Eichler-Shimura theory for general congruence subgroups
Abstract
This article aims to extend the Eichler-Shimura isomorphism theorem for cusp forms on a congruence subgroup of to the whole space of modular forms. To regularize the Eichler-Shimura integral at infinity we reconstruct the standard formalism of the Eichler-Shimura theory starting from an abstract differential definition of the integral. The article also provides new proof of Hida’s twisted Eichler-Shimura isomorphism for the modular forms with nebentypus using the isomorphism theorem mentioned above. We compute the Eichler-Shimura homomorphism for a family of Eisenstein series on and prove algebraicity results for the cohomology classes attached to a basis of the space of Eisenstein series on a general congruence subgroup.
MSC2020: Primary 11F67; Secondary 11F75
Keywords: Eichler-Shimura isomorphism, congruence subgroups
1 Introduction
Let be an integer and be a finite index subgroup of . Suppose that , resp. , is the -vector space of modular forms, resp. cusp forms, of weight on . Set . Let be the -subspace of polynomials having degree . The classical Eichler-Shimura theory constructs a -linear Hecke equivariant homomorphism that formally extends to a -linear Hecke equivariant map
(1.1) |
The Eichler-Shimura isomorphism theorem for the cusp forms [16] asserts that the map (1.1) is injective and its image equals the parabolic subspace of . Hida [10] employed the Eisenstein series with nebentypus to establish a version of the Eichler-Shimura isomorphism for the spaces of modular forms with nebentypus. In general, his investigations suggest that for and the injection (1.1) should extend to a Hecke equivariant isomorphism
(1.2) |
One can also use Zagier’s approach towards the period polynomials of Eisenstein series [19] to formulate and prove a holomorphic variant of (1.2) on these subgroups [14]. The modular symbols provide a homological approach to this problem which allows us to establish an analogue of these results for [3]. The main result of the article is as follows:
Theorem 1.1.
The statement above furnishes a modular form avatar of the general theorems on the cohomology of arithmetic groups that describe the cohomology as a sum of the cuspidal and Eisenstein subspaces [9]. Our work provides complete proof of the result in the spirit of the Eichler-Shimura theory of cusp forms. A usual treatment of the isomorphism theorem for cusp forms involves Eichler-Shimura integrals with endpoints at the boundary of the upper half plane and its convergence requires the cuspidal decay condition. Thus a naive attempt to generalize the constructions to the whole space of modular forms runs into issues regarding convergence. To circumvent this problem we develop our formalism in terms of a differential variant of the classical Eichler-Shimura integral that appears in literature in the context of weakly holomorphic modular forms [4].
Let be the upper half-plane. Suppose that is the -algebra of the holomorphic functions on . Now suppose is an integer and . A weight Eichler-Shimura integral of is a polynomial of degree with coefficients in so that
(1.3) |
The differential equation above always has a solution determined up to an element in . The familiar Eichler-Shimura integral
(1.4) |
provides a canonical solution of (1.3). One can develop a working Eichler-Shimura theory by considering Eichler-Shimura integrals only of the form (1.4); see [10], [15]. The advantage of the differential definition is that it allows us to talk about Eichler-Shimura integrals with the base point at infinity whenever lies in a suitable growth class containing the weakly holomorphic modular forms. Note that the definition of the Eichler-Shimura integral does not require any transformation property for the test function . However, the action of on the right-hand side of (1.3) gives rise to a transformation formula for the Eichler-Shimura integral on the left-hand side of the equation. This approach leads to a new equivariant formulation of the period functions where the group has a nontrivial action on both the test function and module of coefficients. Our formalism allows us to bypass the subtle convergence issues involved in the change of variables of integrals having endpoints at the boundary.
Let be a finite index subgroup of and be a subset of such that is closed under the action of on . Given a choice of Eichler-Shimura integrals one defines a period function
(1.5) |
where is the collection of all set-theoretic maps. Now has a natural right module structure stemming from -actions on and ; see (1.7). The period function is a cocycle on with values in . Moreover the image of in , denoted , is independent of the choice of Eichler-Shimura integrals. We call the Eichler-Shimura class of weight associated with and (Section 2.2). The Eichler-Shimura classes are compatible with the restriction of the subgroup as well as the subset of test functions. In particular, every Eichler-Shimura class arises as the restriction of the global class
Now suppose is a trivial -set, i.e., consists of -invariant functions. Then there is a natural isomorphism
see Section 4.1. We define the Eichler-Shimura homomorphism
to be the image of under the isomorphism mentioned above. This construction furnishes the desired extension of the Eichler-Shimura map promised in Theorem 1.1. The properties of the equivariant period function (1.5) yield a direct proof of the Hecke-equivariance of this map. We also distill another simple yet neglected equivariance property of that plays a crucial role in the proof of the isomorphism theorem. Let the notation be as in the discussion above. Moreover, suppose that is a subgroup of that contains as a normal subgroup and is closed under action of . Note that acts on by conjugation.
Proposition 1.2.
(Proposition 4.7) The Eichler-Shimura map is compatible with the action of on its domain and codomain.
To perform effective computations with the period polynomials one needs to work with a choice of Eichler-Shimura integral that amounts to (1.4) with replaced by . We treat this problem for a class of holomorphic functions that admit decomposition of the form where is an exponential polynomial and is a function with cuspidal decay at (Section 3). Given such a function one can write down a solution of (1.3) as follows:
(1.6) |
Let be a -stable subset consisting of functions with a suitable growth rate. Then our general formalism ensures that the choice of Eichler-Shimura integrals (1.6) recovers the same Eichler-Shimura homomorphism. However to determine the period function attached to (1.6) one needs to discuss the transformation of the integrals having an endpoint on the boundary of the upper-half plane. We handle the convergence problem at the boundary by introducing a family of punctured neighborhoods around boundary points that resemble the Satake topology for the Baily-Borel compactification and transform well under the action of . This strategy to regularize the Eichler-Shimura integral provides an upper half-plane analog of Brown’s construction of period polynomials where he interprets (1.6) in the light of a tangential basepoint [5, Part I]. Readers may also note that our choice of compactification differs from the standard automorphic method [9] that prefers the Borel-Serre compactification over the Baily-Borel approach.
We end the introduction with a glimpse into the isomorphism theorem and its consequences. Let the notation be as in the statement of Theorem 1.1. Recall that where is the space of Eisenstein series. To prove the theorem it suffices to check that the Eichler-Shimura map attached to is injective and the image of in is a linear complement of the parabolic subspace. The main ingredient of the argument is an explicit basis for the space of Eisenstein series that is parameterized by the cusps of and satisfies an appropriate nonvanishing property at the cusps (Section 5.3). Our treatment makes free use of the abstract group cohomological techniques developed in the books of Shimura [16] and Hida [10], but, we completely bypass the Haberland pairing between cocycles. There is a perspective towards the Eichler-Shimura isomorphism that appeals to Shapiro’s lemma to transform the problem into an assertion regarding the cohomology of the full modular group with coefficients in an induced module [14]. Our version of the Eichler-Shimura theorem directly implies the corresponding isomorphism in the induced picture.
Theorem 1.3.
(Theorem 7.5) Let be a congruence subgroup of . There exists an induced Eichler-Shimura homomorphism
extending the map for the cusp forms so that the extended map gives rise to a Hecke equivariant isomorphism .
Hida developed a theory of twisted Eichler-Shimura isomorphism [10, 6.3] in the context of modular forms with nebentypus. The equivariant formalism proposed in this article provides a natural framework for the twisted Eichler-Shimura theory (Section 7.2). Let be a Dirichlet character modulo . Suppose that , resp. , is the space of all modular forms, resp. cusp forms, on that transform according to the nebentypus character under . As before set . Here one needs to replace the module of coefficients by its twisted counterpart, denoted , on which action is twisted by the character . We use the equivariant period function (1.5) to extend the Eichler-Shimura isomorphism for cusp forms and provide the following isomorphism statement strengthening Hida’s earlier result for primitive nebentypus character and prime power level (loc.cit., Theorem 3). Our proof of the result only requires Theorem 1.1 and we do not appeal to the existence Eisenstein series with nebentypus contrary to Hida’s approach; cf. [15].
Theorem 1.4.
(Theorem 7.8) Let the notation be as above. The twisted Eichler-Shimura map for cusp forms extends to a Hecke equivariant isomorphism
The final topic of the article is an explicit computation of the period polynomials attached to the Eisenstein series (Section 8). Recall that if is a subfield of and is a finite index subgroup then the natural inclusion induces an isomorphism ; see (5.4). A class is definable over the subfield if lies in the canonical -form . Now suppose is a congruence subgroup of . Let denote the basis of the space of Eisenstein series described in Section 5.3.
Theorem 1.5.
(Theorem 8.8) Let be a congruence subgroup containing and . Then the image of under the Eichler-Shimura map for is definable over .
The algebraicity theorem above apparently contradicts the fact that the coefficients of the period polynomial of an Eisenstein series involve transcendental expressions closely related to the zeta values; see [19] and Lemma 8.5. However, an observation of Gangl and Zagier suggests that one can modify the period polynomial of the Eisenstein series on by an appropriate coboundary so that the modified cocycle has coefficients in [20, 7]. We first establish a generalization of their observation for and then use the equivariance property in Proposition 1.2 to deduce the general statement. We also prove a similar rationality result (Theorem 8.11) for the twisted Eichler-Shimura isomorphism in Theorem 1.4.
1.1 Notation and conventions
Throughout the article and are variables with values in the upper half-plane . Let denote the imaginary unit. For one writes where and are real numbers with . We normalize the complex exponential function as . One uses to refer to a field of characteristic which in some places is assumed to be a subfield of . For a positive integer set
Let , resp. , be the space of all -valued holomorphic functions, resp. holomorphic differential -forms, on . The constant functions define a natural inclusion which turns into a commutative -algebra. We have a -linear differential operator
whose kernel equals the subspace of constant functions in . During the calculations, one considers integrals along paths on . All paths are piecewise smooth unless otherwise stated.
We abbreviate as and write
for the generators of this group. Sometimes one may refer to the subgroups of . For a subring of set . The letter stands for an arbitrary matrix and we put where take values in an appropriate ring. One uses the notation , resp. , to denote the set of cusps, resp. regular cusps, attached to a finite index subgroup of . Our treatment also makes use of the standard congruence subgroups namely , , and [7, 1.2]. We call the principal congruence subgroup of level . For a subgroup of let
denote the projectivization of .
The group actions relevant to us are usually right actions. In the context of group cohomology , resp. , denotes the collection of -cocycles, -coboundaries. If are two sets then is the set of all set-theoretic functions from to . Suppose that is a -vector space for some field . Then possesses a -linear structure defined by pointwise addition and scalar multiplication. Now suppose is an abstract group and , are right -sets. The set admits a right -action given by
(1.7) |
If is a right -module then the resulting -linear structure and -structure turn into a right -module.
2 Construction of period cocycle
We begin by recalling some background relevant to the Eichler-Shimura theory. The following subsection generalizes the notion of the period polynomial to not necessarily invariant functions. This abstract formulation leads us to the concept of the Eichler-Shimura class that effectively captures our equivariant period function.
2.1 Preliminaries
Group action on .
The group acts on from left by holomorphic automorphisms via fractional linear transformations;
(2.1) |
The restriction of this action to is our primary interest. The -action on descends to a -action on , i.e, acts trivially on each point. Note that the fractional linear transformations (2.1) define a holomorphic left -action on the Riemann sphere . The boundary of in equals and the -action on preserves this boundary. We define the extended upper-half plane as
For let denote the stabilizer of in . With this notation . If is a subgroup of then set . Now suppose is a finite index subgroup. Then a cusp of refers to a -orbit in .
Group action on function spaces.
Let . We define the right action of on by
(2.2) |
As before one primarily uses restriction of this action to . If remains fixed throughout the discussion then we drop it from the notation.
Representations of .
Let be a nonnegative integer. Set . The group acts on from right by
(2.3) |
We omit the index of weight from the subscript if it is clear from the context. More generally, let be a commutative ring and be a commutative -algebra. Write . Then (2.3) defines a right action on . The pair yields a module structure on . In this article, is typically a field of characteristic zero and we consider the cohomology vector spaces for the subgroups of .
Differential operators.
The action of on the coefficients gives rise to a -linear differential operator
(2.4) |
where . The kernel of (2.4) is equal to .
2.2 Period cocycle and Eichler-Shimura class
Definition 2.1.
Let be an integer and . A weight Eichler-Shimura integral of is a polynomial that satisfies .
Now suppose is a collection of holomorphic functions on . A choice of weight Eichler-Shimura integrals for is a set-theoretic function
Note that two choice functions, say and , give rise to a set map
Example 2.2.
Let . The canonical choice for with respect to is
This choice function plays a key role throughout the paper; cf. (1.4).
Let be a finite index subgroup of . Suppose that is a subset of which is stable under the action of . The -actions on and induce a right module structure on according to the recipe (1.7). Let be a choice of weight Eichler-Shimura integrals for . For and put
(2.5) |
We now have the following result:
Lemma 2.3.
.
Proof.
Definition 2.4.
Let the notation be as above. The period function of the choice of Eichler-Shimura integrals is a function on given by
Example 2.5.
Let be the period function of . Then (2.6) amounts to the identity
(2.9) |
The following lemma checks that the period function is a cocycle.
Lemma 2.6.
Let be a choice of Eichler-Shimura integrals for and be the period function of . Then is a cocycle on with values in , i.e.,
(2.10) |
Proof.
We start by proving the statement for . Let and . One employs (2.9) and the change of variable formula to discover that
The identity above together with (2.9) implies that
(2.11) |
Therefore (2.10) holds for . Now the equation (2.8) translates into
(2.12) |
The assertion is a direct consequence of (2.11) and (2.12). ∎
We next study the effect of changing the choice function.
Proposition 2.7.
The image of in is independent of the choice function .
Proof.
Let be a fixed base point and be an arbitrary choice function. The identity (2.12) implies that and differ by the coboundary defined by . Therefore the images of and in cohomology are equal. ∎
Proposition 2.7 proves that the pair determines the image of .
Definition 2.8.
The Eichler-Shimura class of weight attached to , denoted , is the image of in .
The Eichler-Shimura class exhibits good functorial properties concerning restriction of the corresponding set of functions and subgroup; cf. [18, 3]. Let be two -stable subsets of . Now restriction induces a -linear homomorphism that gives rise to a -linear map
Let be a choice of weight Eichler-Shimura integrals for . Set . One computes the corresponding Eichler-Shimura classes with these compatible choice functions to discover that
(2.13) |
Now suppose are two finite index subgroups of . Let be a -stable subset of . The restriction yields change of group map
As before one calculates the Eichler-Shimura classes using compatible choice functions to find
(2.14) |
Global Eichler-Shimura class.
2.3 Variations of formalism
2.3.1 General subgroups of
The constructions of this section are completely algebraic and continue to work for an arbitrary subgroup of . Let be a subgroup of and be stable under action of . Note that (2.3) defines a right -module structure on . Therefore has a natural right -module structure. A straightforward calculation using (2.2) demonstrates that (2.6) and Lemma 2.3 hold in this setup. As a consequence, one can define the period function using (2.5). The rest of the treatment goes through without any change and we obtain a well-defined Eichler-Shimura class in . The general setup described above yields a formalism for Fuchsian groups, cf. [16, 8]. The subgroups of appear in the context of the action of double coset operators. The cocycle formula for the subgroups of is relevant to our treatment of Hecke theory.
2.3.2 Linear cocycles and projectivization
Our formalism of period function allows the choice of Eichler-Shimura integrals to be a set-theoretic function. However, the standard definitions of the Eichler-Shimura integral in literature force to be linear. We now sketch a variant of our construction that incorporates the linearity condition. Let the notation be as in Section 2.2. Suppose that is a subfield of and is a -submodule of . A choice of Eichler-Shimura integrals for , say , is -linear if is a -linear map.
Example 2.9.
The canonical choice is a -linear choice of Eichler-Shimura integrals.
If and are two -linear choices for then is a -linear map. Let be a -linear choice of Eichler-Shimura integrals and be the period function of . It is clear that
for each . |
Note that is a -submodule of . Hence is a cocycle with values in . The image of in is independent of the -linear choice function and determines the -linear Eichler-Shimura class . The natural homomorphism in cohomology induced by the inclusion of into maps to .
For the classical Eichler-Shimura theory of modular forms the linear formalism adds no new structure; see Proposition 4.2. The advantage of linear theory is that it allows us to construct the equivariant period cocycle on the projectivization of . If is odd then the action of on does not descend to . On the other hand the action of on always descends to . Moreover, we have the following lemma:
Lemma 2.10.
Assume that . Let be a -linear choice function for and be the period function of . Then
Proof.
If then the natural projection map identifies the group with its projectivization. On the other hand, if then Lemma 2.10 demonstrates that the linear cocycles on gives rise to well-defined cocycles on . Thus, in both the cases, one obtains a projective Eichler-Shimura class
so that its pullback under equals .
3 Base point at the boundary
The current section constructs Eichler-Shimura integrals with the base point at infinity for a class of holomorphic functions that satisfies an appropriate growth rate at the boundary of the upper half-plane.
3.1 Growth conditions at the boundary
The closure of in the Riemann sphere equals . We endow the extended upper half-plane with the subspace topology arising from because it is favorable for discussing path integrals. In the theory of Baily-Borel compactification, one extends the topology of to by prescribing special types of neighborhoods for the points of that define the so-called Satake topology [8, 2.1]. We next introduce the notion of triangular neighborhoods of the points of which capture the punctured neighborhoods for the Satake topology. Our definition of the growth classes relies on the behavior of test functions on these subsets.
An admissible triangle based at refers to a geodesic triangle on whose one vertex is at and two other vertices lie in .
Definition 3.1.
A triangular neighborhood of is the intersection of an open neighborhood of in with the interior of an admissible triangle based at .
Note that if is an admissible triangle based at then the interior of , denoted , is a triangular neighborhood of . The action of maps a triangular neighborhood of onto a triangular neighborhood of .
Example 3.2.
This example describes the triangular neighborhoods of .
-
(i)
Figure 1: The triangular neighborhood -
(ii)
Let and consider the geodesic triangle with vertices at . Write . If then is degenerate and . Without loss of generality assume that . The geodesic arc joining and is unique. Therefore there exists such that for each where is a relatively compact subset of , i.e., the closure of in is compact; cf. Figure 1. Now suppose is an open neighborhood in . Choose with . It follows that . Thus each nonempty triangular neighborhood of is a union of and a relatively compact subset of for suitable with .
Growth rate at infinity.
We first introduce a scale for the growth of a function in the triangular neighborhoods of . Let be a fixed real number. A function decays with exponent at if is bounded on each triangular neighborhood of . Note that the definition does not require the bound to be uniform. Write
It is clear that is a -subspace of . Next, we provide an equivalent version of the decay condition at that is useful for practical applications.
Lemma 3.3.
The following are equivalent:
-
(i)
decays with exponent at .
-
(ii)
is bounded on each of for some fixed .
Proof.
Given with and there exists a positive constant so that on . Since continuous functions are bounded on relatively compact subsets the assertion is a consequence of Example 3.2 and the definition above. ∎
One employs the equivalence in Lemma 3.3 with to discover that whenever . Set
Growth near the real line.
The boundary points in naturally play a role in the theory due to the action of on the extended upper half plane. Let be a real number and . A function grows with exponent at if is bounded on each triangular neighborhood of . Define
It is clear that is a -subspace of . Note that in a small enough neighborhood of . Therefore if . We write
Exponential polynomials.
The Fourier expansion of a weakly holomorphic modular form at involves a principal part of the Laurent expansion along with a constant term (Section 3.3). Let
For define to be the -subspace of spanned by . An exponential polynomial at of index is an element of . The subspace consists of the constant functions. Set
Since is totally ordered is a -subspace of . Note that the polynomials in have unique holomorphic extensions to .
Lemma 3.4.
With notation above
Proof.
Suppose . Then as . It follows that must be the zero exponential polynomial. Therefore . ∎
Let . Define
Lemma 3.4 shows that . Suppose that and are the projection maps. For brevity one writes and .
The final topic of the current subsection is the action of conformal automorphisms on growth families. With we attach a conformal automorphism
Lemma 3.3 implies that whenever for some . It is clear that if then . Therefore is also closed under .
Lemma 3.5.
Let , , and .
-
(i)
If then both and are stable under . Moreover
(3.1) -
(ii)
Now suppose and . Then .
Proof.
- (i)
-
(ii)
Suppose and . Here and . Assume that . Let be a triangular neighborhood of . Then is a triangular neighborhood of . Note that is bounded away from zero on . We use the decay estimates for to obtain so that for each . Since the statement holds for each it is true for also.
∎
3.2 Eichler-Shimura integrals with base point at
Our discussion primarily concerns the following classes of continuous paths on the extended upper half-plane:
-
•
Let . A good path between and refers to a piecewise smooth path on that maps inside .
-
•
The unique good path joining and is the vertical ray originating at .
In the sequel the path integrals are along good paths joining the endpoints.
Lemma 3.6.
Let and . Suppose that , i.e., is bounded on each triangular neighborhood of . Then
does not depend on the choice of good path joining the endpoints.
Proof.
Let be two good paths joining and . We use piecewise smoothness to obtain an admissible triangle at and so that for each . Note that is geodesically convex. Set
where length is measured with respect to the underlying metric on . Suppose that for all . Then
We let and conclude that the above inequality’s LHS is zero. ∎
Corollary 3.7.
Let . Suppose that . Then
does not depend on the choice of good path joining the endpoints.
Proof.
The assertion is a simple consequence of Lemma 3.6. ∎
We next record an observation that is useful for the subsequent discussion.
Lemma 3.8.
Let . Then converges absolutely to define a holomorphic function on and
Proof.
The hypothesis on amounts to the fact that on every vertical strip bounded below. Now the assertion is an easy exercise in standard techniques of complex analysis. ∎
Let and . Suppose . Write . Set
It is clear that . Also, we have for each . One employs Lemma 3.8 to conclude that
and satisfies . Put
(3.2) |
The differential relations for and together imply that is a weight Eichler-Shimura integral of . We refer to as the Eichler-Shimura integral of with base point at .
Now suppose is a finite index subgroup of and is stable under the action of . Let
be the choice function attached to (3.2) and is the period function of . The Eichler-Shimura integrals with base point at behave like a canonical choice to some extent. In particular is linear whenever is a linear subspace of . Next, we provide an explicit formula for the period function. For convenience write
Proposition 3.9.
Let and . Then
(3.3) | ||||
for each .
Proof.
Let . One uses (3.2) with to conclude that
(3.4) | ||||
To deduce (3.3) from (3.4) one must show that the change of variable transforms the first, resp. third, term in RHS of (3.4) into the first, resp. third, term in RHS of (3.3). We assume that the path of integration in the first integral of the RHS is the geodesic arc joining and which ensures that the transformed path is a good path between the endpoints whenever . Now a routine computation using the definition of growth classes verifies that the boundary contribution in each of the improper integrals is negligible and the change of variable goes through without any problem. ∎
The following specialization of Proposition 3.9 is particularly useful to us.
Corollary 3.10.
Let and . Then
(3.5) |
3.3 Connections with the theory of modular forms
Let be an integer and be a finite index subgroup . Suppose that , resp. , is the space of weight weakly holomorphic, resp. holomorphic, modular forms on . We choose a positive integer so that . A weakly holomorphic modular form admits a -series expansion
(3.6) |
where is a sequence of complex numbers so that for and has radius of convergence on the open unit disc. If then for each .
Lemma 3.11.
.
Proof.
Observe that the space of constants is contained in . Set
The following result is an immediate consequence of the lemma above.
Corollary 3.12.
.
The formula (3.5) simplifies for the class of functions . Let the notation be as in Corollary 3.10. Moreover assume that . Write where and .
Lemma 3.13.
.
Proof.
Note that . Moreover is a constant function. Thus the desired identity follows from the integral formula (3.5). ∎
Our equivariant period function is particularly appropriate for the induced period polynomial perspective towards Eichler-Shimura theory since here one can always evaluate the period function at even if is not -invariant; see Proposition 7.6. The results above also verify that the formalism of the current section applies to the weakly holomorphic modular forms. Thus the theory developed in this article offers a cohomological vista of the recent developments on weakly holomorphic modular forms [4].
4 Eichler-Shimura homomorphism
This section aims to retrieve the setup of Eichler-Shimura theory for invariant functions from the equivariant formalism developed in Section 2.
4.1 Definition and basic properties
We start with a simple algebraic observation that allows us to translate our theory of period functions into the classical language. Let be an abstract group and be a trivial right -set. Suppose that is a field and is a right -module. There is a canonical -linear isomorphism
(4.1) | ||||
Since acts trivially on the map gives rise to -linear isomorphisms at the level of cocyles and cohomology:
Let the notation be as in Section 2.2 and assume that . Now suppose is a choice function for . With one associates an Eichler-Shimura cocycle map given by
(4.2) |
where is the period function of . This cocycle map naturally descends to a map into the cohomology module.
Definition 4.1.
The Eichler-Shimura homomorphism attached to is the image of the Eichler-Shimura class under , i.e.,
Proposition 2.7 ensures that is determined by and does not depend on the choice of the Eichler-Shimura cocycle map. If is clear from context then we drop and from the notation. If consists of functions lying in suitable growth classes, say , then the cocycle function (4.2) determined by base point at retrieves the period cocycle used by Zagier and subsequent authors [14]. However, at the level of cohomology, the corresponding Eichler-Shimura map coincides with the map defined by a canonical choice .
The following result records the basic properties of the Eichler-Shimura map.
Proposition 4.2.
Let be a collection of -invariant functions.
-
(i)
The class vanishes if and only if is the zero map.
-
(ii)
Let be a -subspace of for some subfield of . Then is automatically -linear.
Proof.
Part (i) follows from the observation that is an isomorphism. To prove (ii) let be a canonical choice function for . We know that is a -linear choice function, i.e.,
One now uses the definition of in (4.1) to discover that the cocycle map attached to is also -linear. But descends to the map . Therefore is -linear. ∎
4.2 The action of double coset operators
Next, we verify the Hecke equivariance of the Eichler-Shimura homomorphism defined in the previous subsection. Recall that the double coset operators generate the Hecke algebra of a congruence subgroup. In this article, the term Hecke operator refers to a general double coset operator unless otherwise specified.
The following paragraphs digress from our convention and deal with the action of subgroups of on function spaces and module of coefficients (Section 2.3.1). Let and , be two finite index subgroups of . Suppose that is a set of representatives for the orbits of the left -action on . For each we have a -linear double coset operator on functions given by
The operator is independent of the choice of -s. There is also a notion of double coset operators for group cohomology. Let denote the subgroup of generated by , , and . Suppose that is a right -module where is a field of characteristic . Given and we obtain unique and so that . Define
(4.3) |
The operator sends cocycles to cocycles and descends to a -linear map in cohomology. We denote the map in cohomology by the same notation and often drop from the subscript. The map at the level of cohomology does not depend on the choice of representatives for -orbits. Readers may like to notice that in our convention the double coset operators act on the right of the cohomology module exactly like function spaces.
Proposition 4.3.
(cf. [16], Proposition 8.5) Let and be two subsets of functions so that . Then the Eichler-Shimura maps commute with the action of double coset operators, i.e., .
Proof.
Let be the subgroup of generated by . Suppose that is a -stable subset of containing . We fix a choice function and consider the period function . Restriction yields choice functions for and . Let and . Then
Summing over one finds that
equals a coboundary on . ∎
4.3 Equivariance under group action
We begin with a recap of the conjugation action on the group cohomology [6, III.8]. Let be an abstract group and be a subgroup of . Suppose that is a field of characteristic and is a right -module. For each conjugation gives rise to an isomorphism of pairs
that yields a pullback isomorphism . For set
(4.4) |
Suppose that is a normal subgroup of . Then (4.4) defines a right -module structure on . The subgroup acts trivially on and the -action descends to an action of .
Lemma 4.4.
Let the notation be as above. Suppose that is a finite index normal subgroup of . Then the restriction homomorphism
maps isomorphically onto .
Proof.
See Proposition 10.4 in [6, III.10]. ∎
We next explicate the action of at the level of -cocycles; cf. [6, p.79]. Assume that is a normal subgroup of . The (right) conjugation action of on turns into a -module as follows:
(4.5) |
Note that and are -submodules of . Thus inherits a -action that coincides with (4.4).
Example 4.5.
Let be a subgroup of containing and assume that acts on as . Then, by (4.5), the conjugation action of on equals . But each element of must act trivially on the cohomology. It follows that must be zero. In particular, if is odd and then is zero.
Remark 4.6.
We put ourselves in the setting of Section 4.1 and suppose that is a subgroup of that contains as a normal subgroup. Moreover assume that is stable under the -action of . The discussion of this subsection yields a canonical right -module structure on .
Proposition 4.7.
The Eichler-Shimura homomorphism
is a morphism of -sets.
Proof.
We fix a choice function for and consider the period function and the cocycle map attached to that choice. Let and . Then
Therefore in as desired. ∎
Remark 4.8.
With the notation above let be a subgroup of containing . Set . Then using compatible choice functions for and (cf. (2.13) and (2.14)) one discovers that the following diagram commutes:
Here the top arrow is inclusion and the bottom arrow is the restriction map in group cohomology. Note that the bottom arrow maps isomorphically onto the -invariant subspace of .
5 Preparations for the main theorem
5.1 Parabolic cohomology
We follow the treatment given in the book of Shimura [16] and first introduce the notion of parabolic cohomology for the subgroups of the projective modular group . As per standard convention, we denote the cusp of a finite index subgroup of or by choice of representative for the equivalence class. In particular, the stabilizer of a cusp refers to the stabilizer of a fixed choice of representative for the cusp.
Let be a finite index subgroup of . Suppose that is the set of cusps attached to . For let denote the stablizer of in . Write . We also fix a set of representatives of the elliptic elements in . Let be a right -module where is a field of characteristic . For a subset of one defines the cohomology groups vanishing on , denoted , by imposing constraints on -cochains [16, p.224]. Suppose that is the collection of all parabolic elements in . We refer to as the parabolic cohomology of . However, for practical purposes, it is more convenient to work with . The space of -cocycles vanishing on and coincide and there is a canonical identification . This description of provides an exact sequence of -vector spaces
(5.1) |
where the first map is inclusion and the second map is componentwise restriction. The parabolic subspace is compatible with the restriction map.
Lemma 5.1.
-
Let be two finite index subgroups of and be a -module. Suppose that . Then
if and only if .
Proof.
Let . The natural inclusions give rise to a commutative diagram
where all the arrows are restriction maps. By Remark 4.6 the bottom arrow in the diagram above is injective. Therefore the image of in is zero if and only if the image of in is zero. Since this assertion holds for each the statement follows from (5.1). ∎
We can also compute in terms of a -dimensional simplicial complex that contains the fixed points of -s as -simplices and admits a simplicial -action [16, p.225]. With one associates a complex of finitely presented -modules so that the complex
computes . Readers may like to note that Shimura defines as a complex of left modules which we endow with the canonical right action . Moreover also provides an alternate description of . In particular, there is an isomorphism
(5.2) |
Lemma 5.2.
Let be a field containing . Then the natural change of scalar map induces isomorphisms
Now suppose is a finite-dimensional -vector space. Write
Proposition 5.3.
Let the notation be as above and denote the genus of the canonical compact Riemann surface attached to . Then
-
(i)
, -
(ii)
.
Proof.
The second identity is standard and we refer the readers to [16, p. 229] for a demonstration. The first identity is a consequence of the arguments employed to establish the other identity. Let be the number of -inequivalent -simplices of . Here , resp. , is a free -module of rank , resp. but the elliptic fixed points may prevent from being a free module. A straightforward analysis demonstrates that
But
and . We now substitute the values of in the identity above to arrive at (i). ∎
Lemma 5.4.
Let be a finite index subgroup of and be a right -module. Then .
Proof.
Let be a finite index subgroup contained in so that is torsion-free. Then the arguments given in Proposition 1 of [10, 6.1] demonstrates that . But the natural restriction map from to is injective (Remark 4.6). Therefore is also zero. Note that Hida originally proves the result for congruence subgroups but his proof goes through for all torsion-free finite index subgroups. ∎
Next, we introduce a few conventions to facilitate the transition between the subgroups of and . Let be a subgroup of and be a right module. If is a finite index subgroup and is the subset of all parabolic elements of then one can introduce the notion of parabolic subspace, denoted , using the same formalism mentioned before. The subspace of is stable under the action of double coset operators on cohomology [16, 8.3]. We call a descent module if either or acts trivially on . For example, the familiar pair is a descent module if either is even or . Let be a descent module. Then there exists a natural -structure on so that the -structure on arises from the projection . Moreover, the pullback of the projection map identifies and [16, 8.2]. Now suppose is a finite index subgroup of . Note that the image of in , denoted , equals the collection of parabolic elements of and there is an identification . The procedure described above allows us to rewrite (5.1) as
(5.3) |
And furthermore, the change of scalar property of Lemma 5.2 lifts to in the following manner
(5.4) |
whenever is a field extension of .
Remark 5.5.
The final topic of this subsection is the third term appearing in (5.3). We first consider the subgroup for some positive integer where refers to the subgroup generated by the element. Since is an infinite cyclic group it follows that
(5.5) |
Here the last isomorphism stems from a linear functional described as the coefficient of in . Let be a finite index subgroup and . Choose so that . Then . In more detail, there exists a positive integer so that
where is the collection of regular cusps. The following result is certainly well-known among the experts; cf. [10, 6.3].
Proposition 5.6.
Let be as above and assume that if is odd. Then
Proof.
We use the conjugation by isomorphism (Section 4.3) to obtain
If then . Now suppose . Observe that
Here is even and acts trivially on ; see (4.5). As a consequence . It remains to consider the case . In this situation, there is an isomorphism
One again employs (4.5) to conclude that acts on by . Hence if is even and equals zero if is odd. ∎
5.2 Isomorphism theorem for cusp forms
We first describe the action of complex conjugation on our cohomology spaces. Let be an integer and be a finite index subgroup of . The complex conjugation on gives rise to an conjugate linear automorphism of described as where . This automorphism descends to a complex conjugation on . Now suppose is the space of all weight cusp form on and is the space of antiholomorphic cusp forms on . Let
be the Eichler-Shimura homomorphism constructed in Section 4. One defines an antiholomorphic avatar of by
It is clear that is -linear. We extend the action of a double coset operator to the space of antiholomorphic cusp form using the formula . Since the action of on a cocycle preserves the decomposition into real and imaginary parts it follows that is also Hecke equivariant. Hence and together provide a Hecke-equivariant -linear map
The Eichler-Shimura isomorphism theorem for the cusp forms is as follows:
Theorem 5.7.
[10, p.171] The map is an injection whose image equals the parabolic subspace of .
Proof.
For write and . The -linear version of the isomorphism theorem given in [16, 8.2] asserts that the assignment , , is an isomorphism of real vector spaces. One twists the map above by the -linear automorphism of to deduce that the imaginary part map also defines a real linear isomorphism between and . Now the assertion is a consequence of (5.4). ∎
The injectivity statement of the theorem combined with the functorial properties of the Eichler-Shimura class leads to the following result.
Corollary 5.8.
Let be an integer and be a finite index subgroup of . Then is nonzero.
5.3 The space of Eisenstein series
Let be an integer and be a congruence subgroup of . Suppose that is the space of weight modular forms on . Then the space of Eisenstein series on , denoted , is the unique linear complement of in that is orthogonal to with respect to the Petersson inner product [7, 5.11]. The equivariance properties of the inner product imply that the decomposition is stable under the action of a general double coset operator. Next, we explain how to use the spectral Eisenstein series on to construct a basis for the space of Eisenstein series following the author’s recent account of the theory [15].
Let and be a scaling matrix for , i.e., . If is odd then additionally assume that and is a regular cusp. The spectral Eisenstein series of weight attached to and the scaling matrix is
where for . If is odd then our Eisenstein series may depend on the choice of scaling matrix. The series above admits analytic continuation to the larger domain allowing us to specialize at . For simplicity we abbreviate as . An explicit calculation with the Fourier series shows that
where is a nonzero rational number. Moreover, we have
(5.6) |
Set
where in the second case we write the series for a fixed choice of scaling matrix for each regular cusp. Then is a basis for the space of Eisenstein series . For a subfield of we write for the span of . This -structure on the space of Eisenstein series is compatible with restriction to a principal level. In more detail, if then
(5.7) |
To write down uniform statements for all weights we introduce an extended space of Eisenstein series
It is clear that and is a subspace of codimension inside . The -rational structure on the extended space of weight is also compatible with restriction to a principal level in the sense described above.
Remark 5.9.
The author expects that there are spectral Eisenstein series of weight for every finite index subgroup so that their specializations at give rise to a basis of the space of Eisenstein series. Such a construction would directly lead to a generalization of Theorem 1.1 to finite index subgroups. However, in this article, we restrict our attention to congruence subgroups with a view towards arithmetic application.
6 Proof of Theorem 1.1
Let be an integer and be a congruence subgroup of . Recall the decomposition from Section 5.3. Suppose that
is the Eichler-Shimura homomorphism attached to the space of modular forms on . Then is a Hecke equivariant extension of the map on the space of cusp forms. We need to prove that
is an isomorphism. If is odd and then the domain and codomain of the map above are zero (Example 4.5). Assume that whenever is odd. Let denote the restriction of to the space of Eisenstein series on . In the light of Theorem 5.7 it suffices to verify that is injective and the image of is a linear complement of in .
6.1 Restriction to the cusps
With each one can associate a cuspidal Eichler-Shimura map
where is the restriction map from to . Proposition 5.6 furnishes a complete description of the target of . The subsequent discussion aims to find the kernel of this map.
Let and choose so that . Now suppose with . Note that the restriction maps fit into a commutative diagram
(6.1) |
exactly as in the proof of Lemma 5.1. Here is a normal subgroup of and acts by conjugation on (Section 4.3). The conjugation action of yields another commutative diagram described by
(6.2) |
where the vertical arrows are restriction maps. Next, we use our equivariant machinery to verify the key property of the map .
Proposition 6.1.
Suppose that . Then if and only if .
Proof.
We identify , resp. , with the -invariant subspace of , resp. , and utilize
to describe (Remark 4.8). First, one invokes Remark 4.6 to conclude that the bottom arrow of (6.1) is injective. Hence if and only if the restriction of to is zero. But is -equivariant (Proposition 4.7). Therefore . One now employs (6.2) to conclude that if and only if the restriction of to is zero. Note that . From Lemma 3.13 it follows that
Therefore, by (5.5), if and only if . ∎
One uses the theorem above to determine the cuspidal behavior of the cohomology classes attached to the basis for the space of Eisenstein series. Let the notation be as in the discussion around Section 5.3. For convenience write if is even and if is odd.
Corollary 6.2.
Let and assume that . Then if and only if . Moreover, let and . Then
if and only if .
6.2 Completion of the proof
The cuspidal Eichler-Shimura maps together give rise to a homomorphism
We use this homomorphism to extract information about the Eichler-Shimura map attached to the space of Eisenstein series.
Case I: .
Proposition 5.6 shows that if and only if . Therefore maps our basis of to a basis of (Corollary 6.2). In particular, the map is injective. It follows that is also injective. Moreover where refers to the image a map. The surjectivity of demonstrates that (5.3) is exact at the right end. Therefore is a linear complement of as desired.
Case II: .
7 Versions of the isomorphism theorem
7.1 The induced picture
We begin with a convenient description of Shapiro’s lemma [12, p.62]. Let be an abstract group and be a subgroup of . Suppose that is a right -module where is a field of characteristics zero. Write
and endow it with the right -action . If the -module structure on extends to a -structure then as -module via the assignment where . Shapiro’s lemma states that the canonical homomorphism of -modules described by gives rise to a -linear isomorphism
Let be a finite index subgroup of . Suppose that is a right -module. We consider the Shapiro isomorphism
explicitly described by
(7.1) |
In practice working with the cocycles on is more convenient since each cocycle is determined by its value on . One pulls back the action of the double coset operators on to obtain a -Hecke action on ; cf. [14, 5]. The -action on need not descend to an action of . Define
It is clear that as -modules. There is a natural projection
Note that for each . Therefore the natural inclusion
is an isomorphism. The action of on descends to a -action and one identifies and as in Section 5.1. Thus one can use the isomorphism above to deduce that the induced module satisfies the change of scalar properties given in (5.4). The experts are presumably familiar with the following result.
Proposition 7.1.
[14, 2] Let be a descent module. Then the Shapiro map induces an isomorphism . As a consequence the parabolic subspace is stable under the -Hecke action.
For a finite index subgroup the normal core of , denoted , is
This subgroup is the largest finite index normal subgroup of that is contained in .
Proof.
Let be a positive integer so that . Then for each . Suppose that is a cocycle on with values in . A straightforward calculation yields
Therefore if and only if . It follows that
But , resp. , is a finite index subgroup of the abelian group , resp. . The equivalence above along with the injectivity of restriction argument and (5.3) implies that if and only if where denotes the image in cohomology. Hence induces an isomorphism between and . The second part of the assertion is clear from the definition of -Hecke action. ∎
We next explicitly describe the preimage under of the cohomology classes attached to modular forms. Write . Note that . In particular and one can invoke the theory of Eichler-Shimura integrals with base-point at infinity for this collection of functions. Let
be the Eichler-Shimura cocycle map attached to for the choice of base point at . Define the induced Eichler-Shimura cocycle map by
(7.2) | ||||
Our theory of equivariant period functions for the pair directly verifies that is indeed a cocycle on . Moreover, there is a commutative diagram
(7.3) |
where is as in (7.1). Note that the cocycle takes values in . The induced Eichler-Shimura homomorphism attached to is
We denote the restriction of to by .
Lemma 7.2.
.
Proof.
The assertion is an easy consequence of (7.3). ∎
Remark 7.3.
The -modules and inherit real structures from and in (7.3) is compatible with the associated real structures. As before the real structures descend to the corresponding cohomology modules and is compatible with complex conjugations in the source and target. Let
be the induced antiholomorphic Eichler-Shimura map where is the conjugation in cohomology.
Corollary 7.4.
.
Proof.
Follows from Lemma 7.2 and the discussion above. ∎
Theorem 7.5.
Let be a congruence subgroup of . The map
is a -Hecke equivariant isomorphism. Moreover, this isomorphism maps onto the parabolic subspace .
Proof.
The first part of the assertion is a consequence of Theorem 1.1, Lemma 7.2, and Corollary 7.4. The latter half follows from Theorem 5.7 and Proposition 7.1. Note that if is odd and then the induced cohomology vector space is zero (Example 4.5). Thus the second half of the assertion automatically holds in this case. ∎
We conclude the discussion with the explicit expression of the period cocycles at and . For define the completed -function of by
(7.4) |
The integral defining is absolutely convergent in the region . Moreover, we have the following identity involving entire integrals:
In particular extends to a meromorphic on the whole complex plane with at most two simple poles supported at . It also satisfies a functional equation . The next result provides a complete description of the induced cocycle map (7.2). As before we work with the -stable set and consider it as a subset of .
Proposition 7.6.
7.2 Modular forms with nebentypus
The present subsection aims to retrieve the twisted Eichler-Shimura isomorphism theorem for the spaces of modular forms with nebentypus from Theorem 1.1. Let be a positive integer and be a finite index subgroup of . We say a character is defined over a subfield of if . Let be a character on that is defined over . If be a right -module then denotes a right -module whose underlying vector space equals and the -action is given by
Now suppose is an arbitrary character on . Set
Let be a -stable subset of holomorphic functions so that . Suppose that is a choice function for . With one associates a -twisted Eichler-Shimura cocycle map given by
where is the period function of . The general transformation formula (2.10) ensures that the recipe above indeed defines a cocycle with values in . As a consequence we obtain the -twisted Eichler-Shimura homomorphism:
Since the period functions attached to two distinct choice functions differ by a coboundary in a straightforward calculation verifies that depends only on and is independent of .
Let be a positive integer and consider . Recall that the natural map
induces an isomorphism between and . We use this isomorphism to identify each character on with a character on . For simplicity write . Our treatment of twisted Eichler-Shimura theory relies on the formalism developed in the paragraph above for the pair where . In this setting there is a notion of twisted Hecke action on the functions and the cohomology module [16, 3.5]. Put
where refers to the ring of matrices. Let be a character on defined over . With we associate a homomorphism of monoids by setting . If then . Let . Write as in Section 4.2. Define
The double coset operator also acts on the cocycle model of cohomology. Let be a right -module. Set
where -s are as in (4.3) and acts on by the -action on . This map descends to an endomorphism of that depends only on the double coset . An argument similar to Proposition 4.3 verifies that the twisted Eichler-Shimura map is a homomorphism for the twisted Hecke action. Moreover, a straightforward calculation imitating the untwisted counterpart shows that the twisted Hecke action preserves the parabolic subspace of cohomology. Note that . Therefore is a descent module if and only if acts on by the scalar . In particular is a descent module if and only if .
We next relate the -eigenspace in cohomology with the cohomology of the -twisted module using the conjugation action of on the cohomology of . This result allows us to bypass the use of trace maps in Hida’s theory [10, p.177]. Let be a subfield of and be a character defined over . Suppose that is a -module. Since is a normal subgroup of the canonical restriction map induces an isomorphism
(7.5) |
Lemma 7.7.
Let denote the eigenspace for the -action on . Then
-
(i)
.
-
(ii)
Suppose that is a descent module. Then the image of the parabolic subspace under (7.5) equals .
Proof.
The action of diamond operators on the spaces of modular forms on yields decompositions of the form
where the sum is over all the Dirichlet character modulo . The space , resp. , is zero unless . Attached to these spaces we have Hecke-equivariant twisted Eichler-Shimura homomorphisms
for each . The action of diamond operators on the space of anti-holomorphic cusp forms yields a similar eigenspace decomposition
It is clear that . We transport the -Hecke action on via complex-conjugation to obtain a -Hecke action on . This procedure yields a well-defined -Hecke equivariant homomorphism
where is the complex conjugation defined in Section 5.2.
Theorem 7.8.
Let be an integer and be a character on . The -Hecke-equivariant homomorphism
(7.6) |
is an isomorphism. Moreover the image of under this map equals .
Proof.
We start with the untwisted Eichler-Shimura isomorphism
(7.7) |
One considers the -eigenspaces for the action of diamond operators to discover that the isomorphism (7.7) maps isomorphically onto . Now the Eichler-Shimura homomorphism and its twisted avatar together yield a commutative diagram
(7.8) |
where the vertical arrow is the restriction map. Note that the vertical arrow is injective and its image equals (Lemma 7.7). Therefore the twisted homomorphism (7.6) is an isomorphism. If then the domain and codomain of (7.6) are zero (Example 4.5). Now suppose . Observe that the image of under (7.7) is . Hence, by Lemma 7.7(ii), the isomorphism (7.6) must map onto . ∎
The space of Eisenstein series with nebentypus is given by
(7.9) |
If is trivial then . Without loss of generality assume that is a nontrivial character with . In this setting, the author has recently constructed an explicit basis that is parametrized by the cusps of and established a weaker version of Theorem 7.8 [15, 7.3]. By construction the basis functions in lie inside if . Here is Euler’s totient function and is the -th cyclotomic extension of in . Recall the spectral basis from Section 5.3. For an arbitrary character set
(7.10) |
The compatibility with restriction property described in (5.7) shows that for each . Therefore for all whenever . We study rationality of the image of this basis under in Section 8.2.
8 Computation of Eichler-Shimura map
Next, we use the induced picture, in particular, Proposition 7.6 to compute the Eichler-Shimura map for the space of Eisenstein series on principal congruence subgroups. This calculation enables us to prove algebraicity results regarding the image of the Eichler-Shimura map attached to the space of Eisenstein series on an arbitrary congruence subgroup.
8.1 Eisenstein series on principal congruence subgroups
8.1.1 Eisenstein series on the universal elliptic curve
This subsection aims to study a special family of Eisenstein series arising from the -torsion points of the universal elliptic curve that exhibits remarkable rationality properties regarding period polynomials. We begin with a brief review of Hecke’s theory of Eisenstein series on .
Let be an integer and . With each one associates an Eisenstein series described as
where is lift of and is the imaginary part of . This series converges uniformly on the compact subsets of to define an analytic function of in this region that admits an analytic continuation to the whole plane. Set . As a function on the Fourier expansion of the -series is where
(8.1) |
and is a lift of to [17, 9.1]. Recall the rational structure on the space of Eisenstein series from Section 5.3 arising from the spectral basis. Let denote the -th cyclotomic field . Suppose that is a subfield of containing . Then the -span of equals and its intersection with is [15, 6.2]. The explicit Fourier expansion makes -series accessible for the computation of -function and -values. A standard approach towards this problem uses a cyclotomic linear combination of -series to write down a Hecke eigenfunction on whose -function is given by a product of two Dirichlet -functions [13, IV-39]. However, it is difficult to study the rationality properties of special values of such product -functions.
One defines the Eisenstein series on the universal elliptic curve as
where and [11, 4]. This series converges uniformly on compact subsets to define a real analytic function that transforms like a Jacobi form of weight and index . In particular for each . With one associates the division value
Let be the nondegenerate bilinear form defined by
Then the relation between -series and -series is as follows:
(8.2) |
Let be an integer and . We define the elliptic Eisenstein series of weight and parameter , denoted , using the identity (8.2). Note that is compatible with the natural right action of on , i.e., for each . Thus . In particular is invariant under . One uses the orthogonality of characters for to invert the linear relation (8.2) and discovers that
Therefore the -span of equals the -span of whenever .
We next record the Fourier expansion formula for the elliptic Eisenstein series. Let be the Bernoulli polynomial defined by the generating function identity
Our calculation requires the Fourier expansion formula for the Bernoulli polynomials [17, 4.5]:
(8.3) |
The formula is valid even for if and one interprets the sum as .
Lemma 8.1.
Let be an integer and . Suppose that is a lift of that satisfies . The Fourier series for is
Proof.
Corollary 8.2.
Let . Then is holomorphic if and only if . Moreover is a spanning set for the holomorphic subspace whenever .
Proof.
The first part of the assertion is obvious. The extended space of -series contains as a subspace of codimension . But is already a subset of holomorphic subspace. Therefore this collection spans . ∎
For convenience, we introduce the notation
8.1.2 -function and -values of the elliptic Eisenstein series
Our next job is explicitly writing down the completed -function (7.4) attached to the elliptic Eisenstein series and determining its values at . Let and . One defines the Lerch zeta function [1] by
(8.4) |
The series above converges uniformly on the compact subsets of to define an analytic function in this region. The Lerch series admits a meromorphic continuation to the whole -plane with a simple pole at whenever is an integer. If is not an integer then extends to an entire function on the -plane. Our treatment requires two special avatars of the Lerch function. The Lerch series reduces to the familiar Hurwitz zeta function for , i.e., . The value of the Hurwitz zeta function at nonnegative integers [2, 12.11] is given by
(8.5) |
On the other hand setting in (8.4) yields the polylogarithm function given by the convergent series
in the region . The meromorphic continuation of the Lerch function provides a meromorphic continuation for Li. If is not an integer then equals the conditionally convergent logarithmic series . The Fourier expansion formula for the Bernoulli polynomials (8.3) shows that
(8.6) |
for each . The identity above is valid even for if .
Lemma 8.3.
Let and . Suppose that is a lift of that satisfies . Then
(8.7) | ||||
where
Proof.
We first assume that . Then the completed -function admits convergent integral representation and the identity (8.7) is a straightforward consequence of Lemma 8.1. Observe that both the sides of (8.7) meromorphically continue to the whole plane. Thus the identity holds due to the principle of analytic continuation. ∎
Proposition 8.4.
Let the notation be as in Lemma 8.3. We further assume that . Suppose that . Then
The proof of the identity involves the following standard properties of the Bernoulli polynomials [17, 4.5]:
(8.8) |
In particular
Proof.
If then the identity is an easy consequence of the formulas given in (8.5)-(8.8):
The extremal case of (8.8) together with (8.6) guarantee that the calculation for extends to the following situations without any difficulty:
-
•
all if ,
-
•
if , if .
Suppose that and . The hypothesis on implies that if then is automatically nonzero. In this situation contributes instead of and this discrepancy gives rise to the desired term involving . It remains to verify the identity for and . As before, if then . The functional equation for the completed -function implies that
Here, in the last step, we are using (8.6) to simplify the expression. ∎
8.2 Algebraicity of cohomology classes
Let the notation be as in Section 8.1. The explicit expressions for the special values of the completed function in Proposition 8.4 motivates us to normalize the elliptic Eisenstein series as
The following lemma computes the value of the period function attached to the -stable subset at the normalized elliptic Eisenstein series.
Lemma 8.5.
Let and . Suppose that is a lift of satisfying . Then
Proof.
Note that the value of the period polynomial at involves polylogarithm functions. Thus it is difficult to examine the rationality properties of the period polynomials described above. We next explain how to change the Eichler-Shimura cocycle attached to a function by a coboundary so that the resulting cocycle has coefficients in ; cf. [20, 7]. Let
(8.9) |
be the Eichler-Shimura map attached to the space of modular forms.
Proposition 8.6.
Let and . Then the image of under the Eichler-Shimura map (8.9) is definable over .
Proof.
We use the induced picture of Section 7.1 to compute the Eichler-Shimura map. In the light of the explicit Shapiro map (7.1) it suffices to check that the image of under lies inside the rational subspace . For let denote the unique lift of in . Note that depends only on the coset of in .
Case I. .
Define by setting
A straight-forward computation using the formulas in Proposition 8.4 and Lemma 8.5 shows that the modified cocycle
takes values in at and .
Case II. .
The problem, in this case, is that both the transcendental terms in the special -value may simultaneously contribute a term to the calculation. We construct another function by putting
Then the modified cocycle
again takes values in at and . ∎
Let be a subfield of containing . Recall that the -span of equals . Therefore the collection
is a -spanning set of .
Corollary 8.7.
The homomorphism maps inside the canonical -form .
Proof.
Proposition 8.6 demonstrates that
for each . Thus the assertion follows from the discussion above. ∎
Suppose that is a congruence subgroup containing the principal level . One considers the -invariants of the domain and codomain of (8.9) to retrieve the Eichler-Shimura homomorphism for :
(8.10) |
Theorem 8.8.
Let be a subfield of containing . Then the image of under the Eichler-Shimura map (8.10) lies inside .
Proof.
Proof of Theorem 1.5:
Let the notation be as in Theorem 1.5. Suppose that . From the definition of rational structure it is clear that . Hence the assertion is a consequence of Theorem 8.8.
Remark 8.9.
Observe that the -subspace of is stable under the action of a double coset operator (Section 4.2). Thus one can use Theorem 8.8 to conclude that the action of a double coset operator is definable on our rational spaces of Eisenstein series; cf. Theorem 7.3 in [15].
Corollary 8.10.
Let the notation be as in Theorem 8.8. Then is stable under the action of double coset operators.
Proof.
Follows from the Hecke equivariance of the Eichler-Shimura isomorphism and Theorem 8.8. ∎
Next, we discuss the rationality of the twisted Eichler-Shimura homomorphism on the space of Eisenstein series with nebentypus . Let be an integer and . Suppose that is defined over a subfield of . Then is a -form of and the inclusion induces a change of scalar isomorphism:
Note that is a -form of . Since is defined over the -eigenspace for the on is also defined over , i.e., . Lemma 7.7 shows that the natural restriction map induces an isomorphism
that maps onto . Recall the basis of defined in (7.10).
Theorem 8.11.
Let the notation be as above. The image of under lies inside where .
Proof.
One can extract a Hecke stability statement from the theorem above.
Corollary 8.12.
Let be a subfield of containing . Then the -span of is stable under the action of twisted double coset operators defined in Section 7.2.
Proof.
Follows from the -Hecke equivariance of the twisted Eichler-Shimura map and Theorem 8.11. ∎
Acknowledgments.
The author wishes to thank Don Zagier for introducing him to the Eichler-Shimura theory and for lots of valuable advice. He is grateful to the Max Planck Institute for Mathematics, Bonn for hospitality and financial support during the visit (2018-2019). The author is also indebted to the School of Mathematics at Tata Institute of Fundamental Research for providing an exciting environment to study and work.
References
- [1] T.M. Apostol, On the Lerch zeta function, Pacific J. Math. 1(2), 161-167 (1951). https://doi.org/10.2140/pjm.1951.1.161
- [2] T.M. Apostol, Introduction to analytic number theory, Undergraduate texts in mathematics, Springer, New York (1976)
- [3] J. Bellaïche, The Eigenbook, Pathways in mathematics, Birkhäuser, Springer Nature, Switzwerland (2021)
- [4] K. Bringmann, P. Guerzhoy, Z. Kent et al.; Eichler–Shimura theory for mock modular forms, Math. Ann. 355, 1085–1121 (2013). https://doi.org/10.1007/s00208-012-0816-y
- [5] F. Brown, Multiple modular values and the relative completion of the fundamental group of , Arxiv preprint, Version 4 (2017) https://doi.org/10.48550/arXiv.1407.5167
- [6] K.S. Brown, Cohomology of groups, Graduate texts in mathematics, Vol. 87, Springer, New York (1982)
- [7] F. Diamond and J. Shurman, A first course in modular forms, Corrected 4th printing, Graduate texts in mathematics, Vol. 228, Springer, New York (2016)
- [8] M. Goresky, Compactifications and cohomology of modular varieties, In: J. Arthur, D. Ellwood, R. Kottwitz (eds.) Harmonic analysis, the trace formula, and Shimura varieties; 551-582, Clay mathematical proceedings, Vol. 4, American mathematical society (2005)
- [9] G. Harder, Eisenstein cohomology of arithmetic groups. The case , Invent. math. 89, 37-118 (1987). https://doi.org/10.1007/BF01404673
- [10] H. Hida, Elementary theory of -functions and Eisenstein series, London mathematical society student texts, Vol. 26, Cambridge university press, Cambridge (1993)
- [11] A. Levin, Elliptic polylogarithms: An analytic theory, Compos. Math. 106, 267-282 (1997).
- [12] J. Neukirch, A. Schmidt, and K. Wingberg; Cohomology of Number fields, Second edition, Corrected second printing, A series of comprehensive studies in mathematics, Vol. 322, Springer-Verlag, Berlin Heidelberg (2013)
- [13] A. Ogg, Modular forms and Dirichlet series, W.A. Benjamin, New York (1969).
- [14] V. Paşol and A.A. Popa, Modular forms and period polynomials, Proc. Lond. Math. Soc. (3) 107(4), 713-743 (2013). https://doi.org/10.1112/plms/pdt003
- [15] S. Sahu, A note on the spaces of Eisenstein series on general congruence subgroups, Arxiv preprint, Version 3 (2024). https://doi.org/10.48550/arXiv.2312.10627
- [16] G. Shimura, Introduction to the arithmetic theory of automorphic functions, Princeton University Press, Princeton (1994)
- [17] G. Shimura, Elementary Dirichlet series and modular forms, Springer monographs in mathematics, Springer, New York (2007)
- [18] J.L. Verdier, Sur les intégrales attachées aux formes automorphes, In: Séminaire Bourbaki: années 1960/61, exposés 205-222, Séminaire Bourbaki, No.6 (1961), Talk no. 216, 27p. http://www.numdam.org/item/SB_1960-1961__6__149_0/
- [19] D. Zagier, Periods of modular forms and Jacobi theta functions, Invent. math. 104, 449–465 (1991). https://doi.org/10.1007/BF01245085
- [20] D. Zagier and H. Gangl, Classical and Elliptic Polylogarithms and Special Values of L-Series, In: B.B. Gordon, J.D. Lewis, S. Müller-Stach, S. Saito,, N. Yui (eds) The Arithmetic and Geometry of Algebraic Cycles, NATO Science Series, Vol. 548, Springer, Dordrecht (2000). https://doi.org/10.1007/978-94-011-4098-0_21