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

On \ell-torsion in degree \ell superelliptic Jacobians over 𝐅q\mathbf{F}_{q}

Wanlin Li Department of Mathematics, Washington University in St. Louis, 1 Brookings Dr, St.Louis, MO 63105, USA [email protected] Jonathan Love Mathematics Institute, Leiden University, Leiden, the Netherlands [email protected]  and  Eric Stubley [email protected]
(Date: December 29, 2024)
Abstract.

We study the \ell-torsion subgroup in Jacobians of curves of the form y=f(x)y^{\ell}=f(x) for irreducible f(x)f(x) over a finite field 𝐅q\mathbf{F}_{q} of characteristic pp\neq\ell. This is a function field analogue of the study of \ell-torsion subgroups of ideal class groups of number fields 𝐐(N)\mathbf{Q}(\sqrt[\ell]{N}). We establish an upper bound, lower bound, and parity constraint on the rank of the \ell-torsion which depend only on the parameters \ell, qq, and degf\deg f. Using tools from class field theory, we show that additional criteria depending on congruence conditions involving the polynomial f(x)f(x) can be used to refine the upper and lower bounds. For certain values of the parameters ,q,degf\ell,q,\deg f, we determine the \ell-torsion of the Jacobian for all curves with the given parameters.

Key words and phrases:
ideal class group, superelliptic curve, Jacobian, Weil pairing, Galois cohomology
2020 Mathematics Subject Classification:
Primary 11R29, 11R58; Secondary 11R34, 11R37, 11G20, 11G45

1. Introduction

The ideal class group of a number field is one of the central topics of interest in algebraic number theory. If we consider the collection of degree nn extensions K/𝐐K/\mathbf{Q} with some fixed Galois group, then for all but finitely many primes \ell, the \ell-torsion of the class group of KK is conjectured to be distributed according to the Cohen-Lenstra heuristics [3]. If n\ell\mid n, however, the \ell-torsion is expected to have qualitatively different behavior. For instance, in the case =2\ell=2 and KK is an imaginary quadratic field, Gauss’ genus theory completely describes the 22-torsion of the class group of KK in terms of the number of ramified primes. In general, if 3\ell\geq 3 divides nn, then the \ell-torsion structure can be considerably more mysterious.

In [12], the authors used Galois cohomology to study the \ell-torsion of the ideal class groups of the degree \ell number fields 𝐐(N)\mathbf{Q}(\sqrt[\ell]{N}) for prime NN; see Section 1.4 for more on the history of this problem. In this paper, we study an analogous problem over global function fields, namely the divisor class groups of fields of the form 𝐅q(f,x)\mathbf{F}_{q}(\sqrt[\ell]{f},x) for f(x)𝐅q[x]f(x)\in\mathbf{F}_{q}[x] irreducible. In this setting, we are able to utilize both Galois cohomology inspired by [12] and tools from arithmetic geometry to obtain more refined constraints on the \ell-torsion, and we encounter interesting behavior which does not occur in the number field setting.

Computing the \ell-torsion structure of the divisor class group of a function field is typically a computationally intensive problem that requires first finding the full class group. We produce constraints on the \ell-torsion using data that are much easier to compute, and in some cases, these constraints uniquely determine the \ell-torsion. The full results are discussed in Section 1.1, but we give one example application here.

Theorem 1.1.

Let 3\ell\geq 3 be prime, qq a prime power with q21modq^{2}\equiv 1\bmod\ell, and f(x)𝐅q[x]f(x)\in\mathbf{F}_{q}[x] irreducible with degf\deg f coprime to \ell. The \ell-torsion of the divisor class group of 𝐅q(f,x)\mathbf{F}_{q}(\sqrt[\ell]{f},x) is isomorphic to (𝐙/𝐙)(1)/2(\mathbf{Z}/\ell\mathbf{Z})^{(\ell-1)/2} if q1modq\equiv-1\bmod\ell and degf\deg f is even, and is trivial otherwise.

If =3\ell=3, and qq and degf\deg f are coprime to 33, Theorem 1.1 shows that the 33-torsion can be determined using no information about ff other than its degree. If =5\ell=5, and qq and degf\deg f are coprime to 55, then we can completely determine the 55-torsion structure using easily computable conditions depending on ff (Corollary 1.7).

77-torsion:    0 (𝐙/7𝐙)(\mathbf{Z}/7\mathbf{Z})^{\phantom{1}} (𝐙/7𝐙)2(\mathbf{Z}/7\mathbf{Z})^{2} (𝐙/7𝐙)3(\mathbf{Z}/7\mathbf{Z})^{3} (𝐙/7𝐙)4(\mathbf{Z}/7\mathbf{Z})^{4} (𝐙/7𝐙)5(\mathbf{Z}/7\mathbf{Z})^{5}
count: 0 55525552 0 18401840 0 1212
Table 1. The number of isomorphism classes of fields 𝐅3(f7,x)\mathbf{F}_{3}(\sqrt[7]{f},x) attaining each possible 77-torsion structure in its divisor class group, with f(x)𝐅3[x]f(x)\in\mathbf{F}_{3}[x] irreducible of degree 1212.

For 7\ell\geq 7, the \ell-torsion structure is typically not fully determined by the easily computable conditions mentioned above, but we prove a parity constraint which gives us a better understanding of the \ell-torsion. To illustrate this phenomenon, up to isomorphism there are 74047404 function fields of the form 𝐅3(f7,x)\mathbf{F}_{3}(\sqrt[7]{f},x) with f(x)𝐅3[x]f(x)\in\mathbf{F}_{3}[x] irreducible of degree 1212. For each of these fields, the authors used Magma to compute the divisor class group and recorded the 77-torsion structure of each; see Table 1. While the relative distribution of curves across the possible torsion structures is a subject of future exploration, our results explain the zeroes in the table. More generally, we will see that the largest power of (𝐙/𝐙)(\mathbf{Z}/\ell\mathbf{Z}) occurring as a subgroup of the divisor class group of 𝐅q(f,x)\mathbf{F}_{q}(\sqrt[\ell]{f},x) must be odd whenever 3\ell\geq 3, qq is a primitive root mod \ell, and degf\deg f is even and coprime to \ell (Theorem 1.3). It seems as though this phenomenon is unique to the function field setting and does not arise for number fields.

1.1. Main results

For all the results that follow, we assume 3\ell\geq 3 is prime, qq is a prime power coprime to \ell, and f(x)𝐅q[x]f(x)\in\mathbf{F}_{q}[x] is an irreducible polynomial with d:=degfd:=\deg f coprime to \ell. Let CC be the smooth projective curve with affine equation given by y=f(x)y^{\ell}=f(x); such a curve is an example of a “superelliptic curve.” Let JJ be the Jacobian of CC, so the degree 0 subgroup of the divisor class group of CC is isomorphic to J(𝐅q)J(\mathbf{F}_{q}). The \ell-torsion of J(𝐅q)J(\mathbf{F}_{q}) can be equipped with the structure of a vector space over 𝐅\mathbf{F}_{\ell}, and we define the \ell-rank of CC to be the dimension of this 𝐅\mathbf{F}_{\ell} vector space,

r(C):=dim𝐅J[](𝐅q).r_{\ell}(C):=\dim_{\mathbf{F}_{\ell}}J[\ell](\mathbf{F}_{q}).

The function field of CC is isomorphic to 𝐅q(f,x)\mathbf{F}_{q}(\sqrt[\ell]{f},x), and up to isomorphism CC is the only smooth projective curve with this function field. We define the divisor class group of 𝐅q(f,x)\mathbf{F}_{q}(\sqrt[\ell]{f},x) to be the divisor class group of CC. Then an equivalent definition for r(C)r_{\ell}(C) is that it is the largest power of 𝐙/𝐙\mathbf{Z}/\ell\mathbf{Z} that occurs as a subgroup of the divisor class group of the function field 𝐅q(f,x)\mathbf{F}_{q}(\sqrt[\ell]{f},x).

Remark 1.2.

The above definitions are valid also for =2\ell=2, but in this case we always have r2(C)=0r_{2}(C)=0, because a hyperelliptic curve y2=f(x)y^{2}=f(x) has no 𝐅q\mathbf{F}_{q}-rational 22-torsion in its Jacobian when ff is irreducible.

Let γ=ord(q)\gamma=\textnormal{ord}_{\ell}(q) be the multiplicative order of qmodq\bmod\ell in (𝐙/𝐙)×(\mathbf{Z}/\ell\mathbf{Z})^{\times}, that is, the smallest positive integer such that qγ1modq^{\gamma}\equiv 1\bmod\ell. This is an important invariant for this problem because 𝐅qγ\mathbf{F}_{q^{\gamma}} is the smallest extension of 𝐅q\mathbf{F}_{q} containing \ell-th roots of unity, and hence the Galois closure of 𝐅q(f,x)/𝐅q(x)\mathbf{F}_{q}(\sqrt[\ell]{f},x)/\mathbf{F}_{q}(x) is a degree γ\gamma extension field, namely 𝐅qγ(f,x)\mathbf{F}_{q^{\gamma}}(\sqrt[\ell]{f},x).

Theorem 1.3.

Set

B:=(gcd(d,γ)1)1γ.B:=(\gcd(d,\gamma)-1)\frac{\ell-1}{\gamma}.

Then the \ell-rank r(C)r_{\ell}(C) satisfies min{B,1}r(C)B\min\{B,1\}\leq r_{\ell}(C)\leq B and r(C)Bmod2r_{\ell}(C)\equiv B\bmod 2.

The parity constraint r(C)Bmod2r_{\ell}(C)\equiv B\bmod 2 is proved using the Weil pairing on J[]J[\ell]. This phenomenon does not appear to occur in the analogous situation in number fields, namely ideal class groups of cyclic extensions 𝐐(Np)\mathbf{Q}(\sqrt[p]{N}) for NN prime discussed in [12]; see Section 1.4 for a discussion of the number field case.

Example 1.4.

Consider the case =3\ell=3. If q1mod3q\equiv 1\bmod 3 or if dd is odd, then gcd(d,γ)=1\gcd(d,\gamma)=1, so Theorem 1.3 implies that r(C)=0r_{\ell}(C)=0. Otherwise, if degf\deg f is even and q2mod3q\equiv 2\bmod 3, we have r3(C)=1r_{3}(C)=1, and we recover Theorem 1.1 for =3\ell=3. Compare [12, Theorem 6.1.1] which addresses extensions 𝐐(N3)/𝐐\mathbf{Q}(\sqrt[3]{N})/\mathbf{Q} for prime N1mod3N\equiv 1\bmod 3.

In one special case, we can prove a lower bound that equals the upper bound in Theorem 1.3, allowing us to construct families of curves with large \ell-torsion subgroups in their divisor class groups.

Proposition 1.5.

If gcd(d,γ)=2\gcd(d,\gamma)=2, then r(C)=1γr_{\ell}(C)=\frac{\ell-1}{\gamma}.

Theorem 1.1 follows immediately from Theorem 1.3 and Proposition 1.5. Both Theorem 1.3 and Proposition 1.5 can be proven with linear algebra, using linear maps on J[]J[\ell] defined using endomorphisms of JJ. The parity constraint r(C)Bmod2r_{\ell}(C)\equiv B\bmod 2 is proved using the Weil pairing on J[]J[\ell]. These topics are summarized in Section 1.2 and discussed in depth in Section 2 and Section 3. The proofs of Theorem 1.3 and Proposition 1.5 are then completed in Section 7.

If γ2\gamma\leq 2 then r(C)r_{\ell}(C) is completely determined by Theorem 1.1, so for the remainder of this section we assume γ3\gamma\geq 3. We can compute more refined bounds on r(C)r_{\ell}(C) if we additionally assume γd\gamma\mid d. This constraint ensures that ff totally splits in the extension 𝐅qγ(x)/𝐅q(x)\mathbf{F}_{q^{\gamma}}(x)/\mathbf{F}_{q}(x); this is analogous to the constraint N1modpN\equiv 1\bmod p in [12] which guarantees that NN totally splits in 𝐐(ζp)/𝐐\mathbf{Q}(\operatorname{\zeta}_{p})/\mathbf{Q}. Over 𝐅qγ[x]\mathbf{F}_{q^{\gamma}}[x], f(x)f(x) splits into γ\gamma irreducible factors, which we label f1(x),f2(x),,fγ(x)f_{1}(x),f_{2}(x),\ldots,f_{\gamma}(x) in such a way that the Frobenius automorphism on 𝐅qγ\mathbf{F}_{q^{\gamma}} sends fi(x)f_{i}(x) to fi+1(x)f_{i+1}(x) for all ii (and fγ(x)f_{\gamma}(x) to f1(x)f_{1}(x)). Set

hn(x)\displaystyle h_{n}(x) :=i=1γfi(x)q(i1)(γn)1,n{2,,γ1},\displaystyle:=\prod_{i=1}^{\gamma}f_{i}(x)^{q^{(i-1)(\gamma-n)}-1},\qquad n\in\{2,\ldots,\gamma-1\},
(1) 𝒯\displaystyle\mathcal{T} :={1}{n{2,,γ1}:hn(x) is an th power in 𝐅qγ[x]/(f1(x))}.\displaystyle:=\{1\}\cup\{n\in\{2,\ldots,\gamma-1\}:h_{n}(x)\text{ is an $\ell^{\text{th}}$ power in }\mathbf{F}_{q^{\gamma}}[x]/(f_{1}(x))\}.

The polynomials hn(x)h_{n}(x) are associated via Kummer theory to certain cyclic degree \ell extensions of 𝐅qγ(x)\mathbf{F}_{q^{\gamma}}(x); see Section 6.3 for more on how these polynomials arise.

Theorem 1.6.

Let f(x)𝐅q[x]f(x)\in\mathbf{F}_{q}[x] be irreducible of degree dd, with dd coprime to \ell and 3γd3\leq\gamma\mid d. Set

B:=|𝒯|1γ.B^{\prime}:=|\mathcal{T}|\frac{\ell-1}{\gamma}.

Then the \ell-rank r(C)r_{\ell}(C) satisfies min{B,2}r(C)B\min\{B^{\prime},2\}\leq r_{\ell}(C)\leq B^{\prime} and r(C)Bmod2r_{\ell}(C)\equiv B^{\prime}\bmod 2.

If in addition γ\gamma is even and 1+γ2𝒯1+\frac{\gamma}{2}\in\mathcal{T}, then r(C)3r_{\ell}(C)\geq 3.

Since |𝒯|γ1=gcd(d,γ)1|\mathcal{T}|\leq\gamma-1=\gcd(d,\gamma)-1 we have BBB^{\prime}\leq B, and from min{B,1}r(C)B\min\{B,1\}\leq r_{\ell}(C)\leq B^{\prime} we can conclude min{B,2}min{B,1}\min\{B^{\prime},2\}\geq\min\{B,1\}. Thus Theorem 1.6 gives both upper and lower bounds that are at least as strong as those in Theorem 1.3.

In addition to the linear algebra on J[]J[\ell] discussed in Section 2 and Section 3, the proof of Theorem 1.6 requires techniques from Kummer theory and Galois cohomology. These techniques are introduced in Section 1.3 and discussed in depth in Sections 4, 5, and 6. The proof of Theorem 1.6 is then completed in Section 7.

In some cases, Theorem 1.3 and Theorem 1.6 are sufficient to determine r(C)r_{\ell}(C) precisely.

Corollary 1.7.

Suppose =5\ell=5. If γ=4d\gamma=4\mid d then r5(C)=Br_{5}(C)=B^{\prime}, and otherwise r5(C)=Br_{5}(C)=B.

Proof.

If γ2\gamma\leq 2 then Theorem 1.1 implies r5(C)=Br_{5}(C)=B, so the only remaining option to consider is γ=4\gamma=4. If gcd(d,γ)=1\gcd(d,\gamma)=1 then B=0B=0, and if gcd(d,γ)=2\gcd(d,\gamma)=2 then B=1B=1; in both cases we must have r5(C)=Br_{5}(C)=B by Theorem 1.3. So we may now assume 4d4\mid d. If 𝒯{1,2,3}\mathcal{T}\neq\{1,2,3\}, then B=|𝒯|B^{\prime}=|\mathcal{T}| is either 11 or 22. In either case min{B,2}=B\min\{B^{\prime},2\}=B^{\prime}, so r5(C)=Br_{5}(C)=B^{\prime} by Theorem 1.6. If 𝒯={1,2,3}\mathcal{T}=\{1,2,3\}, then r5(C)B=3r_{5}(C)\leq B^{\prime}=3, but we also have 1+γ2=3𝒯1+\frac{\gamma}{2}=3\in\mathcal{T} and so r5(C)3r_{5}(C)\geq 3, so again r5(C)=Br_{5}(C)=B^{\prime} by Theorem 1.6. ∎

For larger values of \ell, Theorem 1.3 and Theorem 1.6 are not sufficient to determine r(C)r_{\ell}(C). For example, we have the following options when =7\ell=7:

r7(C)={2 or 4,if γ=3d and |𝒯|=2,3 or 5,if γ=6d and |𝒯|=5,B, if 3γd but not the above cases,B, otherwise.r_{7}(C)=\left\{\begin{array}[]{ll}2\text{ or }4,&\text{if }\gamma=3\mid d\text{ and }|\mathcal{T}|=2,\\ 3\text{ or }5,&\text{if }\gamma=6\mid d\text{ and }|\mathcal{T}|=5,\\ B^{\prime},&\text{ if }3\leq\gamma\mid d\text{ but not the above cases,}\\ B,&\text{ otherwise.}\end{array}\right.

For the first two rows, we can exhibit curves attaining both possible values of r7(C)r_{7}(C), demonstrating that the parameters ,q,d,|𝒯|\ell,q,d,|\mathcal{T}| are not sufficient to fully determine the value of r(C)r_{\ell}(C) in general. For instance, consider the case =7\ell=7, q=3q=3 (so γ=6\gamma=6), and d=12d=12 from the introduction, summarized in Table 1. We may categorize these function fields further by the sets 𝒯\mathcal{T} associated to each. See Table 2, and note in particular the last two columns, consisting of curves with the same 𝒯\mathcal{T} but different values of r(C)r_{\ell}(C). 111The astute reader may notice in Table 2 that 𝒯{1}\mathcal{T}\setminus\{1\} is closed under n1nmodγn\mapsto 1-n\bmod\gamma. This symmetry does hold in general, following from Theorem 1.11 and Theorem 1.13 below, and can be used to cut down the number of computations needed in order to find the set 𝒯\mathcal{T}.

r7(C)r_{7}(C): 11 33 33 33 55
𝒯\mathcal{T}: {1}\{1\} {1,3,4}\{1,3,4\} {1,2,5}\{1,2,5\} {1,2,3,4,5}\{1,2,3,4,5\} {1,2,3,4,5}\{1,2,3,4,5\}
count: 55525552 852852 810810 178178 1212
Table 2. The number of isomorphism classes of fields 𝐅3(f7,x)\mathbf{F}_{3}(\sqrt[7]{f},x) attaining each possible 77-rank and set 𝒯\mathcal{T}, with f(x)𝐅3[x]f(x)\in\mathbf{F}_{3}[x] irreducible of degree 1212.

1.2. Results from the linear algebra of Frobenius eigenvectors

The most important feature of working with function fields is that we can represent elements of the ideal class group of 𝐅q(f,x)\mathbf{F}_{q}(\sqrt[\ell]{f},x) using geometric objects, as described in Section 1.1. This allows us to use morphisms from CC to itself to study r(C)r_{\ell}(C).

The qq-power Frobenius map (x,y)(xq,yq)(x,y)\to(x^{q},y^{q}) on C(𝐅¯q)C(\overline{\mathbf{F}}_{q}) induces a linear operator Frob\operatorname{Frob} on the 𝐅\mathbf{F}_{\ell}-vector space J[]J[\ell]. The eigenspace of eigenvalue 11 for this action is J[](𝐅q)J[\ell](\mathbf{F}_{q}), so r(C)r_{\ell}(C) can be recovered as the dimension of this eigenspace. The primary difficulty we will encounter is that the action of Frob\operatorname{Frob} on J[]J[\ell] is not semi-simple in general. The action of Frobenius on the \ell-adic Tate module J[]J[\ell^{\infty}] is semi-simple, but this property does not descend to the mod \ell reduction. So even though we can determine the full characteristic polynomial of Frob\operatorname{Frob} acting on J[]J[\ell] with relatively little work (Remark 2.10), this is not enough to determine the dimension of any particular eigenspace.

To study r(C)r_{\ell}(C), we use a filtration coming from the automorphism (x,y)(x,ζy)(x,y)\mapsto(x,\operatorname{\zeta}_{\ell}y) on the space J[]J[\ell] denoted as

0=V0V1V2V1=J[]0=V^{0}\subseteq V^{1}\subseteq V^{2}\subseteq\cdots\subseteq V^{\ell-1}=J[\ell]

that is preserved by Frob\operatorname{Frob}, and we consider the intersection of generalized eigenspaces for Frob\operatorname{Frob} with this filtration.

Definition 1.8.

For 1k11\leq k\leq\ell-1 and n𝐙/γ𝐙n\in\mathbf{Z}/\gamma\mathbf{Z}, let FnkF_{n}^{k} denote the set of vVkVk1v\in V^{k}\setminus V^{k-1} for which

(Frobqnk+1)iv=0for some i1.(\operatorname{Frob}-q^{n-k+1})^{i}v=0\qquad\text{for some }i\geq 1.

Note that for n𝐙/γ𝐙n\in\mathbf{Z}/\gamma\mathbf{Z}, multiplication by qnk+1q^{n-k+1} is a well-defined scalar operator on J[]J[\ell] because qγ=1q^{\gamma}=1 in 𝐅\mathbf{F}_{\ell}. We will show that J[](𝐅q)J[\ell](\mathbf{F}_{q}) has a basis formed by taking one true eigenvector of Frob\operatorname{Frob} from each Fk1kF_{k-1}^{k}, whenever such an eigenvector exists. Thus r(C)r_{\ell}(C) is directed related to FnkF_{n}^{k} in the following result.

Theorem 1.9.

The \ell-rank of the divisor class group r(C)r_{\ell}(C) equals the number of values 1k11\leq k\leq\ell-1 such that Fk1kF_{k-1}^{k} contains an eigenvector of Frob\operatorname{Frob}.

For a visual interpretation, see LABEL:fig:lifting_charts. Each of the three figures represents a different possible curve. If the cell with coordinates (n,k)(n,k) is shaded dark gray, this means that there exists a Frob\operatorname{Frob} eigenvector with eigenvalue qnk+1q^{n-k+1} in VkVk1V^{k}\setminus V^{k-1}. The circles correspond to cells (n,k)(n,k) with qnk+1=1q^{n-k+1}=1 in 𝐅\mathbf{F}_{\ell}, so any dark gray circle corresponds to a Frob\operatorname{Frob} eigenvector of eigenvalue 11, that is, an element of J[](𝐅q)J[\ell](\mathbf{F}_{q}). The number of dark gray circles equals to r(C)r_{\ell}(C). See Remark 2.7 for a more thorough guide to reading these diagrams.

We prove several constraints that determine when FnkF_{n}^{k} contains an eigenvector of Frob\operatorname{Frob}. A number of relations follow fairly directly from linear algebra on J[]J[\ell]. We say that FnkF_{n}^{k} is a “rooftop” if FnkF_{n}^{k} has a Frob\operatorname{Frob} eigenvector but there is no k<k1k<k^{\prime}\leq\ell-1 such that FnkF_{n}^{k^{\prime}} has an eigenvector. This notion is justified by statement (b) of the following Proposition. We say a rooftop FnkF_{n}^{k} is “non-maximal” if k1k\neq\ell-1.

Proposition 1.10.

Let n𝐙/γ𝐙n\in\mathbf{Z}/\gamma\mathbf{Z} and 1k11\leq k\leq\ell-1. We have the following:

  1. (a)

    Fn1F_{n}^{1} has a Frob\operatorname{Frob} eigenvector if and only if γdn\gamma\mid dn and γn\gamma\nmid n.

  2. (b)

    If FnkF_{n}^{k} has a Frob\operatorname{Frob} eigenvector, then FnkF_{n}^{k^{\prime}} has a Frob\operatorname{Frob} eigenvector for all 1kk1\leq k^{\prime}\leq k.

  3. (c)

    If FnkF_{n}^{k} is a rooftop, then Fnk1F_{n-k}^{1} has a Frob\operatorname{Frob} eigenvector (that is, γd(nk)\gamma\mid d(n-k) and γ(nk)\gamma\nmid(n-k)).

  4. (d)

    If FnkF_{n}^{k} is a non-maximal rooftop, there is no rooftop of the form Fn+ik+iF_{n+i}^{k+i} for i0i\neq 0 and 1k+i21\leq k+i\leq\ell-2.

These relations are proven in Section 2.5. Part (a) is obtained by constructing an explicit basis of Frob\operatorname{Frob} eigenvectors for V1V^{1}, and part (b) follows from a relation between Frob\operatorname{Frob} and the map used to define the filtration. If FnkF_{n}^{k} is a non-maximal rooftop, we will show that applying (Frobqnk)(\operatorname{Frob}-q^{n-k}) to a vector in Fnk+1F_{n}^{k+1} yields an eigenvector in Fnk1F_{n-k}^{1} (a “Jordan chain” of length 22), giving part (c). Part (d) comes from that if two such Jordan chains exist, a linear combination of the two generalized eigenvectors will produce an eigenvector.

The remaining three constraints Theorem 1.11, Theorem 1.12, and Theorem 1.13 are less straightforward, and form the main technical contributions of this paper. The first two can be proven using the Weil pairing on J[]J[\ell], by proving a numerical relation between pairs of Jordan chains (Lemma 3.3).

Theorem 1.11.

Suppose 1k21\leq k\leq\ell-2 and n𝐙/γ𝐙n\in\mathbf{Z}/\gamma\mathbf{Z}. Then FnkF_{n}^{k} is a rooftop if and only if FknkF_{k-n}^{k} is a rooftop.

When k2nmodγk\equiv 2n\bmod\gamma, the above result is vacuous; however in this case we will see that the numerical relation gives us the following result.

Theorem 1.12.

Suppose 1k21\leq k\leq\ell-2 and n𝐙/γ𝐙n\in\mathbf{Z}/\gamma\mathbf{Z}. If k2nmodγk\equiv 2n\bmod\gamma and kk is even then FnkF_{n}^{k} is not a rooftop.

Both Theorem 1.11 and Theorem 1.12 are proved in Section 3. Theorem 1.11 says that FnkFknkF_{n}^{k}\mapsto F_{k-n}^{k} defines an involution on the set of non-maximal rooftops, and Theorem 1.12 imposes a parity constraint on the fixed points of this involution; these two observations will be used together in Section 7.3 to prove the parity constraint r(C)Bmod2r_{\ell}(C)\equiv B\bmod 2.

1.3. Results from Galois cohomology

Another method we use to study the \ell-rank of J[](𝐅q)J[\ell](\mathbf{F}_{q}) comes from Galois cohomology. This approach is more closely related to the approach used to study the number field version of this problem, where there is no direct analogue of the geometric \ell-torsion subgroup J[](𝐅¯q)J[\ell](\overline{\mathbf{F}}_{q}) (see Section 1.4). In Section 4 and Section 5 we use Kummer theory to relate the two perspectives. The culmination of these sections is Proposition 5.7, which relates the existence of an eigenvector of Frob\operatorname{Frob} in FnkF_{n}^{k} to the existence of a certain cohomology class in a Selmer group associated to a (k+1)(k+1)-dimensional representation of Gal(𝐅q(x)sep/𝐅q(x))\textnormal{Gal}(\mathbf{F}_{q}(x)^{\textnormal{sep}}/\mathbf{F}_{q}(x)).

In Section 6 we show that FnkF_{n}^{k} is a non-maximal rooftop if and only if a certain cup product of cohomology classes does not vanish. The vanishing of this cup product can be determined by a residue field calculation, leading us to the last constraint.

Theorem 1.13.

Suppose 2γd2\leq\gamma\mid d and n𝐙/γ𝐙n\in\mathbf{Z}/\gamma\mathbf{Z}. Then Fn2F_{n}^{2} has a Frob\operatorname{Frob} eigenvector if and only if n𝒯n\in\mathcal{T} (as in Eq. 1).

We note that Proposition 1.10 and Theorem 1.11 can also be proven entirely using this Galois cohomology framework: for Theorem 1.11, instead of the Weil pairing we one can use Poitou-Tate duality. On the other hand, we have not yet found a way to prove Theorem 1.12 using this framework. The key difficulty comes from determining whether a Selmer class associated to a self-dual representation lifts to a Selmer class associated to a higher-dimensional representation. These self-dual representations are quite difficult to work with compared to their non-self-dual counterparts, so our geometric proof of Theorem 1.12 using the Weil pairing illustrates a method that can be used to work with them in the function field setting. On the other hand, we were only able to prove Theorem 1.13 using cohomological techniques. Thus, using both the geometric approach (Sections 23) and the cohomological approach (Sections 46) allows us to prove stronger results than any one approach individually.

Together with Proposition 1.10 and Theorem 1.9, constraints on r(C)r_{\ell}(C) can be obtained by counting arguments, analyzing the various restrictions on pairs (n,k)(n,k). Section 7 contains proofs of some such constraints, including all the results stated in Section 1.1.

1.4. Prior work

The study of \ell-torsion in divisor class groups of superelliptic extensions K(f,x)/K(x)K(\sqrt[\ell]{f},x)/K(x) (for some field KK) has been explored in many other contexts; for some examples [14, 7]. Most of these explorations are largely independent from the content of this paper; for instance, some take K=𝐐K=\mathbf{Q} instead of K=𝐅qK=\mathbf{F}_{q}, and they impose different conditions on \ell and f(x)f(x). Further, these works typically focus on a particular subgroup of the \ell-torsion generated by divisors supported at the ramification locus of f(x)f(x), which is the first stage V1V^{1} in the filtration of J[]J[\ell] discussed in Section 1.2.

In the case of hyperelliptic function fields 𝐅q(f,x)/𝐅q(x)\mathbf{F}_{q}(\sqrt{f},x)/\mathbf{F}_{q}(x), this first stage V1V^{1} contains the entirety of the 22-torsion; Cornelissen uses this to compute the 22-rank of J[2](𝐅q)J[2](\mathbf{F}_{q}) for arbitrary hyperelliptic curves over 𝐅q\mathbf{F}_{q} (allowing f(x)f(x) to be reducible) [4]. However, for >2\ell>2, there is more to the filtration than this first stage, and these deeper filtration stages are one of the primary focuses of this paper. The primary difficulty we face is that unlike the action of Frobenius on V1V^{1}, the action of Frob\operatorname{Frob} on J[]J[\ell] as a whole is not semi-simple. See Remark 2.10 for a discussion.

The aforementioned filtration can be defined using an endomorphism (x,y)(x,ζy)(x,y)\mapsto(x,\operatorname{\zeta}_{\ell}y) on CC, where ζ\operatorname{\zeta}_{\ell} is some th\ell^{\text{th}} root of unity in 𝐅¯q\overline{\mathbf{F}}_{q}. This endomorphism and the filtration it defines were used by Poonen–Schaefer [11] and were further explored by Arul [1].

Other authors have studied the \ell-torsion subgroups of divisor class groups of different kinds of degree \ell extensions of 𝐅q(x)\mathbf{F}_{q}(x). For instance, Wittmann considered degree \ell Galois extensions K/𝐅q(x)K/\mathbf{F}_{q}(x), and studied the Galois module structure of the \ell-torsion in the divisor class group of KK [15].

The question this paper is exploring has a direct analogue in number fields: namely, to study the pp-rank of the ideal class group of 𝐐(N1/p)\mathbf{Q}(N^{1/p}) for distinct primes NN and pp. Several authors have studied this question under the assumption N1modpN\equiv 1\bmod p, which is analogous to the assumption γd\gamma\mid d we make in Theorem 1.6. Using deformations of Galois representations, Calegari–Emerton [2] determined conditions under which the pp-part is cyclic (i.e. pp-rank 11). For instance, one of their results is that if i=1(N1)/2ii\prod_{i=1}^{(N-1)/2}i^{i} is a pp-th power modulo NN, then the pp-rank of Cl(𝐐(N1/p))\textnormal{Cl}(\mathbf{Q}(N^{1/p})) is at least 22 [2, Theorem 1.3(ii)]. These results were generalized by Wake–Wang-Erickson [13, Proposition 11.1.1]; in particular, they interpreted the congruence condition as a cup product on Galois cohomology. The techniques of Wake–Wang-Erickson were used by Karl Schaefer and the third author to prove a full converse of Calegari–Emerton’s result by imposing additional congruence conditions.

The cohomological methods used in Sections 46 of this paper closely follow the work of Schaefer and the third author, using the Galois cohomology framework developed by Wake–Wang-Erickson. In particular, the upper bound in Theorem 1.6 is directly analogous to [12, Theorem 1.1.1]. On the other hand, [12, Table 3] shows that there is no parity constraint on the pp-rank in the number field setting; the parity constraint on r(C)r_{\ell}(C) appears to be a phenomenon unique to function fields.

1.5. Acknowledgements

The bulk of this research was conducted while all three authors held a CRM-ISM postdoctoral fellowship. The first author was partially supported by NSF grant DMS-2302511, and the second author was partially supported by ERC Starting Grant 101076941 (‘Gagarin’).

The authors thank Patrick Allen, Jordan Ellenberg, Jaclyn Lang, Bjorn Poonen, Karl Schaefer, Jacob Stix, Yunqing Tang, Carl Wang-Erickson for conversations that pointed them in helpful directions. The first two authors want to thank the third author for suggesting this project and for introducing them to the technical details of Galois cohomology needed for this paper, and the third author wishes to thank his collaborators for seeing this project through after he left academia.

2. Structure of the \ell-torsion subgroup

In this section, we introduce our setup and prove Theorem 1.9 and Proposition 1.10.

2.1. Notation and Setup

We use the following notation throughout the paper.

  • 3\ell\geq 3 is a prime.

  • qq is a prime power coprime to \ell.

  • γ\gamma is the multiplicative order of qq in (𝐙/𝐙)×(\mathbf{Z}/\ell\mathbf{Z})^{\times}.

  • ζ𝐅qγ\operatorname{\zeta}\in\mathbf{F}_{q^{\gamma}} is a fixed nontrivial th\ell^{\text{th}} root of unity.

  • f(x)𝐅q[x]f(x)\in\mathbf{F}_{q}[x] is an irreducible polynomial with d:=degfd:=\deg f coprime to \ell.

  • C/𝐅qC/\mathbf{F}_{q} is the smooth projective curve with affine equation y=f(x)y^{\ell}=f(x), and C𝐅¯qC_{\overline{\mathbf{F}}_{q}} its base change to 𝐅¯q\overline{\mathbf{F}}_{q}.

  • J/𝐅qJ/\mathbf{F}_{q} is the Jacobian of CC, and J𝐅¯qJ_{\overline{\mathbf{F}}_{q}} its base change to 𝐅¯q\overline{\mathbf{F}}_{q}.

  • For a field extension 𝐅/𝐅q\mathbf{F}/\mathbf{F}_{q}, J(𝐅)J(\mathbf{F}) denotes the 𝐅\mathbf{F}-points of JJ. Elements of J(𝐅¯q)J(\overline{\mathbf{F}}_{q}) can be interpreted as divisors on C𝐅¯qC_{\overline{\mathbf{F}}_{q}} modulo linear equivalence, and if 𝐅/𝐅q\mathbf{F}/\mathbf{F}_{q} is a finite extension, elements of J(𝐅)J(\mathbf{F}) correspond to divisor classes in J(𝐅¯q)J(\overline{\mathbf{F}}_{q}) that are invariant under Gal(𝐅¯q/𝐅)\textnormal{Gal}(\overline{\mathbf{F}}_{q}/\mathbf{F}).

  • J[](𝐅)J[\ell](\mathbf{F}) denotes the \ell-torsion subgroup of J(𝐅)J(\mathbf{F}), and J[]:=J[](𝐅¯q)J[\ell]:=J[\ell](\overline{\mathbf{F}}_{q}) the geometric \ell-torsion group.

  • The \ell-torsion rank of CC is defined to be

    r(C):=dim𝐅(J[](𝐅q)).r_{\ell}(C):=\dim_{\mathbf{F}_{\ell}}(J[\ell](\mathbf{F}_{q})).

We also define two morphisms of C𝐅qγC_{\mathbf{F}_{q^{\gamma}}} by giving their actions on geometric points (x,y)C(𝐅¯q)(x,y)\in C(\overline{\mathbf{F}}_{q}). Using the th\ell^{\text{th}} root of unity ζ𝐅qγ\operatorname{\zeta}\in\mathbf{F}_{q^{\gamma}} chosen above, by abuse of notation we also let ζ:C𝐅qγC𝐅qγ\operatorname{\zeta}:C_{\mathbf{F}_{q^{\gamma}}}\to C_{\mathbf{F}_{q^{\gamma}}} denote the morphism defined by

ζ:(x,y)(x,ζy).\operatorname{\zeta}:(x,y)\mapsto(x,\operatorname{\zeta}y).

We let Frob:C𝐅qγC𝐅qγ\operatorname{Frob}:C_{\mathbf{F}_{q^{\gamma}}}\to C_{\mathbf{F}_{q^{\gamma}}} denote the relative Frobenius map,

Frob:(x,y)(xq,yq).\operatorname{Frob}:(x,y)\mapsto(x^{q},y^{q}).

Note that ζ\operatorname{\zeta} is an automorphism of C𝐅qγC_{\mathbf{F}_{q^{\gamma}}}, while Frob\operatorname{Frob} is a degree qq endomorphism; both act invertibly on C(𝐅¯q)C(\overline{\mathbf{F}}_{q}). On C(𝐅¯q)C(\overline{\mathbf{F}}_{q}), we also have the relation

Frobζ=ζqFrob.\operatorname{Frob}\circ\operatorname{\zeta}=\operatorname{\zeta}^{q}\circ\operatorname{Frob}.

By further abuse of notation, we also let ζ\operatorname{\zeta} and Frob\operatorname{Frob} denote the respective endomorphisms of J𝐅¯qJ_{\overline{\mathbf{F}}_{q}} induced by their namesakes, as well as the induced linear maps on J[]J[\ell] considered as a vector space over 𝐅\mathbf{F}_{\ell}.

Convention 2.1.

When not otherwise specified, an “eigenvector” will refer to an eigenvector of Frob\operatorname{Frob} acting as a linear map on the 𝐅\mathbf{F}_{\ell}–vector space J[]J[\ell], i.e. a nonzero vJ[]v\in J[\ell] satisfying Frobv=cv\operatorname{Frob}v=cv for some c𝐅c\in\mathbf{F}_{\ell}. Likewise, a “generalized eigenvector” will refer to a generalized eigenvector of Frob\operatorname{Frob}, i.e. a nonzero vJ[]v\in J[\ell] satisfying (Frobc)iv=0(\operatorname{Frob}-c)^{i}v=0 for some c𝐅c\in\mathbf{F}_{\ell} and i0i\geq 0).

2.2. The 1ζ1-\operatorname{\zeta} Filtration of J[]J[\ell]

The automorphism ζ\operatorname{\zeta} and endomorphism 1ζ1-\operatorname{\zeta} on J𝐅¯qJ_{\overline{\mathbf{F}}_{q}} were discussed in detail in [1, Section 2.3]. Here we use them to construct a filtration on J[]J[\ell].

Noting that the endomorphism ζ\operatorname{\zeta} is annihilated by the th\ell^{\text{th}} cyclotomic polynomial, we can derive the relation

i=11(1ζi)=\prod_{i=1}^{\ell-1}(1-\operatorname{\zeta}^{i})=\ell

in the endomorphism ring End(J𝐅¯q)\textnormal{End}(J_{\overline{\mathbf{F}}_{q}}). For each 1i11\leq i\leq\ell-1, 1ζi1-\operatorname{\zeta}^{i} is equal to 1ζ1-\operatorname{\zeta} times a unit, and so (1ζ)1(1-\operatorname{\zeta})^{\ell-1} is \ell times a unit. We can conclude that the kernel of (1ζ)1(1-\operatorname{\zeta})^{\ell-1} on J𝐅¯qJ_{\overline{\mathbf{F}}_{q}} is exactly J[]J[\ell].

Define VkV^{k} to be the 𝐅¯q\overline{\mathbf{F}}_{q}-points of ker(1ζ)k\ker\,(1-\operatorname{\zeta})^{k}; these VkV^{k} then give a filtration

0=V0V1V1=J[](𝐅¯q).0=V^{0}\subset V^{1}\subset\ldots\subset V^{\ell-1}=J[\ell](\overline{\mathbf{F}}_{q}).

Note that each subgroup VkV^{k} has the structure of a 𝐅\mathbf{F}_{\ell}-vector space.

Lemma 2.2.

For each k=2,,1k=2,\ldots,\ell-1, the endomorphism 1ζ1-\operatorname{\zeta} induces an isomorphism Vk/Vk1Vk1/Vk2V^{k}/V^{k-1}\to V^{k-1}/V^{k-2} of (d1)(d-1)-dimensional vector spaces over 𝐅\mathbf{F}_{\ell}.

Proof.

We have N𝐐(ζ)/𝐐(1ζ)=N_{\mathbf{Q}(\operatorname{\zeta})/\mathbf{Q}}(1-\operatorname{\zeta})=\ell, [𝐐(ζ):𝐐]=1[\mathbf{Q}(\operatorname{\zeta}):\mathbf{Q}]=\ell-1, and the curve CC has genus g=12(1)(d1)g=\frac{1}{2}(\ell-1)(d-1) by Riemann-Hurwitz. So for all 0k10\leq k\leq\ell-1,

deg(1ζ)k=k(d1)\deg(1-\operatorname{\zeta})^{k}=\ell^{k(d-1)}

by [8, Proposition 12.12]. These endomorphisms are all separable and so ker(1ζ)k\ker\,(1-\operatorname{\zeta})^{k} has k(d1)\ell^{k(d-1)} points in J(𝐅¯q)J(\overline{\mathbf{F}}_{q}). This implies dim𝐅Vk=k(d1)\dim_{\mathbf{F}_{\ell}}V^{k}=k(d-1). For 2k12\leq k\leq\ell-1, the kernel of the map VkVk1/Vk2V^{k}\to V^{k-1}/V^{k-2} induced by 1ζ1-\operatorname{\zeta} is Vk1V^{k-1}, so (1ζ):Vk/Vk1Vk1/Vk2(1-\operatorname{\zeta}):V^{k}/V^{k-1}\to V^{k-1}/V^{k-2} is an isomorphism by dimension considerations. ∎

2.3. A modification of 1ζ1-\operatorname{\zeta}

Recall from Section 2.1 that the relative Frobenius map Frob:J𝐅¯qJ𝐅¯q\operatorname{Frob}:J_{\overline{\mathbf{F}}_{q}}\to J_{\overline{\mathbf{F}}_{q}} is induced by the action (x,y)(xq,yq)(x,y)\mapsto(x^{q},y^{q}) on C(𝐅¯q)C(\overline{\mathbf{F}}_{q}). The maps Frob\operatorname{Frob} and 1ζ1-\operatorname{\zeta} on J𝐅¯qJ_{\overline{\mathbf{F}}_{q}} satisfy the relation

(2) Frob(1ζ)=(1ζq)Frob.\displaystyle\operatorname{Frob}\circ(1-\operatorname{\zeta})=(1-\operatorname{\zeta}^{q})\circ\operatorname{Frob}.

Since 1ζq1-\operatorname{\zeta}^{q} and 1ζ1-\operatorname{\zeta} are associates in End(J𝐅¯q)\textnormal{End}(J_{\overline{\mathbf{F}}_{q}}), this identity shows that the action of Frob\operatorname{Frob} on J[](𝐅¯q)J[\ell](\overline{\mathbf{F}}_{q}) preserves the filtration stages VkV^{k}. However, the automorphism 1ζq1ζ\frac{1-\operatorname{\zeta}^{q}}{1-\operatorname{\zeta}} of J𝐅¯qJ_{\overline{\mathbf{F}}_{q}} does not preserve the generalized eigenspaces of Frob\operatorname{Frob}. To account for this, we introduce a modification of 1ζ1-\operatorname{\zeta} that interacts in a more predictable way with the Frobenius map.

Definition 2.3.

Let ηEnd(J𝐅¯q)\operatorname{\eta}\in\textnormal{End}(J_{\overline{\mathbf{F}}_{q}}) be defined by

η:=i=12i1(1ζ)i,\operatorname{\eta}:=-\sum_{i=1}^{\ell-2}i^{-1}(1-\operatorname{\zeta})^{i},

where i1𝐙i^{-1}\in\mathbf{Z} denotes an inverse of ii modulo \ell.

While the endomorphism η\operatorname{\eta} depends on the choice of inverses mod \ell, the action of η\operatorname{\eta} on J[]J[\ell] is well-defined, independent of the choice of i1i^{-1} for each ii. We have the following two important facts about η\operatorname{\eta}, which both capture the idea that η\operatorname{\eta} behaves like a “logarithm” of ζ\operatorname{\zeta}. The first statement in the following lemma says that η\operatorname{\eta} acts like ζ1\operatorname{\zeta}-1 up to higher-order terms.

Lemma 2.4.

We have V1=kerηJ[]V^{1}=\ker\operatorname{\eta}\cap J[\ell], and for each 2k12\leq k\leq\ell-1, η\operatorname{\eta} and ζ1\operatorname{\zeta}-1 are equal as isomorphisms Vk/Vk1Vk1/Vk2V^{k}/V^{k-1}\to V^{k-1}/V^{k-2}.

Proof.

This follows from Lemma 2.2 and the fact that η+(1ζ)\operatorname{\eta}+(1-\operatorname{\zeta}) is in the ideal generated by \ell and (1ζ)2(1-\operatorname{\zeta})^{2}. ∎

The second statement about η\operatorname{\eta} in the following lemma can be thought of as a linearization of the relation Frobζ=ζqFrob\operatorname{Frob}\circ\operatorname{\zeta}=\operatorname{\zeta}^{q}\circ\operatorname{Frob}. As a note of caution, the following relation does not hold when Frob\operatorname{Frob} and η\operatorname{\eta} are considered as endomorphisms of J𝐅¯qJ_{\overline{\mathbf{F}}_{q}}; we obtain the desired equality only when we restrict to the actions on J[]J[\ell].

Lemma 2.5.

As linear maps on J[]J[\ell],

Frobη=qηFrob.\operatorname{Frob}\circ\operatorname{\eta}=q\operatorname{\eta}\circ\operatorname{Frob}.
Proof.

This can be proven by formal manipulation of polynomials; see Appendix A. ∎

A consequence of this result is that if vJ[]v\in J[\ell] is a (generalized) eigenvector of Frob\operatorname{Frob}, then so is ηv\operatorname{\eta}v.

2.4. Generalized eigenvectors of Frob\operatorname{Frob}

Any vJ[](𝐅q)v\in J[\ell](\mathbf{F}_{q}) lies in some filtration stage vVkVk1v\in V^{k}\setminus V^{k-1}. Then ηk1vV1\operatorname{\eta}^{k-1}v\in V^{1}, and by Lemma 2.5 we have

Frob(ηk1v)=qk1ηk1(Frobv)=qk1ηk1v,\operatorname{Frob}(\operatorname{\eta}^{k-1}v)=q^{k-1}\operatorname{\eta}^{k-1}(\operatorname{Frob}v)=q^{k-1}\operatorname{\eta}^{k-1}v,

so ηk1v\operatorname{\eta}^{k-1}v is an eigenvector of eigenvalue qk1q^{k-1}. So our first goal is to identify which powers of qq arise as eigenvalues of Frob\operatorname{Frob} acting on V1V^{1}.

To start, we find a basis of V1V^{1} which is most suitable for our study of the Frob\operatorname{Frob} action. The dimension of V1V^{1} is d1d-1 for d:=degfd:=\deg f. The following Lemma, which is well-known in the literature (see e.g. the proof of [5, Theorem 1.7]), shows that the action of Frob\operatorname{Frob} on V1V^{1} is diagonalizable over 𝐅¯\overline{\mathbf{F}}_{\ell}.

Lemma 2.6.

The action of Frob\operatorname{Frob} on V1𝐅¯V^{1}\otimes\overline{\mathbf{F}}_{\ell} has a basis

{uβ:β𝐅¯,βd=1,β1}\{u_{\beta}:\beta\in\overline{\mathbf{F}}_{\ell},\;\beta^{d}=1,\;\beta\neq 1\}

where uβu_{\beta} is an eigenvector of Frob\operatorname{Frob} with eigenvalue β\beta.

(Caution: recall that Frob\operatorname{Frob} is induced by the qq-power Frobenius map on J𝐅¯qJ_{\overline{\mathbf{F}}_{q}}, not an \ell-power Frobenius map on 𝐅¯\overline{\mathbf{F}}_{\ell}.)

Proof.

Suppose that f(x)f(x) factors over 𝐅¯q\overline{\mathbf{F}}_{q} as (xx1)(xxd)(x-x_{1})\ldots(x-x_{d}), where Frob(xi)=xi+1\operatorname{Frob}(x_{i})=x_{i+1}. The curve CC has a unique point above 𝐏1(𝐅¯q)\infty\in\mathbf{P}^{1}(\overline{\mathbf{F}}_{q}) by the assumption d\ell\nmid d, which we also call \infty. Define the points Pi:=[(xi,0)][]J(𝐅¯q)P_{i}:=[(x_{i},0)]-[\infty]\in J(\overline{\mathbf{F}}_{q}) for i=1,,di=1,\ldots,d. We have the relation P1+Pd=0P_{1}+\cdots P_{d}=0; by [1, Proposition 2.3.1], the points P1,,Pd1P_{1},\ldots,P_{d-1} form a basis for V1V^{1} (see also [5, Proof of Theorem 1.7]).

Given β𝐅¯\beta\in\overline{\mathbf{F}}_{\ell} satisfying βd=1\beta^{d}=1 and β1\beta\neq 1, set

(3) uβ:=i=1dβiPiV1𝐅¯.\displaystyle u_{\beta}:=\sum_{i=1}^{d}\beta^{-i}P_{i}\in V^{1}\otimes\overline{\mathbf{F}}_{\ell}.

Since β1\beta\neq 1, the coefficients of P1P_{1} and PdP_{d} are distinct and so uβ0u_{\beta}\neq 0. We have

Frobuβ\displaystyle\operatorname{Frob}u_{\beta} =i=1d1βiPi+1+βdP1\displaystyle=\sum_{i=1}^{d-1}\beta^{-i}P_{i+1}+\beta^{-d}P_{1}
=β(i=1d1βi1Pi+1+β1P1)\displaystyle=\beta\left(\sum_{i=1}^{d-1}\beta^{-i-1}P_{i+1}+\beta^{-1}P_{1}\right)
=βuβ.\displaystyle=\beta u_{\beta}.

Thus we have d1d-1 eigenvectors with distinct eigenvalues, so these form a basis for the (d1)(d-1)-dimensional vector space V1𝐅¯V^{1}\otimes\overline{\mathbf{F}}_{\ell}. ∎

For 2k12\leq k\leq\ell-1, an eigenvector in V1V^{1} with eigenvalue qnq^{n} lifts under ηk1\operatorname{\eta}^{k-1} to an eigenvector of Frob\operatorname{Frob} in Vk/Vk1V^{k}/V^{k-1} with eigenvalue qnk+1q^{n-k+1} (by Lemma 2.4 and Lemma 2.5). This lift is a priori only an eigenvector in the quotient space, but we show in Lemma 2.8 that we can always take the lift to be a generalized eigenvector of Frob\operatorname{Frob} acting on J[]J[\ell] using properties of the operator η\operatorname{\eta}.

Recall the definition of the sets FnkF_{n}^{k}.

See 1.8

Namely FnkF_{n}^{k} is the set of generalized eigenvectors of Frob\operatorname{Frob} in the kk-th filtration stage with eigenvalue qnk+1q^{n-k+1}.

Remark 2.7.

By this point we have developed a lot of notation, so it may be helpful to have a picture in mind as we proceed. Examples are provided in LABEL:fig:lifting_charts. To each curve CC, we associate a grid of cells (n,k)(n,k) with n𝐙/γ𝐙n\in\mathbf{Z}/\gamma\mathbf{Z} and k=1,,1k=1,\ldots,\ell-1. Roughly speaking, the shaded cells can be matched bijectively with an independent set of vectors in J[]J[\ell]; rows (indexed by kk) correspond to the filtration stages VkV^{k}; columns (indexed by nn) are η\operatorname{\eta}-invariant subspaces; and each diagonal with nk𝐙/γ𝐙n-k\in\mathbf{Z}/\gamma\mathbf{Z} constant corresponds to a distinct Frob\operatorname{Frob} eigenvalue.

More precisely, we shade the cell with coordinates (n,k)(n,k) light gray if the set FnkF_{n}^{k} is nonempty: that is, if there exists a generalized Frob\operatorname{Frob} eigenvector with eigenvalue qnk+1q^{n-k+1} in VkVk1V^{k}\setminus V^{k-1}. Lemma 2.8(a) says that η\operatorname{\eta} acts on the grid by shifting everything down one cell, taking FnkF_{n}^{k} to Fnk1F_{n}^{k-1} and annihilating the bottom layer k=1k=1. Lemma 2.8(b) tells us that we can also go backwards: any shaded cell has a shaded cell above it. Thus each column is either entirely shaded or entirely empty. Corollary 2.9 tells us precisely which columns are shaded or empty. Cells (n,k)(n,k) and (n,k)(n^{\prime},k^{\prime}) correspond to the same generalized Frob\operatorname{Frob} eigenvector if qnk+1=qnk+1q^{n-k+1}=q^{n^{\prime}-k^{\prime}+1}, or equivalently, if nknkmodγn-k\equiv n^{\prime}-k^{\prime}\bmod\gamma.

In Section 2.5 we will see that true eigenvectors of Frob\operatorname{Frob} form “towers” in these grids, and in Section 3 we will see that the Weil pairing imposes a kind of rotational symmetry on these grids.

Lemma 2.8.

Let n𝐙/γ𝐙n\in\mathbf{Z}/\gamma\mathbf{Z} and 1k11\leq k\leq\ell-1. Suppose wFnkw\in F_{n}^{k}.

  1. (a)

    If k2k\geq 2, then ηwFnk1\operatorname{\eta}w\in F_{n}^{k-1}.

  2. (b)

    If k2k\leq\ell-2, there exists vFnk+1v\in F_{n}^{k+1} with ηv=w\operatorname{\eta}v=w.

Proof.

By Lemma 2.4 we have ηwVk1Vk2\operatorname{\eta}w\in V^{k-1}\setminus V^{k-2}. If i1i\geq 1 is such that (Frobqnk+1)iw=0(\operatorname{Frob}-q^{n-k+1})^{i}w=0, then by Lemma 2.5 we have

(Frobqnk+2)iηw=qiη(Frobqnk+1)iw=0,(\operatorname{Frob}-q^{n-k+2})^{i}\operatorname{\eta}w=q^{i}\operatorname{\eta}(\operatorname{Frob}-q^{n-k+1})^{i}w=0,

proving (a).

To set up the proof of (b), we first note that the action of Frob\operatorname{Frob} on V1V^{1} is semisimple (for instance by recalling from Lemma 2.6 that V1𝐅¯V^{1}\otimes\overline{\mathbf{F}}_{\ell} splits into a direct sum of Frob\operatorname{Frob} eigenspaces). Let UU be the qnkq^{n-k} eigenspace of Frob\operatorname{Frob} acting on V1V^{1} (note that UU may be 0- or 11-dimensional), and let WW be the Frob\operatorname{Frob}-invariant complementary subspace of UU in V1V^{1}.

Now suppose k2k\leq\ell-2 and wFnkw\in F_{n}^{k}. Since η\operatorname{\eta} maps Vk+1V^{k+1} surjectively onto VkV^{k}, there exists vVk+1Vkv\in V^{k+1}\setminus V^{k} with η(v)=w\operatorname{\eta}(v)=w. We have

(4) qiη(Frobqnk)iv=(Frobqnk+1)iηv=0,\displaystyle q^{i}\operatorname{\eta}(\operatorname{Frob}-q^{n-k})^{i}v=(\operatorname{Frob}-q^{n-k+1})^{i}\operatorname{\eta}v=0,

so (Frobqnk)ivkerηJ[]=V1(\operatorname{Frob}-q^{n-k})^{i}v\in\ker\operatorname{\eta}\cap J[\ell]=V^{1}. Write (Frobqnk)iv=u+s(\operatorname{Frob}-q^{n-k})^{i}v=u+s for uUu\in U and sWs\in W. Since WW is preserved by Frob\operatorname{Frob} and does not contain any Frob\operatorname{Frob} eigenvectors of eigenvalue qnkq^{n-k}, there exists zWz\in W such that (Frobqnk)iz=s(\operatorname{Frob}-q^{n-k})^{i}z=s. Therefore

(5) (Frobqnk)i(vz)=u,\displaystyle(\operatorname{Frob}-q^{n-k})^{i}(v-z)=u,

so (Frobqnk)i+1(vz)=0(\operatorname{Frob}-q^{n-k})^{i+1}(v-z)=0. This proves vzFnk+1v-z\in F_{n}^{k+1} and η(vz)=w\operatorname{\eta}(v-z)=w, so (b) holds. ∎

Corollary 2.9.

Let n𝐙/γ𝐙n\in\mathbf{Z}/\gamma\mathbf{Z} and 1k11\leq k\leq\ell-1. Then FnkF_{n}^{k} is nonempty if and only if γdn\gamma\mid dn and γn\gamma\nmid n.

Proof.

By Lemma 2.8, FnkF_{n}^{k} is nonempty if and only if Fn1F_{n}^{1} is nonempty, which holds if and only if qnq^{n} is an eigenvalue of Frob\operatorname{Frob} on V1V^{1}. By Lemma 2.6, this holds if and only if (qn)d=1(q^{n})^{d}=1 and qn1q^{n}\neq 1 in 𝐅¯\overline{\mathbf{F}}_{\ell}. ∎

Remark 2.10.

One can generalize the above discussion to determine a basis for J[]𝐅¯J[\ell]\otimes\overline{\mathbf{F}}_{\ell} consisting of generalized Frob\operatorname{Frob} eigenvectors. More precisely, for all 1k11\leq k\leq\ell-1 and all β𝐅¯\beta\in\overline{\mathbf{F}}_{\ell} with βd=1β\beta^{d}=1\neq\beta, there exists vβkJ[]𝐅¯v_{\beta}^{k}\in J[\ell]\otimes\overline{\mathbf{F}}_{\ell} satisfying the following conditions:

  • vβk(Vk𝐅¯)(Vk1𝐅¯)v_{\beta}^{k}\in(V^{k}\otimes\overline{\mathbf{F}}_{\ell})\setminus(V^{k-1}\otimes\overline{\mathbf{F}}_{\ell}).

  • For k2k\geq 2, η(vβk)=vβk1\operatorname{\eta}(v_{\beta}^{k})=v_{\beta}^{k-1}.

  • vβkv_{\beta}^{k} is a generalized Frob\operatorname{Frob} eigenvector with eigenvalue βq1k\beta q^{1-k}.

Further, any set {vβk}\{v_{\beta}^{k}\} satisfying the above conditions is a basis for J[]𝐅¯J[\ell]\otimes\overline{\mathbf{F}}_{\ell}. We can use this to explicitly determine all the diagonal entries of the Jordan canonical form of Frob\operatorname{Frob} acting on J[]𝐅¯J[\ell]\otimes\overline{\mathbf{F}}_{\ell}, and hence compute the characteristic polynomial of Frob\operatorname{Frob} acting on J[]J[\ell]. Thus the only real obstacle remaining is the failure of Frob\operatorname{Frob} to be diagonalizable over 𝐅¯\overline{\mathbf{F}}_{\ell}: the rest of this paper can be thought of as a study of the Jordan blocks in the Jordan canonical form of Frob\operatorname{Frob}.

Since our interest lies with J[](𝐅q)J[\ell](\mathbf{F}_{q}), we will typically restrict our attention to the subspace of J[]J[\ell] generated by generalized Frob\operatorname{Frob} eigenvectors with eigenvalues equal to a power of qq, that is, the Frob\operatorname{Frob}-invariant and η\operatorname{\eta}-invariant subspace spanned by all the sets FnkF_{n}^{k}.

2.5. Basic counts and lifting results

The primary goal of this section is to determine, given some FnkF_{n}^{k} containing a Frob\operatorname{Frob} eigenvector, whether Fnk+1F_{n}^{k+1} also contains a Frob\operatorname{Frob} eigenvector; that is, whether the property of containing a Frob\operatorname{Frob} eigenvector “lifts” from FnkF_{n}^{k} to Fnk+1F_{n}^{k+1}. Having an understanding of when this lifting occurs will help us to compute r(C)r_{\ell}(C) because of Theorem 1.9, which we recall and prove below.

See 1.9

Proof.

For each 1k11\leq k\leq\ell-1 such that Fk1kF_{k-1}^{k} contains a Frob\operatorname{Frob} eigenvector, let vkv_{k} denote such an eigenvector. We claim that the set of all such vkv_{k} is a basis for J[](𝐅q)J[\ell](\mathbf{F}_{q}). First observe that by definition of Fk1kF_{k-1}^{k}, vkv_{k} has Frob\operatorname{Frob} eigenvalue 11, so vkJ[](𝐅q)v_{k}\in J[\ell](\mathbf{F}_{q}). Further, the set of all vkv_{k} is linearly independent because each lies in a distinct filtration stage. So it just remains to prove that these vectors span J[](𝐅q)J[\ell](\mathbf{F}_{q}).

We will prove by induction on kk that if vJ[](𝐅q)Vkv\in J[\ell](\mathbf{F}_{q})\cap V^{k} then vv is in the span of the eigenvectors viv_{i} with iki\leq k. If k=1k=1, then we must have v=0v=0, as there is no eigenvector of eigenvalue 11 in V1V^{1} by Lemma 2.6. Now let k2k\geq 2. If vVk1v\in V^{k-1} then the result follows by the induction hypothesis, so we can assume vVkVk1v\in V^{k}\setminus V^{k-1}. This means that vFk1kv\in F_{k-1}^{k} is an eigenvector, and so vkv_{k} must be defined. Now vv and vkv_{k} define nonzero elements of the quotient Vk/Vk1V^{k}/V^{k-1}, and the 11-eigenspace of Frob\operatorname{Frob} in this quotient is at most one-dimensional by Lemma 2.4 and Lemma 2.5. So for some c𝐅c\in\mathbf{F}_{\ell} we have vcvkVk1v-cv_{k}\in V^{k-1}. By the induction hypothesis, vcvkv-cv_{k} is a linear combination of viv_{i} for i<ki<k, so vv is a linear combination of viv_{i} for iki\leq k. ∎

In the remainder of this section we will prove Proposition 1.10. We first note the following important fact about the generalized Frob\operatorname{Frob} eigenvector lifts defined in Lemma 2.8(b).

Lemma 2.11.

Let n𝐙/γ𝐙n\in\mathbf{Z}/\gamma\mathbf{Z}, 1k21\leq k\leq\ell-2, and suppose wFnkw\in F_{n}^{k} is a Frob\operatorname{Frob} eigenvector. Exactly one of the following holds:

  • For all vFnk+1v\in F_{n}^{k+1} with ηv=w\operatorname{\eta}v=w, vv is an eigenvector.

  • For all vFnk+1v\in F_{n}^{k+1} with ηv=w\operatorname{\eta}v=w, (Frobqnk)vFnk1(\operatorname{Frob}-q^{n-k})v\in F_{n-k}^{1} is an eigenvector.

Proof.

We first make the following observation: if uJ[]u\in J[\ell] satisfies ηu=0\operatorname{\eta}u=0, and uu is in the generalized Frob\operatorname{Frob} eigenspace with eigenvalue qnkq^{n-k}, then either u=0u=0 or uFnk1u\in F_{n-k}^{1}. If v,vFnk+1v,v^{\prime}\in F_{n}^{k+1} with ηv=ηv=w\operatorname{\eta}v=\operatorname{\eta}v^{\prime}=w, we apply this observation to vvv-v^{\prime} to conclude that the value of (Frobqnk)v(\operatorname{Frob}-q^{n-k})v does not depend on the choice of vv. Since ww is an eigenvector we have

qη(Frobqnk)v=(Frobqnk+1)ηv=0,q\operatorname{\eta}(\operatorname{Frob}-q^{n-k})v=(\operatorname{Frob}-q^{n-k+1})\operatorname{\eta}v=0,

so we apply the same observation to (Frobqnk)v(\operatorname{Frob}-q^{n-k})v to reach the desired conclusion. ∎

Recall that FnkF_{n}^{k} is a rooftop if FnkF_{n}^{k} has a Frob\operatorname{Frob} eigenvector but there is no k<k1k<k^{\prime}\leq\ell-1 for which FnkF_{n}^{k} has a Frob\operatorname{Frob} eigenvector, and that FnkF_{n}^{k} is a non-maximal rooftop if k1k\neq\ell-1. The non-maximal rooftops are exactly the sets FnkF_{n}^{k} where the property of having a Frob\operatorname{Frob} eigenvector fails to lift to Fnk+1F_{n}^{k+1}.

See 1.10

Remark 2.12.

Following from Remark 2.7, we give a brief visual explanation of each of these conditions. We shade a cell dark gray if FnkF_{n}^{k} contains a true Frob\operatorname{Frob} eigenvector. (a) says that in the bottom layer k=1k=1, if a cell is shaded at all then it is shaded dark gray. (b) says that the dark gray cells form “towers:” any cell below a dark gray cell must also be dark gray. (c) and (d) both place limitations on which cells can contain the top cells of towers (i.e. the rooftops). (c) says that if a diagonal intersects the k=1k=1 layer in an empty cell, then the diagonal below it cannot contain any non-maximal rooftops. (d) says that no diagonal can contain two non-maximal rooftops. These constraints place some limitations on the possible “skylines” that can occur as in LABEL:fig:lifting_charts.

Proof of Proposition 1.10.
  1. (a)

    It follows from Corollary 2.9 and that all elements of Fn1F_{n}^{1} are eigenvectors.

  2. (b)

    If vv is an eigenvector in FnkF_{n}^{k}, then ηkkvFnk\operatorname{\eta}^{k-k^{\prime}}v\in F_{n}^{k^{\prime}} by Lemma 2.8, and this is an eigenvector by Lemma 2.5.

  3. (c)

    If FnkF_{n}^{k} is a maximal rooftop (k=1k=\ell-1), then γk\gamma\mid k. Now Fn1F_{n}^{1} has an eigenvector by (b) and so γdn\gamma\mid dn and γn\gamma\nmid n by (a); this implies γd(nk)\gamma\mid d(n-k) and γ(nk)\gamma\nmid(n-k). So again by (a), Fnk1F_{n-k}^{1} has an eigenvector.

    Now suppose FnkF_{n}^{k} is a non-maximal rooftop, so 1k21\leq k\leq\ell-2. By Lemma 2.11, Fnk1F_{n-k}^{1} is nonempty, so by (a), γd(nk)\gamma\mid d(n-k) and γ(nk)\gamma\nmid(n-k).

  4. (d)

    Suppose FnkF_{n}^{k} and Fn+ik+iF_{n+i}^{k+i} are both non-maximal rooftops, where k<k+i2k<k+i\leq\ell-2 (the case 1k+i<k1\leq k+i<k follows by symmetry). By Lemma 2.11, there exist vVk+i+1Vk+iv\in V^{k+i+1}\setminus V^{k+i} and wVk+1Vkw\in V^{k+1}\setminus V^{k} for which both vv and ww map under (Frobqnk)(\operatorname{Frob}-q^{n-k}) to Fnk1F_{n-k}^{1}, the set of nonzero vectors in a one-dimensional eigenspace. In particular, there exists d𝐅×d\in\mathbf{F}_{\ell}^{\times} such that

    (Frobqnk)v=d(Frobqnk)w.(\operatorname{Frob}-q^{n-k})v=d(\operatorname{Frob}-q^{n-k})w.

    Then vdwVk+i+1Vk+iv-dw\in V^{k+i+1}\setminus V^{k+i} is an eigenvector of eigenvalue qnkq^{n-k}, contradicting the assumption that Fn+ik+iF_{n+i}^{k+i} is a rooftop. ∎

The remaining proofs require more setup. Lemma 2.11 tells us that the obstruction to lifting an eigenvector in FnkF_{n}^{k} to an eigenvector in Fnk+1F_{n}^{k+1} is given by an element of Fnk1F_{n-k}^{1}. Our next goal is to establish relations between these obstructions for different values of nn (with the same kk).

3. The Weil pairing

In this section, we will use the Weil pairing to prove Theorem 1.11 and Theorem 1.12. These will then be used in Section 7 to prove Theorem 1.3.

The Weil pairing is a non-degenerate alternating Gal(𝐅¯q/𝐅q)\textnormal{Gal}(\overline{\mathbf{F}}_{q}/\mathbf{F}_{q})-equivariant bilinear form

e:J[]×J[]μ\displaystyle e:J[\ell]\times J[\ell]\to\mu_{\ell}

with the property that

e(f(u),f(v))=e(u,v)degfe(f(u),f(v))=e(u,v)^{\deg f}

for any endomorphism f:JJf:J\to J and u,vJ[]u,v\in J[\ell].

For the purposes of this paper we will only need the existence of a pairing satisfying the properties listed above; in particular we will never need to compute the pairing explicitly. For the definition of the Weil pairing and proofs of the stated properties, see for example [8, Section 16]. Note that what we call ee is obtained by taking what Milne calls eλe_{\ell}^{\lambda} (with λ\lambda the canonical principal polarization of JJ) and restricting to J[]×J[]J[\ell]\times J[\ell]. The alternating property follows from Milne’s Lemma 16.2(e), and the endomorphism property follows from Lemma 16.2(c).

In the remainder of this section we will prove that the Weil pairing interacts with the (1ζ)(1-\operatorname{\zeta}) filtration and with Frobenius in a compatible way.

Lemma 3.1.

Let 1k,k11\leq k,k^{\prime}\leq\ell-1, and let vVkv\in V^{k} and vVkv^{\prime}\in V^{k^{\prime}}.

  1. (a)

    If k+k1k+k^{\prime}\leq\ell-1, then e(v,v)=1e(v,v^{\prime})=1.

  2. (b)

    If k+k+1k+k^{\prime}\leq\ell+1, then

    e(η(v),v)=e(v,η(v))1.e(\operatorname{\eta}(v),v^{\prime})=e(v,\operatorname{\eta}(v^{\prime}))^{-1}.
  3. (c)

    If k+k=k+k^{\prime}=\ell, and vFnkv\in F_{n}^{k} and vFnkv^{\prime}\in F_{n^{\prime}}^{k^{\prime}} for some n,n𝐙/γ𝐙n,n^{\prime}\in\mathbf{Z}/\gamma\mathbf{Z}, then e(v,v)1e(v,v^{\prime})\neq 1 if and only if n+n=0n+n^{\prime}=0.

In light of Lemma 3.1(c), we make the following definition.

Definition 3.2.

We say the sets FnkF_{n}^{k} and FnkF_{-n}^{\ell-k} are dual. If vFnkv\in F_{n}^{k} and vFnkv^{\prime}\in F_{-n}^{\ell-k}, then we say that (v,v)(v,v^{\prime}) form a dual pair. See Fig. 1.

In terms of the visual interpretation as described in Remark 2.7, each dual pair vFnkv\in F_{n}^{k} and vFnkv^{\prime}\in F_{-n}^{\ell-k} corresponds to cells (n,k)(n,k) and (n,k)(-n,\ell-k) that are related by rotating the grid 180180^{\circ}. Point (b) relates the Weil pairing of vectors at cells (n,k1)(n,k-1) and (n,k)(n^{\prime},k^{\prime}) to the Weil pairing of vectors at cells (n,k)(n,k) and (n,k1)(n^{\prime},k^{\prime}-1), moving one cell up and the other one down. This mirroring effect of the Weil pairing will play an important role in what follows.

Proof of Lemma 3.1.

All three statements depend on the following calculation. Let u,vJ[]u,v\in J[\ell]. Since ζ\operatorname{\zeta} is an automorphism of JJ, we have e(ζu,v)=e(u,ζ1v)e(\operatorname{\zeta}u,v)=e(u,\operatorname{\zeta}^{-1}v). Since ζ1=ζ1\operatorname{\zeta}^{-1}=\operatorname{\zeta}^{\ell-1} and the Weil pairing is bilinear, we have

(6) e((1ζ)u,v)=e(u,(1ζ1)v)=e(u,(ζ2++ζ+1)(1ζ)v).\displaystyle e((1-\operatorname{\zeta})u,v)=e(u,(1-\operatorname{\zeta}^{\ell-1})v)=e(u,(\operatorname{\zeta}^{\ell-2}+\cdots+\operatorname{\zeta}+1)(1-\operatorname{\zeta})v).

Further, since ee is alternating, the same relation holds if we swap the entries on both sides.

We begin by proving (a). If k+k1k+k^{\prime}\leq\ell-1, then v=(1ζ)kwv^{\prime}=(1-\operatorname{\zeta})^{k}w^{\prime} for some wVk+kw^{\prime}\in V^{k^{\prime}+k}. Then

e(v,v)=e((ζ2++ζ+1)k(1ζ)kv,w)=e(0,w)=1,e(v,v^{\prime})=e((\operatorname{\zeta}^{\ell-2}+\cdots+\operatorname{\zeta}+1)^{k}(1-\operatorname{\zeta})^{k}v,w^{\prime})=e(0,w^{\prime})=1,

because (1ζ)k(1-\operatorname{\zeta})^{k} annihilates VkV^{k}.

We now prove (b). By definition of η\operatorname{\eta} (Definition 2.3), we can write η(v)=(ζ1)v+w\operatorname{\eta}(v)=(\operatorname{\zeta}-1)v+w for some wVk2w\in V^{k-2}. So

e(η(v),v)=e((ζ1)v,v)e(w,v)=e((1ζ)v,v)1,e(\operatorname{\eta}(v),v^{\prime})=e((\operatorname{\zeta}-1)v,v^{\prime})e(w,v^{\prime})=e((1-\operatorname{\zeta})v,v^{\prime})^{-1},

since e(w,v)=1e(w,v^{\prime})=1 by part (a). By a symmetric argument we have e(v,η(v))=e(v,(1ζ)v)1e(v,\operatorname{\eta}(v^{\prime}))=e(v,(1-\operatorname{\zeta})v^{\prime})^{-1}, so it suffices to show that

e((1ζ)v,v)=e(v,(1ζ)v)1.e((1-\operatorname{\zeta})v,v^{\prime})=e(v,(1-\operatorname{\zeta})v^{\prime})^{-1}.

Now note that ζ2++ζ+1\operatorname{\zeta}^{\ell-2}+\cdots+\operatorname{\zeta}+1 can be written as an integer polynomial in 1ζ1-\operatorname{\zeta} with constant term 1\ell-1. So applying Eq. 6,

e((1ζ)v,v)=e(v,(ζ2++ζ+1)(1ζ)v)=e(v,(1)(1ζ)v+w)e((1-\operatorname{\zeta})v,v^{\prime})=e(v,(\operatorname{\zeta}^{\ell-2}+\cdots+\operatorname{\zeta}+1)(1-\operatorname{\zeta})v^{\prime})=e(v,(\ell-1)(1-\operatorname{\zeta})v^{\prime}+w^{\prime})

for some wVk2w^{\prime}\in V^{k^{\prime}-2}; again by part (a), e(v,w)=1e(v,w^{\prime})=1. Therefore

e((1ζ)v,v)=e(v,(1ζ)v)1=e(v,(1ζ)v)1.e((1-\operatorname{\zeta})v,v^{\prime})=e(v,(1-\operatorname{\zeta})v^{\prime})^{\ell-1}=e(v,(1-\operatorname{\zeta})v^{\prime})^{-1}.

In order to deduce (c) we will prove a more general result. As in the proposition statement, let k+k=k+k^{\prime}=\ell, and vFnkv\in F_{n}^{k}, meaning vVkVk1v\in V^{k}\setminus V^{k-1} and Frobv=qn(k1)v\operatorname{Frob}\ v=q^{n-(k-1)}v as an element of the quotient Vk/Vk1V^{k}/V^{k-1}. But now let V¯k:=Vk𝐅¯\overline{V}^{k^{\prime}}:=V^{k^{\prime}}\otimes\overline{\mathbf{F}}_{\ell} (and similarly for V¯k1\overline{V}^{k^{\prime}-1}), and let uV¯kV¯k1u\in\overline{V}^{k^{\prime}}\setminus\overline{V}^{k^{\prime}-1} be any vector which reduces to an eigenvector of Frob\operatorname{Frob} in the quotient V¯k/V¯k1\overline{V}^{k^{\prime}}/\overline{V}^{k^{\prime}-1}. In particular, the eigenvalue β𝐅¯\beta\in\overline{\mathbf{F}}_{\ell} of uu does not a priori need to be a power of qq. The Weil pairing extends in a natural way to J[]¯:=J[]𝐅¯\overline{J[\ell]}:=J[\ell]\otimes\overline{\mathbf{F}}_{\ell}, and under these weaker assumptions we will show that e(v,u)1e(v,u)\neq 1 if and only if uFnku\in F_{-n}^{k^{\prime}}.

Since Frob\operatorname{Frob} is an endomorphism of degree qq, we have the following for any i0i\geq 0:

e(v,u)qi\displaystyle e(v,u)^{q^{i}} =e(Frobiv,Frobiu)=e(q(n(k1))iv+w,βiu+w)\displaystyle=e(\operatorname{Frob}^{i}v,\operatorname{Frob}^{i}u)=e(q^{(n-(k-1))i}v+w,\beta^{i}u+w^{\prime})

for some wVk1w\in V^{k-1} and wVk1w^{\prime}\in V^{k^{\prime}-1}. By part (a), we can eliminate ww and ww^{\prime}. Thus

e(v,u)qi\displaystyle e(v,u)^{q^{i}} =e(q(n(k1))iv,βiu)=e(v,βiu)q(n(k1))i\displaystyle=e(q^{(n-(k-1))i}v,\beta^{i}u)=e(v,\beta^{i}u)^{q^{(n-(k-1))i}}

so solving for e(v,βiu)e(v,\beta^{i}u) we find

e(v,βiu)=e(v,u)q(kn)i.\displaystyle e(v,\beta^{i}u)=e(v,u)^{q^{(k-n)i}}.

Writing the minimal polynomial of β\beta as

h(x)=adxd++a1x+a0𝐅[x],h(x)=a_{d}x^{d}+\cdots+a_{1}x+a_{0}\in\mathbf{F}_{\ell}[x],

we have

1\displaystyle 1 =e(v,h(β)u)=i=0de(v,βiu)ai=e(v,u)h(qkn).\displaystyle=e(v,h(\beta)u)=\prod_{i=0}^{d}e(v,\beta^{i}u)^{a_{i}}=e(v,u)^{h(q^{k-n})}.

If e(v,u)1e(v,u)\neq 1, we must have h(qkn)=0h(q^{k-n})=0 in 𝐅\mathbf{F}_{\ell}. Since h(x)h(x) is irreducible but has a root in 𝐅\mathbf{F}_{\ell}, it must be linear; hence β=qkn=q(1)+kn\beta=q^{k-n}=q^{(1-\ell)+k-n}, showing that in fact uFnku\in F_{-n}^{\ell-k}.

Conversely, assume vFnkv\in F_{n}^{k} and vFnkv^{\prime}\in F_{-n}^{\ell-k}. First consider the case k=1k=1. By (a), we know e(v,w)=1e(v,w)=1 for all wV2w\in V^{\ell-2}. By the proof of the reverse direction above, we know that for any uJ[]¯u\in\overline{J[\ell]}, if uu reduces to a Frob\operatorname{Frob} eigenvector in J[]¯/V¯2\overline{J[\ell]}/\overline{V}^{\ell-2} with any eigenvalue other than qknq^{k-n}, then e(v,u)=1e(v,u)=1. We also know that J[]¯/V2\overline{J[\ell]}/V^{\ell-2} is spanned by its eigenspaces, which are all one-dimensional. So if e(v,v)=1e(v,v^{\prime})=1, then we would have e(v,w)=1e(v,w)=1 for all ww in a basis of J[]¯\overline{J[\ell]}, contradicting non-degeneracy of ee. Hence e(v,v)1e(v,v^{\prime})\neq 1.

For k1k\geq 1, note that ηk1(v)Fn1\operatorname{\eta}^{k-1}(v)\in F_{n}^{1}, and we can write v=ηk1(u)v^{\prime}=\operatorname{\eta}^{k-1}(u^{\prime}) for some uFn1u^{\prime}\in F_{-n}^{\ell-1}. So applying the k=1k=1 case and part (b),

e(v,v)=e(ηk1(v),u)(1)k11.e(v,v^{\prime})=e(\operatorname{\eta}^{k-1}(v),u^{\prime})^{(-1)^{k-1}}\neq 1.\qed

3.1. Lifting relations with dual pairs

For all n𝐙/γ𝐙n\in\mathbf{Z}/\gamma\mathbf{Z} with γdn\gamma\mid dn and γn\gamma\nmid n, fix once and for all a Frob\operatorname{Frob} eigenvector unFn1u_{n}\in F_{n}^{1} (which exists by Proposition 1.10(a)). Now suppose that FnkF_{n}^{k} has an eigenvector for some n𝐙/γ𝐙n\in\mathbf{Z}/\gamma\mathbf{Z} and 1k21\leq k\leq\ell-2. By Lemma 2.11, there exists vFnk+1v\in F_{n}^{k+1} mapping to this eigenvector under η\operatorname{\eta}, such that either vv is an eigenvector itself, or (Frobqnk)vFnk1(\operatorname{Frob}-q^{n-k})v\in F_{n-k}^{1} is an eigenvector. In other words, there exists a constant c𝐅c\in\mathbf{F}_{\ell} such that

(7) (Frobqnk)v=cunk,(\operatorname{Frob}-q^{n-k})v=cu_{n-k},

and vv is an eigenvector if and only if c=0c=0. In the same way, if FknkF_{k-n}^{k} has an eigenvector then there exists vFknk+1v^{\prime}\in F_{k-n}^{k+1} and d𝐅d\in\mathbf{F}_{\ell} such that

(8) (Frobqn)v=dun.(\operatorname{Frob}-q^{-n})v^{\prime}=du_{-n}.

The key observation behind the following lemma is that if we lift Eq. 8 along powers of η\operatorname{\eta} until the preimage of vv^{\prime} reaches the very top of the filtration, then the lifts of unu_{-n} and vv^{\prime} form dual pairs with the vectors vv and unku_{n-k} appearing in Eq. 7. See Fig. 1 for a summary of this setup following the visual interpretation laid out in Remark 2.7.

Lemma 3.3.

Assume FnkF_{n}^{k} and FknkF_{k-n}^{k} both contain Frob\operatorname{Frob} eigenvectors, and let v,c,v,dv,c,v^{\prime},d be as above. Then there exist v^Fkn1\widehat{v^{\prime}}\in F_{k-n}^{\ell-1} and un^Fn1k\widehat{u_{-n}}\in F_{-n}^{\ell-1-k} such that η2k(v^)=v\operatorname{\eta}^{\ell-2-k}(\widehat{v^{\prime}})=v^{\prime}, η2k(un^)=un\operatorname{\eta}^{\ell-2-k}(\widehat{u_{-n}})=u_{-n}, and

e(v^,unk)cqne(un^,v)dqkn=1.e(\widehat{v^{\prime}},u_{n-k})^{cq^{n}}e(\widehat{u_{-n}},v)^{dq^{k-n}}=1.
vvunku_{n-k}un^\widehat{u_{-n}}v^\widehat{v}^{\prime}unu_{-n}vv^{\prime}nkn-knnn-nknk-n11k+1k+11k\ell-1-k1\ell-1
Figure 1. An illustration of the setup (vectors v,vv,v^{\prime}) and conclusion (v^,un^\widehat{v^{\prime}},\widehat{u_{-n}}) of Lemma 3.3. Vertical arrows point from a vector to its image under η2k\operatorname{\eta}^{\ell-2-k}. A diagonal arrow from ww to uu means that ww is a generalized eigenvector for Frob\operatorname{Frob} with some eigenvalue β\beta and (Frobβ)w(\operatorname{Frob}-\beta)w is in the span of uu. Dashed curves connect dual pairs.

Since (v^,unk)(\widehat{v^{\prime}},u_{n-k}) and (un^,v)(\widehat{u_{-n}},v) are dual pairs, their Weil pairings are both nontrivial by Lemma 3.1(c). This will allow us to make conclusions about the constants cc and dd.

Proof.

By Lemma 2.8, there exist v^Fkn1\widehat{v^{\prime}}\in F_{k-n}^{\ell-1} and un^Fn1k\widehat{u_{-n}}\in F_{-n}^{\ell-1-k} that are preimages of vv^{\prime} and unu_{-n}, respectively, under η2k\operatorname{\eta}^{\ell-2-k}. Then

(Frobqk+2n)v^=qk+2dun^+w(\operatorname{Frob}-q^{k+2-\ell-n})\widehat{v^{\prime}}=q^{k+2-\ell}d\widehat{u_{-n}}+w

for some wV2kw\in V^{\ell-2-k}, which can be checked by showing that both sides have the same image under η2k\operatorname{\eta}^{\ell-2-k}. Therefore,

e(v^,v)\displaystyle e(\widehat{v^{\prime}},v) =e(v^,v)q1\displaystyle=e(\widehat{v^{\prime}},v)^{q^{\ell-1}}
=e(Frobv^,Frobv)q2\displaystyle=e(\operatorname{Frob}\widehat{v^{\prime}},\operatorname{Frob}v)^{q^{\ell-2}}
=e(qk+2nv^+qk+2dun^+w,qnkv+cunk)q2\displaystyle=e(q^{k+2-\ell-n}\widehat{v^{\prime}}+q^{k+2-\ell}d\widehat{u_{-n}}+w,\;q^{n-k}v+cu_{n-k})^{q^{\ell-2}}
=e(qknv^+qkdun^+q2w,qnkv+cunk).\displaystyle=e(q^{k-n}\widehat{v^{\prime}}+q^{k}d\widehat{u_{-n}}+q^{\ell-2}w,\;q^{n-k}v+cu_{n-k}).

We now apply bilinearity. Since v,unkVk+1v,u_{n-k}\in V^{k+1}, we can use Lemma 3.1(a) to eliminate all pairings involving ww, as well as the pairing of un^\widehat{u_{-n}} with unku_{n-k}. We obtain

e(v^,v)=e(v^,v)e(v^,unk)cqkne(un^,v)dqne(\widehat{v^{\prime}},v)=e(\widehat{v^{\prime}},v)e(\widehat{v^{\prime}},u_{n-k})^{cq^{k-n}}e(\widehat{u_{-n}},v)^{dq^{n}}

which implies the desired result. ∎

With this we can now give proofs of Theorem 1.11 and Theorem 1.12.

See 1.11

Proof.

We first prove by induction on kk that if FnkF_{n}^{k} is a non-maximal rooftop, then so is FknkF_{k-n}^{k}. We must first show that FknkF_{k-n}^{k} contains a Frob\operatorname{Frob} eigenvector. Since FnkF_{n}^{k} is a rooftop, we have that Fkn1F_{k-n}^{1} is nonempty by Proposition 1.10(c). For the base case k=1k=1 this already establishes that FknkF_{k-n}^{k} has a Frob\operatorname{Frob} eigenvector. Otherwise, for the sake of contradiction, suppose FknkF_{k-n}^{k^{\prime}} is a rooftop for some 1k<k1\leq k^{\prime}<k. By the induction hypothesis, Fk(kn)k=Fn(kk)k(kk)F_{k^{\prime}-(k-n)}^{k^{\prime}}=F_{n-(k-k^{\prime})}^{k-(k-k^{\prime})} is also a rooftop. But since FnkF_{n}^{k} is a non-maximal rooftop, this contradicts Proposition 1.10(d). Hence FknkF_{k-n}^{k} has an eigenvector.

Now since FnkF_{n}^{k} is a non-maximal rooftop, we can take vFnk+1v\in F_{n}^{k+1} and c𝐅c\in\mathbf{F}_{\ell} as in Eq. 7 with c0c\neq 0. If we assume for the sake of contradiction that FknkF_{k-n}^{k} is not a rooftop, then we can take vFknk+1v^{\prime}\in F_{k-n}^{k+1} and d=0d=0 in Eq. 8. By Lemma 3.3 we have

e(v^,unk)cqne(un^,v)dqkn=1,e(\widehat{v^{\prime}},u_{n-k})^{cq^{n}}e(\widehat{u_{-n}},v)^{dq^{k-n}}=1,

where (v^,w)(\widehat{v^{\prime}},w) and (w^,v)(\widehat{w^{\prime}},v) are dual pairs; in particular, we have e(v^,unk)1e(\widehat{v^{\prime}},u_{n-k})\neq 1 by Lemma 3.1(c). Since cqn0modcq^{n}\not\equiv 0\bmod\ell and d=0d=0 we obtain a contradiction. Hence FknkF_{k-n}^{k} must be a rooftop. This concludes the proof for k2k\leq\ell-2. ∎

See 1.12

Proof.

Note that nknmodγn\equiv k-n\bmod\gamma, so we can take v=vFnk+1v=v^{\prime}\in F_{n}^{k+1}, and c=d𝐅c=d\in\mathbf{F}_{\ell} satisfying Eq. 7 and Eq. 8. By Lemma 3.3, there exist v^Fn1\widehat{v}\in F_{n}^{\ell-1} and unk^Fn1k\widehat{u_{n-k}}\in F_{-n}^{\ell-1-k} such that

e(v^,unk)cqne(unk^,v)cqkn=1,e(\widehat{v},u_{n-k})^{cq^{n}}e(\widehat{u_{n-k}},v)^{cq^{k-n}}=1,

where η2k(v^)=v\operatorname{\eta}^{\ell-2-k}(\widehat{v})=v and η2k(unk^)=unk\operatorname{\eta}^{\ell-2-k}(\widehat{u_{n-k}})=u_{n-k}. Hence, by Lemma 3.1(b) and the fact that ee is alternating,

e(unk^,v)=e(unk,v^)(1)2k=e(v^,unk)(1)k.e(\widehat{u_{n-k}},v)=e(u_{n-k},\widehat{v})^{(-1)^{\ell-2-k}}=e(\widehat{v},u_{n-k})^{(-1)^{k}}.

Since qnqknmodq^{n}\equiv q^{k-n}\bmod\ell we have

e(v^,unk)cqn(1+(1)k)=1.e(\widehat{v},u_{n-k})^{cq^{n}(1+(-1)^{k})}=1.

Since e(v^,w)1e(\widehat{v},w)\neq 1 by Lemma 3.1, and kk is even by assumption, this is only possible if c=0c=0, so that vFnk+1v\in F_{n}^{k+1} is an eigenvector. ∎

For the proofs of Theorem 1.3 and Proposition 1.5, the reader may skip ahead to Section 7. The intervening sections on Galois representations and Galois cohomology are only required for the proof of Theorem 1.6, though they can also be used to provide an alternate cohomological interpretation of some of the preceding results.

4. From Frobenius eigenvectors to Galois representations

The primary goal of this section is to show that the existence of an eigenvector of Frob\operatorname{Frob} in FnkF_{n}^{k} is equivalent to the existence of a certain kk-dimensional representation ψ\psi of the absolute Galois group

G𝐅q(x):=Gal(𝐅q(x)sep/𝐅q(x)),G_{\mathbf{F}_{q}(x)}\colon=\textnormal{Gal}(\mathbf{F}_{q}(x)^{\textnormal{sep}}/\mathbf{F}_{q}(x)),

where 𝐅q(x)sep\mathbf{F}_{q}(x)^{\textnormal{sep}} denotes the separable closure of 𝐅q(x)\mathbf{F}_{q}(x). A precise statement is given in Proposition 4.14. In the following sections we use Galois cohomology to analyze conditions under which such a Galois representation can occur, with the goal of proving Theorem 1.13.

4.1. Automorphisms of function fields

Our first step is to translate our setup to the function field setting. The function field associated to the curve CC is 𝐅q(C)=𝐅q(x,y)\mathbf{F}_{q}(C)=\mathbf{F}_{q}(x,y), where yy satisfies the equation y=f(x)y^{\ell}=f(x).Since \ell is coprime to qq, the extension 𝐅q(C)/𝐅q(x)\mathbf{F}_{q}(C)/\mathbf{F}_{q}(x) is separable. The Galois closure of 𝐅q(C)/𝐅q(x)\mathbf{F}_{q}(C)/\mathbf{F}_{q}(x) is the field 𝐅qγ(C)=𝐅qγ(x,y)\mathbf{F}_{q^{\gamma}}(C)=\mathbf{F}_{q^{\gamma}}(x,y), since 𝐅qγ/𝐅q\mathbf{F}_{q^{\gamma}}/\mathbf{F}_{q} is generated by th\ell^{\text{th}} roots of unity. We define two automorphisms of 𝐅qγ(C)\mathbf{F}_{q^{\gamma}}(C) by their actions on xx, on yy, and on c𝐅qγc\in\mathbf{F}_{q^{\gamma}}:

ξ:x\displaystyle\xi:\qquad x x,\displaystyle\mapsto x, y\displaystyle y ζ1y,\displaystyle\mapsto\operatorname{\zeta}^{-1}y, c\displaystyle c c for c𝐅qγ,\displaystyle\mapsto c\text{ for }c\in\mathbf{F}_{q^{\gamma}},
frob:x\displaystyle\operatorname{frob}:\qquad x x,\displaystyle\mapsto x, y\displaystyle y y,\displaystyle\mapsto y, c\displaystyle c cq for c𝐅qγ.\displaystyle\mapsto c^{q}\text{ for }c\in\mathbf{F}_{q^{\gamma}}.

Both of these automorphisms preserve 𝐅qγ(C)\mathbf{F}_{q^{\gamma}}(C) and fix 𝐅q(x)\mathbf{F}_{q}(x) and so define elements of Gal(𝐅qγ(C)/𝐅q(x))\textnormal{Gal}(\mathbf{F}_{q^{\gamma}}(C)/\mathbf{F}_{q}(x)). More precisely, ξ\xi is a generator of the subgroup Gal(𝐅qγ(C)/𝐅qγ(x))𝐙/𝐙\textnormal{Gal}(\mathbf{F}_{q^{\gamma}}(C)/\mathbf{F}_{q^{\gamma}}(x))\simeq\mathbf{Z}/\ell\mathbf{Z}, and frob\operatorname{frob} maps to a generator of the quotient group Gal(𝐅qγ(x)/𝐅q(x))𝐙/γ𝐙\textnormal{Gal}(\mathbf{F}_{q^{\gamma}}(x)/\mathbf{F}_{q}(x))\simeq\mathbf{Z}/\gamma\mathbf{Z}. As discussed in the following Lemma 4.1, we can pick a section Gal(𝐅qγ(x)/𝐅q(x))Gal(𝐅qγ(C)/𝐅q(x))\textnormal{Gal}(\mathbf{F}_{q^{\gamma}}(x)/\mathbf{F}_{q}(x))\to\textnormal{Gal}(\mathbf{F}_{q^{\gamma}}(C)/\mathbf{F}_{q}(x)) and thus we will use frob\operatorname{frob} to denote both the element in Gal(𝐅qγ(C)/𝐅q(x))\textnormal{Gal}(\mathbf{F}_{q^{\gamma}}(C)/\mathbf{F}_{q}(x)) and its image in Gal(𝐅qγ(x)/𝐅q(x))\textnormal{Gal}(\mathbf{F}_{q^{\gamma}}(x)/\mathbf{F}_{q}(x)). In summary, we have the following structure.

Lemma 4.1.

The Galois group GC:=Gal(𝐅qγ(C)/𝐅q(x))G_{C}:=\textnormal{Gal}(\mathbf{F}_{q^{\gamma}}(C)/\mathbf{F}_{q}(x)) is isomorphic to (𝐙/𝐙)(𝐙/γ𝐙)(\mathbf{Z}/\ell\mathbf{Z})\rtimes(\mathbf{Z}/\gamma\mathbf{Z}), satisfying the exact sequence

1ξGCfrob11\to\langle\xi\rangle\to G_{C}\to\langle\operatorname{frob}\rangle\to 1

and with the semi-direct product structure given by frobξ=ξqfrob\operatorname{frob}\circ\xi=\xi^{q}\circ\operatorname{frob}.

The following diagram summarizes the fields we are considering so far. All pictured extensions are Galois except for 𝐅q(C)/𝐅q(x)\mathbf{F}_{q}(C)/\mathbf{F}_{q}(x).

𝐅qγ(C){\mathbf{F}_{q^{\gamma}}(C)}𝐅q(C){\mathbf{F}_{q}(C)}𝐅qγ(x){\mathbf{F}_{{q^{\gamma}}}(x)}𝐅q(x){\mathbf{F}_{q}(x)}ξ𝐙/𝐙\scriptstyle{\langle\xi\rangle\simeq\mathbf{Z}/\ell\mathbf{Z}}GC\scriptstyle{G_{C}}frob𝐙/γ𝐙\scriptstyle{\langle\operatorname{frob}\rangle\simeq\mathbf{Z}/\gamma\mathbf{Z}}

We briefly discuss how these function field maps relate to the morphisms ζ,Frob:C𝐅qγC𝐅qγ\operatorname{\zeta},\operatorname{Frob}:C_{\mathbf{F}_{q^{\gamma}}}\to C_{\mathbf{F}_{q^{\gamma}}} from Section 2.1. Let ζ\operatorname{\zeta}^{\sharp} and Frob\operatorname{Frob}^{\sharp} denote the respective maps of function fields induced by the maps ζ\operatorname{\zeta} and Frob\operatorname{Frob}. Then we can immediately see that ζ=ξ1\operatorname{\zeta}^{\sharp}=\xi^{-1}. On the other hand, the maps frob\operatorname{frob} and Frob\operatorname{Frob}^{\sharp} are quite different. The arithmetic Frobenius map frob\operatorname{frob} does not fix the base field 𝐅qγ\mathbf{F}_{q^{\gamma}}, and so is not induced by a morphism of 𝐅qγ\mathbf{F}_{q^{\gamma}}-varieties. The relative Frobenius map Frob\operatorname{Frob}^{\sharp} – which is the map from 𝐅qγ(C)\mathbf{F}_{q^{\gamma}}(C) to itself that sends xxqx\mapsto x^{q}, yyqy\mapsto y^{q}, and fixes 𝐅qγ\mathbf{F}_{q^{\gamma}} – is induced by a morphism of 𝐅qγ\mathbf{F}_{q^{\gamma}}-varieties, but is not a field automorphism; the image is a degree qq subfield of 𝐅qγ(C)\mathbf{F}_{q^{\gamma}}(C). The composition Frobfrob\operatorname{Frob}^{\sharp}\circ\operatorname{frob} is the absolute Frobenius map, which acts on any h𝐅qγ(C)h\in\mathbf{F}_{q^{\gamma}}(C) by hhqh\mapsto h^{q}. See [10, Page 93] for more on the decomposition of absolute Frobenius into relative Frobenius and arithmetic Frobenius. These observations give us the following relations for all geometric points PC(𝐅¯q)P\in C(\overline{\mathbf{F}}_{q}) and rational functions h𝐅qγ(C)h\in\mathbf{F}_{q^{\gamma}}(C):

ξ(h)(ζ(P))=h(P),frob(h)(Frob(P))=h(P)q.\xi(h)(\operatorname{\zeta}(P))=h(P),\qquad\operatorname{frob}(h)(\operatorname{Frob}(P))=h(P)^{q}.

As a consequence, on the divisor classes we have

(9) div(ξ(h))=ζ(div(h)),div(frob(h))=Frob(div(h)).\displaystyle\textnormal{div}(\xi(h))=\operatorname{\zeta}(\textnormal{div}(h)),\qquad\textnormal{div}(\operatorname{frob}(h))=\operatorname{Frob}(\textnormal{div}(h)).

4.2. Constructing Galois representations

We first define a few explicit representations that will be used to identify Frob\operatorname{Frob} eigenvectors in FnkF_{n}^{k}. Recall that GC=Gal(𝐅qγ(C)/𝐅q(x))G_{C}=\textnormal{Gal}(\mathbf{F}_{q^{\gamma}}(C)/\mathbf{F}_{q}(x)) is generated by ξ\xi and frob\operatorname{frob}, and there is a natural quotient map G𝐅q(x)GCG_{\mathbf{F}_{q}(x)}\to G_{C}.

Definition 4.2.

Let χ:G𝐅q(x)𝐅\chi:G_{\mathbf{F}_{q}(x)}\to\mathbf{F}_{\ell}^{*} be the representation given by the action on μ𝐅qγ\mu_{\ell}\subset\mathbf{F}_{q^{\gamma}}. That is, χ\chi factors through GCG_{C} where it acts by

χ(frob)=q,χ(ξ)=1.\chi(\operatorname{frob})=q,\qquad\chi(\xi)=1.
Definition 4.3.

Let ρ:G𝐅q(x)GL2(𝐅)\rho:G_{\mathbf{F}_{q}(x)}\to\textnormal{GL}_{2}(\mathbf{F}_{\ell}) be the representation that factors through GCG_{C} where it acts by

ρ(frob)=(q001),ρ(ξ)=(1101)\rho(\operatorname{frob})=\begin{pmatrix}q&0\\ 0&1\end{pmatrix},\qquad\rho(\xi)=\begin{pmatrix}1&1\\ 0&1\end{pmatrix}

for a choice of basis {u,v}𝐅2\{u,v\}\subset\mathbf{F}_{\ell}^{2}.

The span of uu is the unique one-dimensional subrepresentation of ρ\rho, which is isomorphic to χ\chi as can be seen by considering the upper left entry of the matrix form of ρ\rho. Moreover, for any σG𝐅q(x)\sigma\in G_{\mathbf{F}_{q}(x)}, we have ρ(σ)=(χ(σ)b(σ)01)\rho(\sigma)=\begin{pmatrix}\chi(\sigma)&b(\sigma)\\ 0&1\end{pmatrix} for some b(σ)𝐅b(\sigma)\in\mathbf{F}_{\ell}; in particular, bb is a crossed homomorphism G𝐅q(x)𝐅G_{\mathbf{F}_{q}(x)}\to\mathbf{F}_{\ell} representing a class in H1(G𝐅q(x),χ)H^{1}(G_{\mathbf{F}_{q}(x)},\chi). Using the cohomological setup we will introduce in Section 5.1, this is equivalent to saying that ρ\rho is the extension of 11 by χ\chi associated to the function bH1(G𝐅q(x),χ)b\in H^{1}(G_{\mathbf{F}_{q}(x)},\chi).

For 0k10\leq k\leq\ell-1, the kk-th symmetric power of ρ\rho, Symkρ:G𝐅q(x)GLk+1(𝐅)\textnormal{Sym}^{k}\rho:G_{\mathbf{F}_{q}(x)}\to\textnormal{GL}_{k+1}(\mathbf{F}_{\ell}), is defined as follows. Taking the basis {u,v}\{u,v\} for 𝐅2\mathbf{F}_{\ell}^{2} as in Definition 4.3, a basis for 𝐅k+1\mathbf{F}_{\ell}^{k+1} is given by formal monomials ei:=1i!ukivie_{i}:=\frac{1}{i!}u^{k-i}v^{i} for i=0,,ki=0,\ldots,k (note that i!i! is invertible over 𝐅\mathbf{F}_{\ell}). For each σG𝐅q(x)\sigma\in G_{\mathbf{F}_{q}(x)}, define

Symkρ(σ)(ei)=1i!(ρ(σ)(u))i(ρ(σ)(v))j,\displaystyle\textnormal{Sym}^{k}\rho(\sigma)(e_{i})=\frac{1}{i!}(\rho(\sigma)(u))^{i}(\rho(\sigma)(v))^{j},

computed by expanding the right-hand side as a sum of monomials. Then Symkρ\textnormal{Sym}^{k}\rho factors through GCG_{C}, and has the following matrix representation on the generators of GCG_{C} with respect to the basis e0,,eke_{0},\ldots,e_{k}.

(10) Symkρ(frob)=(qk000qk100qk201),Symkρ(ξ)=(1112!1k!111(k1)!11(k2)!1).\textnormal{Sym}^{k}\rho(\operatorname{frob})=\begin{pmatrix}q^{k}&0&0&\cdots&0\\ &q^{k-1}&0&\cdots&0\\ &&q^{k-2}&\cdots&0\\ &&&\ddots&\vdots\\ &&&&1\end{pmatrix},\qquad\qquad\textnormal{Sym}^{k}\rho(\xi)=\begin{pmatrix}1&1&\frac{1}{2!}&\cdots&\frac{1}{k!}\\ &1&1&\cdots&\frac{1}{(k-1)!}\\ &&1&\cdots&\frac{1}{(k-2)!}\\ &&&\ddots&\vdots\\ &&&&1\end{pmatrix}.

The matrix Symkρ(ξ)\textnormal{Sym}^{k}\rho(\xi) is similar to a Jordan block, so there is no nontrivial decomposition of 𝐅k+1\mathbf{F}_{\ell}^{k+1} into a direct sum of two subspaces invariant under Symkρ(ξ)\textnormal{Sym}^{k}\rho(\xi). We can conclude that Symkρ\textnormal{Sym}^{k}\rho is indecomposable for 0k10\leq k\leq\ell-1.

Lemma 4.4.

For any n𝐙/γ𝐙n\in\mathbf{Z}/\gamma\mathbf{Z} and 2k12\leq k\leq\ell-1, there is a short exact sequence of G𝐅q(x)G_{\mathbf{F}_{q}(x)}-representations

(11) 0Symk1ρχn+1Symkρχnχn0.0\to\textnormal{Sym}^{k-1}\rho\otimes\chi^{n+1}\to\textnormal{Sym}^{k}\rho\otimes\chi^{n}\to\chi^{n}\to 0.
Proof.

The matrix representation for Symkρχn\textnormal{Sym}^{k}\rho\otimes\chi^{n} is given by the matrix representation for Symρ\textnormal{Sym}^{\rho} with every entry multiplied by χn(σ)\chi^{n}(\sigma). The span of e0,,ek1e_{0},\ldots,e_{k-1} is a kk-dimensional invariant subspace. The action of Symkρχn\textnormal{Sym}^{k}\rho\otimes\chi^{n} on this subspace given by the k×kk\times k upper left submatrix, giving an explicit isomorphism to Symk1ρχn+1\textnormal{Sym}^{k-1}\rho\otimes\chi^{n+1}. The corresponding quotient representation is given by the entry at the bottom-right, which is isomorphic to χn\chi^{n}. ∎

Under the same assumptions as Lemma 4.4, we also have a short exact sequence of the form

(12) 0χk+nSymkρχnSymk1ρχn0.0\to\chi^{k+n}\to\textnormal{Sym}^{k}\rho\otimes\chi^{n}\to\textnormal{Sym}^{k-1}\rho\otimes\chi^{n}\to 0.

We will also compute the dual representations of the representations defined above. Let VV be a vector space over 𝐅\mathbf{F}_{\ell}, and V:=Hom(V,𝐅)V^{\vee}:=\textnormal{Hom}(V,\mathbf{F}_{\ell}) its dual space, so that there is a natural perfect pairing V×V𝐅V\times V^{\vee}\to\mathbf{F}_{\ell}.

Definition 4.5.

The linear dual (or contragredient) of a representation θ:G𝐅q(x)GL(V)\theta:G_{\mathbf{F}_{q}(x)}\to\textnormal{GL}(V) is a representation θ:G𝐅q(x)GL(V)\theta^{\vee}:G_{\mathbf{F}_{q}(x)}\to\textnormal{GL}(V^{\vee}) characterized by the condition that V×V𝐅V\times V^{\vee}\to\mathbf{F}_{\ell} induces a G𝐅q(x)G_{\mathbf{F}_{q}(x)}-equivariant homomorphism θ×θ1\theta\times\theta^{\vee}\to 1.

The cohomological dual of a Galois representation θ:G𝐅q(x)GL(V)\theta:G_{\mathbf{F}_{q}(x)}\to\textnormal{GL}(V) is defined as

θ:=HomG𝐅q(x)(θ,μ)θχ.\theta^{*}:=\textnormal{Hom}_{G_{\mathbf{F}_{q}(x)}}(\theta,\mu_{\ell})\simeq\theta^{\vee}\otimes\chi.

If we identify the vector space 𝐅n\mathbf{F}_{\ell}^{n} with its dual via the pairing (u,v)uTv(u,v)\mapsto u^{T}v, we can check that for each σG𝐅q(x)\sigma\in G_{\mathbf{F}_{q}(x)}, the matrix representing θ(σ)\theta^{\vee}(\sigma) must be the transpose of the matrix representing θ(σ)1\theta(\sigma)^{-1}, in order to guarantee that (θ(σ)u)T(θ(σ)v)=uTv(\theta^{\vee}(\sigma)u)^{T}(\theta(\sigma)v)=u^{T}v for all u,v𝐅nu,v\in\mathbf{F}_{\ell}^{n}.

Lemma 4.6.

Let 0k10\leq k\leq\ell-1 and n𝐙/γ𝐙n\in\mathbf{Z}/\gamma\mathbf{Z}. The linear dual of Symkρχn\textnormal{Sym}^{k}\rho\otimes\chi^{n} is isomorphic to Symkρχkn\textnormal{Sym}^{k}\rho\otimes\chi^{-k-n}.

Proof.

Since Symkρχn\textnormal{Sym}^{k}\rho\otimes\chi^{n} factors through GCG_{C}, the same must be true of its linear dual, so we can restrict our attention to GCG_{C}. Let θ=Symkρχkn\theta=\textnormal{Sym}^{k}\rho\otimes\chi^{-k-n}. Using the change of basis determined by the matrix

(13) B=(111(1)k),\displaystyle B=\begin{pmatrix}&&&&-1\\ &&&1&\\ &&-1&&\\ &\iddots&&&\\ (-1)^{k}&&&&\\ \end{pmatrix},

we compute

(Bθ(frob)B1)T=(qnkq1nkqn),(Bθ(ξ)B1)T=(1112!(1)kk!11(1)k1(k1)!1(1)k2(k2)!1).\displaystyle(B\theta(\operatorname{frob})B^{-1})^{T}=\begin{pmatrix}q^{-n-k}&&&\\ &q^{1-n-k}&&\\ &&\ddots&\\ &&&q^{-n}\end{pmatrix},\qquad(B\theta(\xi)B^{-1})^{T}=\begin{pmatrix}1&-1&\frac{1}{2!}&\cdots&\frac{(-1)^{k}}{k!}\\ &1&-1&\cdots&\frac{(-1)^{k-1}}{(k-1)!}\\ &&1&\cdots&\frac{(-1)^{k-2}}{(k-2)!}\\ &&&\ddots&\vdots\\ &&&&1\end{pmatrix}.

These are the inverses of (Symkρχn)(frob)(\textnormal{Sym}^{k}\rho\otimes\chi^{n})(\operatorname{frob}) and (Symkρχn)(ξ)(\textnormal{Sym}^{k}\rho\otimes\chi^{n})(\xi), respectively. The former assertion is clear, and the latter uses the observation that for all 0ijk0\leq i\leq j\leq k,

t=ij(1)jt(ti)!(jt)!=1(ji)!t=0ji(1)(jit)={1if j=i,0otherwise.\displaystyle\sum_{t=i}^{j}\frac{(-1)^{j-t}}{(t-i)!(j-t)!}=\frac{1}{(j-i)!}\sum_{t=0}^{j-i}(-1)^{\ell}\binom{j-i}{t}=\left\{\begin{array}[]{ll}1&\text{if }j=i,\\ 0&\text{otherwise.}\end{array}\right.

We can conclude that the change of basis determined by BB takes Symkρχnk\textnormal{Sym}^{k}\rho\otimes\chi^{-n-k} to the linear dual of Symkρχn\textnormal{Sym}^{k}\rho\otimes\chi^{n}. ∎

4.3. J[](𝐅qγ)J[\ell](\mathbf{F}_{q^{\gamma}}) as a Galois representation

Note that any Frob\operatorname{Frob} eigenvector vFnkv\in F_{n}^{k} is automatically in J[](𝐅qγ)J[\ell](\mathbf{F}_{{q^{\gamma}}}) since qγ=1𝐅q^{\gamma}=1\in\mathbf{F}_{\ell}^{*} and

Frobγv=q(nk+1)γv=(qγ)nk+1v=v.\operatorname{Frob}^{\gamma}v=q^{(n-k+1)\gamma}v=(q^{\gamma})^{n-k+1}v=v.

We therefore restrict our attention to the structure of J[](𝐅qγ)J[\ell](\mathbf{F}_{q^{\gamma}}) as a G𝐅q(x)G_{\mathbf{F}_{q}(x)} Galois module.

We will define an action of G𝐅q(x)G_{\mathbf{F}_{q}(x)} on J[](𝐅qγ)J[\ell](\mathbf{F}_{q^{\gamma}}) via Kummer theory. Let HH denote the subgroup of 𝐅qγ(C)×/𝐅qγ(C)×\mathbf{F}_{q^{\gamma}}(C)^{\times}/\mathbf{F}_{q^{\gamma}}(C)^{\times\ell} defined by the property that hHh\in H if and only if div(h)\textnormal{div}(h) is a multiple of \ell in Div(C𝐅qγ)\textnormal{Div}(C_{\mathbf{F}_{q^{\gamma}}}).

Lemma 4.7.

There is a split exact sequence

0μHJ[](𝐅qγ)0.0\to\mu_{\ell}\to H\to J[\ell](\mathbf{F}_{q^{\gamma}})\to 0.
Proof.

Define the map HJ[](𝐅qγ)H\to J[\ell](\mathbf{F}_{q^{\gamma}}) by h1div(h)h\mapsto\frac{1}{\ell}\textnormal{div}(h). If hh is in the kernel of this map then there exists g𝐅qγ(C)×g\in\mathbf{F}_{q^{\gamma}}(C)^{\times} with div(g)=div(h)\ell\textnormal{div}(g)=\textnormal{div}(h), so h=cgh=cg^{\ell} for some constant c𝐅qγ×c\in\mathbf{F}_{q^{\gamma}}^{\times}. This implies that h=ch=c as elements of the quotient group 𝐅qγ(C)×/𝐅qγ(C)×\mathbf{F}_{q^{\gamma}}(C)^{\times}/\mathbf{F}_{q^{\gamma}}(C)^{\times\ell}. Since qγ1modq^{\gamma}\equiv 1\bmod\ell, we have 𝐅qγ×/𝐅qγ×μ\mathbf{F}_{q^{\gamma}}^{\times}/\mathbf{F}_{q^{\gamma}}^{\times\ell}\simeq\mu_{\ell}.

It suffices to define a right inverse for the map h1div(h)h\mapsto\frac{1}{\ell}\textnormal{div}(h). Pick an arbitrary place QQ of 𝐅qγ(C)\mathbf{F}_{q^{\gamma}}(C) with corresponding uniformizer πQ\pi_{Q} and valuation ordQ\textnormal{ord}_{Q}, and define the set of “monic” rational functions

K1:={h𝐅qγ(C)×:(hπQordQ(h))(Q)=1}.K^{1}:=\left\{h\in\mathbf{F}_{q^{\gamma}}(C)^{\times}:\left(\frac{h}{\pi_{Q}^{\textnormal{ord}_{Q}(h)}}\right)(Q)=1\right\}.

For every h𝐅qγ(C)×h\in\mathbf{F}_{q^{\gamma}}(C)^{\times}, there exists α𝐅qγ×\alpha\in\mathbf{F}_{q^{\gamma}}^{\times} such that αhK1\alpha h\in K^{1}. Then K1K^{1} is a subgroup of 𝐅qγ(C)×\mathbf{F}_{q^{\gamma}}(C)^{\times}, and we have an exact sequence

0K1divDiv(C𝐅qγ)J(𝐅qγ)0.0\to K^{1}\xrightarrow{\textnormal{div}}\textnormal{Div}(C_{\mathbf{F}_{q^{\gamma}}})\to J(\mathbf{F}_{q^{\gamma}})\to 0.

Now for any DJ[](𝐅qγ)D\in J[\ell](\mathbf{F}_{q^{\gamma}}), there exists hDK1h_{D}\in K^{1} satisfying div(hD)=D\textnormal{div}(h_{D})=\ell D. Thus hDHh_{D}\in H maps to DD. ∎

The action of G𝐅q(x)G_{\mathbf{F}_{q}(x)} on 𝐅q(x)sep\mathbf{F}_{q}(x)^{\textnormal{sep}} induces actions on μ\mu_{\ell} and HH, and the map μH\mu_{\ell}\to H from Lemma 4.7 is equivariant under these actions. This induces a quotient representation structure on J[](𝐅qγ)J[\ell](\mathbf{F}_{q^{\gamma}}). Explicitly, for each DJ[](𝐅qγ)D\in J[\ell](\mathbf{F}_{q^{\gamma}}), pick a preimage hDHh_{D}\in H. By Eq. 9, we have ξ(hD)=hζ(D)\xi(h_{D})=h_{\operatorname{\zeta}(D)} and frob(hD)=hFrob(D)\operatorname{frob}(h_{D})=h_{\operatorname{Frob}(D)} as elements of H/μH/\mu_{\ell}. Thus the structure of J[](𝐅qγ)J[\ell](\mathbf{F}_{q^{\gamma}}) as a G𝐅q(x)G_{\mathbf{F}_{q}(x)}-representation is determined by stipulating that the map G𝐅q(x)GL(J[](𝐅qγ))G_{\mathbf{F}_{q}(x)}\to\textnormal{GL}(J[\ell](\mathbf{F}_{q^{\gamma}})) factors through GCG_{C}, where it acts by

frobD\displaystyle\operatorname{frob}\cdot D =Frob(D),\displaystyle=\operatorname{Frob}(D),
ξD\displaystyle\xi\cdot D =ζ(D).\displaystyle=\operatorname{\zeta}(D).

Using this setup, we can show that each Frob\operatorname{Frob} eigenvector in FnkF_{n}^{k} generates a G𝐅q(x)G_{\mathbf{F}_{q}(x)}-subrepresentation of J[](𝐅qγ)J[\ell](\mathbf{F}_{q^{\gamma}}) that is isomorphic to one of the representations constructed in the previous section.

Lemma 4.8.

There exists an eigenvector of Frob\operatorname{Frob} in FnkF_{n}^{k} if and only if there exists a G𝐅q(x)G_{\mathbf{F}_{q}(x)}-subrepresentation of J[](𝐅qγ)J[\ell](\mathbf{F}_{q^{\gamma}}) isomorphic to Symk1ρχn+1k\textnormal{Sym}^{k-1}\rho\otimes\chi^{n+1-k}.

Proof.

As was discussed above, the structure of J[](𝐅qγ)J[\ell](\mathbf{F}_{q^{\gamma}}) as a G𝐅q(x)G_{\mathbf{F}_{q}(x)}-representation is determined by the fact that the representation factors through GCG_{C}, where frob\operatorname{frob} acts by Frob\operatorname{Frob} and ξ\xi acts by ζ\operatorname{\zeta}. So to prove the lemma it suffices to consider the actions of Frob\operatorname{Frob} and ζ\operatorname{\zeta}.

Suppose vFnkv\in F_{n}^{k} is a Frob\operatorname{Frob} eigenvector. Since vVkVk1v\in V_{k}\setminus V_{k-1}, the vectors ηk1v,ηk2v,,v\operatorname{\eta}^{k-1}v,\operatorname{\eta}^{k-2}v,\cdots,v are linearly independent and span a kk-dimensional 𝐅\mathbf{F}_{\ell} vector space which we call WW. By Lemma 2.5, the matrix representing Frob\operatorname{Frob} acting on the basis {ηk1v,ηk2v,,v}\{\operatorname{\eta}^{k-1}v,\operatorname{\eta}^{k-2}v,\cdots,v\} is

(14) (qnqn1qnk+1).\begin{pmatrix}q^{n}&&&\\ &q^{n-1}&&\\ &&\ddots&\\ &&&q^{n-k+1}\end{pmatrix}.

Since ηkv=0\operatorname{\eta}^{k}v=0, WW is also stable under the action of η\operatorname{\eta}. By Definition 2.3, the actions of η\operatorname{\eta} and ζ\operatorname{\zeta} satisfy the equation

ζ=1+η+η22!++ηk1(k1)!\operatorname{\zeta}=1+\operatorname{\eta}+\frac{\operatorname{\eta}^{2}}{2!}+\cdots+\frac{\operatorname{\eta}^{k-1}}{(k-1)!}

on WW. The matrix representing ζ\operatorname{\zeta} acting on the basis {ηk1v,ηk2v,,v}\{\operatorname{\eta}^{k-1}v,\operatorname{\eta}^{k-2}v,\cdots,v\} is therefore

(15) (1112!1(k1)!111(k2)!11(k3)!1).\begin{pmatrix}1&1&\frac{1}{2!}&\cdots&\frac{1}{(k-1)!}\\ &1&1&\cdots&\frac{1}{(k-2)!}\\ &&1&\cdots&\frac{1}{(k-3)!}\\ &&&\ddots&\vdots\\ &&&&1\par\end{pmatrix}.

Comparing the two matrices with Equations Eq. 10, we conclude that the G𝐅q(x)G_{\mathbf{F}_{q}(x)} action on WW (with frob\operatorname{frob} acting via Frob\operatorname{Frob} and ξ\xi acting via ζ\operatorname{\zeta}) is isomorphic to Symk1ρχnk+1\textnormal{Sym}^{k-1}\rho\otimes\chi^{n-k+1}.

Conversely, suppose WW is a subgroup of J[](𝐅qγ)J[\ell](\mathbf{F}_{q^{\gamma}}) that is isomorphic as a G𝐅q(x)G_{\mathbf{F}_{q}(x)}-representation to Symk1ρχnk+1\textnormal{Sym}^{k-1}\rho\otimes\chi^{n-k+1}. Let vWv\in W correspond to the vector (0,,0,1)(0,\ldots,0,1). Then we have Frobv=qnk+1v\operatorname{Frob}v=q^{n-k+1}v, (ζ1)k1v0(\operatorname{\zeta}-1)^{k-1}v\neq 0, and (ζ1)kv=0(\operatorname{\zeta}-1)^{k}v=0, establishing that vv is a Frob\operatorname{Frob} eigenvector in FnkF_{n}^{k} as desired. ∎

4.4. Galois representations from unramified extensions

Recall that H𝐅qγ(C)×/𝐅qγ(C)×H\leq\mathbf{F}_{q^{\gamma}}(C)^{\times}/\mathbf{F}_{q^{\gamma}}(C)^{\times\ell} is defined by the property that hHh\in H if div(h)Div(C𝐅qγ)\textnormal{div}(h)\in\ell\ \textnormal{Div}(C_{\mathbf{F}_{q^{\gamma}}}). By Lemma 4.7 HH is a finite group. Given a subgroup ΓH\Gamma\leq H, we define

KΓ:=𝐅qγ(C)(h:hΓ).K_{\Gamma}:=\mathbf{F}_{q^{\gamma}}(C)(\sqrt[\ell]{h}:h\in\Gamma).
Lemma 4.9.

The maximal unramified elementary \ell-extension of 𝐅qγ(C)\mathbf{F}_{q^{\gamma}}(C) is KHK_{H}.

Proof.

By Kummer theory, subgroups of 𝐅qγ(C)×/𝐅qγ(C)×\mathbf{F}_{q^{\gamma}}(C)^{\times}/\mathbf{F}_{q^{\gamma}}(C)^{\times\ell} correspond bijectively with abelian extensions of 𝐅qγ(C)\mathbf{F}_{q^{\gamma}}(C) of exponent \ell by taking \ell-th roots of all elements of the subgroup. Adjoining an \ell-th root of h𝐅qγ(C)×h\in\mathbf{F}_{q^{\gamma}}(C)^{\times} results in an unramified extension if and only if for every place vv of 𝐅qγ(C)\mathbf{F}_{q^{\gamma}}(C), h=uvπvnvh=u_{v}\pi_{v}^{\ell n_{v}} for some unit uv𝒪v×u_{v}\in\mathcal{O}_{v}^{\times} and nv𝐙n_{v}\in\mathbf{Z}, where πv\pi_{v} is a choice of uniformizer. This is equivalent to requiring div(h)\ell\mid\textnormal{div}(h), that is, hHh\in H. ∎

As a consequence of Lemma 4.9, every subextension KΓ/𝐅qγ(C)K_{\Gamma}/\mathbf{F}_{q^{\gamma}}(C) of KH/𝐅qγ(C)K_{H}/\mathbf{F}_{q^{\gamma}}(C) is an abelian extension of 𝐅qγ(C)\mathbf{F}_{q^{\gamma}}(C). Let AΓ:=Gal(KΓ/𝐅qγ(C))A_{\Gamma}:=\textnormal{Gal}(K_{\Gamma}/\mathbf{F}_{q^{\gamma}}(C)). See the following diagram for a summary of the fields involved, and the Galois groups corresponding to some of the extensions.

KΓ{K_{\Gamma}}𝐅qγ(C){\mathbf{F}_{q^{\gamma}}(C)}𝐅q(C){\mathbf{F}_{q}(C)}𝐅qγ(x){\mathbf{F}_{{q^{\gamma}}}(x)}𝐅q(x){\mathbf{F}_{q}(x)}AΓ\scriptstyle{A_{\Gamma}}ξ𝐙/𝐙\scriptstyle{\langle\xi\rangle\simeq\mathbf{Z}/\ell\mathbf{Z}}GC\scriptstyle{G_{C}}frob𝐙/γ𝐙\scriptstyle{\langle\operatorname{frob}\rangle\simeq\mathbf{Z}/\gamma\mathbf{Z}}
Proposition 4.10.

Let ΓH\Gamma\leq H be a G𝐅q(x)G_{\mathbf{F}_{q}(x)}-invariant subgroup. Then we have an isomorphism AΓΓA_{\Gamma}\simeq\Gamma^{*} of GCG_{C}-representations, where Γ=HomG𝐅q(x)(Γ,μ)\Gamma^{*}=\textnormal{Hom}_{G_{\mathbf{F}_{q}(x)}}(\Gamma,\mu_{\ell}) is the cohomological dual of Γ\Gamma defined in Definition 4.5.

Proof.

By Kummer theory, we have an isomorphism of groups

ΓHom(AΓ,μ),\Gamma\simeq\textnormal{Hom}(A_{\Gamma},\mu_{\ell}),

where fΓf\in\Gamma is associated to the homomorphism ττ(f)/f\tau\mapsto\tau(\sqrt[\ell]{f})/\sqrt[\ell]{f} for any choice of \ell-th root of ff. It suffices to check that this isomorphism is G𝐅q(x)G_{\mathbf{F}_{q}(x)}-equivariant. For σG𝐅q(x)\sigma\in G_{\mathbf{F}_{q}(x)}, τAΓ\tau\in A_{\Gamma}, and fΓf\in\Gamma, we have σ(f)=ζtσ(f)\sigma(\sqrt[\ell]{f})=\operatorname{\zeta}^{t}\sqrt[\ell]{\sigma(f)} for some integer tt, and so

σ(τ(f)f)=στσ1(ζtσ(f))ζtσ(f)=στσ1(σ(f))σ(f),\sigma\left(\frac{\tau(\sqrt[\ell]{f})}{\sqrt[\ell]{f}}\right)=\frac{\sigma\tau\sigma^{-1}(\operatorname{\zeta}^{t}\sqrt[\ell]{\sigma(f)})}{\operatorname{\zeta}^{t}\sqrt[\ell]{\sigma(f)}}=\frac{\sigma\tau\sigma^{-1}(\sqrt[\ell]{\sigma(f)})}{\sqrt[\ell]{\sigma(f)}},

the last equality following because στσ1\sigma\tau\sigma^{-1} is in AΓA_{\Gamma} and therefore fixes ζ𝐅qγ\operatorname{\zeta}\in\mathbf{F}_{q^{\gamma}}. ∎

Lemma 4.11.

Let ΓH\Gamma\leq H be a G𝐅q(x)G_{\mathbf{F}_{q}(x)}-invariant subgroup. The map Gal(KΓ/𝐅q(x))GC=Gal(𝐅qγ(C)/𝐅q(x))\textnormal{Gal}(K_{\Gamma}/\mathbf{F}_{q}(x))\to G_{C}=\textnormal{Gal}(\mathbf{F}_{q^{\gamma}}(C)/\mathbf{F}_{q}(x)) has a splitting; equivalently,

Gal(KΓ/𝐅q(x))AΓGC.\textnormal{Gal}(K_{\Gamma}/\mathbf{F}_{q}(x))\simeq A_{\Gamma}\rtimes G_{C}.
Proof.

We follow the proof of Lemma 3.1.3 of [12]. The extension

(16) 1AΓGal(KΓ/𝐅q(x))GC11\to A_{\Gamma}\to\textnormal{Gal}(K_{\Gamma}/\mathbf{F}_{q}(x))\to G_{C}\to 1

determines a cohomology class [Gal(KΓ/𝐅q(x))]H2(GC,AΓ)[\textnormal{Gal}(K_{\Gamma}/\mathbf{F}_{q}(x))]\in H^{2}(G_{C},A_{\Gamma}), and it suffices to determine whether this class is 0. To do this, we consider the subgroup

ξ=Gal(𝐅qγ(C)/𝐅qγ(x))GC,\langle\xi\rangle=\textnormal{Gal}(\mathbf{F}_{q^{\gamma}}(C)/\mathbf{F}_{q^{\gamma}}(x))\leq G_{C},

which determines a restriction map H2(GC,AΓ)H2(ξ,AΓ)H^{2}(G_{C},A_{\Gamma})\to H^{2}(\langle\xi\rangle,A_{\Gamma}). The image of [Gal(KΓ/𝐅q(x))][\textnormal{Gal}(K_{\Gamma}/\mathbf{F}_{q}(x))] under this restriction map corresponds to the sequence

1AΓGal(KΓ/𝐅qγ(x))Gal(𝐅qγ(C)/𝐅qγ(x))1.1\to A_{\Gamma}\to\textnormal{Gal}(K_{\Gamma}/\mathbf{F}_{q^{\gamma}}(x))\to\textnormal{Gal}(\mathbf{F}_{q^{\gamma}}(C)/\mathbf{F}_{q^{\gamma}}(x))\to 1.

We show that this map has a splitting. Let f1f_{1} be a place of KΓK_{\Gamma} lying above ff, the place determined by the irreducible polynomial f(x)f(x). Since ff is totally ramified in 𝐅qγ(C)/𝐅qγ(x)\mathbf{F}_{q^{\gamma}}(C)/\mathbf{F}_{q^{\gamma}}(x), but the extension KΓ/𝐅qγ(C)K_{\Gamma}/\mathbf{F}_{q^{\gamma}}(C) is unramified by Lemma 4.9, the inertia group at f1f_{1} is a copy of 𝐙/𝐙\mathbf{Z}/\ell\mathbf{Z} in Gal(KΓ/𝐅qγ(x))\textnormal{Gal}(K_{\Gamma}/\mathbf{F}_{q^{\gamma}}(x)) that maps isomorphically to Gal(𝐅qγ(C)/𝐅qγ(x))\textnormal{Gal}(\mathbf{F}_{q^{\gamma}}(C)/\mathbf{F}_{q^{\gamma}}(x)), defining a splitting as desired.

Hence [Gal(KΓ/𝐅q(x))][\textnormal{Gal}(K_{\Gamma}/\mathbf{F}_{q}(x))] maps to the zero class under the restriction map. But by the Lyndon-Hoschild-Serre spectral sequence we have an inflation-restriction exact sequence

H2(GC/ξ,AΓξ)H2(GC,AΓ)H2(ξ,AΓ).H^{2}(G_{C}/\langle\xi\rangle,A_{\Gamma}^{\langle\xi\rangle})\to H^{2}(G_{C},A_{\Gamma})\to H^{2}(\langle\xi\rangle,A_{\Gamma}).

Since GC/ξ=Gal(𝐅qγ(x)/𝐅q(x))G_{C}/\langle\xi\rangle=\textnormal{Gal}(\mathbf{F}_{q^{\gamma}}(x)/\mathbf{F}_{q}(x)) has order γ\gamma (coprime to \ell), while AΓA_{\Gamma} is a vector space over 𝐅\mathbf{F}_{\ell}, the first term of this sequence is 0. Hence the restriction map is injective, proving that [Gal(KΓ/𝐅q(x))]=0[\textnormal{Gal}(K_{\Gamma}/\mathbf{F}_{q}(x))]=0 as desired. ∎

Lemma 4.12.

Suppose ΓH\Gamma\leq H is isomorphic as a G𝐅q(x)G_{\mathbf{F}_{q}(x)}-representation to Symkρχn\textnormal{Sym}^{k}\rho\otimes\chi^{n} for k2k\geq 2. Then Γ\Gamma maps isomorphically onto its image in J[](𝐅qγ)J[\ell](\mathbf{F}_{q^{\gamma}}) under the map defined in Lemma 4.7.

Proof.

If Γ\Gamma contains ζμ\operatorname{\zeta}\in\mu_{\ell}, then ζ\operatorname{\zeta} spans a one-dimensional G𝐅q(x)G_{\mathbf{F}_{q}(x)}-subrepresentation of Γ\Gamma isomorphic to χ\chi. The unique one-dimensional subrepresentation of Symkρχn\textnormal{Sym}^{k}\rho\otimes\chi^{n} is χn+k\chi^{n+k}, so we can conclude n=1kn=1-k. Since k2k\geq 2, there is a two-dimensional subrepresentation of Symkρχ1k\textnormal{Sym}^{k}\rho\otimes\chi^{1-k} isomorphic to ρ\rho. Hence there exists hΓh\in\Gamma with frob(h)=h\operatorname{frob}(h)=h and ξ(h)=ζh\xi(h)=\operatorname{\zeta}h as elements of HH. This implies that the map σσ(h)/h\sigma\mapsto\sigma(h)/h is a homomorphism G𝐅q(x)μG_{\mathbf{F}_{q}(x)}\to\mu_{\ell} that factors through GCG_{C} and sends frob1\operatorname{frob}\mapsto 1 and ξζ\xi\mapsto\operatorname{\zeta}. But the map σσ(y1)/y1\sigma\mapsto\sigma(y^{-1})/y^{-1} is identical, so by Kummer theory hh is equal to y1y^{-1} as elements of 𝐅qγ(C)×/𝐅qγ(C)×\mathbf{F}_{q^{\gamma}}(C)^{\times}/\mathbf{F}_{q^{\gamma}}(C)^{\times\ell}. This is a contradiction because div(y)Div(C𝐅qγ)\textnormal{div}(y)\notin\ell\ \textnormal{Div}(C_{\mathbf{F}_{q^{\gamma}}}) and so y1Hy^{-1}\notin H. We can conclude that Γμ={1}\Gamma\cap\mu_{\ell}=\{1\} and so Γ\Gamma maps isomorphically to its image in J[](𝐅qγ)J[\ell](\mathbf{F}_{q^{\gamma}}). ∎

4.5. Relating eigenvectors to Galois representations

As stated in Lemma 4.8, the existence of a Frob\operatorname{Frob} eigenvector in FnkF_{n}^{k} is equivalent to the existence of a G𝐅q(x)G_{\mathbf{F}_{q}(x)}-subrepresentation of J[](𝐅qγ)J[\ell](\mathbf{F}_{q^{\gamma}}) isomorphic to Symk1ρχn+1k\textnormal{Sym}^{k-1}\rho\otimes\chi^{n+1-k}. Recall the representation Symk1ρχn+1k\textnormal{Sym}^{k-1}\rho\otimes\chi^{n+1-k} has kernel Gal(𝐅q(x)sep/𝐅qγ(C))\textnormal{Gal}(\mathbf{F}_{q}(x)^{\textnormal{sep}}/\mathbf{F}_{q^{\gamma}}(C)). We will transform this condition to the existence of a G𝐅q(x)G_{\mathbf{F}_{q}(x)} representation related to a field extension of 𝐅qγ(C))\mathbf{F}_{q^{\gamma}}(C)).

Definition 4.13.

Let G=Gal(L/F)G=\textnormal{Gal}(L/F) for some separable field extension L/FL/F, and θ\theta a representation of GG. The kernel field of θ\theta, denoted KθK^{\theta}, is the fixed field in LL of kerθ\ker\theta.

If θ\theta is a finite-dimensional representation over 𝐅\mathbf{F}_{\ell}, then kerθ\ker\theta is finite index in GG, and so KθK^{\theta} is a finite extension of the base field FF. The kernel field is a Galois extension of the base field FF, and by the first isomorphism theorem, θ\theta descends to a faithful representation of Gal(Kθ/F)\textnormal{Gal}(K^{\theta}/F). In the following statement, we consider the case G=G𝐅q(x)=Gal(𝐅q(x)sep/𝐅q(x))G=G_{\mathbf{F}_{q}(x)}=\textnormal{Gal}(\mathbf{F}_{q}(x)^{\textnormal{sep}}/\mathbf{F}_{q}(x)).

Proposition 4.14.

Let 2k12\leq k\leq\ell-1. There exists an eigenvector of Frob\operatorname{Frob} in FnkF_{n}^{k} if and only if there exists a representation ψ:G𝐅q(x)GLk+1(𝐅)\psi:G_{\mathbf{F}_{q}(x)}\to\textnormal{GL}_{k+1}(\mathbf{F}_{\ell}) satisfying the following conditions:

  1. (a)

    there is an exact sequence of G𝐅q(x)G_{\mathbf{F}_{q}(x)}-representations

    0Symk1ρχ1nψ𝐅0,0\to\textnormal{Sym}^{k-1}\rho\otimes\chi^{1-n}\to\psi\to\mathbf{F}_{\ell}\to 0,

    where 𝐅\mathbf{F}_{\ell} denotes the one-dimensional trivial representation;

  2. (b)

    the kernel field KψK^{\psi} of ψ\psi is an unramified extension of 𝐅qγ(C)\mathbf{F}_{q^{\gamma}}(C) with Gal(Kψ/𝐅qγ(C))𝐅k\textnormal{Gal}(K^{\psi}/\mathbf{F}_{q^{\gamma}}(C))\simeq\mathbf{F}_{\ell}^{k}.

Proof.

Recall GC=Gal(𝐅qγ(C)/𝐅q(x))G_{C}=\textnormal{Gal}(\mathbf{F}_{q^{\gamma}}(C)/\mathbf{F}_{q}(x)). Given a Frobenius eigenvector wFnkw\in F_{n}^{k}, let GCwJ[](𝐅qγ)\langle G_{C}w\rangle\leq J[\ell](\mathbf{F}_{q^{\gamma}}) be the span of the GCG_{C}-orbit of ww, and let ΓH\Gamma\leq H be the image of this subgroup under the map DhDD\mapsto h_{D} from the proof of Lemma 4.7. Then as G𝐅q(x)G_{\mathbf{F}_{q}(x)}-representations we have ΓGCwSymk1ρχnk+1\Gamma\simeq\langle G_{C}w\rangle\simeq\textnormal{Sym}^{k-1}\rho\otimes\chi^{n-k+1} by Lemma 4.8, and hence AΓSymk1ρχ1nA_{\Gamma}\simeq\textnormal{Sym}^{k-1}\rho\otimes\chi^{1-n} by Proposition 4.10 and Lemma 4.6.

We can use the splitting Lemma 4.11 to construct a representation ψ:G𝐅q(x)GLk+1(𝐅)\psi:G_{\mathbf{F}_{q}(x)}\to\textnormal{GL}_{k+1}(\mathbf{F}_{\ell}). We assert that this representation factors through Gal(KΓ/𝐅q(x))\textnormal{Gal}(K_{\Gamma}/\mathbf{F}_{q}(x)) and for an arbitrary element (τ,σ)AΓGC(\tau,\sigma)\in A_{\Gamma}\rtimes G_{C} we define

ψ(τ,σ):=(θ(σ)τ01),\psi(\tau,\sigma):=\begin{pmatrix}\theta(\sigma)&\tau\\ 0&1\end{pmatrix},

where θSymk1ρχ1n\theta\simeq\textnormal{Sym}^{k-1}\rho\otimes\chi^{1-n} is the G𝐅q(x)G_{\mathbf{F}_{q}(x)}-representation on AΓA_{\Gamma}. Since kerψkerθ\ker\psi\leq\ker\theta, the kernel field KψK^{\psi} contains KθK^{\theta}, which is equal to KρK^{\rho} since k2k\geq 2; hence KψK^{\psi} is an extension of 𝐅qγ(C)\mathbf{F}_{q^{\gamma}}(C). By construction, ψ\psi satisfies the desired exact sequence in condition (a). The fact that ψ\psi factors through Gal(KΓ/𝐅q(x))\textnormal{Gal}(K_{\Gamma}/\mathbf{F}_{q}(x)) implies KψK^{\psi} is contained in KΓK_{\Gamma}, which is an unramified extension of 𝐅qγ(C)\mathbf{F}_{q^{\gamma}}(C) by Lemma 4.9. In fact Kψ=KΓK^{\psi}=K_{\Gamma} because ψ\psi acts faithfully on AΓA_{\Gamma}, so Gal(Kψ/𝐅qγ(C))=AΓ𝐅k\textnormal{Gal}(K^{\psi}/\mathbf{F}_{q^{\gamma}}(C))=A_{\Gamma}\simeq\mathbf{F}_{\ell}^{k}.

Conversely, suppose there exists ψ:G𝐅q(x)GLk+1(𝐅)\psi:G_{\mathbf{F}_{q}(x)}\to\textnormal{GL}_{k+1}(\mathbf{F}_{\ell}) with the given properties. Let θ\theta be the subrepresentation isomorphic to Symk1ρχ1n\textnormal{Sym}^{k-1}\rho\otimes\chi^{1-n}. The exact sequence implies that with respect to an appropriate basis, ψ\psi can be written in the form

ψ(σ)=(θ(σ)a(σ)01)\psi(\sigma)=\begin{pmatrix}\theta(\sigma)&a(\sigma)\\ 0&1\end{pmatrix}

for some a(σ)𝐅ka(\sigma)\in\mathbf{F}_{\ell}^{k}. Condition (b) says that Kψ/𝐅qγ(C)K^{\psi}/\mathbf{F}_{q^{\gamma}}(C) is an unramified elementary \ell-extension, so by Lemma 4.9, Kψ=KΓK^{\psi}=K_{\Gamma} for some ΓH\Gamma\leq H.

From the matrix form for ψ\psi we see that for σ,τG𝐅q(x)\sigma,\tau\in G_{\mathbf{F}_{q}(x)} with τ\tau mapping into Gal(KΓ/𝐅qγ(C))\textnormal{Gal}(K_{\Gamma}/\mathbf{F}_{q^{\gamma}}(C)), we have a(στσ1)=θ(σ)a(τ)a(\sigma\tau\sigma^{-1})=\theta(\sigma)a(\tau), so Gal(KΓ/𝐅qγ(C))\textnormal{Gal}(K_{\Gamma}/\mathbf{F}_{q^{\gamma}}(C)) is isomorphic to θ\theta as a G𝐅q(x)G_{\mathbf{F}_{q}(x)}-representation. So by Proposition 4.10, Γ\Gamma is isomorphic as a G𝐅q(x)G_{\mathbf{F}_{q}(x)}-representation to Symk1ρχnk+1\textnormal{Sym}^{k-1}\rho\otimes\chi^{n-k+1}, which maps isomorphically to a subrepresentation of J[](𝐅qγ)J[\ell](\mathbf{F}_{q^{\gamma}}) by Lemma 4.12. The structure of Symk1ρχnk+1\textnormal{Sym}^{k-1}\rho\otimes\chi^{n-k+1} implies existence of a vector vWv\in W such that Frob(v)=qnk+1v\operatorname{Frob}(v)=q^{n-k+1}v and GCvG_{C}\cdot v spans WW; by considering the dimension of WW we can conclude vVkVk1v\in V_{k}\setminus V_{k-1} and therefore vFnkv\in F_{n}^{k}. ∎

In the next section we will relate the existence of this representation ψ\psi to the existence of a certain cohomology class in H1(G𝐅q(x),AΓ)H1(G𝐅q(x),Symk1ρχ1n)H^{1}(G_{\mathbf{F}_{q}(x)},A_{\Gamma})\simeq H^{1}(G_{\mathbf{F}_{q}(x)},\textnormal{Sym}^{k-1}\rho\otimes\chi^{1-n}).

Remark 4.15.

A version of Proposition 4.14 holds also for k=1k=1, but this requires a different set of conditions on ψ\psi. First, the fixed field of the kernel of Symk1ρχ1n=χ1n\textnormal{Sym}^{k-1}\rho\otimes\chi^{1-n}=\chi^{1-n} is not 𝐅qγ(C)\mathbf{F}_{q^{\gamma}}(C), but rather some subfield of 𝐅qγ(x)\mathbf{F}_{q^{\gamma}}(x) depending on the value of nn; thus the fixed field KψK^{\psi} of kerψ\ker\psi may not be an extension of 𝐅qγ(C)\mathbf{F}_{q^{\gamma}}(C). We must replace condition (b) with the condition that Kψ𝐅qγ(C)/𝐅qγ(C)K^{\psi}\cdot\mathbf{F}_{q^{\gamma}}(C)/\mathbf{F}_{q^{\gamma}}(C) is unramified, and then we may continue the proof as above but with KΓ:=Kψ𝐅qγ(C)K_{\Gamma}:=K^{\psi}\cdot\mathbf{F}_{q^{\gamma}}(C). Second, if n1modγn\equiv 1\bmod\gamma then there may exist a representation ψ:G𝐅q(x)GL2(𝐅)\psi:G_{\mathbf{F}_{q}(x)}\to\textnormal{GL}_{2}(\mathbf{F}_{\ell}) satisfying all the conditions, but for which the fixed field KψK^{\psi} is the degree \ell base field extension 𝐅q(x)\mathbf{F}_{q^{\ell}}(x); then Kψ𝐅qγ(C)K^{\psi}\cdot\mathbf{F}_{q^{\gamma}}(C) does not correspond to any nontrivial subspace of J[](𝐅qγ)J[\ell](\mathbf{F}_{q^{\gamma}}). To obtain the desired equivalence we must impose the condition that KψK^{\psi} is not unramified over 𝐅q(x)\mathbf{F}_{q}(x). While it is possible to keep track of these additional constraints, we consider only k2k\geq 2 for convenience, since we already have a criterion for the existence of a Frob\operatorname{Frob} eigenvector in Fn1F_{n}^{1} by Proposition 1.10(a).

5. Galois cohomology

Throughout this section and the next we will rely on many facts about Galois cohomology of global fields, most of which can be found in Neukirch, Schmidt, and Wingberg [9].

5.1. Cohomology classes and kernel fields

Let GG be a group. Given an nn-dimensional 𝐅\mathbf{F}_{\ell} representation θ:GGLn(𝐅)\theta:G\to\textnormal{GL}_{n}(\mathbf{F}_{\ell}), an element aH1(G,θ)a\in H^{1}(G,\theta) can be represented by a crossed homomorphism α:G𝐅n\alpha:G\to\mathbf{F}_{\ell}^{n} satisfying α(στ)=θ(σ)α(τ)+α(σ)\alpha(\sigma\tau)=\theta(\sigma)\alpha(\tau)+\alpha(\sigma) for σ,τG\sigma,\tau\in G. Any different representative α\alpha^{\prime} for the same cohomology class aa differs from α\alpha by a coboundary. That is, there exists an element v𝐅nv\in\mathbf{F}_{\ell}^{n} such that α(σ)α(σ)=θ(σ)vv\alpha^{\prime}(\sigma)-\alpha(\sigma)=\theta(\sigma)v-v for all σG\sigma\in G.

Definition 5.1.

With a representation θ:GGLn(𝐅)\theta:G\to\textnormal{GL}_{n}(\mathbf{F}_{\ell}) and a crossed homomorphism α:G𝐅n\alpha:G\to\mathbf{F}_{\ell}^{n} as above, define a n+1n+1 dimensional 𝐅\mathbf{F}_{\ell} representation θ[α]:GGLn+1(𝐅)\theta[\alpha]:G\to\textnormal{GL}_{n+1}(\mathbf{F}_{\ell}) by

θ[α](σ):𝐅n×𝐅\displaystyle\theta[\alpha](\sigma):\mathbf{F}_{\ell}^{n}\times\mathbf{F}_{\ell} 𝐅n×𝐅\displaystyle\to\mathbf{F}_{\ell}^{n}\times\mathbf{F}_{\ell}
(v,c)\displaystyle(v,c) (θ(σ)v+cα(σ),c).\displaystyle\mapsto(\theta(\sigma)v+c\alpha(\sigma),\,c).

We say that θ[α]\theta[\alpha] is the extension of 11 by θ\theta associated to α\alpha.

This definition can be summarized using matrix notation:

(21) θ[α]:=(θα\hdashline[2pt/3pt]01).\displaystyle\theta[\alpha]:=\left(\begin{array}[]{ccc;{2pt/3pt}c}&&&\\ &\theta&&\alpha\\ &&&\\ \hdashline[2pt/3pt]&0&&1\end{array}\right).

If α\alpha^{\prime} is a different crossed homomorphism which represents the same class aH1(G,θ)a\in H^{1}(G,\theta), then we have

θ[α]=(I-v\hdashline[2pt/3pt]01)(θα\hdashline[2pt/3pt]01)(Iv\hdashline[2pt/3pt]01)\theta[\alpha^{\prime}]=\left(\begin{array}[]{ccc;{2pt/3pt}c}&&&\\ &I&&-v\\ &&&\\ \hdashline[2pt/3pt]&0&&1\end{array}\right)\left(\begin{array}[]{ccc;{2pt/3pt}c}&&&\\ &\theta&&\alpha\\ &&&\\ \hdashline[2pt/3pt]&0&&1\end{array}\right)\left(\begin{array}[]{ccc;{2pt/3pt}c}&&&\\ &I&&v\\ &&&\\ \hdashline[2pt/3pt]&0&&1\end{array}\right)

where v𝐅nv\in\mathbf{F}_{\ell}^{n} is the vector satisfying α(σ)α(σ)=θ(σ)vv\alpha^{\prime}(\sigma)-\alpha(\sigma)=\theta(\sigma)v-v. Thus, different representatives of a cocycle class give rise to representations that are equivalent up to conjugation. From now on, we will denote this representation class by θ[a]\theta[a] since it only depends on θ\theta and the cocycle class aa.

It is straightforward to check that θ[a]\theta[a] is a GG-representation and fits into an exact sequence

0θθ[a]𝐅00\to\theta\to\theta[a]\to\mathbf{F}_{\ell}\to 0

of GG-representations, where 𝐅\mathbf{F}_{\ell} represents the one-dimensional trivial representation. This exact sequence induces a long exact sequence in cohomology which begins

0H0(G,θ)H0(G,θ[a])H0(G,𝐅)𝛿H1(G,θ),0\to H^{0}(G,\theta)\to H^{0}(G,\theta[a])\to H^{0}(G,\mathbf{F}_{\ell})\xrightarrow{\delta}H^{1}(G,\theta),

and we have a=δ(1)a=\delta(1). Conversely, for any exact sequence 0θκ𝐅00\to\theta\to\kappa\to\mathbf{F}_{\ell}\to 0 of GG-representations there is a class aH1(G,θ)a\in H^{1}(G,\theta) with κθ[a]\kappa\simeq\theta[a].

Recall the definition of the kernel field of a Galois representation in Definition 4.13 and for aH1(G,θ)a\in H^{1}(G,\theta), we will denote the kernel field of θ[a]\theta[a] as KaK^{a}. Since kerθ[a]kerθ\ker\theta[a]\leq\ker\theta, KaK^{a} is necessarily an extension of KθK^{\theta}. One can check that the kernel field is well-defined on cohomology classes (in particular, if aa is a coboundary then Ka=KθK^{a}=K^{\theta}, though the converse does not necessarily hold), and in fact that the kernel field is invariant under scaling:

Lemma 5.2.

If aH1(G,θ)a\in H^{1}(G,\theta) and c𝐅×c\in\mathbf{F}_{\ell}^{\times} then Kca=KaK^{ca}=K^{a}.

5.2. Selmer conditions

Let \mathcal{M} denote the set of all places of 𝐅q(x)\mathbf{F}_{q}(x). For each vv\in\mathcal{M}, let 𝐅q(x)v\mathbf{F}_{q}(x)_{v} denote the localization of 𝐅q(x)\mathbf{F}_{q}(x) at vv, and pick once and for all an inclusion 𝐅q(x)sep𝐅q(x)vsep\mathbf{F}_{q}(x)^{\textnormal{sep}}\hookrightarrow\mathbf{F}_{q}(x)_{v}^{\textnormal{sep}} of separable closures. This is equivalent to picking a prime above vv in 𝐅q(x)sep\mathbf{F}_{q}(x)^{\textnormal{sep}}, or equivalently a compatible system of one place above vv in each finite extension of 𝐅q(x)\mathbf{F}_{q}(x). We define

G𝐅q(x)v\displaystyle G_{\mathbf{F}_{q}(x)_{v}} :=Gal(𝐅q(x)vsep/𝐅q(x)v),\displaystyle:=\textnormal{Gal}(\mathbf{F}_{q}(x)_{v}^{\textnormal{sep}}/\mathbf{F}_{q}(x)_{v}),

and the inclusion 𝐅q(x)sep𝐅q(x)vsep\mathbf{F}_{q}(x)^{\textnormal{sep}}\hookrightarrow\mathbf{F}_{q}(x)_{v}^{\textnormal{sep}} induces an inclusion G𝐅q(x)vG𝐅q(x)G_{\mathbf{F}_{q}(x)_{v}}\hookrightarrow G_{\mathbf{F}_{q}(x)} by restriction. The image of G𝐅q(x)vG_{\mathbf{F}_{q}(x)_{v}} is the decomposition group of the prime above vv in 𝐅q(x)sep\mathbf{F}_{q}(x)^{\textnormal{sep}}.

Given a Galois representation θ:G𝐅q(x)GLn(𝐅)\theta:G_{\mathbf{F}_{q}(x)}\to\textnormal{GL}_{n}(\mathbf{F}_{\ell}), we define θv\theta_{v} to be its restriction to the decomposition group G𝐅q(x)vG_{\mathbf{F}_{q}(x)_{v}}. For convenience we will define the notation

Hi(θ):=Hi(G𝐅q(x),θ)andHi(v,θ):=Hi(G𝐅q(x)v,θv).H^{i}(\theta):=H^{i}(G_{\mathbf{F}_{q}(x)},\theta)\qquad\text{and}\qquad H^{i}(v,\theta):=H^{i}(G_{\mathbf{F}_{q}(x)_{v}},\theta_{v}).

By restriction to the decomposition group G𝐅q(x)vG_{\mathbf{F}_{q}(x)_{v}}, we obtain a map

resv:Hi(θ)Hi(v,θ).\textnormal{res}_{v}:H^{i}(\theta)\to H^{i}(v,\theta).

For any vv\in\mathcal{M}, let kvk_{v} denote the residue field of 𝐅q(x)\mathbf{F}_{q}(x) at the place vv, so that GkvG_{k_{v}} is the quotient of G𝐅q(x)vG_{\mathbf{F}_{q}(x)_{v}} by the inertia group IvI_{v} above vv. Let S={f,}S=\{f,\infty\}\subseteq\mathcal{M}, where ff denotes the place determined by the irreducible polynomial f(x)f(x) defining the curve CC, and \infty is the place determined by 1x\frac{1}{x}.

For each vv\in\mathcal{M} we define a subgroup LvH1(v,θ)L_{v}\subseteq H^{1}(v,\theta):

(22) Lv:={H1(v,θ)vS,H1(Gkv,θ)vS.L_{v}:=\left\{\begin{array}[]{ll}H^{1}(v,\theta)&v\in S,\\ H^{1}(G_{k_{v}},\theta)&v\in\mathcal{M}\setminus S.\end{array}\right.

The subgroup LvL_{v} for vSv\notin S is the “unramified subspace”

H1(Gkv,θ)=ker(H1(G𝐅q(x)v,θ)H1(Iv,θ)),H^{1}(G_{k_{v}},\theta)=\ker(H^{1}(G_{\mathbf{F}_{q}(x)_{v}},\theta)\to H^{1}(I_{v},\theta)),

which is equal to the group defined in [9, Definition 7.2.14] by the inflation-restriction exact sequence. The unramified subspace is so called because kernel fields of classes in the unramified subspace introduce no new ramification at vv: if IvkerθI_{v}\leq\ker\theta (so the kernel field of θ\theta is unramified over 𝐅q(x)\mathbf{F}_{q}(x) at vv), and if aH1(G𝐅q(x),θ)a\in H^{1}(G_{\mathbf{F}_{q}(x)},\theta) satisfies resv(a)H1(Gkv,θ)\textnormal{res}_{v}(a)\in H^{1}(G_{k_{v}},\theta), then in fact we also have Ivkerθ[a]I_{v}\leq\ker\theta[a] (the kernel field of aa is unramified over 𝐅q(x)\mathbf{F}_{q}(x) at vv). With this setup, we can define the Selmer group

HS1(θ)\displaystyle H^{1}_{S}(\theta) :={aH1(θ):resv(a)Lv for all v}\displaystyle:=\{a\in H^{1}(\theta):\textnormal{res}_{v}(a)\in L_{v}\text{ for all }v\in\mathcal{M}\}
=ker(H1(θ)resvH1(v,θ)/Lv).\displaystyle=\ker\left(H^{1}(\theta)\overset{\textnormal{res}}{\to}\prod_{v\in\mathcal{M}}H^{1}(v,\theta)/L_{v}\right).

5.3. A basis for local cohomology groups

Recall that γ\gamma is the order of qq in 𝐅×\mathbf{F}_{\ell}^{\times} and d=degfd=\deg f. In Section 4.2 we defined G𝐅q(x)G_{\mathbf{F}_{q}(x)}-representations χ\chi and ρ\rho which factor through GC=Gal(𝐅qγ(C)/𝐅q(x))G_{C}=\textnormal{Gal}(\mathbf{F}_{q^{\gamma}}(C)/\mathbf{F}_{q}(x)).

Let

θ=Symkρχn,\theta=\textnormal{Sym}^{k}\rho\otimes\chi^{n},

and let hi(θ)h^{i}(\theta) denote the 𝐅\mathbf{F}_{\ell}-dimension of cohomology group Hi(θ)H^{i}(\theta). Since θ(ξ)\theta(\xi) is similar to a Jordan block as is described in Eq. 10, the dimension of H0(θ)H^{0}(\theta) depends only on whether the top-left entry of θ(frob)\theta(\operatorname{frob}) equals 11. That is,

(23) h0(θ)={1χn+k=10otherwise={1γn+k0otherwise.h^{0}(\theta)=\begin{cases}1&\chi^{n+k}=1\\ 0&\text{otherwise}\end{cases}=\begin{cases}1&\gamma\mid n+k\\ 0&\text{otherwise.}\end{cases}

We now consider θv\theta_{v}, the restriction of θ\theta to the decomposition group G𝐅q(x)vG_{\mathbf{F}_{q}(x)_{v}}, for vS={f,}v\in S=\{f,\infty\}. The polynomial f(x)f(x) splits over 𝐅qγ\mathbf{F}_{q^{\gamma}} into gcd(d,γ)\gcd(d,\gamma) factors which are cyclically permuted by frobGal(𝐅qγ(x)/𝐅q(x))\operatorname{frob}\in\textnormal{Gal}(\mathbf{F}_{q^{\gamma}}(x)/\mathbf{F}_{q}(x)). The decomposition group G𝐅q(x)fG_{\mathbf{F}_{q}(x)_{f}} fixes these factors, so the only powers of frob\operatorname{frob} lying in the image of G𝐅q(x)fGCG_{\mathbf{F}_{q}(x)_{f}}\to G_{C} are powers of frobd\operatorname{frob}^{d}. On the other hand \infty has a unique prime above it in 𝐅qγ(x)\mathbf{F}_{q^{\gamma}}(x), so frob\operatorname{frob} is in the image of G𝐅q(x)G_{\mathbf{F}_{q}(x)_{\infty}}. We can conclude that

(24) h0(f,θ)\displaystyle h^{0}(f,\theta) ={1γd(n+k)0otherwise,,\displaystyle=\begin{cases}1&\gamma\mid d(n+k)\\ 0&\text{otherwise,}\end{cases}, h0(,θ)\displaystyle\qquad h^{0}(\infty,\theta) ={1γ(n+k)0otherwise.\displaystyle=\begin{cases}1&\gamma\mid(n+k)\\ 0&\text{otherwise.}\end{cases}

By Tate duality [9, (7.2.6)] and Lemma 4.6, this implies

(25) h2(f,θ)\displaystyle h^{2}(f,\theta) ={1γd(1n)0otherwise,,\displaystyle=\begin{cases}1&\gamma\mid d(1-n)\\ 0&\text{otherwise,}\end{cases}, h2(,θ)\displaystyle\qquad h^{2}(\infty,\theta) ={1γ(1n)0otherwise.\displaystyle=\begin{cases}1&\gamma\mid(1-n)\\ 0&\text{otherwise.}\end{cases}

Finally, since \ell is coprime to qq, the local Euler characteristic of θ\theta at any vv\in\mathcal{M} is trivial [9, (7.3.2)] and so

(26) h1(v,θ)\displaystyle h^{1}(v,\theta) =h0(v,θ)+h2(v,θ).\displaystyle=h^{0}(v,\theta)+h^{2}(v,\theta).

In particular, we observe that H1(v,θ)H^{1}(v,\theta) is at most 22-dimensional for vSv\in S.

We now give an explicit basis for H1(v,θ)H^{1}(v,\theta) and discuss their corresponding kernel fields.

Lemma 5.3.

Let v=fv=f or \infty, and let δ=d\delta=d or 11 respectively. Let θ=Symkρχn\theta=\textnormal{Sym}^{k}\rho\otimes\chi^{n} for 0k10\leq k\leq\ell-1, and θv\theta_{v} the restriction to G𝐅q(x)vG_{\mathbf{F}_{q}(x)_{v}}.

  1. (a)

    If γδ(n+k)\gamma\mid\delta(n+k) then there exists a nonzero element 𝐮𝐫H1(v,θ)\mathbf{ur}\in H^{1}(v,\theta) such that the kernel field K𝐮𝐫K^{\mathbf{ur}} is the degree \ell unramified extension of KθvK^{\theta_{v}}.

  2. (b)

    If γδ(1n)\gamma\mid\delta(1-n) then there exists a nonzero element 𝐛H1(v,θ)\mathbf{b}\in H^{1}(v,\theta) such that the kernel field K𝐛K^{\mathbf{b}} is 𝐅qγ(C)v\mathbf{F}_{q^{\gamma}}(C)_{v}.

Further, H1(v,θ)H^{1}(v,\theta) has a basis consisting of whichever of 𝐛\mathbf{b} and 𝐮𝐫\mathbf{ur} it contains.

Here 𝐅qγ(C)v\mathbf{F}_{q^{\gamma}}(C)_{v} means the completion of 𝐅qγ(C)\mathbf{F}_{q^{\gamma}}(C) at the place above vv determined by the decomposition group G𝐅q(x)vG_{\mathbf{F}_{q}(x)_{v}}.

Remark 5.4.

If k1k\geq 1 then the Kθv=𝐅qγ(C)vK^{\theta_{v}}=\mathbf{F}_{q^{\gamma}}(C)_{v}, so when 𝐛\mathbf{b} exists in H1(v,θ)H^{1}(v,\theta), K𝐛K^{\mathbf{b}} is not a nontrivial extension of KθvK^{\theta_{v}}; this serves as a caution that a nontrivial cohomology class may define a trivial extension of kernel fields. However, if k=0k=0 then KθvK^{\theta_{v}} is a subfield of 𝐅qγ(x)v\mathbf{F}_{q^{\gamma}}(x)_{v}, so K𝐛/KθvK^{\mathbf{b}}/K^{\theta_{v}} is a nontrivial extension in this case.

Proof.

We start by defining an unramified class 𝐮𝐫H1(v,𝐅)\mathbf{ur}\in H^{1}(v,\mathbf{F}_{\ell}). Since 𝐅\mathbf{F}_{\ell} is the trivial representation, a class in H1(v,𝐅)H^{1}(v,\mathbf{F}_{\ell}) is represented by a group homomorphism. Let 𝐮𝐫H1(v,𝐅)\mathbf{ur}\in H^{1}(v,\mathbf{F}_{\ell}) be defined by the property that it factors through Gal(𝐅q(x)v/𝐅q(x)v)\textnormal{Gal}(\mathbf{F}_{q^{\ell}}(x)_{v}/\mathbf{F}_{q}(x)_{v}), where 𝐮𝐫\mathbf{ur} is represented by Gal(𝐅q(x)v/𝐅q(x)v)𝐅\textnormal{Gal}(\mathbf{F}_{q^{\ell}}(x)_{v}/\mathbf{F}_{q}(x)_{v})\to\mathbf{F}_{\ell} given by sending the Frobenius map ccqc\mapsto c^{q} to 11.

If γδ(n+k)\gamma\mid\delta(n+k), then following Eq. 23 𝐅\mathbf{F}_{\ell} is a subrepresentation of θv\theta_{v}, and the image of 𝐮𝐫H1(v,𝐅)\mathbf{ur}\in H^{1}(v,\mathbf{F}_{\ell}) under H1(v,𝐅)H1(v,θv)H^{1}(v,\mathbf{F}_{\ell})\to H^{1}(v,\theta_{v}) we also denote by 𝐮𝐫\mathbf{ur}. The kernel field of 𝐮𝐫H1(v,θv)\mathbf{ur}\in H^{1}(v,\theta_{v}) is the compositum 𝐅qKθv\mathbf{F}_{q^{\ell}}\cdot K^{\theta_{v}}; as KθvK^{\theta_{v}} is the composite of an unramified 𝐙/γ𝐙\mathbf{Z}/\gamma\mathbf{Z} extension with a ramified 𝐙/𝐙\mathbf{Z}/\ell\mathbf{Z} extension, it never contains 𝐅q\mathbf{F}_{q^{\ell}}. Thus K𝐮𝐫/KθvK^{\mathbf{ur}}/K^{\theta_{v}} is a nontrivial extension and so the class 𝐮𝐫\mathbf{ur} is nonzero.

If γδ(1n)\gamma\mid\delta(1-n), then χvn=χv\chi^{n}_{v}=\chi_{v}, so the localizations at vv of Symkρχn\textnormal{Sym}^{k}\rho\otimes\chi^{n} and Symkρχ\textnormal{Sym}^{k}\rho\otimes\chi are isomorphic. The representation Symk+1ρ\textnormal{Sym}^{k+1}\rho is an extension of 11 by Symkρχ\textnormal{Sym}^{k}\rho\otimes\chi following Lemma 4.4, so we can define 𝐛H1(v,θv)\mathbf{b}\in H^{1}(v,\theta_{v}) to be the class corresponding to this extension by 11, or equivalently the image of 1H0(v,𝐅)1\in H^{0}(v,\mathbf{F}_{\ell}) under δ\delta in the long exact sequence

0H0(v,Symkρχ)H0(v,Symk+1ρ)H0(v,𝐅)𝛿H1(v,Symkρχ).0\to H^{0}(v,\textnormal{Sym}^{k}\rho\otimes\chi)\to H^{0}(v,\textnormal{Sym}^{k+1}\rho)\to H^{0}(v,\mathbf{F}_{\ell})\xrightarrow{\delta}H^{1}(v,\textnormal{Sym}^{k}\rho\otimes\chi).

The injection H0(v,Symkρχ)H0(v,Symk+1ρ)H^{0}(v,\textnormal{Sym}^{k}\rho\otimes\chi)\to H^{0}(v,\textnormal{Sym}^{k+1}\rho) is an isomorphism because both groups have the same dimension, and therefore 𝐛=δ(1)0\mathbf{b}=\delta(1)\neq 0. The kernel field of 𝐛\mathbf{b} is equal to the fixed field of kerρv\ker\rho_{v}. Note that this equals the fixed field of kerθv\ker\theta_{v} provided k1k\geq 1, which is why we can’t use the kernel field to deduce that 𝐛\mathbf{b} defines a nonzero cohomology class.

If both γδ(n+k)\gamma\mid\delta(n+k) and γδ(1n)\gamma\mid\delta(1-n), then 𝐛\mathbf{b} and 𝐮𝐫\mathbf{ur} define independent classes in H1(v,Symkρχn)H^{1}(v,\textnormal{Sym}^{k}\rho\otimes\chi^{n}) because they define distinct kernel fields over 𝐅q(x)v\mathbf{F}_{q}(x)_{v} (Lemma 5.2). So whichever of the elements 𝐛\mathbf{b} and 𝐮𝐫\mathbf{ur} exist in H1(v,θv)H^{1}(v,\theta_{v}), they span a subspace that matches the dimension h1(v,θ)h^{1}(v,\theta) computed above, and therefore they must form a basis. ∎

5.4. Selmer groups of characters

Using the computations from Section 5.3, we can compute the dimension of global Selmer groups of characters.

Lemma 5.5.

Recall hS1(χn)h^{1}_{S}(\chi^{n}) denotes the dimension of the cohomology group HS1(χn)H^{1}_{S}(\chi^{n}). We have

hS1(χn)={1γd(1n)0otherwise+{1γn0otherwise.h^{1}_{S}(\chi^{n})=\begin{cases}1&\gamma\mid d(1-n)\\ 0&\text{otherwise}\end{cases}+\begin{cases}1&\gamma\mid n\\ 0&\text{otherwise.}\end{cases}
Proof.

We begin by defining an alternate Selmer group. Let θ\theta be a G𝐅q(x)G_{\mathbf{F}_{q}(x)}-representation and θ\theta^{*} its cohomological dual (Definition 4.5). For each place vv\in\mathcal{M} we define a subgroup of H1(v,θ)H^{1}(v,\theta^{*}) by

Lv:={0vS,H1(Gkv,θ)vS,L_{v}^{\perp}:=\left\{\begin{array}[]{ll}0&v\in S,\\ H^{1}(G_{k_{v}},\theta^{*})&v\in\mathcal{M}\setminus S,\end{array}\right.

where H1(Gkv,θ)H^{1}(G_{k_{v}},\theta^{*}) is the unramified subspace as defined in Eq. 22. For all places vv of 𝐅q(x)\mathbf{F}_{q}(x), the subgroup LvH1(v,θ)L_{v}^{\perp}\leq H^{1}(v,\theta^{*}) is precisely the annihilator under the local Tate pairing of the subgroup Lv(θ)H1(v,θ)L_{v}(\theta)\leq H^{1}(v,\theta) as defined in Eq. 22 [9, Theorem 7.2.15]. Using these subgroups we define

HS1(θ)\displaystyle H^{1}_{S^{*}}(\theta^{*}) :={aH1(θ):resv(a)Lv for all v}.\displaystyle:=\{a\in H^{1}(\theta^{*}):\textnormal{res}_{v}(a)\in L_{v}^{\perp}\text{ for all }v\}.

In this setting, the Greenberg-Wiles formula [9, Theorem 8.7.9] reduces to

#HS1(χn)#HS1(χ1n)=#H0(χn)#H0(χ1n)#H1(f,χn)#H0(f,χn)#H1(,χn)#H0(,χn),\frac{\#H^{1}_{S}(\chi^{n})}{\#H^{1}_{S^{*}}(\chi^{1-n})}=\frac{\#H^{0}(\chi^{n})}{\#H^{0}(\chi^{1-n})}\cdot\frac{\#H^{1}(f,\chi^{n})}{\#H^{0}(f,\chi^{n})}\cdot\frac{\#H^{1}(\infty,\chi^{n})}{\#H^{0}(\infty,\chi^{n})},

because for all v{f,}v\notin\{f,\infty\}, we have #Lv=#H0(v,θ)\#L_{v}=\#H^{0}(v,\theta) (see the proof of [9, (8.7.9)]). The right-hand side of this equation can be determined using the dimension computations in Section 5.3, and equals the right-hand side of the statement of the lemma. So it suffices to show that HS1(χ1n)H^{1}_{S^{\ast}}(\chi^{1-n}) is trivial.

Now suppose there exists a nonzero class aHS1(χ1n)a\in H^{1}_{S^{\ast}}(\chi^{1-n}) and let KaK^{a} denote the kernel field of aa. For vSv\in S, the Selmer condition LvL_{v}^{\perp} implies that the G𝐅q(x)vG_{\mathbf{F}_{q}(x)_{v}}-representations θv\theta^{*}_{v} and θ[a]v=θv[resv(a)]\theta^{*}[a]_{v}=\theta^{*}_{v}[\textnormal{res}_{v}(a)] have the same kernel field, so the extension Ka/KθK^{a}/K^{\theta} is totally split at all places over vSv\in S. Since aa is not a coboundary, the extension of 11 determined by aa must have a nontrivial unipotent element in its image, so there is an element of order \ell in Gal(Ka/𝐅q(x))\textnormal{Gal}(K^{a}/\mathbf{F}_{q}(x)). Since Gal(𝐅qγ(x)/𝐅q(x))\textnormal{Gal}(\mathbf{F}_{q^{\gamma}}(x)/\mathbf{F}_{q}(x)) has order coprime to \ell, this implies that KaK^{a} is a 𝐙/𝐙\mathbf{Z}/\ell\mathbf{Z} extension of some subfield of 𝐅qγ(x)\mathbf{F}_{q^{\gamma}}(x) that is unramified everywhere and split at ff and \infty. No such extensions exist, so in fact HS1(χ1n)H^{1}_{S^{\ast}}(\chi^{1-n}) is trivial. Hence the dimension of HS1(χn)H^{1}_{S}(\chi^{n}) is exactly as predicted by the statement of the lemma. ∎

5.5. From Frob\operatorname{Frob} eigenvectors to cohomology

Using the cohomology computations above, we can now complete the work we began in Section 4 of relating the existence of eigenvectors of Frob\operatorname{Frob} in FnkF_{n}^{k} to the existence of certain cohomology classes.

Lemma 5.6.

Let aHS1(Symkρχn)a\in H^{1}_{S}(\textnormal{Sym}^{k}\rho\otimes\chi^{n}) for some 1k11\leq k\leq\ell-1, and let L/𝐅q(x)L/\mathbf{F}_{q}(x) be the kernel field of aa. Then LL is an unramified extension of 𝐅qγ(C)\mathbf{F}_{q^{\gamma}}(C).

Proof.

The kernel field contains the fixed field of kerSymkρχn\ker\textnormal{Sym}^{k}\rho\otimes\chi^{n}, which is 𝐅qγ(C)\mathbf{F}_{q^{\gamma}}(C) because k1k\geq 1. The Selmer condition ensures LL is unramified over 𝐅qγ(C)\mathbf{F}_{q^{\gamma}}(C) at all vSv\notin S, so it suffices to check the ramification at vSv\in S. By Lemma 5.3, for vSv\in S, resv(a)\textnormal{res}_{v}(a) is a linear combination of 𝐮𝐫\mathbf{ur} and 𝐛\mathbf{b} (allowing the coefficient to be 0 if the corresponding class is not in H1(v,θ)H^{1}(v,\theta)). Since these classes both define unramified extensions of 𝐅qγ(C)v\mathbf{F}_{q^{\gamma}}(C)_{v}, the corresponding extensions of 11 both vanish on the inertia group of 𝐅qγ(C)v\mathbf{F}_{q^{\gamma}}(C)_{v}, so the same is true of any linear combination. This implies that the kernel field of resv(a)\textnormal{res}_{v}(a) is an unramified extension of 𝐅qγ(C)v\mathbf{F}_{q^{\gamma}}(C)_{v}. Since the kernel field of resv(a)\textnormal{res}_{v}(a) is the completion at a prime above vv of the kernel field of aa, we can conclude that LL is unramified over 𝐅qγ(C)\mathbf{F}_{q^{\gamma}}(C) at vSv\in S, and also at all vSv\notin S by the Selmer condition. ∎

Proposition 5.7.

Let 2k12\leq k\leq\ell-1 and n𝐙/γ𝐙n\in\mathbf{Z}/\gamma\mathbf{Z}. The following are equivalent:

  • There exists an eigenvector of Frob\operatorname{Frob} in FnkF_{n}^{k}.

  • γn\gamma\nmid n, and there exists a class aHS1(Symk1ρχ1n)a\in H^{1}_{S}(\textnormal{Sym}^{k-1}\rho\otimes\chi^{1-n}) that maps to a nonzero class aHS1(χ1n)a^{\prime}\in H^{1}_{S}(\chi^{1-n}) under the map induced by Eq. 11.

Proof.

Given a Frobenius eigenvector wFnkw\in F_{n}^{k}, we obtain a representation ψ\psi as in Proposition 4.14, which is an extension of 11 by θ:=Symk1ρχ1n\theta:=\textnormal{Sym}^{k-1}\rho\otimes\chi^{1-n} and therefore corresponds to a class aH1(θ)a\in H^{1}(\theta). Letting KψK^{\psi} denote the kernel field, by Proposition 4.14 we have Kψ/𝐅qγ(C)K^{\psi}/\mathbf{F}_{q^{\gamma}}(C) unramified and therefore aHS1(θ)a\in H^{1}_{S}(\theta). If we write ψ\psi in matrix form by picking a cocycle α\alpha as

(27) ψ(τ)=(θ(τ)α(τ)01),\psi(\tau)=\begin{pmatrix}\theta(\tau)&\alpha(\tau)\\ 0&1\end{pmatrix},

then τGal(𝐅q(x)sep/𝐅qγ(C))\tau\in\textnormal{Gal}(\mathbf{F}_{q}(x)^{\textnormal{sep}}/\mathbf{F}_{q^{\gamma}}(C)) maps to a matrix of the form (Ikα(τ)01)\begin{pmatrix}I_{k}&\alpha(\tau)\\ 0&1\end{pmatrix}. Since ψ\psi descends to a faithful representation of Gal(Kψ/𝐅qγ(C))\textnormal{Gal}(K^{\psi}/\mathbf{F}_{q^{\gamma}}(C)), the map τα(τ)\tau\mapsto\alpha(\tau) is an isomorphism onto 𝐅k\mathbf{F}_{\ell}^{k} by Proposition 4.14. So letting aHS1(χ1n)a^{\prime}\in H^{1}_{S}(\chi^{1-n}) be the class given by restriction of α\alpha to the bottom entry, there exists τGal(Kψ/𝐅qγ(C))\tau\in\textnormal{Gal}(K^{\psi}/\mathbf{F}_{q^{\gamma}}(C)) for which a(τ)0a^{\prime}(\tau)\neq 0; this proves that the kernel field of aa^{\prime} is strictly larger than 𝐅qγ(C)\mathbf{F}_{q^{\gamma}}(C), so aa^{\prime} must define a nonzero class in HS1(χ1n)H^{1}_{S}(\chi^{1-n}).

We also have γn\gamma\nmid n. For if γn\gamma\mid n, then χ1n=χ\chi^{1-n}=\chi and hS1(χ1n)=1h^{1}_{S}(\chi^{1-n})=1 by Lemma 5.5. This implies that up to scalar multiple, aHS1(χ1n)a^{\prime}\in H^{1}_{S}(\chi^{1-n}) is the class defining ρ\rho as an extension of 11 by χ\chi, and therefore has kernel field 𝐅qγ(C)\mathbf{F}_{q^{\gamma}}(C), a contradiction.

Conversely, suppose we are given an element aHS1(θ)a\in H^{1}_{S}(\theta) satisfying the described condition with γn\gamma\nmid n. This class corresponds to an extension of 11 by θ\theta which we denote ψ\psi. Let KψK^{\psi} denote the kernel field of ψ\psi. By Lemma 5.6, KψK^{\psi} is an unramified extension of 𝐅qγ(C)\mathbf{F}_{q^{\gamma}}(C).

We can pick a cocycle α\alpha such that the matrix form ψ\psi has the form as in Eq. 27 where the last entry α\alpha^{\prime} of α\alpha represents the nonzero class aHS1(χ1n)a^{\prime}\in H^{1}_{S}(\chi^{1-n}) given by the assumption.

Now we claim that with this matrix form, there must exist τGal(𝐅q(x)sep/𝐅qγ(C))\tau\in\textnormal{Gal}(\mathbf{F}_{q}(x)^{\textnormal{sep}}/\mathbf{F}_{q^{\gamma}}(C)) for which α(τ)0\alpha^{\prime}(\tau)\neq 0. If not, the fact α(τ)=0\alpha^{\prime}(\tau)=0 for all τGal(𝐅q(x)sep/𝐅qγ(C))\tau\in\textnormal{Gal}(\mathbf{F}_{q}(x)^{\textnormal{sep}}/\mathbf{F}_{q^{\gamma}}(C)) implies that the kernel field of aa^{\prime} is contained in 𝐅qγ(C)\mathbf{F}_{q^{\gamma}}(C). Thus the representation (χ1nα01)\begin{pmatrix}\chi^{1-n}&\alpha^{\prime}\\ 0&1\end{pmatrix} factors through GC=Gal(𝐅qγ(C)/𝐅q(x))=ξ,frobG_{C}=\textnormal{Gal}(\mathbf{F}_{q^{\gamma}}(C)/\mathbf{F}_{q}(x))=\langle\xi,\operatorname{frob}\rangle. Since

qα(ξ)=α(ξq)=α(frobξfrob1)=q1nα(ξ),\displaystyle q\alpha^{\prime}(\xi)=\alpha^{\prime}(\xi^{q})=\alpha^{\prime}(\operatorname{frob}\xi\operatorname{frob}^{-1})=q^{1-n}\alpha^{\prime}(\xi),

and γn\gamma\nmid n by assumption, we must have α(ξ)=0\alpha^{\prime}(\xi)=0, so the representation factors through Gal(𝐅qγ(x)/𝐅q(x))\textnormal{Gal}(\mathbf{F}_{q^{\gamma}}(x)/\mathbf{F}_{q}(x)) which is a cyclic group. This contradicts the assumption that aa^{\prime} is a nonzero class in HS1(χ1n)H^{1}_{S}(\chi^{1-n}).

Hence there exists τGal(𝐅q(x)sep/𝐅qγ(C))\tau\in\textnormal{Gal}(\mathbf{F}_{q}(x)^{\textnormal{sep}}/\mathbf{F}_{q^{\gamma}}(C)) for which α(τ)\alpha(\tau) has nonzero bottom entry. Now for arbitrary σG𝐅q(x)\sigma\in G_{\mathbf{F}_{q}(x)} we have α(στσ1)=θ(σ)α(τ)\alpha(\sigma\tau\sigma^{-1})=\theta(\sigma)\alpha(\tau), so every vector in the orbit of α(τ)\alpha(\tau) under the G𝐅q(x)G_{\mathbf{F}_{q}(x)}-representation θ\theta is obtained as α(τ)\alpha(\tau^{\prime}) for some τGal(𝐅q(x)sep/𝐅qγ(C))\tau^{\prime}\in\textnormal{Gal}(\mathbf{F}_{q}(x)^{\textnormal{sep}}/\mathbf{F}_{q^{\gamma}}(C)). The orbit under θ\theta of a vector with nonzero bottom entry has full dimension kk, so the restriction of ψ\psi to Gal(𝐅q(x)sep/𝐅qγ(C))\textnormal{Gal}(\mathbf{F}_{q}(x)^{\textnormal{sep}}/\mathbf{F}_{q^{\gamma}}(C)) has image isomorphic to 𝐅k\mathbf{F}_{\ell}^{k}; in particular this implies Gal(Kψ/𝐅qγ(C))𝐅k\textnormal{Gal}(K^{\psi}/\mathbf{F}_{q^{\gamma}}(C))\simeq\mathbf{F}_{\ell}^{k}. We thus obtain a representation ψ\psi as in Proposition 4.14, from which we obtain an eigenvector in FnkF_{n}^{k}. ∎

6. Cup products

Suppose 0k30\leq k\leq\ell-3 and we want to determine the existence of Frob\operatorname{Frob} eigenvectors in F1nk+2F^{k+2}_{1-n} for n𝐙/γ𝐙n\in\mathbf{Z}/\gamma\mathbf{Z}. In light of Proposition 5.7, we are led to consider when HS1(Symk+1ρχn)H^{1}_{S}(\textnormal{Sym}^{k+1}\rho\otimes\chi^{n}) contains an element that maps to a nontrivial element of HS1(χn)H^{1}_{S}(\chi^{n}) under the map induced by Eq. 11. The map Symk+1ρχnχn\textnormal{Sym}^{k+1}\rho\otimes\chi^{n}\to\chi^{n} factors through the map Symk+1ρχnSymkρχn\textnormal{Sym}^{k+1}\rho\otimes\chi^{n}\to\textnormal{Sym}^{k}\rho\otimes\chi^{n} coming from Eq. 12, so a necessary condition is that HS1(Symkρχn)H^{1}_{S}(\textnormal{Sym}^{k}\rho\otimes\chi^{n}) contains a class that maps to a nontrivial element of HS1(χn)H^{1}_{S}(\chi^{n}). So we will now assume we are given a class in HS1(Symkρχn)H^{1}_{S}(\textnormal{Sym}^{k}\rho\otimes\chi^{n}), and want to know when it lifts to an element in HS1(Symk+1ρχn)H^{1}_{S}(\textnormal{Sym}^{k+1}\rho\otimes\chi^{n}). We will see that this condition is equivalent to the vanishing of a local cup product, and explain how to determine this vanishing condition explicitly in the case k=0k=0.

6.1. Lifting cohomology classes

Let 𝐅q(x)S\mathbf{F}_{q}(x)_{S} denote the maximal separable extension of 𝐅q(x)\mathbf{F}_{q}(x) that is unramified outside SS. Then 𝐅q(x)S\mathbf{F}_{q}(x)_{S} contains 𝐅qγ(C)\mathbf{F}_{q^{\gamma}}(C), and so for any G𝐅q(x)G_{\mathbf{F}_{q}(x)}-representation θ\theta that factors through GCG_{C}, θ\theta descends to a representation of G𝐅q(x),S:=Gal(𝐅q(x)S/𝐅q(x))G_{\mathbf{F}_{q}(x),S}:=\textnormal{Gal}(\mathbf{F}_{q}(x)_{S}/\mathbf{F}_{q}(x)). For the Selmer conditions defined in Section 5.2, we can write the corresponding Selmer group as a full cohomology group:

HS1(θ)\displaystyle H^{1}_{S}(\theta) =ker(H1(G𝐅q(x),θ)H1(Gal(𝐅q(x)sep/𝐅q(x)S),θ))\displaystyle=\ker\left(H^{1}(G_{\mathbf{F}_{q}(x)},\theta)\to H^{1}(\textnormal{Gal}(\mathbf{F}_{q}(x)^{\textnormal{sep}}/\mathbf{F}_{q}(x)_{S}),\theta)\right)
=H1(G𝐅q(x),S,θ).\displaystyle=H^{1}(G_{\mathbf{F}_{q}(x),S},\theta).

This will allow us to locate HS1(θ)H^{1}_{S}(\theta) within a long exact sequence in cohomology. We also define HiS(θ):=Hi(G𝐅q(x),S,θ)H^{i}_{S}(\theta):=H^{i}(G_{\mathbf{F}_{q}(x),S},\theta) for all i0i\geq 0.

Let 0k20\leq k\leq\ell-2. Take the short exact sequence of representations

0χ1+k+nSymk+1ρχnSymkρχn00\to\chi^{1+k+n}\to\textnormal{Sym}^{k+1}\rho\otimes\chi^{n}\to\textnormal{Sym}^{k}\rho\otimes\chi^{n}\to 0

from Eq. 12, but now considered as representations of G𝐅q(x),SG_{\mathbf{F}_{q}(x),S}. The corresponding long exact sequence has a portion given by

(28) H1S(Symk+1ρχn)H1S(Symkρχn)𝛿H2S(χ1+k+n).\displaystyle\cdots\to H^{1}_{S}(\textnormal{Sym}^{k+1}\rho\otimes\chi^{n})\to H^{1}_{S}(\textnormal{Sym}^{k}\rho\otimes\chi^{n})\overset{\delta}{\to}H^{2}_{S}(\chi^{1+k+n})\to\cdots.

From this we see that a class aH1S(Symkρχn)a\in H^{1}_{S}(\textnormal{Sym}^{k}\rho\otimes\chi^{n}) lifts to a class aH1S(Symk+1ρχn)a^{\prime}\in H^{1}_{S}(\textnormal{Sym}^{k+1}\rho\otimes\chi^{n}) if and only if the image of aa under the connecting homomorphism δ\delta to H2S(χ1+k+n)H^{2}_{S}(\chi^{1+k+n}) is 0. We will show that this condition can be detected by the vanishing of a cup product

H1S(Symkρχ)H1S(Symkρχn)H2S(χ1+k+n)\displaystyle H^{1}_{S}(\textnormal{Sym}^{k}\rho\otimes\chi)\otimes H^{1}_{S}(\textnormal{Sym}^{k}\rho\otimes\chi^{n})\xrightarrow{\cup}H^{2}_{S}(\chi^{1+k+n})

induced by the dual pairing (Symkρ)(Symkρ)𝐅(\textnormal{Sym}^{k}\rho)\otimes(\textnormal{Sym}^{k}\rho)^{\vee}\to\mathbf{F}_{\ell} together with the calculation in Lemma 4.6.

Note that Symk+1ρ\textnormal{Sym}^{k+1}\rho is an extension of 11 by Symkρχ\textnormal{Sym}^{k}\rho\otimes\chi, unramified away from SS, and therefore corresponds to a class in H1S(Symkρχ)H^{1}_{S}(\textnormal{Sym}^{k}\rho\otimes\chi). Under the map H1S(Symkρχ)χH^{1}_{S}(\textnormal{Sym}^{k}\rho\otimes\chi)\to\chi as in Eq. 11, this class maps to the class bH1S(χ)b\in H^{1}_{S}(\chi) that determines ρ\rho as an extension of 11, as discussed in Section 4.2. We therefore denote this class by b[k]b^{[k]}. As was discussed in Section 5.1, we can also identify this class as the image of 1H0S(𝐅)1\in H^{0}_{S}(\mathbf{F}_{\ell}) under the connecting homomorphism H0S(𝐅)H1S(Symkρχ)H^{0}_{S}(\mathbf{F}_{\ell})\to H^{1}_{S}(\textnormal{Sym}^{k}\rho\otimes\chi) induced by the following short exact sequence obtained from Eq. 11 by taking n=0n=0,

0SymkρχSymk+1ρ𝐅0.0\to\textnormal{Sym}^{k}\rho\otimes\chi\to\textnormal{Sym}^{k+1}\rho\to\mathbf{F}_{\ell}\to 0.
Lemma 6.1.

The image of a class aH1S(Symkρχn)a\in H^{1}_{S}(\textnormal{Sym}^{k}\rho\otimes\chi^{n}) under the connecting homomorphism in Eq. 28

δ:H1S(Symkρχn)H2S(χ1+k+n)\delta:H^{1}_{S}(\textnormal{Sym}^{k}\rho\otimes\chi^{n})\to H^{2}_{S}(\chi^{1+k+n})

equals b[k]ab^{[k]}\cup a.

Proof.

This follows from the formal properties of cup products and connecting homomorphisms in group cohomology. To be precise, if we let θ=Symkρχn\theta=\textnormal{Sym}^{k}\rho\otimes\chi^{n} then we have a GCG_{C}-equivariant map of short exact sequences

0{0}(Symkρχ)θ{(\textnormal{Sym}^{k}\rho\otimes\chi)\otimes\theta}(Symk+1ρ)θ{(\textnormal{Sym}^{k+1}\rho)\otimes\theta}𝐅θ{\mathbf{F}_{\ell}\otimes\theta}0{0}0{0}χ1+k+n{\chi^{1+k+n}}Symk+1ρχn{\textnormal{Sym}^{k+1}\rho\otimes\chi^{n}}Symkρχn{\textnormal{Sym}^{k}\rho\otimes\chi^{n}}0.{0.}

The left vertical arrow is induced by the dual pairing; the right vertical arrow is the isomorphism 𝐅θθ\mathbf{F}_{\ell}\otimes\theta\simeq\theta; then there is a unique choice of middle arrow that makes the diagram commute. Then [9, (1.4.3.i)] says that the following diagram commutes:

H0S(𝐅)H1S(Symkρχn){H^{0}_{S}(\mathbf{F}_{\ell})\otimes H^{1}_{S}(\textnormal{Sym}^{k}\rho\otimes\chi^{n})}H1S(Symkρχn){H^{1}_{S}(\textnormal{Sym}^{k}\rho\otimes\chi^{n})}H1S(Symkρχ)H1S(Symkρχn){H^{1}_{S}(\textnormal{Sym}^{k}\rho\otimes\chi)\otimes H^{1}_{S}(\textnormal{Sym}^{k}\rho\otimes\chi^{n})}H2S(χ1+k+n).{H^{2}_{S}(\chi^{1+k+n}).}\scriptstyle{\cup}δ1\scriptstyle{\delta\otimes 1}δ\scriptstyle{\delta}\scriptstyle{\cup}

Note in particular that the top cup product sends 1aa1\cup a\mapsto a. Therefore we have

δ(a)=δ(1a)=δ(1)a=b[k]a.\delta(a)=\delta(1\cup a)=\delta(1)\cup a=b^{[k]}\cup a.\qed

6.2. Local cup product

In the previous section we showed that a class in H1S(Symkρχn)H^{1}_{S}(\textnormal{Sym}^{k}\rho\otimes\chi^{n}) lifts to a class in H1S(Symk+1ρχn)H^{1}_{S}(\textnormal{Sym}^{k+1}\rho\otimes\chi^{n}) if and only if its cup product with b[k]b^{[k]} vanishes. Our next goal is to show that the vanishing of the cup product of b[k]b^{[k]} and a class in H1S(Symkρχn)H^{1}_{S}(\textnormal{Sym}^{k}\rho\otimes\chi^{n}) can be detected locally at the place ff (Proposition 6.4).

Lemma 6.2.

Recall h2S(χn)h^{2}_{S}(\chi^{n}) denotes the dimension of the cohomology group H2S(χn)H^{2}_{S}(\chi^{n}). We have

h2S(χn)={1γd(1n)0otherwise.h^{2}_{S}(\chi^{n})=\begin{cases}1&\gamma\mid d(1-n)\\ 0&\text{otherwise.}\end{cases}
Proof.

This formula follows from Eq. 23, Lemma 5.5, and the triviality of the global Euler characteristic over function fields [9, (8.7.4)]. ∎

Lemma 6.3.

The restriction map resf:H2S(χn)H2(f,χn)\textnormal{res}_{f}:H^{2}_{S}(\chi^{n})\to H^{2}(f,\chi^{n}) is an isomorphism.

Proof.

By Eq. 25 and Lemma 6.2, if γd(1n)\gamma\nmid d(1-n) then both groups are trivial, so we may suppose γd(1n)\gamma\mid d(1-n). Then both groups have dimension 11, so it suffices to show that resf\textnormal{res}_{f} is nonzero. We will extract this from the end of the Poitou-Tate exact sequence [9, (8.6.10)]. Recall since χ\chi is finite and unramified, we have Hom(χ,𝒪S×)=χ=χ1n\textnormal{Hom}(\chi,\mathcal{O}_{S}^{\times})=\chi^{*}=\chi^{1-n} ([9, page 387]). The relevant portion of the sequence is

(29) H2S(χn)resvvSH2(v,χn)H0S(χ1n)0.\displaystyle\to H^{2}_{S}(\chi^{n})\xrightarrow{\oplus\textnormal{res}_{v}}\bigoplus_{v\in S}H^{2}(v,\chi^{n})\to H^{0}_{S}(\chi^{1-n})^{\vee}\to 0.

Our ramification set SS contains only two places, ff and \infty. We know the dimensions of every term appearing in this sequence: the global H2SH^{2}_{S} by Lemma 6.2, the middle terms by Eq. 25, and the term H0S(χ1n)H^{0}_{S}(\chi^{1-n})^{\vee} has dimension 11 when n1modγn\equiv 1\bmod\gamma and dimension 0 otherwise.

When n1modγn\not\equiv 1\bmod\gamma the only non-zero terms are H2S(χn)H^{2}_{S}(\chi^{n}) and H2(f,χn)H^{2}(f,\chi^{n}). Exactness of this sequence implies that the map resf\textnormal{res}_{f} is not the zero map, and is therefore an isomorphism.

When n1modγn\equiv 1\bmod\gamma the groups H2S(χn)H^{2}_{S}(\chi^{n}), H2(f,χn)H^{2}(f,\chi^{n}), H2(,χn)H^{2}(\infty,\chi^{n}) and H0S(χ1n)H^{0}_{S}(\chi^{1-n})^{\vee} are all 11-dimensional. The second map

H2(f,χn)H2(,χn)H0S(χ1n)H^{2}(f,\chi^{n})\oplus H^{2}(\infty,\chi^{n})\to H^{0}_{S}(\chi^{1-n})^{\vee}

is by definition (e.g. [9, page 495]) the linear dual of the restriction map

H0S(χ1n)resvH0(f,χ1n)H0(,χ1n)(H2(f,χn)H2(,χn)),H^{0}_{S}(\chi^{1-n})\xrightarrow{\oplus\textnormal{res}_{v}}H^{0}(f,\chi^{1-n})\oplus H^{0}(\infty,\chi^{1-n})\simeq(H^{2}(f,\chi^{n})\oplus H^{2}(\infty,\chi^{n}))^{\vee},

with the isomorphism following by Tate duality [9, (7.2.6)] and Lemma 4.6. Since the image of this restriction map has non-zero projection in both local factors, the kernel of the linear dual map necessarily also has non-zero projection on both local factors. From this, together with exactness of Eq. 29, we conclude that the map resf\textnormal{res}_{f} necessarily has non-trivial image in H2(f,χn)H^{2}(f,\chi^{n}), and so again must be an isomorphism. ∎

Proposition 6.4.

Let k2k\leq\ell-2. A class aH1S(Symkρχn)a\in H^{1}_{S}(\textnormal{Sym}^{k}\rho\otimes\chi^{n}) lifts to a class in H1S(Symk+1ρχn)H^{1}_{S}(\textnormal{Sym}^{k+1}\rho\otimes\chi^{n}) if and only if the local cup product

resf(b[k])resf(a)H2(f,χ1+k+n)\textnormal{res}_{f}(b^{[k]})\cup\textnormal{res}_{f}(a)\in H^{2}(f,\chi^{1+k+n})

vanishes.

Proof.

From Eq. 28 we see that aa lifts to a class aH1S(Symk+1ρχn)a^{\prime}\in H^{1}_{S}(\textnormal{Sym}^{k+1}\rho\otimes\chi^{n}) exactly when δ(a)=0H2S(χ1+k+n)\delta(a)=0\in H^{2}_{S}(\chi^{1+k+n}). Lemma 6.3 shows that resf:H2S(χ1+k+n)H2(f,χ1+k+n)\textnormal{res}_{f}:H^{2}_{S}(\chi^{1+k+n})\to H^{2}(f,\chi^{1+k+n}) is injective, so the vanishing of δ(a)\delta(a) can be checked locally at ff. Cup products and connecting homomorphism commute with restriction maps, so by Lemma 6.1, we have δ(a)=0\delta(a)=0 exactly when

0=resf(δ(a))=resf(b[k]a)=resf(b[k])resf(a).0=\textnormal{res}_{f}(\delta(a))=\textnormal{res}_{f}(b^{[k]}\cup a)=\textnormal{res}_{f}(b^{[k]})\cup\textnormal{res}_{f}(a).\qed

Since we’ve reduced our lifting question to one about the vanishing of local cup products, it is now in our interest to do a detailed analysis of these local cup products. Recall from Lemma 5.3 that H1(f,Symkρχn)H^{1}(f,\textnormal{Sym}^{k}\rho\otimes\chi^{n}) contains a nonzero element 𝐮𝐫\mathbf{ur} when γd(n+k)\gamma\mid d(n+k), and contains a nonzero element 𝐛\mathbf{b} when γd(1n)\gamma\mid d(1-n), and whichever of these classes exist form a basis of H1(f,Symkρχn)H^{1}(f,\textnormal{Sym}^{k}\rho\otimes\chi^{n}). Note that b[k]H1S(Symkρχ)b^{[k]}\in H^{1}_{S}(\textnormal{Sym}^{k}\rho\otimes\chi) maps under resf\textnormal{res}_{f} to 𝐛H1(f,Symkρχ)\mathbf{b}\in H^{1}(f,\textnormal{Sym}^{k}\rho\otimes\chi).

Lemma 6.5.

Let 0k10\leq k\leq\ell-1. Under the cup product pairing

H1(f,Symkρχm)×H1(f,Symkρχn)H2(f,χk+m+n),H^{1}(f,\textnormal{Sym}^{k}\rho\otimes\chi^{m})\times H^{1}(f,\textnormal{Sym}^{k}\rho\otimes\chi^{n})\xrightarrow{\cup}H^{2}(f,\chi^{k+m+n}),

we have

𝐛𝐛\displaystyle\mathbf{b}\cup\mathbf{b} =0,\displaystyle=0, 𝐛𝐮𝐫\displaystyle\mathbf{b}\cup\mathbf{ur} 0,\displaystyle\neq 0,
𝐮𝐫𝐮𝐫\displaystyle\mathbf{ur}\cup\mathbf{ur} =0,\displaystyle=0, 𝐮𝐫𝐛\displaystyle\mathbf{ur}\cup\mathbf{b} 0\displaystyle\neq 0

whenever each exist in their respective cohomology group. In particular we have that 𝐛a=0\mathbf{b}\cup a=0 if and only if aa is in the span of 𝐛\mathbf{b}.

Proof.

We divide into cases depending on which of 𝐮𝐫,𝐛\mathbf{ur},\mathbf{b} exist in each of the two groups H1(f,Symkρχm)H^{1}(f,\textnormal{Sym}^{k}\rho\otimes\chi^{m}) and H1(f,Symkρχn)H^{1}(f,\textnormal{Sym}^{k}\rho\otimes\chi^{n}).

  • If the groups are 11-dimensional and spanned by the same class (for example γd(n1)\gamma\mid d(n-1) and γd(m1)\gamma\mid d(m-1), but γd(n+k)\gamma\nmid d(n+k) and γd(m+k)\gamma\nmid d(m+k)), then the representations Symkρχn\textnormal{Sym}^{k}\rho\otimes\chi^{n} and Symkρχm\textnormal{Sym}^{k}\rho\otimes\chi^{m} have the same restriction to G𝐅q(x)fG_{\mathbf{F}_{q}(x)_{f}}. The cup product is alternating on H1H^{1} [9, (1.4.4)], so the cup product of a class with itself equals zero.

  • If the groups are 11-dimensional but spanned by different classes (for example γd(n1)\gamma\mid d(n-1) and γd(m+k)\gamma\mid d(m+k), but γd(n+k)\gamma\nmid d(n+k) and γd(m1)\gamma\nmid d(m-1)), then the restrictions of Symkρχn\textnormal{Sym}^{k}\rho\otimes\chi^{n} and Symkρχm\textnormal{Sym}^{k}\rho\otimes\chi^{m} to G𝐅q(x)fG_{\mathbf{F}_{q}(x)_{f}} are cohomological duals of each other. In this case the cup product is the local Tate pairing [9, (7.2.6)], which is non-degenerate; hence the cup product of the respective generators is nonzero.

  • If both groups are 22-dimensional, then the representations Symkρχn\textnormal{Sym}^{k}\rho\otimes\chi^{n} and Symkρχm\textnormal{Sym}^{k}\rho\otimes\chi^{m} have the same self-dual restriction to G𝐅q(x)fG_{\mathbf{F}_{q}(x)_{f}}, so the cup product is an alternating perfect pairing: classes pair with themselves to be zero, and independent elements pair to be nonzero.

Note that it is impossible for one group to be 11-dimensional and the other to be 22-dimensional; for instance, if γd(n1)\gamma\mid d(n-1), γd(n+k)\gamma\mid d(n+k), and γd(m1)\gamma\mid d(m-1), then

γd(m1)+d(n+k)d(n1)=d(m+k).\gamma\mid d(m-1)+d(n+k)-d(n-1)=d(m+k).

Hence the only case remaining is when H1(f,Symkρχn)H^{1}(f,\textnormal{Sym}^{k}\rho\otimes\chi^{n}) or H1(f,Symkρχm)H^{1}(f,\textnormal{Sym}^{k}\rho\otimes\chi^{m}) is 0-dimensional, in which case the claim is vacuously true. ∎

6.3. Detecting vanishing of local cup product

Suppose γd\gamma\mid d, and f(x)f(x) factors in 𝐅qγ[x]\mathbf{F}_{q^{\gamma}}[x] as f1(x)fγ(x)f_{1}(x)\cdots f_{\gamma}(x), arranged so that frobfi=fi+1\operatorname{frob}f_{i}=f_{i+1} for all ii. For n1n\geq 1 define

gn(x)=i=1γfi(x)q(i1)(n1),g_{n}(x)=\prod_{i=1}^{\gamma}f_{i}(x)^{q^{(i-1)(n-1)}},

and set

Kn:=𝐅qγ(gn).K_{n}:=\mathbf{F}_{q^{\gamma}}(\sqrt[\ell]{g_{n}}).

Then KnK_{n} is a (𝐙/𝐙)(\mathbf{Z}/\ell\mathbf{Z})-extension of 𝐅qγ(x)\mathbf{F}_{q^{\gamma}}(x), Galois over 𝐅q(x)\mathbf{F}_{q}(x), and unramified away from ff and \infty. Note that g1(x)=f(x)g_{1}(x)=f(x) and so K1=𝐅qγ(C)K_{1}=\mathbf{F}_{q^{\gamma}}(C). Also,

gn+γ(x)gn(x)=i=1γfi(x)(qγ(i1)1)q(i1)(n1)\frac{g_{n+\gamma}(x)}{g_{n}(x)}=\prod_{i=1}^{\gamma}f_{i}(x)^{(q^{\gamma(i-1)}-1)q^{(i-1)(n-1)}}

is an \ell-th power in 𝐅qγ(x)\mathbf{F}_{q^{\gamma}}(x) because qγ(i1)1modq^{\gamma(i-1)}\equiv 1\bmod\ell. Hence Kn=Kn+γK_{n}=K_{n+\gamma}, so KnK_{n} is well-defined for n𝐙/γ𝐙n\in\mathbf{Z}/\gamma\mathbf{Z}. The fields KnK_{n} can also be produced using explicit class field theory for the rational function field 𝐅qγ(x)\mathbf{F}_{q^{\gamma}}(x), as in [6] for example.

Lemma 6.6.

Suppose γd\gamma\mid d. For all nn, there exists a nonzero cocycle anH1(χn)a_{n}\in H^{1}(\chi^{n}) with kernel field KnK_{n}.

Proof.

We have

frob(gn)qn1=(f1q(γ1)(n1)i=2γfiq(i2)(n1))qn1=f1qγ(n1)1gn.\operatorname{frob}(g_{n})^{q^{n-1}}=\left(f_{1}^{q^{(\gamma-1)(n-1)}}\prod_{i=2}^{\gamma}f_{i}^{q^{(i-2)(n-1)}}\right)^{q^{n-1}}=f_{1}^{q^{\gamma(n-1)}-1}g_{n}.

Writing qγ(n1)1=jq^{\gamma(n-1)}-1=j\ell for some integer jj, we have frob(gn)qn1=f1jgn\operatorname{frob}(g_{n})^{q^{n-1}}=f_{1}^{j\ell}g_{n}. We therefore have frob(gn)qn1=ζtf1jgn\operatorname{frob}(\sqrt[\ell]{g_{n}})^{q^{n-1}}=\operatorname{\zeta}^{t}f_{1}^{j}\sqrt[\ell]{g_{n}} for some integer tt. If we let τGal(Kn/𝐅qγ(x))\tau\in\textnormal{Gal}(K_{n}/\mathbf{F}_{q^{\gamma}}(x)) denote the automorphism sending gnζgn\sqrt[\ell]{g_{n}}\mapsto\operatorname{\zeta}\sqrt[\ell]{g_{n}} then

frob(τ(gn))qn1=frob(ζgn)qn1=ζqn+tf1jgn=τqn(ζtf1jgn)=τqn(frob(gn))qn1,\operatorname{frob}(\tau(\sqrt[\ell]{g_{n}}))^{q^{n-1}}=\operatorname{frob}(\operatorname{\zeta}\sqrt[\ell]{g_{n}})^{q^{n-1}}=\operatorname{\zeta}^{q^{n}+t}f_{1}^{j}\sqrt[\ell]{g_{n}}=\tau^{q^{n}}(\operatorname{\zeta}^{t}f_{1}^{j}\sqrt[\ell]{g_{n}})=\tau^{q^{n}}(\operatorname{frob}(\sqrt[\ell]{g_{n}}))^{q^{n-1}},

and since gcd(q,)=1\gcd(q,\ell)=1, we can conclude frobτ=τqnfrob\operatorname{frob}\circ\tau=\tau^{q^{n}}\circ\operatorname{frob}.

Analogously to Definition 4.3, let ρn:Gal(Kn/𝐅q(x))GL2(𝐅)\rho_{n}:\textnormal{Gal}(K_{n}/\mathbf{F}_{q}(x))\to\textnormal{GL}_{2}(\mathbf{F}_{\ell}) be the representation defined by

ρn(frob)=(qn001),ρn(τ)=(1101)\rho_{n}(\operatorname{frob})=\begin{pmatrix}q^{n}&0\\ 0&1\end{pmatrix},\qquad\rho_{n}(\tau)=\begin{pmatrix}1&1\\ 0&1\end{pmatrix}

for τ\tau as above; this gives a well-defined representation because frobτ=τqnfrob\operatorname{frob}\circ\tau=\tau^{q^{n}}\circ\operatorname{frob}. This extends to a representation of G𝐅q(x)G_{\mathbf{F}_{q}(x)} that factors through Gal(Kn/𝐅q(x))\textnormal{Gal}(K_{n}/\mathbf{F}_{q}(x)). Then ρn\rho_{n} is an extension of 11 by χn\chi^{n} corresponding to a cocycle anH1(χn)a_{n}\in H^{1}(\chi^{n}) with kernel field KnK_{n}. Since Kn/𝐅q(x)K_{n}/\mathbf{F}_{q}(x) is unramified away from ff and \infty we in fact have aH1S(χn)a\in H^{1}_{S}(\chi^{n}). ∎

We finally reach the proof of Theorem 1.13. Recall that for n=2,,γ1n=2,\ldots,\gamma-1 we have

hn(x):=i=1γfi(x)q(i1)(γn)1=gγ+1n(x)f(x).h_{n}(x):=\prod_{i=1}^{\gamma}f_{i}(x)^{q^{(i-1)(\gamma-n)}-1}=\frac{g_{\gamma+1-n}(x)}{f(x)}.

See 1.13

Proof.

Note that F02F_{0}^{2} does not have a Frob\operatorname{Frob} eigenvector by Proposition 1.10(a) and (b). Since we assume γd\gamma\mid d, F11F_{1}^{1} has a Frob\operatorname{Frob} eigenvector by Proposition 1.10(a); if it were a rooftop then F10F^{1}_{0} would contain a Frob\operatorname{Frob} eigenvector by Proposition 1.10(c), but this contradicts Proposition 1.10(a). Hence F12F_{1}^{2} has a Frob\operatorname{Frob} eigenvector. Correspondingly, we have 0𝒯0\notin\mathcal{T} and 1𝒯1\in\mathcal{T}.

Now suppose n0,1modγn\not\equiv 0,1\bmod\gamma. By Lemma 5.5, H1S(χ1n)H^{1}_{S}(\chi^{1-n}) is one-dimensional, and by Lemma 6.6 it is spanned by a cocycle a1na_{1-n} with kernel field K1nK_{1-n}. By Proposition 5.7, Fn2F_{n}^{2} has an eigenvector if and only if H1S(ρχ1n)H^{1}_{S}(\rho\otimes\chi^{1-n}) has a nontrivial element mapping to a1na_{1-n}. By Proposition 6.4, such an element exists if and only if resf(a1n)𝐛=0\textnormal{res}_{f}(a_{1-n})\cup\mathbf{b}=0, and by Lemma 6.5 it suffices to check whether resf(a1n)\textnormal{res}_{f}(a_{1-n}) is in the span of 𝐛\mathbf{b}. By Lemma 5.3, we know that H1(f,ρχ1n)H^{1}(f,\rho\otimes\chi^{1-n}) is spanned by 𝐮𝐫\mathbf{ur}, which defines an extension with nontrivial residue field degree, and by 𝐛\mathbf{b}, which has kernel field 𝐅qγ(C)f\mathbf{F}_{q^{\gamma}}(C)_{f}. Hence we can detect whether resf(a1n)\textnormal{res}_{f}(a_{1-n}) is in the span of 𝐛\mathbf{b} by checking whether K1nK_{1-n} and 𝐅qγ(C)\mathbf{F}_{q^{\gamma}}(C) have the same completion at a prime over ff. Taking the place f1f_{1} above ff in 𝐅qγ(x)\mathbf{F}_{q^{\gamma}}(x), this is equivalent by Kummer theory to checking that hn(x)=gγ+1n(x)/f(x)h_{n}(x)=g_{\gamma+1-n}(x)/f(x) is an \ell-th power in the residue field 𝐅qγ[x]/(f1(x))\mathbf{F}_{q^{\gamma}}[x]/(f_{1}(x)). ∎

Remark 6.7.

Proposition 1.10 and Theorem 1.11 may also be proven using an argument similar to that of Theorem 1.13, that is, by combining Proposition 5.7 with the cohomological lifting conditions developed in Section 6.2. While the current proof of Theorem 1.11 invokes the Weil pairing, the cohomological proof instead uses the Poitou-Tate exact sequence [9, (8.6.10)] together with the observation that Symk1ρχ1n\textnormal{Sym}^{k-1}\rho\otimes\chi^{1-n} and Symk1ρχn+1k\textnormal{Sym}^{k-1}\rho\otimes\chi^{n+1-k} are cohomological duals.

7. Consequences of lifting conditions

Using the relations set up in the previous sections, we are now reduced to an essentially combinatorial problem: under the constraints described in Section 1.2, what are the possibilities for the set of pairs (n,k)(n,k) such that FnkF_{n}^{k} has an eigenvector? In particular, following Theorem 1.9, we are interested in the \ell-rank of the divisor class group r(C)r_{\ell}(C), which equals the number of kk for which Fk1kF_{k-1}^{k} contains a Frob\operatorname{Frob} eigenvector. In this section we prove all of the constraints on r(C)r_{\ell}(C) mentioned in Section 1.1.

For n𝐙/γ𝐙n\in\mathbf{Z}/\gamma\mathbf{Z}, let 0kn10\leq k_{n}\leq\ell-1 denote the smallest integer such that for all kn<k1k_{n}<k^{\prime}\leq\ell-1, there is no Frob\operatorname{Frob} eigenvector in FnkF_{n}^{k^{\prime}}. In other words, if kn1k_{n}\geq 1 then FnknF_{n}^{k_{n}} is a rooftop, and if kn=0k_{n}=0 then there is no value of kk for which FnkF_{n}^{k} has a Frob\operatorname{Frob} eigenvector. We will say that knk_{n} is the “rooftop height” over nn. If we take an integer representative n{0,1,,γ1}n\in\{0,1,\ldots,\gamma-1\}, then the number of 1kkn1\leq k\leq k_{n} with k1nmodγk-1\equiv n\bmod\gamma is equal to

c(n):=kn1nγ+1.c(n):=\left\lfloor\frac{k_{n}-1-n}{\gamma}\right\rfloor+1.

Following the visual interpretation as described in Remark 2.7, c(n)c(n) counts the number of dark grey circles in column nn. Since k0=0k_{0}=0 by Proposition 1.10(a), we have

r(C)=c(1)+c(2)++c(γ1)r_{\ell}(C)=c(1)+c(2)+\cdots+c(\gamma-1)

by Theorem 1.9, so c(n)c(n) counts the number of contributions to r(C)r_{\ell}(C) from nn. For all 0kn10\leq k_{n}\leq\ell-1 and 0nγ10\leq n\leq\gamma-1, we have 0c(n)1γ0\leq c(n)\leq\frac{\ell-1}{\gamma}; specifically, if kn=0k_{n}=0 then c(n)=0c(n)=0, and if kn=1k_{n}=\ell-1 then c(n)=1γc(n)=\frac{\ell-1}{\gamma}.

The proofs of Theorem 1.3 and Theorem 1.6 have many similar features: we will first prove the upper bounds in both theorems, then the lower bounds, then the parity constraints.

7.1. Upper bounds

We first prove the upper bound in Theorem 1.3, namely

r(C)B:=(gcd(d,γ)1)1γ.r_{\ell}(C)\leq B:=(\gcd(d,\gamma)-1)\frac{\ell-1}{\gamma}.

We have kn1k_{n}\geq 1 if and only if γdn\gamma\mid dn and γn\gamma\nmid n by Proposition 1.10(a). There are gcd(d,γ)1\gcd(d,\gamma)-1 such values of n𝐙/γ𝐙n\in\mathbf{Z}/\gamma\mathbf{Z}, and for each of these we have c(n)1γc(n)\leq\frac{\ell-1}{\gamma}. For all other values of nn we have kn=0k_{n}=0 and therefore c(n)=0c(n)=0. Combining these bounds gives the desired result. (BB counts the set of pairs (n,k)(n,k) for which FnkF_{n}^{k} is non-empty and nk1modγn\equiv k-1\bmod\gamma; in the visual interpretation of Remark 2.7, this corresponds to circles in cells that are either light or dark gray.)

We can similarly prove the upper bound in Theorem 1.6: if 3γd3\leq\gamma\mid d then

r(C)B:=|𝒯|1γ.r_{\ell}(C)\leq B^{\prime}:=|\mathcal{T}|\frac{\ell-1}{\gamma}.

We have kn2k_{n}\geq 2 if and only if n𝒯n\in\mathcal{T} by Theorem 1.13, and for each of these nn we have c(n)1γc(n)\leq\frac{\ell-1}{\gamma} as above. We have c(0)=0c(0)=0 as before, and for all other n𝒯n\notin\mathcal{T} we have n1n\geq 1 and kn1k_{n}\leq 1 so that c(n)=0c(n)=0. Combining these bounds gives the desired result.

7.2. Lower bounds

We begin by proving a general lemma that will be useful in producing lower bounds. For fixed n,kn,k, the diagonal containing FnkF_{n}^{k} is the set of all Fn+ik+iF_{n+i}^{k+i} for i𝐙i\in\mathbf{Z} with 1k+i11\leq k+i\leq\ell-1. Equivalently, each diagonal is determined by a constant value of kn𝐙/γ𝐙k-n\in\mathbf{Z}/\gamma\mathbf{Z}. We have strong constraints on how rooftops can be arranged across diagonals: Proposition 1.10(d) states that any given diagonal can contain at most one non-maximal rooftop, and Proposition 1.10(c) limits which diagonals are allowed to contain rooftops at all. So if we can find many sets FnkF_{n}^{k} that contain Frob\operatorname{Frob} eigenvectors among a small collection of diagonals right below the main diagonal Fk1kF_{k-1}^{k} (the circles in LABEL:fig:lifting_charts), the pigeonhole principle will ensure that only a few of them can be rooftops; the rest must all lift past this main diagonal and hence contribute to r(C)r_{\ell}(C).

Lemma 7.1.

Let m{0,,γ1}m\in\{0,\ldots,\gamma-1\}, and suppose there are rr values of n{m,,γ1}n\in\{m,\ldots,\gamma-1\} such that Fnnm+1F_{n}^{n-m+1} has a Frob\operatorname{Frob} eigenvector. Then r(C)rr_{\ell}(C)\geq r.

Proof.

If m=0m=0 this is immediate from Theorem 1.9 (we have no Frob\operatorname{Frob} eigenvectors in F01F_{0}^{1} by Proposition 1.10(a)), so from now on assume m1m\geq 1. Define the sets

R\displaystyle R :={n{m,,γ1}:Fnnm+1 contains a Frob eigenvector},\displaystyle:=\{n\in\{m,\ldots,\gamma-1\}:F_{n}^{n-m+1}\text{ contains a }\operatorname{Frob}\text{ eigenvector}\},
S\displaystyle S :={n{1,,m1}:γdn}.\displaystyle:=\{n\in\{1,\ldots,m-1\}:\gamma\mid dn\}.

Finally, let LRSL\subseteq R\cup S be the set of all nRSn\in R\cup S for which nkn>1n-k_{n}>-1, where knk_{n} is the rooftop height over nn; equivalently, the set LL consists of all nRSn\in R\cup S with c(n)=0c(n)=0. We will prove that |L||S||L|\leq|S|.

To this end, let nLn\in L. By Proposition 1.10(c) we cannot have nkn=0n-k_{n}=0, so in fact nkn1n-k_{n}\geq 1. We also have nknm1n-k_{n}\leq m-1: for nRn\in R this follows from knnm+1k_{n}\geq n-m+1, and for nSn\in S this follows from kn0k_{n}\geq 0 and nm1n\leq m-1. Again by Proposition 1.10(c) we have γd(nkn)\gamma\mid d(n-k_{n}). Taken together, we can conclude that nknSn-k_{n}\in S. In other words, all nLn\in L lie on one of a set of |S||S| diagonals.

Now for each nLn\in L, we have kn1k_{n}\geq 1: this follows by Proposition 1.10(a) if nSn\in S, and by knnm+1k_{n}\geq n-m+1 if nRn\in R. So consider the rooftop FnknF_{n}^{k_{n}}. Since nkn>1n-k_{n}>-1 and nγ1n\leq\gamma-1 we have kn<γ1k_{n}<\gamma\leq\ell-1, so FnknF_{n}^{k_{n}} is a non-maximal rooftop. By Proposition 1.10(d), it is impossible to have nkn=nknn-k_{n}=n^{\prime}-k_{n^{\prime}} for any distinct n,nLn,n^{\prime}\in L. So by the pigeonhole principle, there are at most |S||S| elements in LL.

In conclusion, there are at least rr elements in RSR\cup S satisfying knn+1k_{n}\geq n+1, so c(n)1c(n)\geq 1. We can conclude that r(C)rr_{\ell}(C)\geq r. ∎

Lower bound of Theorem 1.3: We will show that B1B\geq 1 implies r(C)1r_{\ell}(C)\geq 1. If B1B\geq 1 then gcd(d,γ)>1\gcd(d,\gamma)>1, so there exists n𝐙/γ𝐙n\in\mathbf{Z}/\gamma\mathbf{Z} with γdn\gamma\mid dn and γn\gamma\nmid n. For this value of nn, Fn1F_{n}^{1} has an eigenvector by Proposition 1.10(a). Taking m=nm=n in Lemma 7.1 proves r(C)1r_{\ell}(C)\geq 1.222Alternatively, simply use the fact that any generalized eigenspace must contain at least one true eigenvector.

First lower bound of Theorem 1.6: Suppose 3γd3\leq\gamma\mid d. If B=1B^{\prime}=1, then since BBB\geq B^{\prime} we have r(C)Br_{\ell}(C)\geq B^{\prime} by the lower bound of Theorem 1.3. So it is sufficient to assume

B:=|𝒯|1γ2B^{\prime}:=|\mathcal{T}|\frac{\ell-1}{\gamma}\geq 2

and prove r(C)2r_{\ell}(C)\geq 2. Note that the assumption γd\gamma\mid d implies Fn1F_{n}^{1} has an eigenvector for all 1nγ11\leq n\leq\gamma-1 by Proposition 1.10(a). We always have 1𝒯1\in\mathcal{T} by definition of 𝒯\mathcal{T}, which implies by Theorem 1.13 that F12F_{1}^{2} has an eigenvector.

First consider the case |𝒯|=1|\mathcal{T}|=1, so that Fn1F_{n}^{1} is a rooftop for all 2nγ12\leq n\leq\gamma-1. Since 2=212=2\cdot 1, we can apply Theorem 1.12 to conclude that F12F_{1}^{2} is not a rooftop and thus F13F_{1}^{3} has a Frob\operatorname{Frob} eigenvector. Now since B2B^{\prime}\geq 2 but |𝒯|=1|\mathcal{T}|=1 we can conclude 12γ\ell-1\geq 2\gamma. So for each 2iγ12\leq i\leq\gamma-1, if F1i+1F_{1}^{i+1} were a rooftop, then it would be a non-maximal rooftop; since F1i1F_{1-i}^{1} is also a non-maximal rooftop, this contradicts Proposition 1.10(d). Therefore F1γ+1F_{1}^{\gamma+1} has an eigenvector, but is not a rooftop by Proposition 1.10(c). Hence k1γ+2k_{1}\geq\gamma+2, which implies c(1)2c(1)\geq 2 and hence r(C)2r_{\ell}(C)\geq 2.

Now consider the case |𝒯|2|\mathcal{T}|\geq 2, so Fn2F_{n}^{2} has an eigenvector for some n{2,,γ1}n\in\{2,\ldots,\gamma-1\}. Since Fn2F_{n}^{2} and Fn11F_{n-1}^{1} both have eigenvectors, we have r(C)2r_{\ell}(C)\geq 2 by Lemma 7.1, taking m=n1m=n-1. Together with the previous case, we see that r(C)2r_{\ell}(C)\geq 2 whenever B2B^{\prime}\geq 2.

Second lower bound of Theorem 1.6: Assuming γ\gamma is even and 1+γ2𝒯1+\frac{\gamma}{2}\in\mathcal{T}, we will show that r(C)3r_{\ell}(C)\geq 3. Since 1+γ2𝒯1+\frac{\gamma}{2}\in\mathcal{T}, we can conclude that F1+γ/22F_{1+\gamma/2}^{2} has a Frob\operatorname{Frob} eigenvector by Theorem 1.13. Since 22(1+γ2)modγ2\equiv 2(1+\frac{\gamma}{2})\bmod\gamma, we are in the setting of Theorem 1.12, and can conclude that F1+γ/22F_{1+\gamma/2}^{2} is not a rooftop: thus F1+γ/23F_{1+\gamma/2}^{3} also has a Frob\operatorname{Frob} eigenvector. Further, since F1+γ/21F_{1+\gamma/2}^{1} is not a rooftop, neither is Fγ/21F_{\gamma/2}^{1} by Theorem 1.11. Hence F1+γ/23F_{1+\gamma/2}^{3}, Fγ/22F_{\gamma/2}^{2}, and Fγ/211F_{\gamma/2-1}^{1} all contain eigenvectors, so r(C)3r_{\ell}(C)\geq 3 by Lemma 7.1.

7.3. Parity

We first prove the parity constraint of Theorem 1.3. Let

S:={n{1,,γ1}:γdn,kn1}.S:=\{n\in\{1,\ldots,\gamma-1\}:\gamma\mid dn,\ k_{n}\neq\ell-1\}.

By Proposition 1.10(a), it is equivalent to say that SS is the set of n{0,,γ1}n\in\{0,\ldots,\gamma-1\} with 1kn21\leq k_{n}\leq\ell-2. For each nSn\in S, let nn^{\vee} denote the unique value in {0,,γ1}\{0,\ldots,\gamma-1\} satisfying nknnmodγn^{\vee}\equiv k_{n}-n\bmod\gamma. By Theorem 1.11 we have kn=knk_{n^{\vee}}=k_{n} for all nSn\in S, so nnn\mapsto n^{\vee} is an involution on SS and c(n)=c(n)c(n)=c(n^{\vee}) for all nSn\in S. We can write

r(C)\displaystyle r_{\ell}(C) =n{1,,γ1},γdn(1γ(1γc(n)))\displaystyle=\sum_{\begin{subarray}{c}n\in\{1,\ldots,\gamma-1\},\\ \gamma\mid dn\end{subarray}}\left(\frac{\ell-1}{\gamma}-\left(\frac{\ell-1}{\gamma}-c(n)\right)\right)
=BnS(1γc(n)),\displaystyle=B-\sum_{n\in S}\left(\frac{\ell-1}{\gamma}-c(n)\right),

since in the first equality we only remove terms with c(n)=0c(n)=0, and in the second equality we only remove terms with c(n)=1γc(n)=\frac{\ell-1}{\gamma}. If nnn\neq n^{\vee}, then nn and nn^{\vee} contribute equal quantities to the sum. So to prove r(C)Bmod2r_{\ell}(C)\equiv B\bmod 2, it suffices to show that the terms indexed by fixed points of the involution are even.

Suppose n=nn=n^{\vee}, so that kn2nmodγk_{n}\equiv 2n\bmod\gamma. By definition of c(n)c(n) we have

(c(n)1)γkn1n<c(n)γ.(c(n)-1)\gamma\leq k_{n}-1-n<c(n)\gamma.

Since we additionally have 1n<γ1\leq n<\gamma we can conclude

(c(n)2)γ+2n+1<kn<c(n)γ+2n,(c(n)-2)\gamma+2n+1<k_{n}<c(n)\gamma+2n,

so that in fact kn=(c(n)1)γ+2nk_{n}=(c(n)-1)\gamma+2n. We can therefore write

1γc(n)=1kn+2nγ+1.\frac{\ell-1}{\gamma}-c(n)=\frac{\ell-1-k_{n}+2n}{\gamma}+1.

By Theorem 1.12, knk_{n} is odd, and therefore 1kn+2n\ell-1-k_{n}+2n is odd. Thus 1γc(n)\frac{\ell-1}{\gamma}-c(n) is even as desired.

We now assume 3γd3\leq\gamma\mid d and prove the parity constraint of Theorem 1.6, namely that r(C)Bmod2r_{\ell}(C)\equiv B^{\prime}\bmod 2. Following Theorem 1.3, it suffices to prove that BBmod2B^{\prime}\equiv B\bmod 2. If γ\gamma is odd, then BB and BB^{\prime} are both even because 1γ\frac{\ell-1}{\gamma} is even. So suppose instead that γ\gamma is even. By Theorem 1.11, Fn1F_{n}^{1} is a rooftop if and only if F1n1F_{1-n}^{1} is a rooftop; since γ\gamma is even, we always have 1nnmodγ1-n\not\equiv n\bmod\gamma, and so the rooftops FnkF_{n}^{k} with k=1k=1 always come in pairs. Now |𝒯||\mathcal{T}| counts the number of nn for which Fn2F_{n}^{2} contains an eigenvector, which equals γ1\gamma-1 minus the number of rooftops with k=1k=1. Hence |𝒯|γ1mod2|\mathcal{T}|\equiv\gamma-1\bmod 2. Since we are assuming γd\gamma\mid d we have |𝒯|gcd(d,γ)1mod2|\mathcal{T}|\equiv\gcd(d,\gamma)-1\bmod 2, which implies BBmod2B^{\prime}\equiv B\bmod 2.

7.4. Proof of Proposition 1.5

Finally, we assume gcd(d,γ)=2\gcd(d,\gamma)=2 and prove that r(C)=1γr_{\ell}(C)=\frac{\ell-1}{\gamma}. By Proposition 1.10(a), Fn1F_{n}^{1} has an eigenvector only for n=γ2n=\frac{\gamma}{2}. We will prove that the rooftop height over nn is kn=1k_{n}=\ell-1, which will imply r(C)=c(n)=1γr_{\ell}(C)=c(n)=\frac{\ell-1}{\gamma} as desired.

Suppose 1k21\leq k\leq\ell-2 is such that FnkF_{n}^{k} is a rooftop. By Proposition 1.10(c), this implies k0modγ2k\equiv 0\bmod{\frac{\gamma}{2}} and kγ2modγk\not\equiv\frac{\gamma}{2}\bmod{\gamma}; equivalently, kk is a multiple of γ\gamma. But then k2nmodγk\equiv 2n\bmod\gamma and kk is even, so FnkF_{n}^{k} is not a rooftop by Theorem 1.12. Hence FnkF_{n}^{k} is not a rooftop for any 1k21\leq k\leq\ell-2.

Appendix A Proof of Lemma 2.5

We prove that Frobη=qηFrob\operatorname{Frob}\circ\operatorname{\eta}=q\operatorname{\eta}\circ\operatorname{Frob}.

Lemma A.1.

For all integers qq and 0k10\leq k\leq\ell-1,

i=11j=0i(1)j+1i(ij)(qjk)={(1)k+1qkk>00k=0.\sum_{i=1}^{\ell-1}\sum_{j=0}^{i}\frac{(-1)^{j+1}}{i}\binom{i}{j}\binom{qj}{k}=\left\{\begin{array}[]{ll}\displaystyle\frac{(-1)^{k+1}q}{k}&k>0\\ 0&k=0.\end{array}\right.
Proof.

For k=0k=0, this follows from the fact that the alternating sum of (ij)\binom{i}{j} for 0ji0\leq j\leq i is equal to 0. So from now on we assume k1k\geq 1. We will first prove the identity

(30) i=01j=0i(1)ij(ij)(qjk)(xi)=(qxk)\displaystyle\sum_{i=0}^{\ell-1}\sum_{j=0}^{i}(-1)^{i-j}\binom{i}{j}\binom{qj}{k}\binom{x}{i}=\binom{qx}{k}

as an equality in 𝐐[x]\mathbf{Q}[x]. First consider the value of

j=0i(1)ij(ij)(qjk)\sum_{j=0}^{i}(-1)^{i-j}\binom{i}{j}\binom{qj}{k}

for i>ki>k. This equals the coefficient of tkt^{k} in ((t+1)q1)i((t+1)^{q}-1)^{i}. But (t+1)q1(t+1)^{q}-1 is a multiple of tt and so ((t+1)q1)i((t+1)^{q}-1)^{i} is a multiple of tit^{i}, and hence the coefficient of tkt^{k} is zero. Hence only terms with iki\leq k contribute to the left-hand side of Eq. 30, so both sides of the desired identity are polynomials of degree at most kk.

If x{0,1,,1}x\in\{0,1,\ldots,\ell-1\}, we can compute the coefficient of tkt^{k} in (((t+1)q1)+1)x=(t+1)qx(((t+1)^{q}-1)+1)^{x}=(t+1)^{qx}, obtaining

i=0x(xi)j=0i(1)ij(ij)(qjk)=(qxk).\sum_{i=0}^{x}\binom{x}{i}\sum_{j=0}^{i}(-1)^{i-j}\binom{i}{j}\binom{qj}{k}=\binom{qx}{k}.

Since x1x\leq\ell-1, and (xi)=0\binom{x}{i}=0 for i>xi>x, we can sum ii from 0 to 1\ell-1 without changing the value, giving us the desired equality for these specified values of xx. We therefore have two polynomials of degree at most kk in 𝐐[x]\mathbf{Q}[x] with equal values at >k\ell>k points, and therefore the polynomials are equal.

Now compare the coefficient of xx in each side of Eq. 30. The coefficient of xx in (xi)\binom{x}{i} is 0 if i=0i=0 and (1)i+1/i(-1)^{i+1}/i if i1i\geq 1, so we obtain the desired result. ∎

Lemma A.1 is an identity over 𝐐\mathbf{Q}, but the only denominators that occur are coprime to \ell; it therefore descends to an identity over 𝐅\mathbf{F}_{\ell}. It is in this form that we apply it below.

Proof of Lemma 2.5.

Using the definition

η=i=12i1(1ζ)i=i=12(1)i+1i(ζ1)i,\operatorname{\eta}=-\sum_{i=1}^{\ell-2}i^{-1}(1-\operatorname{\zeta})^{i}=\sum_{i=1}^{\ell-2}\frac{(-1)^{i+1}}{i}(\operatorname{\zeta}-1)^{i},

we have

Frobη\displaystyle\operatorname{Frob}\circ\operatorname{\eta} =i=11(1)i+1iFrob(ζ1)i\displaystyle=\sum_{i=1}^{\ell-1}\frac{(-1)^{i+1}}{i}\operatorname{Frob}\circ(\operatorname{\zeta}-1)^{i}
=i=11(1)i+1i(ζq1)iFrob\displaystyle=\sum_{i=1}^{\ell-1}\frac{(-1)^{i+1}}{i}(\operatorname{\zeta}^{q}-1)^{i}\circ\operatorname{Frob}
=i=11(1)i+1ij=0i(ij)(1)ij(ζ1+1)qjFrob\displaystyle=\sum_{i=1}^{\ell-1}\frac{(-1)^{i+1}}{i}\sum_{j=0}^{i}\binom{i}{j}(-1)^{i-j}(\operatorname{\zeta}-1+1)^{qj}\circ\operatorname{Frob}
=i=11j=0i(1)j+1i(ij)k=0qj(qjk)(ζ1)kFrob.\displaystyle=\sum_{i=1}^{\ell-1}\sum_{j=0}^{i}\frac{(-1)^{j+1}}{i}\binom{i}{j}\sum_{k=0}^{qj}\binom{qj}{k}(\operatorname{\zeta}-1)^{k}\circ\operatorname{Frob}.

Now we have (ζ1)k=0(\operatorname{\zeta}-1)^{k}=0 for all k>1k>\ell-1 and (qjk)=0\binom{qj}{k}=0 for all k>qjk>qj, so the sum over kk can be indexed from 0 to 1\ell-1 without changing the value (in both cases, only terms with kmin{qj,1}k\leq\min\{qj,\ell-1\} contribute). Using Lemma A.1, we can conclude:

Frobη\displaystyle\operatorname{Frob}\circ\operatorname{\eta} =k=01i=11j=0i(1)j+1i(ij)(qjk)(ζ1)kFrob\displaystyle=\sum_{k=0}^{\ell-1}\sum_{i=1}^{\ell-1}\sum_{j=0}^{i}\frac{(-1)^{j+1}}{i}\binom{i}{j}\binom{qj}{k}(\operatorname{\zeta}-1)^{k}\circ\operatorname{Frob}
=k=01((1)k+1qk)(ζ1)kFrob\displaystyle=\sum_{k=0}^{\ell-1}\left(\frac{(-1)^{k+1}q}{k}\right)(\operatorname{\zeta}-1)^{k}\circ\operatorname{Frob}
=qηFrob.\displaystyle=q\operatorname{\eta}\circ\operatorname{Frob}.\qed

References

  • [1] Vishal Arul “Explicit Division and Torsion Points on Superelliptic Curves and Jacobians” Thesis (Ph.D.)–Massachusetts Institute of Technology ProQuest LLC, Ann Arbor, MI, 2020, pp. (no paging) URL: http://gateway.proquest.com/openurl?url_ver=Z39.88-2004&rft_val_fmt=info:ofi/fmt:kev:mtx:dissertation&res_dat=xri:pqm&rft_dat=xri:pqdiss:28216982
  • [2] Frank Calegari and Matthew Emerton “On the ramification of Hecke algebras at Eisenstein primes” In Invent. Math. 160.1, 2005, pp. 97–144 DOI: 10.1007/s00222-004-0406-z
  • [3] H. Cohen and H. W. Lenstra “Heuristics on class groups of number fields” In Number Theory Noordwijkerhout 1983 Berlin, Heidelberg: Springer Berlin Heidelberg, 1984, pp. 33–62
  • [4] Gunther Cornelissen “Two-torsion in the Jacobian of hyperelliptic curves over finite fields” In Arch. Math. (Basel) 77.3, 2001, pp. 241–246 DOI: 10.1007/PL00000487
  • [5] Jordan S. Ellenberg, Wanlin Li and Mark Shusterman “Nonvanishing of hyperelliptic zeta functions over finite fields” In Algebra Number Theory 14.7, 2020, pp. 1895–1909 DOI: 10.2140/ant.2020.14.1895
  • [6] D. R. Hayes “Explicit class field theory for rational function fields” In Trans. Amer. Math. Soc. 189, 1974, pp. 77–91 DOI: 10.2307/1996848
  • [7] Tomasz Jędrzejak “On the torsion of the Jacobians of superelliptic curves yq=xp+ay^{q}=x^{p}+a In J. Number Theory 145, 2014, pp. 402–425 DOI: 10.1016/j.jnt.2014.06.013
  • [8] J. S. Milne “Abelian Varieties” In Arithmetic Geometry New York, NY: Springer New York, 1986, pp. 103–150 DOI: 10.1007/978-1-4613-8655-1_5
  • [9] J. Neukirch, A. Schmidt and K. Wingberg “Cohomology of Number Fields”, Grundlehren der mathematischen Wissenschaften Springer Berlin Heidelberg, 2013 URL: https://books.google.ca/books?id=ZX8CAQAAQBAJ
  • [10] Bjorn Poonen “Rational points on varieties” 186, Graduate Studies in Mathematics American Mathematical Society, Providence, RI, 2017, pp. xv+337 DOI: 10.1090/gsm/186
  • [11] Bjorn Poonen and Edward F. Schaefer “Explicit descent for Jacobians of cyclic covers of the projective line” In J. Reine Angew. Math. 488, 1997, pp. 141–188
  • [12] Karl Schaefer and Eric Stubley “Class groups of Kummer extensions via cup products in Galois cohomology” In Trans. Amer. Math. Soc. 372, 2019, pp. 6927–6980 DOI: https://doi.org/10.1090/tran/7746
  • [13] Preston Wake and Carl Wang-Erickson “The rank of Mazur’s Eisenstein ideal” In Duke Math. J. 169.1, 2020, pp. 31–115 DOI: 10.1215/00127094-2019-0039
  • [14] Wojciech Wawrów “On torsion of superelliptic Jacobians” In J. Théor. Nombres Bordeaux 33.1, 2021, pp. 223–235 URL: http://jtnb.cedram.org/item?id=JTNB_2021__33_1_223_0
  • [15] Christian Wittmann “l-Class groups of cyclic function fields of degree l” In Finite Fields and Their Applications 13.2, 2007, pp. 327–347 DOI: https://doi.org/10.1016/j.ffa.2005.09.001