Finiteness of CAT group actions
Abstract.
We prove some finiteness results for discrete isometry groups of uniformly packed CAT-spaces with uniformly bounded codiameter (up to group isomorphism), and for CAT-orbispaces (up to equivariant homotopy equivalence or equivariant diffeomorphism); these results generalize, in nonpositive curvature, classical finiteness theorems of Riemannian geometry. As a corollary, the order of every torsion subgroup of is bounded above by a universal constant only depending on the packing constants and the codiameter. The main tool is a splitting theorem for sufficiently collapsed actions: namely we show that if a geodesically complete, packed, CAT-space admits a discrete, cocompact group of isometries with sufficiently small systole then it necessarily splits a non-trivial Euclidean factor.
1. Introduction
This is the first of two papers devoted to the theory of convergence for groups acting geometrically (that is discretely, by isometries and with compact quotient) on CAT-spaces . In this one, we will mainly focus on finiteness and splitting results. In our work, all CAT-spaces are assumed to be proper and geodesically complete, which ensures many desirable geometric properties, such as the equality of topological dimension and Hausdorff dimension, the existence of a canonical measure , etc. (see [Kle99], [LN19] and Section 2 for fundamentals on CAT-spaces).
The following is the first main finding of this paper:
Theorem A (Finiteness).
Given , there exist only finitely many groups
acting
discretely and nonsingularly by isometries on some proper, geodesically complete, -packed, -space with quotient of diameter at most , up to isomorphism of abstract groups.
(We stress the fact that, in the above theorem, the CAT-spaces on which the groups are supposed to act are not fixed a-priori).
Here is a quick explanation of the terms used in the statement above.
We say that a group acts nonsingularly on if there exists at least one point such that is trivial (in particular, the action of on is faithful).
This condition has many natural consequences (e.g., the existence of a fundamental domain)
and is a mild assumption on the action, which is automatically satisfied for instance when the action is faithful and either the group is torsion-free or is a homology manifold (see Section 3 and [CS22]).
We stress the fact that nonsingularity is essential for Theorem A: in [BK90],
Bass and Kulkarni exhibit an infinite family of non-isomorphic, discrete groups acting (singularly) on the same CAT space (a regular tree with bounded valency) with diam for all ; we thank P.E. Caprace for pointing out to us this example (see Remark 5.3). We will say that acts -cocompactly on if diam.
On the other hand, a metric space is -packed if all balls of radius in contain at most points that are -separated. This condition should be thought as weak form of lower curvature bound at scale ; however, it is much weaker than assuming the curvature bounded below in the sense of Alexandrov, or a lower bound on the Ricci curvature for Riemanniann manifolds, or the CD condition for metric measure spaces. Indeed, the Bishop-Gromov’s comparison theorem for Riemannian -manifolds with Ric (or its generalization to CD spaces) yields a doubling condition
(1) |
for the measure of all -balls with (for the Riemannian measure in case of manifolds, and for the reference measure in case of CD-spaces, cp. for instance [Stu06]), from which the packing condition at scale easily follows. See [BCGS17] for a detailed comparison of packing, doubling, entropy and curvature conditions. Let us just recall here that the packing condition for a CAT-space is also strictly weaker than the doubling condition (1) for the natural measure since a local doubling condition implies that the space has pure dimension (i.e. the tangent cones at all points have the same geometric dimension , see [CS21, Theorem C]), which restricts considerably the class of spaces under consideration.
Moreover, for the spaces we are interested in, a packing condition is a natural, and in some sense minimal, assumption. Indeed, every metric space with a compact quotient is -packed for suitable constants (cp. [Cav22a, Proof of Lemma 5.4]). Namely, let be the class of all isometric actions where is a proper, geodesically complete, CAT-space, and is a -cocompact, discrete subgroup of ; then the -packing conditions define a filtration
where is the subset of made of the actions such that is, moreover, -packed.
As proved in [CS21], for geodesically complete CAT-spaces, a packing condition at some scale is equivalent to a uniform upper bound of the canonical measure of all -balls (a condition sometimes called macroscopic scalar curvature bounded below cp. [Gut10], [Sab20]). Namely, there exist functions depending only on such that for all and we have (cp. Proposition 2.3):
(2) |
Moreoever, for geodesically complete CAT-spaces the -packing condition yields an important generalization of the classical Margulis’ Lemma, due to Breuillard-Green-Tao: there exists a constant such that for every discrete group of isometries of the -almost stabilizer of any point is virtually nilpotent (cp. [BGT11] and Section 2.2 for details). We call this the Margulis constant, since it plays the role of the classical Margulis constant in our metric setting.
We will see that the finiteness up to isomorphism of Theorem A
can be improved to finiteness up to equivariant homotopy equivalence of the pairs
(see Section 5.2 for the precise definition), and then deduce from Theorem A
corresponding finiteness results for the quotient spaces . We will call such spaces
CAT-orbispaces
111When restricting our attention to groups acting rigidly on (i.e. such that every acting as the identity on an open subset is trivial) then this definition is equivalent to the notion of rigid, developable orbispace as defined in [DLHG90, Ch.11] (with CAT universal covering in the sense of orbispaces). However, our results apply to all nonsingular CAT-orbispaces.,
using this term as a shortening for quotient of a (proper, geodesically complete) CAT-space by a discrete, isometry group ;
we will also say that the CAT-orbispace is nonsingular if acts nonsingularly on .
Notice that if is torsion-free then is a locally CAT-space.
For the following, let us define the class of all CAT-orbispaces
where is in CAT, and the subclasses
of those which are, respectively, locally CAT-spaces and non-singular.
Then we obtain:
Corollary B.
There are finitely many
spaces in the class up to homotopy equivalence, and finitely many spaces in , up to equivariant homotopy equivalence.
(Two orbispaces and are equivariantly homotopy equivalent if there exists a group isomorphism and a -equivariant homotopy equivalence , that is such that for all and all .)
A particular case of Theorem A, declined in the Riemannian setting, is:
Corollary C.
For every fixed , there exist only finitely many compact -dimensional Riemannian orbifolds with curvature and , up to equivariant diffeomorphisms.
(Two Riemannian orbifolds , are equivariantly diffeomorphic if there exists a diffeomorphism which is equivariant with respect to some isomorphism .)
In particular, for torsion-free orbifolds, this gives a new proof of the finiteness of compact Riemannian -manifolds with curvature and diam, modulo diffeomorphisms (this result was announced without proof by Gromov in [Gro78] and proved by Buyalo, up to homeomorphisms, in [Buy83]; the orbifold version was proved by Fukaya [Fuk86] in strictly negative curvature only).
Finiteness theorems in the spirit of Corollary C have been proved in different contexts in literature, the archetype of all of them being of course Weinstein’s theorem for pinched, positively curved manifolds [Wei67] and Cheeger’s finiteness theorems in bounded sectional curvature of variable sign (see [Che70], [Gro07], [Pet84], [Yam85]). In all these theorems, the finiteness is obtained by assuming a positive lower bound on the injectivity radius, or deducing such a lower bound from the combination of geometric and topological assumptions (see for instance [PT99] for simply connected manifolds with finite second homotopy groups).
A result similar to Theorem A was recently proved in [BCGS21], where the authors obtain the finiteness of torsionless groups acting faithfully, discretely and -cocompactly on -hyperbolic spaces with entropy . They achieve this by proving a positive, universal lower bound of the systole of , similar to the classical Heintze-Margulis’ Lemma (holding for manifolds with pinched, strictly negative curvature). Namely, for a discrete isometry group of let us call, respectively,
the systole of at and the (global) systole of (notice that the systole of the fundamental group of a nonpositively curved manifold , acting on its Riemannian universal covering , precisely equals twice the injectivity radius of ). Then, in [BCGS21], the authors prove that is bounded below by a positive constant only depending on the hyperbolicity constant, the entropy of and on the diameter of .This bound is, clearly, consequence of the Gromov hyperbolicity of the space , which is a form of strictly negative curvature at macroscopic scale. In contrast, for the group actions in the classes considered in Theorem A and in the above corollaries, the systole may well be arbitrarily small, since the curvature is only assumed to be non-positive. The main difficulty in Theorem A and in the corollaries above precisely boils down in understanding what happens when the systole or the injectivity radius tend to zero.
In fact, the problem of collapsing will be of primary interest in this work. Recall that a Riemannian manifold is called -collapsed if the injectivity radius is smaller than at every point. The theory of collapsing for Riemannian manifolds with bounded sectional curvature was developed by Cheeger, Gromov and Fukaya: for a differentiable manifold , the existence of a Riemannian metric with bounded sectional curvature sufficiently collapsed imposes strong restrictions to its topology. Namely, there exists an such that if a Riemannian -manifold with is -collapsed with , then admits a so-called -structure of positive rank, cp. [CG86]-[CG90] and [CFG92] (see also the works of Fukaya [Fuk87]-[Fuk88] for collapsible manifolds with uniformly bounded diameter).
More specifically about nonpositively curved geometry, Buyalo [Buy90a]-[Buy90b] first, in dimension smaller than , and Cao-Cheeger-Rong [CCR01] later, in any dimension, studied the possibility of collapsing compact, -dimensional manifolds with bounded sectional curvature . They proved that either the injectivity radius at some point is bounded below by a universal positive constant , or admits a so-called abelian local splitting structure: this is, roughly speaking, a decomposition of the universal covering into a union of minimal sets of hyperbolic isometries with the additional property that if two minimal sets intersect then the corresponding isometries commute. A prototypical example of collapsing with bounded, non-positive curvature is the following, which might be useful to have in mind for the sequel (cp. [Gro78], Section 5 and [Buy81], Section 4): consider two copies of the same hyperbolic surface with connected, geodesic boundary of length , then take the products with a circle of length , and glue to by means of an isometry of the boundaries which interchanges the circles with . This yields a nonpositively curved -manifold (a graph manifold) with sectional curvature , whose injectivity radius at every point is arbitrarily small provided that .
Coming back to our metric setting, where we also allow groups with torsion and isometries with fixed points, it is useful to distinguish between systole and free systole: we define the free systole of at and the (global) free systole of respectively as
where denotes the subset of torsion-free elements of .
The first non-trivial finding for non-singular actions is that, when assuming a bound on the diameter, then
the smallness of the systole at every point is quantitatively equivalent to the smallness of the global free systole (see Theorem 3.1).
Therefore, we will say that a -cocompact action of a group on a CAT-space (or, equivalently, the quotient space ) is -collapsed if .
The following theorem, which is the key to our finiteness theorems, shows that if admits a discrete, -cocompact action which is -collapsed for sufficiently small , then necessarily splits a non-trivial Euclidean factor:
Theorem D (Splitting of an Euclidean factor under -collapsed actions).
Let . There exists such that if is a proper, geodesically complete, -packed, -space admitting a discrete, -cocompact group of isometries with
then splits isometrically as , with .
Notice that, as the prototypical example above shows already for manifolds, the splitting of does not hold without an a priori bound on the diameter.
As proved by Buyalo in [Buy83], when is a manifold with pinched, nonpositive sectional curvature and diameter bounded by ,
then there exists a positive constant such that the condition of being -collapsed for implies the existence of a normal, free abelian subgroup of rank ; then,
virtually splits , and the existence of a non-trivial Euclidean factor for can be deduced from classical splitting theorems for non-positively curved manifolds (see [Ebe88], or [BH13, Prop.6.23]). The proof of Buyalo uses a stability property of hyperbolic isometries based on the fact that two isometries which coincide on an open subset are equal, a fact which is drastically false in our metric context (see the discussion in [CS22, Example 1.2]).
Also, notice that in Theorem D we do not assume any bound on the order of torsion elements (as opposite to [Fuk86]):
actually, bounding the order of the torsion elements of is one of the main conclusions of this work, see Corollary F below.
Our proof is more inspired to [CM09b]-[CM19]: we do not prove the existence of a normal free abelian subgroup, rather we find a free abelian, commensurated subgroup of , from which we construct a -invariant closed convex subset which splits as , and then we use the minimality of the action to deduce that (see Section 4 for the proof). Actually, in our metric setting it can happen that does not have any non-trivial, normal, free abelian subgroup at all (see [CS23]).
The final step for Theorem A is realizing that the -collapsing of an action of on can occur on different subspaces of at different scales; a careful analysis of this phenomenon allows us to renormalize, in a precise sense, the metric of , keeping the diameter of the quotient bounded, and obtaining the following result, similar to Buyalo’s [Buy83, Theorem 1.1], which we believe is of independent interest (see Section 5 and Remark 5.2 for a more precise statement):
Theorem E (Renormalization).
Given , there exist and such that the following holds.
Let be a discrete and -cocompact isometry group of
a proper, geodesically complete, -packed, -space : then, admits also a faithful, discrete, -cocompact action by isometries on a -space isometric to , such that
.
Moreover, the action of on is nonsingular if and only if the action on is nonsingular.
This theorem allows us to deduce the finiteness of nonsinngular groups in the class CAT, using Serre’s classical presentation of groups acting on simply connected spaces, and to obtain Corollary C from Cheeger’s and Fukaya’s finiteness theorems.
Remark. Explicitly, we may take , where bounds the dimension of every in CAT (see Proposition 2.3). Also the constant could be made explicit in terms of and of the Margulis’ constant , following the proof of Theorem E and using the function appearing in Theorem D (itself explicitable in terms of and of the universal bound of the index of the lattice of translations in any crystallographic group of dimension ). The only quantity which is not explicit is the Margulis constant provided by [BGT11].
Another immediate (although a-priori highly non-trivial) consequence of Theorem A is that there exists a uniform bound on the order of the finite subgroups of the groups under consideration. We stress that the existence of such a bound is new also for isometry groups of Hadamard manifolds; for instance, this is a key-assumption in the theory of convergence of Riemannian orbifolds of Fukaya. A similar bound is proved in [Fuk86] under the additional assumption of a lower bound on the volume of . More explicitly, there exists a constant (whose geometric meaning is explained in Proposition 3.2) such that the following holds:
Corollary F (Universal bound of the order of finite subgroups).
Let . Then for every nonsingular
in CAT,
every finite subgroup has order
Acknowledgments. The authors thank P.E.Caprace, S.Gallot and A.Lytchak for many interesting discussions during the preparation of this paper, and D.Semola for pointing us to interesting references.
Notation. Throughout the paper, we will adopt the following convention: a letter with the subscript 0 will always denote a fixed constant or a universal function depending only on the parameters and, possibly, on . For instance, in the above theorems, we used the constants with this meaning.
2. Preliminaries on CAT-spaces
We fix here some notation and recall some facts about CAT-spaces.
Throughout the paper will be a proper metric space with distance .
The open (resp. closed) ball in of radius , centered at , will be denoted by (resp. ); we will often drop the subscript when the space is clear from the context.
A geodesic in a metric space is an isometry , where is an interval of . The endpoints of the geodesic are the points and ; a geodesic with endpoints is also denoted by . A geodesic ray is an isometry and a geodesic line is an isometry . A metric space is called geodesic if for every two points there is a geodesic with endpoints and .
A metric space is called CAT if it is geodesic and every geodesic triangle is thinner than its Euclidean comparison triangle : that is, for any couple of points and we have where are the corresponding points in (see for instance [BH13] for the basics of CAT(0)-geometry). As a consequence, every CAT-space is uniquely geodesic: for every two points there exists a unique geodesic with endpoints and .
A CAT-metric space is geodesically complete if evvery geodesic can be extended to a geodesic line. For instance, if a CAT-space is a homology manifold then it is always geodesically complete, see [BH13, Proposition II.5.12]).
The boundary at infinity of a CAT-space (that is, the set of equivalence classes of geodesic rays, modulo the relation of being asymptotic), endowed with the Tits distance, will be denoted by , see [BH13, Chapter II.9].
A subset of is said to be convex if for all the geodesic is contained in . Given a subset we denote by the convex closure of , that is the smallest closed convex subset containing .If is a convex subset of a CAT-space then it is itself CAT, and its boundary at infinity naturally and isometrically embeds in .
We will denote by HD and TD the Hausdorff and the topological dimension of a metric space , respectively. By [LN19] we know that if is a proper and geodesically complete CAT-space then every point has a well defined integer dimension in the following sense: there exists such that every small enough ball around has Hausdorff dimension equal to . This defines a stratification of into pieces of different integer dimensions: namely, if denotes the subset of points of with dimension , then
The dimension of is the supremum of the dimensions of its points: it coincides with the topological dimension of , cp. [LN19, Theorem 1.1].
Calling the -dimensional Hausdorff measure, the formula
defines a canonical measure on which is locally positive and locally finite.
2.1. Discrete isometry groups
Let be the group of isometries of , endowed with the compact-open topology: as is proper, it is a topological, locally compact group.
The translation length of is by definition
When the infimum is realized, the isometry is called elliptic if and hyperbolic otherwise. The minimal set of , , is defined as the subset of points of where realizes its translation length; notice that if is elliptic then is the subset of points fixed by . An isometry is called semisimple if it is either elliptic or hyperbolic; a subgroup of Isom is called semisimple if all of its elements are semisimple.
Let be a subgroup of Isom. For and we set
(3) |
(4) |
When the context is clear we will simply write and .
The subgroup is discrete if it is discrete as a subset of Isom(X) (with respect to the compact-open topology). Since is assumed to be proper and acts by isometries, this condition is the same as asking that the orbit is discrete and is finite for some (or, equivalently, for all) . This is in turn equivalent to asking that the sets are finite for all and all .
When dealing with isometry groups of CAT-spaces with torsion, a difficulty is that there may exist nontrivial elliptic isometries which act as the identity on open sets.
Following [DLHG90, Chapter 11], a subgroup of Isom will be called rigid (or slim, with the terminology used in [CS22]) if for all the subset has empty interior. Every torsion-free group is trivially rigid, as well as any discrete group acting on a CAT-homology manifold, as proved in [CS22, Lemma 2.1].
A CAT-orbispace
(in the sense of [Fuk86])
is the quotient of a (proper, geodesically complete) CAT-space by a discrete isometry group .
One might define the notion of orbispace in terms of an orbifold atlas, that is a collection of uniformizing charts covering (where is a locally compact space endowed with the action of a finite group such that induces a homeomorphism ) and a pseudogroup of local homeomorphisms given by changing charts.
This is for instance the approach of Haefliger in [DLHG90, Ch.11], where rigid orbispaces are defined; rigid here means that the actions of the groups on are all supposed to be rigid.
Every quotient of a CAT-space by a discrete and rigid group has a structure of rigid orbispace; reciprocally, every rigid orbispace with nonpositive curvature (that is, such that the domains of the uniformizing charts are locally CAT-spaces) is developable, which means that , for some rigid, discrete isometry group acting on a suitable CAT-space , see [DLHG90, Ch.11, Théorème 8].
A group is said to be cocompact if the quotient metric space is compact; in this case, we call codiameter of the diameter of the quotient, and we will say that is -cocompact if it has codiameter at most . Notice that the codiameter of coincides with
(5) |
It is well-known, and we will consistently use it in the paper, that if is geodesic and is discrete and -cocompact, then for all the subset is a generating set for , that is , for every ; we call this a -short generating set of at .
Moreover, if is simply connected, then admits a finite presentation as
where is a subset of words of length on , see [Ser03, App., Ch.3]. For later use we also record the following general fact.
Lemma 2.1.
Let be a geodesic metric space and let be a discrete, -cocompact group. If is normal subgroup of with index , then is at most -cocompact.
Proof.
As is discrete, the space is geodesic and the group acts by isometries on it with codiameter at most . Take arbitrary and connect them by a geodesic . Let , as far as is defined for integers , and let be such that . By construction, the points are all distinct since they are all -separated, so the ’s are distinct. Since has cardinality at most , we conclude that . By the arbitrariness of , and we deduce that the diameter of is at most . ∎
In general the full isometry group Isom of a CAT-space is not a Lie group, for instance in the case of regular trees. When a CAT-space admits a cocompact, discrete group of isometries then Isom is known to have more structure, as proved by P.-E.Caprace and N.Monod.
Proposition 2.2 ([CM09b, Thm.1.6 & Add.1.8], [CM09a, Cor.3.12]).
Let be a proper, geodesically complete, -space, admitting a discrete, cocompact group of isometries. Then splits isometrically as , where is a symmetric space of noncompact type and is totally disconnected. Moreover
where is a semi-simple Lie group with trivial center and without compact factors and .
2.2. The packing condition and Margulis’ Lemma
Let be a metric space and .
A subset of is called -separated if for all . Given and we denote by Pack the maximal cardinality of a -separated subset of . Moreover we denote by Pack the supremum of Pack among all points of .
Given we say that is -packed at scale (or -packed, for short) if Pack. We will simply say that is packed if it is -packed at scale for some .
The packing condition should be thought as a metric, weak replacement of
a Ricci curvature lower bound: for more details and examples see [CS21]. Actually, by Bishop-Gromov’s Theorem, for a -dimensional Riemannian manifold a lower bound on the Ricci curvature implies a uniform estimate of the packing function at any fixed scale , that is
(6) |
where is the volume of a ball of radius in the -dimensional space form with constant curvature .
Also remark that every metric space admitting a cocompact action is packed (for some ), see the proof of [Cav22a, Lemma 5.4].
The packing condition has many interesting geometric consequences for complete, geodesically complete CAT-spaces, as showed in [CS20], [Cav21] and [Cav22b]. The first one is a uniform estimate of the measure of balls of any radius and an upper bound on the dimension, which we summarize here.
Proposition 2.3 ([CS21, Thms. 3.1, 4.2, 4.9], [Cav22b, Lemma 3.3]).
Let be a complete, geodesically complete, -packed, -space. Then is proper and
-
(i)
for all ;
-
(ii)
the dimension of is at most ;
-
(iii)
there exist functions depending only on such that for all and we have
(7) -
(iv)
The entropy of is bounded above in terms of and , namely
In particular, for a geodesically complete, CAT space which is -packed, the assumptions complete and proper are interchangeable.
Also, property (i) shows that, for complete and geodesically complete CAT(0)-spaces, a packing condition at some scale yields an explicit, uniform control of the packing function at any other scale : therefore, for these spaces, this condition is equivalent to similar conditions which have been considered by other authors with different names (“uniform compactness of the family of -balls” in [Gro81]; “geometrical boundedness” in [DY05], etc.).
The following remarkable version of the Margulis’ Lemma, due to Breuillard-Green-Tao, is another important consequence of a packing condition at some fixed scale. It clarifies the structure of the groups for small , which are sometimes called the “almost stabilizers”. We decline it for geodesically complete CAT-spaces.
Proposition 2.4 ([BGT11, Corollary 11.17]).
Given , there exists such that the following holds.
Let be a proper, geodesically complete, -packed, -space and let be a discrete subgroup of : then, for every and every , the almost stabilizer is virtually nilpotent.
2.3. Crystallographic groups in the Euclidean space
We will denote points in by a bold letter v, and the origin by .
Among CAT-spaces, the Euclidean space and its discrete groups play a special role.
A crystallographic group is a discrete, cocompact group of isometries of some . The simplest and most important of them, in view of Bieberbach’s Theorem, are Euclidean lattices: i.e.
free abelian crystallographic groups.
It is well known that a lattice must act by translations on (see for instance [Far81]); so, alternatively, a lattice can be seen as the set of linear combinations with integer coefficients of independent vectors (we will make no difference between a lattice and this representation).
The integer is also called the rank of the lattice.
A basis of a lattice is a set of independent vectors that generate as a group. There are many geometric invariants classically associated to a lattice , we will need just two of them:
– the covering radius, which is defined as
– the shortest generating radius, that is
Notice that, by the triangle inequality, any lattice is -cocompact. By definition it is always possible to find a basis of such that ; this is called a shortest basis of . The shortest generating radius and the covering radius are related as follows:
(8) |
For our purposes, the content of the famous Bieberbach’s Theorems can be stated as follows.
Proposition 2.5 (Bieberbach’s Theorem).
There exists , only depending on , such that the following holds true. For every crystallographic group of the subgroup is a normal subgroup of index at most , in particular a lattice.
Here denotes the normal subgroup of translations of . The subgroup is called the maximal lattice of . It is well-known that every lattice of rank has finite index in .
2.4. Virtually abelian groups
Recall that the (abelian, or Prüfer) rank of an abelian group , denoted , is the maximal cardinality of a subset of -linear independent elements. We extend this definition to virtually abelian groups , defining rk as the rank of every free abelian subgroup of finite index in : notice that if is a finite index subgroup of an abelian group , then and have same rank, so rk is well defined. One can equivalenty define rk as the rank of every normal, free abelian subgroup of finite index in , since every finite index subgroup of contains a normal, finite index subgroup. It is easy to show that the abelian rank is monotone on subgroups.
If is a discrete, finitely generated, semisimple free abelian group of isometries of a CAT-space , then its minimal set
is not empty and splits isometrically as where . This is the main content of the Flat Torus Theorem (see [BH13, Theorem II.7.1]).
We recall some additional facts about the identification , which we will freely use later:
(a) the abelian group acts as the identity on the factor , and cocompactly by translations on the Euclidean factor ;
(b) writing , the slice coincides with the convex closure of the orbit ;
(c) one has for every finite index subgroup and every .
The last assertion follows from the fact that and are both isometric to , so they necessarily coincide.
As a direct consequence of the Flat Torus Theorem we have the following property for virtually abelian isometry groups of CAT-spaces, that we will often use.
Lemma 2.6.
Let be a proper -space and let be discrete, semisimple, virtually abelian groups of isometries of . Then is finite if and only if and have same rank.
Proof.
The implication is trivial, as every free abelian finite index subgroup is also a finite index subgroup of . To show the converse implication, assume that , and let be a rank , free abelian, finite index normal subgroup of . Consider the (free abelian) subgroup of , and notice that we have , since also . Now, both and act faithfully on the Euclidean factor of (they do not contain elliptics since they are free, and act as the identity on ). Therefore their projections on are both rank Euclidean lattices, hence . But then we deduce that . ∎
The following generalization of the Flat Torus Theorem is classical.
Proposition 2.7 ([BH13, Corollary II.7.2]).
Let be a proper -space, and let be a discrete, semisimple, virtually abelian group of isometries of of rank . Then, there exists a closed, convex, -invariant subset of which splits as , satisfying the following properties:
-
(i)
every preserves the product decomposition and acts as the identity on the first component;
-
(ii)
the image of under the projection is a crystallographic group.
Given a generating set of , the following statement explains how to construct a generating set for the maximal lattice with words on of bounded length. This will be used later in the proof of Theorem 4.1:
Lemma 2.8.
Here denotes the subset of all products of at most elements of ; by definition, this coincides with , where is the closed ball of radius , centered at the identity , in the Cayley graph .
Proof of Lemma 2.8.
Call the projection on Isom. By Proposition 2.5 the normal subgroup has index at most in , so the normal subgroup has index at most in . The group acts discretely by isometries on Cay with codiameter . So, by Lemma 2.1 the group acts on Cay with codiameter at most . As recalled before Lemma 2.1, is therefore generated by the subset , made of the elements of displacing the point in the Cayley graph Cay at most by . It follows that the projection generates . ∎
Finally, remark that given a discrete, semisimple, virtually abelian group of isometries of a proper CAT-space, there can be several closed, convex, -invariant subsets satisfying the conclusions of Proposition 2.7. What is uniquely associated to is a subset in the boundary of .
Proposition 2.9.
Same assumptions as in Proposition 2.7.
Then there exists a closed, convex, -invariant subset of , denoted ,
which is isometric to and has the following properties:
-
(i)
, for every free abelian finite index subgroup of and every ;
-
(ii)
for every subgroup we have ; moreover if .
The closed subset provided by this proposition will be called the trace at infinity of the virtually abelian group .
Proof.
We fix a free abelian, normal subgroup with finite index and . By the Flat Torus Theorem, is isometric to . We set . Clearly is closed, convex, isometric to . Notice that the set does not depend on the choice of , since for the subsets and can be identified, by the Flat Torus Theorem, to two parallel slices and , which have the same boundary. Also, the subset is -invariant: in fact, is normal in , so is -invariant, therefore
because .
Again by the Flat Torus Theorem (namely, property (c) recalled before), if is another free abelian subgroup of finite index and then
, so we have too. This proves (i).
To show (ii), it is enough to consider free abelian subgroups and of finite index. We can even suppose up to replacing by . If then , and by the first part of the statement we have
If moreover and have the same rank then is a finite index subgroup of by Lemma 2.6, and .
∎
3. Systole and diastole
In this section we compare some different invariants of an action of group on a CAT-space , which are related to the problem of collapsing: the systole and the diastole of the action (and their corresponding free analogues), which play the role of the injectivity radius.
Recall that the systole and the free-systole of at a point are defined respectively as
where and is the subset of all elliptic isometries of .
The (global) systole and the free-systole of are accordingly defined as
Similarly, the diastole and the free-diastole of are defined as
In [CS22] the authors showed that dias if and only if there exists a fundamental domain for ; that is, if and only if there exists a point such that the pointwise stabilizer is trivial. The actions satisfying this property will be called nonsingular, and singular when dias (with abuse of language, we will often say that the group itself is singular or nonsingular). For a nonsingular action, the Dirichlet domain at is defined as
and is always a fundamental domain for the action of , see [CS22, Prop. 2.9]. Notice that a fundamental domain exists when, for instance, is torsion-free, or when is a homology manifold, as follows by the combination of Lemma 2.1 and Proposition 2.9 of [CS22].
By definition, we have the trivial inequalities:
The following result shows that the free systole and the free diastole are for small values quantitatively equivalent, provided one knows an a priori bound on the diameter of the quotient. Moreover, for nonsingular actions, both are quantitatively equivalent to the diastole.
Theorem 3.1.
Given , there exists such that the following holds true. Let be a proper, geodesically complete, -packed, -space and let be discrete and -cocompact. Then:
-
(i)
.
-
(ii)
If moreover the action of on is nonsingular then
(Here, is the Margulis’ constant given by Proposition 2.4).
Dropping the assumption of nonsingularity. (ii) is no longer true: see [CS22, Example 1.4], where dias while the free systole is positive. Also, it is easy to convince oneself that the usual systole is not equivalent, for small values, to the other three invariants: for instance, for every discrete, cocompact action of a group on a proper CAT-space , one has if there exists some point trivial stabilizer, but clearly if has torsion. Finally, notice that the inequality (ii) holds for a constant depending only on and , and not on ; it is not difficult to show that the same is not true for (i).
To show the above equivalences, we need two auxiliary facts.
The first one is a generalization of Buyalo’s and Cao-Cheeger-Rong’s theorem about the existence of abelian, local splitting structure, for groups acting faithfully and geometrically on packed, CAT-spaces which are -thin for sufficiently small .
Proposition 3.2 ([CS22, Theorem A & Remark 3.5]).
Let and fix , where is given by Proposition 2.4. There exists a constant such that the following holds true. Let be a proper, geodesically complete, -packed, -space and let be discrete and cocompact. If dias then
The second fact is the following result, which propagates the smallness of the systole for torsion-free cyclic groups from point to point.
Proposition 3.3.
Let and . Then there exists with the following property. Let be a proper, geodesically complete, -packed, CAT-space. If is an isometry of with infinite order and is a point of such that , then for every with there exists such that .
(We can choose, explicitely, ).
Proof.
This follows immediately from [CS20, Proposition 4.5.(ii)], where we expressed this property in terms of the distance between generalized Margulis domains222Notice that in [CS20] we were in the setting of Gromov-hyperbolic spaces with a geodesically complete convex geodesic bicombing (in particular, Gromov-hyperbolic, geodesically complete CAT-spaces); but in that proof we never used the hyperbolicity.. ∎
Proof of Theorem 3.1.
Assume that
.
By definition there exists such that for every hyperbolic isometry one has .
Now, if sys, we could find and a hyperbolic such that .
By Proposition 3.3, for every there would exists a non trivial power satisfying .
But then, since the action is -cocompact we could find a conjugate of (thus, a hyperbolic isometry) such that , a contradiction. The conclusion follows by the arbitrariness of .
To see (ii), take any . If , where is the constant given by Proposition 3.2, we conclude that
where runs over all nontrivial elements with .
But by definition for all hyperbolic , so
Hence, , contradicting the nonsingularity of . As is arbitrary, this proves (ii).
∎
4. The splitting theorem
In this section we will prove Theorem D, actually a stronger, parametric version given by Theorem 4.1 below. To set the notation, recall that for a proper, geodesically complete, CAT-space which is -packed we have an upper bound on the dimension of given by Proposition 2.3
and a Margulis’s constant (only depending on ) given by Proposition 2.4. Also, recall the constant given by Proposition 2.5, and define
which also clearly depends only on , so ultimately only on .
Finally, for a discrete subgroup
recall the definition (4) of the subgroup generated by given in Section 2.1.
Theorem 4.1.
Given positive constants , there exists a function (depending only on the parameters ) such that the following holds. Let be a proper, geodesically complete, -packed, -space, and be discrete and -cocompact. For every chosen , if then:
-
(i)
the space splits isometrically as , with , and this splitting is -invariant;
-
(ii)
there exists such that the rank of the virtually abelian subgroups is exactly , for all ;
-
(iii)
the traces at infinity equal the boundary of the convex subsets , for all and all ;
-
(iv)
for every there exists such that preserves ;
-
(v)
the projection of on is a crystallographic group, whose maximal lattice is generated by the projection of a subset ;
-
(vi)
the closure of the projection of on is compact and totally disconnected.
Here, by -invariant splitting we mean that every isometry of preserves the product decomposition. By of [BH13, Proposition I.5.3.(4)] we can see as a subgroup of . In particular it is meaningful to talk about the projection of on and .
Observe that Theorem D is a special case of Theorem 4.1.(i), for , which yields the constant . We call the integer given by (ii) the -splitting rank of .
To prove Theorem 4.1 we need a little of preparation.
The first step will be to exhibit a free abelian subgroup of rank which is commensurated in .
Recall that two subgroups are called commensurable in if the intersection has finite index in both and . A subgroup is said to be commensurated in if and are commensurable in for every .
Now, we fix as in the assumptions of Theorem 4.1, and we define inductively the sequence of subgroups associated to positive numbers
as follows:
– first, we apply Proposition 3.3 to and to obtain a smaller , and we set ;
– then, we define inductively by repeatedly applying Proposition 3.3 to and , and we set .
Notice that, by construction, each depends only on and .
By Proposition 2.4, the subgroups form a decreasing sequence of virtually abelian, semisimple subgroups for every .
We set and we will show that this is the function of for which Theorem 4.1 holds; it clearly depends only on and . In what follows, we will write for short .
Lemma 4.2.
If then there exists such that .
Proof.
The subgroups are virtually abelian with , for all , since they contain . Moreover, by Proposition 2.7, the rank of each cannot exceed the dimension of , which is at most . Since the rank decreases as increases we conclude that for some we have . ∎
Proposition 4.3.
If then there exists a free abelian subgroup of rank , which has finite index in and is commensurated in , for some .
Proof.
Consider the virtually abelian groups which have the same rank , given by Lemma 4.2, for some and let be free abelian, finite index subgroups of rank , for . Notice that has finite index also in , by Lemma 2.6. Then, let be the -invariant, convex subset of given by Proposition 2.7, applied to , and call the image of under the projection on the second factor of . Finally, denote by the maximal Euclidean lattice of the crystallograhic group .
By Lemma 2.8 we can find a subset of
whose projection on Isom generates the lattice .
In particular every non-trivial element of is hyperbolic.
Moreover, by the definition of ,
the following holds:
and there exists such that
(in fact, for every , so by Proposition 3.3 there exists such that , that is ).
Since and are finite sets, we can then find a positive integer such that for all and all .
Moreover, we can even choose so that and
the elements and belong, respectively, to the free abelian, finite index subgroups and of .
Now, let
be the (free abelian) subgroup generated by the subset .
We claim that is commensurated in .
Actually, notice first that has rank equal to , since its projection is a subgroup of finite index of the lattice of . Therefore, for all , the free abelian group has also rank , and is contained in the free abelian group of same rank. This implies, again by Lemma 2.6, that and have finite index in , and that has finite index in both and . Hence and are commensurable for every .
Now, given , we can write it as for some and set . We clearly have
with of finite index in both factors; so and are commensurable. Since commensurability in a group is a transitive relation, this shows that and are commensurable for all .
Finally, observe that and these groups have same rank, so has finite index also in , again by Lemma 2.6.
∎
We need now to recall an additional notion. Given we say that a subset is -boundary-minimal if it is closed, convex, and is minimal with this properties. The union of all the -boundary-minimal sets is denoted by Bd-Min. A particular case of [CM09b, Proposition 3.6] reads as follows.
Lemma 4.4.
Let be a proper -space and let be a closed, convex subset of which is isometric to . Then each -boundary-minimal subset of is isometric to and Bd-Min is a closed, convex subset of which splits isometrically as . Moreover coincides with the boundary at infinity of all the slices , for .
A consequence of Lemma 4.4 for a commensurated group of is the following.
Proposition 4.5.
Same assumptions as in Theorem 4.1.
If is a free abelian, commensurated subgroup of of rank then we have and splits isometrically and -invariantly as . Moreover, the projection of on is a lattice and the closure of the projection of on is compact and totally disconnected.
The splitting satisfies the following properties:
-
(i)
the trace at infinity is -invariant and coincides with the boundary of each slice , for all ;
-
(ii)
if is another free abelian, commensurated subgroup of of rank then the splittings and associated respectively to and are compatible, i.e. is isometric to .
Proof.
The trace at infinity of is isometric to , by Proposition 2.9. We claim that it is -invariant. Indeed, if then has finite index in both and . Then, the characterization of the trace at infinity given in Proposition 2.9 implies that
By Lemma 4.4 applied to we deduce that is a closed, convex subset of which splits isometrically as , and that coincides with the boundary at infinity of all sets .
Now, each element of sends a -boundary-minimal subset into a -boundary-minimal subset because is -invariant, therefore itself is -invariant. Since is cocompact, the action of on is minimal: that is, if is a closed, convex, -invariant subset of then (see [CM09b, Lemma 3.13]). Therefore we deduce that , and so splits isometrically and -invariantly as , which proves the first assertion and (i).
The fact that the projection of on is a lattice follows from the Flat Torus Theorem. To study the projection of on , recall that by Proposition 2.2 we can split as , for some , and as , where is a semi-simple Lie group with trivial center and without compact factors, and is totally disconnected.
Therefore and splits as .
Moreover, with .
Since acts as the identity on , it follows that the projection of on fixes some point , hence its closure is a compact group.
Finally, Theorem 2.(i) of [CM19] and the beginning of the proof therein show that the projection of on is finite and the projection of on is discrete; as the projection of on is a lattice, then also the projection on is discrete. This, combined with the fact that is totally disconnected, implies that the closure of the projection of on is totally disconnected.
Suppose now to have another abelian subgroup of rank which is commensurated in .
Let be the splitting associated to . Let be a point and write it as and . Then, by the first part of the proof and by Proposition 2.9 we have
It follows that , so the parallel slices associated to are contained in the parallel slices associated to . Decomposing as the orhogonal sum of and , we also deduce that the sets are parallel for all (since the slices of are all parallel). Therefore, is also isometric to , which implies that is isometric to and proves (ii). ∎
Putting the ingredients all together we can give the
Proof of Theorem 4.1.
We show that the statement holds for
, where is given by Proposition 4.3. In fact, since , then there exists with ; in particular, contains a hyperbolic isometry, hence . Then, we can apply Proposition 4.3 and find a free abelian, commensurated subgroup of rank , with and with finite index in ,
for some ; Proposition 4.5 now implies that splits isometrically and -invariantly as , proving (i).
Moreover, we know that coincides with the boundary at infinity of each slice .
Let us now study the properties (ii)-(vi) for the groups .
We start proving them for every such that .
By Proposition 3.3 and by definition of , for every hyperbolic there exists such that ,
that is .
Let be free abelian of finite index, so by definition. Since the set is finite, we can find such that for every hyperbolic .
So, if we define the subgroup
we have . Notice that is again free abelian of rank , hence it has finite index in and , by Lemma 2.6. Thus
Moreover by Proposition 2.9 we have
.
Reversing the roles of and and starting from we obtain the opposite estimate
and , which proves (ii) and (iii) in this case.
By construction has finite index in both and , so also the projection on is discrete, and the closure of the projection of on is compact and totally disconnected. Notice that is cocompact, so it is a crystallographic group of ; moreover, as generates , then the maximal lattice of is generated by a subset of , by Lemma 2.8. This proves (v) and (vi). Moreover, the precompact group has a fixed point ([BH13, Corollary II.2.8.(1)]), so is preserved by , proving (iv). Finally, assume that is a point of , say with .
Observe that
, so the rank does not change and
the conditions (iv) and (v) continue to hold since
the splitting is -invariant.
Moreover
, because is -invariant, so (iii) still holds.
Finally, if the group preserves , then clearly preserves , which proves (iv).
∎
5. The finiteness Theorems
The goal of this section is to prove Theorem A and its corollaries. The work is divided into two steps: we first prove the renormalization Theorem E, from which we immediately deduce the finiteness up to group isomorphism, and Corollary F. Then, we will improve the result showing the finiteness of the class (and of as a particular case) up to equivariant homotopy equivalence of orbispaces, that is Corollary B). Finally, we will deduce Corollary C from the renormalization Theorem E combined with Cheeger’s and Fukaya’s finiteness theorems.
5.1. Finiteness up to group isomorphism
Recall that a marked group is a group endowed with a generating set .
Two marked groups and are equivalent if there exists a group isomorphism such that ; notice that such a induces an isometry between the respective Cayley graphs.
The first step for Theorem A is the following:
Proposition 5.1.
Let be given. For every fixed and , there exist only finitely many marked groups where:
-
–
is a discrete, -cocompact isometry group of a proper, geodesically complete, -packed, CAT-space satisfying for some ,
-
–
is a -short generating set of at .
(Finiteness here is meant up to equivalence of marked groups.)
Proof.
Recall from Section 2.1 that for any the group admits a presentation as , where is a subset of words of length on the alphabet . Therefore the number of equivalence classes of such marked groups , with , is bounded above if we are able to bound uniformly the cardinality of . But this is an immediate consequence of Proposition 2.3; actually, using the fact that the points in the orbit of are -separated, we get
Now, the main idea to prove Theorem A is to use the Splitting Theorem 4.1 to show that, up to increasing in a controlled way the codiameter, we can always suppose that the free-systole is bounded away from zero by a universal positive constant: this is the content of the renormalization Theorem E, which is proved below.
The proof will show that the space is isometric to (though generally the quotients and are not isometric, since the first one can be -collapsed for arbitrarily small , while the second one is not collapsed, by construction).
Naively, one can think that if is too small then, as we know that splits -invariantly as by Theorem 4.1, the new space is , obtained by dilating the Euclidean factor of a suitable , and the action of is the natural one induced on it.
The construction of is however a bit more complicate, since this naïf renormalization is not sufficient in general to enlarge the free systole while keeping the diameter bounded. In order to make things work we need to take into account that can be collapsed on different subsets at different scales, and an algorithm allowing us to detect them.
We refer to Remark 5.2 for a more precise statement.
Recall the constants , which bounds the dimension of every proper, geodesically complete, CAT-space which is -packed, and , introduced at the beginning of Section 4. Also recall the Margulis constant given by Proposition 2.4, which we will always assume smaller than in the sequel.
Proof of Theorem E.
Recall the function of Theorem 4.1. Then we define inductively , and
We claim that and satisfy the thesis; notice that both depend only on and .
We now describe a process which takes the CAT-space and produces a new proper, geodesically complete, -packed, CAT-space on which still acts faithfully and discretely by isometries, still satisfying all the assumptions of the theorem, except that ; and we will show that, repeating again and again this process, we end up with a CAT-space with , for some .
If , there is nothing to do, and we just set .
Otherwise, , and we apply
Theorem 4.1
with .
Then, there exists such that the groups
have all rank for every . We then fix .
By Proposition 4.3 there exists
a free abelian, finite index subgroup of of rank , which is commensurated in ; and we have that splits isometrically and -invariantly as by Proposition 4.5.
Moreover, always by Theorem 4.1, there exists such that preserves , and there exists a subset of whose projection
on generates the maximal lattice of the crystallographic group .
So, we can find
a shortest basis of the lattice whose vectors have all length at most ; without loss of generality, we may suppose that , where denotes the Euclidean norm of .
By (8), we also know that the covering radius of is at most .
Now, we define the metric space
. This is again a proper, geodesically complete, CAT-space, still -packed, on which still acts discretely by isometries (because the splitting of is -invariant).We claim that the action of on is -cocompact. In fact, let be a point of . Since the action of on is -cocompact, we know that there exists such that ; moreover,
as preserves , we can compose with elements of this group in order to find such that and
. Therefore
As was arbitrary, we then deduce that
so is -cocompact.
If now , we stop the process and set : this space has all the desired properties.
Otherwise, we have
and we can apply again Theorem 4.1,
Proposition 4.3 and Proposition 4.5
to , with . Then, there exists
such that the groups
have rank for every , in particular . Moreover, there exists
a free abelian, finite index subgroup of rank which is commensurated in , the space splits isometrically and -invariantly as , and the trace at infinity coincides with the boundary at infinity of all the sets ; and there exists a subset of whose projection on generates the maximal lattice of .
Therefore, we can find
a shortest basis of with lengths (with respect to the Euclidean norm of )
Observe that, as by construction the factor is bigger than , we have
(9) |
so
,
and . In particular, the metric splittings of as and determined, respectively, by the groups and satisfy and , because of the second part of Proposition 4.5.
We will now show that .
Actually, suppose that .
Then,
the groups and
split the same Euclidean factor and .
The lengths of the basis with respect to the Euclidean norm of the Euclidean factor of are
for every .
But then, since by (9), we would be able to find independent vectors of of length less than its shortest generating radius, which is impossible.
This shows that .
We can now define , on which acts faithfully, discretely by isometries, -cocompactly, for computed as before.
We can repeat this process to get a sequence of proper, geodesically complete, -packed, CAT-spaces on which always acts faithfully, discretely and -cocompactly by isometries. Moreover at each step
either or the -splitting rank of provided by Theorem 4.1 is strictly smaller than the -splitting rank of . Since the splitting rank of is at most there must exist such that . The proof then ends by setting .
Clearly, we may take, explicitely, .
Finally, observe that, by construction, is isometric to the initial space , and that the action of on is nonsingular if and only if the action of on , and by induction on , was nonsingular.
∎
Remark 5.2.
The proof of Theorem E actually gives us something more. Indeed we produce a sequence of free abelian commensurated subgroups of
for some , such that:
-
(a)
denoting by the rank of , then ;
-
(b)
setting then there is a corresponding isometric and -invariant splitting of as . Moreover, there exist such that the natural action of on the space
(10) is -cocompact with free-systole at least .
The finiteness Theorem A is now an immediate consequence of the above renormalization result, combined with Proposition 5.1 and Theorem 3.1:
Proof of Theorem A.
By Theorem E every group under consideration acts discretely by isometries and -cocompactly on a proper, geodesically complete, -packed, CAT-space with . The action is still nonsingular, so by Theorem 3.1 we can find with , a positive lower bound depending only on and . Then, applying Proposition 5.1 with and , we conclude the proof.
∎
Remark 5.3.
The assumption that the groups act nonsingularly is necessary for Theorem A. Actually, in [BK90, Theorem 7.1], Bass and Kulkarni exhibit an infinite ascending family of discrete groups acting (singularly) on a regular tree with bounded valency, with same compact quotient, in particular with diam for all . Moreover these groups are lattices satisfying Vol, where the volume of in is defined as
This family contains infinitely many different groups, since the minimal order of a torsion subgroup in tends to infinity as the volume goes to zero (every torsion subgroup of stabilizes some point, as is CAT, hence is the dominant term of the sum yielding ).
Proof of Corollary F.
The number of isomorphism types of possible group is finite by Theorem B, so the thesis is a trivial consequence of this theorem. However we can give a more costructive proof that gives an explicit upper bound for the order of finite subgroups. By Theorem E we can suppose that is -cocompact and . Let be a finite subgroup. It fixes a point (cp. [BH13, Corollary II.2.8]). Let be the Dirichlet domain of at , which is clearly -invariant. By Theorem 3.1 there exists some point where . Therefore the orbit is a -separated subset of and the balls are all disjoint. Using the fact that is contained in we get
by Proposition 2.3.(iii). This yields the explicit bound for the cardinality of , only depending only on and . ∎
5.2. Finiteness up to equivariant homotopy equivalence
We have proved so far that the number of discrete, nonsingular and -cocompact groups of isometries of proper, geodesically complete, -packed, CAT-spaces is finite up to isomorphism of abstract groups.
In this section we will show the finiteness up to equivariant homotopy equivalence of orbispaces. That is, we can upgrade every
isomorphism of groups , acting discretely, nonsingularly and cocompactly by isometries respectively on CAT-spaces and , to a homotopy equivalence which is -equivariant, i.e. for all and all . Namely we show:
Proposition 5.4.
Let and be discrete, cocompact, nonsingular isometry groups of two proper, geodesically complete, -packed, -spaces and , respectively. If there exists a group isomorphism , then there exists a -equivariant homotopy equivalence .
This is a CAT(0) version of a result about the realization of orbifold group-isomorphisms by orbi-maps equivalences, see [Yam90, Theorem 2.5]. We will give a direct proof of this fact, using the barycenter technique introduced by Besson-Courtois-Gallot [BCG95] for strictly negatively curved manifolds (cp. [Sam99] for an approach similar to the one we follow here).
Let be a proper, CAT-metric space. Consider the space of positive, finite, Borel measures on with finite second moment, i.e. such that the distance function is -square integrable for some (hence every) . The barycenter of is defined as the unique point of minimum of the function
Notice that tends to for since for every fixed in we have, by triangular and Schwarz inequalities,
which diverges as . Moreover, by standard comparison with the Euclidean space, the function is also strictly convex, namely -convex in the sense of [Kle99]: that is,
is a convex function, for every geodesic .
It follows that the function is -convex as well, which implies that has a unique point of minimum (cp. [Kle99, Lemma 2.3]).
Therefore is well-defined.
It is straightforward to check that the barycenter is equivariant with respect to the natural actions of on and , that is:
(11) |
We can now give the Proof of Proposition 5.4:
Proof of Proposition 5.4..
As and are nonsingular, we can choose and such that the stabilizers , are trivial. Then there exists a unique -equivariant map sending to , namely . Now, by varc-Milnor Lemma, the spaces and are respectively quasi-isometric to the orbits and and also to the marked groups and endowed with their word metrics. Moreover, since the groups and are isomorphic, and any two word metrics associated to finite generating sets on the same group are equivalent, we deduce that the map is an -quasi-isometry for suitable , in particular
(12) |
We now define a -equivariant homotopy equivalence as follows: choose any (the upper bound of the entropy of given by Proposition 2.3.(iv) and consider the family of measures , indexed by and supported by the orbit of , given by
where is the Dirac measure at , and then define
Notice that the total mass of coincides with the value of the Poincaré series of the group acting on for , which is finite since is chosen greater than the critical exponent of the series (recall that equals the critical exponent of the group acting on , since the action is cocompact, see for instance [BCGS17] or [Cav22a, Proposition 5.7]). Moreover, the function is square summable with respect to , since we have by (12)
which is finite as the Poincaré series converges exponentially fast for .
Therefore the map is well-defined.
Step 1: is continuous.
First, we show that converge uniformly on compacts to for in .
Actually, let be a fixed compact subset. Since the action of on is discrete, the subset
is finite. Then for every choose some with such that . From (12) and the triangular inequality we find that, for , it holds
for . Now, the series in parentheses is bounded above independently of and , while (for fixed ) the term is uniformly bounded on since belong to the finite subset . The above estimate then implies that uniformily on when .
Secondly, we call and show that there exists such that for every the functions are greater than on .
Actually, let be such that outside (recall that is proper, since we showed that
for all ), and assume that we have infinitely many points with such that . Then, calling the point at distance from ,
we would have, by convexity,
Since uniformly on we would deduce that , which contradicts the choice of , since .
Now, since on for all , the uniform convergence
on implies that the sequence of (unique) minimum points of converge to the (unique) minimum point of . In fact, we have that for all , and
(13) |
so if accumulates to a point , passing to the limit in (13) yields . This implies that , by unicity of the minimum point of .
Step 2: is a -equivariant homotopy equivalence.
Firstly, the map is -equivariant, since for all we have
and by (11) we deduce
.
Then, from the inverse map
we can analogously construct a -equivariant, continuous map .
Now consider the homotopy map defined by the formula (where denotes the point on the geodesic segment at distance from ). The composition is -equivariant, so we deduce that is a -equivariant homotopy between and .
Similarly, one proves that the map is a -equivariant homotopy between and .
∎
5.3. Finiteness of nonpositively curved orbifolds
We restrict here our attention to CAT-orbispaces which are quotients, by a discrete isometry group , of a Hadamard manifold (that is, a complete, connected and simply connected, nonpositively curved Riemannian manifold) with pinched sectional curvature : we call such a quotient a nonpositively curved Riemannian orbifold with curvature .
Recall that, in this case, any discrete isometry group of is rigid, in the sense explained in Section 2.1.
For compact, non-positively curved Riemannian orbifolds, the finiteness up to equivariant homotopy equivalence can be improved to finiteness up to equivariant diffeomorphisms: recall that an equivariant diffeomorphism between two Riemannian orbifolds , is a diffeomorphism which is equivariant with respect to some group isomorphism , i.e. for all and all .
Proof of Corollary C.
First notice that the splitting Theorem D is true in the manifold category.
Actually, let be the De Rham decomposition of a Hadamard manifold , where is the product of all non-Euclidean factors. The factor coincides with the Euclidean factor splitted by as a CAT-space (since the decomposition of a finite dimensional geodesic space into flat and irreducible factors is invariant by isometries, cp. [FL08]).
Then, the factor splitted by Theorem D under an -collapsed action, with , is isometrically immersed in the Euclidean de Rham factor of the manifold decomposition of ; hence it is -embedded as a submanifold, as well as its orthogonal complement .Then, also the factor of Theorem D is -embedded in , because .
From this, we deduce that also the renormalization Theorem E holds in the manifold category: that is, every splitting considered in the proof is smooth, and the resulting decomposition (10) yields a smooth Riemannian structure on (such that as a differentiable manifold).
Moreover, if , the new Riemannian manifold satisfies the same curvature bounds
(since the metric is dilated on the Euclidean factors by the constants , as explained in Remark 5.2).
It then follows that every -dimensional Riemannian orbifold
with and diam is equivariantly diffeomorphic to
a Riemannian orbifold still satisfying , but with diameter bounded above by and with sys; where the constants and only depend on and on the packing constants of (and then, ultimately, only on and on the dimension , by (6)).Moreover, by Theorem 3.1, the systole of acting on is bounded below by at some point , where again only depend on .
We then deduce the finiteness of these orbifolds from Fukaya’s [Fuk86, Theorem 8.1] (or from Cheeger’s finiteness theorem, as completed in [Pet84], [Yam85], in the torsion-free case).
∎
References
- [BCG95] G. Besson, G. Courtois, and S. Gallot. Entropies et rigidités des espaces localement symétriques de courbure strictement négative. Geometric and Functional Analysis, 5:731–799, 09 1995.
- [BCGS17] G. Besson, G. Courtois, S. Gallot, and A. Sambusetti. Curvature-free Margulis lemma for Gromov-hyperbolic spaces. arXiv preprint arXiv:1712.08386, 2017.
- [BCGS21] G. Besson, G. Courtois, S. Gallot, and A. Sambusetti. Finiteness theorems for Gromov-hyperbolic spaces and groups. arXiv preprint arXiv:2109.13025, 2021.
- [BGT11] E. Breuillard, B. Green, and T. Tao. The structure of approximate groups. Publications mathématiques de l’IHÉS, 116, 10 2011.
- [BH13] M. Bridson and A. Haefliger. Metric spaces of non-positive curvature, volume 319. Springer Science & Business Media, 2013.
- [BK90] H. Bass and R. Kulkarni. Uniform tree lattices. Journal of the American Mathematical Society, 3(4):843–902, 1990.
- [Buy81] S. Buyalo. Manifolds of nonpositive curvature with small volume. Mathematical notes of the Academy of Sciences of the USSR, 29(2):125–130, 1981.
- [Buy83] S. Buyalo. Volume and fundamental group of a manifold of nonpositive curvature. Mat. Sb. (N.S.), 122(164)(2):142–156, 1983.
- [Buy90a] S. Buyalo. Collapsing manifolds of non-positive curvature I. Leningrad Math. J., 1, No. 5:1135?1155, 1990.
- [Buy90b] S. Buyalo. Collapsing manifolds of non-positive curvature II. Leningrad Math. J., 1, No. 6:1371?1399, 1990.
- [Cav21] N. Cavallucci. Topological entropy of the geodesic flow of non-positively curved metric spaces. arXiv preprint arXiv:2105.11774, 2021.
- [Cav22a] N. Cavallucci. Continuity of critical exponent of quasiconvex-cocompact groups under Gromov–Hausdorff convergence. Ergodic Theory and Dynamical Systems, pages 1–33, 2022.
- [Cav22b] N. Cavallucci. Entropies of non-positively curved metric spaces. Geometriae Dedicata, 216(5):54, 2022.
- [CCR01] J. Cao, J. Cheeger, and X. Rong. Splittings and cr-structures for manifolds with nonpositive sectional curvature. Inventiones mathematicae, 144(1):139–167, 2001.
- [CFG92] J. Cheeger, K. Fukaya, and M. Gromov. Nilpotent structures and invariant metrics on collapsed manifolds. J. Amer. Math. Soc., 5(2):327–372, 1992.
- [CG86] J. Cheeger and M. Gromov. Collapsing Riemannian manifolds while keeping their curvature bounded I. J. Diff. Geom., 23:309–364, 1986.
- [CG90] J. Cheeger and M. Gromov. Collapsing Riemannian manifolds while keeping their curvature bounded. II. J. Differential Geom., 32(1):269–298, 1990.
- [Che70] J. Cheeger. Finiteness theorems for Riemannian manifolds. Amer. J. Math., 92:61–74, 1970.
- [CM09a] P.E. Caprace and N. Monod. Isometry groups of non-positively curved spaces: discrete subgroups. J. Topol., 2(4):701–746, 2009.
- [CM09b] P.E. Caprace and N. Monod. Isometry groups of non-positively curved spaces: structure theory. Journal of topology, 2(4):661–700, 2009.
- [CM19] P.E. Caprace and N. Monod. Erratum and addenda to "Isometry groups of non-positively curved spaces: discrete subgroups". arXiv preprint arXiv:1908.10216, 2019.
- [CS20] N. Cavallucci and A. Sambusetti. Discrete groups of packed, non-positively curved, Gromov hyperbolic metric spaces. arXiv preprint arXiv:2102.09829, 2020.
- [CS21] N. Cavallucci and A. Sambusetti. Packing and doubling in metric spaces with curvature bounded above. Mathematische Zeitschrift, pages 1–46, 2021.
- [CS22] N. Cavallucci and A. Sambusetti. Thin actions on CAT(0) spaces. arXiv preprint arXiv:2210.01085, 2022.
- [CS23] N. Cavallucci and A. Sambusetti. Convergence and collapsing of CAT-group actions. arXiv preprint arXiv:2304.10763, 2023.
- [DLHG90] P. De La Harpe and E. Ghys. Espaces métriques hyperboliques. Sur les groupes hyperboliques d’après Mikhael Gromov, pages 27–45, 1990.
- [DY05] F. Dahmani and A. Yaman. Bounded geometry in relatively hyperbolic groups. New York J. Math., 11:89–95, 2005.
- [Ebe88] P. Eberlein. Symmetry diffeomorphism group of a manifold of nonpositive curvature. Trans. Amer. Math. Soc., 309(1):355–374, 1988.
- [Far81] D.R. Farkas. Crystallographic groups and their mathematics. Rocky Mountain J. Math., 11(4):511–551, 1981.
- [FL08] T. Foertsch and A. Lytchak. The de Rham decomposition theorem for metric spaces. Geom. Funct. Anal., 18:120–143, 2008.
- [Fuk86] K. Fukaya. Theory of convergence for Riemannian orbifolds. Japanese journal of mathematics. New series, 12(1):121–160, 1986.
- [Fuk87] K. Fukaya. Collapsing Riemannian manifolds to ones of lower dimensions. J. Differential Geom., 25(1):139–156, 1987.
- [Fuk88] K. Fukaya. A boundary of the set of the Riemannian manifolds with bounded curvatures and diameters. J. Differential Geom., 28(1):1–21, 1988.
- [Gro78] M. Gromov. Manifolds of negative curvature. Journal of differential geometry, 13(2):223–230, 1978.
- [Gro81] M. Gromov. Groups of polynomial growth and expanding maps (with an appendix by Jacques Tits). Publications Mathématiques de l’IHÉS, 53:53–78, 1981.
- [Gro07] M. Gromov. Metric structures for Riemannian and non-Riemannian spaces. Transl. from the French by Sean Michael Bates. With appendices by M. Katz, P. Pansu, and S. Semmes. Edited by J. LaFontaine and P. Pansu. 3rd printing. Basel: Birkhäuser, 2007.
- [Gut10] L. Guth. Metaphors in systolic geometry. In Proceedings of the International Congress of Mathematicians. Volume II, pages 745–768. Hindustan Book Agency, New Delhi, 2010.
- [Kle99] B. Kleiner. The local structure of length spaces with curvature bounded above. Mathematische Zeitschrift, 231, 01 1999.
- [LN19] A. Lytchak and K. Nagano. Geodesically complete spaces with an upper curvature bound. Geometric and Functional Analysis, 29(1):295–342, Feb 2019.
- [Pet84] S. Peters. Cheeger’s finiteness theorem for diffeomorphism classes of Riemannian manifolds. J. Reine Angew. Math., 349:77–82, 1984.
- [PT99] A. Petrunin and W. Tuschmann. Diffeomorphism finiteness, positive pinching, and second homotopy. Geom. Funct. Anal., 9(4):736–774, 1999.
- [Sab20] S. Sabourau. Macroscopic scalar curvature and local collapsing. arXiv preprint arXiv:2006.00663, 2020.
- [Sam99] A. Sambusetti. Minimal entropy and simplicial volume. Manuscripta Mathematica, 99(4):541–560, 1999.
- [Ser03] J.P. Serre. Trees. Springer Monographs in Mathematics. Springer-Verlag, Berlin, 2003. Translated from the French original by John Stillwell, Corrected 2nd printing of the 1980 English translation.
- [Stu06] Karl-Theodor Sturm. On the geometry of metric measure spaces. II. Acta Math., 196(1):133–177, 2006.
- [Wei67] A. Weinstein. On the homotopy type of positively-pinched manifolds. Arch. Math. (Basel), 18:523–524, 1967.
- [Yam85] T. Yamaguchi. On the number of diffeomorphism classes in a certain class of Riemannian manifolds. Nagoya Mathematical Journal, 97(none):173 – 192, 1985.
- [Yam90] M. Yamasaki. Maps between orbifolds. Proc. Amer. Math. Soc., 109(1):223–232, 1990.