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

Restrictions on Weil polynomials of Jacobians of hyperelliptic curves

Edgar Costa Department of Mathematics, Massachusetts Institute of Technology, Cambridge, MA 02139, USA [email protected] https://edgarcosta.org Ravi Donepudi Department of Mathematics, University of Illinois at Urbana-Champaign, Urbana, IL, 61801, USA [email protected] https://faculty.math.illinois.edu/~donepud2/ Ravi Fernando Department of Mathematics, University of California — Berkeley, Berkeley, CA 94720, USA [email protected] https://math.berkeley.edu/~fernando/ Valentijn Karemaker Mathematical Institute, Utrecht University, 3508 TA Utrecht, the Netherlands [email protected] http://www.staff.science.uu.nl/~karem001/ Caleb Springer Department of Mathematics, The Pennsylvania State University, University Park, PA 16802, USA [email protected] http://personal.psu.edu/cks5320/  and  Mckenzie West Department of Mathematics, University of Wisconsin-Eau Claire, Eau Claire, WI 54701, USA [email protected] https://people.uwec.edu/westmr/
Abstract.

Inspired by experimental data, we investigates which isogeny classes of abelian varieties defined over a finite field of odd characteristic contain the Jacobian of a hyperelliptic curve. We provide a necessary condition by demonstrating that the Weil polynomial of a hyperelliptic Jacobian must have a particular form modulo 2. For fixed g1{g\geq 1}, the proportion of isogeny classes of gg-dimensional abelian varieties defined over 𝔽q\mathbb{F}_{q} which fail this condition is 1Q(2g+2)/2g1-Q(2g+2)/2^{g} as qq\to\infty ranges over odd prime powers, where Q(n)Q(n) denotes the number of partitions of nn into odd parts.

1. Introduction

The question of which abelian varieties arise as Jacobians of curves has a long and rich history. It has classically been investigated over the complex numbers as the Schottky Problem, using techniques from differential geometry and Hodge theory. In positive characteristic, some of these tools are no longer available, but instead the Frobenius endomorphism becomes a formidable weapon.

In this article, we study gg-dimensional abelian varieties which are defined over a finite field 𝔽q\mathbb{F}_{q} of odd cardinality. The 𝔽q\mathbb{F}_{q}-isogeny class of such an abelian variety AA is uniquely determined by the characteristic polynomial of its Frobenius endomorphism. This polynomial is called the Weil polynomial of the abelian variety and has the following form:

ZA(t)=t2g+a1t2g1++ag1tg+1+agtg+ag1qtg1++a1qg1t+qg[t].Z_{A}(t)=t^{2g}+a_{1}t^{2g-1}+\cdots+a_{g-1}t^{g+1}+a_{g}t^{g}+a_{g-1}qt^{g-1}+\cdots+a_{1}q^{g-1}t+q^{g}\in\mathbb{Z}[t].

Given ZA(t)Z_{A}(t), one would like to to determine if the isogeny class of AA contains the Jacobian of a smooth curve over 𝔽q\mathbb{F}_{q}.

In genus one, the problem is straightforward, as every one-dimensional abelian variety is an elliptic curve, which is isomorphic to its Jacobian. For genus two, the problem was solved by Howe, Nart, and Ritzenthaler [4] who give an explicit classification using only elementary restrictions involving the integers a1,a2a_{1},a_{2}, and qq. Their method relies on three key facts:

  1. (1)

    the Jacobian of a curve is canonically principally polarized;

  2. (2)

    a principally polarized abelian surface is isomorphic to exactly one of the following: a Jacobian of a genus two curve, a product of two elliptic curves, or the restriction of scalars of an elliptic curve from a quadratic extension of the ground field;

  3. (3)

    every genus two curve is hyperelliptic.

In particular, they take advantage of the canonical involution associated with any genus two curve.

Solving this problem in higher genus appears to be significantly more complicated. For example, every curve of genus three is isomorphic to either a hyperelliptic curve or to a smooth plane quartic curve. Curves of the latter type generically do not possess any non-trivial automorphisms, so the arguments of Howe, Nart and Ritzenthaler for genus two cannot easily be extended to non-hyperelliptic genus three curves.

In this paper, we focus on a more accessible question, namely, whether a given isogeny class of abelian varieties contains the Jacobian of a hyperelliptic curve. We do this by studying the geometric configurations of the Weierstrass points of hyperelliptic curves, as is done in the g=2g=2 case, e.g., [11, Appendix]. In doing so, we obtain parity conditions on the coefficients of the Weil polynomials which prevent certain isogeny classes from containing the Jacobian of a hyperelliptic curve. For instance, in genus three we obtain:

Theorem 2.8. Let qq be an odd prime power. The isogeny classes of three-dimensional abelian varieties corresponding to Weil polynomials of the form

t6+a1t5+a2t4+a3t3+qa2t2+q2a1t+q3t^{6}+a_{1}t^{5}+a_{2}t^{4}+a_{3}t^{3}+qa_{2}t^{2}+q^{2}a_{1}t+q^{3}

with a20(mod2)a_{2}\equiv 0\pmod{2} and a31(mod2)a_{3}\equiv 1\pmod{2} do not contain the Jacobian of a hyperelliptic curve over 𝔽q\mathbb{F}_{q}.

In fact, we can perform the same analysis for abelian varieties of any dimension; as far as we know, this is the first non-existence result for Jacobians in isogeny classes when g>2g>2 and qq is any odd prime power. Indeed, we demonstrate that the Weil polynomial of the Jacobian of a hyperelliptic curve of genus gg over a finite field of odd characteristic must be congruent modulo 2 to a polynomial of the form i=1r(tdi1)/(t1)2𝔽2[t]\prod_{i=1}^{r}(t^{d_{i}}-1)/(t-1)^{2}\in\mathbb{F}_{2}[t] where 2g+2=d1++dr2g+2=d_{1}+\dots+d_{r} is a partition. We call a Weil polynomial admissible if it takes this form modulo 2, and inadmissible otherwise. Notice that it is easy to test whether a polynomial is admissible by explicitly trying all possible partitions, and an isogeny class is guaranteed to not contain a hyperelliptic Jacobian if its Weil polynomial is inadmissible.

By using asymptotic results on the number of Weil polynomials of a fixed degree whose coefficients lie in prescribed congruence classes modulo an integer, it is possible to determine the asymptotic proportion of isogeny classes with admissible, or inadmissible, Weil polynomials.

Theorem 3.3. Let c(q,g)c(q,g) be the proportion of isogeny classes of gg-dimensional abelian varieties over 𝔽q\mathbb{F}_{q} with admissible Weil polynomial. For all g2g\geq 2 we have

limqc(q,g)=Q(2g+2)2g,\lim_{q\to\infty}c(q,g)=\frac{Q(2g+2)}{2^{g}},

as qq ranges over odd prime powers, where Q(2g+2)Q(2g+2) is the number of partitions of 2g+22g+2 into distinct parts.

For comparison, notice that for fixed g>6g>6, it is possible to see by simply counting the number of isogeny classes of gg-dimensional abelian varieties over 𝔽q\mathbb{F}_{q} versus the number of hyperelliptic curves of genus gg over 𝔽q\mathbb{F}_{q} that actually 0%0\% of gg-dimensional abelian varieties over 𝔽q\mathbb{F}_{q} contain a hyperelliptic Jacobian as qq\to\infty ranges over odd prime powers. However, this counting argument does not provide a way to identify which isogeny classes do or do not contain a hyperelliptic Jacobian. In contrast, for any fixed g1g\geq 1, any isogeny class with an inadmissible Weil polynomial is explicitly known to not contain a hyperelliptic Jacobian. Rephrasing the theorem above for fixed g>6g>6, the proportion of isogeny classes of gg-dimensional abelian varieties over 𝔽q\mathbb{F}_{q} which do not contain a hyperelliptic Jacobian but are not found by testing for inadmissible Weil polynomials is Q(2g+2)/2gQ(2g+2)/2^{g} as qq\to\infty ranges over odd prime powers. As noted in Remark 3.4, this discrepancy approaches zero as gg grows.

On the other hand, over the algebraic closure 𝔽¯q\overline{\mathbb{F}}_{q}, determining which isogeny classes contain Jacobians remains an open question. In order for our methods to shed light in this context, the behavior of admissibility under finite field extensions must first be understood. We leave this question for future work.

The outline of the paper is as follows. In Section 2, we recall some results on the geometry of hyperelliptic curves and their Jacobians and prove Theorem 2.8 and its generalizations to higher genus. In Section 3, we study asymptotic consequences of these non-existence results, deriving Theorem 3.3. In Section 4, we determine restrictions on point counts of curves, allowing us to reprove Theorem 2.8 by elementary methods. Finally, in Section 5, we present experimental data on the optimality of our results.

Acknowledgements

This project started at the AMS MRC workshop “Explicit methods in arithmetic geometry in characteristic pp” held at Whispering Pines (RI). The authors thank Andrew Sutherland for suggesting this problem, and Jeff Achter, Everett Howe, Bjorn Poonen, Christophe Ritzenthaler, David Roe, Andrew Sutherland, and Christelle Vincent for helpful discussions. They also thank the anonymous referees for careful reading and useful feedback. The first author was supported by the Simons Collaboration in Arithmetic Geometry, Number Theory, and Computation via Simons Foundation grant 550033. The third author was partially supported by NSF RTG grant DMS-1646385. The fourth author was supported by an AMS–Simons Travel Grant. The fifth author was partially supported by National Science Foundation award CNS-1617802.

2. Weil polynomials mod 2

Let CC be a hyperelliptic curve of genus gg defined over 𝔽q\mathbb{F}_{q} (of characteristic p2p\neq 2) with canonical degree two map π:C1\pi\colon C\to\mathbb{P}^{1}. Let W\colonequals{α1,α2,,α2g+2}W\colonequals\{\alpha_{1},\alpha_{2},\dots,\alpha_{2g+2}\} be the support of ramification divisor of π\pi, i.e., the geometric Weierstrass points of CC. The 𝔽q\mathbb{F}_{q}-Frobenius endomorphism Frob\operatorname{Frob} acts on WW by permuting its elements; we denote the cardinalities of the orbits of this action by did_{i}. (Alternatively, WW consists of some number, rr, of 𝔽q\mathbb{F}_{q}-places with respective degrees did_{i}, where d1++dr=2g+2d_{1}+\cdots+d_{r}=2g+2.)

The multiset dC{di}d_{C}\coloneqq\{d_{i}\} forms a partition of the integer 2g+22g+2; we call this the degree set of the curve CC, and by convention we order the did_{i} so that d1d2drd_{1}\leq d_{2}\leq\cdots\leq d_{r}.

Explicitly, CC can be given in coordinates by y2=f(x)y^{2}=f(x), where ff is a squarefree polynomial of degree either 2g+12g+1 or 2g+22g+2. Then WW consists of the points (α,0)(\alpha,0), where α\alpha runs over the roots of ff, together with the point at infinity if deg(f)=2g+1\deg(f)=2g+1. Then the did_{i} are precisely the degrees of the irreducible factors of ff, along with an extra 1 in the case deg(f)=2g+1\deg(f)=2g+1.

Remark 2.1.

If qq is small compared to the genus, not all partitions of 2g+22g+2 may arise as degree sets of genus gg hyperelliptic curves, due to the finite number of irreducible polynomials of any fixed degree in 𝔽q[x]\mathbb{F}_{q}[x]. For example, if 2g+2>q+12g+2>q+1, then the partition 2g+2=1++12g+2=1+\cdots+1 cannot be realized, as there are not enough 𝔽q\mathbb{F}_{q}-points for CC to ramify over.

Let JJ denote the Jacobian variety of CC whose elements are degree-zero divisors on CC modulo linear equivalence. Then the group J[2]J[2] of (geometric) 22-torsion of JJ is an 𝔽2\mathbb{F}_{2}-vector space that admits the following explicit description.

Lemma 2.2.

The group J[2]J[2] of 22-torsion elements of JJ forms a 2g2g-dimensional vector space over 𝔽2\mathbb{F}_{2}. Explicitly, this group can be expressed as the vector space obtained from 𝔽22g+2\mathbb{F}_{2}^{2g+2} by considering all vectors with an even number of non-zero entries and forming the quotient by (1,1,,1)\langle(1,1,\ldots,1)\rangle.

Proof.

The first claim follows since p2p\neq 2. For the second, we argue as follows; cf. [12, Corollary 2.11] over \mathbb{C}, which extends to any algebraically closed field of characteristic unequal to two. Every element of J[2]J[2] can be explicitly represented either by a divisor

eUPUP|U|(),e_{U}\coloneqq\sum_{P\in U}P-|U|(\infty),

if the map π:C1\pi\colon C\to\mathbb{P}^{1} is ramified at infinity, in which case the symbol \infty slightly abusively denotes π1()C\pi^{-1}(\infty)\in C, or by a divisor

eUPUP|U|2(1+2),e_{U}\coloneqq\sum_{P\in U}P-\frac{|U|}{2}(\infty_{1}+\infty_{2}),

if π\pi is split at infinity, in which case π1()={1,2}\pi^{-1}(\infty)=\{\infty_{1},\infty_{2}\}. Either way, the set UWU\subseteq W is a subset of even cardinality, and two divisors, eUe_{U} and eUe_{U^{\prime}}, represent the same element of J[2]J[2] if either U=UU=U^{\prime} or U=WUU=W\setminus U^{\prime}. In the 𝔽2\mathbb{F}_{2}-vector space VV whose standard basis is indexed by WW, the above set of representatives determine the subspace of vectors with an even number of non-zero entries. Two vectors v1,v2Vv_{1},v_{2}\in V yield the same element of J[2]J[2] precisely when their sum is contained in (1,1,,1)\langle(1,1,\dots,1)\rangle. The second claim follows. ∎

We will also need the following standard fact about vector spaces, whose proof is omitted.

Lemma 2.3.

Consider an exact sequence of vector spaces

0W1W2W300\to W_{1}\to W_{2}\to W_{3}\to 0

and a linear map T:W2W2T\colon W_{2}\to W_{2} such that T(W1)W1T(W_{1})\subseteq W_{1}. We will also denote the induced map T:W3W3T\colon W_{3}\to W_{3}. For i{1,2,3}i\in\{1,2,3\}, denote the characteristic polynomial of TT on WiW_{i} by χ(T,Wi)\chi(T,W_{i}). Then we have

χ(T,W2)=χ(T,W1)χ(T,W3).\chi(T,W_{2})=\chi(T,W_{1})\chi(T,W_{3}).

Denote the base-change of CC (resp. JJ) to 𝔽¯q\overline{\mathbb{F}}_{q} by CalgC^{\operatorname{{alg}}} (resp. JalgJ^{\operatorname{{alg}}}). Exploiting the action of Frob\operatorname{Frob} on the geometric Weierstrass points WW, we obtain the following result on the characteristic polynomial of Frob\operatorname{Frob} modulo 22.

Proposition 2.4.

Let CC be a hyperelliptic curve of genus gg defined over 𝔽q\mathbb{F}_{q}. Let {di}i=1r\{d_{i}\}_{i=1}^{r} be the partition of 2g+22g+2 which records the sizes of the orbits of Frobenius acting on the 2g+22g+2 geometric Weierstrass points. For any prime p\ell\neq p we have

(2.5) det(1Frobt|He´t1(Calg,))(i=1rtdi1)/(t1)2(mod2).\det\left(1-\operatorname{Frob}t\,|\,H_{\mathrm{\acute{e}t}}^{1}(C^{\operatorname{{alg}}},\mathbb{Q}_{\ell})\right)\equiv\Big{(}\prod_{i=1}^{r}{t}^{d_{i}}-1\Big{)}/{(t-1)}^{2}\pmod{2}.
Proof.

We have (cf. [10, Corollary 9.6]) that

det(1Frobt|He´t1(Calg,))\displaystyle\det\left(1-\operatorname{Frob}t\,|\,H_{\mathrm{\acute{e}t}}^{1}(C^{\operatorname{{alg}}},\mathbb{Q}_{\ell})\right) =det(1Frobt|He´t1(Jalg,))\displaystyle=\det\left(1-\operatorname{Frob}t\,|\,H_{\mathrm{\acute{e}t}}^{1}(J^{\operatorname{{alg}}},\mathbb{Q}_{\ell})\right)
=det(1Frobt|V(J)),\displaystyle=\det\left(1-\operatorname{Frob}t\,|\,V_{\ell}(J)\right),

where V(J)=T(J)V_{\ell}(J)=T_{\ell}(J)\otimes_{\mathbb{Z}_{\ell}}\mathbb{Q}_{\ell}, and where T(J)T_{\ell}(J) is the \ell-adic Tate module of JJ. By the Weil conjectures the polynomials above all have integer coefficients. By taking =2\ell=2, we further have

det(1Frobt|V2(J))det(1Frobt|J[2])(mod2).\det\left(1-\operatorname{Frob}t\,|\,V_{2}(J)\right)\equiv\det\left(1-\operatorname{Frob}t\,|\,J[2]\right)\pmod{2}.

As before, consider the 𝔽2\mathbb{F}_{2}-vector space W2\colonequals𝔽22g+2W_{2}\colonequals\mathbb{F}_{2}^{2g+2} whose standard basis is indexed by WW and which is acted on by the 𝔽q\mathbb{F}_{q}-Frobenius endomorphism Frob\operatorname{Frob}. By assumption, the action of Frob\operatorname{Frob} on W2W_{2} can be represented by a block-diagonal matrix with rr blocks, whose iith block (of order did_{i}) is a cyclic permutation of basis vectors. Since the characteristic polynomial of a cyclic permutation of order nn is tn1t^{n}-1, it follows that

χ(Frob,W2)=i=1r(tdi1).\chi(\operatorname{Frob},W_{2})=\prod_{i=1}^{r}\bigl{(}t^{d_{i}}-1\bigr{)}.

Furthermore, let W1W_{1} be the codimension-one subspace of vectors with an even number of non-zero entries. Then W1W_{1} is stable under Frob\operatorname{Frob}. Moreover, Frob\operatorname{Frob} acts invertibly on W1W_{1} and hence also on the one-dimensional quotient W3\colonequalsW2/W1W_{3}\colonequals W_{2}/W_{1}, so that χ(Frob,W3)=t1\chi(\operatorname{Frob},W_{3})=t-1. Applying Lemma 2.3 to the short exact sequence 0W1W2W2/W100\to W_{1}\to W_{2}\to W_{2}/W_{1}\to 0 yields

χ(Frob,W1)=i=1r(tdi1)t1.\chi(\operatorname{Frob},W_{1})=\frac{\prod_{i=1}^{r}(t^{d_{i}}-1)}{t-1}.

Now let W4W_{4} be the one-dimensional subspace of (1,1,,1)W1\langle(1,1,\ldots,1)\rangle\subseteq W_{1} and define W5W1/W4W_{5}\coloneqq W_{1}/W_{4}. Similarly, W4W_{4} is stable under Frob\operatorname{Frob} with χ(Frob,W4)=t1\chi(\operatorname{Frob},W_{4})=t-1 since Frob\operatorname{Frob} acts invertibly. Applying Lemma 2.3 to the short exact sequence 0W4W1W500\to W_{4}\to W_{1}\to W_{5}\to 0, we find

χ(Frob,W5)=i=1r(tdi1)(t1)2.\chi(\operatorname{Frob},W_{5})=\frac{\prod_{i=1}^{r}\bigr{(}t^{d_{i}}-1\bigl{)}}{(t-1)^{2}}.

Since, by Lemma 2.2, Frob\operatorname{Frob} acts on W5W_{5} as it does on J[2]J[2], the result follows. ∎

Remark 2.6.

It is possible to prove a congruence modulo 2 similar to (2.5) in Proposition 2.4 also when the finite field 𝔽q\mathbb{F}_{q} is allowed to have characteristic 2. In this case, Frobenius acts on Weierstrass points as it does on the étale part of J[2]J[2]. Since we require odd characteristic throughout the rest of the paper, notably in Theorem 3.1, we do not pursue this further.

Definition 2.7.

We call any polynomial of the form (2.5) an admissible Weil polynomial modulo 2. Note that the notion of admissibility is independent of qq.

By applying Proposition 2.4 to all possible partitions of 88 we obtain our main theorem.

Theorem 2.8.

Let qq be an odd prime power. The isogeny classes of three-dimensional abelian varieties corresponding to Weil polynomials of the form

t6+a1t5+a2t4+a3t3+qa2t2+q2a1t+q3t^{6}+a_{1}t^{5}+a_{2}t^{4}+a_{3}t^{3}+qa_{2}t^{2}+q^{2}a_{1}t+q^{3}

with a20(mod2)a_{2}\equiv 0\pmod{2} and a31(mod2)a_{3}\equiv 1\pmod{2} do not contain the Jacobian of a hyperelliptic curve over 𝔽q\mathbb{F}_{q}.

Proof.

For each of the twenty-two partitions {di}\{d_{i}\} of 2g+2=82g+2=8, we compute the corresponding polynomial (t1)2i=1r(tdi1)(t-1)^{-2}\prod_{i=1}^{r}(t^{d_{i}}-1). By Proposition 2.4, these are all of the admissible Weil polynomials modulo 22 for Jacobians of hyperelliptic curves of genus three. In Table 2.9, we tabulate the coefficients (a1,a2,a3)(a_{1},a_{2},a_{3}) of the resulting Weil polynomials modulo 22 with the corresponding partitions {di}\{d_{i}\}.

Coefficients (a1,a2,a3)(mod2)(a_{1},a_{2},a_{3})\pmod{2} Partition of 8
(0,1,1)(0,1,1) {3,5}\{3,5\}
(1,1,0)(1,1,0) {1,1,1,1,1,3}\{1,1,1,1,1,3\}, {1,1,1,2,3}\{1,1,1,2,3\}, {1,2,2,3}\{1,2,2,3\}, {1,3,4}\{1,3,4\}
(1,0,0)(1,0,0) {1,1,1,5}\{1,1,1,5\}, {1,2,5}\{1,2,5\}
(0,0,0)(0,0,0) {1,1,3,3}\{1,1,3,3\}, {1,1,6}\{1,1,6\}, {2,3,3}\{2,3,3\}, {2,6}\{2,6\}
(0,1,0)(0,1,0) {1,1,1,1,1,1,1,1}\{1,1,1,1,1,1,1,1\}, {1,1,1,1,1,1,2}\{1,1,1,1,1,1,2\}, {1,1,1,1,2,2}\{1,1,1,1,2,2\},
{1,1,1,1,4}\{1,1,1,1,4\}, {1,1,2,2,2}\{1,1,2,2,2\}, {1,1,2,4}\{1,1,2,4\},
{2,2,2,2}\{2,2,2,2\}, {2,2,4}\{2,2,4\}, {4,4}\{4,4\}, {8}\{8\}
(1,1,1)(1,1,1) {1,7}\{1,7\}
Table 2.9. Weil coefficients modulo 2 and corresponding partitions for threefolds.

Although Theorem 2.8 only considers the case of three-dimensional abelian varieties, Proposition 2.4 applies much more generally. Indeed, for any g1g\geq 1, we can produce a list of admissible Weil polynomials modulo 2 for Jacobians of hyperelliptic curves of genus gg, independent of qq. The following result counts the number of admissible polynomials in terms of partitions.

Proposition 2.10.

The number of admissible Weil polynomials modulo 2 for Jacobians of hyperelliptic curves of genus gg over finite fields of odd characteristic is equal to Q(2g+2)Q(2g+2), the number of partitions of 2g+22g+2 into distinct parts, or equivalently, the number of partitions of 2g+22g+2 into odd parts.

Proof.

It is well-known that the number of partitions of 2g+22g+2 into distinct parts is equal to the number of partitions of 2g+22g+2 into odd parts; see [14, Theorem 10.2]. For the remainder of this proof, our perspective will focus on partitions into distinct parts.

Proposition 2.4 shows how to compute admissible Weil polynomials modulo 2 using partitions, although many different partitions can correspond to a single admissible polynomial. For the purposes of this proof, we will call two partitions {d1,,dr}\{d_{1},\dots,d_{r}\} and {e1,,es}\{e_{1},\dots,e_{s}\} of an integer nn equivalent if

i=1r(tdi1)j=1s(tej1)(mod2).\prod_{i=1}^{r}\bigl{(}t^{d_{i}}-1\bigr{)}\equiv\prod_{j=1}^{s}\bigl{(}t^{e_{j}}-1\bigr{)}\pmod{2}.

Using Equation (2.5), it suffices to prove that every equivalence class of partitions of 2g+22g+2 contains precisely one partition with distinct parts.

Observe that every partition of 2g+22g+2 is equivalent to a partition with distinct parts. Indeed, if {d1,,dr+1}\{d_{1},\dots,d_{r+1}\} is any partition of 2g+22g+2 where dr=dr+1d_{r}=d_{r+1}, then we can construct another partition {d1,,dr1,dr}\{d_{1},\dots,d_{r-1},d^{\prime}_{r}\} with dr=2drd^{\prime}_{r}=2d_{r}. These two partitions are equivalent since t2dr1(tdr1)2mod2t^{2d_{r}}-1\equiv(t^{d_{r}}-1)^{2}\mod 2. Thus by induction on the number of equal parts, we conclude that every admissible Weil polynomial modulo 2 arises as a partition of 2g+22g+2 into distinct parts.

Now suppose that {d1,,dr}\{d_{1},\dots,d_{r}\} and {e1,,es}\{e_{1},\dots,e_{s}\} are two equivalent partitions of 2g+22g+2 into distinct parts. Without loss of generality, we order the parts so that d1d_{1} and e1e_{1} are the smallest parts of their respective partitions. After expanding the polynomials above, we see that the non-constant monomials of smallest degree are td1t^{d_{1}} and te1t^{e_{1}}, respectively. This implies that d1=e1d_{1}=e_{1}, hence {d2,,dr}\{d_{2},\dots,d_{r}\} and {e2,,es}\{e_{2},\dots,e_{s}\} are equivalent partitions of 2g+2d12g+2-d_{1}. Thus by induction on rr, we conclude that two equivalent partitions of 2g+22g+2 into distinct parts are equal. ∎

Remark 2.11.

In contrast with Remark 2.1, every admissible Weil polynomial mod 2 arises as the reduction mod 2 of a Weil polynomial for a hyperelliptic curve over 𝔽q\mathbb{F}_{q}, for each qq: Since there is at least one irreducible polynomial of every degree, we can construct a hyperelliptic curve ramified at points of degree did_{i}, where {di}\{d_{i}\} is the partition of 2g+22g+2 into distinct parts.

In Table 2.12, we tabulate the number of admissible and inadmissible Weil polynomials modulo 22 for Jacobians of hyperelliptic curves of genus g7g\leq 7, i.e., the numbers Q(2g+2)Q(2g+2) and 2gQ(2g+2)2^{g}-Q(2g+2).

gg 1 2 3 4 5 6 7
Q(2g+2)Q(2g+2) 2 4 6 10 15 22 32
100.00% 100.00% 75.00% 62.50% 46.88% 34.38% 25.00%
2gQ(2g+2)2^{g}-Q(2g+2) 0 0 2 6 17 42 96
0.00% 0.00% 25.00% 37.50% 53.12% 65.62% 75.00%
Table 2.12. The number of admissible and inadmissible Weil polynomials modulo 22 for hyperelliptic curves of small genus.

3. Asymptotics

The results in the previous section show that an isogeny class cannot contain a hyperelliptic Jacobian if the coefficients of the corresponding Weil polynomial satisfy a certain parity condition. Therefore, it is natural to ask how many Weil polynomials satisfy any given parity condition, or more generally, how many Weil polynomials are congruent to a given fixed polynomial of the correct form modulo an integer mm. The following theorem answers this question asymptotically for abelian varieties of fixed dimension.

Theorem 3.1.

Let g2g\geq 2 and mm be fixed integers, and let f(t)[t]f(t)\in\mathbb{Z}[t] be a fixed polynomial of the form

f(t)=t2g+a1t2g1++ag1tg+1+agtg+ag1qtg1++a1qg1t+qg.f(t)=t^{2g}+a_{1}t^{2g-1}+\cdots+a_{g-1}t^{g+1}+a_{g}t^{g}+a_{g-1}qt^{g-1}+\cdots+a_{1}q^{g-1}t+q^{g}.

For a prime power qq coprime to mm, write ef,m(q,g)e_{f,m}(q,g) for the proportion of isogeny classes of gg-dimensional abelian varieties over 𝔽q\mathbb{F}_{q} whose Weil polynomial is congruent to f(x)f(x) modulo mm. Then

limqef,m(q,g)=1mg,\lim_{q\to\infty}e_{f,m}(q,g)=\frac{1}{m^{g}},

where the limit is taken over all prime powers qq coprime to mm.

Proof.

This is essentially a theorem of Holden [5, Theorem 5], although we present a slightly more general statement here. Specifically, Holden’s theorem only considers the case when m=m=\ell is a prime, and the limit is taken over q=prq=p^{r} for a single fixed prime pp. However, the more general version of the theorem presented above is obtained immediately from Holden’s methods, as follows. Denote by I(q,g)I(q,g) the number of isogeny classes of gg-dimensional abelian varieties over 𝔽q\mathbb{F}_{q} and write If,m(q,g)I_{f,m}(q,g) for the number of such isogeny classes whose Weil polynomial is congruent to f(x)f(x) modulo mm. With this notation we may write ef,m(q,g)=If,m(q,g)I(q,g)e_{f,m}(q,g)=\frac{I_{f,m}(q,g)}{I(q,g)}. Using lattices, DiPippo and Howe obtained upper and lower bounds for I(q,g)I(q,g); see [1, Theorem 1.2] and [2]. Holden employed analogous techniques to obtain similar bounds for If,m(q,g)I_{f,m}(q,g); see [5, Proposition 2.2]. Using these bounds, as in the proof of [5, Theorem 5], we find

vgr(q)qg(g+1)/4mg2c(g,m)qg(g+1)/41/2m1gvgr(q)qg(g+1)/4+(vg+3c(g,1))qg(g+1)/41/2\displaystyle\frac{v_{g}r(q)q^{g(g+1)/4}m^{-g}-2c(g,m)q^{g(g+1)/4-1/2}m^{1-g}}{v_{g}r(q)q^{g(g+1)/4}+(v_{g}+3c(g,1))q^{g(g+1)/4-1/2}}
If,m(q,g)I(q,g)\displaystyle\hskip 56.9055pt\leq\frac{I_{f,m}(q,g)}{I(q,g)}
vgr(q)qg(g+1)/4mg+(vg+3c(g,1))qg(g+1)/41/2m1gvgr(q)qg(g+1)/42c(g,m)qg(g+1)/41/2,\displaystyle\hskip 56.9055pt\leq\frac{v_{g}r(q)q^{g(g+1)/4}m^{-g}+(v_{g}+3c(g,1))q^{g(g+1)/4-1/2}m^{1-g}}{v_{g}r(q)q^{g(g+1)/4}-2c(g,m)q^{g(g+1)/4-1/2}},

where vgv_{g} is a constant depending gg, and c(g,m)c(g,m) is a constant depending on both gg and mm, and r(q)=φ(q)/qr(q)=\varphi(q)/q, where φ\varphi denotes Euler totient function. Letting qq\to\infty, the theorem follows. ∎

Recall that, in light of Theorem 2.8 and Proposition 2.10, we say that the Weil polynomial of an isogeny class of gg-dimensional abelian varieties is admissible if it is congruent modulo 2 to a polynomial of the form i=1r(tdi1)/(t21)𝔽2[t]\prod_{i=1}^{r}(t^{d_{i}}-1)/(t^{2}-1)\in\mathbb{F}_{2}[t] for some partition 2g+2=d1++dr2g+2=d_{1}+\dots+d_{r}, and is inadmissible otherwise. By combining Theorems 2.8 and 3.1, we obtain the following corollary.

Corollary 3.2.

Denote by c(q)c(q) the proportion of isogeny classes of abelian threefolds over 𝔽q\mathbb{F}_{q} with admissible Weil polynomials. Then

limqc(q)=75%,\lim_{q\to\infty}c(q)=75\%,

where the limit is taken over all odd prime powers qq.

More generally, by combining Theorem 2.8 and Proposition 2.10, we can determine the proportion of isogeny classes of gg-dimensional abelian varieties over 𝔽q\mathbb{F}_{q} with admissible Weil polynomials in terms of the number of partitions of 2g+22g+2 into distinct parts.

Theorem 3.3.

Let c(q,g)c(q,g) be the proportion of isogeny classes of gg-dimensional abelian varieties over 𝔽q\mathbb{F}_{q} with admissible Weil polynomial. For all g2g\geq 2 we have

limqc(q,g)=Q(2g+2)2g,\lim_{q\to\infty}c(q,g)=\frac{Q(2g+2)}{2^{g}},

as qq ranges over odd prime powers, where Q(2g+2)Q(2g+2) is the number of partitions of 2g+22g+2 into distinct parts.

Proof.

By Proposition 2.10, we know there is a set of polynomials S={f1,,fQ(2g+2)}S=\{f_{1},\dots,f_{Q(2g+2)}\} such that the Weil polynomial of the Jacobian of a hyperelliptic curve of genus gg over any finite field of odd characteristic is equivalent modulo 2 to some polynomial in SS. Thus, the result follows immediately from Theorem 3.1. ∎

Remark 3.4.

For fixed g2g\geq 2, the number of isogeny classes of gg-dimensional abelian varieties over 𝔽q\mathbb{F}_{q} and the number of hyperelliptic curves of genus gg over 𝔽q\mathbb{F}_{q} are asymptotically bounded by qg(g+1)/4q^{g(g+1)/4} and q2g1q^{2g-1} , respectively, as qq\to\infty ranges over odd prime powers; see [1, Theorem 1.1], and [13, Table 1, Corollary 3.4]. For g>6g>6, this means that asymptotically 0%0\% of the isogeny classes of gg-dimensional abelian varieties over 𝔽q\mathbb{F}_{q} contain a hyperelliptic Jacobian as qq\to\infty ranges over odd prime powers. Comparing this reality to Theorem 3.3, we see that the proportion of isogeny classes of gg-dimensional abelian varieties that have an admissible Weil polynomial but still do not contain a hyperelliptic Jacobian is Q(2g+2)2g\frac{Q(2g+2)}{2^{g}} as q{q\to\infty}. Notice that this discrepancy gets smaller as gg grows. Indeed, Q(N)33/412N3/4exp(πN/3)Q(N)\sim\frac{3^{3/4}}{12N^{3/4}}\exp(\pi\sqrt{N/3}) as NN\to\infty [3, Figure 1.9], so it follows that Q(2g+2)2g0\frac{Q(2g+2)}{2^{g}}\to 0 as gg\to\infty.

4. Point counts

In Section 2 we determined restrictions modulo 22 for Weil polynomials of hyperelliptic Jacobians. These Weil polynomials govern the point counts of the corresponding hyperelliptic curve C/𝔽qC/\mathbb{F}_{q} over all extensions of 𝔽q\mathbb{F}_{q}. In this section, we determine 22-adic restrictions on the point counts of hyperelliptic curves over extensions of 𝔽q\mathbb{F}_{q} by more elementary means. Rather than studying the action of Frobenius on J[2]J[2], we study its action on the Weierstrass points of the curve directly. In particular, this provides an alternative proof of Theorem 2.8.

4.1. Restrictions on parity of point counts

As before, let qq be an odd prime power, C/𝔽qC/\mathbb{F}_{q} be a hyperelliptic curve of genus g>1g>1. Recall that we denote the support of the ramification divisor by WW and we denote by {di}\{d_{i}\} a partition of 2g+22g+2, corresponding to the decomposition of WW into Frobenius orbits. We begin with the following observations on the point counts of CC over extensions of 𝔽q\mathbb{F}_{q}.

Lemma 4.1.

For each n1n\geq 1, we have #W(𝔽qn)=i:di|ndi\#W(\mathbb{F}_{q^{n}})=\sum_{i:d_{i}|n}d_{i}.

Proof.

A point of degree dd contributes dd 𝔽qn\mathbb{F}_{q^{n}}-points if d|nd|n, and none otherwise. ∎

Lemma 4.2.

For each n1n\geq 1, we have #C(𝔽qn)#W(𝔽qn)(mod2)\#C(\mathbb{F}_{q^{n}})\equiv\#W(\mathbb{F}_{q^{n}})\pmod{2}.

Proof.

The 𝔽qn\mathbb{F}_{q^{n}}-points of WW are precisely the 𝔽qn\mathbb{F}_{q^{n}}-points of CC that are fixed by the hyperelliptic involution. Since all other points appear in pairs, it follows that #C(𝔽qn)#W(𝔽qn)(mod2)\#C(\mathbb{F}_{q^{n}})\equiv\#W(\mathbb{F}_{q^{n}})\pmod{2}. ∎

Corollary 4.3.

For each n1n\geq 1, we have #C(𝔽qn)#C(𝔽q2n)(mod2)\#C(\mathbb{F}_{q^{n}})\equiv\#C(\mathbb{F}_{q^{2n}})\pmod{2}.

Proof.

The numbers nn and 2n2n share the same set of odd divisors, therefore by Lemmas 4.1 and 4.2

#C(𝔽qn)#W(𝔽qn)=i:di|ndii:di|2ndi=#W(𝔽q2n)#C(𝔽q2n)(mod2).\#C(\mathbb{F}_{q^{n}})\equiv\#W(\mathbb{F}_{q^{n}})=\sum_{i:d_{i}|n}d_{i}\equiv\sum_{i:d_{i}|2n}d_{i}=\#W(\mathbb{F}_{q^{2n}})\equiv\#C(\mathbb{F}_{q^{2n}})\pmod{2}.\qed
Corollary 4.4.

For each nn, we have

n#{i:di=n}d|nμ(n/d)#C(𝔽qd)(mod2),\displaystyle n\cdot\#\{i:d_{i}=n\}\equiv\sum_{d|n}\mu(n/d)\#C(\mathbb{F}_{q^{d}})\pmod{2},

where μ\mu is the Möbius function.

Proof.

Using Lemma 4.1 we can restate Lemma 4.2 as

#C(𝔽qn)d|nd#{i:di=d}(mod2).\#C(\mathbb{F}_{q^{n}})\equiv\sum_{d|n}d\cdot\#\{i:d_{i}=d\}\pmod{2}.

The result follows by applying Möbius inversion. ∎

Remark 4.5.

For CC as above, consider the binary sequence

aC(n)#C(𝔽qn)(mod2).a_{C}(n)\coloneqq\#C(\mathbb{F}_{q^{n}})\pmod{2}.

Lemma 4.2 and Corollary 4.4 imply that aCa_{C} determines, and is determined by, the numbers n#{i:di=n}(mod2)n\cdot\#\{i:d_{i}=n\}\pmod{2} for all nn. These numbers are encoded in the degree set dCd_{C}; they tell us the parity of #(i:di=n)\#(i:d_{i}=n) when nn is odd, and give no information when nn is even.

More precisely, Lemma 4.2 and Corollary 4.4 give us a dictionary between the sequence aCa_{C} of parities of point counts and the set

{d:d is odd and appears an odd number of times in dC}.\{d:d\text{ is odd and appears an odd number of times in }d_{C}\}.

This is consistent with Proposition 2.10.

This is already enough to prove that some sequences of point count parities are inconsistent with the identity di=2g+2\sum d_{i}=2g+2. For example:

Lemma 4.6.

If the genus of CC is 3, then we cannot have #C(𝔽q)0,#C(𝔽q3)1\#C(\mathbb{F}_{q})\equiv 0,\#C(\mathbb{F}_{q^{3}})\equiv~{}1, and #C(𝔽q5)0(mod2)\#C(\mathbb{F}_{q^{5}})\equiv 0\pmod{2}.

Proof.

If such a curve CC existed, Corollary 4.4 implies that #{di=1}\#\{d_{i}=1\} is even, #{di=3}\#\{d_{i}=3\} is odd, and #{di=5}\#\{d_{i}=5\} is even. Thus #{di=3}=1\#\{d_{i}=3\}=1 and #{di=5}=0\#\{d_{i}=5\}=0, which contradicts #{di=1}\#\{d_{i}=1\} being even. ∎

4.2. Restrictions on point counts modulo powers of 2

We have just seen some restrictions on the point counts modulo 22 of hyperelliptic curves. In this section we will obtain further restrictions modulo higher powers of 2.

Fix an integer m1m\geq 1. Let GG be the group {±1}×Gal(𝔽q2m/𝔽q)\{\pm 1\}\times\operatorname{Gal}(\mathbb{F}_{q^{2^{m}}}/\mathbb{F}_{q}) of order 2m+12^{m+1}. This acts on the 𝔽q2m\mathbb{F}_{q^{2^{m}}}-points of CC, where the first factor acts by the hyperelliptic involution and the second by field automorphisms. We will study #C(𝔽q2m)(mod2m+1)\#C(\mathbb{F}_{q^{2^{m}}})\pmod{2^{m+1}} by examining the orbits of the action of GG.

Proposition 4.7.

If CC is a genus gg hyperelliptic curve over 𝔽q\mathbb{F}_{q}, then

#C(𝔽qn)2(qn+1)#W(𝔽qn)(mod2a+1),\#C(\mathbb{F}_{q^{n}})\equiv 2(q^{n}+1)-\#W(\mathbb{F}_{q^{n}})\pmod{2^{a+1}},

where n=2am1n=2^{a}\cdot m\geq 1 with mm odd.

Proof.

We will count the 𝔽qn\mathbb{F}_{q^{n}}-points of CC by considering the fibers of the hyperelliptic map π:C(𝔽qn)1(𝔽qn)\pi\colon C(\mathbb{F}_{q^{n}})\to\mathbb{P}^{1}(\mathbb{F}_{q^{n}}). If x1(𝔽qn)x\in\mathbb{P}^{1}(\mathbb{F}_{q^{n}}), then π1(x)\pi^{-1}(x) can be computed by extracting the square roots of f(x)f(x), where ff is a polynomial of degree 2g+12g+1 or 2g+22g+2. Accordingly, π1(x)\pi^{-1}(x) contains either zero, one, or two 𝔽qn\mathbb{F}_{q^{n}}-points, and a preimage of size one occurs if and only if xWx\in W. Further, if nn is even and xx is defined over 𝔽qn/2\mathbb{F}_{q^{n/2}} (equivalently, if its degree is not divisible by 2a2^{a}), then the preimage must have size two, since quadratic equations y2=f(x)y^{2}=f(x) over 𝔽qn/2\mathbb{F}_{q^{n/2}} can be solved over 𝔽qn\mathbb{F}_{q^{n}}. Thus, the preimage may have size zero only if the degree of xx is divisible by 2a2^{a}. In this case, the Galois orbit of xx has size divisible by 2a2^{a}, so if there are 𝔽qn\mathbb{F}_{q^{n}}-points above xx, they occur in an orbit whose size is divisible by 2a+12^{a+1}. Hence, to count modulo 2a+12^{a+1}, we may assume that all unramified 𝔽qn\mathbb{F}_{q^{n}}-points of 1\mathbb{P}^{1} have exactly two preimages defined over 𝔽qn\mathbb{F}_{q^{n}}, and subtract the number of 𝔽qn\mathbb{F}_{q^{n}}-points of WW to obtain the result. ∎

As with Proposition 2.4, this implies that certain sequences of point counts of hyperelliptic curves (equivalently, certain Weil polynomials of their Jacobians) cannot be realized.

Corollary 4.8.

Assume that the genus of CC is 33. If #C(𝔽q)1(mod2)\#C(\mathbb{F}_{q})\equiv 1\pmod{2} and #C(𝔽q3)0(mod2)\#C(\mathbb{F}_{q^{3}})\equiv 0\pmod{2}, then we have #C(𝔽q2m)1(mod2m+1)\#C(\mathbb{F}_{q^{2^{m}}})\equiv-1\pmod{2^{m+1}} for all m1m\geq 1.

Proof.

Since #C(𝔽q)\#C(\mathbb{F}_{q}) is odd, we have that #{di=1}\#\{d_{i}=1\} is odd by Lemma 4.2. Analogously, since #{di=1}\#\{d_{i}=1\} is odd and #C(𝔽q3)\#C(\mathbb{F}_{q^{3}}) is even, it follows that #{di=3}\#\{d_{i}=3\} is odd and hence must equal one. Altogether, we have dj=3d_{j}=3 for a unique jj and dj{1,2,4}d_{j^{\prime}}\in\{1,2,4\} for jjj^{\prime}\neq j. Hence #(k1W(𝔽q2k))=5\#\left(\bigcup_{k\geq 1}W(\mathbb{F}_{q^{2^{k}}})\right)=5, and so #W(𝔽q2m)5(mod2m+1)\#W(\mathbb{F}_{q^{2^{m}}})\equiv 5\pmod{2^{m+1}}. The result follows by applying Proposition 4.7 to n=2mn=2^{m}, since we have q2m1(mod2m+1)q^{2^{m}}\equiv 1\pmod{2^{m+1}} for qq odd. ∎

4.3. Reproving Theorem 2.8

The results in the previous subsections can be used to reprove Theorem 2.8 as follows:

Alternative proof of Theorem 2.8.

Suppose that the statement is false, so that there is some hyperelliptic curve CC of genus three whose Jacobian has the prescribed Weil polynomial. By using Newton’s formulae and [10, Theorem 11.1], we can write the point counts #C(𝔽qk)\#C(\mathbb{F}_{q^{k}}) in terms of the coefficients of the Weil polynomial for all k1k\geq 1, as follows:

#C(𝔽q)\displaystyle\#C(\mathbb{F}_{q}) =q+1+a1;\displaystyle=q+1+a_{1};
#C(𝔽q2)\displaystyle\#C(\mathbb{F}_{q^{2}}) =q2+1a12+2a2;\displaystyle=q^{2}+1-a_{1}^{2}+2a_{2};
#C(𝔽q3)\displaystyle\#C(\mathbb{F}_{q^{3}}) =q3+1+a133a1a2+3a3;\displaystyle=q^{3}+1+a_{1}^{3}-3a_{1}a_{2}+3a_{3};
#C(𝔽q4)\displaystyle\#C(\mathbb{F}_{q^{4}}) =q4+1a14+4a12a24a1a32a22+4qa2;\displaystyle=q^{4}+1-a_{1}^{4}+4a_{1}^{2}a_{2}-4a_{1}a_{3}-2a_{2}^{2}+4qa_{2};
#C(𝔽q5)\displaystyle\#C(\mathbb{F}_{q^{5}}) =q5+1+a155a13a2+5a1a225qa1a25a2a3+5q2a1.\displaystyle=q^{5}+1+a_{1}^{5}-5a_{1}^{3}a_{2}+5a_{1}a_{2}^{2}-5qa_{1}a_{2}-5a_{2}a_{3}+5q^{2}a_{1}.

We will complete our proof by showing that the restrictions a20(mod2)a_{2}\equiv 0\pmod{2} and a31(mod2)a_{3}\equiv 1\pmod{2} on the coefficients force the point counts to fall into the impossible cases listed in the lemmas above. First consider the case when a10(mod2)a_{1}\equiv 0\pmod{2}. Expanding the equations above leads to

#C(𝔽q)\displaystyle\#C(\mathbb{F}_{q}) 0(mod2),\displaystyle\equiv 0\pmod{2}, #C(𝔽q2)\displaystyle\#C(\mathbb{F}_{q^{2}}) 2(mod4),\displaystyle\equiv 2\pmod{4},
#C(𝔽q3)\displaystyle\#C(\mathbb{F}_{q^{3}}) 1(mod2),\displaystyle\equiv 1\pmod{2}, #C(𝔽q4)\displaystyle\#C(\mathbb{F}_{q^{4}}) 2(mod8),\displaystyle\equiv 2\pmod{8},
#C(𝔽q5)\displaystyle\#C(\mathbb{F}_{q^{5}}) 0(mod2).\displaystyle\equiv 0\pmod{2}.

In particular, Lemma 4.6 applies, and we reach a contradiction.

Now assume that a11(mod2)a_{1}\equiv 1\pmod{2}. The analogous computation gives

#C(𝔽q)\displaystyle\#C(\mathbb{F}_{q}) 1(mod2),\displaystyle\equiv 1\pmod{2}, #C(𝔽q2)\displaystyle\#C(\mathbb{F}_{q^{2}}) 1(mod4),\displaystyle\equiv 1\pmod{4},
#C(𝔽q3)\displaystyle\#C(\mathbb{F}_{q^{3}}) 0(mod2),\displaystyle\equiv 0\pmod{2}, #C(𝔽q4)\displaystyle\#C(\mathbb{F}_{q^{4}}) 5(mod8),\displaystyle\equiv 5\pmod{8},
#C(𝔽q5)\displaystyle\#C(\mathbb{F}_{q^{5}}) 0(mod2).\displaystyle\equiv 0\pmod{2}.

Corollary 4.8 applies, and we again reach a contradiction. ∎

Remark 4.9.

The argument above can be automated, and by iterating over partitions of 2g+22g+2, and applying Proposition 4.7, one can rule out Weil polynomials modulo 22, similarly to Proposition 2.10. We have verified that the sets of unrealizable Weil polynomials obtained by these two different procedures agree up to genus g10g\leq 10 by using the Newton identities. Indeed, we expect both sets to match for any gg, although we are unable to prove it in general.

5. Experimental data

The main results of this paper were inspired by experimental data, which we include here to illustrate the phenomena. By performing an exhaustive search, the number of Jacobians of hyperelliptic curves in each isogeny class over 𝔽q\mathbb{F}_{q} (up to isomorphism of principally polarized abelian varieties) has been computed for threefolds and prime powers q13q\leq 13 that are either prime or odd. Similarly, by iterating over the isomorphism classes of smooth plane quartic curves, the number of isogeny classes of abelian threefolds which contain the Jacobian of a smooth plane quartic curve has been computed for q=2,3q=2,3, and 55. Both searches were done by Andrew Sutherland using the techniques developed in [7, 8], and the data have been incorporated in the LL-functions and Modular Forms Database [9]; see www.LMFDB.org/Variety/Abelian/Fq/.

By combining these data sets, one can also deduce which isogeny classes of abelian threefolds do not contain a Jacobian for q=2,3q=2,3, and 55, see Table 5.1.

qq Contains Jacobian Total
Yes No Hyperelliptic Quartic curve
2 108 107 59 73 215
50.23% 49.77% 27.44% 33.95%
3 479 198 297 389 677
70.75% 29.25% 43.87% 57.46%
5 2 611 342 1 723 2 471 2 953
88.42% 11.58% 58.35% 83.68%
Table 5.1. Types of Jacobians per isogeny classes of abelian threefolds.

These data sets, and more precisely the multisets of virtual point counts modulo 2 and 4 that can be extracted from this data, provide motivation for Theorem 2.8.

In Figure 5.2, we compare the proportion of isogeny classes of abelian threefolds over 𝔽q\mathbb{F}_{q} which do not contain a hyperelliptic Jacobian for q13q\leq 13 with the proportion of such isogeny classes which are ruled out via Theorem 2.8, giving us some insight into the efficiency of Theorem 2.8 as qq grows.

Refer to caption
Figure 5.2. Effectiveness of Theorem 2.8.

Examples of isogeny classes which do not contain a (hyperelliptic) Jacobian but are not ruled out by Theorem 2.8 are given by the following Weil polynomials, cf. isogeny classes 3.3.a_ab_ac and 3.5.ac_j_aw on the LMFDB [9]:

Z1(t)\displaystyle Z_{1}(t) =t6t42t33t2+27,\displaystyle=t^{6}-t^{4}-2t^{3}-3t^{2}+27,
Z2(t)\displaystyle Z_{2}(t) =t62t5+9t422t3+45t250t+125.\displaystyle=t^{6}-2t^{5}+9t^{4}-22t^{3}+45t^{2}-50t+125.

In fact, neither isogeny class contains the Jacobian of any curve. These isogeny classes correspond to the row of Table 2.9 with the most partitions.

In Table 5.3, we present the number of isogeny classes which are ruled out by Theorem 2.8, for 17q3117\leq q\leq 31. The data show that the proportion of isogeny classes ruled out by Theorem 2.8 quickly approaches 25%, as expected; see Corollary 3.2. Furthermore, naive extrapolation seems to indicate that the proportion of isogeny classes of abelian threefolds over 𝔽q\mathbb{F}_{q} which do not contain a hyperelliptic Jacobian approaches 25%25\% from above as qq\to\infty ranges over odd prime powers, while the proportion of the isogeny classes that are ruled out via Theorem 2.8 approaches 25%25\% from below. Similarly, in Tables 5.4 and 5.5 we display the number of isogeny classes which cannot contain a hyperelliptic Jacobian by Proposition 2.4, and as before, we observe proportions which are already very close to the ones attained in the qq-limit; see Corollary 3.2.

qq 17 19 23 25 27 29 31
Number of isogeny classes 112 283 156 589 277 517 332 166 333 695 555 843 678 957
Inadmissible by Thm. 2.8 27 974 39 034 69 268 82 564 83 350 138 730 169 574
24.91% 24.93% 24.96% 24.86% 24.98% 24.96% 24.98%
Table 5.3. Isogeny classes ruled out by Proposition 2.4 for genus 3 and small qq.
qq 3 5 7 9 11 13 qq\rightarrow\infty
Number of isogeny classes 10 963 132 839 705 593 2 232 114 6 718 947 15 477 119
Inadmissible by Prop. 2.4 3 856 48 910 262 564 829 189 2 513 570 5 794 772
35.17% 36.82% 37.21% 37.15% 37.41% 37.44% 37.50%
Table 5.4. Isogeny classes ruled out by Proposition 2.4 for genus 4 and small qq.
qq 3 5 qq\rightarrow\infty
Number of isogeny classes 267 465 11 902 325
Inadmissible by Prop. 2.4 137 866 6 286 570
51.55% 52.82% 53.12%
Table 5.5. Isogeny classes ruled out by Proposition 2.4 in genus 5 and small qq.

To generate Tables 5.35.4 and 5.5, we enumerated all isogeny classes through their Weil polynomials: First, we enumerated Weil polynomials of degree 2g2g, and then we filtered by the Honda–Tate condition on its factors, see [15, Chapter 2], to only keep the ones that correspond to an isogeny class of abelian varieties of dimension gg. We enumerated Weil polynomials using root-unitary111https://github.com/kedlaya/root-unitary, which implements the techniques introduced in [6].

References

  • DH [98] Stephen A. DiPippo and Everett W. Howe. Real polynomials with all roots on the unit circle and abelian varieties over finite fields. J. Number Theory, 73(2):426–450, 1998.
  • DH [00] Stephen A. DiPippo and Everett W. Howe. Corrigendum: “Real polynomials with all roots on the unit circle and abelian varieties over finite fields” [J. Number Theory 73 (1998), no. 2, 426–450; MR1657992 (2000c:11101)]. J. Number Theory, 83(1):182, 2000.
  • FS [09] Philippe Flajolet and Robert Sedgewick. Analytic combinatorics. Cambridge University Press, Cambridge, 2009.
  • HNR [09] Everett W. Howe, Enric Nart, and Christophe Ritzenthaler. Jacobians in isogeny classes of abelian surfaces over finite fields. Ann. Inst. Fourier (Grenoble), 59(1):239–289, 2009.
  • Hol [04] Joshua Holden. Abelian varieties over finite fields with a specified characteristic polynomial modulo ll. J. Théor. Nombres Bordeaux, 16(1):173–178, 2004.
  • Ked [08] Kiran S. Kedlaya. Search techniques for root-unitary polynomials. In Computational arithmetic geometry, volume 463 of Contemp. Math., pages 71–81. Amer. Math. Soc., Providence, RI, 2008.
  • KS [08] Kiran S. Kedlaya and Andrew V. Sutherland. Computing LL-series of hyperelliptic curves. In Algorithmic number theory, volume 5011 of Lecture Notes in Comput. Sci., pages 312–326. Springer, Berlin, 2008.
  • KS [16] Kiran S. Kedlaya and Andrew V. Sutherland. A census of zeta functions of quartic K3 surfaces over 𝔽2\mathbb{F}_{2}. LMS J. Comput. Math., 19(suppl. A):1–11, 2016.
  • LMF [19] The LMFDB Collaboration. The L-functions and modular forms database. http://www.lmfdb.org, 2019. [Online; accessed 30 October 2019].
  • Mil [86] J. S. Milne. Jacobian varieties. In Arithmetic geometry (Storrs, Conn., 1984), pages 167–212. Springer, New York, 1986.
  • MN [02] Daniel Maisner and Enric Nart. Abelian surfaces over finite fields as Jacobians. Experiment. Math., 11(3):321–337, 2002. With an appendix by Everett W. Howe.
  • Mum [07] David Mumford. Tata lectures on theta. II. Modern Birkhäuser Classics. Birkhäuser Boston, Inc., Boston, MA, 2007. Jacobian theta functions and differential equations, With the collaboration of C. Musili, M. Nori, E. Previato, M. Stillman and H. Umemura, Reprint of the 1984 original.
  • Nar [09] Enric Nart. Counting hyperelliptic curves. Adv. Math., 221(3):774–787, 2009.
  • NZM [91] Ivan Niven, Herbert S. Zuckerman, and Hugh L. Montgomery. An introduction to the theory of numbers. John Wiley & Sons, Inc., New York, fifth edition, 1991.
  • Wat [69] William C. Waterhouse. Abelian varieties over finite fields. Ann. Sci. École Norm. Sup. (4), 2:521–560, 1969.