Hida theory for special orders
Abstract.
This note is devoted to the study of families of quaternionic modular forms arising from orders defined by Pizer. In this situation, the Hecke-eigenspaces are 2-dimensional contrary to the classical case of Eichler orders. The main result is a Control Theorem in the spirit of Hida, interpolating these 2-dimensional Hecke-eigenspaces. We restrict our attention to a definite rational quaternion algebra ramified at a single odd prime .
2010 Mathematics Subject Classification: 11F11, 11R52.
Key words: Quaternion algebras, Hijikata–Pizer–Shemanske orders, Hida families, control theorems.
1. Introduction and statement of the result
Let be an odd prime, an integer prime to and an integer. Let be the unique, up to isomorphism, quaternion algebra over ramified exactly at and . Take to be an Eichler order of level in and consider the space of -valued cuspidal quaternionic newforms with level , denoted by ; we recall its precise definition in Section 2.6 but, roughly speaking, it represents the space of non-Eisenstein quaternionic modular forms which do not satisfy a lower level structure. The Jacquet–Langlands correspondence ensures an injective transfer between automorphic representations of the algebraic group associated with and , but at the level of automorphic forms it takes an explicit realization, often referred to as the Eichler–Jacquet–Langlands correspondence. More precisely, one obtains the Hecke-equivariant isomorphism . On the other hand, in order to study modular forms with higher level structure at , one needs more general orders, namely the special orders defined by Pizer and Hijikta–Pizer–Shemanske. As the precise definition of special orders does not contribute necessarily to the understanding of this introduction, we prefer to avoid technicalities and postpone it to Definition 2.1.1. In the case where is such a special order in , corresponding to level structure with , the relation between automorphic forms changes considerably. Before stating the precise relation, in order to be consistent with the notation present in [15], let us introduce the following (unfortunately unconventional but rather useful) piece of notation. Throughout the whole document, for any module , we write for the direct sum . The Jacquet–Langlands correspondence, for , takes the explicit form of the following Hecke-isomorphism
| (1.1) |
where two copies of must be taken into account and the sum of the twisted spaces (see beginning of Section 3.3) runs over certain primitive characters modulo . Equation (1.1) has a slightly more complicated expression for , but the above situation is already explanatory of the general phenomenon; an exhaustive statement can be found in Theorem 3.3.2. For any choice of an isomorphism in Eq. (1.1) we see that any classical newform of level , which is twist-minimal at (as in Definition 3.3.4), lifts to two linearly independent quaternionic newforms with the same Hecke-eigenvalues away from the level. The above-mentioned situation has been extensively studied in a series of works by H. Hijikata, A. Pizer, and T. Shemanske, most notably [26] and [15]. In the light of this multiplicity, it is natural to ask whether these quaternionic modular forms still live in -adic families, for an odd prime different from and prime to . In this note we provide a positive answer to this question.
As introduced above, let , and be, in the order, two distinct odd primes and a positive integer prime to both and . For the sake of simplicity, we restrict here to the case of trivial character at and exponent ; we refer the reader to Theorem 4.2.4 for the general statement. As in Definition 2.4 of [9], we consider to be the universal ordinary -adic Hecke algebra of tame level . Considering the Iwasawa algebra , represents the -algebra of Hecke-operators acting on Hida families of tame level . For any continuous group homomorphism , we say that is an arithmetic homomorphism if its restriction to defines a character of the form , for and a -adic character of conductor , . We recall that these homomorphisms are identified with points on the so-called weight space. As usual, we associate to each arithmetic homomorphism the couple . For any we also denote its kernel by and the corresponding localization of by . Let be a classical modular newform in and assume that is twist-minimal at . Moreover, if is -ordinary, we can consider the unique Hida family associated with by the works of Hida and Wiles. For each arithmetic homomorphism we denote by the specialization of at . We set (resp. ) to be the field extension of generated by the Fourier coefficients of (resp. ) and take (resp. ) to be its ring of integers. Fix now to be a maximal order in the quaternion algebra which contains the family of nested orders ,
| (1.2) |
At all places , we fix isomorphisms such that is identified with the upper triangular matrices modulo . For each we consider the compact open subgroup defined as
| (1.3) |
Let be the dual of , the space of homogeneous polynomials in of degree , endowed with the action of , induced by the left multiplication . Denoting for the finite adèles of , we consider, as in Definition 2.4.3, the space of quaternionic -adic modular forms of weight , character (of conductor ) and level ,
| (1.4) |
More generally, let be the subset of primitive vectors in , namely the subset of vectors with at least one component which is not divisible by , and consider the space of -valued measures on . We construct, following Definition 4.1.1, the space of measure-valued quaternionic modular forms
| (1.5) |
for the action of induced by the left multiplication on the variables. By integration, we induce, for any arithmetic homomorphism , a specialization map
| (1.6) |
such that
| (1.7) |
for and as above, and any ; all details can be found in Section 4.1. Considering the ordinary component of , which we denote by , the specialization maps descend to maps between the ordinary components
| (1.8) |
for the subspace of -ordinary quaternionic forms in . As the algebra acts on the space of Hida families, we can consider the -isotypic component
| (1.9) |
that is, the component of where the Hecke-operators act with the same -eigenvalues of . Up to a mild condition on the level , as explained in Remark 4.1.5, one can assume the -module to be free, as we do here. We can hence state our main result under the above simplifying restrictions for the character and the power of ; the general statements are the content of Theorems 4.2.4 and 4.2.5.
Theorem A (Control theorem for special orders).
With the above notation, suppose that is twist-minimal at . For any arithmetic homomorphism , the map (of Proposition 4.2.1, induced by the specialization map) induces an isomorphism of -dimensional -vector spaces
| (1.10) |
The two main ingredients needed for the proof of Theorem A are the above isomorphism (1.1) and the seminal paper [13]. The results proved in [13] for definite quaternion algebras over totally real fields different from remain true in the case of definite quaternion algebras over , as already noticed in Section 3 of [21] and Section 4 of [17], and in the case of special orders, as remarked in Remarks 4.1.5 and 4.2.3. The strategy of the proof is then a generalization of the work [21], from which we take inspiration.
Theorem A extends the foundational results of Hida theory to the case of quaternionic modular forms with special level structure, allowing to consider quaternionic -adic families with tame level over the quaternion algebra , which, we remark, is ramified at . We highlight again that the situation discussed here differs markedly from the classical case of Eichler orders, where all the local non-archimedean automorphic representations at are 1-dimensional. In particular, the novelty of this result lies in the rank of the Hecke-eigenspaces being and no more 1 as in the classical Eichler case.
An additional motivation for such a Control Theorem originates from the desire to study points outside the interpolation region of the triple product -adic -function, as treated in [6]. More precisely, the present study together with [6] are motivated by a conjecture of Bertolini–Seveso–Venerucci and by the wish to provide computational support to it.
The present work leaves open some questions which we briefly discuss in Section 4.3. The main one is whether it is possible to distinguish the 2-dimensional spaces of quaternionic modular forms attached to special orders, first in the case of a classical eigenspace, and then for families.
Acknowledgments: This note presents the content of a section of the author’s doctoral dissertation [7]; he expresses his gratitude to his supervisor Massimo Bertolini. Many thanks go also to Matteo Longo for several helpful discussions, among the others on [21] and [13]. The author is grateful to the anonymous referee for the valuable comments, suggestions, and questions which led to a significant improvement of the exposition. The author also wishes to thank Matteo Tamiozzo, for numerous mathematical conversations, and Jonas Franzel, for reading an early draft of this work and finding out various typos. The author is grateful to the Universität Duisburg–Essen, where the main part of this work has been carried out, as well as to the Università degli Studi di Padova (Research Grant funded by PRIN 2017 “Geometric, algebraic and analytic methods in arithmetic”) for their financial support.
2. Quaternionic orders and modular forms
We begin by recalling the definitions of the various quaternionic orders as well as the definition of quaternionic modular forms, both -adic and classical, for a definite quaternion algebra over . Special orders are a generalization of the classical Eichler orders, which are needed for studying both higher ramification and the presence of a character at primes where the quaternion algebra is ramified. We refer the reader to [15] and [16] for all the details that we are not recalling here.
We fix, once and for all, a choice of field embeddings . For any prime we denote -adic valuation by , normalized so that .
2.1. Special orders
Let the unique, up to isomorphism, quaternion algebra over with discriminant . Fix an isomorphism for each . We denote the reduced norm of by . Let be an odd rational prime, and fix to be a quadratic non residue modulo . The local field has a unique quadratic unramified extension and two quadratic ramified ones, and . For one of these quadratic extensions, we denote by its ring of integers. Set
| (2.1) |
and, for ,
| (2.2) |
We notice that, for , is the unique maximal ideal of .
Definition 2.1.1 ([16], Def. 6.1).
An order in is said to be a special order of level if
-
(i)
is conjugate to by an element of (via ), for each ;
-
(ii)
there exists a quadratic extension of such that is conjugate to , for each .
In the following, we choose for each the isomorphism such that . If in the above definition we take the level to be such that , we obtain the usual definition of an Eichler order of level (see [25], page 344).
From now on, we assume to be odd and we fix a particular choice of special orders and quadratic field extensions. We follow the thorough summary given in Section of [19], which is based on a careful analysis of [15] and takes into account more general orders. Let be any newform in . In order to be able to lift to a quaternionic modular form, we fix the choice of the special order such that the quadratic extensions of and the exponents are as follows. For any (odd) prime , if
-
(1)
is odd, we take to be the unramified extension of and ;
-
(2)
is even, we take to be one of the two ramified extension of and .
The case of even discriminant presents some further difficulties and, as our main case of interest is the case of odd, we omit it and refer the reader to [15] and [19].
We recall a part of the notation already used in the Introduction.
Notation 2.1.2.
Let be the profinite completion of . We set and , where are the finite adèles of .
Special orders satisfy properties similar to the Eichler ones.
Lemma 2.1.3.
is a compact open subgroup of . In fact, this is true for each one of its local components.
Proof.
This lemma is a classical result whenever is an Eichler order (see e.g. Sections 5.1 and 5.2 of [23]). We consider the case of a special order. Lemma 5.1.1 of [23] tells us that is compact in , independently of the order. By definition of special order, it is enough to consider . Since the reduced norm is continuous, is open and thus, as the sum is a continuous homomorphism, we deduce the claim. ∎
Proposition 2.1.4 ([24], Proposition 2.13).
All special orders have finite class number. Moreover, it depends only on the level and not on the specific choice of the special order.
Lemma 2.1.5 ([16], Lemma 7.4).
Let be a special order of level . Then there exists a set of ideal class representatives for the left -ideal classes, such that for all dividing the level.
2.2. Characters
Let be a special order of level and let be a Dirichlet character with conductor .
Assumption 2.2.1.
Assume that for all and, for each , that
| (2.3) |
The choice of character in Assumption 2.2.1 is motivated by the explicit form of the Jacquet–Langlands correspondence recalled in the Introduction (see Section 3.3 for details).
We want to extend to a character of and for this purpose we must deal with several sub-cases. First of all, we decompose by the Chinese Reminder Theorem and we define each as follows.
-
(1)
If and , we set for each .
-
(2)
If and , we set for .
-
(3)
If is odd, and we deal with three further sub-cases:
-
(i)
If and is odd, we can always find a lift to and thus to , which we call ; here is the maximal ideal of . Because of Assumption 2.2.1, is contained in (if is unramified) or in , hence we set for each (and ).
-
(ii)
If , with , and is even, we can always find a character such that and with conductor . As remarked in Section 7.2 of [16], the choice of this character is not important, but the fact that a particular choice is fixed is. We set .
-
(iii)
If , with , and is odd, we write for a fixed choice of characters odd and with , and even. Thus, proceeding analogously to the previous sub-cases, we set .
-
(i)
-
(4)
If , and , one proceeds in a similar way as in case .
Patching together the local lifts, we define
| (2.4) |
for any . In particular, if is a lattice in such that for each , and , we have
| (2.5) |
We refer to [16], Section 7.2 for all the details.
2.3. Quaternionic modular forms of weight
Take as in the above Section 2.1, with a special order of level . Recall that we fixed isomorphisms for each , such that . Set and
| (2.6) |
We extend any character as in the above Section 2.2 to imposing , where is the finite part of .
Definition 2.3.1.
We define the space of weight- quaternionic modular forms with level structure , character satisfying Assumption 2.2.1 and -coefficients, as the -vector space of all continuous functions satisfying
| (2.7) |
for all , and .
As in Chapter 5 of [15], we can decompose as a finite union of distinct double cosets
| (2.8) |
where is the class number of . Since is definite, the analogous decomposition holds for , namely , with . By the above Lemma 2.1.5, the representatives can be taken to lie in , in particular for each prime . If we fix the representatives in this way, we have an explicit description of quaternionic modular forms. By the definition of a quaternionic modular forms and the double coset decomposition, a quaternionic modular form is uniquely determined by its values on the representatives. More precisely, for , let and define
| (2.9) |
As thoroughly explained in loc.cit., the above observations yield the identification
| (2.10) |
given by . We are allowed to consider different coefficients, in fact the above identification still holds when we replace by , the field extension of generated by the values of the character . By extension of scalars we recover and we can consider -adic coefficients , for a prime which does not divide the reduced discriminant of .
Remark 2.3.2.
All the above constructions and definitions are, up to isomorphism, independent of the specific choice of the special order. Moreover, fixing compatible choices of the lifting characters , all the constructions are compatible with respect to the inclusion of special orders.
We end this section with the following fact: often the groups have cardinality , i.e. .
Proposition 2.3.3 ([26], Proposition 5.12).
Let be a special order of level in the quaternion algebra over ramified exactly at and . Then
| (2.11) |
Moreover, if and or is divisible by a prime , then .
2.4. -adic quaternionic modular forms for special orders
Let and be two distinct rational odd primes. From now on, we denote by the (unique up to isomorphism) definite quaternion algebra over with discriminant and we fix to be a maximal order in . For a fixed positive integer, prime to both and , consider a family of nested special orders satisfying
| (2.12) |
where . Up to conjugation, we can suppose that the orders are all canonical orders of level , that is, as in Definition 2.1.1. For any prime different from , we assume that the fixed isomorphism satisfies and
| (2.13) |
Definition 2.4.1.
We define (cf. Lemma 2.1.3) open compact subgroups ,
| (2.14) |
By construction, . Given any special order , we denote by the corresponding compact open subgroup defined analogously to .
For any commutative ring , we consider the left action of on the polynomial ring , defined as
| (2.15) |
for and . We denote by the submodule of consisting of homogeneous polynomials of degree ; by definition, is stable under the action of . Its dual module is endowed with the right action
| (2.16) |
for any and .
We take now to be a finite flat extension of , which we assume to contain all the -th roots of unity, where is Euler’s totient function. Given an -algebra , any -valued Dirichlet character modulo , can be lifted to its adèlization, that is, the unique finite order Hecke character
| (2.17) |
such that , for every and .
We fix such a Dirichlet character modulo with small conductor at . More precisely, as in [15], we enforce the following assumption.
Assumption 2.4.2.
The -component of , , is either the trivial character modulo or an odd character of conductor exactly .
Definition 2.4.3.
Let be an integer and let be as in Definition 2.4.1. For an -algebra and an -valued Dirichlet character (for modulo and modulo ) we define the space of -adic quaternionic modular forms of weight , level and character as
| (2.18) |
This space can be identified with the space of functions satisfying
| (2.19) |
for and , and such that for any and any .
Remark 2.4.4.
The term -adic quaternionic modular forms refers to the fact that we are considering -adic coefficients, and the action at the place instead of the action at infinity (see the next Section 2.5).
2.5. Quaternionic modular forms of higher weight
The definition of classical quaternionic modular forms for higher weight is similar to the one for weight 2. We fix an identification in order to compare -adic and classical quaternionic modular forms.
Definition 2.5.1.
We define the space of weight-, , quaternionic modular forms with level structure , character satisfying Assumption 2.2.1 and -coefficients, as the -vector space of all continuous functions satisfying
| (2.20) |
for all , , and .
2.6. Quaternionic Eisenstein series and newforms
We recall the notions of quaternionic Eisenstein series and quaternionic newforms as presented in [15], Chapters 5 and 7. For this section, we take to be a -module. We begin with the Eisenstein series part of , that is
| (2.22) |
where is the extension of the quaternionic norm to . In other words, is the space of quaternionic modular forms factoring through the reduced norm map. As proved by Propositions 5.2, 5.3 and the discussion after Proposition 5.4 in loc.cit., this space is often trivial, in fact
| (2.23) |
In particular, has at most rank 2. Defining the Petersson inner product as in [27] or [10], one can consider the orthogonal complement of in , namely
| (2.24) |
Definition 2.6.1.
We call the space of -valued cuspidal quaternionic modular forms of level and character .
Inside of this space, we find the so-called space of old forms, , defined to be the subspace of spanned by all for each special order for which makes sense. One should pay attention to fix a suitable ramified extension of , but we point the reader to Remarks 7.13 and 7.14 of [15] for further details. Finally, we define the space of quaternionic newforms as the orthogonal complement of , with respect to the Petersson inner product, inside .
3. Hecke algebras and lifts to quaternionic modular forms
One of Hida’s main results is the extension of the classical duality between the Hecke algebra and the space of classical modular forms to -adic families. The analogous result can be recovered in the quaternionic setting, when one considers Eichler orders or special orders with odd exponent at the primes of ramification, but in the case of special orders with even exponent, this is no more true (see Remark 3.3.3). Even though one cannot speak about duality anymore, it is indeed possible to recover the correct dimension result for proving a rank-2 Hida theory.
3.1. Hecke operators
For any prime , recall the element introduced in Section 2.4, such that and otherwise. Let be again an -algebra and take . On this quaternionic space we have (for any ) the Hecke operators ,
| (3.1) |
and the Hecke operators ,
| (3.2) |
We also consider the quaternionic operator at , which is defined as
| (3.3) |
for such that is a units in the maximal order at of norm , and elsewhere. For each , we also recall the diamond operator with its usual definition on classical modular forms and straightforwardly extended to the -adic quaternionic case. On the space of classical modular forms we have the usual operators with a similar expression to the quaternionic ones except at , where the definition of is analogous to the above operators.
3.2. Hecke algebras
For each , let be the Hecke algebra generated over by all Hecke operators and the diamond operators away from the level, which acts on . We denote by the direct summand of acting on and by the Hecke algebra acting on the space of newforms . For each we have the projection maps and the same holds true for the subalgebras and . We construct the projective limits with respect to these maps,
| (3.4) | and |
together with the projection maps . For any , we define to be ordinary part of , namely the product of all the localizations of on which is invertible, and denote by the corresponding projector . Similarly we define and , together with the corresponding ordinary projectors, which we denote by the same symbol . Passing to the limit we obtain , and , each of them equipped with the corresponding ordinary projector .
Remark 3.2.1.
On the quaternionic side we proceed similarly. For each , let be the Hecke algebra acting on , generated over by the Hecke and diamond operators. We denote by the component acting on the space of newforms . For each we have the projection maps and we construct the projective limits with respect to these maps,
| (3.6) | and |
together with the projection map . In the end, we define as above the ordinary Hecke algebras and , and obtain and as inverse limit of the and respectively.
The Jacquet–Langlands correspondence provides a compatible morphism between the classical and the quaternionic side, that is
| (3.7) |
which preserves the Hecke and diamond operators away from the discriminant of the quaternion algebra.
Let be the finite flat extension of the Iwasawa algebra obtained from . We remark that by construction, the algebra is naturally a -algebra; moreover, one can prove that it is finitely generated over . We define the two universal -adic Hecke algebras
| (3.8) | ||||
| (3.9) |
and, as in [21], we obtain the compatible morphisms
| (3.10) | and |
Definition 3.2.2.
It is useful to formally introduce the notation considered in Remark 3.2.1. More precisely, let be any one of the Hecke algebras defined above and let be any positive integer. We denote by the Hecke-subalgebra of generated by the Hecke and diamond operators away from .
3.3. Quaternionic lifts of modular forms and the failure of the duality
We analyze more carefully [15], recalling the results which we need. Let be the Kronecker character at and , a field extension of . For any space of modular forms and each Dirichlet character modulo , we denote by the space of all the modular forms which are twists by of modular forms in .
Theorem 3.3.1 ([15], Theorem 7.10).
Let be a special order of level (so is the unramified quadratic extension of ). Let be a character modulo such that is either the trivial character modulo or an odd character modulo . Suppose moreover that is even and that . Then there exist a Hecke–equivariant isomorphism
| (3.11) |
The above theorem proves that, as in the case of Eichler orders, there is a one-to-one correspondence for special orders with odd exponent at . The situation for even exponent is more complicated.
Theorem 3.3.2 ([15], Theorems 7.16 & 7.17).
Let be an odd prime, and let and be integers. Let be a character modulo such that , it satisfies Assumptions 2.2.1 and 2.4.2, and such that . Then the following decomposition of -modules holds true.
-
(a)
If and is the trivial character:
(3.12) where the sum runs over all the classes of primitive characters modulo excepting , modulo the equivalence .
-
(b)
If and is a odd character modulo :
(3.13) where is a lift of as in Section 2.2 and the sum runs over all the classes of primitive characters modulo excepting , modulo the equivalence .
-
(c)
If and is either trivial or odd of conductor :
(3.14) where is a lift of as in Section 2.2 and the sum runs over all the classes of primitive characters modulo , modulo the equivalence .
Remark 3.3.3.
-
(a)
In the above theorem the decomposition is given as -modules, but strong multiplicity one for classical modular newforms guarantees the decomposition to hold (at least) as -modules. As already noticed in Remark 3.2.1, the Hecke algebra coincides with since the Hecke operator is the -operator on this space.
-
(b)
The theorem implies that the duality between the Hecke algebra and the space of modular forms does not necessarily hold true for special orders with level . This situation represents the main difference between this setting and the case of classical modular forms (and special orders with an odd power of ). We recall that, on the contrary, the Jacquet–Langlands correspondence does hold true, as well as the multiplicity one result for automorphic representations. This phenomenon is purely local, as already remarked in Example 2.6 of [19]. More precisely, the dimension of the local automorphic representation at is bigger than in the case of level and determined by the minimal conductor of the modular forms. We refer to Section 5 of [4] for all the related details.
We recall the definition of a twist-minimal modular form.
Definition 3.3.4.
A modular form is twist-miminal if for all -adic characters , where is the automorphic representation attached to and for any automophic representation we denote the conductor of .
Corollary 3.3.5.
Each twist-minimal modular eigenform in lifts to (up to linear combinations) exactly two linearly independent quaternionic modular eigenforms in with the same Hecke eigenvalues for .
Regardless of Remark 3.3.3.(b), one can still obtain an isomorphism between the space of quaternionic modular forms and the square of a suitable Hecke algebra, as in the following proposition.
Proposition 3.3.6.
Under the hypotheses of Theorem 3.3.2, there exists a -vector subspace of , which is a -submodule satisfying
| (3.15) |
Moreover, for the Hecke-subalgebra of acting on , we have an isomorphism of -modules,
| (3.16) |
Proof.
The first statement follows directly from Theorem 3.3.2 as noticed in Chapter 8 of [15]. The second part follows from where the first isomorphism is due to the properties of and the second is the Hecke-duality for classical modular forms restricted to (since the decomposition is Hecke-equivariant away from ). ∎
As in Section 3.2, taken an -algebra, we define and . We obtain injective homomorphisms and .
Definition 3.3.7.
For any module with an action of a suitable Hecke algebra, and any classical eigenform , we denote by the -isotypic component of , i.e. the biggest submodule of on which the Hecke algebra acts with the same eigenvalues of .
Proposition 3.3.8.
Let be a newform in with and . Write for and the component of , respectively, modulo and .
-
(a)
If and is the trivial character modulo ,
(3.17) -
(b)
If either is a non-trivial character modulo or ,
(3.18)
Proof.
This is a straightforward consequence of Theorem 3.3.2 combined with the fact that strong multiplicity one applies to . ∎
3.4. Choice of a modular form
Let be a fixed -ordinary newform, for and with a Dirichlet character modulo satisfying Assumptions 2.2.1 and 2.4.2. In this way, the automorphic representation associated with admits a Jacquet–Langlands lift. Moreover, we assume that the -adic Galois representation associated with is residually absolutely irreducible and -distinguished. Let be the finite extension of defined by and take to be its ring of integers; note that is a finite flat extension of . We denote by the unique Hida family passing through . By duality with the ordinary Hecke algebra, we know that defines a character, which we denote with the same symbol ,
| (3.19) |
where is the universal ordinary -adic Hecke algebra of tame level as in Definition 2.4 of [9]. The Jacquet–Langlands correspondence ensures that such character factors through the morphism to ; we keep denoting the corresponding map by .
4. The control theorem
In this last section, we prove a control theorem for special orders of even conductor at . We introduce a space that is suitable for the -adic interpolation and we define some specialization maps. We consider again to be the ring of integers of a fixed finite extension of and we take an -algebra which we denote again by .
4.1. Specialization maps
Let be the set of primitive row vectors, that is, the vectors in which have at least one component not divisible by . Denote by the space of -valued continuous functions on and by the space of -valued measures on . We have a left -action on via
| (4.1) |
for and , and the induced right action on as
| (4.2) |
for . Considering the action by , with , we can notice that the subspace satisfies
| (4.3) |
Definition 4.1.1.
This space can be identified with the space of functions satisfying
| (4.5) |
for and , and such that for any and any .
Take and let be any character which factors through . We extend multiplicatively to imposing . We define the specialization map
| (4.6) |
such that
| (4.7) |
where , , and .
Proposition 4.1.2.
The specialization maps are well-defined and Hecke-equivariant for , where the equivariance at is meant as .
Proof.
Let . Then, for any , , and , we have
| (4.8) |
and, since , is extended to and , we obtain
| (4.9) |
The equivariance with respect to the operators is obvious, as well as that for the operators with (also for ). To prove the equivariance at it is enough to note that we have
| (4.10) |
for the characteristic function of . ∎
We must now investigate the properties of the space . We begin by noticing that the action on is exactly the one induced by the right action, defined by right multiplication, of on . We proceed similarly to Proposition 7.5 of [20] or Chapter 6 of [9] and, denoting by the set of primitive vectors in , we recover , with respect to the canonical projection maps. We obtain then (see e.g. Section 7 of [22]). Since is a finite set, is identified with the space of step functions. The action of on is transitive and the stabilizer of is
| (4.11) |
This shows that and then that . We denote by the profinite double quotient associated with , for the ring of finite adèles away from . By Shapiro’s Lemma we obtain
| (4.12) |
as well as the analogous isomorphism when we consider a character . Equation (4.12) implies that
| (4.13) |
where the identification is -equivariant. We hence deduce that is a compact -module, since is -adically complete and each is a finitely generated free -module. This allows us to define its ordinary part as usual (see Section 2.4 of [21] and the references therein) as its direct summand on which the Hecke operator acts invertibly. We shorten the notation and denote by the space . In particular, the Hecke-equivariance in the inverse limit construction of , implies that , where is replaced by on each component of the inverse limit. Proposition 4.1.2 shows that the specialization maps descend to Hecke-equivariant specialization maps between the ordinary components,
| (4.14) |
with the same definition of and where is the finite extension of generated by the values of .
Notation 4.1.3.
We need to introduce some more notation.
-
(a)
For any and any character , let such that
(4.15) In particular, is homogeneous of degree for each .
-
(b)
Let be the extension of the classical Iwasawa algebra , obtained by . We set for any -algebra .
-
(c)
We say that a homomorphism is an arithmetic homomorphism if its restriction to is of the form for and a character which factors through , with minimal. In this situation, we say that has weight and character of conductor .
Lemma 4.1.4.
Let be an arithmetic homomorphism of weight and character of conductor . Let be the field extension of containing the values of . The map induces the injective Hecke-equivariant homomorphism
| (4.16) |
where is the kernel of in .
Proof.
We begin noting that , as it can be seen by applying twice Lemma 1.2 of [1] to . We prove now that . Let ; therefore lies in for any . Lemma 6.3 of [9] shows that if and only if for each homogeneous function of degree . For each , is homogeneous of degree and hence for each and . Take now and let . Since is invertible, let be such that . Let for and . The definition of the Hecke operator in Section 3.1 does come from the coset decomposition . Then corresponds to a decomposition of the form , where each is a product of matrices , for . We compute
| (4.17) |
For each , thus, whenever contains a copy of . Therefore, only the matrices contribute to the integral and we recognize that
| (4.18) |
By construction, and since is invertible on the space , and hence . Lemma 6.3 of [9] implies that . ∎
Remark 4.1.5.
As in the case of Eichler orders, the space of quaternionic modular forms is often finitely generated over and free, as it follows from the discussion in Section 2.3. In particular, this holds true under the conditions discussed in Proposition 2.3.3. Therefore, Lemma 4.1.4 implies that is -finitely generated and free. The discussion in Section 2.3 shows also that is often -free and finitely generated, once again, for example under the conditions in Proposition 2.3.3. One can argue as in the proofs of Theorem 10.1, Corollary 10.3, and Corollary 10.4 of [13], since the results proved there for quaternionic modular forms over definite quaternion algebras hold in more generality for special orders which are split at the interpolation prime (see also Remark 4.2.3).
4.2. The proof of the control theorem
As in Section 3.4, we fix a -ordinary newform in , for and with a Dirichlet character modulo satisfying Assumptions 2.2.1 and 2.4.2. We also assume that its associated -adic Galois representations is residually absolutely irreducible and -distinguished. Let the homomorphism associated in Section 3.4 with the Hida family passing through . Recall that and that is its ring of integers. For any as in the above Lemma 4.1.4, we denote by the composition
| (4.19) |
where the first map is the compatible morphism of Section 3.2 and the last map is the one to the localization of at the prime . We recall that is a -algebra; we identify any -algebra as a -algebra via the inclusion and write
| (4.20) |
for the isotypic component of the -module , where the Hecke operators act as determined by .
Proposition 4.2.1.
With the notation of Lemma 4.1.4, there is an induced injective homomorphism
| (4.21) |
for the weight- specialization of .
Proof.
The proof is the same as of Proposition 3.5 in [21], since it does not depend on the choice of the quaternionic order. ∎
As one can note from Theorem 3.3.2, the case of level and trivial character has to be handled with more care. The theory of Hida families for classical modular forms is well known and we can restrict our attention to the Hecke-submodules with a Dirichlet character modulo , with . We do not provide details here, but we refer to Chapter 7 of [14] and Section 2 of [21]. We construct the space of -adic modular newforms, level and character , as . Moreover, we can twist its Hecke action by the character obtaining the corresponding space . As in Section 3.2 we have an action of the universal Hecke algebra on . In particular, taking in , the module is a free rank-1 -module (see Proposition 2.17 and the proof of Theorem 2.18 in [21]).
Lemma 4.2.2.
Assume to be -free and finitely generated (see Remark 4.1.5). Suppose that and set, for any arithmetic homomorphism ,
| (4.22) |
where we let act on via the homomorphism induced by the Jacquet–Langlands correspondence. Then is a free rank-2 -module.
Proof.
We start dealing with the case . We consider the -divisible abelian group (cf. Remark 4.1.5 and Section 2.3) , where the inductive limit is taken with respect to the restriction maps induced by the inclusions . The Hecke and diamond operators (at least away from ) act on since, as in the case of Eichler orders, the restriction maps in [13] (see Eqs. (2.9a), (2.9b) and (3.5)) are compatible with the Hecke action. Taking the Pontryagin dual of we obtain the Hecke-equivariant isomorphism (cf. Eqs. 4.12 and 4.13), which shows it is a free -module of finite rank. We denote by the field extension of generated by the values of and by its ring of integers, which we can assume to be finite flat over . Up to a scalar and up to taking the tensor product by , we can suppose to have coefficients in . We observe that , as the action of the Hecke algebra is on the first component and the tensor product is just an extension of scalars. We can hence apply Theorem 9.4 of [13] (cf. Remark 4.2.3) to and obtain the isomorphism of -modules,
| (4.23) |
We remark that the last Hecke-equivariant isomorphism in the above Eq. (4.23) (as well as in Eq. (4.24)), comes from the restriction to of the Pontryagin duality established in Lemma 7.1 of [11]; under the hypotheses of Lemma 2.3.3 one has the isomorphism , as in the proof of Theorem 10.1 in [13], and then Proposition 3.3.6 recovers the needed Hecke-isomorphism for quaternionic modular forms. Similarly to the above discussion for , we can follow Section 2 of [21] and construct the interpolation module , relative to the ordinary subspaces . We notice that under the hypothesis of Proposition 2.3.3, the space is free of finite rank. In particular, we can reproduce the above chain of isomorphisms and obtain -isomorphisms
| (4.24) |
Applying Propositions 3.3.6 and 3.3.8, we deduce the isomorphism of -modules,
| (4.25) |
Tensoring over with , we obtain the isomorphism of -modules,
| (4.26) |
As in the proof of Theorem 2.18 of [21], Proposition 2.17 of loc.cit. guarantees that is a free -module of rank 1, therefore is a free -module of rank 2.
The case of and trivial character at is carried out similarly, once we define the -divisible abelian group
| (4.27) |
whose Pontryagin dual is . ∎
Remark 4.2.3.
-
(a)
The congruence subgroup we consider, away from , is the one denoted by in [13] and one passes from this choice to the one used there by changing all the actions via .
-
(b)
We point out that Theorem 9.4 of [13] is stated under more strict hypotheses but, in the case of definite quaternion algebras, such hypotheses can be relaxed; this has been already noticed in [21] and [17] in order to work with Eichler orders for algebras over , but Theorem 9.4 of [13] holds true also for special orders. This is due to the degree of generality in which the results of Chapter 8 of [13] are proved (as well as Lemma 2.1.3), together with the necessity of a controlled behavior only at the interpolation prime . Let us remark that we did not take into account the case of indefinite algebras, but that it seems to require a generalization of the spectral sequences approach contained in Chapter 9 of [13].
We can finally state the sought for Hida control theorem in the case of special orders of level .
Theorem 4.2.4 (Control theorem for special orders).
With the above notation, suppose to be twist-minimal at . For any arithmetic homomorphism , the map of Proposition 4.2.1 induces an isomorphism of -dimensional -vector spaces
| (4.28) |
If , has trivial character at and lies in (in particular, it is not twist-minimal at ), then the above isomorphism still holds, but the -vector spaces are -dimensional.
Proof.
Suppose to be twist-minimal. Because of Propositions 4.2.1 and 3.3.8 we know that
| (4.29) |
and thus it is enough to prove the opposite inequality. Lemma 4.2.2 shows that is a free -module of rank . The case of , trivial character at and follows similarly. The remaining case accounts to the fact that the Jacquet–Langlands correspondence preserves twists. ∎
We can consider the finitely generated -module
| (4.30) |
Proceeding similarly as in the proof of the above theorem we notice that is a 2-dimensional -vector space, where is the finite field extension of called the primitive component associated with the Hida family (see Section 3 in [12] in particular, Theorem 3.5 and also Theorem 2.6a of [9]). As noticed in Section 2.2 of [21], we point out that is the integral closure of in . We can then formulate Theorem 4.2.4 highlighting this global -module.
Theorem 4.2.5.
With the above notation, suppose to be twist-minimal at . For any arithmetic homomorphism , the map of Proposition 4.2.1 induces an isomorphism of -dimensional -vector spaces
| (4.31) |
If , has trivial character at and lies in (in particular, it is not twist-minimal at ), then the isomorphism of -dimensional -vector spaces holds:
| (4.32) |
Corollary 4.2.6.
Let be a primitive Hida family of tame level , , tame character with its -component, , as in Assumption 2.4.2. Suppose moreover to be twist-minimal at . Then there exist two -linearly independent elements and in , which form a basis for . Moreover, for any arithmetic homomorphism , and form a -basis for .
Definition 4.2.7.
We denote by the -linear span of and and call it the subspace of special quaternionic Hida families associated with .
4.3. A small remark on related works and open questions
The mathematical literature about this situation of higher ramification at the primes at which the quaternion algebra ramifies is quite meager. Excluding the (singular and collective) works of Pizer, Hijikata and Shemanske, there are few other works considering special orders and they all share working with indefinite algebras. We already referred to [19], but we wish to point the reader’s attention also to the two works [5] and [8]. In particular, in [26], Pizer defines certain local operators acting on the quaternionic modular forms. The present note leaves unanswered whether the two linearly independent quaternionic modular forms, and then the two Hida families, can be distinguished via some of these local operators. We wish to address carefully this question in the near future.
References
- [1] Anver Ash and Glenn Stevens. p-adic deformations of cohomology classes of subgroups of . Collectanea Mathematica, 48:1–30, 1997.
- [2] Massimo Bertolini, Henri Darmon, and Adrian Iovita. Families of automorphic forms on definite quaternion algebras and Teitelbaum’s conjecture. Astérisque, (331):29–64, 2010.
- [3] Kevin Buzzard. On p-adic Families of Automorphic Forms, pages 23–44. Birkhäuser Basel, Basel, 2004.
- [4] Henri Carayol. Représentations cuspidales du groupe linéaire. Annales scientifiques de l’École Normale Supérieure, 4e série, 17(2):191–225, 1984.
- [5] Miriam Ciavarella. Congruences between modular forms and related modules. Funct. Approx. Comment. Math., 41(part 1):55–70, 2009.
- [6] Luca Dall’Ava. Approximations of the balanced triple product -adic -function. submitted, 2022.
- [7] Luca Dall’Ava. Quaternionic Hida families and the triple product -adic -function. PhD thesis, Universtät Duisburg-Essen, 2021. available at https://duepublico2.uni-due.de/receive/duepublico_mods_00074866.
- [8] Carlos de Vera-Piquero. The Shimura covering of a Shimura curve: automorphisms and étale subcoverings. J. Number Theory, 133(10):3500–3516, 2013.
- [9] Ralph Greenberg and Glenn Stevens. -adic -functions and -adic periods of modular forms. Inventiones mathematicae, 111(2):407–448, 1993.
- [10] Benedict H. Gross. Heights and the special values of -series. In Number theory (Montreal, Que., 1985), volume 7 of CMS Conf. Proc., pages 115–187. Amer. Math. Soc., Providence, RI, 1987.
- [11] Haruzo Hida. Galois representations into attached to ordinary cusp forms. Inventiones mathematicae, 85:545–614, 1986.
- [12] Haruzo Hida. Iwasawa modules attached to congruences of cusp forms. Annales scientifiques de l’École Normale Supérieure, Ser. 4, 19(2):231–273, 1986.
- [13] Haruzo Hida. On -adic Hecke algebras for over totally real fields. Annals of Mathematics, 128(2):295–384, 1988.
- [14] Haruzo Hida. Elementary Theory of L-functions and Eisenstein Series. London Mathematical Society Student Texts. Cambridge University Press, 1993.
- [15] Hiroaki Hijikata, Arnold Pizer, and Thomas R. Shemanske. The basis problem for modular forms on . Mem. Amer. Math. Soc., 82(418):vi+159, 1989.
- [16] Hiroaki Hijikata, Arnold Pizer, and Thomas R. Shemanske. Orders in quaternion algebras. Journal für die reine und angewandte Mathematik, 394:59–106, 1989.
- [17] Ming-Lun Hsieh. Hida families and -adic triple product -functions. Amer. J. Math., 143(2):411–532, 2021.
- [18] Wen Ch’ing Winnie Li. Newforms and functional equations. Math. Ann., 212:285–315, 1975.
- [19] Matteo Longo, Víctor Rotger, and Carlos de Vera-Piquero. Heegner points on Hijikata-Pizer-Shemanske curves and the Birch and Swinnerton-Dyer conjecture. Publ. Mat., 62(2):355–396, 2018.
- [20] Matteo Longo, Victor Rotger, and Stefano Vigni. On rigid analytic uniformizations of Jacobians of Shimura curves. American Journal of Mathematics, 134(5):1197–1246, 2012.
- [21] Matteo Longo and Stefano Vigni. A note on control theorems for quaternionic Hida families of modular forms. Int. J. Number Theory, 8(6):1425–1462, 2012.
- [22] Barry C. Mazur and Peter Swinnerton-Dyer. Arithmetic of Weil curves. Inventiones mathematicae, 25(1):1–61, Mar 1974.
- [23] Toshitsune Miyake and Yoshitaka Maeda. Modular Forms. Springer Monographs in Mathematics. Springer Berlin Heidelberg, 2006.
- [24] Arnold Pizer. The action of the canonical involution on modular forms of weight on . Math. Ann., 226(2):99–116, 1977.
- [25] Arnold Pizer. An algorithm for computing modular forms on . Journal of Algebra, 64(2):340 – 390, 1980.
- [26] Arnold Pizer. Theta series and modular forms of level . Compositio Math., 40(2):177–241, 1980.
- [27] Hideo Shimizu. On zeta functions of quaternion algebras. Annals of Mathematics, 81(1):166–193, 1965.
L. Dall’Ava, Dipartimento di Matematica, Università degli Studi di Padova, Padova, Italy.
E-mail address: luca.dallava@math.unipd.it