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

Four-dimensional compact Clifford–Klein forms of pseudo-Riemannian symmetric spaces with signature (2, 2)

Keiichi Maeta
Abstract

We give a classification of irreducible four-dimensional symmetric spaces which admit compact Clifford–Klein forms. For this, we develop a method that applies to particular 1-connected solvable symmetric spaces.

We also examine a ‘solvable analogue’ of Kobayashi’s conjecture for reductive groups and find an evidence that the reductive assumption in Kobayashi’s conjecture is crucial.

1 Introduction

We are interested in the classification of indecomposable pseudo-Riemannian symmetric spaces which admit compact Clifford–Klein forms. In this paper, we classify the spaces whose dimensions are up to four and transvection groups are solvable. In the following, we review a background of our problem from two different viewpoints.

First, we review a classification of pseudo-Riemannian symmetric spaces. After É. Cartan ([9, 10]) classified Riemannian symmetric spaces, Berger ([4]) gave a classification theory of irreducible symmetric spaces. However, since pseudo-Riemannian symmetric spaces have a degenerate subspace in their tangent spaces, they are not necessarily decomposed into irreducible symmetric spaces. A ‘minimum unit’ of pseudo-Riemannian symmetric space is said to be indecomposable. Therefore, one may expect to classify indecomposable symmetric spaces. Indecomposable symmetric spaces with signature (n,1)(n,1) and (n,2)(n,2) are classified by Cahen, Wallach, Parker, Kath and Olbrich ([8, 7, 15]).

Second, we review the existence problem of compact Clifford–Klein forms. For a Lie group GG, its closed subgroup HH assume a discrete subgroup ΓG\Gamma\subset G acts on G/HG/H properly discontinuously. We say the quotient space Γ\G/H\Gamma\backslash G/H is a Clifford–Klein form (see Definition 2.21). In the late 1980s, a systematic study of Clifford–Klein forms for non-Riemannian homogeneous spaces was initiated by T. Kobayashi ([17]). The following problem is one of the central problems in this field, but the final answer remains open.

Problem 1.1 ([17]).

Classify homogeneous spaces G/HG/H which admit compact Clifford–Klein forms.

Any Riemannian spaces of reductive type admit compact Clifford–Klein forms ([5]). However, this problem for non-Riemannian spaces is open. By the classification of Berger ([4]), irreducible symmetric spaces are of reductive type, and most of the works on this problem have focused on classifying symmetric spaces of reductive type which admits compact Clifford–Klein forms (see [3, 19, 21, 25, 26, 28, 29] and so on). Five series and seven sporadic irreducible symmetric spaces have been found to admit compact Clifford–Klein forms so far ([26]).

In view of the above two prospects, we are interested in the following subproblem of Problem 1.1.

Problem 1.2 ([16, §1]).

Classify reducible and indecomposable pseudo-Riemannian symmetric spaces G/HG/H which admit compact Clifford–Klein forms.

Since different groups act on a symmetric space transitively, symmetric spaces can be written as different forms. Therefore we can consider Problem 1.2 for each group. For example, the following two groups act on a symmetric space transitively.

Definition 1.3 (isometry group, transvection group, [7]).

The isometry group GiG_{\text{i}} of a pseudo-Riemannian symmetric space is the group which consists of all transformations which preserve the pseudo-Riemannian metric on the space. The transvection group GtG_{\text{t}} of a pseudo-Riemmanian symmetric space is the closed and connected subgroup of the isometry group generated by the products of two geodesic symmetries.

The transvection group GtG_{\text{t}} is a normal subgroup of the isometry group GiG_{\text{i}}. Roughly speaking, the transvection group is the “smallest” group which acts on the symmetric space transitively. For example,

  • for Euclid space n{\mathbb{R}}^{n}, we have Gi=O(n)nG_{\text{i}}=O(n)\ltimes{\mathbb{R}}^{n} and Gt=nG_{\text{t}}={\mathbb{R}}^{n},

  • for the sphere SnS^{n}, we have Gi=O(n+1)G_{\text{i}}=O(n+1) and Gt=SO(n+1)G_{\text{t}}=SO(n+1).

Problem 1.2 was studied by Kath–Olbrich ([16]). They found a necessary and sufficient condition for the existence of compact Clifford–Klein forms in the case of indecomposable Lorentzian symmetric spaces. In this paper, we are interested in Problem 1.2 for symmetric spaces with signature (n,2)(n,2).

For the first step of the problem, we classify four-dimensional indecomposable symmetric spaces MM which admit compact Clifford–Klein forms whose transvection group is solvable. Since Kath–Olbrich ([16]) classified Lorentzian spaces which admit compact Clifford–Klein forms, we discuss the space with signature (2,2)(2,2).

The indecomposable but reducible pseudo-Riemannian symmetric spaces with signature (2,2) are classified by Kath–Olbrich ([15]). Since a 1-connected pseudo-Riemannian symmetric space (M,σ,g)(M,\sigma,g) uniquely corresponds to a symmetric triple (𝔤,σ,g)(\mathfrak{g},\sigma,g), they classified its symmetric triples.

Fact 1.4 ([15, Theorem 7.1]).

Let (𝔤,σ,g)(\mathfrak{g},\sigma,g) be the symmetric triple corresponding to a 1-connected four-dimensional reducible and indecomposable pseudo-Riemannian symmetric spaces with signature (2,2)(2,2), and assume that its transvection group is solvable. Then the symmetric triple (𝔤,σ,g)(\mathfrak{g},\sigma,g) is isometric to one in the following list.

  1. Case (I)

    Nilpotent symmetric triples (𝔤nil,σ,g±)(\mathfrak{g}_{\text{nil}},\sigma,g_{\pm}) (see Definition 4.15),

  2. Case (II)

    Solvable symmetric triples (𝔤D,D,σ,g)(\mathfrak{g}_{D,D^{\prime}},\sigma,g) (see Definition 4.4), where

    1. (a)

      (D,D)=(±diag(1,ν),diag(1,ν))(ν>0)(D,D^{\prime})=(\pm\operatorname{diag}\left(1,\nu\right),\operatorname{diag}\left(1,-\nu\right))\quad(\nu>0),
      (D,D)=(±diag(1,ν),diag(1,ν))(ν>0,ν1)(D,D^{\prime})=(\pm\operatorname{diag}\left(1,-\nu\right),\operatorname{diag}\left(1,-\nu\right))\quad(\nu>0,\ \nu\neq 1),

    2. (b)

      (D,D)=(Qν,Qν)(ν>0)(D,D^{\prime})=\left(Q_{\nu},Q_{-\nu}\right)\quad(\nu>0) (see Notation 2.1),

    3. (c)

      (D,D)=((±1110),(011±1)),((±1110),(0111))(D,D^{\prime})=\left(\begin{pmatrix}\pm 1&-1\\ -1&0\end{pmatrix},\begin{pmatrix}0&-1\\ -1&\pm 1\end{pmatrix}\right),\ \left(\begin{pmatrix}\pm 1&-1\\ -1&0\end{pmatrix},\begin{pmatrix}0&1\\ 1&\mp 1\end{pmatrix}\right),

    4. (d)

      (D,D)=(±I1,1,I1,1)(D,D^{\prime})=(\pm I_{1,1},I_{1,1}).

Here, 𝔤nil\mathfrak{g}_{\rm nil} is a 3-step nilpotent Lie algebra given in Definition 4.15. On the other hand, 𝔤D,D\mathfrak{g}_{D,D^{\prime}} denotes an extension of the Heisenberg Lie algebra 𝔥2\mathfrak{h}_{2} by (DD)𝔰𝔭(2,)\begin{pmatrix}&D^{\prime}\\ D&\end{pmatrix}\in\mathfrak{s}\mathfrak{p}(2,{\mathbb{R}}) (see Definition 4.4). The isometry class of the triple (𝔤D,D,σ,g)(\mathfrak{g}_{D,D^{\prime}},\sigma,g) has continuous parameters (see also Figure 4.1).

Our main result in this paper is:

Theorem 1.5.

Let (M,σ,g)(M,\sigma,g) be a four-dimensional reducible and indecomposable 1-connected pseudo-Riemannian symmetric space with signature (2,2)(2,2) whose transvection group is solvable. We denote by GtG_{\text{t}} and GiG_{\text{i}} its transvection group and isometry group respectively, and set closed subgroups HtGtH_{\text{t}}\subset G_{\text{t}} and HiGiH_{\text{i}}\subset G_{\text{i}} by MGt/HtGi/HiM\simeq G_{\text{t}}/H_{\text{t}}\simeq G_{\text{i}}/H_{\text{i}}.

  1. (1)

    Gt/HtG_{\text{t}}/H_{\text{t}} admits compact Clifford–Klein forms if and only if its corresponding symmetric triple is (𝔤±I1,1,I1,1,σ,g)(\mathfrak{g}_{\pm I_{1,1},I_{1,1}},\sigma,g).

  2. (2)

    Gi/HiG_{\text{i}}/H_{\text{i}} admits compact Clifford–Klein forms if and only if its corresponding symmetric triple is (𝔤±I1,1,I1,1,g,σ)(\mathfrak{g}_{\pm I_{1,1},I_{1,1}},g,\sigma) or (𝔤nil,σ,g±)(\mathfrak{g}_{\rm nil},\sigma,g_{\pm}).

symmetric triple (𝔤D,D,σ,g)(\mathfrak{g}_{D,D^{\prime}},\sigma,g) (𝔤nil,σ,g±)(\mathfrak{g}_{\rm nil},\sigma,g_{\pm})
Gt/HtG_{\text{t}}/H_{\text{t}} (D,D)=(±I1,1,I1,1)(D,D^{\prime})=(\pm I_{1,1},I_{1,1}) Never
Gi/HiG_{\text{i}}/H_{\text{i}} (D,D)=(±I1,1,I1,1)(D,D^{\prime})=(\pm I_{1,1},I_{1,1}) Always
Table 1: Symmetric spaces which admit compact Clifford–Klein forms.

To prove this theorem, we use two strategies, the constructors (see Definition 2.34) and intermediate syndetic hulls (see Definition 5.27). The idea of the constructor for reductive case was introduced by T. Kobayashi ([17]), and the following conjecture remains open.

Conjecture 1.6 ([26, Conjecture 3.3.10]).

If a homogeneous space G/HG/H of reductive type admits compact Clifford–Klein forms, then G/HG/H admits a reductive constructor, that is, there exists a subgroup LL which is reductive in GG and acts properly and cocompactly on G/HG/H.

Remark that the conjecture above does not assert that for a compact Clifford–Klein form Γ\G/H\Gamma\backslash G/H of a homogeneous space G/HG/H of reductive type, there exists a reductive constructor LL containing Γ\Gamma cocompactly. A Clifford–Klein form Γ\G/H\Gamma\backslash G/H is standard ([14]) if Γ\Gamma contained in some reductive subgroup LL of GG acting properly on XX. In some cases, we obtain a non-standard compact Clifford–Klein form by deforming standard one (see [12, 23]).

We show the assumption ‘reductive type’ in this conjecture is crucial by showing a ‘solvable analogue’ of the conjecture does not hold (Example 7.1).

Organizations of this paper. Section 2 gives basic concepts of pseudo-Riemannian symmetric spaces and Clifford–Klein forms. In Section 3, we show some general properties about properness and freeness in 1-connected solvable Lie groups. Then we define a class of symmetric spaces GD,D/HG_{D,D^{\prime}}/H in Section 4. We prove the main theorem in Section 6 using the necessary and sufficient condition for the existence of compact Clifford–Klein forms of GD,D/HG_{D,D^{\prime}}/H given in Section 5. Finally, we show a ‘solvable analogue’ of the Kobayashi’s conjecture does not hold in Section 7.

2 Preliminaries

In this section, we review some basic concepts of pseudo-Riemannian symmetric spaces and Clifford–Klein forms.

2.1 Notation

In this subsection, we prepare notation used in this paper.

Notation 2.1.
  • ×{\mathbb{R}}^{\times} := {0}{\mathbb{R}}-\{0\},

  • ee : the identity element of a group,

  • g\mathcal{I}_{g} : the inner automorphism with respect to an element gg of a group,

  • SH:={sh|sS,hH}\mathcal{I}_{S}H:=\left\{\mathcal{I}_{s}h\ \left|\vphantom{\mathcal{I}_{s}h}\ s\in S,h\in H\right.\right\} for subsets S,HS,H of a group,

  • 𝒵G\mathcal{Z}_{G} : the center of a group GG,

  • 𝒵G(g)\mathcal{Z}_{G}(g) : the centralizer of an element gGg\in G in a group GG,

  • 𝒩G(L)\mathcal{N}_{G}(L) : the normalizer of a subgroup LGL\subset G in a group GG,

  • Der𝔤\operatorname{Der}\mathfrak{g} : the derivation algebra of a Lie algebra 𝔤\mathfrak{g},

  • Ip,q:=(IpIq)GL(p+q,)I_{p,q}:=\begin{pmatrix}I_{p}&\\ &-I_{q}\end{pmatrix}\in GL(p+q,{\mathbb{R}}),

  • MTM^{T} : the transposed matrix of a matrix MM,

  • diag(ai):=(a1a2an)\operatorname{diag}\left(a_{i}\right):=\begin{pmatrix}a_{1}&&&\\ &a_{2}&&\\ &&\ddots&\\ &&&a_{n}\end{pmatrix},

  • Sym(n,):={MM(n,)|M is symmetric}\mathrm{Sym}(n,{\mathbb{R}}):=\left\{M\in M(n,{\mathbb{R}})\ \left|\vphantom{M\in M(n,{\mathbb{R}})}\ M\text{ is symmetric}\right.\right\},

  • Sym(reg)(n,):={MSym(n,)|detM0}\mathrm{Sym}^{(\text{reg})}(n,{\mathbb{R}}):=\left\{M\in\mathrm{Sym}(n,{\mathbb{R}})\ \left|\vphantom{M\in\mathrm{Sym}(n,{\mathbb{R}})}\ \det M\neq 0\right.\right\},

  • Qν:=(ν11ν)M(2,)Q_{\nu}:=\begin{pmatrix}\nu&1\\ 1&-\nu\end{pmatrix}\in M(2,{\mathbb{R}}).

In this paper, we use the terminology inner product as a non-degenerate symmetric bilinear form (not necessarily positive definite) and Lie algebras are real and finite dimensional.

2.2 Symmetric triples and pseudo-Riemannian symmetric spaces

In this subsection, we recall a correspondence between 1-connected pseudo-Riemannian symmetric spaces and symmetric triples.

Definition 2.2 (metric Lie algebra with involution, [8, 15]).

Let 𝔤\mathfrak{g} be a Lie algebra, σ\sigma an involution on 𝔤\mathfrak{g} and gg an (indefinite) inner product on 𝔤\mathfrak{g}. We say (𝔤,σ,g)(\mathfrak{g},\sigma,g) is a metric Lie algebra with involution if 𝔤\mathfrak{g}, σ\sigma and gg are mutually compatible, that is, satisfy the following conditions:

  1. (1)

    the inner product gg is σ\sigma-invariant,

  2. (2)

    the inner product gg is 𝔤\mathfrak{g}-invariant, namely,

    g([X,Y],Z)+g(Y,[X,Z])=0(X,Y,Z𝔤).g([X,Y],Z)+g(Y,[X,Z])=0\quad(\forall X,Y,Z\in\mathfrak{g}).
Definition 2.3 (symmetric triple, [8, 15]).

A metric Lie algebra with involution (𝔤,σ,g)(\mathfrak{g},\sigma,g) is called a symmetric triple if the subspace 𝔮𝔤σ\mathfrak{q}\coloneqq\mathfrak{g}^{-\sigma} satisfies [𝔮,𝔮]=𝔤σ[\mathfrak{q},\mathfrak{q}]=\mathfrak{g}^{\sigma}.

Definition 2.4 (homomorphism on symmetric triple, [8, 15]).

For two symmetric triples (𝔤1,σ1,g1)(\mathfrak{g}_{1},\sigma_{1},g_{1}) and (𝔤2,σ2,g2)(\mathfrak{g}_{2},\sigma_{2},g_{2}), a Lie algebra homomorphism ϕ:𝔤1𝔤2\phi:\mathfrak{g}_{1}\rightarrow\mathfrak{g}_{2} is said to be a homomorphism of symmetric triple if ϕ\phi is compatible with the involutions and the inner products, that is, satisfies the following conditions:

  1. (1)

    σ2ϕ=ϕσ1\sigma_{2}\circ\phi=\phi\circ\sigma_{1},

  2. (2)

    g2(ϕ(X1),ϕ(X2))=g1(X1,X2)g_{2}(\phi(X_{1}),\phi(X_{2}))=g_{1}(X_{1},X_{2})  (X1,X2𝔤1\forall X_{1},X_{2}\in\mathfrak{g}_{1}).

Note 2.5.

Let 𝔤\mathfrak{g} be a Lie algebra and 𝔥\mathfrak{h} its subalgebra. The following correspondence is bijective:

{involutions σ on 𝔤 satisfying 𝔤σ=𝔥}\displaystyle\{\text{involutions $\sigma$ on $\mathfrak{g}$ satisfying $\mathfrak{g}^{\sigma}=\mathfrak{h}$}\}
{complementary spaces 𝔮𝔤 of 𝔥 satisfying [𝔮,𝔥]𝔮 and [𝔮,𝔮]𝔥}\displaystyle\simeq\{\text{complementary spaces $\mathfrak{q}\subset\mathfrak{g}$ of $\mathfrak{h}$ satisfying $[\mathfrak{q},\mathfrak{h}]\subset\mathfrak{q}$ and $[\mathfrak{q},\mathfrak{q}]\subset\mathfrak{h}$}\}
σ𝔤σ.\displaystyle\sigma\mapsto\mathfrak{g}^{-\sigma}.
Fact 2.6 ([7]).

Let 𝔤\mathfrak{g} be a Lie algebra and σ\sigma its involution. Put 𝔥:=𝔤σ\mathfrak{h}:=\mathfrak{g}^{\sigma} and 𝔮:=𝔤σ\mathfrak{q}:=\mathfrak{g}^{-\sigma}. If [𝔮,𝔮]=𝔥[\mathfrak{q},\mathfrak{q}]=\mathfrak{h}, then the following restriction is bijective:

{𝔤-invariant inner product on 𝔤}\displaystyle\{\text{$\mathfrak{g}$-invariant inner product on $\mathfrak{g}$}\} {𝔥-invariant inner product on 𝔮}\displaystyle\simeq\{\text{$\mathfrak{h}$-invariant inner product on $\mathfrak{q}$}\}
g\displaystyle g g|𝔮×𝔮.\displaystyle\mapsto g|_{\mathfrak{q}\times\mathfrak{q}}.

Especially, any 𝔤\mathfrak{g}-invariant inner product is also σ\sigma-invariant in this case.

Note 2.7.

For a Lie algebra 𝔤\mathfrak{g} and its subalgebra 𝔥\mathfrak{h}, assume a subspace 𝔮𝔤\mathfrak{q}\subset\mathfrak{g} and an inner product gg on 𝔮\mathfrak{q} satisfy the following conditions:

  • 𝔤=𝔮𝔥\mathfrak{g}=\mathfrak{q}\oplus\mathfrak{h},

  • [𝔮,𝔥]𝔮[\mathfrak{q},\mathfrak{h}]\subset\mathfrak{q} and [𝔮,𝔮]=𝔥[\mathfrak{q},\mathfrak{q}]=\mathfrak{h},

  • gg is 𝔥\mathfrak{h}-invariant.

Then a symmetric triple (𝔤,σ,g)(\mathfrak{g},\sigma,g) is uniquely determined by gg and 𝔮\mathfrak{q} (see Note 2.5 and Fact 2.6).

Definition 2.8 ([8, 15]).

For a symmetric triple (𝔤,σ,g)(\mathfrak{g},\sigma,g), we call the signature of gg (on 𝔮\mathfrak{q}) the signature of the symmetric triple.

In the following, we review the correspondence between symmetric triples and pseudo-Riemannian symmetric spaces.

Fact 2.9 ([7, Ch.I Section 2][15]).

There is a bijection between the isomorphic classes of 11-connected pseudo-Riemannian symmetric spaces with signature (p,q)(p,q) and the isomorphic classes of symmetric triples with signature (p,q)(p,q). Let (𝔤,σ,g)(\mathfrak{g},\sigma,g) be a symmetric triple and MM its corresponding pseudo-Riemannian symmetric space, then 𝔤\mathfrak{g} is the Lie algebra of the transvection group of MM. The 1-connected Lie group GG with Lie algebra 𝔤\mathfrak{g} is the transvection group of MM.

Like the case of Riemannian symmetric spaces, the goal of the classification problem of pseudo-Riemannian symmetric spaces is to classify their ‘minimum units’, which are indecomposable. For Riemannian spaces, they are irreducible symmetric spaces, but are not necessarily for pseudo-Riemannian spaces. We define reducibilities and decomposabilities of symmetric triples and symmetric spaces.

Definition 2.10 ([32]).

We say a symmetric triple (𝔤,σ,g)(\mathfrak{g},\sigma,g) is reducible if the isotropy representation ad:𝔥𝔤𝔩(𝔮){\rm ad}:\mathfrak{h}\to\mathfrak{g}\mathfrak{l}(\mathfrak{q}) is reducible for 𝔥:=𝔤σ\mathfrak{h}:=\mathfrak{g}^{\sigma} and 𝔮:=𝔤σ\mathfrak{q}:=\mathfrak{g}^{-\sigma}. A 1-connected pseudo-Riemannian symmetric space G/HG/H is said to be reducible if its triple is reducible.

Definition 2.11 ([8, 15]).

For two symmetric triples 𝔱1=(𝔤1,σ1,g1)\mathfrak{t}_{1}=(\mathfrak{g}_{1},\sigma_{1},g_{1}) and 𝔱2=(𝔤2,σ2,g2)\mathfrak{t}_{2}=(\mathfrak{g}_{2},\sigma_{2},g_{2}), the triple 𝔱1𝔱2(𝔤1𝔤2,σ1σ2,g1g2)\mathfrak{t}_{1}\oplus\mathfrak{t}_{2}\coloneqq(\mathfrak{g}_{1}\oplus\mathfrak{g}_{2},\sigma_{1}\oplus\sigma_{2},g_{1}\oplus g_{2}) is also a symmetric triple. We say 𝔱1𝔱2\mathfrak{t}_{1}\oplus\mathfrak{t}_{2} is the direct sum of 𝔱1\mathfrak{t}_{1} and 𝔱2\mathfrak{t}_{2}. A symmetric triple is said to be decomposable if it is written as the direct sum of two non-trivial symmetric triples.

Definition 2.12 ([8, 15]).

A pseudo-Riemannian symmetric space is said to be decomposable if the space is isomorphic to the direct product of two non-trivial pseudo-Riemannian symmetric spaces.

The decomposability of pseudo-Riemannian symmetric spaces corresponds to that of symmetric triples.

Proposition 2.13 ([7, Proposition 4.4]).

Let MM be a 1-connected pseudo-Riemannian symmetric space, and 𝔱(𝔤,σ,g)\mathfrak{t}\coloneqq(\mathfrak{g},\sigma,g) the corresponding symmetric triple. Then the following correspondence is one to one.

{decompositions of 𝔱}{decompositions of M}𝔱1𝔱2M1×M2,\{\text{decompositions of $\mathfrak{t}$}\}\to\{\text{decompositions of $M$}\}\quad\mathfrak{t}_{1}\oplus\mathfrak{t}_{2}\mapsto M_{1}\times M_{2},

where M1M_{1} and M2M_{2} are the corresponding 1-connected pseudo-Riemannian symmetric spaces of symmetric triples 𝔱1\mathfrak{t}_{1} and 𝔱2\mathfrak{t}_{2}, respectively.

2.3 Isometry groups

In this subsection, we prepare some lemmas which are used for calculating isometry group of pseudo-Riemannian symmetric spaces. For this, we use:

Notation 2.14.

For a 1-connected pseudo-Riemannian symmetric space (M,S,g)(M,S,g), we set the origin point oMo\in M. We put

  • GiG_{\text{i}} : the isometry group, HiH_{\text{i}} : the stabilizer of oo,

  • GtG_{\text{t}} : the transvection group, HtH_{\text{t}} : the stabilizer of oo,

  • 𝔤i=𝔥i𝔮i\mathfrak{g}_{\text{i}}=\mathfrak{h}_{\text{i}}\oplus\mathfrak{q}_{\text{i}} : the Lie algebra of GiG_{\text{i}} and its eigenspace deconposition with respect to AdS0{\rm Ad}_{S_{0}}.

  • 𝔤t=𝔥t𝔮t\mathfrak{g}_{\text{t}}=\mathfrak{h}_{\text{t}}\oplus\mathfrak{q}_{\text{t}} : the Lie algebra of GtG_{\text{t}} and its eigenspace deconposition with respect to AdS0{\rm Ad}_{S_{0}}.

Lemma 2.15.

For a 1-connected pseudo Riemannian symmetric space (M,S,g)(M,S,g), the map HiGtGi,(h,g)ghH_{\text{i}}\ltimes G_{\text{t}}\to G_{\text{i}},\ (h,g)\mapsto gh is surjective group homomorphism and its kernel is:

Ht~:={(g,g1)HiGt|gHt}.\widetilde{H_{\text{t}}}:=\left\{(g,g^{-1})\in H_{\text{i}}\ltimes G_{\text{t}}\ \left|\vphantom{(g,g^{-1})\in H_{\text{i}}\ltimes G_{\text{t}}}\ g\in H_{\text{t}}\right.\right\}.

Especially, we have a Lie group isomorphism Gi(HiGt)/Ht~G_{\text{i}}\simeq(H_{\text{i}}\ltimes G_{\text{t}})/\widetilde{H_{\text{t}}}. Moreover, if there exists a closed subgroup THiT\subset H_{\text{i}} satisfying HiTHtH_{\text{i}}\simeq T\ltimes H_{\text{t}}, we have a Lie group isomorphism GiTGtG_{\text{i}}\simeq T\ltimes G_{\text{t}}.

Proof..

Since the GtG_{\text{t}}-action on MM is transitive, for any gGig\in G_{\text{i}}, there exists gGtg^{\prime}\in G_{\text{t}} satisfying g1gHi{g^{\prime}}^{-1}g\in H_{\text{i}}. Therefore, the map HiGtGi,(h,g)ghH_{\text{i}}\ltimes G_{\text{t}}\to G_{\text{i}},\ (h,g)\mapsto gh is surjective group homomorphism. The latter statement follows from the following Note with A:=T,B:=HtA:=T,\ B:=H_{\text{t}} and C:=GtC:=G_{\text{t}}. ∎

Note 2.16.

Let A,BA,B and CC be Lie groups. Assume that there exists a map ϕ:AAut(C)\phi:A\to\operatorname{Aut}(C) such that BCB\subset C is a ϕ\phi-invariant closed subgroup. We set ψ:ABAut(C)\psi:A\ltimes B\to\operatorname{Aut}(C) by ψ(a,b)(c)=bϕa(c)\psi(a,b)(c)=\mathcal{I}_{b}\phi_{a}(c). Then ψ\psi is a Lie group homomorphism. Moreover the following map ff is also a Lie group homomorphism and its kernel is {(e,b,b1)|bB}=:B~\left\{(e,b,b^{-1})\ \left|\vphantom{(e,b,b^{-1})}\ b\in B\right.\right\}=:\widetilde{B}.

f:(AB)ψCAC,(a,b,c)(a,cb).f:(A\ltimes B)\ltimes_{\psi}C\to A\ltimes C,\quad(a,b,c)\mapsto(a,cb).

Especially, we have ((AB)C)/B~AC((A\ltimes B)\ltimes C)/\widetilde{B}\simeq A\ltimes C.

Lemma 2.17.

Let (𝔤t,σ,g)(\mathfrak{g}_{\text{t}},\sigma,g) is the symmetric triple which corresponds to 1-connected pseudo-Riemannian symmetric space (M,S,g)(M,S,g). Then the map Φ:HiAut(𝔤t,σ,g),hAdh|𝔤t\Phi:H_{\text{i}}\to\operatorname{Aut}(\mathfrak{g}_{\text{t}},\sigma,g),\quad h\mapsto{\rm Ad}_{h}|_{\mathfrak{g}_{\text{t}}} is a Lie group isomorphism.

Proof..

First, we check that Φ\Phi is well-defined. Since GtGiG_{\text{t}}\subset G_{\text{i}} is a normal subgroup, we have h(Gt)Gt\mathcal{I}_{h}(G_{\text{t}})\subset G_{\text{t}}. By HiGiσH_{\text{i}}\subset G_{\text{i}}^{\sigma}, we have S0h=h\mathcal{I}_{S_{0}}h=h. Therefore AdS0{\rm Ad}_{S_{0}} and Adh{\rm Ad}_{h} are commutative, namely, Adh{\rm Ad}_{h} preserves the decomposition 𝔤i=𝔥i𝔮i\mathfrak{g}_{\text{i}}=\mathfrak{h}_{\text{i}}\oplus\mathfrak{q}_{\text{i}} and 𝔤t=𝔥t𝔮t\mathfrak{g}_{\text{t}}=\mathfrak{h}_{\text{t}}\oplus\mathfrak{q}_{\text{t}}. On the other hand, Adh{\rm Ad}_{h} preserves the inner product on 𝔮i=𝔮t\mathfrak{q}_{\text{i}}=\mathfrak{q}_{\text{t}}. Since it is commutative with σ\sigma and a Lie algebra homomorphism, it also preserves the inner product on 𝔥t\mathfrak{h}_{\text{t}}, and so gg on 𝔤t\mathfrak{g}_{\text{t}}. Then we have Adh|𝔤tAut(𝔤t,σ,g){\rm Ad}_{h}|_{\mathfrak{g}_{\text{t}}}\in\operatorname{Aut}(\mathfrak{g}_{\text{t}},\sigma,g).

We set a map Ψ:Aut(𝔤t,σ,g)Hi,AΨ(A):=fA¯\Psi:\operatorname{Aut}(\mathfrak{g}_{\text{t}},\sigma,g)\to H_{\text{i}},\ A\mapsto\Psi(A):=\overline{f_{A}} as follows. Let (Gt,Ht,σ,g)(G_{\text{t}},H_{\text{t}},\sigma,g) be the symmetric triple which corresponds to the symmetric triple (𝔤t,σ,g)(\mathfrak{g}_{\text{t}},\sigma,g). Put fA:GtGtf_{A}:G_{\text{t}}\to G_{\text{t}} the induced Lie group homomorphism by the Lie group homomorphism A:𝔤t𝔤tA:\mathfrak{g}_{\text{t}}\to\mathfrak{g}_{\text{t}}. Note that fAf_{A} is a Lie group isomorphism. Since fAf_{A} preserves HtH_{\text{t}}, it induces fA¯:Gt/HtGt/Ht\overline{f_{A}}:G_{\text{t}}/H_{\text{t}}\to G_{\text{t}}/H_{\text{t}}, and it is an isometry by Note 2.18. Therefore we have fA¯Hi\overline{f_{A}}\in H_{\text{i}}.

Next we check that Ψ\Psi is the inverse map of Φ\Phi. First we show ΨΦ=id\Psi\circ\Phi=\operatorname{id}. Let hHih\in H_{\text{i}}. It is enough to show Ψ(Φ(h))=fAdh¯\Psi(\Phi(h))=\overline{f_{{\rm Ad}_{h}}} is coincides to hh. It follows from fAdh=hf_{{\rm Ad}_{h}}=\mathcal{I}_{h}, and h:MMh:M\to M and fAdh¯:MM\overline{f_{{\rm Ad}_{h}}}:M\to M make the following diagram commutative:

Gt\textstyle{G_{\text{t}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}fAdh\scriptstyle{f_{{\rm Ad}_{h}}}π\scriptstyle{\pi}Gt\textstyle{G_{\text{t}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}π\scriptstyle{\pi}M\textstyle{M\ignorespaces\ignorespaces\ignorespaces\ignorespaces}M.\textstyle{M.}

Then we check ΦΨ=id\Phi\circ\Psi=\operatorname{id}. Let AAut(𝔤t,σ,g)A\in\operatorname{Aut}(\mathfrak{g}_{\text{t}},\sigma,g) and fA¯Hi\overline{f_{A}}\in H_{\text{i}}. We are enough to show AdfA¯=A{\rm Ad}_{\overline{f_{A}}}=A, which is equivalent to AdfA¯|𝔮t=A|𝔮t{\rm Ad}_{\overline{f_{A}}}|_{\mathfrak{q}_{\text{t}}}=A|_{\mathfrak{q}_{\text{t}}}. For X𝔮X\in\mathfrak{q}, we have:

A(X)=d(fA¯)eH(X)=ddt|t=0fA¯(exptX)Hi=AdfA¯(X).A(X)=d(\overline{f_{A}})_{eH}(X)=\dfrac{d}{dt}\Big{|}_{t=0}\overline{f_{A}}(\exp tX)H_{\text{i}}={\rm Ad}_{\overline{f_{A}}}(X).

Here, A(X)=d(fA¯)A(X)=d(\overline{f_{A}}) follows from Note 2.19.

Note 2.18.

Let GG be a Lie group, HH its closed subgroup and gg a GG-invariant inner product on G/HG/H. Suppose that a group isomorphism f:GGf:G\to G preserves HH and put f¯:G/HG/H\overline{f}:G/H\to G/H the induced map. If df¯eHd\overline{f}_{eH} preserves inner product geHg_{eH} on TeH(G/H)T_{eH}(G/H), then f¯\overline{f} is an isometry.

Note 2.19.

Let ff be an automorphism on (G,H,σ)(G,H,\sigma). Put 𝔤:=Lie(G)\mathfrak{g}:=\operatorname{Lie}(G) and 𝔮:=𝔤σ\mathfrak{q}:=\mathfrak{g}^{\sigma}. Then we have df¯|eH=dfe|𝔮d\overline{f}|_{eH}=df_{e}|_{\mathfrak{q}}, where f¯:G/HG/H\overline{f}:G/H\to G/H is the induced map by ff and we identify TeHG/H𝔮T_{eH}G/H\simeq\mathfrak{q}.

Note 2.20.

Let GG be a Lie group, AGA\subset G a closed subgroup and NGN\subset G a normal subgroup. If G=ANG=AN and AN={e}A\cap N=\{e\}, then we have GANG\simeq A\ltimes N, where AA acts on NN as an inner automorphism.

2.4 Clifford–Klein forms

In this subsection, we review Clifford–Klein forms following [17] and [22].

Definition 2.21 ([17]).

Let GG be a Lie group, HH its closed subgroup, and Γ\Gamma its discrete subgroup. Assume the Γ\Gamma-action on G/HG/H is (fixed point) free and properly discontinuous. Then the quotient space Γ\G/H\Gamma\backslash G/H has the unique manifold structure such that the natural surjection π:G/HΓ\G/H\pi:G/H\to\Gamma\backslash G/H is a CC^{\infty}-covering map. The manifold Γ\G/H\Gamma\backslash G/H is said to be a Clifford–Klein form of G/HG/H.

In the study of Clifford–Klein forms, Problem 1.1 is a significant open question. Let us recall basic terminologies for Problem 1.1.

Definition 2.22 ([17]).

Suppose a locally compact group LL acts on a locally compact space XX. The LL-action is said to be proper if {L|SS}\left\{\ell\in L\ \left|\vphantom{\ell\in L}\ \ell S\cap S\neq\emptyset\right.\right\} is compact for any compact subset SXS\subset X.

It is easy to check the following:

Note 2.23.

In the setting of Definition 2.21, if the LL-action on XX is proper, any LL-orbit is closed in XX.

Fact 2.24 ([17]).

Let LL be a locally compact group and XX a locally compact space. Assume LL acts on XX and Γ\Gamma is a uniform lattice (cocompact discrete subgroup) of LL. Then the following statements hold.

  1. (1)

    The Γ\Gamma-action on XX is cocompact if and only if so is the LL-action.

  2. (2)

    The Γ\Gamma-action on XX is properly discontinuous if and only if the LL-action is proper.

We recall some definitions and properties.

Definition 2.25 ([18, Definition 6], [21, Definition 2.1.1]).

Let GG be a locally compact group, and LL and HH its subsets.

  1. (1)

    We say the pair (L,H)(L,H) is proper in GG, denoted by LHL\pitchfork H in GG, if the set LSHSL\cap SHS is relatively compact in GG for any compact set SGS\subset G.

  2. (2)

    We say the pair (L,H)(L,H) has the property (CI) in GG, if the set LgHg1L\cap gHg^{-1} is relatively compact in GG for any gGg\in G.

  3. (3)

    We say the pair (L,H)(L,H) is free if the condition LgHg1={e}L\cap gHg^{-1}=\{e\} holds for any gGg\in G.

  4. (4)

    We denote by LHL\sim H in GG the existence of a compact set SGS\subset G satisfying LSHSL\subset SHS and HSLSH\subset SLS.

Remark 2.26.

In [18], Kobayashi defined the property (CI) for subgroups LL and HH, but here we define it for subsets LL and HH for the sake of Lemma 2.30.

Property 2.27 ([18, 21]).

Let GG be a locally compact group, and H,HH,H^{\prime} and LL its subsets.

  1. (1)

    The pair (L,H)(L,H) is proper (resp. has the property (CI), is free) in GG if and only if so is (H,L)(H,L) in GG.

  2. (2)

    The relation \sim is an equivalence relation.

  3. (3)

    If HHH\sim H^{\prime} in GG, then LHL\pitchfork H if and only if LHL\pitchfork H^{\prime} in GG.

  4. (4)

    If the pair (L,H)(L,H) is proper, then (L,H)(L,H) has the property (CI).

Property 2.28 ([21, Observation 2.13]).

Let GG be a locally compact group, and LL and HH its closed subgroups.

  1. (1)

    The LL-action on G/HG/H is proper if and only if LHL\pitchfork H in GG.

  2. (2)

    The LL-action on G/HG/H is free if and only if the pair (L,H)(L,H) is free in GG.

Property 2.29.

Let GG be a locally compact group and NN its closed normal subgroup, and HH, LL closed subgroups of GG. We denote by G~\widetilde{G}, H~\widetilde{H} and L~\widetilde{L} the image of GG, LL and HH, respectively, by the natural projection π:GG/N\pi:G\to G/N. Then we have:

  1. (1)

    If N/(LN)N/(L\cap N) is compact and LL is discrete, L~G~\widetilde{L}\subset\widetilde{G} is also discrete.

  2. (2)

    If N/(LN)N/(L\cap N) is compact, the condition LHL\pitchfork H in GG implies L~H~\widetilde{L}\pitchfork\widetilde{H} in G~\widetilde{G}.

  3. (3)

    If L\G/HL\backslash G/H is compact, so is L~\G~/H~\widetilde{L}\backslash{\widetilde{G}}/\widetilde{H}.

The statement (1) follows from [11, Lemma 5.1.4]. The statement (2) follows from [20, Lemma 1.3(2)]. The statement (3) is easy.

Finally, we prepare some easy lemmas which are used in Sections 3 and 5.

Lemma 2.30.

Let C1C_{1} and C2C_{2} be two closed cones in n{\mathbb{R}}^{n}. Then the following conditions are equivalent:

  1. (a)

    the pair (C1,C2)(C_{1},C_{2}) has the (CI) property, namely, C1C2={0}C_{1}\cap C_{2}=\{0\},

  2. (b)

    the pair (C1,C2)(C_{1},C_{2}) is proper in n{\mathbb{R}}^{n}.

Proof..

Since the implication (b)\Rightarrow(a) is easy, we prove the implication (a)\Rightarrow(b). We take any R>0R\in{\mathbb{R}}_{>0} and denote by B(R)B(R) the closed ball in n{\mathbb{R}}^{n}. It is enough to show that (C1+B(R))C2(C_{1}+B(R))\cap C_{2} is relatively compact. Set d0>0d_{0}\in{\mathbb{R}}_{>0} as the distance between C1C_{1} and C2Sn1C_{2}\cap S^{n-1}, where Sn1S^{n-1} is the unit sphere in n{\mathbb{R}}^{n}. Take any x(C1+B(R))C2x\in(C_{1}+B(R))\cap C_{2}, then we have Rd(C1,x)xd0R\geq d(C_{1},x)\geq\|x\|d_{0}, and so (C1+B(R))C2B(R/d0)(C_{1}+B(R))\cap C_{2}\subset B(R/d_{0}). ∎

Lemma 2.31.

Let GG be a locally compact group and NN its closed normal subgroup. Let L0L_{0} and HH be subsets of NN, and L1L_{1}\neq\emptyset a subset of GG satisfying L1NL_{1}\pitchfork N in GG. Set L:=L1L0L:=L_{1}L_{0}, then the following conditions are equivalent:

  1. (a)

    (SL0)H(\mathcal{I}_{S}L_{0})\pitchfork H in NN for any compact set SGS\subset G,

  2. (b)

    L0HL_{0}\pitchfork H in GG,

  3. (c)

    LHL\pitchfork H in GG.

Proof..

Since the implications (c)\Rightarrow(b)\Rightarrow(a) are easy, we prove the implication (a)\Rightarrow(c). Let SGS\subset G be a compact set. We have:

SLS1HSL1S1(SL0)H=(SL1S1N)SL0H.SLS^{-1}\cap H\subset SL_{1}S^{-1}(\mathcal{I}_{S}L_{0})\cap H=(SL_{1}S^{-1}\cap N)\mathcal{I}_{S}L_{0}\cap H.

By the assumption L1NL_{1}\pitchfork N in GG, we take a compact set KGK\subset G satisfying SL1S1NKSL_{1}S^{-1}\cap N\subset K. Then we have:

(SL1S1N)SL0HK(SL0)H.(SL_{1}S^{-1}\cap N)\mathcal{I}_{S}L_{0}\cap H\subset K(\mathcal{I}_{S}L_{0})\cap H.

By the condition (a), the subset K(SL0)HK(\mathcal{I}_{S}L_{0})\cap H is relatively compact in GG. Therefore the condition (c) follows. ∎

Lemma 2.32.

Let LL and NN be locally compact groups. Assume LL acts on NN continuously as group automorphisms. Put G:=LNG:=L\ltimes N, then we have LNL\pitchfork N in GG.

Proof..

It is enough to show L(S1×S2)N(S1×S2)L\cap(S_{1}\times S_{2})N(S_{1}\times S_{2}) is relatively compact for any compact subsets S1LS_{1}\subset L and S2NS_{2}\subset N. This follows from:

L(S1×S2)N(S1×S2)LS1S1=S1S1.L\cap(S_{1}\times S_{2})N(S_{1}\times S_{2})\subset L\cap S_{1}S_{1}=S_{1}S_{1}.

Note 2.33.

Let GG be a Lie group with finite connected component and HH its closed subgroup. Put G0GG_{0}\subset G the identity component. Put H0:=G0HH_{0}:=G_{0}\cap H. Then G/HG/H admits compact Clifford–Klein forms if and only if so does G0/H0G_{0}/H_{0}.

2.5 Constructors

By Fact 2.24, it is natural to define the following subgroups called constructors, of which the terminology is introduced in [27]111see also http://coe.math.sci.hokudai.ac.jp/sympo/ccyr/2006/pdf/TaroYOSHINO.pdf. In this subsection, we define constructors and see some basic properties.

Definition 2.34 ([27]).

Let G/HG/H be a homogeneous space of a Lie group GG. A closed and connected subgroup LGL\subset G is said to be a constructor of G/HG/H if the natural action of LL on G/HG/H is proper and cocompact.

We think constructors for homogeneous spaces of solvable type. We note:

Fact 2.35 ([13]).

A connected subgroup of a 1-connected solvable Lie group is closed.

It is important to consider the existence of a constructor for the existence problem of compact Clifford–Klein forms.

Definition 2.36 ([35]).

Let GG be a Lie group and Γ\Gamma its closed subgroup. We say a closed and connected subgroup LGL\subset G a syndetic hull of Γ\Gamma if LL includes Γ\Gamma cocompactly.

Fact 2.37 ([34, 1]).

Let GG be a 1-connected completely solvable Lie group and ΓG\Gamma\subset G a closed subgroup. Then there exists a unique syndetic hull LL of Γ\Gamma. Especially, if the space G/HG/H has a compact Clifford–Klein form Γ\G/H\Gamma\backslash G/H, the space G/HG/H has a constructor LL which is the syndetic hull of Γ\Gamma.

Remark 2.38.

The assumption of complete solvability in the above fact is crucial. In fact, a solvable Lie group GG may have a discrete subgroup Γ\Gamma without its syndetic hulls (see Example 7.1).

3 Properness and cocompactness in solvable Lie groups

In this section, we show some criterions to check properness and cocompactness in 1-connected solvable Lie groups, which are used to show the main theorem. The main results in this section are Propositions 3.7, 3.15 and 3.16.

3.1 Freeness and the property (CI) in solvable Lie groups

In this subsection, we review some criterions for freeness and the property (CI) in solvable Lie groups.

First, the following note gives a criterion of the property (CI) for 1-connected nilpotent Lie groups.

Note 3.1 ([24]).

Let GG be a 1-connected nilpotent Lie group, and LL and HH its connected subgroups. Then the following conditions are equivalent:

  1. (a)

    The pair (L,H)(L,H) has the property (CI),

  2. (b)

    AdG𝔩𝔥={0}{\rm Ad}_{G}\mathfrak{l}\cap\mathfrak{h}=\{0\},

  3. (c)

    𝔩AdG𝔥={0}\mathfrak{l}\cap{\rm Ad}_{G}\mathfrak{h}=\{0\}.

Here, 𝔩\mathfrak{l} and 𝔥\mathfrak{h} are the Lie algebras of LL and HH, respectively.

This note is easily shown by using the diffeomorphism exp:𝔤G\exp:\mathfrak{g}\to G. It is easy to show the following note in the same way.

Note 3.2.

Note 3.1 also holds under the assumptions that GG is an arbitrary Lie group and there exists a 1-connected closed normal nilpotent subgroup NGN\subset G satisfying L,HNL,H\subset N.

Next, we review the following:

Fact 3.3 ([13, Theorem 2.3]).

A compact subgroup of a 1-connected solvable Lie group is trivial.

We have two corollaries from this fact.

Corollary 3.4.

Let GG be a 1-connected solvable Lie group, and LL and HH closed subgroups of GG. Then the following conditions are equivalent.

  1. (a)

    The pair (L,H)(L,H) is free in GG.

  2. (b)

    The pair (L,H)(L,H) has the property (CI) in GG.

Remark 3.5.

For a 1-connected exponential solvable Lie group GG, the above statement was proven by Baklouti and Kédim [1].

By using Corollary 3.4 and Property 2.27, we have:

Note 3.6.

Let GG be a 1-connected solvable Lie group, and LL and HH its closed subgroups.

  1. (1)

    If the pair (L,H)(L,H) is proper, the quotient space L\G/HL\backslash G/H has a manifold structure.

  2. (2)

    Let ΓL\Gamma\subset L be a uniform lattice. Assume the action LG/HL\curvearrowright G/H is proper and cocompact, then Γ\G/H\Gamma\backslash G/H is a compact Clifford–Klein form.

3.2 Constructors in solvable homogeneous space

In this subsection, we show some propositions for the existence of constructors in solvable homogeneous spaces. First, we see a criterion of the cocompactness of the LL-action.

Proposition 3.7.

Let GG be a 1-connected solvable Lie group, and LL and HH its connected subgroups. Assume the LL-action on G/HG/H is proper. Then the following conditions are equivalent:

  1. (a)

    the space L\G/HL\backslash G/H is compact,

  2. (b)

    G=LHG=LH,

  3. (c)

    𝔤=𝔩𝔥\mathfrak{g}=\mathfrak{l}\oplus\mathfrak{h} as a linear space.

Here, 𝔤\mathfrak{g}, 𝔥\mathfrak{h} and 𝔩\mathfrak{l} are Lie algebras of G,HG,H and LL, respectively.

Proof..

Since the implications (b)\Rightarrow(a) is clear, we first show the implication (a)\Rightarrow(b). The condition (b) is equivalent to the transitivity of the LL-action on G/HG/H, so we are enough to show that the space L\G/HL\backslash G/H consists of one point. Since GG is a 1-connected solvable Lie group, and HH and LL are connected subgroups, G/HG/H and LL are contractible by Note 3.8 and Lemma 3.11 below. By Note 3.6 (1), the quotient space L\G/HL\backslash G/H has a manifold structure, it is one point by Lemma 3.10. Next, we show the implication (b)\Rightarrow(c). Since G/HG/H is an LL-orbit, we have dim(G/H)dimL\dim(G/H)\leq\dim L. On the other hand, by the properness of the LL-action we have 𝔩𝔥={0}\mathfrak{l}\cap\mathfrak{h}=\{0\} (Note 3.9). Then we have dimGdimH+dimL\dim G\geq\dim H+\dim L and so we obtain dimG=dimH+dimL\dim G=\dim H+\dim L, which implies 𝔤=𝔩𝔥\mathfrak{g}=\mathfrak{l}\oplus\mathfrak{h}.

Finally, we check the implication (c)\Rightarrow(b). We consider the LL-orbit of the origin point eHeH. The orbit is closed since the LL-action on G/HG/H is proper. On the other hand, since the LL-action is free by Corollary 3.4, the dimension of LL-orbit equals to dimL=dim(G/H)\dim L=\dim(G/H). Hence the LL-orbit is open. Since G/HG/H is connected, G/HG/H coincides with the LL-orbit. ∎

Note 3.8 ([6]).

Any 1-connected solvable Lie group is diffeomorphic to a Euclidian space.

Note 3.9.

Let HH and LL be closed subgroups of a 1-connected solvable Lie group GG. Assume LHL\pitchfork H in GG, then we have 𝔩𝔥={0}\mathfrak{l}\cap\mathfrak{h}=\{0\}.

Lemma 3.10.

Suppose a contractible Lie group GG acts on a contractible manifold MM. If the quotient space G\MG\backslash M is a compact manifold, it consists of one point.

This lemma is an immediate consequence of the following two lemmas.

Lemma 3.11.

Let GG be a contractible topological group and act on a contractible space MM. Then G\MG\backslash M is contractible.

Proof..

By the homotopy exact sequence of the fiber bundle GMG\MG\to M\to G\backslash M, we have πi(G\M)=0(i)\pi_{i}(G\backslash M)=0\ (\forall i\in{\mathbb{N}}). Then we have G\MG\backslash M is contractible by J. H. C. Whitehead’s theorem. ∎

Lemma 3.12.

A contractible and closed manifold consists of one point.

Proof..

Let MM be a contractible and closed manifold. Since an arbitrary vector bundle over MM is trivial, MM is orientable. For a volume element ω\omega of MM, we have Mω=0\int_{M}\omega=0 by Stokes’ theorem. ∎

3.3 Properness, the property (CI) and cocompactness

To check the property (CI) is easier than the properness. The property (CI) was introduced by T. Kobayashi and the equivalence of properness and the property (CI) was shown for any pair of closed reductive subgroups of linear reductive Lie groups [18]. Lipsman considered an extension of Kobayashi’s theory to non-reductive case [30]. For 1-connected nilpotent Lie groups, the equivalence of the properties is known as Lipsman’s conjecture. About this conjecture, the following results have been obtained so far. The properness and the property (CI) are equivalent for less than or equal to 3-step nilpotent Lie groups [31, 36, 2] and not necessarily equivalent for 4-step nilpotent Lie groups [36]. In this subsection, we generalize the following Nasrin’s result (Fact 3.13) in Proposition 3.15 and introduce a criterion of cocompactness in a similar setting in Proposition 3.16.

Fact 3.13 ([31] for 2-step nilpotent Lie groups, [36, 2] for 3-step nilpotent Lie groups).

Let GG be a 1-connected 3-step nilpotent Lie group, and LL and HH its connected subgroups. Then LHL\pitchfork H in GG if and only if the pair (L,H)(L,H) has the (CI) property in GG.

Setting 3.14.

Let GG be a Lie group, and NN its closed normal subgroup. Assume NN is 1-connected nilpotent. Let L0L_{0} and HH be connected subgroups of NN, and L1L_{1} a closed subgroup of 𝒩G(L0)\mathcal{N}_{G}(L_{0}) (see Notation 2.1) satisfying L1NL_{1}\pitchfork N in GG. Set L:=L1L0=L0L1L:=L_{1}L_{0}=L_{0}L_{1}.

Proposition 3.15.

Under Setting 3.14, we additionally assume NN is 2-step nilpotent. Then the following conditions are equivalent:

  1. (a)

    the pair (L0,H)(L_{0},H) has the property (CI) in GG,

  2. (b)

    LHinGL\pitchfork H\ \text{in}\ G.

Proof..

It is enough to show that the following four conditions are equivalent:

  1. (i)

    the pair (L0,H)(L_{0},H) has the property (CI) in GG,

  2. (ii)

    AdG𝔩0𝔥={0}{\rm Ad}_{G}\mathfrak{l}_{0}\cap\mathfrak{h}=\{0\},

  3. (iii)

    (SL0)H(\mathcal{I}_{S}L_{0})\pitchfork H in NN for any compact set SGS\subset G,

  4. (iv)

    LHL\pitchfork H in GG.

Here, 𝔩0\mathfrak{l}_{0}, 𝔥\mathfrak{h} and 𝔫\mathfrak{n} are the Lie algebras of L0L_{0}, HH and NN, respectively. Since the exponential map exp:𝔫N\exp:\mathfrak{n}\to N is diffeomorphism, we denote by log\log its inverse.

The implication (i)\Rightarrow(ii) comes from Note 3.2, the implication (iii)\Rightarrow(iv) holds by Lemma 2.31, and the implication (iv)\Rightarrow(i) follows from Property 2.27(4).

Then we show the implication (ii)\Rightarrow(iii). Take any compact sets SGS\subset G and T𝔫T\subset\mathfrak{n}. It is enough to show that the subset (expT(SL0)expT)H(\exp T(\mathcal{I}_{S}L_{0})\exp T)\cap H is compact. Since NN is 2-step nilpotent, for X𝔫X\in\mathfrak{n} we have:

expTexpXexpT\displaystyle\exp T\exp X\exp T exp((id+12ad(TT))X+T+T+12[T,T])\displaystyle\subset\exp\left(\left(\operatorname{id}+\dfrac{1}{2}{\rm ad}(T-T)\right)X+T+T+\dfrac{1}{2}[T,T]\right)
=exp(AdSX+T),\displaystyle=\exp({\rm Ad}_{S^{\prime}}X+T^{\prime}),

where S:=exp((TT)/2)NS^{\prime}:=\exp((T-T)/2)\subset N and T:=T+T+[T,T]/2𝔫T^{\prime}:=T+T+[T,T]/2\subset\mathfrak{n}. Then we have:

log(expT(SL0)expT)AdSAdS𝔩0+T=AdSS𝔩0+T=C+T,\log(\exp T(\mathcal{I}_{S}L_{0})\exp T)\subset{\rm Ad}_{S^{\prime}}{\rm Ad}_{S}\mathfrak{l}_{0}+T^{\prime}={\rm Ad}_{S^{\prime}S}\mathfrak{l}_{0}+T^{\prime}=C+T^{\prime},

where C:=AdSS𝔩0C:={\rm Ad}_{S^{\prime}S}\mathfrak{l}_{0} is a closed cone. On the other hand, by the condition (ii), we have C𝔥={0}C\cap\mathfrak{h}=\{0\}. Hence the pair of the closed cones (C,𝔥)\left(C,\mathfrak{h}\right) is proper by Lemma 2.30. Then the subset (C+T)𝔥\left(C+T^{\prime}\right)\cap\mathfrak{h} is compact and so is the subset (expT(SL0)expT)H(\exp T(\mathcal{I}_{S}L_{0})\exp T)\cap H. ∎

Finally, we introduce criterion of cocompactness under a similar situation in Proposition 3.15.

Proposition 3.16.

Under Setting 3.14, assume L0HL_{0}\pitchfork H in GG. Then the following conditions are equivalent:

  1. (a)

    The LL-action on G/NG/N and the L0L_{0}-action on N/HN/H are cocompact.

  2. (b)

    The LL-action on G/HG/H is cocompact.

Proof..

First, we show the implication (a)\Rightarrow(b). By the assumption L0HL_{0}\pitchfork H in GG and the cocompactness of the L0L_{0}-action, for any gGg\in G, we have g(L0)H\mathcal{I}_{g}(L_{0})\pitchfork H in NN and Adg𝔩0𝔥=𝔫{\rm Ad}_{g}\mathfrak{l}_{0}\oplus\mathfrak{h}=\mathfrak{n}, and so g(L0)H=N\mathcal{I}_{g}(L_{0})H=N (Proposition 3.7). Take a compact subset CGC\subset G satisfying G=LCNG=LCN. We are enough to show G=LCHG=LCH. Then we get:

G=LCN=cCLcN=cCLcc1(L0)H=cCLL0cH=LCH.G=LCN=\bigcup_{c\in C}LcN=\bigcup_{c\in C}Lc\mathcal{I}_{c^{-1}}(L_{0})H=\bigcup_{c\in C}LL_{0}cH=LCH.

Next, we show the implication (b)\Rightarrow(a). By HNH\subset N, the LL-action on G/NG/N is cocompact, so we show that the L0L_{0}-action is compact. Take a compact subset CGC\subset G satisfying G=LCHG=LCH. Then we have:

N=L0L1CHN=L0(L1CN)H.N=L_{0}L_{1}CH\cap N=L_{0}(L_{1}C\cap N)H.

By the condition L1NL_{1}\pitchfork N, we have L1CNL_{1}C\cap N is compact, and so have the L0L_{0}-action on N/HN/H is cocompact. ∎

4 Indecomposable symmetric triples with signature (2,2)(2,2)

In this section, we review the classification of symmetric triples with signature (2,2)(2,2) given by Kath–Olbrich (Fact 1.4). We use another notation for our calculations.

We introduce solvable Lie algebras 𝔤D,D\mathfrak{g}_{D,D^{\prime}} and indecomposable symmetric triples 𝔱D,D:=(𝔤D,D,σ,g)\mathfrak{t}_{D,D^{\prime}}:=(\mathfrak{g}_{D,D^{\prime}},\sigma,g), which are most part of the pseudo-Riemannian symmetric triple with signature (2,2)(2,2) (see Fact 1.4). We also define a symmetric triples (𝔤nil,σ,g±)(\mathfrak{g}_{\rm nil},\sigma,g_{\pm}) in Subsection 4.3.

In this section, we use the following:

Notation 4.1.
  • 𝔥n:=2n\mathfrak{h}_{n}:={\mathbb{R}}^{2n}\oplus{\mathbb{R}} : the (2n+1)(2n+1)-dimensional Heisenberg Lie algebra equipped with the following non-trivial brackets:

    [X,Y]:=ω(X,Y),X,Y2n[X,Y]:=\omega(X,Y),\quad X,Y\in{\mathbb{R}}^{2n}

    where ω\omega is a symplectic form on 2n{\mathbb{R}}^{2n}. Put 𝔥0:=\mathfrak{h}_{0}:={\mathbb{R}}.

  • 𝔷:=\mathfrak{z}:={\mathbb{R}} : the center of 𝔥n\mathfrak{h}_{n}, Z:=1𝔷Z:=1\in\mathfrak{z}.

  • HnH_{n} : Heisenberg Lie group, namely, 1-connected Lie group whose Lie algebra is 𝔥n\mathfrak{h}_{n}.

Remark 4.2.

The correspondence between the spaces of Fact 1.4 and the list of Kath–Olbrich [15, Theorem 7.1] is given by the following table:

[15, Theorem 7.1] (1) (a)(b) (1)(c) (2)(a)(b) (3) (4)
Fact 1.4 II(a),(d) II(b) I II(c) D=DεD^{\prime}=D^{\prime}_{\varepsilon} II(c) D=DεD^{\prime}=-D^{\prime}_{\varepsilon}

,

where Dε:=(011ε)D^{\prime}_{\varepsilon}:=\begin{pmatrix}0&1\\ 1&\varepsilon\end{pmatrix} and ε=±1\varepsilon=\pm 1.

4.1 Definition of symmetric triples 𝔱D,D\mathfrak{t}_{D,D^{\prime}}

In this subsection, we define symmetric triples 𝔱D,D\mathfrak{t}_{D,D^{\prime}} and see some properties.

Definition 4.3.

We think 𝔰𝔭(n,)\mathfrak{s}\mathfrak{p}(n,{\mathbb{R}}) as a subalgebra of Der𝔥n\operatorname{Der}\mathfrak{h}_{n}. For W𝔰𝔭(n,)W\in\mathfrak{s}\mathfrak{p}(n,{\mathbb{R}}), we define:

𝔤W:=W𝔥n𝔰𝔭(n,)𝔥n.\mathfrak{g}_{W}:={\mathbb{R}}W\ltimes\mathfrak{h}_{n}\subset\mathfrak{s}\mathfrak{p}(n,{\mathbb{R}})\ltimes\mathfrak{h}_{n}.

In the following, we use a basis (X1,,Xn,Y1,,Yn)(X_{1},\cdots,X_{n},Y_{1},\cdots,Y_{n}) of 2n{\mathbb{R}}^{2n} such that:

ω=(InIn).\omega=\begin{pmatrix}&I_{n}\\ -I_{n}&\end{pmatrix}.

Using this basis, we identify 𝔰𝔭(n,){WM(2n,)|ωW+WTω=0}\mathfrak{s}\mathfrak{p}(n,{\mathbb{R}})\simeq\left\{W\in M(2n,{\mathbb{R}})\ \left|\vphantom{W\in M(2n,{\mathbb{R}})}\ \omega W+W^{T}\omega=0\right.\right\}.

Definition 4.4 (𝔤D,D\mathfrak{g}_{D,D^{\prime}}).

For matrices D,DSym(reg)(n,)D,D^{\prime}\in\mathrm{Sym}^{(\text{reg})}(n,{\mathbb{R}}), we put W:=(DD)𝔰𝔭(n,)W:=\begin{pmatrix}&D^{\prime}\\ D&\end{pmatrix}\in\mathfrak{s}\mathfrak{p}(n,{\mathbb{R}}). We define a subalgebra 𝔤D,D𝔰𝔭(n,)𝔥n\mathfrak{g}_{D,D^{\prime}}\subset\mathfrak{s}\mathfrak{p}(n,{\mathbb{R}})\ltimes\mathfrak{h}_{n} by 𝔤D,D:=𝔤W\mathfrak{g}_{D,D^{\prime}}:=\mathfrak{g}_{W}.

Lemma 4.5.

For D,DSym(reg)(n,)D,D^{\prime}\in\mathrm{Sym}^{(\text{reg})}(n,{\mathbb{R}}), we have the following statements.

  1. (1)

    The Lie algebra 𝔤D,D\mathfrak{g}_{D,D^{\prime}} is solvable and dim𝔤D,D=2n+2\dim\mathfrak{g}_{D,D^{\prime}}=2n+2.

  2. (2)

    The eigenvalues of WW are square roots of the eigenvalues of the product DDDD^{\prime}.

  3. (3)

    The Lie algebra 𝔤D,D\mathfrak{g}_{D,D^{\prime}} is completely solvable if and only if all the eigenvalues of the product DDDD^{\prime} are positive real numbers.

Proof..

Since the statements (1) and (2) are clear, we show the statement (3). In general, a Lie algebra over {\mathbb{R}} is completely solvable if and only if all the eigenvalues of the adjoint representations are real. In our case, since 𝔥n\mathfrak{h}_{n} is a nilpotent ideal, it is equivalent to the eigenvalues of ad(W+U){\rm ad}(W+U) on 𝔤D,D\mathfrak{g}_{D,D^{\prime}} are real for any U𝔥nU\in\mathfrak{h}_{n}. By an easy calculation, with respect to the base (W,X1,,Xn,Y1,,Yn,Z)(W,X_{1},\cdots,X_{n},Y_{1},\cdots,Y_{n},Z), we have:

ad(W+U)=(0DD00).{\rm ad}(W+U)=\begin{pmatrix}0&&&\\ *&&D^{\prime}&\\ *&D&&\\ 0&*&*&0\end{pmatrix}.

Hence, 𝔤D,D\mathfrak{g}_{D,D^{\prime}} is completely solvable if and only if all the eigenvalues of W=(0DD0)W=\begin{pmatrix}0&D^{\prime}\\ D&0\end{pmatrix} are real. By the statement (2), it is equivalent to the condition that all the eigenvalues of the product DDDD^{\prime} are positive. ∎

Proposition and Definition 4.6.

For D,DSym(reg)(n,)D,D^{\prime}\in\mathrm{Sym}^{(\text{reg})}(n,{\mathbb{R}}), set a subalgebra 𝔥Y1,,Yn\mathfrak{h}\coloneqq\left\langle Y_{1},\cdots,Y_{n}\right\rangle_{{\mathbb{R}}} and a subspace 𝔮WX1,,Xn𝔷𝔤D,D\mathfrak{q}\coloneqq{\mathbb{R}}W\oplus\left\langle X_{1},\cdots,X_{n}\right\rangle_{{\mathbb{R}}}\oplus\mathfrak{z}\subset\mathfrak{g}_{D,D^{\prime}}. Let gg be the inner product on 𝔮𝔤D,D\mathfrak{q}\subset\mathfrak{g}_{D,D^{\prime}} defined by the following Gram matrix with respect to the basis (W,X1,,Xn,Z)(W,X_{1},\cdots,X_{n},Z):

g=(01D110).g=\begin{pmatrix}0&&-1\\ &D^{\prime^{-1}}&\\ -1&&0\end{pmatrix}.

Then we have:

  1. (1)

    𝔤D,D=𝔮𝔥\mathfrak{g}_{D,D^{\prime}}=\mathfrak{q}\oplus\mathfrak{h},

  2. (2)

    [𝔮,𝔥]𝔮[\mathfrak{q},\mathfrak{h}]\subset\mathfrak{q} and [𝔮,𝔮]=𝔥[\mathfrak{q},\mathfrak{q}]=\mathfrak{h},

  3. (3)

    gg is 𝔥\mathfrak{h}-invariant.

Especially, by Note 2.7, we construct a symmetric triple 𝔱D,D:=(𝔤D,D,σ,g)\mathfrak{t}_{D,D^{\prime}}:=(\mathfrak{g}_{D,D^{\prime}},\sigma,g) with signature (p+1,q+1)(p+1,q+1), where (p,q)(p,q) is the signature of DD^{\prime}.

Proof..

The statements (1) and (2) [𝔮,𝔥]𝔮[\mathfrak{q},\mathfrak{h}]\subset\mathfrak{q} are clear. Since DD is invertible, we have [𝔮,𝔮]=𝔥[\mathfrak{q},\mathfrak{q}]=\mathfrak{h}. Then we show the statement (3), namely,

g((adv)X,Y)+g(X,(adv)Y)=0(X,Y𝔮,v𝔥).g(({\rm ad}v)X,Y)+g(X,({\rm ad}v)Y)=0\quad(\forall X,Y\in\mathfrak{q},\forall v\in\mathfrak{h}).

Regard gg as the representation matrix of the inner product on 𝔮=WX1,,Xn𝔷\mathfrak{q}={\mathbb{R}}W\oplus\left\langle X_{1},\cdots,X_{n}\right\rangle_{{\mathbb{R}}}\oplus\mathfrak{z} and let AM(2n+2,)A\in M(2n+2,{\mathbb{R}}) denote the representation matrix of the linear translation advEnd(𝔮){\rm ad}v\in\operatorname{End}(\mathfrak{q}). Then this condition is equivalent to the condition gA+(gA)T=0gA+(gA)^{T}=0, namely, gAgA is skew symmetric. By an easy calculation, we have:

A=(DvvT),gA=(vTv).A=\begin{pmatrix}&&\\ -D^{\prime}v&&\\ &-v^{T}&\quad\\ \end{pmatrix},\quad gA=\begin{pmatrix}&v^{T}&\\ -v&&\\ &&\end{pmatrix}.

We denote by GD,DG_{D,D^{\prime}} the 1-connected Lie group with the Lie algebra 𝔤D,D\mathfrak{g}_{D,D^{\prime}} and by HGD,DH\subset G_{D,D^{\prime}} the analytic subgroup with respect to 𝔥\mathfrak{h}. By Fact 2.9, GD,DG_{D,D^{\prime}} is the transvection group of GD,D/HG_{D,D^{\prime}}/H.

Proposition 4.7.

For any D,DSym(reg)(n,)D,D^{\prime}\in\mathrm{Sym}^{(\text{reg})}(n,{\mathbb{R}}), the symmetric triple 𝔱D,D\mathfrak{t}_{D,D^{\prime}} is reducible and indecomposable.

Proof..

Since the subspace 𝔷𝔮\mathfrak{z}\subset\mathfrak{q} is ad𝔥{\rm ad}\mathfrak{h}-invariant, the symmetric triple 𝔱D,D\mathfrak{t}_{D,D^{\prime}} is reducible. Then we show the indecomposability. Let 𝔤D,D=𝔤1𝔤2\mathfrak{g}_{D,D^{\prime}}=\mathfrak{g}_{1}\oplus\mathfrak{g}_{2} be a non-trivial decomposition. Then we have a 𝔥\mathfrak{h}-invariant decomposition 𝔮=𝔮1𝔮2\mathfrak{q}=\mathfrak{q}_{1}\oplus\mathfrak{q}_{2}. By [𝔮i,𝔮i]=𝔥i[\mathfrak{q}_{i},\mathfrak{q}_{i}]=\mathfrak{h}_{i} for i=1,2i=1,2, the subspace 𝔮1\mathfrak{q}_{1} is non-trivial. Then we are enough to show that 𝔮1\mathfrak{q}_{1} is degenerate. By Note 4.8 below, it is enough to show that 𝔷𝔮1𝒳𝔷\mathfrak{z}\subset\mathfrak{q}_{1}\subset\mathcal{X}\oplus\mathfrak{z}, where 𝒳:=X1,,Xn\mathcal{X}:=\left\langle X_{1},\cdots,X_{n}\right\rangle_{{\mathbb{R}}}. For the decomposition 𝔮=W𝒳𝔷\mathfrak{q}={\mathbb{R}}W\oplus\mathcal{X}\oplus\mathfrak{z}, we set the projections pr1:𝔮W,pr2:𝔮𝒳\operatorname{pr}_{1}:\mathfrak{q}\to{\mathbb{R}}W,\ \operatorname{pr}_{2}:\mathfrak{q}\to\mathcal{X} and pr3:𝔮𝔷\operatorname{pr}_{3}:\mathfrak{q}\to\mathfrak{z}, respectively. First, we show 𝔮1𝒳𝔷\mathfrak{q}_{1}\subset\mathcal{X}\oplus\mathfrak{z}. Let v𝔮1v\in\mathfrak{q}_{1} and assume pr1(v)0\operatorname{pr}_{1}(v)\neq 0. Since DD is invertible, we obtain 𝒳(ad𝔥)v+𝔷\mathcal{X}\subset({\rm ad}\mathfrak{h})v+\mathfrak{z} and 𝔷(ad𝔥)2v\mathfrak{z}\subset({\rm ad}\mathfrak{h})^{2}v, and so we have 𝔮1=𝔮\mathfrak{q}_{1}=\mathfrak{q}, which contradicts the non-triviality of 𝔮1\mathfrak{q}_{1}. Therefore, pr1(v)=0\operatorname{pr}_{1}(v)=0 for any v𝔮1v\in\mathfrak{q}_{1} and so we have 𝔮1𝒳𝔷\mathfrak{q}_{1}\subset\mathcal{X}\oplus\mathfrak{z}. Next, we show 𝔷𝔮1\mathfrak{z}\subset\mathfrak{q}_{1}. If pr2(v)=0\operatorname{pr}_{2}(v)=0 for any v𝔮1v\in\mathfrak{q}_{1}, we have 𝔷𝔮1\mathfrak{z}\subset\mathfrak{q}_{1}, so we take v𝔮1v\in\mathfrak{q}_{1} satisfying pr2(v)0\operatorname{pr}_{2}(v)\neq 0. Then we have 𝔷(ad𝔥)v\mathfrak{z}\subset({\rm ad}\mathfrak{h})v, and so 𝔷𝔮1\mathfrak{z}\subset\mathfrak{q}_{1}. ∎

Note 4.8.

Put 𝔷:={x𝔤D,D|g(x,𝔷)={0}}\mathfrak{z}^{\perp}:=\left\{x\in\mathfrak{g}_{D,D^{\prime}}\ \left|\vphantom{x\in\mathfrak{g}_{D,D^{\prime}}}\ g(x,\mathfrak{z})=\{0\}\right.\right\}, then we have 𝔷=X1,,Xn𝔷\mathfrak{z}^{\perp}=\left\langle X_{1},\cdots,X_{n}\right\rangle_{{\mathbb{R}}}\oplus\mathfrak{z}.

4.2 Isomorphic classes of the triples 𝔱D,D\mathfrak{t}_{D,D^{\prime}}

In this subsection, we see an isomorphic classes of the triple 𝔱D,D\mathfrak{t}_{D,D^{\prime}} (Proposition 4.9) and give the classification for triples with signature (2,2)(2,2) (Proposition 6.13).

Proposition 4.9.

For D1,D2,D1,D2Sym(reg)(n,)D_{1},D_{2},D_{1}^{\prime},D_{2}^{\prime}\in\mathrm{Sym}^{(\text{reg})}(n,{\mathbb{R}}), two symmetric triples 𝔱D1,D1\mathfrak{t}_{D_{1},D^{\prime}_{1}} and 𝔱D2,D2\mathfrak{t}_{D_{2},D^{\prime}_{2}} are isomorphic if and only if there exists (P,k)GL(n,)×>0(P,k)\in GL(n,{\mathbb{R}})\times{\mathbb{R}}_{>0} satisfying:

PD1PT=D2andkPTD2P=D1.PD_{1}^{\prime}P^{T}=D_{2}^{\prime}\ {\rm and}\ kP^{T}D_{2}P=D_{1}.
Proof..

Take an isometric Lie algebra homomorphism ϕ:𝔤D1,D1𝔤D2,D2\phi:\mathfrak{g}_{D_{1},D_{1}^{\prime}}\to\mathfrak{g}_{D_{2},D_{2}^{\prime}}, which is compatible with the involutions. Then ϕ\phi preserves:

  • the decomposition 𝔥𝔮\mathfrak{h}\oplus\mathfrak{q},

  • the center 𝔷\mathfrak{z},

  • its orthogonal subspace 𝔷\mathfrak{z}^{\perp}.

Therefore, with a basis (W,X1,,Xn,Y1,,Yn,Z)(W,X_{1},\cdots,X_{n},Y_{1},\cdots,Y_{n},Z), the map ϕ\phi is written of the form:

ϕ=(1PQ2),\displaystyle\phi=\begin{pmatrix}\ell_{1}&&&\\ *&P&&\\ &&Q&\\ *&*&&\ell_{2}\end{pmatrix},

where 1,2×\ell_{1},\ell_{2}\in{\mathbb{R}}^{\times} and P,QGL(n,)P,Q\in GL(n,{\mathbb{R}}). Since ϕ\phi preserves the inner products, we have:

g(ϕ(Xi),ϕ(Xj))\displaystyle g(\phi(X_{i}),\phi(X_{j})) =g(Xi,Xj)(i,j=1,2,,n)\displaystyle=g(X_{i},X_{j})\quad(i,j=1,2,\cdots,n) PTD21P=D11,\displaystyle\iff P^{T}{D_{2}^{\prime}}^{-1}P={D_{1}^{\prime}}^{-1},
g(ϕ(W),ϕ(Z))\displaystyle g(\phi(W),\phi(Z)) =g(W,Z)\displaystyle=g(W,Z) 12=1.\displaystyle\iff-\ell_{1}\ell_{2}=-1.

Since ϕ\phi is a Lie algebra homomorphism, we have

ϕ([Xi,Yj])\displaystyle\phi([X_{i},Y_{j}]) =[ϕ(Xi),ϕ(Yj)](i,j=1,2,,n)\displaystyle=[\phi(X_{i}),\phi(Y_{j})]\quad(i,j=1,2,\cdots,n) 2In=PQT,\displaystyle\iff\ell_{2}I_{n}=PQ^{T},
ϕ([W,Xi])\displaystyle\phi([W,X_{i}]) =[ϕ(W),ϕ(Xi)](i=1,2,,n)\displaystyle=[\phi(W),\phi(X_{i})]\quad(i=1,2,\cdots,n) QD1=1D2P.\displaystyle\iff QD_{1}=\ell_{1}D_{2}P.

Set k:=12k:=\ell_{1}^{2}, then the conditions PD1PT=D2PD_{1}^{\prime}P^{T}=D_{2}^{\prime} and kPTD2P=D1kP^{T}D_{2}P=D_{1} follow from the above calculations. Conversely, if there exists (P,k)GL(n,)×>0(P,k)\in GL(n,{\mathbb{R}})\times{\mathbb{R}}_{>0} satisfying the condition, a direct calculation leads us that the homomorphism ϕ\phi obtained by putting =0*=0 in the matrix representation above is an isomorphism of symmetric triples. ∎

Definition 4.10.

We denote by (D1,D1)(D2,D2)(D_{1},D^{\prime}_{1})\sim(D_{2},D^{\prime}_{2}) the condition of Proposition 4.9. It is easy to check that \sim is an equivalence relation on Sym(reg)(n,)×Sym(reg)(n,)\mathrm{Sym}^{(\text{reg})}(n,{\mathbb{R}})\times\mathrm{Sym}^{(\text{reg})}(n,{\mathbb{R}}).

In the rest of this subsection, we see the isomorphic classes of the triples 𝔱D,D\mathfrak{t}_{D,D^{\prime}} with signature (2,2)(2,2) with respect to this equivalence relation.

Proposition 4.11.

For matrices D,DSym(reg)(2,)D,D^{\prime}\in\mathrm{Sym}^{(\text{reg})}(2,{\mathbb{R}}), assume the signature of DD^{\prime} is (1,1)(1,1). Then the following list gives a complete class representatives of symmetric triple (𝔤D,D,σ,g)(\mathfrak{g}_{D,D^{\prime}},\sigma,g).

  1. (1)

    (D,D)=(±diag(1,ν),diag(1,ν))(ν>0)(D,D^{\prime})=(\pm\operatorname{diag}\left(1,\nu\right),\operatorname{diag}\left(1,-\nu\right))\quad(\nu>0),
    (D,D)=(±diag(1,ν),diag(1,ν))(ν>0,ν1)(D,D^{\prime})=(\pm\operatorname{diag}\left(1,-\nu\right),\operatorname{diag}\left(1,-\nu\right))\quad(\nu>0,\ \nu\neq 1),

  2. (2)

    (D,D)=(Qν,Qν)(ν>0)(D,D^{\prime})=\left(Q_{\nu},Q_{-\nu}\right)\quad(\nu>0) (see Notation 2.1),

  3. (3)

    (D,D)=((±1110),(011±1)),((±1110),(0111))(D,D^{\prime})=\left(\begin{pmatrix}\pm 1&-1\\ -1&0\end{pmatrix},\begin{pmatrix}0&-1\\ -1&\pm 1\end{pmatrix}\right),\ \left(\begin{pmatrix}\pm 1&-1\\ -1&0\end{pmatrix},\begin{pmatrix}0&1\\ 1&\mp 1\end{pmatrix}\right),

  4. (4)

    (D,D)=(±I1,1,I1,1)(D,D^{\prime})=(\pm I_{1,1},I_{1,1}).

Proof of Proposition 6.13.

By Proposition 4.9, we may and do assume D=I1,1D^{\prime}=I_{1,1}. For a basis (I2,Q0,I1,1)(I_{2},Q_{0},I_{1,1}) of Sym(2,)\mathrm{Sym}(2,{\mathbb{R}}), we put D(x,y,z):=xI2+yQ0+zI1,1Sym(2,)D(x,y,z):=xI_{2}+yQ_{0}+zI_{1,1}\in\mathrm{Sym}(2,{\mathbb{R}}). Note that detD(x,y,z)=x2y2z2\det D(x,y,z)=x^{2}-y^{2}-z^{2}. Then we are enough to consider the orbit of D(x,y,z)D(x,y,z) with respect to the action O(1,1)×>0Isom(Sym(reg)(2,)),(P,k)(DkPDPT)O(1,1)\times{\mathbb{R}}_{>0}\to{\rm Isom}(\mathrm{Sym}^{(\text{reg})}(2,{\mathbb{R}})),\ (P,k)\mapsto(D\mapsto kPDP^{T}). Note that:

O(1,1)\displaystyle O(1,1) =H(t):=(coshtsinhtsinhtcosht),I1,1,±I2t,\displaystyle=\Big{\langle}H(t):=\begin{pmatrix}\cosh t&\sinh t\\ \sinh t&\cosh t\end{pmatrix},\ I_{1,1},\ \pm I_{2}\Big{\rangle}_{t\in{\mathbb{R}}},
H(t)D(x,y,z)H(t)T\displaystyle H(t)D(x,y,z)H(t)^{T} =D(xcosh2t+ysinh2t,xsinh2t+ycosh2t,z),\displaystyle=D(x\cosh 2t+y\sinh 2t,x\sinh 2t+y\cosh 2t,z),
I1,1D(x,y,z)I1,1T\displaystyle I_{1,1}D(x,y,z)I_{1,1}^{T} =D(x,y,z).\displaystyle=D(x,-y,z).

Set G:=diag(H(2t),1),diag(1,±1,1),kI3t,k>0G:=\langle{\operatorname{diag}\left(H(2t),1\right),\operatorname{diag}\left(1,\pm 1,1\right),kI_{3}}\rangle_{t\in{\mathbb{R}},k\in{\mathbb{R}}_{>0}}. Then we are enough to consider the orbit space of the GG-action on {(x,y,z)3|x2y2z20}3{0}\left\{(x,y,z)\in{\mathbb{R}}^{3}\ \left|\vphantom{(x,y,z)\in{\mathbb{R}}^{3}}\ x^{2}-y^{2}-z^{2}\neq 0\right.\right\}\subset{\mathbb{R}}^{3}-\{0\}. By an easy calculation, we have:

Note 4.12.

The orbit space of the GG-action on 3{0}{\mathbb{R}}^{3}-\{0\} is:

{[(±1,0,z)],[(0,1,z)],[(1,1,±1)],[(1,1,±1)],[(0,0,±1)]}z.\{[(\pm 1,0,z)],[(0,1,z)],[(1,1,\pm 1)],[(-1,1,\pm 1)],[(0,0,\pm 1)]\}_{z\in{\mathbb{R}}}.
  1. (1)

    The orbits [(±1,0,z)][(\pm 1,0,z)] (z,z±1z\in{\mathbb{R}},\ z\neq\pm 1).
    Here, the constraint z±1z\neq\pm 1 comes from the condition x2y2z20x^{2}-y^{2}-z^{2}\neq 0. We have (D,D)(diag(1+z,1z),I1,1),(diag(1+z,1z),I1,1)(D,D^{\prime})\sim(\operatorname{diag}\left(1+z,1-z\right),I_{1,1}),\ (\operatorname{diag}\left(-1+z,-1-z\right),I_{1,1}) and:

    (diag(1+z,1z),I1,1)\displaystyle(\operatorname{diag}\left(1+z,1-z\right),I_{1,1}) {(diag(1,μ),I1,1)(1+z>0,μ>1,μ0)(diag(1,μ),I1,1)(1+z<0,μ>1),\displaystyle\sim\begin{cases}(\operatorname{diag}\left(1,\mu\right),I_{1,1})\quad(1+z>0,\mu>-1,\mu\neq 0)\\ (\operatorname{diag}\left(-1,\mu\right),I_{1,1})\quad(1+z<0,\mu>1)\end{cases},
    (diag(1+z,1z),I1,1)\displaystyle(\operatorname{diag}\left(-1+z,-1-z\right),I_{1,1}) {(diag(1,μ),I1,1)(1+z>0,μ<1)(diag(1,μ),I1,1)(1+z<0,μ<1,μ0).\displaystyle\sim\begin{cases}(\operatorname{diag}\left(1,\mu\right),I_{1,1})\quad(-1+z>0,\mu<-1)\\ (\operatorname{diag}\left(-1,\mu\right),I_{1,1})\quad(-1+z<0,\mu<1,\mu\neq 0)\end{cases}.

    Then we get (D,D)(±diag(1,μ),I1,1)(D,D^{\prime})\sim(\pm\operatorname{diag}\left(1,\mu\right),I_{1,1}) for some μ×\mu\in{\mathbb{R}}^{\times} (μ1\mu\neq-1). Putting ν:=|μ|\nu:=\sqrt{|\mu|} and P:=diag(1,ν)P:=\operatorname{diag}\left(1,\sqrt{\nu}\right), we obtain (diag(±1,μ),I1,1)(±diag(1,ν),diag(1,ν))(\operatorname{diag}\left(\pm 1,\mu\right),I_{1,1})\sim(\pm\operatorname{diag}\left(1,\nu\right),\operatorname{diag}\left(1,-\nu\right)) or (±diag(1,ν),diag(1,ν)),ν1(\pm\operatorname{diag}\left(1,-\nu\right),\operatorname{diag}\left(1,-\nu\right)),\nu\neq 1.

  2. (2)

    The orbits [(0,1,z)][(0,1,z)] (zz\in{\mathbb{R}}).
    In this case, we have:

    (D,D)((z11z),I1,1)((1ν2νν1+ν2),I1,1)(D,D^{\prime})\sim\left(\begin{pmatrix}z&1\\ 1&-z\end{pmatrix},I_{1,1}\right)\sim\left(\begin{pmatrix}1-\nu^{2}&\nu\\ \nu&-1+\nu^{2}\end{pmatrix},I_{1,1}\right)

    for some ν>0\nu\in{\mathbb{R}}_{>0}. Take tt\in{\mathbb{R}} satisfying sinht=ν\sinh t=\nu and put:

    P:=et/22(et11et),P:=\dfrac{e^{-t/2}}{\sqrt{2}}\begin{pmatrix}e^{t}&1\\ 1&-e^{t}\end{pmatrix},

    then we have the equivalence (D,D)(Qν,Qν)(D,D^{\prime})\sim\left(Q_{\nu},Q_{-\nu}\right).

  3. (3)

    The orbits [(1,1,±1)],[(1,1,±1)][(1,1,\pm 1)],[(-1,1,\pm 1)].
    In this case, we have:

    (D,D)((0112),I1,1),((2110),I1,1),((2110),I1,1),((0112),I1,1).\displaystyle(D,D^{\prime})\sim\left(\begin{pmatrix}0&1\\ 1&2\end{pmatrix},I_{1,1}\right),\left(\begin{pmatrix}-2&1\\ 1&0\end{pmatrix},I_{1,1}\right),\left(\begin{pmatrix}2&1\\ 1&0\end{pmatrix},I_{1,1}\right),\left(\begin{pmatrix}0&1\\ 1&-2\end{pmatrix},I_{1,1}\right).

    By a direct calculation, we have the equivalence to the following classes, respectively. (For example, we use P:=122(2213)P:=\dfrac{1}{2\sqrt{2}}\begin{pmatrix}2&2\\ 1&-3\end{pmatrix} for the first pair.)

    ((1110),(0111)),((1110),(0111)),\displaystyle\left(\begin{pmatrix}1&-1\\ -1&0\end{pmatrix},\begin{pmatrix}0&1\\ 1&-1\end{pmatrix}\right),\left(\begin{pmatrix}-1&-1\\ -1&0\end{pmatrix},\begin{pmatrix}0&1\\ 1&1\end{pmatrix}\right),
    ((1110),(0111)),((1110),(0111)).\displaystyle\left(\begin{pmatrix}1&-1\\ -1&0\end{pmatrix},\begin{pmatrix}0&-1\\ -1&1\end{pmatrix}\right),\left(\begin{pmatrix}-1&-1\\ -1&0\end{pmatrix},\begin{pmatrix}0&-1\\ -1&-1\end{pmatrix}\right).
  4. (4)

    The orbits [(0,0,±1)][(0,0,\pm 1)].
    In this case, we have (D,D)(±I1,1,I1,1)(D,D^{\prime})\sim\left(\pm I_{1,1},I_{1,1}\right).

Remark 4.13.

With the natural identification (3{0})/>0𝕊2({\mathbb{R}}^{3}-\{0\})/{\mathbb{R}}_{>0}\simeq\mathbb{S}^{2} (the 2-dimensional unit sphere), a picture of the parameter spaces of GD,I1,1G_{D,I_{1,1}} is given as Figure 4.1.

Refer to caption
Figure 4.1: A picture of the parameter space of GD,I1,1G_{D,I_{1,1}}
Remark 4.14.

For a symmetric triple (𝔤D,D,σ,g)(\mathfrak{g}_{D,D^{\prime}},\sigma,g), put α,β\alpha,\beta\in{\mathbb{C}} eigenvalues of the product DDDD^{\prime}. Then the following table shows which class of Proposition 6.13 the symmetric triple (𝔤D,D,σ,g)(\mathfrak{g}_{D,D^{\prime}},\sigma,g) belongs to.

class (1) (2) (3) (4)
eigenvalues real not real real real
relation αβ\alpha\neq\beta αβ\alpha\neq\beta α=β\alpha=\beta α=β\alpha=\beta
DDDD^{\prime} is diagonalizable yes yes no yes

4.3 Definition of symmetric triples (𝔤nil,σ,g±)(\mathfrak{g}_{\rm nil},\sigma,g_{\pm})

In this subsection, we define a symmetric triples (𝔤nil,σ,g±)(\mathfrak{g}_{\rm nil},\sigma,g_{\pm}), which is in the list of the classification (Fact 1.4).

Definition 4.15 (𝔤nil,Gnil/H\mathfrak{g}_{\text{nil}},G_{\text{nil}}/H).

We define a nilpotent Lie algebra
𝔤nilA1,A2,B,C1,C2\mathfrak{g}_{\rm nil}\coloneqq\left\langle A_{1},A_{2},B,C_{1},C_{2}\right\rangle_{{\mathbb{R}}} as follows:

[A1,A2]=B,[B,A1]=C1,[B,A2]=C2,\displaystyle[A_{1},A_{2}]=B,\quad[B,A_{1}]=C_{1},\quad[B,A_{2}]=C_{2},
the other brackets are trivial.

Put 𝔥B\mathfrak{h}\coloneqq{\mathbb{R}}{B} and 𝔮A1,A2,C1,C2\mathfrak{q}\coloneqq\left\langle A_{1},A_{2},C_{1},C_{2}\right\rangle_{{\mathbb{R}}}, then we have 𝔤nil=𝔮𝔥\mathfrak{g}_{\text{nil}}=\mathfrak{q}\oplus\mathfrak{h}, [𝔮,𝔥]𝔮[\mathfrak{q},\mathfrak{h}]\subset\mathfrak{q} and [𝔮,𝔮]=𝔥[\mathfrak{q},\mathfrak{q}]=\mathfrak{h}. We define a 𝔥\mathfrak{h}-invariant inner product gg on 𝔮\mathfrak{q} as follows:

g±(±J±J),g_{\pm}\coloneqq\begin{pmatrix}&\pm J\\ \\ \pm J&\end{pmatrix},

where J:=(11)J:=\begin{pmatrix}&-1\\ 1&\end{pmatrix}. By Note 2.7, the triples (𝔤nil,σ,g±)(\mathfrak{g}_{\text{nil}},\sigma,g_{\pm}) are indecomposable symmetric triples with signature (2,2)(2,2). We denote by GnilG_{\rm nil} the 1-connected nilpotent Lie group with the Lie algebra 𝔤nil\mathfrak{g}_{\rm nil} and by HH the analytic subgroup of GnilG_{\rm nil} with respect to 𝔥\mathfrak{h}. The Lie group GnilG_{\rm nil} is the transvection group of Gnil/HG_{\rm nil}/H (Fact 2.9).

5 Criterions of the existence of compact Clifford–Klein forms for spaces GD,D/HG_{D,D^{\prime}}/H

To prove the main theorem, we prepare some criterions for the existence of compact Clifford–Klein forms of the symmetric space GD,D/HG_{D,D^{\prime}}/H (Propositions 5.25 and 5.26). In this section, we use Notation 4.1 and the following:

Notation 5.1.
  • 𝔥:=Y1,,Yn𝔥n\mathfrak{h}:=\left\langle Y_{1},\cdots,Y_{n}\right\rangle_{{\mathbb{R}}}\subset\mathfrak{h}_{n},

  • D,DSym(reg)(n,)D,D^{\prime}\in\mathrm{Sym}^{(\text{reg})}(n,{\mathbb{R}}),

  • W(DD)M(2n,),Wt:=exptW=(AtBt)GL(2n,)W\coloneqq\begin{pmatrix}&D^{\prime}\\ D&\end{pmatrix}\in M(2n,{\mathbb{R}}),\quad W_{t}:=\exp tW=\begin{pmatrix}A_{t}&B_{t}\\ *&*\end{pmatrix}\in GL(2n,{\mathbb{R}}),

  • pr1:GD,D=Hn the first projection\operatorname{pr}_{1}:G_{D,D^{\prime}}={\mathbb{R}}\ltimes H_{n}\to{\mathbb{R}}\text{ the first projection},

  • pr2:GD,D=HnHn the second projection\operatorname{pr}_{2}:G_{D,D^{\prime}}={\mathbb{R}}\ltimes H_{n}\to H_{n}\text{ the second projection}.

In this section, we identify X1,,Xn,Y1,,Yn\left\langle X_{1},\cdots,X_{n},Y_{1},\cdots,Y_{n}\right\rangle_{{\mathbb{R}}} as 2n{\mathbb{R}}^{2n} and 𝔥n\mathfrak{h}_{n} as 2n+1{\mathbb{R}}^{2n+1}. Especially, we think M(2n,)End(X1,,Xn,Y1,Yn)M(2n,{\mathbb{R}})\simeq\operatorname{End}({\left\langle X_{1},\cdots,X_{n},Y_{1}\cdots,Y_{n}\right\rangle_{{\mathbb{R}}}}) and M(2n+1,)End(X1,,Xn,Y1,Yn,Z)M(2n+1,{\mathbb{R}})\simeq\operatorname{End}({\left\langle X_{1},\cdots,X_{n},Y_{1}\cdots,Y_{n},Z\right\rangle_{{\mathbb{R}}}}).

Note 5.2.

By a direct calculation, we have:

Adg|𝔥n=(Wpr1(g)1),(gGD,D).{\rm Ad}_{g}|_{\mathfrak{h}_{n}}=\begin{pmatrix}W_{\operatorname{pr}_{1}(g)}&\\ *&1\end{pmatrix},\quad(g\in G_{D,D^{\prime}}).

5.1 Subgroups LCL_{C} and LC,wL_{C,w}

In Subsection 5.3, we see criterions for the existence of compact Clifford–Klein forms. Before that, we show two basic Propositions 5.12 and 5.14. In Proposition 5.14, we classify constructors of GD,D/HG_{D,D^{\prime}}/H. To do this, we introduce some subgroups of GD,DG_{D,D^{\prime}} and show their basic properties.

Definition 5.3.

For CM(n,)C\in M(n,{\mathbb{R}}) and w𝔥w\in\mathfrak{h}, we denote by 𝔩C2n\mathfrak{l}^{\prime}_{C}\subset{\mathbb{R}}^{2n} the image of the linear transform defined by (In0C0)\begin{pmatrix}I_{n}&0\\ C&0\end{pmatrix}. Then we put:

𝔩C\displaystyle\mathfrak{l}_{C} :=𝔩C𝔷subalgebra𝔥n,\displaystyle:=\mathfrak{l}_{C}^{\prime}\oplus\mathfrak{z}\underset{\rm subalgebra}{\subset}\mathfrak{h}_{n},
𝔩C,w\displaystyle\mathfrak{l}_{C,w} :=(W+w)𝔩Csubspace𝔤D,D.\displaystyle:={\mathbb{R}}(W+w)\oplus\mathfrak{l}_{C}\underset{\rm subspace}{\subset}\mathfrak{g}_{D,D^{\prime}}.

We denote by LCL_{C} the analytic subgroup in GD,DG_{D,D^{\prime}} with respect to 𝔩C\mathfrak{l}_{C}. If 𝔩C,w\mathfrak{l}_{C,w} is a subalgebra of 𝔤D,D\mathfrak{g}_{D,D^{\prime}}, we denote by LC,wL_{C,w} its analytic subgroup in GD,DG_{D,D^{\prime}}.

We see the criterion of 𝔩C,w\mathfrak{l}_{C,w} to be a subalgebra of 𝔤D,D\mathfrak{g}_{D,D^{\prime}}.

Proposition 5.4.

For CM(n,)C\in M(n,{\mathbb{R}}) and w𝔥w\in\mathfrak{h}, the following conditions are equivalent:

  1. (a)

    the subspace 𝔩C,w𝔤D,D\mathfrak{l}_{C,w}\subset\mathfrak{g}_{D,D^{\prime}} is a subalgebra,

  2. (b)

    [W+w,𝔩C]𝔩C[W+w,\mathfrak{l}_{C}]\subset\mathfrak{l}_{C},

  3. (c)

    the subspace 𝔩C\mathfrak{l}^{\prime}_{C} is WW-invariant,

  4. (d)

    the subspace 𝔩C\mathfrak{l}^{\prime}_{C} is WtW_{t}-invariant (t\forall t\in{\mathbb{R}}),

  5. (e)

    the subalgebra 𝔩C\mathfrak{l}_{C} is AdGD,D{\rm Ad}_{G_{D,D^{\prime}}}-invariant,

  6. (f)

    CDC=DCD^{\prime}C=D.

Remark 5.5.

By this proposition, the conditions (a) and (b) do not depend on w𝔥w\in\mathfrak{h}.

Proof..

Since 𝔩C\mathfrak{l}_{C} is a subalgebra of 𝔥n\mathfrak{h}_{n}, the condition (a) is equivalent to the condition (b). By [w,𝔩C]𝔷[w,\mathfrak{l}_{C}]\subset\mathfrak{z}, we have the equivalence (b)\Leftrightarrow(c) by a direct calculation. The equivalence (c)\Leftrightarrow(d) is easy and we have the implication (d)\Leftrightarrow(e) by Note 5.2. By Note 5.6 below, we show the equivalence (c)\Leftrightarrow(f) as follows.

(c)(C,In)(DD)(InC)=0CDCD=0(f).\displaystyle\text{(c)}\iff(C,-I_{n})\begin{pmatrix}&D^{\prime}\\ D&\end{pmatrix}\begin{pmatrix}I_{n}\\ C\end{pmatrix}=0\iff CD^{\prime}C-D=0\iff\text{(f)}.

Note 5.6.

For CM(n,)C\in M(n,{\mathbb{R}}) and AM(2n,)A\in M(2n,{\mathbb{R}}), the subspace Im(InC)2n{\rm Im}\begin{pmatrix}I_{n}\\ C\end{pmatrix}\subset{\mathbb{R}}^{2n} is AA-invariant if and only if (C,In)A(InC)=O(C,\ -I_{n})A\begin{pmatrix}I_{n}\\ C\end{pmatrix}=O.

We give fundamental properties of the subspaces 𝔩C\mathfrak{l}_{C} and 𝔩C,w\mathfrak{l}_{C,w}.

Note 5.7.

There is a Lie algebra isomorphism 𝔩C𝔥kn2k\mathfrak{l}_{C}\simeq\mathfrak{h}_{k}\oplus{\mathbb{R}}^{n-2k}, where k(rank(CCT))/2k\coloneqq(\operatorname{rank}(C-C^{T}))/2.

Note 5.8.

For CM(n,)C\in M(n,{\mathbb{R}}) and w𝔥w\in\mathfrak{h}, we have decompositions 2n=𝔩C𝔥{\mathbb{R}}^{2n}=\mathfrak{l}_{C}^{\prime}\oplus\mathfrak{h}, 𝔥n=𝔩C𝔥\mathfrak{h}_{n}=\mathfrak{l}_{C}\oplus\mathfrak{h} and 𝔤D,D=𝔩C,w𝔥\mathfrak{g}_{D,D^{\prime}}=\mathfrak{l}_{C,w}\oplus\mathfrak{h} as linear spaces.

The “converse of Note 5.8” also holds. In fact, we have:

Proposition 5.9.
  1. (1)

    For any subspace 𝔩2n\mathfrak{l}\subset{\mathbb{R}}^{2n} satisfying 2n=𝔩𝔥{\mathbb{R}}^{2n}=\mathfrak{l}\oplus\mathfrak{h}, there exists CM(n,)C\in M(n,{\mathbb{R}}) satisfying 𝔩=𝔩C\mathfrak{l}=\mathfrak{l}^{\prime}_{C}.

  2. (2)

    For any subalgebra 𝔩𝔥n\mathfrak{l}\subset\mathfrak{h}_{n} satisfying 𝔥n=𝔩𝔥\mathfrak{h}_{n}=\mathfrak{l}\oplus\mathfrak{h}, there exists CM(n,)C\in M(n,{\mathbb{R}}) satisfying 𝔩=𝔩C\mathfrak{l}=\mathfrak{l}_{C}.

  3. (3)

    For any subalgebra 𝔩𝔤D,D\mathfrak{l}\subset\mathfrak{g}_{D,D^{\prime}} satisfying 𝔤D,D=𝔩𝔥\mathfrak{g}_{D,D^{\prime}}=\mathfrak{l}\oplus\mathfrak{h}, there exist CM(n,)C\in M(n,{\mathbb{R}}) and w𝔥w\in\mathfrak{h} satisfying CDC=DCD^{\prime}C=D and 𝔩=𝔩C,w\mathfrak{l}=\mathfrak{l}_{C,w}.

To prove this proposition, we use the following fact and note.

Fact 5.10 ([16, Lemma 4.3]).

If a subalgebra 𝔩𝔥n\mathfrak{l}\subset\mathfrak{h}_{n} satisfies 𝔥n=𝔩𝔥\mathfrak{h}_{n}=\mathfrak{l}\oplus\mathfrak{h} as a linear space, we have 𝔷𝔩\mathfrak{z}\subset\mathfrak{l}.

Note 5.11.

Let V=UWV=U\oplus W be a linear space decomposition. For a subspace V1VV_{1}\subset V satisfying WV1W\subset V_{1}, we have V1=(V1U)WV_{1}=(V_{1}\cap U)\oplus W.

Proof of Proposition 5.9.
  1. (1)

    Let 𝔩2n\mathfrak{l}\subset{\mathbb{R}}^{2n} be a subspace satisfying 2n=𝔩𝔥{\mathbb{R}}^{2n}=\mathfrak{l}\oplus\mathfrak{h}. Then there exists CM(n,)C\in M(n,{\mathbb{R}}) such that:

    𝔩=Im(I0C0)=𝔩C.\mathfrak{l}={\rm Im}\begin{pmatrix}I&0\\ C&0\end{pmatrix}=\mathfrak{l}^{\prime}_{C}.
  2. (2)

    Let 𝔩𝔥n\mathfrak{l}\subset\mathfrak{h}_{n} be a subalgebra satisfying 𝔥n=𝔩𝔥\mathfrak{h}_{n}=\mathfrak{l}\oplus\mathfrak{h}. By using Note 5.11 for the decomposition 𝔥n=𝔩𝔥\mathfrak{h}_{n}=\mathfrak{l}\oplus\mathfrak{h} and the subspace 2n𝔥n{\mathbb{R}}^{2n}\subset\mathfrak{h}_{n}, we have 2n=(2n𝔩)𝔥{\mathbb{R}}^{2n}=({\mathbb{R}}^{2n}\cap\mathfrak{l})\oplus\mathfrak{h}. By the statement (1), there exists CM(n,)C\in M(n,{\mathbb{R}}) satisfying 2n𝔩=𝔩C{\mathbb{R}}^{2n}\cap\mathfrak{l}=\mathfrak{l}^{\prime}_{C}. By Fact 5.10, we get 𝔷𝔩\mathfrak{z}\subset\mathfrak{l}. Then by using Note 5.11 again for the decomposition 𝔥n=2n𝔷\mathfrak{h}_{n}={\mathbb{R}}^{2n}\oplus\mathfrak{z} and the subspace 𝔩𝔥n\mathfrak{l}\subset\mathfrak{h}_{n}, we have 𝔩=(𝔩2n)𝔷=𝔩C𝔷=𝔩C\mathfrak{l}=(\mathfrak{l}\cap{\mathbb{R}}^{2n})\oplus\mathfrak{z}=\mathfrak{l}^{\prime}_{C}\oplus\mathfrak{z}=\mathfrak{l}_{C}.

  3. (3)

    Let 𝔩𝔤D,D\mathfrak{l}\subset\mathfrak{g}_{D,D^{\prime}} be a subalgebra satisfying 𝔤D,D=𝔩𝔥\mathfrak{g}_{D,D^{\prime}}=\mathfrak{l}\oplus\mathfrak{h}. By using Note 5.11 for the decomposition 𝔤D,D=𝔩𝔥\mathfrak{g}_{D,D^{\prime}}=\mathfrak{l}\oplus\mathfrak{h} and the subspace 𝔥n𝔤D,D\mathfrak{h}_{n}\subset\mathfrak{g}_{D,D^{\prime}}, we get 𝔥n=(𝔥n𝔩)𝔥\mathfrak{h}_{n}=(\mathfrak{h}_{n}\cap\mathfrak{l})\oplus\mathfrak{h}. By the statement (2), we have 𝔥n𝔩=𝔩C\mathfrak{h}_{n}\cap\mathfrak{l}=\mathfrak{l}_{C} for some CM(n,)C\in M(n,{\mathbb{R}}). Then we have 𝔩=(W+w)𝔩C=𝔩C,w\mathfrak{l}={\mathbb{R}}(W+w)\oplus\mathfrak{l}_{C}=\mathfrak{l}_{C,w} for some w𝔥w\in\mathfrak{h}. Since 𝔩C,w\mathfrak{l}_{C,w} is subalgebra, we obtain CDC=DCD^{\prime}C=D (Proposition 5.4).

Proposition 5.12.

For CM(n,)C\in M(n,{\mathbb{R}}), the following conditions are equivalent:

  1. (a)

    the pair (LC,H)(L_{C},H) satisfies the property (CI) in GD,DG_{D,D^{\prime}},

  2. (b)

    AdGD,D𝔩C𝔥={0}{\rm Ad}_{G_{D,D^{\prime}}}\mathfrak{l}_{C}\cap\mathfrak{h}=\{0\},

  3. (c)

    Wt𝔩C𝔥={0}W_{t}\mathfrak{l}^{\prime}_{C}\cap\mathfrak{h}=\{0\}  (t\forall t\in{\mathbb{R}}),

  4. (d)

    the matrix At+BtCA_{t}+B_{t}C is invertible  (t\forall t\in{\mathbb{R}}).

Proof..

The equivalence (a)\Leftrightarrow(b) comes from Note 3.2. We have Adg𝔩C=Wpr1(g)𝔩C𝔷{\rm Ad}_{g}\mathfrak{l}_{C}=W_{\operatorname{pr}_{1}(g)}\mathfrak{l}^{\prime}_{C}\oplus\mathfrak{z} for gGD,Dg\in G_{D,D^{\prime}} by Note 5.2, so the equivalence (b)\Leftrightarrow(c) holds.

Then we show the equivalence (c)\Leftrightarrow(d). Take any tt\in{\mathbb{R}}. Since Wt𝔩CW_{t}\mathfrak{l}_{C}^{\prime} is the image of the linear map (AtBt)(In0C0)\begin{pmatrix}A_{t}&B_{t}\\ *&*\end{pmatrix}\begin{pmatrix}I_{n}&0\\ C&0\end{pmatrix}, we have:

Wt𝔩C𝔥={0}det(At+BtC)0.W_{t}\mathfrak{l}^{\prime}_{C}\cap\mathfrak{h}=\{0\}\Leftrightarrow\det(A_{t}+B_{t}C)\neq 0.

Then we get:

Lemma 5.13.

Suppose CM(n,)C\in M(n,{\mathbb{R}}) satisfies CDC=DCD^{\prime}C=D, then we have LC,wHL_{C,w}\pitchfork H in GD,DG_{D,D^{\prime}}.

Proof..

By Note 3.2 and Propositions 5.4 and 5.12, the pair (LC,H)(L_{C},H) has the property (CI) in HnH_{n}. By Fact 3.13, we have LCHL_{C}\pitchfork H in HnH_{n}. By the condition gLC=LC\mathcal{I}_{g}L_{C}=L_{C} for any gGD,Dg\in G_{D,D^{\prime}}, we have SLCH\mathcal{I}_{S}L_{C}\pitchfork H in HnH_{n} for any compact set SGD,DS\subset G_{D,D^{\prime}}. By using Lemma 2.31 with (G,N,L)=(GD,D,Hn,L1LC)(G,N,L)=(G_{D,D^{\prime}},H_{n},L_{1}L_{C}), we obtain LC,wHL_{C,w}\pitchfork H in GG, where L1L_{1} is the analytic subgroup of GD,DG_{D,D^{\prime}} with respect to (W+w){\mathbb{R}}(W+w) and the condition L1HnL_{1}\pitchfork H_{n} comes from Lemma 2.32. ∎

Finally, we classify the constructors of GD,D/HG_{D,D^{\prime}}/H, namely, we have:

Proposition 5.14.

For a connected subgroup LGD,DL\subset G_{D,D^{\prime}}, the following conditions are equivalent.

  1. (a)

    The subgroup LL is a constructor of GD,D/HG_{D,D^{\prime}}/H.

  2. (b)

    There exist CM(n,)C\in M(n,{\mathbb{R}}) and w𝔥w\in\mathfrak{h} satisfying CDC=DCD^{\prime}C=D and L=LC,wL=L_{C,w}.

Proof..

First, we show the implication (a)\Rightarrow(b). We denote by 𝔩\mathfrak{l} the Lie algebra of LL. By Proposition 3.7, we have 𝔤D,D=𝔩𝔥\mathfrak{g}_{D,D^{\prime}}=\mathfrak{l}\oplus\mathfrak{h}, and so the condition (b) follows from Proposition 5.9.

Next, we show the implication (b)\Rightarrow(a). Take CM(n,)C\in M(n,{\mathbb{R}}) and w𝔥w\in\mathfrak{h} satisfying CDC=DCD^{\prime}C=D. We are enough to show the properness and cocompactness of the LC,wL_{C,w}-action. By Lemma 5.13, the LC,wL_{C,w}-action on GD,D/HG_{D,D^{\prime}}/H is proper. By Note 5.8, we have 𝔤D,D=𝔩C,w𝔥\mathfrak{g}_{D,D^{\prime}}=\mathfrak{l}_{C,w}\oplus\mathfrak{h}, which implies the cocompactness by Proposition 3.7. ∎

5.2 Uniform lattices of LCL_{C} and LC,wL_{C,w}

In this subsection, we discuss necessary conditions for the existence of a uniform lattice in LCL_{C} and LC,wL_{C,w}, namely, we show the following two propositions.

Proposition 5.15.

Assume there exists GD,DHn\ell\in G_{D,D^{\prime}}-H_{n} such that the subgroup LCL_{C} has an \mathcal{I}_{\ell}-invariant uniform lattice. Then the subspace 𝔩C\mathfrak{l}^{\prime}_{C} is Wt0W_{t_{0}}-invariant and we have det(Wt0|𝔩C)=±1\det(W_{t_{0}}|_{\mathfrak{l}^{\prime}_{C}})=\pm 1, where t0:=pr1()×t_{0}:=\operatorname{pr}_{1}(\ell)\in{\mathbb{R}}^{\times}.

Proposition 5.16.

Suppose CM(n,)C\in M(n,{\mathbb{R}}) and w𝔥w\in\mathfrak{h} satisfy the condition in Proposition 5.4. If the subgroup LC,wL_{C,w} has a uniform lattice, then the condition trDC=0\operatorname{tr}{D^{\prime}C}=0 holds.

To prove these propositions, we define a solvable Lie group.

Definition 5.17.

For MM(m,)M\in M(m,{\mathbb{R}}), we consider the {\mathbb{R}}-action on m{\mathbb{R}}^{m}, ϕ:GL(m,),texptM\phi:{\mathbb{R}}\to GL(m,{\mathbb{R}}),\ t\mapsto\exp tM, and denote by SMS_{M} the semidirect product ϕm{\mathbb{R}}\ltimes_{\phi}{\mathbb{R}}^{m}. We denote by pr:SM\operatorname{pr}_{\mathbb{R}}:S_{M}\to{\mathbb{R}} the first projection. We regard {\mathbb{R}} as a subgroup of SMS_{M} by the injection SM,t(t,0){\mathbb{R}}\to S_{M},\ t\mapsto(t,0).

We see some basic properties of SMS_{M} (Lemma 5.18, Note 5.19 and Proposition 5.20).

Lemma 5.18.

Let MM(m,)M\in M(m,{\mathbb{R}}) and t0t_{0}\in{\mathbb{R}}. If expt0M\exp t_{0}M does not have an eigenvalue 11, then we have 𝒵SM(t0)=\mathcal{Z}_{S_{M}}(t_{0})={\mathbb{R}}.

Proof..

For an element (t1,v1)SM(t_{1},v_{1})\in S_{M}, we have:

(t1,v1)𝒵SM(t0)(t0,0)(t1,v1)=(t1,v1)(expt0M)v1=v1v1=0.(t_{1},v_{1})\in\mathcal{Z}_{S_{M}}(t_{0})\Leftrightarrow\mathcal{I}_{(t_{0},0)}(t_{1},v_{1})=(t_{1},v_{1})\Leftrightarrow(\exp t_{0}M)v_{1}=v_{1}\Leftrightarrow v_{1}=0.

Note 5.19.

Some Lie groups in this paper are isomorphic to SMS_{M}. Let CM(n,)C\in M(n,{\mathbb{R}}) and w𝔥w\in\mathfrak{h} satisfy the condition of Proposition 5.4.

  • If CC is symmetric, LC,wSML_{C,w}\simeq S_{M}, where M:=(DC0wT0)M:=\begin{pmatrix}D^{\prime}C&0\\ -w^{T}&0\end{pmatrix},

  • LC,w/𝒵HnSDCL_{C,w}/\mathcal{Z}_{H_{n}}\simeq S_{D^{\prime}C},

  • GD,D/𝒵HnSWG_{D,D^{\prime}}/\mathcal{Z}_{H_{n}}\simeq S_{W}.

Proposition 5.20.

For MM(m,)M\in M(m,{\mathbb{R}}), we have trM=0\operatorname{tr}M=0 if the group SMS_{M} has a uniform lattice.

To prove this proposition, we prepare some facts and a proposition.

Fact 5.21 ([33, Theorem 3.3]).

Let GG be a connected solvable Lie group and NN its maximum connected (closed) normal nilpotent subgroup. Let HH be a cocompact closed subgroup of GG. Assume that HNH\cap N contains no non-trivial connected (closed) Lie subgroups which are normal in GG. Then N/(HN)N/(H\cap N) is compact.

Fact 5.22 ([11]).

Let GG be a 1-connected nilpotent Lie group and ΓG\Gamma\subset G a uniform lattice. Then GG is commutative if and only if so is Γ\Gamma.

Proposition 5.23.

For MM(m,)M\in M(m,{\mathbb{R}}), we consider the following conditions.

  1. (1)

    There exists a cocompact discrete subgroup Γm\Gamma\subset{\mathbb{R}}^{m} satisfying MΓΓM\Gamma\subset\Gamma.

  2. (2)

    There exists a cocompact discrete subgroup Γm\Gamma\subset{\mathbb{R}}^{m} satisfying MΓ=ΓM\Gamma=\Gamma.

  3. (3)

    All the coefficients of the characteristic polynomial of MM are integers.

  4. (4)

    All the coefficients of the characteristic polynomial of MM are integers and detM=±1\det M=\pm 1.

Then the implications (1)\Rightarrow(3) and (2)\Rightarrow(4) hold. Moreover, the equivalences (1)\Leftrightarrow(3) and (2)\Leftrightarrow(4) also hold if the eigenvalues of MM are distinct.

Proof..

First, we prove the implication (1)\Rightarrow(3). Since the uniform lattice of n{\mathbb{R}}^{n} is isomorphic to n{\mathbb{Z}}^{n}, MM is similar to an element of M(n,)M(n,{\mathbb{Z}}). Therefore, all the coefficients of characteristic polynomial are integers.

Next, we show the implication (2)\Rightarrow(4). Since MM is invertible and both detM\det M and detM1\det M^{-1} are integers, then we have detM=±1\det M=\pm 1.

Finally, we prove the inverse implications (3)\Rightarrow(1) and (4)\Rightarrow(2). Assume the eigenvalues of MM are distinct, and all the coefficients of the characteristic polynomial of MM are integers. Since the eigenvalues of MM are distinct, there exists vnv\in{\mathbb{R}}^{n} such that (Miv)i=0,1,,n1(M^{i}v)_{i=0,1,\cdots,n-1} is a basis of n{\mathbb{R}}^{n}. Then ΓMiv|i=0,1,,n1\Gamma\coloneqq\left\langle M^{i}v\ \left|\vphantom{M^{i}v}\ i=0,1,\cdots,n-1\right.\right\rangle_{{\mathbb{Z}}} satisfies the condition (1). Actually, by Cayley–Hamilton’s theorem, MnvM^{n}v is written as an linear combination of (Miv)i=0,1,,n1(M^{i}v)_{i=0,1,\cdots,n-1} with integer coefficients, so we have MnvΓM^{n}v\in\Gamma. Especially, in the case detM=±1\det M=\pm 1, since M1vM^{-1}v is also written as a linear combination of (Miv)i=0,1,,n1(M^{i}v)_{i=0,1,\cdots,n-1}, we have MΓ=ΓM\Gamma=\Gamma. ∎

Example 5.24.

Put M:=diag(1,1,0)M:=\operatorname{diag}\left(1,-1,0\right) and let us see that the group SMS_{M} admits a uniform lattice. Set t0:=log(2+3)t_{0}:=\log(2+\sqrt{3}). Since the characteristic polynomial of expt0M\exp t_{0}M is t35t2+5t1t^{3}-5t^{2}+5t-1 and its roots are distinct, there exists a uniform lattice Γ03\Gamma_{0}\subset{\mathbb{R}}^{3} which is expt0M\exp t_{0}M-invariant by Proposition 5.23. The subgroup t0Γ0SM\langle t_{0}\rangle\Gamma_{0}\subset S_{M} is a uniform lattice.

Proof of Proposition 5.20.

If MM is nilpotent, we already have trM=0\operatorname{tr}M=0. On the other hand, if MM is not nilpotent, n{\mathbb{R}}^{n} is the maximum connected normal nilpotent subgroup of SMS_{M}. In this case, let Γ\Gamma be a uniform lattice of SMS_{M}. Then Γn\Gamma\cap{\mathbb{R}}^{n} is a uniform lattice of n{\mathbb{R}}^{n} by Fact 5.21. By Proposition 5.23, we have det(exptM)=±1\det(\exp tM)=\pm 1 for some tpr(Γ){0}t\in\operatorname{pr}_{\mathbb{R}}(\Gamma)-\{0\}. Then we obtain trM=0\operatorname{tr}M=0. ∎

Finally, we prove Propositions 5.15 and 5.16.

Proof of Proposition 5.15.

Let Γ\Gamma be an \mathcal{I}_{\ell}-invariant uniform lattice in LCL_{C}. Since LCL_{C} is \mathcal{I}_{\ell}-invariant, the subspace 𝔩C\mathfrak{l}^{\prime}_{C} is Wt0W_{t_{0}}-invariant by Note 5.2. Now we are enough to show det(Wt0|𝔩C)=±1\det(W_{t_{0}}|_{\mathfrak{l}^{\prime}_{C}})=\pm 1. We denote by log:Hn𝔥n\log:H_{n}\to\mathfrak{h}_{n} the inverse of the exponential map exp:𝔥nHn\exp:\mathfrak{h}_{n}\to H_{n} (diffeomorphism).
(a) The case where LCL_{C} is commutative.

Since log(Γ)\log(\Gamma) is an Ad{\rm Ad}_{\ell}-invariant uniform lattice in 𝔩C\mathfrak{l}_{C}, we obtain det(Ad|𝔩C)=±1\det({\rm Ad}_{\ell}|_{\mathfrak{l}_{C}})=\pm 1 by Proposition 5.23, and so det(Wt0|𝔩C)=det(Ad|𝔩C)=±1\det(W_{t_{0}}|_{\mathfrak{l}^{\prime}_{C}})=\det({\rm Ad}_{\ell}|_{\mathfrak{l}_{C}})=\pm 1.
(b) The case where LCL_{C} is not commutative.

By Fact 5.22, the uniform lattice Γ\Gamma is also non-commutative. Since the quotient 𝒵Hn/(Γ𝒵Hn)\mathcal{Z}_{H_{n}}/(\Gamma\cap\mathcal{Z}_{H_{n}}) is compact, we apply Property 2.29 to the natural surjection π:LCLC/𝒵Hn\pi:L_{C}\to L_{C}/\mathcal{Z}_{H_{n}}, then Γ~π(Γ)\widetilde{\Gamma}\coloneqq\pi(\Gamma) is a discrete subgroup of LC/𝒵Hn2nL_{C}/\mathcal{Z}_{H_{n}}\simeq{\mathbb{R}}^{2n}. Hence, the subset log(Γ~)\log(\widetilde{\Gamma}) is a Wt0W_{t_{0}}-invariant uniform lattice of 𝔩C/𝔷\mathfrak{l}_{C}/\mathfrak{z}. Then we have det(Wt0|𝔩C)=det(Wt0|𝔩/𝔷)=±1\det(W_{t_{0}}|_{\mathfrak{l}^{\prime}_{C}})=\det(W_{t_{0}}|_{\mathfrak{l}/\mathfrak{z}})=\pm 1 by Proposition 5.23. ∎

Proof of Proposition 5.16.

Let Γ\Gamma be a uniform lattice in LC,wL_{C,w}.
(a) The case where LCL_{C} is commutative.

Since the matrix CC is symmetric by Note 5.7, the condition trDC=0\operatorname{tr}D^{\prime}C=0 follows from Note 5.19 and Proposition 5.20.
(b) The case where LCL_{C} is not commutative.

Since LCL_{C} is the maximum connected normal nilpotent Lie subgroup of LC,wL_{C,w}, the subgroup ΓLC\Gamma\cap L_{C} is a uniform lattice of LCL_{C} by Fact 5.21. By Fact 5.22, the lattice ΓLC\Gamma\cap L_{C} is not commutative, either. Hence 𝒵Hn/(Γ𝒵Hn)\mathcal{Z}_{H_{n}}/(\Gamma\cap\mathcal{Z}_{H_{n}}) is compact. By applying Property 2.29 to the natural surjection π:LC,wLC,w/𝒵Hn\pi:L_{C,w}\to L_{C,w}/\mathcal{Z}_{H_{n}}, we find Γ~:=π(Γ)\widetilde{\Gamma}:=\pi(\Gamma) is a uniform lattice of the group LC,w~:=π(LC,w)\widetilde{L_{C,w}}:=\pi(L_{C,w}). Since we have LC,w~SDC\widetilde{L_{C,w}}\simeq S_{D^{\prime}C} by Note 5.19, the condition trDC=0\operatorname{tr}D^{\prime}C=0 follows from Proposition 5.20. ∎

5.3 Criterions of the existence of compact Clifford–Klein forms

In this subsection, we give the following criterions for the existence of compact Clifford–Klein forms of GD,D/HG_{D,D^{\prime}}/H. If DD and DD^{\prime} are diagonal, this proposition also follows from [16, Proposition 4.8].

Proposition 5.25.

The following conditions are equivalent.

  1. (a)

    The symmetric space GD,D/HG_{D,D^{\prime}}/H admits compact Clifford–Klein forms.

  2. (b)

    There exists CM(n,)C\in M(n,{\mathbb{R}}) satisfying the following conditions.

    1. (i)

      The matrix At+BtCA_{t}+B_{t}C is invertible for any tt\in{\mathbb{R}}.

    2. (ii)

      The subgroup LCL_{C} has an \mathcal{I}_{\ell}-invariant uniform lattice for some GD,DHn\ell\in G_{D,D^{\prime}}-H_{n}.

Moreover, the following condition is a necessary condition of the above condition (ii).

  1. (ii’)

    The subspace 𝔩C\mathfrak{l}^{\prime}_{C} is Wt0W_{t_{0}}-invariant and det(Wt0|𝔩C)=±1\det(W_{t_{0}}|_{\mathfrak{l}^{\prime}_{C}})=\pm 1 for some t0×t_{0}\in{\mathbb{R}}^{\times}.

We prove this in Subsection 5.5. If GD,DG_{D,D^{\prime}} is completely solvable, we have an easier criterion.

Proposition 5.26.

Assume GD,DG_{D,D^{\prime}} is completely solvable. The following conditions are equivalent.

  1. (a)

    The symmetric space GD,D/HG_{D,D^{\prime}}/H admits compact Clifford–Klein forms.

  2. (b)

    There exist CM(n,)C\in M(n,{\mathbb{R}}) and w𝔥w\in\mathfrak{h} satisfying the following conditions.

    1. (i)

      CDC=DCD^{\prime}C=D.

    2. (ii)

      The subgroup LC,wL_{C,w} admits a uniform lattice.

Moreover, the following condition is a necessary condition of the above condition (ii).

  1. (ii’)

    trDC=0\operatorname{tr}D^{\prime}C=0.

Proof..

The implication (b)(ii)\Rightarrow(ii’) follows from Proposition 5.16, so we are enough to show the equivalence (a)\Leftrightarrow(b), which equivalent to check that the following conditions are equivalent:

  1. (A)

    there exists a discrete subgroup ΓGD,D\Gamma\subset G_{D,D^{\prime}} which acts on GD,D/HG_{D,D^{\prime}}/H properly discontinuously, cocompactly and freely,

  2. (B)

    there exists a constructor LL of GD,D/HG_{D,D^{\prime}}/H and LL has a uniform lattice,

  3. (C)

    there exists CM(n,)C\in M(n,{\mathbb{R}}) and w𝔥w\in\mathfrak{h} such that CDC=DCD^{\prime}C=D and LC,wL_{C,w} has a uniform lattice.

The implication (B)\Rightarrow(A) comes from Note 3.6 (2). Since GD,DG_{D,D^{\prime}} is completely solvable, the implication (A)\Rightarrow(B) follows from Fact 2.24. The equivalence (B)\Leftrightarrow(C) follows from Proposition 5.14. ∎

5.4 Intermediate syndetic hulls

In this subsection, we consider the existence problem of compact Clifford–Klein forms for solvable homogeneous spaces. A difficulty arises when GD,DG_{D,D^{\prime}} is not completely solvable. In fact, in this case, a discrete subgroup ΓGD,D\Gamma\subset G_{D,D^{\prime}} may fail to have its syndetic hull (Remark 2.38). To overcome this difficulty, we introduce intermediate syndetic hulls which play a similar role to syndetic hulls.

Definition 5.27 (intermediate syndetic hull).

For a closed subgroup ΓGD,D\Gamma\subset G_{D,D^{\prime}}, a closed subgroup LGD,DL^{\prime}\subset G_{D,D^{\prime}} is called an intermediate syndetic hull of Γ\Gamma if LL^{\prime} satisfies the following conditions with L0:=LHnL_{0}:=L^{\prime}\cap H_{n}:

  1. (i)

    L0L_{0} is connected,

  2. (ii)

    there exists ΓHn\ell\in\Gamma-H_{n} satisfying L=L0L^{\prime}=\langle{\ell}\rangle L_{0},

  3. (iii)

    Γ\Gamma is a cocompact subgroup of LL^{\prime}.

Note 5.28.

The condition (iii) above is equivalent to:

  1. (iii’)

    Γ\Gamma is a subgroup of LL^{\prime} and ΓHn\Gamma\cap H_{n} is cocompact in L0L_{0}.

In this subsection, we show the next:

Proposition 5.29.

Let ΓGD,D\Gamma\subset G_{D,D^{\prime}} be a discrete subgroup acting on GD,D/HG_{D,D^{\prime}}/H cocompactly. Then Γ\Gamma has an intermediate syndetic hull.

To prove this proposition, we use Lemmas 5.30 and 5.32.

Lemma 5.30 (Criterion of the existence of an intermediate syndetic hull).

Let ΓGD,D\Gamma\subset G_{D,D^{\prime}} be a discrete subgroup and put Γ0:=ΓHn\Gamma_{0}:=\Gamma\cap H_{n}. Then the following conditions are equivalent.

  1. (a)

    The subgroup pr1(Γ)\operatorname{pr}_{1}(\Gamma)\subset{\mathbb{R}} is non-trivial and discrete.

  2. (b)

    There exists γΓHn\gamma\in\Gamma-H_{n} satisfying Γ=γΓ0\Gamma=\langle\gamma\rangle\Gamma_{0}.

  3. (c)

    Γ\Gamma has an intermediate syndetic hull.

Proof..

The implication (c)\Rightarrow(a) is easy. We show the implication (a)\Rightarrow(b). Let t0pr1(Γ)t_{0}\in\operatorname{pr}_{1}(\Gamma) be a generator of pr1(Γ)\operatorname{pr}_{1}(\Gamma). We take γΓ\gamma\in\Gamma satisfying pr1(γ)=t0\operatorname{pr}_{1}(\gamma)=t_{0}. Then the following exact sequence splits by the group homomorphism s:pr1(Γ)Γ,t0γs:\operatorname{pr}_{1}(\Gamma)\to\Gamma,\quad t_{0}\mapsto\gamma.

{e}Γ0Γpr1pr1(Γ)s{e}.\lx@xy@svg{\hbox{\raise 0.0pt\hbox{\kern 10.32814pt\hbox{\ignorespaces\ignorespaces\ignorespaces\hbox{\vtop{\kern 0.0pt\offinterlineskip\halign{\entry@#!@&&\entry@@#!@\cr&&&&\crcr}}}\ignorespaces{\hbox{\kern-10.32814pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{\{e\}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces{\hbox{\kern 34.32814pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}{\hbox{\kern 34.32814pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{\Gamma_{0}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces{\hbox{\kern 73.37816pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}{\hbox{\kern 73.37816pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{\Gamma\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces\ignorespaces\ignorespaces{\hbox{\kern 95.3254pt\raise 5.1875pt\hbox{{}\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\hbox{\hbox{\kern 0.0pt\raise-0.8264pt\hbox{$\scriptstyle{\operatorname{pr}_{1}}$}}}\kern 3.0pt}}}}}}\ignorespaces{\hbox{\kern 109.62816pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}{\hbox{\kern 109.62816pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{\operatorname{pr}_{1}(\Gamma)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces{\hbox{\kern 165.9282pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{}{}{{}{}{}{}{}{}{}{}}{}\ignorespaces\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{{}{}{{}}{{}{}{}}{}}}}\ignorespaces{}\ignorespaces{}{}{}{{}{}}\ignorespaces\ignorespaces{\hbox{\kern 98.00005pt\raise-6.87415pt\hbox{{}\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\hbox{\hbox{\kern 0.0pt\raise-1.50694pt\hbox{$\scriptstyle{s}$}}}\kern 3.0pt}}}}}}\ignorespaces{}{}{}{{}{}{}{}{}{}{}{}{}{}{}{}{}{}{}{}{}{}{}{}{}{}{}}{\hbox{\kern 82.78517pt\raise-2.99979pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}\ignorespaces\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{{}{}{}{{}}{{}{}{}\lx@xy@spline@}{}}}}\ignorespaces{}\ignorespaces\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{{}{}{{}}{{}{}{}}{}}}}\ignorespaces{}{\hbox{\kern 165.9282pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{\{e\}}$}}}}}}}\ignorespaces}}}}\ignorespaces.

Then we have Γ=s(pr1(Γ))Γ0=γΓ0\Gamma=s(\operatorname{pr}_{1}(\Gamma))\Gamma_{0}=\langle\gamma\rangle\Gamma_{0} and so the condition (b).

Then we see the implication (b)\Rightarrow(c). By Fact 2.37, the discrete subgroup Γ0\Gamma_{0} has the syndetic hull L0HnL_{0}\subset H_{n}. Then the closed subgroup L=γL0L^{\prime}=\langle{\gamma}\rangle L_{0} clearly satisfies the conditions (i) and (ii) in Definition 5.27 and (iii’) in Note 5.28. ∎

In the rest of this subsection, we use the following:

Notation and Setting 5.31.

Let NN be a 1-connected 2-step nilpotent Lie group, 𝔫\mathfrak{n} its Lie algebra.

  • 𝔫:=𝔫\mathfrak{n}_{{\mathbb{C}}}:=\mathfrak{n}\otimes{\mathbb{C}}

  • 𝔫(A,λ)𝔫\mathfrak{n}(A,\lambda)\subset\mathfrak{n}_{\mathbb{C}} is the generalized eigenspace of AEnd(𝔫)A\in\operatorname{End}(\mathfrak{n}) with respect to an eigenvalue λ\lambda\in{\mathbb{C}}.

  • 𝔫(A,λ):=kλ𝔫(A,k)\mathfrak{n}(A,\lambda)^{\perp}:=\bigoplus_{k\neq\lambda}\mathfrak{n}(A,k).

Lemma 5.32.

Let MDer𝔫M\in\operatorname{Der}\mathfrak{n}. we put ϕ:Aut(𝔫),texptM\phi:{\mathbb{R}}\to\operatorname{Aut}(\mathfrak{n}),\ t\mapsto\exp tM. By identifying Aut(N)Aut(𝔫)\operatorname{Aut}(N)\simeq\operatorname{Aut}(\mathfrak{n}), put G:=ϕNG:={\mathbb{R}}\ltimes_{\phi}N and pr:G\operatorname{pr}_{\mathbb{R}}:G\to{\mathbb{R}} the first projection. Let ΓG\Gamma\subset G be a discrete subgroup. Then pr(Γ)\operatorname{pr}_{\mathbb{R}}(\Gamma) is discrete if the following conditions are satisfied.

  1. (1)

    𝔫(M,0)=[𝔫,𝔫]\mathfrak{n}(M,0)=[\mathfrak{n}_{\mathbb{C}},\mathfrak{n}_{\mathbb{C}}].

  2. (2)

    Γ0:=ΓN\Gamma_{0}:=\Gamma\cap N is commutative.

  3. (3)

    𝔩0[𝔫,𝔫]\mathfrak{l}_{0}\not\subset[\mathfrak{n},\mathfrak{n}], where 𝔩0\mathfrak{l}_{0} is the Lie algebra of the syndetic hull of Γ0\Gamma_{0}.

To prove this lemma, we use the following lemma.

Lemma 5.33.

In the setting in Lemma 5.32 with the assumptions (1), (2) and (3), put a finite set FF as follows, where λk\lambda_{k}\in{\mathbb{C}} are eigenvalues of MM.

F:={fI(t):=kIetλk|I{1,2,,dim𝔫},fI is not constant}C().F:=\left\{f_{I}(t):=\displaystyle\sum_{k\in I}e^{t\lambda_{k}}\ \left|\vphantom{f_{I}(t):=\displaystyle\sum_{k\in I}e^{t\lambda_{k}}}\ I\subset\{1,2,\cdots,\dim\mathfrak{n}\},\ f_{I}\text{ is not constant}\right.\right\}\subset C^{\infty}({\mathbb{R}}).

Fix γ=(t0,expv0)Γ\gamma=(t_{0},\exp v_{0})\in\Gamma and assume 𝔫(ϕt0,1)=[𝔫,𝔫]\mathfrak{n}(\phi_{t_{0}},1)=[\mathfrak{n}_{\mathbb{C}},\mathfrak{n}_{\mathbb{C}}]. We think AdγGL(𝔫){\rm Ad}_{\gamma}\in GL(\mathfrak{n}_{\mathbb{C}}). Then there exists fFf\in F satisfying trAdγ|𝔩0=f(t0)\operatorname{tr}{\rm Ad}_{\gamma}|_{\mathfrak{l}_{0}}=f(t_{0}).

Proof..

Since 𝔫\mathfrak{n} is 2-step nilpotent, we have:

AdγX=ϕt0X+[v0,ϕt0X](X𝔫).{\rm Ad}_{\gamma}X=\phi_{t_{0}}X+[v_{0},\phi_{t_{0}}X]\quad(\forall X\in\mathfrak{n}_{\mathbb{C}}).

Put V:=𝔫(ϕt0,1)V:=\mathfrak{n}(\phi_{t_{0}},1)^{\perp}, then we have a decomposition 𝔫=V[𝔫,𝔫]\mathfrak{n}_{\mathbb{C}}=V\oplus[\mathfrak{n}_{\mathbb{C}},\mathfrak{n}_{\mathbb{C}}] and the following matrix representation:

Adγ|𝔫=(ϕt0|V0ϕt0|[𝔫,𝔫]).{\rm Ad}_{\gamma}|_{\mathfrak{n}_{\mathbb{C}}}=\begin{pmatrix}\phi_{t_{0}}|_{V}&0\\ *&{\phi_{t_{0}}}|_{[\mathfrak{n}_{\mathbb{C}},\mathfrak{n}_{\mathbb{C}}]}\end{pmatrix}.

Therefore, the eigenvalues of Adγ|𝔫{\rm Ad}_{\gamma}|_{\mathfrak{n}_{\mathbb{C}}} coincides with them of ϕt0\phi_{t_{0}}. Since 𝔩0\mathfrak{l}_{0} is Adγ{\rm Ad}_{\gamma}-invariant, there exists I{1,2,,dim𝔫}I\subset\{1,2,\cdots,\dim\mathfrak{n}\} such that trAdγ|𝔩0=fI(t0)\operatorname{tr}{\rm Ad}_{\gamma}|_{\mathfrak{l}_{0}}=f_{I}(t_{0}). Then we only have to show that fIf_{I} is not constant. It is enough to show that there exists kIk\in I satisfying λk0\lambda_{k}\neq 0. Assume λk=0\lambda_{k}=0 for any kIk\in I then the eigenvalues of Adγ|𝔩0{\rm Ad}_{\gamma}|_{\mathfrak{l}_{0}} are all 1. Then by the matrix representation Adγ|𝔫{\rm Ad}_{\gamma}|_{\mathfrak{n}_{\mathbb{C}}}, we have 𝔩0[𝔫,𝔫]\mathfrak{l}_{0}\otimes{\mathbb{C}}\subset[\mathfrak{n}_{\mathbb{C}},\mathfrak{n}_{\mathbb{C}}], which contradicts the condition 𝔩0[𝔫,𝔫]\mathfrak{l}_{0}\not\subset[\mathfrak{n},\mathfrak{n}]. Therefore, we obtain fIFf_{I}\in F. ∎

Proof of Lemma 5.32..

It is enough to show that pr(Γ)\operatorname{pr}_{\mathbb{R}}(\Gamma) is included in a countable and closed subset of {\mathbb{R}}. We put a finite set FF as in Lemma 5.33, and put subsets A,BA,B\subset{\mathbb{R}} as follows:

A\displaystyle A :={t|𝔫(ϕt,1)=[𝔫,𝔫]},\displaystyle:=\left\{t\in{\mathbb{R}}\ \left|\vphantom{t\in{\mathbb{R}}}\ \mathfrak{n}(\phi_{t},1)=[\mathfrak{n}_{\mathbb{C}},\mathfrak{n}_{\mathbb{C}}]\right.\right\},
B\displaystyle B :=fFf1().\displaystyle:=\bigcup_{f\in F}f^{-1}({\mathbb{Z}}).

Then Ac:=AA^{c}:={\mathbb{R}}-A and BB are countable and closed. Take any γ=(t,v)Γ\gamma=(t,v)\in\Gamma. We are enough to show that tAcBt\in A^{c}\cup B. Assume tAt\in A, then there exists fFf\in F such that trAdγ|𝔩0=f(t)\operatorname{tr}{\rm Ad}_{\gamma}|_{\mathfrak{l}_{0}}=f(t) by Lemma 5.33. Since the subalgebra 𝔩0\mathfrak{l}_{0} is abelian and has an Adγ{\rm Ad}_{\gamma}-invariant uniform lattice, trAdγ|𝔩0\operatorname{tr}{\rm Ad}_{\gamma}|_{\mathfrak{l}_{0}} must be integer by Proposition 5.23. Therefore we have tf1()Bt\in f^{-1}({\mathbb{Z}})\subset B. ∎

Finally, we prove Proposition 5.29.

Proof of Proposition 5.29.

Take a discrete subgroup ΓGD,D\Gamma\subset G_{D,D^{\prime}} acting on GD,D/HG_{D,D^{\prime}}/H cocompactly. By Lemma 5.30, we are enough to show that pr1(Γ)\operatorname{pr}_{1}(\Gamma) is discrete. We consider the natural surjection π:GD,DGD,D/𝒵HnSW\pi:G_{D,D^{\prime}}\to G_{D,D^{\prime}}/\mathcal{Z}_{H_{n}}\simeq S_{W} (see Note 5.19) and put GD,D~,H~\widetilde{G_{D,D^{\prime}}},\ \widetilde{H} and Γ~\widetilde{\Gamma} the image by π\pi of GD,D,HG_{D,D^{\prime}},H and Γ\Gamma, respectively. By Property 2.29, Γ~\widetilde{\Gamma} acts on GD,D~/H~\widetilde{G_{D,D^{\prime}}}/\widetilde{H} cocompactly. We consider Γ~\widetilde{\Gamma} as a subgroup of SWS_{W}. We put Γ0:=ΓHn\Gamma_{0}:=\Gamma\cap H_{n}.

  1. (A)

    The case Γ0𝒵Hn\Gamma_{0}\subset\mathcal{Z}_{H_{n}}.

    We are enough to show pr1(Γ)A\operatorname{pr}_{1}(\Gamma)\subset A for A:={t|Wt has an eigenvalue 1}A:=\left\{t\in{\mathbb{R}}\ \left|\vphantom{t\in{\mathbb{R}}}\ W_{t}\text{ has an eigenvalue }1\right.\right\}. Here, note that AA\subset{\mathbb{R}} is closed and countable. Assume pr1(Γ)A\operatorname{pr}_{1}(\Gamma)\not\subset A. Then there exists (t0,v0)Γ~(t_{0},v_{0})\in\widetilde{\Gamma}, where t0At_{0}\in{\mathbb{R}}-A and v02nv_{0}\in{\mathbb{R}}^{2n}. By [Γ,Γ]Γ0𝒵Hn[\Gamma,\Gamma]\subset\Gamma_{0}\subset\mathcal{Z}_{H_{n}}, Γ~\widetilde{\Gamma} is abelian. Especially, we have Γ~𝒵SW((t0,v0))=x𝒵SW(t0)\widetilde{\Gamma}\subset\mathcal{Z}_{S_{W}}((t_{0},v_{0}))=\mathcal{I}_{x}\mathcal{Z}_{S_{W}}(t_{0}), where x:=(idWt0)1v0x:=(\operatorname{id}-W_{t_{0}})^{-1}v_{0}. By Lemma 5.18, we have:

    Γ~x𝒵SW(t0)x.\widetilde{\Gamma}\subset\mathcal{I}_{x}\mathcal{Z}_{S_{W}}(t_{0})\subset\mathcal{I}_{x}{\mathbb{R}}.

    On the other hand, since the {\mathbb{R}}-action on GD,D~/H~\widetilde{G_{D,D^{\prime}}}/\widetilde{H} is not cocompact, neither is the Γ~\widetilde{\Gamma}-action, which contradicts the cocompactness of Γ~\widetilde{\Gamma}-action (Property 2.29). Then we have pr1(Γ)A\operatorname{pr}_{1}(\Gamma)\subset A.

  2. (B)

    The case Γ0𝒵Hn\Gamma_{0}\not\subset\mathcal{Z}_{H_{n}}.

    1. (a)

      Γ0\Gamma_{0} is commutative.
      We denote by 𝔩0\mathfrak{l}_{0} the Lie algebra of the syndetic hull of Γ0\Gamma_{0}. By using Lemma 5.32 for (N,M,Γ)=(Hn,diag(W,0),Γ)(N,M,\Gamma)=(H_{n},\operatorname{diag}\left(W,0\right),\Gamma), we get pr1(Γ)\operatorname{pr}_{1}(\Gamma) is discrete. Actually, it is easy to check the conditions in Lemma 5.32 as follows.

      1. (1)

        𝔫(M,0)=𝔷(=[𝔥n,𝔥n])\mathfrak{n}(M,0)=\mathfrak{z}\otimes{\mathbb{C}}(=[\mathfrak{h}_{n},\mathfrak{h}_{n}]\otimes{\mathbb{C}}).

      2. (2)

        Γ0\Gamma_{0} is commutative by the assumption (a).

      3. (3)

        By the assumption Γ0𝒵Hn\Gamma_{0}\not\subset\mathcal{Z}_{H_{n}}, we have 𝔩0[𝔥n,𝔥n]=𝔷\mathfrak{l}_{0}\not\subset[\mathfrak{h}_{n},\mathfrak{h}_{n}]=\mathfrak{z}.

    2. (b)

      Γ0\Gamma_{0} is non-commutative.
      We have [Γ0,Γ0][\Gamma_{0},\Gamma_{0}] is non-trivial and so 𝒵Hn/(𝒵HnΓ0)\mathcal{Z}_{H_{n}}/(\mathcal{Z}_{H_{n}}\cap\Gamma_{0}) is compact. By Property 2.29, Γ~\widetilde{\Gamma} is a discrete subgroup of GD,D~\widetilde{G_{D,D^{\prime}}} and acts on GD,D~/H~\widetilde{G_{D,D^{\prime}}}/\widetilde{H} cocompactly. By using Lemma 5.32 for (N,M,Γ)=(2n,W,Γ~)(N,M,\Gamma)=({\mathbb{R}}^{2n},W,\widetilde{\Gamma}), we have the discreteness of pr1(Γ)=pr(Γ~)\operatorname{pr}_{1}(\Gamma)=\operatorname{pr}_{\mathbb{R}}(\widetilde{\Gamma}). Actually, it is easy to check the conditions in Lemma 5.32 as follows.

      1. (1)

        𝔫(M,0)={0}(=[2n,2n])\mathfrak{n}(M,0)=\{0\}(=[{\mathbb{R}}^{2n},{\mathbb{R}}^{2n}]\otimes{\mathbb{C}}).

      2. (2)

        Since NN is commutative, so is Γ0~:=Γ~N\widetilde{\Gamma_{0}}:=\widetilde{\Gamma}\cap N.

      3. (3)

        By the assumption Γ0𝒵Hn\Gamma_{0}\not\subset\mathcal{Z}_{H_{n}}, we have 𝔩0~[2n,2n]={0}\widetilde{\mathfrak{l}_{0}}\not\subset[{\mathbb{R}}^{2n},{\mathbb{R}}^{2n}]=\{0\}, where 𝔩0~\widetilde{\mathfrak{l}_{0}} is the Lie algebra of the syndetic hull of Γ0~\widetilde{\Gamma_{0}}.

5.5 (L) condition

In this subsection, our goal is to prove Proposition 5.25. To do this, we introduce (L) condition.

Definition 5.34 ((L) condition).

We say a closed subgroup LGD,DL^{\prime}\subset G_{D,D^{\prime}} satisfies (L) condition if the following conditions are satisfied.

  1. (1)

    LL^{\prime} is unimodular,

  2. (2)

    L0=LHnL_{0}=L^{\prime}\cap H_{n} is connected,

  3. (3)

    there exists LL0\ell\in L^{\prime}-L_{0} satisfying L=L0L^{\prime}=\langle\ell\rangle L_{0}.

Clearly, intermediate syndetic hulls satisfy (L) condition. We see a fundamental property of (L) condition, namely, we have:

Proposition 5.35 (Criterion of properness and cocompactness).

Suppose a closed subgroup LGD,DL^{\prime}\subset G_{D,D^{\prime}} satisfies (L) condition and put L0:=LHnL_{0}:=L^{\prime}\cap H_{n}. Then the following conditions are equivalent:

  1. (a)

    the LL^{\prime}-action on GD,D/HG_{D,D^{\prime}}/H is proper and cocompact,

  2. (b)

    the pair (L0,H)(L_{0},H) satisfies the property (CI) in GD,DG_{D,D^{\prime}} and the L0L_{0}-action on Hn/HH_{n}/H is cocompact,

  3. (c)

    there exists a matrix CM(n,)C\in M(n,{\mathbb{R}}) such that L0=LCL_{0}=L_{C} and At+BtCA_{t}+B_{t}C is invertible for any tt\in{\mathbb{R}} (see Notation 5.1).

Proof..

Take L\ell\in L^{\prime} satisfying L=L0L^{\prime}=\langle{\ell}\rangle L_{0}. First we show the equivalence (a)\Leftrightarrow(b). Since the LL^{\prime}-action on GD,D/HnG_{D,D^{\prime}}/H_{n} is cocompact, it follows from Propositions 3.15 and 3.16 by putting (G,N,L0,L1,H)=(GD,D,Hn,L0,,H)(G,N,L_{0},L_{1},H)=(G_{D,D^{\prime}},H_{n},L_{0},\langle{\ell}\rangle,H). Here we need to check that the tuple satisfies the condition of Setting 3.14. The condition L1𝒩G(L0)L_{1}\subset\mathcal{N}_{G}(L_{0}) is clear. Take T𝔤D,DT\in\mathfrak{g}_{D,D^{\prime}} satisfying expT=\exp T=\ell, we have expTH\exp{\mathbb{R}}T\pitchfork H in GD,DG_{D,D^{\prime}} by Lemma 2.32, and so L1NL_{1}\pitchfork N.

Next we show the equivalence (b)\Leftrightarrow(c). By Proposition 3.15 for G=N=HnG=N=H_{n}, the condition that (L0,H)(L_{0},H) satisfies the property (CI) in GD,DG_{D,D^{\prime}} implies L0HL_{0}\pitchfork H in HnH_{n}. Then the equivalence (b)\Leftrightarrow(c) follows from Propositions 3.7, 5.9 and 5.12. ∎

Finally, we prove Proposition 5.25.

Proof of Proposition 5.25.

First, we show the implication (a)\Rightarrow(b). Take a discrete subgroup ΓGD,D\Gamma\subset G_{D,D^{\prime}} such that Γ\GD,D/H\Gamma\backslash G_{D,D^{\prime}}/H is a compact Clifford–Klein form. By Proposition 5.29, we take an intermediate syndetic hull LGD,DL^{\prime}\subset G_{D,D^{\prime}} of Γ\Gamma. Put L0:=LHnL_{0}:=L^{\prime}\cap H_{n} and take ΓH\ell\in\Gamma-H satisfying L=L0L^{\prime}=\langle{\ell}\rangle L_{0}. By Fact 2.24, the LL^{\prime}-action on GD,D/HG_{D,D^{\prime}}/H is proper and cocompact. Take CM(n,)C\in M(n,{\mathbb{R}}) such that L0=LCL_{0}=L_{C} and At+BtCA_{t}+B_{t}C is invertible for any tt\in{\mathbb{R}} by Proposition 5.35. Hence the condition (b)(i) holds. Moreover, by Proposition 5.29(iii), LCL_{C} admits an \mathcal{I}_{\ell}-invariant uniform lattice ΓHn\Gamma\cap H_{n} and so the condition (b)(ii) holds.

Next, we check the implication (b)\Rightarrow(a). Take CM(n,)C\in M(n,{\mathbb{R}}) and GD,DHn\ell\in G_{D,D^{\prime}}-H_{n} satisfying the condition (b) and let Γ0LC\Gamma_{0}\subset L_{C} be an \mathcal{I}_{\ell}-invariant uniform lattice. Put LLCL^{\prime}\coloneqq\langle\ell\rangle L_{C} and ΓΓ0\Gamma\coloneqq\langle\ell\rangle\Gamma_{0}. Note that LL^{\prime} satisfies the (L) condition and Γ\Gamma is a uniform lattice of LL^{\prime}. The LL^{\prime}-action is proper and cocompact by Proposition 5.35. Therefore, Γ\GD,D/H\Gamma\backslash G_{D,D^{\prime}}/H is a compact Clifford–Klein form by Note 3.6.

Finally, the implication (b)(ii)\Rightarrow(ii’) comes from Proposition 5.15. ∎

6 Proof of the main theorem

In this section, we give a proof of the main theorem (Theorem 1.5).

We check the existence of compact Clifford–Klein forms for the spaces with signature (2,2)(2,2) which correspond to each case in Fact 1.4.

6.1 On the spaces which correspond to the triples (𝔤nil,σ,g±)(\mathfrak{g}_{\rm nil},\sigma,g_{\pm})

We consider the spaces which correspond to Case (I) in Fact 1.4.

6.1.1 transvection group

Note that GnilG_{\rm nil} is the transvection group of the symmetric spaces (Gnil/H,σ,g±)(G_{\rm nil}/H,\sigma,g_{\pm}).

Proposition 6.1.

The symmetric space Gnil/HG_{\rm nil}/H does not admit compact Clifford–Klein forms.

Proof..

Assume Gnil/HG_{\rm nil}/H admits a compact Clifford–Klein form Γ\Gnil/H\Gamma\backslash G_{\rm nil}/H. Since the Lie group GnilG_{\rm nil} is 1-connected and nilpotent, there exists a constructor LL including Γ\Gamma cocompactly by Fact 2.37. By Proposition 3.7, we get 𝔤nil=𝔩𝔥\mathfrak{g}_{\rm nil}=\mathfrak{l}\oplus\mathfrak{h}, where 𝔩\mathfrak{l} and 𝔥\mathfrak{h} are the Lie algebras of LL and HH, respectively. On the other hand, there is no such a subalgebra 𝔩\mathfrak{l} by Lemma 6.2. ∎

In this proof, we use the following lemmas.

Lemma 6.2.

For 𝔤nil\mathfrak{g}_{\text{nil}} and 𝔥\mathfrak{h} in Definition 4.15, no subalgebra 𝔩𝔤nil\mathfrak{l}\subset\mathfrak{g}_{\rm nil} satisfies 𝔤nil=𝔩𝔥\mathfrak{g}_{\rm{nil}}=\mathfrak{l}\oplus\mathfrak{h}.

Proof..

Take 𝔮\mathfrak{q} as in Definition 4.15. Suppose a subalgebra 𝔩𝔤nil\mathfrak{l}\subset\mathfrak{g}_{\rm nil} satisfies 𝔤nil=𝔩𝔥\mathfrak{g}_{\rm{nil}}=\mathfrak{l}\oplus\mathfrak{h}. We denote by 𝔷0\mathfrak{z}_{0} the center of 𝔤nil\mathfrak{g}_{\rm nil}. Since the element C1,C2𝔷0𝔮C_{1},C_{2}\in\mathfrak{z}_{0}\cap\mathfrak{q} satisfy the assumption of Lemma 6.3 below, we have 𝔷0𝔩\mathfrak{z}_{0}\subset\mathfrak{l}. Take k1,k2k_{1},k_{2}\in{\mathbb{R}} satisfying A1+k1B,A2+k2B𝔩A_{1}+k_{1}B,A_{2}+k_{2}B\in\mathfrak{l}. Then [A1+k1B,A2+k2B]B=k1C1+k2C2𝔷𝔩[A_{1}+k_{1}B,A_{2}+k_{2}B]-B=k_{1}C_{1}+k_{2}C_{2}\in\mathfrak{z}\subset\mathfrak{l} and so we have B𝔩B\in\mathfrak{l}, which is a contradiction. ∎

Lemma 6.3.

Let 𝔤\mathfrak{g} be a Lie algebra, 𝔷0\mathfrak{z}_{0} its center and 𝔤=𝔮𝔥=𝔩𝔥\mathfrak{g}=\mathfrak{q}\oplus\mathfrak{h}=\mathfrak{l}\oplus\mathfrak{h} its decompositions as a linear space. Assume 𝔥\mathfrak{h} and 𝔩\mathfrak{l} are subalgebras of 𝔤\mathfrak{g} and 𝔥\mathfrak{h} is abelian. For Z0𝔷0𝔮Z_{0}\in\mathfrak{z}_{0}\cap\mathfrak{q} satisfying the following condition, we have Z0𝔩Z_{0}\in\mathfrak{l}.

Y(𝔥{0}),X𝔮s.t.[X,Y]=Z0.\forall Y\in(\mathfrak{h}-\{0\}),\ \exists X\in\mathfrak{q}\ \text{s.t.}\ [X,Y]=Z_{0}.
Proof..

Take the linear map ϕ:𝔮𝔥\phi:\mathfrak{q}\to\mathfrak{h} satisfying 𝔩={x+ϕ(x)|x𝔮}\mathfrak{l}=\left\{x+\phi(x)\ \left|\vphantom{x+\phi(x)}\ x\in\mathfrak{q}\right.\right\}. Since ϕ(Z0)=0\phi(Z_{0})=0 implies Z0𝔩Z_{0}\in\mathfrak{l}, we assume ϕ(Z0)0\phi(Z_{0})\neq 0. Then there exists X𝔮X\in\mathfrak{q} satisfying [X,ϕ(Z0)]=Z0[X,\phi(Z_{0})]=Z_{0}, so we have 𝔩[X+ϕ(X),Z0+ϕ(Z0)]=[X,ϕ(Z0)]=Z0\mathfrak{l}\ni[X+\phi(X),Z_{0}+\phi(Z_{0})]=[X,\phi(Z_{0})]=Z_{0}. ∎

6.1.2 isometry group

First, we calculate isometry group of the symmetric spaces.

Proposition 6.4.

We have a Lie group isomorphism Gi(/2O(1,1))GtG_{\text{i}}\simeq({\mathbb{Z}}/2{\mathbb{Z}}\ltimes O(1,1))\ltimes G_{\text{t}}. Especially, GiG_{\text{i}} is completely solvable.

In the following, we calculate the group HiH_{\text{i}}.

Lemma 6.5.

Using the basis (A1,A2,B,C1,C2)(A_{1},A_{2},B,C_{1},C_{2}) of 𝔤nil\mathfrak{g}_{\rm nil}, we have:

Aut(𝔤nil,σ,g±)\displaystyle\operatorname{Aut}(\mathfrak{g}_{\rm nil},\sigma,g_{\pm}) =(I1,11I1,1),(P1P),(I21aI2I2):a,PPt,J,I2\displaystyle=\left\langle\begin{pmatrix}I_{1,1}&&\\ &-1&\\ &&I_{1,1}\\ \end{pmatrix},\begin{pmatrix}P&&\\ &1&\\ &&P\\ \end{pmatrix},\begin{pmatrix}I_{2}&&\\ &1&\\ aI_{2}&&I_{2}\\ \end{pmatrix}:\begin{gathered}a\in{\mathbb{R}},\\ P\in\langle{P_{t},J,-I_{2}}\rangle\end{gathered}\right\rangle
(/2O(1,1)),\displaystyle\simeq({\mathbb{Z}}/2{\mathbb{Z}}\ltimes O(1,1))\ltimes{\mathbb{R}},

where Pt:=exp(tI1,1)=diag(et0,et0)P_{t}:=\exp(-tI_{1,1})=\operatorname{diag}\left(e^{-t_{0}},e^{t_{0}}\right) and J:=(11)J:=\begin{pmatrix}&-1\\ 1\end{pmatrix}.

Proof..

Let ϕAut(𝔤t,σ,g±)\phi\in\operatorname{Aut}(\mathfrak{g}_{\text{t}},\sigma,g_{\pm}). We use (A1,A2,B,C1,C2)(A_{1},A_{2},B,C_{1},C_{2}) as the basis of 𝔤t\mathfrak{g}_{\text{t}}. Then the representation matrix of ϕ\phi is:

ϕ=(PkQR)GL(5,),\phi=\begin{pmatrix}P&&\\ &k&\\ Q&&R\\ \end{pmatrix}\in GL(5,{\mathbb{R}}),

where k×k\in{\mathbb{R}}^{\times} and P,Q,RGL(2,)P,Q,R\in GL(2,{\mathbb{R}}). Then we have:

g± is ϕ invariantϕTg±ϕ=g±\displaystyle\text{$g_{\pm}$ is $\phi$ invariant}\iff\phi^{T}g_{\pm}\phi=g_{\pm} {PTJR=JQTJP+PTJQ=0\displaystyle\iff\begin{cases}P^{T}JR=J\\ Q^{T}JP+P^{T}JQ=0\end{cases} (1)
[ϕ(A1),ϕ(A2)]=ϕ([A1,A2])\displaystyle[\phi(A_{1}),\phi(A_{2})]=\phi([A_{1},A_{2}]) detP=k,\displaystyle\iff\det P=k, (2)
[ϕ(Ai),ϕ(B)]=ϕ([Ai,B])(i=1,2)\displaystyle[\phi(A_{i}),\phi(B)]=\phi([A_{i},B])\ (i=1,2) kP=R,\displaystyle\iff kP=R, (3)

By the first identity of (0.7) and (0,5), we have kPTJP=JkP^{T}JP=J. Then we have k2(detP)2=1k^{2}(\det P)^{2}=1 and by (0.6), we have k4=1k^{4}=1, and so k=±1k=\pm 1. By the second identity of (0.5), PTJQP^{T}JQ is skew symmetric. We put aa\in{\mathbb{R}} by PTJQ=aJP^{T}JQ=aJ, then we have Q=akPQ=akP. Moreover, a direct calculation leads us ϕAut(𝔤t,σ,g)\phi\in\operatorname{Aut}(\mathfrak{g}_{\text{t}},\sigma,g) if ϕ\phi satisfies the conditions of Lemma 6.5.

Lemma 6.6.

By the identification HiAut(𝔤t,σ,g)H_{\text{i}}\simeq\operatorname{Aut}(\mathfrak{g}_{\text{t}},\sigma,g), we have:

Ht={(I21aI2I2)|a}.H_{\text{t}}=\left\{\begin{pmatrix}I_{2}&&\\ &1&\\ aI_{2}&&I_{2}\end{pmatrix}\ \left|\vphantom{\begin{pmatrix}I_{2}&&\\ &1&\\ aI_{2}&&I_{2}\end{pmatrix}}\ a\in{\mathbb{R}}\right.\right\}\simeq{\mathbb{R}}.

Then we prove:

Proposition 6.7.

Gi/HiG_{\text{i}}/H_{\text{i}} admits compact Clifford–Klein forms.

Proof..

We are enough to show that (Gi)0/(Hi)0(G_{\text{i}})_{0}/(H_{\text{i}})_{0} admits compact Clifford–Klein forms. Note that 𝔤i𝔤t\mathfrak{g}_{\text{i}}\simeq{\mathbb{R}}\ltimes\mathfrak{g}_{\text{t}}. Let XX be the generator of {\mathbb{R}}, we have:

[X,A1]=A2,[X,A2]=A1,[X,B]=0,[X,C1]=C2,[X,C2]=C1.[X,A_{1}]=A_{2},\ [X,A_{2}]=A_{1},\ [X,B]=0,\ [X,C_{1}]=C_{2},\ [X,C_{2}]=C_{1}.

Then we put T1:=A1A2X,T2:=A1A2+2B2C12C2T_{1}:=-A_{1}-A_{2}-X,\ T_{2}:=A_{1}-A_{2}+2B-2C_{1}-2C_{2} and:

𝔩:=T1,T2,C1,C2,L:=expL.\mathfrak{l}:=\left\langle T_{1},T_{2},C_{1},C_{2}\right\rangle_{{\mathbb{R}}},\ L:=\exp L.

Note that [T1,T2]=T2[T_{1},T_{2}]=T_{2}. Then 𝔩\mathfrak{l} is an ideal of 𝔤i\mathfrak{g}_{\text{i}} and we have 𝔤i=𝔩𝔥\mathfrak{g}_{\text{i}}=\mathfrak{l}\oplus\mathfrak{h}. We are enough to show that:

  1. (1)

    LHiL\pitchfork H_{\text{i}} in GiG_{\text{i}},

  2. (2)

    LL acts on Gi/HiG_{\text{i}}/H_{\text{i}} cocompactly,

  3. (3)

    LL admits a uniform lattice Γ\Gamma.

We check the condition (1). Note that 𝔤i\mathfrak{g}_{\text{i}} is 3-step nilpotent. By Fact 3.13, it is equivalent to the condition that (L,H)(L,H) is (CI) in GiG_{\text{i}}. By Note 3.1, we are enough to check that Adg𝔩𝔥={0}{\rm Ad}_{g}\mathfrak{l}\cap\mathfrak{h}=\{0\} for all gGig\in G_{\text{i}}. Since 𝔩\mathfrak{l} is an ideal, we have Adg𝔩𝔥=𝔩𝔥={0}{\rm Ad}_{g}\mathfrak{l}\cap\mathfrak{h}=\mathfrak{l}\cap\mathfrak{h}=\{0\}. The condition (2) follows from Proposition 3.7. Finally we check the condition (3). We put:

𝔩0:=T1,T2,C1,C2,Γ:=exp𝔩0.\mathfrak{l}_{0}:=\left\langle T_{1},T_{2},C_{1},C_{2}\right\rangle_{{\mathbb{Z}}},\ \Gamma:=\exp\mathfrak{l}_{0}.

Then Γ\Gamma is a lattice of LL, which construct a compact Clifford–Klein form. ∎

6.2 On the spaces which correspond to the triples (𝔤D,D,σ,g)(\mathfrak{g}_{D,D^{\prime}},\sigma,g)

We consider the spaces which correspond to Case (II) in Fact 1.4. These spaces are written as GD,D/HG_{D,D^{\prime}}/H for some matrices D,DSym(reg)(n,)D,D^{\prime}\in\mathrm{Sym}^{(\text{reg})}(n,{\mathbb{R}}).

6.2.1 transvection group

Note that the transvection group of (GD,D/H,σ,g)(G_{D,D^{\prime}}/H,\sigma,g) is GD,DG_{D,D^{\prime}}. We check the existence of compact Clifford–Klein forms by using Propositions 5.25 and 5.26. To do this, we introduce subsets of M(n,)M(n,{\mathbb{R}}).

Definition 6.8.

We define the following sets:

  • 𝒫D,D:={CM(n,)|C satisfies the condition (b) (i) of Proposition 5.25}={CM(n,)|At+BtC is invertible for any t}\mathcal{P}_{D,D^{\prime}}:=\left\{C\in M(n,{\mathbb{R}})\ \left|\vphantom{C\in M(n,{\mathbb{R}})}\ \text{$C$ satisfies the condition (b) (i) of Proposition~{}}\ref{Prop:mainprop}\right.\right\}\\ \hskip 28.45274pt=\left\{C\in M(n,{\mathbb{R}})\ \left|\vphantom{C\in M(n,{\mathbb{R}})}\ A_{t}+B_{t}C\text{ is invertible for any }t\in{\mathbb{R}}\right.\right\},

  • D,D:={CM(n,)|C satisfies the condition (b) (ii’) of Proposition 5.25}={CM(n,)|t0×s.t.𝔩C is Wt0-invariant and det(Wt0|𝔩C)=±1.}\mathcal{L}_{D,D^{\prime}}:=\left\{C\in M(n,{\mathbb{R}})\ \left|\vphantom{C\in M(n,{\mathbb{R}})}\ \text{$C$ satisfies the condition (b) (ii') of Proposition~{}}\ref{Prop:mainprop}\right.\right\}\\ \hskip 28.45274pt=\left\{C\in M(n,{\mathbb{R}})\ \left|\vphantom{C\in M(n,{\mathbb{R}})}\ \exists t_{0}\in{\mathbb{R}}^{\times}\ \text{s.t.}\ \mathfrak{l}^{\prime}_{C}\text{ is $W_{t_{0}}$-invariant and $\det(W_{t_{0}}|_{\mathfrak{l}^{\prime}_{C}})=\pm 1$.}\right.\right\},

  • 𝒫D,Dc:={CM(n,)|C satisfies the condition (b) (i) of Proposition 5.26}={CM(n,)|CDC=D}\mathcal{P}^{c}_{D,D^{\prime}}:=\left\{C\in M(n,{\mathbb{R}})\ \left|\vphantom{C\in M(n,{\mathbb{R}})}\ \text{$C$ satisfies the condition (b) (i) of Proposition~{}}\ref{Prop:csmainprop}\right.\right\}\\ \hskip 28.45274pt=\left\{C\in M(n,{\mathbb{R}})\ \left|\vphantom{C\in M(n,{\mathbb{R}})}\ CD^{\prime}C=D\right.\right\},

  • D,Dc:={CM(n,)|C satisfies the condition (b) (ii’) of Proposition 5.26}={CM(n,)|trDC=0}.\mathcal{L}^{c}_{D,D^{\prime}}:=\left\{C\in M(n,{\mathbb{R}})\ \left|\vphantom{C\in M(n,{\mathbb{R}})}\ \text{$C$ satisfies the condition (b) (ii') of Proposition~{}}\ref{Prop:csmainprop}\right.\right\}\\ \hskip 28.45274pt=\left\{C\in M(n,{\mathbb{R}})\ \left|\vphantom{C\in M(n,{\mathbb{R}})}\ \operatorname{tr}D^{\prime}C=0\right.\right\}.

In this subsection, we put n=2n=2 and denote the five-dimensional Heisenberg Lie algebra by 𝔥2=X1,X2,Y1,Y2,Z\mathfrak{h}_{2}=\left\langle X_{1},X_{2},Y_{1},Y_{2},Z\right\rangle_{{\mathbb{R}}} and the Heisenberg Lie group by H2H_{2}. We also use the notation W,WtM(2n,)W,W_{t}\in M(2n,{\mathbb{R}}), At,BtM(n,)A_{t},B_{t}\in M(n,{\mathbb{R}}) as in Notation 5.1.

Remark 6.9.

The condition 𝒫D,DD,D=\mathcal{P}_{D,D^{\prime}}\cap\mathcal{L}_{D,D^{\prime}}=\emptyset is a sufficient condition for the non-existence of compact Clifford–Klein forms by Proposition 5.25. In the completely solvable case, so is the condition 𝒫D,DcD,Dc=\mathcal{P}^{c}_{D,D^{\prime}}\cap\mathcal{L}_{D,D^{\prime}}^{c}=\emptyset by Proposition 5.26.

\diamondsuit The spaces which correspond to Case (II)(a) in Fact 1.4

We show that the spaces do not admit compact Clifford–Klein forms in this case.

  1. (i)

    The case (D,D)=(±diag(1,ν),diag(1,ν))(D,D^{\prime})=(\pm\operatorname{diag}\left(1,\nu\right),\operatorname{diag}\left(1,-\nu\right)) (ν>0\nu>0).
    Claim. D,D=\mathcal{L}_{D,D^{\prime}}=\emptyset.
    Take any CD,DC\in\mathcal{L}_{D,D^{\prime}} and put V1:=X1,Y1𝔥2V_{1}:=\left\langle X_{1},Y_{1}\right\rangle_{{\mathbb{R}}}\subset\mathfrak{h}_{2} and V2:=X2,Y2𝔥2V_{2}:=\left\langle X_{2},Y_{2}\right\rangle_{{\mathbb{R}}}\subset\mathfrak{h}_{2}. Note that the eigenvalues of WtW_{t} is {e±νit,e±t}\{e^{\pm\nu it},e^{\pm t}\} or {e±νt,e±it}\{e^{\pm\nu t},e^{\pm it}\}. Since 𝔩C\mathfrak{l}^{\prime}_{C} is Wt0W_{t_{0}}-invariant for some t0×t_{0}\in{\mathbb{R}}^{\times} and Wt0|V1,Wt0|V2W_{t_{0}}|_{V_{1}},W_{t_{0}}|_{V_{2}} do not have common eigenvalues, we have 𝔩C=(𝔩CV1)(𝔩CV2)\mathfrak{l}^{\prime}_{C}=(\mathfrak{l}^{\prime}_{C}\cap V_{1})\oplus(\mathfrak{l}^{\prime}_{C}\cap V_{2}) (see Note 6.10). On the other hand, V1V_{1} or V2V_{2} does not admit non-trivial Wt0W_{t_{0}}-invariant subspaces. Then we have D,D=\mathcal{L}_{D,D^{\prime}}=\emptyset.

    Note 6.10.

    For 𝕂=\mathbb{K}={\mathbb{R}} or {\mathbb{C}}, let AM(n,𝕂)A\in M(n,\mathbb{K}) be a matrix. Suppose 𝕂n=V1V2\mathbb{K}^{n}=V_{1}\oplus V_{2} is an AA-invariant decomposition such that A|V1A|_{V_{1}} and A|V2A|_{V_{2}} do not have common complex eigenvalues. For an AA-invariant subspace V𝕂nV\subset\mathbb{K}^{n}, we have V=(VV1)(VV2)V=(V\cap V_{1})\oplus(V\cap V_{2}).

  2. (ii)

    The case (D,D)=(diag(1,ν),diag(1,ν))(D,D^{\prime})=(\operatorname{diag}\left(1,-\nu\right),\operatorname{diag}\left(1,-\nu\right)) (ν>0\nu>0ν1\nu\neq 1).
    In this case, 𝔤D,D\mathfrak{g}_{D,D^{\prime}} is completely solvable (Lemma 4.5(3)).
    Claim. 𝒫D,DcD,Dc=\mathcal{P}^{c}_{D,D^{\prime}}\cap\mathcal{L}^{c}_{D,D^{\prime}}=\emptyset.
    A direct calculation leads us 𝒫D,Dc={±I2,±I1,1}\mathcal{P}^{c}_{D,D^{\prime}}=\{\pm I_{2},\pm I_{1,1}\}. Therefore, we have trDC=±(1±ν)0\operatorname{tr}D^{\prime}C=\pm(1\pm\nu)\neq 0 for any C𝒫D,DcC\in\mathcal{P}^{c}_{D,D^{\prime}} and so Claim holds.

  3. (iii)

    The case (D,D)=(diag(1,ν),diag(1,ν))(D,D^{\prime})=(\operatorname{diag}\left(-1,\nu\right),\operatorname{diag}\left(1,-\nu\right)) (ν>0\nu>0ν1\nu\neq 1).
    Claim. 𝒫D,D=\mathcal{P}_{D,D^{\prime}}=\emptyset.
    In this case, for tt\in{\mathbb{R}} we have:

    At=diag(cost,cosνt),Bt=diag(sint,sinνt).A_{t}=\operatorname{diag}\left(\cos t,\cos\nu t\right),\quad B_{t}=\operatorname{diag}\left(\sin t,-\sin\nu t\right).

    Therefore, for any C=(abcd)M(2,)C=\begin{pmatrix}a&b\\ c&d\end{pmatrix}\in M(2,{\mathbb{R}}), a direct calculation implies:

    det(At+BtC)=\displaystyle\det(A_{t}+B_{t}C)= 12(1+(adbc))cos(t+νt)+12(a+d)sin(t+νt)\displaystyle\dfrac{1}{2}\left(1+(ad-bc)\right)\cos(t+\nu t)+\dfrac{1}{2}\left(a+d\right)\sin(t+\nu t)
    +12(1(adbc))cos(tνt)+12(ad)sin(tνt).\displaystyle+\dfrac{1}{2}\left(1-(ad-bc)\right)\cos(t-\nu t)+\dfrac{1}{2}\left(a-d\right)\sin(t-\nu t).

    Then Claim is a consequence of the following:

Note 6.11.

For A,B,b,dA,B,b,d\in{\mathbb{R}}, and a,c×a,c\in{\mathbb{R}}^{\times}, put f(t):=Asin(at+b)+Bsin(ct+d)f(t):=A\sin(at+b)+B\sin(ct+d). Then we have f(t)=0f(t)=0 for some tt\in{\mathbb{R}}.

\diamondsuit The spaces which correspond to Case (II)(b) in Fact 1.4

We show that the spaces do not admit compact Clifford–Klein forms in this case.

Lemma 6.12.

Put (D,D)=(Qν,Qν)(D,D^{\prime})=(Q_{\nu},Q_{-\nu}) for ν>0\nu>0. Then we have 𝒫D,DD,D=\mathcal{P}_{D,D^{\prime}}\cap\mathcal{L}_{D,D^{\prime}}=\emptyset.

Proof..

First, for tt\in{\mathbb{R}} we have:

Wt=(coshtsinhtcoshtsinhtsinhtcoshtsinhtcosht)(cosνtsinνtcosνtsinνtsinνtcosνtsinνtcosνt).W_{t}=\begin{pmatrix}\cosh t&&&\sinh t\\ &\cosh t&\sinh t&\\ &\sinh t&\cosh t&\\ \sinh t&&&\cosh t\end{pmatrix}\begin{pmatrix}\cos\nu t&&-\sin\nu t&\\ &\cos\nu t&&\sin\nu t\\ \sin\nu t&&\cos\nu t&\\ &-\sin\nu t&&\cos\nu t\end{pmatrix}.

Set V±:=𝔩±Q0V_{\pm}:=\mathfrak{l}^{\prime}_{\pm Q_{0}}, namely, V+=X1+Y2,X2+Y1V_{+}=\left\langle X_{1}+Y_{2},X_{2}+Y_{1}\right\rangle_{{\mathbb{R}}} and V=X1Y2,X2Y1V_{-}=\left\langle X_{1}-Y_{2},X_{2}-Y_{1}\right\rangle_{{\mathbb{R}}}. Note that trW|V±=±2\operatorname{tr}W|_{V_{\pm}}=\pm 2 and that two dimensional subspace V2nV\subset{\mathbb{R}}^{2n} is WW-invariant if and only if V=V±V=V_{\pm}.

To prove this lemma, we show the following:
Claim. D,D{(abcd)SL(2,)|b+c=0}\mathcal{L}_{D,D^{\prime}}\subset\left\{\begin{pmatrix}a&b\\ c&d\end{pmatrix}\in SL(2,{\mathbb{R}})\ \left|\vphantom{\begin{pmatrix}a&b\\ c&d\end{pmatrix}\in SL(2,{\mathbb{R}})}\ b+c=0\right.\right\}.
\because) Let CD,DC\in\mathcal{L}_{D,D^{\prime}} and take t0×t_{0}\in{\mathbb{R}}^{\times} such that 𝔩C\mathfrak{l}^{\prime}_{C} is Wt0W_{t_{0}}-invariant and det(Wt0|𝔩C)=±1\det(W_{t_{0}}|_{\mathfrak{l}^{\prime}_{C}})=\pm 1.
Subclaim 1. C±Q0C\neq\pm Q_{0}.

This subclaim follows from det(Wt0|V±)=e±2t0\det(W_{t_{0}}|_{V_{\pm}})=e^{\pm 2{t_{0}}}.
Subclaim 2. νt0π\nu t_{0}\in\pi{\mathbb{Z}}.

Assume νt0π\nu t_{0}\not\in\pi{\mathbb{Z}}. Since the eigenvalues of WW are {1±νi,1±νi}\{1\pm\nu i,-1\pm\nu i\} by Lemma 4.5(2), the eigenvalues of Wt0W_{t_{0}} are distinct, and so 𝔩C\mathfrak{l}^{\prime}_{C} is WW-invariant. Since 𝔩C\mathfrak{l}^{\prime}_{C} is two-dimensional, we have 𝔩C=V+\mathfrak{l}^{\prime}_{C}=V_{+} or VV_{-}, which contradicts Subclaim 1 and so we have proven the Subclaim 2.

Since 𝔩C\mathfrak{l}^{\prime}_{C} is Wt0W_{t_{0}}-invariant, by Note 5.6 and Subclaim 2, we have:

𝔩C\mathfrak{l}^{\prime}_{C} is Wt0W_{t_{0}}-invariant (C,I2)Wt0(I2C)=O\displaystyle\iff(C,-I_{2})W_{t_{0}}\begin{pmatrix}I_{2}\\ C\end{pmatrix}=O
(C,I2)((cosht0)I4+(sinht0)(0Q0Q00))(I2C)=O\displaystyle\iff(C,-I_{2})\left((\cosh t_{0})I_{4}+(\sinh t_{0})\begin{pmatrix}0&Q_{0}\\ Q_{0}&0\end{pmatrix}\right)\begin{pmatrix}I_{2}\\ C\end{pmatrix}=O
(C,I2)(0Q0Q00)(I2C)=O\displaystyle\iff(C,-I_{2})\begin{pmatrix}0&Q_{0}\\ Q_{0}&0\end{pmatrix}\begin{pmatrix}I_{2}\\ C\end{pmatrix}=O
CQ0C=Q0\displaystyle\iff CQ_{0}C=Q_{0}
C{±Q0}{(abcd)SL(2,)|b+c=0}.\displaystyle\iff C\in\{\pm Q_{0}\}\sqcup\left\{\begin{pmatrix}a&b\\ c&d\end{pmatrix}\in SL(2,{\mathbb{R}})\ \left|\vphantom{\begin{pmatrix}a&b\\ c&d\end{pmatrix}\in SL(2,{\mathbb{R}})}\ b+c=0\right.\right\}.

By Subclaim 1, we have shown Claim.

Finally, we prove the lemma. Take any C=(abcd)D,DC=\begin{pmatrix}a&b\\ c&d\end{pmatrix}\in\mathcal{L}_{D,D^{\prime}}. By a direct calculation, we have:

det(At+BtC)\displaystyle\det(A_{t}+B_{t}C) =(b+c)2sinh2t+(ad)2sin2νt+cos2νt(cosh2t(adbc)sinh2t)\displaystyle=\dfrac{(b+c)}{2}\sinh 2t+\dfrac{(a-d)}{2}\sin 2\nu t+\cos 2\nu t(\cosh^{2}t-(ad-bc)\sinh^{2}t)
=cos2νt+ad2sin2νt.\displaystyle=\cos 2\nu t+\frac{a-d}{2}\sin 2\nu t.

Therefore, det(At+BtC)=0\det(A_{t}+B_{t}C)=0 for some tt\in{\mathbb{R}}, so we obtain 𝒫D,DD,D=\mathcal{P}_{D,D^{\prime}}\cap\mathcal{L}_{D,D^{\prime}}=\emptyset

\diamondsuit The spaces which correspond to Case (II)(c) in Fact 1.4

In this case, the spaces do not admit compact Clifford–Klein forms.

  1. (i)

    The case (D,D)=((ε110),(011ε))(D,D^{\prime})=\left(\begin{pmatrix}\varepsilon&-1\\ -1&0\end{pmatrix},\begin{pmatrix}0&-1\\ -1&\varepsilon\end{pmatrix}\right), where ε=±1\varepsilon=\pm 1.
    In this case, 𝔤D,D\mathfrak{g}_{D,D^{\prime}} is completely solvable (Lemma 4.5(3)).

    Claim. 𝒫D,DcD,Dc=\mathcal{P}^{c}_{D,D^{\prime}}\cap\mathcal{L}^{c}_{D,D^{\prime}}=\emptyset.
    By a direct calculation, we have 𝒫D,Dc={±Q0}\mathcal{P}^{c}_{D,D^{\prime}}=\{\pm Q_{0}\}. Thus, the claim follows from trDC=±2\operatorname{tr}D^{\prime}C=\pm 2 for C𝒫D,DcC\in\mathcal{P}_{D,D^{\prime}}^{c}.

  2. (ii)

    The case (D,D)=((ε110),(011ε))(D,D^{\prime})=\left(\begin{pmatrix}\varepsilon&-1\\ -1&0\end{pmatrix},\begin{pmatrix}0&1\\ 1&-\varepsilon\end{pmatrix}\right), where ε=±1\varepsilon=\pm 1.
    Claim. 𝒫D,D=\mathcal{P}_{D,D^{\prime}}=\emptyset.
    We have:

    At=(cost0εtsintcost),Bt=(0sintsintεtcost).A_{t}=\begin{pmatrix}\cos t&0\\ \varepsilon t\sin t&\cos t\end{pmatrix},\quad B_{t}=\begin{pmatrix}0&\sin t\\ \sin t&-\varepsilon t\cos t\end{pmatrix}.

    Let C=(abcd)M(2,)C=\begin{pmatrix}a&b\\ c&d\end{pmatrix}\in M(2,{\mathbb{R}}). If d=0d=0, we have det(At+BtC)=(cost+csint)(cost+bsint)\det(A_{t}+B_{t}C)=(\cos t+c\sin t)(\cos t+b\sin t). Hence det(At+BtC)=0\det(A_{t}+B_{t}C)=0 holds for some tt\in{\mathbb{R}} and so C𝒫D,DC\not\in\mathcal{P}_{D,D^{\prime}}. Then we assume d0d\neq 0. For mm\in{\mathbb{Z}}, we have det(A2mπ+B2mπC)=12mπdε\det(A_{2m\pi}+B_{2m\pi}C)=1-2m\pi d\varepsilon. Then there exist t1,t2t_{1},t_{2}\in{\mathbb{R}} satisfying det(At1+Bt1C)>0,det(At2+Bt2C)<0\det(A_{t_{1}}+B_{t_{1}}C)>0,\ \det(A_{t_{2}}+B_{t_{2}}C)<0. Hence we have det(At+BtC)=0\det(A_{t}+B_{t}C)=0 for some tt\in{\mathbb{R}} by the intermediate value theorem and so C𝒫D,DC\not\in\mathcal{P}_{D,D^{\prime}}. Therefore we have 𝒫D,D=\mathcal{P}_{D,D^{\prime}}=\emptyset.

\diamondsuit The spaces which correspond to Case (II)(b) in Fact 1.4

In this case, the spaces admit compact Clifford–Klein forms.

  1. (i)

    The case (D,D)=(I1,1,I1,1)(D,D^{\prime})=(I_{1,1},I_{1,1}).
    In this case, 𝔤D,D\mathfrak{g}_{D,D^{\prime}} is completely solvable (Lemma 4.5(3)), so we are enough to show that the space satisfies the condition (b) in Proposition 5.26. Set (C,w):=(I2,0)M(2,)×𝔥(C,w):=(I_{2},0)\in M(2,{\mathbb{R}})\times\mathfrak{h}, then the condition (b)(i) CDC=DCD^{\prime}C=D is clear. By Note 5.19, we have LC,wSML_{C,w}\simeq S_{M} for M:=diag(1,1,0)M:=\operatorname{diag}\left(1,-1,0\right). By Example 5.24, LC,wL_{C,w} admits a uniform lattice and so the condition (b)(ii) holds.

  2. (ii)

    The case (D,D)=(I1,1,I1,1)(D,D^{\prime})=(-I_{1,1},I_{1,1}).
    It is enough to show the conditions (b) in Proposition 5.25. Put C:=Q0C:=Q_{0}. Since we have At=(cost)I2A_{t}=(\cos t)I_{2} and Bt=(sint)I1,1B_{t}=(\sin t)I_{1,1}, we get det(At+BtC)=sin2t+cos2t=1\det(A_{t}+B_{t}C)=\sin^{2}t+\cos^{2}t=1, which implies the condition (b)(i). By Note 5.7, we have LC3L_{C}\simeq{\mathbb{R}}^{3} and =id\mathcal{I}_{\ell}=\operatorname{id}, where =(2π,e)GD,D\ell=(2\pi,e)\in G_{D,D^{\prime}}. Then the subgroup LCL_{C} has an \mathcal{I}_{\ell}-invariant uniform lattice Γ3\Gamma\simeq{\mathbb{Z}}^{3}. Then the condition (b)(ii) holds.

6.2.2 isometry group

First, we calculate isometry group of the symmetric spaces.

Proposition 6.13.

For matrices D,DSym(reg)(2,)D,D^{\prime}\in\mathrm{Sym}^{(\text{reg})}(2,{\mathbb{R}}), assume the signature of DD^{\prime} is (1,1)(1,1). Then the following list gives a complete class representatives of symmetric triple (𝔤D,D,σ,g)(\mathfrak{g}_{D,D^{\prime}},\sigma,g).

  1. (1)

    (D,D)=(±diag(1,ν),diag(1,ν))(ν>0)(D,D^{\prime})=(\pm\operatorname{diag}\left(1,\nu\right),\operatorname{diag}\left(1,-\nu\right))\quad(\nu>0),
    (D,D)=(±diag(1,ν),diag(1,ν))(ν>0,ν1)(D,D^{\prime})=(\pm\operatorname{diag}\left(1,-\nu\right),\operatorname{diag}\left(1,-\nu\right))\quad(\nu>0,\ \nu\neq 1),

  2. (2)

    (D,D)=(Qν,Qν)(ν>0)(D,D^{\prime})=\left(Q_{\nu},Q_{-\nu}\right)\quad(\nu>0),

  3. (3)

    (D,D)=((±1110),(011±1)),((±1110),(0111))(D,D^{\prime})=\left(\begin{pmatrix}\pm 1&-1\\ -1&0\end{pmatrix},\begin{pmatrix}0&-1\\ -1&\pm 1\end{pmatrix}\right),\ \left(\begin{pmatrix}\pm 1&-1\\ -1&0\end{pmatrix},\begin{pmatrix}0&1\\ 1&\mp 1\end{pmatrix}\right),

  4. (4)

    (D,D)=(±I1,1,I1,1)(D,D^{\prime})=(\pm I_{1,1},I_{1,1}).

Proposition 6.14.

For a pseudo-Riemannian symmetric space (GD,D/H,σ,g)(G_{D,D^{\prime}}/H,\sigma,g), we have:

Gi{(I2,I1,1×/2)Gt(the spaces in the list (1))(I2×/2)Gt(the spaces in the list (2) and (3))(O(1,1)×/2)Gt(the spaces in the list (4))G_{\text{i}}\simeq\begin{cases}(\langle{-I_{2},I_{1,1}}\rangle\times{\mathbb{Z}}/2{\mathbb{Z}})\ltimes G_{\text{t}}\quad\text{(the spaces in the list (1))}\\ (\langle{-I_{2}}\rangle\times{\mathbb{Z}}/2{\mathbb{Z}})\ltimes G_{\text{t}}\quad\text{(the spaces in the list (2) and (3))}\\ (O(1,1)\times{\mathbb{Z}}/2{\mathbb{Z}})\ltimes G_{\text{t}}\quad\text{(the spaces in the list (4))}\\ \end{cases}
Proof..

By Lemma 2.15, it follows from Lemma 6.17. ∎

In the following, we calculate the group HiH_{\text{i}}.

Lemma 6.15.

Using the basis (W,X1,X2,Y1,Y2,Z)(W,X_{1},X_{2},Y_{1},Y_{2},Z) of 𝔤t𝔥2\mathfrak{g}_{\text{t}}\simeq{\mathbb{R}}\ltimes\mathfrak{h}_{2}, we have:

Aut(𝔤D,D,σ,g)={(kvPkPT1swk)GL(6,)|PO(1,1)k{±1},v2PDPT=D},\operatorname{Aut}(\mathfrak{g}_{D,D^{\prime}},\sigma,g)=\left\{\begin{pmatrix}k&&&\\ v&P&&\\ &&{kP^{T}}^{-1}&\\ s&w&&k\end{pmatrix}\in GL(6,{\mathbb{R}})\ \left|\vphantom{\begin{pmatrix}k&&&\\ v&P&&\\ &&{kP^{T}}^{-1}&\\ s&w&&k\end{pmatrix}\in GL(6,{\mathbb{R}})}\ \begin{gathered}P\in O(1,1)\\ k\in\{\pm 1\},\ v\in{\mathbb{R}}^{2}\\ PDP^{T}=D\end{gathered}\right.\right\},

where s:=(k/2)vTI1,1v,w:=kvTI1,1Ps:=(k/2)v^{T}I_{1,1}v,\ w:=kv^{T}I_{1,1}P.

Proof..

Let ϕAut(𝔤t,σ,g)\phi\in\operatorname{Aut}(\mathfrak{g}_{\text{t}},\sigma,g). We use (W,X1,X2,Y1,Y2,Z)(W,X_{1},X_{2},Y_{1},Y_{2},Z) as the basis of 𝔤t\mathfrak{g}_{\text{t}}. Then the representation matrix of ϕ\phi is:

ϕ=(1vPQsw2)GL(6,),\phi=\begin{pmatrix}\ell_{1}&&&\\ v&P&&\\ &&Q&\\ s&w&&\ell_{2}\end{pmatrix}\in GL(6,{\mathbb{R}}),

where v,wT2,s,1,2,P,QGL(2,)v,w^{T}\in{\mathbb{R}}^{2},\ s,\ell_{1},\ell_{2}\in{\mathbb{R}},\ P,Q\in GL(2,{\mathbb{R}}). Then we have:

g is ϕ invariantϕTgϕ=g\displaystyle\text{$g$ is $\phi$ invariant}\iff\phi^{T}g\phi=g {PTI1,1P=I1,1,12=1,vTI1,1P=1w,vTI1,1v=21s,\displaystyle\iff\begin{cases}P^{T}I_{1,1}P=I_{1,1},\\ \ell_{1}\ell_{2}=1,\\ v^{T}I_{1,1}P=\ell_{1}w,\\ v^{T}I_{1,1}v=2\ell_{1}s,\\ \end{cases} (4)
[ϕ(Xi),ϕ(Yj)]=ϕ([Xi,Yj])(i,j=1,2)\displaystyle[\phi(X_{i}),\phi(Y_{j})]=\phi([X_{i},Y_{j}])\ (i,j=1,2) PTQ=2I2,\displaystyle\iff P^{T}Q=\ell_{2}I_{2}, (5)
[ϕ(W),ϕ(Yi)]=ϕ([W,Yi])(i=1,2)\displaystyle[\phi(W),\phi(Y_{i})]=\phi([W,Y_{i}])\ (i=1,2) I1,1P=1QI1,1,vTQ=wI1,1,\displaystyle\iff I_{1,1}P=\ell_{1}QI_{1,1},\ v^{T}Q=wI_{1,1}, (6)
[ϕ(W),ϕ(Xi)]=ϕ([W,Xi])(i=1,2)\displaystyle[\phi(W),\phi(X_{i})]=\phi([W,X_{i}])\ (i=1,2) 1PD=DQ,\displaystyle\iff\ell_{1}PD=DQ, (7)

By the first identity of (0.1), we have PO(1,1)P\in O(1,1). On the other hand Q=2PT1Q=\ell_{2}{P^{T}}^{-1} by (5). By the determinant relation of (0.3) and (0.4), we have 1=±1\ell_{1}=\pm 1. By the second identity of (0.1), we have 1=2=±1\ell_{1}=\ell_{2}=\pm 1. We put k:=1k:=\ell_{1}, then s:=(k/2)vTI1,1vs:=(k/2)v^{T}I_{1,1}v and w=kvTI1,1Pw=kv^{T}I_{1,1}P follows from (0.1). Moreover, a direct calculation leads us ϕAut(𝔤t,σ,g)\phi\in\operatorname{Aut}(\mathfrak{g}_{\text{t}},\sigma,g) if ϕ\phi satisfies the conditions of Lemma 6.15. ∎

Lemma 6.16.

By the identification HiAut(𝔤t,σ,g)H_{\text{i}}\simeq\operatorname{Aut}(\mathfrak{g}_{\text{t}},\sigma,g), we have:

Ht={(1I1,1yI2I2(yTI1,1y)/2yT1)|yn}H_{\text{t}}=\left\{\begin{pmatrix}1&&&\\ -I_{1,1}y&I_{2}&&\\ &&I_{2}&\\ (y^{T}I_{1,1}y)/2&-y^{T}&&1\end{pmatrix}\ \left|\vphantom{\begin{pmatrix}1&&&\\ -I_{1,1}y&I_{2}&&\\ &&I_{2}&\\ (y^{T}I_{1,1}y)/2&-y^{T}&&1\end{pmatrix}}\ y\in{\mathbb{R}}^{n}\right.\right\}
Proof..

Note that Ht={exp(i=1nyiYi)|yn}H_{\text{t}}=\left\{\exp(\sum_{i=1}^{n}y_{i}Y_{i})\ \left|\vphantom{\exp(\sum_{i=1}^{n}y_{i}Y_{i})}\ y\in{\mathbb{R}}^{n}\right.\right\}. The adjoint action of exp(i=1nyiYi)\exp(\sum_{i=1}^{n}y_{i}Y_{i}) is:

ad(Yi)=(0I1,1ei00eiT0){\rm ad}(Y_{i})=\begin{pmatrix}0&&&\\ -I_{1,1}e_{i}&0&&\\ &&0&\\ &-e_{i}^{T}&&0\end{pmatrix}

This lemma follows from Lemma 2.17. ∎

Lemma 6.17.

We put A:={diag(k,P,kPT1,k)|PO(1,1),k{±1},PDPT=D}HiA:=\left\{\operatorname{diag}\left(k,P,{kP^{T}}^{-1},k\right)\ \left|\vphantom{\operatorname{diag}\left(k,P,{kP^{T}}^{-1},k\right)}\ P\in O(1,1),k\in\{\pm 1\},PDP^{T}=D\right.\right\}\subset H_{\text{i}}. Then AA is a closed subgroup of HiH_{\text{i}} and we have Hi=AHtH_{\text{i}}=A\ltimes H_{\text{t}}.

Proof..

By a direct calculation, we see AA is a closed subgroup of HiH_{\text{i}}. Note that HtH_{\text{t}} is a closed normal subgroup of HiH_{\text{i}}. By Lemma 6.16, we have HtA={e},Hi=AHtH_{\text{t}}\cap A=\{e\},\ H_{\text{i}}=AH_{\text{t}}. Then this lemma follows from Note 2.20. ∎

Lemma 6.18.

We have:

{diag(k,P,kPT1,k)|PO(1,1),PDPT=D,k{±1}}\displaystyle\left\{\operatorname{diag}\left(k,P,{kP^{T}}^{-1},k\right)\ \left|\vphantom{\operatorname{diag}\left(k,P,{kP^{T}}^{-1},k\right)}\ P\in O(1,1),PDP^{T}=D,k\in\{\pm 1\}\right.\right\}
{/2×O(1,1)(the spaces in the list (1))/2×I2(the spaces in the list (2) and (3))/2×I2,I1,1(the spaces in the list (4))\displaystyle\simeq\begin{cases}{\mathbb{Z}}/2{\mathbb{Z}}\times O(1,1)\quad\text{(the spaces in the list (1))}\\ {\mathbb{Z}}/2{\mathbb{Z}}\times\langle{-I_{2}}\rangle\quad\text{(the spaces in the list (2) and (3))}\\ {\mathbb{Z}}/2{\mathbb{Z}}\times\langle{-I_{2},I_{1,1}}\rangle\quad\text{(the spaces in the list (4))}\end{cases}
Proof..

In the case D=±I1,1D=\pm I_{1,1}, it is clear. In the case DI1,1D\neq I_{1,1}, it follows from PDPT=DP=±I2PDP^{T}=D\Leftrightarrow P=\pm I_{2}. ∎

Corollary 6.19.

The space Gt/HG_{\text{t}}/H admits compact Clifford–Klein forms if and only if so does Gi/HG_{\text{i}}/H.

Proof..

In the case D=I1,1D=I_{1,1}, the space Gt/HtG_{\text{t}}/H_{\text{t}} admits compact Clifford–Klein forms. In the case D±I1,1D\neq\pm I_{1,1}, we have Lie(Gi)=Lie(Gt)\operatorname{Lie}(G_{\text{i}})=\operatorname{Lie}(G_{\text{t}}) by Proposition 6.14, so it does not admit compact Clifford–Klein forms by Note 2.33 and . ∎

7 Kobayashi’s conjecture about standard quotients

There have been attempts to extend Kobayashi’s theory on discontinuous groups for reductive cases [17-23] to non-reductive cases such as Baklouti-Kédim[1], Kath-Olbrich[16], Kobayashi-Nasrin[24], Lipsman[30], Nasrin[31], Yoshino[36] and so on. In this section, we examine a ‘solvable analogue’ of Kobayashi’s conjecture (Conjecture 1.6) and see an evidence that the assumption ‘reductive type’ in Kobayashi’s conjecture is crucial.

Example 7.1.

We put n=3n=3 and (D,D):=(diag(1,1,2),diag(1,1,2))(D,D^{\prime}):=(\operatorname{diag}\left(-1,-1,2\right),\operatorname{diag}\left(1,1,-2\right)). Then GD,D/HG_{D,D^{\prime}}/H admits compact Clifford–Klein forms and does not admit constructors.

Proof..

First, we check GD,D/HG_{D,D^{\prime}}/H does not admit constructors. Assume GD,D/HG_{D,D^{\prime}}/H admits constructors. Then there exists CM(3,)C\in M(3,{\mathbb{R}}) such that CDC=DCD^{\prime}C=D by Proposition 5.14. However, we obtain (detC)2=1(\det C)^{2}=-1, which contradicts CM(3,)C\in M(3,{\mathbb{R}}).

Next, we check GD,D/HG_{D,D^{\prime}}/H admits compact Clifford–Klein forms by using Proposition 5.25. We set:

C(010102010).C\coloneqq\begin{pmatrix}0&1&0\\ 1&0&-2\\ 0&-1&0\end{pmatrix}.

It is enough to check that the conditions (b)(i) and (ii) in Proposition 5.25. A direct calculation leads us that At=diag(cost,cost,cos2t)A_{t}=\operatorname{diag}\left(\cos t,\cos t,\cos 2t\right), Bt=diag(sint,sint,sin2t)B_{t}=\operatorname{diag}\left(\sin t,\sin t,-\sin 2t\right), det(At+BtC)=cos2(2t)+sin2(2t)=1\det(A_{t}+B_{t}C)=\cos^{2}(2t)+\sin^{2}(2t)=1 and so the condition (i) holds. Set t02πt_{0}\coloneqq 2\pi\in{\mathbb{R}} and :=(t0,e)GD,D\ell:=(t_{0},e)\in G_{D,D^{\prime}}. Then we have Wt0=I6W_{t_{0}}=I_{6} and so =id\mathcal{I}_{\ell}=\operatorname{id}. Then LCH1×L_{C}\simeq H_{1}\times{\mathbb{R}} (Note 5.7) has an \mathcal{I}_{\ell}-invariant uniform lattice ΓH1()×\Gamma\simeq H_{1}({\mathbb{Z}})\times{\mathbb{Z}}, and so the condition (ii) is satisfied. ∎

Acknowledgments

The author would like to thank his supervisors Professor Toshiyuki Kobayashi and Professor Taro Yoshino for many constructive comments. This work was supported by the Program for Leading Graduate Schools, MEXT, Japan.

References

  • [1] A. Baklouti and I. Kédim. On non-abelian discontinuous subgroups acting on exponential solvable homogeneous spaces. Int. Math. Res. Not. IMRN, (7):1315–1345, 2010.
  • [2] A Baklouti and F Khlif. Proper actions and some exponential solvable homogeneous spaces. Internat. J. Math., 16(9):941–955, 2005.
  • [3] Y. Benoist. Actions propres sur les espaces homogènes réductifs. Ann. of Math. (2), 144(2):315–347, 1996.
  • [4] M. Berger. Les espaces symétriques noncompacts. Ann. Sci. École Norm. Sup. (3), 74:85–177, 1957.
  • [5] A. Borel. Compact Clifford-Klein forms of symmetric spaces. Topology, 2:111–122, 1963.
  • [6] N. Bourbaki. Éléments de mathématique. Fasc. XXXVII. Groupes et algèbres de Lie. Chapitre II: Algèbres de Lie libres. Chapitre III: Groupes de Lie. Hermann, Paris, 1972. Actualités Scientifiques et Industrielles, No. 1349.
  • [7] M. Cahen and M. Parker. Pseudo-Riemannian symmetric spaces. Memoirs of the AMS, vol.24, No.229, American Mathematical Society, 1980.
  • [8] M. Cahen and N. Wallach. Lorentzian symmetric spaces. Bull. Amer. Math. Soc., 76:585–591, 1970.
  • [9] É. Cartan. Sur une classe remarquable d’espaces de Riemann. Bull. Soc. Math. France, 54:214–264, 1926.
  • [10] É. Cartan. Sur une classe remarquable d’espaces de Riemann. II. Bull. Soc. Math. France, 55:114–134, 1927.
  • [11] L. J. Corwin and F. P. Greenleaf. Representations of nilpotent Lie groups and their applications. Part I, volume 18 of Cambridge Studies in Advanced Mathematics. Cambridge University Press, Cambridge, 1990. Basic theory and examples.
  • [12] W. M. Goldman. Nonstandard Lorentz space forms. J. Differential Geom., 21(2):301–308, 1985.
  • [13] G. Hochschild. The structure of Lie groups. Holden-Day, Inc., San Francisco-London-Amsterdam, 1965.
  • [14] F. Kassel and T. Kobayashi. Poincaré series for non-Riemannian locally symmetric spaces. Adv. Math., 287:123–236, 2016.
  • [15] I. Kath and M. Olbrich. On the structure of pseudo-Riemannian symmetric spaces. Transform. Groups, 14(4):847–885, 2009.
  • [16] I. Kath and M. Olbrich. Compact quotients of Cahen-Wallach spaces. Mem. Amer. Math. Soc., 262(1264):v+84, 2019.
  • [17] T. Kobayashi. Proper action on a homogeneous space of reductive type. Math. Ann., 285(2):249–263, 1989.
  • [18] T. Kobayashi. Discontinuous groups acting on homogeneous spaces of reductive type. In Representation theory of Lie groups and Lie algebras (Fuji-Kawaguchiko, 1990), pages 59–75. World Sci. Publ., River Edge, NJ, 1992.
  • [19] T. Kobayashi. A necessary condition for the existence of compact Clifford-Klein forms of homogeneous spaces of reductive type. Duke Math. J., 67(3):653–664, 1992.
  • [20] T. Kobayashi. On discontinuous groups acting on homogeneous spaces with noncompact isotropy subgroups. J. Geom. Phys., 12(2):133–144, 1993.
  • [21] T. Kobayashi. Criterion for proper actions on homogeneous spaces of reductive groups. J. Lie Theory, 6(2):147–163, 1996.
  • [22] T. Kobayashi. Discontinuous groups and Clifford-Klein forms of pseudo-Riemannian homogeneous manifolds. In Algebraic and analytic methods in representation theory (Sønderborg, 1994), volume 17 of Perspect. Math., pages 99–165. Academic Press, San Diego, CA, 1997.
  • [23] T. Kobayashi. Deformation of compact Clifford-Klein forms of indefinite-Riemannian homogeneous manifolds. Math. Ann., 310(3):395–409, 1998.
  • [24] T. Kobayashi and S. Nasrin. Deformation of properly discontinuous actions of k\mathbb{Z}^{k} on k+1\mathbb{R}^{k+1}. Internat. J. Math., 17(10):1175–1193, 2006.
  • [25] T. Kobayashi and K. Ono. Note on Hirzebruch’s proportionality principle. J. Fac. Sci. Univ. Tokyo Sect. IA Math., 37(1):71–87, 1990.
  • [26] T. Kobayashi and T. Yoshino. Compact Clifford-Klein forms of symmetric spaces—revisited. Pure Appl. Math. Q., 1(3, Special Issue: In memory of Armand Borel. Part 2):591–663, 2005.
  • [27] T. Kobayashi and T Yoshino. Uniform lattices for tangential symmetric spaces. unpublished manuscript, 2007.
  • [28] R. S. Kulkarni. Proper actions and pseudo-Riemannian space forms. Adv. in Math., 40(1):10–51, 1981.
  • [29] F. Labourie. Quelques résultats récents sur les espaces localement homogènes compacts. In Manifolds and geometry (Pisa, 1993), Sympos. Math., XXXVI, pages 267–283. Cambridge Univ. Press, Cambridge, 1996.
  • [30] R. L. Lipsman. Proper actions and a compactness condition. J. Lie Theory, 5(1):25–39, 1995.
  • [31] S. Nasrin. Criterion of proper actions for 2-step nilpotent Lie groups. Tokyo J. Math., 24(2):535–543, 2001.
  • [32] Katsumi Nomizu. Invariant affine connections on homogeneous spaces. Amer. J. Math., 76:33–65, 1954.
  • [33] M. S. Raghunathan. Discrete subgroups of Lie groups. Springer-Verlag, New York-Heidelberg, 1972. Ergebnisse der Mathematik und ihrer Grenzgebiete, Band 68.
  • [34] M. Saito. Sur certains groupes de Lie résolubles. II. Sci. Papers Coll. Gen. Ed. Univ. Tokyo, 7:157–168, 1957.
  • [35] D. Witte. Superrigidity of lattices in solvable Lie groups. Invent. Math., 122(1):147–193, 1995.
  • [36] T. Yoshino. Criterion of proper actions for 3-step nilpotent Lie groups. Internat. J. Math., 18(7):783–795, 2007.