has exponential growth. acts on the Bass-Seree tree of this free product. Then for any finite generating set , the set contains a hyperbolic element on , [Se]. The action of on is -acylindrical, so that the action is uniformly -WPD for some . Since is hyperbolic, Theorem 3.1 applies with .
(2) In this case, let be the JSJ-tree of . Then the action of on is -acylindrical, so that it is uniformly -WPD for some . Theorem 3.1 applies with .
(3) If is hyperbolic, then is a non-elementary, hyperbolic group. Then is well-ordered by [FS].
(4) In this case, we have the following exact sequence:
where is the orbifold-fundamental group of the base 2-orbifold, and is the fundamental group of the fiber circle. We denote this subgroup . By assumption, is a non-elementary, hyperbolic group.
We claim that . To see it, let be a finite generating set of . Let be the image of by the projection . We have . But it is known that the subgroup is not distorted in in the sense that there is a constant such that for any , we have , [NS, Proposition 1.2 (2)]. It implies that for all , we have . It follows that , therefore . On the other hand, if is a finite generating set of , then there is a finite generating set of which projects to , by lifting each element of to , then adding a generator of , which gives . Then as we saw, , which implies . We proved the claim.
But by [FS], is well-ordered, therefore so is . ∎
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 -manifolds with fundamental groups of exponential growth except that has Sol-geometry. We leave it as a question.
Question 5.8.
Let a closed, orientable -manifold that has the geometry of three dimensional solvable group, Sol. Then is well-ordered ? Also, in view of Proposition 2.10, is it true that if then ?
6. The set of growth of subgroups
We discuss the set of growth of subgroups in a finitely generated group . Define
The set is countable and contains as a subset. If is a hyperbolic group, it is known by [FS, Section 5] that is well-ordered.
6.1. Subgroups with hyperbolic elements
As usual, suppose acts on a -hyperbolic space . We introduce a subset in as follows:
where in addition we only consider such that contains a hyperbolic element on . This set depends on the action on .
Theorem 6.1 (cf. Theorem 3.1).
Suppose acts on a -hyperbolic space , and is not virtually cyclic. Let be a function for WPD. Assume that there exists a constant such that if contains a hyperbolic element on for a finite subset , then contains a hyperbolic element that is -WPD. Assume that is equationally Noetherian. Then, is a well-ordered set.
Proof.
The proof is nearly identical to Theorem 3.1.
Let be a sequence of finite generating sets of subgroups of exponential growth in , , such that each contains a hyperbolic element on and that is a strictly decreasing sequence with , for some .
By our assumption, Proposition 2.10 applies to . Therefore, is uniformly bounded from above. By passing to a subsequence we may assume that is constant, , for the entire sequence.
Let . Let be the free group of rank with a free generating set: . For each index , we define a map: , by setting: . By construction: .
Then, as before, the sequence subconverges to a surjective homomorphism , where is a limit group over .
By assumption, is equationally Noetherian. By the general principle (Lemma 1.11), there exists an epimorphism such that by passing to a subsequence we may assume that all the homomorphisms factor through the limit epimorphism: , ie, .
Then, we have
Proposition 6.2 (cf. Proposition 3.2).
Remark 6.3.
This proposition is a special case of Proposition 4.9, when all -hyperbolic spaces are .
6.2. Relatively hyperbolic groups
We apply Theorem 6.1 relatively hyperbolic groups. Let be a relatively hyperbolic group with finitely generated. Define
where in addition we only consider that is not conjugate into any . This is a subset in .
We first characterize the subgroups that appear in the definition in terms of the action on a -hyperbolic space that we described in Section 5.1. Fix such and an action by .
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 be a subgroup. Then the following two are equivalent:
-
(1)
has an element that is hyperbolic on , and is not virtually .
-
(2)
is infinite, not virtually and not conjugate into any .
The lemma implies:
We prove:
Theorem 6.5.
Let be a group that is hyperbolic relative to a collection of subgroups . Suppose is not virtually cyclic, and not equal to for any . Suppose each is finitely generated and equationally Noetherian. Then is well-ordered.
Proof.
It suffices to argue that is well-ordered. For that, we apply Theorem 6.1 to the action of on . 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 such that for any finite set such that contains a hyperbolic isometry on , then 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 be a group in Theorem 5.5. Then is a well-ordered set.
Proof.
As we said in the proof of Theorem 5.5, is relatively hyperbolic w.r.t. the parabolic subgroups , which are associated to the cusps, and are virtually nilpotent. By applying Theorem 6.5 to , we have that is well-ordered. But if a subgroup is conjugate into one of , then it is virtually nilpotent, so that has polynomial growth. It follows that holds for , so that is well-ordered. ∎
6.3. Subgroups in mapping class groups
We discuss mapping class groups. A subgroup is called large if it contains two independent pseudo-Anosov elements. Such has exponential growth.
We define:
where in addition is large. Note that .
Theorem 6.7.
Assume that is equationally Noetherian. Then is well-ordered.
Proof.
We suppress and denote . Let the curve graph of . Then, as we said, Theorem 6.1 applies to the action of on . We conclude that is well-ordered. Now, we claim . Indeed, given , if contains a hyperbolic element on , then it is a pseudo-Anosov element, and moreover, from , must be large. We showed . On the other hand, for if is large in , then contains hyperbolic isometries on , so that . ∎
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 well-ordered ?
7. Finiteness
7.1. Finiteness of equal growth generating sets
If is a hyperbolic group, it is known by [FS, Section 3] that for , there are only finitely many generating sets of , up to , such that . We discuss this issue.
Theorem 7.1 (Finiteness. cf. Theorem 3.1 in [FS]).
Suppose a finitely generated group acts on a -hyperbolic space and is not virtually cyclic. Let be a WPD-function. Assume that there exists a constant such that if is a finite generating set of , then contains a hyperbolic element that is -WPD. Assume that is equationally Noetherian.
Then for any , up to the action on , there are at most finitely many finite generating set such that .
Proof.
We argue by contradiction. Suppose that there are infinitely many finite sets of generators that satisfy: , and no pair of generating sets is equivalent under the action of the automorphism group . As in the proof of Theorem 3.1, by Proposition 2.10, the cardinality of the generating sets is bounded, so we may pass to a subsequence that have a fixed cardinality . Hence, each generating set corresponds to an epimorphism, , where is a fixed free generating set of , and .
By passing to a further subsequence, we may assume that the sequence of epimorphisms converges to a limit group with the associated quotient map. As in the proof of Theorem 3.1, since is equationally Noetherian, by Lemma 1.11, for large , , where is an epimorphism. In particular, . We pass to a further subsequence such that for every , .
Since for every index , is an epimorphism from onto that maps to , . We prove:
Proposition 7.2 (cf. Proposition 3.2 [FS]).
If is infinite for some , then .
We postpone the proof of the proposition until the next section, and proceed. We prove a lemma.
Lemma 7.3.
The group contains a finite normal subgroup that contains all finite normal subgroups in , such that .
We recall one fact we use in the proof. If a finitely generated group acts on a -hyperbolic space such that is not virtually cyclic and contains a hyperbolic element on that is -WPD, then contains a maximal finite normal subgroup . Moreover, . We sometimes denote by .
The existence of such is known for an acylindrically hyperbolic group [DGO, Theorem 6.14], and the same proof applies to our setting, which we briefly recall. Indeed, is the intersection of for all hyperbolic elements on . It is obvious that is normal. By assumption, there must be a hyperbolic and WPD element, . Also, there is another element such that and are independent. Then, by Proposition 2.6, is finite, so that is finite. On the other hand, if is a finite normal subgroup in , then for every hyperbolic element , there is such that is contained in the centralizer of , so that . It implies that . We showed that is maximal.
Lastly, to see , consider the exact sequence for the hyperbolic and -WPD element . Recall that if is cyclic. From this we have .
We prove the lemma.
Proof.
Let be a finite normal subgroup. Since is surjective, is a finite normal subgroup, therefore for any . Also, for sufficiently large , the surjection is injective on . But since , we have .
If are two finite normal subgroups, then is a finite normal subgroup. Combined with the fact in the previous paragraph, there must be the maximal finite normal subgroup in with . ∎
By Proposition 3.2, . By our assumption, for every index , . Hence, , so that for every , .
It follows from Proposition 7.2 that for every , is finite. Since is a normal subgroup in , by Lemma 7.3, . Since is a finite group, there are only finitely many possibilities for . It follows that there must be such that for infinitely many .
The map induces an isomorphism from to . Notice that this gives an isomorphism from to since gives a bijection between and . (Here, we may assume that each consists of distinct elements, so that no two elements in are identifies by .) But this implies that is isomorphic to for infinitely many by , ie, those 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 on a limit object, called a limit tree, while in our paper we use the actions of on induced from the maps . 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 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 through the argument.
Strategy of the proof. We start with an informal description of the idea. The constant is given as the assumption, which gives the homomorphism with an infinite kernel. To show that , we will produce not only infinitely many (which is obvious by the assumption), but “exponentially many” elements in the preimage of by for each . Those elements are given by the map . They are exponentially many in terms of the word length of w.r.t. . See the estimate (6).
In the proof two constants and appear. They will be chosen and fixed around the end of the proof. The constant is first chosen. It will be used to measure the gap between and . The constant depends on , so that is injective on . Also, a positive integer is used to make the word length of , which is longer and take a limit at the end of the proof.
We now explain more concretely. Similar to , the ball of radius in centered at the identity, let denote the sphere of radius . It is an elementary fact that if , then
Given , for a large enough depending on , we will define a “map” :
for all , where is a constant that does not depend on . The map is similar to the map in the section 4.2, but strictly speaking is not a map, but is a finite set of elements in for each . But we abuse the notation and call them maps in the following account.
It satisfies for every to have the following two properties. Set .
-
(i).
see the estimate (5).
-
(ii).
which implies that for , we have .
Once we have such a map , we argue as follows: fix a (large) . Since , we have from (i) and (ii) that
Taking , dividing by , and letting , we have as the limsup
Since does not depend on , choosing large enough, this inequality implies
Roughly speaking, the construction of is as follows. As in the construction of , we first construct separators. To define separators in , we use a non-trivial element . Separators will be products of conjugates of , so that they are also in , which will imply the property (ii) in the above. The separators depend on .
For each , the map gives a canonical bijection between and . This gives a bijection between the words on (not elements in ) and the words on .
Let be integers. We will fix and let later. Given an element , we choose a word of length on that represents . We divide into subwords of length . As we said, each subword of length canonically gives a word of length on via the map . We further subdivide each of the subwords of length on into two words of length and . We choose to satisfy . In this way, for each choice of a -tuple of such ’s, we divided the word on corresponding to into subwords. There are ways to subdivide it.
To each of such subdivision, we insert separators to the break points and obtain an element in since the word length of each separator is at most . We obtain such elements. Since separators are in , those elements (words) are mapped to by .
But we do not know if they are all distinct as elements in , but we will show that there are at least elements that are distinct. They are called feasible elements. This collection of feasible elements is denoted by . They are significantly many, so that is strictly larger than as we computed in the above.
Lastly, to show that those feasible elements in defined for each are distinct, as in Section 4, we let them act on the space via the map for a large enough (We choose such that is injective on .) We then argue that the images of the base point by those elements are distinct. In the paper [FS], they use a limit tree on which acts to argue that feasible elements are distinct. Here, we use the action on . This difference already appeared in Section 4.
7.3. Proof
We prove Proposition 7.2.
Proof.
As in Section 4.1, set , and for each let be a point where is achieved and put
We consider germs w.r.t. .
As we did in Lemma 4.2, in the next lemma we construct separators , which give by pulling them back by . In addition to the properties in Lemma 4.2 they satisfy in . We state it as a lemma. The constant is different from the constant in Lemma 4.2, and it depends on but not on .
Lemma 7.4 (cf Lemma 3.3 and Lemma 3.9 in [FS]).
Suppose is infinite. Then there exists a constant with the following property: if is sufficiently large, then there are elements that satisfy the conditions (i), (ii) and (iii) in Lemma 4.2, and in addition to that, all satisfy
-
(iv)
.
Moreover, those elements are such that there are elements with and for each .
We remark that the moreover part is rather than an additional property, but immediate from the construction, ie, we first construct then map them by . We will call both and separators.
Proof.
In the proof there will be several constants , for which we do not try to give explicit values. The important property of those constants is that they do not depend on .
First, since is infinite, choose distinct elements that are in the kernel. Let be the maximum of the word lengths of the in terms of . If is large enough, then the image of those elements by are all distinct. From now on, we only consider such . In the following we fix each such and argue. We have distinct elements and the word lengths of those w.r.t. are bounded by .
Secondly, choose an element that is hyperbolic on such that its -axis is at distance at most from the point . By assumption, such an element exists. Also there is such that and are independent. Choose such .
Thirdly, we choose one element, , from the ’s as follows: if there is such that
(I) The element is hyperbolic on ,
then choose one of such and set . Otherwise choose with
(II)
and set .
This is clearly possible since .
Note that the element depends on . From now on we suppress and write as .
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 with , then we have (see Lemma 2.8)
Note that we have
This is because since and , we have
The last inequality is by and
Recall that the axes of are at at most from .
In the proof of Lemma 4.2, we set and produce as words on , but this time we take into account the germs of the element and choose as follows.
Notice that 6 elements define six germs at that are mutually opposite since as we noted. From the six, choose four distinct germs that are opposite to the germs for and . If those two germs are empty, then ignore this condition. Denote those four germs as . Now choose such that the germ for is equivalent to , respectively, such that the axes of are at at most from .
We also arrange that the axes of are not parallel to each other, and no element of flips the axes of or . 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 that depends only on and such that the word lengths of w.r.t. are bounded by . The constant does not depend on , nor the choice of .
Now, choose with , whose word length w.r.t. are also bounded by . Using we define as follows:
The word length of in terms of is at most , which does not depend on . Indeed, we may set .
Clearly, are in since they are products of conjugates of . Finally, define . 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 is the germ of , the germ is the germ of , the germ is the germ of , and the germ is the germ of . Then satisfy (i). The property (iii) is a consequence of that the axes of and 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 are defined or empty. In the empty case, we use the property that the axes of are not flipped by any element of .
In conclusion, those are desired elements, so in this case we take and we are done.
Suppose we are in the case (II) when we chose . In this case, we replace with another element such that is hyperbolic on . We explain how we produce such an . First choose an element with , where the word length of is at most in terms of . Then we consider an element of the form:
We will show that if , is hyperbolic. See Lemma 7.6. Also, since , we have . 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 in terms of is at most , where , which does not depend on .
With the element we repeat the same argument as we did for in the case (I) and obtain desired with a bound on the word length, uniformly over all , which will finish the proof of Lemma 7.4.
We now explain some details on how to produce from . For an element , define a set
This is a -invariant set. We state a few standard facts on this set.
-
(M1)
If is not hyperbolic then is not empty, since if it is empty, then , which implies is hyperbolic, a contradiction.
-
(M2)
If is not hyperbolic, then for any point , we have
We prove (M2). If , then nothing to show so suppose . Let be a point with .
Let be the point with . We claim that and . To prove the first claim, suppose not, ie, . Let be with and . Then we have since and . Then we have since is not hyperbolic. (Otherwise, is “pushed” along the geodesic by at least by , which implies that is hyperbolic, impossible.) But it implies that , which is a contradiction. We showed .
By the same argument, we can show .
Having those two claims, we have
We showed (M2).
We go back to the explanation. For a hyperbolic isometry and its -axis , we consider the nearest points projection in to . We denote the projection by . For every point , although is not a point, the diameter of is bounded by . The following lemma is well-known. See Figure 5.
Lemma 7.5 (Bounded projection).
If is hyperbolic and -WPD; and and is not hyperbolic, then the image of in by the projection is bounded by in diameter.
We prove the lemma for readers’ convenience.
Proof.
Let and suppose . Assume that is larger than , since otherwise there is nothing to show. Then is a uniform quasi-geodesic and every point on it is moved by by at most . Moreover, each point on with is moved by by at most since is not hyperbolic.
Now, consider the points with . Then since is -WPD, we have by Lemma 2.5 (2). This is because, otherwise, , impossible. It follows that . ∎

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 as in the lemma 7.5, the element is hyperbolic if .
We also give a brief proof.
Proof.
Set . (Or one can set as usual. It does not matter since .) In general, is not exactly -invariant, but if by definition. We also know that by Lemma 2.2 (1) for our . So, if necessary, by replacing by , we may assume that is -invariant. For simplicity, in the following argument, we assume that is -invariant.
Consider the set , which is equal to the set . We consider the projection of those two sets by . Then since is -invariant, the projection is -equivariant. It implies that .
Then by Lemma 7.5, the distance between and is at least
since . It follows that the distance between and is at least . It follows that the product of and is hyperbolic (this is a well-known fact in -hyperbolic geometry, ie, if the distance between and is at least , then is hyperbolic since both satisfy the property (M2)). ∎
This finishes the explanation for the part to produce from , 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 that is large enough to apply Lemma 7.4. We explain how we define the map .
Let . Choose a shortest representative word of length on for . By the bijection between and , canonically gives a word of length on .
From this word , we construct a collection of elements in . Given a positive integer with , we divide the word into a prefix of length , and a suffix of length . The prefix corresponds to an element in that we denote , and the suffix corresponds to an element in that we denote .
Now, from the four separators we constructed in Lemma 7.4 we choose a separator for such that is admissible for and is admissible for , after we map them to by . We are writing instead of to indicate that the separator is in .
To the pair , we associate the following element in : .
Note that for all since .
In this way, we obtain “words” in from , but possibly, some of them represent the same elements in . 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 , take large enough such that is injective on from now on.
Definition 7.7 (Forbidden words. cf. Def 3.4 in [FS]).
Let be a word of length on in the above explanation. We say that a word , from the collection that is built from , is forbidden if there exists , such that:
We give a bound on the number of forbidden words.
Lemma 7.8 (cf. Lemma 3.5, [FS]).
For and each word as in Definition 7.7, there are at most forbidden words of the form: for .
In the proof of this lemma we use the assumption that is injective on 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 , but here, we do it in the set . Namely, if is forbidden for some , then there are two other elements in that are candidates for non-forbidden elements. Then if at least one of them is forbidden, then there are two other elements in 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 and the word as above, the non-forbidden words: , for all , , are distinct elements in .
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 . We choose a maximal subset in the set of non-forbidden elements in such that for any two distinct elements in the subset, are not in the same coset w.r.t. for any separator . We call those elements adequate elements. This notion depends on . Then, as before,
This is because, as we explained, there are only four separators, , and . In conclusion, since ,
(5) |
Using adequate elements, we construct a collection of feasible words in .
Definition 7.10 (Feasible words in . cf. Definition 3.7 [FS]).
Let be positive integers. Let be a word of length on that is associated to an element as in the above discussion. We present as a concatenation of subwords of length : .
Then for any choice of integers: with , and , for which all the elements, , are adequate (here we drop the “hat” from deliberately although they are in to avoid confusion since we want to use it right in the below) , we associate a feasible word (of type ) on (in ):
where for each , , is one of the separators from Lemma 4.2 such that is admissible for and is admissible for .
Finally we define the map . Suppose positive integers are given. Then for , choose one shortest representative of length on , which defines a word of length on as in the definition 7.10. From we produce feasible words on , which define feasible elements in . Note that those elements are in and mapped to by . We denote this collection as .
We have :
Lemma 7.11 (cf. Lemma 3.8 [FS]).
For any positive integers , the feasible elements in the collection we obtain for each are all distinct in .
Moreover, all the feasible elements obtained from all the elements in are all distinct in .
In this lemma, must be large enough in the sense that Lemma 7.4 applies and also, for the given , the map is injective on (cf. Lemma 4.5). We will choose such 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 , we have
(6) |
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 is given to start with. We want to show . Then the constant is given by Lemma 7.4, which does not depend on . Choose large enough so that we can apply Lemma 7.4.
Then choose such that
This implies:
Then choose larger if necessary such that is injective on . We need this to apply Lemma 7.9.
7.4. Family version
We state a family version of Proposition 7.2. Let be constants and a function for WPD. Suppose is -hyperbolic, a group acts on and is a finite generating set of such that contains a hyperbolic element on that is -WPD.
Suppose that for all , and let be a free group of rank with a free generating set . Let be a surjection with a bijection . Assume that the sequence converges to .
Also, assume that for all , there is a surjection with . 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 , there is a surjection with . If is infinite for some , then
7.5. Finiteness for
A finiteness result similar to Theorem 7.1 holds for subgroups. It is known for hyperbolic groups, [FS, Theorem 5.3].
Let be two finite subsets. Let be the subgroup generated by . We say and are isomorphic if there is a bijection between that induces an isomorphism between .
Theorem 7.13 (Finiteness in the subgroups case).
Assume the same condition on as in Theorem 7.1. Moreover, we assume that if is a finite set of such that contains a hyperbolic isometry on , then contains a hyperbolic isometry that is -WPD. Let , Then there are at most finitely many , up to isomorphism, such that is finite, , contains a hyperbolic element on , and .
The proof is nearly identical to the proof of Theorem 7.1 and we only need to modify the setting from the entire group to subgroups.
Proof.
To argue by contradiction, let and suppose that there are infinitely many distinct, up to isomorphism, with such that contains a hyperbolic isometry on . Note that by assumption, contains a hyperbolic isometry that is -WPD.
As before, by Proposition 2.10, passing to a subsequence, one may assume that there is with for all . Then we obtain with , where is the free group on with . Then, passing to a subsequence again, converges to a limit group with . Since is equationally Noetherian, by Lemma 1.11, passing to a further subsequence, we may assume that there are with for all .
First, by Proposition 6.2, we have that for all .
On the other hand, we prove a version of Proposition 7.2 for subgroups: if is infinite for some , then . The proof is same and we only outline it. As before set . First, since contains a hyperbolic isometry on that is -WPD, each contains the maximal finite normal subgroup, which we denote by with , (by Lemma 7.3).
The key step is to prove a lemma similar to Lemma 7.4. The argument is same. We use that has a hyperbolic and -WPD element for all . Then, as before, for all sufficiently large , there exists an element such that is hyperbolic on , and that the word length of in terms of is bounded uniformly on . 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 are common. But we did not describe the details of the argument for that.)
Combining those two, we conclude that is finite for all .
Finally, there are only finitely many possibilities for since it is contained in and . It implies that the desired finiteness for holds as before. ∎
7.6. Examples
Theorem 7.14 (Finiteness for relatively hyperbolic groups).
Let be a group that is hyperbolic relative to a collection of subgroups . Suppose is not virtually cyclic, and not equal to for any . Suppose each is finitely generated and equationally Noetherian. Then for each there are at most finitely many finite generating sets of , up to , s.t. .
Moreover, for each , there are at most finitely many , up to isomorphism, s.t. , where is finite and is not conjugate into any .
Proof.
Let be a hyperbolic space on which acts as we explained in Section 5.1. We also verified that all the assumption of Theorem 7.1 for and the action of on . Recall that the action of on is uniformly WPD by Lemma 5.2. It implies the first part of the theorem.
Theorem 7.15 (Finiteness for lattices).
Let be a group in Theorem 5.5. Then for each there are at most finitely many finite generating sets , up to , s.t. .
Moreover, for each , there are at most finitely many , up to isomorphism, s.t. , where is finite and .
Lastly we record the following (potential) example:
Theorem 7.16 (Finiteness for MCG).
Let be the mapping class group of a compact orientable surface . Assume that it is equationally Noetherian.
Then for each there are at most finitely many finite generating sets , up to , such that .
Moreover, for each , there are at most finitely many , up to isomorphism, such that , where is finite and is a large subgroup.
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, -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.