Invariant integral structures in pseudo -type Lie algebras: construction and classification
Abstract.
Pseudo -type Lie algebras are a special class of 2-step nilpotent metric Lie algebras, intimately related to Clifford algebras . In this work we propose the classification method for integral orthonormal structures of pseudo -type Lie algebras. We apply this method for the full classification of these structures for , and irreducible Clifford modules. The latter cases form the basis for the further extensions by making use of the Atiyah-Bott periodicity. The existence of integral structures gives rise to the integral discrete uniform subgroups of the pseudo -type Lie groups.
Key words and phrases:
nilpotent Lie group, rational structure, integral basis, uniform discrete subgroup, Clifford algebra, pseudo type Lie group, admissible module2010 Mathematics Subject Classification:
Primary 22E40, 22E25; Secondary 20H051. Introduction
Two-step nilpotent Lie algebras attracted the attention of G. Métivier [M8́0] in an attempt to describe hypoelliptic operators in a non-Euclidean setting. The condition of hypo-ellipticity required the adjoint map with the value on the center to be surjective. This type of Lie algebras was studied under different names and for different purposes, for instance, in [Ebe94, LT99, MS04, OW10, GMKMV18]. A. Kaplan [Kap80] showed that if the adjoint operator is an isometry, then the sub-Laplacian on two-step nilpotent Lie groups, admits a fundamental solution, reminiscent of that in Euclidean space. His result extended a theorem obtained by G. Folland on the Heisenberg group [Fol73]. Therefore, the class of these Lie algebras received the name (eisenberg)-type Lie algebras. The -type Lie algebras are in a bijective relation to Clifford algebras , generated by the Euclidean space [Rei01a]. The definition of -type Lie algebras related to Clifford algebras , , generated by pseudo Euclidean spaces was extended by P. Ciatti [Cia00] and received the name pseudo -type Lie algebras, see also [GMKM13]. The pseudo -type Lie algebras, which will be denoted by is a fruitful source for studies of Damek-Ricci spaces [BTV95], Iwasawa decomposition of symmetric spaces [CDKR98], Riemannian nilmanifolds [Kap81], rigidity problems [Rei01b], properties of PDE on Lie groups [CS12, MR92, BFM20] and many others topics in geometry, analysis, and geometric measure theory. The classification of the pseudo -type Lie algebras was completed in [FM17, FM19].
Our work is motivated by the study of uniform discrete subgroups on nilpotent Lie groups, which are crucial for the study of homogeneous spaces, compact nilmanifolds, and spectral problems. The existence of a uniform subgroup is guaranteed by a presence of a rational structure on the associated Lie algebra by seminal work of A. I. Malčev [Mc49]. The existence of rational structures on pseudo -type Lie algebras was proved in [CD02, Ebe03, FM14]. A complete classification of rational structures in the class of pseudo -type Lie algebras exists only on the Heisenberg algebra (related to the Clifford algebra ) [GW86]. Some progress in the study of lattices can be found in [CP08].
In the present work, we describe the set of invariant integral structures, which are at the core of rational structures of the Lie algebras. An invariant integral structure is a span over of an orthonormal basis, constructed as an action of a subgroup of the invertible elements in the Clifford algebra on a suitably chosen normal vector in the Clifford module , see Section 3 and Section 3.2. As a result, the basis of the Clifford module is invariant under the action of and the non-vanishing structure constants of the -type Lie algebra are equal to . We emphasize that invariant integral structures are particular cases of integral structures (having structure constants ) that are included in a general class of rational structures on a Lie algebra (having rational structure constants). Two invariant integral structures are orthogonally isomorphic, if and only if the isotropy subgroups and of belongs to the same equivalence class, see Definition 4.3 in Section 4. Section 6 is dedicated to showing the isomorphism properties of invariant integral structures on the -type Lie algebras concerning the equivalence of the isotropy subgroups. The isomorphism of invariant integral structures of the Lie algebras leads to the isomorphism of uniform discrete subgroups on the corresponding Lie groups, which is always extended to an automorphism of ambient pseudo -type Lie groups, see [Rag72].
We apply the classification algorithm to isotropy groups for parameters and in Section 5. We note that the restricted range of and in the construction of the list of non-equivalent isotropy groups corresponds to the first and the second period in of pseudo -type Lie groups concerning the Atiyah-Bott periodicity of the Clifford algebras. The reader can notice that the second period contains more non-equivalent subgroups with phenomena, such as disconnectedness, that can not appear in the first period due to the lack of dimension of the center of the Lie algebra. The forthcoming paper will be dedicated to the study of new features in the increasing of the parameter and the study of the periodicity in both and of the construction of non-equivalent isotropy groups. Despite this, most of the theorems and the characterizations proved in Sections 3, 4, and 6 are valid for arbitrary parameters .
2. Clifford algebras and pseudo -type Lie algebras
In this section we remind some classical objects and introduce the main ones of our interest.
2.1. Clifford algebras
We denote by the pseudo Euclidean space, that is the vector space endowed with the non-degenerate symmetric bilinear form
Let be a Clifford algebra over generated by . Remind that is a quotient of the tensor algebra
by a two sided ideal generated by elements of the form
and is the identity element of the Clifford algebra . Consider a representation of on a real vector space
We call the -module, or simply module if we do not want to specify the signature , and will denote by the action of on . Assume also that the module is equipped with a non-degenerate symmetric bilinear form satisfying the condition
(2.1) |
We call such a module an admissible module of the Clifford algebra . We write or simply for an admissible -module of the minimal dimension and call it a minimal admissible module. The reader can find more about analogous constructions of 2 step nilpotent Lie algebras, not related to representation of Clifford algebras in [Ebe04].
We emphasise the difference between an irreducible Clifford module and a minimal admissible module. Not all irreducible modules can be equipped with a non-degenerate bilinear symmetric form, satisfying (2.1). For instance, lack of dimension of an irreducible module can make any bilinear symmetric form degenerate. An admissible module of has an even dimension . It is isometric to if and it is isometric to if , see [Cia00, Theorem 3.1] and [FM17, Proposition 1]. Any admissible -module can be decomposed into an orthogonal direct sum of minimal admissible modules [FM19, Proposition 2.3 (2)].
2.2. Pseudo -type Lie algebras and Lie groups
Definition 2.1.
Let be an admissible module of a Clifford algebra with the representation map . Define the Lie bracket on by
(2.2) |
The pseudo -type Lie algebra is a Lie algebra whose non-vanishing Lie bracket is defined in (2.2).
Note that the Lie algebra is 2-step nilpotent where is the centre. Property (2.1) and the representation property for imply
(2.3) |
The connected simply connected Lie group of the Lie algebra is called the pseudo -type Lie group. The exponential map is a global analytic diffeomorphism [CG90, Theorem 1.2.1]. It allows to induce the coordinates on the Lie group from the Lie algebra by means of Backer-Campbell-Hausdroff formula. Points are considered as vectors . The group product on is given by
2.3. Automorphisms of pseudo -type Lie algebras
Since automorphisms of a Lie algebra define the automorphisms of its connected simply connected Lie group, we consider only the automorphisms of Lie algebras. The complete description of the group of automorphisms of pseudo -type Lie algebras can be found in [Rie82, Saa96, FM21], see also [AS14].
The automorphisms of pseudo -type Lie algebras are generated by the following ones:
[1] The transformations , calling the dilations.
[2] Let be a nonsingular linear map and an orthogonal transformation of . Then the map is a pseudo -type Lie algebra automorphism, if and only if
(2.4) |
where , are transpose maps with respect to the respective bilinear forms
[3] Let be a linear map, then is an automorphism.
2.4. Rational structures, uniform discrete subgroups, lattices
Definition 2.2.
A Lie algebra over rational numbers is called the rational structure of a real Lie algebra if is isomorphic to .
A real Lie algebra has a rational structure if and only if there is a basis for such that the structure constants of the Lie algebra are rational numbers.
Definition 2.3.
Let be a Lie group. A subgroup is called uniform subgroup if is discrete and is a compact space.
Definition 2.4.
Let be a Lie group with a measure . A subgroup is called lattice if .
Let be a nilpotent Lie group and the Haar measure on it. Then a discrete subgroup is lattice if and only if it is a uniform subgroup, i.e implies that is compact. From now on we will not distinguish the lattices and uniform subgroups. A result from [Mc49] can be formulated as follows.
-
•
If is a uniform subgroup of , then has a rational structure such that .
-
•
If has a rational structure , then has a uniform subgroup such that .
Theorem 2.5.
[Rag72] Let , be uniform subgroups of simply connected nilpotent Lie groups . An isomorphism of discrete subgroups, can be extended to the smooth isomorphism of the Lie groups.
3. Invariant basis of a Clifford module
3.1. Definition of invariant integral structure and uniform subgroups
From now on we will consider only minimal admissible modules of Clifford algebras , denoting them either by or simply by . Let be a pseudo -type Lie algebra with a basis for and a basis for . Note that is the centre of . We write the structure constants for with respect to bases and by
(3.1) |
Definition 3.1.
A basis for is called integral if the structure constants in (3.1) take the values in .
We want to study a special class of integral bases of . To describe it, we fix an orthonormal basis of , where
(3.2) |
Denote by a finite subgroup of the Pin group in defined by
Thus the generators of the group are . Elements satisfy the properties: either or .
We proceed to the construction of bases for the minimal admissible module . In Table (1) the reader finds dimensions of . We indicated by red colour the Clifford algebras, where the minimal admissible modules differ from the irreducible modules. With the subscript we indicated the presence of two non-equivalent minimal admissible modules.
8 | 16 | 32 | 64 | 64×2 | 128 | 128 | 128 | 128×2 | 256 |
---|---|---|---|---|---|---|---|---|---|
7 | 16 | 32 | 64 | 64 | 128 | 128 | 128 | 128 | 256 |
6 | 16 | 16×2 | 32 | 32 | 64 | 64×2 | 128 | 128 | 256 |
5 | 16 | 16 | 16 | 16 | 32 | 64 | 128 | 128 | 256 |
4 | 8 | 8 | 8 | 32 | 64 | 128 | |||
3 | 8 | 8 | 8 | 64 | 128 | ||||
2 | 4 | 64 | |||||||
1 | 32 | ||||||||
0 | |||||||||
s/r | 0 | 1 | 2 | 3 | 4 | 5 | 6 | 7 | 8 |
(1) If a minimal admissible module is irreducible, then the set
(3.3) |
contains a basis for any non-zero vector .
(2) If a minimal admissible module is reducible, then set (3.3) contains for any non-zero and non-null vector .
Thus we obtain that . If is a null vector, then the orbit depends on the choice of , but even in this case, one can make a special choice of a null vector , that generates an entire orbit including . From the other side if is a decomposition of a minimal admissible module on irreducible modules, then the bilinear form vanishes identically on , . In this case only the union contains a basis , where one needs to choose two non-zero vectors .
Based on the latter discussions we restrict ourselves at bases consisting of non-null vectors and make the following definition.
Definition 3.2.
Fix an orthonormal basis of . An orthonormal basis of a minimal admissible module is called invariant basis if it is invariant under the action of ; that is for any and , there exists such that or .
Definition 3.2 requires that the maps , act on an invariant basis by permutations up to the sign .
Remark 3.1.
We emphasise that we require bases to be both orthonormal and invariant.
Example A. Consider the Heisenberg Lie algebra with the normalized basis for the centre and . Set , and . Consider also
where is an orthogonal transformation of . Then the basis is orthonormal. The basis will be invariant under the action of if and only if commutes with . Thus we see that a basis can be orthonormal, but not invariant under the action of .
Example B. Consider the Lie algebra with an orthonormal basis for the centre and a minimal admissible module of the Clifford algebra . We take , such that . The eight vectors
(3.4) |
are linearly independent, have square of the norm equal to , and invariant under the action of . Note that the value is arbitrary and basis (3.4) is orthogonal if and only if . Nevertheless, the vector always can be chosen to make , see [FM14, Lemmas 2.8, 2.9]. This is an example, when the basis can be invariant, but not necessary orthonormal.
Proposition 3.3.
Let be an invariant basis. Then it is an integral basis.
Proof.
We claim that for any with we have:
(3.5) |
Indeed, (3.5) implies and therefore . Assume by contrary that . Suppose first that both and are positive or negative. Then , which is a contradiction. From the other side, if and are opposite, then
by (2.3), and must be a null vector, which is again a contradiction.
The definition of an invariant basis leads to the definition of an invariant integral structure on pseudo -type Lie algebras and (invariant) integral uniform subgroup on the respective pseudo -type Lie groups.
Definition 3.4.
Let be an orthonormal basis for and an invariant basis for a minimal admissible module . An invariant integral structure on the pseudo -type Lie algebra is the vector space over given by
An (invariant) integral uniform subgroup on the pseudo -type Lie group is given by the coordinates
The main goal of the present work is the classification of invariant integral structures on pseudo -type Lie algebras that give rise to classification of integral uniform subgroups on the corresponding pseudo -type Lie groups. Note that invariant integral structures is a subclass of integral (not necessary invariant and/or orthonormal) structures on pseudo -type Lie algebras. In the present work we make a first step and classify only invariant integral structures. Classification of general integral structures and more general rational structures is postponed for the future works. In the article [GW86] the authors made a classification of rational uniform subgroups on the Heisenberg groups, where the starting point was a unique invariant integral basis of the Heisenberg algebra. Thus, in an essence, we make a first step towards the full classification of rational structures on two step nilpotent Lie algebras related to Clifford algebras.
Remark 3.2.
We remark that in the cases of , the invariant integral structures are unique. If and is a vector for with , then is an invariant basis of the minimal admissible module for any choice of a vector with . Thus gives rise to an invariant integral structure of as in Definition 3.4. The Lie algebras and are not isometric, but they are both isomorphic to the Heisenberg Lie algebra.
If and is an orthonormal basis of , then is an invariant basis of the minimal admissible module for any choice of , . The bases generate a unique invariant integral structure of the respective -type Lie algebras. By uniqueness we mean that for any choice of orthonormal basis and any as above the invariant integral structures of the pseudo -type Lie algebras will give the isomorphic invariant uniform subgroups in the pseudo -type Lie groups. The proof is a simplified version of Theorem 6.2.
3.2. A subgroup of positive involutions
In the present section we study subgroups of which will be a core for the construction of invariant bases . Some of the properties of can be learned from the definition of the subgroups , but some of them became clear by considering their action on minimal admissible modules .
Recall that the group consists of elements of the Clifford algebra of the form
(3.6) |
The subgroup is generated by the even number of elements in (3.6). Thus the group is a finite subgroup of .
Definition 3.5.
We denote by a subgroup of satisfying the conditions
-
;
-
and
-
.
Elements are called positive involutions.
The name positive involution refers to the action of on : if () then (). We denote by (or just ), the set of all subgroups of satisfying Definition 3.5. This set is a partially ordered set with respect to the inclusion relation among subsets.
Remark 3.3.
The groups are necessarily commutative.
Example 3.1.
Consider . Then the example of possible subgroups are
and
The first four groups are isomorphic under the action of the orthogonal group . A map permutes the basis vectors , or change their sign. All five groups are isomorphic as abelian groups of order 2. However, the roles of the first four and the last one are different in construction of an invariant basis for .
To avoid the ambiguity occurring with the very similar groups and , we define a bigger group.
Definition 3.6.
Let be a group from Definition 3.5. We denote by the extended group
In Example 3.1 we have subgroups of , where we fix the basis . The subgroups are different, nevertheless
3.3. Generators for a group of positive involutions
In this section, we study groups by describing their generating sets.
Definition 3.7.
We denote by , is the cardinality of the set , a subset in satisfying the conditions:
-
, for , and satisfy in Definition 3.5 for all .
-
The vectors
(3.7) are linearly independent in the vector space .
Proposition 3.8.
The condition is equivalent to
-
non of the products , , , is equal to .
Proof.
Recall that the elements
(3.8) |
, , where and can be chosen to be “” or “”, form a basis for .
It is obvious that implies . Assume that the condition is fulfild. Then the collection in is a reduced collection of linearly independent basis vectors from (3.8), and therefore they are linearly independent. ∎
As an example of a set we present the minimal length positive involutions, which can be classified in the following types:
An easy combinatorial computation shows that generally positive involutions can contain either or basis vectors. This observation inspires us to make a more general definition.
Definition 3.9.
A positive involution containing basis vectors is called a type involution. A positive involution containing basis vectors is called a type involution.
Notation 3.1.
For an element , we denote by the set of the vectors in the product , and by we denote the number of the vectors in . Analogously, () is the set of positive negative vectors in and is the cardinality of the respective sets.
Proposition 3.10.
The following properties can be easily verified
-
(A)
Two type involutions and commute if the number is even. The product is an involution of type .
-
(B)
A type involution and a type involution commute if the number is even. The product is an involution of type .
-
(C)
Two type involutions and commute if the number is odd. The product is an involution of type .
Proof.
The proof is based on the Clifford algebra property
which for orthogonal vectors and leads to . ∎
Notation 3.2.
We denote by the collection of sets satisfying Definition 3.7. The set is partially ordered by the inclusion relation similar to . If , then we denote by a group generated by the set .
Proposition 3.11.
-
(1)
Let . Then
(3.9) is a group of order in and .
-
(2)
Conversely, let . Then there is a (non unique) set such that .
-
(3)
Let be a tuple consisting of , and . Then and .
Proof.
Set in (3.7) is linearly independent and coincides with in (3.9), therefore . If is in the set (3.7), then is not in the set (3.7), which implies that . Any is a positive involution by definition of the set . We showed (1).
The second property will be proved by induction arguments with respect to the order of the group . Let be given. Assume and if there are no elements in other than , then we can put and .
Assume now that there is a set satisfying Definition 3.7. If
is a proper subset of , then there is a positive involution such that , and . Consider the set of commuting involutions
If , then , as a product of involutions and from . Thus non of the elements in can be written in the form for . If
then . So the set satisfies Definition 3.7.
Continuing the procedure, we find in finitely many steps a set satisfying Definition 3.7 such that .
The proof of the last assertion is easily follows from Definition 3.7. ∎
3.4. Relation of and an isotropy subgroup
Now we relate a group with the isotropy subgroup for some and show that they are in a close relation.
Proposition 3.12.
Let be a non-null vector and let denote the isotropy subgroup in of the vector :
Then satisfies Definition 3.5.
Proof.
It is clear that . To check the second property we take and assume by contrary that is a product containing an odd number of negative basis vectors from . Then for with we obtain
by (2.3), which is a contradiction. Similar argument is applied for a vector with . Hence .
The square of every element in equal either or . If , then . Hence . ∎
The relation of an arbitrary to an isotropy group for some is given in the following statement.
Proposition 3.13.
Let and be such that . Let . Then the intersection contains a non-null vector . Moreover, the group is the isotropy subgroup of the vector , and .
If , and there is such that acts as on the minimal admissible module , then the change to leads to the above statement.
Proof.
Let and let , be the eigenspaces of an involution with eigenvalue and , respectively. If one of the spaces is trivial, then the symmetric bi-linear form on the non-trivial subspace is non-degenerate. If both of are non-trivial spaces, then they are orthogonal with respect to and the restriction of onto is non-degenerate too.
Assume . Then the space is invariant under the action of the involution . Therefore, . By repeating the procedures we get that and the restriction of onto is non-degenerate. Thus there is a non-null vector such that for all . Hence .
If , then without loss of generality we can assume that acts as . We change to to get and continue the proof as above. ∎
Corollary 3.14.
Let , and let be an isotropy subgroup of as in Proposition 3.13. The orbit , defined in (3.3), contains an invariant basis of the minimal admissible module . There is no canonical way to prescribe the direction or for a basis vector in . Therefore is a set of basis vectors counted with signes . Hence and .
Proof.
If the group is an isotropy subgroup of an invariant basis, then
(3.10) |
Since the module is minimal admissible and the basis vectors are counted twice (with plus and minus signs), we conclude . ∎
Remark 3.4.
We denote by the subset in consisting of subgroups satisfying (3.10). Furthermore denotes the maximal set of : that is if and only if , see Proposition 3.11. Note that the correspondence from to , assigning is surjective but not necessarily injective.
In Proposition 3.13, if , then . Indeed, since if and only if , we obtain .
Notation 3.3.
We denote by the maximal number of involutions in a set . The value depends only on the signature and it satisfies by Corollary 3.14.
The orbit gives the invariant basis for up to a sign. Since the elements in either commute or anti-commute with elements in , we can more precisely describe the construction of an invariant basis for a minimal admissible module .
Theorem 3.15.
Let be a unit vector from Proposition 3.13. There is a set such that the family is an invariant basis of .
Proof.
Let . We fix a maximal set such that and write , where is either or . We denote and define
(3.11) |
Before we continue the proof we note that , and either or . In the latter case, one involution acts as Id or on , which happens if , see details in [FM21]. Thus
Let be the centralizer of the subgroup in . Then by choosing a unit vector , we can find representatives , and such that
These and form the set . ∎
Proposition 3.16.
Fix the group and the representatives
Assume that generate two sets of invariant bases
as in Theorem 3.15. Then the invariant integral structures
(3.12) |
are isomorphic.
Proof.
We define the correspondence by
(3.13) |
and extend it by linearity over . Then the map is an automorphism of invariant integral structures (3.12). To show that is an isomoprhism, we denote the basis vectors from by and the basis vectors from by , where . Then we note that the bases and are invariant, which means that for any and any there is such that
(3.14) |
The correspondence (3.13) and (3.14) imply that for chosen and we have
Note also that since it maps an othonormal basis to an orthonormal basis. Then we have
∎
4. Equivalence and connectedness of groups
We define an equivalence relation between groups that will descend to the equivalence of their generating sets . We also introduce parameters to distinguish sets for a fixed value . Different sets of parameters will lead to non-equivalent generating sets and the groups. Our aim is to show that equivalent groups lead to the isomorphic invariant integral structures on .
4.1. Equivalence of groups
We recall Notation 3.1 and extend it to the sets .
Notation 4.1.
Let . We denote
We set also , for the cardinality of the respective set, and .
Definition 4.1.
A set consisting only of the involutions of type will be called -type set. A set consisting of the involutions of type and having at least one involution of type will be called -type set.
Proposition 4.2.
Any -type set can be reduced to -type set containing at most one involution of type and the rest of involutions will be of type .
Proof.
The proof follows directly from Proposition 3.10. ∎
Notation 4.2.
If , then we denote by the same letter its natural extension to the action on the group of invertible elements .
Let be a basis as in (3.2). Let . Then is a signed permutation matrix for having only one nonzero component in each column. We call such a map (signed) re-ordering of . If , then defines an element . Since a re-ordering matrix maps positive basis vectors to positive vectors and negative basis vectors to negative basis vectors, it induces a map . For the particular case the map can be chosen also to map positive basis vectors to negative vectors and vice versa. The changes for will be discussed separately in a forthcoming paper.
Definition 4.3.
We say that the groups and are equivalent, writing , if there is a map such that its natural extention to gives the isomorphism between the extended groups and ; that is .
Definition 4.4.
Let and be two sets of involutions. Then we say that and are equivalent, writing , if is equivalent to in the sense of Definition 4.3.
Example 4.1.
Recall Example 3.1 and consider . Let and . Then , since the groups
in are isomorphic under which permutes the basis vectors and , fixing and . Nevertheless, is not equivalent to , since there is no extention of to which maps to .
Example 4.2.
In this example we present a construction of a sequence of subgroups that will be important in Section 5. We call these subgroups standard. Let be an orthonormal basis of . We form a set of mutually different pairs
(4.1) |
(4.2) |
and
The cardinalities of the sets of pairs are
Now we form a set of involutions of type , which from now on will be denoted always by . For any positive integers and we make a product of pairs:
(4.3) |
We denote by the group generated by involutions (4.3).
Proposition 4.5.
In the notation above the groups have the following properties.
-
(i)
is a subgroup of for any and ;
-
(ii)
is a subgroup of for any ;
-
(iii)
is a subgroup of for any ;
-
(iv)
is a subgroup of for any and ;
-
(v)
The standard groups are equivalent for fixed in the sense of Definition 4.3;
- (vi)
-
(vii)
Pairs and commute with all elements in ;
- (viii)
Proof.
Properties (i)-(iv) are obvious. Statements (v) and (vi) follows from the fact the pairs can be chosen up to a sign permutation of the basis in . Properties (vii) and (viii) are the consequence of the facts that pairs , , and the product will have even number of common elements and that the number of vectors in any element of the group is also even. ∎
Example 4.3.
Consider with the basis . The first six elements of the basis are positive and the last three are negative. We can choose the pairs
(4.4) |
up to the sign permutation. They generate a group of cardinality . A possible choice of -type set of involutions generating is
(4.5) |
Any pair from (4.4) will commute with involutions in (4.5) and therefore with all elements in the group . Furthermore, , which is chosen up to a sign permutation, commutes with elements in the group as well. The pairs
Likewise the pairs
Each of the subgroups and is a representative in its class of equivalence. Nevertheless, the groups and are not equivalent.
4.2. Connectivity of groups
Here we introduce another tool of detecting non-equivalent subgroups , that we call “connectedness” for .
Definition 4.6.
A group is called connected if there is no two subgroups , such that is isomorphic to with . We write in this case .
If a group admits the decomposition into subgroups with and for any , then we say that has connected components and we write .
Lemma 4.7.
Let , and . Assume that there is such that , and moreover, there is no such that for any . Then .
Proof.
Note that any product of odd number contains . Let us assume that is a non-trivial decomposition.
If both subgroups include a product of odd number of involutions , , then . Therefore should be connected.
Assume the subgroup consists of only even products of involutions in . We write one of these products in the form , where is one of the generators from the set and is a product of odd number of some involutions in . It implies that . By the assumption for any , there exists a basis vector such that . This implies that and therefore . This shows that the group is connected. ∎
Example 4.4.
The standard subgroups constructed in Example 4.2 are connected for any .
Proposition 4.8.
Let be two generating sets. If , then .
Proof.
We write , and . By the assumption there exists a re-ordering map of the basis such that . If
with
and
then . The re-ordering map will map the non intersecting sets and onto non intersecting sets and . The set (with possible change of signs) will form the set and the set (again with possible change of signs) will form the set . Thus we obtain . ∎
We describe how the -graded product of Clifford algebras can lead to the construction of disconnected subgroups . Consider the following decompositions of an orthonormal basis :
We put and and consider the decomposition , where we assume . This decomposition leads to the isomorphism , where denotes the -graded tensor product of Clifford algebras, see [LM89, Proposition 1.5]. For each of the Clifford algebras , , we consider the minimal admissible modules and the corresponding sets . For and , we have . Let and satisfy
and . We assume also that each set contains at most one type involution , . Then by non-commutativity of and it is easy to see the following properties:
-
If one of the sets or is -type set, then
This implies
(4.6) -
If both and are -type sets, containing type involutions and , then
This implies
(4.7)
One can state similar properties for any number of components in a decomposition .
Remark 4.1.
Proposition 4.9.
The number has three periodicities:
5. Construction of subgroups in , ,
5.1. General method of the construction
In this section we apply the previous theory for the classification of groups and perform the exact construction of non-equivalent subgroups. We restrict ourself to and because we want to illustrate the main features that appears in classification without diving into technical details. The classification for arbitrary is postponed for the forthcoming paper.
We start from and the classification for will be the strait forward generalisation. We classify groups according to parameters: , , and the type or of the set generating the group . We use the standard groups and notations introduced in Example 4.2. For the standard group we will add from none to two additional involutions, see Step 1 below for details. To distinguish the groups, where all previous parameters coincide, we assign the following signature about -type sets, :
(5.1) |
We formulate the results in 15 theorems following the dimension and illustrate each case by a table. We list the set of generators for each group. The group itself and the set of generators will be given up to a sign permutation. The word unique is understood in the sense of equivalence relation of Definition 4.3 or Definition 4.4.
5.1.1. Main steps of the construction of for a fixed .
We divide the construction into three steps.
Step 1. We start from a group satisfying and . First we find standard subgroup and complement it (if necessary) by involutions to reach the maximal number of involutions in generating . The additional involutions will be formed by checking whether the product of and/or by are involutions commuting with . Then we consider a smaller standard subgroup and complement it by a careful choice of involutions to reach the maximal number for , checking whether the connectivity is not violated. We can repeat the last step several times if the condition still holds.
Step 2. We continue to look on and . In most cases it will be a simple step back from to as, for example, for reduction from to .
Step 3. Next we check and . This step is reduced to combinations of the previous 2 steps. If needs, we can proceed to higher number of connected components.
The equivalence of the groups constructed in the previous three steps is summarised in the following proposition.
Proposition 5.1.
Let , with and . Then, the maximal standard subgroups, included in a given group , are equivalent modulo reordering by induction arguments with respect to the dimension , see also Proposition 4.5, item (v).
Lemma 5.2.
If for , then sets satisfying and are always of -type.
Proof.
We start from . For the case there is only one type involution. Let and assume, by contrary, that there is a -type set . We have . The standard subgroup , , does not contain , since is odd. Let will be involutions generating , then . It implies
This contradicts to .
The arguments for the cases , and are similar to the case .
Let . We assume that there is a -type set . We have . The standard subgroup , , contains . Let be involutions generating , where we can assume that and is the last type involution.
-
(1)
If , then
This contradicts to .
-
(2)
If , then we replace by another type involution . In this case and the situation is reduced to the previous step (1). Note that the group is equivalent .
We also note that for and the volume forms which are type involutions can be included to . It justifies the -type set of s in cases and . ∎
5.2. Constructions of groups for
Theorem 5.3.
There is a unique group . It is generated by type involution . Thus we have
Proof.
The group is unique up to reordering. ∎
Theorem 5.4.
There are two non-equivalent groups in .
Signature | |||||
---|---|---|---|---|---|
Proof.
The proof is obvious. ∎
Notation 5.1.
From now on we write to indicate that product in starts from and ends with containing all for odd between and . We have
Theorem 5.5.
There is unique group in .
Signature | |||||
---|---|---|---|---|---|
Proof.
We start from the standard subgroup of the maximal group . The products and commute with the involution . To complete the standard subgroup to the maximal group we add a type involutions
Both choices lead to the equivalent subgroups
by permutation . ∎
Theorem 5.6.
There is unique group in .
Signature | |||||
---|---|---|---|---|---|
Proof.
Theorem 5.7.
There is unique group in .
Signature | |||||
---|---|---|---|---|---|
Proof.
The standard subgroup is generated by involutions (5.2). We need to add two involutions since , at least one of which must contain . We observe that the products and are both involutions commuting with generators (5.2) with each other. We append them both to reach . The reductions to is not possible due to . We finish the proof. ∎
Theorem 5.8.
There are two non-equivalent groups in .
Sinature | |||||
---|---|---|---|---|---|
Proof.
The standard subgroup is generated by involutions
(5.3) |
We need to add one involution since . It is easy to see that only commutes with generators (5.3).
Theorem 5.9.
There are three non-equivalent groups in .
Signature | |||||
---|---|---|---|---|---|
Proof.
The standard subgroup is generated by involutions in (5.3). We need to add one involution containing since and . We add .
We release and consider . It is easy to see that is isomorphic to .
Consider standard subgroup generated by (5.2). This case is reduced to and it is indicated in the table. We finish the proof. ∎
Theorem 5.10.
There are four connected non-equivalent and two disconnected non-equivalent groups in .
Signature | |||||
---|---|---|---|---|---|
Proof.
The standard subgroup is generated by involutions
(5.4) |
We do not need to add any involution, since .
The rest of the connected groups comes from lower dimensions.
To construct the disconnected subgroup corresponding to the -graded tensor product of the Clifford algebras we consider standard subgroup generated by (5.2) and add type involution . Then .
To obtain corresponding to the -graded tensor product of the Clifford algebras we take standard subgroup generated by (5.2) and add type involution . Then .
∎
Theorem 5.11.
There are one connected and two disconnected non-equivalent subgroups in .
Signature | |||||
---|---|---|---|---|---|
Proof.
The standard subgroup is generated by involutions (5.4). We need to add one involution, since . We add . A reduction to the cases is not possible due to .
To construct the disconnected subgroup corresponding to the -graded tensor product of the Clifford algebras we start from the standard subgroup generated by (5.3) and add type involution . Then .
To obtain corresponding to the -graded tensor product of the Clifford algebras we consider standard subgroup generated by (5.2) and add type involution and type involution . Then . ∎
Theorem 5.12.
There are three connected non-equivalent and five disconnected non-equivalent subgroups in .
Signature | |||||
---|---|---|---|---|---|
Proof.
The standard subgroup is generated by involutions
(5.5) |
and it coincides with .
Consider the standard subgroup generated by involutions (5.4). We need to add one involution containing and . We see that commutes with all involutions in (5.4). Adding as the type involution will finish the construction of the maximal group , see Table 10.
To construct the disconnected subgroup corresponding to the -graded tensor product of the Clifford algebras we consider standard subgroup generated by (5.3). We add the involution to the set of generators for and generate the first component in the product . Then we set .
Analogously we construct the disconnected subgroups related to the decomposition of the Clifford algebras and . In both of these cases we remove one of the type involutions and obtain 5 involutions in the total set . Note also that if in the decomposition for we take the set to be of -type set generating and for to be of -type set, then we obtain a group isomorphic to .
If , then the constructions reduce to the case of . ∎
Theorem 5.13.
There are three connected non-equivalent and three disconnected non-equivalent subgroups in .
Signature | |||||
---|---|---|---|---|---|
Proof.
The standard subgroup is generated by involutions (5.5). We add either or as type involutions. We obtain two connected groups and .
Consider the standard subgroup generated by involutions (5.4). We need to add two involutions containing and . We see that type involution commutes with all involutions in (5.4). Adding as the type involution and as type involution, we obtain the maximal group , see Table 11.
To construct the disconnected subgroup corresponding to the - graded tensor product of the Clifford algebras we consider standard subgroup generated by (5.3). We add the involutions to the set of generators for and generate the first component in the product . Then we set generated by the set of .
Analogously we construct disconnected subgroups , , corresponding to the -graded tensor product . For we choose for the group to be -type set and two standard involutions in for the groups to be -type set. For we change the type of the sets .
There are no groups with because . ∎
Theorem 5.14.
There are two connected non-equivalent and two disconnected non-equivalent subgroups in .
Signature | |||||
---|---|---|---|---|---|
Proof.
The standard subgroup is generated by involutions
(5.6) |
We add type involution and obtain the connected group .
Next we consider the standard subgroup generated by involutions (5.5). We need to add two involutions containing and . We see that type involution commutes with involutions in (5.5). Adding either or as type involution, we obtain the maximal group , see Table 12. Adding or , we create equivalent groups .
To construct the disconnected subgroup corresponding to the -graded tensor product of the Clifford algebras we consider standard subgroup generated by (5.3). We add the involutions to the set of generators for and generate the first component in the product . Then we set to be generated by
The disconnected subgroup corresponding to the -graded tensor product of the Clifford algebras , generated similarly. We remove the type involution from one of the sets generating , or in order to get a commutative set for with .
There are no groups with because . ∎
Theorem 5.15.
There are one connected and one disconnected subgroups in .
Signature | |||||
---|---|---|---|---|---|
Proof.
The standard subgroup is generated by involutions (5.6) We add type involution and type involution . We obtain the connected group .
To construct the disconnected subgroup corresponding to the -graded tensor product of the Clifford algebras we proceed as in the case for , and set to be generated by
There are no groups with because . ∎
Theorem 5.16.
There are two connected non-equivalent and two disconnected non-equivalent subgroups in .
Signature | |||||
---|---|---|---|---|---|
Proof.
The standard subgroup is generated by involutions
(5.7) |
We add type involution and obtain the connected group .
To construct the disconnected subgroups corresponding to the -graded tensor product of the Clifford algebras and we proceed as in the previous cases.
The group with coincides with the group . ∎
Proof.
For , the negative basis vector plays no role in forming the involutions, see Definition 3.5. ∎
1 | 2 | 3 | 4 | 5 | 6 | 7 | 8 | |
0 | 0 | 1 | 2 | 1 | 1 | 1 | 2 | |
0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | |
9 | 10 | 11 | 12 | 13 | 14 | 15 | 16 | |
3 | 4 | 1 | 3 | 3 | 2 | 1 | 2 | |
0 | 2 | 2 | 5 | 3 | 2 | 1 | 2 |
6. Isomorphism of invariant integral structures
Theorem 6.1.
If
(6.1) |
then for any orthonormal basis and , with the invariant orthonormal structures spanned by bases as in Table 16 are isomorphic.
Proof.
6.1. Isomorphic invariant integral structures.
We fix an orthonormal basis and a group . Recall the construction of an invariant basis on the minimal admissible module from Theorem 3.15, which used the centraliser of the isotropy group of a unit vector . The invariant integral structure on the Lie algebra given by will be denoted by
Theorem 6.2.
If two groups and are equivalent; that is there exists a map such that , then the invariant integral structures and are isomorphic under a map , where is an orthogonal map with respect to ; that is .
Proof.
The proof is a light generalisation of Proposition 3.16. Let and be equivalent groups. It imply that there is such that where we denoted by the same letter the extention of the orthogonal map to the group of invertible elements of the Clifford algebra . Let
(6.2) |
be the invariant basis, constructed in Theorem 3.15 by making use the eigenspaces of involutions from . The set is equivalent to under . We use the method of Theorem 3.15 and obtain a basis
(6.3) | |||||
where and the set was replaced by . Note that since we also have .
We construct a correspondence by
and . The correspondence extended to a linear map over or is an orthogonal map on since it maps orthonormal basis (6.2) to orthonormal basis (6.3). To show that the linear map is an isomorphism of invariant integral structures, we argue as in Proposition 3.16. By the invariance of the bases and we have
for any , , and for some . It implies
for any and . ∎
Theorem 6.3.
Let and , be the corresponding invariant integral structures. If there is an isomorphism
(6.4) |
with such that , then and are equivalent in the sense of Definition 4.3.
Proof.
Let
be the invariant integral srtuctures generated by the groups and . Here we also assume that is the isotropy subgroup of a unit vector and is the isotropy subgroup of a unit vector . Since is an isomorphism, we obtain . By noting that , we deduce that .
We denote by the same letter the automorphism of which restriction to gives map (6.4). The properties and imply for and , the latter one being an orthogonal transformation over as well. For we find a basis vector such that . If there holds , then the proof is similar. By renumbering the basis vectors we can assume that . We have for the stationary group of
(6.5) | |||||
Since , and , we have
This implies for any . Thus we conclude that if , for then . Thus the groups and are equivalent. The equalities (6.5) shows that and are equivalent. ∎
Table 17 shows the classical groups such that the map with is the automorphism of -type Lie algebras , see also [FM21, Table 3] for non-minimal admissible modules. The groups are subgroups of orthogonal transformations.
8 | |||||||||
---|---|---|---|---|---|---|---|---|---|
7 | |||||||||
6 | |||||||||
5 | |||||||||
4 | |||||||||
3 | |||||||||
2 | |||||||||
1 | |||||||||
0 | |||||||||
0 | 1 | 2 | 3 | 4 | 5 | 6 | 7 | 8 |
Theorem 6.4.
Proof.
If as in the statement of Theorem 6.4 then for an automorphism of we have . It implies that the general automorphisms of also satisfies , see [FM21, Section 3.2].
Thus if the invariant integral structures and are isomorphic, then they will be isomorphic under a map with . It implies that the group and are equivalent by Theorem 6.3.
Conversely, if we assume now that the groups and are equivalent, then by Theorem 6.2 the corresponding invariant integral structures will be isomorphic. ∎
6.2. Non-isomorphic invariant integral structures
Theorem 6.5.
Let and be non-equivalent groups such that there is a type involution in and an involution such that . Then the invariant integral structures and are not isomorphic.
Proof.
Let be a pseudo -type Lie algebra and , as in the statement of Theorem 6.5. We denote by the eigenspace of type involution and by
the non-trivial eigen spaces of . Then the subspaces in the direct sum are orthogonal.
Let us assume that there exists an isomorphism and write
Note the following: since , we obtain that . The map , extended to the Clifford algebra , satisfies . Therefore
(6.6) |
where are non-trivial orthogonal vector spaces.
Let and put , where and . We also have
Since
we obtain and by the uniqueness of the decomposition into a direct sum of vector spaces. We conclude
Let be an orthonormal basis of the space , which is a part of the invariant basis on defined by the . The matrix components of the operator with respect to the basis have the form
where we used the orthogonality of the vector spaces and in (6.6).
Hence the non-vanishing components of the matrix are always even numbers, so that can not be invertible in . It implies that there are no an isomorphism between the invariant integral structures and . ∎
Corollary 6.6.
Proof.
Since a generating set of consists only of involutions of type , the non-existence of an involution such that for any requires that by the maximality of the groups and . But then which is a contradiction. ∎
There are 3 pairs consisting of non-equivalent groups for , which does not satisfies the conditions of Theorem 6.5 For we have two non-equivalent groups and violating the conditions of Theorem 6.5, see Table 10. The generating set is presented here
For there are two sets of pairs of non-equivalent groups violating the conditions of Theorem 6.5, see Table 11. The first collection contains the groups , which are all connected. The second collection contains the groups , which are products of two smaller subgroups. The generating sets for the first collection are:
The generating sets for the second collection are:
We formulate three theorems and prove them. The method is essentially the same and differs only by a choice of a convenient basis for the space invariant under the action of type involutions. We start from since the dimension of is equal to four and the calculations are more transparent.
Theorem 6.7.
Let . The invariant orthogonal lattices and defined by non-equivalent groups and are not isomorphic.
Proof.
The minimal admissible module is isometric to . Let be the eigenspace of involutions of type . Then and , there are the eigenspaces of . Let , . The vectors
form an orthonormal basis of . In fact,
and analogously . Furthermore, from one side
(6.7) |
But from other side
(6.8) |
The equalities (6.7) and(6.8) imply the orthogonality of and . The orthogonality of the rest of vectors are reduced to the calculations as in (6.7) and(6.8), where we only used that the skew symmetry of with respect to product and skew symmetry of the Clifford product .
Assume that there exists an isomorphism between the invariant orthogonal lattices to .
We show that is an orthogonal transformation. In fact, we have
Furthermore, by making use of the fact that the product contains 6 numbers of different , we get
(6.9) | |||||
In the last step we used the skew symmetry of with respect to and skew symmetry . It shows and are orthogonal. Analogously we obtain .
Finally,
This shows that and then requires . Hence by Theorem 6.3, the groups and are equivalent, that is a contradiction. ∎
Theorem 6.8.
Let . The invariant orthogonal lattices and defined by non-equivalent groups and are not isomorphic.
Proof.
As in Theorem 6.7 we define and . Let , . Note that and the product contains six different maps . We show as in Theorem 6.7 that the vectors
form an orthonormal basis of . Assuming that there is an isomorphism mapping the invariant orthogonal lattices to we show that and obtain the contradiction as in Theorem 6.7. ∎
Theorem 6.9.
Let . The invariant orthogonal lattices and defined by non-equivalent groups and are not isomorphic.
Proof.
The minimal admissible module is isometric to . Let be the eigenspace of involutions of type . Then and , there are the eigenspaces of . Let , . The vectors
form an orthonormal basis of by making of calculations as in (6.7) and (6.8). Note that the space is two dimensional quaternion space with the quaternion structure
Then we continue the proof as in Theorem 6.7 and obtain a contradiction. ∎
References
- [AS14] Andrea Altomani and Andrea Santi. Tanaka structures modeled on extended Poincaré algebras. Indiana Univ. Math. J., 63(1):91–117, 2014.
- [BFM20] Wolfram Bauer, André Froehly, and Irina Markina. Fundamental solutions of a class of ultra-hyperbolic operators on pseudo -type groups. Adv. Math., 369:107186, 46, 2020.
- [BTV95] Jürgen Berndt, Franco Tricerri, and Lieven Vanhecke. Generalized Heisenberg groups and Damek-Ricci harmonic spaces, volume 1598 of Lecture Notes in Mathematics. Springer-Verlag, Berlin, 1995.
- [CD02] Gordon Crandall and Józef Dodziuk. Integral structures on -type Lie algebras. J. Lie Theory, 12(1):69–79, 2002.
- [CDKR98] Michael Cowling, Anthony Dooley, Adam Korányi, and Fulvio Ricci. An approach to symmetric spaces of rank one via groups of Heisenberg type. J. Geom. Anal., 8(2):199–237, 1998.
- [CG90] Laurence Corwin and Frederick P Greenleaf. Representations of nilpotent Lie groups and their applications: Volume 1, Part 1, Basic theory and examples, volume 18. Cambridge university press, 1990.
- [Cia00] Paolo Ciatti. Scalar products on Clifford modules and pseudo--type Lie algebras. Ann. Mat. Pura Appl. (4), 178:1–31, 2000.
- [CP08] Luis A. Cordero and Phillip E. Parker. Lattices and periodic geodesics in pseudoriemannian 2-step nilpotent Lie groups. Int. J. Geom. Methods Mod. Phys., 5(1):79–99, 2008.
- [CS12] Isolda Cardoso and Linda Saal. Explicit fundamental solutions of some second order differential operators on Heisenberg groups. Colloq. Math., 129(2):263–288, 2012.
- [Ebe94] Patrick Eberlein. Geometry of -step nilpotent groups with a left invariant metric. II. Trans. Amer. Math. Soc., 343(2):805–828, 1994.
- [Ebe03] Patrick Eberlein. Riemannian submersions and lattices in 2-step nilpotent Lie groups. Comm. Anal. Geom., 11(3):441–488, 2003.
- [Ebe04] Patrick Eberlein. Geometry of 2-step nilpotent Lie groups. In Modern dynamical systems and applications, pages 67–101. Cambridge Univ. Press, Cambridge, 2004.
- [FM14] Kenro Furutani and Irina Markina. Existence of lattices on general -type groups. J. Lie Theory, 24(4):979–1011, 2014.
- [FM17] Kenro Furutani and Irina Markina. Complete classification of pseudo -type Lie algebras: I. Geom. Dedicata, 190:23–51, 2017.
- [FM19] Kenro Furutani and Irina Markina. Complete classification of pseudo -type algebras: II. Geom. Dedicata, 202:233–264, 2019.
- [FM21] Kenro Furutani and Irina Markina. Automorphism groups of pseudo -type algebras. J. Algebra, 568:91–138, 2021.
- [Fol73] G. B. Folland. A fundamental solution for a subelliptic operator. Bull. Amer. Math. Soc., 79:373–376, 1973.
- [GMKM13] Mauricio Godoy Molina, Anna Korolko, and Irina Markina. Sub-semi-Riemannian geometry of general -type groups. Bull. Sci. Math., 137(6):805–833, 2013.
- [GMKMV18] Mauricio Godoy Molina, Boris Kruglikov, Irina Markina, and Alexander Vasil’ev. Rigidity of 2-step Carnot groups. J. Geom. Anal., 28(2):1477–1501, 2018.
- [GW86] Carolyn S. Gordon and Edward N. Wilson. The spectrum of the Laplacian on Riemannian Heisenberg manifolds. Michigan Math. J., 33(2):253–271, 1986.
- [Kap80] Aroldo Kaplan. Fundamental solutions for a class of hypoelliptic PDE generated by composition of quadratic forms. Trans. Amer. Math. Soc., 258(1):147–153, 1980.
- [Kap81] Aroldo Kaplan. Riemannian nilmanifolds attached to Clifford modules. Geom. Dedicata, 11(2):127–136, 1981.
- [LM89] H. Blaine Lawson, Jr. and Marie-Louise Michelsohn. Spin geometry, volume 38 of Princeton Mathematical Series. Princeton University Press, Princeton, NJ, 1989.
- [LT99] Fernando Levstein and Alejandro Tiraboschi. Classes of -step nilpotent Lie algebras. Comm. Algebra, 27(5):2425–2440, 1999.
- [M8́0] Guy Métivier. Hypoellipticité analytique sur des groupes nilpotents de rang . Duke Math. J., 47(1):195–221, 1980.
- [Mc49] A. I. Mal′ cev. On a class of homogeneous spaces. Izvestiya Akad. Nauk. SSSR. Ser. Mat., 13:9–32, 1949.
- [MR92] D. Müller and F. Ricci. Analysis of second order differential operators on Heisenberg groups. II. J. Funct. Anal., 108(2):296–346, 1992.
- [MS04] Detlef Müller and Andreas Seeger. Singular spherical maximal operators on a class of two step nilpotent Lie groups. Israel J. Math., 141:315–340, 2004.
- [OW10] Alessandro Ottazzi and Ben Warhurst. Rigidity of Iwasawa nilpotent Lie groups via Tanaka’s theory. Note Mat., 30(1):141–146, 2010.
- [Rag72] M. S. Raghunathan. Discrete subgroups of Lie groups. Ergebnisse der Mathematik und ihrer Grenzgebiete, Band 68. Springer-Verlag, New York-Heidelberg, 1972.
- [Rei01a] H. M. Reimann. -type groups and Clifford modules. Adv. Appl. Clifford Algebras, 11(S2):277–287, 2001.
- [Rei01b] Hans Martin Reimann. Rigidity of -type groups. Math. Z., 237(4):697–725, 2001.
- [Rie82] C. Riehm. The automorphism group of a composition of quadratic forms. Trans. Amer. Math. Soc., 269(2):403–414, 1982.
- [Saa96] L. Saal. The automorphism group of a Lie algebra of Heisenberg type. Rend. Sem. Mat. Univ. Politec. Torino, 54(2):101–113, 1996.