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

Invariant integral structures in pseudo HH-type Lie algebras: construction and classification

Kenro Furutani and Irina Markina K. Furutani. Osaka Central Advanced Mathematical Institute, Osaka Metropolitan University, Sugimoto, Sumiyoshi-ku, Osaka 558-8585, Japan [email protected] I. Markina. Department of Mathematics, University of Bergen, P.O. Box 7803, Bergen N-5020, Norway [email protected]
Abstract.

Pseudo HH-type Lie algebras are a special class of 2-step nilpotent metric Lie algebras, intimately related to Clifford algebras Clr,s\text{Cl}_{r,s}. In this work we propose the classification method for integral orthonormal structures of pseudo HH-type Lie algebras. We apply this method for the full classification of these structures for r{1,,16}r\in\{1,\ldots,16\}, s{0,1}s\in\{0,1\} 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 HH-type Lie groups.

Key words and phrases:
nilpotent Lie group, rational structure, integral basis, uniform discrete subgroup, Clifford algebra, pseudo HH type Lie group, admissible module
2010 Mathematics Subject Classification:
Primary 22E40, 22E25; Secondary 20H05
The work of authors was partially supported by the project Pure Mathematics in Norway, funded by Trond Mohn Foundation and Tromsø Research Foundation, by JSPS fund No. 20K03662, and Osaka Metropolitan University, Central Advanced Mathematical Institute (MEXT Joint Usage/Research Center on Mathematics and Theoretical Physics (JPMXP0619217849)).

1. 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 HH(eisenberg)-type Lie algebras. The HH-type Lie algebras are in a bijective relation to Clifford algebras Clr,0\text{Cl}_{r,0}, generated by the Euclidean space r\mathbb{R}^{r} [Rei01a]. The definition of HH-type Lie algebras related to Clifford algebras Clr,s\text{Cl}_{r,s}, s>0s>0, generated by pseudo Euclidean spaces r,s\mathbb{R}^{r,s} was extended by P. Ciatti [Cia00] and received the name pseudo HH-type Lie algebras, see also [GMKM13]. The pseudo HH-type Lie algebras, which will be denoted by 𝔫r,s\mathfrak{n}_{r,s} 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 HH-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 HH-type Lie algebras was proved in [CD02, Ebe03, FM14]. A complete classification of rational structures in the class of pseudo HH-type Lie algebras exists only on the Heisenberg algebra (related to the Clifford algebra Cl1,0\text{Cl}_{1,0}[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 \mathbb{Z} of an orthonormal basis, constructed as an action of a subgroup G(Br,s)G(B_{r,s}) of the invertible elements Pin(r,s)\text{\rm Pin}(r,s) in the Clifford algebra Clr,s\text{Cl}_{r,s} on a suitably chosen normal vector vVv\in V in the Clifford module VV, see Section 3 and Section 3.2. As a result, the basis of the Clifford module VV is invariant under the action of G(Br,s)G(B_{r,s}) and the non-vanishing structure constants of the HH-type Lie algebra are equal to ±1\pm 1. We emphasize that invariant integral structures are particular cases of integral structures (having structure constants ±1\pm 1) 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 𝒮v(1)Clr,s\mathcal{S}^{(1)}_{v}\subset\text{Cl}_{r,s} and 𝒮v(2)Clr,s\mathcal{S}^{(2)}_{v}\subset\text{Cl}_{r,s} of vVv\in V 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 HH-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 HH-type Lie groups, see [Rag72].

We apply the classification algorithm to isotropy groups 𝒮v\mathcal{S}_{v} for parameters r{3,,16}r\in\{3,\ldots,16\} and s{0,1}s\in\{0,1\} in Section 5. We note that the restricted range of rr and ss in the construction of the list of non-equivalent isotropy groups corresponds to the first and the second period in rr of pseudo HH-type Lie groups concerning the Atiyah-Bott periodicity of the Clifford algebras. The reader can notice that the second period r{9,,16}r\in\{9,\ldots,16\} contains more non-equivalent subgroups with phenomena, such as disconnectedness, that can not appear in the first period r{3,,8}r\in\{3,\ldots,8\} 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 ss and the study of the periodicity in both rr and ss of the construction of non-equivalent isotropy groups. Despite this, most of the theorems and the characterizations proved in Sections 34, and 6 are valid for arbitrary parameters (r,s)(r,s).

2. Clifford algebras and pseudo HH-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 r,s\mathbb{R}^{r,s} the pseudo Euclidean space, that is the vector space r+s\mathbb{R}^{r+s} endowed with the non-degenerate symmetric bilinear form

x,yr,s=k=1rxkykk=r+1r+sxkyk.\langle x,y\rangle_{r,s}=\sum_{k=1}^{r}x_{k}y_{k}-\sum_{k=r+1}^{r+s}x_{k}y_{k}.

Let Clr,s\text{Cl}_{r,s} be a Clifford algebra over \mathbb{R} generated by r,s\mathbb{R}^{r,s}. Remind that Clr,s\text{Cl}_{r,s} is a quotient of the tensor algebra

𝒯(U):=r,s(2r,s)(3r,s)(4r,s)\mathcal{T}(U):=\mathbb{R}\oplus\mathbb{R}^{r,s}\oplus\Big{(}\stackrel{{\scriptstyle 2}}{{\otimes}}\mathbb{R}^{r,s}\Big{)}\oplus\Big{(}\stackrel{{\scriptstyle 3}}{{\otimes}}\mathbb{R}^{r,s}\Big{)}\oplus\Big{(}\stackrel{{\scriptstyle 4}}{{\otimes}}\mathbb{R}^{r,s}\Big{)}\oplus\ldots

by a two sided ideal Ir,sI_{r,s} generated by elements of the form

xx+x,xr,s𝟏,xr+s,x\otimes x+\langle x,x\rangle_{r,s}{\mathbf{1}},\quad x\in\mathbb{R}^{r+s},

and 𝟏{\mathbf{1}} is the identity element of the Clifford algebra Clr,s\text{Cl}_{r,s}. Consider a representation of Clr,s\text{Cl}_{r,s} on a real vector space VV

J:Clr,sEnd(V).J\colon\text{Cl}_{r,s}\to\text{\rm End}(V).

We call VV the Clr,s\text{Cl}_{r,s}-module, or simply module if we do not want to specify the signature (r,s)(r,s), and will denote by JzvJ_{z}v the action of zr,sz\in\mathbb{R}^{r,s} on vVv\in V. Assume also that the module VV is equipped with a non-degenerate symmetric bilinear form .,.V\langle.\,,.\rangle_{V} satisfying the condition

(2.1) Jzu,vV+u,JzvV=0for anyzr,sandu,vV.\langle J_{z}u,v\rangle_{V}+\langle u,J_{z}v\rangle_{V}=0\quad\text{for any}\quad z\in\mathbb{R}^{r,s}\quad\text{and}\quad u,v\in V.

We call such a module V=(V,.,.V)V=(V,\langle.\,,.\rangle_{V}) an admissible module of the Clifford algebra Clr,s\text{Cl}_{r,s}. We write Vmin=(Vmin,.,.V)V_{min}=(V_{min},\langle.\,,.\rangle_{V}) or simply VminV_{min} for an admissible Clr,s\text{Cl}_{r,s}-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 VV of Clr,s\text{Cl}_{r,s} has an even dimension dim(V)=2n=N\dim(V)=2n=N. It is isometric to n,n\mathbb{R}^{n,n} if s>0s>0 and it is isometric to ±N,0\mathbb{R}^{\pm N,0} if s=0s=0, see [Cia00, Theorem 3.1] and [FM17, Proposition 1]. Any admissible Clr,s\text{Cl}_{r,s}-module can be decomposed into an orthogonal direct sum of minimal admissible modules [FM19, Proposition 2.3 (2)].

2.2. Pseudo HH-type Lie algebras and Lie groups

Definition 2.1.

Let (V,.,.V)(V,\langle.\,,.\rangle_{V}) be an admissible module of a Clifford algebra Clr,s\text{Cl}_{r,s} with the representation map JJ. Define the Lie bracket on V×r,sV\times\mathbb{R}^{r,s} by

(2.2) Jzu,vV=z,[u,v]r,s,zr,s,u,vV.\langle J_{z}u,v\rangle_{V}=\langle z,[u,\,v]\rangle_{r,s},\quad z\in\mathbb{R}^{r,s},\quad u,v\in V.

The pseudo HH-type Lie algebra 𝔫r,s(V)=(Vr,s,[.,.])\mathfrak{n}_{r,s}(V)=(V\oplus\mathbb{R}^{r,s},[.\,,.]) is a Lie algebra whose non-vanishing Lie bracket is defined in (2.2).

Note that the Lie algebra 𝔫r,s(V)\mathfrak{n}_{r,s}(V) is 2-step nilpotent where r,s\mathbb{R}^{r,s} is the centre. Property (2.1) and the representation property Jz2v=z,zr,svJ_{z}^{2}v=-\langle z,z\rangle_{r,s}v for vVv\in V imply

(2.3) Jzu,Jzvr,s=z,zr,su,vV,Jzu,Jwur,s=z,wr,su,uV.\langle J_{z}u,J_{z}v\rangle_{r,s}=\langle z,z\rangle_{r,s}\langle u,v\rangle_{V},\quad\langle J_{z}u,J_{w}u\rangle_{r,s}=\langle z,w\rangle_{r,s}\langle u,u\rangle_{V}.

The connected simply connected Lie group r,s(V)\mathbb{N}_{r,s}(V) of the Lie algebra 𝔫r,s(V)\mathfrak{n}_{r,s}(V) is called the pseudo HH-type Lie group. The exponential map exp:𝔫r,s(V)r,s(V)\exp\colon\mathfrak{n}_{r,s}(V)\to\mathbb{N}_{r,s}(V) 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 gr,s(V)g\in\mathbb{N}_{r,s}(V) are considered as vectors g=(u,z)Vr,s=𝔫r,s(V)g=(u,z)\in V\oplus\mathbb{R}^{r,s}=\mathfrak{n}_{r,s}(V). The group product \ast on r,s(V)\mathbb{N}_{r,s}(V) is given by

:r,s(V)\displaystyle\ast\colon\mathbb{N}_{r,s}(V) ×r,s(V)r,s(V),\displaystyle\times\mathbb{N}_{r,s}(V)\to\mathbb{N}_{r,s}(V),
(u1,z1)\displaystyle(u_{1},z_{1}) (u2,z2)=(u1+u2,z1+z2+12[u1,u2]).\displaystyle\ast(u_{2},z_{2})=\Big{(}u_{1}+u_{2},\ z_{1}+z_{2}+\frac{1}{2}[u_{1},u_{2}]\Big{)}.

2.3. Automorphisms of pseudo HH-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 HH-type Lie algebras can be found in [Rie82, Saa96, FM21], see also [AS14].

The automorphisms of pseudo HH-type Lie algebras are generated by the following ones:

[1] The transformations δλ(u,z)=(λu,λ2z)\delta_{\lambda}(u,z)=(\lambda u,\lambda^{2}z), calling the dilations.

[2] Let A:VVA\colon V\to V be a nonsingular linear map and CO(r,s)C\in\text{\rm O}(r,s) an orthogonal transformation of r,s\mathbb{R}^{r,s}. Then the map ACA\oplus C is a pseudo HH-type Lie algebra automorphism, if and only if

(2.4) AτJzA=JCτ(z),zr,s,A^{\tau}\circ J_{z}\circ A=J_{C^{\tau}(z)},\quad z\in\mathbb{R}^{r,s},

where AτA^{\tau}, CτC^{\tau} are transpose maps with respect to the respective bilinear forms

Aτu,vV=u,AvV,Cτz,wr,s=z,Cwr,s.\langle A^{\tau}u,v\rangle_{V}=\langle u,Av\rangle_{V},\quad\langle C^{\tau}z,w\rangle_{r,s}=\langle z,Cw\rangle_{r,s}.

[3] Let B:Vr,sB\colon V\to\mathbb{R}^{r,s} be a linear map, then (v,z)(v,z+Bv)(v,z)\mapsto\big{(}v,z+Bv\big{)} is an automorphism.

2.4. Rational structures, uniform discrete subgroups, lattices

We refer to [Rag72, CG90] for the details discussed in this section.

Definition 2.2.

A Lie algebra 𝔤\mathfrak{g}_{\mathbb{Q}} over rational numbers \mathbb{Q} is called the rational structure of a real Lie algebra 𝔤\mathfrak{g} if 𝔤\mathfrak{g} is isomorphic to 𝔤\mathfrak{g}_{\mathbb{Q}}\otimes\mathbb{R}.

A real Lie algebra 𝔤\mathfrak{g} has a rational structure if and only if there is a basis for 𝔤\mathfrak{g} such that the structure constants of the Lie algebra are rational numbers.

Definition 2.3.

Let GG be a Lie group. A subgroup Γ\Gamma is called uniform subgroup if Γ\Gamma is discrete and G/ΓG/\Gamma is a compact space.

Definition 2.4.

Let GG be a Lie group with a measure μ\mu. A subgroup Λ\Lambda is called lattice if μ(G/Λ)<\mu(G/\Lambda)<\infty.

Let GG be a nilpotent Lie group and μ\mu the Haar measure on it. Then a discrete subgroup Γ\Gamma is lattice if and only if it is a uniform subgroup, i.e μ(G/Γ)<\mu(G/\Gamma)<\infty implies that G/ΓG/\Gamma is compact. From now on we will not distinguish the lattices and uniform subgroups. A result from [Mc49] can be formulated as follows.

  • If Γ\Gamma is a uniform subgroup of GG, then 𝔤\mathfrak{g} has a rational structure 𝔤\mathfrak{g}_{\mathbb{Q}} such that 𝔤=span{log\mathfrak{g}_{\mathbb{Q}}=\text{\rm span}\,_{\mathbb{Q}}\{\log(Γ)}(\Gamma)\}.

  • If 𝔤\mathfrak{g} has a rational structure 𝔤\mathfrak{g}_{\mathbb{Q}}, then GG has a uniform subgroup Γ\Gamma such that log(Γ)𝔤\log(\Gamma)\subseteq\mathfrak{g}_{\mathbb{Q}}.

Theorem 2.5.

[Rag72] Let ΓiGi\Gamma_{i}\subset G_{i}, i=1,2i=1,2 be uniform subgroups of simply connected nilpotent Lie groups GiG_{i}. An isomorphism φ:Γ1Γ2\varphi\colon\Gamma_{1}\to\Gamma_{2} of discrete subgroups, can be extended to the smooth isomorphism φ~:G1G2\tilde{\varphi}\colon G_{1}\to G_{2} 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 Clr,s\text{Cl}_{r,s}, denoting them either by Vr,sV^{r,s} or simply by VV. Let 𝔫r,s(V)=(Vr,s,[.,.])\mathfrak{n}_{r,s}(V)=(V\oplus\mathbb{R}^{r,s},[.\,,.]) be a pseudo HH-type Lie algebra with Br,sB_{r,s} a basis for r,s\mathbb{R}^{r,s} and 𝔅(V)\mathfrak{B}(V) a basis for VV. Note that r,s\mathbb{R}^{r,s} is the centre of 𝔫r,s(V)\mathfrak{n}_{r,s}(V). We write the structure constants cijlc_{ij}^{l} for 𝔫r,s(V)\mathfrak{n}_{r,s}(V) with respect to bases 𝔅(V)\mathfrak{B}(V) and Br,sB_{r,s} by

(3.1) [vi,vj]=l=1r+scijlzl.[v_{i},v_{j}]=\sum_{l=1}^{r+s}c_{ij}^{l}z_{l}.
Definition 3.1.

A basis {𝔅(V),Br,s}\{\mathfrak{B}(V),B_{r,s}\} for 𝔫r,s(V)\mathfrak{n}_{r,s}(V) is called integral if the structure constants cijlc_{ij}^{l} in (3.1) take the values in {1,0,1}\{-1,0,1\}.

We want to study a special class of integral bases of 𝔫r,s(V)\mathfrak{n}_{r,s}(V). To describe it, we fix an orthonormal basis Br,s={z1,,zr,zr+1,,zr+s}B_{r,s}=\{z_{1},\ldots,z_{r},z_{r+1},\ldots,z_{r+s}\} of r,s\mathbb{R}^{r,s}, where

(3.2) {z1,,zrare positive, i.e.,zi,zir,s=1,zr+1,,zr+sare negative, i.e.,zi,zir,s=1.\begin{cases}&z_{1},\ldots,z_{r}\quad\text{are positive, i.e.,}~{}\langle z_{i},\,z_{i}\rangle_{r,s}=1,\\ &z_{r+1},\ldots,z_{r+s}~{}\text{are negative, i.e.,}~{}\langle z_{i},\,z_{i}\rangle_{r,s}=-1.\end{cases}

Denote by G(Br,s)G(B_{r,s}) a finite subgroup of the Pin group in Clr,s\text{Cl}_{r,s} defined by

G(Br,s)={±𝟏,±z1,,±zr+s,±zi1zik1i1<<ikr+s,k=2,,r+s}.\begin{array}[]{lll}G(B_{r,s})=\big{\{}&\pm{\mathbf{1}},\ \pm z_{1},\ \ldots,\ \pm z_{r+s},\ \pm z_{i_{1}}\cdots z_{i_{k}}\mid\\ &1\leq i_{1}<\cdots<i_{k}\leq r+s,\quad k=2,\ldots,r+s\big{\}}.\end{array}

Thus the generators of the group G(Br,s)G(B_{r,s}) are {1,Br,s}\{-1,B_{r,s}\}. Elements σG(Br,s)\sigma\in G(B_{r,s}) satisfy the properties: either σ2=𝟏\sigma^{2}={\mathbf{1}} or σ2=𝟏\sigma^{2}=-{\mathbf{1}}.

We proceed to the construction of bases 𝔅(Vr,s)\mathfrak{B}(V^{r,s}) for the minimal admissible module Vr,sV^{r,s}. In Table (1) the reader finds dimensions of Vr,sV^{r,s}. We indicated by red colour the Clifford algebras, where the minimal admissible modules differ from the irreducible modules. With the subscript ×2\ {}_{\times 2} we indicated the presence of two non-equivalent minimal admissible modules.

Table 1. Dimensions of 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 8×28_{\times 2} 1616 32 64 64×2{\text{\small 64}_{\times 2}} 128
3 8 8 8 88 1616 3232 64 6464 128
2 4 4×2{\color[rgb]{1,0,0}4_{\times 2}} 8{\color[rgb]{1,0,0}8} 88 1616 16×216_{\times 2} 3232 3232 64
1 2{\color[rgb]{1,0,0}2} 4{\color[rgb]{1,0,0}4} 8{\color[rgb]{1,0,0}8} 88 16{\color[rgb]{1,0,0}16} 1616 1616 1616 32
0 11 22 44 4×24_{\times 2} 88 88 88 8×28_{\times 2} 1616
s/r 0 1 2 3 4 5 6 7 8

(1) If a minimal admissible module Vr,sV^{r,s} is irreducible, then the set

(3.3) Ov=G(Br,s).v:={JσvσG(Br,s)}O_{v}=G(B_{r,s}).v:=\{J_{\sigma}v\mid\ \sigma\in G(B_{r,s})\}

contains a basis 𝔅(Vr,s)\mathfrak{B}(V^{r,s}) for any non-zero vector vVr,sv\in V^{r,s}.

(2) If a minimal admissible module Vr,sV^{r,s} is reducible, then set (3.3) contains 𝔅(Vr,s)\mathfrak{B}(V^{r,s}) for any non-zero and non-null vector vVr,sv\in V^{r,s}.

Thus we obtain that Vr,s=span{Ov}=span{𝔅(Vr,s)}V^{r,s}=\text{\rm span}\,_{\mathbb{R}}\{O_{v}\}=\text{\rm span}\,_{\mathbb{R}}\{\mathfrak{B}(V^{r,s})\}. If vVr,sv\in V^{r,s} is a null vector, then the orbit OvO_{v} depends on the choice of vv, but even in this case, one can make a special choice of a null vector vVr,sv\in V^{r,s}, that generates an entire orbit OvO_{v} including 𝔅(Vr,s)\mathfrak{B}(V^{r,s}). From the other side if Vr,s=V1r,sV2r,sV^{r,s}=V_{1}^{r,s}\oplus V_{2}^{r,s} is a decomposition of a minimal admissible module on irreducible modules, then the bilinear form .,.Vr,s\langle.\,,.\rangle_{V^{r,s}} vanishes identically on Vir,sV_{i}^{r,s}, i=1,2i=1,2. In this case only the union i=12{JσviσG(Br,s)}\bigcup_{i=1}^{2}\,\big{\{}J_{\sigma}v_{i}\mid\sigma\in G(B_{r,s})\big{\}} contains a basis 𝔅(Vr,s)\mathfrak{B}(V^{r,s}), where one needs to choose two non-zero vectors viVir,sv_{i}\in V_{i}^{r,s}.

Based on the latter discussions we restrict ourselves at bases 𝔅(Vr,s)\mathfrak{B}(V^{r,s}) consisting of non-null vectors and make the following definition.

Definition 3.2.

Fix an orthonormal basis Br,sB_{r,s} of r,s\mathbb{R}^{r,s}. An orthonormal basis 𝔅(Vr,s)\mathfrak{B}(V^{r,s}) of a minimal admissible module Vr,sV^{r,s} is called invariant basis if it is invariant under the action of G(Br,s)G(B_{r,s}); that is for any vi𝔅(Vr,s)v_{i}\in\mathfrak{B}(V^{r,s}) and zjBr,sz_{j}\in B_{r,s}, there exists vk𝔅(Vr,s)v_{k}\in\mathfrak{B}(V^{r,s}) such that Jzjvi=vkJ_{z_{j}}v_{i}=v_{k} or Jzjvi=vkJ_{z_{j}}v_{i}=-v_{k}.

Definition 3.2 requires that the maps JzjJ_{z_{j}}, zjBr,sz_{j}\in B_{r,s} act on an invariant basis 𝔅(Vr,s)\mathfrak{B}(V^{r,s}) by permutations up to the sign ±\pm.

Remark 3.1.

We emphasise that we require bases 𝔅(Vr,s)\mathfrak{B}(V^{r,s}) to be both orthonormal and invariant.

Example A. Consider the Heisenberg Lie algebra 𝔫1,0(V)\mathfrak{n}_{1,0}(V) with the normalized basis B1,0={z}B_{1,0}=\{z\} for the centre and V1,02V^{1,0}\cong\mathbb{R}^{2}. Set v1V1,02,0v_{1}\in V^{1,0}\cong\mathbb{R}^{2,0}, and v2=Jzv1v_{2}=J_{z}v_{1}. Consider also

u1=Av1,u2=Av2,u_{1}=Av_{1},\quad u_{2}=Av_{2},

where AA is an orthogonal transformation of V1,0V^{1,0}. Then the basis (V1,0)={u1,u2}\mathfrak{(}V^{1,0})=\{u_{1},u_{2}\} is orthonormal. The basis (V1,0)={u1,u2}\mathfrak{(}V^{1,0})=\{u_{1},u_{2}\} will be invariant under the action of G(B1,0)G(B_{1,0}) if and only if JzJ_{z} commutes with AA. Thus we see that a basis 𝔅(V1,0)\mathfrak{B}(V^{1,0}) can be orthonormal, but not invariant under the action of G(B1,0)G(B_{1,0}).

Example B. Consider the Lie algebra 𝔫0,3(V)\mathfrak{n}_{0,3}(V) with an orthonormal basis B0,3={z1,z2,z3}B_{0,3}=\{z_{1},z_{2},z_{3}\} for the centre and a minimal admissible module V0,34,4V^{0,3}\cong\mathbb{R}^{4,4} of the Clifford algebra Cl 0,3\text{Cl}_{\,0,3}. We take vV0,3v\in V^{0,3}, such that v,vV0,3=1\langle v,v\rangle_{V^{0,3}}=1. The eight vectors

(3.4) v,Jz1v,Jz2v,Jz3v,Jz1Jz2v,Jz1Jz3v,Jz2Jz3v,Jz1Jz2Jz3vv,~{}J_{z_{1}}v,~{}J_{z_{2}}v,~{}J_{z_{3}}v,~{}J_{z_{1}}J_{z_{2}}v,~{}J_{z_{1}}J_{z_{3}}v,~{}J_{z_{2}}J_{z_{3}}v,~{}J_{z_{1}}J_{z_{2}}J_{z_{3}}v

are linearly independent, have square of the norm equal to ±1\pm 1, and invariant under the action of G(B0,3)G(B_{0,3}). Note that the value v,Jz1Jz2Jz3vV0,3=α\langle v,J_{z_{1}}J_{z_{2}}J_{z_{3}}v\rangle_{V^{0,3}}=\alpha is arbitrary and basis (3.4) is orthogonal if and only if α=0\alpha=0. Nevertheless, the vector vV0,3v\in V^{0,3} always can be chosen to make α=0\alpha=0, see [FM14, Lemmas 2.8, 2.9]. This is an example, when the basis 𝔅(V0,3)\mathfrak{B}(V^{0,3}) can be invariant, but not necessary orthonormal.

Proposition 3.3.

Let 𝔅(Vr,s)\mathfrak{B}(V^{r,s}) be an invariant basis. Then it is an integral basis.

Proof.

We claim that for any vVr,sv\in V^{r,s} with v,vVr,s0\langle v,v\rangle_{V^{r,s}}\neq 0 we have:

(3.5) Jziv=±Jzjv,zi=zj.J_{z_{i}}v=\pm J_{z_{j}}v,\quad\Longrightarrow\quad z_{i}=z_{j}.

Indeed, (3.5) implies JziJzjv=±vJ_{z_{i}}J_{z_{j}}v=\pm v and therefore (JziJzj)2v=v\big{(}J_{z_{i}}J_{z_{j}}\big{)}^{2}v=v. Assume by contrary that zizjz_{i}\neq z_{j}. Suppose first that both ziz_{i} and zjz_{j} are positive or negative. Then (JziJzj)2=Jzi2Jzj2=Id\big{(}J_{z_{i}}J_{z_{j}}\big{)}^{2}=-J_{z_{i}}^{2}J_{z_{j}}^{2}=-\text{\rm Id}, which is a contradiction. From the other side, if ziz_{i} and zjz_{j} are opposite, then

±v,±vVr,s=JziJzjv,JziJzjvVr,s=zi,zir,szj,zjr,sv,vVr,s=v,vVr,s\langle\pm v,\pm v\rangle_{V^{r,s}}=\langle J_{z_{i}}J_{z_{j}}v,J_{z_{i}}J_{z_{j}}v\rangle_{V^{r,s}}=\langle z_{i},z_{i}\rangle_{{r,s}}\langle z_{j},z_{j}\rangle_{{r,s}}\langle v,v\rangle_{V^{r,s}}=-\langle v,v\rangle_{V^{r,s}}

by (2.3), and vv must be a null vector, which is again a contradiction.

Assume now that 𝔅(Vr,s)={vj}\mathfrak{B}(V^{r,s})=\{v_{j}\} is an invariant basis for Vr,sV^{r,s} and that Jzvi=±vkJ_{z_{\ell}}v_{i}=\pm v_{k}. Then by definition of the Lie bracket (2.2) we obtain

z,[vi,vj]r,s=Jzvi,vjVr,s=±vk,vjVr,s=±δkj.\langle z_{\ell},[v_{i},v_{j}]\rangle_{r,s}=\langle J_{z_{\ell}}v_{i},v_{j}\rangle_{V^{r,s}}=\langle\pm v_{k},v_{j}\rangle_{V^{r,s}}=\pm\delta_{kj}.

If k=jk=j, then the orthonormality of Br,sB_{r,s} and z,[vi,vj]r,s=±1\langle z_{\ell},[v_{i},v_{j}]\rangle_{r,s}=\pm 1 imply that [vi,vj]=±z[v_{i},v_{j}]=\pm z_{\ell}, and the structure constants in (3.1) are such that cij=±1c_{ij}^{\ell}=\pm 1. If kjk\neq j then cij=0c_{ij}^{\ell}=0. ∎

The definition of an invariant basis leads to the definition of an invariant integral structure on pseudo HH-type Lie algebras and (invariant) integral uniform subgroup on the respective pseudo HH-type Lie groups.

Definition 3.4.

Let Br,s={zk}k=1r+sB_{r,s}=\{z_{k}\}_{k=1}^{r+s} be an orthonormal basis for r,s\mathbb{R}^{r,s} and 𝔅(Vr,s)={vi}i=1N\mathfrak{B}(V^{r,s})=\{v_{i}\}_{i=1}^{N} an invariant basis for a minimal admissible module Vr,sV^{r,s}. An invariant integral structure on the pseudo HH-type Lie algebra 𝔫r,s(V)\mathfrak{n}_{r,s}(V) is the vector space over \mathbb{Z} given by

span{𝔅(Vr,s)}span{Br,s}={i=1Nnivik=1r+smkzk|ni,mk}.\text{\rm span}\,_{\mathbb{Z}}\{\mathfrak{B}(V^{r,s})\}\oplus\text{\rm span}\,_{\mathbb{Z}}\{B_{r,s}\}=\Big{\{}\sum_{i=1}^{N}n_{i}v_{i}\oplus\sum_{k=1}^{r+s}m_{k}z_{k}~{}\Big{|}~{}n_{i},m_{k}\in\mathbb{Z}\Big{\}}.

An (invariant) integral uniform subgroup on the pseudo HH-type Lie group r,s(V)={(v,z)vVr,s,zr,s}\mathbb{N}_{r,s}(V)=\{(v,z)\mid v\in V^{r,s},z\in\mathbb{R}^{r,s}\} is given by the coordinates

((i=1Nnivini),(12k=1r+smkzkmk)).\left(\Big{(}\sum_{i=1}^{N}n_{i}v_{i}\mid n_{i}\in\mathbb{Z}\Big{)},\ \ \Big{(}\frac{1}{2}\sum_{k=1}^{r+s}m_{k}z_{k}\mid m_{k}\in\mathbb{Z}\Big{)}\right).

The main goal of the present work is the classification of invariant integral structures on pseudo HH-type Lie algebras that give rise to classification of integral uniform subgroups on the corresponding pseudo HH-type Lie groups. Note that invariant integral structures is a subclass of integral (not necessary invariant and/or orthonormal) structures on pseudo HH-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 r+s2r+s\leq 2, the invariant integral structures are unique. If (r,s){(1,0),(0,1)}(r,s)\in\{(1,0),(0,1)\} and z1z_{1} is a vector for r,s\mathbb{R}^{r,s} with |z1,z1r,s|=1|\langle z_{1},z_{1}\rangle_{r,s}|=1, then (Vr,s)={v,Jz1v}\mathcal{B}(V^{r,s})=\{v,J_{z_{1}}v\} is an invariant basis of the minimal admissible module Vr,sV^{r,s} for any choice of a vector vVr,sv\in V^{r,s} with v,vVr,s=1\langle v,v\rangle_{V^{r,s}}=1. Thus {z1,v,Jz1v}\{z_{1},v,J_{z_{1}}v\} gives rise to an invariant integral structure of 𝔫r,s(Vr,s)\mathfrak{n}_{r,s}(V^{r,s}) as in Definition 3.4. The Lie algebras 𝔫1,0\mathfrak{n}_{1,0} and 𝔫0,1\mathfrak{n}_{0,1} are not isometric, but they are both isomorphic to the Heisenberg Lie algebra.

If (r,s){(2,0),(1,1),(0,2)}(r,s)\in\{(2,0),(1,1),(0,2)\} and Br,s={z1,z2}B_{r,s}=\{z_{1},z_{2}\} is an orthonormal basis of r,s\mathbb{R}^{r,s}, then (Vr,s)={v,Jz1v,Jz2v,Jz1Jz2v}\mathcal{B}(V^{r,s})=\{v,J_{z_{1}}v,J_{z_{2}}v,J_{z_{1}}J_{z_{2}}v\} is an invariant basis of the minimal admissible module Vr,sV^{r,s} for any choice of vVr,sv\in V^{r,s}, v,vVr,s=1\langle v,v\rangle_{V^{r,s}}=1. The bases {Br,s,(Vr,s)}\{B_{r,s},\mathcal{B}(V^{r,s})\} generate a unique invariant integral structure of the respective HH-type Lie algebras. By uniqueness we mean that for any choice of orthonormal basis Br,sB_{r,s} and any vVr,sv\in V^{r,s} as above the invariant integral structures of the pseudo HH-type Lie algebras will give the isomorphic invariant uniform subgroups in the pseudo HH-type Lie groups. The proof is a simplified version of Theorem 6.2.

3.2. A subgroup 𝒮G(Br,s)\mathcal{S}\subset G(B_{r,s}) of positive involutions

In the present section we study subgroups 𝒮\mathcal{S} of G(Br,s)Clr,sG(B_{r,s})\subset\text{Cl}_{r,s} which will be a core for the construction of invariant bases 𝔅(Vr,s)\mathfrak{B}(V^{r,s}). Some of the properties of 𝒮\mathcal{S} can be learned from the definition of the subgroups 𝒮\mathcal{S}, but some of them became clear by considering their action on minimal admissible modules Vr,sV^{r,s}.

Recall that the group Pin(r,s)\text{\rm Pin}(r,s) consists of elements of the Clifford algebra of the form

(3.6) σ=xi1xik,xij,xijr,s=±1.\sigma=x_{i_{1}}\cdots x_{i_{k}},\quad\langle x_{i_{j}},x_{i_{j}}\rangle_{r,s}=\pm 1.

The subgroup Spin(r,s)Pin(r,s)\text{\rm Spin}(r,s)\subset\text{\rm Pin}(r,s) is generated by the even number of elements in (3.6). Thus the group G(Br,s)G(B_{r,s}) is a finite subgroup of Pin(r,s)\text{\rm Pin}(r,s).

Definition 3.5.

We denote by 𝒮\mathcal{S} a subgroup of G(Br,s)G(B_{r,s}) satisfying the conditions

  • (S1)(S1)

    𝟏𝒮-{\mathbf{1}}\notin\mathcal{S};

  • (S2)(S2)

    pPin(r,0)×Spin(0,s)p\in\text{\rm Pin}(r,0)\times\text{\rm Spin}(0,s) and

  • (S3)(S3)

    p2=𝟏p^{2}={\mathbf{1}}.

Elements p𝒮p\in\mathcal{S} are called positive involutions.

The name positive involution refers to the action of p𝒮p\in\mathcal{S} on Vr,sV^{r,s}: if v,vVr,s>0\langle v,v\rangle_{V^{r,s}}>0 (v,vVr,s<0\langle v,v\rangle_{V^{r,s}}<0) then Jpv,JpvVr,s>0\langle J_{p}v,J_{p}v\rangle_{V^{r,s}}>0 (Jpv,JpvVr,s<0\langle J_{p}v,J_{p}v\rangle_{V^{r,s}}<0). We denote by 𝕊r,s\mathbb{S}_{r,s} (or just 𝕊\mathbb{S}), the set of all subgroups of G(Br,s)G(B_{r,s}) satisfying Definition 3.5. This set is a partially ordered set with respect to the inclusion relation among subsets.

Remark 3.3.

The groups 𝒮𝕊r,s\mathcal{S}\in\mathbb{S}_{r,s} are necessarily commutative.

Example 3.1.

Consider G(B4,0)G(B_{4,0}). Then the example of possible subgroups 𝒮\mathcal{S} are

𝒮1={𝟏,z1z2z3},𝒮2={𝟏,z1z2z4},𝒮3={𝟏,z1z3z4},𝒮4={𝟏,z1z2z4}\mathcal{S}_{1}=\{{\mathbf{1}},z_{1}z_{2}z_{3}\},\ \mathcal{S}_{2}=\{{\mathbf{1}},z_{1}z_{2}z_{4}\},\ \mathcal{S}_{3}=\{{\mathbf{1}},z_{1}z_{3}z_{4}\},\ \mathcal{S}_{4}=\{{\mathbf{1}},-z_{1}z_{2}z_{4}\}

and

𝒮5={𝟏,z1z2z3z4}.\mathcal{S}_{5}=\{{\mathbf{1}},z_{1}z_{2}z_{3}z_{4}\}.

The first four groups are isomorphic under the action of the orthogonal group O(4)\text{\rm O}(4). A map CO(4)C\in\text{\rm O}(4) permutes the basis vectors {zi}\{z_{i}\}, i=1,2,3,4i=1,2,3,4 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 𝔅(V4,0)\mathfrak{B}(V^{4,0}).

To avoid the ambiguity occurring with the very similar groups 𝒮2\mathcal{S}_{2} and 𝒮4\mathcal{S}_{4}, we define a bigger group.

Definition 3.6.

Let 𝒮\mathcal{S} be a group from Definition 3.5. We denote by 𝒮^G(Br,s)\mathaccent 866{\mathcal{S}}\subset G(B_{r,s}) the extended group

𝒮^=𝒮{σ:σ𝒮}.\mathaccent 866{\mathcal{S}}=\mathcal{S}\cup\{-\sigma:\ \sigma\in\mathcal{S}\}.

In Example 3.1 we have 𝒮2,𝒮4\mathcal{S}_{2},\mathcal{S}_{4} subgroups of G(B4,0)G(B_{4,0}), where we fix the basis {z1,z2,z3,z4}\{z_{1},z_{2},z_{3},z_{4}\}. The subgroups 𝒮2,𝒮4\mathcal{S}_{2},\mathcal{S}_{4} are different, nevertheless

𝒮4^=𝒮2^={±𝟏,±z1z2z4}.\mathaccent 866{\mathcal{S}_{4}}=\mathaccent 866{\mathcal{S}_{2}}=\{\pm{\bf 1},\pm z_{1}z_{2}z_{4}\}.

3.3. Generators for a group 𝒮\mathcal{S} of positive involutions

In this section, we study groups 𝒮𝕊\mathcal{S}\in\mathbb{S} by describing their generating sets.

Definition 3.7.

We denote by PI={pi}i=1PI=\{\,p_{i}\,\}_{i=1}^{\ell}, =#[PI]\ell=\#[PI] is the cardinality of the set PIPI, a subset in G(Br,s)G(B_{r,s}) satisfying the conditions:

  • (PI1)(PI1)

    𝟏PI{\mathbf{1}}\notin PI, pipj=pjpip_{i}p_{j}=p_{j}p_{i} for iji\neq j, and piPIp_{i}\in PI satisfy (S2)(S3)(S2)-(S3) in Definition 3.5 for all i=1,,i=1,\ldots,\ell.

  • (PI2)(PI2)

    The vectors

    (3.7) { 1,p1,,p,pi1pik 1i1<<ik,k=1,,}\{\,{\mathbf{1}},p_{1},\ \ldots,\ p_{\ell},\ \ p_{i_{1}}\cdots p_{i_{k}}\mid\ 1\leq i_{1}<\ldots<i_{k}\leq\ell,\ k=1,\ldots,\ell\}

    are linearly independent in the vector space Clr,s\text{Cl}_{\,r,s}.

Proposition 3.8.

The condition (PI2)(PI2) is equivalent to

  • (PI2)(PI2)^{\prime}

    non of the products pi1pikp_{i_{1}}\cdots p_{i_{k}}, 1i1<<ik=#[PI]1\leq i_{1}<\cdots<i_{k}\leq\ell=\#[PI], k=1,,k=1,\ldots,\ell, is equal to ±𝟏\pm\mathbf{1}.

Proof.

Recall that the elements

(3.8) {ϵ0𝟏,ϵi1,,ikzi1zik}Clr,s,\{\epsilon_{0}\mathbf{1},\ \epsilon_{i_{1},\ldots,i_{k}}z_{i_{1}}\cdots z_{i_{k}}\}\ \subset\text{Cl}_{r,s},

1i1<<ikr+s1\leq i_{1}<\cdots<i_{k}\leq r+s, k=1,,r+sk=1,\ldots,r+s, where ϵ0\epsilon_{0} and ϵi1,,ik\epsilon_{i_{1},\ldots,i_{k}} can be chosen to be “++” or “-”, form a basis for Clr,s\text{Cl}_{r,s}.

It is obvious that (PI2)(PI2) implies (PI2)(PI2)^{\prime}. Assume that the condition (PI2)(PI2)^{\prime} is fulfild. Then the collection in (PI2)(PI2)^{\prime} is a reduced collection of linearly independent basis vectors from (3.8), and therefore they are linearly independent. ∎

As an example of a set PIPI we present the minimal length positive involutions, which can be classified in the following types:

T1{p=zi1zi2zi3zi4,where all zik are positive basis vectors;p=zi1zi2zi3zi4,where all zik are negative basis vectors;p=zi1zi2zi3zi4,where two zik are positive and two zil are negative basis vectors;\displaystyle T_{1}\,\,\begin{cases}p=z_{i_{1}}z_{i_{2}}z_{i_{3}}z_{i_{4}},~{}\text{where all $z_{i_{k}}$ are positive basis vectors;}\\ p=z_{i_{1}}z_{i_{2}}z_{i_{3}}z_{i_{4}},~{}\text{where all $z_{i_{k}}$ are negative basis vectors;}\\ p=z_{i_{1}}z_{i_{2}}z_{i_{3}}z_{i_{4}},~{}\text{where two $z_{i_{k}}$ are positive and two $z_{i_{l}}$}\\ \text{\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad are negative basis vectors;}\end{cases}
T2{q=zi1zi2zi3,where all zik are positive basis vectors;q=zi1zi2zi3,where one zik is positive and two zil are negative basis vectors.\displaystyle T_{2}\,\,\begin{cases}q=z_{i_{1}}z_{i_{2}}z_{i_{3}},~{}\text{where all $z_{i_{k}}$ are positive basis vectors;}\\ q=z_{i_{1}}z_{i_{2}}z_{i_{3}},~{}\text{where one $z_{i_{k}}$ is positive and two $z_{i_{l}}$}\\ \text{\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad are negative basis vectors.}\end{cases}

An easy combinatorial computation shows that generally positive involutions can contain either 3mod 43\,{\rm mod}\,4 or 4mod 44\,{\rm mod}\,4 basis vectors. This observation inspires us to make a more general definition.

Definition 3.9.

A positive involution containing 4mod 44\,{\rm mod}\,4 basis vectors is called a type T1T_{1} involution. A positive involution containing 3mod 43\,{\rm mod}\,4 basis vectors is called a type T2T_{2} involution.

Notation 3.1.

For an element σ=±zi1zikG(Br,s)\sigma=\pm\,z_{i_{1}}\cdots z_{i_{k}}\in G(B_{r,s}), we denote by 𝔟(σ)={zi1,,zik}\mathfrak{b}(\sigma)=\{z_{i_{1}},\ldots,z_{i_{k}}\} the set of the vectors in the product σ\sigma, and by |𝔟(σ)||\mathfrak{b}(\sigma)| we denote the number of the vectors in 𝔟(σ)\mathfrak{b}(\sigma). Analogously, 𝔟+(σ)\mathfrak{b^{+}}(\sigma) (𝔟(σ)\mathfrak{b^{-}}(\sigma)) is the set of positive ((negative)) vectors in σ\sigma and |𝔟+(σ)||\mathfrak{b^{+}}(\sigma)| (|𝔟(σ)|)(|\mathfrak{b^{-}}(\sigma)|) is the cardinality of the respective sets.

Proposition 3.10.

The following properties can be easily verified

  • (A)

    Two type T1T_{1} involutions p1p_{1} and p2p_{2} commute if the number |𝔟(p1)𝔟(p2)||\mathfrak{b}(p_{1})\cap\mathfrak{b}(p_{2})| is even. The product p1p2p_{1}p_{2} is an involution of type T1T_{1}.

  • (B)

    A type T1T_{1} involution pp and a type T2T_{2} involution qq commute if the number |𝔟(p)𝔟(q)||\mathfrak{b}(p)\cap\mathfrak{b}(q)| is even. The product pqpq is an involution of type T2T_{2}.

  • (C)

    Two type T2T_{2} involutions q1q_{1} and q2q_{2} commute if the number |𝔟(q1)𝔟(q2)||\mathfrak{b}(q_{1})\cap\mathfrak{b}(q_{2})| is odd. The product q1q2q_{1}q_{2} is an involution of type T1T_{1}.

Proof.

The proof is based on the Clifford algebra property

z1z2+z2z1=2z1,z2r,s𝟏,z1,z2r,s,z_{1}z_{2}+z_{2}z_{1}=-2\langle z_{1},z_{2}\rangle_{r,s}{\mathbf{1}},\quad z_{1},z_{2}\in\mathbb{R}^{r,s},

which for orthogonal vectors z1z_{1} and z2z_{2} leads to z1z2=z2z1z_{1}z_{2}=-z_{2}z_{1}. ∎

Notation 3.2.

We denote by 𝕀r,s\mathbb{PI}_{r,s} the collection of sets PIPI satisfying Definition 3.7. The set 𝕀r,s\mathbb{PI}_{r,s} is partially ordered by the inclusion relation similar to 𝕊r,s\mathbb{S}_{r,s}. If PI𝕀r,sPI\in\mathbb{PI}_{r,s}, then we denote by 𝒮(PI)\mathcal{S}(PI) a group generated by the set PIPI.

Proposition 3.11.
  1. (1)

    Let PI𝕀PI\in\mathbb{PI}. Then

    (3.9) 𝒮(PI)\displaystyle\mathcal{S}(PI) =\displaystyle= {𝟏,p1,,p,pi1pik| 1i1<<ik\displaystyle\{{\mathbf{1}},\ p_{1},\ldots,p_{\ell},\ p_{i_{1}}\cdots p_{i_{k}}|\ 1\leq i_{1}<\cdots<i_{k}\leq\ell
    1i1<<ik=#[PI]}\displaystyle 1\leq i_{1}<\cdots<i_{k}\leq\ell=\#[PI]\}

    is a group of order #[𝒮(PI)]=2#[PI]\#[\,\mathcal{S}(PI)\,]=2^{\#[PI]} in G(Br,s)G(B_{r,s}) and 𝒮(PI)𝕊\mathcal{S}(PI)\in\mathbb{S}.

  2. (2)

    Conversely, let 𝒮𝕊\mathcal{S}\in\mathbb{S}. Then there is a (non unique) set PI𝕀PI\in\mathbb{PI} such that 𝒮(PI)=𝒮\mathcal{S}(PI)=\mathcal{S}.

  3. (3)

    Let ε=(ε1,,ε)\varepsilon=(\varepsilon_{1},\ldots,\varepsilon_{\ell}) be a tuple consisting of ±1\pm 1, and PI={pi}i=1𝕀r,sPI=\{p_{i}\}_{i=1}^{\ell}\in\mathbb{PI}_{r,s}. Then εPI={ε1p1,,εp}𝕀r,s\varepsilon\cdot PI=\{\varepsilon_{1}p_{1},\ldots,\varepsilon_{\ell}p_{\ell}\}\in\mathbb{PI}_{r,s} and 𝒮(PI)^=𝒮(εPI)^\mathaccent 1371{\mathcal{S}(PI)}=\mathaccent 1371{\mathcal{S}(\varepsilon\cdot PI)}.

Proof.

Set in (3.7) is linearly independent and coincides with 𝒮(PI)\mathcal{S}(PI) in (3.9), therefore #[𝒮(PI)]=2#[PI]\#[\mathcal{S}(PI)]=2^{\#[PI]}. If pp is in the set (3.7), then p-p is not in the set (3.7), which implies that 𝟏𝒮(PI)-{\mathbf{1}}\notin\mathcal{S}(PI). Any p𝒮(PI)p\in\mathcal{S}(PI) is a positive involution by definition of the set PIPI. We showed (1).

The second property will be proved by induction arguments with respect to the order of the group 𝒮\mathcal{S}. Let 𝒮𝕊r,s\mathcal{S}\in\mathbb{S}_{r,s} be given. Assume p1𝒮p_{1}\in\mathcal{S} and if there are no elements in 𝒮\mathcal{S} other than 𝟏,p1{\mathbf{1}},p_{1}, then we can put PI={p1}PI=\{p_{1}\} and 𝒮(PI)=𝒮\mathcal{S}(PI)=\mathcal{S}.

Assume now that there is a set PI={p1,,p}2PI^{\prime}=\{p_{1},\ldots,p_{\ell}\}_{\ell\geq 2} satisfying Definition 3.7. If

𝒮(PI)={𝟏,p1,,p,pi1pik 1i1<<ik,k=1,,},\mathcal{S}(PI^{\prime})=\{{\mathbf{1}},\ p_{1},\ldots,p_{\ell},\ \ p_{i_{1}}\cdots p_{i_{k}}\mid\ 1\leq i_{1}<\cdots<i_{k}\leq\ell,\ k=1,\ldots,\ell\},

is a proper subset of 𝒮\mathcal{S}, then there is a positive involution q𝒮q\in\mathcal{S} such that q𝒮(PI)q\not\in\mathcal{S}(PI^{\prime}), and q± 1q\not=\pm\,{\mathbf{1}}. Consider the set of commuting involutions

𝒮(PI)q={q,p1q,,pq,pi1pikq 1i1<<ik,k=1,,}.\mathcal{S}(PI^{\prime})\cdot q=\{q,\ p_{1}q,\ldots,p_{\ell}q,\ \ p_{i_{1}}\cdots p_{i_{k}}q\mid\ 1\leq i_{1}<\cdots<i_{k}\leq\ell,\ k=1,\ldots,\ell\}.

If pi1pim=pj1pjmqp_{i_{1}}\cdots p_{i_{m}}=p_{j_{1}}\cdots p_{j_{m^{\prime}}}q, then q𝒮(PI)q\in\mathcal{S}(PI^{\prime}), as a product of involutions pj1pjmp_{j_{1}}\cdots p_{j_{m^{\prime}}} and pi1pimp_{i_{1}}\cdots p_{i_{m}} from 𝒮(PI)\mathcal{S}(PI^{\prime}). Thus non of the elements in 𝒮(PI)\mathcal{S}(PI^{\prime}) can be written in the form pj1pjmqp_{j_{1}}\cdots p_{j_{m^{\prime}}}q for pj1pjm𝒮(PI)p_{j_{1}}\cdots p_{j_{m^{\prime}}}\in\mathcal{S}(PI^{\prime}). If

pi1pikpj1pjkforpi1pik,pj1pjk𝒮(PI),p_{i_{1}}\cdots p_{i_{k}}\neq p_{j_{1}}\cdots p_{j_{k^{\prime}}}\quad\text{for}\quad p_{i_{1}}\cdots p_{i_{k}},~{}p_{j_{1}}\cdots p_{j_{k^{\prime}}}\in\mathcal{S}(PI^{\prime}),

then pi1pikqpj1pjkqp_{i_{1}}\cdots p_{i_{k}}q\neq p_{j_{1}}\cdots p_{j_{k^{\prime}}}q. So the set PI′′=PI{q}PI^{\prime\prime}=PI^{\prime}\cup\{q\} satisfies Definition 3.7.

Continuing the procedure, we find in finitely many steps a set PIPI satisfying Definition 3.7 such that 𝒮(PI)=𝒮\mathcal{S}(PI)=\mathcal{S}.

The proof of the last assertion is easily follows from Definition 3.7. ∎

3.4. Relation of 𝒮\mathcal{S} and an isotropy subgroup 𝒮v\mathcal{S}_{v}

Now we relate a group 𝒮\mathcal{S} with the isotropy subgroup 𝒮v\mathcal{S}_{v} for some vVr,sv\in V^{r,s} and show that they are in a close relation.

Proposition 3.12.

Let vVr,sv\in V^{r,s} be a non-null vector and let 𝒮v\mathcal{S}_{v} denote the isotropy subgroup in G(Br,s)G(B_{r,s}) of the vector vv:

𝒮v={σG(Br,s)Jσv=v}.\mathcal{S}_{v}=\{\sigma\in G(B_{r,s})\mid J_{\sigma}v=v\}.

Then 𝒮v\mathcal{S}_{v} satisfies Definition 3.5.

Proof.

It is clear that 𝟏𝒮v-{\mathbf{1}}\notin\mathcal{S}_{v}. To check the second property we take σ𝒮vG(Br,s)\sigma\in\mathcal{S}_{v}\subset G(B_{r,s}) and assume by contrary that σ\sigma is a product containing an odd number of negative basis vectors from Br,sB_{r,s}. Then for vVr,sv\in V^{r,s} with v,vVr,s>0\langle v,v\rangle_{V^{r,s}}>0 we obtain

0<v,vVr,s=Jσv,JσvVr,s<00<\langle v,v\rangle_{V^{r,s}}=\langle J_{\sigma}v,J_{\sigma}v\rangle_{V^{r,s}}<0

by (2.3), which is a contradiction. Similar argument is applied for a vector vVr,sv\in V^{r,s} with v,vVr,s<0\langle v,v\rangle_{V^{r,s}}<0. Hence σPin(r,0)×Spin(0,s)\sigma\in\text{\rm Pin}(r,0)\times\text{\rm Spin}(0,s).

The square of every element in G(Br,s)G(B_{r,s}) equal either 𝟏{\mathbf{1}} or 𝟏-{\mathbf{1}}. If σ𝒮v\sigma\in\mathcal{S}_{v}, then Jσ2=IdVr,sJ_{\sigma}^{2}=\text{\rm Id}_{V^{r,s}}. Hence σ2=𝟏{\sigma}^{2}={\mathbf{1}}. ∎

The relation of an arbitrary 𝒮\mathcal{S} to an isotropy group 𝒮v\mathcal{S}_{v} for some vVr,sv\in V^{r,s} is given in the following statement.

Proposition 3.13.

Let 𝒮𝕊r,s\mathcal{S}\in\mathbb{S}_{r,s} and PI={p1,,p}𝕀r,sPI=\{p_{1},\ldots,p_{\ell}\}\in\mathbb{PI}_{r,s} be such that 𝒮(PI)=𝒮\mathcal{S}(PI)=\mathcal{S}. Let E+1(pk)={uVr,sJpku=u}E^{+1}(p_{k})=\{u\in V^{r,s}\mid J_{p_{k}}u=u\}. Then the intersection k=1E+1(pk)\bigcap_{k=1}^{\ell}E^{+1}(p_{k}) contains a non-null vector vv. Moreover, the group 𝒮(PI)\mathcal{S}(PI) is the isotropy subgroup 𝒮v\mathcal{S}_{v} of the vector vv, and #[𝒮]=#[𝒮v]=2#[PI]\#[\mathcal{S}]=\#[\mathcal{S}_{v}]=2^{\#[PI]}.

If rs=3mod 4r-s=3\,{\rm mod}\,4, and there is piPIp_{i}\in PI such that JpiJ_{p_{i}} acts as Id-\text{\rm Id} on the minimal admissible module Vr,sV^{r,s}, then the change pip_{i} to pi-p_{i} leads to the above statement.

Proof.

Let rs3mod 4r-s\neq 3\,{\rm mod}\,4 and let E+1(pk)E^{+1}(p_{k}), E1(pk)E^{-1}(p_{k}) be the eigenspaces of an involution JpkJ_{p_{k}} with eigenvalue 11 and 1-1, respectively. If one of the spaces E±1(pk)E^{\pm 1}(p_{k}) is trivial, then the symmetric bi-linear form .,.Vr,s\langle.\,,.\rangle_{V^{r,s}} on the non-trivial subspace is non-degenerate. If both of E±1(pk)E^{\pm 1}(p_{k}) are non-trivial spaces, then they are orthogonal with respect to .,.Vr,s\langle.\,,.\rangle_{V^{r,s}} and the restriction of .,.Vr,s\langle.\,,.\rangle_{V^{r,s}} onto E±1(pk)E^{\pm 1}(p_{k}) is non-degenerate too.

Assume E+1(p1){0}E^{+1}(p_{1})\not=\{0\}. Then the space E+1(p1)E^{+1}(p_{1}) is invariant under the action of the involution Jp2J_{p_{2}}. Therefore, E+1(p1)E+1(p2){0}E^{+1}(p_{1})\bigcap E^{+1}(p_{2})\neq\{0\}. By repeating the procedures we get that E=k=1E+1(pk){0}E=\bigcap_{k=1}^{\ell}\,E^{+1}(p_{k})\not=\{0\} and the restriction of .,.Vr,s\langle.\,,.\rangle_{V^{r,s}} onto EE is non-degenerate. Thus there is a non-null vector vEv\in E such that Jpkv=vJ_{p_{k}}v=v for all k=1,,k=1,\ldots,\ell. Hence 𝒮(PI)=𝒮v\mathcal{S}(PI)=\mathcal{S}_{v}.

If rs=3mod 4r-s=3\,{\rm mod}\,4, then without loss of generality we can assume that Jp1J_{p_{1}} acts as Id-\text{\rm Id}. We change p1p_{1} to p1-p_{1} to get E+1(p1)={uVr,sJp1u=u}E^{+1}(p_{1})=\{u\in V^{r,s}\mid J_{p_{1}}u=u\} and continue the proof as above. ∎

Corollary 3.14.

Let 𝒮𝕊r,s\mathcal{S}\in\mathbb{S}_{r,s}, and let 𝒮v=𝒮\mathcal{S}_{v}=\mathcal{S} be an isotropy subgroup of vv as in Proposition 3.13. The orbit Ov=G(Br,s).vO_{v}={{G(B_{r,s}).v}}, defined in (3.3), contains an invariant basis 𝔅(Vr,s)\mathfrak{B}(V^{r,s}) of the minimal admissible module Vr,sV^{r,s}. There is no canonical way to prescribe the direction uu or u-u for a basis vector in 𝔅(Vr,s)\mathfrak{B}(V^{r,s}). Therefore OvO_{v} is a set of basis vectors counted with signes ±\pm. Hence G(Br,s)/SvG(Br,s).vG(B_{r,s})/S_{v}\cong{G(B_{r,s}).v} and dim(Vr,s)=12#[G(Br,s).v]\dim(V^{r,s})=\frac{1}{2}\#[G(B_{r,s}).v].

Proof.

If the group 𝒮v\mathcal{S}_{v} is an isotropy subgroup of an invariant basis, then

(3.10) #[𝒮v]#[G(Br,s).v]=2r+s+1=#[G(Br,s)].\#[\mathcal{S}_{v}]\cdot\#[G(B_{r,s}).v]=2^{r+s+1}=\#[G(B_{r,s})].

Since the module is minimal admissible and the basis vectors are counted twice (with plus and minus signs), we conclude #[G(Br,s).v]=2dim(Vr,s)\#[G(B_{r,s}).v]=2\dim(V^{r,s}). ∎

Remark 3.4.

We denote by 𝕊r,sM\mathbb{S}^{M}_{r,s} the subset in 𝕊r,s\mathbb{S}_{r,s} consisting of subgroups 𝒮=𝒮(PI)\mathcal{S}=\mathcal{S}(PI) satisfying (3.10). Furthermore 𝕀r,sM\mathbb{PI}^{M}_{r,s} denotes the maximal set of PIPI: that is 𝒮(PI)𝕊r,sM\mathcal{S}(PI)\in\mathbb{S}^{M}_{r,s} if and only if PI𝕀r,sMPI\in\mathbb{PI}_{r,s}^{M}, see Proposition 3.11. Note that the correspondence from 𝕀r,s\mathbb{PI}_{r,s} to 𝕊r,s\mathbb{S}_{r,s}, assigning PI𝒮(PI)PI\mapsto\mathcal{S}(PI) is surjective but not necessarily injective.

In Proposition 3.13, if 𝒮(PI)𝕊r,sM\mathcal{S}(PI)\in\mathbb{S}^{M}_{r,s}, then 𝒮(PI)=𝒮v𝕊r,sM\mathcal{S}(PI)=\mathcal{S}_{v}\in\mathbb{S}^{M}_{r,s}. Indeed, since PI𝕀r,sMPI\in\mathbb{PI}^{M}_{r,s} if and only if 𝒮=𝒮(PI)𝕊r,sM\mathcal{S}=\mathcal{S}(PI)\in\mathbb{S}^{M}_{r,s}, we obtain 𝒮(PI)=𝒮v𝕊r,sM\mathcal{S}(PI)=\mathcal{S}_{v}\in\mathbb{S}^{M}_{r,s}.

Notation 3.3.

We denote by (r,s)\ell(r,s) the maximal number of involutions in a set PIr,s𝕀r,sMPI_{r,s}\in\mathbb{PI}_{r,s}^{M}. The value (r,s)\ell(r,s) depends only on the signature (r,s)(r,s) and it satisfies 2(r,s)=2r+sdim(Vr,s)2^{\ell(r,s)}=\frac{2^{r+s}}{\dim(V^{r,s})} by Corollary 3.14.

The orbit Ov=G(Br,s).vO_{v}={G(B_{r,s}).v} gives the invariant basis for Vr,sV^{r,s} up to a sign. Since the elements in G(Br,s)G(B_{r,s}) either commute or anti-commute with elements in 𝒮v\mathcal{S}_{v}, we can more precisely describe the construction of an invariant basis for a minimal admissible module Vr,sV^{r,s}.

Theorem 3.15.

Let vVr,sv\in V^{r,s} be a unit vector from Proposition 3.13. There is a set ΣG(Br,s)\Sigma\subset G(B_{r,s}) such that the family {Jσv}σΣ\{J_{\sigma}v\}_{\sigma\in\Sigma} is an invariant basis of Vr,sV^{r,s}.

Proof.

Let 𝒮v𝕊r,sM\mathcal{S}_{v}\in\mathbb{S}_{r,s}^{M}. We fix a maximal set PIr,s={pi}i=1(r,s)PI_{r,s}=\{p_{i}\}_{i=1}^{\ell(r,s)} such that 𝒮(PIr,s)=𝒮v\mathcal{S}(PI_{r,s})=\mathcal{S}_{v} and write Eεi(pi)={vVr,sJpiv=εiv}E^{\varepsilon_{i}}(p_{i})=\{v\in V^{r,s}\mid J_{p_{i}}v=\varepsilon_{i}v\}, where εi\varepsilon_{i} is either +1+1 or 1-1. We denote ε=(ε1,,ε(r,s))\varepsilon=(\varepsilon_{1},\ldots,\varepsilon_{\ell(r,s)}) and define

(3.11) E=i=1(r,s)E+1(pi),Eε1,,ε(r,s)=i=1(r,s)Eεi(pi).E=\bigcap_{i=1}^{\ell(r,s)}\,E^{+1}(p_{i}),\quad E^{\varepsilon_{1},\ldots,\varepsilon_{\ell(r,s)}}=\bigcap_{i=1}^{\ell(r,s)}\,E^{\varepsilon_{i}}(p_{i}).

Before we continue the proof we note that dim(E){1,2,4,8}\dim(E)\in\{1,2,4,8\}, and either dim(Vr,s)=dim(E)×2(r,s)\dim(V^{r,s})=\dim(E)\times 2^{\ell(r,s)} or dim(Vr,s)=dim(E)×2(r,s)1\dim(V^{r,s})=\dim(E)\times 2^{\ell(r,s)-1}. In the latter case, one involution JpiJ_{p_{i}} acts as Id or Id-\text{\rm Id} on Vr,sV^{r,s}, which happens if rs=3mod4r-s=3\mod 4, see details in [FM21]. Thus

dim(E)=2r+s2(r,s)ordim(E)=2r+s2((r,s)1).\dim(E)=2^{r+s-2\ell(r,s)}\quad\text{or}\quad\dim(E)=2^{r+s-2(\ell(r,s)-1)}.

Let 𝐂G(Br,s)(𝒮(PIr,s)){\mathbf{C}}_{G(B_{r,s})}(\mathcal{S}(PI_{r,s})) be the centralizer of the subgroup 𝒮(PIr,s)\mathcal{S}(PI_{r,s}) in G(Br,s)G(B_{r,s}). Then by choosing a unit vector vEv\in E, we can find representatives {σi}i=1dim(E)𝐂G(Br,s)(𝒮(PIr,s))/𝒮(PIr,s)^\{\sigma_{i}\}_{i=1}^{\dim(E)}\in{\mathbf{C}}_{G(B_{r,s})}\big{(}\mathcal{S}(PI_{r,s})\big{)}\big{/}\mathaccent 1371{\mathcal{S}(PI_{r,s})}, and {τj}j=12(r,s)G(Br,s)/𝐂G(Br,s)(𝒮(PIr,s))\{\tau_{j}\}_{j=1}^{2^{\ell(r,s)}}\in G(B_{r,s})\big{/}{\mathbf{C}}_{G(B_{r,s})}\big{(}\mathcal{S}(PI_{r,s})\big{)} such that

the vectors{Jσiv}i=1dim(E)form an orthonormal basis for E,the vectors{JτjJσiv}i=1dim(E)j=12(r,s) form an orthonormal basis for Vr,s.\begin{split}&\text{the vectors}\quad\{J_{\sigma_{i}}v\}_{i=1}^{\dim(E)}\quad\text{form an orthonormal basis for $E$},\\ &\text{the vectors}\quad\{J_{\tau_{j}}J_{\sigma_{i}}v\}_{i=1}^{\dim(E)}~{}_{j=1}^{2^{\ell(r,s)}}\quad\text{ form an orthonormal basis for $V^{r,s}$.}\end{split}

These {σi}i=1dim(E)\{\sigma_{i}\}_{i=1}^{\dim(E)} and {τj}j=12(r,s)\{\tau_{j}\}_{j=1}^{2^{\ell(r,s)}} form the set Σ\Sigma. ∎

Proposition 3.16.

Fix the group 𝒮(PIr,s)\mathcal{S}(PI_{r,s}) and the representatives

{σi}i=1dim(E)𝐂G(Br,s)(𝒮(PIr,s))/𝒮(PIr,s)^,\{\sigma_{i}\}_{i=1}^{\dim(E)}\in{\mathbf{C}}_{G(B_{r,s})}\big{(}\mathcal{S}(PI_{r,s})\big{)}\big{/}\mathaccent 1371{\mathcal{S}(PI_{r,s})},
{τj}j=12(r,s)G(Br,s)/𝐂G(Br,s)(𝒮(PIr,s)).\{\tau_{j}\}_{j=1}^{2^{\ell(r,s)}}\in G(B_{r,s})\big{/}{\mathbf{C}}_{G(B_{r,s})}\big{(}\mathcal{S}(PI_{r,s})\big{)}.

Assume that v1,v2Ev_{1},v_{2}\in E generate two sets of invariant bases

𝔅vk(Vr,s)={vk,Jσivk,Jτjvk,JτjJσivk}i=1dim(E),j=12(r,s)k=1,2,\mathfrak{B}_{v_{k}}(V^{r,s})=\{v_{k},\ J_{\sigma_{i}}v_{k},\ J_{\tau_{j}}v_{k},\ J_{\tau_{j}}J_{\sigma_{i}}v_{k}\}_{i=1}^{\dim(E)}\,{}_{j=1}^{2^{\ell(r,s)}},\quad k=1,2,

as in Theorem 3.15. Then the invariant integral structures

(3.12) span{𝔅v1(Vr,s)}span{Br,s}span{𝔅v2(Vr,s)}span{Br,s}\begin{array}[]{lll}&\text{\rm span}\,_{\mathbb{Z}}\{\mathfrak{B}_{v_{1}}(V^{r,s})\}\oplus\text{\rm span}\,_{\mathbb{Z}}\{B_{r,s}\}\\ \\ &\text{\rm span}\,_{\mathbb{Z}}\{\mathfrak{B}_{v_{2}}(V^{r,s})\}\oplus\text{\rm span}\,_{\mathbb{Z}}\{B_{r,s}\}\end{array}

are isomorphic.

Proof.

We define the correspondence A:𝔅v1(Vr,s)𝔅v2(Vr,s)A\colon\mathfrak{B}_{v_{1}}(V^{r,s})\to\mathfrak{B}_{v_{2}}(V^{r,s}) by

(3.13) v1v2,Jσiv1Jσiv2,Jτjv1Jτjv2,JτjJσiv1JτjJσiv2,\begin{array}[]{lllll}&v_{1}\mapsto v_{2},&J_{\sigma_{i}}v_{1}\mapsto J_{\sigma_{i}}v_{2},\\ &J_{\tau_{j}}v_{1}\mapsto J_{\tau_{j}}v_{2},&J_{\tau_{j}}J_{\sigma_{i}}v_{1}\mapsto J_{\tau_{j}}J_{\sigma_{i}}v_{2},\end{array}

and extend it by linearity over \mathbb{Z}. Then the map AIdA\oplus\text{\rm Id} is an automorphism of invariant integral structures (3.12). To show that AIdA\oplus\text{\rm Id} is an isomoprhism, we denote the basis vectors from 𝔅v1(Vr,s)\mathfrak{B}_{v_{1}}(V^{r,s}) by {uα}α=1dim(Vr,s)\{u_{\alpha}\}_{\alpha=1}^{\dim(V^{r,s})} and the basis vectors from 𝔅v2(Vr,s)\mathfrak{B}_{v_{2}}(V^{r,s}) by {wα}α=1dim(Vr,s)\{w_{\alpha}\}_{\alpha=1}^{\dim(V^{r,s})}, where wα=Auαw_{\alpha}=Au_{\alpha}. Then we note that the bases 𝔅v1(Vr,s)\mathfrak{B}_{v_{1}}(V^{r,s}) and 𝔅v2(Vr,s)\mathfrak{B}_{v_{2}}(V^{r,s}) are invariant, which means that for any uα𝔅v1(Vr,s)u_{\alpha}\in\mathfrak{B}_{v_{1}}(V^{r,s}) and any zkBr,sz_{k}\in B_{r,s} there is uβ𝔅v1(Vr,s)u_{\beta}\in\mathfrak{B}_{v_{1}}(V^{r,s}) such that

(3.14) Jzkuα=±uβ=±Jϰv1,for someϰΣ={σi,τj,τjσi}J_{z_{k}}u_{\alpha}=\pm u_{\beta}=\pm J_{\varkappa}v_{1},\quad\text{for some}\quad\varkappa\in\Sigma=\{\sigma_{i},\tau_{j},\tau_{j}\sigma_{i}\}

The correspondence (3.13) and (3.14) imply that for chosen uα𝔅v1(Vr,s)u_{\alpha}\in\mathfrak{B}_{v_{1}}(V^{r,s}) and zkBr,sz_{k}\in B_{r,s} we have

JzkAuα=Jzkwα=±wβ=±Jϰv2=±AJϰv1=AJzkuα.J_{z_{k}}Au_{\alpha}=J_{z_{k}}w_{\alpha}=\pm w_{\beta}=\pm J_{\varkappa}v_{2}=\pm AJ_{\varkappa}v_{1}=AJ_{z_{k}}u_{\alpha}.

Note also that AτA=IdVr,sA^{\tau}A=\text{\rm Id}_{V^{r,s}} since it maps an othonormal basis to an orthonormal basis. Then we have

[Auα,Auβ],zkr,s\displaystyle\langle[Au_{\alpha},Au_{\beta}],z_{k}\rangle_{r,s} =\displaystyle= JzkAuα,AuβVr,s=AJzkuα,AuβVr,s\displaystyle\langle J_{z_{k}}Au_{\alpha},Au_{\beta}\rangle_{V^{r,s}}=\langle AJ_{z_{k}}u_{\alpha},Au_{\beta}\rangle_{V^{r,s}}
=\displaystyle= AτAJzkuα,uβVr,s=Jzkuα,uβVr,s\displaystyle\langle A^{\tau}AJ_{z_{k}}u_{\alpha},u_{\beta}\rangle_{V^{r,s}}=\langle J_{z_{k}}u_{\alpha},u_{\beta}\rangle_{V^{r,s}}
=\displaystyle= [uα,uβ],zkr,s.\displaystyle\langle[u_{\alpha},u_{\beta}],z_{k}\rangle_{r,s}.

4. Equivalence and connectedness of groups 𝒮\mathcal{S}

We define an equivalence relation between groups 𝒮G(Br,s)\mathcal{S}\subset G(B_{r,s}) that will descend to the equivalence of their generating sets PIr,sPI_{r,s}. We also introduce parameters to distinguish sets PIr,sPI_{r,s} for a fixed value (r,s)(r,s). Different sets of parameters will lead to non-equivalent generating sets and the groups. Our aim is to show that equivalent groups 𝒮\mathcal{S} lead to the isomorphic invariant integral structures on 𝔫r,s\mathfrak{n}_{r,s}.

4.1. Equivalence of groups 𝒮\mathcal{S}

We recall Notation 3.1 and extend it to the sets PIPI.

Notation 4.1.

Let PI𝕀r,sPI\in\mathbb{PI}_{r,s}. We denote

𝔟+(PI)={ziziis a positive vector in some piPI},\displaystyle\mathfrak{b}^{+}(PI)=\{z_{i}\mid\ z_{i}\ \text{is a positive vector in some $p_{i}\in PI$}\},
𝔟(PI)={ziziis a negative vector in some piPI}.\displaystyle\mathfrak{b}^{-}(PI)=\{z_{i}\mid\ z_{i}\ \text{is a negative vector in some $p_{i}\in PI$}\}.

We set also |𝔟+(PI)||\mathfrak{b}^{+}(PI)|, |𝔟(PI)||\mathfrak{b}^{-}(PI)| for the cardinality of the respective set, and |𝔟(PI)|=|𝔟+(PI)|+|𝔟(PI)||\mathfrak{b}(PI)|=|\mathfrak{b}^{+}(PI)|+|\mathfrak{b}^{-}(PI)|.

Definition 4.1.

A set PIPI consisting only of the involutions of type T1T_{1} will be called (T1)(T1)-type set. A set PIPI consisting of the involutions of type T1T_{1} and having at least one involution of type T2T_{2} will be called (T2)(T2)-type set.

Proposition 4.2.

Any (T2)(T2)-type set can be reduced to (T2)(T2)-type set containing at most one involution of type T2T_{2} and the rest of involutions will be of type T1T_{1}.

Proof.

The proof follows directly from Proposition 3.10. ∎

Notation 4.2.

If CO(r,s)C\in\text{\rm O}(r,s), then we denote by the same letter CC its natural extension C:Clr,sClr,sC\colon\text{Cl}_{\,r,s}^{\,*}\to\text{Cl}_{\,r,s}^{\,*} to the action on the group of invertible elements Clr,sClr,s\text{Cl}_{\,r,s}^{\,*}\subset\text{Cl}_{\,r,s}.

Let Br,sB_{r,s} be a basis as in (3.2). Let CO(r,s)C\in\text{\rm O}(r,s). Then CC is a signed permutation matrix for Br,sB_{r,s} having only one nonzero component "±1""\pm 1" in each column. We call such a map (signed) re-ordering of Br,sB_{r,s}. If σ=zi1zikG(Br,s)\sigma=z_{i_{1}}\cdots z_{i_{k}}\in G(B_{r,s}), then CC defines an element C(σ):=C(zi1)C(zik)G(Br,s)C(\sigma):=C(z_{i_{1}})\cdots C(z_{i_{k}})\in G(B_{r,s}). Since a re-ordering matrix CC maps positive basis vectors to positive vectors and negative basis vectors to negative basis vectors, it induces a map C:𝕀r,s𝕀r,sC\colon\mathbb{PI}_{r,s}\to\mathbb{PI}_{r,s}. For the particular case (r,r)(r,r) the map CC can be chosen also to map positive basis vectors to negative vectors and vice versa. The changes for (r,r)(r,r) will be discussed separately in a forthcoming paper.

Definition 4.3.

We say that the groups 𝒮1\mathcal{S}_{1} and 𝒮2\mathcal{S}_{2} are equivalent, writing 𝒮1𝒮2\mathcal{S}_{1}\sim\mathcal{S}_{2}, if there is a map CO(r,s)C\in\text{\rm O}(r,s) such that its natural extention to Clr,sClr,s\text{Cl}_{\,r,s}^{*}\subset\text{Cl}_{\,r,s} gives the isomorphism between the extended groups 𝒮1^\mathaccent 866{\mathcal{S}_{1}} and 𝒮2^\mathaccent 866{\mathcal{S}_{2}}; that is C(𝒮1^)=𝒮2^C(\mathaccent 866{\mathcal{S}_{1}})=\mathaccent 866{\mathcal{S}_{2}}.

Definition 4.4.

Let PI1PI_{1} and PI2PI_{2} be two sets of involutions. Then we say that PI1PI_{1} and PI2PI_{2} are equivalent, writing PI1PI2PI_{1}\sim PI_{2}, if 𝒮(PI1)\mathcal{S}(PI_{1}) is equivalent to 𝒮(PI2)\mathcal{S}(PI_{2}) in the sense of Definition 4.3.

Example 4.1.

Recall Example 3.1 and consider G(B4,0)G(B_{4,0}). Let PI1={z1z2z3}PI_{1}=\{z_{1}z_{2}z_{3}\} and PI2={z1z2z4}PI_{2}=\{z_{1}z_{2}z_{4}\}. Then PI1PI2PI_{1}\sim PI_{2}, since the groups

𝒮(PI1)^={±𝟏,±z1z2z3}and𝒮(PI2)^={±𝟏,±z1z2z4}\mathaccent 1371{\mathcal{S}(PI_{1})}=\{\pm\mathbf{1},\ \pm z_{1}z_{2}z_{3}\}\quad\text{and}\quad\mathaccent 1371{\mathcal{S}(PI_{2})}=\{\pm\mathbf{1},\ \pm z_{1}z_{2}z_{4}\}

in Cl4,0\text{Cl}_{4,0} are isomorphic under O(4,0)\text{\rm O}(4,0) which permutes the basis vectors z3z_{3} and z4z_{4}, fixing z1z_{1} and z2z_{2}. Nevertheless, PI1PI_{1} is not equivalent to PI3={z1z2z3z4}PI_{3}=\{z_{1}z_{2}z_{3}z_{4}\}, since there is no extention of CO(4,0)C\in\text{\rm O}(4,0) to Clr,s\text{Cl}_{\,r,s}^{*} which maps 𝒮(PI1)^\mathaccent 1371{\mathcal{S}(PI_{1})} to 𝒮(PI3)^={±𝟏,±z1z2z3z4}Cl4,0\mathaccent 1371{\mathcal{S}(PI_{3})}=\{\pm\mathbf{1},\ \pm z_{1}z_{2}z_{3}z_{4}\}\subset\text{Cl}_{4,0}^{*}.

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 Br,sB_{r,s} be an orthonormal basis of r,s\mathbb{R}^{r,s}. We form a set of mutually different pairs

(4.1) πi,j=zizj,i<j,i,j{{1,,r}if r is even{1,,r1}if r is odd,\pi_{i,j}=z_{i}z_{j},\ \ i<j,\ \ i,j\in\begin{cases}\{1,\ldots,r\}&\quad\text{if\ $r$\ is even}\\ \{1,\ldots,r-1\}&\quad\text{if\ $r$\ is odd}\end{cases},
(4.2) νk,l=zkzl,k<l,k,l{{r+1,,s}if s is even{r+1,,s1}if s is odd,\nu_{k,l}=z_{k}z_{l},\ \ k<l,\ \ k,l\in\begin{cases}\{r+1,\ldots,s\}&\quad\text{if\ $s$\ is even}\\ \{r+1,\ldots,s-1\}&\quad\text{if\ $s$\ is odd}\end{cases},

and

𝔟(πi1,j1)𝔟(πi2,j2)=,𝔟(νk1,l1)𝔟(νk2,l2)=,\mathfrak{b}(\pi_{i_{1},j_{1}})\cap\mathfrak{b}(\pi_{i_{2},j_{2}})=\emptyset,\quad\mathfrak{b}(\nu_{k_{1},l_{1}})\cap\mathfrak{b}(\nu_{k_{2},l_{2}})=\emptyset,

The cardinalities of the sets of pairs are

𝐩=#{πi,j}={r2if r is evenr12if r is odd,𝐧=#{νkl}={s2if s is evens12if s is odd.{\bf p}=\#\{\pi_{i,j}\}=\begin{cases}\frac{r}{2}&\quad\text{if\ $r$\ is even}\\ \frac{r-1}{2}&\quad\text{if\ $r$\ is odd}\end{cases},\quad{\bf n}=\#\{\nu_{kl}\}=\begin{cases}\frac{s}{2}&\quad\text{if\ $s$\ is even}\\ \frac{s-1}{2}&\quad\text{if\ $s$\ is odd}\end{cases}.

Now we form a set of involutions of type T1T_{1}, which from now on will be denoted always by pip_{i}. For any positive integers p¯{1,,𝐩}\bar{p}\in\{1,\ldots,{\bf p}\} and n¯{1,,𝐧}\bar{n}\in\{1,\ldots,{\bf n}\} we make a product of pairs:

(4.3) πiα,jαπiβ,jβ,πiα,jανkγ,lγ,νkγ,lγνkδ,lδ,α,β{1,,p¯},γ,δ{1,,n¯}.\pi_{i_{\alpha},j_{\alpha}}\pi_{i_{\beta},j_{\beta}},\quad\pi_{i_{\alpha},j_{\alpha}}\nu_{k_{\gamma},l_{\gamma}},\quad\nu_{k_{\gamma},l_{\gamma}}\nu_{k_{\delta},l_{\delta}},\quad\alpha,\beta\in\{1,\ldots,{\bar{p}}\},\ \gamma,\delta\in\{1,\ldots,{\bar{n}}\}.

We denote by 𝒮p¯,n¯\mathcal{S}^{\bar{p},\bar{n}} the group generated by involutions (4.3).

Proposition 4.5.

In the notation above the groups 𝒮p¯,n¯\mathcal{S}^{\bar{p},\bar{n}} have the following properties.

  • (i)

    𝒮p¯,n¯\mathcal{S}^{\bar{p},\bar{n}} is a subgroup of G(Br,s)G(B_{r,s}) for any p¯{1,,𝐩}\bar{p}\in\{1,\ldots,{\bf p}\} and n¯{1,,𝐧}\bar{n}\in\{1,\ldots,{\bf n}\};

  • (ii)

    𝒮p¯k,n¯\mathcal{S}^{\bar{p}-k,\bar{n}} is a subgroup of 𝒮p¯,n¯\mathcal{S}^{\bar{p},\bar{n}} for any k=0,1,,p¯k=0,1,\ldots,\bar{p};

  • (iii)

    𝒮p¯,n¯k\mathcal{S}^{\bar{p},\bar{n}-k} is a subgroup of 𝒮p¯,n¯\mathcal{S}^{\bar{p},\bar{n}} for any k=0,1,,n¯k=0,1,\ldots,\bar{n};

  • (iv)

    𝒮p¯k1,n¯k2\mathcal{S}^{\bar{p}-k_{1},\bar{n}-k_{2}} is a subgroup of 𝒮p¯,n¯\mathcal{S}^{\bar{p},\bar{n}} for any k1=0,1,,p¯k_{1}=0,1,\ldots,\bar{p} and k2=0,1,,n¯k_{2}=0,1,\ldots,\bar{n};

  • (v)

    The standard groups 𝒮p¯,n¯\mathcal{S}^{\bar{p},\bar{n}} are equivalent for fixed (p¯,n¯)(\bar{p},\bar{n}) in the sense of Definition 4.3;

  • (vi)

    Any set PIr,sPI_{r,s} satisfying Definition 3.7 and such that 𝒮𝐩,𝐧=𝒮(PIr,s)\mathcal{S}^{\bf p,\bf n}=\mathcal{S}(PI_{r,s}) will be equivalent in the sense of Definition 4.4;

  • (vii)

    Pairs πi,j\pi_{i,j} and νk,l\nu_{k,l} commute with all elements in 𝒮𝐩,𝐧\mathcal{S}^{\bf p,\bf n};

  • (viii)

    Let θ=zi1zi𝐩+𝐧\theta=z_{i_{1}}\cdots z_{i_{\bf p+\bf n}} be a product such that each zitz_{i_{t}}, t=1,,𝐩+𝐧t=1,\ldots,\bf p+\bf n belongs only to one pair from (4.1) or (4.2). Then θ\theta commutes with all elements in 𝒮𝐩,𝐧\mathcal{S}^{\bf p,\bf n}.

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 r,s\mathbb{R}^{r,s}. Properties (vii) and (viii) are the consequence of the facts that pairs πi,j\pi_{i,j}, νk,l\nu_{k,l}, and the product θ\theta will have even number of common elements and that the number of vectors ziz_{i} in any element of the group 𝒮𝐩,𝐧G(Br,s)\mathcal{S}^{\bf p,\bf n}\subset G(B_{r,s}) is also even. ∎

Example 4.3.

Consider 6,3\mathbb{R}^{6,3} with the basis B6,3={z1,,z9}B_{6,3}=\{z_{1},\ldots,z_{9}\}. The first six elements of the basis are positive and the last three are negative. We can choose the pairs

(4.4) π1,2=z1z2,π3,4=z3z4,π5,6=z5z6,ν78=z7z8,\pi_{1,2}=z_{1}z_{2},\quad\pi_{3,4}=z_{3}z_{4},\quad\pi_{5,6}=z_{5}z_{6},\quad\nu_{78}=z_{7}z_{8},

up to the sign permutation. They generate a group 𝒮𝟑,𝟏G(B6,3)\mathcal{S}^{{\bf 3},{\bf 1}}\subset G(B_{6,3}) of cardinality #𝒮𝟑,𝟏=8\#\mathcal{S}^{{\bf 3},{\bf 1}}=8. A possible choice of (T1)(T1)-type set of involutions PIPI generating 𝒮𝟑,𝟏\mathcal{S}^{{\bf 3},{\bf 1}} is

(4.5) PI6,3={p1=π1,2π3,4,p2=π1,2π5,6,p3=π1,2π7,8}.PI_{6,3}=\{p_{1}=\pi_{1,2}\pi_{3,4},\quad p_{2}=\pi_{1,2}\pi_{5,6},\quad p_{3}=\pi_{1,2}\pi_{7,8}\}.

Any pair from (4.4) will commute with involutions in (4.5) and therefore with all elements in the group 𝒮𝟑,𝟏G(B6,3)\mathcal{S}^{{\bf 3},{\bf 1}}\subset G(B_{6,3}). Furthermore, θ=z1z3z5z7\theta=z_{1}z_{3}z_{5}z_{7}, which is chosen up to a sign permutation, commutes with elements in the group 𝒮𝟑,𝟏G(B6,3)\mathcal{S}^{{\bf 3},{\bf 1}}\subset G(B_{6,3}) as well. The pairs

π1,2,π3,4,π5,6generates the subgroup 𝒮𝟑,𝟎𝒮𝟑,𝟏.\pi_{1,2},\quad\pi_{3,4},\quad\pi_{5,6}\quad\text{generates the subgroup $\mathcal{S}^{{\bf 3},{\bf 0}}\subset\mathcal{S}^{{\bf 3},{\bf 1}}$}.

Likewise the pairs

π1,2,π3,4,π7,8generates the subgroup 𝒮𝟐,𝟏𝒮𝟑,𝟏.\pi_{1,2},\quad\pi_{3,4},\quad\pi_{7,8}\quad\text{generates the subgroup $\mathcal{S}^{{\bf 2},{\bf 1}}\subset\mathcal{S}^{{\bf 3},{\bf 1}}$}.

Each of the subgroups 𝒮𝟑,𝟎\mathcal{S}^{{\bf 3},{\bf 0}} and 𝒮𝟐,𝟏\mathcal{S}^{{\bf 2},{\bf 1}} is a representative in its class of equivalence. Nevertheless, the groups 𝒮𝟑,𝟎\mathcal{S}^{{\bf 3},{\bf 0}} and 𝒮𝟐,𝟏\mathcal{S}^{{\bf 2},{\bf 1}} are not equivalent.

4.2. Connectivity of groups 𝒮\mathcal{S}

Here we introduce another tool of detecting non-equivalent subgroups 𝒮G(Br,s)\mathcal{S}\subset G(B_{r,s}), that we call “connectedness” for 𝒮=𝒮(PIr,s)\mathcal{S}=\mathcal{S}(PI_{r,s}).

Definition 4.6.

A group 𝒮𝕊r,s\mathcal{S}\in\mathbb{S}_{r,s} is called connected if there is no two subgroups 𝒮(1),𝒮(2)𝒮\mathcal{S}_{(1)},\mathcal{S}_{(2)}\subset\mathcal{S}, such that 𝒮\mathcal{S} is isomorphic to 𝒮(1)×𝒮(2)\mathcal{S}_{(1)}\times\mathcal{S}_{(2)} with 𝔟(𝒮(1))𝔟(𝒮(2))=𝟏\mathfrak{b}(\mathcal{S}_{(1)})\cap\mathfrak{b}(\mathcal{S}_{(2)})=\mathbf{1}. We write in this case π0(𝒮)=1\pi_{0}(\mathcal{S})=1.

If a group 𝒮𝕊r,s\mathcal{S}\in\mathbb{S}_{r,s} admits the decomposition into subgroups 𝒮=𝒮(1)××𝒮(k)\mathcal{S}=\mathcal{S}_{(1)}\times\ldots\times\mathcal{S}_{(k)} with π0(𝒮(i))=1\pi_{0}(\mathcal{S}_{(i)})=1 and 𝔟(𝒮(i))𝔟(𝒮(j))=𝟏\mathfrak{b}(\mathcal{S}_{(i)})\cap\mathfrak{b}(\mathcal{S}_{(j)})=\mathbf{1} for any iji\not=j, then we say that 𝒮\mathcal{S} has kk connected components and we write π0(𝒮)=k\pi_{0}(\mathcal{S})=k.

Lemma 4.7.

Let PI={pi}i=1(r,s)𝕀r,sMPI=\{p_{i}\}_{i=1}^{\ell(r,s)}\in\mathbb{PI}^{M}_{r,s}, and |𝔟(PI)|=r+s|\mathfrak{b}(PI)|=r+s. Assume that there is zαG(Br,s)z_{\alpha}\in G(B_{r,s}) such that zαi=1(r,s)𝔟(pi)z_{\alpha}\in\bigcap_{i=1}^{\ell(r,s)}\mathfrak{b}(p_{i}), and moreover, there is no σ𝒮(PI)\sigma\in\mathcal{S}(PI) such that 𝔟(σ)𝔟(pi)\mathfrak{b}(\sigma)\subset\mathfrak{b}(p_{i}) for any piPIp_{i}\in PI. Then π0(𝒮(PI))=1\pi_{0}(\mathcal{S}(PI))=1.

Proof.

Note that any product j2k+1pj\prod_{j}^{2k+1}p_{j} of odd number contains zαz_{\alpha}. Let us assume that 𝒮=𝒮(1)×𝒮(2)\mathcal{S}=\mathcal{S}_{(1)}\times\mathcal{S}_{(2)} is a non-trivial decomposition.

If both subgroups include a product of odd number of involutions j2l+1pj\prod_{j}^{2l+1}p_{j}, pjPIp_{j}\in PI, then zα𝔟(𝒮(1))𝔟(𝒮(2))z_{\alpha}\in\mathfrak{b}(\mathcal{S}_{(1)})\bigcap\mathfrak{b}(\mathcal{S}_{(2)}). Therefore 𝒮\mathcal{S} should be connected.

Assume the subgroup 𝒮(1)\mathcal{S}_{(1)} consists of only even products η=j2kpj\eta=\prod_{j}^{2k}p_{j} of involutions in PIPI. We write one of these products in the form η=pi0σ𝒮(1)\eta=p_{i_{0}}\cdot\sigma\in\mathcal{S}_{(1)}, where pi0p_{i_{0}} is one of the generators from the set PIPI and σ\sigma is a product of odd number of some involutions in PIPI. It implies that σ𝒮(2)\sigma\in\mathcal{S}_{(2)}. By the assumption 𝔟(σ)𝔟(pi)\mathfrak{b}(\sigma)\not\subset\mathfrak{b}(p_{i}) for any piPIp_{i}\in PI, there exists a basis vector zβ𝔟(σ)z_{\beta}\in\mathfrak{b}(\sigma) such that zβ𝔟(pi0)z_{\beta}\notin\mathfrak{b}(p_{i_{0}}). This implies that zβ𝔟(pi0σ)z_{\beta}\in\mathfrak{b}(p_{i_{0}}\cdot\sigma) and therefore zβ𝔟(σ)𝔟(pi0σ)𝔟(𝒮(2))𝔟(𝒮(1))z_{\beta}\in\mathfrak{b}(\sigma)\cap\mathfrak{b}(p_{i_{0}}\cdot\sigma)\subset\mathfrak{b}(\mathcal{S}_{(2)})\bigcap\mathfrak{b}(\mathcal{S}_{(1)}). This shows that the group 𝒮\mathcal{S} is connected. ∎

Example 4.4.

The standard subgroups 𝒮𝐩,𝟎𝕊r,0\mathcal{S}^{\bf p,0}\in\mathbb{S}_{r,0} constructed in Example 4.2 are connected for any r0r\geq 0.

Proposition 4.8.

Let PI1,PI2𝕀r,sMPI_{1},PI_{2}\in\mathbb{PI}_{r,s}^{M} be two generating sets. If PI1PI2PI_{1}~{}\sim PI_{2}, then π0(PI1)=π0(PI2)\pi_{0}(PI_{1})=\pi_{0}(PI_{2}).

Proof.

We write PI1={pk}k=1(r,s)PI_{1}=\{p_{k}\}_{k=1}^{\ell(r,s)}, PI2={qm}m=1(r,s)PI_{2}=\{q_{m}\}_{m=1}^{\ell(r,s)} and |𝔟(PIk)|=t|\mathfrak{b}(PI_{k})|=t. By the assumption there exists a re-ordering map CC of the basis Br,sB_{r,s} such that C(𝒮({pk}k=1(r,s))^)=𝒮({qm}m=1(r,s))^C\Big{(}\mathaccent 1371{\mathcal{S}\big{(}\{p_{k}\}_{k=1}^{\ell(r,s)}\big{)}}\Big{)}=\mathaccent 1371{\mathcal{S}\big{(}\{q_{m}\}_{m=1}^{\ell(r,s)}\big{)}}. If

𝒮(PI1)=𝒮(1)×𝒮(2)=𝒮(1)(PI11)×𝒮(2)(PI12),\mathcal{S}(PI_{1})=\mathcal{S}_{(1)}\times\mathcal{S}_{(2)}=\mathcal{S}_{(1)}(PI_{11})\times\mathcal{S}_{(2)}(PI_{12}),\quad

with

PI11={pik}k=1a,|𝔟({pik}k=1a)|=β,PI_{11}=\{p_{i_{k}}\}_{k=1}^{a},\quad|\mathfrak{b}(\{p_{i_{k}}\}_{k=1}^{a})|=\beta,

and

PI12={pjk}k=a+1(r,s),|𝔟({pjk}k=a+1(r,s))|=tβ,PI_{12}=\{p_{j_{k}}\}_{k=a+1}^{\ell(r,s)},\quad|\mathfrak{b}(\{p_{j_{k}}\}_{k=a+1}^{\ell(r,s)})|=t-\beta,

then 𝔟({pik}k=1a)𝔟({pjk}k=a+1(r,s))=\mathfrak{b}(\{p_{i_{k}}\}_{k=1}^{a})\cap\mathfrak{b}(\{p_{j_{k}}\}_{k=a+1}^{\ell(r,s)})=\emptyset. The re-ordering map CC will map the non intersecting sets 𝔟({pik}k=1a)\mathfrak{b}(\{p_{i_{k}}\}_{k=1}^{a}) and 𝔟({pjk}k=a+1(r,s))\mathfrak{b}(\{p_{j_{k}}\}_{k=a+1}^{\ell(r,s)}) onto non intersecting sets 𝒵1={zi1,,ziβ}\mathcal{Z}_{1}=\{z_{i_{1}},\ldots,z_{i_{\beta}}\} and 𝒵2={zjβ+1,,zit}\mathcal{Z}_{2}=\{z_{j_{\beta+1}},\ldots,z_{i_{t}}\}. The set 𝒵1\mathcal{Z}_{1} (with possible change of signs) will form the set PI21={qik}k=1aPI_{21}=\{q_{i_{k}}\}_{k=1}^{a} and the set 𝒵2\mathcal{Z}_{2} (again with possible change of signs) will form the set PI22={qjk}k=a+1tPI_{22}=\{q_{j_{k}}\}_{k=a+1}^{t}. Thus we obtain 𝒮(PI2)=𝒮(PI21)×𝒮(PI22)\mathcal{S}(PI_{2})=\mathcal{S}(PI_{21})\times\mathcal{S}(PI_{22}). ∎

We describe how the 2\mathbb{Z}_{2}-graded product of Clifford algebras can lead to the construction of disconnected subgroups 𝒮G(Br,s)\mathcal{S}\subset G(B_{r,s}). Consider the following decompositions of an orthonormal basis Br,s={z1,,zr,zr+1,,zr+s}B_{r,s}=\{z_{1},\ldots,z_{r},z_{r+1},\ldots,z_{r+s}\}:

z1,,zr1positive,zr+1,,zr+s1negative,andzr1+1,,zrpositive,zr+s1+1,,zr+snegative.\underbrace{z_{1},\ldots,z_{r_{1}}}_{\text{positive}},\underbrace{z_{r+1},\ldots,z_{r+s_{1}}}_{\text{negative}},\quad\text{and}\quad\underbrace{z_{r_{1}+1},\ldots,z_{r}}_{\text{positive}},\underbrace{z_{r+s_{1}+1},\ldots,z_{r+s}}_{\text{negative}}.

We put r2=rr1r_{2}=r-r_{1} and s2=ss1s_{2}=s-s_{1} and consider the decomposition r,sr1,s1r2,s2\mathbb{R}^{r,s}\cong\mathbb{R}^{r_{1},s_{1}}\oplus\mathbb{R}^{r_{2},s_{2}}, where we assume r1+s1rr1+ss1=r2+s2r_{1}+s_{1}\geq r-r_{1}+s-s_{1}=r_{2}+s_{2}. This decomposition leads to the isomorphism Clr1,s1^Clr2,s2Clr1+r2,s1+s2=Clr,s\text{Cl}_{r_{1},s_{1}}\mathaccent 866{\otimes}\text{Cl}_{r_{2},s_{2}}\cong\text{Cl}_{r_{1}+r_{2},s_{1}+s_{2}}=\text{Cl}_{r,s}, where ^\mathaccent 866{\otimes} denotes the 2\mathbb{Z}_{2}-graded tensor product of Clifford algebras, see [LM89, Proposition 1.5]. For each of the Clifford algebras Clrk,sk\text{Cl}_{r_{k},s_{k}}, k=1,2k=1,2, we consider the minimal admissible modules Vrk,skV^{r_{k},s_{k}} and the corresponding sets PIrk,skPI_{r_{k},s_{k}}. For r=r1+r2r=r_{1}+r_{2} and s=s1+s2s=s_{1}+s_{2}, we have (r1,s1)(r,s)\ell(r_{1},s_{1})\leq\ell(r,s). Let PIr1,s1𝕀r1,s1MPI_{r_{1},s_{1}}\in\mathbb{PI}_{r_{1},s_{1}}^{M} and PIr2,s2𝕀r2,s2MPI_{r_{2},s_{2}}\in\mathbb{PI}_{r_{2},s_{2}}^{M} satisfy

|𝔟+(PIr1,s1)|=r1,|𝔟(PIr1,s1)|=s1,|\mathfrak{b}^{+}(PI_{r_{1},s_{1}})|=r_{1},\quad|\mathfrak{b}^{-}(PI_{r_{1},s_{1}})|=s_{1},
|𝔟+(PIr2,s2)|=r2,|𝔟(PIr2,s2)|=s2,|\mathfrak{b}^{+}(PI_{r_{2},s_{2}})|=r_{2},\quad|\mathfrak{b}^{-}(PI_{r_{2},s_{2}})|=s_{2},

and PIr1,s1PIr2,s2=PI_{r_{1},s_{1}}\bigcap PI_{r_{2},s_{2}}=\emptyset. We assume also that each set contains at most one type T2T_{2} involution qkPIrk,skq_{k}\in PI_{r_{k},s_{k}}, k=1,2k=1,2. Then by non-commutativity of q1q_{1} and q2q_{2} it is easy to see the following properties:

  • If one of the sets PIr1,s1PI_{r_{1},s_{1}} or PIr2,s2PI_{r_{2},s_{2}} is (T1)(T1)-type set, then

    PIr1,s1PIr2,s2𝕀r,s.PI_{r_{1},s_{1}}\bigcup PI_{r_{2},s_{2}}\in\mathbb{PI}_{r,s}.

    This implies

    (4.6) (r1,s1)+(r2,s2)(r,s).\ell(r_{1},s_{1})+\ell(r_{2},s_{2})\leq\ell(r,s).
  • If both PIr1,s1PI_{r_{1},s_{1}} and PIr2,s2PI_{r_{2},s_{2}} are (T2)(T2)-type sets, containing type T2T_{2} involutions q1PIr1,s1q_{1}\in PI_{r_{1},s_{1}} and q2PIr2,s2q_{2}\in PI_{r_{2},s_{2}}, then

    (PIr1,s1\{q1})PIr2,s2𝕀r,sandPIr1,s1(PIr2,s2\{q2})𝕀r,s.\big{(}PI_{r_{1},s_{1}}\backslash\{q_{1}\}\big{)}\bigcup PI_{r_{2},s_{2}}\in\mathbb{PI}_{r,s}\quad\text{and}\quad PI_{r_{1},s_{1}}\bigcup\big{(}PI_{r_{2},s_{2}}\backslash\{q_{2}\}\big{)}\in\mathbb{PI}_{r,s}.

    This implies

    (4.7) (r1,s1)+(r2,s2)1(r,s).\ell(r_{1},s_{1})+\ell(r_{2},s_{2})-1\leq\ell(r,s).

One can state similar properties for any number of components in a decomposition PI=kPIrk,skPI=\cup_{k}PI_{r_{k},s_{k}}.

Remark 4.1.

If the equalities in (4.6) or (4.7) hold, then non-connected subgroups 𝒮(PIr1,s1)\mathcal{S}(PI_{r_{1},s_{1}}) and 𝒮(PIr2,s2)\mathcal{S}(PI_{r_{2},s_{2}}) can be constructed from lower dimensions and

𝒮(PIr,s)=𝒮(PIr1,s1)×𝒮(PIr2,s2).\mathcal{S}(PI_{r,s})=\mathcal{S}(PI_{r_{1},s_{1}})\times\mathcal{S}(PI_{r_{2},s_{2}}).

Particularly, if r9r\leq 9 and s{0,1}s\in\{0,1\}, then all the groups are connected. It follows by showing that the inequalities (4.6) and (4.7) are always strict.

Proposition 4.9.

The number (r,s)\ell(r,s) has three periodicities:

(r+8,s)\displaystyle\ell(r+8,s) =\displaystyle= (r+4,s+4)=(r,s+8)=(r,s)+4\displaystyle\ell(r+4,s+4)=\ell(r,s+8)=\ell(r,s)+4
=\displaystyle= (r,s)+(8,0)=(r,s)+(0,8)=(r,s)+(4,4).\displaystyle\ell(r,s)+\ell(8,0)=\ell(r,s)+\ell(0,8)=\ell(r,s)+\ell(4,4).
Proof.

The number (r,s)\ell(r,s) is determined by 2(r,s)dim(Vr,s)=2r+s2^{\ell(r,s)}\cdot\dim(V^{r,s})=2^{r+s}, see Notation 3.3. Hence,

2(r+8,s)dim(Vr+8,s)=2r+8+s=2r+s28=2(r,s)dim(Vr,s)28.2^{\ell(r+8,s)}\cdot\dim(V^{r+8,s})=2^{r+8+s}=2^{r+s}2^{8}=2^{\ell(r,s)}\cdot\dim(V^{r,s})\cdot 2^{8}.

We know that dim(Vr+8,s)=24dim(Vr,s)\dim(V^{r+8,s})=2^{4}\dim(V^{r,s}) [FM17, Section 4.1]. Hence it holds (r+8,s)=(r,s)+4\ell(r+8,s)=\ell(r,s)+4.

Other equalities hold by the same reason. ∎

5. Construction of subgroups in 𝕊r,sM\mathbb{S}_{r,s}^{M}, r{3,,16}r\in\{3,\ldots,16\}, s{0,1}s\in\{0,1\}

5.1. General method of the construction

In this section we apply the previous theory for the classification of groups 𝒮G(Br,s)\mathcal{S}\subset G(B_{r,s}) and perform the exact construction of non-equivalent subgroups. We restrict ourself to r{1,,16}r\in\{1,\ldots,16\} and s=0,1s=0,1 because we want to illustrate the main features that appears in classification without diving into technical details. The classification for arbitrary 𝒮G(Br,s)\mathcal{S}\subset G(B_{r,s}) is postponed for the forthcoming paper.

We start from s=0s=0 and the classification for s=1s=1 will be the strait forward generalisation. We classify groups 𝒮𝕊r,0M\mathcal{S}\subset\mathbb{S}^{M}_{r,0} according to parameters: π0(𝒮)\pi_{0}(\mathcal{S}), |𝔟(PIr,0)||\mathfrak{b}(PI_{r,0})|, and the type (T1)(T1) or (T2)(T2) of the set PIPI generating the group 𝒮𝕊r,sM\mathcal{S}\in\mathbb{S}^{M}_{r,s}. 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 (TI)(TI)-type sets, I=1,2I=1,2:

(5.1) {(i) We use the signature (TI,π) if an additional involution  is related to product π1,2;(ii) We use the signature (TI,θ) if an additional involution  is related to product θ;(iii) We use the signature (TI,π,θ) if there are two additional  involutions, which are related to both products π1,2 and θ;(iv) Finally we just write (TI) if there is no involutions, except of standards;\left\{\begin{array}[]{l}\text{(i) We use the signature $(TI,\pi)$ if an additional involution }\\ \text{\qquad is related to product $\pi_{1,2}$;}\\ \text{(ii) We use the signature $(TI,\theta)$ if an additional involution }\\ \text{\qquad is related to product $\theta$;}\\ \text{(iii) We use the signature $(TI,\pi,\theta)$ if there are two additional }\\ \text{\qquad involutions, which are related to both products $\pi_{1,2}$ and $\theta$;}\\ \text{(iv) Finally we just write $(TI)$ if there is no involutions,}\\ \text{\qquad except of standards;}\end{array}\right.

We formulate the results in 15 theorems following the dimension rr and illustrate each case by a table. We list the set of generators PIPI 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 𝒮𝕊r,0M\mathcal{S}\in\mathbb{S}_{r,0}^{M} for a fixed r>0r>0.

We divide the construction into three steps.

Step 1. We start from a group satisfying π0(𝒮)=1\pi_{0}(\mathcal{S})=1 and |𝔟(PIr,0)|=r|\mathfrak{b}(PI_{r,0})|=r. First we find standard subgroup 𝒮𝐩,𝟎𝒮\mathcal{S}^{\bf p,0}\subset\mathcal{S} and complement it (if necessary) by involutions to reach the maximal number (r,0)\ell(r,0) of involutions in PIr,0PI_{r,0} generating 𝒮𝕊r,0M\mathcal{S}\in\mathbb{S}_{r,0}^{M}. The additional involutions will be formed by checking whether the product of π1,2\pi_{1,2} and/or θ\theta by zrz_{r} are involutions commuting with 𝒮𝐩,𝟎\mathcal{S}^{\bf p,0}. Then we consider a smaller standard subgroup 𝒮𝐩𝟏,𝟎𝒮𝐩,𝟎\mathcal{S}^{\bf p-1,0}\subset\mathcal{S}^{\bf p,0} and complement it by a careful choice of involutions to reach the maximal number (r,0)\ell(r,0) for 𝒮(PIr,0)\mathcal{S}(PI_{r,0}), checking whether the connectivity π0(𝒮)=1\pi_{0}(\mathcal{S})=1 is not violated. We can repeat the last step several times if the condition π0(𝒮)=1\pi_{0}(\mathcal{S})=1 still holds.

Step 2. We continue to look on π0(𝒮)=1\pi_{0}(\mathcal{S})=1 and |𝔟(PIr,0)|=r1|\mathfrak{b}(PI_{r,0})|=r-1. In most cases it will be a simple step back from (r,0)(r,0) to (r1,0)(r-1,0) as, for example, for reduction from r=4r=4 to r=3r=3.

Step 3. Next we check π0(𝒮)=2\pi_{0}(\mathcal{S})=2 and 𝒮=𝒮(1)×𝒮(2)\mathcal{S}=\mathcal{S}_{(1)}\times\mathcal{S}_{(2)}. 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 𝒮=𝒮(PIr,0)𝕊r,0M\mathcal{S}=\mathcal{S}(PI_{r,0})\in\mathbb{S}^{M}_{r,0}, with |𝔟(PIr,0)|=r|\mathfrak{b}(PI_{r,0})|=r and π0(𝒮)=1\pi_{0}(\mathcal{S})=1. Then, the maximal standard subgroups, included in a given group 𝒮𝕊r,0M\mathcal{S}\in\mathbb{S}_{r,0}^{M}, are equivalent modulo reordering by induction arguments with respect to the dimension (r,0)(r,0), see also Proposition 4.5, item (v).

Moreover, once we fix a standard group with its generators of form (4.3), the maximally complemented sets PIPI obtained by adding involutions as in Step 1, will be equivalent in the sense of Definition 4.4 if they have the same signature described in (5.1) and π0(𝒮(PI))=1\pi_{0}(\mathcal{S}(PI))=1.

Lemma 5.2.

If r=3+8k,5+8k,6+8k,7+8kr=3+8k,5+8k,6+8k,7+8k for k0k\geq 0, then sets PIr,0𝕀r,0MPI_{r,0}\in\mathbb{PI}^{M}_{r,0} satisfying π0(𝒮(PIr,0))=1\pi_{0}(\mathcal{S}(PI_{r,0}))=1 and |𝔟(PIr,0)|=r|\mathfrak{b}(PI_{r,0})|=r are always of (T2)(T2)-type.

Proof.

We start from r=3+8kr=3+8k. For the case r=3r=3 there is only one type T2T_{2} involution. Let k1k\geq 1 and assume, by contrary, that there is a (T1)(T1)-type set PIr,0𝕀r,0MPI_{r,0}\in\mathbb{PI}_{r,0}^{M}. We have (r,0)=(3+8k,0)=1+4k\ell(r,0)=\ell(3+8k,0)=1+4k. The standard subgroup 𝒮𝐩,𝟎𝒮(PIr,0)\mathcal{S}^{\bf p,0}\subset\mathcal{S}(PI_{r,0}), 𝐩=1+4k{\bf p}=1+4k, does not contain zrz_{r}, since rr is odd. Let p1,,p4kp_{1},\ldots,p_{4k} will be involutions generating 𝒮𝐩,𝟎\mathcal{S}^{\bf p,0}, then zr𝔟(p1+4k)z_{r}\in\mathfrak{b}(p_{1+4k}). It implies

{p1,,p4k,zrp1+4k}𝕀r1,0M.\{p_{1},\ldots,p_{4k},\ z_{r}\cdot p_{1+4k}\}\in\mathbb{PI}_{r-1,0}^{M}.

This contradicts to (r1,0)=(2+8k,0)=(3+8k,0)1=(r,0)1\ell(r-1,0)=\ell(2+8k,0)=\ell(3+8k,0)-1=\ell(r,0)-1.

The arguments for the cases r=5+8kr=5+8k, and r=7+8kr=7+8k are similar to the case r=3+8kr=3+8k.

Let r=6+8kr=6+8k. We assume that there is a (T1)(T1)-type set PIr,0𝕀r,0MPI_{r,0}\in\mathbb{PI}_{r,0}^{M}. We have (r,0)=(6+8k,0)=3+4k\ell(r,0)=\ell(6+8k,0)=3+4k. The standard subgroup 𝒮𝐩,𝟎𝒮(PIr,0)\mathcal{S}^{\bf p,0}\subset\mathcal{S}(PI_{r,0}), 𝐩=3+4k{\bf p}=3+4k, contains zrz_{r}. Let p1,,p2+4kp_{1},\ldots,p_{2+4k} be involutions generating 𝒮𝐩,𝟎\mathcal{S}^{\bf p,0}, where we can assume that zr𝔟(p2+4k)z_{r}\in\mathfrak{b}(p_{2+4k}) and p3+4kPI6+8k,0p_{3+4k}\in PI_{6+8k,0} is the last type T1T_{1} involution.

  • (1)

    If zr𝔟(p3+4k)z_{r}\notin\mathfrak{b}(p_{3+4k}), then

    {p1,,p1+4k,zrp2+4k,p3+4k}𝕀r1,0M.\{p_{1},\ldots,p_{1+4k},\ z_{r}\cdot p_{2+4k},\ p_{3+4k}\}\in\mathbb{PI}_{r-1,0}^{M}.

    This contradicts to (r1,0)=(5+8k,0)=(6+8k,0)1=(r,0)1\ell(r-1,0)=\ell(5+8k,0)=\ell(6+8k,0)-1=\ell(r,0)-1.

  • (2)

    If zr𝔟(p3+4k)z_{r}\in\mathfrak{b}(p_{3+4k}), then we replace p3+4kPIr,0p_{3+4k}\in PI_{r,0} by another type T1T_{1} involution p~3+4k=p2+4kp3+4kPIr,0~\tilde{p}_{3+4k}=p_{2+4k}p_{3+4k}\in\mathaccent 869{PI_{r,0}}. In this case zr𝔟(p^3+4k)z_{r}\notin\mathfrak{b}(\hat{p}_{3+4k}) and the situation is reduced to the previous step (1). Note that the group 𝒮(PIr,0)\mathcal{S}(PI_{r,0}) is equivalent 𝒮(PIr,0~)\mathcal{S}(\mathaccent 869{PI_{r,0}}).

We also note that for r=3+8kr=3+8k and r=7+8kr=7+8k the volume forms Ωr=i=1rzi\Omega_{r}=\prod_{i=1}^{r}z_{i} which are type T2T_{2} involutions can be included to PIr,0PI_{r,0}. It justifies the (T2)(T2)-type set of PIPIs in cases r=3+8kr=3+8k and r=7+8kr=7+8k. ∎

5.2. Constructions of groups 𝒮𝕊r,0M\mathcal{S}\in\mathbb{S}^{M}_{r,0} for r{3,,16}r\in\{3,\ldots,16\}

Theorem 5.3.

There is a unique group 𝒮𝕊3,0M\mathcal{S}\subset\mathbb{S}^{M}_{3,0}. It is generated by type T2T_{2} involution p=z1z2z3p=z_{1}z_{2}z_{3}. Thus we have

π0(𝒮)=1,|𝔟(PI3,0)|=3,𝒮={𝟏,p=z1z2z3=π1,2z3}.\pi_{0}(\mathcal{S})=1,\quad|\mathfrak{b}(PI_{3,0})|=3,\quad\mathcal{S}=\{{\bf 1},p=z_{1}z_{2}z_{3}=\pi_{1,2}z_{3}\}.
Proof.

The group 𝒮\mathcal{S} is unique up to reordering. ∎

Theorem 5.4.

There are two non-equivalent groups in 𝕊4,0M\mathbb{S}^{M}_{4,0}.

Table 2. Groups for r=4r=4
(4,0)\ell(4,0) π0(𝒮)\pi_{0}(\mathcal{S}) |𝔟(PI)||\mathfrak{b}(PI)| Signature PIPI
𝒮(1)\mathcal{S}^{(1)} 11 11 44 (T1)(T1) p=π1,2π3,4p=\pi_{1,2}\pi_{3,4}
𝒮(2)\mathcal{S}^{(2)} 11 11 33 (T2,π)(T2,\pi) q=π1,2z3q=\pi_{1,2}z_{3}
Proof.

The proof is obvious. ∎

Notation 5.1.

From now on we write θi,j¯\theta_{\overline{i,j}} to indicate that product in θ\theta starts from ziz_{i} and ends with zjz_{j} containing all zkz_{k} for odd kk between ii and jj. We have

|𝔟(θi,j¯)|=ji2+1.|\mathfrak{b}(\theta_{\overline{i,j}})|=\frac{j-i}{2}+1.
Theorem 5.5.

There is unique group in 𝕊5,0M\mathbb{S}^{M}_{5,0}.

Table 3. Groups for r=5r=5
(5,0)\ell(5,0) π0(𝒮)\pi_{0}(\mathcal{S}) |𝔟(PI)||\mathfrak{b}(PI)| Signature PIPI
𝒮\mathcal{S} 22 11 55 (T2,θ)(T2,\theta) p=π1,2π3,4q=θ1,3¯z5=z1z3z5\begin{array}[]{llllll}p=\pi_{1,2}\pi_{3,4}\\ q=\theta_{\overline{1,3}}z_{5}=z_{1}z_{3}z_{5}\end{array}
Proof.

We start from the standard subgroup 𝒮𝟐,𝟎={𝟏,p=π1,2π3,4}\mathcal{S}^{\bf 2,\bf 0}=\{\mathbf{1},p=\pi_{1,2}\pi_{3,4}\} of the maximal group 𝒮𝕊5,0M\mathcal{S}\subset\mathbb{S}_{5,0}^{M}. The products π1,2=z1z2\pi_{1,2}=z_{1}z_{2} and θ=z1z3\theta=z_{1}z_{3} commute with the involution pp. To complete the standard subgroup 𝒮𝟐,𝟎\mathcal{S}^{\bf 2,\bf 0} to the maximal group 𝒮𝕊5,0M\mathcal{S}\subset\mathbb{S}^{M}_{5,0} we add a type T2T_{2} involutions

eitherq1=π1,2z5=z1z2z5orq2=θz5=z1z3z5.\text{either}\quad q_{1}=\pi_{1,2}z_{5}=z_{1}z_{2}z_{5}\quad\text{or}\quad q_{2}=\theta z_{5}=z_{1}z_{3}z_{5}.

Both choices lead to the equivalent subgroups

{𝟏,p=π1,2π3,4,q1=z1z2z5}and {𝟏,p=π1,2π3,4,q2=z1z3z5}\{\mathbf{1},p=\pi_{1,2}\pi_{3,4},q_{1}=z_{1}z_{2}z_{5}\}\quad\text{and\quad}\{\mathbf{1},p=\pi_{1,2}\pi_{3,4},q_{2}=z_{1}z_{3}z_{5}\}

by permutation z2z3z_{2}\leftrightarrow z_{3}. ∎

Theorem 5.6.

There is unique group in 𝕊6,0M\mathbb{S}^{M}_{6,0}.

Table 4. Groups for r=6r=6
(6,0)\ell(6,0) π0(𝒮)\pi_{0}(\mathcal{S}) |𝔟(PI)||\mathfrak{b}(PI)| Signature PIPI
𝒮\mathcal{S} 33 11 66 (T2,θ)(T2,\theta) p1=π1,2π3,4p2=π1,2π5,6q=θ1,5¯\begin{array}[]{llllll}p_{1}=\pi_{1,2}\pi_{3,4}\\ p_{2}=\pi_{1,2}\pi_{5,6}\\ q=\theta_{\overline{1,5}}\end{array}
Proof.

The standard subgroup 𝒮𝟑,𝟎\mathcal{S}^{\bf 3,\bf 0} is generated by the involutions

(5.2) p1=π1,2π3,4,p2=π1,2π5,6.p_{1}=\pi_{1,2}\pi_{3,4},\quad p_{2}=\pi_{1,2}\pi_{5,6}.

We need to add one involution since (6,0)=3\ell(6,0)=3. We observe that π1,2zj\pi_{1,2}z_{j}, j=1,,6j=1,\ldots,6, does not commute with generators (5.2), but θ=θ1,5¯=z1z3z5\theta=\theta_{\overline{1,5}}=z_{1}z_{3}z_{5} is an involution itself commuting with generators (5.2). Thus we add θ1,5¯\theta_{\overline{1,5}} to make PIPI complete. It finishes the proof. ∎

Theorem 5.7.

There is unique group in 𝕊7,0M\mathbb{S}^{M}_{7,0}.

Table 5. Groups for r=7r=7
(7,0)\ell(7,0) π0(𝒮)\pi_{0}(\mathcal{S}) |𝔟(PI)||\mathfrak{b}(PI)| Signature PIPI
𝒮\mathcal{S} 44 11 77 (T2,π,θ)(T2,\pi,\theta) p1=π1,2π3,4p2=π1,2π5,6p3=θ1,5¯z7q=π1,2z7\begin{array}[]{llllll}p_{1}=\pi_{1,2}\pi_{3,4}\\ p_{2}=\pi_{1,2}\pi_{5,6}\\ p_{3}=\theta_{\overline{1,5}}z_{7}\\ q=\pi_{1,2}z_{7}\end{array}
Proof.

The standard subgroup 𝒮𝟑,𝟎𝒮\mathcal{S}^{\bf 3,\bf 0}\subset\mathcal{S} is generated by involutions (5.2). We need to add two involutions since (7,0)=4\ell(7,0)=4, at least one of which must contain z7z_{7}. We observe that the products π1,2z7\pi_{1,2}z_{7} and θ1,5¯z7=z1z3z5z7\theta_{\overline{1,5}}z_{7}=z_{1}z_{3}z_{5}z_{7} are both involutions commuting with generators (5.2) with each other. We append them both to reach (7,0)=4\ell(7,0)=4. The reductions to |𝔟(PI7,0)|=6|\mathfrak{b}(PI_{7,0})|=6 is not possible due to (6,0)<(7,0)\ell(6,0)<\ell(7,0). We finish the proof. ∎

Theorem 5.8.

There are two non-equivalent groups in 𝕊8,0M\mathbb{S}^{M}_{8,0}.

Table 6. Groups for r=8r=8
(8,0)\ell(8,0) π0(𝒮)\pi_{0}(\mathcal{S}) |𝔟(PI)||\mathfrak{b}(PI)| Sinature PIPI
𝒮(1)\mathcal{S}^{(1)} 44 11 88 (T1,θ)(T1,\theta) p1=π1,2π3,4p2=π1,2π5,6p3=π1,2π7,8p4=θ1,7¯\begin{array}[]{llllll}p_{1}=\pi_{1,2}\pi_{3,4}\\ p_{2}=\pi_{1,2}\pi_{5,6}\\ p_{3}=\pi_{1,2}\pi_{7,8}\\ p_{4}=\theta_{\overline{1,7}}\end{array}
𝒮(2)\mathcal{S}^{(2)} 44 11 77 (T2,π,θ)(T2,\pi,\theta) p1=π1,2π3,4p2=π1,2π5,6p3=θ1,5¯z7q=π1,2z7\begin{array}[]{llllll}p_{1}=\pi_{1,2}\pi_{3,4}\\ p_{2}=\pi_{1,2}\pi_{5,6}\\ p_{3}=\theta_{\overline{1,5}}z_{7}\\ q=\pi_{1,2}z_{7}\end{array}
Proof.

The standard subgroup 𝒮𝟒,𝟎𝒮(1)\mathcal{S}^{\bf 4,\bf 0}\subset\mathcal{S}^{(1)} is generated by involutions

(5.3) p1=π1,2π3,4,p2=π1,2π5,6,p3=π1,2π7,8.p_{1}=\pi_{1,2}\pi_{3,4},\quad p_{2}=\pi_{1,2}\pi_{5,6},\quad p_{3}=\pi_{1,2}\pi_{7,8}.

We need to add one involution since (8,0)=4\ell(8,0)=4. It is easy to see that only θ1,7¯=z1z3z5z7\theta_{\overline{1,7}}=z_{1}z_{3}z_{5}z_{7} commutes with generators (5.3).

Consider standard subgroup 𝒮𝟑,𝟎𝒮(2)\mathcal{S}^{\bf 3,\bf 0}\subset\mathcal{S}^{(2)} generated by (5.2). This case is reduced to r=7r=7 and it is indicated in Table 6. We finish the proof. ∎

Theorem 5.9.

There are three non-equivalent groups in 𝕊9,0M\mathbb{S}^{M}_{9,0}.

Table 7. Groups for r=9r=9
(9,0)\ell(9,0) π0(𝒮)\pi_{0}(\mathcal{S}) |𝔟(PI)||\mathfrak{b}(PI)| Signature PIPI
𝒮(1)\mathcal{S}^{(1)} 44 11 99 (T2,π)(T2,\pi) p1=π1,2π3,4p2=π1,2π5,6p3=π1,2π7,8p4=π1,2z9\begin{array}[]{llllll}p_{1}=\pi_{1,2}\pi_{3,4}\\ p_{2}=\pi_{1,2}\pi_{5,6}\\ p_{3}=\pi_{1,2}\pi_{7,8}\\ p_{4}=\pi_{1,2}z_{9}\end{array}
𝒮(2)\mathcal{S}^{(2)} 44 11 88 (T1,θ)(T1,\theta) p1=π1,2π3,4p2=π1,2π5,6p3=π1,2π7,8p4=θ1,7¯\begin{array}[]{llllll}p_{1}=\pi_{1,2}\pi_{3,4}\\ p_{2}=\pi_{1,2}\pi_{5,6}\\ p_{3}=\pi_{1,2}\pi_{7,8}\\ p_{4}=\theta_{\overline{1,7}}\end{array}
𝒮(3)\mathcal{S}^{(3)} 44 11 77 (T2,π,θ)(T2,\pi,\theta) p1=π1,2π3,4p2=π1,2π5,6p3=θ1,7¯q=π1,2z7\begin{array}[]{llllll}p_{1}=\pi_{1,2}\pi_{3,4}\\ p_{2}=\pi_{1,2}\pi_{5,6}\\ p_{3}=\theta_{\overline{1,7}}\\ q=\pi_{1,2}z_{7}\end{array}
Proof.

The standard subgroup 𝒮𝟒,𝟎𝒮(1)\mathcal{S}^{\bf 4,\bf 0}\subset\mathcal{S}^{(1)} is generated by involutions in (5.3). We need to add one involution containing z9z_{9} since (9,0)=4\ell(9,0)=4 and |𝔟(PI9,0)|=9|\mathfrak{b}(PI_{9,0})|=9. We add q=π1,2z9q=\pi_{1,2}z_{9}.

We release |𝔟(PI9,0)|=9|\mathfrak{b}(PI_{9,0})|=9 and consider |𝔟(PI9,0)|=8|\mathfrak{b}(PI_{9,0})|=8. It is easy to see that 𝒮(2)\mathcal{S}^{(2)} is isomorphic to 𝒮(1)𝕊8,0M\mathcal{S}^{(1)}\in\mathbb{S}^{M}_{8,0}.

Consider standard subgroup 𝒮𝟑,𝟎𝒮(3)\mathcal{S}^{\bf 3,\bf 0}\subset\mathcal{S}^{(3)} generated by (5.2). This case is reduced to r=7r=7 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 𝕊10,0M\mathbb{S}^{M}_{10,0}.

Table 8. Groups for r=10r=10
(10,0)\ell(10,0) π0(𝒮)\pi_{0}(\mathcal{S}) |𝔟(PI)||\mathfrak{b}(PI)| Signature PIPI
𝒮(1)\mathcal{S}^{(1)} 44 11 1010 (T1,π)(T1,\pi) p1=π1,2π3,4p2=π1,2π5,6p3=π1,2π7,8p4=π1,2π9,10\begin{array}[]{llllll}p_{1}=\pi_{1,2}\pi_{3,4}\\ p_{2}=\pi_{1,2}\pi_{5,6}\\ p_{3}=\pi_{1,2}\pi_{7,8}\\ p_{4}=\pi_{1,2}\pi_{9,10}\end{array}
𝒮(2)\mathcal{S}^{(2)} 44 11 99 (T2,π)(T2,\pi) p1=π1,2π3,4p2=π1,2π5,6p3=π1,2π7,8p4=π1,2z9\begin{array}[]{llllll}p_{1}=\pi_{1,2}\pi_{3,4}\\ p_{2}=\pi_{1,2}\pi_{5,6}\\ p_{3}=\pi_{1,2}\pi_{7,8}\\ p_{4}=\pi_{1,2}z_{9}\end{array}
𝒮(3)\mathcal{S}^{(3)} 44 11 88 (T1,θ)(T1,\theta) p1=π1,2π3,4p2=π1,2π5,6p3=π1,2π7,8p4=θ1,7¯\begin{array}[]{llllll}p_{1}=\pi_{1,2}\pi_{3,4}\\ p_{2}=\pi_{1,2}\pi_{5,6}\\ p_{3}=\pi_{1,2}\pi_{7,8}\\ p_{4}=\theta_{\overline{1,7}}\end{array}
𝒮(4)\mathcal{S}^{(4)} 44 11 77 (T2,π,θ)(T2,\pi,\theta) p1=π1,2π3,4p2=π1,2π5,6p3=θ1,5¯z7q=π1,2z7\begin{array}[]{llllll}p_{1}=\pi_{1,2}\pi_{3,4}\\ p_{2}=\pi_{1,2}\pi_{5,6}\\ p_{3}=\theta_{\overline{1,5}}z_{7}\\ q=\pi_{1,2}z_{7}\end{array}
𝒮(5)\mathcal{S}^{(5)} 44 22 77 (T1,θ)×(T2,π)r=7+3\begin{array}[]{llll}(T1,\theta)\times(T2,\pi)\\ \quad r=7+3\end{array} (p1)(1)=π1,2π3,4,(q)(2)=π8,9z10(p2)(1)=π1,2π5,6,(p3)(1)=θ1,7¯,\begin{array}[]{llllll}&(p_{1})_{(1)}=\pi_{1,2}\pi_{3,4},&(q)_{(2)}=\pi_{8,9}z_{10}\\ &(p_{2})_{(1)}=\pi_{1,2}\pi_{5,6},\\ &(p_{3})_{(1)}=\theta_{\overline{1,7}},\end{array}
𝒮(6)\mathcal{S}^{(6)} 44 22 77 (T2,θ)×(T1)r=6+4\begin{array}[]{llll}(T2,\theta)\times(T1)\\ \quad r=6+4\end{array} (p1)(1)=π1,2π3,4,(p1)(2)=π7,8π9,10(p2)(1)=π1,2π5,6,(q)(1)=θ1,5¯,\begin{array}[]{llllll}&(p_{1})_{(1)}=\pi_{1,2}\pi_{3,4},&(p_{1})_{(2)}=\pi_{7,8}\pi_{9,10}\\ &(p_{2})_{(1)}=\pi_{1,2}\pi_{5,6},\\ &(q)_{(1)}=\theta_{\overline{1,5}},\end{array}
Proof.

The standard subgroup 𝒮𝟓,𝟎𝒮(1)\mathcal{S}^{\bf 5,\bf 0}\subset\mathcal{S}^{(1)} is generated by involutions

(5.4) p1=π1,2π3,4,p2=π1,2π5,6,p3=π1,2π7,8,p4=π1,2π9,10.p_{1}=\pi_{1,2}\pi_{3,4},\quad p_{2}=\pi_{1,2}\pi_{5,6},\quad p_{3}=\pi_{1,2}\pi_{7,8},\quad p_{4}=\pi_{1,2}\pi_{9,10}.

We do not need to add any involution, since (10,0)=4\ell(10,0)=4.

The rest of the connected groups comes from lower dimensions.

To construct the disconnected subgroup 𝒮(5)=𝒮(1)(5)×𝒮(2)(5)\mathcal{S}^{(5)}=\mathcal{S}_{(1)}^{(5)}\times\mathcal{S}_{(2)}^{(5)} corresponding to the 2\mathbb{Z}_{2}-graded tensor product of the Clifford algebras Cl10,0Cl7,0^Cl3,0\text{Cl}_{10,0}\cong\text{Cl}_{7,0}\hat{\otimes}\text{Cl}_{3,0} we consider standard subgroup 𝒮(1)𝟑,𝟎𝒮(1)(5)\mathcal{S}_{(1)}^{\bf 3,0}\subset\mathcal{S}_{(1)}^{(5)} generated by (5.2) and add type T1T_{1} involution θ1,7¯\theta_{\overline{1,7}}. Then 𝒮(2)(5)={𝟏,π8,9z10}\mathcal{S}_{(2)}^{(5)}=\{{\bf 1},\pi_{8,9}z_{10}\}.

To obtain 𝒮(6)=𝒮(1)(6)×𝒮(2)(6)\mathcal{S}^{(6)}=\mathcal{S}_{(1)}^{(6)}\times\mathcal{S}_{(2)}^{(6)} corresponding to the 2\mathbb{Z}_{2}-graded tensor product of the Clifford algebras Cl10,0Cl6,0^Cl4,0\text{Cl}_{10,0}\cong\text{Cl}_{6,0}\hat{\otimes}\text{Cl}_{4,0} we take standard subgroup 𝒮(1)𝟑,𝟎𝒮(1)(6)\mathcal{S}_{(1)}^{\bf 3,0}\subset\mathcal{S}_{(1)}^{(6)} generated by (5.2) and add type T2T_{2} involution θ1,5¯\theta_{\overline{1,5}}. Then 𝒮(2)(6)={𝟏,π7,8π9,10}\mathcal{S}_{(2)}^{(6)}=\{{\bf 1},\pi_{7,8}\pi_{9,10}\}.

Theorem 5.11.

There are one connected and two disconnected non-equivalent subgroups in 𝕊11,0M\mathbb{S}^{M}_{11,0}.

Table 9. Groups for r=11r=11
(11,0)\ell(11,0) π0(𝒮)\pi_{0}(\mathcal{S}) |𝔟(PI)||\mathfrak{b}(PI)| Signature PIPI
𝒮(1)\mathcal{S}^{(1)} 55 11 1111 (T2,π)(T2,\pi) p1=π1,2π3,4p2=π1,2π5,6p3=π1,2π7,8p4=π1,2π9,10q=π1,2z11\begin{array}[]{llllll}p_{1}=\pi_{1,2}\pi_{3,4}\\ p_{2}=\pi_{1,2}\pi_{5,6}\\ p_{3}=\pi_{1,2}\pi_{7,8}\\ p_{4}=\pi_{1,2}\pi_{9,10}\\ q=\pi_{1,2}z_{11}\end{array}
𝒮(2)\mathcal{S}^{(2)} 55 22 1111 (T1,θ)×(T2,π)r=8+3\begin{array}[]{llll}(T1,\theta)\times(T2,\pi)\\ \qquad r=8+3\end{array} (p1)(1)=π1,2π3,4,(q)(2)=π9,10z11(p2)(1)=π1,2π5,6,(p3)(1)=π1,2π7,8,(p4)(1)=θ1,7¯,\begin{array}[]{llllll}&(p_{1})_{(1)}=\pi_{1,2}\pi_{3,4},&(q)_{(2)}=\pi_{9,10}z_{11}\\ &(p_{2})_{(1)}=\pi_{1,2}\pi_{5,6},\\ &(p_{3})_{(1)}=\pi_{1,2}\pi_{7,8},\\ &(p_{4})_{(1)}=\theta_{\overline{1,7}},\end{array}
𝒮(3)\mathcal{S}^{(3)} 55 22 1111 (T2,π,θ)×(T1)r=7+4\begin{array}[]{llll}(T2,\pi,\theta)\times(T1)\\ \qquad r=7+4\end{array} (p1)(1)=π1,2π3,4,(p1)(2)=π8,9π10,11(p2)(1)=π1,2π5,6,(p3)(1)=θ1,7¯,(q)(1)=π1,2z7,\begin{array}[]{llllll}&(p_{1})_{(1)}=\pi_{1,2}\pi_{3,4},&(p_{1})_{(2)}=\pi_{8,9}\pi_{10,11}\\ &(p_{2})_{(1)}=\pi_{1,2}\pi_{5,6},\\ &(p_{3})_{(1)}=\theta_{\overline{1,7}},\\ &(q)_{(1)}=\pi_{1,2}z_{7},\end{array}
Proof.

The standard subgroup 𝒮𝟓,𝟎𝒮(1)\mathcal{S}^{\bf 5,\bf 0}\subset\mathcal{S}^{(1)} is generated by involutions (5.4). We need to add one involution, since (11,0)=5\ell(11,0)=5. We add q=π1,2z11q=\pi_{1,2}z_{11}. A reduction to the cases |𝔟(PI11,0)|=10|\mathfrak{b}(PI_{11,0})|=10 is not possible due to (10,0)<(11,0)\ell(10,0)<\ell(11,0).

To construct the disconnected subgroup 𝒮(2)=𝒮(1)(2)×𝒮(2)(2)\mathcal{S}^{(2)}=\mathcal{S}_{(1)}^{(2)}\times\mathcal{S}_{(2)}^{(2)} corresponding to the 2\mathbb{Z}_{2}-graded tensor product of the Clifford algebras Cl11,0Cl8,0^Cl3,0\text{Cl}_{11,0}\cong\text{Cl}_{8,0}\hat{\otimes}\text{Cl}_{3,0} we start from the standard subgroup 𝒮(1)𝟒,𝟎𝒮(1)(2)\mathcal{S}_{(1)}^{\bf 4,0}\subset\mathcal{S}_{(1)}^{(2)} generated by (5.3) and add type T1T_{1} involution θ1,7¯\theta_{\overline{1,7}}. Then 𝒮(2)(2)={𝟏,π9,10z11}\mathcal{S}_{(2)}^{(2)}=\{{\bf 1},\pi_{9,10}z_{11}\}.

To obtain 𝒮(3)=𝒮(1)(3)×𝒮(2)(3)\mathcal{S}^{(3)}=\mathcal{S}_{(1)}^{(3)}\times\mathcal{S}_{(2)}^{(3)} corresponding to the 2\mathbb{Z}_{2}-graded tensor product of the Clifford algebras Cl11,0Cl7,0^Cl4,0\text{Cl}_{11,0}\cong\text{Cl}_{7,0}\hat{\otimes}\text{Cl}_{4,0} we consider standard subgroup 𝒮(1)𝟑,𝟎𝒮(1)(3)\mathcal{S}_{(1)}^{\bf 3,0}\subset\mathcal{S}_{(1)}^{(3)} generated by (5.2) and add type T1T_{1} involution θ1,7¯\theta_{\overline{1,7}} and type T2T_{2} involution π1,2z7\pi_{1,2}z_{7}. Then 𝒮(2)(3)={𝟏,π8,9π10,11}\mathcal{S}_{(2)}^{(3)}=\{{\bf 1},\pi_{8,9}\pi_{10,11}\}. ∎

Theorem 5.12.

There are three connected non-equivalent and five disconnected non-equivalent subgroups in 𝕊12,0M\mathbb{S}^{M}_{12,0}.

Table 10. Groups for r=12r=12
(12,0)\ell(12,0) π0(𝒮)\pi_{0}(\mathcal{S}) |𝔟(PI)||\mathfrak{b}(PI)| Signature PIPI
𝒮(1)\mathcal{S}^{(1)} 55 11 1212 (T1,π)(T1,\pi) p1=π1,2π3,4p2=π1,2π5,6p3=π1,2π7,8p4=π1,2π9,10p5=π1,2π11,12\begin{array}[]{llllll}p_{1}=\pi_{1,2}\pi_{3,4}\\ p_{2}=\pi_{1,2}\pi_{5,6}\\ p_{3}=\pi_{1,2}\pi_{7,8}\\ p_{4}=\pi_{1,2}\pi_{9,10}\\ p_{5}=\pi_{1,2}\pi_{11,12}\end{array}
𝒮(2)\mathcal{S}^{(2)} 55 11 1212 (T2,θ)(T2,\theta) p1=π1,2π3,4p2=π1,2π5,6p3=π1,2π7,8p4=π1,2π9,10q=θ1,9¯z11z12\begin{array}[]{llllll}p_{1}=\pi_{1,2}\pi_{3,4}\\ p_{2}=\pi_{1,2}\pi_{5,6}\\ p_{3}=\pi_{1,2}\pi_{7,8}\\ p_{4}=\pi_{1,2}\pi_{9,10}\\ q=\theta_{\overline{1,9}}z_{11}z_{12}\end{array}
𝒮(3)\mathcal{S}^{(3)} 55 11 1111 (T2,π)(T2,\pi) p1=π1,2π3,4p2=π1,2π5,6p3=π1,2π7,8p4=π1,2π9,10q=π1,2z11\begin{array}[]{llllll}&p_{1}=\pi_{1,2}\pi_{3,4}\\ &p_{2}=\pi_{1,2}\pi_{5,6}\\ &p_{3}=\pi_{1,2}\pi_{7,8}\\ &p_{4}=\pi_{1,2}\pi_{9,10}\\ &q=\pi_{1,2}z_{11}\end{array}
𝒮(4)\mathcal{S}^{(4)} 55 22 1212 (T1,θ)×(T1)r=8+4\begin{array}[]{llll}(T1,\theta)\times(T1)\\ \quad r=8+4\end{array} (p1)(1)=π1,2π3,4,(p1)(2)=π9,10π11,12(p2)(1)=π1,2π5,6,(p3)(1)=π1,2π7,8,(p4)(1)=θ1,7¯,\begin{array}[]{llllll}&(p_{1})_{(1)}=\pi_{1,2}\pi_{3,4},&(p_{1})_{(2)}=\pi_{9,10}\pi_{11,12}\\ &(p_{2})_{(1)}=\pi_{1,2}\pi_{5,6},\\ &(p_{3})_{(1)}=\pi_{1,2}\pi_{7,8},\\ &(p_{4})_{(1)}=\theta_{\overline{1,7}},\end{array}
𝒮(5)\mathcal{S}^{(5)} 55 22 1212 (T1,θ)×(T2,π)r=7+5\begin{array}[]{llll}(T1,\theta)\times(T2,\pi)\\ \quad r=7+5\end{array} (p1)(1)=π1,2π3,4,(p1)(2)=π8,9π10,11(p2)(1)=π1,2π5,6,(q)(2)=π8,9z12(p3)(1)=θ1,7¯,\begin{array}[]{llllll}&(p_{1})_{(1)}=\pi_{1,2}\pi_{3,4},&(p_{1})_{(2)}=\pi_{8,9}\pi_{10,11}\\ &(p_{2})_{(1)}=\pi_{1,2}\pi_{5,6},&(q)_{(2)}=\pi_{8,9}z_{12}\\ &(p_{3})_{(1)}=\theta_{\overline{1,7}},\end{array}
𝒮(6)\mathcal{S}^{(6)} 55 22 1212 (T2,π)×(T1)r=6+6\begin{array}[]{llll}(T2,\pi)\times(T1)\\ \quad r=6+6\end{array} (p1)(1)=π1,2π3,4,(p1)(2)=π7,8π9,10(p2)(1)=π1,2π5,6,(p2)(2)=π7,8π11,12(q)(1)=θ1,5¯,\begin{array}[]{llllll}&(p_{1})_{(1)}=\pi_{1,2}\pi_{3,4},&(p_{1})_{(2)}=\pi_{7,8}\pi_{9,10}\\ &(p_{2})_{(1)}=\pi_{1,2}\pi_{5,6},&(p_{2})_{(2)}=\pi_{7,8}\pi_{11,12}\\ &(q)_{(1)}=\theta_{\overline{1,5}},\end{array}
𝒮(7)\mathcal{S}^{(7)} 55 22 1111 (T1,θ)×(T2,π)r=8+3\begin{array}[]{llll}(T1,\theta)\times(T2,\pi)\\ \quad r=8+3\end{array} (p1)(1)=π1,2π3,4,(q)(1)=π9,10z11(p2)(1)=π1,2π5,6,(p3)(1)=π1,2π7,8,(p4)(1)=θ1,7¯,\begin{array}[]{lll}&(p_{1})_{(1)}=\pi_{1,2}\pi_{3,4},&(q)_{(1)}=\pi_{9,10}z_{11}\\ &(p_{2})_{(1)}=\pi_{1,2}\pi_{5,6},\\ &(p_{3})_{(1)}=\pi_{1,2}\pi_{7,8},\\ &(p_{4})_{(1)}=\theta_{\overline{1,7}},\end{array}
𝒮(8)\mathcal{S}^{(8)} 55 22 1111 (T2,π,θ)×(T1)r=7+4\begin{array}[]{llll}(T2,\pi,\theta)\times(T1)\\ \quad r=7+4\end{array} (p1)(1)=π1,2π3,4,(p1)(2)=π8,9π10,11(p2)(1)=π1,2π5,6,(p3)(1)=θ1,7¯,(q)(1)=π1,2z7,\begin{array}[]{llllll}&(p_{1})_{(1)}=\pi_{1,2}\pi_{3,4},&(p_{1})_{(2)}=\pi_{8,9}\pi_{10,11}\\ &(p_{2})_{(1)}=\pi_{1,2}\pi_{5,6},\\ &(p_{3})_{(1)}=\theta_{\overline{1,7}},\\ &(q)_{(1)}=\pi_{1,2}z_{7},\end{array}
Proof.

The standard subgroup 𝒮𝟔,𝟎𝒮(1)\mathcal{S}^{\bf 6,\bf 0}\subseteq\mathcal{S}^{(1)} is generated by involutions

(5.5) p1=π1,2π3,4,p2=π1,2π5,6,p3=π1,2π7,8,p4=π1,2π9,10,p5=π1,2π11,12.\begin{array}[]{llll}&p_{1}=\pi_{1,2}\pi_{3,4},\quad p_{2}=\pi_{1,2}\pi_{5,6},\quad p_{3}=\pi_{1,2}\pi_{7,8},\quad p_{4}=\pi_{1,2}\pi_{9,10},\\ &p_{5}=\pi_{1,2}\pi_{11,12}.\end{array}

and it coincides with 𝒮(1)𝕊12,0M\mathcal{S}^{(1)}\in\mathbb{S}^{M}_{12,0}.

Consider the standard subgroup 𝒮𝟓,𝟎𝒮(2)\mathcal{S}^{\bf 5,\bf 0}\subseteq\mathcal{S}^{(2)} generated by involutions (5.4). We need to add one involution containing z11z_{11} and z12z_{12}. We see that q=θ1,9¯z11z12q=\theta_{\overline{1,9}}z_{11}z_{12} commutes with all involutions in (5.4). Adding qq as the type T2T_{2} involution will finish the construction of the maximal group 𝒮(2)\mathcal{S}^{(2)}, see Table 10.

To construct the disconnected subgroup 𝒮(3)=𝒮(1)(3)×𝒮(2)(3)\mathcal{S}^{(3)}=\mathcal{S}_{(1)}^{(3)}\times\mathcal{S}_{(2)}^{(3)} corresponding to the 2\mathbb{Z}_{2}-graded tensor product of the Clifford algebras Cl12,0Cl8,0^Cl4,0\text{Cl}_{12,0}\cong\text{Cl}_{8,0}\hat{\otimes}\text{Cl}_{4,0} we consider standard subgroup 𝒮(1)𝟒,𝟎𝒮(1)(3)\mathcal{S}_{(1)}^{\bf 4,0}\subset\mathcal{S}_{(1)}^{(3)} generated by (5.3). We add the involution p4=θ1,7¯p_{4}=\theta_{\overline{1,7}} to the set of generators for 𝒮(1)𝟒,𝟎\mathcal{S}_{(1)}^{\bf 4,0} and generate the first component 𝒮(1)(3)\mathcal{S}_{(1)}^{(3)} in the product 𝒮(3)=𝒮(1)(3)×𝒮(2)(3)\mathcal{S}^{(3)}=\mathcal{S}_{(1)}^{(3)}\times\mathcal{S}_{(2)}^{(3)}. Then we set 𝒮(2)(3)={𝟏,π9,10π11,12}\mathcal{S}_{(2)}^{(3)}=\{{\bf 1},\pi_{9,10}\pi_{11,12}\}.

Analogously we construct the disconnected subgroups related to the decomposition of the Clifford algebras Cl12,0Cl7,0^Cl5,0\text{Cl}_{12,0}\cong\text{Cl}_{7,0}\hat{\otimes}\text{Cl}_{5,0} and Cl12,0Cl6,0^Cl6,0\text{Cl}_{12,0}\cong\text{Cl}_{6,0}\hat{\otimes}\text{Cl}_{6,0}. In both of these cases we remove one of the type T2T_{2} involutions and obtain 5 involutions in the total set PI12,0PI_{12,0}. Note also that if in the decomposition Cl12,0Cl7,0^Cl5,0\text{Cl}_{12,0}\cong\text{Cl}_{7,0}\hat{\otimes}\text{Cl}_{5,0} for 𝒮(5)=𝒮(1)(5)×𝒮(2)(5)\mathcal{S}^{(5)}=\mathcal{S}_{(1)}^{(5)}\times\mathcal{S}_{(2)}^{(5)} we take the set PI7,0PI_{7,0} to be of (T2)(T2)-type set generating 𝒮(1)(5)\mathcal{S}_{(1)}^{(5)} and PI4,0PI_{4,0} for 𝒮(2)(5)\mathcal{S}_{(2)}^{(5)} to be of (T1)(T1)-type set, then we obtain a group isomorphic to 𝒮(8)\mathcal{S}^{(8)}.

If |𝔟(PI12,0)|=11|\mathfrak{b}(PI_{12,0})|=11, then the constructions reduce to the case of 𝒮𝕊11,0M\mathcal{S}\subset\mathbb{S}_{11,0}^{M}. ∎

Theorem 5.13.

There are three connected non-equivalent and three disconnected non-equivalent subgroups in 𝕊13,0M\mathbb{S}^{M}_{13,0}.

Table 11. Groups for r=13r=13
(13,0)\ell(13,0) π0(𝒮)\pi_{0}(\mathcal{S}) |𝔟(PI)||\mathfrak{b}(PI)| Signature PIPI
𝒮(1)\mathcal{S}^{(1)} 66 11 1313 (T2,π)(T2,\pi) p1=π1,2π3,4p2=π1,2π5,6p3=π1,2π7,8p4=π1,2π9,10p5=π1,2π11,12q=π1,2z13\begin{array}[]{llllll}p_{1}=\pi_{1,2}\pi_{3,4}\\ p_{2}=\pi_{1,2}\pi_{5,6}\\ p_{3}=\pi_{1,2}\pi_{7,8}\\ p_{4}=\pi_{1,2}\pi_{9,10}\\ p_{5}=\pi_{1,2}\pi_{11,12}\\ q=\pi_{1,2}z_{13}\end{array}
𝒮(2)\mathcal{S}^{(2)} 66 11 1313 (T2,θ)(T2,\theta) p1=π1,2π3,4p2=π1,2π5,6p3=π1,2π7,8p4=π1,2π9,10p5=π1,2π11,12q=θ1,13¯\begin{array}[]{llllll}p_{1}=\pi_{1,2}\pi_{3,4}\\ p_{2}=\pi_{1,2}\pi_{5,6}\\ p_{3}=\pi_{1,2}\pi_{7,8}\\ p_{4}=\pi_{1,2}\pi_{9,10}\\ p_{5}=\pi_{1,2}\pi_{11,12}\\ q=\theta_{\overline{1,13}}\end{array}
𝒮(3)\mathcal{S}^{(3)} 66 11 1313 (T2,π,θ)(T2,\pi,\theta) p1=π1,2π3,4p2=π1,2π5,6p3=π1,2π7,8p4=π1,2π9,10p5=θ1,9¯z11z12z13q=π1,2z12\begin{array}[]{lll}p_{1}=\pi_{1,2}\pi_{3,4}\\ p_{2}=\pi_{1,2}\pi_{5,6}\\ p_{3}=\pi_{1,2}\pi_{7,8}\\ p_{4}=\pi_{1,2}\pi_{9,10}\\ p_{5}=\theta_{\overline{1,9}}z_{11}z_{12}z_{13}\\ q=\pi_{1,2}z_{12}\end{array}
𝒮(4)\mathcal{S}^{(4)} 66 22 1313 (T1)×(T2)r=8+5\begin{array}[]{llll}(T1)\times(T2)\\ \quad r=8+5\end{array} (p1)(1)=π1,2π3,4,(p1)(2)=π9,10π11,12(p2)(1)=π1,2π5,6,(q)(2)=π9,10z13(p3)(1)=π1,2π7,8,(p4)(1)=θ1,7¯,\begin{array}[]{llllll}&(p_{1})_{(1)}=\pi_{1,2}\pi_{3,4},&(p_{1})_{(2)}=\pi_{9,10}\pi_{11,12}\\ &(p_{2})_{(1)}=\pi_{1,2}\pi_{5,6},&(q)_{(2)}=\pi_{9,10}z_{13}\\ &(p_{3})_{(1)}=\pi_{1,2}\pi_{7,8},\\ &(p_{4})_{(1)}=\theta_{\overline{1,7}},\end{array}
𝒮(5)\mathcal{S}^{(5)} 66 22 1313 (T2,π,θ)×(T1)r=7+6\begin{array}[]{llll}(T2,\pi,\theta)\times(T1)\\ \qquad r=7+6\end{array} (p1)(1)=π1,2π3,4,(p1)(2)=π8,9π10,11(p2)(1)=π1,2π5,6,(p2)(2)=π8,9π12,13(p3)(1)=θ1,7¯,(q)(1)=π1,2z7,\begin{array}[]{llllll}&(p_{1})_{(1)}=\pi_{1,2}\pi_{3,4},&(p_{1})_{(2)}=\pi_{8,9}\pi_{10,11}\\ &(p_{2})_{(1)}=\pi_{1,2}\pi_{5,6},&(p_{2})_{(2)}=\pi_{8,9}\pi_{12,13}\\ &(p_{3})_{(1)}=\theta_{\overline{1,7}},\\ &(q)_{(1)}=\pi_{1,2}z_{7},\end{array}
𝒮(6)\mathcal{S}^{(6)} 66 22 1313 (T1,θ)×(T2,θ)r=7+6\begin{array}[]{llll}(T1,\theta)\times(T2,\theta)\\ \quad r=7+6\end{array} (p1)(1)=π1,2π3,4,(p1)(2)=π8,9π10,11(p2)(1)=π1,2π5,6,(p2)(2)=π8,9π12,13(p3)(1)=θ1,7¯,(q)(2)=z8z10z12\begin{array}[]{llllll}&(p_{1})_{(1)}=\pi_{1,2}\pi_{3,4},&(p_{1})_{(2)}=\pi_{8,9}\pi_{10,11}\\ &(p_{2})_{(1)}=\pi_{1,2}\pi_{5,6},&(p_{2})_{(2)}=\pi_{8,9}\pi_{12,13}\\ &(p_{3})_{(1)}=\theta_{\overline{1,7}},&(q)_{(2)}=z_{8}z_{10}z_{12}\end{array}
Proof.

The standard subgroup 𝒮𝟔,𝟎𝒮𝕊13,0M\mathcal{S}^{\bf 6,\bf 0}\subseteq\mathcal{S}\subset\mathbb{S}^{M}_{13,0} is generated by involutions (5.5). We add either q=π1,2z13q=\pi_{1,2}z_{13} or q=θ1,13¯q=\theta_{\overline{1,13}} as type T2T_{2} involutions. We obtain two connected groups 𝒮(1)\mathcal{S}^{(1)} and 𝒮(2)\mathcal{S}^{(2)}.

Consider the standard subgroup 𝒮𝟓,𝟎𝒮(3)\mathcal{S}^{\bf 5,\bf 0}\subseteq\mathcal{S}^{(3)} generated by involutions (5.4). We need to add two involutions containing z11,z12z_{11},z_{12} and z13z_{13}. We see that type T1T_{1} involution p5=θ1,9¯z11z12z13p_{5}=\theta_{\overline{1,9}}z_{11}z_{12}z_{13} commutes with all involutions in (5.4). Adding p5p_{5} as the type T1T_{1} involution and q=π1,2z13q=\pi_{1,2}z_{13} as type T2T_{2} involution, we obtain the maximal group 𝒮(3)\mathcal{S}^{(3)}, see Table 11.

To construct the disconnected subgroup 𝒮(4)=𝒮(1)(4)×𝒮(2)(4)\mathcal{S}^{(4)}=\mathcal{S}_{(1)}^{(4)}\times\mathcal{S}_{(2)}^{(4)} corresponding to the 2\mathbb{Z}_{2}- graded tensor product of the Clifford algebras Cl13,0Cl8,0^Cl5,0\text{Cl}_{13,0}\cong\text{Cl}_{8,0}\hat{\otimes}\text{Cl}_{5,0} we consider standard subgroup 𝒮(1)𝟒,𝟎𝒮(1)(4)\mathcal{S}_{(1)}^{\bf 4,0}\subset\mathcal{S}_{(1)}^{(4)} generated by (5.3). We add the involutions p4=θ1¯,7p_{4}=\theta_{\overline{1},7} to the set of generators for 𝒮(1)𝟒,𝟎\mathcal{S}_{(1)}^{\bf 4,0} and generate the first component 𝒮(1)(4)\mathcal{S}_{(1)}^{(4)} in the product 𝒮(4)=𝒮(1)(4)×𝒮(2)(4)\mathcal{S}^{(4)}=\mathcal{S}_{(1)}^{(4)}\times\mathcal{S}_{(2)}^{(4)}. Then we set 𝒮(2)(4)\mathcal{S}_{(2)}^{(4)} generated by the set of PI={(p1)(2)=π9,10π11,12,(q)(2)=π9,10z13}PI=\{(p_{1})_{(2)}=\pi_{9,10}\pi_{11,12},(q)_{(2)}=\pi_{9,10}z_{13}\}.

Analogously we construct disconnected subgroups 𝒮(k)=𝒮(1)(k)×𝒮(2)(k)\mathcal{S}^{(k)}=\mathcal{S}_{(1)}^{(k)}\times\mathcal{S}_{(2)}^{(k)}, k=5,6k=5,6, corresponding to the 2\mathbb{Z}_{2}-graded tensor product Cl13,0Cl7,0^Cl6,0\text{Cl}_{13,0}\cong\text{Cl}_{7,0}\hat{\otimes}\text{Cl}_{6,0}. For k=5k=5 we choose PI7,0PI_{7,0} for the group 𝒮(1)(5)\mathcal{S}_{(1)}^{(5)} to be (T2)(T2)-type set and two standard involutions in PI6,0PI_{6,0} for the groups 𝒮(2)(5)\mathcal{S}_{(2)}^{(5)} to be (T1)(T1)-type set. For k=6k=6 we change the type of the sets PIPI.

There are no groups with |𝔟(PI13,0)|=12|\mathfrak{b}(PI_{13,0})|=12 because (13,0)>(12,0)\ell(13,0)>\ell(12,0). ∎

Theorem 5.14.

There are two connected non-equivalent and two disconnected non-equivalent subgroups in 𝕊14,0M\mathbb{S}^{M}_{14,0}.

Table 12. Groups for r=14r=14
(14,0)\ell(14,0) π0(𝒮)\pi_{0}(\mathcal{S}) |𝔟(PI)||\mathfrak{b}(PI)| Signature PIPI
𝒮(1)\mathcal{S}^{(1)} 77 11 1414 (T2,θ)(T2,\theta) p1=π1,2π3,4p2=π1,2π5,6p3=π1,2π7,8p4=π1,2π9,10p5=π1,2π11,12p6=π1,2π13,14q=θ1,13¯\begin{array}[]{llllll}p_{1}=\pi_{1,2}\pi_{3,4}\\ p_{2}=\pi_{1,2}\pi_{5,6}\\ p_{3}=\pi_{1,2}\pi_{7,8}\\ p_{4}=\pi_{1,2}\pi_{9,10}\\ p_{5}=\pi_{1,2}\pi_{11,12}\\ p_{6}=\pi_{1,2}\pi_{13,14}\\ q=\theta_{\overline{1,13}}\end{array}
𝒮(2)\mathcal{S}^{(2)} 77 11 1414 (T2,π,θ)(T2,\pi,\theta) p1=π1,2π3,4p2=π1,2π5,6p3=π1,2π7,8p4=π1,2π9,10p5=π1,2π11,12p6=θ1,11¯z13z14q=π1,2z14\begin{array}[]{llllll}p_{1}=\pi_{1,2}\pi_{3,4}\\ p_{2}=\pi_{1,2}\pi_{5,6}\\ p_{3}=\pi_{1,2}\pi_{7,8}\\ p_{4}=\pi_{1,2}\pi_{9,10}\\ p_{5}=\pi_{1,2}\pi_{11,12}\\ p_{6}=\theta_{\overline{1,11}}z_{13}z_{14}\\ q=\pi_{1,2}z_{14}\end{array}
𝒮(3)\mathcal{S}^{(3)} 77 22 1414 (T1,θ)×(T2,θ)r=8+6\begin{array}[]{llll}(T1,\theta)\times(T2,\theta)\\ \quad r=8+6\end{array} (p1)(1)=π1,2π3,4,(p1)(2)=π9,10π11,12(p2)(1)=π1,2π5,6,(p2)(2)=π9,10π13,14(p3)(1)=π1,2π7,8,(q)(2)=θ9,13¯(p4)(1)=θ1,7¯,\begin{array}[]{llllll}&(p_{1})_{(1)}=\pi_{1,2}\pi_{3,4},&(p_{1})_{(2)}=\pi_{9,10}\pi_{11,12}\\ &(p_{2})_{(1)}=\pi_{1,2}\pi_{5,6},&(p_{2})_{(2)}=\pi_{9,10}\pi_{13,14}\\ &(p_{3})_{(1)}=\pi_{1,2}\pi_{7,8},&(q)_{(2)}=\theta_{\overline{9,13}}\\ &(p_{4})_{(1)}=\theta_{\overline{1,7}},\end{array}
𝒮(4)\mathcal{S}^{(4)} 77 22 1414 (T2,π,θ)×(T1,θ)r=7+7\begin{array}[]{llll}(T2,\pi,\theta)\times(T1,\theta)\\ \quad r=7+7\end{array} (p1)(1)=π1,2π3,4,(p1)(2)=π8,9π10,11(p2)(1)=π1,2π5,6,(p2)(2)=π8,9π12,13(p3)(1)=θ1,7¯,(p3)(2)=θ9,13¯z14(q4)(1)=π1,2z7,\begin{array}[]{llllll}&(p_{1})_{(1)}=\pi_{1,2}\pi_{3,4},&(p_{1})_{(2)}=\pi_{8,9}\pi_{10,11}\\ &(p_{2})_{(1)}=\pi_{1,2}\pi_{5,6},&(p_{2})_{(2)}=\pi_{8,9}\pi_{12,13}\\ &(p_{3})_{(1)}=\theta_{\overline{1,7}},&(p_{3})_{(2)}=\theta_{\overline{9,13}}z_{14}\\ &(q_{4})_{(1)}=\pi_{1,2}z_{7},\end{array}
Proof.

The standard subgroup 𝒮𝟕,𝟎𝒮(1)𝕊14,0M\mathcal{S}^{\bf 7,\bf 0}\subseteq\mathcal{S}^{(1)}\subset\mathbb{S}^{M}_{14,0} is generated by involutions

(5.6) p1=π1,2π3,4,p2=π1,2π5,6,p3=π1,2π7,8,p4=π1,2π9,10,p5=π1,2π11,12,p6=π1,2π13,14.\begin{array}[]{lllll}&p_{1}=\pi_{1,2}\pi_{3,4},\quad p_{2}=\pi_{1,2}\pi_{5,6},\quad p_{3}=\pi_{1,2}\pi_{7,8},\quad p_{4}=\pi_{1,2}\pi_{9,10},\\ &p_{5}=\pi_{1,2}\pi_{11,12},\quad p_{6}=\pi_{1,2}\pi_{13,14}.\end{array}

We add type T2T_{2} involution q=θ1,13¯q=\theta_{\overline{1,13}} and obtain the connected group 𝒮(1)\mathcal{S}^{(1)}.

Next we consider the standard subgroup 𝒮𝟔,𝟎𝒮(2)\mathcal{S}^{\bf 6,\bf 0}\subseteq\mathcal{S}^{(2)} generated by involutions (5.5). We need to add two involutions containing z13z_{13} and z14z_{14}. We see that type T1T_{1} involution p6=θ1,11¯z13z14p_{6}=\theta_{\overline{1,11}}z_{13}z_{14} commutes with involutions in (5.5). Adding either q1=π1,2z13q_{1}=\pi_{1,2}z_{13} or q2=π1,2z14q_{2}=\pi_{1,2}z_{14} as type T2T_{2} involution, we obtain the maximal group 𝒮(2)\mathcal{S}^{(2)}, see Table 12. Adding q1q_{1} or q2q_{2}, we create equivalent groups 𝒮(2)\mathcal{S}^{(2)}.

To construct the disconnected subgroup 𝒮(3)=𝒮(1)(3)×𝒮(2)(3)\mathcal{S}^{(3)}=\mathcal{S}_{(1)}^{(3)}\times\mathcal{S}_{(2)}^{(3)} corresponding to the 2\mathbb{Z}_{2}-graded tensor product of the Clifford algebras Cl14,0Cl8,0^Cl6,0\text{Cl}_{14,0}\cong\text{Cl}_{8,0}\hat{\otimes}\text{Cl}_{6,0} we consider standard subgroup 𝒮(1)𝟒,𝟎𝒮(1)(3)\mathcal{S}_{(1)}^{\bf 4,0}\subset\mathcal{S}_{(1)}^{(3)} generated by (5.3). We add the involutions p4=θ1,7¯p_{4}=\theta_{\overline{1,7}} to the set of generators for 𝒮(1)𝟒,𝟎\mathcal{S}_{(1)}^{\bf 4,0} and generate the first component 𝒮(1)(3)\mathcal{S}_{(1)}^{(3)} in the product 𝒮(3)=𝒮(1)(3)×𝒮(2)(3)\mathcal{S}^{(3)}=\mathcal{S}_{(1)}^{(3)}\times\mathcal{S}_{(2)}^{(3)}. Then we set 𝒮(2)(3)\mathcal{S}_{(2)}^{(3)} to be generated by

PI={(p1)(2)=π9,10π11,12,(p2)(2)=π9,10π13,14,(q)(2)=θ9,13¯}.PI=\{(p_{1})_{(2)}=\pi_{9,10}\pi_{11,12},\ (p_{2})_{(2)}=\pi_{9,10}\pi_{13,14},\ (q)_{(2)}=\theta_{\overline{9,13}}\}.

The disconnected subgroup 𝒮(4)=𝒮(1)(4)×𝒮(2)(4)\mathcal{S}^{(4)}=\mathcal{S}_{(1)}^{(4)}\times\mathcal{S}_{(2)}^{(4)} corresponding to the 2\mathbb{Z}_{2}-graded tensor product of the Clifford algebras Cl14,0Cl7,0^Cl7,0\text{Cl}_{14,0}\cong\text{Cl}_{7,0}\hat{\otimes}\text{Cl}_{7,0}, generated similarly. We remove the type T2T_{2} involution from one of the sets PI7,0PI_{7,0} generating 𝒮(k)(4)\mathcal{S}_{(k)}^{(4)}, k=1k=1 or k=2k=2 in order to get a commutative set for 𝒮(4)\mathcal{S}^{(4)} with (14,0)=7\ell(14,0)=7.

There are no groups with |𝔟(PI14,0)|=13|\mathfrak{b}(PI_{14,0})|=13 because (14,0)>(13,0)\ell(14,0)>\ell(13,0). ∎

Theorem 5.15.

There are one connected and one disconnected subgroups in 𝕊15,0M\mathbb{S}^{M}_{15,0}.

Table 13. Groups for r=15r=15
(15,0)\ell(15,0) π0(𝒮)\pi_{0}(\mathcal{S}) |𝔟(PI)||\mathfrak{b}(PI)| Signature PIPI
𝒮(1)\mathcal{S}^{(1)} 88 11 1515 (T2,π)(T2,\pi) p1=π1,2π3,4p2=π1,2π5,6p3=π1,2π7,8p4=π1,2π9,10p5=π1,2π11,12p6=π1,2π13,14p7=θ1,13¯z15q=π1,2z15\begin{array}[]{llllll}p_{1}=\pi_{1,2}\pi_{3,4}\\ p_{2}=\pi_{1,2}\pi_{5,6}\\ p_{3}=\pi_{1,2}\pi_{7,8}\\ p_{4}=\pi_{1,2}\pi_{9,10}\\ p_{5}=\pi_{1,2}\pi_{11,12}\\ p_{6}=\pi_{1,2}\pi_{13,14}\\ p_{7}=\theta_{\overline{1,13}}z_{15}\\ q=\pi_{1,2}z_{15}\end{array}
𝒮(2)\mathcal{S}^{(2)} 88 22 1515 (T1,θ)×(T2,π,θ)r=8+7\begin{array}[]{llll}(T1,\theta)\times(T2,\pi,\theta)\\ \quad r=8+7\end{array} (p1)(1)=π1,2π3,4,(p1)(2)=π9,10π11,12(p2)(1)=π1,2π5,6,(p2)(2)=π9,10π13,14(p3)(1)=π1,2π7,8,(p3)(2)=θ9,15¯(p4)(1)=θ1,7¯,(q)(2)=π9,10z15\begin{array}[]{llllll}&(p_{1})_{(1)}=\pi_{1,2}\pi_{3,4},&(p_{1})_{(2)}=\pi_{9,10}\pi_{11,12}\\ &(p_{2})_{(1)}=\pi_{1,2}\pi_{5,6},&(p_{2})_{(2)}=\pi_{9,10}\pi_{13,14}\\ &(p_{3})_{(1)}=\pi_{1,2}\pi_{7,8},&(p_{3})_{(2)}=\theta_{\overline{9,15}}\\ &(p_{4})_{(1)}=\theta_{\overline{1,7}},&(q)_{(2)}=\pi_{9,10}z_{15}\end{array}
Proof.

The standard subgroup 𝒮𝟕,𝟎𝒮(1)𝕊15,0M\mathcal{S}^{\bf 7,\bf 0}\subseteq\mathcal{S}^{(1)}\subset\mathbb{S}^{M}_{15,0} is generated by involutions (5.6) We add type T1T_{1} involution p7=θ1,15¯p_{7}=\theta_{\overline{1,15}} and type T2T_{2} involution q=π1,2z15q=\pi_{1,2}z_{15}. We obtain the connected group 𝒮(1)\mathcal{S}^{(1)}.

To construct the disconnected subgroup 𝒮(2)=𝒮(1)(2)×𝒮(2)(2)\mathcal{S}^{(2)}=\mathcal{S}_{(1)}^{(2)}\times\mathcal{S}_{(2)}^{(2)} corresponding to the 2\mathbb{Z}_{2}-graded tensor product of the Clifford algebras Cl15,0Cl8,0^Cl7,0\text{Cl}_{15,0}\cong\text{Cl}_{8,0}\hat{\otimes}\text{Cl}_{7,0} we proceed as in the case r=14r=14 for 𝒮(1)(3)\mathcal{S}_{(1)}^{(3)}, and set 𝒮(2)(3)\mathcal{S}_{(2)}^{(3)} to be generated by

PI={\displaystyle PI=\{ (p1)(2)\displaystyle(p_{1})_{(2)} =π9,10π11,12,(p2)(2)=π9,10π13,14,\displaystyle=\pi_{9,10}\pi_{11,12},\ (p_{2})_{(2)}=\pi_{9,10}\pi_{13,14},\
(p3)(2)\displaystyle(p_{3})_{(2)} =θ9,13¯,(q)(2)=π9,10z15}.\displaystyle=\theta_{\overline{9,13}},\quad(q)_{(2)}=\pi_{9,10}z_{15}\}.

There are no groups with |𝔟(PI15,0)|=14|\mathfrak{b}(PI_{15,0})|=14 because (15,0)>(14,0)\ell(15,0)>\ell(14,0). ∎

Theorem 5.16.

There are two connected non-equivalent and two disconnected non-equivalent subgroups in 𝕊16,0M\mathbb{S}^{M}_{16,0}.

Table 14. Groups for r=16r=16
(16,0)\ell(16,0) π0(𝒮)\pi_{0}(\mathcal{S}) |𝔟(PI)||\mathfrak{b}(PI)| Signature PIPI
𝒮(1)\mathcal{S}^{(1)} 88 11 1616 (T1,π,θ)(T1,\pi,\theta) p1=π1,2π3,4p2=π1,2π5,6p3=π1,2π7,8p4=π1,2π9,10p5=π1,2π11,12p6=π1,2π13,14p7=θ1,13¯z15p8=π1,2z15z16\begin{array}[]{llllll}p_{1}=\pi_{1,2}\pi_{3,4}\\ p_{2}=\pi_{1,2}\pi_{5,6}\\ p_{3}=\pi_{1,2}\pi_{7,8}\\ p_{4}=\pi_{1,2}\pi_{9,10}\\ p_{5}=\pi_{1,2}\pi_{11,12}\\ p_{6}=\pi_{1,2}\pi_{13,14}\\ p_{7}=\theta_{\overline{1,13}}z_{15}\\ p_{8}=\pi_{1,2}z_{15}z_{16}\end{array}
𝒮(2)\mathcal{S}^{(2)} 88 11 1515 (T2,π,θ)(T2,\pi,\theta) p1=π1,2π3,4p2=π1,2π5,6p3=π1,2π7,8p4=π1,2π9,10p5=π1,2π11,12p6=π1,2π13,14p7=θ1,13¯z15q=π1,2z15\begin{array}[]{llllll}p_{1}=\pi_{1,2}\pi_{3,4}\\ p_{2}=\pi_{1,2}\pi_{5,6}\\ p_{3}=\pi_{1,2}\pi_{7,8}\\ p_{4}=\pi_{1,2}\pi_{9,10}\\ p_{5}=\pi_{1,2}\pi_{11,12}\\ p_{6}=\pi_{1,2}\pi_{13,14}\\ p_{7}=\theta_{\overline{1,13}}z_{15}\\ q=\pi_{1,2}z_{15}\end{array}
𝒮(3)\mathcal{S}^{(3)} 88 22 1616 (T1,θ)×(T1,θ)r=8+8\begin{array}[]{llll}(T1,\theta)\times(T1,\theta)\\ \quad r=8+8\end{array} (p1)1=π1,2π3,4,(p1)2=π9,10π11,12(p2)1=π1,2π5,6,(p2)2=π9,10π13,14(p3)1=π1,2π7,8,(p3)2=π9,10π15,16(p4)1=θ1,7¯,(p4)2=θ9,15¯\begin{array}[]{llllll}&(p_{1})_{1}=\pi_{1,2}\pi_{3,4},&(p_{1})_{2}=\pi_{9,10}\pi_{11,12}\\ &(p_{2})_{1}=\pi_{1,2}\pi_{5,6},&(p_{2})_{2}=\pi_{9,10}\pi_{13,14}\\ &(p_{3})_{1}=\pi_{1,2}\pi_{7,8},&(p_{3})_{2}=\pi_{9,10}\pi_{15,16}\\ &(p_{4})_{1}=\theta_{\overline{1,7}},&(p_{4})_{2}=\theta_{\overline{9,15}}\end{array}
𝒮(4)\mathcal{S}^{(4)} 88 22 1515 (T1,θ)×(T2,π,θ)r=8+7\begin{array}[]{llll}(T1,\theta)\times(T2,\pi,\theta)\\ \quad r=8+7\end{array} (p1)1=π1,2π3,4,(p1)2=π9,10π11,12(p2)1=π1,2π5,6,(p2)2=π9,10π13,14(p3)1=π1,2π7,8,(p3)2=θ9,15¯(p4)1=θ1,7¯,(q)2=π9,10z15\begin{array}[]{llllll}&(p_{1})_{1}=\pi_{1,2}\pi_{3,4},&(p_{1})_{2}=\pi_{9,10}\pi_{11,12}\\ &(p_{2})_{1}=\pi_{1,2}\pi_{5,6},&(p_{2})_{2}=\pi_{9,10}\pi_{13,14}\\ &(p_{3})_{1}=\pi_{1,2}\pi_{7,8},&(p_{3})_{2}=\theta_{\overline{9,15}}\\ &(p_{4})_{1}=\theta_{\overline{1,7}},&(q)_{2}=\pi_{9,10}z_{15}\end{array}
Proof.

The standard subgroup 𝒮𝟖,𝟎𝒮(1)𝕊16,0M\mathcal{S}^{\bf 8,\bf 0}\subseteq\mathcal{S}^{(1)}\subset\mathbb{S}^{M}_{16,0} is generated by involutions

(5.7) p1=π1,2π3,4,p2=π1,2π5,6,p3=π1,2π7,8,p4=π1,2π9,10p5=π1,2π11,12,p6=π1,2π13,14,p7=π1,2π15,16.\begin{array}[]{lll}p_{1}=\pi_{1,2}\pi_{3,4},\quad p_{2}=\pi_{1,2}\pi_{5,6},\quad p_{3}=\pi_{1,2}\pi_{7,8},\quad p_{4}=\pi_{1,2}\pi_{9,10}\\ p_{5}=\pi_{1,2}\pi_{11,12},\quad p_{6}=\pi_{1,2}\pi_{13,14},\quad p_{7}=\pi_{1,2}\pi_{15,16}.\end{array}

We add type T1T_{1} involution p7=θ1,15¯p_{7}=\theta_{\overline{1,15}} and obtain the connected group 𝒮(1)\mathcal{S}^{(1)}.

To construct the disconnected subgroups 𝒮(2)=𝒮(1)(2)×𝒮(2)(2)\mathcal{S}^{(2)}=\mathcal{S}_{(1)}^{(2)}\times\mathcal{S}_{(2)}^{(2)} corresponding to the 2\mathbb{Z}_{2}-graded tensor product of the Clifford algebras Cl16,0Cl8,0^Cl8,0\text{Cl}_{16,0}\cong\text{Cl}_{8,0}\hat{\otimes}\text{Cl}_{8,0} and Cl16,0Cl8,0^Cl7,0\text{Cl}_{16,0}\cong\text{Cl}_{8,0}\hat{\otimes}\text{Cl}_{7,0} we proceed as in the previous cases.

The group with |𝔟(PI16,0)|=15|\mathfrak{b}(PI_{16,0})|=15 coincides with the group 𝒮(1)𝕊15,0M\mathcal{S}^{(1)}\in\mathbb{S}^{M}_{15,0}. ∎

Theorem 5.17.

Theorems 5.3 – 5.16 are true for HH-type Lie algebras 𝔫r,1\mathfrak{n}_{r,1}, r{3,,16}r\in\{3,\ldots,16\}.

Proof.

For s=1s=1, the negative basis vector plays no role in forming the involutions, see Definition 3.5. ∎

Table 15. Number of non-equivalent groups
rr 1 2 3 4 5 6 7 8
π0(𝒮)=1\pi_{0}(\mathcal{S})=1 0 0 1 2 1 1 1 2
π0(𝒮)=2\pi_{0}(\mathcal{S})=2 0 0 0 0 0 0 0 0
rr 9 10 11 12 13 14 15 16
π0(𝒮)=1\pi_{0}(\mathcal{S})=1 3 4 1 3 3 2 1 2
π0(𝒮)=2\pi_{0}(\mathcal{S})=2 0 2 2 5 3 2 1 2

6. Isomorphism of invariant integral structures

Theorem 6.1.

If

(6.1) (r,s){(0,0),(1,0),(2,0),(0,1),(1,1),(2,1),(0,2)},(r,s)\in\{(0,0),(1,0),(2,0),(0,1),(1,1),(2,1),(0,2)\},

then for any orthonormal basis Br,s={zj}B_{r,s}=\{z_{j}\} and vVr,sv\in V^{r,s}, with v,vV1,0=±1\langle v,v\rangle_{V^{1,0}}=\pm 1 the invariant orthonormal structures spanned by bases as in Table 16 are isomorphic.

Table 16. Invariant integral structures for (r,s)(r,s) in Theorem 6.1
22 {v,Jz1v,Jz2v,Jz1Jz2v,z1,z2}\{v,J_{z_{1}}v,J_{z_{2}}v,J_{z_{1}}J_{z_{2}}v,z_{1},z_{2}\}
11 {v,Jz1v,z1}\{v,J_{z_{1}}v,z_{1}\} {v,Jz1v,Jz2v,Jz1Jz2v,z1,z2}\{v,J_{z_{1}}v,J_{z_{2}}v,J_{z_{1}}J_{z_{2}}v,z_{1},z_{2}\} {v,Jz1v,Jz2v,Jz1Jz2v,z1,z2}\{v,J_{z_{1}}v,J_{z_{2}}v,J_{z_{1}}J_{z_{2}}v,z_{1},z_{2}\}
0 vv {v,Jz1v,z1}\{v,J_{z_{1}}v,z_{1}\} {v,Jz1v,Jz2v,Jz1Jz2v,z1,z2}\{v,J_{z_{1}}v,J_{z_{2}}v,J_{z_{1}}J_{z_{2}}v,z_{1},z_{2}\}
s/rs/r 0 11 22
Proof.

There are only trivial groups 𝒮𝕊r,sM\mathcal{S}\subset\mathbb{S}^{M}_{r,s} for (r,s)(r,s) as in (6.1) since there are no involutions. The proof of uniqueness is literally repeats the proof of Proposition 3.16. See also discussions in Remark 3.2. ∎

6.1. Isomorphic invariant integral structures.

We fix an orthonormal basis Br,s={z1,,zr+s}B_{r,s}=\{z_{1},\ldots,z_{r+s}\} and a group 𝒮=𝒮(PIr,s)\mathcal{S}=\mathcal{S}(PI_{r,s}). Recall the construction of an invariant basis v(Vr,s)\mathcal{B}_{v}(V^{r,s}) on the minimal admissible module Vr,sV^{r,s} from Theorem 3.15, which used the centraliser of the isotropy group 𝒮=𝒮(PIr,s)=𝒮v\mathcal{S}=\mathcal{S}(PI_{r,s})=\mathcal{S}_{v} of a unit vector vVr,sv\in V^{r,s}. The invariant integral structure on the Lie algebra 𝔫r,s(Vr,s)\mathfrak{n}_{r,s}(V^{r,s}) given by 𝒮\mathcal{S} will be denoted by

(𝒮)=span{v(Vr,s)}span{Br,s}.\mathcal{L}(\mathcal{S})=\text{\rm span}\,_{\mathbb{Z}}\{\mathcal{B}_{v}(V^{r,s})\}\oplus\text{\rm span}\,_{\mathbb{Z}}\{B_{r,s}\}.
Theorem 6.2.

If two groups 𝒮1\mathcal{S}_{1} and 𝒮2\mathcal{S}_{2} are equivalent; that is there exists a map CO(r,s)C\in O(r,s) such that C(S1^))=S2^C(\mathaccent 866{S_{1}}))=\mathaccent 866{S_{2}}, then the invariant integral structures (𝒮1)\mathcal{L}(\mathcal{S}_{1}) and (𝒮2)\mathcal{L}(\mathcal{S}_{2}) are isomorphic under a map ACA\oplus C, where A:Vr,sVr,sA\colon V^{r,s}\to V^{r,s} is an orthogonal map with respect to .,.Vr,s\langle.\,,.\rangle_{V^{r,s}}; that is AτA=IdVr,sA^{\tau}A=\text{\rm Id}_{V^{r,s}}.

Proof.

The proof is a light generalisation of Proposition 3.16. Let 𝒮1=𝒮(PI1)\mathcal{S}_{1}=\mathcal{S}(PI_{1}) and 𝒮2=𝒮(PI2)\mathcal{S}_{2}=\mathcal{S}(PI_{2}) be equivalent groups. It imply that there is CO(r,s)C\in\text{\rm O}(r,s) such that C(S^1)=S^2C(\mathaccent 866{S}_{1})=\mathaccent 866{S}_{2} where we denoted by the same letter CC the extention of the orthogonal map to the group Clr,sClr,s\text{Cl}^{*}_{r,s}\subset\text{Cl}_{r,s} of invertible elements of the Clifford algebra Clr,s\text{Cl}_{r,s}. Let

(6.2) v(Vr,s)={v,Jσi(v),Jτj(v),JτjJσi(v)σi,τj,σiτjΣ(𝒮1)}\mathcal{B}_{v}(V^{r,s})=\Big{\{}v,J_{\sigma_{i}}(v),J_{\tau_{j}}(v),J_{\tau_{j}}J_{\sigma_{i}}(v)\mid\ \sigma_{i},\tau_{j},\sigma_{i}\tau_{j}\in\Sigma(\mathcal{S}_{1})\Big{\}}

be the invariant basis, constructed in Theorem 3.15 by making use the eigenspaces of involutions from PI1PI_{1}. The set PI1PI_{1} is equivalent to PI2PI_{2} under CC. We use the method of Theorem 3.15 and obtain a basis

(6.3) w(Vr,s)\displaystyle\mathcal{B}_{w}(V^{r,s}) =\displaystyle= {w,JC(σi)(w),JC(τj)(w),JC(τj)JC(σi)(w)\displaystyle\Big{\{}w,J_{C(\sigma_{i})}(w),J_{C(\tau_{j})}(w),J_{C(\tau_{j})}J_{C(\sigma_{i})}(w)\mid
C(σi),C(τj),C(σi)C(τj)Σ(𝒮2)},\displaystyle C(\sigma_{i}),C(\tau_{j}),C(\sigma_{i})C(\tau_{j})\in\Sigma(\mathcal{S}_{2})\Big{\}},

where 𝒮2𝒮(PI2)𝒮(C(PI1))\mathcal{S}_{2}\cong\mathcal{S}(PI_{2})\cong\mathcal{S}(C(PI_{1})) and the set PI2PI_{2} was replaced by C(PI1)C(PI_{1}). Note that since C(Br,s)=Br,sC(B_{r,s})=B_{r,s} we also have G(Br,s)=G(C(Br,s))G(B_{r,s})=G\big{(}C(B_{r,s})\big{)}.

We construct a correspondence A:v(Vr,s)w(Vr,s)A\colon\mathcal{B}_{v}(V^{r,s})\to\mathcal{B}_{w}(V^{r,s}) by

vw,Jσi(v)JC(σi)(w),Jτj(v)JC(τj)(w),\displaystyle v\longmapsto w,\quad J_{\sigma_{i}}(v)\longmapsto J_{C(\sigma_{i})}(w),\quad J_{\tau_{j}}(v)\longmapsto J_{C(\tau_{j})}(w),
Jτj(v)Jσi(v)JC(τj)(w)JC(σi)(w),\displaystyle J_{\tau_{j}}(v)J_{\sigma_{i}}(v)\longmapsto J_{C(\tau_{j})}(w)J_{C(\sigma_{i})}(w),

and C:zkC(zk)C\colon z_{k}\longmapsto C(z_{k}). The correspondence ACA\oplus C extended to a linear map over \mathbb{R} or \mathbb{Z} is an orthogonal map on Vr,sV^{r,s} since it maps orthonormal basis  (6.2) to orthonormal basis (6.3). To show that the linear map ACA\oplus C is an isomorphism of invariant integral structures, we argue as in Proposition 3.16. By the invariance of the bases v(Vr,s)\mathcal{B}_{v}(V^{r,s}) and w(Vr,s)\mathcal{B}_{w}(V^{r,s}) we have

JC(zk)Auα=±JC(κ)v2=±AJϰv1=AJzkuαJ_{C(z_{k})}Au_{\alpha}=\pm J_{C(\kappa)}v_{2}=\pm AJ_{\varkappa}v_{1}=AJ_{z_{k}}u_{\alpha}

for any uαv(Vr,s)u_{\alpha}\in\mathcal{B}_{v}(V^{r,s}), zkBr,sz_{k}\in B_{r,s}, and for some ϰΣ={σi,τj,τjσi}\varkappa\in\Sigma=\{\sigma_{i},\tau_{j},\tau_{j}\sigma_{i}\}. It implies

[Auα,Auβ],C(zk)r,s\displaystyle\langle[Au_{\alpha},Au_{\beta}],C(z_{k})\rangle_{r,s} =\displaystyle= JC(zk)Auα,AuβVr,s=AJzkuα,AuβVr,s\displaystyle\langle J_{C(z_{k})}Au_{\alpha},Au_{\beta}\rangle_{V^{r,s}}=\langle AJ_{z_{k}}u_{\alpha},Au_{\beta}\rangle_{V^{r,s}}
=\displaystyle= AτAJzkuα,uβVr,s=Jzkuα,uβVr,s\displaystyle\langle A^{\tau}AJ_{z_{k}}u_{\alpha},u_{\beta}\rangle_{V^{r,s}}=\langle J_{z_{k}}u_{\alpha},u_{\beta}\rangle_{V^{r,s}}
=\displaystyle= [uα,uβ],zkr,s.\displaystyle\langle[u_{\alpha},u_{\beta}],z_{k}\rangle_{r,s}.

for any uα,uβv(Vr,s)u_{\alpha},u_{\beta}\in\mathcal{B}_{v}(V^{r,s}) and zkBr,sz_{k}\in B_{r,s}. ∎

Theorem 6.3.

Let 𝒮1,𝒮2𝕊M\mathcal{S}_{1},\mathcal{S}_{2}\in\mathbb{S}^{M} and (𝒮1)\mathcal{L}(\mathcal{S}_{1}), (𝒮2)\mathcal{L}(\mathcal{S}_{2}) be the corresponding invariant integral structures. If there is an isomorphism

(6.4) AC:(𝒮1)(𝒮2)A\oplus C\colon\mathcal{L}(\mathcal{S}_{1})\to\mathcal{L}(\mathcal{S}_{2})

with A:Vr,sVr,sA\colon V^{r,s}\to V^{r,s} such that AτA=IdVr,sA^{\tau}A=\text{\rm Id}_{V^{r,s}}, then 𝒮1\mathcal{S}_{1} and 𝒮2\mathcal{S}_{2} are equivalent in the sense of Definition 4.3.

Proof.

Let

(𝒮1)=span{v(Vr,s)}span{Br,s}=L1span{Br,s}\mathcal{L}(\mathcal{S}_{1})=\text{\rm span}\,_{\mathbb{Z}}\{\mathcal{B}_{v}(V^{r,s})\}\oplus\text{\rm span}\,_{\mathbb{Z}}\{B_{r,s}\}=L_{1}\oplus\text{\rm span}\,_{\mathbb{Z}}\{B_{r,s}\}
(𝒮2)=span{u(Vr,s)}span{Br,s}=L2span{Br,s}\mathcal{L}(\mathcal{S}_{2})=\text{\rm span}\,_{\mathbb{Z}}\{\mathcal{B}_{u}(V^{r,s})\}\oplus\text{\rm span}\,_{\mathbb{Z}}\{B_{r,s}\}=L_{2}\oplus\text{\rm span}\,_{\mathbb{Z}}\{B_{r,s}\}

be the invariant integral srtuctures generated by the groups 𝒮1\mathcal{S}_{1} and 𝒮2\mathcal{S}_{2}. Here we also assume that 𝒮1=𝒮v\mathcal{S}_{1}=\mathcal{S}_{v} is the isotropy subgroup of a unit vector vVr,sv\in V^{r,s} and 𝒮2=𝒮u\mathcal{S}_{2}=\mathcal{S}_{u} is the isotropy subgroup of a unit vector uVr,su\in V^{r,s}. Since ACA\oplus C is an isomorphism, we obtain A(L1)=L2A(L_{1})=L_{2}. By noting that A1(L2)=Aτ(L2)=L1A^{-1}(L_{2})=A^{\tau}(L_{2})=L_{1}, we deduce that AτA(L1)=L1A^{\tau}A(L_{1})=L_{1}.

We denote by the same letter ACAut(𝔫r,s)A\oplus C\in\text{\rm Aut}(\mathfrak{n}_{r,s}) the automorphism of 𝔫r,s(Vr,s)\mathfrak{n}_{r,s}(V^{r,s}) which restriction to (𝒮1)\mathcal{L}(\mathcal{S}_{1}) gives map (6.4). The properties AτA=IdVr,sA^{\tau}A=\text{\rm Id}_{V^{r,s}} and AτJC(z)A=JzA^{\tau}J_{C(z)}A=J_{z} imply AJzx=JC(z)AxAJ_{z}x=J_{C(z)}Ax for xL1x\in L_{1} and CO(r,s)C\in\text{\rm O}(r,s), the latter one being an orthogonal transformation over \mathbb{Z} as well. For v𝔅v(Vr,s)v\in\mathfrak{B}_{v}(V^{r,s}) we find a basis vector uj𝔅u(Vr,s)u_{j}\in\mathfrak{B}_{u}(V^{r,s}) such that Av=ujAv=u_{j}. If there holds Av=ujAv=-u_{j}, then the proof is similar. By renumbering the basis vectors {uj}\{u_{j}\} we can assume that Av=uAv=u. We have for the stationary group of AvAv

(6.5) 𝒮Av\displaystyle\mathcal{S}_{Av} =\displaystyle= {σ~G(C(Br,s))Jσ~Av=Av}\displaystyle\{\tilde{\sigma}\in G\big{(}C(B_{r,s})\big{)}\mid\ J_{\tilde{\sigma}}Av=Av\}
=\displaystyle= {σ~G(C(Br,s))Jσ~u=u}=𝒮u\displaystyle\{\tilde{\sigma}\in G\big{(}C(B_{r,s})\big{)}\mid\ J_{\tilde{\sigma}}u=u\}=\mathcal{S}_{u}

Since σ~=C(zi1)C(zik)\tilde{\sigma}=C(z_{i_{1}})\ldots C(z_{i_{k}}), and AJzx=JC(z)AxAJ_{z}x=J_{C(z)}Ax, xL1x\in L_{1} we have

Av=Jσ~Av=JC(zi1)JC(zik)Av=AJzi1Jzikv=AJσv.Av=J_{\tilde{\sigma}}Av=J_{C(z_{i_{1}})}\ldots J_{C(z_{i_{k}})}Av=AJ_{z_{i_{1}}}\ldots J_{z_{i_{k}}}v=AJ_{\sigma}v.

This implies v=Jσvv=J_{\sigma}v for any σG(Br,s)\sigma\in G(B_{r,s}). Thus we conclude that if σ~𝒮Av\tilde{\sigma}\in\mathcal{S}_{Av}, for σ~=C(zi1)C(zik)G(C(Br,s))\tilde{\sigma}=C(z_{i_{1}})\ldots C(z_{i_{k}})\in G(C(B_{r,s})) then σ=zi1zik𝒮v\sigma=z_{i_{1}}\ldots z_{i_{k}}\in\mathcal{S}_{v}. Thus the groups 𝒮Av\mathcal{S}_{Av} and 𝒮v\mathcal{S}_{v} are equivalent. The equalities (6.5) shows that 𝒮2=𝒮u=𝒮Av\mathcal{S}_{2}=\mathcal{S}_{u}=\mathcal{S}_{Av} and 𝒮1=𝒮v\mathcal{S}_{1}=\mathcal{S}_{v} are equivalent. ∎

Table 17 shows the classical groups 𝔸\mathbb{A} such that the map AIdA\oplus\text{\rm Id} with A𝔸A\in\mathbb{A} is the automorphism of HH-type Lie algebras 𝔫r,s(Vr,s)\mathfrak{n}_{r,s}(V^{r,s}), see also [FM21, Table 3] for non-minimal admissible modules. The groups Sp(n),O(n,),U(n),O(n)\text{\rm Sp}(n),\text{\rm O}(n,\mathbb{C}),\text{\rm U}(n),\text{\rm O}^{*}(n) are subgroups of orthogonal transformations.

Table 17. Groups 𝔸\mathbb{A}
8 GL(1,)\text{\rm GL}(1,\mathbb{R})
7 O(1,)\text{\rm O}(1,\mathbb{R}) U(1)\text{\rm U}(1) Sp(1)\text{\rm Sp}(1) Sp(1)×Sp(1)\text{\rm Sp}(1)\times\text{\rm Sp}(1)
6 O(2,)\text{\rm O}(2,\mathbb{C}) O(2)\text{\rm O}^{*}(2) GL(1,)\text{\rm GL}(1,\mathbb{H}) Sp(1)\text{\rm Sp}(1)
5 O(4)\text{\rm O}^{*}(4) O(2)×O(2)\text{\rm O}^{*}(2)\times\text{\rm O}^{*}(2) O(2)\text{\rm O}^{*}(2) U(1)\text{\rm U}(1)
4 GL(1,)\text{\rm GL}(1,\mathbb{H}) O(2)\text{\rm O}^{*}(2) O(1,)\text{\rm O}(1,\mathbb{C}) O(1)\text{\rm O}(1\mathbb{R}) GL(1,)\text{\rm GL}(1,\mathbb{R})
3 Sp(1)\text{\rm Sp}(1) U(1)\text{\rm U}(1) O(1,)\text{\rm O}(1,\mathbb{R}) O(1,)×O(1,)\text{\rm O}(1,\mathbb{R})\times\text{\rm O}(1,\mathbb{R}) O(1)\text{\rm O}(1) U(1)\text{\rm U}(1) Sp(1)\text{\rm Sp}(1) Sp(1)×Sp(1)\text{\rm Sp}(1)\times\text{\rm Sp}(1)
2 Sp(2,)\text{\rm Sp}(2,\mathbb{C}) Sp(2,)\text{\rm Sp}(2,\mathbb{R}) GL(2,)\text{\rm GL}(2,\mathbb{R}) O(2)\text{\rm O}(2\mathbb{R}) O(2,)\text{\rm O}(2,\mathbb{C}) O(2)\text{\rm O}^{*}(2) GL(1,)\text{\rm GL}(1,\mathbb{H}) Sp(1)\text{\rm Sp}(1)
1 Sp(2,)\text{\rm Sp}(2,\mathbb{R}) Sp(2,)×Sp(2,)\text{\rm Sp}(2,\mathbb{R})\times\text{\rm Sp}(2,\mathbb{R}) Sp(4,)\text{\rm Sp}(4,\mathbb{R}) U(2)\text{\rm U}(2) O(4)\text{\rm O}^{*}(4) O(2)×O(2)\text{\rm O}^{*}(2)\times\text{\rm O}^{*}(2) O(1)\text{\rm O}^{*}(1) U(1)\text{\rm U}(1)
0 Sp(2,)\text{\rm Sp}(2,\mathbb{R}) Sp(2,)\text{\rm Sp}(2,\mathbb{C}) Sp(1)\text{\rm Sp}(1) GL(1)\text{\rm GL}(1\mathbb{H}) O(2)\text{\rm O}^{*}(2) O(1,)\text{\rm O}(1,\mathbb{C}) O(1,)\text{\rm O}(1,\mathbb{R}) GL(1,)\text{\rm GL}(1,\mathbb{R})
0 1 2 3 4 5 6 7 8
Theorem 6.4.

Let (r,s)(r,s) be such that the groups 𝔸\mathbb{A} in Table 17 is a subgroup of orthogonal transformations. The groups 𝒮1,𝒮2𝕊r,sM\mathcal{S}_{1},\mathcal{S}_{2}\in\mathbb{S}_{r,s}^{M} are equivalent in sense of Definition (4.4), if and only if the corresponding invariant integral structures (𝒮1)\mathcal{L}(\mathcal{S}_{1}) and (𝒮2)\mathcal{L}(\mathcal{S}_{2}) are isomorphic.

Proof.

If (r,s)(r,s) as in the statement of Theorem 6.4 then for an automorphism A~Id\tilde{A}\oplus\text{\rm Id} of 𝔫r,s(Vr,s)\mathfrak{n}_{r,s}(V^{r,s}) we have A~τA~=IdVr,s\tilde{A}^{\tau}\tilde{A}=\text{\rm Id}_{V^{r,s}}. It implies that the general automorphisms ACA\oplus C of 𝔫r,s(Vr,s)\mathfrak{n}_{r,s}(V^{r,s}) also satisfies AτA=IdVr,sA^{\tau}A=\text{\rm Id}_{V^{r,s}}, see [FM21, Section 3.2].

Thus if the invariant integral structures (𝒮1)\mathcal{L}(\mathcal{S}_{1}) and (𝒮2)\mathcal{L}(\mathcal{S}_{2}) are isomorphic, then they will be isomorphic under a map ACA\oplus C with AτA=IdVr,sA^{\tau}A=\text{\rm Id}_{V^{r,s}}. It implies that the group 𝒮1\mathcal{S}_{1} and 𝒮2\mathcal{S}_{2} are equivalent by Theorem 6.3.

Conversely, if we assume now that the groups 𝒮1\mathcal{S}_{1} and 𝒮2\mathcal{S}_{2} 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 𝒮1=𝒮(PI1)𝕊r,sM\mathcal{S}_{1}=\mathcal{S}(PI_{1})\in\mathbb{S}^{M}_{r,s} and 𝒮2=𝒮(PI2)𝕊M\mathcal{S}_{2}=\mathcal{S}(PI_{2})\in\mathbb{S}^{M} be non-equivalent groups such that there is a type T1T_{1} involution in pPI1p\in PI_{1} and an involution qPI2q\in PI_{2} such that pq=qpp\cdot q=-q\cdot p. Then the invariant integral structures (𝒮1)\mathcal{L}(\mathcal{S}_{1}) and (𝒮2)\mathcal{L}(\mathcal{S}_{2}) are not isomorphic.

Proof.

Let 𝔫r,s(Vr,s)\mathfrak{n}_{r,s}(V^{r,s}) be a pseudo HH-type Lie algebra and pPI1p\in PI_{1}, qPI2q\in PI_{2} as in the statement of Theorem 6.5. We denote by E(p)={xVr,sJpx=x}E(p)=\{x\in V^{r,s}\mid\ J_{p}x=x\} the eigenspace of type T1T_{1} involution pPI1p\in PI_{1} and by

E+(q)={xE(p)Jqx=x},E(q)={xE(p)Jqx=x}E_{+}(q)=\{x\in E(p)\mid\ J_{q}x=x\},\quad E_{-}(q)=\{x\in E(p)\mid\ J_{q}x=-x\}

the non-trivial eigen spaces of qPI2q\in PI_{2}. Then the subspaces in the direct sum E(p)=E+(q)E(q)E(p)=E_{+}(q)\oplus E_{-}(q) are orthogonal.

Let us assume that there exists an isomorphism AC:(𝒮1)(𝒮2)A\oplus C\colon\mathcal{L}(\mathcal{S}_{1})\to\mathcal{L}(\mathcal{S}_{2}) and write

F(p)=A(E(p)),F±=F±(C(q))={yF(p)JC(q)y=±y}.F(p)=A\big{(}E(p)\big{)},\quad F_{\pm}=F_{\pm}\big{(}C(q)\big{)}=\{y\in F(p)\mid\ J_{C(q)}y=\pm y\}.

Note the following: since AJp=JpAAJ_{p}=J_{p}A, we obtain that AτA(E(p))=E(p)A^{\tau}A\big{(}E(p)\big{)}=E(p). The map CC, extended to the Clifford algebra Clr,s\text{Cl}_{r,s}, satisfies C(p)C(q)=C(q)C(p)C(p)C(q)=-C(q)C(p). Therefore

(6.6) F(p)=F+F,F(p)=F_{+}\oplus F_{-},

where F+,FF_{+},F_{-} are non-trivial orthogonal vector spaces.

Let xE(p)x\in E(p) and put Ax=y+(x)+y(x)Ax=y_{+}(x)+y_{-}(x), where y+(x)F+y_{+}(x)\in F_{+} and y(x)Fy_{-}(x)\in F_{-}. We also have

Ax=AJpx=JC(p)Ax=JC(p)(y+(x)+y(x))=JC(p)y+(x)+JC(p)y(x).Ax=AJ_{p}x=J_{C(p)}Ax=J_{C(p)}(y_{+}(x)+y_{-}(x))=J_{C(p)}y_{+}(x)+J_{C(p)}y_{-}(x).

Since

JC(p):F+F,andJC(p)y+(x)F,JC(p)y(x)F+J_{C(p)}\colon F_{+}\to F_{-},\quad\text{and}\quad J_{C(p)}y_{+}(x)\in F_{-},\quad J_{C(p)}y_{-}(x)\in F_{+}

we obtain y+(x)=JC(p)y(x)y_{+}(x)=J_{C(p)}y_{-}(x) and y(x)=JC(p)y+(x)y_{-}(x)=J_{C(p)}y_{+}(x) by the uniqueness of the decomposition into a direct sum of vector spaces. We conclude

Ax=y+(x)+JC(p)y+(x).Ax=y_{+}(x)+J_{C(p)}y_{+}(x).

Let {vi}\{v_{i}\} be an orthonormal basis of the space E(p)E(p), which is a part of the invariant basis on Vr,sV^{r,s} defined by the 𝒮1\mathcal{S}_{1}. The matrix components aija_{ij} of the operator AτA:E(p)E(p)A^{\tau}A\colon E(p)\to E(p) with respect to the basis {vi}\{v_{i}\} have the form

aij\displaystyle a_{ij} =AτAvi,vjVr,s=Avi,AvjVr,s\displaystyle=\langle A^{\tau}Av_{i},v_{j}\rangle_{V^{r,s}}=\langle Av_{i},Av_{j}\rangle_{V^{r,s}}
=y+(vi)+JC(p)(y+(vi)),y+(vj)+JC(p)y+(vj)Vr,s\displaystyle=\langle y_{+}(v_{i})+J_{C(p)}(y_{+}(v_{i})),y_{+}(v_{j})+J_{C(p)}y_{+}(v_{j})\rangle_{V^{r,s}}
=2y+(vi),y+(vj)Vr,s,\displaystyle=2\langle y_{+}(v_{i}),y_{+}(v_{j})\rangle_{V^{r,s}},

where we used the orthogonality of the vector spaces F+F_{+} and FF_{-} in (6.6).

Hence the non-vanishing components of the matrix AτAA^{\tau}A are always even numbers, so that AA can not be invertible in SL(n,)SL(n,\mathbb{Z}). It implies that there are no an isomorphism ACA\oplus C between the invariant integral structures (𝒮1)\mathcal{L}(\mathcal{S}_{1}) and (𝒮2)\mathcal{L}(\mathcal{S}_{2}). ∎

Corollary 6.6.

Let 𝒮1\mathcal{S}_{1} and 𝒮2\mathcal{S}_{2} be in 𝕊r,sM\mathbb{S}^{M}_{r,s} and assume

  • (1)

    𝒮1=𝒮1(PI1)\mathcal{S}_{1}=\mathcal{S}_{1}(PI_{1}) and 𝒮2=𝒮2(PI2)\mathcal{S}_{2}=\mathcal{S}_{2}(PI_{2}) are not equivalent in the sense of the Definition 4.4,

  • (2)

    one of the sets PIkPI_{k}, k=1,2k=1,2 is of (T1)(T1)-type.

Then Theorem 6.5 holds.

Proof.

Since a generating set PI1PI_{1} of 𝒮1\mathcal{S}_{1} consists only of involutions of type T1T_{1}, the non-existence of an involution qPI2q\in PI_{2} such that pq=qppq=-qp for any pPI1p\in PI_{1} requires that PI1𝒮2PI_{1}\subset\mathcal{S}_{2} by the maximality of the groups 𝒮1\mathcal{S}_{1} and 𝒮2\mathcal{S}_{2}. But then 𝒮1=𝒮2\mathcal{S}_{1}=\mathcal{S}_{2} which is a contradiction. ∎

There are 3 pairs consisting of non-equivalent groups for (r,0)(r,0), which does not satisfies the conditions of Theorem 6.5 For r=12r=12 we have two non-equivalent groups 𝒮(5)\mathcal{S}^{(5)} and 𝒮(8)\mathcal{S}^{(8)} violating the conditions of Theorem 6.5, see Table 10. The generating set is presented here

PI(5)={p1=z1z2z3z4,p2=z1z2z5z6,p3=z1z3z5z7,p4=z8z9z10z11,ϰ1=z8z9z12},PI(8)={p1=z1z2z3z4,p2=z1z2z5z6,p3=z1z3z5z7,p4=z8z9z10z11,ϰ2=z1z2z7}\begin{array}[]{lllll}&PI^{(5)}=\{&p_{1}=z_{1}z_{2}z_{3}z_{4},p_{2}=z_{1}z_{2}z_{5}z_{6},p_{3}=z_{1}z_{3}z_{5}z_{7},p_{4}=z_{8}z_{9}z_{10}z_{11},\\ &&\varkappa_{1}=z_{8}z_{9}z_{12}\},\\ \\ &PI^{(8)}=\{&p_{1}=z_{1}z_{2}z_{3}z_{4},p_{2}=z_{1}z_{2}z_{5}z_{6},p_{3}=z_{1}z_{3}z_{5}z_{7},p_{4}=z_{8}z_{9}z_{10}z_{11},\\ &&\varkappa_{2}=z_{1}z_{2}z_{7}\}\end{array}

For r=13r=13 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 𝒮(k)\mathcal{S}^{(k)}, k=1,2k=1,2 which are all connected. The second collection contains the groups 𝒮(k)\mathcal{S}^{(k)}, k=5,6k=5,6 which are products of two smaller subgroups. The generating sets for the first collection are:

PI(1)={p1=z1z2z3z4,p2=z1z2z5z6,p3=z1z2z7z8,p4=z1z2z9z10,p5=z1z2z11z12,ρ1=z1z2z13},PI(2)={p1=z1z2z3z4,p2=z1z2z5z6,p3=z1z2z7z8,p4=z1z2z9z10,p5=z1z2z11z12,ρ2=z1z3z5z7z9z11z13},\begin{array}[]{lllll}&PI^{(1)}=\{&p_{1}=z_{1}z_{2}z_{3}z_{4},p_{2}=z_{1}z_{2}z_{5}z_{6},p_{3}=z_{1}z_{2}z_{7}z_{8},p_{4}=z_{1}z_{2}z_{9}z_{10},\\ &&p_{5}=z_{1}z_{2}z_{11}z_{12},\rho_{1}=z_{1}z_{2}z_{13}\},\\ \\ &PI^{(2)}=\{&p_{1}=z_{1}z_{2}z_{3}z_{4},p_{2}=z_{1}z_{2}z_{5}z_{6},p_{3}=z_{1}z_{2}z_{7}z_{8},p_{4}=z_{1}z_{2}z_{9}z_{10},\\ &&p_{5}=z_{1}z_{2}z_{11}z_{12},\rho_{2}=z_{1}z_{3}z_{5}z_{7}z_{9}z_{11}z_{13}\},\end{array}

The generating sets for the second collection are:

PI(5)={p1=z1z2z3z4,p2=z1z2z5z6,p3=z1z3z5z7,p4=z8z9z10z11,p5=z8z9z12z13,τ1=z1z2z7},PI(6)={p1=z1z2z3z4,p2=z1z2z5z6,p3=z1z3z5z7,p4=z8z9z10z11,p5=z8z9z12z13,τ2=z8z10z12},\begin{array}[]{lllll}&PI^{(5)}=\{&p_{1}=z_{1}z_{2}z_{3}z_{4},p_{2}=z_{1}z_{2}z_{5}z_{6},p_{3}=z_{1}z_{3}z_{5}z_{7},p_{4}=z_{8}z_{9}z_{10}z_{11},\\ &&p_{5}=z_{8}z_{9}z_{12}z_{13},\tau_{1}=z_{1}z_{2}z_{7}\},\\ \\ &PI^{(6)}=\{&p_{1}=z_{1}z_{2}z_{3}z_{4},p_{2}=z_{1}z_{2}z_{5}z_{6},p_{3}=z_{1}z_{3}z_{5}z_{7},p_{4}=z_{8}z_{9}z_{10}z_{11},\\ &&p_{5}=z_{8}z_{9}z_{12}z_{13},\tau_{2}=z_{8}z_{10}z_{12}\},\end{array}

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 EE invariant under the action of type T1T_{1} involutions. We start from r=13r=13 since the dimension of EE is equal to four and the calculations are more transparent.

Theorem 6.7.

Let r=13r=13. The invariant orthogonal lattices (𝒮(5))\mathcal{L}(\mathcal{S}^{(5)}) and (𝒮(6))\mathcal{L}(\mathcal{S}^{(6)}) defined by non-equivalent groups 𝒮(5)=𝒮(PI(5))\mathcal{S}^{(5)}=\mathcal{S}(PI^{(5)}) and 𝒮(6)=𝒮(PI(6))\mathcal{S}^{(6)}=\mathcal{S}(PI^{(6)}) are not isomorphic.

Proof.

The minimal admissible module V13,0V^{13,0} is isometric to 128,0\mathbb{R}^{128,0}. Let E={xV13,0Jpi(x)=x,i=1,2,3,4,5}E=\{x\in V^{13,0}\mid\ J_{p_{i}}(x)=x,\ i=1,2,3,4,5\} be the eigenspace of involutions of type T1T_{1}. Then dim(E)=4\dim(E)=4 and E=E+(τ1)E(τ1)E=E_{+}(\tau_{1})\oplus E_{-}(\tau_{1}), there E±(τ1)E_{\pm}(\tau_{1}) are the eigenspaces of τ1\tau_{1}. Let vE+(τ1)v\in E_{+}(\tau_{1}), v,vV13,0=1\langle v,v\rangle_{V^{13,0}}=1. The vectors

v1=v,v2=Jz8Jz9v,v3=Jz8Jz10Jz12v=Jτ2v,v4=Jz9Jz10Jz12v=Jz9Jτ2v,v_{1}=v,\ v_{2}=J_{z_{8}}J_{z_{9}}v,\ v_{3}=J_{z_{8}}J_{z_{10}}J_{z_{12}}v=J_{\tau_{2}}v,\ v_{4}=J_{z_{9}}J_{z_{10}}J_{z_{12}}v=J_{z_{9}}J_{\tau_{2}}v,

form an orthonormal basis of EE. In fact,

v,v1V13,0=v,Jz8Jz9vV13,0=z8,z913,0v,vV13,0=0,\langle v,v_{1}\rangle_{V^{13,0}}=\langle v,J_{z_{8}}J_{z_{9}}v\rangle_{V^{13,0}}=-\langle z_{8},z_{9}\rangle_{\mathbb{R}^{13,0}}\langle v,v\rangle_{V^{13,0}}=0,

and analogously v2,v3V13,0=0\langle v_{2},v_{3}\rangle_{V^{13,0}}=0. Furthermore, from one side

(6.7) v,v3V13,0=Jτ1v,Jτ2vV13,0=v,Jτ1Jτ2vV13,0=v,Jτ2Jτ1vV13,0,\langle v,v_{3}\rangle_{V^{13,0}}=\langle J_{\tau_{1}}v,J_{\tau_{2}}v\rangle_{V^{13,0}}=\langle v,J_{\tau_{1}}J_{\tau_{2}}v\rangle_{V^{13,0}}=-\langle v,J_{\tau_{2}}J_{\tau_{1}}v\rangle_{V^{13,0}},

But from other side

(6.8) v,v3V13,0=Jτ1v,Jτ2vV13,0=Jτ2Jτ1v,vV13,0.\langle v,v_{3}\rangle_{V^{13,0}}=\langle J_{\tau_{1}}v,J_{\tau_{2}}v\rangle_{V^{13,0}}=\langle J_{\tau_{2}}J_{\tau_{1}}v,v\rangle_{V^{13,0}}.

The equalities (6.7) and(6.8) imply the orthogonality of vv and v3v_{3}. 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 JzkJ_{z_{k}} with respect to product .,.V13,0\langle.\,,.\rangle_{V^{13,0}} and skew symmetry of the Clifford product JzkJzl=JzlJzkJ_{z_{k}}J_{z_{l}}=-J_{z_{l}}J_{z_{k}}.

Assume that there exists an isomorphism AC:𝔫13,0𝔫13,0A\oplus C\colon\mathfrak{n}_{13,0}\to\mathfrak{n}_{13,0} between the invariant orthogonal lattices (𝒮(5))\mathcal{L}(\mathcal{S}^{(5)}) to (𝒮(6))\mathcal{L}(\mathcal{S}^{(6)}).

We show that AA is an orthogonal transformation. In fact, we have

Av1,Av2V13,0=Av,JC(z8)JC(z9)AvV13,0=C(z8),C(z9)13,0Av,AvV13,0=0.\begin{split}\langle Av_{1},Av_{2}\rangle_{V^{13,0}}&=\langle Av,J_{C(z_{8})}J_{C(z_{9})}Av\rangle_{V^{13,0}}\\ &=\langle C(z_{8}),C(z_{9})\rangle_{\mathbb{R}^{13,0}}\langle Av,Av\rangle_{V^{13,0}}=0.\end{split}

Furthermore, by making use of the fact that the product Jτ2Jτ1J_{\tau_{2}}J_{\tau_{1}} contains 6 numbers of different JzkJ_{z_{k}}, we get

(6.9) Av1,Av3V13,0\displaystyle\langle Av_{1},Av_{3}\rangle_{V^{13,0}} =\displaystyle= Av,AJτ2vV13,0=Av,AJτ2Jτ1vV13,0\displaystyle\langle Av,AJ_{\tau_{2}}v\rangle_{V^{13,0}}=\langle Av,AJ_{\tau_{2}}J_{\tau_{1}}v\rangle_{V^{13,0}}
=\displaystyle= Av,JC(τ2)JC(τ1)AvV13,0=(1)11JC(τ2)JC(τ1)Av,AvV13,0.\displaystyle\langle Av,J_{C(\tau_{2})}J_{C(\tau_{1})}Av\rangle_{V^{13,0}}=(-1)^{11}\langle J_{C(\tau_{2})}J_{C(\tau_{1})}Av,Av\rangle_{V^{13,0}}.

In the last step we used the skew symmetry of JC(zk)J_{C(z_{k})} with respect to .,.V13,0\langle.\,,.\rangle_{V^{13,0}} and skew symmetry JC(zk)JC(zl)=JC(zl)JC(zk)J_{C(z_{k})}J_{C(z_{l})}=-J_{C(z_{l})}J_{C(z_{k})}. It shows Av1Av_{1} and Av3Av_{3} are orthogonal. Analogously we obtain Av1,Av4V13,0=0\langle Av_{1},Av_{4}\rangle_{V^{13,0}}=0.

Next we show

Av2,Av3V13,0\displaystyle\langle Av_{2},Av_{3}\rangle{V^{13,0}} =AJz8Jz9v,AJτ2vV13,0=JC(z8)JC(z9)Av,JC(τ2)JC(τ1)AvV13,0\displaystyle=\langle AJ_{z_{8}}J_{z_{9}}v,AJ_{\tau_{2}}v\rangle_{V^{13,0}}=\langle J_{C(z_{8})}J_{C(z_{9})}Av,J_{C(\tau_{2})}J_{C(\tau_{1})}Av\rangle_{V^{13,0}}
=Av,JC(z9)JC(z10)JC(z12)JC(τ1)AvV13,0\displaystyle=-\langle Av,J_{C(z_{9})}J_{C(z_{10})}J_{C(z_{12})}J_{C(\tau_{1})}Av\rangle_{V^{13,0}}
=(1)12Av,JC(z9)JC(z10)JC(z12)JC(τ1)AvV13,0=0,\displaystyle=(-1)^{12}\langle Av,J_{C(z_{9})}J_{C(z_{10})}J_{C(z_{12})}J_{C(\tau_{1})}Av\rangle_{V^{13,0}}=0,

by using the same arguments as in (6.9). The value Av2,Av4V13,0=0\langle Av_{2},Av_{4}\rangle_{V^{13,0}}=0 is shown in the same way.

Finally,

Av3,Av4V13,0\displaystyle\langle Av_{3},Av_{4}\rangle_{V^{13,0}} =\displaystyle= JC(τ2)JC(τ1)Av,JC(z9)JC(τ2)JC(τ1)Av,V13,0\displaystyle\langle J_{C(\tau_{2})}J_{C(\tau_{1})}Av,J_{C(z_{9})}J_{C(\tau_{2})}J_{C(\tau_{1})}Av,\rangle_{V^{13,0}}
=\displaystyle= JC(z8)Av,JC(z9)AvV13,0=0.\displaystyle\langle J_{C(z_{8})}Av,J_{C(z_{9})}Av\rangle_{V^{13,0}}=0.

This shows that AτA=λ=A(v)V13,0IdV13,0A^{\tau}A=\lambda=||A(v)||_{V^{13,0}}Id_{V^{13,0}} and then AτASL(4,)A^{\tau}A\in SL(4,\mathbb{Z}) requires A(v)V13,0=1||A(v)||_{V^{13,0}}=1. Hence by Theorem 6.3, the groups 𝒮(5)\mathcal{S}^{(5)} and 𝒮(6)\mathcal{S}^{(6)} are equivalent, that is a contradiction. ∎

Theorem 6.8.

Let r=13r=13. The invariant orthogonal lattices (𝒮(1))\mathcal{L}(\mathcal{S}^{(1)}) and (𝒮(2))\mathcal{L}(\mathcal{S}^{(2)}) defined by non-equivalent groups 𝒮(1)=𝒮(PI(1))\mathcal{S}^{(1)}=\mathcal{S}(PI^{(1)}) and 𝒮(2)=𝒮(PI(2))\mathcal{S}^{(2)}=\mathcal{S}(PI^{(2)}) are not isomorphic.

Proof.

As in Theorem 6.7 we define E={xV13,0Jpi(x)=x,i=1,2,3,4,5}E=\{x\in V^{13,0}\mid\ J_{p_{i}}(x)=x,\ i=1,2,3,4,5\} and E=E+(ρ1)E(ρ1)E=E_{+}(\rho_{1})\oplus E_{-}(\rho_{1}). Let vE+(ρ1)v\in E_{+}(\rho_{1}), v,vV13,0=1\langle v,v\rangle_{V^{13,0}}=1. Note that ρ1ρ2=ρ2ρ1\rho_{1}\rho_{2}=-\rho_{2}\rho_{1} and the product Jρ1Jρ2J_{\rho_{1}}J_{\rho_{2}} contains six different maps JzkJ_{z_{k}}. We show as in Theorem 6.7 that the vectors

v1=v,v2=Jz1Jz2v,v3=Jρ2v,v4=Jz2Jρ2v,v_{1}=v,\quad v_{2}=J_{z_{1}}J_{z_{2}}v,\quad v_{3}=J_{\rho_{2}}v,\quad v_{4}=J_{z_{2}}J_{\rho_{2}}v,

form an orthonormal basis of EE. Assuming that there is an isomorphism AC:𝔫13,0𝔫13,0A\oplus C\colon\mathfrak{n}_{13,0}\to\mathfrak{n}_{13,0} mapping the invariant orthogonal lattices (𝒮(1))\mathcal{L}(\mathcal{S}^{(1)}) to (𝒮(2))\mathcal{L}(\mathcal{S}^{(2)}) we show that AτA=IdV13,0A^{\tau}A=\text{\rm Id}_{V^{13,0}} and obtain the contradiction as in Theorem 6.7. ∎

Theorem 6.9.

Let r=12r=12. The invariant orthogonal lattices (𝒮(5))\mathcal{L}(\mathcal{S}^{(5)}) and (𝒮(8))\mathcal{L}(\mathcal{S}^{(8)}) defined by non-equivalent groups 𝒮(5)=𝒮(PI(5))\mathcal{S}^{(5)}=\mathcal{S}(PI^{(5)}) and 𝒮(8)=𝒮(PI(8))\mathcal{S}^{(8)}=\mathcal{S}(PI^{(8)}) are not isomorphic.

Proof.

The minimal admissible module V12,0V^{12,0} is isometric to 128,0\mathbb{R}^{128,0}. Let E={xV12,0Jpi(x)=x,i=1,2,3,4}E=\{x\in V^{12,0}\mid\ J_{p_{i}}(x)=x,\ i=1,2,3,4\} be the eigenspace of involutions of type T1T_{1}. Then dim(E)=8\dim(E)=8 and E=E+(ϰ2)E(ϰ2)E=E_{+}(\varkappa_{2})\oplus E_{-}(\varkappa_{2}), there E±(ϰ2)E_{\pm}(\varkappa_{2}) are the eigenspaces of Jϰ2=Jz1Jz2Jz7J_{\varkappa_{2}}=J_{z_{1}}J_{z_{2}}J_{z_{7}}. Let vE+(ϰ2)v\in E_{+}(\varkappa_{2}), v,vV12,0=1\langle v,v\rangle_{V^{12,0}}=1. The vectors

v1=v,v2=Jz8Jz9v=𝐈v,v3=Jz8Jz10v=𝐉v,v4=Jz9Jz10v=𝐊v,v5=Jϰ1v=Jz8Jz9Jz12v,v6=𝐈v5=Jz12v,v7=𝐉v5=Jz9Jz10Jz12v,v8=𝐊v5=Jz8Jz10Jz12v,\begin{array}[]{lllll}&v_{1}=v,&v_{2}=J_{z_{8}}J_{z_{9}}v=\mathbf{I}v,\\ &v_{3}=J_{z_{8}}J_{z_{10}}v=\mathbf{J}v,&v_{4}=J_{z_{9}}J_{z_{10}}v=\mathbf{K}v,\\ &v_{5}=J_{\varkappa_{1}}v=J_{z_{8}}J_{z_{9}}J_{z_{12}}v,&v_{6}=\mathbf{I}v_{5}=J_{z_{12}}v,\\ &v_{7}=\mathbf{J}v_{5}=-J_{z_{9}}J_{z_{10}}J_{z_{12}}v,&v_{8}=\mathbf{K}v_{5}=J_{z_{8}}J_{z_{10}}J_{z_{12}}v,\end{array}

form an orthonormal basis of EE by making of calculations as in (6.7) and (6.8). Note that the space EE is two dimensional quaternion space with the quaternion structure

𝐈=Jz8Jz9,𝐉=Jz8Jz10,𝐊=Jz9Jz10,𝐈2=𝐉2=𝐊2=𝐈𝐉𝐊=𝟏.\mathbf{I}=J_{z_{8}}J_{z_{9}},\quad\mathbf{J}=J_{z_{8}}J_{z_{10}},\quad\mathbf{K}=J_{z_{9}}J_{z_{10}},\quad\mathbf{I}^{2}=\mathbf{J}^{2}=\mathbf{K}^{2}=\mathbf{I}\mathbf{J}\mathbf{K}=-\mathbf{1}.

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 HH-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 𝐇\bf H-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-HH-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 22-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 HH-type groups. J. Lie Theory, 24(4):979–1011, 2014.
  • [FM17] Kenro Furutani and Irina Markina. Complete classification of pseudo HH-type Lie algebras: I. Geom. Dedicata, 190:23–51, 2017.
  • [FM19] Kenro Furutani and Irina Markina. Complete classification of pseudo HH-type algebras: II. Geom. Dedicata, 202:233–264, 2019.
  • [FM21] Kenro Furutani and Irina Markina. Automorphism groups of pseudo HH-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 HH-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 22-step nilpotent Lie algebras. Comm. Algebra, 27(5):2425–2440, 1999.
  • [M8́0] Guy Métivier. Hypoellipticité analytique sur des groupes nilpotents de rang 22. 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. HH-type groups and Clifford modules. Adv. Appl. Clifford Algebras, 11(S2):277–287, 2001.
  • [Rei01b] Hans Martin Reimann. Rigidity of HH-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.