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

Finiteness of CAT(0)(0) group actions

Nicola Cavallucci  and  Andrea Sambusetti
Abstract.

We prove some finiteness results for discrete isometry groups Γ\Gamma of uniformly packed CAT(0)(0)-spaces XX with uniformly bounded codiameter (up to group isomorphism), and for CAT(0)(0)-orbispaces M=Γ\XM=\Gamma\backslash X (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 Γ\Gamma 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(0)(0)-space admits a discrete, cocompact group of isometries with sufficiently small systole then it necessarily splits a non-trivial Euclidean factor.

N. Cavallucci has been partially supported by the SFB/TRR 191, funded by the DFG
A. Sambusetti is member of GNSAGA and acknowledges the support of INdAM during the preparation of this work.

1. Introduction

This is the first of two papers devoted to the theory of convergence for groups Γ\Gamma acting geometrically (that is discretely, by isometries and with compact quotient) on CAT(0)(0)-spaces XX. In this one, we will mainly focus on finiteness and splitting results. In our work, all CAT(0)(0)-spaces XX 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 μX\mu_{X}, etc. (see [Kle99], [LN19] and Section 2 for fundamentals on CAT(0)(0)-spaces).

The following is the first main finding of this paper:

Theorem A (Finiteness).


Given P0,r0,D0>0P_{0},r_{0},D_{0}>0, there exist only finitely many groups Γ\Gamma acting discretely and nonsingularly by isometries on some proper, geodesically complete, (P0,r0)(P_{0},r_{0})-packed, CAT(0)\textup{CAT}(0)-space with quotient of diameter at most D0D_{0}, up to isomorphism of abstract groups.
(We stress the fact that, in the above theorem, the CAT(0)(0)-spaces on which the groups Γ\Gamma 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 Γ\Gamma acts nonsingularly on XX if there exists at least one point xXx\in X such that StabΓ(x)\text{Stab}_{\Gamma}(x) is trivial (in particular, the action of Γ\Gamma on XX 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 XX 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 Γj\Gamma_{j} acting (singularly) on the same CAT(0)(0) space XX (a regular tree with bounded valency) with diam(Γj\X)D0(\Gamma_{j}\backslash X)\leq D_{0} for all jj; we thank P.E. Caprace for pointing out to us this example (see Remark 5.3). We will say that Γ\Gamma acts D0D_{0}-cocompactly on XX if diam(Γ\X)D0(\Gamma\backslash X)\leq D_{0}.

On the other hand, a metric space XX is (P0,r0)(P_{0},r_{0})-packed if all balls BX(x,3r0)B_{X}(x,3r_{0}) of radius 3r03r_{0} in XX contain at most P0P_{0} points that are 2r02r_{0}-separated. This condition should be thought as weak form of lower curvature bound at scale r0r_{0}; 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(k,n)(k,n) condition for metric measure spaces. Indeed, the Bishop-Gromov’s comparison theorem for Riemannian nn-manifolds with RicX(n1)k{}_{X}\geq-(n-1)k (or its generalization to CD(k,n)(k,n) spaces) yields a doubling condition

(1) μ(BX(x,2r))μ(BX(x,r))C(n,k,R)\frac{\mu(B_{X}(x,2r))}{\mu(B_{X}(x,r))}\leq C(n,k,R)\hskip 14.22636pt

for the measure of all rr-balls with rRr\leq R (for the Riemannian measure in case of manifolds, and for the reference measure in case of CD(k,n)(k,n)-spaces, cp. for instance [Stu06]), from which the packing condition at scale r0r/4r_{0}\leq r/4 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(0)(0)-space XX is also strictly weaker than the doubling condition (1) for the natural measure μX\mu_{X} since a local doubling condition implies that the space has pure dimension (i.e. the tangent cones at all points xx have the same geometric dimension nn, 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 XX with a compact quotient is (P0,r0)(P_{0},r_{0})-packed for suitable constants P0,r0P_{0},r_{0} (cp. [Cav22a, Proof of Lemma 5.4]). Namely, let CAT0(D0)\textup{CAT}_{0}(D_{0}) be the class of all isometric actions ΓX\Gamma\curvearrowright X where XX is a proper, geodesically complete, CAT(0)(0)-space, and Γ\Gamma is a D0D_{0}-cocompact, discrete subgroup of Isom(X)\text{Isom}(X); then the (P0,r0)(P_{0},r_{0})-packing conditions define a filtration

CAT0(D0)=P0,r0CAT0(P0,r0,D0)\textup{CAT}_{0}(D_{0})=\bigcup_{P_{0},r_{0}}\textup{CAT}_{0}(P_{0},r_{0},D_{0})

where CAT0(P0,r0,D0)\textup{CAT}_{0}(P_{0},r_{0},D_{0}) is the subset of CAT0(D0)\textup{CAT}_{0}(D_{0}) made of the actions ΓX\Gamma\curvearrowright X such that XX is, moreover, (P0,r0)(P_{0},r_{0})-packed.

As proved in [CS21], for geodesically complete CAT(0)(0)-spaces, a packing condition at some scale r0r_{0} is equivalent to a uniform upper bound of the canonical measure of all rr-balls (a condition sometimes called macroscopic scalar curvature bounded below cp. [Gut10], [Sab20]). Namely, there exist functions v,V:(0,+)(0,+)v,V\colon(0,+\infty)\to(0,+\infty) depending only on P0,r0P_{0},r_{0} such that for all xXx\in X and R>0R>0 we have (cp. Proposition 2.3):

(2) v(R)μX(B(x,R))V(R).v(R)\leq\mu_{X}(B(x,R))\leq V(R).

Moreoever, for geodesically complete CAT(0)(0)-spaces the (P0,r0)(P_{0},r_{0})-packing condition yields an important generalization of the classical Margulis’ Lemma, due to Breuillard-Green-Tao: there exists a constant ε0(P0,r0)>0\varepsilon_{0}(P_{0},r_{0})\!>0 such that for every discrete group of isometries Γ\Gamma of XX the ε0\varepsilon_{0}-almost stabilizer Γε0(x)\Gamma_{\varepsilon_{0}}(x) of any point xx is virtually nilpotent (cp. [BGT11] and Section 2.2 for details). We call this ε0=ε0(P0,r0)\varepsilon_{0}=\varepsilon_{0}(P_{0},r_{0}) 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 (X,Γ)(X,\Gamma) (see Section 5.2 for the precise definition), and then deduce from Theorem A corresponding finiteness results for the quotient spaces M=Γ\XM=\Gamma\backslash X. We will call such spaces CAT(0)(0)-orbispaces 111When restricting our attention to groups Γ\Gamma acting rigidly on XX (i.e. such that every gΓg\in\Gamma 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(0)(0) universal covering in the sense of orbispaces). However, our results apply to all nonsingular CAT(0)(0)-orbispaces., using this term as a shortening for quotient of a (proper, geodesically complete) CAT(0)(0)-space XX by a discrete, isometry group Γ\Gamma; we will also say that the CAT(0)(0)-orbispace MM is nonsingular if Γ\Gamma acts nonsingularly on XX. Notice that if Γ\Gamma is torsion-free then MM is a locally CAT(0)(0)-space.
For the following, let us define the class of all CAT(0)(0)-orbispaces M=Γ\XM=\Gamma\backslash X

𝒪-CAT0(P0,r0,D0){\mathcal{O}}\textup{-CAT}_{0}(P_{0},r_{0},D_{0})

where ΓX\Gamma\curvearrowright X is in CAT(P0,r0,D0)0{}_{0}(P_{0},r_{0},D_{0}), and the subclasses

-CAT0(P0,r0,D0)𝒩𝒮-CAT0(P0,r0,D0){\mathcal{L}}\textup{-CAT}_{0}(P_{0},r_{0},D_{0})\hskip 2.84526pt\subset\hskip 2.84526pt{\mathcal{NS}}\textup{-CAT}_{0}(P_{0},r_{0},D_{0})

of those which are, respectively, locally CAT(0)(0)-spaces and non-singular.
Then we obtain:

Corollary B.

There are finitely many spaces in the class -CAT0(P0,r0,D0){\mathcal{L}}\textup{-CAT}_{0}(P_{0},r_{0},D_{0}) up to homotopy equivalence, and finitely many spaces in 𝒩𝒮-CAT0(P0,r0,D0){\mathcal{NS}}\textup{-CAT}_{0}(P_{0},r_{0},D_{0}), up to equivariant homotopy equivalence.
(Two orbispaces M=Γ\XM=\Gamma\backslash X and M=Γ\XM^{\prime}=\Gamma^{\prime}\backslash X^{\prime} are equivariantly homotopy equivalent if there exists a group isomorphism φ:ΓΓ\varphi:\Gamma\rightarrow\Gamma^{\prime} and a φ\varphi-equivariant homotopy equivalence F:XXF:X\rightarrow X^{\prime}, that is such that F(gx)=φ(g)F(x)F(g\cdot x)=\varphi(g)\cdot F(x) for all gΓg\in\Gamma and all xXx\in X.)

A particular case of Theorem A, declined in the Riemannian setting, is:

Corollary C.

For every fixed n,κ,D0n,\kappa,D_{0}, there exist only finitely many compact nn-dimensional Riemannian orbifolds MM with curvature κ2k(M)0-\kappa^{2}\leq k(M)\leq 0 and diam(M)D0\textup{diam}(M)\leq D_{0}, up to equivariant diffeomorphisms.
(Two Riemannian orbifolds M=Γ\XM=\Gamma\backslash X, M=Γ\XM^{\prime}=\Gamma^{\prime}\backslash X^{\prime} are equivariantly diffeomorphic if there exists a diffeomorphism XXX\rightarrow X^{\prime} which is equivariant with respect to some isomorphism φ:ΓΓ\varphi:\Gamma\rightarrow\Gamma^{\prime}.)

In particular, for torsion-free orbifolds, this gives a new proof of the finiteness of compact Riemannian nn-manifolds MM with curvature κ2k(M)0-\kappa^{2}\leq k(M)\leq 0 and diam(M)D0(M)\leq D_{0}, 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 Γ\Gamma acting faithfully, discretely and D0D_{0}-cocompactly on δ0\delta_{0}-hyperbolic spaces with entropy H0\leq H_{0}. They achieve this by proving a positive, universal lower bound of the systole of Γ\Gamma, similar to the classical Heintze-Margulis’ Lemma (holding for manifolds with pinched, strictly negative curvature). Namely, for a discrete isometry group Γ\Gamma of XX let us call, respectively,

sys(Γ,x):=infgΓ{id}d(x,gx)\text{sys}(\Gamma,x):=\inf_{g\in\Gamma\setminus\{\text{id}\}}d(x,gx)
sys(Γ,X):=infxXsys(Γ,x)\text{sys}(\Gamma,X):=\inf_{x\in X}\text{sys}(\Gamma,x)

the systole of Γ\Gamma at xx and the (global) systole of Γ\Gamma (notice that the systole of the fundamental group Γ=π1(M)\Gamma=\pi_{1}(M) of a nonpositively curved manifold MM, acting on its Riemannian universal covering XX, precisely equals twice the injectivity radius of MM). Then, in [BCGS21], the authors prove that sys(Γ,X)\text{sys}(\Gamma,X) is bounded below by a positive constant c0(δ0,H0,D0)c_{0}(\delta_{0},H_{0},D_{0}) only depending on the hyperbolicity constant, the entropy of XX and on the diameter of Γ\X\Gamma\backslash X.This bound is, clearly, consequence of the Gromov hyperbolicity of the space XX, 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 MM is called ε\varepsilon-collapsed if the injectivity radius is smaller than ε\varepsilon 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 MM, the existence of a Riemannian metric with bounded sectional curvature sufficiently collapsed imposes strong restrictions to its topology. Namely, there exists an ε0(n)>0\varepsilon_{0}(n)>0 such that if a Riemannian nn-manifold with |KM|1|K_{M}|\leq 1 is ε\varepsilon-collapsed with ε<ε0(n)\varepsilon<\varepsilon_{0}(n), then MM admits a so-called FF-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 55, and Cao-Cheeger-Rong [CCR01] later, in any dimension, studied the possibility of collapsing compact, nn-dimensional manifolds MM with bounded sectional curvature κ2KM0-\kappa^{2}\leq K_{M}\leq 0. They proved that either the injectivity radius at some point is bounded below by a universal positive constant i0(n,κ)>0i_{0}(n,\kappa)>0, or MM admits a so-called abelian local splitting structure: this is, roughly speaking, a decomposition of the universal covering XX 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 Σ1,Σ2\Sigma_{1},\Sigma_{2} of the same hyperbolic surface with connected, geodesic boundary of length 1n\frac{1}{n}, then take the products Σi=Σi×S1\Sigma^{\prime}_{i}=\Sigma_{i}\times S^{1} with a circle of length 1n\frac{1}{n}, and glue Σ1\Sigma^{\prime}_{1} to Σ2\Sigma^{\prime}_{2} by means of an isometry φ\varphi of the boundaries Σi=Σi×S1\partial\Sigma^{\prime}_{i}=\partial\Sigma_{i}\times S^{1} which interchanges the circles Σi\partial\Sigma_{i} with S1S^{1}. This yields a nonpositively curved 33-manifold MnM_{n} (a graph manifold) with sectional curvature 1KMn0-1\leq K_{M_{n}}\leq 0, whose injectivity radius at every point is arbitrarily small provided that n0n\gg 0.

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 Γ\Gamma at xx and the (global) free systole of Γ\Gamma respectively as

sys(Γ,x)=infgΓd(x,gx)\text{sys}^{\diamond}(\Gamma,x)=\inf_{g\in\Gamma^{\diamond}}d(x,gx)
sys(Γ,X)=infxXsys(Γ,x)\text{sys}^{\diamond}(\Gamma,X)=\inf_{x\in X}\text{sys}^{\diamond}(\Gamma,x)

where Γ\Gamma^{\diamond} denotes the subset of torsion-free elements of Γ\Gamma.
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 D0D_{0}-cocompact action of a group Γ\Gamma on a CAT(0)(0)-space XX (or, equivalently, the quotient space M=Γ\XM=\Gamma\backslash X) is ε\varepsilon-collapsed if sys(Γ,X)ε\textup{sys}^{\diamond}(\Gamma,X)\leq\varepsilon.

The following theorem, which is the key to our finiteness theorems, shows that if XX admits a discrete, D0D_{0}-cocompact action which is ε\varepsilon-collapsed for sufficiently small ε\varepsilon, then XX necessarily splits a non-trivial Euclidean factor:

Theorem D (Splitting of an Euclidean factor under ε\varepsilon-collapsed actions).


Let P0,r0,D0>0P_{0},r_{0},D_{0}>0. There exists σ0=σ0(P0,r0,D0)>0\sigma_{0}=\sigma_{0}(P_{0},r_{0},D_{0})>0 such that if XX is a proper, geodesically complete, (P0,r0)(P_{0},r_{0})-packed, CAT(0)\textup{CAT}(0)-space admitting a discrete, D0D_{0}-cocompact group of isometries Γ\Gamma with sys(Γ,X)σ0\textup{sys}^{\diamond}(\Gamma,X)\leq\sigma_{0} then XX splits isometrically as Y×kY\times\mathbb{R}^{k}, with k1k\geq 1.

Notice that, as the prototypical example above shows already for manifolds, the splitting of XX does not hold without an a priori bound on the diameter.

As proved by Buyalo in [Buy83], when M=Γ\XM=\!\Gamma\!\backslash X is a manifold with pinched, nonpositive sectional curvature 1kM0-1\leq k_{M}\leq\!0 and diameter bounded by DD, then there exists a positive constant ε(n,D)\varepsilon(n,D) such that the condition of being ε\varepsilon-collapsed for ε<ε(n,D)\varepsilon<\varepsilon(n,D) implies the existence of a normal, free abelian subgroup AA of rank k1k\geq 1; then, Γ\Gamma virtually splits AA, and the existence of a non-trivial Euclidean factor for XX 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 gΓg\in\Gamma (as opposite to [Fuk86]): actually, bounding the order of the torsion elements of Γ\Gamma 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 AA^{\prime} of Γ\Gamma, from which we construct a Γ\Gamma-invariant closed convex subset X0XX_{0}\subset X which splits as Y×kY\times{\mathbb{R}}^{k}, and then we use the minimality of the action to deduce that X0=XX_{0}=X (see Section 4 for the proof). Actually, in our metric setting it can happen that Γ\Gamma does not have any non-trivial, normal, free abelian subgroup at all (see [CS23]).

The final step for Theorem A is realizing that the ε\varepsilon-collapsing of an action of Γ\Gamma on XX can occur on different subspaces of XX at different scales; a careful analysis of this phenomenon allows us to renormalize, in a precise sense, the metric of XX, keeping the diameter of the quotient Γ\X\Gamma\backslash X 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 P0,r0,D0P_{0},r_{0},D_{0}, there exist s0=s0(P0,r0,D0)>0s_{0}=s_{0}(P_{0},r_{0},D_{0})>0 and Δ0=Δ0(P0,D0)\Delta_{0}=\Delta_{0}(P_{0},D_{0}) such that the following holds. Let Γ\Gamma be a discrete and D0D_{0}-cocompact isometry group of a proper, geodesically complete, (P0,r0)(P_{0},r_{0})-packed, CAT(0)\textup{CAT}(0)-space XX: then, Γ\Gamma admits also a faithful, discrete, Δ0\Delta_{0}-cocompact action by isometries on a CAT(0)\textup{CAT}(0)-space XX^{\prime} isometric to XX, such that sys(Γ,X)s0\textup{sys}^{\diamond}(\Gamma,X^{\prime})\geq s_{0}.
Moreover, the action of Γ\Gamma on XX is nonsingular if and only if the action on XX^{\prime} is nonsingular.

This theorem allows us to deduce the finiteness of nonsinngular groups Γ\Gamma in the class CAT(P0,r0,D0)0{}_{0}(P_{0},r_{0},D_{0}), 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 Δ0=2n0(D0+n0)\Delta_{0}=2^{n_{0}}(D_{0}+\sqrt{n_{0}}), where n0=P0/2n_{0}=P_{0}/2 bounds the dimension of every XX in CAT(P0,r0,D0)0{}_{0}(P_{0},r_{0},D_{0}) (see Proposition 2.3). Also the constant s0s_{0} could be made explicit in terms of P0,R0,D0P_{0},R_{0},D_{0} and of the Margulis’ constant ε0=ε0(P0,r0)\varepsilon_{0}=\varepsilon_{0}(P_{0},r_{0}), following the proof of Theorem E and using the function σ0(P0,r0,D0)\sigma_{0}(P_{0},r_{0},D_{0}) appearing in Theorem D (itself explicitable in terms of P0,R0,D0P_{0},R_{0},D_{0} and of the universal bound of the index of the lattice of translations (G){\mathcal{L}}(G) in any crystallographic group GG of dimension n0\leq n_{0}). The only quantity which is not explicit is the Margulis constant ε0(P0,r0)\varepsilon_{0}(P_{0},r_{0}) 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 Γ\Gamma under consideration. We stress that the existence of such a bound is new also for isometry groups Γ\Gamma 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 M=Γ\XM=\Gamma\backslash X. More explicitly, there exists a constant b0=b0(P0,r0)>0b_{0}=b_{0}(P_{0},r_{0})>0 (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 P0,r0,D0>0P_{0},r_{0},D_{0}\!>0. Then for every nonsingular ΓX\Gamma\curvearrowright X in CAT(P0,r0,D0)0{}_{0}(P_{0},r_{0},D_{0}), every finite subgroup F<ΓF<\Gamma has order

|F|V(Δ0)v(12b0s0)|F|\leq\frac{V(\Delta_{0})}{v\left(\frac{1}{2}b_{0}s_{0}\right)}

(where v,Vv,V are the universal functions appearing in (2), and Δ0,s0\Delta_{0},s_{0} are the constants of Theorem E).

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 P0,r0P_{0},r_{0} and, possibly, on D0D_{0}. For instance, in the above theorems, we used the constants σ0,s0,Δ0,b0\sigma_{0},s_{0},\Delta_{0},b_{0} with this meaning.

2. Preliminaries on CAT(0)(0)-spaces

We fix here some notation and recall some facts about CAT(0)(0)-spaces.
Throughout the paper XX will be a proper metric space with distance dd. The open (resp. closed) ball in XX of radius rr, centered at xx, will be denoted by BX(x,r)B_{X}(x,r) (resp. B¯X(x,r)\overline{B}_{X}(x,r)); we will often drop the subscript XX when the space is clear from the context.
A geodesic in a metric space XX is an isometry c:[a,b]Xc\colon[a,b]\to X, where [a,b][a,b] is an interval of \mathbb{R}. The endpoints of the geodesic cc are the points c(a)c(a) and c(b)c(b); a geodesic with endpoints x,yXx,y\in X is also denoted by [x,y][x,y]. A geodesic ray is an isometry c:[0,+)Xc\colon[0,+\infty)\to X and a geodesic line is an isometry c:Xc\colon\mathbb{R}\to X. A metric space XX is called geodesic if for every two points x,yXx,y\in X there is a geodesic with endpoints xx and yy.

A metric space XX is called CAT(0)(0) if it is geodesic and every geodesic triangle Δ(x,y,z)\Delta(x,y,z) is thinner than its Euclidean comparison triangle Δ¯(x¯,y¯,z¯)\overline{\Delta}(\bar{x},\bar{y},\bar{z}): that is, for any couple of points p[x,y]p\in[x,y] and q[x,z]q\in[x,z] we have d(p,q)d(p¯,q¯)d(p,q)\leq d(\bar{p},\bar{q}) where p¯,q¯\bar{p},\bar{q} are the corresponding points in Δ¯(x¯,y¯,z¯)\overline{\Delta}(\bar{x},\bar{y},\bar{z}) (see for instance [BH13] for the basics of CAT(0)-geometry). As a consequence, every CAT(0)(0)-space is uniquely geodesic: for every two points x,yx,y there exists a unique geodesic with endpoints xx and yy.

A CAT(0)(0)-metric space XX is geodesically complete if evvery geodesic c:[a,b]Xc\colon[a,b]\to X can be extended to a geodesic line. For instance, if a CAT(0)(0)-space is a homology manifold then it is always geodesically complete, see [BH13, Proposition II.5.12]).

The boundary at infinity of a CAT(0)(0)-space XX (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 X\partial X, see [BH13, Chapter II.9].

A subset CC of XX is said to be convex if for all x,yCx,y\in C the geodesic [x,y][x,y] is contained in CC. Given a subset YXY\subseteq X we denote by Conv(Y)\text{Conv}(Y) the convex closure of YY, that is the smallest closed convex subset containing YY.If CC is a convex subset of a CAT(0)(0)-space XX then it is itself CAT(0)(0), and its boundary at infinity C\partial C naturally and isometrically embeds in X\partial X.

We will denote by HD(X)(X) and TD(X)(X) the Hausdorff and the topological dimension of a metric space XX, respectively. By [LN19] we know that if XX is a proper and geodesically complete CAT(0)(0)-space then every point xXx\in X has a well defined integer dimension in the following sense: there exists nxn_{x}\in\mathbb{N} such that every small enough ball around xx has Hausdorff dimension equal to nxn_{x}. This defines a stratification of XX into pieces of different integer dimensions: namely, if XkX^{k} denotes the subset of points of XX with dimension kk, then

X=kXk.X=\bigcup_{k\in\mathbb{N}}X^{k}.

The dimension of XX is the supremum of the dimensions of its points: it coincides with the topological dimension of XX, cp. [LN19, Theorem 1.1].

Calling k\mathcal{H}^{k} the kk-dimensional Hausdorff measure, the formula

μX:=kkXk\mu_{X}:=\sum_{k\in\mathbb{N}}\mathcal{H}^{k}\llcorner X^{k}

defines a canonical measure on XX which is locally positive and locally finite.

2.1. Discrete isometry groups

Let Isom(X)\text{Isom}(X) be the group of isometries of XX, endowed with the compact-open topology: as XX is proper, it is a topological, locally compact group.
The translation length of gIsom(X)g\in\text{Isom}(X) is by definition (g):=infxXd(x,gx).\ell(g):=\inf_{x\in X}d(x,gx). When the infimum is realized, the isometry gg is called elliptic if (g)=0\ell(g)=0 and hyperbolic otherwise. The minimal set of gg, Min(g)\text{Min}(g), is defined as the subset of points of XX where gg realizes its translation length; notice that if gg is elliptic then Min(g)\text{Min}(g) is the subset of points fixed by gg. An isometry is called semisimple if it is either elliptic or hyperbolic; a subgroup Γ\Gamma of Isom(X)(X) is called semisimple if all of its elements are semisimple.
Let Γ\Gamma be a subgroup of Isom(X)(X). For xXx\in X and r0r\geq 0 we set

(3) Σ¯r(x,X):={gΓ s.t. d(x,gx)r}\overline{\Sigma}_{r}(x,X):=\{g\in\Gamma\text{ s.t. }d(x,gx)\leq r\}
(4) Γ¯r(x,X):=Σ¯r(x,X)\overline{\Gamma}_{r}(x,X):=\langle\overline{\Sigma}_{r}(x,X)\rangle

When the context is clear we will simply write Σ¯r(x)\overline{\Sigma}_{r}(x) and Γ¯r(x)\overline{\Gamma}_{r}(x).
The subgroup Γ\Gamma is discrete if it is discrete as a subset of Isom(X) (with respect to the compact-open topology). Since XX is assumed to be proper and Γ\Gamma acts by isometries, this condition is the same as asking that the orbit Γx\Gamma x is discrete and StabΓ(x)\text{Stab}_{\Gamma}(x) is finite for some (or, equivalently, for all) xXx\in X. This is in turn equivalent to asking that the sets Σ¯r(x)\overline{\Sigma}_{r}(x) are finite for all xXx\in X and all r0r\geq 0.
When dealing with isometry groups Γ\Gamma of CAT(0)(0)-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 Γ\Gamma of Isom(X)(X) will be called rigid (or slim, with the terminology used in [CS22]) if for all gΓg\in\Gamma the subset Fix(g)\text{Fix}(g) has empty interior. Every torsion-free group is trivially rigid, as well as any discrete group acting on a CAT(0)(0)-homology manifold, as proved in [CS22, Lemma 2.1].
A CAT(0)(0)-orbispace (in the sense of [Fuk86]) is the quotient M=Γ\XM=\Gamma\backslash X of a (proper, geodesically complete) CAT(0)(0)-space XX by a discrete isometry group Γ\Gamma. One might define the notion of orbispace MM in terms of an orbifold atlas, that is a collection of uniformizing charts πi:ViUiM\pi_{i}:V_{i}\rightarrow U_{i}\subset M covering MM (where ViV_{i} is a locally compact space endowed with the action of a finite group Γi\Gamma_{i} such that πi\pi_{i} induces a homeomorphism Γi\ViUi\Gamma_{i}\backslash V_{i}\simeq U_{i}) 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 Γi\Gamma_{i} on ViV_{i} are all supposed to be rigid. Every quotient of a CAT(0)(0)-space XX by a discrete and rigid group Γ\Gamma has a structure of rigid orbispace; reciprocally, every rigid orbispace MM with nonpositive curvature (that is, such that the domains ViV_{i} of the uniformizing charts are locally CAT(0)(0)-spaces) is developable, which means that M=Γ\XM=\Gamma\backslash X, for some rigid, discrete isometry group Γ\Gamma acting on a suitable CAT(0)(0)-space XX, see [DLHG90, Ch.11, Théorème 8].
A group Γ<Isom(X)\Gamma<\textup{Isom}(X) is said to be cocompact if the quotient metric space Γ\X\Gamma\backslash X is compact; in this case, we call codiameter of Γ\Gamma the diameter of the quotient, and we will say that Γ\Gamma is D0D_{0}-cocompact if it has codiameter at most D0D_{0}. Notice that the codiameter of Γ\Gamma coincides with

(5) inf{r>0 s.t. ΓB¯(x,r)=XxX}.\inf\{r>0\text{ s.t. }\Gamma\cdot\overline{B}(x,r)=X\,\,\,\,\forall x\in X\}.

It is well-known, and we will consistently use it in the paper, that if XX is geodesic and Γ<Isom(X)\Gamma<\textup{Isom}(X) is discrete and D0D_{0}-cocompact, then for all DD0D\geq D_{0} the subset Σ¯2D(x)\overline{\Sigma}_{2D}(x) is a generating set for Γ\Gamma, that is Γ¯2D(x)=Γ\overline{\Gamma}_{2D}(x)=\Gamma, for every xXx\in X; we call this a 2D2D-short generating set of Γ\Gamma at xx.
Moreover, if XX is simply connected, then Γ\Gamma admits a finite presentation as

Γ=Σ¯2D(x)|2D(x)\Gamma=\langle\overline{\Sigma}_{2D}(x)\hskip 2.84526pt|\hskip 2.84526pt{\mathcal{R}}_{2D}(x)\rangle

where 2D(x){\mathcal{R}}_{2D}(x) is a subset of words of length 33 on Σ¯2D(x)\overline{\Sigma}_{2D}(x), see [Ser03, App., Ch.3]. For later use we also record the following general fact.

Lemma 2.1.

Let XX be a geodesic metric space and let Γ<Isom(X)\Gamma<\textup{Isom}(X) be a discrete, D0D_{0}-cocompact group. If Γ\Gamma^{\prime} is normal subgroup of Γ\Gamma with index [Γ:Γ]J[\Gamma:\Gamma^{\prime}]\leq J, then Γ\Gamma^{\prime} is at most 2D0(J+1)2D_{0}(J+1)-cocompact.

Proof.

As Γ\Gamma^{\prime} is discrete, the space Y=Γ\XY=\Gamma^{\prime}\backslash X is geodesic and the group Γ\Γ\Gamma^{\prime}\backslash\Gamma acts by isometries on it with codiameter at most D0D_{0}. Take arbitrary y,yYy,y^{\prime}\in Y and connect them by a geodesic cc. Let tk=k(2D0+ε)t_{k}=k\cdot(2D_{0}+\varepsilon), as far as c(tk)c(t_{k}) is defined for integers kk, and let gkΓ\Γg_{k}\in\Gamma^{\prime}\backslash\Gamma be such that d(c(tk),gky)D0d(c(t_{k}),g_{k}y)\leq D_{0}. By construction, the points {gky}\{g_{k}y\} are all distinct since they are all ε\varepsilon-separated, so the gkg_{k}’s are distinct. Since Γ\Γ\Gamma^{\prime}\backslash\Gamma has cardinality at most JJ, we conclude that d(y,y)(J+1)(2D0+ε)d(y,y^{\prime})\leq(J+1)\cdot(2D_{0}+\varepsilon). By the arbitrariness of ε\varepsilon, yy and yy^{\prime} we deduce that the diameter of YY is at most 2D0(J+1)2D_{0}(J+1). ∎

In general the full isometry group Isom(X)(X) of a CAT(0)(0)-space is not a Lie group, for instance in the case of regular trees. When a CAT(0)(0)-space XX admits a cocompact, discrete group of isometries Γ\Gamma then Isom(X)(X) 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 XX be a proper, geodesically complete, CAT(0)\textup{CAT}(0)-space, admitting a discrete, cocompact group of isometries. Then XX splits isometrically as M×n×NM\times\mathbb{R}^{n}\times N, where MM is a symmetric space of noncompact type and 𝒟:=Isom(N){\mathcal{D}}:=\textup{Isom}(N) is totally disconnected. Moreover

Isom(X)𝒮×n×𝒟\textup{Isom}(X)\cong{\mathcal{S}}\times{\mathcal{E}}_{n}\times{\mathcal{D}}

where 𝒮{\mathcal{S}} is a semi-simple Lie group with trivial center and without compact factors and nIsom(n){\mathcal{E}}_{n}\cong\textup{Isom}(\mathbb{R}^{n}).

2.2. The packing condition and Margulis’ Lemma

Let XX be a metric space and r>0r>0. A subset YY of XX is called rr-separated if d(y,y)>rd(y,y^{\prime})>r for all y,yYy,y^{\prime}\in Y. Given xXx\in X and 0<rR0<r\leq R we denote by Pack(B¯(x,R),r)(\overline{B}(x,R),r) the maximal cardinality of a 2r2r-separated subset of B¯(x,R)\overline{B}(x,R). Moreover we denote by Pack(R,r)(R,r) the supremum of Pack(B¯(x,R),r)(\overline{B}(x,R),r) among all points of XX. Given P0,r0>0P_{0},r_{0}>0 we say that XX is P0P_{0}-packed at scale r0r_{0} (or (P0,r0)(P_{0},r_{0})-packed, for short) if Pack(3r0,r0)P0(3r_{0},r_{0})\leq P_{0}. We will simply say that XX is packed if it is P0P_{0}-packed at scale r0r_{0} for some P0,r0>0P_{0},r_{0}>0.
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 nn-dimensional Riemannian manifold a lower bound on the Ricci curvature RicX(n1)κ2\text{Ric}_{X}\geq-(n-1)\kappa^{2} implies a uniform estimate of the packing function at any fixed scale r0r_{0}, that is

(6) Pack(3r0,r0)vκn(3r0)vκn(r0)\text{Pack}(3r_{0},r_{0})\leq\frac{v_{{\mathbb{H}}^{n}_{\kappa}}(3r_{0})}{v_{{\mathbb{H}}^{n}_{\kappa}}(r_{0})}

where vκn(r)v_{{\mathbb{H}}^{n}_{\kappa}}(r) is the volume of a ball of radius rr in the nn-dimensional space form with constant curvature κ2-\kappa^{2}.
Also remark that every metric space admitting a cocompact action is packed (for some P0,r0P_{0},r_{0}), see the proof of [Cav22a, Lemma 5.4].

The packing condition has many interesting geometric consequences for complete, geodesically complete CAT(0)(0)-spaces, as showed in [CS20], [Cav21] and [Cav22b]. The first one is a uniform estimate of the measure of balls of any radius RR 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 XX be a complete, geodesically complete, (P0,r0)(P_{0},r_{0})-packed, CAT(0)\textup{CAT}(0)-space. Then XX is proper and

  • (i)

    Pack(R,r)P0(1+P0)Rmin{r,r0}1\textup{Pack}(R,r)\leq P_{0}(1+P_{0})^{\frac{R}{\min\{r,r_{0}\}}-1} for all 0<rR0<r\leq R;

  • (ii)

    the dimension of XX is at most n0:=P0/2n_{0}:=P_{0}/2;

  • (iii)

    there exist functions v,V:(0,+)(0,+)v,V\colon(0,+\infty)\to(0,+\infty) depending only on P0,r0P_{0},r_{0} such that for all xXx\in X and R>0R>0 we have

    (7) v(R)μX(B¯(x,R))V(R);v(R)\leq\mu_{X}(\overline{B}(x,R))\leq V(R);
  • (iv)

    The entropy of XX is bounded above in terms of P0P_{0} and r0r_{0}, namely

    Ent(X):=lim supR+1RlogPack(R,1)log(1+P0)r0.\textup{Ent}(X):=\limsup_{R\to+\infty}\frac{1}{R}\log\textup{Pack}(R,1)\leq\frac{\log(1+P_{0})}{r_{0}}.

In particular, for a geodesically complete, CAT(0)(0) space XX which is (P0,r0)(P_{0},r_{0})-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 r0r_{0} yields an explicit, uniform control of the packing function at any other scale rr: 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 rr-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 Γ¯r(x)\overline{\Gamma}_{r}(x) for small rr, which are sometimes called the “almost stabilizers”. We decline it for geodesically complete CAT(0)(0)-spaces.

Proposition 2.4 ([BGT11, Corollary 11.17]).


Given P0,r0>0P_{0},r_{0}>0, there exists ε0=ε0(P0,r0)>0\varepsilon_{0}=\varepsilon_{0}(P_{0},r_{0})>0 such that the following holds. Let XX be a proper, geodesically complete, (P0,r0)(P_{0},r_{0})-packed, CAT(0)\textup{CAT}(0)-space and let Γ\Gamma be a discrete subgroup of Isom(X)\textup{Isom}(X): then, for every xXx\in X and every 0εε00\leq\varepsilon\leq\varepsilon_{0}, the almost stabilizer Γ¯ε(x)\overline{\Gamma}_{\varepsilon}(x) is virtually nilpotent.

We will often refer to the constant ε0=ε0(P0,r0)\varepsilon_{0}=\varepsilon_{0}(P_{0},r_{0}) as the Margulis’ constant. The conclusion of Proposition 2.4 can be improved for cocompact groups, as in this case the group Γ¯ε(x)\overline{\Gamma}_{\varepsilon}(x) is virtually abelian (cp. [BH13, Theorem II.7.8]; indeed, a cocompact group of a CAT(0)(0)-space is always semisimple).

2.3. Crystallographic groups in the Euclidean space

We will denote points in k\mathbb{R}^{k} by a bold letter v, and the origin by 𝐎\bf O.
Among CAT(0)(0)-spaces, the Euclidean space k{\mathbb{R}}^{k} and its discrete groups play a special role. A crystallographic group is a discrete, cocompact group GG of isometries of some k\mathbb{R}^{k}. 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 k\mathbb{R}^{k} (see for instance [Far81]); so, alternatively, a lattice \mathcal{L} can be seen as the set of linear combinations with integer coefficients of kk independent vectors 𝐛𝟏,,𝐛𝐤\bf b_{1},\ldots,\bf b_{k} (we will make no difference between a lattice and this representation). The integer kk is also called the rank of the lattice.
A basis ={𝐛𝟏,,𝐛𝐤}\mathcal{B}=\{\bf b_{1},\ldots,\bf b_{k}\} of a lattice \mathcal{L} is a set of kk independent vectors that generate \mathcal{L} as a group. There are many geometric invariants classically associated to a lattice \mathcal{L}, we will need just two of them:
– the covering radius, which is defined as

ρ()=inf{r>0 s.t. 𝐯B¯(𝐯,r)=k}\rho(\mathcal{L})=\inf\left\{r>0\text{ s.t. }\bigcup_{\bf v\in\mathcal{L}}\overline{B}({\bf v},r)=\mathbb{R}^{k}\right\}

– the shortest generating radius, that is

λ()=inf{r>0 s.t.  contains k independent vectors of lengthr}.\lambda(\mathcal{L})=\inf\{r>0\text{ s.t. }\mathcal{L}\text{ contains }k\text{ independent vectors of length}\leq r\}.

Notice that, by the triangle inequality, any lattice \mathcal{L} is 2ρ()2\rho(\mathcal{L})-cocompact. By definition it is always possible to find a basis ={𝐛𝟏,,𝐛𝐤}\mathcal{B}=\{\bf b_{1},\ldots,\bf b_{k}\} of \mathcal{L} such that 𝐛𝟏𝐛𝐤=λ()\|\bf b_{1}\|\leq\cdots\leq\|b_{k}\|=\lambda(\mathcal{L}); this is called a shortest basis of \mathcal{L}. The shortest generating radius and the covering radius are related as follows:

(8) ρ()k2λ().\rho(\mathcal{L})\leq\frac{\sqrt{k}}{2}\cdot\lambda(\mathcal{L}).

For our purposes, the content of the famous Bieberbach’s Theorems can be stated as follows.

Proposition 2.5 (Bieberbach’s Theorem).

There exists J(k)J(k), only depending on kk, such that the following holds true. For every crystallographic group GG of k\mathbb{R}^{k} the subgroup (G)=GTransl(k){\mathcal{L}}(G)=G\cap\textup{Transl}({\mathbb{R}}^{k}) is a normal subgroup of index at most J(k)J(k), in particular a lattice.

Here Transl(k)\textup{Transl}({\mathbb{R}}^{k}) denotes the normal subgroup of translations of k=Isom(k){\mathcal{E}}_{k}=\textup{Isom}(\mathbb{R}^{k}). The subgroup (G)\mathcal{L}(G) is called the maximal lattice of GG. It is well-known that every lattice <G{\mathcal{L}}<G of rank kk has finite index in GG.

2.4. Virtually abelian groups

Recall that the (abelian, or Prüfer) rank of an abelian group AA, denoted rk(A)\text{rk}(A), is the maximal cardinality of a subset SAS\subset A of \mathbb{Z}-linear independent elements. We extend this definition to virtually abelian groups GG, defining rk(G)(G) as the rank of every free abelian subgroup AA of finite index in GG: notice that if AA^{\prime} is a finite index subgroup of an abelian group AA, then AA and AA^{\prime} have same rank, so rk(G)(G) is well defined. One can equivalenty define rk(G)(G) as the rank of every normal, free abelian subgroup AA of finite index in GG, since every finite index subgroup of GG contains a normal, finite index subgroup. It is easy to show that the abelian rank is monotone on subgroups.

If AA is a discrete, finitely generated, semisimple free abelian group of isometries of a CAT(0)(0)-space XX, then its minimal set

Min(A):=aAMin(a)\text{Min}(A):=\bigcap_{a\in A}\text{Min}(a)

is not empty and splits isometrically as Z×kZ\times\mathbb{R}^{k} where k=rk(A)k=\textup{rk}(A). This is the main content of the Flat Torus Theorem (see [BH13, Theorem II.7.1]).
We recall some additional facts about the identification Min(A)=Z×k\text{Min}(A)=Z\times\mathbb{R}^{k}, which we will freely use later:

(a) the abelian group AA acts as the identity on the factor ZZ, and cocompactly by translations on the Euclidean factor k\mathbb{R}^{k};

(b) writing x=(z,v)Min(A)x=(z,v)\in\text{Min}(A), the slice {z}×k\{z\}\times\mathbb{R}^{k} coincides with the convex closure Conv(Ax)\textup{Conv}(Ax) of the orbit AxAx;

(c) one has Conv(Ax)=Conv(Ax)\textup{Conv}(A^{\prime}x)=\textup{Conv}(Ax) for every finite index subgroup A<AA^{\prime}<A and every xMin(A)x\in\textup{Min}(A).

The last assertion follows from the fact that Conv(Ax)Conv(Ax)\textup{Conv}(A^{\prime}x)\subseteq\textup{Conv}(Ax) and are both isometric to k\mathbb{R}^{k}, 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(0)(0)-spaces, that we will often use.

Lemma 2.6.

Let XX be a proper CAT(0)\textup{CAT}(0)-space and let G0<GG_{0}<G be discrete, semisimple, virtually abelian groups of isometries of XX. Then [G:G0][G:G_{0}] is finite if and only if GG and G0G_{0} have same rank.

Proof.

The implication [G:G0]<rk(G)=rk(G0)[G:G_{0}]<\infty\Rightarrow\text{rk}(G)=\text{rk}(G_{0}) is trivial, as every free abelian finite index subgroup A<G0A<G_{0} is also a finite index subgroup of GG. To show the converse implication, assume that rk(G0)=rk(G)=k\text{rk}(G_{0})=\text{rk}(G)=k, and let AA be a rank kk, free abelian, finite index normal subgroup of GG. Consider the (free abelian) subgroup A0=AG0A_{0}=A\cap G_{0} of G0G_{0}, and notice that we have rk(A0)=rk(G0)=k\text{rk}(A_{0})=\text{rk}(G_{0})=k, since also [G0:A0]=[G:A]<[G_{0}:A_{0}]=[G:A]<\infty. Now, both A0A_{0} and AA act faithfully on the Euclidean factor of Min(A)=Z×k\text{Min}(A)=Z\times{\mathbb{R}}^{k} (they do not contain elliptics since they are free, and act as the identity on ZZ). Therefore their projections p(A0)<p(A)p(A_{0})<p(A) on Isom(k)\text{Isom}({\mathbb{R}}^{k}) are both rank kk Euclidean lattices, hence [A:A0]=[p(A):p(A0)]<[A:A_{0}]=[p(A):p(A_{0})]<\infty. But then we deduce that [G:G0][G:A0]=[G:A][A:A0]<[G:G_{0}]\leq[G:A_{0}]=[G:A][A:A_{0}]<\infty. ∎

The following generalization of the Flat Torus Theorem is classical.

Proposition 2.7 ([BH13, Corollary II.7.2]).

Let XX be a proper CAT(0)\textup{CAT}(0)-space, and let GG be a discrete, semisimple, virtually abelian group of isometries of XX of rank kk. Then, there exists a closed, convex, GG-invariant subset C(G)C(G) of XX which splits as Z×kZ\times\mathbb{R}^{k}, satisfying the following properties:

  • (i)

    every gGg\in G preserves the product decomposition and acts as the identity on the first component;

  • (ii)

    the image GkG_{\mathbb{R}^{k}} of GG under the projection GIsom(k)G\rightarrow\textup{Isom}(\mathbb{R}^{k}) is a crystallographic group.

Given a generating set SS of GG, the following statement explains how to construct a generating set for the maximal lattice (Gk){\mathcal{L}}(G_{\mathbb{R}^{k}}) with words on SS of bounded length. This will be used later in the proof of Theorem 4.1:

Lemma 2.8.

Same assumptions as in Proposition 2.7 above.
If SS is a symmetric, finite generating set for GG containing the identity, then there exists a subset ΣS4J(k)+2\Sigma\subseteq S^{4J(k)+2} whose projection Σk\Sigma_{\mathbb{R}^{k}} on Isom(k)\textup{Isom}(\mathbb{R}^{k}) generates (Gk)\mathcal{L}(G_{\mathbb{R}^{k}}), where J(k)J(k) is the constant of Proposition 2.5.

Here SnGS^{n}\subset G denotes the subset of all products of at most nn elements of SS; by definition, this coincides with GB¯(e,n)G\cap\overline{B}(e,n), where B¯(e,n)\overline{B}(e,n) is the closed ball of radius nn, centered at the identity ee, in the Cayley graph Cay(G,S)\text{Cay}(G,S).

Proof of Lemma 2.8.

Call p:GGkp\colon G\to G_{\mathbb{R}^{k}} the projection on Isom(k)(\mathbb{R}^{k}). By Proposition 2.5 the normal subgroup (Gk)\mathcal{L}(G_{\mathbb{R}^{k}}) has index at most J(k)J(k) in GkG_{\mathbb{R}^{k}}, so the normal subgroup G:=p1((Gk))G^{\prime}:=p^{-1}(\mathcal{L}(G_{\mathbb{R}^{k}})) has index at most J(k)J(k) in GG. The group GG acts discretely by isometries on Cay(G,S)(G,S) with codiameter 11. So, by Lemma 2.1 the group GG^{\prime} acts on Cay(G,S)(G,S) with codiameter at most 2J(k)+12J(k)+1. As recalled before Lemma 2.1, GG^{\prime} is therefore generated by the subset Σ:=GS4J(k)+2\Sigma:=G^{\prime}\cap S^{4J(k)+2}, made of the elements of GG^{\prime} displacing the point ee in the Cayley graph Cay(G,S)(G,S) at most by 4J(k)+24J(k)+2. It follows that the projection Σk=p(Σ)\Sigma_{\mathbb{R}^{k}}=p(\Sigma) generates (Gk)\mathcal{L}(G_{\mathbb{R}^{k}}). ∎

Finally, remark that given a discrete, semisimple, virtually abelian group GG of isometries of a proper CAT(0)(0)-space, there can be several closed, convex, GG-invariant subsets C(G)C(G) satisfying the conclusions of Proposition 2.7. What is uniquely associated to GG is a subset in the boundary of XX.

Proposition 2.9.

Same assumptions as in Proposition 2.7.
Then there exists a closed, convex, GG-invariant subset of X\partial X, denoted G\partial G, which is isometric to 𝕊k1\mathbb{S}^{k-1} and has the following properties:

  • (i)

    G=Conv(Ax)\partial G=\partial\textup{Conv}(Ax), for every free abelian finite index subgroup AA of GG and every xMin(A)x\in\textup{Min}(A);

  • (ii)

    for every subgroup G<GG^{\prime}<G we have GG\partial G^{\prime}\subseteq\partial G; moreover G=G\partial G^{\prime}=\partial G if rk(G)=rk(G)\textup{rk}(G^{\prime})=\textup{rk}(G).

The closed subset G\partial G provided by this proposition will be called the trace at infinity of the virtually abelian group GG.

Proof.

We fix a free abelian, normal subgroup A0GA_{0}\triangleleft G with finite index and x0Min(A0)x_{0}\in\text{Min}(A_{0}). By the Flat Torus Theorem, Conv(A0x0)\text{Conv}(A_{0}x_{0}) is isometric to k\mathbb{R}^{k}. We set G:=Conv(A0x0)\partial G:=\partial\textup{Conv}(A_{0}x_{0}). Clearly G\partial G is closed, convex, isometric to 𝕊k1\mathbb{S}^{k-1}. Notice that the set G\partial G does not depend on the choice of x0Min(A0)x_{0}\in\text{Min}(A_{0}), since for xMin(A0)x\in\text{Min}(A_{0}) the subsets Conv(A0x)\textup{Conv}(A_{0}x) and Conv(A0x0)\textup{Conv}(A_{0}x_{0}) can be identified, by the Flat Torus Theorem, to two parallel slices {z}×k\{z\}\times\mathbb{R}^{k} and {z0}×k\{z_{0}\}\times\mathbb{R}^{k}, which have the same boundary. Also, the subset G\partial G is GG-invariant: in fact, A0A_{0} is normal in GG, so Min(A0)\text{Min}(A_{0}) is GG-invariant, therefore

gG=(gConv(A0x0))=Conv((gA0g1)gx0)=Conv(A0gx0)=Gg\cdot\partial G=\partial\left(g\cdot\text{Conv}(A_{0}x_{0})\right)=\partial\text{Conv}((gA_{0}g^{-1})gx_{0})=\partial\text{Conv}(A_{0}gx_{0})=\partial G

because gx0Min(A0)gx_{0}\in\text{Min}(A_{0}). Again by the Flat Torus Theorem (namely, property (c) recalled before), if A<GA<G is another free abelian subgroup of finite index and xMin(A)x\in\text{Min}(A) then Conv(Ax)=Conv((AA0)x)=Conv((AA0)x0)=Conv(A0x0)\text{Conv}(Ax)=\text{Conv}((A\cap A_{0})x)=\text{Conv}((A\cap A_{0})x_{0})=\text{Conv}(A_{0}x_{0}), so we have Conv(Ax)=G\partial\text{Conv}(Ax)=\partial G too. This proves (i).
To show (ii), it is enough to consider free abelian subgroups A<GA^{\prime}<G^{\prime} and A<GA<G of finite index. We can even suppose A<AA^{\prime}<A up to replacing AA^{\prime} by AAA^{\prime}\cap A. If xMin(A)x\in\text{Min}(A) then xMin(A)x\in\text{Min}(A^{\prime}), and by the first part of the statement we have G=Conv(Ax)Conv(Ax)=G.\partial G^{\prime}=\partial\text{Conv}(A^{\prime}x)\subseteq\partial\text{Conv}(Ax)=\partial G. If moreover GG and GG^{\prime} have the same rank then AA^{\prime} is a finite index subgroup of GG by Lemma 2.6, and G=Conv(Ax)=G\partial G=\partial\text{Conv}(A^{\prime}x)=\partial G^{\prime}. ∎

3. Systole and diastole

In this section we compare some different invariants of an action of group Γ\Gamma on a CAT(0)(0)-space XX, 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 Γ\Gamma at a point xXx\in X are defined respectively as

sys(Γ,x):=infgΓd(x,gx),sys(Γ,x):=infgΓΓd(x,gx),\text{sys}(\Gamma,x):=\inf_{g\in\Gamma^{*}}d(x,gx),\qquad\text{sys}^{\diamond}(\Gamma,x):=\inf_{g\in\Gamma^{\ast}\setminus\Gamma^{\diamond}}d(x,gx),

where Γ=Γ{id}\Gamma^{*}=\Gamma\setminus\{\text{id}\} and Γ\Gamma^{\diamond} is the subset of all elliptic isometries of Γ\Gamma.
The (global) systole and the free-systole of Γ\Gamma are accordingly defined as

sys(Γ,X)=infxXsys(Γ,x),sys(Γ,X)=infxXsys(Γ,x).\text{sys}(\Gamma,X)=\inf_{x\in X}\text{sys}(\Gamma,x),\qquad\text{sys}^{\diamond}(\Gamma,X)=\inf_{x\in X}\text{sys}^{\diamond}(\Gamma,x).

Similarly, the diastole and the free-diastole of Γ\Gamma are defined as

dias(Γ,X)=supxXsys(Γ,x),dias(Γ,X)=supxXsys(Γ,x).\text{dias}(\Gamma,X)=\sup_{x\in X}\text{sys}(\Gamma,x),\qquad\text{dias}^{\diamond}(\Gamma,X)=\sup_{x\in X}\text{sys}^{\diamond}(\Gamma,x).

In [CS22] the authors showed that dias(Γ,X)>0(\Gamma,X)>0 if and only if there exists a fundamental domain for Γ\Gamma; that is, if and only if there exists a point x0x_{0} such that the pointwise stabilizer StabΓ(x0)\text{Stab}_{\Gamma}(x_{0}) is trivial. The actions satisfying this property will be called nonsingular, and singular when dias(Γ,X)=0(\Gamma,X)=0 (with abuse of language, we will often say that the group Γ\Gamma itself is singular or nonsingular). For a nonsingular action, the Dirichlet domain at x0x_{0} is defined as

Dx0={yX s.t. d(x0,y)<d(x0,gy) for all gΓ}\mathrm{D}_{x_{0}}=\{y\in X\text{ s.t. }d(x_{0},y)<d(x_{0},gy)\text{ for all }g\in\Gamma^{*}\}

and is always a fundamental domain for the action of Γ\Gamma, see [CS22, Prop. 2.9]. Notice that a fundamental domain exists when, for instance, Γ\Gamma is torsion-free, or when XX 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:

sys(Γ,X)sys(Γ,X)dias(Γ,X)\text{sys}(\Gamma,X)\leq\text{sys}^{\diamond}(\Gamma,X)\leq\text{dias}^{\diamond}(\Gamma,X)
sys(Γ,X)dias(Γ,X)dias(Γ,X).\text{sys}(\Gamma,X)\leq\text{dias}(\Gamma,X)\leq\text{dias}^{\diamond}(\Gamma,X).

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 P0,r0,D0P_{0},r_{0},D_{0}, there exists b0=b0(P0,r0)>0b_{0}=b_{0}(P_{0},r_{0})>0 such that the following holds true. Let XX be a proper, geodesically complete, (P0,r0)(P_{0},r_{0})-packed, CAT(0)\textup{CAT}(0)-space and let Γ<Isom(X)\Gamma<\textup{Isom}(X) be discrete and D0D_{0}-cocompact. Then:

  • (i)

    sys(Γ,X)(1+P0)(2D0+1)min{dias(Γ,X),r0}\textup{sys}^{\diamond}(\Gamma,X)\geq\left(1+P_{0}\right)^{-\frac{(2D_{0}+1)}{\min\{\textup{dias}^{\diamond}(\Gamma,X),r_{0}\}}}.

  • (ii)

    If moreover the action of Γ\Gamma on XX is nonsingular then

    dias(Γ,X)b0min{sys(Γ,X),ε0}.\textup{dias}(\Gamma,X)\geq b_{0}\cdot\min\{\textup{sys}^{\diamond}(\Gamma,X),\varepsilon_{0}\}.

(Here, ε0\varepsilon_{0} 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(Γ,X)=0(\Gamma,X)=0 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 Γ\Gamma on a proper CAT(0)(0)-space XX, one has dias(Γ,X)>0\text{dias}(\Gamma,X)>0 if there exists some point trivial stabilizer, but clearly sys(Γ,X)=0\text{sys}(\Gamma,X)=0 if Γ\Gamma has torsion. Finally, notice that the inequality (ii) holds for a constant depending only on P0P_{0} and r0r_{0}, and not on D0D_{0}; 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(0)(0)-spaces which are ε\varepsilon-thin for sufficiently small ε\varepsilon.

Proposition 3.2 ([CS22, Theorem A & Remark 3.5]).

Let P0,r0>0P_{0},r_{0}>0 and fix 0<λε00<\lambda\leq\varepsilon_{0}, where ε0\varepsilon_{0} is given by Proposition 2.4. There exists a constant b0=b0(P0,r0)>0b_{0}=b_{0}(P_{0},r_{0})>0 such that the following holds true. Let XX be a proper, geodesically complete, (P0,r0)(P_{0},r_{0})-packed, CAT(0)\textup{CAT}(0)-space and let Γ<Isom(X)\Gamma<\textup{Isom}(X) be discrete and cocompact. If dias(Γ,X)b0λ(\Gamma,X)\leq b_{0}\cdot\lambda then

X=gΓ,(g)λMin(g).X=\bigcup_{g\in\Gamma^{*},\,\ell(g)\leq\lambda}\textup{Min}(g).

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 P0,r0,R>0P_{0},r_{0},R>0 and 0<εr00<\varepsilon\leq r_{0}. Then there exists δ(P0,r0,R,ε)>0\delta(P_{0},r_{0},R,\varepsilon)>0 with the following property. Let XX be a proper, geodesically complete, (P0,r0)(P_{0},r_{0})-packed, CAT(0)(0)-space. If gg is an isometry of XX with infinite order and xx is a point of XX such that d(x,gx)δ(P0,r0,R,ε)d(x,gx)\leq\delta(P_{0},r_{0},R,\varepsilon), then for every yXy\in X with d(x,y)Rd(x,y)\leq R there exists mm\in\mathbb{Z}^{*} such that d(y,gmy)εd(y,g^{m}y)\leq\varepsilon.
(We can choose, explicitely, δ(P0,r0,R,ε)=(1+P0)e(2R+1)ε\delta(P_{0},r_{0},R,\varepsilon)=(1+P_{0})e^{-\frac{(2R+1)}{\varepsilon}}).

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(0)(0)-spaces); but in that proof we never used the hyperbolicity.. ∎

Proof of Theorem 3.1.

Assume that min{dias(Γ,X),r0}>ε\min\{\text{dias}^{\diamond}(\Gamma,X),r_{0}\}>\varepsilon. By definition there exists x0Xx_{0}\in X such that for every hyperbolic isometry gΓg\in\Gamma one has d(x0,gx0)>εd(x_{0},gx_{0})>\varepsilon. Now, if sys(Γ,X)(1+P0)e(2D0+1)ε=:δ{}^{\diamond}(\Gamma,X)\leq(1+P_{0})e^{-\frac{(2D_{0}+1)}{\varepsilon}}=:\delta, we could find xXx\in X and a hyperbolic gΓg\in\Gamma such that d(x,gx)δd(x,gx)\leq\delta. By Proposition 3.3, for every yB(x,D0)y\in B(x,D_{0}) there would exists a non trivial power gmg^{m} satisfying d(y,gmy)εd(y,g^{m}y)\leq\varepsilon. But then, since the action is D0D_{0}-cocompact we could find a conjugate γ\gamma of gmg^{m} (thus, a hyperbolic isometry) such that d(x0,γx0)εd(x_{0},\gamma x_{0})\leq\varepsilon, a contradiction. The conclusion follows by the arbitrariness of ε\varepsilon.
To see (ii), take any λ<min{sys(Γ,X),ε0}\lambda<\min\{\text{sys}^{\diamond}(\Gamma,X),\varepsilon_{0}\}. If dias(Γ,X)b0λ\text{dias}(\Gamma,X)\leq b_{0}\cdot\lambda, where b0b_{0} is the constant given by Proposition 3.2, we conclude that X=gMin(g)X=\bigcup_{g}\text{Min}(g)where gg runs over all nontrivial elements with (g)λ\ell(g)\leq\lambda. But by definition (g)>λ\ell(g)>\lambda for all hyperbolic gg, so X=gΓMin(g).X=\bigcup_{g\in\Gamma^{\diamond}}\text{Min}(g). Hence, dias(Γ,X)=0\text{dias}(\Gamma,X)=0, contradicting the nonsingularity of Γ\Gamma. As λ\lambda 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(0)(0)-space XX which is (P0,r0)(P_{0},r_{0})-packed we have an upper bound on the dimension of XX given by Proposition 2.3

dim(X)n0=P0/2\dim(X)\leq n_{0}=P_{0}/2

and a Margulis’s constant ε0\varepsilon_{0} (only depending on P0,r0P_{0},r_{0}) given by Proposition 2.4. Also, recall the constant J(k)J(k) given by Proposition 2.5, and define

J0:=maxk{0,,n0}J(k)+1J_{0}:=\max_{k\in\{0,\ldots,n_{0}\}}J(k)+1

which also clearly depends only on n0n_{0}, so ultimately only on P0P_{0}.
Finally, for a discrete subgroup Γ<Isom(X)\Gamma<\textup{Isom}(X) recall the definition (4) of the subgroup Γ¯r(x)<Γ\overline{\Gamma}_{r}(x)<\Gamma generated by Σ¯r(x)\overline{\Sigma}_{r}(x) given in Section 2.1.

Theorem 4.1.

Given positive constants P0,r0,D0P_{0},r_{0},D_{0}, there exists a function σP0,r0,D0:(0,ε0](0,ε0]\sigma_{P_{0},r_{0},D_{0}}:(0,\varepsilon_{0}]\rightarrow(0,\varepsilon_{0}] (depending only on the parameters P0,r0,D0P_{0},r_{0},D_{0}) such that the following holds. Let XX be a proper, geodesically complete, (P0,r0)(P_{0},r_{0})-packed, CAT(0)\textup{CAT}(0)-space, and Γ<Isom(X)\Gamma<\textup{Isom}(X) be discrete and D0D_{0}-cocompact. For every chosen ε(0,ε0]\varepsilon\in(0,\varepsilon_{0}], if sys(Γ,X)σP0,r0,D0(ε)\textup{sys}^{\diamond}(\Gamma,X)\leq\sigma_{P_{0},r_{0},D_{0}}(\varepsilon) then:

  • (i)

    the space XX splits isometrically as Y×kY\times\mathbb{R}^{k}, with k1k\geq 1, and this splitting is Γ\Gamma-invariant;

  • (ii)

    there exists ε(σP0,r0,D0(ε),ε){\varepsilon^{\ast}}\in(\sigma_{P_{0},r_{0},D_{0}}(\varepsilon),\varepsilon) such that the rank of the virtually abelian subgroups Γ¯ε(x)\overline{\Gamma}_{{\varepsilon^{\ast}}}(x) is exactly kk, for all xXx\in X;

  • (iii)

    the traces at infinity Γ¯ε(x)\partial\overline{\Gamma}_{{\varepsilon^{\ast}}}(x) equal the boundary 𝕊k1{\mathbb{S}}^{k-1} of the convex subsets {y}×k\{y\}\times\mathbb{R}^{k}, for all xXx\in X and all yYy\in Y;

  • (iv)

    for every xXx\in X there exists yYy\in Y such that Γ¯ε(x)\overline{\Gamma}_{\varepsilon^{\ast}}(x) preserves {y}×k\{y\}\times\mathbb{R}^{k};

  • (v)

    the projection of Γ¯ε(x)\overline{\Gamma}_{\varepsilon^{\ast}}(x) on Isom(k)\textup{Isom}(\mathbb{R}^{k}) is a crystallographic group, whose maximal lattice is generated by the projection of a subset ΣΣ4J0ε(x)\Sigma\subset\Sigma_{4J_{0}\cdot{\varepsilon^{\ast}}}(x);

  • (vi)

    the closure of the projection of Γ¯ε(x)\overline{\Gamma}_{\varepsilon^{\ast}}(x) on Isom(Y)\textup{Isom}(Y) is compact and totally disconnected.

Here, by Γ\Gamma-invariant splitting we mean that every isometry of Γ\Gamma preserves the product decomposition. By of [BH13, Proposition I.5.3.(4)] we can see Γ\Gamma as a subgroup of Isom(Y)×Isom(k)\textup{Isom}(Y)\times\textup{Isom}(\mathbb{R}^{k}). In particular it is meaningful to talk about the projection of AA on Isom(Y)\text{Isom}(Y) and Isom(k)\text{Isom}(\mathbb{R}^{k}).

Observe that Theorem D is a special case of Theorem 4.1.(i), for ε=ε0\varepsilon=\varepsilon_{0}, which yields the constant σ0=σP0,r0,D0(ε0)\sigma_{0}=\sigma_{P_{0},r_{0},D_{0}}(\varepsilon_{0}). We call the integer 1kn01\leq k\leq n_{0} given by (ii) the ε{\varepsilon^{\ast}}-splitting rank of XX.

To prove Theorem 4.1 we need a little of preparation.
The first step will be to exhibit a free abelian subgroup of rank k1k\geq 1 which is commensurated in Γ\Gamma. Recall that two subgroups G1,G2<ΓG_{1},G_{2}<\Gamma are called commensurable in Γ\Gamma if the intersection G1G2G_{1}\cap G_{2} has finite index in both G1G_{1} and G2G_{2}. A subgroup G<ΓG<\Gamma is said to be commensurated in Γ\Gamma if GG and γGγ1\gamma G\gamma^{-1} are commensurable in Γ\Gamma for every γΓ\gamma\in\Gamma.

Now, we fix 0<εε00<\varepsilon\leq\varepsilon_{0} as in the assumptions of Theorem 4.1, and we define inductively the sequence of subgroups Γ¯εi(x)\overline{\Gamma}_{\varepsilon_{i}}(x) associated to positive numbers

ε1:=ε>ε2>>ε2n0+1>0\varepsilon_{1}:=\varepsilon>\varepsilon_{2}>\ldots>\varepsilon_{2n_{0}+1}>0

as follows:
– first, we apply Proposition 3.3 to ε=ε1\varepsilon=\varepsilon_{1} and R=2D0R=2D_{0} to obtain a smaller δ2:=δ(P0,r0,2D0,ε1)\delta_{2}:=\delta(P_{0},r_{0},2D_{0},\varepsilon_{1}), and we set ε2:=δ2/4J0\varepsilon_{2}:=\delta_{2}/4J_{0};
– then, we define inductively δi+1:=δ(P0,r0,2D0,εi)\delta_{i+1}:=\delta(P_{0},r_{0},2D_{0},\varepsilon_{i}) by repeatedly applying Proposition 3.3 to εi\varepsilon_{i} and R=2D0R=2D_{0}, and we set εi+1=δi+1/4J0\varepsilon_{i+1}=\delta_{i+1}/4J_{0}.
Notice that, by construction, each εi\varepsilon_{i} depends only on P0,r0,D0P_{0},r_{0},D_{0} and ε\varepsilon.
By Proposition 2.4, the subgroups Γ¯εi(x)\overline{\Gamma}_{\varepsilon_{i}}(x) form a decreasing sequence of virtually abelian, semisimple subgroups for every xXx\in X.

We set σP0,r0,D0(ε):=ε2n0+1\sigma_{P_{0},r_{0},D_{0}}(\varepsilon):=\varepsilon_{2n_{0}+1} and we will show that this is the function of ε\varepsilon for which Theorem 4.1 holds; it clearly depends only on P0,r0,D0P_{0},r_{0},D_{0} and ε\varepsilon. In what follows, we will write for short σ:=σP0,r0,D0(ε)\sigma:=\sigma_{P_{0},r_{0},D_{0}}(\varepsilon).

Lemma 4.2.

If rk(Γ¯σ(x))1\textup{rk}(\overline{\Gamma}_{\sigma}(x))\geq 1 then there exists i{2,,2n0}i\in\{2,\ldots,2n_{0}\} such that rk(Γ¯εi+1(x))=rk(Γ¯εi(x))=rk(Γ¯εi1(x))1\textup{rk}(\overline{\Gamma}_{\varepsilon_{i+1}}(x))=\textup{rk}(\overline{\Gamma}_{\varepsilon_{i}}(x))=\textup{rk}(\overline{\Gamma}_{\varepsilon_{i-1}}(x))\geq 1.

Proof.

The subgroups Γ¯εi(x)\overline{\Gamma}_{\varepsilon_{i}}(x) are virtually abelian with rk(Γ¯εi(x))1\textup{rk}(\overline{\Gamma}_{\varepsilon_{i}}(x))\geq 1, for all 1i2n0+11\leq i\leq 2n_{0}+1, since they contain Γ¯σ(x)\overline{\Gamma}_{\sigma}(x). Moreover, by Proposition 2.7, the rank of each Γ¯εi(x)\overline{\Gamma}_{\varepsilon_{i}}(x) cannot exceed the dimension of XX, which is at most n0n_{0}. Since the rank decreases as ii increases we conclude that for some 2i2n02\leq i\leq 2n_{0} we have rk(Γ¯εi+1(x))=rk(Γ¯εi(x))=rk(Γ¯εi1(x))1\textup{rk}(\overline{\Gamma}_{\varepsilon_{i+1}}(x))=\textup{rk}(\overline{\Gamma}_{\varepsilon_{i}}(x))=\textup{rk}(\overline{\Gamma}_{\varepsilon_{i-1}}(x))\geq 1. ∎

Proposition 4.3.

If rk(Γ¯σ(x))1\textup{rk}(\overline{\Gamma}_{\sigma}(x))\geq 1 then there exists a free abelian subgroup A<Γ¯εi(x)A<\overline{\Gamma}_{\varepsilon_{i}}(x) of rank k1k\geq 1, which has finite index in Γ¯εi1(x)\overline{\Gamma}_{\varepsilon_{i-1}}(x) and is commensurated in Γ\Gamma, for some i{2,,2n0}i\in\{2,\ldots,2n_{0}\}.

Proof.

Consider the virtually abelian groups Γ¯εi+1(x)<Γ¯εi(x)<Γ¯εi1(x)\overline{\Gamma}_{\varepsilon_{i+1}}(x)<\overline{\Gamma}_{\varepsilon_{i}}(x)<\overline{\Gamma}_{\varepsilon_{i-1}}(x) which have the same rank k1k\geq 1, given by Lemma 4.2, for some 2i2n02\leq i\leq 2n_{0} and let Aεj(x)<Γ¯εj(x)A_{\varepsilon_{j}}(x)<\overline{\Gamma}_{\varepsilon_{j}}(x) be free abelian, finite index subgroups of rank kk, for j=i1,i,i+1j=i-1,i,i+1. Notice that Aεi+1(x)A_{\varepsilon_{i+1}}(x) has finite index also in Γ¯εi(x)\overline{\Gamma}_{\varepsilon_{i}}(x), by Lemma 2.6. Then, let Ci+1=Zi+1×kC_{i+1}=Z_{i+1}\times{\mathbb{R}}^{k} be the Γ¯εi+1(x)\overline{\Gamma}_{\varepsilon_{i+1}}(x)-invariant, convex subset of XX given by Proposition 2.7, applied to Γ¯εi+1(x)\overline{\Gamma}_{\varepsilon_{i+1}}(x), and call Γ¯i+1\bar{\Gamma}_{i+1} the image of Γ¯εi+1(x)\overline{\Gamma}_{\varepsilon_{i+1}}(x) under the projection pi+1:Γ¯εi+1(x)Isom(k)p_{i+1}:\overline{\Gamma}_{\varepsilon_{i+1}}(x)\rightarrow\text{Isom}(\mathbb{R}^{k}) on the second factor of Ci+1C_{i+1}. Finally, denote by i+1{\mathcal{L}}_{i+1} the maximal Euclidean lattice of the crystallograhic group Γ¯i+1\bar{\Gamma}_{i+1}.

By Lemma 2.8 we can find a subset Σ\Sigma of Σ¯εi+1(x)4J0Σ¯4J0εi+1(x)\overline{\Sigma}_{\varepsilon_{i+1}}(x)^{4J_{0}}\subseteq\overline{\Sigma}_{4J_{0}\cdot\varepsilon_{i+1}}(x) whose projection Σk=pi+1(Σ)\Sigma_{\mathbb{R}^{k}}=p_{i+1}(\Sigma) on Isom(k)(\mathbb{R}^{k}) generates the lattice i+1{\mathcal{L}}_{i+1}. In particular every non-trivial element of Σ\Sigma is hyperbolic.
Moreover, by the definition of εi+1=δi+1/4J0\varepsilon_{i+1}=\delta_{i+1}/4J_{0}, the following holds:

gΣ\forall g\in\Sigma and hΣ¯2D0(x)\forall h\in\overline{\Sigma}_{2D_{0}}(x) there exists m>0m>0 such that h1gmhΣ¯εi(x)h^{-1}g^{m}h\in\overline{\Sigma}_{\varepsilon_{i}}(x)

(in fact, d(x,gx)δi+1=δ(P0,r0,2D0,εi)d(x,gx)\leq\delta_{i+1}=\delta(P_{0},r_{0},2D_{0},{\varepsilon_{i}}) for every gΣg\in\Sigma, so by Proposition 3.3 there exists m>0m>0 such that d(x,h1gmhx)=d(hx,gmhx)εid(x,h^{-1}g^{m}hx)=d(hx,g^{m}hx)\leq\varepsilon_{i}, that is h1gmhΣ¯εi(x)h^{-1}g^{m}h\in\overline{\Sigma}_{\varepsilon_{i}}(x)).
Since Σ\Sigma and Σ¯2D0(x)\overline{\Sigma}_{2D_{0}}(x) are finite sets, we can then find a positive integer MM such that hgMh1Γ¯εi(x)hg^{M}h^{-1}\in{\overline{\Gamma}_{\varepsilon_{i}}}(x) for all gΣg\in\Sigma and all hΣ¯2D0(x)h\in\overline{\Sigma}_{2D_{0}}(x).
Moreover, we can even choose M>0M>0 so that gΣ\forall g\in\Sigma and hΣ¯2D0(x)\forall h\in\overline{\Sigma}_{2D_{0}}(x) the elements gMg^{M} and hgMh1hg^{M}h^{-1} belong, respectively, to the free abelian, finite index subgroups Aεi+1(x)A_{\varepsilon_{i+1}}(x) and Aεi(x)A_{\varepsilon_{i}}(x) of Γ¯εi(x)\overline{\Gamma}_{\varepsilon_{i}}(x).
Now, let A<Aεi+1(x)A<A_{\varepsilon_{i+1}}(x) be the (free abelian) subgroup generated by the subset S={gM|gΣ}S=\{g^{M}\,\;|\;\,g\in\Sigma\}. We claim that AA is commensurated in Γ\Gamma.
Actually, notice first that AA has rank equal to kk, since its projection pi+1(A)p_{i+1}(A) is a subgroup of finite index of the lattice i+1{\mathcal{L}}_{i+1} of k{\mathbb{R}^{k}}. Therefore, for all hΣ¯2D0(x)h\in\overline{\Sigma}_{2D_{0}}(x), the free abelian group hAh1hAh^{-1} has also rank kk, and is contained in the free abelian group Aεi(x)A_{\varepsilon_{i}}(x) of same rank. This implies, again by Lemma 2.6, that AA and hAh1hAh^{-1} have finite index in Aεi(x)A_{\varepsilon_{i}}(x), and that AhAh1A\cap hAh^{-1} has finite index in both AA and hAh1hAh^{-1}. Hence AA and hAh1hAh^{-1} are commensurable for every hΣ¯2D0(x)h\in\overline{\Sigma}_{2D_{0}}(x).

Now, given gΓg\in\Gamma, we can write it as g=h1hng=h_{1}\cdots h_{n} for some hiΣ¯2D0(x)h_{i}\in\overline{\Sigma}_{2D_{0}}(x) and set gk:=h1hkg_{k}:=h_{1}\cdots h_{k}. We clearly have

gkAgk1gk+1Agk+11=gk(Ahk+1Ahk+11)gk1g_{k}Ag_{k}^{-1}\cap g_{k+1}Ag_{k+1}^{-1}=g_{k}(A\cap h_{k+1}Ah_{k+1}^{-1})g_{k}^{-1}

with Ahk+1Ahk+11A\cap h_{k+1}Ah_{k+1}^{-1} of finite index in both factors; so gkAgk1g_{k}Ag_{k}^{-1} and gk+1Agk+11g_{k+1}Ag_{k+1}^{-1} are commensurable. Since commensurability in a group is a transitive relation, this shows that AA and gAg1gAg^{-1} are commensurable for all gΓg\in\Gamma.
Finally, observe that A<Γ¯εi1(x)A<\overline{\Gamma}_{\varepsilon_{i-1}}(x) and these groups have same rank, so AA has finite index also in Γ¯εi1(x)\overline{\Gamma}_{\varepsilon_{i-1}}(x), again by Lemma 2.6. ∎

We need now to recall an additional notion. Given ZXZ\subseteq\partial X we say that a subset YXY\subseteq X is ZZ-boundary-minimal if it is closed, convex, Y=Z\partial Y=Z and YY is minimal with this properties. The union of all the ZZ-boundary-minimal sets is denoted by Bd-Min(Z)(Z). A particular case of [CM09b, Proposition 3.6] reads as follows.

Lemma 4.4.

Let XX be a proper CAT(0)\textup{CAT}(0)-space and let ZZ be a closed, convex subset of X\partial X which is isometric to 𝕊k1\mathbb{S}^{k-1}. Then each ZZ-boundary-minimal subset of XX is isometric to k\mathbb{R}^{k} and Bd-Min(Z)(Z) is a closed, convex subset of XX which splits isometrically as Y×kY\times\mathbb{R}^{k}. Moreover ZZ coincides with the boundary at infinity of all the slices {y}×k\{y\}\times\mathbb{R}^{k}, for yYy\in Y.

A consequence of Lemma 4.4 for a commensurated group AA of Γ\Gamma is the following.

Proposition 4.5.

Same assumptions as in Theorem 4.1.
If AA is a free abelian, commensurated subgroup of Γ\Gamma of rank kk then we have X=Bd-Min(A)X=\textup{Bd-Min}(\partial A) and XX splits isometrically and Γ\Gamma-invariantly as Y×kY\times\mathbb{R}^{k}. Moreover, the projection of AA on Isom(k)\textup{Isom}(\mathbb{R}^{k}) is a lattice and the closure of the projection of AA on Isom(Y)\textup{Isom}(Y) is compact and totally disconnected.
The splitting X=Y×kX=Y\times\mathbb{R}^{k} satisfies the following properties:

  • (i)

    the trace at infinity A\partial A is Γ\Gamma-invariant and coincides with the boundary of each slice {y}×k\{y\}\times\mathbb{R}^{k}, for all yYy\in Y;

  • (ii)

    if A<AA^{\prime}<A is another free abelian, commensurated subgroup of Γ\Gamma of rank kk^{\prime} then the splittings X=Y×kX=Y\times\mathbb{R}^{k} and X=Y×kX=Y^{\prime}\times\mathbb{R}^{k^{\prime}} associated respectively to AA and AA^{\prime} are compatible, i.e. YY^{\prime} is isometric to Y×kkY\times\mathbb{R}^{k-k^{\prime}}.

Proof.

The trace at infinity A\partial A of AA is isometric to 𝕊k1\mathbb{S}^{k-1}, by Proposition 2.9. We claim that it is Γ\Gamma-invariant. Indeed, if gΓg\in\Gamma then AgAg1A\cap gAg^{-1} has finite index in both AA and gAg1gAg^{-1}. Then, the characterization of the trace at infinity given in Proposition 2.9 implies that

A=(AgAg1)=(gAg1)=gA.\partial A=\partial\left(A\cap gAg^{-1}\right)=\partial\left(gAg^{-1}\right)=g\partial A.

By Lemma 4.4 applied to Z=AZ=\partial A we deduce that Bd-Min(A)\text{Bd-Min}(\partial A) is a closed, convex subset of XX which splits isometrically as Y×kY\times\mathbb{R}^{k}, and that A\partial A coincides with the boundary at infinity of all sets {y}×k\{y\}\times\mathbb{R}^{k}.
Now, each element of Γ\Gamma sends a A\partial A-boundary-minimal subset into a A\partial A-boundary-minimal subset because A\partial A is Γ\Gamma-invariant, therefore Bd-Min(A)\text{Bd-Min}(\partial A) itself is Γ\Gamma-invariant. Since Γ\Gamma is cocompact, the action of Γ\Gamma on XX is minimal: that is, if CC\neq\emptyset is a closed, convex, Γ\Gamma-invariant subset of XX then C=XC=X (see [CM09b, Lemma 3.13]). Therefore we deduce that X=Bd-Min(A)X=\text{Bd-Min}(\partial A), and so XX splits isometrically and Γ\Gamma-invariantly as Y×kY\times\mathbb{R}^{k}, which proves the first assertion and (i).
The fact that the projection of AA on Isom(k)\textup{Isom}(\mathbb{R}^{k}) is a lattice follows from the Flat Torus Theorem. To study the projection of AA on Isom(Y)\text{Isom}(Y), recall that by Proposition 2.2 we can split XX as M×n×NM\times\mathbb{R}^{n}\times N, for some nkn\geq k, and Isom(X)\textup{Isom}(X) as 𝒮×n×𝒟{\mathcal{S}}\times{\mathcal{E}}_{n}\times{\mathcal{D}}, where 𝒮{\mathcal{S}} is a semi-simple Lie group with trivial center and without compact factors, nIsom(n){\mathcal{E}}_{n}\cong\text{Isom}(\mathbb{R}^{n}) and 𝒟{\mathcal{D}} is totally disconnected. Therefore Y=M×nk×NY=M\times\mathbb{R}^{n-k}\times N and Isom(Y)\textup{Isom}(Y) splits as 𝒮×nk×𝒟{\mathcal{S}}\times{\mathcal{E}}_{n-k}\times{\mathcal{D}}. Moreover, Min(A)=Z×k\text{Min}(A)=Z\times\mathbb{R}^{k} with ZYZ\subset Y. Since AA acts as the identity on ZZ, it follows that the projection of AA on Isom(Y)\textup{Isom}(Y) fixes some point zYz\in Y, 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 AA on 𝒮{\mathcal{S}} is finite and the projection of AA on n{\mathcal{E}}_{n} is discrete; as the projection of AA on k{\mathcal{E}}_{k} is a lattice, then also the projection on nk{\mathcal{E}}_{n-k} is discrete. This, combined with the fact that 𝒟{\mathcal{D}} is totally disconnected, implies that the closure of the projection of AA on Isom(Y)\text{Isom}(Y) is totally disconnected.
Suppose now to have another abelian subgroup A<AA^{\prime}<A of rank kk^{\prime} which is commensurated in Γ\Gamma. Let X=Y×kX=Y^{\prime}\times\mathbb{R}^{k^{\prime}} be the splitting associated to AA^{\prime}. Let xXx\in X be a point and write it as (y,𝐯)Y×k(y,{\bf v})\in Y\times\mathbb{R}^{k} and (y,𝐯)Y×k(y^{\prime},{\bf v^{\prime}})\in Y^{\prime}\times\mathbb{R}^{k^{\prime}}. Then, by the first part of the proof and by Proposition 2.9 we have

({y}×k)=AA=({y}×k).\partial(\{y^{\prime}\}\times\mathbb{R}^{k^{\prime}})=\partial A^{\prime}\subseteq\partial A=\partial(\{y\}\times\mathbb{R}^{k}).

It follows that {y}×k{y}×k\{y^{\prime}\}\times\mathbb{R}^{k^{\prime}}\subseteq\{y\}\times\mathbb{R}^{k}, so the parallel slices associated to AA^{\prime} are contained in the parallel slices associated to AA. Decomposing k\mathbb{R}^{k} as the orhogonal sum of k\mathbb{R}^{k^{\prime}} and kk\mathbb{R}^{k-k^{\prime}}, we also deduce that the sets {y}×kk\{y\}\times\mathbb{R}^{k-k^{\prime}} are parallel for all yYy\in Y (since the slices of AA^{\prime} are all parallel). Therefore, XX is also isometric to (Y×kk)×k(Y\times\mathbb{R}^{k-k^{\prime}})\times\mathbb{R}^{k^{\prime}}, which implies that YY^{\prime} is isometric to Y×kkY\times\mathbb{R}^{k-k^{\prime}} and proves (ii). ∎

Putting the ingredients all together we can give the

Proof of Theorem 4.1.

We show that the statement holds for ε=εi(σ,ε){\varepsilon^{\ast}}=\varepsilon_{i}\in(\sigma,\varepsilon), where εi\varepsilon_{i} is given by Proposition 4.3. In fact, since sys(Γ,X)σ\textup{sys}^{\diamond}(\Gamma,X)\leq\sigma, then there exists x0Xx_{0}\in X with sys(Γ,x0)σ\textup{sys}^{\diamond}(\Gamma,x_{0})\leq\sigma; in particular, Γ¯σ(x0)\overline{\Gamma}_{\sigma}(x_{0}) contains a hyperbolic isometry, hence rk(Γ¯σ(x0))1\text{rk}(\overline{\Gamma}_{\sigma}(x_{0}))\geq 1. Then, we can apply Proposition 4.3 and find a free abelian, commensurated subgroup A0<ΓA_{0}<\Gamma of rank k1k\geq 1, with A0Γ¯εi+1(x0)A_{0}\subset\overline{\Gamma}_{\varepsilon_{i+1}}(x_{0}) and with finite index in Γ¯εi1(x0)\overline{\Gamma}_{\varepsilon_{i-1}}(x_{0}), for some 2i2n02\leq i\leq 2n_{0}; Proposition 4.5 now implies that XX splits isometrically and Γ\Gamma-invariantly as Y×kY\times\mathbb{R}^{k}, proving (i). Moreover, we know that A0\partial A_{0} coincides with the boundary at infinity of each slice {y}×k\{y\}\times\mathbb{R}^{k}.
Let us now study the properties (ii)-(vi) for the groups Γ¯εi(x)\overline{\Gamma}_{{\varepsilon_{i}}}(x).
We start proving them for every xXx\in X such that d(x,x0)D0d(x,x_{0})\leq D_{0}. By Proposition 3.3 and by definition of εi\varepsilon_{i}, for every hyperbolic gΣ¯εi(x)g\in\overline{\Sigma}_{\varepsilon_{i}}(x) there exists mm\in\mathbb{Z}^{*} such that d(x0,gmx0)εi1d(x_{0},g^{m}x_{0})\leq\varepsilon_{i-1}, that is gmΓ¯εi1(x0)g^{m}\in\overline{\Gamma}_{\varepsilon_{i-1}}(x_{0}). Let Aεi(x)<Γ¯εi(x)A_{\varepsilon_{i}}(x)<\overline{\Gamma}_{\varepsilon_{i}}(x) be free abelian of finite index, so rk(Aεi(x))=rk(Γ¯εi(x))\text{rk}(A_{\varepsilon_{i}}(x))=\text{rk}\left(\overline{\Gamma}_{\varepsilon_{i}}(x)\right) by definition. Since the set Σ¯εi(x)\overline{\Sigma}_{\varepsilon_{i}}(x) is finite, we can find MM\in\mathbb{Z}^{*} such that gMAεi(x)Γ¯εi1(x0)g^{M}\in A_{\varepsilon_{i}}(x)\cap\overline{\Gamma}_{\varepsilon_{i-1}}(x_{0}) for every hyperbolic gΣ¯εi(x)g\in\overline{\Sigma}_{\varepsilon_{i}}(x). So, if we define the subgroup

A:=gM|gΣ¯εi(x) hyperbolicA:=\langle g^{M}\,\;|\,\;g\in\overline{\Sigma}_{\varepsilon_{i}}(x)\text{ hyperbolic}\rangle

we have A<Aεi(x)Γ¯εi1(x0)A<A_{\varepsilon_{i}}(x)\cap\overline{\Gamma}_{\varepsilon_{i-1}}(x_{0}). Notice that AA is again free abelian of rank kk, hence it has finite index in Aεi(x)A_{\varepsilon_{i}}(x) and Γ¯εi(x)\overline{\Gamma}_{\varepsilon_{i}}(x), by Lemma 2.6. Thus

rk(Γ¯εi(x))=rk(A)rk(Γ¯εi1(x0))=rk(A0)=k.\text{rk}(\overline{\Gamma}_{\varepsilon_{i}}(x))=\text{rk}(A)\leq\text{rk}(\overline{\Gamma}_{\varepsilon_{i-1}}(x_{0}))=\text{rk}(A_{0})=k.

Moreover by Proposition 2.9 we have Γ¯εi(x)=AΓ¯εi1(x0)=A0\partial\overline{\Gamma}_{\varepsilon_{i}}(x)=\partial A\subseteq\partial\overline{\Gamma}_{\varepsilon_{i-1}}(x_{0})=\partial A_{0}. Reversing the roles of xx and x0x_{0} and starting from Γ¯εi+1(x0)\overline{\Gamma}_{\varepsilon_{i+1}}(x_{0}) we obtain the opposite estimate k=rk(A0)=rk(Γ¯εi+1(x0))rk(Γ¯εi(x))k=\text{rk}(A_{0})=\text{rk}(\overline{\Gamma}_{\varepsilon_{i+1}}(x_{0}))\leq\text{rk}(\overline{\Gamma}_{\varepsilon_{i}}(x)) and A0Γ¯εi(x)\partial A_{0}\subseteq\partial\overline{\Gamma}_{\varepsilon_{i}}(x), which proves (ii) and (iii) in this case.
By construction Γ¯εi(x)A\overline{\Gamma}_{\varepsilon_{i}}(x)\cap A has finite index in both Γ¯εi(x)\overline{\Gamma}_{\varepsilon_{i}}(x) and AA, so also the projection pk(Γ¯εi(x))p_{\mathbb{R}^{k}}(\overline{\Gamma}_{\varepsilon_{i}}(x)) on Isom(k)\textup{Isom}(\mathbb{R}^{k}) is discrete, and the closure of the projection pY(Γ¯ε(x))p_{Y}(\overline{\Gamma}_{\varepsilon^{\ast}}(x)) of Γ¯εi(x)\overline{\Gamma}_{\varepsilon_{i}}(x) on Isom(Y)\text{Isom}(Y) is compact and totally disconnected. Notice that pk(Γ¯εi(x))p_{\mathbb{R}^{k}}(\overline{\Gamma}_{\varepsilon_{i}}(x)) is cocompact, so it is a crystallographic group of k\mathbb{R}^{k}; moreover, as Σεi(x)\Sigma_{\varepsilon_{i}}(x) generates Γ¯εi(x)\overline{\Gamma}_{\varepsilon_{i}}(x), then the maximal lattice of pk(Γ¯εi(x))p_{\mathbb{R}^{k}}(\overline{\Gamma}_{\varepsilon_{i}}(x)) is generated by a subset of Σ4J0εi(x)\Sigma_{4J_{0}\cdot\varepsilon_{i}}(x), by Lemma 2.8. This proves (v) and (vi). Moreover, the precompact group pY(Γ¯εi(x))p_{Y}(\overline{\Gamma}_{\varepsilon_{i}}(x)) has a fixed point yYy\in Y ([BH13, Corollary II.2.8.(1)]), so {y}×k\{y\}\times\mathbb{R}^{k} is preserved by Γ¯εi(x)\overline{\Gamma}_{\varepsilon_{i}}(x), proving (iv). Finally, assume that xx^{\prime} is a point of XX, say x=gxx^{\prime}=gx with d(x,x0)D0d(x,x_{0})\leq D_{0}. Observe that Γ¯εi(x)=gΓ¯εi(x)g1\overline{\Gamma}_{\varepsilon_{i}}(x^{\prime})=g\overline{\Gamma}_{\varepsilon_{i}}(x)g^{-1}, so the rank does not change and the conditions (iv) and (v) continue to hold since the splitting is Γ\Gamma-invariant. Moreover Γ¯εi(x)=gΓ¯εi(x)=gA=A\partial\overline{\Gamma}_{\varepsilon_{i}}(x^{\prime})=g\cdot\partial\overline{\Gamma}_{\varepsilon_{i}}(x)=g\cdot\partial A=A, because A\partial A is Γ\Gamma-invariant, so (iii) still holds. Finally, if the group pY(Γ¯εi(x))p_{Y}(\overline{\Gamma}_{\varepsilon_{i}}(x)) preserves {y}×k\{y\}\times\mathbb{R}^{k}, then pY(Γ¯εi(x))p_{Y}(\overline{\Gamma}_{\varepsilon_{i}}(x^{\prime})) clearly preserves {gy}×k\{gy\}\times\mathbb{R}^{k}, 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 𝒪-CAT0(P0,r0,D0){\mathcal{O}}\text{-CAT}_{0}(P_{0},r_{0},D_{0}) (and of -CAT0(P0,r0,D0){\mathcal{L}}\text{-CAT}_{0}(P_{0},r_{0},D_{0}) 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 Γ\Gamma endowed with a generating set Σ\Sigma. Two marked groups (Γ,Σ)(\Gamma,\Sigma) and (Γ,Σ)(\Gamma^{\prime},\Sigma^{\prime}) are equivalent if there exists a group isomorphism ϕ:ΓΓ\phi:\Gamma\rightarrow\Gamma^{\prime} such that ϕ(Σ)=Σ\phi(\Sigma)=\Sigma^{\prime}; notice that such a ϕ\phi induces an isometry between the respective Cayley graphs.
The first step for Theorem A is the following:

Proposition 5.1.

Let P0,r0,D0P_{0},r_{0},D_{0} be given. For every fixed DD0D\geq D_{0} and s>0s>0, there exist only finitely many marked groups (Γ,Σ)(\Gamma,\Sigma) where:

  • Γ\Gamma is a discrete, D0D_{0}-cocompact isometry group of a proper, geodesically complete, (P0,r0)(P_{0},r_{0})-packed, CAT(0)(0)-space XX satisfying sys(Γ,x)s\textup{sys}(\Gamma,x)\geq s for some xXx\in X,

  • Σ=Σ¯2D(x)\Sigma=\overline{\Sigma}_{2D}(x) is a 2D2D-short generating set of Γ\Gamma at xx.

(Finiteness here is meant up to equivalence of marked groups.)

Proof.

Recall from Section 2.1 that for any DD0D\geq D_{0} the group Γ\Gamma admits a presentation as Γ=Σ¯2D(x)|2D(x)\Gamma=\langle\overline{\Sigma}_{2D}(x)\hskip 2.84526pt|\hskip 2.84526pt{\mathcal{R}}_{2D}(x)\rangle, where 2D(x){\mathcal{R}}_{2D}(x) is a subset of words of length 33 on the alphabet Σ¯2D(x)\overline{\Sigma}_{2D}(x). Therefore the number of equivalence classes of such marked groups (Γ,Σ)(\Gamma,\Sigma), with Σ=Σ¯2D(x)\Sigma=\overline{\Sigma}_{2D}(x), is bounded above if we are able to bound uniformly the cardinality of Σ¯2D(x)\overline{\Sigma}_{2D}(x). But this is an immediate consequence of Proposition 2.3; actually, using the fact that the points in the orbit of xx are s2\frac{s}{2}-separated, we get

#Σ¯2D(x)Pack(2D,s4)P0(1+P0)24Ds1.\#\overline{\Sigma}_{2D}(x)\leq\text{Pack}\left(2D,\frac{s}{4}\right)\leq P_{0}(1+P_{0})^{\frac{24D}{s}-1}.\qed

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 XX^{\prime} is isometric to XX (though generally the quotients Γ\X\Gamma\backslash X and Γ\X\Gamma\backslash X^{\prime} are not isometric, since the first one can be ε\varepsilon-collapsed for arbitrarily small ε\varepsilon, while the second one is not collapsed, by construction).
Naively, one can think that if sys(Γ,X)\textup{sys}^{\diamond}(\Gamma,X) is too small then, as we know that XX splits Γ\Gamma-invariantly as Y×kY\times\mathbb{R}^{k} by Theorem 4.1, the new space is X=Y×skX^{\prime}=Y\times s\cdot\mathbb{R}^{k}, obtained by dilating the Euclidean factor of a suitable s>0s>0, and the action of Γ\Gamma is the natural one induced on it. The construction of XX^{\prime} 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 XX 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 n0=P0/2n_{0}=P_{0}/2, which bounds the dimension of every proper, geodesically complete, CAT(0)(0)-space XX which is (P0,r0)(P_{0},r_{0})-packed, and J0=J0(P0)J_{0}=J_{0}(P_{0}), introduced at the beginning of Section 4. Also recall the Margulis constant ε0=ε0(P0,r0)\varepsilon_{0}=\varepsilon_{0}(P_{0},r_{0}) given by Proposition 2.4, which we will always assume smaller than 11 in the sequel.

Proof of Theorem E.

Recall the function σP0,r0,D0(ε)\sigma_{P_{0},r_{0},D_{0}}(\varepsilon) of Theorem 4.1. Then we define inductively D1=2D0+n0D_{1}=2D_{0}+\sqrt{n_{0}}, σ1=σP0,r0,D1(ε04J0)\sigma_{1}=\sigma_{P_{0},r_{0},D_{1}}(\frac{\varepsilon_{0}}{4J_{0}}) and

Dj=2Dj1+n0,σj=σP0,r0,Dj(σj1)>0.D_{j}=2D_{j-1}+\sqrt{n_{0}}\;\,,\hskip 28.45274pt\sigma_{j}=\sigma_{P_{0},r_{0},D_{j}}(\sigma_{j-1})>0.

We claim that Δ0:=Dn01\Delta_{0}:=D_{n_{0}-1} and s0:=σn0s_{0}:=\sigma_{n_{0}} satisfy the thesis; notice that both depend only on P0,r0P_{0},r_{0} and D0D_{0}.
We now describe a process which takes the CAT(0)(0)-space X0:=XX_{0}:=X and produces a new proper, geodesically complete, (P0,r0)(P_{0},r_{0})-packed, CAT(0)(0)-space X1X_{1} on which Γ\Gamma still acts faithfully and discretely by isometries, still satisfying all the assumptions of the theorem, except that diam(Γ\X1)D1\text{diam}(\Gamma\backslash X_{1})\leq D_{1}; and we will show that, repeating again and again this process, we end up with a CAT(0)(0)-space XjX_{j} with sys(Γ,Xj)>σn0\textup{sys}^{\diamond}(\Gamma,X_{j})>\sigma_{n_{0}}, for some jn0j\leq n_{0}.
If sys(Γ,X0)>σn0\textup{sys}^{\diamond}(\Gamma,X_{0})>\sigma_{n_{0}}, there is nothing to do, and we just set X=X0X^{\prime}=X_{0}.
Otherwise, sys(Γ,X)σn0<σ1=σP0,r0,D0(ε04J0)\textup{sys}^{\diamond}(\Gamma,X)\leq\sigma_{n_{0}}<\sigma_{1}=\sigma_{P_{0},r_{0},D_{0}}(\frac{\varepsilon_{0}}{4J_{0}}), and we apply Theorem 4.1 with ε=ε04J0\varepsilon=\frac{\varepsilon_{0}}{4J_{0}}. Then, there exists ε0:=ε(σ1,ε04J0)\varepsilon^{\ast}_{0}:=\varepsilon^{\ast}\in(\sigma_{1},\frac{\varepsilon_{0}}{4J_{0}}) such that the groups Γ¯ε0(x,X0)\overline{\Gamma}_{\varepsilon^{\ast}_{0}}(x,X_{0}) have all rank k01k_{0}\geq 1 for every xX0x\in X_{0}. We then fix x0X0x_{0}\in X_{0}. By Proposition 4.3 there exists a free abelian, finite index subgroup A0A_{0} of Γ¯ε0(x0,X)\overline{\Gamma}_{\varepsilon^{\ast}_{0}}(x_{0},X) of rank k01k_{0}\geq 1, which is commensurated in Γ\Gamma; and we have that X0=Bd-Min(A0)X_{0}=\textup{Bd-Min}(\partial A_{0}) splits isometrically and Γ\Gamma-invariantly as Y0×k0Y_{0}\times\mathbb{R}^{k_{0}} by Proposition 4.5. Moreover, always by Theorem 4.1, there exists y0Y0y_{0}\in Y_{0} such that Γ¯ε0(x0,X0)\overline{\Gamma}_{\varepsilon^{\ast}_{0}}(x_{0},X_{0}) preserves {y0}×k0\{y_{0}\}\times\mathbb{R}^{k_{0}}, and there exists a subset of Σ¯4J0ε0(x0,X0)\overline{\Sigma}_{4J_{0}\cdot\varepsilon^{\ast}_{0}}(x_{0},X_{0}) whose projection on Isom(k0)\textup{Isom}(\mathbb{R}^{k_{0}}) generates the maximal lattice ε0(x0,X0)\mathcal{L}_{\varepsilon^{\ast}_{0}}(x_{0},X_{0}) of the crystallographic group pk0(Γ¯ε0(x0,X0))p_{\mathbb{R}^{k_{0}}}(\overline{\Gamma}_{\varepsilon^{\ast}_{0}}(x_{0},X_{0})). So, we can find a shortest basis 0={𝐛𝟏𝟎,,𝐛𝐤𝟎𝟎}\mathcal{B}_{0}=\{\bf b^{0}_{1},\ldots,\bf b^{0}_{k_{0}}\} of the lattice ε0(x0,X0)\mathcal{L}_{\varepsilon^{\ast}_{0}}(x_{0},X_{0}) whose vectors have all length at most 4J0ε0<ε0<14J_{0}\cdot\varepsilon^{\ast}_{0}<\varepsilon_{0}<1; without loss of generality, we may suppose that 𝐛𝟏𝟎0𝐛𝐤𝟎𝟎0=λ(ε0(x0,X0))=:0<1\|{\bf b^{0}_{1}}\|_{0}\leq\ldots\leq\|{\bf b^{0}_{k_{0}}}\|_{0}=\lambda\left(\mathcal{L}_{\varepsilon^{\ast}_{0}}(x_{0},X_{0})\right)=:\ell_{0}<1, where 0\|\hskip 5.69054pt\|_{0} denotes the Euclidean norm of k0\mathbb{R}^{k_{0}}. By (8), we also know that the covering radius of ε0(x0,X0)\mathcal{L}_{\varepsilon^{\ast}_{0}}(x_{0},X_{0}) is at most k020n020\frac{\sqrt{k_{0}}}{2}\cdot\ell_{0}\leq\frac{\sqrt{n_{0}}}{2}\cdot\ell_{0}.
Now, we define the metric space X1:=Y0×(10k0)X_{1}:=Y_{0}\times\left(\frac{1}{\ell_{0}}\cdot\mathbb{R}^{k_{0}}\right). This is again a proper, geodesically complete, CAT(0)(0)-space, still (P0,r0)(P_{0},r_{0})-packed, on which Γ\Gamma still acts discretely by isometries (because the splitting of X0X_{0} is Γ\Gamma-invariant).We claim that the action of Γ\Gamma on X1X_{1} is D1D_{1}-cocompact. In fact, let x=(y,𝐯)x=(y,{\bf v}) be a point of X1X_{1}. Since the action of Γ\Gamma on X0X_{0} is D0D_{0}-cocompact, we know that there exists gΓg\in\Gamma such that dX0(x,g(y0,𝐎))D0d_{X_{0}}\left(x,g\cdot(y_{0},{\bf O})\right)\leq D_{0}; moreover, as Γ¯ε0(x0,X)\overline{\Gamma}_{\varepsilon^{\ast}_{0}}(x_{0},X) preserves {y0}×k0\{y_{0}\}\times\mathbb{R}^{k_{0}}, we can compose gg with elements of this group in order to find g=(g1,g2)Γg^{\prime}=(g^{\prime}_{1},g^{\prime}_{2})\in\Gamma such that dX0(x,g(y0,𝐎))D0d_{X_{0}}(x,g^{\prime}\cdot(y_{0},{\bf O}))\leq D_{0} and 𝐯g2𝐎0n020\|{\bf v}-g^{\prime}_{2}\cdot{\bf O}\|_{0}\leq\frac{\sqrt{n_{0}}}{2}\cdot\ell_{0}. Therefore

dX1((x,g(y0,𝐎))\displaystyle d_{X_{1}}((x,g^{\prime}\cdot(y_{0},{\bf O})) dY0(y,g1y0)2+(10)2𝐯g2𝐎02\displaystyle\leq\sqrt{d_{Y_{0}}(y,g_{1}^{\prime}\cdot y_{0})^{2}+\left(\frac{1}{\ell_{0}}\right)^{2}\cdot\|{\bf v}-g^{\prime}_{2}\cdot{\bf O}\|_{0}^{2}}
D02+n04D0+n02=D12.\displaystyle\leq\sqrt{D_{0}^{2}+\frac{n_{0}}{4}}\leq D_{0}+\frac{\sqrt{n_{0}}}{2}=\frac{D_{1}}{2}.

As (y,𝐯)X1(y,{\bf v})\in X_{1} was arbitrary, we then deduce that X1=ΓB¯X1((y0,𝐎),D12),X_{1}=\Gamma\cdot\overline{B}_{X_{1}}\left((y_{0},{\bf O}),\frac{D_{1}}{2}\right), so Γ<Isom(X1)\Gamma<\text{Isom}(X_{1}) is D1D_{1}-cocompact. If now sys(Γ,X1)>σn0\textup{sys}^{\diamond}(\Gamma,X_{1})>\sigma_{n_{0}}, we stop the process and set X=X1X^{\prime}=X_{1}: this space has all the desired properties.
Otherwise, we have sys(Γ,X1)σn0<σ2=σP0,r0,D1(σ1)\textup{sys}^{\diamond}(\Gamma,X_{1})\leq\sigma_{n_{0}}<\sigma_{2}=\sigma_{P_{0},r_{0},D_{1}}(\sigma_{1}) and we can apply again Theorem 4.1, Proposition 4.3 and Proposition 4.5 to X1X_{1}, with ε=σ1\varepsilon=\sigma_{1}. Then, there exists ε1(σ2,σ1)\varepsilon^{\ast}_{1}\in(\sigma_{2},\sigma_{1}) such that the groups Γ¯ε1(x,X1)\overline{\Gamma}_{\varepsilon^{\ast}_{1}}(x,X_{1}) have rank k11k_{1}\geq 1 for every xX1x\in X_{1}, in particular rk(Γ¯ε1(x0,X1))=k1\text{rk}(\overline{\Gamma}_{\varepsilon^{\ast}_{1}}(x_{0},X_{1}))=k_{1}. Moreover, there exists a free abelian, finite index subgroup A1<Γ¯ε1(x0,X1)A_{1}<\overline{\Gamma}_{\varepsilon^{\ast}_{1}}(x_{0},X_{1}) of rank k1k_{1} which is commensurated in Γ\Gamma, the space X1=Bd-Min(A1)X_{1}=\textup{Bd-Min}(\partial A_{1}) splits isometrically and Γ\Gamma-invariantly as Y1×k1Y_{1}\times\mathbb{R}^{k_{1}}, and the trace at infinity Γ¯ε1(x0,X1)\partial\overline{\Gamma}_{\varepsilon^{\ast}_{1}}(x_{0},X_{1}) coincides with the boundary at infinity of all the sets {y}×k1\{y\}\times\mathbb{R}^{k_{1}}; and there exists a subset of Σ¯4J0ε1(x0,X1)\overline{\Sigma}_{4J_{0}\cdot\varepsilon^{\ast}_{1}}(x_{0},X_{1}) whose projection on Isom(k1)\textup{Isom}(\mathbb{R}^{k_{1}}) generates the maximal lattice ε1(x0,X1)\mathcal{L}_{\varepsilon^{\ast}_{1}}(x_{0},X_{1}) of pk1(Γ¯ε1(x0,X1))p_{\mathbb{R}^{k_{1}}}(\overline{\Gamma}_{\varepsilon^{\ast}_{1}}(x_{0},X_{1})). Therefore, we can find a shortest basis 1={𝐛𝟏𝟏,,𝐛𝐤𝟏𝟏}\mathcal{B}^{1}=\{\bf b_{1}^{1},\ldots,\bf b_{k_{1}}^{1}\} of ε1(x0,X1)\mathcal{L}_{\varepsilon^{\ast}_{1}}(x_{0},X_{1}) with lengths (with respect to the Euclidean norm 1\|\hskip 5.69054pt\|_{1} of k1\mathbb{R}^{k_{1}})

𝐛𝟏𝟏1𝐛𝐤𝟏𝟏1=λ(ε1(x0,X1))=:14J0σ1<ε0<1.\|{\bf b_{1}^{1}}\|_{1}\leq\ldots\leq\|{\bf b_{k_{1}}^{1}}\|_{1}=\lambda\left(\mathcal{L}_{\varepsilon^{\ast}_{1}}(x_{0},X_{1})\right)=:\ell_{1}\leq 4J_{0}\cdot\sigma_{1}<\varepsilon_{0}<1.

Observe that, as by construction the factor (10)2\left(\frac{1}{\ell_{0}}\right)^{2} is bigger than 11, we have

(9) Γ¯ε1(x0,X1)<Γ¯σ1(x0,X1)<Γ¯σ1(x0,X0)<Γ¯ε0(x0,X0),\overline{\Gamma}_{\varepsilon^{\ast}_{1}}(x_{0},X_{1})<\overline{\Gamma}_{\sigma_{1}}(x_{0},X_{1})<\overline{\Gamma}_{\sigma_{1}}(x_{0},X_{0})<\overline{\Gamma}_{\varepsilon^{\ast}_{0}}(x_{0},X_{0}),

so k1=rk(Γ¯ε1(x0,X1))k0=rk(Γ¯ε0(x0,X0))k_{1}=\text{rk}(\overline{\Gamma}_{\varepsilon^{\ast}_{1}}(x_{0},X_{1}))\leq k_{0}=\text{rk}(\overline{\Gamma}_{\varepsilon^{\ast}_{0}}(x_{0},X_{0})), and A1<A0A_{1}<A_{0}. In particular, the metric splittings of X1X_{1} as Y0×(10k0)Y_{0}\times\left(\frac{1}{\ell_{0}}{\mathbb{R}}^{k_{0}}\right) and Y1×k1Y_{1}\times{\mathbb{R}}^{k_{1}} determined, respectively, by the groups A0A_{0} and A1A_{1} satisfy k110k0{\mathbb{R}}^{k_{1}}\subset\frac{1}{\ell_{0}}{\mathbb{R}}^{k_{0}} and Y0Y1Y_{0}\subset Y_{1}, because of the second part of Proposition 4.5.
We will now show that k1k0k_{1}\lneq k_{0}. Actually, suppose that k1=k0=:kk_{1}=k_{0}=:k. Then, the groups A0A_{0} and A1A_{1} split the same Euclidean factor and Y0=Y1Y_{0}=Y_{1}. The lengths of the basis 1\mathcal{B}^{1} with respect to the Euclidean norm 0\|\hskip 5.69054pt\|_{0} of the Euclidean factor k0{\mathbb{R}}^{k_{0}} of X0X_{0} are 𝐛𝐢𝟏0=0𝐛𝐢𝟏1<0\|{\bf b_{i}^{1}}\|_{0}=\ell_{0}\cdot\|{\bf b_{i}^{1}}\|_{1}<\ell_{0} for every i=1,,ki=1,\ldots,k. But then, since ε1(x0,X1)<ε0(x0,X0)\mathcal{L}_{\varepsilon^{\ast}_{1}}(x_{0},X_{1})<\mathcal{L}_{\varepsilon^{\ast}_{0}}(x_{0},X_{0}) by (9), we would be able to find kk independent vectors of ε0(x0,X0)\mathcal{L}_{\varepsilon^{\ast}_{0}}(x_{0},X_{0}) of length less than its shortest generating radius, which is impossible. This shows that k1<k0k_{1}<k_{0}.
We can now define X2:=Y1×(11k1)X_{2}:=Y_{1}\!\times\left(\frac{1}{\ell_{1}}\!\cdot\mathbb{R}^{k_{1}}\right), on which Γ\Gamma acts faithfully, discretely by isometries, D2D_{2}-cocompactly, for D2=2D1+n0D_{2}=2D_{1}+\sqrt{n_{0}} computed as before. We can repeat this process to get a sequence of proper, geodesically complete, (P0,r0)(P_{0},r_{0})-packed, CAT(0)(0)-spaces XjX_{j} on which Γ\Gamma always acts faithfully, discretely and DjD_{j}-cocompactly by isometries. Moreover at each step either sys(Γ,Xj)σn0\textup{sys}^{\diamond}(\Gamma,X_{j})\geq\sigma_{n_{0}} or the εj\varepsilon^{\ast}_{j}-splitting rank kjk_{j} of XjX_{j} provided by Theorem 4.1 is strictly smaller than the εj1\varepsilon^{\ast}_{j-1}-splitting rank kj1k_{j-1} of Xj1X_{j-1}. Since the splitting rank of X0X_{0} is at most n0n_{0} there must exist j{1,,n0}j\in\{1,\ldots,n_{0}\} such that sys(Γ,Xj)σn0\textup{sys}^{\diamond}(\Gamma,X_{j})\geq\sigma_{n_{0}}. The proof then ends by setting X=XjX^{\prime}=X_{j}.
Clearly, we may take, explicitely, Δ0=2n0D0+(2n01)n0\Delta_{0}=2^{n_{0}}D_{0}+(2^{n_{0}}-1)\sqrt{n_{0}}.
Finally, observe that, by construction, XX^{\prime} is isometric to the initial space XX, and that the action of Γ\Gamma on XjX_{j} is nonsingular if and only if the action of Γ\Gamma on Xj1X_{j-1}, and by induction on XX, 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 Γ\Gamma

{id}A0A1Am\{\textup{id}\}\lneqq A_{0}\lneqq A_{1}\ldots\lneqq A_{m}

for some mn01m\leq n_{0}-1, such that:

  • (a)

    denoting by kjk_{j} the rank of AjA_{j}, then 1k0<k1<km1\leq k_{0}<k_{1}\ldots<k_{m};

  • (b)

    setting hj=kjkj1h_{j}=k_{j}-k_{j-1} then there is a corresponding isometric and Γ\Gamma-invariant splitting of XX as Y×h0×h1××hmY\times\mathbb{R}^{h_{0}}\times\mathbb{R}^{h_{1}}\times\cdots\times\mathbb{R}^{h_{m}}. Moreover, there exist 0<Lj<10<L_{j}<1 such that the natural action of Γ\Gamma on the space

    (10) X:=Y×(1L0h0)×(1L1h1)××(1Lmhm)X^{\prime}:=Y\times\left(\frac{1}{L_{0}}\cdot\mathbb{R}^{h_{0}}\right)\times\left(\frac{1}{L_{1}}\cdot\mathbb{R}^{h_{1}}\right)\times\cdots\times\left(\frac{1}{L_{m}}\cdot\mathbb{R}^{h_{m}}\right)

    is Δ0\Delta_{0}-cocompact with free-systole at least s0s_{0}.

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 Γ\Gamma under consideration acts discretely by isometries and Δ0\Delta_{0}-cocompactly on a proper, geodesically complete, (P0,r0)(P_{0},r_{0})-packed, CAT(0)(0)-space XX^{\prime} with sys(Γ,X)s0\textup{sys}^{\diamond}(\Gamma,X^{\prime})\geq s_{0}. The action is still nonsingular, so by Theorem 3.1 we can find xXx^{\prime}\in X^{\prime} with sys(Γ,x)b0s0\textup{sys}(\Gamma,x^{\prime})\geq b_{0}\cdot s_{0}, a positive lower bound depending only on P0,r0P_{0},r_{0} and D0D_{0}. Then, applying Proposition 5.1 with D0=Δ0=DD_{0}=\Delta_{0}=D and s=b0s0s=b_{0}\cdot s_{0}, we conclude the proof.

Remark 5.3.

The assumption that the groups Γ\Gamma act nonsingularly is necessary for Theorem A. Actually, in [BK90, Theorem 7.1], Bass and Kulkarni exhibit an infinite ascending family of discrete groups Γ1<Γ2<\Gamma_{1}<\Gamma_{2}<\cdots acting (singularly) on a regular tree XX with bounded valency, with same compact quotient, in particular with diam(Γj\X)2(\Gamma_{j}\backslash X)\leq 2 for all jj. Moreover these groups are lattices satisfying Vol(Γi\\X)0(\Gamma_{i}\backslash\backslash X)\rightarrow 0, where the volume of Γ\Gamma in XX is defined as

Vol(Γ\\X)=xΓ\X1|StabΓ(x)|.\text{Vol}(\Gamma\backslash\backslash X)=\sum_{x\in\Gamma\backslash X}\frac{1}{|\text{Stab}_{\Gamma}(x)|}.

This family contains infinitely many different groups, since the minimal order σi\sigma_{i} of a torsion subgroup in Γi\Gamma_{i} tends to infinity as the volume goes to zero (every torsion subgroup of Γi\Gamma_{i} stabilizes some point, as XX is CAT(0)(0), hence 1σi\frac{1}{\sigma_{i}} is the dominant term of the sum yielding Vol(Γi\\X)\text{Vol}(\Gamma_{i}\backslash\backslash X)).

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 Γ\Gamma is Δ0\Delta_{0}-cocompact and sys(Γ,X)s0\text{sys}^{\diamond}(\Gamma,X)\geq s_{0}. Let F<ΓF<\Gamma be a finite subgroup. It fixes a point x0Xx_{0}\in X (cp. [BH13, Corollary II.2.8]). Let Dx0\mathrm{D}_{x_{0}} be the Dirichlet domain of Γ\Gamma at x0x_{0}, which is clearly FF-invariant. By Theorem 3.1 there exists some point xDx0x\in\mathrm{D}_{x_{0}} where sys(Γ,x)b0s0>0\text{sys}(\Gamma,x)\geq b_{0}\cdot s_{0}>0. Therefore the orbit {fx}fF\{fx\}_{f\in F} is a b0s02\frac{b_{0}\cdot s_{0}}{2}-separated subset of Dx0\mathrm{D}_{x_{0}} and the balls B(fx,b0s04)B(fx,\frac{b_{0}\cdot s_{0}}{4}) are all disjoint. Using the fact that Dx0\mathrm{D}_{x_{0}} is contained in B¯(x0,Δ0)\overline{B}(x_{0},\Delta_{0}) we get

V(Δ0)μX(Dx0)#Fv(b0s04),V(\Delta_{0})\geq\mu_{X}(\mathrm{D}_{x_{0}})\geq\#F\cdot v\left(\frac{b_{0}\cdot s_{0}}{4}\right),

by Proposition 2.3.(iii). This yields the explicit bound #FV(Δ0)v(b0s04)\#F\leq\frac{V(\Delta_{0})}{v\left(\frac{b_{0}\cdot s_{0}}{4}\right)} for the cardinality of FF, only depending only on P0,r0P_{0},r_{0} and D0D_{0}. ∎

5.2. Finiteness up to equivariant homotopy equivalence


We have proved so far that the number of discrete, nonsingular and D0D_{0}-cocompact groups of isometries Γ\Gamma of proper, geodesically complete, (P0,r0)(P_{0},r_{0})-packed, CAT(0)(0)-spaces XX 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 φ:ΓΓ\varphi:\Gamma\rightarrow\Gamma^{\prime}, acting discretely, nonsingularly and cocompactly by isometries respectively on CAT(0)(0)-spaces XX and XX^{\prime}, to a homotopy equivalence F:XXF:X\rightarrow X^{\prime} which is φ\varphi-equivariant, i.e. F(gx)=φ(g)F(x)F(g\cdot x)=\varphi(g)\cdot F(x) for all gΓg\in\Gamma and all xXx\in X. Namely we show:

Proposition 5.4.

Let Γ\Gamma and Γ\Gamma^{\prime} be discrete, cocompact, nonsingular isometry groups of two proper, geodesically complete, (P0,r0)(P_{0},r_{0})-packed, CAT(0)\textup{CAT}(0)-spaces XX and XX^{\prime}, respectively. If there exists a group isomorphism φ:ΓΓ\varphi:\Gamma\rightarrow\Gamma^{\prime}, then there exists a φ\varphi-equivariant homotopy equivalence f:XXf:X\rightarrow X^{\prime}.

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 XX be a proper, CAT(0)(0)-metric space. Consider the space 2(X){\mathcal{M}}_{2}(X) of positive, finite, Borel measures μ\mu on XX with finite second moment, i.e. such that the distance function d(x,)d(x,\cdot) is μ\mu-square integrable for some (hence every) xXx\in X. The barycenter of μ2(X)\mu\in{\mathcal{M}_{2}}(X) is defined as the unique point of minimum of the function

μ(x)=Xd(x,y)2𝑑μ(y).\mathcal{B}_{\mu}(x)=\int_{X}d(x,y)^{2}\,d\mu(y).

Notice that μ(x)\mathcal{B}_{\mu}(x) tends to ++\infty for xx\rightarrow\infty since for every fixed x0x_{0} in XX we have, by triangular and Schwarz inequalities,

μ(x)d(x,x0)2μ(X)2d(x,x0)μ(X)μ(x0)+μ(x0)\mathcal{B}_{\mu}(x)\geq d(x,x_{0})^{2}\mu(X)-2d(x,x_{0})\mu(X)\mathcal{B}_{\mu}(x_{0})+\mathcal{B}_{\mu}(x_{0})

which diverges as d(x,x0)+d(x,x_{0})\rightarrow+\infty. Moreover, by standard comparison with the Euclidean space, the function d(x,)2d(x,\cdot)^{2} is also strictly convex, namely 22-convex in the sense of [Kle99]: that is, d(x,c(t))t2d(x,c(t))-t^{2} is a convex function, for every geodesic c:[a,b]Xc:[a,b]\rightarrow X. It follows that the function μ\mathcal{B}_{\mu} is 2μ(X)2\mu(X)-convex as well, which implies that μ\mathcal{B}_{\mu} has a unique point of minimum (cp. [Kle99, Lemma 2.3]). Therefore bar[μ]:=argmin(μ)\texttt{bar}[\mu]:=\text{arg}\,\text{min}(\mathcal{B}_{\mu}) is well-defined.
It is straightforward to check that the barycenter is equivariant with respect to the natural actions of Isom(X)\text{Isom}(X) on XX and 2(X){\mathcal{M}}_{2}(X), that is:

(11) bar[gμ]=gbar[μ],gIsom(X),μ2(X)\texttt{bar}[g_{\ast}\mu]=g\cdot\texttt{bar}[\mu],\hskip 5.69054pt\forall g\in\text{Isom}(X),\forall\mu\in{\mathcal{M}}_{2}(X)

We can now give the Proof of Proposition 5.4:

Proof of Proposition 5.4..

As (Γ,X)(\Gamma,X) and (Γ,X)(\Gamma^{\prime},X^{\prime}) are nonsingular, we can choose x0Xx_{0}\in X and x0Xx_{0}^{\prime}\in X^{\prime} such that the stabilizers StabΓ(x0)\text{Stab}_{\Gamma}(x_{0}), StabΓ(x0)\text{Stab}_{\Gamma^{\prime}}(x_{0}^{\prime}) are trivial. Then there exists a unique φ\varphi-equivariant map ϕ:Γx0Γx0\phi:\Gamma x_{0}\rightarrow\Gamma^{\prime}x_{0}^{\prime} sending x0x_{0} to x0x_{0}^{\prime}, namely ϕ(gx0)=φ(g)x0\phi(g\cdot x_{0})=\varphi(g)\cdot x_{0}^{\prime}. Now, by Sˇ\check{\textup{S}}varc-Milnor Lemma, the spaces XX and XX^{\prime} are respectively quasi-isometric to the orbits Γx0\Gamma x_{0} and Γx0\Gamma^{\prime}x_{0}^{\prime} and also to the marked groups (Γ,Σ2D(x0))(\Gamma,\Sigma_{2D}(x_{0})) and (Γ,Σ2D(x0))(\Gamma^{\prime},\Sigma_{2D}(x_{0}^{\prime})) endowed with their word metrics. Moreover, since the groups Γ\Gamma and Γ\Gamma^{\prime} are isomorphic, and any two word metrics associated to finite generating sets on the same group are equivalent, we deduce that the map ϕ\phi is an (a,b)(a,b)-quasi-isometry for suitable a,ba,b, in particular

(12) d(φ(g1)x0,φ(g2)x0)ad(g1x0,g2x0)+b for all g1,g2Γd(\varphi(g_{1})x_{0}^{\prime},\varphi(g_{2})x_{0}^{\prime})\leq ad(g_{1}x_{0},g_{2}x_{0})+b\hskip 14.22636pt\text{ for all }g_{1},g_{2}\in\Gamma

We now define a φ\varphi-equivariant homotopy equivalence f:XXf:X\rightarrow X^{\prime} as follows: choose any h>log(1+P0)r0h>\frac{\log(1+P_{0})}{r_{0}} (the upper bound of the entropy of XX given by Proposition 2.3.(iv) and consider the family of measures μx2(X)\mu_{x}\in{\mathcal{M}}_{2}(X^{\prime}), indexed by xXx\in X and supported by the orbit of x0x_{0}^{\prime}, given by

μx:=gΓehd(x,gx0)δφ(g)x0,\mu_{x}:=\sum_{g\in\Gamma}e^{-hd(x,gx_{0})}\delta_{\varphi(g)x_{0}^{\prime}},

where δφ(g)x0\delta_{\varphi(g)x_{0}^{\prime}} is the Dirac measure at φ(g)x0\varphi(g)x_{0}^{\prime}, and then define

f(x):=bar[μx].f(x):=\texttt{bar}[\mu_{x}].

Notice that the total mass of μx\mu_{x} coincides with the value of the Poincaré series PΓ(x,x0,s)=gΓesd(x,gx0)P_{\Gamma}(x,x_{0},s)=\sum_{g\in\Gamma}e^{-sd(x,gx_{0})} of the group Γ\Gamma acting on XX for s=hs=h, which is finite since hh is chosen greater than the critical exponent of the series (recall that Ent(X)\text{Ent}(X) equals the critical exponent of the group Γ\Gamma acting on XX, since the action is cocompact, see for instance [BCGS17] or [Cav22a, Proposition 5.7]). Moreover, the function d(x0,)d(x_{0}^{\prime},\cdot) is square summable with respect to μx\mu_{x}, since we have by (12)

Xd(x0,y)2𝑑μx(y)\displaystyle\int_{X^{\prime}}d(x_{0}^{\prime},y)^{2}d\mu_{x}(y) =gΓd(x0,φ(g)x0)2ehd(x,gx0)\displaystyle=\sum_{g\in\Gamma}d(x_{0}^{\prime},\varphi(g)x_{0}^{\prime})^{2}e^{-hd(x,gx_{0})}
gΓ(ad(x0,gx0)+b)2ehd(x,gx0)\displaystyle\leq\sum_{g\in\Gamma}(a\cdot d(x_{0},gx_{0})+b)^{2}e^{-hd(x,gx_{0})}

which is finite as the Poincaré series converges exponentially fast for s=hs=h.
Therefore the map ff is well-defined.

Step 1: ff is continuous.
First, we show that μxn\mathcal{B}_{\mu_{x_{n}}} converge uniformly on compacts to μx\mathcal{B}_{\mu_{x}} for xnxx_{n}\rightarrow x in XX. Actually, let KXK\subset X^{\prime} be a fixed compact subset. Since the action of Γ\Gamma^{\prime} on XX^{\prime} is discrete, the subset

S={gΓ|gB(x0,2D0)K}S=\{g^{\prime}\in\Gamma^{\prime}\hskip 2.84526pt|\hskip 2.84526ptg^{\prime}B(x_{0}^{\prime},2D_{0})\cap K\neq\emptyset\}

is finite. Then for every xKx^{\prime}\in K choose some gxΓg_{x^{\prime}}\in\Gamma with φ(gx)S\varphi(g_{x^{\prime}})\in S such that d(x,φ(gx)x0)2D0d(x^{\prime},\varphi(g_{x^{\prime}})x_{0}^{\prime})\leq 2D_{0}. From (12) and the triangular inequality we find that, for d(xn,x)<εd(x_{n},x)<\varepsilon, it holds

|μxn(x)μx(x)|gΓd(x,φ(g)x0)2|ehd(xn,gx0)ehd(x,gx0)|\left|\mathcal{B}_{\mu_{x_{n}}}(x^{\prime})-\mathcal{B}_{\mu_{x}}(x^{\prime})\right|\leq\sum_{g\in\Gamma}d(x^{\prime},\varphi(g)\cdot x_{0}^{\prime})^{2}\left|e^{-hd(x_{n},g\cdot x_{0})}-e^{-hd(x,g\cdot x_{0})}\right|\hskip 28.45274pt
h(gΓ(d(φ(gx)x0,φ(g)x0)+2D0)2eh(d(x,gx0)ε))d(xn,x)\hskip 42.67912pt\leq h\left(\sum_{g\in\Gamma}\left(d(\varphi(g_{x^{\prime}})x_{0}^{\prime},\varphi(g)x_{0}^{\prime})+2D_{0}\right)^{2}e^{-h\left(d(x,gx_{0})-\varepsilon\right)}\right)d(x_{n},x)
h(gΓ(ad(x0,gx0)+b)2eh(d(x,gxgx0)ε))d(xn,x)\leq h\left(\sum_{g\in\Gamma}\left(a\cdot d(x_{0},g\cdot x_{0})+b^{\prime}\right)^{2}e^{-h\left(d(x,g_{x^{\prime}}gx_{0})-\varepsilon\right)}\right)d(x_{n},x)
hehεehd(x,gxx)(gΓ(ad(x0,gx0)+b)2ehd(x,gx0))d(xn,x)\hskip 36.98857pt\leq he^{h\varepsilon}e^{hd(x,g_{x^{\prime}}x)}\left(\sum_{g\in\Gamma}\left(a\cdot d(x_{0},g\cdot x_{0})+b^{\prime}\right)^{2}e^{-hd(x,g\cdot x_{0})}\right)d(x_{n},x)

for b=b+2D0b^{\prime}=b+2D_{0}. Now, the series in parentheses is bounded above independently of nn and xx^{\prime}, while (for fixed xx) the term ehd(x,gxx)e^{hd(x,g_{x^{\prime}}\cdot x)} is uniformly bounded on KK since gxg_{x^{\prime}} belong to the finite subset φ1(S)\varphi^{-1}(S). The above estimate then implies that μxnμx\mathcal{B}_{\mu_{x_{n}}}\rightarrow\mathcal{B}_{\mu_{x}} uniformily on KK when xnxx_{n}\rightarrow x.
Secondly, we call m=minμxm=\min\mathcal{B}_{\mu_{x}} and show that there exists RR such that for every n0n\gg 0 the functions μxn\mathcal{B}_{\mu_{x_{n}}} are greater than 2m2m on XB¯(bar[μx],R)X^{\prime}\setminus\overline{B}(\texttt{bar}[\mu_{x}],R).
Actually, let RR be such that μx>2m\mathcal{B}_{\mu_{x}}>2m outside B(bar[μx],R/2)B(\texttt{bar}[\mu_{x}],R/2) (recall that μx\mathcal{B}_{\mu_{x}} is proper, since we showed that limxμ(x)=+\lim_{x^{\prime}\rightarrow\infty}\mathcal{B}_{\mu}(x^{\prime})=+\infty for all μ2(X)\mu\in{\mathcal{M}}_{2}(X)), and assume that we have infinitely many points xnx_{n}^{\prime} with d(bar[μx],xn)>Rd(\texttt{bar}[\mu_{x}],x_{n}^{\prime})>R such that μxn(xn)2m\mathcal{B}_{\mu_{x_{n}}}(x_{n}^{\prime})\leq 2m. Then, calling yn[bar[μx],xn]y^{\prime}_{n}\in[\texttt{bar}[\mu_{x}],x^{\prime}_{n}] the point at distance RR from bar[μx]\texttt{bar}[\mu_{x}], we would have, by convexity,

μxn(yn)max{μxn(bar[μx]),2m}.\mathcal{B}_{\mu_{x_{n}}}(y^{\prime}_{n})\leq\max\{\mathcal{B}_{\mu_{x_{n}}}(\texttt{bar}[\mu_{x}]),2m\}.

Since μxnμx\mathcal{B}_{\mu_{x_{n}}}\rightarrow\mathcal{B}_{\mu_{x}} uniformly on B¯(bar[μx],R)\overline{B}(\texttt{bar}[\mu_{x}],R) we would deduce that μx(yn)2m\mathcal{B}_{\mu_{x}}(y_{n}^{\prime})\leq 2m, which contradicts the choice of RR, since d(bar[μx],yn)>R/2d(\texttt{bar}[\mu_{x}],y_{n}^{\prime})>R/2.
Now, since μxn>2m\mathcal{B}_{\mu_{x_{n}}}>2m on XB¯(bar[μx],R)X^{\prime}\setminus\overline{B}(\texttt{bar}[\mu_{x}],R) for all n0n\gg 0, the uniform convergence μxnμx\mathcal{B}_{\mu_{x_{n}}}\rightarrow\mathcal{B}_{\mu_{x}} on B¯(bar[μx],R)\overline{B}(\texttt{bar}[\mu_{x}],R) implies that the sequence f(xn)=bar[μxn]f(x_{n})=\texttt{bar}[\mu_{x_{n}}] of (unique) minimum points of μxn\mathcal{B}_{\mu_{x_{n}}} converge to the (unique) minimum point f(x)=bar[μx]f(x)=\texttt{bar}[\mu_{x}] of μx\mathcal{B}_{\mu_{x}}. In fact, we have that bar[μxn]B¯(bar[μx],R)\texttt{bar}[\mu_{x_{n}}]\in\overline{B}(\texttt{bar}[\mu_{x}],R) for all n0n\gg 0, and

(13) μxn(bar[μx])μxn(bar[μxn])\mathcal{B}_{\mu_{x_{n}}}(\texttt{bar}[\mu_{x}])\geq\mathcal{B}_{\mu_{x_{n}}}(\texttt{bar}[\mu_{x_{n}}])

so if bar[μxn]\texttt{bar}[\mu_{x_{n}}] accumulates to a point bb_{\infty}, passing to the limit in (13) yields μx(bar[μx])μx(b)\mathcal{B}_{\mu_{x}}(\texttt{bar}[\mu_{x}])\geq\mathcal{B}_{\mu_{x}}(b_{\infty}). This implies that b=bar[μx]b_{\infty}=\texttt{bar}[\mu_{x}], by unicity of the minimum point of μx\mathcal{B}_{\mu_{x}}.

Step 2: ff is a φ\varphi-equivariant homotopy equivalence.
Firstly, the map ff is φ\varphi-equivariant, since for all gΓg\in\Gamma we have

φ(g)μx=γΓehd(x,γx0)δφ(gγ)x0=γΓehd(x,g1γx0)δφ(γ)x0=μgx\varphi(g)_{\!\ast}\,\mu_{x}=\sum_{\gamma\in\Gamma}e^{-hd(x,\gamma x_{0})}\delta_{\varphi(g\gamma)\cdot x_{0}^{\prime}}=\sum_{\gamma\in\Gamma}e^{-hd(x,g^{-1}\gamma x_{0})}\delta_{\varphi(\gamma)x_{0}^{\prime}}=\mu_{gx}

and by (11) we deduce f(gx)=bar[φ(g)μx]=φ(g)bar[μx]=φ(g)f(x)f(g\cdot x)=\texttt{bar}[\varphi(g)_{\ast}\mu_{x}]=\varphi(g)\cdot\texttt{bar}[\mu_{x}]=\varphi(g)\cdot f(x).
Then, from the inverse map ϕ1:Γx0Γx0\phi^{-1}:\Gamma^{\prime}x_{0}^{\prime}\rightarrow\Gamma x_{0} we can analogously construct a φ1\varphi^{-1}-equivariant, continuous map f:XXf^{\prime}:X^{\prime}\rightarrow X. Now consider the homotopy map H:X×[0,1]XH:X\times[0,1]\rightarrow X defined by the formula H(x,t)=tx+(1t)ff(x)H(x,t)=tx+(1-t)\,f^{\prime}\!\circ\!f(x) (where tx+(1t)ytx+(1-t)y denotes the point on the geodesic segment [x,y][x,y] at distance t/d(x,y)t/d(x,y) from xx). The composition fff^{\prime}\circ f is Γ\Gamma-equivariant, so we deduce that HH is a Γ\Gamma-equivariant homotopy between fff^{\prime}\circ f and idX\text{id}_{X}.
Similarly, one proves that the map H(x,t)=tx+(1t)ff(x)H^{\prime}(x^{\prime},t)=tx^{\prime}+(1-t)\,f\!\circ\!f^{\prime}(x) is a Γ\Gamma^{\prime}-equivariant homotopy between fff\circ f^{\prime} and idX\text{id}_{X^{\prime}}. ∎

5.3. Finiteness of nonpositively curved orbifolds


We restrict here our attention to CAT(0)(0)-orbispaces which are quotients, by a discrete isometry group Γ\Gamma, of a Hadamard manifold XX (that is, a complete, connected and simply connected, nonpositively curved Riemannian manifold) with pinched sectional curvature κ2k(X)0-\kappa^{2}\leq k(X)\leq 0: we call such a quotient M=Γ\XM=\Gamma\backslash X a nonpositively curved Riemannian orbifold with curvature k(M)κ2k(M)\geq-\kappa^{2}. Recall that, in this case, any discrete isometry group Γ\Gamma of XX 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 M=Γ\XM=\Gamma\backslash X, M=Γ\XM^{\prime}=\Gamma^{\prime}\backslash X^{\prime} is a diffeomorphism F:XXF:X\rightarrow X^{\prime} which is equivariant with respect to some group isomorphism φ:ΓΓ\varphi:\Gamma\rightarrow\Gamma^{\prime}, i.e. F(gx)=φ(g)F(x)F(g\cdot x)=\varphi(g)\cdot F(x) for all gΓg\in\Gamma and all xXx\in X.

Proof of Corollary C.

First notice that the splitting Theorem D is true in the manifold category. Actually, let X=n×X0X={\mathbb{R}}^{n}\times X_{0} be the De Rham decomposition of a Hadamard manifold XX, where X0X_{0} is the product of all non-Euclidean factors. The factor n{\mathbb{R}}^{n} coincides with the Euclidean factor splitted by XX as a CAT(0)(0)-space (since the decomposition of a finite dimensional geodesic space into flat and irreducible factors is invariant by isometries, cp. [FL08]). Then, the k{\mathbb{R}}^{k} factor splitted by Theorem D under an ε\varepsilon-collapsed action, with εσ0\varepsilon\leq\sigma_{0}, is isometrically immersed in the Euclidean de Rham factor n{\mathbb{R}}^{n} of the manifold decomposition of XX; hence it is CC^{\infty}-embedded as a submanifold, as well as its orthogonal complement mk{\mathbb{R}}^{m-k}.Then, also the factor YY of Theorem D is CC^{\infty}-embedded in XX, because Y=X0×mkY=X_{0}\times{\mathbb{R}}^{m-k}. From this, we deduce that also the renormalization Theorem E holds in the manifold category: that is, every splitting Yi×kiY_{i}\times{\mathbb{R}}^{k_{i}} considered in the proof is smooth, and the resulting decomposition (10) yields a smooth Riemannian structure on XX^{\prime} (such that X=XX^{\prime}=X as a differentiable manifold). Moreover, if κ2k(X)0-\kappa^{2}\leq k(X)\leq 0, the new Riemannian manifold XX^{\prime} satisfies the same curvature bounds (since the metric is dilated on the Euclidean factors by the constants 1Li>1\frac{1}{L_{i}}>1, as explained in Remark 5.2).
It then follows that every nn-dimensional Riemannian orbifold M=Γ\XM=\Gamma\backslash X with κ2k(M)0-\kappa^{2}\leq k(M)\leq 0 and diam(M)D0(M)\leq D_{0} is equivariantly diffeomorphic to a Riemannian orbifold M=Γ\XM^{\prime}=\Gamma\backslash X^{\prime} still satisfying κ2k(M)0-\kappa^{2}\leq k(M^{\prime})\leq 0, but with diameter bounded above by Δ0\Delta_{0} and with sys(Γ,X)s0{}^{\diamond}(\Gamma,X^{\prime})\geq s_{0}; where the constants Δ0\Delta_{0} and s0s_{0} only depend on D0D_{0} and on the packing constants (P0,r0)(P_{0},r_{0}) of XX (and then, ultimately, only on κ\kappa and on the dimension nn, by (6)).Moreover, by Theorem 3.1, the systole of Γ\Gamma acting on XX^{\prime} is bounded below by b0min{sys(Γ,X),ε0}b_{0}\cdot\min\{\text{sys}^{\diamond}(\Gamma,X^{\prime}),\varepsilon_{0}\} at some point xx^{\prime}, where again b0,ε0b_{0},\varepsilon_{0} only depend on (P0,r0)(P_{0},r_{0}). 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(0)(0)-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.