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

On Mordell-Weil groups
and congruences between derivatives
of twisted Hasse-Weil LL-functions

David Burns, Daniel Macias Castillo and Christian Wuthrich King’s College London, Department of Mathematics, London WC2R 2LS, U.K. david.burns@kcl.ac.uk Instituto de Ciencias Matemáticas (ICMAT), 28049 Madrid, Spain. daniel.macias@icmat.es School of Math. Sciences, University of Nottingham, Nottingham NG7 2RD, U.K. christian.wuthrich@nottingham.ac.uk
Abstract.

Let AA be an abelian variety defined over a number field kk and let FF be a finite Galois extension of kk. Let pp be a prime number. Then under certain not-too-stringent conditions on AA and FF we compute explicitly the algebraic part of the pp-component of the equivariant Tamagawa number of the pair (h1(A/F)(1),[Gal(F/k)])\bigl{(}h^{1}(A_{/F})(1),\mathbb{Z}[{\rm Gal}(F/k)]\bigr{)}. By comparing the result of this computation with the theorem of Gross and Zagier we are able to give the first verification of the pp-component of the equivariant Tamagawa number conjecture for an abelian variety in the technically most demanding case in which the relevant Mordell-Weil group has strictly positive rank and the relevant field extension is both non-abelian and of degree divisible by pp. More generally, our approach leads us to the formulation of certain precise families of conjectural pp-adic congruences between the values at s=1s=1 of derivatives of the Hasse-Weil LL-functions associated to twists of AA, normalised by a product of explicit equivariant regulators and periods, and to explicit predictions concerning the Galois structure of Tate-Shafarevich groups. In several interesting cases we provide theoretical and numerical evidence in support of these more general predictions.

1. Introduction

Let AA be an abelian variety defined over a number field kk. Then for any finite Galois extension FF of kk with group GG the equivariant Tamagawa number conjecture for the pair (h1(A/F)(1),[G])\bigl{(}h^{1}(A_{/F})(1),\mathbb{Z}[G]\bigr{)} is formulated in [16, Conjecture 4(iv)] as an equality in a relative algebraic KK-group. This conjectural equality is rather technical to state and very inexplicit in nature but is known to constitute a strong and simultaneous refinement of the Birch and Swinnerton-Dyer conjecture for AA over each of the intermediate fields of F/kF/k.

The refined conjecture also naturally decomposes into ‘components’, one for each rational prime pp, in a way that will be made precise in §5.1 below, and each such pp-component (which for convenience we refer to as ‘eTNCp’ in the remainder of this introduction) is itself of some interest. We recall, for example, that if AA has good ordinary reduction at pp, then the compatibility result proved by Venjakob and the first named author in [19, Theorem 8.4] shows that eTNCp is very closely related to the main conjecture of non-commutative Iwasawa theory for AA with respect to any compact pp-adic Lie extension of kk that contains FF.

If pp does not divide the order of GG, then, without any hypothesis on the reduction of AA at pp, it is straightforward to use the techniques developed in [15, §1.7] to give an explicit interpretation of eTNCp (although, of course, obtaining a full proof in this case still remains a very difficult problem). However, if pp divides the order of GG, then even obtaining an explicit interpretation of eTNCp has hitherto seemed to be a very difficult problem - see, for example, the considerable efforts made by Bley [3, 4] in this direction.

One of the main goals of the present article is therefore to develop a general approach which can be used to obtain (both theoretical and numerical) verifications of eTNCp in the technically most demanding case in which the abelian variety AA has strictly positive rank over FF and the Galois group GG is both non-abelian and of order divisible by pp.

To give a concrete example of the results we shall obtain in this way, we note that in §6.1 our techniques will be combined with the theorem of Gross and Zagier to prove the following result (and to help put this result into context see Remarks 1.2 and 1.3 below).

Theorem 1.1.

Let pp be an odd prime, KK an imaginary quadratic field in which pp is unramified and FF its Hilbert pp-classfield. Write GG for the generalised dihedral group Gal(F/)\operatorname{Gal}(F/\mathbb{Q}).

Let AA be an elliptic curve over \mathbb{Q} which satisfies, with respect to the field KK, the hypotheses (a), (b), (c), (e) and (f) that are listed in §3 and for which the Tate-Shafarevich group X(AF)\hbox{\russ\char 88\relax}(A_{F}) is finite.

Assume in addition that each of the following four hypotheses is satisfied.

  1. (i)

    all primes of bad reduction for AA split in K/K/\mathbb{Q};

  2. (ii)

    the Galois group of K(A[p])/KK\bigl{(}A[p]\bigr{)}/K is isomorphic to GL2(𝔽p)\operatorname{GL}_{2}(\mathbb{F}_{p});

  3. (iii)

    the Hasse-Weil LL-function L(A/K,s)L(A_{/K},s) has a simple zero at s=1s=1;

  4. (iv)

    the trace to A(K)A(K) of the Heegner point is not divisible by pp.

Then the pp-component of the equivariant Tamagawa number conjecture is valid for the pair (h1(A/F)(1),[G])\bigl{(}h^{1}(A_{/F})(1),\mathbb{Z}[G]\bigr{)}.

Remark 1.2.

For a fixed elliptic curve AA over \mathbb{Q} and imaginary quadratic field KK, it will be clear that the hypotheses (a), (b) and (c) that occur in the statement of Theorem 1.1 are satisfied by all but finitely many primes pp. In addition, it will be clear that for a fixed AA and pp the hypotheses (e) and (f) only constitute a mild restriction on the ramification of the extension K/K/\mathbb{Q}. Finally we recall that, by a famous result of Serre [46], the hypothesis (ii) holds for almost all pp if AA does not have complex multiplication.

Remark 1.3.

In the context of Theorem 1.1 we also note that for any natural number nn there are infinitely many imaginary quadratic fields KK in which pp does not ramify and for which the group Gal(F/K)\operatorname{Gal}(F/K) has exponent divisible by pnp^{n}. (For details of an explicit, and infinite, family of such fields see for example Louboutin [36].) In this context it therefore seems worth noting that the only previous verification of eTNCp for an abelian variety AA and a Galois extension of degree divisible by pp is given by Bley in [5, Corollary 1.4 and Remark 4.1] where it is assumed, amongst other things, that AA is an elliptic curve, FF is an abelian extension of \mathbb{Q} of exponent pp and, critically, that the rank of AA over FF is equal to zero.

In addition to the proof of Theorem 1.1, in §6 we will also provide further explicit examples in which our approach leads to a proof of eTNCp in cases for which pp divides the degree of the relevant extension. More precisely, we show that for certain elliptic curves AA the validity of eTNCp with p=3p=3 follows from that of the relevant cases of the Birch and Swinnerton-Dyer conjecture for a family of S3S_{3}-extensions of number fields (see Corollary 6.2) and provide examples (with p=5p=5 and p=7p=7) in which our approach allows eTNCp to be verified by numerical computations (see §6.3).

In order to prove the above results we must first establish several intermediate results which are both more general and also we feel of some independent interest. To briefly discuss these results, we fix an abelian variety AA over a number field kk, a finite Galois extension FF of kk with Galois group GG and an odd prime pp.

In §4 we shall first compute, under some not-too-stringent conditions on AA and FF, the ‘algebraic part’ of the pp-component of the equivariant Tamagawa number of the pair (h1(A/F)(1),[G])\bigl{(}h^{1}(A_{/F})(1),\mathbb{Z}[G]\bigr{)}. This computation requires a close analysis of certain refined Euler characteristics (in the sense reviewed in §4.1) that are constructed by combining the finite support cohomology complex of Bloch and Kato for the base change through F/kF/k of the pp-adic Tate module of the dual AtA^{t} of AA together with the Néron-Tate height of AA relative to the field FF. We note that the main result of this analysis constitutes a natural equivariant refinement and/or generalisation of several earlier computations in this area including those that are made by Venjakob in [50, §3.1], by Bley in [3], by Kings in [35, Lecture 3] and by the first named author in [14, §12].

The detailed computation made in §4 then allows us in §5 to interpret the relevant case of eTNCp as a family of pp-adic congruence relations between the values at s=1s=1 of higher derivatives of the Hasse-Weil LL-functions of twists of AA by irreducible complex characters of GG, suitably normalised by a product of explicit equivariant regulators and periods.

We next describe several interesting (conjectural) consequences of this reinterpretation of eTNCp, including the predicted annihilation as a Galois module of the pp-primary Tate-Shafarevich of AtA^{t} over FF by elements which interpolate the (suitably normalised) values at s=1s=1 of higher derivatives of twisted Hasse-Weil LL-functions (see Proposition 5.2 and Proposition 5.6).

By using results on the explicit Galois structure of Selmer groups that are obtained in [18], we then show that for generalised dihedral groups GG the necessary pp-adic congruence relations can be made very explicit even in the case that AA has strictly positive rank over FF (see Theorem 5.8).

The latter explicit interpretation is then used as a key step in the proof of Theorem 1.1.

We note finally that there are several ways in which it seems reasonable to expect that some of the explicit computations made in §4 could (with perhaps considerably more effort) be extended and our overall approach thereby generalised and that we hope such possibilities will be further explored in subsequent articles.

1.1. Acknowledgements

The authors are very grateful to Werner Bley and Stefano Vigni for helpful discussions and correspondence and to the anonymous referee for making several useful suggestions.

2. Notations and setting

For any finite group Γ\Gamma we use the following notation. We write Ir(Γ)\operatorname{Ir}(\Gamma) for the set of irreducible EE-valued characters of Γ\Gamma, where EE denotes either \mathbb{C} or p\mathbb{C}_{p} (and the intended meaning will always be clear from the context). We also write 𝟏Γ\boldsymbol{1}_{\Gamma} for the trivial character of Γ\Gamma and ψˇ\check{\psi} for the contragredient of each ψ\psi in Ir(Γ)\operatorname{Ir}(\Gamma). For each ψ\psi in Ir(Γ)\operatorname{Ir}(\Gamma) we write

eψ=ψ(1)|Γ|γΓψ(γ1)γe_{\psi}=\frac{\psi(1)}{|\Gamma|}\,\sum_{\gamma\in\Gamma}\psi(\gamma^{-1})\,\gamma

for the primitive idempotent of the centre ζ(E[Γ])\zeta\bigl{(}E[\Gamma]\bigr{)} of the group ring E[Γ]E[\Gamma]. For each EE-valued character ψ\psi we also fix an E[Γ]E[\Gamma]-module VψV_{\psi} of character ψ\psi.

For any abelian group MM we write MtorM_{\rm tor} for its torsion subgroup and MtfM_{\rm tf} for the quotient M/MtorM/M_{\rm tor}, which we often regard as a subgroup of M\mathbb{Q}\otimes_{\mathbb{Z}}M. For any prime pp we write M[p]M[p] for the subgroup {mM:pm=0}\{m\in M:pm=0\} of the Sylow pp-subgroup M[p]M[p^{\infty}] of MtorM_{\rm tor}. We also set Mp:=pMM_{p}:=\mathbb{Z}_{p}\otimes_{\mathbb{Z}}M and write MpM^{\wedge}_{p} for the pro-pp-completion limnM/pnM\varprojlim_{n}M/p^{n}M. If MM is finitely generated, then we set rk(M):=dim(M)\operatorname{rk}(M):=\dim_{\mathbb{Q}}(\mathbb{Q}\otimes_{\mathbb{Z}}M).

For any p[Γ]\mathbb{Z}_{p}[\Gamma]-module MM we write MM^{\vee} for the Pontryagin dual Homp(M,p/p)\operatorname{Hom}_{\mathbb{Z}_{p}}(M,\mathbb{Q}_{p}/\mathbb{Z}_{p}) and MM^{*} for the linear dual Homp(M,p)\operatorname{Hom}_{\mathbb{Z}_{p}}(M,\mathbb{Z}_{p}), each endowed with the natural contragredient action of Γ\Gamma. If MM is finitely generated, then for any field extension FF of p\mathbb{Q}_{p} we set FM:=FpMF\cdot M:=F\otimes_{\mathbb{Z}_{p}}M.

For any Galois extension of fields L/KL/K we write GL/KG_{L/K} in place of Gal(L/K)\operatorname{Gal}(L/K). We also fix an algebraic closure KcK^{c} of KK and abbreviate GKc/KG_{K^{c}/K} to GKG_{K}. For each non-archimedean place vv of a number field we write κv\kappa_{v} for its residue field.

Throughout this paper, we will consider the following situation. We have a fixed odd prime pp and a Galois extension F/kF/k of number fields with Galois group G=GF/kG=G_{F/k}. We choose a pp-Sylow subgroup PP in GG and set K:=FPK:=F^{P}. We give ourselves an abelian variety AA of dimension dd defined over kk. We write AtA^{t} for the dual abelian variety of AA.

For each intermediate field LL of F/kF/k we write SpLS_{p}^{L}, SrLS^{L}_{\rm r} and SbLS^{L}_{\rm b} for the set of non-archimedean places of LL that are pp-adic, which ramify in F/LF/L and at which A/LA_{/L} has bad reduction respectively. Similarly, we write SLS_{\infty}^{L}, SLS^{L}_{\mathbb{R}} and SLS^{L}_{\mathbb{C}} for the set of archimedean, real and complex places of LL respectively. The notation S?S_{\text{?}} will stand for S?kS_{\text{?}}^{k} when “?” is b, r, pp, \infty, \mathbb{R} or \mathbb{C}.

For each such field LL we also set rk(AL):=rk(A(L))\operatorname{rk}(A_{L}):=\operatorname{rk}(A(L)) and write Xp(AL)\hbox{\russ\char 88\relax}_{p}(A_{L}) and Selp(AL)\operatorname{Sel}_{p}(A_{L}) for the pp-primary Tate-Shafarevich and Selmer groups of A/LA_{/L}. We recall that there exists a canonical exact sequence of the form

(1) 0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Xp(AL)\textstyle{\hbox{\russ\char 88\relax}_{p}(A_{L})^{\vee}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Selp(AL)\textstyle{\operatorname{Sel}_{p}(A_{L})^{\vee}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}A(L)p\textstyle{A(L)_{p}^{*}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0.\textstyle{0.}

3. The hypotheses

Fix an odd prime pp, number fields k,Fk,F and K=FPK=F^{P} and an abelian variety AA as described above.

Then throughout this article we will find it convenient to assume that this data satisfies the following hypotheses:

  • (a)

    A(K)[p]=0A(K)[p]=0 and At(K)[p]=0A^{t}(K)[p]=0;

  • (b)

    The Tamagawa number of A/KA_{/K} at each place in SbKS_{\rm b}^{K} is not divisible by pp;

  • (c)

    A/KA_{/K} has good reduction at all pp-adic places;

  • (d)

    For all pp-adic places vv that ramify in F/KF/K, the reduction is ordinary and A(κv)[p]=0A(\kappa_{v})[p]=0;

  • (e)

    No place of bad reduction for A/kA_{/k} is ramified in F/kF/k, i.e. SbSr=S_{\rm b}\cap S_{\rm r}=\emptyset.

  • (f)

    For any place vv in KK such that the primes above it in FF are ramified in F/kF/k, we have A(κv)[p]=0A(\kappa_{v})[p]=0;

  • (g)

    The Tate-Shafarevich group X(AF)\hbox{\russ\char 88\relax}(A_{F}) is finite.

Remark 3.1.

For a fixed abelian variety AA over kk and extension K/kK/k the hypotheses (a), (b) and (c) are clearly satisfied by all but finitely many primes pp (which do not divide the degree of K/kK/k), the hypotheses (e) and (f) constitute a mild restriction on the ramification of F/kF/k and the hypothesis (g) is famously conjectured to be true in all cases. However, the hypothesis (d) excludes the case that is called ‘anomalous’ by Mazur in [38] and, for a given AA, there may be infinitely many primes pp for which there are pp-adic places vv at which AA has good ordinary reduction but A(κv)[p]A(\kappa_{v})[p] does not vanish. Nevertheless, it is straightforward to describe examples of abelian varieties AA for which there are only finitely many such anomalous places – see, for example, the result of Mazur and Rubin in [39, Lemma A.5].

Remark 3.2.

In the analysis of finite support cohomology complexes that is given in Lemma 4.1 below we find it convenient to adopt the convention of considering the pp-adic Tate module of AtA^{t} rather than that of AA itself. Applications of the hypotheses (b)(f) in this article will therefore often take place with AA replaced by AtA^{t}. In this regard, we note that the validity of each of these hypotheses as currently formulated is equivalent to the validity of the corresponding hypothesis with AA replaced by AtA^{t} and we will henceforth use this fact without further explicit comment.

4. Canonical Euler characteristics

In most of this section we assume that the fixed data p,k,F,Kp,k,F,K and AA satisfy all of the hypotheses (a)(g) that are listed in §3.

Our main aim is to compute explicitly the natural (refined) Euler characteristic that is associated to the pair comprising the Bloch-Kato finite support cohomology complex of the base-change through F/kF/k of the pp-adic Tate module of the dual abelian variety AtA^{t} of AA and the Néron-Tate height of AA relative to the field FF.

4.1. Euler characteristics

We first quickly review the definition of refined Euler characteristics that will play an essential role in the sequel. For convenience, we shall only give an explicit construction in the relevant special case rather than discussing the general approach (which is given in [13])

For any finite group Γ\Gamma we write D(p[Γ])D\bigl{(}\mathbb{Z}_{p}[\Gamma]\bigr{)} for the derived category of complexes of (left) p[Γ]\mathbb{Z}_{p}[\Gamma]-modules. We also write Dp(p[Γ])D^{\rm p}\bigl{(}\mathbb{Z}_{p}[\Gamma]\bigr{)} for the full triangulated subcategory of D(p[Γ])D\bigl{(}\mathbb{Z}_{p}[\Gamma]\bigr{)} comprising complexes that are ‘perfect’ (that is, isomorphic in D(p[Γ])D\bigl{(}\mathbb{Z}_{p}[\Gamma]\bigr{)} to a bounded complex of finitely generated projective p[Γ]\mathbb{Z}_{p}[\Gamma]-modules).

We assume to be given a complex CC in Dp(p[G])D^{\rm p}\bigl{(}\mathbb{Z}_{p}[G]\bigr{)} which is acyclic outside degrees aa and a+1a+1 for any given integer aa and such that Ha(C)H^{a}(C) is p\mathbb{Z}_{p}-free. We also assume to be given an isomorphism of p[G]\mathbb{C}_{p}[G]-modules λ:ppHa(C)ppHa+1(C)\lambda:\mathbb{C}_{p}\otimes_{\mathbb{Z}_{p}}H^{a}(C)\cong\mathbb{C}_{p}\otimes_{\mathbb{Z}_{p}}H^{a+1}(C). Then, under these hypotheses, it can be shown that there exists an isomorphism ι:PC\iota:P^{\bullet}\cong C in Dp(p[G])D^{\rm p}\bigl{(}\mathbb{Z}_{p}[G]\bigr{)} where PP^{\bullet} is a complex of p[G]\mathbb{Z}_{p}[G]-modules of the form P𝑑PP\xrightarrow{d}P where PP is finitely generated and projective and the first term occurs in degree aa. We then consider the following auxiliary composite isomorphism of p[G]\mathbb{C}_{p}[G]-modules

λP:ppP(ppHa(P))(ppim(d))(ppHa+1(P))(ppim(d))ppP\lambda_{P^{\bullet}}:\mathbb{C}_{p}\otimes_{\mathbb{Z}_{p}}P\cong\bigl{(}\mathbb{C}_{p}\otimes_{\mathbb{Z}_{p}}H^{a}(P^{\bullet})\bigr{)}\oplus\bigl{(}\mathbb{C}_{p}\otimes_{\mathbb{Z}_{p}}\operatorname{im}(d)\bigr{)}\\ \cong\bigl{(}\mathbb{C}_{p}\otimes_{\mathbb{Z}_{p}}H^{a+1}(P^{\bullet})\bigr{)}\oplus\bigl{(}\mathbb{C}_{p}\otimes_{\mathbb{Z}_{p}}\operatorname{im}(d)\bigr{)}\cong\mathbb{C}_{p}\otimes_{\mathbb{Z}_{p}}P

where the first and third maps are obtained by choosing p[G]\mathbb{C}_{p}[G]-equivariant splittings of the tautological surjections ppPppim(d)\mathbb{C}_{p}\otimes_{\mathbb{Z}_{p}}P\to\mathbb{C}_{p}\otimes_{\mathbb{Z}_{p}}\operatorname{im}(d) and ppPppHa+1(P)\mathbb{C}_{p}\otimes_{\mathbb{Z}_{p}}P\to\mathbb{C}_{p}\otimes_{\mathbb{Z}_{p}}H^{a+1}(P^{\bullet}) respectively and the second is ((ppHa+1(ι))1λ(ppHa(ι)),id)\bigl{(}(\mathbb{C}_{p}\otimes_{\mathbb{Z}_{p}}H^{a+1}(\iota))^{-1}\circ\lambda\circ(\mathbb{C}_{p}\otimes_{\mathbb{Z}_{p}}H^{a}(\iota)),{\mathrm{id}}\bigr{)}.

We now write K0(p[G],p[G])K_{0}\bigl{(}\mathbb{Z}_{p}[G],\mathbb{C}_{p}[G]\bigr{)} for the relative algebraic KK-group of the inclusion p[G]p[G]\mathbb{Z}_{p}[G]\subset\mathbb{C}_{p}[G] and recall that this group is generated by elements of the form [Q,μ,Q][Q,\mu,Q^{\prime}] where QQ and QQ^{\prime} are finitely generated projective p[G]\mathbb{Z}_{p}[G]-modules and μ\mu is an isomorphism of p[G]\mathbb{C}_{p}[G]-modules ppQppQ\mathbb{C}_{p}\otimes_{\mathbb{Z}_{p}}Q\cong\mathbb{C}_{p}\otimes_{\mathbb{Z}_{p}}Q^{\prime}.

In particular, in terms of this description, it can be shown that the element of K0(p[G],p[G])K_{0}\bigl{(}\mathbb{Z}_{p}[G],\mathbb{C}_{p}[G]\bigr{)} obtained by setting

χG,p(C,λ):=(1)a[P,λP,P]\chi_{G,p}(C,\lambda):=(-1)^{a}\,[P,\lambda_{P^{\bullet}},P]

is independent of the choice of isomorphism ι\iota (and hence of the module PP) and of the splittings made when defining the isomorphism λP\lambda_{P^{\bullet}}. This element will be referred to as the ‘refined Euler characteristic’ of the pair (C,λ)(C,\lambda) in the sequel.

4.2. Cohomology with finite support

We now turn to consider the finite support cohomology complex of the pp-adic Tate module Tp(A)T_{p}(A) of AtA^{t}.

To do this we write Σk(F)\Sigma_{k}(F) for the set of kk-embeddings FkcF\to k^{c} and YF/k,pY_{F/k,p} for the module Σk(F)p\prod_{\Sigma_{k}(F)}\mathbb{Z}_{p}, endowed with the obvious action of G×GkG\times G_{k}.

Then the pp-adic Tate module of the base change of AtA^{t} through F/kF/k is equal to

Tp,F(A):=YF/k,ppTp(A),T_{p,F}(A):=Y_{F/k,p}\otimes_{\mathbb{Z}_{p}}T_{p}(A),

where GG acts naturally on the first factor and GkG_{k} acts diagonally. We set Vp,F(A):=pTp,F(A)V_{p,F}(A):=\mathbb{Q}_{p}\,T_{p,F}(A).

For each place vv of kk we set Gv:=GkvG_{v}:=G_{k_{v}} and write IvI_{v} for the inertia subgroup of GvG_{v}. We also fix a place ww of FF above vv and a corresponding embedding FkvcF\to k_{v}^{c} and write G¯v\overline{G}_{v} and I¯v\overline{I}_{v} for the images of GvG_{v} and IvI_{v} under the induced homomorphism GvGG_{v}\to G.

For each non-archimedean place vv we then define a complex of p[G]\mathbb{Z}_{p}[G]-modules by setting

RΓf(kv,Tp,F(A)):={Tp,F(A)Iv1Frv1Tp,F(A)Iv, if vpHf1(kv,Tp,F(A))[1], if vpR\Gamma_{f}\bigl{(}k_{v},T_{p,F}(A)\bigr{)}:=\begin{cases}T_{p,F}(A)^{I_{v}}\xrightarrow{1-\operatorname{Fr}_{v}^{-1}}T_{p,F}(A)^{I_{v}},&\text{ if $v\nmid p$}\\ H^{1}_{f}\bigl{(}k_{v},T_{p,F}(A)\bigr{)}[-1],&\text{ if $v\mid p$}\end{cases}

where, if vpv\nmid p, the first occurrence of Tp,F(A)IvT_{p,F}(A)^{I_{v}} is placed in degree zero and Frv\operatorname{Fr}_{v} is the natural Frobenius in Gv/IvG_{v}/I_{v} and, if vpv\mid p, then Hf1(kv,Tp,F(A))H^{1}_{f}\bigl{(}k_{v},T_{p,F}(A)\bigr{)} denotes the full pre-image in H1(kv,Tp,F(A))H^{1}\bigl{(}k_{v},T_{p,F}(A)\bigr{)} of the ‘finite part’ subspace Hf1(kv,Vp,F(A))H^{1}_{f}\bigl{(}k_{v},V_{p,F}(A)\bigr{)} of H1(kv,Vp,F(A))H^{1}\bigl{(}k_{v},V_{p,F}(A)\bigr{)} that is defined by Bloch and Kato in [8, (3.7.2)].

We next define RΓ/f(kv,Tp,F(A))R\Gamma_{/f}\bigl{(}k_{v},T_{p,F}(A)\bigr{)} to be a complex of p[G]\mathbb{Z}_{p}[G]-modules which lies in an exact triangle in D(p[G])D\bigl{(}\mathbb{Z}_{p}[G]\bigr{)} of the form

(2) RΓf(kv,Tp,F(A))\textstyle{R\Gamma_{f}\bigl{(}k_{v},T_{p,F}(A)\bigr{)}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ϖv\scriptstyle{\varpi_{v}}RΓ(kv,Tp,F(A))\textstyle{R\Gamma\bigl{(}k_{v},T_{p,F}(A)\bigr{)}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ϖv\scriptstyle{\varpi^{\prime}_{v}}RΓ/f(kv,Tp,F(A))\textstyle{R\Gamma_{/f}\bigl{(}k_{v},T_{p,F}(A)\bigr{)}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}

where ϖv\varpi_{v} is the natural inflation morphism if vpv\nmid p, and is induced by the inclusion Hf1(kv,Tp,F(A))H1(kv,Tp,F(A))H^{1}_{f}\bigl{(}k_{v},T_{p,F}(A)\bigr{)}\subseteq H^{1}\bigl{(}k_{v},T_{p,F}(A)\bigr{)} and the fact that RΓ(kv,Tp,F(A))R\Gamma\bigl{(}k_{v},T_{p,F}(A)\bigr{)} is acyclic in degrees less than one, if vpv\mid p. We finally set S:=SrSbS:=S_{\rm r}\cup S_{\rm b}, write 𝒪k,S\mathcal{O}_{k,S} for the subring of kk comprising elements that are integral at all non-archimedean places outside SS and define the complex RΓf(k,Tp,F(A))R\Gamma_{f}\bigl{(}k,T_{p,F}(A)\bigr{)} so that it lies in an exact triangle in D(p[G])D\bigl{(}\mathbb{Z}_{p}[G]\bigr{)} of the form

(3) RΓf(k,Tp,F(A))RΓ(𝒪k,S[1p],Tp,F(A))ϖvSSpRΓ/f(kv,Tp,F(A)).\lx@xy@svg{\hbox{\raise 2.5pt\hbox{\kern 39.87648pt\hbox{\ignorespaces\ignorespaces\ignorespaces\hbox{\vtop{\kern 0.0pt\offinterlineskip\halign{\entry@#!@&&\entry@@#!@\cr&&&\crcr}}}\ignorespaces{\hbox{\kern-39.87648pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise-2.5pt\hbox{$\textstyle{R\Gamma_{f}\bigl{(}k,T_{p,F}(A)\bigr{)}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces{\hbox{\kern 55.2654pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}{\hbox{\kern 55.2654pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise-2.5pt\hbox{$\textstyle{R\Gamma\Bigl{(}\mathcal{O}_{k,S}\bigl{[}\tfrac{1}{p}\bigr{]},T_{p,F}(A)\Bigr{)}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces\ignorespaces\ignorespaces{\hbox{\kern 158.60619pt\raise 5.8978pt\hbox{{}\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\hbox{\hbox{\kern 0.0pt\raise-2.8978pt\hbox{$\scriptstyle{\varpi^{\prime}}$}}}\kern 3.0pt}}}}}}\ignorespaces{\hbox{\kern 173.91304pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}{\hbox{\kern 173.91304pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise-2.5pt\hbox{$\textstyle{\bigoplus\limits_{v\in S\cup S_{p}}R\Gamma_{/f}\bigl{(}k_{v},T_{p,F}(A)\bigr{)}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces{\hbox{\kern 319.21706pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}{\hbox{\kern 319.21706pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise-2.5pt\hbox{$\textstyle{}$}}}}}}}\ignorespaces}}}}\ignorespaces.

Here each vv-component of ϖ\varpi^{\prime} is equal to the composite of the natural localisation morphism RΓ(𝒪k,S[1p],Tp,F(A))RΓ(kv,Tp,F(A))R\Gamma\bigl{(}\mathcal{O}_{k,S}[\tfrac{1}{p}],T_{p,F}(A)\bigr{)}\to R\Gamma\bigl{(}k_{v},T_{p,F}(A)\bigr{)} in étale cohomology and the morphism ϖv\varpi^{\prime}_{v}.

In the sequel we also set

Hfi(k,Tp,F(A)):=Hi(RΓf(k,Tp,F(A)))H^{i}_{f}\bigl{(}k,T_{p,F}(A)\bigr{)}:=H^{i}\bigl{(}R\Gamma_{f}(k,T_{p,F}(A))\bigr{)}

in each degree ii.

Lemma 4.1.
  1. (i)

    Assume that AA and FF satisfy the hypotheses (c)(f). Then RΓf(k,Tp,F(A))R\Gamma_{f}\bigl{(}k,T_{p,F}(A)\bigr{)} belongs to Dp(p[G])D^{\rm p}\bigl{(}\mathbb{Z}_{p}[G]\bigr{)}. In addition, for any non-archimedean place vv of kk which ramifies in F/kF/k and does not divide pp, the complex RΓf(kv,Tp,F(A))R\Gamma_{f}\bigl{(}k_{v},T_{p,F}(A)\bigr{)} is acyclic.

  2. (ii)

    Assume that AA and FF satisfy the hypotheses (a), (b), (e) and (g). Then the complex RΓf(k,Tp,F(A))R\Gamma_{f}\bigl{(}k,T_{p,F}(A)\bigr{)} is acyclic outside degrees one and two and there are canonical identifications of Hf1(k,Tp,F(A))H^{1}_{f}\bigl{(}k,T_{p,F}(A)\bigr{)} and Hf2(k,Tp,F(A))H^{2}_{f}\bigl{(}k,T_{p,F}(A)\bigr{)} with At(F)pA^{t}(F)_{p} and Selp(AF)\operatorname{Sel}_{p}(A_{F})^{\vee} respectively.

Proof.

First, since pp is odd and Tp,F(A)T_{p,F}(A) is a (finitely generated) free p[G]\mathbb{Z}_{p}[G]-module the complexes RΓ(𝒪k,S[1p],Tp,F(A))R\Gamma\bigl{(}\mathcal{O}_{k,S}\bigl{[}\tfrac{1}{p}\bigr{]},T_{p,F}(A)\bigr{)} and RΓ(kv,Tp,F(A))R\Gamma\bigl{(}k_{v},T_{p,F}(A)\bigr{)} for each place vv of kk belong to Dp(p[G])D^{\rm p}\bigl{(}\mathbb{Z}_{p}[G]\bigr{)}. In view of the exact triangles (2) and (3), claim (i) will therefore follow if we can show that for each vv in SSpS\cup S_{p} the complex RΓf(kv,Tp,F(A))R\Gamma_{f}\bigl{(}k_{v},T_{p,F}(A)\bigr{)} belongs to Dp(p[G])D^{\rm p}\bigl{(}\mathbb{Z}_{p}[G]\bigr{)}, or equivalently (as PP is a Sylow pp-subgroup) that it belongs to Dp(p[P])D^{\rm p}\bigl{(}\mathbb{Z}_{p}[P]\bigr{)}.

We first consider the case vpv\nmid p. In this case RΓf(kv,Tp,F(A))R\Gamma_{f}\bigl{(}k_{v},T_{p,F}(A)\bigr{)} is isomorphic in D(p[P])D\bigl{(}\mathbb{Z}_{p}[P]\bigr{)} to RΓf(Kw,Tp)R\Gamma_{f}\bigl{(}K_{w^{\prime}},T_{p}\bigr{)} where we write ww^{\prime} for a place in KK below ww and set Tp:=p[P]pTp(A/K)T_{p}:=\mathbb{Z}_{p}[P]\otimes_{\mathbb{Z}_{p}}T_{p}(A_{/K}).

In particular, even without assuming our hypotheses, if the ramification index of vv in F/kF/k is not divisible by pp, then ww^{\prime} is unramified in F/KF/K so TpIwp[P]pTp(A)IwT_{p}^{I_{w^{\prime}}}\cong\mathbb{Z}_{p}[P]\otimes_{\mathbb{Z}_{p}}T_{p}(A)^{I_{w^{\prime}}} is a free p[P]\mathbb{Z}_{p}[P]-module and so RΓf(Kw,Tp)R\Gamma_{f}(K_{w^{\prime}},T_{p}) belongs to Dp(p[P])D^{\rm p}\bigl{(}\mathbb{Z}_{p}[P]\bigr{)}.

Next we show that hypotheses (e) and (f) imply that if ww^{\prime} ramifies in F/kF/k and vpv\nmid p, then RΓf(Kw,Tp)R\Gamma_{f}(K_{w^{\prime}},T_{p}) (and therefore also RΓf(kv,Tp,F(A))R\Gamma_{f}\bigl{(}k_{v},T_{p,F}(A)\bigr{)}) is acyclic. We write d0d^{0} for the differential in degree 0 of RΓf(Kw,Tp)R\Gamma_{f}(K_{w^{\prime}},T_{p}). Then it is clear that ker(d0)=TpGw\ker(d^{0})=T_{p}^{G_{w^{\prime}}} vanishes and so it suffices to prove cok(d0)\operatorname{cok}(d^{0}) vanishes or equivalently, since PP and cok(d0)\operatorname{cok}(d^{0}) are finite pp-groups, that cok(d0)P\operatorname{cok}(d^{0})_{P} vanishes. But (e) implies TpIw=p[P/I¯w]pTp(A/K)T_{p}^{I_{w^{\prime}}}=\mathbb{Z}_{p}[P/\overline{I}_{w^{\prime}}]\otimes_{\mathbb{Z}_{p}}T_{p}(A_{/K}) and so the p\mathbb{Z}_{p}-module cok(d0)P\operatorname{cok}(d^{0})_{P} is isomorphic to the cokernel of the action of 1Frw11-\operatorname{Fr}_{w^{\prime}}^{-1} on Tp(A/K)T_{p}(A_{/K}) and hence of cardinality the maximum power of pp that divides detp(1Frw1|Tp(A/K))\det_{\mathbb{Z}_{p}}\bigl{(}1-\operatorname{Fr}_{w^{\prime}}^{-1}\bigl{|}T_{p}(A_{/K})\bigr{)} in p\mathbb{Z}_{p}. The module cok(d0)P\operatorname{cok}(d^{0})_{P} therefore vanishes since the latter determinant is equal to |At(κw)|/|κw|d|A^{t}(\kappa_{w^{\prime}})|/|\kappa_{w^{\prime}}|^{d} and this is a unit at pp by (f).

Finally we fix a pp-adic place vv of kk and recall from [8, after (3.2)] that the group Hf1(kv,Tp,F(A))H^{1}_{f}\bigl{(}k_{v},T_{p,F}(A)\bigr{)} is the image in H1(kv,Tp,F(A))p[G]p[Gw]H1(Fw,Tp(A))H^{1}\bigl{(}k_{v},T_{p,F}(A)\bigr{)}\cong\mathbb{Z}_{p}[G]\otimes_{\mathbb{Z}_{p}[G_{w}]}H^{1}\bigl{(}F_{w},T_{p}(A)\bigr{)} of p[G]p[Gw]At(Fw)p\mathbb{Z}_{p}[G]\otimes_{\mathbb{Z}_{p}[G_{w}]}A^{t}(F_{w})^{\wedge}_{p} under the natural (injective) Kummer map. Since the group Hf1(kv,Tp,F(A))H^{1}_{f}\bigl{(}k_{v},T_{p,F}(A)\bigr{)} is finitely generated over p\mathbb{Z}_{p} it is thus enough to show that each module At(Fw)pA^{t}(F_{w})^{\wedge}_{p} is cohomologically-trivial over GwG_{w}.

This is clear if the order of GwG_{w} is prime to pp and true in any other case provided that At(Fw)pA^{t}(F_{w})^{\wedge}_{p} is cohomologically-trivial over each subgroup CC of GwG_{w} that has order pp. The point here is that, for a given pp-Sylow subgroup PwP_{w} of GwG_{w}, there exists a normal subgroup CC of PwP_{w} which has order pp and hence also a Hochschild-Serre spectral sequence in Tate cohomology H^a(Pw/C,H^b(C,At(Fw)p))H^a+b(Pw,At(Fw)p)\hat{H}^{a}\bigl{(}P_{w}/C,\hat{H}^{b}(C,A^{t}(F_{w})^{\wedge}_{p})\bigr{)}\Longrightarrow\hat{H}^{a+b}\bigl{(}P_{w},A^{t}(F_{w})^{\wedge}_{p}\bigr{)}. Thus if At(Fw)pA^{t}(F_{w})^{\wedge}_{p} is a cohomologically-trivial CC-module, then H^m(Pw,At(Fw)p)\hat{H}^{m}\bigl{(}P_{w},A^{t}(F_{w})^{\wedge}_{p}\bigr{)} is trivial for all integers mm, and then [12, Chapter VI, (8.7) and (8.8)] combine to imply that At(Fw)pA^{t}(F_{w})^{\wedge}_{p} is a cohomologically-trivial GwG_{w}-module.

We hence fix a subgroup CC of GwG_{w} of order pp. Now cohomology over CC is periodic of order 22 and pAt(Fw)p\mathbb{Q}_{p}\cdot A^{t}(F_{w})^{\wedge}_{p} is isomorphic (via the formal group logarithm) to the free p[Gw]\mathbb{Q}_{p}[G_{w}]-module FwdF^{d}_{w}, so [2, Corollary to Proposition 11] implies that the Herbrand quotient h(C,At(Fw)p)h(C,A^{t}(F_{w})^{\wedge}_{p}) is equal to 1.

It is thus enough to prove that the group H^0(C,At(Fw)p)\hat{H}^{0}\bigl{(}C,A^{t}(F_{w})^{\wedge}_{p}\bigr{)} vanishes, or equivalently that the natural norm map At(Fw)pAt(FwC)pA^{t}(F_{w})^{\wedge}_{p}\to A^{t}(F_{w}^{C})^{\wedge}_{p} is surjective. But the hypotheses (c) and (d) imply that A/FCtA^{t}_{/F^{C}} has good reduction at all pp-adic places and further that at any pp-adic place ww^{\prime} of FCF^{C} which ramifies in F/FCF/F^{C} the reduction is ordinary and such that At(κw)A^{t}(\kappa_{w^{\prime}}) has no pp-torsion, and so the argument of Mazur in [38, §4] implies that the norm map At(Fw)pAt(FwC)pA^{t}(F_{w})^{\wedge}_{p}\to A^{t}(F_{w}^{C})^{\wedge}_{p} is surjective, as required to complete the proof of claim (i). (We note in passing that if pp is unramified in F/F/\mathbb{Q}, as will be the case in applications of Lemma 4.1 in the present article, then the relevant argument is given by [38, Corollary 4.4].)

Turning to claim (ii) we note that, under the hypothesis (g), the argument of [15, p. 86-87] implies directly that Hfi(k,Tp,F(A))H^{i}_{f}\bigl{(}k,T_{p,F}(A)\bigr{)} vanishes if i{1,2,3}i\notin\{1,2,3\}, that Hf3(k,Tp,F(A))H^{3}_{f}\bigl{(}k,T_{p,F}(A)\bigr{)} is isomorphic to (A(F)p,tor)(A(F)_{p,{\rm tor}})^{\vee} and that there is a canonical exact sequence of p[G]\mathbb{Z}_{p}[G]-modules

0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Hf1(k,Tp,F(A))\textstyle{H^{1}_{f}\bigl{(}k,T_{p,F}(A)\bigr{)}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}At(F)p\textstyle{A^{t}(F)_{p}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}vSrSbH0(kv,H1(Iv,Tp,F(A))tor)\textstyle{{\bigoplus\limits_{v\in S_{\rm r}\cup S_{\rm b}}}H^{0}\Bigl{(}k_{v},H^{1}\bigl{(}I_{v},T_{p,F}(A)\bigr{)}_{\rm tor}\Bigr{)}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Hf2(k,Tp,F(A))\textstyle{H^{2}_{f}\bigl{(}k,T_{p,F}(A)\bigr{)}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Selp(AF)\textstyle{\operatorname{Sel}_{p}(A_{F})^{\vee}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0.\textstyle{0.}

Now hypothesis (a) implies that the pp-group PP acts on A(F)[p]A(F)[p] with a single fixed point and hence that Hf3(k,Tp,F(A))H^{3}_{f}\bigl{(}k,T_{p,F}(A)\bigr{)} vanishes. Claim (ii) will therefore follow if we can show that, under the stated hypotheses, the group H0(kv,H1(Iv,Tp,F(A))tor)H^{0}\bigl{(}k_{v},H^{1}(I_{v},T_{p,F}(A))_{\rm tor}\bigr{)} vanishes for all vSrSbv\in S_{\rm r}\cup S_{\rm b}. In view of the natural isomorphism

H0(kv,H1(Iv,Tp,F(A))tor)p[G]p[Gw]H0(Fw,H1(Iw,Tp(A))tor)H^{0}\Bigl{(}k_{v},H^{1}\bigl{(}I_{v},T_{p,F}(A)\bigr{)}_{\rm tor}\Bigr{)}\cong\mathbb{Z}_{p}[G]\otimes_{\mathbb{Z}_{p}[G_{w}]}H^{0}\Bigl{(}F_{w},H^{1}\bigl{(}I_{w},T_{p}(A)\bigr{)}_{\rm tor}\Bigr{)}

it is thus enough to show that H0(Fw,H1(Iw,Tp(A))tor)H^{0}\bigl{(}F_{w},H^{1}(I_{w},T_{p}(A))_{\rm tor}\bigr{)} vanishes for all vSrSbv\in S_{\rm r}\cup S_{\rm b}. On the one hand, if vSbv\notin S_{\rm b}, then H1(Iw,Tp(A))=Homcont(Iw,Tp(A))H^{1}\bigl{(}I_{w},T_{p}(A)\bigr{)}=\operatorname{Hom}_{\rm cont}\bigl{(}I_{w},T_{p}(A)\bigr{)} is torsion-free and so the result is clear. On the other hand, if vSbv\in S_{\rm b}, then (e) implies that vSrv\notin S_{\rm r} so Iw=IwI_{w}=I_{w^{\prime}} and hence the group

H0(Fw,H1(Iw,Tp(A))tor)G¯w=H0(Kw,H1(Iw,Tp(A))tor)H^{0}\Bigl{(}F_{w},H^{1}\bigl{(}I_{w},T_{p}(A)\bigr{)}_{\rm tor}\Bigr{)}^{\overline{G}_{w^{\prime}}}=H^{0}\Bigl{(}K_{w^{\prime}},H^{1}\bigl{(}I_{w^{\prime}},T_{p}(A)\bigr{)}_{\rm tor}\Bigr{)}

vanishes as a consequence of (b) and the fact that H0(Kw,H1(Iw,Tp(A))tor)H^{0}\Bigl{(}K_{w^{\prime}},H^{1}\bigl{(}I_{w^{\prime}},T_{p}(A)\bigr{)}_{\rm tor}\Bigr{)} has cardinality equal to the maximal power of pp that divides the Tamagawa number of A/KtA^{t}_{/K} at ww^{\prime}. Since G¯w\overline{G}_{w^{\prime}} is a pp-group, this implies that H0(Fw,H1(Iw,Tp(A))tor)H^{0}\bigl{(}F_{w},H^{1}(I_{w},T_{p}(A))_{\rm tor}\bigr{)} vanishes, as required to complete the proof of claim (ii). ∎

In the sequel we assume that AA, FF and pp satisfy the hypotheses (a)(g). Then Lemma 4.1 implies that the complex

CA,Ff,:=RΓf(k,Tp,F(A))C^{f,\bullet}_{A,F}:=R\Gamma_{f}\bigl{(}k,T_{p,F}(A)\bigr{)}

belongs to Dp(p[G])D^{\rm p}\bigl{(}\mathbb{Z}_{p}[G]\bigr{)}, is acyclic outside degrees one and two and is such that for each isomorphism of fields j:pj:\mathbb{C}\cong\mathbb{C}_{p} there exists a canonical composite isomorphism of p[G]\mathbb{C}_{p}[G]-modules of the form

λA,FNT,j:ppH1(CA,Ff,)ppAt(F)pp,j(At(F))p,jHom(A(F),)ppHomp(A(F)p,p)ppH2(CA,Ff,)\lambda^{{\rm NT},j}_{A,F}\colon\mathbb{C}_{p}\otimes_{\mathbb{Z}_{p}}H^{1}\bigl{(}C^{f,\bullet}_{A,F}\bigr{)}\cong\mathbb{C}_{p}\otimes_{\mathbb{Z}_{p}}A^{t}(F)_{p}\\ \cong\mathbb{C}_{p}\otimes_{\mathbb{C},j}(\mathbb{C}\otimes_{\mathbb{Z}}A^{t}(F))\cong\mathbb{C}_{p}\otimes_{\mathbb{C},j}\operatorname{Hom}_{\mathbb{C}}\bigl{(}\mathbb{C}\otimes_{\mathbb{Z}}A(F),\mathbb{C}\bigr{)}\\ \cong\mathbb{C}_{p}\otimes_{\mathbb{Z}_{p}}\operatorname{Hom}_{\mathbb{Z}_{p}}\bigl{(}A(F)_{p},\mathbb{Z}_{p}\bigr{)}\cong\mathbb{C}_{p}\otimes_{\mathbb{Z}_{p}}H^{2}\bigl{(}C^{f,\bullet}_{A,F}\bigr{)}

in which the central isomorphism is induced by the \mathbb{C}-linear extension to A(F)\mathbb{C}\otimes_{\mathbb{Z}}A(F) of the Néron-Tate height of AA relative to the field FF.

Hypothesis (a) implies that the pp-group PP acts on A(F)[p]A(F)[p] with a single fixed point and hence that the module At(F)pH1(CA,Ff,)A^{t}(F)_{p}\cong H^{1}(C^{f,\bullet}_{A,F}) is p\mathbb{Z}_{p}-free and so the above observations imply that the construction in §4.1 applies to the pair (CA,Ff,,λA,FNT,j)\bigl{(}C^{f,\bullet}_{A,F},\lambda^{{\rm NT},j}_{A,F}\bigr{)} to give a canonical Euler characteristic

χj(A,F/k):=χG,p(CA,Ff,,λA,FNT,j)\chi_{j}(A,F/k):=-\chi_{G,p}\bigl{(}C^{f,\bullet}_{A,F},\lambda^{{\rm NT},j}_{A,F}\bigr{)}

in the relative algebraic KK-group K0(p[G],p[G])K_{0}\bigl{(}\mathbb{Z}_{p}[G],\mathbb{C}_{p}[G]\bigr{)}.

This element χj(A,F/k)\chi_{j}(A,F/k) encodes detailed information about a range of aspects of the arithmetic of A/FA_{/F} and in the remainder of §4 we shall explicitly relate it to the equivariant Tamagawa numbers that are defined in [16]. In §5 this comparison result plays a key role in the formulation of a precise conjectural description of a pre-image of χj(A,F/k)\chi_{j}(A,F/k) under the boundary homomorphism K1(p[G])K0(p[G],p[G])K_{1}\bigl{(}\mathbb{C}_{p}[G]\bigr{)}\to K_{0}\bigl{(}\mathbb{Z}_{p}[G],\mathbb{C}_{p}[G]\bigr{)}.

4.3. The element RΩj(h1(A/F)(1),[G])R\Omega_{j}\bigl{(}h^{1}(A_{/F})(1),\mathbb{Z}[G]\bigr{)}.

For each isomorphism j:pj:\mathbb{C}\cong\mathbb{C}_{p} as above there is an induced composite homomorphism of abelian groups

(4) jG,:K0([G],[G])K0([G],[G])K0([G],p[G])K0(p[G],p[G])j_{G,*}:K_{0}\bigl{(}\mathbb{Z}[G],\mathbb{R}[G]\bigr{)}\to K_{0}\bigl{(}\mathbb{Z}[G],\mathbb{C}[G]\bigr{)}\cong K_{0}\bigl{(}\mathbb{Z}[G],\mathbb{C}_{p}[G]\bigr{)}\to K_{0}\bigl{(}\mathbb{Z}_{p}[G],\mathbb{C}_{p}[G]\bigr{)}

(where the first and third arrows are induced by the inclusions [G][G]\mathbb{R}[G]\subset\mathbb{C}[G] and [G]p[G]\mathbb{Z}[G]\subset\mathbb{Z}_{p}[G] respectively).

For each such jj we set

RΩj(h1(A/F)(1),[G]):=jG,(RΩ(h1(A/F)(1),[G]))R\Omega_{j}\bigl{(}h^{1}(A_{/F})(1),\mathbb{Z}[G]\bigr{)}:=j_{G,*}\Bigl{(}R\Omega\bigl{(}h^{1}(A_{/F})(1),\mathbb{Z}[G]\bigr{)}\Bigr{)}

where RΩ(h1(A/F)(1),[G])R\Omega\bigl{(}h^{1}(A_{/F})(1),\mathbb{Z}[G]\bigr{)} is the ‘algebraic part’ of the equivariant Tamagawa number for the pair (h1(A/F)(1),[G])\bigl{(}h^{1}(A_{/F})(1),\mathbb{Z}[G]\bigr{)}, as defined (unconditionally under the assumed validity of hypothesis (g)) in [16, §3.4].

In order to relate this element to the Euler characteristic χj(A,F/k)\chi_{j}(A,F/k) defined above we must first introduce an auxiliary element of K0(p[G],p[G])K_{0}\bigl{(}\mathbb{Z}_{p}[G],\mathbb{C}_{p}[G]\bigr{)}.

To do this we define a complex

CA,Floc,:=vSRΓ(kv,Tp,F(A))vSSpRΓf(kv,Tp,F(A)),C^{{\rm loc},\bullet}_{A,F}:=\bigoplus_{v\in S_{\infty}}R\Gamma\bigl{(}k_{v},T_{p,F}(A)\bigr{)}\oplus\bigoplus_{v\in S\cup S_{p}}R\Gamma_{f}\bigl{(}k_{v},T_{p,F}(A)\bigr{)},

where, as before, SS denotes SrSbS_{\rm r}\cup S_{\rm b}. It is then clear that CA,Floc,C^{{\rm loc},\bullet}_{A,F} is acyclic outside degrees zero and one, that H0(CA,Floc,)=vSH0(kv,Tp,F(A))H^{0}\bigl{(}C^{{\rm loc},\bullet}_{A,F}\bigr{)}=\bigoplus_{v\in S_{\infty}}H^{0}(k_{v},T_{p,F}(A)) and that there is a canonical identification of ppH1(CA,Floc,)\mathbb{Q}_{p}\otimes_{\mathbb{Z}_{p}}H^{1}\bigl{(}C^{{\rm loc},\bullet}_{A,F}\bigr{)} with ppAt(Fp)p\mathbb{Q}_{p}\otimes_{\mathbb{Z}_{p}}A^{t}(F_{p})^{\wedge}_{p}, where we set Fp:=FpF_{p}:=F\otimes_{\mathbb{Q}}\mathbb{Q}_{p}.

We next use this description of the cohomology of CA,Floc,C^{{\rm loc},\bullet}_{A,F} to define a canonical isomorphism of p[G]\mathbb{C}_{p}[G]-modules

λA,Fexp,j:ppH0(CA,Floc,)ppH1(CA,Floc,).\lambda_{A,F}^{{\rm exp},j}:\mathbb{C}_{p}\otimes_{\mathbb{Z}_{p}}H^{0}\bigl{(}C^{{\rm loc},\bullet}_{A,F}\bigr{)}\cong\mathbb{C}_{p}\otimes_{\mathbb{Z}_{p}}H^{1}\bigl{(}C^{{\rm loc},\bullet}_{A,F}\bigr{)}.

For this we write Σ(k)\Sigma(k) for the set of embeddings kk\to\mathbb{C} and Σσ(F)\Sigma_{\sigma}(F), for each σ\sigma in Σ(k)\Sigma(k), for the set of embeddings FF\to\mathbb{C} that extend σ\sigma. For vv in SS_{\infty} we fix a corresponding element σv\sigma_{v} of Σ(k)\Sigma(k) and consider the \mathbb{C}-linear map

πv:H1(σv(At)(),)Hom(k,σvH0(At,ΩAt1),)=k,σvHomk(H0(At,ΩAt1),k)\pi_{v}\colon\ \mathbb{C}\otimes_{\mathbb{Z}}H_{1}\bigl{(}\sigma_{v}(A^{t})(\mathbb{C}),\mathbb{Z}\bigr{)}\\ \longrightarrow\operatorname{Hom}_{\mathbb{C}}\Bigl{(}\mathbb{C}\otimes_{k,\sigma_{v}}H^{0}\bigl{(}A^{t},\Omega^{1}_{A^{t}}\bigr{)},\,\mathbb{C}\Bigr{)}=\mathbb{C}\otimes_{k,\sigma_{v}}\operatorname{Hom}_{k}\bigl{(}H^{0}(A^{t},\Omega^{1}_{A^{t}}),k\bigr{)}

that sends the image γ\gamma in H1(σv(At)(),)H_{1}\bigl{(}\sigma_{v}(A^{t})(\mathbb{C}),\mathbb{Z}\bigr{)} of a cycle γ^\hat{\gamma} to the map induced by sending a differential ω\omega to γω:=γ^ω\int_{\gamma}\omega:=\int_{\hat{\gamma}}\omega.

For each place vv in SS_{\infty} we also write Yv,FY_{v,F} for the module Σσv(F)\prod_{\Sigma_{\sigma_{v}}(F)}\mathbb{Z}, endowed with its natural action of the direct product G×GvG\times G_{v}, and then define λA,Fexp,j\lambda_{A,F}^{{\rm exp},j} to be the following composite isomorphism of p[G]\mathbb{C}_{p}[G]-modules

ppH0(CA,Floc,)=\displaystyle\mathbb{C}_{p}\otimes_{\mathbb{Z}_{p}}H^{0}\bigl{(}C^{{\rm loc},\bullet}_{A,F}\bigr{)}= ppvSH0(kv,Tp,F(A))\displaystyle\,\mathbb{C}_{p}\otimes_{\mathbb{Z}_{p}}\bigoplus_{v\in S_{\infty}}H^{0}\bigl{(}k_{v},T_{p,F}(A)\bigr{)}
\displaystyle\cong pvSH0(kv,Yv,FH1(σv(At)(),))\displaystyle\,\mathbb{C}_{p}\otimes_{\mathbb{Z}}\bigoplus_{v\in S_{\infty}}H^{0}\Bigl{(}k_{v},Y_{v,F}\otimes_{\mathbb{Z}}H_{1}(\sigma_{v}(A^{t})(\mathbb{C}),\mathbb{Z})\Bigr{)}
\displaystyle\cong p(FkHomk(H0(At,ΩAt1),k))\displaystyle\,\mathbb{C}_{p}\otimes_{\mathbb{Q}}\Bigl{(}F\otimes_{k}\operatorname{Hom}_{k}\bigl{(}H^{0}(A^{t},\Omega^{1}_{A^{t}}),k\bigr{)}\Bigr{)}
\displaystyle\cong ppAt(Fp)p\displaystyle\,\mathbb{C}_{p}\otimes_{\mathbb{Z}_{p}}A^{t}(F_{p})^{\wedge}_{p}
=\displaystyle= ppH1(CA,Floc,).\displaystyle\,\mathbb{C}_{p}\otimes_{\mathbb{Z}_{p}}H^{1}\bigl{(}C^{{\rm loc},\bullet}_{A,F}\bigr{)}.

Here the first isomorphism is induced by the canonical comparison isomorphisms

(5) Tp,F(A)=YF/k,ppTp(A)Yv,F,ppH1(σv(At)(),)p,T_{p,F}(A)=Y_{F/k,p}\otimes_{\mathbb{Z}_{p}}T_{p}(A)\cong Y_{v,F,p}\otimes_{\mathbb{Z}_{p}}H_{1}\bigl{(}\sigma_{v}(A^{t})(\mathbb{C}),\mathbb{Z}\bigr{)}_{p},

the second by the maps p,jπv\mathbb{C}_{p}\otimes_{\mathbb{C},j}\pi_{v} and the canonical decomposition

(6) (Fk\displaystyle\mathbb{R}\otimes_{\mathbb{Q}}\Bigl{(}F\otimes_{k} Homk(H0(At,ΩAt1),k))\displaystyle\operatorname{Hom}_{k}\bigl{(}H^{0}(A^{t},\Omega^{1}_{A^{t}}),k\bigr{)}\Bigr{)}
\displaystyle\cong vSH0(kv,(k,σF)(k,σHomk(H0(At,ΩAt1),k)))\displaystyle\bigoplus_{v\in S_{\infty}}H^{0}\Bigl{(}k_{v},\bigl{(}\mathbb{C}\otimes_{k,\sigma}F\bigr{)}\otimes_{\mathbb{C}}\bigl{(}\mathbb{C}\otimes_{k,\sigma}\operatorname{Hom}_{k}(H^{0}(A^{t},\Omega^{1}_{A^{t}}),k)\bigr{)}\Bigr{)}
\displaystyle\cong vSH0(kv,Yv,F(k,σHomk(H0(At,ΩAt1),k))),\displaystyle\bigoplus_{v\in S_{\infty}}H^{0}\Bigl{(}k_{v},\,\mathbb{C}Y_{v,F}\otimes_{\mathbb{C}}\bigl{(}\mathbb{C}\otimes_{k,\sigma}\operatorname{Hom}_{k}\bigl{(}H^{0}(A^{t},\Omega^{1}_{A^{t}}),k\bigr{)}\bigr{)}\Bigr{)},

and the third by the sum over places vv in SpS_{p} of the classical exponential map

(7) FvkHomk(H0(At,ΩAt1),k)HomFv(H0(A/Fvt,ΩA/Fvt1),Fv)pAt(Fv)pF_{v}\otimes_{k}\operatorname{Hom}_{k}\bigl{(}H^{0}(A^{t},\Omega^{1}_{A^{t}}),k\bigr{)}\cong\operatorname{Hom}_{F_{v}}\bigl{(}H^{0}(A^{t}_{/F_{v}},\Omega^{1}_{A^{t}_{/F_{v}}}),F_{v}\bigr{)}\cong\mathbb{Q}_{p}A^{t}(F_{v})^{\wedge}_{p}

where we set Fv:=FkkvF_{v}:=F\otimes_{k}k_{v}.

Under our stated hypotheses the complex CA,Floc,C^{{\rm loc},\bullet}_{A,F} belongs to Dp(p[G])D^{\rm p}\bigl{(}\mathbb{Z}_{p}[G]\bigr{)}. It is also acyclic outside degrees zero and one and such that H0(CA,Floc,)H^{0}\bigl{(}C^{{\rm loc},\bullet}_{A,F}\bigr{)} is p\mathbb{Z}_{p}-free and so we may use the construction of §4.1 to define an element of K0(p[G],p[G])K_{0}\bigl{(}\mathbb{Z}_{p}[G],\mathbb{C}_{p}[G]\bigr{)} by setting

χjloc(A,F/k):=χG,p(CA,Floc,,λA,Fexp,j).\chi_{j}^{\rm loc}(A,F/k):=\chi_{G,p}\Bigl{(}C^{{\rm loc},\bullet}_{A,F},\,\lambda_{A,F}^{{\rm exp},j}\Bigr{)}.

We are now ready to state the main result of this section. In this result, and the sequel, we shall use the composite homomorphism

δG,p:ζ(p[G])×K1(p[G])K0(p[G],p[G])\delta_{G,p}:\zeta\bigl{(}\mathbb{C}_{p}[G]\bigr{)}^{\times}\to K_{1}\bigl{(}\mathbb{C}_{p}[G]\bigr{)}\to K_{0}\bigl{(}\mathbb{Z}_{p}[G],\mathbb{C}_{p}[G]\bigr{)}

where the first arrow is the inverse of the (bijective) reduced norm homomorphism Nrdp[G]\operatorname{Nrd}_{\mathbb{C}_{p}[G]} and the second is the standard boundary homomorphism G,p\partial_{G,p}.

Proposition 4.2.

Assume that AA and FF satisfy the hypotheses (a)(g). Then in K0(p[G],p[G])K_{0}\bigl{(}\mathbb{Z}_{p}[G],\mathbb{C}_{p}[G]\bigr{)} one has

RΩj(h1(A/F)(1),[G])=χj(A,F/k)χjloc(A,F/k)+vSSpδG,p(Lv(A,F/k)).R\Omega_{j}\bigl{(}h^{1}(A_{/F})(1),\mathbb{Z}[G]\bigr{)}=-\chi_{j}(A,F/k)-\chi_{j}^{\rm loc}(A,F/k)+\sum_{v\in S\cup S_{p}}\delta_{G,p}\bigl{(}L_{v}(A,F/k)\bigr{)}.

Here for each place vv in SSpS\cup S_{p} we set

Lv(A,F/k):={Nrdp[G](1Frv1|Vp,F(A)Iv),if vpNrdp[G](1φv|Dcr,v(Vp,F(A))),if vp,L_{v}(A,F/k):=\begin{cases}\operatorname{Nrd}_{\mathbb{Q}_{p}[G]}\bigl{(}1-\operatorname{Fr}_{v}^{-1}\bigm{|}V_{p,F}(A)^{I_{v}}\bigr{)},&\text{if $v\nmid p$}\\ \operatorname{Nrd}_{\mathbb{Q}_{p}[G]}\bigl{(}1-\varphi_{v}\bigm{|}D_{{\rm cr},v}(V_{p,F}(A))\bigr{)},&\text{if $v\mid p$,}\end{cases}

where φv\varphi_{v} is the crystalline Frobenius at vv.

Proof.

To discuss RΩj(h1(A/F)(1),[G])R\Omega_{j}\bigl{(}h^{1}(A_{/F})(1),\mathbb{Z}[G]\bigr{)} we first recall relevant facts concerning the formalism of virtual objects introduced by Deligne in [24] (for more details see [16, §2]).

With RR denoting either p[G]\mathbb{Z}_{p}[G] or p[G]\mathbb{C}_{p}[G] we write V(R)V(R) for the category of virtual objects over RR and [C]R[C]_{R} for the object of V(R)V(R) associated to each CC in Dp(R)D^{\rm p}(R). Then V(R)V(R) is a Picard category with π1(V(R))\pi_{1}(V(R)) naturally isomorphic to K1(R)K_{1}(R) (see [16, (2)]) and we write (X,Y)XY(X,Y)\mapsto X\cdot Y for its product and 𝟏R\boldsymbol{1}_{R} for the unit object [0]R[0]_{R}. Writing 𝒫0\mathcal{P}_{0} for the Picard category with unique object 𝟏𝒫0\boldsymbol{1}_{\mathcal{P}_{0}} and Aut𝒫0(𝟏𝒫0)=0\operatorname{Aut}_{\mathcal{P}_{0}}(\boldsymbol{1}_{\mathcal{P}_{0}})=0 we use the isomorphism of abelian groups

(8) π0(V(p[G])×V(p[G])𝒫0)K0(p[G],p[G])\pi_{0}\Bigl{(}V\bigl{(}\mathbb{Z}_{p}[G]\bigr{)}\times_{V(\mathbb{C}_{p}[G])}\mathcal{P}_{0}\Bigr{)}\cong K_{0}\bigl{(}\mathbb{Z}_{p}[G],\mathbb{C}_{p}[G]\bigr{)}

that is described in [16, Proposition 2.5]. In particular, via this isomorphism, each pair comprising an object XX of V(p[G])V\bigl{(}\mathbb{Z}_{p}[G]\bigr{)} and a morphism ι:p[G]p[G]X𝟏p[G]\iota:\mathbb{C}_{p}[G]\otimes_{\mathbb{Z}_{p}[G]}X\to\boldsymbol{1}_{\mathbb{C}_{p}[G]} gives rise to a canonical element [X,ι][X,\iota] of K0(p[G],p[G])K_{0}\bigl{(}\mathbb{Z}_{p}[G],\mathbb{C}_{p}[G]\bigr{)}.

Now if CC belongs to Dp(p[G])D^{\rm p}\bigl{(}\mathbb{Z}_{p}[G]\bigr{)} and is acyclic outside degrees aa and a+1a+1 (for any integer aa), then any isomorphism λ:Ha(ppC)Ha+1(ppC)\lambda:H^{a}\bigl{(}\mathbb{C}_{p}\otimes_{\mathbb{Z}_{p}}C\bigr{)}\cong H^{a+1}\bigl{(}\mathbb{C}_{p}\otimes_{\mathbb{Z}_{p}}C\bigr{)} of p[G]\mathbb{C}_{p}[G]-modules gives a canonical morphism λTriv:[ppC]p[G]𝟏p[G]\lambda_{\rm Triv}:[\mathbb{C}_{p}\otimes_{\mathbb{Z}_{p}}C]_{\mathbb{C}_{p}[G]}\to\boldsymbol{1}_{\mathbb{C}_{p}[G]} in V(p[G])V\bigl{(}\mathbb{C}_{p}[G]\bigr{)}. The associated element [[C]p[G],λTriv]\bigl{[}[C]_{\mathbb{Z}_{p}[G]},\lambda_{{\rm Triv}}\bigr{]} of K0(p[G],p[G])K_{0}\bigl{(}\mathbb{Z}_{p}[G],\mathbb{C}_{p}[G]\bigr{)} coincides with the Euler characteristic χp[G],p[G](C,λ(1)a)\chi_{\mathbb{Z}_{p}[G],\mathbb{C}_{p}[G]}\bigl{(}C,\lambda^{(-1)^{a}}\bigr{)} defined in [11, Definition 5.5]. In particular, from Proposition 5.6(3) in loc. cit. it follows that [[C[1]]p[G],λTriv]=[[C]p[G],λTriv]\bigl{[}[C[-1]]_{\mathbb{Z}_{p}[G]},\lambda_{{\rm Triv}}\bigr{]}=-\bigl{[}[C]_{\mathbb{Z}_{p}[G]},\lambda_{{\rm Triv}}\bigr{]}, whilst Theorem 6.2 and Lemma 6.3 in loc. cit. combine to imply that if Ha(C)H^{a}(C) is p\mathbb{Z}_{p}-free, then

(9) [[C]p[G],λTriv]=χG,p(C,λ)+δG,p(i1,2mod4Nrdp[G](id|ppHi(C))(1)i),\bigl{[}[C]_{\mathbb{Z}_{p}[G]},\lambda_{{\rm Triv}}\bigr{]}=\chi_{G,p}(C,\lambda)+\delta_{G,p}\biggl{(}\prod_{i\equiv 1,2\bmod{4}}\operatorname{Nrd}_{\mathbb{Q}_{p}[G]}\bigl{(}-{\mathrm{id}}\bigm{|}\mathbb{Q}_{p}\otimes_{\mathbb{Z}_{p}}H^{i}(C)\bigr{)}^{(-1)^{i}}\biggr{)},

where χG,p(,)\chi_{G,p}(-,-) is the explicit Euler characteristic discussed in §4.1.

We now set

CA,Fc,:=RΓc(𝒪k,S[1p],Tp,F(A)),C^{c,\bullet}_{A,F}:=R\Gamma_{c}\bigl{(}\mathcal{O}_{k,S}[\tfrac{1}{p}],T_{p,F}(A)\bigr{)},

which is the complex of the compactly supported étale cohomology of Tp,F(A)T_{p,F}(A) on Spec(𝒪k,S[1p])\operatorname{Spec}\bigl{(}\mathcal{O}_{k,S}[\tfrac{1}{p}]\bigr{)}, regarded as a complex of p[G]\mathbb{Z}_{p}[G]-modules in the natural way.

Then a direct comparison of the definitions of CA,Fc,C^{c,\bullet}_{A,F} and CA,Ff,C^{f,\bullet}_{A,F} shows that there is a canonical exact triangle in Dp(p[G])D^{\rm p}\bigl{(}\mathbb{Z}_{p}[G]\bigr{)}

(10) CA,Floc,[1]\textstyle{C^{{\rm loc},\bullet}_{A,F}[-1]\ignorespaces\ignorespaces\ignorespaces\ignorespaces}CA,Fc,\textstyle{C^{c,\bullet}_{A,F}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}CA,Ff,\textstyle{C^{f,\bullet}_{A,F}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}CA,Floc,.\textstyle{C^{{\rm loc},\bullet}_{A,F}.}

We consider the following diagram in V(p[G])V\bigl{(}\mathbb{C}_{p}[G]\bigr{)}

(11) [ppCA,Fc,]p[G]\textstyle{\bigl{[}\mathbb{C}_{p}\otimes_{\mathbb{Z}_{p}}C^{c,\bullet}_{A,F}\bigr{]}_{\mathbb{C}_{p}[G]}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Δ\scriptstyle{\Delta}α\scriptstyle{\alpha}[ppCA,Floc,[1]]p[G][ppCA,Ff,]p[G]\textstyle{\bigl{[}\mathbb{C}_{p}\otimes_{\mathbb{Z}_{p}}C^{{\rm loc},\bullet}_{A,F}[-1]\bigr{]}_{\mathbb{C}_{p}[G]}\cdot\bigl{[}\mathbb{C}_{p}\otimes_{\mathbb{Z}_{p}}C^{f,\bullet}_{A,F}\bigr{]}_{\mathbb{C}_{p}[G]}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}[(λA,Fexp,j)Triv]p[G][(λA,FNT,j)Triv]p[G]\scriptstyle{\bigl{[}(\lambda_{A,F}^{{\rm exp},j})_{\rm Triv}\bigr{]}_{\mathbb{C}_{p}[G]}\cdot\bigl{[}(\lambda^{{\rm NT},j}_{A,F})_{\rm Triv}\bigr{]}_{\mathbb{C}_{p}[G]}}𝟏p[G]\textstyle{\boldsymbol{1}_{\mathbb{C}_{p}[G]}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ι\scriptstyle{\iota}α\scriptstyle{\alpha^{\prime}}𝟏p[G]𝟏p[G]\textstyle{\boldsymbol{1}_{\mathbb{C}_{p}[G]}\cdot\boldsymbol{1}_{\mathbb{C}_{p}[G]}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}(vSSpLv(A,F/k))id\scriptstyle{\bigl{(}\prod_{v\in S\cup S_{p}}L_{v}(A,F/k)\bigr{)}\cdot{\mathrm{id}}}𝟏p[G]\textstyle{\boldsymbol{1}_{\mathbb{C}_{p}[G]}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ι\scriptstyle{\iota}𝟏p[G]𝟏p[G].\textstyle{\boldsymbol{1}_{\mathbb{C}_{p}[G]}\cdot\boldsymbol{1}_{\mathbb{C}_{p}[G]}.}

Here Δ\Delta is the morphism induced by the scalar extension of (10), ι\iota denotes the canonical identification and the morphisms α\alpha and α\alpha^{\prime} are defined by the condition that the two squares commute.

We claim that this definition of α\alpha and α\alpha^{\prime} implies that

(12) αα=(p,jϑ)(ppϑp)1\alpha^{\prime}\circ\alpha=(\mathbb{C}_{p}\otimes_{\mathbb{R},j}\vartheta_{\infty})\circ(\mathbb{C}_{p}\otimes_{\mathbb{Q}_{p}}\vartheta_{p})^{-1}

where the morphisms ϑ\vartheta_{\infty} and ϑp\vartheta_{p} are as constructed in [16, §3.4]. To verify this we note that the scalar extension of (10) is naturally isomorphic to the exact triangle in Dp(p[G])D^{\rm p}\bigl{(}\mathbb{C}_{p}[G]\bigr{)} induced by the central column of the diagram [16, (26)] and then simply compare the explicit definitions of the morphisms λA,Fexp,j\lambda_{A,F}^{{\rm exp},j} and ϑp\vartheta_{p} and of λA,FNT,j\lambda^{{\rm NT},j}_{A,F} and ϑ\vartheta_{\infty}. After this it only remains to note the following fact. For each place vSSpv\in S\setminus S_{p}, respectively vSpv\in S_{p}, we write Vp,vV_{p,v} and ϕv\phi_{v} for Vp,F(A)IvV_{p,F}(A)^{I_{v}} and Frv1\operatorname{Fr}_{v}^{-1}, respectively Dcr,v(Vp,F(A))D_{{\rm cr},v}\bigl{(}V_{p,F}(A)\bigr{)} and φv\varphi_{v}, and then Vp,vV_{p,v}^{\bullet} for the complex Vp,v1ϕvVp,vV_{p,v}\xrightarrow{1-\phi_{v}}V_{p,v}, with the first term placed in degree zero. Our definition of λA,Fexp,j\lambda_{A,F}^{{\rm exp},j} implicitly uses the morphism [Vp,v]p[G]𝟏p[G][V^{\bullet}_{p,v}]_{\mathbb{Q}_{p}[G]}\to\boldsymbol{1}_{\mathbb{Q}_{p}[G]} induced by the acyclicity of Vp,vV^{\bullet}_{p,v} whereas the definition of ϑp\vartheta_{p} uses (via [16, (19) and (22)]) the morphism [Vp,v]p[G]𝟏p[G][V^{\bullet}_{p,v}]_{\mathbb{Q}_{p}[G]}\to\boldsymbol{1}_{\mathbb{Q}_{p}[G]} induced by the identity map on Vp,vV_{p,v}; the occurrence of the morphism α\alpha^{\prime} in the equality (12) is thus accounted for by applying the remark made immediately after [16, (24)] to each of the complexes Vp,vV^{\bullet}_{p,v} and noting that Nrdp[G](1ϕvVp,v)=Lv(A,F/k)\operatorname{Nrd}_{\mathbb{Q}_{p}[G]}(1-\phi_{v}\mid V_{p,v})=L_{v}(A,F/k).

Now, taking into account the equality (12), the term RΩj(h1(A/F)(1),[G])R\Omega_{j}\bigl{(}h^{1}(A_{/F})(1),\mathbb{Z}[G]\bigr{)} is defined in [16] to be equal to

[[CA,Fc,]p[G],(p,jϑ)(ppϑp)1]=[[CA,Fc,]p[G],αα].\Bigl{[}\bigl{[}C^{c,\bullet}_{A,F}\bigr{]}_{\mathbb{Z}_{p}[G]},\,(\mathbb{C}_{p}\otimes_{\mathbb{R},j}\vartheta_{\infty})\circ(\mathbb{C}_{p}\otimes_{\mathbb{Q}_{p}}\vartheta_{p})^{-1}\Bigr{]}=\Bigl{[}\bigl{[}C^{c,\bullet}_{A,F}\bigr{]}_{\mathbb{Z}_{p}[G]},\,\alpha^{\prime}\circ\alpha\Bigr{]}.

The product structure of π0(V(p[G])×V(p[G])𝒫0)\pi_{0}\bigl{(}V(\mathbb{Z}_{p}[G])\times_{V(\mathbb{C}_{p}[G])}\mathcal{P}_{0}\bigr{)} then combines with the commutativity of (11) to imply that it is also equal to

[[C\displaystyle\Bigl{[}\bigl{[}C ]p[G]A,Fc,,α]+[𝟏p[G],α]{}^{c,\bullet}_{A,F}\bigr{]}_{\mathbb{Z}_{p}[G]},\,\alpha\Bigr{]}+\bigl{[}\boldsymbol{1}_{\mathbb{Z}_{p}[G]},\,\alpha^{\prime}\bigr{]}
=[[CA,Floc,[1]]p[G],(λA,Fexp,j)Triv]+[[CA,Ff,]p[G],(λA,FNT,j)Triv]+[𝟏p[G],α]\displaystyle=\Bigl{[}\bigl{[}C^{{\rm loc},\bullet}_{A,F}[-1]\bigr{]}_{\mathbb{Z}_{p}[G]},\,\bigl{(}\lambda_{A,F}^{{\rm exp},j}\bigr{)}_{\rm Triv}\Bigr{]}+\Bigl{[}\bigl{[}C^{f,\bullet}_{A,F}\bigr{]}_{\mathbb{Z}_{p}[G]},\,\bigl{(}\lambda^{{\rm NT},j}_{A,F}\bigr{)}_{\rm Triv}\Bigr{]}+\bigl{[}\boldsymbol{1}_{\mathbb{Z}_{p}[G]},\,\alpha^{\prime}\bigr{]}
=[[CA,Floc,]p[G],(λA,Fexp,j)Triv]+[[CA,Ff,]p[G],(λA,FNT,j)Triv]\displaystyle=-\Bigl{[}\bigl{[}C^{{\rm loc},\bullet}_{A,F}\bigr{]}_{\mathbb{Z}_{p}[G]},\,\bigl{(}\lambda_{A,F}^{{\rm exp},j}\bigr{)}_{\rm Triv}\Bigr{]}+\Bigl{[}\bigl{[}C^{f,\bullet}_{A,F}\bigr{]}_{\mathbb{Z}_{p}[G]},\,\bigl{(}\lambda^{{\rm NT},j}_{A,F}\bigr{)}_{\rm Triv}\Bigr{]}
+[𝟏p[G],vSSpLv(A,F/k)].\displaystyle\phantom{=-aa}+\Bigl{[}\boldsymbol{1}_{\mathbb{Z}_{p}[G]},\,\prod_{v\in S\cup S_{p}}L_{v}(A,F/k)\Bigr{]}.

This implies the claimed result since (9) implies that

[[CA,Floc,]p[G],(λA,Fexp,j)Triv]=χjloc(A,F/k)\Bigl{[}\bigl{[}C^{{\rm loc},\bullet}_{A,F}\bigr{]}_{\mathbb{Z}_{p}[G]},\,\bigl{(}\lambda_{A,F}^{{\rm exp},j}\bigr{)}_{\rm Triv}\Bigr{]}=\chi_{j}^{\rm loc}(A,F/k)

(as δG,p(Nrdp[G](id|ppH1(CA,Floc,)))=0\delta_{G,p}\bigr{(}\operatorname{Nrd}_{\mathbb{Q}_{p}[G]}\bigl{(}-{\mathrm{id}}\bigm{|}\mathbb{Q}_{p}\otimes_{\mathbb{Z}_{p}}H^{1}(C^{{\rm loc},\bullet}_{A,F})\bigr{)}\bigr{)}=0 because ppH1(CA,Floc,)\mathbb{Q}_{p}\otimes_{\mathbb{Z}_{p}}H^{1}\bigl{(}C^{{\rm loc},\bullet}_{A,F}\bigr{)} is a free p[G]\mathbb{Q}_{p}[G]-module) and

[[CA,Ff,]p[G],(λA,FNT,j)Triv]=χj(A,F/k)\Bigl{[}\bigl{[}C^{f,\bullet}_{A,F}\bigr{]}_{\mathbb{Z}_{p}[G]},\,\bigl{(}\lambda^{{\rm NT},j}_{A,F}\bigr{)}_{\rm Triv}\Bigr{]}=-\chi_{j}(A,F/k)

(because i=1i=2Nrdp[G](id|ppHi(CA,Ff,))(1)i=1\prod_{i=1}^{i=2}\operatorname{Nrd}_{\mathbb{Q}_{p}[G]}\bigl{(}-{\mathrm{id}}\bigm{|}\mathbb{Q}_{p}\otimes_{\mathbb{Z}_{p}}H^{i}(C^{f,\bullet}_{A,F})\bigr{)}^{(-1)^{i}}=1) whilst the explicit description of the isomorphism (8) implies that

[𝟏p[G],vSSpLv(A,F/k)]=\displaystyle\Bigl{[}\boldsymbol{1}_{\mathbb{Z}_{p}[G]},\,\prod_{v\in S\cup S_{p}}L_{v}(A,F/k)\Bigr{]}= δG,p(vSSpLv(A,F/k))\displaystyle\delta_{G,p}\Bigl{(}\prod_{v\in S\cup S_{p}}L_{v}(A,F/k)\Bigr{)}
=\displaystyle= vSSpδG,p(Lv(A,F/k)).\displaystyle\sum_{v\in S\cup S_{p}}\delta_{G,p}(L_{v}(A,F/k)).

4.4. The local term

In this section we explicitly compute the Euler characteristic χjloc(A,F/k)\chi^{\rm loc}_{j}(A,F/k) that occurs in Proposition 4.2 in terms of both archimedean periods and global Galois-Gauss sums.

In the sequel we sometimes suppress explicit reference to the fixed identification of fields j:pj:\mathbb{C}\cong\mathbb{C}_{p}. In addition, for each natural number mm we will write [m][m] for the set of integers ii which satisfy 1im1\leq i\leq m.

4.4.1. Periods and Galois Gauss sums.

We fix Néron models 𝒜t\mathcal{A}^{t} for AtA^{t} over 𝒪k\mathcal{O}_{k} and 𝒜vt\mathcal{A}^{t}_{v} for A/kvtA^{t}_{/k_{v}} over 𝒪kv\mathcal{O}_{k_{v}} for each vv in SpkS_{p}^{k}, and then fix a kk-basis {ωb}b[d]\{\omega_{b}\}_{b\in[d]} of the space of invariant differentials H0(At,ΩAt1)H^{0}(A^{t},\Omega^{1}_{A^{t}}) which gives 𝒪kv\mathcal{O}_{k_{v}}-bases of H0(𝒜vt,Ω𝒜vt1)H^{0}\bigl{(}\mathcal{A}^{t}_{v},\Omega_{\mathcal{A}^{t}_{v}}^{1}\bigr{)} for each vv in SpkS_{p}^{k} and is also such that each ωb\omega_{b} extends to an element of H0(𝒜t,Ω𝒜t1)H^{0}\bigl{(}\mathcal{A}^{t},\Omega^{1}_{\mathcal{A}^{t}}\bigr{)}.

For each place vv in SS_{\mathbb{R}} we fix \mathbb{Z}-bases {γv,a+}a[d]\{\gamma_{v,a}^{+}\}_{a\in[d]} and {γv,a}a[d]\{\gamma_{v,a}^{-}\}_{a\in[d]} of H1(σv(At)(),)c=1H_{1}\bigl{(}\sigma_{v}(A^{t})(\mathbb{C}),\mathbb{Z}\bigr{)}^{c=1} and H1(σv(At)(),)c=1H_{1}\bigl{(}\sigma_{v}(A^{t})(\mathbb{C}),\mathbb{Z}\bigr{)}^{c=-1}, where cc denotes complex conjugation, and then set

Ωv+(A):=|det(γv,a+ωb)a,b|,Ωv(A):=|det(γv,aωb)a,b|,\Omega_{v}^{+}(A):=\Biggl{|}\det\biggl{(}\int_{\gamma_{v,a}^{+}}\omega_{b}\biggr{)}_{\!a,b}\Biggr{|},\qquad\Omega_{v}^{-}(A):=\Biggl{|}\det\biggl{(}\int_{\gamma_{v,a}^{-}}\omega_{b}\biggr{)}_{\!a,b}\Biggr{|},

where in both matrices (a,b)(a,b) runs over [d]×[d][d]\times[d]. For each vv in SS_{\mathbb{C}} we fix a \mathbb{Z}-basis {γv,a}a[2d]\{\gamma_{v,a}\}_{a\in[2d]} of H1(σv(At)(),)H_{1}\bigl{(}\sigma_{v}(A^{t})(\mathbb{C}),\mathbb{Z}\bigr{)} and set

Ωv(A):=|det(γv,aωb,c(γv,aωb))a,b|\Omega_{v}(A):=\Biggl{|}\det\biggl{(}\int_{\gamma_{v,a}}\omega_{b},c\Bigl{(}\int_{\gamma_{v,a}}\omega_{b}\Bigr{)}\biggr{)}_{\!a,b}\Biggr{|}

where (a,b)(a,b) runs over [2d]×[d][2d]\times[d]. (By explicitly computing integrals these periods can be related to those obtained by integrating measures as occurring in the classical formulation of the Birch and Swinnerton-Dyer conjecture – see, for example, Gross [30, p. 224]).

For any ψIr(G)\psi\in\operatorname{Ir}(G) and vSv\in S_{\mathbb{R}}, we set ψv+(1):=dimVψGkv\psi_{v}^{+}(1):=\dim_{\mathbb{C}}V_{\psi}^{G_{k_{v}}} and ψv(1):=ψ(1)ψv+(1)\psi_{v}^{-}(1):=\psi(1)-\psi_{v}^{+}(1). Then we define the periods

Ωv(A,ψ):={Ωv+(A)ψv+(1)Ωv(A)ψv(1) if vS andΩv(A)ψ(1) if vS;\Omega_{v}(A,\psi):=\begin{cases}\Omega^{+}_{v}(A)^{\psi_{v}^{+}(1)}\cdot\Omega^{-}_{v}(A)^{\psi_{v}^{-}(1)}&\text{ if $v\in S_{\mathbb{R}}$ and}\\ \Omega_{v}(A)^{\psi(1)}&\text{ if $v\in S_{\mathbb{C}}$;}\end{cases}

and Ω(A,ψ):=vSΩv(A,ψ)\Omega(A,\psi):=\prod_{v\in S_{\infty}}\Omega_{v}(A,\psi) and finally set

Ω(A,F/k):=ψIr(G)Ω(A,ψ)eψζ(p[G])×.\Omega(A,F/k):=\sum_{\psi\in\operatorname{Ir}(G)}\Omega(A,\psi)\,e_{\psi}\in\zeta\bigl{(}\mathbb{C}_{p}[G]\bigr{)}^{\times}.

For each vv in SS_{\mathbb{R}}, respectively. in SS_{\mathbb{C}}, we also define wv(ψ)w_{v}(\psi) to be equal to iψv(1)i^{\psi^{-}_{v}(1)}, respectively. iψ(1)i^{\psi(1)}, and then set w(ψ):=vSwv(ψ)w_{\infty}(\psi):=\prod_{v\in S_{\infty}}w_{v}(\psi) and

w(F/k):=ψIr(G)w(ψ)eψζ(p[G])×.w_{\infty}(F/k):=\sum_{\psi\in\operatorname{Ir}(G)}w_{\infty}(\psi)\,e_{\psi}\in\zeta\bigl{(}\mathbb{C}_{p}[G]\bigr{)}^{\times}.

To describe the relevant Galois Gauss sums we first define for each non-archimedean place vv of kk the ‘non-ramified characteristic’ uvu_{v} to be the image under the natural induction map ζ(p[G¯v])×ζ(p[G])×\zeta\bigl{(}\mathbb{C}_{p}[\,\overline{G}_{v}\,]\bigr{)}^{\times}\to\zeta\bigl{(}\mathbb{C}_{p}[G]\bigr{)}^{\times} of the element (1eI¯v)+(Frw1)eI¯v(1-e_{\overline{I}_{v}})+(-\operatorname{Fr}_{w}^{-1})e_{\overline{I}_{v}} of ζ([G¯v])×\zeta(\mathbb{Q}[\overline{G}_{v}])^{\times}.

For each character ψ\psi in Ir(G)\operatorname{Ir}(G) we then define a modified equivariant global Galois-Gauss sum by setting

τ(F/k):=(vSrkuv)ψIr(G)eψτ(,Indkψ)ζ(c[G])×,\tau^{*}(F/k):=\biggl{(}\prod_{v\in S_{\rm r}^{k}}u_{v}\biggr{)}\sum_{\psi\in\operatorname{Ir}(G)}e_{\psi}\,\tau\bigl{(}\mathbb{Q},\operatorname{Ind}_{k}^{\mathbb{Q}}\psi\bigr{)}\in\zeta\bigl{(}\mathbb{Q}^{c}[G]\bigr{)}^{\times},

where the individual Galois Gauss sums τ(,)\tau(\mathbb{Q},-) are as defined by Martinet in [37].

4.4.2. Computation of the local Euler characteristic.

The explicit computation of the Euler characteristic χjloc(A,F/k)\chi_{j}^{\rm loc}(A,F/k) is made considerably more difficult by the presence of pp-adic places which ramify in any of the relevant field extensions. For simplicity in the sequel, and to focus attention on the key ideas in the present article, we therefore impose the following additional hypothesis

  • (h)

    pp is unramified in F/F/\mathbb{Q}.

A full treatment of the terms χjloc(A,F/k)\chi_{j}^{\rm loc}(A,F/k) in the general (ramified) case will then be given in a future article.

In the following result we use the elements Lv(A,F/k)L_{v}(A,F/k) of ζ(p[G])×\zeta(\mathbb{Q}_{p}[G])^{\times} that are defined in Proposition 4.2. We will also write jj_{*} for the isomorphism ζ([G])×ζ(p[G])×\zeta(\mathbb{C}[G])^{\times}\cong\zeta(\mathbb{C}_{p}[G])^{\times} that is induced by jj. We finally recall that dd denotes the dimension of AA.

Theorem 4.3.

If AA and FF satisfy the hypotheses (a)(h), then one has

χjloc(A,F/k)=δG,p(j(w(F/k)dΩ(A,F/k)τ(F/k)d)vSbSpLv(A,F/k)).\chi_{j}^{\rm loc}(A,F/k)=\delta_{G,p}\Bigl{(}j_{*}(\frac{w_{\infty}(F/k)^{d}\cdot\Omega(A,F/k)}{\tau^{*}(F/k)^{d}})\prod_{v\in S_{\rm b}\cup S_{p}}L_{v}(A,F/k)\Bigr{)}.
Proof.

We set HA,vF/k:=H0(kv,Yv,FH1(σv(At)(),))H^{F/k}_{A,v}:=H^{0}\bigl{(}k_{v},Y_{v,F}\otimes_{\mathbb{Z}}H_{1}(\sigma_{v}(A^{t})(\mathbb{C}),\mathbb{Z})\bigr{)} for each vv in SS_{\infty} and then also HAF/k:=vSHA,vF/kH^{F/k}_{A}:=\bigoplus_{v\in S_{\infty}}H^{F/k}_{A,v} and C~A,Floc,:=HA,pF/k[0]At(Fp)p[1]\tilde{C}^{{\rm loc},\bullet}_{A,F}:=H^{F/k}_{A,p}[0]\oplus A^{t}(F_{p})^{\wedge}_{p}[-1]. Then, as pp is odd and RΓf(kv,Tp,F(A))R\Gamma_{f}\bigl{(}k_{v},T_{p,F}(A)\bigr{)} is acyclic for all vv in SrS_{\rm r} (by Lemma 4.1(i)), the comparison isomorphisms (5) induce a natural isomorphism

κ:CA,Floc,C~A,Floc,vSbHf1(kv,Tp,F(A))[1]\kappa:C^{{\rm loc},\bullet}_{A,F}\cong\tilde{C}^{{\rm loc},\bullet}_{A,F}\oplus\bigoplus_{v\in S_{\rm b}}H^{1}_{f}\bigl{(}k_{v},T_{p,F}(A)\bigr{)}[-1]

in Dp(p[G])D^{\rm p}\bigl{(}\mathbb{Z}_{p}[G]\bigr{)} and hence an equality in K0(p[G],p[G])K_{0}\bigl{(}\mathbb{Z}_{p}[G],\mathbb{C}_{p}[G]\bigr{)}

(13) χjloc(A,F/k)χG,p(C~A,Floc,,λ~A,Fexp)\displaystyle\chi_{j}^{\rm loc}(A,F/k)-\chi_{G,p}\bigl{(}\tilde{C}^{{\rm loc},\bullet}_{A,F},\tilde{\lambda}^{\rm exp}_{A,F}\bigr{)} =vSbχG,p(Hf1(kv,Tp,F(A))[1], 0)\displaystyle=\sum_{v\in S_{\rm b}}\chi_{G,p}\Bigl{(}H^{1}_{f}\bigl{(}k_{v},T_{p,F}(A)\bigr{)}[-1],\,0\Bigr{)}
=δG,p(vSbLv(A,F/k)).\displaystyle=\delta_{G,p}\Bigl{(}\prod_{v\in S_{\rm b}}L_{v}(A,F/k)\Bigr{)}.

Here we set λ~A,Fexp:=(ppH1(κ))λA,Fexp(ppH0(κ))1\tilde{\lambda}^{\rm exp}_{A,F}:=\bigl{(}\mathbb{Q}_{p}\otimes_{\mathbb{Z}_{p}}H^{1}(\kappa)\bigr{)}\circ\lambda^{\rm exp}_{A,F}\circ\bigl{(}\mathbb{Q}_{p}\otimes_{\mathbb{Z}_{p}}H^{0}(\kappa)\bigr{)}^{-1} and the second equality follows by combining for each vv in SbS_{\rm b} the explicit definition of the term Lv(A,F/k)L_{v}(A,F/k) together with the facts that there is an exact sequence of p[G]\mathbb{Z}_{p}[G]-modules

0Tp,F(A)Iv1Frv1Tp,F(A)IvHf1(kv,Tp,F(A))0,\lx@xy@svg{\hbox{\raise 2.5pt\hbox{\kern 3.5pt\hbox{\ignorespaces\ignorespaces\ignorespaces\hbox{\vtop{\kern 0.0pt\offinterlineskip\halign{\entry@#!@&&\entry@@#!@\cr&&&&\crcr}}}\ignorespaces{\hbox{\kern-3.5pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 1.0pt\raise-2.5pt\hbox{$\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces{\hbox{\kern 42.5694pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}{\hbox{\kern 42.5694pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 1.0pt\raise-2.5pt\hbox{$\textstyle{T_{p,F}(A)^{I_{v}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces\ignorespaces\ignorespaces{\hbox{\kern 96.22249pt\raise 6.62001pt\hbox{{}\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\hbox{\hbox{\kern 0.0pt\raise-2.62001pt\hbox{$\scriptstyle{1-\operatorname{Fr}_{v}^{-1}}$}}}\kern 3.0pt}}}}}}\ignorespaces{\hbox{\kern 128.0622pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}{\hbox{\kern 128.0622pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 1.0pt\raise-2.5pt\hbox{$\textstyle{T_{p,F}(A)^{I_{v}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces{\hbox{\kern 213.555pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}{\hbox{\kern 213.555pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 1.0pt\raise-2.5pt\hbox{$\textstyle{H^{1}_{f}\bigl{(}k_{v},T_{p,F}(A)\bigr{)}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces{\hbox{\kern 327.63538pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}{\hbox{\kern 327.63538pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 1.0pt\raise-2.5pt\hbox{$\textstyle{0}$}}}}}}}\ignorespaces}}}}\ignorespaces,

and that Tp,F(A)IvT_{p,F}(A)^{I_{v}} is a projective p[G]\mathbb{Z}_{p}[G]-module (as a consequence of hypothesis (e)).

We next note that λ~A,Fexp:=μ2μ1\tilde{\lambda}^{\rm exp}_{A,F}:=\mu_{2}\circ\mu_{1}, where μ1\mu_{1} is the second isomorphism that occurs in the definition of λA,Fexp,j\lambda_{A,F}^{{\rm exp},j} and μ2=vSpμ2v\mu_{2}=\bigoplus_{v\in S_{p}}\mu_{2}^{v} where each μ2v\mu_{2}^{v} is induced by the classical exponential map (7).

For each vv in SpS_{p} we set 𝒪Fv:=𝒪F𝒪k𝒪kv\mathcal{O}_{F_{v}}:=\mathcal{O}_{F}\otimes_{\mathcal{O}_{k}}\mathcal{O}_{k_{v}}. Then hypothesis (h) implies that each such place vv is unramified in Fw/kvF_{w}/k_{v} and hence, by Noether’s Theorem, that the p[G]\mathbb{Z}_{p}[G]-module 𝒪Fvwv𝒪Fw\mathcal{O}_{F_{v}}\cong\prod_{w^{\prime}\mid v}\mathcal{O}_{F_{w^{\prime}}} is free. In particular, the p[G]\mathbb{Z}_{p}[G]-module

JA,Fp:=vSp𝒪Fv𝒪kvHom𝒪kv(H0(𝒜vt,Ω𝒜vt1),𝒪kv),J_{A,F_{p}}:=\bigoplus_{v\in S_{p}}\mathcal{O}_{F_{v}}\otimes_{\mathcal{O}_{k_{v}}}\operatorname{Hom}_{\mathcal{O}_{k_{v}}}\Bigl{(}H^{0}\bigl{(}\mathcal{A}^{t}_{v},\Omega^{1}_{\mathcal{A}^{t}_{v}}\bigr{)},\mathcal{O}_{k_{v}}\Bigr{)},

is free and so in K0(p[G],p[G])K_{0}\bigl{(}\mathbb{Z}_{p}[G],\mathbb{C}_{p}[G]\bigr{)} one has an equality

χG,p(C~A,Floc,,λ~A,Fexp)=χG,p(HA,pF/k[0]JA,Fp[1],μ1)+χG,p(JA,Fp[0]At(Fp)p[1],μ2).\chi_{G,p}\bigl{(}\tilde{C}^{{\rm loc},\bullet}_{A,F},\tilde{\lambda}^{\rm exp}_{A,F}\bigr{)}\\ =\chi_{G,p}\bigl{(}H^{F/k}_{A,p}[0]\oplus J_{A,F_{p}}[-1],\mu_{1}\bigr{)}+\chi_{G,p}\bigl{(}J_{A,F_{p}}[0]\oplus A^{t}(F_{p})^{\wedge}_{p}[-1],\mu_{2}\bigr{)}.

Combining this equality with (13) and the explicit computations in Lemmas 4.4 and 4.5 below one finds that χjloc(A,F/k)\chi_{j}^{\rm loc}(A,F/k) is equal to

δG,p(j(w(F/k)dΩ(A,F/k))vSbSpLv(A,F/k))d(𝒪F,p,πF,YF,p).\delta_{G,p}\Bigl{(}j_{*}(w_{\infty}(F/k)^{d}\Omega(A,F/k))\prod_{v\in S_{\rm b}\cup S_{p}}L_{v}(A,F/k)\Bigr{)}-d\cdot(\mathcal{O}_{F,p},\pi_{F},Y_{F,p}).

Here we have set YF:=Σ(F)Y_{F}:=\prod_{\Sigma(F)}\mathbb{Z}, endowed with its natural action of GG, and written πF:ppFpppYF,p\pi_{F}:\mathbb{C}_{p}\otimes_{\mathbb{Q}_{p}}F_{p}\cong\mathbb{C}_{p}\otimes_{\mathbb{Z}_{p}}Y_{F,p} for the natural isomorphism. To deduce the claimed result from this expression it suffices to show (𝒪F,p,πF,YF,p)\bigl{(}\mathcal{O}_{F,p},\pi_{F},Y_{F,p}\bigr{)} is equal to δG,p(j(τ(F/k)))\delta_{G,p}\bigl{(}j_{*}(\tau^{*}(F/k))\bigr{)} and, since no pp-adic place of kk ramifies in F/kF/k, this follows directly from equation (34), Proposition 7.1 and Corollary 7.6 in [6]. ∎

Lemma 4.4.

If at each place vv in SpS_{p} the extension Fw/pF_{w}/\mathbb{Q}_{p} is unramified and AA has good reduction, then one has

χG,p(JA,Fp[0]At(Fp)p[1],μ2)=δG,p(vSpLv(A,F/k)).\chi_{G,p}\Bigl{(}J_{A,F_{p}}[0]\oplus A^{t}(F_{p})^{\wedge}_{p}[-1],\mu_{2}\Bigr{)}=\delta_{G,p}\Bigl{(}\prod_{v\in S_{p}}L_{v}(A,F/k)\Bigr{)}.
Proof.

We fix vv in SpS_{p} and write 𝒜v,Ft\mathcal{A}^{t}_{v,F} for the Néron model 𝒜vt×𝒪kv𝒪Fv\mathcal{A}^{t}_{v}\times_{\mathcal{O}_{k_{v}}}\mathcal{O}_{F_{v}} of AtA^{t} over 𝒪Fv:=wv𝒪Fw\mathcal{O}_{F_{v}}:=\prod_{w^{\prime}\mid v}\mathcal{O}_{F_{w}^{\prime}}. Then Hom𝒪Fv(H0(𝒜v,Ft,Ω𝒜v,Ft1),𝒪Fv)\operatorname{Hom}_{\mathcal{O}_{F_{v}}}\bigl{(}H^{0}(\mathcal{A}^{t}_{v,F},\Omega^{1}_{\mathcal{A}^{t}_{v,F}}\bigr{)},\mathcal{O}_{F_{v}}) is naturally isomorphic to the free p[G]\mathbb{Z}_{p}[G]-module JA,Fv:=𝒪Fv𝒪kvHom𝒪kv(H0(𝒜vt,Ω𝒜vt1),𝒪kv)J_{A,F_{v}}:=\mathcal{O}_{F_{v}}\otimes_{\mathcal{O}_{k_{v}}}\operatorname{Hom}_{\mathcal{O}_{k_{v}}}\bigl{(}H^{0}(\mathcal{A}^{t}_{v},\Omega^{1}_{\mathcal{A}^{t}_{v}}),\mathcal{O}_{k_{v}}\bigr{)}.

In addition, the stated hypotheses on vv imply that there exists a full free p[G]\mathbb{Z}_{p}[G]-submodule

𝒟v:=Dcr,v(Tp,F(A))𝒪Fv𝒪kvDcr,v(Tp(A))\mathcal{D}_{v}:=D_{{\rm cr},v}\bigl{(}T_{p,F}(A)\bigr{)}\cong\mathcal{O}_{F_{v}}\otimes_{\mathcal{O}_{k_{v}}}D_{{\rm cr},v}\bigl{(}T_{p}(A)\bigr{)}

of Dcr,v(Vp,F(A))D_{{\rm cr},v}(V_{p,F}(A)), where Dcr,vD_{{\rm cr},v} is the quasi-inverse to the functor of Fontaine and Lafaille used by Niziol in [42], and the theory of Fontaine and Messing [28] implies that the canonical comparison isomorphism

HomFv(H0(AFvt,ΩAFvt1),Fv)DdR,v(Vp,F(A))/F0\operatorname{Hom}_{F_{v}}\bigl{(}H^{0}(A^{t}_{F_{v}},\Omega^{1}_{A^{t}_{F_{v}}}),F_{v}\bigr{)}\cong D_{{\rm dR},v}\bigl{(}V_{p,F}(A)\bigr{)}/F^{0}

maps JA,FvJ_{A,F_{v}} to 𝒟v/F0𝒟v\mathcal{D}_{v}/F^{0}\mathcal{D}_{v} (see §5 in [8]).

In this case it is also shown in [8] that there is a natural short exact sequence of perfect complexes of p[G]\mathbb{Z}_{p}[G]-modules (with vertical differentials)

0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝒟v/F0𝒟v\textstyle{\mathcal{D}_{v}/F^{0}\mathcal{D}_{v}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}(0,id)\scriptstyle{(0,\mathrm{id})}𝒟v𝒟v/F0𝒟v\textstyle{\mathcal{D}_{v}\oplus\mathcal{D}_{v}/F^{0}\mathcal{D}_{v}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}(id,0)\scriptstyle{(\mathrm{id},0)}𝒟v\textstyle{\mathcal{D}_{v}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0\textstyle{0}0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝒟v\textstyle{\mathcal{D}_{v}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}id\scriptstyle{\mathrm{id}}(1φv,π)\scriptstyle{(1-\varphi_{v},\pi)}𝒟v\textstyle{\mathcal{D}_{v}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}1φv\scriptstyle{1-\varphi_{v}}0.\textstyle{0.}

Here the term 𝒟v/F0𝒟v\mathcal{D}_{v}/F^{0}\mathcal{D}_{v} in the first complex occurs in degree 11 and π\pi is the tautological projection. Further, the exponential map of Bloch and Kato maps the cohomology in degree 11 of the second complex bijectively to Hf1(kv,Tp,F(A))=At(Fv)pH^{1}_{f}(k_{v},T_{p,F}(A))=A^{t}(F_{v})^{\wedge}_{p} and the differential 1φv1-\varphi_{v} of the third complex is injective. For a proof of all these claims see Lemma 4.5 (together with Example 3.11) in loc. cit.. The long exact sequence of cohomology of the above exact sequence thus gives rise to a short exact sequence of p[G]\mathbb{Z}_{p}[G]-modules 0JA,FvAt(Fv)pcok(1φv)00\to J_{A,F_{v}}\to A^{t}(F_{v})^{\wedge}_{p}\to\operatorname{cok}(1-\varphi_{v})\to 0, in which (following [8, Example 3.11]) the second arrow is equal to μ2v\mu_{2}^{v}. This sequence then implies that the term χG,p(JA,Fv[0]At(Fv)p[1],μ2v)\chi_{G,p}\bigl{(}J_{A,F_{v}}[0]\oplus A^{t}(F_{v})^{\wedge}_{p}[-1],\mu^{v}_{2}\bigr{)} is equal to

χG,p(cok(1φv)[1],0)=G,p((1φvDcr,v(Vp,F(A))))=δG,p(Lv(A,F/k))\chi_{G,p}\bigl{(}\operatorname{cok}(1-\varphi_{v})[-1],0\bigr{)}=\partial_{G,p}\bigl{(}(1-\varphi_{v}\mid D_{{\rm cr},v}(V_{p,F}(A)))\bigr{)}=\delta_{G,p}\bigl{(}L_{v}(A,F/k)\bigr{)}

and by summing over all places vv in SpS_{p} this implies the claimed equality. ∎

For each σΣ(k)\sigma\in\Sigma(k) we fix an embedding σ^Σσ(F)\hat{\sigma}\in\Sigma_{\sigma}(F).

Lemma 4.5.

In K0(p[G],p[G])K_{0}\bigl{(}\mathbb{Z}_{p}[G],\mathbb{C}_{p}[G]\bigr{)} one has

χG,p(HA,pF/k[0]JA,Fp[1],μ1)=d(𝒪F,p,πF,YF,p)+δG,p(j(w(F/k)dΩ(A,F/k))).\chi_{G,p}\Bigl{(}H^{F/k}_{A,p}[0]\oplus J_{A,F_{p}}[-1],\,\mu_{1}\Bigr{)}\\ =-d\cdot(\mathcal{O}_{F,p},\pi_{F},Y_{F,p})+\delta_{G,p}\bigl{(}j_{*}(w_{\infty}(F/k)^{d}\,\Omega(A,F/k))\bigr{)}.
Proof.

For each place vv in SS_{\mathbb{R}} we define τv\tau_{v} in GG via the equality τv(σ^v)=cσ^v\tau_{v}(\hat{\sigma}_{v})=c\circ\hat{\sigma}_{v}. For each integer aa in [d][d] we then define an element

xv,a:=12(1+τv)σ^vγv,a++12(1τv)σ^vγv,ax_{v,a}:=\tfrac{1}{2}(1+\tau_{v})\hat{\sigma}_{v}\otimes_{\mathbb{Z}}\gamma_{v,a}^{+}+\tfrac{1}{2}(1-\tau_{v})\hat{\sigma}_{v}\otimes_{\mathbb{Z}}\gamma_{v,a}^{-}

and we note that, since pp is odd, the set 𝒳v:={xv,a}a[d]\mathcal{X}_{v}:=\{x_{v,a}\}_{a\in[d]} is a p[G]\mathbb{Z}_{p}[G]-basis of HA,v,pF/kH^{F/k}_{A,v,p}.

For vv in SS_{\mathbb{C}} and aa in [2d][2d] we set xv,a:=σ^vγv,ax_{v,a}:=\hat{\sigma}_{v}\otimes_{\mathbb{Z}}\gamma_{v,a} and note that 𝒳v:={xv,a}a[2d]\mathcal{X}_{v}:=\{x_{v,a}\}_{a\in[2d]} is a p[G]\mathbb{Z}_{p}[G]-basis of HA,v,pF/kH^{F/k}_{A,v,p}.

We next set n:=[k:]=|S|+2|S|n:=[k:\mathbb{Q}]=|S_{\mathbb{R}}|+2|S_{\mathbb{C}}| and fix an 𝒪k,p[G]\mathcal{O}_{k,p}[G]-generator ZZ of 𝒪Fp\mathcal{O}_{F_{p}} and a p\mathbb{Z}_{p}-basis {zm}m[n]\{z_{m}\}_{m\in[n]} of 𝒪k,p\mathcal{O}_{k,p} and set tm,b:=Zzmkωbt_{m,b}:=Zz_{m}\otimes_{k}\omega_{b}^{*} with ωb\omega_{b}^{*} the element of Homk(H0(At,ΩAt1),k)\operatorname{Hom}_{k}\bigl{(}H^{0}(A^{t},\Omega^{1}_{A^{t}}),k\bigr{)} that is dual to ωb\omega_{b}. Then the set 𝒯:={tm,b}(m,b)[n]×[d]\mathcal{T}:=\{t_{m,b}\}_{(m,b)\in[n]\times[d]} is a p[G]\mathbb{Z}_{p}[G]-basis of JA,FpJ_{A,F_{p}}.

The key to our argument is to compute the matrix of μ1\mu_{1} with respect to the bases 𝒯\mathcal{T} and 𝒳:=vS𝒳v\mathcal{X}:=\bigcup_{v\in S_{\infty}}\mathcal{X}_{v} of (FkHomk(H0(At,ΩAt),k))\mathbb{R}\otimes_{\mathbb{Q}}\bigl{(}F\otimes_{k}\operatorname{Hom}_{k}(H^{0}(A^{t},\Omega_{A^{t}}),k)\bigr{)} and HAF/k\mathbb{R}\otimes_{\mathbb{Z}}H^{F/k}_{A}. To do this we find it convenient to introduce an auxiliary basis. Thus, for vv in SS_{\mathbb{R}} we set 𝒴v:={yv,b}b[d]\mathcal{Y}_{v}:=\{y_{v,b}\}_{b\in[d]} with yv,b:=σ^vσv,(ωb)y_{v,b}:=\hat{\sigma}_{v}\otimes_{\mathbb{Z}}\sigma_{v,*}(\omega^{*}_{b}), and for vv in SS_{\mathbb{C}} we set 𝒴v:={yv,b}b[2d]\mathcal{Y}_{v}:=\{y_{v,b}\}_{b\in[2d]} with yv,b:=σ^vσv,(ωb)y_{v,b}:=\hat{\sigma}_{v}\otimes_{\mathbb{Z}}\sigma_{v,*}(\omega^{*}_{b}) if b[d]b\in[d] and yv,b:=σ^vc(σv,(ωb))y_{v,b}:=\hat{\sigma}_{v}\otimes_{\mathbb{Z}}c\bigl{(}\sigma_{v,*}(\omega^{*}_{b})\bigr{)} if b[2d][d]b\in[2d]\setminus[d]. Then 𝒴:=vS𝒴v\mathcal{Y}:=\cup_{v\in S_{\infty}}\mathcal{Y}_{v} is an [G]\mathbb{R}[G]-basis of vSkvYv,Fkv(kvk,σvHomk(H0(At,ΩAt1),k))\bigoplus_{v\in S_{\infty}}k_{v}Y_{v,F}\otimes_{k_{v}}\bigl{(}k_{v}\otimes_{k,\sigma_{v}}\operatorname{Hom}_{k}(H^{0}(A^{t},\Omega^{1}_{A^{t}}),k)\bigr{)} and for each index aa one has

πv(xv,a)={j=1d12(Ωv,a,j+(1+τv)+Ωv,a,j(1τv))yv,j,if vS,j=12dΩv,a,jyv,j,if vS\pi_{v}(x_{v,a})=\begin{cases}\sum_{j=1}^{d}\tfrac{1}{2}(\Omega_{v,a,j}^{+}(1+\tau_{v})+\Omega_{v,a,j}^{-}(1-\tau_{v}))y_{v,j},&\text{if $v\in S_{\mathbb{R}}$},\\ \sum_{j=1}^{2d}\Omega_{v,a,j}\,y_{v,j},&\text{if $v\in S_{\mathbb{C}}$}\end{cases}

with Ωv,a,j±:=γv,a±ωj\Omega_{v,a,j}^{\pm}:=\int_{\gamma_{v,a}^{\pm}}\omega_{j} and Ωv,a,j:=γv,aωj\Omega_{v,a,j}:=\int_{\gamma_{v,a}}\omega_{j} if a[d]a\in[d] and Ωv,a,j:=c(γv,aωj)\Omega_{v,a,j}:=c\bigl{(}\int_{\gamma_{v,a}}\omega_{j}\bigr{)} if a[2d][d]a\in[2d]\setminus[d]. This formula implies that the matrix of μ1\mu_{1} with respect to the bases 𝒳\mathcal{X} and 𝒯\mathcal{T} is M1ΦM^{-1}\Phi where Φ=diag((Φv)vS)\Phi=\operatorname{diag}((\Phi_{v})_{v\in S_{\infty}}) is a diagonal block matrix with blocks

Φv:={(Ωv,a,j+(1+τv)/2+Ωv,a,j(1τv)/2)(a,j)[d]×[d]if vS,(Ωv,a,j)(a,j)[2d]×[2d]if vS\Phi_{v}:=\begin{cases}\bigl{(}\Omega_{v,a,j}^{+}(1+\tau_{v})/2+\Omega_{v,a,j}^{-}(1-\tau_{v})/2\bigr{)}_{(a,j)\in[d]\times[d]}&\text{if $v\in S_{\mathbb{R}}$},\\ \bigl{(}\Omega_{v,a,j}\bigr{)}_{(a,j)\in[2d]\times[2d]}&\text{if $v\in S_{\mathbb{C}}$}\end{cases}

and MM is the matrix in GLnd([G])\operatorname{GL}_{nd}\bigl{(}\mathbb{C}[G]\bigr{)} that represents the isomorphism (6) with respect to the bases 𝒯\mathcal{T} and 𝒴\mathcal{Y} so that

M(a,j),(v,α)={σv(za)gGσ^v(g(Z))g1if vS and α=j,σv(za)gGσ^v(g(Z))g1if vS and α=j,cσv(za)gGcσ^v(g(Z))g1if vS and α=j+d,0otherwise.M_{(a,j),(v,\alpha)}=\begin{cases}\sigma_{v}(z_{a})\,\sum_{g\in G}\hat{\sigma}_{v}\bigl{(}g(Z)\bigr{)}g^{-1}&\text{if $v\in S_{\mathbb{R}}$ and $\alpha=j$},\\ \sigma_{v}(z_{a})\,\sum_{g\in G}\hat{\sigma}_{v}\bigl{(}g(Z)\bigr{)}g^{-1}&\text{if $v\in S_{\mathbb{C}}$ and $\alpha=j$},\\ c\circ\sigma_{v}(z_{a})\,\sum_{g\in G}c\circ\hat{\sigma}_{v}\bigl{(}g(Z)\bigr{)}g^{-1}&\text{if $v\in S_{\mathbb{C}}$ and $\alpha=j+d$,}\\ 0&\text{otherwise.}\end{cases}

Writing [M1Φ]\bigl{[}M^{-1}\Phi\bigr{]} for the class of M1ΦM^{-1}\Phi in K1(p[G])K_{1}\bigl{(}\mathbb{C}_{p}[G]\bigr{)} one therefore has

(14) χG,p(HA,pF/k[0]JA,Fp[1],μ1)=G,p([M1Φ])=δG,p(Nrdp[G](M1Φ))=δG,p(Nrdp[G](M))+δG,p(Nrdp[G](Φ)).\chi_{G,p}\Bigl{(}H^{F/k}_{A,p}[0]\oplus J_{A,F_{p}}[-1],\,\mu_{1}\Bigr{)}=\partial_{G,p}\bigl{(}[M^{-1}\Phi]\bigr{)}\\ =\delta_{G,p}\bigl{(}\operatorname{Nrd}_{\mathbb{C}_{p}[G]}(M^{-1}\Phi)\bigr{)}=-\delta_{G,p}\bigl{(}\operatorname{Nrd}_{\mathbb{C}_{p}[G]}(M)\bigr{)}+\delta_{G,p}\bigl{(}\operatorname{Nrd}_{\mathbb{C}_{p}[G]}(\Phi)\bigr{)}.

Now MM is equal to the product M1M2M_{1}M_{2} for the block matrices M1=(σ(za)Id)σΣ(k),a[n]M_{1}=\bigl{(}\sigma(z_{a})I_{d}\bigr{)}_{\sigma\in\Sigma(k),a\in[n]} and M2=diag((gGσ^(g(Z))g1)Id)σΣ(k)M_{2}=\operatorname{diag}\bigl{(}(\sum_{g\in G}\hat{\sigma}(g(Z))g^{-1})I_{d}\bigr{)}_{\sigma\in\Sigma(k)} and so Nrd[G](M)\operatorname{Nrd}_{\mathbb{C}[G]}(M) is equal to the product

Nrd[G](M1)Nrd[G](M2)\displaystyle\operatorname{Nrd}_{\mathbb{C}[G]}(M_{1})\operatorname{Nrd}_{\mathbb{C}[G]}(M_{2}) =Nrd[G]((σ(za))σ,a)dNrd[G](diag(gGσ^(g(Z))g1)σ)d\displaystyle=\operatorname{Nrd}_{\mathbb{C}[G]}\bigl{(}(\sigma(z_{a}))_{\sigma,a}\bigr{)}^{d}\operatorname{Nrd}_{\mathbb{C}[G]}\biggl{(}\operatorname{diag}\Bigl{(}\sum_{g\in G}\hat{\sigma}\bigl{(}g(Z)\bigr{)}g^{-1}\Bigr{)}_{\sigma}\biggr{)}^{d}
=Nrd[G]((σ(za)gGσ^(g(Z))g1)a,σ)d.\displaystyle=\operatorname{Nrd}_{\mathbb{C}[G]}\biggl{(}\Bigl{(}\sigma(z_{a})\sum_{g\in G}\hat{\sigma}\bigl{(}g(Z)\bigr{)}g^{-1}\Bigr{)}_{a,\sigma}\biggr{)}^{d}.

In addition, (σ(za)gGσ^(g(Z))g1)a,σ\bigl{(}\sigma(z_{a})\sum_{g\in G}\hat{\sigma}(g(Z))g^{-1}\bigr{)}_{a,\sigma} is the matrix of the natural isomorphism πF:ppFpppYF,p\pi_{F}:\mathbb{C}_{p}\otimes_{\mathbb{Q}_{p}}F_{p}\cong\mathbb{C}_{p}\otimes_{\mathbb{Z}_{p}}Y_{F,p} with respect to the p[G]\mathbb{Z}_{p}[G]-bases {Zza}a[n]\{Zz_{a}\}_{a\in[n]} and {σ^}σΣ(k)\{\hat{\sigma}\}_{\sigma\in\Sigma(k)} of 𝒪F,p\mathcal{O}_{F,p} and YF,pY_{F,p} and so

(15) δG,p(Nrdp[G](M))=dG,p([(σ(za)gGσ^(g(Z))g1)a,σ])=d(𝒪F,p,πF,YF,p).\delta_{G,p}\bigl{(}\operatorname{Nrd}_{\mathbb{C}_{p}[G]}(M)\bigr{)}=d\cdot\partial_{G,p}\biggl{(}\biggl{[}\Bigl{(}\sigma(z_{a})\sum_{g\in G}\hat{\sigma}\bigl{(}g(Z)\bigr{)}g^{-1}\Bigr{)}_{a,\sigma}\biggr{]}\biggr{)}=d\cdot\!\bigl{(}\mathcal{O}_{F,p},\pi_{F},Y_{F,p}\bigr{)}.

Next we note that Nrdp[G](Φ)=vSNrdp[G](Φv)\operatorname{Nrd}_{\mathbb{C}_{p}[G]}(\Phi)=\prod_{v\in S_{\infty}}\operatorname{Nrd}_{\mathbb{C}_{p}[G]}(\Phi_{v}) and that for each vv in SS_{\infty} also Nrdp[G](Φv)=ψIr(G)eψdetp(Φv|Vψd[kv:])\operatorname{Nrd}_{\mathbb{C}_{p}[G]}(\Phi_{v})=\sum_{\psi\in\operatorname{Ir}(G)}e_{\psi}\det_{\mathbb{C}_{p}}\bigl{(}\Phi_{v}\bigm{|}V_{\psi}^{d[k_{v}:\mathbb{R}]}\bigr{)} with

detp(Φv|Vψd[kv:])\displaystyle\det\nolimits_{\mathbb{C}_{p}}\Bigl{(}\Phi_{v}\Bigm{|}V_{\psi}^{d\,[k_{v}:\mathbb{R}]}\Bigr{)} ={det((Ωv,a,j+)a,j)ψv+(1)det((Ωv,a,j)a,j)ψv(1)if vS,det((Ωv,a,j)a,j)ψ(1)if vS.\displaystyle=\begin{cases}\det\bigl{(}(\Omega^{+}_{v,a,j})_{a,j}\bigr{)}^{\psi^{+}_{v}(1)}\cdot\det\bigl{(}(\Omega^{-}_{v,a,j})_{a,j}\bigr{)}^{\psi^{-}_{v}(1)}&\text{if $v\in S_{\mathbb{R}}$},\\ \det\bigl{(}(\Omega_{v,a,j})_{a,j}\bigr{)}^{\psi(1)}&\text{if $v\in S_{\mathbb{C}}$.}\end{cases}

Now if vv is real, then det(Ωv,a,j+)a,j=ϖv+Ωv+(A)\det(\Omega^{+}_{v,a,j})_{a,j}=\varpi_{v}^{+}\cdot\Omega_{v}^{+}(A) and det(Ωv,a,j)a,j=ϖvidΩv(A)\det(\Omega^{-}_{v,a,j})_{a,j}=\varpi_{v}^{-}\cdot i^{d}\cdot\Omega_{v}^{-}(A) with ϖv+\varpi_{v}^{+} and ϖv\varpi_{v}^{-} in {±1}\{\pm 1\}, whilst if vv is complex, then det(Ωv,a,j)a,j=ϖvidΩv(A)\det(\Omega_{v,a,j})_{a,j}=\varpi_{v}\cdot i^{d}\cdot\Omega_{v}(A) with ϖv\varpi_{v} in {±1}\{\pm 1\}. Thus, since each of the terms

ψIrp(G)eψ(ϖv+)ψv+(1)(ϖv)ψv(1)=Nrdp[G](ϖv+(1+τv)/2+ϖv(1τv)/2)\sum_{\psi\in\operatorname{Ir}_{p}(G)}e_{\psi}(\varpi_{v}^{+})^{\psi^{+}_{v}(1)}\,(\varpi^{-}_{v})^{\psi^{-}_{v}(1)}=\operatorname{Nrd}_{\mathbb{Q}_{p}[G]}\bigl{(}\varpi_{v}^{+}(1+\tau_{v})/2+\varpi_{v}^{-}(1-\tau_{v})/2\bigr{)}

if vv is real, and ψIrp(G)eψϖvψ(1)=Nrdp[G](ϖv)\sum_{\psi\in\operatorname{Ir}_{p}(G)}e_{\psi}\,\varpi_{v}^{\psi(1)}=\operatorname{Nrd}_{\mathbb{Q}_{p}[G]}(\varpi_{v}) if vv is complex, belong to the kernel of δG,p\delta_{G,p}, we conclude that δG,p(Nrdp[G](Φ))\delta_{G,p}\bigl{(}\operatorname{Nrd}_{\mathbb{C}_{p}[G]}(\Phi)\bigr{)} is equal to

δG,p(ψIr(G)eψvS(Ωv+(A)ψv+(1)(idΩv(A))ψv(1))vS(idΩv(A))ψ(1))\displaystyle\delta_{G,p}\biggl{(}\sum_{\psi\in\operatorname{Ir}(G)}e_{\psi}\prod_{v\in S_{\mathbb{R}}}\Bigl{(}\Omega^{+}_{v}(A)^{\psi^{+}_{v}(1)}\bigl{(}i^{d}\cdot\Omega^{-}_{v}(A)\bigr{)}^{\psi^{-}_{v}(1)}\Bigr{)}\prod_{v\in S_{\mathbb{C}}}\bigl{(}i^{d}\cdot\Omega_{v}(A)\bigr{)}^{\psi(1)}\biggr{)}
=\displaystyle= δG,p((ψIr(G)w(ψ)deψ)(ψIr(G)Ω(A,ψ)eψ))\displaystyle\delta_{G,p}\biggl{(}\Bigl{(}\sum_{\psi\in\operatorname{Ir}(G)}w_{\infty}(\psi)^{d}\,e_{\psi}\Bigr{)}\Bigl{(}\sum_{\psi\in\operatorname{Ir}(G)}\Omega(A,\psi)e_{\psi}\Bigr{)}\biggr{)}
=\displaystyle= δG,p(w(F/k)dΩ(A,F/k)).\displaystyle\delta_{G,p}\Bigl{(}w_{\infty}(F/k)^{d}\cdot\Omega(A,F/k)\Bigr{)}.

This formula combines with (14) and (15) to give the claimed formula for the term χG,p(HA,pF/k[0]JA,Fp[1],μ1)\chi_{G,p}\bigl{(}H^{F/k}_{A,p}[0]\oplus J_{A,F_{p}}[-1],\,\mu_{1}\bigr{)}. ∎

This completes the proof of Theorem 4.3.

5. Congruences between derivatives

We assume in the sequel that, in addition to the hypotheses (a)(h), the following standard conjecture is also valid.

  • (i)

    For each finite set of places Σ\Sigma of kk and each character ψ\psi in Ir(G)\operatorname{Ir}(G) the Σ\Sigma-truncated ψ\psi-twisted Hasse-Weil LL-function LΣ(A,ψ,s)L_{\Sigma}(A,\psi,s) of AA has an analytic continuation to s=1s=1 where it has a zero of order rψ:=dim(VψA(F))Gr_{\psi}:=\dim_{\mathbb{C}}\bigl{(}V_{\psi}\otimes_{\mathbb{Z}}A(F)\bigr{)}^{G}.

Here GG acts diagonally on the tensor product and so rψr_{\psi} is equal to the multiplicity with which ψ\psi, and hence also ψˇ\check{\psi}, appears in the representation A(F)\mathbb{C}\otimes_{\mathbb{Z}}A(F).

For each ψ\psi in Ir(G)\operatorname{Ir}(G) we then write LΣ(A,ψ,1)L^{\raisebox{2.0pt}{$\scriptstyle\star$}}_{\Sigma}(A,\psi,1) for the the coefficient of (s1)rψ(s-1)^{r_{\psi}} in the Taylor expansion at s=1s=1 of LΣ(A,ψ,s)L_{\Sigma}(A,\psi,s).

In this section we use the computations of §4 to give a reinterpretation, under the hypotheses (a)(i), of the appropriate case of the equivariant Tamagawa number conjecture of [16, Conjecture 4] as a family of congruence relations between the leading terms LSr(A,ψ,1)L^{\raisebox{2.0pt}{$\scriptstyle\star$}}_{S_{\rm r}}(A,\psi,1) as ψ\psi varies over Ir(G)\operatorname{Ir}(G). We next discuss several consequences of this reinterpretation and then, motivated by the results of [18], we specialise to the case that the (pp-completed) Mordell-Weil group A(F)pA(F)_{p} is a projective p[G]\mathbb{Z}_{p}[G]-module. In this case we prove results that will subsequently enable us in §6 to give some important new theoretical and numerical verifications of [16, Conjecture 4].

5.1. The general case

If AA and FF satisfy the hypotheses (g) and (i), then [16, Conjecture 4] for the pair (h1(A/F)(1),[G])\bigl{(}h^{1}(A_{/F})(1),\mathbb{Z}[G]\bigr{)} asserts the validity of an equality in K0([G],[G])K_{0}\bigl{(}\mathbb{Z}[G],\mathbb{R}[G]\bigr{)}. In such a case we shall say that the ‘eTNCp for (h1(A/F)(1),[G])\bigl{(}h^{1}(A_{/F})(1),\mathbb{Z}[G]\bigr{)} is valid’ if for every isomorphism of fields j:pj:\mathbb{C}\cong\mathbb{C}_{p} the predicted equality is valid after projection under the homomorphism jG,j_{G,*} defined in (4).

If pp does not divide |G||G|, then the algebra p[G]\mathbb{Z}_{p}[G] is regular and it is straightforward to use the techniques described in [15, §1.7] to give an explicit interpretation of these projections. If pp divides |G||G|, however, then obtaining an explicit interpretation is in general very difficult (see, for example, the efforts made by Bley in [3, 4]).

The following result is thus of some interest since, as we shall see later, it can be combined with the structure results obtained in [18] to show that under certain natural conditions one can explicitly interpret, and verify, the eTNCp for (h1(A/F)(1),[G])\bigl{(}h^{1}(A_{/F})(1),\mathbb{Z}[G]\bigr{)} even if GG is non-abelian, pp divides |G||G| and the rank of A(F)A(F) is strictly positive.

For each non-archimedean place vv of kk, we decompose the non-ramified characteristic uvu_{v} as ψIr(G)uv,ψeψ\sum_{\psi\in\operatorname{Ir}(G)}u_{v,\psi}\cdot e_{\psi} with each uv,ψu_{v,\psi} in ×\mathbb{C}^{\times}. For each ψ\psi in Ir(G)\operatorname{Ir}(G) we then define a modified global Galois-Gauss sum by setting

(16) τ(,Indkψ):=uψτ(,Indkψ).\tau^{*}\bigl{(}\mathbb{Q},\operatorname{Ind}_{k}^{\mathbb{Q}}\psi\bigr{)}:=u_{\psi}\cdot\tau\bigl{(}\mathbb{Q},\operatorname{Ind}_{k}^{\mathbb{Q}}\psi\bigr{)}.

with uψ:=vSrkuv,ψu_{\psi}:=\prod_{v\in S_{\rm r}^{k}}u_{v,\psi}.

Theorem 5.1.

Assume that AA and FF satisfy the hypotheses (a)(i). Then the element

A,F/k:=ψIr(G)eψLSr(A,ψˇ,1)τ(,Indkψ)dΩ(A,ψ)w(ψ)d\mathcal{L}^{\raisebox{1.0pt}{$\scriptstyle\star$}}_{A,F/k}:=\sum_{\psi\in\operatorname{Ir}(G)}e_{\psi}\,\frac{L^{\raisebox{2.0pt}{$\scriptstyle\star$}}_{S_{\rm r}}(A,\check{\psi},1)\cdot\tau^{*}\bigl{(}\mathbb{Q},\operatorname{Ind}_{k}^{\mathbb{Q}}\psi\bigr{)}^{d}}{\Omega(A,\psi)\cdot w_{\infty}(\psi)^{d}}

belongs to ζ([G])×\zeta\bigl{(}\mathbb{C}[G]\bigr{)}^{\times} and the eTNCp for (h1(A/F)(1),[G])\bigl{(}h^{1}(A_{/F})(1),\mathbb{Z}[G]\bigr{)} is valid if and only if for every isomorphism of fields j:pj:\mathbb{C}\cong\mathbb{C}_{p} one has

(17) δG,p(j(A,F/k))=χj(A,F/k)\delta_{G,p}\bigl{(}j_{*}(\mathcal{L}^{\raisebox{1.0pt}{$\scriptstyle\star$}}_{A,F/k})\bigr{)}=\chi_{j}(A,F/k)
Proof.

Since LSr(A,ψˇ,1)0L^{\raisebox{2.0pt}{$\scriptstyle\star$}}_{S_{\rm r}}(A,\check{\psi},1)\neq 0 for all ψ\psi in Ir(G)\operatorname{Ir}(G) the element A,F/k\mathcal{L}^{\raisebox{1.0pt}{$\scriptstyle\star$}}_{A,F/k} clearly belongs to ζ([G])×\zeta(\mathbb{C}[G])^{\times}.

We next recall that [16, Conjecture 4(iv)] for the pair (h1(A/F)(1),[G])\bigl{(}h^{1}(A_{/F})(1),\mathbb{Z}[G]\bigr{)} asserts

RΩ(h1(A/F)(1),[G])=δG(L([G]h1(A/F)(1),0))R\Omega\bigl{(}h^{1}(A_{/F})(1),\mathbb{Z}[G]\bigr{)}=-\delta_{G}\bigl{(}L^{\raisebox{2.0pt}{$\scriptstyle\star$}}(_{\mathbb{Q}[G]}h^{1}(A_{/F})(1),0)\bigr{)}

where δG\delta_{G} is the ‘extended boundary homomorphism’ ζ([G])×K0([G],[G])\zeta\bigl{(}\mathbb{R}[G]\bigr{)}^{\times}\to K_{0}\bigl{(}\mathbb{Z}[G],\mathbb{R}[G]\bigr{)} defined in [16, Lemma 9] and L([G]h1(A/F)(1),0)L^{\raisebox{2.0pt}{$\scriptstyle\star$}}\bigl{(}_{\mathbb{Q}[G]}h^{1}(A_{/F})(1),0\bigr{)} is the leading term at s=0s=0 of the ζ([G])\zeta\bigl{(}\mathbb{C}[G]\bigr{)}-valued LL-function L([G]h1(A/F)(1),s)L\bigl{(}_{\mathbb{Q}[G]}h^{1}(A_{/F})(1),s\bigr{)} defined in [16, §4.1].

Noting that for each isomorphism j:pj:\mathbb{C}\cong\mathbb{C}_{p} one has jG,δG=δG,pjj_{G,*}\circ\delta_{G}=\delta_{G,p}\circ j_{*}, and recalling the result of Proposition 4.2, it follows that the eTNCp for (h1(A/F)(1),[G])\bigl{(}h^{1}(A_{/F})(1),\mathbb{Z}[G]\bigr{)} is valid if and only if for every isomorphism jj one has

δG,p(j(L([G]h1(A/F)(1),0)))=χj(A,F/k)+χjloc(A,F/k)vSSpδG,p(Lv(A,F/k)).\delta_{G,p}\Bigl{(}j_{*}\bigl{(}L^{\raisebox{2.0pt}{$\scriptstyle\star$}}(_{\mathbb{Q}[G]}h^{1}(A_{/F})(1),0)\bigr{)}\Bigr{)}\\ =\chi_{j}(A,F/k)+\chi_{j}^{\rm loc}(A,F/k)-\sum_{v\in S\cup S_{p}}\delta_{G,p}\bigl{(}L_{v}(A,F/k)\bigr{)}.

Now hypotheses (e) and (h) combine to imply that Sr=(SSp)(SbSp)S_{\rm r}=(S\cup S_{p})\setminus(S_{\rm b}\cup S_{p}) and for each place vv in this set the term Lv(A,F/k)L_{v}(A,F/k) is equal to the value at s=0s=0 of the Euler factor at vv that occurs in the definition of L([G]h1(A/F)(1),s)L\bigl{(}_{\mathbb{Q}[G]}h^{1}(A_{/F})(1),s\bigr{)} (by [16, Remark 7]). In view of the formula for χjloc(A,F/k)\chi_{j}^{\rm loc}(A,F/k) that is given in Theorem 4.3, one therefore finds that the above equality is valid if and only if

δG,p(j(LSr([G]h1(A/F)(1),0)τ(F/k)dΩ(A,F/k)w(F/k)d))=χj(A,F/k).\delta_{G,p}\biggl{(}j_{*}\Bigl{(}\frac{L^{\raisebox{2.0pt}{$\scriptstyle\star$}}_{S_{\rm r}}\bigl{(}_{\mathbb{Q}[G]}h^{1}(A_{/F})(1),0\bigr{)}\cdot\tau^{*}(F/k)^{d}}{\Omega(A,F/k)\cdot w_{\infty}(F/k)^{d}}\Bigr{)}\biggr{)}=\chi_{j}(A,F/k).

It now suffices to show that the quotient on the left hand side of this equality is equal to A,F/k\mathcal{L}^{\raisebox{1.0pt}{$\scriptstyle\star$}}_{A,F/k} and this follows from a straightforward comparison of all of the terms involved and then noting that the definition of the truncated LL-function LSr([G]h1(A/F)(1),s)L_{S_{\rm r}}\bigl{(}_{\mathbb{Q}[G]}h^{1}(A_{/F})(1),s\bigr{)} implies LSr([G]h1(A/F)(1),0)=ψIr(G)eψLSr(A,ψˇ,1)L^{\raisebox{2.0pt}{$\scriptstyle\star$}}_{S_{\rm r}}\bigl{(}_{\mathbb{Q}[G]}h^{1}(A_{/F})(1),0\bigr{)}=\sum_{\psi\in\operatorname{Ir}(G)}e_{\psi}L^{\raisebox{2.0pt}{$\scriptstyle\star$}}_{S_{\rm r}}(A,\check{\psi},1). ∎

If it is valid, then the equality (17) can be combined with the theory of organising matrices developed in [17] to extract from the element A,F/k\mathcal{L}^{\raisebox{1.0pt}{$\scriptstyle\star$}}_{A,F/k} a range of detailed information about the arithmetic of AA over FF. To give an explicit example of such an implication we assume, motivated by the results of Theorem 2.7 and Corollary 2.10 in [18], that there exists a surjective homomorphism of p[G]\mathbb{Z}_{p}[G]-modules of the form

π:Selp(AF)Π=ΠprΠnpr\pi:\operatorname{Sel}_{p}(A_{F})^{\vee}\to\Pi=\Pi^{\rm pr}\oplus\Pi^{\rm npr}

where Π\Pi is a trivial source p[G]\mathbb{Z}_{p}[G]-module, Πpr\Pi^{\rm pr} is a projective p[G]\mathbb{Z}_{p}[G]-module and π\pi induces, upon passage to GG-coinvariants, an isomorphism A(k)pΠGA(k)^{*}_{p}\cong\Pi_{G} (via the relevant canonical short exact sequence of the form (1)). The notion of ‘trivial source p[G]\mathbb{Z}_{p}[G]-module’ that we use here corresponds to the one defined in [18, §2.3.2]. For the purpose of applying the result [17, Corollary 2.13] to prove Proposition 5.2 below we however warn the reader that this terminology differs from the one introduced in §2.2 in loc. cit., where trivial source modules are instead called ‘permutation modules’. In this case we also write Υpr\Upsilon_{\rm pr} for the subset of Ir(G)\operatorname{Ir}(G) comprising characters for which the homomorphism eψ(π)e_{\psi}(\mathbb{C}\otimes_{\mathbb{Z}}\pi) is bijective and then define an idempotent in ζ([G])\zeta\bigl{(}\mathbb{Q}[G]\bigr{)} by setting epr:=ψΥpreψe_{\rm pr}:=\sum_{\psi\in\Upsilon_{\rm pr}}e_{\psi}.

Given a natural number mm and a matrix MM in Mm(p[G])\operatorname{M}_{m}\bigl{(}\mathbb{Z}_{p}[G]\bigr{)} we write MM^{\prime} for the corresponding ‘generalised adjoint matrix’ in Mm(p[G])\operatorname{M}_{m}\bigl{(}\mathbb{Q}_{p}[G]\bigr{)} that is defined by Johnston and Nickel in [34, §3.6]. We then write 𝒜p(G)\mathcal{A}_{p}(G) for the ζ(p[G])\zeta\bigl{(}\mathbb{Z}_{p}[G]\bigr{)}-submodule of ζ(p[G])\zeta\bigl{(}\mathbb{Q}_{p}[G]\bigr{)} given by the set

{xζ(p[G]): if m>0 and MMm(p[G]) then xMMm(p[G])}.\Bigl{\{}x\in\zeta\bigl{(}\mathbb{Q}_{p}[G]\bigr{)}:\text{ if }m>0\text{ and }M\in\operatorname{M}_{m}\bigl{(}\mathbb{Z}_{p}[G]\bigr{)}\text{ then }xM^{\prime}\in\operatorname{M}_{m}\bigl{(}\mathbb{Z}_{p}[G]\bigr{)}\Bigr{\}}.

For more details about this module see Remark 5.4 below.

Proposition 5.2.

Assume that AA and FF satisfy the hypotheses (a)(i) and that the equality (17) is valid. Fix a surjective homomorphism of p[G]\mathbb{Z}_{p}[G]-modules

π:Selp(AF)Π\pi:\operatorname{Sel}_{p}(A_{F})^{\vee}\to\Pi

as above and an element aa of 𝒜p(G)\mathcal{A}_{p}(G). For each homomorphism θ:At(F)pA(F)p\theta:A^{t}(F)_{p}\to A(F)_{p}^{*} of p[G]\mathbb{Z}_{p}[G]-modules and each isomorphism of fields j:pj:\mathbb{C}\cong\mathbb{C}_{p} set

Rj(θ):=Nrdp[G]((ppθ)(λA,FNT,j)1).R_{j}(\theta):=\operatorname{Nrd}_{\mathbb{C}_{p}[G]}\bigl{(}(\mathbb{C}_{p}\otimes_{\mathbb{Z}_{p}}\theta)\circ(\lambda_{A,F}^{{\rm NT},j})^{-1}\bigr{)}.

Then for each such θ\theta and jj the element aRj(θ)j(A,F/k)epra\,R_{j}(\theta)\,j_{*}(\mathcal{L}^{\raisebox{1.0pt}{$\scriptstyle\star$}}_{A,F/k})\,e_{\rm pr} belongs to p[G]\mathbb{Z}_{p}[G] and annihilates both modules Xp(AFt)\hbox{\russ\char 88\relax}_{p}(A^{t}_{F}) and Πnpr\Pi^{\rm npr}. In particular, the element

A,F/k:=ψIr(G)eψLSr(A,ψˇ,1)τ(,Indkψ)dΩ(A,ψ)w(ψ)d\mathcal{L}_{A,F/k}:=\sum_{\psi\in\operatorname{Ir}(G)}e_{\psi}\,\frac{L_{S_{\rm r}}(A,\check{\psi},1)\cdot\tau^{*}\bigl{(}\mathbb{Q},\operatorname{Ind}_{k}^{\mathbb{Q}}\psi\bigr{)}^{d}}{\Omega(A,\psi)\cdot w_{\infty}(\psi)^{d}}

is such that aj(A,F/k)a\cdot j_{*}(\mathcal{L}_{A,F/k}) belongs to p[G]\mathbb{Z}_{p}[G] and annihilates Xp(AFt)\hbox{\russ\char 88\relax}_{p}(A^{t}_{F}).

Proof.

Hypothesis (a) implies that the module At(F)pA^{t}(F)_{p} is p\mathbb{Z}_{p}-free. Lemma 4.1 therefore implies, in the terminology of [17], that RΓf(k,Tp,F(A))R\Gamma_{f}\bigl{(}k,T_{p,F}(A)\bigr{)} is an admissible complex of p[G]\mathbb{Z}_{p}[G]-modules and equation (17) asserts that j(A,F/k)j_{*}(\mathcal{L}^{\raisebox{1.0pt}{$\scriptstyle\star$}}_{A,F/k}) is a characteristic element for the pair (RΓf(k,Tp,F(A)),(λA,FNT,j)1)\bigl{(}R\Gamma_{f}(k,T_{p,F}(A)),\,(\lambda^{{\rm NT},j}_{A,F})^{-1}\bigr{)}. The first claim is therefore a direct consequence of [17, Corollary 2.13] and the isomorphisms

Xp(AFt)Xp(AF)(Selp(AF))torH2(RΓf(k,Tp,F(A)))tor,\hbox{\russ\char 88\relax}_{p}(A^{t}_{F})\cong\hbox{\russ\char 88\relax}_{p}(A_{F})^{\vee}\cong\bigl{(}\operatorname{Sel}_{p}(A_{F})^{\vee}\bigr{)}_{\rm tor}\cong H^{2}\bigl{(}R\Gamma_{f}(k,T_{p,F}(A))\bigr{)}_{\rm tor},

where the first isomorphism is induced by the Cassels-Tate pairing, the second follows from the short exact sequence (1) (with L=FL=F) and the last is a consequence of Lemma 4.1.

To deduce the second claim we note that, under our hypotheses (a)(i), the natural projection map

Selp(AF)(Selp(AF))GSelp(Ak)A(k)p=:Π\operatorname{Sel}_{p}(A_{F})^{\vee}\to\bigl{(}\operatorname{Sel}_{p}(A_{F})^{\vee}\bigr{)}_{G}\cong\operatorname{Sel}_{p}(A_{k})^{\vee}\to A(k)_{p}^{*}=:\Pi

is a homomorphism π\pi of the required type. Further, in this case one has Πpr=0\Pi^{\rm pr}=0 and so hypothesis (i) implies Υpr\Upsilon_{\rm pr} is equal to {ψIr(G):LSr(A,ψˇ,1)0}\bigl{\{}\psi\in\operatorname{Ir}(G):L_{S_{\rm r}}(A,\check{\psi},1)\not=0\bigr{\}} and hence that A,F/k=A,F/kepr=A,F/kepr\mathcal{L}_{A,F/k}=\mathcal{L}_{A,F/k}\,e_{\rm pr}=\mathcal{L}^{\raisebox{1.0pt}{$\scriptstyle\star$}}_{A,F/k}\,e_{\rm pr}. It therefore suffices to show that Rj(θ)epr=eprR_{j}(\theta)e_{\rm pr}=e_{\rm pr} and this is true because in this case the space eprHom(A(F),)e_{\rm pr}\operatorname{Hom}_{\mathbb{C}}\bigl{(}\mathbb{C}\cdot A(F),\mathbb{C}\bigr{)} vanishes. ∎

Remark 5.3.

The conjectural equality (17) implies, via Proposition 5.2, a family of explicit congruence relations between the complex numbers A(ψ)A(\psi) that are defined by the equalities A(ψ)eψ=aRj(θ)j(A,F/k)eψA(\psi)\,e_{\psi}=a\,R_{j}(\theta)\,j_{*}(\mathcal{L}^{\raisebox{1.0pt}{$\scriptstyle\star$}}_{A,F/k})\,e_{\psi} as ψ\psi varies over Υpr\Upsilon_{\rm pr}. This is because

aRj(θ)j(A,F/k)epr=ψΥprA(ψ)eψ=|G|1gGgψΥprψ(1)ψˇ(g)A(ψ)a\,R_{j}(\theta)\,j_{*}(\mathcal{L}^{\raisebox{1.0pt}{$\scriptstyle\star$}}_{A,F/k})\,e_{\rm pr}=\sum_{\psi\in\Upsilon_{\rm pr}}A(\psi)\,e_{\psi}=\,|G|^{-1}\sum_{g\in G}g\sum_{\psi\in\Upsilon_{\rm pr}}\psi(1)\,\check{\psi}(g)\,A(\psi)

and this sum belongs to ζ(p[G])\zeta\bigl{(}\mathbb{Z}_{p}[G]\bigr{)} if and only if the elements A(ψ)A(\psi) satisfy all of the following conditions:

  • (i)

    A(ψ)p[ψ]A(\psi)\in\mathbb{Z}_{p}[\psi] for all ψΥpr\psi\in\Upsilon_{\rm pr};

  • (ii)

    α(A(ψ))=A(αψ)\alpha(A(\psi))=A(\alpha\circ\psi) for all ψΥpr\psi\in\Upsilon_{\rm pr} and αGp(ψ)/p\alpha\in G_{\mathbb{Q}_{p}(\psi)/\mathbb{Q}_{p}};

  • (iii)

    ψΥprψ(1)ψˇ(g)A(ψ)0(mod|G|p)\sum_{\psi\in\Upsilon_{\rm pr}}\psi(1)\check{\psi}(g)A(\psi)\equiv 0\pmod{|G|\mathbb{Z}_{p}} for all gGg\in G.

Remark 5.4.

Ideals of the form 𝒜p(G)\mathcal{A}_{p}(G) were introduced by Nickel in [41] and have been computed extensively by Johnston and Nickel in [34]. For example, if M=ImM=I_{m}, then M=ImM^{\prime}=I_{m} so that 𝒜p(G)ζ(p[G])\mathcal{A}_{p}(G)\subseteq\zeta\bigl{(}\mathbb{Z}_{p}[G]\bigr{)} and it is shown in loc. cit. that this inclusion is an equality if and only if the order of the commutator subgroup of GG is not divisible by pp. More generally, for each MM in Md(p[G])\operatorname{M}_{d}\bigl{(}\mathbb{Z}_{p}[G]\bigr{)} the matrix MM^{\prime} belongs to Mm()\operatorname{M}_{m}(\mathcal{M}) for any maximal order \mathcal{M} in p[G]\mathbb{Q}_{p}[G] that contains p[G]\mathbb{Z}_{p}[G] (cf. [41, Lemma. 4.1]) and so Jacobinski’s description in [33] of the central conductor of \mathcal{M} in p[G]\mathbb{Z}_{p}[G] implies, for example, that for any pc\mathbb{Q}_{p}^{c}-valued character ψ\psi of GG the element ψ(1)1|G|eψ\psi(1)^{-1}|G|e_{\psi} belongs to p[ψ]p𝒜p(G)\mathbb{Z}_{p}[\psi]\otimes_{\mathbb{Z}_{p}}\mathcal{A}_{p}(G) where p[ψ]\mathbb{Z}_{p}[\psi] is the subring of pc\mathbb{Q}_{p}^{c} that is generated over p\mathbb{Z}_{p} by the values of ψ\psi. This gives an easy ‘lower bound’ on 𝒜p(G)\mathcal{A}_{p}(G) (but which is, in most cases, not best possible).

Remark 5.5.

For the explicit computation of terms of the form χj(A,F/k)\chi_{j}(A,F/k) and Rj(θ)R_{j}(\theta) occurring in Theorem 5.1 and Proposition 5.2 respectively in the case that GG is cyclic of pp-power order we refer the reader to [7] where, in addition, Theorem 5.1 is used in order to discuss further explicit (conjectural) properties of elements of the form aRj(θ)j(A,F/k)epra\,R_{j}(\theta)\,j_{*}(\mathcal{L}^{\raisebox{1.0pt}{$\scriptstyle\star$}}_{A,F/k})\,e_{\rm pr} in such settings.

5.2. The case of projective Mordell-Weil groups

In the rest of this article we consider the conjectural equality (17) in the case that A(F)pA(F)_{p} and At(F)pA^{t}(F)_{p} are both projective p[G]\mathbb{Z}_{p}[G]-modules. (In this regard, note that if AA is principally polarised, then the p[G]\mathbb{Z}_{p}[G]-modules A(F)pA(F)_{p} and At(F)pA^{t}(F)_{p} are isomorphic.) This corresponds to taking the module Π\Pi in Proposition 5.2 to be equal to Πpr=A(F)p\Pi^{\rm pr}=A(F)_{p}^{*} and the more general case that Π=A(F)p\Pi=A(F)_{p}^{*} is a trivial source p[G]\mathbb{Z}_{p}[G]-module will be considered in a future article.

In the following result we use the notion of non-commutative Fitting invariant that was introduced by Parker [44] and studied further by Nickel [41].

Proposition 5.6.

Assume that AA and FF satisfy (a)(i). Assume also that A(F)pA(F)_{p} and At(F)pA^{t}(F)_{p} are both projective p[G]\mathbb{Z}_{p}[G]-modules.

Then the projective dimension of the p[G]\mathbb{Z}_{p}[G]-module Xp(AFt)\hbox{\russ\char 88\relax}_{p}(A^{t}_{F}) is at most one and there exists an isomorphism ιp:A(F)pAt(F)p\iota_{p}:A(F)^{*}_{p}\cong A^{t}(F)_{p} of p[G]\mathbb{Z}_{p}[G]-modules.

For each ψIr(G)\psi\in\operatorname{Ir}(G) now set

Rψ,ιpNT,j(A/F):=det((ppιp)λA,FNT,j|(VjψAt(F))G)pp×.R_{\psi,\iota_{p}}^{{\rm NT},j}(A_{/F}):=\det{}_{\mathbb{C}_{p}}\Bigl{(}(\mathbb{C}_{p}\otimes_{\mathbb{Z}_{p}}\iota_{p})\circ\lambda^{{\rm NT},j}_{A,F}\Bigm{|}\bigl{(}V_{j\circ\psi}\otimes_{\mathbb{Z}}A^{t}(F)\bigr{)}^{G}\Bigr{)}\in\mathbb{C}_{p}^{\times}.

Then the equality (17) is valid if and only if the (non-commutative) Fitting invariant of the p[G]\mathbb{Z}_{p}[G]-module Xp(AFt)\hbox{\russ\char 88\relax}_{p}(A^{t}_{F}) is generated by the set of elements of the form uA,F/k,j,ιpu\cdot\mathcal{L}^{\raisebox{1.0pt}{$\scriptstyle\star$}}_{A,F/k,j,\iota_{p}} where uu is in Nrdp[G](K1(p[G]))\operatorname{Nrd}_{\mathbb{Q}_{p}[G]}(K_{1}(\mathbb{Z}_{p}[G])) and

A,F/k,j,ιp:=ψIr(G)j(eψLSr(A,ψˇ,1)τ(,Indkψ)dΩ(A,ψ)w(ψ)d)1Rψ,ιpNT,j(A/F).\mathcal{L}^{\raisebox{1.0pt}{$\scriptstyle\star$}}_{A,F/k,j,\iota_{p}}:=\sum_{\psi\in\operatorname{Ir}(G)}j_{*}\biggl{(}e_{\psi}\cdot\frac{L^{\raisebox{2.0pt}{$\scriptstyle\star$}}_{S_{\rm r}}(A,\check{\psi},1)\cdot\tau^{*}\bigl{(}\mathbb{Q},\operatorname{Ind}_{k}^{\mathbb{Q}}\psi\bigr{)}^{d}}{\Omega(A,\psi)\cdot w_{\infty}(\psi)^{d}}\biggr{)}\frac{1}{R_{\psi,\iota_{p}}^{{\rm NT},j}(A_{/F})}.
Proof.

The assumed projectivity of A(F)pA(F)_{p} implies that its p\mathbb{Z}_{p}-linear dual A(F)pA(F)_{p}^{*} is also a projective p[G]\mathbb{Z}_{p}[G]-module. The existence of an isomorphism ιp\iota_{p} therefore follows from Swan’s Theorem [22, Theorem 32.1] and the fact that λA,FNT,j\lambda^{{\rm NT},j}_{A,F} induces an isomorphism of p[G]\mathbb{C}_{p}[G]-modules pAt(F)pHomp(pA(F),p)=pA(F)p\mathbb{C}_{p}\cdot A^{t}(F)_{p}\cong\operatorname{Hom}_{\mathbb{C}_{p}}\bigl{(}\mathbb{C}_{p}\otimes_{\mathbb{Z}}A(F),\mathbb{C}_{p}\bigr{)}=\mathbb{C}_{p}\cdot A(F)^{*}_{p}.

We now write CC^{\bullet} for the complex At(F)p0A(F)pA^{t}(F)_{p}\xrightarrow{0}A(F)^{*}_{p}, where the first term occurs in degree one. Then, since A(F)pA(F)_{p}^{*} is a projective p[G]\mathbb{Z}_{p}[G]-module, we may choose a p[G]\mathbb{Z}_{p}[G]-equivariant section to the natural surjection

H2(RΓf(k,Tp,F(A)))Selp(AF)A(F)p=H2(C).H^{2}\bigl{(}R\Gamma_{f}(k,T_{p,F}(A))\bigr{)}\cong\operatorname{Sel}_{p}(A_{F})^{\vee}\to A(F)^{*}_{p}=H^{2}(C^{\bullet}).

Since the kernel of this surjection is isomorphic (via the Cassels-Tate pairing) to Xp(AFt)\hbox{\russ\char 88\relax}_{p}(A^{t}_{F}), such a section induces a short exact sequence of p[G]\mathbb{Z}_{p}[G]-modules

0H2(C)H2(RΓf(k,Tp,F(A)))Xp(AFt)00\to H^{2}(C^{\bullet})\to H^{2}\bigl{(}R\Gamma_{f}(k,T_{p,F}(A))\bigr{)}\to\hbox{\russ\char 88\relax}_{p}(A^{t}_{F})\to 0

and hence also an exact triangle in Dp(p[G])D^{\rm p}\bigl{(}\mathbb{Z}_{p}[G]\bigr{)} of the form

CRΓf(k,Tp,F(A))Xp(AFt)[2]C[1].C^{\bullet}\to R\Gamma_{f}(k,T_{p,F}(A))\to\hbox{\russ\char 88\relax}_{p}(A^{t}_{F})[-2]\to C^{\bullet}[1].

In particular, since Xp(AFt)[2]\hbox{\russ\char 88\relax}_{p}(A^{t}_{F})[-2] belongs to Dp(p[G])D^{\rm p}\bigl{(}\mathbb{Z}_{p}[G]\bigr{)}, the projective dimension of the finite p[G]\mathbb{Z}_{p}[G]-module Xp(AFt)\hbox{\russ\char 88\relax}_{p}(A^{t}_{F}) is finite, and hence at most one (by [12, Chapter VI, (8.12)]), as claimed.

Further, if we use the cohomology sequence of the above triangle to identify pHi(C)\mathbb{C}_{p}\cdot H^{i}(C^{\bullet}) and pHi(RΓf(k,Tp,F(A))[1])\mathbb{C}_{p}\cdot H^{i}\bigl{(}R\Gamma_{f}(k,T_{p,F}(A))[1]\bigr{)} in all degrees ii, then this triangle combines with the additivity criterion of [11, Corollary 6.6] to imply that the element χj(A,F/k)=χG,p(RΓf(k,Tp,F(A)),λA,FNT,j)\chi_{j}(A,F/k)=-\chi_{G,p}\bigl{(}R\Gamma_{f}(k,T_{p,F}(A)),\lambda^{{\rm NT},j}_{A,F}\bigr{)} is equal to

χG,p(\displaystyle-\chi_{G,p}\bigl{(} C,λA,FNT,j)χG,p(Xp(AFt)[2],0)\displaystyle C^{\bullet},\lambda^{{\rm NT},j}_{A,F}\bigr{)}-\chi_{G,p}\bigl{(}\hbox{\russ\char 88\relax}_{p}(A^{t}_{F})[-2],0\bigr{)}
=χG,p(C,λA,FNT,j)+χG,p(C,(ppιp)1)χG,p(Xp(AFt)[2],0)\displaystyle=-\chi_{G,p}\bigl{(}C^{\bullet},\lambda^{{\rm NT},j}_{A,F}\bigr{)}+\chi_{G,p}\bigl{(}C^{\bullet},(\mathbb{C}_{p}\otimes_{\mathbb{Z}_{p}}\iota_{p})^{-1}\bigr{)}-\chi_{G,p}\bigl{(}\hbox{\russ\char 88\relax}_{p}(A^{t}_{F})[-2],0\bigr{)}
=δG,p(Nrdp[G]((ppιp)λA,FNT,j))χG,p(Xp(AFt)[2],0).\displaystyle=\delta_{G,p}\bigl{(}\operatorname{Nrd}_{\mathbb{C}_{p}[G]}((\mathbb{C}_{p}\otimes_{\mathbb{Z}_{p}}\iota_{p})\circ\lambda^{{\rm NT},j}_{A,F})\bigr{)}-\chi_{G,p}\bigl{(}\hbox{\russ\char 88\relax}_{p}(A^{t}_{F})[-2],0\bigr{)}.

Here the first displayed equality is valid as χG,p(C,(ppιp)1)=0\chi_{G,p}(C^{\bullet},(\mathbb{C}_{p}\otimes_{\mathbb{Z}_{p}}\iota_{p})^{-1})=0 and the second because the difference χG,p(C,λA,FNT,j)+χG,p(C,(ppιp)1)-\chi_{G,p}(C^{\bullet},\lambda^{{\rm NT},j}_{A,F})+\chi_{G,p}(C^{\bullet},(\mathbb{C}_{p}\otimes_{\mathbb{Z}_{p}}\iota_{p})^{-1}) is represented by the triple (At(F)p,(ppιp)λA,FNT,j,At(F)p)\bigl{(}A^{t}(F)_{p},(\mathbb{C}_{p}\otimes_{\mathbb{Z}_{p}}\iota_{p})\circ\lambda^{{\rm NT},j}_{A,F},A^{t}(F)_{p}\bigr{)}.

The equality (17) is therefore valid if and only if one has

δG,p(j(A,F/k)\displaystyle\delta_{G,p}\Bigl{(}j_{*}(\mathcal{L}^{\raisebox{1.0pt}{$\scriptstyle\star$}}_{A,F/k})\cdot Nrdp[G]((ppιp)λA,FNT,j)1)\displaystyle\operatorname{Nrd}_{\mathbb{C}_{p}[G]}\bigl{(}(\mathbb{C}_{p}\otimes_{\mathbb{Z}_{p}}\iota_{p})\circ\lambda^{{\rm NT},j}_{A,F}\bigr{)}^{-1}\Bigr{)}
=δG,p(j(A,F/k))δG,p(Nrdp[G]((ppιp)λA,FNT,j))\displaystyle=\delta_{G,p}\bigl{(}j_{*}(\mathcal{L}^{\raisebox{1.0pt}{$\scriptstyle\star$}}_{A,F/k})\bigr{)}-\delta_{G,p}\Bigl{(}\operatorname{Nrd}_{\mathbb{C}_{p}[G]}\bigl{(}(\mathbb{C}_{p}\otimes_{\mathbb{Z}_{p}}\iota_{p})\circ\lambda^{{\rm NT},j}_{A,F}\bigr{)}\Bigr{)}
=χj(A,F/k)δG,p(Nrdp[G]((ppιp)λA,FNT,j))\displaystyle=\chi_{j}(A,F/k)-\delta_{G,p}\Bigl{(}\operatorname{Nrd}_{\mathbb{C}_{p}[G]}\bigl{(}(\mathbb{C}_{p}\otimes_{\mathbb{Z}_{p}}\iota_{p})\circ\lambda^{{\rm NT},j}_{A,F}\bigr{)}\Bigr{)}
=χG,p(Xp(AFt)[2],0).\displaystyle=-\chi_{G,p}\bigl{(}\hbox{\russ\char 88\relax}_{p}(A^{t}_{F})[-2],0\bigr{)}.

In addition, one has Nrdp[G]((ppιp)λA,FNT,j)ejψ=Rψ,ιpNT,j(A/F)ejψ\operatorname{Nrd}_{\mathbb{C}_{p}[G]}\bigl{(}(\mathbb{C}_{p}\otimes_{\mathbb{Z}_{p}}\iota_{p})\circ\lambda^{{\rm NT},j}_{A,F}\bigr{)}e_{j\circ\psi}=R_{\psi,\iota_{p}}^{{\rm NT},j}(A_{/F})e_{j\circ\psi} for each ψ\psi in Ir(G)\operatorname{Ir}(G) and so

j(A,F/k)Nrdp[G]((ppιp)λA,FNT,j)1=A,F/k,j,ιp.j_{*}(\mathcal{L}^{\raisebox{1.0pt}{$\scriptstyle\star$}}_{A,F/k})\cdot\operatorname{Nrd}_{\mathbb{C}_{p}[G]}\bigl{(}(\mathbb{C}_{p}\otimes_{\mathbb{Z}_{p}}\iota_{p})\circ\lambda^{{\rm NT},j}_{A,F}\bigr{)}^{-1}=\mathcal{L}^{\raisebox{1.0pt}{$\scriptstyle\star$}}_{A,F/k,j,\iota_{p}}.

Given this, a straightforward exercise (comparing the explicit definitions of refined Euler characteristic and non-commutative Fitting invariants) shows that the above formula for χG,p(Xp(AFt)[2],0)\chi_{G,p}\bigl{(}\hbox{\russ\char 88\relax}_{p}(A^{t}_{F})[-2],0\bigr{)} is valid if and only if Fitp[G](Xp(AFt))\operatorname{Fit}_{\mathbb{Z}_{p}[G]}\bigl{(}\hbox{\russ\char 88\relax}_{p}(A^{t}_{F})\bigr{)} is generated by the set of elements uA,F/k,j,ιpu\cdot\mathcal{L}^{\raisebox{1.0pt}{$\scriptstyle\star$}}_{A,F/k,j,\iota_{p}} with uu in Nrdp[G](K1(p[G]))\operatorname{Nrd}_{\mathbb{Q}_{p}[G]}(K_{1}(\mathbb{Z}_{p}[G])), as claimed. ∎

5.3. Dihedral congruences for elliptic curves

We now investigate the criterion of Proposition 5.6 in the case that AA is an elliptic curve (so A=AtA=A^{t}) and F/kF/k is dihedral (in the sense of Mazur and Rubin [40]).

Thus, as before, we have an odd prime pp and a Galois extension F/kF/k of group GG with pp-Sylow subgroup PP and we assume that PP is an abelian (normal) subgroup of GG of index two and that the conjugation action of any lift to GG of the generator of G/PG/P inverts elements of PP. In particular, the degree of F/kF/k is equal to 2pn2p^{n} for some n1n\geq 1 and K/kK/k is a quadratic extension. We fix an element τ\tau of order 22 in GG. We set 𝟏:=𝟏G\boldsymbol{1}:=\boldsymbol{1}_{G} and write ϵ\epsilon for the unique non-trivial linear character of GG.

In the following result, we will be interested in the case when AA has rank one over KK. In this case we write ρA\rho_{A} for the unique linear character which does not occur in the [G]\mathbb{C}[G]-module A(K)\mathbb{C}\otimes_{\mathbb{Z}}A(K). Hence ρA=𝟏\rho_{A}=\boldsymbol{1} if rk(Ak)=0\operatorname{rk}(A_{k})=0 and ρA=ϵ\rho_{A}=\epsilon otherwise.

We also set :=[12]\mathbb{Z}^{\prime}:=\mathbb{Z}\bigl{[}\tfrac{1}{2}\bigr{]}.

Proposition 5.7.

Assume that the elliptic curve AA, odd prime pp and dihedral extension F/kF/k satisfy the hypotheses (a)(e) and (g), that all places above pp split in K/kK/k, that rk(AK)=1\operatorname{rk}(A_{K})=1 and that Xp(AK)=0\hbox{\russ\char 88\relax}_{p}(A_{K})=0.

Then Xp(AF)=0\hbox{\russ\char 88\relax}_{p}(A_{F})=0 and there is a point QQ in A(F)A(F) with τ(Q)=(1)rk(Ak)Q\tau(Q)=-(-1)^{\operatorname{rk}(A_{k})}\,Q which generates a [G]\mathbb{Z}^{\prime}[G]-module that is isomorphic to [G](1(1)rk(Ak)τ)\mathbb{Z}^{\prime}[G]\bigl{(}1-(-1)^{\operatorname{rk}(A_{k})}\,\tau\bigr{)} and has finite, prime-to-pp, index in A(F)\mathbb{Z}^{\prime}\otimes_{\mathbb{Z}}A(F).

In particular, one has rρA=0r_{\rho_{A}}=0 and rψ=1r_{\psi}=1 for all ψIr(G){ρA}\psi\in\operatorname{Ir}(G)\setminus\{\rho_{A}\}.

Proof.

Under the stated hypotheses, [18, Corollary 2.10(ii)] implies that Xp(AF)\hbox{\russ\char 88\relax}_{p}(A_{F}) vanishes, that Selp(AF)\operatorname{Sel}_{p}(A_{F})^{\vee} is a projective p[G]\mathbb{Z}_{p}[G]-module and that the multiplicity with which each ρ\rho in Ir(P)\operatorname{Ir}(P) occurs in the representation pSelp(AF)\mathbb{C}_{p}\cdot\operatorname{Sel}_{p}(A_{F})^{\vee} is equal to one.

Roiter’s Lemma (cf. [22, (31.6)]) therefore implies (via the exact sequence (1)) the existence of an exact sequence of [G]\mathbb{Z}^{\prime}[G]-modules

0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}[G](1(1)rk(Ak)τ)\textstyle{\mathbb{Z}^{\prime}[G](1-(-1)^{\operatorname{rk}(A_{k})}\tau)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}A(F)tf\textstyle{\mathbb{Z}^{\prime}\otimes_{\mathbb{Z}}A(F)_{\rm tf}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}X\textstyle{X\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0,\textstyle{0,}

where the group XX is both finite and of order prime to pp.

Since the group A(F)torA(F)_{\rm tor} is also finite of order prime to pp (by hypothesis (a)) it follows that any point QQ of A(F)A(F) whose projection in A(F)tfA(F)_{\rm tf} is equal to the image of 1(1)rk(Ak)τ1-(-1)^{\operatorname{rk}(A_{k})}\tau multiplied by a large enough power of 2 has the properties described above.

The above description of A(F)\mathbb{Z}^{\prime}\otimes_{\mathbb{Z}}A(F) also implies that the [G]\mathbb{C}[G]-module A(F)\mathbb{C}\otimes_{\mathbb{Z}}A(F) is isomorphic to [G](1(1)rk(Ak)τ)\mathbb{C}[G]\bigl{(}1-(-1)^{\operatorname{rk}(A_{k})}\tau\bigr{)} and the claimed formulas for rψr_{\psi} are then easily verified by explicit computation. ∎

For each subgroup HH of GG and character ρ\rho in Ir(H)\operatorname{Ir}(H) we set

Tρ:=hHρ(h1)hζ([H]).T_{\rho}:=\sum_{h\in H}\rho(h^{-1})h\in\zeta\bigl{(}\mathbb{C}[H]\bigr{)}.

For any point RR of A(F)A(F) and any ψIr(G)\psi\in\operatorname{Ir}(G) we then define a non-zero complex number

hF,ψ(R):=ψ(1)2|G|Tψ(R),Tψˇ(R)Fh_{F,\psi}(R):=\frac{\psi(1)}{2|G|}\cdot\Bigl{\langle}T_{\psi}(R),\,T_{\check{\psi}}(R)\Bigr{\rangle}_{F}

where ,F\langle\cdot,\cdot\rangle_{F} is the \mathbb{C}-linear extension of the Néron-Tate height on AA, defined relative to the field FF.

In the next result we assume the hypotheses of Proposition 5.7 to be satisfied and fix a point QQ as in that result. For each ψIr(G)\psi\in\operatorname{Ir}(G) we then obtain a non-zero complex number 𝒬ψ=𝒬Q,ψ\mathcal{Q}_{\psi}=\mathcal{Q}_{Q,\psi} by setting

(18) 𝒬ψ=uψdψLSr(A,ψ,1)Ω(A,ψ)Hψ\mathcal{Q}_{\psi}=u_{\psi}\cdot\frac{\sqrt{d_{\psi}}\cdot L^{\raisebox{2.0pt}{$\scriptstyle\star$}}_{S_{\rm r}}(A,\psi,1)}{\Omega(A,\psi)\cdot H_{\psi}}

where the quantity uψu_{\psi} is as in (16). Further we define

d𝟏=|dk|,dϵ=|dK/dk|, and dIndPG(χ)=|dK|Nf(χ)d_{\boldsymbol{1}}=|d_{k}|,\quad d_{\epsilon}=|d_{K}/d_{k}|,\quad\text{ and }\quad d_{\operatorname{Ind}^{G}_{P}(\chi)}=|d_{K}|\cdot N\!f(\chi)

for all χIr(P){𝟏P}\chi\in\operatorname{Ir}(P)\setminus\{\boldsymbol{1}_{P}\} and

HρA=1 and Hψ=hF,ψ(Q),for all ψρA.H_{\rho_{A}}=1\qquad\text{ and }\qquad H_{\psi}=h_{F,\psi}(Q),\qquad\text{for all $\psi\neq\rho_{A}$.}

Here dEd_{E} denotes the absolute discriminant of a number field EE and for any finite dimensional complex character ψ\psi of GEG_{E} we write Nf(ψ)N\!f(\psi) for the absolute norm of its Artin conductor.

We also note that hypothesis (i) combines with Proposition 5.7 to imply that the leading term LSr(A,ψ,1)L^{\star}_{S_{\rm r}}(A,\psi,1) is equal to the value LSr(A,ψ,1)L_{S_{\rm r}}(A,\psi,1) for ψ=ρA\psi=\rho_{A} and to the first derivative LSr(A,ψ,1)L^{\prime}_{S_{\rm r}}(A,\psi,1) for ψ\psi in Ir(G){ρA}\operatorname{Ir}(G)\setminus\{\rho_{A}\}.

Theorem 5.8.

Fix an odd prime pp, a dihedral extension F/kF/k of degree 2pn2p^{n} and an elliptic curve AA over kk. Assume that AA and FF satisfy the hypotheses (a)(h), that all places above pp split in K/kK/k, that rk(AK)=1\operatorname{rk}(A_{K})=1 and that Xp(AK)=0\hbox{\russ\char 88\relax}_{p}(A_{K})=0. Fix a point QQ in A(F)A(F) as given by Proposition 5.7.

Then the equality (17) is valid if and only if the following conditions are satisfied.

  1. (i)

    For each ψ\psi in Ir(G)\operatorname{Ir}(G) the number 𝒬ψ\mathcal{Q}_{\psi} defined above belongs to (ψ)\mathbb{Q}(\psi), is a unit at all primes above pp and satisfies (𝒬ψ)α=𝒬ψα(\mathcal{Q}_{\psi})^{\alpha}=\mathcal{Q}_{\psi^{\alpha}} for all α\alpha in Gal((ψ)/)\operatorname{Gal}\bigl{(}{\mathbb{Q}(\psi)/\mathbb{Q}}\bigr{)}.

  2. (ii)

    For all πP\pi\in P one has a congruence

    (19) 𝒬𝟏G𝒬ϵχIr(P){𝟏P}χˇ(π)𝒬IndPG(χ)(modpn(p))\mathcal{Q}_{\boldsymbol{1}_{G}}\cdot\mathcal{Q}_{\epsilon}\equiv-\sum_{\chi\in\operatorname{Ir}(P)\setminus\{\boldsymbol{1}_{P}\}}\check{\chi}(\pi)\,\mathcal{Q}_{\operatorname{Ind}^{G}_{P}(\chi)}\pmod{p^{n}\mathbb{Z}_{(p)}}

    where (p)\mathbb{Z}_{(p)} denotes the localisation of \mathbb{Z} at pp.

Remark 5.9.

If QQ^{\prime} is any element of A(F)A(F) with Tψ(Q)0T_{\psi}(Q^{\prime})\neq 0 for all ψIr(G){ρA}\psi\in\operatorname{Ir}(G)\setminus\{\rho_{A}\}, then for each ψIr(G)\psi\in\operatorname{Ir}(G) one can define a non-zero complex number 𝒬ψ=𝒬Q,ψ\mathcal{Q}^{\prime}_{\psi}=\mathcal{Q}_{Q^{\prime},\psi} just as in (18). Our proof of Theorem 5.8 will actually show that if the conditions (i) and (ii) are valid with each 𝒬ψ\mathcal{Q}_{\psi} replaced by 𝒬ψ\mathcal{Q}^{\prime}_{\psi}, then QQ^{\prime} is a p[G]\mathbb{Z}_{p}[G]-generator of A(F)pA(F)_{p} and the equality (17) is valid.

Remark 5.10.

If n=1n=1, so the extension F/kF/k in Theorem 5.8 is dihedral of degree 2p2p, then each term χˇ(π)\check{\chi}(\pi) is congruent to 11 modulo the unique prime above pp in (χ)\mathbb{Q}(\chi) and so the congruences (19) reduce to a single congruence

(20) 𝒬𝟏G𝒬ϵ2ψIr(G)dim(ψ)=2𝒬ψ(modp(p)).\mathcal{Q}_{\boldsymbol{1}_{G}}\cdot\mathcal{Q}_{\epsilon}\equiv-2\sum_{\begin{subarray}{c}\psi\in\operatorname{Ir}(G)\\ \dim(\psi)=2\end{subarray}}\mathcal{Q}_{\psi}\pmod{p\mathbb{Z}_{(p)}}.
Proof of Theorem 5.8.

It suffices to prove that the criterion of Proposition 5.6 is valid if and only if both the condition (i) and the congruences in (19) are valid. Now, since Xp(AF)=0\hbox{\russ\char 88\relax}_{p}(A_{F})=0 (by [18, Corollary 2.10(ii)]), the criterion of Proposition 5.6 is equivalent to asking that the element A,F/k,j,ιp\mathcal{L}^{\raisebox{1.0pt}{$\scriptstyle\star$}}_{A,F/k,j,\iota_{p}} belongs to ker(δG,p)\ker(\delta_{G,p}). In view of Lemma 5.11 below it is thus enough to show that the condition (i) and the congruences in (19) are valid if and only if the element :=j(ψIr(G)𝒬ψeψ)\mathcal{L}^{\prime}:=j_{*}(\sum_{\psi\in\operatorname{Ir}(G)}\mathcal{Q}_{\psi}e_{\psi}) belongs to ker(δG,p)\ker(\delta_{G,p}).

To investigate the element δG,p()\delta_{G,p}(\mathcal{L}^{\prime}) we fix a maximal p\mathbb{Z}_{p}-order 𝔐p\mathfrak{M}_{p} in p[G]\mathbb{Q}_{p}[G] that contains p[G]\mathbb{Z}_{p}[G]. Then δG,p()\delta_{G,p}(\mathcal{L}^{\prime}) belongs to the subgroup K0(p[G],p[G])torK_{0}\bigl{(}\mathbb{Z}_{p}[G],\mathbb{Q}_{p}[G]\bigr{)}_{\rm tor} of K0(p[G],p[G])K_{0}\bigl{(}\mathbb{Z}_{p}[G],\mathbb{C}_{p}[G]\bigr{)} if and only if Nrdp[G](𝔐p×)\mathcal{L}^{\prime}\in\operatorname{Nrd}_{\mathbb{Q}_{p}[G]}\bigl{(}\mathfrak{M}_{p}^{\times}\bigr{)} and this condition is satisfied if and only if the conditions of Theorem 5.8(i) are valid (for more details of these equivalences see the proof of [16, Lemma 11]).

It thus suffices to show that an element =ψIr(G)ψeψ\mathcal{E}=\sum_{\psi\in\operatorname{Ir}(G)}\mathcal{E}_{\psi}e_{\psi} of Nrdp[G](𝔐p×)ψIr(G)p(ψ)\operatorname{Nrd}_{\mathbb{Q}_{p}[G]}\bigl{(}\mathfrak{M}_{p}^{\times}\bigr{)}\subset\bigoplus_{\psi\in\operatorname{Ir}(G)}\mathbb{Q}_{p}(\psi) belongs to ker(δG,p)\ker(\delta_{G,p}) if and only if the congruences in (19) are valid with each term 𝒬ψ\mathcal{Q}_{\psi} replaced by ψ\mathcal{E}_{\psi} and with (p)\mathbb{Z}_{(p)} replaced by p\mathbb{Z}_{p}. But, from the results of Lemma 5.12(i) and (ii) below, one has δG,p()=0\delta_{G,p}(\mathcal{E})=0 if and only if res()p[P]×\operatorname{res}(\mathcal{E})\in\mathbb{Z}_{p}[P]^{\times}, and by Lemma 5.12(iii) this is true if and only if the congruences in (19) are valid after making the changes described above.

This therefore completes the proof of Theorem 5.8. ∎

Lemma 5.11.

There exists an isomorphism of p[G]\mathbb{Z}_{p}[G]-modules ιp:A(F)pA(F)p\iota_{p}:A(F)^{*}_{p}\to A(F)_{p} such that A,F/k,j,ιp\mathcal{L}^{\raisebox{1.0pt}{$\scriptstyle\star$}}_{A,F/k,j,\iota_{p}} is equal to the element :=j(ψIr(G)𝒬ψeψ)\mathcal{L}^{\prime}:=j_{*}(\sum_{\psi\in\operatorname{Ir}(G)}\mathcal{Q}_{\psi}e_{\psi}) defined above.

Proof.

We claim first that for each ρIr(G)\rho\in\operatorname{Ir}(G) there is an equality

(21) τ(,Indkρ)w(ρ)={u𝟏G|dk|=(1)|Sr||dk|,if ρ=𝟏Guϵ|dK/dk|=(1)|Sr||dK/dk|,if ρ=ϵuρ|dK|Nf(χ),if ρ=IndPGχ\frac{\tau^{*}\bigl{(}\mathbb{Q},\operatorname{Ind}_{k}^{\mathbb{Q}}\rho\bigr{)}}{w_{\infty}(\rho)}=\begin{cases}u_{{\boldsymbol{1}}_{G}}\,\sqrt{|d_{k}|}=(-1)^{|S_{\rm r}|}\,\sqrt{|d_{k}|},&\text{if $\rho={\boldsymbol{1}}_{G}$}\\ u_{\epsilon}\,\sqrt{|d_{K}/d_{k}|}=(-1)^{|S_{\rm r}^{\prime}|}\,\sqrt{|d_{K}/d_{k}|},&\text{if $\rho=\epsilon$}\\ u_{\rho}\,\sqrt{|d_{K}|\,Nf(\chi)},&\text{if $\rho=\operatorname{Ind}_{P}^{G}\chi$}\end{cases}

where χIr(P){𝟏P}\chi\in\operatorname{Ir}(P)\setminus\{\boldsymbol{1}_{P}\} and with SrS_{\rm r}^{\prime} denoting the subset of SrS_{\rm r} comprising places which split in K/kK/k.

To prove this we note that for each ρ\rho in Ir(G)\operatorname{Ir}(G) one has

(22) uρ1τ(,Indkρ)=τ(,Indkρ)=(i|S||dk|)ρ(1)τ(k,ρ),u_{\rho}^{-1}\cdot\tau^{*}\bigl{(}\mathbb{Q},\operatorname{Ind}_{k}^{\mathbb{Q}}\rho\bigr{)}=\tau\bigl{(}\mathbb{Q},\operatorname{Ind}_{k}^{\mathbb{Q}}\rho\bigr{)}=\Bigl{(}i^{|S_{\mathbb{C}}|}\,\sqrt{|d_{k}|}\Bigr{)}^{\rho(1)}\cdot\tau(k,\rho),

where the first equality follows straight from the definition (16) and the second from the result of [37, Theorem 8.1(iii)].

Now if ρ=𝟏G\rho=\boldsymbol{1}_{G}, then τ(k,ρ)=1\tau(k,\rho)=1 and VρIw=pV_{\rho}^{I_{w}}=\mathbb{C}_{p} for all vv in SrS_{\rm r} so uρ=(1)|Sr|u_{\rho}=(-1)^{|S_{\rm r}|}, whilst it is clear that w(ρ)=i|S|w_{\infty}(\rho)=i^{|S_{\mathbb{C}}|}, and so in this case the equality (21) is an immediate consequence of (22).

Next we note that ϵ=IndPG(𝟏P)𝟏G\epsilon=\operatorname{Ind}_{P}^{G}({\boldsymbol{1}}_{P})-{\boldsymbol{1}}_{G} and hence that [37, Theorem 8.1(iii)] implies

τ(k,ϵ)=τ(k,IndPG(𝟏P))τ(k,𝟏G)=τ(k,IndPG(𝟏P))=i|Sr|dK/k=i|Sr||dK||dk|,\tau(k,\epsilon)=\frac{\tau\bigl{(}k,\operatorname{Ind}_{P}^{G}({\boldsymbol{1}}_{P})\bigr{)}}{\tau(k,{\boldsymbol{1}}_{G})}=\tau\bigl{(}k,\operatorname{Ind}_{P}^{G}({\boldsymbol{1}}_{P})\bigr{)}=i^{|S_{\mathbb{R}}^{\rm r}|}\sqrt{d_{K/k}}=i^{|S_{\mathbb{R}}^{\rm r}|}\cdot\frac{\sqrt{|d_{K}|}}{|d_{k}|},

where SrS_{\mathbb{R}}^{\rm r} is the set of real places of kk that ramify in K/kK/k and dK/kd_{K/k} the absolute norm of the different of K/kK/k. One then obtains the claimed equality (21) by substituting this formula for τ(k,ϵ)\tau(k,\epsilon) into (22) and then using that fact that w(ϵ)=i|S|+|Sr|w_{\infty}(\epsilon)=i^{|S_{\mathbb{C}}|+|S_{\mathbb{R}}^{\rm r}|} whilst uϵ=(1)|Sr|u_{\epsilon}=(-1)^{|S^{\prime}_{\rm r}|} since det(FrwVϵIw)\det\bigl{(}-\operatorname{Fr}_{w}\mid V_{\epsilon}^{I_{w}}\bigr{)} is equal to 1-1 for vSrv\in S_{\rm r}^{\prime} and to 11 for vSrSrv\in S_{\rm r}\setminus S_{\rm r}^{\prime}.

Finally we assume that ρ=IndPGχ\rho=\operatorname{Ind}_{P}^{G}\chi with χIr(P){𝟏P}.\chi\in\operatorname{Ir}(P)\setminus\{{\boldsymbol{1}}_{P}\}. Then, just as above, one finds that

τ(k,ρ)=\displaystyle\tau(k,\rho)= τ(k,IndPG(𝟏P))τ(K,χ)\displaystyle\tau(k,\operatorname{Ind}_{P}^{G}({\boldsymbol{1}}_{P}))\cdot\tau(K,\chi)
=\displaystyle= i|Sr||dK||dk|τ(K,χ)\displaystyle i^{|S_{\mathbb{R}}^{\rm r}|}\,\frac{\sqrt{|d_{K}|}}{|d_{k}|}\,\tau(K,\chi)
=\displaystyle= i|Sr||dK||dk|W(χˇ)W(χ)Nf(χ),\displaystyle i^{|S_{\mathbb{R}}^{\rm r}|}\,\frac{\sqrt{|d_{K}|}}{|d_{k}|}\,\frac{W(\check{\chi})}{W_{\infty}(\chi)}\,\sqrt{Nf(\chi)},

where W(χˇ)W(\check{\chi}) is the Artin root number of χˇ\check{\chi} and W(χ)W_{\infty}(\chi) the infinite part of the Artin root number of χ\chi and the final equality follows from the very definition of τ(K,χ)\tau(K,\chi). Now W(χ)=1W_{\infty}(\chi)=1 since no archimedean place ramifies in F/KF/K. In addition, the inductivity of Artin root numbers implies W(χˇ)=W(ρˇ)W(\check{\chi})=W(\check{\rho}) and so, since ρˇ\check{\rho} is an orthogonal character, the main result of Fröhlich and Queyrut in [29] implies that W(χˇ)=1W(\check{\chi})=1. Thus, upon substituting the last displayed expression into (22), and noting that w(ρ)=iρ(1)|S|+|Sr|w_{\infty}(\rho)=i^{\rho(1)|S_{\mathbb{C}}|+|S_{\mathbb{R}}^{\rm r}|}, one obtains the claimed equality (21) in this case.

Having proved (21), we now write θQ\theta_{Q} for the generator of the p[G]\mathbb{Z}_{p}[G]-module A(F)pA(F)_{p}^{*} that sends QQ to 11 and g(Q)g(Q) to zero for each non-trivial element gg of PP. We then define ιp\iota_{p} to be the isomorphism of p[G]\mathbb{Z}_{p}[G]-modules A(F)pA(F)pA(F)^{*}_{p}\to A(F)_{p} that sends θQ\theta_{Q} to QQ.

One computes that ((ppιp)λA,FNT,j)(Q)=j(RQ)(Q)\bigl{(}(\mathbb{C}_{p}\otimes_{\mathbb{Z}_{p}}\iota_{p})\circ\lambda^{{\rm NT},j}_{A,F}\bigr{)}(Q)=j_{*}(R_{Q})(Q), where RQR_{Q} is the resolvent element gPQ,g(Q)Fg\sum_{g\in P}\bigl{\langle}Q,g(Q)\bigr{\rangle}_{F}\cdot g in p[G]\mathbb{C}_{p}[G], and also that θQ(ιp(ϑ))=ϑ(Q)\theta_{Q}\bigl{(}\iota_{p}(\vartheta)\bigr{)}=\vartheta(Q) for all ϑA(F)p\vartheta\in A(F)_{p}^{*}.

Proposition 5.7 implies rρA=0r_{\rho_{A}}=0 so RρA,ιpNT,j(A/F)=1=HρAR_{\rho_{A},\iota_{p}}^{{\rm NT},j}(A_{/F})=1=H_{\rho_{A}} and also that for each ψIr(G){ρA}\psi\in\operatorname{Ir}(G)\setminus\{\rho_{A}\}, the complex vector space (VψA(F))G=[G]Tψ(Q)(V_{\psi}\otimes_{\mathbb{Z}}A(F))^{G}=\mathbb{C}[G]\cdot T_{\psi}(Q) has dimension one and hence Rψ,ιpNT,j(A/F)Tψ(Q)=((ppιp)λA,FNT,j)(Tψ(Q)).R_{\psi,\iota_{p}}^{{\rm NT},j}(A_{/F})\cdot T_{\psi}(Q)=\bigl{(}(\mathbb{C}_{p}\otimes_{\mathbb{Z}_{p}}\iota_{p})\circ\lambda^{{\rm NT},j}_{A,F}\bigr{)}\bigl{(}T_{\psi}(Q)\bigr{)}.

Now, for any ϑA(F)p\vartheta\in A(F)_{p}^{*} and PA(F)P\in A(F) one has Tψˇ(ϑ)(P)=ϑ(Tψˇ#(P))T_{\check{\psi}}(\vartheta)(P)=\vartheta\bigl{(}T_{\check{\psi}}^{\#}(P)\bigr{)} where we write xx#x\mapsto x^{\#} for the p\mathbb{Z}_{p}-linear involution on p[G]\mathbb{Z}_{p}[G] that inverts elements of GG. In addition, for each ψρA\psi\neq\rho_{A} one has

Tψˇ(θQ)(Tψ(Q))\displaystyle T_{\check{\psi}}(\theta_{Q})\Bigl{(}T_{\psi}(Q)\Bigr{)} =gGhGψˇ(g1)ψ(h1)θQ(g1hQ)\displaystyle=\sum_{g\in G}\sum_{h\in G}\check{\psi}(g^{-1})\,\psi(h^{-1})\,\theta_{Q}\Bigl{(}g^{-1}\,h\,Q\Bigr{)}
=gGψˇ(g1)ψ(g1)+(1)1rk(Ak)gGψˇ(g1)ψ(τg1)\displaystyle=\sum_{g\in G}\check{\psi}(g^{-1})\psi(g^{-1})+(-1)^{1-\operatorname{rk}(A_{k})}\cdot\sum_{g\in G}\check{\psi}(g^{-1})\,\psi(\tau g^{-1})

because g1hQg^{-1}hQ is a multiple of QQ only when h=gh=g or h=gτh=g\tau. The first sum here is always equal to |G||G|, while the second is equal to |G||G| for ψ=𝟏\psi=\boldsymbol{1}, to |G|-|G| for ψ=ϵ\psi=\epsilon and to 0 otherwise. Hence, writing ϑ\vartheta for λA,FNT,j(Tψ(Q))\lambda^{{\rm NT},j}_{A,F}\bigl{(}T_{\psi}(Q)\bigr{)}, we find that

2|G|ψ(1)Rψ,ιpNT,j\displaystyle\frac{2|G|}{\psi(1)}\,R_{\psi,\iota_{p}}^{{\rm NT},j} (A/F)=Tψˇ(θQ)((ppιp)λA,FNT,j(Tψ(Q)))\displaystyle(A_{/F})=T_{\check{\psi}}(\theta_{Q})\Bigl{(}(\mathbb{C}_{p}\otimes_{\mathbb{Z}_{p}}\iota_{p})\circ\lambda^{{\rm NT},j}_{A,F}\bigl{(}T_{\psi}(Q)\bigr{)}\Bigr{)}
=θQ(Tψˇ#((ppιp)(ϑ)))=θQ((ppιp)(Tψˇ#(ϑ)))\displaystyle=\theta_{Q}\Bigl{(}T_{\check{\psi}}^{\#}\bigl{(}(\mathbb{C}_{p}\otimes_{\mathbb{Z}_{p}}\iota_{p})(\vartheta)\bigr{)}\Bigr{)}=\theta_{Q}\Bigl{(}(\mathbb{C}_{p}\otimes_{\mathbb{Z}_{p}}\iota_{p})\bigl{(}T_{\check{\psi}}^{\#}(\vartheta)\bigr{)}\Bigr{)}
=(Tψˇ#(ϑ))(Q)=ϑ(Tψˇ(Q))=j(Tψ(Q),Tψˇ(Q)F)=2|G|ψ(1)j(hF,ψ(Q)).\displaystyle=\bigl{(}T_{\check{\psi}}^{\#}(\vartheta)\bigr{)}(Q)=\vartheta\bigl{(}T_{\check{\psi}}(Q)\bigr{)}=j(\bigl{\langle}T_{\psi}(Q),T_{\check{\psi}}(Q)\bigr{\rangle}_{F})=\frac{2|G|}{\psi(1)}\,j(h_{F,\psi}(Q)).

Upon substituting this equality and (21) into the definition of A,F/k,j,ιp\mathcal{L}^{\raisebox{1.0pt}{$\scriptstyle\star$}}_{A,F/k,j,\iota_{p}} one obtains the element \mathcal{L}^{\prime}, as required. ∎

In the next result we write 𝒪L\mathcal{O}_{L} for the valuation ring of a finite extension LL of p\mathbb{Q}_{p}.

Lemma 5.12.

Write 𝔐p\mathfrak{M}_{p}^{\prime} for the integral closure of p\mathbb{Z}_{p} in p[P]\mathbb{Q}_{p}[P].

  1. (i)

    The following diagram commutes

    Nrdp[G](𝔐p×)\textstyle{\operatorname{Nrd}_{\mathbb{Q}_{p}[G]}\bigl{(}\mathfrak{M}_{p}^{\times}\bigr{)}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}res\scriptstyle{\operatorname{res}}𝔐p×\textstyle{\mathfrak{M}^{\prime\times}_{p}}im(α:K1(𝔐p)K1(p[G]))\textstyle{\operatorname{im}\Bigl{(}\alpha:K_{1}\bigl{(}\mathfrak{M}_{p}\bigr{)}\to K_{1}\bigl{(}\mathbb{Q}_{p}[G]\bigr{)}\Bigr{)}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}res\scriptstyle{\operatorname{res}^{\prime}}Nrdp[G]\scriptstyle{\operatorname{Nrd}_{\mathbb{Q}_{p}[G]}}G,p\scriptstyle{\partial_{G,p}^{\prime}}im(β:K1(𝔐p)K1(p[P]))\textstyle{\operatorname{im}\Bigl{(}\beta:K_{1}\bigl{(}\mathfrak{M}^{\prime}_{p}\bigr{)}\to K_{1}\bigl{(}\mathbb{Q}_{p}[P]\bigr{)}\Bigr{)}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Nrdp[P]\scriptstyle{\operatorname{Nrd}_{\mathbb{Q}_{p}[P]}}P,p\scriptstyle{\partial_{P,p}^{\prime}}K0(p[G],p[G])tor\textstyle{K_{0}\bigl{(}\mathbb{Z}_{p}[G],\mathbb{Q}_{p}[G]\bigr{)}_{\rm tor}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}resP,0G\scriptstyle{\operatorname{res}^{G}_{P,0}}K0(p[P],p[P])tor.\textstyle{K_{0}\bigl{(}\mathbb{Z}_{p}[P],\mathbb{Q}_{p}[P]\bigr{)}_{\rm tor}.}

    Here res\operatorname{res} is the homomorphism that sends an element ψIr(G)ξψeψ\sum_{\psi\in\operatorname{Ir}(G)}\xi_{\psi}e_{\psi} in Nrdp[G](𝔐p×)ζ(p[G])×=ψp×eψ\operatorname{Nrd}_{\mathbb{Q}_{p}[G]}(\mathfrak{M}_{p}^{\times})\subset\zeta\bigl{(}\mathbb{C}_{p}[G]\bigr{)}^{\times}=\sum_{\psi}\mathbb{C}_{p}^{\times}e_{\psi} to ξ𝟏Gξϵe𝟏P+χIr(P){𝟏P}ξIndPGχeχ\xi_{\boldsymbol{1}_{G}}\xi_{\epsilon}e_{\boldsymbol{1}_{P}}+\sum_{\chi\in\operatorname{Ir}(P)\setminus\{\boldsymbol{1}_{P}\}}\xi_{\operatorname{Ind}_{P}^{G}\chi}e_{\chi}, next α\alpha and β\beta are the natural scalar extension homomorphisms, res\operatorname{res}^{\prime} and resP,0G\operatorname{res}^{G}_{P,0} the natural restriction homomorphisms and G,p\partial_{G,p}^{\prime} and P,p\partial_{P,p}^{\prime} the restrictions of the connecting homomorphisms G,p\partial_{G,p} and P,p\partial_{P,p}.

  2. (ii)

    The homomorphism resP,0G\operatorname{res}^{G}_{P,0} is injective.

  3. (iii)

    Fix :=χIr(P)χeχ\mathcal{E}:=\sum_{\chi\in\operatorname{Ir}(P)}\mathcal{E}_{\chi}e_{\chi} inside 𝔐p×=χIr(P)𝒪p(χ)×eχ\mathfrak{M}^{\prime\times}_{p}=\sum_{\chi\in\operatorname{Ir}(P)}\mathcal{O}_{\mathbb{Q}_{p}(\chi)}^{\times}e_{\chi}. Then \mathcal{E} belongs to p[P]×\mathbb{Z}_{p}[P]^{\times} if and only if χIr(P)χ(π)1χ\sum_{\chi\in\operatorname{Ir}(P)}\chi(\pi)^{-1}\mathcal{E}_{\chi} belongs to |P|p|P|\cdot\mathbb{Z}_{p} for all π\pi in PP.

Proof.

To prove claim (i) we fix a set of representatives Ir(P)\operatorname{Ir}(P)^{\dagger} of the orbits of the action of GpG_{\mathbb{Q}_{p}} on Ir(P)\operatorname{Ir}(P) and abbreviate the functor IndPG()\operatorname{Ind}^{G}_{P}(-) to IPG()\operatorname{I}^{G}_{P}(-).

The commutativity of the lower square of the diagram follows from the naturality of the long exact sequences of relative KK-theory. To consider the upper square we use the p\mathbb{Q}_{p}-algebra isomorphisms

ωG:p[G]p×p×χIr(P){𝟏P}M2(p(IPG(χ)))\omega_{G}:\mathbb{Q}_{p}[G]\cong\mathbb{Q}_{p}\times\mathbb{Q}_{p}\times\prod_{\chi\in\operatorname{Ir}(P)^{\dagger}\setminus\{\boldsymbol{1}_{P}\}}{\rm M}_{2}\bigl{(}\mathbb{Q}_{p}(\operatorname{I}^{G}_{P}(\chi))\bigr{)}

and ωP:p[P]χIr(P)p(χ)\omega_{P}:\mathbb{Q}_{p}[P]\cong\prod_{\chi\in\operatorname{Ir}(P)^{\dagger}}\mathbb{Q}_{p}(\chi) where ωG(g)=(1,ϵ(g),(SIPG(χ)(g))χ)\omega_{G}(g)=\bigl{(}1,\epsilon(g),(S_{\operatorname{I}_{P}^{G}(\chi)}(g))_{\chi}\bigr{)}, with SIPG(χ)S_{\operatorname{I}_{P}^{G}(\chi)} a representation GM2(p(IPG(χ))G\to{\rm M}_{2}\bigl{(}\mathbb{Q}_{p}(\operatorname{I}^{G}_{P}(\chi)) of character IPG(χ)\operatorname{I}_{P}^{G}(\chi), and ωP(π)=(χ(π))χ\omega_{P}(\pi)=(\chi(\pi))_{\chi} for each gGg\in G and πP\pi\in P. Taken together these isomorphisms induce a diagram

p××p××χIr(P)χ𝟏Pp(IPG(χ))×\textstyle{\mathbb{Q}_{p}^{\times}\times\mathbb{Q}_{p}^{\times}\times\prod\limits_{\begin{subarray}{c}\chi\in\operatorname{Ir}(P)^{\dagger}\\ \chi\neq\boldsymbol{1}_{P}\end{subarray}}\mathbb{Q}_{p}(\operatorname{I}^{G}_{P}(\chi))^{\times}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}res~\scriptstyle{\widetilde{\operatorname{res}}}χIr(P)p(χ)×\textstyle{\prod\limits_{\chi\in\operatorname{Ir}(P)^{\dagger}}\mathbb{Q}_{p}(\chi)^{\times}}K1(p)×K1(p)×χIr(P)χ𝟏PK1(M2(p(IPG(χ))))\textstyle{K_{1}(\mathbb{Q}_{p})\times K_{1}(\mathbb{Q}_{p})\times\prod\limits_{\begin{subarray}{c}\chi\in\operatorname{Ir}(P)^{\dagger}\\ \chi\neq\boldsymbol{1}_{P}\end{subarray}}K_{1}\Bigl{(}{\rm M}_{2}\bigl{(}\mathbb{Q}_{p}(\operatorname{I}^{G}_{P}(\chi))\bigr{)}\Bigr{)}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}(det,det,(det)χ)\scriptstyle{(\det,\det,(\det)_{\chi})}χIr(P)K1(p(χ))\textstyle{\prod\limits_{\chi\in\operatorname{Ir}(P)^{\dagger}}K_{1}(\mathbb{Q}_{p}(\chi))\ignorespaces\ignorespaces\ignorespaces\ignorespaces}(det)χ\scriptstyle{(\det)_{\chi}}K1(p[G])\textstyle{K_{1}\bigl{(}\mathbb{Q}_{p}[G]\bigr{)}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}resP,1G\scriptstyle{\operatorname{res}^{G}_{P,1}}K1(ωG)\scriptstyle{K_{1}(\omega_{G})}K1(p[P])\textstyle{K_{1}\bigl{(}\mathbb{Q}_{p}[P]\bigr{)}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}K1(ωP)\scriptstyle{K_{1}(\omega_{P})}

in which the left and right hand composite vertical arrows are equal to Nrdp[G]\operatorname{Nrd}_{\mathbb{Q}_{p}[G]} and Nrdp[P]\operatorname{Nrd}_{\mathbb{Q}_{p}[P]} and res~\widetilde{\operatorname{res}} is defined to make the diagram commute. To complete the proof of claim (i) it thus suffices to show res~\widetilde{\operatorname{res}} sends each element (α,α~,(aχ)χ)(\alpha,\tilde{\alpha},(a_{\chi})_{\chi}) to (αα~,(aχ)χ)(\alpha\tilde{\alpha},(a_{\chi})_{\chi}) and this follows from the argument used by Breuning in [10, Lemma 3.9].

To prove claim (ii) we recall a group is said to be p\mathbb{Q}_{p}-elementary if it is isomorphic to a group (/n)J(\mathbb{Z}/n\mathbb{Z})\rtimes J where JJ is an \ell-group for some prime \ell that is coprime to nn and the image of the homomorphism JAut(/n)Gal((e2πin)/)J\to\operatorname{Aut}(\mathbb{Z}/n\mathbb{Z})\cong\operatorname{Gal}(\mathbb{Q}(e^{\frac{2\pi i}{n}})/\mathbb{Q}) belongs to the decomposition subgroup of pp. This means that a subgroup of GG is p\mathbb{Q}_{p}-elementary if it is either a subgroup of PP or of the form C,τ\langle C,\tau^{\prime}\rangle where CC is a cyclic subgroup of PP and τ\tau^{\prime} an element of order 22. We consider the exact commutative diagram

K1(p[G])\textstyle{K_{1}\bigl{(}\mathbb{Z}_{p}[G]\bigr{)}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}α1\scriptstyle{\alpha_{1}}K1(p[G])\textstyle{K_{1}\bigl{(}\mathbb{Q}_{p}[G]\bigr{)}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}α2\scriptstyle{\alpha_{2}}K0(p[G],p[G])\textstyle{K_{0}\bigl{(}\mathbb{Z}_{p}[G],\mathbb{Q}_{p}[G]\bigr{)}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}α3\scriptstyle{\alpha_{3}}0\textstyle{0}limHK1(p[H])\textstyle{\varprojlim_{H}K_{1}\bigl{(}\mathbb{Z}_{p}[H]\bigr{)}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}limHK1(p[H])\textstyle{\varprojlim_{H}K_{1}\bigl{(}\mathbb{Q}_{p}[H]\bigr{)}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}limHK0(p[H],p[H])\textstyle{\varprojlim_{H}K_{0}\bigl{(}\mathbb{Z}_{p}[H],\mathbb{Q}_{p}[H]\bigr{)}}

where the limits are over all p\mathbb{Q}_{p}-elementary subgroups HH of GG. The transition maps are the homomorphisms induced by inclusions HHH\subseteq H^{\prime} and by maps of the form HgHg1H\to gHg^{-1} for gGg\in G and all vertical arrows are the natural restriction maps. By a theorem of Dress [27] (see also [43, Theorem 11.2]), the maps α1\alpha_{1} and α2\alpha_{2} are bijective and hence α3\alpha_{3} is injective. Now K0(p[H],p[H])torK_{0}\bigl{(}\mathbb{Z}_{p}[H],\mathbb{Q}_{p}[H]\bigr{)}_{\rm tor} is trivial whenever HH has order prime to pp and in [10, Proposition 3.2.(2)] Breuning has shown that the restriction map K0(p[H],p[H])torK0(p[C],p[C])torK_{0}\bigl{(}\mathbb{Z}_{p}[H],\mathbb{Q}_{p}[H]\bigr{)}_{\rm tor}\to K_{0}\bigl{(}\mathbb{Z}_{p}[C],\mathbb{Q}_{p}[C]\bigr{)}_{\rm tor} is injective for any subgroup HH of the form C,τ\langle C,\tau^{\prime}\rangle. The map α3\alpha_{3} therefore restricts to give an injective homomorphism

K0(p[G],p[G])torlimJK0(p[J],p[J])torK0(p[P],p[P])tor,K_{0}\bigl{(}\mathbb{Z}_{p}[G],\mathbb{Q}_{p}[G]\bigr{)}_{\rm tor}\to\varprojlim_{J}K_{0}\bigl{(}\mathbb{Z}_{p}[J],\mathbb{Q}_{p}[J]\bigr{)}_{\rm tor}\cong K_{0}\bigl{(}\mathbb{Z}_{p}[P],\mathbb{Q}_{p}[P]\bigr{)}_{\rm tor},

where JJ runs over all subgroups of PP, as required to prove claim (ii).

Claim (iii) follows from the equalities

=χIr(P)χπP|P|1χ(π)1π=πP(|P|1χIr(P)χ(π)1χ)π\mathcal{E}=\sum_{\chi\in\operatorname{Ir}(P)}\mathcal{E}_{\chi}\sum_{\pi\in P}|P|^{-1}\chi(\pi)^{-1}\pi=\sum_{\pi\in P}\biggl{(}|P|^{-1}\sum_{\chi\in\operatorname{Ir}(P)}\chi(\pi)^{-1}\mathcal{E}_{\chi}\biggr{)}\pi

and the fact that p[P]×=p[P]𝔐p×\mathbb{Z}_{p}[P]^{\times}=\mathbb{Z}_{p}[P]\cap\mathfrak{M}_{p}^{\prime\times}. ∎

6. Special cases

In this section we use the criteria of Theorem 5.8 to give both theoretical and numerical verifications of the pp-part of the equivariant Tamagawa number conjecture for pairs (h1(A/F)(1),[G])\bigl{(}h^{1}(A_{/F})(1),\mathbb{Z}[G]\bigr{)} where AA is an elliptic curve for which A(F)A(F) has strictly positive rank and GG is both non-abelian and of order divisible by pp.

We believe that, apart from the recent results of Bley in [5], where the group A(F)pA(F)_{p} is assumed to be trivial and the field FF to be an abelian extension of \mathbb{Q} of exponent pp, these results constitute the first verifications of the pp-part of the equivariant Tamagawa number conjecture for any elliptic curve and any Galois extension of degree divisible by pp.

We begin by giving the proof of Theorem 1.1, which relies on the theory of Heegner points and makes crucial use of the theorem of Gross and Zagier. We note that the additional hypotheses in Theorem 1.1 imply the validity of hypothesis (i). In addition, Kolyvagin [31, Proposition 2.1] shows that in this case one has Xp(AK)=0\hbox{\russ\char 88\relax}_{p}(A_{K})=0. In particular one knows that the pp-primary part of the Birch and Swinnerton-Dyer conjecture holds for A/KA_{/K}. Furthermore, the hypotheses (d) and (h) are obviously satisfied in the setting of Theorem 1.1.

6.1. The proof of Theorem 1.1

In view of Theorems 5.1 and 5.8 we are reduced to verifying the conditions (i) and (ii) that occur in the latter result.

To do this we fix a modular parametrisation φ:X0(N)A\varphi\colon X_{0}(N)\to A of smallest degree. We also denote by cc the Manin constant of φ\varphi and write QQ for the trace in A(F)A(F) of the Heegner point that is defined over the Hilbert class field of KK.

Set C:=4cc2|𝒪K×|2C:=4\,c_{\infty}\cdot c^{-2}\cdot\bigl{|}\mathcal{O}_{K}^{\times}\bigr{|}^{-2} and c=[A():A0()]c_{\infty}=\bigl{[}A(\mathbb{R}):A^{0}(\mathbb{R})\bigr{]} is the number of connected components of A()A(\mathbb{R}). Under our hypotheses, the theorem of Gross and Zagier (see, in particular, [32, §I, (6.5) and the discussion on p. 310]) implies that for each χIr(P)\chi\in\operatorname{Ir}(P) one has

(23) 0L(A/K,χ,1)|dK|Ω+(A)Ω(A)\displaystyle 0\neq\frac{L^{\prime}(A_{/K},\chi,1)\sqrt{|d_{K}|}}{\Omega^{+}_{\infty}(A)\Omega^{-}_{\infty}(A)} =C1|P|Tχ(Q),Tχˇ(Q)F\displaystyle=C\cdot\frac{1}{|P|}\bigl{\langle}T_{\chi}(Q),T_{\check{\chi}}(Q)\bigr{\rangle}_{F}
={ChF,IndPG(χ)(Q),if χ𝟏PChF,ρA(Q),if χ=𝟏P,\displaystyle=\begin{cases}C\cdot h_{F,\operatorname{Ind}^{G}_{P}(\chi)}(Q),\quad&\text{if $\chi\not=\boldsymbol{1}_{P}$}\\ C\cdot h_{F,\rho_{A}^{\prime}}(Q),&\text{if $\chi=\boldsymbol{1}_{P}$,}\end{cases}

where ρA\rho^{\prime}_{A} is the linear character of GG appearing in A(K)A(K) and \infty denotes the archimedean place of \mathbb{Q}. The second equality here follows from the equality TIndGPχ=Tχ+TχˇT_{\operatorname{Ind}_{G}^{P}\chi}=T_{\chi}+T_{\check{\chi}} and the fact that Tχ(Q),Tχ(Q)F=0\langle T_{\chi}(Q),T_{\chi}(Q)\rangle_{F}=0 (since the height pairing is GG-invariant).

For each ψ\psi in Ir(G)\operatorname{Ir}(G) we set

(24) 𝒬^ψ:=L(A,ψ,1)dψΩ(A,ψ)Hψ\hat{\mathcal{Q}}_{\psi}:=\frac{L^{\raisebox{2.0pt}{$\scriptstyle\star$}}(A,\psi,1)\cdot\sqrt{d_{\psi}}}{\Omega(A,\psi)\cdot H_{\psi}}

where the quantities dψd_{\psi} and HψH_{\psi} are as defined just before Theorem 5.8. We also write

(25) tψ=LSr(A,ψ,1)L(A,ψ,1)t_{\psi}=\frac{L^{\raisebox{2.0pt}{$\scriptstyle\star$}}_{S_{\rm r}}(A,\psi,1)}{L^{\raisebox{2.0pt}{$\scriptstyle\star$}}(A,\psi,1)}

for the correction term that accounts for the SrS_{\rm r}-truncation in the leading terms for each ψIr(G)\psi\in\operatorname{Ir}(G). Then the non-zero complex number 𝒬ψ:=uψtψ𝒬^ψ\mathcal{Q}_{\psi}:=u_{\psi}\cdot t_{\psi}\cdot\hat{\mathcal{Q}}_{\psi} is as in Theorem 5.8 using our Heegner point QQ. By using (23), and the fact that dψ=d𝟏dϵ=|dK|d_{\psi}=d_{\boldsymbol{1}}d_{\epsilon}=|d_{K}| for all ψ\psi of dimension two (as F/KF/K is unramified), one then finds that

(26) 𝒬𝟏𝒬ϵ\displaystyle\mathcal{Q}_{\boldsymbol{1}}\cdot\mathcal{Q}_{\epsilon} =u𝟏t𝟏uϵtϵC,\displaystyle=u_{\boldsymbol{1}}\,t_{\boldsymbol{1}}\,u_{\epsilon}\,t_{\epsilon}\cdot C,\qquad and
𝒬ψ\displaystyle\mathcal{Q}_{\psi} =uψtψC\displaystyle=u_{\psi}\,t_{\psi}\cdot C\qquad if ψ=IndPGχ\psi=\operatorname{Ind}^{G}_{P}\chi for χIr(P){𝟏P}\chi\in\operatorname{Ir}(P)\setminus\{\boldsymbol{1}_{P}\}.

Our hypotheses imply that A(K)A(K) has rank one and Xp(AK)\hbox{\russ\char 88\relax}_{p}(A_{K}) vanishes by Kolyvagin [31] and hence all of the hypotheses of Theorem 5.8 are satisfied except for the requirement that pp splits in K/K/\mathbb{Q}. However, the sole purpose of the latter hypothesis is to ensure (via the proof of Proposition 5.7) that the [P]\mathbb{Q}[P]-module A(F)\mathbb{Q}\otimes_{\mathbb{Z}}A(F) has a free rank one direct summand and so, since (23) implies that the point QQ generates such a summand, this hypothesis can be ignored in our case. Following Remark 5.9, it therefore suffices for us to prove that the terms 𝒬ψ\mathcal{Q}_{\psi} satisfy both the conditions of Theorem 5.8(i) and the congruences (19).

Since the pp-part of the Birch and Swinnerton-Dyer conjecture holds for A/KA_{/K}, the hypotheses (a) and (b) and the known vanishing of Xp(AK)\hbox{\russ\char 88\relax}_{p}(A_{K}) combine to imply that 𝒬^𝟏\hat{\mathcal{Q}}_{\boldsymbol{1}} and 𝒬^ϵ\hat{\mathcal{Q}}_{\epsilon} are both pp-units. The hypothesis (c) implies that AA has good reduction at pp and so the hypothesis (ii) combines with  [1, Theorem 2.7] to imply that the Manin constant cc is a pp-unit. Using also the fact that pp is unramified, we find that CC is a pp-unit too.

We proceed to compute the values of uψu_{\psi} and tψt_{\psi}. Since F/KF/K is unramified, we have that each ramified place vv in SrS_{\rm r} has ramification index 22 and vv does not split in K/K/\mathbb{Q}. It is easy to see that uϵ=1u_{\epsilon}=1 and that u𝟏=uψ=(1)|Sr|u_{\boldsymbol{1}}=u_{\psi}=(-1)^{|S_{\rm r}|} for all ψ\psi of dimension two. Next, for any ψ\psi in Ir(G)\operatorname{Ir}(G), we have

tψ=vSrNrdp[G](1Frw1|κv|1|(Tp(A)Vψ)Iw)t_{\psi}=\prod_{v\in S_{\rm r}}\operatorname{Nrd}_{\mathbb{C}_{p}[G]}\Bigl{(}1-\operatorname{Fr}_{w}^{-1}|\kappa_{v}|^{-1}\Bigl{|}\bigl{(}T_{p}(A)\otimes V_{\psi}\bigr{)}^{I_{w}}\Bigr{)}

and since no place of bad reduction is allowed to ramify by hypothesis (e), one has (Tp(A)Vψ)Iw=Tp(A)VψIw\bigl{(}T_{p}(A)\otimes V_{\psi}\bigr{)}^{I_{w}}=T_{p}(A)\otimes V_{\psi}^{I_{w}}. Hence tϵ=1t_{\epsilon}=1 and t𝟏=tψt_{\boldsymbol{1}}=t_{\psi} for all ψ\psi of dimension two. Moreover, this last value t𝟏t_{\boldsymbol{1}} is equal to vSr|A(κv)|/|κv|\prod_{v\in S_{\rm r}}|A(\kappa_{v})|/|\kappa_{v}|, which is a pp-unit by hypothesis (f). Therefore 𝒬ψ\mathcal{Q}_{\psi} is a pp-unit for all ψIr(G)\psi\in\operatorname{Ir}(G).

Next we note that 𝒬𝟏𝒬ϵ=𝒬ψ\mathcal{Q}_{\boldsymbol{1}}\cdot\mathcal{Q}_{\epsilon}=\mathcal{Q}_{\psi} for any ψ\psi of dimension two as we have shown that u𝟏t𝟏uϵtϵ=uψtψu_{\boldsymbol{1}}\,t_{\boldsymbol{1}}\,u_{\epsilon}\,t_{\epsilon}=u_{\psi}\,t_{\psi}. Since each element 𝒬ψ\mathcal{Q}_{\psi} also belongs to \mathbb{Q}, this formula makes it clear that they satisfy the condition of Theorem 5.8(i). In addition, it shows that for any πP\pi\in P one has

χIr(P){𝟏P}χˇ(π)𝒬IndPGχ=𝒬𝟏𝒬ϵ(1+χIr(P)χˇ(π)).-\sum_{\chi\in\operatorname{Ir}(P)\setminus\{\boldsymbol{1}_{P}\}}\check{\chi}(\pi)\mathcal{Q}_{\operatorname{Ind}^{G}_{P}\chi}=-\mathcal{Q}_{\boldsymbol{1}}\,\mathcal{Q}_{\epsilon}\cdot\Bigl{(}-1+\sum_{\chi\in\operatorname{Ir}(P)}\check{\chi}(\pi)\Bigr{)}.

Finally we note that the last sum in the above expression is equal to 0 if π1\pi\neq 1 and to |P||P| otherwise. Hence the expression is in both cases congruent to 𝒬𝟏𝒬ϵ\mathcal{Q}_{\boldsymbol{1}}\,\mathcal{Q}_{\epsilon} modulo p(p)p\mathbb{Z}_{(p)}, as required to complete the proof of Theorem 1.1.

6.2. S3S_{3}-extensions

In this section we investigate the case that F/KF/K is a cyclic extension of degree p=3p=3 and hence F/kF/k is a non-abelian extension of degree six. We show that the conjecture of Birch and Swinnerton-Dyer (or ‘BSD’ for short) implies an explicit congruence modulo rational squares and then combine this congruence with Theorem 5.8 to prove the equality (17) for a natural family of examples.

We assume throughout that GG is a non-abelian group of order six. In this case Ir(G)\operatorname{Ir}(G) comprises 𝟏:=𝟏G\boldsymbol{1}:=\boldsymbol{1}_{G}, ϵ\epsilon and ψ=IndPGχ\psi=\operatorname{Ind}_{P}^{G}\chi for a fixed χ\chi in Ir(P){𝟏P}\operatorname{Ir}(P)\setminus\{\boldsymbol{1}_{P}\}. We fix a subfield LL of FF that is of degree 33 over kk and set H:=GF/LH:=G_{F/L}.

6.2.1. An arithmetic congruence

For each character η\eta in Ir(G)\operatorname{Ir}(G) we define

(27) 𝒬~η:=L(A,η,1)dηΩ(A,η)H~η.\tilde{\mathcal{Q}}_{\eta}:=\frac{L^{\raisebox{2.0pt}{$\scriptstyle\star$}}(A,\eta,1)\cdot\sqrt{d_{\eta}}}{{\Omega(A,\eta)}\cdot\tilde{H}_{\eta}}.

Here the quantities dηd_{\eta} are as defined just before Theorem 5.8 and we have set

(28) H~𝟏:=Reg(A/k),H~ϵ:=Reg(A/K)Reg(A/k),andH~ψ:=Reg(A/L)Reg(A/k)\tilde{H}_{\boldsymbol{1}}:=\operatorname{Reg}(A_{/k}),\qquad\tilde{H}_{\epsilon}:=\frac{\operatorname{Reg}(A_{/K})}{\operatorname{Reg}(A_{/k})},\qquad\text{and}\qquad\tilde{H}_{\psi}:=\frac{\operatorname{Reg}(A_{/L})}{\operatorname{Reg}(A_{/k})}

with Reg(A/E)\operatorname{Reg}(A_{/E}) denoting the regulator of ,E\langle\cdot,\cdot\rangle_{E} on the Mordell-Weil group A(E)A(E) for any field EE. These slight variations of the regulators that we used earlier fit well with BSD and do not require any knowledge of the explicit Galois structure of A(F)pA(F)_{p}. In fact, BSD predicts that each expression 𝒬~η\tilde{\mathcal{Q}}_{\eta} is a non-zero rational number.

In the following result we include the assumed analytic continuation of the LL-series and finiteness of the Tate-Shafarevich group in the assumption that BSD holds but do not assume any of the hypotheses (a)(d) and (f)(h). This result may therefore suggest one sort of congruence relation that might be expected to hold when our hypotheses fail. We note also that the remark made by Dokchitser and Dokchitser in the fourth paragraph after Conjecture 1.4 of [26] hints at the possibility of this sort of result in a more restrictive setting.

Theorem 6.1.

Let AA be an elliptic curve over kk and assume no place at which AA has bad reduction ramifies in F/kF/k. Then if the Birch and Swinnerton-Dyer conjecture holds for A/kA_{/k}, A/KA_{/K} and A/LA_{/L} one has a congruence modulo non-zero rational squares

(29) 𝒬~𝟏𝒬~ϵ𝒬~ψ(mod).\tilde{\mathcal{Q}}_{\boldsymbol{1}}\cdot\tilde{\mathcal{Q}}_{\epsilon}\equiv\tilde{\mathcal{Q}}_{\psi}\pmod{\square}.
Proof.

Fix a finite field extension EE of kk. If vv is a (finite or infinite) place of EE, write c(A/Ev)c(A/E_{v}) for the number of connected components [A(Ev):A(Ev)0]\bigl{[}A(E_{v}):A(E_{v})^{0}\bigr{]} of A(Ev)A(E_{v}). For a finite place vv in EE, also write ωv\omega_{v}^{\text{N\'{e}}} for a Néron differential of A/EvA_{/E_{v}}. Then BSD for the field EE asserts that

(30) L(A/E,1)|dE|Ω(A/E)Reg(A/E)=all vc(A/Ev)finite v|ωAωv|v|X(AE)||A(E)tor|2\frac{L^{\raisebox{2.0pt}{$\scriptstyle\star$}}(A_{/E},1)\,\sqrt{|d_{E}|}}{\Omega(A_{/E})\cdot\operatorname{Reg}(A_{/E})}=\prod_{\text{all }v}c(A/{E_{v}})\cdot\prod_{\text{finite }v}\biggl{|}\frac{\omega_{A}}{\omega_{v}^{\text{N\'{e}}}}\biggr{|}_{v}\cdot\frac{|\hbox{\russ\char 88\relax}(A_{E})|}{|A(E)_{\rm tor}|^{2}}

where we write Ω(A/E)\Omega(A_{/E}) for the period vSEΩv+(A)vSEΩv(A)\prod_{v\in S_{\mathbb{R}}^{E}}\Omega_{v}^{+}(A)\cdot\prod_{v\in S_{\mathbb{C}}^{E}}\Omega_{v}(A) of AA over EE, as defined in §4.4.1 with respect to a fixed invariant differential ωA\omega_{A} of A/kA_{/k}.

The term 𝒬~𝟏\tilde{\mathcal{Q}}_{\boldsymbol{1}} is equal to the left hand side of (30) with E=kE=k. Furthermore the products 𝒬~𝟏𝒬~ϵ\tilde{\mathcal{Q}}_{\boldsymbol{1}}\tilde{\mathcal{Q}}_{\epsilon} and 𝒬~𝟏𝒬~ψ\tilde{\mathcal{Q}}_{\boldsymbol{1}}\tilde{\mathcal{Q}}_{\psi} link to the left hand sides of (30) for the fields KK and LL respectively. More precisely, one has

𝒬~𝟏𝒬~ϵ=L(A/K,1)|dK|Ω(A/K)Reg(A/K)vSrc(A/kv)\tilde{\mathcal{Q}}_{\boldsymbol{1}}\cdot\tilde{\mathcal{Q}}_{\epsilon}=\frac{L^{\raisebox{2.0pt}{$\scriptstyle\star$}}(A_{/K},1)\,\sqrt{|d_{K}|}}{\Omega(A_{/K})\cdot\operatorname{Reg}(A_{/K})}\cdot\prod_{v\in S_{\mathbb{R}}^{\rm r}}c(A/k_{v})

where, as before, SrS_{\mathbb{R}}^{\rm r} denotes the set of real places of kk that become complex places in KK. The formula for 𝒬~𝟏𝒬~ψ\tilde{\mathcal{Q}}_{\boldsymbol{1}}\tilde{\mathcal{Q}}_{\psi} is modified by the same factor as there is exactly one complex place in LL above each place in SrS_{\mathbb{R}}^{\rm r}. In proving this last formula one also uses the fact that dψ=|dK|Nf(χ)d_{\psi}=|d_{K}|\cdot Nf(\chi) is equal to |dL/dk||d_{L}/d_{k}| since ψ=IndHG𝟏H𝟏\psi=\operatorname{Ind}_{H}^{G}\boldsymbol{1}_{H}-\boldsymbol{1}.

In addition, since we are working modulo squares, we may neglect the terms |X(AE)||\hbox{\russ\char 88\relax}(A_{E})| and |A(E)tor|2|A(E)_{\rm tor}|^{2} which occur on the right hand side of the formula (30). The required congruence (29) will therefore be proved if we can show for each place vv in kk that

(31) c(A/kv)wvin Kc(A/Kw)wvin Lc(A/Lw)(mod)c(A/k_{v})\cdot\prod_{\begin{subarray}{c}w\mid v\\ \text{in }K\end{subarray}}c(A/K_{w})\equiv\prod_{\begin{subarray}{c}w\mid v\\ \text{in }L\end{subarray}}c(A/L_{w})\pmod{\square}

and for each finite place vv in kk that

(32) |ωAωv|vwvin K|ωAωw|w=wvin L|ωAωw|w.\biggl{|}\frac{\omega_{A}}{\omega_{v}^{\text{N\'{e}}}}\biggr{|}_{v}\cdot\prod_{\begin{subarray}{c}w\mid v\\ \text{in }K\end{subarray}}\biggl{|}\frac{\omega_{A}}{\omega_{w}^{\text{N\'{e}}}}\biggr{|}_{w}=\prod_{\begin{subarray}{c}w\mid v\\ \text{in }L\end{subarray}}\biggl{|}\frac{\omega_{A}}{\omega_{w}^{\text{N\'{e}}}}\biggr{|}_{w}.

Now, by our assumption, no place at which AA has bad reduction is ramified in F/kF/k and so the Néron differential for AA over kvk_{v} remains a Néron differential for AA over both of the fields KwK_{w} and LwL_{w}. Hence the equation (32) is valid for all finite places, because ωA/ωv\omega_{A}/\omega^{\text{N\'{e}}}_{v} is in kvk_{v} and one has [k:k]+[K:k]=[L:k][k:k]+[K:k]=[L:k].

Next we note that the congruence (31) only needs to be checked at places at which AA has bad reduction (and which therefore do not ramify in F/kF/k) and at infinite places. If the decomposition group GvG_{v} at a place above vv in FF is trivial, then both sides of this congruence are equal to c(A/kv)3c(A/k_{v})^{3}. If GvG_{v} is cyclic of order 22, then there is one place ww^{\prime} in LL with Lw=kvL_{w^{\prime}}=k_{v} and one place w′′w^{\prime\prime} with Lw′′=KwL_{w^{\prime\prime}}=K_{w} for the unique place ww above vv in KK and so both sides of (31) are equal to c(A/kv)c(A/Kw)c(A/k_{v})\cdot c(A/K_{w}). Finally we have to treat the case when GvG_{v} is cyclic of order 33 and hence the place vv is finite. In this case the left hand side of (31) is equal to c(A/kv)3c(A/k_{v})^{3} while the right hand side is equal to c(A/Lw)c(A/L_{w}) where LwL_{w} is an unramified cubic extension of kvk_{v}. If c(A/Lw)=c(A/kv)c(A/L_{w})=c(A/k_{v}) then we have indeed a congruence modulo squares. However it is possible that the Tamagawa numbers change in an unramified extension. Luckily, the only possibility for this to happen in a cubic extension is when the Kodaira type is I0{}_{0}^{*} and then the change is from c(A/kv)=1c(A/k_{v})=1 to c(A/Lw)=4c(A/L_{w})=4, see for instance Step 6 in [47] on page 367, and the congruence (31) holds in all cases. ∎

6.2.2. The connection to eTNCp

We now use Theorem 6.1 to show that, under the hypotheses of Theorem 5.8, the relevant cases of BSD imply the equality (17).

Corollary 6.2.

We assume that the elliptic curve AA and field FF satisfy the hypotheses (a)(h). We assume also that GG and LL are as in Theorem 6.1 (so p=3p=3) and that the Birch and Swinnerton-Dyer conjecture holds for AA over each of the fields kk, KK and LL.

Then the equality (17) is valid provided that Xp(AK)=0\hbox{\russ\char 88\relax}_{p}(A_{K})=0, rk(AK)=1\operatorname{rk}(A_{K})=1 and there exists a point QQ in A(F)A(F) which generates a GG-module of finite prime-to-pp index in A(F)A(F) that is isomorphic to [G](1(1)rk(Ak)τ)\mathbb{Z}[G]\bigl{(}1-(-1)^{\operatorname{rk}(A_{k})}\tau\bigr{)}.

Proof.

First, we note that given our hypotheses (a), (b) and (e), our choice of ωA\omega_{A} (as in §4.4.1) and our assumption that Xp(AK)\hbox{\russ\char 88\relax}_{p}(A_{K}) vanishes, the validity of BSD implies that the term 𝒬~η\tilde{\mathcal{Q}}_{\eta} is a pp-adic unit for all ηIr(G)\eta\in\operatorname{Ir}(G).

Next we link 𝒬~η\tilde{\mathcal{Q}}_{\eta} to 𝒬η\mathcal{Q}_{\eta} in Theorem 5.8, the difference being the terms H~η\tilde{H}_{\eta} versus HηH_{\eta} and the terms uηu_{\eta} and tηt_{\eta} (as defined in (25)). Using the explicit structure of A(F)pA(F)_{p}, it is easy to show that

HψH~ψ,Hϵ{H~ϵ if ρA=𝟏12H~ϵ if ρA=ϵandH𝟏{H~𝟏 if ρA=𝟏2H~𝟏 if ρA=ϵ.H_{\psi}\equiv\tilde{H}_{\psi},\qquad H_{\epsilon}\equiv\begin{cases}\tilde{H}_{\epsilon}&\text{ if $\rho_{A}=\boldsymbol{1}$}\\ \frac{1}{2}\,\tilde{H}_{\epsilon}&\text{ if $\rho_{A}=\epsilon$}\end{cases}\qquad\text{and}\qquad H_{\boldsymbol{1}}\equiv\begin{cases}\tilde{H}_{\boldsymbol{1}}&\text{ if $\rho_{A}=\boldsymbol{1}$}\\ 2\,\tilde{H}_{\boldsymbol{1}}&\text{ if $\rho_{A}=\epsilon$.}\end{cases}

where all congruences are modulo squares in (3)×\mathbb{Z}_{(3)}^{\times}; namely the quotients are squares of indices, like the index of [G]Q\mathbb{Z}[G]Q in A(F)A(F). Up to squares in (3)×\mathbb{Z}_{(3)}^{\times} we therefore have a congruence

1u𝟏t𝟏uϵtϵ𝒬𝟏𝒬ϵ1uψtψ𝒬ψ.\frac{1}{u_{\boldsymbol{1}}\,t_{\boldsymbol{1}}\,u_{\epsilon}\,t_{\epsilon}}\mathcal{Q}_{\boldsymbol{1}}\,\mathcal{Q}_{\epsilon}\equiv\frac{1}{u_{\psi}\,t_{\psi}}\mathcal{Q}_{\psi}.

An argument similar to the one that concludes the proof of Theorem 1.1 hence implies that we will have verified the criterion in Theorem 5.8 if we show that u𝟏t𝟏uϵtϵu_{\boldsymbol{1}}\,t_{\boldsymbol{1}}\,u_{\epsilon}\,t_{\epsilon} and uψtψu_{\psi}\,t_{\psi} are 33-adic units that are congruent modulo 33. Writing Nv=|A(κv)|N_{v}=|A(\kappa_{v})| for the number of points in the reduction at a place vv and qv=|κv|q_{v}=|\kappa_{v}|, we can summarise the computations of the local contribution to these terms at a place vSrv\in S_{\rm r} in the following table according to the type of ramification. Here eve_{v} stands for the ramification index at a place in FF above vv and fvf_{v} for the residual degree.

det(Frw|VηIw)\det\bigl{(}-Fr_{w}\bigl{|}V_{\eta}^{I_{w}}\bigr{)} det(1Frw1qv1|(VηTp(A))Iw)\det\bigl{(}1-Fr_{w}^{-1}\,q_{v}^{-1}\bigl{|}(V_{\eta}\otimes T_{p}(A))^{I_{w}}\bigr{)}
eve_{v} fvf_{v} u𝟏u_{\boldsymbol{1}} uϵu_{\epsilon} uψu_{\psi} t𝟏t_{\boldsymbol{1}} tϵt_{\epsilon} tψt_{\psi}
2 1 1-1 1 1-1 Nv/qvN_{v}/q_{v} 1 Nv/qvN_{v}/q_{v}
3 1 1-1 1-1 1 Nv/qvN_{v}/q_{v} Nv/qvN_{v}/q_{v} 1
3 2 1-1 1 1 Nv/qvN_{v}/q_{v} Nw/qwqv/NvN_{w}/q_{w}\cdot q_{v}/N_{v} 1

In the last line ww denotes the unique place ww in KK above vv. If a place vv were totally ramified it must be above 22 or 33 as the tame inertia group is cyclic and it can not be above 22 because the wild inertia group is normal in the inertia group. Hence there was no need to list the totally ramified case as all places above p=3p=3 were assumed to be unramified by (h). From the table we can conclude that all the terms uηu_{\eta} and tηt_{\eta} are indeed 33-units by (f) and that

u𝟏uϵuψt𝟏tϵtψvSr′′(Nwqw)vSr′′(Nw)(mod)\frac{u_{\boldsymbol{1}}\,u_{\epsilon}}{u_{\psi}}\cdot\frac{t_{\boldsymbol{1}}\,t_{\epsilon}}{t_{\psi}}\equiv\prod_{v\in S^{\prime\prime}_{\rm r}}\biggl{(}-\frac{N_{w}}{q_{w}}\biggr{)}\equiv\prod_{v\in S^{\prime\prime}_{\rm r}}\bigl{(}-N_{w}\bigr{)}\pmod{\square}

with Sr′′:={v|ev=3 and fv=2}S^{\prime\prime}_{\rm r}:=\{v\,|\,e_{v}=3\text{ and }f_{v}=2\}.

Hence we are reduced to showing that Nw1(mod3)-N_{w}\equiv 1\pmod{3} when ev=3e_{v}=3 and fv=2f_{v}=2. We first note that for such a place we must have qv1(mod3)q_{v}\equiv-1\pmod{3}. Indeed, this is true because the map θ0:Gal(Fw/Kw)μ3(κw)\theta_{0}:\operatorname{Gal}\bigr{(}F_{w^{\prime}}/K_{w}\bigr{)}\to\mu_{3}(\kappa_{w}) in Corollaire IV.1 in [45] is GG-equivariant and, since Fw/kvF_{w^{\prime}}/k_{v} is dihedral, the action of τ\tau on μ3(κw)\mu_{3}(\kappa_{w}) is non-trivial.

Finally, we have the equality Nw=Nv(2qv+2Nv)N_{w}=N_{v}\cdot(2q_{v}+2-N_{v}) valid for all quadratic extension of finite fields. Hence NwNv21(mod3)N_{w}\equiv-N_{v}^{2}\equiv-1\pmod{3}. ∎

Remark 6.3.

A closer analysis of the above argument shows that the hypotheses of Corollary 6.2 may be weakened a little. One can allow places above 33 to be tamely ramified in F/kF/k and can omit any assumption about the reduction of AA at such places. In addition, one need only assume that the group A(κv)[p]A(\kappa_{v})[p] vanishes for places vv that are both inert in K/kK/k and ramify in F/KF/K.

Remark 6.4.

For any prime p>3p>3, the methods used in the proofs of Theorem 6.1 and Corollary 6.2 enable one to deduce from the assumed validity of suitable cases of the Birch-Swinnerton-Dyer Conjecture a congruence for the product dim(ψ)=2𝒬^ψ\prod_{\dim(\psi)=2}\hat{\mathcal{Q}}_{\psi}, rather than for the sum dim(ψ)=2𝒬^ψ\sum_{\dim(\psi)=2}\hat{\mathcal{Q}}_{\psi} that occurs in (17).

6.3. Numerical examples

In this final section we describe two numerical examples to further illustrate the predicted congruences in Theorem 5.8 and to explain how one can check these congruences for numerous examples. It is comparatively straightforward to give examples with p=3p=3 but Corollary 6.2 implies that there is limited interest in doing so. We therefore discuss examples with p=7p=7 and p=5p=5.

Our numerical computations were done using Sage [48], which uses underlying Pari-GP [49]. The computations of the LL-values was done in Magma [9] which contains an implementation of [25]. The code can be obtained from the last named author’s webpage.

6.3.1. A Stark-Heegner point example

We consider the example of the elliptic curve labelled 37a1 in Cremona’s tables [20]

A:y2+y=x3xA\colon\qquad y^{2}\ +\ y\ =\ x^{3}\ -\ x

over the Hilbert class field FF of K=(577)K=\mathbb{Q}\bigl{(}\sqrt{577}\bigr{)}. The curve has rank 11 over \mathbb{Q} and KK. The extension F/KF/K is of degree p=7p=7 defined by a root ξ\xi of the polynomial

x72x67x5+10x4+13x310x2x+1.x^{7}-2\,x^{6}-7\,x^{5}+10\,x^{4}+13\,x^{3}-10\,x^{2}-x+1.

All hypotheses (a)(h) except the finiteness of X(AF)\hbox{\russ\char 88\relax}(A_{F}) in (g) can be verified easily. We find in [23] that there is a point

Q=(2ξ64ξ514ξ4+17ξ3+30ξ27ξ3, 2ξ619ξ42ξ3+32ξ2+ξ5)Q=\Bigl{(}2\,\xi^{6}-4\,\xi^{5}-14\,\xi^{4}+17\,\xi^{3}+30\,\xi^{2}-7\,\xi-3,\ 2\,\xi^{6}-19\,\xi^{4}-2\,\xi^{3}+32\,\xi^{2}+\xi-5\Bigr{)}

of infinite order on AA defined over L=(ξ)L=\mathbb{Q}(\xi) obtained from Darmon’s construction of modular points. The trace of QQ in L/L/\mathbb{Q} is equal to the generator R=(0,0)A()R=(0,0)\in A(\mathbb{Q}). Let σ\sigma be a generator of PP. It is easy to check that

RQ(σ+σ1)Q(σ2+σ2)Q\mathbb{Z}\,R\oplus\mathbb{Z}\,Q\oplus\mathbb{Z}\,(\sigma+\sigma^{-1})Q\oplus\mathbb{Z}\,(\sigma^{2}+\sigma^{-2})Q

has finite index coprime to pp in A(L)A(L). Hence we can take QQ as the point whose existence is predicted by Proposition 5.7. (In the general case, we may have to take a linear combination of a new point in A(L)A(L) and the generator in A()A(\mathbb{Q}) to assure that the trace generates A()A(\mathbb{Q}).) In fact, in our case, we can check that A(F)=[G]QA(F)=\mathbb{Z}[G]Q by using the bound given in [21].

Using modular symbols for the character ϵ\epsilon and a Heegner point computation for 𝟏\boldsymbol{1}, we can prove that the formulae (24) evaluate to

𝒬^𝟏=L(A,1)Ω+(A)H𝟏=1 and 𝒬^ϵ=L(A,ϵ,1)dKΩ+(A)=4\hat{\mathcal{Q}}_{\boldsymbol{1}}=\frac{L^{\prime}(A,1)}{\Omega^{+}(A)\cdot H_{\boldsymbol{1}}}=1\qquad\text{ and }\qquad\hat{\mathcal{Q}}_{\epsilon}=\frac{L(A,\epsilon,1)\cdot\sqrt{d_{K}}}{\Omega^{+}(A)}=4

and hence conclude that BSD holds for A/A_{/\mathbb{Q}} and A/KA_{/K} with X(A)=X(AK)=0\hbox{\russ\char 88\relax}(A_{\mathbb{Q}})=\hbox{\russ\char 88\relax}(A_{K})=0. Next, we compute a numerical approximation to L(A,ψ,1)L^{\prime}(A,\psi,1) for a 22-dimensional representation ψIr(G)\psi\in\operatorname{Ir}(G). The corresponding value of 𝒬^ψ\hat{\mathcal{Q}}_{\psi} is equal to 4.00000000000004.0000000000000 for all such ψ\psi, but we know of no means of proving that this value is indeed algebraic and equal to the value 44. It predicts with good accuracy that BSD for A/LA_{/L} would imply that X(AL)\hbox{\russ\char 88\relax}(A_{L}) and X(AF)\hbox{\russ\char 88\relax}(A_{F}) are trivial. However assuming that 𝒬^ψ=4\hat{\mathcal{Q}}_{\psi}=4, we compute the SrS_{\rm r}-truncated version

𝒬𝟏=578577 and 𝒬ϵ=4 and 𝒬ψ=2312577 for all dim(ψ)=2\mathcal{Q}_{\boldsymbol{1}}=-\frac{578}{577}\ \text{ and }\ \mathcal{Q}_{\epsilon}=4\ \text{ and }\ \mathcal{Q}_{\psi}=-\frac{2312}{577}\ \text{ for all $\dim(\psi)=2$}

and hence find that the congruence (20) holds modulo p=7p=7. Note that in this particular case, the values 𝒬^ψ\hat{\mathcal{Q}}_{\psi} were all in \mathbb{Q}. In other words we have found convincing numerical evidence that a Gross-Zagier formula

L(A/K,χ,1)dKΩ+(A)2=c2c21|P|Tχ(Q),Tχˇ(Q)F\frac{L^{\prime}(A_{/K},\chi,1)\sqrt{d_{K}}}{\Omega^{+}(A)^{2}}=\frac{c_{\infty}^{2}}{c^{2}}\cdot\frac{1}{|P|}\bigl{\langle}T_{\chi}(Q),T_{\check{\chi}}(Q)\bigr{\rangle}_{F}

analogous to (23) should hold for all χIr(P)\chi\in\operatorname{Ir}(P) because c=2c_{\infty}=2 and c=1c=1.

6.3.2. A quintic example

As a second example, we consider the curve

A:y2+xy=x3 4x 1A\colon\qquad y^{2}\ +\ x\,y\ =\ x^{3}\ -\ 4\,x\ -\ 1

labelled 21a1 in [20]. It has rank 0 over \mathbb{Q}, but rank 11 over K=(i)K=\mathbb{Q}(i) and the group A(K)A(K) is generated by the point

R=(32,3+7i4).R=\bigl{(}\tfrac{3}{2},\ \tfrac{-3+7\,i}{4}\bigr{)}.

Now we consider the extension F/KF/K given by a solution ξ\xi of the polynomial

x52x46x3+10x2+17x12x^{5}-2\,x^{4}-6\,x^{3}+10\,x^{2}+17\,x-12

The extension F/KF/K is only ramified at the place 19[i]19\,\mathbb{Z}[i]. All of our hypotheses except (g) can be verified to hold in this example.

By a simple search for points, we find the point

T=(18(ξ4ξ212ξ8),132(39ξ47ξ3213ξ2127ξ+196))T=\Bigl{(}\tfrac{1}{8}\bigl{(}-\xi^{4}-\xi^{2}-12\,\xi-8\bigr{)},\ \tfrac{1}{32}\big{(}39\xi^{4}-7\,\xi^{3}-213\xi^{2}-127\,\xi+196\bigr{)}\Bigr{)}

of infinite order defined over L=(ξ)L=\mathbb{Q}(\xi). The bounds in [21] can then be used to prove that T(σ+σ1)T\mathbb{Z}\,T\oplus\mathbb{Z}\,(\sigma+\sigma^{-1})T generates A(L)tfA(L)_{\rm tf} where σ\sigma is a generator of PP. To find the point QQ is a bit more elaborate than in the previous example as ρA=𝟏\rho_{A}=\boldsymbol{1}. In fact the [G]\mathbb{Z}[G]-module generated by TT and RR will have index pp in A(F)A(F) because [18, Corollary 2.5] tells us that A(F)pA(F)_{p} is projective. We are going to use the relation

pσ(p+1)/2=TrF/K+(σ1)i=1(p1)/2i(σiσi)p\cdot\sigma^{(p+1)/2}=\operatorname{Tr}_{F/K}\ +\ (\sigma-1)\sum_{i=1}^{(p-1)/2}i\,\bigl{(}\sigma^{i}-\sigma^{-i}\bigr{)}

in [G]\mathbb{Z}[G] which is reminiscent of Kolyvagin’s derivative construction. We now try to find a point T=aT+b(σ+σ1)TT^{\prime}=a\,T+b\,(\sigma+\sigma^{-1})T in A(L)A(L) with 0a,b<p0\leq a,b<p such that OQ=R+(σ1)(T)O\neq Q^{\prime}=R+(\sigma-1)(T^{\prime}) is divisible by pp in A(F)A(F). Then we can take QQ such that pσ(p+1)/2(Q)=Qp\sigma^{(p+1)/2}(Q)=Q^{\prime} to be the point predicted by Proposition 5.7. In our concrete case this works with (a,b)=(4,3)(a,b)=(4,3). It can be shown that A(F)=/2/4[G]QA(F)={}^{\mathbb{Z}}\!/\!{}_{2\mathbb{Z}}\oplus{}^{\mathbb{Z}}\!/\!{}_{4\mathbb{Z}}\oplus\mathbb{Z}[G]Q as a [G]\mathbb{Z}[G]-module.

Using modular symbols and Heegner points, we can provably compute that

𝒬^𝟏=14 and 𝒬^ϵ=12\hat{\mathcal{Q}}_{\boldsymbol{1}}=\frac{1}{4}\qquad\text{ and }\qquad\hat{\mathcal{Q}}_{\epsilon}=\frac{1}{2}

and hence deduce that BSD holds for A/A_{/\mathbb{Q}} and A/KA_{/K} with trivial Tate-Shafarevich groups in both cases. We compute to a high precision the derivatives of L(A,ψ,s)L(A,\psi,s) for the representations ψ=ψ1\psi=\psi_{1} and ψ2\psi_{2} of dimension two and we find, with an error less than 102810^{-28}, that 𝒬^ψ1𝒬^ψ2256\hat{\mathcal{Q}}_{\psi_{1}}\cdot\hat{\mathcal{Q}}_{\psi_{2}}\approx 256, predicting that |X(AL)|=4|\hbox{\russ\char 88\relax}(A_{L})|=4 and 𝒬^ψ1+𝒬^ψ248\hat{\mathcal{Q}}_{\psi_{1}}+\hat{\mathcal{Q}}_{\psi_{2}}\approx 48. It also predicts that X(AF)\hbox{\russ\char 88\relax}(A_{F}) has order 3232. We will now assume that these are actually equalities and conclude that 𝒬^ψi=8(3±5)\hat{\mathcal{Q}}_{\psi_{i}}=8\cdot\bigl{(}3\pm\sqrt{5}\bigr{)}. One then computes the SrS_{\rm r}-truncated values to be

𝒬𝟏=819,𝒬ϵ=2419 and 𝒬ψ1+𝒬ψ2=96\mathcal{Q}_{\boldsymbol{1}}=\frac{8}{19},\qquad\mathcal{Q}_{\epsilon}=\frac{24}{19}\qquad\text{ and }\qquad\mathcal{Q}_{\psi_{1}}+\mathcal{Q}_{\psi_{2}}=-96

and this shows that the congruence (20) holds modulo p=5p=5.

References

  • [1] A. Agashe, K. Ribet, and W. A. Stein, The Manin constant, Pure Appl. Math. Q. 2 (2006), no. 2, part 2, 617–636.
  • [2] M. F. Atiyah and C. T. C. Wall, Cohomology of groups, Algebraic Number Theory (Proc. Instructional Conf., Brighton, 1965), Thompson, Washington, D.C., 1967, pp. 94–115.
  • [3] W. Bley, Numerical evidence for the equivariant Birch and Swinnerton-Dyer conjecture, Exp. Math. 20 (2011), no. 4, 426–456.
  • [4] by same author, Numerical evidence for the equivariant Birch and Swinnerton-Dyer conjecture (Part II), Math. Comp. 81 (2012), 1681–1705.
  • [5] by same author, The equivariant Tamagawa number conjecture and modular symbols, to appear in Math. Ann., 2013.
  • [6] W. Bley and D. Burns, Equivariant epsilon constants, discriminants and étale cohomology, Proc. London Math. Soc. (3) 87 (2003), no. 3, 545–590.
  • [7] W. Bley and D. Macias Castillo, Congruences for critical values of higher derivatives of twisted Hasse-Weil LL-functions, to appear in J. reine u. angew. Math., 2015.
  • [8] S. Bloch and K. Kato, LL-functions and Tamagawa numbers of motives, The Grothendieck Festschrift, Vol. I, Progr. Math., vol. 86, Birkhäuser Boston, Boston, MA, 1990, pp. 333–400.
  • [9] W. Bosma, J. Cannon, and C. Playoust, The Magma algebra system. I. The user language, J. Symbolic Comput. 24 (1997), no. 3-4, 235–265, Computational algebra and number theory (London, 1993).
  • [10] M. Breuning, On equivariant global epsilon constants for certain dihedral extensions, Math. Comp. 73 (2004), no. 246, 881–898.
  • [11] M. Breuning and D. Burns, Additivity of Euler characteristics in relative algebraic KK-groups, Homology, Homotopy Appl. 7 (2005), no. 3, 11–36.
  • [12] K. S. Brown, Cohomology of groups, Graduate Texts in Mathematics, vol. 87, Springer-Verlag, New York, 1982.
  • [13] D. Burns, Equivariant Whitehead torsion and refined Euler characteristics, Number theory, CRM Proc. Lecture Notes, vol. 36, Amer. Math. Soc., Providence, RI, 2004, pp. 35–59.
  • [14] by same author, On leading terms and values of equivariant motivic LL-functions, Pure Appl. Math. Q. 6 (2010), no. 1, Special Issue: In honor of John Tate. Part 2, 83–172.
  • [15] D. Burns and M. Flach, Motivic LL-functions and Galois module structures, Math. Ann. 305 (1996), no. 1, 65–102.
  • [16] by same author, Tamagawa numbers for motives with (non-commutative) coefficients, Doc. Math. 6 (2001), 501–570.
  • [17] D. Burns and D. Macias Castillo, Organising matrices for arithmetic complexes, to appear in Int. Math. Res. Not., 2013.
  • [18] D. Burns, D. Macias Castillo, and C. Wuthrich, On the Galois structure of Selmer groups, to appear in Int. Math. Res. Not., 2015.
  • [19] D. Burns and O. Venjakob, On descent theory and main conjectures in non-commutative Iwasawa theory, J. Inst. Math. Jussieu 10 (2011), 59–118.
  • [20] J. E. Cremona, Algorithms for modular elliptic curves, second ed., Cambridge University Press, Cambridge, 1997.
  • [21] J. E. Cremona, M. Prickett, and S. Siksek, Height difference bounds for elliptic curves over number fields, J. Number Theory 116 (2006), no. 1, 42–68.
  • [22] C. W. Curtis and I. Reiner, Methods of representation theory. Vol. I, Wiley Classics Library, John Wiley & Sons Inc., New York, 1990, With applications to finite groups and orders, Reprint of the 1981 original, A Wiley-Interscience Publication.
  • [23] H. Darmon and R. Pollack, Efficient calculation of Stark-Heegner points via overconvergent modular symbols, Israel J. Math. 153 (2006), 319–354.
  • [24] P. Deligne, Le déterminant de la cohomologie, Current trends in arithmetical algebraic geometry (Arcata, Calif., 1985), Contemp. Math., vol. 67, Amer. Math. Soc., Providence, RI, 1987, pp. 93–177.
  • [25] T. Dokchitser, Computing special values of motivic LL-functions, Experiment. Math. 13 (2004), no. 2, 137–149.
  • [26] T. Dokchitser and V. Dokchitser, Computations in non-commutative Iwasawa theory, Proc. Lond. Math. Soc. (3) 94 (2007), no. 1, 211–272, With an appendix by J. Coates and R. Sujatha.
  • [27] A. W. M. Dress, Induction and structure theorems for orthogonal representations of finite groups, Ann. of Math. (2) 102 (1975), no. 2, 291–325.
  • [28] J-M. Fontaine and W. Messing, pp-adic periods and pp-adic étale cohomology, Current trends in arithmetical algebraic geometry (Arcata, Calif., 1985), Contemp. Math., vol. 67, Amer. Math. Soc., Providence, RI, 1987, pp. 179–207.
  • [29] A. Fröhlich and J. Queyrut, On the functional equation of the Artin LL-function for characters of real representations, Invent. Math. 20 (1973), 125–138.
  • [30] B. H. Gross, On the conjecture of Birch and Swinnerton-Dyer for elliptic curves with complex multiplication, Number theory related to Fermat’s last theorem (Cambridge, Mass., 1981), Progr. Math., vol. 26, Birkhäuser Boston, Mass., 1982, pp. 219–236.
  • [31] by same author, Kolyvagin’s work on modular elliptic curves, LL-functions and arithmetic (Durham, 1989), London Math. Soc. Lecture Note Ser., vol. 153, Cambridge Univ. Press, Cambridge, 1991, pp. 235–256.
  • [32] B. H. Gross and D. B. Zagier, Heegner points and derivatives of LL-series, Invent. Math. 84 (1986), no. 2, 225–320.
  • [33] H. Jacobinski, On extensions of lattices, Michigan Math. J. 13 (1966), 471–475.
  • [34] H. Johnston and A. Nickel, Noncommutative Fitting invariants and improved annihilation results, to appear in J. London Math. Soc., 2013.
  • [35] G. Kings, The equivariant Tamagawa number conjecture and the Birch-Swinnerton-Dyer conjecture, Arithmetic of LL-functions, IAS/Park City Math. Ser., vol. 18, Amer. Math. Soc., Providence, RI, 2011, pp. 315–349.
  • [36] S. R. Louboutin, On the divisibility of the class number of imaginary quadratic number fields, Proc. Amer. Math. Soc. 137 (2009), no. 12, 4025–4028.
  • [37] J. Martinet, Character theory and Artin LL-functions, Algebraic number fields: LL-functions and Galois properties (Proc. Sympos., Univ. Durham, Durham, 1975), Academic Press, London, 1977, pp. 1–87.
  • [38] B. Mazur, Rational points of abelian varieties with values in towers of number fields, Invent. Math. 18 (1972), 183–266.
  • [39] B. Mazur and K. Rubin, Organizing the arithmetic of elliptic curves, Adv. Math. 198 (2005), no. 2, 504–546.
  • [40] by same author, Finding large Selmer rank via an arithmetic theory of local constants, Ann. of Math. (2) 166 (2007), no. 2, 579–612.
  • [41] A. Nickel, Non-commutative Fitting invariants and annihilation of class groups, J. Algebra 323 (2010), no. 10, 2756–2778.
  • [42] W. Nizioł, Cohomology of crystalline representations, Duke Math. J. 71 (1993), no. 3, 747–791.
  • [43] R. Oliver, Whitehead groups of finite groups, London Mathematical Society Lecture Note Series, vol. 132, Cambridge University Press, Cambridge, 1988.
  • [44] A. Parker, Equivariant Tamagawa numbers and non-commutative Fitting invariants, Ph.D. thesis, King’s College London, 2007, available at http://www.mth.kcl.ac.uk/staff/dj_burns/former_students/andy_parker/The%sisAndrewParker.pdf.
  • [45] J-P. Serre, Corps locaux, Hermann, Paris, 1968, Deuxième édition, Publications de l’Université de Nancago, No. VIII.
  • [46] by same author, Propriétés galoisiennes des points d’ordre fini des courbes elliptiques, Invent. Math. 15 (1972), no. 4, 259–331.
  • [47] J. H. Silverman, Advanced topics in the arithmetic of elliptic curves, Graduate Texts in Mathematics, vol. 151, Springer-Verlag, New York, 1994.
  • [48] W. A. Stein et al., Sage Mathematics Software (Version 5.0), The Sage Development Team, 2011, http://www.sagemath.org.
  • [49] The PARI Group, Bordeaux, PARI/GP, version 2.5.0, 2011, available from http://pari.math.u-bordeaux.fr/.
  • [50] O. Venjakob, From the Birch and Swinnerton-Dyer conjecture to non-commutative Iwasawa theory via the equivariant Tamagawa number conjecture—a survey, LL-functions and Galois representations, London Math. Soc. Lecture Note Ser., vol. 320, Cambridge Univ. Press, Cambridge, 2007, pp. 333–380.