Four-dimensional compact Clifford–Klein forms of pseudo-Riemannian symmetric spaces with signature (2, 2)
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 and 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 , its closed subgroup assume a discrete subgroup acts on properly discontinuously. We say the quotient space 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 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 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 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 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 is a normal subgroup of the isometry group . Roughly speaking, the transvection group is the “smallest” group which acts on the symmetric space transitively. For example,
-
•
for Euclid space , we have and ,
-
•
for the sphere , we have and .
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 .
For the first step of the problem, we classify four-dimensional indecomposable symmetric spaces 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 .
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 uniquely corresponds to a symmetric triple , they classified its symmetric triples.
Fact 1.4 ([15, Theorem 7.1]).
Let be the symmetric triple corresponding to a 1-connected four-dimensional reducible and indecomposable pseudo-Riemannian symmetric spaces with signature , and assume that its transvection group is solvable. Then the symmetric triple is isometric to one in the following list.
-
Case (I)
Nilpotent symmetric triples (see Definition 4.15),
- Case (II)
Here, is a 3-step nilpotent Lie algebra given in Definition 4.15. On the other hand, denotes an extension of the Heisenberg Lie algebra by (see Definition 4.4). The isometry class of the triple has continuous parameters (see also Figure 4.1).
Our main result in this paper is:
Theorem 1.5.
Let be a four-dimensional reducible and indecomposable 1-connected pseudo-Riemannian symmetric space with signature whose transvection group is solvable. We denote by and its transvection group and isometry group respectively, and set closed subgroups and by .
-
(1)
admits compact Clifford–Klein forms if and only if its corresponding symmetric triple is .
-
(2)
admits compact Clifford–Klein forms if and only if its corresponding symmetric triple is or .
symmetric triple | ||
---|---|---|
Never | ||
Always |
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 of reductive type admits compact Clifford–Klein forms, then admits a reductive constructor, that is, there exists a subgroup which is reductive in and acts properly and cocompactly on .
Remark that the conjecture above does not assert that for a compact Clifford–Klein form of a homogeneous space of reductive type, there exists a reductive constructor containing cocompactly. A Clifford–Klein form is standard ([14]) if contained in some reductive subgroup of acting properly on . 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 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 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.
-
•
:= ,
-
•
: the identity element of a group,
-
•
: the inner automorphism with respect to an element of a group,
-
•
for subsets of a group,
-
•
: the center of a group ,
-
•
: the centralizer of an element in a group ,
-
•
: the normalizer of a subgroup in a group ,
-
•
: the derivation algebra of a Lie algebra ,
-
•
,
-
•
: the transposed matrix of a matrix ,
-
•
,
-
•
,
-
•
,
-
•
.
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 be a Lie algebra, an involution on and an (indefinite) inner product on . We say is a metric Lie algebra with involution if , and are mutually compatible, that is, satisfy the following conditions:
-
(1)
the inner product is -invariant,
-
(2)
the inner product is -invariant, namely,
Definition 2.3 (symmetric triple, [8, 15]).
A metric Lie algebra with involution is called a symmetric triple if the subspace satisfies .
Definition 2.4 (homomorphism on symmetric triple, [8, 15]).
For two symmetric triples and , a Lie algebra homomorphism is said to be a homomorphism of symmetric triple if is compatible with the involutions and the inner products, that is, satisfies the following conditions:
-
(1)
,
-
(2)
().
Note 2.5.
Let be a Lie algebra and its subalgebra. The following correspondence is bijective:
Fact 2.6 ([7]).
Let be a Lie algebra and its involution. Put and . If , then the following restriction is bijective:
Especially, any -invariant inner product is also -invariant in this case.
Note 2.7.
Definition 2.8 ([8, 15]).
For a symmetric triple , we call the signature of (on ) 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 -connected pseudo-Riemannian symmetric spaces with signature and the isomorphic classes of symmetric triples with signature . Let be a symmetric triple and its corresponding pseudo-Riemannian symmetric space, then is the Lie algebra of the transvection group of . The 1-connected Lie group with Lie algebra is the transvection group of .
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 is reducible if the isotropy representation is reducible for and . A 1-connected pseudo-Riemannian symmetric space is said to be reducible if its triple is reducible.
Definition 2.11 ([8, 15]).
For two symmetric triples and , the triple is also a symmetric triple. We say is the direct sum of and . 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 be a 1-connected pseudo-Riemannian symmetric space, and the corresponding symmetric triple. Then the following correspondence is one to one.
where and are the corresponding 1-connected pseudo-Riemannian symmetric spaces of symmetric triples and , 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 , we set the origin point . We put
-
•
: the isometry group, : the stabilizer of ,
-
•
: the transvection group, : the stabilizer of ,
-
•
: the Lie algebra of and its eigenspace deconposition with respect to .
-
•
: the Lie algebra of and its eigenspace deconposition with respect to .
Lemma 2.15.
For a 1-connected pseudo Riemannian symmetric space , the map is surjective group homomorphism and its kernel is:
Especially, we have a Lie group isomorphism . Moreover, if there exists a closed subgroup satisfying , we have a Lie group isomorphism .
Proof..
Since the -action on is transitive, for any , there exists satisfying . Therefore, the map is surjective group homomorphism. The latter statement follows from the following Note with and . ∎
Note 2.16.
Let and be Lie groups. Assume that there exists a map such that is a -invariant closed subgroup. We set by . Then is a Lie group homomorphism. Moreover the following map is also a Lie group homomorphism and its kernel is .
Especially, we have .
Lemma 2.17.
Let is the symmetric triple which corresponds to 1-connected pseudo-Riemannian symmetric space . Then the map is a Lie group isomorphism.
Proof..
First, we check that is well-defined. Since is a normal subgroup, we have . By , we have . Therefore and are commutative, namely, preserves the decomposition and . On the other hand, preserves the inner product on . Since it is commutative with and a Lie algebra homomorphism, it also preserves the inner product on , and so on . Then we have .
We set a map as follows. Let be the symmetric triple which corresponds to the symmetric triple . Put the induced Lie group homomorphism by the Lie group homomorphism . Note that is a Lie group isomorphism. Since preserves , it induces , and it is an isometry by Note 2.18. Therefore we have .
Next we check that is the inverse map of . First we show . Let . It is enough to show is coincides to . It follows from , and and make the following diagram commutative:
∎
Then we check . Let and . We are enough to show , which is equivalent to . For , we have:
Here, follows from Note 2.19.
Note 2.18.
Let be a Lie group, its closed subgroup and a -invariant inner product on . Suppose that a group isomorphism preserves and put the induced map. If preserves inner product on , then is an isometry.
Note 2.19.
Let be an automorphism on . Put and . Then we have , where is the induced map by and we identify .
Note 2.20.
Let be a Lie group, a closed subgroup and a normal subgroup. If and , then we have , where acts on as an inner automorphism.
2.4 Clifford–Klein forms
Definition 2.21 ([17]).
Let be a Lie group, its closed subgroup, and its discrete subgroup. Assume the -action on is (fixed point) free and properly discontinuous. Then the quotient space has the unique manifold structure such that the natural surjection is a -covering map. The manifold is said to be a Clifford–Klein form of .
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 acts on a locally compact space . The -action is said to be proper if is compact for any compact subset .
It is easy to check the following:
Note 2.23.
In the setting of Definition 2.21, if the -action on is proper, any -orbit is closed in .
Fact 2.24 ([17]).
Let be a locally compact group and a locally compact space. Assume acts on and is a uniform lattice (cocompact discrete subgroup) of . Then the following statements hold.
-
(1)
The -action on is cocompact if and only if so is the -action.
-
(2)
The -action on is properly discontinuous if and only if the -action is proper.
We recall some definitions and properties.
Definition 2.25 ([18, Definition 6], [21, Definition 2.1.1]).
Let be a locally compact group, and and its subsets.
-
(1)
We say the pair is proper in , denoted by in , if the set is relatively compact in for any compact set .
-
(2)
We say the pair has the property (CI) in , if the set is relatively compact in for any .
-
(3)
We say the pair is free if the condition holds for any .
-
(4)
We denote by in the existence of a compact set satisfying and .
Remark 2.26.
Property 2.27 ([18, 21]).
Let be a locally compact group, and and its subsets.
-
(1)
The pair is proper (resp. has the property (CI), is free) in if and only if so is in .
-
(2)
The relation is an equivalence relation.
-
(3)
If in , then if and only if in .
-
(4)
If the pair is proper, then has the property (CI).
Property 2.28 ([21, Observation 2.13]).
Let be a locally compact group, and and its closed subgroups.
-
(1)
The -action on is proper if and only if in .
-
(2)
The -action on is free if and only if the pair is free in .
Property 2.29.
Let be a locally compact group and its closed normal subgroup, and , closed subgroups of . We denote by , and the image of , and , respectively, by the natural projection . Then we have:
-
(1)
If is compact and is discrete, is also discrete.
-
(2)
If is compact, the condition in implies in .
-
(3)
If is compact, so is .
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.
Lemma 2.30.
Let and be two closed cones in . Then the following conditions are equivalent:
-
(a)
the pair has the (CI) property, namely, ,
-
(b)
the pair is proper in .
Proof..
Since the implication (b)(a) is easy, we prove the implication (a)(b). We take any and denote by the closed ball in . It is enough to show that is relatively compact. Set as the distance between and , where is the unit sphere in . Take any , then we have , and so . ∎
Lemma 2.31.
Let be a locally compact group and its closed normal subgroup. Let and be subsets of , and a subset of satisfying in . Set , then the following conditions are equivalent:
-
(a)
in for any compact set ,
-
(b)
in ,
-
(c)
in .
Proof..
Since the implications (c)(b)(a) are easy, we prove the implication (a)(c). Let be a compact set. We have:
By the assumption in , we take a compact set satisfying . Then we have:
By the condition (a), the subset is relatively compact in . Therefore the condition (c) follows. ∎
Lemma 2.32.
Let and be locally compact groups. Assume acts on continuously as group automorphisms. Put , then we have in .
Proof..
It is enough to show is relatively compact for any compact subsets and . This follows from:
∎
Note 2.33.
Let be a Lie group with finite connected component and its closed subgroup. Put the identity component. Put . Then admits compact Clifford–Klein forms if and only if so does .
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 be a homogeneous space of a Lie group . A closed and connected subgroup is said to be a constructor of if the natural action of on 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 be a Lie group and its closed subgroup. We say a closed and connected subgroup a syndetic hull of if includes cocompactly.
Fact 2.37 ([34, 1]).
Let be a 1-connected completely solvable Lie group and a closed subgroup. Then there exists a unique syndetic hull of . Especially, if the space has a compact Clifford–Klein form , the space has a constructor which is the syndetic hull of .
Remark 2.38.
The assumption of complete solvability in the above fact is crucial. In fact, a solvable Lie group may have a discrete subgroup 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 be a 1-connected nilpotent Lie group, and and its connected subgroups. Then the following conditions are equivalent:
-
(a)
The pair has the property (CI),
-
(b)
,
-
(c)
.
Here, and are the Lie algebras of and , respectively.
This note is easily shown by using the diffeomorphism . It is easy to show the following note in the same way.
Note 3.2.
Note 3.1 also holds under the assumptions that is an arbitrary Lie group and there exists a 1-connected closed normal nilpotent subgroup satisfying .
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 be a 1-connected solvable Lie group, and and closed subgroups of . Then the following conditions are equivalent.
-
(a)
The pair is free in .
-
(b)
The pair has the property (CI) in .
Remark 3.5.
For a 1-connected exponential solvable Lie group , the above statement was proven by Baklouti and Kédim [1].
Note 3.6.
Let be a 1-connected solvable Lie group, and and its closed subgroups.
-
(1)
If the pair is proper, the quotient space has a manifold structure.
-
(2)
Let be a uniform lattice. Assume the action is proper and cocompact, then 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 -action.
Proposition 3.7.
Let be a 1-connected solvable Lie group, and and its connected subgroups. Assume the -action on is proper. Then the following conditions are equivalent:
-
(a)
the space is compact,
-
(b)
,
-
(c)
as a linear space.
Here, , and are Lie algebras of and , respectively.
Proof..
Since the implications (b)(a) is clear, we first show the implication (a)(b). The condition (b) is equivalent to the transitivity of the -action on , so we are enough to show that the space consists of one point. Since is a 1-connected solvable Lie group, and and are connected subgroups, and are contractible by Note 3.8 and Lemma 3.11 below. By Note 3.6 (1), the quotient space has a manifold structure, it is one point by Lemma 3.10. Next, we show the implication (b)(c). Since is an -orbit, we have . On the other hand, by the properness of the -action we have (Note 3.9). Then we have and so we obtain , which implies .
Finally, we check the implication (c)(b). We consider the -orbit of the origin point . The orbit is closed since the -action on is proper. On the other hand, since the -action is free by Corollary 3.4, the dimension of -orbit equals to . Hence the -orbit is open. Since is connected, coincides with the -orbit. ∎
Note 3.8 ([6]).
Any 1-connected solvable Lie group is diffeomorphic to a Euclidian space.
Note 3.9.
Let and be closed subgroups of a 1-connected solvable Lie group . Assume in , then we have .
Lemma 3.10.
Suppose a contractible Lie group acts on a contractible manifold . If the quotient space is a compact manifold, it consists of one point.
This lemma is an immediate consequence of the following two lemmas.
Lemma 3.11.
Let be a contractible topological group and act on a contractible space . Then is contractible.
Proof..
By the homotopy exact sequence of the fiber bundle , we have . Then we have is contractible by J. H. C. Whitehead’s theorem. ∎
Lemma 3.12.
A contractible and closed manifold consists of one point.
Proof..
Let be a contractible and closed manifold. Since an arbitrary vector bundle over is trivial, is orientable. For a volume element of , we have 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 be a 1-connected 3-step nilpotent Lie group, and and its connected subgroups. Then in if and only if the pair has the (CI) property in .
Setting 3.14.
Let be a Lie group, and its closed normal subgroup. Assume is 1-connected nilpotent. Let and be connected subgroups of , and a closed subgroup of (see Notation 2.1) satisfying in . Set .
Proposition 3.15.
Under Setting 3.14, we additionally assume is 2-step nilpotent. Then the following conditions are equivalent:
-
(a)
the pair has the property (CI) in ,
-
(b)
.
Proof..
It is enough to show that the following four conditions are equivalent:
-
(i)
the pair has the property (CI) in ,
-
(ii)
,
-
(iii)
in for any compact set ,
-
(iv)
in .
Here, , and are the Lie algebras of , and , respectively. Since the exponential map is diffeomorphism, we denote by its inverse.
The implication (i)(ii) comes from Note 3.2, the implication (iii)(iv) holds by Lemma 2.31, and the implication (iv)(i) follows from Property 2.27(4).
Then we show the implication (ii)(iii). Take any compact sets and . It is enough to show that the subset is compact. Since is 2-step nilpotent, for we have:
where and . Then we have:
where is a closed cone. On the other hand, by the condition (ii), we have . Hence the pair of the closed cones is proper by Lemma 2.30. Then the subset is compact and so is the subset . ∎
Finally, we introduce criterion of cocompactness under a similar situation in Proposition 3.15.
Proposition 3.16.
Under Setting 3.14, assume in . Then the following conditions are equivalent:
-
(a)
The -action on and the -action on are cocompact.
-
(b)
The -action on is cocompact.
Proof..
First, we show the implication (a)(b). By the assumption in and the cocompactness of the -action, for any , we have in and , and so (Proposition 3.7). Take a compact subset satisfying . We are enough to show . Then we get:
Next, we show the implication (b)(a). By , the -action on is cocompact, so we show that the -action is compact. Take a compact subset satisfying . Then we have:
By the condition , we have is compact, and so have the -action on is cocompact. ∎
4 Indecomposable symmetric triples with signature
In this section, we review the classification of symmetric triples with signature given by Kath–Olbrich (Fact 1.4). We use another notation for our calculations.
We introduce solvable Lie algebras and indecomposable symmetric triples , which are most part of the pseudo-Riemannian symmetric triple with signature (see Fact 1.4). We also define a symmetric triples in Subsection 4.3.
In this section, we use the following:
Notation 4.1.
-
•
: the -dimensional Heisenberg Lie algebra equipped with the following non-trivial brackets:
where is a symplectic form on . Put .
-
•
: the center of , .
-
•
: Heisenberg Lie group, namely, 1-connected Lie group whose Lie algebra is .
Remark 4.2.
4.1 Definition of symmetric triples
In this subsection, we define symmetric triples and see some properties.
Definition 4.3.
We think as a subalgebra of . For , we define:
In the following, we use a basis of such that:
Using this basis, we identify .
Definition 4.4 ().
For matrices , we put . We define a subalgebra by .
Lemma 4.5.
For , we have the following statements.
-
(1)
The Lie algebra is solvable and .
-
(2)
The eigenvalues of are square roots of the eigenvalues of the product .
-
(3)
The Lie algebra is completely solvable if and only if all the eigenvalues of the product are positive real numbers.
Proof..
Since the statements (1) and (2) are clear, we show the statement (3). In general, a Lie algebra over is completely solvable if and only if all the eigenvalues of the adjoint representations are real. In our case, since is a nilpotent ideal, it is equivalent to the eigenvalues of on are real for any . By an easy calculation, with respect to the base , we have:
Hence, is completely solvable if and only if all the eigenvalues of are real. By the statement (2), it is equivalent to the condition that all the eigenvalues of the product are positive. ∎
Proposition and Definition 4.6.
For , set a subalgebra and a subspace . Let be the inner product on defined by the following Gram matrix with respect to the basis :
Then we have:
-
(1)
,
-
(2)
and ,
-
(3)
is -invariant.
Especially, by Note 2.7, we construct a symmetric triple with signature , where is the signature of .
Proof..
The statements (1) and (2) are clear. Since is invertible, we have . Then we show the statement (3), namely,
Regard as the representation matrix of the inner product on and let denote the representation matrix of the linear translation . Then this condition is equivalent to the condition , namely, is skew symmetric. By an easy calculation, we have:
∎
We denote by the 1-connected Lie group with the Lie algebra and by the analytic subgroup with respect to . By Fact 2.9, is the transvection group of .
Proposition 4.7.
For any , the symmetric triple is reducible and indecomposable.
Proof..
Since the subspace is -invariant, the symmetric triple is reducible. Then we show the indecomposability. Let be a non-trivial decomposition. Then we have a -invariant decomposition . By for , the subspace is non-trivial. Then we are enough to show that is degenerate. By Note 4.8 below, it is enough to show that , where . For the decomposition , we set the projections and , respectively. First, we show . Let and assume . Since is invertible, we obtain and , and so we have , which contradicts the non-triviality of . Therefore, for any and so we have . Next, we show . If for any , we have , so we take satisfying . Then we have , and so . ∎
Note 4.8.
Put , then we have .
4.2 Isomorphic classes of the triples
In this subsection, we see an isomorphic classes of the triple (Proposition 4.9) and give the classification for triples with signature (Proposition 6.13).
Proposition 4.9.
For , two symmetric triples and are isomorphic if and only if there exists satisfying:
Proof..
Take an isometric Lie algebra homomorphism , which is compatible with the involutions. Then preserves:
-
•
the decomposition ,
-
•
the center ,
-
•
its orthogonal subspace .
Therefore, with a basis , the map is written of the form:
where and . Since preserves the inner products, we have:
Since is a Lie algebra homomorphism, we have
Set , then the conditions and follow from the above calculations. Conversely, if there exists satisfying the condition, a direct calculation leads us that the homomorphism obtained by putting in the matrix representation above is an isomorphism of symmetric triples. ∎
Definition 4.10.
We denote by the condition of Proposition 4.9. It is easy to check that is an equivalence relation on .
In the rest of this subsection, we see the isomorphic classes of the triples with signature with respect to this equivalence relation.
Proposition 4.11.
For matrices , assume the signature of is . Then the following list gives a complete class representatives of symmetric triple .
-
(1)
,
, -
(2)
(see Notation 2.1),
-
(3)
,
-
(4)
.
Proof of Proposition 6.13.
By Proposition 4.9, we may and do assume . For a basis of , we put . Note that . Then we are enough to consider the orbit of with respect to the action . Note that:
Set . Then we are enough to consider the orbit space of the -action on . By an easy calculation, we have:
Note 4.12.
The orbit space of the -action on is:
-
(1)
The orbits ().
Here, the constraint comes from the condition . We have and:Then we get for some (). Putting and , we obtain or .
-
(2)
The orbits ().
In this case, we have:for some . Take satisfying and put:
then we have the equivalence .
-
(3)
The orbits .
In this case, we have:By a direct calculation, we have the equivalence to the following classes, respectively. (For example, we use for the first pair.)
-
(4)
The orbits .
In this case, we have .
∎
Remark 4.13.
With the natural identification (the 2-dimensional unit sphere), a picture of the parameter spaces of is given as Figure 4.1.

Remark 4.14.
For a symmetric triple , put eigenvalues of the product . Then the following table shows which class of Proposition 6.13 the symmetric triple belongs to.
class | (1) | (2) | (3) | (4) |
eigenvalues | real | not real | real | real |
relation | ||||
is diagonalizable | yes | yes | no | yes |
4.3 Definition of symmetric triples
In this subsection, we define a symmetric triples , which is in the list of the classification (Fact 1.4).
Definition 4.15 ().
We define a nilpotent Lie algebra
as follows:
the other brackets are trivial. |
Put and , then we have , and . We define a -invariant inner product on as follows:
where . By Note 2.7, the triples are indecomposable symmetric triples with signature . We denote by the 1-connected nilpotent Lie group with the Lie algebra and by the analytic subgroup of with respect to . The Lie group is the transvection group of (Fact 2.9).
5 Criterions of the existence of compact Clifford–Klein forms for spaces
To prove the main theorem, we prepare some criterions for the existence of compact Clifford–Klein forms of the symmetric space (Propositions 5.25 and 5.26). In this section, we use Notation 4.1 and the following:
Notation 5.1.
-
•
,
-
•
,
-
•
,
-
•
,
-
•
.
In this section, we identify as and as . Especially, we think and .
Note 5.2.
By a direct calculation, we have:
5.1 Subgroups and
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 . To do this, we introduce some subgroups of and show their basic properties.
Definition 5.3.
For and , we denote by the image of the linear transform defined by . Then we put:
We denote by the analytic subgroup in with respect to . If is a subalgebra of , we denote by its analytic subgroup in .
We see the criterion of to be a subalgebra of .
Proposition 5.4.
For and , the following conditions are equivalent:
-
(a)
the subspace is a subalgebra,
-
(b)
,
-
(c)
the subspace is -invariant,
-
(d)
the subspace is -invariant (),
-
(e)
the subalgebra is -invariant,
-
(f)
.
Remark 5.5.
By this proposition, the conditions (a) and (b) do not depend on .
Proof..
Note 5.6.
For and , the subspace is -invariant if and only if .
We give fundamental properties of the subspaces and .
Note 5.7.
There is a Lie algebra isomorphism , where .
Note 5.8.
For and , we have decompositions , and as linear spaces.
The “converse of Note 5.8” also holds. In fact, we have:
Proposition 5.9.
-
(1)
For any subspace satisfying , there exists satisfying .
-
(2)
For any subalgebra satisfying , there exists satisfying .
-
(3)
For any subalgebra satisfying , there exist and satisfying and .
To prove this proposition, we use the following fact and note.
Fact 5.10 ([16, Lemma 4.3]).
If a subalgebra satisfies as a linear space, we have .
Note 5.11.
Let be a linear space decomposition. For a subspace satisfying , we have .
Proof of Proposition 5.9.
-
(1)
Let be a subspace satisfying . Then there exists such that:
- (2)
- (3)
∎
Proposition 5.12.
For , the following conditions are equivalent:
-
(a)
the pair satisfies the property (CI) in ,
-
(b)
,
-
(c)
(),
-
(d)
the matrix is invertible ().
Proof..
The equivalence (a)(b) comes from Note 3.2. We have for by Note 5.2, so the equivalence (b)(c) holds.
Then we show the equivalence (c)(d). Take any . Since is the image of the linear map , we have:
∎
Then we get:
Lemma 5.13.
Suppose satisfies , then we have in .
Proof..
By Note 3.2 and Propositions 5.4 and 5.12, the pair has the property (CI) in . By Fact 3.13, we have in . By the condition for any , we have in for any compact set . By using Lemma 2.31 with , we obtain in , where is the analytic subgroup of with respect to and the condition comes from Lemma 2.32. ∎
Finally, we classify the constructors of , namely, we have:
Proposition 5.14.
For a connected subgroup , the following conditions are equivalent.
-
(a)
The subgroup is a constructor of .
-
(b)
There exist and satisfying and .
5.2 Uniform lattices of and
In this subsection, we discuss necessary conditions for the existence of a uniform lattice in and , namely, we show the following two propositions.
Proposition 5.15.
Assume there exists such that the subgroup has an -invariant uniform lattice. Then the subspace is -invariant and we have , where .
Proposition 5.16.
Suppose and satisfy the condition in Proposition 5.4. If the subgroup has a uniform lattice, then the condition holds.
To prove these propositions, we define a solvable Lie group.
Definition 5.17.
For , we consider the -action on , , and denote by the semidirect product . We denote by the first projection. We regard as a subgroup of by the injection .
Lemma 5.18.
Let and . If does not have an eigenvalue , then we have .
Proof..
For an element , we have:
∎
Note 5.19.
Some Lie groups in this paper are isomorphic to . Let and satisfy the condition of Proposition 5.4.
-
•
If is symmetric, , where ,
-
•
,
-
•
.
Proposition 5.20.
For , we have if the group has a uniform lattice.
To prove this proposition, we prepare some facts and a proposition.
Fact 5.21 ([33, Theorem 3.3]).
Let be a connected solvable Lie group and its maximum connected (closed) normal nilpotent subgroup. Let be a cocompact closed subgroup of . Assume that contains no non-trivial connected (closed) Lie subgroups which are normal in . Then is compact.
Fact 5.22 ([11]).
Let be a 1-connected nilpotent Lie group and a uniform lattice. Then is commutative if and only if so is .
Proposition 5.23.
For , we consider the following conditions.
-
(1)
There exists a cocompact discrete subgroup satisfying .
-
(2)
There exists a cocompact discrete subgroup satisfying .
-
(3)
All the coefficients of the characteristic polynomial of are integers.
-
(4)
All the coefficients of the characteristic polynomial of are integers and .
Then the implications (1)(3) and (2)(4) hold. Moreover, the equivalences (1)(3) and (2)(4) also hold if the eigenvalues of are distinct.
Proof..
First, we prove the implication (1)(3). Since the uniform lattice of is isomorphic to , is similar to an element of . Therefore, all the coefficients of characteristic polynomial are integers.
Next, we show the implication (2)(4). Since is invertible and both and are integers, then we have .
Finally, we prove the inverse implications (3)(1) and (4)(2). Assume the eigenvalues of are distinct, and all the coefficients of the characteristic polynomial of are integers. Since the eigenvalues of are distinct, there exists such that is a basis of . Then satisfies the condition (1). Actually, by Cayley–Hamilton’s theorem, is written as an linear combination of with integer coefficients, so we have . Especially, in the case , since is also written as a linear combination of , we have . ∎
Example 5.24.
Put and let us see that the group admits a uniform lattice. Set . Since the characteristic polynomial of is and its roots are distinct, there exists a uniform lattice which is -invariant by Proposition 5.23. The subgroup is a uniform lattice.
Proof of Proposition 5.20.
Proof of Proposition 5.15.
Let be an -invariant uniform lattice in . Since is -invariant, the subspace is -invariant by Note 5.2. Now we are enough to show . We denote by the inverse of the exponential map (diffeomorphism).
(a) The case where is commutative.
Since is an -invariant uniform lattice in , we obtain by Proposition 5.23, and so .
(b) The case where is not commutative.
Proof of Proposition 5.16.
Let be a uniform lattice in .
(a) The case where is commutative.
Since the matrix is symmetric by Note 5.7, the condition follows from Note 5.19 and Proposition 5.20.
(b) The case where is not commutative.
Since is the maximum connected normal nilpotent Lie subgroup of , the subgroup is a uniform lattice of by Fact 5.21. By Fact 5.22, the lattice is not commutative, either. Hence is compact. By applying Property 2.29 to the natural surjection , we find is a uniform lattice of the group . Since we have by Note 5.19, the condition 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 . If and are diagonal, this proposition also follows from [16, Proposition 4.8].
Proposition 5.25.
The following conditions are equivalent.
-
(a)
The symmetric space admits compact Clifford–Klein forms.
-
(b)
There exists satisfying the following conditions.
-
(i)
The matrix is invertible for any .
-
(ii)
The subgroup has an -invariant uniform lattice for some .
-
(i)
Moreover, the following condition is a necessary condition of the above condition (ii).
-
(ii’)
The subspace is -invariant and for some .
We prove this in Subsection 5.5. If is completely solvable, we have an easier criterion.
Proposition 5.26.
Assume is completely solvable. The following conditions are equivalent.
-
(a)
The symmetric space admits compact Clifford–Klein forms.
-
(b)
There exist and satisfying the following conditions.
-
(i)
.
-
(ii)
The subgroup admits a uniform lattice.
-
(i)
Moreover, the following condition is a necessary condition of the above condition (ii).
-
(ii’)
.
Proof..
The implication (b)(ii)(ii’) follows from Proposition 5.16, so we are enough to show the equivalence (a)(b), which equivalent to check that the following conditions are equivalent:
-
(A)
there exists a discrete subgroup which acts on properly discontinuously, cocompactly and freely,
-
(B)
there exists a constructor of and has a uniform lattice,
-
(C)
there exists and such that and has a uniform lattice.
The implication (B)(A) comes from Note 3.6 (2). Since is completely solvable, the implication (A)(B) follows from Fact 2.24. The equivalence (B)(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 is not completely solvable. In fact, in this case, a discrete subgroup 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 , a closed subgroup is called an intermediate syndetic hull of if satisfies the following conditions with :
-
(i)
is connected,
-
(ii)
there exists satisfying ,
-
(iii)
is a cocompact subgroup of .
Note 5.28.
The condition (iii) above is equivalent to:
-
(iii’)
is a subgroup of and is cocompact in .
In this subsection, we show the next:
Proposition 5.29.
Let be a discrete subgroup acting on cocompactly. Then has an intermediate syndetic hull.
Lemma 5.30 (Criterion of the existence of an intermediate syndetic hull).
Let be a discrete subgroup and put . Then the following conditions are equivalent.
-
(a)
The subgroup is non-trivial and discrete.
-
(b)
There exists satisfying .
-
(c)
has an intermediate syndetic hull.
Proof..
The implication (c)(a) is easy. We show the implication (a)(b). Let be a generator of . We take satisfying . Then the following exact sequence splits by the group homomorphism .
Then we have and so the condition (b).
In the rest of this subsection, we use the following:
Notation and Setting 5.31.
Let be a 1-connected 2-step nilpotent Lie group, its Lie algebra.
-
•
-
•
is the generalized eigenspace of with respect to an eigenvalue .
-
•
.
Lemma 5.32.
Let . we put . By identifying , put and the first projection. Let be a discrete subgroup. Then is discrete if the following conditions are satisfied.
-
(1)
.
-
(2)
is commutative.
-
(3)
, where is the Lie algebra of the syndetic hull of .
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 as follows, where are eigenvalues of .
Fix and assume . We think . Then there exists satisfying .
Proof..
Since is 2-step nilpotent, we have:
Put , then we have a decomposition and the following matrix representation:
Therefore, the eigenvalues of coincides with them of . Since is -invariant, there exists such that . Then we only have to show that is not constant. It is enough to show that there exists satisfying . Assume for any then the eigenvalues of are all 1. Then by the matrix representation , we have , which contradicts the condition . Therefore, we obtain . ∎
Proof of Lemma 5.32..
It is enough to show that is included in a countable and closed subset of . We put a finite set as in Lemma 5.33, and put subsets as follows:
Then and are countable and closed. Take any . We are enough to show that . Assume , then there exists such that by Lemma 5.33. Since the subalgebra is abelian and has an -invariant uniform lattice, must be integer by Proposition 5.23. Therefore we have . ∎
Finally, we prove Proposition 5.29.
Proof of Proposition 5.29.
Take a discrete subgroup acting on cocompactly. By Lemma 5.30, we are enough to show that is discrete. We consider the natural surjection (see Note 5.19) and put and the image by of and , respectively. By Property 2.29, acts on cocompactly. We consider as a subgroup of . We put .
-
(A)
The case .
We are enough to show for . Here, note that is closed and countable. Assume . Then there exists , where and . By , is abelian. Especially, we have , where . By Lemma 5.18, we have:
On the other hand, since the -action on is not cocompact, neither is the -action, which contradicts the cocompactness of -action (Property 2.29). Then we have .
-
(B)
The case .
- (a)
-
(b)
is non-commutative.
We have is non-trivial and so is compact. By Property 2.29, is a discrete subgroup of and acts on cocompactly. By using Lemma 5.32 for , we have the discreteness of . Actually, it is easy to check the conditions in Lemma 5.32 as follows.-
(1)
.
-
(2)
Since is commutative, so is .
-
(3)
By the assumption , we have , where is the Lie algebra of the syndetic hull of .
-
(1)
∎
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 satisfies (L) condition if the following conditions are satisfied.
-
(1)
is unimodular,
-
(2)
is connected,
-
(3)
there exists satisfying .
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 satisfies (L) condition and put . Then the following conditions are equivalent:
-
(a)
the -action on is proper and cocompact,
-
(b)
the pair satisfies the property (CI) in and the -action on is cocompact,
-
(c)
there exists a matrix such that and is invertible for any (see Notation 5.1).
Proof..
Take satisfying . First we show the equivalence (a)(b). Since the -action on is cocompact, it follows from Propositions 3.15 and 3.16 by putting . Here we need to check that the tuple satisfies the condition of Setting 3.14. The condition is clear. Take satisfying , we have in by Lemma 2.32, and so .
Finally, we prove Proposition 5.25.
Proof of Proposition 5.25.
First, we show the implication (a)(b). Take a discrete subgroup such that is a compact Clifford–Klein form. By Proposition 5.29, we take an intermediate syndetic hull of . Put and take satisfying . By Fact 2.24, the -action on is proper and cocompact. Take such that and is invertible for any by Proposition 5.35. Hence the condition (b)(i) holds. Moreover, by Proposition 5.29(iii), admits an -invariant uniform lattice and so the condition (b)(ii) holds.
Next, we check the implication (b)(a). Take and satisfying the condition (b) and let be an -invariant uniform lattice. Put and . Note that satisfies the (L) condition and is a uniform lattice of . The -action is proper and cocompact by Proposition 5.35. Therefore, is a compact Clifford–Klein form by Note 3.6.
Finally, the implication (b)(ii)(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 which correspond to each case in Fact 1.4.
6.1 On the spaces which correspond to the triples
We consider the spaces which correspond to Case (I) in Fact 1.4.
6.1.1 transvection group
Note that is the transvection group of the symmetric spaces .
Proposition 6.1.
The symmetric space does not admit compact Clifford–Klein forms.
Proof..
Assume admits a compact Clifford–Klein form . Since the Lie group is 1-connected and nilpotent, there exists a constructor including cocompactly by Fact 2.37. By Proposition 3.7, we get , where and are the Lie algebras of and , respectively. On the other hand, there is no such a subalgebra by Lemma 6.2. ∎
In this proof, we use the following lemmas.
Lemma 6.2.
For and in Definition 4.15, no subalgebra satisfies .
Proof..
Lemma 6.3.
Let be a Lie algebra, its center and its decompositions as a linear space. Assume and are subalgebras of and is abelian. For satisfying the following condition, we have .
Proof..
Take the linear map satisfying . Since implies , we assume . Then there exists satisfying , so we have . ∎
6.1.2 isometry group
First, we calculate isometry group of the symmetric spaces.
Proposition 6.4.
We have a Lie group isomorphism . Especially, is completely solvable.
In the following, we calculate the group .
Lemma 6.5.
Using the basis of , we have:
where and .
Proof..
Let . We use as the basis of . Then the representation matrix of is:
where and . Then we have:
(1) | ||||
(2) | ||||
(3) |
By the first identity of (0.7) and (0,5), we have . Then we have and by (0.6), we have , and so . By the second identity of (0.5), is skew symmetric. We put by , then we have . Moreover, a direct calculation leads us if satisfies the conditions of Lemma 6.5.
∎
Lemma 6.6.
By the identification , we have:
Then we prove:
Proposition 6.7.
admits compact Clifford–Klein forms.
Proof..
We are enough to show that admits compact Clifford–Klein forms. Note that . Let be the generator of , we have:
Then we put and:
Note that . Then is an ideal of and we have . We are enough to show that:
-
(1)
in ,
-
(2)
acts on cocompactly,
-
(3)
admits a uniform lattice .
We check the condition (1). Note that is 3-step nilpotent. By Fact 3.13, it is equivalent to the condition that is (CI) in . By Note 3.1, we are enough to check that for all . Since is an ideal, we have . The condition (2) follows from Proposition 3.7. Finally we check the condition (3). We put:
Then is a lattice of , which construct a compact Clifford–Klein form. ∎
6.2 On the spaces which correspond to the triples
We consider the spaces which correspond to Case (II) in Fact 1.4. These spaces are written as for some matrices .
6.2.1 transvection group
Note that the transvection group of is . We check the existence of compact Clifford–Klein forms by using Propositions 5.25 and 5.26. To do this, we introduce subsets of .
Definition 6.8.
We define the following sets:
-
•
,
-
•
,
-
•
,
-
•
In this subsection, we put and denote the five-dimensional Heisenberg Lie algebra by and the Heisenberg Lie group by . We also use the notation , as in Notation 5.1.
Remark 6.9.
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.
-
(i)
The case ().
Claim. .
Take any and put and . Note that the eigenvalues of is or . Since is -invariant for some and do not have common eigenvalues, we have (see Note 6.10). On the other hand, or does not admit non-trivial -invariant subspaces. Then we have .Note 6.10.
For or , let be a matrix. Suppose is an -invariant decomposition such that and do not have common complex eigenvalues. For an -invariant subspace , we have .
-
(ii)
The case (, ).
In this case, is completely solvable (Lemma 4.5(3)).
Claim. .
A direct calculation leads us . Therefore, we have for any and so Claim holds. -
(iii)
The case (, ).
Claim. .
In this case, for we have:Therefore, for any , a direct calculation implies:
Then Claim is a consequence of the following:
Note 6.11.
For , and , put . Then we have for some .
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 for . Then we have .
Proof..
First, for we have:
Set , namely, and . Note that and that two dimensional subspace is -invariant if and only if .
To prove this lemma, we show the following:
Claim. .
) Let and take such that is -invariant and .
Subclaim 1. .
This subclaim follows from .
Subclaim 2. .
Assume . Since the eigenvalues of are by Lemma 4.5(2), the eigenvalues of are distinct, and so is -invariant. Since is two-dimensional, we have or , which contradicts Subclaim 1 and so we have proven the Subclaim 2.
Since is -invariant, by Note 5.6 and Subclaim 2, we have:
is -invariant | |||
By Subclaim 1, we have shown Claim.
Finally, we prove the lemma. Take any . By a direct calculation, we have:
Therefore, for some , so we obtain ∎
The spaces which correspond to Case (II)(c) in Fact 1.4
In this case, the spaces do not admit compact Clifford–Klein forms.
-
(i)
The case , where .
In this case, is completely solvable (Lemma 4.5(3)).Claim. .
By a direct calculation, we have . Thus, the claim follows from for . -
(ii)
The case , where .
Claim. .
We have:Let . If , we have . Hence holds for some and so . Then we assume . For , we have . Then there exist satisfying . Hence we have for some by the intermediate value theorem and so . Therefore we have .
The spaces which correspond to Case (II)(b) in Fact 1.4
In this case, the spaces admit compact Clifford–Klein forms.
-
(i)
The case .
In this case, 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 , then the condition (b)(i) is clear. By Note 5.19, we have for . By Example 5.24, admits a uniform lattice and so the condition (b)(ii) holds. - (ii)
6.2.2 isometry group
First, we calculate isometry group of the symmetric spaces.
Proposition 6.13.
For matrices , assume the signature of is . Then the following list gives a complete class representatives of symmetric triple .
-
(1)
,
, -
(2)
,
-
(3)
,
-
(4)
.
Proposition 6.14.
For a pseudo-Riemannian symmetric space , we have:
In the following, we calculate the group .
Lemma 6.15.
Using the basis of , we have:
where .
Proof..
Let . We use as the basis of . Then the representation matrix of is:
where . Then we have:
(4) | ||||
(5) | ||||
(6) | ||||
(7) |
By the first identity of (0.1), we have . On the other hand by (5). By the determinant relation of (0.3) and (0.4), we have . By the second identity of (0.1), we have . We put , then and follows from (0.1). Moreover, a direct calculation leads us if satisfies the conditions of Lemma 6.15. ∎
Lemma 6.16.
By the identification , we have:
Proof..
Lemma 6.17.
We put . Then is a closed subgroup of and we have .
Proof..
Lemma 6.18.
We have:
Proof..
In the case , it is clear. In the case , it follows from . ∎
Corollary 6.19.
The space admits compact Clifford–Klein forms if and only if so does .
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 and . Then admits compact Clifford–Klein forms and does not admit constructors.
Proof..
First, we check does not admit constructors. Assume admits constructors. Then there exists such that by Proposition 5.14. However, we obtain , which contradicts .
Next, we check admits compact Clifford–Klein forms by using Proposition 5.25. We set:
It is enough to check that the conditions (b)(i) and (ii) in Proposition 5.25. A direct calculation leads us that , , and so the condition (i) holds. Set and . Then we have and so . Then (Note 5.7) has an -invariant uniform lattice , 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 on . 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.