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

Algebraic Groups Constructed From Rings with Involution

Arseniy (Senia) Sheydvasser Department of Mathematics, Graduate Center at CUNY, 365 5th Ave, New York, NY 10016 ssheydvasser@gc.cuny.edu
Abstract.

We define a class of groups constructed from rings equipped with an involution. We show that under suitable conditions, these groups are either algebraic or arithmetic, including as special cases the orientation-preserving isometry group of hyperbolic 4-space, SL(2,R)SL(2,R) for any commutative ring RR, various symplectic and orthogonal groups, and an important class of arithmetic subgroups of SO+(4,1)SO^{+}(4,1). We investigate when such groups are isomorphic and conjugate, and relate this to problem of determining when hyperbolic 44-orbifolds are homotopic.

Key words and phrases:
Algebraic groups, arithmetic groups, quaternion algebras, central simple algebras, involutions
2010 Mathematics Subject Classification:
Primary 20G15, 11R52, 11E04, 16H10

1. Introduction:

The idea of representing elements of Möb(n)\text{M\"{o}b}(\mathbb{R}^{n}) with 2×22\times 2 matrices with entries in a Clifford algebra goes back at least to Vahlen [Vah02], and was later popularized by Ahlfors [Ahl86]. More recently, this approach was used by the author to construct explicit examples of integral, crystallographic sphere packings [She19]; briefly, these are generalizations of the classical Apollonian gasket which arise from hyperbolic lattices. Such packings were formally defined by Kontorovich and Nakamura [KN19], although they were studied in various forms previously [GLM+05, GM10, Sta15]. How to define Möb()\text{M\"{o}b}(\mathbb{R}) and Möb(2)\text{M\"{o}b}(\mathbb{R}^{2}) in terms of the real and complex matrices is well-known. In order to describe Möb(3)\text{M\"{o}b}(\mathbb{R}^{3}) in terms of 2×22\times 2 matrices, we proceed as follows: let HH_{\mathbb{R}} be the standard Hamilton quaternions, and define an involution (x+yi+zj+tk)=x+yi+zjtk(x+yi+zj+tk)^{\ddagger}=x+yi+zj-tk. One can then define the set

SL(2,H)={(abcd)Mat(2,H)|ab=ba,cd=dc,adbc=1}.\displaystyle SL^{\ddagger}(2,H_{\mathbb{R}})=\left\{\begin{pmatrix}a&b\\ c&d\end{pmatrix}\in\text{Mat}(2,H_{\mathbb{R}})\middle|ab^{\ddagger}=ba^{\ddagger},\ cd^{\ddagger}=dc^{\ddagger},\ ad^{\ddagger}-bc^{\ddagger}=1\right\}.

One checks that this is a group, and that SL(2,H)/{±I} Möb(3)SL^{\ddagger}(2,H_{\mathbb{R}})/\{\pm I\}\cong\text{ M\"{o}b}(\mathbb{R}^{3})—or, if one prefers, since  Möb(3)SO+(4,1)\text{ M\"{o}b}(\mathbb{R}^{3})\cong SO^{+}(4,1), SL(2,H)Spin(4,1)SL^{\ddagger}(2,H_{\mathbb{R}})\cong\text{Spin}(4,1), the universal cover of SO+(4,1)SO^{+}(4,1). However, one observes that there is nothing in the definition of SL(2,H)SL^{\ddagger}(2,H_{\mathbb{R}}) that is specific to the quaternions: one can just as well choose any ring equipped with an involution σ\sigma and this will allow you to define a group SLσ(2,R)SL^{\sigma}(2,R). Our goal shall be describe what these groups are and under what circumstances they are isomorphic to one another. For example, if RR is an associative algebra over a field FF and σ:RR\sigma:R\rightarrow R is a morphism of affine FF-varieties, then this an algebraic group. In fact, our first major result will be the following.

Theorem 1.1.

Let FF be a field of characteristic not 22, AA a central simple algebra over FF of dimension n2n^{2}, and σ:AA\sigma:A\rightarrow A an FF-linear involution. Then SLσ(2,A)SL^{\sigma}(2,A) is a linear algebraic group over FF. Specifically, it is either a symplectic group of dimension 2n2+n2n^{2}+n, or an orthogonal group of dimension 2n2n2n^{2}-n.

Section 3 proves this result along with more detailed information, such as working out the Lie algebra of this algebraic group. We primarily consider the case where SLσ(2,A)SL^{\sigma}(2,A) is symplectic since in that case it is simply-connected and we can get very detailed information about when such groups are isomorphic. In Section 4, we restrict to looking at algebraic number fields, as then we can consider orders 𝒪\mathcal{O} of the central simple algebra AA. If 𝒪\mathcal{O} is closed under σ\sigma, then SLσ(2,𝒪)SL^{\sigma}(2,\mathcal{O}) is a well-defined group; in fact, it is an arithmetic subgroup of SLσ(2,𝒪)SL^{\sigma}(2,\mathcal{O}). Assuming SLσ(2,A)SL^{\sigma}(2,A) is symplectic, we can give a nice description of when such groups are isomorphic to one another in a sense that behaves well with respect to algebraic groups.

Theorem 1.2.

Let KK be an algebraic number field, AA a central simple algebra over KK, σ:AA\sigma:A\rightarrow A an FF-linear involution, and 𝒪1,𝒪2\mathcal{O}_{1},\mathcal{O}_{2} orders of AA closed under σ\sigma. If SLσ(2,A)SL^{\sigma}(2,A) is a symplectic group, then there exists a group isomorphism Ψ:SLσ(2,𝒪1)SLσ(2,𝒪2)\Psi:SL^{\sigma}(2,\mathcal{O}_{1})\rightarrow SL^{\sigma}(2,\mathcal{O}_{2}) which lifts to an automorphism of SLσ(2,A)SL^{\sigma}(2,A) if and only if there exists a ring isomorphism Φ:Mat(2,𝒪1)Mat(2,𝒪2)\Phi:\text{Mat}(2,\mathcal{O}_{1})\rightarrow\text{Mat}(2,\mathcal{O}_{2}) such that Φσ^=σ^Φ\Phi\circ\hat{\sigma}=\hat{\sigma}\circ\Phi, where

σ^:Mat(2,A)\displaystyle\hat{\sigma}:\text{Mat}(2,A) Mat(2,A)\displaystyle\rightarrow\text{Mat}(2,A)
(abcd)\displaystyle\begin{pmatrix}a&b\\ c&d\end{pmatrix} (σ(d)σ(b)σ(c)σ(a)).\displaystyle\mapsto\begin{pmatrix}\sigma(d)&-\sigma(b)\\ -\sigma(c)&\sigma(a)\end{pmatrix}.

If we restrict to the case of rational quaternion algebras, then we can remove the requirement that the group isomorphism Ψ:SLσ(2,𝒪1)SLσ(2,𝒪2)\Psi:SL^{\sigma}(2,\mathcal{O}_{1})\rightarrow SL^{\sigma}(2,\mathcal{O}_{2}) extends to an automorphism of SLσ(2,A)SL^{\sigma}(2,A), which is proved in Section 6. In this setting, SLσ(2,H)SL^{\sigma}(2,H) is symplectic group if and only if σ\sigma is FF-linear and is not quaternion conjugation—such involutions are called orthogonal and we shall always denote such an involution by \ddagger. We then have the following.

Theorem 1.3.

Let H1,H2H_{1},H_{2} be rational quaternion algebras with orthogonal involutions 1,2\ddagger_{1},\ddagger_{2}. Let 𝒪1,𝒪2\mathcal{O}_{1},\mathcal{O}_{2} be orders of H1,H2H_{1},H_{2} closed under 1,2\ddagger_{1},\ddagger_{2}, respectively. Then SL1(2,𝒪1)SL2(2,𝒪2)SL^{\ddagger_{1}}(2,\mathcal{O}_{1})\cong SL^{\ddagger_{2}}(2,\mathcal{O}_{2}) as a group if and only if there exists a ring isomorphism Φ:Mat(2,𝒪1)Mat(2,𝒪2)\Phi:\text{Mat}(2,\mathcal{O}_{1})\rightarrow\text{Mat}(2,\mathcal{O}_{2}) such that Φ^1=^2Φ\Phi\circ\hat{\ddagger}_{1}=\hat{\ddagger}_{2}\circ\Phi, where

^i:Mat(2,Hi)\displaystyle\hat{\ddagger}_{i}:\text{Mat}(2,H_{i}) Mat(2,Hi)\displaystyle\rightarrow\text{Mat}(2,H_{i})
(abcd)\displaystyle\begin{pmatrix}a&b\\ c&d\end{pmatrix} (dibiciai).\displaystyle\mapsto\begin{pmatrix}d^{\ddagger_{i}}&-b^{\ddagger_{i}}\\ -c^{\ddagger_{i}}&a^{\ddagger_{i}}\end{pmatrix}.

This is interesting from a geometric point of view due to an accidental isomorphism between the symplectic groups SL(2,H)SL^{\ddagger}(2,H) quotiented by {±1}\{\pm 1\} and orthogonal groups of indefinite, quinary quadratic forms, which we prove in Section 5.

Theorem 1.4.

Let FF be a characteristic 0 field. Then there is a bijection

{Isomorphism classes ofquaternion algebras over F}\displaystyle\left\{\begin{subarray}{c}\text{Isomorphism classes of}\\ \text{quaternion algebras over }F\end{subarray}\right\} {Isomorphism classes oforthogonal groups of indefinite,quinary quadratic forms over F}\displaystyle\rightarrow\left\{\begin{subarray}{c}\text{Isomorphism classes of}\\ \text{orthogonal groups of indefinite,}\\ \text{quinary quadratic forms over }F\end{subarray}\right\}
[H]\displaystyle[H] [SL(2,H)/{±1}].\displaystyle\mapsto\left[SL^{\ddagger}(2,H)/\{\pm 1\}\right].

In particular, this implies that if HH is a rational, definite quaternion algebra, \ddagger an orthogonal involution on HH, and 𝒪\mathcal{O} is a \ddagger-order, then SL(2,𝒪)/{±1}SL^{\ddagger}(2,\mathcal{O})/\{\pm 1\} is an arithmetic subgroup of SO(4,1)Isom(4)SO(4,1)\cong\text{Isom}(\mathbb{H}^{4}), the isometry group of hyperbolic 44-space—this makes SL(2,𝒪)/{±1}SL^{\ddagger}(2,\mathcal{O})/\{\pm 1\} a Kleinian group, and so have an immediate corollary to Theorem 1.3 as a consequence of Mostow rigidity.

Corollary 1.1.

Let H1,H2H_{1},H_{2} be rational quaternion algebras with orthogonal involutions 1,2\ddagger_{1},\ddagger_{2}. Let 𝒪1,𝒪2\mathcal{O}_{1},\mathcal{O}_{2} be orders of H1,H2H_{1},H_{2} closed under 1,2\ddagger_{1},\ddagger_{2}, respectively. For i=1,2i=1,2, let Γi=SLi(2,𝒪i)/{±1}\Gamma_{i}=SL^{\ddagger_{i}}(2,\mathcal{O}_{i})/\{\pm 1\}, which we think of as subgroups of Isom(4)\text{Isom}(\mathbb{H}^{4}). The following are equivalent.

  1. (1)

    The orbifolds 4/Γ1,4/Γ2\mathbb{H}^{4}/\Gamma_{1},\mathbb{H}^{4}/\Gamma_{2} are homotopic.

  2. (2)

    The orbifolds 4/Γ1,4/Γ2\mathbb{H}^{4}/\Gamma_{1},\mathbb{H}^{4}/\Gamma_{2} are isometric.

  3. (3)

    There exists a ring isomorphism Φ:Mat(2,𝒪1)Mat(2,𝒪2)\Phi:\text{Mat}(2,\mathcal{O}_{1})\rightarrow\text{Mat}(2,\mathcal{O}_{2}) such that Φ^1=^2Φ\Phi\circ\hat{\ddagger}_{1}=\hat{\ddagger}_{2}\circ\Phi, where ^i\hat{\ddagger}_{i} is defined as in Theorem 1.3.

This property of the groups SL(2,𝒪)SL^{\ddagger}(2,\mathcal{O}) is analogous to the Bianchi groups SL(2,𝔬K)SL(2,\mathfrak{o}_{K}), where KK is an imaginary quadratic field and 𝔬K\mathfrak{o}_{K} is its ring of integers. Recall that PSL(2,𝔬K)PSL(2,\mathfrak{o}_{K}) can be viewed as an arithmetic subgroup of Isom(3)\text{Isom}(\mathbb{H}^{3}), and the corresponding orbifolds 3/PSL(2,𝔬K)\mathbb{H}^{3}/PSL(2,\mathfrak{o}_{K}) are homotopic if and only if the rings of integers are isomorphic. The Bianchi groups have the additional property that they are in some sense maximal—SL(2,𝔬K)SL(2,\mathfrak{o}_{K}) is not contained inside any larger arithmetic subgroup of SL(2,K)SL(2,K). As it happens, our new groups also have this property, which shall be shown in Section 4.

Theorem 1.5.

Let KK be an algebraic number field, HH a quaternion algebra over KK, \ddagger an orthogonal involution on HH, and 𝒪\mathcal{O} an order of HH that is closed under \ddagger but is not contained inside any larger order closed under \ddagger. Then SL(2,𝒪)SL^{\ddagger}(2,\mathcal{O}) is a maximal arithmetic subgroup of SL(2,H)SL^{\ddagger}(2,H) in the sense that it is not contained inside any larger arithmetic subgroup of SL(2,H)SL^{\ddagger}(2,H).

In Section 6, we conclude with a variety of examples and counter-examples demonstrating that our results are in some sense sharp—for example, we show that in Theorem 1.3, one cannot replace Mat(2,𝒪)\text{Mat}(2,\mathcal{O}) with 𝒪\mathcal{O} instead, which is a major discrepancy from the comparatively simpler theory for commutative rings.

Acknowledgements:

The author would like to thank Joseph Quinn for a very productive conversation about invariants and isomorphisms of hyperbolic quotient manifolds, which inspired many of the approaches used in this paper, as well as Ara Basmajian and Abhijit Champanerkar, for asking pointed questions.

2. General Rings:

Let RR be a ring. An involution on RR is a map σ:RR\sigma:R\rightarrow R such that for all x,yRx,y\in R,

  1. (1)

    σ(σ(x))=x\sigma(\sigma(x))=x,

  2. (2)

    σ(x+y)=σ(x)+σ(y)\sigma(x+y)=\sigma(x)+\sigma(y), and

  3. (3)

    σ(xy)=σ(y)σ(x)\sigma(xy)=\sigma(y)\sigma(x).

Given two rings with involution (R1,σ1)(R_{1},\sigma_{1}), (R2,σ2)(R_{2},\sigma_{2}), a morphism between them is a ring homomorphism φ:R1R2\varphi:R_{1}\rightarrow R_{2} such that φ(σ1(x))=σ2(φ(x))\varphi(\sigma_{1}(x))=\sigma_{2}(\varphi(x)) for all xR1x\in R_{1}. There are countless standard examples of involutions; for instance, if FF is a field, we can consider M=Mat(n,F)M=\text{Mat}(n,F), the ring of n×nn\times n matrices with coefficients in FF, together with the adjugate map, usually written as \dagger. For our purposes, a particularly useful example is when n=2n=2—in that case, R=Mat(2,F)R=\text{Mat}(2,F) is a quaternion algebra over FF, and the adjugate is

(abcd)=(dbca),\displaystyle\begin{pmatrix}a&b\\ c&d\end{pmatrix}^{\dagger}=\begin{pmatrix}d&-b\\ -c&a\end{pmatrix},

which happens to be the standard involution, also known in this context as quaternion conjugation. More generally, for any commutative ring SS, R=Mat(2,S)R=\text{Mat}(2,S) equipped with the adjugate is a ring with involution. The same is not true if the base ring is not commutative, but there certainly exists a way to fix this in the n=2n=2 case, as follows.

Lemma 2.1.

Let (R,σ)(R,\sigma) be a ring with involution. Then Mat(2,R)\text{Mat}(2,R) together with the map

σ^:Mat(2,R)\displaystyle\hat{\sigma}:\text{Mat}(2,R) Mat(2,R)\displaystyle\rightarrow\text{Mat}(2,R)
(abcd)\displaystyle\begin{pmatrix}a&b\\ c&d\end{pmatrix} (σ(d)σ(b)σ(c)σ(a))\displaystyle\mapsto\begin{pmatrix}\sigma(d)&-\sigma(b)\\ -\sigma(c)&\sigma(a)\end{pmatrix}

is a ring with involution.

Proof.

It is easy to see that σ^\hat{\sigma} squares to the identity and that it preserves addition. It remains to prove that it reverses multiplication. This is a straightforward computation. On the one hand,

σ^((a1b1c1d1)(a2b2c2d2))\displaystyle\hat{\sigma}\left(\begin{pmatrix}a_{1}&b_{1}\\ c_{1}&d_{1}\end{pmatrix}\begin{pmatrix}a_{2}&b_{2}\\ c_{2}&d_{2}\end{pmatrix}\right) =σ^((a1a2+b1c2a1b2+b1d2c1a2+d1c2c1b2+d1d2))\displaystyle=\hat{\sigma}\left(\begin{pmatrix}a_{1}a_{2}+b_{1}c_{2}&a_{1}b_{2}+b_{1}d_{2}\\ c_{1}a_{2}+d_{1}c_{2}&c_{1}b_{2}+d_{1}d_{2}\end{pmatrix}\right)
=(σ(c1b2+d1d2)σ(a1b2+b1d2)σ(c1a2+d1c2)σ(a1a2+b1c2))\displaystyle=\begin{pmatrix}\sigma(c_{1}b_{2}+d_{1}d_{2})&-\sigma(a_{1}b_{2}+b_{1}d_{2})\\ -\sigma(c_{1}a_{2}+d_{1}c_{2})&\sigma(a_{1}a_{2}+b_{1}c_{2})\end{pmatrix}

On the other hand,

σ^((a2b2c2d2))\displaystyle\hat{\sigma}\left(\begin{pmatrix}a_{2}&b_{2}\\ c_{2}&d_{2}\end{pmatrix}\right) σ^((a1b1c1d1))\displaystyle\hat{\sigma}\left(\begin{pmatrix}a_{1}&b_{1}\\ c_{1}&d_{1}\end{pmatrix}\right)
=(σ(d2)σ(b2)σ(c2)σ(a2))(σ(d1)σ(b1)σ(c1)σ(a1))\displaystyle=\begin{pmatrix}\sigma\left(d_{2}\right)&-\sigma\left(b_{2}\right)\\ -\sigma\left(c_{2}\right)&\sigma\left(a_{2}\right)\end{pmatrix}\begin{pmatrix}\sigma\left(d_{1}\right)&-\sigma\left(b_{1}\right)\\ -\sigma\left(c_{1}\right)&\sigma\left(a_{1}\right)\end{pmatrix}
=(σ(d2)σ(d1)+σ(b2)σ(c1)σ(d2)σ(b1)σ(b2)σ(a1)σ(c2)σ(d1)σ(a2)σ(c1)σ(c2)σ(b1)+σ(a2)σ(a1)).\displaystyle=\begin{pmatrix}\sigma(d_{2})\sigma(d_{1})+\sigma(b_{2})\sigma(c_{1})&-\sigma(d_{2})\sigma(b_{1})-\sigma(b_{2})\sigma(a_{1})\\ -\sigma(c_{2})\sigma(d_{1})-\sigma(a_{2})\sigma(c_{1})&\sigma(c_{2})\sigma(b_{1})+\sigma(a_{2})\sigma(a_{1})\end{pmatrix}.

By the properties of σ\sigma, these are in fact the same. ∎

Remark 2.1.

As an aside, it is entirely possible that there is a nice generalization of this argument for n>2n>2—however, it is not clear to the author how to do this, since in general the adjugate is defined in terms of the minors of the matrix, whereas there are many different generalizations of the determinant for non-commutative rings. This is left as a question for the reader to ponder.

Remark 2.2.

A slightly more standard way to define an involution on Mat(n,R)\text{Mat}(n,R) is by composing the transpose map with element-wise application of σ:RR\sigma:R\rightarrow R—one checks that this works for any nn. (Consult [KMRT98] for more details and examples.) For n=2n=2, this is related to σ^\hat{\sigma} in the following way:

σ^((abcd))=(0110)(σ(a)σ(c)σ(b)σ(d))(0110)1.\displaystyle\hat{\sigma}\left(\begin{pmatrix}a&b\\ c&d\end{pmatrix}\right)=\begin{pmatrix}0&1\\ -1&0\end{pmatrix}\begin{pmatrix}\sigma(a)&\sigma(c)\\ \sigma(b)&\sigma(d)\end{pmatrix}\begin{pmatrix}0&1\\ -1&0\end{pmatrix}^{-1}.

In any case, this lemma allows us to make the following definition.

Definition 2.1.

Let (R,σ)(R,\sigma) be a ring with involution. By the special linear group on R2R^{2} twisted by σ\sigma, we shall mean the group

SLσ(2,R):={MMat(2,R)|Mσ^(M)=1}.\displaystyle SL^{\sigma}(2,R):=\left\{M\in\text{Mat}(2,R)\middle|M\hat{\sigma}(M)=1\right\}.

It is easy to see that this really is a group under matrix multiplication. The name of this group is motivated as follows. For convenience, define

R+:\displaystyle R^{+}: ={xR|σ(x)=x}\displaystyle=\left\{x\in R\middle|\sigma(x)=x\right\}
R:\displaystyle R^{-}: ={xR|σ(x)=x}.\displaystyle=\left\{x\in R\middle|\sigma(x)=-x\right\}.

Then we have the following lemma.

Lemma 2.2.

Let (R,σ)(R,\sigma) be a ring with involution. Then

SLσ(2,R)={(abcd)Mat(2,R)|aσ(b),cσ(d)R+,aσ(d)bσ(c)=1}.\displaystyle SL^{\sigma}(2,R)=\left\{\begin{pmatrix}a&b\\ c&d\end{pmatrix}\in\text{Mat}(2,R)\middle|a\sigma(b),c\sigma(d)\in R^{+},\ a\sigma(d)-b\sigma(c)=1\right\}.
Proof.

Choose any matrix

M=(abcd)Mat(2,R)\displaystyle M=\begin{pmatrix}a&b\\ c&d\end{pmatrix}\in\text{Mat}(2,R)

and compute

Mσ^(M)\displaystyle M\hat{\sigma}(M) =(abcd)(σ(d)σ(b)σ(c)σ(a))\displaystyle=\begin{pmatrix}a&b\\ c&d\end{pmatrix}\begin{pmatrix}\sigma(d)&-\sigma(b)\\ -\sigma(c)&\sigma(a)\end{pmatrix}
=(aσ(d)bσ(c)aσ(b)+bσ(a)cσ(d)dσ(c)cσ(b)+dσ(a)).\displaystyle=\begin{pmatrix}a\sigma(d)-b\sigma(c)&-a\sigma(b)+b\sigma(a)\\ c\sigma(d)-d\sigma(c)&-c\sigma(b)+d\sigma(a)\end{pmatrix}.

This matrix is equal to II if and only if

aσ(d)bσ(c)\displaystyle a\sigma(d)-b\sigma(c) =1\displaystyle=1
aσ(b)\displaystyle a\sigma(b) =σ(aσ(b))\displaystyle=\sigma\left(a\sigma(b)\right)
cσ(d)\displaystyle c\sigma(d) =σ(cσ(d)),\displaystyle=\sigma\left(c\sigma(d)\right),

which is to say that

M{(abcd)Mat(2,R)|aσ(b),cσ(d)R+,aσ(d)bσ(c)=1}.\displaystyle M\in\left\{\begin{pmatrix}a&b\\ c&d\end{pmatrix}\in\text{Mat}(2,R)\middle|a\sigma(b),c\sigma(d)\in R^{+},\ a\sigma(d)-b\sigma(c)=1\right\}.

Ergo, these two sets are in fact one and the same. ∎

It is now easy to see that Definition 2.1 has two important special cases:

  1. (1)

    If RR is commutative, then the identity map id:RRid:R\rightarrow R is an involution, and it is easy to see that

    SLid(2,R)=SL(2,R).\displaystyle SL^{id}(2,R)=SL(2,R).

    That is, if the twist is trivial, then we simply reduce to the ordinary special linear group.

  2. (2)

    If R=HR=H_{\mathbb{R}}, the Hamiltonian quaternions, and :HH\ddagger:H_{\mathbb{R}}\rightarrow H_{\mathbb{R}} is the map (x+wi+yj+zk)=x+wi+yjzk(x+wi+yj+zk)^{\ddagger}=x+wi+yj-zk, then SL(2,H)SL^{\ddagger}(2,H_{\mathbb{R}}) is exactly the group we saw in the introduction, and so

    SL(2,H)/{±I}SO+(3,1)Isom0(4),\displaystyle SL^{\ddagger}(2,H_{\mathbb{R}})/\{\pm I\}\cong SO^{+}(3,1)\cong\text{Isom}^{0}(\mathbb{H}^{4}),

    the orientation-preserving isometry group of hyperbolic 44-space.

3. Central Simple Algebras:

Generalizing the example of SL(2,H)SL^{\ddagger}(2,H_{\mathbb{R}}), one could take any field FF, a central simple algebra AA over FF, and an involution σ:AA\sigma:A\rightarrow A. This group is also known under another name: it is Isom(Mat(2,A),σ^)\text{Isom}(\text{Mat}(2,A),\hat{\sigma}), the collection of elements MMat(2,A)M\in\text{Mat}(2,A) with the property that Aσ^(A)=1A\hat{\sigma}(A)=1, which are known as isometries. (Consult [KMRT98] for more details.) It is easily checked that for any central simple algebra AA over a field FF, any involution σ\sigma preserves its center; consequently, the restriction of σ\sigma to FF is either the identity or an automorphism of order two. Involutions that fix FF are known as involutions of the first kind; the other type are involutions of the second kind. We will not consider involutions of the second kind here for a variety of reasons; ultimately, we will be interested in the case where F=F=\mathbb{Q}, in which case all involutions are of the first kind. It is easy to see that for any λF\lambda\in F, σ^(λ)=σ(λ)\hat{\sigma}(\lambda)=\sigma(\lambda), so σ^\hat{\sigma} is of the first kind if and only if σ\sigma itself is.

Involutions of the first kind are further split into symplectic and orthogonal involutions. These can defined as follows: any involution of the first kind σ:AA\sigma:A\rightarrow A can be extended to an involution on AFF¯A\otimes_{F}\overline{F}, where F¯\overline{F} is the algebraic closure of FF. However, by the Artin-Wedderburn theorem, AFF¯Mat(n,F¯)A\otimes_{F}\overline{F}\cong\text{Mat}(n,\overline{F}) for some nn. It readily checked that on such an algebra, any involution of the first kind σ\sigma corresponds to a bilinear form bσb_{\sigma} with the defining property that for all v,wF¯nv,w\in\overline{F}^{n} and all MMat(n,F¯)M\in\text{Mat}(n,\overline{F})

bσ(v,Mw)=bσ(σ(M)v,w).\displaystyle b_{\sigma}(v,Mw)=b_{\sigma}(\sigma(M)v,w).

This form is unique up to multiplication by scalars, and it is either symmetric or alternating. If it is symmetric, we say that σ\sigma is orthogonal; it is alternating, we say that it is symplectic.

Lemma 3.1.

Let FF be a field, AA a central simple algebra over FF, and σ:AA\sigma:A\rightarrow A be an involution of the first kind. If σ\sigma is symplectic, then σ^\hat{\sigma} is orthogonal, and vice versa.

Proof.

Let bσb_{\sigma} be a bilinear form on AFF¯A\otimes_{F}\overline{F} such that for all v,wF¯nv,w\in\overline{F}^{n} and all MMat(n,F¯)M\in\text{Mat}(n,\overline{F})

bσ(v,Mw)=bσ(σ(M)v,w).\displaystyle b_{\sigma}(v,Mw)=b_{\sigma}(\sigma(M)v,w).

We shall construct a corresponding bilinear form for σ^\hat{\sigma}. Specifically, note that Mat(2,AFF¯)Mat(2n,F¯)\text{Mat}(2,A\otimes_{F}\overline{F})\cong\text{Mat}(2n,\overline{F}), and so we shall want a bilinear form bσ^b_{\hat{\sigma}} on F¯2n=F¯n×F¯n\overline{F}^{2n}=\overline{F}^{n}\times\overline{F}^{n}. We do this by

bσ^((v1,v2),(w1,w2))\displaystyle b_{\hat{\sigma}}\left((v_{1},v_{2}),(w_{1},w_{2})\right) =bσ(v1,w2)bσ(v2,w1),\displaystyle=b_{\sigma}(v_{1},w_{2})-b_{\sigma}(v_{2},w_{1}),

in which case for any

γ=(M1M2M3M4)Mat(2,Mat(n,F¯))Mat(2n,F¯),\displaystyle\gamma=\begin{pmatrix}M_{1}&M_{2}\\ M_{3}&M_{4}\end{pmatrix}\in\text{Mat}(2,\text{Mat}(n,\overline{F}))\cong\text{Mat}(2n,\overline{F}),

we have

bσ^((v1,v2),γ(w1,w2))\displaystyle b_{\hat{\sigma}}\left((v_{1},v_{2}),\gamma(w_{1},w_{2})\right) =bσ^((v1,v2),(M1w1+M2w2,M3w1+M4w2))\displaystyle=b_{\hat{\sigma}}\left((v_{1},v_{2}),(M_{1}w_{1}+M_{2}w_{2},M_{3}w_{1}+M_{4}w_{2})\right)
=bσ(v1,M3w1+M4w2)bσ(v2,M1w1+M2w2)\displaystyle=b_{\sigma}(v_{1},M_{3}w_{1}+M_{4}w_{2})-b_{\sigma}(v_{2},M_{1}w_{1}+M_{2}w_{2})
=bσ(σ(M3)v1,w1)+bσ(σ(M4)v1,w2)\displaystyle=b_{\sigma}(\sigma(M_{3})v_{1},w_{1})+b_{\sigma}(\sigma(M_{4})v_{1},w_{2})
bσ(σ(M1)v2,w1)bσ(σ(M2)v2,w2)\displaystyle-b_{\sigma}(\sigma(M_{1})v_{2},w_{1})-b_{\sigma}(\sigma(M_{2})v_{2},w_{2})
=bσ(σ(M4)v1σ(M2)v2,w2)bσ(σ(M3)v1+σ(M1)v2,w1)\displaystyle=b_{\sigma}(\sigma(M_{4})v_{1}-\sigma(M_{2})v_{2},w_{2})-b_{\sigma}(-\sigma(M_{3})v_{1}+\sigma(M_{1})v_{2},w_{1})
=bσ^((σ(M4)v1σ(M2)v2,σ(M3)v1+σ(M1)v2),(w1,w2))\displaystyle=b_{\hat{\sigma}}\left((\sigma(M_{4})v_{1}-\sigma(M_{2})v_{2},-\sigma(M_{3})v_{1}+\sigma(M_{1})v_{2}),(w_{1},w_{2})\right)
=bσ^(σ^(γ)(v1,v2),(w1,w2)),\displaystyle=b_{\hat{\sigma}}\left(\hat{\sigma}(\gamma)(v_{1},v_{2}),(w_{1},w_{2})\right),

as desired. Note that if σ\sigma is orthogonal, then

bσ^((v1,v2),(v1,v2))\displaystyle b_{\hat{\sigma}}\left((v_{1},v_{2}),(v_{1},v_{2})\right) =bσ(v1,v2)bσ(v2,v1)=0,\displaystyle=b_{\sigma}(v_{1},v_{2})-b_{\sigma}(v_{2},v_{1})=0,

so the form is alternating and σ^\hat{\sigma} is symplectic. On the other hand, if σ\sigma is symplectic, then

bσ^((v1,v2),(w1,w2))\displaystyle b_{\hat{\sigma}}\left((v_{1},v_{2}),(w_{1},w_{2})\right) bσ^((w1,w2),(v1,v2))\displaystyle-b_{\hat{\sigma}}\left((w_{1},w_{2}),(v_{1},v_{2})\right)
=bσ(v1,w2)bσ(v2,w1)bσ(w1,v2)+bσ(w2,v1)\displaystyle=b_{\sigma}(v_{1},w_{2})-b_{\sigma}(v_{2},w_{1})-b_{\sigma}(w_{1},v_{2})+b_{\sigma}(w_{2},v_{1})
=bσ(v1+w2,v1+w2)bσ(v2+w1,v2+w1)=0,\displaystyle=b_{\sigma}(v_{1}+w_{2},v_{1}+w_{2})-b_{\sigma}(v_{2}+w_{1},v_{2}+w_{1})=0,

so the form is symmetric and σ^\hat{\sigma} is orthogonal. ∎

Remark 3.1.

Notice that because did not identify alternating and skew-symmetric forms, this proof applies perfectly well in characteristic 22, where those two types of forms are distinct.

Corollary 3.1.

Let FF be a field, AA a central simple algebra over FF, and σ:AA\sigma:A\rightarrow A be an involution of the first kind. If σ\sigma is orthogonal, then SLσ(2,A)SL^{\sigma}(2,A) is a symplectic group. Otherwise, it is an orthogonal group.

Proof.

As was shown in the proof of Lemma 3.1, SLσ(2,AFF¯)SL^{\sigma}(2,A\otimes_{F}\overline{F}) is the group of linear transformations that preserve a bilinear form bσ^b_{\hat{\sigma}} which is alternating if σ\sigma is orthogonal, and symmetric otherwise. In the first case, we have that SLσ(2,AFF¯)=Sp(bσ^)SL^{\sigma}(2,A\otimes_{F}\overline{F})=\text{Sp}(b_{\hat{\sigma}}), the symplectic group of bσ^b_{\hat{\sigma}}; in the second case, we have that SLσ(2,AFF¯)=O(bσ^)SL^{\sigma}(2,A\otimes_{F}\overline{F})=O(b_{\hat{\sigma}}), the orthogonal group of bσ^b_{\hat{\sigma}}. ∎

Since SLσ(2,A)SL^{\sigma}(2,A) is an algebraic group, we can work out its Lie algebra. In this context, a Lie algebra is a subspace of an associative FF-algebra that is closed under the bracket [x,y]=xyyx[x,y]=xy-yx. Any algebraic group GG has an associated Lie algebra consisting of its left invariant derivations—that is, if R=F[G]R=F[G] is the space of regular functions on GG, then it is the subspace

(G)\displaystyle\mathcal{L}(G) ={δDer(R)|δλx=λxδ,xG}\displaystyle=\left\{\delta\in\text{Der}(R)\middle|\delta\circ\lambda_{x}=\lambda_{x}\circ\delta,\ \forall x\in G\right\}

where Der(R)\text{Der}(R) is the set of derivations on RR and (λxf)(y)=f(x1y)(\lambda_{x}f)(y)=f(x^{-1}y) for any fRf\in R. Of course, the Lie algebras of the symplectic and orthogonal groups are known, but here we can give a nice description in terms of the central simple algebra AA.

Definition 3.1.

Let FF be a field, AA a central simple algebra over FF, σ:AA\sigma:A\rightarrow A an involution of the first kind. Then we define

𝔰𝔩σ(2,A)={XMat(2,A)|σ^(X)=X}.\displaystyle\mathfrak{sl}^{\sigma}(2,A)=\left\{X\in\text{Mat}(2,A)\middle|\hat{\sigma}(X)=-X\right\}.
Lemma 3.2.

𝔰𝔩σ(2,A)\mathfrak{sl}^{\sigma}(2,A) is a Lie algebra.

Proof.

Since σ\sigma is FF-linear, σ^\hat{\sigma} is FF-linear. Therefore 𝔰𝔩σ(2,A)\mathfrak{sl}^{\sigma}(2,A) is an FF-vector space. Checking that it is closed under the bracket is also easy, since

σ^(XYYX)\displaystyle\hat{\sigma}(XY-YX) =σ^(Y)σ^(X)σ^(X)σ^(Y)\displaystyle=\hat{\sigma}(Y)\hat{\sigma}(X)-\hat{\sigma}(X)\hat{\sigma}(Y)
=(XYYX),\displaystyle=-(XY-YX),

concluding the proof. ∎

Theorem 3.1.

Let FF be a field, AA a central simple algebra over FF, σ:AA\sigma:A\rightarrow A an involution of the first kind. Then 𝔰𝔩σ(2,A)\mathfrak{sl}^{\sigma}(2,A) is isomorphic to the Lie algebra of SLσ(2,A)SL^{\sigma}(2,A).

Proof.

The group SLσ(2,A)SL^{\sigma}(2,A) is a closed subgroup of GL(2,A)GL(2,A). The Lie algebra of GL(2,A)GL(2,A), 𝔤𝔩(2,A)\mathfrak{gl}(2,A), is isomorphic to Mat(2,A)\text{Mat}(2,A) via the map

Mat(2,A)\displaystyle\text{Mat}(2,A) 𝔤𝔩(2,A)\displaystyle\rightarrow\mathfrak{gl}(2,A)
M\displaystyle M (f(Xddtf(X(I+tM))|t=0)).\displaystyle\mapsto\left(f\mapsto\left(X\mapsto\left.\frac{d}{dt}f(X(I+tM))\right|_{t=0}\right)\right).

However, it is more convenient to use the fact that the Lie algebra can be identified with the tangent space at II—this we can define as the ideal of rational functions p:FGL(2,A)p:F\rightarrow GL(2,A) such that p(0)=Ip(0)=I, quotiented by the square of this ideal. In that case, the Lie algebra of SLσ(2,A)SL^{\sigma}(2,A) can be viewed as the sub-ideal consisting of all rational functions p:FSLσ(2,A)p:F\rightarrow SL^{\sigma}(2,A). By our isomorphism, this is equivalent to finding all MMat(2,A)M\in\text{Mat}(2,A) such that

(I+tM)σ^(I+tM)=I+O(t2).\displaystyle(I+tM)\hat{\sigma}(I+tM)=I+O(t^{2}).

However,

(I+tM)σ^(I+tM)\displaystyle(I+tM)\hat{\sigma}(I+tM) =(I+tM)(I+tσ^(M))\displaystyle=(I+tM)(I+t\hat{\sigma}(M))
=I+tM+tσ^(M)+t2Mσ^(M),\displaystyle=I+tM+t\hat{\sigma}(M)+t^{2}M\hat{\sigma}(M),

so this actually happens if and only if σ^(M)=M\hat{\sigma}(M)=-M. ∎

One benefit of having this explicit description of the Lie algebra is that it makes it very easy to see what the dimension of SLσ(2,A)SL^{\sigma}(2,A) is.

Corollary 3.2.

Let FF be a field, AA a central simple algebra over FF, σ:AA\sigma:A\rightarrow A an involution of the first kind. Then SLσ(2,A)SL^{\sigma}(2,A) is a linear algebraic group over FF of dimension

dimF(A)+2dimF(A+).\displaystyle\dim_{F}(A)+2\dim_{F}(A^{+}).
Proof.

The dimension of SLσ(2,A)SL^{\sigma}(2,A) is the dimension of 𝔰𝔩σ(2,A)\mathfrak{sl}^{\sigma}(2,A). But

(abcd)=σ^((abcd))=(σ(a)σ(b)σ(c)σ(d))\displaystyle\begin{pmatrix}a&b\\ c&d\end{pmatrix}=-\hat{\sigma}\left(\begin{pmatrix}a&b\\ c&d\end{pmatrix}\right)=\begin{pmatrix}-\sigma(a)&\sigma(b)\\ \sigma(c)&-\sigma(d)\end{pmatrix}

if and only if d=σ(a)d=-\sigma(a) and b,cA+b,c\in A^{+}. ∎

It is known that if the characteristic is not 22, then

dimF(A+)={n(n+1)2if n is orthogonaln(n1)2if n is symplectic,\displaystyle\dim_{F}(A^{+})=\begin{cases}\frac{n(n+1)}{2}&\text{if $n$ is orthogonal}\\ \frac{n(n-1)}{2}&\text{if $n$ is symplectic},\end{cases}

(see [KMRT98]), and so Theorem 1.1 is an immediate corollary of this and Corollary 3.1.

4. Orders and Arithmetic Groups:

If AA is a central simple algebra over an algebraic number field KK, then we can consider orders of AA—in this context, an order is a sub-ring 𝒪A\mathcal{O}\subset A that is finitely-generated as a 𝔬K\mathfrak{o}_{K}-module, where 𝔬K\mathfrak{o}_{K} is the ring of integers of KK, and such that K𝒪=AK\mathcal{O}=A. If σ:AA\sigma:A\rightarrow A is an involution and σ(𝒪)=𝒪\sigma(\mathcal{O})=\mathcal{O}, then we say that 𝒪\mathcal{O} is a σ\sigma-order of AA. If it is not contained inside any larger σ\sigma-order, then we say that it is a maximal σ\sigma-order. Such orders were studied by Scharlau [Sch74] in the 1970s and then generalized to Azumaya algebras by Saltmann [Sal78]. More recently, the author demonstrated how to classify the σ\sigma-orders of a quaternion algebra over local and global fields, which we will discuss in Section 6. In the meantime, we consider some of the broader theory.

If σ\sigma is an involution of the first kind on AA and 𝒪\mathcal{O} is a σ\sigma-order of AA, then it is easy to see that SLσ(2,𝒪)SL^{\sigma}(2,\mathcal{O}) is an arithmetic subgroup of SLσ(2,A)SL^{\sigma}(2,A). If σ\sigma is an orthogonal involution, then we can say significantly more; this is because SLσ(2,A)SL^{\sigma}(2,A) is a symplectic group, and it is therefore an almost simple, simply-connected group. Moreover, it is easy to see that if ν\nu is an infinite place of 𝔬K\mathfrak{o}_{K} then SLσ(2,Aν)SL^{\sigma}(2,A_{\nu}) is not a compact group since none of the symplectic groups over \mathbb{R} or \mathbb{C} are; here Aν=AKKνA_{\nu}=A\otimes_{K}K_{\nu}, where KνK_{\nu} is the localization of KK at the place ν\nu. Therefore, we can use the strong approximation theorem proved by Knesser and Platonov [Kne65, Pla69].

Theorem 4.1.

Let KK be an algebraic number field, AA a central simple algebra over KK, σ\sigma an orthogonal involution on AA, and SS be the set of infinite places of KK. Then SLσ(2,A)SL^{\sigma}(2,A) has strong approximation with respect to SS.

The version of strong approximation that we will need can be summarized as follows.

Corollary 4.1.

Let KK be an algebraic number field, AA a central simple algebra over KK, σ\sigma an orthogonal involution on AA, 𝒪\mathcal{O} a σ\sigma-order of AA, and SS be the set of infinite places of KK. Then SLσ(2,𝒪)SL^{\sigma}(2,\mathcal{O}) is dense inside

𝔭 primeSLσ(2,𝒪𝔭)\displaystyle\prod_{\mathfrak{p}\text{ prime}}SL^{\sigma}(2,\mathcal{O}_{\mathfrak{p}})

given the direct product topology, where 𝒪𝔭=𝒪𝔬K𝔬K,ν\mathcal{O}_{\mathfrak{p}}=\mathcal{O}\otimes_{\mathfrak{o}_{K}}\mathfrak{o}_{K,\nu}, 𝔬K\mathfrak{o}_{K} is the ring of integers of KK, and 𝔬K,ν\mathfrak{o}_{K,\nu} is the ring of integers of KνK_{\nu}.

The reason why this is helpful is that it will allow us to compute an important invariant, the matrix ring of the group.

Definition 4.1.

Let KK be an algebraic number field, AA a central simple algebra over KK, and σ\sigma an orthogonal involution on AA. Let ΓSLσ(2,A)\Gamma\subset SL^{\sigma}(2,A) be a subgroup. The matrix ring 𝔬K[Γ]\mathfrak{o}_{K}[\Gamma] of Γ\Gamma is the subring of Mat(2,A)\text{Mat}(2,A) generated by 𝔬K\mathfrak{o}_{K} and the elements of Γ\Gamma.

Remark 4.1.

One might ask in what sense this is an invariant. If Γ,Γ\Gamma,\Gamma^{\prime} are conjugate to one another in GL(2,AKK¯)GL(2,A\otimes_{K}\overline{K}), then it is clear that this extends to an isomorphism

𝔬K[Γ]\displaystyle\mathfrak{o}_{K}[\Gamma] 𝔬K[Γ]\displaystyle\rightarrow\mathfrak{o}_{K}[\Gamma^{\prime}]
M\displaystyle M γMγ1\displaystyle\mapsto\gamma M\gamma^{-1}

for some γGL(2,HKK¯)\gamma\in GL(2,H\otimes_{K}\overline{K}). Thus, in some sense any isomorphism of arithmetic groups leaves the matrix rings of the groups invariant. We shall see later that in special cases we can generalize this to group isomorphisms.

Lemma 4.1.

Let KK be an algebraic number field, AA a central simple algebra over KK, σ\sigma an orthogonal involution on AA, and 𝒪\mathcal{O} a σ\sigma-order of AA. Then 𝔬K[SLσ(2,𝒪)]=Mat(2,𝒪)\mathfrak{o}_{K}\left[SL^{\sigma}(2,\mathcal{O})\right]=\text{Mat}(2,\mathcal{O}).

Proof.

Obviously, 𝔬K[SLσ(2,𝒪)]\mathfrak{o}_{K}\left[SL^{\sigma}(2,\mathcal{O})\right] is contained inside of Mat(2,𝒪)\text{Mat}(2,\mathcal{O}), so it shall suffice to show that every element of Mat(2,𝒪)\text{Mat}(2,\mathcal{O}) is contained in 𝔬K[SLσ(2,𝒪)]\mathfrak{o}_{K}\left[SL^{\sigma}(2,\mathcal{O})\right]. Since 𝔬K𝒪\mathfrak{o}_{K}\subset\mathcal{O}, we know that Mat(2,𝔬K)𝔬K[SLσ(2,𝒪)]\text{Mat}(2,\mathfrak{o}_{K})\subset\mathfrak{o}_{K}\left[SL^{\sigma}(2,\mathcal{O})\right]. Now, let 𝔬^K\hat{\mathfrak{o}}_{K} be the inverse limit

𝔬^K=lim𝔞 ideal𝔬K/𝔞=𝔭 prime𝔬K,𝔭.\displaystyle\hat{\mathfrak{o}}_{K}=\varprojlim_{\mathfrak{a}\text{ ideal}}\mathfrak{o}_{K}/\mathfrak{a}=\prod_{\mathfrak{p}\text{ prime}}\mathfrak{o}_{K,\mathfrak{p}}.

This is the completion of 𝔬K\mathfrak{o}_{K} given the structure of a topological ring by specifying that the prime ideals 𝔭\mathfrak{p} form a basis of open sets for 0. Define

𝒪^=𝔭 prime𝒪𝔭=𝔭 prime𝒪𝔬K𝔬K,𝔭=𝒪𝔬K𝔬^K,\displaystyle\hat{\mathcal{O}}=\prod_{\mathfrak{p}\text{ prime}}\mathcal{O}_{\mathfrak{p}}=\prod_{\mathfrak{p}\text{ prime}}\mathcal{O}\otimes_{\mathfrak{o}_{K}}\mathfrak{o}_{K,\mathfrak{p}}=\mathcal{O}\otimes_{\mathfrak{o}_{K}}\hat{\mathfrak{o}}_{K},

giving it the structure of a topological ring via the product topology. Thinking of 𝒪^\hat{\mathcal{O}} as a group, we can take the quotient 𝒪^/𝔬^K\hat{\mathcal{O}}/\hat{\mathfrak{o}}_{K}, which inherits the quotient topology. All of this allows us to define the following continuous projection.

Ψ:𝔭 primeSL(2,𝒪𝔭)\displaystyle\Psi:\prod_{\mathfrak{p}\text{ prime}}SL^{\ddagger}(2,\mathcal{O}_{\mathfrak{p}}) 𝒪^/𝔬^K\displaystyle\rightarrow\hat{\mathcal{O}}/\hat{\mathfrak{o}}_{K}
(abcd)\displaystyle\begin{pmatrix}a&b\\ c&d\end{pmatrix} a+𝔬^K.\displaystyle\mapsto a+\hat{\mathfrak{o}}_{K}.

This projection is surjective—this is because for any z𝔭𝒪𝔭z_{\mathfrak{p}}\in\mathcal{O}_{\mathfrak{p}}, there exists some λ𝔭𝔬K,𝔭\lambda_{\mathfrak{p}}\in\mathfrak{o}_{K,\mathfrak{p}} such that λ𝔭+z𝔭𝒪𝔭×\lambda_{\mathfrak{p}}+z_{\mathfrak{p}}\in\mathcal{O}_{\mathfrak{p}}^{\times}. So, defining λ𝔬^K\lambda\in\hat{\mathfrak{o}}_{K}, z𝒪^z\in\hat{\mathcal{O}} such that their 𝔭\mathfrak{p}-th coordinates are λ𝔭\lambda_{\mathfrak{p}} and z𝔭z_{\mathfrak{p}} respectively, we have

(λ+z00(λ+σ(z))1)SLσ(2,𝒪^)z+𝔬^K.\displaystyle\underbrace{\begin{pmatrix}\lambda+z&0\\ 0&\left(\lambda+\sigma(z)\right)^{-1}\end{pmatrix}}_{\in SL^{\sigma}(2,\hat{\mathcal{O}})}\mapsto z+\hat{\mathfrak{o}}_{K}.

However, by Corollary 4.1, we know that SLσ(2,𝒪)SL^{\sigma}(2,\mathcal{O}) is dense, hence its image is also dense, which is to say that for any a𝒪/𝔬Ka\in\mathcal{O}/\mathfrak{o}_{K} and any prime ideal 𝔭\mathfrak{p}, there exists γSLσ(2,𝒪)\gamma\in SL^{\sigma}(2,\mathcal{O}) such that Ψ(γ)a𝔭𝒪\Psi(\gamma)-a\in\mathfrak{p}\mathcal{O}. Choose a basis {ei}i=1n\{e_{i}\}_{i=1}^{n} for 𝒪\mathcal{O} and an ideal 𝔞𝔬K\mathfrak{a}\subset\mathfrak{o}_{K} such that {ei+αi}i=1n\{e_{i}+\alpha_{i}\}_{i=1}^{n} is a basis for all αi𝔞𝒪\alpha_{i}\in\mathfrak{a}\mathcal{O}. Then there exist λi𝔬K\lambda_{i}\in\mathfrak{o}_{K} and αi𝔞𝒪\alpha_{i}\in\mathfrak{a}\mathcal{O} such that there exists γiSLσ(2,𝒪)\gamma_{i}\in SL^{\sigma}(2,\mathcal{O}) such Ψ(γi)=λi+ei+αi\Psi(\gamma_{i})=\lambda_{i}+e_{i}+\alpha_{i}, and therefore

(1000)γi(1000)(λi000)=(ei+αi000)𝔬K[SLσ(2,𝒪)].\displaystyle\begin{pmatrix}1&0\\ 0&0\end{pmatrix}\gamma_{i}\begin{pmatrix}1&0\\ 0&0\end{pmatrix}-\begin{pmatrix}\lambda_{i}&0\\ 0&0\end{pmatrix}=\begin{pmatrix}e_{i}+\alpha_{i}&0\\ 0&0\end{pmatrix}\in\mathfrak{o}_{K}\left[SL^{\sigma}(2,\mathcal{O})\right].

However, since {ei+αi}i=1n\{e_{i}+\alpha_{i}\}_{i=1}^{n} is a basis, it follows that

(𝒪000)𝔬K[SLσ(2,𝒪)].\displaystyle\begin{pmatrix}\mathcal{O}&0\\ 0&0\end{pmatrix}\subset\mathfrak{o}_{K}\left[SL^{\sigma}(2,\mathcal{O})\right].

By conjugating with elements in Mat(2,𝔬K)\text{Mat}(2,\mathfrak{o}_{K}), we get that Mat(2,𝒪)𝔬K[SLσ(2,𝒪)]\text{Mat}(2,\mathcal{O})\subset\mathfrak{o}_{K}\left[SL^{\sigma}(2,\mathcal{O})\right], as desired. ∎

More generally, for other arithmetic subgroups of SLσ(2,A)SL^{\sigma}(2,A), the matrix ring is an order of Mat(2,A)\text{Mat}(2,A).

Lemma 4.2.

Let KK be an algebraic number field, AA a central simple algebra over KK, σ\sigma an orthogonal involution on AA, and Γ\Gamma an arithmetic subgroup of SLσ(2,A)SL^{\sigma}(2,A). Then 𝔬K[Γ]\mathfrak{o}_{K}\left[\Gamma\right] is an order of the central simple algebra Mat(2,A)\text{Mat}(2,A).

Proof.

First, note that since ΓSLσ(2,A)\Gamma\subset SL^{\sigma}(2,A), 𝔬K[Γ]𝔬K[SLσ(2,A)]=Mat(2,A)\mathfrak{o}_{K}\left[\Gamma\right]\subset\mathfrak{o}_{K}\left[SL^{\sigma}(2,A)\right]=\text{Mat}(2,A). Since Γ\Gamma is an arithmetic group, for some integer ll, there is a morphism Ψ:ΓSL(l,𝔬K)\Psi:\Gamma\rightarrow SL(l,\mathfrak{o}_{K}) with a finite kernel. It is easy to see that 𝔬K[SL(l,𝔬K)]=Mat(l,𝔬K)\mathfrak{o}_{K}\left[SL(l,\mathfrak{o}_{K})\right]=\text{Mat}(l,\mathfrak{o}_{K}) is a finitely-generated, Noetherian 𝔬K\mathfrak{o}_{K}-module. Therefore, the sub-module 𝔬K[Ψ(Γ)]\mathfrak{o}_{K}\left[\Psi(\Gamma)\right] is finitely-generated. This, in turn, means that 𝔬K[Γ]\mathfrak{o}_{K}\left[\Gamma\right] is finitely-generated as an 𝔬K\mathfrak{o}_{K}-module. Since it is a subring of the finite-dimensional algebra Mat(2,A)\text{Mat}(2,A), it is therefore an order. ∎

We can use this invariant to characterize when arithmetic subgroups are conjugate to each other.

Theorem 4.2.

Let KK be an algebraic number field, AA a central simple algebra over KK, σ\sigma an orthogonal involution on AA. Let Γ1,Γ2\Gamma_{1},\Gamma_{2} be arithmetic subgroups of SLσ(2,A)SL^{\sigma}(2,A) such that Γi=𝔬K[Γi]SLσ(2,A)\Gamma_{i}=\mathfrak{o}_{K}[\Gamma_{i}]\cap SL^{\sigma}(2,A). The following are equivalent.

  1. (1)

    There exists an element γSLσ(2,AKK¯)\gamma\in SL^{\sigma}(2,A\otimes_{K}\overline{K}) such that Γ2=γΓ1γ1\Gamma_{2}=\gamma\Gamma_{1}\gamma^{-1}.

  2. (2)

    There exists a group isomorphism Ψ:Γ1Γ2\Psi:\Gamma_{1}\rightarrow\Gamma_{2} that extends to an automorphism of SLσ(2,A)SL^{\sigma}(2,A) as an algebraic group.

  3. (3)

    (𝔬K[Γ1],σ^)(𝔬K[Γ2],σ^)(\mathfrak{o}_{K}[\Gamma_{1}],\hat{\sigma})\cong(\mathfrak{o}_{K}[\Gamma_{2}],\hat{\sigma}).

Proof.

Since Γ1,Γ2\Gamma_{1},\Gamma_{2} are arithmetic groups, if Γ2=γΓ1γ1\Gamma_{2}=\gamma\Gamma_{1}\gamma^{-1} then this extends to an automorphism of SLσ(2,A)SL^{\sigma}(2,A); thus, the first condition certainly implies the second. On the other hand, any automorphism of SLσ(2,A)SL^{\sigma}(2,A) must come from an automorphism of the Lie algebra 𝔰𝔩σ(2,A)\mathfrak{sl}^{\sigma}(2,A), since the group is simply-connected. Since the Lie algebra is of type CnC_{n}, all of its automorphisms are inner, which is to say that they arise as conjugation by elements in SLσ(2,AKK¯)SL^{\sigma}(2,A\otimes_{K}\overline{K})—therefore, the second condition implies the first. Now, if Γ2=γΓ1γ1\Gamma_{2}=\gamma\Gamma_{1}\gamma^{-1}, then this extends to a ring isomorphism

𝔬K[Γ1]\displaystyle\mathfrak{o}_{K}[\Gamma_{1}] 𝔬K[Γ1]\displaystyle\rightarrow\mathfrak{o}_{K}[\Gamma_{1}]
M\displaystyle M γMγ1.\displaystyle\mapsto\gamma M\gamma^{-1}.

If γSLσ(2,AKK¯)\gamma\in SL^{\sigma}(2,A\otimes_{K}\overline{K}), then we see that

σ(γMγ1)\displaystyle\sigma\left(\gamma M\gamma^{-1}\right) =σ(γ1)σ(M)σ(γ)\displaystyle=\sigma(\gamma^{-1})\sigma(M)\sigma(\gamma)
=γσ(M)γ1,\displaystyle=\gamma\sigma(M)\gamma^{-1},

hence this is actually an isomorphism of rings with involution. On the other hand, since Γ1,Γ2\Gamma_{1},\Gamma_{2} are arithmetic, any isomorphism (𝔬K[Γ1],σ^)(𝔬K[Γ2],σ^)(\mathfrak{o}_{K}[\Gamma_{1}],\hat{\sigma})\cong(\mathfrak{o}_{K}[\Gamma_{2}],\hat{\sigma}) extends to an automorphism of (Mat(2,A),σ^)(\text{Mat}(2,A),\hat{\sigma}). Since Mat(2,A)\text{Mat}(2,A) is a central simple algebra, there exists γGL(2,A)\gamma\in GL(2,A) such that this automorphism has the form MγMγ1M\mapsto\gamma M\gamma^{-1}. Since this is an automorphism of rings with involution, it must be that

σ(γMγ1)=σ(γ1)σ(M)σ(γ)=γσ(M)γ1\displaystyle\sigma(\gamma M\gamma^{-1})=\sigma(\gamma^{-1})\sigma(M)\sigma(\gamma)=\gamma\sigma(M)\gamma^{-1}

for all MMat(2,A)M\in\text{Mat}(2,A), which implies that γσ(γ)\gamma\sigma(\gamma) is in the center of Mat(2,A)\text{Mat}(2,A), which is KK. Therefore, there exists λK¯\lambda\in\overline{K} such that γ=λγ\gamma^{\prime}=\lambda\gamma satisfies the property γσ(γ)=1\gamma^{\prime}\sigma(\gamma^{\prime})=1, which is to say that γSLσ(2,AKK¯)\gamma^{\prime}\in SL^{\sigma}(2,A\otimes_{K}\overline{K}). Therefore, we can take our desired isomorphism to be of the form

𝔬K[Γ1]\displaystyle\mathfrak{o}_{K}[\Gamma_{1}] 𝔬K[Γ1]\displaystyle\rightarrow\mathfrak{o}_{K}[\Gamma_{1}]
M\displaystyle M γMγ1.\displaystyle\mapsto\gamma^{\prime}M{\gamma^{\prime}}^{-1}.

Since 𝔬K[Γi]SLσ(2,A)=Γi\mathfrak{o}_{K}[\Gamma_{i}]\cap SL^{\sigma}(2,A)=\Gamma_{i}, this isomorphism restricts to a group isomorphism from Γ1\Gamma_{1} to Γ2\Gamma_{2} by looking at the subgroup of elements MM such that Mσ^(M)=1M\hat{\sigma}(M)=1. ∎

Remark 4.2.

Theorem 1.2 is an immediate corollary of this result and Lemma 4.1.

We shall see later in the context of quaternion algebras, that while (𝒪1,σ)(𝒪2,σ)(\mathcal{O}_{1},\sigma)\cong(\mathcal{O}_{2},\sigma) certainly implies (Mat(2,𝒪1),σ^)(Mat(2,𝒪2),σ^)(\text{Mat}(2,\mathcal{O}_{1}),\hat{\sigma})\cong(\text{Mat}(2,\mathcal{O}_{2}),\hat{\sigma}), the converse is false; in fact, there are examples where 𝒪1𝒪2\mathcal{O}_{1}\ncong\mathcal{O}_{2}, but (Mat(2,𝒪1),σ^)(Mat(2,𝒪2),σ^)(\text{Mat}(2,\mathcal{O}_{1}),\hat{\sigma})\cong(\text{Mat}(2,\mathcal{O}_{2}),\hat{\sigma}). Requiring (𝒪1,σ)(𝒪2,σ)(\mathcal{O}_{1},\sigma)\cong(\mathcal{O}_{2},\sigma) actually corresponds to a significantly more stringent notion of isomorphism.

Theorem 4.3.

Let KK be an algebraic number field, AA a central simple algebra over KK, σ\sigma an orthogonal involution on AA, and 𝒪1,𝒪2\mathcal{O}_{1},\mathcal{O}_{2} be σ\sigma-orders of AA. The following are equivalent.

  1. (1)

    There exists γSLσ(2,AKK¯)\gamma\in SL^{\sigma}(2,A\otimes_{K}\overline{K}) such that γSLσ(2,𝒪1)γ1=SLσ(2,𝒪2)\gamma SL^{\sigma}(2,\mathcal{O}_{1})\gamma^{-1}=SL^{\sigma}(2,\mathcal{O}_{2}) and γMγ1=M\gamma M\gamma^{-1}=M for all MSL(2,K)M\in SL(2,K).

  2. (2)

    (𝒪1,)(𝒪2,)(\mathcal{O}_{1},\ddagger)\cong(\mathcal{O}_{2},\ddagger).

Proof.

If (𝒪1,σ)(𝒪2,σ)(\mathcal{O}_{1},\sigma)\cong(\mathcal{O}_{2},\sigma), then this isomorphism extends to an automorphism of (A,σ)(A,\sigma). Since AA is a central simple algebra, this automorphism must take the form xuxu1x\mapsto uxu^{-1} for some uA×u\in A^{\times}. Furthermore, it must be that uσ(u)K×u\sigma(u)\in K^{\times}. Therefore, there exists some λK¯\lambda\in\overline{K} such that v=λuv=\lambda u satisfies vσ(v)=1v\sigma(v)=1 and

𝒪1\displaystyle\mathcal{O}_{1} 𝒪2\displaystyle\rightarrow\mathcal{O}_{2}
x\displaystyle x vxv1\displaystyle\mapsto vxv^{-1}

is the desired isomorphism of rings with involution. Now, we have that

γ=(v00σ(v)1)SLσ(2,AKK¯)\displaystyle\gamma=\begin{pmatrix}v&0\\ 0&\sigma(v)^{-1}\end{pmatrix}\in SL^{\sigma}(2,A\otimes_{K}\overline{K})

and therefore

SLσ(2,A)\displaystyle SL^{\sigma}(2,A) SLσ(2,A)\displaystyle\mapsto SL^{\sigma}(2,A)
M\displaystyle M γMγ1\displaystyle\mapsto\gamma M\gamma^{-1}

is an automorphism. However,

γ(abcd)γ1=(vav1vbv1vcv1vdv1),\displaystyle\gamma\begin{pmatrix}a&b\\ c&d\end{pmatrix}\gamma^{-1}=\begin{pmatrix}vav^{-1}&vbv^{-1}\\ vcv^{-1}&vdv^{-1}\end{pmatrix},

and therefore γSLσ(2,𝒪1)γ1=SLσ(2,𝒪2)\gamma SL^{\sigma}(2,\mathcal{O}_{1})\gamma^{-1}=SL^{\sigma}(2,\mathcal{O}_{2}) and this map acts as the identity on SL(2,K)SL(2,K). In the other direction, suppose that there exists an element γSLσ(2,HKK¯)\gamma\in SL^{\sigma}(2,H\otimes_{K}\overline{K}) such that γSLσ(2,𝒪1)γ1=SLσ(2,𝒪2)\gamma SL^{\sigma}(2,\mathcal{O}_{1})\gamma^{-1}=SL^{\sigma}(2,\mathcal{O}_{2}) and γMγ1=M\gamma M\gamma^{-1}=M for all MSL(2,K)M\in SL(2,K). Then, by Lemma 4.1, the conjugation map extends to a ring isomorphism

Mat(2,𝒪1)\displaystyle\text{Mat}(2,\mathcal{O}_{1}) Mat(2,𝒪2)\displaystyle\rightarrow\text{Mat}(2,\mathcal{O}_{2})
M\displaystyle M γMγ1.\displaystyle\mapsto\gamma M\gamma^{-1}.

This map restricts to the identity on Mat(2,𝔬K)\text{Mat}(2,\mathfrak{o}_{K}). This means that if we define subrings

Ui={MMat(2,𝒪i)|M(1101)=(1101)M},\displaystyle U_{i}=\left\{M\in\text{Mat}(2,\mathcal{O}_{i})\middle|M\begin{pmatrix}1&1\\ 0&1\end{pmatrix}=\begin{pmatrix}1&1\\ 0&1\end{pmatrix}M\right\},

then we are guaranteed that the ring isomorphism between the Mat(2,𝒪i)\text{Mat}(2,\mathcal{O}_{i}) restricts to a ring isomorphism between the UiU_{i}. However, it is easy to see that MUiM\in U_{i} if and only if it is of the form

(st0s)\displaystyle\begin{pmatrix}s&t\\ 0&s\end{pmatrix}

for some s,t𝒪is,t\in\mathcal{O}_{i}. Therefore, for any z𝒪1z\in\mathcal{O}_{1},

γ(1z01)γ1U2.\displaystyle\gamma\begin{pmatrix}1&z\\ 0&1\end{pmatrix}\gamma^{-1}\in U_{2}.

Write

γ=(abcd)SLσ(2,AKK¯),\displaystyle\gamma=\begin{pmatrix}a&b\\ c&d\end{pmatrix}\in SL^{\sigma}(2,A\otimes_{K}\overline{K}),

and note that

γ(0z00)γ1=(czσ(c)),\displaystyle\gamma\begin{pmatrix}0&z\\ 0&0\end{pmatrix}\gamma^{-1}=\begin{pmatrix}*&*\\ cz\sigma(c)&*\end{pmatrix},

hence czσ(c)=0cz\sigma(c)=0. Since czσ(c)=0cz\sigma(c)=0 for all z𝒪1z\in\mathcal{O}_{1}, it must be that czσ(c)=0cz\sigma(c)=0 for all zAKK¯z\in A\otimes_{K}\overline{K}. Since AKK¯Mat(n,K¯)A\otimes_{K}\overline{K}\cong\text{Mat}(n,\overline{K}) for some nn, we can think of c,zc,z as linear transformations. Suppose that there exists vK¯nv\in\overline{K}^{n} such that cv0cv\neq 0. Then there must also exist wK¯nw\in\overline{K}^{n} such that σ(c)w0\sigma(c)w\neq 0. Therefore, there exists zMat(n,K¯)z\in\text{Mat}(n,\overline{K}) such that zσ(c)w=vz\sigma(c)w=v, which means that czσ(v)w0cz\sigma(v)w\neq 0. This is contradicted by the fact that czσ(c)=0cz\sigma(c)=0 identically, which means that c=0c=0. By a similar argument with the sub-ring

Li={MMat(2,𝒪i)|M(1011)=(1011)M},\displaystyle L_{i}=\left\{M\in\text{Mat}(2,\mathcal{O}_{i})\middle|M\begin{pmatrix}1&0\\ 1&1\end{pmatrix}=\begin{pmatrix}1&0\\ 1&1\end{pmatrix}M\right\},

we can prove that b=0b=0 as well. Thus γ\gamma is a diagonal matrix, which is to say that

γ=(u00σ(u)1)\displaystyle\gamma=\begin{pmatrix}u&0\\ 0&\sigma(u)^{-1}\end{pmatrix}

for some u(AKK¯)×u\in\left(A\otimes_{K}\overline{K}\right)^{\times}. Thus 𝒪2=u𝒪1u1\mathcal{O}_{2}=u\mathcal{O}_{1}u^{-1}. Since

(u00σ(u)1)(1101)(u100σ(u))\displaystyle\begin{pmatrix}u&0\\ 0&\sigma(u)^{-1}\end{pmatrix}\begin{pmatrix}1&1\\ 0&1\end{pmatrix}\begin{pmatrix}u^{-1}&0\\ 0&\sigma(u)\end{pmatrix} =(1uσ(u)01)=(1101),\displaystyle=\begin{pmatrix}1&u\sigma(u)\\ 0&1\end{pmatrix}=\begin{pmatrix}1&1\\ 0&1\end{pmatrix},

we have that uσ(u)=1u\sigma(u)=1, which means that the map

Mat(2,𝒪1)\displaystyle\text{Mat}(2,\mathcal{O}_{1}) Mat(2,𝒪2)\displaystyle\rightarrow\text{Mat}(2,\mathcal{O}_{2})
M\displaystyle M uMu1\displaystyle\mapsto uMu^{-1}

is an isomorphism of rings with involution. ∎

We end this section by proving that if 𝒪\mathcal{O} happens to be a maximal σ\sigma-order, then SLσ(2,𝒪)SL^{\sigma}(2,\mathcal{O}) is also a maximal arithmetic group in some sense.

Theorem 4.4.

Let KK be an algebraic number field, AA a central simple algebra over KK, σ\sigma an orthogonal involution on AA, and 𝒪\mathcal{O} a maximal σ\sigma-order of AA. Then SLσ(2,𝒪)SL^{\sigma}(2,\mathcal{O}) is a maximal arithmetic subgroup of SLσ(2,A)SL^{\sigma}(2,A) in the sense that it is not contained inside any larger arithmetic subgroup of SLσ(2,A)SL^{\sigma}(2,A).

Proof.

Suppose that Γ\Gamma is an arithmetic group containing SLσ(2,𝒪)SL^{\sigma}(2,\mathcal{O}). By Lemma 4.2, we know that 𝔬K[Γ]\mathfrak{o}_{K}\left[\Gamma\right] is an order of Mat(2,A)\text{Mat}(2,A) which, by Lemma 4.1 contains Mat(2,𝒪)\text{Mat}(2,\mathcal{O}). Choose any element γΓ\gamma\in\Gamma, and choose any one of its coordinates xx. Since the matrix ring contains

(1000),(0100),(0010),(0001),\displaystyle\begin{pmatrix}1&0\\ 0&0\end{pmatrix},\begin{pmatrix}0&1\\ 0&0\end{pmatrix},\begin{pmatrix}0&0\\ 1&0\end{pmatrix},\begin{pmatrix}0&0\\ 0&1\end{pmatrix},

it is clear that 𝔬K[Γ]\mathfrak{o}_{K}\left[\Gamma\right] must contain Mat(2,𝒪[x])\text{Mat}(2,\mathcal{O}[x]). However, since

(abcd)1=(σ(d)σ(b)σ(c)σ(a)),\displaystyle\begin{pmatrix}a&b\\ c&d\end{pmatrix}^{-1}=\begin{pmatrix}\sigma(d)&-\sigma(b)\\ -\sigma(c)&\sigma(a)\end{pmatrix},

we see that it must actually contain Mat(2,𝒪[x,σ(x)])\text{Mat}(2,\mathcal{O}[x,\sigma(x)]). Since Mat(2,𝒪[x,σ(x)])\text{Mat}(2,\mathcal{O}[x,\sigma(x)]) is a subring of the matrix ring, it must also be an order, from which we get that 𝒪[x,σ(x)]\mathcal{O}[x,\sigma(x)] is an order. However, 𝒪[x,σ(x)]\mathcal{O}[x,\sigma(x)] is clearly closed under σ\sigma, hence it is a σ\sigma-order. Since 𝒪\mathcal{O} is a maximal σ\sigma-order, it follows that 𝒪[x,σ(x)]=𝒪\mathcal{O}[x,\sigma(x)]=\mathcal{O}. Therefore, ΓMat(2,𝒪)\Gamma\subset\text{Mat}(2,\mathcal{O}). However, SLσ(2,𝒪)=Mat(2,𝒪)SLσ(2,A)SL^{\sigma}(2,\mathcal{O})=\text{Mat}(2,\mathcal{O})\cap SL^{\sigma}(2,A), therefore Γ=SLσ(2,𝒪)\Gamma=SL^{\sigma}(2,\mathcal{O}). ∎

Remark 4.3.

Theorem 1.5 is nothing more than a special case of this result.

5. Quaternion Algebras:

We shall now explore the special case where A=HA=H is a quaternion algebra over a field FF—that is, a central simple algebra over FF such that every element has degree at most 22 over FF. For simplicity, we shall only consider the case where char(F)2\text{char}(F)\neq 2, in which case we can equivalently describe a quaternion algebra as an FF-algebra generated by two elements i,ji,j subject to the relations i2=a,j2=b,ij=jii^{2}=a,j^{2}=b,ij=-ji for some a,bF×a,b\in F^{\times}—we typically denote such an algebra by

(a,bF).\displaystyle\left(\frac{a,b}{F}\right).

It is easy to check that such an algebra is dimension 44—each element can be written in the form x+yi+zj+tijx+yi+zj+tij for some x,y,z,tFx,y,z,t\in F—and it has an involution x+yi+zj+tij¯=xyizjtij\overline{x+yi+zj+tij}=x-yi-zj-tij known as the standard involution or quaternion conjugation.The subspace on which the standard involution acts as the identity is just FF; the subspace on which it acts as multiplication by 1-1 is three-dimensional and will be denoted by H0H^{0}. It is common to define the (reduced) trace and (reduced) norm in terms of the standard involution as

tr(x)\displaystyle\text{tr}(x) =x+x¯\displaystyle=x+\overline{x}
nrm(x)\displaystyle\text{nrm}(x) =xx¯,\displaystyle=x\overline{x},

respectively. The standard involution is clearly an involution of the first kind. Any other involution of the first kind will be of the form

σ:H\displaystyle\sigma:H H\displaystyle\rightarrow H
x\displaystyle x ax¯a1\displaystyle\mapsto a\overline{x}a^{-1}

for some aH×H0a\in H^{\times}\cap H^{0}. One can check that quaternion conjugation is the unique symplectic involution on the quaternion algebra, whereas all the other involutions are orthogonal. In fact, one can show that for any orthogonal involution, if one correctly chooses a basis 1,i,j,ij1,i,j,ij for HH, then the involution will be of the form

(x+yi+zj+tij)\displaystyle(x+yi+zj+tij)^{\ddagger} =x+yi+zjtij\displaystyle=x+yi+zj-tij

Clearly, any such involution will act as the identity on a subspace of dimension 33, which we shall denote by H+H^{+}, and act as multiplication by 1-1 on a subspace of dimension 11, which we shall denote by HH^{-}.

Remark 5.1.

An objection might be raised to the notation \ddagger to denote an orthogonal involution. The motivation for this notation is simple: Mat(2,F)\text{Mat}(2,F) is a quaternion algebra and one can check that in that case

(abcd)¯=(dbca).\displaystyle\overline{\begin{pmatrix}a&b\\ c&d\end{pmatrix}}=\begin{pmatrix}d&-b\\ -c&a\end{pmatrix}.

This is the adjugate, which is usually denoted by \dagger. Since orthogonal involutions are related to the standard involution but are nevertheless distinct, we choose the notation \ddagger to represent them.

Since orthogonal involutions act as the identity on a space of dimension 11, we can define their discriminant as follows:

disc:{orthogonal involutions}\displaystyle\text{disc}:\left\{\text{orthogonal involutions}\right\} F×/(F×)2\displaystyle\rightarrow F^{\times}/\left(F^{\times}\right)^{2}
\displaystyle\ddagger x2(F×)2,\displaystyle\mapsto x^{2}\left(F^{\times}\right)^{2},

where xx is any element of H×HH^{\times}\cap H^{-}. The discriminant uniquely determines the involution in the sense that if 1,2\ddagger_{1},\ddagger_{2} are orthogonal involutions on HH then there exists an isomorphism of rings with involutions (H,1)(H,2)(H,\ddagger_{1})\cong(H,\ddagger_{2}) if and only if disc(1)=disc(2)\text{disc}(\ddagger_{1})=\text{disc}(\ddagger_{2}). In general, a quaternion algebra admits many inequivalent orthogonal involutions; however, the linear algebraic groups SL(2,H)SL^{\ddagger}(2,H) that arise as a result are all essentially the same.

Lemma 5.1.

Let FF be a field of characteristic not 22. Let HH be a quaternion algebra over FF, and let 1,2\ddagger_{1},\ddagger_{2} be orthogonal involutions on HH. Then SL1(2,H)SL^{\ddagger_{1}}(2,H) and SL2(2,H)SL^{\ddagger_{2}}(2,H) are conjugate inside GL(2,HFF¯)GL(2,H\otimes_{F}\overline{F}), where F¯\overline{F} is the algebraic closure of FF.

Proof.

Both 1\ddagger_{1} and 2\ddagger_{2} can be extended to orthogonal involution on HFF¯H\otimes_{F}\overline{F}—however, since every element is a square in F¯\overline{F}, it follows that (HFF¯,1)(HF(¯F),2)(H\otimes_{F}\overline{F},\ddagger_{1})\cong(H\otimes_{F}\overline{(}F),\ddagger_{2}). Since HFF¯H\otimes_{F}\overline{F} is a central simple algebra, by the Skolem-Noether theorem there must exist uHFF¯u\in H\otimes_{F}\overline{F} such that the desired isomorphism of the form xuxu1x\mapsto uxu^{-1} for all xHFF¯x\in H\otimes_{F}\overline{F}. Now, choose any

γ=(abcd)SL1(2,H)\displaystyle\gamma=\begin{pmatrix}a&b\\ c&d\end{pmatrix}\in SL^{\ddagger_{1}}(2,H)

and note that

((u00u)γ(u00u)1)^2\displaystyle\left(\begin{pmatrix}u&0\\ 0&u\end{pmatrix}\gamma\begin{pmatrix}u&0\\ 0&u\end{pmatrix}^{-1}\right)^{\hat{\ddagger}_{2}} =(uau1ubu1ucu1udu1)2\displaystyle=\begin{pmatrix}uau^{-1}&ubu^{-1}\\ ucu^{-1}&udu^{-1}\end{pmatrix}^{\ddagger_{2}}
=((udu1)2(ubu1)2(ucu1)2(uau1)2)\displaystyle=\begin{pmatrix}\left(udu^{-1}\right)^{\ddagger_{2}}&-\left(ubu^{-1}\right)^{\ddagger_{2}}\\ -\left(ucu^{-1}\right)^{\ddagger_{2}}&\left(uau^{-1}\right)^{\ddagger_{2}}\end{pmatrix}
=(ud1u1ub1u1uc1u1ua1u1)\displaystyle=\begin{pmatrix}ud^{\ddagger_{1}}u^{-1}&-ub^{\ddagger_{1}}u^{-1}\\ -uc^{\ddagger_{1}}u^{-1}&ua^{\ddagger_{1}}u^{-1}\end{pmatrix}
=(u00u)γ^1(u00u)1\displaystyle=\begin{pmatrix}u&0\\ 0&u\end{pmatrix}\gamma^{\hat{\ddagger}_{1}}\begin{pmatrix}u&0\\ 0&u\end{pmatrix}^{-1}
=(u00u)γ1(u00u)1,\displaystyle=\begin{pmatrix}u&0\\ 0&u\end{pmatrix}\gamma^{-1}\begin{pmatrix}u&0\\ 0&u\end{pmatrix}^{-1},

from which we conclude that we have constructed a map

SL1(2,H)\displaystyle SL^{\ddagger_{1}}(2,H) SL2(2,H)\displaystyle\rightarrow SL^{\ddagger_{2}}(2,H)
γ\displaystyle\gamma (u00u)γ(u00u)1.\displaystyle\mapsto\begin{pmatrix}u&0\\ 0&u\end{pmatrix}\gamma\begin{pmatrix}u&0\\ 0&u\end{pmatrix}^{-1}.

It is easy to see that this map is the desired isomorphism. ∎

Remark 5.2.

Since these groups are conjugate inside GL(2,HFF¯)GL(2,H\otimes_{F}\overline{F}), they are isomorphic as algebraic groups. In contrast, purely by dimensional considerations, we can see that SL(2,H)SL^{\dagger}(2,H) is not isomorphic to these algebraic groups. Furthermore, if we restrict from HH to a sub-ring, we shall again find that these groups are not isomorphic in general.

We know that if \ddagger is an orthogonal involution, then SL(2,H)SL^{\ddagger}(2,H) is a symplectic group. However, in low dimensions, there is an accidental isomorphism between symplectic groups and spin groups. With our machinery, we can work out this isomorphism very explicitly.

Theorem 5.1.

Let HH be a quaternion algebra over a field FF not characteristic 22, with orthogonal involution \ddagger. Define a quadratic form qHq_{H} on F2H+F^{2}\oplus H^{+} by

qH(s,t,z)=stnrm(z).\displaystyle q_{H}(s,t,z)=st-\text{nrm}(z).

Then there is an exact sequence of algebraic groups

1{±1}SL(2,H)O0(qH)1,\displaystyle 1\rightarrow\left\{\pm 1\right\}\rightarrow SL^{\ddagger}(2,H)\rightarrow O^{0}(q_{H})\rightarrow 1,

where O0(qH)O^{0}(q_{H}) is the connected component of the orthogonal group of qHq_{H}.

Remark 5.3.

The special case where K=K=\mathbb{Q} and HH is positive definite was worked out in [She19]. We follow mostly the same argument.

Proof.

Define a set

H={M=(abcd)Mat(2,H)|M¯T=M,ab,cdH+}.\displaystyle\mathcal{M}_{H}=\left\{M=\begin{pmatrix}a&b\\ c&d\end{pmatrix}\in\text{Mat}(2,H)\middle|\overline{M}^{T}=M,\ ab^{\ddagger},cd^{\ddagger}\in H^{+}\right\}.

It is easy to see that there is a bijective linear map

F2H+\displaystyle F^{2}\oplus H^{+} H\displaystyle\rightarrow\mathcal{M}_{H}
(s,t,z)\displaystyle(s,t,z) (szz¯t),\displaystyle\mapsto\begin{pmatrix}s&z\\ \overline{z}&t\end{pmatrix},

taking the quadratic norm to the quasi-determinant stzz¯=stnrm(z)st^{\ddagger}-z\overline{z}^{\ddagger}=st-\text{nrm}(z)—thus, we can identify these two sets. On the other hand, if γSL(2,H)\gamma\in SL^{\ddagger}(2,H) and MHM\in\mathcal{M}_{H}, then it is easy to check that γMγ¯TH\gamma M\overline{\gamma}^{T}\in\mathcal{M}_{H} as well, and has the same quasi-determinant as γ\gamma. Therefore, we have defined a morphism of algebraic groups SL(2,H)O(qH)SL^{\ddagger}(2,H)\rightarrow O(q_{H}). For any element of the kernel,

(1000)\displaystyle\begin{pmatrix}1&0\\ 0&0\end{pmatrix} =(abcd)(1000)(a¯c¯b¯d¯)\displaystyle=\begin{pmatrix}a&b\\ c&d\end{pmatrix}\begin{pmatrix}1&0\\ 0&0\end{pmatrix}\begin{pmatrix}\overline{a}&\overline{c}\\ \overline{b}&\overline{d}\end{pmatrix}
=(a0c0)(a¯c¯b¯d¯)\displaystyle=\begin{pmatrix}a&0\\ c&0\end{pmatrix}\begin{pmatrix}\overline{a}&\overline{c}\\ \overline{b}&\overline{d}\end{pmatrix}
=(nrm(a)ac¯ca¯nrm(c)),\displaystyle=\begin{pmatrix}\text{nrm}(a)&a\overline{c}\\ c\overline{a}&\text{nrm}(c)\end{pmatrix},

from which we conclude that c=0c=0 and nrm(a)=1\text{nrm}(a)=1. Similarly, the relation

(0001)\displaystyle\begin{pmatrix}0&0\\ 0&1\end{pmatrix} =(ab0d)(0001)(a¯0b¯d¯)\displaystyle=\begin{pmatrix}a&b\\ 0&d\end{pmatrix}\begin{pmatrix}0&0\\ 0&1\end{pmatrix}\begin{pmatrix}\overline{a}&0\\ \overline{b}&\overline{d}\end{pmatrix}
=(0b0d)(a¯0b¯d¯)\displaystyle=\begin{pmatrix}0&b\\ 0&d\end{pmatrix}\begin{pmatrix}\overline{a}&0\\ \overline{b}&\overline{d}\end{pmatrix}
=(nrm(b)bd¯db¯nrm(d))\displaystyle=\begin{pmatrix}\text{nrm}(b)&b\overline{d}\\ d\overline{b}&\text{nrm}(d)\end{pmatrix}

gives us that b=0b=0 and nrm(d)=1\text{nrm}(d)=1. Finally, we note that

(0zz¯0)\displaystyle\begin{pmatrix}0&z\\ \overline{z}&0\end{pmatrix} =(a00d)(0zz¯0)(a¯00d¯)\displaystyle=\begin{pmatrix}a&0\\ 0&d\end{pmatrix}\begin{pmatrix}0&z\\ \overline{z}&0\end{pmatrix}\begin{pmatrix}\overline{a}&0\\ 0&\overline{d}\end{pmatrix}
=(0azdz¯0)(a¯00d¯)\displaystyle=\begin{pmatrix}0&az\\ d\overline{z}&0\end{pmatrix}\begin{pmatrix}\overline{a}&0\\ 0&\overline{d}\end{pmatrix}
=(0azd¯dz¯a¯0)\displaystyle=\begin{pmatrix}0&az\overline{d}\\ d\overline{z}\,\overline{a}&0\end{pmatrix}

implies azd¯=zaz\overline{d}=z for all zH+z\in H^{+}. Since nrm(d)=1\text{nrm}(d)=1, this is just to say that az=zdaz=zd for all zH+z\in H^{+}, and since ad=1ad^{\ddagger}=1, this is the same as saying that az=za¯az=z\overline{a^{\ddagger}} for all zH+z\in H^{+}. It is easy to check this equation is satisfied only if aFa\in F, but since nrm(a)=1\text{nrm}(a)=1, we see that a2=1a^{2}=1, and therefore the kernel actually just consists of ±1\pm 1, as claimed. Since the kernel has dimension 0, the dimension of the image is dim(SL(2,H))=10\dim\left(SL^{\ddagger}(2,H)\right)=10 by Corollary 3.2, which is the same as the dimension of O(qH)O(q_{H}). Since SL(2,H)SL^{\ddagger}(2,H) is a symplectic group by Corollary 3.1, it is connected, and so its image must be O0(qH)O^{0}(q_{H}), the connected component of the identity. ∎

In characteristic 0, this sets up a correspondence between the groups SL(2,H)SL^{\ddagger}(2,H) and spin groups of quadratic forms.

Theorem 5.2.

Let FF be a characteristic 0 field. Then there is a bijection

{Isomorphism classes ofquaternion algebras over F}\displaystyle\left\{\begin{subarray}{c}\text{Isomorphism classes of}\\ \text{quaternion algebras over }F\end{subarray}\right\} {Isomorphism classes ofspin groups of indefinite,quinary quadratic forms over F}\displaystyle\rightarrow\left\{\begin{subarray}{c}\text{Isomorphism classes of}\\ \text{spin groups of indefinite,}\\ \text{quinary quadratic forms over }F\end{subarray}\right\}
[H]\displaystyle[H] [SL(2,H)].\displaystyle\mapsto\left[SL^{\ddagger}(2,H)\right].
Proof.

By Theorem 5.1, we know that SL(2,H)SL^{\ddagger}(2,H) is a double-cover of an orthogonal group, which is to say that it is a spin group. By Lemma 5.1, we know that this map is well-defined in that the choice of orthogonal involution \ddagger does not change the isomorphism class of SL(2,H)SL^{\ddagger}(2,H). It is easy to check that this map is surjective. Choose any indefinite, quinary quadratic form qq over FF. Since it is indefinite, we can decompose it as 1,1a,b,c\langle 1,-1\rangle\oplus\langle a,b,c\rangle, for some a,b,cF×a,b,c\in F^{\times}. In fact, since scaling the quadratic form does not change the spin group, we can assume that the quadratic form is 1,11,b,c\langle 1,-1\rangle\oplus\langle 1,b,c\rangle. In that case, it is clear that the image of

H=(b,cF)\displaystyle H=\left(\frac{-b,-c}{F}\right)

will be the desired spin group. So, we are finally left with checking that the map is injective, which is to say that if SL1(2,H1)SL^{\ddagger_{1}}(2,H_{1}) is isomorphic to SL2(2,H2)SL^{\ddagger_{2}}(2,H_{2}), then H1H2H_{1}\cong H_{2}. An isomorphism of the algebraic groups induces an isomorphism of the Lie algebras, which we know are

𝔰𝔩i(2,Hi)={MMat(2,Hi)|M^i=M}\displaystyle\mathfrak{sl}^{\ddagger_{i}}(2,H_{i})=\left\{M\in\text{Mat}(2,H_{i})\middle|M^{\hat{\ddagger}_{i}}=-M\right\}

by Theorem 3.1. This extends to an isomorphism

𝔰𝔩1(2,H1FF¯)𝔰𝔩2(2,H2FF¯).\displaystyle\mathfrak{sl}^{\ddagger_{1}}(2,H_{1}\otimes_{F}\overline{F})\cong\mathfrak{sl}^{\ddagger_{2}}(2,H_{2}\otimes_{F}\overline{F}).

However, we know that (H1FF¯,1)(H2FF¯,2)(H_{1}\otimes_{F}\overline{F},\ddagger_{1})\cong(H_{2}\otimes_{F}\overline{F},\ddagger_{2}), and in fact we can view (H1,1)(H_{1},\ddagger_{1}) and (H2,2)(H_{2},\ddagger_{2}) as embedded inside of a ring with involution (H,)(H,\ddagger) where HH is the unique quaternion algebra over F¯\overline{F}. Thus, the isomorphism of Lie algebras can be considered as coming from an automorphism of 𝔰𝔩(2,H)\mathfrak{sl}^{\ddagger}(2,H). However, by Theorem 1.1, we know that this is isomorphic to 𝔰𝔭(4)\mathfrak{sp}(4) which is a simple Lie algebra of type CnC_{n}. In characteristic 0, the outer automorphisms of such Lie algebras correspond to graph automorphisms of their corresponding Dynkin diagrams—however, there are no such automorphisms for the CnC_{n} type, and therefore the automorphism of 𝔰𝔩(2,H)\mathfrak{sl}^{\ddagger}(2,H) must be inner. That is to say, there exists some element γSL(2,H)\gamma\in SL^{\ddagger}(2,H) such that

𝔰𝔩1(2,H1)\displaystyle\mathfrak{sl}^{\ddagger_{1}}(2,H_{1}) 𝔰𝔩2(2,H2)\displaystyle\rightarrow\mathfrak{sl}^{\ddagger_{2}}(2,H_{2})
M\displaystyle M γMγ1.\displaystyle\mapsto\gamma M\gamma^{-1}.

This extends to an isomorphism of Mat(2,H1)\text{Mat}(2,H_{1}) with Mat(2,H2)\text{Mat}(2,H_{2}). Why is this? Well, any element MMat(2,H1)M\in\text{Mat}(2,H_{1}) can be written uniquely as a sum M=M+M′′M=M^{\prime}+M^{\prime\prime} such that M^=M{M^{\prime}}^{\hat{\ddagger}}=M^{\prime} and M′′^=M′′{M^{\prime\prime}}^{\hat{\ddagger}}=-M^{\prime\prime}. Clearly, M′′𝔰𝔩1(2,H1)M^{\prime\prime}\in\mathfrak{sl}^{\ddagger_{1}}(2,H_{1}) and there exists some M′′′𝔰𝔩1(2,H1)M^{\prime\prime\prime}\in\mathfrak{sl}^{\ddagger_{1}}(2,H_{1}) such that

M=(1001)𝔰𝔩1(2,H1)M′′′.\displaystyle M^{\prime}=\underbrace{\begin{pmatrix}1&0\\ 0&-1\end{pmatrix}}_{\in\mathfrak{sl}^{\ddagger_{1}}(2,H_{1})}M^{\prime\prime\prime}.

Therefore, γMγ1,γM′′γ1Mat(2,H2)\gamma M^{\prime}\gamma^{-1},\gamma M^{\prime\prime}\gamma^{-1}\in\text{Mat}(2,H_{2}), so γMγ1Mat(2,H2)\gamma M\gamma^{-1}\in\text{Mat}(2,H_{2}) for all MMat(2,H1)M\in\text{Mat}(2,H_{1}), and so we have our desired map

Mat(2,H1)\displaystyle\text{Mat}(2,H_{1}) Mat(2,H2)\displaystyle\rightarrow\text{Mat}(2,H_{2})
M\displaystyle M γMγ1.\displaystyle\mapsto\gamma M\gamma^{-1}.

However, Mat(2,H1)\text{Mat}(2,H_{1}) and Mat(2,H2)\text{Mat}(2,H_{2}) are central simple algebras and so there is an isomorphism between them if and only if H1H2H_{1}\cong H_{2}. ∎

Remark 5.4.

Theorem 1.4 is an immediate consequence of Theorem 5.2.

6. Orders of Quaternion Algebras:

We shall now consider some of the special features that are true for orders of quaternion algebras. First, note that all orders of a quaternion algebra are automatically closed under quaternion conjugation; therefore, there is only interest in looking at orders closed under an orthogonal involution. We previously noted that orthogonal involutions are classified by their discriminant. As it happens, if \ddagger is an orthogonal involution on a quaternion algebra, then we can compute an important algebraic invariant of the maximal \ddagger-orders in terms of this discriminant. However, to state our desired result, we shall need two other notions of discriminant as well. First, any quaternion algebra HH over KK is either a division algebra or isomorphic to Mat(2,K)\text{Mat}(2,K). For any place ν\nu of KK, we say that HH ramifies if HνH_{\nu} is a division algebra, and we say that it splits otherwise. Recalling that the finite places of KK correspond to its prime ideals, we define the discriminant of HH to be the ideal

disc(H)=𝔭 a prime idealH𝔭 ramifies𝔭.\displaystyle\text{disc}(H)=\prod_{\begin{subarray}{c}\mathfrak{p}\text{ a prime ideal}\\ H_{\mathfrak{p}}\text{ ramifies}\end{subarray}}\mathfrak{p}.

The places at which HH ramify uniquely determine it up to isomorphism; thus, knowing the infinite places where HH ramifies and the discriminant uniquely determines HH. In fact, since the number of ramified places is always even, over \mathbb{Q} the discriminant uniquely determines the isomorphism class. There is also the related notion of the discriminant of an order 𝒪\mathcal{O}, which we shall define as

disc(𝒪)2=det(tr(eiej¯))0i,j3𝔬K,\displaystyle\text{disc}(\mathcal{O})^{2}=\det\left(\text{tr}(e_{i}\overline{e_{j}})\right)_{0\leq i,j\leq 3}\mathfrak{o}_{K},

where e0,e1,e2,e3e_{0},e_{1},e_{2},e_{3} is any basis of 𝒪\mathcal{O}. One checks that the expression on the right is always a square ideal. One also checks that 𝒪\mathcal{O} is a maximal order if and only if disc(𝒪)=disc(H)\text{disc}(\mathcal{O})=\text{disc}(H). A similar characterization applies to \ddagger-orders as well.

Theorem 6.1.

[She18, Theorem 1.1] Given a quaternion algebra HH over a local or global field FF of characteristic not 22 and with an orthogonal involution \ddagger, the maximal \ddagger-orders of HH are exactly the orders of the form 𝒪𝒪\mathcal{O}\cap\mathcal{O}^{\ddagger} with discriminant

disc(H)ι(disc()),\displaystyle\text{disc}(H)\cap\iota(\text{disc}(\ddagger)),

where 𝒪\mathcal{O} is a maximal order and ι\iota is the map

ι:F×/(F×)2\displaystyle\iota:F^{\times}/\left(F^{\times}\right)^{2} {square-free ideals of 𝔬F}\displaystyle\rightarrow\left\{\text{square-free ideals of }\mathfrak{o}_{F}\right\}
[λ]\displaystyle[\lambda] λ[λ]𝔬λ𝔬F.\displaystyle\mapsto\bigcup_{\lambda\in[\lambda]\cap\mathfrak{o}}\lambda\mathfrak{o}_{F}.

Over local fields, one can obtain more detailed information about isomorphism classes. Surprisingly, unlike maximal orders, maximal \ddagger-orders are not necessarily all of the same isomorphism class over a local field; if the maximal ideal of the ring of integers contains 22, there can be multiple isomorphism classes—precise statements can be found in [She18]. Depending on the choice of involution, maximal \ddagger-orders can be maximal (in the usual sense) or strictly smaller. The number of isomorphism classes depends both on the class number and the number of local isomorphism classes.

Example 1.

Define (x+yi+zj+tij)=x+yizj+tij(x+yi+zj+tij)^{\ddagger}=x+yi-zj+tij. Then

𝒪1\displaystyle\mathcal{O}_{1} =i1+j2i+ij2(1,23)\displaystyle=\mathbb{Z}\oplus\mathbb{Z}i\oplus\mathbb{Z}\frac{1+j}{2}\oplus\mathbb{Z}\frac{i+ij}{2}\subset\left(\frac{-1,-23}{\mathbb{Q}}\right)
𝒪2\displaystyle\mathcal{O}_{2} =3i1+j211i+ij6(1,23)\displaystyle=\mathbb{Z}\oplus 3\mathbb{Z}i\oplus\mathbb{Z}\frac{1+j}{2}\oplus\mathbb{Z}\frac{11i+ij}{6}\subset\left(\frac{-1,-23}{\mathbb{Q}}\right)

are both easily checked to be \ddagger-orders. Both of them have discriminant (23)(23), which is the discriminant of the quaternion algebra; consequently, they are both maximal and \ddagger-maximal. All of the localizations of 𝒪1\mathcal{O}_{1} and 𝒪2\mathcal{O}_{2} are isomorphic as algebras with involution; this can be seen from the fact that there is only one isomorphism class for each localization [She18]. However, not only is it true that (𝒪1,)(𝒪2,)(\mathcal{O}_{1},\ddagger)\ncong(\mathcal{O}_{2},\ddagger), in fact 𝒪1𝒪2\mathcal{O}_{1}\ncong\mathcal{O}_{2}, since 𝒪1×={1,1,i,i}\mathcal{O}_{1}^{\times}=\{1,-1,i,-i\}, whereas 𝒪2×={1,1}\mathcal{O}_{2}^{\times}=\{1,-1\}. This is an easy computation using the fact that u𝒪i×u\in\mathcal{O}_{i}^{\times} if and only if nrm(u)=1\text{nrm}(u)=1; since both 𝒪i\mathcal{O}_{i} are contained inside a definite quaternion algebra, there are only finitely many possibilities for the units and it is easy to enumerate all of them.

Example 2.

Define (x+yi+zj+tij)=x+yi+zjtij(x+yi+zj+tij)^{\ddagger}=x+yi+zj-tij. Then

𝒪1\displaystyle\mathcal{O}_{1} =ii+j21+ij2(1,3)\displaystyle=\mathbb{Z}\oplus\mathbb{Z}i\oplus\mathbb{Z}\frac{i+j}{2}\oplus\mathbb{Z}\frac{1+ij}{2}\subset\left(\frac{-1,-3}{\mathbb{Q}}\right)
𝒪2\displaystyle\mathcal{O}_{2} =i1+j2i+ij2(1,3)\displaystyle=\mathbb{Z}\oplus\mathbb{Z}i\oplus\mathbb{Z}\frac{1+j}{2}\oplus\mathbb{Z}\frac{i+ij}{2}\subset\left(\frac{-1,-3}{\mathbb{Q}}\right)

are both \ddagger-orders, and both have discriminant (3)(3). Thus, they are both maximal and \ddagger-maximal. It is true that 𝒪1𝒪2\mathcal{O}_{1}\cong\mathcal{O}_{2}—this can be seen from a computation of the class number, which is 11—but (𝒪1,)(𝒪2,)(\mathcal{O}_{1},\ddagger)\ncong(\mathcal{O}_{2},\ddagger). This is because tr(𝒪1+)=(2)\text{tr}(\mathcal{O}_{1}^{+})=(2), but tr(𝒪2+)=(1)\text{tr}(\mathcal{O}_{2}^{+})=(1).

Example 3.

Define (x+yi+zj+tij)=x+yi+zjtij(x+yi+zj+tij)^{\ddagger}=x+yi+zj-tij. Then

𝒪=i1+i+j21+i+ij2(1,6)\displaystyle\mathcal{O}=\mathbb{Z}\oplus\mathbb{Z}i\oplus\mathbb{Z}\frac{1+i+j}{2}\oplus\mathbb{Z}\frac{1+i+ij}{2}\subset\left(\frac{-1,-6}{\mathbb{Q}}\right)

is a \ddagger-order with discriminant (6)(6). The discriminant of the quaternion algebra is 33, so this order is not maximal. However, it is \ddagger-maximal since disc()=6(×)2\text{disc}(\ddagger)=-6\left(\mathbb{Q}^{\times}\right)^{2}, and therefore disc(H)ι(disc())=(6)\text{disc}(H)\cap\iota(\text{disc}(\ddagger))=(6).

We already know that the groups SL(2,𝒪)SL^{\ddagger}(2,\mathcal{O}) are arithmetic subgroups of symplectic groups or, equivalently by Theorem 5.1, arithmetic subgroups of spin groups. The case of the greatest interest to the author is when HH is a definite, rational quaternion algebra as then SL(2,𝒪)SL^{\ddagger}(2,\mathcal{O}) will be an arithmetic subgroup of Isom0(4)SO+(4,1)\text{Isom}^{0}(\mathbb{H}^{4})\cong SO^{+}(4,1). However, we can state our result a little more generally and just consider the case where HH is a rational quaternion algebra; then SL(2,𝒪)SL^{\ddagger}(2,\mathcal{O}) will either be an arithmetic group of SO+(4,1)SO^{+}(4,1) if HH is definite or SO+(3,2)SO^{+}(3,2) if it is indefinite. In either case, the matrix ring is now an invariant under group isomorphism.

Lemma 6.1.

Let H1,H2H_{1},H_{2} be rational quaternion algebras with orthogonal involutions 1,2\ddagger_{1},\ddagger_{2}. Let Γ1,Γ2\Gamma_{1},\Gamma_{2} be lattices of SL1(2,H1)SL^{\ddagger_{1}}(2,H_{1}), SL2(2,H2)SL^{\ddagger_{2}}(2,H_{2}) such that their centers are {±1}\{\pm 1\}. If Γ1\Gamma_{1} and Γ2\Gamma_{2} are isomorphic as groups, then ([Γ1],^1)([Γ2],^2)(\mathbb{Z}[\Gamma_{1}],\hat{\ddagger}_{1})\cong(\mathbb{Z}[\Gamma_{2}],\hat{\ddagger}_{2}).

Proof.

Let ϕ:Γ1Γ2\phi:\Gamma_{1}\rightarrow\Gamma_{2} be the group isomorphism. Note that ϕ(I)=I\phi(-I)=-I, since I-I is the unique non-identity element of the centers of Γi\Gamma_{i}. Therefore, it induces an isomorphism ϕ¯:Γ¯1Γ¯2\overline{\phi}:\overline{\Gamma}_{1}\rightarrow\overline{\Gamma}_{2} between the images of the Γi\Gamma_{i} inside SLi(2,Hi)/{±I}SL^{\ddagger_{i}}(2,H_{i})/\{\pm I\}. Note that

SLi(2,Hi)/{±I}{PSO+(4,1)if HiHPSO+(3,2)if HiMat(2,),\displaystyle SL^{\ddagger_{i}}(2,H_{i}\otimes_{\mathbb{Q}}\mathbb{R})/\{\pm I\}\cong\begin{cases}PSO^{+}(4,1)&\text{if }H_{i}\otimes_{\mathbb{Q}}\mathbb{R}\cong H_{\mathbb{R}}\\ PSO^{+}(3,2)&\text{if }H_{i}\otimes_{\mathbb{Q}}\mathbb{R}\cong\text{Mat}(2,\mathbb{R}),\end{cases}

and so we can apply the Mostow rigidity theorem to conclude that both Γ1¯\overline{\Gamma_{1}} and Γ2¯\overline{\Gamma_{2}} can be viewed as being lattices of the same Lie group GG, namely either PSO+(3,2)PSO^{+}(3,2) or PSO+(4,1)PSO^{+}(4,1). In either case, GG is simple, and therefore by Mostow rigidity Γ1¯\overline{\Gamma_{1}} and Γ2¯\overline{\Gamma_{2}} are conjugate in GG. Let gGg\in G be such that Γ2¯=gΓ1¯g1\overline{\Gamma_{2}}=g\overline{\Gamma_{1}}g^{-1}, and choose an element gSL(2,Hi)g^{\prime}\in SL^{\ddagger}(2,H_{i}\otimes_{\mathbb{R}}\mathbb{R}) such that its image in the quotient is gg. Then we have a well-defined ring isomorphism

Φ:[Γ1]\displaystyle\Phi:\mathbb{Z}[\Gamma_{1}] [Γ2]\displaystyle\rightarrow\mathbb{Z}[\Gamma_{2}]
M\displaystyle M gMg1.\displaystyle\mapsto g^{\prime}M{g^{\prime}}^{-1}.

Letting \ddagger be the orthogonal involution on H1H2H_{1}\otimes_{\mathbb{Q}}\mathbb{R}\cong H_{2}\otimes_{\mathbb{Q}}\mathbb{R}, we see that

(gMg1)^=gM^g1\displaystyle\left(g^{\prime}M{g^{\prime}}^{-1}\right)^{\hat{\ddagger}}=g^{\prime}M^{\hat{\ddagger}}{g^{\prime}}^{-1}

since gSL(2,Hi)g^{\prime}\in SL^{\ddagger}(2,H_{i}\otimes_{\mathbb{R}}\mathbb{R}). Therefore, the map we have constructed is an isomorphism of rings with involution. ∎

Remark 6.1.

The use of the Mostow rigidity theorem in Lemma 6.1 makes clear why we restrict to the case where HH is a quaternion algebra over \mathbb{Q}—over other number fields, the set of infinite places Ω\Omega_{\infty} has more than one element, and so rather than SL(2,𝒪)/{±I}SL^{\ddagger}(2,\mathcal{O})/\{\pm I\} injecting as a lattice into PSO0(4,1)PSO^{0}(4,1) or PSO0(3,2)PSO^{0}(3,2), it will instead inject into some non-simple Lie group.

Remark 6.2.

In light of Theorem 4.2, Lemma 6.1 immediately implies Theorem 1.3.

For commutative rings, it is true that Mat(2,R)Mat(2,S)\text{Mat}(2,R)\cong\text{Mat}(2,S) implies RSR\cong S. This is false in general for non-commutative rings. We know that SL1(2,𝒪1)SL2(2,𝒪2)SL^{\ddagger_{1}}(2,\mathcal{O}_{1})\cong SL^{\ddagger_{2}}(2,\mathcal{O}_{2}) if and only if (Mat(2,𝒪1),^1)(Mat(2,𝒪2),^1)(\text{Mat}(2,\mathcal{O}_{1}),\hat{\ddagger}_{1})\cong(\text{Mat}(2,\mathcal{O}_{2}),\hat{\ddagger}_{1}); we now give a couple of examples showing that this latter condition could not be replaced with (𝒪1,1)(𝒪2,2)(\mathcal{O}_{1},\ddagger_{1})\cong(\mathcal{O}_{2},\ddagger_{2}) or Mat(2,𝒪1)Mat(2,𝒪2)\text{Mat}(2,\mathcal{O}_{1})\cong\text{Mat}(2,\mathcal{O}_{2}).

Example 4.

Let

𝒪1\displaystyle\mathcal{O}_{1} =i1+j2i+ij2(1,7)\displaystyle=\mathbb{Z}\oplus\mathbb{Z}i\oplus\mathbb{Z}\frac{1+j}{2}\oplus\mathbb{Z}\frac{i+ij}{2}\subset\left(\frac{-1,-7}{\mathbb{Q}}\right)
𝒪2\displaystyle\mathcal{O}_{2} =ii+j21+ij2(1,7).\displaystyle=\mathbb{Z}\oplus\mathbb{Z}i\oplus\mathbb{Z}\frac{i+j}{2}\oplus\mathbb{Z}\frac{1+ij}{2}\subset\left(\frac{-1,-7}{\mathbb{Q}}\right).

Both of these are maximal \ddagger-orders, if we take the usual involution (x+yi+zj+tij)=x+yi+zjtij\left(x+yi+zj+tij\right)^{\ddagger}=x+yi+zj-tij. Since tr(𝒪1H+)=(1)\text{tr}(\mathcal{O}_{1}\cap H^{+})=(1) and tr(𝒪2H+)=(2)\text{tr}(\mathcal{O}_{2}\cap H^{+})=(2), we see that (𝒪1,)(𝒪2,)(\mathcal{O}_{1},\ddagger)\ncong(\mathcal{O}_{2},\ddagger). However, (1+i)𝒪1(1+i)1=𝒪2(1+i)\mathcal{O}_{1}(1+i)^{-1}=\mathcal{O}_{2}, and therefore

(1+i2001+i2)SL(2,𝒪1)(1+i2001+i2)1\displaystyle\begin{pmatrix}\frac{1+i}{\sqrt{2}}&0\\ 0&\frac{-1+i}{\sqrt{2}}\end{pmatrix}SL^{\ddagger}(2,\mathcal{O}_{1})\begin{pmatrix}\frac{1+i}{\sqrt{2}}&0\\ 0&\frac{-1+i}{\sqrt{2}}\end{pmatrix}^{-1} =SL(2,𝒪2).\displaystyle=SL^{\ddagger}(2,\mathcal{O}_{2}).
Remark 6.3.

This example was originally worked out in [She19].

Example 5.

Take 𝒪1,𝒪2\mathcal{O}_{1},\mathcal{O}_{2} as in Example 1. We proved that 𝒪1𝒪2\mathcal{O}_{1}\ncong\mathcal{O}_{2}; however, we claim that SL(2,𝒪1)SL(2,𝒪2)SL^{\ddagger}(2,\mathcal{O}_{1})\cong SL^{\ddagger}(2,\mathcal{O}_{2}).

Proof.

Define

γ=(16i+j231+j231+6i+j233i)SL(2,H(3)).\displaystyle\gamma=\begin{pmatrix}\frac{1-6i+j}{2\sqrt{3}}&\frac{1+j}{2\sqrt{3}}\\ \frac{1+6i+j}{2\sqrt{3}}&\sqrt{3}i\end{pmatrix}\in SL^{\ddagger}\left(2,H\otimes_{\mathbb{Q}}\mathbb{Q}(\sqrt{3})\right).

By an easy computation,

γ(0100)γ1\displaystyle\gamma\begin{pmatrix}0&1\\ 0&0\end{pmatrix}\gamma^{-1} Mat(2,𝒪2)\displaystyle\in\text{Mat}(2,\mathcal{O}_{2})
γ(0i00)γ1\displaystyle\gamma\begin{pmatrix}0&i\\ 0&0\end{pmatrix}\gamma^{-1} Mat(2,𝒪2)\displaystyle\in\text{Mat}(2,\mathcal{O}_{2})
γ(01+j200)γ1\displaystyle\gamma\begin{pmatrix}0&\frac{1+j}{2}\\ 0&0\end{pmatrix}\gamma^{-1} Mat(2,𝒪2)\displaystyle\in\text{Mat}(2,\mathcal{O}_{2})
γ(0i+ij200)γ1\displaystyle\gamma\begin{pmatrix}0&\frac{i+ij}{2}\\ 0&0\end{pmatrix}\gamma^{-1} Mat(2,𝒪2)\displaystyle\in\text{Mat}(2,\mathcal{O}_{2})
γ(0010)γ1\displaystyle\gamma\begin{pmatrix}0&0\\ 1&0\end{pmatrix}\gamma^{-1} Mat(2,𝒪2).\displaystyle\in\text{Mat}(2,\mathcal{O}_{2}).

However, the elements

(0100),(0i00),(01+j200),(0i+ij200),(0010)\displaystyle\begin{pmatrix}0&1\\ 0&0\end{pmatrix},\begin{pmatrix}0&i\\ 0&0\end{pmatrix},\begin{pmatrix}0&\frac{1+j}{2}\\ 0&0\end{pmatrix},\begin{pmatrix}0&\frac{i+ij}{2}\\ 0&0\end{pmatrix},\begin{pmatrix}0&0\\ 1&0\end{pmatrix}

generate Mat(2,𝒪1)\text{Mat}(2,\mathcal{O}_{1}) as a \mathbb{Z}-algebra, so we have a well-defined, injective ring homomorphism

Mat(2,𝒪1)\displaystyle\text{Mat}(2,\mathcal{O}_{1}) Mat(2,𝒪2)\displaystyle\rightarrow\text{Mat}(2,\mathcal{O}_{2})
M\displaystyle M γMγ1.\displaystyle\mapsto\gamma M\gamma^{-1}.

This map must be surjective, as can be checked either from computing the discriminants of Mat(2,𝒪1)\text{Mat}(2,\mathcal{O}_{1}) and Mat(2,𝒪2)\text{Mat}(2,\mathcal{O}_{2}), or simply by noting that SL(2,𝒪1),SL(2,𝒪2)SL^{\ddagger}(2,\mathcal{O}_{1}),SL^{\ddagger}(2,\mathcal{O}_{2}) are maximal arithmetic groups. In any case, this is an isomorphism of rings with involution, and therefore by Theorem 1.3, SL(2,𝒪1)SL(2,𝒪2)SL^{\ddagger}(2,\mathcal{O}_{1})\cong SL^{\ddagger}(2,\mathcal{O}_{2}). ∎

Remark 6.4.

Note that we have effectively produced a proof that one can find two non-isomorphic rings 𝒪1,𝒪2\mathcal{O}_{1},\mathcal{O}_{2} such that Mat(2,𝒪1)Mat(2,𝒪2)\text{Mat}(2,\mathcal{O}_{1})\cong\text{Mat}(2,\mathcal{O}_{2}). While this is technically a new proof of that fact, this example was in fact already worked out by Chatters [Cha96, Example 5.1].

Example 6.

Let

𝒪1\displaystyle\mathcal{O}_{1} =ij1+i+j+ij2(1,5)\displaystyle=\mathbb{Z}\oplus\mathbb{Z}i\oplus\mathbb{Z}j\oplus\mathbb{Z}\frac{1+i+j+ij}{2}\subset\left(\frac{-1,-5}{\mathbb{Q}}\right)
𝒪2\displaystyle\mathcal{O}_{2} =i1+i+j21+i+ij2(1,10).\displaystyle=\mathbb{Z}\oplus\mathbb{Z}i\oplus\mathbb{Z}\frac{1+i+j}{2}\oplus\mathbb{Z}\frac{1+i+ij}{2}\subset\left(\frac{-1,-10}{\mathbb{Q}}\right).

In both cases, define (x+yi+zj+tij)=x+yi+zjtij(x+yi+zj+tij)^{\ddagger}=x+yi+zj-tij—this is slight abuse of notation, since these two orders have entirely different bases. However, 𝒪1𝒪2\mathcal{O}_{1}\cong\mathcal{O}_{2}—this is because both of their ambient quaternion algebras have discriminant (2)(2) and so are isomorphic, they both have discriminant (10)(10), and their class numbers are 11. Thus, they are both Eichler orders of the same level and must be conjugate to one another. This means that Mat(2,𝒪1)Mat(2,𝒪2)\text{Mat}(2,\mathcal{O}_{1})\cong\text{Mat}(2,\mathcal{O}_{2}). However, we claim that (Mat(2,𝒪1),^)(Mat(2,𝒪2),^)(\text{Mat}(2,\mathcal{O}_{1}),\hat{\ddagger})\ncong(\text{Mat}(2,\mathcal{O}_{2}),\hat{\ddagger}), and consequently SL(2,𝒪1)SL(2,𝒪2)SL^{\ddagger}(2,\mathcal{O}_{1})\ncong SL^{\ddagger}(2,\mathcal{O}_{2}).

Proof.

Define two lattices

Mi={MMat(2,𝒪i)+|tr(M)=0}\displaystyle M_{i}=\left\{M\in\text{Mat}(2,\mathcal{O}_{i})^{+}\middle|\text{tr}(M)=0\right\}

and consider the integral quadratic forms

qi:Mi\displaystyle q_{i}:M_{i} \displaystyle\rightarrow\mathbb{Z}
M\displaystyle M tr(M2).\displaystyle\mapsto\text{tr}(M^{2}).

What do these quadratic forms look like? First, note that

Mat(2,𝒪i)+\displaystyle\text{Mat}(2,\mathcal{O}_{i})^{+} ={(abca)Mat(2,𝒪i)|b,c𝒪i}\displaystyle=\left\{\begin{pmatrix}a&b\\ c&a^{\ddagger}\end{pmatrix}\in\text{Mat}(2,\mathcal{O}_{i})\middle|b,c\in\mathcal{O}_{i}^{-}\right\}
Mi\displaystyle M_{i} ={(abca)Mat(2,𝒪i)|b,c𝒪i,tr(a)=0}.\displaystyle=\left\{\begin{pmatrix}a&b\\ c&a^{\ddagger}\end{pmatrix}\in\text{Mat}(2,\mathcal{O}_{i})\middle|b,c\in\mathcal{O}_{i}^{-},\ \text{tr}(a)=0\right\}.

Furthermore,

tr((abca)2)\displaystyle\text{tr}\left(\begin{pmatrix}a&b\\ c&a^{\ddagger}\end{pmatrix}^{2}\right) =tr((a2+bcab+bacb+accb+(a)2))\displaystyle=\text{tr}\left(\begin{pmatrix}a^{2}+bc&ab+ba^{\ddagger}\\ cb+a^{\ddagger}c&cb+\left(a^{\ddagger}\right)^{2}\end{pmatrix}\right)
=2tr(bc)+2tr(a2).\displaystyle=2\text{tr}(bc)+2\text{tr}(a^{2}).

Since tr(a)=0\text{tr}(a)=0, a2=nrm(a)a^{2}=-\text{nrm}(a). Furthermore, any element x𝒪ix\in\mathcal{O}_{i}^{-} with be of the form nijnij for some integer nn. Therefore, our quadratic forms actually look like 4nrm(ij)st4nrm(a)4\text{nrm}(ij)st-4\text{nrm}(a). Working out exactly what this is in coordinates, we have

q1(s,t,x,y,z)\displaystyle q_{1}(s,t,x,y,z) =5stx25y2xz5yz3z2\displaystyle=5st-x^{2}-5y^{2}-xz-5yz-3z^{2}
q2(s,t,x,y,z)\displaystyle q_{2}(s,t,x,y,z) =10stx2xy3y2xzyz3z2.\displaystyle=10st-x^{2}-xy-3y^{2}-xz-yz-3z^{2}.

One can check that these quadratic forms are not equivalent. However, if there was an isomorphism (Mat(2,𝒪1),^)(Mat(2,𝒪2),^)(\text{Mat}(2,\mathcal{O}_{1}),\hat{\ddagger})\rightarrow(\text{Mat}(2,\mathcal{O}_{2}),\hat{\ddagger}) then it would give a polynomial map between M1M_{1} and M2M_{2}, and thus an equivalence between q1q_{1} and q2q_{2}. ∎

We include one final example showing that conjugacy in SL(2,HKK¯)SL^{\ddagger}(2,H\otimes_{K}\overline{K}) cannot be replaced with conjugacy in SL(2,H)SL^{\ddagger}(2,H), even if (𝒪1,)(𝒪2,)(\mathcal{O}_{1},\ddagger)\cong(\mathcal{O}_{2},\ddagger).

Example 7.

Let KK be any characteristic 0 local field with maximal ideal 𝔭\mathfrak{p}. Choose any λ𝔬F\lambda\in\mathfrak{o}_{F} such that 1+λ𝔭\𝔭21+\lambda\in\mathfrak{p}\backslash\mathfrak{p}^{2}. Then H=Mat(2,K)H=\text{Mat}(2,K) is a quaternion algebra over FF, and

(abcd)=(ac/λbλd)\displaystyle\begin{pmatrix}a&b\\ c&d\end{pmatrix}^{\ddagger}=\begin{pmatrix}a&c/\lambda\\ b\lambda&d\end{pmatrix}

defines an orthogonal involution. Since λ𝔬×\lambda\in\mathfrak{o}^{\times}, 𝒪1=Mat(2,𝔬K)\mathcal{O}_{1}=\text{Mat}(2,\mathfrak{o}_{K}) is a maximal \ddagger-order. Similarly,

𝒪2=(11λ1):=u𝒪1(11λ1)1\displaystyle\mathcal{O}_{2}=\underbrace{\begin{pmatrix}1&-1\\ \lambda&1\end{pmatrix}}_{:=u}\mathcal{O}_{1}\begin{pmatrix}1&-1\\ \lambda&1\end{pmatrix}^{-1}

must be a maximal \ddagger-order; indeed, (𝒪1,)(𝒪2,)(\mathcal{O}_{1},\ddagger)\cong(\mathcal{O}_{2},\ddagger). However, SL(2,𝒪1)SL^{\ddagger}(2,\mathcal{O}_{1}) and SL(2,𝒪2)SL^{\ddagger}(2,\mathcal{O}_{2}) are not conjugate in SL(2,H)SL^{\ddagger}(2,H).

Proof.

Suppose that there exist a,b,c,dHa,b,c,d\in H such that

γ:=(abcd)SL(2,H)\displaystyle\gamma:=\begin{pmatrix}a&b\\ c&d\end{pmatrix}\in SL^{\ddagger}(2,H)

and γSL(2,𝒪1)γ1=SL(2,𝒪2)\gamma SL^{\ddagger}(2,\mathcal{O}_{1})\gamma^{-1}=SL^{\ddagger}(2,\mathcal{O}_{2}). Since

(abcd)(1z01)(abcd)1\displaystyle\begin{pmatrix}a&b\\ c&d\end{pmatrix}\begin{pmatrix}1&z\\ 0&1\end{pmatrix}\begin{pmatrix}a&b\\ c&d\end{pmatrix}^{-1} =(azaczc)\displaystyle=\begin{pmatrix}*&aza^{\ddagger}\\ -czc^{\ddagger}&*\end{pmatrix}
(abcd)(10z1)(abcd)1\displaystyle\begin{pmatrix}a&b\\ c&d\end{pmatrix}\begin{pmatrix}1&0\\ z&1\end{pmatrix}\begin{pmatrix}a&b\\ c&d\end{pmatrix}^{-1} =(bzbdzd),\displaystyle=\begin{pmatrix}*&-bzb^{\ddagger}\\ dzd^{\ddagger}&*\end{pmatrix},

we wish to determine for which vHv\in H v𝒪1+v𝒪2v\mathcal{O}_{1}^{+}v^{\ddagger}\subset\mathcal{O}_{2}. This is the same as determining all vHv\in H such that u1v𝒪1+vu𝒪1u^{-1}v\mathcal{O}_{1}^{+}v^{\ddagger}u\subset\mathcal{O}_{1}. We shall show that this is possible only if v𝒪1v\in\mathcal{O}_{1}, proving that γSL(2,𝒪1)\gamma\in SL^{\ddagger}(2,\mathcal{O}_{1}). Write

v=(v1v2v3v4),\displaystyle v=\begin{pmatrix}v_{1}&v_{2}\\ v_{3}&v_{4}\end{pmatrix},

so

u1v(1000)vu\displaystyle u^{-1}v\begin{pmatrix}1&0\\ 0&0\end{pmatrix}v^{\ddagger}u =((v1+v3)2λ+1(λv1v3)(v1+v3)λ(λ+1)(λv1v3)(v1+v3)λ+1(λv1v3)2λ(λ+1))Mat(2,𝔬K)\displaystyle=\begin{pmatrix}\frac{\left(v_{1}+v_{3}\right)^{2}}{\lambda+1}&-\frac{\left(\lambda v_{1}-v_{3}\right)\left(v_{1}+v_{3}\right)}{\lambda(\lambda+1)}\\ -\frac{\left(\lambda v_{1}-v_{3}\right)\left(v_{1}+v_{3}\right)}{\lambda+1}&\frac{\left(\lambda v_{1}-v_{3}\right)^{2}}{\lambda(\lambda+1)}\end{pmatrix}\in\text{Mat}(2,\mathfrak{o}_{K})
u1v(0001)vu\displaystyle u^{-1}v\begin{pmatrix}0&0\\ 0&1\end{pmatrix}v^{\ddagger}u =(λ(v2+v4)2λ+1(λv2v4)(v2+v4)λ+1λ(λv2v4)(v2+v4)λ+1(λv2v4)2λ+1)Mat(2,𝔬K).\displaystyle=\begin{pmatrix}\frac{\lambda\left(v_{2}+v_{4}\right)^{2}}{\lambda+1}&-\frac{\left(\lambda v_{2}-v_{4}\right)\left(v_{2}+v_{4}\right)}{\lambda+1}\\ -\frac{\lambda\left(\lambda v_{2}-v_{4}\right)\left(v_{2}+v_{4}\right)}{\lambda+1}&\frac{\left(\lambda v_{2}-v_{4}\right)^{2}}{\lambda+1}\end{pmatrix}\in\text{Mat}(2,\mathfrak{o}_{K}).

Note that this is only possible if v1+v3,λv1v3,v2+v4,λv2v4𝔭v_{1}+v_{3},\lambda v_{1}-v_{3},v_{2}+v_{4},\lambda v_{2}-v_{4}\in\mathfrak{p}. From it follows that (1+λ)v1,(1+λ)v3𝔭(1+\lambda)v_{1},(1+\lambda)v_{3}\in\mathfrak{p}, hence v1,v3𝔬Fv_{1},v_{3}\in\mathfrak{o}_{F}. Therefore, v3,v4𝔬Fv_{3},v_{4}\in\mathfrak{o}_{F}. We conclude that v𝒪1v\in\mathcal{O}_{1}. Ergo, γSL(2,𝒪1)\gamma\in SL^{\ddagger}(2,\mathcal{O}_{1}), and so γSL(2,𝒪1)γ1=SL(2,𝒪1)=SL(2,𝒪2)\gamma SL^{\ddagger}(2,\mathcal{O}_{1})\gamma^{-1}=SL^{\ddagger}(2,\mathcal{O}_{1})=SL^{\ddagger}(2,\mathcal{O}_{2}). However, since

u(1000)u1=(11+λ11+λλ1+λλ1+λ)𝒪1,\displaystyle u\begin{pmatrix}1&0\\ 0&0\end{pmatrix}u^{-1}=\begin{pmatrix}\frac{1}{1+\lambda}&\frac{1}{1+\lambda}\\ \frac{\lambda}{1+\lambda}&\frac{\lambda}{1+\lambda}\end{pmatrix}\notin\mathcal{O}_{1},

𝒪1𝒪2\mathcal{O}_{1}\neq\mathcal{O}_{2}, and so SL(2,𝒪1)SL(2,𝒪2)SL^{\ddagger}(2,\mathcal{O}_{1})\neq SL^{\ddagger}(2,\mathcal{O}_{2}). This is a contradiction, and so we are done. ∎

References

  • [Ahl86] Lars V. Ahlfors. Möbius transformations in expressed through 2×22\times 2 matrices of Clifford numbers. Complex Variables, Theory and Application: An International Journal, 5(2-4):215–224, 1986.
  • [Cha96] A.W. Chatters. Matrix-isomorphic maximal ZZ-orders. Journal of Algebra, 181(2):593–600, 1996.
  • [GLM+05] Ronald L. Graham, Jeffrey C. Lagarias, Colin L. Mallows, Allan R. Wilks, and Catherine H. Yan. Apollonian circle packings: Geometry and group theory ii. super-Apollonian group and integral packings. Discrete & Computational Geometry, 35(1):1–36, 2005.
  • [GM10] Gerhard Guettler and Colin Mallows. A generalization of Apollonian packing of circles. Journal of Combinatorics, 1(1):1–27, 2010.
  • [KMRT98] Max-Albert Knus, Alexander Merkurjev, Markus Rost, and Jean-Pierre Tignol. The Book of Involutions. American Mathematical Society, June 1998.
  • [KN19] Alex Kontorovich and Kei Nakamura. Geometry and arithmetic of crystallographic sphere packings. Proceedings of the National Academy of Sciences, 116(2):436–441, 2019.
  • [Kne65] Martin Kneser. Starke approximation in algebraischen Gruppen. i. Journal für die reine und angewandte Mathematik, 218:190–203, 1965.
  • [Pla69] Vladimir P. Platonov. The problem of strong approximation and the Kneser-Tits conjecture for algebraic groups. Mathematics of the USSR-Izvestiya, 3(6):1139, 1969.
  • [Sal78] David J. Saltman. Azumaya algebras with involution. Journal of Algebra, 52(2):526 – 539, 1978.
  • [Sch74] Winfried Scharlau. Involutions on orders. i. Journal für die reine und angewandte Mathematik, 0268_0269:190–202, 1974.
  • [She18] Arseniy Sheydvasser. Orders of quaternion algebras with involution. Journal of Number Theory, 183:249 – 268, 2018.
  • [She19] Arseniy Sheydvasser. Quaternion orders and sphere packings. Journal of Number Theory, 204:41 – 98, 2019.
  • [Sta15] Katherine E. Stange. The Apollonian structure of Bianchi groups. Transactions of the American Mathematical Society, 5 2015.
  • [Vah02] Karl Theodor Vahlen. Ueber bewegungen und complexe Zahlen. Mathematische Annalen, 55(4):585–593, Dec 1902.