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

On Superspecial abelian surfaces over finite fields III

Jiangwei Xue (Xue) Collaborative Innovation Center of Mathematics, School of Mathematics and Statistics, Wuhan University, Luojiashan, 430072, Wuhan, Hubei, P.R. China (Xue) Hubei Key Laboratory of Computational Science (Wuhan University), Wuhan, Hubei, 430072, P.R. China. xue_j@whu.edu.cn Chia-Fu Yu (Yu) Institute of Mathematics, Academia Sinica and NCTS, Astronomy-Mathematics Building, No. 1, Sec. 4, Roosevelt Road, Taipei 10617, TAIWAN. chiafu@math.sinica.edu.tw  and  Yuqiang Zheng (Zheng) School of Mathematics and Statistics, Wuhan University, Luojiashan, 430072, Wuhan, Hubei, P.R. China zhhhhxhq@whu.edu.cn (Zheng) Academy of Mathematics and Systems Science, Chinese Academy of Science, No. 55, Zhongguancun East Road, Beijing 100190, China zhengyq@amss.ac.cn
Abstract.

In the paper [On superspecial abelian surfaces over finite fields II. J. Math. Soc. Japan, 72(1):303–331, 2020], Tse-Chung Yang and the first two current authors computed explicitly the number |SSp2(𝔽q)|\lvert\mathrm{SSp}_{2}(\mathbb{F}_{q})\rvert of isomorphism classes of superspecial abelian surfaces over an arbitrary finite field 𝔽q\mathbb{F}_{q} of even degree over the prime field 𝔽p\mathbb{F}_{p}. There it was assumed that certain commutative p\mathbb{Z}_{p}-orders satisfy an étale condition that excludes the primes p=2,3,5p=2,3,5. We treat these remaining primes in the present paper, where the computations are more involved because of the ramifications. This completes the calculation of |SSp2(𝔽q)|\lvert\mathrm{SSp}_{2}(\mathbb{F}_{q})\rvert in the even degree case. The odd degree case was previous treated by Tse-Chung Yang and the first two current authors in [On superspecial abelian surfaces over finite fields.Doc. Math., 21:1607–1643, 2016]. Along the proof of our main theorem, we give the classification of lattices over local quaternion Bass orders, which is a new input to our previous works.

Key words and phrases:
superspecial abelian surfaces, quaternion algebra, Bass order, conjugacy classes of arithmetic subgroups
2020 Mathematics Subject Classification:
11R52, 11G10

1. Introduction

Throughout this paper, pp denotes a prime number, q=paq=p^{a} a power of pp, and 𝔽q\mathbb{F}_{q} the finite field of qq-elements. We reserve {\mathbb{N}} for the set of strictly positive integers. Let kk be a field of characteristic pp, and k¯\bar{k} an algebraic closure of kk. An abelian variety over kk is said to be supersingular if it is isogenous to a product of supersingular elliptic curves over k¯\bar{k}; it is said to be superspecial if it is isomorphic to a product of supersingular elliptic curves over k¯\bar{k}. For any dd\in{\mathbb{N}}, denote by SSpd(𝔽q)\operatorname{SSp}_{d}(\mathbb{F}_{q}) the set of 𝔽q\mathbb{F}_{q}-isomorphism classes of dd-dimensional superspecial abelian varieties over 𝔽q\mathbb{F}_{q}. The classification of supersingular elliptic curves (namely, the d=1d=1 case) over finite fields were carried out by Deuring [8, 7], Eichler [10], Igusa[11], Waterhouse [24] and many others since the 1930s.

In a series of papers [26, 28, 27, 25], Tse-Chung Yang and the first two current authors attempt to calculate the cardinality |SSpd(𝔽q)|\lvert\operatorname{SSp}_{d}(\mathbb{F}_{q})\rvert explicitly in the case d=2d=2. More precisely, it is shown in [26] that for every fixed d>1d>1, |SSpd(𝔽q)|\lvert\operatorname{SSp}_{d}(\mathbb{F}_{q})\rvert depends only on the parity of the degree a=[𝔽q:𝔽p]a=[\mathbb{F}_{q}:\mathbb{F}_{p}], and an explicit formula of |SSp2(𝔽q)|\lvert\operatorname{SSp}_{2}(\mathbb{F}_{q})\rvert is provided for the odd degree case. The most involving part of this explicit calculation is carried out prior in [27, 25], which counts the number of isomorphism classes of abelian surfaces over 𝔽p\mathbb{F}_{p} within the simple isogeny class corresponding to the Weil pp-numbers ±p\pm\,\sqrt[]{p}\,. For the even degree case, an explicit formula of |SSp2(𝔽q)|\lvert\operatorname{SSp}_{2}(\mathbb{F}_{q})\rvert is obtained in [28] under a mild condition on pp (see Remark 3.7 of loc. cit.), which holds for all p7p\geq 7. We treat the remaining primes p{2,3,5}p\in\{2,3,5\} in the present paper, thus completing the calculation of |SSp2(𝔽q)|\lvert\operatorname{SSp}_{2}(\mathbb{F}_{q})\rvert in the even degree case.

For the rest of the paper, we assume that q=paq=p^{a} is an even power of pp. All isogenies and isomorphisms are over the base field 𝔽q\mathbb{F}_{q} unless specified otherwise. The set SSp2(𝔽q)\operatorname{SSp}_{2}(\mathbb{F}_{q}) naturally partitions into subsets by isogeny equivalence, which can be parametrized by (multiple) Weil numbers (see [26, §4.1]). For each integer nn\in{\mathbb{N}}, let ζn\zeta_{n} be a primitive nn-th root of unity, and πn\pi_{n} be the Weil qq-number (p)a/2ζn(-p)^{a/2}\zeta_{n}. By the Honda-Tate theorem, there is a unique simple abelian variety Xn/𝔽qX_{n}/\mathbb{F}_{q} up to isogeny corresponding to the Gal(¯/)\operatorname{Gal}(\bar{\mathbb{Q}}/\mathbb{Q})-conjugacy class of πn\pi_{n}. Moreover, the XnX_{n}’s are mutually non-isogenous for distinct nn. Thanks to the Manin-Oort Theorem [29, Theorem 2.9], a simple abelian variety over 𝔽q\mathbb{F}_{q} is supersingular if and only if it is isogenous to XnX_{n} for some nn. Let d(n)d(n) be the dimension of XnX_{n}. The formula for d(n)d(n) is given in [26, §3]. When p{2,3,5}p\in\{2,3,5\}, we have

  • d(n)=1d(n)=1 if and only if n{1,2,3,6}n\in\{1,2,3,6\} or (n,p){(4,2),(4,3)}(n,p)\in\{(4,2),(4,3)\};

  • d(n)=2d(n)=2 if and only if (n,p)=(4,5)(n,p)=(4,5) or n{5,8,10,12}n\in\{5,8,10,12\}.

Given a superspecial abelian surface X/𝔽qX/\mathbb{F}_{q}, we have two cases to consider:

  1. (I)

    the isotypic case where XX is isogenous to Xn2/d(n)X_{n}^{2/d(n)} for some nn\in{\mathbb{N}} with d(n)2d(n)\leq 2;

  2. (II)

    the non-isotypic case where XX is isogenous to Xn¯:=Xn1×Xn2X_{\underline{n}}:=X_{n_{1}}\times X_{n_{2}} for a pair n¯=(n1,n2)2{\underline{n}}=(n_{1},n_{2})\in{\mathbb{N}}^{2} with n1<n2n_{1}<n_{2} and d(n1)=d(n2)=1d(n_{1})=d(n_{2})=1.

Let o(n)o(n) (resp. o(n¯)o({\underline{n}})) denote the number of isomorphism classes of superspecial abelian surfaces over 𝔽q\mathbb{F}_{q} that are isogenous to Xn2/d(n)X_{n}^{2/d(n)} (resp. Xn¯X_{\underline{n}}). It was shown in [28, §3.2] that

(1.1) o(1)=o(2)=1,o(3)=o(6),o(5)=o(10);\displaystyle o(1)=o(2)=1,\ o(3)=o(6),\ o(5)=o(10);
(1.2) o(1,3)=o(2,6),o(1,4)=o(2,4),o(1,6)=o(2,3),o(3,4)=o(4,6).\displaystyle o(1,3)=o(2,6),\ o(1,4)=o(2,4),\ o(1,6)=o(2,3),\ o(3,4)=o(4,6).

Thus we have

(1.3) |SSp2(𝔽q)|=2+2o(3)+o(4)+2o(5)+o(8)+o(12)+o(1,2)+2o(2,3)+2o(2,4)+2o(2,6)+2o(3,4)+o(3,6).\begin{split}\lvert\operatorname{SSp}_{2}(\mathbb{F}_{q})\rvert=&2+2o(3)+o(4)+2o(5)+o(8)+o(12)\\ &+o(1,2)+2o(2,3)+2o(2,4)+2o(2,6)+2o(3,4)+o(3,6).\end{split}

As mentioned before, the value of each o(n)o(n) or o(n¯)o({\underline{n}}) in (1.3) has been worked out in [28] conditionally on pp. To make explicit this condition, we uniformize the notation. For each rr\in{\mathbb{N}}, let us denote

˘r:={n¯=(n1,,nr)r0<n1<<nr}.\breve{\mathbb{N}}^{r}:=\{{\underline{n}}=(n_{1},\cdots,n_{r})\in{\mathbb{N}}^{r}\mid 0<n_{1}<\cdots<n_{r}\}.

In particular, if r=1r=1, then ˘r=\breve{\mathbb{N}}^{r}={\mathbb{N}} and we drop the underline from n¯{\underline{n}}. For each n¯˘r{\underline{n}}\in\breve{\mathbb{N}}^{r} with rr arbitrary, we define

(1.4) An¯:=[T](i=1rΦni(T)),Kn¯:=[T](i=1rΦni(T))i=1r(ζni),A_{\underline{n}}:=\frac{\mathbb{Z}[T]}{\left(\prod_{i=1}^{r}\Phi_{n_{i}}(T)\right)},\qquad K_{\underline{n}}:=\frac{\mathbb{Q}[T]}{\left(\prod_{i=1}^{r}\Phi_{n_{i}}(T)\right)}\simeq\prod_{i=1}^{r}\mathbb{Q}(\zeta_{n_{i}}),

where Φn(T)[T]\Phi_{n}(T)\in\mathbb{Z}[T] is the nn-th cyclotomic polynomial. Clearly, An¯A_{\underline{n}} is a \mathbb{Z}-order in Kn¯K_{\underline{n}}, so it is contained in the unique maximal order OKn¯:=i=1r[T]/(Φni(T))O_{K_{\underline{n}}}:=\prod_{i=1}^{r}\mathbb{Z}[T]/(\Phi_{n_{i}}(T)). Let n¯˘r{\underline{n}}\in\breve{\mathbb{N}}^{r} with r{1,2}r\in\{1,2\} be an rr-tuple appearing in (1.3). In the proof of [28, Theorem 3.3], it is assumed that

(1.5) An¯,p:=An¯pis an étale p-order.A_{{\underline{n}},p}:=A_{\underline{n}}\otimes\mathbb{Z}_{p}\quad\text{is an \'{e}tale $\mathbb{Z}_{p}$-order}.

This condition fails precisely in the following two situations:

  1. (C1)

    pp is ramified in (ζni)\mathbb{Q}(\zeta_{n_{i}}) for some 1ir1\leq i\leq r, or

  2. (C2)

    pp divides the index [OKn¯:An¯][O_{K_{\underline{n}}}:A_{\underline{n}}].

If r=1r=1, then AnA_{n} coincides with OKnO_{K_{n}}, so (C2) is possible only if r=2r=2. For the reader’s convenience, we list the indices i(n¯):=[OKn¯:An¯]i({\underline{n}}):=[O_{K_{\underline{n}}}:A_{\underline{n}}] from [28, Table 1]:

n¯{\underline{n}} (1,2)(1,2) (2,3)(2,3) (2,4)(2,4) (2,6)(2,6) (3,4)(3,4) (3,6)(3,6)
i(n¯)i({\underline{n}}) 22 11 22 33 11 44
Theorem 1.1.

Let n¯˘r{\underline{n}}\in\breve{\mathbb{N}}^{r} be an rr-tuple appearing in (1.3), and pp be a prime satisfying (C1) or (C2). Then the values of o(n¯)o({\underline{n}}) for each pp are given by the following table

n¯{\underline{n}} 33 44 55 88 1212 (1,2)(1,2) (2,3)(2,3) (2,4)(2,4) (2,6)(2,6) (3,4)(3,4) (3,6)(3,6)
pp 33 22 55 22 2,32,3 22 33 22 33 2,32,3 22 33
o(n¯)o({\underline{n}}) 22 22 11 11 33 33 11 22 33 22 88 22

Moreover, the number of isomorphism classes of superspecial abelian surfaces over a finite field 𝔽q\mathbb{F}_{q} of even degree over 𝔽p\mathbb{F}_{p} with p{2,3,5}p\in\{2,3,5\} is given by

(1.6) |SSp2(𝔽q)|={49if p=2,45if p=3,47if p=5.\lvert\operatorname{SSp}_{2}(\mathbb{F}_{q})\rvert=\begin{cases}49&\text{if }p=2,\\ 45&\text{if }p=3,\\ 47&\text{if }p=5.\\ \end{cases}
Remark 1.2.

We provide an arithmetic interpretation of the values o(n¯)o({\underline{n}}). Let D=Dp,D=D_{p,\infty} be the unique quaternion \mathbb{Q}-algebra up to isomorphism ramified precisely at pp and \infty, and Mat2(D)\operatorname{Mat}_{2}(D) be the algebra of 2×22\times 2 matrices over DD. Fix a maximal \mathbb{Z}-order 𝒪{\mathcal{O}} in DD. As explained in [28, §1, p. 304], up to isomorphism, the arithmetic group GL2(𝒪)\operatorname{GL}_{2}({\mathcal{O}}) depends only on pp and not on the choice of 𝒪{\mathcal{O}}. An element xGL2(𝒪)x\in\operatorname{GL}_{2}({\mathcal{O}}) of finite group order111Unfortunately, the word “order” plays double duties in this paragraph: for the order inside an algebra and also for the order of a group element. To make a distinction, we always insert the word “group” when the second meaning applies. is semisimple, so its minimal polynomial over \mathbb{Q} in Mat2(D)\operatorname{Mat}_{2}(D) is of the form Pn¯(T):=i=1rΦni(T)P_{{\underline{n}}}(T):=\prod_{i=1}^{r}\Phi_{n_{i}}(T) for some n¯=(n1,,nr)˘r{\underline{n}}=(n_{1},\cdots,n_{r})\in\breve{\mathbb{N}}^{r}. It is not hard to show that r2r\leq 2 (see [28, §3.1]). A Galois cohomological argument shows that o(n¯)o({\underline{n}}) counts the number of conjugacy classes of elements of GL2(𝒪)\operatorname{GL}_{2}({\mathcal{O}}) with minimal polynomial Pn¯(T)P_{{\underline{n}}}(T), and |SSp2(𝔽q)|\lvert\operatorname{SSp}_{2}(\mathbb{F}_{q})\rvert is equal to the total number of conjugacy classes of elements of finite group order in GL2(𝒪)\operatorname{GL}_{2}({\mathcal{O}}) (see [28, Proposition 1.1]). Actually, this arithmetic interpretation works for GLd(𝒪)\operatorname{GL}_{d}({\mathcal{O}}) with any d2d\geq 2, not just for d=2d=2.

The proof of Theorem 1.1 will occupy the remaining part of the paper. In Section 2, we recall from [28, §3.1] the general strategy for computing o(n¯)o({\underline{n}}). The isotypic case (i.e. r=1r=1) will be treated in Section 3, and the non-isotypic case (i.e. r=2r=2) will be treated in Section 4.

2. General strategy for computing o(n¯)o({\underline{n}})

Keep the notation and the assumptions of the previous section. We recall from [28, §3.1] and [26, §6.4] the general strategy for calculating o(n¯)o({\underline{n}}) with n¯˘r{\underline{n}}\in\breve{\mathbb{N}}^{r} for r2r\leq 2. Based on the arithmetic interpretation of o(n¯)o({\underline{n}}) in Remark 1.2, we further provide a lattice description of o(n¯)o({\underline{n}}). Indeed, it is via this lattice description that the value of each o(n¯)o({\underline{n}}) is calculated.

Let V=D2V=D^{2} be the unique simple left Mat2(D)\operatorname{Mat}_{2}(D)-module, which is at the same time a right DD-vector space of dimension 22. Let M0:=𝒪2M_{0}:={\mathcal{O}}^{2} be the standard right 𝒪{\mathcal{O}}-lattice in VV, whose endomorphism ring End𝒪(M0)\operatorname{End}_{{\mathcal{O}}}(M_{0}) is just Mat2(𝒪)\operatorname{Mat}_{2}({\mathcal{O}}). For each element xGL2(𝒪)x\in\operatorname{GL}_{2}({\mathcal{O}}) of finite group order with minimal polynomial Pn¯(T)[T]P_{{\underline{n}}}(T)\in\mathbb{Z}[T], there is a canonical embedding An¯=[T]/(Pn¯(T))Mat2(𝒪)A_{\underline{n}}=\mathbb{Z}[T]/(P_{\underline{n}}(T))\hookrightarrow\operatorname{Mat}_{2}({\mathcal{O}}) sending TT to xx. This embedding equips M0M_{0} with an (An¯,𝒪)(A_{\underline{n}},{\mathcal{O}})-bimodule structure, or equivalently, a faithful left An¯𝒪oppA_{\underline{n}}\otimes_{\mathbb{Z}}{\mathcal{O}}^{\mathrm{opp}}-module structure. Similarly, VV is equipped with a faithful left Kn¯DoppK_{\underline{n}}\otimes_{\mathbb{Q}}D^{\mathrm{opp}}-module structure. The canonical involution induces an isomorphism between the opposite ring 𝒪opp{\mathcal{O}}^{\mathrm{opp}} and 𝒪{\mathcal{O}} itself (and similarly between DoppD^{\mathrm{opp}} and DD), so we put

(2.1) 𝒜n¯:=An¯𝒪,and𝒦n¯:=Kn¯D.{\mathscr{A}}_{\underline{n}}:=A_{\underline{n}}\otimes_{\mathbb{Z}}{\mathcal{O}},\quad\text{and}\quad{\mathscr{K}}_{\underline{n}}:=K_{\underline{n}}\otimes_{\mathbb{Q}}D.

Clearly, 𝒜n¯{\mathscr{A}}_{\underline{n}} is a \mathbb{Z}-order in the semisimple \mathbb{Q}-algebra 𝒦n¯{\mathscr{K}}_{\underline{n}}. It has been shown in [28, p. 309] that the 𝒦n¯{\mathscr{K}}_{\underline{n}}-module structure on VV is uniquely determined by the rr-tuple n¯{\underline{n}}. From [26, Theorem 6.11] (see also [28, Lemma 3.1]), the above construction induces a bijection between the following two finite sets:

(2.2) {conjugacy classes of elements of GL2(𝒪) with minimal polynomial Pn¯(x)}{isomorphism classes of 𝒜n¯-lattices in the left 𝒦n¯-module V}\left\{\parbox{122.85876pt}{conjugacy classes of elements of $\operatorname{GL}_{2}({\mathcal{O}})$ with minimal polynomial $P_{\underline{n}}(x)$}\right\}\longleftrightarrow\left\{\parbox{86.72377pt}{isomorphism classes of ${\mathscr{A}}_{\underline{n}}$-lattices in the left ${\mathscr{K}}_{\underline{n}}$-module $V$}\right\}

Therefore, we have o(n¯)=|(n¯)|o({\underline{n}})=\lvert{\mathscr{L}}({\underline{n}})\rvert, where (n¯){\mathscr{L}}({\underline{n}}) denote the set on the right.

Now fix a pair (n¯,p)({\underline{n}},p) in Theorem 1.1 and in turn a left 𝒦n¯{\mathscr{K}}_{\underline{n}}-module VV. Given an 𝒜n¯{\mathscr{A}}_{\underline{n}}-lattices ΛV\Lambda\subset V, we write [Λ][\Lambda] for its isomorphism class, and OΛO_{\Lambda} for its endomorphism ring End𝒜n¯(Λ)End𝒦n¯(V)\operatorname{End}_{{\mathscr{A}}_{\underline{n}}}(\Lambda)\subset\operatorname{End}_{{\mathscr{K}}_{\underline{n}}}(V). As a convention, the endomorphism algebra n¯:=End𝒦n¯(V){\mathscr{E}}_{\underline{n}}:=\operatorname{End}_{{\mathscr{K}}_{\underline{n}}}(V) acts on VV from the left, so it coincides with the centralizer of Kn¯K_{\underline{n}} in Mat2(D)\operatorname{Mat}_{2}(D). Two 𝒜n¯{\mathscr{A}}_{\underline{n}}-lattices Λ1\Lambda_{1} and Λ2\Lambda_{2} in VV are isomorphic if and only if there exists gn¯×g\in{\mathscr{E}}_{\underline{n}}^{\times} such that Λ1=gΛ2\Lambda_{1}=g\Lambda_{2}.

For each prime \ell\in{\mathbb{N}}, we use the subscript to indicate \ell-adic completion. For example, 𝒜n¯,{\mathscr{A}}_{{\underline{n}},\ell} (the \ell-adic completion of 𝒜n¯{\mathscr{A}}_{\underline{n}}) is a \mathbb{Z}_{\ell}-order in the semisimple \mathbb{Q}_{\ell}-algebra 𝒦n¯,{\mathscr{K}}_{{\underline{n}},\ell}, and Λ\Lambda_{\ell} is an 𝒜n¯,{\mathscr{A}}_{{\underline{n}},\ell}-lattice in VV_{\ell}. For each prime \ell, let (n¯){\mathscr{L}}_{\ell}({\underline{n}}) denote the set of isomorphism classes of 𝒜n¯,{\mathscr{A}}_{{\underline{n}},\ell}-lattices in the left 𝒦n¯,{\mathscr{K}}_{{\underline{n}},\ell}-module VV_{\ell}. For almost all primes \ell, the \mathbb{Z}_{\ell}-order 𝒜n¯,{\mathscr{A}}_{{\underline{n}},\ell} is maximal in 𝒦n¯,{\mathscr{K}}_{{\underline{n}},\ell}, in which case both of the following hold by [6, Theorem 26.24]:

  1. (i)

    Λ\Lambda_{\ell} is uniquely determined up to isomorphism (i.e. |(n¯)|=1\lvert{\mathscr{L}}_{\ell}({\underline{n}})\rvert=1), and

  2. (ii)

    OΛ,O_{\Lambda,\ell} is maximal in n¯,{\mathscr{E}}_{{\underline{n}},\ell}.

Let S(n¯,p)S({\underline{n}},p) be the finite set of primes \ell for which 𝒜n¯,{\mathscr{A}}_{{\underline{n}},\ell} is non-maximal. The profinite completion ΛΛ^:=Λ\Lambda\mapsto\widehat{\Lambda}:=\prod_{\ell}\Lambda_{\ell} induces a surjective map

(2.3) Ψ:(n¯)(n¯)S(n¯,p)(n¯).\Psi:{\mathscr{L}}({\underline{n}})\to\prod_{\ell}{\mathscr{L}}_{\ell}({\underline{n}})\simeq\prod_{\ell\in S({\underline{n}},p)}{\mathscr{L}}_{\ell}({\underline{n}}).

Two 𝒜n¯{\mathscr{A}}_{\underline{n}}-lattices Λ1\Lambda_{1} and Λ2\Lambda_{2} in VV are said to be in the same genus if Ψ([Λ1])=Ψ([Λ2])\Psi([\Lambda_{1}])=\Psi([\Lambda_{2}]), or equivalently, (Λ1)(Λ2)(\Lambda_{1})_{\ell}\simeq(\Lambda_{2})_{\ell} for every prime \ell. The fibers of Ψ\Psi partition (n¯){\mathscr{L}}({\underline{n}}) into a disjoint union of genera. Let 𝒢(Λ):=Ψ1(Ψ([Λ]))(n¯){\mathscr{G}}(\Lambda):=\Psi^{-1}(\Psi([\Lambda]))\subseteq{\mathscr{L}}({\underline{n}}) be the fiber of Ψ\Psi over Ψ([Λ])\Psi([\Lambda]), that is, the set of isomorphism classes of 𝒜n¯{\mathscr{A}}_{\underline{n}}-lattices in the genus of Λ\Lambda. From [21, Proposition 1.4], we have

(2.4) |𝒢(Λ)|=h(OΛ),\lvert{\mathscr{G}}(\Lambda)\rvert=h(O_{\Lambda}),

where h(OΛ)h(O_{\Lambda}) denote the class number of OΛO_{\Lambda}. In other words, h(OΛ)h(O_{\Lambda}) is the number of locally principal right (or equivalently, left) ideal classes of OΛO_{\Lambda}.

Therefore, the computation of o(n¯)o({\underline{n}}) can be carried out in the following two steps:

  1. (Step 1)

    Classify the genera of 𝒜n¯{\mathscr{A}}_{{\underline{n}}}-lattices in the left 𝒦n¯{\mathscr{K}}_{{\underline{n}}}-module VV. Equivalently, classify the isomorphism classes of 𝒜n¯,{\mathscr{A}}_{{\underline{n}},\ell}-lattices in VV_{\ell} for each S(n¯,p)\ell\in S({\underline{n}},p).

  2. (Step 2)

    Pick a lattice Λ\Lambda in each genus and write down its endomorphism ring OΛO_{\Lambda} (at least locally at each prime \ell). The number o(n¯)o({\underline{n}}) is obtained by summing up the class numbers h(OΛ)h(O_{\Lambda}) over all genera.

Remark 2.1.

The reason that condition (1.5) is assumed throughout the calculations in [28] is to make sure that the p\mathbb{Z}_{p}-order 𝒜n¯,p{\mathscr{A}}_{{\underline{n}},p} is a product of Eichler orders [28, Remark 3.7]. In our setting, pp satisfies condition (C1) or (C2), so 𝒜n¯,p{\mathscr{A}}_{{\underline{n}},p} becomes more complicated. This is precisely why the primes p{2,3,5}p\in\{2,3,5\} are treated separately from the rest of the primes. Luckily for us, many 𝒜n¯,p{\mathscr{A}}_{{\underline{n}},p} turn out to be Bass orders (see Definition 3.1 below), which makes the classification of 𝒜n¯,p{\mathscr{A}}_{{\underline{n}},p}-lattices more manageable.

3. The isotypic case

In this section, we calculate the values of o(n)o(n) for n{3,4,5,8,12}n\in\{3,4,5,8,12\} and p|np|n. Keep the notation of previous sections. In particular, D=Dp,D=D_{p,\infty} is the unique quaternion \mathbb{Q}-algebra ramified precisely at pp and \infty, and 𝒪{\mathcal{O}} is a maximal \mathbb{Z}-order in DD. Since AnA_{n} is the maximal order in the nn-th cyclotomic field KnK_{n}, and 𝒪{\mathcal{O}} has reduced discriminant pp, we have S(n,p)={p}S(n,p)=\{p\}. In other words, the \ell-adic completion 𝒜n,{\mathscr{A}}_{n,\ell} is non-maximal in 𝒦n,{\mathscr{K}}_{n,\ell} if and only if =p\ell=p. It turns out that 𝒜n,p{\mathscr{A}}_{n,p} is always a Bass order in the quaternion Kn,pK_{n,p}-algebra 𝒦n,p{\mathscr{K}}_{n,p}. Therefore, the classification of genera of the lattice set (n){\mathscr{L}}(n) is then reduced to the classification of lattices over local quaternion Bass orders.

3.1. Classification of lattices over local quaternion Bass orders

The main references for this section are [2, 4] and [6, §37]. Let FF be a nonarchimedean local field, and OFO_{F} be its ring of integers. Fix a uniformizer ϖ\varpi of FF and denote the the finite residue field OF/ϖOFO_{F}/\varpi O_{F} by 𝔨{\mathfrak{k}}. Let BB be a finite dimensional separable FF-algebra [6, Definition 7.1 and Corollary 7.6], and 𝒪{\mathscr{O}} be an OFO_{F}-order (of full rank) in BB. We write Ov(𝒪)\operatorname{Ov}({\mathscr{O}}) for the finite set of overorders of 𝒪{\mathscr{O}}, i.e. OFO_{F}-orders in BB containing 𝒪{\mathscr{O}}. A minimal overorder of 𝒪{\mathscr{O}} is a minimal member of Ov(𝒪){𝒪}\operatorname{Ov}({\mathscr{O}})\smallsetminus\{{\mathscr{O}}\} with respect to inclusion.

Definition 3.1.

An OFO_{F}-order 𝒪{\mathscr{O}} in BB is Gorenstein if its dual lattice 𝒪:=HomOF(𝒪,OF){\mathscr{O}}^{\vee}:=\operatorname{Hom}_{O_{F}}({\mathscr{O}},O_{F}) is projective as a left (or right) 𝒪{\mathscr{O}}-module. It is called a Bass order if every member of Ov(𝒪)\operatorname{Ov}({\mathscr{O}}) is Gorenstein. It is called a hereditary order if every left ideal of 𝒪{\mathscr{O}} is projective as a left 𝒪{\mathscr{O}}-module. If 𝒪{\mathscr{O}} is the intersection of two maximal orders, then it is called an Eichler order.

We have the following inclusions of orders:

(maximal)(herediary)(Eichler)(Bass)(Gorenstein).\text{(maximal)}\subset\text{(herediary)}\subset\text{(Eichler)}\subset\text{(Bass)}\subset\text{(Gorenstein)}.

If BB is division, then Eichler orders are also maximal. Bass notes in [1] that Gorenstein orders are ubiquitous.

Let II be a fractional left 𝒪{\mathscr{O}}-ideal (of full rank) in BB. We say II is proper over 𝒪{\mathscr{O}} if its associated left order Ol(I):={xBxII}O_{l}(I):=\{x\in B\mid xI\subseteq I\} coincides with 𝒪{\mathscr{O}}. From [3, Example 2.6 and Corollary 2.7], the following lemma provides an equivalent characterization of Gorenstein orders in certain types of FF-algebras:

Lemma 3.2.

Suppose that BB is either a commutative algebra or a quaternion FF-algebra. Then 𝒪{\mathscr{O}} is Gorenstein if and only if every proper fractional left 𝒪{\mathscr{O}}-ideal IBI\subset B is principal (i.e. there exists xB×x\in B^{\times} such that I=𝒪xI={\mathscr{O}}x).

In the quaternion case, the above lemma can also be obtained by combining [9, Condition G4 or G4’, p. 1364] and [12, Theorem 1]. The lemma no longer holds in general for orders in more complicated algebras. If every proper fractional left 𝒪{\mathscr{O}}-ideal II is principal, then 𝒪{\mathscr{O}} is Gorenstein, but the converse is not necessarily true. See [12, p. 220] and [3, p. 535] for some examples. Nevertheless, a proper fractional left ideal over a Gorenstein order is always left projective according to [18, Theorem 5.3, pp. 253–255]. However, unlike the situation over commutative rings, a projective module over a non-commutative ring may not be locally free. Brzezinski [3, Proposition 2.3] gave a precise characterization of the orders 𝒪{\mathscr{O}} such that every proper fractional left 𝒪{\mathscr{O}}-ideal IBI\subset B is principal (Such orders are called strongly Gorenstein by him).

For the rest of Section 3.1, we assume that char(F)2\operatorname{char}(F)\neq 2 and BB is a quaternion FF-algebra. The reduced trace and reduced norm maps of BB are denoted by Tr:BF\operatorname{Tr}:B\to F and Nr:BF\operatorname{Nr}:B\to F respectively. We write 𝔡(𝒪){\mathfrak{d}}({\mathscr{O}}) for the reduced discriminant of 𝒪{\mathscr{O}}, which is a nonzero integral ideal of OFO_{F}. From [2, Proposition 1.2], 𝒪{\mathscr{O}} is hereditary if and only if 𝔡(𝒪){\mathfrak{d}}({\mathscr{O}}) is square-free. Thus if BB is division, then 𝒪{\mathscr{O}} is hereditary if and only if 𝒪{\mathscr{O}} is the unique maximal order of BB; if BMat2(F)B\simeq\operatorname{Mat}_{2}(F), then 𝒪{\mathscr{O}} is hereditary if and only if 𝒪{\mathscr{O}} is isomorphic to Mat2(OF)\operatorname{Mat}_{2}(O_{F}) or [OFOFϖOFOF]\left[\begin{smallmatrix}O_{F}&O_{F}\\ \varpi O_{F}&O_{F}\end{smallmatrix}\right].

Theorem 3.3.

The following are equivalent:

  1. (a)

    every left 𝒪{\mathscr{O}}-ideal is generated by at most 2 elements;

  2. (b)

    𝒪{\mathscr{O}} is Bass;

  3. (c)

    every indecomposable 𝒪{\mathscr{O}}-lattice is isomorphic to an ideal of 𝒪{\mathscr{O}};

  4. (d)

    𝒪OL{\mathscr{O}}\supseteq O_{L} for some semisimple quadratic FF-subalgebra LBL\subseteq B.

Indeed, the implications (a)(b)(c)(a)\Rightarrow(b)\Rightarrow(c) hold in much more general settings according to [6, §37]. The implication (c)(a)(c)\Rightarrow(a) is proved by Drozd, Kirichenko and Roiter [9] (see [6, p. 790]). Lastly, the equivalence (b)(d)(b)\Leftrightarrow(d) is proved by Brzezinski [4, Proposition 1.12]. See Chari et al. [5] for more characterization of quaternion Bass orders.

We recall the notion of Eichler invariant following [2, Definition 1.8].

Definition 3.4.

Let 𝔨/𝔨{\mathfrak{k}}^{\prime}/{\mathfrak{k}} be the unique quadratic field extension. When 𝒪≄Mat2(OF){\mathscr{O}}\not\simeq\operatorname{Mat}_{2}(O_{F}), the quotient of 𝒪{\mathscr{O}} by its Jacobson radical 𝔍(𝒪){\mathfrak{J}}({\mathscr{O}}) falls into the following three cases:

𝒪/𝔍(𝒪)𝔨×𝔨,𝔨,or𝔨,{\mathscr{O}}/{\mathfrak{J}}({\mathscr{O}})\simeq{\mathfrak{k}}\times{\mathfrak{k}},\qquad{\mathfrak{k}},\quad\text{or}\quad{\mathfrak{k}}^{\prime},

and the Eichler invariant e(𝒪)e({\mathscr{O}}) is defined to be 1,0,11,0,-1 accordingly. As a convention, if 𝒪Mat2(OF){\mathscr{O}}\simeq\operatorname{Mat}_{2}(O_{F}), then its Eichler invariant is defined to be 22.

For example, if BB is division and 𝒪{\mathscr{O}} is the unique maximal order, then e(𝒪)=1e({\mathscr{O}})=-1. It is shown in [2, Proposition 2.1] that e(𝒪)=1e({\mathscr{O}})=1 if and only if 𝒪{\mathscr{O}} is a non-maximal Eichler order. Note that e(𝒪)=1e({\mathscr{O}})=1 can only occur when BMat2(F)B\simeq\operatorname{Mat}_{2}(F). Moreover, if e(𝒪)0e({\mathscr{O}})\neq 0, then 𝒪{\mathscr{O}} is automatically Bass by [2, Corollary 2.4 and Propoisition 3.1]. The classification of lattices over Eichler orders is well known (see [28, p. 315] for example), which we recall as follows.

Lemma 3.5.

Suppose BMat2(F)B\simeq\operatorname{Mat}_{2}(F) and let 𝒪[OFOFϖeOFOF]{\mathscr{O}}\simeq\begin{bmatrix}O_{F}&O_{F}\\ \varpi^{e}O_{F}&O_{F}\end{bmatrix} be an Eichler order. Let MM be an 𝒪{\mathscr{O}}-lattice in a finite left BB-module WW. Then

(3.1) W[FF]u,andMi=1u[OFϖeiOF],W\simeq\begin{bmatrix}F\\ F\end{bmatrix}^{\oplus u},\quad\text{and}\quad M\simeq\bigoplus_{i=1}^{u}\begin{bmatrix}O_{F}\\ \varpi^{e_{i}}O_{F}\end{bmatrix},

where the eie_{i}’s are integers such that 0eie0\leq e_{i}\leq e and eiei+1e_{i}\leq e_{i+1} for all ii. Moreover, the isomorphism class of MM is uniquely determined by these eie_{i}’s.

Henceforth we assume that e(𝒪){0,1}e({\mathscr{O}})\in\{0,-1\}. Let n(𝒪)n({\mathscr{O}}) be the unique non-negative integer such that 𝔡(𝒪)=(ϖn(𝒪)){\mathfrak{d}}({\mathscr{O}})=(\varpi^{n({\mathscr{O}})}). Suppose that 𝒪{\mathscr{O}} is Bass but non-hereditary. From [2, Proposition 1.12], 𝒪{\mathscr{O}} has a unique minimal overorder (𝒪){\mathcal{M}}({\mathscr{O}}), which is also Bass by definition. According to [2, Propositions 3.1 and 4.1],

(3.2) n((𝒪))={n(𝒪)2if e(𝒪)=1,n(𝒪)1if e(𝒪)=0,n({\mathcal{M}}({\mathscr{O}}))=\begin{cases}n({\mathscr{O}})-2&\text{if }e({\mathscr{O}})=-1,\\ n({\mathscr{O}})-1&\text{if }e({\mathscr{O}})=0,\end{cases}

and e((𝒪))=e(𝒪)e({\mathcal{M}}({\mathscr{O}}))=e({\mathscr{O}}) if (𝒪){\mathcal{M}}({\mathscr{O}}) is also non-hereditary. Thus starting from 0(𝒪):=𝒪{\mathcal{M}}^{0}({\mathscr{O}}):={\mathscr{O}}, we define i(𝒪):=(i1(𝒪)){\mathcal{M}}^{i}({\mathscr{O}}):={\mathcal{M}}({\mathcal{M}}^{i-1}({\mathscr{O}})) recursively to obtain a unique chain of Bass orders terminating at a hereditary order m(𝒪){\mathcal{M}}^{m}({\mathscr{O}}):

(3.3) 𝒪=0(𝒪)1(𝒪)2(𝒪)m1(𝒪)m(𝒪),{\mathscr{O}}={\mathcal{M}}^{0}({\mathscr{O}})\subset{\mathcal{M}}^{1}({\mathscr{O}})\subset{\mathcal{M}}^{2}({\mathscr{O}})\subset\cdots\subset{\mathcal{M}}^{m-1}({\mathscr{O}})\subset{\mathcal{M}}^{m}({\mathscr{O}}),

where mm is given as follows

  • m=n(𝒪)1m=n({\mathscr{O}})-1 if e(𝒪)=0e({\mathscr{O}})=0; and

  • m=n(𝒪)/2m=\left\lfloor n({\mathscr{O}})/2\right\rfloor if e(𝒪)=1e({\mathscr{O}})=-1, where xxx\mapsto\left\lfloor x\right\rfloor is the floor function on \mathbb{R}.

The order m(𝒪){\mathcal{M}}^{m}({\mathscr{O}}) is called the hereditary closure of 𝒪{\mathscr{O}} and will henceforth be denoted by (𝒪){\mathcal{H}}({\mathscr{O}}). If e(𝒪)=1e({\mathscr{O}})=-1, then (𝒪){\mathcal{H}}({\mathscr{O}}) is always a maximal order by [2, Proposition 3.1]. Thus when e(𝒪)=1e({\mathscr{O}})=-1, n(𝒪)n({\mathscr{O}}) is even if BMat2(F)B\simeq\operatorname{Mat}_{2}(F), and n(𝒪)n({\mathscr{O}}) is odd if BB is division. If e(𝒪)=0e({\mathscr{O}})=0, then

  • (𝒪)[OFOFϖOFOF]{\mathcal{H}}({\mathscr{O}})\simeq\left[\begin{smallmatrix}O_{F}&O_{F}\\ \varpi O_{F}&O_{F}\end{smallmatrix}\right] if BMat2(F)B\simeq\operatorname{Mat}_{2}(F), and

  • (𝒪){\mathcal{H}}({\mathscr{O}}) is the unique maximal order if BB is division.

Note that 𝒪{\mathscr{O}} is hereditary (i.e. m=0m=0) if and only if e(𝒪)=1e({\mathscr{O}})=-1 and BB is division, so mm is strictly positive in the remaining cases.

From [4, Proposition 1.12], there exists a quadratic field extension L/FL/F such that OLO_{L} embeds222From the proof of [4, Theorem 3.3 and 3.10], any two embeddings of OLO_{L} into 𝒪{\mathscr{O}} are conjugate by an element of the normalizer of 𝒪{\mathscr{O}}, thus expression (3.4) does not depend on the choice of the embedding OL𝒪O_{L}\hookrightarrow{\mathscr{O}}. into 𝒪{\mathscr{O}}, and

(3.4) 𝒪\displaystyle{\mathscr{O}} =OL+𝔍((𝒪))c,where\displaystyle=O_{L}+{\mathfrak{J}}({\mathcal{H}}({\mathscr{O}}))^{c},\qquad\text{where}
(3.5) c\displaystyle c ={n(𝒪)/2if e(𝒪)=1 and BMat2(F),n(𝒪)1otherwise.\displaystyle=\begin{cases}n({\mathscr{O}})/2&\text{if }e({\mathscr{O}})=-1\text{ and }B\simeq\operatorname{Mat}_{2}(F),\\ n({\mathscr{O}})-1&\text{otherwise.}\end{cases}

In fact, L/FL/F is the unique unramified quadratic field extension if e(𝒪)=1e({\mathscr{O}})=-1, and it is a ramified quadratic field extension if e(𝒪)=0e({\mathscr{O}})=0. In the latter case, the ramified quadratic extension L/FL/F can be arbitrary if n(𝒪)=2n({\mathscr{O}})=2 according to [4, (3.14)]; and it is uniquely determined by 𝒪{\mathscr{O}} if n(𝒪)3n({\mathscr{O}})\geq 3 and FF is nondyadic according to [17, Lemma 3.5].

Lemma 3.6.

Suppose that 𝒪B{\mathscr{O}}\subset B is a Bass order with e(𝒪){0,1}e({\mathscr{O}})\in\{0,-1\}. Let NN be an indecomposable left 𝒪{\mathscr{O}}-lattice.

  1. (1)

    If BB is division, then

    (3.6) Ni(𝒪)for some 0im.N\simeq{\mathcal{M}}^{i}({\mathscr{O}})\qquad\text{for some }0\leq i\leq m.
  2. (2)

    Suppose that BB is split, i.e. BMat2(F)B\simeq\operatorname{Mat}_{2}(F). Fix an identification of (𝒪){\mathcal{H}}({\mathscr{O}}) with [OFOFϖOFOF]\left[\begin{smallmatrix}O_{F}&O_{F}\\ \varpi O_{F}&O_{F}\end{smallmatrix}\right] (resp. Mat2(OF)\operatorname{Mat}_{2}(O_{F})) if e(𝒪)=0e({\mathscr{O}})=0 (resp. 1-1).

    1. (a)

      If e(𝒪)=0e({\mathscr{O}})=0, then NN is isomorphic to one of the following 𝒪{\mathscr{O}}-lattices:

      (3.7) [OFϖOF],[OFOF],ori(𝒪)with0im1.\begin{bmatrix}O_{F}\\ \varpi O_{F}\end{bmatrix},\quad\begin{bmatrix}O_{F}\\ O_{F}\end{bmatrix},\quad\text{or}\quad{\mathcal{M}}^{i}({\mathscr{O}})\quad\text{with}\quad 0\leq i\leq m-1.
    2. (b)

      If e(𝒪)=1e({\mathscr{O}})=-1, then NN is isomorphic to one of the following 𝒪{\mathscr{O}}-lattices:

      (3.8) [OFOF]ori(𝒪)with0im1.\begin{bmatrix}O_{F}\\ O_{F}\end{bmatrix}\quad\text{or}\quad{\mathcal{M}}^{i}({\mathscr{O}})\quad\text{with}\quad 0\leq i\leq m-1.
Proof.

According to the Drozd-Krichenko-Roiter Theorem [6, Theorem 37.16],

(3.9) NOFF{Bif B is division,F2 or Mat2(F)if BMat2(F).N\otimes_{O_{F}}F\simeq\begin{cases}B&\text{if $B$ is division},\\ F^{2}\text{ or }\operatorname{Mat}_{2}(F)&\text{if }B\simeq\operatorname{Mat}_{2}(F).\end{cases}

First, suppose that BMat2(F)B\simeq\operatorname{Mat}_{2}(F) and NOFFF2N\otimes_{O_{F}}F\simeq F^{2}. Then NOF2N\simeq O_{F}^{2} as an OFO_{F}-module, and EndOF(N)\operatorname{End}_{O_{F}}(N) is a maximal order in BB containing 𝒪{\mathscr{O}}. It follows from (3.3) that EndOF(N)\operatorname{End}_{O_{F}}(N) contains (𝒪){\mathcal{H}}({\mathscr{O}}), which equips NN with a canonical (𝒪){\mathcal{H}}({\mathscr{O}})-module structure. Therefore, if e(𝒪)=1e({\mathscr{O}})=-1, then (𝒪)=Mat2(OF){\mathcal{H}}({\mathscr{O}})=\operatorname{Mat}_{2}(O_{F}), and hence NN is homothetic to [OFOF]\left[\begin{smallmatrix}O_{F}\\ O_{F}\end{smallmatrix}\right]. Similarly, if e(𝒪)=0e({\mathscr{O}})=0, then (𝒪)=[OFOFϖOFOF]{\mathcal{H}}({\mathscr{O}})=\left[\begin{smallmatrix}O_{F}&O_{F}\\ \varpi O_{F}&O_{F}\end{smallmatrix}\right], and hence NN is homothetic to [OFOF]\left[\begin{smallmatrix}O_{F}\\ O_{F}\end{smallmatrix}\right] or [OFϖOF]\left[\begin{smallmatrix}O_{F}\\ \varpi O_{F}\end{smallmatrix}\right].

Next, suppose that NOFFBN\otimes_{O_{F}}F\simeq B. Then we regard NN as a fractional left ideal of 𝒪{\mathscr{O}}. Let Ol(N)={xBxNN}O_{l}(N)=\{x\in B\mid xN\subseteq N\} be the associated left order of NN. Clearly, Ol(N)O_{l}(N) contains 𝒪{\mathscr{O}}, so Ol(N)=i(𝒪)O_{l}(N)={\mathcal{M}}^{i}({\mathscr{O}}) for some 0im0\leq i\leq m. In particular, Ol(N)O_{l}(N) is Gorenstein. It follows from Lemma 3.2 that NOl(N)N\simeq O_{l}(N) as 𝒪{\mathscr{O}}-lattices.

Clearly, if BB is division, then i(𝒪){\mathcal{M}}^{i}({\mathscr{O}}) is indecomposable for every 0im0\leq i\leq m. On the other hand, if BMat2(F)B\simeq\operatorname{Mat}_{2}(F), then the hereditary closure (𝒪)=m(𝒪){\mathcal{H}}({\mathscr{O}})={\mathcal{M}}^{m}({\mathscr{O}}) is decomposable as an 𝒪{\mathscr{O}}-lattice. Thus N≄m(𝒪)N\not\simeq{\mathcal{M}}^{m}({\mathscr{O}}) in this case. It remains to show that i(𝒪){\mathcal{M}}^{i}({\mathscr{O}}) is indecomposable for the remaining ii’s. Suppose otherwise so that i(𝒪)=N1N2{\mathcal{M}}^{i}({\mathscr{O}})=N_{1}\oplus N_{2}, where each NjN_{j} is an 𝒪{\mathscr{O}}-lattice in NjOFFF2N_{j}\otimes_{O_{F}}F\simeq F^{2}. Then

i(𝒪)=Ol(i(𝒪))=Ol(N1N2)=Ol(N1)Ol(N2).{\mathcal{M}}^{i}({\mathscr{O}})=O_{l}({\mathcal{M}}^{i}({\mathscr{O}}))=O_{l}(N_{1}\oplus N_{2})=O_{l}(N_{1})\cap O_{l}(N_{2}).

Since Ol(Ni)O_{l}(N_{i}) is a maximal order in Mat2(F)\operatorname{Mat}_{2}(F) for each ii, this would imply that i(𝒪)){\mathcal{M}}^{i}({\mathscr{O}})) is an Eichler order (i.e. e(i(𝒪)){1,2}e({\mathcal{M}}^{i}({\mathscr{O}}))\in\{1,2\}), contradicting to the fact that e(i(𝒪))=e(𝒪){0,1}e({\mathcal{M}}^{i}({\mathscr{O}}))=e({\mathscr{O}})\in\{0,-1\} for 0im10\leq i\leq m-1. ∎

Applying the Krull-Schmidt-Azumaya Theorem [6, Theorem 6.12], we immediately obtain the following proposition.

Proposition 3.7.

Suppose that 𝒪B{\mathscr{O}}\subset B is a Bass order with e(𝒪){0,1}e({\mathscr{O}})\in\{0,-1\}. Let MM be an 𝒪{\mathscr{O}}-lattice in a finite left BB-module WW.

  1. (1)

    If BB is division, then Mi=0mi(𝒪)tiM\simeq\bigoplus_{i=0}^{m}{\mathcal{M}}^{i}({\mathscr{O}})^{\oplus t_{i}} with (t0,,tm)0m+1(t_{0},\cdots,t_{m})\in\mathbb{Z}_{\geq 0}^{m+1} and i=0mti=dimBW\sum_{i=0}^{m}t_{i}=\dim_{B}W.

  2. (2)

    If BMat2(F)B\simeq\operatorname{Mat}_{2}(F), then WMat2,u(F)W\simeq\operatorname{Mat}_{2,u}(F) for some u0u\geq 0. There are two cases to consider:

    1. (2a)

      if e(𝒪)=0e({\mathscr{O}})=0, then

      (3.10) M[OFϖOF]r[OFOF]si=0m1i(𝒪)tiM\simeq\begin{bmatrix}O_{F}\\ \varpi O_{F}\end{bmatrix}^{\oplus r}\bigoplus\begin{bmatrix}O_{F}\\ O_{F}\end{bmatrix}^{\oplus s}\bigoplus\bigoplus_{i=0}^{m-1}{\mathcal{M}}^{i}({\mathscr{O}})^{\oplus t_{i}}

      with (r,s,t0,,tm1)0m+2(r,s,t_{0},\cdots,t_{m-1})\in\mathbb{Z}_{\geq 0}^{m+2} and r+s+2i=0m1ti=ur+s+2\sum_{i=0}^{m-1}t_{i}=u;

    2. (2b)

      if e(𝒪)=1e({\mathscr{O}})=-1, then

      (3.11) M[OFOF]si=0m1i(𝒪)ti,M\simeq\begin{bmatrix}O_{F}\\ O_{F}\end{bmatrix}^{\oplus s}\bigoplus\bigoplus_{i=0}^{m-1}{\mathcal{M}}^{i}({\mathscr{O}})^{\oplus t_{i}},

      with (s,t0,,tm1)0m+1(s,t_{0},\cdots,t_{m-1})\in\mathbb{Z}_{\geq 0}^{m+1} and s+2i=0m1ti=us+2\sum_{i=0}^{m-1}t_{i}=u.

In all cases, the isomorphism class of MM is uniquely determined by the numerical invariants r,sr,s (if applicable) and the tit_{i}’s.

3.2. Explicit computations

Recall that our goal is to compute the value of o(n)o(n) for n{3,4,5,8,12}n\in\{3,4,5,8,12\} and p|np|n. As explained in Section 2, o(n)o(n) coincides with the number of isomorphism classes of 𝒜n{\mathscr{A}}_{n}-lattices in the left 𝒦n{\mathscr{K}}_{n}-module V=D2V=D^{2} (See (1.4) and (2.1) for the definition of 𝒜n{\mathscr{A}}_{n} and 𝒦n{\mathscr{K}}_{n}). From [23, Theorem 11.1], the nn-th cyclotomic field KnK_{n} has class number 11 for each n{3,4,5,8,12}n\in\{3,4,5,8,12\}.

Since KnK_{n} is totally imaginary and pp does not split completely in KnK_{n}, we have 𝒦n=KnD=Mat2(Kn){\mathscr{K}}_{n}=K_{n}\otimes_{\mathbb{Q}}D=\operatorname{Mat}_{2}(K_{n}). Thus as a left Mat2(Kn)\operatorname{Mat}_{2}(K_{n})-module,

(3.12) V{Mat2(Kn)if n{3,4},Kn2if n{5,8,12}.V\simeq\begin{cases}\operatorname{Mat}_{2}(K_{n})&\text{if }n\in\{3,4\},\\ K_{n}^{2}&\text{if }n\in\{5,8,12\}.\end{cases}

From this, we can easily write down its endomorphism algebra

(3.13) n=End𝒦n(V)={Mat2(Kn)if n{3,4},Knif n{5,8,12}.{\mathscr{E}}_{n}=\operatorname{End}_{{\mathscr{K}}_{n}}(V)=\begin{cases}\operatorname{Mat}_{2}(K_{n})&\text{if }n\in\{3,4\},\\ K_{n}&\text{if }n\in\{5,8,12\}.\end{cases}

In particular, we see that every arithmetic subgroup of n×{\mathscr{E}}_{n}^{\times} is infinite, and hence the abelian surfaces in these isogeny classes have infinite automorphism groups.

Given an 𝒜n{\mathscr{A}}_{n}-lattice ΛV\Lambda\subset V, its endomorphism ring OΛ=End𝒜n(Λ)O_{\Lambda}=\operatorname{End}_{{\mathscr{A}}_{n}}(\Lambda) is an AnA_{n}-order in n{\mathscr{E}}_{n}. Therefore, if n{5,8,12}n\in\{5,8,12\}, then OΛ=AnO_{\Lambda}=A_{n} and h(OΛ)=1h(O_{\Lambda})=1. Now suppose that n{3,4}n\in\{3,4\}. We are going to show in (3.19) that det(OΛ,p×)=An,p×\det(O_{\Lambda,p}^{\times})=A_{n,p}^{\times}. On the other hand, at each prime p\ell\neq p, we have OΛ,Mat2(An,)O_{\Lambda,\ell}\simeq\operatorname{Mat}_{2}(A_{n,\ell}) since OΛ,O_{\Lambda,\ell} is maximal. Thus if we write O^Λ\widehat{O}_{\Lambda} (resp. A^n\widehat{A}_{n}) for the profinite completion of OΛO_{\Lambda} (resp. AnA_{n}), then det(O^Λ×)=A^n×\det(\widehat{O}_{\Lambda}^{\times})=\widehat{A}_{n}^{\times}. The same proof of [22, Corollaire III.5.7(1)] shows that h(OΛ)=h(An)=1h(O_{\Lambda})=h(A_{n})=1. In conclusion, for every pair (n,p)(n,p) with n{3,4,5,8,12}n\in\{3,4,5,8,12\} and p|np|n, we have

(3.14) o(n)=|p(n)|,o(n)=\lvert{\mathscr{L}}_{p}(n)\rvert,

which is consistent with [28, (4.3)].

Lemma 3.8.

Suppose that n{3,4,5,8}n\in\{3,4,5,8\} and p|np|n. Let Up:=Kn,p2U_{p}:=K_{n,p}^{2} be the unique simple 𝒦n,p{\mathscr{K}}_{n,p}-module. Then up to isomorphism, there is a unique 𝒜n,p{\mathscr{A}}_{n,p}-lattice in UpU_{p}.

Proof.

For each pair (n,p)(n,p) under consideration, pp is totally ramified in KnK_{n}. It follows from [15, §2.4] that the Eichler invariants

(3.15) e(𝒜n,p)=e(𝒪p)=1.e({\mathscr{A}}_{n,p})=e({\mathcal{O}}_{p})=-1.

From [2, Proposition 3.1], 𝒜n,p{\mathscr{A}}_{n,p} is a Bass order. Thus the lemma is a direct application of Corollary 3.7. ∎

Lemma 3.9.

Suppose that n{3,4,5,8}n\in\{3,4,5,8\} and p|np|n. Then

(3.16) o(n)={2if n{3,4},1if n{5,8}.o(n)=\begin{cases}2&\text{if }n\in\{3,4\},\\ 1&\text{if }n\in\{5,8\}.\\ \end{cases}
Proof.

First, suppose that n{5,8}n\in\{5,8\}. Then VpKn,p2V_{p}\simeq K_{n,p}^{2} is a simple Mat2(Kn,p)\operatorname{Mat}_{2}(K_{n,p})-module. From Lemma 3.8, |p(n)|=1\lvert{\mathscr{L}}_{p}(n)\rvert=1, and hence o(n)=1o(n)=1 by (3.14).

Next, suppose that n{3,4}n\in\{3,4\}. We have already seen in the proof of Lemma 3.8 that 𝒜n,p{\mathscr{A}}_{n,p} is a Bass order with Eichler invariant 1-1. Let ϖn=1ζn\varpi_{n}=1-\zeta_{n} be the uniformizer of the local field Kn,pK_{n,p}. The reduced discriminant 𝔡(𝒜n,p){\mathfrak{d}}({\mathscr{A}}_{n,p}) is given by

(3.17) 𝔡(𝒜n,p)=𝔡(An,pp𝒪p)=𝔡(𝒪p)An,p=pAn,p=ϖn2An,p.{\mathfrak{d}}({\mathscr{A}}_{n,p})={\mathfrak{d}}(A_{n,p}\otimes_{\mathbb{Z}_{p}}{\mathcal{O}}_{p})={\mathfrak{d}}({\mathcal{O}}_{p})A_{n,p}=pA_{n,p}=\varpi_{n}^{2}A_{n,p}.

Thus the chain of Bass orders in (3.3) reduces to 𝒜n,p(𝒜n,p){\mathscr{A}}_{n,p}\subset{\mathcal{M}}({\mathscr{A}}_{n,p}), where (𝒜n,p)=Mat2(An,p){\mathcal{M}}({\mathscr{A}}_{n,p})=\operatorname{Mat}_{2}(A_{n,p}) under a suitable identification 𝒦n,p=Mat2(Kn,p){\mathscr{K}}_{n,p}=\operatorname{Mat}_{2}(K_{n,p}). In this case, VpV_{p} is a free Mat2(Kn,p)\operatorname{Mat}_{2}(K_{n,p})-module of rank 11. From (3.11), every 𝒜n,p{\mathscr{A}}_{n,p}-lattice Λp\Lambda_{p} in VpV_{p} is isomorphic to either 𝒜n,p{\mathscr{A}}_{n,p} or Mat2(An,p)\operatorname{Mat}_{2}(A_{n,p}). Correspondingly, the endomorphism ring OΛ,pO_{\Lambda,p} is given by

(3.18) OΛ,p=End𝒜n,p(Λp){𝒜n,pif Λp𝒜n,p,Mat2(An,p)if ΛpMat2(An,p).O_{\Lambda,p}=\operatorname{End}_{{\mathscr{A}}_{n,p}}(\Lambda_{p})\simeq\begin{cases}{\mathscr{A}}_{n,p}&\text{if }\Lambda_{p}\simeq{\mathscr{A}}_{n,p},\\ \operatorname{Mat}_{2}(A_{n,p})&\text{if }\Lambda_{p}\simeq\operatorname{Mat}_{2}(A_{n,p}).\\ \end{cases}

In both cases, we have

(3.19) det(OΛ,p×)=An,p×.\det(O_{\Lambda,p}^{\times})=A_{n,p}^{\times}.

Indeed, this is clear if OΛ,pMat2(An,p)O_{\Lambda,p}\simeq\operatorname{Mat}_{2}(A_{n,p}). In the case OΛ,p𝒜n,pO_{\Lambda,p}\simeq{\mathscr{A}}_{n,p}, let LL be the unique unramified quadratic field extension of Kn,pK_{n,p}. From (3.4), OΛ,pO_{\Lambda,p} contains a copy of OLO_{L}, which implies that det(OΛ,p×)NL/Kn,p(OL×)=An,p×\det(O_{\Lambda,p}^{\times})\supseteq\operatorname{N}_{L/K_{n,p}}(O_{L}^{\times})=A_{n,p}^{\times}. On the other hand, det(OΛ,p×)\det(O_{\Lambda,p}^{\times}) is obviously contained in An,p×A_{n,p}^{\times}, so equality (3.19) holds in this case as well. We conclude that o(n)=|p(n)|=2o(n)=\lvert{\mathscr{L}}_{p}(n)\rvert=2 if n{3,4}n\in\{3,4\} and p|np|n. ∎

Lemma 3.10.

o(12)=3o(12)=3 if p{2,3}p\in\{2,3\}.

Proof.

Since 22 and 33 are ramified in KnK_{n}, 22 is inert in (ζ3)\mathbb{Q}(\zeta_{3}) and 33 is inert in (ζ4)\mathbb{Q}(\zeta_{4}), the pp-adic completion Kn,pK_{n,p} is a field extension of degree 44 over p\mathbb{Q}_{p} with residue degree 22, so e(𝒜n,p)=e(An,pp𝒪p)=1e({\mathscr{A}}_{n,p})=e(A_{n,p}\otimes_{\mathbb{Z}_{p}}{\mathcal{O}}_{p})=1 by [15, §2.4]. A similar calculation as (3.17) shows that 𝔡(𝒜n,p)=ϖp2An,p{\mathfrak{d}}({\mathscr{A}}_{n,p})=\varpi_{p}^{2}A_{n,p}, where ϖp\varpi_{p} denotes a uniformizer of Kn,pK_{n,p}. From [2, Proposition 2.1], we may identify 𝒜n,p{\mathscr{A}}_{n,p} with the Eichler order [An,pAn,pϖp2An,pAn,p]\begin{bmatrix}A_{n,p}&A_{n,p}\\ \varpi_{p}^{2}A_{n,p}&A_{n,p}\end{bmatrix}. Since VpKn,p2V_{p}\simeq K_{n,p}^{2} is a simple Mat2(Kn,p)\operatorname{Mat}_{2}(K_{n,p})-module, every 𝒜n,p{\mathscr{A}}_{n,p}-lattice in VpV_{p} is isomorphic to one of the following lattices:

(3.20) [An,pϖp2An,p],[An,pϖpAn,p],[An,pAn,p].\begin{bmatrix}A_{n,p}\\ \varpi_{p}^{2}A_{n,p}\end{bmatrix},\qquad\begin{bmatrix}A_{n,p}\\ \varpi_{p}A_{n,p}\end{bmatrix},\qquad\begin{bmatrix}A_{n,p}\\ A_{n,p}\end{bmatrix}.

Therefore, o(12)=|p(12)|=3o(12)=\lvert{\mathscr{L}}_{p}(12)\rvert=3 by (3.14). ∎

Remark 3.11.

We have proved (cf. (3.14)) that in the isotypic case, every genus in the set (n){\mathscr{L}}(n) of lattice classes has class number one. This holds also in the case pnp\nmid n according to [28, (4.3)].

4. The non-isotypic case

In this section, we compute the values of o(n¯)o({\underline{n}}) for n¯=(n1,n2)˘2{\underline{n}}=(n_{1},n_{2})\in\breve{\mathbb{N}}^{2} and pp satisfying condition (C1) or (C2) (or both). More explicitly, the pairs (n¯,p)({\underline{n}},p) are listed in the following table:

n¯=(n1,n2){\underline{n}}=(n_{1},n_{2}) (1,2)(1,2) (2,3)(2,3) (2,4)(2,4) (2,6)(2,6) (3,4)(3,4) (3,6)(3,6)
pp 22 33 22 33 2,32,3 2,32,3
i(n¯)i({\underline{n}}) 22 11 22 33 11 44

Here we have also included the index i(n¯)=[OKn¯:An¯]i({\underline{n}})=[O_{K_{\underline{n}}}:A_{{\underline{n}}}], where OKn¯=An1×An2O_{K_{\underline{n}}}=A_{n_{1}}\times A_{n_{2}} is the unique maximal order of Kn¯K_{\underline{n}}.

Since Kn¯=Kn1×Kn2K_{{\underline{n}}}=K_{n_{1}}\times K_{n_{2}}, we have 𝒦n¯=𝒦n1×𝒦n2{\mathscr{K}}_{\underline{n}}={\mathscr{K}}_{n_{1}}\times{\mathscr{K}}_{n_{2}}. Consequently, the left 𝒦n¯{\mathscr{K}}_{\underline{n}}-module V=D2V=D^{2} decomposes into a product Vn1×Vn2V_{n_{1}}\times V_{n_{2}}, where each VniV_{n_{i}} is a simple left 𝒦ni{\mathscr{K}}_{n_{i}}-module with dimDVni=1\dim_{D}V_{n_{i}}=1. In turn, n¯=n1×n2{\mathscr{E}}_{\underline{n}}={\mathscr{E}}_{n_{1}}\times{\mathscr{E}}_{n_{2}} with ni:=End𝒦ni(Vni){\mathscr{E}}_{n_{i}}:=\operatorname{End}_{{\mathscr{K}}_{n_{i}}}(V_{n_{i}}). If ni{1,2}n_{i}\in\{1,2\}, then Kni=K_{n_{i}}=\mathbb{Q}, so we have

(4.1) 𝒦ni=D,VniD,niDifni{1,2}.{\mathscr{K}}_{n_{i}}=D,\quad V_{n_{i}}\simeq D,\quad{\mathscr{E}}_{n_{i}}\simeq D\quad\text{if}\quad n_{i}\in\{1,2\}.

If ni{3,4,6}n_{i}\in\{3,4,6\}, then KniK_{n_{i}} is an imaginary quadratic extension of \mathbb{Q}, and pp does not split completely in KniK_{n_{i}}. Thus

(4.2) 𝒦niMat2(Kni),VniKni2,ni=Kniifni{3,4,6}.{\mathscr{K}}_{n_{i}}\simeq\operatorname{Mat}_{2}(K_{n_{i}}),\quad V_{n_{i}}\simeq K_{n_{i}}^{2},\quad{\mathscr{E}}_{n_{i}}=K_{n_{i}}\quad\text{if}\quad n_{i}\in\{3,4,6\}.

To avoid conflict of notations between VniV_{n_{i}} and VV_{\ell}, we will always write the full expression VV\otimes\mathbb{Q}_{\ell} instead of VV_{\ell} for the \ell-adic completion of VV. On the other hand, the subscript n¯{}_{\underline{n}} will never be expanded out explicitly as (n1,n2){}_{(n_{1},n_{2})} nor n1,n2{}_{n_{1},n_{2}}, so there should be no ambiguity about 𝒜ni,:=𝒜ni{\mathscr{A}}_{n_{i},\ell}:={\mathscr{A}}_{n_{i}}\otimes_{\mathbb{Z}}\mathbb{Z}_{\ell}.

If \ell\in{\mathbb{N}} is a prime with i(n¯)\ell\nmid i({\underline{n}}) and p\ell\neq p, then An¯,=OKn¯,A_{{\underline{n}},\ell}=O_{K_{\underline{n}},\ell}, and 𝒪Mat2(){\mathcal{O}}_{\ell}\simeq\operatorname{Mat}_{2}(\mathbb{Z}_{\ell}), which implies that

𝒜n¯,=An¯,𝒪Mat2(OKn¯,).{\mathscr{A}}_{{\underline{n}},\ell}=A_{{\underline{n}},\ell}\otimes_{\mathbb{Z}_{\ell}}{\mathcal{O}}_{\ell}\simeq\operatorname{Mat}_{2}(O_{K_{\underline{n}},\ell}).

Thus 𝒜n¯,{\mathscr{A}}_{{\underline{n}},\ell} is maximal in 𝒦n¯,{\mathscr{K}}_{{\underline{n}},\ell} for such an \ell. From this, we can easily write the set S(n¯,p)S({\underline{n}},p) of primes at which 𝒜n¯{\mathscr{A}}_{\underline{n}} is non-maximal:

(4.3) S(n¯,p)={{2,3}if n¯=(3,6) and p=3;{p}otherwise.S({\underline{n}},p)=\begin{cases}\{2,3\}&\text{if }{\underline{n}}=(3,6)\text{ and }p=3;\\ \{p\}&\text{otherwise}.\end{cases}

Recall that the class number h(𝒪)h({\mathcal{O}}) is given by the following formula [22, Proposition V.3.2]

(4.4) h(𝒪)=p112+13(1(3p))+14(1(4p)),h({\mathcal{O}})=\frac{p-1}{12}+\frac{1}{3}\left(1-\genfrac{(}{)}{}{}{-3}{p}\right)+\frac{1}{4}\left(1-\genfrac{(}{)}{}{}{-4}{p}\right),

where (p)\genfrac{(}{)}{}{}{\cdot}{p} denotes the Legendre symbol. In particular, h(𝒪)=1h({\mathcal{O}})=1 if p=2,3p=2,3. We also note that h(Ani)=1h(A_{n_{i}})=1 for every ni{3,4,6}n_{i}\in\{3,4,6\}. Given dd\in{\mathbb{N}}, we write d\mathbb{Q}_{\ell^{d}} for the unique unramified extension of degree dd over \mathbb{Q}_{\ell}, and d\mathbb{Z}_{\ell^{d}} for its ring of integers.

Lemma 4.1.

o(2,3)=1o(2,3)=1 if p=3p=3, and o(3,4)=2o(3,4)=2 if p{2,3}p\in\{2,3\}.

Proof.

For n¯=(2,3){\underline{n}}=(2,3) or (3,4)(3,4), we have An¯=An1×An2A_{\underline{n}}=A_{n_{1}}\times A_{n_{2}}, and hence 𝒜n¯=𝒜n1×𝒜n2{\mathscr{A}}_{\underline{n}}={\mathscr{A}}_{n_{1}}\times{\mathscr{A}}_{n_{2}}. Any 𝒜n¯{\mathscr{A}}_{{\underline{n}}}-lattice ΛV\Lambda\subset V decomposes into a product Λn1×Λn2\Lambda_{n_{1}}\times\Lambda_{n_{2}}, where each Λni\Lambda_{n_{i}} is an 𝒜ni{\mathscr{A}}_{n_{i}}-lattice in VniV_{n_{i}}. Thus the techniques developed in Section 3 applies here.

First suppose that n¯=(2,3){\underline{n}}=(2,3) and p=3p=3. In this case, 𝒜n1=𝒪=𝒪{\mathscr{A}}_{n_{1}}=\mathbb{Z}\otimes_{\mathbb{Z}}{\mathcal{O}}={\mathcal{O}}, and Vn1DV_{n_{1}}\simeq D by (4.1). Since h(𝒪)=1h({\mathcal{O}})=1, there is a unique isomorphism class of 𝒪{\mathcal{O}}-lattices in Vn1V_{n_{1}}. As S(n¯,p)={3}S({\underline{n}},p)=\{3\} in this case, we consider the 𝒜n2,3{\mathscr{A}}_{n_{2},3}-lattices in Vn23Kn2,32V_{n_{2}}\otimes\mathbb{Q}_{3}\simeq K_{n_{2},3}^{2}. From Lemma 3.8, there is a unique 𝒜n2,3{\mathscr{A}}_{n_{2},3}-lattice up to isomorphism in Vn23V_{n_{2}}\otimes\mathbb{Q}_{3}. This implies that there is only a single genus of 𝒜n2{\mathscr{A}}_{n_{2}}-lattices in Vn2V_{n_{2}}. Note that End𝒜n2(Λn2)=An2[ζ3]\operatorname{End}_{{\mathscr{A}}_{n_{2}}}(\Lambda_{n_{2}})=A_{n_{2}}\simeq\mathbb{Z}[\zeta_{3}], which has class number 11. Thus there is a unique isomorphism class of 𝒜n2{\mathscr{A}}_{n_{2}}-lattices in Vn2V_{n_{2}}. As a result, o(2,3)=11=1o(2,3)=1\cdot 1=1 if p=3p=3.

Next, suppose that n¯=(3,4){\underline{n}}=(3,4) and p=2p=2. Since 22 is ramified in Kn2=(ζ4)K_{n_{2}}=\mathbb{Q}(\zeta_{4}), the same proof as above shows that there is a unique isomorphism class of 𝒜n2{\mathscr{A}}_{n_{2}}-lattices in Vn2V_{n_{2}} in this case. On the other hand, 22 is inert in Kn1=(ζ3)K_{n_{1}}=\mathbb{Q}(\zeta_{3}), so An1,2=4A_{n_{1},2}=\mathbb{Z}_{4}. It follows from [15, Lemma 2.10] that

(4.5) 𝒜n1,2=An1,22𝒪2=42𝒪2[44244].{\mathscr{A}}_{n_{1},2}=A_{n_{1},2}\otimes_{\mathbb{Z}_{2}}{\mathcal{O}}_{2}=\mathbb{Z}_{4}\otimes_{\mathbb{Z}_{2}}{\mathcal{O}}_{2}\simeq\begin{bmatrix}\mathbb{Z}_{4}&\mathbb{Z}_{4}\\ 2\mathbb{Z}_{4}&\mathbb{Z}_{4}\end{bmatrix}.

Hence every 𝒜n1,2{\mathscr{A}}_{n_{1},2}-lattices in Vn12V_{n_{1}}\otimes_{\mathbb{Q}}\mathbb{Q}_{2} is isomorphic to either [424]\left[\begin{smallmatrix}\mathbb{Z}_{4}\\ 2\mathbb{Z}_{4}\end{smallmatrix}\right] or [44]\left[\begin{smallmatrix}\mathbb{Z}_{4}\\ \mathbb{Z}_{4}\end{smallmatrix}\right]. We find that there are two genera of 𝒜n1{\mathscr{A}}_{n_{1}}-lattices in Vn1V_{n_{1}}, each consisting of a unique isomorphism class since h(An1)=h([ζ3])=1h(A_{n_{1}})=h(\mathbb{Z}[\zeta_{3}])=1. Therefore, o(3,4)=(1+1)1=2o(3,4)=(1+1)\cdot 1=2 if p=2p=2.

Lastly, the value of o(3,4)o(3,4) for p=3p=3 can be computed in exactly the same way as above, since 33 is ramified in Kn1K_{n_{1}} and inert in Kn2K_{n_{2}}. ∎

Lemma 4.2.

o(1,2)=3o(1,2)=3 if p=2p=2.

Proof.

Set n¯=(1,2){\underline{n}}=(1,2) throughout this proof. In this case, 𝒜n¯{\mathscr{A}}_{{\underline{n}}} is non-maximal only at p=2p=2, and 𝒪2{\mathcal{O}}_{2} is the unique maximal order in the division quaternion 2\mathbb{Q}_{2}-algebra D2D_{2}. From [28, (5.4)], we have

An¯={(a,b)×ab(mod2)},A_{\underline{n}}=\{(a,b)\in\mathbb{Z}\times\mathbb{Z}\mid a\equiv b\pmod{2}\},

which implies that

(4.6) 𝒜n¯,2=An¯𝒪2={(x,y)𝒪2×𝒪2xy(mod2𝒪2)}.{\mathscr{A}}_{{\underline{n}},2}=A_{\underline{n}}\otimes_{\mathbb{Z}}{\mathcal{O}}_{2}=\{(x,y)\in{\mathcal{O}}_{2}\times{\mathcal{O}}_{2}\mid x\equiv y\pmod{2{\mathcal{O}}_{2}}\}.

In particular, 𝒜n¯,2{\mathscr{A}}_{{\underline{n}},2} is a subdirect sum333Let {Ri1in}\{R_{i}\mid 1\leq i\leq n\} be a finite set of (unital) rings, and R:=i=1nRiR:=\prod_{i=1}^{n}R_{i} be their direct product. A ring TT is called a subdirect sum of the RiR_{i}’s if there exists an embedding ρ:TR\rho:T\to R such that every canonical projection pri:RRi\operatorname{pr}_{i}:R\to R_{i} maps ρ(T)\rho(T) surjectively onto RiR_{i}. of two copies of 𝒪2{\mathcal{O}}_{2}, so it is a Bass order by [9, Proposition 12.3].

Let 𝔓{\mathfrak{P}} be the unique two-sided prime ideal of 𝒪{\mathcal{O}} above p=2p=2. We put

(4.7) :={(x,y)𝒪×𝒪xy(mod𝔓)},{\mathscr{R}}:=\{(x,y)\in{\mathcal{O}}\times{\mathcal{O}}\mid x\equiv y\pmod{{\mathfrak{P}}}\},

which has index 44 in the maximal order 𝕆:=𝒪×𝒪{\mathbb{O}}:={\mathcal{O}}\times{\mathcal{O}} in 𝒦n¯=D×D{\mathscr{K}}_{{\underline{n}}}=D\times D. Indeed, /(𝔓×𝔓)𝒪/𝔓𝔽4{\mathscr{R}}/({\mathfrak{P}}\times{\mathfrak{P}})\simeq{\mathcal{O}}/{\mathfrak{P}}\simeq\mathbb{F}_{4} while 𝕆/(𝔓×𝔓)𝔽4×𝔽4{\mathbb{O}}/({\mathfrak{P}}\times{\mathfrak{P}})\simeq\mathbb{F}_{4}\times\mathbb{F}_{4}. Clearly, both {\mathscr{R}} and 𝕆{\mathbb{O}} are overorders of 𝒜n¯{\mathscr{A}}_{{\underline{n}}}. We claim that there are no other overorders except 𝒜n¯{\mathscr{A}}_{{\underline{n}}} itself. It is enough to prove this locally at p=2p=2. From 𝒜n¯,2/(2𝒪2×2𝒪2)𝒪2/2𝒪2{\mathscr{A}}_{{\underline{n}},2}/(2{\mathcal{O}}_{2}\times 2{\mathcal{O}}_{2})\simeq{\mathcal{O}}_{2}/2{\mathcal{O}}_{2}, we find that

(4.8) 𝒜n¯,2/𝔍(𝒜n¯,2)𝒪2/𝔍(𝒪2)𝔽4,{\mathscr{A}}_{{\underline{n}},2}/{\mathfrak{J}}({\mathscr{A}}_{{\underline{n}},2})\simeq{\mathcal{O}}_{2}/{\mathfrak{J}}({\mathcal{O}}_{2})\simeq\mathbb{F}_{4},

where 𝔍(){\mathfrak{J}}(\cdot) denotes the Jacboson radical. Hence 𝒜n¯,2{\mathscr{A}}_{{\underline{n}},2} is completely primary in the sense of [18, p. 262]. Now according to [18, Lemma 6.6], every non-maximal completely primary Gorenstein order has a unique minimal overorder. As 𝒜n¯,2{\mathscr{A}}_{{\underline{n}},2} has index 44 in 2{\mathscr{R}}_{2}, the left 𝒜n¯,2{\mathscr{A}}_{{\underline{n}},2}-module 2/𝒜n¯,2{\mathscr{R}}_{2}/{\mathscr{A}}_{{\underline{n}},2} is isomorphic to the unique simple left 𝒜n¯,2{\mathscr{A}}_{{\underline{n}},2}-module 𝔽4\mathbb{F}_{4}. Thus 2{\mathscr{R}}_{2} coincides with the unique minimal overorder of 𝒜n¯,2{\mathscr{A}}_{{\underline{n}},2}. Similarly, 2{\mathscr{R}}_{2} is also completely primary, and 𝕆2{\mathbb{O}}_{2} is the unique minimal overorder of 2{\mathscr{R}}_{2} by the same argument. This verifies the claim about the overorders of 𝒜n¯{\mathscr{A}}_{{\underline{n}}}.

In this case, V2V\otimes\mathbb{Q}_{2} is a free left module of rank 11 over 𝒦n¯,2{\mathscr{K}}_{{\underline{n}},2}. For any 𝒜n¯,2{\mathscr{A}}_{{\underline{n}},2}-lattice in V2V\otimes\mathbb{Q}_{2}, its associated left order necessarily coincides with one of the three overorders of 𝒜n¯,2{\mathscr{A}}_{{\underline{n}},2}. Since 𝒜n¯,2{\mathscr{A}}_{{\underline{n}},2} is completely primary, it is indecomposable as a left module over itself. Taking into account that 𝒜n¯,2{\mathscr{A}}_{{\underline{n}},2} is Gorenstein, it follows from [3, Proposition 2.3] that every proper 𝒜n¯,2{\mathscr{A}}_{{\underline{n}},2}-lattice in the V2V\otimes\mathbb{Q}_{2} is principal. This holds for 2{\mathscr{R}}_{2} as well by the same token. On the other hand, every proper 𝕆2{\mathbb{O}}_{2}-lattice in the V2V\otimes\mathbb{Q}_{2} is principal since 𝕆2{\mathbb{O}}_{2} is maximal. Therefore, every 𝒜n¯,2{\mathscr{A}}_{{\underline{n}},2}-lattice in V2V\otimes\mathbb{Q}_{2} is isomorphic to one of the following

(4.9) 𝒜n¯,2,2,𝕆2.{\mathscr{A}}_{{\underline{n}},2},\qquad{\mathscr{R}}_{2},\qquad{\mathbb{O}}_{2}.

We find that there are three genera of 𝒜n¯{\mathscr{A}}_{\underline{n}}-lattices in VV represented by 𝒜n¯,{\mathscr{A}}_{\underline{n}},{\mathscr{R}} and 𝕆{\mathbb{O}} respectively. Clearly,

End𝒜n¯(𝒜n¯)=𝒜n¯opp,End𝒜n¯()=opp,End𝒜n¯(𝕆)=𝕆opp.\operatorname{End}_{{\mathscr{A}}_{\underline{n}}}({\mathscr{A}}_{\underline{n}})={\mathscr{A}}_{\underline{n}}^{\mathrm{opp}},\quad\operatorname{End}_{{\mathscr{A}}_{\underline{n}}}({\mathscr{R}})={\mathscr{R}}^{\mathrm{opp}},\quad\operatorname{End}_{{\mathscr{A}}_{\underline{n}}}({\mathbb{O}})={\mathbb{O}}^{\mathrm{opp}}.

In each case, the opposite ring can be canonically identified with the original ring itself, so we drop the superscript opp henceforth.

It remains to show that h(𝒜n¯)=h()=h(𝕆)=1h({\mathscr{A}}_{\underline{n}})=h({\mathscr{R}})=h({\mathbb{O}})=1. This clearly holds true for the maximal order 𝕆=𝒪×𝒪{\mathbb{O}}={\mathcal{O}}\times{\mathcal{O}} since h(𝕆)=h(𝒪)h(𝒪)=11=1h({\mathbb{O}})=h({\mathcal{O}})\cdot h({\mathcal{O}})=1\cdot 1=1. If we prove h(𝒜n¯)=1h({\mathscr{A}}_{\underline{n}})=1, then h()=1h({\mathscr{R}})=1 since h(𝒜n¯)h()h({\mathscr{A}}_{\underline{n}})\geq h({\mathscr{R}}) by [28, (6.1)]. Let 𝕆^\widehat{\mathbb{O}} (resp. 𝒜^n¯\widehat{\mathscr{A}}_{\underline{n}}) be the profinite completion of 𝕆{\mathbb{O}} (resp. 𝒜n¯{\mathscr{A}}_{\underline{n}}). Since h(𝕆)=1h({\mathbb{O}})=1, it follows from [28, (6.3)] that

(4.10) h(𝒜n¯)=|𝕆×\𝕆^×/(𝒜^n¯)×|.h({\mathscr{A}}_{\underline{n}})=\lvert{\mathbb{O}}^{\times}\backslash\widehat{\mathbb{O}}^{\times}/(\widehat{\mathscr{A}}_{\underline{n}})^{\times}\rvert.

Clearly, 𝒜n¯,2×1+2𝕆2{\mathscr{A}}_{{\underline{n}},2}^{\times}\supseteq 1+2{\mathbb{O}}_{2} and 𝒜n¯,×=𝕆^×{\mathscr{A}}_{{\underline{n}},\ell}^{\times}=\widehat{\mathbb{O}}_{\ell}^{\times} for every prime 2\ell\neq 2, so there is an 𝕆^×\widehat{\mathbb{O}}^{\times}-equivariant projection

(𝕆/2𝕆)×𝕆2×/(1+2𝕆2)𝕆^×/(𝒜^n¯)×.({\mathbb{O}}/2{\mathbb{O}})^{\times}\simeq{\mathbb{O}}_{2}^{\times}/(1+2{\mathbb{O}}_{2})\twoheadrightarrow\widehat{\mathbb{O}}^{\times}/(\widehat{\mathscr{A}}_{\underline{n}})^{\times}.

Hence to prove h(𝒜n¯)=1h({\mathscr{A}}_{\underline{n}})=1, it is enough to show that the canonical projection 𝕆×(𝕆/2𝕆)×{\mathbb{O}}^{\times}\to({\mathbb{O}}/2{\mathbb{O}})^{\times} is surjective. Since 𝕆=𝒪×𝒪{\mathbb{O}}={\mathcal{O}}\times{\mathcal{O}}, this amounts to show that

(4.11) φ:𝒪×(𝒪/2𝒪)×is surjective.\varphi:{\mathcal{O}}^{\times}\to({\mathcal{O}}/2{\mathcal{O}})^{\times}\quad\text{is surjective}.

From [22, Proposition V.3.1], 𝒪×SL2(𝔽3){\mathcal{O}}^{\times}\simeq\operatorname{SL}_{2}(\mathbb{F}_{3}), which has order 2424. On the other hand,

|(𝒪/2𝒪)×|=|(𝒪/𝔓2)×|=|𝔽4×||𝔽4|=34=12.\lvert({\mathcal{O}}/2{\mathcal{O}})^{\times}\rvert=\lvert({\mathcal{O}}/{\mathfrak{P}}^{2})^{\times}\rvert=\lvert\mathbb{F}_{4}^{\times}\rvert\cdot\lvert\mathbb{F}_{4}\rvert=3\cdot 4=12.

Thus to prove the surjectivity of φ\varphi, it suffices to show that ker(φ)={±1}\ker(\varphi)=\{\pm 1\}. Since every α𝒪×\alpha\in{\mathcal{O}}^{\times} has finite group order, this follows from a well-known lemma of Serre [19, Theorem, p. 17–19] (See also [16, Lemma, p. 192], [20] and [13, Lemma 7.2] for some variations and generalizations). Therefore, h(𝒜n¯)=1h({\mathscr{A}}_{\underline{n}})=1 as claimed.

In conclusion, we have

o(1,2)=h(𝒜n¯)+h()+h(𝕆)=1+1+1=3.o(1,2)=h({\mathscr{A}}_{\underline{n}})+h({\mathscr{R}})+h({\mathbb{O}})=1+1+1=3.\qed

For n¯=(3,6){\underline{n}}=(3,6), the values of o(n¯)o({\underline{n}}) for p{2,3}p\in\{2,3\} will be calculated in a few steps.

Lemma 4.3.

o(3,6)=S(n¯,p)|(n¯)|o(3,6)=\prod_{\ell\in S({\underline{n}},p)}\lvert{\mathscr{L}}_{\ell}({\underline{n}})\rvert for both p{2,3}p\in\{2,3\}, where the set S(n¯,p)S({\underline{n}},p) is given in (4.3).

Proof.

We identify both A3A_{3} and A6A_{6} with [ζ3]\mathbb{Z}[\zeta_{3}] via the following maps:

[T]/(T2+T+1)[ζ3],Tζ3,and[T]/(T2T+1)[ζ3],Tζ3.\mathbb{Z}[T]/(T^{2}+T+1)\to\mathbb{Z}[\zeta_{3}],\quad T\mapsto\zeta_{3},\quad\text{and}\quad\mathbb{Z}[T]/(T^{2}-T+1)\to\mathbb{Z}[\zeta_{3}],\quad T\mapsto-\zeta_{3}.

As a result, the maximal order OKn¯O_{K_{\underline{n}}} of Kn¯K_{\underline{n}} is identified with [ζ3]×[ζ3]\mathbb{Z}[\zeta_{3}]\times\mathbb{Z}[\zeta_{3}]. From [28, (5.7)], we have

(4.12) An¯={(a,b)OKn¯ab(mod2[ζ3])}.A_{\underline{n}}=\{(a,b)\in O_{K_{\underline{n}}}\mid a\equiv b\pmod{2\mathbb{Z}[\zeta_{3}]}\}.

In particular, An¯/(2OKn¯)𝔽4A_{\underline{n}}/(2O_{K_{\underline{n}}})\simeq\mathbb{F}_{4}. Hence [OKn¯:An¯]=4[O_{K_{\underline{n}}}:A_{\underline{n}}]=4, and the overorders of An¯A_{\underline{n}} are precisely OKn¯O_{K_{\underline{n}}} and An¯A_{\underline{n}} itself. From (4.2), n¯=Kn¯{\mathscr{E}}_{\underline{n}}=K_{{\underline{n}}}, so the endomorphism ring OΛO_{\Lambda} of any 𝒜n¯{\mathscr{A}}_{\underline{n}}-lattice ΛV\Lambda\subset V is an overorder of An¯A_{\underline{n}}. Thus OΛO_{\Lambda} is equal to either An¯A_{\underline{n}} or OKn¯O_{K_{\underline{n}}}. Since h(An¯)=h(OKn¯)=1h(A_{\underline{n}})=h(O_{K_{{\underline{n}}}})=1 by [28, (5.8)], we conclude that

(4.13) o(3,6)=S(n¯,p)|(n¯)|.o(3,6)=\prod_{\ell\in S({\underline{n}},p)}\lvert{\mathscr{L}}_{\ell}({\underline{n}})\rvert.\qed
Lemma 4.4.

o(3,6)=2o(3,6)=2 if p=3p=3.

Proof.

From (4.3), S(n¯,3)={2,3}S({\underline{n}},3)=\{2,3\}, so we need to classify 𝒜n¯,{\mathscr{A}}_{{\underline{n}},\ell}-lattices in VV\otimes\mathbb{Q}_{\ell} for both =2,3\ell=2,3. For =2\ell=2, we have 𝒪2=Mat2(2){\mathcal{O}}_{2}=\operatorname{Mat}_{2}(\mathbb{Z}_{2}). Hence

(4.14) 𝒜n¯,2=An¯,22𝒪2=Mat2(An¯,2).{\mathscr{A}}_{{\underline{n}},2}=A_{{\underline{n}},2}\otimes_{\mathbb{Z}_{2}}{\mathcal{O}}_{2}=\operatorname{Mat}_{2}(A_{{\underline{n}},2}).

Recall that V=Kn¯2V=K_{\underline{n}}^{2} by (4.2). Applying Morita’s equivalence, we reduce the classification of 𝒜n¯,2{\mathscr{A}}_{{\underline{n}},2}-lattices in V2V\otimes\mathbb{Q}_{2} to that of An¯,2A_{{\underline{n}},2}-lattices in Kn¯,2K_{{\underline{n}},2}. Now An¯,2A_{{\underline{n}},2} is a commutative Bass order over 2\mathbb{Z}_{2}, so it follows from Lemma 3.2 that each An¯,2A_{{\underline{n}},2}-lattice in Kn¯,2K_{{\underline{n}},2} is isomorphic to an overorder of An¯,2A_{{\underline{n}},2} (namely, either An¯,2A_{{\underline{n}},2} or OKn¯,2O_{K_{\underline{n}},2}). Therefore, |2(n¯)|=2\lvert{\mathscr{L}}_{2}({\underline{n}})\rvert=2.

Next, we compute |3(n¯)|\lvert{\mathscr{L}}_{3}({\underline{n}})\rvert. Since 33 is coprime to [OKn¯:An¯]=4[O_{K_{\underline{n}}}:A_{\underline{n}}]=4, we have

(4.15) An¯,3\displaystyle A_{{\underline{n}},3} =OKn¯3=An1,3×An2,3,and hence\displaystyle=O_{K_{\underline{n}}}\otimes_{\mathbb{Z}}\mathbb{Z}_{3}=A_{n_{1},3}\times A_{n_{2},3},\quad\text{and hence}
(4.16) 𝒜n¯,3\displaystyle{\mathscr{A}}_{{\underline{n}},3} =An¯,33𝒪3=𝒜n1,3×𝒜n2,3.\displaystyle=A_{{\underline{n}},3}\otimes_{\mathbb{Z}_{3}}{\mathcal{O}}_{3}={\mathscr{A}}_{n_{1},3}\times{\mathscr{A}}_{n_{2},3}.

Here 𝒜ni,3=3[ζ3]3𝒪3{\mathscr{A}}_{n_{i},3}=\mathbb{Z}_{3}[\zeta_{3}]\otimes_{\mathbb{Z}_{3}}{\mathcal{O}}_{3} for each i=1,2i=1,2. On the other hand, Vni3Kni,32V_{n_{i}}\otimes\mathbb{Q}_{3}\simeq K_{n_{i},3}^{2} by (4.2). It follows from Lemma 3.8 that |3(n¯)|=1\lvert{\mathscr{L}}_{3}({\underline{n}})\rvert=1.

We conclude from Lemma 4.3 that o(3,6)=21=2o(3,6)=2\cdot 1=2 when p=3p=3. ∎

Proposition 4.5.

o(3,6)=8o(3,6)=8 if p=2p=2.

Proof.

In this case, S(n¯,2)={2}S({\underline{n}},2)=\{2\} by (4.3). Let 4\mathbb{Q}_{4} be the unique unramified quadratic extension of 2\mathbb{Q}_{2}, and 4\mathbb{Z}_{4} be its ring of integers. Then Ani,2=2[ζ3]=4A_{n_{i},2}=\mathbb{Z}_{2}[\zeta_{3}]=\mathbb{Z}_{4} for both i=1,2i=1,2. The same calculation as in (4.5) shows that

(4.17) OKn¯,22𝒪2=(4×4)2𝒪2[44244]×[44244].O_{K_{\underline{n}},2}\otimes_{\mathbb{Z}_{2}}{\mathcal{O}}_{2}=(\mathbb{Z}_{4}\times\mathbb{Z}_{4})\otimes_{\mathbb{Z}_{2}}{\mathcal{O}}_{2}\simeq\begin{bmatrix}\mathbb{Z}_{4}&\mathbb{Z}_{4}\\ 2\mathbb{Z}_{4}&\mathbb{Z}_{4}\end{bmatrix}\times\begin{bmatrix}\mathbb{Z}_{4}&\mathbb{Z}_{4}\\ 2\mathbb{Z}_{4}&\mathbb{Z}_{4}\end{bmatrix}.

For simplicity, put 𝔼:=[44244]{\mathbb{E}}:=\left[\begin{smallmatrix}\mathbb{Z}_{4}&\mathbb{Z}_{4}\\ 2\mathbb{Z}_{4}&\mathbb{Z}_{4}\end{smallmatrix}\right] and identify (OKn¯,22𝒪2)(O_{K_{\underline{n}},2}\otimes_{\mathbb{Z}_{2}}{\mathcal{O}}_{2}) with :=𝔼×𝔼{\mathscr{B}}:={\mathbb{E}}\times{\mathbb{E}}. Then it follows from (4.12) that

(4.18) 𝒜n¯,2={(x,y)ab(mod2𝔼)}.{\mathscr{A}}_{{\underline{n}},2}=\{(x,y)\in{\mathscr{B}}\mid a\equiv b\pmod{2{\mathbb{E}}}\}.

In particular, 22{\mathscr{B}} is a two-sided ideal of 𝒜n¯,2{\mathscr{A}}_{{\underline{n}},2} contained in 𝔍(𝒜n¯,2){\mathfrak{J}}({\mathscr{A}}_{{\underline{n}},2}). We put

(4.19) 𝒜¯n¯,2:=𝒜n¯,2/(2)=𝔼/2𝔼,¯:=/2=(𝔼/2𝔼)×(𝔼/2𝔼),\bar{\mathscr{A}}_{{\underline{n}},2}:={\mathscr{A}}_{{\underline{n}},2}/(2{\mathscr{B}})={\mathbb{E}}/2{\mathbb{E}},\qquad\bar{\mathscr{B}}:={\mathscr{B}}/2{\mathscr{B}}=({\mathbb{E}}/2{\mathbb{E}})\times({\mathbb{E}}/2{\mathbb{E}}),

where 𝒜¯n¯,2\bar{\mathscr{A}}_{{\underline{n}},2} embeds into ¯\bar{\mathscr{B}} diagonally.

From (4.2), Vni2=Kni,22=[44]V_{n_{i}}\otimes\mathbb{Q}_{2}=K_{n_{i},2}^{2}=\left[\begin{smallmatrix}\mathbb{Q}_{4}\\ \mathbb{Q}_{4}\end{smallmatrix}\right] for each i=1,2i=1,2. For simplicity, we identify V2V\otimes\mathbb{Q}_{2} with Mat2(4)\operatorname{Mat}_{2}(\mathbb{Q}_{4}) and regard the ii-th column as a 𝒦ni,2{\mathscr{K}}_{n_{i},2}-module for each ii. Any {\mathscr{B}}-lattice in V2V\otimes\mathbb{Q}_{2} is isomorphic to one of the following

(4.20) [442424],[4444],[44244],[44424].\begin{bmatrix}\mathbb{Z}_{4}&\mathbb{Z}_{4}\\ 2\mathbb{Z}_{4}&2\mathbb{Z}_{4}\end{bmatrix},\qquad\begin{bmatrix}\mathbb{Z}_{4}&\mathbb{Z}_{4}\\ \mathbb{Z}_{4}&\mathbb{Z}_{4}\end{bmatrix},\qquad\begin{bmatrix}\mathbb{Z}_{4}&\mathbb{Z}_{4}\\ 2\mathbb{Z}_{4}&\mathbb{Z}_{4}\end{bmatrix},\qquad\begin{bmatrix}\mathbb{Z}_{4}&\mathbb{Z}_{4}\\ \mathbb{Z}_{4}&2\mathbb{Z}_{4}\end{bmatrix}.

From Lemma 4.3, we are only concerned with 𝒜n¯,2{\mathscr{A}}_{{\underline{n}},2}-lattices in V2V\otimes\mathbb{Q}_{2}, so for ease of notation we drop the subscript 2 from Λ2\Lambda_{2} and write Λ\Lambda for an 𝒜n¯,2{\mathscr{A}}_{{\underline{n}},2}-lattice in V2V\otimes\mathbb{Q}_{2}. Replacing Λ\Lambda by gΛg\Lambda for a suitable gn¯,2×g\in{\mathscr{E}}_{{\underline{n}},2}^{\times} if necessary, we assume that Λ{\mathscr{B}}\Lambda is equal to one of the {\mathscr{B}}-lattices Δ\Delta in (4.20). Clearly 2ΔΛΔ2\Delta\subseteq\Lambda\subseteq\Delta since 2𝒜n¯,22{\mathscr{B}}\subseteq{\mathscr{A}}_{{\underline{n}},2}. Thus Λ¯:=Λ/(2Δ)\bar{\Lambda}:=\Lambda/(2\Delta) is an 𝒜¯n¯,2\bar{\mathscr{A}}_{{\underline{n}},2}-submodule of Δ¯:=Δ/2Δ\bar{\Delta}:=\Delta/2\Delta. Moreover, as a ¯\bar{\mathscr{B}}-module, Δ¯\bar{\Delta} is spanned by Λ¯\bar{\Lambda}. Fix one Δ\Delta in (4.20). Let 𝔖(Δ){\mathfrak{S}}(\Delta) be the set of 𝒜n¯,2{\mathscr{A}}_{{\underline{n}},2}-sublattices Λ\Lambda of Δ\Delta satisfying Λ=Δ{\mathscr{B}}\Lambda=\Delta, and 𝔖(Δ¯){\mathfrak{S}}(\bar{\Delta}) be the set of 𝒜¯n¯,2\bar{\mathscr{A}}_{{\underline{n}},2}-submodules M¯\bar{M} of Δ¯\bar{\Delta} satisfying ¯M¯=Δ¯\bar{\mathscr{B}}\bar{M}=\bar{\Delta}. For any M¯𝔖(Δ¯)\bar{M}\in{\mathfrak{S}}(\bar{\Delta}), there is a unique 𝒜n¯,2{\mathscr{A}}_{{\underline{n}},2}-sublattice Λ\Lambda of Δ\Delta satisfying Λ2Δ\Lambda\supseteq 2\Delta and Λ¯=M¯\bar{\Lambda}=\bar{M}. Moreover, Λ=Δ{\mathscr{B}}\Lambda=\Delta by Nakayama’s lemma. Hence the association ΛΛ¯\Lambda\mapsto\bar{\Lambda} induces a bijective map:

(4.21) 𝔖(Δ)𝔖(Δ¯),ΛΛ¯.{\mathfrak{S}}(\Delta)\xrightarrow{\sim}{\mathfrak{S}}(\bar{\Delta}),\qquad\Lambda\mapsto\bar{\Lambda}.

The group End(Δ)×\operatorname{End}_{{\mathscr{B}}}(\Delta)^{\times} acts on both 𝔖(Δ){\mathfrak{S}}(\Delta) and 𝔖(Δ¯){\mathfrak{S}}(\bar{\Delta}), and the map in (4.21) is End(Δ)×\operatorname{End}_{{\mathscr{B}}}(\Delta)^{\times}-equivariant as well. Two members Λ,Λ\Lambda,\Lambda^{\prime} of 𝔖(Δ){\mathfrak{S}}(\Delta) are 𝒜n¯,2{\mathscr{A}}_{{\underline{n}},2}-isomorphic if and only if there exists an αEnd(Δ)×\alpha\in\operatorname{End}_{{\mathscr{B}}}(\Delta)^{\times} such that αΛ=Λ\alpha\Lambda=\Lambda^{\prime}. Therefore, we have established the following bijection

(4.22) {Isomorphism classes of 𝒜n¯,2-lattices Λ in V2 with ΛΔ}{End(Δ)×-equivalent classes of 𝒜¯n¯,2-submodules M¯ in Δ¯ such that ¯M¯=Δ¯}.\left\{\parbox{115.63243pt}{Isomorphism classes of ${\mathscr{A}}_{{\underline{n}},2}$-lattices $\Lambda$ in $V\otimes\mathbb{Q}_{2}$ with ${\mathscr{B}}\Lambda\simeq\Delta$}\right\}\longleftrightarrow\left\{\parbox{130.08621pt}{$\operatorname{End}_{{\mathscr{B}}}(\Delta)^{\times}$-equivalent classes of $\bar{\mathscr{A}}_{{\underline{n}},2}$-submodules $\bar{M}$ in $\bar{\Delta}$ such that $\bar{\mathscr{B}}\bar{M}=\bar{\Delta}$}\right\}.

Note that End(Δ)=4×4\operatorname{End}_{{\mathscr{B}}}(\Delta)=\mathbb{Z}_{4}\times\mathbb{Z}_{4} for every Δ\Delta in (4.20), so the action of End(Δ)×\operatorname{End}_{{\mathscr{B}}}(\Delta)^{\times} on 𝔖(Δ¯){\mathfrak{S}}(\bar{\Delta}) factors through 𝔽4××𝔽4×\mathbb{F}_{4}^{\times}\times\mathbb{F}_{4}^{\times}.

Recall from (4.19) that 𝒜¯n¯,2=𝔼/2𝔼\bar{\mathscr{A}}_{{\underline{n}},2}={\mathbb{E}}/2{\mathbb{E}}, and ¯=(𝔼/2𝔼)2\bar{\mathscr{B}}=({\mathbb{E}}/2{\mathbb{E}})^{2}. We describe the 44-dimensional 𝔽4\mathbb{F}_{4}-algebra 𝔼/2𝔼{\mathbb{E}}/2{\mathbb{E}} more concretely. Let RR be the commutative 𝔽4\mathbb{F}_{4}-algebra 𝔽4×𝔽4\mathbb{F}_{4}\times\mathbb{F}_{4}. Regard Q:=RQ:=R as an (R,R)(R,R)-bimodule such that for each (a,d)R(a,d)\in R and (b,c)Q(b,c)\in Q, the multiplications are given by the following rules:

(4.23) (a,d)(b,c)=(ab,dc),(b,c)(a,d)=(bd,ca).(a,d)\cdot(b,c)=(ab,dc),\qquad(b,c)\cdot(a,d)=(bd,ca).

We form the trivial extension [14, Example 1.14] of RR by QQ and denote it as 𝔼¯:=𝔽4𝔽4𝔽4𝔽4\bar{{\mathbb{E}}}:=\left<\begin{smallmatrix}\mathbb{F}_{4}&\mathbb{F}_{4}\\ \mathbb{F}_{4}&\mathbb{F}_{4}\end{smallmatrix}\right>, where RR is identified with the diagonal of 𝔽4𝔽4𝔽4𝔽4\left<\begin{smallmatrix}\mathbb{F}_{4}&\mathbb{F}_{4}\\ \mathbb{F}_{4}&\mathbb{F}_{4}\end{smallmatrix}\right> and QQ is identified with the anti-diagonal. More explicitly, the product in 𝔼¯\bar{\mathbb{E}} is defined by the following rule:

(4.24) abcdabcd:=aaab+bdca+dcdd.\left<\begin{matrix}a&b\\ c&d\end{matrix}\right>\cdot\left<\begin{matrix}a^{\prime}&b^{\prime}\\ c^{\prime}&d^{\prime}\end{matrix}\right>:=\left<\begin{matrix}aa^{\prime}&ab^{\prime}+bd^{\prime}\\ ca^{\prime}+dc^{\prime}&dd^{\prime}\end{matrix}\right>.

For each x4x\in\mathbb{Z}_{4}, let x¯\bar{x} be its canonical image in 𝔽4=4/24\mathbb{F}_{4}=\mathbb{Z}_{4}/2\mathbb{Z}_{4}. One easily checks that the following map induces an isomorphism between 𝔼/2𝔼{\mathbb{E}}/2{\mathbb{E}} and 𝔼¯\bar{\mathbb{E}}:

[44244]𝔽4𝔽4𝔽4𝔽4,[xy2zw]x¯y¯z¯w¯.\begin{bmatrix}\mathbb{Z}_{4}&\mathbb{Z}_{4}\\ 2\mathbb{Z}_{4}&\mathbb{Z}_{4}\end{bmatrix}\to\left<\begin{matrix}\mathbb{F}_{4}&\mathbb{F}_{4}\\ \mathbb{F}_{4}&\mathbb{F}_{4}\end{matrix}\right>,\qquad\begin{bmatrix}x&y\\ 2z&w\end{bmatrix}\mapsto\left<\begin{matrix}\bar{x}&\bar{y}\\ \bar{z}&\bar{w}\end{matrix}\right>.

Henceforth 𝔼/2𝔼{\mathbb{E}}/2{\mathbb{E}} will be identified with 𝔼¯\bar{\mathbb{E}} via this induced isomorphism. Each column of 𝔼¯\bar{\mathbb{E}} admits a canonical 𝔼¯\bar{\mathbb{E}}-module structure. These two 𝔼¯\bar{\mathbb{E}}-modules will be denoted by [𝔽4𝔽4]\left[\begin{smallmatrix}\mathbb{F}_{4}\\ \mathbb{F}_{4}\end{smallmatrix}\right]^{\dagger} and [𝔽4𝔽4]\left[\begin{smallmatrix}\mathbb{F}_{4}\\ \mathbb{F}_{4}\end{smallmatrix}\right]^{\ddagger} respectively. It is clear from (4.24) that [0𝔽4]\left[\begin{smallmatrix}0\\ \mathbb{F}_{4}\end{smallmatrix}\right]^{\dagger} (resp. [𝔽40]\left[\begin{smallmatrix}\mathbb{F}_{4}\\ 0\end{smallmatrix}\right]^{\ddagger}) is the unique 11-dimensional (as an 𝔽4\mathbb{F}_{4}-vector space) 𝔼¯\bar{\mathbb{E}}-submodule of [𝔽4𝔽4]\left[\begin{smallmatrix}\mathbb{F}_{4}\\ \mathbb{F}_{4}\end{smallmatrix}\right]^{\dagger} (resp. [𝔽4𝔽4]\left[\begin{smallmatrix}\mathbb{F}_{4}\\ \mathbb{F}_{4}\end{smallmatrix}\right]^{\ddagger}). We leave it as a simple exercise to check that [𝔽4𝔽4]\left[\begin{smallmatrix}\mathbb{F}_{4}\\ \mathbb{F}_{4}\end{smallmatrix}\right]^{\dagger} and [𝔽4𝔽4]\left[\begin{smallmatrix}\mathbb{F}_{4}\\ \mathbb{F}_{4}\end{smallmatrix}\right]^{\ddagger} are non-isomorphic 𝔼¯\bar{\mathbb{E}}-modules.

Now for each Δ\Delta in (4.20), we classify the (𝔽4××𝔽4×)(\mathbb{F}_{4}^{\times}\times\mathbb{F}_{4}^{\times})-orbits of 𝔖(Δ¯){\mathfrak{S}}(\bar{\Delta}). For i=1,2i=1,2, let Δi\Delta_{i} be the ii-th column of Δ\Delta so that Δ¯=Δ¯1×Δ¯2\bar{\Delta}=\bar{\Delta}_{1}\times\bar{\Delta}_{2}. Let M¯Δ¯\bar{M}\subseteq\bar{\Delta} be an 𝔼¯\bar{\mathbb{E}}-submodule, and pri:M¯Δ¯i\operatorname{pr}_{i}:\bar{M}\to\bar{\Delta}_{i} be the projection map to the factor Δ¯i\bar{\Delta}_{i}. Then (𝔼¯×𝔼¯)M¯=Δ¯(\bar{\mathbb{E}}\times\bar{\mathbb{E}})\bar{M}=\bar{\Delta} if and only if both pri\operatorname{pr}_{i} are surjective. Thus

(4.25) dim𝔽4M¯2for everyM¯𝔖(Δ¯).\dim_{\mathbb{F}_{4}}\bar{M}\geq 2\quad\text{for every}\quad\bar{M}\in{\mathfrak{S}}(\bar{\Delta}).

Suppose that M¯𝔖(Δ¯)\bar{M}\in{\mathfrak{S}}(\bar{\Delta}) from now on.

If dim𝔽4M¯=2\dim_{\mathbb{F}_{4}}\bar{M}=2, then both pri\operatorname{pr}_{i} are 𝔼¯\bar{\mathbb{E}}-isomorphisms, and M¯\bar{M} is the graph of the isomorphism pr2pr11:Δ¯1Δ¯2\operatorname{pr}_{2}\circ\operatorname{pr}_{1}^{-1}:\bar{\Delta}_{1}\to\bar{\Delta}_{2}. Necessarily, Δ\Delta is equal to either [442424]\left[\begin{smallmatrix}\mathbb{Z}_{4}&\mathbb{Z}_{4}\\ 2\mathbb{Z}_{4}&2\mathbb{Z}_{4}\end{smallmatrix}\right] or [4444]\left[\begin{smallmatrix}\mathbb{Z}_{4}&\mathbb{Z}_{4}\\ \mathbb{Z}_{4}&\mathbb{Z}_{4}\end{smallmatrix}\right]. In these two cases, any isomorphism Δ¯1Δ¯2\bar{\Delta}_{1}\to\bar{\Delta}_{2} is a scalar multiplication by 𝔽4×\mathbb{F}_{4}^{\times}. After a suitable multiplication by an element of 𝔽4××𝔽4×\mathbb{F}_{4}^{\times}\times\mathbb{F}_{4}^{\times}, we may identify M¯\bar{M} with the diagonal of Δ¯1×Δ¯2\bar{\Delta}_{1}\times\bar{\Delta}_{2}.

Next, suppose that dim𝔽4M¯=3\dim_{\mathbb{F}_{4}}\bar{M}=3. Then ker(pr1)\ker(\operatorname{pr}_{1}) is a 11-dimensional submodule of M¯Δ¯2\bar{M}\cap\bar{\Delta}_{2}, so it must coincide with the unique 11-dimensional submodule of Δ¯2\bar{\Delta}_{2}. A similar result holds for ker(pr2)\ker(\operatorname{pr}_{2}). We claim that Δ¯1Δ¯2\bar{\Delta}_{1}\simeq\bar{\Delta}_{2} in this case as well. If not, without lose of generality, we may assume that Δ¯1=[𝔽4𝔽4]\bar{\Delta}_{1}=\left[\begin{smallmatrix}\mathbb{F}_{4}\\ \mathbb{F}_{4}\end{smallmatrix}\right]^{\dagger} and Δ¯2=[𝔽4𝔽4]\bar{\Delta}_{2}=\left[\begin{smallmatrix}\mathbb{F}_{4}\\ \mathbb{F}_{4}\end{smallmatrix}\right]^{\ddagger} so that Δ¯=𝔼¯\bar{\Delta}=\bar{\mathbb{E}}. By the above discussion, M¯0𝔽4𝔽40\bar{M}\supseteq\left<\begin{smallmatrix}0&\mathbb{F}_{4}\\ \mathbb{F}_{4}&0\end{smallmatrix}\right>. Since dim𝔽4M¯=3\dim_{\mathbb{F}_{4}}\bar{M}=3, there exists u,v𝔽4×u,v\in\mathbb{F}_{4}^{\times} such that

M¯=𝔽4u00v+𝔽40010+𝔽40100.\bar{M}=\mathbb{F}_{4}\left<\begin{matrix}u&0\\ 0&v\end{matrix}\right>+\mathbb{F}_{4}\left<\begin{matrix}0&0\\ 1&0\end{matrix}\right>+\mathbb{F}_{4}\left<\begin{matrix}0&1\\ 0&0\end{matrix}\right>.

Indeed, both uu and vv have to be nonzero since pri\operatorname{pr}_{i} is surjective for each i=1,2i=1,2. However, u00v𝔼¯×\left<\begin{smallmatrix}u&0\\ 0&v\end{smallmatrix}\right>\in\bar{\mathbb{E}}^{\times}, which implies that M¯=𝔼¯=Δ¯\bar{M}=\bar{\mathbb{E}}=\bar{\Delta}. This contradicts dim𝔽4M¯=3\dim_{\mathbb{F}_{4}}\bar{M}=3 and verifies our claim. It remains to consider the cases Δ¯=[𝔽4𝔽4]×[𝔽4𝔽4]\bar{\Delta}=\left[\begin{smallmatrix}\mathbb{F}_{4}\\ \mathbb{F}_{4}\end{smallmatrix}\right]^{\dagger}\times\left[\begin{smallmatrix}\mathbb{F}_{4}\\ \mathbb{F}_{4}\end{smallmatrix}\right]^{\dagger} or Δ¯=[𝔽4𝔽4]×[𝔽4𝔽4]\bar{\Delta}=\left[\begin{smallmatrix}\mathbb{F}_{4}\\ \mathbb{F}_{4}\end{smallmatrix}\right]^{\ddagger}\times\left[\begin{smallmatrix}\mathbb{F}_{4}\\ \mathbb{F}_{4}\end{smallmatrix}\right]^{\ddagger}. For simplicity, we write them as Mat2(𝔽4)\operatorname{Mat}_{2}(\mathbb{F}_{4})^{\dagger} and Mat2(𝔽4)\operatorname{Mat}_{2}(\mathbb{F}_{4})^{\ddagger} respectively. Suppose that Δ¯=Mat2(𝔽4)\bar{\Delta}=\operatorname{Mat}_{2}(\mathbb{F}_{4})^{\dagger}. Then there exists u,v𝔽4×u,v\in\mathbb{F}_{4}^{\times} such that

M¯=𝔽4[uv00]+𝔽4[0010]+𝔽4[0001].\bar{M}=\mathbb{F}_{4}\begin{bmatrix}u&v\\ 0&0\end{bmatrix}+\mathbb{F}_{4}\begin{bmatrix}0&0\\ 1&0\end{bmatrix}+\mathbb{F}_{4}\begin{bmatrix}0&0\\ 0&1\end{bmatrix}.

Multiplication by (u1,v1)𝔽4××𝔽4×(u^{-1},v^{-1})\in\mathbb{F}_{4}^{\times}\times\mathbb{F}_{4}^{\times} sends M¯\bar{M} to

M¯0:={[abcd]Mat2(𝔽4)|a=b}.\bar{M}_{0}:=\left\{\begin{bmatrix}a&b\\ c&d\end{bmatrix}\in\operatorname{Mat}_{2}(\mathbb{F}_{4})^{\dagger}\,\middle|\,a=b\right\}.

Thus all 3-dimensional members of 𝔖(Mat2(𝔽4)){\mathfrak{S}}(\operatorname{Mat}_{2}(\mathbb{F}_{4})^{\dagger}) are in the same (𝔽4××𝔽4×)(\mathbb{F}_{4}^{\times}\times\mathbb{F}_{4}^{\times})-orbit. A similar result holds for Δ¯=Mat2(𝔽4)\bar{\Delta}=\operatorname{Mat}_{2}(\mathbb{F}_{4})^{\ddagger}.

Lastly, if dim𝔽4M¯=4\dim_{\mathbb{F}_{4}}\bar{M}=4, then M¯=Δ¯\bar{M}=\bar{\Delta}.

Combining (4.20), (4.22) and the above classifications, we find that

|2(n¯)|=Δ|(𝔽4××𝔽4×)\𝔖(Δ¯)|=3+3+1+1=8.\lvert{\mathscr{L}}_{2}({\underline{n}})\rvert=\sum_{\Delta}\lvert(\mathbb{F}_{4}^{\times}\times\mathbb{F}_{4}^{\times})\backslash{\mathfrak{S}}(\bar{\Delta})\rvert=3+3+1+1=8.

Thus o(3,6)=8o(3,6)=8 if p=2p=2 according to Lemma 4.3. ∎

Proposition 4.6.

Suppose that p{2,3}p\in\{2,3\}. Then

(4.26) o(2,2p)={2if p=2,3if p=3.o(2,2p)=\begin{cases}2&\text{if }p=2,\\ 3&\text{if }p=3.\end{cases}
Proof.

Let n¯=(n1,n2)=(2,2p){\underline{n}}=(n_{1},n_{2})=(2,2p) for p{2,3}p\in\{2,3\}. Then Kn¯=×Kn2K_{\underline{n}}=\mathbb{Q}\times K_{n_{2}}, and 𝒦n¯D×Mat2(Kn2){\mathscr{K}}_{\underline{n}}\simeq D\times\operatorname{Mat}_{2}(K_{n_{2}}) by (4.1) and (4.2). Here Kn2K_{n_{2}} is equal to (1)\mathbb{Q}(\,\sqrt[]{-1}\,) or (3)\mathbb{Q}(\,\sqrt[]{-3}\,) according to whether p=2p=2 or p=3p=3. Let 𝔭{\mathfrak{p}} be the unique ramified prime ideal of An2A_{n_{2}} above pp. From [28, (5.5) and (5.6)], we have

(4.27) An¯={(a,b)×An2ab(mod𝔭)},A_{\underline{n}}=\{(a,b)\in\mathbb{Z}\times A_{n_{2}}\mid a\equiv b\pmod{{\mathfrak{p}}}\},

which is a suborder of index pp in OKn¯=×An2O_{K_{\underline{n}}}=\mathbb{Z}\times A_{n_{2}}.

From (4.3), S(n¯,p)={p}S({\underline{n}},p)=\{p\}. Clearly, 𝒪p{\mathcal{O}}_{p} is canonically a subring of 𝒜n2,p=An2𝒪p{\mathscr{A}}_{n_{2},p}=A_{n_{2}}\otimes_{\mathbb{Z}}{\mathcal{O}}_{p}, and (𝔭𝒜n2,p)𝒪p=p𝒪p({\mathfrak{p}}{\mathscr{A}}_{n_{2},p})\cap{\mathcal{O}}_{p}=p{\mathcal{O}}_{p}. It follows from (4.27) that

(4.28) 𝒜n¯,p={(x,y)𝒪p×𝒜n2,pxy(mod𝔭𝒜n2,p)}.{\mathscr{A}}_{{\underline{n}},p}=\{(x,y)\in{\mathcal{O}}_{p}\times{\mathscr{A}}_{n_{2},p}\mid x\equiv y\pmod{{\mathfrak{p}}{\mathscr{A}}_{n_{2},p}}\}.

For simplicity, let us put 𝒞:=𝒪p×𝒜n2,p{\mathscr{C}}:={\mathcal{O}}_{p}\times{\mathscr{A}}_{n_{2},p}, and J:=𝔍(OKn¯,p)=pp×𝔭J:={\mathfrak{J}}(O_{K_{\underline{n}},p})=p\mathbb{Z}_{p}\times{\mathfrak{p}}, where 𝔭{\mathfrak{p}} denotes the maximal ideal of An2,pA_{n_{2},p} by an abuse of notation. Then J𝒞𝒜n¯,pJ{\mathscr{C}}\subseteq{\mathscr{A}}_{{\underline{n}},p}, so we further define three quotient rings

(4.29) 𝒪¯p\displaystyle\bar{\mathcal{O}}_{p} :=𝒪p/p𝒪p,\displaystyle:={\mathcal{O}}_{p}/p{\mathcal{O}}_{p},
(4.30) 𝒞¯\displaystyle\bar{\mathscr{C}} :=𝒞/(J𝒞)=(𝒪p/p𝒪p)×(𝒜n2,p/𝔭𝒜n2,p)=𝒪¯p×𝒪¯p,\displaystyle:={\mathscr{C}}/(J{\mathscr{C}})=({\mathcal{O}}_{p}/p{\mathcal{O}}_{p})\times({\mathscr{A}}_{n_{2},p}/{\mathfrak{p}}{\mathscr{A}}_{n_{2},p})=\bar{\mathcal{O}}_{p}\times\bar{\mathcal{O}}_{p},
(4.31) 𝒜¯n¯,p\displaystyle\bar{\mathscr{A}}_{{\underline{n}},p} :=𝒜n¯,p/(J𝒞)=𝒪¯p,\displaystyle:={\mathscr{A}}_{{\underline{n}},p}/(J{\mathscr{C}})=\bar{\mathcal{O}}_{p},

where 𝒜¯n¯,p\bar{\mathscr{A}}_{{\underline{n}},p} embeds into 𝒞¯\bar{\mathscr{C}} diagonally by (4.28). From [22, Corollary II.1.7], 𝒪p{\mathcal{O}}_{p} contains a copy of p2\mathbb{Z}_{p^{2}}, and there exists η𝒪p\eta\in{\mathcal{O}}_{p} such that

(4.32) 𝒪p=p2+p2η,η2=p,andxη=ηx~,xp2.{\mathcal{O}}_{p}=\mathbb{Z}_{p^{2}}+\mathbb{Z}_{p^{2}}\eta,\qquad\eta^{2}=p,\quad\text{and}\quad x\eta=\eta\tilde{x},\quad\forall x\in\mathbb{Z}_{p^{2}}.

Here xx~x\mapsto\tilde{x} denotes the unique nontrivial p\mathbb{Q}_{p}-automorphism of p2\mathbb{Q}_{p^{2}}. If we write η¯\bar{\eta} for the canonical image of η\eta in 𝒪¯p\bar{\mathcal{O}}_{p}, then

(4.33) 𝒪¯p=𝔽p2+𝔽p2η¯.\bar{\mathcal{O}}_{p}=\mathbb{F}_{p^{2}}+\mathbb{F}_{p^{2}}\bar{\eta}.

In particular, dim𝔽p2𝒪¯p=2\dim_{\mathbb{F}_{p^{2}}}\bar{\mathcal{O}}_{p}=2, and the Jacobson radical 𝔍(𝒪¯p)=𝔽p2η¯{\mathfrak{J}}(\bar{\mathcal{O}}_{p})=\mathbb{F}_{p^{2}}\bar{\eta} is the unique 11-dimensional submodule of 𝒪¯p\bar{\mathcal{O}}_{p}.

From (4.1) and (4.2), Vn1V_{n_{1}} is a free module of rank 11 over 𝒦n1D{\mathscr{K}}_{n_{1}}\simeq D, and Vn2=Kn22V_{n_{2}}=K_{n_{2}}^{2} is a simple module over 𝒦n2Mat2(Kn2){\mathscr{K}}_{n_{2}}\simeq\operatorname{Mat}_{2}(K_{n_{2}}). Fix a suitable identification of 𝒦n2,p=Mat2(Kn2,p){\mathscr{K}}_{n_{2},p}=\operatorname{Mat}_{2}(K_{n_{2},p}) so that the hereditary closure (𝒜n2,p){\mathcal{H}}({\mathscr{A}}_{n_{2},p}) is equal to Mat2(An2,p)\operatorname{Mat}_{2}(A_{n_{2},p}). From Lemma 3.8, every 𝒞{\mathscr{C}}-lattice in VpV\otimes\mathbb{Q}_{p} is isomorphic to

(4.34) Δ:=Δ1×Δ2=𝒪p×[An2,pAn2,p].\Delta:=\Delta_{1}\times\Delta_{2}={\mathcal{O}}_{p}\times\begin{bmatrix}A_{n_{2},p}\\ A_{n_{2},p}\end{bmatrix}.

Put Δ¯:=Δ/JΔ=Δ¯1×Δ¯2\bar{\Delta}:=\Delta/J\Delta=\bar{\Delta}_{1}\times\bar{\Delta}_{2}, where

(4.35) Δ¯1=𝒪p/p𝒪p=𝒪¯p,andΔ¯2=Δ2/𝔭Δ2.\bar{\Delta}_{1}={\mathcal{O}}_{p}/p{\mathcal{O}}_{p}=\bar{\mathcal{O}}_{p},\quad\text{and}\quad\bar{\Delta}_{2}=\Delta_{2}/{\mathfrak{p}}\Delta_{2}.

The same proof as (4.22) shows that there is a bijection

(4.36) {Isomorphism classes of 𝒜n¯,p-lattices in Vp}{End𝒞(Δ)×-equivalent classes of 𝒜¯n¯,p-submodules M¯ in Δ¯ such that 𝒞¯M¯=Δ¯}.\left\{\parbox{86.72377pt}{Isomorphism classes of ${\mathscr{A}}_{{\underline{n}},p}$-lattices in $V\otimes\mathbb{Q}_{p}$}\right\}\longleftrightarrow\left\{\parbox{130.08621pt}{$\operatorname{End}_{{\mathscr{C}}}(\Delta)^{\times}$-equivalent classes of $\bar{\mathscr{A}}_{{\underline{n}},p}$-submodules $\bar{M}$ in $\bar{\Delta}$ such that $\bar{\mathscr{C}}\bar{M}=\bar{\Delta}$}\right\}.

Clearly, dim𝔽pΔ¯2=2\dim_{\mathbb{F}_{p}}\bar{\Delta}_{2}=2. When regarded as an 𝒜n2,p{\mathscr{A}}_{n_{2},p}-module, Δ¯2\bar{\Delta}_{2} is isomorphic to the unique simple 𝒜n2,p{\mathscr{A}}_{n_{2},p}-module 𝔽p2\mathbb{F}_{p^{2}} by (4.30) and (4.33). We fix such an isomorphism and write Δ¯2=𝔽p2\bar{\Delta}_{2}=\mathbb{F}_{p^{2}}. Note that End𝒞(Δ)=𝒪p×An2,p\operatorname{End}_{{\mathscr{C}}}(\Delta)={\mathcal{O}}_{p}\times A_{n_{2},p}, so the action of End𝒞(Δ)×\operatorname{End}_{{\mathscr{C}}}(\Delta)^{\times} on Δ¯\bar{\Delta} factors through 𝒪¯p××𝔽p×\bar{\mathcal{O}}_{p}^{\times}\times\mathbb{F}_{p}^{\times}. Recall that 𝒞¯=𝒪¯p×𝒪¯p\bar{\mathscr{C}}=\bar{\mathcal{O}}_{p}\times\bar{\mathcal{O}}_{p} and 𝒜¯n¯,p=𝒪¯p\bar{\mathscr{A}}_{{\underline{n}},p}=\bar{\mathcal{O}}_{p} by (4.30) and (4.31).

Let M¯Δ¯\bar{M}\subseteq\bar{\Delta} be an 𝒪¯p\bar{\mathcal{O}}_{p}-submodule, and pri:M¯Δi\operatorname{pr}_{i}:\bar{M}\to\Delta_{i} be the canonical projections for i=1,2i=1,2. Then (𝒪¯p×𝒪¯p)M¯=Δ¯(\bar{\mathcal{O}}_{p}\times\bar{\mathcal{O}}_{p})\bar{M}=\bar{\Delta} if and only if both pri\operatorname{pr}_{i} are surjective. Suppose that this is the case. Then necessarily

dim𝔽p2M¯dim𝔽p2Δ¯1=2.\dim_{\mathbb{F}_{p^{2}}}\bar{M}\geq\dim_{\mathbb{F}_{p^{2}}}\bar{\Delta}_{1}=2.

If dim𝔽p2M¯=3\dim_{\mathbb{F}_{p^{2}}}\bar{M}=3, then M¯=Δ¯\bar{M}=\bar{\Delta}.

Now suppose that dim𝔽p2M¯=2\dim_{\mathbb{F}_{p^{2}}}\bar{M}=2. Then ker(pr2)\ker(\operatorname{pr}_{2}) is a 11-dimensional submodule of Δ¯1\bar{\Delta}_{1}, so ker(pr2)=𝔽p2η¯\ker(\operatorname{pr}_{2})=\mathbb{F}_{p^{2}}\bar{\eta}. Therefore, there exists a,c𝔽p2×a,c\in\mathbb{F}_{p^{2}}^{\times} such that

(4.37) M¯=𝔽p2(η¯,0)+𝔽p2(a,c)Δ¯=(𝔽p2+𝔽p2η¯)×𝔽p2.\bar{M}=\mathbb{F}_{p^{2}}(\bar{\eta},0)+\mathbb{F}_{p^{2}}(a,c)\subseteq\bar{\Delta}=(\mathbb{F}_{p^{2}}+\mathbb{F}_{p^{2}}\bar{\eta})\times\mathbb{F}_{p^{2}}.

Indeed, both a,ca,c have to be nonzero since pri\operatorname{pr}_{i} is surjective for each i=1,2i=1,2. Multiplication by (ca1,1)𝒪¯p××𝔽p×(ca^{-1},1)\in\bar{\mathcal{O}}_{p}^{\times}\times\mathbb{F}_{p}^{\times} sends M¯\bar{M} to the following submodule of Δ¯\bar{\Delta}:

(4.38) Γ¯:={(a+bη¯,c)Δ¯a=c}.\bar{\Gamma}:=\{(a+b\bar{\eta},c)\in\bar{\Delta}\mid a=c\}.

Let Γ\Gamma be the unique 𝒜n¯,p{\mathscr{A}}_{{\underline{n}},p}-sublattice of Δ\Delta such that ΓJΔ\Gamma\supseteq J\Delta and Γ/JΔ=Γ¯\Gamma/J\Delta=\bar{\Gamma}. Then every 𝒜n¯,p{\mathscr{A}}_{{\underline{n}},p}-lattice in VpV\otimes\mathbb{Q}_{p} is isomorphic to either Δ\Delta or Γ\Gamma. Let 𝔓{\mathfrak{P}} be the unique two-sided prime ideal of 𝒪p{\mathcal{O}}_{p}. We compute

(4.39) End𝒜n¯,p(Γ)={(x,y)𝒪p×An2,p(x,y)Γ¯Γ¯}={(x,y)𝒪p×An2,pxy(mod𝔓)}.\begin{split}\operatorname{End}_{{\mathscr{A}}_{{\underline{n}},p}}(\Gamma)&=\{(x,y)\in{\mathcal{O}}_{p}\times A_{n_{2},p}\mid(x,y)\bar{\Gamma}\subseteq\bar{\Gamma}\}\\ &=\{(x,y)\in{\mathcal{O}}_{p}\times A_{n_{2},p}\mid x\equiv y\pmod{{\mathfrak{P}}}\}.\end{split}

Here 𝔓An2,p=𝔭{\mathfrak{P}}\cap A_{n_{2},p}={\mathfrak{p}} for any embedding of An2,pA_{n_{2},p} into 𝒪p{\mathcal{O}}_{p}, so the congruence relation does not depend on the choice of such an embedding.

We have seen from above that there are two genera of 𝒜n¯{\mathscr{A}}_{\underline{n}}-lattices in VV. According to [22, Proposition V.3.1], h(𝒪)=1h({\mathcal{O}})=1 for p{2,3}p\in\{2,3\}, so 𝒪{\mathcal{O}} is the unique maximal order in DD up to D×D^{\times}-conjugation. If Δ~\widetilde{\Delta} is an 𝒜n¯{\mathscr{A}}_{\underline{n}}-lattices in VV with Δ~p=Δ\widetilde{\Delta}\otimes\mathbb{Z}_{p}=\Delta, then

(4.40) End𝒜n¯(Δ~)𝒪×An2.\operatorname{End}_{{\mathscr{A}}_{\underline{n}}}(\widetilde{\Delta})\simeq{\mathcal{O}}\times A_{n_{2}}.

Similarly, if Γ~\widetilde{\Gamma} is the unique 𝒜n¯{\mathscr{A}}_{\underline{n}}-sublattice of Δ~\widetilde{\Delta} such that Γ~p=Γ\widetilde{\Gamma}\otimes\mathbb{Z}_{p}=\Gamma, then End𝒜n¯(Γ~)\operatorname{End}_{{\mathscr{A}}_{\underline{n}}}(\widetilde{\Gamma}) is the unique suborder of End𝒜n¯(Δ~)\operatorname{End}_{{\mathscr{A}}_{\underline{n}}}(\widetilde{\Delta}) satisfying

(4.41) End𝒜n¯(Γ~)={End𝒜n¯,p(Γ)if =p,End𝒜n¯(Δ~)otherwise.\operatorname{End}_{{\mathscr{A}}_{\underline{n}}}(\widetilde{\Gamma})\otimes\mathbb{Z}_{\ell}=\begin{cases}\operatorname{End}_{{\mathscr{A}}_{\underline{n}},p}(\Gamma)&\text{if }\ell=p,\\ \operatorname{End}_{{\mathscr{A}}_{\underline{n}}}(\widetilde{\Delta})\otimes\mathbb{Z}_{\ell}&\text{otherwise}.\end{cases}

For simplicity, let us put 𝔒:=End𝒜n¯(Δ~){\mathfrak{O}}:=\operatorname{End}_{{\mathscr{A}}_{\underline{n}}}(\widetilde{\Delta}) and =End𝒜n¯,p(Γ){\mathfrak{R}}=\operatorname{End}_{{\mathscr{A}}_{\underline{n}},p}(\Gamma). From the general strategy explained in Section 2, we have

(4.42) o(2,2p)=h(𝔒)+h()for p=2,3.o(2,2p)=h({\mathfrak{O}})+h({\mathfrak{R}})\qquad\text{for }p=2,3.

Here h(𝔒)=h(𝒪)h(An2)=11=1h({\mathfrak{O}})=h({\mathcal{O}})h(A_{n_{2}})=1\cdot 1=1. It remains to compute h()h({\mathfrak{R}}).

Replacing Δ~\widetilde{\Delta} by gΔ~g\widetilde{\Delta} for a suitable gn¯×g\in{\mathscr{E}}_{\underline{n}}^{\times} if necessary, we assume that 𝔒=𝒪×An2{\mathfrak{O}}={\mathcal{O}}\times A_{n_{2}}. Let 𝔒^\widehat{\mathfrak{O}} (resp. ^\widehat{\mathfrak{R}}) be the profinite completion of 𝔒{\mathfrak{O}} (resp. {\mathfrak{R}}). We apply [28, (6.3)] to obtain

(4.43) h()=|𝔒×\𝔒^×/^×|=|𝔒×\𝔒p×/p×|.h({\mathfrak{R}})=\lvert{\mathfrak{O}}^{\times}\backslash\widehat{\mathfrak{O}}^{\times}/\widehat{\mathfrak{R}}^{\times}\rvert=\lvert{\mathfrak{O}}^{\times}\backslash{\mathfrak{O}}_{p}^{\times}/{\mathfrak{R}}_{p}^{\times}\rvert.

From (4.39), 𝔓×𝔭p{\mathfrak{P}}\times{\mathfrak{p}}\subseteq{\mathfrak{R}}_{p}, and

𝔒p/(𝔓×𝔭)=𝔽p2×𝔽p,p/(𝔓×𝔭)=𝔽p.{\mathfrak{O}}_{p}/({\mathfrak{P}}\times{\mathfrak{p}})=\mathbb{F}_{p^{2}}\times\mathbb{F}_{p},\qquad{\mathfrak{R}}_{p}/({\mathfrak{P}}\times{\mathfrak{p}})=\mathbb{F}_{p}.

where 𝔽p\mathbb{F}_{p} embeds into 𝔽p2×𝔽p\mathbb{F}_{p^{2}}\times\mathbb{F}_{p} diagonally. It follows that 𝔒p×/p×{\mathfrak{O}}_{p}^{\times}/{\mathfrak{R}}_{p}^{\times} can be further simplified into (𝔽p2××𝔽p×)/diag(𝔽p×)(\mathbb{F}_{p^{2}}^{\times}\times\mathbb{F}_{p}^{\times})/\operatorname{diag}(\mathbb{F}_{p}^{\times}). On the other hand, 𝔒×=𝒪××An2×{\mathfrak{O}}^{\times}={\mathcal{O}}^{\times}\times A_{n_{2}}^{\times}. Since p{2,3}p\in\{2,3\}, the natural map 𝔒×𝔽p2××𝔽p×{\mathfrak{O}}^{\times}\to\mathbb{F}_{p^{2}}^{\times}\times\mathbb{F}_{p}^{\times} sends {±1}×An2×\{\pm 1\}\times A_{n_{2}}^{\times} onto 𝔽p××𝔽p×\mathbb{F}_{p}^{\times}\times\mathbb{F}_{p}^{\times}, which contains diag(𝔽p×)\operatorname{diag}(\mathbb{F}_{p}^{\times}). Therefore,

(4.44) h()=|𝔒×\𝔒p×/p×|=|(𝒪××An2×)\(𝔽p2××𝔽p×)/diag(𝔽p×)|=|(𝒪××{1})\(𝔽p2××𝔽p×)/(𝔽p××𝔽p×)|=|𝒪×\𝔽p2×/𝔽p×|=|𝒪×\𝔽p2×|.\begin{split}h({\mathfrak{R}})&=\lvert{\mathfrak{O}}^{\times}\backslash{\mathfrak{O}}_{p}^{\times}/{\mathfrak{R}}_{p}^{\times}\rvert=\lvert({\mathcal{O}}^{\times}\times A_{n_{2}}^{\times})\backslash(\mathbb{F}_{p^{2}}^{\times}\times\mathbb{F}_{p}^{\times})/\operatorname{diag}(\mathbb{F}_{p}^{\times})\rvert\\ &=\lvert({\mathcal{O}}^{\times}\times\{1\})\backslash(\mathbb{F}_{p^{2}}^{\times}\times\mathbb{F}_{p}^{\times})/(\mathbb{F}_{p}^{\times}\times\mathbb{F}_{p}^{\times})\rvert=\lvert{\mathcal{O}}^{\times}\backslash\mathbb{F}_{p^{2}}^{\times}/\mathbb{F}_{p}^{\times}\rvert=\lvert{\mathcal{O}}^{\times}\backslash\mathbb{F}_{p^{2}}^{\times}\rvert.\end{split}

Here we have used freely the commutativity of 𝔽p2××𝔽p×\mathbb{F}_{p^{2}}^{\times}\times\mathbb{F}_{p}^{\times}.

If p=2p=2, we have already shown in the proof of Lemma 4.2 that the canonical map 𝒪×(𝒪/2𝒪)×{\mathcal{O}}^{\times}\to({\mathcal{O}}/2{\mathcal{O}})^{\times} is surjective (see (4.11)). Consequently, 𝒪×(𝒪p/𝔓)×=𝔽4×{\mathcal{O}}^{\times}\to({\mathcal{O}}_{p}/{\mathfrak{P}})^{\times}=\mathbb{F}_{4}^{\times} is surjective as well. Thus h()=1h({\mathfrak{R}})=1 if p=2p=2.

Next, suppose that p=3p=3. According to [22, Exercise III.5.2],

D=(1,3),and𝒪=+i+1+j2+i(1+j)2.D=\left(\frac{-1,-3}{\mathbb{Q}}\right),\quad\text{and}\quad{\mathcal{O}}=\mathbb{Z}+\mathbb{Z}i+\mathbb{Z}\frac{1+j}{2}+\mathbb{Z}\frac{i(1+j)}{2}.

Here {1,i,j,ij}\{1,i,j,ij\} is the standard \mathbb{Q}-basis of DD. We have

𝒪×={±1,±i,±1±j2,±i(1±j)2}.{\mathcal{O}}^{\times}=\left\{\pm 1,\quad\pm i,\quad\pm\frac{1\pm j}{2},\quad\pm\frac{i(1\pm j)}{2}\right\}.

Note that j𝔓j\in{\mathfrak{P}}, so the image of 𝒪×{\mathcal{O}}^{\times} in (𝒪p/𝔓)×=𝔽9×({\mathcal{O}}_{p}/{\mathfrak{P}})^{\times}=\mathbb{F}_{9}^{\times} is equal to {±1¯,±i¯}\{\pm\bar{1},\pm\bar{i}\}, which has index 22 in 𝔽9×\mathbb{F}_{9}^{\times}. Therefore, h()=2h({\mathfrak{R}})=2 if p=3p=3.

In conclusion, we find that o(2,4)=1+1=2o(2,4)=1+1=2 if p=2p=2, and o(2,6)=1+2=3o(2,6)=1+2=3 if p=3p=3. ∎

Remark 4.7.

Assume that p{2,3}p\in\{2,3\}. We have shown that except for (n¯,p)=((2,6),3)({\underline{n}},p)=((2,6),3) (and ((1,3),3)((1,3),3) by [28, Remark 3.2(i)]), every genus in the set (n¯){\mathscr{L}}({\underline{n}}) of lattice classes has class number one. For the exceptional cases, (n¯){\mathscr{L}}({\underline{n}}) consists of two genera; one genus has class number one while the other one has class number two.

Concluding the Proof of Theorem 1.1.

For p=2,3,5p=2,3,5, the value of o(n¯)o({\underline{n}}) is listed in the following table (see [28] for the values not covered in the present paper):

n¯{\underline{n}} 33 44 55 88 1212 (1,2)(1,2) (2,3)(2,3) (2,4)(2,4) (2,6)(2,6) (3,4)(3,4) (3,6)(3,6)
p=2p=2 33 22 22 11 33 33 22 22 44 22 88
p=3p=3 22 33 22 44 33 33 11 44 33 22 22
p=5p=5 33 11 11 44 44 44 22 0 66 0 88

Formula (1.6) is obtained by plugging in the above values in (1.3). ∎

Acknowledgments

J. Xue is partially supported by Natural Science Foundation of China grant #11601395. C.-F. Yu is partially supported by the MoST grant 109-2115-M-001-002-MY3.

References

  • [1] Hyman Bass. On the ubiquity of Gorenstein rings. Math. Z., 82:8–28, 1963.
  • [2] J. Brzeziński. On orders in quaternion algebras. Comm. Algebra, 11(5):501–522, 1983.
  • [3] J. Brzezinski. Riemann-Roch theorem for locally principal orders. Math. Ann., 276(4):529–536, 1987.
  • [4] Juliusz Brzezinski. On automorphisms of quaternion orders. J. Reine Angew. Math., 403:166–186, 1990.
  • [5] Sara Chari, Daniel Smertnig, and John Voight. On basic and Bass quaternion orders. arXiv e-prints, page arXiv:1903.00560, March 2019, arXiv:1903.00560.
  • [6] Charles W. Curtis and Irving Reiner. Methods of representation theory. Vol. I. Wiley Classics Library. John Wiley & Sons, Inc., New York, 1990. With applications to finite groups and orders, Reprint of the 1981 original, A Wiley-Interscience Publication.
  • [7] Max Deuring. Die Typen der Multiplikatorenringe elliptischer Funktionenkörper. Abh. Math. Sem. Hansischen Univ., 14:197–272, 1941.
  • [8] Max Deuring. Die Anzahl der Typen von Maximalordnungen einer definiten Quaternionenalgebra mit primer Grundzahl. Jber. Deutsch. Math. Verein., 54:24–41, 1950.
  • [9] Ju. A. Drozd, V. V. Kiričenko, and A. V. Roĭter. Hereditary and Bass orders. Izv. Akad. Nauk SSSR Ser. Mat., 31:1415–1436, 1967.
  • [10] Martin Eichler. Über die Idealklassenzahl total definiter Quaternionenalgebren. Math. Z., 43(1):102–109, 1938.
  • [11] Jun-ichi Igusa. Class number of a definite quaternion with prime discriminant. Proc. Nat. Acad. Sci. U.S.A., 44:312–314, 1958.
  • [12] Irving Kaplansky. Submodules of quaternion algebras. Proc. London Math. Soc. (3), 19:219–232, 1969.
  • [13] Valentijn Karemaker, Fuetaro Yobuko, and Chia-Fu Yu. Mass formula and oort’s conjecture for supersingular abelian threefolds, 2020, arXiv:2002.10960.
  • [14] T. Y. Lam. A first course in noncommutative rings, volume 131 of Graduate Texts in Mathematics. Springer-Verlag, New York, 1991.
  • [15] Qun Li, Jiangwei Xue, and Chia-Fu Yu. Unit groups of maximal orders in totally definite quaternion algebras over real quadratic fields. ArXiv e-prints, July 2018, arXiv:1807.04736. To appear in Trans. Amer. Math. Soc.
  • [16] David Mumford. Abelian varieties, volume 5 of Tata Institute of Fundamental Research Studies in Mathematics. Published for the Tata Institute of Fundamental Research, Bombay, 2008. With appendices by C. P. Ramanujam and Yuri Manin, Corrected reprint of the second (1974) edition.
  • [17] Deke Peng and Jiangwei Xue. Optimal spinor selectivity for quaternion Bass orders. arXiv e-prints, December 2020, arXiv:2012.01117.
  • [18] Klaus W. Roggenkamp. Lattices over orders. II. Lecture Notes in Mathematics, Vol. 142. Springer-Verlag, Berlin-New York, 1970.
  • [19] Jean-Pierre Serre. Rigidité du foncteur de jacobi d’échelon n3n\geq 3, Appendix to A. Grothendieck, Techniques de construction en géométrie analytique. X. Construction de l’espace de Teichmüller. Séminaire Henri Cartan, 13(2), 1960-1961. talk:17.
  • [20] A. Silverberg and Yu. G. Zarhin. Variations on a theme of Minkowski and Serre. J. Pure Appl. Algebra, 111(1-3):285–302, 1996.
  • [21] Richard G. Swan. Torsion free cancellation over orders. Illinois J. Math., 32(3):329–360, 1988.
  • [22] Marie-France Vignéras. Arithmétique des algèbres de quaternions, volume 800 of Lecture Notes in Mathematics. Springer, Berlin, 1980.
  • [23] Lawrence C. Washington. Introduction to cyclotomic fields, volume 83 of Graduate Texts in Mathematics. Springer-Verlag, New York, second edition, 1997.
  • [24] William C. Waterhouse. Abelian varieties over finite fields. Ann. Sci. École Norm. Sup. (4), 2:521–560, 1969.
  • [25] Jiangwei Xue, Tse-Chung Yang, and Chia-Fu Yu. Numerical invariants of totally imaginary quadratic [p]\mathbb{Z}[\,\sqrt[]{p}\,]-orders. Taiwanese J. Math., 20(4):723–741, 2016.
  • [26] Jiangwei Xue, Tse-Chung Yang, and Chia-Fu Yu. On superspecial abelian surfaces over finite fields. Doc. Math., 21:1607–1643, 2016.
  • [27] Jiangwei Xue, Tse-Chung Yang, and Chia-Fu Yu. Supersingular abelian surfaces and Eichler’s class number formula. Asian J. Math., 23(4):651–680, 2019.
  • [28] Jiangwei Xue, Tse-Chung Yang, and Chia-Fu Yu. On superspecial abelian surfaces over finite fields II. J. Math. Soc. Japan, 72(1):303–331, 2020.
  • [29] Chia-Fu Yu. Endomorphism algebras of QM abelian surfaces. J. Pure Appl. Algebra, 217(5):907–914, 2013.