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

Reflection theorems for number rings

Evan O’Dorney
Abstract

The Ohno-Nakagawa reflection theorem is an unexpectedly simple identity relating the number of GL2\mathrm{GL}_{2}\mathbb{Z}-classes of binary cubic forms (equivalently, cubic rings) of two different discriminants DD, 27D-27D; it generalizes cubic reciprocity and the Scholz reflection theorem. In this paper, we provide a framework for generalizing this theorem using a global and local step. The global step uses Fourier analysis on the adelic cohomology H1(𝔸K,M)H^{1}(\mathbb{A}_{K},M) of a finite Galois module, modeled after the celebrated Fourier analysis on 𝔸K\mathbb{A}_{K} used in Tate’s thesis. The local step is combinatorial, more elementary but much more mysterious. We establish reflection theorems for binary quadratic forms over number fields of class number 11, and for cubic and quartic rings over arbitrary number fields, as well as binary quartic forms over \mathbb{Z}; the quartic results are conditional on some computational algebraic identities that are probabilistically true. Along the way, we find elegant new results on Igusa zeta functions of conics and the average value of a quadratic character over a box in a local field.

Part I Introduction

1 Introduction

1.1 Historical background

In 1932, using the then-new machinery of class field theory, Scholz [ScholzRefl] proved that the class groups of the quadratic fields (D)\mathbb{Q}(\sqrt{D}) and (3D)\mathbb{Q}(\sqrt{-3D}), whose discriminants are in the ratio 3-3, have 33-ranks differing by at most 11. This is a remarkable early example of a reflection theorem. A generalization due to Leopoldt [Leopoldt] relates different components of the pp-torsion of the class group of a number field containing μp\mu_{p} when decomposed under the Galois group of that field. Applications of such reflection theorems are far-ranging: for instance, Ellenberg and Venkatesh [EV] use reflection theorems of Scholz type to prove upper bounds on \ell-torsion in class groups of number fields, while Mihăilescu [MihCat2] uses Leopoldt’s generalization to simplify a step of his monumental proof of the Catalan conjecture that 88 and 99 are the only consecutive perfect powers. Through the years, numerous reflection principles for different generalizations of ideal class groups have come into print. A very general reflection theorem for Arakelov class groups is due by Gras [Gras].

A quite different direction of generalization was discovered by accident in 1997: The following relation was conjectured by Ohno [Ohno] on the basis of numerical data and proved by Nakagawa [Nakagawa], for which reason we will call it the Ohno-Nakagawa (O-N) reflection theorem:

Theorem 1.1 (Ohno–Nakagawa).

For a nonzero integer DD, let h(D)h(D) be the number of GL2()\mathrm{GL}_{2}(\mathbb{Z})-orbits of binary cubic forms

f(x,y)=ax3+bx2y+cxy2+dy3f(x,y)=ax^{3}+bx^{2}y+cxy^{2}+dy^{3}

of discriminant DD, each orbit weighted by the reciprocal of its number of symmetries (i.e. stabilizer in GL2()\mathrm{GL}_{2}(\mathbb{Z})). Let h3(D)h_{3}(D) be the number of such orbits f(x,y)f(x,y) such that the middle two coefficients b,cb,c are multiples of 33, weighted in the same way.

Then for every nonzero integer DD, we have the exact identity

h3(27D)={3h(D),D>0h(D),D<0.h_{3}(-27D)=\begin{cases}3h(D),&D>0\\ h(D),&D<0.\end{cases} (1)

By the well-known index-form parametrization (see 6.9 below), h(D)h(D) also counts the cubic rings of discriminant DD over \mathbb{Z}, weighted by the reciprocal of the order of the automorphism group. It turns out that h3(D)h_{3}(D) counts those rings CC for which 3|trC/ξ3|\operatorname{tr}_{C/\mathbb{Z}}\xi for every ξC\xi\in C. When DD is a fundamental discriminant, the corresponding cubic extensions are closely related, via class field theory, to the 33-class group of (D)\mathbb{Q}(\sqrt{D}) and we get back Scholz’s reflection theorem, as Nakagawa points out ([Nakagawa], Remark 0.9).

Theorem 1.1 was quite unexpected, because GL2()\mathrm{GL}_{2}(\mathbb{Z})-orbits of binary cubics have been tabulated since Eisenstein without unearthing any striking patterns. Even the exact normalizations h(D)h(D), h3(D)h_{3}(D) had been in use for over two decades. They appear in the Shintani zeta functions

ζ±(s)\displaystyle\zeta^{\pm}(s) =n1h(±n)ns\displaystyle=\sum_{n\geq 1}\frac{h(\pm n)}{n^{s}}
ζ^±(s)\displaystyle\hat{\zeta}^{\pm}(s) =n1h3(±27n)ns,\displaystyle=\sum_{n\geq 1}\frac{h_{3}(\pm 27n)}{n^{s}},

a family of Dirichlet series which play a prominent role in understanding the distribution of cubic number fields, similar to how the famous Riemann zeta function controls the distribution of primes. As Shintani proved as early as 1972 [Shintani], the Shintani zeta functions satisfy a matrix functional equation (see Nakagawa [Nakagawa], eq. (0.1))

[ζ+(1s)ζ(1s)]=2133s2π4sΓ(s16)Γ(s)2Γ(s+16)[sin2πssinπs3sinπssin2πs][ζ^+(s)ζ^(s)]\begin{bmatrix}\zeta^{+}(1-s)\\ \zeta^{-}(1-s)\end{bmatrix}=2^{-1}3^{3s-2}\pi^{-4s}\Gamma\left(s-\frac{1}{6}\right)\Gamma(s)^{2}\Gamma\left(s+\frac{1}{6}\right)\begin{bmatrix}\sin 2\pi s&\sin\pi s\\ 3\sin\pi s&\sin 2\pi s\end{bmatrix}\begin{bmatrix}\hat{\zeta}^{+}(s)\\ \hat{\zeta}^{-}(s)\end{bmatrix} (2)

The condition that 33 divide bb and cc is equivalent to requiring that the cubic form ff is integer-matrix, that is, its corresponding symmetric trilinear form

b/3\textstyle{b/3\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}c/3\textstyle{c/3\ignorespaces\ignorespaces\ignorespaces\ignorespaces}a\textstyle{a\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}b/3\textstyle{b/3\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}c/3\textstyle{c/3\ignorespaces\ignorespaces\ignorespaces\ignorespaces}d\textstyle{d}b/3\textstyle{b/3\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}c/3\textstyle{c/3\ignorespaces\ignorespaces\ignorespaces\ignorespaces}

has integer entries. This condition arose in Shintani’s work by taking the dual lattice to 4\mathbb{Z}^{4} under the pairing

(a,b,c,d),(a,b,c,d)=ad13bc+13cbda,\left\langle(a,b,c,d),(a^{\prime},b^{\prime},c^{\prime},d^{\prime})\right\rangle=ad^{\prime}-\frac{1}{3}bc^{\prime}+\frac{1}{3}cb^{\prime}-da^{\prime}, (3)

which plays a central role in proving the functional equation. However, as we will find, the pairing (3) does not figure in the proof of our reflection theorems, which indeed often relate lattices that are not dual under it.

Using the functional equation, Shintani proved that the ζ±\zeta^{\pm} admit meromorphic continuations to the complex plane with simple poles at 11 and 5/65/6, inspiring him to conjecture that the number N±(X)N_{\pm}(X) of cubic fields of positive or negative discriminant up to XX has the shape

N±(X)=a±X+b±X5/6+o(X5/6)N_{\pm}(X)=a_{\pm}X+b_{\pm}X^{5/6}+o(X^{5/6})

for suitable constants a±a_{\pm} and b±b_{\pm}. This conjecture was proven by Bhargava, Shankar, and Tsimerman [BST-2ndOrd] and independently by Taniguchi and Thorne [TT_rc]. Neither proof needs the Ohno-Nakagawa reflection theorem (Theorem 1.1), which appears in the notation of Shintani zeta functions in the succinct form

ζ^+(s)=ζ(s)andζ^(s)=3ζ+(s).\hat{\zeta}^{+}(s)=\zeta^{-}(s)\quad\text{and}\quad\hat{\zeta}^{-}(s)=3\zeta^{+}(s). (4)
Remark 1.2.

In the earlier papers, the term “Ohno-Nakagawa identities” was used, referring to the pair (4). Our work confirms the intuition that, despite the different scalings, both identities are essentially one theorem.

1.2 Methods

Several proofs of O-N are now in print ([Nakagawa, Marinescu, OOnARemarkable, Gao]), all of which consist of two main steps:

  • A “global” step that uses global class field theory to understand cubic fields, equivalently GL2()\mathrm{GL}_{2}(\mathbb{Q})-orbits of cubic forms;

  • A “local” step to count the rings in each cubic field, equivalently the GL2()\mathrm{GL}_{2}(\mathbb{Z})-orbits in each GL2()\mathrm{GL}_{2}(\mathbb{Q})-orbit, and put the result in a usable form.

In this paper, the distinction between these steps will be formalized and clarified.

For the global step, we take inspiration from Tate’s celebrated thesis [Tate_thesis], which uses Fourier analysis on the adeles to give illuminating new proofs of the functional equations for the Riemann ζ\zeta-function and various LL-functions. Taniguchi and Thorne (see [TT_oexp]) used Fourier analysis on the space of binary cubic forms over 𝔽q\mathbb{F}_{q} to get the functional equation for the Shintani zeta function of forms satisfying local conditions at primes. Despite the similarities, their work is essentially independent from ours. We are also inspired by a remark due to Calegari in a paper of Cohen, Rubinstein-Salzedo, and Thorne ([CohON], Remark 1.6), pointing out that their reflection theorem counting dihedral fields of prime order can also be derived from a theorem of Greenberg and Wiles for the sizes of Selmer groups in Galois cohomology.

We present a notion of composed variety, a scheme 𝒱\mathcal{V} over the ring of integers 𝒪K\mathcal{O}_{K} of a number field admitting an action of an algebraic group 𝒢\mathcal{G} over 𝒪K\mathcal{O}_{K}. Our guiding example is the scheme 𝒱\mathcal{V} of binary cubic forms of discriminant DD with its action of 𝒱=SL2\mathcal{V}=\mathrm{SL}_{2}. The term “composed” refers to the presence of a composition law on the orbits, which relate naturally to a Galois cohomology group H1(K,M)H^{1}(K,M). Our (global) reflection theorems can be stated as saying that two composed varieties 𝒱(1)\mathcal{V}^{(1)}, 𝒱(2)\mathcal{V}^{(2)} have the same number of 𝒪K\mathcal{O}_{K}-points, with a suitable weighting. Introducing a new technique of Fourier analysis on the adelic cohomology group H1(𝔸K,M)=vH1(Kv,M)H^{1}(\mathbb{A}_{K},M)=\prod^{\prime}_{v}H^{1}(K_{v},M), based on Poitou-Tate duality, we present a generalized reflection engine (Theorems 8.12 and LABEL:thm:main_compose_multi) that reduces global reflection theorems to local reflection theorems, that is, statements involving only the 𝒪Kv\mathcal{O}_{K_{v}}-points of 𝒱(1)\mathcal{V}^{(1)} and 𝒱(2)\mathcal{V}^{(2)} for a single place vv of KK. A typical case is Theorem LABEL:thm:O-N_cubic_local.

These local reflection theorems are approachable by elementary methods but can be difficult to prove. We present two kinds of proofs. The first is a bijective argument involving Bhargava’s self-balanced ideals that is very clean but has only been discovered at the “tame primes” (𝔭3\mathfrak{p}\nmid 3 in the cubic case, 𝔭2\mathfrak{p}\nmid 2 in the quartic). The second is by explicitly computing the number of orders of given resolvent in a cubic or quartic algebra. We express it as a generating function in a number of variables depending on the splitting type of the resolvent. The generating function is rational, and local reflection can be written as an equality between two rational functions; but these functions are so complicated that the best approximation to a proof of the identity that we can find is a Monte Carlo proof, namely, substituting random values for the variables in some large finite field and verifying that the equality holds. The reader is invited to recheck this verification using the source code in Sage that will be made available with the final version of this paper.

1.3 Results

We are able to prove O-N for binary cubic forms over all number fields KK, verifying and extending the conjectures of Dioses [Dioses, Conjecture 1.1]. However, we go further and ask whether every SL2(𝒪K)\mathrm{SL}_{2}(\mathcal{O}_{K})-invariant lattice within the space V(K)V(K) of binary cubic forms admits an O-N-style reflection theorem. Over \mathbb{Z}, this question was answered affirmatively for each of the ten invariant lattices by Ohno and Taniguchi [10lat]. Over 𝒪K\mathcal{O}_{K}, such lattices were classified by Osborne [Osborne], and they differ from one another only at the primes dividing 22 and 33. The lattices at 33 yield an elegant reflection theorem (Theorem LABEL:thm:O-N_traced) in which the condition b,c𝔱b,c\in\mathfrak{t}, where 𝔱\mathfrak{t} is an ideal dividing 33 in 𝒪K\mathcal{O}_{K}, reflects to b,c3𝔱1b,c\in 3\mathfrak{t}^{-1}, the complementary divisor. At 22, the corresponding reflection theorems still exist, though they become difficult to write explicitly: see Theorem LABEL:thm:invarlat.

We also find a new reflection theorem (Theorem LABEL:thm:O-N_quad) counting binary quadratic forms, not by discriminant, but by a curious invariant: the product a(b24ac)a(b^{2}-4ac) of the discriminant and the leading coefficient. Over \mathbb{Z}, the reflection theorem (Theorem LABEL:thm:O-N_quad_Z) has the potential to be proved simply using quadratic reciprocity, eschewing the machinery of Galois cohomology, though it seems unlikely that the theorem would have ever been discovered without it.

Nakagawa has also conjectured [NakPairs] a reflection theorem for pairs of ternary quadratic forms, which parametrize quartic rings. The natural invariant to count by is the discriminant, but it is more natural from our perspective to subdivide further and ask for a reflection theorem for rings with fixed cubic resolvent, which holds in the known cases [NakPairs, Theorem 1]. Here our global framework applies without change, but the local enumeration of orders in a quartic field presents formidable combinatorial difficulties, especially in the wildly ramified (22-adic) setting, which have been attacked in another work of Nakagawa [NakOrders]. Our methods have the potential to finish this work, but because we count by resolvent rather than discriminant, our answers do not directly match his.

The process of proving local quartic O-N leads us down some fruitful routes that do not at first sight have any connection to reflection theorems or to the enumeration of quartic rings. These include new cases of the Igusa zeta functions of conics (Lemmas LABEL:lem:conic_1 and LABEL:lem:conic_pi) and a result on the average value of a quadratic character on a box in a local field (Theorem LABEL:thm:char_box). If quartic O-N holds true in all cases, it implies that the cubic resolvent ring (in the sense of Bhargava) of a maximal quartic order has a second natural characterization: it is the “conductor ring” for which the Galois-naturally attached extension K6/K3K_{6}/K_{3} is a ring class field (Theorem* LABEL:thm*:cond_ring).

1.4 Outline of the paper

In Section 2, we state and give examples of the main global reflection theorems of the paper over \mathbb{Z}, in a fashion that requires a minimum of prior knowledge, for the end of further diffusing interest in, and appreciation of, the beauty of number theory.

In Part II, we lay out preliminary matter, much of which is closely related to results that have appeared in the literature but under different guises. It includes a simple characterization (Proposition 4.21) of Galois H1H^{1} in terms of étale algebras whose Galois group is a semidirect product. It also includes a theorem (Theorem 7.1) on the structure of H1(K,M)H^{1}(K,M) in the case that KK is local and M𝒞pM\cong\mathcal{C}_{p} (with any Galois structure), which will be invaluable in what follows.

In Part III, we lay out the framework of composed varieties, on which we perform the novel technique of Fourier analysis of the local and global Tate pairings to get our main local-to-global reflection engine (Theorems 8.12 and LABEL:thm:main_compose_multi). The remainder of the paper will concern applications of this engine.

In Part LABEL:part:first, we prove two relatively simple reflection theorems: one for quadratic forms (Theorem LABEL:thm:O-N_quad), and a version of the Scholz reflection principle for class groups of quadratic orders (Theorem LABEL:thm:Scholz_for_locally_dual_orders).

In Part LABEL:part:cubic, we prove our extensions of Ohno-Nakagawa for cubic forms and rings.

The quartic case is dealt with in Parts LABEL:part:quartic and LABEL:part:quartic_count: the first part dealing with the bijective methods, and the second with the (long) work of explicitly counting orders in each quartic algebra. The case of partially ramified cubic resolvent (splitting type 1211^{2}1) is still in progress, so we restrict our attention to the four tamely splitting types in the present version.

We conclude the paper with some unanswered questions engendered by this research.

1.5 Acknowledgements

For fruitful discussions, I would like to thank (in no particular order): Manjul Bhargava, Xiaoheng Jerry Wang, Fabian Gundlach, Levent Alpöge, Melanie Matchett Wood, Kiran Kedlaya, Alina Bucur, Benedict Gross, Sameera Vemulapalli, Brandon Alberts, Peter Sarnak, and Jack Thorne.

2 Examples for the lay reader

Fortunately for the non-specialist reader, the statements (though not the proofs) of the main results in this thesis can be stated in a way requiring little more than high-school algebra. We here present these statements and some examples to illustrate them.

2.1 Reflection for quadratic equations

Definition 2.1.

Let f(x)=ax2+bx+cf(x)=ax^{2}+bx+c be a quadratic polynomial, where the coefficients aa, bb, cc are integers. The superdiscriminant of ff is the product

I=a(b24ac)I=a\cdot(b^{2}-4ac)

of the leading coefficient with the usual discriminant.

Lemma 2.2.

If we replace xx by x+tx+t in a quadratic polynomial ff, where tt is a fixed integer, then the superdiscriminant does not change.

Proof.

This can be verified by brute-force calculation, but the following method is more illuminating. The discriminant is classically related to the two roots of ff,

x1=b+b24ac2aandx2=bb24ac2a,x_{1}=\frac{-b+\sqrt{b^{2}-4ac}}{2a}\quad\text{and}\quad x_{2}=\frac{-b-\sqrt{b^{2}-4ac}}{2a},

through their difference:

x1x2\displaystyle x_{1}-x_{2} =2b24ac2a=b24aca\displaystyle=\frac{2\sqrt{b^{2}-4ac}}{2a}=\frac{\sqrt{b^{2}-4ac}}{a}
(x1x2)2\displaystyle(x_{1}-x_{2})^{2} =b24aca2\displaystyle=\frac{b^{2}-4ac}{a^{2}}
a3(x1x2)2\displaystyle a^{3}(x_{1}-x_{2})^{2} =a(b24ac)=I.\displaystyle=a\cdot(b^{2}-4ac)=I.

If we replace xx by x+tx+t, then aa does not change, and both roots x1,x2x_{1},x_{2} are decreased by tt, so their difference x1x2x_{1}-x_{2} is unchanged. Therefore II is unchanged. ∎

Definition 2.3.

Call two quadratics f1f_{1}, f2f_{2} equivalent if they are related by a translation f2(x)=f1(x+t)f_{2}(x)=f_{1}(x+t). If II is a nonzero integer, let q(I)q(I) be the number of quadratics of superdiscriminant II, up to equivalence. Let q2(I)q_{2}(I), q+(I)q^{+}(I), q2+(I)q_{2}^{+}(I) be the number of such quadratics that satisfy certain added conditions:

  • For q2(I)q_{2}(I), we require that the middle coefficient bb be even.

  • For q+(I)q^{+}(I), we require that the roots be real, that is, that b24ac>0b^{2}-4ac>0.

  • For q2+(I)q_{2}^{+}(I), we impose both of the last two conditions.

We are now ready to state a quadratic reflection theorem, the main result of this section.

Theorem 2.4 (“Quadratic O-N”).

For every nonzero integer nn,

q2+(4n)\displaystyle q_{2}^{+}(4n) =q(n)\displaystyle=q(n)
q2(4n)\displaystyle q_{2}(4n) =2q+(n).\displaystyle=2q^{+}(n).
Proof.

The proof is not easy. See Theorem LABEL:thm:O-N_quad_Z. ∎

It’s not hard to compute all quadratics of a fixed superdiscriminant II. The leading coefficient aa must be a divisor of II (possibly negative), and there are only finitely many of these. Then, by replacing xx by x+tx+t where tt is an integer nearest to b/(2a)-b/(2a), we can assume that bb lies in the window |a|<b|a|-|a|<b\leq|a|. We can try each of the integer values in this window, checking whether

c=ab2I4a2c=\frac{ab^{2}-I}{4a^{2}}

comes out to an integer.

Example 2.5.

There are five quadratics of superdiscriminant 1515:

f(x)f(x) qq q+q^{+} q2q_{2} q2+q_{2}^{+}
x2+x4{-x^{2}}+x-4 \checkmark
15x2+x15x^{2}+x \checkmark \checkmark
15x2x15x^{2}-x \checkmark \checkmark
15x2+11x+215x^{2}+11x+2 \checkmark \checkmark
15x211x+215x^{2}-11x+2 \checkmark \checkmark

You might think we left out x2x4-x^{2}-x-4, but it is equivalent to another quadratic on the list:

x2x4=(x+1)2+(x+1)4.-x^{2}-x-4=-(x+1)^{2}+(x+1)-4.

So we get the totals

q(15)=5andq+(15)=4.q(15)=5\quad\text{and}\quad q^{+}(15)=4.

There are 1818 quadratics of superdiscriminant 6060:

f(x)f(x) qq q+q^{+} q2q_{2} q2+q_{2}^{+}
x215{x^{2}-15} \checkmark \checkmark \checkmark \checkmark
x215{-x^{2}-15} \checkmark \checkmark
3x2+2x2{-3x^{2}+2x-2} \checkmark \checkmark
3x22x2{-3x^{2}-2x-2} \checkmark \checkmark
4x2+x1{-4x^{2}+x-1} \checkmark
4x2x1{-4x^{2}-x-1} \checkmark
15x2+2x{15x^{2}+2x} \checkmark \checkmark \checkmark \checkmark
15x22x{15x^{2}-2x} \checkmark \checkmark \checkmark \checkmark
15x2+8x+1{15x^{2}+8x+1} \checkmark \checkmark \checkmark \checkmark
15x28x+1{15x^{2}-8x+1} \checkmark \checkmark \checkmark \checkmark
60x2+x{60x^{2}+x} \checkmark \checkmark
60x2x{60x^{2}-x} \checkmark \checkmark
60x2+31x+4{60x^{2}+31x+4} \checkmark \checkmark
60x231x+4{60x^{2}-31x+4} \checkmark \checkmark
60x2+41x+7{60x^{2}+41x+7} \checkmark \checkmark
60x241x+7{60x^{2}-41x+7} \checkmark \checkmark
60x2+49x+10{60x^{2}+49x+10} \checkmark \checkmark
60x249x+10{60x^{2}-49x+10} \checkmark \checkmark

Counting carefully, we get

q(60)=18,q2(60)=8,q+(60)=13,q2+(60)=5.q(60)=18,\quad q_{2}(60)=8,\quad q^{+}(60)=13,\quad q_{2}^{+}(60)=5.

The equalities

q2+(60)=5=q(15)andq2(60)=8=24=2q+(15)q_{2}^{+}(60)=5=q(15)\quad\text{and}\quad q_{2}(60)=8=2\cdot 4=2q^{+}(15)

are instances of Theorem 2.4. From the same theorem, we derive, without computation, that

q2+(240)=q(60)=18andq2(240)=2q+(60)=26.q_{2}^{+}(240)=q(60)=18\quad\text{and}\quad q_{2}(240)=2q^{+}(60)=26.

This short investigation raises many questions. The superdiscriminant I=a(b24ac)I=a(b^{2}-4ac) does not seem to have been considered before. Is there an explicit formula for q(I)q(I)? Is there an elementary proof of Theorem 2.4? See Example LABEL:ex:QR for a connection to Gauss’s celebrated law of quadratic reciprocity.

2.2 Reflection for cubic equations

Definition 2.6.

For a cubic polynomial

f(x)=ax3+bx2+cx+d,f(x)=ax^{3}+bx^{2}+cx+d,

we define the discriminant to be

discf=a4(x1x2)2(x1x3)2(x2x3)2,\operatorname{disc}f=a^{4}(x_{1}-x_{2})^{2}(x_{1}-x_{3})^{2}(x_{2}-x_{3})^{2}, (5)

where x1,x2,x3x_{1},x_{2},x_{3} are the roots. Explicitly,

discf=b2c24ac34b3d27a2d2+18abcd.\operatorname{disc}f=b^{2}c^{2}-4ac^{3}-4b^{3}d-27a^{2}d^{2}+18abcd. (6)

There are many transformations of a cubic polynomial that don’t change the discriminant. One is changing xx to x+tx+t, where tt is a constant. Another is reversing the coefficients,

f(x)=ax3+bx2+cx+dx3f(1x)=dx3+cx2+bx+a.f(x)=ax^{3}+bx^{2}+cx+d\longmapsto x^{3}f\left(\frac{1}{x}\right)=dx^{3}+cx^{2}+bx+a.

Both of these are special cases of the following construction.

Definition 2.7.

Two cubic polynomials f1f_{1}, f2f_{2} with integer coefficients are equivalent if there is a matrix

[pqrs]\begin{bmatrix}p&q\\ r&s\end{bmatrix}

whose determinant psqrps-qr is ±1\pm 1 such that

f2(x)=(rx+s)3f1(px+qrx+s).f_{2}(x)=(rx+s)^{3}\cdot f_{1}\left(\frac{px+q}{rx+s}\right).

A matrix that makes ff equivalent to itself, that is,

f(x)=(rx+s)3f(px+qrx+s),f(x)=(rx+s)^{3}\cdot f\left(\frac{px+q}{rx+s}\right),

is called a symmetry of ff. The number of symmetries of ff is denoted by s(f)s(f).

Definition 2.8.

If DD is a nonzero integer, define h(D)h(D) to be the number of cubic polynomials

f(x)=ax3+bx2+cx+df(x)=ax^{3}+bx^{2}+cx+d

of discriminant DD, up to equivalence, each ff counted not once but 1/s(f)1/s(f) times, where s(f)s(f) is the number of symmetries. Define h3(D)h_{3}(D) to be the number of cubics of discriminant DD for which the middle two coefficients, bb and cc, are multiples of 33, up to equivalence, each ff counted 1/s(f)1/s(f) times as before.

We can now state the Ohno-Nakagawa reflection theorem that got this research project started:

Theorem 2.9 (Ohno-Nakagawa; Theorem 1.1).

For every nonzero integer DD,

h3(27D)={3h(D),D>0h(D),D<0.h_{3}(-27D)=\begin{cases}3h(D),&D>0\\ h(D),&D<0.\end{cases}
Proof.

Several proofs are in print (see the Introduction). In this paper, we prove this theorem as a special case of Theorem LABEL:thm:O-N_traced. ∎

Example 2.10.

Take D=1D=1. There is just one cubic with integer coefficients and discriminant 11, namely

f(x)=x(x+1)=x2+x.f(x)=x(x+1)=x^{2}+x.

The reader may balk at considering a quadratic polynomial as a “cubic” with leading coefficient 0, but the polynomial can be replaced by any number of equivalent forms, for instance

(x1)3f(xx1)=x(x1)(2x1).(x-1)^{3}\cdot f\left(\frac{x}{x-1}\right)=x(x-1)(2x-1).

We will suppress this detail in subsequent examples. (A program for computing all cubics of a given discriminant is found in the attached file cubics.sage, based on an algorithm of Cremona [Crem_Redn, Crem_Redn_2]). The cubic ff has six symmetries, which is related to the fact that three linear factors can be permuted in 3!=63!=6 ways. In terms of f(x)=x(x+1)f(x)=x(x+1), the symmetries are

[1001],[1101],[0110],[1110],[1011],[0111].\begin{bmatrix}1&0\\ 0&1\end{bmatrix},\begin{bmatrix}-1&-1\\ 0&1\end{bmatrix},\begin{bmatrix}0&1\\ 1&0\end{bmatrix},\begin{bmatrix}-1&-1\\ 1&0\end{bmatrix},\begin{bmatrix}1&0\\ -1&-1\end{bmatrix},\begin{bmatrix}0&1\\ -1&-1\end{bmatrix}.

So h(1)=1/6h(1)=1/6.

Correspondingly, we look at cubics of discriminant 27-27. There are two:

f(x)=x2+x+7andf(x)=x3+1.f(x)=x^{2}+x+7\quad\text{and}\quad f(x)=x^{3}+1.

Each admits two symmetries: the first has

[1001],[1101],\begin{bmatrix}1&0\\ 0&1\end{bmatrix},\begin{bmatrix}-1&-1\\ 0&1\end{bmatrix},

and the second has

[1001],[0110].\begin{bmatrix}1&0\\ 0&1\end{bmatrix},\begin{bmatrix}0&1\\ 1&0\end{bmatrix}.

So h(27)=1/2+1/2=1h(-27)=1/2+1/2=1 and h3(27)=1/2h_{3}(-27)=1/2. In particular,

h3(27)=3h3(1),h_{3}(-27)=3h_{3}(1),

in conformity with Theorem 2.9.

2.3 Reflection for 2×n×n2\times n\times n boxes

Bhargava [B3] studied 2×3×32\times 3\times 3 boxes as a visual representation for quartic rings, as cubic polynomials do for cubic rings. We think that reflection holds not only for 2×3×32\times 3\times 3 boxes but for 2×5×52\times 5\times 5, 2×7×72\times 7\times 7, and so on. We nearly prove the 2×3×32\times 3\times 3 case in this paper. We are quite far from proving it for the larger boxes.

Definition 2.11.

A box is a pair (A,B)(A,B) of n×nn\times n integer symmetric matrices. The resolvent of a box is the polynomial

f(x)=det(AxB).f(x)=\det(Ax-B).

It is a polynomial in xx, of degree at most nn. If AA is the identity matrix, the resolvent devolves into the standard characteristic polynomial.

Definition 2.12.

Two boxes (A1,B1)(A_{1},B_{1}) and (A2,B2)(A_{2},B_{2}) are equivalent if there is an integer n×nn\times n matrix XX, whose inverse X1X^{-1} also has integer entries, such that

A2=XA1XandB2=XB1X.A_{2}=XA_{1}X^{\top}\quad\text{and}\quad B_{2}=XB_{1}X^{\top}.

If (A2,B2)=(A1,B1)=(A,B)(A_{2},B_{2})=(A_{1},B_{1})=(A,B) are the same pair, then XX is called a symmetry of (A,B)(A,B). The number of symmetries of (A,B)(A,B) will be denoted by s(A,B)s(A,B).

Conjecture 2.13 (“O-N for 2×n×n2\times n\times n boxes”).

Let nn be a positive odd integer. Let ff be a polynomial of degree nn with no multiple roots and only one real root. Denote by h(f)h(f) the number of 2×n×n2\times n\times n boxes with resolvent ff, up to equivalence, each box weighted by the reciprocal of its number of symmetries. Denote by h2(f)h_{2}(f) the number of such boxes with even numbers along the main diagonals of AA and BB, weighted the same way. Then

h2(2n1f)=2n12h(f).h_{2}(2^{n-1}f)=2^{\frac{n-1}{2}}\cdot h(f). (7)
Remark 2.14.

The condition that ff have no multiple roots (even complex ones) is needed to ensure that there are only finitely many boxes with ff as a resolvent. The condition that ff have no more than one real root can be eliminated, but then we must impose conditions on the real behavior of the boxes that are difficult to state succinctly.

Example 2.15.

Take as resolvent f(x)=x3x1f(x)=x^{3}-x-1, the simplest irreducible cubic. It has one real root ξ1.3247\xi\approx 1.3247 and discriminant 23-23. There are two boxes with resolvent ff, up to equivalence:

[(001010101),(010101011)],[(010101011),(101011111)].\left[\left(\begin{array}[]{rrr}0&0&-1\\ 0&-1&0\\ -1&0&-1\end{array}\right),\left(\begin{array}[]{rrr}0&-1&0\\ -1&0&-1\\ 0&-1&-1\end{array}\right)\right],\left[\left(\begin{array}[]{rrr}0&-1&0\\ -1&0&-1\\ 0&-1&-1\end{array}\right),\left(\begin{array}[]{rrr}-1&0&-1\\ 0&-1&-1\\ -1&-1&-1\end{array}\right)\right].

(These were computed from the balanced pairs (𝒪R,1)(\mathcal{O}_{R},1) and (𝒪R,ξ)(\mathcal{O}_{R},\xi) in the number field R=[ξ]/(ξ3ξ1)R=\mathbb{Z}[\xi]/(\xi^{3}-\xi-1) corresponding to ff.) Neither has any symmetries besides the two trivial ones, the identity matrix and its negative, so

h(f)=12+12=1.h(f)=\frac{1}{2}+\frac{1}{2}=1.

There are many boxes with resolvent 2f2f, but just one with even numbers all along the main diagonals of AA and BB, namely

[(001020102),(010100002)].\left[\left(\begin{array}[]{rrr}0&0&1\\ 0&-2&0\\ 1&0&2\end{array}\right),\left(\begin{array}[]{rrr}0&1&0\\ 1&0&0\\ 0&0&-2\end{array}\right)\right].

(This was computed from the unique quartic ring 𝒪=×𝒪R\mathcal{O}=\mathbb{Z}\times\mathcal{O}_{R} with resolvent 𝒪R\mathcal{O}_{R}.) It too has only the trivial symmetries, to h2(2f)=1/2h_{2}(2f)=1/2, in accord with Conjecture 2.13.

2.4 Reflection for quartic equations

There are also reflection theorems that appear when counting quartic polynomials.

Definition 2.16.

If

f(x)=ax4+bx3+cx2+dx+ef(x)=ax^{4}+bx^{3}+cx^{2}+dx+e

is a quartic polynomial with integer coefficients, its resolvent is

g(y)=y3cy2+(bd4ae)y+4aceb2ead2;g(y)=y^{3}-cy^{2}+(bd-4ae)y+4ace-b^{2}e-ad^{2}; (8)

equivalently, if

f(x)=a(xx1)(xx2)(xx3)(xx4),f(x)=a(x-x_{1})(x-x_{2})(x-x_{3})(x-x_{4}),

then

g(y)=(ya(x1x2+x3x4))(ya(x1x3+x2x4))(ya(x1x4+x2x3)).g(y)=\big{(}y-a(x_{1}x_{2}+x_{3}x_{4})\big{)}\big{(}y-a(x_{1}x_{3}+x_{2}x_{4})\big{)}\big{(}y-a(x_{1}x_{4}+x_{2}x_{3})\big{)}.
Remark 2.17.

Cubic resolvents of this type have been used since the 16th century as a step in solving quartic equations. For instance, it is well known that if f(x)f(x) factors as the product of two quadratics with integer coefficients, then g(y)g(y) has a rational root (the converse is not true).

Analogously to Definition 2.7, we put:

Definition 2.18.

Two quartic polynomials f1f_{1}, f2f_{2} with integer coefficients are equivalent if there is a matrix

[pqrs]\begin{bmatrix}p&q\\ r&s\end{bmatrix}

whose determinant psqrps-qr is ±1\pm 1 such that

f2(x)=(rx+s)4f1(px+qrx+s).f_{2}(x)=(rx+s)^{4}\cdot f_{1}\left(\frac{px+q}{rx+s}\right).

A matrix that makes ff equivalent to itself, that is,

f(x)=(rx+s)4f(px+qrx+s),f(x)=(rx+s)^{4}\cdot f\left(\frac{px+q}{rx+s}\right),

is called a symmetry of ff. The number of symmetries of ff is denoted by s(f)s(f).

We have:

Lemma 2.19.
  1. ((a))

    If two quartics f1f_{1}, f2f_{2} are equivalent, then their resolvents g1g_{1}, g2g_{2} are related by a translation

    g2(x)=g1(x+t)g_{2}(x)=g_{1}(x+t)

    for some integer tt.

  2. ((b))

    A quartic and its resolvent have the same discriminant

    discf=discg=b2c2d24ac3d24b3d3+18abcd327a2d44b2c3e+16ac4e+18b3cde80abc2de6ab2d2e+144a2cd2e27b4e2+144ab2ce2128a2c2e2192a2bde2+256a3e3.?\operatorname{disc}f=\operatorname{disc}g=\parbox[t]{303.53267pt}{ $b^{2}c^{2}d^{2}-4ac^{3}d^{2}-4b^{3}d^{3}+18abcd^{3}-27a^{2}d^{4}-4b^{2}c^{3}e+16ac^{4}e+18b^{3}cde-80abc^{2}de-6ab^{2}d^{2}e+144a^{2}cd^{2}e-27b^{4}e^{2}+144ab^{2}ce^{2}-128a^{2}c^{2}e^{2}-192a^{2}bde^{2}+256a^{3}e^{3}.$ }
Proof.

Exercise. ∎

As before, our reflection theorem will relate general quartics to quartics satisfying certain divisibility relations. Here the relations are quite peculiar:

Definition 2.20.

A quartic polynomial

f(x)=ax4+bx3+cx2+dx+ef(x)=ax^{4}+bx^{3}+cx^{2}+dx+e

is called supereven if bb, cc, and ee are multiples of 44 and dd is a multiple of 88.

Not every quartic equivalent to a super-even quartic is itself supereven. (For instance, f1=x4+4f_{1}=x^{4}+4 and f2=4x4+1f_{2}=4x^{4}+1 are equivalent under the flip [0110]\big{[}\begin{smallmatrix}0&1\\ 1&0\end{smallmatrix}\big{]} , but f2f_{2} is not supereven.) We therefore make the following definition.

Definition 2.21.

Two quartic polynomials f1f_{1}, f2f_{2} with integer coefficients are evenly equivalent if there is a matrix

[pqrs]\begin{bmatrix}p&q\\ r&s\end{bmatrix}

whose determinant psqrps-qr is ±1\pm 1, and r is even, such that

f2(x)=(rx+s)4f1(px+qrx+s).f_{2}(x)=(rx+s)^{4}\cdot f_{1}\left(\frac{px+q}{rx+s}\right).

Such a matrix that makes ff equivalent to itself, that is,

f(x)=(rx+s)4f(px+qrx+s),f(x)=(rx+s)^{4}\cdot f\left(\frac{px+q}{rx+s}\right),

is called an even symmetry of ff. The number of even symmetries of ff is denoted by s2(f)s_{2}(f).

Theorem 2.22 (“Quartic O-N”).

Let gg be an integer cubic with leading coefficient 11, no multiple roots, and odd discriminant. Denote by h(g)h(g) the number of quartics whose resolvent is g(y+t)g(y+t) for some tt, up to equivalence and weighted by the reciprocal of the number of symmetries. Denote by h2(g)h_{2}(g) the number of supereven quartics whose resolvent is g(y+t)g(y+t) for some tt, up to even equivalence and weighted by the reciprocal of the number of even symmetries. Define g2g_{2} by

g2(y)=64g(y4)g_{2}(y)=64g\left(\frac{y}{4}\right)

Then:

  • If gg has one real root, then

    4h(g)=h2(g2).4h(g)=h_{2}(g_{2}).
  • If gg has three real roots, then we subdivide

    h(g)=h+(g)+h(g)+h±(g)h(g)=h^{+}(g)+h^{-}(g)+h^{\pm}(g)

    where the respective terms count only quartic functions that are always positive, always negative, and have four real roots. We subdivide

    h2(g)=h+2(g)+h2(g)+h±2(g).h_{2}(g)=h^{+}_{2}(g)+h^{-}_{2}(g)+h^{\pm}_{2}(g).

    Then:

    2h(g)\displaystyle 2h(g) =h2±(g2)\displaystyle=h_{2}^{\pm}(g_{2})
    4(h+(g)+h±(g))\displaystyle 4\big{(}h^{+}(g)+h^{\pm}(g)\big{)} =h2+(g2)+h2±(g2)\displaystyle=h_{2}^{+}(g_{2})+h_{2}^{\pm}(g_{2})
    4(h(g)+h±(g))\displaystyle 4\big{(}h^{-}(g)+h^{\pm}(g)\big{)} =h2(g2)+h2±(g2)\displaystyle=h_{2}^{-}(g_{2})+h_{2}^{\pm}(g_{2})

    Also, denote by k(g)k(g) the number of integral 3×33\times 3 symmetric matrices of characteristic polynomial gg. Then

    k(g)=24(h±(g)h+(g)h(g)).k(g)=24\big{(}h^{\pm}(g)-h^{+}(g)-h^{-}(g)\big{)}.
Proof.

See Theorem LABEL:thm:BQ. ∎

Remark 2.23.

We think that the hypothesis of odd discriminant is removable, but we have not yet finished the proof.

Example 2.24.

Let g(y)=y3y1g(y)=y^{3}-y-1. By techniques presented in Section LABEL:sec:bq, it is possible to transform the boxes found in example 2.15 into binary quartic forms. We find that there is only one quartic with resolvent gg, namely

f(x)=x3x1f(x)=x^{3}-x-1

(which, as before, can be transformed by an equivalence to one with nonzero leading coefficient); and four supereven binary quartics with resolvent g2(y)=y316y64g_{2}(y)=y^{3}-16y-64, namely

f(x)\displaystyle f(x) =4x3+12x2+8x4=4((x+1)3(x+1)1)\displaystyle=4x^{3}+12x^{2}+8x-4=4\big{(}(x+1)^{3}-(x+1)-1\big{)}
f(x)\displaystyle f(x) =x4+4x3+12x2+8x\displaystyle=-x^{4}+4x^{3}+12x^{2}+8x
f(x)\displaystyle f(x) =x4+8x4\displaystyle=-x^{4}+8x-4
f(x)\displaystyle f(x) =x4+4x34.\displaystyle=-x^{4}+4x^{3}-4.

All these have one pair of complex roots (as must occur for a resolvent with negative discriminant) and only the trivial symmetries ±[1001]\pm\big{[}\begin{smallmatrix}1&0\\ 0&1\end{smallmatrix}\big{]}, so

h(g)=12andh2(g2)=2=412,h(g)=\frac{1}{2}\quad\text{and}\quad h_{2}(g_{2})=2=4\cdot\frac{1}{2},

in accord with the first part of the theorem.

Example 2.25.

Consider g(y)=y32y23y+6=(y2)(y+3)(y3)g(y)=y^{3}-2y^{2}-3y+6=(y-2)(y+\sqrt{3})(y-\sqrt{3}), a cubic with three real roots. The quartics with resolvent gg are

f(x)=x(2x1)(3x21),f(x)=-x(2x-1)(3x^{2}-1),

which has four real roots, and

f(x)=(x2+x+1)(x2+1),f(x)=(x^{2}+x+1)(x^{2}+1),

which has no real roots and is positive for all real xx. Each has only the trivial symmetries, so

h±(g)=12,h+(g)=12,h(g)=0.h^{\pm}(g)=\frac{1}{2},\quad h^{+}(g)=\frac{1}{2},\quad h^{-}(g)=0.

(Note the discrepancy between h+h^{+} and hh^{-}.) Correspondingly, there are eight supereven binary quartics with resolvent g2(y)=(y8)(y+43)(y43)g_{2}(y)=(y-8)(y+4\sqrt{3})(y-4\sqrt{3}):

f(x)\displaystyle f(x) =2x48x34x2+8x=2x(x+2)(x2+2x2)\displaystyle=-2x^{4}-8x^{3}-4x^{2}+8x=-2x(x+2)(x^{2}+2x-2)
f(x)\displaystyle f(x) =4x34x216x8=4(x+1)(x22x2)\displaystyle=4x^{3}-4x^{2}-16x-8=4(x+1)(x^{2}-2x-2)
f(x)\displaystyle f(x) =8x416x3+20x212x+4=4(x2x+1)(2x22x+1)\displaystyle=8x^{4}-16x^{3}+20x^{2}-12x+4=4(x^{2}-x+1)(2x^{2}-2x+1)
f(x)\displaystyle f(x) =x46x3+20x232x+32=(x24x+8)(x22x+4)\displaystyle=x^{4}-6x^{3}+20x^{2}-32x+32=(x^{2}-4x+8)(x^{2}-2x+4)
f(x)\displaystyle f(x) =x4+8x212=(x22)(x26)\displaystyle=-x^{4}+8x^{2}-12=-(x^{2}-2)(x^{2}-6)
f(x)\displaystyle f(x) =3x4+8x24=(x22)(3x22)\displaystyle=-3x^{4}+8x^{2}-4=-(x^{2}-2)(3x^{2}-2)
f(x)\displaystyle f(x) =x4+8x2+12=(x2+2)(x2+6)\displaystyle=x^{4}+8x^{2}+12=(x^{2}+2)(x^{2}+6)
f(x)\displaystyle f(x) =3x4+8x2+4=(x2+2)(3x2+2).\displaystyle=3x^{4}+8x^{2}+4=(x^{2}+2)(3x^{2}+2).

Thus

h2±(g2)=2,h2+(g2)=2,h2(g2)=0.h_{2}^{\pm}(g_{2})=2,\quad h_{2}^{+}(g_{2})=2,\quad h_{2}^{-}(g_{2})=0.

This is in accord with the theorem, from which we also learn that

k(g)=48(h±(g)h+(g)h(g))=0,k(g)=48\big{(}h^{\pm}(g)-h^{+}(g)-h^{-}(g)\big{)}=0,

so gg is not the characteristic polynomial of any integer 3×33\times 3 symmetric matrix, despite having three real roots (which is a necessary, but not a sufficient, condition).

Example 2.26.

Let g(y)=y3yg(y)=y^{3}-y. Knowing that f(x)=x3xf(x)=x^{3}-x is the only quartic with cubic resolvent ff, and it has four symmetries, the powers of [0110]\big{[}\begin{smallmatrix}0&-1\\ 1&0\end{smallmatrix}\big{]}, we get

k(g)=24(h±(g)h+(g)h(g))=24(1400)=6.k(g)=24\left(h^{\pm}(g)-h^{+}(g)-h^{-}(g)\right)=24\left(\frac{1}{4}-0-0\right)=6.

So there are six symmetric matrices with characteristic polynomial y3yy^{3}-y. Indeed, they are the diagonal matrices with 11, 0, and 1-1 along the diagonal in any of the 3!=63!=6 possible orders.

3 Notation

The following conventions will be observed in the remainder of the paper.

We denote by \mathbb{N} and +\mathbb{N}^{+}, respectively, the sets of nonnegative and of positive integers.

If PP is a statement, then

𝟏P={1P is true0P is false.\mathbf{1}_{P}=\begin{cases}1&\text{$P$ is true}\\ 0&\text{$P$ is false}.\end{cases}

If SS is a set, then 𝟏S\mathbf{1}_{S} denotes the characteristic function 𝟏S(x)=𝟏xS\mathbf{1}_{S}(x)=\mathbf{1}_{x\in S}.

An algebra will always be commutative and of finite rank over a field, while a ring or order will be a finite-dimensional, torsion-free ring over a Dedekind domain, containing 11. An order need not be a domain.

If a,bLa,b\in L are elements of a local or global field, a separable closure thereof, or a finite product of the preceding, we write a|ba|b to mean that b=cab=ca for some cc in the appropriate ring of integers 𝒪L\mathcal{O}_{L}. If a|ba|b and b|ab|a, we say that aa and bb are associates and write aba\sim b. Note that aa and bb may be zero-divisors.

If SS is a finite set, we let Sym(S)\operatorname{Sym}(S) denote the set of permutations of SS; thus Sn=Sym({1,,n})S_{n}=\operatorname{Sym}(\{1,\ldots,n\}). If |S|=|T|\lvert S\rvert=\lvert T\rvert, and if gSym(S)g\in\operatorname{Sym}(S), hSym(T)h\in\operatorname{Sym}(T) are elements, we say that gg and hh are conjugate if there is a bijection between SS and TT under which they correspond. Likewise when we say that two subgroups GSym(S)G\subseteq\operatorname{Sym}(S), HSym(T)H\subseteq\operatorname{Sym}(T) are conjugate.

We will use the semicolon to separate the coordinates of an element of a product of rings. For instance, in ×\mathbb{Z}\times\mathbb{Z}, the nontrivial idempotents are (1;0)(1;0) and (0;1)(0;1).

If nn is a positive integer, then ζn\zeta_{n} denotes a primitive nnth root of unity in ¯\bar{\mathbb{Q}}, while ζ¯n\bar{\zeta}_{n} denotes the nnth root of unity

ζ¯n=(1;ζn;ζn2;;ζnn1)¯n.\bar{\zeta}_{n}=\left(1;\zeta_{n};\zeta_{n}^{2};\ldots;\zeta_{n}^{n-1}\right)\in\bar{\mathbb{Q}}^{n}.

Throughout the proofs of the local reflection theorems, we will fix a local field KK, its valuation v=vKv=v_{K}, its residue field kKk_{K} of order qq, and a uniformizer π=πK\pi=\pi_{K}. The letter ee will denote the absolute ramification index (e=vK(2)e=v_{K}(2) in the quadratic and quartic cases, vK(3)v_{K}(3) in the cubic). We let 𝔪K\mathfrak{m}_{K} denote the maximal ideal, and likewise 𝔪K¯\mathfrak{m}_{\bar{K}} be the maximal ideal of the ring 𝒪K¯\mathcal{O}_{\bar{K}} of algebraic integers over KK; note that 𝔪K¯\mathfrak{m}_{\bar{K}} is not finitely generated. We also allow v=vKv=v_{K} to be applied to elements of K¯\bar{K}, the valuation being scaled so that its restriction to KK has value group \mathbb{Z}. We use the absolute value bars ||\lvert\bullet\rvert for the corresponding metric, whose normalization will be left undetermined.

If KK is a local field, an mm-pixel is a subset of an affine or projective space over 𝒪K\mathcal{O}_{K} defined by requiring the coordinates to lie in specified congruence classes modulo πn\pi^{n}. For instance, in 2(𝒪K)\mathbb{P}^{2}(\mathcal{O}_{K}), a 0-pixel is the whole space, which is subdivided into (q2+q+1)q2n2(q^{2}+q+1)q^{2n-2}-many nn-pixels for each n1n\geq 1.

If R/KR/K is a finite-dimensional, locally free algebra over a ring, we denote by RN=1R^{N=1} the subgroup of units of norm 11. The group operation is implicitly multiplication, so RN=1[n]R^{N=1}[n], for instance, denotes the nnth roots of unity of norm 11.

Part II Galois cohomology

4 Étale algebras and their Galois groups

4.1 Étale algebras

If KK is a field, an étale algebra over KK is a finite-dimensional separable commutative algebra over KK, or equivalently, a finite product of finite separable extension fields of KK. A treatment of étale algebras is found in Milne ([MilneFields], chapter 8): here we summarize this theory and prove a few auxiliary results that will be of use.

An étale algebra LL of rank nn admits exactly nn maps ι1,,ιn\iota_{1},\ldots,\iota_{n} (of KK-algebras) to a fixed separable closure K¯\bar{K} of KK. We call these the coordinates of LL; the set of them will be called Coord(L/K)\operatorname{Coord}(L/K) or simply Coord(L)\operatorname{Coord}(L). Together, the coordinates define an embedding of LL into K¯n\bar{K}^{n}, which we call the Minkowski embedding because it subsumes as a special case the embedding of a degree-nn number field into n\mathbb{C}^{n}, which plays a major role in algebraic number theory, as in Delone-Faddeev [DF].

For any element γ\gamma of the absolute Galois group GKG_{K}, the composition γιi\gamma\circ\iota_{i} with any coordinate is also a coordinate ιj\iota_{j}, so we get a homomorphism ϕ=ϕL:GKSym(Coord(L))\phi=\phi_{L}:G_{K}\mathop{\rightarrow}\limits\operatorname{Sym}(\operatorname{Coord}(L))) such that

γ(ι(x))=(ϕγι)(x)\gamma(\iota(x))=(\phi_{\gamma}\iota)(x)

for all xL,ιCoord(L)x\in L,\iota\in\operatorname{Coord}(L). This gives a functor from étale KK-algebras to GKG_{K}-sets (sets with a GKG_{K}-action), which is denoted \mathcal{F} in Milne’s terminology. A functor going the other way, which Milne calls 𝒜\mathcal{A}, takes ϕ:GKSn\phi:G_{K}\mathop{\rightarrow}\limits S_{n} to

L={(x1,,xn)K¯nγ(xi)=xϕγ(i)γGK,i}L=\{(x_{1},\ldots,x_{n})\in\bar{K}^{n}\mid\gamma(x_{i})=x_{\phi_{\gamma}(i)}\,\forall\gamma\in G_{K},\forall i\} (9)
Proposition 4.1 ([MilneFields], Theorem 7.29).

The functors \mathcal{F} and 𝒜\mathcal{A} establish a bijection between

  • étale extensions L/KL/K of degree nn, up to isomorphism, and

  • GKG_{K}-sets of size nn up to isomorphism; that is to say, homomorphisms ϕ:GKSn\phi:G_{K}\mathop{\rightarrow}\limits S_{n}, up to conjugation in SnS_{n}.

Moreover, the bijection respects base change, in the following way:

Proposition 4.2.

Let K1/KK_{1}/K be a field extension, not necessarily algebraic, and let L/KL/K be an étale extension of degree nn. Then L1=LKK1L_{1}=L\otimes_{K}K_{1} is étale over K1K_{1}, and the associated Galois representations ϕL/K\phi_{L/K}, ϕL1/K1\phi_{L_{1}/K_{1}} are related by the commutative diagram

GK1\textstyle{G_{K_{1}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}|K¯\scriptstyle{\bullet|_{\bar{K}}}ϕL1/K1\scriptstyle{\phi_{L_{1}/K_{1}}}GK\textstyle{G_{K}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ϕL/K\scriptstyle{\phi_{L/K}}Sym(CoordK1(L1))\textstyle{\operatorname{Sym}(\operatorname{Coord}_{K_{1}}(L_{1}))\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\sim}Sym(CoordK(L))\textstyle{\operatorname{Sym}(\operatorname{Coord}_{K}(L))} (10)
Proof.

That L1/K1L_{1}/K_{1} is étale is standard (see Milne [MilneFields], Prop. 8.10). For the second claim, consider the natural restriction map r:CoordK1(L1)CoordK(L)r:\operatorname{Coord}_{K_{1}}(L_{1})\mathop{\rightarrow}\limits\operatorname{Coord}_{K}(L). It is injective, since a K1K_{1} linear map out of L1L_{1} is determined by its values on LL; and since both sets have the same size, rr is surjective and is hence an isomorphism of GK1G_{K_{1}}-sets (the GK1G_{K_{1}}-structure on CoordK(L)\operatorname{Coord}_{K}(L) arising by restriction from the GKG_{K}-structure). ∎

We will use this proposition most frequently in the case that KK is a global field and K1=KvK_{1}=K_{v} one of its completions. The resulting L1L_{1} is then the product Lvw|vLwL_{v}\cong\prod_{w|v}L_{w} of the completions of LL at the places dividing vv. Note the departure from the classical habit of studying the completion LwL_{w} at each place individually. The preservation of degrees, [L1:K1]=[L:K][L_{1}:K_{1}]=[L:K] will be important for our applications.

4.2 The Galois group of an étale algebra

Define the Galois group G(L/K)G(L/K) of an étale algebra to be the image of its associated Galois representation ϕ:GKSym(Coord(L))\phi:G_{K}\mathop{\rightarrow}\limits\operatorname{Sym}(\operatorname{Coord}(L)). It transitively permutes the coordinates corresponding to each field factor. For example, if LL is a quartic field, then G(L/K)G(L/K) is one of the five (up to conjugacy) transitive subgroups of 𝒮4\mathcal{S}_{4}, which (to use the traditional names) are 𝒮4\mathcal{S}_{4}, 𝒜4\mathcal{A}_{4}, 𝒟4\mathcal{D}_{4}, 𝒱4\mathcal{V}_{4}, and 𝒞4\mathcal{C}_{4}. Galois groups in this sense are used in the tables of cubic and quartic fields in Delone-Faddeev [DF] and the Number Field Database [NFDB]. Note that the Galois group G(L/K)G(L/K) is defined whether or not LL is a Galois extension. If it is, then the Galois group is simply transitive and coincides with the Galois group in the sense of Galois theory.

Important for us will be two notions pertaining to the Galois group.

Definition 4.3.

Let G𝒮nG\subseteq\mathcal{S}_{n} be a subgroup. A GG-extension of KK is a degree-nn étale algebra LL with a choice of subgroup GSym(Coord(L))G^{\prime}\subseteq\operatorname{Sym}(\operatorname{Coord}(L)) that is conjugate to GG and contains G(L/K)G(L/K), plus a conjugacy class of isomorphisms GGG^{\prime}\cong G: the conjugacy being in GG, not in 𝒮n\mathcal{S}_{n}. The added data is called a GG-structure on LL.

Proposition 4.4.

GG-extensions L/KL/K up to isomorphism are in bijection with homomorphisms ϕ:GKG\phi:G_{K}\mathop{\rightarrow}\limits G, up to conjugation in GG.

Proof.

Immediate from Proposition 4.1. ∎

Example 4.5.

L=(ζ5)L=\mathbb{Q}(\zeta_{5}) is a 𝒞4\mathcal{C}_{4}-extension (taking 𝒞4=(1234)𝒮4\mathcal{C}_{4}=\left\langle(1234)\right\rangle\subseteq\mathcal{S}_{4}), indeed its Galois group is isomorphic to 𝒞4\mathcal{C}_{4}; and LL admits two distinct 𝒞4\mathcal{C}_{4}-structures, as there are two ways to identify 𝒞4\mathcal{C}_{4} with its image in 𝒮4\mathcal{S}_{4}, which are conjugate in 𝒮4\mathcal{S}_{4} but not in 𝒞4\mathcal{C}_{4}. Likewise, L=×××L=\mathbb{Q}\times\mathbb{Q}\times\mathbb{Q}\times\mathbb{Q} admits six 𝒞4\mathcal{C}_{4}-structures, one for each embedding of 𝒞4\mathcal{C}_{4} into 𝒮4\mathcal{S}_{4}, as its Galois group is trivial.

4.3 Resolvents

This will be an important notion.

Definition 4.6.

Let G𝒮nG\subseteq\mathcal{S}_{n}, H𝒮mH\subseteq\mathcal{S}_{m} be subgroups and ρ:GH\rho:G\mathop{\rightarrow}\limits H be a homomorphism. Then for every GG-extension L/KL/K, the corresponding ϕL:GKG\phi_{L}:G_{K}\mathop{\rightarrow}\limits G may be composed with ρ\rho to yield a map ϕR:GKH\phi_{R}:G_{K}\mathop{\rightarrow}\limits H, which defines an étale extension R/KR/K of degree mm. This RR is called the resolvent of LL under the map ρ\rho.

Example 4.7.

Since there is a surjective map ρ4,3:𝒮4𝒮3\rho_{4,3}:\mathcal{S}_{4}\mathop{\rightarrow}\limits\mathcal{S}_{3}, every quartic étale algebra L/KL/K has a cubic resolvent RR. This resolvent appears in Bhargava [B3], but it is much older than that. It is generated by a formal root of the resolvent cubic that appears when a general quartic equation is to be solved by radicals.

Example 4.8.

Likewise, the sign map can be viewed as a homomorphism sgn:𝒮n𝒮2\operatorname{sgn}:\mathcal{S}_{n}\mathop{\rightarrow}\limits\mathcal{S}_{2}, attaching to every étale algebra LL a quadratic resolvent TT. If L=K[θ]/f(θ)L=K[\theta]/f(\theta) is generated by a polynomial ff, and if charK2\operatorname{char}K\neq 2, then it is not hard to see that T=K[discf]T=K[\sqrt{\operatorname{disc}f}] where discf\operatorname{disc}f is the polynomial discriminant. Note that TT still exists even if charK=2\operatorname{char}K=2. We have that TK×KT\cong K\times K is split if and only if the Galois group G(L/K)G(L/K) is contained in the alternating group 𝒜n\mathcal{A}_{n}.

Example 4.9.

The dihedral group 𝒟4\mathcal{D}_{4} has an outer automorphism, because rotating a square in the plane by 4545^{\circ} does not preserve the square but does preserve every symmetry of the square. This map ρ:𝒟4𝒟4\rho:\mathcal{D}_{4}\mathop{\rightarrow}\limits\mathcal{D}_{4} associates to each 𝒟4\mathcal{D}_{4}-algebra LL a new 𝒟4\mathcal{D}_{4}-algebra LL^{\prime}, not in general isomorphic. This is the classical phenomenon of the mirror field. For instance, if L=[1+2]L=\mathbb{Q}[\sqrt{1+\sqrt{2}}], then

L=[1+2+12]=[2+21].L^{\prime}=\mathbb{Q}\left[\sqrt{1+\sqrt{2}}+\sqrt{1-\sqrt{2}}\right]=\mathbb{Q}\left[\sqrt{2+2\sqrt{-1}}\right].

Both LL and LL^{\prime} have the same Galois closure, a 𝒟4\mathcal{D}_{4}-octic extension of \mathbb{Q}. Likewise, the outer automorphism of 𝒮6\mathcal{S}_{6} permits the association to each sextic étale algebra L/KL/K a mirror sextic étale algebra LL^{\prime}.

Example 4.10.

The Cayley embedding is an embedding of any group GG into Sym(G)\operatorname{Sym}(G), acting by left multiplication. The Cayley embedding ρ:𝒮n𝒮n!\rho:\mathcal{S}_{n}\hookrightarrow\mathcal{S}_{n!} attaches to every étale algebra LL of degree nn an algebra L~\tilde{L} of degree n!n! with an 𝒮n\mathcal{S}_{n}-torsor structure. This is none other than the 𝒮n\mathcal{S}_{n}-closure of LL, constructed by Bhargava in a quite different way in [B3, Section 2].

More generally, for any GSnG\subseteq S_{n}, the Cayley embedding GSym(G)G\hookrightarrow\operatorname{Sym}(G) allows one to associate to each GG-extension LL a GG-torsor TT, which we may call the GG-closure of LL. The name “closure” is justified by the following observation: if GSnG\subseteq S_{n} is a transitive subgroup, then, since any transitive GG-set is a quotient of the simply transitive one, we can embed LL into TT by Proposition 4.11 below. More generally, GG-closures of ring extensions, not necessarily étale or even reduced, have been constructed and studied by Biesel [Biesel_thesis, Biesel].

If ρ:GH\rho:G\mathop{\rightarrow}\limits H is invertible, as in many of the above examples, then the map from GG-extensions to HH-extensions is also invertible: we say that the two extensions are mutual resolvents.

4.4 Subextensions and automorphisms

The Galois group holds the answers to various natural questions about an étale algebra. The next two propositions are given without proof, since they follow immediately from the functorial character of the correspondence in Proposition 4.1

Proposition 4.11.

The subextensions LLL^{\prime}\subseteq L of an étale extension L/KL/K, correspond to the equivalence relations \sim on Coord(L)\operatorname{Coord}(L) stable under permutation by G(L/K)G(L/K), under the bijection

L={xL:ι(x)=ι(x) whenever ιι}.\mathord{\sim}\mapsto L^{\prime}=\{x\in L:\iota(x)=\iota^{\prime}(x)\text{ whenever }\iota\sim\iota^{\prime}\}.
Remark 4.12.

Note that if LL is a Galois field extension, the image of ϕL\phi_{L} is a simply transitive subgroup Γ\mathcal{\Gamma}, and identifying Coord(L)\operatorname{Coord}(L) with Γ\mathcal{\Gamma}, the stable equivalence relations are just right congruences modulo subgroups of Γ\mathcal{\Gamma}: so we recover the Galois correspondence between subgroups and subfields.

The Galois group is not a group of automorphisms of LL. However, the automorphisms of LL as a KK-algebra can be described in terms of the Galois group readily.

Proposition 4.13.

Let LL be Minkowski-embedded by its coordinates ι1,,ιn\iota_{1},\ldots,\iota_{n}. Then the automorphism group Aut(L/K)\operatorname{Aut}(L/K) is given by permutations of coordinates,

τπ(x1;;xn)=xπ1(1);;xπ1(n),\tau_{\pi}(x_{1};\ldots;x_{n})=x_{\pi^{-1}(1)};\ldots;x_{\pi^{-1}(n)},

for π\pi in the centralizer C(Sn,G(L/K))C(S_{n},G(L/K)) of the Galois group.

(For HGH\subseteq G groups, the centralizer C(G,H)C(G,H) of HH in GG is the subgroup of elements of GG that commute with every element of HH.)

This provides a characterization, in terms of the Galois group, of rings having various kinds of automorphisms.

  • Since S2S_{2} is abelian, any étale algebra LL of rank 22 has a unique non-identity automorphism, the conjugation x¯=trxx\bar{x}=\operatorname{tr}x-x.

  • If LL has rank 44, automorphisms τ\tau of LL of order 22 whose fixed algebra is of rank 22 are in bijection with D4D_{4}-structures on LL. Indeed, the conditions force τ\tau to correspond to the permutation π=(12)(34)\pi=(12)(34) or one of its conjugates, and the centralizer of this permutation is D4D_{4}.

  • Particularly relevant is the case that LL has a complete set of automorphisms that permute the coordinates simply transitively: this is a generalization of a Galois field extension called a torsor. This case is sufficiently important to merit its own subsection.

4.5 Torsors

Definition 4.14.

Let GG be a finite group. A GG-torsor over KK is an étale algebra LL over KK equipped with an action of GG by automorphisms {τg}gG\{\tau_{g}\}_{g\in G} that permute the coordinates simply transitively, that is, such that LKK¯L\otimes_{K}\bar{K} is isomorphic to

gGK¯\bigoplus_{g\in G}\bar{K}

with GG acting by right multiplication on the indices.

Proposition 4.15.

Let GG be a group of order nn. An étale algebra LL is a GG-torsor if and only if it is a GG-extension, where GG is embedded into SnS_{n} by the Cayley embedding (GG acting on itself by left multiplication). Moreover, there is a bijection between

  • GG-torsor structures on LL, up to conjugation in GG, and

  • GG-structures on LL.

The bijection is given in the following way: there is a labeling {ιg}\{\iota_{g}\} of the coordinates of LL with the elements of GG such that the Galois action is by left multiplication

g(ιh(x))=ιϕgh(x)g(\iota_{h}(x))=\iota_{\phi_{g}h}(x) (11)

while the torsor action is by right multiplication

ιg(τh(x))=ιgh1(x).\iota_{g}(\tau_{h}(x))=\iota_{gh^{-1}}(x). (12)
Proof.

We first claim that the only elements of Sym(G)\operatorname{Sym}(G) commuting with all right multiplications are left multiplications, and vice versa. If π:GG\pi:G\mathop{\rightarrow}\limits G is a permutation commuting with left multiplications, then

π(g)=π(gidG)=gπ(idG),\pi(g)=\pi(g\cdot\operatorname{id}_{G})=g\cdot\pi(\operatorname{id}_{G}),

so π\pi is a right multiplication. So the embedded images of GG in Sym(G)\operatorname{Sym}(G) given by left and right multiplication (which are conjugate under the inversion permutation 1Sym(G)\bullet^{-1}\in\operatorname{Sym}(G)) are centralizers of one another. It is then clear that conjugates GG^{\prime} of GG in Sym(Coord(L))\operatorname{Sym}(\operatorname{Coord}(L)) that contain G(L/K)G(L/K) are in bijection with conjugates GG^{\prime\prime} that commute with G(L/K)G(L/K). This establishes the first assertion. For the bijection of structures, if an embedding GGSym(Coord(L))G\cong G^{\prime}\subseteq\operatorname{Sym}(\operatorname{Coord}(L)) is given, then we can label the coordinates with elements of GG so that GG acts on them by multiplication; then GG^{\prime\prime} gets identified with GG by the corresponding right action. The only ambiguity is in which embedding is labeled with the identity element; if this is changed, one computes that the resulting identification of GG^{\prime\prime} with GG is merely conjugated, so the map is well defined. The reverse map is constructed in exactly the same way. ∎

Here is another perspective on torsors.

Proposition 4.16.

GG-torsors over a field KK, up to isomorphism, are determined by their field factor, a Galois extension L1/KL_{1}/K equipped with an embedding Gal(L/K)G\operatorname{Gal}(L/K)\hookrightarrow G up to conjugation in GG.

Proof.

If TT is a GG-torsor, then since GG permutes the coordinates simply transitively, all the coordinates have the same image; that is, the field factors of GG are all isomorphic to a Galois extension L/KL/K. The torsor operations fixing one field factor LiL_{i} of TT realize the Galois group Gal(L/K)\operatorname{Gal}(L/K) as a subgroup of GG; changing the field factor LiL_{i} and/or the identification LiLL_{i}\cong L corresponds to conjugating the map Gal(L/K)G\operatorname{Gal}(L/K)\hookrightarrow G by an element of GG.

Conversely, suppose LL and an embedding

Gal(L/K)HG\operatorname{Gal}(L/K)\mathop{\longrightarrow}\limits^{\sim}H\subseteq G

are given. Let 1=g1,,gr1=g_{1},\ldots,g_{r} be coset representatives for G/HG/H. Then g2,,grg_{2},\ldots,g_{r} must map any field factor L1LL_{1}\cong L isomorphically onto the remaining field factors L2,,LrL_{2},\ldots,L_{r}, each LiL_{i} occurring once. To finish specifying the GG-action on TL1××LrT\cong L_{1}\times\cdots\times L_{r}, it suffices to determine g|Lig|_{L_{i}} for each gGg\in G. Factor ggi=gjhgg_{i}=g_{j}h for some j{1,,r}j\in\{1,\ldots,r\}, hHh\in H. Then for each xL1x\in L_{1}, g(gi(x))=gj(h(x))g(g_{i}(x))=g_{j}(h(x)), and the value of this is known because the HH-action on L1L_{1} is known. It is easy to see that we get one and only one consistent GG-torsor action in this way. ∎

Because all field factors of a torsor are isomorphic, we will sometimes speak of “the” field factor of a torsor.

4.5.1 Torsors over étale algebras

On occasion, we will speak of a GG-torsor over LL, where LL is itself a product K1××KrK_{1}\times\cdots\times K_{r} of fields. By this we simply mean a product T1××TrT_{1}\times\cdots\times T_{r} where each TiT_{i} is a GG-torsor over KiK_{i}. This case is without conceptual difficulty, and some theorems on torsors will be found to extend readily to it, such as the following variant of the fundamental theorem of Galois theory:

Theorem 4.17.

Let TT be a GG-torsor over an étale algebra LL. For each subgroup HGH\subseteq G,

  1. (a)

    The fixed algebra THT^{H} is uniformly of degree [G:H][G:H] over LL (that is, of this same degree over each field factor of LL);

  2. (b)

    TT is an HH-torsor over THT^{H}, under the same action;

  3. (c)

    If HH is normal, then THT^{H} is also a G/HG/H-torsor over LL, under the natural action.

Proof.

Adapt the relevant results from Galois theory. ∎

4.6 A fresh look at Galois cohomology

Galois cohomology is one of the basic tools in the development of class field theory. It is usually presented in a highly abstract fashion, but certain Galois cohomology groups, specifically H1(K,M)H^{1}(K,M) for finite MM, have explicit meaning in terms of field extensions of MM. It seems that this interpretation is well known but has not yet been written down fully, a gap that we fill in here. We begin by describing Galois modules.

Proposition 4.18 (a description of Galois modules).

Let MM be a finite abelian group, and let KK be a field. Let MM^{-} denote the subset of elements of MM of maximal order mm, the exponent of MM. The following objects are in bijection:

  1. ((a))

    Galois module structures on MM over KK, that is, continuous homomorphisms ϕ:GKAutM\phi:G_{K}\mathop{\rightarrow}\limits\operatorname{Aut}M;

  2. ((b))

    (AutM)(\operatorname{Aut}M)-torsors T/KT/K;

  3. ((c))

    (AutM)(\operatorname{Aut}M)-extensions L0/KL_{0}/K, where AutMSymM\operatorname{Aut}M\hookrightarrow\operatorname{Sym}M in the natural way;

  4. ((d))

    (AutM)(\operatorname{Aut}M)-extensions L/KL^{-}/K, where AutMSymM\operatorname{Aut}M\hookrightarrow\operatorname{Sym}M^{-} in the natural way.

Proof.

For item d to make sense, we need that MM^{-} generates MM; this follows easily from the classification of finite abelian groups.

The bijections are immediate from Propositions 4.4 and 4.15. ∎

We will denote MM with its Galois-module structure coming from these bijections by MϕM_{\phi}, MTM_{T}, or ML0M_{L_{0}}. Note that TT, L0L_{0}, and LL^{-} are mutual resolvents.

Example 4.19.

For example (and we will return to this case frequently), if we let M=𝒞3M=\mathcal{C}_{3} be the smallest group with nontrivial automorphism group: AutM𝒞2\operatorname{Aut}M\cong\mathcal{C}_{2}. Then the Galois module structures on MM are in natural bijection with 𝒞2\mathcal{C}_{2}-torsors over KK, that is, quadratic étale extensions T/KT/K. If charK2\operatorname{char}K\neq 2, these can be parametrized by Kummer theory as T=K[D]T=K[\sqrt{D}], DK×/(K×)2D\in K^{\times}/\left(K^{\times}\right)^{2}. The value D=1D=1 corresponds to the split algebra T=K×KT=K\times K and to the module MM with trivial action. We have an isomorphism

MT{0,D,D}M_{T}\cong\{0,\sqrt{D},-\sqrt{D}\}

of GKG_{K}-sets, and of Galois modules if the right-hand side is given the appropriate group structure with 0 as identity.

In particular, the Galois-module structures on 𝒞3\mathcal{C}_{3} form a group Hom(GK,𝒞2)K×/(K×)2\operatorname{Hom}(G_{K},\mathcal{C}_{2})\cong K^{\times}/\left(K^{\times}\right)^{2}: the group operation can also be viewed as tensor product of one-dimensional 𝔽3\mathbb{F}_{3}-vector spaces with Galois action.

4.6.1 Galois cohomology

Note that the zeroth cohomology group H0(K,M)H^{0}(K,M) has a ready parametrization:

Proposition 4.20.

Let M=ML0M=M_{L_{0}} be a Galois module. The elements of H0(K,M)H^{0}(K,M) are in bijection with the degree-11 field factors of L0L_{0}.

Proof.

Proposition 4.18 establishes an isomorphism of GKG_{K}-sets between the coordinates of L0L_{0} and the points of MM. A degree-11 field factor corresponds to an orbit of GKG_{K} on Coord(L0)\operatorname{Coord}(L_{0}) of size 11, which corresponds exactly to a fixed point of GKG_{K} on MM. ∎

Deeper and more useful is a description of H1H^{1}. For an abelian group MM, let 𝒢𝒜(M)=MAutM\mathcal{GA}(M)=M\rtimes\operatorname{Aut}M be the semidirect product under the natural action of AutM\operatorname{Aut}M on MM. We can describe 𝒢𝒜(M)\mathcal{GA}(M) more explicitly as the group of affine-linear transformations of MM; that is, maps

ag,t(x)=gx+t,gAutM,tMa_{g,t}(x)=gx+t,\quad g\in\operatorname{Aut}M,t\in M

composed of an automorphism and a translation, the group operation being composition. In particular, we have an embedding

𝒢𝒜(M)Sym(M).\mathcal{GA}(M)\hookrightarrow\operatorname{Sym}(M).
Proposition 4.21 (a description of H1H^{1}).

Let M=Mϕ=ML0M=M_{\phi}=M_{L_{0}} be a Galois module.

  1. ((a))

    Z1(K,M)Z^{1}(K,M) is in natural bijection with the set of continuous homomorphisms ψ:GK𝒢𝒜(M)\psi:G_{K}\mathop{\rightarrow}\limits\mathcal{GA}(M) such that the following triangle commutes:

    GK\textstyle{G_{K}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ψ\scriptstyle{\psi}ϕ\scriptstyle{\phi}𝒢𝒜(M)\textstyle{\mathcal{GA}(M)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}AutM\textstyle{\operatorname{Aut}M} (13)
  2. ((b))

    H1(K,M)H^{1}(K,M) is in natural bijection with the set of such ψ:GK𝒢𝒜(M)\psi:G_{K}\mathop{\rightarrow}\limits\mathcal{GA}(M) up to conjugation by M𝒢𝒜(M)M\subseteq\mathcal{GA}(M).

  3. ((c))

    H1(K,M)H^{1}(K,M) is also in natural bijection with the set of 𝒢𝒜(M)\mathcal{GA}(M)-extensions L/KL/K (with respect to the embedding 𝒢𝒜(M)Sym(M)\mathcal{GA}(M)\hookrightarrow\operatorname{Sym}(M)) equipped with an isomorphism from their resolvent (AutM)(\operatorname{Aut}M)-torsor to TT.

Proof.

By the standard construction of group cohomology, Z1Z^{1} is the group of continuous crossed homomorphisms

Z1(K,M)={σ:GKMσ(γδ)=σ(γ)+ϕ(γ)σ(δ)}.Z^{1}(K,M)=\{\sigma:G_{K}\mathop{\rightarrow}\limits M\mid\sigma(\gamma\delta)=\sigma(\gamma)+\phi(\gamma)\sigma(\delta)\}.

Send each σ\sigma to the map

ψ:GK\displaystyle\psi:G_{K} 𝒢𝒜(M)\displaystyle\mathop{\rightarrow}\limits\mathcal{GA}(M)
γ\displaystyle\gamma aϕ(γ),σ(γ).\displaystyle\mapsto a_{\phi(\gamma),\sigma(\gamma)}.

It is easy to see that the conditions for ψ\psi to be a homomorphism are exactly those for σ\sigma to be a crossed homomorphism, establishing a. For b, we observe that adding a coboundary σa(γ)=γ(a)a\sigma_{a}(\gamma)=\gamma(a)-a to a crossed homomorphism σ\sigma is equivalent to post-conjugating the associated map ψ:GK𝒢𝒜(M)\psi:G_{K}\mathop{\rightarrow}\limits\mathcal{GA}(M) by aa. As to c, a 𝒢𝒜(M)\mathcal{GA}(M)-extension carries the same information as a map ψ\psi up to conjugation by the whole of 𝒢𝒜(M)\mathcal{GA}(M). Specifying the isomorphism from the resolvent (AutM)(\operatorname{Aut}M)-torsor to TT means that the map πψ=ϕ:GKAut(M)\pi\circ\psi=\phi:G_{K}\mathop{\rightarrow}\limits\operatorname{Aut}(M) is known exactly, not just up to conjugation. Hence ψ\psi is known up to conjugation by MM. ∎

Remark 4.22.

The zero cohomology class corresponds to the extension L0L_{0}, with its structure given by the embedding AutM𝒢𝒜(M)\operatorname{Aut}M\hookrightarrow\mathcal{GA}(M). This can be seen to be the unique cohomology class whose corresponding 𝒢𝒜(X)\mathcal{GA}(X)-extension has a field factor of degree 11.

If KK is a local field, a cohomology class αH1(K,M)\alpha\in H^{1}(K,M) is called unramified if it is represented by a cocycle α:Gal(K¯/K)M\alpha:\operatorname{Gal}(\bar{K}/K)\mathop{\rightarrow}\limits M that factors through the unramified Galois group Gal(Kur/K)\operatorname{Gal}(K^{\mathrm{ur}}/K). The subgroup of unramified coclasses is denoted by H1ur(K,M)H^{1}_{\mathrm{ur}}(K,M). If MM itself is unramified (and we will never have to think about unramified cohomology in any other case), this is equivalent to the associated étale algebra LL being unramified.

If X=MTX=M_{T} is a Galois module and σZ1(K,M)\sigma\in Z^{1}(K,M) is the Galois module corresponding to a 𝒢𝒜(X)\mathcal{GA}(X)-extension L/KL/K, we can also take the 𝒢𝒜(X)\mathcal{GA}(X)-closure of LL, a 𝒢𝒜(X)\mathcal{GA}(X)-torsor EE which fits into the following diagram:

ETLK\begin{gathered}\lx@xy@svg{\hbox{\raise 0.0pt\hbox{\kern 36.20668pt\hbox{\ignorespaces\ignorespaces\ignorespaces\hbox{\vtop{\kern 0.0pt\offinterlineskip\halign{\entry@#!@&&\entry@@#!@\cr&\\&\crcr}}}\ignorespaces{\hbox{\kern-6.97916pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{E\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces{}\ignorespaces\ignorespaces{\hbox{\lx@xy@drawline@}}\ignorespaces{}\ignorespaces\ignorespaces{\hbox{\lx@xy@drawline@}}\ignorespaces{\hbox{\lx@xy@drawline@}}\ignorespaces\ignorespaces\ignorespaces{}\ignorespaces\ignorespaces{\hbox{\lx@xy@drawline@}}\ignorespaces{}\ignorespaces\ignorespaces{\hbox{\lx@xy@drawline@}}\ignorespaces{\hbox{\lx@xy@drawline@}}{\hbox{\kern 23.18758pt\raise-29.8039pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{T\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces{}\ignorespaces\ignorespaces{\hbox{\lx@xy@drawline@}}\ignorespaces{}\ignorespaces\ignorespaces{\hbox{\lx@xy@drawline@}}\ignorespaces{\hbox{\lx@xy@drawline@}}{\hbox{\kern-36.20668pt\raise-29.8039pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{L\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces{}\ignorespaces\ignorespaces{\hbox{\lx@xy@drawline@}}\ignorespaces{}\ignorespaces\ignorespaces{\hbox{\lx@xy@drawline@}}\ignorespaces{\hbox{\lx@xy@drawline@}}{\hbox{\kern-7.60416pt\raise-59.60779pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{K}$}}}}}}}\ignorespaces}}}}\ignorespaces\end{gathered} (14)

Because of the semidirect product structure of 𝒢𝒜(X)\mathcal{GA}(X), we have ELKTE\cong L\otimes_{K}T. It is also worth tabulating the permutation representations of finite groups that yield each of the étale algebras discussed here:

0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}M\textstyle{M\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝒢𝒜(M)\textstyle{\mathcal{GA}(M)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}yields LLyields EEAut(M)\textstyle{\operatorname{Aut}(M)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}yields TT0\textstyle{0}Sym(M)\textstyle{\operatorname{Sym}(M)}Sym(𝒢𝒜(M))\textstyle{\operatorname{Sym}(\mathcal{GA}(M))}Sym(Aut(M))\textstyle{\operatorname{Sym}(\operatorname{Aut}(M))}
(15)
4.6.2 The Tate dual

If MM is a Galois module and the exponent mm of MM is not divisible by charK\operatorname{char}K, then

M=Hom(M,μm)M^{\prime}=\operatorname{Hom}(M,\mu_{m})

is also a Galois module, called the Tate dual of MM. The modules MM and MM^{\prime} have the same order and are isomorphic as abstract groups, though not canonically; as Galois modules, they are frequently not isomorphic at all.

Example 4.23.

If M=MK[D]M=M_{K[\sqrt{D}]} is one of the order-33 modules studied in Example 4.19, then the relevant μm\mu_{m} is

μ3MK[3].\mu_{3}\cong M_{K[\sqrt{-3}]}.

Examining the Galois actions (here it helps to use the theory of GKG_{K}-sets of size 22 presented in Knus and Tignol [QuarticExercises]), we see that

M=MK[3D].M^{\prime}=M_{K[\sqrt{-3D}]}.

This explains the D3DD\mapsto-3D pattern in the Scholz reflection theorem and its generalizations, including cubic Ohno-Nakagawa.

Example 4.24.

A module MM of underlying group 𝒞2×𝒞2\mathcal{C}_{2}\times\mathcal{C}_{2} is always self-dual, regardless of what Galois-module structure is placed on it. This can be proved by noting that MM has a unique alternating bilinear form

B:M×M\displaystyle B:M\times M μ2\displaystyle\mathop{\rightarrow}\limits\mu_{2}
(x,y)\displaystyle(x,y) {1,x=0,y=0, or x=y1,otherwise.\displaystyle\mapsto\begin{cases}1,&x=0,y=0,\text{ or }x=y\\ -1,&\text{otherwise.}\end{cases}

Being unique, it is Galois-stable and induces an isomorphism MMM^{\prime}\cong M.

Particularly notable for us are the cases when 𝒢𝒜(M)\mathcal{GA}(M) is the full symmetric group Sym(M)\operatorname{Sym}(M), for then every étale algebra L/KL/K of degree |M|\lvert M\rvert has a (unique) 𝒢𝒜(M)\mathcal{GA}(M)-affine structure. It is easy to see that there are only four such cases:

  • M={1}M=\{1\}, 𝒢𝒜(M)𝒮1\mathcal{GA}(M)\cong\mathcal{S}_{1}

  • M=/2M=\mathbb{Z}/2\mathbb{Z}, 𝒢𝒜(M)𝒮2\mathcal{GA}(M)\cong\mathcal{S}_{2}

  • M=/3M=\mathbb{Z}/3\mathbb{Z}, 𝒢𝒜(M)𝒞3𝒞2𝒮3\mathcal{GA}(M)\cong\mathcal{C}_{3}\rtimes\mathcal{C}_{2}\cong\mathcal{S}_{3}

  • M=/2×/2M=\mathbb{Z}/2\mathbb{Z}\times\mathbb{Z}/2\mathbb{Z}, 𝒢𝒜(M)(𝒞2×𝒞2)𝒮3𝒮4\mathcal{GA}(M)\cong(\mathcal{C}_{2}\times\mathcal{C}_{2})\rtimes\mathcal{S}_{3}\cong\mathcal{S}_{4}.

For degree exceeding 44, not every étale algebra arises from Galois cohomology, a restriction that plays out in the existing literature on reflection theorems. For instance, Cohen, Rubinstein-Salzedo, and Thorne [CohON] prove a reflection theorem in which one side counts 𝒟p\mathcal{D}_{p}-dihedral fields of prime degree p3p\geq 3. From our perspective, these correspond to cohomology classes of an M=𝒞pM=\mathcal{C}_{p} whose Galois action is by ±1\pm 1. The Tate dual of such an MM can have Galois action by the full (/p)×(\mathbb{Z}/p\mathbb{Z})^{\times}, and indeed they count extensions of Galois group 𝒢𝒜(𝒞p)\mathcal{GA}(\mathcal{C}_{p}) on the other side of the reflection theorem. This will appear inevitable in light of the motivations elucidated in Part III.

5 Extensions of Kummer theory to explicitize Galois cohomology

Now that Galois cohomology groups H1(K,M)H^{1}(K,M) have been parametrized by étale algebras, can invoke parametrizations of étale algebras by even more explicit objects. The most familiar instance of this is Kummer theory, an isomorphism

H1(K,μm)K×/(K×)mH^{1}(K,\mu_{m})\cong K^{\times}/(K^{\times})^{m}

coming from the long exact sequence associated to the Kummer sequence

0μmK¯×mK¯×0.0\mathop{\longrightarrow}\limits\mu_{m}\mathop{\longrightarrow}\limits\bar{K}^{\times}\mathop{\longrightarrow}\limits^{\bullet^{m}}\bar{K}^{\times}\mathop{\longrightarrow}\limits 0.

In favorable cases, the cohomology H1(K,M)H^{1}(K,M) of other Galois modules MM can be embedded into R×/(R×)mR^{\times}/(R^{\times})^{m} for some finite extension RR of KK.

We first state the hypothesis we need:

Definition 5.1.

Let MM be a finite Galois module of exponent mm over a field KK, and let XX be a Galois-stable generating set of MM. We say that MM equipped with XX is a good module if the natural map of Galois modules

𝔛=xX(/m)\displaystyle\mathfrak{X}=\bigoplus_{x\in X}(\mathbb{Z}/m\mathbb{Z}) M\displaystyle\mathop{\rightarrow}\limits M
ex\displaystyle e_{x} x\displaystyle\mapsto x

is split, that is, its kernel admits a Galois-stable complementary direct summand M~\tilde{M}. Such a direct summand is known as a good structure on MM.

Proposition 5.2.

The following examples of a Galois module MM with generating set XX are good:

  1. ((a))

    M𝒞pM\cong\mathcal{C}_{p}, with any action, and X=M{0}X=M\setminus\{0\}.

  2. ((b))

    M𝒞mnM\cong\mathcal{C}_{m}^{n}, with any action preserving a basis XX.

  3. ((c))

    M𝒞mn1M\cong\mathcal{C}_{m}^{n-1}, n2n\geq 2 with gcd(m,n)\gcd(m,n), with an action that preserves a hyperbasis XX, that is, a generating set of nn elements with sum 0.

Proof.
  1. ((a))

    Here the Galois modules are representations of 𝔽p×𝒞p1\mathbb{F}_{p}^{\times}\cong\mathcal{C}_{p-1} over 𝔽p\mathbb{F}_{p}. Since the group and field are of coprime order, complete reducibility holds: any subrepresentation is a direct summand. In fact, 𝔛\mathfrak{X} is the regular representation, MM is the tautological representation in which each λ𝔽p×\lambda\in\mathbb{F}_{p}^{\times} acts by multiplication by λ\lambda, and M~\tilde{M} can be taken (uniquely in general) to be the product of all the other isotypical components of 𝔛\mathfrak{X}.

  2. ((b))

    Here the natural map 𝔛M\mathfrak{X}\mathop{\rightarrow}\limits M is an isomorphism, so M~=𝔛\tilde{M}=\mathfrak{X}.

  3. ((c))

    Here the natural map 𝔛M\mathfrak{X}\mathop{\rightarrow}\limits M is the quotient by the one-dimensional space

    xXex.\left\langle\sum_{x\in X}e_{x}\right\rangle.

    This space has a Galois-stable direct complement, namely the kernel M~\tilde{M} of the linear functional

    𝔛\displaystyle\mathfrak{X} 𝔽2\displaystyle\mathop{\rightarrow}\limits\mathbb{F}_{2}
    ex\displaystyle e_{x} 1.\displaystyle\mathop{\rightarrow}\limits 1.\qed
Proposition 5.3.

Let MM be a Galois module with a good structure (X,M~)(X,\tilde{M}), and let RR be the resolvent algebra corresponding to the GKG_{K}-set XX. For any Galois module AA with underlying group /m\mathbb{Z}/m\mathbb{Z}, there is a natural injection

H1(K,MA)H1(R,A)H^{1}(K,M\otimes A)\mathop{\rightarrow}\limits H^{1}(R,A)

as a direct summand. The cokernel is naturally isomorphic to

H1(K,(𝔛/M~)A).H^{1}(K,(\mathfrak{X}/\tilde{M})\otimes A).
Proof.

We use the good structure

MM~𝔛M\cong\tilde{M}\hookrightarrow\mathfrak{X}

to embed

H1(K,M~A)H1(𝔛A).H^{1}(K,\tilde{M}\otimes A)\hookrightarrow H^{1}(\mathfrak{X}\otimes A).

Since M~\tilde{M} is a direct summand, this is an injection with cokernel naturally isomorphic to H1(K,(𝔛/M~)A)H^{1}(K,(\mathfrak{X}/\tilde{M})\otimes A). It remains to construct an isomorphism

H1(𝔛A)H1(R,A).H^{1}(\mathfrak{X}\otimes A)\stackrel{{\scriptstyle\sim}}{{\longrightarrow}}H^{1}(R,A).

If RR decomposes as a product

RR1××RsR\cong R_{1}\times\cdots\times R_{s}

of field factors corresponding to the orbits X=iXiX=\bigsqcup_{i}X_{i} of GKG_{K} on XX, then 𝔛\mathfrak{X} has a corresponding decomposition

𝔛=i=1s𝔛i\mathfrak{X}=\bigoplus_{i=1}^{s}\mathfrak{X}_{i}

where 𝔛i=ex:xXi\mathfrak{X}_{i}=\left\langle e_{x}:x\in X_{i}\right\rangle is none other than the induced module IndKRi/m\operatorname{Ind}_{K}^{R_{i}}\mathbb{Z}/m\mathbb{Z}. Its cohomology is computed by Shapiro’s lemma:

H1(K,𝔛A)i=1sH1(K,𝔛iA)=i=1sH1(K,IndKRiA)i=1sH1(Ri,A)=H1(R,A).H^{1}(K,\mathfrak{X}\otimes A)\cong\bigoplus_{i=1}^{s}H^{1}(K,\mathfrak{X}_{i}\otimes A)=\bigoplus_{i=1}^{s}H^{1}(K,\operatorname{Ind}_{K}^{R_{i}}A)\cong\bigoplus_{i=1}^{s}H^{1}(R_{i},A)=H^{1}(R,A).

This is the desired isomorphism. ∎

We can harness Kummer theory to parametrize cohomology of other modules as follows.

Theorem 5.4 (an extension of Kummer theory).

Let MM be a finite Galois module, and assume that m=expMm=\exp M is not divisible by charK\operatorname{char}K. Let GKG_{K} act on the set MM^{\prime-} of surjective characters χ:Mμm\chi:M\twoheadrightarrow\mu_{m} through its actions on MM and μm\mu_{m}, and let FF be the étale algebra corresponding to this GKG_{K}-set.

  1. ((a))

    There is a natural group homomorphism

    Kum:H1(K,M)F×/(F×)m.\operatorname{Kum}:H^{1}(K,M)\mathop{\rightarrow}\limits F^{\times}/(F^{\times})^{m}.
  2. ((b))

    If M𝒞pM\cong\mathcal{C}_{p} is cyclic of prime order, then Kum\operatorname{Kum} is injective, FF is naturally a (/p)×(\mathbb{Z}/p\mathbb{Z})^{\times}-torsor, and

    im(Kum)={αF×/(F×)p:τc(α)=αcc(/p)×}.\operatorname{im}(\operatorname{Kum})=\left\{\alpha\in F^{\times}/(F^{\times})^{p}:\tau_{c}(\alpha)=\alpha^{c}\,\forall c\in(\mathbb{Z}/p\mathbb{Z})^{\times}\right\}.

    If p=3p=3, then the image simplifies to

    im(Kum)={αF×/(F×)3:NF/K(α)=1},\operatorname{im}(\operatorname{Kum})=\left\{\alpha\in F^{\times}/(F^{\times})^{3}:N_{F/K}(\alpha)=1\right\},

    and the 𝒢𝒜(𝒞3)S3\mathcal{GA}(\mathcal{C}_{3})\cong S_{3}-extension LL corresponding to a given αF×/(F×)3\alpha\in F^{\times}/(F^{\times})^{3} of norm 11 can be described as follows: Define a KK-linear map

    κ:K\displaystyle\kappa:K K¯3\displaystyle\mathop{\rightarrow}\limits\bar{K}^{3}
    ξ\displaystyle\xi (trK¯2/Kξωδ3)ω,\displaystyle\mapsto\left(\operatorname{tr}_{\bar{K}^{2}/K}\xi\omega\sqrt[3]{\delta}\right)_{\omega},

    where δ3K¯2\sqrt[3]{\delta}\in\bar{K}^{2} is chosen to have norm 11, and ω\omega ranges through the set

    {(1;1),(ζ3;ζ32);(ζ32;ζ3)}\{(1;1),(\zeta_{3};\zeta_{3}^{2});(\zeta_{3}^{2};\zeta_{3})\}

    of cube roots of 11 in K¯2\bar{K}^{2} of norm 11. Then

    L=K+κ(F).L=K+\kappa(F).
  3. ((c))

    If M𝒞2×𝒞2M\cong\mathcal{C}_{2}\times\mathcal{C}_{2}, then Kum\operatorname{Kum} is injective and

    im(Kum)={αF×/(F×)2:NF/K(α)=1}.\operatorname{im}(\operatorname{Kum})=\left\{\alpha\in F^{\times}/(F^{\times})^{2}:N_{F/K}(\alpha)=1\right\}.

    Moreover, the 𝒢𝒜(M)S4\mathcal{GA}(M)\cong S_{4}-extension corresponding to a given αF×/(F×)2\alpha\in F^{\times}/(F^{\times})^{2} of norm 11 can be described as follows: Define a KK-linear map

    κ:K\displaystyle\kappa:K K¯4\displaystyle\mathop{\rightarrow}\limits\bar{K}^{4}
    ξ\displaystyle\xi (trK¯3/Kξωδ)ω,\displaystyle\mapsto\left(\operatorname{tr}_{\bar{K}^{3}/K}\xi\omega\sqrt{\delta}\right)_{\omega},

    where δK¯3\sqrt{\delta}\in\bar{K}^{3} is chosen to have norm 11, and ω\omega ranges through the set

    {(1;1;1),(1;1;1);(1;1;1);(1;1;1)}\{(1;1;1),(1;-1;-1);(-1;1;-1);(-1;-1;1)\}

    of square roots of 11 in K¯3\bar{K}^{3} of norm 11. Then

    L=K+κ(F).L=K+\kappa(F).
Proof.

If χ:Mμm\chi:M\twoheadrightarrow\mu_{m} is a surjective character, let FχF_{\chi} be the fixed field of the stabilizer of χ\chi; thus FχF_{\chi} is the field factor of FF corresponding to the GKG_{K}-orbit of χ\chi. If χ1,,χ\chi_{1},\ldots,\chi_{\ell} are orbit representatives, we can map

H1(K,M)resiH1(Fχi,M)χiiH1(Fχi,μm)iFχi×/(Fχi×)mF×/(F×)m.H^{1}(K,M)\mathop{\longrightarrow}\limits^{\prod\operatorname{res}}\prod_{i}H^{1}(F_{\chi_{i}},M)\mathop{\longrightarrow}\limits^{\prod\chi_{i*}}\prod_{i}H^{1}(F_{\chi_{i}},\mu_{m})\cong\prod_{i}F_{\chi_{i}}^{\times}/(F_{\chi_{i}}^{\times})^{m}\cong F^{\times}/(F^{\times})^{m}.

This yields our map Kum\operatorname{Kum}. Alternatively, note that by Shapiro’s lemma,

iH1(Fχi,μm)iH1(K,IndFχiKμm)H1(K,I),\prod_{i}H^{1}(F_{\chi_{i}},\mu_{m})\cong\prod_{i}H^{1}(K,\operatorname{Ind}_{F_{\chi_{i}}}^{K}\mu_{m})\cong H^{1}(K,I),

where

IM=IndFKμm=χ:Mμmμm,I_{M}=\operatorname{Ind}_{F}^{K}\mu_{m}=\bigoplus_{\chi:M\twoheadrightarrow\mu_{m}}\mu_{m},

a Galois module under the action

g((aχ)i)=(g(cg1(χ))χ)=(g(cχ(g))χ).g\big{(}(a_{\chi})_{i}\big{)}=\big{(}g(c_{g^{-1}(\chi)})_{\chi}\big{)}=\big{(}g(c_{\chi(g\bullet)})_{\chi}\big{)}.

Under this identification, it is not hard to check that Kum=j\operatorname{Kum}=j_{*}, where jj is the inclusion MIMM\hookrightarrow I_{M} given by

a(χ(a))χ.a\mapsto(\chi(a))_{\chi}.

Although jj is injective (because the characters of maximal order mm generate the group of all characters), it is not obvious whether jj induces an injection on cohomology, nor what the image is. What makes the modules MM in parts b and c tractable is that, in these cases, M\{0}M\backslash\{0\} is a good generating set for MM, so MM is a direct summand of IMI_{M}. In part b, we can identify

IMInd{1}(/p)×𝔽p𝔽pμmI_{M}\cong\operatorname{Ind}_{\{1\}}^{(\mathbb{Z}/p\mathbb{Z})^{\times}}\mathbb{F}_{p}\otimes_{\mathbb{F}_{p}}\mu_{m}

as a twist of the regular representation of (/p)×(\mathbb{Z}/p\mathbb{Z})^{\times} over 𝔽p\mathbb{F}_{p}. Since 𝔽p\mathbb{F}_{p} has a complete set of (p1)(p-1)st roots of unity, this representation splits completely into one-dimensional subrepresentations. The image of jj is the eigenspace generated by (c)c/p×(c)_{c\in\mathbb{Z}/p\mathbb{Z}^{\times}}, so Kum\operatorname{Kum} is injective and its image is the subspace of F×/(F×)pF^{\times}/(F^{\times})^{p} cut out by the same relations τc(x)=cx\tau_{c}(x)=cx (where τc\tau_{c} is the torsor operation on FF, resp. the automorphism of IMI_{M}, indexed by cc) that cut out j(M)j(M) in IMI_{M}.

As to part c, since /2×/2\mathbb{Z}/2\mathbb{Z}\times\mathbb{Z}/2\mathbb{Z} has three surjective characters whose product is 11, we have IM/Mμ2I_{M}/M\cong\mu_{2} with the map η:IMμ2\eta:I_{M}\mathop{\rightarrow}\limits\mu_{2} given by multiplying the coordinates. Since μ2\mu_{2} also injects diagonally into IMI_{M}, we easily get a direct sum decomposition, which shows that Kum\operatorname{Kum} is injective. As to the image, it is not hard to show that the diagram

F×/(F×)2\textstyle{F^{\times}/(F^{\times})^{2}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\sim}N\scriptstyle{N}H1(F,𝒞2)\textstyle{H^{1}(F,\mathcal{C}_{2})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\sim}cor\scriptstyle{\operatorname{cor}}H1(K,I)\textstyle{H^{1}(K,I)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}η\scriptstyle{\eta_{*}}K×/(K×)2\textstyle{K^{\times}/(K^{\times})^{2}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\sim}H1(K,C2)\textstyle{H^{1}(K,C_{2})}

commutes, establishing the desired norm characterization of im(Kum)\operatorname{im}(\operatorname{Kum}).

The formulas by radicals for the cubic and quartic algebras corresponding to a Kummer element follow easily by chasing through the Galois actions on the appropriate étale algebras. The quartic case is also considered by Knus and Tignol, where a closely related description of LL is given ([QuarticExercises], Proposition 5.13). ∎

Remark 5.5.

For general MM, the map H1(K,M)F×/(F×)mH^{1}(K,M)\mathop{\rightarrow}\limits F^{\times}/(F^{\times})^{m} may be made by the construction in Theorem 5.4, but its image is hard to characterize, and it may not even be injective: for instance, when M𝒞4M\cong\mathcal{C}_{4}, coclasses correspond to 𝒟4\mathcal{D}_{4}-extensions, and Kum\operatorname{Kum} conflates each extension with its mirror extension (compare Example 4.9).

Though it will not be used in the sequel, it is worth noting that Artin-Schreyer theory is amenable to the same treatment.

Theorem 5.6.

Let Let MM be a finite Galois module with underlying abelian group AA of exponent m=p=charKm=p=\operatorname{char}K.

  1. ((a))

    There is a natural map

    AS:H1(K,M)F/(F).\operatorname{AS}:H^{1}(K,M)\mathop{\rightarrow}\limits F/\wp(F).
  2. ((b))

    If A𝒞pA\cong\mathcal{C}_{p}, then AS\operatorname{AS} is injective, FF is naturally a (/p)×(\mathbb{Z}/p\mathbb{Z})^{\times}-torsor, and

    im(AS)={αF/(F):τc(α)=cαc(/p)×}.\operatorname{im}(\operatorname{AS})=\left\{\alpha\in F/\wp(F):\tau_{c}(\alpha)=c\alpha\;\forall c\in(\mathbb{Z}/p\mathbb{Z})^{\times}\right\}.
  3. ((c))

    If p=2p=2 and A𝒞2×𝒞2A\cong\mathcal{C}_{2}\times\mathcal{C}_{2}, then AS\operatorname{AS} is injective and

    im(AS)={αF/(F):trF/K(α)=0}.\operatorname{im}(\operatorname{AS})=\left\{\alpha\in F/\wp(F):\operatorname{tr}_{F/K}(\alpha)=0\right\}.

5.1 The Tate pairing and the Hilbert symbol

Assume now that KK is a local field. Our next step will be to understand the (local) Tate pairing, which is given by a cup product

,T:H1(K,M)×H1(K,M)H2(K,μm)μm.\langle\,,\,\rangle_{T}:H^{1}(K,M)\times H^{1}(K,M^{\prime})\mathop{\rightarrow}\limits H^{2}(K,\mu_{m})\cong\mu_{m}.

As we were able to parametrize the cohomology groups H1(K,M)H^{1}(K,M) in favorable cases, it should not come as a surprise that we can often describe the Tate pairing with similar explicitness.

Recall the definitions of the Artin and Hilbert symbols. If M/mM\cong\mathbb{Z}/m\mathbb{Z} has trivial GKG_{K}-action, then MμmM^{\prime}\cong\mu_{m}, and we have a Tate pairing

,T:H1(K,/m)×H1(K,μm)μm\langle\,,\,\rangle_{T}:H^{1}(K,\mathbb{Z}/m\mathbb{Z})\times H^{1}(K,\mu_{m})\mathop{\rightarrow}\limits\mu_{m}

Now H1(K,/m)Hom(K,/m)H^{1}(K,\mathbb{Z}/m\mathbb{Z})\cong\operatorname{Hom}(K,\mathbb{Z}/m\mathbb{Z}) parametrizes /m\mathbb{Z}/m\mathbb{Z}-torsors, while by Kummer theory, H1(K,μm)K×/(K×)mH^{1}(K,\mu_{m})\cong K^{\times}/(K^{\times})^{m}. The Tate pairing in this case is none other than the Artin symbol (or norm-residue symbol) ϕL(x)\phi_{L}(x) which attaches to a cyclic extension LL, of degree dividing mm, a mapping ϕL:K×Gal(L/K)μm\phi_{L}:K^{\times}\mathop{\rightarrow}\limits\operatorname{Gal}(L/K)\mathop{\rightarrow}\limits\mu_{m} whose kernel is the norm group NL/K(L×)N_{L/K}(L^{\times}) (see Neukirch [NeukirchCoho], Prop. 7.2.13). If, in addition, μmK\mu_{m}\subseteq K, then H1(K,/m)H^{1}(K,\mathbb{Z}/m\mathbb{Z}) is also isomorphic to K×/(K×)mK^{\times}/\left(K^{\times}\right)^{m}, and the Tate pairing is an alternating pairing

,:K×/(K×)m×K×/(K×)mμm\langle\,,\,\rangle:K^{\times}/\left(K^{\times}\right)^{m}\times K^{\times}/\left(K^{\times}\right)^{m}\mathop{\rightarrow}\limits\mu_{m}

classically called the Hilbert symbol (or Hilbert pairing). It is defined in terms of the Artin symbol by

a,b=ϕK[bm](a).\left\langle a,b\right\rangle=\phi_{K[\sqrt[m]{b}]}(a). (16)

In particular, a,b=1\left\langle a,b\right\rangle=1 if and only if aa is the norm of an element of K[bm]K[\sqrt[m]{b}]. This can also be described in terms of the splitting of an appropriate Severi-Brauer variety; for instance, if m=2m=2, we have a,b=1\left\langle a,b\right\rangle=1 exactly when the conic

ax2+by2=z2ax^{2}+by^{2}=z^{2}

has a KK-rational point. See also Serre ([SerreLF], §§XIV.1–2). (All identifications between pairings here are up to sign; the signs are not consistent in the literature and are totally irrelevant for this paper.) Pleasantly, for the types of MM featured in Theorem 5.4, the Tate pairing can be expressed simply in terms of the Hilbert pairing.

We extend the Hilbert pairing to étale algebras in the obvious way: if L=K1××KsL=K_{1}\times\cdots\times K_{s}, then

(a1;;as),(b1;;bs)L:=a1,b1K1as,bsKs.\left\langle(a_{1};\ldots;a_{s}),(b_{1};\ldots;b_{s})\right\rangle_{L}:=\left\langle a_{1},b_{1}\right\rangle_{K_{1}}\cdot\cdots\cdot\left\langle a_{s},b_{s}\right\rangle_{K_{s}}.

Note that if aa is a norm from L[bm]L[\sqrt[m]{b}] to LL, then a,bL=1\left\langle a,b\right\rangle_{L}=1, but the converse no longer holds. We then have the following:

Theorem 5.7 (a formula for the local Tate pairing).

Let KK be a local field. For MM, FF as in Theorem 5.4, let MM^{\prime} be the Tate dual of MM, and let FF^{\prime} be the corresponding étale algebra, corresponding to the GKG_{K}-set MM^{-} of elements of maximal order in MM, just as FF corresponds to MM^{\prime-}. The Tate pairing

,:H1(K,M)×H1(K,M)H2(K,μm)𝒞m\left\langle\bullet,\bullet\right\rangle:H^{1}(K,M)\times H^{1}(K,M^{\prime})\mathop{\rightarrow}\limits H^{2}(K,\mu_{m})\cong\mathcal{C}_{m}

can be described in terms of the Hilbert pairing in the following cases:

  1. (a)

    If A𝒞pA\cong\mathcal{C}_{p}, then both FF and FF^{\prime} embed naturally into F:=F[μp]F^{\prime\prime}:=F[\mu_{p}], and the Tate pairing is the restriction of the Hilbert pairing on FF^{\prime\prime}.

  2. (b)

    If A(/2)2A\cong(\mathbb{Z}/2\mathbb{Z})^{2}, then we have natural isomorphisms MMM\cong M^{\prime}, FFF\cong F^{\prime}, and the Tate pairing is the restriction of the Hilbert pairing on FF.

Proof.

In case b, set F=F=F[μ2]F^{\prime\prime}=F=F[\mu_{2}]. We will do the two cases largely in parallel.

Let Surj(A,B)Hom(A,B)\operatorname{Surj}(A,B)\subseteq\operatorname{Hom}(A,B) denote the set of surjections between two groups A,BA,B. Note that if A,BA,B are Galois modules, then Surj(A,B)\operatorname{Surj}(A,B) is a GKG_{K}-set. Note that FF^{\prime\prime} is the étale algebra corresponding to the GKG_{K}-set

Z=Surj(M,μp)×Surj(μp,/m).Z=\operatorname{Surj}(M,\mu_{p})\times\operatorname{Surj}(\mu_{p},\mathbb{Z}/m\mathbb{Z}).

There is an obvious map ZSurj(M,μp)Z\mathop{\rightarrow}\limits\operatorname{Surj}(M,\mu_{p}) given by projection to the first factor, which allows us to recover the identification F=F[μm]F^{\prime\prime}=F[\mu_{m}]. There is also a map of GKG_{K}-sets

Ψ:ZSurj(M,μp)\Psi:Z\mathop{\rightarrow}\limits\operatorname{Surj}(M^{\prime},\mu_{p})

which sends a pair (χ,u)(\chi,u) (where χ:Mμm\chi:M\twoheadrightarrow\mu_{m}, u:μmu:\mu_{m}) to the unique surjective ψ:Mμp\psi:M^{\prime}\mathop{\rightarrow}\limits\mu_{p} satisfying

{ψ(χ)=u1(1),A𝒞pψ(χ)=1,A𝒞2×𝒞2.\begin{cases}\psi(\chi)=u^{-1}(1),&A\cong\mathcal{C}_{p}\\ \psi(\chi)=1,&A\cong\mathcal{C}_{2}\times\mathcal{C}_{2}.\end{cases}

This allows us to embed FF^{\prime} into FF^{\prime\prime}. It is worth noting that when A𝒞2×𝒞2A\cong\mathcal{C}_{2}\times\mathcal{C}_{2}, uu carries no information and FFFF\cong F^{\prime}\cong F^{\prime\prime}.

Let F1,,FF_{1}^{\prime\prime},\ldots,F_{\ell}^{\prime\prime} be the field factors of FF^{\prime\prime}; each FiF_{i}^{\prime\prime} corresponds to an orbit GK(χi,ui)G_{K}(\chi_{i},u_{i}) on ZZ. Let ψi=Ψ(χi,ui).\psi_{i}=\Psi(\chi_{i},u_{i}). Then for σH1(K,M)\sigma\in H^{1}(K,M), τH1(K,M)\tau\in H^{1}(K,M^{\prime}),

σ,τHilb\displaystyle\left\langle\sigma,\tau\right\rangle_{\text{Hilb}} =Kum(σ),Kum(τ)Hilb; F\displaystyle=\left\langle\operatorname{Kum}(\sigma),\operatorname{Kum}(\tau)\right\rangle_{\text{Hilb; }F^{\prime\prime}}
=FiinvFi(χiresKFiσψiresKFiτ).\displaystyle=\prod_{F^{\prime\prime}_{i}}\operatorname{inv}_{F^{\prime\prime}_{i}}\big{(}\chi_{i*}\operatorname{res}^{K}_{F_{i}^{\prime\prime}}\sigma\cup\psi_{i*}\operatorname{res}^{K}_{F_{i}^{\prime\prime}}\tau\big{)}.

Since invFi=invKcorKFi\operatorname{inv}_{F^{\prime\prime}_{i}}=\operatorname{inv}_{K}\circ\operatorname{cor}^{K}_{F^{\prime\prime}_{i}} (a standard fact), we have

σ,τHilb\displaystyle\left\langle\sigma,\tau\right\rangle_{\text{Hilb}} =invKFicorKFi(χiresKFiσψiresKFiτ)\displaystyle=\operatorname{inv}_{K}\sum_{F_{i}^{\prime\prime}}\operatorname{cor}^{K}_{F_{i}^{\prime\prime}}\big{(}\chi_{i*}\operatorname{res}^{K}_{F_{i}^{\prime\prime}}\sigma\cup\psi_{i*}\operatorname{res}^{K}_{F_{i}^{\prime\prime}}\tau\big{)}
=invKFicorKFi(ev(χiψi))resKFi(στ)\displaystyle=\operatorname{inv}_{K}\sum_{F_{i}^{\prime\prime}}\operatorname{cor}^{K}_{F_{i}^{\prime\prime}}\big{(}\operatorname{ev}\circ(\chi_{i}\otimes\psi_{i})\big{)}_{*}\operatorname{res}^{K}_{F_{i}^{\prime\prime}}(\sigma\cup\tau)

where ev:MMμm\operatorname{ev}:M\otimes M^{\prime}\mathop{\rightarrow}\limits\mu_{m} is the evaluation map. We now apply the following lemma, which slightly generalizes results seen in the literature.

Lemma 5.8.

Let HGH\subseteq G be a subgroup of finite index. Let XX and YY be GG-modules, and let f:XYf:X\mathop{\rightarrow}\limits Y be a map that is HH-linear (but not necessarily GG-linear). Denote by f~\tilde{f} the GG-linear map

f~(x)=gHG/Hgfg1(x).\tilde{f}(x)=\sum_{gH\in G/H}gfg^{-1}(x).

Let σHn(H,Y)\sigma\in H^{n}(H,Y). Then

corHG(fresHGσ)=f~σ.\operatorname{cor}_{H}^{G}(f_{*}\operatorname{res}_{H}^{G}\sigma)=\tilde{f}_{*}\sigma.
Proof.

Since we are concerned with the equality of a pair of δ\delta-functors, we can apply dimension shifting to assume that n=0n=0. The proof is now straightforward. ∎

Applying with G=GKG=G_{K}, H=GFiH=G_{F^{\prime\prime}_{i}}, and f=ev(χiψi):MMμmf=\operatorname{ev}\circ(\chi_{i}\otimes\psi_{i}):M\otimes M^{\prime}\mathop{\rightarrow}\limits\mu_{m}, we get

σ,τHilb\displaystyle\left\langle\sigma,\tau\right\rangle_{\text{Hilb}} =invKFi(gGK/GFigev(χig1ψig1))\displaystyle=\operatorname{inv}_{K}\sum_{F_{i}^{\prime\prime}}\Big{(}\sum_{g\in G_{K}/G_{F_{i}^{\prime\prime}}}g\circ\operatorname{ev}\circ(\chi_{i}\circ g^{-1}\otimes\psi_{i}\circ g^{-1})\Big{)}
=invKFi(gGK/GFi(ev(χi,gψi,g)))(στ),\displaystyle=\operatorname{inv}_{K}\sum_{F_{i}^{\prime\prime}}\Big{(}\sum_{g\in G_{K}/G_{F_{i}^{\prime\prime}}}\big{(}\operatorname{ev}\circ(\chi_{i,g}\otimes\psi_{i,g})\big{)}\Big{)}_{*}(\sigma\cup\tau),

where χi,g=g(χi)\chi_{i,g}=g(\chi_{i}) and ψi,g=g(ψi)\psi_{i,g}=g(\psi_{i}) are given by the natural action. Now the outer sum runs over all GKG_{K}-orbits of ZZ while the inner sum runs over the elements of each orbit, so we simply get

σ,τHilb=invK((χ,u)Zev(χΨ(χ,u)))(στ).\left\langle\sigma,\tau\right\rangle_{\text{Hilb}}=\operatorname{inv}_{K}\Big{(}\sum_{(\chi,u)\in Z}\operatorname{ev}\circ(\chi\otimes\Psi(\chi,u))\Big{)}_{*}(\sigma\cup\tau).

Since the Tate pairing is given by

σ,τTate=invKev(στ),\left\langle\sigma,\tau\right\rangle_{\text{Tate}}=\operatorname{inv}_{K}\operatorname{ev}_{*}(\sigma\cup\tau),

it remains to check that

(χ,u)Zev(χΨ(χ,u))=ev\sum_{(\chi,u)\in Z}\operatorname{ev}\circ(\chi\otimes\Psi(\chi,u))=\operatorname{ev}

as maps from MMM\otimes M^{\prime} to μm\mu_{m}. In the case M𝒞pM\cong\mathcal{C}_{p}, each term is actually equal to ev\operatorname{ev}, and there are (p1)21(p-1)^{2}\equiv 1 mod pp terms. In the case M𝒞2×𝒞2M\cong\mathcal{C}_{2}\times\mathcal{C}_{2}, a direct verification on a basis of MMM\otimes M^{\prime} is not difficult. ∎

6 Rings over a Dedekind domain

Thus far, we have been considering étale algebras LL over a field KK. We now suppose that KK is the fraction field of a Dedekind domain 𝒪K\mathcal{O}_{K} (not of characteristic 22), which for us will usually be a number field or a completion thereof, although there is no need to be so restrictive. Our topic of study will be the subrings of LL that are lattices of full rank over 𝒪K\mathcal{O}_{K}—the orders, to use the standard but unfortunately overloaded word.

There is always a unique maximal order 𝒪L\mathcal{O}_{L}, the integral closure of 𝒪K\mathcal{O}_{K} in LL. If L=L1××LrL=L_{1}\times\cdots\times L_{r} is a product of field factors, we have 𝒪L=𝒪L1××𝒪Lr\mathcal{O}_{L}=\mathcal{O}_{L_{1}}\times\cdots\times\mathcal{O}_{L_{r}}.

6.1 Indices of lattices

There is one piece of notation that we explain here to avoid confusion. If VV is an nn-dimensional vector space over KK and A,BVA,B\subseteq V are two full-rank lattices, we denote by the index [A:B][A:B] the unique fractional ideal 𝔠\mathfrak{c} such that

ΛnA=𝔠ΛnB\Lambda^{n}A=\mathfrak{c}\Lambda^{n}B

as 𝒪K\mathcal{O}_{K}-submodules of the top exterior power ΛnV\Lambda^{n}V. Alternatively, if ABA\supseteq B, then the classification theorem for finitely generated modules over 𝒪K\mathcal{O}_{K} lets us write

A/B𝒪K/𝔠1𝒪K/𝔠r,A/B\cong\mathcal{O}_{K}/\mathfrak{c}_{1}\oplus\cdots\oplus\mathcal{O}_{K}/\mathfrak{c}_{r},

and the index equals

𝔠1𝔠2𝔠r.\mathfrak{c}_{1}\mathfrak{c}_{2}\cdots\mathfrak{c}_{r}.

The index satisfies the following basic properties:

  • [A:B][B:C]=[A:C][A:B][B:C]=[A:C];

  • If VV is a vector space over both KK and a finite extension LL, and AA and BB are two 𝒪L\mathcal{O}_{L}-sublattices, then [A:B]K=NL/K[A:B]L[A:B]_{K}=N_{L/K}[A:B]_{L};

  • If V=LV=L is a KK-algebra and αL×\alpha\in L^{\times}, then [A:αA]=NL/K(α)[A:\alpha A]=N_{L/K}(\alpha).

Despite the apparent abstractness of its definition, the index [A:B][A:B] is not hard to compute in particular cases: localizing at a prime ideal, we can assume 𝒪K\mathcal{O}_{K} is a PID, and then it is the determinant of the matrix expressing any basis of BB in terms of a basis of AA.

If LL is a KK-algebra, 𝒪L\mathcal{O}\subseteq L is an order, and 𝔞L\mathfrak{a}\subseteq L is a fractional ideal, the index [𝒪:𝔞][\mathcal{O}:\mathfrak{a}] is called the norm of 𝔞\mathfrak{a} and will be denoted by N𝒪(𝔞)N_{\mathcal{O}}(\mathfrak{a}) or, when the context is clear, by N(𝔞)N(\mathfrak{a}). Note the following basic properties:

  • If 𝔞=α𝒪\mathfrak{a}=\alpha\mathcal{O} is principal, then N𝒪(𝔞)=NL/K(α)N_{\mathcal{O}}(\mathfrak{a})=N_{L/K}(\alpha).

  • If 𝔞\mathfrak{a} and 𝔟\mathfrak{b} are two 𝒪\mathcal{O}-ideals and 𝔞\mathfrak{a} is invertible, then N(𝔞𝔟)=N(𝔞)N(𝔟)N(\mathfrak{a}\mathfrak{b})=N(\mathfrak{a})N(\mathfrak{b}). This is easily derived from the theorem that an invertible ideal is locally principal (Lemma LABEL:lem:inv=pri). It is false for two arbitrary 𝒪\mathcal{O}-ideals.

  • If 𝒪K=\mathcal{O}_{K}=\mathbb{Z} or p\mathbb{Z}_{p}, then for any LL and 𝒪\mathcal{O}, the norm N𝒪(𝔞)N_{\mathcal{O}}(\mathfrak{a}) of an integral ideal is the ideal generated by the absolute norm |𝒪/𝔞|.\lvert\mathcal{O}/\mathfrak{a}\rvert.

6.2 Discriminants

As is standard, we define the discriminant ideal of an order 𝒪\mathcal{O} in an étale algebra LL to be the ideal 𝔡\mathfrak{d} generated by the trace pairing

τ:𝒪2n\displaystyle\tau:\mathcal{O}^{2n} 𝒪K\displaystyle\mathop{\rightarrow}\limits\mathcal{O}_{K} (17)
(ξ1,ξ2,,ξn,η1,η2,,ηn)\displaystyle(\xi_{1},\xi_{2},\ldots,\xi_{n},\eta_{1},\eta_{2},\ldots,\eta_{n}) det[trξiηj]i,j=1n.\displaystyle\mapsto\det[\operatorname{tr}\xi_{i}\eta_{j}]_{i,j=1}^{n}.

The trace pairing is nondegenerate, that is, 𝔡0\mathfrak{d}\neq 0 (this is one equivalent definition of étale). The primes dividing 𝔡\mathfrak{d} are those at which LL is ramified and/or 𝒪\mathcal{O} is nonmaximal. This notion is standard and widely used. However, it does not quite extend the (also standard) notion of the discriminant of a \mathbb{Z}-algebra over \mathbb{Z}, which has a distinction between positive and negative discriminants. The Ohno-Nakagawa theorem involves this distinction prominently; Dioses [Dioses] and Cohen–Rubinstein-Salzedo–Thorne [CohON] each frame their extensions of O-N in terms of an ad-hoc notion of discriminant that incorporates the splitting data of an order at the infinite primes. Here we explain the variant that we will use.

Since the trace pairing τ\tau is alternating in the ξ\xi’s and also in the η\eta’s, it can be viewed as a bilinear form on the rank-11 lattice Λn𝒪\Lambda^{n}\mathcal{O}. Identifying Λn𝒪\Lambda^{n}\mathcal{O} with a (fractional) ideal 𝔠\mathfrak{c} of 𝒪K\mathcal{O}_{K} (whose class is often called the Steinitz class of 𝒪\mathcal{O}), we can write

τ(ξ)=Dξ2\tau(\xi)=D\xi^{2}

for some nonzero D𝔠2D\in\mathfrak{c}^{-2}. Had we rescaled the identification Λn𝒪𝔠\Lambda^{n}\mathcal{O}\mathop{\rightarrow}\limits\mathfrak{c} by λK×\lambda\in K^{\times}, DD would be multiplied by λ2\lambda^{2}. We call the pair (𝔠,D)(\mathfrak{c},D), up to the equivalence (𝔠,D)(λ𝔠,λ2D)(\mathfrak{c},D)\sim(\lambda\mathfrak{c},\lambda^{-2}D), the discriminant of 𝒪\mathcal{O} and denote it by Disc𝒪\operatorname{Disc}\mathcal{O}.

There is another perspective on the discriminant Disc𝒪\operatorname{Disc}\mathcal{O}. Let L~\tilde{L} be the SnS_{n}-torsor corresponding to LL, which comes with nn embeddings κ1,,κn:LL~\kappa_{1},\ldots,\kappa_{n}:L\mathop{\rightarrow}\limits\tilde{L} freely permuted by the SnS_{n}-action (not to be confused with the nn coordinates of LL). Noting that, for any αL\alpha\in L,

tr(α)=iκi(α),\operatorname{tr}(\alpha)=\sum_{i}\kappa_{i}(\alpha),

we can factor the trace pairing matrix:

[trξiηj]i,j=[κh(ξi)]i,h[κh(ηj)]h,j.[\operatorname{tr}\xi_{i}\eta_{j}]_{i,j}=[\kappa_{h}(\xi_{i})]_{i,h}\cdot[\kappa_{h}(\eta_{j})]_{h,j}.

Define

τ0(ξ1,,ξn)=det[κh(ξi)]i,h,\tau_{0}(\xi_{1},\ldots,\xi_{n})=\det[\kappa_{h}(\xi_{i})]_{i,h},

so that

τ(ξ1,,ξn,η1,,ηn)=τ0(ξ1,,ξn)τ0(η1,,ηn).\tau(\xi_{1},\ldots,\xi_{n},\eta_{1},\ldots,\eta_{n})=\tau_{0}(\xi_{1},\ldots,\xi_{n})\cdot\tau_{0}(\eta_{1},\ldots,\eta_{n}).

Now look more carefully at the map τ0\tau_{0}. First, τ0\tau_{0} is alternating under permutations of the ξi\xi_{i}’s, so it defines a linear map

τ0:𝔠L~.\tau_{0}:\mathfrak{c}\mathop{\rightarrow}\limits\tilde{L}.

Moreover, τ0\tau_{0} is alternating under postcomposition by the torsor action of SnS_{n} on L~\tilde{L}, which permutes the κh\kappa_{h} freely. Thus the image of τ0\tau_{0} lies in the C2C_{2}-torsor T2=L~AnT_{2}=\tilde{L}^{A_{n}}, which we call the discriminant torsor of LL, and even more specifically in the (1)(-1)-eigenspace of the nontrivial element of C2C_{2}. By (the simplest case of) Kummer theory, we may write T2=K[D]T_{2}=K[\sqrt{D^{\prime}}]. If (ξ1,,ξn)(\xi_{1},\ldots,\xi_{n}) is any KK-basis of 𝒪L\mathcal{O}_{L}, so that ξ1ξn\xi_{1}\wedge\cdots\wedge\xi_{n} corresponds to some nonzero element c𝔠c\in\mathfrak{c}, then

Dc2=τ(c,c)=τ0(c)2=(aD)2=Da2.Dc^{2}=\tau(c,c)=\tau_{0}(c)^{2}=\left(a\sqrt{D^{\prime}}\right)^{2}=D^{\prime}a^{2}.

Thus T2=K[D]T_{2}=K[\sqrt{D}]. We summarize this result in a proposition.

Proposition 6.1.

If LL is an étale algebra over KK of discriminant (𝔠,D)(\mathfrak{c},D), then K[D]K[\sqrt{D}] is the discriminant torsor of LL; that is, the diagram of Galois structure maps

GK\textstyle{G_{K}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ϕL\scriptstyle{\phi_{L}}ϕK[D]\scriptstyle{\phi_{K[\sqrt{D}]}}Sn\textstyle{S_{n}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}sgn\scriptstyle{\operatorname{sgn}}S2\textstyle{S_{2}}

commutes.

There is notable integral structure on DD as well.

Lemma 6.2 (Stickelberger’s theorem over Dedekind domains).

If (𝔠,D)(\mathfrak{c},D) is the discriminant of an order 𝒪\mathcal{O}, then Dt2D\equiv t^{2} mod 4𝔠24\mathfrak{c}^{-2} for some t𝔠1t\in\mathfrak{c}^{-1}.

Remark 6.3.

When 𝒪K=\mathcal{O}_{K}=\mathbb{Z}, Lemma 6.2 states that the discriminant of an order is congruent to 0 or 11 mod 44: a nontrivial and classical theorem due to Stickelberger. Our proof is a generalization of the most familiar one for Stickelberger’s theorem, due to Schur [Schur1929].

Proof.

Since D𝔠2D\in\mathfrak{c}^{-2}, the conclusion can be checked locally at each prime dividing 22 in 𝒪K\mathcal{O}_{K}. We can thus assume that 𝒪K\mathcal{O}_{K} is a DVR and in particular that 𝔠=(1)\mathfrak{c}=(1). Now there is a simple tensor ξ1ξn\xi_{1}\wedge\cdots\wedge\xi_{n} that corresponds to the element 1𝔠1\in\mathfrak{c}. By definition,

D=τ0(ξ1,,ξn)=det[κh(ξi)]i,h=πSn(sgn(σ)iκπ(i)(ξi))=ρρ¯,\sqrt{D}=\tau_{0}(\xi_{1},\ldots,\xi_{n})=\det[\kappa_{h}(\xi_{i})]_{i,h}=\sum_{\pi\in S_{n}}\Big{(}\operatorname{sgn}(\sigma)\prod_{i}\kappa_{\pi(i)}(\xi_{i})\Big{)}=\rho-\bar{\rho}, (18)

where

ρ=πAniκπ(i)(ξi)\rho=\sum_{\pi\in A_{n}}\prod_{i}\kappa_{\pi(i)}(\xi_{i})

lies in T2T_{2} by symmetry and ρ¯\bar{\rho} is its conjugate. By construction, ρ\rho is integral over 𝒪K\mathcal{O}_{K}, that is to say, ρ+ρ¯\rho+\bar{\rho} and ρρ¯\rho\bar{\rho} lie in 𝒪K\mathcal{O}_{K}. Now

D=(ρρ¯)2=(ρ+ρ¯)24ρρ¯D=(\rho-\bar{\rho})^{2}=(\rho+\bar{\rho})^{2}-4\rho\bar{\rho}

is the sum of a square and a multiple of 44 in 𝒪K\mathcal{O}_{K}. ∎

Remark 6.4.

One can write (18) in the suggestive form

det[κh(ξi)]i,h=det[θ1(1)θ2(1)θ1(ρ)θ2(ρ)]\det[\kappa_{h}(\xi_{i})]_{i,h}=\det\begin{bmatrix}\theta_{1}(1)&\theta_{2}(1)\\ \theta_{1}(\rho)&\theta_{2}(\rho)\end{bmatrix}

where θ1,θ2\theta_{1},\theta_{2} are the two automorphisms of T2T_{2}. This equates discriminants of orders in LL with those of orders in T2T_{2}. Equalities of determinants of this sort reappear in Bhargava’s parametrizations of quartic and quintic rings and appear to be a common feature of many types of resolvent fields.

We can now state the notion of discriminant as we would like to use it.

Definition 6.5.

A discriminant over 𝒪K\mathcal{O}_{K} is an equivalence class of pairs (𝔠,D)(\mathfrak{c},D), with D𝔠2D\in\mathfrak{c}^{-2} and Dt2D\equiv t^{2} mod 4𝔠24\mathfrak{c}^{2} for some t𝔠1t\in\mathfrak{c}^{-1}, up to the equivalence relation

(𝔠,D)(λ𝔠,λ2D).(\mathfrak{c},D)\sim(\lambda\mathfrak{c},\lambda^{-2}D).

If 𝒪\mathcal{O} is an étale order, the discriminant Disc𝒪\operatorname{Disc}\mathcal{O} is defined as follows: Pick any representation ϕ:Λn𝒪𝔠\phi:\Lambda^{n}\mathcal{O}\mathop{\rightarrow}\limits\mathfrak{c} of the Steinitz class as an ideal class; then Disc𝒪\operatorname{Disc}\mathcal{O} is the unique pair (𝔠,D)(\mathfrak{c},D) such that

det[trξiηj]i,j=1n=Dϕ(ξ1ξn)ϕ(η1ηn).\det[\operatorname{tr}\xi_{i}\eta_{j}]_{i,j=1}^{n}=D\cdot\phi(\xi_{1}\wedge\cdots\wedge\xi_{n})\cdot\phi(\eta_{1}\wedge\cdots\wedge\eta_{n}).

Note the following points.

  • The discriminant recovers the discriminant ideal via 𝔡=D𝔠2\mathfrak{d}=D\mathfrak{c}^{2}.

  • If LL has degree 33, the discriminant also contains the splitting information of LL at the infinite primes. Namely, for each real place ι\iota of KK, if ι(D)>0\iota(D)>0 then Lv××L_{v}\cong\mathbb{R}\times\mathbb{R}\times\mathbb{R}, while if ι(D)<0\iota(D)<0 then Lv×L_{v}\cong\mathbb{R}\times\mathbb{C}.

  • By a usual abuse of language, if LL is an étale algebra over a number field KK, its discriminant is the discriminant of the ring of integers 𝒪L\mathcal{O}_{L} over 𝒪K\mathcal{O}_{K}.

  • The discriminants over 𝒪K\mathcal{O}_{K} form a cancellative semigroup under the multiplication law

    (𝔠1,D1)(𝔠2,D2)=(𝔠1𝔠2,D1D2).(\mathfrak{c}_{1},D_{1})(\mathfrak{c}_{2},D_{2})=(\mathfrak{c}_{1}\mathfrak{c}_{2},D_{1}D_{2}).
  • If 𝒪K\mathcal{O}_{K} is a PID, then we can take 𝔠=𝒪K\mathfrak{c}=\mathcal{O}_{K}, and then the discriminants are simply nonzero elements D𝒪KD\in\mathcal{O}_{K} congruent to a square mod 44, up to multiplication by squares of units.

  • We will often denote a discriminant by a single letter, such as 𝒟\mathcal{D}. When elements or ideals of 𝒪K\mathcal{O}_{K} appear in discriminants, they are to be understood as follows:

    D(D𝒪K)\displaystyle D\quad(D\in\mathcal{O}_{K})\quad means((1),D)\displaystyle\text{means}\quad((1),D) (19)
    𝔠2(𝔠K)\displaystyle\mathfrak{c}^{2}\quad(\mathfrak{c}\subseteq K)\quad means(𝔠,1).\displaystyle\text{means}\quad(\mathfrak{c},1). (20)

    The seemingly counterintuitive convention (20) is motivated by the fact that, if 𝔠=(c)\mathfrak{c}=(c) is principal, then (𝔠,1)(\mathfrak{c},1) is the same discriminant as ((1),c2)((1),c^{2}).

With these remarks in place, the reader should not have difficulty reading and proving the following relation:

Proposition 6.6.

If 𝒪𝒪\mathcal{O}\supseteq\mathcal{O}^{\prime} are two orders in an étale algebra LL, then

Disc𝒪=[𝒪:𝒪]2Disc𝒪.\operatorname{Disc}\mathcal{O}^{\prime}=[\mathcal{O}:\mathcal{O}^{\prime}]^{2}\cdot\operatorname{Disc}\mathcal{O}.

6.3 Quadratic rings

We will spend a lot of time investigating the number of rings over 𝒪K\mathcal{O}_{K} of given degree nn and discriminant 𝒟=(𝔠,D)\mathcal{D}=(\mathfrak{c},D). For quadratic rings, the problem has a complete answer:

Proposition 6.7 (the parametrization of quadratic rings).

Let 𝒪K\mathcal{O}_{K} be a Dedekind domain of characteristic not 22. For every discriminant 𝒟\mathcal{D}, there is a unique quadratic étale order 𝒪𝒟\mathcal{O}_{\mathcal{D}} having discriminant 𝒟\mathcal{D}.

Proof.

Note first that the theorem is true when 𝒪K=K\mathcal{O}_{K}=K is a field: by Kummer theory, quadratic étale algebras over KK are parametrized by K×/(K×)2K^{\times}/\left(K^{\times}\right)^{2}, as are discriminants; and it is a simple matter to check that DiscK[D]=D\operatorname{Disc}K[\sqrt{D}]=D. We proceed to the general case.

For existence, let 𝒟=(𝔠,D)\mathcal{D}=(\mathfrak{c},D) be given. By definition, DD is congruent to a square t2t^{2} mod 4𝔠24\mathfrak{c}^{2}, t𝔠1t\in\mathfrak{c}^{-1}. Consider the lattice

𝒪=𝒪K𝔠ξ,ξ=t+D2L=K[D].\mathcal{O}=\mathcal{O}_{K}\oplus\mathfrak{c}\xi,\quad\xi=\frac{t+\sqrt{D}}{2}\in L=K[\sqrt{D}].

To prove that 𝒪\mathcal{O} is an order in LL, it is enough to verify that (cξ)(dξ)𝒪(c\xi)(d\xi)\in\mathcal{O} for any c,d𝔠c,d\in\mathfrak{c}, and this follows from the computation

ξ2=ξ(tξ¯)=tξ(t2D4)\xi^{2}=\xi(t-\bar{\xi})=t\xi-\left(\frac{t^{2}-D}{4}\right)

and the conditions t𝔠1,t2D4𝔠2t\in\mathfrak{c}^{-1},t^{2}-D\in 4\mathfrak{c}^{-2}.

Now suppose that 𝒪1\mathcal{O}_{1} and 𝒪2\mathcal{O}_{2} are two orders with the same discriminant 𝒟=(𝔠,D)\mathcal{D}=(\mathfrak{c},D). Their enclosing KK-algebras L1L_{1}, L2L_{2} have the same discriminant DD over KK, and hence we can identify L1=L2=LL_{1}=L_{2}=L. Now project each 𝒪i\mathcal{O}_{i} along π:LL/K\pi:L\mathop{\rightarrow}\limits L/K is an 𝒪K\mathcal{O}_{K}-lattice 𝔠i\mathfrak{c}_{i} in L/KL/K, which is a one-dimensional KK-vector space: indeed, we naturally have L/KΛ2LL/K\cong\Lambda^{2}L, and upon computation, we find that Disc𝒪i=(𝔠i,D)\operatorname{Disc}\mathcal{O}_{i}=(\mathfrak{c}_{i},D). Consequently 𝔠1=𝔠2=𝔠\mathfrak{c}_{1}=\mathfrak{c}_{2}=\mathfrak{c}. Now, for each β𝔠\beta\in\mathfrak{c}, the fiber π1(𝔠)𝒪i\pi^{-1}(\mathfrak{c})\cap\mathcal{O}_{i} is of the form βi+𝒪K\beta_{i}+\mathcal{O}_{K} for some βi\beta_{i}. The element β1β2\beta_{1}-\beta_{2} is integral over 𝒪K\mathcal{O}_{K} and lies in KK, hence in 𝒪K\mathcal{O}_{K}. Thus 𝒪1=𝒪2\mathcal{O}_{1}=\mathcal{O}_{2}. ∎

If 𝒪\mathcal{O} is any order in an étale algebra L/KL/K (charK2\operatorname{char}K\neq 2), the quadratic order B=𝒪Disc𝒪B=\mathcal{O}_{\operatorname{Disc}\mathcal{O}} having the same discriminant as 𝒪\mathcal{O} is called the quadratic resolvent ring of 𝒪\mathcal{O}. It embeds into the discriminant torsor T2T_{2}, in two conjugate ways. Indeed, it is not hard to show that BB is generated by the elements

ρ(ξ1,,ξn)=πAniκπ(i)(ξi)T2\rho(\xi_{1},\ldots,\xi_{n})=\sum_{\pi\in A_{n}}\prod_{i}\kappa_{\pi(i)}(\xi_{i})\in T_{2}

appearing in the proof of Lemma 6.2.

Remark 6.8.

The notion of a quadratic resolvent ring extends to characteristic 22, being always an order in the quadratic resolvent algebra constructed in Example 4.8. We omit the details.

6.4 Cubic rings

Cubic and quartic rings have parametrizations, known as higher composition laws, linking them to certain forms over 𝒪K\mathcal{O}_{K} and also to ideals in resolvent rings. The study of higher composition laws was inaugurated by Bhargava in his celebrated series of papers ([B1, B2, B3, B4]), although the gist of the parametrization of cubic rings goes back to work of F.W. Levi [Levi]. Later work by Deligne and by Wood [WQuartic, W2xnxn] has extended much of Bhargava’s work from \mathbb{Z} to an arbitrary base scheme. In a previous paper [ORings], the author explained how a representative sample of these higher composition laws extend to the case when the base ring AA is a Dedekind domain. In the present work, we will need a few more; fortunately, there are no added difficulties, and we will briefly run through the statements and the methods of proof.

Theorem 6.9 (the parametrization of cubic rings).

Let AA be a Dedekind domain with field of fractions KK, charA3\operatorname{char}A\neq 3.

  1. ((a))

    Cubic rings 𝒪\mathcal{O} over AA, up to isomorphism, are in bijection with cubic maps

    Φ:MΛ2M\Phi:M\mathop{\rightarrow}\limits\Lambda^{2}M

    between a two-dimensional AA-lattice MM and its own Steinitz class, up to isomorphism, in the obvious sense of a commutative square

    M1\textstyle{M_{1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\sim}i\scriptstyle{i}Φ1\scriptstyle{\Phi_{1}}M2\textstyle{M_{2}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Φ2\scriptstyle{\Phi_{2}}Λ2M1\textstyle{\Lambda^{2}M_{1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\sim}deti\scriptstyle{\det i}Λ2M2.\textstyle{\Lambda^{2}M_{2}.}

    The bijection sends a ring 𝒪\mathcal{O} to the index form Φ:𝒪/AΛ2(𝒪/A)\Phi:\mathcal{O}/A\mathop{\rightarrow}\limits\Lambda^{2}(\mathcal{O}/A) given by

    xxx2.x\mapsto x\wedge x^{2}.
  2. ((b))

    If 𝒪\mathcal{O} is nondegenerate, that is, the corresponding cubic KK-algebra L=KA𝒪L=K\otimes_{A}\mathcal{O} is étale, then the map Φ\Phi is the restriction, under the Minkowski embedding, of the index form of K¯3\bar{K}^{3}, which is

    Φ:K¯3/K¯\displaystyle\Phi:\bar{K}^{3}/\bar{K} K¯2/K¯\displaystyle\mathop{\rightarrow}\limits\bar{K}^{2}/\bar{K} (21)
    (x;y;z)\displaystyle(x;y;z) ((xy)(yz)(zx),0).\displaystyle\mapsto\big{(}(x-y)(y-z)(z-x),0\big{)}.
  3. ((c))

    Conversely, let LL be a cubic étale algebra over KK. If 𝒪¯L/K\bar{\mathcal{O}}\subseteq L/K is a lattice such that Φ\Phi sends 𝒪¯\bar{\mathcal{O}} into Λ2𝒪¯\Lambda^{2}\bar{\mathcal{O}}, then there is a unique cubic ring 𝒪L\mathcal{O}\subseteq L such that, under the natural identifications, 𝒪/A=𝒪¯\mathcal{O}/A=\bar{\mathcal{O}}.

Proof.
  1. ((a))

    The proof is quite elementary, involving merely solving for the coefficients of the unknown multiplication table of 𝒪\mathcal{O}. The case where AA is a PID is due to Gross ([cubquat], Section 2): the cubic ring having index form

    f(xξ+yη)=ax3+bx2y+cxy2+dy3f(x\xi+y\eta)=ax^{3}+bx^{2}y+cxy^{2}+dy^{3}

    has multiplication table

    ξη=ad,ξ2=ac+bξaη,η2=bd+dξcη.\displaystyle\xi\eta=-ad,\xi^{2}=-ac+b\xi-a\eta,\eta^{2}=-bd+d\xi-c\eta. (22)

    For the general Dedekind case, see my [ORings], Theorem 7.1. It is also subsumed by Deligne’s work over an arbitrary base scheme; see Wood [WQuartic] and the references therein.

  2. ((b))

    This follows from the fact that the index form respects base change. The index form of K¯3/K¯\bar{K}^{3}/\bar{K} is a Vandermonde determinant that can easily be written in the stated form.

  3. ((c))

    We have an integral cubic map Φ|𝒪¯:𝒪¯Λ2𝒪¯\Phi\big{|}_{\bar{\mathcal{O}}}:\bar{\mathcal{O}}\mathop{\rightarrow}\limits\Lambda^{2}\bar{\mathcal{O}}, which is the index form Φ𝒪\Phi_{\mathcal{O}} of a unique cubic ring 𝒪\mathcal{O} over 𝒪K\mathcal{O}_{K}. But over KK, Φ𝒪\Phi_{\mathcal{O}} is isomorphic to the index form of LL. Since LL (as a cubic ring over KK) is determined by its index form, we obtain an identification 𝒪𝒪KKL\mathcal{O}\otimes_{\mathcal{O}_{K}}K\cong L for which 𝒪/A\mathcal{O}/A, the projection of 𝒪\mathcal{O} onto L/KL/K, coincides with 𝒪¯\bar{\mathcal{O}}. The uniqueness of 𝒪\mathcal{O} is obvious, as 𝒪\mathcal{O} must lie in the integral closure 𝒪L\mathcal{O}_{L} of KK in LL.

In this paper we only deal with nondegenerate rings, that is, those of nonzero discriminant, or equivalently, those that lie in an étale KK-algebra. Consequently, all index forms Φ\Phi that we will see are restrictions of (21). When cubic algebras are parametrized Kummer-theoretically, the resolvent map becomes very explicit and simple:

Proposition 6.10 (explicit Kummer theory for cubic algebras).

Let RR be a quadratic étale algebra over KK (charK3\operatorname{char}K\neq 3), and let

L=K+κ(R)L=K+\kappa(R)

be the cubic algebra of resolvent R=RK[μ3]R^{\prime}=R\odot K[\mu_{3}] (the Tate dual of RR) corresponding to an element δK×\delta\in K^{\times} of norm 11 in Theorem 5.4b, where

κ(ξ)=(trK¯2/Kξωδ3)ω(K¯2)N=1[3]K¯3\kappa(\xi)=\left(\operatorname{tr}_{\bar{K}^{2}/K}\xi\omega\sqrt[3]{\delta}\right)_{\omega\in\left(\bar{K}^{2}\right)^{N=1}[3]}\in\bar{K}^{3}

so κ\kappa maps RR bijectively onto the traceless plane in LL. Then the index form of LL is given explicitly by

Φ:L/K\displaystyle\Phi:L/K Λ2(L/K)\displaystyle\mathop{\rightarrow}\limits\Lambda^{2}(L/K) (23)
κ(ξ)\displaystyle\kappa(\xi) 33δξ31,\displaystyle\mapsto 3\sqrt{-3}\delta\xi^{3}\wedge 1,

where we identify

Λ2L/KΛ3LΛ2RR/K3R/K\Lambda^{2}L/K\cong\Lambda^{3}L\cong\Lambda^{2}R^{\prime}\cong R^{\prime}/K\cong\sqrt{-3}\cdot R/K

using the fact that RR^{\prime} is the discriminant resolvent of LL.

Proof.

Direct calculation, after reducing to the case K=K¯K=\bar{K}. ∎

Theorem 6.11 (self-balanced ideals in the cubic case).

Let 𝒪K\mathcal{O}_{K} be a Dedekind domain, charK3\operatorname{char}K\neq 3, and let RR be a quadratic étale extension. A self-balanced triple in RR is a triple (B,I,δ)(B,I,\delta) consisting of a quadratic order BRB\subseteq R, a fractional ideal II of BB, and a scalar δ(KB)×\delta\in(KB)^{\times} satisfying the conditions

δI3B,N(I)=(t) is principal,andN(δ)t3=1,\delta I^{3}\subseteq B,\quad N(I)=(t)\text{ is principal},\quad\text{and}\quad N(\delta)t^{3}=1, (24)
  1. ((a))

    Fix BB and δR×\delta\in R^{\times} with N(δ)N(\delta) a cube t3t^{-3}. Then the mapping

    I𝒪=𝒪K+κ(I)I\mapsto\mathcal{O}=\mathcal{O}_{K}+\kappa(I) (25)

    defines a bijection between

    • self-balanced triples of the form (B,I,δ)(B,I,\delta), and

    • subrings 𝒪L\mathcal{O}\subseteq L of the cubic algebra L=K+κ(R)L=K+\kappa(R) corresponding to the Kummer element δ\delta, such that 𝒪\mathcal{O} is 33-traced, that is, tr(ξ)3𝒪K\operatorname{tr}(\xi)\in 3\mathcal{O}_{K} for every ξL\xi\in L.

  2. ((b))

    Under this bijection, we have the discriminant relation

    discC=27discB.\operatorname{disc}C=-27\operatorname{disc}B. (26)
Proof.

The mapping κ\kappa defines a bijection between lattices IRI\subseteq R and κ(I)L/K\kappa(I)\subseteq L/K. The difficult part is showing that II fits into a self-balanced triple (B,I,δ)(B,I,\delta) if and only if κ(I)\kappa(I) is the projection of a 33-traced order 𝒪\mathcal{O}. Note that if (B,I,δ)(B,I,\delta) exists, it is unique, as the requirement [B:I]=(t)[B:I]=(t) pins down BB.

Rather than establish this equivalence directly, we will show that both conditions are equivalent to the symmetric trilinear form

β:I×I×I\displaystyle\beta:I\times I\times I Λ2R\displaystyle\mathop{\rightarrow}\limits\Lambda^{2}R
(α1,α2,α3)\displaystyle(\alpha_{1},\alpha_{2},\alpha_{3}) δα1α2α3\displaystyle\mathop{\rightarrow}\limits\delta\alpha_{1}\alpha_{2}\alpha_{3}

taking values in t1Λ2It^{-1}\cdot\Lambda^{2}I.

In the case of self-balanced ideals, this was done over \mathbb{Z} by Bhargava [B1, Theorem 3]. Over a Dedekind domain, it follows from the parametrization of balanced triples of ideals over BB [ORings, Theorem 5.3], after specializing to the case that all three ideals are identified with one ideal II. It also follows from the corresponding results over an arbitrary base in Wood [W2xnxn, Theorem 1.4].

In the case of rings, we compute by Proposition 6.10 that β\beta is the trilinear form attached to the index form of κ(I)\kappa(I). By Theorem 6.9c, the diagonal restriction β(α,α,α)\beta(\alpha,\alpha,\alpha) takes values in t1Λ2(I)t^{-1}\Lambda^{2}(I) if and only if κ(I)\kappa(I) lifts to a ring 𝒪\mathcal{O}. We wish to prove that β\beta itself takes values in t1Λ2(I)t^{-1}\Lambda^{2}(I) if and only if 𝒪\mathcal{O} is 33-traced. Note that both conditions are local at the primes dividing 22 and 33, so we may assume that 𝒪K\mathcal{O}_{K} is a DVR. With respect to a basis (ξ,η)(\xi,\eta) of II and a generator of t1Λ2(I)t^{-1}\Lambda^{2}(I), the index form of 𝒪\mathcal{O} has the form

f(x,y)=ax3+bx2y+cxy2+dy3,a,,d𝒪K.f(x,y)=ax^{3}+bx^{2}y+cxy^{2}+dy^{3},\quad a,\ldots,d\in\mathcal{O}_{K}.

If this is the diagonal restriction of β\beta, then β\beta itself can be represented as a 33-dimensional matrix

b/3c/3ab/3c/3db/3c/3
,
\begin{minipage}{43.36464pt} \lx@xy@svg{\hbox{\raise 0.0pt\hbox{\kern 12.36801pt\hbox{\ignorespaces\ignorespaces\ignorespaces\hbox{\vtop{\kern 0.0pt\offinterlineskip\halign{\entry@#!@&&\entry@@#!@\cr&&&\\&&\\&&&\\&&\crcr}}}\ignorespaces{\hbox{\kern-3.0pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{}$}}}}}}}{\hbox{\kern 36.38596pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{b/3\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces{}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces{}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}{\hbox{\kern 94.52588pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{}$}}}}}}}{\hbox{\kern 133.91183pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{c/3\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces{}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}{\hbox{\kern-5.64294pt\raise-40.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{a\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces{}\ignorespaces\ignorespaces{\hbox{\lx@xy@drawline@}}\ignorespaces{}\ignorespaces\ignorespaces{\hbox{\lx@xy@drawline@}}\ignorespaces{\hbox{\lx@xy@drawline@}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces{}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces{}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}{\hbox{\kern 45.75397pt\raise-40.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{}$}}}}}}}{\hbox{\kern 85.15787pt\raise-40.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{b/3\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces{}\ignorespaces\ignorespaces{\hbox{\lx@xy@drawline@}}\ignorespaces{}\ignorespaces\ignorespaces{\hbox{\lx@xy@drawline@}}\ignorespaces{\hbox{\lx@xy@drawline@}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces{}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}{\hbox{\kern-3.0pt\raise-80.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{}$}}}}}}}{\hbox{\kern 36.36801pt\raise-80.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{c/3\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces{}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}{\hbox{\kern 94.52588pt\raise-80.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{}$}}}}}}}{\hbox{\kern 140.69536pt\raise-80.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{d}$}}}}}}}{\hbox{\kern-12.36801pt\raise-120.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{b/3\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces{}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}\ignorespaces\ignorespaces\ignorespaces{}\ignorespaces\ignorespaces{\hbox{\lx@xy@drawline@}}\ignorespaces{}\ignorespaces\ignorespaces{\hbox{\lx@xy@drawline@}}\ignorespaces{\hbox{\lx@xy@drawline@}}{\hbox{\kern 45.75397pt\raise-120.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{}$}}}}}}}{\hbox{\kern 85.13992pt\raise-120.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{c/3\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces{}\ignorespaces\ignorespaces{\hbox{\lx@xy@drawline@}}\ignorespaces{}\ignorespaces\ignorespaces{\hbox{\lx@xy@drawline@}}\ignorespaces{\hbox{\lx@xy@drawline@}}\ignorespaces}}}}\ignorespaces\end{minipage},

which is integral exactly when b,c3𝒪Kb,c\in 3\mathcal{O}_{K}. Since the trace ideal of 𝒪\mathcal{O} is generated by

tr(1)=3,tr(ξ)=b,tr(η)=c\operatorname{tr}(1)=3,\quad\operatorname{tr}(\xi)=-b,\quad\operatorname{tr}(\eta)=c

(by reference to the multiplication table (22)), this is also the condition for 𝒪\mathcal{O} to be 33-traced, establishing the equivalence.

The discriminant relation (26) follows easily from the definition of κ\kappa. ∎

6.5 Quartic rings and their cubic resolvent rings

The basic method for parametrizing quartic orders is by means of cubic resolvent rings, introduced by Bhargava in [B3] and developed by Wood in [WQuartic] and the author in [ORings].

Definition 6.12 ([ORings], Definition 8.1; also a special case of [WQuartic], p. 1069).

Let AA be a Dedekind domain, and let 𝒪\mathcal{O} be a quartic algebra over AA. A resolvent for 𝒪\mathcal{O} (“numerical resolvent” in [ORings]) consists of a rank-22 AA-lattice YY, an AA-module isomorphism Θ:Λ2YΛ3(𝒪/A)\Theta:\Lambda^{2}Y\mathop{\rightarrow}\limits\Lambda^{3}(\mathcal{O}/A), and a quadratic map Φ:𝒪/AY\Phi:\mathcal{O}/A\mathop{\rightarrow}\limits Y such that there is an identity of biquadratic maps

xyxy=Θ(Φ(x)Φ(y))x\wedge y\wedge xy=\Theta(\Phi(x)\wedge\Phi(y)) (27)

from 𝒪×𝒪\mathcal{O}\times\mathcal{O} to Λ3(𝒪/A)\Lambda^{3}(\mathcal{O}/A).

We collect some basic facts about these resolvents.

Theorem 6.13 (the parametrization of quartic rings).

The notion of resolvent for quartic rings has the following properties.

  1. ((a))

    If XX is a rank-33 AA-lattice and Θ:Λ2YΛ3X\Theta:\Lambda^{2}Y\mathop{\rightarrow}\limits\Lambda^{3}X, Φ:XY\Phi:X\mathop{\rightarrow}\limits Y satisfy (27), then there is a unique (up to isomorphism) quartic ring 𝒪\mathcal{O} equipped with an identification 𝒪/AX\mathcal{O}/A\cong X making (𝒪,Y,Θ,Φ)(\mathcal{O},Y,\Theta,\Phi) a resolvent.

  2. ((b))

    There is a canonical (in particular, base-change-respecting) way to associate to a resolvent (𝒪,Y,Θ,Φ)(\mathcal{O},Y,\Theta,\Phi) a cubic ring CC and an identification C/AXC/A\cong X with the following property: For any element x𝒪x\in\mathcal{O} and any lift yCy\in C of the element Φ(x)C/A\Phi(x)\in C/A, we have the equality

    xx2x3=Θ(yy2).x\wedge x^{2}\wedge x^{3}=\Theta(y\wedge y^{2}).

    It satisfies

    DiscC=Disc𝒪.\operatorname{Disc}C=\operatorname{Disc}\mathcal{O}.

    (Here the discriminants are to be seen as quadratic resolvent rings, as in [ORings]; this implies the corresponding identity of discriminant ideals.) If 𝒪\mathcal{O} is nondegenerate, then CC is unique.

  3. ((c))

    Any quartic ring 𝒪\mathcal{O} has at least one resolvent.

  4. ((d))

    If 𝒪\mathcal{O} is maximal, the resolvent is unique (but need not be maximal).

  5. ((e))

    The number of resolvents of 𝒪\mathcal{O} is the sum of the absolute norms of the divisors of the content of 𝒪\mathcal{O}, the smallest ideal 𝔠\mathfrak{c} such that 𝒪=𝒪K+𝔠𝒪\mathcal{O}=\mathcal{O}_{K}+\mathfrak{c}\mathcal{O}^{\prime} for some order 𝒪\mathcal{O}^{\prime}.

  6. ((f))

    Let (Y,Θ,Φ)(Y,\Theta,\Phi) be a resolvent of 𝒪\mathcal{O} with associated cubic ring CC, and let K=FracAK=\operatorname{Frac}A. If the corresponding quartic KK-algebra L=KA𝒪L=K\otimes_{A}\mathcal{O} is étale, then the cubic KK-algebra R=CA𝒪R=C\otimes_{A}\mathcal{O} is none other than the cubic resolvent of LL, as defined in Example 4.7. The maps Θ\Theta and Φ\Phi are the restrictions, under the Minkowski embedding, of the unique resolvent of K¯4\bar{K}^{4}, which is K¯3\bar{K}^{3} with the maps

    Θ:Λ2(K¯3/K¯)\displaystyle\Theta:\Lambda^{2}(\bar{K}^{3}/\bar{K}) Λ3(K¯4/K¯)\displaystyle\mathop{\rightarrow}\limits\Lambda^{3}(\bar{K}^{4}/\bar{K}) (28)
    (0;1;0)(0;0;1)\displaystyle(0;1;0)\wedge(0;0;1) (0;1;0;0)(0;0;1;0)(0;0;0;1)\displaystyle\mapsto(0;1;0;0)\wedge(0;0;1;0)\wedge(0;0;0;1)

    and

    Φ:K¯4/K¯\displaystyle\Phi:\bar{K}^{4}/\bar{K} K¯3/K¯\displaystyle\mathop{\rightarrow}\limits\bar{K}^{3}/\bar{K} (29)
    (x;y;z;w)\displaystyle(x;y;z;w) (xy+zw;xz+yw;xw+yz).\displaystyle\mapsto(xy+zw;xz+yw;xw+yz).
  7. ((g))

    Conversely, let LL be a quartic étale algebra over KK and RR its cubic resolvent. Let

    ΘK:Λ2R/KΛ3L/K,ΦK:L/KR/K\Theta_{K}:\Lambda^{2}R/K\mathop{\rightarrow}\limits\Lambda^{3}L/K,\quad\Phi_{K}:L/K\mathop{\rightarrow}\limits R/K

    be the resolvent data of LL as a (maximal) quartic ring over KK. Suppose 𝒪¯L/K\bar{\mathcal{O}}\subseteq L/K, C¯R/K\bar{C}\subseteq R/K are lattices such that

    • ΦK\Phi_{K} sends 𝒪¯\bar{\mathcal{O}} into C¯\bar{C},

    • ΘK\Theta_{K} maps Λ3𝒪¯\Lambda^{3}\bar{\mathcal{O}} isomorphically onto Λ2C¯\Lambda^{2}\bar{C}.

    Then there are unique quartic rings 𝒪L\mathcal{O}\subseteq L, CRC\subseteq R such that, under the natural identifications, 𝒪/A=𝒪¯\mathcal{O}/A=\bar{\mathcal{O}}, C/A=C¯C/A=\bar{C}, and C¯\bar{C} is a resolvent with the restrictions of ΘK\Theta_{K} and ΦK\Phi_{K}.

Proof.
  1. ((a))

    See [ORings], Theorem 8.3.

  2. ((b))

    See [ORings], Theorems 8.7 and 8.8.

  3. ((c))

    See [ORings], Corollary 8.6.

  4. ((d))

    This is a special case of the following part.

  5. ((e))

    See [ORings], Corollary 8.5.

  6. ((f))

    By base-changing to KK, we see that YAK=R/KY\otimes_{A}K=R/K is a resolvent for LL. Since the resolvent is unique, it suffices to show that the cubic resolvent RR^{\prime} from Example 4.7 is a resolvent for LL also. The maps Θ\Theta and Φ\Phi defined in the theorem statement are seen, by symmetry, to restrict to maps of the appropriate KK-modules. The verification of (27) and of the fact that the multiplicative structure on RR^{\prime} is the right one can be checked at the level of K¯\bar{K}-algebras.

  7. ((g))

    Letting X=𝒪¯X=\bar{\mathcal{O}}, Y=C¯Y=\bar{C} in part a, we construct the desired 𝒪\mathcal{O} and CC. By comparison to the situation under base-change to KK, we see that 𝒪\mathcal{O}, CC naturally inject into LL, RR respectively. Uniqueness is obvious, as 𝒪\mathcal{O} must lie in the integral closure 𝒪L\mathcal{O}_{L}.

In this paper we only deal with nondegenerate rings, that is, those of nonzero discriminant, or equivalently, those that lie in an étale KK-algebra. Consequently, all resolvent maps Θ\Theta, Φ\Phi that we will see are restrictions of (28) and (29). When quartic algebras are parametrized Kummer-theoretically, the resolvent map becomes very explicit and simple:

Proposition 6.14 (explicit Kummer theory for quartic algebras).

Let RR be a cubic étale algebra over KK (charK2\operatorname{char}K\neq 2), and let

L=K+κ(R)L=K+\kappa(R)

be the quartic algebra of resolvent RR corresponding to an element δK×\delta\in K^{\times} of norm 11 in Theorem 5.4c, where

κ(ξ)=(trK¯3/Kξωδ)ω(K¯3)N=1[2]K¯4\kappa(\xi)=\left(\operatorname{tr}_{\bar{K}^{3}/K}\xi\omega\sqrt{\delta}\right)_{\omega\in\left(\bar{K}^{3}\right)^{N=1}[2]}\in\bar{K}^{4}

so κ\kappa maps RR bijectively onto the traceless hyperplane in LL. Then the resolvent of LL is given explicitly by

Θ:Λ3(R)\displaystyle\Theta:\Lambda^{3}(R) Λ3(L/K)\displaystyle\mathop{\rightarrow}\limits\Lambda^{3}(L/K) (30)
αβγ\displaystyle\alpha\wedge\beta\wedge\gamma 116N(δ)κ(α)κ(β)κ(γ)\displaystyle\mapsto\frac{1}{16\sqrt{N(\delta)}}\cdot\kappa(\alpha)\wedge\kappa(\beta)\wedge\kappa(\gamma)
Φ:L/K\displaystyle\Phi:L/K R/K\displaystyle\mathop{\rightarrow}\limits R/K (31)
κ(ξ)\displaystyle\kappa(\xi) 4δξ2\displaystyle\mapsto 4\delta\xi^{2}
Proof.

Since the resolvent is unique (over a field, any étale extension has content 11), it suffices to prove that (30) and (31) define a resolvent. This can be done after extension to K¯\bar{K}, and then it is enough to prove that (30) and (31) agree with the standard resolvent on K¯4\bar{K}^{4}, given in Theorem 6.13f.

As to (30), since both sides are alternating in α\alpha, β\beta, and γ\gamma, it suffices to prove it in the case that

α=(1;0;0),β=(0;1;0)γ=(0;0;1)\alpha=(1;0;0),\quad\beta=(0;1;0)\quad\gamma=(0;0;1)

form the standard basis of R=K¯3R=\bar{K}^{3}. Let δ=(δ(1),δ(2),δ(3))\delta=(\delta^{(1)},\delta^{(2)},\delta^{(3)}). Then

κ(α)\displaystyle\kappa(\alpha) =(δ(1),δ(1),δ(1),δ(1))\displaystyle=\left(\sqrt{\delta^{(1)}},\sqrt{\delta^{(1)}},-\sqrt{\delta^{(1)}},-\sqrt{\delta^{(1)}}\right)
κ(β)\displaystyle\kappa(\beta) =(δ(2),δ(2),δ(2),δ(2))\displaystyle=\left(\sqrt{\delta^{(2)}},-\sqrt{\delta^{(2)}},\sqrt{\delta^{(2)}},-\sqrt{\delta^{(2)}}\right)
κ(γ)\displaystyle\kappa(\gamma) =(δ(3),δ(3),δ(3),δ(3))\displaystyle=\left(\sqrt{\delta^{(3)}},-\sqrt{\delta^{(3)}},-\sqrt{\delta^{(3)}},\sqrt{\delta^{(3)}}\right)

and hence the wedge product of these differs from the standard generator of Λ3(L/K)\Lambda^{3}(L/K) by a factor of

|1111δ(1)δ(1)δ(1)δ(1)δ(2)δ(2)δ(2)δ(2)δ(3)δ(3)δ(3)δ(3)|\displaystyle\begin{vmatrix}1&1&1&1\\ \sqrt{\delta^{(1)}}&\sqrt{\delta^{(1)}}&-\sqrt{\delta^{(1)}}&-\sqrt{\delta^{(1)}}\\ \sqrt{\delta^{(2)}}&-\sqrt{\delta^{(2)}}&\sqrt{\delta^{(2)}}&-\sqrt{\delta^{(2)}}\\ \sqrt{\delta^{(3)}}&-\sqrt{\delta^{(3)}}&-\sqrt{\delta^{(3)}}&\sqrt{\delta^{(3)}}\end{vmatrix}
=N(δ)|1111111111111111|\displaystyle=\sqrt{N(\delta)}\begin{vmatrix}1&1&1&1\\ 1&1&-1&-1\\ 1&-1&1&-1\\ 1&-1&-1&1\end{vmatrix}
=16N(δ).\displaystyle=16\cdot\sqrt{N(\delta)}.

The calculation for (31) is even more routine. ∎

Remark 6.15.

The datum Θ\Theta of a resolvent carries no information, in the following sense. It is unique up to scaling by cA×c\in A^{\times}, and the resolvent data (X,Y,Θ,Φ)(X,Y,\Theta,\Phi) and (X,Y,cΘ,Φ)(X,Y,c\Theta,\Phi) are isomorphic under multiplication by c1c^{-1} on XX and by c2c^{-2} on YY. If AA is a PID, indeed, neither XX nor YY carries any information, and the entire data of the resolvent is encapsulated in Φ\Phi, a pair of 3×33\times 3 symmetric matrices over AA (with formal factors of 1/21/2 off the diagonal) defined up to the natural action of GL3A×GL2A\mathrm{GL}_{3}A\times\mathrm{GL}_{2}A. This establishes the close kinship with Bhargava’s parametrization of quartic rings in [B3]. However, it is useful to keep Θ\Theta around.

6.5.1 Traced resolvents

Just as we found it natural to study not just binary cubic 11111111-forms, but also 13311331-forms and their analogue for each divisor of the ideal (3)(3), so too we study not just quartic rings in general but those satisfying a natural condition at the primes dividing 22.

Definition 6.16.

Let AA be a Dedekind domain, charA2\operatorname{char}A\neq 2, and let 𝔱\mathfrak{t} be an ideal dividing (2)(2) in AA. A resolvent (𝒪,Y,Θ,Φ)(\mathcal{O},Y,\Theta,\Phi) over AA is called 𝔱\mathfrak{t}-traced if, for all xx and yy in AA, the associated bilinear form

Φ(x,y)=Φ(x+y)Φ(x)Φ(y)2\Phi(x,y)=\frac{\Phi(x+y)-\Phi(x)-\Phi(y)}{2}

whose diagonal restriction is Φ(x,x)=Φ(x)\Phi(x,x)=\Phi(x) takes values in 21𝔱Y2^{-1}\mathfrak{t}Y. If AA is a PID, this is equivalent to saying that the off-diagonal entries in the matrix representation of Φ\Phi, which a priori live in 12A\frac{1}{2}A, actually belong to 𝔱2A\frac{\mathfrak{t}}{2}A. We say that AA is 𝔱\mathfrak{t}-traced if it admits a 𝔱\mathfrak{t}-traced resolvent.

Here are some facts about traced resolvents:

Proposition 6.17.

Let 𝒪\mathcal{O} be a quartic ring over a Dedekind domain 𝒪K\mathcal{O}_{K}.

  1. (a)(a)

    𝒪\mathcal{O} is 𝔱\mathfrak{t}-traced if and only if

    1. (i)(i)

      𝔱2|trx\mathfrak{t}^{2}|\operatorname{tr}x for all x𝒪x\in\mathcal{O};

    2. (ii)(ii)

      x2A+𝔱𝒪x^{2}\in A+\mathfrak{t}\mathcal{O} for all x𝒪x\in\mathcal{O}.

  2. (b)(b)

    If 𝒪\mathcal{O} is not an order in the trivial algebra K[ε1,ε2,ε3]/(εiεj)i,j=13K[\varepsilon_{1},\varepsilon_{2},\varepsilon_{3}]/(\varepsilon_{i}\varepsilon_{j})_{i,j=1}^{3}, the number of 𝔱\mathfrak{t}-traced resolvents of 𝒪\mathcal{O} is the sum of the absolute norms of the divisors of its 𝔱\mathfrak{t}-traced content, which is the smallest ideal 𝔠\mathfrak{c} such that 𝒪=A+𝔠𝒪\mathcal{O}=A+\mathfrak{c}\mathcal{O}^{\prime} and 𝒪\mathcal{O}^{\prime} is also 𝔱\mathfrak{t}-traced.

  3. (c)(c)

    If (𝒪,Y,Θ,Φ)(\mathcal{O},Y,\Theta,\Phi) is a 𝔱\mathfrak{t}-traced resolvent with associated cubic ring CC, then 𝔱2|ct(C)\mathfrak{t}^{2}|\operatorname{ct}(C), that is, C=A+𝔱2CC=A+\mathfrak{t}^{2}C^{\prime} for some cubic ring CC^{\prime}. We call CC^{\prime} a “reduced resolvent” of the 𝔱\mathfrak{t}-traced ring AA. Also, 𝔱8|discA\mathfrak{t}^{8}|\operatorname{disc}A.

Proof.
  1. (a)(a)

    Since both statements are local at the primes dividing 22, we can assume that 𝒪K\mathcal{O}_{K} is a DVR, and thus that 𝔱=(t)\mathfrak{t}=(t) is principal. With respect to bases (1=ξ0,ξ1,ξ2,ξ3)(1=\xi_{0},\xi_{1},\xi_{2},\xi_{3}) for 𝒪\mathcal{O} and (1=η0,η1,η2)(1=\eta_{0},\eta_{1},\eta_{2}) for a resolvent CC, the structure constants cijkc_{ij}^{k} of the ring 𝒪\mathcal{O}, defined by

    ξiξj=kcijkξk,\xi_{i}\xi_{j}=\sum_{k}c_{ij}^{k}\xi_{k},

    are determined by the entries of the resolvent

    Φ=([aij],[bij])\Phi=\left([a_{ij}],[b_{ij}]\right)

    via the determinants

    λijk=2𝟏ij+𝟏k|aijakbijbk|\lambda^{ij}_{k\ell}=2^{\mathbf{1}_{i\neq j}+\mathbf{1}_{k\neq\ell}}\begin{vmatrix}a_{ij}&a_{k\ell}\\ b_{ij}&b_{k\ell}\end{vmatrix}

    and a set of formulas appearing in Bhargava [B3, equation (21)] and over a Dedekind domain by the author [ORings, equation (12)]:

    ciij\displaystyle c_{ii}^{j} =ελiiik\displaystyle=-\varepsilon\lambda^{ii}_{ik} (32)
    cijk\displaystyle c_{ij}^{k} =ελjjii\displaystyle=\varepsilon\lambda^{jj}_{ii}
    cijjcikk\displaystyle c_{ij}^{j}-c_{ik}^{k} =ελjkii\displaystyle=\varepsilon\lambda^{jk}_{ii}
    ciiicijjcikk\displaystyle c_{ii}^{i}-c_{ij}^{j}-c_{ik}^{k} =ελijik,\displaystyle=\varepsilon\lambda^{ij}_{ik},

    where (i,j,k)(i,j,k) denotes any permutation of (1,2,3)(1,2,3) and ε=±1\varepsilon=\pm 1 its sign. (Here the nonappearance of some of the individual cijkc_{ij}^{k} on the left-hand side of (32) stems from the ambiguity of translating each ξi\xi_{i} by 𝒪K\mathcal{O}_{K}, which does not change the matrix of Φ\Phi.)

    Assume first that Φ:𝒪/𝒪KC/𝒪K\Phi:\mathcal{O}/\mathcal{O}_{K}\mathop{\rightarrow}\limits C/\mathcal{O}_{K} is 𝔱\mathfrak{t}-traced. Then

    λijk𝔱𝟏ij+𝟏k.\lambda^{ij}_{k\ell}\in\mathfrak{t}^{\mathbf{1}_{i\neq j}+\mathbf{1}_{k\neq\ell}}. (33)

    We then prove that the conditions ai and aii must hold:

    1. (i)(i)

      The trace

      tr(ξ1)\displaystyle\operatorname{tr}(\xi_{1}) =c111+c122+c133\displaystyle=c_{11}^{1}+c_{12}^{2}+c_{13}^{3}
      =λ1213+2λ2311+4c133\displaystyle=\lambda^{12}_{13}+2\lambda^{23}_{11}+4c_{13}^{3}
      0mod𝔱2,\displaystyle\equiv 0\mod\mathfrak{t}^{2},

      and likewise tr(ξ2),tr(ξ3)𝔱2\operatorname{tr}(\xi_{2}),\operatorname{tr}(\xi_{3})\in\mathfrak{t}^{2}.

    2. (ii)(ii)

      The coefficients c11ic_{11}^{i} of ξ12{\xi_{1}}^{2} satisfy:

      c112\displaystyle c_{11}^{2} =λ1311𝔱\displaystyle=\lambda_{13}^{11}\in\mathfrak{t}

      and likewise for c113c_{11}^{3}; and then c111𝔱c_{11}^{1}\in\mathfrak{t} also, since the trace c111+c122+c133=tr(ξ1)𝔱2𝔱c_{11}^{1}+c_{12}^{2}+c_{13}^{3}=\operatorname{tr}(\xi_{1})\in\mathfrak{t}^{2}\subseteq\mathfrak{t}. So the desired relation ξ2𝒪K+𝔱𝒪\xi^{2}\in\mathcal{O}_{K}+\mathfrak{t}\mathcal{O} holds when ξ=ξ1\xi=\xi_{1}, indeed ξ=a1ξ1\xi=a_{1}\xi_{1} for any a1𝒪Ka_{1}\in\mathcal{O}_{K}. The same proof works for ξ=a2ξ2\xi=a_{2}\xi_{2} or ξ=a3ξ3\xi=a_{3}\xi_{3}. Since the case ξ=a0𝒪K\xi=a_{0}\in\mathcal{O}_{K} is trivial and squaring is a \mathbb{Z}-linear operation modulo 22, we get the result for all ξ𝒪\xi\in\mathcal{O}.

    Conversely, suppose that ai and aii hold. We first establish (33). We have

    • λ1113=c112𝔱\lambda^{11}_{13}=c_{11}^{2}\in\mathfrak{t}

    • λ2311=c122c133=trξ1c1112c133𝔱\lambda^{23}_{11}=c_{12}^{2}-c_{13}^{3}=\operatorname{tr}\xi_{1}-c_{11}^{1}-2c_{13}^{3}\in\mathfrak{t}

    • λ1213=c111c122c133=trξ12λ2311+4c133𝔱2\lambda^{12}_{13}=c_{11}^{1}-c_{12}^{2}-c_{13}^{3}=\operatorname{tr}\xi_{1}-2\lambda^{23}_{11}+4c_{13}^{3}\in\mathfrak{t}^{2}.

    Permuting the indices as needed, this accounts for all the λijk\lambda^{ij}_{k\ell} about which (33) makes a nontrivial assertion.

    Now we work from the λijk\lambda^{ij}_{k\ell} back to the resolvent (𝒜,)(\mathcal{A},\mathcal{B}). We may assume that CC is nontrivial (the trivial rings, one for each Steinitz class, are plainly 22-traced with (𝒜,)=(0,0)(\mathcal{A},\mathcal{B})=(0,0).) Then, in the proof of [ORings], Theorem 8.4, the author established that there are vectors μij\mu_{ij} in a two-dimensional vector space VV over KK, unique up to GL2(V)\mathrm{GL}_{2}(V), such that

    μijμk=λijkω\mu_{ij}\wedge\mu_{k\ell}=\lambda^{ij}_{k\ell}\cdot\omega

    for some fixed generator ωΛ2V\omega\in\Lambda^{2}V. (The proof uses the Plücker relations, which are a consequence of the associative law on 𝒪\mathcal{O}.) This VV is none other than R/KR/K, the resolvent module of the quartic algebra L=𝒪𝒪KKL=\mathcal{O}\otimes_{\mathcal{O}_{K}}K, which admits the unique resolvent

    Φ(a1ξ1+a2ξ2+a3ξ3)\displaystyle\Phi(a_{1}\xi_{1}+a_{2}\xi_{2}+a_{3}\xi_{3}) =i<jaiajμij\displaystyle=\sum_{i<j}a_{i}a_{j}\mu_{ij} (34)
    Θ(ξ1ξ2ξ3)\displaystyle\Theta(\xi_{1}\wedge\xi_{2}\wedge\xi_{3}) =ω.\displaystyle=\omega.

    The resolvents of 𝒪\mathcal{O} were found to be exactly the lattices MM containing the span M0M_{0} of the six μij\mu_{ij}, with the correct index

    [M:M0]=𝔠=(λijk)i,j,k,=(ciij,cijk,cijjcikk,ciiicijjcikk:ijki),[M:M_{0}]=\mathfrak{c}=\left(\lambda^{ij}_{k\ell}\right)_{i,j,k,\ell}=\left(c_{ii}^{j},c_{ij}^{k},c_{ij}^{j}-c_{ik}^{k},c_{ii}^{i}-c_{ij}^{j}-c_{ik}^{k}:i\neq j\neq k\neq i\right),

    the content ideal of 𝒪\mathcal{O}. By inspection of (34) that MM is 𝔱\mathfrak{t}-traced if and only if it actually contains the span M~0\tilde{M}_{0} of the six vectors

    μ~ij=t𝟏ijμij.\tilde{\mu}_{ij}=t^{-\mathbf{1}_{i\neq j}}\mu_{ij}.

    Condition (33) is interpreted as saying that the μ~ijμ~k\tilde{\mu}_{ij}\wedge\tilde{\mu}_{k\ell} are still integer multiples of ω\omega. Then the 𝔱\mathfrak{t}-traced resolvents are the lattices MM~0M\supseteq\tilde{M}_{0}. The needed index

    𝔠~=[M:M~0]=(λ~ijk)i,j,k,\tilde{\mathfrak{c}}=[M:\tilde{M}_{0}]=\left(\tilde{\lambda}^{ij}_{k\ell}\right)_{i,j,k,\ell}

    is an integral ideal, so such MM exists, finishing the proof of a.

  2. (b)(b)

    It suffices to prove that 𝔠~\tilde{\mathfrak{c}} is the 𝔱\mathfrak{t}-traced content of 𝒪\mathcal{O}. To see this, note that if 𝒪=𝒪K+a𝒪\mathcal{O}=\mathcal{O}_{K}+a\mathcal{O}^{\prime} has content divisible by aa, then the structure coefficients cijkc_{ij}^{k} of 𝒪\mathcal{O}^{\prime} are obtained from those of 𝒪\mathcal{O} by dividing by aa. This means that the λijk\lambda^{ij}_{k\ell} and λ~ijk\tilde{\lambda}^{ij}_{k\ell} are divided by aa, and so remain integral (indicating that 𝒪\mathcal{O}^{\prime} is also 𝔱\mathfrak{t}-traced) exactly when a𝔠~a\mid\tilde{\mathfrak{c}}.

  3. (c)(c)

    We can again reduce to the case that 𝒪K\mathcal{O}_{K} is a DVR so 𝒪\mathcal{O} has an 𝒪K\mathcal{O}_{K}-basis. Recall that the index form of the resolvent CC is given by

    f(x,y)=4det(𝒜x+y)f(x,y)=4\det(\mathcal{A}x+\mathcal{B}y)

    ([B3], Proposition 11; [ORings], Theorem 8.7). If 𝒜\mathcal{A} and \mathcal{B} have off-diagonal entries in 21𝔱2^{-1}\mathfrak{t}, it immediately follows that ff is divisible by t2t^{2}, so t2ct(C)t^{2}\mid\operatorname{ct}(C). Consequently disc𝒪=discC\operatorname{disc}\mathcal{O}=\operatorname{disc}C, being quartic in the coefficients of ff, is divisible by t8t^{8}. ∎

Similar to Theorem 6.11, we have the following relation between 22-traced quartic rings and self-balanced ideals:

Theorem 6.18 (self-balanced ideals in the quartic setting).

Let 𝒪K\mathcal{O}_{K} be a Dedekind domain, charK2\operatorname{char}K\neq 2, and let RR be a cubic étale extension. A self-balanced triple in RR is a triple (C,I,δ)(C,I,\delta) consisting of a cubic order CRC\subseteq R, a fractional ideal II of CC, and a scalar δ(KC)×\delta\in(KC)^{\times} satisfying the conditions

δI2C,N(I)=(t) is principal,andN(δ)t2=1,\delta I^{2}\subseteq C,\quad N(I)=(t)\text{ is principal},\quad\text{and}\quad N(\delta)t^{2}=1, (35)

Fix an order CRC\subseteq R and a scalar δR×\delta\in R^{\times} with N(δ)N(\delta) a square t2t^{-2}. Then the mapping

I𝒪=𝒪K+κ(I)I\mapsto\mathcal{O}=\mathcal{O}_{K}+\kappa(I) (36)

defines a bijection between

  • self-balanced triples of the form (C,I,δ)(C,I,\delta), and

  • subrings 𝒪L\mathcal{O}\subseteq L of the quartic algebra L=K+κ(R)L=K+\kappa(R) corresponding to the Kummer element δ\delta, such that 𝒪\mathcal{O} is 22-traced with reduced resolvent CC.

Proof.

The proof is very similar to that of 6.11, so we simply summarize the main points. The linear isomorphism κ\kappa establishes a bijection between lattices IRI\subseteq R and κ(I)L/K\kappa(I)\subseteq L/K. We wish to prove that (C,I,δ)(C,I,\delta) is balanced if and only if κ(I)\kappa(I) is the projection of a 22-traced order with reduced resolvent CC.

First note that either of these conditions uniquely specifies

[C:I]=(t),[C:I]=(t),

the former by the balancing condition N(I)=(t)N(I)=(t), and the latter by the Θ\Theta-condition that 𝒪\mathcal{O} have discriminant 256discC256\operatorname{disc}C.

Once again, it is difficult to proceed directly, and we instead prove that both conditions are equivalent to the bilinear map

Φ:I×I\displaystyle\Phi:I\times I R/K\displaystyle\mathop{\rightarrow}\limits R/K
(α1,α2)\displaystyle(\alpha_{1},\alpha_{2}) 4δα1α2\displaystyle\mathop{\rightarrow}\limits 4\delta\alpha_{1}\alpha_{2}

taking values in 4C/K4C/K. ∎

On the self-balanced ideals side, this follows from the parametrization of balanced pairs of ideals by 2×3×32\times 3\times 3 boxes performed over \mathbb{Z} by Bhargava [B2, Theorem 2] and over a general base by Wood [W2xnxn, Theorem 1.4].

On the quartic rings side, the diagonal restriction of Φ\Phi is precisely the resolvent of κ(I)\kappa(I), by Proposition 6.14. That Φ(α,α)4C/K\Phi(\alpha,\alpha)\in 4C/K for each αI\alpha\in I expresses the one condition remaining for κ(I)\kappa(I) to lift (by Theorem 6.13g) to a quartic ring 𝒪\mathcal{O} with resolvent 𝒪K+4C\mathcal{O}_{K}+4C. Then, by definition, this resolvent is 22-traced exactly when Φ\Phi itself has image in 4C/K4C/K.

7 Cohomology of cyclic modules over a local field

Let MM be a Galois module with underlying group 𝒞p\mathcal{C}_{p} over a local field KpK\supseteq\mathbb{Q}_{p} (that is, a wild local field of characteristic 0). Denote by TT and TT^{\prime}, respectively, the (/p)×(\mathbb{Z}/p\mathbb{Z})^{\times}-torsors corresponding to the action of GKG_{K} on M\{0}M\backslash\{0\} and on

Surj(M,μp)=M{0},\operatorname{Surj}(M,\mu_{p})=M^{\prime}\setminus\{0\},

and denote by τc:TT\tau_{c}:T^{\prime}\mathop{\rightarrow}\limits T^{\prime} the torsor operation corresponding to c(/p)×AutMc\in(\mathbb{Z}/p\mathbb{Z})^{\times}\cong\operatorname{Aut}M. By Theorem 5.4, Kummer theory gives an isomorphism

H1(K,M){αT×/(T×)p:τc(α)=αcc(/p)×}.H^{1}(K,M)\cong\left\{\alpha\in T^{\prime\times}/(T^{\prime\times})^{p}:\tau_{c}(\alpha)=\alpha^{c}\,\forall c\in(\mathbb{Z}/p\mathbb{Z})^{\times}\right\}. (37)

Our objective in this section is to understand the group on the right: that is, to describe a basis of it (a generalization of the well-known Shafarevich basis for T×T^{\prime\times}) and understand how the Tate pairing respects it. Much of our work parallels that of Del Corso and Dvornicich [DCD] and Nguyen-Quang-Do [Nguyen].

If σH1(K,M)\sigma\in H^{1}(K,M), we let L=LσL=L_{\sigma} be the 𝒢𝒜(𝒞p)\mathcal{GA}(\mathcal{C}_{p})-extension of degree pp coming from the affine action of GKG_{K} on MM, while we let E=EσE=E_{\sigma} be the associated 𝒢𝒜(𝒞p)\mathcal{GA}(\mathcal{C}_{p})-torsor. Owing to the semidirect product structure of 𝒢𝒜(𝒞p)\mathcal{GA}(\mathcal{C}_{p}), we get a natural decomposition

ELKT.E\cong L\otimes_{K}T.

Using the division algorithm in \mathbb{Z}, we let =(L)=(σ)\ell=\ell(L)=\ell(\sigma) and θ=θ(L)=θ(σ)\theta=\theta(L)=\theta(\sigma) the integers such that

vK(discL)=p(e)+θ,1θp2.v_{K}(\operatorname{disc}L)=p(e-\ell)+\theta,\quad-1\leq\theta\leq p-2.

We call \ell the level, and θ\theta the offset, of the 𝒢𝒜(𝒞p)\mathcal{GA}(\mathcal{C}_{p})-extension LL or of the coclass σ\sigma. Although these definitions appear strange, they allow us to state concisely the following theorem, which will be the main theorem of this section.

Theorem 7.1 (levels and offsets).

Let MM be a Galois module with underlying group 𝒞p\mathcal{C}_{p} over a local field KK with charKp\operatorname{char}K\neq p.

  1. ((a))

    The level \ell of a coclass determines its offset θ\theta uniquely in the following way:

    1. ((i))

      If =e\ell=e, then θ=vK(discT)\theta=v_{K}(\operatorname{disc}T).

    2. ((ii))

      If 0<e0\leq\ell<e, then θ0\theta\geq 0 and

      θvK(β)modp1,\theta\equiv\ell-v_{K}(\beta)\mod p-1,

      where β\beta is the Kummer element corresponding to the resolvent μp1\mu_{p-1}-torsor of MM.

    3. ((iii))

      If =1\ell=-1, then θ=1\theta=-1.

  2. ((b))

    For all i0i\geq 0, the level space

    i=i(M)={σH1(K,M):(σ)i}\mathcal{L}_{i}=\mathcal{L}_{i}(M)=\{\sigma\in H^{1}(K,M):\ell(\sigma)\geq i\}

    consisting of coclasses of level at least ii is a subgroup of H1(K,M)H^{1}(K,M).

  3. ((c))

    e=H1ur(K,M)\mathcal{L}_{e}=H^{1}_{\mathrm{ur}}(K,M).

  4. ((d))

    For 0ie0\leq i\leq e,

    |i|=qei|H0(K,M)|.\lvert\mathcal{L}_{i}\rvert=q^{e-i}\lvert H^{0}(K,M)\rvert.
  5. ((e))

    1\mathcal{L}_{-1} is the whole of H1(K,M)H^{1}(K,M), and

    |H1(K,M)|=qe|H0(K,M)||H0(K,M)|.\lvert H^{1}(K,M)\rvert=q^{e}\lvert H^{0}(K,M)\rvert\cdot\lvert H^{0}(K,M^{\prime})\rvert.
  6. ((f))

    For d1d\leq 1, a neighborhood

    {[α]H1(K,M):αT×,|α1|d}\left\{[\alpha]\in H^{1}(K,M):\alpha\in T^{\prime\times},\lvert\alpha-1\rvert\leq d\right\}

    is a level space i\mathcal{L}_{i} whose index ii is given by

    i={vK(β)p+p1plogdlog|πK|1vK(β)p1,ddmine+1,d<dmin,i=\begin{cases}\displaystyle\Biggl{\lceil}\frac{v_{K}(\beta)}{p}+\frac{p-1}{p}\left\lceil\frac{\log d}{\log\lvert\pi_{K}\rvert}-1-\frac{v_{K}(\beta)}{p-1}\right\rceil\Biggr{\rceil},&d\geq d_{\min}\\ e+1,&d<d_{\min},\end{cases}

    where

    dmin=|p|p/(p1).d_{\min}=\lvert p\rvert^{p/(p-1)}.
  7. ((g))

    For 0ie0\leq i\leq e, with respect to the Tate pairing between H1(K,M)H^{1}(K,M) and H1(K,M)H^{1}(K,M^{\prime}),

    i(M)=ei(M).\mathcal{L}_{i}(M)^{\perp}=\mathcal{L}_{e-i}(M^{\prime}).

One corollary is sufficiently important that we state it before starting the proof:

Corollary 7.2.

For 0ie0\leq i\leq e, the characteristic function LiL_{i} of the level space i\mathcal{L}_{i} has Fourier transform given by

Li^=qeiLei.\widehat{{L_{i}}}=q^{e-i}{L_{e-i}}. (38)

where q=|kK|q=\lvert k_{K}\rvert.

Proof.

Immediate from Theorem 7.1, parts d and g. ∎

7.1 Discriminants of Kummer and affine extensions

The starting point for our investigation of discriminants is as follows:

Theorem 7.3.

Let KK be a local field with μpK\mu_{p}\subseteq K, and let uK×u\in K^{\times} be a minimal representative of a class in K×/(K×)pK^{\times}/(K^{\times})^{p}. The discriminant ideal of the associated Kummer extension L=K[up]L=K[\sqrt[p]{u}] is given by

Disc(L/K)={ppπKp1(u1)p1𝒪KvK(u1)<peKp1(1)vK(u1)peKp1.\operatorname{Disc}(L/K)=\begin{cases}\dfrac{p^{p}\cdot{\pi_{K}}^{p-1}}{(u-1)^{p-1}}\mathcal{O}_{K}&v_{K}(u-1)<\dfrac{pe_{K}}{p-1}\\ (1)&v_{K}(u-1)\geq\dfrac{pe_{K}}{p-1}.\end{cases}
Proof.

One can find an explicit basis for 𝒪L\mathcal{O}_{L} and compute the discriminant. For details, see Del Corso and Dvornicich [DCD, Lemmas 5, 6, and 7]. ∎

In this section, we will prove the following generalization:

Theorem 7.4.

Let KK be a local field, and let MM be a GKG_{K}-module with underlying group 𝒞p\mathcal{C}_{p}. Let TT^{\prime} be the (/p)×(\mathbb{Z}/p\mathbb{Z})^{\times}-torsor corresponding to the GKG_{K}-set Surj(M,μm)\operatorname{Surj}(M,\mu_{m}), and let T1T^{\prime}_{1} be the field factor of TT^{\prime}. Let uT×1u\in T^{\prime\times}_{1} be a minimal representative for a class in (T×1)ω(T^{\prime\times}_{1})_{\omega} parametrizing, via Theorem 5.4 a coclass σH1(K,M)\sigma\in H^{1}(K,M), and let LL be the corresponding 𝒢𝒜(𝒞p)\mathcal{GA}(\mathcal{C}_{p})-extension. Then

Disc(L/K)𝒪T1={ppπKp1(u1)p1𝒪T1vT1(u1)<peT1p1Disc(T/K)𝒪T1vT1(u1)peT1p1,\operatorname{Disc}(L/K)\mathcal{O}_{T^{\prime}_{1}}=\begin{cases}\dfrac{p^{p}\cdot{\pi_{K}}^{p-1}}{(u-1)^{p-1}}\mathcal{O}_{T^{\prime}_{1}}&v_{T^{\prime}_{1}}(u-1)<\dfrac{pe_{T^{\prime}_{1}}}{p-1}\\ \operatorname{Disc}(T/K)\cdot\mathcal{O}_{T^{\prime}_{1}}&v_{T^{\prime}_{1}}(u-1)\geq\dfrac{pe_{T^{\prime}_{1}}}{p-1},\end{cases} (39)

where TT is the (/p)×(\mathbb{Z}/p\mathbb{Z})^{\times}-torsor corresponding to MM.

Remark 7.5.

Note that (u1)p1𝒪T1(u-1)^{p-1}\mathcal{O}_{T^{\prime}_{1}} is the extension of an ideal of KK, since eT1/Ke_{T^{\prime}_{1}/K} divides p1p-1.

Proof.

If LL is not a field, then the image of GKG_{K} in 𝒢𝒜(𝒞p)\mathcal{GA}(\mathcal{C}_{p}) lies in a nontransitive subgroup (viewing 𝒢𝒜(𝒞p)\mathcal{GA}(\mathcal{C}_{p}) as embedded in Sym(𝒞p)\operatorname{Sym}(\mathcal{C}_{p})). It is not hard to show that every nontransitive subgroup of 𝒢𝒜(𝒞p)\mathcal{GA}(\mathcal{C}_{p}) has a fixed point. Moving this fixed point to 0, we get that σ=0\sigma=0, u=1u=1, and LK×TL\cong K\times T, in accord with the second case of the formula.

We may now assume that LL is a field. Although the extension L/KL/K need not be Galois, we have

T[μp]KLT[μp,up],T[\mu_{p}]\otimes_{K}L\cong T[\mu_{p},\sqrt[p]{u}],

a Kummer extension of T[μp]T[\mu_{p}]. Let EE be a field factor of T[μp]T[\mu_{p}] containing T1T^{\prime}_{1}. Then

EKLE[up]E\otimes_{K}L\cong E[\sqrt[p]{u}]

as extensions of EE. Note that [E:K]|(p1)2[E:K]|(p-1)^{2}; in particular, [E:K][E:K] is prime to pp. So uu remains a minimal representative in E×E^{\times}, and since EE and LL must be linearly disjoint, EKL=ELE\otimes_{K}L=EL is a field unless u=1u=1. So

Disc(EL/E)={ppπEp1(u1)p1𝒪EvK(u1)<peKp1(1)vK(u1)peKp1.\operatorname{Disc}(EL/E)=\begin{cases}\dfrac{p^{p}\cdot{\pi_{E}}^{p-1}}{(u-1)^{p-1}}\mathcal{O}_{E}&v_{K}(u-1)<\dfrac{pe_{K}}{p-1}\\ (1)&v_{K}(u-1)\geq\dfrac{pe_{K}}{p-1}.\end{cases}

We must now relate Disc(EL/E)\operatorname{Disc}(EL/E) to Disc(L/K)\operatorname{Disc}(L/K). If vK(u1)peK/(p1)v_{K}(u-1)\geq{pe_{K}}/(p-1), then EL/EEL/E is unramified, so peEL/Kp\nmid e_{EL/K} and L/KL/K is unramified as well. In particular, L/KL/K is Galois, so MM is trivial, TT is totally split, and the formula again holds.

We are left with the case that vK(u1)<peKp1v_{K}(u-1)<\dfrac{pe_{K}}{p-1}. Here EL/EEL/E, and hence L/KL/K, are totally ramified. We relate their discriminants by the following trick, which also appears in Del Corso and Dvornicich [DCD]. An 𝒪K\mathcal{O}_{K}-basis for 𝒪L\mathcal{O}_{L} is given by

1,πL,,πLp1.1,\pi_{L},\ldots,\pi_{L}^{p-1}. (40)

The same elements form an 𝒪E\mathcal{O}_{E}-basis for an order 𝒪𝒪EL\mathcal{O}\subseteq\mathcal{O}_{EL}, but their ELEL-valuations are 0,e,,(p1)e0,e^{\prime},\ldots,(p-1)e^{\prime}, where e=eE/Ke^{\prime}=e_{E/K}. Divide each basis element by πE\pi_{E} as many times as possible so that it remains integral. We get a new system of elements

1,πLπEa1,,πLp1πEap1.1,\frac{\pi_{L}}{\pi_{E}^{a_{1}}},\ldots,\frac{\pi_{L}^{p-1}}{\pi_{E}^{a_{p-1}}}. (41)

Since ee^{\prime} is coprime to pp, these elements have ELEL-valuations 0,1,,e10,1,\ldots,e^{\prime}-1 in some order and thus form an 𝒪E\mathcal{O}_{E}-basis for 𝒪EL\mathcal{O}_{EL}. We have

vE([𝒪EL:𝒪])\displaystyle v_{E}([\mathcal{O}_{EL}:\mathcal{O}]) =a1++ap1\displaystyle=a_{1}+\cdots+a_{p-1}
=[e++(p1)e][1++(p1)]p\displaystyle=\frac{[e^{\prime}+\cdots+(p-1)e^{\prime}]-[1+\cdots+(p-1)]}{p}
=(p1)(e1)2,\displaystyle=\frac{(p-1)(e^{\prime}-1)}{2},

and hence

Disc(L/K)\displaystyle\operatorname{Disc}(L/K) =Disc(𝒪/𝒪E)\displaystyle=\operatorname{Disc}(\mathcal{O}/\mathcal{O}_{E})
=Disc(𝒪EL/𝒪E)[𝒪EL:𝒪]2\displaystyle=\operatorname{Disc}(\mathcal{O}_{EL}/\mathcal{O}_{E})\cdot[\mathcal{O}_{EL}:\mathcal{O}]^{2}
=ppπEp1(u1)p1πE(p1)(e1)𝒪E\displaystyle=\dfrac{p^{p}\cdot{\pi_{E}}^{p-1}}{(u-1)^{p-1}}\cdot\pi_{E}^{(p-1)(e^{\prime}-1)}\cdot\mathcal{O}_{E}
=ppπEe(p1)(u1)p1𝒪E\displaystyle=\dfrac{p^{p}\cdot{\pi_{E}}^{e^{\prime}(p-1)}}{(u-1)^{p-1}}\cdot\mathcal{O}_{E}
=ppπK(p1)(u1)p1𝒪E.\displaystyle=\dfrac{p^{p}\cdot{\pi_{K}}^{(p-1)}}{(u-1)^{p-1}}\cdot\mathcal{O}_{E}.\qed
Remark 7.6.

Along the lines of the preceding argument, we can prove the following more general result on discriminants in extensions of coprime degree:

Proposition 7.7.

Let LL and MM be two extensions of a local field KK with gcd([L:K],[M:K])=1\gcd([L:K],[M:K])=1. Then

Disc(L/K)=Disc(LM/M)(πMπK)fL/K(eL/K1).\operatorname{Disc}(L/K)=\operatorname{Disc}(LM/M)\cdot\left(\frac{\pi_{M}}{\pi_{K}}\right)^{f_{L/K}\left(e_{L/K}-1\right)}.

7.2 The Shafarevich basis

We start with the following exposition of the Shafarevich basis theorem. Although this theorem has appeared many times in the literature (see Del Corso and Dvornicich, [DCD], Proposition 6), we include a proof here by a method that will establish some important corollaries for us.

Filter U=K×U=K^{\times} by the subgroups

Ui={x𝒪K×:x1modπi},U_{i}=\{x\in\mathcal{O}_{K}^{\times}:x\equiv 1\mod\pi^{i}\},

and let U¯i\bar{U}_{i} be the projection of UiU_{i} onto U¯:=K×/(K×)p\bar{U}:=K^{\times}/(K^{\times})^{p}. Note that U¯i=0\bar{U}_{i}=0 for i>peK/(p1)i>pe_{K}/(p-1), as the Taylor series for xp\sqrt[p]{x} about 11 converges for x1x\equiv 1 mod πpeK/(p1)+1\pi^{\left\lfloor pe_{K}/(p-1)\right\rfloor+1}. So

U¯U¯/U¯00ipeT1p1U¯i/U¯i+1\bar{U}\cong\bar{U}/\bar{U}_{0}\oplus\bigoplus_{0\leq i\leq\frac{pe_{T^{\prime}_{1}}}{p-1}}\bar{U}_{i}/\bar{U}_{i+1} (42)

as 𝔽p\mathbb{F}_{p}-vector spaces, and we can produce a basis for U¯\bar{U} by lifting a basis for each of the composition factors on the right-hand side.

Proposition 7.8 (the Shafarevich basis theorem).

Let KpK\supseteq\mathbb{Q}_{p} be a local field, and let i0i\geq 0. The structure of U¯i/U¯i+1\bar{U}_{i}/\bar{U}_{i+1} is as follows:

  • If

    0<i<pp1eK,pi,0<i<\frac{p}{p-1}e_{K},\quad p\nmid i,

    then U¯i/U¯i+1Ui/Ui+1\bar{U}_{i}/\bar{U}_{i+1}\cong U_{i}/U_{i+1} has a basis of fKf_{K} units of the form 1+xjπi1+x_{j}\pi^{i}, where xjx_{j} ranges over an 𝔽p\mathbb{F}_{p}-basis of kKk_{K}. We call these generic units.

  • if i=pp1eKi=\frac{p}{p-1}e_{K} and μpK\mu_{p}\subseteq K, then U¯i/U¯i+1\bar{U}_{i}/\bar{U}_{i+1} has dimension 11 and is generated by

    u=1+p(ζp1)a,u=1+p(\zeta_{p}-1)a,

    for any a𝒪Ka\in\mathcal{O}_{K} with ptr𝒪K/p(a)p\nmid\operatorname{tr}_{\mathcal{O}_{K}/\mathbb{Q}_{p}}(a). We call such a generator an intimate unit, and we let

    dmin=|p(ζp1)|=|p|p/(p1),d_{\min}=\lvert p(\zeta_{p}-1)\rvert=\lvert p\rvert^{p/(p-1)},

    the distance of an intimate unit to 11.

  • For all other i,i, we have U¯i/U¯i+1=0\bar{U}_{i}/\bar{U}_{i+1}=0.

Proof.

Note that U¯0/U¯1=0\bar{U}_{0}/\bar{U}_{1}=0 because U0/U1kK×U_{0}/U_{1}\cong k_{K}^{\times} has order prime to pp. To compute U¯i/U¯i+1\bar{U}_{i}/\bar{U}_{i+1}, where i1i\geq 1, we must see how many of the congruence classes 1+xπi1+x\pi^{i} mod πi+1\pi^{i+1} (where xkKx\in k_{K}) contain a ppth power.

Consider a general ppth power upu^{p}, 1u𝒪K×1\neq u\in\mathcal{O}_{K}^{\times}. Write u=1+yπju=1+y\pi^{j}, πy\pi\nmid y. By the binomial theorem,

up1+pyπj+ypπpjmodπ2j+eK,u^{p}\equiv 1+py\pi^{j}+y^{p}\pi^{pj}\mod\pi^{2j+e_{K}},

so

v(up1){=pj,j<eKp1,peKp1,j=eKp1,=j+eK,j>eKp1.v(u^{p}-1)\begin{cases}=pj,&j<\frac{e_{K}}{p-1},\\ \geq\frac{pe_{K}}{p-1},&j=\frac{e_{K}}{p-1},\\ =j+e_{K},&j>\frac{e_{K}}{p-1}.\end{cases}

We now perform the needed analysis in each case:

  • If 0<i<peK/(p1)0<i<pe_{K}/(p-1) and pip\nmid i, then v(up1)v(u^{p}-1) can never attain the value ii, so the map Ui/Ui+1U¯i/U¯i+1U_{i}/U_{i+1}\mathop{\rightarrow}\limits\bar{U}_{i}/\bar{U}_{i+1} is an isomorphism, and we must include an entire basis of ff elements in our basis for U¯\bar{U}.

  • If 0<i<peK/(p1)0<i<pe_{K}/(p-1) and p|ip|i, then the ppth powers

    (1+yπi/p)p1+ypπimodπi+1(1+y\pi^{i/p})^{p}\equiv 1+y^{p}\pi^{i}\mod\pi^{i+1}

    cover all the desired congruence classes, as the map yypy\mapsto y^{p} on kKk_{K} is surjective; so U¯i/U¯i+1=0\bar{U}_{i}/\bar{U}_{i+1}=0.

  • If i>peK/(p1)i>pe_{K}/(p-1), then the ppth powers of elements of the form 1+xπieK1+x\pi^{i-e_{K}} surject onto the congruence classes, repeating what we knew from the Taylor series.

  • Finally, if i=peK/(p1)i=pe_{K}/(p-1), then we can only use powers up=(1+πj)pu^{p}=(1+\pi^{j})^{p} where j=eK/(p1)j=e_{K}/(p-1). We have

    (1+yπj)1+pyπj+ypπpj1+(ypcy)πimodπi+1,(1+y\pi^{j})\equiv 1+py\pi^{j}+y^{p}\pi^{pj}\equiv 1+(y^{p}-cy)\pi^{i}\mod\pi^{i+1},

    where

    c=pπ(p1)j𝒪K×.c=\frac{-p}{\pi^{(p-1)j}}\in\mathcal{O}_{K}^{\times}.

    So we must analyze the (clearly linear) map c:kKkK\wp_{c}:k_{K}\mathop{\rightarrow}\limits k_{K} given by c(y)=ypcy\wp_{c}(y)=y^{p}-cy.

    If cc is not a (p1)(p-1)st power in kKk_{K} (or in 𝒪K\mathcal{O}_{K}, which amounts to the same thing by Hensel’s lemma), then c\wp_{c} is injective and hence surjective, so U¯i/U¯i+1=0\bar{U}_{i}/\bar{U}_{i+1}=0. Also, KK has no nontrivial ppth roots of unity, as u=ζpu=\zeta_{p} would yield a nontrivial element of kerc\ker\wp_{c} (since vp(ζp)=eK/(p1)v_{p}(\zeta_{p})=e_{K}/(p-1)).

    If c=bp1c=b^{p-1} is a (p1)(p-1)st power in 𝒪K\mathcal{O}_{K}, then y=by=b is an element of kerc\ker\wp_{c}. Note that u=1+bπju=1+b\pi^{j} lifts to a nontrivial ppth root of unity ζp\zeta_{p}, since up1u^{p}\equiv 1 mod πi+1\pi^{i+1} has (by the Taylor series again) a ppth root that is 11 mod πj+1\pi^{j+1}. Note that kerc\ker\wp_{c} has dimension only 11, since bb is unique up to μp1=𝔽p×\mu_{p-1}=\mathbb{F}_{p}^{\times}. Consequently cokercU¯i/U¯i+1\operatorname{coker}\wp_{c}\cong\bar{U}_{i}/\bar{U}_{i+1} has dimension exactly 11.

    This does not tell us how to find a generator for U¯i/U¯i+1\bar{U}_{i}/\bar{U}_{i+1}. For this, put y=byy=by^{\prime} so

    (1+byπj)1+bp(ypy)πi1+bp(y)modπi+1,(1+by^{\prime}\pi^{j})\equiv 1+b^{p}(y^{p}-y)\pi^{i}\equiv 1+b^{p}\wp(y)\mod\pi^{i+1},

    where (y)=ypy\wp(y)=y^{p}-y is the usual Artin-Schreyer map. Since ypy^{p} and yy are Galois conjugates over 𝔽p\mathbb{F}_{p}, we have trkK/𝔽p(ypy)=0\operatorname{tr}_{k_{K}/\mathbb{F}_{p}}(y^{p}-y)=0, so if akKa\in k_{K} is an element with nonzero trace to 𝔽p\mathbb{F}_{p}, then U¯i/U¯i+1\bar{U}_{i}/\bar{U}_{i+1} is generated by

    1+abpπi1+abcπi1+ap(ζp1)modπi+1.1+ab^{p}\pi^{i}\equiv 1+abc\pi^{i}\equiv 1+ap(\zeta_{p}-1)\mod\pi^{i+1}.\qed

We draw two corollaries of the above method.

Corollary 7.9.

U¯\bar{U} has dimension

[K:p]+1+𝟏μpK[K:\mathbb{Q}_{p}]+1+\mathbf{1}_{\mu_{p}\subseteq K}

with a basis consisting of πK\pi_{K} and (arbitrary lifts of) the elements in the bases of U¯i/U¯i+1\bar{U}_{i}/\bar{U}_{i+1} in Proposition 7.8.

We call this basis the Shafarevich basis for UU.

Remark 7.10.

The dimension of U¯\bar{U} follows also from the Euler-characteristic computation

|H0(K,μp)||H2(K,μp)||H1(K,μp)|=p[K:p]\frac{\lvert H^{0}(K,\mu_{p})\rvert\lvert H^{2}(K,\mu_{p})\rvert}{\lvert H^{1}(K,\mu_{p})\rvert}=p^{-[K:\mathbb{Q}_{p}]}
Proof.

Clearly U¯/U¯0𝒞p\bar{U}/\bar{U}_{0}\cong\mathcal{C}_{p} is generated by πK\pi_{K}. The result follows from the composition series (42). ∎

Corollary 7.11.

An element xU¯x\in\bar{U} belongs to U¯i\bar{U}_{i} if and only if, in the expansion of xx in the Shafarevich basis, only the basis elements in UiU_{i} appear (to nonzero exponents).

Proof.

Filtering U¯i\bar{U}_{i} by the U¯j\bar{U}_{j}, for jij\geq i, we see that a basis for U¯i\bar{U}_{i} is given by the portion of the Shafarevich basis coming from U¯j/U¯j+1\bar{U}_{j}/\bar{U}_{j+1} for jij\geq i. This is just the basis elements that lie in UiU_{i}. ∎

The following simple result is one I have not seen in the literature before:

Corollary 7.12.

The cyclotomic extension p[μp]\mathbb{Q}_{p}[\mu_{p}] is isomorphic to p[pp1]\mathbb{Q}_{p}[\sqrt[p-1]{-p}] and has Kummer element p-p as a μp1\mu_{p-1}-torsor.

Proof.

Let K=p[pp1]K=\mathbb{Q}_{p}[\sqrt[p-1]{-p}], with uniformizer πK=pp1\pi_{K}=\sqrt[p-1]{-p}. In the notation of the intimate unit case of Proposition 7.8, we have i=pi=p, j=1j=1,

c=pπKp1=1,c=\frac{-p}{\pi_{K}^{p-1}}=1,

and we can take b=cp1=1b=\sqrt[p-1]{c}=1. Accordingly, 1+πK1+\pi_{K} is congruent mod πK2\pi_{K}^{2} to a unique ppth root of unity ζp\zeta_{p}. Note that ζpi1+iπK\zeta_{p}^{i}\equiv 1+i\pi_{K} mod πK2\pi_{K}^{2}. Since μp1\mu_{p-1} acts on πK\pi_{K} by multiplication, it must act on the powers of ζp\zeta_{p} by ω\omega. Hence Kp[μp]K\cong\mathbb{Q}_{p}[\mu_{p}] as μp1\mu_{p-1}-torsors. ∎

In the rest of this section we will study how U¯=U¯(K)\bar{U}=\bar{U}(K) behaves under field extension. We use the following notational conventions:

Elements uK×/(K×)pu\in K^{\times}/(K^{\times})^{p} are classified by their distance, by which we mean the closest distance of a representative from 11:

d(u)=dK(u)=minyK×|uyp1|.d(u)=d_{K}(u)=\min_{y\in K^{\times}}\lvert uy^{p}-1\rvert.

Here the absolute value is the local one on KK. (We could choose a normalization of this absolute value, but we prefer to express d(u)d(u) in terms of an undetermined |πK|\lvert\pi_{K}\rvert and |p|=|πK|eK\lvert p\rvert=\lvert\pi_{K}\rvert^{e_{K}}.) Note that d(u)1d(u)\leq 1, since uu can always be taken to have nonnegative valuation. Also, it is easy to see that d(uv)max{d(u),d(v)}d(uv)\leq\max\{d(u),d(v)\}, so dd defines a norm on K×/(K×)pK^{\times}/(K^{\times})^{p}.

For ease in stating theorems involving distances, we note that an ideal (or even a fractional ideal) 𝔞\mathfrak{a} of a local field KK is uniquely determined by the largest absolute value of its elements, which we denote by |𝔞|\lvert\mathfrak{a}\rvert. We have

𝔞={xK:|x||𝔞|}.\mathfrak{a}=\{x\in K:\lvert x\rvert\leq\lvert\mathfrak{a}\rvert\}.

For any real d>0d>0, let BdB_{\leq d} denote the closed ball of radius dd about 11 in K×/(K×)pK^{\times}/(K^{\times})^{p}:

Bd={uK×/(K×)p:d(u)d}B_{\leq d}=\{u\in K^{\times}/(K^{\times})^{p}:d(u)\leq d\}

and likewise for B<dB_{<d}. It is easy to prove that these are subgroups. If 𝔣𝒪K\mathfrak{f}\subsetneq\mathcal{O}_{K} is an ideal, then B|𝔣|B_{\leq\lvert\mathfrak{f}\rvert} is the projection of 𝒰𝔣\mathcal{U}_{\mathfrak{f}}; but this fails for 𝔣=(1)\mathfrak{f}=(1).

Note the following:

Lemma 7.13.

If L/KL/K is an extension of local fields whose degree nn is prime to pp, then the canonical map from K×/(K×)pK^{\times}/(K^{\times})^{p} to L×/(L×)pL^{\times}/(L^{\times})^{p} is injective and preserves distance: that is, for every uK×/(K×)pu\in K^{\times}/(K^{\times})^{p},

dK(u)=dL(u).d_{K}(u)=d_{L}(u).
Proof.

The injectivity follows from the fact that if uKu\in K, then NL/K=unN_{L/K}=u^{n} and nn is prime to pp.

It is obvious that dL(u)dK(u)d_{L}(u)\leq d_{K}(u), so it suffices to prove the opposite inequality. It’s easy to see that if x𝒪Lx\in\mathcal{O}_{L}, then

|NL/K(x)1||x1|.\lvert N_{L/K}(x)-1\rvert\leq\lvert x-1\rvert.

Let yL×y\in L^{\times} achieve |uyp1|=dL(u)\lvert uy^{p}-1\rvert=d_{L}(u). Then

dL(u)=|uyp1||NL/K(uyp)1|=|u[L:K]NL/K(y)p1|dK(u[L:K]).d_{L}(u)=\lvert uy^{p}-1\rvert\geq\lvert N_{L/K}(uy^{p})-1\rvert=\lvert u^{[L:K]}N_{L/K}(y)^{p}-1\rvert\geq d_{K}\left(u^{[L:K]}\right).

But uu is a power of u[L:K]u^{[L:K]} up to ppth powers, so dK(u)dK(u[L:K])d_{K}(u)\leq d_{K}\left(u^{[L:K]}\right), completing the proof. ∎

Assume KpK\supseteq\mathbb{Q}_{p}. We first parametrize MM itself. By (classical) Kummer theory, we have canonical isomorphisms

Hom(GK,Aut(𝒞p))H1(GK,(/p)×)H1(GK,μp1)K×/(K×)p1\operatorname{Hom}(G_{K},\operatorname{Aut}(\mathcal{C}_{p}))\cong H^{1}(G_{K},(\mathbb{Z}/p\mathbb{Z})^{\times})\cong H^{1}(G_{K},\mu_{p-1})\cong K^{\times}/(K^{\times})^{p-1}

where the isomorphism (/p)×μp1(\mathbb{Z}/p\mathbb{Z})^{\times}\cong\mu_{p-1} is given by Teichmüller lift. (Note that we do not need to pick a generator of (/p)×(\mathbb{Z}/p\mathbb{Z})^{\times} to do this.) We have the following:

Lemma 7.14.

If MM is the Galois module with underlying group 𝒞p\mathcal{C}_{p} corresponding to the Kummer element βK×/(K×)p1\beta\in K^{\times}/(K^{\times})^{p-1}, then the Tate dual MM^{\prime} has Kummer element β=p/β\beta^{\prime}=-p/\beta.

Proof.

When MM is trivial, the result was proved as Corollary 7.12. The lemma then follows by noting that if M1M_{1}, M2M_{2} are cyclic Galois modules of order pp with Kummer elements β1\beta_{1}, β2\beta_{2}, respectively, then Hom(M1,M2)\operatorname{Hom}(M_{1},M_{2}) is also cyclic of order pp and has Kummer element β2/β1\beta_{2}/\beta_{1}. ∎

If TT^{\prime} has rr field factors (all necessarily isomorphic to one T1T^{\prime}_{1}), then, by Corollary 7.9,

dim𝔽pT×/(T×)p={n+2r,μpT1n+rotherwise.\dim_{\mathbb{F}_{p}}T^{\prime\times}/(T^{\prime\times})^{p}=\begin{cases}n+2r,&\mu_{p}\subseteq T^{\prime}_{1}\\ n+r&\text{otherwise}.\end{cases}

The group T×/(T×)pT^{\prime\times}/(T^{\prime\times})^{p} is a representation of (/p)×(\mathbb{Z}/p\mathbb{Z})^{\times} over the field 𝔽p\mathbb{F}_{p}. Since 𝔽p\mathbb{F}_{p} has the (p1)(p-1)st roots of unity, such a representation splits as a direct sum of 11-dimensional representations; there are p1p-1 of these, and they are the powers of the standard representation

ω:(/p)×GL1(𝔽p)\omega:(\mathbb{Z}/p\mathbb{Z})^{\times}\mathop{\rightarrow}\limits\mathrm{GL}_{1}(\mathbb{F}_{p})

given by the obvious isomorphism. By Theorem 5.4, H1(K,M)H^{1}(K,M) is parametrized by the ω\omega-isotypical component of T×/(T×)pT^{\prime\times}/(T^{\prime\times})^{p}, which we denote by T×ωT^{\prime\times}_{\omega} for brevity.

We can reduce the problem from TT^{\prime} to T1T^{\prime}_{1} in the following way:

Lemma 7.15.

Let TT^{\prime} be a μt\mu_{t}-torsor, t|p1t|p-1. Let T1T^{\prime}_{1} be the field factor of TT^{\prime}, and let r=[T1:K]r=[T^{\prime}_{1}:K]. Then:

  1. ((a))

    The subgroup fixing T1T^{\prime}_{1} (as a set) is μrμp1(/p)×\mu_{r}\subseteq\mu_{p-1}\cong(\mathbb{Z}/p\mathbb{Z})^{\times}, and T1T^{\prime}_{1} is a μr\mu_{r}-torsor;

  2. ((b))

    Projection to T1T^{\prime}_{1} defines an isomorphism T×ω(T×1)ωT^{\prime\times}_{\omega}\cong(T^{\prime\times}_{1})_{\omega}, where

    T×ω\displaystyle T^{\prime\times}_{\omega} ={αT×/(T×)p:τc(α)=αccμt}\displaystyle=\left\{\alpha\in T^{\prime\times}/(T^{\prime\times})^{p}:\tau_{c}(\alpha)=\alpha^{c}\,\forall c\in\mu_{t}\right\} (43)
    (T×1)ω\displaystyle(T^{\prime\times}_{1})_{\omega} ={αT×1/(T×1)p:τc(α)=αccμr}.\displaystyle=\left\{\alpha\in T^{\prime\times}_{1}/(T^{\prime\times}_{1})^{p}:\tau_{c}(\alpha)=\alpha^{c}\,\forall c\in\mu_{r}\right\}. (44)
  3. ((c))

    More generally, for any ss with r|s|tr|s|t, the orbit μs(T1)\mu_{s}(T^{\prime}_{1}) consists of s/ns/n field factors whose product TsT^{\prime}_{s} is a μs\mu_{s}-torsor. Projection onto TsT^{\prime}_{s} and then onto T1T^{\prime}_{1} defines isomorphisms

    T×ω(T×s)ω(T×1)ω.T^{\prime\times}_{\omega}\cong(T^{\prime\times}_{s})_{\omega}\cong(T^{\prime\times}_{1})_{\omega}.
Remark 7.16.

A result with much the same content, but in a slightly different setting, is proved by Del Corso and Dvornicich ([DCD], Proposition 7).

Proof.

The torsor action must permute the field factors transitively; since μn\mu_{n} is cyclic, a generator ζn\zeta_{n} must cyclically permute them, and the stabilizer of T1T_{1} (as a set) is ζn/r=μr\left\langle\zeta^{n/r}\right\rangle=\mu_{r}, provinga. Since the action simply transitively permutes the coordinates of T1T_{1}, T1T_{1} is a μr\mu_{r}-torsor. If we know the T1T_{1}-component α|T1×\alpha|_{T_{1}^{\times}} of an α(T×/(T×)p)ω\alpha\in\left(T^{\times}/(T^{\times})^{p}\right)_{\omega}, then all the other components are uniquely determined by the eigenvector condition; it is only necessary for α|T1×\alpha|_{T_{1}^{\times}} to behave properly under μr\mu_{r}, namely that αT1×(T1×)ωk\alpha_{T_{1}^{\times}}\in(T_{1}^{\times})_{\omega^{k}}.

This proves b. Also, it is clear from our analysis that μs(T1)\mu_{s}(T_{1}) is a μs\mu_{s}-torsor. Applying b to this torsor proves c. ∎

We now filter T×1T^{\prime\times}_{1} as above to discover its ω\omega-component.

Proposition 7.17 (the Shafarevich basis for T×ωT^{\prime\times}_{\omega}).

As above, let T=K[βp1]T^{\prime}=K[\sqrt[p-1]{\beta}] be the (/p)×(\mathbb{Z}/p\mathbb{Z})^{\times}-torsor corresponding to the Tate dual MM^{\prime} of a cyclic Galois module MM of order pp with Kummer element βK×/(K×)p1\beta\in K^{\times}/(K^{\times})^{p-1}, and let T1T^{\prime}_{1} be the field factor of TT^{\prime}. Filter U¯=T×1/(T×1)p\bar{U}=T^{\prime\times}_{1}/(T^{\prime\times}_{1})^{p} by the subgroups U¯i\bar{U}_{i} as in the previous subsection. Since the μr\mu_{r}-torsor action on T1T^{\prime}_{1} preserves the valuation, each UiU_{i} is a subrepresentation of T×1/(T×1)pT^{\prime\times}_{1}/(T^{\prime\times}_{1})^{p}. Then:

  1. ((a))

    The ω\omega-isotypical component of U¯/U¯0\bar{U}/\bar{U}_{0} has dimension 11 if MM^{\prime} is trivial, 0 otherwise.

  2. ((b))

    The ω\omega-isotypical component of U¯i/U¯i+1\bar{U}_{i}/\bar{U}_{i+1} has dimension

    • ff if

      0<i<peT1/pp1,i0modp,ieT1/KvK(β)p1modeT1/K;0<i<\frac{pe_{T^{\prime}_{1}/\mathbb{Q}_{p}}}{p-1},\quad i\not\equiv 0\mod p,\quad i\equiv\frac{e_{T^{\prime}_{1}/K}v_{K}(\beta)}{p-1}\mod e_{T^{\prime}_{1}/K};
    • 11 if i=peT1/pp1i=\frac{pe_{T^{\prime}_{1}/\mathbb{Q}_{p}}}{p-1} and MM is trivial;

    • 0 otherwise.

Proof.

Note that U/U0=πT1U/U_{0}=\left\langle\pi_{T^{\prime}_{1}}\right\rangle is a copy of the trivial representation and that ω\omega is trivial (as a representation of μr\mu_{r}) exactly when r=1r=1, that is, MM^{\prime} is trivial.

Our convention for the Kummer map is such that

τc(βp1)=c~βp1.\tau_{c}(\sqrt[p-1]{\beta})=\tilde{c}\sqrt[p-1]{\beta}.

By a standard result in Kummer theory, the degree rr is the least integer such that β=β1(p1)/r\beta=\beta_{1}^{(p-1)/r} is a (p1)/r(p-1)/rth power, and T1=K[β1r]T^{\prime}_{1}=K[\sqrt[r]{\beta_{1}}] with

τc(β1r)=c~β1r.\tau_{c}(\sqrt[r]{\beta_{1}})=\tilde{c}\sqrt[r]{\beta_{1}}.

Let f=gcd(vK(β1),r)f^{\prime}=\gcd\big{(}v_{K}(\beta_{1}),r\big{)} and e=r/fe^{\prime}=r/f^{\prime}. We claim that ee^{\prime} and ff^{\prime} are respectively the ramification and inertia indices of T1T^{\prime}_{1} over KK. By the Euclidean algorithm, we may choose integers gg and hh such that

gv(β1)hr=f.gv(\beta_{1})-hr=f^{\prime}. (45)

Construct the elements

π=(β1r)gπKhandu=(β1r)eπKvK(β1)/f.\pi^{\prime}=\frac{\left(\sqrt[r]{\beta_{1}}\right)^{g}}{\pi_{K}^{h}}\quad\text{and}\quad u^{\prime}=\frac{\left(\sqrt[r]{\beta_{1}}\right)^{e^{\prime}}}{\pi_{K}^{v_{K}(\beta_{1})/f^{\prime}}}.

Note that vK(π)=1/ev_{K}(\pi^{\prime})=1/e^{\prime}, so

eT1/Ke.e_{T^{\prime}_{1}/K}\geq e^{\prime}. (46)

On the other hand, vK(u)=0v_{K}(u^{\prime})=0, and uu^{\prime} is an ff^{\prime}th root of the unit

β2=β1πKvK(β1).\beta_{2}=\frac{\beta_{1}}{\pi_{K}^{v_{K}(\beta_{1})}}.

Note that β2\beta_{2} is not an \ellth power for any prime |f\ell|f^{\prime}, as otherwise β\beta would be a (p1)/r(p-1)\ell/rth power, contradicting what we know about rr. So the residue class of uu^{\prime} generates a degree-ff^{\prime} extension of kKk_{K} inside kT1k_{T^{\prime}_{1}}; in particular,

fT1/Kf.f_{T^{\prime}_{1}/K}\geq f^{\prime}. (47)

Equality must hold in (46) and (47), so T1T^{\prime}_{1} has uniformizer π\pi^{\prime} and residue field generator uu^{\prime}. Note that for c𝔽p×c\in\mathbb{F}_{p}^{\times},

τc(π)=c~gπandτc(u)=c~eu.\tau_{c}(\pi^{\prime})=\tilde{c}^{g}\pi^{\prime}\quad\text{and}\quad\tau_{c}(u^{\prime})=\tilde{c}^{e^{\prime}}u^{\prime}.

We now have what we need to compute the Galois action on U¯i/U¯i+1\bar{U}_{i}/\bar{U}_{i+1}. By Proposition 7.8, the space U¯i/U¯i+1\bar{U}_{i}/\bar{U}_{i+1} is nonzero only for

0<i<peT1p1,pi0<i<\frac{pe_{T^{\prime}_{1}}}{p-1},\quad p\nmid i

(the generic units) and possibly for i=peK/(p1)i=pe_{K}/(p-1) also (the intimate units).

We begin with the first case. Here U¯i/U¯i+1Ui/Ui+1\bar{U}_{i}/\bar{U}_{i+1}\cong U_{i}/U_{i+1} has a basis

1+πiuj,0j<f.1+\pi^{\prime i}u^{\prime j},\quad 0\leq j<f^{\prime}.

For cμrc\in\mu_{r},

τc(1+πiuj)\displaystyle\tau_{c}(1+\pi^{\prime i}u^{\prime j}) =1+c~gi+ejπiuj\displaystyle=1+\tilde{c}^{gi+e^{\prime}j}\pi^{\prime i}u^{\prime j}
(1+πiuj)cgi+ejmodπi+1.\displaystyle\equiv(1+\pi^{\prime i}u^{\prime j})^{c^{gi+e^{\prime}j}}\mod\pi^{\prime i+1}.

Thus the basis element 1+πiuj1+\pi^{\prime i}u^{\prime j} generates a 11-dimensional μr\mu_{r}-submodule of Ui/Ui+1U_{i}/U_{i+1} isomorphic to ωgi+ej\omega^{gi+e^{\prime}j}. Accordingly, we select the generic units satisfying

gi+ej1modr.gi+e^{\prime}j\equiv 1\mod r.

Since r=efr=e^{\prime}f^{\prime}, there are exactly ff^{\prime} values of jj satisfying this when

e|gi1,e^{\prime}|gi-1, (48)

and none otherwise. By (45), gg is the multiplicative inverse of v(β1)/f=ev(β)/(p1)v(\beta_{1})/f^{\prime}=e^{\prime}v(\beta)/(p-1) mod ee^{\prime}, so we can rewrite (48) as

iev(β)p1mode,i\equiv\frac{e^{\prime}v(\beta)}{p-1}\mod e^{\prime},

as desired.

As to the case that i=peT1/(p1)i=pe_{T^{\prime}_{1}}/(p-1) (the intimate units), we simply note that, by Proposition 7.4, we have

(U¯i)ωH1ur(K,M),(\bar{U}_{i})_{\omega}\cong H^{1}_{\mathrm{ur}}(K,M),

so

|(U¯i)ω|=|H1ur(K,M)|=|H0(K,M)|.\lvert(\bar{U}_{i})_{\omega}\rvert=\lvert H^{1}_{\mathrm{ur}}(K,M)\rvert=\lvert H^{0}(K,M)\rvert.

(A direct computation of the torsor action on the intimate units is also possible; it turns out that U¯iμm\bar{U}_{i}\cong\mu_{m} as μm1\mu_{m-1}-modules.) ∎

By complete reducibility, we can get a basis for U¯ω\bar{U}_{\omega} from the bases for its composition factors:

Corollary 7.18.

If M𝒞pM\cong\mathcal{C}_{p}, then H1(K,M)H^{1}(K,M) has dimension

[K:p]+dim𝔽pH0(K,M)+dim𝔽pH0(K,M),[K:\mathbb{Q}_{p}]+\dim_{\mathbb{F}_{p}}H^{0}(K,M)+\dim_{\mathbb{F}_{p}}H^{0}(K,M^{\prime}),

with a basis consisting of appropriate lifts of the Shafarevich basis elements picked out by Proposition 7.17.

In particular, we have proved Theorem 7.1e.

7.3 Proof of Theorem 7.1

It now remains to recast the above results in terms of levels and offsets and prove the remaining parts of Theorem 7.1.

Let α(T×1)ω\alpha\in(T^{\prime\times}_{1})_{\omega}, α\alpha being a minimal-distance element. We consider the possibilities for the leading factor in α\alpha with respect to the Shafarevich basis; this determines |α1|\lvert\alpha-1\rvert by Corollary 7.9, and thence d:=vK(Disc(L/K))d:=v_{K}(\operatorname{Disc}(L/K)) and hence the level and offset of α\alpha.

  • If α\alpha is led by the uniformizer, then MM^{\prime} is trivial. From Theorem 7.4, we get d=peK+p1d=pe_{K}+p-1, so (α)=1\ell(\alpha)=-1 and θ(α)=1\theta(\alpha)=-1.

  • If α\alpha is led by a generic unit, then we have vT1(α1)=i,v_{T^{\prime}_{1}}(\alpha-1)=i, where ii is an integer satisfying

    0<i<peT1/pp1,i0modp,ieT1/KvK(β)p1modeT1/K,0<i<\frac{pe_{T^{\prime}_{1}/\mathbb{Q}_{p}}}{p-1},\quad i\not\equiv 0\mod p,\quad i\equiv\frac{e_{T^{\prime}_{1}/K}v_{K}(\beta)}{p-1}\mod e_{T^{\prime}_{1}/K},

    and each of these values is attained by some α\alpha. Using the one-to-one correspondence of Theorem 7.4, we get that vK(discL)v_{K}(\operatorname{disc}L) attains exactly the values dd such that

    p1<d<peK+p1,d1modp,deKvK(β)modp1.p-1<d<pe_{K}+p-1,\quad d\not\equiv-1\mod p,\quad d\equiv e_{K}-v_{K}(\beta)\mod p-1.

    Thus 0(α)<eK0\leq\ell(\alpha)<e_{K}, 0θ(α)p20\leq\theta(\alpha)\leq p-2, and θ\theta is determined by \ell via the condition mod p1.p-1.

  • If α\alpha is led by an intimate unit or α=1\alpha=1, then d=vK(discL)=vK(discT)d=v_{K}(\operatorname{disc}L)=v_{K}(\operatorname{disc}T^{\prime}) was already computed in proving Theorem 7.4. Since TT^{\prime} is a product of tamely ramified extensions, we have d<[T:K]=p1d<[T^{\prime}:K]=p-1, so =0\ell=0.

This proves a and c.

In the case that α\alpha is led by a generic unit of level \ell, 0<e0\leq\ell<e, there are alternative ways to characterize θ\theta. We have

θ={vK(β)p1}(p1),\theta=\left\{\frac{\ell-v_{K}(\beta)}{p-1}\right\}(p-1),

from which

vK(discL)=p(e)+θ=p(e)+{vK(β)p1}(p1)v_{K}(\operatorname{disc}L)=p(e-\ell)+\theta=p(e-\ell)+\left\{\frac{\ell-v_{K}(\beta)}{p-1}\right\}(p-1)

Since vK(discL)v_{K}(\operatorname{disc}L) is in bijection with |α1|\lvert\alpha-1\rvert by Theorem 7.4, we can likewise determine

vK(α1)=pvK(β)p1+1+vK(β)p1.v_{K}(\alpha-1)=\left\lfloor\frac{p\ell-v_{K}(\beta)}{p-1}\right\rfloor+1+\frac{v_{K}(\beta)}{p-1}.

Item f demands that we invert this to determine how the level (α)\ell(\alpha) changes as α\alpha ranges in a ball

|α1|d.\lvert\alpha-1\rvert\leq d.

If d<dmind<d_{\min}, then every α\alpha in this ball is a ppth power (by Proposition 7.8, or simply by noting that the Taylor series for ppth root converges on this ball), so the range of [α][\alpha] is {1}=e+1\{1\}=\mathcal{L}_{e+1}. If ddmind\geq d_{\min}, then the intimate units are certainly included, so the range of [α][\alpha] is at least e\mathcal{L}_{e}; it also includes all generic units whose levels \ell satisfy

vK(α1)\displaystyle v_{K}(\alpha-1) logdlog|πK|\displaystyle\geq\frac{\log d}{\log\lvert\pi_{K}\rvert}
pvK(β)p1+1+vK(β)p1\displaystyle\left\lfloor\frac{p\ell-v_{K}(\beta)}{p-1}\right\rfloor+1+\frac{v_{K}(\beta)}{p-1} logdlog|πK|\displaystyle\geq\frac{\log d}{\log\lvert\pi_{K}\rvert}
pvK(β)p1\displaystyle\left\lfloor\frac{p\ell-v_{K}(\beta)}{p-1}\right\rfloor logdlog|πK|1vK(β)p1.\displaystyle\geq\frac{\log d}{\log\lvert\pi_{K}\rvert}-1-\frac{v_{K}(\beta)}{p-1}.

Using the exchange

xyxyxy,\left\lfloor x\right\rfloor\geq y\iff\left\lfloor x\right\rfloor\geq\left\lceil y\right\rceil\iff x\geq\left\lceil y\right\rceil,

valid for all real numbers xx and yy, we can get this into a form solvable for \ell:

pvK(β)p1\displaystyle\frac{p\ell-v_{K}(\beta)}{p-1} logdlog|πK|1vK(β)p1\displaystyle\geq\left\lceil\frac{\log d}{\log\lvert\pi_{K}\rvert}-1-\frac{v_{K}(\beta)}{p-1}\right\rceil
\displaystyle\ell vK(β)p+p1plogdlog|πK|1vK(β)p1.\displaystyle\geq\frac{v_{K}(\beta)}{p}+\frac{p-1}{p}\left\lceil\frac{\log d}{\log\lvert\pi_{K}\rvert}-1-\frac{v_{K}(\beta)}{p-1}\right\rceil.

So the range of [α][\alpha] is i\mathcal{L}_{i}, where

vK(β)p+p1plogdlog|πK|1vK(β)p1,\Biggl{\lceil}\frac{v_{K}(\beta)}{p}+\frac{p-1}{p}\left\lceil\frac{\log d}{\log\lvert\pi_{K}\rvert}-1-\frac{v_{K}(\beta)}{p-1}\right\rceil\Biggr{\rceil}, (49)

as claimed in f.

For b, we note that as dd decreases from 11 to dmind_{\min}, the corresponding ii in (49) hits every value from 0 to ee, since the argument to the outer ceiling increases by jumps of (p1)/p<1(p-1)/p<1. So each i\mathcal{L}_{i} is a subgroup. For d, we note that |0|=|H0(K,M)|\lvert\mathcal{L}_{0}\rvert=\lvert H^{0}(K,M)\rvert, while for 1ie,1\leq i\leq e,

i+1/iU¯j/U¯j+1\mathcal{L}_{i+1}/\mathcal{L}_{i}\cong\bar{U}_{j}/\bar{U}_{j+1}

has pf=qp^{f}=q elements, where jj is the unique value of vT1(α1)v_{T^{\prime}_{1}}(\alpha-1) for values of α\alpha having level ii.

Finally, we have claimed a relation g regarding how level spaces interact with the Tate pairing. In the case of the Hilbert pairing, the result we need is as follows:

Lemma 7.19 (an explicit reciprocity law).

Let KK be a local field with μpK\mu_{p}\subseteq K. If α,β𝒪K×\alpha,\beta\in\mathcal{O}_{K}^{\times} satisfy

|α1||β1|<dmin=|p|p/(p1),\lvert\alpha-1\rvert\cdot\lvert\beta-1\rvert<d_{\min}=\lvert p\rvert^{p/(p-1)},

then the Hilbert pairing α,βK\left\langle\alpha,\beta\right\rangle_{K} vanishes.

Proof.

This is a consequence of the conductor-discriminant formula (see Neukirch [Neukirch], VII.11.9): For a Galois extension L/KL/K,

Disc(L/K)=χ𝔣(χ)χ(1),\operatorname{Disc}(L/K)=\prod_{\chi}\mathfrak{f}(\chi)^{\chi(1)},

where χ\chi ranges over the irreducible characters of Gal(L/K)\operatorname{Gal}(L/K). Here we apply the formula to L=K[αp]L=K[\sqrt[p]{\alpha}]. Scale α\alpha by ppth powers to be as close to 11 as possible. If α=1\alpha=1 or α\alpha is an intimate unit, the Hilbert pairing clearly vanishes since LL is unramified and β\beta is a unit. So we can assume that LL is a ramified extension of degree pp. Then there are pp-many characters on Gal(L/K)\operatorname{Gal}(L/K), all of dimension 11. One is the trivial character, whose conductor is 11. The others all have the same conductor 𝔣\mathfrak{f}, so

Disc(L/K)=𝔣p1.\operatorname{Disc}(L/K)=\mathfrak{f}^{p-1}.

By Theorem 7.4, we have

Disc(L/K)ppπKp1(α1)p1,\operatorname{Disc}(L/K)\sim\frac{p^{p}\cdot\pi_{K}^{p-1}}{(\alpha-1)^{p-1}},

so 𝔣\mathfrak{f} is generated by any element ff with

|f|=|pp/(p1)πKα1|.\lvert f\rvert=\left\lvert\frac{p^{p/(p-1)}\cdot\pi_{K}}{\alpha-1}\right\rvert.

Note that dmin=|p|p/(p1)d_{\min}=\lvert p\rvert^{p/(p-1)} is actually an attainable norm of an element of KK, namely (ζp1)p(\zeta_{p}-1)^{p}. By the given inequality, β1mod𝔣\beta\equiv 1\mod\mathfrak{f} which implies that the Hilbert symbol

α,βK=ϕL/K(β)\left\langle\alpha,\beta\right\rangle_{K}=\phi_{L/K}(\beta)

vanishes. ∎

If 0ie0\leq i\leq e, αi(M)\alpha\in\mathcal{L}_{i}(M), and βei(M)\beta\in\mathcal{L}_{e-i}(M), then it is easy to verify that the hypothesis of Lemma 7.19 holds in each field factor of T[μm]T[\mu_{m}], in which the Hilbert pairing is being computed. Hence

i(M)ei(M).\mathcal{L}_{i}(M)^{\perp}\supseteq\mathcal{L}_{e-i}(M^{\prime}).

However, since

|i(M)||ei(M)|=qeK|H0(K,M)||H0(K,M)|=|H1(K,M)|,\lvert\mathcal{L}_{i}(M)\rvert\cdot\lvert\mathcal{L}_{e-i}(M^{\prime})\rvert=q^{e_{K}}\cdot\lvert H^{0}(K,M)\rvert\cdot\lvert H^{0}(K,M^{\prime})\rvert=\lvert H^{1}(K,M)\rvert,

equality must hold. ∎

7.4 The tame case

If KK is a tame local field, that is, charkKp\operatorname{char}k_{K}\neq p, the structure of H1(K,M)H^{1}(K,M) is well known. We put

e=0,1={0},0=H1ur(K,M),1=H1(K,M)e=0,\quad\mathcal{L}_{-1}=\{0\},\quad\mathcal{L}_{0}=H^{1}_{\mathrm{ur}}(K,M),\quad\mathcal{L}_{1}=H^{1}(K,M)

and observe that Theorem 7.1c, d, e, g and Corollary 7.2 still hold.

The wild function field case K=𝔽pr((t))K=\mathbb{F}_{p^{r}}(\!(t)\!) admits a similar treatment, but now the number of levels is infinite. We do not address this case here.

Part III Composed varieties

8 Composed varieties

It has long been noted that orbits of certain algebraic group actions on varieties over a field KK parametrize rings of low rank over KK, which can also be identified with the cohomology of small Galois modules over KK. The aim of this section is to explain all this in a level of generality suitable for our applications.

Definition 8.1.

Let KK be a field and K¯\bar{K} its separable closure. A composed variety over KK is a quasi-projective variety VV over KK with an action of a quasi-projective algebraic group Γ\Gamma over KK such that:

  1. ((a))

    VV has a KK-rational point x0x_{0};

  2. ((b))

    the K¯\bar{K}-points of VV consist of just one orbit Γ(K¯)x0\Gamma(\bar{K})x_{0};

  3. ((c))

    the point stabilizer M=StabΓ(K¯)x0M=\operatorname{Stab}_{\Gamma(\bar{K})}x_{0} is a finite abelian subgroup.

The term composed is derived from Gauss composition of binary quadratic forms and the “higher composition laws” of the work of Bhargava and others, from which we derive many of our examples.

Proposition 8.2.
  1. ((a))

    Once a base orbit Γ(K)x0\Gamma(K)x_{0} is fixed, there is a natural injection

    ψ:Γ(K)\V(K)H1(K,M)\psi:\Gamma(K)\backslash V(K)\hookrightarrow H^{1}(K,M)

    by which the orbits Γ(K)\V(K)\Gamma(K)\backslash V(K) parametrize some subset of the Galois cohomology group H1(K,M)H^{1}(K,M).

  2. ((b))

    The Γ(K)\Gamma(K)-stabilizer of every xV(K)x\in V(K) is canonically isomorphic to H0(K,M)H^{0}(K,M).

Proof.
  1. ((a))

    Let xV(K)x\in V(K) be given. Since there is only one Γ(K¯)\Gamma(\bar{K})-orbit, we can find γΓ(K¯)\gamma\in\Gamma(\bar{K}) such that γ(x0)=x\gamma(x_{0})=x. For any gGal(K¯/K)g\in\operatorname{Gal}(\bar{K}/K), g(γ)g(\gamma) also takes x0x_{0} to xx and so differs from γ\gamma by right-multiplication by an element in StabΓ(K¯)x0=M\operatorname{Stab}_{\Gamma(\bar{K})}x_{0}=M. Define a cocycle σx:Gal(K¯/K)M\sigma_{x}:\operatorname{Gal}(\bar{K}/K)\mathop{\rightarrow}\limits M by

    σx(g)=g(γ)γ1.\sigma_{x}(g)=g(\gamma)\cdot\gamma^{-1}.

    It is routine to verify that

    • σx\sigma_{x} satisfies the cocycle condition σx(gh)=σx(g)g(σx(h))\sigma_{x}(gh)=\sigma_{x}(g)\cdot g(\sigma_{x}(h)) and hence defines an element of H1(K,M)H^{1}(K,M);

    • If a different γ\gamma is chosen, then σx\sigma_{x} changes by a coboundary;

    • If xx is replaced by αx\alpha x for some αΓK\alpha\in\Gamma_{K}, the cocycle σx\sigma_{x} is unchanged;

    • If the basepoint x0x_{0} is replaced by αx0\alpha x_{0} for some αΓ(K)\alpha\in\Gamma(K), the cocycle σx\sigma_{x} is unchanged, up to identifying MM with StabΓ(K¯)(αx0)=αMα1\operatorname{Stab}_{\Gamma(\bar{K})}(\alpha x_{0})=\alpha M\alpha^{-1} in the obvious way. (This is why we can fix merely a base orbit instead of a basepoint.)

    So we get a map

    ψ:Γ(K)\V(K)H1(K,M).\psi:\Gamma(K)\backslash V(K)\mathop{\rightarrow}\limits H^{1}(K,M).

    We claim that ψ\psi is injective. Suppose that x1,x2V(K)x_{1},x_{2}\in V(K) map to equivalent cocycles σx1\sigma_{x_{1}}, σx2\sigma_{x_{2}}. Let γiΓ(K¯)\gamma_{i}\in\Gamma(\bar{K}) be the associated transformation that maps x0x_{0} to xix_{i}. By right-multiplying γ1\gamma_{1} by an element of MM, as above, we can remove any coboundary discrepancy and assume that σx1=σx2\sigma_{x_{1}}=\sigma_{x_{2}} on the nose. That is, for every gGal(K¯/K)g\in\operatorname{Gal}(\bar{K}/K),

    g(γ1)γ11=g(γ2)γ21,g(\gamma_{1})\cdot\gamma_{1}^{-1}=g(\gamma_{2})\cdot\gamma_{2}^{-1},

    which can also be written as

    g(γ2γ11)=γ2γ11.g(\gamma_{2}\gamma_{1}^{-1})=\gamma_{2}\gamma_{1}^{-1}.

    Thus, γ2γ11\gamma_{2}\gamma_{1}^{-1} is Galois stable and hence defined over KK. It takes x1x_{1} to x2x_{2}, establishing that these points lie in the same Γ(K)\Gamma(K)-orbit, as desired.

  2. ((b))

    If γ(x0)=x\gamma(x_{0})=x, then the Γ(K¯)\Gamma(\bar{K})-stabilizer of xx is of course γMγ1\gamma M\gamma^{-1}. We claim that the obvious map

    MγMγ1\displaystyle M\mathop{\rightarrow}\limits\gamma M\gamma^{-1}
    μγμγ1\displaystyle\mu\mapsto\gamma\mu\gamma^{-1}

    is an isomorphism of Galois modules. We compute, for gGal(K¯/K)g\in\operatorname{Gal}(\bar{K}/K),

    g(γμγ1)=g(γ)g(μ)g(γ)1=γσx(g)g(μ)σx(g)1γ1=γg(μ)γ1,g\left(\gamma\mu\gamma^{-1}\right)=g(\gamma)g(\mu)g(\gamma)^{-1}=\gamma\sigma_{x}(g)g(\mu)\sigma_{x}(g)^{-1}\gamma^{-1}=\gamma g(\mu)\gamma^{-1},

    establishing the isomorphism. In particular, the Galois-stable points StabΓ(K)x0=H0(K,M)\operatorname{Stab}_{\Gamma(K)}x_{0}=H^{0}(K,M) are the same at xx as at x0x_{0}. Note the crucial way that we used that MM is abelian. By the same token, the identification of stabilizers is independent of γ\gamma and is thus canonical. ∎

The base orbit is distinguished only insofar as it corresponds to the zero element 0H1(K,M)0\in H^{1}(K,M). Changing base orbits changes the parametrization minimally:

Proposition 8.3.

The parametrizations ψx0,ψx1:Γ(K)\V(K)H1(K,M)\psi_{x_{0}},\psi_{x_{1}}:\Gamma(K)\backslash V(K)\mathop{\rightarrow}\limits H^{1}(K,M) corresponding to two basepoints x0,x1V(K)x_{0},x_{1}\in V(K) differ only by translation:

ψx1(x)=ψx0(x)ψx0(x1),\psi_{x_{1}}(x)=\psi_{x_{0}}(x)-\psi_{x_{0}}(x_{1}),

under the isomorphism between the stabilizers MM established in the previous proposition.

Proof.

Routine calculation. ∎

While ψ\psi is always injective, it need not be surjective, as we will see by examples in the following section.

Definition 8.4.
  1. ((a))

    A composed variety is full if ψ\psi is surjective, that is, it includes a Γ(K)\Gamma(K)-orbit for every cohomology class in H1(K,M)H^{1}(K,M).

  2. ((b))

    If KK is a global field, a composed variety is Hasse if for every αH1(K,M)\alpha\in H^{1}(K,M), if the localization αvH1(Kv,M)\alpha_{v}\in H^{1}(K_{v},M) at each place vv lies in the image of the local parametrization

    ψv:V(Kv)\Γ(Kv)H1(Kv,M),\psi_{v}:V(K_{v})\backslash\Gamma(K_{v})\mathop{\rightarrow}\limits H^{1}(K_{v},M),

    then α\alpha also lies in the image of the global parametrization ψ\psi.

8.1 Examples

In this section, KK is any field not of one of finitely many bad characteristics for which the exposition does not make sense.

Example 8.5.

The group Γ=𝔾m\Gamma=\mathbb{G}_{m} can act on the variety V=𝔸1\{0}V=\mathbb{A}^{1}\backslash\{0\}, the punctured affine line, by

λ(x)=λnx.\lambda(x)=\lambda^{n}\cdot x.

There is a unique K¯\bar{K}-orbit. The point stabilizer is μn\mu_{n}, and the parametrization corresponding to this composed variety (choosing basepoint x0=1x_{0}=1) is none other than the Kummer map

K×/(K×)nH1(K,μn).K^{\times}/(K^{\times})^{n}\mathop{\rightarrow}\limits H^{1}(K,\mu_{n}).

That VV is full follows from Hilbert’s Theorem 90.

Example 8.6.

Let VV be the variety of binary cubic forms ff over KK with fixed discriminant D0D_{0}. This has an algebraic action of SL2\mathrm{SL}_{2}, which is transitive over K¯\bar{K} (essentially because PSL2\mathrm{PSL}_{2} carries any three points of 1\mathbb{P}^{1} to any other three), and there is a ready-at-hand basepoint

f0(X,Y)=X2YD4Y3.f_{0}(X,Y)=X^{2}Y-\frac{D}{4}Y^{3}.

The point stabilizer MM is isomorphic to /3\mathbb{Z}/3\mathbb{Z}, but twisted by the character of K(D)K(\sqrt{D}); that is, M{0,D,D}M\cong\{0,\sqrt{D},-\sqrt{D}\} as sets with Galois action. Coupled with the appropriate higher composition law (Theorem 6.9), this recovers the parametrization of cubic étale algebras with fixed quadratic resolvent by H1(K,M)H^{1}(K,M) in Proposition 4.18. To see that it is the same parametrization, note that a γΓ(K¯/K)\gamma\in\Gamma(\bar{K}/K) that takes f0f_{0} to ff is determined by where it sends the rational root [1:0][1:0] of f0f_{0}, so the three γ\gamma’s are permuted by Gal(K¯/K)\operatorname{Gal}(\bar{K}/K) just like the three roots of ff. In particular, VV is full.

Example 8.7.

Continuing with the sequence of known ring parametrizations, we might study the variety VV of pairs of ternary quadratic forms with fixed discriminant D0D_{0}. This has one orbit over K¯\bar{K} under the action of the group Γ=SL2×SL3\Gamma=\mathrm{SL}_{2}\times\mathrm{SL}_{3}; unfortunately, the point stabilizer is isomorphic to the alternating group A4A_{4}, which is not abelian.

So we narrow the group, which widens the ring of invariants and requires us to take a smaller VV. We let Γ=SL3\Gamma=\mathrm{SL}_{3} alone act on pairs (A,B)(A,B) of ternary quadratic forms, which preserves the resolvent

g(X,Y)=4det(AX+BY),g(X,Y)=4\det\left(AX+BY\right),

a binary cubic form. We let VV be the variety of (A,B)(A,B) for which g=g0g=g_{0} is a fixed separable polynomial. These parametrize quartic étale algebras LL over KK whose cubic resolvent RR is fixed. There is a natural base orbit (A0,B0)(A_{0},B_{0}) whose associated LK×RL\cong K\times R has a linear factor. The point stabilizer M/2×/2M\cong\mathbb{Z}/2\mathbb{Z}\times\mathbb{Z}/2\mathbb{Z}, with the three non-identity elements permuted by Gal(K¯/K)\operatorname{Gal}(\bar{K}/K) in the same manner as the three roots of g0g_{0}. We have reconstructed the parametrization of quartic étale algebras with fixed cubic resolvent by H1(K,M)H^{1}(K,M) in Proposition 4.18. In particular, VV is full.

Example 8.8.

Alternatively, we can consider the space VV of binary quartic forms whose invariants I=I0I=I_{0}, J=J0J=J_{0} are fixed. The orbits of this space have been found useful for parametrizing 22-Selmer elements of the elliptic curve E:y2=x3+xI+JE:y^{2}=x^{3}+xI+J, because the point stabilizer is M/2×/2M\cong\mathbb{Z}/2\mathbb{Z}\times\mathbb{Z}/2\mathbb{Z} with the Galois-module structure E[2]E[2]. This space VV embeds into the space of the preceding example via a map which we call the Wood embedding after its prominent role in Wood’s work [WoodBQ]:

f\displaystyle f (A,B)\displaystyle\mapsto(A,B)
ax4+bx3y+cx2y2+dxy3+ey4\displaystyle ax^{4}+bx^{3}y+cx^{2}y^{2}+dxy^{3}+ey^{4} ([1/211/2],[ab/2c/3b/2c/3d/2c/3d/2e]).\displaystyle\mapsto\left(\begin{bmatrix}&&1/2\\ &-1&\\ 1/2&&\end{bmatrix},\begin{bmatrix}a&b/2&c/3\\ b/2&c/3&d/2\\ c/3&d/2&e\end{bmatrix}\right).

In general, VV is not full. For instance, over K=K=\mathbb{R}, if EE has full 22-torsion, there are only three kinds of binary quartics over \mathbb{R} with positive discriminant (positive definite, negative definite, and those with four real roots) which cover three of the four elements in H1(,/2)H^{1}(\mathbb{R},\mathbb{Z}/2\mathbb{Z}). Two of these three (positive definite, four real roots) form the subgroup of elements whose corresponding EE-torsor z2=f(x,y)z^{2}=f(x,y) is soluble at \infty: these are the ones we retain when studying Sel2E\operatorname{Sel}_{2}E. The fourth element of H1(,/2)H^{1}(\mathbb{R},\mathbb{Z}/2\mathbb{Z}) yields étale algebras whose (A,B)(A,B) has

A=[1/211/2],A=\begin{bmatrix}1/2&&\\ &1&\\ &&1/2\end{bmatrix},

a conic with no real points. However, over global fields, it is possible to show that VV is Hasse, using the Hasse-Minkowski theorem for conics.

Remark 8.9.

Because of the extreme flexibility afforded by general varieties, it is reasonable to suppose that any finite KK-Galois module MM appears as the point stabilizer of some full composed variety over KK. However, we do not pursue this question here.

8.2 Integral models; localization of orbit counts

Let KK be a number field and 𝒪K\mathcal{O}_{K} its ring of integers. Let (V,Γ)(V,\Gamma) be a composed variety, and let (𝒱,𝒢)(\mathcal{V},\mathcal{G}) be an integral model, that is, a pair of a flat separated scheme and a flat algebraic group over 𝒪K\mathcal{O}_{K} acting on it, equipped with an identification of the generic fiber with (V,Γ)(V,\Gamma). Then 𝒢(𝒪K)Γ(K)\mathcal{G}(\mathcal{O}_{K})\hookrightarrow\Gamma(K), and the Γ(K)\Gamma(K)-orbits on V(K)V(K) decompose into 𝒢(𝒪K)\mathcal{G}(\mathcal{O}_{K})-orbits.

Lemma 8.10 (localization of global class numbers).

Let (𝒱,𝒢)(\mathcal{V},\mathcal{G}) be an integral model for a composed variety (V,Γ)(V,\Gamma). For each place vv, let

wv:𝒢(𝒪v)\𝒱(𝒪v)w_{v}:\mathcal{G}(\mathcal{O}_{v})\backslash\mathcal{V}(\mathcal{O}_{v})\mathop{\rightarrow}\limits\mathbb{C}

be a function on the local orbits, which we call a local weighting. Suppose that:

  1. ((i))

    (V,Γ)(V,\Gamma) is Hasse.

  2. ((ii))

    𝒢\mathcal{G} has class number one, that is, the natural localization embedding

    𝒢(𝒪K)\Γ(K)v𝒢(𝒪v)\Γ(Kv)\mathcal{G}(\mathcal{O}_{K})\backslash\Gamma(K)\hookrightarrow\bigoplus_{v}\mathcal{G}(\mathcal{O}_{v})\backslash\Gamma(K_{v})

    is surjective.

  3. ((iii))

    For each place vv, there are only finitely many orbits of 𝒢(𝒪v)\mathcal{G}(\mathcal{O}_{v}) on 𝒱(𝒪v)\mathcal{V}(\mathcal{O}_{v}). This ensures that the weighted local orbit counter

    gv,wv:H1(Kv,M)\displaystyle g_{v,w_{v}}:H^{1}(K_{v},M) \displaystyle\mathop{\rightarrow}\limits\mathbb{C}
    α\displaystyle\alpha 𝒢(𝒪Kv)γ𝒢(𝒪v)\Γ(Kv)such that γxα𝒱(𝒪v)wv(γxα)\displaystyle\mapsto\sum_{\begin{subarray}{c}\mathcal{G}(\mathcal{O}_{K_{v}})\gamma\in\mathcal{G}(\mathcal{O}_{v})\backslash\Gamma(K_{v})\\ \text{such that }\gamma x_{\alpha}\in\mathcal{V}(\mathcal{O}_{v})\end{subarray}}w_{v}(\gamma x_{\alpha})

    takes finite values. (Here xαx_{\alpha} is a representative of the Γ(Kv)\Gamma(K_{v})-orbit corresponding to α\alpha. If there is no such orbit because VV is not full, we take gv,wv(α)=0g_{v,w_{v}}(\alpha)=0.)

  4. ((iv))

    For almost all vv, 𝒢(𝒪v)\𝒱(𝒪v)\mathcal{G}(\mathcal{O}_{v})\backslash\mathcal{V}(\mathcal{O}_{v}) consists of at most one orbit in each Γ(Kv)\Gamma(K_{v})-orbit, and wv=1w_{v}=1 identically.

Then the global integral points 𝒱(𝒪K)\mathcal{V}(\mathcal{O}_{K}) consist of finitely many 𝒢(𝒪K)\mathcal{G}(\mathcal{O}_{K})-orbits, and the global weighted orbit count can be expressed in terms of the gv,wvg_{v,w_{v}} by

h{wv}𝒢(𝒪K)x𝒢(𝒪K)\𝒱(𝒪K)vwv(x)|Stab𝒢(𝒪K)x|=1|H0(K,M)|αH1(K,M)vgv,wv(α).h_{\{w_{v}\}}\coloneqq\sum_{\mathcal{G}(\mathcal{O}_{K})x\in\mathcal{G}(\mathcal{O}_{K})\backslash\mathcal{V}(\mathcal{O}_{K})}\frac{\prod_{v}w_{v}(x)}{\lvert\operatorname{Stab}_{\mathcal{G}(\mathcal{O}_{K})}x\rvert}=\frac{1}{\lvert H^{0}(K,M)\rvert}\sum_{\alpha\in H^{1}(K,M)}\prod_{v}g_{v,w_{v}}(\alpha). (50)
Proof.

Grouping the 𝒢(𝒪K)\mathcal{G}(\mathcal{O}_{K})-orbits into Γ(K)\Gamma(K)-orbits, it suffices to prove that for all αH1(K,M)\alpha\in H^{1}(K,M),

𝒢(𝒪K)xΓ(K)xαvwv(x)|Stab𝒢(𝒪K)x|=1|H0(K,M)|vgv,wv(α).\sum_{\mathcal{G}(\mathcal{O}_{K})x\subseteq\Gamma(K)x_{\alpha}}\frac{\prod_{v}w_{v}(x)}{\lvert\operatorname{Stab}_{\mathcal{G}(\mathcal{O}_{K})}x\rvert}=\frac{1}{\lvert H^{0}(K,M)\rvert}\prod_{v}g_{v,w_{v}}(\alpha). (51)

If there is no xαx_{\alpha}, the left-hand side is zero by definition, and at least one of the gv,wv(α)g_{v,w_{v}}(\alpha) is also zero since VV is Hasse. So we fix an xαx_{\alpha}. The right-hand side of (51), which is finite by hypothesis iv since α\alpha is unramified almost everywhere, can be written as

1|H0(K,M)|{𝒢(𝒪v)γv}vvwv(γvxα),\frac{1}{\lvert H^{0}(K,M)\rvert}\sum_{\{\mathcal{G}(\mathcal{O}_{v})\gamma_{v}\}_{v}}\prod_{v}w_{v}(\gamma_{v}x_{\alpha}),

the sum being over systems of γvΓ(Kv)\gamma_{v}\in\Gamma(K_{v}) such that γvxα\gamma_{v}x_{\alpha} is 𝒪v\mathcal{O}_{v}-integral. Since 𝒢\mathcal{G} has class number one, each such system glues uniquely to a global orbit 𝒢(𝒪K)γ,γΓ(K)\mathcal{G}(\mathcal{O}_{K})\gamma,\gamma\in\Gamma(K), for which γxα\gamma x_{\alpha} is 𝒪v\mathcal{O}_{v}-integral for all vv, that is, 𝒪K\mathcal{O}_{K}-integral. Thus the right-hand side of (51) is now transformed to

1|H0(K,M)|𝒢(𝒪K)γγxα𝒱(𝒪K)vwv(γvxα).\frac{1}{\lvert H^{0}(K,M)\rvert}\sum_{\begin{subarray}{c}\mathcal{G}(\mathcal{O}_{K})\gamma\\ \quad\gamma x_{\alpha}\in\mathcal{V}(\mathcal{O}_{K})\end{subarray}}\prod_{v}w_{v}(\gamma_{v}x_{\alpha}).

Now each γ\gamma corresponds to a term of the left-hand side of (51) under the map

𝒢(𝒪K)\Γ(K)\displaystyle\mathcal{G}(\mathcal{O}_{K})\backslash\Gamma(K) 𝒢(𝒪K)\V(K)\displaystyle\mathop{\rightarrow}\limits\mathcal{G}(\mathcal{O}_{K})\backslash V(K)
𝒢(𝒪K)γ\displaystyle\mathcal{G}(\mathcal{O}_{K})\gamma 𝒢(𝒪K)γxα.\displaystyle\mapsto\mathcal{G}(\mathcal{O}_{K})\gamma x_{\alpha}.

The fiber of each 𝒢(𝒪K)x\mathcal{G}(\mathcal{O}_{K})x has size

[StabΓ(K)x:Stab𝒢(𝒪K)x]=|H0(K,M)||Stab𝒢(𝒪K)x|.[\operatorname{Stab}_{\Gamma(K)}x:\operatorname{Stab}_{\mathcal{G}(\mathcal{O}_{K})}x]=\frac{\lvert H^{0}(K,M)\rvert}{\lvert\operatorname{Stab}_{\mathcal{G}(\mathcal{O}_{K})}x\rvert}.

So we match up one term of the left-hand side, having value

vwv(x)/|Stab𝒢(𝒪K)x|,\prod_{v}w_{v}(x)/\lvert\operatorname{Stab}_{\mathcal{G}(\mathcal{O}_{K})}x\rvert,

with |H0(K,M)|/|Stab𝒢(𝒪K)x|\lvert H^{0}(K,M)\rvert/\lvert\operatorname{Stab}_{\mathcal{G}(\mathcal{O}_{K})}x\rvert-many elements on the right-hand side. In view of the outlying factor 1/|H0(K,M)|1/\lvert H^{0}(K,M)\rvert, this completes the proof. ∎

8.3 Fourier analysis of the local and global Tate pairings

We now introduce the main innovative technique of this thesis: Fourier analysis of local and global Tate duality. In structure we are indebted to Tate’s celebrated thesis [Tate_thesis], in which he

  1. 1.

    constructs a perfect pairing on the additive group of a local field KK, taking values in the unit circle N=1\mathbb{C}^{N=1}, and thus furnishing a notion of Fourier transform for \mathbb{C}-valued L1L^{1} functions on KK;

  2. 2.

    derives thereby a pairing and Fourier transform on the adele group 𝔸K\mathbb{A}_{K} of a global field KK;

  3. 3.

    proves that the discrete subgroup K𝔸KK\subseteq\mathbb{A}_{K} is a self-dual lattice and that the Poisson summation formula

    xKf(x)=xKf^(x)\sum_{x\in K}f(x)=\sum_{x\in K}\hat{f}(x) (52)

    holds for all ff satisfying reasonable integrability conditions.

In this paper, we work not with the additive group KK but with a Galois cohomology group H1(K,M)H^{1}(K,M). The needed theoretical result is Poitou-Tate duality, a nine-term exact sequence of which the middle three terms are of main interest to us:

(finite kernel)H1(K,M)vH1(Kv,M)H1(K,M)(finite cokernel).(\text{finite kernel})\mathop{\rightarrow}\limits H^{1}(K,M)\mathop{\rightarrow}\limits\sideset{}{{}^{\prime}}{\bigoplus}_{v}H^{1}(K_{v},M)\mathop{\rightarrow}\limits H^{1}(K,M^{\prime})^{\vee}\mathop{\rightarrow}\limits(\text{finite cokernel}).

This can be interpreted as saying that H1(K,M)H^{1}(K,M) and H1(K,M)H^{1}(K,M^{\prime}) (where M=Hom(M,μ)M^{\prime}=\operatorname{Hom}(M,\mu) is the Tate dual) map to dual lattices in the respective adelic cohomology groups

H1(𝔸K,M)=vH1(Kv,M)andH1(𝔸K,M)=vH1(Kv,M),H^{1}(\mathbb{A}_{K},M)=\sideset{}{{}^{\prime}}{\bigoplus}_{v}H^{1}(K_{v},M)\quad\text{and}\quad H^{1}(\mathbb{A}_{K},M^{\prime})=\sideset{}{{}^{\prime}}{\bigoplus}_{v}H^{1}(K_{v},M^{\prime}),

which are mutually dual under the product of the local Tate pairings

{αv},{βv}=vαv,βvμ.\left\langle\{\alpha_{v}\},\{\beta_{v}\}\right\rangle=\prod_{v}\left\langle\alpha_{v},\beta_{v}\right\rangle\in\mu.

Here, for KK a local field, the local Tate pairing is given by the cup product

,:H1(K,M)×H1(K,M)H2(K,μ)μ.\left\langle\bullet,\bullet\right\rangle:H^{1}(K,M)\times H^{1}(K,M^{\prime})\mathop{\rightarrow}\limits H^{2}(K,\mu)\cong\mu.

It is well known that this pairing is perfect. (The Brauer group H2(K,μ)H^{2}(K,\mu) is usually described as being /\mathbb{Q}/\mathbb{Z} but, having no need for a Galois action on it, we identify it with μ\mu to avoid the need to write an exponential in the Fourier transform.) Now, for any sufficiently nice function f:H1(𝔸K,M)f:H^{1}(\mathbb{A}_{K},M)\mathop{\rightarrow}\limits\mathbb{C} (locally constant and compactly supported is more than enough), we have Poisson summation

αH1(K,M)f(α)=cMβH1(K,M)f^(β)\sum_{\alpha\in H^{1}(K,M)}f(\alpha)=c_{M}\sum_{\beta\in H^{1}(K,M^{\prime})}\hat{f}(\beta)

for some constant cMc_{M} which we think of as the covolume of H1(K,M)H^{1}(K,M) as a lattice in the adelic cohomology. (In fact, by examining the preceding term in the Poitou-Tate sequence, H1(K,M)H^{1}(K,M) need not inject into H1(𝔸K,M)H^{1}(\mathbb{A}_{K},M), but maps in with finite kernel; but this subtlety can be absorbed into the constant cMc_{M}.)

We apply Poisson summation to the local orbit counters gvg_{v} defined in the preceding subsection and get a very general reflection theorem.

Definition 8.11.

Let KK be a local field. Let (V(1),Γ(1))(V^{(1)},\Gamma^{(1)}) and (V(2),Γ(2))(V^{(2)},\Gamma^{(2)}) be a pair of composed varieties over KK whose associated point stabilizers M(1)M^{(1)}, M(2)M^{(2)} are Tate duals of one another, and let (𝒱(i),𝒢(i))(\mathcal{V}^{(i)},\mathcal{G}^{(i)}) be an integral model of (V(i),Γ(i))(V^{(i)},\Gamma^{(i)}). Two weightings on orbits

w(i):𝒢(𝒪K)\𝒱(𝒪K)w^{(i)}:\mathcal{G}(\mathcal{O}_{K})\backslash\mathcal{V}(\mathcal{O}_{K})\mathop{\rightarrow}\limits\mathbb{C}

are called (mutually) dual with duality constant cc\in\mathbb{Q} if their local orbit counters gw(i)g_{w^{(i)}} are mutual Fourier transforms:

g(2)=cg^(1).g^{(2)}=c\cdot\hat{g}^{(1)}. (53)

where the Fourier transform is scaled by

f^(β)=1H0(K,M)αH1(K,M)f(α).\hat{f}(\beta)=\frac{1}{H^{0}(K,M)}\sum_{\alpha\in H^{1}(K,M)}f(\alpha).

An equation of the form (53) is called a local reflection theorem. If the constant weightings w(i)=1w^{(i)}=1 are mutually dual, we say that the two integral models (𝒱(i),𝒢(i))(\mathcal{V}^{(i)},\mathcal{G}^{(i)}) are naturally dual.

Theorem 8.12 (local-to-global reflection engine).

Let KK be a number field. Let (V(1),Γ(1))(V^{(1)},\Gamma^{(1)}) and (V(2),Γ(2))(V^{(2)},\Gamma^{(2)}) be a pair of composed varieties over KK whose associated point stabilizers M(1)M^{(1)}, M(2)M^{(2)} are Tate duals of one another. Let (𝒱(i),𝒢(i))(\mathcal{V}^{(i)},\mathcal{G}^{(i)}) be an integral model for each (V(i),Γ(i))(V^{(i)},\Gamma^{(i)}), and let

wv(i):𝒢(i)(𝒪v)\𝒱(i)(𝒪v)w_{v}^{(i)}:\mathcal{G}^{(i)}(\mathcal{O}_{v})\backslash\mathcal{V}^{(i)}(\mathcal{O}_{v})\mathop{\rightarrow}\limits\mathbb{C}

be a local weighting on each integral model. Suppose that each integral model and local weighting satisfies the hypotheses of Lemma 8.10, and suppose that at each place vv, the two integral models are dual with some duality constant cvc_{v}\in\mathbb{Q}. Then the weighted global class numbers are in a simple ratio:

h{wv(2)}=vcvh{wv(1)}.h_{\left\{w_{v}^{(2)}\right\}}=\prod_{v}c_{v}\cdot h_{\left\{w_{v}^{(1)}\right\}}.
Proof.

By Lemma 8.10,

h{wv(i)}=1|H0(K,M(i))|αH1(K,M(i))vgv,wv(i)(α).h_{\left\{w_{v}^{(i)}\right\}}=\frac{1}{\lvert H^{0}(K,M^{(i)})\rvert}\sum_{\alpha\in H^{1}(K,M^{(i)})}\prod_{v}g_{v,w_{v}^{(i)}}(\alpha).

At almost all vv, each gv,wv(i)g_{v,w_{v}^{(i)}} is supported on the unramified cohomology, and must be constant there because otherwise its Fourier transform would not be supported on the unramified cohomology. However, gv,wv(i)g_{v,w_{v}^{(i)}} cannot be identically 0 because of the existance of a global basepoint. So for such vv,

gv,wv(i)=𝟏H1ur(K,M(i))andcv=1.g_{v,w_{v}^{(i)}}=\mathbf{1}_{H^{1}_{\mathrm{ur}}(K,M^{(i)})}\quad\text{and}\quad c_{v}=1.

In particular, the product vgv,wv(i)\prod_{v}g_{v,w_{v}^{(i)}} is a locally constant, compactly supported function on H1(K,𝔸K)H^{1}(K,\mathbb{A}_{K}), which is more than enough for Poisson summation to be valid.

Since the pairing between the adelic cohomology groups H1(𝔸K,M(i))H^{1}(\mathbb{A}_{K},M^{(i)}) is made by multiplying the local Tate pairings, a product of local factors has a Fourier transform with a corresponding product expansion:

vgv,wv(1)^=vg^v,wv(1)=vcvvgv,wv(2).