Elliptic Curves of Odd Modular Degree
Frank Calegari
*
*
Supported in part by the American Institute of Mathematics
Matthew Emerton
†
†
Supported in part by NSF grant DMS-0401545
November 27, 2006
1 Introduction
Let
be an elliptic curve over
of conductor
. Since
is modular [3] , there exists a surjective map
defined over
. There is a unique such map of minimal degree (up to composing with automorphisms of
), and its degree
is known as the modular degree of
. In [24] the parity of
was determined for a very particular explicit class of elliptic curves, namely, the Neumann–Setzer curves, which have prime conductor and a rational
-torsion point. (See also the remark following theorem 5.1 below.) We study the question more generally for arbitrary elliptic curves
, and prove the following theorem:
1.1 Theorem.
If
is an elliptic curve of odd modular degree then:
-
1.
the conductor
of
is divisible by at most two odd primes,
-
2.
is of even analytic rank, and
-
3.
either (a)
has a rational point of order
(or equivalently, admits a rational
-isogeny ), (b)
has prime conductor and supersingular reduction at
, and
is totally complex (equivalently,
is connected ), or (c)
has complex multiplication, and
.
1.2 Example.
The following examples of elliptic curves with odd modular degree should serve to illustrate conditions (3a), (3b) and (3c). The curve
has modular degree one and a rational two torsion point, and thus satisfies condition (3a). Another example is given by the curve
(
in Cremona's tables) of conductor
with modular degree
and torsion subgroup
. The curves
and
both have modular degree one and satisfy condition (3b). An example of larger conductor is given by
of conductor
and modular degree
. Finally, there are exactly four curves of odd modular degree with complex multiplication, namely
,
,
(all of modular degree one) and
of modular degree
, conductor
and
-invariant
.
1.4 Remark.
Cremona and Watkins have computed the modular degree of all optimal elliptic curves of conductor
25,000 [
7]
and these computations suggest that there may be even stronger limitations on the conductor of a curve of odd modular degree than those imposed by theorem 1.1 . Indeed, in the range of Watkins' computations, every curve of odd modular degree has conductor divisible by at most two primes, and the conductor always has one of the following forms:
,
, or
, where
and
are odd primes.
1.5 Remark.
The statement of the theorem regarding the analytic rank of
is consistent with the rank conjecture of Watkins [
27,conj. 4.1]
(and the conjecture of Birch and Swinnerton-Dyer).
The following result, conjectured by Watkins ([24,conj. 4.3] , [27,conj. 4.2] ), is a simple consequence of theorem 1.1 (see lemma 3.5 ):
1.6 Theorem.
Let
be an elliptic curve of prime conductor
, and suppose that
is neither a Neumann–Setzer curve, nor
(equivalently,
does not have a rational
-torsion point ). If
is odd, then
.
One technique for proving that an elliptic curve
has even modular degree is to show that the map
factors through
for some non-trivial Atkin–Lehner involution
.
We use this approach in section 2 to prove theorem 2.1 , which in turn implies parts (1) and (2) of theorem 1.1 , and shows that (3a) holds if
is divisible by at least two primes.
It remains to prove (3) in the case when
is a prime power. The most difficult case to handle is when
is actually prime, and in this case we deduce theorem 1.1 from the following result, proved in section 3.
1.7 Theorem.
Let
be prime, let
denote the Hecke algebra over
acting on weight two cuspforms on
, and let
be a maximal ideal of
such that
, and such that the associated semi-simple Galois representation
is irreducible. If the completion
, then
-
1.
is supersingular at
, and
-
2.
is totally complex.
The relevance of this result to theorem 1.1 is that, since
is prime in the context of theorem 1.7 , a result of Ribet [30] shows (assuming, as we may, that
is optimal) that the modular degree of
is even if and only if
is a congruence prime for the newform of level
attached to
.
The proof of theorem 1.7 is motivated by the following considerations: If
is an odd prime and
is a surjective modular representation, then theorems of Wiles and Taylor–Wiles [29, 26] show that the universal minimal deformation ring
attached to
is isomorphic to the universal minimal modular deformation ring
(
, since
is prime). Since
is a finite
algebra with residue field
, it is exactly equal to
if and only if it is an étale
-algebra. On the other hand, since
, this is equivalent to
being étale over
, which is in turn equivalent to the reduced Zariski cotangent space of
being trivial. Since by construction
represents the minimal deformation functor, its reduced Zariski cotangent space considered as a set has cardinality equal to the number of minimal deformations
of
. Thus to prove that
it suffices to show that there exists a non-trivial minimal deformation of
to
.
In spirit, the proof of theorem 1.7 follows this strategy; in other words, we determine whether or not
by a calculation on tangent spaces. A significant problem arises, however, since we are working in the case
, whilst the method of Wiles and Taylor–Wiles applies only to
. This is not a mere technical obstruction; many phenomena can occur when
that do not occur for odd
. To name two such: the possible failure of
to be Gorenstein and the consequent failure of multiplicity one [16] , and the fact that
can arise from a totally real extension of
. Calculations in the second case suggest that the Taylor–Wiles strategy for proving
in the minimal case will not work without some significant new idea, since the numerical coincidences that occur for odd
whilst balancing the Selmer and dual Selmer groups in the Greenberg–Wiles product formula (see for example the remarks of de Shalit [8] , top of p. 442) do not occur in the case
. Mark Dickinson [10] has proved an
theorem for
; however, his result requires many non-trivial hypotheses, and indeed does not apply to any of the representations considered in theorem 1.7 . (Its main application to date has been to representations with image
.) The issue here is that the Taylor–Wiles auxiliary prime arguments fail when
and the image of
is dihedral.
Thus, instead of appealing to any general modularity results, we show that
is bigger than
by explicitly constructing (in certain situations) non-trivial deformations of
to
that are demonstrably modular (and hence contribute to the reduced cotangent space of
). The most difficult point is to show that these deformations are modular of the correct (minimal) level. We prove this via a level-lowering result for modular forms with values in Artinian
-algebras (theorem 3.18 below). This level lowering result maybe of independent interest; for example, it provides evidence that an
theorem should hold for those
of characteristic two to which it applies.
The proof of (3) when
is a prime power (but is not actually prime) is given in section 4. In section 5 we make some concluding remarks.
Let us close this introduction by pointing out that recently Dummigan [11] has provided a heuristic explanation for Watkins' rank conjecture that also relies on a hypothetical
theorem for the residual Galois representation
arising from the
-torsion on an elliptic curve
: he uses the symmetric square map from
to
to lift elements from
-Selmer group of
to the tangent space to the deformation ring of
. He also shows that the resulting tangent space elements can never be “trapped” (in the words of [29,p. 450] ) by the Taylor–Wiles method of introducing auxiliary primes.
Altogether, the experimental work of Watkins on the parity of modular degrees, taken together with the results of this paper and of [11] , suggests the validity of an
theorem for (at least certain) residual Galois representations arising from the
-torsion on elliptic curves, a theorem whose proof, however, seems out of reach of the current techniques.
2
composite with at least two distinct prime factors
In this section we prove the following theorem.
2.1 Theorem.
If
is an elliptic curve of odd modular degree then the conductor
of
is divisible by at most two odd primes, and
is of even analytic rank. Furthermore, if
is divisible by at least two primes, then
contains a rational
-torsion point.
We begin with a preliminary lemma. Let
be an elliptic curve over a field
; let
denote the origin of
. We let
denote the group of automorphisms of
as a curve over
(i.e.
-rational automorphisms of
that do not necessarily fix
), and suppose that
is a finite elementary abelian
-subgroup of
.
2.2 Lemma.
The order of
divides twice the order of
.
Proof. Let
denote the subgroup of
consisting of automorphisms of
as an elliptic curve over
(i.e.
-rational automorphisms of
that do fix
). The action of
on
via translation realizes
as a normal subgroup of
which has trivial intersection with
, and which together with
generates
. Thus
sits in the split short exact sequence of groups
|
(1)
|
(This is of course well known. The surjection
may also be regarded as the map
induced by the functoriality of the formation of Picard varieties — the target being the group of automorphisms of
as a group variety — once we identify
and
as group varieties in the usual way.) The short exact sequence ( 1 ) induces a short exact sequence
where
denotes the projection of
onto
. The known structure of
shows that
is either trivial or of order
. Since
the lemma follows.
Proof of theorem 2.1 . The discussion of remark 1.3 shows that it suffices to prove the theorem under the additional assumption that
is an optimal quotient of
.
Let
denote the group of automorphisms of
generated by the Atkin–Lehner involutions;
is an elementary abelian
-group of rank equal to the number of primes dividing
. Since
is an optimal quotient of
, the action of
on
descends to an action on
. If
were to act trivially on
, then the quotient map
would factor through
, contradicting our assumption that
is of odd modular degree. Thus lemma 2.2 shows that
has order at most
, and hence that
is divisible by at most
primes. Furthermore, if
is divisible by more than one prime, then it shows that
is non-trivial. Suppose now that
is odd, so that
and
both have good reduction at
. We may then apply the argument of the preceding paragraph over
, and so conclude from lemma 2.2 that
has order at most
. Hence
is divisible by at most two primes.
If
is of odd analytic rank, and if
denotes the normalized newform of level
attached to
, then
, and so the automorphism of
induced by
has trivial image in
. Thus
acts on
via translation by an element
. Since
interchanges the cusps
and
on
, we see that
(where
is a modular parameterization of
, chosen so that
).
The assumption that
has odd analytic rank also implies that
. Since this
-value can be computed (up to a non-zero factor) by integrating
from
to
in the upper half-plane, we conclude that
and thus that
acts trivially on
. Hence
factors through the quotient
of
, and so must be of even modular degree, a contradiction.
3
prime
3.1 Reductions
3.2 Lemma.
Theorem 1.7 implies part (3) of theorem 1.1 for
prime.
Proof. Suppose that
is an elliptic curve of conductor
, assumed to be optimal in its isogeny class. Let
be the associated Hecke eigenform of level
and weight
. From a theorem of Ribet [
30]
,
if and only if
satisfies a congruence with another cuspidal eigenform of level
. The set of cuspidal eigenforms (in characteristic zero) congruent to
is indexed by
. Thus
satisfies no non-trivial congruences if and only if
, or equivalently if and only if
.
The following lemma shows that it is possible to detect congruences from modular forms modulo two.
3.3 Lemma.
Let
be the maximal ideal in
associated to
, and suppose there exist two distinct non-zero
-expansions
,
with coefficients in
such that
. Then
.
Proof. Since
is irreducible,
and
are both cuspidal. If
, then
. Thus
and
are both
-expansions which are killed by
. By multiplicity one for
-expansions [
18,prop. 9.3]
it follows that
.
Of use to us will be the following theorem of Grothendieck on Abelian varieties with semistable reduction [15,ExposéIX,prop. 3.5] :
3.4 Theorem (Grothendieck).
Let
be an Abelian variety over
with semistable reduction at
. Let
denote a choice of inertia group at
. Then the action of
on the
-division points of
for
is rank two unipotent; i.e., as an endomorphism, for
,
In particular,
acts through its maximal pro-
quotient, which is procyclic.
Shimura proved that a modular form
of weight
and level
gives rise to a modular Abelian variety
in such a way that the
-adic representations
attached to
arise from the torsion points of
. For prime
, these varieties are semistable at
, and so we may apply the theorem above to deduce that for
, such representations
restricted to
factor through a pro-cyclic
-group. For representations
with image inside
, this means in particular that the order of inertia at
is either
or
.
Let us now consider a Galois representation
, arising from a cuspidal Hecke eigenform of level
, whose image is not contained in a Borel subgroup. (This is equivalent to
being irreducible, and also to the image of
having order divisible by
.) Let
be the fixed field of the kernel of
; the extension
is unramified outside
and
.
If
is unramified at
then
has Serre conductor
, contradicting a theorem of Tate[25] . Thus by the discussion above we see that inertia at
factors through a group of order
, that
is an
-extension, and (hence) that
is absolutely irreducible. Let
be a cubic subfield of
, and let
be the quadratic extension inside
. Since
is finite flat at
, it follows from Fontaine's discriminant bounds [13] that the power of
dividing the discriminant of
is at most
. Thus
must be
(as it is ramified at
).
The extension
is unramified at the prime above
and is ramified at
if and only if
is supersingular.
3.5 Lemma.
If
is supersingular at two and not totally real then
. In particular, theorem 1.1 implies theorem 1.6 .
Proof. By class field theory the quadratic field
admits a degree three extension ramified precisely at
only if
is unramified and inert in
. This occurs if and only if
and
, or
and
. Moreover if
then
and
are totally real.
We shall prove theorem 1.7 by showing in the following subsections that if
satisfies at least one of the following conditions:
-
1.
is totally real;
-
2.
is unramified at
;
-
3.
is ordinary, complex, and ramified at
;
then
.
3.6
is totally real
The theory of modular deformations is not well understood when
is totally real. Thus our arguments in this section are geometric. We use the following theorem, due to Merel [20,prop. 5] . (This interpretation of Merel's result is due to Agashe [1,cor. 3.2.9] ).
3.7 Theorem.
Let
be prime. Then
is connected.
If we let
denote the dimension of
it follows that
,
and
.
Let
. Then
is a
-divisible group over
admitting an action of
.
Since
is finite and flat over the complete local ring
there exists a decomposition
where the product is taken over the maximal ideals
of
of residue characteristic two.
If
denotes the rank of
over
, then
If
, then
(compare [18] §7, p. 91). From lemma 7.7 of [18] we see that
is free of rank two over
(where
is the
-adic Tate module of
), and thus that
|
(2)
|
Let
be the
-torsion subgroup scheme of
.
3.8 Lemma.
For all maximal ideals
of residue characteristic two there is an equality
Proof. The isomorphism ( 2 ) induces an isomorphism
Let
denote complex conjugation. Then
. Thus
(which is the kernel of
) has dimension at least
. If
for some
, then since
and since (as was noted above)
, we would deduce the inequality:
which is absurd.
Now let
be a totally real (absolutely) irreducible continuous modular representation of
into
of level
, and let
be the corresponding maximal ideal of
. The main result of [2] shows that the
-representation
is a direct sum of copies of
. Thus, since
is totally real, we find that
Combining this inequality with the inclusion
and lemma 3.8 we find that
, and thus that
.
3.9
is unramified at
Suppose that
is unramified at
. This forces
to be ordinary. By the theory of companion forms [14] one expects that
arises from a mod
form of level
and weight
. Although the results of [14] do not apply in this case, Wiese [28] explicitly constructs such forms when the image of
is dihedral, as it is in our situation. (In fact the only difficult point of Wiese's construction is the case when
is totally real, and this case of theorem 1.7 is already covered by section 3.6 .) Let
be this companion (Katz) modular form of weight one modulo
. Let
be the Hasse invariant modulo
, which is a modular form of level one with
expansion given by
. Then
and
are two distinct
-expansions modulo
of weight
and level
. Moreover, one sees that
for all odd
, and that
and
. Thus
and therefore by lemma 3.3 ,
and we are done.
3.10
is ordinary, complex, and ramified at
Suppose that
is ordinary, complex, and ramified at
. It follows that
is complex and ramified at
, and thus that
for some
. Moreover, the extension
is unramified everywhere. Since
it follows that
is also unramified everywhere over
. The field
is Galois over
, and clearly
|
(3)
|
We may embed
into
by fixing an identification of
with
, and mapping a generator of
to the matrix
Composing the isomorphism ( 3 ) with this embedding yields a representation:
The representation
has trivial determinant (equivalently, determinant equal to the mod
cyclotomic character). We also claim that
is finite flat at two. To check this, it suffices to prove this over
. The representation
factors through a group of order
, and one explicitly sees that it arises as the generic fibre of the group scheme
, where
is the non-trivial extension of
by
considered in [18] (Prop 4.2, p. 58).
Thus one expects
to arise from an
-valued modular form of weight two and level
, corresponding to a surjective map of rings
. This would follow if we knew that
coincided with the minimal deformation ring associated to
. Rather than proving this, however, we shall use weight one forms to explicitly construct a weight two modular form giving rise to
.
Let
be the character of conductor
associated to
. Consider two faithful representations
Since
is complex, these dihedral representations are odd and therefore give rise to weight one modular forms
,
in
.
3.11 Lemma.
The modular forms
,
are ordinary at
, have coefficients in
, and are congruent modulo
. Let
Then
is non-zero, and the form
is an eigenform for all the Hecke operators, including
. The associated
Galois representation attached to
is
.
Proof. The modular forms are both ordinary at
because the representations
and
have non-trivial subspaces on which inertia at two is trivial (since
acts through a group of order
). They both have coefficients in
, since
. The congruence
follows from the fact that both are ordinary-at-
Hecke eigenforms, and that
for all odd primes
, where
is the character of conductor
.
From this one also sees that
is non-trivial modulo two. The claims about
follow formally from the fact that
and
are Hecke eigenforms that are congruent modulo
, and the definition of
.
Now that we have constructed the weight one form
of level
giving rise to
, we would like to construct a corresponding weight two form of level
. Multiplying
by the Eisenstein series in
, we see that
is the
-expansion of a modular form in
. Since
is ordinary and is a
eigenform, we may apply the
operator to deduce that
. Applying theorem 3.18 (proved in the following subsection) we then deduce that in fact
, and (thus) that
. An easy calculation shows that if
then
, whilst
, and so also
. Thus
, by lemma 3.3 .
3.12 Level-lowering for modular deformations
The goal of this section is to prove a level-lowering result for modular forms with coefficients in Artinian rings that strengthens the case
of [12,thm. 2.8] (which in turn extends a level lowering result proved by Mazur [22,thm. 6.1] in the odd prime case).
We first establish a version of the multiplicity one theorem [29,thm. 2.1] for
.
Under the additional assumption that
is not finite at
, this theorem was proved in [4,§2] (as was the corresponding result for odd level). Thus the key point in our theorem is that
is allowed to be finite at
, even though the level is taken to be even.
3.13 Theorem.
Let
be an odd natural number, and let
denote the full
-algebra of Hecke operators acting on weight two cuspforms of level
. If
is a maximal ideal in
whose residue field
is of characteristic
, and for which the associated residual Galois representation
is (absolutely ) irreducible, ordinary, and ramified at
, then
(the
-adic Tate module of
) is free of rank two over the completion
.
To be clear, the condition “ordinary at
” means that the image of a decomposition group at
under
lies in a Borel subgroup of
. Since
is of characteristic
, we see that (for an appropriate choice of basis) the restriction of
to an inertia group at
may be written in the form
The assumption that
is ramified at
then implies that
is not identically zero. Thus the representation space of
has a unique line invariant under
, and so
is irreducible if and only if it is absolutely irreducible.
3.14 Lemma.
Let
be a finite field of characteristic
. If
is a continuous representation that is finite, ordinary, and ramified at
, then
has a unique finite flat prolongation over
(up to unique isomorphism). Furthermore, this prolongation is an extension of a rank one étale
-vector space scheme by a rank one multiplicative
-vector space scheme.
Proof. Any finite flat group scheme that prolongs an unramified continuous representation of
on a one-dimensional
-vector space is either étale or multiplicative. Thus there are a priori four possible structures for a finite flat prolongation of
: étale extended by étale; multiplicative extended by multiplicative; multiplicative extended by étale; or étale extended by multiplicative. However, all but the last possibility necessarily gives riseto an unramified generic fibre (note that any extension of multiplicative by étale must split, by a consideration of the connected étale sequence). Thus, since
is assumed ramified, we see that any prolongation of
must be an extension of a rank one étale
-vector space scheme scheme by a rank one multiplicative
-vector space scheme.
To see that such a prolongation is unique, consider the minimal and maximal prolongations
and
of
to a finite flat group scheme [
21,cor. 2.2.3]
. The result of the preceding paragraph shows that the natural morphism
necessarily induces an isomorphism on the connected components of the identity, and on the corresponding groups of connected components. By the
-lemma, this morphism is thus an isomorphism, and the lemma follows.
We now show that certain results of Mazur [19] cited in the proof of [29,thm. 2.1] extend to the case
. We put ourselves in the context of [19,§1] , and use the notation introduced there. Namely, we let
denote a finite extension of
for some prime
, and let
denote the ring of integers of
. If
is an abelian variety over
, then we let
denote the connected component of the identity of the Néron model of
over
. For any power
of
, the
-torsion subgroup scheme
of
is then a quasi-finite flat group scheme over
; we let
denote its maximal finite flat subgroup scheme, and
denote the maximal connected closed subgroup scheme of
.
Since we took
to be the connected component of the Néron model of
, the inductive limit
is a
-divisible group, and
is the maximal connected
-divisible subgroup of
.
The following proposition is a variation on [19,prop. 1.3] , in which we allow the ramification of
over
to be unrestricted, at the expense of imposing more restrictive hypotheses on the reduction of the abelian varieties appearing in the exact sequence under consideration.
3.15 Proposition.
Let
be an exact sequence of abelian varieties over
such that
has purely toric reduction, whilst
has good reduction. Then the induced sequence of
-divisible groups
is a short exact sequence of
-divisible groups over
. Equivalently, for any power
of
, the induced sequence
is a short exact sequence of finite flat group schemes over
.
Proof. Since
has purely toric reduction, the group scheme
is of multiplicative type for each
. Thus it necessarily maps isomorphically onto its scheme theoretic image in
, and thus the induced map
is a closed embedding.
Let
be an abelian subvariety chosen so that the induced map
is an isogeny. Then
also has good reduction, and so
is an epimorphism of
-divisible groups over
. Thus the induced map
is also an epimorphism of
-divisible groups. A consideration of generic fibres shows that the kernel of this surjection coincides with the scheme-theoretic image of
in
, and so the proposition is proved.
Proof of theorem 3.13 . We closely follow the method of proof of [29,thm. 2.1 (ii)] in the case when “
is trivial
” (in the terminology of that proof; see [29,pp. 485–488] ).
If we let
denote the connected part of the kernel of the map
induced by Albanese functoriality applied to the two “degeneracy maps” from level
to level
, then
is an abelian subvariety of
having purely toric reduction at
, whilst the quotient
of
by
has good reduction at
. From proposition 3.15 we obtain (for any
) the short exact sequence
of connected finite flat group schemes over
. By functoriality of the formation of this short exact sequence, and since
is a
-invariant subvariety of
, we see that this is in fact a short exact sequence of
-module schemes. Localizing at
induces the corresponding short exact sequence
|
(4)
|
Passing to
-valued points induces a short exact sequence of
-modules
|
(5)
|
which is a subexact sequence of the short exact sequence of
-modules
|
(6)
|
Let
(respectively
, respectively
) denote the maximal
-subrepresentation of
(respectively
, respectively
) on which the inertia group acts through the
-adic cyclotomic character
.
The short exact sequence ( 6 ) induces an exact sequence
|
(7)
|
3.16 Lemma.
Each of the groups schemes appearing in the exact sequence ( 4 ) is of multiplicative type, and the exact sequences ( 5 ) and ( 7 ) coincide (as subsequences of ( 6 )).
Proof. We first remark that ( 6 ) is the exact sequence of
-modules underlying the the corresponding exact sequence of
-modules
Since
is assumed irreducible as a
-representation, each of the modules appearing in this exact sequence is a successive extension of copies of
. The same is thus true of each of the modules appearing in the exact sequence ( 6 ).
Since
has purely toric reduction, it is clear that
is of multiplicative type, and so
|
(8)
|
Fix a filtration
of
for which the successive quotients
are isomorphic to
. Since
has purely toric reduction the quotient
is Cartier dual to
(where
is the dual abelian variety to
), and so
is an unramified
-representation. Since
is assumed ramified at
, this implies that
for each
. Furthermore,
for each
, because
is trivial. Since
is two dimensional over
for each
we conclude by induction on
that
for each
. Taking
then shows that the inclusion ( 8 ) is in fact an equality.
Since
has good reduction at
, we have equality
. As noted above, any Jordan–Hölder filtration of the localization
as a
-modulehas all its associated graded pieces isomorphic to
. Taking scheme-theoretic closures of such a filtration in
we obtain a filtration of the localization
by finite flat closed subgroup schemes, whose associated graded pieces are prolongations of
. Now lemma 3.14 shows that the connected component of any such finite flat prolongation is multiplicative. Thus
is indeed multiplicative, whilst
is an unramified
-module. Arguing as in the preceding paragraph gives the required equality
Since any extension of multiplicative type groups is again of multiplicative type, we see that
is also of multiplicative type, and that the exact sequence ( 5 ) is a subsequence of the exact sequence ( 7 ). We have furthermore shown that first and third non-trivial terms of these two sequences coincide. This implies that these exact sequences do indeed coincide.
Specializing lemma 3.16 to the case
shows that
is the maximal unramified
-subrepresentation of
(since
is trivial).
Recall that there is a natural isomorphism
(indeed, this is true with
replaced by any group scheme over
), and also a natural isomorphism
(as follows from the discussion on [29,p. 488] ). Localizing at
, and taking into account [29,lem. 2.2] , which is valid for
, we find that
is a cyclic
-module, and thus that the maximal unramified
-subrepresentation of
is a cyclic
-module.
Let
denote the Galois representation associated to
by [6,thm. 3] . Carayol has proved [6,thm. 4] that there is an isomorphism
for some ideal
in
, and thus an isomorphism
. We conclude that
is a cyclic
-module, and hence that
is a principal ideal in
.
The discussion of [6,3.3.2] shows that in fact
and that
is free of rank two over
, as claimed.
3.17 Corollary.
In the situation of theorem 3.13 , the completion
is a Gorenstein
-algebra.
Proof. This follows from the theorem together with the self-duality of the
-adic Tate module under the Weil pairing.
We now prove our level lowering result. Let
be an Artinian ring with finite residue field
of characteristic
, and suppose given a continuous representation
that is modular of level
for some odd natural number
, in the sense that it arises from a Hecke eigenform
. Let
denote the residual representationattached to
(so
arises from the Hecke eigenform
obtained by reducing
modulo the maximal ideal of
).
3.18 Theorem.
If
is a modular Galois representation of level
as above, such that
-
1.
is (absolutely ) irreducible,
-
2.
is ordinary and ramified at
, and
-
3.
is finite flat at
,
then
arises from an
-valued Hecke eigenform of level
.
Proof. The Hecke eigenform
corresponds to a ring homomorphism
. Since
is local of residue characteristic
, the map
factors through the completion
of
at some maximal ideal
of residue characteristic
, and the residual representation
is the residual Galois representation attached to the maximal ideal
. We let
denote the Galois representation
attached to
by [
6,thm. 3]
. The Galois representation
attached to
coincides with the pushforward of
via
.
Replacing
by the image of
, we may and do assume from now on that
is surjective.
We let
denote the kernel of
. Since
is Artinian, we may choose
so that
. Theorem 3.13 shows that the
-adic Tate module
is isomorphic as a
-representation to
; thus
is isomorphic to the reduction mod
of
. Since
is a Gorenstein
-algebra, by corollary 3.17 , we see that
is a Gorenstein
-algebra, and thus that there is an isomorphism
of
-modules. In particular,
is a faithful
-module, isomorphic as an
-module to
To simplify notation, we will write
|
(9)
|
By assumption,
prolongs to a finite flat group scheme
over
. If we fix a Jordan–Hölder filtration of
as an
-module, then the associated graded pieces are each isomorphic to
, and so lemma 3.14 and [
5,Prop. 2.5]
together imply that
is uniquely determined by
, whilst [
5,Lem. 2.4]
then implies that
is naturally an
-module scheme. From ( 9 ) we see that
also prolongs to a finite flat
-module scheme
over
. Again, lemma 3.14 and [
5,Prop. 2.5]
show that
is the unique finite flat prolongation of
.
Lemma 3.14 furthermore implies that
is the extension of an étale
-module scheme
by a multiplicative
-module scheme
, each of which is free of rank one as an
-module scheme. Thus
is also an extension of an étale
-module scheme
by multiplicative
-module scheme
, each of which is faithful as an
-module scheme.
Let
and
denote the generic fibres of these schemes.
We write
to denote the Néron model of
over
. For a scheme over
, use the subscript “
” to denote its special fibre over
. The special fibre
admits the following filtration by
-invariant closed subgroups:
where
is the maximal torus contained in
and
is the connected component of the identity of
. The quotient
is an abelian variety on which
acts through its quotient
(where
denotes the quotient of
that acts faithfully on the space of
-old forms of level
). The connected component group
is Eisenstein [
22,Thm. 3.12]
.
The following lemma provides an analogue in our situation of [
22,lem. 6.2]
(and generalizes one step of the argument in the proof of [
12,thm. 2.8]
).
3.19 Lemma.
The Zariski closure of
in
is a finite flat
-module scheme over
(which is thus isomorphic to
).
Proof. Since
is a multiplicative type group scheme, inertia at
acts on
through the cyclotomic character. It follows from lemma 3.16 that
is contained in the generic fibre of
, and thus that the Zariski closure of
in
is indeed finite flat, and in fact of multiplicative type. Thus it coincides with
, and so we see that the embedding of
in
prolongs to an embedding of
in
. Since the quotient
is étale, lemma 5.9.2 of [
15,ExposéIX]
serves to complete the proof of the lemma.
Lemma 3.19 allows us to regard
as a closed
-submodule scheme of
, and thus to regard
as a closed
-submodule scheme of
. Since
is irreducible and
is Eisenstein, we see that
is in fact contained in
. On the other hand, since
is a torus, we see that
. Thus
appears as a subquotient of
, and in particular the
-action on
factors through the quotient
of
. Since
is a faithful
-module scheme, we see that the map
factors through
, completing the proof of the theorem.
We remark that the obvious analogue of theorem 3.18 in the case of odd residue characteristic is also true. The proof is similar but easier, relying on the uniqueness results on finite flat models due to Raynaud [21] . Of course, in those cases when the
theorem of Wiles, Taylor–Wiles, and Diamond [9, 26, 29] applies, it is an immediate consequence of that theorem. (Thus theorem 3.18 may be regarded as evidence for an
theorem for those
of residue characteristic
that satisfy its hypotheses.)
4
a proper prime power
There are only finitely many elliptic curves of conductor
for all
, and we may explicitly determine which have odd modular degree. Therefore we assume that
has conductor
, where
with
and
. Let
be the unique quadratic character of conductor
. Let
be the elliptic curve
twisted by
. The curve
also has conductor
, and moreover the associated modular forms
and
are congruent modulo
, since twisting by quadratic characters preserves
. Since
is odd, any non-trivial congruence modulo
between
and other forms in
forces the modular degree
to be even [30] . Thus we are done unless
. In particular, the representations associated to
must be induced from a quadratic field, and thus
has complex multiplication. (Alternatively, the equality
implies that
is isogenous to its twist and deduce this way that
has CM.) If
has CM and prime power conductor then
is one of finitely many well known elliptic curves, for which we can directly determine the modular degree by consulting current databases (for
, we use the elliptic curve database of Stein–Watkins, described in [23] ).
5 Further remarks
Certainly not every
satisfying the conditions of theorem 1.1 will actually have odd modular degree, and one could try to refine this result by deducing additional necessary conditions that
must satisfy in order to have odd modular degree. In this section we say a little about the related question of whether or not
is a congruence prime for the associated modular form
, when
satisfies either of conditions (3a) or (3b) of the theorem.
For curves
with a rational two torsion point, the modular form
automatically satisfies a mod two congruence with an Eisenstein series, and so detecting whether
satisfies a congruence with a cuspform is a more subtle phenomenon than in the non-Eisenstein situation. One approach might be to relate the Hecke algebra to an appropriate universal deformation ring (if the latter exists). If
is prime this can be done [5] , and this enables one to determine when
for such representations. The specific determination of when
, however, was already achieved (for
prime and
reducible) by Merel in [20] :
5.1 Theorem.
Let
be prime, and let
be the localization of the Eisenstein prime at
. Then
if and only if
and
.
If
is a Neumann–Setzer curve, then
for some
. The result of Merel above then clearly implies that the optimal Neumann–Setzer curve
has odd modular degree if and only if
. (An alternative proof of this fact, relying on the results of [18] , is given in [24] , thm. 2.1.) If
has composite conductor, then one might try to generalize the results of [20] or [5] to this setting.
Suppose now that
has prime conductor, that
is irreducible and supersingular, and that
is totally complex. Let
and
be the extensions of
attached to
as in the discussion of section 3 . If one had an
result of the type discussed in the introduction, then to obtain further necessary conditions for
to have odd modular degree, it would suffice to establish sufficient conditions for the existence of an appropriate non-trivial minimal deformation
lifting
. For representations
that were complex and ramified at
, but ordinary, we constructed such a
directly in subsection 3.10 by considering a quadratic genus field extension of
. When
is supersingular, such deformations
(when they exist) may be more subtle and can not necessarily be constructed so directly. One can however prove the following result.
Recall that if
is supersingular and totally complex, then
is totally ramified at
and
has exactly two complex embeddings.
5.2 Proposition.
Suppose
is supersingular at
and totally complex. If either:
-
1.
the class number of
is even, or
-
2.
the fundamental unit
of
satisfies
, where
is the unique prime above
in
,
then there exists a non-trivial minimal deformation
Note that
. The idea behind the proof of this result is to study
extensions
containing
that are minimally ramified at
and unramified at
. Such extensions can be obtained by considering the Galois closure over
of certain quadratic extensions of
. One obtains a suitable such extension either by considering an unramified extension of
, in case (1), or the extension
, in case (2).
References
-
A. Agashe, The Birch and Swinnerton-Dyer formula for modular abelian varieties of analytic rank zero, Berkeley Ph.D. Thesis, 2000.
-
N. Boston, H. Lenstra, K. Ribet, Quotients of group rings arising from two-dimensional representations, C. R. Acad. Sci. Paris, Série I 312 (1991), 323–328.
-
C. Breuil, B. Conrad, F. Diamond, R. Taylor, On the modularity of elliptic curves over
: wild
-adic exercises, J. Amer. Math. Soc. 14 (2001), no. 4, 843–939.
-
K. Buzzard, On level-lowering for mod 2 representations, Math. Res. Lett. 7 (2000), no. 1, 95–110.
-
F. Calegari, M. Emerton, On the ramification of Hecke algebras at Eisenstein primes, to appear in Invent. Math.
-
H. Carayol, Formes modulaires et répresentations Galoisiennes á valeurs dans un anneau local complet, in
-adic monodromy and the Birch and Swinnerton-Dyer conjecture (B. Mazur, G. Stevens, eds.), Contemp. Math. 165 (1994), 213–235.
-
J. Cremona, M. Watkins, data available at http://www.maths.nott.ac.uk/personal/jec/ftp/data
-
E. de Shalit, Hecke rings and universal deformation rings, in Modular forms and Fermat's last theorem (Boston, MA, 1995), 421–445, Springer.
-
F. Diamond, On deformation rings and Hecke rings, Ann. Math. 144 (1996), no. 1, 137–166.
-
M. Dickinson, On the modularity of certain
-adic Galois representations, Duke Math. J. 109 (2001), no. 2, 319–382.
-
N. Dummigan On a conjecture of Watkins, preprint, 2004.
-
B. Edixhoven, The weight in Serre's conjectures on modular forms, Invent. Math. 109 (1992), no. 3, 563–594.
-
J. Fontaine, Il n'y a pas de variété abélienne sur
, Invent. Math. 81, 1985, no. 3, 515–538.
-
B. Gross, A tameness criterion for Galois representations associated to modular forms (mod
), Duke Math. J. 61 (1990), no. 2, 445–517.
-
A. Grothendieck, Groups de monodromie en géométrie algébrique I, (SGA 7), Séminaire de Géométrie Algébrique du Bois-Marie, 1967-1969, Lecture Notes in Math., vol. 288 , Springer-Verlag, Berlin and New York, 1972, 313–523.
-
L. Kilford, Some non-Gorenstein Hecke algebras attached to spaces of modular forms, J. Number Theory 97 (2002), no. 1, 157–164.
-
W. Li, Newforms and functional equations, Math. Ann. 212 (1975), 285–315.
-
B. Mazur, Modular curves and the Eisenstein ideal, Publ. Math. IHES 47 (1977), 33–186.
-
B. Mazur Rational isogenies of prime degree, Invent. Math. 44 (1978), no. 2, 129–162.
-
L. Merel, L'accouplement de Weil entre le sous-groupe de Shimura et le sous-groupe cuspidal de
, J. Reine Angew. Math. 477 (1996), 71–115.
-
M. Raynaud, Schémas en groupes de type
, Bull. Soc. Math. France 102 (1974), 241–280.
-
K. Ribet, On modular representations of
arising from modular forms, Invent. Math. 100, (1990), no. 2, 431–476.
-
W. Stein, M. Watkins, A Database of Elliptic Curves—First Report, Algorithmic number theory (Sydney, 2002), 267–275, Lecture Notes in Comput. Sci., 2369, Springer, Berlin, 2002.
-
W. Stein, M. Watkins, Modular parametrizations of Neumann–Setzer elliptic curves, IMRN, no. 27 (2004), 1395–1405.
-
J. Tate, The non-existence of certain Galois extensions of
unramified outside
, in Arithmetic Geometry (N. Childress, J. Jones, eds.), Contemp. Math. 174 (1994), 153–156.
-
R. Taylor, A. Wiles, Ring-theoretic properties of certain Hecke algebras, Ann. Math. 141 (1995), no. 3, 553–572.
-
M. Watkins, Computing the modular degree of an elliptic curve, Experiment. Math. 11 (2002), no. 4, 487–502.
-
G. Wiese, Dihedral Galois representations and Katz modular forms, Doc. Math. 9 (2004), 123–133.
-
A. Wiles, Modular elliptic curves and Fermat's last theorem, Ann. Math. 141 (1995), no. 3, 443–551.
-
D. Zagier, Modular parametrizations of elliptic curves, Canad. Math. Bull. 28 (1985), no. 3, 372–384.
Email addresses: fcale@math.harvard.edu emerton@math.northwestern.edu