On Mordell-Weil groups
and congruences between derivatives
of twisted Hasse-Weil -functions
Abstract.
Let be an abelian variety defined over a number field and let be a finite Galois extension of . Let be a prime number. Then under certain not-too-stringent conditions on and we compute explicitly the algebraic part of the -component of the equivariant Tamagawa number of the pair . By comparing the result of this computation with the theorem of Gross and Zagier we are able to give the first verification of the -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 . More generally, our approach leads us to the formulation of certain precise families of conjectural -adic congruences between the values at of derivatives of the Hasse-Weil -functions associated to twists of , 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 be an abelian variety defined over a number field . Then for any finite Galois extension of with group the equivariant Tamagawa number conjecture for the pair is formulated in [16, Conjecture 4(iv)] as an equality in a relative algebraic -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 over each of the intermediate fields of .
The refined conjecture also naturally decomposes into ‘components’, one for each rational prime , in a way that will be made precise in §5.1 below, and each such -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 has good ordinary reduction at , 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 with respect to any compact -adic Lie extension of that contains .
If does not divide the order of , then, without any hypothesis on the reduction of at , 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 divides the order of , 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 has strictly positive rank over and the Galois group is both non-abelian and of order divisible by .
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 be an odd prime, an imaginary quadratic field in which is unramified and its Hilbert -classfield. Write for the generalised dihedral group .
Let be an elliptic curve over which satisfies, with respect to the field , the hypotheses (a), (b), (c), (e) and (f) that are listed in §3 and for which the Tate-Shafarevich group is finite.
Assume in addition that each of the following four hypotheses is satisfied.
-
(i)
all primes of bad reduction for split in ;
-
(ii)
the Galois group of is isomorphic to ;
-
(iii)
the Hasse-Weil -function has a simple zero at ;
-
(iv)
the trace to of the Heegner point is not divisible by .
Then the -component of the equivariant Tamagawa number conjecture is valid for the pair .
Remark 1.2.
For a fixed elliptic curve over and imaginary quadratic field , 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 . In addition, it will be clear that for a fixed and the hypotheses (e) and (f) only constitute a mild restriction on the ramification of the extension . Finally we recall that, by a famous result of Serre [46], the hypothesis (ii) holds for almost all if does not have complex multiplication.
Remark 1.3.
In the context of Theorem 1.1 we also note that for any natural number there are infinitely many imaginary quadratic fields in which does not ramify and for which the group has exponent divisible by . (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 and a Galois extension of degree divisible by is given by Bley in [5, Corollary 1.4 and Remark 4.1] where it is assumed, amongst other things, that is an elliptic curve, is an abelian extension of of exponent and, critically, that the rank of over 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 divides the degree of the relevant extension. More precisely, we show that for certain elliptic curves the validity of eTNCp with follows from that of the relevant cases of the Birch and Swinnerton-Dyer conjecture for a family of -extensions of number fields (see Corollary 6.2) and provide examples (with and ) 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 over a number field , a finite Galois extension of with Galois group and an odd prime .
In §4 we shall first compute, under some not-too-stringent conditions on and , the ‘algebraic part’ of the -component of the equivariant Tamagawa number of the pair . 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 of the -adic Tate module of the dual of together with the Néron-Tate height of relative to the field . 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 -adic congruence relations between the values at of higher derivatives of the Hasse-Weil -functions of twists of by irreducible complex characters of , 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 -primary Tate-Shafarevich of over by elements which interpolate the (suitably normalised) values at of higher derivatives of twisted Hasse-Weil -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 the necessary -adic congruence relations can be made very explicit even in the case that has strictly positive rank over (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 we use the following notation. We write for the set of irreducible -valued characters of , where denotes either or (and the intended meaning will always be clear from the context). We also write for the trivial character of and for the contragredient of each in . For each in we write
for the primitive idempotent of the centre of the group ring . For each -valued character we also fix an -module of character .
For any abelian group we write for its torsion subgroup and for the quotient , which we often regard as a subgroup of . For any prime we write for the subgroup of the Sylow -subgroup of . We also set and write for the pro--completion . If is finitely generated, then we set .
For any -module we write for the Pontryagin dual and for the linear dual , each endowed with the natural contragredient action of . If is finitely generated, then for any field extension of we set .
For any Galois extension of fields we write in place of . We also fix an algebraic closure of and abbreviate to . For each non-archimedean place of a number field we write for its residue field.
Throughout this paper, we will consider the following situation. We have a fixed odd prime and a Galois extension of number fields with Galois group . We choose a -Sylow subgroup in and set . We give ourselves an abelian variety of dimension defined over . We write for the dual abelian variety of .
For each intermediate field of we write , and for the set of non-archimedean places of that are -adic, which ramify in and at which has bad reduction respectively. Similarly, we write , and for the set of archimedean, real and complex places of respectively. The notation will stand for when “?” is b, r, , , or .
For each such field we also set and write and for the -primary Tate-Shafarevich and Selmer groups of . We recall that there exists a canonical exact sequence of the form
(1) |
3. The hypotheses
Fix an odd prime , number fields and and an abelian variety as described above.
Then throughout this article we will find it convenient to assume that this data satisfies the following hypotheses:
-
(a)
and ;
-
(b)
The Tamagawa number of at each place in is not divisible by ;
-
(c)
has good reduction at all -adic places;
-
(d)
For all -adic places that ramify in , the reduction is ordinary and ;
-
(e)
No place of bad reduction for is ramified in , i.e. .
-
(f)
For any place in such that the primes above it in are ramified in , we have ;
-
(g)
The Tate-Shafarevich group is finite.
Remark 3.1.
For a fixed abelian variety over and extension the hypotheses (a), (b) and (c) are clearly satisfied by all but finitely many primes (which do not divide the degree of ), the hypotheses (e) and (f) constitute a mild restriction on the ramification of 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 , there may be infinitely many primes for which there are -adic places at which has good ordinary reduction but does not vanish. Nevertheless, it is straightforward to describe examples of abelian varieties 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 -adic Tate module of rather than that of itself. Applications of the hypotheses (b)–(f) in this article will therefore often take place with replaced by . 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 replaced by 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 and 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 of the -adic Tate module of the dual abelian variety of and the Néron-Tate height of relative to the field .
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 we write for the derived category of complexes of (left) -modules. We also write for the full triangulated subcategory of comprising complexes that are ‘perfect’ (that is, isomorphic in to a bounded complex of finitely generated projective -modules).
We assume to be given a complex in which is acyclic outside degrees and for any given integer and such that is -free. We also assume to be given an isomorphism of -modules . Then, under these hypotheses, it can be shown that there exists an isomorphism in where is a complex of -modules of the form where is finitely generated and projective and the first term occurs in degree . We then consider the following auxiliary composite isomorphism of -modules
where the first and third maps are obtained by choosing -equivariant splittings of the tautological surjections and respectively and the second is .
We now write for the relative algebraic -group of the inclusion and recall that this group is generated by elements of the form where and are finitely generated projective -modules and is an isomorphism of -modules .
In particular, in terms of this description, it can be shown that the element of obtained by setting
is independent of the choice of isomorphism (and hence of the module ) and of the splittings made when defining the isomorphism . This element will be referred to as the ‘refined Euler characteristic’ of the pair in the sequel.
4.2. Cohomology with finite support
We now turn to consider the finite support cohomology complex of the -adic Tate module of .
To do this we write for the set of -embeddings and for the module , endowed with the obvious action of .
Then the -adic Tate module of the base change of through is equal to
where acts naturally on the first factor and acts diagonally. We set .
For each place of we set and write for the inertia subgroup of . We also fix a place of above and a corresponding embedding and write and for the images of and under the induced homomorphism .
For each non-archimedean place we then define a complex of -modules by setting
where, if , the first occurrence of is placed in degree zero and is the natural Frobenius in and, if , then denotes the full pre-image in of the ‘finite part’ subspace of that is defined by Bloch and Kato in [8, (3.7.2)].
We next define to be a complex of -modules which lies in an exact triangle in of the form
(2) |
where is the natural inflation morphism if , and is induced by the inclusion and the fact that is acyclic in degrees less than one, if . We finally set , write for the subring of comprising elements that are integral at all non-archimedean places outside and define the complex so that it lies in an exact triangle in of the form
(3) |
Here each -component of is equal to the composite of the natural localisation morphism in étale cohomology and the morphism .
In the sequel we also set
in each degree .
Lemma 4.1.
- (i)
- (ii)
Proof.
First, since is odd and is a (finitely generated) free -module the complexes and for each place of belong to . In view of the exact triangles (2) and (3), claim (i) will therefore follow if we can show that for each in the complex belongs to , or equivalently (as is a Sylow -subgroup) that it belongs to .
We first consider the case . In this case is isomorphic in to where we write for a place in below and set .
In particular, even without assuming our hypotheses, if the ramification index of in is not divisible by , then is unramified in so is a free -module and so belongs to .
Next we show that hypotheses (e) and (f) imply that if ramifies in and , then (and therefore also ) is acyclic. We write for the differential in degree of . Then it is clear that vanishes and so it suffices to prove vanishes or equivalently, since and are finite -groups, that vanishes. But (e) implies and so the -module is isomorphic to the cokernel of the action of on and hence of cardinality the maximum power of that divides in . The module therefore vanishes since the latter determinant is equal to and this is a unit at by (f).
Finally we fix a -adic place of and recall from [8, after (3.2)] that the group is the image in of under the natural (injective) Kummer map. Since the group is finitely generated over it is thus enough to show that each module is cohomologically-trivial over .
This is clear if the order of is prime to and true in any other case provided that is cohomologically-trivial over each subgroup of that has order . The point here is that, for a given -Sylow subgroup of , there exists a normal subgroup of which has order and hence also a Hochschild-Serre spectral sequence in Tate cohomology . Thus if is a cohomologically-trivial -module, then is trivial for all integers , and then [12, Chapter VI, (8.7) and (8.8)] combine to imply that is a cohomologically-trivial -module.
We hence fix a subgroup of of order . Now cohomology over is periodic of order and is isomorphic (via the formal group logarithm) to the free -module , so [2, Corollary to Proposition 11] implies that the Herbrand quotient is equal to 1.
It is thus enough to prove that the group vanishes, or equivalently that the natural norm map is surjective. But the hypotheses (c) and (d) imply that has good reduction at all -adic places and further that at any -adic place of which ramifies in the reduction is ordinary and such that has no -torsion, and so the argument of Mazur in [38, §4] implies that the norm map is surjective, as required to complete the proof of claim (i). (We note in passing that if is unramified in , 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 vanishes if , that is isomorphic to and that there is a canonical exact sequence of -modules
Now hypothesis (a) implies that the -group acts on with a single fixed point and hence that vanishes. Claim (ii) will therefore follow if we can show that, under the stated hypotheses, the group vanishes for all . In view of the natural isomorphism
it is thus enough to show that vanishes for all . On the one hand, if , then is torsion-free and so the result is clear. On the other hand, if , then (e) implies that so and hence the group
vanishes as a consequence of (b) and the fact that has cardinality equal to the maximal power of that divides the Tamagawa number of at . Since is a -group, this implies that vanishes, as required to complete the proof of claim (ii). ∎
In the sequel we assume that , and satisfy the hypotheses (a)–(g). Then Lemma 4.1 implies that the complex
belongs to , is acyclic outside degrees one and two and is such that for each isomorphism of fields there exists a canonical composite isomorphism of -modules of the form
in which the central isomorphism is induced by the -linear extension to of the Néron-Tate height of relative to the field .
Hypothesis (a) implies that the -group acts on with a single fixed point and hence that the module is -free and so the above observations imply that the construction in §4.1 applies to the pair to give a canonical Euler characteristic
in the relative algebraic -group .
This element encodes detailed information about a range of aspects of the arithmetic of 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 under the boundary homomorphism .
4.3. The element .
For each isomorphism as above there is an induced composite homomorphism of abelian groups
(4) |
(where the first and third arrows are induced by the inclusions and respectively).
For each such we set
where is the ‘algebraic part’ of the equivariant Tamagawa number for the pair , as defined (unconditionally under the assumed validity of hypothesis (g)) in [16, §3.4].
In order to relate this element to the Euler characteristic defined above we must first introduce an auxiliary element of .
To do this we define a complex
where, as before, denotes . It is then clear that is acyclic outside degrees zero and one, that and that there is a canonical identification of with , where we set .
We next use this description of the cohomology of to define a canonical isomorphism of -modules
For this we write for the set of embeddings and , for each in , for the set of embeddings that extend . For in we fix a corresponding element of and consider the -linear map
that sends the image in of a cycle to the map induced by sending a differential to .
For each place in we also write for the module , endowed with its natural action of the direct product , and then define to be the following composite isomorphism of -modules
Here the first isomorphism is induced by the canonical comparison isomorphisms
(5) |
the second by the maps and the canonical decomposition
(6) | ||||
and the third by the sum over places in of the classical exponential map
(7) |
where we set .
Under our stated hypotheses the complex belongs to . It is also acyclic outside degrees zero and one and such that is -free and so we may use the construction of §4.1 to define an element of by setting
We are now ready to state the main result of this section. In this result, and the sequel, we shall use the composite homomorphism
where the first arrow is the inverse of the (bijective) reduced norm homomorphism and the second is the standard boundary homomorphism .
Proposition 4.2.
Proof.
To discuss we first recall relevant facts concerning the formalism of virtual objects introduced by Deligne in [24] (for more details see [16, §2]).
With denoting either or we write for the category of virtual objects over and for the object of associated to each in . Then is a Picard category with naturally isomorphic to (see [16, (2)]) and we write for its product and for the unit object . Writing for the Picard category with unique object and we use the isomorphism of abelian groups
(8) |
that is described in [16, Proposition 2.5]. In particular, via this isomorphism, each pair comprising an object of and a morphism gives rise to a canonical element of .
Now if belongs to and is acyclic outside degrees and (for any integer ), then any isomorphism of -modules gives a canonical morphism in . The associated element of coincides with the Euler characteristic defined in [11, Definition 5.5]. In particular, from Proposition 5.6(3) in loc. cit. it follows that , whilst Theorem 6.2 and Lemma 6.3 in loc. cit. combine to imply that if is -free, then
(9) |
where is the explicit Euler characteristic discussed in §4.1.
We now set
which is the complex of the compactly supported étale cohomology of on , regarded as a complex of -modules in the natural way.
Then a direct comparison of the definitions of and shows that there is a canonical exact triangle in
(10) |
We consider the following diagram in
(11) |
Here is the morphism induced by the scalar extension of (10), denotes the canonical identification and the morphisms and are defined by the condition that the two squares commute.
We claim that this definition of and implies that
(12) |
where the morphisms and 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 induced by the central column of the diagram [16, (26)] and then simply compare the explicit definitions of the morphisms and and of and . After this it only remains to note the following fact. For each place , respectively , we write and for and , respectively and , and then for the complex , with the first term placed in degree zero. Our definition of implicitly uses the morphism induced by the acyclicity of whereas the definition of uses (via [16, (19) and (22)]) the morphism induced by the identity map on ; the occurrence of the morphism in the equality (12) is thus accounted for by applying the remark made immediately after [16, (24)] to each of the complexes and noting that .
4.4. The local term
In this section we explicitly compute the Euler characteristic 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 . In addition, for each natural number we will write for the set of integers which satisfy .
4.4.1. Periods and Galois Gauss sums.
We fix Néron models for over and for over for each in , and then fix a -basis of the space of invariant differentials which gives -bases of for each in and is also such that each extends to an element of .
For each place in we fix -bases and of and , where denotes complex conjugation, and then set
where in both matrices runs over . For each in we fix a -basis of and set
where runs over . (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 and , we set and . Then we define the periods
and and finally set
For each in , respectively. in , we also define to be equal to , respectively. , and then set and
To describe the relevant Galois Gauss sums we first define for each non-archimedean place of the ‘non-ramified characteristic’ to be the image under the natural induction map of the element of .
For each character in we then define a modified equivariant global Galois-Gauss sum by setting
where the individual Galois Gauss sums are as defined by Martinet in [37].
4.4.2. Computation of the local Euler characteristic.
The explicit computation of the Euler characteristic is made considerably more difficult by the presence of -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)
is unramified in .
A full treatment of the terms in the general (ramified) case will then be given in a future article.
In the following result we use the elements of that are defined in Proposition 4.2. We will also write for the isomorphism that is induced by . We finally recall that denotes the dimension of .
Proof.
We set for each in and then also and . Then, as is odd and is acyclic for all in (by Lemma 4.1(i)), the comparison isomorphisms (5) induce a natural isomorphism
in and hence an equality in
(13) | ||||
Here we set and the second equality follows by combining for each in the explicit definition of the term together with the facts that there is an exact sequence of -modules
and that is a projective -module (as a consequence of hypothesis (e)).
We next note that , where is the second isomorphism that occurs in the definition of and where each is induced by the classical exponential map (7).
For each in we set . Then hypothesis (h) implies that each such place is unramified in and hence, by Noether’s Theorem, that the -module is free. In particular, the -module
is free and so in one has an equality
Combining this equality with (13) and the explicit computations in Lemmas 4.4 and 4.5 below one finds that is equal to
Here we have set , endowed with its natural action of , and written for the natural isomorphism. To deduce the claimed result from this expression it suffices to show is equal to and, since no -adic place of ramifies in , this follows directly from equation (34), Proposition 7.1 and Corollary 7.6 in [6]. ∎
Lemma 4.4.
If at each place in the extension is unramified and has good reduction, then one has
Proof.
We fix in and write for the Néron model of over . Then is naturally isomorphic to the free -module .
In addition, the stated hypotheses on imply that there exists a full free -submodule
of , where 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
maps to (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 -modules (with vertical differentials)
Here the term in the first complex occurs in degree and is the tautological projection. Further, the exponential map of Bloch and Kato maps the cohomology in degree of the second complex bijectively to and the differential 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 -modules , in which (following [8, Example 3.11]) the second arrow is equal to . This sequence then implies that the term is equal to
and by summing over all places in this implies the claimed equality. ∎
For each we fix an embedding .
Lemma 4.5.
In one has
Proof.
For each place in we define in via the equality . For each integer in we then define an element
and we note that, since is odd, the set is a -basis of .
For in and in we set and note that is a -basis of .
We next set and fix an -generator of and a -basis of and set with the element of that is dual to . Then the set is a -basis of .
The key to our argument is to compute the matrix of with respect to the bases and of and . To do this we find it convenient to introduce an auxiliary basis. Thus, for in we set with , and for in we set with if and if . Then is an -basis of and for each index one has
with and if and if . This formula implies that the matrix of with respect to the bases and is where is a diagonal block matrix with blocks
and is the matrix in that represents the isomorphism (6) with respect to the bases and so that
Writing for the class of in one therefore has
(14) |
Now is equal to the product for the block matrices and and so is equal to the product
In addition, is the matrix of the natural isomorphism with respect to the -bases and of and and so
(15) |
Next we note that and that for each in also with
Now if is real, then and with and in , whilst if is complex, then with in . Thus, since each of the terms
if is real, and if is complex, belong to the kernel of , we conclude that is equal to
This formula combines with (14) and (15) to give the claimed formula for the term . ∎
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 of and each character in the -truncated -twisted Hasse-Weil -function of has an analytic continuation to where it has a zero of order .
Here acts diagonally on the tensor product and so is equal to the multiplicity with which , and hence also , appears in the representation .
For each in we then write for the the coefficient of in the Taylor expansion at of .
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 as varies over . We next discuss several consequences of this reinterpretation and then, motivated by the results of [18], we specialise to the case that the (-completed) Mordell-Weil group is a projective -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 and satisfy the hypotheses (g) and (i), then [16, Conjecture 4] for the pair asserts the validity of an equality in . In such a case we shall say that the ‘eTNCp for is valid’ if for every isomorphism of fields the predicted equality is valid after projection under the homomorphism defined in (4).
If does not divide , then the algebra is regular and it is straightforward to use the techniques described in [15, §1.7] to give an explicit interpretation of these projections. If divides , 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 even if is non-abelian, divides and the rank of is strictly positive.
For each non-archimedean place of , we decompose the non-ramified characteristic as with each in . For each in we then define a modified global Galois-Gauss sum by setting
(16) |
with .
Theorem 5.1.
Proof.
Since for all in the element clearly belongs to .
We next recall that [16, Conjecture 4(iv)] for the pair asserts
where is the ‘extended boundary homomorphism’ defined in [16, Lemma 9] and is the leading term at of the -valued -function defined in [16, §4.1].
Noting that for each isomorphism one has , and recalling the result of Proposition 4.2, it follows that the eTNCp for is valid if and only if for every isomorphism one has
Now hypotheses (e) and (h) combine to imply that and for each place in this set the term is equal to the value at of the Euler factor at that occurs in the definition of (by [16, Remark 7]). In view of the formula for that is given in Theorem 4.3, one therefore finds that the above equality is valid if and only if
It now suffices to show that the quotient on the left hand side of this equality is equal to and this follows from a straightforward comparison of all of the terms involved and then noting that the definition of the truncated -function implies . ∎
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 range of detailed information about the arithmetic of over . 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 -modules of the form
where is a trivial source -module, is a projective -module and induces, upon passage to -coinvariants, an isomorphism (via the relevant canonical short exact sequence of the form (1)). The notion of ‘trivial source -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 for the subset of comprising characters for which the homomorphism is bijective and then define an idempotent in by setting .
Given a natural number and a matrix in we write for the corresponding ‘generalised adjoint matrix’ in that is defined by Johnston and Nickel in [34, §3.6]. We then write for the -submodule of given by the set
For more details about this module see Remark 5.4 below.
Proposition 5.2.
Assume that and satisfy the hypotheses (a)–(i) and that the equality (17) is valid. Fix a surjective homomorphism of -modules
as above and an element of . For each homomorphism of -modules and each isomorphism of fields set
Then for each such and the element belongs to and annihilates both modules and . In particular, the element
is such that belongs to and annihilates .
Proof.
Hypothesis (a) implies that the module is -free. Lemma 4.1 therefore implies, in the terminology of [17], that is an admissible complex of -modules and equation (17) asserts that is a characteristic element for the pair . The first claim is therefore a direct consequence of [17, Corollary 2.13] and the isomorphisms
where the first isomorphism is induced by the Cassels-Tate pairing, the second follows from the short exact sequence (1) (with ) 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
is a homomorphism of the required type. Further, in this case one has and so hypothesis (i) implies is equal to and hence that . It therefore suffices to show that and this is true because in this case the space vanishes. ∎
Remark 5.3.
The conjectural equality (17) implies, via Proposition 5.2, a family of explicit congruence relations between the complex numbers that are defined by the equalities as varies over . This is because
and this sum belongs to if and only if the elements satisfy all of the following conditions:
-
(i)
for all ;
-
(ii)
for all and ;
-
(iii)
for all .
Remark 5.4.
Ideals of the form were introduced by Nickel in [41] and have been computed extensively by Johnston and Nickel in [34]. For example, if , then so that and it is shown in loc. cit. that this inclusion is an equality if and only if the order of the commutator subgroup of is not divisible by . More generally, for each in the matrix belongs to for any maximal order in that contains (cf. [41, Lemma. 4.1]) and so Jacobinski’s description in [33] of the central conductor of in implies, for example, that for any -valued character of the element belongs to where is the subring of that is generated over by the values of . This gives an easy ‘lower bound’ on (but which is, in most cases, not best possible).
Remark 5.5.
For the explicit computation of terms of the form and occurring in Theorem 5.1 and Proposition 5.2 respectively in the case that is cyclic of -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 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 and are both projective -modules. (In this regard, note that if is principally polarised, then the -modules and are isomorphic.) This corresponds to taking the module in Proposition 5.2 to be equal to and the more general case that is a trivial source -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.
Then the projective dimension of the -module is at most one and there exists an isomorphism of -modules.
For each now set
Then the equality (17) is valid if and only if the (non-commutative) Fitting invariant of the -module is generated by the set of elements of the form where is in and
Proof.
The assumed projectivity of implies that its -linear dual is also a projective -module. The existence of an isomorphism therefore follows from Swan’s Theorem [22, Theorem 32.1] and the fact that induces an isomorphism of -modules .
We now write for the complex , where the first term occurs in degree one. Then, since is a projective -module, we may choose a -equivariant section to the natural surjection
Since the kernel of this surjection is isomorphic (via the Cassels-Tate pairing) to , such a section induces a short exact sequence of -modules
and hence also an exact triangle in of the form
In particular, since belongs to , the projective dimension of the finite -module 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 and in all degrees , then this triangle combines with the additivity criterion of [11, Corollary 6.6] to imply that the element is equal to
Here the first displayed equality is valid as and the second because the difference is represented by the triple .
The equality (17) is therefore valid if and only if one has
In addition, one has for each in and so
Given this, a straightforward exercise (comparing the explicit definitions of refined Euler characteristic and non-commutative Fitting invariants) shows that the above formula for is valid if and only if is generated by the set of elements with in , as claimed. ∎
5.3. Dihedral congruences for elliptic curves
We now investigate the criterion of Proposition 5.6 in the case that is an elliptic curve (so ) and is dihedral (in the sense of Mazur and Rubin [40]).
Thus, as before, we have an odd prime and a Galois extension of group with -Sylow subgroup and we assume that is an abelian (normal) subgroup of of index two and that the conjugation action of any lift to of the generator of inverts elements of . In particular, the degree of is equal to for some and is a quadratic extension. We fix an element of order in . We set and write for the unique non-trivial linear character of .
In the following result, we will be interested in the case when has rank one over . In this case we write for the unique linear character which does not occur in the -module . Hence if and otherwise.
We also set .
Proposition 5.7.
Assume that the elliptic curve , odd prime and dihedral extension satisfy the hypotheses (a)–(e) and (g), that all places above split in , that and that .
Then and there is a point in with which generates a -module that is isomorphic to and has finite, prime-to-, index in .
In particular, one has and for all .
Proof.
Under the stated hypotheses, [18, Corollary 2.10(ii)] implies that vanishes, that is a projective -module and that the multiplicity with which each in occurs in the representation 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 -modules
where the group is both finite and of order prime to .
Since the group is also finite of order prime to (by hypothesis (a)) it follows that any point of whose projection in is equal to the image of multiplied by a large enough power of 2 has the properties described above.
The above description of also implies that the -module is isomorphic to and the claimed formulas for are then easily verified by explicit computation. ∎
For each subgroup of and character in we set
For any point of and any we then define a non-zero complex number
where is the -linear extension of the Néron-Tate height on , defined relative to the field .
In the next result we assume the hypotheses of Proposition 5.7 to be satisfied and fix a point as in that result. For each we then obtain a non-zero complex number by setting
(18) |
where the quantity is as in (16). Further we define
for all and
Here denotes the absolute discriminant of a number field and for any finite dimensional complex character of we write 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 is equal to the value for and to the first derivative for in .
Theorem 5.8.
Fix an odd prime , a dihedral extension of degree and an elliptic curve over . Assume that and satisfy the hypotheses (a)–(h), that all places above split in , that and that . Fix a point in as given by Proposition 5.7.
Then the equality (17) is valid if and only if the following conditions are satisfied.
-
(i)
For each in the number defined above belongs to , is a unit at all primes above and satisfies for all in .
-
(ii)
For all one has a congruence
(19) where denotes the localisation of at .
Remark 5.9.
Remark 5.10.
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 (by [18, Corollary 2.10(ii)]), the criterion of Proposition 5.6 is equivalent to asking that the element belongs to . 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 belongs to .
To investigate the element we fix a maximal -order in that contains . Then belongs to the subgroup of if and only if 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 of belongs to if and only if the congruences in (19) are valid with each term replaced by and with replaced by . But, from the results of Lemma 5.12(i) and (ii) below, one has if and only if , 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 -modules such that is equal to the element defined above.
Proof.
We claim first that for each there is an equality
(21) |
where and with denoting the subset of comprising places which split in .
To prove this we note that for each in one has
(22) |
where the first equality follows straight from the definition (16) and the second from the result of [37, Theorem 8.1(iii)].
Now if , then and for all in so , whilst it is clear that , and so in this case the equality (21) is an immediate consequence of (22).
Next we note that and hence that [37, Theorem 8.1(iii)] implies
where is the set of real places of that ramify in and the absolute norm of the different of . One then obtains the claimed equality (21) by substituting this formula for into (22) and then using that fact that whilst since is equal to for and to for .
Finally we assume that with Then, just as above, one finds that
where is the Artin root number of and the infinite part of the Artin root number of and the final equality follows from the very definition of . Now since no archimedean place ramifies in . In addition, the inductivity of Artin root numbers implies and so, since is an orthogonal character, the main result of Fröhlich and Queyrut in [29] implies that . Thus, upon substituting the last displayed expression into (22), and noting that , one obtains the claimed equality (21) in this case.
Having proved (21), we now write for the generator of the -module that sends to and to zero for each non-trivial element of . We then define to be the isomorphism of -modules that sends to .
One computes that , where is the resolvent element in , and also that for all .
Proposition 5.7 implies so and also that for each , the complex vector space has dimension one and hence
Now, for any and one has where we write for the -linear involution on that inverts elements of . In addition, for each one has
because is a multiple of only when or . The first sum here is always equal to , while the second is equal to for , to for and to otherwise. Hence, writing for , we find that
Upon substituting this equality and (21) into the definition of one obtains the element , as required. ∎
In the next result we write for the valuation ring of a finite extension of .
Lemma 5.12.
Write for the integral closure of in .
-
(i)
The following diagram commutes
Here is the homomorphism that sends an element in to , next and are the natural scalar extension homomorphisms, and the natural restriction homomorphisms and and the restrictions of the connecting homomorphisms and .
-
(ii)
The homomorphism is injective.
-
(iii)
Fix inside . Then belongs to if and only if belongs to for all in .
Proof.
To prove claim (i) we fix a set of representatives of the orbits of the action of on and abbreviate the functor to .
The commutativity of the lower square of the diagram follows from the naturality of the long exact sequences of relative -theory. To consider the upper square we use the -algebra isomorphisms
and where , with a representation of character , and for each and . Taken together these isomorphisms induce a diagram
in which the left and right hand composite vertical arrows are equal to and and is defined to make the diagram commute. To complete the proof of claim (i) it thus suffices to show sends each element to 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 -elementary if it is isomorphic to a group where is an -group for some prime that is coprime to and the image of the homomorphism belongs to the decomposition subgroup of . This means that a subgroup of is -elementary if it is either a subgroup of or of the form where is a cyclic subgroup of and an element of order . We consider the exact commutative diagram
where the limits are over all -elementary subgroups of . The transition maps are the homomorphisms induced by inclusions and by maps of the form for and all vertical arrows are the natural restriction maps. By a theorem of Dress [27] (see also [43, Theorem 11.2]), the maps and are bijective and hence is injective. Now is trivial whenever has order prime to and in [10, Proposition 3.2.(2)] Breuning has shown that the restriction map is injective for any subgroup of the form . The map therefore restricts to give an injective homomorphism
where runs over all subgroups of , as required to prove claim (ii).
Claim (iii) follows from the equalities
and the fact that . ∎
6. Special cases
In this section we use the criteria of Theorem 5.8 to give both theoretical and numerical verifications of the -part of the equivariant Tamagawa number conjecture for pairs where is an elliptic curve for which has strictly positive rank and is both non-abelian and of order divisible by .
We believe that, apart from the recent results of Bley in [5], where the group is assumed to be trivial and the field to be an abelian extension of of exponent , these results constitute the first verifications of the -part of the equivariant Tamagawa number conjecture for any elliptic curve and any Galois extension of degree divisible by .
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 . In particular one knows that the -primary part of the Birch and Swinnerton-Dyer conjecture holds for . 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 of smallest degree. We also denote by the Manin constant of and write for the trace in of the Heegner point that is defined over the Hilbert class field of .
Set and is the number of connected components of . 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 one has
(23) | ||||
where is the linear character of appearing in and denotes the archimedean place of . The second equality here follows from the equality and the fact that (since the height pairing is -invariant).
For each in we set
(24) |
where the quantities and are as defined just before Theorem 5.8. We also write
(25) |
for the correction term that accounts for the -truncation in the leading terms for each . Then the non-zero complex number is as in Theorem 5.8 using our Heegner point . By using (23), and the fact that for all of dimension two (as is unramified), one then finds that
(26) | and | |||||
if for . |
Our hypotheses imply that has rank one and vanishes by Kolyvagin [31] and hence all of the hypotheses of Theorem 5.8 are satisfied except for the requirement that splits in . However, the sole purpose of the latter hypothesis is to ensure (via the proof of Proposition 5.7) that the -module has a free rank one direct summand and so, since (23) implies that the point 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 satisfy both the conditions of Theorem 5.8(i) and the congruences (19).
Since the -part of the Birch and Swinnerton-Dyer conjecture holds for , the hypotheses (a) and (b) and the known vanishing of combine to imply that and are both -units. The hypothesis (c) implies that has good reduction at and so the hypothesis (ii) combines with [1, Theorem 2.7] to imply that the Manin constant is a -unit. Using also the fact that is unramified, we find that is a -unit too.
We proceed to compute the values of and . Since is unramified, we have that each ramified place in has ramification index and does not split in . It is easy to see that and that for all of dimension two. Next, for any in , we have
and since no place of bad reduction is allowed to ramify by hypothesis (e), one has . Hence and for all of dimension two. Moreover, this last value is equal to , which is a -unit by hypothesis (f). Therefore is a -unit for all .
Next we note that for any of dimension two as we have shown that . Since each element also belongs to , this formula makes it clear that they satisfy the condition of Theorem 5.8(i). In addition, it shows that for any one has
Finally we note that the last sum in the above expression is equal to if and to otherwise. Hence the expression is in both cases congruent to modulo , as required to complete the proof of Theorem 1.1.
6.2. -extensions
In this section we investigate the case that is a cyclic extension of degree and hence 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 is a non-abelian group of order six. In this case comprises , and for a fixed in . We fix a subfield of that is of degree over and set .
6.2.1. An arithmetic congruence
For each character in we define
(27) |
Here the quantities are as defined just before Theorem 5.8 and we have set
(28) |
with denoting the regulator of on the Mordell-Weil group for any field . 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 . In fact, BSD predicts that each expression is a non-zero rational number.
In the following result we include the assumed analytic continuation of the -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 be an elliptic curve over and assume no place at which has bad reduction ramifies in . Then if the Birch and Swinnerton-Dyer conjecture holds for , and one has a congruence modulo non-zero rational squares
(29) |
Proof.
Fix a finite field extension of . If is a (finite or infinite) place of , write for the number of connected components of . For a finite place in , also write for a Néron differential of . Then BSD for the field asserts that
(30) |
where we write for the period of over , as defined in §4.4.1 with respect to a fixed invariant differential of .
The term is equal to the left hand side of (30) with . Furthermore the products and link to the left hand sides of (30) for the fields and respectively. More precisely, one has
where, as before, denotes the set of real places of that become complex places in . The formula for is modified by the same factor as there is exactly one complex place in above each place in . In proving this last formula one also uses the fact that is equal to since .
In addition, since we are working modulo squares, we may neglect the terms and 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 in that
(31) |
and for each finite place in that
(32) |
Now, by our assumption, no place at which has bad reduction is ramified in and so the Néron differential for over remains a Néron differential for over both of the fields and . Hence the equation (32) is valid for all finite places, because is in and one has .
Next we note that the congruence (31) only needs to be checked at places at which has bad reduction (and which therefore do not ramify in ) and at infinite places. If the decomposition group at a place above in is trivial, then both sides of this congruence are equal to . If is cyclic of order , then there is one place in with and one place with for the unique place above in and so both sides of (31) are equal to . Finally we have to treat the case when is cyclic of order and hence the place is finite. In this case the left hand side of (31) is equal to while the right hand side is equal to where is an unramified cubic extension of . If 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 I and then the change is from to , 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 and field satisfy the hypotheses (a)–(h). We assume also that and are as in Theorem 6.1 (so ) and that the Birch and Swinnerton-Dyer conjecture holds for over each of the fields , and .
Then the equality (17) is valid provided that , and there exists a point in which generates a -module of finite prime-to- index in that is isomorphic to .
Proof.
First, we note that given our hypotheses (a), (b) and (e), our choice of (as in §4.4.1) and our assumption that vanishes, the validity of BSD implies that the term is a -adic unit for all .
Next we link to in Theorem 5.8, the difference being the terms versus and the terms and (as defined in (25)). Using the explicit structure of , it is easy to show that
where all congruences are modulo squares in ; namely the quotients are squares of indices, like the index of in . Up to squares in we therefore have a congruence
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 and are -adic units that are congruent modulo . Writing for the number of points in the reduction at a place and , we can summarise the computations of the local contribution to these terms at a place in the following table according to the type of ramification. Here stands for the ramification index at a place in above and for the residual degree.
2 | 1 | 1 | 1 | ||||
---|---|---|---|---|---|---|---|
3 | 1 | 1 | 1 | ||||
3 | 2 | 1 | 1 | 1 |
In the last line denotes the unique place in above . If a place were totally ramified it must be above or as the tame inertia group is cyclic and it can not be above 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 were assumed to be unramified by (h). From the table we can conclude that all the terms and are indeed -units by (f) and that
with .
Hence we are reduced to showing that when and . We first note that for such a place we must have . Indeed, this is true because the map in Corollaire IV.1 in [45] is -equivariant and, since is dihedral, the action of on is non-trivial.
Finally, we have the equality valid for all quadratic extension of finite fields. Hence . ∎
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 to be tamely ramified in and can omit any assumption about the reduction of at such places. In addition, one need only assume that the group vanishes for places that are both inert in and ramify in .
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 but Corollary 6.2 implies that there is limited interest in doing so. We therefore discuss examples with and .
Our numerical computations were done using Sage [48], which uses underlying Pari-GP [49]. The computations of the -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]
over the Hilbert class field of . The curve has rank over and . The extension is of degree defined by a root of the polynomial
All hypotheses (a)–(h) except the finiteness of in (g) can be verified easily. We find in [23] that there is a point
of infinite order on defined over obtained from Darmon’s construction of modular points. The trace of in is equal to the generator . Let be a generator of . It is easy to check that
has finite index coprime to in . Hence we can take 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 and the generator in to assure that the trace generates .) In fact, in our case, we can check that by using the bound given in [21].
Using modular symbols for the character and a Heegner point computation for , we can prove that the formulae (24) evaluate to
and hence conclude that BSD holds for and with . Next, we compute a numerical approximation to for a -dimensional representation . The corresponding value of is equal to for all such , but we know of no means of proving that this value is indeed algebraic and equal to the value . It predicts with good accuracy that BSD for would imply that and are trivial. However assuming that , we compute the -truncated version
and hence find that the congruence (20) holds modulo . Note that in this particular case, the values were all in . In other words we have found convincing numerical evidence that a Gross-Zagier formula
analogous to (23) should hold for all because and .
6.3.2. A quintic example
As a second example, we consider the curve
labelled 21a1 in [20]. It has rank over , but rank over and the group is generated by the point
Now we consider the extension given by a solution of the polynomial
The extension is only ramified at the place . All of our hypotheses except (g) can be verified to hold in this example.
By a simple search for points, we find the point
of infinite order defined over . The bounds in [21] can then be used to prove that generates where is a generator of . To find the point is a bit more elaborate than in the previous example as . In fact the -module generated by and will have index in because [18, Corollary 2.5] tells us that is projective. We are going to use the relation
in which is reminiscent of Kolyvagin’s derivative construction. We now try to find a point in with such that is divisible by in . Then we can take such that to be the point predicted by Proposition 5.7. In our concrete case this works with . It can be shown that as a -module.
Using modular symbols and Heegner points, we can provably compute that
and hence deduce that BSD holds for and with trivial Tate-Shafarevich groups in both cases. We compute to a high precision the derivatives of for the representations and of dimension two and we find, with an error less than , that , predicting that and . It also predicts that has order . We will now assume that these are actually equalities and conclude that . One then computes the -truncated values to be
and this shows that the congruence (20) holds modulo .
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 -functions, to appear in J. reine u. angew. Math., 2015.
- [8] S. Bloch and K. Kato, -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 -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 -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 -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 -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, -adic periods and -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 -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, -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 -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 -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 -functions, Algebraic number fields: -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, -functions and Galois representations, London Math. Soc. Lecture Note Ser., vol. 320, Cambridge Univ. Press, Cambridge, 2007, pp. 333–380.