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

Rigidity of actions on metric spaces close to three dimensional manifolds

Noé Bárcenas  and  Manuel Sedano-Mendoza [email protected]
[email protected]
Centro de Ciencias Matemáticas. UNAM
Ap.Postal 61-3 Xangari. Morelia, Michoacán MEXICO 58089
Abstract.

In this paper we propose a metric variation on the C0C^{0}-version of the Zimmer program for three manifolds. After a reexamination of the isometry groups of geometric three-manifolds, we consider homomorphisms defined on higher rank lattices to them and establish a dichotomy betweeen finite image or infinite volume of the quotient. Along the way, we enumerate classification results for actions of finite groups on three manifolds where available, and we give an answer to a metric variation on topological versions of the Zimmer program for aspherical three-manifolds, as asked by Weinberger and Ye, which are based on the dichotomy stablished in this work and known topological rigidity phenomena for three manifolds. Using results by John Pardon and Galaz-García-Guijarro, the dichotomy for homomorphisms of higher rank lattices to isometry groups of three manifolds implies that a C0C^{0}-isometric version of the Zimmer program is also true for singular geodesic spaces closely related to three dimensional manifolds, namely three dimensional geometric orbifolds and Alexandrov spaces. A topological version of the Zimmer Program is seen to hold in dimension 3 for Alexandrov spaces using Pardon’s ideas.

Key words and phrases:
Zimmer Program, 3-manifolds, Alexandrov Spaces, Hilbert Smith Conjecture. 2020 AMS Classification: 57S05, 57S20, 22E40, 57M60, 53C23, 53C30.

1. Structure of the paper

2. Introduction

Zimmers program

The question on the nature of group homomorphisms ρ:ΓDiff(M)\rho:\Gamma\to{\rm Diff}(M), between a finitely generated group and the group of diffeomorphisms of a compact, nn-dimensional, smooth manifold is interesting in many contexts. Particularly, in a series of conjectures known as the Zimmer program [Zim87], [Fis11], [Fis20], concerning the question on whether the group homomorphism cannot have large image if the dimension of the manifold is small, relative to the rank of the group. More precisely, the Zimmer program deals with groups Γ\Gamma which are lattices in a semisimple algebraic Lie group of rank at least 2, which we will refer to in this introduction as higher rank lattices see 3. As an example of this, in the recent result [BFH20], it is found that a homomorphism ρ:SLk+1()Diff(M)\rho:SL_{k+1}(\mathbb{Z})\rightarrow{\rm Diff(M)} must have finite image when k>nk>n and k2k\geq 2, in this case the parameter kk is the rank of the Lie group SLk+1()SL_{k+1}(\mathbb{R}). This result is greatly generalized for other higher rank semisimple lattices on [BFH21]. Another instance of the Zimmer program is the complete characterization of the action in critical dimensions, c.f. Conjectures 1.2 and 1.3 in [Fis11]. For example, when the dimension hypothesis is modified in the previous context to n=k+1n=k+1, and the hypothesis that the action preserves a finite volume and an affine connection is added, then [Fis11, Theorem 6.9] tells us that the action is conjugated to the standard linear action of SLn()SL_{n}(\mathbb{Z}) on n/n\mathbb{R}^{n}/\mathbb{Z}^{n}.

The C0C^{0}-version of the Zimmer Program, as suggested in [Wei11], and [Ye20], [Ye19], asks roughly for changing the category of manifolds and morphisms in the Zimmer Program, from the smooth setting into a topological setting, that is, by considering a group homomorphism from a finitely generated group, and specifically a higher rank lattice, onto the group of homeomorphisms within a prescribed category (topological, smooth, piecewise linear, etc.). The following Conjecture is an example of a problem stated in this setting, found in [Ye20]:

Problem 2.1.

Any group action of SLk+1()SL_{k+1}(\mathbb{Z}), with k2k\geq 2, on a closed, aspherical nn-manifold by homeomorphisms factors trough a finite group if n<kn<k.

This variation of Zimmer Program is, as expected, much harder than the original one, and it has shown only small advances, such as the solution in the one-dimensional case by Hurtado and Deroin [DH20], and in the context of infinite homological actions on three-manifolds [FS00].

Variations of the problem

It is natural to explore analogous rigidity results outside of the differentiable category and into the category of metric spaces endowed with extra structure. As an example of this, in [Hae20] it is proved that any action of a higher rank lattice in a Gromov δ\delta-hyperbolic metric space must be elementary. Among many things, such result implies that any homomorphism ρ:ΓMod(S)\rho:\Gamma\rightarrow\textrm{Mod}(S) from a higher rank lattice onto the mapping class group of a compact surface is finite, a result first proved in [FM98].

The notion closer to manifolds for which we explore these rifidity results is that of Alexandrov spaces, which are metric spaces with a synthetic notion of curvature bounded from below. Alexandrov spaces include compact Riemannian manifolds and non-trivial modifications of them, such as orbifold quotients and Gromov-Hausdorff limits (see Section 3.2. In this paper we propose the following variation of Problem 2.1

Problem 2.2.

Let XX be an n-dimensional, compact, Alexandrov space. Does a group homomorphism

ρ:SLk+1()Homeo(X),\rho:SL_{k+1}(\mathbb{Z})\to{\rm Homeo}(X),

factor through a finite group if k>nk>n? Can we obtain a classification of such actions in the case k=nk=n or k=n+1k=n+1 if we ask the action to be isometric?

Of course, Problem 2.2 can be stated for general higher rank lattices with comparison bounds on the dimension and rank as in [Fis11, Conjecture 4.12].

Alexandrov spaces are rigid (in a sense analogous to Gromov’s rigid geometric structures [Fis11]) as their isometry groups are Lie groups with bounded dimension in terms of the dimension of the space [GGG13]. Moreover, 3-dimensional Alexandrov spaces which are sufficiently collapsed (c.f. section Alex), are in fact orbifolds over one of the eight Thurston geometries [GGGNnZ20]. These reduction phenomenae together with classification results for isometric actions on Thurston geometries, lead us to the following much more tractable problem:

Problem 2.3.

Let XX be a 33-dimensional, compact Alexandrov space. Characterize any homomorphism

ρ:ΓIso(X),\rho:\Gamma\to{\rm Iso}(X),

where Γ\Gamma is a higher rank lattice.

It is worth mentioning other rigidity results obtained for 33-dimensional Alexandrov spaces such as the proof of the Borel conjeture for sufficiently collapsed Alexandrov spaces [BNnZ21]. Finally, John Pardon’s proof of the Hilbert-Smith conjecture for three manifolds [Par13], can be extended to the singular case in the setting of Alexandrov spaces as it can be reduced to a local behaviour, leading to the following result:

Theorem 2.4.

If GG is a locally compact, topological group, acting faithfully on a three dimensional Alexandrov space by homeomorphisms, then GG is a Lie group.

This result lead to the natural generalization of the Hilbert-Smith conjecture, which simply would ask if Theorem 2.4 is valid for nn-dimensional Alexandrov spaces). A first approach to this generalizationis to extend the result of [RS97], proving Hilbert-Smith conjecture for Lipschitz actions, where the difficulty lies on the extension of Yang’s Theorem (on the increase of dimension in the quotient for pp-adic actions [Yan60] [BRW61]) to Alexandrov spaces.

Main results and related discussions

The main result of this paper, concerning Problem 2.3 is

Theorem 2.5.

Let X~\widetilde{X} be a simply connected, homogeneous 3-dimensional manifold and let HH be a discrete group of isometries of X~\widetilde{X}, such that X=X~/HX=\widetilde{X}/H has finite volume, then XX admits an infinite isometric action of a higher rank lattice ΓG\Gamma\subset G if and only if the group Iso(X)Iso(X) contains the group SO(3)SO(3). Moreover, the semisimple Lie group GG is isotypic of type SO(3)SO(3), the lattice is uniform and XX is a orbifold over either S3S^{3} or ×S2\mathbb{R}\times S^{2}.

Recall that an isotypic group of type SO(3)SO(3) is an algebraic group which is, up to finite covers and connected components, a product of copies of SO(3)SO(3) and SO(2,1)SO(2,1) (see Section 3). In [BNnZ21], a three dimensional Alexandrov space XX of curv1{\rm curv}\geq-1 is said to be sufficiently collapsed, if there exist D>0D>0 and ϵ>0\epsilon>0 such that the diameter of XX is less or equal to DD, and the volume is strictly less than ϵ\epsilon. We include as a corollary of the results here a classification of the discrete groups acting by isometries on a three dimensional Alexandrov space with a sufficiently collapsed quotient

Corollary 2.6.

Assume that a discrete group Γ\Gamma acts by isometries on the three dimensional Alexandrov space XX such that the quotient X/ΓX/\Gamma is sufficiently collapsed with parameters dd, and ϵ\epsilon. Then, Theorem 2.8, together with the geometrization of 33-dimensional Alexandrov spaces provide a classification of the possible such Γ\Gamma within the lattices in the isometry groups.

As an immediate consequence of this theorem, we get the following corollary in the spirit of the Zimmer’s problem

Corollary 2.7.

Let Γ\Gamma be a higher rank lattice acting by isometries on a finite volume, three dimensional orbifold XX (modelled over a homogeneous 3-manifold XX), then the action factors through a finite group if either:

  • XX is aspherical or,

  • Γ\Gamma is non-uniform.

As an example of this, we have Γ=SLr()\Gamma=SL_{r}(\mathbb{Z}) with r3r\geq 3.

The proof of Theorem 2.5 relies on close, case by case examination of Thurston’s 33-dimensional geometries, their finite volume quotients and their corresponding isometry groups. The computations of such groups can be summarized in the following

Theorem 2.8.

Let X~\widetilde{X} be a simply connected, homogeneous 3-dimensional manifold and let GG be a discrete group of isometries of X~\widetilde{X}, such that X~/G\widetilde{X}/G has finite volume. Then the isometry group Iso(X~/G)Iso(\widetilde{X}/G) has finitely many connected components, such that its connected component of the identity is isomorphic to

  • a closed subgroup of 𝕊1{\mathbb{S}}^{1}, if X~\widetilde{X} is either 2×{\mathbb{H}}^{2}\times{\mathbb{R}}, SL~2()\widetilde{SL}_{2}({\mathbb{R}}) or NilNil;

  • a closed subgroup of SO(3)×S1SO(3)\times S^{1}, if X~=S2×\widetilde{X}=S^{2}\times{\mathbb{R}};

  • a closed subgroup of 3/3{\mathbb{R}}^{3}/{\mathbb{Z}}^{3}, if X~=3\widetilde{X}={\mathbb{R}}^{3};

  • a closed subgroup of SO(4)SO(4), if X~=S3\widetilde{X}=S^{3}.

Moreover, Iso(X~/G)Iso(\widetilde{X}/G) is finite if X~\widetilde{X} is either 3{\mathbb{H}}^{3} or SolSol.

Strategy of the proof and structure of the paper.

Main Theorem 2.5 is proved in Section 3 using the classification of isometry groups of orbifolds given by Theorem 2.8, together with rigidity results of semisimple Lie groups. To prove Theorem 2.8, one first need to understand finite volume quotients of Thurston’s three-dimensional geometries. Among such geometries, the most homogeneous ones are S3S^{3}, 3{\mathbb{H}}^{3} and 3{\mathbb{R}}^{3}; and the remaining five present a more flexible description as fiber bundles

FXB,F\rightarrow X\rightarrow B,

where BB is a two dimensional homogeneous geometry for XX either 2×{\mathbb{H}}^{2}\times{\mathbb{R}}, SL~2()\widetilde{SL}_{2}({\mathbb{R}}) or Nil and B=B={\mathbb{R}} for XX either SolSol or S2×S^{2}\times{\mathbb{R}}. In this context, a discrete group acting on the homogeneous space XX, acts on the base space of the corresponding fiber bundle as well. The induced action on the base space of the fiber bundle presents a dual behaviour given by the following Theorem, whose proof can be derived from the discussions on [Thu97].

Theorem 2.9.

Let GG be a discrete group of isometries of any of the 3-dimensional geometric manifolds 2×{\mathbb{H}}^{2}\times{\mathbb{R}}, SL~2()\widetilde{SL}_{2}({\mathbb{R}}) or NilNil; then either

  • GG projects to a discrete group of isometries of the base BB of the fiber bundle, or

  • the orbifold X~/G\widetilde{X}/G has infinite volume.

Moreover, in the cases SolSol and S2×S^{2}\times{\mathbb{R}}, the projection to the base space is always discrete.

For the sake of completeness we present here a proof of Theorem 2.9. The proofs of Theorem 2.8 and Theorem 2.9 are carried out in a case by case setting on each Thurston geometry.

The structure of the paper is as follows: In Section 3 we present background material on three dimensional Alexandrov spaces, the Hilbert Smith conjecture and semisimple Lie groups and their lattices. Sections 5 through 13 cover the proof of Theorem 2.9 and Theorem 2.8 on each individual three dimensional geometry.

Theorem 2.9 can be used to obtain explicit characterizations of discrete groups acting on the corresponding three-dimensional geometry (in particular within the NilNil and SolSol cases), which lead to the proof of Theorem 2.8 in each case.

2.1. Concluding remarks and open questions

In this work we proposed a metric variation on the Zimmer program. The variation consisted in

  • Strengthening the category of automorphisms of the action from C0C^{0} to isometries.

  • Relaxing the topological type of the spaces considered from smooth manifolds to Alexandrov spaces, which include three dimensional geometric orbifolds.

While Alexandrov spaces have an open dense subset which is a topological manifold, results related to Zimmer’s conjeture do not apply directly because the topological manifold is open, and the only rigidity results for actions on open manifolds in the spirit of the Zimmer program which are known to the authors are restricted to the one dimensional case [DH20]. In another instance of a complication, the manifold is not Riemannian, as Otsu-Shioya’s example shows [OS94].

There exist several instances of families of homeomorphisms of metric geodesic spaces for which rigidity results of actions of discrete groups can be proved. Among them we can consider, for a strengthening with respect to homeomorphisms and a weakening with respect to isometries:

  • Quasiconformal homeomorphisms, as in the alternative proof of Mostow Rigidity Theorem by [Bou09].

  • Bilipschitz homeomorphisms, as in the proof of the Hilbert-Smith Conjecture mentioned before [RS97].

  • Quasimöbius homeomorphisms as in [BK02], where rigidity results for them have as a consequence the rigidity of actions of quasi convex cocompact actions on CAT(1){\rm CAT(-1)}-spaces.

Moreover, the specific analytic and geometric characteristics of the class of homeomorphisms are exploited in the process of proving an action rigidity result in an analogous way to how we used the geometric structure of Alexandrov three spaces in this work, inspired by the proofs of Zimmer program results in the diffeomorphism case.

Let us introduce the notation

AUTα(X){\rm AUT}^{\alpha}(X)

for homeomorphisms of a metric space with a metric property α\alpha and let us refer to a metric condition α\alpha as a decoration in analogy with surgery theory, having at least the following examples in mind:

  1. (i)

    The smooth case α=Diff\alpha={\rm Diff}, for diffeomorphisms of the smooth structure associated to a Riemannian metric on a smooth manifold with fixed metric.

  2. (ii)

    The topological case α=Top\alpha={\rm Top}, refering to homeomorphisms of a topological manifold.

  3. (iii)

    The isometric case Iso\rm Iso, meaning isometries of the geodesic metric space associated to a geometric three manifold, orbifold or Alexandrov space.

  4. (iv)

    The quasiconformal case α=QC\alpha=QC, the referring to quasiconformal homemomorphisms, α=QM\alpha=QM, associated to quasimöbius homeomorphisms and α=BI\alpha=BI for bilipschitz homemorphisms of a geodesic length metric space as discussed in the paragraph above.

We can consider the problem of describing the behaviour of a group homomorphism

ΛAUTα(X)\Lambda\longrightarrow{\rm AUT}^{\alpha}(X)

to a homeomorphism group with decorations as described. Notice that α=Diff\alpha={\rm Diff} is the Zimmer program as described for instance in [Fis11],[Fis20], [Zim84]. On the other hand α=Top\alpha={\rm Top} is the C0C^{0}-Zimmer program as described in [Wei11], [Ye20], [Ye19]. Finally α=Iso\alpha={\rm Iso} was the point of view adopted in this note, and α=QC,QM,BI\alpha=QC,QM,BI are as before.

We would like to finish the present note with the following two questions:

  • To what extent a condition of prescribed curvature in metric spaces, such as the Alexandrov condition, or a choice of QCQC, QMQM or BIBI-structures can be seen as a rigid structure, in the sense of Gromov [Fis11]?

  • For which decorations of homeomorphisms α\alpha is it possible to prove that homeomorphisms of a higher rank lattice Λ\Lambda with respect to the dimension of an Alexandrov space XX

    ΛAUTα(X)\Lambda\longrightarrow{\rm AUT}^{\alpha}(X)

    or in general a metric measure space XX with finite Hausdorff dimension less than the rank of Λ\Lambda either factorize trough a finite quotient of Λ\Lambda or produce a quotient of infinite volume?

Acknowledgments

The first author thanks DGAPA Project IN100221. Parts of this project were written during a Sabbatical Stay at the Universität des Saarlandes, with support of DGAPA-UNAM Sabbatical Program and the SFB TRR 195 Symbolic Tools in Mathematics and their Application, and CONACYT trough grant CF 217392.

The second named author thanks CONACYT Grant CB2016-283988-F, CONACYT Grant CB2016-283960, as well as a DGAPA-UNAM Postdoctoral Scholarship.

3. Preliminaries

3.1. Three Dimensional Manifolds

Recall that a closed three dimensional manifold MM is prime if it cannot be written as a connected sum with summands not homeomorphic to the three dimensional sphere.

By the prime decomposition Theorem, 3.15 page 31 in [Hem04], any closed three dimensional manifold can be written as a connected sum of prime factors.

Recall that a model geometry is a simply connected smooth three manifold XX together with a transitive action of a Lie group GG on XX with compact stabilizers such that GG is maximal among groups acting smoothly and transitively on XX with compact stabilizers.

A geometric structure on a three dimensional manifold MM is a diffeomorphism from MM to XΓX\diagup\Gamma for some model geometry XX, where Γ\Gamma is a discrete subgroup of GG acting freely on X ;

It is a consequence of Thurston geometrization theorem due to Perelman, that any prime closed three dimensional manifold can be cut along three dimensional tori such that the interiors of the resulting manifold carries a geometric structure of finite volume.

The eight three dimensional geometries which admit at least one compact three manifold are :

  1. (i)

    Spherical, where the three dimensional sphere is a representative.

  2. (ii)

    Euclidean, where the flat three dimensional torus is an example,

  3. (iii)

    Hyperbolic, where the three dimensional hyperbolic space is an example.

  4. (iv)

    SL~2()\widetilde{SL}_{2}(\mathbb{R}), where an example is given by the unit tangent bundle in a Riemannian metric of the tangent bundle over a genus two surface.

  5. (v)

    NilNil, where an example is given by the mapping torus of a Dehn twist on the two dimensional torus.

  6. (vi)

    SolSol, where an example is given by a manifold which fibers over the line with fiber the plane.

  7. (vii)

    S2×S^{2}\times\mathbb{R}.

  8. (viii)

    S1×2S^{1}\times\mathbb{H}^{2}.

A closed 3-manifold has a geometric structure of at most one of the 8 types above.

Remark 3.1.

(Non-uniqueness of Geometric Structures on three manifolds)

  • Finite volume non-compact manifolds may have more than one type of geometric structure. An example is the complement of the trefoil knot, which has hyperbolic structure and Sl~2\widetilde{Sl}_{2} structure.

  • If the three manifold has infinite volume, it might carry many geometric structures, for example 3\mathbb{R}^{3} is diffeomorphic to all aspherical model geometries.

  • There exists an infinite number of geometric structures with no compact models; for example, the geometry of many non-unimodular 3-dimensional Lie groups, see remark 9.1 below.

3.2. Three dimensional Alexandrov Spaces

We will need some preliminaries on three dimensional Alexandrov spaces. For a general reference see [BBI01].

For the purposes of this work, Alexandrov spaces are well behaved metric spaces which have three main properties that we want to highlight here:

  1. (i)

    They have an open dense subset which is topological manifold.

  2. (ii)

    Their isometry groups are Lie groups.

  3. (iii)

    The class of Alexandrov spaces is stable under Gromov-Hausdorff convergence.

  4. (iv)

    They include orbifolds over Riemannian manifolds.

In slightly more detail, Alexandrov spaces are a synthetic generalization of complete Riemannian manifolds with a lower bound on sectional curvature. The generalization uses comparison triangles with respect to the model spaces Sk2S_{k}^{2}, which are simply connected, two dimensional complete Riemannian manifolds of constant curvature kk. More precisely, for k>0k>0, Sk2S_{k}^{2} is the sphere of radius 1k\frac{1}{\sqrt{k}}, for k<0k<0, Sk2S_{k}^{2} is the hyperbolic plane 2(1k)\mathbb{H}^{2}(\frac{1}{\sqrt{-k}}) of constant curvature kk, and for k=0k=0, Sk2S_{k}^{2} is the euclidean space 2\mathbb{R}^{2}.

Given a geodesic triangle in a geodesic length space (X,d)(X,d), with vertices p,q,rXp,q,r\in X, a comparison triangle in Sk2S_{k}^{2} is a geodesic triangle p¯q¯r¯\bar{p}\bar{q}\bar{r} having the same side lengths. The geodesic length space (X,d)(X,d) is said to satisfy the Topogonov property for kk\in\mathbb{R}, if for each triple p,q,rXp,q,r\in X of vertices of a geodesic triangle, and each point ss on the geodesic from qq to rr, the inequality d(p,s)d(p¯,s¯)d(p,s)\geq d(\bar{p},\bar{s}) holds, where s¯\bar{s} is the point on the geodesic side q¯r¯\bar{q}\bar{r} of the comparison triangle with d(p¯s¯)=d(p,s)d(\bar{p}\bar{s})=d(p,s).

Definition 3.2.

A nn-dimensional kk-Alexandrov space is a complete, locally compact, length space of finite Hausdorff dimension nn, such that the Topogonov Property is satisfied locally for kk.

Topogonov’s globalization theorem tells us that the local and global Toponogov property are equivalent in kk-Alexandrov spaces. By Gromov’s precompactness theorem, Alexandrov nn-dimensional spaces arise as Gromov-Hausdorf limits of compact riemannnian manifolds of dimension nn for which the sectional curvature is bounded below by kk, and the diameter is bounded above by some fixed positive number DD.

The class of kk-Alexandrov spaces includes riemannian manifolds of sectional curvature bounded below by kk, and several constructions including more general geodesic length spaces such as euclidean cones, suspensions, joins, quotients by isometric actions of compact Lie groups, and glueings along a submetry, see [GG16] section 2.2. From now on, we will omit the kk from the notation.

There exists a notion of angle between geodesics of an Alexandrov space, and a space of tangent directions at a given point pp, denoted by Σp\Sigma_{p}, can be defined as the completion of the metric space of equivalence classes of geodesics making a zero angle.

The space of tangent directions at a point pp in an Alexandrov space XX, denoted by Σp\Sigma_{p}, has the structure of a 11-Alexandrov space of Hausdorf dimension dim(X)1{\rm dim}(X)-1. There is a set RXXR_{X}\subset X, called the set of metrically regular points, where a point pp belongs to RXR_{X} if its direction space Σp\Sigma_{p} is isometric to the radius one sphere. The complement is called the set of metrically singular points and denoted by SX=XRXS_{X}=X\setminus R_{X}. There are examples of Alexandrov spaces whose space of metrically singular points is dense, as seen in an example constructed in [OS94] as a limit of Alexandrov spaces, using baricentric subdivisions of a tetrahedron. However, for every Alexandrov space XX, there is a dense subset of topologically regular points, whose space of directions are homeomorphic to a sphere (the set of topologically singular points is the complement of the set of topologically regular points). By Perelman’s conical neighborhood theorem, every point pp in an Alexandrov space has a neighborhood pointed homeomorphic to the euclidean cone over Σp\Sigma_{p}, so that a locally compact, finite dimensional Alexandrov space has a dense subset which is a topological manifold.

In the specific case of dimension three, there are only two possibilities for the homeomorphic type of the space of directions, which is the two sphere S2S^{2}, for the topologically regular points and the real projective space 2\mathbb{RP}^{2} for the topologically singular points. Let us summarize the basic structure of three dimensional Alexandrov spaces due to Galaz-García and Guijarro, compare Theorem 1.1 in page 5561 of [GGG15], and Theorem 3.1 and 3.2 in page 1196 of [GGG13]. See also [HS17].

Theorem 3.3.

Let XX be a three dimensional Alexandrov space.

  • The set of metrically regular points is a Riemannian three manifold.

  • The set of topologically singular points is a discrete subset of XX.

  • If XX is closed, and positively curved Alexandrov space, that contains a topologically singular point, then XX is homeomorphic to the suspension of P2\mathbb{R}P^{2}.

A closed Alexandrov space is geometric if it can be written as a quotient of one of the eight geometries of Thurston under a cocompact lattice. The following theorem was proved as Theorem 1.6 in [GGG15] in page 5563. See also [HS17]

Theorem 3.4.

A three dimensional Alexandrov space admits a geometric decomposition into geometric pieces, along spheres, projective planes, tori and Klein bottles.

We now direct our attention to the isometry group of three dimensional Alexandrov spaces.

Theorem 3.5.

Let XX be an nn-dimensional Alexandrov space of Hausdorff dimension nn.

  • The Isometry group of XX is a Lie Group which is compact if XX is compact as well.

  • The dimension of the group of Isometries of XX is at most

    n(n+1)2,\frac{n(n+1)}{2},

    and the bound is attained if and only if XX is a Riemannian manifold.

Proof.
  • The first part is proved as the main Theorem, 1.1 in [FY94]. The second part follows from the Van Dantzig-Van der Waerden Theorem [DVdW28].

  • This is proved as Theorem 3.1 in page 570.

Remark 3.6.

It is proved in [BZ07] that the same lower bound for the dimension of the isometry group holds in general for Riemannian orbifolds.

3.3. Hilbert-Smith Conjecture

The following conjecture was formulated as an extension of Hilbert’s 5th Problem:

Conjecture 3.7 (Hilbert-Smith conjecture).

If GG is a locally compact, topological group, acting faithfully on a topological manifold, then GG is a Lie group.

See [Tao14] for a modern account.

As a consequence of structural theorems of locally compact groups, such as the Gleason-Yamabe theorem and its predecessor by Von Neumann [vN33]. a counter-example to the Hilbert-Smith conjecture must contain a copy of a p-adic group ^p\widehat{\mathbb{Z}}_{p}, for some pp, see [Lee97], thus giving the equivalent conjecture

Conjecture 3.8 (Hilbert-Smith conjecture pp-adic version).

For every prime pp, there are no faithful actions of the pp-adic group ^p\widehat{\mathbb{Z}}_{p} on a topological manifold.

Conjecture 3.8 has been proven in different contexts. For example, if there is a notion of dimension which must be preserved, such as bi-Lipschitz actions of ^p\widehat{\mathbb{Z}}_{p} on Riemannian manifolds, where three notions of dimension coincide: Hausdorff dimension, cohomological dimension with integer coefficients and topological dimension. In such setting, the bi-Lipschitz condition tells us that the Hausdorff dimension on the quotient cannot decrease, but on the other hand a theorem by Yang [Yan60], tells us that the cohomological dimension of the quotient increases by two, leading to the following result:

Theorem 3.9 (Repovš-Ščepin [RS97]).

There are no faithful actions by bi-Lipschitz maps of the pp-adic group ^p\widehat{\mathbb{Z}}_{p} on a Riemannian manifold.

The stronger setting of topological actions is much harder and has been proven only for small dimensions

Theorem 3.10 ([Par19], [Par13]).

For every prime pp, there are no faithful actions by homeomorphisms of the pp-adic group ^p\widehat{\mathbb{Z}}_{p} on a topological manifold of dimension n3n\leq 3.

Remark 3.11.

The pp-adic group can be described as

^p={n=0anpn:an{0,1,,p1},}\widehat{\mathbb{Z}}_{p}=\left\{\sum_{n=0}^{\infty}a_{n}p^{n}:a_{n}\in\{0,1,\cdots,p-1\},\right\}

so that pk^p^pp^{k}\widehat{\mathbb{Z}}_{p}\subset\widehat{\mathbb{Z}}_{p} is an open, normal subgroup, with ^p/pk^ppk\widehat{\mathbb{Z}}_{p}/p^{k}\widehat{\mathbb{Z}}_{p}\cong\mathbb{Z}_{p^{k}\mathbb{Z}}, giving the inverse limit description ^p=limpk\widehat{\mathbb{Z}}_{p}=\lim_{\leftarrow}\mathbb{Z}_{p^{k}}, moreover, the group ^p\widehat{\mathbb{Z}}_{p} is homeomorphic to the Cantor space {0,,p1}\{0,\cdots,p-1\}^{\mathbb{N}}. Observe that there is a topological 22-manifold with the cantor space 22^{\mathbb{N}} as its ends space, which is Σ=S2C\Sigma=S^{2}\setminus C, where CS2C\subset S^{2} is a closed subset homeomorphic to 22^{\mathbb{N}}. Thus, there is a faithful action of ^2\widehat{\mathbb{Z}}_{2} on End(Σ)2End(\Sigma)\cong 2^{\mathbb{N}} and every homeomorphism of End(Σ)End(\Sigma) extends to a homeomorphism of the surface Σ\Sigma, however, by Theorem 3.10, such extensions cannot be promoted to an action of ^2\widehat{\mathbb{Z}}_{2} on the Freudenthal compactification.

Hence, the weaker version of the pp-adic Hilbert-Smith conjecture for Alexandrov spaces holds, and we can consider the following conjecture:

Conjecture 3.12.

If GG is a locally compact, topological group, acting faithfully on a finite dimensional Alexandrov space by homeomorphisms, then GG is a Lie group.

A consequence of Theorem 3.10, gives us a three dimensional case of this result 14.4

3.4. Lie Groups and Lattices

A real Lie group is a Hausdorff topological group which is a CC^{\infty} smooth manifold for which the multiplication and inversion are smooth maps.

Recall that by Haar’s Theorem there exists up to a positive multiplicative constant, a unique countably additive, nontrivial measure μ\mu on the Borel subsets of GG which is right translation invariant, has finite values on compact subsets, and is inner and outer regular.

Definition 3.13.

A lattice in a Lie Group GG is a discrete group Λ\Lambda for which the quotient space G/ΛG/\Lambda has finite measure.

The right translate μ(g1)\mu(g^{-1}\quad)by an element in the group g1g^{-1} of a right invariant Haar measure μ\mu is a right invariant Haar measure, and hence there exists a real function Δ\Delta satisfying

μ(g1S)=Δ(g)μ(S).\mu(g^{-1}S)=\Delta(g)\mu(S).
Definition 3.14.

A group is said to be unimodular if the function Δ\Delta is the constant function 11.

Example 3.15.

The following families of examples of Lie groups will be the main focus of the article.

  • By the second Myers-Steenrod theorem [MS39], the isometry group of a smooth manifold is a Lie group.

  • By the Montgomery-Zippin theorem [MZ74], if a topological group acts by isometries transitively on a finite dimensional, locally compact, connected and locally connected metric space, then it is a Lie group.

  • By results of Bochner [Boc46], the isometry groups of a smooth manifold of constant negative Ricci curvature is finite.

    On the other hand, a locally compact subgroup of C2C^{2} diffeomorphisms of a C2C^{2} manifold for which the trivial subgroup is the only subgroup with fixed points with nonempty interior must be a Lie group [BM46].

In the subsequent sections of this article, we will examine the isometry groups of three manifolds and their lattices, as well as the isometry groups of orbifolds or Alexandrov spaces.

3.5. Discrete groups of isometries

If X~\widetilde{X} is a complete, simply connected, Riemannian manifold and ΓIso(X~)\Gamma\subset Iso(\widetilde{X}) a discrete subgroup of isometries, then X/ΓX/\Gamma has the structure of a complete Riemannian orbifold. The covering map ρ:X~X\rho:\widetilde{X}\rightarrow X satisfies the property that ρ(x)=ρ(y)\rho(x)=\rho(y) if and only if γx=y\gamma x=y for some γΓ\gamma\in\Gamma. An isometry ϕ:XX\phi:X\rightarrow X lifts to ϕ~:X~X~\widetilde{\phi}:\widetilde{X}\rightarrow\widetilde{X} such that ρϕ~=ϕρ\rho\circ\widetilde{\phi}=\phi\circ\rho and for every γΓ\gamma\in\Gamma and xX~x\in\widetilde{X} we have

ρϕ~(γx)=ϕρ(γx)=ϕρ(x)=ρϕ~(x),\rho\circ\widetilde{\phi}(\gamma x)=\phi\circ\rho(\gamma x)=\phi\circ\rho(x)=\rho\circ\widetilde{\phi}(x),

thus there exist γΓ\gamma^{\prime}\in\Gamma such that

ϕ~(γx)=γϕ~(x),\widetilde{\phi}(\gamma x)=\gamma^{\prime}\widetilde{\phi}(x),

that is ϕ~γϕ~1=γ\widetilde{\phi}\circ\gamma\circ\widetilde{\phi}^{-1}=\gamma^{\prime} and ϕ~N=NIso(X~)(Γ)\widetilde{\phi}\in N=N_{Iso(\widetilde{X})}(\Gamma). This tells us that we have the isomorphism

Iso(X)NIso(X~)(Γ)/Γ.Iso(X)\cong N_{Iso(\widetilde{X})}(\Gamma)/\Gamma.
Proposition 3.16.

If GG is a Lie group and ΓG\Gamma\subset G is a discrete subgroup with associated normalizer and centralizer subgroups

N=NG(Γ),Z=ZG(Γ),N=N_{G}(\Gamma),\qquad Z=Z_{G}(\Gamma),

then the connected components of NN and ZZ coincide. Moreover, if Z0Z_{0} denotes such connected component, the projection π:NN/Γ\pi:N\rightarrow N/\Gamma is a covering Lie group homomorphism such that π(Z0)N/Γ\pi(Z_{0})\subset N/\Gamma is the connected component of the identity.

Proof.

If gtNg_{t}\in N is a 1-parameter subgroup and γΓ\gamma\in\Gamma, then gtγgt=γtg_{t}\gamma g_{-t}=\gamma_{t} is a 1-parameter group in Γ\Gamma, but as Γ\Gamma is discrete, γt=γ\gamma_{t}=\gamma and this tells us that gtZg_{t}\in Z, so that Z0=N0Z_{0}=N_{0}. Now, NN is a Lie group having Γ\Gamma as a normal, discrete subgroup so that the projection map

π:NN/Γ\pi:N\rightarrow N/\Gamma

is a homomorphism of Lie groups and a covering map. In particular, π(N0)\pi(N_{0}) is a connected, open Lie subgroup of the same dimension of N/ΓN/\Gamma and thus it is the connected component of the identity. ∎

3.6. Lattices on semisimple Lie groups of higher rank

Recall that an algebraic \mathbb{R}-group is a subgroup 𝔾GLm()\mathbb{G}_{\mathbb{C}}\subset GL_{m}(\mathbb{C}) obtained as solutions of polynomial equations with coefficients over \mathbb{R} and 𝔾=𝔾GLm()\mathbb{G}_{\mathbb{R}}=\mathbb{G}_{\mathbb{C}}\cap GL_{m}(\mathbb{R}) is a real Lie group. In this context we say that 𝔾\mathbb{G}_{\mathbb{R}} is a real form of 𝔾\mathbb{G}_{\mathbb{C}} or that 𝔾\mathbb{G}_{\mathbb{C}} is a complexification of 𝔾\mathbb{G}_{\mathbb{R}}. The local structure of a Lie group is captured by its Lie algebra, so that two groups are locally isomorphic if and only if they have isomorphic Lie algebras, and thus, they can be obtained one from the other by taking connected components and topological covers.

The class of semisimple Lie groups can be defined as the class of Lie groups which are constructed up to covers and connected components from algebraic \mathbb{R}-groups which split as products G1××GkG_{1}\times\cdots\times G_{k}, where each factor GjG_{j} is simple. This definition is equivalent to other definitions of semisimple Lie groups aviailable in the literature, see [Zim84].

Remark 3.17.

Not every semisimple Lie group is an algebraic group as the group SL2()SL_{2}(\mathbb{R}) has a universal cover, denoted by SL~2()\widetilde{SL}_{2}(\mathbb{R}), which is homeomorphic to 3\mathbb{R}^{3} and it cannot be embedded in any linear group GLm()GL_{m}(\mathbb{C}) as a Lie subgroup. In the same way, not every semisimple Lie group splits as a product of simple Lie groups, as the example SO(4)SO(4) shows, but its universal cover is isomorphic to the product SU(2)×SU(2)SU(2)\times SU(2). In general, given a connected semisimple Lie group GG, with center Z(G)Z(G), then the quotient G/Z(G)G/Z(G) is a connected, linear algebraic group which splits as a product of simple groups and it is locally isomorphic to GG. Thus it is common for some results to ask for the group to be centerless.

In the context of algebraic groups defined over a field kk, the concept of kk-rank is the maximal abelian subgroup which can be diagonalized over kk. Thus, for a complex algebraic group, the \mathbb{C}-rank is the dimension of a maximal subgroup isomorphic to a complex torus ()l(\mathbb{C}^{*})^{l} and we are particularly interested in the real rank of a real form. We can observe that the real rank of a product G1××GkG_{1}\times\cdots\times G_{k} is the sum of the real rank of its factors GjG_{j} and we can give some explicit examples.

Example 3.18.

The following is a complete list, up to local isomorphism, of complex, simple Lie groups and some examples of their real forms:

  1. (i)

    The group SLn()SL_{n}(\mathbb{C}), has \mathbb{C}-rank n1n-1 and has the groups SU(p,q)SU(p,q) and SLn()SL_{n}(\mathbb{R}) as real forms, with real rank equal to min{p,q}\mathrm{min}\{p,q\} and n1n-1 respectively.

  2. (ii)

    The group SO(n,)SO(n,\mathbb{C}) has \mathbb{C}-rank n2\left\lfloor\frac{n}{2}\right\rfloor and has the groups SO(p,q)SO(p,q) as real forms, having real rank equal to min{p,q}\mathrm{min}\{p,q\}.

  3. (iii)

    The group Sp(2n,)Sp(2n,\mathbb{C}) has \mathbb{C}-rank nn and has the groups Sp(p,q)Sp(p,q) and Sp(2n,)Sp(2n,\mathbb{R}) as real forms, with real rank equal to nn and min{p,q}\mathrm{min}\{p,q\} respectively.

  4. (iv)

    The exceptional complex groups G2()G_{2}(\mathbb{C}), F4()F_{4}(\mathbb{C}), E6()E_{6}(\mathbb{C}), E7()E_{7}(\mathbb{C}), E8()E_{8}(\mathbb{C}) have \mathbb{C}-rank determined by the corresponding subindex.

Remark 3.19.

Between the possible real forms of a complex semisimple Lie group, there is one and only one compact real form up to conjugacy and such compact form has a compact universal cover, so the compactness property survives in the process of passing to a cover. We can thus, speak of the compact factors of a real semisimple Lie group. Moreover, the rank of a compact Lie group, defined as the dimension of a maximal torus (S1)l(S^{1})^{l} contained in the group, equals the rank of its complexification and has real rank equal to 0. Finally, given a compact, connected, Lie group CC, there is a finite cover of CC that splits as G×TG\times T, where GG is an algebraic semisimple Lie group, and TT is a torus.

Definition 3.20 (Higher Rank Lattice).

A semisimple Lie group is said to have higher rank if its real rank is greater than or equal to 22. Moreover, if a semisimple Lie group has a complexification whose simple factors are all locally isomorphic, the group is called isotypic.

Isotypic Lie groups are important because we can construct irreducible lattices in them, which don’t split as a product of lattices in the simple factors.

Example 3.21.

If σ:(2)(2)\sigma:\mathbb{Q}(\sqrt{2})\rightarrow\mathbb{Q}(\sqrt{2}) is the non-trivial Galois automorphism and Q(x,y,z,t)=x2+y22z22t2Q(x,y,z,t)=x^{2}+y^{2}-\sqrt{2}z^{2}-\sqrt{2}t^{2}, σ(Q)=x2+y2+2z2+2t2\sigma(Q)=x^{2}+y^{2}+\sqrt{2}z^{2}+\sqrt{2}t^{2}. The groups G=SO(Q,)SO(2,2)G=SO(Q,\mathbb{R})\cong SO(2,2) and K=SO(σ(Q),)SO(4)K=SO(\sigma(Q),\mathbb{R})\cong SO(4) are semisimple Lie groups, with KK compact and GG of real rank equal to 22. If we consider the integral points in GG, that is, the group Γ=SO(Q,(2))G\Gamma=SO(Q,\mathbb{Z}(\sqrt{2}))\subset G, then the group,

Γ^={(g,σ(g))G×K:gΓ}\widehat{\Gamma}=\{(g,\sigma(g))\in G\times K:g\in\Gamma\}

is discrete. In fact there is an \mathbb{R}-group GLm()\mathbb{H}_{\mathbb{C}}\subset GL_{m}(\mathbb{C}) such that =G×K\mathbb{H}_{\mathbb{R}}=G\times K and =Γ^\mathbb{H}_{\mathbb{Z}}=\widehat{\Gamma}, in particular, it is a lattice which is co-compact. As the projection G×KGG\times K\rightarrow G has compact Kernel and maps Γ^\widehat{\Gamma} onto Γ\Gamma, thus ΓG\Gamma\subset G is discrete and thus, a co-compact lattice in GG.

Remark 3.22.

The previous example captures the general behaviour of irreducible lattices in isotypic semisimple Lie groups. In fact, isotypic, semisimple Lie groups are the only cases of semisimple Lie groups admiting irreducible lattices and such lattices are constructed with the method of the previous example, but with higher degree extension fields k/k/\mathbb{Q}. See [Mor15], Section 5.6 for the details of the previous example and the construction in general.

4. Detailed Strategy for the Proof of Main Theorem

We are interested in the particular case where given a three-manifold XX, the universal cover X~\widetilde{X} is a homogeneous space, i.e. its group of isometries acts transitively on X~\widetilde{X} , and XX has finite volume. A general setting where this is achieved is when we consider G=X~G=\widetilde{X} a simply-connected Lie group with a right-invariant (or left-invariant) riemannian metric and X=G/ΓX=G/\Gamma, with ΓG\Gamma\subset G a lattice subgroup (i.e. Γ\Gamma is a discrete subgroup such that G/ΓG/\Gamma has finite, left GG-invariant volume). As there is an embedding GIso(G)G\subset Iso(G), we have that NG(Γ)NIso(G)(Γ)N_{G}(\Gamma)\subset N_{Iso(G)}(\Gamma), but it could happen that NG(Γ)/ΓN_{G}(\Gamma)/\Gamma is strictly smaller than Iso(G/Γ)Iso(G/\Gamma). On the other hand, we can extend Γ\Gamma to a discrete subgroup of Iso(G)Iso(G) which is not completely contained in GG, so that the isometry group Iso(G/Γ)Iso(G/\Gamma) is decreased.

In the following sections, we will examinate this phenomenon in the Thurston Geometries, and determine the possible isometry groups of the corresponding finite-volume orbifolds.

The main result 2.8 is proved by a case by case schema organized around the eight geometries, which we present in a resumed form below.

Within the most homogeneous three of them (spherical, hyperbolic and euclidean), the Theorem is a consequence of classical results, which we gather from references and include the classification of spherical manifolds due to Seifert-Threlfall, Borel’s density Theorem for hyperbolic geometry, and Bieberbach’s theorems for the euclidean case, respectively.

In the cases of NilNil and SolSol, we elaborate arguments in this text which include the restrictive conditions for the existence of lattices in solvable groups, as well as an algebraically rigid classification of discrete subgroups of isometries of nilmanifolds. This is the most original contribution among the proofs presented in this note.

For the geometry SL2(~\widetilde{SL_{2}(\mathbb{R}}, which is given as a non-trivial central extension of PSL2()PSL_{2}(\mathbb{R}) by \mathbb{Z}, the arguments include an analysis of the behaviour of the fixed point set of discrete subgroups of isometries of the visual compactification of the hyperbolic plane.

The products S2×S^{2}\times\mathbb{R} and 2×\mathbb{H}^{2}\times\mathbb{R} exhibit differences in the main argument. For the latter the projection to the hyperbolic factor is analyzed, and the result is reduced to the observation that the isometry groups of a two dimensional hyperbolic orbifold are finite, which are combined together with the fact that discrete subgroups of isometries on 2×\mathbb{H}^{2}\times\mathbb{R} project to isometry groups of 2\mathbb{H}^{2} producing finite volume orbifolds.

For the former, S2×S^{2}\times\mathbb{R}, the key remark is that a discrete cocompact isometry group can be realized as subgroup of SO(3)×S1SO(3)\times S^{1}. This result is followed from Tollefsons’ classification results [Tol74] of the groups acting on three manifolds with that geometry.

4.1. Overview of Classification Results

We notice that the classification of isometry groups presented here has as consequence classification results for finite group actions on three manifolds. There are three general kinds of behaviour:

  • Within the geometries whose isometry groups are extensions involving a discrete and finite volume subgroup of SL2()SL_{2}(\mathbb{R}), any finite group can act by isometries. This concerns the geometries 3\mathbb{H}^{3}, SL2~\widetilde{SL_{2}} and 2×\mathbb{H}^{2}\times\mathbb{R}.

  • For the case of spherical factors and the euclidean case, the classical results by Tollefson, Seifert-Threlfall and Bieberbach theorems exhaust the class of finite groups acting by isometries. This is recorded in [McC02], [Tol74], [Now34]. There exists a classification of free (topological) finite group actions in [LSY93], [HJKL02].

  • For Sol, any cyclic group can be realized as a consequence of the discussion of example 9.8. For the case Nil, 6.9 gives a classification depending on the results developed in this note.

    Notice that there exists a classification of topological free actions of finite groups on Nil and Sol manifolds based on the pp- rank and P.A. Smith theory [Lee97], [LSY93].

4.2. Proof Schema for Theorem 2.8.

In this short subsection we summarize the detailed chain of implications leading to the proof of Theorem 2.8.

Euclidean Geometry

According to the Bieberbach Theorems, there exists a discrete free abelian subgroup of translations TT, which has rank less or equal to three. The assertion of theorem 2.8 for the isometry group of R3/ΓR^{3}/\Gamma will be verified by examining the rank of the translation subgroup TT, and discarding rank two and one by producing a quotient of infinite volume.

Nil Geometry.

For the Nil geometry, Theorem 2.8 is a consequence of 6.10, characterizing lattices of infinite volume, and Proposition 6.9 giving an exact sequence between isometry groups.

Spherical Geometry.

The theorem is a direct consequence of the classification of finite group actions on three -manifolds, as well as the determination of the components of the isometry groups 7.3. Notice that there are neither non-discrete subgroups of isometries nor groups whose quotient shows infinite volume within the spherical geometry.

S2×S^{2}\times\mathbb{R} Geometry.

The theorem is consequence of the splitting of the isometry groups, as well as the characterization of discrete subgroups of isometries in 8.3, finally concluding in 8.4.

Sol Geometry.

The theorem is stated as Corollary 9.7, which depends on the determination of centralizers in 9.6, and the determination of finite volume in 9.4.

Hyperbolic Geometry.

The theorem is direct consequence of Lemma 10.1, which is in turn consequence of Borel Density, or the preceding argumentation there. See also [Boc46].

2×\mathbb{H}^{2}\times\mathbb{R} Geometry. The theorem is stated in 13.6, and it is consequence of Theorem12.1, stating the splitting of isometry groups of the factors.

SL2()~\widetilde{SL_{2}(\mathbb{R})} Geometry.

The Theorem is also stated in 13.6, and it is consequence of Lemma 10.1, and Proposition 13.5.

5. Euclidean Geometry

Recall that a three manifold is Euclidean if it is locally isometric to the Euclidean three dimensional space 3\mathbb{R}^{3}. The isometry group of the three dimensional space is the semidirect product E(3)=3O(3).E(3)=\mathbb{R}^{3}\rtimes O(3).

Let Γ\Gamma be a discrete subgroup of isometries E(3)E(3). It is a consequence of the Bieberbach Theorems, as interpreted by Nowacki [Now34], that there exists a free abelian group TT of rank 3\leq 3 and having finite index in Γ\Gamma.

End of proof of Theorem 2.8 for euclidean geometry.

We will now verify the assertion of theorem 2.8 for the isometry group of R3/ΓR^{3}/\Gamma by examining the rank of the translation subgroup TT.

  • If the rank of TT is one, then Γ\Gamma is a finite extension of \mathbb{Z}, and 3/T\mathbb{R}^{3}/T is either the interior of a solid torus or the topological interior of a solid Klein Bottle, depending on the orientability, where the generator of TT acts as a screwdriver isometry (combination of a rotation around an axis and a translation along a parallel direction). It follows that 3/T\mathbb{R}^{3}/T has infinite volume.

  • If the rank of TT is two, then 3/T\mathbb{R}^{3}/T is the total space of a line bundle over either the torus or the Klein bottle, and 3/T\mathbb{R}^{3}/T has infinite volume.

  • If the rank of TT is three, then the isometry group of E(3)/ΓE(3)/\Gamma is a finite extension of a rank three torus by a finite subgroup.

5.1. Classification

The classification of (topological) finite group actions on the torus by isometries has been concluded by work of Lee, Shin and Yokura [LSY93] and Ha, Jo, Kim and Lee [HJKL02].

It follows from the Bieberbach theorems that any topological action on the three torus is topologically conjugated to an isometry; moreover, by the fact that the three dimensional torus is sufficiently large in the sense of Heil and Waldhausen, [Wal68], any homotopy equivalence is homotopic to a homeomorphism, and any two homotopic homeomorphisms are isotopic.

Connected components

The isometry groups of co-compact euclidean orbifolds have been determined by Ratcliffe and Tschantz [RT15], in Theorem 1 and Corollaries 1 and 2 in pages 46 and 47, which we state now for later reference.

Theorem 5.1.

The isometry group of a cocompact euclidean orbifold 3/Γ\mathbb{R}^{3}/\Gamma is a compact Lie group whose identity component is a Torus of dimension equal the first Betti number of the group Γ\Gamma, which corresponds to the rank of the abelian group Γ/[Γ,Γ]\Gamma/[\Gamma,\Gamma].

5.2. Examples

To understand why a compact quotient 3/Γ\mathbb{R}^{3}/\Gamma could have as isometry group a torus of smaller dimension than 33, we can take a look at two examples in dimension two:

Example 5.2.

The group 2\mathbb{Z}^{2} is a discrete subgroup of Iso(2)Iso(\mathbb{R}^{2}), such that the quotient 2/2\mathbb{R}^{2}/\mathbb{Z}^{2}, has the torus N2(2)/22/2N_{\mathbb{R}^{2}}(\mathbb{Z}^{2})/\mathbb{Z}^{2}\cong\mathbb{R}^{2}/\mathbb{Z}^{2} acting naturally by isometries, however the full isometry group Iso(2/2)(2/2)D4Iso(\mathbb{R}^{2}/\mathbb{Z}^{2})\cong(\mathbb{R}^{2}/\mathbb{Z}^{2})\rtimes D_{4} is bigger.

Example 5.3.

We may extend the previous example to the group Λ=2D4\Lambda=\mathbb{Z}^{2}\rtimes D_{4}, which is a discrete subgroup of Iso(2)Iso(\mathbb{R}^{2}), such that it is not completely contained in 2\mathbb{R}^{2} and produces a compact quotient 2/Λ\mathbb{R}^{2}/\Lambda, homeomorphic to the 22-sphere S2S^{2}. To compute the isometry group, we observe the contentions

N2(Λ)={(n/2,n/2+m):n,m}N2(2)=2,N_{\mathbb{R}^{2}}(\Lambda)=\{(n/2,n/2+m):n,m\in\mathbb{Z}\}\subset N_{\mathbb{R}^{2}}(\mathbb{Z}^{2})=\mathbb{R}^{2},

and NIso(2)(Λ)=Aut(2)N2(Λ)D4N2(Λ)N_{Iso(\mathbb{R}^{2})}(\Lambda)=Aut(\mathbb{Z}^{2})\rtimes N_{\mathbb{R}^{2}}(\Lambda)\cong D_{4}\rtimes N_{\mathbb{R}^{2}}(\Lambda), which gives us

NIso(2)(Λ)/Λ(D4N2(Λ))/(D4×2)2.N_{Iso(\mathbb{R}^{2})}(\Lambda)/\Lambda\cong(D_{4}\rtimes N_{\mathbb{R}^{2}}(\Lambda))/(D_{4}\times\mathbb{Z}^{2})\cong\mathbb{Z}_{2}.

This gives us a finite isometry group Iso(2/Λ)2Iso(\mathbb{R}^{2}/\Lambda)\cong\mathbb{Z}_{2}. Observe that if σD4\sigma\in D_{4} and v2v\in\mathbb{Z}^{2}, then the commutator of these elements is [σ,v]=σ(v)v2[\sigma,v]=\sigma(v)-v\in\mathbb{Z}^{2} and we can see that the commutator group [Γ,Γ][\Gamma,\Gamma] contains a lattice subgroup of 2\mathbb{R}^{2} which implies that Γ/[Γ,Γ]\Gamma/[\Gamma,\Gamma] is finite, verifying Theorem 5.1.

6. Nil geometry

6.1. Riemannian geometry of the Heisenberg group

If 𝔽\mathbb{F} is a commutative ring, denote by H𝔽H_{\mathbb{F}} the group of 3×33\times 3 upper triangular matrices over 𝔽\mathbb{F} with 11 in the diagonal, that is

H𝔽={(1xz01y001):x,y,z𝔽}.H_{\mathbb{F}}=\left\{\left(\begin{array}[]{ccc}1&x&z\\ 0&1&y\\ 0&0&1\end{array}\right):x,y,z\in\mathbb{F}\right\}.

The group HH_{\mathbb{R}} is a Lie group called the three dimensional Heisenberg group that fits into the exact sequence

1H21,1\rightarrow\mathbb{R}\rightarrow H_{\mathbb{R}}\rightarrow\mathbb{R}^{2}\rightarrow 1,

where H\mathbb{R}\subset H_{\mathbb{R}} is its center. The three matrices

e1=(010000000),e2=(000001000),e3=(001000000);e_{1}=\left(\begin{array}[]{ccc}0&1&0\\ 0&0&0\\ 0&0&0\end{array}\right),\quad e_{2}=\left(\begin{array}[]{ccc}0&0&0\\ 0&0&1\\ 0&0&0\end{array}\right),\quad e_{3}=\left(\begin{array}[]{ccc}0&0&1\\ 0&0&0\\ 0&0&0\end{array}\right);

determine a canonical basis of the tangent space at the identity TI(H)T_{I}(H_{\mathbb{R}}), so that its translations by left-multiplications gives us a basis of left invariant vector fields denoted by XjX_{j} with Xj(I)=ejX_{j}(I)=e_{j}. For a fixed element

g=(1xz01y001)H,g=\left(\begin{array}[]{ccc}1&x&z\\ 0&1&y\\ 0&0&1\end{array}\right)\in H_{\mathbb{R}},

the vector fields at Tg(H)T_{g}(H_{\mathbb{R}}) have expresions

X1(g)=(010000000),X2(g)=(00x001000),X3(g)=(001000000).X_{1}(g)=\left(\begin{array}[]{ccc}0&1&0\\ 0&0&0\\ 0&0&0\end{array}\right),\quad X_{2}(g)=\left(\begin{array}[]{ccc}0&0&x\\ 0&0&1\\ 0&0&0\end{array}\right),\quad X_{3}(g)=\left(\begin{array}[]{ccc}0&0&1\\ 0&0&0\\ 0&0&0\end{array}\right).

If we consider the global coordinates

3H,(x,y,z)(1xz01y001),\mathbb{R}^{3}\rightarrow H_{\mathbb{R}},\qquad(x,y,z)\mapsto\left(\begin{array}[]{ccc}1&x&z\\ 0&1&y\\ 0&0&1\end{array}\right),

then a vector vTg(H)v\in T_{g}(H_{\mathbb{R}}) decomposes as

v=v1e1+v2e2+v3e3=v1X1(g)+v2X2(g)+(v3xv2)X3(g),v=v_{1}e_{1}+v_{2}e_{2}+v_{3}e_{3}=v_{1}X_{1}(g)+v_{2}X_{2}(g)+(v_{3}-xv_{2})X_{3}(g),

so that the left-invariant metric in HH_{\mathbb{R}} having Xj(g)X_{j}(g) as an orthonormal basis is given in this coordinates as ds2=dx2+dy2+(dzxdy)2ds^{2}=dx^{2}+dy^{2}+(dz-xdy)^{2}. Being left-invariant, this metric has HH_{\mathbb{R}} as a subgroup of isometries given by left multiplication

Lg:HH,Lg(h)=gh,L_{g}:H_{\mathbb{R}}\rightarrow H_{\mathbb{R}},\qquad L_{g}(h)=gh,

for every gHg\in H_{\mathbb{R}}. Notice that there are other isometries that don’t come from left multiplication of HH_{\mathbb{R}}. Such isometries form a group isometric to the orthogonal group O(2)O(2) generated by the reflection R(x,y,z)=(x,y,z)R(x,y,z)=(x,-y,-z) and the twisted rotations

m:S1×HH,mθ(x,y,z)=(ρθ(x,y),z+ηθ(x,y)),m:S^{1}\times H_{\mathbb{R}}\rightarrow H_{\mathbb{R}},\qquad m_{\theta}(x,y,z)=(\rho_{\theta}(x,y),z+\eta_{\theta}(x,y)),

where ρθ\rho_{\theta} is a rotation in the (x,y)(x,y)-plane with angle θ\theta, ηθ\eta_{\theta} is a polynomial function in xx and yy and trigonometric in θ\theta.

The full isometry group Iso(H)Iso(H_{\mathbb{R}}) can be described as a semi-direct product HO(2)H_{\mathbb{R}}\rtimes O(2), because

mθLgmθ1=Lmθ(g),RLgR=LR(g),m_{\theta}\circ L_{g}\circ m_{\theta}^{-1}=L_{m_{\theta}(g)},\qquad R\circ L_{g}\circ R=L_{R(g)},

meaning that the exact exact sequence

1Iso(H)Iso(2)1,1\rightarrow\mathbb{R}\rightarrow Iso(H_{\mathbb{R}})\rightarrow Iso(\mathbb{R}^{2})\rightarrow 1,

induced by the action on the quotient by the center H/2H_{\mathbb{R}}/\mathbb{R}\cong\mathbb{R}^{2} splits off, see [Sco83] for more details.

6.2. Examples

In this section we describe a series of ilustrative examples that capture the behaviour of every discrete subgroup of Iso(H)Iso(H_{\mathbb{R}}).

Example 6.1.

The group HHH_{\mathbb{Z}}\subset H_{\mathbb{R}} is a discrete subgroup so that the exact sequence determining HH_{\mathbb{R}} induces the fiber-bundle structure

/H/H2/2\mathbb{R}/\mathbb{Z}\rightarrow H_{\mathbb{R}}/H_{\mathbb{Z}}\rightarrow\mathbb{R}^{2}/\mathbb{Z}^{2}

and thus HH_{\mathbb{Z}} is a lattice subgroup of HH_{\mathbb{R}} such that H/HH_{\mathbb{R}}/H_{\mathbb{Z}} is a compact Riemannian manifold. As the conjugation can be computed as

g=(x,y,z),g(n,m,p)g1=(n,m,p+xmyn),g=(x,y,z),\qquad g\ (n,m,p)\ g^{-1}=(n,m,p+xm-yn),

the normalizer in HH_{\mathbb{R}} is NH(H)={(n,m,p):n,m,p}N_{H_{\mathbb{R}}}(H_{\mathbb{Z}})=\left\{\left(n,m,p\right):n,m\in\mathbb{Z},\ p\in\mathbb{R}\right\}. This gives us the isometries in the quotient

S1NH(H)/HIso(H/H).S^{1}\cong N_{H_{\mathbb{R}}}(H_{\mathbb{Z}})/H_{\mathbb{Z}}\hookrightarrow Iso(H_{\mathbb{R}}/H_{\mathbb{Z}}).

We consider now the normalizer of the Heisenberg group in Iso(H)Iso(H_{\mathbb{R}}). This can be determined as

NIso(H)(H)={(n,m,p):n,m,p}D4,N_{Iso(H_{\mathbb{R}})}(H_{\mathbb{Z}})=\left\{\left(n,m,p\right):n,m\in\mathbb{Z},\ p\in\mathbb{R}\right\}\rtimes D_{4},

where the Dihedral group D4D_{4} is generated by the isometries

mπ/2(n,m,p)=(m,n,pnm),R(n,m,p)=(n,m,p),m_{\pi/2}(n,m,p)=(-m,n,p-nm),\qquad R(n,m,p)=(n,-m,-p),

so that what we get is Iso(H/H)S1D4Iso(H_{\mathbb{R}}/H_{\mathbb{Z}})\cong S^{1}\rtimes D_{4}.

We can modify this example by adding the dihedral group to the lattice, so that we have the fiber bundle structure

/H/(HD4)2/(2D4)S2.\mathbb{R}/\mathbb{Z}\rightarrow H_{\mathbb{R}}/(H_{\mathbb{Z}}\rtimes D_{4})\rightarrow\mathbb{R}^{2}/(\mathbb{Z}^{2}\rtimes D_{4})\cong S^{2}.

The discussion on Example 5.3 explains the last quotient in the sequence). Notice that we decreased the normalizer

NIso(H)(HD4)={(n,m,l):n,m,2l}D4,N_{Iso(H_{\mathbb{R}})}(H_{\mathbb{Z}}\rtimes D_{4})=\{(n,m,l):n,m,2l\in\mathbb{Z}\}\rtimes D_{4},

so that Iso(H/(HD4))2Iso(H_{\mathbb{R}}/(H_{\mathbb{Z}}\rtimes D_{4}))\cong\mathbb{Z}_{2}.

Example 6.2.

Fix a positive integer pp\in\mathbb{N} and consider the lattice

Gp={(n,m,lp):n,m,l}H,G_{p}=\left\{\left(n,m,\frac{l}{p}\right):n,m,l\in\mathbb{Z}\right\}\subset H_{\mathbb{R}},

which has as normalizer group in HH_{\mathbb{R}} the group

NH(Gp)={(np,mp,r):n,m,r},N_{H_{\mathbb{R}}}(G_{p})=\left\{\left(\frac{n}{p},\frac{m}{p},r\right):n,m\in\mathbb{Z},\ r\in\mathbb{R}\right\},

and normalizer group in Iso(H)Iso(H_{\mathbb{R}}), the group NH(Gp)D4N_{H_{\mathbb{R}}}(G_{p})\rtimes D_{4}, with Dihedral group D4=mπ/2,RD_{4}=\langle m_{\pi/2},R\rangle as before. The isometry group is characterized by the exact sequence

1S1Iso(H/Gp)D4(p×p)11\rightarrow S^{1}\rightarrow Iso(H_{\mathbb{R}}/G_{p})\rightarrow D_{4}\ltimes(\mathbb{Z}_{p}\times\mathbb{Z}_{p})\rightarrow 1

and we recover the previous example by taking p=1p=1.

Example 6.3.

Fix a positive integer pp\in\mathbb{N} and consider the lattice

Lp={(n2+m,3n2,3l2p):n,m,l}H,L_{p}=\left\{\left(\frac{n}{2}+m,\frac{\sqrt{3}n}{2},\frac{\sqrt{3}l}{2p}\right):n,m,l\in\mathbb{Z}\right\}\subset H_{\mathbb{R}},

so that it has normalizer group in HH_{\mathbb{R}}

NH(Lp)={(n2p+mp,3n2p,r):n,m,r}.N_{H_{\mathbb{R}}}(L_{p})=\left\{\left(\frac{n}{2p}+\frac{m}{p},\frac{\sqrt{3}n}{2p},r\right):n,m\in\mathbb{Z},\ r\in\mathbb{R}\right\}.

As the group LpL_{p} projects to a hexagonal lattice in 2\mathbb{R}^{2}, we should expect to have a Dihedral group D6D_{6} normalizing LpL_{p}, however, the rotation mπ/3:HHm_{\pi/3}:H_{\mathbb{R}}\rightarrow H_{\mathbb{R}} given by

mπ/3(x,y,z)=(12(x3y),12(y+3x),z+38(y2x223xy)),m_{\pi/3}(x,y,z)=\left(\frac{1}{2}\left(x-\sqrt{3}y\right),\frac{1}{2}\left(y+\sqrt{3}x\right),z+\frac{\sqrt{3}}{8}\left(y^{2}-x^{2}-2\sqrt{3}xy\right)\right),

doesn’t preserve LpL_{p}. To fix this, we must add a translation mixed with the rotation.Put g=(18,38,0)Hg=\left(\frac{1}{8},-\frac{\sqrt{3}}{8},0\right)\in H_{\mathbb{R}}, then φ=mπ/3LgNIso(H)(Lp)\varphi=m_{\pi/3}\circ L_{g}\in N_{Iso(H_{\mathbb{R}})}(L_{p}), which can be verified using the relation φLhφ1=Lmπ/3(ghg1)\varphi\circ L_{h}\circ\varphi^{-1}=L_{m_{\pi/3}(ghg^{-1})}. We can describe the normalizer group of LpL_{p} in Iso(H)Iso(H_{\mathbb{R}}) in terms of generators as

NIso(H)(Lp)=Lg,φ,R:gNH(Lp),N_{Iso(H_{\mathbb{R}})}(L_{p})=\langle L_{g},\varphi,R:g\in N_{H_{\mathbb{R}}}(L_{p})\rangle,

where R(x,y,z)=(x,y,z)R(x,y,z)=(x,-y,-z), and so, we have the isometry group

1S1Iso(H/Lp)(p×p)D61,1\rightarrow S^{1}\rightarrow Iso(H_{\mathbb{R}}/L_{p})\rightarrow(\mathbb{Z}_{p}\times\mathbb{Z}_{p})\rtimes D_{6}\rightarrow 1,

where the dihedral group D6D_{6} is generated by φ,R\langle\varphi,R\rangle.

Example 6.4.

All the previous examples can be generalized as follows: Fix pp\in\mathbb{N} and u,v2u,v\in\mathbb{R}^{2} linearly independent, so that Γ={nu+mv:n,m}2\Gamma=\{nu+mv:n,m\in\mathbb{Z}\}\subset\mathbb{R}^{2} is a lattice. If (u,0)×(v,0)=(0,0,λ)3(u,0)\times(v,0)=(0,0,\lambda)\in\mathbb{R}^{3}, then the group

Mp={(nu+mv,λpl):n,m,l}HM_{p}=\left\{\left(nu+mv,\frac{\lambda}{p}l\right):n,m,l\in\mathbb{Z}\right\}\subset H_{\mathbb{R}}

is a lattice having normalizer group in HH_{\mathbb{R}}

NH(Mp)={(npu+mpv,r):n,m,r}.N_{H_{\mathbb{R}}}(M_{p})=\left\{\left(\frac{n}{p}u+\frac{m}{p}v,r\right):n,m\in\mathbb{Z},\ r\in\mathbb{R}\right\}.

The lattice Γ\Gamma has an automorphism group Aut(Γ){0,2,D4,D6}Aut(\Gamma)\in\{0,\mathbb{Z}_{2},D_{4},D_{6}\}, which is, if non-trivial, generated by a rotation with angle θ\theta and a reflection. The whole normalizer group is given in terms of generators as

NIso(H)(Mp)=Lg,φ,R:gNH(Mp),N_{Iso(H_{\mathbb{R}})}(M_{p})=\langle L_{g},\varphi,R:g\in N_{H_{\mathbb{R}}}(M_{p})\rangle,

where φ=mθLw\varphi=m_{\theta}\circ L_{w}. Here, w=(w0,0)Hw=(w_{0},0)\in H_{\mathbb{R}} must be chosen so that if

w×(u,0)=(0,r1),w×(v,0)=(0,r2),w\times(u,0)=(0,r_{1}),\qquad w\times(v,0)=(0,r_{2}),

then mθ(u,r1),mθ(v,r2)Mpm_{\theta}(u,r_{1}),m_{\theta}(v,r_{2})\in M_{p}. Thus, we have an isometry group of the quotient given by the exact sequence

1S1Iso(H/Mp)(p×p)Aut(Γ)1.1\rightarrow S^{1}\rightarrow Iso(H_{\mathbb{R}}/M_{p})\rightarrow(\mathbb{Z}_{p}\times\mathbb{Z}_{p})\rtimes Aut(\Gamma)\rightarrow 1.
Remark 6.5.

The previous examples give us the general strategy to compute the isometry group of a quotient H/GH_{\mathbb{R}}/G, for GIso(H)G\subset Iso(H_{\mathbb{R}}) a discrete group of isometries. This strategy is as follows: GG projects to a discrete subgroup ΓIso(2)\Gamma\subset Iso(\mathbb{R}^{2}) which has a finite index subgroup Γ02\Gamma_{0}\subset\mathbb{R}^{2}, corresponding to a finite index subgroup G0=GHGG_{0}=G\cap H_{\mathbb{R}}\subset G and a lattice in HH_{\mathbb{R}}. The normalizer of G0G_{0} projects again a lattice in 2\mathbb{R}^{2} and thus Iso(H/G0)Iso(H_{\mathbb{R}}/G_{0}) is an extension of a finite group p×p\mathbb{Z}_{p}\times\mathbb{Z}_{p} by S1S^{1}. The isometry group Iso(H/G)Iso(H_{\mathbb{R}}/G) is just the previous group with an extra finite group of isometries, coming from the automorphisms of the lattice Aut(Γ)Aut(\Gamma). This strategy fails if the projection to 2\mathbb{R}^{2} is non-discrete, a possibility shown in the following two examples, however, in the case where the quotient H/GH_{\mathbb{R}}/G has finite volume, we will see that this patological behaviour doesn’t occur.

Here we add two examples of discrete groups whose projected action onto 2\mathbb{R}^{2} is non-discrete, these examples capture the general behaviour of discrete groups having this property as we will see in the next section.

Example 6.6.

Consider φ:S1\varphi:{\mathbb{N}}\rightarrow S^{1}, a homomorphism with dense image and g=(0,0,1)Hg=(0,0,1)\in H_{\mathbb{R}} a generator of the center, so that the group

{(gn,φ(n)):n}HS1Iso(H)\{(g^{n},\varphi(n)):n\in{\mathbb{N}}\}\subset H_{\mathbb{R}}\rtimes S^{1}\cong Iso(H_{\mathbb{R}})

is a discrete subgroup of isometries of HH_{\mathbb{R}} with dense projection onto S1SO(2)S^{1}\cong SO(2) and in particular, with a non-discrete action on 2\mathbb{R}^{2}. In this example, the projected group leaves fixed the point p=11λ2p=\frac{1}{1-\lambda}\in{\mathbb{C}}\cong{\mathbb{R}}^{2}, where λ=φ(1)\lambda=\varphi(1) and in particular, it is a group of rotations around such point.

Example 6.7.

Given a scaling 0<ε<10<\varepsilon<1, consider the group generated by (1,0,1),(ε,0,1)HIso(H)(1,0,1),(\varepsilon,0,1)\in H_{\mathbb{R}}\subset Iso(H_{\mathbb{R}}) and 1𝕊1Iso(H)-1\in{\mathbb{S}}^{1}\subset Iso(H_{\mathbb{R}}). This is a discrete subgroup of Iso(H)Iso(H_{\mathbb{R}}), which projects to a non-discrete subgroup of Iso(2)Iso({\mathbb{R}}^{2}) leaving fixed the line {(x,0):x}2\{(x,0):x\in{\mathbb{R}}\}\subset{\mathbb{R}}^{2}.

Remark 6.8.

The most symmetric lattices in 2\mathbb{R}^{2} are the square and hexagonal lattices, having linear symmetry groups D4D_{4} and D6D_{6}. Theorem 6.12 tells us that the generalizations of these lattices to HH_{\mathbb{R}}, described in Example 6.2 and Example 6.3 are the most symmetric finite volume quotients H/GH_{\mathbb{R}}/G, with isometry groups

1S1Iso(H/G)(n×n)D1,1\rightarrow S^{1}\rightarrow Iso(H_{\mathbb{R}}/G)\rightarrow(\mathbb{Z}_{n}\times\mathbb{Z}_{n})\rtimes D\rightarrow 1,

with DD equal to D4D_{4} and D6D_{6} respectively, and nn\in\mathbb{N}.

6.3. Classification of discrete subgroups of isometries

In this section HH_{\mathbb{R}} denotes the Heisenberg Lie group considered as a Riemannian manifold with respect to the left-invariant metric constructed in the previous section. Here, we describe the conditions on which a discrete group on Iso(H)Iso(H_{\mathbb{R}}) induces a discrete action on the Euclidean plane 2\mathbb{R}^{2}.

Proposition 6.9.

If GG is a discrete subgroup of isometries of HH_{\mathbb{R}}, then the exact sequence

1Iso(H)Iso(2)11\rightarrow\mathbb{R}\rightarrow Iso(H_{\mathbb{R}})\rightarrow Iso(\mathbb{R}^{2})\rightarrow 1

induces an exact sequence

1KGΓ1,1\rightarrow K\rightarrow G\rightarrow\Gamma\rightarrow 1,

where ΓIso(2)\Gamma\subset Iso(\mathbb{R}^{2}) is either discrete or it is an abelian group leaving fixed either a point or a line. Moreover,

  1. (i)

    if ΓIso(2)\Gamma\subset Iso(\mathbb{R}^{2}) has a finite index lattice, then KK\subset\mathbb{R} is a non-trivial discrete subgroup and

  2. (ii)

    if Γ\Gamma is non-discrete and leaves fixed a line, then there is a finite index subgroup of GG which is contained in HH_{\mathbb{R}}.

Proof.

Observe first that K=GK=G\cap\mathbb{R} is a discrete subgroup of isometries of \mathbb{R} and so, if non-trivial, there is an isomorphism /KS1\mathbb{R}/K\cong S^{1}. The exact sequence

1S1H/K211\rightarrow S^{1}\rightarrow H_{\mathbb{R}}/K\rightarrow\mathbb{R}^{2}\rightarrow 1

gives us

1S1Iso(H)/KIso(2)1,1\rightarrow S^{1}\rightarrow Iso(H_{\mathbb{R}})/K\rightarrow Iso(\mathbb{R}^{2})\rightarrow 1,

which has compact Kernel and thus, any discrete group in Iso(H)/KIso(H_{\mathbb{R}})/K projects to a discrete group in Iso(2)Iso(\mathbb{R}^{2}). This argument tells us that if KK is non-trivial then Γ\Gamma is discrete in Iso(2)Iso(\mathbb{R}^{2}), because it is the projection of G/KG/K with compact kernel, and G/KG/K is always discrete in Iso(H)/KIso(H_{\mathbb{R}})/K. Suppose from now on that KK is trivial. If we identify 2\mathbb{R}^{2}\cong\mathbb{C} as a Euclidean space, then we can realize the group of orientation preserving isometries of the plane 2\mathbb{R}^{2} as the matrix group

Iso+(2){(λz01):λ,z,|λ|=1}Iso^{+}(\mathbb{R}^{2})\cong\left\{\left(\begin{array}[]{cc}\lambda&z\\ 0&1\end{array}\right):\lambda,z\in\mathbb{C},\ |\lambda|=1\right\}

with action

(λz01)(w1)=(λw+z1).\left(\begin{array}[]{cc}\lambda&z\\ 0&1\end{array}\right)\left(\begin{array}[]{c}w\\ 1\end{array}\right)=\left(\begin{array}[]{c}\lambda w+z\\ 1\end{array}\right).

Observe that the restriction Iso+(2)Iso(2)Iso^{+}(\mathbb{R}^{2})\subset Iso(\mathbb{R}^{2}) reduces the discusion to a subgroup of index 2, which doesn’t alter the property of discreteness. We recall two important properties on commutators. First, commutators of two isometries give elements of pure translation part

[(λz01),(βw01)]=(1(zw)+(λwβz)01),\left[\left(\begin{array}[]{cc}\lambda&z\\ 0&1\end{array}\right),\left(\begin{array}[]{cc}\beta&w\\ 0&1\end{array}\right)\right]=\left(\begin{array}[]{cc}1&(z-w)+(\lambda w-\beta z)\\ 0&1\end{array}\right),

which tells us that [G,G][G,G] projects to a subgroup of Iso(2)Iso(\mathbb{R}^{2}) with only translation part, and so [G,G]H[G,G]\subset H_{\mathbb{R}}. Second, the commutator in HH_{\mathbb{R}} satisfies the relation

[(1xr01y001),(1us01v001)]=(10xvuy010001),\left[\left(\begin{array}[]{ccc}1&x&r\\ 0&1&y\\ 0&0&1\end{array}\right),\left(\begin{array}[]{ccc}1&u&s\\ 0&1&v\\ 0&0&1\end{array}\right)\right]=\left(\begin{array}[]{ccc}1&0&xv-uy\\ 0&1&0\\ 0&0&1\end{array}\right),

which has the geometric interpretation: if two elements of HH_{\mathbb{R}} project to the vectors (x,y)(x,y) and (u,v)(u,v), then its commutator is an element of the center =Z(H)\mathbb{R}=Z(H_{\mathbb{R}}) whose magnitud is the area of the projected vectors. As we are under the supposition that GG\cap\mathbb{R} is trivial, the two previous relations on commutators tells us that [G,G][G,G] is a commutative group and the corresponding projected group satisfies

[Γ,Γ]{(1rz001):r}[\Gamma,\Gamma]\subset\left\{\left(\begin{array}[]{cc}1&rz_{0}\\ 0&1\end{array}\right):r\in\mathbb{R}\right\}

for some z0z_{0}\in\mathbb{C}. Suppose first that Γ\Gamma is non-commutative. The commutation relation

[(λz01),(1z001)]=(1(λ1)z001),\left[\left(\begin{array}[]{cc}\lambda&z\\ 0&1\end{array}\right),\left(\begin{array}[]{cc}1&z_{0}\\ 0&1\end{array}\right)\right]=\left(\begin{array}[]{cc}1&(\lambda-1)z_{0}\\ 0&1\end{array}\right),

and the hypothesis that all the translation elements of [Γ,Γ][\Gamma,\Gamma] are linearly dependent give us the condition r=(1λ)r=(1-\lambda) for some rr\in\mathbb{R} and as |λ|=1|\lambda|=1, the only options are λ=±1\lambda=\pm 1. As Γ\Gamma is non-commutative, there is at least one element that is not a translation, that is

(1z01)Γ,\left(\begin{array}[]{cc}-1&z\\ 0&1\end{array}\right)\in\Gamma,

and without loss of generality, we can change Γ\Gamma by hΓh1h\Gamma h^{-1} (where hh is the translation by 1/2z1/2z) so that in fact

(1001)Γ,\left(\begin{array}[]{cc}-1&0\\ 0&1\end{array}\right)\in\Gamma,

this conjugation leaves [Γ,Γ][\Gamma,\Gamma] invariant. Observe also that

[(βw01),(1001)]=(12w01)[Γ,Γ],\left[\left(\begin{array}[]{cc}\beta&w\\ 0&1\end{array}\right),\left(\begin{array}[]{cc}-1&0\\ 0&1\end{array}\right)\right]=\left(\begin{array}[]{cc}1&2w\\ 0&1\end{array}\right)\subset[\Gamma,\Gamma],

implies by the same argument that β=±1\beta=\pm 1 and w=sz0w=sz_{0} for some ss\in\mathbb{R}, and thus Γ\Gamma preserves the line generated by z0z_{0}. If on the other hand Γ\Gamma is commutative and contains an element of the form

(λz01),λ1,\left(\begin{array}[]{cc}\lambda&z\\ 0&1\end{array}\right),\qquad\lambda\neq 1,

this element has as a unique fixed point zλ1\frac{-z}{\lambda-1}. As Γ\Gamma is a commutative group, every element of Γ\Gamma must fix zλ1\frac{-z}{\lambda-1}, and thus, it consists of rotations around this point. If no such element exists, Γ\Gamma consists of elements with purely translation part, which tells us that GHG\subset H_{\mathbb{R}}. We observe that in this last case, two elements a,bGa,b\in G which project to two linearly independent vectors in Γ\Gamma must satisfy that e[a,b]GKe\neq[a,b]\in G\cap K, which can’t happen by hypothesis, so Γ\Gamma is a subgroup of the group {rω:r}\{r\omega:r\in\mathbb{R}\} for some ω\omega\in\mathbb{C} and thus Γ\Gamma preserves the line generated by ω\omega. ∎

Lemma 6.10.

Let GG be a discrete subgroup of isometries of HH_{\mathbb{R}} together with the projection to the isometry group of 2\mathbb{R}^{2}

GΓIso(2).G\rightarrow\Gamma\subset Iso(\mathbb{R}^{2}).

If Γ\Gamma preserves either a line or a point in 2\mathbb{R}^{2}, then the orbifold H/GH_{\mathbb{R}}/G has infinite volume.

Proof.

Suppose first that Γ\Gamma preserves the line v\mathbb{R}v, then as a consequence of either Bieberbach’s Theorem if Γ\Gamma is discrete, or as a consequence of the proof of Proposition 6.9 if Γ\Gamma is non-discrete, GG has a finite index subgroup that is contained in HH_{\mathbb{R}}. Passing to a finite index subgroup doesn’t change the property of having finite co-volume so without loss of generality we may suppose that GHG\subset H_{\mathbb{R}}. There is a fundamental domain that has non-empty interior, given for example by the Dirichlet’s fundamental domain {qH:d(q0,q)<d(q0,γ(q)),γG{e}}\{q\in H_{\mathbb{R}}:d(q_{0},q)<d(q_{0},\gamma(q)),\ \gamma\in G\setminus\{e\}\}, with respect to the Riemannian distance dd, see [Rat19]. In particular there is a subset of the form

D={(tv+sv,λr0)×:(s,t,λ)(ε,ε)3+(s0,t0,λ0)}HD=\{(tv+sv^{\perp},\lambda r_{0})\subset\mathbb{C}\times\mathbb{R}:(s,t,\lambda)\in(-\varepsilon,\varepsilon)^{3}+(s_{0},t_{0},\lambda_{0})\}\subset H_{\mathbb{R}}

such that no two elements of DD can be identified with an element of GG. As Γ\Gamma preserves the line v\mathbb{R}v, then we can see that no two elements of D~\widetilde{D} can be identified with an element of GG, where

D~={(tv+sv,λr0)×:(s,t,λ)×(ε,ε)2+(0,t0,λ0)}\widetilde{D}=\{(tv+sv^{\perp},\lambda r_{0})\subset\mathbb{C}\times\mathbb{R}:(s,t,\lambda)\in\mathbb{R}\times(-\varepsilon,\varepsilon)^{2}+(0,t_{0},\lambda_{0})\}

but D~=jDj\widetilde{D}=\bigcup_{j}D_{j}, where

Dj={(tv+sv,λr0)×:(s,t,λ)(ε,ε)3+(sj,t0,λ0)}HD_{j}=\{(tv+sv^{\perp},\lambda r_{0})\subset\mathbb{C}\times\mathbb{R}:(s,t,\lambda)\in(-\varepsilon,\varepsilon)^{3}+(s_{j},t_{0},\lambda_{0})\}\subset H_{\mathbb{R}}

and every DjD_{j} can be obtained by translating DD with an element of HH_{\mathbb{R}}, thus

Vol(H/G)Vol(D~)=jVol(Dj)=jVol(D)=.Vol(H_{\mathbb{R}}/G)\geq Vol(\widetilde{D})=\sum_{j}Vol(D_{j})=\sum_{j}Vol(D)=\infty.

The second possibility is when Γ\Gamma is a commutative group preserving a point, that is, Γ\Gamma is conjugated to a subgroup of SO(2)SO(2). Again there is a fundamental domain of GG with non-empty interior and in particular, there is a subset

Ω={(reiθ,sr0)×:(r,θ,s)(ε,ε)3+(a,b,c)}H,\Omega=\{(re^{i\theta},sr_{0})\subset\mathbb{C}\times\mathbb{R}:(r,\theta,s)\in(-\varepsilon,\varepsilon)^{3}+(a,b,c)\}\subset H_{\mathbb{R}},

such that no two elements of Ω\Omega can be identified with an element of GG. As Γ\Gamma acts only as rotations in the \mathbb{C} plane, we can enlarge as before Ω\Omega to the subset

Ω~={(reiθ,sr0)×:(r,θ,s)>0×(ε,ε)2+(0,b,c)},\widetilde{\Omega}=\{(re^{i\theta},sr_{0})\subset\mathbb{C}\times\mathbb{R}:(r,\theta,s)\in\mathbb{R}_{>0}\times(-\varepsilon,\varepsilon)^{2}+(0,b,c)\},

so that no two elements of Ω~\widetilde{\Omega} can be identified with an element of GG. As before, we have a countable union of disjoints sets contained in Ω~\widetilde{\Omega} that are translated copies of Ω\Omega, that is

iΩ+(ωj,0)Ω~\bigcup_{i}\Omega+(\omega_{j},0)\subset\widetilde{\Omega}

and Vol(H/G)Vol(Ω~)jVol(Ω+(ωj,0))=jVol(Ω)=Vol(H_{\mathbb{R}}/G)\geq Vol(\widetilde{\Omega})\geq\sum_{j}Vol(\Omega+(\omega_{j},0))=\sum_{j}Vol(\Omega)=\infty. ∎

Lemma 6.11.

If u,v2u,v\in\mathbb{R}^{2} are two linearly independent vectors, with (u,0)×(v,0)=(0,0,λ)3(u,0)\times(v,0)=(0,0,\lambda)\in\mathbb{R}^{3} and nn\in\mathbb{N}, r,sr,s\in\mathbb{R}, then the group

G=(u,r),(v,s),(0,0,λn)HG=\left\langle\left(u,r\right),\left(v,s\right),\left(0,0,\frac{\lambda}{n}\right)\right\rangle\subset H_{\mathbb{R}}

is a lattice in HH_{\mathbb{R}}. Conversely, every lattice in HH_{\mathbb{R}} can be obtained like this.

Proof.

Observe that the center of GG is the subgroup K={λpn:p}K=\left\{\frac{\lambda p}{n}:p\in\mathbb{Z}\right\} and if (x,y,z),(x,y,z)G(x,y,z),(x,y,z^{\prime})\in G, then

(x,y,z)1(x,y,z)=(0,0,zz)K,(x,y,z)^{-1}\cdot(x,y,z^{\prime})=(0,0,z^{\prime}-z)\in K,

so that for k,lk,l\in\mathbb{N} fixed, and

(u,r)k(v,s)l=(ku+lv,rk,l)(u,r)^{k}\cdot(v,s)^{l}=(ku+lv,r_{k,l})

the level set

{(w,z)G:w=ku+lv}={(ku+lv,rn,m+λpn):p}\{(w,z)\in G:w=ku+lv\}=\left\{\left(ku+lv,r_{n,m}+\frac{\lambda p}{n}\right):p\in\mathbb{Z}\right\}

is discrete and thus, GG is a discrete subgroup of HH_{\mathbb{R}}. If Γ={ku+lv:k,l}\Gamma=\left\{ku+lv:k,l\in\mathbb{Z}\right\} denotes the projection of GG onto 2\mathbb{R}^{2}, then there is an exact sequence

1KGΓ11\rightarrow K\rightarrow G\rightarrow\Gamma\rightarrow 1

which induces the fiber bundle structure

S1/KH/G2/ΓS1×S1,S^{1}\cong\mathbb{R}/K\rightarrow H_{\mathbb{R}}/G\rightarrow\mathbb{R}^{2}/\Gamma\cong S^{1}\times S^{1},

which tells us that H/GH_{\mathbb{R}}/G is compact and thus, GG is a lattice in HH_{\mathbb{R}}. Suppose now that LHL\subset H_{\mathbb{R}} is a lattice, then by Lemma 6.10, LL projects to a lattice subgroup of 2\mathbb{R}^{2}, generated by two linearly independent vectors u,v2u^{\prime},v^{\prime}\in\mathbb{R}^{2} such that (u,0)×(v,0)=(0,0,λ)(u^{\prime},0)\times(v^{\prime},0)=(0,0,\lambda^{\prime}), with 0λ0\neq\lambda^{\prime}\in\mathbb{R} and observe that if g=(u,r),h=(v,s)Lg=(u^{\prime},r^{\prime}),h=(v^{\prime},s^{\prime})\in L, then their commutator is [g,h]=(0,0,λ)[g,h]=(0,0,\lambda^{\prime}). As the intersection K=GZ(H)K^{\prime}=G\cap Z(H_{\mathbb{R}}) is discrete and contains the non-trivial element (0,0,λ)K(0,0,\lambda^{\prime})\in K^{\prime}, then there is an integer nn^{\prime}\in\mathbb{N} such that K={λpn:p}K^{\prime}=\left\{\frac{\lambda^{\prime}p}{n^{\prime}}:p\in\mathbb{Z}\right\} and thus, the lattice LL is generated by the set {(u,r),(v,s),(0,0,λn)}\left\{(u^{\prime},r^{\prime}),(v^{\prime},s^{\prime}),\left(0,0,\frac{\lambda^{\prime}}{n^{\prime}}\right)\right\}. ∎

Connected components

Theorem 6.12.

If GIso(H)G\subset Iso(H_{\mathbb{R}}) is a discrete subgroup such that H/GH_{\mathbb{R}}/G has finite volume, then there is an exact sequence

1CIso(H/G)F1,1\rightarrow C\rightarrow Iso(H_{\mathbb{R}}/G)\rightarrow F\rightarrow 1,

where FF is a finite group, and CS1C\subset S^{1} is a closed subgroup. In particular, either Iso(H/G)Iso(H_{\mathbb{R}}/G) is finite, or it is a finite extension of S1S^{1}.

Proof.

By proposition 6.9 and Lemma 6.10, the projection of GG to Iso(2)Iso(\mathbb{R}^{2}) has a lattice Γ2\Gamma\subset\mathbb{R}^{2} as a finite index subgroup. This is equivalent to the fact that L=GHL=G\cap H_{\mathbb{R}} is a lattice in HH_{\mathbb{R}} and a finite index subgroup in GG. By Lemma 6.11, there are u,v2u,v\in\mathbb{R}^{2}, λ,r,s\lambda,r,s\in\mathbb{R}, with λ0\lambda\neq 0, and nn\in\mathbb{N}, such that Γ={ku+lv:k,l}\Gamma=\{ku+lv:k,l\in\mathbb{Z}\} and

L=(u,r),(v,s),(0,0,λn)H.L=\left\langle\left(u,r\right),\left(v,s\right),\left(0,0,\frac{\lambda}{n}\right)\right\rangle\subset H_{\mathbb{R}}.

As seen in Example 6.4, the group NH(L)={(npu+mpv,r):n,m,r}N_{H_{\mathbb{R}}}(L)=\left\{\left(\frac{n}{p}u+\frac{m}{p}v,r\right):n,m\in\mathbb{Z},\ r\in\mathbb{R}\right\} is the normalizer of LL in HH_{\mathbb{R}}. Denote by Aut(Γ)O(2)Aut(\Gamma)\subset O(2) the subgroup that preserves the lattice Γ\Gamma and observe that an element φ=σLgIso(H)\varphi=\sigma\circ L_{g}\in Iso(H_{\mathbb{R}}) satisfies that φLhφ1=Lσ(ghg1)\varphi\circ L_{h}\circ\varphi^{-1}=L_{\sigma(ghg^{-1})}. As σ(ghg1)\sigma(ghg^{-1}) and σ(h)\sigma(h) have the same projection onto Γ\Gamma, then if φ\varphi normalizes LL, σAut(Γ)\sigma\in Aut(\Gamma) and we have that

1KGFΓ1,1\rightarrow K\rightarrow G\rightarrow F^{\prime}\ltimes\Gamma\rightarrow 1,

for some subgroup FAut(Γ)F^{\prime}\subset Aut(\Gamma) and K=LK=L\cap\mathbb{R}. As HH_{\mathbb{R}} is normal in Iso(H)Iso(H_{\mathbb{R}}), we see that NIso(H)(G)NH(G)N_{Iso(H_{\mathbb{R}})}(G)\subset N_{H_{\mathbb{R}}}(G), and thus, by applying a trick as in Example 5.3, we may describe the greater normalizer as

1HNIso(H)(G)F′′Λ1,1\rightarrow H\rightarrow N_{Iso(H_{\mathbb{R}})}(G)\rightarrow F^{\prime\prime}\ltimes\Lambda\rightarrow 1,

with F′′Aut(Γ)F^{\prime\prime}\subset Aut(\Gamma) a finite group and Λ2\Lambda\subset\mathbb{R}^{2} a lattice containing Γ\Gamma. Thus, the isometry group is calculated as

1C=H/KNIso(H)(G)/GF=(F′′Λ)/(FΓ)1,1\rightarrow C=H/K\rightarrow N_{Iso(H_{\mathbb{R}})}(G)/G\rightarrow F=\left(F^{\prime\prime}\ltimes\Lambda\right)/\left(F^{\prime}\ltimes\Gamma\right)\rightarrow 1,

so that CC is either finite, cyclic or S1S^{1} and FF is finite. ∎

7. Spherical Geometry

This section is largely expository due to the fact that the verification of 2.8 in the spherical case consists of the comparison of the statement with the (fundamentally algebraic) classification of groups acting by isometries on three dimensional spherical manifolds and orbifolds. This concerns specifically the quotient orbifold of an action of a discrete group on a spherical three-manifold, that is, a quotient of the form

M=S3/Γ,M=S^{3}/\Gamma,

for Γ\Gamma a finite subgroup of O(4)O(4). The crucial point is that the classification of orbifolds up to orientation preserving isometry is equivalent to the classification of subgroups of O(4)O(4).

The following is a consequence of the classification of isometry groups of spherical 33- manifolds in [McC02], tables 2 and 3 in pages 173 and 176, relying on work of Mccullough and collaborators and ultimately going back to Seifert, Threlfall, Hopf and Hattori. See [HKMR12], chapter 1 for an account of these facts.

7.1. Classification

Lemma 7.1.

Up to finite subgroups, the isometry groups of spherical three manifolds are:

  • SO(3)SO(3).

  • O(2)O(2).

  • O(4)O(4).

  • SO(4)SO(4).

  • SO(3)SO(3).

  • O(2)×O(2)O(2)\times O(2)

  • S1×/2S1S^{1}\times_{\mathbb{Z}/2}S^{1}.

In particular, these subgroups can be realized as closed subgroups of O(4)O(4).

For a complete list of isometry groups of spherical orbifolds, see Chapter 3 of [MS19].

Connected components

An important result by Hatcher [Hat83], originally conjectured by Smale states that the inclusion of the isometry group of S3S^{3} into the group of diffeomorphisms is a homotopy equivalence.

The following result with contributions of many persons including (at least) Asano, Boileau, Bonahon, Birman, Cappell, Ivanov, Rubinstein, and Shaneson, is a consequence of research in mapping class groups and three- dimensional spherical manifolds. It is discussed with comments about attribution in [McC02], Theorem 1.1 in page 3.

Theorem 7.2.

Let MM be a spherical manifold, then the inclusion of the group of isometries of MM into the group of diffeomorphisms induces a bijection on path components.

As of 2022, the following result in page 2 of [BK19] is a consequence of the study via Ricci flow methods of the homotopy type of the spaces of positive scalar curvature and the subspace of metrics which are locally isometric to either the round sphere S3S^{3} or the round cylinder S2×S^{2}\times\mathbb{R}.

Theorem 7.3.

Let (M,g)(M,g) be a riemannian manifold which is an isometric quotient of the three dimensional round sphere. Then, the inclusion of the isometry group into the diffeomorphism group is a homotopy equivalence.

The following theorem was proved in [MS19], using previous analysis of the authors of Seifert fibrations for spherical orbifolds. It is a consequence of tables 2 in page 1302, table 3 in page 1304 and table 4 in page 1308.

Theorem 7.4.

Let XX be a spherical three-manifold, and let Γ\Gamma be a discrete group of X.

  • The isometry groups of the orbifold X/ΓX/\Gamma are either closed subgroups of SO4SO_{4} or PSO4PSO_{4}, if the action is orientation preserving.

  • The identity component of the isometry groups are S1S^{1}, S1×S1S^{1}\times S^{1} or trivial for the orientation preserving case.

End of the proof of Theorem 2.8 for the spherical geometry.

T he result thus follows from Lemma 7.1. ∎

8. S2×S^{2}\times\mathbb{R} geometry.

A three dimensional manifold is said to have S2×S^{2}\times\mathbb{R}-geometry if its universal covering is homeomorphic to S2×S^{2}\times\mathbb{R}.

The determination of the discrete isometry groups of spaces with S2×S^{2}\times\mathbb{R} geometry is a consequence of well-known facts, which we will gather here.

We recall the following result, proved in [KN96], Chapter VI, Theorem 3.5.

Theorem 8.1.

Given a product of riemannian manifolds M×NM\times N with MM of constant sectional curvature and NN flat, the isometry group of M×NM\times N decomposes as a direct product Iso(M)×Iso(N).{\rm Iso}(M)\times{\rm Iso}(N). It follows that for a discrete subgroup ΓIso(S2×)O(3)×(×2)\Gamma\leq{\rm Iso}(S^{2}\times\mathbb{R})\cong O(3)\times(\mathbb{R}\times\mathbb{Z}_{2}), the projection onto the second factor π(Γ)/2\pi_{\mathbb{R}}(\Gamma)\leq\mathbb{R}\rtimes\mathbb{Z}/2 is a discrete subgroup.

We will use this splitting and the classification of finite groups acting on S2×S^{2}\times\mathbb{R}, which was proved in by Tollefson in page 61 of [Tol74], as follows:

Theorem 8.2.

There exist only four three-manifolds covered by S2×S^{2}\times\mathbb{R},namely: S2×S1S^{2}\times S^{1}, the non orientable S2S^{2}- bundle over S1S^{1}, P1×S1\mathbb{R}P^{1}\times S^{1}, and P3#P2\mathbb{R}P^{3}\#\mathbb{R}P^{2}. Moreover, the finite groups which act freely on S2×S1S^{2}\times S^{1} are classified in [Tol74], Corollary 2. They are:

  • /p\mathbb{Z}/p, producing quotient spaces homeomorphic to S2×S1S^{2}\times S^{1} in the odd case, and P2\mathbb{R}P^{2} in the even case as quotient space.

  • /p×/2\mathbb{Z}/p\times\mathbb{Z}/2, for pp even, producing a quotient space homeomorphic to P2\mathbb{R}P^{2}, and

  • DnD_{n}, the dihedral group of order 2n2n, producing P3#P3\mathbb{R}P^{3}\#\mathbb{R}P^{3} as quotient space.

8.1. Classification of discrete groups of isometries.

The previous example gives us the general behaviour for discrete groups of isometries on S2×S^{2}\times\mathbb{R} as seen by the following Lemma

Lemma 8.3.

If ΓIso(S2×)\Gamma\subset Iso(S^{2}\times\mathbb{R}) is a discrete subgroup, then there is a finite group FO(3)F\subset O(3) and λ\lambda\in\mathbb{R} such that the exact sequence

1O(3)Iso(S2×)21\rightarrow O(3)\rightarrow Iso(S^{2}\times\mathbb{R})\rightarrow\mathbb{R}\rtimes\mathbb{Z}_{2}

induces an exact sequence 1FΓL1\rightarrow F\rightarrow\Gamma\rightarrow L, where LL is either λ\lambda\mathbb{Z} or λ2\lambda\mathbb{Z}\rtimes\mathbb{Z}_{2}.

Proof.

As the group O(3)O(3) is compact, the projection of the discrete group Γ\Gamma onto Iso()Iso(\mathbb{R}) is discrete, so it is of the form λ\lambda\mathbb{Z} or λ2\lambda\mathbb{Z}\rtimes\mathbb{Z}_{2}, for some λ\lambda\in\mathbb{R}. As O(3)O(3) can be seen as a closed subgroup of Iso(S2×)Iso(S^{2}\times\mathbb{R}), then the intersection of Γ\Gamma with O(3)O(3) is a finite group, which we denote by FF. The result thus follows from the product structure of Iso(S2×)Iso(S^{2}\times\mathbb{R}). In fact, Γ\Gamma is generated by FF, 2Iso()\mathbb{Z}_{2}\subset Iso(\mathbb{R}) and the twisted translation subgroup {(σn,nλ)O(3)×:n}\{(\sigma^{n},n\lambda)\in O(3)\times\mathbb{R}:n\in\mathbb{N}\}, for some σO(3)\sigma\in O(3). ∎

Connected Components

Theorem 8.4.

If ΓIso(S2×)\Gamma\subset Iso(S^{2}\times\mathbb{R}) is a discrete subgroup, such that (S2×)/Γ(S^{2}\times\mathbb{R})/\Gamma is compact, then Iso((S2×)/Γ)Iso((S^{2}\times\mathbb{R})/\Gamma) is up to finite index, a closed subgroup of SO(3)×S1SO(3)\times S^{1}. In particular, the connected component of the identity of the isometry group of the quotient can only be one of the three possibilities:

SO(3)×S1,S1×S1,orS1.SO(3)\times S^{1},\quad S^{1}\times S^{1},\quad\textrm{or}\quad S^{1}.
Proof.

By Lemma 8.3, the discrete group Γ\Gamma is generated by a finite group of O(3)O(3) and a twisted translation as in Example 8.5. The isometry group is compact, so it has a finite number of connected components and by Proposition 3.16, the connected component of the identity can be computed using the centralizer, which always contains the \mathbb{R}-factor, so the result follows by examining the possible connected, closed subgroups of SO(3)SO(3). ∎

8.2. Example of a non discrete subgroup of isometries

We may observe that the projection onto the S2S^{2} factor of a discrete group of isometries need not be discrete as the following example shows:

Example 8.5.

If σSO(3)\sigma\in SO(3) is a rotation with irrational angle along a fixed axis, so that the orbit {σn(p):n}\{\sigma^{n}(p):n\in\mathbb{N}\} is dense in a circle, orthogonal to the rotation axis, for almost every pS2p\in S^{2}, then the group given by twisted translations

{(σn,n)O(3)×:n}\{(\sigma^{n},n)\in O(3)\times\mathbb{R}:n\in\mathbb{N}\}

is a discrete subgroup of Iso(S2×)Iso(S^{2}\times\mathbb{R}) with non-discrete projection on Iso(S2)Iso(S^{2}).

9. Sol Geometry

9.1. Riemannian Geometry of Three Dimensional Sol-manifolds

Sol-geometry is given by the solvable Lie group of upper-triangular matrices

S={(et0x0ety001):x,y,t},S=\left\{\left(\begin{array}[]{ccc}e^{t}&0&x\\ 0&e^{-t}&y\\ 0&0&1\end{array}\right):x,y,t\in\mathbb{R}\right\},

which decomposes as a semidirect product S=[S,S](S/[S,S])2S=[S,S]\rtimes(S/[S,S])\cong\mathbb{R}^{2}\rtimes\mathbb{R}. In global coordinates, the vector fields

X1(x,y,t)=(et,0,0),X2(x,y,t)=(0,et,0),X3(x,y,t)=(0,0,1),X_{1}(x,y,t)=(e^{t},0,0),\quad X_{2}(x,y,t)=(0,e^{-t},0),\quad X_{3}(x,y,t)=(0,0,1),

define a basis of left-invariant vector fields. We choose the left invariant Riemannian metric in SS having this basis as orthonormal basis, so that in our global coordinates, the metric has the expresion

ds2=e2tdx2+e2tdy2+dt2.ds^{2}=e^{-2t}dx^{2}+e^{2t}dy^{2}+dt^{2}.

The isometry group of this metric is generated by left translations

Lg:SS,Lg(h)=gh,L_{g}:S\rightarrow S,\qquad L_{g}(h)=gh,

and the group of reflections (x,y,z)(±x,±y,±z)(x,y,z)\rightarrow(\pm x,\pm y,\pm z), isomorphic to the Dihedral group D4D_{4}. In particular, Iso(S)Iso(S) has eight connected components, with the connected component of the identity isomorphic to SS [Sco83].

9.2. Existence of lattices

Consider the following

Remark 9.1.

a Lie group admits a lattice subgroup if and only if it is unimodular [Rag07], and so, not every solvable group admit lattice subgroups.

Example 9.2.

A solvable group which is closely related to SS considered here, is the group of orientation preserving affine transformations on \mathbb{R}, given by

Aff+(){(etx01):x,t}.\mathrm{Aff}^{+}(\mathbb{R})\cong\left\{\left(\begin{array}[]{cc}e^{t}&x\\ 0&1\end{array}\right):x,t\in\mathbb{R}\right\}.

We could try for example, to exponentiate the set

Λ=exp({(nm00):n,m})={(en(m/n)(en1)01):n,m},\Lambda=exp\left(\left\{\left(\begin{array}[]{cc}n&m\\ 0&0\end{array}\right):n,m\in\mathbb{N}\right\}\right)=\left\{\left(\begin{array}[]{cc}e^{n}&(m/n)(e^{n}-1)\\ 0&1\end{array}\right):n,m\in\mathbb{N}\right\},

however, such discrete set is not a subgroup and the group which generates is not discrete. The problem is that the group Aff+()\mathrm{Aff}^{+}(\mathbb{R}) is not unimodular, and in fact its modular function has the expression

Δ:Aff+()+,Δ(etx01)=et,\Delta:\mathrm{Aff}^{+}(\mathbb{R})\rightarrow\mathbb{R}_{+},\qquad\Delta\left(\begin{array}[]{cc}e^{t}&x\\ 0&1\end{array}\right)=e^{t},

which is non-trivial.

The solvable group SS is unimodular, so that it admits a lattice subgroup and an explicit way to construct a lattice is as follows: Consider a matrix ASL2()A\in SL_{2}(\mathbb{Z}), such that tr(A)>2tr(A)>2 and the group

ΓA={(AnZ01):n,ZM2×1()}2A.\Gamma_{A}=\left\{\left(\begin{array}[]{cc}A^{n}&Z\\ 0&1\end{array}\right):n\in\mathbb{Z},\ Z\in M_{2\times 1}(\mathbb{Z})\right\}\cong\mathbb{Z}^{2}\rtimes_{A}\mathbb{Z}.
Lemma 9.3.

For every ASL2()A\in SL_{2}(\mathbb{Z}), with tr(A)>2\mathrm{tr}(A)>2, ΓA\Gamma_{A} is conjugated in SL3()SL_{3}(\mathbb{R}) to a lattice subgroup in SS, moreover, every lattice subgroup of SS is conjugated to one of such groups.

Proof.

Suppose first that ASL2()A\in SL_{2}(\mathbb{Z}), with tr(A)>2\mathrm{tr}(A)>2, then there is a matrix BSL3()B\in SL_{3}(\mathbb{R}) such that

BAB1=(eλ00eλ),BAB^{-1}=\left(\begin{array}[]{cc}e^{\lambda}&0\\ 0&e^{-\lambda}\end{array}\right),

for some λ0\lambda\neq 0. We may define At=B1(etλ00etλ)BA^{t}=B^{-1}\left(\begin{array}[]{cc}e^{t\lambda}&0\\ 0&e^{-t\lambda}\end{array}\right)B, so that ΓA\Gamma_{A} is a discrete subgroup of the group

SA={(AtZ01):t,ZM2×1()}2,S_{A}=\left\{\left(\begin{array}[]{cc}A^{t}&Z\\ 0&1\end{array}\right):t\in\mathbb{R},\ Z\in M_{2\times 1}(\mathbb{R})\right\}\cong\mathbb{R}\ltimes\mathbb{R}^{2},

such that

12/2SA/ΓA/1,1\rightarrow\mathbb{R}^{2}/\mathbb{Z}^{2}\rightarrow S_{A}/\Gamma_{A}\rightarrow\mathbb{R}/\mathbb{Z}\rightarrow 1,

thus, ΓA\Gamma_{A} is a lattice in SAS_{A}. Observe that we have an isomorphisms of Lie groups via the conjugation

SAS,X(B001)X(B1001),S_{A}\rightarrow S,\qquad X\mapsto\left(\begin{array}[]{cc}B&0\\ 0&1\end{array}\right)X\left(\begin{array}[]{cc}B^{-1}&0\\ 0&1\end{array}\right),

and thus, a lattice in SS.

Before proceeding with the proof of this Lemma, we need to prove the following discrete projection Lemma:

Lemma 9.4.

If ΓS\Gamma\subset S is a discrete subgroup, then its projection Γ¯S/[S,S]\overline{\Gamma}\subset S/[S,S]\cong\mathbb{R} is also discrete.

Proof.

An element γ=(x,y,t)\gamma=(x,y,t), with t0t\neq 0, acts discretely by translations on the line {(x1et,y1et,s):s}S\left\{\left(\frac{x}{1-e^{t}},\frac{y}{1-e^{-t}},s\right):s\in\mathbb{R}\right\}\subset S, as γn(x1et,y1et,s)=(x1et,y1et,s+nt)\gamma^{n}\left(\frac{x}{1-e^{t}},\frac{y}{1-e^{-t}},s\right)=\left(\frac{x}{1-e^{t}},\frac{y}{1-e^{-t}},s+nt\right). Thus, if Γ\Gamma is commutative, either Γ2\Gamma\subset\mathbb{R}^{2} and its projection is trivial, or Γ\Gamma preserves a unique line on which it acts as translations and the action on this line is precisely the action on \mathbb{R} of its projection, which must be discrete. If Γ\Gamma is non-commutative, then at least it has two elements u=(a,b,0)u=(a,b,0) and γ=(x,y,t)\gamma=(x,y,t) with t0t\neq 0. Observe that if b=0b=0, then γuγ1=(etaa,0,0)\gamma u\gamma^{-1}=(e^{t}a-a,0,0) and by iterating conjugation we get a non-discrete subgroup of 2Γ\mathbb{R}^{2}\cap\Gamma which is impossible and the same goes for the case a=0a=0. Thus a,b0a,b\neq 0 and Γ2\Gamma\cap\mathbb{R}^{2} contains two linearly independent vectors, say uu and v=γuγ1=(eta,etb,0)v=\gamma u\gamma^{-1}=(e^{t}a,e^{-t}b,0), which implies that Γ2\Gamma\cap\mathbb{R}^{2} is cocompact in 2\mathbb{R}^{2}. Γ/(Γ2)\Gamma/(\Gamma\cap\mathbb{R}^{2}) is discrete in S/(Γ2)S/(\Gamma\cap\mathbb{R}^{2}) and the projection S/(Γ2)S/2S/(\Gamma\cap\mathbb{R}^{2})\rightarrow S/\mathbb{R}^{2} has compact Kernel 2/(Γ2)S1×S1\mathbb{R}^{2}/(\Gamma\cap\mathbb{R}^{2})\cong S^{1}\times S^{1}, thus the corresponding projection of Γ/(Γ2)\Gamma/(\Gamma\cap\mathbb{R}^{2}) into S/[S,S]S/[S,S] is discrete. ∎

Suppose now that ΓS\Gamma\subset S is a lattice subgroup, then by Lemma 9.4, the Γ\Gamma projects to a non-trivial discrete group \mathbb{R}, generated by an element enβe^{n\beta}, with β0\beta\neq 0. The intersection Γ2\Gamma\cap\mathbb{R}^{2} is a lattice, so that there are u,v2u,v\subset\mathbb{R}^{2} linearly independent, such that Γ2={nu+mv:n,m}\Gamma\cap\mathbb{R}^{2}=\{nu+mv:n,m\in\mathbb{Z}\}. Take CGL2()C\in GL_{2}(\mathbb{R}) the matrix sending Γ2\Gamma\cap\mathbb{R}^{2} onto the canonical lattice 2\mathbb{Z}^{2} and define the matrix A=B(eβ00eβ)B1A^{\prime}=B\left(\begin{array}[]{cc}e^{\beta}&0\\ 0&e^{-\beta}\end{array}\right)B^{-1}. An element g=(B1ABW01)Γg=\left(\begin{array}[]{cc}B^{-1}A^{\prime}B&W\\ 0&1\end{array}\right)\in\Gamma must preserve Γ2\Gamma\cap\mathbb{R}^{2}, so that the element h=(ABW01)Γh=\left(\begin{array}[]{cc}A^{\prime}&BW\\ 0&1\end{array}\right)\in\Gamma must preserve 2\mathbb{Z}^{2}. Observe that the action of hh in an element v=(n,m)2v=(n,m)\in\mathbb{Z}^{2} is Av+BW2A^{\prime}v+BW\in\mathbb{Z}^{2}, this implies that BW2BW\in\mathbb{Z}^{2} and ASL2()A^{\prime}\in SL_{2}(\mathbb{Z}). In particular, the group Γ\Gamma is isomorphic to the group ΓA\Gamma_{A^{\prime}} and the isomorphism is obtained by conjugation. ∎

Remark 9.5.

The existence of lattices in the Lie group SS is related to the existence of a \mathbb{Q}-structure on SS. More precisely, if ASL2()A\in SL_{2}(\mathbb{Z}), with tr(A)>2\mathrm{tr}(A)>2, and c=tr(A)24c=\sqrt{\mathrm{tr}(A)^{2}-4}, then AA is diagonalizable over the field (c)\mathbb{Q}(c), that is, there is a matrix BSL2((c))B\in SL_{2}(\mathbb{Q}(c)) such that BAB1BAB^{-1} is diagonal. If Qij(X)=XijQ_{ij}(X)=X_{ij} is the linear map that gives the (i,j)(i,j)-entry, then the group

𝔾(k)={(XZ01):ZM2×1(k),Qij(BXB1)=0,ij,i,j{1,2}},\mathbb{G}(k)=\left\{\left(\begin{array}[]{cc}X&Z\\ 0&1\end{array}\right):Z\in M_{2\times 1}(k),\ Q_{ij}(BXB^{-1})=0,\ i\neq j,i,j\in\{1,2\}\right\},

is algebraic subgroup of SL3()SL_{3}(\mathbb{R}), defined by polynomial equations with coefficients over (c)\mathbb{Q}(c), such that 𝔾()S\mathbb{G}(\mathbb{R})\cong S and 𝔾()=ΓA\mathbb{G}(\mathbb{Z})=\Gamma_{A}. Moreover, the Galois automorphism σ:(c)(c)\sigma:\mathbb{Q}(c)\rightarrow\mathbb{Q}(c), defined by σ(c)=c\sigma(c)=-c, has a natural extension to automorphisms of matrices and polynomials, so that we have the embedding

SL3((c))SL3()×SL3(),Y(Y,σ(Y)),SL_{3}(\mathbb{Q}(c))\rightarrow SL_{3}(\mathbb{R})\times SL_{3}(\mathbb{R}),\qquad Y\mapsto(Y,\sigma(Y)),

and a polynomial condition Q(Y)=0Q(Y)=0 on YSL3((c))Y\in SL_{3}(\mathbb{Q}(c)) is equivalent to the pair of polynomial conditions Q(Y)+σ(Q)(Y)=0Q(Y)+\sigma(Q)(Y^{\prime})=0 and Q(Y)σ(Q)(Y)=0Q(Y)\sigma(Q)(Y^{\prime})=0 on (Y,Y)SL3()×SL3()(Y,Y^{\prime})\in SL_{3}(\mathbb{R})\times SL_{3}(\mathbb{R}), but the latter are polynomials with coefficients over \mathbb{Q} (this trick is called “restriction of scalars” [Mor15]).

Lemma 9.6.

SS has trivial center and the centralizer of a lattice group ΓS\Gamma\subset S is also trivial.

Proof.

Take γ=(x,y,t)\gamma=(x,y,t) in the centralizer of Γ\Gamma in SS, then as in the previous proposition Γ[S,S]\Gamma\cap[S,S] has a rank two subgroup, thus it contains at least a vector u=(a,b,0)u=(a,b,0) such that a,b0a,b\neq 0 and we have

γuγ1=(eta,etb,0)=(a,b,0),\gamma u\gamma^{-1}=(e^{t}a,e^{-t}b,0)=(a,b,0),

which implies that t=0t=0. As Γ\Gamma projects to a lattice group in S/[S,S]S/[S,S]\cong\mathbb{R}, then there is a βΓ\beta\in\Gamma such that β=(c,d,s)\beta=(c,d,s) with s0s\neq 0 and thus

βγβ1=(esx,esy,0)=γ=(x,y,0)\beta\gamma\beta^{-1}=(e^{s}x,e^{-s}y,0)=\gamma=(x,y,0)

which implies that x=y=0x=y=0 and γ\gamma is the identity. A completely analogous computation shows that SS has trivial center. ∎

Corollary 9.7.

If Γ\Gamma is a discrete group of isometries of SS such that S/ΓS/\Gamma has finite volume, then S/ΓS/\Gamma is compact and has finite isometry group.

Proof.

As the connected component of the isometry group of SS is SS itself acting by left multiplications, Γ\Gamma is modulo a finite index subgroup a lattice in SS and it lies in an exact sequence

1Γ0ΓΓ111\rightarrow\Gamma_{0}\rightarrow\Gamma\rightarrow\Gamma_{1}\rightarrow 1

where Γ0=Γ[S,S]\Gamma_{0}=\Gamma\cap[S,S] and Γ/Γ0Γ1\Gamma/\Gamma_{0}\cong\Gamma_{1}\subset\mathbb{R}. By Proposition 9.4 Γ1\Gamma_{1} is a discrete subgroup, so this exact sequence induces a the fiber bundle

2/Γ0S/ΓR/Γ0,\mathbb{R}^{2}/\Gamma_{0}\rightarrow S/\Gamma\rightarrow R/\Gamma_{0},

so that S/ΓS/\Gamma has finite volume if and only if 2/Γ0\mathbb{R}^{2}/\Gamma_{0} and /Γ1\mathbb{R}/\Gamma_{1} are torus of the corresponding dimension and S/ΓS/\Gamma is compact. The isometry group of S/ΓS/\Gamma is a compact Lie group with connected component of the identity determined by the centralizer of Γ\Gamma in SS (Proposition 3.16) which is the trivial group by Lemma 9.6, thus the isometry group is a compact, zero-dimensional Lie group, i.e. finite. ∎

9.3. Examples

Example 9.8.

For A=(2111)A=\left(\begin{array}[]{cc}2&1\\ 1&1\end{array}\right) and nn\in\mathbb{N}, consider the lattice ΓAn=2An\Gamma_{A^{n}}=\mathbb{Z}^{2}\rtimes_{A^{n}}\mathbb{Z}. A matrix Y=(MW01)GL3()Y=\left(\begin{array}[]{cc}M&W\\ 0&1\end{array}\right)\in GL_{3}(\mathbb{R}) normalizes ΓAn\Gamma_{A^{n}} if and only if M=AkM=A^{k} for some kk\in\mathbb{Z} and (IAn)W2(I-A^{n})W\in\mathbb{Z}^{2}, so that if Λn=(IAn)12\Lambda_{n}=(I-A^{n})^{-1}\mathbb{Z}^{2}, then the normalizer is NIso(S)(ΓAn)=AΛnN_{Iso(S)}(\Gamma_{A^{n}})=\mathbb{Z}\ltimes_{A}\Lambda_{n} and the isometry group is computed as

Iso(S/ΓAn)=(Λn/2)An.Iso(S/\Gamma_{A^{n}})=(\Lambda_{n}/\mathbb{Z}^{2})\rtimes_{A}\mathbb{Z}_{n}.

Three ilustrative cases are

  1. (i)

    Λ1=2\Lambda_{1}=\mathbb{Z}^{2}, so that Iso(S/ΓA)Iso(S/\Gamma_{A}) is trivial;

  2. (ii)

    det(IA2)=5\mathrm{det}(I-A^{2})=-5, so that 2Λ2152\mathbb{Z}^{2}\leq\Lambda_{2}\leq\frac{1}{5}\mathbb{Z}^{2} and each contention is of index 55, in particular we have that Iso(S/ΓA2)=52Iso(S/\Gamma_{A^{2}})=\mathbb{Z}_{5}\rtimes\mathbb{Z}_{2};

  3. (iii)

    Λ5=1112\Lambda_{5}=\frac{1}{11}\mathbb{Z}^{2}, so that Iso(S/ΓA5)=(11×11)A5Iso(S/\Gamma_{A^{5}})=(\mathbb{Z}_{11}\times\mathbb{Z}_{11})\rtimes_{A}\mathbb{Z}_{5}.

9.4. Classification of free actions

Remark 9.9.

The previous family of examples exhibits isometric actions of each finite cyclic group on a three dimensional solvmanifold. Moreover, we should notice that such actions are necesarily not free, since there exists a very rigid classification of free actions of finite groups on three dimensional manifolds with Nil and Sol structure, based on pp-rank estimates and P.A. Smith Theory, [JL10], [KOS17].

10. Hyperbolic geometry

10.1. Normalizers of Fuchsian groups.

Denote by n\mathbb{H}^{n} the n-dimensional hyperbolic space and recall that the isometry group Iso(n)Iso(\mathbb{H}^{n}) is a non-compact semisimple Lie group that can be identified with the group PO(n,1)PO(n,1). We begin by recalling the following properties of normalizers of discrete subgroups of isometries.

Lemma 10.1.

If ΓIso(n)\Gamma\subset Iso(\mathbb{H}^{n}) is a discrete subgroup such that n/Γ\mathbb{H}^{n}/\Gamma has finite volume, then the normalizer group

Λ={gIso(n):ghg1=h,hΓ}Iso(n)\Lambda=\{g\in Iso(\mathbb{H}^{n}):ghg^{-1}=h,\ \forall\ h\in\Gamma\}\subset Iso(\mathbb{H}^{n})

is discrete and Iso(n/Γ)Iso(\mathbb{H}^{n}/\Gamma) is a finite group.

Proof.

Passing to a finite cover doesn’t alter the outcome, so we may suppose that Γ,ΛO(n,1)\Gamma,\Lambda\subset O(n,1). By Proposition 3.16, the connected component of Λ\Lambda lies inside the centralizer of Γ\Gamma in O(n,1)O(n,1). Let gO(n,1)g\in O(n,1) centralizing Γ\Gamma, then the polynomial

Pt:Mn+1()Mn+1(),Pt(X)=gXg1XP_{t}:M_{n+1}(\mathbb{R})\rightarrow M_{n+1}(\mathbb{R}),\qquad P_{t}(X)=gXg^{-1}-X

vanishes at Γ\Gamma but by Borel’s density Theorem (see [Fur76]), Γ\Gamma is Zariski dense in O(n,1)O(n,1) and thus Pt(O(n,1))=0P_{t}(O(n,1))=0 which tells us that gg lies in the center of O(n,1)O(n,1), which is finite. This tells us that Λ\Lambda is a discrete group that contains the lattice Γ\Gamma, so Λ\Lambda is also a lattice in O(n,1)O(n,1). If FΛ,FΓnF_{\Lambda},F_{\Gamma}\subset\mathbb{H}^{n} are fundamental domains of the groups Λ\Lambda and Γ\Gamma correspondingly, so we have that

|Iso(n/Γ)|=|Λ/Γ|=Vol(FΓ)/Vol(FΛ)<.|Iso(\mathbb{H}^{n}/\Gamma)|=|\Lambda/\Gamma|=Vol(F_{\Gamma})/Vol(F_{\Lambda})<\infty.

Remark 10.2.

The previous result is stated for hyperbolic manifolds in Corollary 3, Section 12.7 of [Rat19] and for hyperbolic orbifolds in [Rat99], where the hypotheses are that the discrete group is non elementary, geometrically finite and without fixed mm-planes, for m<n1m<n-1. In Lemma 10.1 we presented an argument using Zariski-density of the lattice group in Iso(n)Iso(\mathbb{H}^{n}), which implies for example the non-existence of fixed mm-planes. As seen in [Gre74], every finite group can be realized as the isometry group of a compact hyperbolic surface as in Lemma 10.1.

10.2. Rank of isometries and Lie groups acting on hyperbolic surfaces.

Lemma 10.3.

If Σ\Sigma is a compact, orientable surface of genus g2g\geq 2, then there are no faithful actions of the compact group S1S^{1} on Σ\Sigma.

Proof.

Suppose there is a faithful action S1×ΣΣS^{1}\times\Sigma\rightarrow\Sigma, then perhaps after an averaging process, we may suppose that the action is isometric with respect to a Riemannian metric hh. The existence of isothermal coordinates [UY17] tells us that there exists a complex structure in Σ\Sigma such that in holomorphic coordinates z=x+iyz=x+iy, the vector fields x\partial_{x} and y\partial_{y} are hh-orthogonal. As the S1S^{1}-action is hh-isometric, it preserves angles and orientation in the isothermal coordinates and thus it is an action by holomorphic transformations. By the uniformization Theorem, the universal cover of Σ\Sigma is the hyperbolic semiplane 2\mathbb{H}^{2}\subset\mathbb{C} and the holomorphic automorphisms of Σ\Sigma lift to holomorphic automorphisms of 2\mathbb{H}^{2} which also are isometric automorphisms with respect to the hyperbolic metric. As a consequence of this, we have that the S1S^{1}-action preserves a hyperbolic metric in Σ\Sigma which has finite volume, because Σ\Sigma is compact, but this contradicts Lemma 10.1. ∎

Corollary 10.4.

If Σ\Sigma is a compact, orientable surface of genus g2g\geq 2 and hh is a Riemannian metric in Σ\Sigma, then the isometry group Iso(Σ,h)Iso(\Sigma,h) is finite.

Proof.

As Σ\Sigma is compact, the isometry group G=Iso(Σ,h)G=Iso(\Sigma,h) is a compact Lie group. If 𝔤\mathfrak{g} denotes the Lie algebra of GG, then for every X𝔤X\in\mathfrak{g}, the one parameter group {exp(tX)}\{exp(tX)\} is a commutative group whose closure is a compact, commutative Lie group with connected component of the identity isomorphic to a product S1××S1S^{1}\times\cdots\times S^{1}. As a consequence of this and the fact that GG has only has finitely many connected components, if GG is infinite, then it has a closed subgroup isomorphic to S1S^{1}, but this is impossible as is shown in Lemma 10.3. ∎

10.3. Non-Classification of finite hyperbolic groups of isometries.

Remark 10.5.

It is proved in [Koj88] that every finite group can be realized as the isometry group of a closed hyperbolic manifold of dimension three.

11. Finer classification of 2-dimensional hyperbolic isometries.

Recall that in dimension two, the group SL2()SL_{2}(\mathbb{R}) acts on 2\mathbb{H}^{2} by isometries in the form of Möbius transformations, so that we have a realization of the orientation preserving isometries as Iso(2)PSL2()Iso(\mathbb{H}^{2})\cong PSL_{2}(\mathbb{R}).

11.1. Classification of elements in SL2SL_{2} according to their fixed point sets on the visual compactification of 2\mathbb{H}^{2}.

We recall the classification of elements in SL2()SL_{2}(\mathbb{R}) An element ASL2()A\in SL_{2}(\mathbb{R}) has as a characteristic polynomial pA(x)=x2tr(A)x+1p_{A}(x)=x^{2}-tr(A)x+1, and discriminant tr(A)24tr(A)^{2}-4. Thus, there are three dynamically different possibilities for the isometry of 2\mathbb{H}^{2} generated by AA, characterized by the sign of tr(A)2tr(A)-2:

  • tr(A)2>0tr(A)-2>0, where the matrix is conjugated to a diagonal matrix over \mathbb{R}, and thus, the conjugated isometry is contained in the one parameter group of isometries generated by

    {exp(t00t)=(et00et):t}.\left\{\mathrm{exp}\left(\begin{array}[]{cc}t&0\\ 0&-t\end{array}\right)=\left(\begin{array}[]{cc}e^{t}&0\\ 0&e^{-t}\end{array}\right):t\in\mathbb{R}\right\}.

    One isometry of this type is called hyperbolic, and the one-parameter group generated by this matrix is characterized by the property of having two fixed points in the boundary S1=2S^{1}=\partial\mathbb{H}^{2} and preserves a foliation determined by the two points and guided by the geodesic that joints the two points (in the case of diagonal matrices, this is {0,}\{0,\infty\}).

  • tr(A)2=0tr(A)-2=0, where the matrix is conjugated over \mathbb{R} to an upper triangular matrix, and thus, the conjugated isometry is contained in the one parameter group of isometries generated by

    {exp(0t00)=(1t01):t}.\left\{\mathrm{exp}\left(\begin{array}[]{cc}0&t\\ 0&0\end{array}\right)=\left(\begin{array}[]{cc}1&t\\ 0&1\end{array}\right):t\in\mathbb{R}\right\}.

    One isometry of this type is called parabolic, and the one-parameter group generated by this matrix is characterized by the property of having one fixed point in the boundary 2\partial\mathbb{H}^{2} and preserving the foliation of horocycles tangent to the fixed point (in the upper triangular case, the horocycles that are tangent to \infty are just horizontal lines).

  • tr(A)2<0tr(A)-2<0, where the matrix is conjugated over \mathbb{R} to a rotation matrix, so that the conjugated isometry is contained in the one parameter group of isometries generated by

    {exp(0tt0)=(cos(t)sin(t)sin(t)cos(t)):t}.\left\{\mathrm{exp}\left(\begin{array}[]{cc}0&-t\\ t&0\end{array}\right)=\left(\begin{array}[]{cr}\cos(t)&-\sin(t)\\ \sin(t)&\cos(t)\end{array}\right):t\in\mathbb{R}\right\}.

    One isometry of this type is called elliptic, and the one-parameter group generated by this matrix is characterized by the property of having one fixed point in the interior of 2\mathbb{H}^{2} and preserving a foliation of circles.

Lemma 11.1.

If α,βPSL2()\alpha,\beta\in PSL_{2}(\mathbb{R}) are two non-trivial elements, then

  1. (i)

    α\alpha and β\beta commute if and only if Fix(α)=Fix(β)Fix(\alpha)=Fix(\beta),

  2. (ii)

    C(α)={βPSL2():αβ=βα}={exp(tX):t}C(\alpha)=\{\beta\in PSL_{2}(\mathbb{R}):\alpha\beta=\beta\alpha\}=\{exp(tX):t\in\mathbb{R}\}, for some X𝔰𝔩2()X\in\mathfrak{sl}_{2}(\mathbb{R}). In particular C(α)C(\alpha) is isomorphic to either \mathbb{R} or S1S^{1}.

Proof.

Suppose αβ=βα\alpha\beta=\beta\alpha, then β(Fix(α))=Fix(α)\beta(Fix(\alpha))=Fix(\alpha) and α(Fix(β))=Fix(β)\alpha(Fix(\beta))=Fix(\beta). If α\alpha is parabolic or elliptic, then it has only one fixed point and thus Fix(α)=Fix(β)Fix(\alpha)=Fix(\beta) and the same applies for β\beta either parabolic or elliptic. In the case where both α\alpha and β\beta are hyperbolic, we observe that β\beta cannot interchange two distinct elements of the boundary S1S^{1}, thus the property β(Fix(α))=Fix(α)\beta(Fix(\alpha))=Fix(\alpha) implies Fix(α)=Fix(β)Fix(\alpha)=Fix(\beta). On the other hand, if α\alpha and β\beta have the same set of fixed points, then they are elements of the same one-parameter group, this is obvious when the fixed points are in standard configuration, that is {0,}\{0,\infty\}, {}\{\infty\} or {i}\{i\} according if the element is hyperbolic, parabolic or elliptic; and in general it can be seen via a conjugation of matrices by sending the fixed points to the standard configuration. In particular αβ=βα\alpha\beta=\beta\alpha, because a one-parameter group is commutative and the result follows. ∎

11.2. Discrete subgroups of Isometries of SL2()SL_{2}(\mathbb{R}).

Corollary 11.2.

If ΓPSL2()\Gamma\subset PSL_{2}(\mathbb{R}) is a subgroup such that it has the identity element as an accumulation point (equivalently Γ\Gamma is not a discrete subgroup) and ΛΓ\Lambda\subset\Gamma is a non-trivial, normal and discrete subgroup, then there exists Γ1Γ\Gamma_{1}\subset\Gamma commutative subgroup of finite index.

Proof.

Λ\Lambda is cyclic. As Λ\Lambda is normal, for every γΓ\gamma\in\Gamma, the conjugation induces an automorphism

ΛΛ,gγgγ1,\Lambda\rightarrow\Lambda,\qquad g\mapsto\gamma g\gamma^{-1},

and as Λ\Lambda is discrete and Γ\Gamma has the identity element as an accumulation point, for every FΛF\subset\Lambda finite set, there exist γΓ\gamma\in\Gamma close enough to the identity such that γe\gamma\neq e and γg=gγ\gamma g=g\gamma, for every gFg\in F. By the Lemma 11.1, the group generated by FF is a discrete subgroup of the one-parameter group C(γ)C(\gamma) and thus it is a cyclic group. For F1F2ΛF_{1}\subset F_{2}\subset\Lambda any two distinct finite subsets, there are elements gjΛg_{j}\in\Lambda such that gj=Fj\langle g_{j}\rangle=\langle F_{j}\rangle and F1F2\langle F_{1}\rangle\subset\langle F_{2}\rangle which implies that g1=g2kg_{1}=g_{2}^{k} for some kk and in particular 0<|g2|<|g1|0<|g_{2}|<|g_{1}|. Now Λ\Lambda must be cyclic because otherwise we would have a sequence {gj}Λ\{g_{j}\}\subset\Lambda obtained as the generators of subgroups generated by an increasing tower of finite subsets of Λ\Lambda that converge to the identity.

Existence of Γ1\Gamma_{1}. Take αΛ\alpha\subset\Lambda a generator of the group and as γαγ1\gamma\alpha\gamma^{-1} is again a generator of Λ\Lambda, for every γΓ\gamma\in\Gamma, then the subgroup

Γ1={γΓ:γαγ1=α}\Gamma_{1}=\{\gamma\in\Gamma:\gamma\alpha\gamma^{-1}=\alpha\}

is a finite index subgroup of Γ\Gamma ([Γ:Γ1]2[\Gamma:\Gamma_{1}]\leq 2 if Λ\Lambda\cong\mathbb{Z}, and [Γ:Γ1]|Λ|[\Gamma:\Gamma_{1}]\leq|\Lambda| if Λ/m\Lambda\cong\mathbb{Z}/m\mathbb{Z}). Finally, by the Lemma 11.1, Γ1\Gamma_{1} is commutative and the result follows. ∎

11.3. Non-classification of finite groups of isometries

Remark 11.3.

It is proved in [Gre74] that every finite group can be realized as the isometry group of a compact hyperbolic surface.

12. 2×\mathbb{H}^{2}\times\mathbb{R}

Recall [KN96], Chapter VI, Theorem 3.5 that given a product of riemannian manifolds M×NM\times N with MM of constant sectional curvature and NN flat, the isometry group of M×NM\times N decomposes as a direct product, Iso(M)×Iso(N).{\rm Iso}(M)\times{\rm Iso}(N). The following result gives us the isometry groups of finite volume quotients of 2×\mathbb{H}^{2}\times\mathbb{R} (see Theorem 13.6 for another proof):

12.1. Isometry groups of finite volume.

Theorem 12.1.

If GIso(2×)G\subset Iso(\mathbb{H}^{2}\times\mathbb{R}) is a discrete subgroup such that (2×)/G(\mathbb{H}^{2}\times\mathbb{R})/G has finite volume, then the group Iso((2×)/G)Iso((\mathbb{H}^{2}\times\mathbb{R})/G) is a finite extension of S1S^{1}

Proof.

Consider the exact sequence

1KGΓ1,1\rightarrow K\rightarrow G\rightarrow\Gamma\rightarrow 1,

where K=GIso()K=G\cap Iso(\mathbb{R}) is a discrete subgroup of GG and ΓG/K\Gamma\cong G/K is a subgroup of isometries of 2\mathbb{H}^{2}. If Γ\Gamma is discrete as a subgroup of Iso(2)Iso(\mathbb{H}^{2}), then 2/Γ\mathbb{H}^{2}/\Gamma is an hyperbolic orbifold such that

/K(2×)/G2/Γ\mathbb{R}/K\rightarrow(\mathbb{H}^{2}\times\mathbb{R})/G\rightarrow\mathbb{H}^{2}/\Gamma

is a locally trivial fiber bundle and as (2×)/G(\mathbb{H}^{2}\times\mathbb{R})/G has finite volume, then /KS1\mathbb{R}/K\cong S^{1} and Γ\Gamma is a Lattice subgroup of Iso(2)Iso(\mathbb{H}^{2}). In this case, we have an exact sequence of isometry groups

1Iso(S1)Iso(2×/G)Iso(2/Γ)1,1\rightarrow Iso(S^{1})\rightarrow Iso(\mathbb{H}^{2}\times\mathbb{R}/G)\rightarrow Iso(\mathbb{H}^{2}/\Gamma)\rightarrow 1,

where Iso(2/Γ)Iso(\mathbb{H}^{2}/\Gamma) is a finite group by Lemma 10.1 and thus Iso(2×/G)Iso(\mathbb{H}^{2}\times\mathbb{R}/G) is a finite extension of S1S^{1}.

If Γ\Gamma is not discrete as a subgroup of Iso(2)Iso(\mathbb{H}^{2}), we can see that the quotient (2×)/G(\mathbb{H}^{2}\times\mathbb{R})/G cannot have finite volume. To see this, first observe that we have another exact sequence

1ΛGL1,1\rightarrow\Lambda\rightarrow G\rightarrow L\rightarrow 1,

where Λ=GIso(2)Γ\Lambda=G\cap Iso(\mathbb{H}^{2})\subset\Gamma is a discrete, normal subgroup and G/ΛLIso()G/\Lambda\cong L\subset Iso(\mathbb{R}). If Λ=0\Lambda=0, then GLG\cong L is commutative and thus Γ\Gamma is commutative. If instead Λ\Lambda is non-trivial, then Corollary 11.2 tells us again that Γ\Gamma is commutative (perhaps after passing to a finite index subgroup). In any case, GG leaves a closed surface ζ×2×\zeta\times\mathbb{R}\subset\mathbb{H}^{2}\times\mathbb{R} fixed, where ζ\zeta is a geodesic, an horocycle or a circle (corresponding to the type of the isometries of Γ\Gamma). If Γ\Gamma consists of parabolic or hyperbolic elements, then Γ\Gamma acts discretely by Euclidean automorphisms in ζ×2\zeta\times\mathbb{R}\cong\mathbb{R}^{2} so that by Bieberbach Theorem [Rat19], Γ\Gamma contains a finite index subgroup isomorphic to a subgroup of 2\mathbb{Z}^{2} and in particular the fundamental domain of the GG-action in 2×\mathbb{H}^{2}\times\mathbb{R} contains a subset isometric to

{x+iy:a<x<b}×[c,d]2×,\{x+iy:a<x<b\}\times[c,d]\subset\mathbb{H}^{2}\times\mathbb{R},

this implies that (2×)/G(\mathbb{H}^{2}\times\mathbb{R})/G doesn’t have finite volume. If Γ\Gamma consists of elliptic elements, then GG acts discretely by Euclidean automorphisms in ζ×S1×\zeta\times\mathbb{R}\cong S^{1}\times\mathbb{R}, and thus as in the previous case, the GG-action has a fundamental domain containing an open subset isomorphic to

{(seiθ,r)𝔻×:a<θ<b,c<r<d},\{(se^{i\theta},r)\in\mathbb{D}\times\mathbb{R}:a<\theta<b,\ c<r<d\},

where 𝔻2\mathbb{D}\cong\mathbb{H}^{2} is the Poincaré disc model of the hyperbolic plane, and again (2×)/G(\mathbb{H}^{2}\times\mathbb{R})/G doesn’t have finite volume. ∎

13. SL2~\widetilde{SL_{2}} Geometry

13.1. Riemannian Geometry of PSL2()PSL_{2}(\mathbb{R})

Riemannian structure of Recall that given a Riemannian manifolds (M,g)(M,g), there is a natural construction of a Riemannian metric tensor on the tangent bundle TMTM constructed as follows: if (p,x)TM(p,x)\in TM, and (c(t),v(t))TM(c(t),v(t))\in TM is a smooth curve such that c(0)=pc(0)=p and v(0)=xv(0)=x, then

(c(0),v(0))(p,x)2=dπ(p,x)((c(0),v(0)))p2+Ddt|t=0v(t)p2,\|(c^{\prime}(0),v^{\prime}(0))\|_{(p,x)}^{2}=\|d\pi_{(p,x)}((c^{\prime}(0),v^{\prime}(0)))\|_{p}^{2}+\left\|\frac{D}{dt}_{|t=0}v(t)\right\|^{2}_{p},

where π:TMM\pi:TM\rightarrow M is the projection, Ddtv(t)\frac{D}{dt}v(t) is the covariant derivative along the curve c(t)c(t) and g(u,u)p=up2g(u,u)_{p}=\|u\|_{p}^{2}. If X=c(0)X=c^{\prime}(0) and Z=v(0)Z=v^{\prime}(0), in local coordinates we have the formula

(X,Z)(p,x)2=Xp2+Z+XjviΓijkkp2.\|(X,Z)\|_{(p,x)}^{2}=\|X\|_{p}^{2}+\left\|Z+X^{j}v^{i}\Gamma_{ij}^{k}\partial_{k}\right\|^{2}_{p}.

The vector (X,Z)(X,Z) is called horizontal if c(t)c(t) is constant, and thus X=0X=0, it is called vertical if it is orthogonal to every horizontal vector in which case Z=XjviΓijkkZ=-X^{j}v^{i}\Gamma_{ij}^{k}\partial_{k}. So, we have a decomposition in horizontal and vertical components as

(X,Z)=(0,Z+XjviΓijkk)+(X,XjviΓijkk).(X,Z)=(0,Z+X^{j}v^{i}\Gamma_{ij}^{k}\partial_{k})+(X,-X^{j}v^{i}\Gamma_{ij}^{k}\partial_{k}).

If we take the global coordinates (x,y)x+iy(x,y)\mapsto x+iy of the hyperbolic plane

2={z:Im(z)>0},\mathbb{H}^{2}=\{z\in\mathbb{C}:Im(z)>0\},

with corresponding metric tensor ds2=dx2+dy2y2ds^{2}=\frac{dx^{2}+dy^{2}}{y^{2}}, then the Christoffel symbols at a point x+iyx+iy are given by

Γ112=Γ222=Γ121=Γ211=1/y.-\Gamma_{11}^{2}=\Gamma_{22}^{2}=\Gamma_{12}^{1}=\Gamma_{21}^{1}=-1/y.

There is a natural identification of the tangent bundle

2×T2,(z,w)ddt|t=0(z+tw)\mathbb{H}^{2}\times\mathbb{C}\cong\textrm{T}\mathbb{H}^{2},\qquad(z,w)\mapsto\frac{d}{dt}_{|t=0}(z+tw)

and so the projection π:T22\pi:T\mathbb{H}^{2}\rightarrow\mathbb{H}^{2} is just given by the projection in the first factor and we have global coordinates in each tangent plane 1=1\partial_{1}=1 and 2=i\partial_{2}=i. If as before, (X,Z)(X,Z) is a tangent vector to T2T\mathbb{H}^{2} at the point (p,v)=(i,1)(p,v)=(i,1), then the orthogonal decomposition in horizontal and vertical components is given by

(X,Z)=(0,ZX2+X1i)+(X,X2X1i).(X,Z)=(0,Z-X^{2}+X^{1}i)+(X,X^{2}-X^{1}i).

The isometric action by Möbius transformations of SL2()SL_{2}(\mathbb{R}) in 2\mathbb{H}^{2}, induces the action in the tangent bundle

SL2()×T2T2,(abcd)(z,w)=(az+bcz+d,w(cz+d)2).SL_{2}(\mathbb{R})\times T\mathbb{H}^{2}\rightarrow T\mathbb{H}^{2},\quad\left(\begin{array}[]{cc}a&b\\ c&d\end{array}\right)\cdot(z,w)=\left(\frac{az+b}{cz+d},\frac{w}{(cz+d)^{2}}\right).

This action is transitive in the unitary tangent bundle T12={(z,w)2:wz=1}T^{1}\mathbb{H}^{2}=\{(z,w)\in\mathbb{H}^{2}:\|w\|_{z}=1\}, so the orbit of the point (i,1)T12(i,1)\in T^{1}\mathbb{H}^{2} induces the diffeomorphism ϕ:PSL2()T12\phi:PSL_{2}(\mathbb{R})\rightarrow T^{1}\mathbb{H}^{2} given explicitly by the formula

ϕ(abcd)=(ai+bci+d,1(ci+d)2).\phi\left(\begin{array}[]{cc}a&b\\ c&d\end{array}\right)=\left(\frac{ai+b}{ci+d},\frac{1}{(ci+d)^{2}}\right).

As this action is also isometric with respect to the previously defined metric, it will define a left invariant metric in PSL2()PSL_{2}(\mathbb{R}) that corresponds to an inner product in its tangent vector to the identity, naturally identified with the Lie algebra

𝔰𝔩2()={AM2():tr(A)=0}.\mathfrak{sl}_{2}(\mathbb{R})=\{A\in M_{2}(\mathbb{R}):\textrm{tr}(A)=0\}.

More precisely, if we consider the derivative dϕd\phi, we get the identification

Ψ:𝔰𝔩2()T(i,1)(T2),Ψ(X)=ddt|t=0ϕ(exp(tX)).\Psi:\mathfrak{sl}_{2}(\mathbb{R})\rightarrow T_{(i,1)}(T\mathbb{H}^{2}),\qquad\Psi(X)=\frac{d}{dt}_{|t=0}\phi(exp(tX)).

A basis of 𝔰𝔩2()\mathfrak{sl}_{2}(\mathbb{R}) is given by

X1=(1001),X2=(0110),X3=(0110).X_{1}=\left(\begin{array}[]{cc}1&0\\ 0&-1\end{array}\right),\quad X_{2}=\left(\begin{array}[]{cc}0&1\\ -1&0\end{array}\right),\quad X_{3}=\left(\begin{array}[]{cc}0&1\\ 1&0\end{array}\right).

If gt,j=exp(tXj)g_{t,j}=exp(tX_{j}), then ϕ(gt,1)=(e2ti,e2t)\phi(g_{t,1})=(e^{2t}i,e^{2t}), ϕ(gt,2)=(i,e2it)\phi(g_{t,2})=(i,e^{2it}) and

ϕ(gt,3)=(ch(t)i+sh(t)ch(t)+ish(t),1(ch(t)+ish(t))2),\phi(g_{t,3})=\left(\frac{ch(t)i+sh(t)}{ch(t)+ish(t)},\frac{1}{(ch(t)+ish(t))^{2}}\right),

where ch(t)ch(t) and sh(t)sh(t) denote the hyperbolic cosine and the hyperbolic sine correspondingly. If X^j=Ψ(Xj)\widehat{X}_{j}=\Psi(X_{j}), we have

X^1=(2i,2),X^2=(0,2i),X^3=(2,2i),\widehat{X}_{1}=(2i,2),\quad\widehat{X}_{2}=(0,2i),\quad\widehat{X}_{3}=(2,-2i),

where we immediatly see that X^2\widehat{X}_{2} is vertical and a direct computation tells us that X^1\widehat{X}_{1} and X^3\widehat{X}_{3} are horizontal and orthogonal. Thus {12X1,12X2,12X3}\{\frac{1}{2}X_{1},\frac{1}{2}X_{2},\frac{1}{2}X_{3}\} is an orthonormal basis in the corresonding inner product in 𝔰𝔩2()\mathfrak{sl}_{2}(\mathbb{R}).

As the PSL2()PSL_{2}(\mathbb{R})-action is given by holomorphic maps, it commutes with the action of S1S^{1} given by rotations in each tangent plane

S1×T12T12,η(z,w)=(z,ηw),S^{1}\times T^{1}\mathbb{H}^{2}\rightarrow T^{1}\mathbb{H}^{2},\qquad\eta\cdot(z,w)=(z,\eta w),

as well as with the map (z,w)(z¯,w¯)(z,w)\mapsto(\overline{z},\overline{w}). It is immediate that the previous maps act by isometries and in fact generate the whole isometry group. Thus, the isometry group Iso(PSL2())Iso(PSL_{2}(\mathbb{R})) is isomorphic to PSL2()×(S12)PSL_{2}(\mathbb{R})\times(S^{1}\rtimes\mathbb{Z}_{2}), see [Sco83].

13.2. Groups of isometries of finite volume

Theorem 13.1.

If ΓIso(PSL2())\Gamma\subset Iso(PSL_{2}(\mathbb{R})) is a discrete group such that PSL2()/ΓPSL_{2}(\mathbb{R})/\Gamma has finite volume, then

Iso(PSL2()/Γ)S1F,Iso(PSL_{2}(\mathbb{R})/\Gamma)\cong S^{1}\rtimes F,

where FF is a finite group.

Proof.

Consider the projection into the simple factor

P:Iso(PSL2())PSL2(),P:Iso(PSL_{2}(\mathbb{R}))\rightarrow PSL_{2}(\mathbb{R}),

as the Kernel of PP is compact and Γ\Gamma is a discrete subgroup, then Γ0=P(Γ)\Gamma_{0}=P(\Gamma) is a discrete subgroup of PSL2()PSL_{2}(\mathbb{R}) and Γ0Γ/F0\Gamma_{0}\cong\Gamma/F_{0}, with F0=Ker(P)ΓF_{0}=Ker(P)\cap\Gamma a finite subgroup of S12S^{1}\rtimes\mathbb{Z}_{2}. Observe that π:PSL2()2\pi:PSL_{2}(\mathbb{R})\rightarrow\mathbb{H}^{2} is a fiber bundle with fiber S1S^{1} such that π(γx)=P(γ)π(x)\pi(\gamma x)=P(\gamma)\pi(x), so we have an induced projection

π:PSL2()/Γ2/Γ0,\pi:PSL_{2}(\mathbb{R})/\Gamma\rightarrow\mathbb{H}^{2}/\Gamma_{0},

which implies that 2/Γ0\mathbb{H}^{2}/\Gamma_{0} has finite hyperbolic area. As we also have the identification π:PSL2()/Γ02/Γ0\pi:PSL_{2}(\mathbb{R})/\Gamma_{0}\rightarrow\mathbb{H}^{2}/\Gamma_{0}, we have that Γ0\Gamma_{0} is a Lattice in PSL2()PSL_{2}(\mathbb{R}). By Lemma 10.1, we have that Γ0\Gamma_{0} has finite index in Λ=NPSL2()(Γ0)\Lambda=N_{PSL_{2}(\mathbb{R})}(\Gamma_{0}). Observe that if F1=Λ/Γ0F_{1}=\Lambda/\Gamma_{0}, then we have that ΓΛ×S12\Gamma\subset\Lambda\times S^{1}\rtimes\mathbb{Z}_{2} and a bijection of sets

(Λ×S12)/Γ(Λ×S12)/F0Γ/F0(Λ×S12)/Γ0FS1,(\Lambda\times S^{1}\rtimes\mathbb{Z}_{2})/\Gamma\cong\frac{(\Lambda\times S^{1}\rtimes\mathbb{Z}_{2})/F_{0}}{\Gamma/F_{0}}\hookrightarrow(\Lambda\times S^{1}\rtimes\mathbb{Z}_{2})/\Gamma_{0}\cong F\ltimes S^{1},

where FF is either F1F_{1}, or F1×2F_{1}\times\mathbb{Z}_{2}, deppending on whether Γ\Gamma contains the map (z,w)(z¯,w¯)(z,w)\mapsto(\overline{z},\overline{w}) or not. Thus, we have that

S1NIso(PSL2())(Γ)/Γ(Λ×S12)/ΓFS1,S^{1}\subset N_{Iso(PSL_{2}(\mathbb{R}))}(\Gamma)/\Gamma\subset(\Lambda\times S^{1}\rtimes\mathbb{Z}_{2})/\Gamma\cong F\ltimes S^{1},

and the result follows. ∎

13.3. Non-classification of finite group actions.

Remark 13.2.

As seen in Remark 10.2, we can obtain every finite group as an isometry group of an hyperbolic surface, so that, the finite factor of the isometry group in Theorem 13.1, can be any finite group.

13.4. Isometries of the universal cover SL2~()\widetilde{SL_{2}}(\mathbb{R}).

The Lie group PSL2(R)PSL_{2}(R) is topologically the product S1×2S^{1}\times{\mathbb{R}}^{2}, so that there is a simply connected Lie group denoted by SL2~()\widetilde{SL_{2}}(\mathbb{R}) which is the topological universal cover of PSL2()PSL_{2}(\mathbb{R}) and algebraically it is a non-split central extension by a cyclic group \mathbb{Z}, more precisely, there is an exact sequence

1SL2~()PSL2()1,1\rightarrow\mathbb{Z}\rightarrow\widetilde{SL_{2}}(\mathbb{R})\rightarrow PSL_{2}(\mathbb{R})\rightarrow 1,

where SL2~()\mathbb{Z}\subset\widetilde{SL_{2}}(\mathbb{R}) lies in the center. We can pull-back the metric tensor of PSL2()PSL_{2}(\mathbb{R}), constructed in the previous section, to SL2~()\widetilde{SL_{2}}(\mathbb{R}) to obtain the model of the homogeneous 3-dimensional geometry denoted by SL2SL_{2}.

Remark 13.3.

The isometry group of SL2~()\widetilde{SL_{2}}(\mathbb{R}) can be characterized in three different ways. First, we have the homomorphism

SL~2()×Iso(SL~2())\widetilde{SL}_{2}(\mathbb{R})\times\mathbb{R}\rightarrow Iso(\widetilde{SL}_{2}(\mathbb{R}))

given by left and right multiplications, here SO~(2)\mathbb{R}\cong\widetilde{SO}(2) is the universal cover of the rotation group SO(2)SL2()SO(2)\subset SL_{2}(\mathbb{R}), with Kernel =SL~2()\mathbb{Z}=\mathbb{R}\cap\widetilde{SL}_{2}(\mathbb{R}) being precisely the center of SL2~()\widetilde{SL_{2}}(\mathbb{R}). The group Iso(SL~2())Iso(\widetilde{SL}_{2}(\mathbb{R})) has two connected components and

Iso(SL~2())0(SL~2()×)/Iso(\widetilde{SL}_{2}(\mathbb{R}))_{0}\cong\big{(}\widetilde{SL}_{2}(\mathbb{R})\times\mathbb{R}\big{)}/\mathbb{Z}

is the component of the identity. In fact, we have an epimorphism

Iso(SL~2())Iso(PSL2())PSL2()×(S12),Iso(\widetilde{SL}_{2}(\mathbb{R}))\rightarrow Iso(PSL_{2}(\mathbb{R}))\cong PSL_{2}(\mathbb{R})\times(S^{1}\rtimes\mathbb{Z}_{2}),

with kernel isomorphic to \mathbb{Z}, however, the group Iso(SL~2())Iso(\widetilde{SL}_{2}(\mathbb{R})) is no longer a product group. The left projection of the previous product gives us the second description in terms of a short exact sequence

1Iso(SL~2())0PSL2()1,1\rightarrow\mathbb{R}\rightarrow Iso(\widetilde{SL}_{2}(\mathbb{R}))_{0}\rightarrow PSL_{2}(\mathbb{R})\rightarrow 1,

and if we consider the groups SL~2()\widetilde{SL}_{2}(\mathbb{R}) and \mathbb{R} as closed subgroups of Iso(SL~2()Iso(\widetilde{SL}_{2}(\mathbb{R}), then we have the third description

Iso(SL~2())0=L(SL~2())R(),Iso(\widetilde{SL}_{2}(\mathbb{R}))_{0}=L(\widetilde{SL}_{2}(\mathbb{R}))R(\mathbb{R}),

where L()L(\cdot) and R()R(\cdot) represent left and right multiplications in the group SL~2()\widetilde{SL}_{2}(\mathbb{R}).

A discrete subgroup ΓIso(PSL2())\Gamma\subset Iso(PSL_{2}(\mathbb{R})) can be lifted to a discrete subgroup Γ~Iso(SL~2())\widetilde{\Gamma}\subset Iso(\widetilde{SL}_{2}(\mathbb{R})), so that SL~2()/Γ~PSL2()/Γ\widetilde{SL}_{2}(\mathbb{R})/\widetilde{\Gamma}\cong PSL_{2}(\mathbb{R})/\Gamma and thus, we can compute Iso(SL~2()/Γ~)Iso(\widetilde{SL}_{2}(\mathbb{R})/\widetilde{\Gamma}) with Theorem 13.1, however, not every discrete group of Iso(SL~2())Iso(\widetilde{SL}_{2}(\mathbb{R})) can be obtained this way. In the next section we discuss the proof in the general setting for discrete groups of isometries in SL~2()\widetilde{SL}_{2}(\mathbb{R}).

The following Lemma is well known and holds for every Lie group, but we include a proof of the case we need for the sake of completeness.

Lemma 13.4.

If GG is a Lie group locally isomorphic to ×SL2()\mathbb{R}\times SL_{2}(\mathbb{R}), for example GG can be the isometry group of SL~2()\widetilde{SL}_{2}(\mathbb{R}) or 2×\mathbb{H}^{2}\times\mathbb{R}, then there exists a neighborhood of the identity UGU\subset G such that [U,U]U[U,U]\subset U.

Proof.

Observe first that this is a local property, so we only need to prove this for linear groups. As the \mathbb{R} factor lies in the center, we have that

[gg0,hh0]=[g,h],g0,h0[gg_{0},hh_{0}]=[g,h],\qquad\forall\ g_{0},h_{0}\in\mathbb{R}

and thus we only need to prove this for SL2()SL_{2}(\mathbb{R}). The commutator

[(axyb),(czwd)]=(t1t3t4t2),\left[\left(\begin{array}[]{cc}a&x\\ y&b\end{array}\right),\left(\begin{array}[]{cc}c&z\\ w&d\end{array}\right)\right]=\left(\begin{array}[]{cc}t_{1}&t_{3}\\ t_{4}&t_{2}\end{array}\right),

is defined by the relations

  • t1=1+xy+zw+xyzw+wxac+w2x2adxwa2zw+bxwdyd2xazydt_{1}=1+xy+zw+xyzw+wxac+w^{2}x^{2}-adxw-a^{2}zw+bxwd-yd^{2}x-azyd,

  • t2=1+xy+zw+xyzwxwbcc2xy+ayzczb2wzbcy+zybd+z2y2t_{2}=1+xy+zw+xyzw-xwbc-c^{2}xy+ayzc-zb^{2}w-zbcy+zybd+z^{2}y^{2},

  • t3=xac(dc)cx2w+acz(ab)xwbz+zydx+z2yat_{3}=xac(d-c)-cx^{2}w+acz(a-b)-xwbz+zydx+z^{2}ya,

  • t4=w2xb+wxcy+bdw(ba)awyz+bdy(cd)dy2zt_{4}=w^{2}xb+wxcy+bdw(b-a)-awyz+bdy(c-d)-dy^{2}z.

So that if 0|x|,|y|,|z|,|w|<ε0\leq|x|,|y|,|z|,|w|<\varepsilon and 1ε<a,b,c,d<1+ε1-\varepsilon<a,b,c,d<1+\varepsilon, then there is a constant C>0C>0 independent of ε\varepsilon such that |t3|,|t4|<Cε2|t_{3}|,|t_{4}|<C\varepsilon^{2} and |t11|,|t21|<Cε2|t_{1}-1|,|t_{2}-1|<C\varepsilon^{2}. Thus, by choosing ε>0\varepsilon>0 such that Cε2<εC\varepsilon^{2}<\varepsilon, the neighborhood

Uε={(axyb):|x|,|y|<ε,|a1|,|b1|<ε}U_{\varepsilon}=\left\{\left(\begin{array}[]{cc}a&x\\ y&b\end{array}\right):|x|,|y|<\varepsilon,\ |a-1|,|b-1|<\varepsilon\right\}

is stable under taking commutators. ∎

13.5. Isometry groups of finite volume

Proposition 13.5.

Let HH be a Lie group which is a central extension of PSL2()PSL_{2}(\mathbb{R}) of the form

1HPSL2()1.1\rightarrow\mathbb{R}\rightarrow H\rightarrow PSL_{2}(\mathbb{R})\rightarrow 1.

If GHG\subset H is a discrete subgroup with induced exact sequence

1KGΓ1,1\rightarrow K\rightarrow G\rightarrow\Gamma\rightarrow 1,

with KK\subset\mathbb{R}, then either ΓPSL2()\Gamma\subset PSL_{2}(\mathbb{R}) is discrete or is an abelian subgroup leaving fixed a point, a geodesic or a horocycle in 2\mathbb{H}^{2}.

Proof.

Denote by p:HPSL2()p:H\rightarrow PSL_{2}(\mathbb{R}) the projection and consider UHU\subset H a neighborhood of the identity such that [U,U]U[U,U]\subset U and UG={e}U\cap G=\{e\}. We have that the group L=p(U)ΓL=\langle p(U)\cap\Gamma\rangle is a commutative subgroup of PSL2()PSL_{2}(\mathbb{R}), to see why this is true take two elements α,βG\alpha,\beta\in G such that p(α),p(β)p(U)p(\alpha),p(\beta)\in p(U), then we may write those elements as α=α0α1\alpha=\alpha_{0}\alpha_{1} and β=β0β1\beta=\beta_{0}\beta_{1}, where

α1,β1,α0,β0U.\alpha_{1},\beta_{1}\in\mathbb{R},\qquad\alpha_{0},\beta_{0}\in U.

As \mathbb{R} lies in the center of HH we have that [α0,β0]=[α,β]GU[\alpha_{0},\beta_{0}]=[\alpha,\beta]\in G\cap U and thus α\alpha and β\beta commute. Now, for every αG\alpha\in G, choose a neighborhood of the identity UαGU_{\alpha}\subset G such that [α,Uα]U[\alpha,U_{\alpha}]\subset U, so that the elements of Γp(Uα)\Gamma\cap p(U_{\alpha}) commute with p(α)p(\alpha) (same argument as with the commutativity of LL). Suppose that Γ\Gamma is non-discrete, then LL is a non-trivial commutative subgroup and for every γ=p(α)Γ\gamma=p(\alpha)\in\Gamma, we have that Γp(Uα)\Gamma\cap p(U_{\alpha}) is a non-trivial subset that generates the group LL and commutes with γ\gamma. So, Γ\Gamma commutes with LL and thus, there exists an element X𝔰𝔩2()X\in\mathfrak{sl}_{2}(\mathbb{R}) such that

ΓL¯={exp(tX):t}\Gamma\subset\overline{L}=\{exp(tX):t\in\mathbb{R}\}

and Γ\Gamma leaves fixed a point, a geodesic or a horocycle, depending on the type of XX. ∎

Theorem 13.6.

If GG is a discrete subgroup of isometries of XX either SL~2()\widetilde{SL}_{2}(\mathbb{R}) or 2×\mathbb{H}^{2}\times\mathbb{R} such that X/GX/G has finite volume, then the isometry group Iso(X/G)Iso(X/G) is a finite extension of S1S^{1}.

Proof.

The exact sequence

1Iso(X)PSL2()11\rightarrow\mathbb{R}\rightarrow Iso(X)\rightarrow PSL_{2}(\mathbb{R})\rightarrow 1

induces the sequence

1KGΓ1,1\rightarrow K\rightarrow G\rightarrow\Gamma\rightarrow 1,

if Γ\Gamma is non-discrete, then it preserves a geodesic, a point or a horocycle by Proposition 13.5 and we can see that X/GX/G doesn’t have finite volume (as we did in Theorem 12.1). So, Γ\Gamma is a discrete subgroup of isometries of the hyperbolic plane and we have a fiber bundle structure

X2\mathbb{R}\rightarrow X\rightarrow\mathbb{H}^{2}

so that the volume form decomposes as

2f𝑑μ𝑑t=Xf𝑑volX\int_{\mathbb{R}}\int_{\mathbb{H}^{2}}fd\mu dt=\int_{X}fdvol_{X}

where dμd\mu is the hyperbolic area form. If D2D\subset\mathbb{H}^{2} is a fundamental domain of Γ\Gamma, then π1(D)=D^\pi^{-1}(D)=\widehat{D} is such that gD^D^g\widehat{D}\cap\widehat{D}\neq\emptyset only for gK=Gg\in K=G\cap\mathbb{R}. Thus for Ω\Omega\subset\mathbb{R} fundamental domain of KK in \mathbb{R} we have that Ω×D\Omega\times D is a fundamental domain for GG which implies that

Vol(X/G)ΩDξ=|Ω|×μ(D),Vol(X/G)\geq\int_{\Omega}\int_{D}\xi=|\Omega|\times\mu(D),

and we have that μ(A)<\mu(A)<\infty and |Ω|<|\Omega|<\infty which implies that K=K=\mathbb{Z}. Take N~=NIso(X)(G)\widetilde{N}=N_{Iso(X)}(G) and N=NPSL2()(Γ)N=N_{PSL_{2}(\mathbb{R})}(\Gamma), so that we have the exact sequence

1N0N~N11\rightarrow N_{0}\rightarrow\widetilde{N}\rightarrow N\rightarrow 1

(because gGg1=GgGg^{-1}=G projects to g¯Γg¯1=Γ\overline{g}\Gamma\overline{g}^{-1}=\Gamma and N0=N_{0}=\mathbb{R} because \mathbb{R} normalizes SL2~()\widetilde{SL_{2}}(\mathbb{R})), this sequence induces the exact sequence

1N0/N~/GN/Γ11\rightarrow N_{0}/\mathbb{Z}\rightarrow\widetilde{N}/G\rightarrow N/\Gamma\rightarrow 1

(to see that this sequence is exact observe that π(gG)=g¯Γ\pi(gG)=\overline{g}\Gamma is well defined and surjective, the condition π(gG)=Γ\pi(gG)=\Gamma holds if and only if g¯Γ\overline{g}\in\Gamma and thus g=[A,r]g=[A,r] with AΓA\in\Gamma, this is because there is an element h=[A,s]Gh=[A,s]\in G, thus gG=gh1GgG=gh^{-1}G and gh1i(N0/)gh^{-1}\in i(N_{0}/\mathbb{Z}). This implies that the kernel of π\pi is i(N0/)i(N_{0}/\mathbb{Z}). Finally i(r)=Gi(r\mathbb{Z})=G if and only if rGr\in G, but G=G\cap\mathbb{R}=\mathbb{Z}, so that r=r\mathbb{Z}=\mathbb{Z} and thus ii is injective). This exact sequence can be written as

1S1Iso(X/G)Iso(2/Γ)11\rightarrow S^{1}\rightarrow Iso(X/G)\rightarrow Iso(\mathbb{H}^{2}/\Gamma)\rightarrow 1

which implies the result because of the Lemma 10.1. ∎

14. Corollaries of the Main Theorem

14.1. Actions of SLk()SL_{k}(\mathbb{Z}) on aspherical three dimensional manifolds by isometries.

We have the following affirmative solution to Problem 2.1.

Theorem 14.1.

Any group action by isometries of SLk+1()SL_{k+1}(\mathbb{Z}), with k3k\geq 3, on a closed, aspherical 33-manifold factors trough a finite group.

14.2. Discrete groups acting with a sufficiently collapsed Alexandrov space as quotient

Theorem 14.2.

Assume that a discrete group Γ\Gamma acts by isometries on the three dimensional Alexandrov space XX such that the quotient X/ΓX/\Gamma is sufficiently collapsed with parameters dd, and ϵ\epsilon. Then, Theorem 2.8, together with the geometrization of 33-dimensional Alexandrov spaces provide a classification of the possible such Γ\Gamma within the lattices in the isometry groups.

14.3. Hilbert Smith Conjecture for three dimensional Alexandrov spaces

Let us recall Theorem 3.10.

Theorem 14.3 ([Par19], [Par13]).

For every prime pp, there are no faithful actions by homeomorphisms of the pp-adic group ^p\widehat{\mathbb{Z}}_{p} on a topological manifold of dimension n3n\leq 3.

As seen in section3.2, an Alexandrov space XX has a closed subset SXS_{X}, corresponding to topologically singular points and such that the set of regular points RX=XSXR_{X}=X\setminus S_{X} is an open-dense subset, having the structure of a topological manifold. An action by homeomorphisms on XX must preserve the decomposition X=SXRXX=S_{X}\cup R_{X} and a continuous action of ^p\widehat{\mathbb{Z}}_{p} which is trivial on the regular points, is trivial on the whole space XX.

Hence, the weaker version of the pp-adic Hilbert-Smith conjecture for Alexandrov spaces holds.

A consequence of Theorem 3.10, gives us

Theorem 14.4.

If GG is a locally compact, topological group, acting faithfully on a three dimensional Alexandrov space by homeomorphisms, then GG is a Lie group.

Remark 14.5.

As observed in previous section, there is a subset of metrically regular points which admits a compatible Riemannian metric, constructed in [OS94]. Thus, we have as a consequence of Theorem 3.9, that the pp-adic group ^p\widehat{\mathbb{Z}}_{p} cannot act faithfully by bi-Lipschitz homeomorphisms. However, we should be careful, as the set of metrically singular points can be dense, as seen in an example constructed in [OS94] as a limit of Alexandrov spaces, using baricentric subdivisions of a tetrahedron.

14.4. Non Existence of actions of Higher Rank Lattices by Isometries on three dimensional Geometric Orbifolds

Theorem 14.6.

Let Γ\Gamma be a higher rank lattice acting by isometries on a finite volume, three dimensional orbifold XX (modelled over a homogeneous 3-manifold XX), then the action factors through a finite group if either:

  • XX is aspherical or,

  • Γ\Gamma is non-uniform.

As an example of this, we have Γ=SLr()\Gamma=SL_{r}(\mathbb{Z}) with r3r\geq 3.

14.5. Characterization of Higher Rank Lattices actions by isometries on three dimensional spherical Orbifolds

Remark 14.7.

As a consequence of the previous discussion, for every semisimple Lie group GG such that G×SO(4)G\times SO(4) is isotypic111For example, any product G=G1××GkG=G_{1}\times\cdots\times G_{k}, where each GjG_{j} is one of SO(3,1)SO(3,1), SO(2,2)SO(2,2) or SO(4,)SO(4,\mathbb{C})., there is an irreducible lattice ΓG\Gamma\subset G and an homomorphism ΓSO(4)\Gamma\rightarrow SO(4) with dense image. In particular, such lattice acts by isometries on the round sphere S3S^{3} with dense orbits. This tells us that there is no restriction on the dimension of the class of higher rank lattices which can act on the round sphere, but the type of such lattice is restricted. The same applies to the 33-orbifolds of the type S2×S1S^{2}\times S^{1}, as the first factor has isometry group SO(3)SO(3) which is simple.

We have in fact a converse of Remark 14.7, given by the following Theorem:

Theorem 14.8.

If ΓG\Gamma\subset G is a lattice in a higher rank, semisimple Lie group, KK is a compact Lie group and φ:ΓK\varphi:\Gamma\rightarrow K is a homomorphism with dense image, then the group G×KG\times K is isotypic and ΓG\Gamma\subset G is cocompact.

Theorem 14.8 is “well known to the experts”, but a sketch of the first part of its proof is made in [BFH16], section 2.3. The fact that the lattice Γ\Gamma is cocompact is a consequence of Godement’s compactness criterion.

Corollary 14.9.

If ΓG\Gamma\subset G is a lattice in a higher rank, simple Lie group, KK is a compact Lie group and φ:ΓK\varphi:\Gamma\rightarrow K is a homomorphism with infinite image, then G×LG\times L is isotypic, with L=φ(Γ)¯L=\overline{\varphi(\Gamma)}. In particular, dim(G)dim(K)\mathrm{dim}(G)\leq\mathrm{dim}(K) and Γ\Gamma is cocompact in GG.

As an immediate consequence, non-cocompact lattices don’t appear in this setting and we have

Corollary 14.10.

Let XX be a geometric 33-orbifold of finite volume, and Γ\Gamma a non-cocompact higher rank lattice in a semisimple Lie group GG, then any action of Γ\Gamma in XX factors through a finite group.

As a particular example of the previous, any action of SLn()SL_{n}(\mathbb{Z}) in a geometric 33-orbifold of finite volume, factors through a finite group.

Corollary 14.11.

Let XX be a geometric 33-orbifold of finite volume, then XX admits an isometric action of a higher rank lattice ΓG\Gamma\subset G if and only if the group Iso(X)Iso(X) contains the group SO(3)SO(3). Moreover, the semisimple Lie group GG is isotypic of type SO(3)SO(3) and the lattice is uniform.

Observe that the group SO(4)SO(4) factors locally as the product SO(3)×SO(3)SO(3)\times SO(3) and in fact, there is a copy of SO(3)SO(3) inside SO(4)SO(4), so that the previous Corollary includes at the same time examples like X=S3/ΛX=S^{3}/\Lambda and X=(S2×)/ΛX=(S^{2}\times\mathbb{R})/\Lambda.

References

  • [BBI01] Dmitri Burago, Yuri Burago, and Sergei Ivanov. A course in metric geometry, volume 33 of Graduate Studies in Mathematics. American Mathematical Society, Providence, RI, 2001.
  • [BFH16] Aaron Brown, David Fisher, and Sebastian Hurtado. Zimmer’s conjecture: subexponential growth, measure rigidity, and strong property (t). Preprint, 2016.
  • [BFH20] Aaron Brown, David Fisher, and Sebastian Hurtado. Zimmer’s conjecture for actions of SL(m, \mathbb{Z}). Invent. Math., 221(3):1001–1060, 2020.
  • [BFH21] Aaron Brown, David Fisher, and Sebastian Hurtado. Zimmer’s conjecture for non-uniform lattices and escape of mass, 2021.
  • [BK02] Mario Bonk and Bruce Kleiner. Rigidity for quasi-Möbius group actions. J. Differential Geom., 61(1):81–106, 2002.
  • [BK19] Richard Bamler and Bruce Kleiner. Ricci flow and contractibility of spaces of metrics. arXiv:1909.08710, 2019.
  • [BM46] Salomon Bochner and Deane Montgomery. Locally compact groups of differentiable transformations. Ann. of Math. (2), 47:639–653, 1946.
  • [BNnZ21] Noé Bárcenas and Jesús Núñez Zimbrón. On topological rigidity of Alexandrov 3-spaces. Rev. Mat. Iberoam., 37(5):1629–1639, 2021.
  • [Boc46] S. Bochner. Vector fields and Ricci curvature. Bull. Amer. Math. Soc., 52:776–797, 1946.
  • [Bou09] Marc Bourdon. Quasi-conformal geometry and Mostow rigidity. In Géométries à courbure négative ou nulle, groupes discrets et rigidités, volume 18 of Sémin. Congr., pages 201–212. Soc. Math. France, Paris, 2009.
  • [BRW61] G. E. Bredon, Frank Raymond, and R. F. Williams. pp-adic groups of transformations. Trans. Amer. Math. Soc., 99:488–498, 1961.
  • [BZ07] A. V. Bagaev and N. I. Zhukova. The isometry groups of Riemannian orbifolds. Sibirsk. Mat. Zh., 48(4):723–741, 2007.
  • [DH20] Bertrand Deroin and Sebastian Hurtado. Non left-orderability of lattices in higher rank semi-simple lie groups, 2020. arXiv:2008.10687.
  • [DVdW28] D. van Dantzig and B. L. Van der Waerden. Über metrisch homogene Räume. Abh. Math. Semin. Univ. Hamb., 6:367–376, 1928.
  • [Fis11] David Fisher. Groups acting on manifolds: around the Zimmer program. In Geometry, rigidity, and group actions, Chicago Lectures in Math., pages 72–157. Univ. Chicago Press, Chicago, IL, 2011.
  • [Fis20] David Fisher. Recent developments in the Zimmer program. Notices Amer. Math. Soc., 67(4):492–499, 2020.
  • [FM98] Benson Farb and Howard Masur. Superrigidity and mapping class groups. Topology, 37(6):1169–1176, 1998.
  • [FS00] Benson Farb and Peter Shalen. Lattice actions, 3-manifolds and homology. Topology, 39(3):573–587, 2000.
  • [Fur76] Harry Furstenberg. A note on Borel’s density theorem. Proc. Amer. Math. Soc., 55(1):209–212, 1976.
  • [FY94] Kenji Fukaya and Takao Yamaguchi. Isometry groups of singular spaces. Math. Z., 216(1):31–44, 1994.
  • [GG16] Fernando Galaz-García. A glance at three-dimensional Alexandrov spaces. Front. Math. China, 11(5):1189–1206, 2016.
  • [GGG13] Fernando Galaz-Garcia and Luis Guijarro. Isometry groups of Alexandrov spaces. Bull. Lond. Math. Soc., 45(3):567–579, 2013.
  • [GGG15] Fernando Galaz-Garcia and Luis Guijarro. On three-dimensional Alexandrov spaces. Int. Math. Res. Not. IMRN, (14):5560–5576, 2015.
  • [GGGNnZ20] Fernando Galaz-García, Luis Guijarro, and Jesús Núñez Zimbrón. Sufficiently collapsed irreducible alexandrov 3-spaces are geometric. Indiana Univ. Math. J., 69(3):977–1005, 2020.
  • [Gre74] Leon Greenberg. Maximal groups and signatures. In Discontinuous groups and Riemann surfaces (Proc. Conf., Univ. Maryland, College Park, Md., 1973), Ann. of Math. Studies, No. 79, pages 207–226. Princeton Univ. Press, Princeton, N.J., 1974.
  • [Hae20] Thomas Haettel. Hyperbolic rigidity of higher rank lattices. Ann. Sci. Éc. Norm. Supér. (4), 53(2):439–468, 2020. With an appendix by Vincent Guirardel and Camille Horbez.
  • [Hat83] Allen E. Hatcher. A proof of the Smale conjecture, Diff(S3)O(4){\rm Diff}(S^{3})\simeq{\rm O}(4). Ann. of Math. (2), 117(3):553–607, 1983.
  • [Hem04] John Hempel. 3-manifolds. AMS Chelsea Publishing, Providence, RI, 2004. Reprint of the 1976 original.
  • [HJKL02] Ku Yong Ha, Jang Hyun Jo, Seung Won Kim, and Jong Bum Lee. Classification of free actions of finite groups on the 3-torus. Topology Appl., 121(3):469–507, 2002.
  • [HKMR12] Sungbok Hong, John Kalliongis, Darryl McCullough, and J. Hyam Rubinstein. Diffeomorphisms of elliptic 3-manifolds, volume 2055 of Lecture Notes in Mathematics. Springer, Heidelberg, 2012.
  • [HS17] John Harvey and Catherine Searle. Orientation and symmetries of Alexandrov spaces with applications in positive curvature. J. Geom. Anal., 27(2):1636–1666, 2017.
  • [JL10] Jang Hyun Jo and Jong Bum Lee. Group extensions and free actions by finite groups on solvmanifolds. Math. Nachr., 283(7):1054–1059, 2010.
  • [KN96] Shoshichi Kobayashi and Katsumi Nomizu. Foundations of differential geometry. Vol. I. Wiley Classics Library. John Wiley & Sons, Inc., New York, 1996. Reprint of the 1963 original, A Wiley-Interscience Publication.
  • [Koj88] Sadayoshi Kojima. Isometry transformations of hyperbolic 33-manifolds. Topology Appl., 29(3):297–307, 1988.
  • [KOS17] Daehwan Koo, Myungsung Oh, and Joonkook Shin. Classification of free actions of finite groups on 3-dimensional nilmanifolds. J. Korean Math. Soc., 54(5):1411–1440, 2017.
  • [Lee97] Joo Sung Lee. Totally disconnected groups, pp-adic groups and the Hilbert-Smith conjecture. Commun. Korean Math. Soc., 12(3):691–699, 1997.
  • [LSY93] Kyung Bai Lee, Joon Kook Shin, and Shoji Yokura. Free actions of finite abelian groups on the 33-torus. Topology Appl., 53(2):153–175, 1993.
  • [McC02] Darryl McCullough. Isometries of elliptic 3-manifolds. J. London Math. Soc. (2), 65(1):167–182, 2002.
  • [Mor15] Dave Witte Morris. Introduction to arithmetic groups. Deductive Press, [place of publication not identified], 2015.
  • [MS39] S. B. Myers and N. E. Steenrod. The group of isometries of a Riemannian manifold. Ann. of Math. (2), 40(2):400–416, 1939.
  • [MS19] Mattia Mecchia and Andrea Seppi. Isometry groups and mapping class groups of spherical 3-orbifolds. Math. Z., 292(3-4):1291–1314, 2019.
  • [MZ74] Deane Montgomery and Leo Zippin. Topological transformation groups. Robert E. Krieger Publishing Co., Huntington, N.Y.,,, 1974. Reprint of the 1955 original.
  • [Now34] Werner Nowacki. Die euklidischen, dreidimensionalen, geschlossenen und offenen Raumformen. Comment. Math. Helv., 7(1):81–93, 1934.
  • [OS94] Yukio Otsu and Takashi Shioya. The Riemannian structure of Alexandrov spaces. J. Differential Geom., 39(3):629–658, 1994.
  • [Par13] John Pardon. The Hilbert-Smith conjecture for three-manifolds. J. Amer. Math. Soc., 26(3):879–899, 2013.
  • [Par19] John Pardon. Totally disconnected groups (not) acting on two-manifolds. In Breadth in contemporary topology, volume 102 of Proc. Sympos. Pure Math., pages 187–193. Amer. Math. Soc., Providence, RI, 2019.
  • [Rag07] M. S. Raghunathan. Discrete subgroups of Lie groups. Math. Student, (Special Centenary Volume):59–70 (2008), 2007.
  • [Rat99] John G. Ratcliffe. On the isometry groups of hyperbolic orbifolds. Geom. Dedicata, 78(1):63–67, 1999.
  • [Rat19] John G. Ratcliffe. Foundations of hyperbolic manifolds, volume 149 of Graduate Texts in Mathematics. Springer, Cham, 2019. Third edition [of 1299730].
  • [RS97] Dusan Repovs and Evgenij Scepin. A proof of the hilbert-smith conjecture for actions by lipschitz maps. Math. Ann., 308(2):361–364, 1997.
  • [RT15] John G. Ratcliffe and Steven T. Tschantz. On the isometry group of a compact flat orbifold. Geom. Dedicata, 177:43–60, 2015.
  • [Sco83] Peter Scott. The geometries of 33-manifolds. Bull. London Math. Soc., 15(5):401–487, 1983.
  • [Tao14] Terence Tao. Hilbert’s fifth problem and related topics, volume 153 of Graduate Studies in Mathematics. American Mathematical Society, Providence, RI, 2014.
  • [Thu97] William P. Thurston. Three-dimensional geometry and topology. Vol. 1, volume 35 of Princeton Mathematical Series. Princeton University Press, Princeton, NJ, 1997. Edited by Silvio Levy.
  • [Tol74] Jeffrey L. Tollefson. The compact 33-manifolds covered by S2×R1S^{2}\times R^{1}. Proc. Amer. Math. Soc., 45:461–462, 1974.
  • [UY17] Masaaki Umehara and Kotaro Yamada. Differential geometry of curves and surfaces. World Scientific Publishing Co. Pte. Ltd., Hackensack, NJ, 2017. Translated from the second (2015) Japanese edition by Wayne Rossman.
  • [vN33] J. von Neumann. Die Einführung analytischer Parameter in topologischen Gruppen. Ann. of Math. (2), 34(1):170–190, 1933.
  • [Wal68] Friedhelm Waldhausen. On irreducible 33-manifolds which are sufficiently large. Ann. of Math. (2), 87:56–88, 1968.
  • [Wei11] Shmuel Weinberger. Some remarks inspired by the c0c^{0} zimmer program. In Geometry, rigidity, and group actions, Chicago Lectures in Math., pages 262–282. Univ. Chicago Press, Chicago, IL, 2011.
  • [Yan60] Chung-Tao Yang. pp-adic transformation groups. Michigan Math. J., 7:201–218, 1960.
  • [Ye19] Shengkui Ye. Symmetries of flat manifolds, jordan property and the general zimmer program. J. Lond. Math. Soc. (2), 100(3):1065–1080, 2019.
  • [Ye20] Shengkui Ye. A survey of topological’s zimmer programm. arXiv: arXiv:2002.01206, 2020.
  • [Zim84] Robert J. Zimmer. Ergodic theory and semisimple groups, volume 81 of Monographs in Mathematics. Birkhäuser Verlag, Basel, 1984.
  • [Zim87] Robert J. Zimmer. Actions of semisimple groups and discrete subgroups. In Proceedings of the International Congress of Mathematicians, Vol. 1, 2 (Berkeley, Calif., 1986), pages 1247–1258. Amer. Math. Soc., Providence, RI, 1987.