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

Frobenius distributions of low dimensional
abelian varieties over finite fields

Santiago Arango-Piñeros Department of Mathematics, Emory University, Atlanta, GA 30322, USA [email protected] https://sarangop1728.github.io/ Deewang Bhamidipati Mathematics Department, University of California, Santa Cruz, CA 95064, USA [email protected] https://bdeewang.com/  and  Soumya Sankar Mathematical Institute, Utrecht University, Hans Freudenthal building, Budapest 6, 3584 CD Utrecht, The Netherlands [email protected] https://sites.google.com/site/soumya3sankar/
Abstract.

Given a gg-dimensional abelian variety AA over a finite field 𝐅q\mathbf{F}_{q}, the Weil conjectures imply that the normalized Frobenius eigenvalues generate a multiplicative group of rank at most gg. The Pontryagin dual of this group is a compact abelian Lie group that controls the distribution of high powers of the Frobenius endomorphism. This group, which we call the Serre–Frobenius group, encodes the possible multiplicative relations between the Frobenius eigenvalues. In this article, we classify all possible Serre–Frobenius groups that occur for g3g\leq 3. We also give a partial classification for simple ordinary abelian varieties of prime dimension g3g\geq 3.

Key words and phrases:
Abelian varieties over finite fields, Frobenius traces, Equidistribution
2020 Mathematics Subject Classification:
11G10, 11G25, 11M38, 14K02, 14K15

1. Introduction

Let EE be an elliptic curve over a finite field 𝐅q\mathbf{F}_{q} of characteristic p>0p>0. The zeros α1,α¯1\alpha_{1},\overline{\alpha}_{1} of the characteristic polynomial of Frobenius acting on the Tate module of EE are complex numbers of absolute value q\sqrt{q}. Consider u1\colonequalsα1/qu_{1}\colonequals\alpha_{1}/\sqrt{q} and u¯1\overline{u}_{1} the normalized zeros in the unit circle 𝖴(1)\mathsf{U}(1). The curve EE is ordinary if and only if u1u_{1} is not a root of unity, and in this case, the sequence (u1r)r=1(u_{1}^{r})_{r=1}^{\infty} is equidistributed in 𝖴(1)\mathsf{U}(1). Further, the normalized Frobenius traces xr\colonequalsu1r+u¯1rx_{r}\colonequals u_{1}^{r}+\overline{u}_{1}^{r} are equidistributed on the interval [2,2][-2,2] with respect to the pushforward of the probability Haar measure on 𝖴(1)\mathsf{U}(1) via uu+u¯u\mapsto u+\overline{u}, namely

(1) λ1(x)\colonequalsdxπ4x2,\lambda_{1}(x)\colonequals\frac{\mathrm{d}x}{\pi\sqrt{4-x^{2}}},

where dx\,\mathrm{d}x is the restriction of the Lebesgue measure to [2,2][-2,2] (see [9, Proposition 2.2]).

In contrast, if EE is supersingular, the sequence (u1r)r=1(u_{1}^{r})_{r=1}^{\infty} generates a finite cyclic subgroup of order mm, Cm𝖴(1)C_{m}\subset\mathsf{U}(1). In this case, the normalized Frobenius traces are equidistributed with respect to the pushforward of the uniform measure on CmC_{m}.

This dichotomy branches out in an interesting way for abelian varieties of higher dimension g>1g>1: potential non-trivial multiplicative relations between the Frobenius eigenvalues α1,α¯1,,αg,α¯g\alpha_{1},\overline{\alpha}_{1},\dots,\alpha_{g},\overline{\alpha}_{g} increase the complexity of the problem of classifying the distribution of normalized traces of high powers of Frobenius,

(2) xr\colonequals(α1r+α¯1r++αgr+α¯gr)/qr/2[2g,2g], for r1.x_{r}\colonequals(\alpha_{1}^{r}+\overline{\alpha}_{1}^{r}+\cdots+\alpha_{g}^{r}+\overline{\alpha}_{g}^{r})/q^{r/2}\in[-2g,2g],\text{ for }r\geq 1.

In analogy with the case of elliptic curves, we identify a compact abelian subgroup of 𝖴𝖲𝗉2g(𝐂)\mathsf{USp}_{2g}(\mathbf{C}) controlling the distribution of Sequence (2) via pushforward of the Haar measure. In this article, we provide a complete classification of the conjugacy class of this subgroup, which we call the Serre–Frobenius group, for abelian varieties of dimension up to 3. We do this by classifying the possible multiplicative relations between the Frobenius eigenvalues. This classification provides a description of all the possible distributions of Frobenius traces in these cases (see 1.1.1). We also provide a partial classification for simple ordinary abelian varieties of odd prime dimension.

Definition 1.0.1 (Serre–Frobenius group).

Let AA be an abelian variety of dimension gg over 𝐅q\mathbf{F}_{q}. Let α1,α2,αg,α¯1,α¯2α¯g\alpha_{1},\alpha_{2}\ldots,\alpha_{g},\overline{\alpha}_{1},\overline{\alpha}_{2}\ldots\overline{\alpha}_{g} denote the eigenvalues of Frobenius. Let ui=αi/qu_{i}=\alpha_{i}/\sqrt{q} denote the normalized Frobenius eigenvalues. The Serre–Frobenius group of AA, denoted by 𝖲𝖥(A)\mathsf{SF}(A), is the closure of the subgroup of 𝖴𝖲𝗉2g(𝐂)\mathsf{USp}_{2g}(\mathbf{C}) generated by the diagonal matrix diag(u1,,ug,u¯1,,u¯g)\mathrm{diag}(u_{1},\dots,u_{g},\overline{u}_{1},\dots,\overline{u}_{g}). This group is well defined up to relabelling of the eigenvalues of Frobenius (see Remark 2.3.1).

The classification of Serre–Frobenius groups relies crucially on the relation between the Serre–Frobenius group and the multiplicative subgroup of UA𝐂×U_{A}\subset\mathbf{C}^{\times} generated by the normalized eigenvalues u1,,ugu_{1},\dots,u_{g}. Indeed, the isomorphism class of the former is determined by the Pontryagin dual of the latter (see Lemma 2.3.2). The rank of the group UAU_{A} is called the angle rank of the abelian variety and the order of the torsion subgroup is called the angle torsion order. The relation between 𝖲𝖥(A)\mathsf{SF}(A) and the group generated by the normalized eigenvalues gives us the following structure theorem.

Theorem 1.0.2.

Let AA be an abelian variety defined over 𝐅q\mathbf{F}_{q}. Then

𝖲𝖥(A)𝖴(1)δ×Cm,\mathsf{SF}(A)\cong\mathsf{U}(1)^{\delta}\times C_{m},

where δ=δA\delta=\delta_{A} is the angle rank and m=mAm=m_{A} is the angle torsion order. Furthermore, the connected component of the identity is 𝖲𝖥(A)=𝖲𝖥(A×𝐅q𝐅qm)\mathsf{SF}(A)^{\circ}=\mathsf{SF}(A\times_{\mathbf{F}_{q}}\mathbf{F}_{q^{m}}).

By definition, the Serre–Frobenius group carries the data of the embedding into the 𝖴𝖲𝗉2g(𝐂)\mathsf{USp}_{2g}(\mathbf{C}), which in turn is captured by the relations among the Frobenius eigenvalues. While in general these relations can be hard to pin down (see for instance, [8, Theorem 3.25]), in our cases, we are able to write them down explicitly and use them to deduce the angle torsion order. In particular, we classify the Serre–Frobenius groups of abelian varieties of dimension g3g\leq 3.

Theorem 1.0.3 (Elliptic curves).

Let EE be an elliptic curve defined over 𝐅q\mathbf{F}_{q}. Then

  1. (1)

    EE is ordinary if and only if 𝖲𝖥(E)=𝖴(1)\mathsf{SF}(E)=\mathsf{U}(1).

  2. (2)

    EE is supersingular if and only if 𝖲𝖥(E){C1,C2,C3,C4,C6,C8,C12}\mathsf{SF}(E)\in\mathopen{}\mathclose{{}\left\{C_{1},C_{2},C_{3},C_{4},C_{6},C_{8},C_{12}}\right\}.

Here, CmC_{m} is the subgroup of 𝖴𝖲𝗉2(𝐂)=𝖲𝖫2(𝐂)\mathsf{USp}_{2}(\mathbf{C})=\mathsf{SL}_{2}(\mathbf{C}) generated by (ζm00ζm1)\bigl{(}\begin{smallmatrix}\zeta_{m}&0\\ 0&{\zeta}^{-1}_{m}\end{smallmatrix}\bigr{)} for ζm\zeta_{m} a primitive mm-th root, and 𝖴(1)={(u00u¯):u𝐂×,|u|=1}\mathsf{U}(1)=\mathopen{}\mathclose{{}\left\{\bigl{(}\begin{smallmatrix}u&0\\ 0&\overline{u}\end{smallmatrix}\bigr{)}:u\in\mathbf{C}^{\times},|u|=1}\right\}. Moreover, each one of these groups is realized for some prime power qq.

We note that the classification of supersingular Serre–Frobenius groups of elliptic curves follows from Deuring [5] and Waterhouse’s [33] classification of Frobenius traces (see also [22, Section 14.6] and [28, Theorem 2.6.1]).

Theorem 1.0.4 (Abelian surfaces).

Let SS be an abelian surface over 𝐅q\mathbf{F}_{q}. Then SS has Serre–Frobenius group according to Figure 3. In particular, the possible options for the connected component of the identity 𝖲𝖥(S)\mathsf{SF}(S)^{\circ}, and the size of the cyclic component group 𝖲𝖥(S)/𝖲𝖥(S)Cm\mathsf{SF}(S)/\mathsf{SF}(S)^{\circ}\cong C_{m} are given below. Moreover, each one of these groups is realized for some prime power qq.

𝖲𝖥(S)\mathsf{SF}(S)^{\circ} mm
11 1,2,3,4,5,6,8,10,12,24
𝖴(1)\mathsf{U}(1) 1,2,3,4,6,8,12
𝖴(1)2\mathsf{U}(1)^{2} 1
Theorem 1.0.5 (Abelian threefolds).

Let XX be an abelian threefold over 𝐅q\mathbf{F}_{q}. Then, XX has Serre–Frobenius group according to Figure 7. In particular, the possible options for the connected component of the identity, 𝖲𝖥(X)\mathsf{SF}(X)^{\circ}, and the size of the cyclic component group 𝖲𝖥(X)/𝖲𝖥(X)Cm\mathsf{SF}(X)/\mathsf{SF}(X)^{\circ}\cong C_{m} are given below. Moreover, each one of these groups is realized for some prime power qq.

𝖲𝖥(X)\mathsf{SF}(X)^{\circ} mm
11 1,2,3,4,5,6,7,8,9,10,12,14,15,18,20,24,28,30,36
𝖴(1)\mathsf{U}(1) 1,2,3,4,5,6,7,8,10,12,24
𝖴(1)2\mathsf{U}(1)^{2} 1,2,3,4,6,8,12,24
𝖴(1)3\mathsf{U}(1)^{3} 1

If gg is an odd prime, we have the following classification for simple ordinary abelian varieties; in the following theorem, we say that an abelian variety AA splits over a field extension 𝐅qm\mathbf{F}_{q^{m}} if AA is isogenous over 𝐅qm\mathbf{F}_{q^{m}} to a product of proper abelian subvarieties.

Theorem 1.0.6 (Prime dimension).

Let AA be a simple ordinary abelian variety defined over 𝐅q\mathbf{F}_{q} of prime dimension g>2g>2. Then, exactly one of the following conditions holds.

  1. (1)

    AA is absolutely simple.

  2. (2)

    AA splits over a degree gg extension of 𝐅q\mathbf{F}_{q} as a power of an elliptic curve, and 𝖲𝖥(A)𝖴(1)×Cg\mathsf{SF}(A)\cong\mathsf{U}(1)\times C_{g}.

  3. (3)

    AA splits over a degree 2g+12g+1 extension of 𝐅q\mathbf{F}_{q} as a power of an elliptic curve, and 𝖲𝖥(A)𝖴(1)×C2g+1\mathsf{SF}(A)\cong\mathsf{U}(1)\times C_{2g+1}. This case only occurs if 2g+12g+1 is also a prime, i.e., if gg is a Sophie Germain prime.

1.1. Application to distributions of Frobenius traces

Our results can be applied to understanding the distribution of Frobenius traces of an abelian variety over 𝐅q\mathbf{F}_{q} as we range over finite extensions of the base field. Indeed, for each integer r1r\geq 1, we may rewrite Equation (2) as

xr=u1r+u¯1r++ugr+u¯gr[2g,2g]x_{r}=u_{1}^{r}+\overline{u}_{1}^{r}+\cdots+u_{g}^{r}+\overline{u}_{g}^{r}\in[-2g,2g]

denote the normalized Frobenius trace of the base change of an abelian variety AA to 𝐅qr\mathbf{F}_{q^{r}}.

In [1], the authors study Jacobians of smooth projective genus gg curves with maximal angle rank111In their notation, this is the condition that the Frobenius angles are linearly independent modulo 1. and show that the sequence (xr/2g)r=1(x_{r}/2g)_{r=1}^{\infty} is equidistributed on [1,1][-1,1] with respect to an explicit measure. The Serre–Frobenius group enables us to remove the assumption of maximal angle rank.

Corollary 1.1.1.

Let AA be a gg-dimensional abelian variety defined over 𝐅q\mathbf{F}_{q}. Then, the sequence (xr)r=1(x_{r})_{r=1}^{\infty} of normalized traces of Frobenius is equidistributed in [2g,2g][-2g,2g] with respect to the pushforward of the Haar measure on 𝖲𝖥(A)𝖴𝖲𝗉2g(𝐂)\mathsf{SF}(A)\subseteq\mathsf{USp}_{2g}(\mathbf{C}) via the trace

(3) 𝖲𝖥(A)𝖴𝖲𝗉2g(𝐂)[2g,2g],MTr(M).\mathsf{SF}(A)\subseteq\mathsf{USp}_{2g}(\mathbf{C})\to[-2g,2g],\quad M\mapsto\operatorname{Tr}(M).

The classification of the Serre–Frobenius groups in our theorems can be used to distinguish between the different Frobenius trace distributions occurring in each dimension.

Example 1.1.2.

Let SS be a simple abelian surface over 𝐅q\mathbf{F}_{q} with Frobenius eigenvalues RS={α1,α2,α¯1,α¯2}R_{S}=\mathopen{}\mathclose{{}\left\{\alpha_{1},\alpha_{2},\overline{\alpha}_{1},\overline{\alpha}_{2}}\right\} and suppose that S(2)\colonequalsS×𝐅q𝐅q2S_{(2)}\colonequals S\times_{\mathbf{F}_{q}}\mathbf{F}_{q^{2}} is isogenous to E2E^{2} for some ordinary elliptic curve E/𝐅q2E/\mathbf{F}_{q^{2}}. In this case, {α12,α¯12}=RE={α22,α¯22}\mathopen{}\mathclose{{}\left\{\alpha_{1}^{2},\overline{\alpha}_{1}^{2}}\right\}=R_{E}=\mathopen{}\mathclose{{}\left\{\alpha_{2}^{2},\overline{\alpha}_{2}^{2}}\right\}. Normalizing, and possibly after re-indexing, we see that either u2=u1u_{2}=u_{1} or u2=u1u_{2}=-u_{1}. Since SS is simple and ordinary, the characteristic polynomial of Frobenius of SS is irreducible (see Remark 2.1.1), and we must have u2=u1u_{2}=-u_{1}. The Serre–Frobenius groups of SS and S(2)S_{(2)} are calculated as follows.

𝖲𝖥(S)\displaystyle\mathsf{SF}(S) ={[u1r(u1)ru¯1r(u¯1)r]:r𝐙}¯={[uuu¯u¯]:u𝖴(1)},\displaystyle=\overline{\mathopen{}\mathclose{{}\left\{\begin{bmatrix}u_{1}^{r}&&&\\ &(-u_{1})^{r}&&\\ &&\overline{u}_{1}^{r}&\\ &&&(-\overline{u}_{1})^{r}\end{bmatrix}:r\in\mathbf{Z}}\right\}}=\mathopen{}\mathclose{{}\left\{\begin{bmatrix}u&&&\\ &-u&&\\ &&\overline{u}&\\ &&&-\overline{u}\end{bmatrix}:u\in\mathsf{U}(1)}\right\},
𝖲𝖥(S(2))\displaystyle\mathsf{SF}(S_{(2)}) ={[u12r(u1)2ru¯12r(u¯1)2r]:r𝐙}¯={[uuu¯u¯]:u𝖴(1)}.\displaystyle=\overline{\mathopen{}\mathclose{{}\left\{\begin{bmatrix}u_{1}^{2r}&&&\\ &(-u_{1})^{2r}&&\\ &&\overline{u}_{1}^{2r}&\\ &&&(-\overline{u}_{1})^{2r}\end{bmatrix}:r\in\mathbf{Z}}\right\}}=\mathopen{}\mathclose{{}\left\{\begin{bmatrix}u&&&\\ &u&&\\ &&\overline{u}&\\ &&&\overline{u}\end{bmatrix}:u\in\mathsf{U}(1)}\right\}.

The sequence of normalized traces, henceforth referred to as the a1a_{1}-sequence, is given by xr(S)=0x_{r}(S)=0 when rr is odd, and xr(S)=2u1r+2u¯1rx_{r}(S)=2u_{1}^{r}+2\overline{u}_{1}^{r} when rr is even. Extending the base field to 𝐅q2\mathbf{F}_{q^{2}} yields the sequence of normalized traces xr(S(2))=x2r(S)=2xr(E)x_{r}(S_{(2)})=x_{2r}(S)=2x_{r}(E). The data of the embedding 𝖲𝖥(S)𝖴𝖲𝗉4(𝐂)\mathsf{SF}(S)\subseteq\mathsf{USp}_{4}(\mathbf{C}) precisely captures the (non-trivial) multiplicative relations between the Frobenius eigenvalues.

The sequence of normalized traces xr(S)x_{r}(S) is equidistributed with respect to the pushforward of the Haar measure under the trace map 𝖲𝖥(S)𝖴𝖲𝗉4(𝐂)[4,4]\mathsf{SF}(S)\subseteq\mathsf{USp}_{4}(\mathbf{C})\to[-4,4] given by diag(z1,z2,z¯1,z¯2)z1+z2+z¯1+z¯2\operatorname{diag}(z_{1},z_{2},\overline{z}_{1},\overline{z}_{2})\mapsto z_{1}+z_{2}+\overline{z}_{1}+\overline{z}_{2}, and similarly for S(2)S_{(2)}. These can be computed explicitly for SS and S(2)S_{(2)} as

(4) 12δ0+dx2π16x2 and dxπ16x2,\tfrac{1}{2}\delta_{0}+\frac{\,\mathrm{d}x}{2\pi\sqrt{16-x^{2}}}\quad\text{ and }\quad\frac{\,\mathrm{d}x}{\pi\sqrt{16-x^{2}}},

where dx\,\mathrm{d}x is the restriction of the Haar measure to [4,4][-4,4], and δ0\delta_{0} is the Dirac measure supported at 0.

For instance, choose the surface SS to be in the isogeny class with LMFDB [17] label222Recall the labelling convention for isogeny classes of abelian varieties over finite fields in the LMFDB is g.q.iso where g is the dimension, q is the cardinality of the base field, and iso specifies the isogeny class by writing the coefficients of the Frobenius polynomial in base 26. 2.5.a_ab and Weil polynomial P(T)=T4T2+25P(T)=T^{4}-T^{2}+25. This isogeny class is ordinary and simple, but not geometrically simple. Indeed, S(2)S_{(2)} is in the isogeny class 1.25.ab2=\texttt{1.25.ab}^{2}= 2.25.ac_bz corresponding to the square of an ordinary elliptic curve. The corresponding a1a_{1}-histograms describing the frequency of the sequence (xr)r=1(x_{r})_{r=1}^{\infty} are depicted in Figure 1. Each graph represents a histogram of 166=1677721616^{6}=16777216 samples placed into 46=40964^{6}=4096 buckets partitioning the interval [2g,2g][-2g,2g]. The vertical axis has been suitably scaled, with the height of the uniform distribution, 1/4g1/4g, indicated by a gray line.

Refer to caption
Refer to caption
Figure 1. a1a_{1}-histograms for 2.5.a_ab and 2.25.ac_bz.

1.2. Relation to other work

The reason for adopting the name “Serre–Frobenius group” is that the Lie group 𝖲𝖥(A)\mathsf{SF}(A) is closely related to Serre’s Frobenius torus [27], as explained in Remark 2.3.4.

1.2.1. Angle rank

In this article, we study multiplicative relations between Frobenius eigenvalues, a subject studied extensively by Zarhin [37, 38, 16, 39, 40]. Our classification relies heavily on being able to understand multiplicative relations in low dimension, and we use results of Zarhin in completing parts of it. The number of multiplicative relations is quantified by the angle rank, an invariant studied in [8], [7] for absolutely simple abelian varieties by elucidating its interactions with the Galois group and Newton polygon of the Frobenius polynomial. We study the angle rank as a stepping stone to classifying the full Serre–Frobenius group. While our perspective differs from that in [8], the same theme is continued here: the Serre–Frobenius groups depend heavily on the Galois group of the Frobenius polynomial. It is worth noting that here that the results about the angle rank in the non-absolutely simple case cannot be pieced together by knowing the results in the absolutely simple cases (for instance, see Zywina’s exposition of Shioda’s example [42, Remark 1.16]).

1.2.2. Sato–Tate groups

The Sato–Tate group of an abelian variety defined over a number field controls the distribution of the Frobenius of the reduction modulo prime ideals, and it is defined via its \ell-adic Galois representation (see [30, Section 3.2]). The Serre–Frobenius group can also be defined via \ell-adic representations in an analogous way: it is conjugate to a maximal compact subgroup of the image of Galois representation ρA,:Gal(𝐅¯q/𝐅q)Aut(VA)𝐂\rho_{A,\ell}\colon\operatorname{Gal}(\overline{\mathbf{F}}_{q}/\mathbf{F}_{q})\to\operatorname{Aut}(V_{\ell}A)\otimes\mathbf{C}, where VAV_{\ell}A is the \ell-adic Tate vector space. Therefore it is natural to expect that the Sato–Tate and the Serre–Frobenius group are related to each other. The following observations support this claim:

  • Assuming standard conjectures, the connected component of the identity of the Sato–Tate group can be recovered from knowing the Frobenius polynomial at two suitably chosen primes ([42, Theorem 1.6]).

  • Several abelian Sato–Tate groups (see [10, 11]) appear as Serre–Frobenius groups of abelian varieties over finite fields. The ones with maximal angle rank are as below.

    • 𝖴(1)\mathsf{U}(1) is the Sato–Tate group of an elliptic curve with complex multiplication over any number field that contains the CM field (see 1.2.B.1.1a). It is also the Serre–Frobenius group of any ordinary elliptic curve (see Figure 2(a)), and the a1a_{1}-moments coincide.

    • 𝖴(1)2\mathsf{U}(1)^{2} is the Sato–Tate group of weight 1 and degree 4 (see 1.4.D.1.1a). It is also the Serre–Frobenius group of an abelian surface with maximal angle rank (see Figure 4(a)), and the a1a_{1}-moments coincide.

    • 𝖴(1)3\mathsf{U}(1)^{3} is the Sato–Tate group of weight 1 and degree 6 (see 1.6.H.1.1a). It is also the Serre-Frobenius group of abelian threefolds with maximal angle rank (see Figure 8), and the a1a_{1}-moments coincide.

1.3. Outline

In Section 2, we give some background on abelian varieties over finite fields, expand on the definition of the Serre–Frobenius group, and describe how it controls the distribution of traces of high powers of Frobenius. In Section 3, we prove some preliminary results on the geometric isogeny types of abelian varieties of dimension g3g\leq 3 and gg odd prime. We also recall some results about Weil polynomials of supersingular abelian varieties, and Zarhin’s notion of neatness. In Section 3.2, we discuss the classification in the case of simple ordinary abelian varieties of odd prime dimension. In Section 4, Section 5, and Section 6, we give a complete classification of the Serre–Frobenius group for dimensions 1, 2, and 3 respectively. A list of tables containing different pieces of the classification follows this section.

1.4. Notation

Throughout this paper, AA will denote a gg-dimensional abelian variety over a finite field 𝐅q\mathbf{F}_{q} of characteristic pp. The polynomial PA(T)=i=02gaiT2giP_{A}(T)=\sum_{i=0}^{2g}a_{i}T^{2g-i} will denote the characteristic polynomial of the qq-Frobenius endomorphism πA\pi_{A} acting on the Tate module of AA, and hA(T)h_{A}(T) its minimal polynomial. The set of roots of PA(T)P_{A}(T) is denoted by RAR_{A}. We usually write α1,α¯1,αg,α¯gRA\alpha_{1},\overline{\alpha}_{1}\dots,\alpha_{g},\overline{\alpha}_{g}\in R_{A} for the Frobenius eigenvalues, where α¯i=q/αi\overline{\alpha}_{i}=q/\alpha_{i}. In the case that PA(T)P_{A}(T) is a power of hA(T)h_{A}(T), we will denote this power by eAe_{A} (See 2.1). The subscript ()(r)(\cdot)_{(r)} will denote the base change of any object or map to 𝐅qr\mathbf{F}_{q^{r}}. The group UAU_{A} will denote the multiplicative group generated by the normalized eigenvalues of Frobenius, δA\delta_{A} its rank and mAm_{A} the order of its torsion subgroup. The group ΓA\Gamma_{A} will denote the multiplicative group generated by {α1,α2αg,q}\{\alpha_{1},\alpha_{2}\ldots\alpha_{g},q\}. In Section 5, SS will be used to denote an abelian surface, while in Section 6, XX will be used to denote a threefold.

2. Frobenius multiplicative groups

In this section we introduce the Serre–Frobenius group of AA and explain how it is related to Serre’s theory of Frobenius tori [27]. We do this from the perspective of the theory of algebraic groups of multiplicative type, as in [19, Chapter 12]. We start by recalling some facts about abelian varieties over finite fields.

2.1. Background on Abelian varieties over finite fields

Fix AA a gg dimensional abelian variety over 𝐅q\mathbf{F}_{q}. A qq-Weil number is an algebraic integer α\alpha such that |ϕ(α)|=q|\phi(\alpha)|=\sqrt{q} for every embedding ϕ:𝐐(α)𝐂\phi\colon\mathbf{Q}(\alpha)\to\mathbf{C}. Let PA(T)P_{A}(T) denote the characteristic polynomial of the Frobenius endomorphism acting on the \ell-adic Tate module of AA. The polynomial PA(T)P_{A}(T) is monic of degree 2g2g, and Weil [34] showed that its roots are qq-Weil numbers; we denote the set of roots of PA(T)P_{A}(T) by RA\colonequals{α1,α2,αg,αg+1,,α2g}R_{A}\colonequals\{\alpha_{1},\alpha_{2}\ldots,\alpha_{g},\alpha_{g+1},\dots,\alpha_{2g}\} with α¯j\colonequalsαg+j=q/αj\overline{\alpha}_{j}\colonequals\alpha_{g+j}=q/\alpha_{j} for j{1,,g}j\in\mathopen{}\mathclose{{}\left\{1,\dots,g}\right\}. The seminal work of Honda [13] and Tate [31, 32] classifies the isogeny decomposition type of AA in terms of the factorization of PA(T)P_{A}(T). In particular, if AA is simple, we have that PA(T)=hA(T)eAP_{A}(T)=h_{A}(T)^{e_{A}} where hA(T)h_{A}(T) is the minimal polynomial of the Frobenius endomorphism and eAe_{A} is the degree, i.e., the square root of the dimension of the central simple algebra End0(A)\colonequalsEnd(A)𝐐\mathrm{End}^{0}(A)\colonequals\mathrm{End}(A)\otimes\mathbf{Q} over its center. The Honda–Tate theorem gives a bijective correspondence between isogeny classes of simple abelian varieties over 𝐅q\mathbf{F}_{q} and conjugacy classes of qq-Weil numbers, sending the isogeny class determined by AA to the set of roots RAR_{A}. Further, the isogeny decomposition AA1×A2×AkA\sim A_{1}\times A_{2}\ldots\times A_{k} can be read from the factorization PA(T)=i=1kPAi(T)P_{A}(T)=\prod_{i=1}^{k}P_{A_{i}}(T).

Writing PA(T)=i=02gaiT2giP_{A}(T)=\sum_{i=0}^{2g}a_{i}T^{2g-i}, the qq-Newton polygon of AA is the lower convex hull of the set of points {(i,ν(ai))𝐑2:ai0}\{(i,\nu(a_{i}))\in\mathbf{R}^{2}:a_{i}\neq 0\} where ν\nu is the pp-adic valuation normalized so that ν(q)=1\nu(q)=1. The Newton polygon is isogeny invariant. Define the pp-rank of AA as the number of slope 0 segments of the Newton polygon. An abelian variety is called ordinary if it has maximal pp-rank, i.e., its pp-rank is equal to gg. It is called almost ordinary if it has pp-rank g1g-1; equivalently, the set of slopes of its Newton polygon is {0,1/2,1}\mathopen{}\mathclose{{}\left\{0,1/2,1}\right\} and the slope 1/21/2 has length 22. An abelian variety is called supersingular if all the slopes of the Newton polygon are equal to 1/21/2. The field L=LA\colonequals𝐐(α1,,αg)L=L_{A}\colonequals\mathbf{Q}(\alpha_{1},\dots,\alpha_{g}) is the splitting field of the Frobenius polynomial. By definition, the Galois group Gal(L/𝐐)\operatorname{Gal}(L/\mathbf{Q}) acts on the roots RAR_{A} by permuting them.

Remark 2.1.1.

When an abelian variety AA over 𝐅q\mathbf{F}_{q} is simple and ordinary, then PA(T)P_{A}(T) is irreducible and its endomorphism algebra is a field ([33, Theorem 7.2]).

Notation.

Whenever AA is fixed or clear from context, we will omit the subscript corresponding to it from the notation described above. In particular, we will use P(T),h(T)P(T),h(T) and ee instead of PA(T),hA(T)P_{A}(T),h_{A}(T) and eAe_{A}.

2.2. Angle groups

Denote by Γ\colonequalsΓA\Gamma\colonequals\Gamma_{A} the multiplicative subgroup of 𝐂×\mathbf{C}^{\times} generated by the set of Frobenius eigenvalues RAR_{A}, and let Γ(r)\colonequalsΓA(r)\Gamma_{(r)}\colonequals\Gamma_{A_{(r)}} for every r1r\geq 1. Since αq/α\alpha\mapsto q/\alpha is a permutation of RAR_{A}, the set {α1,,αg,q}\mathopen{}\mathclose{{}\left\{\alpha_{1},\dots,\alpha_{g},q}\right\} is a set of generators for Γ\Gamma; that is, every γΓ\gamma\in\Gamma can be written as

(5) γ=qkj=1gαjkj\gamma=q^{k}\prod_{j=1}^{g}\alpha_{j}^{k_{j}}

for some (k,k1,,kg)𝐙g+1(k,k_{1},\dots,k_{g})\in\mathbf{Z}^{g+1}.

Since Γ\Gamma is a subgroup of 𝐐¯×\overline{\mathbf{Q}}^{\times}, it is naturally a Gal(𝐐¯/𝐐)\operatorname{Gal}(\overline{\mathbf{Q}}/\mathbf{Q})-module. However, this perspective is not necessary for our applications. This group is denoted as ΦA\Phi_{A} in [42].

Definition 2.2.1.

We define the angle group of AA to be U\colonequalsUAU\colonequals U_{A}, the multiplicative subgroup of 𝖴(1)\mathsf{U}(1) generated by the unitarized eigenvalues {uj\colonequalsαj/q:j=1,,g\{u_{j}\colonequals\alpha_{j}/\sqrt{q}:j=1,\dots,g}. When AA is fixed, for every r1r\geq 1 we abbreviate U(r)\colonequalsUA(r)U_{(r)}\colonequals U_{A_{(r)}}.

Definition 2.2.2.

The angle rank of an abelian variety A/𝐅qA/\mathbf{F}_{q} is the rank of the finitely generated abelian group UAU_{A}. It is denoted by δA\colonequalsrkUA\delta_{A}\colonequals\operatorname{rk}U_{A}. The angle torsion order mAm_{A} is the order of the torsion subgroup of UAU_{A}, so that UA𝐙δA𝐙/mA𝐙U_{A}\cong\mathbf{Z}^{\delta_{A}}\oplus\mathbf{Z}/m_{A}\mathbf{Z}.

The angle rank δ\delta is by definition an integer between 0 and gg. When δ=g\delta=g, there are no multiplicative relations among the normalized eigenvalues. In other words, there are no additional relations among the generators of ΓA\Gamma_{A} apart from the ones imposed by the Weil conjectures. If AA is absolutely simple, the maximal angle rank condition also implies that the Tate conjecture holds for all powers of AA (see Remark 1.3 in [8]). On the other extreme, δ=0\delta=0 if and only if AA is supersingular (See Example 5.1 [7]).

Remark 2.2.3.

The angle rank is invariant under base extension: δ(A)=δ(A(r))\delta(A)=\delta(A_{(r)}) for every r1r\geq 1. Indeed, any multiplicative relation between {u1r,,ugr}\mathopen{}\mathclose{{}\left\{u_{1}^{r},\dots,u_{g}^{r}}\right\} is a multiplicative relation between {u1,,ug}\mathopen{}\mathclose{{}\left\{u_{1},\dots,u_{g}}\right\}. We have that UA/Tors(UA)UA(r)/Tors(UA(r))U_{A}/\mathrm{Tors}(U_{A})\cong U_{A_{(r)}}/\mathrm{Tors}(U_{A_{(r)}}) for every positive integer rr. In particular, UA/Tors(UA)UA(m)U_{A}/\mathrm{Tors}(U_{A})\cong U_{A_{(m)}} where m=mAm=m_{A} is the angle torsion order of AA.

Example 2.2.4 (Extension and restriction of scalars).

Let A/𝐅qA/\mathbf{F}_{q} be an abelian variety with Frobenius polynomial PA(T)=(Tαi)𝐂[T]P_{A}(T)=\prod(T-\alpha_{i})\in\mathbf{C}[T] and angle group UA=u1,,ugU_{A}=\langle u_{1},\dots,u_{g}\rangle. Then, the extension of scalars A(r)A_{(r)} has Frobenius polynomial P(r)(T)=(Tαir)P_{(r)}(T)=\prod(T-\alpha_{i}^{r}) and angle group UA(r)=u1r,,ugrUAU_{A_{(r)}}=\langle u_{1}^{r},\dots,u_{g}^{r}\rangle\subset U_{A}. On the other hand, if B/𝐅qrB/\mathbf{F}_{q^{r}} is an abelian variety for some r1r\geq 1, and A/𝐅qA/\mathbf{F}_{q} is the Weil restriction of BB to 𝐅q\mathbf{F}_{q}, then PA(T)=PB(Tr)P_{A}(T)=P_{B}(T^{r}) and UA=UB,ζrUBU_{A}=\langle U_{B},\zeta_{r}\rangle\supset U_{B}. See [6].

2.3. The Serre–Frobenius group

For every locally compact abelian group GG, denote by G^\widehat{G} its Pontryagin dual; this is the topological group of continuous group homomorphisms G𝖴(1)G\to\mathsf{U}(1). It is well known that GG^G\mapsto\widehat{G} gives an anti-equivalence of categories from the category of locally compact abelian groups to itself. Moreover, this equivalence preserves exact sequences, and every such GG is canonically isomorphic to its double dual via the evaluation isomorphism. See [23] for the original reference and [20] for a gentle introduction.

Recall that we defined the Serre–Frobenius group of AA as the topological group generated by the matrix diag(u1,,ug,u¯1,,u¯g)\operatorname{diag}(u_{1},\dots,u_{g},\overline{u}_{1},\dots,\overline{u}_{g}) inside of 𝖴𝖲𝗉2g(𝐂)\mathsf{USp}_{2g}(\mathbf{C}) (see 1.0.1).

Remark 2.3.1.

The group 𝖴(1)g\mathsf{U}(1)^{g} embeds into 𝖴𝖲𝗉2g(𝐂)\mathsf{USp}_{2g}(\mathbf{C}) as a maximal torus via 𝐳diag(𝐳,𝐳¯)\mathbf{z}\mapsto\mathrm{diag}(\mathbf{z},\overline{\mathbf{z}}), so a different choice of indexing of the Frobenius eigenvalues yields a conjugate subgroup g1𝖲𝖥(A)gg^{-1}\mathsf{SF}(A)g, where gg is an element of the Weyl group N𝖴𝖲𝗉2g(𝐂)(𝖴(1)g)/𝖴(1)gN_{\mathsf{USp}_{2g}(\mathbf{C})}(\mathsf{U}(1)^{g})/\mathsf{U}(1)^{g}. This Weyl group is isomorphic to the group Sg±S_{g}^{\pm} of signed permutation matrices; the generic Galois group of a complex multiplication polynomial of degree 2g2g.

Notation.

In light of Remark 2.3.1, we identify 𝖴(1)g\mathsf{U}(1)^{g} with the group

[diag(z1,,zg)diag(z¯1,,z¯g)]𝖴𝖲𝗉2g(𝐂),\begin{bmatrix}\operatorname{diag}(z_{1},\dots,z_{g})&\\ &\operatorname{diag}(\overline{z}_{1},\dots,\overline{z}_{g})\end{bmatrix}\subset\mathsf{USp}_{2g}(\mathbf{C}),

and the vector 𝐮\colonequals(u1,,ug)\mathbf{u}\colonequals(u_{1},\cdots,u_{g}) with the matrix diag(𝐮,𝐮¯)\operatorname{diag}(\mathbf{u},\overline{\mathbf{u}}). The embedding of 𝖲𝖥(A)\mathsf{SF}(A) into 𝖴𝖲𝗉2g(𝐂)\mathsf{USp}_{2g}(\mathbf{C}) is completely determined by the topological generator of (u1,,ug)(u_{1},\dots,u_{g}), a vector of normalized Frobenius eigenvalues. We will represent the embedding of 𝖲𝖥(A)\mathsf{SF}(A) into 𝖴𝖲𝗉2g(𝐂)\mathsf{USp}_{2g}(\mathbf{C}) (up to conjugation) by giving the topological generator of the Serre–Frobenius group.

The following lemma will help us identify the isomorphism class of 𝖲𝖥(A)\mathsf{SF}(A) as a compact abelian group.

Lemma 2.3.2.

The Serre–Frobenius group of an abelian variety AA has character group UAU_{A}. In particular, 𝖲𝖥(A)U^A\mathsf{SF}(A)\cong\widehat{U}_{A} canonically via the evaluation isomorphism.

Proof.

We have an injection UA𝖲𝖥(A)^U_{A}\to\widehat{\mathsf{SF}(A)} given by mapping γ\gamma to the character ϕγ\phi_{\gamma} that maps the topological generator 𝐮\mathbf{u} to γ\gamma. To see that this map is surjective, observe that by the exactness of Pontryagin duality, the inclusion 𝖲𝖥(A)𝖴(1)g\mathsf{SF}(A)\hookrightarrow\mathsf{U}(1)^{g} induces a surjection 𝐙g=𝖴(1)^g𝖲𝖥(A)^\mathbf{Z}^{g}=\widehat{\mathsf{U}(1)}^{g}\to\widehat{\mathsf{SF}(A)}. Explicitly, this tells us that every character of 𝖲𝖥(A)\mathsf{SF}(A) is given by ϕ(z1,,zg)=z1m1zgmg\phi(z_{1},\dots,z_{g})=z_{1}^{m_{1}}\cdots z_{g}^{m_{g}} for some (m1,,mg)𝐙g(m_{1},\dots,m_{g})\in\mathbf{Z}^{g}. By continuity, every character ϕ\phi of 𝖲𝖥(A)\mathsf{SF}(A) is completely determined by ϕ(𝐮)\phi(\mathbf{u}). In particular, we have that ϕ(𝐮)=u1m1ugmgUA\phi(\mathbf{u})=u_{1}^{m_{1}}\cdots u_{g}^{m_{g}}\in U_{A}. ∎

The following theorem should be compared to [30, Theorem 3.12]

Theorem 2.3.3 (Theorem 1.0.2).

Let AA be an abelian variety defined over 𝐅q\mathbf{F}_{q}. Then

𝖲𝖥(A)𝖴(1)δ×Cm,\mathsf{SF}(A)\cong\mathsf{U}(1)^{\delta}\times C_{m},

where δ=δA\delta=\delta_{A} is the angle rank and m=mAm=m_{A} is the angle torsion order. Furthermore, the connected component of the identity is

𝖲𝖥(A)=𝖲𝖥(A(m)).\mathsf{SF}(A)^{\circ}=\mathsf{SF}(A_{(m)}).
Proof.

Since every finite subgroup of 𝖴(1)\mathsf{U}(1) is cyclic, the torsion part of the finitely generated group UAU_{A} is generated by some primitive mm-th root of unity ζm\zeta_{m}. The group U(m)U_{(m)} is torsion free by Remark 2.2.3. We thus have the split short exact sequence

(6) 1{1}ζm{\langle\zeta_{m}\rangle}UA{U_{A}}U(m){U_{(m)}}1.{1.}uum\scriptstyle{u\mapsto u^{m}}

After dualizing, we get:

(7) 1{1}𝖲𝖥(A(m)){\mathsf{SF}(A_{(m)})}𝖲𝖥(A){\mathsf{SF}(A)}ζm{\langle\zeta_{m}\rangle}1.{1.}

We conclude that 𝖲𝖥(A)=𝖲𝖥(A(m))\mathsf{SF}(A)^{\circ}=\mathsf{SF}(A_{(m)}) and 𝖲𝖥(A)/𝖲𝖥(A)ζm\mathsf{SF}(A)/\mathsf{SF}(A)^{\circ}\cong\langle\zeta_{m}\rangle. ∎

Remark 2.3.4.

By definition, UAU_{A} is the image of ΓA\Gamma_{A} under the radial projection ψ:𝐂×𝖴(1),zz/|z|\psi\colon\mathbf{C}^{\times}\to\mathsf{U}(1),z\mapsto z/|z|. Thus, we have a short exact sequence

(8) 1{1}ΓA𝐑>0{\Gamma_{A}\cap\mathbf{R}_{>0}}ΓA{\Gamma_{A}}UA{U_{A}}1,{1,}ψ|Γ\scriptstyle{\psi|_{\Gamma}}

which is split by the section ujαju_{j}\mapsto\alpha_{j}. The kernel Γ𝐑>0\Gamma\cap\mathbf{R}_{>0} is free of rank 11 and contains the group q𝐙q^{\mathbf{Z}}. The relation between the Serre–Frobenius group 𝖲𝖥(A)\mathsf{SF}(A) and Serre’s Frobenius Torus (see [27, Volume IV, 133.], [4, Section 3]) can be understood via their character groups.

  • The (Pontryagin) character group of 𝖲𝖥(A)\mathsf{SF}(A) is UAU_{A}.

  • The (algebraic) character group of the Frobenius torus of AA is the torsion free part of ΓA\Gamma_{A}.

2.4. Equidistribution results

Let (Y,μ)(Y,\mu) be a measure space in the sense of Serre (see Appendix A.1 in [26]). Recall that a sequence (yr)r=1Y(y_{r})_{r=1}^{\infty}\subset Y is μ\mu-equidistributed if for every continuous function f:Y𝐂f\colon Y\to\mathbf{C} we have that

(9) Yfμ=limn1nr=1nf(yr).\int_{Y}f\mu=\lim_{n\to\infty}\frac{1}{n}\sum_{r=1}^{n}f(y_{r}).

In our setting, YY will be a compact abelian Lie group with probability Haar measure μ\mu. We have the following lemma.

Lemma 2.4.1.

Let GG be a compact group, and hGh\in G. Let HH be the closure of the group generated by hh. Then, the sequence (hr)r=1(h^{r})_{r=1}^{\infty} is equidistributed in HH with respect to the Haar measure μH\mu_{H}.

Proof.

For a non-trivial character ϕ:H𝐂×\phi\colon H\to\mathbf{C}^{\times}, the image of the generator ϕ(h)=u𝖴(1)\phi(h)=u\in\mathsf{U}(1) is non-trivial. We see that

limn1nr=1nϕ(hr)=limn1nr=1nur=0,\lim_{n\to\infty}\frac{1}{n}\sum_{r=1}^{n}\phi(h^{r})=\lim_{n\to\infty}\frac{1}{n}\sum_{r=1}^{n}u^{r}=0,

both when uu has finite or infinite order. The latter case follows from Weyl’s equidistribution theorem in 𝖴(1)\mathsf{U}(1). The result follows from Lemma 1 in [26, I-19] and the Peter–Weyl theorem. ∎

Corollary 2.4.2 (1.1.1).

Let AA be a gg-dimensional abelian variety defined over 𝐅q\mathbf{F}_{q}. Then, the sequence (xr)r=1(x_{r})_{r=1}^{\infty} of normalized traces of Frobenius is equidistributed in [2g,2g][-2g,2g] with respect to the pushforward of the Haar measure on 𝖲𝖥(A)𝖴𝖲𝗉2g(𝐂)\mathsf{SF}(A)\subseteq\mathsf{USp}_{2g}(\mathbf{C}) via

𝖲𝖥(A)𝖴𝖲𝗉2g(𝐂)[2g,2g],MTr(M).\mathsf{SF}(A)\subseteq\mathsf{USp}_{2g}(\mathbf{C})\to[-2g,2g],\quad M\mapsto\operatorname{Tr}(M).
Proof.

By Lemma 2.4.1, the sequence (𝐮r)r=1(\mathbf{u}^{r})_{r=1}^{\infty} is equidistributed in 𝖲𝖥(A)\mathsf{SF}(A) with respect to the Haar measure μ𝖲𝖥(A)\mu_{\mathsf{SF}(A)}. By definition, the sequence (xr)r=1(x_{r})_{r=1}^{\infty} is equidistributed with respect to the pushforward measure, and it is invariant under relabelling of the Frobenius eigenvalues. ∎

Remark 2.4.3 (Maximal angle rank).

When AA has maximal angle rank δ=g\delta=g, the Serre–Frobenius group is the full torus 𝖴(1)g\mathsf{U}(1)^{g}, and the sequence of normalized traces of Frobenius is equidistributed with respect to the pushforward of the measure μ𝖴(1)g\mu_{\mathsf{U}(1)^{g}}; which we denote by λg(x)\lambda_{g}(x) following the notation333Beware of the different choice of normalization. We chose to use the interval [2g,2g][-2g,2g] instead of [1,1][-1,1] to be able to compare our distributions with the Sato–Tate distributions of abelian varieties defined over number fields. in [1].

3. Preliminary Results

For this entire section, we let AA be an abelian variety over 𝐅q\mathbf{F}_{q}, where q=pdq=p^{d} for some prime pp. Recall from Section 1 that an abelian variety AA splits over a field extension 𝐅qm\mathbf{F}_{q^{m}} if A(m)A1×A2A_{(m)}\sim A_{1}\times A_{2} and dimA1,dimA2<dimA\dim A_{1},\dim A_{2}<\dim A, i.e., if AA obtains at least one isogeny factor after extending scalars to 𝐅qm\mathbf{F}_{q^{m}}. We say that AA splits completely over 𝐅qm\mathbf{F}_{q^{m}} if A(m)A1×A2××AkA_{(m)}\sim A_{1}\times A_{2}\times\ldots\times A_{k}, where each AiA_{i} is an absolutely simple abelian variety defined over 𝐅qm\mathbf{F}_{q^{m}}. In other words, AA acquires its geometric isogeny decomposition over 𝐅qm\mathbf{F}_{q^{m}}. We define the splitting degree of AA to be the minimal positive integer mm such that AA splits completely over 𝐅qm\mathbf{F}_{q^{m}}.

3.1. Geometric products of elliptic curves

We begin by stating an important lemma, attributed to Bjorn Poonen in [15].

Lemma 3.1.1 (Poonen).

If E1,,EnE_{1},\dots,E_{n} are nn pairwise geometrically non-isogenous elliptic curves over 𝐅q\mathbf{F}_{q}, then their eigenvalues of Frobenius α1,,αn\alpha_{1},\dots,\alpha_{n} are multiplicatively independent.

In fact, for abelian varieties that split completely as products of elliptic curves, we can explicitly describe the Serre–Frobenius group.

Lemma 3.1.2.

Let B/𝐅qB/\mathbf{F}_{q} be an abelian variety that splits over 𝐅qm\mathbf{F}_{q^{m}} as a power of an ordinary elliptic curve, where m1m\geq 1 is the splitting degree of BB. Then, 𝖲𝖥(B)𝖴(1)×Cm\mathsf{SF}(B)\cong\mathsf{U}(1)\times C_{m}. Furthermore, if m>1m>1, then

𝖲𝖥(B)={(u,ξ1νu,ξ2νu,,ξg1νu):u𝖴(1),ν𝐙/m𝐙}𝖴(1)g,\mathsf{SF}(B)=\mathopen{}\mathclose{{}\left\{(u,\xi_{1}^{\nu}u,\xi_{2}^{\nu}u,\dots,\xi_{g-1}^{\nu}u):u\in\mathsf{U}(1),\nu\in\mathbf{Z}/m\mathbf{Z}}\right\}\subset\mathsf{U}(1)^{g},

with ξ1,,ξg1\xi_{1},\dots,\xi_{g-1} mm-th roots of unity whose orders have least common multiple mm. In particular, when BB is simple then all the ξj\xi_{j} are distinct, primitive and g1φ(m)g-1\leq\varphi(m).

Proof.

Angle rank is invariant under base change, so δB=δEg=1\delta_{B}=\delta_{E^{g}}=1. It remains to show that the angle torsion order mBm_{B} equals mm. If m=1m=1, then B=EgB=E^{g} and there is nothing to show. Assume that m>1m>1. Since B(m)EgB_{(m)}\sim E^{g}, we have that PB,(m)(T)=PE(T)gP_{B,(m)}(T)=P_{E}(T)^{g}. If we denote by γ1,γ¯1,γg,γ¯g\gamma_{1},\overline{\gamma}_{1},\ldots\gamma_{g},\overline{\gamma}_{g} and π1,π¯1\pi_{1},\overline{\pi}_{1} the Frobenius eigenvalues of BB and EE respectively, we have that {γ1m,γ¯1m,,γgm,γ¯gm}={π1,π¯1}\mathopen{}\mathclose{{}\left\{\gamma_{1}^{m},\overline{\gamma}_{1}^{m},\dots,\gamma_{g}^{m},\overline{\gamma}_{g}^{m}}\right\}=\mathopen{}\mathclose{{}\left\{\pi_{1},\overline{\pi}_{1}}\right\}. Possibly after relabelling, we have that γj+1=ξjγ1\gamma_{j+1}=\xi_{j}\gamma_{1} for j=1,,g1j=1,\dots,g-1, where the ξj\xi_{j}’s are mm-th roots of unity and the minimality of mm ensures that the lcm of the orders of the ξj\xi_{j}’s is mm. This shows that CmUBC_{m}\subset U_{B}, so that mmBm\mid m_{B}. On the other hand, we have that 𝖲𝖥(B(m))=𝖲𝖥(Eg)𝖴(1)\mathsf{SF}(B_{(m)})=\mathsf{SF}(E^{g})\cong\mathsf{U}(1) is connected. This implies that mBmm_{B}\mid m and we conclude that 𝖲𝖥(B)𝖴(1)×Cm\mathsf{SF}(B)\cong\mathsf{U}(1)\times C_{m}. Assume now BB is simple, then PB(T)P_{B}(T) is irreducible and hence has no repeated roots, and thus ξjξi\xi_{j}\neq\xi_{i} for every 0<j<i<g0<j<i<g, and every ξi\xi_{i} is primitive. This shows that the set {ξj:j=1,,g1}\mathopen{}\mathclose{{}\left\{\xi_{j}:j=1,\dots,g-1}\right\} has g1g-1 elements, and therefore g1φ(m)g-1\leq\varphi(m). ∎

Lemma 3.1.3.

Let A=B×A1A=B\times A_{1} be an abelian variety over 𝐅q\mathbf{F}_{q} such that A1A_{1} is supersingular with angle torsion order mA1=m1m_{A_{1}}=m_{1} and BB splits over 𝐅qm\mathbf{F}_{q^{m}} as the power of an ordinary elliptic curve, where m1m\geq 1 is the splitting degree of BB. Then, 𝖲𝖥(A)𝖴(1)\mathsf{SF}(A)^{\circ}\cong\mathsf{U}(1) and mA=lcm(m1,m)m_{A}=\operatorname{lcm}(m_{1},m). Furthermore,

𝖲𝖥(A)=diag(𝖲𝖥(B),𝖲𝖥(A1),𝖲𝖥(B)¯,𝖲𝖥(A1)¯)𝖴𝖲𝗉2g(𝐂),\mathsf{SF}(A)=\operatorname{diag}(\mathsf{SF}(B),\mathsf{SF}(A_{1}),\overline{\mathsf{SF}(B)},\overline{\mathsf{SF}(A_{1})})\subset\mathsf{USp}_{2g}(\mathbf{C}),

where 𝖲𝖥(B)¯\overline{\mathsf{SF}(B)} denotes the (pointwise) complex conjugate of 𝖲𝖥(B)\mathsf{SF}(B), and similarly for A1A_{1}.

Proof.

From Lemma 3.1.2, we see that UA=ζm1,ζm,v1U_{A}=\langle\zeta_{m_{1}},\zeta_{m},v_{1}\rangle, where v1=v_{1}= γ1/q\gamma_{1}/\sqrt{q} is a normalized Frobenius eigenvalue of BB and all the other roots γj\gamma_{j} can be written as ζmνjγ1\zeta_{m}^{\nu_{j}}\gamma_{1} with lcmj(ord(ζmνj))=m.\operatorname{lcm}_{j}(\operatorname{ord}(\zeta_{m}^{\nu_{j}}))=m. It follows that UA=Clcm(m1,m)v1U_{A}=C_{\operatorname{lcm}(m_{1},m)}\oplus\langle v_{1}\rangle so that δA=1\delta_{A}=1 and mA=lcm(m1,m)m_{A}=\operatorname{lcm}(m_{1},m). Furthermore, 𝖲𝖥(A)\mathsf{SF}(A) is generated by (v1,ξ1v1,,ξdimB1v1,η1,,ηg1)(v_{1},\xi_{1}v_{1},\dots,\xi_{\dim B-1}v_{1},\eta_{1},\dots,\eta_{g_{1}}) for some ξjμm\xi_{j}\in\mu_{m} with lcmj(ord(ξj))=m\operatorname{lcm}_{j}(\operatorname{ord}(\xi_{j}))=m and ηiμm1\eta_{i}\in\mu_{m_{1}}. ∎

Lemma 3.1.4.

Let BB be an ordinary abelian variety defined over 𝐅q\mathbf{F}_{q} such that BB is geometrically isogenous to a product of elliptic curves. Let mm be the splitting degree of BB, and write B(m)E1g1××EngnB_{(m)}\sim E_{1}^{g_{1}}\times\cdots\times E_{n}^{g_{n}} with EjE_{j} not geometrically isogenous to EiE_{i} for jij\neq i. Then 𝖲𝖥(B)𝖴(1)n×Cm\mathsf{SF}(B)\cong\mathsf{U}(1)^{n}\times C_{m}. Moreover, we can describe the embedding of 𝖲𝖥(B)𝖴(1)g\mathsf{SF}(B)\hookrightarrow\mathsf{U}(1)^{g} as follows:

  1. (1)

    Let r1r\geq 1 be the smallest positive integer such that B(r)B1××BnB_{(r)}\sim B_{1}\times\cdots\times B_{n} decomposes into pairwise non-geometrically isogenous factors.

  2. (2)

    Let mjm_{j} be the splitting degree of BjB_{j}, so that (Bj)(mj)Ejgj(B_{j})_{(m_{j})}\sim E_{j}^{g_{j}}.

Then, m=rlcm(m1,,mn)m=r\operatorname{lcm}(m_{1},\dots,m_{n}) and

𝖲𝖥(B(r))=diag(𝖲𝖥(B1),,𝖲𝖥(Bn),𝖲𝖥(B1)¯,,𝖲𝖥(Bn)¯)𝖴𝖲𝗉2g(𝐂),\mathsf{SF}(B_{(r)})=\operatorname{diag}(\mathsf{SF}(B_{1}),\dots,\mathsf{SF}(B_{n}),\overline{\mathsf{SF}(B_{1})},\dots,\overline{\mathsf{SF}(B_{n})})\subset\mathsf{USp}_{2g}(\mathbf{C}),

where each 𝖲𝖥(Bj)\mathsf{SF}(B_{j}) is as in Lemma 3.1.2, and 𝖲𝖥(Bi)¯\overline{\mathsf{SF}(B_{i})} denotes the (pointwise) complex conjugate of 𝖲𝖥(Bi)\mathsf{SF}(B_{i}).

Proof.

This follows from combining Lemma 3.1.1 with the fact that the Serre–Frobenius group of BB is connected over an extension of degree mm. The proof then proceeds as in Lemma 3.1.2. ∎

3.2. Splitting of simple ordinary abelian varieties of odd prime dimension

In this section, we analyze the splitting behavior of simple ordinary abelian varieties of prime dimension g>2g>2. Our first result is analogous to [14, Theorem 6] for odd primes.

Theorem 3.2.1 (Theorem 1.0.6).

Let AA be a simple ordinary abelian variety defined over 𝐅q\mathbf{F}_{q} of prime dimension g>2g>2. Then, exactly one of the following conditions holds.

  1. (1)

    AA is absolutely simple.

  2. (2)

    AA splits over a degree gg extension of 𝐅q\mathbf{F}_{q} as a power of an elliptic curve, and 𝖲𝖥(A)𝖴(1)×Cg\mathsf{SF}(A)\cong\mathsf{U}(1)\times C_{g}.

  3. (3)

    AA splits over a degree 2g+12g+1 extension of 𝐅q\mathbf{F}_{q} as a power of an elliptic curve, and 𝖲𝖥(A)𝖴(1)×C2g+1\mathsf{SF}(A)\cong\mathsf{U}(1)\times C_{2g+1}. This case only occurs if 2g+12g+1 is also a prime, i.e., if gg is a Sophie Germain prime.

Furthermore, in (2) and (3), we have that

𝖲𝖥(A)={(u,ξ1νu,ξ2νu,,ξg1νu):u𝖴(1),ν𝐙/m𝐙},\mathsf{SF}(A)=\mathopen{}\mathclose{{}\left\{(u,\xi_{1}^{\nu}u,\xi_{2}^{\nu}u,\dots,\xi_{g-1}^{\nu}u):u\in\mathsf{U}(1),\nu\in\mathbf{Z}/m\mathbf{Z}}\right\},

with ξ1,,ξg1\xi_{1},\dots,\xi_{g-1} distinct primitive mm-th roots of unity, for m=gm=g and m=2g+1m=2g+1 respectively.

Proof.

Let α=α1\alpha=\alpha_{1} be a Frobenius eigenvalue of AA, and denote by K=𝐐(α)𝐐[T]/P(T)K=\mathbf{Q}(\alpha)\cong\mathbf{Q}[T]/P(T) the number field generated by α\alpha. Since AA is ordinary, 𝐐(αn)𝐐\mathbf{Q}(\alpha^{n})\neq\mathbf{Q} is a CM-field over 𝐐\mathbf{Q} for every positive integer nn, and P(T)P(T) is irreducible and therefore [𝐐(α):𝐐]=2g[\mathbf{Q}(\alpha):\mathbf{Q}]=2g. Suppose that AA is not absolutely simple, and let mm be the smallest positive integer such that A(m)A_{(m)} splits; by [14, Lemma 4] this is also the smallest mm such that 𝐐(αm)𝐐(α)\mathbf{Q}(\alpha^{m})\subsetneq\mathbf{Q}(\alpha). Since 𝐐(αm)\mathbf{Q}(\alpha^{m}) is also a CM field, it is necessarily an imaginary quadratic number field.

Observe first that mm must be odd. Indeed, if mm was even, then 𝐐(αm/2)=𝐐(α)\mathbf{Q}(\alpha^{m/2})=\mathbf{Q}(\alpha) and [𝐐(αm/2):𝐐(αm)]=2[\mathbf{Q}(\alpha^{m/2}):\mathbf{Q}(\alpha^{m})]=2. This contradicts the fact that [𝐐(α):𝐐]=2g[\mathbf{Q}(\alpha):\mathbf{Q}]=2g, since gg is an odd prime. By [14, Lemma 5], there are two possibilities:

  1. (i)

    P(T)𝐐[Tm]P(T)\in\mathbf{Q}[T^{m}],

  2. (ii)

    K=𝐐(αm,ζm)K=\mathbf{Q}(\alpha^{m},\zeta_{m}).

If (i) holds and P(T)=T2m+bTm+qgP(T)=T^{2m}+bT^{m}+q^{g}, we conclude that m=gm=g and b=agb=a_{g}. In this case, the minimal polynomial of αg\alpha^{g} has degree 2 and is of the form h(g)(T)=(Tαg)(Tα¯g)h_{(g)}(T)=(T-\alpha^{g})(T-\overline{\alpha}^{g}). Note that αg\alpha^{g} and α¯g\overline{\alpha}^{g} are distinct, since AA is ordinary. Thus, P(g)(T)=h(g)(T)gP_{(g)}(T)=h_{(g)}(T)^{g} and AA must split over a degree gg extension as the power of an ordinary elliptic curve.

If (ii) holds, we have that φ(m)2g\varphi(m)\mid 2g. Since m>1m>1 is odd and φ(m)\varphi(m) takes even values, we have two possible options: either φ(m)=2\varphi(m)=2 or φ(m)=2g\varphi(m)=2g. If φ(m)=2\varphi(m)=2, then [K:𝐐(αm)]2[K:\mathbf{Q}(\alpha^{m})]\leq 2 which contradicts the fact that K=𝐐(α)K=\mathbf{Q}(\alpha) is a degree 2g2g extension of 𝐐\mathbf{Q}. Therefore, necessarily, φ(m)=2g\varphi(m)=2g, and 𝐐(α)=𝐐(ζm).\mathbf{Q}(\alpha)=\mathbf{Q}(\zeta_{m}). Recall from elementary number theory that the solutions to this equation are (m,g)=(9,3)(m,g)=(9,3) or (m,g)=(2g+1,g)(m,g)=(2g+1,g) for gg a Sophie Germain prime.

  • (g>3g>3) In this case, (ii) only occurs when 2g+12g+1 is prime.

  • (g=3g=3) In this case, either m=7m=7 or m=9m=9. To conclude the proof, we show that m=9m=9 does not occur. More precisely, we will show that if AA splits over a degree 99 extension, it splits over a degree 33 extension as well. In fact, suppose that K=𝐐(ζ)=𝐐(α)K=\mathbf{Q}(\zeta)=\mathbf{Q}(\alpha) for ζ\zeta a primitive 99th root of unity. The subfield F=𝐐(ζ3)F=\mathbf{Q}(\zeta^{3}) is the only imaginary quadratic subfield of KK, so if a power of α\alpha does not generate KK, it must lie in FF. Suppose α9\alpha^{9} lies in FF. Let σ\sigma be the generator of Gal(K/F)\operatorname{Gal}(K/F) sending ζ\zeta to ζ4\zeta^{4}. The minimal polynomial of α\alpha over FF divides T9α9T^{9}-\alpha^{9}, so σ(α)=αζj\sigma(\alpha)=\alpha\cdot\zeta^{j} for some jj, and σ2(α)=αζ5j\sigma^{2}(\alpha)=\alpha\zeta^{5j}. Since the product of the three conjugates of α\alpha over FF must lie in FF, we have that α3ζ6j=(α)(αζj)(αζ5j)F\alpha^{3}\cdot\zeta^{6j}=(\alpha)(\alpha\cdot\zeta^{j})(\alpha\cdot\zeta^{5j})\in F, which implies that α3F\alpha^{3}\in F and we conclude that AA splits over a degree-33 extension of the base field.

The statement about the structure of the Serre–Frobenius group follows from Lemma 3.1.2. ∎

We thank Everett Howe for explaining to us why the case m=9m=9 above does not occur.

3.3. Zarhin’s notion of neatness

In this section we discuss Zarhin’s notion of neatness, a useful technical definition closely related to the angle rank. Define

(10) RA\colonequals{uj2:αjRA}.R_{A}^{\prime}\colonequals\mathopen{}\mathclose{{}\left\{u_{j}^{2}:\alpha_{j}\in R_{A}}\right\}.

Note that according to our numbering convention, we have that uj1=u¯j=uj+gu_{j}^{-1}=\overline{u}_{j}=u_{j+g} for every j{1,,g}j\in\mathopen{}\mathclose{{}\left\{1,\dots,g}\right\}.

Definition 3.3.1 (Zarhin).

Let AA be an abelian variety defined over 𝐅q\mathbf{F}_{q}. We say that AA is neat if it satisfies the following conditions:

  1. (Na)

    ΓA\Gamma_{A} is torsion free.

  2. (Nb)

    For every function e:RA𝐙e\colon R_{A}^{\prime}\to\mathbf{Z} satisfying

    βRAβe(β)=1,\prod_{\beta\in R_{A}^{\prime}}\beta^{e(\beta)}=1,

    then e(β)=e(β1)e(\beta)=e(\beta^{-1}) for every βRA\beta\in R_{A}^{\prime}.

Remarks 3.3.2.
  1. (3.3.2.a)

    If AA is supersingular and ΓA\Gamma_{A} is torsion free, then AA is neat. Indeed, in this case we have that RA={1}R_{A}^{\prime}=\mathopen{}\mathclose{{}\left\{1}\right\} and condition (Nb) is trivially satisfied.

  2. (3.3.2.b)

    Suppose that the Frobenius eigenvalues of AA are distinct and not supersingular. Some base extension of AA is neat if and only if AA has maximal angle rank.

  3. (3.3.2.c)

    In general, maximal angle rank always implies neatness.

3.4. Supersingular Serre–Frobenius groups

Recall that a qq-Weil number α\alpha is called supersingular if α/q\alpha/\sqrt{q} is a root of unity. In [41, Proposition 3.1], Zhu classified the minimal polynomials h(T)h(T) of supersingular qq-Weil numbers. Let Φr(T)\Phi_{r}(T) denote the rr-th cyclotomic polynomial, φ(r)\colonequalsdegΦr(T)\varphi(r)\colonequals\deg\Phi_{r}(T) the Euler totient function, and (ab)\mathopen{}\mathclose{{}\left(\tfrac{a}{b}}\right) the Jacobi symbol. Then the possibilities for the minimal polynomials of supersingular qq-Weil numbers are given in Table 2.

Table 1. Minimal polynomial of a supersingular qq-Weil number α\alpha.
Type dd h(T)h(T) Roots
Z-1 Even - Φm[q](T)\colonequalsqφ(m)Φm(T/q)\Phi_{m}^{[\sqrt{q}]}(T)\colonequals\sqrt{q}^{\varphi(m)}\Phi_{m}(T/\sqrt{q}) ζmjq for j(𝐙/m𝐙)×\zeta_{m}^{j}\sqrt{q}\text{ for }j\in(\mathbf{Z}/m\mathbf{Z})^{\times}
Z-2 Odd 𝐐(α)𝐐(α2)\mathbf{Q}(\alpha)\neq\mathbf{Q}(\alpha^{2}) Φn[q](T2)\colonequalsqφ(n)Φn(T2/q)\Phi_{n}^{[q]}(T^{2})\colonequals q^{\varphi(n)}\Phi_{n}(T^{2}/q) ±ζ2njq for j(𝐙/n𝐙)×\pm\zeta_{2n}^{j}\sqrt{q}\text{ for }j\in(\mathbf{Z}/n\mathbf{Z})^{\times}
Z-3 Odd 𝐐(α)=𝐐(α2)\mathbf{Q}(\alpha)=\mathbf{Q}(\alpha^{2}) 1jngcd(j,n)=1(T(qj)ζmνjq)\displaystyle\prod_{\begin{subarray}{c}1\leq j\leq n\\ \gcd(j,n)=1\end{subarray}}\mathopen{}\mathclose{{}\left(T-\mathopen{}\mathclose{{}\left(\dfrac{q}{j}}\right)\zeta_{m}^{\nu j}\sqrt{q}}\right) (qj)ζmjq for j(𝐙/n𝐙)×\mathopen{}\mathclose{{}\left(\dfrac{q}{j}}\right){\zeta_{m}^{j}}\sqrt{q}\text{ for }j\in(\mathbf{Z}/n\mathbf{Z})^{\times}
Notation (Table 2).

In case (Z-1), mm is any positive integer. In cases (Z-2) and (Z-3), mm additionally satisfies m2 mod 4m\not\equiv 2\text{ mod }4, and n\colonequalsm/gcd(2,m)n\colonequals m/\gcd(2,m). The symbol ζm\zeta_{m} denotes the a primitive mm-th root of unity. Note that in this case, φ(n)=φ(m)/gcd(2,m)\varphi(n)=\varphi(m)/\gcd(2,m). Following the notation in [29], given a polynomial f(T)K[T]f(T)\in K[T] for some field KK, and a constant aK×a\in K^{\times}, let

f[a](T)\colonequalsadegff(T/a).f^{[a]}(T)\colonequals a^{\deg f}f(T/a).

Given any supersingular abelian variety AA defined over 𝐅q\mathbf{F}_{q}, the Frobenius polynomial PA(T)P_{A}(T) is a power of the minimal polynomial hA(T)h_{A}(T), and this minimal polynomial is of type (Z-1), (Z-2), or (Z-3) as above. We say that AA is of type Z-i if the minimal polynomial hA(T)h_{A}(T) is of type (Z-i) for i=1,2,3i=1,2,3.

Since UAU_{A} is finite in the supersingular case, we have that 𝖲𝖥(A)UACmA\mathsf{SF}(A)\cong U_{A}\cong C_{m_{A}}. Furthermore, we have that

𝖲𝖥(A)={(ξ1ν,ξ2ν,,ξgν):ν𝐙/mA𝐙}𝖴(1)g,\mathsf{SF}(A)=\mathopen{}\mathclose{{}\left\{(\xi_{1}^{\nu},\xi_{2}^{\nu},\dots,\xi_{g}^{\nu}):\nu\in\mathbf{Z}/m_{A}\mathbf{Z}}\right\}\subset\mathsf{U}(1)^{g},

with ξi\xi_{i}’s being mAm_{A}-th roots of unity, whose orders have least common multiple mAm_{A}. In particular, we can read off the character group UAU_{A} from the fourth column in Table 2. For instance, if m=3m=3 and dd is even, then we have a polynomial of type Z-1, and the Serre–Frobenius group is isomorphic to C3C_{3}. On the other hand, if m=3m=3 and we have a polynomial of type Z-2, then the Serre–Frobenius group is isomorphic to C6C_{6}. Given a qq-Weil polynomial f(T)𝐐[T]f(T)\in\mathbf{Q}[T] with roots α1,,α2n\alpha_{1},\cdots,\alpha_{2n}, the associated normalized polynomial f~(T)𝐑[T]\tilde{f}(T)\in\mathbf{R}[T] is the monic polynomial with roots u1=α1/q,,u2n=α2n/qu_{1}=\alpha_{1}/\sqrt{q},\dots,u_{2n}=\alpha_{2n}/\sqrt{q}. Table 2 allows us to go back and forth between qq-Weil polynomials f(T)f(T) and the normalized polynomials f~(T)\tilde{f}(T).

  • If h(T)h(T) is the minimal polynomial of a supersingular qq-Weil number of type Z-1, the normalized polynomial h~(T)\tilde{h}(T) is the cyclotomic polynomial Φm(T)\Phi_{m}(T). Conversely, we have that h(T)=h~[q](T)h(T)=\tilde{h}^{[\sqrt{q}]}(T).

  • If h(T)h(T) is the minimal polynomial of a supersingular qq-Weil number of type Z-2, the normalized polynomial h~(T)\tilde{h}(T) is the polynomial Φn(T2)\Phi_{n}(T^{2}). Conversely, h(T)=h~[q](T)h(T)=\tilde{h}^{[q]}(T).

4. Elliptic Curves

The goal of this section is to prove Theorem 1.0.3. Furthermore, we give a thorough description of the set of possible orders mm for the supersingular Serre–Frobenius groups 𝖲𝖥(E)=Cm\mathsf{SF}(E)=C_{m} in terms of pp and q=pdq=p^{d}.

The isogeny classes of elliptic curves over 𝐅q\mathbf{F}_{q} were classified by Deuring [5] and Waterhouse [33, Theorem 4.1]. Writing the characteristic polynomial of Frobenius as P(T)=T2+a1T+qP(T)=T^{2}+a_{1}T+q, the Weil bounds give |a1|2q|a_{1}|\leq 2\sqrt{q}. Conversely, the integers aa satisfying |a|2q|a|\leq 2\sqrt{q} that correspond to the isogeny class of an elliptic curve are the following.

Theorem 4.0.1 ([28, Theorem 2.6.1]).

Let pp be a prime and q=pdq=p^{d}. Let a𝐙a\in\mathbf{Z} satisfy |a|2q|a|\leq 2\sqrt{q}.

  1. (1)

    If pap\nmid a, then aa is the trace of Frobenius of an elliptic curve over 𝐅q\mathbf{F}_{q}. This is the ordinary case.

  2. (2)

    If pap\mid a, then aa is the trace of Frobenius of an elliptic curve over 𝐅q\mathbf{F}_{q} if and only if one of the following holds:

    1. (i)

      dd is even and a=±2qa=\pm 2\sqrt{q},

    2. (ii)

      dd is even and a=qa=\sqrt{q} with p1 mod 3p\not\equiv 1\text{ mod }3,

    3. (iii)

      dd is even and a=qa=-\sqrt{q} with p1 mod 3p\not\equiv 1\text{ mod }3,

    4. (iv)

      dd is even and a=0a=0 with p1 mod 4p\not\equiv 1\text{ mod }4,

    5. (v)

      dd is odd and a=0a=0,

    6. (vi)

      dd is odd, a=±2qa=\pm\sqrt{2q} with p=2p=2.

    7. (vii)

      dd is odd, a=±3qa=\pm\sqrt{3q} with p=3p=3.

    This is the supersingular case.

In the ordinary case, the normalized Frobenius eigenvalue u1u_{1} is not a root of unity, and thus 𝖲𝖥(E)=𝖴(1)\mathsf{SF}(E)=\mathsf{U}(1). In the supersingular case, the normalized Frobenius eigenvalue u1u_{1} is a root of unity, and thus 𝖲𝖥(E)=Cm\mathsf{SF}(E)=C_{m} is cyclic, with mm equal to the order of u1u_{1}. For each value of qq and aa in Theorem 4.0.1 part (2), we get a right triangle of hypotenuse of length q\sqrt{q} and base a/2a/2, from which we can deduce the angle ϑ1\vartheta_{1} and thus the order mm of the corresponding root of unity u1u_{1}. We thus obtain Theorem 1.0.3 as a restatement of Theorem 4.0.1.

There are eight Serre–Frobenius groups for elliptic curves, sumarized in Table 3, and they correspond to eight possible Frobenius distributions of elliptic curves over finite fields. For ordinary elliptic curves (as explained in Section 1), the sequence of normalized traces (xr)r=1(x_{r})_{r=1}^{\infty} is equidistributed in the interval [2,2][-2,2] with respect to the measure λ1(x)\lambda_{1}(x) (Equation 1) obtained as the pushforward of the Haar measure μ𝖴(1)\mu_{\mathsf{U}(1)} under zz+z¯z\mapsto z+\overline{z}. See Figure 2(a).

The remaining seven Serre–Frobenius groups are finite and cyclic; they correspond to supersingular elliptic curves. For a given Cm=ζm𝖴(1)C_{m}=\langle\zeta_{m}\rangle\subset\mathsf{U}(1), denote by δm\delta_{m} the measure obtained by pushforward along zz+z¯z\mapsto z+\overline{z} of the normalized counting measure,

μCm(f)\colonequalsfμCm\colonequals1mj=1mf(ζmj).\mu_{C_{m}}(f)\colonequals\int f\,\mu_{C_{m}}\colonequals\frac{1}{m}\sum_{j=1}^{m}f(\zeta_{m}^{j}).
Table 2. Serre–Frobenius groups of elliptic curves.
Theorem 4.0.1 pp dd aa 𝖲𝖥(E)\mathsf{SF}(E) Generator Example Figure 2
(1) - - pap\nmid a 𝖴(1)\mathsf{U}(1) u1u_{1} 1.2.ab 2(a)
2-2(i) - Even 2q2\sqrt{q} C1C_{1} 11 1.4.ae 2(b)
2-2(i) - Even 2q-2\sqrt{q} C2C_{2} -11 1.4.e 2(c)
2-2(iii) p1 mod 3p\not\equiv 1\text{ mod }3 Even q-\sqrt{q} C3C_{3} ζ3\zeta_{3} 1.4.c 2(d)
2-2(iv) p1 mod 4p\not\equiv 1\text{ mod }4 Even 0 C4C_{4} ζ4\zeta_{4} 1.4.a 2(e)
2-2(v) - Odd 0 C4C_{4} ζ4\zeta_{4} 1.2.a 2(e)
2-2(ii) p1 mod 3p\not\equiv 1\text{ mod }3 Even q\sqrt{q} C6C_{6} ζ6\zeta_{6} 1.4.ac 2(f)
2-2(vi) 2 Odd ±2q\pm\sqrt{2q} C8C_{8} ζ8\zeta_{8} 1.2.ac 2(g)
2-2(vii) 3 Odd ±3q\pm\sqrt{3q} C12C_{12} ζ12\zeta_{12} 1.3.ad 2(h)
Refer to caption
(a) 𝖲𝖥(E)=𝖴(1)\mathsf{SF}(E)=\mathsf{U}(1)
Refer to caption
(b) 𝖲𝖥(E)=C1\mathsf{SF}(E)=C_{1}
Refer to caption
(c) 𝖲𝖥(E)=C2\mathsf{SF}(E)=C_{2}
Refer to caption
(d) 𝖲𝖥(E)=C3\mathsf{SF}(E)=C_{3}
Refer to caption
(e) 𝖲𝖥(E)=C4\mathsf{SF}(E)=C_{4}
Refer to caption
(f) 𝖲𝖥(E)=C6\mathsf{SF}(E)=C_{6}
Refer to caption
(g) 𝖲𝖥(E)=C8\mathsf{SF}(E)=C_{8}
Refer to caption
(h) 𝖲𝖥(E)=C12\mathsf{SF}(E)=C_{12}
Figure 2. a1a_{1}-histograms of elliptic curves.

5. Abelian Surfaces

The goal of this section is to classify the possible Serre–Frobenius groups of abelian surfaces (Theorem 1.0.4). The proof is a careful case-by-case analysis, described by Flowchart 3.

Refer to caption
Figure 3. Theorem 1.0.4: Classification in dimension 2.

We separate our cases first according to pp-rank, and then according to simplicity. In the supersingular and almost ordinary cases this stratification is enough. In the ordinary case, we have to further consider the geometric isogeny type of the surface. When the angle rank is 22, the Serre–Frobenius group is the full torus 𝖴(1)2\mathsf{U}(1)^{2}. When the angle rank is 11, the Serre–Frobenius group is isomorphic to 𝖴(1)×Cm\mathsf{U}(1)\times C_{m} but there are two non-conjugate ways to embed 𝖲𝖥(S)\mathsf{SF}(S) into 𝖴(1)2\mathsf{U}(1)^{2}; these are determined by the topological generator of the group, which can be either (u1,ζmu1)(u_{1},\zeta_{m}u_{1}) or (u1,ζm)(u_{1},\zeta_{m}).

5.1. Simple ordinary surfaces

We restate a theorem of Howe and Zhu in our notation.

Theorem 5.1.1 ([14, Theorem 6]).

Suppose that P(T)=T4+a1T3+a2T2+qa1T+q2P(T)=T^{4}+a_{1}T^{3}+a_{2}T^{2}+qa_{1}T+q^{2} is the Frobenius polynomial of a simple ordinary abelian surface SS defined over 𝐅q\mathbf{F}_{q}. Then, exactly one of the following conditions holds.

  1. (a)

    SS is absolutely simple.

  2. (b)

    a1=0a_{1}=0 and SS splits over a quadratic extension.

  3. (c)

    a12=q+a2a_{1}^{2}=q+a_{2} and SS splits over a cubic extension.

  4. (d)

    a12=2a2a_{1}^{2}=2a_{2} and SS splits over a quartic extension.

  5. (e)

    a12=3a23qa_{1}^{2}=3a_{2}-3q and SS splits over a sextic extension.

Lemma 5.1.2 (Node S-A in Figure 3).

Let SS be a simple ordinary abelian surface over 𝐅q\mathbf{F}_{q}. Then, exactly one of the following conditions holds.

  1. (a)

    SS is absolutely simple and 𝖲𝖥(S)=𝖴(1)2\mathsf{SF}(S)=\mathsf{U}(1)^{2}.

  2. (b)

    SS splits over a quadratic extension and 𝖲𝖥(S)𝖴(1)×C2\mathsf{SF}(S)\cong\mathsf{U}(1)\times C_{2}.

  3. (c)

    SS splits over a cubic extension and 𝖲𝖥(S)𝖴(1)×C3\mathsf{SF}(S)\cong\mathsf{U}(1)\times C_{3}.

  4. (d)

    SS splits over a quartic extension and 𝖲𝖥(S)𝖴(1)×C4\mathsf{SF}(S)\cong\mathsf{U}(1)\times C_{4}.

  5. (e)

    SS splits over a sextic extension and 𝖲𝖥(S)𝖴(1)×C6\mathsf{SF}(S)\cong\mathsf{U}(1)\times C_{6}.

In cases (b)-(e), we have that 𝖲𝖥(A)={(u,ζmνu):u𝖴(1),ν(𝐙/m𝐙)}\mathsf{SF}(A)=\mathopen{}\mathclose{{}\left\{(u,\zeta_{m}^{\nu}u):u\in\mathsf{U}(1),\,\nu\in(\mathbf{Z}/m\mathbf{Z})}\right\}, for some primitive mm-th root of unity ζm\zeta_{m}.

Proof.
  1. (a)

    From [40, Theorem 1.1], we conclude that some finite base extension of an absolutely simple abelian surface is neat and therefore has maximal angle rank by Remark (3.3.2.c). Alternatively, this also follows from the proof of [1, Theorem 2] for Jacobians of genus 2 curves, which generalizes to any abelian surface. Theorem 1.0.2 then implies that 𝖲𝖥(S)=𝖴(1)2\mathsf{SF}(S)=\mathsf{U}(1)^{2}.

  2. (b,c,d,e)

    Denote by mm the splitting degree of SS. By Theorem 5.1.1 we know that m{2,3,4,6}m\in\mathopen{}\mathclose{{}\left\{2,3,4,6}\right\}. Let α{α1,α¯1,α2,α¯2}\alpha\in\mathopen{}\mathclose{{}\left\{\alpha_{1},\overline{\alpha}_{1},\alpha_{2},\overline{\alpha}_{2}}\right\} be a Frobenius eigenvalue of SS. From [14, Lemma 4] and since SS is ordinary, we have that [𝐐(α):𝐐(αm)]=[𝐐(αm):𝐐]=2[\mathbf{Q}(\alpha):\mathbf{Q}(\alpha^{m})]=[\mathbf{Q}(\alpha^{m}):\mathbf{Q}]=2. In particular, the minimal polynomial h(m)(T)h_{(m)}(T) of αm\alpha^{m} is quadratic, and P(m)(T)=h(m)(T)2P_{(m)}(T)=h_{(m)}(T)^{2}. This implies that {α1m,α¯1m}={α2m,α¯2m}\mathopen{}\mathclose{{}\left\{\alpha_{1}^{m},\overline{\alpha}_{1}^{m}}\right\}=\mathopen{}\mathclose{{}\left\{\alpha_{2}^{m},\overline{\alpha}_{2}^{m}}\right\}, so that (up to relabelling) there is a primitive444Note that ζm\zeta_{m} must be primitive, since otherwise, P(n)(T)P_{(n)}(T) would split for some nmn\leq m, contradicting the minimality of mm. mm-th root of unity ζm\zeta_{m} such that α2=ζmα1\alpha_{2}=\zeta_{m}\alpha_{1}. It follows that

    𝖲𝖥(S)=(u1,ζmu1)¯={(u,ζmνu):u𝖴(1),ν𝐙/m𝐙}𝖴(1)×Cm\mathsf{SF}(S)=\overline{\langle(u_{1},\zeta_{m}u_{1})\rangle}=\mathopen{}\mathclose{{}\left\{(u,\zeta_{m}^{\nu}u):u\in\mathsf{U}(1),\,\nu\in\mathbf{Z}/m\mathbf{Z}}\right\}\cong\mathsf{U}(1)\times C_{m}

    and 𝖲𝖥(S)\mathsf{SF}(S)^{\circ} embeds diagonally in 𝖴(1)2\mathsf{U}(1)^{2}.

Figure 4. a1a_{1}-histograms for simple ordinary abelian surfaces.
Notation.

In Table 4, the Splitting type title refers to that of the abelian variety SS, the \cong class title refers to the Serre–Frobenius group 𝖲𝖥(S)\mathsf{SF}(S), and the Generator title refers to the topological generator of 𝖲𝖥(S)\mathsf{SF}(S); which precisely and succinctly captures the data of the embedding of the Serre–Frobenius group into the ambient unitary symplectic group. We will follow the same conventions in the following tables.

Table 3. Serre–Frobenius groups of simple ordinary surfaces.
Splitting type \cong class Generator Example Figure 4
Absolutely simple 𝖴(1)2\mathsf{U}(1)^{2} (u1,u2)(u_{1},u_{2}) 2.2.ab_b 4(a)
S(2)E2S_{(2)}\sim E^{2} 𝖴(1)×C2\mathsf{U}(1)\times C_{2} (u1,u1)(u_{1},-u_{1}) 2.2.a_ad 4(b)
S(3)E2S_{(3)}\sim E^{2} 𝖴(1)×C3\mathsf{U}(1)\times C_{3} (u1,ζ3u1)(u_{1},\zeta_{3}u_{1}) 2.2.ab_ab 4(c)
S(4)E2S_{(4)}\sim E^{2} 𝖴(1)×C4\mathsf{U}(1)\times C_{4} (u1,ζ4u1)(u_{1},\zeta_{4}u_{1}) 2.3.ac_c 4(d)
S(6)E2S_{(6)}\sim E^{2} 𝖴(1)×C6\mathsf{U}(1)\times C_{6} (u1,ζ6u1)(u_{1},\zeta_{6}u_{1}) 2.2.ad_f 4(e)

5.2. Non-simple ordinary surfaces

Let SS be a non-simple ordinary abelian surface defined over 𝐅q\mathbf{F}_{q}. Then SS is isogenous to a product of two ordinary elliptic curves E1×E2E_{1}\times E_{2}. As depicted in Figure 3, we consider two cases:

  1. (S-B)

    E1E_{1} and E2E_{2} are not isogenous over 𝐅¯q\overline{\mathbf{F}}_{q}.

  2. (S-C)

    E1E_{1} and E2E_{2} become isogenous over some base extension 𝐅qm1𝐅q\mathbf{F}_{q^{m_{1}}}\supseteq\mathbf{F}_{q}, for m11m_{1}\geq 1.

The Serre-Frobenius groups corresponding to these isogeny decomposition types are sumarized in Table 5. The proof of the following lemma is a straightforward application of Lemma 3.1.1.

Lemma 5.2.1 (Node S-B in Figure 3).

Let SS be an abelian surface defined over 𝐅q\mathbf{F}_{q} such that SS is isogenous to E1×E2E_{1}\times E_{2}, for E1E_{1} and E2E_{2} geometrically non-isogenous ordinary elliptic curves. Then SS has maximal angle rank δ=2\delta=2 and 𝖲𝖥(S)=𝖴(1)2\mathsf{SF}(S)=\mathsf{U}(1)^{2}.

Lemma 5.2.2 (Node S-C in Figure 3).

Let SS be an abelian surface defined over 𝐅q\mathbf{F}_{q} such that SS is isogenous to E1×E2E_{1}\times E_{2}, for E1E_{1} and E2E_{2} geometrically isogenous ordinary elliptic curves. Then SS has angle rank δ=1\delta=1 and 𝖲𝖥(S)=𝖴(1)×Cm\mathsf{SF}(S)=\mathsf{U}(1)\times C_{m} for m{1,2,3,4,6}m\in\mathopen{}\mathclose{{}\left\{1,2,3,4,6}\right\}. Furthermore, mm is precisely the degree of the extension of 𝐅q\mathbf{F}_{q} over which E1E_{1} and E2E_{2} become isogenous.

Proof.

Let α1,α¯1\alpha_{1},\overline{\alpha}_{1} and α2,α¯2\alpha_{2},\overline{\alpha}_{2} denote the Frobenius eigenvalues of E1E_{1} and E2E_{2} respectively. Let m1m_{1} be the smallest positive integer such that (E1)(m1)(E2)(m1)(E_{1})_{(m_{1})}\sim(E_{2})_{(m_{1})}. From Lemma 3.1.2, we immediately have that 𝖲𝖥(S)𝖴(1)×Cm\mathsf{SF}(S)\cong\mathsf{U}(1)\times C_{m}, where m=m1m=m_{1}. In order to find the value of mm, observe that {α1m,α¯1m}={α2m,α¯2m}\mathopen{}\mathclose{{}\left\{\alpha_{1}^{m},\overline{\alpha}_{1}^{m}}\right\}=\mathopen{}\mathclose{{}\left\{\alpha_{2}^{m},\overline{\alpha}_{2}^{m}}\right\}, from which we may assume, possibly after relabelling, that α2=ζmα1\alpha_{2}=\zeta_{m}\alpha_{1} for some primitive mm-th root of unity ζm\zeta_{m}. Since the curves E1E_{1} and E2E_{2} are ordinary, the number fields 𝐐(α1)\mathbf{Q}(\alpha_{1}) and 𝐐(α2)\mathbf{Q}(\alpha_{2}) are imaginary quadratic and 𝐐(α1)=𝐐(α1m)=𝐐(α2m)=𝐐(α2)\mathbf{Q}(\alpha_{1})=\mathbf{Q}(\alpha_{1}^{m})=\mathbf{Q}(\alpha_{2}^{m})=\mathbf{Q}(\alpha_{2}). Hence, ζm𝐐(α1)\zeta_{m}\in\mathbf{Q}(\alpha_{1}) and thus φ(m)=[𝐐(ζm):𝐐]{1,2}\varphi(m)=[\mathbf{Q}(\zeta_{m}):\mathbf{Q}]\in\mathopen{}\mathclose{{}\left\{1,2}\right\}; therefore m{1,2,3,4,6}m\in\mathopen{}\mathclose{{}\left\{1,2,3,4,6}\right\}. Finally, we have by definition that 𝖲𝖥(S)={(u,ζmνu):u𝖴(1),ν𝐙/m𝐙}𝖴(1)×Cm\mathsf{SF}(S)=\mathopen{}\mathclose{{}\left\{(u,\zeta_{m}^{\nu}u):u\in\mathsf{U}(1),\,\nu\in\mathbf{Z}/m\mathbf{Z}}\right\}\cong\mathsf{U}(1)\times C_{m}. ∎

Figure 5. a1a_{1}-histograms of non-simple ordinary abelian surfaces.
Table 4. Serre–Frobenius groups of non-simple ordinary surfaces S=E1×E2S=E_{1}\times E_{2}.
Splitting type \cong class Generator Example Figure 5
(E1)𝐅¯p≁(E2)𝐅¯p(E_{1})_{\overline{\mathbf{F}}_{p}}\not\sim(E_{2})_{\overline{\mathbf{F}}_{p}} 𝖴(1)2\mathsf{U}(1)^{2} (u1,u2)(u_{1},u_{2}) 2.3.ad_i 5(a)
E1E2E_{1}\sim E_{2} 𝖴(1)\mathsf{U}(1) (u1,u1)(u_{1},u_{1}) 2.2.ac_f 5(b)
(E1)(2)(E2)(2)(E_{1})_{(2)}\sim(E_{2})_{(2)} 𝖴(1)×C2\mathsf{U}(1)\times C_{2} (u1,u1)(u_{1},-u_{1}) 2.2.a_d 5(c)
(E1)(3)(E2)(3)(E_{1})_{(3)}\sim(E_{2})_{(3)} 𝖴(1)×C3\mathsf{U}(1)\times C_{3} (u1,ζ3u1)(u_{1},\zeta_{3}u_{1}) 2.7.af_s 5(d)
(E1)(4)(E2)(4)(E_{1})_{(4)}\sim(E_{2})_{(4)} 𝖴(1)×C4\mathsf{U}(1)\times C_{4} (u1,ζ4u1)(u_{1},\zeta_{4}u_{1}) 2.5.ag_s 5(e)
(E1)(6)(E2)(6)(E_{1})_{(6)}\sim(E_{2})_{(6)} 𝖴(1)×C6\mathsf{U}(1)\times C_{6} (u1,ζ6u1)(u_{1},\zeta_{6}u_{1}) 2.7.aj_bi 5(f)

5.3. Simple almost ordinary surfaces

In [16] Lenstra and Zarhin carried out a careful study of the multiplicative relations of Frobenius eigenvalues of simple almost ordinary varieties (see section 2.1 for the definition), which was later generalized in [8]. In particular, they prove that even dimensional simple almost ordinary abelian varieties have maximal angle rank ([16, Theorem 5.8]). Since every abelian surface of pp-rank 11 is almost ordinary, their result allows us to deduce the following.

Lemma 5.3.1 (Node S-D in Figure 3).

Let SS be a simple and almost ordinary abelian surface defined over 𝐅q\mathbf{F}_{q}. Then SS has maximal angle rank δ=2\delta=2 and 𝖲𝖥(S)=𝖴(1)2\mathsf{SF}(S)=\mathsf{U}(1)^{2}.

5.4. Non-simple almost ordinary surfaces

If SS is almost ordinary and not simple, then SS is isogenous to the product of an ordinary elliptic curve E1E_{1} and a supersingular elliptic curve E2E_{2}. The corresponding Serre-Frobenius groups are sumarized in Table 6.

Lemma 5.4.1 (Node S-E in Figure 3).

Let SS be a non-simple almost ordinary abelian surface defined over 𝐅q\mathbf{F}_{q}. Then SS has angle rank δ=1\delta=1 and angle torsion order m{1,2,3,4,6,8,12}m\in\mathopen{}\mathclose{{}\left\{1,2,3,4,6,8,12}\right\}. Furthermore, 𝖲𝖥(S)={(u,ζmν):u𝖴(1),ν𝐙/m𝐙}𝖴(1)×Cm\mathsf{SF}(S)=\mathopen{}\mathclose{{}\left\{(u,\zeta_{m}^{\nu}):u\in\mathsf{U}(1),\,\nu\in\mathbf{Z}/m\mathbf{Z}}\right\}\cong\mathsf{U}(1)\times C_{m}.

Proof.

Let E1E_{1} be an ordinary elliptic curve and E2E_{2} a supersingular elliptic curve such that SE1×E2S\sim E_{1}\times E_{2}. By Lemma 3.1.3, 𝖲𝖥(S)=𝖲𝖥(E1)×𝖲𝖥(E2)𝖴(1)×Cm\mathsf{SF}(S)=\mathsf{SF}(E_{1})\times\mathsf{SF}(E_{2})\cong\mathsf{U}(1)\times C_{m} with mm in the list of possible orders of Serre–Frobenius groups of supersingular elliptic curves. ∎

Figure 6. a1a_{1}-distributions of non-simple almost ordinary abelian surfaces.
Table 5. Serre–Frobenius groups of non-simple almost ordinary surfaces.
\cong class Generator Example Figure 6
𝖴(1)\mathsf{U}(1) (u1,1)(u_{1},1) 2.4.ah_u 6(a)
𝖴(1)×C2\mathsf{U}(1)\times C_{2} (u1,1)(u_{1},-1) 2.4.b_ae 6(b)
𝖴(1)×C3\mathsf{U}(1)\times C_{3} (u1,ζ3)(u_{1},\zeta_{3}) 2.4.ab_c 6(c)
𝖴(1)×C4\mathsf{U}(1)\times C_{4} (u1,ζ4)(u_{1},\zeta_{4}) 2.4.ad_i 6(d)
𝖴(1)×C6\mathsf{U}(1)\times C_{6} (u1,ζ6)(u_{1},\zeta_{6}) 2.4.af_o 6(e)
𝖴(1)×C8\mathsf{U}(1)\times C_{8} (u1,ζ8)(u_{1},\zeta_{8}) 2.2.ad_g 6(f)
𝖴(1)×C12\mathsf{U}(1)\times C_{12} (u1,ζ12)(u_{1},\zeta_{12}) 2.3.af_m 6(g)

5.5. Simple supersingular surfaces

Since every supersingular abelian variety is geometrically isogenous to a power of an elliptic curve, the Serre–Frobenius group only depends on the splitting degree. We separate our analysis into the simple and non-simple cases.

The classification of Frobenius polynomials of supersingular abelian surfaces over finite fields was completed by Maisner and Nart [18, Theorem 2.9] building on work of Xing [36] and Rück [24]. Denoting by (a1,a2)(a_{1},a_{2}) the isogeny class of abelian surfaces over 𝐅q\mathbf{F}_{q} with Frobenius polynomial PS(T)=T4+a1T3+a2T2+qa1T+q2P_{S}(T)=T^{4}+a_{1}T^{3}+a_{2}T^{2}+qa_{1}T+q^{2}, the following lemma gives the classification of Serre–Frobenius groups of simple supersingular surfaces.

Lemma 5.5.1 (Node S-F in Figure 3).

Let SS be a simple supersingular abelian surface defined over 𝐅q\mathbf{F}_{q}. The Serre–Frobenius group of SS is classified according to Table 7.

Table 6. Serre–Frobenius groups of simple supersingular surfaces.
(a1,a2)(a_{1},a_{2}) pp dd ee Type h~(T)\tilde{h}(T) 𝖲𝖥(S)\mathsf{SF}(S) Example
(0,0)(0,0) 1 mod 8\not\equiv 1\text{ mod }8 even 11 Z-1 Φ8(T)\Phi_{8}(T) C8C_{8} 2.4.a_a
(0,0)(0,0) 2\neq 2 odd 11 Z-2 Φ8(T)\Phi_{8}(T) C8C_{8} 2.3.a_a
(0,q)(0,q) - odd 11 Z-2 Φ3(T2)\Phi_{3}(T^{2}) C6C_{6} 2.2.a_c
(0,q)(0,-q) 1 mod 12\not\equiv 1\text{ mod }12 even 11 Z-1 Φ12(T)\Phi_{12}(T) C12C_{12} 2.4.a_ae
(0,q)(0,-q) 3\neq 3 odd 11 Z-2 Φ6(T2)=Φ12(T)\Phi_{6}(T^{2})=\Phi_{12}(T) C12C_{12} 2.2.a_ac
(q,q)(\sqrt{q},q) 1 mod 5\not\equiv 1\text{ mod }5 even 11 Z-1 Φ5(T)\Phi_{5}(T) C5C_{5} 2.4.c_e
(q,q)(-\sqrt{q},q) 1 mod 5\not\equiv 1\text{ mod }5 even 11 Z-1 Φ10(T)=Φ5(T)\Phi_{10}(T)=\Phi_{5}(-T) C10C_{10} 2.4.ac_e
(5q,3q)(\sqrt{5q},3q) =5=5 odd 11 Z-3 Ψ5,1(T)\Psi_{5,1}(T) C10C_{10} 2.5.f_p
(5q,3q)(-\sqrt{5q},3q) =5=5 odd 11 Z-3 Ψ5,1(T)\Psi_{5,1}(-T) C10C_{10} 2.5.af_p
(2q,q)(\sqrt{2q},q) =2=2 odd 11 Z-3 Ψ2,3(T)\Psi_{2,3}(T) C24C_{24} 2.2.c_c
(2q,q)(-\sqrt{2q},q) =2=2 odd 11 Z-3 Ψ2,3(T)\Psi_{2,3}(-T) C24C_{24} 2.2.ac_c
(0,2q)(0,-2q) - odd 22 Z-2 Φ1(T2)\Phi_{1}(T^{2}) C2C_{2} 2.2.a_ae
(0,2q)(0,2q) 1 mod 4\equiv 1\text{ mod }4 even 22 Z-1 Φ4(T)\Phi_{4}(T) C4C_{4} 2.25.a_by
(2q,3q)(2\sqrt{q},3q) 1 mod 3\equiv 1\text{ mod }3 even 22 Z-1 Φ3(T)\Phi_{3}(T) C3C_{3} 2.49.o_fr
(2q,3q)(-2\sqrt{q},3q) 1 mod 3\equiv 1\text{ mod }3 even 22 Z-1 Φ6(T)=Φ3(T)\Phi_{6}(T)=\Phi_{3}(-T) C6C_{6} 2.49.ao_fr

The notation for polynomials of type Z-3 is taken from [29], where the authors classify simple supersingular Frobenius polynomials for g7g\leq 7. We have

(11) Ψ5,1(T)\displaystyle\Psi_{5,1}(T) \colonequalsa(𝐙/5)×(T(a5)ζ5a)=T4+5T3+3T2+5T+1,\displaystyle\colonequals\prod_{a\in(\mathbf{Z}/5)^{\times}}\mathopen{}\mathclose{{}\left(T-\mathopen{}\mathclose{{}\left(\tfrac{a}{5}}\right)\zeta_{5}^{a}}\right)=T^{4}+\sqrt{5}T^{3}+3T^{2}+\sqrt{5}T+1,
(12) Ψ2,3(T)\displaystyle\Psi_{2,3}(T) \colonequalsa(𝐙/3)×(Tζ8ζ3a)(Tζ¯8ζ3a)=T4+2T3+T2+2T+1.\displaystyle\colonequals\prod_{a\in(\mathbf{Z}/3)^{\times}}\mathopen{}\mathclose{{}\left(T-\zeta_{8}\zeta_{3}^{a}}\right)\mathopen{}\mathclose{{}\left(T-\overline{\zeta}_{8}\zeta_{3}^{a}}\right)=T^{4}+\sqrt{2}T^{3}+T^{2}+\sqrt{2}T+1.

We exhibit the proof of the second line in Table 7 for exposition. The remaining cases can be checked similarly. If (a1,a2)=(0,0)(a_{1},a_{2})=(0,0), p2p\neq 2 and qq is an odd power of pp: then, P(T)=T4+q2=q4Φ8(T/q)=q2Φ4(T2/q)P(T)=T^{4}+q^{2}=\sqrt{q}^{4}\Phi_{8}(T/\sqrt{q})=q^{2}\Phi_{4}(T^{2}/q) and h~(T)=Φ8(T)\tilde{h}(T)=\Phi_{8}(T). Thus USU_{S} is generated by a primitive 8th root of unity.

5.6. Non-simple supersingular surfaces

If SS is a non-simple supersingular surface, then SS is isogenous to a product of two supersingular elliptic curves E1E_{1} and E2E_{2}. If mE1m_{E_{1}} and mE2m_{E_{2}} denote the torsion orders of E1E_{1} and E2E_{2} respectively, then the extension over which E1E_{1} and E2E_{2} become isogenous is precisely lcm(mE1,mE2)\operatorname{lcm}(m_{E_{1}},m_{E_{2}}). Thus, we have the following result, depending on the values of q=pdq=p^{d} as in Table 3.

Lemma 5.6.1 (Node S-G in Figure 3).

Let SS be a non-simple supersingular abelian surface defined over 𝐅q\mathbf{F}_{q}. Then SS has angle rank δ=0\delta=0 and 𝖲𝖥(S)=Cm\mathsf{SF}(S)=C_{m} for mm in the set M=M(p,d)M=M(p,d) described in Table 8.

Table 7. Angle torsion set for non-simple supersingular surfaces defined over 𝐅q\mathbf{F}_{q}, with q=pdq=p^{d}.
dd pp M(p,d)M(p,d)
Even - {1,2}\{1,2\}
Even p1 mod 3p\not\equiv 1\text{ mod }3 {1,2,3,6}\{1,2,3,6\}
Even p1 mod 4p\not\equiv 1\text{ mod }4 {1,2,4}\{1,2,4\}
Odd - {4}\{4\}
Odd p=2p=2 {4,8}\{4,8\}
Odd p=3p=3 {4,12}\{4,12\}

6. Abelian Threefolds

In this section, we classify the Serre–Frobenius groups of abelian threefolds (see Figure 7). Let XX be an abelian variety of dimension 33 defined over 𝐅q\mathbf{F}_{q}. For our analysis, we will first stratify the cases by pp-rank and then by simplicity. Before we proceed, we make some observations about simple threefolds that will be useful later.

Refer to caption
Figure 7. Theorem 1.0.5: Classification in dimension 3.

6.1. Simple abelian threefolds

If XX is a simple abelian threefold, there are only two possibilities for the Frobenius polynomial PX(T)=hX(T)eP_{X}(T)=h_{X}(T)^{e}:

(13) PX(T)=\displaystyle P_{X}(T)= hX(T)\displaystyle\,h_{X}(T)
(14) PX(T)=\displaystyle P_{X}(T)= hX(T)3.\displaystyle\,h_{X}(T)^{3}.

Indeed, if hX(T)h_{X}(T) were a linear or cubic polynomial, it would have a real root ±q\pm\sqrt{q}. By an argument of Waterhouse [33, Chapter 2], the qq-Weil numbers ±q\pm\sqrt{q} must come from simple abelian varieties of dimension 1 or 2. Further, Xing [35] showed that Equation 14 can only occur in very special cases.

Theorem 6.1.1 ([35], [12, Proposition 1.2]).

Let XX be a simple abelian threefold over 𝐅q\mathbf{F}_{q}. Then, PX(T)=hX(T)3P_{X}(T)=h_{X}(T)^{3} if and only if 33 divides d=logp(q)d=\log_{p}(q) and hX(T)=T2+aq1/3T+qh_{X}(T)=T^{2}+aq^{1/3}T+q with gcd(a,p)=1\gcd(a,p)=1.

When PX(T)P_{X}(T) is a cube as above, since gcd(a,p)=1\gcd(a,p)=1, the qq-adic valuation of its middle coefficient is the same as that of aqaq, which in turn is 11. Thus, XX is non-supersingular of pp-rank 0 and its Newton Polygon has slopes (13,13,13,23,23,23)(\tfrac{1}{3},\tfrac{1}{3},\tfrac{1}{3},\tfrac{2}{3},\tfrac{2}{3},\tfrac{2}{3}). Furthermore, every simple abelian threefold is either absolutely simple or geometrically isogenous to the cube of an elliptic curve. Thus, we have the following.

Lemma 6.1.2.

If XX is an abelian threefold defined over 𝐅q\mathbf{F}_{q} that is not ordinary or supersingular, then XX is simple if and only if it is absolutely simple.

Proof.

Assume XX is a simple abelian threefold that is not ordinary or supersingular. Assume also that XX is not absolutely simple. Let r>1r>1 be the splitting degree of XX. Recall that since XX is simple, one either has PX(T)=hX(T)P_{X}(T)=h_{X}(T) or PX(T)=hX(T)3P_{X}(T)=h_{X}(T)^{3}, where hX(T)h_{X}(T) is irreducible of even degree. We will show that in each case X(r)E3X_{(r)}\sim E^{3}, contradicting the assumption that XX is not ordinary or supersingular.

Assume PX(T)=hX(T)3P_{X}(T)=h_{X}(T)^{3}, then PX(r)(T)=hX,(r)(T)3P_{X_{(r)}}(T)=h_{X,(r)}(T)^{3}. Observe that necessarily X(r)X_{(r)} has an elliptic curve E/𝐅qrE/\mathbf{F}_{q^{r}} as an isogeny factor. Then PE(T)P_{E}(T), a quadratic polynomial, must divide PX(r)(T)P_{X_{(r)}}(T), and we conclude PE(T)=hX,(r)(T)P_{E}(T)=h_{X,(r)}(T). Thus X(r)E3X_{(r)}\sim E^{3}.

Assume instead that PX(T)=hX(T)P_{X}(T)=h_{X}(T), that is, PX(T)P_{X}(T) is irreducible. Therefore 𝐐(πX)\mathbf{Q}(\pi_{X}) is a degree 66 extension and 𝐐(πX)End0(X)End0(X(r))\mathbf{Q}(\pi_{X})\hookrightarrow\mathrm{End}^{0}(X)\hookrightarrow\mathrm{End}^{0}(X_{(r)}). Then by [3, Theorem 1.3.1.1] X(r)X_{(r)} is isotypic, so that X(r)E3X_{(r)}\sim E^{3}. ∎

In each case of the classification that follows, we will denote by MM, the set of possible angle torsion orders that occur for that case. When we want to emphasize the dependence on the prime pp and the power dd, we will denote this by M(p,d)M(p,d).

6.2. Simple ordinary threefolds

In this section, XX will denote a simple ordinary abelian threefold defined over 𝐅q\mathbf{F}_{q}. The corresponding Serre-Frobenius groups are sumarized in Table 9.

As a corollary to Theorem 3.2.1, we have the following.

Proposition 6.2.1.

Let XX be a simple ordinary abelian threefold defined over 𝐅q\mathbf{F}_{q}. Then, exactly one of the following conditions is satisfied.

  1. (1)

    XX is absolutely simple.

  2. (2)

    XX splits over a degree 33 extension and PX(T)=T6+a3T3+q3P_{X}(T)=T^{6}+a_{3}T^{3}+q^{3}.

  3. (3)

    XX splits over a degree 77 extension and the number field of PX(T)P_{X}(T) is 𝐐(ζ7)\mathbf{Q}(\zeta_{7}).

Lemma 6.2.2 (Node X-A in Figure 7).

Let XX be an absolutely simple abelian threefold defined over 𝐅q\mathbf{F}_{q}. Then XX has maximal angle rank δ=3\delta=3 and 𝖲𝖥(X)=𝖴(1)3\mathsf{SF}(X)=\mathsf{U}(1)^{3}.

Proof.

Let m=mXm=m_{X} be the order of the torsion subgroup of ΓX\Gamma_{X}. By [40, Theorem 1.1], we have that X(m)X_{(m)} is neat. Since X(m)X_{(m)} is ordinary and simple, its Frobenius eigenvalues are distinct and non-real. Remark (3.3.2.b) implies that X(m)X_{(m)} has maximal angle rank. Since angle rank is invariant under base extension (Remark 2.2.3) we have that δ(X)=δ(X(m))=3\delta(X)=\delta(X_{(m)})=3 as we wanted to show. ∎

Lemma 6.2.3 (Node X-B in Figure 7).

Let XX be a simple ordinary abelian threefold over 𝐅q\mathbf{F}_{q} that is not absolutely simple. Then XX has angle rank 11 and

  1. (a)

    𝖲𝖥(X)𝖴(1)×C3\mathsf{SF}(X)\cong\mathsf{U}(1)\times C_{3} if XX splits over a degree 33 extension, or

  2. (b)

    𝖲𝖥(X)𝖴(1)×C7\mathsf{SF}(X)\cong\mathsf{U}(1)\times C_{7} if XX splits over a degree 77 extension.

Furthermore, in (a) and (b), we have that

𝖲𝖥(X)={(u,ξ1νu,ξ2νu):u𝖴(1),ν𝐙/m𝐙},\mathsf{SF}(X)=\mathopen{}\mathclose{{}\left\{(u,\xi_{1}^{\nu}u,\xi_{2}^{\nu}u):u\in\mathsf{U}(1),\nu\in\mathbf{Z}/m\mathbf{Z}}\right\},

with ξ1,ξ2\xi_{1},\xi_{2} distinct primitive mm-th roots of unity, for m=3m=3 and m=7m=7 respectively.

Proof.

From the proof of Theorem 3.2.1, we have that the torsion free part of UXU_{X} is generated by a fixed normalized root u1=α1/qu_{1}=\alpha_{1}/\sqrt{q}, and all other roots uju_{j} for 1<jg1<j\leq g are related to u1u_{1} by a primitive root of unity of order 33 or 77 respectively; u2=ξ1u1u_{2}=\xi_{1}u_{1} and u3=ξ2u1u_{3}=\xi_{2}u_{1} with ξ1ξ2\xi_{1}\neq\xi_{2}. ∎

Table 8. Serre–Frobenius groups of simple ordinary threefolds XX.
Splitting type \cong class Generator Example Figure 8
Absolutely simple 𝖴(1)3\mathsf{U}(1)^{3} (u1,u2,u3)(u_{1},u_{2},u_{3}) 3.2.ad_f_ah 8(a)
X(3)E3X_{(3)}\sim E^{3} 𝖴(1)×C3\mathsf{U}(1)\times C_{3} (u1,ζ3u1,ζ3νu1)(u_{1},\zeta_{3}u_{1},\zeta_{3}^{\nu}u_{1}) 3.2.a_a_ad 8(b)
X(7)E3X_{(7)}\sim E^{3} 𝖴(1)×C7\mathsf{U}(1)\times C_{7} (u1,ζ7u1,ζ7νu1)(u_{1},\zeta_{7}u_{1},\zeta_{7}^{\nu}u_{1}) 3.2.ae_j_ap 8(c)
Figure 8. a1a_{1}-distributions for simple ordinary threefolds.

6.3. Non-simple ordinary threefolds

Let XX be a non-simple ordinary abelian threefold defined over 𝐅q\mathbf{F}_{q}. Then XX is isogenous to a product S×ES\times E, for some ordinary surface SS and some ordinary elliptic curve EE.

The Frobenius polynomial of XX is the product of those of SS and EE. Further, exactly one of the following is true for SS: either it is absolutely simple, or it is simple and geometrically isogenous to the power of a single elliptic curve, or it is not simple (see observation after Lemma 5.1.2). The Serre–Frobenius group of XX depends on its geometric isogeny decomposition, of which there are the following four possibilities.

  1. (6.3-a)

    XX is geometrically isogenous to E3E^{3}.

  2. (6.3-b)

    XX is geometrically isogenous to E12×EE_{1}^{2}\times E, for some ordinary elliptic curve E1E_{1}, with (E1)𝐅¯q≁(E)𝐅¯q(E_{1})_{\overline{\mathbf{F}}_{q}}\not\sim(E)_{\overline{\mathbf{F}}_{q}}.

  3. (6.3-c)

    XX is geometrically isogenous to E1×E2×EE_{1}\times E_{2}\times E, for ordinary and pairwise geometrically non-isogenous elliptic curves E1,E2E_{1},E_{2} and EE.

  4. (6.3-d)

    XX is geometrically isogenous to S×ES\times E for an absolutely simple ordinary surface SS and an ordinary elliptic curve EE.

Lemma 6.3.1 (Node X-C in Figure 7).

Let XX be a non-simple ordinary abelian threefold over 𝐅q\mathbf{F}_{q}. The possible Serre–Frobenius groups of XX are given in Table 10.

Table 9. Serre–Frobenius groups of non-simple ordinary threefolds X=S×EX=S\times E.
Splitting type \cong class 𝖲𝖥(X)\mathsf{SF}(X) Generator mMm\in M Examples
(6.3-a) 𝖴(1)×Cm\mathsf{U}(1)\times C_{m} (u1,ζmu1,u1)(u_{1},\zeta_{m}u_{1},u_{1}) {1,2,3,4,6}\mathopen{}\mathclose{{}\left\{1,2,3,4,6}\right\} Example 6.3.2
(6.3-b) 𝖴(1)2×Cm\mathsf{U}(1)^{2}\times C_{m} (u1,ζmu1,u2)(u_{1},\zeta_{m}u_{1},u_{2}) {1,2,3,4,6}\mathopen{}\mathclose{{}\left\{1,2,3,4,6}\right\} Example 6.3.3
(6.3-c) 𝖴(1)3\mathsf{U}(1)^{3} (u1,u2,u3)(u_{1},u_{2},u_{3}) {1}\mathopen{}\mathclose{{}\left\{1}\right\} 3.5.ai_bi_ado
(6.3-d) 𝖴(1)3\mathsf{U}(1)^{3} (u1,u2,u3)(u_{1},u_{2},u_{3}) {1}\mathopen{}\mathclose{{}\left\{1}\right\} 3.2.ad_h_al
Proof.

Recall that XS×EX\sim S\times E over 𝐅q\mathbf{F}_{q}.
(6.3-a) If XX is geometrically isogenous to E3E^{3}, then SS is geometrically isogenous to E2E^{2}. By Lemma 3.1.4 𝖲𝖥(X)𝖴(1)×Cm\mathsf{SF}(X)\cong\mathsf{U}(1)\times C_{m}, where mm is the splitting degree of SS, and so S(m)E2S_{(m)}\sim E^{2}. By [14, Theorem 6], we have that m{1,2,3,4,6}m\in\{1,2,3,4,6\}.

(6.3-b) In this case, by Lemma 3.1.4, 𝖲𝖥(X)𝖴(1)2×Cm\mathsf{SF}(X)\cong\mathsf{U}(1)^{2}\times C_{m}, where mm is the splitting degree of SS and S(m)E12S_{(m)}\sim E_{1}^{2}. As in the previous case, m{1,2,3,4,6}m\in\{1,2,3,4,6\}.

(6.3-c) In this case SE1×E2S\sim E_{1}\times E_{2} over the base field. Lemma 3.1.1 implies δX=3\delta_{X}=3.

(6.3-d) In this case, XS×EX\sim S\times E with SS absolutely simple. By [40, Theorem 1.1], we know that XX is neat. Since XX is ordinary and SS is simple, all Frobenius eigenvalues are distinct and not supersingular. By Remark (3.3.2.b), we conclude that δX=3\delta_{X}=3. ∎

Example 6.3.2 (Non-simple ordinary threefolds of splitting type (6.3-a)).
Figure 9. a1a_{1}-distributions for non-simple ordinary abelian threefolds of splitting type (6.3-a).
Example 6.3.3 (Non-simple ordinary threefolds of splitting type (6.3-b)).
Figure 10. a1a_{1}-distributions for non-simple ordinary abelian threefolds of splitting type (6.3-b).

6.4. Simple almost ordinary threefolds

Let XX be a simple and almost ordinary abelian threefold over 𝐅q\mathbf{F}_{q}. Recall that XX is in fact absolutely simple, so that the Frobenius polynomial P(r)(T)P_{(r)}(T) is irreducible for every positive integer rr.

Lemma 6.4.1 (Node X-D in Figure 7).

Let XX be a simple almost ordinary abelian threefold over 𝐅q\mathbf{F}_{q}. The Serre–Frobenius group of XX can be read from Table 11.

Table 10. Serre–Frobenius groups of simple almost ordinary threefolds.
Def. 3.3.1 q𝐐(πX)\sqrt{q}\in\mathbf{Q}(\pi_{X}) \cong class Generator mMm\in M Example
Neat - 𝖴(1)3\mathsf{U}(1)^{3} (u1,u2,u3)(u_{1},u_{2},u_{3}) {\{1}\} 3.2.ab_ab_c
Not neat Yes 𝖴(1)2×Cm\mathsf{U}(1)^{2}\times C_{m} (u1,u2,ζmu1u2)(u_{1},u_{2},\zeta_{m}u_{1}u_{2}) {\{1, 2, 3, 4, 6}\} Example 6.4.2
Not neat No 𝖴(1)2×Cm\mathsf{U}(1)^{2}\times C_{m} (u1,u2,ζmu1u2)(u_{1},u_{2},\zeta_{m}u_{1}u_{2}) {\{4, 6, 8, 12}\} Example 6.4.2
Proof.

Let m\colonequalsmXm\colonequals m_{X} be the torsion order of UXU_{X}, and consider the base extension Y\colonequalsX(m)Y\colonequals X_{(m)}. By [16, Theorem 5.7], we know that δX=δY2\delta_{X}=\delta_{Y}\geq 2. Furthermore, since YY is absolutely simple, by the discussion in Section 6.1, the roots of PY(T)=P(m)(T)P_{Y}(T)=P_{(m)}(T) are distinct and non-supersingular. If YY is neat, Remark (3.3.2.b) implies that δX=δY=3\delta_{X}=\delta_{Y}=3. Assume then that YY is not neat, so that δX=2\delta_{X}=2. Let α=α1\alpha=\alpha_{1} be a Frobenius eigenvalue of XX. By [40, Theorem 1.1] and the discussion thereafter, we have that the sextic CM-field 𝐐(α)=𝐐(αm)\mathbf{Q}(\alpha)=\mathbf{Q}(\alpha^{m}) contains an imaginary quadratic field BB, and (u1u2u3)2m=Norm𝐐(α)/B(u12m)=1(u_{1}u_{2}u_{3})^{2m}=\mathrm{Norm}_{\mathbf{Q}(\alpha)/B}(u_{1}^{2m})=1. Further, Norm𝐐(α)/B(α1)=α1α2α3\mathrm{Norm}_{\mathbf{Q}(\alpha)/B}(\alpha_{1})=\alpha_{1}\alpha_{2}\alpha_{3}. Since UYU_{Y} has no torsion, this implies that (u1u2u3)m=1(u_{1}u_{2}u_{3})^{m}=1. Moreover, this means that u1u2u3=ζu_{1}u_{2}u_{3}=\zeta for some primitive555The primitivity of ζ\zeta follows from the fact that mm is the minimal positive integer such that UY=U(m)U_{Y}=U_{(m)} is torsion free. mm-th root of unity ζ\zeta. Therefore,

(15) ζ2=Norm𝐐(α)/B(u12)B.\zeta^{2}=\mathrm{Norm}_{\mathbf{Q}(\alpha)/B}(u_{1}^{2})\in B.

If mm is odd, ζ2\zeta^{2} is also primitive, so that φ(m)2\varphi(m)\leq 2 and m{1,3}m\in\mathopen{}\mathclose{{}\left\{1,3}\right\}. If mm is even, then we may distinguish between two cases. If q𝐐(α)\sqrt{q}\in\mathbf{Q}(\alpha), we know that u1𝐐(α)u_{1}\in\mathbf{Q}(\alpha) so that in fact ±ζ=Norm𝐐(α)/B(u1)B\pm\zeta=\mathrm{Norm}_{\mathbf{Q}(\alpha)/B}(u_{1})\in B and φ(m)2\varphi(m)\leq 2 implies that m{2,4,6}m\in\mathopen{}\mathclose{{}\left\{2,4,6}\right\}. If q𝐐(α)\sqrt{q}\not\in\mathbf{Q}(\alpha), then ζ2\zeta^{2} is a primitive m/2m/2-root of unity and m/2{1,2,3,4,6}m/2\in\mathopen{}\mathclose{{}\left\{1,2,3,4,6}\right\}.

In the setting where YY is not neat and q𝐐(α)\sqrt{q}\notin\mathbf{Q}(\alpha), we notice that u1u2u3=±1u_{1}u_{2}u_{3}=\pm 1 implies that q=±α1α2α3/q𝐐(α)\sqrt{q}=\pm\alpha_{1}\alpha_{2}\alpha_{3}/q\in\mathbf{Q}(\alpha), so the cases m=1,2m=1,2 don’t occur when q𝐐(α)\sqrt{q}\not\in\mathbf{Q}(\alpha). Similarly, if u1u2u3=ζ3u_{1}u_{2}u_{3}=\zeta_{3}, then q=(α1α2α3)3/q4𝐐(α)\sqrt{q}=(\alpha_{1}\alpha_{2}\alpha_{3})^{3}/q^{4}\in\mathbf{Q}(\alpha). Thus, the torsion orders m=1,2,3m=1,2,3 do not occur in this case. ∎

Example 6.4.2 (a1a_{1}-distributions of simple almost ordinary abelian threefolds with angle rank 22).

The histograms corresponding to the following examples are presented in Figure 11. In these examples we use SageMath [25] to initialize the degree-6 number field K=𝐐(α)K=\mathbf{Q}(\alpha) corresponding to the Frobenius polynomial, find the corresponding quadratic subfield BB, and check that Norm𝐐(α)/B(u1)\mathrm{Norm}_{\mathbf{Q}(\alpha)/B}(u_{1}) is the root of unity in question. The code for generating the histograms is available on the GitHub repository [2].

Figure 11. a1a_{1}-distributions of simple almost ordinary abelian threefolds of angle rank 22.

6.5. Non-simple almost ordinary threefolds

Since XX is not simple, we have that XS×EX\sim S\times E for some surface SS and some elliptic curve EE. For this section, we let π1,π¯1,π2,π¯2\pi_{1},\overline{\pi}_{1},\pi_{2},\overline{\pi}_{2} and α,α¯\alpha,\overline{\alpha} be the Frobenius eigenvalues of SS and EE respectively. The normalized eigenvalues will be denoted by u1\colonequalsπ1/q,u2=π2/qu_{1}\colonequals\pi_{1}/\sqrt{q},u_{2}=\pi_{2}/\sqrt{q} and u\colonequalsα/qu\colonequals\alpha/\sqrt{q}. If XX has a geometric supersingular factor, by Honda–Tate theory, it must have a supersingular factor over the base field; and without loss of generality we may assume that this factor is EE.

Refer to caption
Figure 12. Serre–Frobenius groups of non-simple almost ordinary threefolds.
Lemma 6.5.1 (Node X-E in Figure 7).

Let XS×EX\sim S\times E be a non-simple almost ordinary abelian threefold over 𝐅q\mathbf{F}_{q}. The Serre–Frobenius group of XX can be read from Flowchart 12. In particular, if XX has no supersingular factor, then δX=3\delta_{X}=3. If EE is supersingular, then δX{1,2}\delta_{X}\in\mathopen{}\mathclose{{}\left\{1,2}\right\} and mX=lcm(mS,mE)m_{X}=\operatorname{lcm}(m_{S},m_{E}). The list of possible torsion orders mXm_{X} in this case is given Table 12.

Proof.

First, suppose that XX has no supersingular factor. Thus EE is ordinary and SS is almost ordinary and absolutely simple. This implies that 𝐐(π1r)\mathbf{Q}(\pi_{1}^{r}) and 𝐐(αr)\mathbf{Q}(\alpha^{r}) are CM-fields of degrees 44 and 22 respectively, for every positive integer rr. In particular, #{π1r,π¯1r,π2r,π¯2r,αr,α¯r}\#\mathopen{}\mathclose{{}\left\{\pi_{1}^{r},\overline{\pi}_{1}^{r},\pi_{2}^{r},\overline{\pi}_{2}^{r},\alpha^{r},\overline{\alpha}^{r}}\right\} =6=6 for every rr. Let m=mXm=m_{X} and consider the base extension X(m)X_{(m)}. Since X(m)X_{(m)} is not simple, [40, Theorem 1.1] implies that X(m)X_{(m)} is neat. The eigenvalues of X(m)X_{(m)} are all distinct and not supersingular, so that δ(X)=δ(X(m))=3\delta(X)=\delta(X_{(m)})=3 by Remark (3.3.2.b). The case where XX has a supersingular factor follows from Lemma 3.1.3. ∎

Table 11. Serre–Frobenius groups of non-simple almost ordinary threefolds X=S×EX=S\times E.
δE\delta_{E} \cong class Generator d=logp(q)d=\log_{p}(q) mM(p,d)m\in M(p,d) Example
11 𝖴(1)3\mathsf{U}(1)^{3} (u1,u2,u3)(u_{1},u_{2},u_{3}) - {1}\mathopen{}\mathclose{{}\left\{1}\right\} 3.3.ac_d_ae
0 𝖴(1)2×Cm\mathsf{U}(1)^{2}\times C_{m} (u1,u2,ζmE)(u_{1},u_{2},\zeta_{m_{E}}) - m=mE{1,2,3,4,6,8,12}m=m_{E}\in\{1,2,3,4,6,8,12\} Figure 13
0 𝖴(1)×Cm\mathsf{U}(1)\times C_{m} (u1,ζmSu1,ζmE)(u_{1},\zeta_{m_{S}}u_{1},\zeta_{m_{E}}) even m=lcm(mS,mE){1,2,3,4,6,12}m=\operatorname{lcm}(m_{S},m_{E})\in\mathopen{}\mathclose{{}\left\{1,2,3,4,6,12}\right\} Figure 14
0 𝖴(1)×Cm\mathsf{U}(1)\times C_{m} (u1,ζmSu1,ζmE)(u_{1},\zeta_{m_{S}}u_{1},\zeta_{m_{E}}) odd m=lcm(mS,mE){4,8,12,24}m=\operatorname{lcm}(m_{S},m_{E})\in\mathopen{}\mathclose{{}\left\{4,8,12,24}\right\} Figure 14
Figure 13. a1a_{1}-distributions of non-simple almost ordinary abelian threefolds of angle rank 22.
Figure 14. a1a_{1}-distributions of non-simple almost ordinary abelian threefolds of angle rank 11.

6.6. Abelian threefolds of K3-type

In this section XX will be an abelian threefold defined over 𝐅q\mathbf{F}_{q} of pp-rank 11. The qq-Newton polygon of such a variety has slopes (0,12,12,12,12,1)(0,\tfrac{1}{2},\tfrac{1}{2},\tfrac{1}{2},\tfrac{1}{2},1). This is the three-dimensional instance of abelian varieties of K3 type, which were studied by Zarhin in [38] and [37].

Definition 6.6.1.

An abelian variety AA defined over 𝐅q\mathbf{F}_{q} is said to be of K3-type if the set of Newton slopes is either {0,1}\mathopen{}\mathclose{{}\left\{0,1}\right\} or {0,1/2,1}\mathopen{}\mathclose{{}\left\{0,1/2,1}\right\}, and the segments of slope 0 and 11 have length one.

By [37, Theorem 5.9], simple abelian varieties of K3-type have maximal angle rank. As a corollary, we have another piece of the classification.

Lemma 6.6.2 (Node X-F in Figure 7).

Let XX be a simple abelian threefold over 𝐅q\mathbf{F}_{q} of pp-rank 11. Then XX has maximal angle rank and 𝖲𝖥(X)𝖴(1)3\mathsf{SF}(X)\cong\mathsf{U}(1)^{3}.

There are several examples of such XX, one of them being 3.2.ab_a_a. Now assume that XX is not simple, so that XS×EX\sim S\times E for some surface SS and elliptic curve EE.

Lemma 6.6.3 (Node X-G in Figure 7).

Let XS×EX\sim S\times E be a non-simple abelian threefold over 𝐅q\mathbf{F}_{q} of pp-rank 11. The Serre–Frobenius group of XX is given by Table 13.

We consider three cases:

  1. (6.6.3-a)

    SS is simple and almost ordinary, and EE is supersingular.

  2. (6.6.3-b)

    SS is non-simple and almost ordinary, and EE is supersingular.

  3. (6.6.3-c)

    SS is supersingular and EE is ordinary.

Proof.

As in Section 6.3, we let π1,π¯1,π2,π¯2\pi_{1},\overline{\pi}_{1},\pi_{2},\overline{\pi}_{2} and α,α¯\alpha,\overline{\alpha} be the Frobenius eigenvalues of SS and EE respectively. Denote the normalized eigenvalues by u1\colonequalsπ1/q,u2\colonequalsπ2/qu_{1}\colonequals\pi_{1}/\sqrt{q},u_{2}\colonequals\pi_{2}/\sqrt{q} and u\colonequalsα/qu\colonequals\alpha/\sqrt{q}.

Suppose first that XX is of type (6.6.3-a). By Lemma 5.3.1, the set {u1,u2}\mathopen{}\mathclose{{}\left\{u_{1},u_{2}}\right\} is multiplicatively independent. Since uu is a root of unity, UX=u1,u2,u=USUE𝐙2CmU_{X}=\langle u_{1},u_{2},u\rangle=U_{S}\oplus U_{E}\cong\mathbf{Z}^{2}\oplus C_{m} for mM={1,2,3,4,6,8,12}m\in M=\mathopen{}\mathclose{{}\left\{1,2,3,4,6,8,12}\right\} the set of possible torsion orders for supersingular elliptic curves. Thus in this case, 𝖲𝖥(X)𝖴(1)2×Cm\mathsf{SF}(X)\cong\mathsf{U}(1)^{2}\times C_{m} and is generated by (u1,u2,ζm)(u_{1},u_{2},\zeta_{m}).

If XX is of type (6.6.3-b), then SE1×E2S\sim E_{1}\times E_{2} with E1E_{1} ordinary and E2E_{2} supersingular. By Lemma 3.1.3, 𝖲𝖥(X)𝖴(1)×Cm\mathsf{SF}(X)\cong\mathsf{U}(1)\times C_{m}, with mm in the set of possible torsion orders of non-simple supersingular surfaces.

If XX is of type (6.6.3-c), we have UX=UEUS𝐙CmU_{X}=U_{E}\oplus U_{S}\cong\mathbf{Z}\oplus C_{m} for mm in the set M={1,2,3,4,5,6,8,10,12,24}M=\mathopen{}\mathclose{{}\left\{1,2,3,4,5,6,8,10,12,24}\right\} of possible torsion orders of supersingular surfaces from Lemma 5.5.1 and Lemma 5.6.1. ∎

Table 12. Serre–Frobenius groups of abelian threefolds of pp-rank 1.
Splitting type \cong class Generator mMm\in M Example
Absolutely simple 𝖴(1)3\mathsf{U}(1)^{3} (u1,u2,u3)(u_{1},u_{2},u_{3}) {1}\mathopen{}\mathclose{{}\left\{1}\right\} 3.2.ab_a_a
(6.6.3-a) 𝖴(1)2×Cm\mathsf{U}(1)^{2}\times C_{m} (u1,u2,ζmE)(u_{1},u_{2},\zeta_{m_{E}}) m=mE{1,2,3,4,6,8,12}m=m_{E}\in\mathopen{}\mathclose{{}\left\{1,2,3,4,6,8,12}\right\} -
(6.6.3-b) 𝖴(1)×Cm\mathsf{U}(1)\times C_{m} (u1,ζm1,ζm2)(u_{1},\zeta_{m_{1}},\zeta_{m_{2}}) m=lcm(m1,m2)m=\operatorname{lcm}(m_{1},m_{2}) in Table 8 -
(6.6.3-c) 𝖴(1)×Cm\mathsf{U}(1)\times C_{m} (u1,ζm1,ζm2)(u_{1},\zeta_{m_{1}},\zeta_{m_{2}}) m=lcm(m1,m2){1,2,3,4,5,6,8,10,12,24}m=\operatorname{lcm}(m_{1},m_{2})\in\mathopen{}\mathclose{{}\left\{1,2,3,4,5,6,8,10,12,24}\right\} Figure 16

The following examples are all of splitting type (6.6.3-c), since this splitting type contains all the new Serre–Frobenius groups appearing in Table 13. The histograms corresponding to these examples are presented in Figure 16.

Figure 16. a1a_{1}-distributions of pp-rank 11 threefolds with angle rank 1

6.7. Absolutely simple p-rank 0 threefolds

In this section, XX will be a non-supersingular pp-rank 0 abelian threefold over 𝐅q\mathbf{F}_{q}. Since the qq-Newton polygon of the Frobenius polynomial P(T)=PX(T)P(T)=P_{X}(T) has slopes 13\tfrac{1}{3} and 23\tfrac{2}{3}, each with multiplicity three, it follows that XX is absolutely simple, since the slope 1/31/3 does not occur for abelian varieties of smaller dimension. Let er2e_{r}^{2} denote the dimension of End0(X(r))\mathrm{End}^{0}(X_{(r)}) over its center. We consider two cases:

  1. (6.7-a)

    There exists r1r\geq 1 such that er=3e_{r}=3. In this case we have P(r)(T)=h(r)(T)3P_{(r)}(T)=h_{(r)}(T)^{3} and h(r)(T)h_{(r)}(T) is as in Theorem 6.1.1, so that 33 divides rlogp(q)r\cdot\log_{p}(q).

  2. (6.7-b)

    er=1e_{r}=1 for every positive integer rr.

Lemma 6.7.1 (Node X-H in Figure 7).

Let XX be an absolutely simple abelian threefold of pp-rank 0 defined over 𝐅q\mathbf{F}_{q}. Then, the Serre–Frobenius group of XX is classified according to Table 14. Furthermore, XX is of type (6.7-a), mXm_{X} is the smallest positive integer rr such that er=3e_{r}=3.

Table 13. Serre–Frobenius groups of absolutely simple abelian threefolds of pp-rank 0.
Case q=pdq=p^{d} \cong class Generator mMm\in M
(6.7-a) 3mXd3\mid m_{X}\cdot d 𝖴(1)×Cm\mathsf{U}(1)\times C_{m} (u1,ζmu1,ζmνu1)(u_{1},\zeta_{m}u_{1},\zeta_{m}^{\nu}u_{1}) {1,3,7}\mathopen{}\mathclose{{}\left\{1,3,7}\right\}
(6.7-b) - 𝖴(1)3\mathsf{U}(1)^{3} (u1,u2,u3)(u_{1},u_{2},u_{3}) {1}\mathopen{}\mathclose{{}\left\{1}\right\}
Remark 6.7.2.

The techniques for proving the Generalized Lenstra–Zarhin result in [8, Theorem 1.5], cannot be applied to this case. Thus, even the angle rank analysis in this case is particularly interesting.

Proof.

Suppose first that XX is of type (6.7-a), and let mm be the minimal positive integer such that em=3e_{m}=3. Maintaining previous notation, P(m)(T)=h(m)(T)3P_{(m)}(T)=h_{(m)}(T)^{3} implies that α2=ζα1\alpha_{2}=\zeta\cdot\alpha_{1} and α3=ξα1\alpha_{3}=\xi\cdot\alpha_{1} for mm-th roots of unity ζ\zeta and ξ\xi, whose orders have lcm mm. By Lemma 2.3.2, this implies that 𝖲𝖥(X)𝖴(1)×Cm\mathsf{SF}(X)\cong\mathsf{U}(1)\times C_{m}. We conclude that δX=1\delta_{X}=1 and m=mXm=m_{X}. To calculate the set MM of possible torsion orders, assume that mX=m>1m_{X}=m>1. Then 𝐐(α1m)\mathbf{Q}(\alpha_{1}^{m}) is a quadratic imaginary subextension of 𝐐(α1)𝐐\mathbf{Q}(\alpha_{1})\supset\mathbf{Q}, and we can argue as in the proof of Theorem 3.2.1 (with =3\ell=3) to conclude that m{3,7}m\in\mathopen{}\mathclose{{}\left\{3,7}\right\}.

Assume now that XX is of type (6.7-b). This implies that 𝐐(α1r)\mathbf{Q}(\alpha_{1}^{r}) is a degree 66 CM-field for every positive integer rr. If m:=mXm:=m_{X}, the base extension X(m)X_{(m)} is neat and the Frobenius eigenvalues are distinct and not supersingular. By Remark (3.3.2.b) we have that δX=3\delta_{X}=3 and m=1m=1. ∎

Example 6.7.3 (a1a_{1}-distribution for pp-rank 0 non-supersingular threefolds of splitting type (6.7-a)).

The histograms corresponding to these examples are presented in Figure 17. Note that the first one already showed up in Figure 9, while the other ones appeared in Figure 8.

  1. (A)

    (m=1m=1) The isogeny class 3.8.ag_bk_aea satisfies mX=1m_{X}=1. Note that 33 divides mXlog2(8)m_{X}\cdot\log_{2}(8).

  2. (B)

    (m=3m=3) The isogeny class 3.2.a_a_ac has angle rank 11 and irreducible Frobenius polynomial P(T)=T62T3+8P(T)=T^{6}-2T^{3}+8. The cubic base extension gives the isogeny class 3.8.ag_bk_aea with reducible Frobenius polynomial P(3)(T)=(T62T3+8)3P_{(3)}(T)=(T^{6}-2T^{3}+8)^{3}. Note that 33 divides mXlog2(2)m_{X}\cdot\log_{2}(2).

  3. (C)

    (m=7m=7) The isogeny class 3.8.ai_bk_aeq has angle rank 11 and irreducible Frobenius polynomial P(T)=T68T5+36T4120T3+288T2512T+512P(T)=T^{6}-8T^{5}+36T^{4}-120T^{3}+288T^{2}-512T+512. Its base change over a degree mX=7m_{X}=7 extension is the isogeny class 3.2097152.ahka_bfyoxc_adesazpwa with Frobenius polynomial

    P(7)(T)=(T21664T+2097152)3.P_{(7)}(T)=(T^{2}-1664T+2097152)^{3}.

    In this example, q=8q=8, so that 33 divides mXlog2(8)m_{X}\cdot\log_{2}(8).

Figure 17. a1a_{1}-distribution for pp-rank 0 non-supersingular threefolds of splitting type (6.7-a).

6.8. Simple supersingular threefolds

Nart and Ritzenthaler [21] showed that the only degree 66 supersingular qq-Weil numbers are the conjugates of

±qζ7,±qζ9, when q is a square, and\displaystyle\pm\sqrt{q}\zeta_{7},\pm\sqrt{q}\zeta_{9},\quad\text{ when }q\text{ is a square, and}
7d/2ζ28,3d/2ζ36, when q is not a square.\displaystyle 7^{d/2}\zeta_{28},3^{d/2}\zeta_{36},\quad\text{ when }q\text{ is not a square.}

Building on their work, Haloui [12, Proposition 1.5] completed the classification of simple supersingular threefolds. This classification is also discussed in [29]; and we adapt their notation for the polynomials of Z-3 type. Denoting by (a1,a2,a3)(a_{1},a_{2},a_{3}) the isogeny class of abelian threefolds over 𝐅q\mathbf{F}_{q} with Frobenius polynomial PX(T)=T6+a1T5+a2T4+a3T3+qa2T2+q2a1T+q3P_{X}(T)=T^{6}+a_{1}T^{5}+a_{2}T^{4}+a_{3}T^{3}+qa_{2}T^{2}+q^{2}a_{1}T+q^{3}, the following lemma gives the classification of Serre–Frobenius groups of simple supersingular threefolds, which is a corollary of Haloui’s result.

Lemma 6.8.1 (Node X-I in Figure 7).

Let XX be a simple supersingular abelian threefold defined over 𝐅q\mathbf{F}_{q}. The Serre–Frobenius group of XX is classified according to Table 15.

Table 14. Serre–Frobenius groups of simple supersingular threefolds.
(a1,a2,a3)(a_{1},a_{2},a_{3}) pp dd Type h~(T)\tilde{h}(T) 𝖲𝖥(X)\mathsf{SF}(X) Example
(q,q,qq)(\sqrt{q},q,q\sqrt{q}) 7(p31)7\nmid(p^{3}-1) even Z-1 Φ7(T)\Phi_{7}(T) C7C_{7} 3.9.d_j_bb
(q,q,qq)(-\sqrt{q},q,-q\sqrt{q}) 7(p31)7\nmid(p^{3}-1) even Z-1 Φ14(T)\Phi_{14}(T) C14C_{14} 3.9.ad_j_abb
(0,0,qq)(0,0,q\sqrt{q}) 1 mod 3\not\equiv 1\text{ mod }3 even Z-1 Φ9(T)\Phi_{9}(T) C9C_{9} 3.4.a_a_i
(0,0,qq)(0,0,-q\sqrt{q}) 1 mod 3\not\equiv 1\text{ mod }3 even Z-1 Φ18(T)\Phi_{18}(T) C18C_{18} 3.4.a_a_ai
(7q,3q,q7q)(\sqrt{7q},3q,q\sqrt{7q}) =7=7 odd Z-3 h7,1(T)h_{7,1}(T) C28C_{28} 3.7.h_v_bx
(7q,3q,q7q)(-\sqrt{7q},3q,-q\sqrt{7q}) =7=7 odd Z-3 h7,1(T)h_{7,1}(-T) C28C_{28} 3.7.ah_v_abx
(0,0,q3q)(0,0,q\sqrt{3q}) =3=3 odd Z-3 h3,3(T)h_{3,3}(T) C36C_{36} 3.3.a_a_j
(0,0,q3q)(0,0,-q\sqrt{3q}) =3=3 odd Z-3 h3,3(T)h_{3,3}(-T) C36C_{36} 3.3.a_a_aj
Proof.

By Theorem 6.1.1 and the discussion following it, we know that the Frobenius polynomial of every supersingular threefold PX(T)P_{X}(T), coincides with the minimal polynomial hX(T)h_{X}(T) with e=1e=1 in some row of the table. The first four rows of Table 15 correspond to isogeny classes of type (Z-1). By the discussion in Section 3.4, the minimal polynomials are of the form666Recall that f[a](T)\colonequalsadegff(T/a)f^{[a]}(T)\colonequals a^{\deg f}f(T/a). Φm[q](T)\Phi_{m}^{[\sqrt{q}]}(T) and the normalized polynomials are just the usual cyclotomic polynomials Φm(T)\Phi_{m}(T).

The last four rows of Table 15 correspond to isogeny classes of type (Z-3). The normalized Frobenius polynomials are h7,1(±T)=T6±7T5+3T4±7T3+3T2±7T+1h_{7,1}(\pm T)=\,T^{6}\pm\sqrt{7}T^{5}+3T^{4}\pm\sqrt{7}T^{3}+3T^{2}\pm\sqrt{7}T+1, and h3,3(±T)=T6±3T3+1.h_{3,3}(\pm T)=\,T^{6}\pm\sqrt{3}T^{3}+1. Noting that h7,1(T)h7,1(T)=Φ28(T)h_{7,1}(T)h_{7,1}(-T)=\Phi_{28}(T) and h3,3(T)h3,3(T)=Φ36(T)h_{3,3}(T)h_{3,3}(-T)=\Phi_{36}(T) we conclude that the unit groups UXU_{X} are generated by ζ28\zeta_{28} and ζ36\zeta_{36} respectively. ∎

6.9. Non-simple supersingular threefolds

If XX is a non-simple supersingular abelian threefold over 𝐅q\mathbf{F}_{q}, then there are two cases:

  1. (6.9.0-a)

    XS×EX\sim S\times E, with SS a simple supersingular surface over 𝐅q\mathbf{F}_{q} and EE a supersingular elliptic curve.

  2. (6.9.0-b)

    XE1×E2×E3X\sim E_{1}\times E_{2}\times E_{3}, where each EiE_{i} is a supersingular elliptic curve.

The classification of the Serre–Frobenius group in these cases can be summarized in the following lemma.

Lemma 6.9.1 (Node X-J in Figure 7).

If XX is a non-simple supersingular abelian threefold as in Case (6.9.0-a), then 𝖲𝖥(X)Cm\mathsf{SF}(X)\cong C_{m}, for mM(p,d)m\in M(p,d), where

  • if dd is even, M(p,d)={3,4,5,6,8,10,12,15,20,24,30}M(p,d)=\{3,4,5,6,8,10,12,15,20,24,30\}, and

  • if dd is odd, M(p,d)={4,8,12,20,24}M(p,d)=\{{4,8,12,20,24}\}.

Proof.

In this case, m=lcm(mS,mE)m=\operatorname{lcm}(m_{S},m_{E}), since this is the degree of the smallest extension over which the Serre–Frobenius group becomes connected. The list of values for mEm_{E} and mSm_{S} come from Table 3 and Table 7. ∎

Lemma 6.9.2 (Node X-J in Figure 7).

If XX is a non-simple supersingular abelian threefold as in Case (6.9.0-b), then 𝖲𝖥(X)Cm\mathsf{SF}(X)\cong C_{m}, for mM(p,d)m\in M(p,d), where

  • if dd is even, M(p,d)={1,2,3,4,6,12}M(p,d)=\{1,2,3,4,6,12\}, and

  • if dd is odd, M(p,d)={4,8,12}M(p,d)=\{4,8,12\}.

Proof.

We know that mm is the degree of the extension over which all the elliptic curve factors EiE_{i} become isogenous. This is precisely the least common multiple of the mEim_{E_{i}}’s. From Table 3, we can calculate the various possibilities for the lcm\operatorname{lcm}’s depending on the parity of dd. ∎

6.10. Acknowledgements

We would like to thank David Zureick-Brown, Kiran Kedlaya, Francesc Fité, Brandon Alberts, Edgar Costa, and Andrew Sutherland for useful conversations about this paper. We thank Yuri Zarhin for providing us with useful references, and Hendrik Lenstra for pointing out one of the missing cases in Section 4. We would also like to thank Everett Howe for helping us with a missing piece of the puzzle in Theorem 3.2.1. This project started as part of the Rethinking Number Theory workshop in 2021. We thank the organizers of the workshop for giving us the opportunity and space to collaborate, and the funding sources for the workshop: AIM, the Number Theory Foundation, and the University of Wisconsin-Eau Claire Department of Mathematics. We are also grateful to Rachel Pries for her guidance at the beginning of the workshop, which helped launch this project. Finally, we thank the anonymous referee for the elucidating and pertinent suggestions that improved the exposition and results in the paper.

References

  • Ahmadi and Shparlinski [2010] Ahmadi, O. and I. E. Shparlinski (2010). On the distribution of the number of points on algebraic curves in extensions of finite fields. Math. Res. Lett. 17(4), 689–699.
  • [2] Arango-Piñeros, S., D. Bhamidipati, and S. Sankar. Code accompanying “Frobenius distributions of abelian varieties over finite fields”. https://github.com/sarangop1728/Frobenius-distributions-AVs-Fq.
  • Chai et al. [2014] Chai, C.-L., B. Conrad, and F. Oort (2014). Complex multiplication and lifting problems, Volume 195 of Mathematical Surveys and Monographs. American Mathematical Society, Providence, RI.
  • Chi [1992] Chi, W. C. (1992). \ell-adic and λ\lambda-adic representations associated to abelian varieties defined over number fields. Amer. J. Math. 114(2), 315–353.
  • Deuring [1941] Deuring, M. (1941). Die Typen der Multiplikatorenringe elliptischer Funktionenkörper. Volume 14, pp.  197–272.
  • Diem and Naumann [2003] Diem, C. and N. Naumann (2003). On the structure of Weil restrictions of abelian varieties. J. Ramanujan Math. Soc. 18(2), 153–174.
  • Dupuy et al. [2021] Dupuy, T., K. Kedlaya, D. Roe, and C. Vincent ([2021] ©2021). Isogeny classes of abelian varieties over finite fields in the LMFDB. In Arithmetic geometry, number theory, and computation, Simons Symp., pp.  375–448. Springer, Cham.
  • Dupuy et al. [2024] Dupuy, T., K. S. Kedlaya, and D. Zureick-Brown (2024). Angle ranks of abelian varieties. Math. Ann. 389, 169–185.
  • Fité [2015] Fité, F. (2015). Equidistribution, LL-functions, and Sato-Tate groups. In Trends in number theory, Volume 649 of Contemp. Math., pp.  63–88. Amer. Math. Soc., Providence, RI.
  • Fité et al. [2012] Fité, F., K. S. Kedlaya, V. Rotger, and A. V. Sutherland (2012). Sato-Tate distributions and Galois endomorphism modules in genus 2. Compos. Math. 148(5), 1390–1442.
  • Fité et al. [2023] Fité, F., K. S. Kedlaya, and A. V. Sutherland (2023). Sato-Tate groups of abelian threefolds.
  • Haloui [2010] Haloui, S. (2010). The characteristic polynomials of abelian varieties of dimensions 3 over finite fields. J. Number Theory 130(12), 2745–2752.
  • Honda [1968] Honda, T. (1968). Isogeny classes of abelian varieties over finite fields. J. Math. Soc. Japan 20, 83–95.
  • Howe and Zhu [2002] Howe, E. W. and H. J. Zhu (2002). On the existence of absolutely simple abelian varieties of a given dimension over an arbitrary field. J. Number Theory 92(1), 139–163.
  • Krajíček and Scanlon [2000] Krajíček, J. and T. Scanlon (2000). Combinatorics with definable sets: Euler characteristics and Grothendieck rings. Bull. Symbolic Logic 6(3), 311–330.
  • Lenstra and Zarhin [1993] Lenstra, Jr., H. W. and Y. G. Zarhin (1993). The Tate conjecture for almost ordinary abelian varieties over finite fields. In Advances in number theory (Kingston, ON, 1991), Oxford Sci. Publ., pp.  179–194. Oxford Univ. Press, New York.
  • LMFDB Collaboration [2024] LMFDB Collaboration, T. (2024). The L-functions and modular forms database. https://www.lmfdb.org. [Online; accessed 15 March 2024].
  • Maisner and Nart [2002] Maisner, D. and E. Nart (2002). Abelian surfaces over finite fields as Jacobians. Experiment. Math. 11(3), 321–337. With an appendix by Everett W. Howe.
  • Milne [2017] Milne, J. S. (2017). Algebraic groups, Volume 170 of Cambridge Studies in Advanced Mathematics. Cambridge University Press, Cambridge. The theory of group schemes of finite type over a field.
  • Morris [1977] Morris, S. A. (1977). Pontryagin duality and the structure of locally compact abelian groups. London Mathematical Society Lecture Note Series, No. 29. Cambridge University Press, Cambridge-New York-Melbourne.
  • Nart and Ritzenthaler [2008] Nart, E. and C. Ritzenthaler (2008). Jacobians in isogeny classes of supersingular abelian threefolds in characteristic 2. Finite Fields Appl. 14(3), 676–702.
  • Oort [2008] Oort, F. (2008). Abelian varieties over finite fields. In Higher-dimensional geometry over finite fields, Volume 16 of NATO Sci. Peace Secur. Ser. D Inf. Commun. Secur., pp.  123–188. IOS, Amsterdam.
  • Pontrjagin [1934] Pontrjagin, L. (1934). The theory of topological commutative groups. Ann. of Math. (2) 35(2), 361–388.
  • Rűck [1990] Rűck, H.-G. (1990). Abelian surfaces and Jacobian varieties over finite fields. Compositio Math. 76(3), 351–366.
  • Sage Developers [2024] Sage Developers, T. (2024). Sagemath, the Sage Mathematics Software System (Version 9.5). https://www.sagemath.org.
  • Serre [1998] Serre, J.-P. (1998). Abelian \ell-adic representations and elliptic curves, Volume 7 of Research Notes in Mathematics. A K Peters, Ltd., Wellesley, MA. With the collaboration of Willem Kuyk and John Labute, Revised reprint of the 1968 original.
  • Serre [2013] Serre, J.-P. (2013). Oeuvres/Collected papers. IV. 1985–1998. Springer Collected Works in Mathematics. Springer, Heidelberg. Reprint of the 2000 edition [MR1730973].
  • Serre [2020] Serre, J.-P. ([2020] ©2020). Rational points on curves over finite fields, Volume 18 of Documents Mathématiques (Paris) [Mathematical Documents (Paris)]. Société Mathématique de France, Paris. With contributions by Everett Howe, Joseph Oesterlé and Christophe Ritzenthaler.
  • Singh et al. [2014] Singh, V., G. McGuire, and A. Zaytsev (2014). Classification of characteristic polynomials of simple supersingular abelian varieties over finite fields. Funct. Approx. Comment. Math. 51(2), 415–436.
  • Sutherland [2019] Sutherland, A. V. ([2019] ©2019). Sato-Tate distributions. In Analytic methods in arithmetic geometry, Volume 740 of Contemp. Math., pp.  197–248. Amer. Math. Soc., [Providence], RI.
  • Tate [1966] Tate, J. (1966). Endomorphisms of abelian varieties over finite fields. Invent. Math. 2, 134–144.
  • Tate [1971] Tate, J. (1971). Classes d’isogénie des variétés abéliennes sur un corps fini (d’après T. Honda). In Séminaire Bourbaki. Vol. 1968/69: Exposés 347–363, Volume 175 of Lecture Notes in Math., pp.  Exp. No. 352, 95–110. Springer, Berlin.
  • Waterhouse [1969] Waterhouse, W. C. (1969). Abelian varieties over finite fields. Ann. Sci. École Norm. Sup. (4) 2, 521–560.
  • Weil [1949] Weil, A. (1949). Numbers of solutions of equations in finite fields. Bull. Amer. Math. Soc. 55, 497–508.
  • Xing [1994] Xing, C. (1994). The characteristic polynomials of abelian varieties of dimensions three and four over finite fields. Sci. China Ser. A 37(2), 147–150.
  • Xing [1996] Xing, C. (1996). On supersingular abelian varieties of dimension two over finite fields. Finite Fields Appl. 2(4), 407–421.
  • Zarhin [1991] Zarhin, Y. G. (1991). Abelian varieties of K3K3 type and \ell-adic representations. In Algebraic geometry and analytic geometry (Tokyo, 1990), ICM-90 Satell. Conf. Proc., pp.  231–255. Springer, Tokyo.
  • Zarhin [1993] Zarhin, Y. G. (1993). Abelian varieties of K3K3 type. In Séminaire de Théorie des Nombres, Paris, 1990–91, Volume 108 of Progr. Math., pp.  263–279. Birkhäuser Boston, Boston, MA.
  • Zarhin [1994] Zarhin, Y. G. (1994). The Tate conjecture for nonsimple abelian varieties over finite fields. In Algebra and number theory (Essen, 1992), pp.  267–296. de Gruyter, Berlin.
  • Zarhin [2015] Zarhin, Y. G. (2015). Eigenvalues of Frobenius endomorphisms of abelian varieties of low dimension. J. Pure Appl. Algebra 219(6), 2076–2098.
  • Zhu [2001] Zhu, H. J. (2001). Supersingular abelian varieties over finite fields. J. Number Theory 86(1), 61–77.
  • Zywina [2022] Zywina, D. (2022). Determining monodromy groups of abelian varieties. Res. Number Theory 8(4), Paper No. 89, 53.