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

Homological stability for generalized Hurwitz spaces and Selmer groups in quadratic twist families over function fields

Jordan S. Ellenberg  and  Aaron Landesman
(Date: August 11, 2025)
Abstract.

We prove a version of the Bhargava-Kane-Lenstra-Poonen-Rains heuristics for Selmer groups of quadratic twist families of abelian varieties over global function fields. As a consequence, we derive a result towards the “minimalist conjecture" on Selmer ranks of abelian varieties in such families. More precisely, we show that the probabilities predicted in these two conjectures are correct to within an error term in the size of the constant field, qq, which goes to 0 as qq grows. Two key inputs are a new homological stability theorem for a generalized version of Hurwitz spaces parameterizing covers of punctured Riemann surfaces of arbitrary genus, and an expression of average sizes of Selmer groups in terms of the number of rational points on these Hurwitz spaces over finite fields.

Key words and phrases:
Bhargava-Kane-Lenstra-Poonen-Rains heuristics, the minimalist conjecture, quadratic twists, homological stability, big monodromy
2020 Mathematics Subject Classification:
Primary 11G05; Secondary 11G10, 14G15, 55N99

1. Introduction

For ν\nu a positive integer and AA an abelian variety over a global field KK, the ν\nu Selmer group of AA, denoted Selν(A)\operatorname{Sel}_{\nu}(A), is a group which sits in an exact sequence between the mod ν\nu Mordell-Weil group A(K)/νA(K)A(K)/\nu A(K) and the ν\nu torsion in the Tate-Shafarevich group (A)[ν]\Sha(A)[\nu]. These Selmer groups, unlike the other two terms in the exact sequence, are computationally approachable, and provide the most tractable means of obtaining information about the rank of AA and of (A)\Sha(A).

The Selmer group of an abelian variety can be thought of as a higher analogue of the class group of a number field. The behavior of the class group of a number field chosen at random from a specified family is the subject of the Cohen-Lenstra conjecture and its many subsequent generalizations. In the same way, the question “what does the ν\nu Selmer group of a random abelian variety look like?" is the subject of a suite of more recent conjectures. Conjectures predicting the distribution of Selmer groups were formulated in [PR12], when ν\nu is prime, and generalized to the case of composite ν\nu in [BKL+15, §5.7], see also [FLR23, §5.3.3]. We call these conjectures the “BKLPR heuristics.” Although the above papers state their conjectures in the context of the universal family parameterizing all elliptic curves, it is also natural to ask under what circumstances they apply to quadratic twist families ( [PR12, Remark 1.9].) Our main result is a proof of these conjectures over function fields of arbitrary genus, up to an error term in qq that approaches 0 as qq grows, in the case where the family of abelian varieties is the family of quadratic twists of a fixed abelian variety.

For \ell a suitably large prime, as an immediate consequence of our main result, we obtain a version of the minimalist conjecture for \ell^{\infty} Selmer ranks, which predicts that quadratic twists of a fixed elliptic curve have \ell^{\infty} Selmer rank 0 half the time, \ell^{\infty} Selmer rank 11 half the time, and \ell^{\infty} Selmer rank at least 22 zero percent of the time.

The approach of this paper is similar to that of [EVW16], which verifies a version of the Cohen-Lenstra heuristics over genus 0 function fields. As in [EVW16], one key input is a new homological stability theorem. This theorem, which is purely topological in nature, is used to bound the étale cohomology of relevant moduli spaces, whose 𝔽q\mathbb{F}_{q} points count elements of Selmer groups of quadratic twists of an abelian variety.

1.1. Main Results

To give an indication of the nature of the results we prove in this paper, we start with a very special case of Theorem 1.1.3 below, see 1.1.6. We now describe this special case informally. Let 𝔽q\mathbb{F}_{q} be a finite field of odd characteristic, AA be an abelian variety over the field 𝔽q(t)\mathbb{F}_{q}(t) with good reduction over \infty, and \ell an odd prime not dividing qq. For any squarefree polynomial f𝔽q[t]f\in\mathbb{F}_{q}[t] of even degree nn,111See 1.4.6 for a discussion on how to generalize this to the case that the degree, nn, is odd. we denote by AfA_{f} the quadratic twist of AA by the quadratic character of 𝔽q(t)\mathbb{F}_{q}(t) associated to ff. Write 𝔼nSelAf\mathbb{E}_{n}\operatorname{Sel}_{\ell}A_{f} for the average size of the \ell Selmer group of AfA_{f} as ff ranges over squarefree polynomials of degree nn which are coprime to the bad reduction locus of AA. Similarly, write 𝔼n,jSelAf\mathbb{E}_{n,j}\operatorname{Sel}_{\ell}A_{f} for the same average obtained from the base change A/𝔽qj(t)A/\mathbb{F}_{q^{j}}(t), so that the average is now over the squarefree polynomials in 𝔽qj(t)\mathbb{F}_{q^{j}}(t) coprime to bad reduction. Then, the Poonen-Rains heuristics assert that limn𝔼n,jSelAf=+1\lim_{n}\mathbb{E}_{n,j}\operatorname{Sel}_{\ell}A_{f}=\ell+1 for all jj. What we prove, subject to some modest conditions on AA and \ell, which will be specified in Theorem 1.1.3, is that

limjlimn𝔼n,jSelAf=+1.\lim_{j}\lim_{n}\mathbb{E}_{n,j}\operatorname{Sel}_{\ell}A_{f}=\ell+1.

We emphasize that the computation that limj𝔼n,jSelAf=+1\lim_{j}\mathbb{E}_{n,j}\operatorname{Sel}_{\ell}A_{f}=\ell+1, without first taking a limit in nn, is substantially easier, see § 1.6 for more on this issue. The contribution of the present paper is to understand, as in the BKLPR heuristics, what happens when nn goes to infinity with jj fixed, or, in other words, AA is defined over a specific global field 𝔽qj(t)\mathbb{F}_{q^{j}}(t).

Before getting to our main result we present yet another special case which is a bit simpler to state, but is already of significant interest. Let CC be a smooth proper geometrically connected curve over a finite field 𝔽q\mathbb{F}_{q} of odd characteristic and let UCU\subset C be a nonempty open subscheme with nonempty complement. Let ν\nu be an odd integer and AUA\to U be a polarized abelian scheme with polarization of degree prime to ν\nu. Let QTwistU/𝔽qn(𝔽qj)\operatorname{QTwist}^{n}_{U/\mathbb{F}_{q}}(\mathbb{F}_{q^{j}}) denote the groupoid of quadratic twists of A×Spec𝔽qSpec𝔽qjA\times_{\operatorname{Spec}\mathbb{F}_{q}}\operatorname{Spec}\mathbb{F}_{q^{j}}, ramified over a degree nn divisor contained in UU with nn even. (See 5.1.4 for a precise definition.) For xQTwistU/𝔽qn(𝔽qj)x\in\operatorname{QTwist}^{n}_{U/\mathbb{F}_{q}}(\mathbb{F}_{q^{j}}), we let AxA_{x} denote the corresponding quadratic twist. We use SelνBKLPR\operatorname{Sel}^{\operatorname{BKLPR}}_{\nu} for the predicted distribution of the ν\nu Selmer group, as given in [BKL+15]; see 2.2.1 for a brief definition. The following consequence of our main result says the BKLPR heuristics hold for quadratic twists of an elliptic curve with squarefree discriminant, up to an error that goes to 0 as qq grows.

Theorem 1.1.1.

With notation as above, suppose AA is a nonconstant elliptic curve with squarefree discriminant. Choose ν\nu and qq so that char𝔽q>3\operatorname{\operatorname{char}}\mathbb{F}_{q}>3 and ν\nu is prime to 6q6q. Let HH be a finitely generated /ν\mathbb{Z}/\nu\mathbb{Z}-module. Then

Prob(SelνBKLPRH)\displaystyle\operatorname{Prob}(\operatorname{Sel}^{\operatorname{BKLPR}}_{\nu}\simeq H) =limjlim supnnevenProb(Selν(Ax)H:xQTwistU/𝔽qn(𝔽qj))\displaystyle=\lim_{j\to\infty}\limsup_{\begin{subarray}{c}n\to\infty\\ n\hskip 2.84544pt\mathrm{even}\end{subarray}}\operatorname{Prob}(\operatorname{Sel}_{\nu}(A_{x})\simeq H:x\in\operatorname{QTwist}^{n}_{U/\mathbb{F}_{q}}(\mathbb{F}_{q^{j}}))
=limjlim infnnevenProb(Selν(Ax)H:xQTwistU/𝔽qn(𝔽qj)).\displaystyle=\lim_{j\to\infty}\liminf_{\begin{subarray}{c}n\to\infty\\ n\hskip 2.84544pt\mathrm{even}\end{subarray}}\operatorname{Prob}(\operatorname{Sel}_{\nu}(A_{x})\simeq H:x\in\operatorname{QTwist}^{n}_{U/\mathbb{F}_{q}}(\mathbb{F}_{q^{j}})).

We next state a more general theorem of which Theorem 1.1.1 is a consequence: Indeed, note that the tameness of A[ν]UA[\nu]\to U which we will assume in Theorem 1.1.2 holds in the setting of Theorem 1.1.1 from the assumption that qq is prime to 66 and the irreducibility assumption in Theorem 1.1.2 holds in the setting of Theorem 1.1.1 by [Zyw14, Proposition 2.7]. The remaining assumptions in Theorem 1.1.2 also automatically hold for any nonconstant elliptic curve of squarefree discriminant. Use notation as prior to Theorem 1.1.1.

Theorem 1.1.2.

With notation as above, choose an abelian scheme AA so that

(1.1) AA has multiplicative reduction with toric part of dimension 11 over some point of CC.

Choose ν\nu so that every prime ν\ell\mid\nu satisfies >2dimA+1\ell>2\dim A+1 and A[]×𝔽q𝔽¯qA[\ell]\times_{\mathbb{F}_{q}}\overline{\mathbb{F}}_{q} corresponds to a irreducible sheaf of /\mathbb{Z}/\ell\mathbb{Z} modules on U×𝔽q𝔽¯qU\times_{\mathbb{F}_{q}}\overline{\mathbb{F}}_{q}, ν\nu is prime to qq, and A[ν]A[\nu] is a tame finite étale cover of UU. Further assume that ν\nu is relatively prime to the order of the geometric component group of the Néron model of AA over CC, as defined in 5.2.2. We have

limjlim supnnevenProb(Selν(Ax)H:xQTwistU/𝔽qn(𝔽qj))\displaystyle\lim_{j\to\infty}\limsup_{\begin{subarray}{c}n\to\infty\\ n\hskip 2.84544pt\mathrm{even}\end{subarray}}\operatorname{Prob}(\operatorname{Sel}_{\nu}(A_{x})\simeq H:x\in\operatorname{QTwist}^{n}_{U/\mathbb{F}_{q}}(\mathbb{F}_{q^{j}})) =Prob(SelνBKLPRH),\displaystyle=\operatorname{Prob}(\operatorname{Sel}^{\operatorname{BKLPR}}_{\nu}\simeq H),

as well as the analogous statement with lim sup\limsup replaced with lim inf\liminf.

Theorem 1.1.2 is proven in § 10.2.2. We also explain in § 10.2.5 how the proof of Theorem 1.1.2 can be somewhat shortened in the case that ν\nu is prime.

Remark 1.1.1.

If we start with an abelian scheme over an affine curve over a number field KK, one can spread it out to an abelian scheme over an affine curve over a sufficiently small nonempty open Spec𝒪Spec𝒪K\operatorname{Spec}\mathscr{O}\subset\operatorname{Spec}\mathscr{O}_{K}. One can then deduce a version of Theorem 1.1.2 where one takes a limit over prime powers with characteristic avoiding finitely many primes, instead of restricting the characteristic to take a single fixed value, as in Theorem 1.1.2. See 9.2.4 for more on this. The key point is that the cohomology groups of the relevant moduli space will be independent of the geometric point of Spec𝒪\operatorname{Spec}\mathscr{O} we choose.

We next include some remarks on the relation between our results, the BKLPR heuristics, and the results of [EVW16].

Remark 1.1.2.

Theorem 1.1.2 can be thought of as a version of the conjectures of [BKL+15] over global function fields for quadratic twist families of abelian varieties. There are two respects in which our result does not precisely say that the BKLPR conjecture holds for such families. The first difference, and the more substantial one, is that we can’t show the probabilities we analyze agree with the BKLPR heuristics exactly, but only up to an error term that shrinks as the finite field gets larger and larger. The second difference is that BKLPR makes conjectures for \ell^{\infty} Selmer groups, while our results apply only to finite order Selmer groups. It seems likely the ideas in this paper could be extended to the case of \ell^{\infty} Selmer groups, and we think it would be quite interesting to do so.

The relationship between the theorems of the present paper and the BKLPR heuristics is analogous to the relationship between the results of [EVW16] and the Cohen-Lenstra heuristics. The connection between the two papers is discussed further in the next remark.

Remark 1.1.3.

We believe the version of the Cohen-Lenstra heuristics proven in [EVW16] should be viewable as a degenerate case of Theorem 1.1.2, where one takes AA to be a 11-dimensional torus, instead of an abelian scheme. The torus may be viewed as a degeneration of an elliptic curve. We note that [EVW16] does not directly follow from the results presented here, but we are hopeful that a modest generalization of the work in this paper could imply both those results and ours.

The next result computes the moments of Selmer groups. To introduce some further notation, if XX and YY are two finite abelian groups, we use #Surj(X,Y)\#\operatorname{Surj}(X,Y) for the number of surjections from XX to YY. We also define Z:=CUZ:=C-U.

Theorem 1.1.3.

With the same hypotheses on AA and ν\nu as in Theorem 1.1.2,

(1.2) limjlim supnnevenxQTwistU/𝔽qn(𝔽qj)#Surj(Selν(Ax),H)xQTwistU/𝔽qn(𝔽qj)1\displaystyle\lim_{j\to\infty}\limsup_{\begin{subarray}{c}n\to\infty\\ n\hskip 2.84544pt\mathrm{even}\end{subarray}}\frac{\sum_{x\in\operatorname{QTwist}^{n}_{U/\mathbb{F}_{q}}(\mathbb{F}_{q^{j}})}\#\operatorname{Surj}(\operatorname{Sel}_{\nu}(A_{x}),H)}{\sum_{x\in\operatorname{QTwist}^{n}_{U/\mathbb{F}_{q}}(\mathbb{F}_{q^{j}})}1} =#Sym2H,\displaystyle=\#\operatorname{Sym}^{2}H,

as well as the analogous statement with lim sup\limsup replaced with lim inf\liminf.

If, moreover, there is some σZ(𝔽q)\sigma\in Z(\mathbb{F}_{q}) over which AA has good reduction,

(1.3) limjlimnnevenxQTwistU/𝔽qn(𝔽qj)#Surj(Selν(Ax),H)xQTwistU/𝔽qn(𝔽qj)1\displaystyle\lim_{j\to\infty}\lim_{\begin{subarray}{c}n\to\infty\\ n\hskip 2.84544pt\mathrm{even}\end{subarray}}\frac{\sum_{x\in\operatorname{QTwist}^{n}_{U/\mathbb{F}_{q}}(\mathbb{F}_{q^{j}})}\#\operatorname{Surj}(\operatorname{Sel}_{\nu}(A_{x}),H)}{\sum_{x\in\operatorname{QTwist}^{n}_{U/\mathbb{F}_{q}}(\mathbb{F}_{q^{j}})}1} =#Sym2H.\displaystyle=\#\operatorname{Sym}^{2}H.

Theorem 1.1.3 is proven in § 10.2.3.

Remark 1.1.4.

An upgraded version of (1.2), bounding the error term as jj\to\infty by a constant (depending on AA and HH) divided by q\sqrt{q} can be deduced from the analogous error term provided in Theorem 9.2.1, following the same proof in § 10.2.3.

Remark 1.1.5.

The condition that there is some σZ(𝔽q)\sigma\in Z(\mathbb{F}_{q}) over which AA has good reduction is fairly easy to arrange, by first passing to an extension where CC has a 𝔽q\mathbb{F}_{q} point of good reduction, and then augmenting ZZ to include that point. Note that this requires us to restrict the class of quadratic twists we consider to those which are unramified at the point we added to Z.

Moreover, it seems likely that the hypothesis that there is σZ(𝔽q)\sigma\in Z(\mathbb{F}_{q}) over which AA has good reduction can be removed. A viable path to doing so would involve two generalizations. First, we would need to carry out the whole paper in a setting where we require our quadratic twists be ramified at specified points in ZZ, as described further in 1.4.6. Second, we would need to carry out Appendix A in the setting where A[ν]A[\nu] has inertia type id-\operatorname{\mathrm{id}} over σ\sigma, see A.1.5. If one were able to verify both these generalizations, one could then show the limit in nn exists over 𝔽qj\mathbb{F}_{q^{j}} for sufficiently large jj where there is a point σC(𝔽qj)\sigma\in C(\mathbb{F}_{q^{j}}) by verifying the limit exists both in the case of quadratic twists ramified at σ\sigma and unramified at σ\sigma, and then adding the two resulting limits. These generalizations both seem quite approachable, and we believe it would be interesting to work this out.

Remark 1.1.6.

As we now explain, the informal example given in the first paragraph of § 1.1 is the special case of of Theorem 1.1.3 where ν=,H=/,\nu=\ell,H=\mathbb{Z}/\ell\mathbb{Z}, and C=𝔽q1C=\mathbb{P}^{1}_{\mathbb{F}_{q}}, and Z:=CUZ:=C-U is the union of the places of bad reduction of the abelian scheme, together with \infty. We will assume AA has good reduction over \infty so that the hypothesis preceding (1.3) is satisfied, although, as mentioned in 1.1.5, this is likely unnecessary. In this case, #Sym2H=\#\operatorname{Sym}^{2}H=\ell, so the average number of surjections from the \ell Selmer group to /\mathbb{Z}/\ell\mathbb{Z} is \ell. Since the \ell Selmer group is a finite dimensional vector space VV over /\mathbb{Z}/\ell\mathbb{Z},

#V=#Hom(/,V)=#Hom(V,/)=#Surj(V,/)+1.\displaystyle\#V=\#\mathrm{Hom}(\mathbb{Z}/\ell\mathbb{Z},V)=\#\mathrm{Hom}(V,\mathbb{Z}/\ell\mathbb{Z})=\#\operatorname{Surj}(V,\mathbb{Z}/\ell\mathbb{Z})+1.

Thus, the average size of the \ell Selmer group is +1\ell+1 as claimed.

It is well-known that bounds for average sizes (or more generally moments) of ν\nu Selmer groups yield interesting bounds on algebraic ranks (also known as Mordell-Weil ranks). Moreover, control of algebraic ranks gets better as ν\nu gets larger. See [BS13a, Proposition 5] and [PR12, p.246-247]. Since the results of the present paper allow ν\nu to be arbitrarily large, they are well-suited for results on algebraic ranks. For AA an abelian variety over a global field, we use rkA\operatorname{rk}_{\ell^{\infty}}A to denote the \ell^{\infty} Selmer rank of AA, which means that we can write Sel(A)(/)rkAG\operatorname{Sel}_{\ell^{\infty}}(A)\simeq(\mathbb{Q}_{\ell}/\mathbb{Z}_{\ell})^{\operatorname{rk}_{\ell^{\infty}}A}\oplus G, for GG a finite group. The minimalist conjecture, a version of which was originally posed by Goldfeld in 1979 [Gol79, Conjecture B], states that for suitable families of elliptic curves, the rank takes the value 0 half the time and 11 half the time. In this direction, we will prove the following version of the minimalist conjecture:

Theorem 1.1.4.

Suppose AA is an abelian scheme over UU satisfying (1.1), and ν=\nu=\ell is a prime satisfying the hypotheses of Theorem 1.1.2. Then,

limjlim supnnevenProb(rkAx=0:xQTwistU/𝔽qn(𝔽qj))\displaystyle\lim_{j\to\infty}\limsup_{\begin{subarray}{c}n\to\infty\\ n\hskip 2.84544pt\mathrm{even}\end{subarray}}\operatorname{Prob}(\operatorname{rk}_{\ell^{\infty}}A_{x}=0:x\in\operatorname{QTwist}^{n}_{U/\mathbb{F}_{q}}(\mathbb{F}_{q^{j}})) =12,\displaystyle=\frac{1}{2},
limjlim supnnevenProb(rkAx=1:xQTwistU/𝔽qn(𝔽qj))\displaystyle\lim_{j\to\infty}\limsup_{\begin{subarray}{c}n\to\infty\\ n\hskip 2.84544pt\mathrm{even}\end{subarray}}\operatorname{Prob}(\operatorname{rk}_{\ell^{\infty}}A_{x}=1:x\in\operatorname{QTwist}^{n}_{U/\mathbb{F}_{q}}(\mathbb{F}_{q^{j}})) =12,\displaystyle=\frac{1}{2},
limjlim supnnevenProb(rkAx2:xQTwistU/𝔽qn(𝔽qj))\displaystyle\lim_{j\to\infty}\limsup_{\begin{subarray}{c}n\to\infty\\ n\hskip 2.84544pt\mathrm{even}\end{subarray}}\operatorname{Prob}(\operatorname{rk}_{\ell^{\infty}}A_{x}\geq 2:x\in\operatorname{QTwist}^{n}_{U/\mathbb{F}_{q}}(\mathbb{F}_{q^{j}})) =0,\displaystyle=0,

as well as the analogous statements with lim sup\limsup replaced with lim inf\liminf.

Theorem 1.1.4 is proven in § 10.2.4.

Remark 1.1.7 (Versions of Theorem 1.1.4 for algebraic and analytic rank).

The \ell^{\infty} Selmer rank is conjecturally independent of \ell and equal to the analytic rank and algebraic rank. Since the Selmer rank is an upper bound for the algebraic rank, we can immediately deduce from Theorem 1.1.4 that the algebraic rank is at most 11 with probability 11, as jj\to\infty. We can also deduce from the parity conjecture [TY14] that the parity of the analytic rank approaches equidistribution as jj\to\infty. If we knew that the parity of the algebraic rank approached equidistribution as jj\to\infty, we could prove a version of the minimalist conjecture above for algebraic rank. Similarly, if we knew the analytic rank is at most 11 with probability 11 as jj\to\infty, we could deduce a version of the minimalist conjecture for analytic rank, and also use this and known relations between analytic and algebraic rank to deduce a version of the minimalist conjecture for algebraic rank.

1.2. Overview of the proof

The method of the proof has similar broad strokes to that of [EVW16]. See also [RW20] for a summary of this method. The loose idea is to construct moduli spaces parameterizing objects associated to the Selmer groups we want to count. We then count 𝔽q\mathbb{F}_{q} points on these moduli spaces using the Grothendieck-Lefschetz trace formula and Deligne’s bounds, which relates these point counts to the cohomology of these moduli spaces. We bound the higher homology groups using a homological stability theorem, and control the 0th homology group via a big monodromy result. Altogether, this gives us enough control on the point counts to estimate the moments. Finally, we show that these moments determine the distribution of Selmer groups, and that the resulting distribution agrees with the predicted one.

Nearly every aspect of this strategy turns out to be trickier in the context of the BKLPR heuristics than it was in the context of the Cohen-Lenstra heuristics. We next outline the additional difficulties.

1.3. Summary of the main innovations

1.3.1. The connection between Selmer groups and Hurwitz stacks

One of the main insights in this paper is that there is a close relation between Selmer groups and Hurwitz stacks. It has been well known for many years that the moduli spaces parameterizing objects in the Cohen-Lenstra heuristics were Hurwitz stacks related to dihedral group covers. However, it seems not to have been previously noticed that the moduli spaces appearing in the BKLPR heuristics are also closely related to Hurwitz stacks. Indeed, in 6.4.5, we relate stacks parameterizing ν\nu Selmer group elements to Hurwitz stacks for the group ASp2r(/ν)\operatorname{ASp}_{2r}(\mathbb{Z}/\nu\mathbb{Z}), where ASp\operatorname{ASp} denotes the affine symplectic group, see 6.3.2.

1.3.2. Homological stability over higher genus punctured curves

A second difficulty is that the above Hurwitz stacks do not occur over compact topological surfaces, but instead occur over punctured surfaces, where the punctures occur at the places of bad reduction of the abelian scheme. This necessitates that we prove a generalization of the topological results of [EVW16] (which only apply to Hurwitz stacks over the disc) to Hurwitz stacks over more general Riemann surfaces which may be punctured and may have positive genus.

The reader familiar with [EVW16] may note the absence of something that plays a crucial role in that paper: a conjugacy class cc in G=ASp2r(/ν)G=\operatorname{ASp}_{2r}(\mathbb{Z}/\nu\mathbb{Z}) which generates the whole group and which satisfies the “non-splitting" condition necessary for that paper. In fact, that role is played in the present work by the conjugacy class in GG consisting of elements whose image in the symplectic group is id.-\operatorname{\mathrm{id}}. This conjugacy class does not, of course, generate the whole of GG, which places us outside the context in which the methods of [EVW16] directly apply. More precisely, a branched GG-cover of the disc, all of whose monodromy lies in cc, is automatically disconnected, consisting of components whose monodromy group is actually the smaller group generated by cc. But, in the generality of the present paper, our Hurwitz spaces will be covers of a Riemann surface with (f+1)+n(f+1)+n punctures, where the monodromies of the relevant Sp2g(/ν)\mathrm{Sp}_{2g}(\mathbb{Z}/\nu\mathbb{Z}) representation around the first f+1f+1 punctures and around loops forming a basis for the homology of the surface are specified in advance, while only the monodromies around the last nn punctures are required to lie in the conjugacy class cc. Such a cover of a Riemann surface can certainly be connected, i.e., have full monodromy group GG. As we will see, it is examples precisely of this kind that will arise when we analyze the moduli stacks attached to variation of Selmer groups in quadratic twist families.

1.3.3. Homological stability for spaces more exotic than Hurwitz stacks

Once one deals with the above issues, one might then expect it to be possible to follow the strategy of [EVW16] to control the cohomology of these spaces, use this to control the finite field point counts via the Grothendieck-Lefschetz trace formula and Deligne’s bounds, and finally deduce the relevant BKLPR conjectures. However, this approach would, at best, only compute the moments of the BKLPR distribution. It turns out that this distribution is not completely determined by its moments, see [FLR23, Example 1.12]. In particular, if one restricts to elliptic curves whose Selmer rank is even, the resulting distribution has the same moments as the full BKLPR distribution. Therefore, at the very least, in order to show these heuristics hold, we need a way of separating out abelian varieties of even and odd Selmer rank. Fortunately, it turns out that there is a certain double cover of the stack of quadratic twists which governs whether the corresponding abelian variety has even or odd Selmer rank. This double cover is not a Hurwitz stack; nonetheless, the new homological stability results proved in this paper are general enough to apply to such covers. In this way, we prove homological stability results not just for Hurwitz stacks over punctured Riemann surfaces, but a more general class of covers of configuration space on these Riemann surfaces. A similar framework, applying to a different class of covers, was developed in [RWW17].

1.3.4. Proving the stabilization maps respect the Frobenius action

One step of this paper whose analog does not appear in [EVW16] is that we prove that the limit in nn exists in (1.3). To show this limit exists, the key point is to show that the homological stabilization maps appearing in our main results respect the action of Frobenius, and hence the traces of Frobenius on these cohomology groups are compatible. This is carried out in Appendix A.

A natural explanation for the equivariance would be that the stabilization map we exhibit topologically is the base change of a map of schemes over 𝔽q\mathbb{F}_{q}; but this appears to be too much to hope for. Instead, we show the map is induced by a map of log schemes over 𝔽q\mathbb{F}_{q}, which is enough to obtain Frobenius equivariance of the stabilization map. This idea was inspired by a similar use of log schemes in [BDPW23, §8]. In that paper, log structures were used not for the purpose of showing stabilization maps are equivariant, but instead for the purpose of showing that the cohomology of the relevant spaces are of Tate type.

In our setting, significant technical care and new ideas are needed to properly construct the stabilization maps and show they are equivariant. First, we need to carefully construct partial compactifications of Selmer spaces. Second, we must endow these spaces with the correct additional data and log structure so that the resulting map of log stacks matches the topological stabilization map over \mathbb{C}.

1.3.5. Proving the stabilization maps have degree 22

Even once the Frobenius equivariance described above was in place, in order to show the limit in (1.3) exists over all even nn, we needed to construct a stabilization map of degree 22. If we only had a degree dd stabilization map, we would only be able to show the limit exists along nn lying a given residue class modulo dd. Previously, as far as we are aware, the general belief of the community seems to have been that the degree of the stabilization map was rather large. However, by using recent work of Wood, we are able to show in § A.3 that there is a stabilization map of degree 22, and so the limit over all even nn exists on the nose.

1.3.6. Working with symplectically self-dual sheaves

Another crucial point is that throughout we work not with ν\nu-torsion in an abelian scheme, but in the more general setting of symplectically self-dual sheaves. This idea is also prominent in many works of Katz, such as [Kat02]. Working in this level of generality is crucial for us, as our topological results only apply in characteristic 0, so if we start with an abelian scheme in positive characteristic, we need some way of lifting it to characteristic 0 in a way compatible with our hypotheses. While we are quite unsure whether this is possible for abelian schemes, it is not too difficult for symplectically self-dual sheaves.

We now explain why we are able to get away with working with symplectically self-dual sheaves, in place of abelian schemes. Under the assumptions of Theorem 1.1.2, the Selν(A)\operatorname{Sel}_{\nu}(A) only depends on A[ν]A[\nu]. Namely, if C,A,ν,C,A,\nu, and qq are as in Theorem 1.1.2, Selν(A)H1(C,𝒜[ν])\operatorname{Sel}_{\nu}(A)\simeq H^{1}(C,\mathscr{A}[\nu]), for 𝒜\mathscr{A} the Néron model of AA over CC. (A similar isomorphism holds in the number field case, see [Ces16, Proposition 5.4(c)].) Hence, Selν(A)\operatorname{Sel}_{\nu}(A) is determined just from the group scheme A[ν]A[\nu] because 𝒜[ν]=jA[ν]\mathscr{A}[\nu]=j_{*}A[\nu] for j:UCj:U\to C the open inclusion. Therefore, we are free to forget that we started with an abelian scheme, so long as we remember this symplectically self-dual étale sheaf A[ν]A[\nu].

1.3.7. Difficulties related to g>0g>0, BKLPR moments, and monodromy

There are several further subtleties, and we now briefly summarize a couple of them. First, unlike the case of genus 0, in higher genus, there may be many quadratic twists with the same ramification divisor. Second, for ν\nu a general composite integer, the the moments of the BKLPR distribution do not seem to be computed in the existing literature. We note that when ν\nu is prime, and more generally when HH is a free /ν\mathbb{Z}/\nu\mathbb{Z} module, these moments were computed in [BKL+15, Theorem 5.10]. We compute the moments of the BKLPR distribution for general composite ν\nu in 2.3.1.

Third, we need to compute the relevant monodromy groups. This too requires additional technical work, where we draw great inspiration from works of Katz [Kat02] and Hall [Hal08], relying on the theory of middle convolution.

1.4. Discussion of equidistribution of parity of rank

We next include a number of remarks relating to our main results and equidistribution of the parity of rank. The following example gives a case where the parity of rank is not equidistributed, and shows that some version of our assumption (1.1) is necessary.

Remark 1.4.1.

Some version of the assumption (1.1) in Theorem 1.1.2 is necessary. Indeed, without (1.1), it is possible that every quadratic twist corresponding to a point of QTwistU/𝔽qn(𝔽qj)\operatorname{QTwist}^{n}_{U/\mathbb{F}_{q}}(\mathbb{F}_{q^{j}}) has Selmer rank of a fixed parity. Hence, quadratic twists of such a curve do not satisfy the minimalist conjecture. A specific example is given by the elliptic curve y2=λ(λ1)x(x1)(xλ)y^{2}=\lambda(\lambda-1)x(x-1)(x-\lambda), over 𝔽q(λ)\mathbb{F}_{q}(\lambda), where qq is a prime which is 1mod41\bmod 4. This is a variant of the Legendre family. Indeed, in [Kat02, 8.6.7], it is shown the relevant arithmetic monodromy group we define in 7.1.1 is contained in the special orthogonal group. (We can also see the geometric monodromy is contained in the special orthogonal group using the methods of this paper, since one can use 8.1.7 to show all generators of the fundamental group of configuration space map to the special orthogonal group.) In this case, the proof of 8.3.1 shows that for all but finitely many primes \ell, the \ell Selmer group of every quadratic twist unramified over the places of bad reduction has even Selmer rank. Note here that assumption (1.1) of Theorem 1.1.2 is not satisfied as each of the three places of bad reduction of the elliptic curve y2=λ(λ1)x(x1)(xλ)y^{2}=\lambda(\lambda-1)x(x-1)(x-\lambda), given by λ=0,λ=1,\lambda=0,\lambda=1, and λ=\lambda=\infty, has additive reduction. For some further related examples, also see [Riz97, Riz99, Riz03].

Remark 1.4.2.

Under the assumptions of Theorem 1.1.2, the parity of the rank of Selmer groups in the quadratic twist families we consider is equidistributed. The proportion of the time the rank takes a given parity in the number field setting has been the object of much study, see for example [KMR13, Conjecture 7.12]. We believe it would be quite interesting to understand better understand the relation between the number field and function field perspectives on this question.

In the example considered in 1.4.1, for sufficiently large qq, the proportion of quadratic twists with Selmer rank 2\geq 2 becomes arbitrarily close to 0. We wonder whether this continues to hold even in the absence of (1.1):

Question 1.4.3.

Suppose AA is any abelian scheme over UU, for UU an affine curve over 𝔽q\mathbb{F}_{q}. What conditions do we need on AA so that the proportion of quadratic twists of A×Spec𝔽q𝔽qjA\times_{\operatorname{Spec}\mathbb{F}_{q}}\mathbb{F}_{q^{j}} with (Selmer) rank 2\geq 2 tend to 0 as jj grows, even in the absence of (1.1)?

We conjecture that an irreducibility condition on the Galois representation associated to AA will suffice. More specifically make the following conjecture, many cases of which are suggested by Theorem 1.1.4. We say a quadratic twist is unramified at a real place if the corresponding double cover has two real places over that real place, and is ramified at a real place if the double has a complex place over that real place.

Conjecture 1.4.4.

Let KK be any global field of characteristic not 22 and AA any abelian variety of dimension rr over KK.

(1.4) Suppose that for some prime \ell, char(K)\ell\neq\operatorname{\operatorname{char}}(K), the identity component of the Zariski
closure of im(Gal(K¯/K)GL(H1(AK¯,¯(1))))\operatorname{im}(\operatorname{Gal}(\overline{K}/K)\to\operatorname{GL}(H^{1}(A_{\overline{K}},\overline{\mathbb{Q}}_{\ell}(1)))) acts irreducibly on H1(AK¯,¯(1))H^{1}(A_{\overline{K}},\overline{\mathbb{Q}}_{\ell}(1)).

Specify divisors Dunram,DramD_{\operatorname{unram}},D_{\operatorname{ram}} whose union contains all places of bad reduction of AA and all real places. The set of quadratic twists of AA unramified over DunramD_{\operatorname{unram}} and ramified over DramD_{\operatorname{ram}} have ranks distributed according to one of the following three possibilities:

  1. (1)

    0%0\% rank >1>1, 50%50\% rank 0, 50%50\% rank 11,

  2. (2)

    0%0\% rank >1>1, 100%100\% rank 0, 0%0\% rank 11,

  3. (3)

    0%0\% rank >1>1, 0%0\% rank 0, 100%100\% rank 11.

We next explain some of our motivation for the above conjecture, especially the hypothesis (1.4).

Remark 1.4.5.

Note that some sort of assumption of the flavor of (1.4) is necessary in 1.4.4, since if A=ErA=E^{r}, for r>1r>1 and EE a generic elliptic curve, we would expect the rank to be 0 half the time and rr half the time.

The reason that we believe (1.4) should be sufficient comes from the big monodromy result of Katz, [Kat02, Proposition 5.4.3]. This essentially says that if, in the function field setting, for AA an abelian scheme over UU and geometric point Spec𝔽¯qx¯U𝔽¯q\operatorname{Spec}\overline{\mathbb{F}}_{q}\simeq\overline{x}\to U_{\overline{\mathbb{F}}_{q}}, H1(A𝔽¯q×U𝔽¯qx¯,¯(1))H^{1}(A_{\overline{\mathbb{F}}_{q}}\times_{U_{\overline{\mathbb{F}}_{q}}}{\overline{x}},\overline{\mathbb{Q}}_{\ell}(1)) corresponds to an irreducible representation of π1(U𝔽¯q,x¯)\pi_{1}(U_{\overline{\mathbb{F}}_{q}},\overline{x}) for some char(K)\ell\neq\operatorname{\operatorname{char}}(K), a certain related monodromy group should be big, i.e., contain the special orthogonal group. It seems to us this should imply that the geometric monodromy representation considered in 7.1.1 for ν=\nu=\ell has index at most 44 in the orthogonal group mod\bmod\ell. We conjecture that in this case the BKLPR conjectures hold, with the possible caveat that the rank may have a fixed parity if the monodromy group is contained in the special orthogonal group. It is not immediately clear how to best generalize the condition that H1(A𝔽¯q×U𝔽¯qx¯,¯(1))H^{1}(A_{\overline{\mathbb{F}}_{q}}\times_{U_{\overline{\mathbb{F}}_{q}}}{\overline{x}},\overline{\mathbb{Q}}_{\ell}(1)) is irreducible to the number field setting, but it seems that (1.4) should imply it, and so (1.4) seems a reasonable sufficient criterion.

Remark 1.4.6.

Throughout this paper, we work with the space of quadratic twists parameterizing double covers whose ramification locus does not intersect the discriminant locus. As a variant, we could work with the space of finite double covers whose ramification locus contains a specified divisor RR (where RR may intersect the discriminant locus) but the ramification locus of the cover does not meet the discriminant locus outside of RR.

Assuming there is a place of multiplicative reduction with toric part of codimension 11 outside of RR, and replacing the space of quadratic twists in our main theorems with the above variant, we believe the conclusions of Theorem 1.1.2, Theorem 1.1.3, and Theorem 1.1.4 should still hold.

In fact, we believe one can make a more precise version of 1.4.4 that predicts which of the three cases we are in based on local data associated to the abelian variety, similarly to the case of elliptic curves which is closely related to [KMR13, Proposition 7.9]. We believe this generalization would lead to a version of [KMR13, Conjecture 7.12] for global arbitrary fields.

It would be quite interesting to work the above claims out precisely.

1.5. Discussion on the presence of limsup and liminf

We conclude our remarks with comments pertaining to the presence of the lim sup\limsup and lim inf\liminf.

Remark 1.5.1.

Previously, it was not even known that the lim sup\limsup and lim inf\liminf appearing in (1.2) of Theorem 1.1.3 even existed, nor that the limit in nn appearing in (1.3) existed, let alone what their limiting value as jj\to\infty was. The fact that these limits exist is an important part of these theorems. We also note that if one only cares about verifying the existence of the lim sup\limsup and lim inf\liminf, without computing the value after taking a further limit in jj, one does not need the full force of our big monodromy results culminating in 9.2.1, which are what enables us to compute these values precisely. Instead, one may use Theorem 4.2.1 and 4.2.4 to obtain an ineffective bound on the relevant number of irreducible components.

Remark 1.5.2.

The reason we cannot propagate this existence of the limit in nn of (1.3) to our other main results such as Theorem 1.1.2 (which only has a lim sup\limsup and a lim inf\liminf) is that we do not know how to rule out the possibility that the moments grow too quickly to determine a distribution for any fixed value of qq.

Even more ambitiously, one might want to know what these limits in nn actually are, and in particular whether they agree with the BKLPR heuristics. For this, one would likely want to know not only that the étale cohomology groups stabilize as Frobenius modules up to Tate twist, but what Frobenius module they stabilize to. For the moment, this appears to be a substantially harder problem. See also 8.2.4 and 9.2.6.

1.6. Past work

As mentioned above, two guiding sets of conjectures in number theory are the Cohen-Lenstra heuristics and the BKLPR heuristics. Focusing on the latter over number fields, very little is known. Over \mathbb{Q}, work by [HB93, HB94, SD08, Kan13] led to a determination of the distribution of 22 Selmer groups in quadratic twist families of elliptic curves. Building on this, Smith described the distribution of 22^{\infty} Selmer groups of elliptic curves over \mathbb{Q} in [Smi22, Theorem 1.5]. Smith is able to use this to deduce the minimalist conjecture in many quadratic twist families over \mathbb{Q} [Smi22, Theorem 1.2]. The reason for this deduction is that Smith’s work, like ours, but unlike the previous papers cited in this paragraph, provides distributional information about ν\nu Selmer groups with ν\nu arbitrarily large. These results for quadratic twist families over number fields nearly exclusively deal with 22-power Selmer groups. Our results are in some sense disjoint, applying only to ν\nu Selmer groups for ν\nu odd.

There is also some work toward understanding 33-isogeny Selmer groups in quadratic twist families ([BKLOS19].) However, the above results are only for 33 Selmer groups, and only when the pertinent curves possess unexpected isogenies. As far as we are aware, our work provides the first results toward describing the distribution of odd order Selmer groups in quadratic twist families when there are no unexpected isogenies.

There is also a growing literature about variation of Selmer groups in the universal family parameterizing all elliptic curves. For this family, Bhargava and Shankar computed the average size of the ν\nu Selmer group for ν5\nu\leq 5 [BS15a, BS15b, BS13a, BS13b], and Bhargava-Shankar-Swaminathan computed the second moment of 22 Selmer groups [BSS21].

Over function fields, much more is known if one permits taking a limit in the finite field order qq before any limit in log-height is taken. (Here, the log-height of a quadratic twist refers to the degree of its ramification locus.) In the context of the Cohen-Lenstra heuristics, [Ach08] established a large qq limit version of the Cohen-Lenstra heuristics, where he took a qq limit before letting the log-height grow.

In the context of the BKLPR heuristics, some results were also known when one takes a large qq limit prior to large log-height limit: The average size of certain Selmer groups in quadratic twist families were computed in [PW23]. In the context of the universal family, [Lan21] computed the average size of Selmer groups, and the full BKLPR distribution was computed in [FLR23].

Closer to the present work are results in which one takes a limit in log-height first, with qq fixed, and only then lets qq increase. De Jong [dJ02] computed the average size of 33 Selmer groups over 𝔽q(t)\mathbb{F}_{q}(t) in the universal family of elliptic curves. Hồ, Lê Hùng, and Ngô [HLHN14] compute the average size of 22 Selmer groups over function fields for the universal family, while [Ach23] carries out a similar program in all characteristics, including characteristic 22. We note that these results both have the same flavor as our main results, in that they only arrive at the predicted value after first taking a large log-height limit, and then taking a large qq limit. Another more recent result of Thorne [Tho19] calculates the average size of 22 Selmer groups in a family of elliptic curves with 22 marked points over genus 0 function fields, and, interestingly, this result does not require taking a large qq limit at the end. We also note that [HLHN14, Theorem 2.2.5] does not require taking a large qq limit if one restricts to elliptic curves with squarefree discriminant.

Since the work [EVW16] proved a homological stability result for Hurwitz stacks, there has also been further activity in this topological direction. The homological stability results of [EVW16] have been employed in a number of arithmetic papers, such as in [LST20], [LT19], and [ELS20]. However, few papers have further developed the homological stability techniques. Some notable examples where these techniques were developed further include [ETW17], proving a version of Malle’s conjecture, a polynomial version of homological stability in [BM23], a verification that stability in [EVW16] holds with period 11 instead of with period degU\deg U in [DS23], and a bound on the ranks of homology groups for Hurwitz spaces associated to punctured genus 0 surfaces in [Hoa23]. Finally, [BDPW23] and [MPPRW24] used homological stability techniques to approach a conjecture on moments of quadratic L-functions, and were able to not only show the relevant cohomology groups stabilize, but even compute their limiting values.

1.7. Outline

The structure of the paper is as follows. We suggest the reader consult Figure 1 for a schematic depiction of the main ingredients in the proof. In § 2 we review background on orthogonal groups, the BKLPR heuristics, and Hurwitz stacks. Next, we continue to the topological part of our paper. In § 3, we set up a general notion of coefficient systems (which include Hurwitz stacks over the complex numbers as a special case) to which the arc complex spectral sequence applies. This is the context in which we prove our main homological stability results in § 4. We next continue to the more algebraic part of the paper, beginning with § 5, where we construct Selmer stacks which parameterize Selmer elements on quadratic twists of our abelian scheme. In § 6, we show that the above constructed Selmer stacks can be identified with Hurwitz stacks over the complex numbers. In order to compute the 0th homology of these spaces, we prove a big monodromy result in § 7. We verify our homological stability results apply to these Selmer stacks, as well as to certain double covers, which control the parity of the \ell^{\infty} Selmer rank of the quadratic twists of our abelian scheme, in § 8. Having controlled the cohomologies of the spaces we care about, we conclude our main results by combining the above with some slightly more analytic computations. In § 9, we compute the moments related to Selmer stacks, as well as fiber products of these with the above mentioned double cover. In § 10, we show these moments determine the distribution, obtaining our main result, Theorem 1.1.2. In Appendix A, we use logarithmic geometry to prove that the stabilization maps on cohomology are equivariant for the action of Frobenius, up to twist, which allows us to show that a limit as nn\to\infty exists in (1.3), instead of only knowing that the lim inf\liminf and lim sup\limsup exist as in (1.2). Finally, in Appendix B, we use logarithmic geometry to prove that configuration spaces and Hurwitz spaces have normal crossings compactifications. This is a crucial ingredient for us to be able to transfer cohomology between the complex numbers and finite fields.

TopologyBig monodromyProbabilityLogarithmic GeometryCombinatorial Group Theory Lem.5.3.2Lem.7.4.6Lem.9.1.5Prop.5.2.6Thm.7.1.1[Hal08, Kat02, Middle convolution]{\underbrace{\text{Thm.}~\ref{theorem:big-monodromy-mod-ell}}_{\text{\cite[cite]{[\@@bibref{}{hall:bigMonodromySympletic,katz:twisted-l-functions-and-monodromy}{}{}, Middle convolution]}}}}Prop.7.3.3Prop.9.2.1Prop.10.1.1Lem.10.2.2   Lem.6.3.7Cor.6.4.7Lem.8.2.3Thm.9.2.1Thm.10.2.1Thm.1.1.2   Thm.4.2.1Cor.4.3.4Thm.1.1.3Prop.3.2.4Thm.4.1.1Lem.A.4.2Thm.A.5.1   Lem.4.2.3Thm.4.2.2Prop.A.3.1[EVW16]

Figure 1. A diagram depicting the structure of the proof of the main result, Theorem 1.1.2.

1.8. Notation

For the reader’s convenience, in Figure 2 we collect some notation introduced throughout the paper.

Notation Description Location defined
ν\nu Odd integer indexing the Selmer group Selν(A)\operatorname{Sel}_{\nu}(A) 2.1.1
DQD_{Q} The Dickson invariant map associated to a quadratic form QQ 2.1.1
SelνBKLPR\operatorname{Sel}^{\operatorname{BKLPR}}_{\nu} The BKLPR predicted distribution of ν\nu Selmer groups 2.2.1
BB Base scheme 2.4.1
CC Smooth proper curve over BB 2.4.1
ZZ Divisor in CC of degree f+1f+1 which twists are unramified along 2.4.1
ConfU/Bn\operatorname{Conf}^{n}_{U/B} Configuration space of degree nn divisors in UU 2.4.1
HurC/BG,n,Z,S\operatorname{Hur}^{G,n,Z,S}_{C/B} Hurwitz space parameterizing GG covers with monodromy in SS 2.4.2
Σg,fb\Sigma^{b}_{g,f} Topological surface of genus gg with bb boundary components and ff punctures 3.1.1
XnAg,fX^{\oplus n}\oplus A_{g,f} nn copies of a marked cylinder glued onto Σg,f1\Sigma^{1}_{g,f} 3.1.1
Bg,fnB^{n}_{g,f} The surface braid group π1(ConfXnAg,fn)\pi_{1}(\operatorname{Conf}^{n}_{X^{\oplus n}\oplus A_{g,f}}) 3.1.1
HTG,c,g,fnH_{T^{n}_{G,c,g,f}}. The nnth vector space from a coefficient system corresponding to a Hurwitz space 3.1.9
RVR^{V} Ring of connected components associated to a coefficient system over Σ0,01\Sigma^{1}_{0,0} 3.2.1
𝒦(M)\mathcal{K}(M) KK-complex associated a graded RVR^{V} module 3.2.1
MpV,FM^{V,F}_{p} n0Hp(Bg,fn,Fn)\oplus_{n\geq 0}H_{p}(B^{n}_{g,f},F_{n}) 3.2.2
\mathscr{F} A tame, symplectically self-dual lcc sheaf of free /ν\mathbb{Z}/\nu\mathbb{Z} modules on UU 5.1.4
QTwistU/Bn\operatorname{QTwist}^{n}_{U/B} A Hurwitz space parameterizing quadratic twists 5.1.4
Bn\mathscr{F}^{n}_{B} The universal degree nn quadratic twist of \mathscr{F} 5.1.4
𝒮eBn{\mathcal{S}e\ell}_{\mathscr{F}^{n}_{B}} The Selmer sheaf, which parameterizes torsors for quadratic twists of \mathscr{F} of log-height nn 5.1.5
SelBn\operatorname{Sel}_{\mathscr{F}^{n}_{B}} The Selmer stack, which is the finite étale cover of QTwistU/Bn\operatorname{QTwist}^{n}_{U/B} corresponding to the Selmer sheaf 5.1.5
Cx,Ux,x,AxC_{x},U_{x},\mathscr{F}_{x},A_{x} The fiber of the relevant object over xQTwistU/Bnx\in\operatorname{QTwist}^{n}_{U/B} 5.1.10
ΦA\Phi_{A^{\prime}} Component group of the abelian scheme AA^{\prime} over UU^{\prime} 5.2.2
ASp2r\operatorname{\mathrm{ASp}}_{2r} The affine symplectic group 6.3.2
AHSp2r\operatorname{\mathrm{A}^{\operatorname{H}}\mathrm{Sp}}_{2r} The HH moment of the affine symplectic group 6.3.4
HurHBn\operatorname{Hur}^{\mathscr{F}^{n}_{B}}_{H} A certain Hurwitz space which is geometrically isomorphic to the Selmer sheaf 6.4.1
VBnV_{\mathscr{F}^{n}_{B}} The vector space corresponding to a geometric fiber of the Selmer sheaf 7.1.1
ρBn\rho_{\mathscr{F}^{n}_{B}} The monodromy representation associated to the Selmer sheaf 7.1.1
XA[ν]𝔽qnX_{A[\nu]^{n}_{\mathbb{F}_{q}}} Probability distribution on ν\nu-Selmer groups of quadratic twists of AA over 𝔽q\mathbb{F}_{q} 7.4.1
XA[ν]𝔽qniX^{i}_{A[\nu]^{n}_{\mathbb{F}_{q}}} Probability distribution on ν\nu-Selmer groups of quadratic twists of AA over 𝔽q\mathbb{F}_{q} with fixed parity of rank 7.4.1
S,H,g,fnS^{n}_{\mathscr{F},H,g,f} Coefficient system associated to HH-moment of the ν\nu Selmer sheaf 8.1.3
S¯,H,0,0nS^{n}_{\underline{\mathscr{F}},H,0,0} The coefficient system for Σ0,01\Sigma^{1}_{0,0} which S,H,g,fnS^{n}_{\mathscr{F},H,g,f} lies over 8.1.3
S,g,fn,rkS^{n,\operatorname{rk}}_{\mathscr{F},g,f} The coefficient system associated to the rank double cover 8.1.4
𝒩\mathcal{N} Finite abelian /ν\mathbb{Z}/\nu\mathbb{Z} modules 7.4.1
𝒩i\mathcal{N}^{i} The subset of objects of NN of the form (/ν)i×G2(\mathbb{Z}/\nu\mathbb{Z})^{i}\times G^{2} 7.4.1

Figure 2. Some notation introduced in the paper.

1.9. Acknowledgements

We thank Craig Westerland for numerous helpful and detailed discussions which were invaluable in pinning down some of the trickiest topological inputs to this paper. We also thank Dori Bejleri for suggesting the idea to prove the stabilization maps respect the Frobenius action, for extensive help with technical aspects of log geometry. We thank Chris Hall for a meticulously close reading and numerous helpful discussions. Thanks to Eric Rains for many helpful exchanges, especially relating to the BKLPR heuristics and Vasiu’s lifting results. We thank Melanie Wood for a number of useful conversations relating to determining the distribution from the moments. Thanks additionally to Levent Alpöge and Bjorn Poonen for help understanding the possible structures of the Tate-Shafarevich group. We also thank Sun Woo Park for a close reading and for numerous detailed and helpful comments. We further thank Dan Abramovich, Niven Achenjang, Andrea Bianchi, Kevin Chang, Qile Chen, Chantal David, Tony Feng, Jeremy Hahn, David Harbater, Anh Trong Nam Hoang, Hyun Jong Kim, Ben Knudsen, Michael Kural, Jef Laga, Peter Landesman, Eric Larson, Robert Lemke Oliver, Ishan Levy, Siyan Daniel Li-Huerta, Daniel Litt, Davesh Maulik, Barry Mazur, Jeremy Miller, Samouil Molcho, Martin Olsson, Dan Petersen, Andy Putman, Oscar Randal-Williams, Zev Rosengarten, Will Sawin, Mark Shusterman, Alex Smith, Salim Tayou, Ravi Vakil, and David Yang. This work also owes a large intellectual debt to a number of others including work of Chris Hall, work of Nick Katz, and work of Oscar Randal-Williams and Nathalie Wahl. JE was supported by the National Science Foundation under Award No. DMS 2301386, and AL was supported by the National Science Foundation under Award No. DMS 2102955.

2. Background

We now review some background on orthogonal groups in § 2.1, background on the BKLPR heuristics in § 2.2, and background on Hurwitz stacks in § 2.4. The one new part of this section is § 2.3, where we compute the moments of the BKLPR distribution.

2.1. Orthogonal groups

We now define some notation we will use relating to orthogonal groups. Throughout, we will be working over base rings RR with 22 invertible on RR, and so we will freely pass between quadratic spaces and spaces with a bilinear pairing. For some additional detail and further references, we refer the reader to [FLR23, §3.2] whose material in turn was largely drawn from [Con14, Appendix C].

Notation 2.1.1.

Let R=/νR=\mathbb{Z}/\nu\mathbb{Z}, for some ν\nu with gcd(ν,2)=1\gcd(\nu,2)=1. Let VV be a free RR module of rank at least 33 with a bilinear pairing B:V×VRB:V\times V\to R. Let Q:VR,Q:V\to R, defined by Q(v):=B(v,v)Q(v):=B(v,v) denote the associated quadratic form. We assume throughout that QQ is nondegenerate, meaning that the quadric associated to QQ is smooth, or equivalently QQ is nondegenerate modulo every prime ν\ell\mid\nu. We let O(Q){\rm{O}}(Q) denote the associated orthogonal group preserving QQ. There is a Dickson invariant map DQ:O(Q)ν prime/2D_{Q}:{\rm{O}}(Q)\to\prod_{\ell\mid\nu\text{ prime}}\mathbb{Z}/2\mathbb{Z} by sending an element to 0 in coordinate \ell if its determinant mod\bmod\ell is 11 and sending it to 11 if its determinant mod\bmod\ell is 1-1. There is also a +1+1-spinor norm map spQ+:O(Q)H1(SpecR,μ2)R×/(R×)2ν prime/2\operatorname{sp}_{Q}^{+}:{\rm{O}}(Q)\to H^{1}(\operatorname{Spec}R,\mu_{2})\simeq R^{\times}/(R^{\times})^{2}\simeq\prod_{\ell\mid\nu\text{ prime}}\mathbb{Z}/2\mathbb{Z}, where the map in cohomology is induced by the boundary map associated to the exact sequence of algebraic groups μ2Pin(Q)O(Q)\mu_{2}\to\operatorname{Pin}(Q)\to{\rm{O}}(Q). The 1-1-spinor norm, spQ:O(Q)ν prime/2\operatorname{sp}_{Q}^{-}:{\rm{O}}(Q)\to\prod_{\ell\mid\nu\text{ prime}}\mathbb{Z}/2\mathbb{Z}, is the composition of spQ+\operatorname{sp}_{Q}^{+} with the identification O(Q)O(Q){\rm{O}}(Q)\simeq{\rm{O}}(-Q), see [Con14, Remark C.4.9, Remark C.5.4, and p.348]. In particular, if rvr_{v} is the reflection about the vector vv, spQ(rv)=[Q(v)]\operatorname{sp}_{Q}^{-}(r_{v})=[-Q(v)], where [x][x] denotes the square class of xx, viewed as an element of ν prime/2\prod_{\ell\mid\nu\text{ prime}}\mathbb{Z}/2\mathbb{Z}.

We define Ω(Q):=kerDQkerspQO(Q)\Omega(Q):=\ker D_{Q}\cap\ker\operatorname{sp}_{Q}^{-}\subset{\rm{O}}(Q). In particular, since ν\nu is odd, Ω(Q)O(Q)\Omega(Q)\subset{\rm{O}}(Q) has index 4ω(ν)4^{\omega(\nu)}, where ω(ν)\omega(\nu) denotes the number of primes dividing ν\nu.

Remark 2.1.2.

It turns out that the map DQ×spQ:O(Q)ν prime(/2×/2)D_{Q}\times\operatorname{sp}_{Q}^{-}:{\rm{O}}(Q)\to\prod_{\ell\mid\nu\text{ prime}}(\mathbb{Z}/2\mathbb{Z}\times\mathbb{Z}/2\mathbb{Z}) can be identified with the abelianization of O(Q){\rm{O}}(Q), assuming QQ is nondegenerate and has rank more than 22.

The following lemma will be useful throughout the paper, and connects the Dickson invariant to the dimension of the 11-eigenspace of an element of the orthogonal group. We will see that the latter is related to Selmer groups via 5.3.2.

Lemma 2.1.3.

Let (V,Q)(V,Q) be a quadratic space over a field and gO(Q)g\in{\rm{O}}(Q). We have

dimker(gid)mod2rkVDQ(g).\displaystyle\dim\ker(g-\operatorname{\mathrm{id}})\bmod 2\equiv\operatorname{rk}V-D_{Q}(g).
Proof.

It follows from [Tay92, p. 160], that dimim(gid)=DQ(g)mod2\dim\operatorname{im}(g-\operatorname{\mathrm{id}})=D_{Q}(g)\bmod 2. We find

(2.1) dimker(gid)mod2\displaystyle\dim\ker(g-\operatorname{\mathrm{id}})\bmod 2 rkVdimim(gid)mod2\displaystyle\equiv\operatorname{rk}V-\dim\operatorname{im}(g-\operatorname{\mathrm{id}})\bmod 2
rkVDQ(g)mod2,\displaystyle\equiv\operatorname{rk}V-D_{Q}(g)\bmod 2,

using the exact sequence relating the kernel and image of gid:VVg-\operatorname{\mathrm{id}}:V\to V. ∎

2.2. Review of the BKLPR distribution

We now give a quick review of the predicted distribution for ν\nu Selmer groups given in [BKL+15]. We also suggest the reader consult [FLR23, §5.3] for a slightly more detailed description of this distribution, geared to the context in which we will use it in this paper.

2.2.1. The \ell^{\infty} Selmer distribution from BKLPR conditioned on rank

Let \ell be a prime. For non-negative integers m,rm,r with mr20m-r\in 2\mathbb{Z}_{\geq 0}, let AA be drawn randomly from the Haar probability measure on the set of alternating m×mm\times m-matrices over \mathbb{Z}_{\ell} having rank mrm-r. Let 𝒯m,r,\mathscr{T}_{m,r,\ell} be the distribution of (cokerA)tors(\operatorname{coker}A)_{\operatorname{tors}}, the torsion in cokerA\operatorname{coker}A. According to [BKL+15, Theorem 1.10], as mm\rightarrow\infty through integers with mr20m-r\in 2\mathbb{Z}_{\geq 0}, the distributions 𝒯m,r,\mathscr{T}_{m,r,\ell} converge to a limit 𝒯r,\mathscr{T}_{r,\ell}.

2.2.2. The BKLPR ν\nu Selmer distribution

We next review the model for ν\nu Selmer elements described at the beginning of [BKL+15, §5.7]. Let 𝒯r,\mathscr{T}_{r,\ell} denote the random variable defined on isomorphism classes of finite abelian \ell groups (notated 𝒯r\mathscr{T}_{r} in [BKL+15]) defined in [BKL+15, Theorem 1.6] and reviewed in § 2.2.1. For GG an abelian group, we let G[ν]G[\nu] denote the ν\nu torsion of GG. For ν1\nu\in\mathbb{Z}_{\geq 1} with prime factorization ν=νa\nu=\prod_{\ell\mid\nu}\ell^{a_{\ell}}, define a distribution 𝒯r,/ν\mathscr{T}_{r,\mathbb{Z}/\nu\mathbb{Z}} on finitely generated /ν\mathbb{Z}/\nu\mathbb{Z} modules by choosing a collection of abelian groups {T}ν\{T_{\ell}\}_{\ell\mid\nu}, with TT_{\ell} drawn from 𝒯r,\mathscr{T}_{r,\ell}, and defining the probability 𝒯r,/ν=G\mathscr{T}_{r,\mathbb{Z}/\nu\mathbb{Z}}=G to be the probability that νT[ν]G\oplus_{\ell\mid\nu}T_{\ell}[\nu]\simeq G.

Given the above predicted distribution for the ν\nu Selmer group of abelian varieties of rank rr, the heuristic that 50%50\% of abelian varieties have rank 0 and 50%50\% have rank 11 leads to the predicted joint distribution of the ν\nu Selmer group and rank given in 2.2.1. We use 𝒯1,/ν/ν\mathscr{T}_{1,\mathbb{Z}/\nu\mathbb{Z}}\oplus\mathbb{Z}/\nu\mathbb{Z} as notation for the random variable so that the probability 𝒯1,/ν/νG/ν\mathscr{T}_{1,\mathbb{Z}/\nu\mathbb{Z}}\oplus\mathbb{Z}/\nu\mathbb{Z}\simeq G\oplus\mathbb{Z}/\nu\mathbb{Z} is equal to the probability that 𝒯1,/νG\mathscr{T}_{1,\mathbb{Z}/\nu\mathbb{Z}}\simeq G.

Definition 2.2.1.

Let 𝒩\mathcal{N} denote the set of finite /ν\mathbb{Z}/\nu\mathbb{Z} modules. Let SelνBKLPR:𝒩0\operatorname{Sel}^{\operatorname{BKLPR}}_{\nu}:\mathcal{N}\to\mathbb{R}_{\geq 0} denote the probability distribution defined by

SelνBKLPR:=12𝒯0,/ν+12(𝒯1,/ν/ν).\displaystyle\operatorname{Sel}^{\operatorname{BKLPR}}_{\nu}:=\frac{1}{2}\mathscr{T}_{0,\mathbb{Z}/\nu\mathbb{Z}}+\frac{1}{2}\left(\mathscr{T}_{1,\mathbb{Z}/\nu\mathbb{Z}}\oplus\mathbb{Z}/\nu\mathbb{Z}\right).

For i{0,1}i\in\{0,1\} let SelνBKLPR,i:𝒩0\operatorname{Sel}^{\operatorname{BKLPR},i}_{\nu}:\mathcal{N}\to\mathbb{R}_{\geq 0} denote the distribution SelνBKLPR\operatorname{Sel}^{\operatorname{BKLPR}}_{\nu} conditioning on rkSelνBKLPRmodimod2,\operatorname{rk}\operatorname{Sel}^{\operatorname{BKLPR}}_{\nu}\bmod\ell\equiv i\bmod 2, for any ν\ell\mid\nu. In particular, SelνBKLPR,0\operatorname{Sel}^{\operatorname{BKLPR},0}_{\nu} is the distribution 𝒯0,/ν\mathscr{T}_{0,\mathbb{Z}/\nu\mathbb{Z}} while SelνBKLPR,1\operatorname{Sel}^{\operatorname{BKLPR},1}_{\nu} is the distribution 𝒯1,/ν/ν\mathscr{T}_{1,\mathbb{Z}/\nu\mathbb{Z}}\oplus\mathbb{Z}/\nu\mathbb{Z}.

Remark 2.2.2.

Note that SelνBKLPR,i\operatorname{Sel}^{\operatorname{BKLPR},i}_{\nu} is independent of ν\ell\mid\nu as follows from the definition of SelνBKLPR,i\operatorname{Sel}^{\operatorname{BKLPR},i}_{\nu}, 2.2.1, so the definition of SelνBKLPR,i\operatorname{Sel}^{\operatorname{BKLPR},i}_{\nu} is independent of the choice of ν\ell\mid\nu.

Remark 2.2.3.

We note that there was a slight error in [FLR23, Definition 5.12]. There, when r=1r=1, the distribution should have been given by 12(𝒯1,/ν/ν)\frac{1}{2}\left(\mathscr{T}_{1,\mathbb{Z}/\nu\mathbb{Z}}\oplus\mathbb{Z}/\nu\mathbb{Z}\right) and not 12(𝒯1,/ν)\frac{1}{2}\left(\mathscr{T}_{1,\mathbb{Z}/\nu\mathbb{Z}}\right) as written there. The latter models [ν]\Sha[\nu] as opposed to Selν\operatorname{Sel}_{\nu}.

2.3. Computing the moments of ν\nu Selmer groups

We next compute moments of the BKLPR distribution. For a distribution XX valued in finite abelian groups, we use the HH-moment of XX as terminology for the expected number of surjections or homomorphisms XHX\to H. Knowing the expected number of homomorphisms for all HH is equivalent to knowing the expected number of surjections for all HH by an inclusion exclusion argument.

The computation of the moments below in the case that H(/j)mH\simeq(\mathbb{Z}/\ell^{j}\mathbb{Z})^{m} was explained in [BKL+15, Theorem 5.10 and Remark 5.11]. Surprisingly, the general case appears to be missing from the literature. We follow a similar method of proof to [BKL+15, Theorem 5.10], though it is somewhat more involved.

Proposition 2.3.1.

We have

#Sym2H\displaystyle\#\operatorname{Sym}^{2}H =𝔼(#Surj(SelνBKLPR,H))\displaystyle=\mathbb{E}(\#\operatorname{Surj}(\operatorname{Sel}^{\operatorname{BKLPR}}_{\nu},H))
=𝔼(#Surj(SelνBKLPR,0,H))\displaystyle=\mathbb{E}(\#\operatorname{Surj}(\operatorname{Sel}^{\operatorname{BKLPR},0}_{\nu},H))
=𝔼(#Surj(SelνBKLPR,1,H)).\displaystyle=\mathbb{E}(\#\operatorname{Surj}(\operatorname{Sel}^{\operatorname{BKLPR},1}_{\nu},H)).
Proof.

We first reduce to the case that ν=j\nu=\ell^{j}, for \ell prime and j1j\geq 1. First, if HH_{\ell} is the Sylow \ell subgroup of HH, we have Sym2H=νSym2H\operatorname{Sym}^{2}H=\prod_{\ell\mid\nu}\operatorname{Sym}^{2}H_{\ell}. Using the universal property of products, we also have that for any abelian group AA, Hom(A,H)=νHom(A,H)\mathrm{Hom}(A,H)=\prod_{\ell\mid\nu}\mathrm{Hom}(A,H_{\ell}). Hence, we may assume that ν=j\nu=\ell^{j}. Instead of counting surjections, we can dually count injections from HH to any of the above three distributions.

Now, write Hi=1m/λiH\simeq\oplus_{i=1}^{m}\mathbb{Z}/\ell^{\lambda_{i}}\mathbb{Z}, so that HH is determined by a partition λ=(λ1,,λm)\lambda=(\lambda_{1},\ldots,\lambda_{m}). Let λ\lambda^{\prime} denote the partition conjugate to λ\lambda so that λi\lambda^{\prime}_{i} is the number of copies of /i\mathbb{Z}/\ell^{i}\mathbb{Z} appearing in HH. We first consider the case of computing injections HSeljBKLPRH\to\operatorname{Sel}^{\operatorname{BKLPR}}_{\ell^{j}}. The number of injective homomorphisms HSeljBKLPRH\to\operatorname{Sel}^{\operatorname{BKLPR}}_{\ell^{j}} can be expressed as the limit as nn\to\infty of the number of injections HZWH\to Z\cap W where Z,WOGrn(/j)Z,W\in\operatorname{OGr}_{n}(\mathbb{Z}/\ell^{j}\mathbb{Z}), for OGrn\operatorname{OGr}_{n} the orthogonal Grassmannian parameterizing nn-dimensional maximal isotropic subspaces in the rank 2n2n quadratic space with the split quadratic form i=1nxixi+n\sum_{i=1}^{n}x_{i}x_{i+n}. (This uses an alternate description of the BKLPR distribution from the one we gave in § 2.2.2, given in [BKL+15, §1.2 and 1.3]; see also [FLR23, §5.3.1] for a summary.) For fixed nn, we can express this as the number of injective homomorphisms h:HWh:H\to W times the probability that a uniformly random ZZ contains im(h)\operatorname{im}(h). We can compute both of these numbers by inductively computing the answer on t\ell^{t} torsion for each tjt\leq j.

First, we compute the number of injective homomorphisms h:HWh:H\to W. In the case t=1t=1, so t=\ell^{t}=\ell, this was shown in the proof of [BKL+15, Theorem 5.10] to be (n)λ1i=0λ11(1in)(\ell^{n})^{\lambda^{\prime}_{1}}\prod_{i=0}^{\lambda^{\prime}_{1}-1}(1-\ell^{i-n}). In general, a map HWH\to W is injective if and only if H[]WH[\ell]\to W is injective, so the number of injective maps H[t]WH[\ell^{t}]\to W lifting a given map H[t1]WH[\ell^{t-1}]\to W for t2t\geq 2 is (n)λt(\ell^{n})^{\lambda^{\prime}_{t}}. Recall we defined mm by Hi=1m/λiH\simeq\oplus_{i=1}^{m}\mathbb{Z}/\ell^{\lambda_{i}}\mathbb{Z}. Then, the total number of injective maps HWH\to W is

(2.2) ntλti=0m1(1in).\displaystyle\ell^{n\cdot\sum_{t}\lambda^{\prime}_{t}}\cdot\prod_{i=0}^{m-1}(1-\ell^{i-n}).

Next, we compute the probability that ZZ contains im(h),\operatorname{im}(h), for h:HWh:H\to W an injective homomorphism. First, the chance that ZZ contains im(H[])\operatorname{im}(H[\ell]) was computed in [BKL+15, Theorem 5.10] and it is

#OGrnm(/)#OGrn(/)=n(n1)2(nm)(nm1)2i=nmn1(1+i).\displaystyle\frac{\#\operatorname{OGr}_{n-m}(\mathbb{Z}/\ell\mathbb{Z})}{\#\operatorname{OGr}_{n}(\mathbb{Z}/\ell\mathbb{Z})}=\ell^{\frac{n(n-1)}{2}-\frac{(n-m)(n-m-1)}{2}}\prod_{i=n-m}^{n-1}(1+\ell^{-i}).

Let VV denote the quadratic space we are working in. Suppose we have fixed the image Z/t1ZV/t1VZ/\ell^{t-1}Z\subset V/\ell^{t-1}V containing h(H[t1])h(H[\ell^{t-1}]). We next compute the chance that Z/tZZ/\ell^{t}Z contains the image of h(H[t])h(H[\ell^{t}]) in V/tVV/\ell^{t}V. Since OGr\operatorname{OGr} is smooth of dimension n(n1)2\frac{n(n-1)}{2}, there are n(n1)2\ell^{\frac{n(n-1)}{2}} lifts of t1Z\ell^{t-1}Z to tZ\ell^{t}Z. The number of these containing imh(H[t])\operatorname{im}h(H[\ell^{t}]) can be identified with lifts of a maximal isotropic subspace of dimension nλtn-\lambda^{\prime}_{t}, since an isotropic subspace of WW containing a rank mm isotropic space TT can be identified with an isotropic subspace of the rank mnm-n space T/TT^{\perp}/T. There are (nλt)(nλt1)2\ell^{\frac{(n-\lambda^{\prime}_{t})(n-\lambda^{\prime}_{t}-1)}{2}} such subspaces. Hence, the chance Z/tZZ/\ell^{t}Z contains the image of h(H[t])h(H[\ell^{t}]) is (nλt)(nλt1)2n(n1)2=(λt)22nλt+λt2\ell^{\frac{(n-\lambda^{\prime}_{t})(n-\lambda^{\prime}_{t}-1)}{2}-\frac{n(n-1)}{2}}=\ell^{\frac{(\lambda^{\prime}_{t})^{2}-2n\lambda^{\prime}_{t}+\lambda^{\prime}_{t}}{2}}. Multiplying these probabilities over all values of tt up to jj, the chance ZZ contains h(H)h(H) is

(2.3) t=1j(λt)22nλt+λt2i=nmn1(1+i).\displaystyle\ell^{\sum_{t=1}^{j}\ell^{\frac{(\lambda^{\prime}_{t})^{2}-2n\lambda^{\prime}_{t}+\lambda^{\prime}_{t}}{2}}}\prod_{i=n-m}^{n-1}(1+\ell^{-i}).

Therefore, the moment we are seeking is the product of (2.2) with (2.3), which gives

ntλti=0m1(1in)t=1j(λt)22nλt+λt2i=nmn1(1+i)\displaystyle\ell^{n\cdot\sum_{t}\lambda^{\prime}_{t}}\cdot\prod_{i=0}^{m-1}(1-\ell^{i-n})\cdot\ell^{\sum_{t=1}^{j}\ell^{\frac{(\lambda^{\prime}_{t})^{2}-2n\lambda^{\prime}_{t}+\lambda^{\prime}_{t}}{2}}}\prod_{i=n-m}^{n-1}(1+\ell^{-i})
=t=1j(λt)2+λt2i=0m1(1in)i=nmn1(1+i).\displaystyle=\ell^{\sum_{t=1}^{j}\ell^{\frac{(\lambda^{\prime}_{t})^{2}+\lambda^{\prime}_{t}}{2}}}\prod_{i=0}^{m-1}(1-\ell^{i-n})\prod_{i=n-m}^{n-1}(1+\ell^{-i}).

As nn\to\infty, this approaches t=1j(λt)2+λt2\ell^{\sum_{t=1}^{j}\ell^{\frac{(\lambda^{\prime}_{t})^{2}+\lambda^{\prime}_{t}}{2}}}. A standard argument shows this agrees with #Sym2H\#\operatorname{Sym}^{2}H. For example, the analogous computation of the size of 2H\wedge^{2}H in place of Sym2H\operatorname{Sym}^{2}H was carried out in [Woo17, §2.4].

The cases of SelνBKLPR,1\operatorname{Sel}^{\operatorname{BKLPR},1}_{\nu} and SelνBKLPR,0\operatorname{Sel}^{\operatorname{BKLPR},0}_{\nu} follow similarly by only taking one of the components of the orthogonal Grassmannian, as also explained in [BKL+15, Remark 5.11]. ∎

2.4. Background on Hurwitz stacks

In this subsection, we give a precise definition of the Hurwitz stacks we will be working with. Throughout the paper, we will employ the following notation.

Notation 2.4.1.

Let BB be a base scheme. Let CBC\to B be a relative curve, which is smooth and proper of genus gg with geometrically connected fibers. Let ZCZ\subset C be a divisor, with ZZ finite étale over BB of degree f+1f+1, for f0f\geq 0. Let U:=CZU:=C-Z. The situation is summarized in the following diagram:

(2.4) Z{Z}C{C}U:=CZ{U:=C-Z}B.{B.}

Let n0n\geq 0 be an integer. Let SymC/Bn\operatorname{Sym}^{n}_{C/B} denote the relative nnth symmetric power of the curve CC over BB. Define ConfU/BnSymC/Bn\operatorname{Conf}^{n}_{U/B}\subset\operatorname{Sym}^{n}_{C/B} to be the open subscheme parameterizing effective divisors on CC which are finite étale of degree nn over BB and disjoint from ZZ. Let 𝒞BnConfU/Bn\>\mathscr{C}^{n}_{B}\to\operatorname{Conf}^{n}_{U/B} denote the universal curve, which has a universal degree nn divisor 𝒟Bn𝒞Bn\mathscr{D}^{n}_{B}\subset\mathscr{C}^{n}_{B} whose fiber over a point [D]ConfU/Bn[D]\in\operatorname{Conf}^{n}_{U/B} is DUD\subset U. Let 𝒰Bn:=𝒞Bn𝒟Bn(𝒞Bn×CZ)\mathscr{U}^{n}_{B}:=\mathscr{C}^{n}_{B}-\mathscr{D}^{n}_{B}-(\mathscr{C}^{n}_{B}\times_{C}Z) and let j:𝒰Bn𝒞Bnj:\mathscr{U}^{n}_{B}\subset\mathscr{C}^{n}_{B} denote the open inclusion. This setup is pictured in the next diagram:

(2.5) 𝒟Bn{\mathscr{D}^{n}_{B}}𝒞Bn{\mathscr{C}^{n}_{B}}𝒰Bn:=𝒞Bn𝒟Bn(𝒞Bn×CZ){\mathscr{U}^{n}_{B}:=\mathscr{C}^{n}_{B}-\mathscr{D}^{n}_{B}-(\mathscr{C}^{n}_{B}\times_{C}Z)}SymC/Bn{\operatorname{Sym}^{n}_{C/B}}ConfU/Bn{\operatorname{Conf}^{n}_{U/B}}B.{B.}j\scriptstyle{j}
Definition 2.4.2.

Keeping notation from 2.4.1, suppose BB is a scheme and GG is a finite group with #G\#G invertible on BB with chosen geometric point b¯B{\overline{b}}\in B. Suppose 𝒮Hom(π1(Σg,n+f+1),G)\mathcal{S}\subset\mathrm{Hom}(\pi_{1}(\Sigma_{g,n+f+1}),G) is a GG conjugation invariant subset preserved by the action of π1(ConfUb¯/b¯n)\pi_{1}(\operatorname{Conf}^{n}_{U_{\overline{b}}/\overline{b}}), acting on the first nn points. Define HurC/BG,n,Z,𝒮\operatorname{Hur}^{G,n,Z,\mathcal{S}}_{C/B} to be the stack over BB whose functor of points is defined as follows: For TT a BB-scheme, HurC/BG,n,Z,𝒮(T)\operatorname{Hur}^{G,n,Z,\mathcal{S}}_{C/B}(T) is the groupoid

(D,i:DCT,X,h:XCT)\displaystyle\left(D,i:D\to C_{T},X,h:X\to C_{T}\right)

satisfying the following conditions:

  1. (1)

    DD is a finite étale cover of TT of degree nn.

  2. (2)

    ii is a closed immersion i:DCTi:D\subset C_{T} which is disjoint from ZTCTZ_{T}\subset C_{T}.

  3. (3)

    XX is a smooth proper relative curve over T,T, not necessarily having geometrically connected fibers.

  4. (4)

    h:XCTh:X\to C_{T} is a finite locally free Galois GG-cover, (meaning that GG acts simply transitively on the geometric generic fiber of hh,) which is étale away from ZTi(D)CTZ_{T}\cup i(D)\subset C_{T}.

  5. (5)

    Let t¯T{\overline{t}}\to T be a fixed geometric point. Let η¯\overline{\eta} denote the geometric generic point of (CT)t¯(C_{T})_{\overline{t}}. Then the representation ρ:π1((UT)t¯i(Dt¯),η¯)G\rho:\pi_{1}((U_{T})_{\overline{t}}-i(D_{\overline{t}}),\overline{\eta})\to G afforded by hh, under the identification of Hom(π1((UT)t¯i(Dt¯),η¯),G)\mathrm{Hom}(\pi_{1}((U_{T})_{\overline{t}}-i(D_{\overline{t}}),\overline{\eta}),G) and Hom(π1(Σg,n+f+1),G)\mathrm{Hom}(\pi_{1}(\Sigma_{g,n+f+1}),G) corresponds to an element of 𝒮\mathcal{S}.

  6. (6)

    Two such covers are considered equivalent if they are related by the GG-conjugation action.

  7. (7)

    The morphisms between two points (Di,ii,Xi,hi)(D_{i},i_{i},X_{i},h_{i}) for i{1,2}i\in\{1,2\} are given by (ϕD,ψX)(\phi_{D},\psi_{X}) where ϕD:D1D2\phi_{D}:D_{1}\simeq D_{2} is an isomorphism so that i2ϕD=i2i_{2}\circ\phi_{D}=i_{2} and ψX:X1X2\psi_{X}:X_{1}\simeq X_{2} is an isomorphism such h2ψX=h1h_{2}\circ\psi_{X}=h_{1} and ψX=g1ψXg\psi_{X}=g^{-1}\psi_{X}g for every gGg\in G.

Remark 2.4.3.

The above Hurwitz stacks are algebraic by [AV02, Theorem 1.4.1]. Specifically, one can construct these Hurwitz stacks as an open substack of the quotient stack [𝒦g,n([C/G],Z,1)/Sn][{\mathcal{K}}_{g,n}([C/G],Z,1)/S_{n}], where 𝒦g,n([C/G],Z,1){\mathcal{K}}_{g,n}([C/G],Z,1) is defined in B.1.1.

Remark 2.4.4.

When GG is center free, the Hurwitz stacks parameterizing connected covers are indeed schemes, see [Wew98, Theorem 4]. However, we will consider Hurwitz stacks parameterizing disconnected covers, and, in this case, it is possible that those components may be stacks which are not schemes, even when GG is center free. This will actually occur in the cases we investigate in this paper.

The following pointed Hurwitz stack, which is a variant of the Hurwitz stack defined above, will be useful in connecting Hurwitz stacks to Hurwitz spaces over the complex numbers, described in terms of tuples of monodromy elements. See 3.1.10. We learned about the following slick construction from [Cha23].

Definition 2.4.5.

With notation as in 2.4.1, suppose there is a section σ:BC\sigma:B\to C with image contained in ZZ. Fix an integer ww and first define 𝒞(σ,w)\mathscr{C}_{(\sigma,w)} to be the root stack of order ww along σ\sigma, as defined in [Cad07, Definition 2.2.4]. The fiber of this root stack over σ\sigma is the stack quotient [(SpecB𝒪B[x]/(xr))/μr][\left(\operatorname{Spec}_{B}\mathscr{O}_{B}[x]/(x^{r})\right)/\mu_{r}] of the relative spectrum SpecB𝒪B[x]/(xr)\operatorname{Spec}_{B}\mathscr{O}_{B}[x]/(x^{r}) by μr\mu_{r}. Let σ~:B𝒞(σ,w)\widetilde{\sigma}:B\to\mathscr{C}_{(\sigma,w)} denote the section over σ\sigma corresponding to map B[(SpecB𝒪B[x]/(xr))/μr]B\to[\left(\operatorname{Spec}_{B}\mathscr{O}_{B}[x]/(x^{r})\right)/\mu_{r}] given by the trivial μr\mu_{r} torsor over BB, μrB\mu_{r}\to B, and the μr\mu_{r} equivariant map μrBSpecB𝒪B[x]/(xr)\mu_{r}\to B\to\operatorname{Spec}_{B}\mathscr{O}_{B}[x]/(x^{r}).

Define the ww-pointed Hurwitz stack, (HurC/BG,n,σZ,𝒮)w\left(\operatorname{Hur}^{G,n,\sigma\subset Z,\mathcal{S}}_{C/B}\right)^{w}, to be the stack whose groupoid of TT points is a setoid parameterizing data of the form

(D,h:X(𝒞(σ,w))T,t:TX×h,(𝒞(σ,w))T,σ~TT,i:DCT,X,h:XCT),\displaystyle\left(D,h^{\prime}:X\to(\mathscr{C}_{(\sigma,w)})_{T},t:T\to X\times_{h^{\prime},(\mathscr{C}_{(\sigma,w)})_{T},\widetilde{\sigma}_{T}}T,i:D\to C_{T},X,h:X\to C_{T}\right),

where D,i,X,D,i,X, and hh are as defined in 2.4.2. We also assume the order of inertia of hh along σ\sigma is ww and define σ~T\widetilde{\sigma}_{T} to be the base change of the section σ~\widetilde{\sigma} defined above to TT. We also impose the condition that hh^{\prime} is a finite locally free GG-cover, étale over σ~\widetilde{\sigma}, such that the composition of h:X(C(σ,w))Th^{\prime}:X\to\mathscr{(}C_{(\sigma,w)})_{T} with the coarse space map (𝒞(σ,w))TCT(\mathscr{C}_{(\sigma,w)})_{T}\to C_{T} is hh, and t:TX×(𝒞(σ,w))T,σ~TTt:T\to X\times_{(\mathscr{C}_{(\sigma,w)})_{T},\widetilde{\sigma}_{T}}T is a section of hh^{\prime} over σ~\widetilde{\sigma}.

In general, we define the pointed Hurwitz stack as HurC/BG,n,σZ,𝒮:=w1(HurC/BG,n,σZ,𝒮)w.\operatorname{Hur}^{G,n,\sigma\subset Z,\mathcal{S}}_{C/B}:=\coprod_{w\geq 1}\left(\operatorname{Hur}^{G,n,\sigma\subset Z,\mathcal{S}}_{C/B}\right)^{w}.

Remark 2.4.6.

It will be useful for later to note that there is a GG action on HurC/BG,n,σZ,𝒮\operatorname{Hur}^{G,n,\sigma\subset Z,\mathcal{S}}_{C/B} obtained by sending tt to gtg\circ t, for g:XXg:X\to X the automorphism corresponding to gGg\in G. By construction, the stack quotient [HurC/BG,n,σZ,𝒮/G][\operatorname{Hur}^{G,n,\sigma\subset Z,\mathcal{S}}_{C/B}/G] is HurC/BG,n,Z,𝒮\operatorname{Hur}^{G,n,Z,\mathcal{S}}_{C/B}.

Although we will not need the next remark it what follows, it may comfort the reader who is less familiar with stacks.

Remark 2.4.7.

In fact, HurC/BG,n,σZ,𝒮\operatorname{Hur}^{G,n,\sigma\subset Z,\mathcal{S}}_{C/B} is a scheme. One may verify this by proving it is a finite étale cover of ConfU/Bn\operatorname{Conf}^{n}_{U/B}.

We will see later that the complex points of Hurwitz stacks as in 2.4.5 admit a purely combinatorial description arising from actions of braid groups on finite sets. We turn to the relevant topology now.

3. The arc complex spectral sequence

In this section, we set up the spectral sequence which will relate various finite index subgroups of surface braid groups corresponding to Hurwitz spaces and allow induction arguments to take place. As usual in arguments of this kind, the decisive fact is the high degree of connectivity of a certain complex, provided to us in this case by a theorem of Hatcher and Wahl. In § 3.1 we define the basic objects, called coefficient systems, we will work with associated to surfaces. In § 4, we will show these coefficient systems have nice homological stability properties. In § 3.2 we set up the spectral sequence coming from the arc complex for these coefficient systems.

3.1. Defining coefficient systems

In this subsection, we define coefficient systems, which correspond to a certain kind of compatible sequence of local systems on the unordered configuration space of nn points on a topological surface with 11 boundary component, as nn varies. Later, we will show these have desirable homological stability properties. We are strongly guided here by the setup in [RWW17].

In order to define coefficient systems, which will be our basic objects guiding our study of homological stability, we begin by introducing some notation for surface braid groups.

Refer to caption
Figure 3. The blue surface with green punctures is a picture of A2,3Σ2,31A_{2,3}\simeq\Sigma^{1}_{2,3} and the black surface is XΣ0,02X\simeq\Sigma^{2}_{0,0}. The yellow circles correspond to the point xx, the red rectangles are the subsurface YY with xYXx\in Y\subset X. We also depict X3X^{\oplus 3} and X3A2,3X^{\oplus 3}\oplus A_{2,3}.
Notation 3.1.1.

Let Σg,fb\Sigma^{b}_{g,f} denote a genus gg topological surface with bb boundary components and ff punctures. For WW a topological space, we use ConfWn\operatorname{Conf}^{n}_{W} for the configuration space parameterizing tuples of nn unordered distinct points on WW. Let Ag,f:=Σg,f1A_{g,f}:=\Sigma^{1}_{g,f}, let X:=Σ0,02X:=\Sigma^{2}_{0,0}, and let xx be a point in the interior of XX. If we think of XX as /×[0,1]\mathbb{R}/\mathbb{Z}\times[0,1], we may place xx at (0,1/2)(0,1/2). With this same identification, we denote by YY the rectangle [1/4,1/4]×[0,1][-1/4,1/4]\times[0,1]. See Figure 3.

For n>0n>0, define the surface XnAg,fX^{\oplus n}\oplus A_{g,f}, which is homeomorphic to Σg,f1\Sigma^{1}_{g,f}, inductively by gluing the first boundary component of XX along a chosen isomorphism to the boundary component of Xn1Ag,fX^{\oplus{n-1}}\oplus A_{g,f}. We suggest the reader consult Figure 3 for a visualization. We denote by xnx^{\oplus n} the nn-element subset of XnAg,fX^{\oplus n}\oplus A_{g,f} obtained as the union of the copy of the point xx in each of the nn copies of XX. We also let XnX^{\oplus n} denote the complement of the interior of Ag,fA_{g,f} in XnAg,fX^{\oplus n}\oplus A_{g,f} and we let YnXnY^{\oplus n}\subset X^{\oplus n} denote the subsurface of XnX^{\oplus n} covered by the nn copies of YXY\subset X. Again, see Figure 3 for a visualization.

Now, let Bg,fn:=π1(ConfXnAg,fn,xn)B^{n}_{g,f}:=\pi_{1}(\operatorname{Conf}^{n}_{X^{\oplus n}\oplus A_{g,f}},x^{\oplus n}) denote the surface braid group. The natural map

Yi(XniAg,f)Xi(XniAg,f)XnAg,fY^{\oplus i}\coprod(X^{\oplus n-i}\oplus A_{g,f})\to X^{\oplus i}\coprod(X^{\oplus n-i}\oplus A_{g,f})\to X^{\oplus n}\oplus A_{g,f}

induces a map ConfYii×ConfXniAg,fniConfXnAg,fn\operatorname{Conf}^{i}_{Y^{\oplus i}}\times\operatorname{Conf}^{n-i}_{X^{\oplus n-i}\oplus A_{g,f}}\to\operatorname{Conf}^{n}_{X^{\oplus n}\oplus A_{g,f}} which sends xixnix^{\oplus i}\coprod x^{\oplus n-i} to xnx^{\oplus n}. We note that YiY^{\oplus i} is homeomorphic to a disc embedded in XiX^{\oplus i}, so the fundamental group of the configuration space ConfYii\operatorname{Conf}^{i}_{Y^{\oplus i}} is just the usual Artin braid group on ii strands. We thus get a map of fundamental groups

π1(ConfYii,xi)×π1(ConfXniAg,fni,xn1)π1(ConfXnAg,fn,xn)\pi_{1}(\operatorname{Conf}^{i}_{Y^{\oplus i}},x^{\oplus i})\times\pi_{1}(\operatorname{Conf}^{n-i}_{X^{\oplus n-i}\oplus A_{g,f}},x^{\oplus n-1})\to\pi_{1}(\operatorname{Conf}^{n}_{X^{\oplus n}\oplus A_{g,f}},x^{\oplus n})

or, in shorter terms, B0,0i×Bg,fniBg,fnB^{i}_{0,0}\times B^{n-i}_{g,f}\to B^{n}_{g,f}.

Remark 3.1.2.

By means of the homeomorphism between XnAg,fX^{\oplus n}\oplus A_{g,f} and Σg,f1\Sigma_{g,f}^{1}, we may think of Bg,fnB^{n}_{g,f} as the usual surface braid group on nn strands in a genus gg surface with ff punctures and a boundary component. We have chosen to define Bg,fnB^{n}_{g,f} in this more specific way because it will help us keep track of the maps between braid groups we will need to invoke.

Remark 3.1.3.

The reason for us introducing YY in 3.1.1, instead of just using XX, is to obtain an inclusion B0,0i=B0,0i×{1}B0,0i×Bg,fniBg,fnB^{i}_{0,0}=B^{i}_{0,0}\times\{1\}\subset B^{i}_{0,0}\times B^{n-i}_{g,f}\to B^{n}_{g,f}, which gives an inclusion from a braid group for a surface with 11 boundary component instead of from a surface with two boundary components. The key point of our homological stability results is that we will view certain systems of representations of Bg,fnB^{n}_{g,f} as modules-like objects for systems of representations of B0,0nB^{n}_{0,0}, and in order to define the module structure, the inclusion B0,0nBg,fnB^{n}_{0,0}\to B^{n}_{g,f} is essential.

We next define coefficient systems. Our definition of coefficient systems is inspired by [RWW17, Definition 4.1], though it is not exactly the same.

Definition 3.1.4.

For kk a field, a coefficient system for Σ0,01\Sigma^{1}_{0,0} is a sequence of kk vector spaces (Vn)n0(V_{n})_{n\geq 0} with actions B0,0n×VnVnB^{n}_{0,0}\times V_{n}\to V_{n} so that V0:=kV_{0}:=k, Vn:=V1nV_{n}:=V_{1}^{\otimes n}, and so that the B0,0nB^{n}_{0,0} action on VnV_{n} satisfies the following condition. For any 0in0\leq i\leq n, the diagram

(3.1) (B0,0i×B0,0ni)×ViVni{(B^{i}_{0,0}\times B^{n-i}_{0,0})\times V_{i}\otimes V_{n-i}}ViVni{V_{i}\otimes V_{n-i}}B0,0n×Vn{B^{n}_{0,0}\times V_{n}}Vn{V_{n}}

commutes, with maps described as follows: the right vertical map is induced by the isomorphism coming from the definition of VnV_{n}, the left vertical map is induced by this isomorphism together with the inclusion B0,0i×B0,0niB0,0nB^{i}_{0,0}\times B^{n-i}_{0,0}\to B^{n}_{0,0} described in 3.1.1, and the horizontal maps are induced by the given actions of B0,0jB^{j}_{0,0} on VjV_{j}.

Remark 3.1.5.

If (Vn)n0(V_{n})_{n\geq 0} is a coefficient system, then V1V_{1} naturally has the structure of a braided vector space coming from the action of a specified generator of B0,02B^{2}_{0,0}\simeq\mathbb{Z} on V2=V1V1V_{2}=V_{1}\otimes V_{1}. For any braided vector space VV, the tensor powers VnV^{\otimes n} acquire actions of B0,0nB^{n}_{0,0} satisfying (3.1). So, the definition of coefficient system for Σ0,01\Sigma^{1}_{0,0} is equivalent to that of a braided vector space.

We chose to set up 3.1.4 as we did so that its structure is analogous to that of coefficient systems for higher genus surfaces, which we define next.

Definition 3.1.6.

Fix a field kk and let VV be a fixed coefficient system for Σ0,01\Sigma^{1}_{0,0}. For g,f0g,f\geq 0, a coefficient system for Σg,f1\Sigma^{1}_{g,f} over VV is a sequence of kk vector spaces (Fn)n0(F_{n})_{n\geq 0} with actions Bg,fn×FnFnB^{n}_{g,f}\times F_{n}\to F_{n} so that Fn:=VnF0F_{n}:=V_{n}\otimes F_{0} and the Bg,fnB^{n}_{g,f} action on FnF_{n} satisfies the following condition. For any 0in0\leq i\leq n, the diagram

(3.2) (B0,0i×Bg,fni)×ViFni{(B^{i}_{0,0}\times B^{n-i}_{g,f})\times V_{i}\otimes F_{n-i}}ViFni{V_{i}\otimes F_{n-i}}Bg,fn×Fn{B^{n}_{g,f}\times F_{n}}Fn{F_{n}}

commutes, with maps described as follows: the right vertical map is an equality coming from the definition of FnF_{n}, the left vertical map is induced by the above equality and the inclusion B0,0i×Bg,fniBg,fnB^{i}_{0,0}\times B^{n-i}_{g,f}\to B^{n}_{g,f} described in 3.1.1, and the horizontal maps are induced by the given actions of B0,0jB^{j}_{0,0} on VjV_{j} and Bg,fjB^{j}_{g,f} on FjF_{j}.

Remark 3.1.7.

It is natural to think of coefficient systems (over VV) as a compatible sequence of local systems on ConfΣg,f1n\operatorname{Conf}^{n}_{\Sigma^{1}_{g,f}}. The compatibility condition amounts to commutativity of the diagram (3.2).

Remark 3.1.8.

Just as a braided vector space is determined by a finite amount of linear algebraic data (an endomorphism of V12V_{1}^{\otimes 2} satisfying a certain identity) it would be interesting to define a coefficient system for Σg,f1\Sigma^{1}_{g,f} over VV in a similar way, in the spirit of the definitions introduced by Hoang in [Hoa23, §3].

We next describe coefficient systems related to Hurwitz spaces, which come from maps from π1(Ag,f)\pi_{1}(A_{g,f}) to a finite group.

Example 3.1.9.

Fix g,f0g,f\geq 0. Let GG be a finite group and cc a conjugacy-closed subset of GG, and use notation as in 3.1.1. Choose a basepoint pg,fp_{g,f} on Ag,fA_{g,f}. Note that Bg,fnB^{n}_{g,f} acts on π1(XnAg,fxn,pg,f)\pi_{1}(X^{\oplus n}\oplus A_{g,f}-x^{\oplus n},p_{g,f}) and hence on Hom(π1(XnAg,fxn,pg,f),G)\mathrm{Hom}(\pi_{1}(X^{\oplus n}\oplus A_{g,f}-x^{\oplus n},p_{g,f}),G). Choose subsets TG,c,g,fnHom(π1(XnAg,fxn,pg,f),G)T^{n}_{G,c,g,f}\subset\mathrm{Hom}(\pi_{1}(X^{\oplus n}\oplus A_{g,f}-x^{\oplus n},p_{g,f}),G) so that TG,c,g,fn=c×TG,c,g,fn1T^{n}_{G,c,g,f}=c\times T^{n-1}_{G,c,g,f} and TG,c,g,fnT^{n}_{G,c,g,f} is closed under the action of Bg,fnB^{n}_{g,f} on Hom(π1(XnAg,fxn,pg,f),G)\mathrm{Hom}(\pi_{1}(X^{\oplus n}\oplus A_{g,f}-x^{\oplus n},p_{g,f}),G). Write HTG,c,g,fnH_{T^{n}_{G,c,g,f}} for the vector space freely spanned over kk by the subset Tg,f,G,cnHom(π1(XnAg,fxn,pg,f),G)T^{n}_{g,f,G,c}\subset\mathrm{Hom}(\pi_{1}(X^{\oplus n}\oplus A_{g,f}-x^{\oplus n},p_{g,f}),G).

Specializing the above to the case g=f=0g=f=0, the action of B0,0nB^{n}_{0,0} on T0,0,G,cnT^{n}_{0,0,G,c} induces an action of B0,0nB^{n}_{0,0} on HTG,c,0,0nH_{T^{n}_{G,c,0,0}}. We denote by HTG,c,0,0H_{T_{G,c,0,0}} the coefficient system VV for Σ0,01\Sigma^{1}_{0,0} given by Vn=HT0,0,G,cnV_{n}=H_{T^{n}_{0,0,G,c}}. This corresponds to the usual action of the Artin braid group on Nielsen tuples that underlies the classical combinatorial description of Hurwitz spaces of covers of the disc. Further, we denote by HTG,c,g,fH_{T_{G,c,g,f}} the coefficient system FF for Σg,f1\Sigma^{1}_{g,f} over HTG,c,0,0H_{T_{G,c,0,0}} given by Fn:=HTG,c,g,fnF_{n}:=H_{T^{n}_{G,c,g,f}}.

Warning 3.1.10.

We note that the cover of configuration space afforded by the coefficient system HTG,c,g,fH_{T_{G,c,g,f}} in 3.1.9 is not exactly the same thing as the space of complex points of the Hurwitz space defined in 2.4.2, but rather it is the complex points of the pointed Hurwitz space from 2.4.5. The sets TG,c,g,fnT^{n}_{G,c,g,f} carry an action of GG by conjugation and, via 2.4.6, the Hurwitz stack as in 2.4.2 is the quotient of the cover afforded by TG,c,g,fnT^{n}_{G,c,g,f} by this GG-action.

The reason we work more with this quotient is that it is easier to access from the point of view of moduli theory in algebraic geometry, while the unquotiented version is more suitable for the topological arguments we will make over the next several sections. This is easiest to see in the case (g,f)=(0,0)(g,f)=(0,0), where an element of TG,c,g,fnT^{n}_{G,c,g,f} is an nn-tuple of elements of cc. Then the concatenation operation cm×cncm+nc^{m}\times c^{n}\rightarrow c^{m+n} plays a key role in our arguments; but there is no well-defined concatenation on cm/G×cn/Gc^{m}/G\times c^{n}/G.

Example 3.1.11.

Take VV to be the coefficient system for Σ0,01\Sigma^{1}_{0,0} with Vi=kV_{i}=k and the trivial action for all ii. We call VV the trivial coefficient system for Σ0,01\Sigma^{1}_{0,0}. Let F0F_{0} be a vector space. Then Fi:=F0F_{i}:=F_{0} defines a coefficient system where the action of Bg,fnB^{n}_{g,f} on F0F_{0} is trivial.

Remark 3.1.12.

A different but related notion of “coefficient system” is considered in [RWW17]. Their precise definition doesn’t concern us here, but a property of the coefficient systems they consider (which they call finite degree) is that in their sequence of vector spaces V={Vn}n0V=\{V_{n}\}_{n\in\mathbb{Z}_{\geq 0}}, dimVn\dim V_{n} is eventually polynomial in nn. The coefficient system HTG,c,g,fH_{T_{G,c,g,f}} considered above, by contrast, have dimHTG,c,g,fn\dim H_{T^{n}_{G,c,g,f}} growing exponentially in nn. More precisely, the dimension grows proportionally to |c|n|c|^{n}. In general, our coefficient systems will have dimension which is bounded by a polynomial in nn only when dimV1=1\dim V_{1}=1, in which case the polynomial must even be a constant polynomial.

3.2. The spectral sequence

Our next main result is 3.2.4, which sets up a spectral sequence coming from the arc complex. In order to describe this, we first describe the 𝒦\mathcal{K} complex associated to a module.

Definition 3.2.1.

Let VV be a coefficient system for Σ0,01\Sigma^{1}_{0,0}. Let RV=n0H0(B0,0n,Vn)R^{V}=\oplus_{n\geq 0}H_{0}(B^{n}_{0,0},V_{n}), as in (4.1), which has the structure of a graded ring induced by the isomorphisms VsVrVr+sV_{s}\otimes V_{r}\to V_{r+s}. Let MM be a graded RVR^{V} module and let {M}n\{M\}_{n} denote the nnth graded part of MM. Let 𝒦(M)\mathcal{K}(M) denote the complex of graded RVR^{V} modules whose qqth term is 𝒦(M)q:=VqM[q]\mathcal{K}(M)_{q}:=V_{q}\otimes M[q]. That is, 𝒦(M)\mathcal{K}(M) is given by

VnM[n]V1M[1]M[0]\displaystyle\cdots\to V_{n}\otimes M[n]\to\cdots\to V_{1}\otimes M[1]\to M[0]

where M[i]M[i] denotes the shift by grading ii so that {M[i]}n={M}i+n\{M[i]\}_{n}=\{M\}_{i+n}. Here we treat ViV_{i} as living in degree 0 for all ii.

To define the differential, we next introduce some notation. Using τ\tau to denote the braiding automorphism of V1V1V_{1}\otimes V_{1} from 3.1.5, for 1i<n1\leq i<n, we let τin:V1nV1n\tau^{n}_{i}:V_{1}^{\otimes n}\to V_{1}^{\otimes n} denote the automorphism τin:=idi1τidni1\tau^{n}_{i}:=\operatorname{\mathrm{id}}^{\otimes i-1}\otimes\tau\otimes\operatorname{\mathrm{id}}^{\otimes n-i-1}, which applies τ\tau to the ii and i+1i+1 factors. For 1ijn1\leq i\leq j\leq n, we define τi,jn:=τj1nτi+1nτin\tau^{n}_{i,j}:=\tau_{j-1}^{n}\cdots\tau_{i+1}^{n}\tau_{i}^{n}. So, in particular, τin=τi,i+1n\tau^{n}_{i}=\tau^{n}_{i,i+1} and id=τi,in.\operatorname{\mathrm{id}}=\tau^{n}_{i,i}. We use μn:V1{M}n{M}n+1\mu_{n}:V_{1}\otimes\{M\}_{n}\to\{M\}_{n+1} to denote the multiplication map coming from the structure of MM as a RVR^{V}-module. As mentioned above, we use {M}n\{M\}_{n} to denote the nnth graded piece of a graded module MM, and then the differential on 𝒦(M)\mathcal{K}(M) is given by

(3.3) {𝒦(M)q+1}n\displaystyle\{\mathcal{K}(M)_{q+1}\}_{n} {𝒦(M)q}n\displaystyle\rightarrow\{\mathcal{K}(M)_{q}\}_{n}
(v0vq)m\displaystyle(v_{0}\otimes\cdots\otimes v_{q})\otimes m i=0q(1)i(idqμq)(τi+1,q+1q+1(v0vq)m).\displaystyle\mapsto\sum_{i=0}^{q}(-1)^{i}(\operatorname{\mathrm{id}}^{\otimes q}\otimes\mu_{q})\left(\tau_{i+1,q+1}^{q+1}(v_{0}\otimes\cdots\otimes v_{q})\otimes m\right).

The main case of 3.2.1 we will be interested in is when our module for RVR^{V} is of the form MpV,FM_{p}^{V,F}, which we now define.

Notation 3.2.2.

Given a coefficient system VV for Σ0,01\Sigma^{1}_{0,0} and a coefficient system FF for Σg,f1\Sigma^{1}_{g,f} over VV, define MpV,F:=n0Hp(Bg,fn,Fn)M_{p}^{V,F}:=\oplus_{n\geq 0}H_{p}(B^{n}_{g,f},F_{n}), where here the homology denotes group homology.

In the case our coefficient system is of the form MpV,FM_{p}^{V,F}, we next describe the map μn\mu_{n} concretely as well as the RVR^{V} module structure on MpV,FM_{p}^{V,F}.

Remark 3.2.3.

In the case we take our module for RVR^{V} in 3.2.1 to be MpV,FM_{p}^{V,F} from 3.2.2, we can describe the map μn:V1{MpV,F}n{MpV,F}n+1\mu_{n}:V_{1}\otimes\{M_{p}^{V,F}\}_{n}\to\{M_{p}^{V,F}\}_{n+1} concretely as follows. The inclusion Bg,fnBg,fn+1B^{n}_{g,f}\to B^{n+1}_{g,f} from 3.1.1 coming from the inclusion XnAg,fXn+1Ag,fX^{\oplus n}\oplus A_{g,f}\to X^{\oplus n+1}\oplus A_{g,f} induces a cup product map

V1Hp(Bg,fn,Fn)=H0(B0,01,V1)Hp(Bg,fn,Fn)Hp(Bg,fn+1,V1Fn)Hp(Bg,fn+1,Fn+1).\displaystyle V_{1}\otimes H_{p}(B^{n}_{g,f},F_{n})=H_{0}(B^{1}_{0,0},V_{1})\otimes H_{p}(B^{n}_{g,f},F_{n})\to H_{p}(B^{n+1}_{g,f},V_{1}\otimes F_{n})\simeq H_{p}(B^{n+1}_{g,f},F_{n+1}).

This composition is μn\mu_{n}. More generally, for nmn\geq m, the inclusions B0,0m×Bg,fnmBg,fnB^{m}_{0,0}\times B^{n-m}_{g,f}\to B^{n}_{g,f} from 3.1.1 give MpV,FM_{p}^{V,F} the structure of a RVR^{V} module via the cup product map

H0(B0,0i,Vi)Hp(Bg,fn,Fn)Hp(Bg,fn+i,ViFn)Hp(Bg,fn+1,Fn+i).\displaystyle H_{0}(B^{i}_{0,0},V_{i})\otimes H_{p}(B^{n}_{g,f},F_{n})\to H_{p}(B^{n+i}_{g,f},V_{i}\otimes F_{n})\simeq H_{p}(B^{n+1}_{g,f},F_{n+i}).

We now describe the spectral sequence coming from the arc complex. For a picture of the E2E^{2} page of this spectral sequence, see Figure 4. (We include the picture only in a later section as we believe it is helpful to see it side by side the proof of Theorem 4.1.1.)

Proposition 3.2.4.

Let VV be a coefficient system for Σ0,01\Sigma^{1}_{0,0} and let FF be a coefficient system for Σg,f1\Sigma^{1}_{g,f} over VV. There is a homological spectral sequence Eq,p1E^{1}_{q,p} converging to 0 in dimensions q+pn1q+p\leq n-1, where the ppth row (E,p1,d1)(E^{1}_{\ast,p},d_{1}) is isomorphic to the nnth graded piece of 𝒦(MpV,F)\mathcal{K}(M_{p}^{V,F}). That is, Eq,p1E^{1}_{q,p} is the nnth graded piece of 𝒦(MpV,F)q\mathcal{K}(M_{p}^{V,F})_{q} for p,q0.p,q\geq 0.

Proof.

The proof is a fairly immediate generalization of [EVW16, Proposition 5.1]. We now fill in some of the details. One minor difference is that we opt to use an augmented version of the arc complex so that the spectral sequence converges to 0, instead of 𝒦(MpV,F)\mathcal{K}(M_{p}^{V,F}) as in [EVW16, Proposition 5.1].

The spectral sequence will be obtained from filtering the arc complex by the dimension of its simplices. We begin by defining a combinatorial version of the arc complex, which we denote 𝔸(g,f,n)\mathbb{A}(g,f,n). For 1qn1-1\leq q\leq n-1, let LqBg,fnL_{q}\subset B^{n}_{g,f} denote the subgroup LqBg,fnq1L_{q}\simeq B^{n-q-1}_{g,f} obtained via the inclusion Bg,fnq1Bg,fnB^{n-q-1}_{g,f}\subset B^{n}_{g,f} coming from 3.1.1. If q=n1q=n-1, LqL_{q} is the trivial group. Define 𝔸(g,f,n)q:=Bg,fn/Lq\mathbb{A}(g,f,n)_{q}:=B^{n}_{g,f}/L_{q} as a Bg,fnB^{n}_{g,f} set. Define the faces of the qq-simplex bLqbL_{q} by i(bLq)=bsq,iLq1\partial_{i}(bL_{q})=bs_{q,i}L_{q-1} for 0iq0\leq i\leq q, where sq,i=σi+1σi+2σqs_{q,i}=\sigma_{i+1}\cdots\sigma_{i+2}\cdots\sigma_{q} and σi\sigma_{i} denotes an elementary transformation moving the iith point counterclockwise around the i+1i+1st point in π1(ConfΣg,f1n)Bg,fn\pi_{1}(\operatorname{Conf}^{n}_{\Sigma^{1}_{g,f}})\simeq B^{n}_{g,f}. Here, sq,q=1s_{q,q}=1. An identical computation to [EVW16, Proposition 5.3] shows ij=j1i\partial_{i}\partial_{j}=\partial_{j-1}\partial_{i} for i<ji<j, implying 𝔸(g,f,n)\mathbb{A}(g,f,n) is a semisimplicial set.

We next define the topological version of the arc complex, which we denote by 𝒜(g,f,n)\mathcal{A}(g,f,n) and we define next. Choose a finite set PP of points p1,,pnp_{1},\ldots,p_{n} in the interior of Σg,f1\Sigma^{1}_{g,f}. Let Σg,f1\star\in\Sigma^{1}_{g,f} denote a fixed basepoint lying on the boundary of Σg,f1\Sigma^{1}_{g,f}. Following Hatcher and Wahl [HW10, §7], we define a complex 𝒜(g,f,n)\mathcal{A}(g,f,n) as follows. A vertex of 𝒜(g,f,n)\mathcal{A}(g,f,n) is an embedded arc in Σg,f1\Sigma^{1}_{g,f} with one endpoint at \star and the other at some pip_{i}. For 1qn1-1\leq q\leq n-1, a qq-simplex of 𝒜(g,f,n)\mathcal{A}(g,f,n) is a collection of (q+1)(q+1) such arcs, which are disjoint away from \star. In particular, there are no simplices of dimension larger than n1n-1. Note that if we omit the 1-1 simplex, and only consider q0q\geq 0, the resulting complex 𝒜(g,f,n)q0\mathcal{A}(g,f,n)_{q\geq 0} is the complex denoted A(S,Λ0,Λn)A(S,\Lambda_{0},\Lambda_{n}) in [HW10, §7], with S=Σg,f1S=\Sigma^{1}_{g,f}, Λ0={}\Lambda_{0}=\{\star\}, and Λn={p1,,pn}.\Lambda_{n}=\{p_{1},\ldots,p_{n}\}. Hatcher and Wahl prove in [HW10, Proposition 7.2] that 𝒜(g,f,n)q0\mathcal{A}(g,f,n)_{q\geq 0} is (n2)(n-2)-connected. (Recall XX is nn-connected if πi(X)=0\pi_{i}(X)=0 for 1in-1\leq i\leq n.) Since the 1-1 skeleton of 𝒜(g,f,n)\mathcal{A}(g,f,n) is a point and 𝒜(g,f,n)q0\mathcal{A}(g,f,n)_{q\geq 0} is (n2)(n-2)-connected, both of these spaces have trivial homotopy groups in degree n2\leq n-2 and so the resulting boundary map 𝒜(g,f,n)q0𝒜(g,f,n)1\mathcal{A}(g,f,n)_{q\geq 0}\to\mathcal{A}(g,f,n)_{-1} yields a homotopy equivalence of spaces in degrees n2\leq n-2. We will soon construct a simplicial chain complex associated to 𝒜(g,f,n)\mathcal{A}(g,f,n), and the above implies it has trivial homology in degrees n2\leq n-2.

We now relate the two models 𝔸(g,f,n)\mathbb{A}(g,f,n) and 𝒜(g,f,n)\mathcal{A}(g,f,n) of the arc complex. In [EVW16, Proposition 5.6] a natural map 𝔸(g,f,n)𝒜(g,f,n)\mathbb{A}(g,f,n)\to\mathcal{A}(g,f,n), identifying these semisimplicial sets, was constructed when g=f=0g=f=0, and this readily generalizes to the case of arbitrary gg and f0f\geq 0.

We next describe the claimed spectral sequence. As in [EVW16, p. 757], for ZZ a space with a Bg,fnB^{n}_{g,f} action, we write Z//Bg,fnZ\mathbin{/\mkern-6.0mu/}B^{n}_{g,f} for the quotient, also known as the Borel construction EBg,fn×Bg,fnZEB^{n}_{g,f}\times_{B^{n}_{g,f}}Z. We will write k{𝔸(g,f,n)}k\{\mathbb{A}(g,f,n)\} to denote the free vector space on the simplices of 𝔸(g,f,n)\mathbb{A}(g,f,n), which is a Bg,fnB^{n}_{g,f} representation. Then, because 𝔸(g,f,n)=(𝔸(g,f,n)q0𝔸(g,f,n)1)\mathbb{A}(g,f,n)=\left(\mathbb{A}(g,f,n)_{q\geq 0}\to\mathbb{A}(g,f,n)_{-1}\right) is an equivalence of (n2)(n-2)-connected spaces, the map

Hp(k{𝔸(g,f,n)}Fn//Bg,fn)0\displaystyle H_{p}(k\{\mathbb{A}(g,f,n)\}\otimes F_{n}\mathbin{/\mkern-6.0mu/}B^{n}_{g,f})\to 0

is an isomorphism in degrees pn2p\leq n-2. That is, the left cohomology group vanishes for pn2p\leq n-2. We can also identify Hp(k{𝔸q(g,f,n)q}Fn//Bg,fn)H_{p}(k\{\mathbb{A}_{q}(g,f,n)_{q}\}\otimes F_{n}\mathbin{/\mkern-6.0mu/}B^{n}_{g,f}) with {𝒦(MpV,F)q+1}n\{\mathcal{K}(M_{p}^{V,F})_{q+1}\}_{n} via the isomorphisms

Hp(Bg,fn,k{𝔸q(g,f,n)}Fn)Hp(Lq,Fn)Hp(Bg,fnq1,V1q+1Fnq1){𝒦(MpV,F)q+1}n.\displaystyle H_{p}\left(B^{n}_{g,f},k\{\mathbb{A}_{q}(g,f,n)\}\otimes F_{n}\right)\simeq H_{p}(L_{q},F_{n})\simeq H_{p}(B^{n-q-1}_{g,f},V_{1}^{q+1}\otimes F_{n-q-1})\simeq\{\mathcal{K}(M_{p}^{V,F})_{q+1}\}_{n}.

Filtering k{𝔸(g,f,n)}Fn//Bg,fnk\{\mathbb{A}(g,f,n)\}\otimes F_{n}\mathbin{/\mkern-6.0mu/}B^{n}_{g,f} by the simplicial structure on 𝔸(g,f,n)\mathbb{A}(g,f,n), we obtain a spectral sequence

(3.4) Eq,p1:=Hp(k{𝔸(g,f,n)q}Fn//Bg,fn)Hp+q((𝔸(g,f,n)Fn)//Bg,fn).\displaystyle E^{1}_{q,p}:=H_{p}(k\{\mathbb{A}(g,f,n)_{q}\}\otimes F_{n}\mathbin{/\mkern-6.0mu/}B^{n}_{g,f})\implies H_{p+q}\left(\left(\mathbb{A}(g,f,n)\otimes F_{n}\right)\mathbin{/\mkern-6.0mu/}B^{n}_{g,f}\right).

Since 𝔸(g,f,n)q0\mathbb{A}(g,f,n)_{q\geq 0} is (n2)(n-2)-connected, 𝔸(g,f,n)\mathbb{A}(g,f,n) has trivial homology in degrees n2\leq n-2 and hence the right hand side of (3.4) vanishes for p+qn2p+q\leq n-2. Analogously to [EVW16, Lemma 5.4], one may verify that the differential d1:Eq,p1Eq1,p1d_{1}:E^{1}_{q,p}\to E^{1}_{q-1,p} is identified with the nnth graded part of the differential 𝒦(MpV,F)q𝒦(MpV,F)q1\mathcal{K}(M_{p}^{V,F})_{q}\to\mathcal{K}(M_{p}^{V,F})_{q-1} as in (3.3). The spectral sequence we have now constructed has bounds 1qn1-1\leq q\leq n-1. Replacing qq by q1q-1 gives 0qn0\leq q\leq n and yields the vanishing in degrees p+(q1)n2p+(q-1)\leq n-2, or equivalently p+qn1p+q\leq n-1. This gives desired spectral sequence, as in the statement. ∎

4. Deducing homological stability results for coefficient systems

In this section, we prove that certain types of coefficient systems have nice homological stability properties, following closely ideas from [EVW16]. In § 4.1 we give a general formulation of this stability property. In § 4.2 we show that finitely generated modules for coefficient systems with a suitable central element satisfy this stability property. Finally, in § 4.3 we put together all the topological material developed in this section and the previous one to arrive at an exponential bound on the cohomology of these coefficient systems.

For the reader primarily interested in our application to Selmer groups, only two results from this section needed in future parts. First, 4.3.4 will be used as a central ingredient in the proof of 8.2.3. Second, Theorem 4.2.2 will be used in the proof of Theorem A.5.1 to show the trace of Frobenius on the cohomology stabilizes.

4.1. Homological stability for 11-controlled coefficient systems

We next prove the main homological stability result of this paper in Theorem 4.1.1, using the arc complex spectral sequence from the previous section. To set things up in a general context, we define the notion of a 11-controlled coefficient system. For M=n{M}nM=\oplus_{n}\{M\}_{n} an object in a category with a \mathbb{Z} grading, we define degM\deg M to be the supremum of all nn such that {M}n0\{M\}_{n}\neq 0. Note that 𝒦(MpV,F)\mathcal{K}(M_{p}^{V,F}) has a grading by the number of points nn, and hence the same is true for Hi(𝒦(MpV,F))H_{i}(\mathcal{K}(M_{p}^{V,F})). The idea is that modules for 11-controlled coefficient systems have degrees of their iith homologies controlled in terms of degrees of their 0th and 11st homologies. For RR a graded ring, we say an element is homogeneous if it lies in a single degree of the grading of RR.

Definition 4.1.1.

Define

(4.1) RV:=n0H0(B0,0n,Vn).R^{V}:=\oplus_{n\geq 0}H_{0}(B^{n}_{0,0},V_{n}).

The monoidal structure of Σ0,01\Sigma^{1}_{0,0} supplies RVR^{V} with the structure of a graded ring supported in nonnegative gradings. Fix a homogeneous element 𝕌RV\mathbb{U}\in R^{V} of positive degree so that left multiplication by 𝕌\mathbb{U} induces a map 𝕌:RVRV\mathbb{U}:R^{V}\to R^{V}. A coefficient system VV for Σ0,01\Sigma^{1}_{0,0} is 11-controlled if degH0(𝒦(RV))\deg H_{0}(\mathcal{K}(R^{V})) and degH1(𝒦(RV))\deg H_{1}(\mathcal{K}(R^{V})) are finite and there exists a constant A0(V)1A_{0}(V)\geq 1 so that for any left RVR^{V}-module MM, the following two properties hold:

  1. (1)

    We have

    degHi(𝒦(M))max(degH0(𝒦(M)),degH1(𝒦(M)))+A0(V)i.\displaystyle\deg H_{i}(\mathcal{K}(M))\leq\max(\deg H_{0}(\mathcal{K}(M)),\deg H_{1}(\mathcal{K}(M)))+A_{0}(V)i.
  2. (2)

    The map induced by left multiplication by 𝕌\mathbb{U}, denoted 𝕌:MM\mathbb{U}:M\to M, induces an isomorphism {M}n{M}n+deg𝕌\{M\}_{n}\simeq\{M\}_{n+\deg\mathbb{U}} for

    nmax(degH0(𝒦(M)),degH1(𝒦(M)))+A0(V).\displaystyle n\geq\max(\deg H_{0}(\mathcal{K}(M)),\deg H_{1}(\mathcal{K}(M)))+A_{0}(V).

Next, we give a crucial example of a 11-controlled coefficient system.

Example 4.1.2.

Let GG be a group and cGc\subset G be a conjugacy class in GG. We will assume (G,c)(G,c) is non-splitting in the sense of [EVW16, Definition 3.1], meaning that cc generates GG and for every subgroup HGH\subset G, HcH\cap c either consists of a single conjugacy class in HH or is empty. Let V:=HTG,c,g,fV:=H_{T_{G,c,g,f}}, as defined in 3.1.9. As described in [EVW16, §3.3], the ring RVR^{V} is generated in degree 11 by elements of the form rgr_{g} (corresponding to right multiplication by gg) for gcg\in c. Consider the map 𝕌:=gcrgord(g)\mathbb{U}:=\sum_{g\in c}r_{g}^{\operatorname{ord}(g)}, where ord(g)\operatorname{ord}(g) denotes the order of gGg\in G. We will show in A.3.1, that kerU,coker𝕌:RVRV\ker U,\operatorname{coker}\mathbb{U}:R^{V}\to R^{V} both have finite degree. This is 11-controlled precisely by [EVW16, Theorem 4.2]. Note that the ring RV=n0H0(B0,0n,Vn)R^{V}=\oplus_{n\geq 0}H_{0}(B^{n}_{0,0},V_{n}) is called RR in [EVW16, Theorem 4.2].

The proof of this next result closely follows the proof of [EVW16, Theorem 6.1].

Theorem 4.1.1.

Suppose VV is a 11-controlled coefficient system for Σ0,01\Sigma^{1}_{0,0} and FF is a coefficient system for Σg,f1\Sigma^{1}_{g,f} over VV. Using notation as in 3.2.2, assume moreover that degH0(𝒦(M0V,F))\deg H_{0}(\mathcal{K}(M_{0}^{V,F})) and degH1(𝒦(M0V,F))\deg H_{1}(\mathcal{K}(M_{0}^{V,F})) are finite. Then, there exist constants I(V),J(F)I(V),J(F) depending on VV and FF but not on nn or pp so that 𝕌\mathbb{U} restricts to an isomorphism {MpV,F}n{MpV,F}n+deg𝕌\{M_{p}^{V,F}\}_{n}\to\{M_{p}^{V,F}\}_{n+\deg\mathbb{U}} whenever n>I(V)p+J(F)n>I(V)p+J(F).

Refer to caption
Figure 4. We depict the spectral sequence coming from the arc complex. To avoid clutter in the picture, we write 𝒦(Mi)\mathcal{K}(M_{i}) where we should write {𝒦(MiV,F)}n\{\mathcal{K}(M_{i}^{V,F})\}_{n}. The entries here start on the E2E^{2} page, so Eq,p2=Hq({𝒦(MpV,F)}n)E^{2}_{q,p}=H_{q}(\{\mathcal{K}(M_{p}^{V,F})\}_{n}). The blue arrows depict the differentials on the E2E^{2} page, the red arrows depict the differentials on the E3E^{3} page, and the green arrow depicts a differential on the E4E^{4} page.
Proof.

By way of induction on pp, we will prove there exist nonnegative constants A0(V)A_{0}(V) and C(F)C(F), independent of p,q,p,q, and nn, so that

(4.2) degHq(𝒦(MpV,F))C(F)+A0(V)(3p+q).\displaystyle\deg H_{q}(\mathcal{K}(M_{p}^{V,F}))\leq C(F)+A_{0}(V)(3p+q).

for all q0q\geq 0. Once we establish (4.2), we will obtain the result because, plugging in the cases q=0q=0 and q=1q=1, we get

degH0(𝒦(MpV,F))\displaystyle\deg H_{0}(\mathcal{K}(M_{p}^{V,F})) C(F)+A0(V)(3p)\displaystyle\leq C(F)+A_{0}(V)(3p)
degH1(𝒦(MpV,F))\displaystyle\deg H_{1}(\mathcal{K}(M_{p}^{V,F})) C(F)+A0(V)(3p+1).\displaystyle\leq C(F)+A_{0}(V)(3p+1).

Hence, by 4.1.1(2), we find 𝕌\mathbb{U} restricts to an isomorphism {MpV,F}n{MpV,F}n+deg𝕌\{M_{p}^{V,F}\}_{n}\to\{M_{p}^{V,F}\}_{n+\deg\mathbb{U}} whenever

nC(F)+1+A0(V)+3A0(V)p,\displaystyle n\geq C(F)+1+A_{0}(V)+3A_{0}(V)p,

and we can then take the constant I(V):=3A0(V)I(V):=3A_{0}(V) and J(F):=C(F)+1+A0(V)J(F):=C(F)+1+A_{0}(V).

We first verify (4.2) for p=0p=0. Indeed, let

C(F):=max(degH0(𝒦(M0V,F)),degH1(𝒦(M0V,F)),1).\displaystyle C(F):=\max(\deg H_{0}(\mathcal{K}(M_{0}^{V,F})),\deg H_{1}(\mathcal{K}(M_{0}^{V,F})),1).

By 4.1.1(1), we have degHq(𝒦(M0V,F))C(F)+A0(V)q\deg H_{q}(\mathcal{K}(M_{0}^{V,F}))\leq C(F)+A_{0}(V)q. This amounts to (4.2) for the case p=0p=0.

We next assume the result holds for p<Pp<P, and aim to show it holds for PP. It suffices to show

(4.3) degH0(𝒦(MPV,F))\displaystyle\deg H_{0}(\mathcal{K}(M_{P}^{V,F})) C(F)+3A0(V)P\displaystyle\leq C(F)+3A_{0}(V)P
degH1(𝒦(MPV,F))\displaystyle\deg H_{1}(\mathcal{K}(M_{P}^{V,F})) C(F)+3A0(V)P,\displaystyle\leq C(F)+3A_{0}(V)P,

as then 4.1.1(1), implies

degHq(𝒦(MP))C(F)+3A0(V)P+A0(V)q=C(F)+A0(V)(3P+q),\displaystyle\deg H_{q}(\mathcal{K}(M_{P}))\leq C(F)+3A_{0}(V)P+A_{0}(V)q=C(F)+A_{0}(V)(3P+q),

which is the inductive claim we wished to prove.

We conclude by proving (4.3). From 3.2.4, we can identify Eq,p2Hq({𝒦(MpV,F)}n)E^{2}_{q,p}\simeq H_{q}(\{\mathcal{K}(M_{p}^{V,F})\}_{n}). Therefore, it is enough to show E0,P2=E1,P2=0E^{2}_{0,P}=E^{2}_{1,P}=0 in degree n>C(F)+3A0(V)Pn>C(F)+3A_{0}(V)P. The differential coming into Eq,P2+iE^{2+i}_{q,P} comes from Eq+2+i,P1i2+iE^{2+i}_{q+2+i,P-1-i}, see Figure 4. By our inductive hypothesis, these vanish in degree n>C(F)+A0(V)(3(P1i)+(q+2+i))n>C(F)+A_{0}(V)(3(P-1-i)+(q+2+i)). When qq is either 0 or 11, we can bound

C(F)+A0(V)(3(P1i)+(q+2+i))\displaystyle C(F)+A_{0}(V)(3(P-1-i)+(q+2+i)) =C(F)+A0(V)(3P3+2+q2i)\displaystyle=C(F)+A_{0}(V)(3P-3+2+q-2i)
C(F)+A0(V)(3P1+q)\displaystyle\leq C(F)+A_{0}(V)(3P-1+q)
C(F)+A0(V)(3P).\displaystyle\leq C(F)+A_{0}(V)(3P).

Hence, once the degree nn satisfies C(F)+A0(V)(3P)<nC(F)+A_{0}(V)(3P)<n, we find Eq,P2=Eq,PE^{2}_{q,P}=E^{\infty}_{q,P} for qq either 0 or 11. Finally, Eq,P=0E^{\infty}_{q,P}=0 so long as P+qn1P+q\leq n-1, for nn the degree, by 3.2.4. Once we verify P+qn1P+q\leq n-1 and C(F)+3A0(V)Pn1C(F)+3A_{0}(V)P\leq n-1, we will conclude Eq,P2=0E^{2}_{q,P}=0. In particular, since we have assumed C(F)1C(F)\geq 1, and A0(V)1A_{0}(V)\geq 1 holds by from 4.1.1, we find P+qC(F)+3A0(V)PP+q\leq C(F)+3A_{0}(V)P, and so (4.3) holds so long as C(F)+3A0(V)P<nC(F)+3A_{0}(V)P<n. ∎

4.2. A sufficient condition for homological stability

We next set out to show that a wide variety of VV and FF satisfy the hypotheses of Theorem 4.1.1. We establish this in Theorem 4.2.2. For the purposes of this paper, our generalization of [EVW16, Theorem 4.2] given in Theorem 4.2.1 is not necessary, as we will only need to apply this to RVR^{V} coming from Hurwitz stacks, which is already proven in [EVW16, Theorem 4.2] applies. However, we include this generalization as we believe it may be useful for approaching similar homological stability problems in the future.

To start, we give a sufficient criterion for a ring to be 11-controlled in terms of a central operator 𝕌RV\mathbb{U}\in R^{V}. The following is the above mentioned generalization of [EVW16, Theorem 4.2].

Theorem 4.2.1.

Suppose VV is a coefficient system for Σ0,01\Sigma^{1}_{0,0} and define RV=n0H0(B0,0n,Vn)R^{V}=\oplus_{n\geq 0}H_{0}(B^{n}_{0,0},V_{n}) as in (4.1). Suppose 𝕌RV\mathbb{U}\in R^{V} is a homogeneous positive degree central element such that degker𝕌\deg\ker\mathbb{U} and degcoker𝕌\deg\operatorname{coker}\mathbb{U} are both finite. Then, VV is 11-controlled.

Proof.

This is essentially proved in [EVW16, Theorem 4.2]. While technically the ring RR used there is for a specific VV, the proof generalizes to the case stated here, as we now explain. Throughout the proof of [EVW16, Theorem 4.2], one may replace k[c]k[c] with V1V_{1}, and, for MM an RVR^{V} module, one may then use our definition of 𝒦(M)\mathcal{K}(M) from 3.2.1 in place of the definition in [EVW16, §4.1]. The two parts of the proof of [EVW16, Theorem 4.2] whose generalization requires some thought are the content of [EVW16, p. 755], where one wishes to establish the bound degTorRV1(k,M)degH1(𝒦(M))\deg\operatorname{Tor}^{1}_{R^{V}}(k,M)\leq\deg H_{1}(\mathcal{K}(M)), as well as [EVW16, Lemma 4.11]. Both of these refer to specific elements of the ring RR in [EVW16], which is related to Hurwitz stacks.

The only step of [EVW16, p. 755] where one cannot easily replace elements of k{c}k\{c\} with elements of V1V_{1} is in the third to last paragraph. To explain why this still holds, let α:V1kM[1]{RV}>0RVM\alpha:V_{1}\otimes_{k}M[1]\to\{R^{V}\}_{>0}\otimes_{R^{V}}M denote the map sending vm[v]mv\otimes m\mapsto[v]\cdot m, where [v][v] denotes the class of vv in {RV}1=H0(B1,V1)V1\{R^{V}\}_{1}=H_{0}(B_{1},V_{1})\simeq V_{1}, and [v]m[v]\cdot m denotes the multiplication using that MM is an RVR^{V} module. For xVnx\in V_{n}, we similarly use [x]m[x]\cdot m to denote the product of the class of xx in {RV}n\{R^{V}\}_{n} with mm. To establish the third to last paragraph of [EVW16, p. 755], we wish to verify that the composite map

V12kM[2]𝑑V1kM[1]𝛼{RV}>0RVM\displaystyle V_{1}^{\otimes 2}\otimes_{k}M[2]\xrightarrow{d}V_{1}\otimes_{k}M[1]\xrightarrow{\alpha}\{R^{V}\}_{>0}\otimes_{R^{V}}M

vanishes. For vwV12v\otimes w\in V_{1}^{\otimes 2}, if τ:V12V12\tau:V_{1}^{\otimes 2}\to V_{1}^{\otimes 2} denotes the isomorphism giving V1V_{1} the structure of a braided vector space, corresponding to a generator of B0,02B^{2}_{0,0}, we obtain that (αd)(vwm)=[vw]m[τ(vw)]m(\alpha\circ d)(v\otimes w\otimes m)=[v\otimes w]\cdot m-[\tau(v\otimes w)]\cdot m. This is equal to 0 because [vw]=[τ(vw)][v\otimes w]=[\tau(v\otimes w)] as elements of {RV}2=H0(B0,02,V2)\{R^{V}\}_{2}=H_{0}(B^{2}_{0,0},V_{2}): Indeed, a generator of B0,02B^{2}_{0,0} acts via τ\tau on V2V12V_{2}\simeq V_{1}^{\otimes 2}, so taking coinvariants via H0H_{0} identifies [vw][v\otimes w] and [τ(vw)][\tau(v\otimes w)].

To conclude, it remains to prove the analog of [EVW16, Lemma 4.11], which we do in 4.2.1. ∎

Lemma 4.2.1.

For VV a coefficient system, the action of {RV}>0\{R^{V}\}_{>0} on Hq(𝒦(RV))H_{q}(\mathcal{K}(R^{V})) is 0.

Proof.

We generalize the proof of the analogous statement given in [EVW16, Lemma 4.11]. Start with some element v1vqs{𝒦(RV)q}n=V1qH0(B0,0n,Vn)v_{1}\otimes\cdots\otimes v_{q}\otimes s\in\{\mathcal{K}(R^{V})_{q}\}_{n}=V_{1}^{\otimes q}\otimes H^{0}(B^{n}_{0,0},V_{n}). Define the linear operator

Sv:𝒦(RV)q\displaystyle S_{v}:\mathcal{K}(R^{V})_{q} 𝒦(RV)q+1\displaystyle\rightarrow\mathcal{K}(R^{V})_{q+1}
v1vqs~\displaystyle v_{1}\otimes\cdots\otimes v_{q}\otimes\widetilde{s} (τ1,q+n+1q+n+1)1(v1vqs~v)¯,\displaystyle\mapsto\overline{(\tau^{q+n+1}_{1,q+n+1})^{-1}(v_{1}\otimes\cdots\otimes v_{q}\otimes\widetilde{s}\otimes v)},

with notation as follows: we use notation as in 3.2.1, we use s~\widetilde{s} to denote a lift of ss from H0(B0,0n,Vn)H^{0}(B^{n}_{0,0},V_{n}) to VnV_{n}, and, for xV1q+1+n,x\in V_{1}^{\otimes q+1+n}, we use x¯\overline{x} for the image in V1q+1H0(B0,0n,Vn)V_{1}^{\otimes q+1}\otimes H^{0}(B^{n}_{0,0},V_{n}). First, we need to verify this map is independent of the choice of lift s~\widetilde{s} of ss. If we chose a different lift s~\widetilde{s}^{\prime}, we can write s~=σs~\widetilde{s}^{\prime}=\sigma\widetilde{s} for some σB0,0n\sigma\in B^{n}_{0,0}. Writing σ\sigma as a product of generators, we may assume σ=(τin)1\sigma=(\tau^{n}_{i})^{-1}. Now, for nmn\leq m and imni\leq m-n, define ιn,m,i:B0,0nB0,0m\iota_{n,m,i}:B^{n}_{0,0}\to B^{m}_{0,0} as the inclusion sending nn strands of B0,0nB^{n}_{0,0} to strands in the range [i+1,,i+n][i+1,\ldots,i+n]. More formally, this can be realized in terms of 3.1.1 as the inclusion

B0,0nB0,0i×B0,0n×B0,0minB0,0i×B0,0miB0,0m,\displaystyle B^{n}_{0,0}\to B^{i}_{0,0}\times B^{n}_{0,0}\times B^{m-i-n}_{0,0}\to B^{i}_{0,0}\times B^{m-i}_{0,0}\to B^{m}_{0,0},

where the first map is the inclusion to the second component, the second map is the product of B0,0iB^{i}_{0,0} with the map of braid groups associated to the inclusion XnXminA0,0XmiA0,0X^{\oplus n}\coprod X^{\oplus m-i-n}\oplus A_{0,0}\to X^{\oplus m-i}\oplus A_{0,0}, and the third map is the map of braid groups associated to the inclusion XiXmiA0,0XmA0,0X^{\oplus i}\coprod X^{\oplus m-i}\oplus A_{0,0}\to X^{\oplus m}\oplus A_{0,0}. The well definedness of SvS_{v} follows from the identity

(τ1,q+n+1q+n+1)1ιn,q+n+1,n((τin)1)\displaystyle(\tau^{q+n+1}_{1,q+n+1})^{-1}\iota_{n,q+n+1,n}((\tau^{n}_{i})^{-1}) =(τ1q+n+1)1(τn+qq+n+1)1(τn+iq+n+1)1\displaystyle=(\tau^{q+n+1}_{1})^{-1}\cdots(\tau^{q+n+1}_{n+q})^{-1}(\tau^{q+n+1}_{n+i})^{-1}
=(τn+i+1q+n+1)1(τ1q+n+1)1(τn+qq+n+1)1\displaystyle=(\tau^{q+n+1}_{n+i+1})^{-1}(\tau^{q+n+1}_{1})^{-1}\cdots(\tau^{q+n+1}_{n+q})^{-1}
=ιn,q+n+1,n+1((τin)1)(τ1,q+n+1q+n+1)1\displaystyle=\iota_{n,q+n+1,n+1}((\tau^{n}_{i})^{-1})(\tau^{q+n+1}_{1,q+n+1})^{-1}

applied to v1vqs~vv_{1}\otimes\cdots\otimes v_{q}\otimes\widetilde{s}\otimes v, as the above computation shows this maps to the same element as v1vqs~vv_{1}\otimes\cdots\otimes v_{q}\otimes\widetilde{s}^{\prime}\otimes v since their images in V1q+1+nV_{1}^{\otimes q+1+n} are related by ιn,q+n+1,n+1((τin)1)\iota_{n,q+n+1,n+1}((\tau^{n}_{i})^{-1}).

Since RVR^{V} is generated in degree 11, it is enough to prove right multiplication by [v][v] nullhomotopic. Having shown that SvS_{v} is well defined, we now compute

(Svd+dSv)(v1vqs)\displaystyle(S_{v}d+dS_{v})(v_{1}\otimes\cdots\otimes v_{q}\otimes s) =(idqμn)(τ1,q+1q+n+1(τ1,q+n+1q+n+1)1(v1vqs~v)¯)\displaystyle=(\operatorname{\mathrm{id}}^{\otimes q}\otimes\mu_{n})\left(\overline{\tau^{q+n+1}_{1,q+1}(\tau^{q+n+1}_{1,q+n+1})^{-1}(v_{1}\otimes\cdots\otimes v_{q}\otimes\widetilde{s}\otimes v)}\right)
=(idqμn)(τq+1,q+n+1q+n+1)1(v1vqs~v)¯\displaystyle=(\operatorname{\mathrm{id}}^{\otimes q}\otimes\mu_{n})\overline{(\tau^{q+n+1}_{q+1,q+n+1})^{-1}(v_{1}\otimes\cdots\otimes v_{q}\otimes\widetilde{s}\otimes v)}
=(v1vq)μn[(τ1,n+1n+1)1(s~v)]\displaystyle=(v_{1}\otimes\cdots\otimes v_{q})\otimes\mu_{n}[(\tau^{n+1}_{1,n+1})^{-1}(\widetilde{s}\otimes v)]
=v1vq(s[v]),\displaystyle=v_{1}\otimes\cdots\otimes v_{q}\otimes(s\cdot[v]),

which shows right multiplication by [v][v] is nullhomotopic. ∎

We next observe that RVR^{V} is noetherian. A similar argument in the context of Hurwitz stacks was given in [DS23, Proposition 3.31] and also [BM23, Lemma 3.3].

Lemma 4.2.2.

Let VV be a coefficient system for Σ0,01\Sigma^{1}_{0,0}. Suppose RV=n0H0(B0,0n,Vn)R^{V}=\oplus_{n\geq 0}H_{0}(B^{n}_{0,0},V_{n}) has some homogeneous positive degree 𝕌RV\mathbb{U}\in R^{V} so that degcoker𝕌\deg\operatorname{coker}\mathbb{U} is finite. Then RVR^{V} is noetherian.

Proof.

Note that RVR^{V} is not commutative. However, we claim RVR^{V} is a finite module over a commutative finitely generated ring, hence noetherian. Let R𝕌RVR_{\mathbb{U}}\subset R^{V} denote the commutative subring generated by 𝕌\mathbb{U} over kk. We claim RVR^{V} is a finite module over R𝕌R_{\mathbb{U}}. We will in fact show that RVR^{V} is generated over R𝕌R_{\mathbb{U}} by all elements of degree at most degcoker𝕌\deg\operatorname{coker}\mathbb{U}. Since each ViV_{i} is finite dimensional, this will imply that RVR^{V} is finitely generated over R𝕌R_{\mathbb{U}}. To prove our claim, by induction on the homogeneous degree of an element, it is enough to show that any homogeneous element rRVr\in R^{V} with degrdegcoker𝕌\deg r\geq\deg\operatorname{coker}\mathbb{U} can be written in the form s+𝕌ts+\mathbb{U}t for degs<degcoker𝕌\deg s<\deg\operatorname{coker}\mathbb{U} and degt<degr\deg t<\deg r. Indeed, consider the image r¯RV/𝕌RV\overline{r}\in R^{V}/\mathbb{U}R^{V}. Because RV/𝕌RV=coker𝕌R^{V}/\mathbb{U}R^{V}=\operatorname{coker}\mathbb{U} has finite degree, there is some element sRVs\in R^{V} of degree at most degcoker𝕌\deg\operatorname{coker}\mathbb{U} so that rs=0RV/𝕌RVr-s=0\in R^{V}/\mathbb{U}R^{V}. This implies rs=𝕌tr-s=\mathbb{U}t for some tRVt\in R^{V}, and hence r=s+𝕌tr=s+\mathbb{U}t with degt<degr\deg t<\deg r and degs<degcoker𝕌\deg s<\deg\operatorname{coker}\mathbb{U}. ∎

Using noetherianness of RVR^{V}, we can also prove the other hypotheses of Theorem 4.1.1 hold for finitely generated RVR^{V} modules.

Lemma 4.2.3.

Let VV be a coefficient system for Σ0,01\Sigma^{1}_{0,0}. Suppose RV=n0H0(B0,0n,Vn)R^{V}=\oplus_{n\geq 0}H_{0}(B^{n}_{0,0},V_{n}) has some homogeneous positive degree central 𝕌RV\mathbb{U}\in R^{V} so that degker𝕌\deg\ker\mathbb{U} and degcoker𝕌\deg\operatorname{coker}\mathbb{U} are both finite. Then, if NN is a finitely generated module over RVR^{V}, both H0(𝒦(N))H^{0}(\mathcal{K}(N)) and H1(𝒦(N))H^{1}(\mathcal{K}(N)) have finite degree.

Proof.

First, since RVR^{V} is generated in degree 11, H0(𝒦(N))=N/im(V1NN)=N/n>0{RV}nNH^{0}(\mathcal{K}(N))=N/\operatorname{im}(V_{1}\otimes N\to N)=N/\oplus_{n>0}\{R^{V}\}_{n}N, and this quotient is supported in the degrees of generators of NN over RVR^{V}. Therefore, NN is finitely generated, with each generator having degree at most dd, if and only if degH0(𝒦(N))d\deg H^{0}(\mathcal{K}(N))\leq d.

Next, we show degH1(𝒦(N))\deg H^{1}(\mathcal{K}(N)) is finite. Since 𝒦(N)=𝒦(RV)RVN\mathcal{K}(N)=\mathcal{K}(R^{V})\otimes_{R^{V}}N, there is a spectral sequence ToriRV(Hj(𝒦(RV))RVN)Hi+j(𝒦(N))\operatorname{Tor}_{i}^{R^{V}}(H_{j}(\mathcal{K}(R^{V}))\otimes_{R^{V}}N)\implies H_{i+j}(\mathcal{K}(N)). By the low degree terms exact sequence coming from the spectral sequence, in order to bound degH1(𝒦(N))\deg H_{1}(\mathcal{K}(N)), it is enough to bound degTor0RV(H1(𝒦(RV)),N)\deg\operatorname{Tor}_{0}^{R^{V}}(H_{1}(\mathcal{K}(R^{V})),N) and degTor1RV(H0(𝒦(RV)),N)\deg\operatorname{Tor}_{1}^{R^{V}}(H_{0}(\mathcal{K}(R^{V})),N). By Theorem 4.2.1, H0(𝒦(RV))H_{0}(\mathcal{K}(R^{V})) and H1(𝒦(RV))H_{1}(\mathcal{K}(R^{V})) have finite degree. In particular, they are finite kk modules. Hence it suffices to show degTor0RV(k,N)\deg\operatorname{Tor}_{0}^{R^{V}}(k,N) and degTor1RV(k,N)\deg\operatorname{Tor}_{1}^{R^{V}}(k,N) are finite. By noetherianness of RVR^{V}, as established in 4.2.2, we may choose a free resolution of the finite RVR^{V} module NN of the form S2S1N\cdots\to S_{2}\to S_{1}\to N where each term SiS_{i} is a finite free RVR^{V} module, hence of finite degree. Applying kRVk\otimes_{R^{V}} to this resolution and taking cohomology shows that ToriRV(k,N)\operatorname{Tor}_{i}^{R^{V}}(k,N) has finite degree for all ii. ∎

We next show that in the case N=M0V,FN=M_{0}^{V,F}, the finite generation hypothesis of 4.2.3 is automatic.

Lemma 4.2.4.

Suppose VV is a coefficient system for Σ0,01\Sigma^{1}_{0,0}. If FF is a coefficient system for Σg,f1\Sigma^{1}_{g,f} over VV, then M0V,FM_{0}^{V,F} is finitely generated as a RVR^{V} module.

Proof.

We may view M0V,FM_{0}^{V,F} as an RVR^{V} module via 3.2.3. Via the inclusion B0,0nBg,fnB^{n}_{0,0}\to B^{n}_{g,f} from 3.1.1, there is a surjection H0(B0,0n,Fn)H0(Bg,fn,Fn)H_{0}(B^{n}_{0,0},F_{n})\to H_{0}(B^{n}_{g,f},F_{n}). We therefore obtain a surjection of graded modules

n0H0(B0,0n,Fn)n0H0(Bg,fn,Fn)M0V,F.\displaystyle\oplus_{n\geq 0}H_{0}(B^{n}_{0,0},F_{n})\to\oplus_{n\geq 0}H_{0}(B^{n}_{g,f},F_{n})\to M_{0}^{V,F}.

Hence, it is enough to show n0H0(B0,0n,Fn)\oplus_{n\geq 0}H_{0}(B^{n}_{0,0},F_{n}) is finitely generated as an RVR^{V} module. Indeed, since B0,0nB^{n}_{0,0} acts trivially on F0F_{0},

n0H0(B0,0n,Fn)(n0H0(B0,0n,Vn))F0=RVF0,\displaystyle\oplus_{n\geq 0}H_{0}(B^{n}_{0,0},F_{n})\simeq\left(\oplus_{n\geq 0}H_{0}(B^{n}_{0,0},V_{n})\right)\otimes F_{0}=R^{V}\otimes F_{0},

and so the desired finite generation holds because F0F_{0} is a finite dimensional vector space. ∎

Combining our work above, we obtain that if we have coefficient systems VV and FF, and RVR^{V} has a central homogeneous element of positive degree with finite degree kernel and cokernel, then Theorem 4.1.1 applies.

Theorem 4.2.2.

Suppose VV is a coefficient system for Σ0,01\Sigma^{1}_{0,0} and 𝕌RV\mathbb{U}\in R^{V} is a homogeneous central element of positive degree such that degker𝕌\deg\ker\mathbb{U} and degcoker𝕌\deg\operatorname{coker}\mathbb{U} are both finite. If FF is a coefficient system for Σg,f1\Sigma^{1}_{g,f} over VV, then there exist constants I(V)I(V) and J(F)J(F) independent of pp and nn so that 𝕌\mathbb{U} induces an isomorphism {MpV,F}n{MpV,F}n+deg𝕌\{M_{p}^{V,F}\}_{n}\to\{M_{p}^{V,F}\}_{n+\deg\mathbb{U}} whenever n>I(V)p+J(F)n>I(V)p+J(F).

Proof.

This follows from Theorem 4.1.1, once we verify its hypotheses. We find RVR^{V} is 11-controlled by Theorem 4.2.1. From 4.2.4, M0V,FM_{0}^{V,F} is finitely generated as an RVR^{V} module. By 4.2.3, it follows that H0(𝒦(M0V,F))H^{0}(\mathcal{K}(M_{0}^{V,F})) and H1(𝒦(M0V,F))H^{1}(\mathcal{K}(M_{0}^{V,F})) both have finite degree. ∎

Remark 4.2.5.

Via private communication with Oscar Randal-Williams, it seems likely that one may be able to prove Theorem 4.2.2 using a setup similar to that in [RW20]. However, this is by no means obvious, and we believe it would be very interesting to work out the details. In particular, one of the trickiest parts to generalize is [RW20, Proposition 8.1] where it is used that B(k,A,A)=kB(k,A,A)=k. In our setting we need to instead analyze B(k,A,M)B(k,A,M), for a suitable value of MM in place of AA.

4.3. An exponential bound on the cohomology

Our main application of the above homological stability results to the BKLPR heuristics comes from the bound on cohomology in 4.3.3, and the corresponding consequence 4.3.4. There are two inputs. The first is our above homological stability results. The other is a bound on the CW structure of configuration space.

We now give this second bound, which nearly appears in [BS23, §4.2] in the case that f=0f=0. We now give the straightforward generalization to the case of arbitrary ff. We will be brief here, but encourage the reader to consult [BS23, §4.2] for further details. We thank Andrea Bianchi for suggesting the following approach.

Lemma 4.3.1.

For g,f,n0g,f,n\geq 0, the space ConfΣg,f1n\operatorname{Conf}^{n}_{\Sigma^{1}_{g,f}}, parameterizing nn unordered points in the interior of Σg,f1\Sigma^{1}_{g,f}, has 11-point compactification with a cell decomposition possessing at most 22g+f+n2^{2g+f+n} cells.

Refer to caption
Figure 5. This picture depicts a cell in the configuration space Conf12Σ1,21\operatorname{Conf}^{12}{\Sigma^{1}_{1,2}}. The boundary component consists of the union of the upper, left, and lower edges. The arrows indicate the orientations of the segments of the edges. Note that the segments of the same color are glued to each other with the orientations indicated, so there are only 44 distinct points represented by the yellow dots on the right boundary despite the fact that there are 88 yellow dots on the right boundary in the picture. The two black dots indicate the two punctures comprising WW. The yellow dots indicate the 1212 points in configuration space. The cell is labeled by the 1212-tuple 𝔱=((3,1,2,2),(2,1),(0,1))\mathfrak{t}=((3,1,2,2),(2,1),(0,1)) with b=4b=4.
Proof.

The idea is to generalize the construction of [BS23, §4.2] to the case that f>0f>0 as follows. We modify their setup so that the right edge of their rectangle 𝐑\mathbf{R} includes the intervals I1,I2,I1,I2,I3,,I2g,I2g1,I2gI_{1},I_{2},-I_{1},-I_{2},I_{3},\ldots,I_{2g},-I_{2g-1},-I_{2g}, as in the case f=0f=0, and then additionally includes the intervals I1,I1,I2,I2,,If,IfI^{\prime}_{1},-I^{\prime}_{1},I^{\prime}_{2},-I^{\prime}_{2},\ldots,I^{\prime}_{f},-I^{\prime}_{f} from bottom to top, see Figure 5.

We now spell this out in some more detail, reviewing the notation of [BS23, §4.2]. First, we describe Σg,f1\Sigma^{1}_{g,f} as a quotient in a particular way, which will be useful for describing a cellular structure on the one point compactification of its configuration space. Let 𝐑:=[0,2]×[0,1]\mathbf{R}:=[0,2]\times[0,1] be a rectangle. Decompose the side {2}×[0,1]\{2\}\times[0,1] into 4g+2f4g+2f consecutive intervals of equal length J1,,J4g,J1,,J2fJ_{1},\ldots,J_{4g},J^{\prime}_{1},\ldots,J^{\prime}_{2f} ordered and oriented with increasing second coordinate, as in Figure 5. Let WW be the set of the ff points consisting of the larger endpoint of J2i+1J^{\prime}_{2i+1} for 0if10\leq i\leq f-1. Let 𝐑W\mathbf{R}-W denote the punctured rectangle where we remove WW. Let \mathcal{M} denote the quotient of 𝐑W\mathbf{R}-W obtained by identifying J4i+1J_{4i+1} with J4i+3J_{4i+3}, J4i+2J_{4i+2} with J4i+4J_{4i+4}, and J2j+1J^{\prime}_{2j+1} with J2j+2J^{\prime}_{2j+2} via their unique orientation reversing isometry for 0ig10\leq i\leq g-1 and 0jf10\leq j\leq f-1 . Let 𝔭:𝐑W\mathfrak{p}:\mathbf{R}-W\to\mathcal{M} denote the quotient map. Then, \mathcal{M} is homeomorphic to Σg,f1\Sigma^{1}_{g,f}.

We next give a description of the cellular structure of \mathcal{M}. Throughout, for XX a topological space, we will use X\accentset{\circ}{X} to denote the interior of XX.

  1. (1)

    The space \mathcal{M} has a single 0 cell p0p_{0}, which is the image of any of the endpoints of the JiJ_{i}, and is also identified with the larger endpoint of J2j+2J^{\prime}_{2j+2}.

  2. (2)

    The space \mathcal{M} has 2g+f+12g+f+1 one-cells, described as follows. There are the 11-cells I2i+jI_{2i+j}, where I2i+j:=𝔭(J4i+j)I_{2i+j}:=\mathfrak{p}\left(\accentset{\circ}{J}_{4i+j}\right) with 0ig10\leq i\leq g-1 and j{1,2}j\in\{1,2\}. There are the 11-cells Ii=𝔭(J2i+1)I^{\prime}_{i}=\mathfrak{p}\left(\accentset{\circ}{J}^{\prime}_{2i+1}\right) for 0if10\leq i\leq f-1. Finally, there is I=𝔭(𝐑{2}×[0,1])I=\mathfrak{p}(\partial\mathbf{R}-\{2\}\times[0,1]).

  3. (3)

    Finally, \mathcal{M} has one 22-cell which is 𝔭(𝐑).\mathfrak{p}(\accentset{\circ}{\mathbf{R}}).

We let ιi:(0,1)\iota_{i}:(0,1)\to\mathcal{M} denote the composition of 𝔭\mathfrak{p} with the linear map sending (0,1)J4i+1(0,1)\to\accentset{\circ}{J}_{4i+1} for 0i2g10\leq i\leq 2g-1. We let ιi:(0,1)\iota^{\prime}_{i}:(0,1)\to\mathcal{M} denote the composition of 𝔭\mathfrak{p} with the linear map sending (0,1)J2i+1(0,1)\to\accentset{\circ}{J}^{\prime}_{2i+1} for 1if11\leq i\leq f-1. (This notation differs from that of [BS23, §4.2], but it is slightly more convenient for our purposes.)

We next introduce notation to define the cells in the CW complex we will construct. For n0n\geq 0, an nn-tuple, which we denote by 𝔱\mathfrak{t}, consists of

  1. (1)

    an integer b0b\geq 0

  2. (2)

    a sequence P¯=(P1,,Pb)\underline{P}=(P_{1},\ldots,P_{b}) of positive integers

  3. (3)

    a sequence 𝔳=(v1,,v2g)\mathfrak{v}=(v_{1},\ldots,v_{2g}) of non-negative integers

  4. (4)

    a sequence 𝔴=(w1,,wf)\mathfrak{w}=(w_{1},\ldots,w_{f}) of non-negative integers

such that P1++Pb+v1+v2g+w1++wf=nP_{1}+\cdots+P_{b}+v_{1}+\cdots v_{2g}+w_{1}+\cdots+w_{f}=n. The above data will index ways to split up nn points, representing a point of ConfΣg,f1n\operatorname{Conf}^{n}_{\Sigma^{1}_{g,f}}, into different cells of Σg,f1\Sigma^{1}_{g,f}.

We next define the cells determining a CW structure for the one point compactification of ConfΣg,f1n\operatorname{Conf}^{n}_{\Sigma^{1}_{g,f}}. We write 𝔱=(b,P¯,𝔳,𝔴)\mathfrak{t}=(b,\underline{P},\mathfrak{v},\mathfrak{w}) and use the notation for our surface \mathcal{M} described above. For 𝔱\mathfrak{t} an nn-tuple, let e𝔱e_{\mathfrak{t}} denote the subset of [S]ConfΣg,f1n[S]\in\operatorname{Conf}^{n}_{\Sigma^{1}_{g,f}} (which we recall parameterizes points in the interior of Σg,f1\Sigma^{1}_{g,f}) which satisfies the following conditions.

  1. (1)

    For 1i2g1\leq i\leq 2g, viv_{i} points lie on IiI_{i}.

  2. (2)

    For 1if1\leq i\leq f, wiw_{i} points lie in IiI^{\prime}_{i}.

  3. (3)

    There are exactly bb real numbers 0<x1<<xb<20<x_{1}<\cdots<x_{b}<2 such that SS admits at least on point in 𝐑\accentset{\circ}{\mathbf{R}} having xix_{i} as a coordinate.

  4. (4)

    For all 1ib1\leq i\leq b, exactly PiP_{i} points of SS which lie in 𝐑\accentset{\circ}{\mathbf{R}} have first coordinate equal to xix_{i}.

Each [S]ConfΣg,f1n[S]\in\operatorname{Conf}^{n}_{\Sigma^{1}_{g,f}} lies in a unique subspace e𝔱e_{\mathfrak{t}}. Given an nn-tuple 𝔱\mathfrak{t}, the space e𝔱e_{\mathfrak{t}} is homeomorphic to an open disc. Let d(𝔱)d(\mathfrak{t}) denote the dimension of this disc. Let Δk\Delta^{k} denote the standard kk-dimensional simplex. Define Δ𝔱:=Δb×i=1bΔPi×i=12gΔvi×i=1fΔwi\Delta^{\mathfrak{t}}:=\Delta^{b}\times\prod_{i=1}^{b}\Delta^{P_{i}}\times\prod_{i=1}^{2g}\Delta^{v_{i}}\times\prod_{i=1}^{f}\Delta^{w_{i}}. Using ConfΣg,f1d(𝔱){}\operatorname{Conf}^{d(\mathfrak{t})}_{\Sigma^{1}_{g,f}}\cup\{\infty\} to denote the 11-point compactification, for 𝔱\mathfrak{t} an nn-tuple, define the map Φt\Phi^{t} given in simplicial coordinates by

Φ𝔱:Δ𝔱ConfΣg,f1d(𝔱){}\displaystyle\Phi^{\mathfrak{t}}:\Delta^{\mathfrak{t}}\to\operatorname{Conf}^{d(\mathfrak{t})}_{\Sigma^{1}_{g,f}}\cup\{\infty\}
((zi)1ib,(sj(i))1ib,1jPi,(tj(i))1i2g,1jvi,(rj(i))1if,1jwi)\displaystyle\left((z_{i})_{1\leq i\leq b},(s^{(i)}_{j})_{1\leq i\leq b,1\leq j\leq P_{i}},(t^{(i)}_{j})_{1\leq i\leq 2g,1\leq j\leq v_{i}},(r^{(i)}_{j})_{1\leq i\leq f,1\leq j\leq w_{i}}\right)
[𝔭(2zj,sj(i)):1ib,1jPi][ιi(tj(i)):1i2g,1jvi]\displaystyle\mapsto\left[\mathfrak{p}(2z_{j},s_{j}^{(i)}):1\leq i\leq b,1\leq j\leq P_{i}\right]\cdot\left[\iota_{i}(t_{j}^{(i)}):1\leq i\leq 2g,1\leq j\leq v_{i}\right]
[ιi(rj(i)):1if,1jwi],\displaystyle\qquad\cdot\left[\iota^{\prime}_{i}(r_{j}^{(i)}):1\leq i\leq f,1\leq j\leq w_{i}\right],

where \cdot denotes the superposition product. The map Φ𝔱\Phi^{\mathfrak{t}} restricts to a homeomorphism sending the Δ𝔱e𝔱\accentset{\circ}{\Delta}^{\mathfrak{t}}\to e_{\mathfrak{t}} and the boundary Δ𝔱\partial\Delta^{\mathfrak{t}} to the union of {}\{\infty\} and some of the subspaces e𝔱e^{\mathfrak{t}^{\prime}} where d(𝔱)<d(𝔱)d(\mathfrak{t}^{\prime})<d(\mathfrak{t}).

As in [BS23, Proposition 4.4], one may verify the e𝔱e_{\mathfrak{t}} together with \infty form a cell decomposition for the one point compactification of ConfΣg,f1n\operatorname{Conf}^{n}_{\Sigma^{1}_{g,f}}.

Finally, we bound the number of cells in this structure by 2n+2g+f2^{n+2g+f}. Note that the number of cells is the same as the number of nn-tuples 𝔱\mathfrak{t}. A cell can equivalently be described by a choice of bb, and a collection of non-negative integers P11,,Pb1,v1,,v2g,w1,,wfP_{1}-1,\ldots,P_{b}-1,v_{1},\ldots,v_{2g},w_{1},\ldots,w_{f} summing to nbn-b. By “stars and bars,” such collections of integers are in bijection with subsets of {1,,(nb)+(b+2g+f)}={1,,n+2g+f}\{1,\ldots,(n-b)+(b+2g+f)\}=\{1,\ldots,n+2g+f\} of size b+2g+fb+2g+f. Varying over different possible values of bb yields that the total number of cells is equal to the number of subsets of {1,,n+2g+f}\{1,\ldots,n+2g+f\} of size at least 2g+f2g+f. This is at most the number of subsets of {1,,n+2g+f}\{1,\ldots,n+2g+f\}, which is 2n+2g+f2^{n+2g+f}, as we wished to show. ∎

As an easy consequence of the above bound on the number of cells, we obtain the following bound on homology.

Lemma 4.3.2.

Suppose VV is a coefficient system for Σ0,01\Sigma^{1}_{0,0} and FF is a coefficient system for Σg,f1\Sigma^{1}_{g,f} over VV. Then, dimHi(Bg,fn,Fn)22g+f+ndimFn.\dim H_{i}(B^{n}_{g,f},F_{n})\leq 2^{2g+f+n}\cdot\dim F_{n}.

Proof.

Since Bg,fnπ1(ConfΣg,f1n)B^{n}_{g,f}\simeq\pi_{1}(\operatorname{Conf}^{n}_{\Sigma^{1}_{g,f}}), the representation FnF_{n} of Bg,fnB^{n}_{g,f} corresponds to a local system 𝔽n\mathbb{F}_{n} on ConfΣg,f1n\operatorname{Conf}^{n}_{\Sigma^{1}_{g,f}} If ConfΣg,f1n{}\operatorname{Conf}^{n}_{\Sigma^{1}_{g,f}}\cup\{\infty\} denotes the 11-point compactification and j:ConfΣg,f1nConfΣg,f1n{}j:\operatorname{Conf}^{n}_{\Sigma^{1}_{g,f}}\to\operatorname{Conf}^{n}_{\Sigma^{1}_{g,f}}\cup\{\infty\}, denotes the inclusion, we have an isomorphism between the compactly supported cohomology and the relative cohomology

(4.4) Hci(ConfΣg,f1n,𝔽n)Hi((ConfΣg,f1n,),j!𝔽n).\displaystyle H^{i}_{\operatorname{c}}(\operatorname{Conf}^{n}_{\Sigma^{1}_{g,f}},\mathbb{F}_{n})\simeq H^{i}((\operatorname{Conf}^{n}_{\Sigma^{1}_{g,f}}\cup\infty,\infty),j_{!}\mathbb{F}_{n}).

We will now bound the dimension of this relative cohomology group. We will use the CW\operatorname{CW} cell structure on ConfΣg,f1n\operatorname{Conf}^{n}_{\Sigma^{1}_{g,f}}\cup\infty from 4.3.1 which has at most 22g+f+n2^{2g+f+n} cells. The cellular cochain complex which computes the iith cohomology group (4.4) has dimension less than rk𝔽n22g+f+n=dimFn22g+f+n\operatorname{rk}\mathbb{F}_{n}\cdot 2^{2g+f+n}=\dim F_{n}\cdot 2^{2g+f+n}. It follows from Poincaré duality that

dimH2dimXni(ConfΣg,f1n,𝔽n)=dimHci(ConfΣg,f1n,𝔽n)dimFn22g+f+n.\dim H_{2\dim X_{n}-i}(\operatorname{Conf}^{n}_{\Sigma^{1}_{g,f}},\mathbb{F}_{n})=\dim H^{i}_{\operatorname{c}}(\operatorname{Conf}^{n}_{\Sigma^{1}_{g,f}},\mathbb{F}_{n})\leq\dim F_{n}\cdot 2^{2g+f+n}.\qed

Combining our homological stability results with the above bounds on homology gives the following bound on cohomology. For the following, we continue to use notation from 3.2.2.

Proposition 4.3.3.

Let \ell^{\prime} be a prime. Suppose VV is a 11-controlled coefficient system for Σ0,01\Sigma^{1}_{0,0} and FF is a coefficient system for Σg,f1\Sigma^{1}_{g,f} over VV. Assume moreover that degH0(𝒦(M0V,F))\deg H_{0}(\mathcal{K}(M_{0}^{V,F})) and degH1(𝒦(M0V,F))\deg H_{1}(\mathcal{K}(M_{0}^{V,F})) are finite. Then, there is a constant KK depending on g,fg,f, and the sequence (Fn)n1(F_{n})_{n\geq 1}, but not on the subscript nn or the index ii so that

(4.5) dimHi(Bg,fn,Fn)Ki+1\displaystyle\dim H^{i}(B^{n}_{g,f},F_{n})\leq K^{i+1}

for all i,ni,n.

Proof.

Since the dimensions of the vector spaces in (4.5) are finite, and we are working with representations over a field, it follows from the universal coefficient theorem that dimHi(Bg,fn,Fn)=dimHi(Bg,fn,Fn)\dim H^{i}(B^{n}_{g,f},F_{n})=\dim H_{i}(B^{n}_{g,f},F_{n}). Hence, it is enough to bound dimHi(Bg,fn,Fn)Ki+1.\dim H_{i}(B^{n}_{g,f},F_{n})\leq K^{i+1}. By Theorem 4.1.1, there are constants I(V)I(V) and J(F)J(F) so that whenever n>I(V)i+J(F),n>I(V)i+J(F), Hi(Bg,fn,Fn)Hi(Bg,fn+deg𝕌,Fn+deg𝕌).H_{i}(B^{n}_{g,f},F_{n})\simeq H_{i}(B^{n+\deg\mathbb{U}}_{g,f},F_{n+\deg\mathbb{U}}). Therefore, applying this repeatedly, it is enough to show Hi(Bg,fn,Fn)Ki+1H_{i}(B^{n}_{g,f},F_{n})\leq K^{i+1} for any nI(V)i+J(F)+deg𝕌n\leq I(V)i+J(F)+\deg\mathbb{U}. By 4.3.2, Hi(Bg,fn,Fn)22g+f+ndimFnH_{i}(B^{n}_{g,f},F_{n})\leq 2^{2g+f+n}\cdot\dim F_{n}. Hence, we only need to produce some constant KK so that

22g+f+I(V)i+J(F)+deg𝕌dimFI(V)i+J(F)+deg𝕌Ki+1.\displaystyle 2^{2g+f+I(V)i+J(F)+\deg\mathbb{U}}\cdot\dim F_{I(V)i+J(F)+\deg\mathbb{U}}\leq K^{i+1}.

We may assume dimV1>0\dim V_{1}>0, as otherwise RV=kR^{V}=k and the statement is trivial. Because FnV1nF0F_{n}\simeq V_{1}^{\otimes n}\otimes F_{0},

22g+f+I(V)i+J(F)+deg𝕌dimFI(V)i+J(F)+deg𝕌\displaystyle 2^{2g+f+I(V)i+J(F)+\deg\mathbb{U}}\cdot\dim F_{I(V)i+J(F)+\deg\mathbb{U}}
=22g+f+J(F)+deg𝕌2I(V)i(dimV1)I(V)i+J(F)+deg𝕌dimF0\displaystyle=2^{2g+f+J(F)+\deg\mathbb{U}}\cdot 2^{I(V)i}\cdot(\dim V_{1})^{I(V)i+J(F)+\deg\mathbb{U}}\cdot\dim F_{0}
(2dimV1)I(V)i(2dimV1)2g+f+J(F)+deg𝕌dimF0.\displaystyle\leq(2\dim V_{1})^{I(V)i}\cdot(2\dim V_{1})^{2g+f+J(F)+\deg\mathbb{U}}\dim F_{0}.

The claim then follows by taking

K>max((2dimV1)I(V),(2dimV1)2g+f+J(F)+deg𝕌dimF0).K>\max((2\dim V_{1})^{I(V)},(2\dim V_{1})^{2g+f+J(F)+\deg\mathbb{U}}\dim F_{0}).\qed

We now reformulate the above in a slightly more convenient form for our applications.

Corollary 4.3.4.

Suppose VV is a 11-controlled coefficient system for Σ0,01\Sigma^{1}_{0,0} and FF is a coefficient system for Σg,f1\Sigma^{1}_{g,f} over VV. Assume that there is a central homogeneous positive degree element 𝕌RV\mathbb{U}\in R^{V} such that degker𝕌\deg\ker\mathbb{U} and degcoker𝕌\deg\operatorname{coker}\mathbb{U} are both finite. Suppose assume FnF_{n} corresponds to a local system 𝔽n\mathbb{F}_{n} on ConfΣg,f1n\operatorname{Conf}^{n}_{\Sigma^{1}_{g,f}} via the identification π1(ConfΣg,f1n)Bg,fn\pi_{1}(\operatorname{Conf}^{n}_{\Sigma^{1}_{g,f}})\simeq B^{n}_{g,f} with 𝔽n=π(/)\mathbb{F}_{n}=\pi_{*}(\mathbb{Z}/\ell^{\prime}\mathbb{Z}) for π:WnConfΣg,f1n\pi:W_{n}\to\operatorname{Conf}^{n}_{\Sigma^{1}_{g,f}} some finite étale cover of spaces over the complex numbers. Then, there is a constant KK depending on the sequence (Wn)n1(W_{n})_{n\geq 1} but not on the subscript nn or index ii so that

dimHi(Wn,/))Ki+1\displaystyle\dim H^{i}(W_{n},\mathbb{Z}/\ell^{\prime}\mathbb{Z}))\leq K^{i+1}

for all i,ni,n.

Proof.

This is an immediate consequence of 4.3.3, upon identifying group cohomology for a finite group with cohomology of the corresponding finite covering space, once we verify that VV is 11-controlled and degH0(𝒦(M0V,F))\deg H_{0}(\mathcal{K}(M_{0}^{V,F})) and degH0(𝒦(M0V,F))\deg H_{0}(\mathcal{K}(M_{0}^{V,F})) are finite. We have that VV is 11-controlled by Theorem 4.2.1. From 4.2.4, we find that M0V,FM_{0}^{V,F} is finitely generated as an RVR^{V} module. By 4.2.3, we find H0(𝒦(M0V,F))H^{0}(\mathcal{K}(M_{0}^{V,F})) and H1(𝒦(M0V,F))H^{1}(\mathcal{K}(M_{0}^{V,F})) both have finite degree. ∎

5. The Selmer stack and its basic properties

In this section, we set up the Selmer stack, which is a finite cover of the stack of quadratic twists of an abelian variety that parameterizes pairs of a quadratic twist and a Selmer element for that quadratic twist. We first define the Selmer stack in § 5.1. In § 5.3 we prove basic properties of the Selmer stack, such as the fact that it is a finite étale cover of the stack of quadratic twists. Since the definition given in § 5.1 is not obviously connected to Selmer groups, in § 5.3 we relate the Selmer stack to Selmer groups. Variants of the Selmer stack for the universal family were studied in [Lan21] and [FLR23], and many of the proofs in this section follow ideas from those articles.

5.1. Definition of the Selmer stack

We now set up notation to define the Selmer stack.

Definition 5.1.1.

Let XX be a Deligne-Mumford stack and ν\nu a positive integer. A locally constant constructible sheaf of free /ν\mathbb{Z}/\nu\mathbb{Z} modules \mathscr{F} on XX is symplectically self-dual if there is an isomorphism (1):=Hom(,μν)\mathscr{F}\simeq\mathscr{F}^{\vee}(1):=\mathrm{Hom}(\mathscr{F},\mu_{\nu}) so that the resulting pairing μν\mathscr{F}\otimes\mathscr{F}\to\mu_{\nu} factors through 2μν\mathscr{F}\otimes\mathscr{F}\to\wedge^{2}\mathscr{F}\to\mu_{\nu}.

Remark 5.1.2.

Sometimes, a symplectically self-dual sheaf is called a weight 11 symplectically self-dual sheaf. Since this is the only kind of symplectically self-dual sheaf we will encounter in our paper, so we omit the “weight 11” adjective. All symplectically self-dual sheaves we encounter will be assumed lcc sheaves of free /ν\mathbb{Z}/\nu\mathbb{Z} modules.

Example 5.1.3.

An important example of a symplectically self-dual sheaf for us will be A[ν]A[\nu], where AUA\to U is an abelian scheme as in 2.4.1 with a polarization of degree prime to ν\nu, for ν\nu invertible on BB.

Notation 5.1.4.

Keep notation for B,C,Z,U,n,fB,C,Z,U,n,f as in 2.4.1. Let \mathscr{F} be a tame symplectically self-dual sheaf on UU.

In order to define a Hurwitz stack for the group /2\mathbb{Z}/2\mathbb{Z}, let 𝒮Hom(π1(Σg,n+f+1),/2)\mathcal{S}\subset\mathrm{Hom}(\pi_{1}(\Sigma_{g,n+f+1}),\mathbb{Z}/2\mathbb{Z}) denote the subset sending loops around the geometric points in the degree f+1f+1 divisor ZZ to the trivial element of /2\mathbb{Z}/2\mathbb{Z} and loops around the nn marked points (corresponding to geometric points of the divisor DD) to the nontrivial element of /2\mathbb{Z}/2\mathbb{Z}. (Since /2\mathbb{Z}/2\mathbb{Z} is abelian, this Hurwitz stack is a /2\mathbb{Z}/2\mathbb{Z} gerbe over its coarse space.) Define QTwistU/Bn\operatorname{QTwist}^{n}_{U/B} to be HurC/B/2,n,Z,𝒮\operatorname{Hur}^{\mathbb{Z}/2\mathbb{Z},n,Z,\mathcal{S}}_{C/B}.

We will assume throughout nn is even, as otherwise there are no such covers by Riemann-Hurwitz. Informally, QTwistU/Bn\operatorname{QTwist}^{n}_{U/B} is a moduli space for finite double covers of CC ramified over a degree nn divisor DD, disjoint from ZZ. Let h:𝒰Bn×ConfU/BnQTwistU/Bn𝒰BnUh:\mathscr{U}^{n}_{B}\times_{\operatorname{Conf}^{n}_{U/B}}\operatorname{QTwist}^{n}_{U/B}\to\mathscr{U}^{n}_{B}\to U denote the composite projection and let λ:𝒞Bn×ConfU/BnQTwistU/BnQTwistU/Bn\lambda:\mathscr{C}^{n}_{B}\times_{\operatorname{Conf}^{n}_{U/B}}\operatorname{QTwist}^{n}_{U/B}\to\operatorname{QTwist}^{n}_{U/B} denote the universal proper curve. The universal open curve 𝒰Bn×ConfU/BnQTwistU/Bn\mathscr{U}^{n}_{B}\times_{\operatorname{Conf}^{n}_{U/B}}\operatorname{QTwist}^{n}_{U/B} possesses a natural finite étale double cover t:𝒳Bn,σ𝒰Bn×ConfU/BnQTwistU/Bnt:\mathscr{X}^{n,\sigma}_{B}\to\mathscr{U}^{n}_{B}\times_{\operatorname{Conf}^{n}_{U/B}}\operatorname{QTwist}^{n}_{U/B}, which is branched precisely along the boundary divisor 𝒟Bn×ConfU/BnQTwistU/Bn\mathscr{D}^{n}_{B}\times_{\operatorname{Conf}^{n}_{U/B}}\operatorname{QTwist}^{n}_{U/B} (but not along the preimage of ZZ).

Define Bn:=tth/h\mathscr{F}^{n}_{B}:=t_{*}t^{*}h^{*}\mathscr{F}/h^{*}\mathscr{F}. This is a sheaf on 𝒰Bn×ConfU/BnQTwistU/Bn\mathscr{U}^{n}_{B}\times_{\operatorname{Conf}^{n}_{U/B}}\operatorname{QTwist}^{n}_{U/B} whose fiber over x:=[(D,ϕ:π1(CD)/2)]QTwistU/Bnx:=[(D,\phi:\pi_{1}(C-D)\to\mathbb{Z}/2\mathbb{Z})]\in\operatorname{QTwist}^{n}_{U/B} is a sheaf on 𝒰Bn×ConfU/BnxU\mathscr{U}^{n}_{B}\times_{\operatorname{Conf}^{n}_{U/B}}x\subset U which is the quadratic twist of \mathscr{F} over UU along the finite étale double cover corresponding to the surjection ϕ\phi, which is branched over DD.

With the above notation in hand, we are now prepared to define the Selmer stack.

Definition 5.1.5.

Maintain notation as in 5.1.4 and let ν\nu be a positive integer. We assume 2ν2\nu is invertible on BB. As in 5.1.4, we have a symplectically self-dual sheaf \mathscr{F} on UU, which we are assuming is an lcc sheaf of free /ν\mathbb{Z}/\nu\mathbb{Z} modules. This gives rise to a symplectically self dual sheaf Bn\mathscr{F}^{n}_{B} on 𝒰Bn\mathscr{U}^{n}_{B} and maps

𝒰Bn𝑗𝒞Bn𝜆QTwistU/Bn.\displaystyle\mathscr{U}^{n}_{B}\xrightarrow{j}\mathscr{C}^{n}_{B}\xrightarrow{\lambda}\operatorname{QTwist}^{n}_{U/B}.

Define the Selmer sheaf of log-height nn associated to \mathscr{F} over BB to be 𝒮eBn:=R1λ(jBn){\mathcal{S}e\ell}_{\mathscr{F}^{n}_{B}}:=R^{1}\lambda_{*}\left(j_{*}\mathscr{F}^{n}_{B}\right). The Selmer stack, SelBn\operatorname{Sel}_{\mathscr{F}^{n}_{B}}, is the algebraic stack representing this étale sheaf.

Remark 5.1.6.

For odd ν\nu, the Selmer stack is never a scheme because QTwistU/Bn\operatorname{QTwist}^{n}_{U/B} is a /2\mathbb{Z}/2\mathbb{Z} gerbe over a scheme, and SelBn\operatorname{Sel}_{\mathscr{F}^{n}_{B}} is an odd degree cover of QTwistU/Bn\operatorname{QTwist}^{n}_{U/B}. Fortunately, since this is a gerbe, its stackiness is rather mild. This will pose some technical, yet overcomable, obstacles.

We next give a couple examples of types of symplectically self-dual sheaves coming from abelian varieties, which will be important for our applications to the BKLPR heuristics.

Example 5.1.7.

Suppose p:AUp:A\to U is a polarized abelian scheme with polarization of degree prime to ν\nu over BB. Take :=A[ν]\mathscr{F}:=A[\nu]. Note A[ν]A[ν]R1pμνA[\nu]\simeq A^{\vee}[\nu]\simeq R^{1}p_{*}\mu_{\nu}, since the polarization has degree prime to ν\nu. Then, the Weil pairing gives A[ν]A[\nu] the structure of a symplectically self-dual sheaf on UU. Further, with notation as in 5.1.4, A[ν]BnA[\nu]^{n}_{B} defines a sheaf on 𝒰Bn\mathscr{U}^{n}_{B}. An important example of a Selmer sheaf for us will be 𝒮eA[ν]Bn=R1λ(jA[ν]Bn){\mathcal{S}e\ell}_{A[\nu]^{n}_{B}}=R^{1}\lambda_{*}\left(j_{*}A[\nu]^{n}_{B}\right).

Example 5.1.8.

A slightly more general setup than 5.1.7 is the following. Suppose we are in the setting of 5.1.4, and bBb\in B is a closed point. Suppose we are given \mathscr{F} a symplectically self-dual sheaf over CC so that the fiber b\mathscr{F}_{b} over UbU_{b} defines a sheaf which is of the form Ab[ν]A_{b}[\nu] for p:AUbp:A\to U_{b} a polarized abelian scheme with polarization degree prime to ν\nu. Then we obtain a Selmer sheaf Bn\mathscr{F}^{n}_{B} over 𝒰Bn\mathscr{U}^{n}_{B} so that bn𝒮eA[ν]n\mathscr{F}^{n}_{b}\simeq{\mathcal{S}e\ell}_{A[\nu]^{n}}. The difference between this and 5.1.7 is that we may not have any abelian scheme over UU restricting to AA over UbU_{b}.

Remark 5.1.9.

In fact, the 5.1.8 will be the setting we work in to prove our main result Theorem 1.1.2 because it is relatively easy to lift symplectically self-dual sheaves from the closed point of a DVR to the whole DVR, as we explain in 10.2.2. However, we are unsure whether it is possible to lift abelian schemes in our setting.

We conclude this subsection with some notation recording data associated to a quadratic twist, which we will use throughout the paper.

Notation 5.1.10.

With notation as in 5.1.4, for xQTwistU/Bnx\in\operatorname{QTwist}^{n}_{U/B} a point or geometric point, let yy denote the image of xx under the map QTwistU/BnConfU/Bn\operatorname{QTwist}^{n}_{U/B}\to\operatorname{Conf}^{n}_{U/B}. We use CxC_{x} to denote the fiber of ξ:𝒞BnConfU/Bn\xi:\mathscr{C}^{n}_{B}\to\operatorname{Conf}^{n}_{U/B} over yy, UxU_{x} to denote the fiber of ξj\xi\circ j over yy, and we use x\mathscr{F}_{x} to denote the fiber of Bn\mathscr{F}^{n}_{B} over the point xx.

Assume we are further in the setup of 5.1.7 or 5.1.8 and xQTwistUb/bnx\in\operatorname{QTwist}^{n}_{U_{b}/b}. We use AxA_{x} to denote the fiber of the abelian scheme tthA/hAt_{*}t^{*}h^{*}A/h^{*}A over xx, where tt^{*} and hh^{*} denote the pullback along tt and hh, and tt_{*} denotes the Weil restriction along tt. Note that AxA_{x} is an abelian scheme over UxU_{x}. We use 𝒜x\mathscr{A}_{x} to denote the Néron model over CxC_{x} of AxUxA_{x}\to U_{x}. We let DxCxUxD_{x}\subset C_{x}-U_{x} denote the divisor associated to yy, the image of xx under the projection QTwistU/BnConfU/Bn\operatorname{QTwist}^{n}_{U/B}\to\operatorname{Conf}^{n}_{U/B}.

5.2. Basic properties of the Selmer stack

We next develop some basic properties of the Selmer stack. The next lemma shows the Selmer sheaf commutes with base change. The proof is similar to [FLR23, Lemma 2.6], though some additional technical difficulties come up related to working over the space of quadratic twists, instead of the universal family.

Lemma 5.2.1.

Use notation as in 5.1.4. In particular, \mathscr{F} is a tame symplectically self-dual sheaf. Suppose 2ν2\nu invertible on BB. Then, the sheaf 𝒮eBn{\mathcal{S}e\ell}_{\mathscr{F}^{n}_{B}} is locally constant constructible and its formation commutes with base change on QTwistU/Bn\operatorname{QTwist}^{n}_{U/B}. Further, for λ¯:=λj\overline{\lambda}:=\lambda\circ j, both Riλ¯(Bn)R^{i}\overline{\lambda}_{*}\left(\mathscr{F}^{n}_{B}\right) and Riλ¯!(Bn)R^{i}\overline{\lambda}_{!}\left(\mathscr{F}^{n}_{B}\right) are locally constant constructible for all i0i\geq 0 and their formation commutes with base change on QTwistU/Bn\operatorname{QTwist}^{n}_{U/B}.

Proof.

In order to prove the result, we first set some notation. We have a natural map ϕ:R1λ¯!Bn𝒮eBn\phi:R^{1}\overline{\lambda}_{!}\mathscr{F}^{n}_{B}\to{\mathcal{S}e\ell}_{\mathscr{F}^{n}_{B}} obtained from the map j!BnjBnj_{!}\mathscr{F}^{n}_{B}\to j_{*}\mathscr{F}^{n}_{B} and the definition R1λ¯!(Bn):=R1λ(j!Bn)R^{1}\overline{\lambda}_{!}(\mathscr{F}^{n}_{B}):=R^{1}\lambda_{*}\left(j_{!}\mathscr{F}^{n}_{B}\right) [FK88, I.8.6]. Similarly, we have a map ψ:𝒮eBnR1λ¯Bn\psi:{\mathcal{S}e\ell}_{\mathscr{F}^{n}_{B}}\to R^{1}\overline{\lambda}_{*}\mathscr{F}^{n}_{B} obtained from the composition of functors spectral sequence for λj\lambda\circ j. Note that ψ\psi is injective by the Leray spectral sequence.

Our first goal is to show 𝒮eBn{\mathcal{S}e\ell}_{\mathscr{F}^{n}_{B}} is the image of ψϕ\psi\circ\phi. Since ψ\psi is injective, it only remains to show ϕ\phi is surjective. Because χ:j!BnjBn\chi:j_{!}\mathscr{F}^{n}_{B}\to j_{*}\mathscr{F}^{n}_{B} is an isomorphism over 𝒰Bn\mathscr{U}^{n}_{B}, cokerχ\operatorname{coker}\chi is supported on 𝒟Bn\mathscr{D}^{n}_{B}, which is finite over ConfU/Bn\operatorname{Conf}^{n}_{U/B}, we find R1λ(cokerχ)=0.R^{1}\lambda_{*}(\operatorname{coker}\chi)=0. This implies ϕ\phi is surjective and so 𝒮eBn{\mathcal{S}e\ell}_{\mathscr{F}^{n}_{B}} is a constructible sheaf.

We conclude by showing R1λ¯!BnR^{1}\overline{\lambda}_{!}\mathscr{F}^{n}_{B} and R1λ¯(Bn)R^{1}\overline{\lambda}_{*}\left(\mathscr{F}^{n}_{B}\right) are both locally constant constructible, and their formation commutes with base change. This will imply SelBn\operatorname{Sel}_{\mathscr{F}^{n}_{B}} is locally constant constructible and its formation commutes with base change, as it is the image of the map ψϕ:R1λ¯!BnR1λ¯(Bn)\psi\circ\phi:R^{1}\overline{\lambda}_{!}\mathscr{F}^{n}_{B}\to R^{1}\overline{\lambda}_{*}\left(\mathscr{F}^{n}_{B}\right).

We first show Riλ¯!BnR^{i}\overline{\lambda}_{!}\mathscr{F}^{n}_{B} is locally constant constructible in the case that ν\nu is prime. Note that its formation commutes with base change by proper base change for any ν\nu. Using [Lau81, Corollaire 2.1.2 and Remarque 2.1.3], it is enough to show the Swan conductor of Bn\mathscr{F}^{n}_{B} is constant. As in [Lau81, Remarque 2.1.3], the Swan conductor over a point [D]ConfU/Bn[D]\in\operatorname{Conf}^{n}_{U/B} is a sum of local contributions, one for each geometric point of DD and one for each geometric point of ZZ over the image of DD in BB. At each geometric point of DD, because we are taking a quadratic twist along DD, the ramification index is 22, and hence the ramification is tame, since 22 is invertible on BB. We are also assuming the ramification along points of ZZ is tame for \mathscr{F}. This is identified with the corresponding ramification for Bn\mathscr{F}^{n}_{B} along points of ZZ, and hence this is tame as well. Therefore, the Swan conductor vanishes identically.

Next, we show Riλ¯!BnR^{i}\overline{\lambda}_{!}\mathscr{F}^{n}_{B} is locally constant constructible for every positive integer ν\nu as in the statement of the lemma, using the case that ν\nu is prime, settled above. As an initial step, we may reduce to the case ν=t\nu=\ell^{t} is a prime power by observing that if ν\nu has prime factorization ν=t\nu=\prod\ell^{t_{\ell}} then μνμt\mu_{\nu}\simeq\oplus\mu_{\ell}^{t_{\ell}}. Now, suppose ν=t\nu=\ell^{t} is a prime power, and inductively assume we have proven Riλ¯!Bn[t1]R^{i}\overline{\lambda}_{!}\mathscr{F}^{n}_{B}[\ell^{t-1}] is locally constant constructible for all ii. Since ν=t\nu=\ell^{t} and Bn\mathscr{F}^{n}_{B} is a locally constant constructible sheaf of free /ν\mathbb{Z}/\nu\mathbb{Z} modules, we have an exact sequence

0{0}Bn[t1]{\mathscr{F}^{n}_{B}[\ell^{t-1}]}Bn{\mathscr{F}^{n}_{B}}Bn[]{\mathscr{F}^{n}_{B}[\ell]}0.{0.}

Applying Rλ¯!R\overline{\lambda}_{!} to the above sequence, we get a long exact sequence on cohomology

Ri1λ¯!Bn[]{R^{i-1}\overline{\lambda}_{!}\mathscr{F}^{n}_{B}[\ell]}Riλ¯!Bn[t1]{R^{i}\overline{\lambda}_{!}\mathscr{F}^{n}_{B}[\ell^{t-1}]}Riλ¯!Bn{R^{i}\overline{\lambda}_{!}\mathscr{F}^{n}_{B}}Riλ¯!Bn[]{R^{i}\overline{\lambda}_{!}\mathscr{F}^{n}_{B}[\ell]}Ri+1λ¯!Bn[t1].{R^{i+1}\overline{\lambda}_{!}\mathscr{F}^{n}_{B}[\ell^{t-1}].}

Since all but the middle term are locally constant constructible by our inductive assumption, it follows that Riλ¯!(Bn[t])R^{i}\overline{\lambda}_{!}(\mathscr{F}^{n}_{B}[\ell^{t}]) is also locally constant constructible by [Sta, Tag 093U].

We conclude by showing R1λ¯(Bn)R^{1}\overline{\lambda}_{*}\left(\mathscr{F}^{n}_{B}\right) is locally constant constructible and its formation commutes with base change. Since Riλ¯!BnR^{i}\overline{\lambda}_{!}\mathscr{F}^{n}_{B} is locally constant constructible, it follows from Poincaré duality [Ver67, Theorem 4.8] and the isomorphism coming from the polarization of degree prime to ν\nu that

om(Riλ¯!(Bn),μν)Ri+2λ¯Rom(Bn,μν)Ri+2λ¯(Bn).\displaystyle\mathscr{H}\kern-0.5ptom\left(R^{-i}\overline{\lambda}_{!}\left(\mathscr{F}^{n}_{B}\right),\mu_{\nu}\right)\xleftarrow{\simeq}R^{i+2}\overline{\lambda}_{*}R\mathscr{H}\kern-0.5ptom(\mathscr{F}^{n}_{B},\mu_{\nu})\simeq R^{i+2}\overline{\lambda}_{*}(\mathscr{F}^{n}_{B}).

Taking i=2+si=-2+s gives (R2sλ¯!(Bn))(1)Rsλ¯(Bn)(R^{2-s}\overline{\lambda}_{!}(\mathscr{F}^{n}_{B}))^{\vee}(1)\simeq R^{s}\overline{\lambda}_{*}\left(\mathscr{F}^{n}_{B}\right). Since we have seen (R2sλ¯!(Bn))(R^{2-s}\overline{\lambda}_{!}(\mathscr{F}^{n}_{B}))^{\vee} is locally constant constructible and its formation commutes with base change, the same holds for Rsλ¯(Bn)R^{s}\overline{\lambda}_{*}\left(\mathscr{F}^{n}_{B}\right). ∎

Notation 5.2.2.

Let kk be a field and let CC be a smooth proper geometrically connected curve over kk of genus gg, with UCU\subset C an open subscheme. Let AA^{\prime} an abelian scheme over UU^{\prime} with Néron model 𝒜C\mathscr{A}^{\prime}\to C. Let 𝒜0\mathscr{A}^{\prime 0} denote the identity component of the Néron model 𝒜\mathscr{A}^{\prime}, meaning that 𝒜0\mathscr{A}^{\prime 0} is the open subscheme of 𝒜\mathscr{A}^{\prime} so that each fiber is nonempty, connected, and contains the identity section, see [BLR90, p. 154]. Let ΦA:=(𝒜/𝒜0)(k)\Phi_{A^{\prime}}:=\left(\mathscr{A}^{\prime}/\mathscr{A}^{\prime 0}\right)(k) denote the component group of the Néron model of AA^{\prime}. We use ΦAk¯=(𝒜k¯/𝒜k¯0)(k¯)\Phi_{A^{\prime}_{\overline{k}}}=\left(\mathscr{A}^{\prime}_{\overline{k}}/\mathscr{A}^{\prime 0}_{\overline{k}}\right)(\overline{k}) to denote the geometric component group.

The following proof is quite similar to [Lan21, Lemma 3.21]. We thank Tony Feng for suggesting the idea that appeared there for bootstrap from the prime case to the general case, which we reuse here. In the next lemma, note that since we are working over an algebraically closed field, the component group is the same as the geometric component group.

Lemma 5.2.3.

Let kk be an algebraically closed field, let CC be a smooth proper geometrically connected curve over kk of genus gg. Let \mathscr{F}^{\prime} be a symplectically self-dual lcc sheaf of free /ν\mathbb{Z}/\nu\mathbb{Z} modules on an open j:UCj:U^{\prime}\subset C. Suppose that

  1. (1)

    for each prime ν\ell\mid\nu, wν\ell^{w}\mid\nu, and twt\leq w, the multiplication by t\ell^{t} map j[w]j[wt]j_{*}\mathscr{F}^{\prime}[\ell^{w}]\to j_{*}\mathscr{F}^{\prime}[\ell^{w-t}] is surjective.

  2. (2)

    j(C)=0j_{*}\mathscr{F}^{\prime}(C)=0.

Then H1(C,j[ν])H^{1}(C,j_{*}\mathscr{F}^{\prime}[\nu]) is a free /ν\mathbb{Z}/\nu\mathbb{Z} module. In the case jj_{*}\mathscr{F}^{\prime} is of the form of A[ν]A^{\prime}[\nu], for AUA^{\prime}\to U^{\prime} an abelian scheme, hypothesis (1)(1) above is satisfied if the geometric component group ΦA\Phi_{A^{\prime}} has order prime to ν\nu.

Proof.

Using the Chinese remainder theorem, we can reduce to the case that ν=w\nu=\ell^{w} is a prime power. Suppose H1(C,j[])(/)r.H^{1}(C,j_{*}\mathscr{F}^{\prime}[\ell])\simeq(\mathbb{Z}/\ell\mathbb{Z})^{r}. We will show by induction on ww that H1(C,j[w])(/w)r.H^{1}(C,j_{*}\mathscr{F}^{\prime}[\ell^{w}])\simeq(\mathbb{Z}/\ell^{w}\mathbb{Z})^{r}.

For 0tw0\leq t\leq w we claim there is an exact sequence

(5.1) 0{0}j[t]{j_{*}\mathscr{F}^{\prime}[\ell^{t}]}j[w]{j_{*}\mathscr{F}^{\prime}[\ell^{w}]}j[wt]{j_{*}\mathscr{F}^{\prime}[\ell^{w-t}]}0.{0.}

This is left exact because the analogous sequence for \mathscr{F}^{\prime} in place of jj_{*}\mathscr{F}^{\prime} is left exact. This sequence is right exact by assumption (1) from the statement of the lemma.

We now prove the final clause of the statement of the lemma: In the case A[ν]\mathscr{F}^{\prime}\simeq A^{\prime}[\nu], the cokernel of the map j[w]j[wt]j_{*}\mathscr{F}^{\prime}[\ell^{w}]\to j_{*}\mathscr{F}^{\prime}[\ell^{w-t}] is identified with ΦA/tΦA\Phi_{A^{\prime}}/\ell^{t}\Phi_{A^{\prime}}. This is trivial by assumption as tν\ell^{t}\mid\nu. Therefore, in this case, (1)(1) holds.

We next claim H0(C,j[t])=H2(C,j[t])=0H^{0}(C,j_{*}\mathscr{F}^{\prime}[\ell^{t}])=H^{2}(C,j_{*}\mathscr{F}^{\prime}[\ell^{t}])=0. The former holds by assumption (2)(2). By [Mil80, V Proposition 2.2(b)] and the polarization ([t])(1)[t](\mathscr{F}^{\prime}[\ell^{t}])^{\vee}(1)\simeq\mathscr{F}^{\prime}[\ell^{t}], we find

H2(C,j[t])H0(C,j(([t])(1)))H0(C,j[t])H0(U,[t])=0.\displaystyle H^{2}(C,j_{*}\mathscr{F}^{\prime}[\ell^{t}])\simeq H^{0}\left(C,j_{*}\left(\left(\mathscr{F}^{\prime}[\ell^{t}]\right)^{\vee}(1)\right)\right)^{\vee}\simeq H^{0}(C,j_{*}\mathscr{F}^{\prime}[\ell^{t}])^{\vee}\simeq H^{0}(U,\mathscr{F}^{\prime}[\ell^{t}])^{\vee}=0.

The long exact sequence associated to (5.1) and the vanishing of the 0th and 22nd cohomology above implies we obtain an exact sequence

(5.2) 0{0}H1(C,j[t]){H^{1}(C,j_{*}\mathscr{F}^{\prime}[\ell^{t}])}H1(C,j[w]){H^{1}(C,j_{*}\mathscr{F}^{\prime}[\ell^{w}])}H1(C,j[wt]){H^{1}(C,j_{*}\mathscr{F}^{\prime}[\ell^{w-t}])}0.{0.}αt\scriptstyle{\alpha^{t}}βt\scriptstyle{\beta^{t}}

Induction on ww implies #H1(C,j[w])=wr\#H^{1}(C,j_{*}\mathscr{F}^{\prime}[\ell^{w}])=\ell^{wr} and we wish to show H1(C,j[w])H^{1}(C,j_{*}\mathscr{F}^{\prime}[\ell^{w}]) is free of rank rr. By the structure theorem for finite abelian groups, it suffices to show the kernel of multiplication by w1\ell^{w-1} on H1(C,j[w])H^{1}(C,j_{*}\mathscr{F}^{\prime}[\ell^{w}]) has order (w1)r\ell^{(w-1)r}. The multiplication by w1\ell^{w-1} map factors as H1(C,j[w])βw1H1(C,j[])α1H1(C,j[w])H^{1}(C,j_{*}\mathscr{F}^{\prime}[\ell^{w}])\xrightarrow{\beta^{w-1}}H^{1}(C,j_{*}\mathscr{F}^{\prime}[\ell])\xrightarrow{\alpha^{1}}H^{1}(C,j_{*}\mathscr{F}^{\prime}[\ell^{w}]). We know from (5.1) that α1\alpha^{1} is injective so

ker(×w1)=ker(βw1α1)=kerβw1=H1(C,j[w1]),\displaystyle\ker(\times\ell^{w-1})=\ker(\beta^{w-1}\circ\alpha^{1})=\ker\beta^{w-1}=H^{1}(C,j_{*}\mathscr{F}^{\prime}[\ell^{w-1}]),

which has size (w1)r\ell^{(w-1)r}, as we wished to show. ∎

We next aim to compute a formula for the rank of the Selmer sheaf, in favorable situations, in 5.2.6. First, we introduce notation needed to state that formula.

Definition 5.2.4.

Suppose ν\nu is a prime number. Given a locally constant constructible sheaf \mathscr{F} of free /ν\mathbb{Z}/\nu\mathbb{Z} modules on an open UCU^{\prime}\subset C of a curve CC, for any point xCUx\in C-U^{\prime}, there is an associated action of the inertia group IxI_{x} at xx on the geometric generic fiber of η¯\mathscr{F}_{\overline{\eta}}, which is well defined up to conjugacy. We use Dropx()\mathrm{Drop}_{x}(\mathscr{F}) to denote the corank of the invariants of IxI_{x}, i.e., Dropx():=rkrkxIx\mathrm{Drop}_{x}(\mathscr{F}):=\operatorname{rk}\mathscr{F}-\operatorname{rk}\mathscr{F}_{x}^{I_{x}}. In general, if ν\nu is not necessarily a prime number, for each prime ν\ell\mid\nu we use Dropx,():=Dropx([])\mathrm{Drop}_{x,\ell}(\mathscr{F}):=\mathrm{Drop}_{x}(\mathscr{F}[\ell]), and if Dropx,()\mathrm{Drop}_{x,\ell}(\mathscr{F}) is independent of \ell, we denote this common value simply by Dropx()\mathrm{Drop}_{x}(\mathscr{F}). Whenever we use the notation Dropx()\mathrm{Drop}_{x}(\mathscr{F}) in the case ν\nu has multiple prime divisors, we are implicitly claiming it is independent of the prime divisor.

Example 5.2.5.

If ν\nu is prime, and A[ν]\mathscr{F}\simeq A[\nu], then for any xCUx\in C-U, Dropx()=0\mathrm{Drop}_{x}(\mathscr{F})=0 if and only if inertia acts trivially at xx, i.e., A[ν]A[\nu] extends over the point xx. If AA is a relative elliptic curve and the order of the geometric component group of the Néron model of AA at xx is prime to ν\nu, then Dropx()=1\mathrm{Drop}_{x}(\mathscr{F})=1 whenever AA has multiplicative reduction at xx and Dropx()=2\mathrm{Drop}_{x}(\mathscr{F})=2 whenever AA has additive reduction at xx.

Proposition 5.2.6.

Maintain notation as in 5.1.4, so, in particular, \mathscr{F} is a tame symplectically self-dual lcc sheaf of free /ν\mathbb{Z}/\nu\mathbb{Z} modules. Suppose ν\nu is odd and n>0n>0. Assume that B=b¯B={\overline{b}} is the spectrum of an algebraically closed field. Assume that

  1. (1)

    for each prime ν\ell\mid\nu, each integer ww with wν\ell^{w}\mid\nu, and each integer twt\leq w, the multiplication by t\ell^{t} map j[w]j[wt]j_{*}\mathscr{F}[\ell^{w}]\to j_{*}\mathscr{F}[\ell^{w-t}] is surjective

  2. (3)

    the sheaf []\mathscr{F}[\ell] is irreducible for each prime ν\ell\mid\nu.

Assume 2ν2\nu is invertible on BB. For each xQTwistU/Bnx\in\operatorname{QTwist}^{n}_{U/B}, consider the following three properties.

  1. (1’)

    for each prime ν\ell\mid\nu with wν\ell^{w}\mid\nu and twt\leq w, the multiplication by t\ell^{t} map jx[w]jx[wt]j_{*}\mathscr{F}_{x}[\ell^{w}]\to j_{*}\mathscr{F}_{x}[\ell^{w-t}] is surjective

  2. (2’)

    jx(Cx)=0j_{*}\mathscr{F}_{x}(C_{x})=0

  3. (3’)

    the sheaf x[]\mathscr{F}_{x}[\ell] is irreducible for each prime ν\ell\mid\nu.

Then, (2)(2^{\prime}) always holds, (1)(1^{\prime}) holds if (1)(1) holds, and (3)(3^{\prime}) holds if (3)(3) holds.

Moreover, assuming (1)(1) and (3)(3), the map π:SelBnQTwistU/Bn\pi:\operatorname{Sel}_{\mathscr{F}^{n}_{B}}\to\operatorname{QTwist}^{n}_{U/B} is finite étale, representing a locally constant constructible sheaf of rank (2g2+n)2r+xZ(B)Dropx()(2g-2+n)\cdot 2r+\sum_{x\in Z(B)}\mathrm{Drop}_{x}(\mathscr{F}) free /ν\mathbb{Z}/\nu\mathbb{Z} modules, whose formation commutes with base change.

Proof.

First, observe that by 5.2.1, π:SelBnQTwistU/Bn\pi:\operatorname{Sel}_{\mathscr{F}^{n}_{B}}\to\operatorname{QTwist}^{n}_{U/B} is finite étale, corresponding to a locally constant sheaf of /ν\mathbb{Z}/\nu\mathbb{Z} modules, and its formation commutes with base change on QTwistU/Bn\operatorname{QTwist}^{n}_{U/B}.

We now verify that condition (1)(1^{\prime}) hold for quadratic twists x\mathscr{F}_{x} of \mathscr{F}, ramified over a divisor DxD_{x} disjoint from ZxZ_{x}, using condition (1)(1). If \mathscr{F} corresponds to a representation of π1(UxDx)\pi_{1}(U_{x}-D_{x}), the quadratic twist corresponds to tensoring this representation with an order 22 character, whose local inertia at any point outside of DxD_{x} is trivial. Surjectivity of the map from (1)(1^{\prime}) can only fail at points pDxZxp\in D_{x}\cup Z_{x}. If pZxp\in Z_{x}, since surjectivity can be verified locally, surjectivity for jxj_{*}\mathscr{F}_{x} at pp follows from the corresponding surjectivity for jj_{*}\mathscr{F} at pp. If pDxp\in D_{x}, the stalk of jx[wt]j_{*}\mathscr{F}_{x}[\ell^{w-t}] is trivial, as it is identified with the invariants of multiplication by 1-1, which is trivial, and so surjectivity at such points is automatic.

Next, we check (2)(2^{\prime}) holds, just using n>0n>0. We wish to show H0(Cx,x)=0H^{0}(C_{x},\mathscr{F}_{x})=0. Thinking of x\mathscr{F}_{x} as a representation of π1(UxDx)\pi_{1}(U_{x}-D_{x}), a section corresponds to an invariant vector. However, since n>0n>0, local inertia at a point of DxD_{x} acts by 1-1, and so there are no invariant vectors.

Third, we show (3)(3^{\prime}) holds for x\mathscr{F}_{x}, assuming (3)(3) holds for \mathscr{F}. Note that the quadratic twist of the sheaf \mathscr{F} is obtained by tensoring the corresponding representation of π1(U)\pi_{1}(U) with a character. This preserves irreducibility.

We next show this π\pi corresponds to a sheaf of free /ν\mathbb{Z}/\nu\mathbb{Z} modules. We may check this at any point of QTwistU/Bn\operatorname{QTwist}^{n}_{U/B} since the formation of SelBn\operatorname{Sel}_{\mathscr{F}^{n}_{B}} commutes with base change on QTwistU/Bn\operatorname{QTwist}^{n}_{U/B} by 5.2.1. It follows that over a geometric point of QTwistU/Bn\operatorname{QTwist}^{n}_{U/B}, the hypotheses (1)(1) and (2)(2) of 5.2.3, which follow from (1)(1^{\prime}) and (2)(2^{\prime}) in the statement of this proposition, are satisfied for any quadratic twist of \mathscr{F}. Therefore, SelBn\operatorname{Sel}_{\mathscr{F}^{n}_{B}} corresponds to a sheaf of free /ν\mathbb{Z}/\nu\mathbb{Z} modules by 5.2.3.

Finally, we compute the rank of this sheaf. Since we have shown \mathscr{F} is an irreducible /ν\mathbb{Z}/\nu\mathbb{Z} locally constant constructible sheaf on 𝒰Bn\mathscr{U}^{n}_{B}, we can compute the formula for its rank after reduction modulo any prime ν\ell\mid\nu, and hence assume that ν\nu is prime. The formula for the rank is given in [Kat02, Lemma 5.1.3]. Technically, the argument is given there for lisse ¯\overline{\mathbb{Q}}_{\ell} sheaves, but the same computation applies to /\mathbb{Z}/\ell\mathbb{Z} sheaves. In particular, with the above assumptions, if B=SpeckB=\operatorname{Spec}k, for kk an algebraically closed field, 𝒮eBn{\mathcal{S}e\ell}_{\mathscr{F}^{n}_{B}} has rank (2g2+n)2r+xZDropx()(2g-2+n)\cdot 2r+\sum_{x\in Z}\mathrm{Drop}_{x}(\mathscr{F}). ∎

5.3. Connecting points of the Selmer stack and Selmer groups

The next two lemmas connect the Selmer stack to the sizes of Selmer groups and their proofs are quite similar to [Lan21, Proposition 3.23] and [Lan21, Corollary 3.24] respectively.

Lemma 5.3.1.

Retaining notation from 5.1.4 and 5.1.10, suppose n>0n>0, 2ν2\nu is invertible on BB, and let π:SelBnQTwistU/Bn\pi:\operatorname{Sel}_{\mathscr{F}^{n}_{B}}\to\operatorname{QTwist}^{n}_{U/B} denote the structure map. Suppose []\mathscr{F}[\ell] is irreducible for each prime ν\ell\mid\nu. Then for xQTwistU/Bn(𝔽q)x\in\operatorname{QTwist}^{n}_{U/B}(\mathbb{F}_{q}),

H1(Cx,x)(π1(x))(𝔽q).\displaystyle H^{1}(C_{x},\mathscr{F}_{x})\simeq\left(\pi^{-1}(x)\right)(\mathbb{F}_{q}).

Note that the right hand (π1(x))(𝔽q)\left(\pi^{-1}(x)\right)(\mathbb{F}_{q}) acquires the structure of an abelian group as the 𝔽q\mathbb{F}_{q} points of a locally constant constructible sheaf.

Proof.

Using 5.2.1, we know the formation of the Selmer sheaf commutes with base change, and hence for x¯\overline{x} a geometric point over xx, the geometric fiber of SelBn\operatorname{Sel}_{\mathscr{F}^{n}_{B}} over x¯\overline{x} is identified with

R1λ(jx¯)\displaystyle R^{1}\lambda_{*}(j_{*}\mathscr{F}_{\overline{x}}) H1(Cx¯,jx¯).\displaystyle\simeq H^{1}(C_{\overline{x}},j_{*}\mathscr{F}_{\overline{x}}).

To distinguish between étale and group cohomology, we use HgrpiH^{i}_{\operatorname{grp}} denote group cohomology and He´tiH^{i}_{\operatorname{\acute{e}t}} to denote étale cohomology. Let Gx:=Aut(Cx¯/Cx)G_{x}:=\operatorname{Aut}(C_{\overline{x}}/C_{x}). The 𝔽q\mathbb{F}_{q} points of π1(x)\pi^{-1}(x) are the GxG_{x} invariants of He´t1(Cx¯,jx¯)H^{1}_{\operatorname{\acute{e}t}}(C_{\overline{x}},j_{*}\mathscr{F}_{\overline{x}}). That is, π1(x)(𝔽q)=Hgrp0(Gx,He´t1(Cx¯,jx¯))\pi^{-1}(x)(\mathbb{F}_{q})=H_{\operatorname{grp}}^{0}(G_{x},H^{1}_{\operatorname{\acute{e}t}}(C_{\overline{x}},j_{*}\mathscr{F}_{\overline{x}})).

We relate this group to H1(Cx,jx)H^{1}(C_{x},j_{*}\mathscr{F}_{x}) using the Leray spectral sequence

(5.3) 0{0}Hgrp1(Gx,He´t0(Cx¯,jx¯)){H^{1}_{\operatorname{grp}}(G_{x},H^{0}_{\operatorname{\acute{e}t}}(C_{\overline{x}},j_{*}\mathscr{F}_{\overline{x}}))}He´t1(Cx,jx){H^{1}_{\operatorname{\acute{e}t}}(C_{x},j_{*}\mathscr{F}_{x})}Hgrp0(Gx,He´t1(Cx¯,jx¯)){H^{0}_{\operatorname{grp}}(G_{x},H^{1}_{\operatorname{\acute{e}t}}(C_{\overline{x}},j_{*}\mathscr{F}_{\overline{x}}))}Hgrp2(Gx,He´t0(Cx¯,jx¯)).{H^{2}_{\operatorname{grp}}(G_{x},H^{0}_{\operatorname{\acute{e}t}}(C_{\overline{x}},j_{*}\mathscr{F}_{\overline{x}})).}θ\scriptstyle{\theta}

When n>0n>0, we want to show θ\theta is an isomorphism, so it suffices to show He´t0(Cx¯,jx¯)=0H^{0}_{\operatorname{\acute{e}t}}(C_{\overline{x}},j_{*}\mathscr{F}_{\overline{x}})=0. This holds using 5.2.6(3’). ∎

Lemma 5.3.2.

With the same assumptions as in 5.3.1, let xQTwistC/Bn(𝔽q)x\in\operatorname{QTwist}^{n}_{C/B}(\mathbb{F}_{q}), and use Selν(Ax)\operatorname{Sel}_{\nu}(A_{x}) to denote the ν\nu Selmer group of the generic fiber of AxA_{x} over UxU_{x}. We have

Selν(Ax)π1(x)(𝔽q).\displaystyle\operatorname{Sel}_{\nu}(A_{x})\simeq\pi^{-1}(x)(\mathbb{F}_{q}).
Proof.

Using 5.2.1, we know the geometric component group ΦAx¯\Phi_{A_{\overline{x}}} has order prime to ν\nu. As we are also assuming qq is prime to ν\nu, it follows from [Ces16, Proposition 5.4(c)], Selν(Ax)Hfppf1(Cx,𝒜x[ν])\operatorname{Sel}_{\nu}(A_{x})\simeq H^{1}_{\operatorname{fppf}}(C_{x},\mathscr{A}_{x}[\nu]). Upon identifying fppf cohomology with étale cohomology [Gro68, Théorème 11.7 11^{\circ}] and combining this with 5.3.1, we obtain the result. ∎

6. Identifying Selmer elements via Hurwitz stacks

Throughout this section, we’ll work over the complex numbers B=SpecB=\operatorname{Spec}\mathbb{C}. One of the main new ideas in this article is that Selmer elements can actually be parameterized by a Hurwitz stack. The reason for doing this is that the topological methods of the first part of the paper can, as in [EVW16], be used to control the number of 𝔽q\mathbb{F}_{q}-points on certain Hurwitz stacks. Using the identification between Selmer stacks and Hurwitz stacks, we will thus be able to count 𝔽q\mathbb{F}_{q}-points on Selmer stacks. These counts underlie our main theorems.

We produce an isomorphism from the Selmer stack to a certain Hurwitz stack over the complex numbers parameterizing ASp2r(/ν)\operatorname{\mathrm{ASp}}_{2r}(\mathbb{Z}/\nu\mathbb{Z}) covers of our base curve CC over \mathbb{C}. This is shown in 6.4.5. Before jumping into the details, we describe the idea of this isomorphism in § 6.1. Continuing to the proof, we give a monodromy theoretic description of torsion sheaves in § 6.2, and give a monodromy theoretic description of torsors for torsion sheaves in § 6.3. Finally, we identify the Selmer stack with certain Hurwitz stacks in § 6.4.

6.1. Idea of the isomorphism

We will now describe the idea of the proof in the context of torsion in abelian varieties, though below the proof is carried out in the more general context of symplectically self-dual sheaves. The basic idea is that ν\nu Selmer elements for an abelian variety AA^{\prime} over UU^{\prime} of relative dimension rr with Néron model jAj_{*}A^{\prime} over CC correspond to torsors for jA[ν]j_{*}A^{\prime}[\nu]. We can identify jA[ν]j_{*}A^{\prime}[\nu] with a Sp2r(/ν)\operatorname{Sp}_{2r}(\mathbb{Z}/\nu\mathbb{Z}) Galois cover of CC via its Galois representation. We can then identify torsors for jA[ν]j_{*}A^{\prime}[\nu] as ASp2r(/ν)\operatorname{\mathrm{ASp}}_{2r}(\mathbb{Z}/\nu\mathbb{Z}) covers of CC, see 6.3.2. This roughly corresponds to the fact that a torsor for jA[ν]j_{*}A^{\prime}[\nu] can translate the monodromy of jA[ν]j_{*}A^{\prime}[\nu] by an element of a geometric fiber of jA[ν]j_{*}A^{\prime}[\nu], which can be identified with (/ν)2r=ker(ASp2r(/ν)Sp2r(/ν))(\mathbb{Z}/\nu\mathbb{Z})^{2r}=\ker\left(\operatorname{\mathrm{ASp}}_{2r}(\mathbb{Z}/\nu\mathbb{Z})\to\mathrm{Sp}_{2r}(\mathbb{Z}/\nu\mathbb{Z})\right). The bulk of this section amounts to working out the precise conditions on the monodromy of these Hurwitz stacks.

6.2. Symplectically self-dual sheaves in terms of monodromy

Recall that throughout this section, we are working over B=SpecB=\operatorname{Spec}\mathbb{C}. As in 2.4.1, we begin with a smooth projective connected CC curve over Spec\operatorname{Spec}\mathbb{C}, and a nonempty open subscheme UCU\subset C. For DUD\subset U a divisor, we work with a sympletically self-dual sheaf \mathscr{F}^{\prime} over UDU-D of rank 2r2r. A useful example to keep in mind will be when we are in the setting of 5.1.8 and there is an abelian scheme AUDA^{\prime}\to U-D and =A[ν]\mathscr{F}=A^{\prime}[\nu]. The main application will occur when \mathscr{F}^{\prime} is a quadratic twist of a sheaf \mathscr{F}, ramified over DD.

We now describe \mathscr{F}^{\prime} in terms of its monodromy. Fix a basepoint pUDp\in U-D and choose an identification [ν]|p(/ν)2r\mathscr{F}^{\prime}[\nu]|_{p}\simeq\left(\mathbb{Z}/\nu\mathbb{Z}\right)^{2r}. Because the fundamental group π1top(UD,p)\pi_{1}^{\mathrm{top}}(U-D,p) acts linearly on |p\mathscr{F}^{\prime}|_{p}, we obtain a map π1top(UD,p)GL(|p)\pi_{1}^{\mathrm{top}}(U-D,p)\rightarrow\operatorname{GL}(\mathscr{F}^{\prime}|_{p}). Because the sheaf is symplectically self-dual, and we are working over \mathbb{C} where the cyclotomic character acts trivially, this representation factors through Sp(|p)\operatorname{Sp}(\mathscr{F}^{\prime}|_{p}). In other words, we obtain a monodromy representation

(6.1) ρ\displaystyle\rho_{\mathscr{F}^{\prime}} :π1top(UD,p)Sp(|p)Sp2r(/ν).\displaystyle:\pi_{1}^{\mathrm{top}}(U-D,p)\rightarrow\operatorname{Sp}(\mathscr{F}^{\prime}|_{p})\simeq\operatorname{Sp}_{2r}(\mathbb{Z}/\nu\mathbb{Z}).

For convenience of notation, label the points of ZZ by s1,,sf+1s_{1},\ldots,s_{f+1}. As in Figure 6, we can draw oriented loops α1,,αg,β1,,βg,γ1,,γn,δ1,,δf+1\alpha_{1},\ldots,\alpha_{g},\beta_{1},\ldots,\beta_{g},\gamma_{1},\ldots,\gamma_{n},\delta_{1},\ldots,\delta_{f+1} based at pp which pairwise intersect only at pp so that

  1. (1)

    α1,,αg,β1,,βg\alpha_{1},\ldots,\alpha_{g},\beta_{1},\ldots,\beta_{g} forms a basis for H1(C,)H_{1}(C,\mathbb{Z}),

  2. (2)

    γi\gamma_{i} is a loop winding once around pip_{i} corresponding to the local inertia at pip_{i}, where p1,,pnp_{1},\ldots,p_{n} are the nn points in DD, and

  3. (3)

    δi\delta_{i} is a loop winding once around sis_{i} corresponding to the local inertia at sis_{i}.

The above loops form generators of π1top(UD,p)\pi_{1}^{\mathrm{top}}(U-D,p) and satisfy the single relation

(6.2) (α1β1α11β11)(αgβgαg1βg1)γ1γnδ1δf+1=id.\displaystyle(\alpha_{1}\beta_{1}\alpha_{1}^{-1}\beta_{1}^{-1})\cdots(\alpha_{g}\beta_{g}\alpha_{g}^{-1}\beta_{g}^{-1})\gamma_{1}\cdots\gamma_{n}\delta_{1}\cdots\delta_{f+1}=\operatorname{\mathrm{id}}.

Since \mathscr{F}^{\prime} is a /ν\mathbb{Z}/\nu\mathbb{Z} local system on UDU-D, the monodromy representation ρ\rho^{\mathscr{F}^{\prime}} determines \mathscr{F}^{\prime}.

Refer to caption
Figure 6. This picture depicts a genus gg surface XX with f+1f+1 punctures and an nn-point configuration. That is it corresponds to a point in ConfXn\operatorname{Conf}^{n}_{X}. It includes the moving points p1,,pnp_{1},\ldots,p_{n}, surrounded by loops γ1,,γn\gamma_{1},\ldots,\gamma_{n}, the fixed punctures s1,,sf+1s_{1},\ldots,s_{f+1} surrounded by loops δ1,,δf+1\delta_{1},\ldots,\delta_{f+1}, and the standard generators for homology of the compact surface α1,β1,,αg,βg\alpha_{1},\beta_{1},\ldots,\alpha_{g},\beta_{g}.

6.3. Torsors for symplectically self-dual sheaves in terms of monodromy

The next result we are aiming toward is 6.3.7, which gives a description of jj_{*}\mathscr{F}^{\prime} torsors.

We retain notation from § 6.2. For DUD\subset U a divisor, we use j:UDCj:U-D\to C to denote the inclusion. As a first observation, we show that any torsor for jj_{*}\mathscr{F}^{\prime} over CC is determined by its restriction to UDU-D.

Lemma 6.3.1.

The restriction map H1(C,j)H1(UD,)H^{1}(C,j_{*}\mathscr{F}^{\prime})\to H^{1}(U-D,\mathscr{F}^{\prime}) is injective. Its image consists of those torsors [𝒮]H1(UD,)[\mathscr{S}]\in H^{1}(U-D,\mathscr{F}^{\prime}) such that for each qDZq\in D\cup Z, there is some sufficiently small complex analytic open neighborhood CWqC\supset W\ni q such that 𝒮|Wq\mathscr{S}|_{W-q} is the restriction of a j|Wj_{*}\mathscr{F}^{\prime}|_{W} torsor to WqW-q.

Proof.

In the étale topology, the spectral sequence associated to the composition UDCSpecU-D\to C\to\operatorname{Spec}\mathbb{C} yields the injection H1(C,j)H1(UD,)H^{1}(C,j_{*}\mathscr{F}^{\prime})\hookrightarrow H^{1}(U-D,\mathscr{F}^{\prime}). Using the comparison between étale and complex analytic sheaf cohomology [SGA72, Exposé XI, Théoréme 4.4(iii)] we may describe elements of H1(UD,)H^{1}(U-D,\mathscr{F}^{\prime}) as torsors in the complex analytic topology for \mathscr{F}^{\prime}. The condition that a torsor [𝒮]H1(UD,)[\mathscr{S}]\in H^{1}(U-D,\mathscr{F}^{\prime}) lies in the image of H1(C,j)H1(UD,)H^{1}(C,j_{*}\mathscr{F}^{\prime})\to H^{1}(U-D,\mathscr{F}^{\prime}) is precisely the condition that it extends to an jj_{*}\mathscr{F}^{\prime} torsor over a sufficiently small neighborhood of each point qDZq\in D\cup Z. ∎

Recall our goal is to give a monodromy theoretic description of jj_{*}\mathscr{F}^{\prime} torsors. Using 6.3.1, we can describe jj_{*}\mathscr{F}^{\prime} torsors as \mathscr{F}^{\prime} torsors which extend over a small neighborhood of each pip_{i} and sis_{i}. We next describe \mathscr{F}^{\prime} torsors, and then, in 6.3.6, give the condition that such a torsor extends over DZD\cup Z. First, we introduce notation used to describe the monodromy representation parameterizing \mathscr{F}^{\prime} torsors.

Definition 6.3.2.

The affine symplectic group is ASp2r(/ν):=(/ν)2rSp2r(/ν),\operatorname{\mathrm{ASp}}_{2r}(\mathbb{Z}/\nu\mathbb{Z}):=\left(\mathbb{Z}/\nu\mathbb{Z}\right)^{2r}\rtimes\mathrm{Sp}_{2r}(\mathbb{Z}/\nu\mathbb{Z}), where the action of Sp2(/ν)\mathrm{Sp}_{2}(\mathbb{Z}/\nu\mathbb{Z}) on (/ν)2r\left(\mathbb{Z}/\nu\mathbb{Z}\right)^{2r} is via the standard action of matrices on their underlying free rank /ν\mathbb{Z}/\nu\mathbb{Z} module of rank 2r2r.

Remark 6.3.3.

By definition, ASp2r(/ν)\operatorname{\mathrm{ASp}}_{2r}(\mathbb{Z}/\nu\mathbb{Z}) sits in an exact sequence

(6.3) 0{0}(/ν)2r{\left(\mathbb{Z}/\nu\mathbb{Z}\right)^{2r}}ASp2r(/ν){\operatorname{\mathrm{ASp}}_{2r}(\mathbb{Z}/\nu\mathbb{Z})}Sp2r(/ν){\mathrm{Sp}_{2r}(\mathbb{Z}/\nu\mathbb{Z})}0{0}ι\scriptstyle{\iota}Π\scriptstyle{\Pi}

with inclusion map ι\iota and quotient map Π\Pi. With this presentation, ASp2r(/ν)\operatorname{\mathrm{ASp}}_{2r}(\mathbb{Z}/\nu\mathbb{Z}) can be explicitly described as those matrices of the form

(6.4) ASp2r(/ν){(Mv01)GL2r+1(/ν):MSp2r(/ν),v(/ν)2r}.\displaystyle\operatorname{\mathrm{ASp}}_{2r}(\mathbb{Z}/\nu\mathbb{Z})\simeq\left\{\begin{pmatrix}M&v\\ 0&1\end{pmatrix}\in\operatorname{GL}_{2r+1}(\mathbb{Z}/\nu\mathbb{Z}):M\in\mathrm{Sp}_{2r}(\mathbb{Z}/\nu\mathbb{Z}),v\in\left(\mathbb{Z}/\nu\mathbb{Z}\right)^{2r}\right\}.
Notation 6.3.4.

Suppose HH is a finite /ν\mathbb{Z}/\nu\mathbb{Z} module of the form Hi=1m/νiH\simeq\prod_{i=1}^{m}\mathbb{Z}/\nu_{i}\mathbb{Z}. Define AHSp2r(/ν)(i=1m(/νi)2r)Sp2r(/ν)\operatorname{\mathrm{A}^{\operatorname{H}}\mathrm{Sp}}_{2r}(\mathbb{Z}/\nu\mathbb{Z})\simeq\left(\prod_{i=1}^{m}\left(\mathbb{Z}/\nu_{i}\mathbb{Z}\right)^{2r}\right)\rtimes\mathrm{Sp}_{2r}(\mathbb{Z}/\nu\mathbb{Z}), with Sp2r(/ν)\mathrm{Sp}_{2r}(\mathbb{Z}/\nu\mathbb{Z}) acting via reduction modulo νi\nu_{i} and the standard representation on each factor (/νi)2r\left(\mathbb{Z}/\nu_{i}\mathbb{Z}\right)^{2r}. In particular, AHSp2r(/ν)\operatorname{\mathrm{A}^{\operatorname{H}}\mathrm{Sp}}_{2r}(\mathbb{Z}/\nu\mathbb{Z}) sits in a split exact sequence

(6.6) 0{0}i=1m(/νi)2r{\prod_{i=1}^{m}\left(\mathbb{Z}/\nu_{i}\mathbb{Z}\right)^{2r}}AHSp2r(/ν){\operatorname{\mathrm{A}^{\operatorname{H}}\mathrm{Sp}}_{2r}(\mathbb{Z}/\nu\mathbb{Z})}Sp2r(/ν){\mathrm{Sp}_{2r}(\mathbb{Z}/\nu\mathbb{Z})}0.{0.}ι\scriptstyle{\iota}Π\scriptstyle{\Pi}

We next describe the condition for a torsor for \mathscr{F}^{\prime} to extend over a puncture, in terms of monodromy. By § 6.2, \mathscr{F}^{\prime} can be described in terms of ρ\rho_{\mathscr{F}^{\prime}}, which has target Sp2r(/ν)\operatorname{Sp}_{2r}(\mathbb{Z}/\nu\mathbb{Z}). A torsor 𝒮\mathscr{S} for \mathscr{F}^{\prime} can be described in terms of \mathscr{F}^{\prime} together with the additional data of transition functions lying in (/ν)2r\left(\mathbb{Z}/\nu\mathbb{Z}\right)^{2r}. In total, 𝒮\mathscr{S} can be described in terms of a monodromy representation

ρ𝒮:π1top(UD,p)ASp2r(/ν).\displaystyle\rho_{\mathscr{S}}:\pi_{1}^{\mathrm{top}}(U-D,p)\to\operatorname{\mathrm{ASp}}_{2r}(\mathbb{Z}/\nu\mathbb{Z}).

A composition of loops in π1top(UD,p)\pi_{1}^{\mathrm{top}}(U-D,p) maps under ρ𝒮\rho_{\mathscr{S}} to the product of their corresponding matrices, viewed as elements of GL2r+1(/ν)\operatorname{GL}_{2r+1}(\mathbb{Z}/\nu\mathbb{Z}) via (6.4).

Remark 6.3.5.

By construction, for Π\Pi as defined in (6.3), Πρ𝒮=ρ\Pi\circ\rho_{\mathscr{S}}=\rho_{\mathscr{F}^{\prime}}.

We now describe the condition that a \mathscr{F}^{\prime} torsor extends to a jj_{*}\mathscr{F}^{\prime} torsor. We note, first of all, that by 6.3.1, we know that this condition only depends on the restriction of ρ𝒮\rho_{\mathscr{S}} to local inertia groups. Since these inertia groups are procyclic, this amounts to specifying some subset of ASp2r(/ν)\operatorname{\mathrm{ASp}}_{2r}(\mathbb{Z}/\nu\mathbb{Z}), necessarily closed under conjugacy, in which the local monodromy groups are constrained to lie. In the following proposition, we work out what these constraints look like in explicit matrix form.

Lemma 6.3.6.

With notation as in § 6.2, let j:UDCj:U-D\to C denote the inclusion. Suppose qZDq\in Z\cup D with η\eta a small loop around qq whose image under ρ\rho_{\mathscr{F}^{\prime}} corresponds to the local inertia at qq. Let d:=Dropq()d:=\mathrm{Drop}_{q}(\mathscr{F}^{\prime}) so that, after choosing a suitable basis p(/ν)2r\mathscr{F}^{\prime}_{p}\simeq(\mathbb{Z}/\nu\mathbb{Z})^{2r}, we may write ρ(η)\rho_{\mathscr{F}^{\prime}}(\eta) in the form

(M1M20id2rd.)\displaystyle\begin{pmatrix}M_{1}&M_{2}\\ 0&\operatorname{\mathrm{id}}_{2r-d}.\end{pmatrix}

Under the identification of ASp2r(/ν)\operatorname{\mathrm{ASp}}_{2r}(\mathbb{Z}/\nu\mathbb{Z}) as in (6.4), we can extend a \mathscr{F}^{\prime} torsor 𝒮\mathscr{S} to an jj_{*}\mathscr{F}^{\prime} torsor in some complex analytic neighborhood WW of qq if and only if

(6.7) ρ𝒮(η)=(M1M20id2rd0001)\displaystyle\rho_{\mathscr{S}}(\eta)=\begin{pmatrix}M_{1}&M_{2}&*\\ 0&\operatorname{\mathrm{id}}_{2r-d}&0\\ 0&0&1\end{pmatrix}

for some vector (/ν)d*\in(\mathbb{Z}/\nu\mathbb{Z})^{d}. Stated more intrinsically, we can extend 𝒮\mathscr{S} to a \mathscr{F}^{\prime} torsor if and only if the vector vv in (6.4) lies in im(1ρ(η))\operatorname{im}(1-\rho_{\mathscr{F}^{\prime}}(\eta))

Proof.

First, 6.3.5 shows all entries of the matrix in (6.7) are necessary and sufficient for 𝒮\mathscr{S} to extend to a jj_{*}\mathscr{F}^{\prime} torsor except the first 2r2r entries of the last column, accounting for the * and the 0.

Choose a simply connected neighborhood WW of qq and fix a basepoint pWp\in W. To conclude the proof, we will show the claimed entries in the last column of (6.7) from rows d+1d+1 to 2r2r are 0 if and only if 𝒮|Wq\mathscr{S}|_{W-q} extends to a jj_{*}\mathscr{F}^{\prime} torsor over WW. We start by assuming the torsor extends, and aim to show the entries mentioned above are 0. Note that we can identify (/ν)2rd|Wj|W(\mathbb{Z}/\nu\mathbb{Z})^{2r-d}|_{W}\subset j_{*}\mathscr{F}^{\prime}|_{W} as a /ν\mathbb{Z}/\nu\mathbb{Z} subsheaf which restricts to Span(ed+1,,e2r)(/ν)2r|q\operatorname{Span}(e_{d+1},\ldots,e_{2r})\subset\left(\mathbb{Z}/\nu\mathbb{Z}\right)^{2r}\simeq\mathscr{F}^{\prime}|_{q} as the inertia invariants. Therefore, any j|Wj_{*}\mathscr{F}^{\prime}|_{W} torsor 𝒯\mathscr{T} has a distinguished (/ν)d(\mathbb{Z}/\nu\mathbb{Z})^{d} subtorsor, which is given as ker(1ρ(η))\ker(1-\rho_{\mathscr{F}^{\prime}}(\eta)). Since WW is simply connected, this (/ν)2rd(\mathbb{Z}/\nu\mathbb{Z})^{2r-d} torsor is trivial, which implies that the local inertia at qq acts trivially on Span(ed+1,,e2r)(/ν)2rj|q\operatorname{Span}(e_{d+1},\ldots,e_{2r})\subset\left(\mathbb{Z}/\nu\mathbb{Z}\right)^{2r}\simeq j_{*}\mathscr{F}^{\prime}|_{q}, and hence there is a 0 in (6.7) as claimed.

Conversely, suppose there is a 0 in the second row of the third column of (6.7). We will conclude by showing the torsor extends over WW. We obtain a section of 𝒮\mathscr{S} over WqW-q corresponding to each element of (/ν)2dr(\mathbb{Z}/\nu\mathbb{Z})^{2d-r}, and hence a subsheaf (/ν)2rd|Wq𝒮|Wq(\mathbb{Z}/\nu\mathbb{Z})^{2r-d}|_{W-q}\subset\mathscr{S}|_{W-q}. By gluing (/ν)2rd|W(\mathbb{Z}/\nu\mathbb{Z})^{2r-d}|_{W} to 𝒮|Wq\mathscr{S}|_{W-q} along (/ν)2rd|Wq(\mathbb{Z}/\nu\mathbb{Z})^{2r-d}|_{W-q}, we obtain an jj_{*}\mathscr{F}^{\prime} torsor 𝒯\mathscr{T}, which is the desired extension of 𝒮\mathscr{S}. ∎

We can now describe jj_{*}\mathscr{F}^{\prime} torsors in terms of monodromy data.

Lemma 6.3.7.

With notation as in § 6.2, let \mathscr{F} be an irreducible symplectically self-dual sheaf on UU. Suppose n>0n>0. Fix some quadratic twist \mathscr{F}^{\prime} of \mathscr{F}, ramified along a degree nn divisor DD, in the sense that \mathscr{F}^{\prime} is some fiber of Bn\mathscr{F}^{n}_{B}, so that we obtain a corresponding monodromy representation ρ\rho_{\mathscr{F}^{\prime}}. Suppose ρ\rho_{\mathscr{F}^{\prime}} satisfies the hypotheses (1)(1) and (3)(3) of 5.2.6. There are precisely ν(2g2+n)2r+xZDropx()\nu^{(2g-2+n)\cdot 2r+\sum_{x\in Z}\mathrm{Drop}_{x}(\mathscr{F})} isomorphism classes of torsors for jj_{*}\mathscr{F}^{\prime}, which can be described in terms of monodromy data by specifying a representation ρ𝒮:π1(UD,p)ASp2r(/ν)\rho_{\mathscr{S}}:\pi_{1}(U-D,p)\to\operatorname{\mathrm{ASp}}_{2r}(\mathbb{Z}/\nu\mathbb{Z}) up to ASp2r(/ν)\operatorname{\mathrm{ASp}}_{2r}(\mathbb{Z}/\nu\mathbb{Z}) conjugacy, satisfying the following conditions:

  1. (1)

    The image of γi\gamma_{i} under ρ𝒮\rho_{\mathscr{S}} is of the form (6.4) with M=idM=-\operatorname{\mathrm{id}}.

  2. (2)

    If Dropsi()=di\mathrm{Drop}_{s_{i}}(\mathscr{F}^{\prime})=d_{i}, the image of δi\delta_{i} under ρ𝒮\rho_{\mathscr{S}} is conjugate to a matrix of the form (6.7), where we take (q,d)(q,d) there to be (si,di)(s_{i},d_{i}) here.

  3. (3)

    We have Πρ𝒮=ρ\Pi\circ\rho_{\mathscr{S}}=\rho_{\mathscr{F}^{\prime}}.

Let j:UDCj:U-D\to C denote the inclusion. As mentioned above, we consider two torsors 𝒯\mathscr{T} and 𝒯\mathscr{T}^{\prime} equivalent if there is some v(/ν)2rv\in\left(\mathbb{Z}/\nu\mathbb{Z}\right)^{2r} so that ρj𝒯()=ι(v)(ρj𝒯())ι(v)1\rho_{j^{*}\mathscr{T}}(\nabla)=\iota(v)\left(\rho_{j^{*}\mathscr{T}^{\prime}}(\nabla)\right)\iota(v)^{-1} for every {α1,,αg,β1,,βg,γ1,,γn,δ1,,δf+1}\nabla\in\{\alpha_{1},\ldots,\alpha_{g},\beta_{1},\ldots,\beta_{g},\gamma_{1},\ldots,\gamma_{n},\delta_{1},\ldots,\delta_{f+1}\}, with ι\iota as in (6.3).

Proof.

Using 6.3.1, we can describe torsors for jj_{*}\mathscr{F}^{\prime} as torsors for \mathscr{F}^{\prime} which extend over a neighborhood of each piDp_{i}\in D. By 6.3.5, condition (3)(3) precisely corresponds to the condition that the associated Sp2r(/ν)\operatorname{Sp}_{2r}(\mathbb{Z}/\nu\mathbb{Z}) local system associated to 𝒮\mathscr{S} on UDU-D is that associated to \mathscr{F}^{\prime}, and hence 𝒮|U\mathscr{S}|_{U} is a \mathscr{F}^{\prime} torsor. By 6.3.6, an \mathscr{F}^{\prime} torsor extend to a jj_{*}\mathscr{F}^{\prime} torsor over p1,,pnp_{1},\ldots,p_{n}, if and only condition (1)(1) holds, and extends over s1,,sf+1s_{1},\ldots,s_{f+1} if and only if condition (2)(2) holds. We consider the representations up to conjugacy, as this corresponds to a change of basepoint of |p(/ν)2r\mathscr{F}^{\prime}|_{p}\simeq\left(\mathbb{Z}/\nu\mathbb{Z}\right)^{2r}, and expresses the usual condition for two torsors to be equivalent.

To conclude, we wish to see that there are ν(2g2+n)2r+xZDropx()\nu^{(2g-2+n)\cdot 2r+\sum_{x\in Z}\mathrm{Drop}_{x}(\mathscr{F})} isomorphism classes of torsors specified by the above data. Indeed, we see there are ν2r\nu^{2r} possible values ρ𝒮\rho_{\mathscr{S}} can take on the loops α1,,αg,β1,,βg\alpha_{1},\ldots,\alpha_{g},\beta_{1},\ldots,\beta_{g} in order to satisfy (3)(3). For each γi\gamma_{i}, there are νDroppi()=ν2r\nu^{\mathrm{Drop}_{p_{i}}(\mathscr{F}^{\prime})}=\nu^{2r} possible values of ρ𝒮\rho_{\mathscr{S}}, because Π(ρ𝒮(γi))=id2r\Pi(\rho_{\mathscr{S}}(\gamma_{i}))=-\operatorname{\mathrm{id}}_{2r}. For each δi\delta_{i}, there are νDropsi()=νDropsi()\nu^{\mathrm{Drop}_{s_{i}}(\mathscr{F}^{\prime})}=\nu^{\mathrm{Drop}_{s_{i}}(\mathscr{F})} possible values of ρ𝒮\rho_{\mathscr{S}}. We additionally must impose the condition that i=1g[αi,βi]i=1nγii=1f+1δi=id\prod_{i=1}^{g}[\alpha_{i},\beta_{i}]\prod_{i=1}^{n}\gamma_{i}\prod_{i=1}^{f+1}\delta_{i}=\operatorname{\mathrm{id}}, from the relation (6.2) defining the fundamental group, and that we consider these torsors up to conjugacy. Before imposing these two conditions, there are ν(2g+n)2r+xZDropx()\nu^{(2g+n)\cdot 2r+\sum_{x\in Z}\mathrm{Drop}_{x}(\mathscr{F})} possible tuples of matrices. The first condition imposes ν2r\nu^{2r} independent constraints on the matrices. Further, the conjugation action always identifies ν2r\nu^{2r} elements since the representation is center free, using that it is irreducible and that ASp2r(/ν)GL2r+1(/ν)\operatorname{\mathrm{ASp}}_{2r}(\mathbb{Z}/\nu\mathbb{Z})\subset\operatorname{GL}_{2r+1}(\mathbb{Z}/\nu\mathbb{Z}) contains no scalars, other than id\operatorname{\mathrm{id}}. Altogether, this yields ν(2g2+n)2r+xZDropx()\nu^{(2g-2+n)\cdot 2r+\sum_{x\in Z}\mathrm{Drop}_{x}(\mathscr{F})} such torsors. ∎

6.4. Identifying Selmer stacks with Hurwitz stacks

We will use the above description of torsors to identify the Selmer stack with a certain Hurwitz stack in 6.4.5. We next define that Hurwitz stack.

Notation 6.4.1.

Let B=SpecB=\operatorname{Spec}\mathbb{C}. Given a symplectically self-dual sheaf \mathscr{F} over UU as in 5.1.4, and fixing values of ν\nu and nn, we now use the notation HurBnH\operatorname{Hur}^{H}_{\mathscr{F}^{n}_{B}} to indicate the stack HurC/BG,n,Z,𝒮\operatorname{Hur}^{G,n,Z,\mathcal{S}}_{C/B} as in 2.4.2, for n,Z,C,Bn,Z,C,B as in 5.1.4 and G,𝒮G,\mathcal{S} as we define next. Let ν1,,νmν\nu_{1},\ldots,\nu_{m}\mid\nu and write Hi=1m/νiH\simeq\prod_{i=1}^{m}\mathbb{Z}/\nu_{i}\mathbb{Z}. Take G:=AHSp2r(/ν)G:=\operatorname{\mathrm{A}^{\operatorname{H}}\mathrm{Sp}}_{2r}(\mathbb{Z}/\nu\mathbb{Z}). Take 𝒮\mathcal{S} to be the orbit under the conjugation action of GG of the following subset of ϕHom(π1(Σg,f+1),G)\phi\in\mathrm{Hom}(\pi_{1}(\Sigma_{g,f+1}),G). Any such ϕ\phi sends a half-twist (moving point ii counterclockwise toward point i+1i+1 and point i+1i+1 counterclockwise toward point ii) to an element gGg\in G so that Π(g)=id\Pi(g)=-\operatorname{\mathrm{id}}, for Π\Pi as defined in (6.6). If α1,,αg,β1,,βgΣg,f+1Σg\alpha_{1},\ldots,\alpha_{g},\beta_{1},\ldots,\beta_{g}\subset\Sigma_{g,f+1}\subset\Sigma_{g} are a fixed set of simple closed curves forming a standard generating set for the first homology of Σg\Sigma_{g}, we require that Π(ϕ(αi)){±ai}\Pi(\phi(\alpha_{i}))\in\{\pm a_{i}\}, Π(ϕ(βj)){±bj}\Pi(\phi(\beta_{j}))\in\{\pm b_{j}\}, where ai:=ρ(αi)a_{i}:=\rho_{\mathscr{F}}(\alpha_{i}) and bj:=ρ(βj)b_{j}:=\rho_{\mathscr{F}}(\beta_{j}). The local inertia around sis_{i}, the iith puncture among the f+1f+1 punctures, maps to (Mi,vi)(M_{i},v_{i}), where MiM_{i} is the given local inertia for \mathscr{F} and viim(Miid)v_{i}\in\operatorname{im}(M_{i}-\operatorname{\mathrm{id}}).

Remark 6.4.2.

The condition in 6.4.1 that the αi\alpha_{i} and βj\beta_{j} map to ±ai\pm a_{i} and ±bj\pm b_{j} under Πϕ\Pi\circ\phi may seem to depend on choices of the αi\alpha_{i} and βj\beta_{j}, but it can be expressed independently of these choices as follows: if ζ:Sp2r(/ν)Sp2r(/ν)/{±1}\zeta:\mathrm{Sp}_{2r}(\mathbb{Z}/\nu\mathbb{Z})\to\mathrm{Sp}_{2r}(\mathbb{Z}/\nu\mathbb{Z})/\{\pm 1\} is the quotient map, ζΠϕ=ζρ\zeta\circ\Pi\circ\phi=\zeta\circ\rho_{\mathscr{F}}.

In order to show the construction in 6.4.1 gives a Hurwitz stack as in 2.4.2, we need to show the set 𝒮\mathcal{S} is invariant under the action of π1(ConfU/Bn)\pi_{1}(\operatorname{Conf}^{n}_{U/B}). We now verify this.

Lemma 6.4.3.

The set 𝒮\mathcal{S} from 6.4.1 is a subset of Hom(π1(Σg,f+1),G)\mathrm{Hom}(\pi_{1}(\Sigma_{g,f+1}),G) which is invariant under the action of π1(ConfU/Bn)\pi_{1}(\operatorname{Conf}^{n}_{U/B}).

Proof.

Throughout this proof, it may help the reader to refer to 8.1.6, which gives an explicit description of the action of π1(ConfU/Bn)\pi_{1}(\operatorname{Conf}^{n}_{U/B}). Recall we use γi\gamma_{i} for the loop giving inertia around pip_{i} for 1in1\leq i\leq n and δi\delta_{i} for the loop giving inertia around sis_{i}, 1if+11\leq i\leq f+1. First, to show the image of γi\gamma_{i} is preserved by the π1(ConfU/Bn)\pi_{1}(\operatorname{Conf}^{n}_{U/B}) action, note that id-\operatorname{\mathrm{id}} preserved by this action. Therefore, the condition that Π(g)=id\Pi(g)=-\operatorname{\mathrm{id}} is preserved by the action as well. Hence, the condition that γi\gamma_{i} has monodromy gg with Π(g)=id\Pi(g)=-\operatorname{\mathrm{id}} is preserved by the action of π1(ConfU/Bn)\pi_{1}(\operatorname{Conf}^{n}_{U/B}). The condition on the αi\alpha_{i} and βj\beta_{j} is invariant as passing one of the nn points across αi\alpha_{i} or βj\beta_{j} has the effect of negating Π(ϕ(αi))\Pi(\phi(\alpha_{i})) or Π(ϕ(αi))\Pi(\phi(\alpha_{i})), since Π(γt)=id\Pi(\gamma_{t})=-\operatorname{\mathrm{id}}. As for the loops δi\delta_{i}, since the loops γi\gamma_{i} have inertia gg with Π(g)=id\Pi(g)=-\operatorname{\mathrm{id}}, which lies in the center of GG, the matrices MiM_{i} defined in 6.4.1 are preserved by conjugation under id-\operatorname{\mathrm{id}}. Therefore, the 11-eigenspace ker(1Mi)\ker(1-M_{i}) is preserved by conjugation under id-\operatorname{\mathrm{id}}, and so the same holds for im(1Mi)\operatorname{im}(1-M_{i}). Thus, the set of such homomorphisms to GG is indeed preserved by the action of π1(ConfU/Bn)\pi_{1}(\operatorname{Conf}^{n}_{U/B}). ∎

Hypotheses 6.4.4.

Suppose n>0n>0, B=SpecB=\operatorname{Spec}\mathbb{C}, and \mathscr{F} is an irreducible symplectically self-dual lcc sheaf of free /ν\mathbb{Z}/\nu\mathbb{Z} modules which satisfies the hypotheses (1)(1) and (3)(3) of 5.2.6. There is a map

θ:SelBnHurBn/ν\displaystyle\theta:\operatorname{Sel}_{\mathscr{F}^{n}_{B}}\to\operatorname{Hur}^{\mathbb{Z}/\nu\mathbb{Z}}_{\mathscr{F}^{n}_{B}}

obtained via the bijection of 6.4.5, which sends a torsor to the corresponding ASp2r(/ν)\operatorname{\mathrm{ASp}}_{2r}(\mathbb{Z}/\nu\mathbb{Z}) cover for some quadratic twist \mathscr{F}^{\prime} of \mathscr{F}.

Proposition 6.4.5.

With hypotheses as in 6.4.4, for n>0n>0, the map θ\theta, defined over B=SpecB=\operatorname{Spec}\mathbb{C}, is an isomorphism.

Proof.

Note that the projection HurBn/νQTwistU/Bn\operatorname{Hur}^{\mathbb{Z}/\nu\mathbb{Z}}_{\mathscr{F}^{n}_{B}}\to\operatorname{QTwist}^{n}_{U/B} sends a point of HurBn/ν\operatorname{Hur}^{\mathbb{Z}/\nu\mathbb{Z}}_{\mathscr{F}^{n}_{B}}, thought of as an ASp2r(/ν)\operatorname{\mathrm{ASp}}_{2r}(\mathbb{Z}/\nu\mathbb{Z}) cover, to the corresponding Sp2r(/ν)\mathrm{Sp}_{2r}(\mathbb{Z}/\nu\mathbb{Z}) cover. The projection SelBnQTwistU/Bn\operatorname{Sel}_{\mathscr{F}^{n}_{B}}\to\operatorname{QTwist}^{n}_{U/B} sends a torsor 𝒯\mathscr{T} for some quadratic twist \mathscr{F}^{\prime} to the corresponding \mathscr{F}^{\prime}. Both HurBn/ν\operatorname{Hur}^{\mathbb{Z}/\nu\mathbb{Z}}_{\mathscr{F}^{n}_{B}} and SelBn\operatorname{Sel}_{\mathscr{F}^{n}_{B}} are finite étale covers of QTwistU/Bn\operatorname{QTwist}^{n}_{U/B}, and by 6.3.7, θ\theta defines a bijection between geometric points over points of QTwistU/Bn\operatorname{QTwist}^{n}_{U/B}, corresponding to a chosen degree nn quadratic twist \mathscr{F}^{\prime} of missingF\mathscr{\mathscr{missing}}F. In order to show θ\theta is an isomorphism, it is enough to show the bijection between two finite étale covers of QTwistU/Bn\operatorname{QTwist}^{n}_{U/B} defines a homeomorphism. Indeed, we may verify this claim locally on QTwistU/Bn\operatorname{QTwist}^{n}_{U/B}, in which case is enough to verify it on sufficiently small analytic open covers of QTwistU/Bn\operatorname{QTwist}^{n}_{U/B}. We can choose a small open neighborhood of some geometric point []QTwistU/Bn[\mathscr{F}^{\prime}]\in\operatorname{QTwist}^{n}_{U/B}, corresponding to varying the points pip_{i}, along with the corresponding double cover, in a small, pairwise disjoint open analytic discs of CC. Since the bijection of 6.3.7 is compatible with such variation in the points pip_{i}, we obtain the desired isomorphism. ∎

Warning 6.4.6.

The Selmer stack SelSpec𝔽qn\operatorname{Sel}_{\mathscr{F}^{n}_{\operatorname{Spec}\mathbb{F}_{q}}} over 𝔽q\mathbb{F}_{q} will not in general be isomorphic to the Hurwitz stack of ASp2r(/ν)\operatorname{\mathrm{ASp}}_{2r}(\mathbb{Z}/\nu\mathbb{Z}) covers we are considering. Rather, they will be twists of each other, and the Hurwitz stack only becomes isomorphic over 𝔽¯q\overline{\mathbb{F}}_{q}. The reason for this is that the monodromy representation associated to \mathscr{F} may fail to be contained in Sp2r(/ν)\operatorname{Sp}_{2r}(\mathbb{Z}/\nu\mathbb{Z}), and in general it will only be contained in GSp2r(/ν)\operatorname{GSp}_{2r}(\mathbb{Z}/\nu\mathbb{Z}), the general symplectic group. However, once one ensures all roots of unity lie in the base field, this issue goes away.

Computing the average size of a Selmer group in a quadratic twist family will come down to counting 𝔽q\mathbb{F}_{q}-rational points on a Selmer stack. We will want to compute not only averages, but higher moments. This will require counting points on fiber products of Selmer stacks. But, as the following corollary shows, these stacks are isomorphic, making them amenable to the methods of this paper.

Corollary 6.4.7.

With hypotheses as in 6.4.4, let Hi=1m/νiH\simeq\prod_{i=1}^{m}\mathbb{Z}/\nu_{i}\mathbb{Z}. The map θ\theta, defined over B=SpecB=\operatorname{Spec}\mathbb{C}, induces an isomorphism

θm:SelBn[ν1]×QTwistU/Bn×QTwistU/BnSelBn[νm]HurBnH.\displaystyle\theta^{m}:\operatorname{Sel}_{\mathscr{F}^{n}_{B}[\nu_{1}]}\times_{\operatorname{QTwist}^{n}_{U/B}}\cdots\times_{\operatorname{QTwist}^{n}_{U/B}}\operatorname{Sel}_{\mathscr{F}^{n}_{B}[\nu_{m}]}\to\operatorname{Hur}^{H}_{\mathscr{F}^{n}_{B}}.
Proof.

It follows from the definition of HurBnH\operatorname{Hur}^{H}_{\mathscr{F}^{n}_{B}} as in 6.4.1 that

HurBnHHurBn[ν1]/ν×QTwistU/Bn×QTwistU/BnHurBn[νm]/ν.\displaystyle\operatorname{Hur}^{H}_{\mathscr{F}^{n}_{B}}\simeq\operatorname{Hur}^{\mathbb{Z}/\nu\mathbb{Z}}_{\mathscr{F}^{n}_{B}[\nu_{1}]}\times_{\operatorname{QTwist}^{n}_{U/B}}\cdots\times_{\operatorname{QTwist}^{n}_{U/B}}\operatorname{Hur}^{\mathbb{Z}/\nu\mathbb{Z}}_{\mathscr{F}^{n}_{B}[\nu_{m}]}.

The map θ\theta from 6.4.5 also induces isomorphisms HurBn[νi]/νiSelBn[νi]\operatorname{Hur}^{\mathbb{Z}/\nu_{i}\mathbb{Z}}_{\mathscr{F}^{n}_{B}[\nu_{i}]}\to\operatorname{Sel}_{\mathscr{F}^{n}_{B}[\nu_{i}]}. For νiν\nu_{i}\mid\nu, we also have HurBn[νi]/νiHurBn[νi]/ν\operatorname{Hur}^{\mathbb{Z}/\nu_{i}\mathbb{Z}}_{\mathscr{F}^{n}_{B}[\nu_{i}]}\simeq\operatorname{Hur}^{\mathbb{Z}/\nu\mathbb{Z}}_{\mathscr{F}^{n}_{B}[\nu_{i}]} from the definition. The result follows from 6.4.5 by taking appropriate fiber products of isomorphisms over QTwistU/Bn\operatorname{QTwist}^{n}_{U/B}. ∎

7. Computing the monodromy of Hurwitz stacks

In this section, we compute the image of the monodromy representation related to Selmer stacks. This will be used later to determine their connected components. We first control the monodromy when ν\nu is prime in § 7.1. We then control the monodromy for prime power ν\nu in § 7.2 and for composite ν\nu in § 7.3. The above shows that the monodromy is sufficiently large, but does not determine it exactly. We will, however, precisely describe the image of the Dickson invariant map in § 7.4.

7.1. Computing the monodromy when ν\nu is a prime

We first consider the case ν=\nu=\ell is prime. The main result in this case is Theorem 7.1.1, which is a generalization of [Hal08, Theorem 6.3] from the case that we have an elliptic curve over a genus 0 base to the case of a general symplectically self-dual sheaf over a base curve of genus gg. We begin with a definition of the monodromy representation for general odd ν\nu.

Definition 7.1.1.

With notation as in 5.1.4, suppose BB is integral, ν\nu is odd, and 2ν2\nu is invertible on BB. Choose a basepoint xQTwistU/Bnx\in\operatorname{QTwist}^{n}_{U/B}. Let VBn:=R1λ(jBn)xV_{\mathscr{F}^{n}_{B}}:=R^{1}\lambda_{*}\left(j_{*}\mathscr{F}^{n}_{B}\right)_{x}. The Selmer sheaf is a finite étale cover of QTwistU/Bn\operatorname{QTwist}^{n}_{U/B} by 5.2.1 and so induces a monodromy representation ρBn:π1(QTwistU/Bn)Aut(VBn)\rho_{\mathscr{F}^{n}_{B}}:\pi_{1}(\operatorname{QTwist}^{n}_{U/B})\to\operatorname{Aut}(V_{\mathscr{F}^{n}_{B}}). For any geometric point b¯B{\overline{b}}\to B, we also obtain a geometric monodromy representation ρb¯n:π1(QTwistUb¯/b¯n)Aut(Vb¯n)\rho_{\mathscr{F}_{\overline{b}}^{n}}:\pi_{1}(\operatorname{QTwist}^{n}_{U_{\overline{b}}/\overline{b}})\to\operatorname{Aut}(V_{\mathscr{F}^{n}_{\overline{b}}}).

Warning 7.1.2.

Note that ρBn\rho_{\mathscr{F}^{n}_{B}} is a representation of the fundamental group of configuration space, while we use ρ\rho_{\mathscr{F}^{\prime}} very differently in (6.1) for a representation of the fundamental group of the curve UDU-D itself.

Remark 7.1.3.

Using that gcd(ν,2)=1\gcd(\nu,2)=1, there is a nondegenerate pairing on VBnV_{\mathscr{F}^{n}_{B}} The pairing is obtained as the composition

H1(C,j(Bn)x)×H1(C,j(Bn)x)\displaystyle H^{1}(C,j_{*}(\mathscr{F}^{n}_{B})_{x})\times H^{1}(C,j_{*}(\mathscr{F}^{n}_{B})_{x}) H2(C,2(jBn)x)\displaystyle\to H^{2}(C,\wedge^{2}(j_{*}\mathscr{F}^{n}_{B})_{x})
H2(C,j(2Bn)x)\displaystyle\to H^{2}(C,j_{*}(\wedge^{2}\mathscr{F}^{n}_{B})_{x})
H2(C,jμν)\displaystyle\to H^{2}(C,j_{*}\mu_{\nu})
/ν\displaystyle\to\mathbb{Z}/\nu\mathbb{Z}

using Poincaré duality [Mil80, V Proposition 2.2(b)], which is preserved by this monodromy representation. The pairing above is symmetric because Poincaré duality on curves is antisymmetric and the pairing on j(Bn)xj_{*}(\mathscr{F}^{n}_{B})_{x} is antisymmetric, coming from the assumption that \mathscr{F} is symplectically self-dual. Let QBnQ_{\mathscr{F}^{n}_{B}} denote the associated quadratic form. Then, ρBn\rho_{\mathscr{F}_{B}^{n}} factors through the orthogonal group O(QBn)\operatorname{O}(Q_{\mathscr{F}^{n}_{B}}) associated to the above symmetric bilinear pairing.

We now set some assumptions, which will serve as our hypotheses going forward.

Hypotheses 7.1.4.

Suppose ν\nu is an odd integer and r>0r\in\mathbb{Z}_{>0} so that every prime ν\ell\mid\nu satisfies >2r+1\ell>2r+1. Suppose we have a rank 2r2r, tame, symplectically self-dual lcc sheaf of free /ν\mathbb{Z}/\nu\mathbb{Z} modules, \mathscr{F}, over UCU\subset C, a nonempty proper open in a smooth proper curve CC with geometrically connected fibers over an integral affine base BB. Suppose Z:=CUZ:=C-U is nonempty and finite étale over BB. Assume further 2ν2\nu is invertible on BB. Fix a geometric point b¯B{\overline{b}}\to B. We assume there is some point xCb¯x\in C_{\overline{b}} at which Dropx(b¯[])=1\mathrm{Drop}_{x}(\mathscr{F}_{\overline{b}}[\ell])=1 for every prime ν\ell\mid\nu. Also suppose b¯[]\mathscr{F}_{\overline{b}}[\ell] is irreducible for each ν\ell\mid\nu, and that the map jb¯[w]jb¯[wt]j_{*}\mathscr{F}_{\overline{b}}[\ell^{w}]\to j_{*}\mathscr{F}_{\overline{b}}[\ell^{w-t}] is surjective for each prime ν\ell\mid\nu such that wν\ell^{w}\mid\nu, and wtw\geq t, as in hypotheses (1)(1) and (3)(3) of 5.2.6. Let f+1:=deg(CU)f+1:=\deg(C-U) and let nn be a positive even integer.

Note that if we are in the situation of 5.1.8, hypothesis 5.2.6(1) in the case b¯=A[ν]\mathscr{F}_{\overline{b}}=A[\nu] is satisfied whenever the geometric component group ΦAb¯\Phi_{A_{\overline{b}}} has order prime to ν\nu, by 5.2.3. If we additionally assume Ab¯A_{\overline{b}} has multiplicative reduction at some point of Ub¯U_{\overline{b}}, with toric part of dimension 11, then Dropx(b¯[])=1\mathrm{Drop}_{x}(\mathscr{F}_{\overline{b}}[\ell])=1 for every prime ν\ell\mid\nu.

Theorem 7.1.1 (Generalization of  [Hal08, Theorem 6.3]).

Suppose ν=>2r+1\nu=\ell>2r+1 is prime. Choose a geometric basepoint xQTwistU/Bnx\in\operatorname{QTwist}^{n}_{U/B} over a geometric point b¯B{\overline{b}}\to B. We next recall our assumptions from 7.1.4: we assume 2ν2\nu is invertible on the integral affine base BB and b¯\mathscr{F}_{\overline{b}} is a rank 2r2r irreducible lcc symplectically self-dual sheaf. We assume there is some point xCb¯x\in C_{\overline{b}} at which Dropx(b¯)=1\mathrm{Drop}_{x}(\mathscr{F}_{\overline{b}})=1, and b¯\mathscr{F}_{\overline{b}} satisfies hypotheses 5.2.6(1) and (3).

For nn an even integer satisfying

n>max(2g,2(2r+1)(f+1)yDx(b¯)Dropy()2r(2g2)),\displaystyle n>\max\left(2g,\frac{2(2r+1)(f+1)-\sum_{y\in D_{x}({\overline{b}})}\mathrm{Drop}_{y}(\mathscr{F})}{2r}-(2g-2)\right),

the geometric monodromy representation ρb¯n:π1(QTwistUb¯/b¯n)Aut(Vb¯n)\rho_{\mathscr{F}_{\overline{b}}^{n}}:\pi_{1}(\operatorname{QTwist}^{n}_{U_{\overline{b}}/\overline{b}})\to\operatorname{Aut}(V_{\mathscr{F}^{n}_{\overline{b}}}) has im(ρb¯n)\operatorname{im}(\rho_{\mathscr{F}^{n}_{\overline{b}}}) of index at most 22 in O(Qb¯n)\operatorname{O}(Q_{\mathscr{F}^{n}_{\overline{b}}}), for Qb¯nQ_{\mathscr{F}^{n}_{\overline{b}}} as in 7.1.3. Moreover, im(ρb¯n)SO(Qb¯n)\operatorname{im}(\rho_{\mathscr{F}^{n}_{\overline{b}}})\neq\operatorname{SO}(Q_{\mathscr{F}^{n}_{\overline{b}}}).

Proof Sketch.

A fair portion of this proof is essentially explained in [Hal08, Theorem 6.3], see also [Zyw14, Theorem 3.4] for an explicit version and [Hal08, §6.6] for the generalization to r>1r>1. We now briefly outline the details needed in the generalization. For the purposes of the proof, we may assume that B=b¯B={\overline{b}}. Since n>2gn>2g, by [Kat02, Theorem 2.2.6], there is a map h:Cx1h:C_{x}\to\mathbb{P}^{1} of degree nn which is simply branched, the branch locus of hh is disjoint from h((ZD)x)h((Z\cup D)_{x}), hh separates points of (ZD)x(Z\cup D)_{x}, and precisely one point δDx\delta\in D_{x} maps to 1\infty\in\mathbb{P}^{1}. Let br(h)\operatorname{br}(h) denote the branch locus of hh. Take W1W\subset\mathbb{P}^{1} to be the complement of br(h)h(ZD)\operatorname{br}(h)\cup h(Z\cup D). Note that W\infty\notin W by assumption. Then, one can show as in [Kat02, Theorem 5.4.1] that there is a map ϕ:WQTwistU/b¯n\phi:W\to\operatorname{QTwist}^{n}_{U/\overline{b}} which we now describe.

In order to specify a finite double cover of C×b¯WC\times_{\overline{b}}W, it is equivalent to specify a rank 11 locally constant constructible /\mathbb{Z}/\ell\mathbb{Z} sheaf on an open of C×b¯WC\times_{\overline{b}}W whose monodromy is trivialized by that double cover. Let \mathscr{F}^{\prime} denote the quadratic twist of \mathscr{F} corresponding to our chosen geometric basepoint xQTwistU/Bnx\in\operatorname{QTwist}^{n}_{U/B}. Then, =𝕍\mathscr{F}^{\prime}=\mathscr{F}\otimes\mathbb{V}, for 𝕍\mathbb{V} the rank 11 locally constant constructible sheaf on UDU-D given by t(/)/(/)t_{*}(\mathbb{Z}/\ell\mathbb{Z})/(\mathbb{Z}/\ell\mathbb{Z}), for t:XUt:X\to U the finite étale double cover associated to xx. We will now find a family of locally constant constructible sheaves (corresponding to quadratic twists) over WW whose fiber over 0W0\in W is 𝕍\mathbb{V}. To this end, let χ\chi denote the rank 11 locally constant constructible sheaf on 𝔾m:=𝔸1{0}\mathbb{G}_{m}:=\mathbb{A}^{1}-\{0\} corresponding to the double cover 𝔾m𝔾m\mathbb{G}_{m}\to\mathbb{G}_{m} via multiplication by 22. There is a map α:𝔸1×𝔸1Δ𝔾m\alpha^{\prime}:\mathbb{A}^{1}\times\mathbb{A}^{1}-\Delta\to\mathbb{G}_{m} given by (x,y)xy(x,y)\mapsto x-y. Consider the map (h,id):C×11×1(h,\operatorname{\mathrm{id}}):C\times\mathbb{P}^{1}\to\mathbb{P}^{1}\times\mathbb{P}^{1} and let Y:=(h,id)1(W×WΔ)Y:=(h,\operatorname{\mathrm{id}})^{-1}(W\times W-\Delta). Let α\alpha denote the composition Y(h,id)𝔸1×𝔸1Δα𝔾mY\xrightarrow{(h,\operatorname{\mathrm{id}})}\mathbb{A}^{1}\times\mathbb{A}^{1}-\Delta\xrightarrow{\alpha^{\prime}}\mathbb{G}_{m} and let 𝕎:=αχ\mathbb{W}:=\alpha^{*}\chi. Let π2:Y𝔸1\pi_{2}:Y\to\mathbb{A}^{1} denote the second projection. Take 𝕍:=𝕎|h1(W0)×0𝕍|h1(W0)\mathbb{V}^{\prime}:=\mathbb{W}|_{h^{-1}(W-0)\times 0}\otimes\mathbb{V}|_{h^{-1}(W-0)}, viewed as a sheaf on h1(W0)Ch^{-1}(W-0)\subset C. Then (𝕍𝕎)|h1(W0)(\mathbb{V}^{\prime}\otimes\mathbb{W}^{\vee})|_{h^{-1}(W-0)} recovers 𝕍|h1(W0)\mathbb{V}|_{h^{-1}(W-0)}. Now, the locally constant constructible sheaf π2𝕍𝕎\pi_{2}^{*}\mathbb{V}^{\prime}\otimes\mathbb{W}^{\vee} determines a locally constant constructible sheaf on YY. The above identifies the fiber of this over the point 0 with a restriction of 𝕍\mathbb{V}. Since both 𝕍\mathbb{V}^{\prime} and 𝕎\mathbb{W} correspond to representations with image /2\mathbb{Z}/2\mathbb{Z}, the same is true of π2𝕍𝕎\pi_{2}^{*}\mathbb{V}^{\prime}\otimes\mathbb{W}^{\vee}, and hence this sheaf corresponds to a finite étale double cover of YY. Overall, this gives a double cover of C×WC\times W, ramified along a degree nn divisor. This divisor is étale and disjoint from ZZ over C×WC\times W, and hence yields a map ϕ:WQTwistU/b¯n\phi:W\to\operatorname{QTwist}^{n}_{U/\overline{b}}, by the universal property of QTwistU/b¯n\operatorname{QTwist}^{n}_{U/\overline{b}} as a moduli stack of finite double covers branched over a divisor disjoint from ZZ. The sheaf ϕ𝒮eb¯n\phi^{*}{\mathcal{S}e\ell}_{\mathscr{F}^{n}_{\overline{b}}} may also be viewed as the middle convolution MCχ((h)|W)\operatorname{MC}_{\chi}((h_{*}\mathscr{F}^{\prime})|_{W}). (See [Kat02, Proposition 5.3.7] for an analogous statement in the \ell-adic setting.)

Since ϕ𝒮eb¯n\phi^{*}{\mathcal{S}e\ell}_{\mathscr{F}^{n}_{\overline{b}}} is the middle convolution MCχ((h)|W)\operatorname{MC}_{\chi}((h_{*}\mathscr{F}^{\prime})|_{W}) of the irreducible sheaf (h)|W(h_{*}\mathscr{F}^{\prime})|_{W}, we obtain that ϕ𝒮eb¯n\phi^{*}{\mathcal{S}e\ell}_{\mathscr{F}^{n}_{\overline{b}}} is irreducible. Here we are using that the middle convolution of an irreducible sheaf is irreducible. This holds because middle convolution is invertible, and hence sends irreducible objects to irreducible objects. A proof is given in [Kat96, Theorem 3.3.3(2d)] for ¯\overline{\mathbb{Q}}_{\ell} sheaves, but the same proof works for sheaves of /\mathbb{Z}/\ell\mathbb{Z} modules. (See also [Det08, Corollary 1.6.4] for a proof in the characteristic 0 setting.)

We may moreover compute the monodromy of ϕ𝒮eb¯n\phi^{*}{\mathcal{S}e\ell}_{\mathscr{F}^{n}_{\overline{b}}} at the geometric points of 𝔸1W\mathbb{A}^{1}-W. At branch points of hh, the monodromy is unipotent via the calculation done in [Kat02, Proposition 5.4.1]. At the other geometric points of 𝔸1W\mathbb{A}^{1}-W the calculation is the same as in the proof of [Hal08, Theorem 6.3 and Lemma 6.5]. In particular, at each of the geometric points of h(D)h(D), the monodromy is also unipotent. This is also explained in [Kat02, Proposition 5.4.1, p. 99, last 3 lines], where it is also shown that Dropy(ϕ𝒮eb¯n)2r\mathrm{Drop}_{y}(\phi^{*}{\mathcal{S}e\ell}_{\mathscr{F}^{n}_{\overline{b}}})\leq 2r at all such geometric points y𝔸1Wy\in\mathbb{A}^{1}-W.

We conclude by verifying the three hypotheses of [Hal08, Theorem 3.1], whose conclusion implies the statement of the theorem we are proving. Note that the monodromy of the sheaf ϕ𝒮eb¯n\phi^{*}{\mathcal{S}e\ell}_{\mathscr{F}^{n}_{\overline{b}}} is generated by the inertia around br(h),h(Zx),\operatorname{br}(h),h(Z_{x}), and h(Dxδ)h(D_{x}-\delta).

We need to verify hypotheses (i),(ii),(i),(ii), and (iii)(iii) [Hal08, Theorem 3.1], as well as show the image of monodromy contains a reflection and an isotropic shear, in the language of [Hal08, p. 185]. We claim the local monodromy around a point of h(Zx)Wh(Z_{x})\subset W over which AxA_{x} has toric part of codimension 11 acts as a reflection, while the local monodromy around a point of h(Dx)h(D_{x}) acts as an isotropic shear. These claims are proven in the case of elliptic curves in [Hal08, Lemma 6.5] and the proof for higher dimensional abelian varieties is analogous.

In order to verify (i)(i), take the value labeled rr in [Hal08, Theorem 3.1] to be what we are calling 2r=2(dimAdimUb¯)2r=2(\dim A-\dim U_{\overline{b}}). Maintaining our notation, we have seen above that the images of inertia around the above mentioned geometric points yS:=𝔸1Wy\in S:=\mathbb{A}^{1}-W generate an irreducible representation, and satisfy Dropy(ϕ𝒮eb¯n)2(dimAdimUb¯)\mathrm{Drop}_{y}(\phi^{*}{\mathcal{S}e\ell}_{\mathscr{F}^{n}_{\overline{b}}})\leq 2(\dim A-\dim U_{\overline{b}}). This verifies [Hal08, Theorem 3.1(i)].

Taking S0SS_{0}\subset S to be the subset of the f+1f+1 geometric points over h(Z)h(Z), we find 2(2r+1)(#S0(b¯))dimV2(2r+1)(\#S_{0}({\overline{b}}))\leq\dim V by rearranging the assumption that

n>2(2r+1)(f+1)yZ(b¯)Dropy()2r(2g2),\displaystyle n>\frac{2(2r+1)(f+1)-\sum_{y\in Z({\overline{b}})}\mathrm{Drop}_{y}(\mathscr{F})}{2r}-(2g-2),

using our computation for the dimension of VV from 5.2.6. This verifies [Hal08, Theorem 3.1(ii)].

Finally, every γSS0\gamma\in S-S_{0} has unipotent monodromy, as we showed above. Hence, every γSS0\gamma\in S-S_{0} has order a power of \ell, so has order prime to (2r+1)!(2r+1)! whenever >2r+1\ell>2r+1. This verifies [Hal08, Theorem 3.1(iii)]. Applying [Hal08, Theorem 3.1] gives result. ∎

7.2. Computing the monodromy for prime-power ν\nu

Our next goal is to generalize Theorem 7.1.1 to prime power ν\nu, and then to general composite ν\nu. We next prove 7.2.2, which will imply that if we have big monodromy mod\bmod\ell, we also have big monodromy modj\bmod\ell^{j} for any integer j>0j>0.

Definition 7.2.1.

Suppose QQ is a quadratic form over /k\mathbb{Z}/\ell^{k}\mathbb{Z}. The lie algebra 𝔰𝔬(Q)(𝔽)\mathfrak{so}(Q)(\mathbb{F}_{\ell}) is by definition ker(SO(Q)(/2)SO(Q)(/))\ker(\operatorname{SO}(Q)(\mathbb{Z}/\ell^{2}\mathbb{Z})\to\operatorname{SO}(Q)(\mathbb{Z}/\ell\mathbb{Z})).

We thank Eric Rains for help with the following proof.

Proposition 7.2.2.

Let s3s\geq 3 and 5\ell\geq 5 a prime. Let (V,Q)(V,Q) be a non-degenerate quadratic space of rank ss over /\mathbb{Z}/\ell\mathbb{Z}. Suppose GΩ(Q)(/j)G\subset\Omega(Q)(\mathbb{Z}/\ell^{j}\mathbb{Z}) is a subgroup so that the composition GΩ(Q)(/j)Ω(Q)(/)G\to\Omega(Q)(\mathbb{Z}/\ell^{j}\mathbb{Z})\to\Omega(Q)(\mathbb{Z}/\ell\mathbb{Z}) is surjective. Then, G=Ω(Q)(/j)G=\Omega(Q)(\mathbb{Z}/\ell^{j}\mathbb{Z}).

Proof.

This is a special case of [Vas03, Theorem 1.3(a)]. Since there are a few mistakes in other parts of that theorem statement (though not in the part relevant to the proposition we’re proving) we spell out a few more details here. The argument proceeds as indicated in the second to last paragraph of [Vas03, p. 327]. First, as in [Vas03, Lemma 4.1.2] we can reduce to the case j=2j=2. To deal with the case j=2j=2, it is enough to show GG meets the Lie algebra 𝔰𝔬(Q)(𝔽)\mathfrak{so}(Q)(\mathbb{F}_{\ell}) nontrivially, as argued in [Vas03, 4.4.1]. Finally, in [Vas03, Theorem 4.5] it is shown that GG meets the Lie algebra nontrivially. ∎

7.3. Bootstrapping to general composite ν\nu

We next collect a few lemmas to bootstrap from showing there is big monodromy modulo prime powers, to showing there is big monodromy modulo composite integers. The main result is 7.3.3. The general strategy will be to apply Goursat’s lemma. A key input in Goursat’s lemma is to understand which simple groups appear as subquotients of orthogonal groups. As a first step, using 7.2.2, we can prove Ω(Q)(/ν)\Omega(Q)(\mathbb{Z}/\nu\mathbb{Z}) is perfect.

Lemma 7.3.1.

For s3s\geq 3, ν\nu a positive integer, and (V,Q)(V,Q) a non-degenerate quadratic space of rank ss over /ν\mathbb{Z}/\nu\mathbb{Z}, Ω(Q)(/ν)\Omega(Q)(\mathbb{Z}/\nu\mathbb{Z}) is perfect. That is, Ω(Q)(/ν)\Omega(Q)(\mathbb{Z}/\nu\mathbb{Z}) is its own commutator.

Proof.

Write ν=i=1tiai\nu=\prod_{i=1}^{t}\ell_{i}^{a_{i}} for i\ell_{i} pairwise distinct primes. Note Ω(Q)(/i)\Omega(Q)(\mathbb{Z}/\ell_{i}\mathbb{Z}) is perfect as shown in [Wil09, p. 73, lines 2-7]. Then, since the commutator subgroup

[Ω(Q)(/iai),Ω(Q)(/iai)]Ω(Q)(/iai)\displaystyle\left[\Omega(Q)(\mathbb{Z}/\ell_{i}^{a_{i}}\mathbb{Z}),\Omega(Q)(\mathbb{Z}/\ell_{i}^{a_{i}}\mathbb{Z})\right]\subset\Omega(Q)(\mathbb{Z}/\ell_{i}^{a_{i}}\mathbb{Z})

is a subgroup of Ω(Q)(/iai)\Omega(Q)(\mathbb{Z}/\ell_{i}^{a_{i}}\mathbb{Z}) surjecting onto Ω(Q)(/i)\Omega(Q)(\mathbb{Z}/\ell_{i}\mathbb{Z}), it must be all of Ω(Q)(/iai)\Omega(Q)(\mathbb{Z}/\ell_{i}^{a_{i}}\mathbb{Z}) by 7.2.2. Finally, as commutators commute with products, and Ω(Q)(/ν)=i=1tΩ(Q)(/iai)\Omega(Q)(\mathbb{Z}/\nu\mathbb{Z})=\prod_{i=1}^{t}\Omega(Q)(\mathbb{Z}/\ell_{i}^{a_{i}}\mathbb{Z}), it follows that Ω(Q)(/ν)\Omega(Q)(\mathbb{Z}/\nu\mathbb{Z}) is its own commutator. ∎

The next result relates monodromy for prime power ν\nu to monodromy for general composite ν\nu.

Proposition 7.3.2.

Let s5s\geq 5. Let (V,Q)(V,Q) be a non-degenerate quadratic space of rank ss over /ν\mathbb{Z}/\nu\mathbb{Z}. Suppose GΩ(Q)(/ν)G\subset\Omega(Q)(\mathbb{Z}/\nu\mathbb{Z}) is a subgroup so that for each prime ν\ell\mid\nu, the composition GΩ(Q)(/ν)Ω(Q)(/)G\to\Omega(Q)(\mathbb{Z}/\nu\mathbb{Z})\to\Omega(Q)(\mathbb{Z}/\ell\mathbb{Z}) is surjective. Then, G=Ω(Q)(/ν)G=\Omega(Q)(\mathbb{Z}/\nu\mathbb{Z}).

Proof.

We have already proven this in the case ν\nu is a prime power in 7.2.2. It now remains to deal with general composite ν\nu.

To this end, write ν=i=1tiai\nu=\prod_{i=1}^{t}\ell_{i}^{a_{i}}, for i\ell_{i} pairwise distinct primes. The proposition follows from an application of Goursat’s lemma, as we now explain. We will show that the groups Ω(Q)(/iai)\Omega(Q)(\mathbb{Z}/\ell_{i}^{a_{i}}\mathbb{Z}) for 1it1\leq i\leq t satisfy the following two properties: (1)(1) they have trivial abelianization and (2)(2) they have no finite non-abelian simple quotients in common. These two facts verify the hypotheses of Goursat’s lemma as stated in [Gre10, Proposition 2.5], which implies that G=i=1tΩ(Q)(/iai)=Ω(Q)(/ν)G=\prod_{i=1}^{t}\Omega(Q)\left(\mathbb{Z}/\ell_{i}^{a_{i}}\mathbb{Z}\right)=\Omega(Q)(\mathbb{Z}/\nu\mathbb{Z}).

It remains to verify (1)(1) and (2)(2). Observe that (1)(1) follows from 7.3.1. To conclude our proof, we only need to check (2)(2): that the groups Ω(Q)(/iai)\Omega(Q)(\mathbb{Z}/\ell_{i}^{a_{i}}\mathbb{Z}) for 1it1\leq i\leq t have no finite non-abelian simple quotients in common. For GG^{\prime} a group, let Quo(G)\operatorname{Quo}(G^{\prime}) denote the set of finite simple non-abelian quotients of GG^{\prime}. To prove (2)(2), it suffices to show Quo(Ω(Q)(/iai))={Ω(Q)(/i)}.\operatorname{Quo}(\Omega(Q)(\mathbb{Z}/\ell_{i}^{a_{i}}\mathbb{Z}))=\left\{\mathbb{P}\Omega(Q)(\mathbb{Z}/\ell_{i}\mathbb{Z})\right\}. Note that the latter group is indeed simple by [Wil09, 3.7.3 and 3.8.2], using that s5s\geq 5.

So, we now check Quo(Ω(Q)(/iai))={Ω(Q)(/i)}.\operatorname{Quo}(\Omega(Q)(\mathbb{Z}/\ell_{i}^{a_{i}}\mathbb{Z}))=\left\{\mathbb{P}\Omega(Q)(\mathbb{Z}/\ell_{i}\mathbb{Z})\right\}. Since every finite simple quotient appears as some Jordan Holder factor, it suffices to check the all simple Jordan Holder factors of Ω(Q)(/iai)\Omega(Q)(\mathbb{Z}/\ell_{i}^{a_{i}}\mathbb{Z}) are contained in {Ω(Q)(/i),/i,/2}.\{\mathbb{P}\Omega(Q)(\mathbb{Z}/\ell_{i}\mathbb{Z}),\mathbb{Z}/\ell_{i}\mathbb{Z},\mathbb{Z}/2\mathbb{Z}\}. To see this, consider the surjections Ω(Q)(/iai)Ω(Q)(/iai1)Ω(Q)(/i2)Ω(Q)(/i){id}\Omega(Q)(\mathbb{Z}/\ell_{i}^{a_{i}}\mathbb{Z})\to\Omega(Q)(\mathbb{Z}/\ell_{i}^{a_{i}-1}\mathbb{Z})\to\cdots\to\Omega(Q)(\mathbb{Z}/\ell_{i}^{2}\mathbb{Z})\to\Omega(Q)(\mathbb{Z}/\ell_{i}\mathbb{Z})\to\left\{\operatorname{\mathrm{id}}\right\}. From these surjections, we obtain an associated filtration. The Jordan Holder factors associated to any refinement of this filtration will all lie in {Ω(Q)(/i),/i,/2}\{\mathbb{P}\Omega(Q)(\mathbb{Z}/\ell_{i}\mathbb{Z}),\mathbb{Z}/\ell_{i}\mathbb{Z},\mathbb{Z}/2\mathbb{Z}\} since the kernels of all maps but the last are products of /i\mathbb{Z}/\ell_{i}\mathbb{Z}. ∎

Proposition 7.3.3.

Keep assumptions as in 7.1.4. Suppose b¯B{\overline{b}}\to B is a geometric point. If

(7.1) n>max(2,2g,2(2r+1)(f+1)yDx(b¯)Dropy()2r(2g2)),\displaystyle n>\max\left(2,2g,\frac{2(2r+1)(f+1)-\sum_{y\in D_{x}({\overline{b}})}\mathrm{Drop}_{y}(\mathscr{F})}{2r}-(2g-2)\right),

then the geometric monodromy representation ρb¯n:π1(QTwistUb¯/b¯n)Aut(Vb¯n)\rho_{\mathscr{F}_{\overline{b}}^{n}}:\pi_{1}(\operatorname{QTwist}^{n}_{U_{\overline{b}}/\overline{b}})\to\operatorname{Aut}(V_{\mathscr{F}_{\overline{b}}^{n}}) satisfies Ω(Qb¯n)im(ρb¯n)O(Qb¯n)\Omega(Q_{\mathscr{F}_{\overline{b}}^{n}})\subset\operatorname{im}(\rho_{\mathscr{F}_{\overline{b}}^{n}})\subset\operatorname{O}(Q_{\mathscr{F}_{\overline{b}}^{n}}) and im(ρb¯n)SO(Qb¯n)\operatorname{im}(\rho_{\mathscr{F}^{n}_{\overline{b}}})\not\subset\operatorname{SO}(Q_{\mathscr{F}^{n}_{\overline{b}}}).

Proof.

We have seen in 7.1.3 that im(ρb¯n)O(Qb¯n)\operatorname{im}(\rho_{\mathscr{F}_{\overline{b}}^{n}})\subset\operatorname{O}(Q_{\mathscr{F}_{\overline{b}}^{n}}) holds. By Theorem 7.1.1, we know Ω(Q[]n)im(ρb¯[]n)\Omega(Q_{\mathscr{F}[\ell]^{n}})\subset\operatorname{im}(\rho_{\mathscr{F}_{\overline{b}}[\ell]^{n}}) for each prime ν\ell\mid\nu. It follows from 7.3.2 that Ω(Qb¯n)im(ρb¯n)\Omega(Q_{\mathscr{F}^{n}_{\overline{b}}})\subset\operatorname{im}(\rho_{\mathscr{F}_{\overline{b}}^{n}}). Note that since n>2n>2, the formula for the rank of Vb¯nV_{\mathscr{F}^{n}_{\overline{b}}} from 5.2.6 shows it is at least 55, so the hypotheses of 7.3.2 are satisfied. From Theorem 7.1.1, we also find that im(ρb¯n)SO(Qb¯n)\operatorname{im}(\rho_{\mathscr{F}^{n}_{\overline{b}}})\not\subset\operatorname{SO}(Q_{\mathscr{F}^{n}_{\overline{b}}}). ∎

7.4. Understanding the image of the Dickson invariant map

Having shown that the image of monodromy is close to the orthogonal group, so in particular contains Ω(QBn)\Omega(Q_{\mathscr{F}^{n}_{B}}), its failure to equal the orthogonal group can be understood in terms of the spinor norm and Dickson invariant. The spinor norm will not have much effect on the distribution of Selmer elements, but the Dickson invariant will have a huge effect, and is closely connected to the parity of the rank of AA in the case bA[ν],\mathscr{F}_{b}\simeq A[\nu], for AUA\to U an abelian scheme as in 5.1.8. In the remainder of this section, specifically 7.4.6, we precisely determine the image of the Dickson invariant, under the arithmetic monodromy representation ρbn\rho_{\mathscr{F}^{n}_{b}}.

Our strategy for determining the arithmetic monodromy will be to use equidistribution of Frobenius elements, and compute images of Frobenius elements by relating them to Selmer groups. The following notation for the distribution of Selmer groups will make it convenient to express the types of Selmer groups which appear.

Definition 7.4.1.

Keep assumptions as in as in 5.1.4 and 5.1.8, and assume that BB is a local scheme so that bBb\in B is the unique closed point and has residue field contained in 𝔽q\mathbb{F}_{q}. In particular, bA[ν]\mathscr{F}_{b}\simeq A[\nu] for AUbA\to U_{b} a polarized abelian scheme with polarization degree prime to ν\nu.

Let 𝒩\mathcal{N} denote the set of isomorphism classes of finite /ν\mathbb{Z}/\nu\mathbb{Z} modules. Let XA[ν]𝔽qnX_{A[\nu]^{n}_{\mathbb{F}_{q}}} denote the probability distribution on 𝒩\mathcal{N} defined by

Prob(XA[ν]𝔽qn=H)=#{xQTwistUb/bn(𝔽q):Selν(Ax)H}#QTwistUb/bn(𝔽q).\displaystyle\operatorname{Prob}\left(X_{A[\nu]^{n}_{\mathbb{F}_{q}}}=H\right)=\frac{\#\{x\in\operatorname{QTwist}^{n}_{U_{b}/b}(\mathbb{F}_{q}):\operatorname{Sel}_{\nu}(A_{x})\simeq H\}}{\#\operatorname{QTwist}^{n}_{U_{b}/b}(\mathbb{F}_{q})}.

Here, as usual, point counts of stacks are weighted inversely proportional to the isotropy group at that point. For i{0,1}i\in\{0,1\}, let 𝒩i𝒩\mathcal{N}^{i}\subset\mathcal{N} denote the subset of 𝒩\mathscr{N} of those HH so that there exists some /ν\mathbb{Z}/\nu\mathbb{Z} module GG such that H(/ν)i×G2H\simeq(\mathbb{Z}/\nu\mathbb{Z})^{i}\times G^{2}. Given H𝒩iH\in\mathcal{N}^{i}, define

Prob(XA[ν]𝔽qni=H)=#{xQTwistUb/bn(𝔽q):Selν(Ax)H}#{QTwistUb/bn(𝔽q):Selν(Ax)𝒩i}.\displaystyle\operatorname{Prob}\left(X^{i}_{A[\nu]^{n}_{\mathbb{F}_{q}}}=H\right)=\frac{\#\{x\in\operatorname{QTwist}^{n}_{U_{b}/b}(\mathbb{F}_{q}):\operatorname{Sel}_{\nu}(A_{x})\simeq H\}}{\#\{\operatorname{QTwist}^{n}_{U_{b}/b}(\mathbb{F}_{q}):\operatorname{Sel}_{\nu}(A_{x})\in\mathcal{N}^{i}\}}.

The next two lemmas give the key constraint on Tate-Shafarevich groups and Selmer groups we will use to determine the image of the Dickson invariant. It is one of the few places in this paper that the arithmetic of abelian varieties comes crucially into play.

Lemma 7.4.2.

Let ν\nu be an odd positive integer. Let KK be the function field of a curve over a finite field, and let AA be an abelian variety over KK with a polarization of degree prime to ν\nu. Then, there is a finite /ν\mathbb{Z}/\nu\mathbb{Z} module GG so that either (A)[ν]G2\Sha(A)[\nu]\simeq G^{2} or (A)[ν]G2/ν\Sha(A)[\nu]\simeq G^{2}\oplus\mathbb{Z}/\nu\mathbb{Z}.

Remark 7.4.3.

If we assume the BSD conjecture, (A)\Sha(A) will be finite and then the assumptions that the polarization has degree prime to ν\nu and ν\nu is odd will imply (A)[ν]\Sha(A)[\nu] has square order.

Remark 7.4.4.

The condition that the polarization has degree prime to ν\nu is important here: In general, even when the Tate-Shafarevich group is known to be finite, it can fail to be a square or twice a square, see [CLQR04, p. 278, Theorem 1.4].

Proof.

To approach this, we first review some general facts about the structure of the Tate-Shafarevich group. We can write (A)[](/)rK\Sha(A)[\ell^{\infty}]\simeq(\mathbb{Q}_{\ell}/\mathbb{Z}_{\ell})^{r_{\ell}}\oplus K_{\ell}, where KK_{\ell} is a finite group and rr_{\ell} is the rank of (A)[]\Sha(A)[\ell^{\infty}]. Note that the BSD conjecture would imply r=0r_{\ell}=0, but we will not use this.

We next claim that νKGnd2\oplus_{\ell\mid\nu}K_{\ell}\simeq G_{\operatorname{nd}}^{2}, for some finite /ν\mathbb{Z}/\nu\mathbb{Z} module GndG_{\operatorname{nd}}. Indeed, let (A)[ν]nd\Sha(A)[\nu]_{\operatorname{nd}} denote the non-divisible part of (A)[ν]\Sha(A)[\nu]. Then, (A)[ν]nd\Sha(A)[\nu]_{\operatorname{nd}} has a nondegenerate pairing, by [Tat63, Theorem 3.2], which is antisymmetric by [Fla90, Theorem 1]. Since ν\nu is odd, any finite /ν\mathbb{Z}/\nu\mathbb{Z} module with an nondegenerate antisymmetric pairing is a square, so there is some /ν\mathbb{Z}/\nu\mathbb{Z} module GndG_{\operatorname{nd}} with (A)[ν]ndGnd2\Sha(A)[\nu]_{\operatorname{nd}}\simeq G_{\operatorname{nd}}^{2}.

We now conclude the proof. By [TY14, Corollary 1.0.3], rr_{\ell} has parity independent of \ell. Write ν=νa\nu=\prod_{\ell\mid\nu}\ell^{a_{\ell}}, and take G=Gnd(ν(/a)r2)G=G_{\operatorname{nd}}\oplus\left(\oplus_{\ell\mid\nu}(\mathbb{Z}/\ell^{a_{\ell}}\mathbb{Z})^{\lfloor\frac{r_{\ell}}{2}\rfloor}\right). We get (A)[ν]G2\Sha(A)[\nu]\simeq G^{2} if rr_{\ell} is even for all ν\ell\mid\nu. Similarly, we get (A)[ν]G2/ν\Sha(A)[\nu]\simeq G^{2}\oplus\mathbb{Z}/\nu\mathbb{Z} if rr_{\ell} is odd for all ν\ell\mid\nu. ∎

Lemma 7.4.5.

Maintain hypotheses from 5.1.4 and notation from 7.4.1. Assume ν\nu is odd, n>0n>0, and BB is an integral affine scheme with 2ν2\nu invertible on BB. Let bBb\in B be a closed point over which bA[ν]\mathscr{F}_{b}\simeq A[\nu], for AUbA\to U_{b} an abelian scheme, as in 5.1.8. The distributions XA[ν]𝔽qnX_{A[\nu]^{n}_{\mathbb{F}_{q}}} are supported on 𝒩0𝒩1\mathcal{N}^{0}\coprod\mathcal{N}^{1}. Hence,

(7.2) XA[ν]𝔽qn=Prob(XA[ν]𝔽qn𝒩0)XA[ν]𝔽qn0+Prob(XA[ν]𝔽qn𝒩1)XA[ν]𝔽qn1.X_{A[\nu]^{n}_{\mathbb{F}_{q}}}=\operatorname{Prob}(X_{A[\nu]^{n}_{\mathbb{F}_{q}}}\in\mathcal{N}^{0})\cdot X^{0}_{A[\nu]^{n}_{\mathbb{F}_{q}}}+\operatorname{Prob}(X_{A[\nu]^{n}_{\mathbb{F}_{q}}}\in\mathcal{N}^{1})\cdot X^{1}_{A[\nu]^{n}_{\mathbb{F}_{q}}}.
Proof.

The claim (7.2) follows from the first claim about the support of XA[ν]𝔽qnX_{A[\nu]^{n}_{\mathbb{F}_{q}}} by the law of total expectation. We now verify XA[ν]𝔽qnX_{A[\nu]^{n}_{\mathbb{F}_{q}}} are supported on 𝒩0𝒩1\mathcal{N}^{0}\coprod\mathcal{N}^{1}.

Using notation as in 5.1.10, it is enough to show the Selmer group of any quadratic twist AxA_{x} of AA lies in 𝒩0\mathcal{N}^{0} or 𝒩1\mathcal{N}^{1}. In general, there is an exact sequence

(7.3) 0{0}Ax(Ux)/νAx(Ux){A_{x}(U_{x})/\nu A_{x}(U_{x})}Selν(Ax){\operatorname{Sel}_{\nu}(A_{x})}(Ax)[ν]{\Sha(A_{x})[\nu]}0.{0.}

By 7.4.2, (Ax)[ν]\Sha(A_{x})[\nu] lies in 𝒩0𝒩1\mathcal{N}^{0}\coprod\mathcal{N}^{1}. By 5.2.6(2’), Ax[ν]=0A_{x}[\nu]=0, which implies that Ax(Ux)/νAx(Ux)A_{x}(U_{x})/\nu A_{x}(U_{x}) is a free /ν\mathbb{Z}/\nu\mathbb{Z} module. Hence, since /ν\mathbb{Z}/\nu\mathbb{Z} is injective as a /ν\mathbb{Z}/\nu\mathbb{Z} module, the exact sequence (7.3) splits and we obtain Selν(Ax)Ax(Ux)/νAx(Ux)(Ax)[ν]\operatorname{Sel}_{\nu}(A_{x})\simeq A_{x}(U_{x})/\nu A_{x}(U_{x})\oplus\Sha(A_{x})[\nu]. Now, we see that since (A)[ν]𝒩0𝒩1\Sha(A)[\nu]\in\mathcal{N}^{0}\coprod\mathcal{N}^{1} and Ax(Ux)/νAx(Ux)A_{x}(U_{x})/\nu A_{x}(U_{x}) is a free /ν\mathbb{Z}/\nu\mathbb{Z} module, Selν(Ax)𝒩0𝒩1\operatorname{Sel}_{\nu}(A_{x})\in\mathcal{N}^{0}\coprod\mathcal{N}^{1}. ∎

Finally, we are prepared to compute the image of the Dickson invariant map.

Lemma 7.4.6.

Assume ν\nu is odd, n>0n>0, and BB is an integral affine base scheme BB with 2ν2\nu invertible on BB. Suppose bBb\in B is a closed point with finite residue field, and keep hypotheses as in 5.1.4 and 7.1.4. Assume there is an abelian scheme AUbA\to U_{b} so that bA[ν]\mathscr{F}_{b}\simeq A[\nu], as in 5.1.8. The Dickson invariant map DQbn:O(Qbn)ν/2D_{Q_{\mathscr{F}^{n}_{b}}}:\operatorname{O}(Q_{\mathscr{F}^{n}_{b}})\to\prod_{\ell\mid\nu}\mathbb{Z}/2\mathbb{Z} sends the arithmetic monodromy group im(ρbn)\operatorname{im}(\rho_{\mathscr{F}^{n}_{b}}) surjectively to the diagonal copy of Δ/2:/2ν/2\Delta_{\mathbb{Z}/2\mathbb{Z}}:\mathbb{Z}/2\mathbb{Z}\subset\prod_{\ell\mid\nu}\mathbb{Z}/2\mathbb{Z}. The same holds for the geometric monodromy group at a geometric point b¯\overline{b} over bb.

Proof.

First, we argue it suffices to show the Dickson invariant of the arithmetic monodromy group satisfies im(DQbnρbn)imΔ/2.\operatorname{im}(D_{Q_{\mathscr{F}^{n}_{b}}}\circ\rho_{\mathscr{F}^{n}_{b}})\subset\operatorname{im}\Delta_{\mathbb{Z}/2\mathbb{Z}}. Indeed, for b¯\overline{b} a geometric point over bb, the image of the arithmetic monodromy group im(DQbn)\operatorname{im}(D_{Q_{\mathscr{F}^{n}_{b}}}) contains the image of the geometric monodromy group im(DQb¯n)\operatorname{im}(D_{Q_{\mathscr{F}^{n}_{\overline{b}}}}). Assuming we have shown the arithmetic monodromy has image the diagonal /2\mathbb{Z}/2\mathbb{Z} under the Dickson invariant map, to show they are equal, it is enough to show the geometric monodromy has nontrivial image under the Dickson invariant map. Equivalently, we wish to show the geometric monodromy is not contained in the special orthogonal group, which follows from Theorem 7.1.1.

We now verify the arithmetic monodromy group has Dickson invariant contained in imΔ/2.\operatorname{im}\Delta_{\mathbb{Z}/2\mathbb{Z}}. The strategy will be to use 7.4.5 to determine the arithmetic monodromy by relating the Dickson invariant map to the parity of the rank of Selmer groups modulo different primes, using equidistribution of Frobenius.

Choose xQTwistUb/bn(𝔽q)x\in\operatorname{QTwist}^{n}_{U_{b}/b}(\mathbb{F}_{q}). As a first step, we identify Selν(Ax)\operatorname{Sel}_{\nu}(A_{x}) with the 11-eigenspace of ρbn(Frobx)\rho_{\mathscr{F}^{n}_{b}}(\operatorname{Frob}_{x}), for Frobx\operatorname{Frob}_{x} the geometric Frobenius at xx. With notation as in 5.3.2, we can identify π1(x)(𝔽q)Selν(Ax)\pi^{-1}(x)(\mathbb{F}_{q})\simeq\operatorname{Sel}_{\nu}(A_{x}). Since π1(x)(𝔽q)\pi^{-1}(x)(\mathbb{F}_{q}) can be identified with the Frobx\operatorname{Frob}_{x} invariants of π1(x)(𝔽¯q)\pi^{-1}(x)(\overline{\mathbb{F}}_{q}), if gx:=ρbn(Frobx)g_{x}:=\rho_{\mathscr{F}^{n}_{b}}(\operatorname{Frob}_{x}), we also have π1(x)(𝔽q)ker(gxid)\pi^{-1}(x)(\mathbb{F}_{q})\simeq\ker(g_{x}-\operatorname{\mathrm{id}}). Combining these two isomorphisms, we obtain ker(gxid)Selν(Ax)\ker(g_{x}-\operatorname{\mathrm{id}})\simeq\operatorname{Sel}_{\nu}(A_{x}). For ν\ell\mid\nu, we use gx,g_{x,\ell} to denote the image of gxg_{x} under the map O(Qbn)O(Qbn[]){\rm{O}}(Q_{\mathscr{F}^{n}_{b}})\to{\rm{O}}(Q_{\mathscr{F}^{n}_{b}[\ell]}). We similarly obtain ker(gx,id)Sel(Ax)\ker(g_{x,\ell}-\operatorname{\mathrm{id}})\simeq\operatorname{Sel}_{\ell}(A_{x}).

We next constrain the image of the Dickson invariant map applied to ρbn(Frobx)\rho_{\mathscr{F}^{n}_{b}}(\operatorname{Frob}_{x}). From 7.4.5, we have seen that ker(gxid)Selν(Ax)𝒩0𝒩1\ker(g_{x}-\operatorname{\mathrm{id}})\simeq\operatorname{Sel}_{\nu}(A_{x})\in\mathcal{N}^{0}\coprod\mathcal{N}^{1}, for 𝒩i\mathcal{N}^{i} defined in 7.4.1. Since the parity of the rank of H/HH/\ell H of any group HH in 𝒩0𝒩1\mathcal{N}^{0}\coprod\mathcal{N}^{1} is independent of the prime ν\ell\mid\nu, it follows that dimker(gx,id)\dim\ker(g_{x,\ell}-\operatorname{\mathrm{id}}) has parity of rank independent of \ell, for ν\ell\mid\nu. By 2.1.3, for any ν\ell\mid\nu,

dimker(gx,id)mod2rkVbn[]DQbn(gx,).\displaystyle\dim\ker(g_{x,\ell}-\operatorname{\mathrm{id}})\bmod 2\equiv\operatorname{rk}V_{\mathscr{F}^{n}_{b}[\ell]}-D_{Q_{\mathscr{F}^{n}_{b}}}(g_{x,\ell}).

Since rkVbn[]\operatorname{rk}V_{\mathscr{F}^{n}_{b}[\ell]} is independent of ν\ell\mid\nu, as VbnV_{\mathscr{F}^{n}_{b}} is a free /ν\mathbb{Z}/\nu\mathbb{Z} module, we also obtain DQbn(gx,)D_{Q_{\mathscr{F}^{n}_{b}}}(g_{x,\ell}) is independent of ν\ell\mid\nu. In other words, the Dickson invariant map factors through the diagonal copy /2\mathbb{Z}/2\mathbb{Z} for each Frobenius element associated to xQTwistUb/bn(𝔽q)x\in\operatorname{QTwist}^{n}_{U_{b}/b}(\mathbb{F}_{q}).

The lemma will now follow from equidistribution of Frobenius elements in the arithmetic fundamental group, as we next explain. At this point, we employ a result on equidistribution of Frobenius, whose precise form we could not find directly in the literature. The result is essentially [Cha97, Theorem 4.1] (see also [Kow06, Theorem 1] and [FLR23, Theorem 3.9]) except that we need a slightly more general statement which also applies to Deligne-Mumford stacks in place of only schemes. The only part of the proof of [Cha97, Theorem 4.1] which does not directly apply to stacks is its use of the Grothendieck-Lefschetz trace formula, but this has been generalized to hold in the context of stacks, see [Sun12, Theorem 4.2]. Using this, we can find a sufficiently large qq and xQTwistUb/bn(𝔽q)x\in\operatorname{QTwist}^{n}_{U_{b}/b}(\mathbb{F}_{q}) with the following property: the generator Frobx\operatorname{Frob}_{x} of π1(x)\pi_{1}(x) is sent to any particular element of im(DQbnρbn)\operatorname{im}(D_{Q_{\mathscr{F}^{n}_{b}}}\circ\rho_{\mathscr{F}^{n}_{b}}) under the composition π1(x)π1(QTwistUb/bn)DQbnρbnν/2\pi_{1}(x)\to\pi_{1}(\operatorname{QTwist}^{n}_{U_{b}/b})\xrightarrow{D_{Q_{\mathscr{F}^{n}_{b}}}\circ\rho_{\mathscr{F}^{n}_{b}}}\prod_{\ell\mid\nu}\mathbb{Z}/2\mathbb{Z}. For our choice of qq above, note that we may need to take qq to be suitably large, and also if q=pjq=p^{j} for p=char𝔽qp=\operatorname{\operatorname{char}}\mathbb{F}_{q} we may need to impose a congruence condition on jj. Therefore, since every Frobx\operatorname{Frob}_{x} has image contained in the diagonal /2\mathbb{Z}/2\mathbb{Z}, the same must be true of im(DQbnρbn)\operatorname{im}(D_{Q_{\mathscr{F}^{n}_{b}}}\circ\rho_{\mathscr{F}^{n}_{b}}). ∎

8. The rank double cover

Perhaps surprisingly, the distribution of Selmer groups of abelian varieties is not determined by its moments. As mentioned in the introduction, if one fixes the parity of the rank of Sel\operatorname{Sel}_{\ell}, this does not change the distribution of Selmer groups. Even more surprisingly, once one does condition on the parity of the rank of Sel\operatorname{Sel}_{\ell}, the BKLPR distribution is determined by its moments. In this section, we investigate the geometry associated to a certain double cover of QTwistU/Bn\operatorname{QTwist}^{n}_{U/B}, which we define in § 8.1. In § 8.2, we will use our homological stability machinery to bound the dimensions of the cohomology of this double cover. In § 8.3, we relate this double cover to the parity of the dimension of Sel\operatorname{Sel}_{\ell} of an abelian variety. Specifically, suppose we are given a symplectically self-dual sheaf \mathscr{F} on UU, and a point bBb\in B with A[ν]\mathscr{F}\simeq A[\nu], for AUbA\to U_{b} an abelian scheme. We will define a particular double cover QTwistrk,n\operatorname{QTwist}^{\operatorname{rk},n}_{\mathscr{F}} of QTwistU/Bn\operatorname{QTwist}^{n}_{U/B} so that the images QTwistbrk,n(𝔽q)QTwistUb/bn(𝔽q)\operatorname{QTwist}^{\operatorname{rk},n}_{\mathscr{F}_{b}}(\mathbb{F}_{q})\to\operatorname{QTwist}^{n}_{U_{b}/b}(\mathbb{F}_{q}) corresponds precisely to abelian varieties whose \ell^{\infty} Selmer rank has parity equal to rkVBnmod2\operatorname{rk}V_{\mathscr{F}^{n}_{B}}\bmod 2.

8.1. The rank double cover and its coefficient system

We now define the rank double cover, and subsequently proceed to show the sequence of rank double covers form a coefficient system.

Definition 8.1.1.

With notation as in 7.1.1 and 2.1.1, suppose the composition DQBnρBn:π1(QTwistC/Bn)ν/2D_{Q_{\mathscr{F}^{n}_{B}}}\circ\rho_{\mathscr{F}^{n}_{B}}:\pi_{1}(\operatorname{QTwist}^{n}_{C/B})\to\prod_{\ell\mid\nu}\mathbb{Z}/2\mathbb{Z} factors through the diagonally embedded Δ/2:/2ν/2\Delta_{\mathbb{Z}/2\mathbb{Z}}:\mathbb{Z}/2\mathbb{Z}\to\prod_{\ell\mid\nu}\mathbb{Z}/2\mathbb{Z}. We define QTwistrk,nQTwistU/Bn\operatorname{QTwist}^{\operatorname{rk},n}_{\mathscr{F}}\to\operatorname{QTwist}^{n}_{U/B} to be the finite étale double cover corresponding to the composition DQBnρBnD_{Q_{\mathscr{F}^{n}_{B}}}\circ\rho_{\mathscr{F}^{n}_{B}}, viewed as a map π1(QTwistC/Bn)/2\pi_{1}(\operatorname{QTwist}^{n}_{C/B})\to\mathbb{Z}/2\mathbb{Z}.

Remark 8.1.2.

When we are in the situation of 7.4.6, it follows from 7.4.6, that the map DQBnρBn:π1(QTwistC/Bn)ν/2D_{Q_{\mathscr{F}^{n}_{B}}}\circ\rho_{\mathscr{F}^{n}_{B}}:\pi_{1}(\operatorname{QTwist}^{n}_{C/B})\to\prod_{\ell\mid\nu}\mathbb{Z}/2\mathbb{Z} takes image in the diagonally embedded copy of /2\mathbb{Z}/2\mathbb{Z}, so the hypothesis in 8.1.1 applies.

Since the rank double cover is a cover of QTwistU/Bn\operatorname{QTwist}^{n}_{U/B}, which is in turn a cover of ConfU/Bn\operatorname{Conf}^{n}_{U/B}, we can ask whether the composition QTwistrk,nConfU/Bn\operatorname{QTwist}^{\operatorname{rk},n}_{\mathscr{F}}\to\operatorname{Conf}^{n}_{U/B} is associated to a coefficient system. There is a technical issue with this question, in that QTwistU/Bn\operatorname{QTwist}^{n}_{U/B} is not a scheme, so the above cover is not representable. However, after suitably rigidifying this cover, we shall see that it indeed is associated to a coefficient system. In order to describe that coefficient system, we will first need to describe the coefficient systems associated to Selmer spaces and to their HH-moments.

Example 8.1.3.

Let B=SpecB=\operatorname{Spec}\mathbb{C} and let \mathscr{F} be a symplectically self-dual sheaf over UU as in 5.1.4. Fix a nontrivial finite /ν\mathbb{Z}/\nu\mathbb{Z} module HH. We now define a coefficient system of the type described in 3.1.9, which we will denote HS,H,g,fH_{S_{\mathscr{F},H,g,f}}. Recall that here we do not quotient by the conjugation action of the relevant group, see 3.1.10. The nnth part of HS,H,g,fH_{S_{\mathscr{F},H,g,f}} is the free vector space generated by a finite set S,H,g,fnS^{n}_{\mathscr{F},H,g,f} which we now define. Take GH:=AHSp2r(/ν)G_{H}:=\operatorname{\mathrm{A}^{\operatorname{H}}\mathrm{Sp}}_{2r}(\mathbb{Z}/\nu\mathbb{Z}), as in (6.6), and, with notation as in (6.6), take cH:=Π1(id)c_{H}:=\Pi^{-1}(-\operatorname{\mathrm{id}}), which is a conjugacy class in GHG_{H}. Take S,H,g,fnHom(π1(XnAg,fxn,pg,f),GH)S^{n}_{\mathscr{F},H,g,f}\subset\mathrm{Hom}(\pi_{1}(X^{\oplus n}\oplus A_{g,f}-x^{\oplus n},p_{g,f}),G_{H}) to be the subset 𝒮\mathcal{S} described in 6.4.1. (So, we are calling S,H,g,fnS^{n}_{\mathscr{F},H,g,f} what we called TGH,cH,g,fnT^{n}_{G_{H},c_{H},g,f} in 3.1.9.) More precisely, S,H,g,fnHom(π1(YnAg,fxn,pg,f),GH)S^{n}_{\mathscr{F},H,g,f}\subset\mathrm{Hom}(\pi_{1}(Y^{\oplus n}\oplus A_{g,f}-x^{\oplus n},p_{g,f}),G_{H}) is the subset consisting of those homomorphisms which send the loops around the nn punctures to cHc_{H}, which send local inertia around the f+1f+1 punctures to the conjugacy class described in 6.4.1, and which have the same projection to Sp2r(/ν)/{±1}\mathrm{Sp}_{2r}(\mathbb{Z}/\nu\mathbb{Z})/\{\pm 1\} as does ρ\rho_{\mathscr{F}}.

So long as we choose the basepoint pg,fp_{g,f} to lie on the boundary of Ag,fA_{g,f}, we can also restrict any homomorphism Hom(π1(YnAg,fxn,pg,f),GH)\mathrm{Hom}(\pi_{1}(Y^{\oplus n}\oplus A_{g,f}-x^{\oplus n},p_{g,f}),G_{H}) to a homomorphism Hom(π1(Ynxn,pg,f),GH)\mathrm{Hom}(\pi_{1}(Y^{\oplus n}-x^{\oplus n},p_{g,f}),G_{H}). We denote by S¯,H,0,0nHom(π1(Ynxn,pg,f),GH)S^{n}_{\underline{\mathscr{F}},H,0,0}\subset\mathrm{Hom}(\pi_{1}(Y^{\oplus n}-x^{\oplus n},p_{g,f}),G_{H}) the restriction of S,H,g,fnS^{n}_{\mathscr{F},H,g,f} to Hom(π1(Ynxn,pg,f),GH)\mathrm{Hom}(\pi_{1}(Y^{\oplus n}-x^{\oplus n},p_{g,f}),G_{H}). Define HS¯,H,0,0H_{S_{\underline{\mathscr{F}},H,0,0}} to be the associated coefficient system, whose nnth piece is HS¯,H,0,0nH_{S^{n}_{\underline{\mathscr{F}},H,0,0}}, the free vector space generated by S¯,H,0,0nS^{n}_{\underline{\mathscr{F}},H,0,0}.

Take V:=HS¯,H,0,0V:=H_{S_{\underline{\mathscr{F}},H,0,0}} and take F:=HS,H,g,fF:=H_{S_{\mathscr{F},H,g,f}}. We claim that VV forms a coefficient system for Σ0,01\Sigma^{1}_{0,0} and FF forms a coefficient system for Σg,f1\Sigma^{1}_{g,f} over VV. Indeed, these sets S,H,g,fnS^{n}_{\mathscr{F},H,g,f} are fixed under the action of Bg,fnB^{n}_{g,f} by 6.4.3. Hence, they form a coefficient system by 3.1.9. We can identify TGH,cH,g,fn+1cH×TGH,cH,g,fnT^{n+1}_{G_{H},c_{H},g,f}\simeq c_{H}\times T^{n}_{G_{H},c_{H},g,f}, where the map to cHc_{H} is given by the local inertia around the added puncture. It follows that Fn+1=k{Tg,fn+1}k{cH}k{TGH,cH,g,fn}=V1FnF_{n+1}=k\{T^{n+1}_{g,f}\}\simeq k\{c_{H}\}\otimes k\{T^{n}_{G_{H},c_{H},g,f}\}=V_{1}\otimes F_{n}. In the case g=f=0g=f=0, we similarly obtain that VV is a coefficient system.

We also define HurS,H,g,f\operatorname{Hur}_{S_{\mathscr{F},H,g,f}} to be the finite covering space of ConfXnAg,fn\operatorname{Conf}^{n}_{X^{\oplus n}\oplus A_{g,f}} associated to the set S,H,g,fnS^{n}_{\mathscr{F},H,g,f}.

Building on 8.1.3, we next describe the coefficient system corresponding to the rank double cover.

Example 8.1.4.

Take G=/2G=\mathbb{Z}/2\mathbb{Z}, c={1}/2c=\{1\}\in\mathbb{Z}/2\mathbb{Z}, corresponding to the nontrivial element, and consider the Hurwitz space coefficient system HT/2,{1},g,fH_{T_{\mathbb{Z}/2\mathbb{Z},\{1\},g,f}}. We recall that in the definition of these coefficient systems, we do not quotient by the conjugation action of the relevant group, see 3.1.10. By 3.1.9, this is a coefficient system for Σg,f1\Sigma^{1}_{g,f} which we claim lies over the trivial coefficient system for Σ0,01\Sigma^{1}_{0,0}. Indeed, HT/2,{1},0,0nH_{T^{n}_{\mathbb{Z}/2\mathbb{Z},\{1\},0,0}} is 11-dimensional because c={1}c=\{1\} has size 11 and moreover the coefficient system is trivial because /2\mathbb{Z}/2\mathbb{Z} is commutative.

We assume \mathscr{F} is a symplectically self dual sheaf as in 8.1.1. We use notation as in 8.1.3, and take the group HH there to be /ν\mathbb{Z}/\nu\mathbb{Z}. For every nn, there is a map of finite sets ϕBn:S,/ν,g,fnT/2,{1},g,fn\phi_{\mathscr{F}^{n}_{B}}:S^{n}_{\mathscr{F},\mathbb{Z}/\nu\mathbb{Z},g,f}\to T^{n}_{\mathbb{Z}/2\mathbb{Z},\{1\},g,f} which induces a map of Bg,fnB^{n}_{g,f} representations. Moreover, the fiber of ϕBn\phi_{\mathscr{F}^{n}_{B}}, which we call WBnW_{\mathscr{F}^{n}_{B}}, over any fixed point of the target can be identified with a free /ν\mathbb{Z}/\nu\mathbb{Z} module which has rank dimVBn+2r\dim V_{\mathscr{F}^{n}_{B}}+2r. There is an action of a fiber of \mathscr{F} on the finite cover of Qg,fnQ^{n}_{g,f} corresponding to WBnW_{\mathscr{F}^{n}_{B}} by conjugation, and we let W¯Bn\overline{W}_{\mathscr{F}^{n}_{B}} denote the quotient of WBnW_{\mathscr{F}^{n}_{B}} by this conjugation action. Since \mathscr{F} acts by conjugation on the corresponding cover, it follows that W¯Bn\overline{W}_{\mathscr{F}^{n}_{B}} inherits the structure of a Bg,fnB^{n}_{g,f} representation. The Bg,fnB^{n}_{g,f} representations corresponding to the sets T/2,{1},g,fT_{\mathbb{Z}/2\mathbb{Z},\{1\},g,f} yield a finite covering space of ConfXnAg,fn\operatorname{Conf}^{n}_{X^{\oplus n}\oplus A_{g,f}} of degree 22g2^{2g}, which we call Qg,fnQ^{n}_{g,f}.

In 8.1.5, we will construct a finite étale double cover RnR_{\mathscr{F}^{n}_{\mathbb{C}}} of Qg,fnQ^{n}_{g,f}. The fiber of RnR_{\mathscr{F}^{n}_{\mathbb{C}}} over a point of ConfXnAg,fn\operatorname{Conf}^{n}_{X^{\oplus n}\oplus A_{g,f}} corresponds to a finite set S,g,fn,rkS^{n,\operatorname{rk}}_{\mathscr{F},g,f} of order 22g+12^{2g+1} and yields a Bg,fnB^{n}_{g,f} representation which we call (H,g,frk)n(H^{\operatorname{rk}}_{\mathscr{F},g,f})_{n}. We will show (H,g,frk)n(H^{\operatorname{rk}}_{\mathscr{F},g,f})_{n} form a coefficient system for Σg,f1\Sigma^{1}_{g,f} over the trivial coefficient system for Σ0,01\Sigma^{1}_{0,0} in 8.1.7 and 8.1.8.

Lemma 8.1.5.

Continuing with notation as in 8.1.4, the action of π1(Qg,fn)\pi_{1}(Q^{n}_{g,f}) on W¯Bn\overline{W}_{\mathscr{F}^{n}_{B}} is /ν\mathbb{Z}/\nu\mathbb{Z} linear and factors through an orthogonal group OBn(/ν)\operatorname{O}_{\mathscr{F}^{n}_{B}}(\mathbb{Z}/\nu\mathbb{Z}). Moreover, the action factors through the preimage of the diagonal /2ν/2\mathbb{Z}/2\mathbb{Z}\subset\prod_{\ell\mid\nu}\mathbb{Z}/2\mathbb{Z} under the Dickson invariant and hence composition with the Dickson invariant defines a map π1(Qg,fn)OBn(/ν)/2\pi_{1}(Q^{n}_{g,f})\to\operatorname{O}_{\mathscr{F}^{n}_{B}}(\mathbb{Z}/\nu\mathbb{Z})\to\mathbb{Z}/2\mathbb{Z}, corresponding to a finite étale double cover RnQg,fnR_{\mathscr{F}^{n}_{\mathbb{C}}}\to Q^{n}_{g,f}.

Proof.

We may identify the Selmer stack SelBn\operatorname{Sel}_{\mathscr{F}^{n}_{B}} with a Hurwitz stack HurBn/ν\operatorname{Hur}^{\mathbb{Z}/\nu\mathbb{Z}}_{\mathscr{F}^{n}_{B}} via 6.4.5. This Hurwitz stack HurBn/ν\operatorname{Hur}^{\mathbb{Z}/\nu\mathbb{Z}}_{\mathscr{F}^{n}_{B}} has a further cover given by a pointed Hurwitz space as in 2.4.5, which is identified with the cover corresponding to the coefficient system whose nnth part is S,/ν,g,fnS^{n}_{\mathscr{F},\mathbb{Z}/\nu\mathbb{Z},g,f}, via 3.1.10. Quotienting WBnW_{\mathscr{F}^{n}_{B}} by the conjugation action of a fiber of \mathscr{F}, we obtain W¯Bn\overline{W}_{\mathscr{F}^{n}_{B}}. This corresponds to quotienting the the pointed Hurwitz space by the conjugation action, which is the Hurwitz space HurBn/ν\operatorname{Hur}^{\mathbb{Z}/\nu\mathbb{Z}}_{\mathscr{F}^{n}_{B}}, so we obtain an identification of W¯Bn\overline{W}_{\mathscr{F}^{n}_{B}} as a π1(Qg,fn)\pi_{1}(Q^{n}_{g,f}) representation with a geometric fiber of HurBn/ν×QTwistU/BnQg,fn\operatorname{Hur}^{\mathbb{Z}/\nu\mathbb{Z}}_{\mathscr{F}^{n}_{B}}\times_{\operatorname{QTwist}^{n}_{U/B}}Q^{n}_{g,f}. over Qg,fnQ^{n}_{g,f}. Hence, we may identify W¯Bn\overline{W}_{\mathscr{F}^{n}_{B}} as a π1(Qg,fn)\pi_{1}(Q^{n}_{g,f}) representation with VBnV_{\mathscr{F}^{n}_{B}} as a π1(Qg,fn)\pi_{1}(Q^{n}_{g,f}) set, viewing π1(Qg,fn)π1(QTwistU/Bn)\pi_{1}(Q^{n}_{g,f})\subset\pi_{1}(\operatorname{QTwist}^{n}_{U/B}), as Qg,fnQ^{n}_{g,f} is a finite étale double cover of QTwistU/Bn\operatorname{QTwist}^{n}_{U/B}. Hence, we obtain the factorization through the claimed group by our assumption on \mathscr{F} from 8.1.1. ∎

Gearing up to explicitly describe the rank double cover as a coefficient system, we next record, in terms of generators, the action of π1(ConfΣg,f1n,[S])\pi_{1}(\operatorname{Conf}^{n}_{\Sigma^{1}_{g,f}},[S]) on π1(Σg,f+n1,p)\pi_{1}(\Sigma^{1}_{g,f+n},p) for [S][S] and pp basepoints. One can prove the description in 8.1.6 by computing where the loops as in § 6.2 are sent under the appropriate Dehn twists or half twists. Also related to this is the explicit presentation for π1(ConfΣg,f1n)\pi_{1}(\operatorname{Conf}^{n}_{\Sigma^{1}_{g,f}}) given in [Bel04, Theorem 1.1], which shows that the four types of loops described in 8.1.6 generate π1(ConfΣg,f1n,[S])\pi_{1}(\operatorname{Conf}^{n}_{\Sigma^{1}_{g,f}},[S]).

Remark 8.1.6.

Use notation as in § 6.2 for loops on Σg,f+n1\Sigma^{1}_{g,f+n}, where the nn punctures correspond to a set SΣg,f1S\subset\Sigma^{1}_{g,f}. We use δf+1\delta_{f+1} for a loop around the boundary component and p:=sf+1p:=s_{f+1}. For nn even, the action of certain generators of π1(ConfΣg,f1n,[S])\pi_{1}(\operatorname{Conf}^{n}_{\Sigma^{1}_{g,f}},[S]) on π1(Σg,f+n1,sf+1)\pi_{1}(\Sigma^{1}_{g,f+n},s_{f+1}) act on the data

(8.1) (α1,β1,,αg,βg,γ1,,γn,δ1,,δf+1)\displaystyle(\alpha_{1},\beta_{1},\ldots,\alpha_{g},\beta_{g},\gamma_{1},\ldots,\gamma_{n},\delta_{1},\ldots,\delta_{f+1})

in the following way:

  1. (1)

    The full twist of pnp_{n} around a loop surrounding s1,s2,,sis_{1},s_{2},\ldots,s_{i} sends (8.1) to

    (α1,,γn1,γnδ1δiγn(δ1δi)1γn1,γnδ1γn1,,γnδiγn1,δi+1,,δf+1).\displaystyle(\alpha_{1},\ldots,\gamma_{n-1},\gamma_{n}\delta_{1}\cdots\delta_{i}\gamma_{n}(\delta_{1}\cdots\delta_{i})^{-1}\gamma_{n}^{-1},\gamma_{n}\delta_{1}\gamma_{n}^{-1},\ldots,\gamma_{n}\delta_{i}\gamma_{n}^{-1},\delta_{i+1},\ldots,\delta_{f+1}).
  2. (2)

    For 1in11\leq i\leq n-1, the half-twist of pip_{i} around pi+1p_{i+1} sends (8.1) to

    (α1,,γi1,γiγi+1γi1,γi,γi+2,γi+3,,γn,δ1,,δn).\displaystyle(\alpha_{1},\ldots,\gamma_{i-1},\gamma_{i}\gamma_{i+1}\gamma_{i}^{-1},\gamma_{i},\gamma_{i+2},\gamma_{i+3},\ldots,\gamma_{n},\delta_{1},\ldots,\delta_{n}).
  3. (3)

    Moving p1p_{1} across αi\alpha_{i} and then in a loop around sf+1,,s1,pn,,p2s_{f+1},\ldots,s_{1},p_{n},\ldots,p_{2} sends (8.1) to

    (α1,,βi1,(j=1i1[αj,βj])1γ1(j=1i1[αj,βj])αi,βi,αi+1,,βg,\displaystyle\left(\alpha_{1},\ldots,\beta_{i-1},\left(\prod_{j=1}^{i-1}[\alpha_{j},\beta_{j}]\right)^{-1}\gamma_{1}\left(\prod_{j=1}^{i-1}[\alpha_{j},\beta_{j}]\right)\alpha_{i},\beta_{i},\alpha_{i+1},\ldots,\beta_{g},\right.
    ((j=1i1[αj,βj])βi1(j=i+1g[αj,βj]))1γ1((j=1i1[αj,βj])βi1(j=i+1g[αj,βj])),γ1γ2γ11,,γ1δf+1γ11).\displaystyle\qquad\left.\left(\left(\prod_{j=1}^{i-1}[\alpha_{j},\beta_{j}]\right)\beta_{i}^{-1}\left(\prod_{j=i+1}^{g}[\alpha_{j},\beta_{j}]\right)\right)^{-1}\gamma_{1}\left(\left(\prod_{j=1}^{i-1}[\alpha_{j},\beta_{j}]\right)\beta_{i}^{-1}\left(\prod_{j=i+1}^{g}[\alpha_{j},\beta_{j}]\right)\right),\gamma_{1}\gamma_{2}\gamma_{1}^{-1},\ldots,\gamma_{1}\delta_{f+1}\gamma_{1}^{-1}\right).
  4. (4)

    Moving p1p_{1} across βi\beta_{i} and then in a loop around sf+1,,s1,pn,,p2s_{f+1},\ldots,s_{1},p_{n},\ldots,p_{2} sends (8.1) to

    (α1,,βi1,αi,αi1(j=1i1[αj,βj])1γ1(j=1i1[αj,βj])αiβi,αi+1,,αg,βg,\displaystyle\left(\alpha_{1},\ldots,\beta_{i-1},\alpha_{i},\alpha_{i}^{-1}\left(\prod_{j=1}^{i-1}[\alpha_{j},\beta_{j}]\right)^{-1}\gamma_{1}\left(\prod_{j=1}^{i-1}[\alpha_{j},\beta_{j}]\right)\alpha_{i}\beta_{i},\alpha_{i+1},\ldots,\alpha_{g},\beta_{g},\right.
    ((j=1i1[αj,βj])αi(j=i+1g[αj,βj]))1γ1((j=1i1[αj,βj])αi(j=i+1g[αj,βj])),γ1γ2γ11,,γ1δf+1γ11).\displaystyle\qquad\left.\left(\left(\prod_{j=1}^{i-1}[\alpha_{j},\beta_{j}]\right)\alpha_{i}\left(\prod_{j=i+1}^{g}[\alpha_{j},\beta_{j}]\right)\right)^{-1}\gamma_{1}\left(\left(\prod_{j=1}^{i-1}[\alpha_{j},\beta_{j}]\right)\alpha_{i}\left(\prod_{j=i+1}^{g}[\alpha_{j},\beta_{j}]\right)\right),\gamma_{1}\gamma_{2}\gamma_{1}^{-1},\ldots,\gamma_{1}\delta_{f+1}\gamma_{1}^{-1}\right).
Lemma 8.1.7.

We use the notation introduced in 8.1.4. For nn even, and B=SpecB=\operatorname{Spec}\mathbb{C}, any element of π1(Qg,fn)\pi_{1}(Q^{n}_{g,f}) mapping to one of the following elements of π1(ConfXnAg,fn)\pi_{1}(\operatorname{Conf}^{n}_{X^{\oplus n}\oplus A_{g,f}}) act on W¯Bn\overline{W}_{\mathscr{F}^{n}_{B}} with trivial Dickson invariant:

  1. (1)

    moving pip_{i} in a half-twist about pi+1p_{i+1}, which is conjugate to 8.1.6(2),

  2. (2)

    moving pip_{i} twice across αi\alpha_{i} or βi\beta_{i}, corresponding to a conjugate of the square of the transformation from 8.1.6(3) or (4).

Elements of π1(ConfXnAg,fn)\pi_{1}(\operatorname{Conf}^{n}_{X^{\oplus n}\oplus A_{g,f}}) sending pjp_{j} around sis_{i} as in 8.1.6(1) may act either with trivial or nontrivial Dickson invariant, where the triviality of the Dickson invariant is a function of ii but not jj.

Proof.

We now explain how the claims can be deduced from the explicit formula for the action of π1(ConfΣg,f1n)\pi_{1}(\operatorname{Conf}^{n}_{\Sigma^{1}_{g,f}}) on π1(Σg,f1)\pi_{1}(\Sigma^{1}_{g,f}) from 8.1.6. We will use this in conjunction with the description of VBnV_{\mathscr{F}^{n}_{B}} in 6.3.7 to verify 8.1.7 with a modicum of computation.

We now describe coordinates for a certain free /ν\mathbb{Z}/\nu\mathbb{Z} module PnP_{n} of rank 2r(2g+n+f+1)2r(2g+n+f+1) of which W¯Bn\overline{W}_{\mathscr{F}^{n}_{B}} is a subquotient. Consider the free /ν\mathbb{Z}/\nu\mathbb{Z} module with the coordinates xjix^{i}_{j} for 1j2r1\leq j\leq 2r, 1i2g+n+(f+1)1\leq i\leq 2g+n+(f+1). Here, the ii indexes the 2g+n+(f+1)2g+n+(f+1) different entries in (8.1) while the jj indexes the coordinate in the vector vv upon plugging in a matrix of the form (6.4) for each such entry.

Let p0p_{0} be a point on UU disjoint from DD and pp. The group π1(UDp0,p)\pi_{1}(U-D-p_{0},p) is free on the generators (α1,β1,,αg,βg,γ1,,γn,δ1,,δf+1)(\alpha_{1},\beta_{1},\ldots,\alpha_{g},\beta_{g},\gamma_{1},\ldots,\gamma_{n},\delta_{1},\ldots,\delta_{f+1}), and also contains an element δf+2\delta_{f+2} which satisfies the relation

(8.2) (α1β1α11β11)(αgβgαg1βg1)γ1γnδ1δf+1=δf+21.\displaystyle(\alpha_{1}\beta_{1}\alpha_{1}^{-1}\beta_{1}^{-1})\cdots(\alpha_{g}\beta_{g}\alpha_{g}^{-1}\beta_{g}^{-1})\gamma_{1}\cdots\gamma_{n}\delta_{1}\cdots\delta_{f+1}=\delta_{f+2}^{-1}.

We can also think of PnP_{n} as the set of group homomorphisms ϕ:π1(UDp0,p)ASp2r(/ν)\phi:\pi_{1}(U-D-p_{0},p)\to\operatorname{\mathrm{ASp}}_{2r}(\mathbb{Z}/\nu\mathbb{Z}) whose projection to Sp2r(/ν)\mathrm{Sp}_{2r}(\mathbb{Z}/\nu\mathbb{Z}) is ρ\rho_{\mathscr{F}^{\prime}}. (In particular, this implies that ϕ(δf+2)\phi(\delta_{f+2}) lies in (/ν)2r(\mathbb{Z}/\nu\mathbb{Z})^{2r}, since ρ(δf+2)=id.)\rho_{\mathscr{F}^{\prime}}(\delta_{f+2})=\operatorname{\mathrm{id}}.) The section s:Sp2r(/ν)ASp2r(/ν)s:\mathrm{Sp}_{2r}(\mathbb{Z}/\nu\mathbb{Z})\to\operatorname{\mathrm{ASp}}_{2r}(\mathbb{Z}/\nu\mathbb{Z}) affords one such homomorphism, namely ϕ0:=sρ\phi_{0}:=s\circ\rho_{\mathscr{F}^{\prime}}. (In the explicit form of (6.4), ϕ0\phi_{0} sends each generator to a pair (M,v)(M,v) with v=0v=0.)

Given such a ϕ\phi, we can attach to each of the 2g+n+f+12g+n+f+1 free generators of π1(UDp0,p)\pi_{1}(U-D-p_{0},p) an element of (/ν)2r(\mathbb{Z}/\nu\mathbb{Z})^{2r}; namely, we send a generator gg to ϕ(g)ϕ01(g)\phi(g)\phi_{0}^{-1}(g). This gives the desired (/ν)(\mathbb{Z}/\nu\mathbb{Z}) module of rank 2r(2g+n+f+1)2r(2g+n+f+1). Equivalently, we can think of PnP_{n} as the space of 11-cocycles from π1(UDp0,p)\pi_{1}(U-D-p_{0},p) to (/ν)2r(\mathbb{Z}/\nu\mathbb{Z})^{2r}, with the group action that given by ρ\rho_{\mathscr{F}^{\prime}}. This description makes it clear that the braids, which are automorphisms of π1(UDp0,p)\pi_{1}(U-D-p_{0},p) fixing ρ\rho_{\mathscr{F}^{\prime}}, act linearly on PnP_{n}. (Of course, this can also be derived from the explicit description of the braid group action.)

Note that W¯Bn\overline{W}_{\mathscr{F}^{n}_{B}} can be identified as a subquotient of PnP_{n} via 6.3.7. The reason for working with PnP_{n} rather than with W¯Bn\overline{W}_{\mathscr{F}^{n}_{B}} directly is that the explicit description of PnP_{n} makes it easier to work out the action of a braid in concrete enough terms to easily compute Dickson invariants.

We first address (2)(2) in the statement. To do so, let TαiT_{\alpha_{i}} denote the transformation described in 8.1.6(3), which moves p1p_{1} across αi\alpha_{i}. We wish to show the Dickson invariant of the transformation induced by Tαi2T_{\alpha_{i}}^{2} on W¯Bn\overline{W}_{\mathscr{F}^{n}_{B}} is trivial. Letting

η\displaystyle\eta :=(j=1i1[αj,βj])βi1(j=i+1g[αj,βj])\displaystyle:=\left(\prod_{j=1}^{i-1}[\alpha_{j},\beta_{j}]\right)\beta_{i}^{-1}\left(\prod_{j=i+1}^{g}[\alpha_{j},\beta_{j}]\right)
ε\displaystyle\varepsilon :=η1γ1η\displaystyle:=\eta^{-1}\gamma_{1}\eta
λ\displaystyle\lambda :=(j=1i1[αj,βj])1γ1(j=1i1[αj,βj])\displaystyle:=\left(\prod_{j=1}^{i-1}[\alpha_{j},\beta_{j}]\right)^{-1}\gamma_{1}\left(\prod_{j=1}^{i-1}[\alpha_{j},\beta_{j}]\right)
λ\displaystyle\lambda^{\prime} :=(j=1i1[αj,βj])1ε(j=1i1[αj,βj])\displaystyle:=\left(\prod_{j=1}^{i-1}[\alpha_{j},\beta_{j}]\right)^{-1}\varepsilon\left(\prod_{j=1}^{i-1}[\alpha_{j},\beta_{j}]\right)

and using the formula from 8.1.6(3), the transformation Tαi2T_{\alpha_{i}}^{2} sends (8.1) to

(8.3) (α1,,βi1,λλαi,βi,αi+1,,βg,η2γ1η2,εγ1γ2γ11ε1,,εγ1δf+1γ11ε1).\displaystyle\left(\alpha_{1},\ldots,\beta_{i-1},\lambda^{\prime}\lambda\alpha_{i},\beta_{i},\alpha_{i+1},\ldots,\beta_{g},\eta^{-2}\gamma_{1}\eta^{2},\varepsilon\gamma_{1}\gamma_{2}\gamma_{1}^{-1}\varepsilon^{-1},\ldots,\varepsilon\gamma_{1}\delta_{f+1}\gamma_{1}^{-1}\varepsilon^{-1}\right).

First, we will show the action of Tαi2T_{\alpha_{i}}^{2} on the free module PnP_{n} has square determinant. The key calculation which we will use repeatedly is the following. Given a matrix MSp2r(/ν)M\in\mathrm{Sp}_{2r}(\mathbb{Z}/\nu\mathbb{Z}) and v(/ν)2rv\in\left(\mathbb{Z}/\nu\mathbb{Z}\right)^{2r} we use (M,v)(M,v) to denote the element of ASp2r(/ν)\operatorname{\mathrm{ASp}}_{2r}(\mathbb{Z}/\nu\mathbb{Z}) as in (6.4). Then,

(8.4) (N,w)1(M,v)(N,w)\displaystyle\left(N,w\right)^{-1}\cdot(M,v)\cdot(N,w) =(N1,N1w)(MN,Mw+v)\displaystyle=(N^{-1},-N^{-1}w)\cdot(MN,Mw+v)
=(N1MN,N1w+N1v+N1Mw).\displaystyle=(N^{-1}MN,-N^{-1}w+N^{-1}v+N^{-1}Mw).

Now, Tαi2T_{\alpha_{i}}^{2} acting on PnP_{n} can be expressed as the composite of several transformations. It is the composite of the maps induced by sending γ2,,δf+1\gamma_{2},\ldots,\delta_{f+1} to their conjugate by εγ1\varepsilon\gamma_{1}, followed by the map induced by αiλλαi\alpha_{i}\mapsto\lambda^{\prime}\lambda\alpha_{i} followed by the map induced by γ1η2γ1η2\gamma_{1}\mapsto\eta^{-2}\gamma_{1}\eta^{2}.

We claim that each of these transformations have square determinant 11, and hence the composite will have square 11. First, we show the determinant is a square for each transformation induced by sending a matrix (M,v)(M,v) associated to one of γ2,,δf+1\gamma_{2},\ldots,\delta_{f+1} to its conjugate by a matrix (N,w)(N,w), associated to εγ1\varepsilon\gamma_{1}. In this case, N=idN=\operatorname{\mathrm{id}} and so the output of conjugation is (M,w+v+Mw)(M,-w+v+Mw) by (8.4). Since the xjix^{i}_{j} entries appearing in ww are disjoint from those appearing in vv, this transformation is unipotent, so has determinant 11.

We next consider the transformation coming from αiλλαi\alpha_{i}\mapsto\lambda^{\prime}\lambda\alpha_{i}. Since λ\lambda^{\prime} and λ\lambda are both conjugate to γ1\gamma_{1}, the element λλ\lambda^{\prime}\lambda corresponds to a pair of the form (id,w)(\operatorname{\mathrm{id}},w). So if (M,v)(M,v) corresponds to αi\alpha_{i}, then the transformation in question sends (M,v)(M,v) to (id,w)(M,v)=(M,v+w)(\operatorname{\mathrm{id}},w)\cdot(M,v)=(M,v+w). Once again, the xjix^{i}_{j} entries appearing in ww are disjoint from those appearing in vv, so this transformation is unipotent and has determinant 11.

Third, to calculate the determinant of the map induced by γ1η2γ1η2\gamma_{1}\mapsto\eta^{-2}\gamma_{1}\eta^{2} we use (8.4) with (N,w)(N,w) corresponding to η\eta and (M,v)(M,v) corresponding to γ1\gamma_{1}. Since M=1M=-1, the output of the transformation is (id,N1v2N1w)(-\operatorname{\mathrm{id}},N^{-1}v-2N^{-1}w). Since ww only involves xjix^{i}_{j} which are disjoint from those associated to γ1\gamma_{1}, appearing in vv, the determinant of this transformation agrees with that of NN. Since we are conjugating by η2\eta^{2}, the resulting NN has square determinant.

In order to conclude the Dickson invariant associated to the action of Tαi2T_{\alpha_{i}}^{2} on W¯Bn\overline{W}_{\mathscr{F}^{n}_{B}} is trivial, it remains to check that this operator still has square determinant upon passing to the subquotient W¯Bn\overline{W}_{\mathscr{F}^{n}_{B}} of PnP_{n}. To do so, note that W¯Bn\overline{W}_{\mathscr{F}^{n}_{B}} of PnP_{n} is obtained from PnP_{n} in three steps:

  1. (A)

    We first take the subspace dictated by the drop condition 6.3.7(2),

  2. (B)

    we then take the the subspace where the product (6.2) is satisfied, upon plugging in matrices for the generators of the fundamental group,

  3. (C)

    and finally we pass to the quotient by the conjugation action as described at the end of 6.3.7.

To prove Tαi2T_{\alpha_{i}}^{2} has square determinant (and hence trivial Dickson invariant) on W¯Bn\overline{W}_{\mathscr{F}^{n}_{B}}, we will show that for each of these steps, the induced operator on the associated subspace or quotient space has trivial determinant.

To simplify our calculations, we may and shall assume for the remainder of this proof that ν\nu is prime; this does not restrict the conclusion of the theorem, because it follows from 7.4.6 that, for any prime ν\ell\mid\nu, the Dickson invariant of the action of an automorphism of W¯Bn\overline{W}_{\mathscr{F}^{n}_{B}} can be computed on W¯Bn/ν/.\overline{W}_{\mathscr{F}^{n}_{B}}\otimes_{\mathbb{Z}/\nu\mathbb{\mathbb{Z}}}\mathbb{Z}/\ell\mathbb{Z}. Under this hypothesis, all our /ν\mathbb{Z}/\nu\mathbb{Z}-modules are now vector spaces over a field and the Dickson invariant is additive in exact sequences, a fact we will use repeatedly in the argument that follows.

For step (A)(A), we note that there is a homomorphism

L:Pni=1f+1[(/ν)2r/im(ρ(δi)id)]L:P_{n}\rightarrow\bigoplus_{i=1}^{f+1}\left[\left(\mathbb{Z}/\nu\mathbb{Z}\right)^{2r}/\operatorname{im}(\rho_{\mathscr{F}}(\delta_{i})-\operatorname{\mathrm{id}})\right]

whose kernel is the subspace specified by condition 6.3.7(2). The map LL is defined by projection onto the δ1,,δf+1\delta_{1},\ldots,\delta_{f+1} coordinates of PnP_{n} followed by projection of the δi\delta_{i} coordinate onto the quotient by im(ρ(δi)id)\operatorname{im}(\rho_{\mathscr{F}}(\delta_{i})-\operatorname{\mathrm{id}}).

We also know that Tαi2T_{\alpha_{i}}^{2} acts on the δi\delta_{i} coordinate by conjugation by εγ1ASp2r(/ν)\varepsilon\gamma_{1}\in\operatorname{\mathrm{ASp}}_{2r}(\mathbb{Z}/\nu\mathbb{Z}); since εγ1\varepsilon\gamma_{1} is an element of the form (id,w)(\operatorname{\mathrm{id}},w), conjugation by εγ1\varepsilon\gamma_{1} modifies the δi\delta_{i} coordinate by adding an element of the form (ρ(δi)ww)(\rho_{\mathscr{F}}(\delta_{i})w-w). Via a computation analogous to (8.4), for any xPnx\in P_{n}, we have L(x)=L(Tαi2x)L(x)=L(T_{\alpha_{i}}^{2}x). We conclude that Tαi2T_{\alpha_{i}}^{2} acts trivially (and a fortiori with trivial Dickson invariant) on Pn/L(Pn)P_{n}/L(P_{n}).

For step (B)(B), we observe that Tαi2T_{\alpha_{i}}^{2}, considered as an automorphism of π1(UDp0,p)\pi_{1}(U-D-p_{0},p), preserves the element δf+2\delta_{f+2}, and so it preserves the left-hand side of (8.2). In particular, if

M:Pn(/ν)2rM:P_{n}\to\left(\mathbb{Z}/\nu\mathbb{Z}\right)^{2r}

is the map provided by the left-hand side of (8.2), whose kernel is the subspace of PnP_{n} obeying (6.2), then the induced action of Tαi2T_{\alpha_{i}}^{2} on Pn/M(Pn)P_{n}/M(P_{n}) is trivial.

Finally, for step (C)(C), we wish to show Tαi2T_{\alpha_{i}}^{2} acts trivially on the subspace generated by coboundaries, corresponding to changing the basepoint of the original matrix. It is possible to compute this directly using the formula (8.3), but we will provide a more conceptual explanation. First, Tαi2T_{\alpha_{i}}^{2} acts linearly on PnP_{n} and thus fixes the zero element. But Tαi2T_{\alpha_{i}}^{2} commutes with the operation of conjugating all 2g+n+f+12g+n+f+1 coordinates by a matrix of the form (id,v)(\operatorname{\mathrm{id}},v), for any v(/ν)2rv\in(\mathbb{Z}/\nu\mathbb{Z})^{2r}. This operation is not linear on PnP_{n}, but it is affine-linear, acting by translation by a coboundary cvc_{v}. Since Tαi2T_{\alpha_{i}}^{2} commutes with translations by all coboundaries in PnP_{n} and fixes 0, it also fixes all coboundaries in PnP_{n}.

The combination of the three steps above allows us to conclude that Tαi2T_{\alpha_{i}}^{2} acts with square determinant on W¯Bn\overline{W}_{\mathscr{F}^{n}_{B}}, and hence has trivial Dickson invariant. Similar reasoning shows the action in 8.1.6(4) is trivial, concluding the verification of (2)(2).

Next, we will show the elements from (1)(1) of the statement act trivially on the module PnP_{n}. Using the description of the half-twist from 8.1.6(2), we can study the resulting 2r(2g+n+(f+1))×2r(2g+n+(f+1))2r(2g+n+(f+1))\times 2r(2g+n+(f+1)) matrix associated to how the transformation in 8.1.6(2) acts on PnP_{n}, upon plugging in matrices of the form (6.4) for each entry in 8.1.6(2). Because the MM from (6.4) associated to each γi\gamma_{i} is id-\operatorname{\mathrm{id}}, we find the above 2r(2g+n+(f+1))×2r(2g+n+(f+1))2r(2g+n+(f+1))\times 2r(2g+n+(f+1)) matrix is a block diagonal matrix, consisting of 2r2r blocks of size 2g+n+(f+1)2g+n+(f+1). In particular, the determinant of this matrix acting on PnP_{n} is a 2r2rth power, so it is a square. To conclude the Dickson invariant is trivial, it remains to justify why passing W¯Bn\overline{W}_{\mathscr{F}^{n}_{B}} of PnP_{n}, preserves the condition that the determinant is a square. The action on the quotient space to the drop condition from (A)(A) above is trivial, because the δi\delta_{i} are preserved by this transformation. Triviality of the action on the quotient associated to (B)(B) and subspace associated to (C)(C) follow in the same way as for Tαi2T_{\alpha_{i}}^{2} in the proof of (2)(2) from the statement of the lemma above.

The final part of the statement of this lemma, regarding the action in 8.1.6(1) holds because one can express the generator sending pjp_{j} around sis_{i} as a product of half twists permuting the pjp_{j} with the loop sending pnp_{n} around sis_{i}, and all these half twists have trivial image under the Dickson invariant map, as we have shown above. ∎

Lemma 8.1.8.

With notation as in 8.1.4, the sequence (H,g,frk)n(H^{\operatorname{rk}}_{\mathscr{F},g,f})_{n} of Bg,fnB^{n}_{g,f} representations defines a coefficient system for Σg,f1\Sigma^{1}_{g,f} over the trivial coefficient system VV for Σ0,01\Sigma^{1}_{0,0}.

Proof.

Using the explicit description of the double cover RnQg,fnR_{\mathscr{F}^{n}_{\mathbb{C}}}\to Q^{n}_{g,f} from 8.1.7(2) we first claim the double cover is in fact the base change of a double cover of ConfU/Bn\operatorname{Conf}^{n}_{U/B}. Indeed, to show this, it is equivalent to show we can extend the homomorphism π1(Qg,fn)/2\pi_{1}(Q^{n}_{g,f})\to\mathbb{Z}/2\mathbb{Z} to a homomorphism π1(ConfU/Bn)/2\pi_{1}(\operatorname{Conf}^{n}_{U/B})\to\mathbb{Z}/2\mathbb{Z}. We can extend this homomorphism, for example, by sending loops corresponding to moving pjp_{j} across αi\alpha_{i} or βi\beta_{i}, as in 8.1.6(3) and (4), to the trivial element of /2\mathbb{Z}/2\mathbb{Z}.

This shows the cover S,g,fn,rkS^{n,\operatorname{rk}}_{\mathscr{F},g,f} is then the product of a two element set {an,bn}\{a_{n},b_{n}\} corresponding to the above double cover of ConfU/Bn\operatorname{Conf}^{n}_{U/B} with the set T/2,{1},g,fnT^{n}_{\mathbb{Z}/2\mathbb{Z},\{1\},g,f}. Therefore, it suffices to show the free vector space on both of these sets form coefficient systems. First, T/2,{1},g,fnT^{n}_{\mathbb{Z}/2\mathbb{Z},\{1\},g,f} forms a coefficient system over the trivial coefficient system for Σ0,01\Sigma^{1}_{0,0} by 3.1.9.

Second, we explain why the explicit description of the action of π1(ConfU/Bn)\pi_{1}(\operatorname{Conf}^{n}_{U/B}) and {an,bn}\{a_{n},b_{n}\} obtained from 8.1.7 shows the free vector space on this collection of these two element sets, {an,bn}n1\{a_{n},b_{n}\}_{n\geq 1}, forms a coefficient system over the trivial coefficient system for Σ0,01\Sigma^{1}_{0,0}. Indeed, the description of half-twists from 8.1.7(1) shows this lies over the trivial coefficient system for Σ0,01\Sigma^{1}_{0,0}. The condition to be a coefficient system over the trivial coefficient system amounts to checking that the action of id×Bg,fniB0,0i×Bg,fniBg,fn\operatorname{\mathrm{id}}\times B^{n-i}_{g,f}\subset B^{i}_{0,0}\times B^{n-i}_{g,f}\subset B^{n}_{g,f}, on {an,bn}\{a_{n},b_{n}\} can be identified with the action of Bg,fniB^{n-i}_{g,f} on {ani,bni}\{a_{n-i},b_{n-i}\} via the map of sets {ani,bni}{an,bn}\{a_{n-i},b_{n-i}\}\to\{a_{n},b_{n}\} given by anian,bnibna_{n-i}\mapsto a_{n},b_{n-i}\mapsto b_{n}. This indeed holds since Bg,fniBg,fnB^{n-i}_{g,f}\subset B^{n}_{g,f}, is generated by a subset of the transformations described in 8.1.7, and the action of each generator of Bg,fniBg,fnB^{n-i}_{g,f}\subset B^{n}_{g,f} on {an,bn}\{a_{n},b_{n}\} acts in the same way on {ani,bni}\{a_{n-i},b_{n-i}\}, via 8.1.7. ∎

Remark 8.1.9.

In fact, the proof of 8.1.8 shows that the nnth graded part of the rank double cover can be identified with the a free vector space on a set VV of size 22g+12^{2g+1}, for VV an explicit quotient H1(Σg,f1,/2)VH^{1}(\Sigma^{1}_{g,f},\mathbb{Z}/2\mathbb{Z})\to V. The Bg,fnB^{n}_{g,f} action is obtained via a surjection Bg,fn(Bg,fn)ab(Bg,f1)abH1(Σg,f1,)H1(Σg,f1,/2)VB^{n}_{g,f}\to(B^{n}_{g,f})^{\operatorname{ab}}\simeq(B^{1}_{g,f})^{\operatorname{ab}}\simeq H^{1}(\Sigma^{1}_{g,f},\mathbb{Z})\to H^{1}(\Sigma^{1}_{g,f},\mathbb{Z}/2\mathbb{Z})\to V.

Remark 8.1.10.

It should come as no surprise that the result of the explicit topological computation of 8.1.7 ends up having a rather simple form, as described in 8.1.9. When =A[ν]\mathscr{F}=A[\nu] is the ν\nu-torsion of an abelian variety AA, it is possible to show that the Dickson invariant is determined by the root number of the quadratic twist AχA_{\chi} of AA, which in turn can be explicitly computed by the formula

W(Aχ)=W(A)χ(NA),W(A_{\chi})=W(A)\chi(N_{A}),

where NAN_{A} is the conductor of AA. See [Bis19, Corollary 6.12] and [Sab13, Proposition 1].

Finally, we construct coefficient systems associated to the fiber product of covers associated to HH moments and the rank double cover.

Example 8.1.11.

Continuing with notation as in 8.1.4, the coefficient systems H,g,frk,HT/2,1,g,f,HS,H,g,fH^{\operatorname{rk}}_{\mathscr{F},g,f},H_{T_{\mathbb{Z}/2\mathbb{Z},1,g,f}},H_{S_{\mathscr{F},H,g,f}} all correspond respectively to the finite covers of ConfXnAg,fn\operatorname{Conf}^{n}_{X^{\oplus n}\oplus A_{g,f}} for varying nn: RBn,Qg,fnR_{\mathscr{F}^{n}_{B}},Q^{n}_{g,f}, and HurS,H,g,fn\operatorname{Hur}_{S^{n}_{\mathscr{F},H,g,f}}. Define HurS,H,g,fnrk:=RBn×Qg,fnHurS,H,g,fn\operatorname{Hur}^{\operatorname{rk}}_{S^{n}_{\mathscr{F},H,g,f}}:=R_{\mathscr{F}^{n}_{B}}\times_{Q^{n}_{g,f}}\operatorname{Hur}_{S^{n}_{\mathscr{F},H,g,f}} and let HS,H,g,frkH^{\operatorname{rk}}_{S_{\mathscr{F},H,g,f}} be the corresponding coefficient system. Take V:=HS¯,H,0,0V:=H_{S_{\underline{\mathscr{F}},H,0,0}} and F:=HS,H,g,frkF:=H^{\operatorname{rk}}_{S_{\mathscr{F},H,g,f}}. Then, FF is a coefficient system over VV because HS,H,g,fH_{S_{\mathscr{F},H,g,f}} is a coefficient system over VV and both HT/2,{1},g,fH_{T_{\mathbb{Z}/2\mathbb{Z},\{1\},g,f}} and H,g,frkH^{\operatorname{rk}}_{\mathscr{F},g,f} are coefficient system over the trivial coefficient system, the latter by 8.1.8 and the former as explained in 8.1.4.

8.2. Homological stability of the rank double cover

We next set out to prove the main homological stability properties for the spaces related to Selmer groups we are interested in. Namely, in 8.2.3 we will prove these results for the Selmer stacks, the rank double cover, and moments associated to both of these.

Notation 8.2.1.

Let HH be a finite /ν\mathbb{Z}/\nu\mathbb{Z} module of the form Hi=1m/νiH\simeq\prod_{i=1}^{m}\mathbb{Z}/\nu_{i}\mathbb{Z}. Let \mathscr{F} be a lcc symplectically self-dual sheaf of free /ν\mathbb{Z}/\nu\mathbb{Z} modules, and maintain hypotheses as in 5.1.4 and 5.1.5. We then have Selmer spaces Sel[νi]Bn\operatorname{Sel}_{\mathscr{F}[\nu_{i}]^{n}_{B}} obtained from the νi\nu_{i}-torsion sheaves [νi]\mathscr{F}[\nu_{i}] in place of \mathscr{F}. Define

SelBnH:=Sel[ν1]Bn×QTwistU/BnSel[ν2]Bn×QTwistU/BnSel[νm]Bn.\displaystyle\operatorname{Sel}_{\mathscr{F}^{n}_{B}}^{H}:=\operatorname{Sel}_{\mathscr{F}[\nu_{1}]^{n}_{B}}\times_{\operatorname{QTwist}^{n}_{U/B}}\operatorname{Sel}_{\mathscr{F}[\nu_{2}]^{n}_{B}}\times_{\operatorname{QTwist}^{n}_{U/B}}\cdots\operatorname{Sel}_{\mathscr{F}[\nu_{m}]^{n}_{B}}.

Also define SelBnH,rk:=SelBnH×QTwistU/BnQTwistrk,n\operatorname{Sel}^{H,\operatorname{rk}}_{\mathscr{F}^{n}_{B}}:=\operatorname{Sel}_{\mathscr{F}^{n}_{B}}^{H}\times_{\operatorname{QTwist}^{n}_{U/B}}\operatorname{QTwist}^{\operatorname{rk},n}_{\mathscr{F}} and define HurBnH,rk:=HurBnH×QTwistU/BnQTwistrk,n\operatorname{Hur}^{H,\operatorname{rk}}_{\mathscr{F}^{n}_{B}}:=\operatorname{Hur}^{H}_{\mathscr{F}^{n}_{B}}\times_{\operatorname{QTwist}^{n}_{U/B}}\operatorname{QTwist}^{\operatorname{rk},n}_{\mathscr{F}}.

Lemma 8.2.2.

The hypotheses of 4.3.4 are satisfied if V=HS¯,H,0,0V=H_{S_{\underline{\mathscr{F}},H,0,0}} and FF is either HS,H,g,fH_{S_{\mathscr{F},H,g,f}} or F=HS,H,g,frkF=H^{\operatorname{rk}}_{S_{\mathscr{F},H,g,f}}.

Proof.

We consider two cases:

  1. (1)

    V=HS¯,H,0,0V=H_{S_{\underline{\mathscr{F}},H,0,0}} and F=HS,H,g,fF=H_{S_{\mathscr{F},H,g,f}},

  2. (2)

    V=HS¯,H,0,0V=H_{S_{\underline{\mathscr{F}},H,0,0}} and F=HS,H,g,frkF=H^{\operatorname{rk}}_{S_{\mathscr{F},H,g,f}}.

Note that by 8.1.3 and 8.1.11, VV and FF are indeed coefficient systems. We will first consider case (1)(1) and show the existence of a homogeneous central 𝕌\mathbb{U} in RVR^{V} of positive degree with kernel and cokernel of finite degree. Note that cHc_{H} does not generate GH=AHSp2r(/ν)G_{H}=\operatorname{\mathrm{A}^{\operatorname{H}}\mathrm{Sp}}_{2r}(\mathbb{Z}/\nu\mathbb{Z}) but instead generates the preimage of {±1}Sp2r(/ν)\{\pm 1\}\subset\mathrm{Sp}_{2r}(\mathbb{Z}/\nu\mathbb{Z}) in GHG_{H}. Let SHGHS_{H}\subset G_{H} denote the subgroup generated by cHc_{H}. Note that SHS_{H} has order 2mod42\bmod 4 because ν\nu is odd. Then, (SH,cH)(S_{H},c_{H}) is non-splitting in the sense of [EVW16, Definition 3.1] by [EVW16, Lemma 3.2]. It then follows from [EVW16, Lemma 3.5] that there is a homogeneous central 𝕌\mathbb{U} of positive degree with finite degree kernel and cokernel.

We can deduce case (2) from case (1). Namely, taking the same operator 𝕌\mathbb{U} as in part (1)(1), we can view F=HS,H,g,frkF=H^{\operatorname{rk}}_{S_{\mathscr{F},H,g,f}} as two copies of HS,H,g,fH_{S_{\mathscr{F},H,g,f}} Since we have already shown in the first case that the action of 𝕌\mathbb{U} on HS,H,g,fH_{S_{\mathscr{F},H,g,f}} has kernel and cokernel of finite degree, the same holds for the action of 𝕌\mathbb{U} on F=HS,H,g,frkF=H^{\operatorname{rk}}_{S_{\mathscr{F},H,g,f}}. ∎

Lemma 8.2.3.

Let HH be a finite /ν\mathbb{Z}/\nu\mathbb{Z} module and B=SpecB=\operatorname{Spec}\mathbb{C}. We work with coefficient systems over the field /\mathbb{Z}/\ell^{\prime}\mathbb{Z}, for \ell^{\prime} relatively prime to 2,q2,q, and #ASp2r(/ν)\#\operatorname{\mathrm{ASp}}_{2r}(\mathbb{Z}/\nu\mathbb{Z}). There is a constant KK depending on HH but not on nn, for nn even, so that

(8.5) dimHi(π1(ConfXnAg,fn,xn),HS,H,g,fn)\displaystyle\dim H^{i}(\pi_{1}(\operatorname{Conf}^{n}_{X^{\oplus n}\oplus A_{g,f}},x^{\oplus n}),H_{S^{n}_{\mathscr{F},H,g,f}}) <Ki+1 and\displaystyle<K^{i+1}\text{ and}
dimHi(π1(ConfXnAg,fn,xn),HS,H,g,fnrk)\displaystyle\dim H^{i}(\pi_{1}(\operatorname{Conf}^{n}_{X^{\oplus n}\oplus A_{g,f}},x^{\oplus n}),H^{\operatorname{rk}}_{S^{n}_{\mathscr{F},H,g,f}}) <Ki+1.\displaystyle<K^{i+1}.

Suppose \mathscr{F} is as in 6.4.4. Then,

(8.6) dimHi(SelnH,/)\displaystyle\dim H^{i}(\operatorname{Sel}_{\mathscr{F}^{n}_{\mathbb{C}}}^{H},\mathbb{Z}/\ell^{\prime}\mathbb{Z}) <Ki+1 and\displaystyle<K^{i+1}\text{ and }
dimHi(SelnH,rk,/)\displaystyle\dim H^{i}(\operatorname{Sel}^{H,\operatorname{rk}}_{\mathscr{F}^{n}_{\mathbb{C}}},\mathbb{Z}/\ell^{\prime}\mathbb{Z}) <Ki+1.\displaystyle<K^{i+1}.
Proof.

First, the bound (8.5) follows from 4.3.4 whose hypotheses are verified by 8.2.2.

For (8.6), note that in order to bound the homology of SelnH\operatorname{Sel}_{\mathscr{F}^{n}_{\mathbb{C}}}^{H}, by transfer and the assumption that 2\ell^{\prime}\neq 2, it suffices to bound the homology of its finite étale double cover SelnH,rk\operatorname{Sel}^{H,\operatorname{rk}}_{\mathscr{F}^{n}_{\mathbb{C}}}.

Recall that we use the notation HurS,H,g,fn\operatorname{Hur}_{S^{n}_{\mathscr{F},H,g,f}} and HurS,H,g,fnrk\operatorname{Hur}^{\operatorname{rk}}_{S^{n}_{\mathscr{F},H,g,f}} for the finite unramified covering space over ConfXnAg,fn\operatorname{Conf}^{n}_{X^{\oplus n}\oplus A_{g,f}} corresponding to the action of of π1(ConfXnAg,fn,xn)\pi_{1}(\operatorname{Conf}^{n}_{X^{\oplus n}\oplus A_{g,f}},x^{\oplus n}) on HS,H,g,fnH_{S^{n}_{\mathscr{F},H,g,f}} and π1(ConfXnAg,fn,xn)\pi_{1}(\operatorname{Conf}^{n}_{X^{\oplus n}\oplus A_{g,f}},x^{\oplus n}) on HS,H,g,fnrkH^{\operatorname{rk}}_{S^{n}_{\mathscr{F},H,g,f}}. It follows from these definitions that

Hi(HurS,H,g,fnrk,/)Hi(π1(ConfXnAg,fn,xn),HS,H,g,fnrk).\displaystyle H^{i}(\operatorname{Hur}^{\operatorname{rk}}_{S^{n}_{\mathscr{F},H,g,f}},\mathbb{Z}/\ell^{\prime}\mathbb{Z})\simeq H^{i}(\pi_{1}(\operatorname{Conf}^{n}_{X^{\oplus n}\oplus A_{g,f}},x^{\oplus n}),H^{\operatorname{\operatorname{rk}}}_{S^{n}_{\mathscr{F},H,g,f}}).

To conclude the final statement for bounding the homology of SelnH,rk\operatorname{Sel}^{H,\operatorname{rk}}_{\mathscr{F}^{n}_{\mathbb{C}}}, by transfer, it suffices to show HurS,H,g,fnrk\operatorname{Hur}^{\operatorname{rk}}_{S^{n}_{\mathscr{F},H,g,f}} defines a finite étale cover of SelnH,rk\operatorname{Sel}^{H,\operatorname{rk}}_{\mathscr{F}^{n}_{\mathbb{C}}}. We next use the isomorphism SelnHHurnH\operatorname{Sel}_{\mathscr{F}^{n}_{\mathbb{C}}}^{H}\to\operatorname{Hur}^{H}_{\mathscr{F}^{n}_{\mathbb{C}}} from 6.4.7 over QTwistU/Bn\operatorname{QTwist}^{n}_{U/B}, which also yields the identification SelnH,rkHurnH,rk\operatorname{Sel}^{H,\operatorname{rk}}_{\mathscr{F}^{n}_{\mathbb{C}}}\simeq\operatorname{Hur}^{H,\operatorname{rk}}_{\mathscr{F}^{n}_{\mathbb{C}}}. It therefore suffices to show HurnH,rk\operatorname{Hur}^{H,\operatorname{rk}}_{\mathscr{F}^{n}_{\mathbb{C}}} has a finite covering space by HurS,H,g,fnrk\operatorname{Hur}^{\operatorname{rk}}_{S^{n}_{\mathscr{F},H,g,f}}. There is an action of the group GHG_{H} as in 8.1.3 on the latter (acting via conjugation on HurS,H,g,fn\operatorname{Hur}_{S^{n}_{\mathscr{F},H,g,f}} and trivially on RnR_{\mathscr{F}^{n}_{\mathbb{C}}}). The quotient by this GHG_{H} action is precisely HurnH,rk\operatorname{Hur}^{H,\operatorname{rk}}_{\mathscr{F}^{n}_{\mathbb{C}}}, as follows from 2.4.6, since HurS,H,g,fn\operatorname{Hur}_{S^{n}_{\mathscr{F},H,g,f}} is the fiber product of the pointed Hurwitz space with the rank double cover, while HurnH,rk\operatorname{Hur}^{H,\operatorname{rk}}_{\mathscr{F}^{n}_{\mathbb{C}}} is the fiber product of the usual Hurwitz space with the rank double cover. ∎

Remark 8.2.4.

The Hurwitz stacks and Selmer stacks, whose cohomology we analyze in 8.2.3 have (up to finite index issues) an action of Modg,f\operatorname{Mod}_{g,f} the mapping class group of a genus gg, ff-punctured surface. Hence, their stable cohomology groups are virtual Modg,f\operatorname{Mod}_{g,f} representations. It would be extremely interesting to determine which representations these are. A precursor to doing so would be to compute the dimension of these representations. We also cannot rule out the possibility these dimensions are 0, and so the representations are not particularly interesting. See also 9.2.6

8.3. Relation between the rank double cover and parity of rank

Our main reason for introducing the rank double cover is that it tells us about the parity of the rank of Sel\operatorname{Sel}_{\ell}, as we next explain. For the next statement, recall the definition of 𝒩i\mathcal{N}^{i} from 7.4.1.

Lemma 8.3.1.

Assume ν\nu is odd, n>0n>0 is even, and BB is an integral affine scheme with 2ν2\nu invertible on BB. Let bBb\in B a closed point with residue field 𝔽q0\mathbb{F}_{q_{0}}. Let 𝔽q\mathbb{F}_{q} be a finite extension of 𝔽q0\mathbb{F}_{q_{0}}. Use hypotheses as in 5.1.4, 7.1.4, and 5.1.8, so bA[ν]\mathscr{F}_{b}\simeq A[\nu]. Let ν\ell\mid\nu and i:=rkVbn[]mod2{0,1}i:=\operatorname{rk}V_{\mathscr{F}^{n}_{b}[\ell]}\bmod 2\in\{0,1\}. Then, for xQTwistUb/bn(𝔽q)x\in\operatorname{QTwist}^{n}_{U_{b}/b}(\mathbb{F}_{q}), Selν(Ax)𝒩i\operatorname{Sel}_{\nu}(A_{x})\in\mathcal{N}^{i} if and only if xx lies in the image of QTwistbrk,n(𝔽q)QTwistUb/bn(𝔽q)\operatorname{QTwist}^{\operatorname{rk},n}_{\mathscr{F}_{b}}(\mathbb{F}_{q})\to\operatorname{QTwist}^{n}_{U_{b}/b}(\mathbb{F}_{q}).

Proof.

Let gx:=ρbn(Frobx)g_{x}:=\rho_{\mathscr{F}^{n}_{b}}(\operatorname{Frob}_{x}). For ν\ell\mid\nu, we use gx,g_{x,\ell} to denote the image of gxg_{x} under the map O(Qbn)O(Qbn[]){\rm{O}}(Q_{\mathscr{F}^{n}_{b}})\to{\rm{O}}(Q_{\mathscr{F}^{n}_{b}[\ell]}). First, (2.1) yields

dimker(gx,id)mod2rkVbn[]DQbn(gx,)mod2.\displaystyle\dim\ker(g_{x,\ell}-\operatorname{\mathrm{id}})\bmod 2\equiv\operatorname{rk}V_{\mathscr{F}^{n}_{b}[\ell]}-D_{Q_{\mathscr{F}^{n}_{b}}}(g_{x,\ell})\bmod 2.

Next, 5.3.2 gives ker(gxid)Selν(Ax)\ker(g_{x}-\operatorname{\mathrm{id}})\simeq\operatorname{Sel}_{\nu}(A_{x}). Combining these, we find

DQbn(gx,)rkVbn[]dimker(gx,id)rkVbn[]dimSel(Ax)mod2.\displaystyle D_{Q_{\mathscr{F}^{n}_{b}}}(g_{x,\ell})\equiv\operatorname{rk}V_{\mathscr{F}^{n}_{b}[\ell]}-\dim\ker(g_{x,\ell}-\operatorname{\mathrm{id}})\equiv\operatorname{rk}V_{\mathscr{F}^{n}_{b}[\ell]}-\dim\operatorname{Sel}_{\ell}(A_{x})\bmod 2.

Since this holds for every ν\ell\mid\nu, we find that DQbn(gx,)D_{Q_{\mathscr{F}^{n}_{b}}}(g_{x,\ell}) takes the value 0 if and only if rkVbn[]dimSel(Ax)mod2\operatorname{rk}V_{\mathscr{F}^{n}_{b}[\ell]}\equiv\dim\operatorname{Sel}_{\ell}(A_{x})\bmod 2. Since the finite étale double cover QTwistrk,nQTwistUb/bn\operatorname{QTwist}^{\operatorname{rk},n}_{\mathscr{F}}\to\operatorname{QTwist}^{n}_{U_{b}/b} is trivial over each 𝔽q\mathbb{F}_{q} point with trivial Dickson invariant, DQbn(gx,)D_{Q_{\mathscr{F}^{n}_{b}}}(g_{x,\ell}) takes the value 0 if and only if xQTwistUb/bn(𝔽q)x\in\operatorname{QTwist}^{n}_{U_{b}/b}(\mathbb{F}_{q}) is in the image of QTwistrk,n(𝔽q)\operatorname{QTwist}^{\operatorname{rk},n}_{\mathscr{F}}(\mathbb{F}_{q}). We conclude the result because rkVbn[]dimSel(Ax)\operatorname{rk}V_{\mathscr{F}^{n}_{b}[\ell]}\equiv\dim\operatorname{Sel}_{\ell}(A_{x}) can be restated as Selν(Ax)𝒩i\operatorname{Sel}_{\nu}(A_{x})\in\mathcal{N}^{i}, with i=rkVbn[]mod2i=\operatorname{rk}V_{\mathscr{F}^{n}_{b}[\ell]}\bmod 2. ∎

We now use the previous lemma to show that the distribution of Selmer elements on the double cover controlling the parity of the rank agrees with the locus of points on the base where the rank of Sel\operatorname{Sel}_{\ell} has a specified parity. This is a fairly trivial observation, but allows us to connect moments of the rank double cover to moments of the space of quadratic twists with specified parity of rank of Sel\operatorname{Sel}_{\ell}. This plays a key role in proving our main theorem, Theorem 1.1.2. For this, recall the definition of XA[ν]𝔽qniX^{i}_{A[\nu]^{n}_{\mathbb{F}_{q}}} from 7.4.1.

Lemma 8.3.2.

With assumptions and notation as in 8.3.1, so, in particular, i:=rkVbn[]mod2{0,1}i:=\operatorname{rk}V_{\mathscr{F}^{n}_{b}[\ell]}\bmod 2\in\{0,1\} for every ν\ell\mid\nu, we have

(8.7) #SelbnH,rk(𝔽q)#QTwistbrk,n(𝔽q)=𝔼(#Hom(XA[ν]𝔽qni,H)).\displaystyle\frac{\#\operatorname{Sel}^{H,\operatorname{rk}}_{\mathscr{F}^{n}_{b}}(\mathbb{F}_{q})}{\#\operatorname{QTwist}^{\operatorname{rk},n}_{\mathscr{F}_{b}}(\mathbb{F}_{q})}=\mathbb{E}(\#\mathrm{Hom}(X^{i}_{A[\nu]^{n}_{\mathbb{F}_{q}}},H)).
Proof.

Using 8.3.1, the distribution XA[ν]𝔽qniX^{i}_{A[\nu]^{n}_{\mathbb{F}_{q}}} agrees with the distribution of Selmer groups at points xQTwistUb/bn(𝔽q)x\in\operatorname{QTwist}^{n}_{U_{b}/b}(\mathbb{F}_{q}) in the image of QTwistbrk,n(𝔽q)QTwistUb/bn(𝔽q)\operatorname{QTwist}^{\operatorname{rk},n}_{\mathscr{F}_{b}}(\mathbb{F}_{q})\to\operatorname{QTwist}^{n}_{U_{b}/b}(\mathbb{F}_{q}). Since QTwistbrk,nQTwistUb/bn\operatorname{QTwist}^{\operatorname{rk},n}_{\mathscr{F}_{b}}\to\operatorname{QTwist}^{n}_{U_{b}/b} is a finite étale double cover, each 𝔽q\mathbb{F}_{q} point of QTwistUb/bn\operatorname{QTwist}^{n}_{U_{b}/b} in the image of a 𝔽q\mathbb{F}_{q} point of QTwistbrk,n\operatorname{QTwist}^{\operatorname{rk},n}_{\mathscr{F}_{b}} has exactly two 𝔽q\mathbb{F}_{q} points in its preimage. This means that, for yy varying over points of QTwistbrk,n(𝔽q)\operatorname{QTwist}^{\operatorname{rk},n}_{\mathscr{F}_{b}}(\mathbb{F}_{q}) and K𝒩K\in\mathcal{N} a finite /ν\mathbb{Z}/\nu\mathbb{Z} module,

Prob(XA[ν]𝔽qniK)\displaystyle\operatorname{Prob}(X^{i}_{A[\nu]^{n}_{\mathbb{F}_{q}}}\simeq K) =Prob(Selν(Ax)K|xim(QTwistbrk,n(𝔽q)QTwistUb/bn(𝔽q)))\displaystyle=\operatorname{Prob}\left(\operatorname{Sel}_{\nu}(A_{x})\simeq K|x\in\operatorname{im}(\operatorname{QTwist}^{\operatorname{rk},n}_{\mathscr{F}_{b}}(\mathbb{F}_{q})\to\operatorname{QTwist}^{n}_{U_{b}/b}(\mathbb{F}_{q}))\right)
=Prob(Selν(Ay)K).\displaystyle=\operatorname{Prob}(\operatorname{Sel}_{\nu}(A_{y})\simeq K).

Taking the expectation of the number of maps to HH, which is the same as the number of maps from HH, it is enough to show the left hand side of (8.7) is the expected number of maps from HH to Selν(Ay)\operatorname{Sel}_{\nu}(A_{y}). This follows from 5.3.2 and the definition of SelbnH,rk\operatorname{Sel}^{H,\operatorname{rk}}_{\mathscr{F}^{n}_{b}} as a fiber product. ∎

9. Computing the moments

The purpose of this section is to combine our homological stability results with our big monodromy results to determine the moments of Selmer groups in quadratic twist families. The analogous problem of determining the moments in the context of Cohen-Lenstra was approached in [EVW16], where the problem was much easier as the relevant big monodromy result was already available in the literature. In § 9.1, we compute various statistics associated to kernels of random elements of orthogonal groups. Via equidistribution of Frobenius elements we then relate this to components of Selmer stacks in § 9.2.

9.1. Moments related to random elements of orthogonal groups

We next compute statistics associated to random elements of orthogonal groups. In 9.1.5, we compute the distributions of 11-eigenspaces of random elements of orthogonal group, and show that these limit to the BKLPR distribution as the size of the matrix grows. Moreover, we show this in a strong enough sense so that the limit of the moments is the moment of the limit.

Our next computation is quite analogous to that of [FLR23, Proposition 4.13], except that here we work over /ν\mathbb{Z}/\nu\mathbb{Z} for general ν\nu, instead of the case that ν\nu is prime covered in [FLR23].

For what follows, we use the notation of [FLR23, §4.2.1]. In the case ν\nu is prime, we let A,B,CA,B,C be the three nontrivial cosets of Ω(Q)\Omega(Q) in O(Q){\rm{O}}(Q) so that spQ\operatorname{sp}^{-}_{Q} is nontrivial on AA and CC, while DQD_{Q} is nontrivial on BB and CC. For ZZ a nonnegative integer-valued random variable, we let GZ(t)=iProb(dimZ=i)tiG_{Z}(t)=\sum_{i\in\mathbb{N}}\operatorname{Prob}(\dim Z=i)t^{i}. As in [FLR23, §4.2.1], for {Ω,A,B,C}\bullet\in\{\Omega,A,B,C\}, we use RSelV\operatorname{RSel}_{V}^{\bullet} to denote the random variable given as ker(gid)\ker(g-\operatorname{\mathrm{id}}) for gg a uniform random element of the coset \bullet.

Lemma 9.1.1.

Let (Q,V)(Q,V) be a quadratic space over /\mathbb{Z}/\ell\mathbb{Z}, with \ell an odd prime. When dimV=2s\dim V=2s is even,

GRSelVB\displaystyle G_{\operatorname{RSel}^{B}_{V}} =GRSelVC,\displaystyle=G_{\operatorname{RSel}^{C}_{V}},
GRSelVΩ\displaystyle G_{\operatorname{RSel}^{\Omega}_{V}} =GRSelVA+1#Ω(Q)i=0s1(t22i).\displaystyle=G_{\operatorname{RSel}^{A}_{V}}+\frac{1}{\#\Omega(Q)}\prod_{i=0}^{s-1}(t^{2}-\ell^{2i}).

For a𝔽×a\in\mathbb{F}_{\ell}^{\times}, let sgn(a)\operatorname{sgn}(a) denote 11 if aa is a square mod\bmod\ell and 1-1 otherwise. When dimV=2s+1\dim V=2s+1 is odd,

GRSelVB\displaystyle G_{\operatorname{RSel}^{B}_{V}} =GRSelVC+2sgn(1)s#Ω(Q)i=1s1(t22i),\displaystyle=G_{\operatorname{RSel}^{C}_{V}}+\frac{2\operatorname{sgn}(-1)\ell^{s}}{\#\Omega(Q)}\prod_{i=1}^{s-1}(t^{2}-\ell^{2i}),
GRSelVΩ\displaystyle G_{\operatorname{RSel}^{\Omega}_{V}} =GRSelVA+t#Ω(Q)i=0s1(t22i).\displaystyle=G_{\operatorname{RSel}^{A}_{V}}+\frac{t}{\#\Omega(Q)}\prod_{i=0}^{s-1}(t^{2}-\ell^{2i}).
Proof.

For the proof when dimV=2s\dim V=2s, note that [FLR23, Lemma 4.7] easily generalizes to show that for any coset HH of Ω(Q)\Omega(Q) in O(Q){\rm{O}}(Q), GRSelVH(i)=GRSelVΩ(i)G_{\operatorname{RSel}^{H}_{V}}(\ell^{i})=G_{\operatorname{RSel}^{\Omega}_{V}}(\ell^{i}) whenever 2i+2dimV2i+2\leq\dim V. When dimV\dim V is even, the proof proceeds mutatis mutandis as in [FLR23, Theorem 4.4].

Therefore, it remains to prove the case that dimV=2s+1\dim V=2s+1 is odd. We again proceed following the proof strategy of [FLR23, Theorem 4.4]. By 2.1.3, only even powers of tt can appear in GRSelVB(t)G_{\operatorname{RSel}^{B}_{V}}(t) and GRSelVC(t)G_{\operatorname{RSel}^{C}_{V}}(t). These are therefore even polynomials of degree at most dimV\dim V and agree at the dimV1\dim V-1 values ±1,±,,±dimV32\pm 1,\pm\ell,\ldots,\pm\ell^{\frac{\dim V-3}{2}} by [FLR23, Lemma 4.5]. Since dimV\dim V is odd and the polynomials are even, the polynomials in fact have degree at most dimV1\dim V-1, and hence are determined up to a scalar. That is, GRSelVB(t)GRSelVC(t)G_{\operatorname{RSel}^{B}_{V}}(t)-G_{\operatorname{RSel}^{C}_{V}}(t) is a scalar multiple of i=1dimV32(t22i)\prod_{i=1}^{\frac{\dim V-3}{2}}(t^{2}-\ell^{2i}). To pin that scalar multiple down, we can examine the coefficient of tdimV1t^{\dim V-1} in GRSelV(t)G_{\operatorname{RSel}^{\bullet}_{V}}(t), for {B,C}\bullet\in\{B,C\}. This coefficient is #R(Q)#Ω(Q)\frac{\#R_{\bullet}(Q)}{\#\Omega(Q)}, where R(Q)R_{\bullet}(Q) is the set of reflections in \bullet, since any non-identity element of the orthogonal group fixing a codimension 11 plane is a reflection. Since there are 2s+qs\ell^{2s}+q^{s} reflections with value α\alpha for any square α𝔽×\alpha\in\mathbb{F}_{\ell}^{\times}, and 2ss\ell^{2s}-\ell^{s} reflections with value β\beta for any for any nonsquare β𝔽×\beta\in\mathbb{F}_{\ell}^{\times}, the definition of spQ\operatorname{sp}^{-}_{Q} yields that

GRSelVBGRSelVC=2sgn(1)s#Ω(Q)i=1dimV32(t22i)=2sgn(1)s#Ω(Q)i=1s1(t22i).\displaystyle G_{\operatorname{RSel}^{B}_{V}}-G_{\operatorname{RSel}^{C}_{V}}=\frac{2\operatorname{sgn}(-1)\ell^{s}}{\#\Omega(Q)}\prod_{i=1}^{\frac{\dim V-3}{2}}(t^{2}-\ell^{2i})=\frac{2\operatorname{sgn}(-1)\ell^{s}}{\#\Omega(Q)}\prod_{i=1}^{s-1}(t^{2}-\ell^{2i}).

Finally, the remaining two cosets satisfy the relation GRSelVΩ=GRSelVA+1#Ω(Q)i=0s1(t22i)G_{\operatorname{RSel}^{\Omega}_{V}}=G_{\operatorname{RSel}^{A}_{V}}+\frac{1}{\#\Omega(Q)}\prod_{i=0}^{s-1}(t^{2}-\ell^{2i}) by an argument analogous to the last paragraph of the proof of [FLR23, Theorem 4.4]: Indeed, GRSelVB(t)G_{\operatorname{RSel}^{B}_{V}}(t) and GRSelVC(t)G_{\operatorname{RSel}^{C}_{V}}(t) are two odd degree dimV\dim V polynomials agreeing on the dimV\dim V values 0,±1,±,,±dimV320,\pm 1,\pm\ell,\ldots,\pm\ell^{\frac{\dim V-3}{2}}, so their difference is divisible by ti=1dimV32(t22i)t\prod_{i=1}^{\frac{\dim V-3}{2}}(t^{2}-\ell^{2i}), and the constant of proportionality can be determined using that the identity is the only element with a dimV\dim V dimensional fixed space. ∎

We next define a notion of mm-total variation distance, which will be useful for proving moments of two distributions converge, see 9.1.4.

Definition 9.1.2.

Let 𝒩\mathcal{N} denote the set of isomorphism classes of finite /ν\mathbb{Z}/\nu\mathbb{Z} modules. Let X,YX,Y be two 𝒩\mathcal{N} valued random variables. For m0m\in\mathbb{Z}_{\geq 0}, we define the mm-total variation distance or dTVm(X,Y)d^{m}_{\operatorname{TV}}(X,Y)

dTVm(X,Y):=H𝒩(#H)m|Prob(X=H)Prob(Y=H)|.\displaystyle d^{m}_{\operatorname{TV}}(X,Y):=\sum_{H\in\mathcal{N}}(\#H)^{m}\left|\operatorname{Prob}(X=H)-\operatorname{Prob}(Y=H)\right|.
Remark 9.1.3.

When m=0m=0, and the random variable is real valued instead of valued in 𝒩\mathcal{N}, this is twice the usual notion of total variation distance, see [LPW09, §4.1 and Proposition 4.2]. We claim that a sequence of random variables (Xn)n0(X_{n})_{n\geq 0} converges to YY in distribution if the total variation distance between XnX_{n} and YY tends to 0 in nn: Indeed, convergence in distribution simply means pointwise convergence for distributions on a discrete probability space.

Remark 9.1.4.

The point of the definition of mm-total variation distance is that if a sequence of random variables XnX_{n} converges to YY in mm-total variation distance then the mmth moment of XnX_{n} converges to the mmth moment of YY. This follows directly from the definition of mm-total variation distance.

With the above definition in hand, we are prepared to show the distribution of 11-eigenspaces of random orthogonal group matrices converges in a strong sense to the BKLPR distribution, as the size of the matrix grows.

Lemma 9.1.5.

Let (Vν,n,Qν,n)n>0(V_{\nu,n},Q_{\nu,n})_{n\in\mathbb{Z}_{>0}} be a sequence of nondegenerate quadratic spaces over /ν\mathbb{Z}/\nu\mathbb{Z}, for ν\nu odd. Suppose rkVν,nn\operatorname{rk}V_{\nu,n}\geq n.

  1. (1)

    Suppose Gν,nO(Qν,n)G_{\nu,n}\subset{\rm{O}}(Q_{\nu,n}) is a subgroup containing Ω(Qν,n)\Omega(Q_{\nu,n}) and not contained in SO(Qν,n)\operatorname{SO}(Q_{\nu,n}). Let Rν,nR_{\nu,n} denote the distribution of ker(gid)\ker(g-\operatorname{\mathrm{id}}) for gGν,ng\in G_{\nu,n} a uniform random element.

    For any m0m\in\mathbb{Z}_{\geq 0}, the limit limnRν,n\lim_{n\to\infty}R_{\nu,n} converges in mm-total variation distance to a distribution which agrees with SelνBKLPR\operatorname{Sel}^{\operatorname{BKLPR}}_{\nu}.

  2. (2)

    Suppose Gν,nSO(Qν,n)G_{\nu,n}\subset\operatorname{SO}(Q_{\nu,n}) is a subgroup containing Ω(Qν,n)\Omega(Q_{\nu,n}). Let Rν,nrkR_{\nu,n}^{\operatorname{rk}} denote the distribution of ker(gid)\ker(g-\operatorname{\mathrm{id}}) for gGν,ng\in G_{\nu,n} a uniform random element.

    For any m0m\in\mathbb{Z}_{\geq 0}, the limit limnRν,nrk\lim_{n\to\infty}R_{\nu,n}^{\operatorname{rk}} converges in mm-total variation distance to a distribution which agrees with SelνBKLPR,rkVmod2\operatorname{Sel}^{\operatorname{BKLPR},\operatorname{rk}V\bmod 2}_{\nu}.

Proof sketch.

We start by verifying (1)(1). The argument closely follows [FLR23, Theorem 6.4]. We now provide some more details on the changes one must make.

We first claim the result holds when ν=\nu=\ell is an odd prime. For sνs\mid\nu, we use Qs,nQ_{s,n} and Rs,nR_{s,n} for the reduction mod ss of Qν,nQ_{\nu,n} and Rν,nR_{\nu,n}. As an initial step in our argument, we next verify in 9.1.6 that when ν=\nu=\ell is prime, limndTVm(R,n,SelBKLPR)((n/2)2ε)\lim_{n\to\infty}d_{\operatorname{TV}}^{m}(R_{\ell,n},\operatorname{Sel}^{\operatorname{BKLPR}}_{\ell})\ll\ell^{-((n/2)^{2}-\varepsilon)}.

Lemma 9.1.6.

With notation as in 9.1.5, for \ell an odd prime,

limndTVm(R,n,SelBKLPR)((n/2)2ε).\displaystyle\lim_{n\to\infty}d_{\operatorname{TV}}^{m}(R_{\ell,n},\operatorname{Sel}^{\operatorname{BKLPR}}_{\ell})\ll\ell^{-((n/2)^{2}-\varepsilon)}.
Proof.

For GG a finite group, we use RGR^{G} to denote the distribution the dimension of the 11-eigenspace of a uniformly random element of GG. We can first bound dTVm(R,n,ROdimV)d^{m}_{\operatorname{TV}}(R_{\ell,n},R^{{\rm{O}}_{\dim V}}), where we use OdimV{\rm{O}}_{\dim V} to denote the orthogonal group over a finite field of dimension dimV\dim V, which Gν,nG_{\nu,n} is a subset of. Note, by convention, ndimVn\leq\dim V. The proof of this bound on mm-total variation distance is quite similar to that of [FLR23, Theorem 4.23], except that we replace the input of [FLR23, Theorem 4.4] with that of 9.1.1, and note that since these probability distributions are both supported on {0,,dimV}\{0,\ldots,\dim V\}, dTVm(R,n,ROdimV)(dimV)mdTV0(R,n,ROdimV)d^{m}_{\operatorname{TV}}(R_{\ell,n},R^{{\rm{O}}_{\dim V}})\leq(\dim V)^{m}\cdot d^{0}_{\operatorname{TV}}(R_{\ell,n},R^{{\rm{O}}_{\dim V}}). Now, dTV0(R,n,ROdimV)d^{0}_{\operatorname{TV}}(R_{\ell,n},R^{{\rm{O}}_{\dim V}}) was shown to be (dimV2)2\ll\ell^{-(\frac{\dim V}{2})^{2}} in [FLR23, Theorem 4.23] when dimV\dim V is even dimensional with discriminant 11, and, as mentioned, an analogous proof applies here. We conclude that dTVm(R,n,ROdimV)((n/2)2ε)d^{m}_{\operatorname{TV}}(R_{\ell,n},R^{{\rm{O}}_{\dim V}})\ll\ell^{-((n/2)^{2}-\varepsilon)}.

Hence, to show dTVm(R,n,SelBKLPR)((n/2)2ε)d^{m}_{\operatorname{TV}}(R_{\ell,n},\operatorname{Sel}^{\operatorname{BKLPR}}_{\ell})\ll\ell^{-((n/2)^{2}-\varepsilon)}, it suffices to bound dTVm(ROdimV,SelνBKLPR)(dimV/2)2d^{m}_{\operatorname{TV}}(R^{{\rm{O}}_{\dim V}},\operatorname{Sel}^{\operatorname{BKLPR}}_{\nu})\ll\ell^{-(\lfloor\dim V/2\rfloor)^{2}}. To this end, let 2s2s denote the smallest even integer with 2sdimV2s\leq\dim V. We use O+(2s,){\rm{O}}^{+}(2s,\ell) to denote the discriminant 11 orthogonal group over 𝔽\mathbb{F}_{\ell} of rank 2s2s. The formulas in [FS16, Theorem 2.7 and 2.9], which give the dimension of fixed spaces of elements of orthogonal groups, show

(9.1) dTVm(ROdimV,RO+(2s,))\displaystyle d^{m}_{\operatorname{TV}}(R^{{\rm{O}}_{\dim V}},R^{{\rm{O}}^{+}(2s,\ell)})
(9.2) k=0s(2k)m(2ks+s2k2+(sk))+k=0s(2k+1)m(2ks+s2k2+(sk))\displaystyle\leq\sum_{k=0}^{s}(2k)^{m}\ell^{-(2ks+s^{2}-k^{2}+(s-k))}+\sum_{k=0}^{s}(2k+1)^{m}\ell^{-(2ks+s^{2}-k^{2}+(s-k))}
(9.3) s2.\displaystyle\ll\ell^{-s^{2}}.

The first sum in (9.2) is accounted for by the second line of [FS16, Theorem 2.9(1)] (and this is the only one that appears in the case dimV\dim V is even) and the second sum is accounted for by the i=nki=n-k term in the sum appearing in [FS16, Theorem 2.7(2)].

To conclude the bound limndTVm(R,n,SelBKLPR)((n/2)2ε)\lim_{n\to\infty}d_{\operatorname{TV}}^{m}(R_{\ell,n},\operatorname{Sel}^{\operatorname{BKLPR}}_{\ell})\ll\ell^{-((n/2)^{2}-\varepsilon)}, it remains to bound

dTVm(RO+(2s,),SelBKLPR)s2.\displaystyle d^{m}_{\operatorname{TV}}(R^{{\rm{O}}^{+}(2s,\ell)},\operatorname{Sel}^{\operatorname{BKLPR}}_{\ell})\ll\ell^{-s^{2}}.

This was essentially done in the last paragraph of the proof of [FLR23, Theorem 4.23] combined with [FLR23, Corollary 4.24], and we now give a slightly more direct argument. First, dTVm(RO+(2s,),RO+(2s+2,))s2d^{m}_{\operatorname{TV}}(R^{{\rm{O}}^{+}(2s,\ell)},R^{{\rm{O}}^{+}(2s+2,\ell)})\ll\ell^{-s^{2}}, using the formulas in [FS16, Theorem 2.9], similarly to the preceding paragraph. This implies that dTVm(RO+(2s,),limsRO+(2s,))s2d^{m}_{\operatorname{TV}}(R^{{\rm{O}}^{+}(2s,\ell)},\lim_{s\to\infty}R^{{\rm{O}}^{+}(2s,\ell)})\ll\ell^{-s^{2}}. An explicit formula for this limiting distribution is given in [FS16, Theorem 2.9(3)]. Note that in the case where \ell is prime, which we are currently considering, the “BKLPR heuristic” first appeared as the “Poonen-Rains heuristic” [PR12], whose explicit formula is given by [PR12, Conjecture 1.1(a)]. By inspection, this agrees with the distribution appearing in [FS16, Theorem 2.9(3)], yielding our claim that dTVm(R,n,SelBKLPR)((n/2)2ε)d^{m}_{\operatorname{TV}}(R_{\ell,n},\operatorname{Sel}^{\operatorname{BKLPR}}_{\ell})\ll\ell^{-((n/2)^{2}-\varepsilon)}. ∎

Proceeding with the proof of 9.1.5, we next explain why the Markov properties established in [FLR23, Theorem 5.1 and Theorem 5.13] for the Rj,nR_{\ell^{j},n} and SeljBKLPR\operatorname{Sel}^{\operatorname{BKLPR}}_{\ell^{j}} imply that we also obtain convergence in mm-total variation distance limnRj,nSeljBKLPR\lim_{n\to\infty}R_{\ell^{j},n}\to\operatorname{Sel}^{\operatorname{BKLPR}}_{\ell^{j}}. Technically, [FLR23, Theorem 5.1] is only stated in the case the quadratic space has even rank. However, the proof for \ell odd does not use the assumption that the rank is even. Although the BKLPR distribution only varies over even dimensional vector spaces, we have showed above that limnR,n=SelBKLPR\lim_{n\to\infty}R_{\ell,n}=\operatorname{Sel}^{\operatorname{BKLPR}}_{\ell}. Since both distributions satisfy the same Markov property relating the modj\mod\ell^{j} and the modj1\mod\ell^{j-1} versions, the mm-total variation distance also tends to 0 between the modj\bmod\ell^{j} distributions, and so limnRj,nSeljBKLPR\lim_{n\to\infty}R_{\ell^{j},n}\to\operatorname{Sel}^{\operatorname{BKLPR}}_{\ell^{j}} in mm-total variation distance.

To obtain the case of general ν\nu, write ν=a\nu=\prod_{\ell}\ell^{a_{\ell}}. The various distributions SelaBKLPR\operatorname{Sel}^{\operatorname{BKLPR}}_{\ell^{a_{\ell}}} are not in general independent, but they are independent after conditioning on the parity of the rank of their reduction mod\bmod\ell. Similarly, the distributions Ra,nR_{\ell^{a_{\ell}},n} are not independent, but they are independent after conditioning on the value of the coset of Ω(Qν,n)\Omega(Q_{\nu,n}) in Gν,nG_{\nu,n}, as Ω(Qν,n)=prime νΩ(Qa,n)\Omega(Q_{\nu,n})=\prod_{\text{prime }\ell\mid\nu}\Omega(Q_{\ell^{a_{\ell}},n}). We therefore obtain that the distribution of any specified coset of Ω(Qν,n)\Omega(Q_{\nu,n}) with specified value of DQν,nD_{Q_{\nu,n}} approaches the distribution SelaBKLPR\operatorname{Sel}^{\operatorname{BKLPR}}_{\ell^{a_{\ell}}}, conditioned on the parity of the rank as nn\to\infty, in mm-total variation distance. Summing over different cosets on both sides gives the claimed convergence in mm-total variation distance limnRν,nSelνBKLPR\lim_{n\to\infty}R_{\nu,n}\to\operatorname{Sel}^{\operatorname{BKLPR}}_{\nu}.

To conclude, it remains to deal with (2)(2). This is completely analogous to the proof of (1)(1), but where one compares distributions to random kernels of special orthogonal groups at each step. The distribution of dimker(gid)\dim\ker(g-\operatorname{\mathrm{id}}) for gSO(Q)g\in\operatorname{SO}(Q), for (V,Q)(V,Q) over 𝔽\mathbb{F}_{\ell} can be deduced from the distribution over gO(Q)g\in{\rm{O}}(Q) using 2.1.3. Namely, 2.1.3 shows that dimker(gid)dimVmod2\dim\ker(g-\operatorname{\mathrm{id}})\equiv\dim V\bmod 2 for gSO(Q)g\in\operatorname{SO}(Q). Since of O(Q){\rm{O}}(Q) elements are equally likely to lie in SO(Q)\operatorname{SO}(Q) and O(Q)SO(Q){\rm{O}}(Q)-\operatorname{SO}(Q), we find

Prob(dimker(gid)=s|gO(Q))=12Prob(dimker(gid)=s|gSO(Q))\displaystyle\operatorname{Prob}(\dim\ker(g-\operatorname{\mathrm{id}})=s|g\in{\rm{O}}(Q))=\frac{1}{2}\operatorname{Prob}(\dim\ker(g-\operatorname{\mathrm{id}})=s|g\in\operatorname{SO}(Q))

when sdimVmod2.s\equiv\dim V\bmod 2.

One can then obtain analogous asymptotic bounds on dTVm(R,nrk,SelBKLPR,dimVmod2)d_{\operatorname{TV}}^{m}(R^{\operatorname{rk}}_{\ell,n},\operatorname{Sel}^{\operatorname{BKLPR},\dim V\bmod 2}_{\ell}) to those proven in 9.1.6, using these explicit formulas. Next one can use the Markov property to obtain analogous bounds on dTVm(Rj,nrk,SeljBKLPR,dimVmod2)d_{\operatorname{TV}}^{m}(R^{\operatorname{rk}}_{\ell^{j},n},\operatorname{Sel}^{\operatorname{BKLPR},\dim V\bmod 2}_{\ell^{j}}). Finally, one can use the Chinese remainder theorem to obtain analogous bounds on dTVm(Rν,nrk,SelνBKLPR,dimVmod2)d_{\operatorname{TV}}^{m}(R^{\operatorname{rk}}_{\nu,n},\operatorname{Sel}^{\operatorname{BKLPR},\dim V\bmod 2}_{\nu}). ∎

9.2. Connected components of Selmer stacks

We are now ready to prove the key input to a “qq\to\infty first, then nn\to\infty“ version of our main result, which amounts to counting connected components of Selmer stacks.

In 9.2.1, we combine the above to compute the number of components of Selmer stacks. To compute this number of connected components, we will combine our big monodromy result from 7.3.3 with the convergence result of 9.1.5 to deduce that the number of components agrees with moments of the BKLPR distribution. Following this, in Theorem 9.2.1 we combine the above with our main homological stability theorem to compute the moments of Selmer groups in quadratic twist families.

Proposition 9.2.1.

Maintain hypotheses as in 5.1.4, 7.1.4, 8.2.1, and 5.1.8, so that bA[ν]\mathscr{F}_{b}\simeq A[\nu]. Take b=Spec𝔽qBb=\operatorname{Spec}\mathbb{F}_{q}\in B a closed point, and suppose the bound on nn from (7.1) is satisfied.

  1. (1)

    Every connected component of SelbnH\operatorname{Sel}_{\mathscr{F}^{n}_{b}}^{H} is geometrically connected and the number of such connected components is equal to 𝔼(#Hom(SelνBKLPR,H))\mathbb{E}(\#\mathrm{Hom}(\operatorname{Sel}^{\operatorname{BKLPR}}_{\nu},H)) for nn sufficiently large, depending on HH.

  2. (2)

    Every connected component of SelbnH,rk\operatorname{Sel}^{H,\operatorname{rk}}_{\mathscr{F}^{n}_{b}} is geometrically connected and the number of such connected components is equal to 𝔼(#Hom(SelνBKLPR,rkVbnmod2,H))\mathbb{E}(\#\mathrm{Hom}(\operatorname{Sel}^{\operatorname{BKLPR},\operatorname{rk}V_{\mathscr{F}^{n}_{b}}\bmod 2}_{\nu},H)) for nn sufficiently large, depending on HH.

Remark 9.2.2.

There has been much recent work, notably [LST20] and [SW23], studying versions of the Cohen-Lenstra heuristics in the presence of roots of unity. When working over function fields, the difference in behavior of the Cohen-Lenstra heuristics when the base field has certain roots of unity, can be traced back to a certain moduli space whose connected components are not all geometrically connected. However, in the context of the BKLPR heuristics, 9.2.1 shows the connected components are always geometrically connected. This explains why the BKLPR heuristics are not sensitive to roots of unity in the base field.

Remark 9.2.3.

We note that 9.2.1 is quite closely related to the main results of [PW23]. Although it is not exactly stated in this language, it follows from the Lang-Weil bounds that they prove a version of 9.2.1 in the special case that HH is of the form /\mathbb{Z}/\ell\mathbb{Z} for 5\ell\geq 5 a prime, and AA an elliptic curve. Both of our proofs follow a similar approach, and their proof is essentially a special case of ours.

Proof.

As a first step, note that the monodromy representation DQbnρbnD_{Q_{\mathscr{F}^{n}_{b}}}\circ\rho_{\mathscr{F}^{n}_{b}} surjects onto the diagonal copy of /2\mathbb{Z}/2\mathbb{Z} by 7.4.6. We first deal with case (1)(1). Let b¯\overline{b} denote a geometric point over bb. Take Gν,nG_{\nu,n} to be the arithmetic monodromy group at bb, imρbn\operatorname{im}\rho_{\mathscr{F}^{n}_{b}}.

This is a union of cosets of the geometric monodromy imρb¯n\operatorname{im}\rho_{\mathscr{F}^{n}_{\overline{b}}} in the orthogonal group, so is not contained in the special orthogonal group by 7.3.3, as we are assuming nn satisfies the bound of (7.1). Therefore, Gν,nG_{\nu,n} satisfies the hypotheses of 9.1.5(1). Let Rν,nR_{\nu,n} denote the distribution of ker(gid)\ker(g-\operatorname{\mathrm{id}}) for gGν,ng\in G_{\nu,n} a uniform random element. In what follows, we will show 𝔼(#Hom(Rν,n,H))\mathbb{E}(\#\mathrm{Hom}(R_{\nu,n},H)) agrees with the number of connected components of SelbnH\operatorname{Sel}_{\mathscr{F}^{n}_{b}}^{H}. Granting this, and using 9.1.5, which shows that the Rν,nR_{\nu,n} converge in mm-total variation distance to SelνBKLPR\operatorname{Sel}^{\operatorname{BKLPR}}_{\nu}, we find limn𝔼(#Hom(Rν,n,H))\lim_{n\to\infty}\mathbb{E}(\#\mathrm{Hom}(R_{\nu,n},H)) converges to 𝔼(#Hom(SelνBKLPR,H))\mathbb{E}(\#\mathrm{Hom}(\operatorname{Sel}^{\operatorname{BKLPR}}_{\nu},H)), whenever HH is a free /ν\mathbb{Z}/\nu\mathbb{Z} module of rank mm.

Having shown the desired convergence for free HH, we claim that the general case that HH is a /ν\mathbb{Z}/\nu\mathbb{Z} module with mm generators follows from the case that HH is a free module with mm generators. Indeed, it suffices to show the postulation that homomorphisms to such HH form a subset of homomorphisms to (/ν)m(\mathbb{Z}/\nu\mathbb{Z})^{m}. For this choose an injection H(/ν)mH\to\left(\mathbb{Z}/\nu\mathbb{Z}\right)^{m}. For any finite group KK, Hom(K,H)Hom(K,(/ν)m)\mathrm{Hom}(K,H)\hookrightarrow\mathrm{Hom}(K,\left(\mathbb{Z}/\nu\mathbb{Z}\right)^{m}) is injective. Hence we obtain the postulation, and therefore the claim.

It remains to show 𝔼(#Hom(Rν,n,H))\mathbb{E}(\#\mathrm{Hom}(R_{\nu,n},H)) agrees with the number of connected components of SelbnH\operatorname{Sel}_{\mathscr{F}^{n}_{b}}^{H}, all of which are geometrically connected. This follows from a standard monodromy argument and Burnside’s lemma, as we now explain. The action of Gν,nG_{\nu,n} on VbnV_{\mathscr{F}^{n}_{b}} is via the standard representation of the orthogonal group on its underlying vector space. Let H=i=1m/νiH=\prod_{i=1}^{m}\mathbb{Z}/\nu_{i}\mathbb{Z}. Then, the action ϕbn,H:Gν,nAut(i=1mV[νi]bn)\phi_{\mathscr{F}^{n}_{b},H}:G_{\nu,n}\to\operatorname{Aut}\left(\prod_{i=1}^{m}V_{\mathscr{F}[\nu_{i}]^{n}_{b}}\right) is via the diagonal action of the orthogonal group on i=1mV[νi]bn\prod_{i=1}^{m}V_{\mathscr{F}[\nu_{i}]^{n}_{b}}: ϕbn,H(g)(v1,,vm)=(gv1,,gvm)\phi_{\mathscr{F}^{n}_{b},H}(g)(v_{1},\ldots,v_{m})=(gv_{1},\ldots,gv_{m}), where gGν,ng\in G_{\nu,n}, viV[νi]bnv_{i}\in V_{\mathscr{F}[\nu_{i}]^{n}_{b}}, and gvigv_{i} denotes the standard action of an element of an orthogonal group on its underlying free module. Hence, the number of connected components of SelbnH\operatorname{Sel}_{\mathscr{F}^{n}_{b}}^{H} is equal to the number of orbits of Gν,nG_{\nu,n} on i=1mV[νi]bn\prod_{i=1}^{m}V_{\mathscr{F}[\nu_{i}]^{n}_{b}} under the above diagonal action ϕbn,H\phi_{\mathscr{F}^{n}_{b},H}. Now, using Burnside’s lemma, this number of orbits is equal to 1#Gν,ngGν,n#ker(ϕbn,H(g)id)\frac{1}{\#G_{\nu,n}}\sum_{g\in G_{\nu,n}}\#\ker(\phi_{\mathscr{F}^{n}_{b},H}(g)-\operatorname{\mathrm{id}}). Noting that an element in ker(ϕbn,H(g)id)\ker(\phi_{\mathscr{F}^{n}_{b},H}(g)-\operatorname{\mathrm{id}}) is a tuple (v1,,vm)(v_{1},\ldots,v_{m}) so that gvi=vigv_{i}=v_{i} and νivi=0\nu_{i}v_{i}=0, we can identify ker(ϕbn,H(g)id)Hom(H,kerϕbn,/ν(g)id)\ker(\phi_{\mathscr{F}^{n}_{b},H}(g)-\operatorname{\mathrm{id}})\simeq\mathrm{Hom}(H,\ker\phi_{\mathscr{F}^{n}_{b},\mathbb{Z}/\nu\mathbb{Z}}(g)-\operatorname{\mathrm{id}}). Hence,

1#Gν,ngGν,n#ker(ϕbn,H(g)id)\displaystyle\frac{1}{\#G_{\nu,n}}\sum_{g\in G_{\nu,n}}\#\ker(\phi_{\mathscr{F}^{n}_{b},H}(g)-\operatorname{\mathrm{id}}) =1#Gν,ngGν,n#Hom(H,kerϕbn,/ν(g)id)\displaystyle=\frac{1}{\#G_{\nu,n}}\sum_{g\in G_{\nu,n}}\#\mathrm{Hom}(H,\ker\phi_{\mathscr{F}^{n}_{b},\mathbb{Z}/\nu\mathbb{Z}}(g)-\operatorname{\mathrm{id}})
=1#Gν,ngGν,n#Hom(kerϕbn,/ν(g)id,H)\displaystyle=\frac{1}{\#G_{\nu,n}}\sum_{g\in G_{\nu,n}}\#\mathrm{Hom}(\ker\phi_{\mathscr{F}^{n}_{b},\mathbb{Z}/\nu\mathbb{Z}}(g)-\operatorname{\mathrm{id}},H)
=𝔼(#Hom(Rν,n,H)).\displaystyle=\mathbb{E}(\#\mathrm{Hom}(R_{\nu,n},H)).

The same argument as above goes through if one replaces G,nG_{\ell,n} with the geometric monodromy group. This shows the number of components over 𝔽¯q\overline{\mathbb{F}}_{q} is also 𝔼(#Hom(SelνBKLPR,H))\mathbb{E}(\#\mathrm{Hom}(\operatorname{Sel}^{\operatorname{BKLPR}}_{\nu},H)) for nn sufficiently large, and so the number of components over 𝔽¯q\overline{\mathbb{F}}_{q} agrees with the number of connected components over 𝔽q\mathbb{F}_{q}. Therefore, every connected component is geometrically connected.

To conclude, it remains to deal with case (2)(2). This is completely analogous to (1)(1), but one uses 9.1.5(2) in place of 9.1.5(1), and therefore as output obtains the number of components agrees with 𝔼(#Hom(SelνBKLPR,rkVbn,H))\mathbb{E}(\#\mathrm{Hom}(\operatorname{Sel}^{\operatorname{BKLPR},\operatorname{rk}V_{\mathscr{F}^{n}_{b}}}_{\nu},H)) instead of 𝔼(#Hom(SelνBKLPR,H))\mathbb{E}(\#\mathrm{Hom}(\operatorname{Sel}^{\operatorname{BKLPR}}_{\nu},H)). ∎

Using the above computation of the connected components of our space, we are able to combine it with our topological tools, the Grothendieck-Lefschetz trace formula, and Deligne’s bounds to deduce the HH-moments of the distribution of Selmer groups in quadratic twist families.

Theorem 9.2.1.

Suppose B=SpecRB=\operatorname{Spec}R for RR a DVR of generic characteristic 0 with closed point bb with residue field 𝔽q0\mathbb{F}_{q_{0}} and geometric point b¯\overline{b} over bb. Keep hypotheses as in 7.1.4: Namely, suppose ν\nu is an odd integer and r>0r\in\mathbb{Z}_{>0} so that every prime ν\ell\mid\nu satisfies >2r+1\ell>2r+1. Let BB be an integral affine base scheme, CC a smooth proper curve with geometrically connected fibers over BB, ZCZ\subset C finite étale nonempty over BB, and U:=CZU:=C-Z. Let \mathscr{F} be a rank 2r2r, tame, locally constant constructible, symplectically self-dual sheaf of free /ν\mathbb{Z}/\nu\mathbb{Z} modules over UU. We assume there is some point xCb¯x\in C_{\overline{b}} at which Dropx(b¯[])=1\mathrm{Drop}_{x}(\mathscr{F}_{\overline{b}}[\ell])=1 for every prime ν\ell\mid\nu. Also suppose b¯[]\mathscr{F}_{\overline{b}}[\ell] is irreducible for each ν\ell\mid\nu, and that the map jb¯[w]jb¯[wt]j_{*}\mathscr{F}_{\overline{b}}[\ell^{w}]\to j_{*}\mathscr{F}_{\overline{b}}[\ell^{w-t}] is surjective for each prime ν\ell\mid\nu such that wν\ell^{w}\mid\nu, and wtw\geq t. Fix AUbA\to U_{b} as in 5.1.8 and suppose the tame irreducible locally constant constructible symplectically self-dual sheaf \mathscr{F} satisfies bA[ν]\mathscr{F}_{b}\simeq A[\nu]. For any finite /ν\mathbb{Z}/\nu\mathbb{Z} module HH, and any finite field extension 𝔽q0𝔽q\mathbb{F}_{q_{0}}\subset\mathbb{F}_{q}, there are constants C(H,)C(H,\mathscr{F}) depending on HH and \mathscr{F}, but not on qq or nn, so that

(9.4) |#SelBnH(𝔽q)qn#Hom(SelνBKLPR,H)|\displaystyle\left|\frac{\#\operatorname{Sel}_{\mathscr{F}^{n}_{B}}^{H}(\mathbb{F}_{q})}{q^{n}}-\#\mathrm{Hom}(\operatorname{Sel}^{\operatorname{BKLPR}}_{\nu},H)\right| C(H,)q\displaystyle\leq\frac{C(H,\mathscr{F})}{\sqrt{q}}
(9.5) |#SelBnH,rk(𝔽q)qn#Hom(SelνBKLPR,rkVBnmod2,H)|\displaystyle\left|\frac{\#\operatorname{Sel}^{H,\operatorname{rk}}_{\mathscr{F}^{n}_{B}}(\mathbb{F}_{q})}{q^{n}}-\#\mathrm{Hom}(\operatorname{Sel}^{\operatorname{BKLPR},\operatorname{rk}V_{\mathscr{F}^{n}_{B}}\bmod 2}_{\nu},H)\right| C(H,)q\displaystyle\leq\frac{C(H,\mathscr{F})}{\sqrt{q}}

for all even n>C(H,)n>C(H,\mathscr{F}), and all qq with q>C(H,)\sqrt{q}>C(H,\mathscr{F}).

Moreover, suppose there is a point σZ(B)\sigma\in Z(B) over which \mathscr{F} has trivial inertia. There are functions fH,(q)f_{H,\mathscr{F}}(q), and positive constants II, C(H,)C(H,\mathscr{F}), and J(,H)J(\mathscr{F},H) so that

(9.6) |#SelBnH(𝔽q)qnfH,(q)|\displaystyle\left|\frac{\#\operatorname{Sel}_{\mathscr{F}^{n}_{B}}^{H}(\mathbb{F}_{q})}{q^{n}}-f_{H,\mathscr{F}}(q)\right| (C(H,)q)nJ(,H)2I\displaystyle\leq\left(\frac{C(H,\mathscr{F})}{\sqrt{q}}\right)^{\frac{n-J(\mathscr{F},H)}{2I}}

for all even n>C(H,)n>C(H,\mathscr{F}), and all qq with q>2C(H,)\sqrt{q}>2C(H,\mathscr{F}).

Proof.

This follows from preceding results in our paper, together with the Grothendieck-Lefschetz trace formula and Deligne’s bounds, much in the same way that [EVW16, Theorem 8.8] follows from [EVW16, Proposition 7.8]. The remainder of the proof is somewhat standard, but we spell out the details for completeness.

We first explain (9.4) and (9.5). Fix a point bBb\in B with residue field 𝔽q\mathbb{F}_{q} with geometric point b¯\overline{b} over bb. Let (Yn)n1(Y_{n})_{n\geq 1} be a sequence of stacks over BB which is either either a sequence of the form (SelBnH)n1(\operatorname{Sel}_{\mathscr{F}^{n}_{B}}^{H})_{n\geq 1} or (SelBnH,rk)n1(\operatorname{Sel}^{H,\operatorname{rk}}_{\mathscr{F}^{n}_{B}})_{n\geq 1}. Define the sequence (Wn)n1(W_{n})_{n\geq 1} to be Wn:=(Yn)W_{n}:=(Y_{n})_{\mathbb{C}}, for some map SpecB\operatorname{Spec}\mathbb{C}\to B.

We next bound the cohomology groups of the geometric fiber of YnY_{n} over b¯\overline{b}, via comparison to the cohomology of WnW_{n}. Note that the YnY_{n} have coarse spaces which are finite étale covers of ConfU/Bn\operatorname{Conf}^{n}_{U/B}. Note that there is a normal crossings compactification of ConfU/Bn\operatorname{Conf}^{n}_{U/B} by B.1.3. It follows from [EVW16, Proposition 7.7] that the geometric generic fiber of YnY_{n} over BB has isomorphic cohomology to the geometric special fiber of YnY_{n} over BB. Now, we will choose \ell^{\prime} to be a sufficiently large prime, which may even depend on nn. We will see in the course of the proof how large \ell^{\prime} needs to be. (It is enough to take \ell^{\prime} to be prime to q,n!,#ASp2r(/ν),q,n!,\#\operatorname{\mathrm{ASp}}_{2r}(\mathbb{Z}/\nu\mathbb{Z}), and 22.) In other words, if we use Xn:=(Yn)𝔽¯qX_{n}:=(Y_{n})_{\overline{\mathbb{F}}_{q}} for the geometric special fiber, we obtain Hi(Xn,/)Hi(Wn,/)H^{i}(X_{n},\mathbb{Z}/\ell^{\prime}\mathbb{Z})\simeq H^{i}(W_{n},\mathbb{Z}/\ell^{\prime}\mathbb{Z}). By 8.2.3, the latter has dimension bounded by Ki+1K^{i+1}, for some constant KK independent of nn. Note that dimHi(Xn,/)rkHi(Xn,)dimHi(Xn,)\dim H^{i}(X_{n},\mathbb{Z}/\ell^{\prime}\mathbb{Z})\geq\operatorname{rk}H^{i}(X_{n},\mathbb{Z}_{\ell^{\prime}})\geq\dim H^{i}(X_{n},\mathbb{Q}_{\ell^{\prime}}), so we also have that Hi(Xn,)H^{i}(X_{n},\mathbb{Q}_{\ell^{\prime}}) is bounded by Ki+1K^{i+1}.

Since YnY_{n} is a finite étale cover of the smooth Deligne-Mumford stack QTwistU/bn\operatorname{QTwist}^{n}_{U/b}, every connected component is smooth and hence irreducible. Let ZnZ_{n} denote the number of connected components of XnX_{n}. Since all the connected components of XnX_{n} are base changed from 𝔽q\mathbb{F}_{q}, by 9.2.1, proving (9.4) and (9.5) amounts to proving

|#Yn(𝔽q)qdimXnZn|Cq\displaystyle\left|\frac{\#Y_{n}(\mathbb{F}_{q})}{q^{\dim X_{n}}}-Z_{n}\right|\leq\frac{C}{\sqrt{q}}

where CC is a constant depending on the sequence (Xn)n1(X_{n})_{n\geq 1}, but not the subscript nn.

Since XnX_{n} is smooth, using Poincaré duality, dimHc2ni(Xn,)=dimHi(Xn,)\dim H_{\operatorname{c}}^{2n-i}(X_{n},\mathbb{Q}_{\ell^{\prime}})=\dim H^{i}(X_{n},\mathbb{Q}_{\ell^{\prime}}). We may then produce a a constant DD, depending on the sequence (Xn)n1(X_{n})_{n\geq 1}, but not nn, such that dimHc2ni(Xn,)=dimHi(Xn,)Di\dim H_{\operatorname{c}}^{2n-i}(X_{n},\mathbb{Q}_{\ell^{\prime}})=\dim H^{i}(X_{n},\mathbb{Q}_{\ell^{\prime}})\leq D^{i}. For example, we can take D=K2D=K^{2}.

Since every eigenvalue of geometric Frobenius Frobq\operatorname{Frob}_{q} acting on the compactly supported cohomology group Hcj(Xn,)H_{\operatorname{c}}^{j}(X_{n},\mathbb{Q}_{\ell^{\prime}}) of the stack XnX_{n} is bounded in absolute value by qj/2q^{j/2}, using Sun’s generalization of Deligne’s bounds to algebraic stacks [Sun12, Theorem 1.4], we find

(9.7) |qdimXnj<2dimXn(1)jtr(FrobqHcj(Xn,))|\displaystyle\left|q^{-\dim X_{n}}\sum_{j<2\dim X_{n}}(-1)^{j}\operatorname{tr}\left(\operatorname{Frob}_{q}|H^{j}_{\operatorname{c}}(X_{n},\mathbb{Q}_{\ell^{\prime}})\right)\right|
qdimXnj=02dimXn1qj/2dimHcj(Xn,)\displaystyle\leq q^{-\dim X_{n}}\sum_{j=0}^{2\dim X_{n}-1}q^{j/2}\dim H^{j}_{c}(X_{n},\mathbb{Q}_{\ell^{\prime}})
qdimXnj=02dimXn1qj/2D2dimXnj\displaystyle\leq q^{-\dim X_{n}}\sum_{j=0}^{2\dim X_{n}-1}q^{j/2}D^{2\dim X_{n}-j}
k=1(Dq)k.\displaystyle\leq\sum_{k=1}^{\infty}\left(\frac{D}{\sqrt{q}}\right)^{k}.

This is bounded by 2D/q2D/\sqrt{q} whenever D/q1/2D/\sqrt{q}\leq 1/2. Hence, taking C:=2DC:=2D, we obtain

|qdimXnj<2dimXn(1)jtr(FrobqHcj(Xn,))|Cq\displaystyle\left|q^{-\dim X_{n}}\sum_{j<2\dim X_{n}}(-1)^{j}\operatorname{tr}\left(\operatorname{Frob}_{q}|H^{j}_{\operatorname{c}}(X_{n},\mathbb{Q}_{\ell^{\prime}})\right)\right|\leq\frac{C}{\sqrt{q}}

whenever CqC\leq\sqrt{q}. Therefore, using the Grothendieck-Lefschetz trace formula, it is enough to show tr(Frobq|Hc2dimXn(Xn,))=ZnqdimXn\operatorname{tr}\left(\operatorname{Frob}_{q}|H^{2\dim X_{n}}_{\operatorname{c}}(X_{n},\mathbb{Q}_{\ell^{\prime}})\right)=Z_{n}q^{\dim X_{n}} for nn sufficiently large, say larger than some constant C1C_{1}. By Poincaré duality, this is equivalent to showing that there are ZnZ_{n} connected components of XnX_{n}, all of which are defined over 𝔽q\mathbb{F}_{q}. Indeed, this was shown in 9.2.1. Finally, we then take C(H,)C(H,\mathscr{F}) in the statement to be max(C,C1)\max(C,C_{1}), which proves (9.4) and (9.5).

We conclude by briefly outlining how one may similarly obtain (9.6) by additionally using Theorem A.5.1. We assume qq is sufficiently large so that the hypotheses of Theorem A.5.1 are satisfied; namely, that Sel𝒢Bd,0,(𝔽q)\operatorname{Sel}_{\mathscr{G}^{d,0,\infty}_{B}}(\mathbb{F}_{q})\neq\emptyset. This will happen for all sufficiently large qq by the Lefschetz trace formula. We maintain the notation set up earlier in the proof. We use JJ to denote J(,H)J(\mathscr{F},H) from the theorem statement. By Theorem A.5.1, when n>Ip+Jn>Ip+J is even, tr(Frobq1|Hp(Xn,))\operatorname{tr}(\operatorname{Frob}_{q}^{-1}|H^{p}(X_{n},\mathbb{Q}_{\ell^{\prime}})) takes on a value independent of nn. Fixing nn even with n>Ip+Jn>Ip+J, let tp(q):=tr(Frobq1|Hp(Xn,))t_{p}(q):=\operatorname{tr}(\operatorname{Frob}_{q}^{-1}|H^{p}(X_{n},\mathbb{Q}_{\ell^{\prime}})), where as usual Frobq\operatorname{Frob}_{q} denotes geometric frobenius. Then define f(q):=j=0(1)jtj(q)f(q):=\sum_{j=0}^{\infty}(-1)^{j}t_{j}(q). (We use ff in place of the function fH,f_{H,\mathscr{F}} as in the theorem statement.) We claim that ff converges as a function in qq for qq sufficiently large. Indeed, 8.2.3 and Deligne’s bounds on the eigenvalues of Frobenius acting on cohomology yield |tp(q)|<Kp+1qp/2|t_{p}(q)|<\frac{K^{p+1}}{q^{p/2}}, and so f(q)f(q) is bounded by a geometric series; see (9.7) and the surrounding paragraphs for a similar bounding argument which is spelled out in more detail.

We now conclude (9.6) by applying the Grothendieck-Lefschetz trace formula. Note that the condition n>Ip+Jn>Ip+J is equivalent to the condition p<nJIp<\frac{n-J}{I}. Note here we are using that the Galois representation Hc2ni(Xn,)H^{2n-i}_{\operatorname{c}}(X_{n},\mathbb{Q}_{\ell^{\prime}}) is identified with the Galois representation Hi(Xn,)(n)H^{i}(X_{n},\mathbb{Q}_{\ell^{\prime}})^{\vee}(-n) via Poincaré duality, and so qntr(Frobq1|Hi(Xn,))=tr(Frobq|Hc2ni(Xn,))q^{n}\cdot\operatorname{tr}(\operatorname{Frob}_{q}^{-1}|H^{i}(X_{n},\mathbb{Q}_{\ell^{\prime}}))=\operatorname{tr}\left(\operatorname{Frob}_{q}|H^{2n-i}_{\operatorname{c}}(X_{n},\mathbb{Q}_{\ell^{\prime}})\right). From this, it follows that 1qntr(Frobq|Hc2nj(Xn,))=tj(q)\frac{1}{q^{n}}\operatorname{tr}\left(\operatorname{Frob}_{q}|H^{2n-j}_{\operatorname{c}}(X_{n},\mathbb{Q}_{\ell^{\prime}})\right)=t_{j}(q). Using the above observation combined with the Grothendieck Lefschetz trace formula, the difference |#Xn(𝔽q)qnf(q)|\left|\frac{\#X_{n}(\mathbb{F}_{q})}{q^{n}}-f(q)\right| can be bounded by the sum of j=nJI(1)jtj(q)\sum_{j=\frac{n-J}{I}}^{\infty}(-1)^{j}t_{j}(q) and

(9.8) 1qnj2nnJI(1)jtr(Frobq|Hcj(Xn,)).\displaystyle\frac{1}{q^{n}}\sum_{j\leq 2n-\frac{n-J}{I}}(-1)^{j}\operatorname{tr}\left(\operatorname{Frob}_{q}|H^{j}_{\operatorname{c}}(X_{n},\mathbb{Q}_{\ell^{\prime}})\right).

By a computation analogous to (9.7), we can bound (9.8) in absolute value by CqnJ2I\frac{C}{q^{\frac{n-J}{2I}}}, for an appropriate constant CC not depending on qq or nn, once nn is sufficiently large and q>2C\sqrt{q}>2C. ∎

Remark 9.2.4.

Suppose one started with a setup as in Theorem 9.2.1, but where BB is a nonempty open in Spec𝒪K\operatorname{Spec}\mathscr{O}_{K}, for KK a number field. (Note that if one starts with this setup over SpecK\operatorname{Spec}K, one can spread it out to such a BB.) For any geometric point Spec𝔽¯qB\operatorname{Spec}\overline{\mathbb{F}}_{q}\to B, we can identify the cohomology groups of the relevant moduli spaces (labeled XnX_{n} in the proof of Theorem 9.2.1) over Spec𝔽¯q\operatorname{Spec}\overline{\mathbb{F}}_{q} with the corresponding cohomology groups over the geometric generic point SpecB\operatorname{Spec}\mathbb{C}\to B, (which are the cohomology of WnW_{n} in the proof of Theorem 9.2.1,) independently of the choice of geometric point above. Then, one could prove a result as in Theorem 9.2.1, but with the limit in qq ranging over primes of all but finitely many characteristics, instead of only powers of a given prime power q0q_{0}.

Remark 9.2.5.

Although the constants C(H,)C(H,\mathscr{F}) in Theorem 9.2.1 depend on \mathscr{F} and HH as stated, they can in fact be chosen to be functions of ν\nu, the rank 2r2r of \mathscr{F} and the degree f+1f+1 of ZZ, and the genus gg of CC, as we next explain.

One way to see this is via comparison to the complex numbers. Then, over the complex numbers, the constants only depend on the topological type of the finite covering space associated to \mathscr{F} over UU. There are only finitely many such topological types once we fix r,νr,\nu, and ff, since the number of these types is bounded by the number of homomorphisms π1(Σg,f+1)ASp2r(/ν)\pi_{1}(\Sigma_{g,f+1})\to\operatorname{\mathrm{ASp}}_{2r}(\mathbb{Z}/\nu\mathbb{Z}), of which there are only finitely many. Hence, the relevant constants C(H,)C(H,\mathscr{F}) can be taken to only depend on r,ν,f,gr,\nu,f,g, and HH.

Remark 9.2.6.

Suppose the stable cohomology groups of spaces appearing in the proof of Theorem 9.2.1, which are not in the top degree, vanish. Then, via the Grothendieck-Lefschetz trace formula, one could deduce that the constants C(H,)C(H,\mathscr{F}) actually vanish. This would imply some of our main results, such as Theorem 1.1.3, hold on the nose for fixed, sufficiently large qq, depending on HH, without the need for taking a large qq limit.

Remark 9.2.7.

It seems likely one could additionally find a function fH,rk(q)f_{H,\mathscr{F}}^{\operatorname{rk}}(q) as in the statement of Theorem 9.2.1 and positive constants II, C(H,)C(H,\mathscr{F}), and J(,H)J(\mathscr{F},H) so that

(9.9) |#SelBnH,rk(𝔽q)qnfH,rk(q)|\displaystyle\left|\frac{\#\operatorname{Sel}^{H,\operatorname{rk}}_{\mathscr{F}^{n}_{B}}(\mathbb{F}_{q})}{q^{n}}-f_{H,\mathscr{F}}^{\operatorname{rk}}(q)\right| (C(H,)q)nJ(,H)2I\displaystyle\leq\left(\frac{C(H,\mathscr{F})}{\sqrt{q}}\right)^{\frac{n-J(\mathscr{F},H)}{2I}}

for all even n>C(H,)n>C(H,\mathscr{F}), and all qq with q>2C(H,)\sqrt{q}>2C(H,\mathscr{F}). For this, one would only need to generalize Theorem A.5.1 to also work for the rank double cover. This seems quite doable, but we have opted not to carry it out as it was not required for our main theorems. We do, however, believe it would be quite interesting to work out.

We conclude with a variant of Theorem 9.2.1, where the powers of qq appearing in the denominators of (9.4) and (9.5) are replaced by the number of points of the stack of quadratic twists.

Corollary 9.2.8.

With notation and hypotheses as in Theorem 9.2.1, after suitably changing the constants C(H,)C(H,\mathscr{F}), we also have

(9.10) |#SelBnH(𝔽q)#QTwistU/Bn(𝔽q)#Hom(SelνBKLPR,H)|\displaystyle\left|\frac{\#\operatorname{Sel}_{\mathscr{F}^{n}_{B}}^{H}(\mathbb{F}_{q})}{\#\operatorname{QTwist}^{n}_{U/B}(\mathbb{F}_{q})}-\#\mathrm{Hom}(\operatorname{Sel}^{\operatorname{BKLPR}}_{\nu},H)\right| C(H,)q\displaystyle\leq\frac{C(H,\mathscr{F})}{\sqrt{q}}
(9.11) |#SelBnH,rk(𝔽q)#QTwistrk,n(𝔽q)#Hom(SelνBKLPR,rkVBnmod2,H)|\displaystyle\left|\frac{\#\operatorname{Sel}^{H,\operatorname{rk}}_{\mathscr{F}^{n}_{B}}(\mathbb{F}_{q})}{\#\operatorname{QTwist}^{\operatorname{rk},n}_{\mathscr{F}}(\mathbb{F}_{q})}-\#\mathrm{Hom}(\operatorname{Sel}^{\operatorname{BKLPR},\operatorname{rk}V_{\mathscr{F}^{n}_{B}}\bmod 2}_{\nu},H)\right| C(H,)q.\displaystyle\leq\frac{C(H,\mathscr{F})}{\sqrt{q}}.

for all even n>C(H,)n>C(H,\mathscr{F}), and all qq with q>C(H,)\sqrt{q}>C(H,\mathscr{F}), and gcd(q,2ν)=1\gcd(q,2\nu)=1.

Proof.

First, applying Theorem 9.2.1 in the case HH is the identity group gives that both #QTwistU/Bn(𝔽q)\#\operatorname{QTwist}^{n}_{U/B}(\mathbb{F}_{q}) and #QTwistrk,n(𝔽q)\#\operatorname{QTwist}^{\operatorname{rk},n}_{\mathscr{F}}(\mathbb{F}_{q}) have qnq^{n} points, up to an error of C(id)/qC(\operatorname{\mathrm{id}})/\sqrt{q}.

Hence, in Theorem 9.2.1, after adjusting the constant C(H,)C(H,\mathscr{F}), we can freely replace qnq^{n} appearing in the denominator in (9.4) and (9.5) with #QTwistU/Bn(𝔽q)\#\operatorname{QTwist}^{n}_{U/B}(\mathbb{F}_{q}) and #QTwistrk,n(𝔽q)\#\operatorname{QTwist}^{\operatorname{rk},n}_{\mathscr{F}}(\mathbb{F}_{q}). ∎

10. Determining the distribution from the moments

In this section, we complete the proof of our main result. In § 10.1 we prove a probabilistic result, which we use to show that the distributions we are studying are determined by their moments, conditioned on the parity of the \ell^{\infty} Selmer rank. Then, in § 10.2, we put everything together, proving our main results in § 10.2.2, § 10.2.3, and § 10.2.4.

10.1. Approximating distributions by approximating moments

In Theorem 9.2.1, we determined the moments of distributions relating to Selmer groups, after taking appropriate limits. We would like to show these moments determine the distribution. If we knew the moments exactly, without taking a qq\to\infty limit, we could appeal to [NW22, Theorem 4.1] to show the distribution is also determined. The next general result will allow us to deal with this issue of taking the qq\to\infty limit. We thank Melanie Wood for pointing out the following argument, which simplifies our previous approach.

Proposition 10.1.1.

Let 𝒩\mathcal{N} denote the set of isomorphism classes of finite abelian /ν\mathbb{Z}/\nu\mathbb{Z} modules and let 𝒮𝒩\mathcal{S}\subset\mathcal{N} denote a subset. Suppose (Xji)iI,jJ(X^{i}_{j})_{i\in I,j\in J} form a set of 𝒮\mathcal{S}-valued random variables, for I,JI,J two infinite subsets of the positive integers. Suppose there is some 𝒮\mathcal{S}-valued random variable YY so that

  1. (1)

    for every H𝒩H\in\mathcal{N} and for any fixed sufficiently large value of ii depending on HH,

    limj𝔼(#Surj(Xji,H))=𝔼(#Surj(Y,H)),\displaystyle\lim_{j\to\infty}\mathbb{E}\left(\#\operatorname{Surj}(X^{i}_{j},H)\right)=\mathbb{E}\left(\#\operatorname{Surj}(Y,H)\right),

    and

  2. (2)

    for any sequence (Ys)s1(Y_{s})_{s\geq 1} of 𝒮\mathcal{S}-valued random variables such that

    lims𝔼(#Surj(Ys,H))=𝔼(#Surj(Y,H)),\displaystyle\lim_{s\to\infty}\mathbb{E}\left(\#\operatorname{Surj}(Y_{s},H)\right)=\mathbb{E}\left(\#\operatorname{Surj}(Y,H)\right),

    we have limsProb(YsA)=Prob(YA)\lim_{s\to\infty}\operatorname{Prob}(Y_{s}\simeq A)=\operatorname{Prob}(Y\simeq A) for every A𝒩A\in\mathcal{N}.

Then, both

limjlim supiProb(XjiA) and limjlim infiProb(XjiA)\displaystyle\lim_{j\to\infty}\limsup_{i\to\infty}\operatorname{Prob}(X^{i}_{j}\simeq A)\text{ and }\lim_{j\to\infty}\liminf_{i\to\infty}\operatorname{Prob}(X^{i}_{j}\simeq A)

exist, and are equal to Prob(YA)\operatorname{Prob}(Y\simeq A).

Proof.

Place a total ordering on the countable set 𝒩\mathcal{N}, so that HtH_{t} is the ttth element of 𝒩\mathcal{N}. By our first assumption, for fixed sufficiently large ii depending on HH, limj𝔼(#Surj(Xji,H))=𝔼(#Surj(Y,H))\lim_{j\to\infty}\mathbb{E}(\#\operatorname{Surj}(X^{i}_{j},H))=\mathbb{E}(\#\operatorname{Surj}(Y,H)). This implies we can find a sequence of pairs (is,js)s1(i_{s},j_{s})_{s\geq 1} so that for every s1s\geq 1 and every tst\leq s,

|𝔼(#Surj(Xjsis,Ht))𝔼(#Surj(Y,Ht))|<2s.\displaystyle\left|\mathbb{E}(\#\operatorname{Surj}(X^{i_{s}}_{j_{s}},H_{t}))-\mathbb{E}(\#\operatorname{Surj}(Y,H_{t}))\right|<2^{-s}.

This implies that lims𝔼(#Surj(Xjsis,H))=𝔼(#Surj(Y,H))\lim_{s\to\infty}\mathbb{E}(\#\operatorname{Surj}(X^{i_{s}}_{j_{s}},H))=\mathbb{E}\left(\#\operatorname{Surj}\left(Y,H\right)\right) for every H𝒩H\in\mathcal{N}. Hence, by our second assumption, applied to the sequence (Ys)s1(Y_{s})_{s\geq 1} defined by Ys:=XjsisY_{s}:=X^{i_{s}}_{j_{s}}, we find limsProb(XjsisA)=Prob(YA)\lim_{s\to\infty}\operatorname{Prob}(X^{i_{s}}_{j_{s}}\simeq A)=\operatorname{Prob}(Y\simeq A). Using [Saw20, Lemma 2.22], we find

lim supjlim supiProb(XjiA)=lim infjlim infiProb(XjiA)=Prob(YA).\displaystyle\limsup_{j\to\infty}\limsup_{i\to\infty}\operatorname{Prob}(X^{i}_{j}\simeq A)=\liminf_{j\to\infty}\liminf_{i\to\infty}\operatorname{Prob}(X^{i}_{j}\simeq A)=\operatorname{Prob}(Y\simeq A).

To conclude, note that

lim supjlim supiProb(XjiA)lim infjlim supiProb(XjiA)lim infjlim infiProb(XjiA),\displaystyle\limsup_{j\to\infty}\limsup_{i\to\infty}\operatorname{Prob}(X^{i}_{j}\simeq A)\geq\liminf_{j\to\infty}\limsup_{i\to\infty}\operatorname{Prob}(X^{i}_{j}\simeq A)\geq\liminf_{j\to\infty}\liminf_{i\to\infty}\operatorname{Prob}(X^{i}_{j}\simeq A),

and since the outer two limits are equal, they also agree with the middle one. This implies limjlim supiProb(XjiA)\lim_{j\to\infty}\limsup_{i\to\infty}\operatorname{Prob}(X^{i}_{j}\simeq A) exists and agrees with Prob(YA)\operatorname{Prob}(Y\simeq A). Analogously, we also find limjlim infiProb(XjiA)\lim_{j\to\infty}\liminf_{i\to\infty}\operatorname{Prob}(X^{i}_{j}\simeq A) exists and agrees with Prob(YA)\operatorname{Prob}(Y\simeq A). ∎

10.2. Proving the main result

We can now prove our main result. To set up notation, suppose we are in the setting of 5.1.8, so that AUbA\to U_{b} is an abelian scheme with bA[ν]\mathscr{F}_{b}\simeq A[\nu]. For xQTwistUb/bnx\in\operatorname{QTwist}^{n}_{U_{b}/b}, and AxUxA_{x}\to U_{x} the corresponding abelian scheme over a curve, we use Selν(Ax)\operatorname{Sel}_{\nu}(A_{x}) to denote the ν\nu Selmer group of the generic fiber of AxA_{x} over UxU_{x}. In the following theorem, we use the standard convention that the 𝔽q\mathbb{F}_{q} points of a stack, such as QTwistU/Bn\operatorname{QTwist}^{n}_{U/B}, are counted weighted by the inverse of the size of the automorphism group of that point. Also recall the notation introduced in 7.4.1 for the distributions of Selmer groups. The following statement is nearly our main result, but here we start out over a dvr, instead of a finite field. Following the proof of this, we will need to lift all our data from a finite field to a dvr in order to deduce Theorem 1.1.2.

Theorem 10.2.1.

Suppose B=SpecRB=\operatorname{Spec}R for RR a dvr of generic characteristic 0 with closed point bb with residue field 𝔽q0\mathbb{F}_{q_{0}} and geometric closed point b¯\overline{b} over bb. Keep hypotheses as in 7.1.4: Namely, suppose ν\nu is an odd integer and r>0r\in\mathbb{Z}_{>0} so that every prime ν\ell\mid\nu satisfies >2r+1\ell>2r+1. Let BB be an integral affine base scheme, CC a smooth proper curve with geometrically connected fibers over BB, ZCZ\subset C finite étale nonempty over BB, and U:=CZU:=C-Z. Let \mathscr{F} be a rank 2r2r, tame, locally constant constructible, symplectically self-dual sheaf of free /ν\mathbb{Z}/\nu\mathbb{Z} modules over UU. We assume there is some point xCb¯x\in C_{\overline{b}} at which Dropx(b¯[])=1\mathrm{Drop}_{x}(\mathscr{F}_{\overline{b}}[\ell])=1 for every prime ν\ell\mid\nu. Also suppose b¯[]\mathscr{F}_{\overline{b}}[\ell] is irreducible for each ν\ell\mid\nu, and that the map jb¯[w]jb¯[wt]j_{*}\mathscr{F}_{\overline{b}}[\ell^{w}]\to j_{*}\mathscr{F}_{\overline{b}}[\ell^{w-t}] is surjective for each prime ν\ell\mid\nu such that wν\ell^{w}\mid\nu, and wtw\geq t. Fix AUbA\to U_{b} as in 5.1.8 and suppose the tame irreducible locally constant constructible symplectically self-dual sheaf of free /ν\mathbb{Z}/\nu\mathbb{Z} modules \mathscr{F} satisfies bA[ν]\mathscr{F}_{b}\simeq A[\nu]. With notation as in 7.4.1, we have that, for each /ν\mathbb{Z}/\nu\mathbb{Z} module HH,

(10.1) limq𝔽q0𝔽qlim supnnevenProb(XA[ν]𝔽qnH)\displaystyle\lim_{\begin{subarray}{c}q\to\infty\\ \mathbb{F}_{q_{0}}\subset\mathbb{F}_{q}\end{subarray}}\limsup_{\begin{subarray}{c}n\to\infty\\ n\hskip 2.84544pt\mathrm{even}\end{subarray}}\operatorname{Prob}(X_{A[\nu]^{n}_{\mathbb{F}_{q}}}\simeq H)
limq𝔽q0𝔽qlim infnnevenProb(XA[ν]𝔽qnH)\displaystyle\lim_{\begin{subarray}{c}q\to\infty\\ \mathbb{F}_{q_{0}}\subset\mathbb{F}_{q}\end{subarray}}\liminf_{\begin{subarray}{c}n\to\infty\\ n\hskip 2.84544pt\mathrm{even}\end{subarray}}\operatorname{Prob}(X_{A[\nu]^{n}_{\mathbb{F}_{q}}}\simeq H)

exist and agree with Prob(SelνBKLPRH)\operatorname{Prob}(\operatorname{Sel}^{\operatorname{BKLPR}}_{\nu}\simeq H). Similarly, for i{0,1}i\in\{0,1\},

(10.2) limq𝔽q0𝔽qlim supnnevenProb(XA[ν]𝔽qniH)\displaystyle\lim_{\begin{subarray}{c}q\to\infty\\ \mathbb{F}_{q_{0}}\subset\mathbb{F}_{q}\end{subarray}}\limsup_{\begin{subarray}{c}n\to\infty\\ n\hskip 2.84544pt\mathrm{even}\end{subarray}}\operatorname{Prob}(X^{i}_{A[\nu]^{n}_{\mathbb{F}_{q}}}\simeq H)
limq𝔽q0𝔽qlim infnnevenProb(XA[ν]𝔽qniH)\displaystyle\lim_{\begin{subarray}{c}q\to\infty\\ \mathbb{F}_{q_{0}}\subset\mathbb{F}_{q}\end{subarray}}\liminf_{\begin{subarray}{c}n\to\infty\\ n\hskip 2.84544pt\mathrm{even}\end{subarray}}\operatorname{Prob}(X^{i}_{A[\nu]^{n}_{\mathbb{F}_{q}}}\simeq H)

exist and agree with Prob(SelνBKLPR,iH)\operatorname{Prob}(\operatorname{Sel}^{\operatorname{BKLPR},i}_{\nu}\simeq H).

Proof.

First, take i0:=rkVABnmod2{0,1}i_{0}:=\operatorname{rk}V_{A^{n}_{B}}\bmod 2\in\{0,1\}. We will apply 10.1.1 with 𝒮=𝒩i0,Y=SelνBKLPR,i0,Xqn=XA[ν]𝔽qni0\mathcal{S}=\mathcal{N}^{i_{0}},Y=\operatorname{Sel}^{\operatorname{BKLPR},i_{0}}_{\nu},X^{n}_{q}=X^{i_{0}}_{A[\nu]^{n}_{\mathbb{F}_{q}}} to prove (LABEL:equation:parity-limit) for i=i0i=i_{0}. (Here, we use XqnX^{n}_{q} in place of the notation XjiX^{i}_{j} from 10.1.1.)

We will now check the hypotheses of 10.1.1. We need to check the XqnX^{n}_{q} and YY are both supported on 𝒮\mathcal{S}, as well as the two enumerated hypotheses of 10.1.1. The XqnX^{n}_{q} are supported on 𝒮\mathcal{S} by 7.4.5. To show YY is supported on 𝒮\mathcal{S}, from the definition in § 2.2.2, it is enough to show the distribution 𝒯r,/ν\mathcal{T}_{r,\mathbb{Z}/\nu\mathbb{Z}} defined there is supported on abelian groups which are squares, i.e., abelian groups of the form K2K^{2} for KK an abelian group. For this, it is enough to show that for any prime ν\ell\mid\nu, 𝒯r,/\mathcal{T}_{r,\mathbb{Z}/\ell\mathbb{Z}} is supported on squares. This follows because it is supported on groups with a nondegenerate alternating pairing by [BKL+15, Proposition 5.5], using that groups with a nondegenerate alternating pairing are squares.

We next check the enumerated hypotheses of 10.1.1. The first enumerated hypothesis of 10.1.1 follows from combining 8.3.2 and (9.11), together with an inclusion exclusion argument allows us to replace the Hom\mathrm{Hom} appearing in these results with Surj\operatorname{Surj}. In order to verify the second enumerated hypothesis of 10.1.1, we use 2.3.1, which bounds the moments of Y=SelνBKLPR,i0Y=\operatorname{Sel}^{\operatorname{BKLPR},i_{0}}_{\nu}. The second hypothesis then follows from [NW22, Theorem 4.1]. This verifies the hypotheses of 10.1.1, and its conclusion implies (LABEL:equation:parity-limit) for i=i0i=i_{0}.

Having proven (LABEL:equation:parity-limit) for i=i0i=i_{0}, we next aim to prove it for i=1i0i=1-i_{0}. In this case, note that for any H𝒩H\in\mathcal{N}, #Surj(XA[ν]𝔽qn,H)\#\operatorname{Surj}(X_{A[\nu]^{n}_{\mathbb{F}_{q}}},H) and #Surj(XA[ν]𝔽qni0,H)\#\operatorname{Surj}(X^{i_{0}}_{A[\nu]^{n}_{\mathbb{F}_{q}}},H) take on the same value, up to an error of C(H,)/qC(H,\mathscr{F})/\sqrt{q}, by combining 8.3.2, Theorem 9.2.1, and 2.3.1. It follows that #Surj(XA[ν]𝔽qn1i0,H)\#\operatorname{Surj}(X^{1-i_{0}}_{A[\nu]^{n}_{\mathbb{F}_{q}}},H) also takes on this same value, up to an error of 2C(H,)/q2C(H,\mathscr{F})/\sqrt{q}. Hence, an analogous argument to the one above for the case i=i0i=i_{0}, this time applying 10.1.1 with 𝒮=𝒩1i0,Y=SelνBKLPR,1i0,Xqn=XA[ν]𝔽qn1i0\mathcal{S}=\mathcal{N}^{1-i_{0}},Y=\operatorname{Sel}^{\operatorname{BKLPR},1-i_{0}}_{\nu},X^{n}_{q}=X^{1-i_{0}}_{A[\nu]^{n}_{\mathbb{F}_{q}}} proves (LABEL:equation:parity-limit) for i=1i0i=1-i_{0}.

Finally, it remains to prove (LABEL:equation:total-distribution-limit). By 7.4.5, the distribution XA[ν]𝔽qnX_{A[\nu]^{n}_{\mathbb{F}_{q}}} is supported on 𝒩0𝒩1\mathcal{N}^{0}\coprod\mathcal{N}^{1}, and so both limits in (LABEL:equation:total-distribution-limit) exist by summing the limits in (LABEL:equation:parity-limit) in the cases i=0i=0 and i=1i=1. Since

XA[ν]𝔽qn=XA[ν]𝔽qn0Prob(XA[ν]𝔽qn𝒩0)+XA[ν]𝔽qn1Prob(XA[ν]𝔽qn𝒩1),\displaystyle X_{A[\nu]^{n}_{\mathbb{F}_{q}}}=X^{0}_{A[\nu]^{n}_{\mathbb{F}_{q}}}\cdot\operatorname{Prob}(X_{A[\nu]^{n}_{\mathbb{F}_{q}}}\in\mathcal{N}^{0})+X^{1}_{A[\nu]^{n}_{\mathbb{F}_{q}}}\cdot\operatorname{Prob}(X_{A[\nu]^{n}_{\mathbb{F}_{q}}}\in\mathcal{N}^{1}),

it is enough to show

(10.3) 1/2=limq𝔽q0𝔽qlim supnnevenProb(XA[ν]𝔽qn𝒩i0),\displaystyle 1/2=\lim_{\begin{subarray}{c}q\to\infty\\ \mathbb{F}_{q_{0}}\subset\mathbb{F}_{q}\end{subarray}}\limsup_{\begin{subarray}{c}n\to\infty\\ n\hskip 2.84544pt\mathrm{even}\end{subarray}}\operatorname{Prob}(X_{A[\nu]^{n}_{\mathbb{F}_{q}}}\in\mathcal{N}^{i_{0}}),

and the analogous statement for lim inf\liminf in place of lim sup\limsup. Indeed, by 8.3.1, the probability Prob(XA[ν]𝔽qn𝒩i0)\operatorname{Prob}(X_{A[\nu]^{n}_{\mathbb{F}_{q}}}\in\mathcal{N}^{i_{0}}) is exactly the probability that an 𝔽q\mathbb{F}_{q} point of QTwistU/Bn\operatorname{QTwist}^{n}_{U/B} is in the image of an 𝔽q\mathbb{F}_{q} point of QTwistrk,n\operatorname{QTwist}^{\operatorname{rk},n}_{\mathscr{F}}. Note that for n>0n>0, QTwistU/Bn\operatorname{QTwist}^{n}_{U/B} and QTwistrk,n\operatorname{QTwist}^{\operatorname{rk},n}_{\mathscr{F}} are both geometrically irreducible; the latter uses Theorem 7.1.1, which implies that the geometric monodromy is nontrivial under the Dickson invariant map. Using (9.4) for H=idH=\operatorname{\mathrm{id}} and (9.5) for H=idH=\operatorname{\mathrm{id}} we find both QTwistU/Bn\operatorname{QTwist}^{n}_{U/B} and QTwistrk,n\operatorname{QTwist}^{\operatorname{rk},n}_{\mathscr{F}} have qdimQTwistrk,n+O(1/q)q^{\dim\operatorname{QTwist}^{\operatorname{rk},n}_{\mathscr{F}}}+O(1/\sqrt{q}) points, where the implicit constant is independent of nn. This implies (10.3) because the number of 𝔽q\mathbb{F}_{q} points in the image of QTwistrk,n(𝔽q)QTwistU/Bn(𝔽q)\operatorname{QTwist}^{\operatorname{rk},n}_{\mathscr{F}}(\mathbb{F}_{q})\to\operatorname{QTwist}^{n}_{U/B}(\mathbb{F}_{q}) is half the number of 𝔽q\mathbb{F}_{q} points of QTwistrk,n(𝔽q)\operatorname{QTwist}^{\operatorname{rk},n}_{\mathscr{F}}(\mathbb{F}_{q}), since this map is a finite étale double cover. ∎

We have nearly proven our main result, Theorem 1.1.2, except that Theorem 10.2.1 begins over a base BB of generic characteristic 0, while Theorem 1.1.2 begins over a finite field. It remains to show that if one starts over a finite field, one can lift the relevant data to a dvr with generic characteristic 0. This is essentially the content of the next lemma, for which we use the following definition.

Definition 10.2.1.

Given a base scheme BB, a symplectic sheaf data over BB is a quadruple (C,U,Z,)(C,U,Z,\mathscr{F}) over BB, where CC is a relative smooth proper curve with geometrically connected fibers over BB, UCU\subset C is a nonempty open, Z=CUZ=C-U is a nonempty divisor which is finite étale over BB, and \mathscr{F} is a tame symplectically self-dual sheaf of /ν\mathbb{Z}/\nu\mathbb{Z} modules on UU.

Lemma 10.2.2.

Suppose we are given a symplectic sheaf data (C0,U0,Z0,0)(C_{0},U_{0},Z_{0},\mathscr{F}_{0}) over Spec𝔽q\operatorname{Spec}\mathbb{F}_{q}. If BB is the spectrum of a complete dvr with residue field 𝔽q\mathbb{F}_{q}, there exists a symplectic sheaf data (C,U,Z,)(C,U,Z,\mathscr{F}) over BB whose restriction to bb, (Cb,Ub,Zb,b)(C_{b},U_{b},Z_{b},\mathscr{F}_{b}), is isomorphic to (C0,U0,Z0,0)(C_{0},U_{0},Z_{0},\mathscr{F}_{0}).

Proof.

The general strategy of the proof will be to show we can lift (C0,U0,Z0,0)(C_{0},U_{0},Z_{0},\mathscr{F}_{0}) to arbitrary neighborhoods of bBb\in B and then algebraize this data. If B=SpecSB=\operatorname{Spec}S, with SS a complete dvr and uniformizer π\pi, let bn:=SpecS/πn+1b_{n}:=\operatorname{Spec}S/\pi^{n+1}. If (Ci,Zi)(C_{i},Z_{i}) is some lifting of (C0,Z0)(C_{0},Z_{0}) to bib_{i}, then the obstruction to further lifting it to bi+1b_{i+1} vanishes because it lies in the coherent cohomology group H2(C0,ΩC0/b(logZ0))=0H^{2}(C_{0},\Omega_{C_{0}/b}(\operatorname{log}Z_{0}))=0. By [FGI+05, Theorem 8.4.10], we can lift CiC_{i} to C0C_{0} over BB using the ample line bundle 𝒪Ci(Zi)\mathscr{O}_{C_{i}}(Z_{i}) on CiC_{i}. Using [FGI+05, Corollary 8.4.5], we obtain a closed subscheme ZCZ\subset C restricting to ZiCiZ_{i}\subset C_{i} over bib_{i}. Note that ZZ is finite étale over BB because it dominates BB and Zb=Z0Z_{b}=Z_{0} is geometrically reduced, (as the residue field is assumed to be perfect,) hence smooth over bb.

Next, we wish to show 0\mathscr{F}_{0} over U0U_{0} lifts to \mathscr{F} over UU. In fact, 0\mathscr{F}_{0} has a unique lift by [Wew99, Corollary 3.1.3], which we note uses our tameness assumption on 0\mathscr{F}_{0}. Note there that 0\mathscr{F}_{0} is a locally constant constructible sheaf with finite coefficients, and when applying the above, we are viewing it as a finite étale cover of U0U_{0}. The lift \mathscr{F} corresponds to a locally constant constructible sheaf, using the uniqueness of the lift. Moreover, by uniqueness of the lift above, the isomorphism 00(1)\mathscr{F}_{0}\simeq\mathscr{F}_{0}^{\vee}(1) giving 0\mathscr{F}_{0} its symplectic self-dual structure lifts to an isomorphism (1)\mathscr{F}\simeq\mathscr{F}^{\vee}(1), giving \mathscr{F} a self-dual structure. Since 00μν\mathscr{F}_{0}\otimes\mathscr{F}_{0}\to\mu_{\nu} factors through 20\wedge^{2}\mathscr{F}_{0}, we also obtain that μν\mathscr{F}\otimes\mathscr{F}\to\mu_{\nu} factors through 2\wedge^{2}\mathscr{F}, implying \mathscr{F} is symplectically self-dual. ∎

10.2.2. Proof of Theorem 1.1.2

We first explain the proof of Theorem 1.1.2. Let b=Spec𝔽qb=\operatorname{Spec}\mathbb{F}_{q}, and (C,U,Z,A[ν])(C,U,Z,A[\nu]) be our given symplectic sheaf data over bb as in Theorem 1.1.2. Let BB be a complete dvr with closed point bb and generic characteristic 0. By 10.2.2, we can realize (C,U,Z,A[ν])(C,U,Z,A[\nu]) as the restriction along bBb\to B of some symplectic sheaf data (CB,UB,ZB,B)(C_{B},U_{B},Z_{B},\mathscr{F}_{B}) on BB. Note that the hypotheses of Theorem 10.2.1 imply those of Theorem 1.1.2 as mentioned in the last paragraph of 7.1.4. Hence, Theorem 1.1.2 follows from Theorem 10.2.1. ∎

10.2.3. Proof of Theorem 1.1.3

As in the proof of Theorem 1.1.2 in § 10.2.2 above, we may lift all our symplectic sheaf data over 𝔽q\mathbb{F}_{q} to symplectic sheaf data over the spectrum of a complete dvr BB, with residue field 𝔽q\mathbb{F}_{q} and generic characteristic 0, using 10.2.2. To obtain (1.2) of Theorem 1.1.3, we note that Sym2H\operatorname{Sym}^{2}H is the HH-surjection moment of the BKLPR distribution by 2.3.1. Hence, (1.2) follows from Theorem 9.2.1, together with an inclusion-exclusion to show that points on a certain subset of the components of SelbnH\operatorname{Sel}_{\mathscr{F}^{n}_{b}}^{H} correspond to surjections onto HH, in place of all homomorphisms.

For establishing (1.3), we only need show that the limit

limnnevenxQTwistU/𝔽qn(𝔽qj)#Surj(Selν(Ax),H)xQTwistU/𝔽qn(𝔽qj)1\displaystyle\lim_{\begin{subarray}{c}n\to\infty\\ n\hskip 2.84544pt\mathrm{even}\end{subarray}}\frac{\sum_{x\in\operatorname{QTwist}^{n}_{U/\mathbb{F}_{q}}(\mathbb{F}_{q^{j}})}\#\operatorname{Surj}(\operatorname{Sel}_{\nu}(A_{x}),H)}{\sum_{x\in\operatorname{QTwist}^{n}_{U/\mathbb{F}_{q}}(\mathbb{F}_{q^{j}})}1}

exists, as then (1.2) yields what the limit as jj\to\infty of this value must be. The limit exists by (9.6), together with an inclusion-exclusion to show that points on a certain subset of the components of SelbnH\operatorname{Sel}_{\mathscr{F}^{n}_{b}}^{H} correspond to surjections onto HH, in place of all homomorphisms. ∎

10.2.4. Proof of Theorem 1.1.4

We next explain the proof of Theorem 1.1.4. Choose ν=\nu=\ell a prime as in Theorem 1.1.2. Note that this only excludes finitely many possibilities for \ell, so any sufficiently large \ell works. By Theorem 10.2.1, together with 10.2.2 as in § 10.2.2 above, we obtain equidistribution of the parity of the dimension of the \ell Selmer group in the quadratic twist family, since the BKLPR distribution predicts the parity of the rank of the \ell Selmer group of the abelian variety is even half the time and odd half the time. It follows from 7.4.2 that the parity of rkA\operatorname{rk}_{\ell^{\infty}}A agrees with the parity of the rank of Sel(A)\operatorname{Sel}_{\ell}(A). Therefore, the parity of rk\operatorname{rk}_{\ell^{\infty}} is also equidistributed.

To conclude the result, we only need to prove that the probability that \ell^{\infty} Selmer rank is 2\geq 2 is 0. It follows from Theorem 1.1.3 (and an inclusion exclusion to relate surjections to homomorphisms) that the average size of the ν\nu Selmer group is σνσ\sum_{\sigma\mid\nu}\sigma. Therefore, the same argument as in [BS13a, Proposition 5] (see also [PR12, p.246-247]) implies that the probability that the \ell^{\infty} Selmer rank is 2\geq 2 is 0. ∎

10.2.5. Proof of Theorem 1.1.2 in the special case ν=\nu=\ell

We will now give a somewhat shorter proof of Theorem 1.1.2 in the special case that ν\nu is a prime \ell. In particular, we need only the first lines of (8.5) and (8.6), and not the second lines of these equations. As explained in § 10.2.2, the case ν=\nu=\ell is all that is necessary for the application to Theorem 1.1.4. We sketch below how to handle this case for the convenience of those readers who are not in need of the full generality of Theorem 1.1.2.

The main difference when ν=\nu=\ell is that in this case it is easier to recover the distribution from the moments. In the proof of Theorem 1.1.2 we need to compute the moments of the variables XA[ν]𝔽qniX^{i}_{A[\nu]^{n}_{\mathbb{F}_{q}}} for i=0,1i=0,1; in other words, we need the moments of the mod \ell Selmer rank conditional on parity. But when ν=\nu=\ell, we can get by with less; by [Woo22, Thm 2.10, Cor 2.12], the distribution of Selmer ranks converges to the BKLPR distribution if the moments converge to the BKLPR moments, and if the parity of the mod \ell Selmer rank is equidistributed between odd and even. The former statement is what we have proved in (9.4).

It remains to show that the parity of the mod \ell Selmer rank is equidistributed. First of all, it follows from the discussion in 7.4.2 that the mod \ell Selmer rank of a quadratic twist AχA_{\chi} has the same parity as the \ell^{\infty}-Selmer rank of AχA_{\chi}. By [TY14, Theorem 1.1], this parity is determined by the root number W(Aχ)W(A_{\chi}). Let NAN_{A} denote the conductor of AA. By [Bis19, Cor. 6.12], we have

W(Aχ)=W(A)χ(NA).W(A_{\chi})=W(A)\chi(N_{A}).

So what remains is to show that the average of χ(NA)\chi(N_{A}) as χ\chi ranges over quadratic characters of discriminant BB approaches 0 as BB goes to infinity. This follows from [BSW15, Theorem 2] upon noting that the contribution to the local term m𝔭(Σ𝔭)m_{\mathfrak{p}}(\Sigma_{\mathfrak{p}}) in [BSW15, Theorem 2], for 𝔭\mathfrak{p} the specified point from (1.1), over which the cover is trivial, is equal to the local contribution from which the cover is étale but nontrivial, and χ(NA)\chi(N_{A}) has opposite signs in these two cases. ∎

Appendix A Frobenius equivariance
By Aaron Landesman

Throughout this section, we will use notation as in 5.1.4. Additionally, we will assume there exists some section σZ(B)\sigma\in Z(B) over which our symplectically self-dual sheaf \mathscr{F} on U=CZU=C-Z has trivial inertia at σ\sigma, and hence its pushforward along UCU\to C is lcc in a neighborhood of σ\sigma. The main result is Theorem A.5.1, which shows that the stabilization isomorphisms on the cohomology of Selmer spaces are equivariant for the action of Frobenius. The only part of our paper this appendix comes into play is to prove (1.3) (and (9.6) along the way). The consequence of this is that we prove (1.3), instead of only knowing that the lim inf\liminf and lim sup\limsup exist as in (1.2).

In order to prove Theorem A.5.1, we first set up notation to describe a compactified version of Selmer spaces in § A.1. Then, we introduce log structures and a logarithmic version of the stabilization map in § A.2. In § A.3, we show that we may take the topological stabilization map to have degree 22. Next, we show this logarithmic stabilization map agrees with the topological stabilization map in § A.4. Finally, in § A.5, we prove Theorem A.5.1.

We thank Dori Bejleri for suggesting that the general strategy taken here could work. We would also like to mention that the idea of viewing these sort of stabilization maps in algebraic geometry as coming from log geometry is not new. Variants have been studied in, for example [ACGS20], [Gro23], [HS23], [Par12], and [BDPW23].

A.1. Notation for the compactified selmer space

We next set up notation for a partially compactified version of Selmer spaces. First we define a partially compactified version of configuration space in § A.1.1, then we define a partially compactified version of the space of quadratic twists in § A.1.2, and finally we define a partially compactified version of the selmer space in § A.1.3.

A.1.1. Defining a partially compactified configuration space with sections

Let 𝒦n+f+1,g(C,1)\mathcal{K}_{n+f+1,g}(C,1) denote the moduli stack (which is in fact a scheme) of n+f+1n+f+1-pointed stable maps of degree 11 over BB to our given curve CC.

Remark A.1.1.

The above stack of stable maps parameterizes curves, one of whose components is CC, and all other components have genus 0, and are contracted under the map to CC. We next construct a locally closed substack of a quotient stack of this which corresponds to only allowing dd of the nn points to simultaneously collide with σ\sigma.

We suppose that σZ\sigma\subset Z and let U=CZU=C-Z as usual. Suppose that ZZ has connected components of degrees 1,f1,,fk1,f_{1},\ldots,f_{k}, the first 11 corresponding to σ\sigma, so that 1+(i=1kfi)=1+f=degZ1+(\sum_{i=1}^{k}f_{i})=1+f=\deg Z. There is an action of Sn×Sf1××SfkS_{n}\times S_{f_{1}}\times\cdots\times S_{f_{k}} on 𝒦n+f+1,g(C,1)\mathcal{K}_{n+f+1,g}(C,1) by permuting the n+f+1n+f+1 points. There is an evaluation map ev:[𝒦n+f+1,g(C,1)/Sn×Sf1××Sfk]C×[Cf1/Sf1]××[Cfk/Sfk]\operatorname{ev}:[\mathcal{K}_{n+f+1,g}(C,1)/S_{n}\times S_{f_{1}}\times\cdots\times S_{f_{k}}]\to C\times[C^{f_{1}}/S_{f_{1}}]\times\cdots\times[C^{f_{k}}/S_{f_{k}}]. Define StU/Bn,σ¯\overline{\operatorname{St}^{n,\sigma}_{U/B}} to be the fiber of the map ev\operatorname{ev} over the point of C×[Cf1/Sf1]××[Cfk/Sfk]C\times[C^{f_{1}}/S_{f_{1}}]\times\cdots\times[C^{f_{k}}/S_{f_{k}}] corresponding to the divisor ZZ.

We further fix an integer dd. There is an open substack StU/Bn,σStU/Bn,σ¯\operatorname{St}^{n,\sigma}_{U/B}\subset\overline{\operatorname{St}^{n,\sigma}_{U/B}} which is set theoretically supported on the locus where the universal curve is either irreducible (hence isomorphic to CC) or a union of CC and 1\mathbb{P}^{1} where a subset of dd of the nn points collide into the point σ\sigma. More precisely, this open substack can be described as the complement of the following divisors: first, the divisor parameterizing two points, neither of which is σ\sigma, colliding, and, second, the divisor where dd^{\prime} points collide with σ\sigma for ddd^{\prime}\neq d.

Remark A.1.2.

We observe that StU/Bn,σ\operatorname{St}^{n,\sigma}_{U/B} is smooth and the complement of ConfU/BnStU/Bn,σ\operatorname{Conf}^{n}_{U/B}\subset\operatorname{St}^{n,\sigma}_{U/B} is a smooth divisor. This follows from Theorem B.1.1 since one can use this to realize StU/Bn,σ\operatorname{St}^{n,\sigma}_{U/B} as an open in a compactification of ConfU/Bn\operatorname{Conf}^{n}_{U/B} with normal crossings boundary.

A.1.2. Defining a partially compactified space of quadratic twists with sections

We next define a space of quadratic twists over StU/Bn,σ\operatorname{St}^{n,\sigma}_{U/B} which we can think of as partially compactifying QTwistU/Bn\operatorname{QTwist}^{n}_{U/B} (though in actuality it will partially compactify a double cover of QTwistU/Bn\operatorname{QTwist}^{n}_{U/B}; the double cover corresponding to specifying a point in the universal double cover over σ\sigma). To generalize the Selmer space to stable curves, we use similar notation to our definition of Selmer space, but include the subscript St\operatorname{St} throughout.

Assume that ZZ is the disjoint union of multisections of degrees f0:=1,f1,,fkf_{0}:=1,f_{1},\ldots,f_{k} the first multi-section corresponding to a section σC(B)\sigma\in C(B). There is a universal schematic proper curve 𝒞St,Bn,σC×BStU/Bn,σStU/Bn,σ\mathscr{C}^{n,\sigma}_{\operatorname{St},B}\to C\times_{B}\operatorname{St}^{n,\sigma}_{U/B}\to\operatorname{St}^{n,\sigma}_{U/B}. There is a universal degree nn divisor 𝒟St,Bn,σ𝒞St,Bn,σ\mathscr{D}^{n,\sigma}_{\operatorname{St},B}\subset\mathscr{C}^{n,\sigma}_{\operatorname{St},B}.

Remark A.1.3.

The first map 𝒞St,Bn,σC×BStU/Bn,σ\mathscr{C}^{n,\sigma}_{\operatorname{St},B}\to C\times_{B}\operatorname{St}^{n,\sigma}_{U/B} is an isomorphism over ConfU/BnStU/Bn,σ\operatorname{Conf}^{n}_{U/B}\subset\operatorname{St}^{n,\sigma}_{U/B}, but in general may contain additional genus 0 fibers corresponding to locations where dd of the nn points collide with σ\sigma.

We next define an extension of a variant of the Selmer sheaf over StU/Bn,σ\operatorname{St}^{n,\sigma}_{U/B}. The informal idea is that QTwistSt,U/Bn,σ\operatorname{QTwist}^{n,\sigma}_{\operatorname{St},U/B} parameterizes double covers of curves with a degree 11 stable map to CC, branched over a degree nn divisor in the smooth locus of the nodal curve, together with a trivialization of this double cover at σ\sigma. We now give a more formal definition.

Let [C/(/2)]=C×B(/2)[C/(\mathbb{Z}/2\mathbb{Z})]=C\times B(\mathbb{Z}/2\mathbb{Z}) denote the stack quotient of CC by the trivial /2\mathbb{Z}/2\mathbb{Z} action. (Recall we are assuming 22 is invertible on BB.) Next, QTwistSt,U/Bn,σ\operatorname{QTwist}^{n,\sigma}_{\operatorname{St},U/B} can be constructed from 𝒦n+f+1,g([C/(/2)],1)\mathcal{K}_{n+f+1,g}([C/(\mathbb{Z}/2\mathbb{Z})],1) much the same way StU/Bn,σ\operatorname{St}^{n,\sigma}_{U/B} was constructed from 𝒦g,n+f+1(C,1)\mathcal{K}_{g,n+f+1}(C,1). Here, 𝒦n+f+1,g([C/(/2)],1)\mathcal{K}_{n+f+1,g}([C/(\mathbb{Z}/2\mathbb{Z})],1) denotes twisted stable maps from a genus gg twisted curve 𝒳\mathcal{X} with n+f+1n+f+1 marked points to [C/(/2)][C/(\mathbb{Z}/2\mathbb{Z})] such that the composition to the coarse space 𝒳[C/(/2)]C\mathcal{X}\to[C/(\mathbb{Z}/2\mathbb{Z})]\to C has degree 11, in the sense that the line bundle 𝒪C(σ)\mathscr{O}_{C}(\sigma) on CC pulls back to a degree 11 line bundle on 𝒳\mathcal{X}. Namely, we first form the quotient of 𝒦g,n+f+1([C/(/2)],2)\mathcal{K}_{g,n+f+1}([C/(\mathbb{Z}/2\mathbb{Z})],2) by the action of Sn×Sf1××SfkS_{n}\times S_{f_{1}}\times\cdots\times S_{f_{k}}. We next construct the fiber of the evaluation map ev:[𝒦g,n+f+1([C/(/2)],2)/Sn×Sf1××Sfk]C×[Cf1/Sf1]××[Cfk/Sfk]\operatorname{ev}:[\mathcal{K}_{g,n+f+1}([C/(\mathbb{Z}/2\mathbb{Z})],2)/S_{n}\times S_{f_{1}}\times\cdots\times S_{f_{k}}]\to C\times[C^{f_{1}}/S_{f_{1}}]\times\cdots\times[C^{f_{k}}/S_{f_{k}}] over the point corresponding to the divisor ZZ. We let QTwistSt,U/Bn,σ¯\overline{\operatorname{QTwist}^{n,\sigma}_{\operatorname{St},U/B}} be the double cover of this fiber, obtained by specifying a point in the fiber of the double cover over the pullback of σ\sigma. We let QTwistSt,U/Bn,σ\operatorname{QTwist}^{n,\sigma}_{\operatorname{St},U/B} denote the open substack parameterizing double covers which are balanced in the sense of [ACV03, §2.1.3], which also map to StU/Bn,σStU/Bn,σ¯\operatorname{St}^{n,\sigma}_{U/B}\subset\overline{\operatorname{St}^{n,\sigma}_{U/B}}. We define QTwistU/Bn,σ:=QTwistSt,U/Bn,σ×StU/Bn,σConfU/Bn\operatorname{QTwist}^{n,\sigma}_{U/B}:=\operatorname{QTwist}^{n,\sigma}_{\operatorname{St},U/B}\times_{\operatorname{St}^{n,\sigma}_{U/B}}\operatorname{Conf}^{n}_{U/B}.

It will be useful to additionally specify a slight variant of the above construction, where one marks 22 (or more) sections, instead of just a single section. Namely, if σ1,,σt\sigma_{1},\ldots,\sigma_{t} are tt sections in Z(B)C(B)Z(B)\subset C(B), we use QTwistSt,U/Bn,σ1,,σt\operatorname{QTwist}^{n,\sigma_{1},\ldots,\sigma_{t}}_{\operatorname{St},U/B} to denote the analogous construction, but where one additionally marks a point of the double cover over each of σ1,,σt\sigma_{1},\ldots,\sigma_{t}. In particular, QTwistSt,U/Bn,σ1,,σt\operatorname{QTwist}^{n,\sigma_{1},\ldots,\sigma_{t}}_{\operatorname{St},U/B} is a finite étale cover of degree 2t12^{t-1} over QTwistSt,U/Bn,σ1\operatorname{QTwist}^{n,\sigma_{1}}_{\operatorname{St},U/B}.

Remark A.1.4.

In what follows, we will only apply this construction with multiple sections in the case C=1C=\mathbb{P}^{1}, t=2t=2, and {σ1,σ2}={0,}\{\sigma_{1},\sigma_{2}\}=\{0,\infty\}.

A.1.3. Defining a partially compactified Selmer sheaf

Using the description of twisted stable maps, there is a universal schematic curve Bn,σ:=𝒞St,Bn,σ×StU/Bn,σQTwistSt,U/Bn,σ\mathscr{R}^{n,\sigma}_{B}:=\mathscr{C}^{n,\sigma}_{\operatorname{St},B}\times_{\operatorname{St}^{n,\sigma}_{U/B}}\operatorname{QTwist}^{n,\sigma}_{\operatorname{St},U/B} over QTwistSt,U/Bn,σ\operatorname{QTwist}^{n,\sigma}_{\operatorname{St},U/B} with a finite degree 22 cover branched over the universal degree nn divisor 𝒟St,Bn×StU/Bn,σQTwistSt,U/Bn,σ\mathscr{D}^{n}_{\operatorname{St},B}\times_{\operatorname{St}^{n,\sigma}_{U/B}}\operatorname{QTwist}^{n,\sigma}_{\operatorname{St},U/B}. There is also an universal evaluation map hSt:Bn,σCh_{\operatorname{St}}:\mathscr{R}^{n,\sigma}_{B}\to C coming from the definition of stable maps to CC. Let j:UCj:U\to C denote the inclusion. Define St,σ:=hSt(j)\mathscr{F}_{\operatorname{St},\sigma}:=h_{\operatorname{St}}^{*}(j_{*}\mathscr{F}) to be the resulting étale sheaf on Bn,σ\mathscr{R}^{n,\sigma}_{B}.

Remark A.1.5.

In some sense the following alternate definition of St,σ\mathscr{F}_{\operatorname{St},\sigma} might be preferred, because it will also work when there is nontrivial ramification of \mathscr{F} at σ\sigma. However, the following construction will agree with our construction above when \mathscr{F} has trivial ramification at σ\sigma, which is the only case we will need.

Let j:U×BQTwistSt,U/Bn,σBnj:U\times_{B}\operatorname{QTwist}^{n,\sigma}_{\operatorname{St},U/B}\to\mathscr{R}^{n}_{B} denote the open immersion and let πU:U×BQTwistSt,U/Bn,σU\pi_{U}:U\times_{B}\operatorname{QTwist}^{n,\sigma}_{\operatorname{St},U/B}\to U denote the projection. Then, one may alternatively define St,σ\mathscr{F}_{\operatorname{St},\sigma} to be j(πU)j_{*}(\pi_{U}^{*}\mathscr{F}). As mentioned above, this can be shown to agree with hSth_{\operatorname{St}}^{*}\mathscr{F} when \mathscr{F} is unramified over σ\sigma, but we will not need this fact.

We believe it would be quite interesting to work out the analog of Theorem A.5.1 for the above mentioned generalization. In particular, it would have the application mentioned in 1.1.5.

Let χBn\chi^{n}_{B} denote the nontrivial rank 11 local system on Bn,σ\mathscr{R}^{n,\sigma}_{B} which is trivialized on the universal double cover of Bn,σ\mathscr{R}^{n,\sigma}_{B}. Define St,Bn,σ:=χBnSt,Bσ\mathscr{F}^{n,\sigma}_{\operatorname{St},B}:=\chi^{n}_{B}\otimes\mathscr{F}^{\sigma}_{\operatorname{St},B}, which can be thought of as the universal quadratic twist of jSt,Bσ.j_{*}\mathscr{F}^{\sigma}_{\operatorname{St},B}. Define SelSt,Bn,σ\operatorname{Sel}_{\mathscr{F}^{n,\sigma}_{\operatorname{St},B}} to be the algebraic space represented by the étale sheaf parameterizing torsors for St,Bn\mathscr{F}^{n}_{\operatorname{St},B} over Bn,σ\mathscr{R}^{n,\sigma}_{B} together with a trivialization of the torsor over Bn,σ\mathscr{R}^{n,\sigma}_{B} at the section σ\sigma. We also use SelBn,σ\operatorname{Sel}_{\mathscr{F}^{n,\sigma}_{B}} for the restriction of SelSt,Bn,σ\operatorname{Sel}_{\mathscr{F}^{n,\sigma}_{\operatorname{St},B}} along the open immersion ConfU/BnStU/Bn,σ\operatorname{Conf}^{n}_{U/B}\subset\operatorname{St}^{n,\sigma}_{U/B}. Note that the fiber (j)σ(j_{*}\mathscr{F})_{\sigma} is finite étale of degree νrk\nu^{\operatorname{rk}\mathscr{F}} over σ\sigma by the assumption that the inertia of \mathscr{F} at σ\sigma is trivial.

We also define St,Bn,σ1,,σt\mathscr{F}^{n,\sigma_{1},\ldots,\sigma_{t}}_{\operatorname{St},B} and SelSt,Bn,σ1,,σt\operatorname{Sel}_{\mathscr{F}^{n,\sigma_{1},\ldots,\sigma_{t}}_{\operatorname{St},B}} over QTwistSt,U/Bn,σ1,,σt\operatorname{QTwist}^{n,\sigma_{1},\ldots,\sigma_{t}}_{\operatorname{St},U/B} as the pullbacks of St,Bn,σ1\mathscr{F}^{n,\sigma_{1}}_{\operatorname{St},B} and SelSt,Bn,σ1\operatorname{Sel}_{\mathscr{F}^{n,\sigma_{1}}_{\operatorname{St},B}} along QTwistSt,U/Bn,σ1,,σtQTwistSt,U/Bn,σ1\operatorname{QTwist}^{n,\sigma_{1},\ldots,\sigma_{t}}_{\operatorname{St},U/B}\to\operatorname{QTwist}^{n,\sigma_{1}}_{\operatorname{St},U/B}. We define Bn,σ1,,σn\mathscr{F}^{n,\sigma_{1},\ldots,\sigma_{n}}_{B} and SelBn,σ1,,σn\operatorname{Sel}_{\mathscr{F}^{n,\sigma_{1},\ldots,\sigma_{n}}_{B}} as the further restrictions to QTwistU/Bn,σ1,,σtQTwistSt,U/Bn,σ1,,σt\operatorname{QTwist}^{n,\sigma_{1},\ldots,\sigma_{t}}_{U/B}\subset\operatorname{QTwist}^{n,\sigma_{1},\ldots,\sigma_{t}}_{\operatorname{St},U/B}

Remark A.1.6.

We will only need this variant with t>1t>1 in the case t=2,C=1,{σ1,σ2}={0,}t=2,C=\mathbb{P}^{1},\{\sigma_{1},\sigma_{2}\}=\{0,\infty\} and =𝒢\mathscr{F}=\mathscr{G} is a trivial sheaf on 𝔾m\mathbb{G}_{m}. Note that the sheaf is only trivialized at the first marked section, so the universal sheaf over SelBn,0,\operatorname{Sel}_{\mathscr{F}^{n,0,\infty}_{B}} is trivialized at 0 while the universal sheaf over SelBn,,0\operatorname{Sel}_{\mathscr{F}^{n,\infty,0}_{B}} is trivialized at \infty.

A.2. The gluing map with log structures

In this subsection, we define the gluing map, which joins the Selmer space of degree ndn-d over CC with a trivialization at pp to the Selmer space of degree dd over 1\mathbb{P}^{1} with a trivialization at 0 and \infty, and sends it to the partially compactified Selmer space of degree nn with a trivialization at pp. We first define the gluing map in § A.2.1 and A.2.1. We then briefly review relevant parts of logarithmic algebraic geometry in § A.2.2. In A.2.2, we define version of the gluing map for logarithmic stacks.

A.2.1. Defining the gluing map

Fix an even positive integer dd and let 𝔾m1\mathbb{G}_{m}\subset\mathbb{P}^{1} over our base BB denote the complement of the sections :B1\infty:B\to\mathbb{P}^{1} and 0:B10:B\to\mathbb{P}^{1}. For even n>dn>d, there is a gluing map Δ:Conf𝔾m/Bd×ConfU/BndStU/Bn,σ\Delta:\operatorname{Conf}^{d}_{\mathbb{G}_{m}/B}\times\operatorname{Conf}^{n-d}_{U/B}\to\operatorname{St}^{n,\sigma}_{U/B} which glues the 0 section on 1\mathbb{P}^{1} to the specified section σC(B)\sigma\in C(B).

Over Δ\Delta, there is another gluing map

(A.1) Γ:QTwist𝔾m/Bd,0,×QTwistU/Bnd,σQTwistSt,U/Bn,σ\displaystyle\Gamma:\operatorname{QTwist}^{d,0,\infty}_{\mathbb{G}_{m}/B}\times\operatorname{QTwist}^{n-d,\sigma}_{U/B}\to\operatorname{QTwist}^{n,\sigma}_{\operatorname{St},U/B}

which we define next. A point of QTwistU/Bnd,σ\operatorname{QTwist}^{n-d,\sigma}_{U/B} can be described as [XC,z][X\to C,z], for XCX\to C a double cover unramified over σ\sigma, and zz a choice of point in the preimage of σ\sigma. We use zz^{\prime} to denote the remaining section of XX in the preimage of σ\sigma. A point of QTwist𝔾m/Bd,0,\operatorname{QTwist}^{d,0,\infty}_{\mathbb{G}_{m}/B} can be described as [Y1,v,s][Y\to\mathbb{P}^{1},v,s] where Y1Y\to\mathbb{P}^{1} is a double cover, vv is a choice of point in the preimage of 0, and ss is a choice of point over \infty. We also use vv^{\prime} to denote the remaining section of YY over 0. The map is then given by gluing vv to zz and gluing vv^{\prime} to zz^{\prime} to obtain a double cover [Xvz,vzYCσ01,s][X\coprod_{v\sim z,v^{\prime}\sim z^{\prime}}Y\to C\coprod_{\sigma\sim 0}\mathbb{P}^{1},s], which we view as a point of QTwistSt,U/Bn,σ\operatorname{QTwist}^{n,\sigma}_{\operatorname{St},U/B}, for ss the choice of section over the marked section 1Cσ01\infty\in\mathbb{P}^{1}\subset C\coprod_{\sigma\sim 0}\mathbb{P}^{1}. (We note here that the universal section σ\sigma on Bn,σ\mathscr{R}^{n,\sigma}_{B} restricts to 1\infty\in\mathbb{P}^{1}.) Via the description above, the map Γ\Gamma from (A.1) is induced by a gluing map on the universal curves

Θ:(C×BQTwistU/Bnd,σ)×(B1×BQTwist𝔾m/Bd,0,)Bn,σ\displaystyle\Theta:(C\times_{B}\operatorname{QTwist}^{n-d,\sigma}_{U/B})\times(\mathbb{P}^{1}_{B}\times_{B}\operatorname{QTwist}^{d,0,\infty}_{\mathbb{G}_{m}/B})\to\mathscr{R}^{n,\sigma}_{B}

together with a gluing map on double covers of these universal curves (which we will not need to distinguish with further notation). We also define 𝒢\mathscr{G} to be the constant sheaf of /ν\mathbb{Z}/\nu\mathbb{Z} modules of rank 2r=rk2r=\operatorname{rk}\mathscr{F} on 𝔾m1\mathbb{G}_{m}\subset\mathbb{P}^{1}. There are maps

Θ\displaystyle\Theta^{\prime} :(B1×BSel𝒢Bd,0,)×B(C×BQTwistU/Bnd,σ)Bn,σ\displaystyle:(\mathbb{P}^{1}_{B}\times_{B}\operatorname{Sel}_{\mathscr{G}^{d,0,\infty}_{B}})\times_{B}(C\times_{B}\operatorname{QTwist}^{n-d,\sigma}_{U/B})\to\mathscr{R}^{n,\sigma}_{B}
Γ\displaystyle\Gamma^{\prime} :Sel𝒢Bd,0,×BQTwistU/Bnd,σQTwistSt,U/Bn,σ\displaystyle:\operatorname{Sel}_{\mathscr{G}^{d,0,\infty}_{B}}\times_{B}\operatorname{QTwist}^{n-d,\sigma}_{U/B}\to\operatorname{QTwist}^{n,\sigma}_{\operatorname{St},U/B}

where Θ\Theta^{\prime} and Γ\Gamma^{\prime} are obtained from Θ\Theta and Γ\Gamma by precomposing these with the projection Sel𝒢Bd,0,QTwist𝔾m/Bd,0,\operatorname{Sel}_{\mathscr{G}^{d,0,\infty}_{B}}\to\operatorname{QTwist}^{d,0,\infty}_{\mathbb{G}_{m}/B}. Define the projections

πC\displaystyle\pi_{C} :Sel𝒢Bd,0,×BQTwistU/Bnd,σQTwistU/Bnd,σ\displaystyle:\operatorname{Sel}_{\mathscr{G}^{d,0,\infty}_{B}}\times_{B}\operatorname{QTwist}^{n-d,\sigma}_{U/B}\to\operatorname{QTwist}^{n-d,\sigma}_{U/B}
π1\displaystyle\pi_{\mathbb{P}^{1}} :Sel𝒢Bd,0,×BQTwistU/Bnd,σSel𝒢Bd,0,.\displaystyle:\operatorname{Sel}_{\mathscr{G}^{d,0,\infty}_{B}}\times_{B}\operatorname{QTwist}^{n-d,\sigma}_{U/B}\to\operatorname{Sel}_{\mathscr{G}^{d,0,\infty}_{B}}.
Lemma A.2.1.

There is an isomorphism of lcc étale sheaves on Sel𝒢Bd,0,×QTwistU/Bnd,σ\operatorname{Sel}_{\mathscr{G}^{d,0,\infty}_{B}}\times\operatorname{QTwist}^{n-d,\sigma}_{U/B},

(A.2) πC𝒮eSt,Bnd,σπ1𝒮e𝒢Bd,,0(Γ)𝒮eSt,Bn,σ.\displaystyle\pi_{C}^{*}{\mathcal{S}e\ell}_{\mathscr{F}^{n-d,\sigma}_{\operatorname{St},B}}\oplus\pi_{\mathbb{P}^{1}}^{*}{\mathcal{S}e\ell}_{\mathscr{G}^{d,\infty,0}_{B}}\simeq(\Gamma^{\prime})^{*}{\mathcal{S}e\ell}_{\mathscr{F}^{n,\sigma}_{\operatorname{St},B}}.
Proof.

First note that (Θ)St,Bσ(Θ)hSt,B(j)(\Theta^{\prime})^{*}\mathscr{F}^{\sigma}_{\operatorname{St},B}\simeq(\Theta^{\prime})^{*}h_{\operatorname{St},B}^{*}(j_{*}\mathscr{F}) by definition of St,Bσ\mathscr{F}^{\sigma}_{\operatorname{St},B} as hSt(j)h_{\operatorname{St}}^{*}(j_{*}\mathscr{F}). Then, (Γ)𝒮eSt,Bn,σ(\Gamma)^{*}{\mathcal{S}e\ell}_{\mathscr{F}^{n,\sigma}_{\operatorname{St},B}} is a universal torsor for quadratic twists of (Θ)hSt(j)(\Theta^{\prime})^{*}h_{\operatorname{St}}^{*}(j_{*}\mathscr{F}) together with a trivialization over 1Cσ0B1\infty\in\mathbb{P}^{1}\subset C\coprod_{\sigma\sim 0}\mathbb{P}^{1}_{B} (since \infty is the pullback of the section corresponding to σ\sigma over StU/Bn,σ\operatorname{St}^{n,\sigma}_{U/B}). It follows that (Γ)𝒮eSt,Bn,σ(\Gamma^{\prime})^{*}{\mathcal{S}e\ell}_{\mathscr{F}^{n,\sigma}_{\operatorname{St},B}} which is the pullback of (Γ)𝒮eSt,Bn,σ(\Gamma)^{*}{\mathcal{S}e\ell}_{\mathscr{F}^{n,\sigma}_{\operatorname{St},B}} along Sel𝒢Bd,0,QTwist𝔾m/Bd,0,\operatorname{Sel}_{\mathscr{G}^{d,0,\infty}_{B}}\to\operatorname{QTwist}^{d,0,\infty}_{\mathbb{G}_{m}/B}, is a universal torsor for quadratic twists of (Θ)hSt(j)(\Theta^{\prime})^{*}h_{\operatorname{St}}^{*}(j_{*}\mathscr{F}) which has a specified trivialization over 1Cσ0B1\infty\in\mathbb{P}^{1}\subset C\coprod_{\sigma\sim 0}\mathbb{P}^{1}_{B}. This sheaf is also trivializable at 0 because we pulled it back to Sel𝒢Bd,0,\operatorname{Sel}_{\mathscr{G}^{d,0,\infty}_{B}}, although no trivialization is specified at 0. We can choose an étale path so as to make an identification of the fiber of \mathscr{F} over 0 and the fiber over \infty, and thereby transfer the trivialization at \infty to a trivialization at 0. Specifying a torsor for a quadratic twist of (Θ)St,Bσ(\Theta^{\prime})^{*}\mathscr{F}^{\sigma}_{\operatorname{St},B} which is trivialized in this way at σ0\sigma\sim 0 is equivalent to specifying a torsor for the restriction of the quadratic twist to 1\mathbb{P}^{1} with a trivialization at \infty, together with a torsor for the restriction to CC with a trivialization at σ\sigma. In other words, this gives the desired isomorphism (A.2). ∎

A.2.2. Background on logarithmic geometry

For a general reference on cohomology of log schemes, we recommend [Ill02]. We will work with a subcategory of log stacks called Deligne-Faltings log stacks. Their basic properties are described in [BDPW23, §8]. We next recall notation for Deligne-Faltings log stacks (which we henceforth simply refer to as log stacks), closely following [BDPW23, §8.2].

For XX a Deligne-Mumford stack, a log structure 𝔏=(σi:𝒪XLi)i=1k\mathfrak{L}=\left(\sigma_{i}:\mathscr{O}_{X}\to L_{i}\right)_{i=1}^{k} is a kk-tuple of invertible sheaves on XX with sections. If 𝔐\mathfrak{M} and 𝔏\mathfrak{L} are two log structures on XX, a morphism of log structures 𝔏:=(σi:𝒪XLi)i=1k𝔐:=(τj:𝒪XMi)j=1l\mathfrak{L}:=\left(\sigma_{i}:\mathscr{O}_{X}\to L_{i}\right)_{i=1}^{k}\to\mathfrak{M}:=\left(\tau_{j}:\mathscr{O}_{X}\to M_{i}\right)_{j=1}^{l} is a tuple of nonnegative integers eij,1ik,1jle_{ij},1\leq i\leq k,1\leq j\leq l and isomorphisms ϕi:Lij=1lMjeij\phi_{i}:L_{i}\simeq\otimes_{j=1}^{l}M_{j}^{\otimes e_{ij}} so that ϕiσi=j=1lτjeij\phi_{i}\circ\sigma_{i}=\otimes_{j=1}^{l}\tau_{j}^{e_{ij}}. A morphism of log stacks (X,𝔏)(Y,𝔐)(X,\mathfrak{L})\to(Y,\mathfrak{M}) is a morphism of stacks f:XYf:X\to Y together with a morphism of log structures f𝔐𝔏f^{*}\mathfrak{M}\to\mathfrak{L}. Here if 𝔐:=(τj:𝒪YMi)j=1l\mathfrak{M}:=\left(\tau_{j}:\mathscr{O}_{Y}\to M_{i}\right)_{j=1}^{l}, f𝔐f^{*}\mathfrak{M} denotes (fτj:𝒪XfMi)j=1l\left(f^{*}\tau_{j}:\mathscr{O}_{X}\to f^{*}M_{i}\right)_{j=1}^{l}. A morphism of log structures is strict if the map f𝔐𝔏f^{*}\mathfrak{M}\to\mathfrak{L} is an isomorphism of log structures.

We will only need three types of log structures: the log structure defined by a divisor, the standard log structure, and the trivial log structure, and we define these three log structures next. For XX a stack and DXD\subset X a Cartier divisor, the log structure defined by DD corresponds to the line bundle 𝒪X(D)\mathscr{O}_{X}(D) with the tautological section 𝒪X𝒪X(D)\mathscr{O}_{X}\to\mathscr{O}_{X}(D). The standard log structure corresponds to the trivial line bundle 𝒪X\mathscr{O}_{X} with the 0 section. The trivial log structure corresponds to no line bundles, meaning that k=0k=0 in the definition of log structure above.

A.2.3. Defining our log stacks

We now define the log stacks we will work with. Let Sel\operatorname{Sel}_{\mathscr{H}} denote the stack associated to the sheaf 𝒮e{\mathcal{S}e\ell}_{\mathscr{H}}. Let ESelSt,Bn,σE\subset\operatorname{Sel}_{\mathscr{F}^{n,\sigma}_{\operatorname{St},B}} denote the divisor associated to the preimage of StU/Bn,σConfU/BnStU/Bn,σ\operatorname{St}^{n,\sigma}_{U/B}-\operatorname{Conf}^{n}_{U/B}\subset\operatorname{St}^{n,\sigma}_{U/B}. Let SelSt,Bn,σlog\operatorname{Sel}^{\operatorname{log}}_{\mathscr{F}^{n,\sigma}_{\operatorname{St},B}} denote the log stack with underlying space SelSt,Bn,σ\operatorname{Sel}_{\mathscr{F}^{n,\sigma}_{\operatorname{St},B}} with the log structure defined by the divisor EE, as in § A.2.2. Define a log scheme ((Sel𝒢Bd,,0×QTwist𝔾m/Bd,0,Sel𝒢Bd,0,)×BSelBnd,σ)log\left(\left(\operatorname{Sel}_{\mathscr{G}^{d,\infty,0}_{B}}\times_{\operatorname{QTwist}^{d,0,\infty}_{\mathbb{G}_{m}/B}}\operatorname{Sel}_{\mathscr{G}^{d,0,\infty}_{B}}\right)\times_{B}\operatorname{Sel}_{\mathscr{F}^{n-d,\sigma}_{B}}\right)^{\operatorname{log}} with underlying scheme

(A.3) (Sel𝒢Bd,,0×QTwist𝔾m/Bd,0,Sel𝒢Bd,0,)×BSelBnd,σ\displaystyle\left(\operatorname{Sel}_{\mathscr{G}^{d,\infty,0}_{B}}\times_{\operatorname{QTwist}^{d,0,\infty}_{\mathbb{G}_{m}/B}}\operatorname{Sel}_{\mathscr{G}^{d,0,\infty}_{B}}\right)\times_{B}\operatorname{Sel}_{\mathscr{F}^{n-d,\sigma}_{B}}

with the standard log structure, as defined in § A.2.2.

Lemma A.2.2.

Suppose BB is the spectrum of a complete dvr (or, more generally, has trivial Picard group). The isomorphism of A.2.1 yields a strict map of log stacks

(A.4) α:((Sel𝒢Bd,,0×QTwist𝔾m/Bd,0,Sel𝒢Bd,0,)×BSelBnd,σ)logSelSt,Bn,σlog.\displaystyle\alpha:\left(\left(\operatorname{Sel}_{\mathscr{G}^{d,\infty,0}_{B}}\times_{\operatorname{QTwist}^{d,0,\infty}_{\mathbb{G}_{m}/B}}\operatorname{Sel}_{\mathscr{G}^{d,0,\infty}_{B}}\right)\times_{B}\operatorname{Sel}_{\mathscr{F}^{n-d,\sigma}_{B}}\right)^{\operatorname{log}}\to\operatorname{Sel}^{\operatorname{log}}_{\mathscr{F}^{n,\sigma}_{\operatorname{St},B}}.
Proof.

The map of underlying stacks of these log stacks is obtained from A.2.1. Note that ατ\alpha^{*}\tau is the zero section because τ\tau vanishes on the image of α\alpha. It remains to show that the line bundle 𝒪SelSt,Bn,σ(E)\mathscr{O}_{\operatorname{Sel}_{\mathscr{F}^{n,\sigma}_{\operatorname{St},B}}}(E) pulls back to the trivial bundle.

First, we define another line bundle \mathscr{L} on (A.3). Then, we show \mathscr{L} is isomorphic to the trivial bundle on (A.3). Finally, we will show 𝒪SelSt,Bn,σ(E)\mathscr{O}_{\operatorname{Sel}_{\mathscr{F}^{n,\sigma}_{\operatorname{St},B}}}(E) pulls back to \mathscr{L}.

Let π1\pi_{1} and π2\pi_{2} denote the two projections from (A.3) onto its two factors. Let 𝕋σ\mathbb{T}_{\sigma} denote the line bundle on SelBnd,σ\operatorname{Sel}_{\mathscr{F}^{n-d,\sigma}_{B}} which is the restriction to σ\sigma of the relative tangent bundle for the universal curve over SelBnd,σ\operatorname{Sel}_{\mathscr{F}^{n-d,\sigma}_{B}}. Similarly, let 𝕋0\mathbb{T}_{0} denote the line bundle on (Sel𝒢Bd,,0×QTwist𝔾m/Bd,0,Sel𝒢Bd,0,)\left(\operatorname{Sel}_{\mathscr{G}^{d,\infty,0}_{B}}\times_{\operatorname{QTwist}^{d,0,\infty}_{\mathbb{G}_{m}/B}}\operatorname{Sel}_{\mathscr{G}^{d,0,\infty}_{B}}\right) which is the restriction to 0 of the relative tangent bundle for the universal curve over (Sel𝒢Bd,,0×QTwist𝔾m/Bd,0,Sel𝒢Bd,0,)\left(\operatorname{Sel}_{\mathscr{G}^{d,\infty,0}_{B}}\times_{\operatorname{QTwist}^{d,0,\infty}_{\mathbb{G}_{m}/B}}\operatorname{Sel}_{\mathscr{G}^{d,0,\infty}_{B}}\right) which is pulled back from the universal curve over Sel𝒢Bd,0,\operatorname{Sel}_{\mathscr{G}^{d,0,\infty}_{B}}. Then, define :=π1𝕋0π2𝕋σ\mathscr{L}:=\pi_{1}^{*}\mathbb{T}_{0}\otimes\pi_{2}^{*}\mathbb{T}_{\sigma} and take τ\tau to be the zero section of this line bundle.

We claim that \mathscr{L} is isomorphic to the trivial bundle. It suffices to show both 𝕋0\mathbb{T}_{0} and 𝕋σ\mathbb{T}_{\sigma} are trivial. These are pulled back from a line bundle on BB since the sections σ\sigma and 0 are pulled back from CC. Hence, these bundles are trivial because BB has trivial Picard group.

It remains to show 𝒪SelSt,Bn,σ𝒪SelSt,Bn,σ(E)\mathscr{O}_{\operatorname{Sel}_{\mathscr{F}^{n,\sigma}_{\operatorname{St},B}}}\to\mathscr{O}_{\operatorname{Sel}_{\mathscr{F}^{n,\sigma}_{\operatorname{St},B}}}(E) pulls back to the zero section of π1𝕋0π2𝕋σ\pi_{1}^{*}\mathbb{T}_{0}\otimes\pi_{2}^{*}\mathbb{T}_{\sigma}. This can be proven via an argument analogous to [ACG11, p. 346, line 2], as we next further expand on. A minor technicality is that the gluing map α\alpha joining CC and 1\mathbb{P}^{1} factors as the composition of a positive dimensional smooth map of relative dimension 11 and an étale map. Namely, it factors through a gluing map joining CC and PP, where PP is a genus 0 curve with 0 and \infty marked. Because a third point is not marked on PP, PP may not be isomorphic to 1\mathbb{P}^{1}. The latter gluing map, described by gluing CC to PP, does define an étale map to SelSt,Bn,σlog\operatorname{Sel}^{\operatorname{log}}_{\mathscr{F}^{n,\sigma}_{\operatorname{St},B}}. Since the normal bundle of an étale morphism is identified with the pullback of the ideal sheaf of its image, an argument similar to [ACG11, p. 346, line 2] shows that 𝒪SelSt,Bn,σ(E)\mathscr{O}_{\operatorname{Sel}_{\mathscr{F}^{n,\sigma}_{\operatorname{St},B}}}(E) pulls back to the tensor product of the tangent line bundles on CC and the genus 0 curve. When one further pulls this line bundle back to (Sel𝒢Bd,,0×QTwist𝔾m/Bd,0,Sel𝒢Bd,0,)×BSelBnd,σ\left(\operatorname{Sel}_{\mathscr{G}^{d,\infty,0}_{B}}\times_{\operatorname{QTwist}^{d,0,\infty}_{\mathbb{G}_{m}/B}}\operatorname{Sel}_{\mathscr{G}^{d,0,\infty}_{B}}\right)\times_{B}\operatorname{Sel}_{\mathscr{F}^{n-d,\sigma}_{B}}, we obtain the claim. ∎

A.3. A fun, combinatorial group theory interlude

Taking a break from the heavy machinery of log geometry, we will need a result from combinatorial group theory which strengthens [EVW16, Lemma 3.5]. This will show the degree of 𝕌\mathbb{U} may be taken to be 22 in our setting above.

Let GG be a group and cGc\subset G be a conjugacy class. Recall that (G,c)(G,c) is non-splitting if, for any subgroup KGK\subset G, cKc\cap K is either empty or a single conjugacy class. I.e., cc does not split into multiple conjugacy classes. Consider the coefficient system VnV_{n} for Σ0,01\Sigma^{1}_{0,0}, as in 3.1.9, associated to the group GG, and cGc\subset G a specified conjugacy class. We use RVR^{V} to denote n0H0(B0,0n,Vn)\oplus_{n\geq 0}H_{0}(B^{n}_{0,0},V_{n}), rhr_{h} denote right multiplication by hh on RVR^{V}, and ord(h)\operatorname{ord}(h) to denote the order of hh.

Proposition A.3.1.

Let (G,c)(G,c) be non-splitting and DD any positive integer. Then 𝕌:=hcrhDord(h)RV\mathbb{U}:=\sum_{h\in c}r_{h}^{D\operatorname{ord}(h)}\in R^{V} is a homogeneous central element with finite degree kernel and cokernel. Hence, 𝕌\mathbb{U} satisfies the hypotheses of Theorem 4.2.2.

Proof.

In [EVW16, Lemma 3.5] it was shown that there exists some integer DD so that 𝕌D:=hcrhDord(h)RV\mathbb{U}_{D}:=\sum_{h\in c}r_{h}^{D\operatorname{ord}(h)}\in R^{V} satisfies the conclusion of the theorem statement. We want to show DD can be taken to be any positive integer. Let St(K)S_{t}(K) denote the subset of quotient set ct/Btc^{t}/B_{t}, via the standard braid group action of BtB_{t} on tt-tuples of elements in cc, consisting of those tt-tuples of elements which generate KK. The proof of [EVW16, Lemma 3.5] shows that we may take DD to be any positive integer so that for every subgroup KGK\subset G, and for tt sufficiently large, the map rhord(h)D:St(K)St+ord(h)D(K)r_{h}^{\operatorname{ord}(h)D}:S_{t}(K)\to S_{t+\operatorname{ord}(h)D}(K) is a bijection which is independent of choice of hcKh\in c\cap K. Therefore, by possibly replacing (G,c)(G,c) with (K,cK)(K,c\cap K), to complete the proof, it suffices to show that for any non-splitting (G,c)(G,c) for tt sufficiently large, and for any h,kch,k\in c, rhord(h),rkord(k):St(G)St+2(G)r_{h}^{\operatorname{ord}(h)},r_{k}^{\operatorname{ord}(k)}:S_{t}(G)\to S_{t+2}(G) induce the same bijection.

Let tt be sufficiently large and xSt(G)x\in S_{t}(G). We wish to show the class of rhord(h)(x)r_{h}^{\operatorname{ord}(h)}(x) agrees with the class of rkord(k)(x)r_{k}^{\operatorname{ord}(k)}(x). Let G^\widehat{G} denote the group Sc×GabS_{c}\times_{G^{\operatorname{ab}}}\mathbb{Z}, where GabG^{\operatorname{ab}} denotes the abelianization of GG and ScS_{c} is a reduced Schur cover for (G,c)(G,c), as defined in [Woo21, Definition, p. 21]; for the reader’s benefit, we next review this notation. A Schur cover SGS\to G is a central extension of GG by some group KK so that the class of the extension in H2(G,K)H^{2}(G,K) maps to an isomorphism in Hom(H2(G,),K)\mathrm{Hom}(H_{2}(G,\mathbb{Z}),K) under the map from the universal coefficients exact sequence. A reduced Schur cover ScSS_{c}\subset S is a particular subgroup which is surjects onto GG. Following the notation of [Woo21], we notate the element of G^\hat{G} corresponding to hch\in c as (h^,eh)Sc×Gab=G^(\hat{h},e_{h})\in S_{c}\times_{G^{\operatorname{ab}}}\mathbb{Z}=\hat{G}, for eh=1e_{h}=1\in\mathbb{Z} and h^Sc\hat{h}\in S_{c} projecting to hGh\in G under the map ScGS_{c}\to G coming from the definition of a reduced Schur cover. For all other kck\in c, if k=shs1k=shs^{-1}, we can choose any lift s~Sc\tilde{s}\in S_{c} of ss and take k^:=s~h^s~1\hat{k}:=\tilde{s}\hat{h}\tilde{s}^{-1}. This is independent of the choice of ss and s~\tilde{s} by [Woo21, Lemma 2.3]. Write x=(h1,,ht)x=(h_{1},\ldots,h_{t}). It follows from [Woo21, Theorem 3.1 and Theorem 2.5] that showing rhord(h)xr_{h}^{\operatorname{ord}(h)}x lies in the same orbit as rkord(k)xr_{k}^{\operatorname{ord}(k)}x is equivalent to showing (h^,eh)ord(h)(h^1,eh1)(h^t,eht)=(k^,ek)ord(k)(h^1,eh1)(h^t,eht).(\hat{h},e_{h})^{\operatorname{ord}(h)}\cdot(\hat{h}_{1},e_{h_{1}})\cdots(\hat{h}_{t},e_{h_{t}})=(\hat{k},e_{k})^{\operatorname{ord(k)}}\cdot(\hat{h}_{1},e_{h_{1}})\cdots(\hat{h}_{t},e_{h_{t}}). Equivalently, we wish to show (h^,eh)ord(h)=(k^,ek)ord(k)(\hat{h},e_{h})^{\operatorname{ord}(h)}=(\hat{k},e_{k})^{\operatorname{ord(k)}}. We are assuming hh and kk lie in the same conjugacy class, and hence have the same order. Thus, the second coordinates of the above products agree, and it is enough to show h^ord(h)k^ord(k)=id\hat{h}^{\operatorname{ord}(h)}\hat{k}^{-\operatorname{ord(k)}}=\operatorname{\mathrm{id}}. Writing k=shs1k=shs^{-1}, and using the relation from [Woo21, Lemma 2.3] that for s~\tilde{s} any lift of ss, s~h^s~1=shs1^\tilde{s}\hat{h}\tilde{s}^{-1}=\widehat{shs^{-1}}, we find

h^ord(h)k^ord(h)=h^ord(h)s~h^ord(h)s~1=[h^ord(h),s~].\displaystyle\hat{h}^{\operatorname{ord}(h)}\hat{k}^{\operatorname{ord}(h)}=\hat{h}^{\operatorname{ord}(h)}\tilde{s}\hat{h}^{-\operatorname{ord}(h)}\tilde{s}^{-1}=[\hat{h}^{\operatorname{ord}(h)},\tilde{s}].

However, h^ord(h)\hat{h}^{\operatorname{ord}(h)} lies in the center of ScS_{c}, since its image in GG is hord(h)=idh^{\operatorname{ord}(h)}=\operatorname{\mathrm{id}} and ScS_{c} is a central extension of GG by ker(ScG)\ker(S_{c}\to G). Therefore, h^ord(h)\hat{h}^{\operatorname{ord}(h)} commutes with s~\tilde{s}, so [h^ord(h),s~]=id[\hat{h}^{\operatorname{ord}(h)},\tilde{s}]=\operatorname{\mathrm{id}}, as desired. ∎

A.4. Relating the gluing map to the stabilization map

In this subsection, we compare the logarithmic gluing map constructed in (A.4) to the stabilization map on cohomology. The main result is A.4.3, which shows they can be identified in a suitable sense.

One can show in a fashion analogous to 8.1.3 that the sequence of spaces Seln,σ\operatorname{Sel}_{\mathscr{F}^{n,\sigma}_{\mathbb{C}}} correspond to a coefficient system FnF_{n} for Σg,f1\Sigma_{g,f}^{1} over the same coefficient system VnV_{n} for Σ0,01\Sigma^{1}_{0,0} as in 8.1.3. Now, let cASp2g(/ν)c\subset\operatorname{\mathrm{ASp}}_{2g}(\mathbb{Z}/\nu\mathbb{Z}) denote the conjugacy class of elements projecting to id-\operatorname{\mathrm{id}} in Sp2g(/ν)\mathrm{Sp}_{2g}(\mathbb{Z}/\nu\mathbb{Z}) and let GASp2g(/ν)G\subset\operatorname{\mathrm{ASp}}_{2g}(\mathbb{Z}/\nu\mathbb{Z}) denote the subgroup generated by cc.

In the case B=Spec,B=\operatorname{Spec}\mathbb{C}, we obtain a stabilization map 𝕌\mathbb{U} on the cohomology over Spec\operatorname{Spec}\mathbb{C} as follows: Following [EVW16, Lemma 3.5], define 𝕌:=hcrhDord(h)RV\mathbb{U}:=\sum_{h\in c}r_{h}^{D\operatorname{ord}(h)}\in R^{V} for rhr_{h} right multiplication by hh, ord(h)\operatorname{ord}(h) the order of hh. Let d=Dord(h)d=D\cdot\mbox{ord}(h) denote the degree of 𝕌\mathbb{U}. (In our case, hh will always have order 22, so d=2Dd=2D, and ultimately we will take D=1D=1, but we will continue to use dd as we believe it is somewhat clarifying.) Then, 𝕌\mathbb{U} is a homogeneous central element with finite degree kernel and cokernel by A.3.1, and hence satisfies the hypotheses of Theorem 4.2.2. The map 𝕌\mathbb{U} on homology can be reexpressed in terms of a map on compactly supported cohomology which we continue to call 𝕌\mathbb{U}. We take \ell^{\prime} to be a prime invertible on BB. We may identify this 𝕌\mathbb{U} operator over the complex numbers with an operator 𝕌𝔽¯q\mathbb{U}_{\overline{\mathbb{F}}_{q}} on the 𝔽¯q\overline{\mathbb{F}}_{q} cohomology via the following commutative diagram

(A.5) Hci(Selnd,σ,(nd)){H_{\operatorname{c}}^{i}(\operatorname{Sel}_{\mathscr{F}^{n-d,\sigma}_{\mathbb{C}}},\mathbb{Q}_{\ell^{\prime}}(n-d))}Hci+2d(Seln,σ,(n)){H_{\operatorname{c}}^{i+2d}(\operatorname{Sel}_{\mathscr{F}^{n,\sigma}_{\mathbb{C}}},\mathbb{Q}_{\ell^{\prime}}(n))}Hci(Sel𝔽¯qnd,σ,(nd)){H_{\operatorname{c}}^{i}(\operatorname{Sel}_{\mathscr{F}^{n-d,\sigma}_{\overline{\mathbb{F}}_{q}}},\mathbb{Q}_{\ell^{\prime}}(n-d))}Hci+2d(Sel𝔽¯qn,σ,(n)),{H_{\operatorname{c}}^{i+2d}(\operatorname{Sel}_{\mathscr{F}^{n,\sigma}_{\overline{\mathbb{F}}_{q}}},\mathbb{Q}_{\ell^{\prime}}(n)),}𝕌\scriptstyle{\mathbb{U}}𝕌𝔽¯q\scriptstyle{\mathbb{U}_{\overline{\mathbb{F}}_{q}}}

where the vertical isomorphisms are obtained via the specialization maps (there are isomorphisms by [EVW16, Proposition 7.7]) and the map 𝕌𝔽¯q\mathbb{U}_{\overline{\mathbb{F}}_{q}} is the unique map making the diagram commute.

We will next define another map coming from logarithmic geometry. In order to define that map, we need the following result.

Lemma A.4.1.

If we are given xSel𝒢Bd,0,(𝔽q)x\in\operatorname{Sel}_{\mathscr{G}^{d,0,\infty}_{B}}(\mathbb{F}_{q}), we may identify the 𝔽¯q\overline{\mathbb{F}}_{q} points of

Sel𝒢Bd,,0×QTwist𝔾m/Bd,0,Sel𝒢Bd,0,\displaystyle\operatorname{Sel}_{\mathscr{G}^{d,\infty,0}_{B}}\times_{\operatorname{QTwist}^{d,0,\infty}_{\mathbb{G}_{m}/B}}\operatorname{Sel}_{\mathscr{G}^{d,0,\infty}_{B}}

over xx with tuples in GdG^{d} whose product is idG\operatorname{\mathrm{id}}\in G. This fiber has an action of Frobenius, Frobq\operatorname{Frob}_{q}. For dd even, under this bijection, the set of elements {(h,,h):hc}\{(h,\ldots,h):h\in c\} in the fiber over xx constitutes a union of Frobq\operatorname{Frob}_{q} orbits.

Proof.

Since Frobenius must preserve the conjugacy class of an element in GG, as the conjugacy class can be read from from the inertia of the corresponding cover, the orbit of (h,,h)(h,\ldots,h) must consist of elements of the form (h1,,hn)(h_{1},\ldots,h_{n}) where each hich_{i}\in c.

It remains to show that any such element in this orbit satisfies hi=hjh_{i}=h_{j} for 1in1\leq i\leq n. We have that Sel𝒢Bd,,0×QTwist𝔾m/Bd,0,Sel𝒢Bd,0,\operatorname{Sel}_{\mathscr{G}^{d,\infty,0}_{B}}\times_{\operatorname{QTwist}^{d,0,\infty}_{\mathbb{G}_{m}/B}}\operatorname{Sel}_{\mathscr{G}^{d,0,\infty}_{B}} is a finite étale cover of Sel𝒢Bd,0,\operatorname{Sel}_{\mathscr{G}^{d,0,\infty}_{B}}, and hence we obtain an action of the fundamental group of Sel𝒢Bd,0,\operatorname{Sel}_{\mathscr{G}^{d,0,\infty}_{B}} on the geometric fiber of Sel𝒢Bd,,0×QTwist𝔾m/Bd,0,Sel𝒢Bd,0,\operatorname{Sel}_{\mathscr{G}^{d,\infty,0}_{B}}\times_{\operatorname{QTwist}^{d,0,\infty}_{\mathbb{G}_{m}/B}}\operatorname{Sel}_{\mathscr{G}^{d,0,\infty}_{B}} over a given point xSel𝒢Bd,0,x\in\operatorname{Sel}_{\mathscr{G}^{d,0,\infty}_{B}}. The above sheaf over Sel𝒢xd,0,\operatorname{Sel}_{\mathscr{G}^{d,0,\infty}_{x}} is obtained as the base change of a sheaf over Conf1/xd\operatorname{Conf}^{d}_{\mathbb{P}^{1}/x} using that 𝒢\mathscr{G} is trivial, so extends over \infty and 0.

We next conclude the proof by showing the set {(h,,h):hG}\{(h,\ldots,h):h\in G\} over the image of xx form a union of Frobq\operatorname{Frob}_{q} orbits. Let x¯\overline{x} denote a geometric point over xx. Note that this set {(h,,h):hG}\{(h,\ldots,h):h\in G\} now inherits an action of the fundamental group of Conf1/xd\operatorname{Conf}^{d}_{\mathbb{P}^{1}/x}, which is a semidirect product of its geometric fundamental group, π1(Conf1/x¯d)\pi_{1}(\operatorname{Conf}^{d}_{\mathbb{P}^{1}/\overline{x}}), the profinite completion of the braid group, and π1(x)^\pi_{1}(x)\simeq\widehat{\mathbb{Z}}, generated by Frobenius. Hence, for any ηπ1(Conf1/x¯d)\eta\in\pi_{1}(\operatorname{Conf}^{d}_{\mathbb{P}^{1}/\overline{x}}) there is some ηπ1(Conf1/x¯d)\eta^{\prime}\in\pi_{1}(\operatorname{Conf}^{d}_{\mathbb{P}^{1}/\overline{x}}) with ηFrobq(h,,h)=Frobqη(h,,h)\eta\operatorname{Frob}_{q}(h,\ldots,h)=\operatorname{Frob}_{q}\eta^{\prime}(h,\ldots,h). Since the braid group fixes (h,,h)(h,\ldots,h), we find ηFrobq(h,,h)=Frobq(h,,h)\eta\operatorname{Frob}_{q}(h,\ldots,h)=\operatorname{Frob}_{q}(h,\ldots,h), and hence Frobq(h,,h)\operatorname{Frob}_{q}(h,\ldots,h) is fixed by the action of the profinite completion of the Braid group. Since elements of the form (k,,k)Gn(k,\ldots,k)\in G^{n} are the only elements fixed by the profinite completion of the Braid group, we must have Frobq(h,,h)=(k,,k)\operatorname{Frob}_{q}(h,\ldots,h)=(k,\ldots,k) for some kGk\in G. ∎

A.4.1. Defining a stabilization map from logarithmic geometry

Let BB be the spectrum of a complete dvr with residue field 𝔽q\mathbb{F}_{q} and generic characteristic 0. Suppose there exists xSel𝒢Bd,0,(𝔽q)x\in\operatorname{Sel}_{\mathscr{G}^{d,0,\infty}_{B}}(\mathbb{F}_{q}). One can lift this to a section of Sel𝒢Bd,0,(B)\operatorname{Sel}_{\mathscr{G}^{d,0,\infty}_{B}}(B) over BB using smoothness of Sel𝒢Bd,0,\operatorname{Sel}_{\mathscr{G}^{d,0,\infty}_{B}} to lift the point over any power of the maximal ideal of the dvr corresponding to BB, which then algebraizes to a BB-point by [FGI+05, Corollary 8.4.6]. Let ι:S𝔽qSel𝒢Bd,,0×QTwist𝔾m/Bd,0,Sel𝒢Bd,0,\iota:S_{\mathbb{F}_{q}}\subset\operatorname{Sel}_{\mathscr{G}^{d,\infty,0}_{B}}\times_{\operatorname{QTwist}^{d,0,\infty}_{\mathbb{G}_{m}/B}}\operatorname{Sel}_{\mathscr{G}^{d,0,\infty}_{B}} denote the reduced closed subscheme over 𝔽q\mathbb{F}_{q} whose base change to 𝔽¯q\overline{\mathbb{F}}_{q} corresponds to the set of 𝔽¯q\overline{\mathbb{F}}_{q} points hc(h,,h)\cup_{h\in c}(h,\ldots,h). This is a well defined subscheme by A.4.1. Let SBS_{B} denote a lift of S𝔽qS_{\mathbb{F}_{q}} over the given lift of xx, which exists and is unique because the cover Sel𝒢Bd,,0×QTwist𝔾m/Bd,0,Sel𝒢Bd,0,Sel𝒢Bd,0,\operatorname{Sel}_{\mathscr{G}^{d,\infty,0}_{B}}\times_{\operatorname{QTwist}^{d,0,\infty}_{\mathbb{G}_{m}/B}}\operatorname{Sel}_{\mathscr{G}^{d,0,\infty}_{B}}\to\operatorname{Sel}_{\mathscr{G}^{d,0,\infty}_{B}} is finite étale. One may verify the complement of SelTn,σSelSt,Tn,σ\operatorname{Sel}_{\mathscr{F}^{n,\sigma}_{T}}\subset\operatorname{Sel}_{\mathscr{F}^{n,\sigma}_{\operatorname{St},T}} is a smooth divisor, using its description as a finite cover of ConfU/BnStU/Bn,σ\operatorname{Conf}^{n}_{U/B}\subset\operatorname{St}^{n,\sigma}_{U/B}, which has complement a smooth divisor by A.1.2. (The cover is not étale over the boundary, but it is branched over the boundary if a fixed degree, which is enough to guarantee the smoothness above.) For TBT\to B the spectrum of a field, using smoothness of the complement SelTn,σSelSt,Tn,σ\operatorname{Sel}_{\mathscr{F}^{n,\sigma}_{T}}\subset\operatorname{Sel}_{\mathscr{F}^{n,\sigma}_{\operatorname{St},T}} mentioned above, we also obtain the identification

δ:Hi(SelTn,σ,(n))Hi(SelSt,Tn,σlog,(n)).\displaystyle\delta:H^{i}\left(\operatorname{Sel}_{\mathscr{F}^{n,\sigma}_{T}},\mathbb{Q}_{\ell^{\prime}}(n)\right)\simeq H^{i}\left(\operatorname{Sel}^{\operatorname{log}}_{\mathscr{F}^{n,\sigma}_{\operatorname{St},T}},\mathbb{Q}_{\ell^{\prime}}(n)\right).

by [BDPW23, §8.5.3].

Next, the inclusion

(A.6) β:SBSB×BBι×{1}Sel𝒢Bd,,0×QTwist𝔾m/Bd,0,Sel𝒢Bd,0,×B𝔾m,\displaystyle\beta:S_{B}\simeq S_{B}\times_{B}B\xrightarrow{\iota\times\{1\}}\operatorname{Sel}_{\mathscr{G}^{d,\infty,0}_{B}}\times_{\operatorname{QTwist}^{d,0,\infty}_{\mathbb{G}_{m}/B}}\operatorname{Sel}_{\mathscr{G}^{d,0,\infty}_{B}}\times_{B}\mathbb{G}_{m},

induces a strict map of log stacks

(SB×BSelBnd,σ)log((Sel𝒢Bd,,0×QTwist𝔾m/Bd,0,Sel𝒢Bd,0,)×BSelBnd,σ)log\displaystyle\left(S_{B}\times_{B}\operatorname{Sel}_{\mathscr{F}^{n-d,\sigma}_{B}}\right)^{\operatorname{log}}\to\left(\left(\operatorname{Sel}_{\mathscr{G}^{d,\infty,0}_{B}}\times_{\operatorname{QTwist}^{d,0,\infty}_{\mathbb{G}_{m}/B}}\operatorname{Sel}_{\mathscr{G}^{d,0,\infty}_{B}}\right)\times_{B}\operatorname{Sel}_{\mathscr{F}^{n-d,\sigma}_{B}}\right)^{\operatorname{log}}

where we endow SB×BSelBnd,σS_{B}\times_{B}\operatorname{Sel}_{\mathscr{F}^{n-d,\sigma}_{B}} with the standard log structure, consisting of the trivial line bundle with the 0 section.

Using the above described maps along with the map α\alpha from (A.4), and base changing along some spectrum of a field TBT\to B, we obtain a map on cohomology

(A.7) Hi(SelTn,σ,)\displaystyle H^{i}\left(\operatorname{Sel}_{\mathscr{F}^{n,\sigma}_{T}},\mathbb{Q}_{\ell^{\prime}}\right) 𝛿Hi(SelSt,Tn,σlog,)\displaystyle\xrightarrow{\delta}H^{i}\left(\operatorname{Sel}^{\operatorname{log}}_{\mathscr{F}^{n,\sigma}_{\operatorname{St},T}},\mathbb{Q}_{\ell^{\prime}}\right)
αHi(((Sel𝒢Td,,0×QTwist𝔾m/Td,0,Sel𝒢Td,0,)×TSelTnd,σ)log,)\displaystyle\xrightarrow{\alpha^{*}}H^{i}\left(\left(\left(\operatorname{Sel}_{\mathscr{G}^{d,\infty,0}_{T}}\times_{\operatorname{QTwist}^{d,0,\infty}_{\mathbb{G}_{m}/T}}\operatorname{Sel}_{\mathscr{G}^{d,0,\infty}_{T}}\right)\times_{T}\operatorname{Sel}_{\mathscr{F}^{n-d,\sigma}_{T}}\right)^{\operatorname{log}},\mathbb{Q}_{\ell^{\prime}}\right)
βHi((ST×TSelTnd,σ)log,).\displaystyle\xrightarrow{\beta^{*}}H^{i}\left(\left(S_{T}\times_{T}\operatorname{Sel}_{\mathscr{F}^{n-d,\sigma}_{T}}\right)^{\operatorname{log}},\mathbb{Q}_{\ell^{\prime}}\right).
Lemma A.4.2.

The map αδ\alpha^{*}\circ\delta in (A.7) over T=SpecT=\operatorname{Spec}\mathbb{C} can be identified with the map induced on cohomology of the gluing map described as follows. The map takes in the following data:

  1. (1)

    a direction τ\tau on the unit circle,

  2. (2)

    an ASp2g(/ν)\operatorname{\mathrm{ASp}}_{2g}(\mathbb{Z}/\nu\mathbb{Z}) cover of Σg,f1\Sigma^{1}_{g,f},

  3. (3)

    an ASp2g(/ν)\operatorname{\mathrm{ASp}}_{2g}(\mathbb{Z}/\nu\mathbb{Z}) cover of Σ0,02\Sigma^{2}_{0,0},

  4. (4)

    a specified identification of the boundary of Σg,f1\Sigma^{1}_{g,f} with S1S^{1},

  5. (5)

    a specified identification of one of the boundary components of Σ0,02\Sigma^{2}_{0,0}, corresponding to the point 010\in\mathbb{P}^{1}, with S1S^{1}.

The gluing map then glues the two copies of S1S^{1} in (4)(4) and (5)(5) via a rotation by τ\tau, and glues the boundary components of the covers by the pullback of this identification.

Proof.

We let T=SpecT=\operatorname{Spec}\mathbb{C} and verify the explicit description of the map. Identifying the log schemes in the source and target of the map α\alpha^{*} in (A.7) with their corresponding Kato-Nakayama spaces as in [BDPW23, Examples 8.4.4 and 8.4.5], we find the map α\alpha^{*} from (A.4) is obtained from the map of underlying stacks from A.2.1 together with the map of logarithmic structures from A.2.2. Observe that we have a commutative square of log stacks

(A.8) ((Sel𝒢Td,,0×QTwist𝔾m/Td,0,Sel𝒢Td,0,)×TSelTnd,σ)log{\left(\left(\operatorname{Sel}_{\mathscr{G}^{d,\infty,0}_{T}}\times_{\operatorname{QTwist}^{d,0,\infty}_{\mathbb{G}_{m}/T}}\operatorname{Sel}_{\mathscr{G}^{d,0,\infty}_{T}}\right)\times_{T}\operatorname{Sel}_{\mathscr{F}^{n-d,\sigma}_{T}}\right)^{\operatorname{log}}}SelSt,Tn,σlog{\operatorname{Sel}^{\operatorname{log}}_{\mathscr{F}^{n,\sigma}_{\operatorname{St},T}}}(Conf𝔾m/Bd×ConfU/Bnd)log{\left(\operatorname{Conf}^{d}_{\mathbb{G}_{m}/B}\times\operatorname{Conf}^{n-d}_{U/B}\right)^{\operatorname{log}}}(StU/Bn,σ)log,{\left(\operatorname{St}^{n,\sigma}_{U/B}\right)^{\operatorname{log}},}

where (Conf𝔾m/Bd×ConfU/Bnd)log\left(\operatorname{Conf}^{d}_{\mathbb{G}_{m}/B}\times\operatorname{Conf}^{n-d}_{U/B}\right)^{\operatorname{log}} has the standard log structure (a trivial line bundle with the 0 section, and (StU/Bn,σ)log\left(\operatorname{St}^{n,\sigma}_{U/B}\right)^{\operatorname{log}} has the log structure defined by the boundary divisor which is the complement of ConfU/Bn\operatorname{Conf}^{n}_{U/B}. The pull back of the line bundle giving the log structure defined by the boundary divisor on StU/Bn,σ\operatorname{St}^{n,\sigma}_{U/B} can be identified with the trivial bundle on Conf𝔾m/Bd×ConfU/Bnd\operatorname{Conf}^{d}_{\mathbb{G}_{m}/B}\times\operatorname{Conf}^{n-d}_{U/B}, which is more canonically the tensor product of the tangent bundles at 0 and σ\sigma, by a proof analogous to the proof of A.2.2. Hence, the gluing map associated to the bottom map in (A.8) can be described as choosing a unit tangent vector over 0 and a unit tangent vector over σ\sigma and then gluing the unit tangent spaces so as to identify those unit tangent vectors. This results in a point of the Kato-Nakayama space of (StU/Bn,σ)log\left(\operatorname{St}^{n,\sigma}_{U/B}\right)^{\operatorname{log}}. Choose an identification of UU with the interior of Σg,f1\Sigma^{1}_{g,f}, with one boundary component corresponding to σ\sigma, and an identification of 𝔾m\mathbb{G}_{m} with the interior of Σ0,11\Sigma^{1}_{0,1}, with a boundary component at 0 and a puncture at \infty. Topologically, we can further identify the above map with a map gluing Σg,f1\Sigma^{1}_{g,f} to Σ0,02\Sigma^{2}_{0,0} via altering Σ0,11\Sigma^{1}_{0,1} to Σ0,02\Sigma^{2}_{0,0} by replacing a puncture at \infty with a boundary component. The above yields a description of the gluing map on configuration spaces analogous to that in the statement. Using the commutative square (A.8), the map of line bundles associated to the map of Selmer spaces is pulled back from the corresponding map of line bundles on configuration spaces, yielding the identification we wished to show. ∎

Additionally, there is a map of log schemes (ST×TSelTnd,σ)logγ~ST×TSelTnd,σ\left(S_{T}\times_{T}\operatorname{Sel}_{\mathscr{F}^{n-d,\sigma}_{T}}\right)^{\operatorname{log}}\xrightarrow{\widetilde{\gamma}}S_{T}\times_{T}\operatorname{Sel}_{\mathscr{F}^{n-d,\sigma}_{T}} where we use ST×TSelTnd,σS_{T}\times_{T}\operatorname{Sel}_{\mathscr{F}^{n-d,\sigma}_{T}} to denote the log scheme with trivial log structure (corresponding to no line bundles). This induces a map on cohomology

(A.9) Hi(ST×TSelTnd,σ,)𝛾Hi((ST×TSelTnd,σ)log,).\displaystyle H^{i}(S_{T}\times_{T}\operatorname{Sel}_{\mathscr{F}^{n-d,\sigma}_{T}},\mathbb{Q}_{\ell^{\prime}})\xrightarrow{\gamma}H^{i}\left(\left(S_{T}\times_{T}\operatorname{Sel}_{\mathscr{F}^{n-d,\sigma}_{T}}\right)^{\operatorname{log}},\mathbb{Q}_{\ell^{\prime}}\right).
Proposition A.4.3.

Assume BB is the spectrum of a complete dvr with residue field 𝔽q\mathbb{F}_{q} and generic characteristic 0. Suppose Sel𝒢Bd,0,(𝔽q)\operatorname{Sel}_{\mathscr{G}^{d,0,\infty}_{B}}(\mathbb{F}_{q})\neq\emptyset. If TT is either Spec𝔽¯q\operatorname{Spec}\overline{\mathbb{F}}_{q} or Spec\operatorname{Spec}\mathbb{C}, there is a canonical splitting

(A.10) Hi((ST×TSelTnd,σ)log,)𝜀Hi(ST×TSelTnd,σ,)\displaystyle H^{i}\left(\left(S_{T}\times_{T}\operatorname{Sel}_{\mathscr{F}^{n-d,\sigma}_{T}}\right)^{\operatorname{log}},\mathbb{Q}_{\ell^{\prime}}\right)\xrightarrow{\varepsilon}H^{i}(S_{T}\times_{T}\operatorname{Sel}_{\mathscr{F}^{n-d,\sigma}_{T}},\mathbb{Q}_{\ell^{\prime}})

of γ\gamma, i.e., εγ=id\varepsilon\circ\gamma=\operatorname{\mathrm{id}}. Additionally, if η:Hi(ST×TSelTnd,σ,)Hi(SelTnd,σ,)\eta:H^{i}(S_{T}\times_{T}\operatorname{Sel}_{\mathscr{F}^{n-d,\sigma}_{T}},\mathbb{Q}_{\ell^{\prime}})\to H^{i}(\operatorname{Sel}_{\mathscr{F}^{n-d,\sigma}_{T}},\mathbb{Q}_{\ell^{\prime}}) is the summation map obtained by identifying STS_{T} with a disjoint union of points and summing the resulting cohomology elements, the composition of (A.7) with ηε\eta\circ\varepsilon is Poincaré dual to a map

(A.11) Hci(SelTnd,σ,(nd))Hci+2d(SelTn,σ,(n)).\displaystyle H^{i}_{\operatorname{c}}(\operatorname{Sel}_{\mathscr{F}^{n-d,\sigma}_{T}},\mathbb{Q}_{\ell^{\prime}}(n-d))\to H^{i+2d}_{\operatorname{c}}\left(\operatorname{Sel}_{\mathscr{F}^{n,\sigma}_{T}},\mathbb{Q}_{\ell^{\prime}}(n)\right).

which agrees with 𝕌\mathbb{U} when T=SpecT=\operatorname{Spec}\mathbb{C} and agrees with 𝕌𝔽¯q\mathbb{U}_{\overline{\mathbb{F}}_{q}} with T=Spec𝔽¯q.T=\operatorname{Spec}\overline{\mathbb{F}}_{q}.

Proof.

First, we explain how to deduce the final statement when T=Spec𝔽¯qT=\operatorname{Spec}\overline{\mathbb{F}}_{q} from the case that T=SpecT=\operatorname{Spec}\mathbb{C} using the specialization map. For the final statement with T=Spec𝔽¯qT=\operatorname{Spec}\overline{\mathbb{F}}_{q}, we wish to prove the surjective specialization map is an isomorphism, and so we wish to prove the constructible cohomology sheaves on BB corresponding to each of the terms in (A.7) and (A.9) are locally constant on BB. Local constancy of the cohomology of SelBnd,σ\operatorname{Sel}_{\mathscr{F}^{n-d,\sigma}_{B}}, and hence also of its finite cover, SB×BSelBnd,σS_{B}\times_{B}\operatorname{Sel}_{\mathscr{F}^{n-d,\sigma}_{B}}, follows from [EVW16, Proposition 7.7]. Hence, by functoriality of the specialization map, it is enough to verify local constancy of the cohomology for the projection (SB×BSelBnd,σ)logB\left(S_{B}\times_{B}\operatorname{Sel}_{\mathscr{F}^{n-d,\sigma}_{B}}\right)^{\operatorname{log}}\to B and that the splitting ε\varepsilon from (A.10) is compatible with the specialization map.

We first verify local constancy of the cohomology. Observe that we can write (SB×BSelBnd,σ)log\left(S_{B}\times_{B}\operatorname{Sel}_{\mathscr{F}^{n-d,\sigma}_{B}}\right)^{\operatorname{log}} as the fiber product (SB×BSelBnd,σ)×BBlog\left(S_{B}\times_{B}\operatorname{Sel}_{\mathscr{F}^{n-d,\sigma}_{B}}\right)\times_{B}B^{\operatorname{log}} where here we give (SB×BSelBnd,σ)\left(S_{B}\times_{B}\operatorname{Sel}_{\mathscr{F}^{n-d,\sigma}_{B}}\right) and BB the trivial log structures, corresponding to no line bundles, and BlogB^{\operatorname{log}} the standard log structure, corresponding to 𝒪B\mathscr{O}_{B} with the 0 section. By the Künneth theorem, whose log cohomology version in our setting follows from proper base change [Ill02, Proposition 6.3] and the projection formula, it is enough to show the cohomology sheaves associated to both (SB×BSelBnd,σ)\left(S_{B}\times_{B}\operatorname{Sel}_{\mathscr{F}^{n-d,\sigma}_{B}}\right) and BlogB^{\operatorname{log}} are locally constant. We have already verified the former above, while the latter follows from [Ill02, Theorem 5.2], assuming \ell^{\prime} is invertible on BB. Moreover the cohomology BlogB^{\operatorname{log}} is isomorphic to that of 𝔾m\mathbb{G}_{m}.

We next define the splitting ε\varepsilon in (A.10) and verify it is compatible with the specialization map. Notice that the above description of the cohomology of (ST×TSelTnd,σ)log\left(S_{T}\times_{T}\operatorname{Sel}_{\mathscr{F}^{n-d,\sigma}_{T}}\right)^{\operatorname{log}} gives an isomorphism of cohomology rings with Frobenius action

H((ST×TSelTnd,σ)log,)\displaystyle H^{\bullet}\left(\left(S_{T}\times_{T}\operatorname{Sel}_{\mathscr{F}^{n-d,\sigma}_{T}}\right)^{\operatorname{log}},\mathbb{Q}_{\ell^{\prime}}\right) H((ST×TSelTnd,σ)×T𝔾m,)\displaystyle\simeq H^{\bullet}\left((S_{T}\times_{T}\operatorname{Sel}_{\mathscr{F}^{n-d,\sigma}_{T}})\times_{T}\mathbb{G}_{m},\mathbb{Q}_{\ell^{\prime}}\right)
H(ST×TSelTnd,σ,)H(𝔾m,),\displaystyle\simeq H^{\bullet}\left(S_{T}\times_{T}\operatorname{Sel}_{\mathscr{F}^{n-d,\sigma}_{T}},\mathbb{Q}_{\ell^{\prime}}\right)\otimes H^{\bullet}(\mathbb{G}_{m},\mathbb{Q}_{\ell^{\prime}}),

the latter isomorphism via the Künneth isomorphism. Hence, we can identify

Hi(ST×TSelTnd,σ,)H0(𝔾m,)Hi((ST×TSelTnd,σ)log,),\displaystyle H^{i}\left(S_{T}\times_{T}\operatorname{Sel}_{\mathscr{F}^{n-d,\sigma}_{T}},\mathbb{Q}_{\ell^{\prime}}\right)\otimes H^{0}(\mathbb{G}_{m},\mathbb{Q}_{\ell^{\prime}})\simeq H^{i}\left(\left(S_{T}\times_{T}\operatorname{Sel}_{\mathscr{F}^{n-d,\sigma}_{T}}\right)^{\operatorname{log}},\mathbb{Q}_{\ell^{\prime}}\right),

where the isomorphism is equivariant for the Frobenius action when T=Spec𝔽¯qT=\operatorname{Spec}\overline{\mathbb{F}}_{q}. This gives the desired splitting ε\varepsilon from (A.10). Moreover, the above subspace is compatible with the specialization map, as we wished to show. Overall, this reduces us to verifying the final claim when T=SpecT=\operatorname{Spec}\mathbb{C}.

We conclude by verifying the final statement when T=SpecT=\operatorname{Spec}\mathbb{C}. On the level of Kato-Nakayama spaces, the splitting ε\varepsilon defined above can be obtained from the inclusion ST×TSelTnd,σS1×(ST×TSelTnd,σ)S_{T}\times_{T}\operatorname{Sel}_{\mathscr{F}^{n-d,\sigma}_{T}}\to S^{1}\times(S_{T}\times_{T}\operatorname{Sel}_{\mathscr{F}^{n-d,\sigma}_{T}}), coming from choosing a fixed direction τS1\tau\in S^{1}. If we compose with the inclusion ιh:ST×TSelTnd,σSelTnd,σ\iota_{h}:S_{T}\times_{T}\operatorname{Sel}_{\mathscr{F}^{n-d,\sigma}_{T}}\to\operatorname{Sel}_{\mathscr{F}^{n-d,\sigma}_{T}} associated to a particular tuple (h,,h)(h,\ldots,h) over xSel𝒢Bd,0,(𝔽q)x\in\operatorname{Sel}_{\mathscr{G}^{d,0,\infty}_{B}}(\mathbb{F}_{q}), the description from A.4.2 implies that the map ιhεβαδ\iota_{h}^{*}\varepsilon\circ\beta^{*}\alpha^{*}\circ\delta on cohomology is induced by the map of Kato-Nakayama spaces described as follows: start with an ASp2r(/ν)\operatorname{\mathrm{ASp}}_{2r}(\mathbb{Z}/\nu\mathbb{Z}) cover of Σg,f1\Sigma^{1}_{g,f} and glue on a disc with dd punctures having monodromy around each such puncture given by hh. The map 𝕌\mathbb{U} is the sum over hch\in c of the Poincaré duals of these maps on cohomology, and hence the composite of ηε\eta\circ\varepsilon with (A.7) is Poincaré dual to 𝕌\mathbb{U}. ∎

By A.4.3, the map 𝕌𝔽¯q\mathbb{U}_{\overline{\mathbb{F}}_{q}} is identified with a map

(A.12) Hci(Sel𝔽¯qnd,σ,(nd))Hci+2d(Sel𝔽¯qn,σ,(n)).\displaystyle H^{i}_{\operatorname{c}}(\operatorname{Sel}_{\mathscr{F}^{n-d,\sigma}_{{\overline{\mathbb{F}}_{q}}}},\mathbb{Q}_{\ell^{\prime}}(n-d))\to H^{i+2d}_{\operatorname{c}}\left(\operatorname{Sel}_{\mathscr{F}^{n,\sigma}_{{\overline{\mathbb{F}}_{q}}}},\mathbb{Q}_{\ell^{\prime}}(n)\right).

A.5. Proving the main Frobenius equivariance result

In this subsection, we prove our main result, Theorem A.5.1, that the stabilization map is equivariant for the Frobenius action.

As a preliminary step to connect the version of Selmer spaces where we mark extra data over σ\sigma to the version without such marked data, we need to understand the group action relating these two spaces. Note that there is an action of /2(j)σ\mathbb{Z}/2\mathbb{Z}\ltimes(j_{*}\mathscr{F})_{\sigma} on SelBnd,σ\operatorname{Sel}_{\mathscr{F}^{n-d,\sigma}_{B}} where the /2\mathbb{Z}/2\mathbb{Z} acts by negation on the fiber (j)σ(j_{*}\mathscr{F})_{\sigma} and the copy of (j)σ(j_{*}\mathscr{F})_{\sigma} acts by translation. The quotient of SelBnd,σ\operatorname{Sel}_{\mathscr{F}^{n-d,\sigma}_{B}} by this action is SelBnd\operatorname{Sel}_{\mathscr{F}^{n-d}_{B}}, where we no longer include the marked point σ\sigma.

Lemma A.5.1.

Assume BB is the spectrum of a complete dvr with residue field 𝔽q\mathbb{F}_{q} and generic characteristic 0. Suppose Sel𝒢Bd,0,(𝔽q)\operatorname{Sel}_{\mathscr{G}^{d,0,\infty}_{B}}(\mathbb{F}_{q})\neq\emptyset. The map (A.12) is equivariant for the actions of Frobenius and the actions of /2(j)σ\mathbb{Z}/2\mathbb{Z}\ltimes(j_{*}\mathscr{F})_{\sigma} on both sides.

Proof.

The map (A.12) is equivariant for the action of Frobenius since it is the composite of the dual map (A.7) with the Frobenius equivariant maps ε\varepsilon and η\eta in (A.4.3). Note here we are using that maps of log schemes induce functorial maps on their cohomology, as follows from functoriality of the Kummer étale topology, see [Ill02, §2.1]. The composite map (A.12) is then also equivariant for the action of Frobenius because S𝔽¯qS_{\overline{\mathbb{F}}_{q}} is defined over 𝔽q\mathbb{F}_{q} by A.4.1.

We conclude by arguing that the action of /2(j)σ\mathbb{Z}/2\mathbb{Z}\ltimes(j_{*}\mathscr{F})_{\sigma} is also equivariant for the map (A.12). One can identify the action of this group with the action on the fiber over σ\sigma. The gluing map 𝕌\mathbb{U} in topology induces an equivariant map on cohomology for this group action, and the algebraic map (A.12) is identified with the map 𝕌\mathbb{U} via A.4.3 and (A.5). ∎

We are now ready to deduce our main result of this section.

Theorem A.5.1.

Assume BB is a complete dvr with residue field 𝔽q\mathbb{F}_{q} and generic characteristic 0. Suppose Sel𝒢Bd,0,(𝔽q)\operatorname{Sel}_{\mathscr{G}^{d,0,\infty}_{B}}(\mathbb{F}_{q})\neq\emptyset. Suppose ZZ as in 2.4.1 has a section σ:BZ\sigma:B\to Z and \mathscr{F} as in 5.1.4 has trivial inertia along σ\sigma. Suppose HH is a finite /ν\mathbb{Z}/\nu\mathbb{Z} module. There is a positive integer constant II, as well as a positive integer constant J(,H)J(\mathscr{F},H) depending on \mathscr{F} and HH so that, for any positive even integer nn, there is a map

(A.13) Hc2np(Sel𝔽¯qnH,(n))\displaystyle H^{2n-p}_{\operatorname{c}}(\operatorname{Sel}_{\mathscr{F}^{n}_{\overline{\mathbb{F}}_{q}}}^{H},\mathbb{Q}_{\ell^{\prime}}(n)) Hc2np+4(Sel𝔽¯qn+2H,(n+2))\displaystyle\to H^{2n-p+4}_{\operatorname{c}}(\operatorname{Sel}_{\mathscr{F}^{n+2}_{\overline{\mathbb{F}}_{q}}}^{H},\mathbb{Q}_{\ell^{\prime}}(n+2))

which is equivariant for the action of Frobenius. Moreover, this map is an isomorphism whenever n>Ip+J(,H)n>Ip+J(\mathscr{F},H).

Remark A.5.2.

The map (A.13) is induced from the map (A.12), with d=2d=2 and nn replaced by n+2n+2, via transfer.

Proof.

First, by A.3.1, since we are working with cG=H(/2)c\subset G=H\rtimes(\mathbb{Z}/2\mathbb{Z}) corresponding to the elements of order 22, we may take the operator 𝕌\mathbb{U} to have degree 22.

We will only explain the proof in the case that H=/νH=\mathbb{Z}/\nu\mathbb{Z}. The case of general HH, where one takes iterated fiber products of the Selmer space over the space of quadratic twists, is quite analogous. However, we opt to just explain the case that H=/νH=\mathbb{Z}/\nu\mathbb{Z} to avoid introducing an onslaught of additional notation that does not require any new ideas.

First, we note that the map 𝕌𝔽¯q\mathbb{U}_{\overline{\mathbb{F}}_{q}} is equivariant for the action of Frobenius by A.5.1. When n>Ip+Jn>Ip+J, commutativity of (A.5) implies that 𝕌𝔽¯q\mathbb{U}_{\overline{\mathbb{F}}_{q}} is an isomorphism

Hc2np(Sel𝔽¯qn,σ,(n))Hc2np+4(Sel𝔽¯qn+2,σ,(n+2))\displaystyle H^{2n-p}_{\operatorname{c}}(\operatorname{Sel}_{\mathscr{F}^{n,\sigma}_{\overline{\mathbb{F}}_{q}}},\mathbb{Q}_{\ell^{\prime}}(n))\to H^{2n-p+4}_{\operatorname{c}}(\operatorname{Sel}_{\mathscr{F}^{n+2,\sigma}_{\overline{\mathbb{F}}_{q}}},\mathbb{Q}_{\ell^{\prime}}(n+2))

when n>Ip+J(,H)n>Ip+J(\mathscr{F},H), since the corresponding map 𝕌\mathbb{U} over \mathbb{C} is an isomorphism by Poincaré duality and Theorem 4.2.2. See 8.1.3 and 8.1.11 for why the relevant representations of Bg,fnB^{n}_{g,f} form coefficient systems.

Finally, using that the map (A.12) is equivariant for the action of /2(j)σ\mathbb{Z}/2\mathbb{Z}\ltimes(j_{*}\mathscr{F})_{\sigma} by A.5.1, we obtain an induced map on the cohomology of the quotient space by this action of /2(j)σ\mathbb{Z}/2\mathbb{Z}\ltimes(j_{*}\mathscr{F})_{\sigma}. By transfer, since we are assuming 2ν2\nu is invertible on BB, the cohomology of the quotient is also equivariant for the action of Frobenius. Since the quotient of Sel𝔽¯qn,σ\operatorname{Sel}_{\mathscr{F}^{n,\sigma}_{\overline{\mathbb{F}}_{q}}} by this /2(j)σ\mathbb{Z}/2\mathbb{Z}\ltimes(j_{*}\mathscr{F})_{\sigma} action is Sel𝔽¯qn\operatorname{Sel}_{\mathscr{F}^{n}_{\overline{\mathbb{F}}_{q}}}, (without trivializations over σ\sigma,) we obtain the maps

Hc2np(Sel𝔽¯qn,(n))Hc2np+4(Sel𝔽¯qn+2,(n+2))\displaystyle H^{2n-p}_{\operatorname{c}}(\operatorname{Sel}_{\mathscr{F}^{n}_{\overline{\mathbb{F}}_{q}}},\mathbb{Q}_{\ell^{\prime}}(n))\to H^{2n-p+4}_{\operatorname{c}}\left(\operatorname{Sel}_{\mathscr{F}^{n+2}_{\overline{\mathbb{F}}_{q}}},\mathbb{Q}_{\ell^{\prime}}(n+2)\right)

are Frobenius equivariant and moreover are isomorphisms when n>Ip+J(,H)n>Ip+J(\mathscr{F},H). ∎

Appendix B A normal crossings compactification of Hurwitz spaces
By Dori Bejleri and Aaron Landesman

The main consequence of this appendix, B.1.3, proves that configuration spaces of points on a pointed smooth curve, considered earlier in this paper, have normal crossing compactifications. This was crucially used to compare the cohomology of Hurwitz spaces over 𝔽¯q\overline{\mathbb{F}}_{q} with the cohomology of Hurwitz spaces over \mathbb{C}. Because it is little extra work, we also show that Hurwitz spaces, which are finite étale covers of these configurations spaces, have normal crossing compactifications. In order to achieve this comparison between \mathbb{C} and 𝔽¯q\overline{\mathbb{F}}_{q}, when dealing with a Hurwitz space for a finite group GG, we work over the base [1/|G|]\mathbb{Z}[1/|G|]. In particular, our results hold over mixed characteristic bases. Additionally, we allow the base curve to be semistable, and do not require that it is smooth. We begin by constructing the normal crossings compactifications of configuration spaces and Hurwitz spaces in § B.1. We next introduce various notation for log covers in § B.2. We then reduce our task to proving a certain map is log smooth in § B.3. Finally, we verify the above mentioned map is log smooth in § B.4.

B.1. The normal crossings compactification via twisted stable maps

In order to prove the Hurwitz spaces we consider have a normal crossings compactification, we first define the relevant compactification, in terms of twisted stable maps.

Notation B.1.1.

Let BB be a Deligne-Mumford stack and let π:CB\pi:C\to B be a projective family of nodal curves with geometrically connected fibers of genus gg. For each geometric point bBb\to B, let [Cb][C_{b}] denote the fundamental class of a fiber of π\pi viewed as a 11-cycle.

Fix a divisor ZCZ\subset C which is finite étale of degree dd over BB and contained in the smooth locus of CBC\to B. Fix a finite group GG whose order is invertible on BB and let [C/G][C/G] denote the stack quotient of CC by the trivial GG action. The reader may wish to recall the notion of a twisted stable map being balanced as defined in [AV02, Definition 3.2.4]; colloquially this means the stabilizer action on smoothing parameters on each side of a twisted node are inverse to each other. Let 𝒦g,n+d([C/G],1){\mathcal{K}}_{g,n+d}([C/G],1) denote the moduli stack of balanced twisted stable maps whose SS-points described as follows: Given a map SBS\to B for SS a scheme, 𝒦g,n+d([C/G],1)(S){\mathcal{K}}_{g,n+d}([C/G],1)(S) is the groupoid of representable maps h:𝒳[C/G]h:\mathcal{X}\to[C/G] from an (n+d)(n+d)-pointed balanced twisted curve 𝒳\mathcal{X} such that

  1. (1)

    XSX\to S is the coarse space of 𝒳\mathcal{X} with map f:XCf:X\to C induced by hh,

  2. (2)

    the fibers of XSX\to S have genus gg, and

  3. (3)

    (fs)[Xs]=[Cs](f_{s})_{*}[X_{s}]=[C_{s}] for each geometric point sSs\to S, where [Xs][X_{s}] is the fundamental class of the fiber over sSs\to S.

We note that 𝒦g,n+d([C/G],1){\mathcal{K}}_{g,n+d}([C/G],1) is an algebraic stack proper over BB by [AV02, §8.3 and §8.4].

There is an action of SdS_{d} permuting the final dd marked points of the curve 𝒳\mathcal{X}. The quotient stack [𝒦g,n+d([C/G],1)/Sd][{\mathcal{K}}_{g,n+d}([C/G],1)/S_{d}] parameterizes stable maps with nn marked sections as well as an étale degree dd divisor contained in the smooth locus and disjoint from the nn marked sections. There is an evaluation map 𝒦g,n+d([C/G],1)CBd{\mathcal{K}}_{g,n+d}([C/G],1)\to C^{d}_{B} to the dd-fold fiber product over BB sending an (n+d)(n+d)-pointed map to the image of the final dd sections, and hence we obtain a map π:[𝒦g,n+d([C/G],1)/Sd][CBd/Sd]\pi:[{\mathcal{K}}_{g,n+d}([C/G],1)/S_{d}]\to[C^{d}_{B}/S_{d}]. If [Z]:SpecB[CBd/Sd][Z]:\operatorname{Spec}B\to[C^{d}_{B}/S_{d}] denotes the BB point of ConfC/Bd[CBd/Sd]\operatorname{Conf}_{C/B}^{d}\subset[C^{d}_{B}/S_{d}] corresponding to the finite étale degree dd divisor ZZ, we then define

𝒦g,n([C/G],Z,1):=[𝒦g,n+d([C/G],1)/Sd]×π,[CBd/Sd],[Z]B.\displaystyle{\mathcal{K}}_{g,n}([C/G],Z,1):=[{\mathcal{K}}_{g,n+d}([C/G],1)/S_{d}]\times_{\pi,[C^{d}_{B}/S_{d}],[Z]}B.

In other words, 𝒦g,n([C/G],Z,1){\mathcal{K}}_{g,n}([C/G],Z,1) is the closed substack of [𝒦g,n+d([C/G],1)/Sd][{\mathcal{K}}_{g,n+d}([C/G],1)/S_{d}] so that the degree dd marked divisor maps to ZCZ\subset C.

The following is the main result of this section, which will lead to a normal crossing compactification of Hurwitz space in B.1.3. We will later generalize Theorem B.1.1 to nodal curves in Theorem B.3.1.

Theorem B.1.1.

Let BB be a regular locally noetherian scheme, CBC\to B a smooth projective curve with geometrically connected fibers. Let ZCZ\subset C be a degree dd divisor which is finite étale over BB. The Deligne-Mumford stack 𝒦g,n([C/G],Z,1){\mathcal{K}}_{g,n}([C/G],Z,1) is smooth and proper over BB. Moreover, the locus of points in 𝒦g,n([C/G],Z,1){\mathcal{K}}_{g,n}([C/G],Z,1) corresponding to stable maps with smooth source forms a dense open substack of 𝒦g,n([C/G],Z,1){\mathcal{K}}_{g,n}([C/G],Z,1) with complement a normal crossings divisor.

We will prove this in § B.3.2.

Remark B.1.2.

In the case ZZ is a disjoint union of sections, (which always holds if ZZ has degree 0 or 11,) BB is a scheme, and GG is trivial, one can verify that 𝒦g,n([C/G],Z,1){\mathcal{K}}_{g,n}([C/G],Z,1) is in fact a projective scheme, and not just an algebraic stack. This amounts to verifying that the inertia stack is trivial, which then implies it is projective because it is known the coarse moduli space is projective [AV02, Theorem 1.4.1].

One of our main motivations for proving Theorem B.1.1 is that it provides a normal crossings compactification of Hurwitz spaces of GG covers of CC. In particular, if we take GG to be trivial, it provides a normal crossings compactification of a configuration space of points in CZC-Z. A normal crossings compactification in the case C=1C=\mathbb{P}^{1} and Z=Z=\infty and GG is trivial was given in an ad hoc fashion in [EVW16, Lemma 7.6]. When ZZ is empty and GG is trivial, this normal crossings compactification was given in [FM94]. However, even in the case GG is trivial, CC is arbitrary, and ZZ is nonempty, which is the most important case for the present paper, we do not know of a reference. A normal crossings compactification of a variant of our Hurwitz spaces was constructed in [Moc95, Corollary p. 390-391], also using log geometry.

Corollary B.1.3.

With notation as in 2.4.2 and 2.4.5, both the Hurwitz stack HurC/BG,n,Z,𝒮\operatorname{Hur}^{G,n,Z,\mathcal{S}}_{C/B} and the pointed Hurwitz scheme HurC/BG,n,σZ,𝒮\operatorname{Hur}^{G,n,\sigma\subset Z,\mathcal{S}}_{C/B} are dense opens inside a Deligne Mumford stack which is smooth and proper over BB, such that the complementary divisor is a normal crossings divisor. In particular, taking U:=CZU:=C-Z, the scheme ConfU/Bn\operatorname{Conf}^{n}_{U/B} as defined in 2.4.1 is a dense open subscheme of a smooth proper Deligne-Mumford stack, such that the complement is a normal crossings divisor.

Proof.

There is an action of SnS_{n} on the stack 𝒦g,n([C/G],Z,1){\mathcal{K}}_{g,n}([C/G],Z,1) which permutes the nn marked points. Consider the quotient stack [𝒦g,n([C/G],Z,1)/Sn].[{\mathcal{K}}_{g,n}([C/G],Z,1)/S_{n}]. An appropriate union of components of this quotient stack contains a dense open substack parameterizing those smooth covers of CC, which precisely correspond to points of HurC/BG,n,Z,𝒮\operatorname{Hur}^{G,n,Z,\mathcal{S}}_{C/B}. The complement is a normal crossings divisor by Theorem B.1.1. In the case we mark a section σZ\sigma\subset Z and mark a point of the cover over σ\sigma, we can form an appropriate finite étale cover of [𝒦g,n([C/G],Z,1)/Sn][{\mathcal{K}}_{g,n}([C/G],Z,1)/S_{n}] corresponding to marking a section over σ\sigma, (similar to the construction in 2.4.5), and a union of components of this cover contains HurC/BG,n,σZ,𝒮\operatorname{Hur}^{G,n,\sigma\subset Z,\mathcal{S}}_{C/B} as a dense open substack with complement a normal crossings divisor.

As a special case, taking G=idG=\operatorname{\mathrm{id}}, we obtain that ConfU/Bn\operatorname{Conf}^{n}_{U/B} forms a dense open subscheme of [𝒦g,n(C,Z,1)/Sn][{\mathcal{K}}_{g,n}(C,Z,1)/S_{n}], whose complement is a normal crossings divisor. ∎

B.2. Notation for log covers

In order to prove Theorem B.1.1, we use log deformation theory. The starting point is the observation that every twisted stable map as in B.1.1 can be endowed with the structure of a map of log stacks and this induces a log structure on the space of twisted stable maps itself. To carefully describe these log structures, we require a hefty amount of notation. We begin by describing a log structure on the moduli stack of curves, which corresponds to the divisor parameterizing singular curves. Throughout this section, we will assume all log structures appearing are fine.

Notation B.2.1.

Let ¯g,n+d¯log\overline{\mathscr{M}}_{g,n+\underline{d}}^{\operatorname{log}} denote the log stack whose underlying stack is [¯g,n+d/Sd][\overline{\mathscr{M}}_{g,n+d}/S_{d}], where SdS_{d} acts on the final dd marked points, over Spec[1/|G|]\operatorname{Spec}\mathbb{Z}[1/|G|]; the log structure on ¯g,n+d¯log\overline{\mathscr{M}}_{g,n+\underline{d}}^{\operatorname{log}} is given by the reduced divisor parameterizing singular curves. We note that the points of the underlying stack ¯g,n+d¯\overline{\mathscr{M}}_{g,n+\underline{d}} of ¯g,n+d¯log\overline{\mathscr{M}}_{g,n+\underline{d}}^{\operatorname{log}} parameterize tuples (C,p1,,pn,Z)(C,p_{1},\ldots,p_{n},Z), where CC is a nodal curve, pip_{i} are marked smooth points, and ZZ is a degree dd étale divisor contained in the smooth locus such that KC+Z+piK_{C}+Z+\sum p_{i} is ample. When n=0n=0, we let 𝒞\mathscr{C} denote the universal curve over ¯g,d¯\overline{\mathscr{M}}_{g,\underline{d}}, and let 𝒵𝒞\mathscr{Z}\subset\mathscr{C} denote the distinguished degree dd divisor. We let the finite group GG act trivially on 𝒞\mathscr{C} and [𝒞/G][\mathscr{C}/G] denote the quotient stack.

We next introduce notation to describe various aspects of the geometric points of the stack 𝒦g,n([𝒞/G],𝒵,1){\mathcal{K}}_{g,n}([\mathscr{C}/G],\mathscr{Z},1). See Figure 7 for a picture depicting some of this notation.

Notation B.2.2.

Let SS be a scheme. Let [h:𝒳[C/G],𝒟+]𝒦g,n([𝒞/G],𝒵,1)(S)[h:\mathcal{X}\to[C/G],\mathcal{D}+\mathcal{E}]\in{\mathcal{K}}_{g,n}([\mathscr{C}/G],\mathscr{Z},1)(S) be a point; here we use CC and ZZ denote the pullbacks of 𝒞\mathscr{C} and 𝒵\mathscr{Z} to SS, 𝒳\mathcal{X} to denote the twisted curve, 𝒟𝒳\mathcal{D}\subset\mathcal{X} is a closed substack which is a gerbe over the nn sections in the smooth locus of 𝒳\mathcal{X}, and 𝒳\mathcal{E}\subset\mathcal{X} a substack in the smooth locus of 𝒳\mathcal{X} which is a gerbe over the degree dd divisor mapping to ZCZ\subset C. We also use XX to denote the coarse space of 𝒳\mathcal{X}, and we will write EXE\subset X for the degree dd subscheme of XX corresponding to 𝒳\mathcal{E}\subset\mathcal{X} and DXD\subset X the subscheme corresponding to 𝒟𝒳\mathcal{D}\subset\mathcal{X}. These both lie in the smooth locus of XX and EE maps isomorphically to ZZ.

We use π:𝒳X\pi:\mathcal{X}\to X and ψ:[C/G]C\psi:[C/G]\to C to denote the coarse space maps, and f:XCf:X\to C to denote the map on coarse spaces induced by hh.

Remark B.2.3.

From now on, following B.2.2, we will use the notation CSC\to S for the target of an SS-point of our stable maps. (In particular, this is not to be confused with CBC\to B, which we are replacing by 𝒞¯g,d¯\mathscr{C}\to\overline{\mathscr{M}}_{g,\underline{d}} and CC is the pullback of 𝒞\mathscr{C} to SS.) We note that this is a slight conflict of notation with B.1.1, but the notation CBC\to B there will not come up for us again in the remainder of this section.

Notation B.2.4.

Continuing to use notation as in B.2.2, we suppose SS is of the form V=SpeckV=\operatorname{Spec}k, for kk an algebraically closed field. By B.2.5, we can write XX in the form X=PC~X=P\cup\widetilde{C} satisfying the conditions from B.2.5. In particular, PP is the union of the irreducible components contracted under ff.

We also let WPW\subset P denote the union of irreducible components of PP, whose connected components consist of WjPjW_{j}\subset P_{j} defined as follows: Let PjPP_{j}\subset P denote a connected component of PP which joins s,tC~s,t\in\widetilde{C} mapping to a node in CC. We take WjPjW_{j}\subset P_{j} to be the union of irreducible components of PjP_{j} which are not directly between uu and vv; more formally, we can say these irreducible components of PjP_{j} in WjW_{j} do not correspond to the vertices of the dual graph of PjP_{j} which lie in a minimal path joining the irreducible component meeting uu to the irreducible component meeting vv. For each PjPP_{j}\subset P a connected component of PP mapping to a smooth point of CC, we take WjPjW_{j}\subset P_{j} to be the union of the irreducible components of PjP_{j} which are not directly between the irreducible component on which EE lies and the irreducible component meeting C~\widetilde{C}; more formally the components of WjW_{j} do not correspond to the vertices of the dual graph of PjP_{j} which lie in a minimal path joining the component on which EE lies to the component meeting C~\widetilde{C}.

We define YXY\subset X to denote the union of irreducible components of XX which are not contained in WW. Define ρ:XY\rho:X\to Y and t:YCt:Y\to C so that f=tρf=t\circ\rho and let i:WXi:W\to X denote the inclusion For pXp\in X, let kpk_{p} denote the skyscraper sheaf at a point pp.

Refer to caption
Figure 7. A picture depicting the names we have given to various parts of the coarse space XX of 𝒳\mathcal{X}.

The following lemma was to make sense of B.2.4 above.

Lemma B.2.5.

Using notation for h:𝒳[C/G],ψ:[C/G]C,π:𝒳Xh:\mathcal{X}\to[C/G],\psi:[C/G]\to C,\pi:\mathcal{X}\to X and f:XCf:X\to C as in B.2.4, we have fπ=ψhf\circ\pi=\psi\circ h. Moreover, XX is of the form X=PC~X=P\cup\widetilde{C}, where PP and C~\widetilde{C} satisfy the following conditions:

  1. (1)

    C~\widetilde{C} is a partial normalization of CC at a finite set NN of its nodes,

  2. (2)

    PP is a genus 0 semistable curve,

  3. (3)

    PP is contracted under the map XCX\to C, and

  4. (4)

    any connected component of PP is either contracted to a smooth point of CC, in which case it meets C~\widetilde{C} at a single smooth point, or the component is contracted to a node of CC, in which case it meets C~\widetilde{C} at both preimages of the node.

Proof.

We have that fπ=ψhf\circ\pi=\psi\circ h by the universal property of the coarse space XX. We now show X=C~PX=\widetilde{C}\cup P satisfies the conditions as in the statement. Since XCX\to C has degree 11 on each component of CC, it must be the union of a birational map with several contracted components, and hence must be of the form C~P\widetilde{C}\cup P for C~\widetilde{C} a partial normalization of CC and PP the components which are contracted under the map. To conclude, we wish to show PP has properties (2)(2) and (4)(4). First, since the genus of XX agrees with the genus of CC, each connected component of PP must have genus 0. Continuing to use that the genus of XX agrees with the genus of CC, if a connected component of PP is contracted to a node in CC, it must meet the two preimages of the node in C~\widetilde{C} nodally. Similarly, if a connected component PP is contracted to a smooth point, the only way XX has the same genus as CC is if that component of PP meets the preimage of that point in a single node, as claimed. ∎

Using the preceding notation, we are now ready to describe the relevant log structures on our twisted curves.

Notation B.2.6.

Using B.1.1, B.2.1, and B.2.2, let 𝒦g,n([𝒞/G],𝒵,1)log{\mathcal{K}}_{g,n}([\mathscr{C}/G],\mathscr{Z},1)^{\operatorname{log}}, denote the log stack whose underlying stack is 𝒦g,n([𝒞/G],𝒵,1){\mathcal{K}}_{g,n}([\mathscr{C}/G],\mathscr{Z},1) with the log structure we describe next. For a scheme SS, suppose we have an SS point of 𝒦g,n([𝒞/G],𝒵,1){\mathcal{K}}_{g,n}([\mathscr{C}/G],\mathscr{Z},1), corresponding to a twisted stable map 𝒳[C/G]\mathcal{X}\to[C/G] over SS. We endow (𝒳,𝒳)(S,S)(\mathcal{X},\mathcal{M}_{\mathcal{X}})\to(S,\mathcal{M}_{S}) with the log structure described in [Ols07, §3.10] obtained by viewing 𝒳\mathcal{X} as an nn pointed twisted curve together with a degree dd divisor (so that, in particular, there is a copy of \mathbb{N} in M𝒳M_{\mathcal{X}} over the nn marked gerbes and the degree dd marked gerbe on 𝒳\mathcal{X}). Similarly, CSC\to S has a canonical log structure from [Ols07, §3.10], and we endow [C/G][C/G] with the pullback of this log structure along [C/G]C[C/G]\to C amalgamated with the log structure induced by the Cartier divisor ZZ (so in particular there is a copy of \mathbb{N} along the preimage of ZZ in [C/G][C/G]). We denote this log structure by ([C/G],[C/G])(S,S)([C/G],\mathcal{M}^{\prime}_{[C/G]})\to(S,\mathcal{M}^{\prime}_{S}).

In general S\mathcal{M}^{\prime}_{S} may be different from S\mathcal{M}_{S} when 𝒳\mathcal{X} has more nodes than CC or has twisted nodes lying over the nodes of CC. If XX denotes the coarse space of 𝒳\mathcal{X} with its log structure X\mathcal{M}_{X} (including the nn points and degree dd divisor), and CC has log structure C\mathcal{M}^{\prime}_{C} (including the degree dd divisor), then ff has the structure of a log map (f,f):(X,X)(C,C)(f,f^{\flat}):(X,\mathcal{M}_{X})\to(C,\mathcal{M}^{\prime}_{C}). We now describe the structure of this log map, see also [AMW14, Theorem B.6]. First, after replacing SS with an étale cover, so that ff factors as XYCX\to Y\to C where YCY\to C is a composition of log blowups of CC and XYX\to Y is a contraction of trees of rational curves lying over smooth unmarked points of YY. Using notation which restricts on geometric fibers to that in B.2.4 and Figure 7, YCY\to C is a sequence of log blowups of nodes and expansions of marked sections which contracts the chains of rational curves denoted PYP\cap Y and XYX\to Y contracts the trees of rational curves denoted WW. Now YCY\to C is a morphism of log schemes by construction and XYX\to Y is a morphism of log schemes since WW lies over the strict locus of YY and XYX\to Y is an isomorphism away from WW. Thus the composition is a morphism of log schemes. Then, by composing the coarse space map (𝒳,𝒳)(X,X)(\mathcal{X},\mathcal{M}_{\mathcal{X}})\to(X,\mathcal{M}_{X}) with the above maps, we have a map of log stacks (𝒳,𝒳)(C,C)(\mathcal{X},\mathcal{M}_{\mathcal{X}})\to(C,\mathcal{M}^{\prime}_{C}) over (S,S)(S,S)(S,\mathcal{M}_{S})\to(S,\mathcal{M}^{\prime}_{S}).

Since the log structure on [C/G][C/G] is pulled back from the log structure on CC, we obtain a corresponding commutative diagram

(B.1) (𝒳,𝒳){(\mathcal{X},\mathcal{M}_{\mathcal{X}})}([C/G],[C/G]){([C/G],\mathcal{M}^{\prime}_{[C/G]})}(S,S){(S,\mathcal{M}_{S})}(S,S).{(S,\mathcal{M}^{\prime}_{S}).}

The SS points of 𝒦g,n([𝒞/G],𝒵,1)log{\mathcal{K}}_{g,n}([\mathscr{C}/G],\mathscr{Z},1)^{\operatorname{log}}, comprise all of the above data, with the log structure on SS for such an SS point given by (S,S)(S,\mathcal{M}_{S}).

The injective map SS\mathcal{M}_{S}^{\prime}\to\mathcal{M}_{S} of locally free log structures is not necessarily saturated due to the presence of twisted nodes in 𝒳\mathcal{X} lying over nodes of CC. We let SS′′S\mathcal{M}_{S}^{\prime}\rightarrow\mathcal{M}_{S}^{\prime\prime}\to\mathcal{M}_{S} denote its saturation. Then SS′′\mathcal{M}_{S}^{\prime}\hookrightarrow\mathcal{M}_{S}^{\prime\prime} is a simple extension (see, for example, [Ols07, Definition 1.5]) of locally free log structures.

B.3. Reducing to log smoothness

In this subsection we will show how log smoothness of the map 𝒦g,n([𝒞/G],𝒵,1)log¯g,d¯log{\mathcal{K}}_{g,n}([\mathscr{C}/G],\mathscr{Z},1)^{\operatorname{log}}\to\overline{\mathscr{M}}_{g,\underline{d}}^{\operatorname{log}} implies our main result, Theorem B.1.1. We also deduce a generalization of Theorem B.1.1 where we allow the curve there to be nodal.

Proposition B.3.1.

With notation as in B.2.6, the log algebraic stack 𝒦g,n([𝒞/G],𝒵,1)log{\mathcal{K}}_{g,n}([\mathscr{C}/G],\mathscr{Z},1)^{\operatorname{log}}, is log smooth over ¯g,d¯log\overline{\mathscr{M}}_{g,\underline{d}}^{\operatorname{log}}.

We will return to the proof of B.3.1 in § B.4.1.

We next record a version of Theorem B.1.1 for nodal curves. The reader may refer to B.2.1 and B.2.6 for notation used in the next statement. We say that a log smooth morphism is semistable if it is saturated and the source and target are regular with log structure given by a normal crossings divisors, see [IT14, Remark 3.6.6]. The reader may also wish to consult [AK00, Definition 0.1] and [ALT19, Subsection 4.2.1].

Theorem B.3.1.

𝒦g,n([𝒞/G],𝒵,1)log{\mathcal{K}}_{g,n}([\mathscr{C}/G],\mathscr{Z},1)^{\operatorname{log}} is a normal crossings compactification of the locus of points corresponding to stable maps with smooth source and the log structure induced by the complementary divisor. Moreover, there is a factorization 𝒦g,n([𝒞/G],𝒵,1)log𝛼~𝛽¯g,d¯log{\mathcal{K}}_{g,n}([\mathscr{C}/G],\mathscr{Z},1)^{\operatorname{log}}\xrightarrow{\alpha}\widetilde{\mathscr{M}}\xrightarrow{\beta}\overline{\mathscr{M}}_{g,\underline{d}}^{\operatorname{log}}, where α\alpha is semistable and β\beta is proper, quasifinite, log étale, and birational on each component of the source.

Remark B.3.2.

In the statement of Theorem B.3.1, ~\widetilde{\mathscr{M}} is a union of components of the stack of simple extensions of log structures over ¯g,d¯\overline{\mathscr{M}}_{g,\underline{d}} as in [Ols07, Section 5.2] and it parameterizes certain twisted curves by the proof of [Ols07, Theorem 1.10].

Remark B.3.3.

We note that when |G|=1|G|=1, Theorem B.3.1 reduces to the well known statement that the forgetful map ¯g,n+dlog¯g,dlog\overline{\mathscr{M}}_{g,n+d}^{\operatorname{log}}\to\overline{\mathscr{M}}_{g,d}^{\operatorname{log}} is semistable, where the moduli spaces of curves are equipped with their boundary log structures, parameterizing singular curves. The fiber over a geometric point representing a curve (C,pn+1,,pn+d)(C,p_{n+1},\ldots,p_{n+d}) is a log smooth compactification of the configuration space of nn points on Csm{pn+1,,pn+d}C^{sm}\setminus\{p_{n+1},\ldots,p_{n+d}\}. This agrees with the Fulton-MacPherson compactification given in [FM94] when CC is smooth and d=0d=0.

B.3.2. Proof of Theorem B.1.1, Theorem B.3.1

We begin by explaining why Theorem B.1.1 and the first part of Theorem B.3.1 follow from B.3.1. Let CSC\to S denote either

  1. (1)

    the family from Theorem B.1.1 where S=BS=B is regular and has the trivial log structure or

  2. (2)

    the pullback of the universal family over ¯g,d¯log\overline{\mathscr{M}}_{g,\underline{d}}^{\operatorname{log}} along some strict map S¯g,d¯logS\to\overline{\mathscr{M}}_{g,\underline{d}}^{\operatorname{log}} from a log scheme SS whose map of underlying stacks is étale.

We now verify that SS is log regular in the above two cases. Using [Ill02, 7.3(b)], the log scheme SS log regular in case (1)(1). In case (2)(2), note that Spec[1/|G|]\operatorname{Spec}\mathbb{Z}[1/|G|] with the trivial log structure is log regular by [Ill02, 7.3(b)]. Since ¯g,d¯\overline{\mathscr{M}}_{g,\underline{d}} is log smooth over Spec[1/|G|]\operatorname{Spec}\mathbb{Z}[1/|G|], we obtain SS is also log smooth over Spec[1/|G|]\operatorname{Spec}\mathbb{Z}[1/|G|]. Hence SS is log regular by [Ill02, 7.3(c)].

We next show stable maps to such CSC\to S as above form a normal crossings compactification of the locus of such maps with smooth source. Using B.3.1, we find that 𝒦g,n([C/G],Z,1)log{\mathcal{K}}_{g,n}([C/G],Z,1)^{\operatorname{log}} is log smooth over SS. By [Ill02, 7.3(c)], 𝒦g,n([C/G],Z,1)log{\mathcal{K}}_{g,n}([C/G],Z,1)^{\operatorname{log}} is log regular. Note also that the log structure defined on 𝒦g,n([C/G],Z,1)log{\mathcal{K}}_{g,n}([C/G],Z,1)^{\operatorname{log}} coming from the divisor parameterizing singular covers is pulled back from that on ¯g,n+d¯\overline{\mathscr{M}}_{g,n+\underline{d}}, as follows from [Ols07, Theorem 1.10] and the proof of [Ols07, Lemma 5.1]. At a geometric point xx of 𝒦g,n([C/G],Z,1)log{\mathcal{K}}_{g,n}([C/G],Z,1)^{\operatorname{log}}, the characteristic monoid of the log structure described in [Ols07, §3.10] and the log structure on SS is described two lines before [Ols07, (3.6.6)]. This log structure is identified with n(x)\mathbb{N}^{n(x)}, where n(x)n(x) is the number of nodes of the twisted curve 𝒳\mathcal{X} corresponding to the point xx, so this log structure is locally free. Then, by [Ill02, 7.3(b)], we obtain that the log structure on 𝒦g,n([C/G],Z,1)log{\mathcal{K}}_{g,n}([C/G],Z,1)^{\operatorname{log}} is defined by a normal crossings divisor whose complement is the locus of triviality of the log structure and 𝒦g,n([C/G],Z,1){\mathcal{K}}_{g,n}([C/G],Z,1) is regular. Since the open subset of triviality of the log structure on 𝒦g,n([C/G],Z,1)log{\mathcal{K}}_{g,n}([C/G],Z,1)^{\operatorname{log}} is precisely the locus of covers of curves with smooth source, we find that the above normal crossings divisor is that parameterizing the locus of covers where the source is singular. Finally, the fact that the locus of points in 𝒦g,n([C/G],Z,1){\mathcal{K}}_{g,n}([C/G],Z,1) corresponding to stable maps with smooth source forms a dense open of 𝒦g,n([C/G],Z,1){\mathcal{K}}_{g,n}([C/G],Z,1) follows from [Ill02, 7.3(d)]. This completes the proof of Theorem B.1.1 and the first part of Theorem B.3.1.

We conclude by now proving the second part of Theorem B.3.1. By the last paragraph of B.2.6, any SS-point of 𝒦g,n([C/G],Z,1){\mathcal{K}}_{g,n}([C/G],Z,1) induces a simple extension SS′′\mathcal{M}_{S}^{\prime}\hookrightarrow\mathcal{M}_{S}^{\prime\prime} where S\mathcal{M}_{S}^{\prime} is the pullback of the log structure of ¯g,d¯log\overline{\mathscr{M}}_{g,\underline{d}}^{\operatorname{log}} along S¯g,d¯logS\to\overline{\mathscr{M}}_{g,\underline{d}}^{\operatorname{log}}. Thus there is a map 𝒦g,n([C/G],Z,1)log{\mathcal{K}}_{g,n}([C/G],Z,1)^{\operatorname{log}} to a union of connected components of the stack of simple extensions of log structures ([Ols07, Section 5.2]) over ¯g,d¯log\overline{\mathscr{M}}_{g,\underline{d}}^{\operatorname{log}}, which we will denote by ~\widetilde{\mathscr{M}}. Note that by representability of the map 𝒳[C/G]\mathcal{X}\to[C/G], the order of the simple extension is bounded by |G||G|. Then ~¯g,d¯log\widetilde{\mathscr{M}}\to\overline{\mathscr{M}}_{g,\underline{d}}^{\operatorname{log}} is proper, quasi-finite, log étale, and birational on each component of the source by [Ols07, Lemma 5.3(ii)] and [Kat89, Proposition 3.4]. It follows that 𝒦g,n([C/G],Z,1)log~{\mathcal{K}}_{g,n}([C/G],Z,1)^{\operatorname{log}}\to\widetilde{\mathscr{M}} is a log smooth and saturated morphism where the source and target are regular with normal crossings log structures by the previous paragraph. This completes the proof. ∎

B.4. Verifying log smoothness

In the remainder of this section, we prove B.3.1, stating that 𝒦g,n([𝒞/G],𝒵,1)log¯g,d¯log{\mathcal{K}}_{g,n}([\mathscr{C}/G],\mathscr{Z},1)^{\operatorname{log}}\to\overline{\mathscr{M}}_{g,\underline{d}}^{\operatorname{log}} is log smooth, which will also complete the proof of Theorem B.1.1.

We will approach B.3.1 via deformation theory. To begin understanding the deformation theory, we next describe the map on log cotangent sheaves associated to XCX\to C.

Remark B.4.1.

Given a geometric point VV of 𝒦g,n([C/G],Z,1)log\mathcal{K}_{g,n}([C/G],Z,1)^{\operatorname{log}}, using notation from B.2.4 and B.2.6, (where we use VV for what is called SS there,) there is an associated map (f,f):(X,X)(C,C)(f,f^{\flat}):(X,\mathcal{M}_{X})\to(C,\mathcal{M}^{\prime}_{C}). This induces a map fΩC/VlogΩX/Vlogf^{*}\Omega^{\operatorname{log}}_{C/V}\to\Omega^{\operatorname{log}}_{X/V} which we now describe as a composition of three maps. First, recall that (C,C)(C,\mathcal{M}^{\prime}_{C}) is a log curve over (V,V)(V,\mathcal{M}^{\prime}_{V}). Let C\mathcal{M}_{C} denote the pullback of C\mathcal{M}^{\prime}_{C} along (V,V)(V,V)(V,\mathcal{M}_{V})\to(V,\mathcal{M}^{\prime}_{V}). Then, the map (X,X)(C,C)(X,\mathcal{M}_{X})\to(C,\mathcal{M}^{\prime}_{C}) factors through (X,X)(C,C)(X,\mathcal{M}_{X})\to(C,\mathcal{M}_{C}) and the two versions of the relative logarithmic sheaf of differentials ΩC/Vlog\Omega^{\operatorname{log}}_{C/V} for these two log structures C\mathcal{M}_{C} and C\mathcal{M}^{\prime}_{C} are isomorphic. Hence, to describe fΩC/VlogΩX/Vlogf^{*}\Omega^{\operatorname{log}}_{C/V}\to\Omega^{\operatorname{log}}_{X/V} we can endow CC with the log structure C\mathcal{M}_{C} so that the map of log schemes (X,X)(C,C)(X,\mathcal{M}_{X})\to(C,\mathcal{M}_{C}) is over the fixed base (V,V)(V,\mathcal{M}_{V}). First, there is a map (a,a):(X,X)(X,Xd)(a,a^{\flat}):(X,\mathcal{M}_{X})\to(X,\mathcal{M}_{X}^{d}), where Xd\mathcal{M}_{X}^{d} is the log structure on XX from [Ols07, §3.10] obtained by forgetting the nn marked points and only remembering the degree dd divisor. Next, there is a map (ρ,ρ):(X,Xd)(Y,Y)(\rho,\rho^{\flat}):(X,\mathcal{M}_{X}^{d})\to(Y,\mathcal{M}_{Y}), for YY as in B.2.4 and Y\mathcal{M}_{Y} the log structure on YY, including the degree dd divisor, but not including any of the nn marked points. And finally, we have a map (t,t):(Y,Y)(C,C)(t,t^{\flat}):(Y,\mathcal{M}_{Y})\to(C,\mathcal{M}_{C}). Note that f=tρaf=t\circ\rho\circ a. We can now identify the map fΩC/VlogΩX/Vlogf^{*}\Omega^{\operatorname{log}}_{C/V}\to\Omega^{\operatorname{log}}_{X/V} as the composite map

(B.2) aρtΩC/VlogaρΩY/VlogaΩX/Vlog,dΩX/Vlog\displaystyle a^{*}\rho^{*}t^{*}\Omega^{\operatorname{log}}_{C/V}\to a^{*}\rho^{*}\Omega^{\operatorname{log}}_{Y/V}\to a^{*}\Omega^{\operatorname{log},d}_{X/V}\to\Omega^{\operatorname{log}}_{X/V}

where ΩX/Vlog,d\Omega^{\operatorname{log},d}_{X/V} denotes the relative sheaf of logarithmic differentials associated to the log structure Xd\mathcal{M}_{X}^{d}. Using [Kat00, Proposition 1.13], and the fact that the identification there is functorial for maps of log schemes, we can identify (B.2) with the sequence of maps

(B.3) fωC/V(Z)=ρtωC/V(Z)𝛼ρωY/V(E)𝜀ωX/V(E)𝛿ωX/V(D+E).f^{*}\omega_{C/V}(Z)=\rho^{*}t^{*}\omega_{C/V}(Z)\xrightarrow{\alpha}\rho^{*}\omega_{Y/V}(E)\xrightarrow{\varepsilon}\omega_{X/V}(E)\xrightarrow{\delta}\omega_{X/V}(D+E).

We denote the composite map in (B.3) by ϕ\phi.

To better understand the deformation theory associated to a stable map, our first step will be to understand the map tωC/V(Z)ωY/V(E)t^{*}\omega_{C/V}(Z)\to\omega_{Y/V}(E), whose pullback under ρ\rho is the first map in (B.3).

Lemma B.4.2.

For t:YCt:Y\to C as in B.2.4, there is an isomorphism ωC/V(Z)tωY/V(E)\omega_{C/V}(Z)\simeq t_{*}\omega_{Y/V}(E) as well as an isomorphism tωC/V(Z)ωY/V(E)t^{*}\omega_{C/V}(Z)\simeq\omega_{Y/V}(E).

Proof.

Write Y=C~QY=\widetilde{C}\cup Q, for QQ the union of components of YY not contained in C~\widetilde{C}, for C~\widetilde{C} as in B.2.4. For s:C~QYs:\widetilde{C}\coprod Q\to Y, we have an exact sequence

(B.4) 0{0}ωY/V(E){\omega_{Y/V}(E)}s(ωY/V(E)|C~)s(ωY/V(E)|Q){s_{*}(\omega_{Y/V}(E)|_{\widetilde{C}})\oplus s_{*}(\omega_{Y/V}(E)|_{Q})}𝒪|C~Q{\mathscr{O}|_{\widetilde{C}\cap Q}}0.{0.}

We can think of this sequence as expressing a local section of ωY/V(E)\omega_{Y/V}(E) as a log differential on the normalization of YY with poles along EE and poles along the preimages of the nodes whose corresponding residues sum to zero. Observe that ωY/V|C~ωC~/V(C~Q)\omega_{Y/V}|_{\widetilde{C}}\simeq\omega_{\widetilde{C}/V}(\widetilde{C}\cap Q) and ωY/V|QωQ/V(C~Q)\omega_{Y/V}|_{Q}\simeq\omega_{Q/V}(\widetilde{C}\cap Q). Hence, pushing forward (B.4) along tt, we get an exact sequence

(B.5) 0{0}tωY/V(E){t_{*}\omega_{Y/V}(E)}ts(ωC~/V(C~Q+E|C~))tsωQ/V(C~Q+E|Q){t_{*}s_{*}(\omega_{\widetilde{C}/V}(\widetilde{C}\cap Q+E|_{\widetilde{C}}))\oplus t_{*}s_{*}\omega_{Q/V}(\widetilde{C}\cap Q+E|_{Q})}t𝒪|C~Q.{t_{*}\mathscr{O}|_{\widetilde{C}\cap Q}.}

Now, let M:=t(Q)M:=t(Q). Note that ts(ωQ/V(C~Q+E|Q))t_{*}s_{*}(\omega_{Q/V}(\widetilde{C}\cap Q+E|_{Q})) is supported on MM, which is a disjoint union of points. By construction of QQ, using B.2.5, each connected component of QQ is a chain of 1\mathbb{P}^{1}’s and C~Q+E|Q\widetilde{C}\cap Q+E|_{Q} consists of a degree two subscheme on each such connected component, with a degree 11 point on each component on either end of the chain. Since the dualizing sheaf of 1\mathbb{P}^{1} has degree 2-2, this allows us to identify ωQ/V(C~Q)𝒪Q\omega_{Q/V}(\widetilde{C}\cap Q)\simeq\mathscr{O}_{Q} and hence ts(ωQ/V(C~Q))t_{*}s_{*}(\omega_{Q/V}(\widetilde{C}\cap Q)) is identified with st𝒪Qs_{*}t_{*}\mathscr{O}_{Q}. This is a skyscraper sheaf supported on MM, which we denote kMk_{M}. Hence, the above sequence (B.5) becomes

(B.6) 0{0}tωY/V(E){t_{*}\omega_{Y/V}(E)}tsωC~/V(C~Q+E|C~)kM{t_{*}s_{*}\omega_{\widetilde{C}/V}(\widetilde{C}\cap Q+E|_{\widetilde{C}})\oplus k_{M}}t𝒪C~Q{t_{*}\mathscr{O}_{\widetilde{C}\cap Q}}μ\scriptstyle{\mu}

We claim that this sequence expresses the condition that tωY/V(E)t_{*}\omega_{Y/V}(E) is the subsheaf of tsωC~/V(C~Q+E|C~)t_{*}s_{*}\omega_{\widetilde{C}/V}(\widetilde{C}\cap Q+E|_{\widetilde{C}}) whose poles at preimages of a given node along the normalization map ts|C~:C~Ct\circ s|_{\widetilde{C}}:\widetilde{C}\to C agree. Since ωC/V(Z)\omega_{C/V}(Z) also has this description, this will yield an identification tωY/V(E)ωC/V(Z)t_{*}\omega_{Y/V}(E)\simeq\omega_{C/V}(Z). To verify our claim above, there are two cases. The easier case occurs in the neighborhood of a point of C~Q\widetilde{C}\cap Q mapping to a smooth point of CC. Then, the map locally in a small neighborhood UU of such a point pp is identified with tsωC~/V(C~Q+E|C~)|Ukpkp,(a,b)abt_{*}s_{*}\omega_{\widetilde{C}/V}(\widetilde{C}\cap Q+E|_{\widetilde{C}})|_{U}\oplus k_{p}\to k_{p},(a,b)\mapsto a-b, and the kernel is tsωC~/V(C~Q+E|C~)|Ut_{*}s_{*}\omega_{\widetilde{C}/V}(\widetilde{C}\cap Q+E|_{\widetilde{C}})|_{U}, as claimed. The more difficult case is to compute the kernel at a nodal point of CC. Here, the fiber of μ\mu is identified with a map k2kk2k^{\oplus 2}\oplus k\to k^{\oplus 2} given by (a,c,b)(ab,bc)(a,c,b)\mapsto(a-b,b-c). The first two copies of kk on the source correspond to the residues of the sheaf on the two preimages of the node in C~\widetilde{C} and the third copy of kk corresponds to the section on the contracted component of QQ. Lying in the kernel of this map expresses the condition that the residues on each side of C~\widetilde{C} agree with the value on the contracted component of QQ. Said another way, the values of the residues on each side of C~\widetilde{C} agree. This verifies our claim.

Finally, since we showed above the restriction of ωY/V(E)\omega_{Y/V}(E) to any fiber of YCY\to C is the structure sheaf, the adjoint tωC/V(Z)ωY/V(E)t^{*}\omega_{C/V}(Z)\to\omega_{Y/V}(E) to our isomorphism ωC/V(Z)tωY/V(E)\omega_{C/V}(Z)\simeq t_{*}\omega_{Y/V}(E) restricts to an isomorphism on each contracted fiber of YCY\to C. Since tωC/V(Z)ωY/V(E)t^{*}\omega_{C/V}(Z)\to\omega_{Y/V}(E) also restricts to an isomorphism on C~\widetilde{C}, it is an isomorphism. ∎

We will see later that the log cotangent complex associated to a geometric point of 𝒦g,n([𝒞/G],𝒵,1)log{\mathcal{K}}_{g,n}([\mathscr{C}/G],\mathscr{Z},1)^{\operatorname{log}} as in B.2.4 can be identified with the two-term complex f(ωC/V(Z))ϕωX/V(D+E)f^{*}(\omega_{C/V}(Z))\xrightarrow{\phi}\omega_{X/V}(D+E). The following lemma will therefore help us analyze the deformation theory of 𝒦g,n([𝒞/G],𝒵,1)log{\mathcal{K}}_{g,n}([\mathscr{C}/G],\mathscr{Z},1)^{\operatorname{log}}.

Lemma B.4.3.

We use notation as in B.2.4, where i:WCi:W\to C is the inclusion. With ϕ\phi as defined in B.4.1, kerϕi𝒪W((YW))\ker\phi\simeq i_{*}\mathscr{O}_{W}(-(Y\cap W)).

Proof.

We first describe the map ϕ\phi, which was defined as a composition δεα\delta\circ\varepsilon\circ\alpha in B.4.1, in a more concrete fashion. Let j:YXj:Y\to X denote the inclusion and ρ:XY\rho:X\to Y the map contracting WW. The following statements can be obtained by unwinding the definitions of the maps induced on log differentials, used to define (B.3). The map α:ρtωC/V(Z)ρωY/V(E)\alpha:\rho^{*}t^{*}\omega_{C/V}(Z)\to\rho^{*}\omega_{Y/V}(E) in (B.3) is obtained as the pullback under ρ\rho of the isomorphism tωC/V(Z)ωY/V(E)t^{*}\omega_{C/V}(Z)\to\omega_{Y/V}(E) from B.4.2. The map δ:ωX/V(E)ωX/V(E+D)\delta:\omega_{X/V}(E)\to\omega_{X/V}(E+D) in (B.3) is obtained from twisting the inclusion 𝒪X𝒪X(D)\mathscr{O}_{X}\to\mathscr{O}_{X}(D) by ωX/V(E)\omega_{X/V}(E). Finally, it remains to describe the map ε:ρωY/V(E)ωX/V(E)\varepsilon:\rho^{*}\omega_{Y/V}(E)\to\omega_{X/V}(E) in (B.3). Since ρj=idY\rho\circ j=\operatorname{\mathrm{id}}_{Y}, There is an isomorphism ωY/VρjωY/V\omega_{Y/V}\simeq\rho_{*}j_{*}\omega_{Y/V} which yields by adjunction a map β:ρωY/V(E)jωY/V(E)\beta:\rho^{*}\omega_{Y/V}(E)\to j_{*}\omega_{Y/V}(E). Define the map γ:jωY/V(E)ωX/V(E)\gamma:j_{*}\omega_{Y/V}(E)\to\omega_{X/V}(E) as that obtained via the inclusion j(ωY/V(E))ωX/V(E(YW))ωX/V(E)j_{*}(\omega_{Y/V}(E))\simeq\omega_{X/V}(E-(Y\cap W))\hookrightarrow\omega_{X/V}(E). Then, ε=γβ\varepsilon=\gamma\circ\beta and so the map ϕ\phi is the composite of the maps ρtωC/V(Z)𝛼ρωY/V(E)𝛽jωY/V(E)𝛾ωX/V(E)𝛿ωX/V(E+D).\rho^{*}t^{*}\omega_{C/V}(Z)\xrightarrow{\alpha}\rho^{*}\omega_{Y/V}(E)\xrightarrow{\beta}j_{*}\omega_{Y/V}(E)\xrightarrow{\gamma}\omega_{X/V}(E)\xrightarrow{\delta}\omega_{X/V}(E+D).

We now wish to identify the kernel of ϕ\phi. First, the map α\alpha is an isomorphism by B.4.2. The maps γ\gamma and δ\delta are both injective maps of locally free sheaves by construction. Therefore, we can identify the kernel of ϕ\phi with the kernel of β:ρωY/V(E)jωY/V(E)\beta:\rho^{*}\omega_{Y/V}(E)\to j_{*}\omega_{Y/V}(E). This map β\beta is an isomorphism away from WW, so we only need compute the kernel restricted to WW. On WW the map β\beta restricts to the map 𝒪W𝒪WY|W\mathscr{O}_{W}\to\mathscr{O}_{W\cap Y}|_{W} and so the kernel is indeed 𝒪W(WY)\mathscr{O}_{W}(-W\cap Y). Hence, the kernel of ϕ\phi is i𝒪W(WY)i_{*}\mathscr{O}_{W}(-W\cap Y), as claimed. ∎

We will see that the obstructions to deforming a point of 𝒦g,n([𝒞/G],𝒵,1)log{\mathcal{K}}_{g,n}([\mathscr{C}/G],\mathscr{Z},1)^{\operatorname{log}} as in B.2.4 lie in Ext2(𝕃hlog,𝒪𝒳)\operatorname{Ext}^{2}(\mathbb{L}_{h}^{\operatorname{log}},\mathscr{O}_{\mathcal{X}}), for 𝕃hlog=[h(ω[C/G]/V([Z/G]))ω𝒳/V(𝒟+)]\mathbb{L}_{h}^{\operatorname{log}}=[h^{*}(\omega_{[C/G]/V}([Z/G]))\to\omega_{\mathcal{X}/V}(\mathcal{D}+\mathcal{E})]. Therefore, the next lemma will verify that deformations are unobstructed and hence be used to show 𝒦g,n([𝒞/G],𝒵,1){\mathcal{K}}_{g,n}([\mathscr{C}/G],\mathscr{Z},1) is log smooth over (¯g,n+d¯)log(\overline{\mathscr{M}}_{g,n+\underline{d}})^{\operatorname{log}}.

Lemma B.4.4.

With notation as in B.2.4, let 𝕃hlog=[h(ω[C/G]/V([Z/G]))ω𝒳/V(𝒟+)]\mathbb{L}_{h}^{\operatorname{log}}=[h^{*}(\omega_{[C/G]/V}([Z/G]))\to\omega_{\mathcal{X}/V}(\mathcal{D}+\mathcal{E})] denote the two term complex on 𝒳\mathcal{X} where the first term lies in degree 1-1 and the second in degree 0. Then Ext2(𝕃hlog,𝒪𝒳)=0\operatorname{Ext}^{2}(\mathbb{L}_{h}^{\operatorname{log}},\mathscr{O}_{\mathcal{X}})=0.

Proof.

First, we identify Ext2(𝕃hlog,𝒪𝒳)Ext2(𝕃flog,𝒪X)\operatorname{Ext}^{2}(\mathbb{L}_{h}^{\operatorname{log}},\mathscr{O}_{\mathcal{X}})\simeq\operatorname{Ext}^{2}(\mathbb{L}_{f}^{\operatorname{log}},\mathscr{O}_{X}) where 𝕃flog=[f(ωC/V(Z))ωX/V(D+E)]\mathbb{L}_{f}^{\operatorname{log}}=[f^{*}(\omega_{C/V}(Z))\to\omega_{X/V}(D+E)], also in degrees [1,0][-1,0]. For π:𝒳X\pi:\mathcal{X}\to X the coarse space, and any line bundle \mathscr{L} on XX the adjunction map ππ\mathscr{L}\to\pi_{*}\pi^{*}\mathscr{L} is an isomorphism, as can be verified locally using that 𝒪Xππ𝒪X\mathscr{O}_{X}\to\pi_{*}\pi^{*}\mathscr{O}_{X} is an isomorphism. Hence, because π(ωX/V(D+E))ω𝒳/V(𝒟+)\pi^{*}(\omega_{X/V}(D+E))\simeq\omega_{\mathcal{X}/V}(\mathcal{D}+\mathcal{E}) by [AB23, Proposition 3.11], we find ωX/V(DE)π(ω𝒳/V(𝒟))\omega_{X/V}^{\vee}(-D-E)\simeq\pi_{*}(\omega_{\mathcal{X}/V}^{\vee}(-\mathcal{D}-\mathcal{E})). We also have

πhω[C/G]/VπhψωC/VππfωC/VfωC/V,\displaystyle\pi_{*}h^{*}\omega_{[C/G]/V}^{\vee}\simeq\pi_{*}h^{*}\psi^{*}\omega_{C/V}^{\vee}\simeq\pi_{*}\pi^{*}f^{*}\omega_{C/V}^{\vee}\simeq f^{*}\omega_{C/V}^{\vee},

using that C[C/G]C\to[C/G] is étale, that fπ=ψhf\circ\pi=\psi\circ h by B.2.5, and that ππ𝒪X𝒪X\pi_{*}\pi^{*}\mathscr{O}_{X}\simeq\mathscr{O}_{X}. The above observations yield the third isomorphism in the below chain of isomorphisms:

Ext2(𝕃hlog,𝒪𝒳)\displaystyle\operatorname{Ext}^{2}(\mathbb{L}_{h}^{\operatorname{log}},\mathscr{O}_{\mathcal{X}}) H2(𝒳,ω𝒳/V(𝒟)h(ω[C/G]/V([Z/G])))\displaystyle\simeq H^{2}(\mathcal{X},\omega_{\mathcal{X}/V}^{\vee}(-\mathcal{D}-\mathcal{E})\to h^{*}(\omega_{[C/G]/V}^{\vee}(-[Z/G])))
H2(X,π(ω𝒳/V(𝒟))πh(ω[C/G]/V([Z/G])))\displaystyle\simeq H^{2}(X,\pi_{*}(\omega_{\mathcal{X}/V}^{\vee}(-\mathcal{D}-\mathcal{E}))\to\pi_{*}h^{*}(\omega_{[C/G]/V}^{\vee}(-[Z/G])))
H2(X,ωX/V(DE)fωC/V(Z))\displaystyle\simeq H^{2}(X,\omega_{X/V}^{\vee}(-D-E)\to f^{*}\omega_{C/V}^{\vee}(-Z))
Ext2(𝕃flog,𝒪X).\displaystyle\simeq\operatorname{Ext}^{2}(\mathbb{L}_{f}^{\operatorname{log}},\mathscr{O}_{X}).

Therefore, it suffices to prove Ext2(𝕃flog,𝒪X)=0\operatorname{Ext}^{2}(\mathbb{L}_{f}^{\operatorname{log}},\mathscr{O}_{X})=0.

It follows from B.4.3 that 𝕃flog=[f(ωC/V(Z))ϕωX/V(D+E)]\mathbb{L}_{f}^{\operatorname{log}}=[f^{*}(\omega_{C/V}(Z))\xrightarrow{\phi}\omega_{X/V}(D+E)] sits in the following exact triangle

(B.7) i𝒪W((YW))[1]𝕃flog𝒬[0]\displaystyle i_{*}\mathscr{O}_{W}(-(Y\cap W))[1]\to\mathbb{L}_{f}^{\operatorname{log}}\to\mathscr{Q}[0]\to\qquad

where 𝒬\mathscr{Q} the cokernel of the map f(ωC/V(Z))ϕωX/V(D+E)f^{*}(\omega_{C/V}(Z))\xrightarrow{\phi}\omega_{X/V}(D+E). Applying Hom(,𝒪X)\mathrm{Hom}(\bullet,\mathscr{O}_{X}) to (B.7) and taking the long exact sequence yields the exact sequence

(B.8)   Ext2(𝒬,𝒪X){\operatorname{Ext}^{2}(\mathscr{Q},\mathscr{O}_{X})}Ext2(𝕃flog,𝒪X){\operatorname{Ext}^{2}(\mathbb{L}_{f}^{\operatorname{log}},\mathscr{O}_{X})}Ext2(i𝒪W((YW))[1],𝒪X).{\operatorname{Ext}^{2}(i_{*}\mathscr{O}_{W}(-(Y\cap W))[1],\mathscr{O}_{X}).}

It is therefore enough to show that the first and third terms of (B.8) vanish. In general, by Serre duality, for \mathscr{F} a coherent sheaf on a Gorenstein curve XX, Exti(,𝒪X)\operatorname{Ext}^{i}(\mathscr{F},\mathscr{O}_{X}) is dual to Ext1i(𝒪,ωX/V)H1i(ωX/V)\operatorname{Ext}^{1-i}(\mathscr{O},\mathscr{F}\otimes\omega_{X/V})\simeq H^{1-i}(\mathscr{F}\otimes\omega_{X/V}). From this, it follows that Ext2(𝒬,𝒪X)=0\operatorname{Ext}^{2}(\mathscr{Q},\mathscr{O}_{X})=0, as the 1-1st cohomology of any coherent sheaf vanishes.

To complete the proof, it remains only to show Ext2(i𝒪W((YW))[1],𝒪X)\operatorname{Ext}^{2}(i_{*}\mathscr{O}_{W}(-(Y\cap W))[1],\mathscr{O}_{X}) vanishes. Using Serre duality,

Ext2(i𝒪W((YW))[1],𝒪X)\displaystyle\operatorname{Ext}^{2}(i_{*}\mathscr{O}_{W}(-(Y\cap W))[1],\mathscr{O}_{X}) Ext1(i𝒪W((YW)),𝒪X)\displaystyle\simeq\operatorname{Ext}^{1}(i_{*}\mathscr{O}_{W}(-(Y\cap W)),\mathscr{O}_{X})
H0(X,i𝒪W((YW))ωX/V)\displaystyle\simeq H^{0}(X,i_{*}\mathscr{O}_{W}(-(Y\cap W))\otimes\omega_{X/V})^{\vee}
H0(W,𝒪W((YW))ωX/V|W)\displaystyle\simeq H^{0}(W,\mathscr{O}_{W}(-(Y\cap W))\otimes\omega_{X/V}|_{W})^{\vee}
H0(W,𝒪W((YW))ωW/V(YW))\displaystyle\simeq H^{0}(W,\mathscr{O}_{W}(-(Y\cap W))\otimes\omega_{W/V}(Y\cap W))^{\vee}
H1(W,𝒪W)\displaystyle\simeq H^{1}(W,\mathscr{O}_{W})
=0.\displaystyle=0.

In the final step, we are using that each connected component of WW has arithmetic genus 0, since it is a union of irreducible components of the arithmetic genus 0 curve PP, so H1(W,𝒪W)=0H^{1}(W,\mathscr{O}_{W})=0. ∎

To prove B.3.1, we will discuss the deformation theory needed to deduce log smoothness of 𝒦g,n([𝒞/G],𝒵,1)log(¯g,n+d¯)log{\mathcal{K}}_{g,n}([\mathscr{C}/G],\mathscr{Z},1)^{\operatorname{log}}\to(\overline{\mathscr{M}}_{g,n+\underline{d}})^{\operatorname{log}} from the vanishing demonstrated in B.4.4. Note that 𝒦g,n([𝒞/G],𝒵,1)log{\mathcal{K}}_{g,n}([\mathscr{C}/G],\mathscr{Z},1)^{\operatorname{log}} parameterizes certain log structures on covers of curves, and we next introduce a stack 𝒦g,n([𝒞/G],𝒵,1)\mathcal{L}{\mathcal{K}}_{g,n}([\mathscr{C}/G],\mathscr{Z},1) parameterizing all fine log structures.

Notation B.4.5.

Using notation as in B.2.6 let 𝒦g,n([𝒞/G],𝒵,1)\mathcal{L}{\mathcal{K}}_{g,n}([\mathscr{C}/G],\mathscr{Z},1) denote the stack whose SS-points are tuples (S,(π,π):(𝒳,𝒳)(S,S),(h,h):(𝒳,𝒳)([C/G],[C/G]))(\mathcal{M}_{S},(\pi,\pi^{\flat}):(\mathcal{X},\mathcal{M}_{\mathcal{X}})\to(S,\mathcal{M}_{S}),(h,h^{\flat}):(\mathcal{X},\mathcal{M}_{\mathcal{X}})\to([C/G],\mathcal{M}^{\prime}_{[C/G]})) where S\mathcal{M}_{S} is a fine log structure on SS, π\pi is a family of log twisted curves of type (g,n+d)(g,n+d) and (h,h)(h,h^{\flat}) is a log map such that hh is as in B.1.1. There is a map ι:𝒦g,n([𝒞/G],𝒵,1)log𝒦g,n([𝒞/G],𝒵,1)\iota:{\mathcal{K}}_{g,n}([\mathscr{C}/G],\mathscr{Z},1)^{\log}\to\mathcal{L}{\mathcal{K}}_{g,n}([\mathscr{C}/G],\mathscr{Z},1) which sends an SS point of the source, thought of as a map 𝒳[C/G]\mathcal{X}\to[C/G] with their log structures, as described in B.2.6, to the corresponding point of 𝒦g,n([𝒞/G],𝒵,1)\mathcal{L}{\mathcal{K}}_{g,n}([\mathscr{C}/G],\mathscr{Z},1).

Combining the above lemmas with some deformation theory, we deduce B.3.1.

B.4.1. Proof of B.3.1

We note that 𝒦g,n+d([𝒞/G],1){\mathcal{K}}_{g,n+d}([\mathscr{C}/G],1) is a proper algebraic stack by [AV02, Theorem 1.4.1], and hence 𝒦g,n([𝒞/G],𝒵,1){\mathcal{K}}_{g,n}([\mathscr{C}/G],\mathscr{Z},1) is also a proper algebraic stack. To show 𝒦g,n([𝒞/G],𝒵,1){\mathcal{K}}_{g,n}([\mathscr{C}/G],\mathscr{Z},1) is Deligne-Mumford, it suffices to show 𝒦g,n+d([𝒞/G],1){\mathcal{K}}_{g,n+d}([\mathscr{C}/G],1) is Deligne-Mumford, which follows from [Ols07, Theorem 1.16].

To conclude the proof, we only need to verify that 𝒦g,n([𝒞/G],𝒵,1)log{\mathcal{K}}_{g,n}([\mathscr{C}/G],\mathscr{Z},1)^{\operatorname{log}} is log smooth over (¯g,d¯)log(\overline{\mathscr{M}}_{g,\underline{d}})^{\operatorname{log}}. Let S=SpecAS=\operatorname{Spec}A denote a local Artin scheme over [1/|G|]\mathbb{Z}[1/|G|]. Fix a point [h:𝒳[C/G],𝒟+]𝒦g,n([𝒞/G],𝒵,1)log(S)[h:\mathcal{X}\to[C/G],\mathcal{D}+\mathcal{E}]\in{\mathcal{K}}_{g,n}([\mathscr{C}/G],\mathscr{Z},1)^{\operatorname{log}}(S). Note that by B.2.6, SS has an induced log structure coming from pulling back the log structure from the associated map S(¯g,n+d¯)logS\to(\overline{\mathscr{M}}_{g,n+\underline{d}})^{\operatorname{log}} classifying XX, the coarse space of 𝒳\mathcal{X}. Let T=SpecAT^{\prime}=\operatorname{Spec}A^{\prime} denote a thickening of SS with I:=kerAAI:=\ker A^{\prime}\to A. Suppose AA^{\prime} has residue field κ\kappa, maximal ideal 𝔪\mathfrak{m}, and assume 𝔪I=0\mathfrak{m}I=0. In order to verify formal smoothness, we wish to extend the above SS point to a SS^{\prime} point compatible with the above extension of log structure. First, we claim the obstruction to deforming our SS point above, viewed as a map of log stacks, lies in the hypercohomology group Ext2(𝕃hlog,IA𝒪𝒳)\operatorname{Ext}^{2}(\mathbb{L}_{h}^{\operatorname{log}},I\otimes_{A^{\prime}}\mathscr{O}_{\mathcal{X}}) for 𝕃hlog:=[hΩ[C/G]/SlogΩ𝒳/Slog]\mathbb{L}_{h}^{\operatorname{log}}:=[h^{*}\Omega_{[C/G]/S}^{\operatorname{log}}\to\Omega^{\operatorname{log}}_{\mathcal{X}/S}] in degrees [1,0][-1,0]. (We will soon show this is isomorphic to the complex 𝕃hlog\mathbb{L}_{h}^{\operatorname{log}} as defined in B.4.4.) Indeed by [Ols05, Theorem 8.36(i)], there is a canonical obstruction in Ext2(𝕃hG,IA𝒪𝒳)\operatorname{Ext}^{2}(\mathbb{L}_{h}^{G},I\otimes_{A^{\prime}}\mathscr{O}_{\mathcal{X}}) where 𝕃hG\mathbb{L}_{h}^{G} is Gabber’s cotangent complex, as defined in [Ols05, Definition 8.5]. By [Ols05, Section 8.29] there is a transitivity triangle

Lh𝕃[C/G]/SG𝕃𝒳/SG𝕃hGLh^{*}\mathbb{L}_{[C/G]/S}^{G}\to\mathbb{L}_{\mathcal{X}/S}^{G}\to\mathbb{L}_{h}^{G}

and by [Ols05, Corollary 8.34 and Theorem 1.1(iii)], we can identify the map Lh𝕃[C/G]/SG𝕃𝒳/SGLh^{*}\mathbb{L}_{[C/G]/S}^{G}\to\mathbb{L}_{\mathcal{X}/S}^{G} with 𝕃hlog\mathbb{L}_{h}^{\operatorname{log}}; here we use that log smooth curves are integral and that Lh=hLh^{*}=h^{*} for a locally free sheaf. We next wish to show Ext2(𝕃hlog,IA𝒪𝒳)=0\operatorname{Ext}^{2}(\mathbb{L}_{h}^{\operatorname{log}},I\otimes_{A^{\prime}}\mathscr{O}_{\mathcal{X}})=0. There is an identification Ext2(𝕃hlog,IA𝒪𝒳)Ext2(𝕃h0log,Iκ𝒪𝒳0)\operatorname{Ext}^{2}(\mathbb{L}_{h}^{\operatorname{log}},I\otimes_{A^{\prime}}\mathscr{O}_{\mathcal{X}})\simeq\operatorname{Ext}^{2}(\mathbb{L}_{h_{0}}^{\operatorname{log}},I\otimes_{\kappa}\mathscr{O}_{\mathcal{X}_{0}}), where 𝒳0{\mathcal{X}_{0}} is the base change of 𝒳\mathcal{X} along SpecκSpecA\operatorname{Spec}\kappa\to\operatorname{Spec}A^{\prime} and h0h_{0} is the base change of hh along SpecκSpecA\operatorname{Spec}\kappa\to\operatorname{Spec}A, since II is killed by 𝔪\mathfrak{m}. In order to show this κ\kappa vector space vanishes, we are free to base change to the algebraic closure of κ\kappa. Hence, for the remainder of the proof, we can assume S=V=SpeckS=V=\operatorname{Spec}k is a geometric point as in B.2.4, and we aim to show Ext2(𝕃hlog,𝒪X)=0\operatorname{Ext}^{2}(\mathbb{L}_{h}^{\operatorname{log}},\mathscr{O}_{X})=0.

To verify Ext2(𝕃hlog,𝒪X)=0\operatorname{Ext}^{2}(\mathbb{L}_{h}^{\operatorname{log}},\mathscr{O}_{X})=0, we next claim we can identify Ω𝒳/Vlogω𝒳/V(𝒟+)\Omega^{\operatorname{log}}_{\mathcal{X}/V}\simeq\omega_{\mathcal{X}/V}(\mathcal{D}+\mathcal{E}) and Ω[C/G]/Vlogω[C/G]/V([Z/G])\Omega_{[C/G]/V}^{\operatorname{log}}\simeq\omega_{[C/G]/V}([Z/G]) so that 𝕃hlog[h(ω[C/G]/V([Z/G]))ω𝒳/V(𝒟+)]\mathbb{L}_{h}^{\operatorname{log}}\simeq[h^{*}(\omega_{[C/G]/V}([Z/G]))\to\omega_{\mathcal{X}/V}(\mathcal{D}+\mathcal{E})]. By [Kat00, Proposition 1.13] we can identify ΩX/VlogωX/V(D+E)\Omega^{\operatorname{log}}_{X/V}\simeq\omega_{X/V}(D+E). Then, by [AB23, Proposition 3.11], if π:𝒳X\pi:\mathcal{X}\to X denotes the coarse space map, Ω𝒳/VlogπΩX/VlogπωX/V(D+E)ω𝒳/V(𝒟+)\Omega^{\operatorname{log}}_{\mathcal{X}/V}\simeq\pi^{*}\Omega^{\operatorname{log}}_{X/V}\simeq\pi^{*}\omega_{X/V}(D+E)\simeq\omega_{\mathcal{X}/V}(\mathcal{D}+\mathcal{E}). Arguing similarly, we also obtain Ω[C/G]/Vlogω[C/G]/V([Z/G])\Omega_{[C/G]/V}^{\operatorname{log}}\simeq\omega_{[C/G]/V}([Z/G]). Therefore, the contangent complex 𝕃hlog\mathbb{L}_{h}^{\operatorname{log}} is identified with [h(ω[C/G]/V([Z/G]))ω𝒳/V(𝒟+)][h^{*}(\omega_{[C/G]/V}([Z/G]))\to\omega_{\mathcal{X}/V}(\mathcal{D}+\mathcal{E})]. Now, note that Ext2(𝕃hlog,𝒪𝒳)=0\operatorname{Ext}^{2}(\mathbb{L}_{h}^{\operatorname{log}},\mathscr{O}_{\mathcal{X}})=0, by B.4.4.

We are nearly done, and it only remains to explain why the vanishing of the obstruction space Ext2(𝕃hlog,𝒪𝒳)\operatorname{Ext}^{2}(\mathbb{L}_{h}^{\operatorname{log}},\mathscr{O}_{\mathcal{X}}) actually implies log smoothness of 𝒦g,n([𝒞/G],𝒵,1)log(¯g,d¯)log{\mathcal{K}}_{g,n}([\mathscr{C}/G],\mathscr{Z},1)^{\operatorname{log}}\to(\overline{\mathscr{M}}_{g,\underline{d}})^{\operatorname{log}}. To this end, let og¯g,d¯\mathcal{L}og_{\overline{\mathscr{M}}_{g,\underline{d}}} denote the algebraic stack classifying fine log schemes over ¯g,d¯\overline{\mathscr{M}}_{g,\underline{d}}, as defined in [Ols03, Section 5], and, in particular, [Ols03, Proposition 5.9]. Using B.4.5, the log structure on 𝒦g,n([𝒞/G],𝒵,1)log{\mathcal{K}}_{g,n}([\mathscr{C}/G],\mathscr{Z},1)^{\operatorname{log}} induces maps 𝒦g,n([𝒞/G],𝒵,1)𝜄𝒦g,n([𝒞/G],𝒵,1)𝜁og¯g,d¯{\mathcal{K}}_{g,n}([\mathscr{C}/G],\mathscr{Z},1)\xrightarrow{\iota}{\mathcal{L}}{\mathcal{K}}_{g,n}([\mathscr{C}/G],\mathscr{Z},1)\xrightarrow{\zeta}\mathcal{L}og_{\overline{\mathscr{M}}_{g,\underline{d}}}. The vanishing of Ext2(𝕃hlog,𝒪𝒳)\operatorname{Ext}^{2}(\mathbb{L}_{h}^{\operatorname{log}},\mathscr{O}_{\mathcal{X}}) implies the map ζ\zeta above is formally smooth. The log structure from B.2.6 is the minimal log structure of the log map in the sense of [Wis16, p. 724] and so [Wis16, Theorem B.2] implies ι\iota above is an open embedding. Therefore, the composite ζι\zeta\circ\iota is formally smooth, hence smooth. It is shown in [Ols03, Theorem 4.6(ii) and (iii)] that if (W,W)(W,\mathcal{M}_{W}) is a scheme with fine log structure then (W,W)(¯g,d¯,¯g,d¯)(W,\mathcal{M}_{W})\to(\overline{\mathscr{M}}_{g,\underline{d}},\mathcal{M}_{\overline{\mathscr{M}}_{g,\underline{d}}}) is log smooth if Wog¯g,d¯W\to\mathcal{L}og_{\overline{\mathscr{M}}_{g,\underline{d}}} is smooth. From this, one can easily deduce the same holds in the case that (W,W)(W,\mathcal{M}_{W}) is an algebraic stack with fine log structure by passing to a smooth cover of WW by a scheme. Hence, we obtain that 𝒦g,n([𝒞/G],𝒵,1)log{\mathcal{K}}_{g,n}([\mathscr{C}/G],\mathscr{Z},1)^{\operatorname{log}} is log smooth over (¯g,d¯)log(\overline{\mathscr{M}}_{g,\underline{d}})^{\operatorname{log}}, completing the proof. ∎

References

  • [AB23] Kenneth Ascher and Dori Bejleri. Smoothability of relative stable maps to stacky curves. Épijournal Géom. Algébrique, 7:Art. 2, 22, 2023.
  • [ACG11] Enrico Arbarello, Maurizio Cornalba, and Pillip A. Griffiths. Geometry of algebraic curves. Volume II, volume 268 of Grundlehren der Mathematischen Wissenschaften [Fundamental Principles of Mathematical Sciences]. Springer, Heidelberg, 2011. With a contribution by Joseph Daniel Harris.
  • [ACGS20] Dan Abramovich, Qile Chen, Mark Gross, and Bernd Siebert. Punctured logarithmic maps. arXiv preprint arXiv:2009.07720v2, 2020.
  • [Ach08] Jeffrey D. Achter. Results of Cohen-Lenstra type for quadratic function fields. In Computational arithmetic geometry, volume 463 of Contemp. Math., pages 1–7. Amer. Math. Soc., Providence, RI, 2008.
  • [Ach23] Niven Achenjang. The average size of 2-selmer groups of elliptic curves in characteristic 2. arXiv preprint arXiv:2310.08493v2, 2023.
  • [ACV03] Dan Abramovich, Alessio Corti, and Angelo Vistoli. Twisted bundles and admissible covers. volume 31, pages 3547–3618. 2003. Special issue in honor of Steven L. Kleiman.
  • [AK00] D. Abramovich and K. Karu. Weak semistable reduction in characteristic 0. Invent. Math., 139(2):241–273, 2000.
  • [ALT19] Karim Adiprasito, Gaku Liu, and Michael Temkin. Semistable reduction in characteristic 0, 2019.
  • [AMW14] Dan Abramovich, Steffen Marcus, and Jonathan Wise. Comparison theorems for Gromov-Witten invariants of smooth pairs and of degenerations. Ann. Inst. Fourier (Grenoble), 64(4):1611–1667, 2014.
  • [AV02] Dan Abramovich and Angelo Vistoli. Compactifying the space of stable maps. J. Amer. Math. Soc., 15(1):27–75, 2002.
  • [BDPW23] Jonas Bergström, Adrian Diaconu, Dan Petersen, and Craig Westerland. Hyperelliptic curves, the scanning map, and moments of families of quadratic l-functions. arXiv preprint arXiv:2302.07664v2, 2023.
  • [Bel04] Paolo Bellingeri. On presentations of surface braid groups. J. Algebra, 274(2):543–563, 2004.
  • [Bis19] Matthew Bisatt. Explicit root numbers of abelian varieties. Trans. Amer. Math. Soc., 372(11):7889–7920, 2019.
  • [BKL+15] Manjul Bhargava, Daniel M. Kane, Hendrik W. Lenstra, Jr., Bjorn Poonen, and Eric Rains. Modeling the distribution of ranks, Selmer groups, and Shafarevich-Tate groups of elliptic curves. Camb. J. Math., 3(3):275–321, 2015.
  • [BKLOS19] Manjul Bhargava, Zev Klagsbrun, Robert J. Lemke Oliver, and Ari Shnidman. 3-isogeny Selmer groups and ranks of abelian varieties in quadratic twist families over a number field. Duke Math. J., 168(15):2951–2989, 2019.
  • [BLR90] Siegfried Bosch, Werner Lütkebohmert, and Michel Raynaud. Néron models, volume 21 of Ergebnisse der Mathematik und ihrer Grenzgebiete (3) [Results in Mathematics and Related Areas (3)]. Springer-Verlag, Berlin, 1990.
  • [BM23] Andrea Bianchi and Jeremy Miller. Polynomial stability of the homology of hurwitz spaces. arXiv preprint arXiv:2303.11194v1, 2023.
  • [BS13a] Manjul Bhargava and Arul Shankar. The average number of elements in the 4-selmer groups of elliptic curves is 7. arXiv preprint arXiv:1312.7333v1, 2013.
  • [BS13b] Manjul Bhargava and Arul Shankar. The average size of the 5-selmer group of elliptic curves is 6, and the average rank is less than 1. arXiv preprint arXiv:1312.7859v1, 2013.
  • [BS15a] Manjul Bhargava and Arul Shankar. Binary quartic forms having bounded invariants, and the boundedness of the average rank of elliptic curves. Ann. of Math. (2), 181(1):191–242, 2015.
  • [BS15b] Manjul Bhargava and Arul Shankar. Ternary cubic forms having bounded invariants, and the existence of a positive proportion of elliptic curves having rank 0. Ann. of Math. (2), 181(2):587–621, 2015.
  • [BS23] Andrea Bianchi and Andreas Stavrou. Homology of configuration spaces of surfaces modulo an odd prime. arXiv preprint arXiv:2307.08664v1, 2023.
  • [BSS21] Manjul Bhargava, Arul Shankar, and Ashvin Swaminathan. The second moment of the size of the 22-selmer group of elliptic curves. arXiv preprint arXiv:2110.09063v1, 2021.
  • [BSW15] Manjul Bhargava, Arul Shankar, and Xiaoheng Wang. Geometry-of-numbers methods over global fields I: Prehomogeneous vector spaces. arXiv preprint arXiv:1512.03035v1, 2015.
  • [Cad07] Charles Cadman. Using stacks to impose tangency conditions on curves. Amer. J. Math., 129(2):405–427, 2007.
  • [Ces16] Kestutis Cesnavicius. Selmer groups as flat cohomology groups. J. Ramanujan Math. Soc., 31(1):31–61, 2016.
  • [Cha97] Nick Chavdarov. The generic irreducibility of the numerator of the zeta function in a family of curves with large monodromy. Duke Math. J., 87(1):151–180, 1997.
  • [Cha23] Kevin Chang. Hurwitz spaces, nichols algebras, and igusa zeta functions. arXiv preprint arXiv:2306.10446, 2023.
  • [CLQR04] John Cremona, Joan-Carles Lario, Jordi Quer, and Kenneth Ribet, editors. Modular curves and abelian varieties, volume 224 of Progress in Mathematics. Birkhäuser Verlag, Basel, 2004. Papers from the conference held in Bellaterra, July 15–18, 2002.
  • [Con14] Brian Conrad. Reductive group schemes. In Autour des schémas en groupes. Vol. I, volume 42/43 of Panor. Synthèses, pages 93–444. Soc. Math. France, Paris, 2014.
  • [Det08] Michael Dettweiler. On the middle convolution of local systems. with an appendix by m. dettweiler and s. reiter. arXiv preprint arXiv:0810.3334v1, 2008.
  • [dJ02] A. J. de Jong. Counting elliptic surfaces over finite fields. Mosc. Math. J., 2(2):281–311, 2002. Dedicated to Yuri I. Manin on the occasion of his 65th birthday.
  • [DS23] Ariel Davis and Tomer M Schlank. The hilbert polynomial of quandles and colorings of random links. arXiv preprint arXiv:2304.08314v1, 2023.
  • [ELS20] Jordan S. Ellenberg, Wanlin Li, and Mark Shusterman. Nonvanishing of hyperelliptic zeta functions over finite fields. Algebra Number Theory, 14(7):1895–1909, 2020.
  • [ETW17] Jordan S Ellenberg, TriThang Tran, and Craig Westerland. Fox-neuwirth-fuks cells, quantum shuffle algebras, and malle’s conjecture for function fields. arXiv preprint arXiv:1701.04541v2, 2017.
  • [EVW16] Jordan S. Ellenberg, Akshay Venkatesh, and Craig Westerland. Homological stability for Hurwitz spaces and the Cohen-Lenstra conjecture over function fields. Ann. of Math. (2), 183(3):729–786, 2016.
  • [FGI+05] Barbara Fantechi, Lothar Göttsche, Luc Illusie, Steven L. Kleiman, Nitin Nitsure, and Angelo Vistoli. Fundamental algebraic geometry, volume 123 of Mathematical Surveys and Monographs. American Mathematical Society, Providence, RI, 2005. Grothendieck’s FGA explained.
  • [FK88] Eberhard Freitag and Reinhardt Kiehl. Étale cohomology and the Weil conjecture, volume 13 of Ergebnisse der Mathematik und ihrer Grenzgebiete (3) [Results in Mathematics and Related Areas (3)]. Springer-Verlag, Berlin, 1988. Translated from the German by Betty S. Waterhouse and William C. Waterhouse, With an historical introduction by J. A. Dieudonné.
  • [Fla90] Matthias Flach. A generalisation of the Cassels-Tate pairing. J. Reine Angew. Math., 412:113–127, 1990.
  • [FLR23] Tony Feng, Aaron Landesman, and Eric M. Rains. The geometric distribution of Selmer groups of elliptic curves over function fields. Math. Ann., 387(1-2):615–687, 2023.
  • [FM94] William Fulton and Robert MacPherson. A compactification of configuration spaces. Ann. of Math. (2), 139(1):183–225, 1994.
  • [FS16] Jason Fulman and Dennis Stanton. On the distribution of the number of fixed vectors for the finite classical groups. Ann. Comb., 20(4):755–773, 2016.
  • [Gol79] Dorian Goldfeld. Conjectures on elliptic curves over quadratic fields. In Number theory, Carbondale 1979 (Proc. Southern Illinois Conf., Southern Illinois Univ., Carbondale, Ill., 1979), volume 751 of Lecture Notes in Math, pages pp 108–118. Springer, Berlin, 1979.
  • [Gre10] Aaron Greicius. Elliptic curves with surjective adelic Galois representations. Experiment. Math., 19(4):495–507, 2010.
  • [Gro68] Alexander Grothendieck. Le groupe de Brauer. III. Exemples et compléments. In Dix exposés sur la cohomologie des schémas, volume 3 of Adv. Stud. Pure Math., pages 88–188. North-Holland, Amsterdam, 1968.
  • [Gro23] Mark Gross. Remarks on gluing punctured logarithmic maps. arXiv preprint arXiv:2306.02661v1, 2023.
  • [Hal08] Chris Hall. Big symplectic or orthogonal monodromy modulo ll. Duke Math. J., 141(1):179–203, 2008.
  • [HB93] D. R. Heath-Brown. The size of Selmer groups for the congruent number problem. Invent. Math., 111(1):171–195, 1993.
  • [HB94] D. R. Heath-Brown. The size of Selmer groups for the congruent number problem. II. Invent. Math., 118(2):331–370, 1994. With an appendix by P. Monsky.
  • [HLHN14] Q. P. Hồ, V. B. Lê Hùng, and B. C. Ngô. Average size of 2-Selmer groups of elliptic curves over function fields. Math. Res. Lett., 21(6):1305–1339, 2014.
  • [Hoa23] Anh Trong Nam Hoang. Fox-neuwirth cells, quantum shuffle algebras, and character sums of the resultant. arXiv preprint arXiv:2308.01410v1, 2023.
  • [HS23] David Holmes and Pim Spelier. Logarithmic cohomological field theories. arXiv preprint arXiv:2308.01099v2, 2023.
  • [HW10] Allen Hatcher and Nathalie Wahl. Stabilization for mapping class groups of 3-manifolds. Duke Math. J., 155(2):205–269, 2010.
  • [Ill02] Luc Illusie. An overview of the work of K. Fujiwara, K. Kato, and C. Nakayama on logarithmic étale cohomology. Number 279, pages 271–322. 2002. Cohomologies pp-adiques et applications arithmétiques, II.
  • [IT14] Luc Illusie and Michael Temkin. Exposé X. Gabber’s modification theorem (log smooth case). Number 363-364, pages 167–212. 2014. Travaux de Gabber sur l’uniformisation locale et la cohomologie étale des schémas quasi-excellents.
  • [Kan13] Daniel Kane. On the ranks of the 2-Selmer groups of twists of a given elliptic curve. Algebra Number Theory, 7(5):1253–1279, 2013.
  • [Kat89] K. Kato. Logarithmic structures of Fontaine-Illusie. In Algebraic analysis, geometry, and number theory (Baltimore, MD, 1988), pages 191–224. Johns Hopkins Univ. Press, Baltimore, MD, 1989.
  • [Kat96] Nicholas M. Katz. Rigid local systems, volume 139 of Annals of Mathematics Studies. Princeton University Press, Princeton, NJ, 1996.
  • [Kat00] Fumiharu Kato. Log smooth deformation and moduli of log smooth curves. Internat. J. Math., 11(2):215–232, 2000.
  • [Kat02] Nicholas M. Katz. Twisted LL-functions and monodromy, volume 150 of Annals of Mathematics Studies. Princeton University Press, Princeton, NJ, 2002.
  • [KMR13] Zev Klagsbrun, Barry Mazur, and Karl Rubin. Disparity in Selmer ranks of quadratic twists of elliptic curves. Ann. of Math. (2), 178(1):287–320, 2013.
  • [Kow06] E. Kowalski. On the rank of quadratic twists of elliptic curves over function fields. Int. J. Number Theory, 2(2):267–288, 2006.
  • [Lan21] Aaron Landesman. The geometric average size of Selmer groups over function fields. Algebra Number Theory, 15(3):673–709, 2021.
  • [Lau81] G. Laumon. Semi-continuité du conducteur de Swan (d’après P. Deligne). In The Euler-Poincaré characteristic (French), volume 83 of Astérisque, pages 173–219. Soc. Math. France, Paris, 1981.
  • [LPW09] David A. Levin, Yuval Peres, and Elizabeth L. Wilmer. Markov chains and mixing times. American Mathematical Society, Providence, RI, 2009. With a chapter by James G. Propp and David B. Wilson.
  • [LST20] Michael Lipnowski, Will Sawin, and Jacob Tsimerman. Cohen-lenstra heuristics and bilinear pairings in the presence of roots of unity. arXiv preprint arXiv:2007.12533v1, 2020.
  • [LT19] Michael Lipnowski and Jacob Tsimerman. Cohen-Lenstra heuristics for étale group schemes and symplectic pairings. Compos. Math., 155(4):758–775, 2019.
  • [Mil80] James S. Milne. Étale cohomology, volume 33 of Princeton Mathematical Series. Princeton University Press, Princeton, N.J., 1980.
  • [Moc95] Shinichi Mochizuki. The geometry of the compactification of the Hurwitz scheme. Publ. Res. Inst. Math. Sci., 31(3):355–441, 1995.
  • [MPPRW24] Jeremy Miller, Peter Patzt, Dan Petersen, and Oscar Randal-Williams. Uniform twisted homological stability. arXiv preprint arXiv:2402.00354v1, 2024.
  • [NW22] Hoi H Nguyen and Melanie Matchett Wood. Local and global universality of random matrix cokernels. arXiv preprint arXiv:2210.08526v1, 2022.
  • [Ols03] Martin C. Olsson. Logarithmic geometry and algebraic stacks. Ann. Sci. École Norm. Sup. (4), 36(5):747–791, 2003.
  • [Ols05] Martin C. Olsson. The logarithmic cotangent complex. Math. Ann., 333(4):859–931, 2005.
  • [Ols07] Martin C. Olsson. (Log) twisted curves. Compos. Math., 143(2):476–494, 2007.
  • [Par12] Brett Parker. Log geometry and exploded manifolds. Abh. Math. Semin. Univ. Hambg., 82(1):43–81, 2012.
  • [PR12] Bjorn Poonen and Eric Rains. Random maximal isotropic subspaces and Selmer groups. J. Amer. Math. Soc., 25(1):245–269, 2012.
  • [PW23] Sun Woo Park and Niudun Wang. On the Average of p-Selmer Ranks in Quadratic Twist Families of Elliptic Curves Over Global Function Fields. International Mathematics Research Notices, page rnad095, 05 2023.
  • [Riz97] Ottavio Giulio Rizzo. On the variation of root numbers in families of elliptic curves. ProQuest LLC, Ann Arbor, MI, 1997. Thesis (Ph.D.)–Brown University.
  • [Riz99] Ottavio G. Rizzo. Average root numbers in families of elliptic curves. Proc. Amer. Math. Soc., 127(6):1597–1603, 1999.
  • [Riz03] Ottavio G. Rizzo. Average root numbers for a nonconstant family of elliptic curves. Compositio Math., 136(1):1–23, 2003.
  • [RW20] Oscar Randal-Williams. Homology of Hurwitz spaces and the Cohen-Lenstra heuristic for function fields [after Ellenberg, Venkatesh, and Westerland]. Astérisque, (422):Exp. No. 1164, 469–497, 2020.
  • [RWW17] Oscar Randal-Williams and Nathalie Wahl. Homological stability for automorphism groups. Adv. Math., 318:534–626, 2017.
  • [Sab13] Maria Sabitova. Twisted root numbers and ranks of abelian varieties. J. Comb. Number Theory, 5(1):25–30, 2013.
  • [Saw20] Will Sawin. Identifying measures on non-abelian groups and modules by their moments via reduction to a local problem. arXiv preprint arXiv:2006.04934v3, 2020.
  • [SD08] Peter Swinnerton-Dyer. The effect of twisting on the 2-Selmer group. Math. Proc. Cambridge Philos. Soc., 145(3):513–526, 2008.
  • [SGA72] Théorie des topos et cohomologie étale des schémas. Lecture Notes in Mathematics, Vol. 269. Springer-Verlag, Berlin, 1972. Séminaire de Géométrie Algébrique du Bois-Marie 1963–1964 (SGA 4), Dirigé par M. Artin, A. Grothendieck, et J. L. Verdier. Avec la collaboration de N. Bourbaki, P. Deligne et B. Saint-Donat.
  • [Smi22] Alexander Smith. The distribution of \ell^{\infty}-selmer groups in degree \ell twist families i. arXiv preprint arXiv:2207.05674v2, 2022.
  • [Sta] The Stacks Project Authors. Stacks Project. http://stacks.math.columbia.edu.
  • [Sun12] Shenghao Sun. LL-series of Artin stacks over finite fields. Algebra Number Theory, 6(1):47–122, 2012.
  • [SW23] Will Sawin and Melanie Matchett Wood. Conjectures for distributions of class groups of extensions of number fields containing roots of unity. arXiv preprint arXiv:2301.00791v1, 2023.
  • [Tat63] John Tate. Duality theorems in Galois cohomology over number fields. In Proc. Internat. Congr. Mathematicians (Stockholm, 1962), pages 288–295. Inst. Mittag-Leffler, Djursholm, 1963.
  • [Tay92] Donald E. Taylor. The geometry of the classical groups, volume 9 of Sigma Series in Pure Mathematics. Heldermann Verlag, Berlin, 1992.
  • [Tho19] Jack A. Thorne. On the average number of 2-Selmer elements of elliptic curves over 𝔽q(X)\mathbb{F}_{q}(X) with two marked points. Doc. Math., 24:1179–1223, 2019.
  • [TY14] Fabien Trihan and Seidai Yasuda. The \ell-parity conjecture for abelian varieties over function fields of characteristic p>0p>0. Compos. Math., 150(4):507–522, 2014.
  • [Vas03] A. Vasiu. Surjectivity criteria for pp-adic representations. I. Manuscripta Math., 112(3):325–355, 2003.
  • [Ver67] J.-L Verdier. A duality theorem in the etale cohomology of schemes. In Proc. Conf. Local Fields (Driebergen, 1966), pages 184–198. Springer, Berlin, 1967.
  • [Wew98] Stefan Wewers. Construction of Hurwitz spaces. Institut für Experimentelle Mathematik Essen, Ph.D. thesis, 1998.
  • [Wew99] Stefan Wewers. Deformation of tame admissible covers of curves. In Aspects of Galois theory (Gainesville, FL, 1996), volume 256 of London Math. Soc. Lecture Note Ser., pages 239–282. Cambridge Univ. Press, Cambridge, 1999.
  • [Wil09] Robert A. Wilson. The finite simple groups, volume 251 of Graduate Texts in Mathematics. Springer-Verlag London, Ltd., London, 2009.
  • [Wis16] Jonathan Wise. Moduli of morphisms of logarithmic schemes. Algebra Number Theory, 10(4):695–735, 2016.
  • [Woo17] Melanie Matchett Wood. The distribution of sandpile groups of random graphs. J. Amer. Math. Soc., 30(4):915–958, 2017.
  • [Woo21] Melanie Matchett Wood. An algebraic lifting invariant of Ellenberg, Venkatesh, and Westerland. Res. Math. Sci., 8(2):Paper No. 21, 13, 2021.
  • [Woo22] Melanie Matchett Wood. Probability theory for random groups arising in number theory. In Proc. Int. Cong. Math, volume 6, pages 4476–4508, 2022.
  • [Zyw14] David Zywina. The inverse galois problem for orthogonal groups. arXiv preprint arXiv:1409.1151v1, 2014.