Startseite Congruences for critical values of higher derivatives of twisted Hasse–Weil L-functions
Artikel Öffentlich zugänglich

Congruences for critical values of higher derivatives of twisted Hasse–Weil L-functions

  • Werner Bley EMAIL logo und Daniel Macias Castillo
Veröffentlicht/Copyright: 7. Oktober 2014

Abstract

Let A be an abelian variety over a number field k and let F be a finite cyclic extension of k of p-power degree for an odd prime p. Under certain technical hypotheses, we obtain a reinterpretation of the equivariant Tamagawa number conjecture (‘eTNC’) for A, F/k and p as an explicit family of p-adic congruences involving values of derivatives of the Hasse–Weil L-functions of twists of A, normalised by completely explicit twisted regulators. This reinterpretation makes the eTNC amenable to numerical verification and furthermore leads to explicit predictions which refine well-known conjectures of Mazur and Tate.

1 Introduction

Let A be an abelian variety of dimension d defined over a number field k. We write At for the dual abelian variety. Let F/k be a finite Galois extension with group G:=Gal(F/k). We let AF denote the base change of A through F/k and consider the motive MF:=h1(AF)(1) as a motive over k with a natural action of the semi-simple -algebra [G].

We will study the equivariant Tamagawa number conjecture as formulated by Burns and Flach in [9] for the pair (MF,[G]). This conjecture asserts the validity of an equality in the relative algebraic K-group K0([G],[G]). If p is a prime, we refer to the image of this equality in K0(p[G],p[G]) as the ‘eTNCp for (MF,[G])’ (here p denotes the completion of an algebraic closure of p). If p does not divide the order of G then the ring p[G] is regular and one can use the techniques described in [8, Section 1.7] to give an explicit interpretation of this projection. In this manuscript we will focus on primes p dividing the order of G, for which such an interpretation is in general very difficult to obtain.

In [12], a close analysis of the finite support cohomology of Bloch and Kato for the base change of the p-adic Tate module of At is carried out under certain technical hypotheses on A and F. A consequence of this analysis is an explicit reinterpretation of the eTNCp in terms of a natural equivariant regulator (see [12, Theorem 5.1]). The main results of the present manuscript are based on the computation of this equivariant regulator in the special case where F/k is cyclic of degree pn for an odd prime p. Under certain additional hypotheses on the structure of Tate–Shafarevich groups of A over the intermediate fields of F/k we obtain a completely explicit interpretation of the eTNCp (see Theorem 2.9). Whilst this is of independent theoretical interest, it also makes the eTNCp amenable to numerical verifications.

One of the main motivations behind our study of the equivariant Tamagawa number conjecture for the pair (MF,[G]) is the hope that this conjecture may provide a coherent overview of and a systematic approach to the study of properties of leading terms and values at s=1 of Hasse–Weil L-functions. In order to describe our current steps in this direction, we first recall the general philosophy of ‘refined conjectures of the Birch and Swinnerton-Dyer type’ that originates in the work of Mazur and Tate in [20]. These conjectures concern, for elliptic curves A defined over and certain abelian groups G, the properties of modular elementsθA,G belonging a priori to the rational group ring [G] and constructed from the modular symbols associated to A, therefore interpolating the values at s=1 of the twisted Hasse–Weil L-functions associated to A and G. More precisely, the aim is to predict the precise power r (possibly infinite) of the augmentation ideal I of the integral group ring [G] with the property that θA,G belongs to Ir but not to Ir+1, and furthermore to describe the image of θA,G in the quotient Ir/Ir+1 (whenever such an integer r exists). In the process of studying the modular element θA,G, Mazur and Tate also predict that it should belong to the Fitting ideal over [G] of their integral Selmer groupS(A/F) (and refer to such a statement as a weak main conjecture) and ask for a strong main conjecture predicting a generator of the Fitting ideal of an explicitly described natural modification of S(A/F) (see [20, Remark after Conjecture 3]).

However, it is well known that in many cases of interest the modular element θA,G vanishes, thus rendering any such properties trivial, and it would therefore be desirable to carry out an analogous study for elements interpolating leading terms rather than values at s=1 of the relevant Hasse–Weil L-functions, normalised by appropriate regulators. Although the aim to study such elements already underlies the results of [12], one of the main advantages of confining ourselves to the special case in which the given extension of number fields F/k is cyclic of prime-power degree is that we are led to defining completely explicit twisted regulators from our computation of the aforementioned equivariant regulator of [12]. Furthermore, we arrive at very explicit statements without having to restrict ourselves to situations in which the relevant Mordell–Weil groups are projective when considered as Galois modules. In particular, we derive predictions of the following nature for such an element that interpolates leading terms at s=1 of twisted Hasse–Weil L-functions normalised by our twisted regulators from the assumed validity of the eTNCp for (MF,[G]):

  1. a formula for the precise power h0 of the augmentation ideal IG,p of the integral group ring p[G] with the property that belongs to IG,ph but not to IG,ph+1 (expressed in terms of the ranks of the Mordell–Weil groups of A over the intermediate fields of F/k), and a formula for the image of in IG,ph/IG,ph+1 (see Corollary 2.11);

  2. the statement that the element of p[G] (resp. a straightforward modification of ) annihilates the p-primary Tate–Shafarevich group of At (resp. A) over F as a Galois module (see Theorem 2.12 and Corollary 2.14);

  3. the explicit description of a natural quotient of (the Pontryagin dual of) the p-primary Selmer group of A over F whose Fitting ideal is generated by (see Theorem 2.12).

The structure of the paper is as follows. In Section 2 we present our main results and in Section 4 we supply the proofs. In order to prepare for the proofs we recall in Section 3 the relevant material from [12]. In the final Section 5 we present some numerical computations.

1.1 Notations and setting

We mostly adapt the notations from [12].

For a finite group Γ we write D(p[Γ]) for the derived category of complexes of left p[Γ]-modules. We also write Dp(p[Γ]) for the full triangulated subcategory of D(p[Γ]) comprising complexes that are perfect (that is, isomorphic in D(p[Γ]) to a bounded complex of finitely generated projective p[Γ]-modules).

We write Γ^ for the set of irreducible E-valued characters of Γ, where E denotes either or p (we will throughout our arguments have fixed an isomorphism of fields j:p and use it to implicitly identify both sets, with the intended meaning of Γ^ always clear from the context). We let 𝟏Γ denote the trivial character of Γ and write ψˇ for the contragredient character of each ψΓ^. We write

eψ=ψ(1)|Γ|γΓψ(γ)γ-1

for the idempotent associated with ψΓ^ and also set

TrΓ:=γΓγ.

For any abelian group M we let Mtor denote its torsion subgroup and we denote by Mtf the torsion-free quotient M/Mtor. We also set Mp:=pM and, if M is finitely generated, we set rk(M):=dim(M).

For any p[Γ]-module M we write M for the Pontryagin dual Homp(M,p/p) and M* for the linear dual Homp(M,p), each endowed with the natural contragredient action of Γ. Explicitly, for a homomorphism f and elements mM and γΓ, one has (γf)(m)=f(γ-1m).

For any Galois extension of fields L/K we abbreviate Gal(L/K) to GL/K. We fix an algebraic closure Kc of K and abbreviate GKc/K to GK. For each non-archimedean place v of a number field we write κv for the residue field.

Throughout this paper, we will consider the following situation. We have fixed an odd prime p and a Galois extension F/k of number fields with group G=GF/k. Except in Section 3, the extension F/k will always be cyclic of degree pn. We give ourselves an abelian variety A of dimension d defined over k. For each intermediate field L of F/k we write SpL, SrL and SbL for the set of non-archimedean places of L which are p-adic, which ramify in F/L and at which A/L has bad reduction respectively. Similarly, we write SL, SL and SL for the sets of archimedean, real and complex places of L respectively. If L=k, we simply write Sp, Sr, Sb, S, S and S.

Finally, we write A(L) for the Mordell–Weil group and Шp(AL) for the p-primary Tate–Shafarevich group of A over L.

2 Statement of the main results

Recall that A is an abelian variety of dimension d defined over the number field k. Furthermore, F/k is cyclic of degree pn where p is an odd prime.

We assume throughout this section that A/k and F/k are such that

  1. p|A(k)tor||At(k)tor|,

  2. pvSbcv(A,k), where cv(A,k) denotes the Tamagawa number of A at v,

  3. A has good reduction at all p-adic places of k,

  4. p is unramified in F/,

  5. no place of bad reduction for A is ramified in F/k, i.e. SbSr=,

  6. pvSr|A(κv)|,

  7. the Tate–Shafarevich group Ш(AF) is finite,

  8. Шp(AFH)=0 for all non-trivial subgroups H of G.

Remarks 2.1

Our assumptions (a)–(g) recover [12, hypotheses (a)–(h)]. For a fixed abelian variety A/k, the hypotheses (a), (b) and (c) clearly exclude only finitely many choices of odd prime p, while the additional hypotheses (d), (e) and (f) constitute a mild restriction on the choice of cyclic field extension F of k of odd, prime-power degree. In order to further illustrate this point, we let S denote any finite set of places of k at which A has good reduction. We then define a set Σ(S) of rational primes as the union of the set of all prime divisors of

|A(k)tor||At(k)tor|vSbcv(A,k)vS|A(κv)|

and the set of all primes with the property that A has bad reduction at an -adic place of k. The set Σ(S) is then clearly finite and, for any odd prime pΣ(S) and any cyclic field extension F of k of p-power degree which is unramified outside S and with the property that p is unramified in F/, all of the hypotheses (a)–(f) are satisfied.

The hypothesis (g) is famously conjectured to be true in all cases, and it is straightforward to produce specific examples for which all of the other hypotheses, including the additional hypothesis (h), are satisfied (see also Section 5 and Remark 2.13 in this regard). We emphasize that in (h) we allow Шp(AF) to be non-trivial.

An understanding of the G-module structure of the relevant Mordell–Weil groups is key to our approach. We hence begin by applying a result of Yakovlev [22] in order to obtain such explicit descriptions. This approach is inspired by work of Burns, who first obtained a similar result in [7, Proposition 7.2.6 (i)]. For a non-negative integer m and a p[G]-module M we write Mm for the direct sum of m copies of M. Furthermore, we set [m]:={1,,m}.

Proposition 2.2

There exist isomorphisms of Zp[G]-modules of the form

A(F)pJGp[G/J]mJAt(F)p,

for a set of non-negative integers {mJ:JG}.

Proposition 2.2 has the following immediate consequence for the ranks of the relevant Mordell–Weil groups.

Corollary 2.3

For any subgroup H of G we have

rk(A(FH))=rk(At(FH))
=J>H|G/J|mJ+|G/H|JHmJ
|G/H|rk(A(k)).

Proposition 2.2 combines with Roiter’s lemma (see [13, (31.6)]) to imply the existence of points P(J,j)A(F) and P(J,j)tAt(F) for JG and j[mJ] with the property that

(2.1)A(F)p=JGj[mJ]p[G/J]P(J,j),p[G/J]P(J,j)p[G/J],
At(F)p=JGj[mJ]p[G/J]P(J,j)t,p[G/J]P(J,j)tp[G/J].

Furthermore, our choice of points as in (2.1) guarantees that one also has

(2.2)A(F)=JGj[mJ][G/J]P(J,j),[G/J]P(J,j)[G/J],
At(F)=JGj[mJ][G/J]P(J,j)t,[G/J]P(J,j)t[G/J].

We now fix sets

𝒫={P(J,j)A(F):JG,j[mJ]},
𝒫t={P(J,j)tAt(F):JG,j[mJ]}

such that (2.2) holds. For 0tn we write Ht for the (unique) subgroup of G of order pn-t and set

P(t,j):=P(Ht,j),P(t,j)t:=P(Ht,j)t.

We also put mt:=mHt and

eHt:=1|Ht|TrHt=1|Ht|gHtg.

We write ,F for the Néron–Tate height pairing A(F)×At(F) defined relative to the field F and define a matrix with entries in [G] by setting

R(𝒫,𝒫t):=(1|Hu|τG/HuτP(u,k),P(t,j)tF(τeHu))(u,k),(t,j),

where (u,k) is the row index with 0un, k[mu], and (t,j) is the column index with 0tn, j[mt] (we always order sets of the form {(t,j):0tn,j[mt]} lexicographically). We note that, since each point P(u,k) belongs to A(FHu), the action of G/Hu on P(u,k) is well-defined.

For any matrix A=(a(u,k),(t,j))(u,k),(t,j) indexed as above we define

At0:=(a(u,k),(t,j))(u,k),(t,j),u,tt0

with the convention At0=1 whenever no entries a(u,k),(t,j) with u,tt0 exist. If A is a matrix with coefficients aij in [G] or p[G], then for any ψG^ we write ψ(A) for the matrix with coefficients ψ(aij). We also set Rt0(𝒫,𝒫t)=R(𝒫,𝒫t)t0.

Definition 2.4

For each character ψG^ we define tψ{0,,n} by the equality

ker(ψ)=Htψ

and call

λψ(𝒫,𝒫t):=det(ψ(Rtψ(𝒫,𝒫t)))

the lower ψ-minor of R(𝒫,𝒫t).

Remark 2.5

It is easy to see that the element

ψG^λψ(𝒫,𝒫t)eψ[G]

depends upon the choice of points 𝒫 and 𝒫t satisfying (2.2) only modulo [G]×. Similarly, for any given isomorphism of fields j:p, it is clear that the element

ψG^j(λψ(𝒫,𝒫t))eψp[G]

depends upon the choice of points 𝒫 and 𝒫t satisfying (2.1) only modulo p[G]×.

For any order Λ in [G] that contains [G] we let C(A,Λ) denote the integrality part of the equivariant Tamagawa number conjecture (‘eTNC’ for brevity) for the pair (h1(AF)(1),Λ) as formulated by Burns and Flach in [9, Conjecture 4 (iv)]. Similarly, we let C(A,[G]) denote the rationality part as formulated in [9, Conjecture 4 (iii) or Conjecture 5]. We recall that, under the assumed validity of hypothesis (g), C(A,Λ) takes the form of an equality in the relative K-group K0(Λ,[G]). For each embedding j:p we denote by Cp,j(A,Λ) the image of this conjectural equality under the induced map K0(Λ,[G])K0(Λp,p[G]). We then say that Cp(A,Λ) is valid if Cp,j(A,Λ) is valid for every isomorphism j:p.

The eTNC is an equality between analytic and algebraic invariants associated with A/k and F/k. In the following we describe and define the analytic part. We first recall the definition of periods and Galois–Gauss sums of [12, Section 4.4]. We fix Néron models 𝒜t for At over 𝒪k and 𝒜vt for Akvt over 𝒪kv for each v in Sp and then fix a k-basis {ωb}b[d] of the space of invariant differentials H0(At,ΩAt1) which gives 𝒪kv-bases of H0(𝒜vt,Ω𝒜vt1) for each such v and is also such that each ωb extends to an element of H0(𝒜t,Ω𝒜t1).

For each v in S we fix a -basis {γv,a}a[2d] of H1(σv(At)(),). For each v in S we let c denote complex conjugation and fix a -basis {γv,a+}a[d] of H1(σv(At)(),)c=1. For each v in S, resp. S, we then define periods by setting

Ωv(A/k):=|det(γv,a+ωb)a,b|,resp.Ωv(A/k):=|det(γv,aωb,c(γv,aωb))a,b|,

where in the first matrix (a,b) runs over [d]×[d] and in the second matrix (a,b) runs over [2d]×[d].

In our special case all characters are one-dimensional and, moreover, |G| is odd. Therefore the definitions of [12] simplify and we set

Ω(A/k):=vSΩv(A/k),
w(k):=i|S|.

For each place v in Sr we write I¯vG for the inertia group of v and Frv for the natural Frobenius in G/I¯v. We define the non-ramified characteristicuv by

uv(ψ):={-ψ(Frv-1),ψ|I¯v=1,1,ψ|I¯v1,

and

u(ψ):=vSruv(ψ).

For each character ψG^ we then define the modified Galois–Gauss sum by setting

τ*(,indk(ψ)):=u(ψ)τ(,indk(ψ))(c)×,

where each individual Galois–Gauss sum τ(,) is as defined by Martinet in [19]. For each character ψG^ we set

ψ*=A,F/k,ψ*:=LSr*(A,ψˇ,1)τ*(,indk(ψ))dΩ(A/k)w(k)d×,

where here for each finite set Σ of places of k we write LΣ*(A,ψ,1) for the leading term in the Taylor expansion at s=1 of the Σ-truncated ψ-twisted Hasse–Weil-L-function of A. Without any further mention we will always assume that the functions LΣ(A,ψ,s) have analytic continuation to s=1 (as conjectured in [9, Conjecture 4 (i)]) and recall that they are then expected to have a zero of order rψ:=dim(eψ(A(F))) (this is the rank conjecture [9, Conjecture 4 (ii)]).

We finally define

*=A,F/k*:=ψG^A,F/k,ψ*eψ[G]×

and note that the element * defined in [12, Theorem 5.1] specialises precisely to our definition.

Theorem 2.6

Conjecture C(A,Q[G]) is valid if and only if

ψ*λψ(𝒫,𝒫t)-1(ψ)

for all ψG^ and furthermore, for any γGal(Q(ψ)/Q),

ψγ*λψγ(𝒫,𝒫t)-1=γ(ψ*λψ(𝒫,𝒫t)-1)

for any, or equivalently every, choice of points P and Pt such that (2.2) holds.

Remarks 2.7

(i) From the definitions of u(ψ), w(k) and the definition of local Euler factors it is immediately clear that in the statement of Theorem 2.6 we can replace ψ* by

~ψ*:=L*(A,ψˇ,1)τ(,indk(ψ))dΩ(A/k).

(ii) The explicit conditions on elements of the form ψ*λψ(𝒫,𝒫t)-1 given in Theorem 2.6 generalise and refine the predictions given by Fearnley and Kisilevsky in [16, 17]. For details see [2, Example 5.2]. In particular, we note that the numerical computations performed by Fearnley and Kisilevsky can be interpreted via Theorem 2.6 as supporting evidence for conjecture C(A,[G]).

We fix a generator σ of G and define Σ to be the diagonal matrix indexed by pairs (t,j) and (s,i) with σpt-1 at the diagonal entry associated to (t,j) and zeros elsewhere. For any matrix A=(a(u,k),(t,j))(u,k),(t,j) indexed by tuples (u,k) and (t,j) as above we define

At0:=(a(u,k),(t,j))(u,k),(t,j),u,tt0,

once again with the convention At0=1 whenever no entries a(u,k),(t,j) with u,tt0 exist. We recall that for each character ψG^ we defined tψ such that ker(ψ)=Htψ. We define the upper ψ-minor of Σ by

δψ:=det(ψ(Σtψ-1)).

It is easy to see that for another choice of generator of G, say τ, one has

ψG^δψ(σ)δψ(τ)eψp[G]×.

Under our current hypotheses on the data (A,F/k,p) and the additional hypothesis that Шp(AF)=0, and for any intermediate field L of F/k, we shall say that BSDp(L) holds if, for any choice of -bases {Qi} and {Rj} of A(L) and At(L) respectively and of isomorphism j:p, one has that

j(L*(A/L,1)(|dL|)ddet(Qi,RjL)vSLΩv(A/L))p×.

Here dL denotes the discriminant of the field L and each period Ωv(A/L) is as defined above but relative to the field L rather than k. It will become apparent in the proof of Theorem 2.8 below that the validity of BSDp(L) is equivalent to the validity of the p-part of the eTNC for the pair (h1(AL)(1),). We recall that hypotheses (a), (b) and (h) justify the fact that no orders of torsion subgroups of Mordell–Weil groups, Tamagawa numbers or orders of Tate–Shafarevich groups occur in this formulation, and furthermore note that, by explicitly computing integrals, the periods Ωv(A/L) 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 [18, p. 224].

For the remainder of this section, we assume that C(A,[G]) is valid. It is then easy to see that, for any order Λ in [G] that contains [G], the validity of Cp,j(A,Λ) is independent of the choice of isomorphism j:p, and so we fix such a j for the remainder of this section. In fact, all relevant elements of [G] appearing in the statements of our results will actually belong to [G] (as a consequence of an easy application of Theorem 2.6) and so we will consider them simultaneously as elements of p[G]p[G] in the natural way without any explicit mention of j.

Let denote the maximal -order in [G]. For any ψG^, let 𝒪ψ be the valuation ring of p(ψ). Let 𝔭ψ be the (unique) prime ideal of 𝒪ψ above p. We write v𝔭ψ for the normalised valuation defined by 𝔭ψ.

Theorem 2.8

Let P and Pt be any choice of points such that (2.1) holds. We assume that Шp(AF)=0. Then the following are equivalent.

  1. Cp(A,) is valid.

  2. BSDp(L) is valid for all intermediate fields L of F/k.

  3. For each ψG^ one has

    v𝔭ψ(ψ*λψ(𝒫,𝒫t))=bψwhere bψ:=s=0tψ-1psms.
  4. One has

    ψG^ψ*λψ(𝒫,𝒫t)δψeψp×.

To describe the full range of implications of the validity of Cp(A,[G]) requires yet more work and some further notations.

For each finite extension L/k and natural number n we write Sel(pn)(AL) for the Selmer group associated to the isogeny [pn]. We define the p-primary Selmer group by

Selp(AL):=limSel(pn)(AL).

We recall that one then obtains a canonical short exact sequence

0p/pA(F)Selp(AF)Шp(AF)0

of p[G]-modules, from which upon taking Pontryagin duals one derives a canonical short exact sequence

(2.3)0Шp(AF)Selp(AF)A(F)p*0.

We will throughout use this canonical short exact sequence to fix identifications of the torsion subgroup (Selp(AF))tor with Шp(AF) and of (Selp(AF))tf with A(F)p*.

In [12] a suitable integral model RΓf(k,Tp,F(A)) of the finite support cohomology of Bloch and Kato for the base change through F/k of the p-adic Tate module of At is defined and then used in order to define an equivariant regulator which is essential to the explicit reformulation of Cp(A,[G]) (see [12, Theorem 5.1]). We will recall this reformulation in Section 3.

By [12, Lemma 4.1], RΓf(k,Tp,F(A)) is under our current hypotheses a perfect complex of p[G]-modules which is acyclic outside degrees 1 and 2 and whose cohomology groups in degrees 1 and 2 canonically identify with At(F)p and Selp(AF) respectively. We recall that given any complex E with just two non-zero cohomology modules Hm(E) and Hn(E), n>m, the complex τnτmE represents an element in the Yoneda ext-group Extp[G]n-m+1(Hn(E),Hm(E)). Here τ is the truncation of complexes preserving cohomology in the indicated degrees. In this way, RΓf(k,Tp,F(A)) uniquely determines a class δA,K,p in Extp[G]2(Selp(AF),At(F)p). The element δA,K,p is furthermore perfect, meaning that it can be represented as a Yoneda 2-extension by a four term exact sequence in which each of the two middle modules is perfect when considered as an object of D(p[G]). We will use Proposition 2.2 to fix an explicit 2-syzygy of the form

(2.4)0MιF0F1A(F)p*0,

in which we set

M:=(t,j)p[G/Ht]

and both F0 and F1 are finitely generated free p[G]-modules and then use the exact sequence (2.4) to compute Extp[G]2(A(F)p*,At(F)p) via the canonical isomorphism

Extp[G]2(A(F)p*,At(F)p)Homp[G](M,At(F)p)/ι*(Homp[G](F0,At(F)p)).

If we now assume that Шp(AF) vanishes, we may identify Selp(AF) and A(F)p*, so that δA,F,p uniquely determines an element of the above quotient. We will prove (see Lemmas 4.2 and 4.3 below) that we may choose a representative ΦHomp[G](M,At(F)p) of δA,F,p with the following properties:

  1. Φ is bijective,

  2. for every j[mn], Φ restricts to send an element x(n,j) of the (n,j)-th direct summand p[G] to x(n,j)P(n,j)t.

For a fixed choice of points 𝒫 and 𝒫t such that the conditions in (2.1) hold and of a representative ΦHomp[G](M,At(F)p) as above, we fix a canonical p[G/Ht]-basis element e(t,j) of each direct summand p[G/Ht] of M and fix any elements Φ(t,j),(s,i) of p[G] with the property that

(2.5)Φ(e(s,i))=(t,j)Φ(t,j),(s,i)P(t,j)t.

We thus obtain an invertible matrix (Φ(t,j),(s,i))(t,j),(s,i) with entries in p[G], which by abuse of notation we shall also denote by Φ. The matrix Φ is of the form

(2.6)((Φ(t,j),(s,i))t,s<n0000Imn)

with Imn denoting the identity mn×mn matrix.

Recall the definition of tψ in Definition 2.4. We define the lower ψ-minor of Φ by setting

εψ(Φ):=det(ψ(Φtψ)).

We note firstly that, since the chosen points P(t,j)t satisfy (2.1), each element εψ(Φ) (and, indeed, even the matrix ψ(Φtψ)) is independent of our particular choice of Φ(t,j),(s,i)p[G] with the property that (2.5) holds.

Theorem 2.9

Let P and Pt be any choice of points such that (2.1) holds. Assume that Шp(AF)=0. Let ΦHomZp[G](M,At(F)p) be any representative of δA,F,p such that (P1) and (P2) hold. Then Cp(A,Z[G]) is valid if and only if

(2.7)ψG^ψ*λψ(𝒫,𝒫t)εψ(Φ)δψeψp[G]×.
Remark 2.10

Theorem 2.9 can be reformulated in terms of explicit congruences.

Via Theorem 2.9, we now obtain completely explicit predictions concerning congruences in the augmentation filtration of the integral group ring p[G] for leading terms at s=1 of the relevant Hasse–Weil-L-functions of A normalised by our twisted regulators. We recall that such predictions constitute a refinement and generalisation of the congruences for modular symbols that are conjectured by Mazur and Tate in [20].

In order to state such conjectural congruences, we require the following notation: if the inequality rk(A(FJ))|G/J|rk(A(k)) of Corollary 2.3 is strict for some subgroup J of G, we may and will denote by H=Ht0 the smallest non-trivial subgroup of G with the property that mH0. Hence t0 is the maximal index with the properties mt00 and t0<n. We then define

:={ψG^ψ*det(ψ(R(𝒫,𝒫t)))eψ,if rk(A(FJ))=|G/J|rk(A(k)) for every J,ψ|H1ψ*λψ(𝒫,𝒫t)eψ,otherwise.

We also let IG,p denote the kernel of the augmentation map p[G]p.

Corollary 2.11

Let P and Pt be any choice of points such that (2.1) holds. Assume that Шp(AF)=0. Let ΦHomZp[G](M,At(F)p) be any representative of δA,F,p such that (P1) and (P2) hold. If Cp(A,Z[G]) is valid, then the following hold:

  1. belongs to the ideal IG,ph of p[G], where h:=t<nmt.

  2. One has ϵ:=det(𝟏G(Φ))p×.

  3. One has

    v:=(-1)d|Sr|LSr*(A/k,1)(|dk|)dΩ(A)det(𝟏G(R(𝒫,𝒫t)))p×.
  4. One has

    vϵt<n(σpt-1)mt(modIG,ph+1).

The theory of organising matrices developed by Burns and the second named author in [10] allows one to derive the containment IG,ph of Corollary 2.11 (i) from the assumed validity of conjecture Cp(A,[G]) in situations in which Шp(AF) is non-trivial. In this greater level of generality, it furthermore leads to explicit statements concerning annihilation of Tate–Shafarevich groups and (generalised) strong main conjectures of the kind that Mazur and Tate ask for in [20, Remark after Conjecture 3]. Namely, we obtain the following result.

Theorem 2.12

Let P,Pt be any choice of points such that (2.1) holds. If Cp(A,Z[G]) is valid, then the following hold:

  1. belongs to the ideal IG,ph of p[G], where h:=t<nmt.

  2. annihilates the p[G]-module Шp(AFt).

  3. There exists a (finitely generated) free p[G]-submodule Π of Selp(AF) of (maximal) rank mn with the property that generates the Fitting ideal of Selp(AF)/Π.

Remark 2.13

It will become clear in the course of the proof that, provided that there exist sets of points 𝒫 and 𝒫t such that (2.1) holds from which one may construct the element , Theorem 2.12 remains valid even if hypothesis (h) fails to hold. This fact is relevant because, as we will see in Section 5, it allows us to obtain numerical supporting evidence for Cp(A,[G]) (via verifying the explicit assertions of Theorem 2.12) in a wider range of situations.

Let #:p[G]p[G] denote the involution induced by gg-1. Recalling that the Cassels–Tate pairing induces a canonical isomorphism between Шp(AF) and Шp(AFt), we immediately obtain the following corollary:

Corollary 2.14

Under the assumptions of Theorem 2.12 one has that the element L# of IG,ph annihilates the Zp[G]-module Шp(AF).

3 An explicit reformulation of conjecture Cp(A,[G])

3.1 K-theory and refined Euler characteristics

Let R be either or p and, for the moment, let G be any finite group. We write K for the quotient field of R and let 𝔼 be a field extension of K. Let Λ be an R-order in K[G]. We recall that there is a canonical exact sequence of algebraic K-groups

(3.1)K1(Λ)K1(𝔼[G])Λ,𝔼1K0(Λ,𝔼[G])K0(Λ)K0(𝔼[G])

where K0(Λ,𝔼[G]) is the relative algebraic K-group, as defined by Swan in [21, p. 215], associated to the ring inclusion Λ𝔼[G].

For any ring Σ we write ζ(Σ) for its center. We let nr𝔼[G]:K1(𝔼[G])ζ(𝔼[G])× denote the (injective) homomorphism induced by the reduced norm map. If Λ is a -order in [G] we write

δG:ζ([G])×K0(Λ,[G]),δG,p:ζ(p[G])×K0(Λp,p[G])

for the extended boundary homomorphisms as defined in [9, Section 4.2]. Recall that

δGnr[G]=Λ,1,δG,pnrp[G]=Λp,p1.

By the general construction described in [9, Proposition 2.5] (and [5, Lemma 5.1]) each pair (C,λ) consisting of a complex CDp(Λp) and an isomorphism of p[G]-modules

λ:pp(iH2i(C))pp(iH2i+1(C))

gives rise to a refined Euler characteristic χG,p(C,λ)K0(Λp,p[G]). For an explicit example of the computation of χG,p(C,λ) in a special case, which is also relevant for the computations in this paper, we refer the reader to [3, Section 3].

It is well known that Λp,p1 is onto and that nrp[G] is an isomorphism. We therefore deduce from (3.1) that

(3.2)K0(Λp,p[G])ζ(p[G])×/nrp[G](K1(Λp)).

Since Λp is semilocal, we can replace K1(Λp) by Λp× in (3.2). Moreover, it follows from (3.2) that for an element ξζ(p[G])× one has that δG,p(ξ)=0 if and only if ξnrp[G](Λp×). Finally, if G is abelian, we have that

K0(Λp,p[G])p[G]×/Λp×,

and hence δG,p(ξ)=0 if and only if ξΛp×.

In this context we shall also recall [2, Lemma 2.5]. We naturally interpret K0(Λ,[G]) and K0(Λp,p[G]) as subgroups of K0(Λ,[G]) and K0(Λp,p[G]) respectively, and recall that if ξζ([G])×, then

δG(ξ)K0(Λ,[G])ξζ([G])×

while if ξζ(p[G])×, then

δG,p(ξ)K0(Λp,p[G])ξζ(p[G])×.

We finally recall that, for any isomorphism j:p, there is an induced composite homomorphism of abelian groups

jG,*:K0(Λ,[G])K0(Λ,[G])K0(Λ,p[G])K0(Λp,p[G])

(where the first and third arrows are induced by the inclusions [G][G] and ΛΛp respectively). We also write j*:ζ([G])×ζ(p[G])× for the obvious map induced by j, and note that it is straightforward to check that one has

jG,*δG=δG,pj*.

3.2 Relevant results from [12]

Conjecture C(A,[G]) is formulated as the vanishing of the equivariant Tamagawa numberTΩ(h1(AF)(1),[G]) of K0([G],[G]) that is defined in [9, Conjecture 4] and constructed (unconditionally under the assumed validity of hypothesis (g)) via the formalism of virtual objects from the various canonical comparison morphisms between the relevant realisations and cohomology spaces associated to the motive h1(AF)(1) (for more details see [9]).

Motivated by work of Bloch and Kato, and in order to isolate the main arithmetic difficulties involved in making TΩ(h1(AF)(1),[G]) explicit, the approach of [12] relies upon the definition of a suitable (global) finite support cohomology complex

CA,Ff,:=RΓf(k,Tp,F(A))

(see [12, Section 4.2]). Under the hypotheses of [12] the complex CA,Ff, is perfect and acyclic outside degrees 1 and 2. Moreover, there are canonical identifications of H1(CA,Ff,) and H2(CA,Ff,) with At(F)p and Selp(AF) respectively (see [12, Lemma 4.1]). Hence, for a given isomorphism j:p, the -linear extension of the Néron–Tate height pairing of A defined relative to the field F induces a canonical trivialisation

λA,FNT,j:ppH1(CA,Ff,)ppAt(F)p
p,j(At(F))
p,jHom(A(F),)
ppHomp(A(F)p,p)
ppH2(CA,Ff,).

It is finally proved in [12, Theorem 5.1] that

(3.3)jG,*(TΩ(h1(A/F)(1),[G]))=δG,p(j*(A,F/k*))+χG,p(CA,Ff,,(λA,FNT,j)-1),

or equivalently that conjecture Cp,j(A,[G]) is valid if and only if

(3.4)δG,p(j*(A,F/k*))=-χG,p(CA,Ff,,(λA,FNT,j)-1).

In order to prove our results stated in Section 2 above we must therefore compute the refined Euler characteristic χG,p(CA,Ff,,(λA,FNT,j)-1) in terms of the heights of the chosen sets of points 𝒫 and 𝒫t.

4 The proofs

4.1 The proof of Proposition 2.2

In this subsection we will prove Proposition 2.2. The existence of global points P(t,j) and P(t,j)t such that (2.1) holds is then an immediate consequence of Roiter’s lemma (see [13, (31.6)]).

To ease notation we set

H1:=H1(CA,Ff,)=At(F)p

and

H2:=H2(CA,Ff,)=Selp(AF).

We recall that, for any intermediate field L of F/k, we may and will use the relevant canonical short exact sequence of the form (2.3) to identify (Selp(AL))tor with Шp(AL) and (Selp(AL))tf with A(L)p*.

Under the assumed validity of hypotheses (a)–(e), the result of [11, Proposition 3.1] directly combines with hypothesis (h) to imply that, for every non-trivial subgroup J of G, the Tate cohomology group H^-1(J,Htf2) vanishes and the module (H2)J is torsionfree. By the definition of Tate cohomology, we have that the finite group H^-1(J,H2) identifies with a submodule of (H2)J and therefore vanishes too.

Furthermore, since the complex CA,Ff, is perfect and acyclic outside degrees 1 and 2, for each subgroup J of G the group H^1(J,H1) is isomorphic to H^-1(J,H2). In addition, since G is cyclic, the Tate cohomology of each J is periodic of order 2 and so H^-1(J,H1) also vanishes.

We next note that, since G is a p-group, hypothesis (a) implies that At(F)p=H1 is torsion-free.

We apply the main result [22, Theorem 2.4] of Yakovlev to see that both At(F)p=H1 and A(F)p*=Htf2 are p[G]-permutation modules, that is, that there exist isomorphisms of the form

At(F)pJGp[G/J]rJ,A(F)p*JGp[G/J]sJ

for some sets of non-negative integers {rJ} and {sJ}. But the Néron–Tate height pairing induces an isomorphism of p[G]-modules between ppAt(F)p and ppA(F)p* and so by rank considerations we find that rJ=sJ=:mJ for every J. Finally, it is easy to see that the p-linear dual of a permutation module is again a permutation module of the same form. Therefore the canonical isomorphism A(F)p**A(F)p shows that one also has that

A(F)pJGp[G/J]mJ.

4.2 The proof of Theorem 2.9

Recall that in addition to our running hypothesis (a)–(h) we also assume that Шp(AF)=0. In particular, we identify Selp(AF) with A(F)p* via the canonical map in (2.3).

We fix an isomorphism of fields j:p. From formula (3.4) and the discussion in Section 3.1 it is clear that it will be enough to show that

(4.1)-χG,p(CA,Ff,,(λA,FNT,j)-1)=δG,p(ψG^j(λψ(𝒫,𝒫t))ϵψ(Φ)δψeψ)

(we recall that, since we have assumed the validity of C(A,[G]), one actually has that the validity of Cp,j(A,[G]) is equivalent to the validity of Cp(A,[G])).

We begin by defining, for every pair (s,i), an element P(s,i)*A(F)p* by setting, for every pair (t,j) and element τ of G,

(4.2)P(s,i)*(τP(t,j))={1,if s=t,i=j and τHs,0,otherwise.
Lemma 4.1

We have

A(F)p*=(s,i)p[G/Hs]P(s,i)*

with each summand Zp[G/Hs]P(s,i)* isomorphic to Zp[G/Hs].

Proof.

If γG, then

(4.3)(γP(s,i)*)(τP(t,j))=P(s,i)*(γ-1τP(t,j))=1
s=t,i=j and γτ(modHs).

Hence we have

γP(s,i)*=P(s,i)*

for γHs. Moreover, it easily follows that the maps γP(s,i)* with γG/Hs form a p-basis of A(F)p* (actually the p-dual basis of τP(t,j) with τG/Ht). ∎

We now proceed to fix an explicit 2-syzygy of the form (2.4). For this purpose, we first recall that Ht=σpt. For each pair (t,j) corresponding to the subgroup Ht of G and j[mt] we hence have a 2-extension

0p[G/Ht]ιtp[G]σpt-1p[G]πt,jp[G/Ht]P(t,j)*0.

In this sequence we let ιt denote the (well-defined) map which sends the image of an element xp[G] under the natural surjection p[G]p[G/Ht] to the element TrHtx of p[G], while πt,j sends the element 1 of p[G] to the element P(t,j)*A(F)p* defined in (4.2). Lemma 4.1 then implies that, summing over all pairs (t,j) we obtain a 2-extension

0MιF0ΘF1πA(F)p*0.

with

F0=F1=X:=(t,j)p[G],
M:=(t,j)p[G/Ht].

We now recall that we have a canonical isomorphism

Extp[G]2(A(F)p*,At(F)p)Homp[G](M,At(F)p)/ι*(Homp[G](F0,At(F)p))

under which an element ϕ of Homp[G](M,At(F)p) corresponds to the element ϵ(ϕ) of the group Extp[G]2(A(F)p*,At(F)p) which has the bottom row of the commutative diagram with exact rows

(4.4)
(4.4)

as a representative. In this diagram X(ϕ) is defined as the push-out of ι and ϕ.

We now proceed to prove that, when considering perfect elements

ϵ(ϕ)Extp[G]2(A(F)p*,At(F)p),

one may without loss of generality restrict attention to a special class of elements

ϕHomp[G](M,At(F)p).
Lemma 4.2

For all subgroups J of G one has

  1. Extp[G]2(p[G],p[G/J])=0,

  2. Extp[G]2(p[G/J],p[G])=0.

Proof.

Claim (i) is clear. Concerning claim (ii), we first note that since p[G/J] is p-torsion-free, there is an isomorphism of the form

Extp[G]2(p[G/J],p[G])H2(G,Homp(p[G/J],p[G])).

But the G-module Homp(p[G/J],p[G]) is cohomologically-trivial and therefore

Extp[G]2(p[G/J],p[G])=0,

as required. ∎

Lemma 4.2 now implies that we can without loss of generality restrict attention to those elements ϕ of Homp[G](M,At(F)p) which satisfy (P2) and, in addition, by the argument of [3, Lemma 4.3], which are furthermore injective.

Lemma 4.3

Suppose ϕHomZp[G](M,At(F)p) has all of the properties described in the previous paragraph. Then the element ϵ(ϕ) of ExtZp[G]2(A(F)p*,At(F)p) is perfect if and only if ϕ is an isomorphism.

Proof.

The fact that ϕ restricts to send an element x(n,j) of the (n,j)-th direct summand p[G] to x(n,j)P(n,j)t immediately implies that cok(ϕ)=cok(ϕ) where

ϕ:(t,j),t<np[G/Ht](t,j),t<np[G/Ht]P(t,j)t

is the map obtained by restriction of ϕ. Since ϕ is injective, the commutative diagram (4.4) implies that the 2-extension ϵ(ϕ) is perfect if and only cok(ϕ)=cok(ϕ) is cohomologically trivial. Note that Hn-1 clearly acts trivially on cok(ϕ). So, if cok(ϕ) is cohomologically trivial, then

cok(ϕ)/pcok(ϕ)=H^0(Hn-1,cok(ϕ))=0.

It then follows that cok(ϕ) must itself vanish, as required. ∎

We henceforth fix ΦHomp[G](M,At(F)p) representing the element

δA,F,pExtp[G]2(Selp(AF),At(F)p)

which is specified by the complex CA,Ff,. Recall that by our current assumption Шp(AF)=0 we identify Selp(AF) and A(F)p*. By Lemmas 4.2 and 4.3 we may and will assume that Φ is an isomorphism and furthermore that the matrix defined in (2.5) is of the form (2.6).

Having justified our choice of homomorphism Φ, we now proceed to compute the term

-χG,p(CA,Ff,,(λA,FNT,j)-1)

that occurs in (4.1) via a generalisation of the computations done in [3, Section 4]. For brevity, given any p[G]-module N, resp. p[G]-homomorphism h, we set

Np:=ppN,resp.hp:=pph.

For any choice of respective splittings

s1:XpMpim(Θ)pands2:Xpker(π)pA(F)p*

of the short exact sequences induced by scalar extension of

0M𝜄XΘim(Θ)0and0ker(π)X𝜋A(F)p*0

respectively, we write λA,FNT,jΦp,Θ,s1,s2 for the composite p[G]-automorphism of Xp given by

Xps1Mpim(Θ)p
(Φp,id)At(F)pim(Θ)p
(λA,FNT,j,id)A(F)p*im(Θ)p=A(F)p*ker(π)ps2-1Xp.

We also write X for the perfect complex of p[G] modules XΘX with the first term placed in degree 1 and the modules H1(X) and H2(X) identified with M and A(F)p* respectively via the top row of diagram (4.4). By unwinding the definition of refined Euler characteristics and using their basic functoriality properties one then finds that, independently of the choice of splittings s1 and s2, one has

-χG,p(CA,Ff,,(λA,FNT,j)-1)=-χG,p(X,Φp-1(λA,FNT,j)-1)
=δG,p(detp[G](λA,FNT,jΦp,Θ,s1,s2)).

The proof of equality (4.1), and hence of Theorem 2.9, will thus be achieved by the following explicit computation.

Proposition 4.4

There exist splittings s1 and s2 as above with the property that

detp[G](λA,FNT,jΦp,Θ,s1,s2)=ψG^j(λψ(𝒫,𝒫t))ϵψ(Φ)δψeψ.

Proof.

Let {w(s,i):s=0,,n,i[ms]} be the standard basis of X. For each (s,i) we write Ws=W(s,i) for the kernel of the canonical map

p[G]p[G/Hs],

so that Ws=(σps-1)p[G]=(1-eHs)p[G]. We then have a commutative diagram

(4.5)
(4.5)

with furthermore (s,i)W(s,i) equal to im(Θ)p=ker(π)p. We now fix the required splittings s1 and s2 by summing over all pairs (s,i) the splittings of the short exact sequences in (4.5) given by

(4.6)p[G]p[G/Hs]Ws,1(1|Hs|,σps-1),
(4.7)p[G]p[G/Hs]P(s,i)*Ws,1(P(s,i)*,1-eHs),

respectively. Note that for the inverse map in (4.7) we have

(P(s,i)*,0)eHsand(0,σps-1)σps-1.

After these preparations we proceed to compute the matrix ΛNT(Φ) which represents λA,FNT,jΦp,Θ,s1,s2 with respect to the fixed p[G]-basis {w(s,i)} of Xp. From (4.6) and (2.5) it follows easily that the composite of s1 and (Φp,id) maps w(s,i) to

(1|Hs|(t,j)Φ(t,j),(s,i)P(t,j)t,(,σps-1,))

in

At(F)pim(Θ)p=((t,j)p[G/Ht]P(t,j)t)((t,j)Wt)

with the only non-zero component in (t,j)Wt at the (s,i)-spot. By Lemma 4.5 below this is further mapped by (λA,FNT,j,id) to

(1|Hs|(t,j)Φ(t,j),(s,i)(u,k)(τG/Huj(τP(u,k),P(t,j)tF)τeHu)P(u,k)*,(,σps-1,)).

Rearranging the summation and applying the map s2-1 as described in (4.7) we obtain

(u,k)(1|Hs|(t,j)Φ(t,j),(s,i)τG/Huj(τP(u,k),P(t,j)tF)τeHu)w(u,k)+(σps-1)w(s,i).

We now fix a character ψG^. We have that

ψ(1|Hs|(t,j)Φ(t,j),(s,i)τG/Huj(τP(u,k),P(t,j)tF)τeHu)={1|Hs|(t,j)Φ(t,j),(s,i)τG/Huj(τP(u,k),P(t,j)t)Fψ(τ),utψ,0,u<tψ,

while ψ(σps-1) is equal to 0 if and only if stψ.

We immediately obtain that

det(ψ(ΛNT(Φ)))=j(λψ(𝒫,𝒫t))εψ(Φ)δψ,

as required. ∎

We finally provide the relevant lemma used in the course of the above proof.

Lemma 4.5

We have

λA,FNT,j(P(t,j)t)=(u,k)(τG/Huj(τP(u,k),P(t,j)tF)τeHu)P(u,k)*.

Proof.

Recall that λA,FNT,j is induced by ,F:A(F)×At(F). For PtAt(F) we explicitly have

λA,FNT,j(Pt)=j(,PtF).

Let fA(F)p* denote the map defined by the right-hand side of the equation in Lemma (4.5). From (4.3) we immediately see that eHuP(u,k)*=P(u,k)*. For each pair (v,l) and γG/Hv we hence have that

f(γP(v,l))=(u,k)τG/Huj(τP(u,k),P(t,j)tF)(τP(u,k)*)(γP(v,l))
=j(γP(v,l),P(t,j)tF)
=(λA,FNT,j(P(t,j)t))(γP(v,l)).

4.3 The proof of Theorem 2.6

We set MA,F=h1(AF)(1) and let us recall that the approach used in [9] in order to formulate the conjecture C(A,[G]) relies upon the theory of categories of virtual objects. Although focused on the study of p-parts of the relevant equivariant Tamagawa numbers for prime numbers p, the exact same techniques involved in the proof of [12, Proposition 4.2] allow one to translate this more technical language into the one of refined Euler characteristics employed throughout this article. For this reason, we will avoid any explicit mention of categories of virtual objects throughout this proof. Furthermore, since we only need to consider the relevant realisations and cohomology spaces associated to MA,F as modules over the semisimple algebras [G] or [G], we may directly reformulate C(A,[G]) in terms of the determinants of certain endomorphisms of free [G]-modules, which is what we do in the sequel.

The leading term L*(MA,F,0) at s=0 of the [G]-equivariant motivic L-function of MA,F is given by χIr(G)eχL*(A,χˇ,1). We let once again

λA,FNT:At(F)A(F)*

denote the canonical isomorphism induced by the Néron–Tate height pairing and

αA,F:vSFH0(Fv,Hv(MA,F))HdR(MA,F)/F0

denote the canonical period isomorphism described by Deligne in [14] (see also [9, Section 3] for a general description of the modules involved and [12, Section 4.3] for more details in the relevant special case).

We now note that the [G]-modules

X:=At(F)andY:=A(F)*,

resp.

Z:=vSFH0(Fv,Hv(MA,F))andW:=HdR(MA,F)/F0,

are isomorphic (as a consequence, for instance, of [1, p. 110]) and hence, since [G] is semisimple, there exist [G]-modules M and N with the property that both XMYM and ZNWN are free [G]-modules. In the sequel we (choose bases and so) fix identifications of XM, YM, ZN and WN with direct sums of copies of [G] and hence regard λA,FNTidM and αA,FidN as elements of K1([G]).

The validity of conjecture C(A,[G]) is then equivalent to the containment

L*(MA,F,0)/(det[G](αA,FidN)det[G](λA,FNTidM))[G]×.

But it is clear from the proof of Lemma 4.5 that, independently of our choice of fixed identifications,

(4.8)det[G](λA,FNTidM)/χIr(G)eχλχ(𝒫,𝒫t)[G]×,

and it is straightforward to deduce from the proof of [12, Lemma 4.5] that, independently of our choice of fixed identifications,

(4.9)det[G](αA,FidN)/χIr(G)eχw(k)dΩ(A/k)τ*(,indk(χ))d[G]×.

Equalities (4.8) and (4.9), combined with the fact that the Euler factors involved in the truncation of each of the leading terms LSr*(A,ψˇ,1) live by definition in [G]×, therefore imply that the validity of C(A,[G]) is equivalent to the containment

ψG^ψ*λψ(𝒫,𝒫t)eψ[G]×.

By [2, Lemma 2.9] this containment is equivalent to the explicit condition described in Theorem 2.6, as required.

4.4 The proof of Theorem 2.8

We assume now that C(A,[G]) is valid and proceed to prove the explicit interpretation of Cp(A,) claimed in Theorem 2.8. We begin by noting that, for any fixed isomorphism of fields j:p, the respective maps jG,* restrict to give the vertical arrows in a natural commutative diagram with exact rows of the form

(4.10)
(4.10)

We note that the exactness of the rows follows from [9, Lemma 11]. We now proceed to prove several useful results.

Lemma 4.6

Conjecture Cp(A,M) holds if and only if

jG,*(TΩ(h1(AF)(1),[G]))K0(p[G],p[G])tor.

Proof.

The equality

TΩ(h1(AF)(1),)=μ(TΩ(h1(AF)(1),[G]))

proved in [9, Theorem 4.1] combines with the commutativity of the right-hand square of diagram (4.10) to imply that

jG,*(TΩ(h1(AF)(1),))=μp(jG,*(TΩ(h1(AF)(1),[G]))).

The exactness of the bottom row of diagram (4.10) thus completes the proof. ∎

Lemma 4.7

Conjecture Cp(A,M) holds if and only if

ψG^ψ*j(λψ(𝒫,𝒫t))ϵψ(Φ)δψeψp×.

Proof.

Lemma 4.6 combines with equalities (3.3) and (4.1) to imply that Cp(A,) holds if and only if

μp(δG,p(ψG^ψ*j(λψ(𝒫,𝒫t))ϵψ(Φ)δψeψ))=0.

We next note that the respective maps δG,p induce vertical (bijective) arrows in a commutative diagram of the form

This completes the proof of the lemma. ∎

Lemma 4.8

We have εψ(Φ)Zp[ψ]×.

Proof.

The map

Φp:Mp[G]pAt(F)pp[G]p

is an isomorphism of p-modules. Since p contains the p[G]-rational idempotents,

Φp[ψ]:Mp[G]p[ψ]At(F)pp[G]p[ψ]

is an isomorphism of p[ψ]-modules. It is now easy to see that Φp[ψ] is represented by ψ(Φtψ). ∎

We now proceed to give the proof of Theorem 2.8.

The equivalence of (i) and (iv) follows directly upon combining Lemmas 4.7 and 4.8.

Furthermore it is straightforward to compute the valuation of each element δψ. One has

v𝔭ψ(δψ)=bψ

with bψ defined as in Theorem 2.8, and hence (iii) and (iv) are clearly equivalent.

In order to prove the equivalence of (i) and (ii), we will use (a special case of) a general fact which we now describe. If H is any subgroup of G, we write

ρHG:K0(p[G],p[G])K0(p[H],p[H])

for the natural restriction map and

q0H:K0(p[H],p[H])K0(p,p)

for the natural map induced by sending an element [P,ϕ,Q] of K0(p[H],p[H]) to the element [PH,ϕH,QH] of K0(p,p). By [6, Theorem 4.1] one then has that

(4.11)K0(p[G],p[G])tor=HGker(q0HρHG).

The functoriality properties of the element TΩ(h1(AF)(1),[G]) with respect to the maps ρHG and q0H proved in [9, Proposition 4.1] then imply that, for any subgroup H of G,

(q0HρHG)(jG,*(TΩ(h1(AF)(1),[G])))=j0,*(TΩ(h1(AFH)(1),)),

and so Lemma 4.6 combines with (4.11) to imply that Cp(A,) holds if and only if, for every intermediate field L of F/k, the element jGL/L,*(TΩ(h1(AL)(1),)) vanishes, that is, if and only if the p-part of the eTNC holds for the pair (h1(AL)(1),). Noting that it is easy to check that the set of data (A/L,L/L,p) satisfies all the hypotheses of Theorem 2.9 for any such field L (see for instance [11, Lemma 3.4] for a proof of a more general assertion), all that is left to do in order to prove the equivalence of (i) and (ii) is to apply Theorem 2.9. Indeed, any choice of -bases {Qi} and {Rj} of A(L) and At(L) respectively satisfy condition (2.1) for the set of data (A/L,L/L,p), while an explicit computation proves that

τ*(,indL(𝟏GL/L))w(L)=|dL|.

4.5 The proof of Corollary 2.11

For brevity we set

λψ:=λψ(𝒫,𝒫t),ϵψ:=ϵψ(Φ),u:=ψG^ψ*λψϵψδψeψ.

By Theorem 2.9 the validity of Cp(A,[G]) is equivalent to the containment up[G]×, which we assume holds throughout the proof. We also let ε:p[G]p denote the augmentation map.

We begin by noting that claim (ii) is just the ψ=𝟏G special case of Lemma 4.8, and proceed now to deduce claim (iii) from it. One clearly has that 𝟏G*/(λ𝟏Gϵ𝟏Gδ𝟏G)=ε(u)p× with δ𝟏G equal by definition to 1, while a straightforward computation shows that

τ*(,indk(𝟏G))w(k)=(-1)|Sr||dk|.

Claim (ii) therefore indeed implies that

(4.12)v=𝟏G*/λ𝟏G=ε(u)ϵ𝟏G=ε(u)ϵ

belongs to p×, as required.

In order to prove the remaining claims, we first note that, if rk(A(FJ))=|G/J|rk(A(k)) for every subgroup J of G, then h=0 by Proposition 2.2 while Φ can be chosen to be the identity matrix by property (P2) and each element δψ is simply equal to 1 by convention. In any such case, claim (i) therefore reduces to the trivial statement =up[G] while claim (iv) simply reads uv(modIG,p) and follows directly from (4.12). We therefore may and will henceforth assume that the inequality rk(A(FJ))|G/J|rk(A(k)) of Corollary 2.3 is strict for some subgroup J of G. We recall that H=Ht0 denotes the smallest non-trivial subgroup of G with the property that mH0.

In order to prove claim (i), we note first that for each ψG^ we have

ψ|H1ker(ψ)Ht0 and ker(ψ)Ht0tψ>t0.

From the definitions of ϵψ and δψ we immediately deduce that, for each character ψG^ such that ψ|H1,

ϵψ=1,δψ=δ:=j=0t0(σpj-1)mj.

Since δeψ=0 for each ψ such that ψ|H=1, we deduce that =δuδp[G]IG,ph, as required.

Finally, claim (iv) follows from (4.12) because u is clearly congruent to ε(u)=v/ϵ modulo IG,p and therefore =δu is congruent to δvϵ modulo IG,ph+1, as required.

4.6 The proof of Theorem 2.12

We begin by defining a (free) p[G]-submodule

P:=j[mn]p[G]P(n,j)*

of A(F)p* and then fix, as we may, an injective lift

κ:PSelp(AF)

of the inclusion PA(F)p* through the canonical projection of (2.3). We also fix, as we may, a representative of the perfect complex CA,Ff, of the form C1C2 in which both C1 and C2 are finitely generated, cohomologically-trivial p[G]-modules. We then obtain a commutative diagram with exact rows and columns of the form

(4.13)
(4.13)

in which we have set

N:=t<n,j[mt]p[G/Ht]

and we simply define the arrow im(κ)C2, as we may since im(κ) is a free p[G]-module, by the commutativity of the upper right square.

The p[G]-modules D1 and D2 are finitely generated and cohomologically-trivial, and hence the central arrow of the bottom row of this diagram defines an object D of Dp(p[G]) which is acyclic outside of degrees 1 and 2 and has identifications of H1(D) with N and of H2(D) with cok(κ). We analogously define an object B of Dp(p[G]) represented by the perfect complex of p[G]-modules

jp[G]P(n,j)t0im(κ).

Following [10, Section 2.1.4] we next define an idempotent

eN:=ψΥNeψ

in p[G] by letting ΥN be the subset of G^ comprising characters ψ with the property that

eψ(ppN)=0.

For any object C of Dp(p[G]) we then obtain an object

eNC:=eNp[G]p[G]𝕃C

of Dp(eNp[G]). In particular, the exact triangle represented by diagram (4.13) induces an exact triangle in Dp(eNp[G]) of the form

(4.14)eNBeNCA,Ff,eNDeNB[1].

But eNDeNp[G]eNp[G] is acyclic and an immediate application of the additivity criterion of [5, Corollary 6.6] to triangle (4.14) implies that one has

(4.15)-χeNp[G],eNp[G](eND,0)
=-χeNp[G],eNp[G](eNCA,Ff,,eN(λA,FNT,j)-1)
+χeNp[G],eNp[G](eNB,λ),

where λ denotes the canonical isomorphism

eN(ppim(κ))=eN(ppSelp(AF))
eN(λA,FNT,j)-1eN(ppAt(F)p)=eN(ppjp[G]P(n,j)t).

If we now write

φ:jp[G]P(n,j)tjp[G]P(n,j)*

for the canonical isomorphism that maps an element P(n,j)t to the element P(n,j)*, then one finds that

(4.16)χeNp[G],eNp[G](eNB,λ)
=δeNp[G],eNp[G](deteNp[G](λeN(pp(κφ))))
=-δeNp[G],eNp[G](ψΥNj(λψ(𝒫,𝒫t))eψ),

where the last equality follows from Lemma 4.5.

The assumed validity of Cp(A,[G]) therefore combines via (3.4) with equalities (4.15) and (4.16) to imply that, in the terminology of [10, Section 2.3.2], the element

j*(ψΥNψ*λψ(𝒫,𝒫t)-1eψ)=j*()

of eNp[G] is a characteristic element for eND. The result [10, Lemma 2.6] therefore implies that there exists a characteristic element for D in p[G] with the property that

eN=j*().

Since D is clearly an admissible complex of p[G]-modules (in the terminology of [10, Section 2.1.1]), the results of [10, Corollary 3.3] therefore imply that the element j*() belongs to the ideal IG,ph~ of p[G], with h~:=dimp(ppcok(κ)G), and furthermore generates Fittp[G](cok(κ)). To proceed with the proof, we first note that

h~=dimp(ppcok(κ)G)=dimp(ppt<np[G/Ht]Gmt)=t<nmt=h.

We have hence proved that j*() belongs to IG,ph, as stated in claim (i) of Theorem 2.12. Furthermore, Π:=im(κ) is clearly a finitely generated, free p[G]-submodule of Selp(AF) of maximal rank mn, and so the fact that the element j*() generates Fittp[G](cok(κ)) proves claim (iii) of Theorem 2.12. To complete the proof, it is enough to note that, since im(κ) is torsion-free, the canonical composite homomorphism

Шp(AF)(Selp(AF))torSelp(AF)cok(κ)

is injective and hence that one has that

Fittp[G](cok(κ))Annp[G](Шp(AF)).

Recalling finally that the Cassels–Tate pairing induces a canonical isomorphism between Шp(AF) and Шp(AFt) completes the proof of claim (ii) and thus of Theorem 2.12.

5 Examples

In this section we gather some evidence, mostly numerical, in support of conjecture Cp(A,[G]). Our aim is to verify statements that would not follow in an straightforward manner from the validity of the Birch and Swinnerton-Dyer conjecture for all intermediate fields of F/k. Because of the equivalence of statements (i)–(ii) in Theorem 2.8 we therefore choose not to focus on presenting evidence for conjecture Cp(A,) (although we also used our MAGMA programs to produce numerical evidence for Cp(A,) by verifying statement (iii) of Theorem 2.8).

Throughout this section A will always denote an elliptic curve.

5.1 Verifications of conjecture Cp(A,[G])

For the verification of Cp(A,[G]) using Theorem 2.9 it is necessary to have explicit knowledge of a map Φ that represents the extension class δA,F,p. Whenever A(F)p is not projective as a p[G]-module, we are currently not able to numerically compute Φ, so we only deal with examples in which A(F)p is projective. To the best of our knowledge there are currently three instances of theoretical evidence (in situations in which our fixed cyclic extension F/k is not trivial):

  1. In [4], it is shown that for each elliptic curve A/ with the property that L(A/,1)0 there are infinitely many primes p and for each such prime p infinitely many (cyclic) p-extensions F/ such that Cp(A,[Gal(F/)]) holds. All of these examples satisfy our hypotheses and are such that A(F)p vanishes.

  2. In [12, Theorem 1.1], conjecture Cp(A,[Gal(F/)]) is proved for certain elliptic curves A/, where F denotes the Hilbert p-classfield of an imaginary quadratic field k. This result combines with the functoriality properties of the eTNC to imply the validity of conjecture Cp(A,[Gal(F/k)]). In these examples one has that A(F)p is a free p[Gal(F/k)]-module of rank one.

  3. In [12, Corollary 6.2], certain S3-extensions F/K are considered. Let k and L denote the quadratic and cubic subfield of F/K respectively. Under certain additional assumptions it is then shown that the validity of the Birch and Swinnerton-Dyer conjecture for A over the fields k,K and L implies the validity of Cp(A,[Gal(F/K)]). Again by functoriality arguments, the validity of Cp(A,[Gal(F/k)]) follows. We note that the assumptions are such that one again has that A(F)p is a free p[Gal(F/k)]-module of rank one.

In the rest of this section we are concerned with numerical evidence. In [2, Section 6] there is a list of examples of elliptic curves A/ and dihedral extensions F/ of order 2p for which conjecture Cp(A,[Gal(F/)]) is numerically verified. Here the quadratic subfield k is real and A(F)p vanishes. Again by functoriality arguments we obtain examples where conjecture Cp(A,p[Gal(F/k)]) is numerically verified. There are two further analogous numerical verifications in dihedral examples in [12, Section 6.3], one of degree 10 and one of degree 14, both of them with the property that A(F)p is a free p[Gal(F/k)]-module of rank one.

In the following we fix an odd prime p and let q denote a prime such that q1(modp). Let F denote the unique subfield of (ζq)/ of degree p and take k to be . For p{3,5,7} and q<50 we went through the list of semistable elliptic curves of rank one and conductor N<200 and checked numerically whether L(A/,χ,1)=0 and L(A/,χ,1)0 for a non-trivial character χ of G, and in addition, whether our hypotheses are satisfied. This resulted in a list of 50 examples (27 for p=3, 20 for p=5 and 3 for p=7). In each of these examples we could find a point R such that A(F)p=p[G]R and numerically verify conjecture Cp(A,[G]).

We now describe in detail an example with [F:]=7. Let A be the elliptic curve

A:y2+xy+y=x3+x2-2x.

This is the curve 79a1 in Cremona’s notation. It is known that A() is free of rank one generated by P1=(0,0) and that Ш(A)=0. Moreover it satisfies the hypotheses used throughout the paper.

We take p=7 and let F be the unique subfield of (ζ29) of degree 7. Explicitly, F is the splitting field of

f(x)=x7+x6-12x5-7x4+28x3+14x2-9x+1

and we let α denote a root of f. Using the MAGMA command Points it is easy to find a point R of infinite order in A(F)A(),

R=(117(31α6+23α5-373α4-135α3+814α2+372α-86),
117(-35α6-83α5+380α4+771α3-811α2-1321α+232)).

By Proposition 2.2 we know that A(F)p is a permutation module, hence A(F)pp[G]. Furthermore, [11, Proposition 3.1] now implies that Шp(AF)=0.

We set Q1:=TrF/(R)=(34,-38) and easily verify that Q1=-4P1. We checked numerically that p[G]R=A(F)p.

Computing numerical approximations to the leading terms using Dokchitser’s MAGMA implementation of [15] we obtain the following vector for (χ*/λχ(𝒫,𝒫t))χG^:

(-0.077586206896551724152,
-0.49999999999999999992+2.1906431337674115362i,
-0.49999999999999999996+0.62698016883135191886i,
-0.49999999999999999998-0.24078730940376432202i,
-0.49999999999999999992-2.1906431337674115362i,
-0.49999999999999999996-0.62698016883135191886i,
-0.49999999999999999998+0.24078730940376432202i).

This is very close to

(-9116,ζ73+ζ72+ζ7,-ζ75-ζ74-ζ7-1,-ζ75-ζ73-ζ7-1,
-ζ73-ζ72-ζ7-1,ζ75+ζ74+ζ7,ζ75+ζ73+ζ7).

It is now easy to verify the rationality conjecture C(A,[G]) by the criterion of Theorem 2.6. Moreover, the valuations of -9/116 and ζ73+ζ72+ζ7 at 𝔭χ are 0, so that by Theorem 2.8 we deduce the validity of Cp(A,). Finally, one easily checks that

-9116ζ73+ζ72+ζ7(mod(1-ζ7)),

so that the element in (2.7) is actually a unit in p[G], thus (numerically) proving Cp(A,[G]).

5.2 Evidence in support of conjecture Cp(A,[G])

In this subsection we collect evidence for statements that we have shown to follow from the validity of Cp(A,[G]) and focus on situations in which A(F)p is not p[G]-projective. In particular, we aim to verify claim (i) of Theorem 2.12. Since we can neither compute the module Шp(AF) nor a map Φ as required, we are not able to verify any other claim of either Corollary 2.11 or Theorem 2.12.

Again we want to focus on evidence which goes beyond implications of the Birch and Swinnerton-Dyer conjecture for A over all intermediate fields of F/k. We assume the notation of Theorem 2.12, so in particular set h=t<nmt. If k= and mn=0, then the element is essentially the Mazur–Tate modular element (see [20]) and the validity of the Birch and Swinnerton-Dyer conjecture would imply that it belongs to IG,p in the type of situations under consideration. Hence, if mn=0, we only searched for examples where h>1.

If F/k is cyclic of order p, then IG,ph=(σ-1)hp for all h1. Letting u and δ denote the elements defined in the proof of Corollary 2.11, we hence note that, if Cp(A,) is valid, then u× and the proof of Corollary 2.11 clearly shows that =δu is contained in δp=IG,p. We therefore further restricted our search for interesting examples to cases where [F:k]=pn with n2.

Restricted by the complexity of the computations and the above considerations we are therefore lead to consider the following types of examples:

  1. A(F)ppm0p[G/H1]m1p[G]m2, [F:]=32, (m0,m1,m2)=(1,1,0),

  2. A(F)ppm0p[G/H1]m1p[G]m2, [F:]=32, (m0,m1,m2)=(m0,0,0), m02.

We note that, whenever Шp(AF) is trivial, the validity of Cp(A,[G]) implies via Corollary 2.11 that h is the exact order of vanishing, i.e., that

IG,phIG,ph+1.

However, this need not be true if Шp(AF) is non-trivial. In such cases, by Theorem 2.12 (iii), Cp(A,[G]) does predict that generates the Fitting ideal of Selp(AF) since mn=0 immediately implies Π=0.

Let q denote a prime such that q1(mod32). We let F denote the unique subfield of (ζq)/ of degree 9 and take k to be .

We checked two examples of type (i), namely those given by the pairs

(A,q){(681c1,19),(1070a1,19)}.

In both cases we were able to find a points P0 and P1 such that A(F)p=pP0p[G/H]P1, where H denotes the subgroup of order 3. Each time we numerically found that Шp(AF)=0 (predicted by the Birch and Swinnerton-Dyer conjecture for A over F) and verified that h is the precise order of vanishing, as predicted by Corollary 2.11.

Concerning examples of type (ii) went through the list of semistable elliptic curves of rank two and conductor N<750 and produced by numerically checking L-values and derivatives a list of twelve examples satisfying the necessary hypotheses. In each of these examples we had h=m0=2 and could numerically verify the containment IG,p2. Whenever Шp(AF) was trivial, we also checked that IG,p3.

We finally present one example in detail. Let A be the elliptic curve

A:y2+y=x3+x2-2x.

This is the curve 389a1 in Cremona’s notation. It is known that A() is free of rank two generated by P1=(0,0) and P2=(-1,1) and that Ш(A)=0. Moreover it satisfies the hypotheses required to apply Theorem 2.12 (see Remark 2.13).

Computing numerical approximations to the leading terms we find that the order of vanishing at each non-trivial character is 0. The rank part of the Birch and Swinnerton-Dyer conjecture therefore predicts that rk(A(F))=2. We checked that P1,P2 is 3-saturated in A(F) and therefore (conjecturally) conclude that A(F)p=P1,P2pp2.

The Birch and Swinnerton-Dyer conjecture predicts that |Шp(AF)|=81. We therefore cannot test for the precise order of vanishing.

Computing leading terms, periods and regulators we find the following numerical approximations to (χ*/λχ(𝒫,𝒫t))χG^:

(-1.243243,1.500000+2.598076i,1.500000-2.598076i,
0.358440+2.032818i,0.286988-0.104455i,-3.645429+3.058878i,
0.358440-2.032818i,0.286988+0.104455i,-3.645429-3.058878i).

The actual computation was done with a precision of 30 decimal digits. This is very close to

(-4637,3ζ3+3,-3ζ3,
2ζ93-ζ92+2ζ9,-ζ94-2ζ93+2ζ92-2,ζ95+2ζ94+2ζ93+ζ92,
-2ζ95+ζ94-2ζ93-2ζ92+ζ9-2,-ζ95-2ζ94+2ζ93-2ζ9,2ζ95-2ζ93-ζ9-2).

It is now easy to verify the rationality conjecture C(A,[G]) by the criterion of Theorem 2.6. We finally find that

=-σ+2σ2-σ3+2σ5-2σ6-2σ7+2σ8

and easily check that IG,p2.

Acknowledgements

We would like to thank David Burns and Christian Wuthrich for some helpful discussions concerning this project, and the referee for making several useful suggestions.

References

[1] Atiyah M. F. and Wall C. T. C., Cohomology of groups, Algebraic number theory, Academic Press, London (1967), 94–115. Suche in Google Scholar

[2] Bley W., Numerical evidence for the equivariant Birch and Swinnerton-Dyer conjecture, Exp. Math. 20 (2011), 426–456. 10.1080/10586458.2011.565259Suche in Google Scholar

[3] Bley W., Numerical evidence for the equivariant Birch and Swinnerton-Dyer conjecture (part II), Math. Comp. 81 (2012), 1681–1705. 10.1090/S0025-5718-2012-02572-5Suche in Google Scholar

[4] Bley W., The equivariant Tamagawa number conjecture and modular symbols, Math. Ann. 356 (2013), 179–190. 10.1007/s00208-012-0837-6Suche in Google Scholar

[5] Breuning M. and Burns D., Additivity of Euler characteristics in relative algebraic K-groups, Homology Homotopy Appl. 7 (2005), 11–36. 10.4310/HHA.2005.v7.n3.a2Suche in Google Scholar

[6] Burns D., Equivariant Whitehead torsion and refined Euler characteristics, Number theory (Montreal 2002), CRM Proc. Lecture Notes 36, American Mathematical Society, Providence (2004), 35–59. 10.1090/crmp/036/04Suche in Google Scholar

[7] Burns D., Leading terms and values of equivariant motivic L-functions, Pure Appl. Math. Q. 6 (2010), 83–172. 10.4310/PAMQ.2010.v6.n1.a4Suche in Google Scholar

[8] Burns D. and Flach M., Motivic L-functions and Galois module structures, Math. Ann. 305 (1996), 65–102. 10.1007/BF01444212Suche in Google Scholar

[9] Burns D. and Flach M., Tamagawa numbers for motives with (non-commutative) coefficients, Doc. Math. 6 (2001), 501–570. 10.4171/dm/113Suche in Google Scholar

[10] Burns D. and Macias Castillo D., Organising matrices for arithmetic complexes, Int. Math. Res. Not. IMRN 2014 (2014), no. 10, 2814–2883. Suche in Google Scholar

[11] Burns D., Macias Castillo D. and Wuthrich C., A note on Selmer structures, preprint 2014, http://www.mth.kcl.ac.uk/staff/dj_burns/selmerstructures.pdf. Suche in Google Scholar

[12] Burns D., Macias Castillo D. and Wuthrich C., On Mordell–Weil groups and congruences between derivatives of Hasse–Weil L-functions, preprint 2014, http://www.mth.kcl.ac.uk/staff/dj_burns/mordellweilgroupandcongruences.pdf. 10.1515/crelle-2014-0153Suche in Google Scholar

[13] Curtis C. W. and Reiner I., Methods of representation theory. Vol. I and II, John Wiley and Sons, New York 1987. Suche in Google Scholar

[14] Deligne P., Valeurs de fonctions L et périodes d’intégrales, Automorphic forms, representations and L-functions (Corvallis 1977), Proc. Sympos. Pure Math. 33 no. 2, American Mathematical Society, Providence (1979), 313–346. 10.1090/pspum/033.2/546622Suche in Google Scholar

[15] Dokchitser T., Computing special values of motivic L-functions, Exp. Math. 13 (2004), no. 2, 137–149. 10.1080/10586458.2004.10504528Suche in Google Scholar

[16] Fearnley J. and Kisilevsky H., Critical values of derivatives of twisted elliptic L-functions, Exp. Math. 19 (2010), 149–160. 10.1080/10586458.2010.10129069Suche in Google Scholar

[17] Fearnley J. and Kisilevsky H., Critical values of higher derivatives of twisted elliptic L-functions, Exp. Math. 21 (2012), 213–222. 10.1080/10586458.2012.676522Suche in Google Scholar

[18] Gross B. H., On the conjecture of Birch and Swinnerton-Dyer for elliptic curves with complex multiplication, Number theory related to Fermat’s last theorem (Cambridge, MA, 1981), Progr. Math. 26, Birkhäuser-Verlag, Boston (1982), 219–236. 10.1007/978-1-4899-6699-5_14Suche in Google Scholar

[19] Martinet J., Character theory and Artin L-functions, Algebraic number fields, Academic Press, London (1977), 1–87. Suche in Google Scholar

[20] Mazur B. and Tate J., Refined conjectures of the Birch and Swinnerton-Dyer type, Duke Math. J. 54 (1987), 711–750. 10.1215/S0012-7094-87-05431-7Suche in Google Scholar

[21] Swan R. G., Algebraic K-theory, Springer-Verlag, New York 1978. Suche in Google Scholar

[22] Yakovlev A. V., Homological definability of p-adic representations of a ring with power basis (Russian), Izv. Akad. Nauk SSSR Ser. Math. 34 (1970), 321–342. Suche in Google Scholar

Received: 2013-6-11
Revised: 2014-5-20
Published Online: 2014-10-7
Published in Print: 2017-1-1

© 2017 by De Gruyter

Heruntergeladen am 30.9.2025 von https://www.degruyterbrill.com/document/doi/10.1515/crelle-2014-0081/html
Button zum nach oben scrollen