This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Eichler-Shimura theory for general congruence subgroups

Soumyadip Sahu
School of Mathematics, Tata Institute of Fundamental Research,
1 Homi Bhabha Road, Mumbai, 400005, Maharashtra, India,
Email: [email protected]
Abstract

This article aims to extend the Eichler-Shimura isomorphism theorem for cusp forms on a congruence subgroup of SL2()\text{SL}_{2}(\mathbb{Z}) 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 Γ(N)\Gamma(N) 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 k2k\geq 2 be an integer and 𝒢\mathcal{G} be a finite index subgroup of SL2()\text{SL}_{2}(\mathbb{Z}). Suppose that k(𝒢)\mathcal{M}_{k}(\mathcal{G}), resp. 𝒮k(𝒢)\mathcal{S}_{k}(\mathcal{G}), is the \mathbb{C}-vector space of modular forms, resp. cusp forms, of weight kk on 𝒢\mathcal{G}. Set 𝒮ka(𝒢)={f¯f𝒮k(𝒢)}\mathcal{S}^{a}_{k}(\mathcal{G})=\{\bar{f}\mid f\in\mathcal{S}_{k}(\mathcal{G})\}. Let Vk2,[X]V_{k-2,\mathbb{C}}\subseteq\mathbb{C}[X] be the \mathbb{C}-subspace of polynomials having degree k2\leq k-2. The classical Eichler-Shimura theory constructs a \mathbb{C}-linear Hecke equivariant homomorphism 𝒮k(𝒢)H1(𝒢,Vk2,)\mathcal{S}_{k}(\mathcal{G})\to H^{1}(\mathcal{G},V_{k-2,\mathbb{C}}) that formally extends to a \mathbb{C}-linear Hecke equivariant map

(1.1) 𝒮k(𝒢)𝒮ka(𝒢)H1(𝒢,Vk2,).\mathcal{S}_{k}(\mathcal{G})\oplus\mathcal{S}^{a}_{k}(\mathcal{G})\to H^{1}(\mathcal{G},V_{k-2,\mathbb{C}}).

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 H1(𝒢,Vk2,)H^{1}(\mathcal{G},V_{k-2,\mathbb{C}}). 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 Γ1(N)\Gamma_{1}(N) and Γ0(N)\Gamma_{0}(N) the injection (1.1) should extend to a Hecke equivariant isomorphism

(1.2) k(𝒢)𝒮ka(𝒢)H1(𝒢,Vk2,).\mathcal{M}_{k}(\mathcal{G})\oplus\mathcal{S}^{a}_{k}(\mathcal{G})\xrightarrow{\cong}H^{1}(\mathcal{G},V_{k-2,\mathbb{C}}).

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 Γ1(N)\Gamma_{1}(N) [3]. The main result of the article is as follows:

Theorem 1.1.

Let k2k\geq 2 be an integer and 𝒢\mathcal{G} be an arbitrary congruence subgroup of SL2()\emph{SL}_{2}(\mathbb{Z}). Then the Eichler-Shimura homomorphism for 𝒮k(𝒢)\mathcal{S}_{k}(\mathcal{G}) admits a natural Hecke equivariant extension

k(𝒢)H1(𝒢,Vk2,)\mathcal{M}_{k}(\mathcal{G})\to H^{1}(\mathcal{G},V_{k-2,\mathbb{C}})

so that (1.1) extends to an isomorphism (1.2).

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 \mathbb{H} be the upper half-plane. Suppose that Hol()\text{Hol}(\mathbb{H}) is the \mathbb{C}-algebra of the holomorphic functions on \mathbb{H}. Now suppose kk is an integer 2\geq 2 and fHol()f\in\text{Hol}(\mathbb{H}). A weight kk Eichler-Shimura integral of ff is a polynomial 𝕀k(τ,f,X)\mathbb{I}_{k}(\tau,f,X) of degree k2k-2 with coefficients in Hol()\text{Hol}(\mathbb{H}) so that

(1.3) τ(𝕀k(τ,f,X))=f(τ)(Xτ)k2dτ.\partial_{\tau}\big{(}\mathbb{I}_{k}(\tau,f,X)\big{)}=f(\tau)(X-\tau)^{k-2}d\tau.

The differential equation above always has a solution determined up to an element in Vk2,V_{k-2,\mathbb{C}}. The familiar Eichler-Shimura integral

(1.4) τ0τf(ξ)(Xξ)k2𝑑ξ(τ0)\int^{\tau}_{\tau_{0}}f(\xi)(X-\xi)^{k-2}d\xi\hskip 14.22636pt(\begin{subarray}{c}\tau_{0}\in\mathbb{H}\end{subarray})

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 ff 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 ff. However, the action of SL2()\text{SL}_{2}(\mathbb{Z}) 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 𝒢\mathcal{G} be a finite index subgroup of SL2()\text{SL}_{2}(\mathbb{Z}) and 𝒳\mathcal{X} be a subset of Hol()\text{Hol}(\mathbb{H}) such that 𝒳\mathcal{X} is closed under the |k\lvert_{k} action of 𝒢\mathcal{G} on Hol()\text{Hol}(\mathbb{H}). Given a choice of Eichler-Shimura integrals {𝕀k(τ,f,X)f𝒳}\{\mathbb{I}_{k}(\tau,f,X)\mid f\in\mathcal{X}\} one defines a period function

(1.5) 𝗉k,𝒳:𝒢Fun(𝒳,Vk2,)\mathsf{p}_{k,\mathcal{X}}:\mathcal{G}\to\text{Fun}(\mathcal{X},V_{k-2,\mathbb{C}})

where Fun(,)\text{Fun}(\cdot,\cdot) is the collection of all set-theoretic maps. Now Fun(𝒳,Vk2,)\text{Fun}(\mathcal{X},V_{k-2,\mathbb{C}}) has a natural right [𝒢]\mathbb{C}[\mathcal{G}] module structure stemming from 𝒢\mathcal{G}-actions on 𝒳\mathcal{X} and Vk2,V_{k-2,\mathbb{C}}; see (1.7). The period function 𝗉k,𝒳\mathsf{p}_{k,\mathcal{X}} is a cocycle on 𝒢\mathcal{G} with values in Fun(𝒳,Vk2,)\text{Fun}(\mathcal{X},V_{k-2,\mathbb{C}}). Moreover the image of 𝗉k,𝒳\mathsf{p}_{k,\mathcal{X}} in H1(𝒢,Fun(𝒳,Vk2,))H^{1}\big{(}\mathcal{G},\text{Fun}(\mathcal{X},V_{k-2,\mathbb{C}})\big{)}, denoted [𝗉k]𝒢,𝒳[\mathsf{p}_{k}]_{\mathcal{G},\mathcal{X}}, is independent of the choice of Eichler-Shimura integrals. We call [𝗉k]𝒢,𝒳[\mathsf{p}_{k}]_{\mathcal{G},\mathcal{X}} the Eichler-Shimura class of weight kk associated with 𝒢\mathcal{G} and 𝒳\mathcal{X} (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

[𝗉k]SL2(),Hol().[\mathsf{p}_{k}]_{\text{SL}_{2}(\mathbb{Z}),\text{Hol}(\mathbb{H})}.

Now suppose 𝒳\mathcal{X} is a trivial 𝒢\mathcal{G}-set, i.e., 𝒳\mathcal{X} consists of 𝒢\mathcal{G}-invariant functions. Then there is a natural isomorphism

H1(𝒢,Fun(𝒳,Vk2,))Fun(𝒳,H1(𝒢,Vk2,));H^{1}\big{(}\mathcal{G},\text{Fun}(\mathcal{X},V_{k-2,\mathbb{C}})\big{)}\xrightarrow{\cong}\text{Fun}\big{(}\mathcal{X},H^{1}(\mathcal{G},V_{k-2,\mathbb{C}})\big{)};

see Section 4.1. We define the Eichler-Shimura homomorphism

[ESk]𝒢,𝒳:𝒳H1(𝒢,Vk2,)[\text{ES}_{k}]_{\mathcal{G},\mathcal{X}}:\mathcal{X}\to H^{1}(\mathcal{G},V_{k-2,\mathbb{C}})

to be the image of [𝗉k]𝒢,𝒳[\mathsf{p}_{k}]_{\mathcal{G},\mathcal{X}} 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 [ESk]𝒢,𝒳[\text{ES}_{k}]_{\mathcal{G},\mathcal{X}} that plays a crucial role in the proof of the isomorphism theorem. Let the notation be as in the discussion above. Moreover, suppose that \mathcal{H} is a subgroup of SL2()\text{SL}_{2}(\mathbb{Z}) that contains 𝒢\mathcal{G} as a normal subgroup and 𝒳\mathcal{X} is closed under |k\lvert_{k} action of \mathcal{H}. Note that \mathcal{H} acts on H1(𝒢,Vk2,)H^{1}(\mathcal{G},V_{k-2,\mathbb{C}}) by conjugation.

Proposition 1.2.

(Proposition 4.7) The Eichler-Shimura map [ESk]𝒢,𝒳[\emph{ES}_{k}]_{\mathcal{G},\mathcal{X}} is compatible with the action of \mathcal{H} 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 τ0\tau_{0} replaced by \infty. We treat this problem for a class of holomorphic functions that admit decomposition of the form f=f+fcf=f_{\infty}+f_{c} where ff_{\infty} is an exponential polynomial and fcf_{c} is a function with cuspidal decay at \infty (Section 3). Given such a function ff one can write down a solution of (1.3) as follows:

(1.6) 0τf(ξ)(Xξ)k2𝑑ξτfc(ξ)(Xξ)k2𝑑ξ.\int^{\tau}_{0}f_{\infty}(\xi)(X-\xi)^{k-2}d\xi-\int^{\infty}_{\tau}f_{c}(\xi)(X-\xi)^{k-2}d\xi.

Let 𝒳\mathcal{X} be a 𝒢\mathcal{G}-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 SL2()\text{SL}_{2}(\mathbb{Z}). 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 k(𝒢)=k(𝒢)𝒮k(𝒢)\mathcal{M}_{k}(\mathcal{G})=\mathcal{E}_{k}(\mathcal{G})\oplus\mathcal{S}_{k}(\mathcal{G}) where k(𝒢)\mathcal{E}_{k}(\mathcal{G}) is the space of Eisenstein series. To prove the theorem it suffices to check that the Eichler-Shimura map attached to k(𝒢)\mathcal{E}_{k}(\mathcal{G}) is injective and the image of k(𝒢)\mathcal{E}_{k}(\mathcal{G}) in H1(𝒢,Vk2,)H^{1}(\mathcal{G},V_{k-2,\mathbb{C}}) 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 𝒢\mathcal{G} 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 Γ\Gamma 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 𝒢\mathcal{G} be a congruence subgroup of Γ\Gamma. There exists an induced Eichler-Shimura homomorphism

k(𝒢)H1(Γ,Fun(𝒢\Γ,Vk2,))\mathcal{M}_{k}(\mathcal{G})\to H^{1}\big{(}\Gamma,\emph{Fun}(\mathcal{G}\backslash\Gamma,V_{k-2,\mathbb{C}})\big{)}

extending the map for the cusp forms so that the extended map gives rise to a Hecke equivariant isomorphism k(𝒢)𝒮ka(𝒢)H1(Γ,Fun(𝒢\Γ,Vk2,))\mathcal{M}_{k}(\mathcal{G})\oplus\mathcal{S}^{a}_{k}(\mathcal{G})\xrightarrow{\cong}H^{1}\big{(}\Gamma,\emph{Fun}(\mathcal{G}\backslash\Gamma,V_{k-2,\mathbb{C}})\big{)}.

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 χ\chi be a Dirichlet character modulo NN. Suppose that k(N,χ)\mathcal{M}_{k}(N,\chi), resp. 𝒮k(N,χ)\mathcal{S}_{k}(N,\chi), is the space of all modular forms, resp. cusp forms, on Γ1(N)\Gamma_{1}(N) that transform according to the nebentypus character χ\chi under Γ0(N)\Gamma_{0}(N). As before set 𝒮ka(N,χ):={f¯f𝒮k(N,χ¯)}\mathcal{S}_{k}^{a}(N,\chi):=\{\bar{f}\mid f\in\mathcal{S}_{k}(N,\bar{\chi})\}. Here one needs to replace the module of coefficients by its twisted counterpart, denoted Vk2,χV_{k-2,\mathbb{C}}^{\chi}, on which Γ0(N)\Gamma_{0}(N) action is twisted by the character χ1\chi^{-1}. 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

k(N,χ)𝒮ka(N,χ)H1(Γ0(N),Vk2,χ).\mathcal{M}_{k}(N,\chi)\oplus\mathcal{S}^{a}_{k}(N,\chi)\xrightarrow{\cong}H^{1}\big{(}\Gamma_{0}(N),V_{k-2,\mathbb{C}}^{\chi}\big{)}.

The final topic of the article is an explicit computation of the period polynomials attached to the Eisenstein series (Section 8). Recall that if 𝕂\mathbb{K} is a subfield of \mathbb{C} and 𝒢\mathcal{G} is a finite index subgroup then the natural inclusion Vk2,𝕂Vk2,V_{k-2,\mathbb{K}}\hookrightarrow V_{k-2,\mathbb{C}} induces an isomorphism H1(𝒢,Vk2,𝕂)𝕂H1(𝒢,Vk2,)H^{1}(\mathcal{G},V_{k-2,\mathbb{K}})\otimes_{\mathbb{K}}\mathbb{C}\cong H^{1}(\mathcal{G},V_{k-2,\mathbb{C}}); see (5.4). A class αH1(𝒢,Vk2,)\mathcal{\alpha}\in H^{1}(\mathcal{G},V_{k-2,\mathbb{C}}) is definable over the subfield 𝕂\mathbb{K} if α\alpha lies in the canonical 𝕂\mathbb{K}-form H1(𝒢,Vk2,𝕂)H^{1}(\mathcal{G},V_{k-2,\mathbb{K}}). Now suppose 𝒢\mathcal{G} is a congruence subgroup of SL2()\text{SL}_{2}(\mathbb{Z}). Let 𝒢,k\mathcal{B}_{\mathcal{G},k} denote the basis of the space of Eisenstein series k(𝒢)\mathcal{E}_{k}(\mathcal{G}) described in Section 5.3.

Theorem 1.5.

(Theorem 8.8) Let 𝒢\mathcal{G} be a congruence subgroup containing Γ(N)\Gamma(N) and f𝒢,kf\in\mathcal{B}_{\mathcal{G},k}. Then the image of ff under the Eichler-Shimura map for 𝒢\mathcal{G} is definable over (𝐞(1N))\mathbb{Q}(\mathbf{e}(\frac{1}{N})).

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 SL2()\text{SL}_{2}(\mathbb{Z}) by an appropriate coboundary so that the modified cocycle has coefficients in \mathbb{Q} [20, 7]. We first establish a generalization of their observation for Γ(N)\Gamma(N) 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 ξ\xi and τ\tau are variables with values in the upper half-plane \mathbb{H}. Let ii denote the imaginary unit. For τ\tau\in\mathbb{H} one writes τ=u(τ)+iv(τ)\tau=u(\tau)+iv(\tau) where u(τ)u(\tau) and v(τ)v(\tau) are real numbers with v(τ)>0v(\tau)>0. We normalize the complex exponential function as e(z):=exp(2πiz)\textbf{e}(z):=\exp(2\pi iz). One uses 𝕂\mathbb{K} to refer to a field of characteristic 0 which in some places is assumed to be a subfield of \mathbb{C}. For a positive integer NN set

N:=(𝐞(1N)).\mathbb{Q}_{N}:=\mathbb{Q}(\mathbf{e}(\frac{1}{N})).

Let Hol()\text{Hol}(\mathbb{H}), resp. Ωhol1()\Omega^{1}_{\text{hol}}(\mathbb{H}), be the space of all \mathbb{C}-valued holomorphic functions, resp. holomorphic differential 11-forms, on \mathbb{H}. The constant functions define a natural inclusion Hol()\mathbb{C}\hookrightarrow\text{Hol}(\mathbb{H}) which turns Hol()\text{Hol}(\mathbb{H}) into a commutative \mathbb{C}-algebra. We have a \mathbb{C}-linear differential operator

τ:Hol()Ωhol1(),ffτdτ\partial_{\tau}:\text{Hol}(\mathbb{H})\to\Omega^{1}_{\text{hol}}(\mathbb{H}),\hskip 8.5359ptf\to\frac{\partial f}{\partial\tau}d\tau

whose kernel equals the subspace of constant functions in Hol()\text{Hol}(\mathbb{H}). During the calculations, one considers integrals along paths on \mathbb{H}. All paths are piecewise smooth unless otherwise stated.

We abbreviate SL2()\text{SL}_{2}(\mathbb{Z}) as Γ\Gamma and write

T:=(1101),S:=(0110)T:=\begin{pmatrix}1&1\\ 0&1\end{pmatrix},\hskip 8.5359ptS:=\begin{pmatrix}0&-1\\ 1&0\end{pmatrix}

for the generators of this group. Sometimes one may refer to the subgroups of GL2()\text{GL}_{2}(\mathbb{R}). For a subring RR of \mathbb{R} set GL2+(R):={γGL2(R)det(γ)>0}\text{GL}^{+}_{2}(R):=\{\gamma\in\text{GL}_{2}(R)\mid\text{det}(\gamma)>0\}. The letter γ\gamma stands for an arbitrary 2×22\times 2 matrix and we put γ=(abcd)\gamma=\begin{pmatrix}a&b\\ c&d\end{pmatrix} where a,b,c,da,b,c,d take values in an appropriate ring. One uses the notation 𝒞()\mathcal{C}(\cdot), resp. 𝒞()\mathcal{C}_{\infty}(\cdot), to denote the set of cusps, resp. regular cusps, attached to a finite index subgroup of Γ\Gamma. Our treatment also makes use of the standard congruence subgroups namely Γ(N)\Gamma(N), Γ1(N)\Gamma_{1}(N), and Γ0(N)\Gamma_{0}(N) [7, 1.2]. We call Γ(N)\Gamma(N) the principal congruence subgroup of level NN. For a subgroup 𝒢\mathcal{G} of SL2()\text{SL}_{2}(\mathbb{Z}) let

P𝒢:=𝒢/𝒢{±Id}\text{P}\mathcal{G}:=\mathcal{G}/\mathcal{G}\cap\{\pm\text{Id}\}

denote the projectivization of 𝒢\mathcal{G}.

The group actions relevant to us are usually right actions. In the context of group cohomology Z1Z^{1}, resp. B1B^{1}, denotes the collection of 11-cocycles, 11-coboundaries. If 𝒳,𝒴\mathcal{X},\mathcal{Y} are two sets then Fun(𝒳,𝒴)\text{Fun}(\mathcal{X},\mathcal{Y}) is the set of all set-theoretic functions from 𝒳\mathcal{X} to 𝒴\mathcal{Y}. Suppose that 𝒴\mathcal{Y} is a 𝔽\mathbb{F}-vector space for some field 𝔽\mathbb{F}. Then Fun(𝒳,𝒴)\text{Fun}(\mathcal{X},\mathcal{Y}) possesses a 𝔽\mathbb{F}-linear structure defined by pointwise addition and scalar multiplication. Now suppose GG is an abstract group and 𝒳\mathcal{X}, 𝒴\mathcal{Y} are right GG-sets. The set Fun(𝒳,𝒴)\text{Fun}(\mathcal{X},\mathcal{Y}) admits a right GG-action given by

(1.7) (F.g)(x)=F(xg1)g.(FFun(𝒳,𝒴),x𝒳,gG)(F.g)(x)=F(xg^{-1})g.\hskip 28.45274pt\big{(}\begin{subarray}{c}F\in\text{Fun}(\mathcal{X},\mathcal{Y}),x\in\mathcal{X},\,g\in G\end{subarray}\big{)}

If 𝒴\mathcal{Y} is a right 𝔽[G]\mathbb{F}[G]-module then the resulting 𝔽\mathbb{F}-linear structure and GG-structure turn Fun(𝒳,𝒴)\text{Fun}(\mathcal{X},\mathcal{Y}) into a right 𝔽[G]\mathbb{F}[G]-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 \mathbb{H}.

The group GL2+()\text{GL}_{2}^{+}(\mathbb{R}) acts on \mathbb{H} from left by holomorphic automorphisms via fractional linear transformations;

(2.1) (γ,τ)γτ:=aτ+bcτ+d,γ=(abcd)GL2+().(\gamma,\tau)\to\gamma\tau:=\frac{a\tau+b}{c\tau+d},\hskip 8.5359pt\gamma=\begin{pmatrix}a&b\\ c&d\end{pmatrix}\in\text{GL}_{2}^{+}(\mathbb{R}).

The restriction of this action to Γ\Gamma is our primary interest. The Γ\Gamma-action on \mathbb{H} descends to a PΓ\text{P}\Gamma-action on \mathbb{H}, i.e, IdΓ-\text{Id}\in\Gamma acts trivially on each point. Note that the fractional linear transformations (2.1) define a holomorphic left Γ\Gamma-action on the Riemann sphere ^={}\widehat{\mathbb{C}}=\mathbb{C}\cup\{\infty\}. The boundary of \mathbb{H} in ^\widehat{\mathbb{C}} equals ={}=1()\partial\mathbb{H}=\{\infty\}\cup\mathbb{R}=\mathbb{P}^{1}(\mathbb{R}) and the Γ\Gamma-action on ^\widehat{\mathbb{C}} preserves this boundary. We define the extended upper-half plane as

:=1().\mathbb{H}^{\ast}:=\mathbb{H}\cup\mathbb{P}^{1}(\mathbb{Q}).

For x1()x\in\mathbb{P}^{1}(\mathbb{Q}) let Γx\Gamma_{x} denote the stabilizer of xx in Γ\Gamma. With this notation Γ={±Tnn}\Gamma_{\infty}=\{\pm T^{n}\mid n\in\mathbb{Z}\}. If 𝒢\mathcal{G} is a subgroup of Γ\Gamma then set 𝒢x:=𝒢Γx\mathcal{G}_{x}:=\mathcal{G}\cap\Gamma_{x}. Now suppose 𝒢\mathcal{G} is a finite index subgroup. Then a cusp of 𝒢\mathcal{G} refers to a 𝒢\mathcal{G}-orbit in 1()\mathbb{P}^{1}(\mathbb{Q}).

Group action on function spaces.

Let kk\in\mathbb{Z}. We define the |k\lvert_{k} right action of GL2+()\text{GL}_{2}^{+}(\mathbb{R}) on Hol()\text{Hol}(\mathbb{H}) by

(2.2) f|k,γ(τ):=det(γ)k1(cτ+d)kf(γτ).(fHol(),γGL2+())f\lvert_{k,\gamma}(\tau):=\text{det}(\gamma)^{k-1}(c\tau+d)^{-k}f(\gamma\tau).\hskip 14.22636pt\big{(}\begin{subarray}{c}f\in\text{Hol}(\mathbb{H}),\gamma\in\text{GL}_{2}^{+}(\mathbb{R})\end{subarray}\big{)}

As before one primarily uses restriction of this action to Γ\Gamma. If kk remains fixed throughout the discussion then we drop it from the notation.

Representations of GL2()\text{GL}_{2}(\mathbb{Z}).

Let nn be a nonnegative integer. Set Vn:={P[X]deg(P)n}V_{n}:=\{P\in\mathbb{Z}[X]\mid\text{deg}(P)\leq n\}. The group GL2()\text{GL}_{2}(\mathbb{Z}) acts on VnV_{n} from right by

(2.3) P(X)|n,γ:=(cX+d)nP(aX+bcX+d).(P(X)Vn,γGL2())P(X)\lvert_{-n,\gamma}:=(cX+d)^{n}P\Big{(}\frac{aX+b}{cX+d}\Big{)}.\hskip 8.5359pt\big{(}\begin{subarray}{c}P(X)\in V_{n},\gamma\in\text{GL}_{2}(\mathbb{Z})\end{subarray}\big{)}

We omit the index of weight from the subscript if it is clear from the context. More generally, let RR be a commutative ring and CC be a commutative RR-algebra. Write Vn,C=CVnV_{n,C}=C\otimes_{\mathbb{Z}}V_{n}. Then (2.3) defines a right C[GL2(R)]C[\text{GL}_{2}(R)] action on Vn,CV_{n,C}. The pair (,C)(\mathbb{Z},C) yields a C[GL2()]C[\text{GL}_{2}(\mathbb{Z})] module structure on Vn,CV_{n,C}. In this article, CC is typically a field of characteristic zero and we consider the cohomology vector spaces H(𝒢,Vk2,C)H^{\ast}(\mathcal{G},V_{k-2,C}) for the subgroups 𝒢\mathcal{G} of Γ\Gamma.

Differential operators.

The action of τ\partial_{\tau} on the coefficients gives rise to a \mathbb{C}-linear differential operator

(2.4) τ:Hol()[X]nΩhol1()Vn,(n0)\partial_{\tau}:\text{Hol}(\mathbb{H})[X]_{n}\to\Omega_{\text{hol}}^{1}(\mathbb{H})\otimes_{\mathbb{C}}V_{n,\mathbb{C}}\hskip 28.45274pt(\begin{subarray}{c}n\in\mathbb{Z}_{\geq 0}\end{subarray})

where Hol()[X]n={PHol()[X]deg(P)n}\text{Hol}(\mathbb{H})[X]_{n}=\{P\in\text{Hol}(\mathbb{H})[X]\mid\text{deg}(P)\leq n\}. The kernel of (2.4) is equal to Vn,Hol()[X]nV_{n,\mathbb{C}}\subseteq\text{Hol}(\mathbb{H})[X]_{n}.

2.2 Period cocycle and Eichler-Shimura class

Definition 2.1.

Let kk be an integer 2\geq 2 and fHol()f\in\text{Hol}(\mathbb{H}). A weight kk Eichler-Shimura integral of ff is a polynomial 𝕀k(τ,f,X)Hol()[X]k2\mathbb{I}_{k}(\tau,f,X)\in\text{Hol}(\mathbb{H})[X]_{k-2} that satisfies τ(𝕀k(τ,f,X))=f(τ)(Xτ)k2dτ\partial_{\tau}\big{(}\mathbb{I}_{k}(\tau,f,X)\big{)}=f(\tau)(X-\tau)^{k-2}d\tau.

Now suppose 𝒳\mathcal{X} is a collection of holomorphic functions on \mathbb{H}. A choice of weight kk Eichler-Shimura integrals for 𝒳\mathcal{X} is a set-theoretic function

𝕀k,𝒳:𝒳Hol()[X]k2,f𝕀k,𝒳(τ,f,X);τ(𝕀k,𝒳(τ,f,X))=f(τ)(Xτ)k2dτ.\begin{gathered}\mathbb{I}_{k,\mathcal{X}}:\mathcal{X}\to\text{Hol}(\mathbb{H})[X]_{k-2},\hskip 8.5359ptf\to\mathbb{I}_{k,\mathcal{X}}(\tau,f,X);\\ \partial_{\tau}\big{(}\mathbb{I}_{k,\mathcal{X}}(\tau,f,X)\big{)}=f(\tau)(X-\tau)^{k-2}d\tau.\end{gathered}

Note that two choice functions, say 𝕀k,𝒳1\mathbb{I}^{1}_{k,\mathcal{X}} and 𝕀k,𝒳2\mathbb{I}^{2}_{k,\mathcal{X}}, give rise to a set map

(𝕀k,𝒳1𝕀k,𝒳2):𝒳Vk2,,f𝕀k,𝒳1(τ,f,X)𝕀k,𝒳2(τ,f,X).\big{(}\mathbb{I}^{1}_{k,\mathcal{X}}-\mathbb{I}^{2}_{k,\mathcal{X}}\big{)}:\mathcal{X}\to V_{k-2,\mathbb{C}},\hskip 8.5359ptf\to\mathbb{I}^{1}_{k,\mathcal{X}}(\tau,f,X)-\mathbb{I}^{2}_{k,\mathcal{X}}(\tau,f,X).
Example 2.2.

Let τ0\tau_{0}\in\mathbb{H}. The canonical choice for 𝒳\mathcal{X} with respect to τ0\tau_{0} is

𝕀k,𝒳τ0:𝒳Hol()[X]k2,fτ0τf(ξ)(Xξ)k2𝑑ξ.\mathbb{I}_{k,\mathcal{X}}^{\tau_{0}}:\mathcal{X}\to\text{Hol}(\mathbb{H})[X]_{k-2},\hskip 8.5359ptf\to\int_{\tau_{0}}^{\tau}f(\xi)(X-\xi)^{k-2}d\xi.

This choice function plays a key role throughout the paper; cf. (1.4).

Let 𝒢\mathcal{G} be a finite index subgroup of Γ\Gamma. Suppose that 𝒳\mathcal{X} is a subset of Hol()\text{Hol}(\mathbb{H}) which is stable under the |k\lvert_{k} action of 𝒢\mathcal{G}. The 𝒢\mathcal{G}-actions on 𝒳\mathcal{X} and Vk2,V_{k-2,\mathbb{C}} induce a right [𝒢]\mathbb{C}[\mathcal{G}] module structure on Fun(𝒳,Vk2,)\text{Fun}(\mathcal{X},V_{k-2,\mathbb{C}}) according to the recipe (1.7). Let 𝕀k,𝒳\mathbb{I}_{k,\mathcal{X}} be a choice of weight kk Eichler-Shimura integrals for 𝒳\mathcal{X}. For γ𝒢\gamma\in\mathcal{G} and f𝒳f\in\mathcal{X} put

(2.5) 𝗉k,𝒳(γ,f,X):=(cX+d)k2𝕀k,𝒳(γτ,f|γ1,aX+bcX+d)𝕀k,𝒳(τ,f,X).\mathsf{p}_{k,\mathcal{X}}(\gamma,f,X):=(cX+d)^{k-2}\mathbb{I}_{k,\mathcal{X}}(\gamma\tau,f\lvert_{\gamma^{-1}},\frac{aX+b}{cX+d})-\mathbb{I}_{k,\mathcal{X}}(\tau,f,X).

We now have the following result:

Lemma 2.3.

𝗉k,𝒳(γ,f,X)Vk2,\mathsf{p}_{k,\mathcal{X}}(\gamma,f,X)\in V_{k-2,\mathbb{C}}.

Proof.

One first verifies the assertion for the canonical choice with respect to τ0\tau_{0}\in\mathbb{H}. A direct calculation using the change of variable formula demonstrates that

(2.6) (cX+d)k2𝕀k,𝒳τ0(γτ,f|γ1,aX+bcX+d)𝕀k,𝒳τ0(τ,f,X)(f𝒳,γ𝒢)\displaystyle(cX+d)^{k-2}\mathbb{I}^{\tau_{0}}_{k,\mathcal{X}}(\gamma\tau,f\lvert_{\gamma^{-1}},\frac{aX+b}{cX+d})-\mathbb{I}^{\tau_{0}}_{k,\mathcal{X}}(\tau,f,X)\hskip 14.22636pt\big{(}\begin{subarray}{c}f\in\mathcal{X},\gamma\in\mathcal{G}\end{subarray}\big{)}
=γ1τ0τ0f(ξ)(Xξ)k2𝑑ξVk2,.\displaystyle=\int_{\gamma^{-1}\tau_{0}}^{\tau_{0}}f(\xi)(X-\xi)^{k-2}d\xi\in V_{k-2,\mathbb{C}}.

Therefore the statement holds for 𝕀k,𝒳τ0\mathbb{I}_{k,\mathcal{X}}^{\tau_{0}}. Note that

(2.7) (cX+d)k2𝕀k,𝒳(γτ,f|γ1,aX+bcX+d)(f𝒳,γ𝒢)\displaystyle(cX+d)^{k-2}\mathbb{I}_{k,\mathcal{X}}(\gamma\tau,f\lvert_{\gamma^{-1}},\frac{aX+b}{cX+d})\hskip 85.35826pt\big{(}\begin{subarray}{c}f\in\mathcal{X},\gamma\in\mathcal{G}\end{subarray}\big{)}
(cX+d)k2𝕀k,𝒳τ0(γτ,f|γ1,aX+bcX+d)\displaystyle\hskip 85.35826pt-(cX+d)^{k-2}\mathbb{I}^{\tau_{0}}_{k,\mathcal{X}}(\gamma\tau,f\lvert_{\gamma^{-1}},\frac{aX+b}{cX+d})
=(𝕀k,𝒳𝕀k,𝒳τ0)(f|γ1)|γVk2,.\displaystyle=(\mathbb{I}_{k,\mathcal{X}}-\mathbb{I}^{\tau_{0}}_{k,\mathcal{X}})(f\lvert_{\gamma^{-1}})\bigl{\lvert}_{\gamma}\in V_{k-2,\mathbb{C}}.

The identities (2.6) and (2.7) together imply

(2.8) 𝗉k,𝒳(γ,f,X)\displaystyle\mathsf{p}_{k,\mathcal{X}}(\gamma,f,X) =γ1τ0τ0f(ξ)(Xξ)k2dξ+(𝕀k,𝒳𝕀k,𝒳τ0)(f|γ1)|γ\displaystyle=\int^{\tau_{0}}_{\gamma^{-1}\tau_{0}}f(\xi)(X-\xi)^{k-2}d\xi+(\mathbb{I}_{k,\mathcal{X}}-\mathbb{I}^{\tau_{0}}_{k,\mathcal{X}})(f\lvert_{\gamma^{-1}})\bigl{\lvert}_{\gamma}
(𝕀k,𝒳𝕀k,𝒳τ0)(f)Vk2,.\displaystyle\hskip 113.81102pt-(\mathbb{I}_{k,\mathcal{X}}-\mathbb{I}^{\tau_{0}}_{k,\mathcal{X}})(f)\in V_{k-2,\mathbb{C}}.

Definition 2.4.

Let the notation be as above. The period function of the choice of Eichler-Shimura integrals 𝕀k,𝒳\mathbb{I}_{k,\mathcal{X}} is a function on 𝒢\mathcal{G} given by

𝗉k,𝒳:𝒢Fun(𝒳,Vk2,),γ𝗉k,𝒳(γ);𝗉k,𝒳(γ)(f):=𝗉k,𝒳(γ,f,X).\mathsf{p}_{k,\mathcal{X}}:\mathcal{G}\to\text{Fun}(\mathcal{X},V_{k-2,\mathbb{C}}),\hskip 5.69046pt\gamma\to\mathsf{p}_{k,\mathcal{X}}(\gamma);\hskip 8.5359pt\mathsf{p}_{k,\mathcal{X}}(\gamma)(f):=\mathsf{p}_{k,\mathcal{X}}(\gamma,f,X).
Example 2.5.

Let 𝗉k,𝒳τ0\mathsf{p}^{\tau_{0}}_{k,\mathcal{X}} be the period function of 𝕀k,𝒳τ0\mathbb{I}_{k,\mathcal{X}}^{\tau_{0}}. Then (2.6) amounts to the identity

(2.9) 𝗉k,𝒳τ0(γ,f,X)=γ1τ0τ0f(ξ)(Xξ)k2dξ.(f𝒳,γ𝒢)\mathsf{p}^{\tau_{0}}_{k,\mathcal{X}}(\gamma,f,X)=\int_{\gamma^{-1}\tau_{0}}^{\tau_{0}}f(\xi)(X-\xi)^{k-2}d\xi.\hskip 28.45274pt\big{(}\begin{subarray}{c}f\in\mathcal{X},\gamma\in\mathcal{G}\end{subarray}\big{)}

The following lemma checks that the period function is a cocycle.

Lemma 2.6.

Let 𝕀k,𝒳\mathbb{I}_{k,\mathcal{X}} be a choice of Eichler-Shimura integrals for 𝒳\mathcal{X} and 𝗉k,𝒳\mathsf{p}_{k,\mathcal{X}} be the period function of 𝕀k,𝒳\mathbb{I}_{k,\mathcal{X}}. Then 𝗉k,𝒳\mathsf{p}_{k,\mathcal{X}} is a cocycle on 𝒢\mathcal{G} with values in Fun(𝒳,Vk2,)\emph{Fun}(\mathcal{X},V_{k-2,\mathbb{C}}), i.e.,

(2.10) 𝗉k,𝒳(γ1γ2)=𝗉k,𝒳(γ1)|γ2+𝗉k,𝒳(γ2).(γ1,γ2𝒢)\mathsf{p}_{k,\mathcal{X}}(\gamma_{1}\gamma_{2})=\mathsf{p}_{k,\mathcal{X}}(\gamma_{1})\lvert_{\gamma_{2}}+\mathsf{p}_{k,\mathcal{X}}(\gamma_{2}).\hskip 28.45274pt(\begin{subarray}{c}\gamma_{1},\gamma_{2}\in\mathcal{G}\end{subarray})
Proof.

We start by proving the statement for 𝕀k,𝒳τ0\mathbb{I}_{k,\mathcal{X}}^{\tau_{0}}. Let γ1,γ2𝒢\gamma_{1},\gamma_{2}\in\mathcal{G} and f𝒳f\in\mathcal{X}. One employs (2.9) and the change of variable formula to discover that

𝗉k,𝒳τ0(γ1)|γ2(f)=γ21γ11τ0γ21τ0f(ξ)(Xξ)k2dξ.\mathsf{p}_{k,\mathcal{X}}^{\tau_{0}}(\gamma_{1})\lvert_{\gamma_{2}}(f)=\int^{\gamma_{2}^{-1}\tau_{0}}_{\gamma_{2}^{-1}\gamma_{1}^{-1}\tau_{0}}f(\xi)(X-\xi)^{k-2}d\xi.

The identity above together with (2.9) implies that

(2.11) 𝗉k,𝒳τ0(γ1γ2)(f)=𝗉k,𝒳τ0(γ1)|γ2(f)+𝗉k,𝒳τ0(γ2)(f).\mathsf{p}_{k,\mathcal{X}}^{\tau_{0}}(\gamma_{1}\gamma_{2})(f)=\mathsf{p}_{k,\mathcal{X}}^{\tau_{0}}(\gamma_{1})\lvert_{\gamma_{2}}(f)+\mathsf{p}_{k,\mathcal{X}}^{\tau_{0}}(\gamma_{2})(f).

Therefore (2.10) holds for 𝕀k,𝒳τ0\mathbb{I}_{k,\mathcal{X}}^{\tau_{0}}. Now the equation (2.8) translates into

(2.12) 𝗉k,𝒳(γ)=𝗉k,𝒳τ0(γ)+(𝕀k,𝒳𝕀k,𝒳τ0)|γ(𝕀k,𝒳𝕀k,𝒳τ0).(γ𝒢)\mathsf{p}_{k,\mathcal{X}}(\gamma)=\mathsf{p}_{k,\mathcal{X}}^{\tau_{0}}(\gamma)+\big{(}\mathbb{I}_{k,\mathcal{X}}-\mathbb{I}_{k,\mathcal{X}}^{\tau_{0}}\big{)}\bigl{\lvert}_{\gamma}-\big{(}\mathbb{I}_{k,\mathcal{X}}-\mathbb{I}_{k,\mathcal{X}}^{\tau_{0}}\big{)}.\hskip 28.45274pt(\begin{subarray}{c}\gamma\in\mathcal{G}\end{subarray})

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 𝗉k,𝒳\mathsf{p}_{k,\mathcal{X}} in H1(𝒢,Fun(𝒳,Vk2,))H^{1}\big{(}\mathcal{G},\emph{Fun}(\mathcal{X},V_{k-2,\mathbb{C}})\big{)} is independent of the choice function 𝕀k,𝒳\mathbb{I}_{k,\mathcal{X}}.

Proof.

Let τ0\tau_{0}\in\mathbb{H} be a fixed base point and 𝕀k,𝒳\mathbb{I}_{k,\mathcal{X}} be an arbitrary choice function. The identity (2.12) implies that 𝗉k,𝒳\mathsf{p}_{k,\mathcal{X}} and 𝗉k,𝒳τ0\mathsf{p}_{k,\mathcal{X}}^{\tau_{0}} differ by the coboundary defined by (𝕀k,𝒳𝕀k,𝒳τ0)(\mathbb{I}_{k,\mathcal{X}}-\mathbb{I}^{\tau_{0}}_{k,\mathcal{X}}). Therefore the images of 𝗉k,𝒳\mathsf{p}_{k,\mathcal{X}} and 𝗉k,𝒳τ0\mathsf{p}_{k,\mathcal{X}}^{\tau_{0}} in cohomology are equal. ∎

Proposition 2.7 proves that the pair (𝒢,𝒳)(\mathcal{G},\mathcal{X}) determines the image of 𝗉k,𝒳\mathsf{p}_{k,\mathcal{X}}.

Definition 2.8.

The Eichler-Shimura class of weight kk attached to (𝒢,𝒳)(\mathcal{G},\mathcal{X}), denoted [𝗉k]𝒢,𝒳[\mathsf{p}_{k}]_{\mathcal{G},\mathcal{X}}, is the image of 𝗉k,𝒳\mathsf{p}_{k,\mathcal{X}} in H1(𝒢,Fun(𝒳,Vk2,))H^{1}\big{(}\mathcal{G},\text{Fun}(\mathcal{X},V_{k-2,\mathbb{C}})\big{)}.

The Eichler-Shimura class exhibits good functorial properties concerning restriction of the corresponding set of functions and subgroup; cf. [18, 3]. Let 𝒳2𝒳1\mathcal{X}_{2}\subseteq\mathcal{X}_{1} be two 𝒢\mathcal{G}-stable subsets of Hol()\text{Hol}(\mathbb{H}). Now restriction induces a [𝒢]\mathbb{C}[\mathcal{G}]-linear homomorphism Res𝒳2𝒳1:Fun(𝒳1,Vk2,)Fun(𝒳2,Vk2,)\text{Res}^{\mathcal{X}_{1}}_{\mathcal{X}_{2}}:\text{Fun}(\mathcal{X}_{1},V_{k-2,\mathbb{C}})\to\text{Fun}(\mathcal{X}_{2},V_{k-2,\mathbb{C}}) that gives rise to a \mathbb{C}-linear map

R𝒳2𝒳1:H1(𝒢,Fun(𝒳1,Vk2,))H1(𝒢,Fun(𝒳2,Vk2,)).\text{R}^{\mathcal{X}_{1}}_{\mathcal{X}_{2}}:H^{1}\big{(}\mathcal{G},\text{Fun}(\mathcal{X}_{1},V_{k-2,\mathbb{C}})\big{)}\to H^{1}\big{(}\mathcal{G},\text{Fun}(\mathcal{X}_{2},V_{k-2,\mathbb{C}})\big{)}.

Let 𝕀k,𝒳1\mathbb{I}_{k,\mathcal{X}_{1}} be a choice of weight kk Eichler-Shimura integrals for 𝒳1\mathcal{X}_{1}. Set 𝕀k,𝒳2:=𝕀k,𝒳1|𝒳2\mathbb{I}_{k,\mathcal{X}_{2}}:=\mathbb{I}_{k,\mathcal{X}_{1}}\bigl{\lvert}_{\mathcal{X}_{2}}. One computes the corresponding Eichler-Shimura classes with these compatible choice functions to discover that

(2.13) [𝗉k]𝒢,𝒳2=R𝒳2𝒳1([𝗉k]𝒢,𝒳1).[\mathsf{p}_{k}]_{\mathcal{G},\mathcal{X}_{2}}=\text{R}^{\mathcal{X}_{1}}_{\mathcal{X}_{2}}\big{(}[\mathsf{p}_{k}]_{\mathcal{G},\mathcal{X}_{1}}\big{)}.

Now suppose 𝒢2𝒢1\mathcal{G}_{2}\subseteq\mathcal{G}_{1} are two finite index subgroups of Γ\Gamma. Let 𝒳\mathcal{X} be a 𝒢1\mathcal{G}_{1}-stable subset of Hol()\text{Hol}(\mathbb{H}). The restriction Res𝒢2𝒢1:Fun(𝒢1,Fun(𝒳,Vk2,))Fun(𝒢2,Fun(𝒳,Vk2,))\text{Res}^{\mathcal{G}_{1}}_{\mathcal{G}_{2}}:\text{Fun}\big{(}\mathcal{G}_{1},\text{Fun}(\mathcal{X},V_{k-2,\mathbb{C}})\big{)}\to\text{Fun}\big{(}\mathcal{G}_{2},\text{Fun}(\mathcal{X},V_{k-2,\mathbb{C}})\big{)} yields change of group map

R𝒢2𝒢1:H1(𝒢1,Fun(𝒳,Vk2,))H1(𝒢2,Fun(𝒳,Vk2,)).\text{R}^{\mathcal{G}_{1}}_{\mathcal{G}_{2}}:H^{1}\big{(}\mathcal{G}_{1},\text{Fun}(\mathcal{X},V_{k-2,\mathbb{C}})\big{)}\to H^{1}\big{(}\mathcal{G}_{2},\text{Fun}(\mathcal{X},V_{k-2,\mathbb{C}})\big{)}.

As before one calculates the Eichler-Shimura classes using compatible choice functions to find

(2.14) [𝗉k]𝒢2,𝒳=R𝒢2𝒢1([𝗉k]𝒢1,𝒳).[\mathsf{p}_{k}]_{\mathcal{G}_{2},\mathcal{X}}=\text{R}^{\mathcal{G}_{1}}_{\mathcal{G}_{2}}([\mathsf{p}_{k}]_{\mathcal{G}_{1},\mathcal{X}}).
Global Eichler-Shimura class.

Let [𝗉k]Γ,Hol()[\mathsf{p}_{k}]_{\Gamma,\text{Hol}(\mathbb{H})} be the Eichler-Shimura class attached to the Γ\Gamma-stable set Hol()\text{Hol}(\mathbb{H}). Then, using (2.13) and (2.14), we conclude that

[𝗉k]𝒢,𝒳=R𝒳Hol()R𝒢Γ([𝗉k]Γ,Hol()).[\mathsf{p}_{k}]_{\mathcal{G},\mathcal{X}}=\text{R}^{\text{Hol}(\mathbb{H})}_{\mathcal{X}}\circ\text{R}^{\Gamma}_{\mathcal{G}}\big{(}[\mathsf{p}_{k}]_{\Gamma,\text{Hol}(\mathbb{H})}\big{)}.

The identity above shows that each Eichler-Shimura class arises as a restriction of the global class [𝗉k]Γ,Hol()[\mathsf{p}_{k}]_{\Gamma,\text{Hol}(\mathbb{H})}.

2.3 Variations of formalism

2.3.1 General subgroups of GL2+()\text{GL}_{2}^{+}(\mathbb{R})

The constructions of this section are completely algebraic and continue to work for an arbitrary subgroup of GL2+()\text{GL}_{2}^{+}(\mathbb{R}). Let 𝒢\mathcal{G} be a subgroup of GL2+()\text{GL}_{2}^{+}(\mathbb{R}) and 𝒳Hol()\mathcal{X}\subseteq\text{Hol}(\mathbb{H}) be stable under |k\lvert_{k} action of 𝒢\mathcal{G}. Note that (2.3) defines a right [GL2()]\mathbb{C}[\text{GL}_{2}(\mathbb{C})]-module structure on Vk2,V_{k-2,\mathbb{C}}. Therefore Fun(𝒳,Vk2,)\text{Fun}(\mathcal{X},V_{k-2,\mathbb{C}}) has a natural right [𝒢]\mathbb{C}[\mathcal{G}]-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 [𝗉k]𝒢,𝒳[\mathsf{p}_{k}]_{\mathcal{G},\mathcal{X}} in H1(𝒢,Fun(𝒳,Vk2,))H^{1}\big{(}\mathcal{G},\text{Fun}(\mathcal{X},V_{k-2,\mathbb{C}})\big{)}. The general setup described above yields a formalism for Fuchsian groups, cf. [16, 8]. The subgroups of GL2+()\text{GL}_{2}^{+}(\mathbb{Q}) appear in the context of the action of double coset operators. The cocycle formula for the subgroups of GL2+()\text{GL}_{2}^{+}(\mathbb{Q}) 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 f𝕀k(τ,f,X)f\to\mathbb{I}_{k}(\tau,f,X) 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 𝕂\mathbb{K} is a subfield of \mathbb{C} and 𝒳\mathcal{X} is a 𝕂[𝒢]\mathbb{K}[\mathcal{G}]-submodule of Hol()\text{Hol}(\mathbb{H}). A choice of Eichler-Shimura integrals for 𝒳\mathcal{X}, say 𝕀k,𝒳\mathbb{I}_{k,\mathcal{X}}, is 𝕂\mathbb{K}-linear if 𝕀k,𝒳\mathbb{I}_{k,\mathcal{X}} is a 𝕂\mathbb{K}-linear map.

Example 2.9.

The canonical choice 𝕀k,𝒳τ0\mathbb{I}_{k,\mathcal{X}}^{\tau_{0}} is a 𝕂\mathbb{K}-linear choice of Eichler-Shimura integrals.

If 𝕀k,𝒳1\mathbb{I}_{k,\mathcal{X}}^{1} and 𝕀k,𝒳2\mathbb{I}_{k,\mathcal{X}}^{2} are two 𝕂\mathbb{K}-linear choices for 𝒳\mathcal{X} then (𝕀k,𝒳1𝕀k,𝒳2):𝒳Vk2,\big{(}\mathbb{I}_{k,\mathcal{X}}^{1}-\mathbb{I}_{k,\mathcal{X}}^{2}\big{)}:\mathcal{X}\to V_{k-2,\mathbb{C}} is a 𝕂\mathbb{K}-linear map. Let 𝕀k,𝒳\mathbb{I}_{k,\mathcal{X}} be a 𝕂\mathbb{K}-linear choice of Eichler-Shimura integrals and 𝗉k,𝒳\mathsf{p}_{k,\mathcal{X}} be the period function of 𝕀k,𝒳\mathbb{I}_{k,\mathcal{X}}. It is clear that

𝗉k,𝒳(γ)Hom𝕂(𝒳,Vk2,)\mathsf{p}_{k,\mathcal{X}}(\gamma)\in\text{Hom}_{\mathbb{K}}(\mathcal{X},V_{k-2,\mathbb{C}}) for each γ𝒢\gamma\in\mathcal{G}.

Note that Hom𝕂(𝒳,Vk2,)\text{Hom}_{\mathbb{K}}(\mathcal{X},V_{k-2,\mathbb{C}}) is a [𝒢]\mathbb{C}[\mathcal{G}]-submodule of Fun(𝒳,Vk2,)\text{Fun}(\mathcal{X},V_{k-2,\mathbb{C}}). Hence 𝗉k,𝒳\mathsf{p}_{k,\mathcal{X}} is a cocycle with values in Hom𝕂(𝒳,Vk2,)\text{Hom}_{\mathbb{K}}(\mathcal{X},V_{k-2,\mathbb{C}}). The image of 𝗉k,𝒳\mathsf{p}_{k,\mathcal{X}} in H1(𝒢,Hom𝕂(𝒳,Vk2,))H^{1}\big{(}\mathcal{G},\text{Hom}_{\mathbb{K}}(\mathcal{X},V_{k-2,\mathbb{C}})\big{)} is independent of the 𝕂\mathbb{K}-linear choice function 𝕀k,𝒳\mathbb{I}_{k,\mathcal{X}} and determines the 𝕂\mathbb{K}-linear Eichler-Shimura class [𝗉k]𝒢,𝒳𝕂[\mathsf{p}_{k}]_{\mathcal{G},\mathcal{X}}^{\mathbb{K}}. The natural homomorphism in cohomology induced by the inclusion of Hom𝕂(𝒳,Vk2,)\text{Hom}_{\mathbb{K}}(\mathcal{X},V_{k-2,\mathbb{C}}) into Fun(𝒳,Vk2,)\text{Fun}(\mathcal{X},V_{k-2,\mathbb{C}}) maps [𝗉k]𝒢,𝒳𝕂[\mathsf{p}_{k}]_{\mathcal{G},\mathcal{X}}^{\mathbb{K}} to [𝗉k]𝒢,𝒳[\mathsf{p}_{k}]_{\mathcal{G},\mathcal{X}}.

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 𝒢\mathcal{G}. If kk is odd then the action of 𝒢\mathcal{G} on Fun(𝒳,Vk2,)\text{Fun}\big{(}\mathcal{X},V_{k-2,\mathbb{C}}\big{)} does not descend to P𝒢\text{P}\mathcal{G}. On the other hand the action of 𝒢\mathcal{G} on Hom𝕂(𝒳,Vk2,)\text{Hom}_{\mathbb{K}}\big{(}\mathcal{X},V_{k-2,\mathbb{C}}\big{)} always descends to P𝒢\text{P}\mathcal{G}. Moreover, we have the following lemma:

Lemma 2.10.

Assume that Id𝒢-\emph{Id}\in\mathcal{G}. Let 𝕀k,𝒳\mathbb{I}_{k,\mathcal{X}} be a 𝕂\mathbb{K}-linear choice function for 𝒳\mathcal{X} and 𝗉k,𝒳\mathsf{p}_{k,\mathcal{X}} be the period function of 𝕀k,𝒳\mathbb{I}_{k,\mathcal{X}}. Then

𝗉k,𝒳(γ)=𝗉k,𝒳(γ).(γ𝒢)\mathsf{p}_{k,\mathcal{X}}(\gamma)=\mathsf{p}_{k,\mathcal{X}}(-\gamma).\hskip 8.5359pt(\begin{subarray}{c}\gamma\in\mathcal{G}\end{subarray})
Proof.

The canonical choice function 𝕀k,𝒳τ0\mathbb{I}_{k,\mathcal{X}}^{\tau_{0}} is 𝕂\mathbb{K}-linear and (𝕀k,𝒳𝕀k,𝒳τ0)Hom𝕂(𝒳,Vk2,)(\mathbb{I}_{k,\mathcal{X}}-\mathbb{I}_{k,\mathcal{X}}^{\tau_{0}})\in\text{Hom}_{\mathbb{K}}(\mathcal{X},V_{k-2,\mathbb{C}}). But (2.9) implies that the identity holds for 𝗉k,𝒳τ0\mathsf{p}_{k,\mathcal{X}}^{\tau_{0}}. Now the assertion is a consequence of (2.12). ∎

If Id𝒢-\text{Id}\notin\mathcal{G} then the natural projection map 𝒢P𝒢\mathcal{G}\twoheadrightarrow\text{P}\mathcal{G} identifies the group with its projectivization. On the other hand, if Id𝒢-\text{Id}\in\mathcal{G} then Lemma 2.10 demonstrates that the linear cocycles on 𝒢\mathcal{G} gives rise to well-defined cocycles on P𝒢\text{P}\mathcal{G}. Thus, in both the cases, one obtains a projective Eichler-Shimura class

[𝗉k]P𝒢,𝒳𝕂H1(P𝒢,Hom𝕂(𝒳,Vk2,))[\mathsf{p}_{k}]_{\text{P}\mathcal{G},\mathcal{X}}^{\mathbb{K}}\in H^{1}\big{(}\text{P}\mathcal{G},\text{Hom}_{\mathbb{K}}(\mathcal{X},V_{k-2,\mathbb{C}})\big{)}

so that its pullback under 𝒢P𝒢\mathcal{G}\twoheadrightarrow\text{P}\mathcal{G} equals [𝗉k]𝒢,𝒳𝕂[\mathsf{p}_{k}]^{\mathbb{K}}_{\mathcal{G},\mathcal{X}}.

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 \mathbb{H} in the Riemann sphere equals ¯:={}\overline{\mathbb{H}}:=\mathbb{H}\cup\mathbb{R}\cup\{\infty\}. We endow the extended upper half-plane ¯\mathbb{H}^{\ast}\subseteq\overline{\mathbb{H}} with the subspace topology arising from ^\widehat{\mathbb{C}} because it is favorable for discussing path integrals. In the theory of Baily-Borel compactification, one extends the topology of \mathbb{H} to \mathbb{H}^{\ast} by prescribing special types of neighborhoods for the points of 1()\mathbb{P}^{1}(\mathbb{Q}) that define the so-called Satake topology [8, 2.1]. We next introduce the notion of triangular neighborhoods of the points of 1()\mathbb{P}^{1}(\mathbb{Q}) 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 x1()x\in\mathbb{P}^{1}(\mathbb{Q}) refers to a geodesic triangle on \mathbb{H}^{\ast} whose one vertex is at xx and two other vertices lie in \mathbb{H}.

Definition 3.1.

A triangular neighborhood of x1()x\in\mathbb{P}^{1}(\mathbb{Q}) is the intersection of an open neighborhood of xx in \mathbb{H}^{\ast} with the interior of an admissible triangle based at xx.

Note that if 𝒯\mathcal{T} is an admissible triangle based at xx then the interior of 𝒯\mathcal{T}, denoted int(𝒯)\text{int}(\mathcal{T}), is a triangular neighborhood of xx. The action of γΓ\gamma\in\Gamma maps a triangular neighborhood of xx onto a triangular neighborhood of γx\gamma x.

Example 3.2.

This example describes the triangular neighborhoods of \infty.

  1. (i)

    Let t1,t2,ht_{1},t_{2},h\in\mathbb{R} so that t1<t2t_{1}<t_{2} and h>0h>0. Set

    𝒩t1,t2h={τv(τ)>h,t1<u(τ)<t2}.\mathcal{N}^{h}_{t_{1},t_{2}}=\{\tau\in\mathbb{H}\mid v(\tau)>h,t_{1}<u(\tau)<t_{2}\}.

    If h>t2t12h>\frac{t_{2}-t_{1}}{2} then 𝒩t1,t2h\mathcal{N}^{h}_{t_{1},t_{2}} is a triangular neighborhood of \infty (Figure 1).

    Refer to caption
    Figure 1: The triangular neighborhood 𝒩t1,t2h\mathcal{N}^{h}_{t_{1},t_{2}}
  2. (ii)

    Let τ1,τ2\tau_{1},\tau_{2}\in\mathbb{H} and consider the geodesic triangle 𝒯\mathcal{T} with vertices at τ1,τ2,\tau_{1},\tau_{2},\infty. Write t1=u(τ1),t2=u(τ2)t_{1}=u(\tau_{1}),t_{2}=u(\tau_{2}). If t1=t2t_{1}=t_{2} then 𝒯\mathcal{T} is degenerate and int(𝒯)=\text{int}(\mathcal{T})=\emptyset. Without loss of generality assume that t1<t2t_{1}<t_{2}. The geodesic arc joining τ1\tau_{1} and τ2\tau_{2} is unique. Therefore there exists h0t2t12h_{0}\geq\frac{t_{2}-t_{1}}{2} such that int(𝒯)=𝒩t1,t2h𝒞h\text{int}(\mathcal{T})=\mathcal{N}_{t_{1},t_{2}}^{h}\cup\mathcal{C}_{h} for each h>h0h>h_{0} where 𝒞h\mathcal{C}_{h} is a relatively compact subset of \mathbb{H}, i.e., the closure of 𝒞h\mathcal{C}_{h} in \mathbb{H} is compact; cf. Figure 1. Now suppose WW is an open neighborhood \infty in \mathbb{H}^{\ast}. Choose h>h0h>h_{0} with 𝒩t1,t2hW\mathcal{N}_{t_{1},t_{2}}^{h}\subseteq W. It follows that int(𝒯)W=𝒩t1,t2h(𝒞hW)\text{int}(\mathcal{T})\cap W=\mathcal{N}_{t_{1},t_{2}}^{h}\cup(\mathcal{C}_{h}\cap W). Thus each nonempty triangular neighborhood of \infty is a union of 𝒩t1,t2h\mathcal{N}_{t_{1},t_{2}}^{h} and a relatively compact subset of \mathbb{H} for suitable t1,t2,ht_{1},t_{2},h\in\mathbb{R} with h>t2t12>0h>\frac{t_{2}-t_{1}}{2}>0.

Growth rate at infinity.

We first introduce a scale for the growth of a function in the triangular neighborhoods of \infty. Let α\alpha be a fixed real number. A function fHol()f\in\text{Hol}(\mathbb{H}) decays with exponent α\alpha at \infty if |f(τ)||τ|α\lvert f(\tau)\rvert\lvert\tau\rvert^{\alpha} is bounded on each triangular neighborhood of \infty. Note that the definition does not require the bound to be uniform. Write

𝔉α:={fHol()f decays with exponent α at }.\mathfrak{F}_{\alpha}:=\{f\in\text{Hol}(\mathbb{H})\mid\text{$f$ decays with exponent $\alpha$ at $\infty$}\}.

It is clear that 𝔉α\mathfrak{F}_{\alpha} is a \mathbb{C}-subspace of Hol()\text{Hol}(\mathbb{H}). Next, we provide an equivalent version of the decay condition at \infty that is useful for practical applications.

Lemma 3.3.

The following are equivalent:

  1. (i)

    ff decays with exponent α\alpha at \infty.

  2. (ii)

    |f(τ)|v(τ)α\lvert f(\tau)\rvert v(\tau)^{\alpha} is bounded on each of {𝒩t1,t2ht1<t2h>h0}\big{\{}\mathcal{N}^{h}_{t_{1},t_{2}}\mid\text{$t_{1}<t_{2}$, $h>h_{0}$}\big{\}} for some fixed h00h_{0}\geq 0.

Proof.

Given t1,t2,ht_{1},t_{2},h\in\mathbb{R} with t1<t2t_{1}<t_{2} and h>0h>0 there exists a positive constant C=C(t1,t2,h)>1C=C(t_{1},t_{2},h)>1 so that v(τ)|τ|Cv(τ)v(\tau)\leq\lvert\tau\rvert\leq Cv(\tau) on 𝒩t1,t2h\mathcal{N}^{h}_{t_{1},t_{2}}. 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 h0=1h_{0}=1 to discover that 𝔉α2𝔉α1\mathfrak{F}_{\alpha_{2}}\subseteq\mathfrak{F}_{\alpha_{1}} whenever α1α2\alpha_{1}\leq\alpha_{2}. Set

𝔉:=α𝔉α.\mathfrak{F}_{\infty}:=\bigcap_{\alpha\in\mathbb{R}}\mathfrak{F}_{\alpha}.
Growth near the real line.

The boundary points in 1(){}\mathbb{P}^{1}(\mathbb{Q})-\{\infty\} naturally play a role in the theory due to the action of Γ\Gamma on the extended upper half plane. Let α\alpha be a real number and x01(){}x_{0}\in\mathbb{P}^{1}(\mathbb{Q})-\{\infty\}. A function fHol()f\in\text{Hol}(\mathbb{H}) grows with exponent α\alpha at x0x_{0} if |f(τ)||τx0|α\lvert f(\tau)\rvert\lvert\tau-x_{0}\rvert^{\alpha} is bounded on each triangular neighborhood of x0x_{0}. Define

𝔉αx0:={fHol()f grows with exponent α at x0}.\mathfrak{F}_{\alpha}^{x_{0}}:=\{f\in\text{Hol}(\mathbb{H})\mid\text{$f$ grows with exponent $\alpha$ at $x_{0}$}\}.

It is clear that 𝔉αx0\mathfrak{F}^{x_{0}}_{\alpha} is a \mathbb{C}-subspace of Hol()\text{Hol}(\mathbb{H}). Note that |τx0|1\lvert\tau-x_{0}\rvert\leq 1 in a small enough neighborhood of x0x_{0}. Therefore 𝔉α1x0𝔉α2x0\mathfrak{F}_{\alpha_{1}}^{x_{0}}\subseteq\mathfrak{F}_{\alpha_{2}}^{x_{0}} if α1α2\alpha_{1}\leq\alpha_{2}. We write

𝔉x0:=α𝔉αx0.\mathfrak{F}_{-\infty}^{x_{0}}:=\bigcap_{\alpha\in\mathbb{R}}\mathfrak{F}_{\alpha}^{x_{0}}.
Exponential polynomials.

The Fourier expansion of a weakly holomorphic modular form at \infty involves a principal part of the Laurent expansion along with a constant term (Section 3.3). Let

𝒥n:={(κ1,,κn)0nκ1<<κn}.(n1)\mathcal{J}_{n}:=\{(\kappa_{1},\cdots,\kappa_{n})\in\mathbb{R}_{\geq 0}^{n}\mid\kappa_{1}<\cdots<\kappa_{n}\}.\hskip 8.5359pt(\begin{subarray}{c}n\geq 1\end{subarray})

For κ=(κ1,,κn)𝒥n\kappa=(\kappa_{1},\cdots,\kappa_{n})\in\mathcal{J}_{n} define 𝖤𝖯(κ)\mathsf{EP}(\kappa) to be the \mathbb{C}-subspace of Hol()\text{Hol}(\mathbb{H}) spanned by {e(κrτ)1rn}\big{\{}\textbf{e}(-\kappa_{r}\tau)\mid 1\leq r\leq n\big{\}}. An exponential polynomial at \infty of index κ\kappa is an element of 𝖤𝖯(κ)\mathsf{EP}(\kappa). The subspace 𝖤𝖯(0)\mathsf{EP}(0) consists of the constant functions. Set

𝖤𝖯:=κ𝒥n,n1𝖤𝖯(κ).\mathsf{EP}:=\bigcup_{\begin{subarray}{c}\kappa\in\mathcal{J}_{n},\\ n\geq 1\end{subarray}}\mathsf{EP}(\kappa).

Since 0\mathbb{R}_{\geq 0} is totally ordered 𝖤𝖯\mathsf{EP} is a \mathbb{C}-subspace of Hol()\text{Hol}(\mathbb{H}). Note that the polynomials in 𝖤𝖯\mathsf{EP} have unique holomorphic extensions to \mathbb{C}.

Lemma 3.4.

With notation above

𝖤𝖯𝔉α={0}.(α(0,])\mathsf{EP}\cap\mathfrak{F}_{\alpha}=\{0\}.\hskip 28.45274pt(\begin{subarray}{c}\alpha\in(0,\infty]\end{subarray})
Proof.

Suppose f𝖤𝖯𝔉αf\in\mathsf{EP}\cap\mathfrak{F}_{\alpha}. Then f(iv)0f(iv)\to 0 as vv\to\infty. It follows that ff must be the zero exponential polynomial. Therefore 𝖤𝖯𝔉α={0}\mathsf{EP}\cap\mathfrak{F}_{\alpha}=\{0\}. ∎

Let α(0,]\alpha\in(0,\infty]. Define

𝖤𝖯𝔉α:=𝖤𝖯+𝔉αHol().\mathsf{EP}\mathfrak{F}_{\alpha}:=\mathsf{EP}+\mathfrak{F}_{\alpha}\subseteq\text{Hol}(\mathbb{H}).

Lemma 3.4 shows that 𝖤𝖯𝔉α=𝖤𝖯𝔉α\mathsf{EP}\mathfrak{F}_{\alpha}=\mathsf{EP}\oplus\mathfrak{F}_{\alpha}. Suppose that π:𝖤𝖯𝔉α𝖤𝖯\pi_{\infty}:\mathsf{EP}\mathfrak{F}_{\alpha}\to\mathsf{EP} and πc:𝖤𝖯𝔉α𝔉α\pi_{c}:\mathsf{EP}\mathfrak{F}_{\alpha}\to\mathfrak{F}_{\alpha} are the projection maps. For brevity one writes f:=π(f)f_{\infty}:=\pi_{\infty}(f) and fc:=πc(f)f_{c}:=\pi_{c}(f).

The final topic of the current subsection is the action of conformal automorphisms on growth families. With (λ,μ)>0×(\lambda,\mu)\in\mathbb{R}_{>0}\times\mathbb{R} we attach a conformal automorphism

ϕλ,μ:,τλτ+μ.\phi_{\lambda,\mu}:\mathbb{H}\to\mathbb{H},\hskip 8.5359pt\tau\to\lambda\tau+\mu.

Lemma 3.3 implies that fϕλ,μ𝔉αf\circ\phi_{\lambda,\mu}\in\mathfrak{F}_{\alpha} whenever f𝔉αf\in\mathfrak{F}_{\alpha} for some α(,]\alpha\in(-\infty,\infty]. It is clear that if f𝖤𝖯(κ)f\in\mathsf{EP}(\kappa) then fϕλ,μ𝖤𝖯(λκ)f\circ\phi_{\lambda,\mu}\in\mathsf{EP}(\lambda\kappa). Therefore 𝖤𝖯\mathsf{EP} is also closed under ϕλ,μ\phi_{\lambda,\mu}.

Lemma 3.5.

Let kk\in\mathbb{Z}, α(0,]\alpha\in(0,\infty], and γΓ\gamma\in\Gamma.

  1. (i)

    If γΓ\gamma\in\Gamma_{\infty} then both 𝖤𝖯\mathsf{EP} and 𝔉α\mathfrak{F}_{\alpha} are stable under |k,γ\lvert_{k,\gamma}. Moreover

    (3.1) π(f|k,γ)=π(f)|k,γ,πc(f|k,γ)=πc(f)|k,γ.(γΓ,f𝖤𝖯𝔉α)\pi_{\infty}(f\lvert_{k,\gamma})=\pi_{\infty}(f)\lvert_{k,\gamma},\hskip 8.5359pt\pi_{c}(f\lvert_{k,\gamma})=\pi_{c}(f)\lvert_{k,\gamma}.\hskip 14.22636pt(\begin{subarray}{c}\gamma\in\Gamma_{\infty},f\in\mathsf{EP}\mathfrak{F}_{\alpha}\end{subarray})
  2. (ii)

    Now suppose γΓΓ\gamma\in\Gamma-\Gamma_{\infty} and f𝔉αf\in\mathfrak{F}_{\alpha}. Then f|k,γ𝔉kαdcf\lvert_{k,\gamma}\in\mathfrak{F}^{-\frac{d}{c}}_{k-\alpha}.

Proof.
  1. (i)

    Let γΓ\gamma\in\Gamma_{\infty} and fHol()f\in\text{Hol}(\mathbb{H}). Then f|k,γ(τ)=dkfϕ1,bd(τ)f\lvert_{k,\gamma}(\tau)=d^{-k}f\circ\phi_{1,\frac{b}{d}}(\tau). We conclude that both 𝖤𝖯\mathsf{EP} and 𝔉α\mathfrak{F}_{\alpha} are stable under |k,γ\lvert_{k,\gamma}. Now (3.1) is a consequence of Lemma 3.4.

  2. (ii)

    Suppose γΓΓ\gamma\in\Gamma-\Gamma_{\infty} and f𝔉αf\in\mathfrak{F}_{\alpha}. Here c0c\neq 0 and γ1()=dc\gamma^{-1}(\infty)=-\frac{d}{c}\in\mathbb{Q}. Assume that α>0\alpha\in\mathbb{R}_{>0}. Let 𝒩\mathcal{N} be a triangular neighborhood of dc-\frac{d}{c}. Then γ𝒩\gamma\mathcal{N} is a triangular neighborhood of \infty. Note that |aτ+b|\lvert a\tau+b\rvert is bounded away from zero on 𝒩\mathcal{N}. We use the decay estimates for γ𝒩\gamma\mathcal{N} to obtain C>0C>0 so that |f|k,γ(τ)||τ+dc|kαC\lvert f\lvert_{k,\gamma}(\tau)\rvert\lvert\tau+\frac{d}{c}\rvert^{k-\alpha}\leq C for each τ𝒩\tau\in\mathcal{N}. Since the statement holds for each α>0\alpha\in\mathbb{R}_{>0} it is true for α=\alpha=\infty also.

3.2 Eichler-Shimura integrals with base point at \infty

Our discussion primarily concerns the following classes of continuous paths on the extended upper half-plane:

  • Let τ1,τ2{}\tau_{1},\tau_{2}\in\mathbb{H}^{\ast}-\{\infty\}. A good path between τ1\tau_{1} and τ2\tau_{2} refers to a piecewise smooth path on \mathbb{H}^{\ast} that maps (0,1)(0,1) inside \mathbb{H}.

  • The unique good path joining τ{}\tau\in\mathbb{H}^{\ast}-\{\infty\} and \infty is the vertical ray originating at τ\tau.

In the sequel the path integrals are along good paths joining the endpoints.

Lemma 3.6.

Let τ\tau\in\mathbb{H} and x1(){}x\in\mathbb{P}^{1}(\mathbb{Q})-\{\infty\}. Suppose that f𝔉0xf\in\mathfrak{F}^{x}_{0}, i.e., ff is bounded on each triangular neighborhood of xx. Then

xτf(ξ)𝑑ξ\int^{\tau}_{x}f(\xi)d\xi

does not depend on the choice of good path joining the endpoints.

Proof.

Let γ1,γ2\gamma_{1},\gamma_{2} be two good paths joining τ\tau and xx. We use piecewise smoothness to obtain an admissible triangle 𝒯\mathcal{T} at xx and 0<ϵ010<\epsilon_{0}\ll 1 so that γ1([1ϵ,1)),γ2([1ϵ,1))int(𝒯)\gamma_{1}\big{(}[1-\epsilon,1)\big{)},\gamma_{2}\big{(}[1-\epsilon,1)\big{)}\subseteq\text{int}(\mathcal{T}) for each 0<ϵϵ00<\epsilon\leq\epsilon_{0}. Note that int(𝒯)\text{int}(\mathcal{T}) is geodesically convex. Set

1(ϵ)=length of γ1|[1ϵ,1],2(ϵ)=length of γ2|[1ϵ,1],12(ϵ)=length of the geodesic segment joining γ1(1ϵ) and γ2(1ϵ)\begin{gathered}\ell_{1}(\epsilon)=\text{length of $\gamma_{1}\bigl{\lvert}_{[1-\epsilon,1]}$},\;\ell_{2}(\epsilon)=\text{length of $\gamma_{2}\bigl{\lvert}_{[1-\epsilon,1]}$},\\ \ell_{12}(\epsilon)=\text{length of the geodesic segment joining $\gamma_{1}(1-\epsilon)$ and $\gamma_{2}(1-\epsilon)$}\end{gathered}

where length is measured with respect to the underlying metric on \mathbb{C}. Suppose that |f(τ)|M\lvert f(\tau)\rvert\leq M for all τint(𝒯)\tau\in\text{int}(\mathcal{T}). Then

|γ1f(ξ)dξγ2f(ξ)dξ|M(1(ϵ)+2(ϵ)+12(ϵ)).(0<ϵϵ0)\bigl{\lvert}\int_{\gamma_{1}}f(\xi)d\xi-\int_{\gamma_{2}}f(\xi)d\xi\bigl{\rvert}\leq M\big{(}\ell_{1}(\epsilon)+\ell_{2}(\epsilon)+\ell_{12}(\epsilon)\big{)}.\hskip 8.5359pt(\begin{subarray}{c}0<\epsilon\leq\epsilon_{0}\end{subarray})

We let ϵ0\epsilon\to 0 and conclude that the above inequality’s LHS is zero. ∎

Corollary 3.7.

Let x1,x21(){}x_{1},x_{2}\in\mathbb{P}^{1}(\mathbb{Q})-\{\infty\}. Suppose that f𝔉0x1𝔉0x2f\in\mathfrak{F}^{x_{1}}_{0}\cap\mathfrak{F}^{x_{2}}_{0}. Then

x1x2f(ξ)𝑑ξ\int^{x_{2}}_{x_{1}}f(\xi)d\xi

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 f𝔉2f\in\mathfrak{F}_{2}. Then τf(ξ)𝑑ξ\int_{\tau}^{\infty}f(\xi)d\xi converges absolutely to define a holomorphic function on \mathbb{H} and

ddττf(ξ)𝑑ξ=f(τ).\frac{d}{d\tau}\int_{\tau}^{\infty}f(\xi)d\xi=-f(\tau).
Proof.

The hypothesis on ff amounts to the fact that |f(ξ)|Cv(ξ)2\lvert f(\xi)\rvert\leq\frac{C}{v(\xi)^{2}} on every vertical strip bounded below. Now the assertion is an easy exercise in standard techniques of complex analysis. ∎

Let k2k\geq 2 and α[k,]\alpha\in[k,\infty]. Suppose f𝖤𝖯𝔉αf\in\mathsf{EP}\mathfrak{F}_{\alpha}. Write f=f+fcf=f_{\infty}+f_{c}. Set

𝕀k(τ,f,X):=0τf(ξ)(Xξ)k2𝑑ξHol()[X]k2.\mathbb{I}^{\infty}_{k}(\tau,f,X):=\int^{\tau}_{0}f_{\infty}(\xi)(X-\xi)^{k-2}d\xi\in\text{Hol}(\mathbb{H})[X]_{k-2}.

It is clear that τ(𝕀k(τ,f,X))=f(τ)(Xτ)k2dτ\partial_{\tau}\big{(}\mathbb{I}^{\infty}_{k}(\tau,f,X)\big{)}=f_{\infty}(\tau)(X-\tau)^{k-2}d\tau. Also, we have f(τ)τr𝔉2f(\tau)\tau^{r}\in\mathfrak{F}_{2} for each 0rk20\leq r\leq k-2. One employs Lemma 3.8 to conclude that

𝕀kc(τ,f,X):=τfc(ξ)(Xξ)k2𝑑ξHol()[X]k2\mathbb{I}^{c}_{k}(\tau,f,X):=-\int^{\infty}_{\tau}f_{c}(\xi)(X-\xi)^{k-2}d\xi\in\text{Hol}(\mathbb{H})[X]_{k-2}

and satisfies τ(𝕀kc(τ,f,X))=fc(τ)(Xτ)k2dτ\partial_{\tau}\big{(}\mathbb{I}^{c}_{k}(\tau,f,X)\big{)}=f_{c}(\tau)(X-\tau)^{k-2}d\tau. Put

(3.2) 𝕀k(τ,f,X)=𝕀k(τ,f,X)+𝕀kc(τ,f,X)Hol()[X]k2.\mathbb{I}^{*}_{k}(\tau,f,X)=\mathbb{I}^{\infty}_{k}(\tau,f,X)+\mathbb{I}^{c}_{k}(\tau,f,X)\in\text{Hol}(\mathbb{H})[X]_{k-2}.

The differential relations for 𝕀k\mathbb{I}^{\infty}_{k} and 𝕀kc\mathbb{I}^{c}_{k} together imply that 𝕀k(τ,f,X)\mathbb{I}^{*}_{k}(\tau,f,X) is a weight kk Eichler-Shimura integral of ff. We refer to 𝕀k(τ,f,X)\mathbb{I}^{*}_{k}(\tau,f,X) as the Eichler-Shimura integral of ff with base point at \infty.

Now suppose 𝒢\mathcal{G} is a finite index subgroup of Γ\Gamma and 𝒳𝖤𝖯𝔉α\mathcal{X}\subseteq\mathsf{EP}\mathfrak{F}_{\alpha} is stable under the |k\lvert_{k} action of 𝒢\mathcal{G}. Let

𝕀k,𝒳:𝒳Hol()[X]k2\mathbb{I}_{k,\mathcal{X}}^{*}:\mathcal{X}\to\text{Hol}(\mathbb{H})[X]_{k-2}

be the choice function attached to (3.2) and 𝗉k,𝒳\mathsf{p}^{*}_{k,\mathcal{X}} is the period function of 𝕀k,𝒳\mathbb{I}^{*}_{k,\mathcal{X}}. The Eichler-Shimura integrals with base point at \infty behave like a canonical choice to some extent. In particular 𝕀k,𝒳\mathbb{I}_{k,\mathcal{X}}^{*} is linear whenever 𝒳\mathcal{X} is a linear subspace of Hol()\text{Hol}(\mathbb{H}). Next, we provide an explicit formula for the period function. For convenience write

f(γ):=π(f|γ),fc(γ):=πc(f|γ).(f𝒳,γ𝒢)f_{\infty}(\gamma):=\pi_{\infty}(f\lvert_{\gamma}),\hskip 5.69046ptf_{c}(\gamma):=\pi_{c}(f\lvert_{\gamma}).\hskip 8.5359pt(\begin{subarray}{c}f\in\mathcal{X},\gamma\in\mathcal{G}\end{subarray})
Proposition 3.9.

Let f𝒳f\in\mathcal{X} and γ𝒢\gamma\in\mathcal{G}. Then

(3.3) 𝗉k,𝒳(γ,f,X)\displaystyle\mathsf{p}_{k,\mathcal{X}}^{*}(\gamma,f,X)
=γ10τ0f(γ1)|γ(ξ)(Xξ)k2dξ0τ0f(ξ)(Xξ)k2dξ\displaystyle=\int_{\gamma^{-1}0}^{\tau_{0}}f_{\infty}(\gamma^{-1})\lvert_{\gamma}(\xi)(X-\xi)^{k-2}d\xi-\int_{0}^{\tau_{0}}f_{\infty}(\xi)(X-\xi)^{k-2}d\xi
τ0γ1fc(γ1)|γ(ξ)(Xξ)k2dξ+τ0fc(ξ)(Xξ)k2dξ\displaystyle-\int_{\tau_{0}}^{\gamma^{-1}\infty}f_{c}(\gamma^{-1})\lvert_{\gamma}(\xi)(X-\xi)^{k-2}d\xi+\int_{\tau_{0}}^{\infty}f_{c}(\xi)(X-\xi)^{k-2}d\xi

for each τ0\tau_{0}\in\mathbb{H}.

Proof.

Let τ0\tau_{0}\in\mathbb{H}. One uses (3.2) with τ=τ0\tau=\tau_{0} to conclude that

(3.4) 𝗉k,𝒳(γ,f,X)\displaystyle\mathsf{p}_{k,\mathcal{X}}^{*}(\gamma,f,X)
=(cX+d)k20γτ0f(γ1)(ξ)(aX+bcX+dξ)k2𝑑ξ𝕀k(τ0,f,X)\displaystyle=(cX+d)^{k-2}\int^{\gamma\tau_{0}}_{0}f_{\infty}(\gamma^{-1})(\xi)(\frac{aX+b}{cX+d}-\xi)^{k-2}d\xi-\mathbb{I}^{\infty}_{k}(\tau_{0},f,X)
(cX+d)k2γτ0fc(γ1)(ξ)(aX+bcX+dξ)k2𝑑ξ𝕀kc(τ0,f,X).\displaystyle-(cX+d)^{k-2}\int_{\gamma\tau_{0}}^{\infty}f_{c}(\gamma^{-1})(\xi)(\frac{aX+b}{cX+d}-\xi)^{k-2}d\xi-\mathbb{I}^{c}_{k}(\tau_{0},f,X).

To deduce (3.3) from (3.4) one must show that the change of variable ξγξ\xi\to\gamma\xi 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 0 and γτ0\gamma\tau_{0} which ensures that the transformed path is a good path between the endpoints whenever γ10=\gamma^{-1}0=\infty. 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 γ𝒢\gamma\in\mathcal{G}_{\infty} and f𝒳f\in\mathcal{X}. Then

(3.5) 𝗉k,𝒳(γ,f,X)=γ100f(ξ)(Xξ)k2𝑑ξ.\mathsf{p}_{k,\mathcal{X}}^{*}(\gamma,f,X)=\int_{\gamma^{-1}0}^{0}f_{\infty}(\xi)(X-\xi)^{k-2}d\xi.
Proof.

We have γ1=\gamma^{-1}\infty=\infty. Moreover a0a\neq 0 and γ10=ba\gamma^{-1}0=-\frac{b}{a}\in\mathbb{Q}. Lemma 3.5(i) implies that f(γ1)|γ=ff_{\infty}(\gamma^{-1})\lvert_{\gamma}=f_{\infty} and fc(γ1)|γ=fcf_{c}(\gamma^{-1})\lvert_{\gamma}=f_{c}. Therefore the assertion is a direct consequence of (3.3). ∎

3.3 Connections with the theory of modular forms

Let k2k\geq 2 be an integer and 𝒢\mathcal{G} be a finite index subgroup Γ\Gamma. Suppose that k!(𝒢)\mathcal{M}_{k}^{!}(\mathcal{G}), resp. k(𝒢)\mathcal{M}_{k}(\mathcal{G}), is the space of weight kk weakly holomorphic, resp. holomorphic, modular forms on 𝒢\mathcal{G}. We choose a positive integer NN so that TN=(1N01)𝒢T^{N}=\begin{pmatrix}1&N\\ 0&1\end{pmatrix}\in\mathcal{G}. A weakly holomorphic modular form fk!(𝒢)f\in\mathcal{M}_{k}^{!}(\mathcal{G}) admits a qq-series expansion

(3.6) f(τ)=n=a(n)qNn,qN=e(τ/N),f(\tau)=\sum_{n=-\infty}^{\infty}a(n)q_{N}^{n},\hskip 14.22636ptq_{N}=\textbf{e}(\tau/N),

where (a(n))n\big{(}a(n)\big{)}_{n\in\mathbb{Z}} is a sequence of complex numbers so that a(n)=0a(n)=0 for n<<0n<<0 and n1a(n)zn\sum_{n\geq 1}a(n)z^{n} has radius of convergence 11 on the open unit disc. If fk(𝒢)f\in\mathcal{M}_{k}(\mathcal{G}) then a(n)=0a(n)=0 for each n<0n<0.

Lemma 3.11.

k!(𝒢)𝖤𝖯𝔉\mathcal{M}_{k}^{!}(\mathcal{G})\subseteq\mathsf{EP}\mathfrak{F}_{\infty}.

Proof.

Let fk!(𝒢)f\in\mathcal{M}_{k}^{!}(\mathcal{G}) and consider the Fourier expansion (3.6). The condition on the radius of convergence yields positive constants C,MC,M with M>1M>1 so that |a(n)|CMn\lvert a(n)\rvert\leq CM^{n} for each n1n\geq 1. Set h0=NlogM2πh_{0}=\frac{N\log M}{2\pi} and let t1,t2,ht_{1},t_{2},h\in\mathbb{R} with t1<t2t_{1}<t_{2} and h>h0h>h_{0}. Then

n1|a(n)qNn|CM1Mexp2πhNexp(2πv(τ)N).(τ𝒩t1,t2h)\sum_{n\geq 1}\lvert a(n)q_{N}^{n}\rvert\leq\frac{CM}{1-\frac{M}{\exp{\frac{2\pi h}{N}}}}\exp{(-\frac{2\pi v(\tau)}{N})}.\hskip 14.22636pt(\begin{subarray}{c}\tau\in\mathcal{N}_{t_{1},t_{2}}^{h}\end{subarray})

Thus |n1a(n)qNn|v(τ)α\lvert\sum_{n\geq 1}a(n)q_{N}^{n}\rvert v(\tau)^{\alpha} is bounded on 𝒩t1,t2h\mathcal{N}_{t_{1},t_{2}}^{h} for each α>0\alpha>0. This calculation shows that n1a(n)qNn𝔉\sum_{n\geq 1}a(n)q_{N}^{n}\in\mathfrak{F}_{\infty} (Lemma 3.3). Therefore f𝖤𝖯𝔉f\in\mathsf{EP}\mathfrak{F}_{\infty}. ∎

Observe that the space of constants \mathbb{C} is contained in 𝖤𝖯\mathsf{EP}. Set

𝔉:=+𝔉𝖤𝖯𝔉.\mathbb{C}\mathfrak{F}_{\infty}:=\mathbb{C}+\mathfrak{F}_{\infty}\subseteq\mathsf{EP}\mathfrak{F}_{\infty}.

The following result is an immediate consequence of the lemma above.

Corollary 3.12.

k(𝒢)𝔉\mathcal{M}_{k}(\mathcal{G})\subseteq\mathbb{C}\mathfrak{F}_{\infty}.

The formula (3.5) simplifies for the class of functions 𝔉\mathbb{C}\mathfrak{F}_{\infty}. Let the notation be as in Corollary 3.10. Moreover assume that 𝒳𝔉\mathcal{X}\subseteq\mathbb{C}\mathfrak{F}_{\infty}. Write γ=ϵTr\gamma=\epsilon T^{r} where ϵ{±1}\epsilon\in\{\pm 1\} and rr\in\mathbb{Z}.

Lemma 3.13.

𝗉k,𝒳(γ,f,X)=π(f)(X+r)k1Xk1k1\mathsf{p}_{k,\mathcal{X}}^{*}(\gamma,f,X)=\pi_{\infty}(f)\frac{(X+r)^{k-1}-X^{k-1}}{k-1}.

Proof.

Note that γ10=r\gamma^{-1}0=-r. Moreover f(ξ)f_{\infty}(\xi) 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 γ=S\gamma=S even if ff is not SS-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 𝒢\mathcal{G} be an abstract group and 𝒳\mathcal{X} be a trivial right 𝒢\mathcal{G}-set. Suppose that 𝕂\mathbb{K} is a field and VV is a right 𝕂[𝒢]\mathbb{K}[\mathcal{G}]-module. There is a canonical 𝕂\mathbb{K}-linear isomorphism

(4.1) Φ:Fun(𝒢,Fun(𝒳,V))Fun(𝒳,Fun(𝒢,V));\displaystyle\Phi:\text{Fun}\big{(}\mathcal{G},\text{Fun}(\mathcal{X},V)\big{)}\xrightarrow{\cong}\text{Fun}\big{(}\mathcal{X},\text{Fun}(\mathcal{G},V)\big{)};
FΦ(F),(Φ(F)(f))(γ):=F(γ)(f).(f𝒳,γ𝒢)\displaystyle F\to\Phi(F),\hskip 8.5359pt\big{(}\Phi(F)(f)\big{)}(\gamma):=F(\gamma)(f).\hskip 14.22636pt(\begin{subarray}{c}f\in\mathcal{X},\gamma\in\mathcal{G}\end{subarray})

Since 𝒢\mathcal{G} acts trivially on 𝒳\mathcal{X} the map Φ\Phi gives rise to 𝕂\mathbb{K}-linear isomorphisms at the level of cocyles and cohomology:

Φcy:Z1(𝒢,Fun(𝒳,V))Fun(𝒳,Z1(𝒢,V)),\displaystyle\Phi_{\text{cy}}:Z^{1}\big{(}\mathcal{G},\text{Fun}(\mathcal{X},V)\big{)}\xrightarrow{\cong}\text{Fun}\big{(}\mathcal{X},Z^{1}(\mathcal{G},V)\big{)},
Φ¯:H1(𝒢,Fun(𝒳,V))Fun(𝒳,H1(𝒢,V)).\displaystyle\overline{\Phi}:H^{1}\big{(}\mathcal{G},\text{Fun}(\mathcal{X},V)\big{)}\xrightarrow{\cong}\text{Fun}\big{(}\mathcal{X},H^{1}(\mathcal{G},V)\big{)}.

Let the notation be as in Section 2.2 and assume that 𝒳Hol()𝒢\mathcal{X}\subseteq\text{Hol}(\mathbb{H})^{\mathcal{G}}. Now suppose 𝕀k,𝒳\mathbb{I}_{k,\mathcal{X}} is a choice function for 𝒳\mathcal{X}. With 𝕀k,𝒳\mathbb{I}_{k,\mathcal{X}} one associates an Eichler-Shimura cocycle map given by

(4.2) ESk,𝒳:𝒳Z1(𝒢,Vk2,),ESk,𝒳:=Φcy(𝗉k,𝒳)\text{ES}_{k,\mathcal{X}}:\mathcal{X}\to Z^{1}\big{(}\mathcal{G},V_{k-2,\mathbb{C}}\big{)},\hskip 8.5359pt\text{ES}_{k,\mathcal{X}}:=\Phi_{\text{cy}}(\mathsf{p}_{k,\mathcal{X}})

where 𝗉k,𝒳\mathsf{p}_{k,\mathcal{X}} is the period function of 𝕀k,𝒳\mathbb{I}_{k,\mathcal{X}}. This cocycle map naturally descends to a map into the cohomology module.

Definition 4.1.

The Eichler-Shimura homomorphism attached to (𝒢,𝒳)(\mathcal{G},\mathcal{X}) is the image of the Eichler-Shimura class [𝗉k]𝒢,𝒳[\mathsf{p}_{k}]_{\mathcal{G},\mathcal{X}} under Φ¯\overline{\Phi}, i.e.,

[ESk]𝒢,𝒳:𝒳H1(𝒢,Vk2,),[ESk]𝒢,𝒳:=Φ¯([𝗉k]𝒢,𝒳).[\text{ES}_{k}]_{\mathcal{G},\mathcal{X}}:\mathcal{X}\to H^{1}\big{(}\mathcal{G},V_{k-2,\mathbb{C}}\big{)},\hskip 8.5359pt[\text{ES}_{k}]_{\mathcal{G},\mathcal{X}}:=\overline{\Phi}([\mathsf{p}_{k}]_{\mathcal{G},\mathcal{X}}).

Proposition 2.7 ensures that [ESk]𝒢,𝒳[\text{ES}_{k}]_{\mathcal{G},\mathcal{X}} is determined by (𝒢,𝒳)(\mathcal{G},\mathcal{X}) and does not depend on the choice of the Eichler-Shimura cocycle map. If (𝒢,𝒳)(\mathcal{G},\mathcal{X}) is clear from context then we drop 𝒢\mathcal{G} and 𝒳\mathcal{X} from the notation. If 𝒳\mathcal{X} consists of functions lying in suitable growth classes, say 𝒳𝔉\mathcal{X}\subseteq\mathbb{C}\mathfrak{F}_{\infty}, then the cocycle function (4.2) determined by base point at \infty 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 𝕀k,𝒳τ0\mathbb{I}_{k,\mathcal{X}}^{\tau_{0}}.

The following result records the basic properties of the Eichler-Shimura map.

Proposition 4.2.

Let 𝒳Hol()\mathcal{X}\subseteq\emph{Hol}(\mathbb{H}) be a collection of 𝒢\mathcal{G}-invariant functions.

  1. (i)

    The class [𝗉k]𝒢,𝒳[\mathsf{p}_{k}]_{\mathcal{G},\mathcal{X}} vanishes if and only if [ESk][\emph{ES}_{k}] is the zero map.

  2. (ii)

    Let 𝒳\mathcal{X} be a 𝕂\mathbb{K}-subspace of Hol()\emph{Hol}(\mathbb{H}) for some subfield 𝕂\mathbb{K} of \mathbb{C}. Then [ESk][\emph{ES}_{k}] is automatically 𝕂\mathbb{K}-linear.

Proof.

Part (i) follows from the observation that Φ¯\overline{\Phi} is an isomorphism. To prove (ii) let 𝕀k,𝒳τ0\mathbb{I}_{k,\mathcal{X}}^{\tau_{0}} be a canonical choice function for 𝒳\mathcal{X}. We know that 𝕀k,𝒳τ0\mathbb{I}_{k,\mathcal{X}}^{\tau_{0}} is a 𝕂\mathbb{K}-linear choice function, i.e.,

Im(𝗉k,𝒳τ0)Hom𝕂(𝒳,Vk2,)Fun(𝒳,Vk2,).\text{Im}\big{(}\mathsf{p}^{\tau_{0}}_{k,\mathcal{X}}\big{)}\subseteq\text{Hom}_{\mathbb{K}}(\mathcal{X},V_{k-2,\mathbb{C}})\subseteq\text{Fun}(\mathcal{X},V_{k-2,\mathbb{C}}).

One now uses the definition of Φ\Phi in (4.1) to discover that the cocycle map attached to 𝕀k,𝒳τ0\mathbb{I}_{k,\mathcal{X}}^{\tau_{0}} is also 𝕂\mathbb{K}-linear. But ESk,𝒳\text{ES}_{k,\mathcal{X}} descends to the map [ESk][\text{ES}_{k}]. Therefore [ESk][\text{ES}_{k}] is 𝕂\mathbb{K}-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 GL2+()\text{GL}_{2}^{+}(\mathbb{Q}) on function spaces and module of coefficients (Section 2.3.1). Let αGL2+()\alpha\in\text{GL}_{2}^{+}(\mathbb{Q}) and 𝒢1\mathcal{G}_{1}, 𝒢2\mathcal{G}_{2} be two finite index subgroups of Γ\Gamma. Suppose that {α1,,αn}\{\alpha_{1},\ldots,\alpha_{n}\} is a set of representatives for the orbits of the left 𝒢1\mathcal{G}_{1}-action on 𝒢1α𝒢2\mathcal{G}_{1}\alpha\mathcal{G}_{2}. For each kk\in\mathbb{Z} we have a \mathbb{C}-linear double coset operator on functions given by

[𝒢1α𝒢2]k:Hol()𝒢1Hol()𝒢2;f|[𝒢1α𝒢2]k:=j=1nf|k,αj.[\mathcal{G}_{1}\alpha\mathcal{G}_{2}]_{k}:\text{Hol}(\mathbb{H})^{\mathcal{G}_{1}}\to\text{Hol}(\mathbb{H})^{\mathcal{G}_{2}};\hskip 8.5359ptf\lvert_{[\mathcal{G}_{1}\alpha\mathcal{G}_{2}]_{k}}:=\sum_{j=1}^{n}f|_{k,\alpha_{j}}.

The operator [𝒢1α𝒢2]k[\mathcal{G}_{1}\alpha\mathcal{G}_{2}]_{k} is independent of the choice of α\alpha-s. There is also a notion of double coset operators for group cohomology. Let 𝒢\mathcal{G} denote the subgroup of GL2+()\text{GL}_{2}^{+}(\mathbb{Q}) generated by 𝒢1\mathcal{G}_{1}, 𝒢2\mathcal{G}_{2}, and α\alpha. Suppose that VV is a right 𝕂[𝒢]\mathbb{K}[\mathcal{G}]-module where 𝕂\mathbb{K} is a field of characteristic 0. Given γ𝒢2\gamma\in\mathcal{G}_{2} and 1jn1\leq j\leq n we obtain unique γj𝒢1\gamma_{j}\in\mathcal{G}_{1} and j(γ){1,,n}j(\gamma)\in\{1,\ldots,n\} so that αjγ=γjαj(γ)\alpha_{j}\gamma=\gamma_{j}\alpha_{j(\gamma)}. Define

(4.3) :VFun(𝒢1,V)Fun(𝒢2,V),P|[𝒢1α𝒢2]V(γ):=j=1nP(γj)αj(γ).\begin{gathered}{}_{V}:\text{Fun}(\mathcal{G}_{1},V)\to\text{Fun}(\mathcal{G}_{2},V),\\ P\bigl{\lvert}_{[\mathcal{G}_{1}\alpha\mathcal{G}_{2}]_{V}}(\gamma):=\sum_{j=1}^{n}P(\gamma_{j})\alpha_{j(\gamma)}.\end{gathered}

The operator [𝒢1α𝒢2]V[\mathcal{G}_{1}\alpha\mathcal{G}_{2}]_{V} sends cocycles to cocycles and descends to a 𝕂\mathbb{K}-linear map in cohomology. We denote the map in cohomology by the same notation and often drop VV from the subscript. The map at the level of cohomology does not depend on the choice of representatives for 𝒢1\mathcal{G}_{1}-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 𝒳1Hol()𝒢1\mathcal{X}_{1}\subseteq\emph{Hol}(\mathbb{H})^{\mathcal{G}_{1}} and 𝒳2Hol()𝒢2\mathcal{X}_{2}\subseteq\emph{Hol}(\mathbb{H})^{\mathcal{G}_{2}} be two subsets of functions so that [𝒢1α𝒢2]k(𝒳1)𝒳2[\mathcal{G}_{1}\alpha\mathcal{G}_{2}]_{k}(\mathcal{X}_{1})\subseteq\mathcal{X}_{2}. Then the Eichler-Shimura maps commute with the action of double coset operators, i.e., [𝒢1α𝒢2][ESk]𝒢1,𝒳1=[ESk]𝒢2,𝒳2[𝒢1α𝒢2]k[\mathcal{G}_{1}\alpha\mathcal{G}_{2}]\circ[\emph{ES}_{k}]_{\mathcal{G}_{1},\mathcal{X}_{1}}=[\emph{ES}_{k}]_{\mathcal{G}_{2},\mathcal{X}_{2}}\circ[\mathcal{G}_{1}\alpha\mathcal{G}_{2}]_{k}.

Proof.

Let 𝒢\mathcal{G} be the subgroup of GL2+()\text{GL}_{2}^{+}(\mathbb{Q}) generated by 𝒢1𝒢2{α}\mathcal{G}_{1}\cup\mathcal{G}_{2}\cup\{\alpha\}. Suppose that 𝒳\mathcal{X} is a 𝒢\mathcal{G}-stable subset of Hol()\text{Hol}(\mathbb{H}) containing 𝒳1𝒳2\mathcal{X}_{1}\cup\mathcal{X}_{2}. We fix a choice function 𝕀k,𝒳\mathbb{I}_{k,\mathcal{X}} and consider the period function 𝗉k,𝒳\mathsf{p}_{k,\mathcal{X}}. Restriction yields choice functions for 𝒳1\mathcal{X}_{1} and 𝒳2\mathcal{X}_{2}. Let f𝒳1f\in\mathcal{X}_{1} and γ𝒢2\gamma\in\mathcal{G}_{2}. Then

𝗉k,𝒳(γj,f,X)|αj(γ)=𝗉k,𝒳(γjαj(γ),f|αj(γ),X)𝗉k,𝒳(αj(γ),f|αj(γ),X)(1jn)\displaystyle\mathsf{p}_{k,\mathcal{X}}(\gamma_{j},f,X)\lvert_{\alpha_{j(\gamma)}}=\mathsf{p}_{k,\mathcal{X}}(\gamma_{j}\alpha_{j(\gamma)},f\lvert_{\alpha_{j(\gamma)}},X)-\mathsf{p}_{k,\mathcal{X}}(\alpha_{j(\gamma)},f\lvert_{\alpha_{j(\gamma)}},X)\hskip 5.69046pt(\begin{subarray}{c}1\leq j\leq n\end{subarray})
=𝗉k,𝒳(γ,f|αj(γ),X)+𝗉k,𝒳(αj,f|αj,X)|γ𝗉k,𝒳(αj(γ),f|αj(γ),X).\displaystyle=\mathsf{p}_{k,\mathcal{X}}(\gamma,f\lvert_{\alpha_{j(\gamma)}},X)+\mathsf{p}_{k,\mathcal{X}}(\alpha_{j},f\lvert_{\alpha_{j}},X)\bigl{\lvert}_{\gamma}-\mathsf{p}_{k,\mathcal{X}}(\alpha_{j(\gamma)},f\lvert_{\alpha_{j(\gamma)}},X).

Summing over jj one finds that

[𝒢1α𝒢2]ESk,𝒳1(f)ESk,𝒳2[𝒢1α𝒢2]k(f)[\mathcal{G}_{1}\alpha\mathcal{G}_{2}]\circ\text{ES}_{k,\mathcal{X}_{1}}(f)-\text{ES}_{k,\mathcal{X}_{2}}\circ[\mathcal{G}_{1}\alpha\mathcal{G}_{2}]_{k}(f)

equals a coboundary on 𝒢2\mathcal{G}_{2}. ∎

4.3 Equivariance under group action

We begin with a recap of the conjugation action on the group cohomology [6, III.8]. Let \mathcal{H} be an abstract group and 𝒢\mathcal{G} be a subgroup of \mathcal{H}. Suppose that 𝕂\mathbb{K} is a field of characteristic 0 and VV is a right 𝕂[]\mathbb{K}[\mathcal{H}]-module. For each hh\in\mathcal{H} conjugation gives rise to an isomorphism of pairs

c(h):(𝒢,V)(h1𝒢h,V);(gh1gh,mmh)c(h):(\mathcal{G},V)\to(h^{-1}\mathcal{G}h,V);\hskip 8.5359pt(g\mapsto h^{-1}gh,m\mapsto mh)

that yields a pullback isomorphism c(h):H(h1𝒢h,V)H(𝒢,V)c(h)^{*}:H^{*}(h^{-1}\mathcal{G}h,V)\to H^{*}(\mathcal{G},V). For αH(𝒢,V)\alpha\in H^{*}(\mathcal{G},V) set

(4.4) α|h:=(c(h))1(α)H(h1𝒢h,V).\alpha\lvert_{h}:=\big{(}c(h)^{*}\big{)}^{-1}(\alpha)\in H^{*}(h^{-1}\mathcal{G}h,V).

Suppose that 𝒢\mathcal{G} is a normal subgroup of \mathcal{H}. Then (4.4) defines a right 𝕂[]\mathbb{K}[\mathcal{H}]-module structure on H(𝒢,V)H^{*}(\mathcal{G},V). The subgroup 𝒢\mathcal{G} acts trivially on H(𝒢,V)H^{*}(\mathcal{G},V) and the \mathcal{H}-action descends to an action of /𝒢\mathcal{H}/\mathcal{G}.

Lemma 4.4.

Let the notation be as above. Suppose that 𝒢\mathcal{G} is a finite index normal subgroup of \mathcal{H}. Then the restriction homomorphism

res𝒢:H(,V)H(𝒢,V)\emph{res}^{\mathcal{H}}_{\mathcal{G}}:H^{*}(\mathcal{H},V)\to H^{*}(\mathcal{G},V)

maps H(,V)H^{*}(\mathcal{H},V) isomorphically onto H(𝒢,V)H^{*}(\mathcal{G},V)^{\mathcal{H}}.

Proof.

See Proposition 10.4 in [6, III.10]. ∎

We next explicate the action of \mathcal{H} at the level of 11-cocycles; cf. [6, p.79]. Assume that 𝒢\mathcal{G} is a normal subgroup of \mathcal{H}. The (right) conjugation action of \mathcal{H} on 𝒢\mathcal{G} turns Fun(𝒢,V)\text{Fun}(\mathcal{G},V) into a 𝕂[]\mathbb{K}[\mathcal{H}]-module as follows:

(4.5) (P,h)P|h;P|h(g)=P(hgh1)h.(P,h)\mapsto P\lvert_{h};\hskip 8.5359ptP\lvert_{h}(g)=P(hgh^{-1})h.

Note that Z1(𝒢,V)Z^{1}(\mathcal{G},V) and B1(𝒢,V)B^{1}(\mathcal{G},V) are 𝕂[]\mathbb{K}[\mathcal{H}]-submodules of Fun(𝒢,V)\text{Fun}(\mathcal{G},V). Thus H1(𝒢,V)H^{1}(\mathcal{G},V) inherits a \mathcal{H}-action that coincides with (4.4).

Example 4.5.

Let 𝒢\mathcal{G} be a subgroup of Γ\Gamma containing Id-\text{Id} and assume that Id-\text{Id} acts on VV as IdV-\text{Id}_{V}. Then, by (4.5), the conjugation action of Id-\text{Id} on H1(𝒢,V)H^{1}(\mathcal{G},V) equals IdH1(𝒢,V)-\text{Id}_{H^{1}(\mathcal{G},V)}. But each element of 𝒢\mathcal{G} must act trivially on the cohomology. It follows that H1(𝒢,V)H^{1}(\mathcal{G},V) must be zero. In particular, if kk is odd and Id𝒢-\text{Id}\in\mathcal{G} then H1(𝒢,Vk2,𝕂)H^{1}(\mathcal{G},V_{k-2,\mathbb{K}}) is zero.

Remark 4.6.

Let the notation be as in Lemma 4.4 but we drop the assumption that 𝒢\mathcal{G} is normal in \mathcal{H}. By general theory

cores𝒢res𝒢=multiplication by [:𝒢]\text{cores}^{\mathcal{H}}_{\mathcal{G}}\circ\text{res}^{\mathcal{H}}_{\mathcal{G}}=\text{multiplication by $[\mathcal{H}:\mathcal{G}]$}

where res𝒢\text{res}^{\mathcal{H}}_{\mathcal{G}} is the restriction map and cores𝒢:H(𝒢,V)H(,V)\text{cores}^{\mathcal{H}}_{\mathcal{G}}:H^{*}(\mathcal{G},V)\to H^{*}(\mathcal{H},V) is the corestriction map [6, III.9]. Therefore the restriction map is injective.

We put ourselves in the setting of Section 4.1 and suppose that \mathcal{H} is a subgroup of Γ\Gamma that contains 𝒢\mathcal{G} as a normal subgroup. Moreover assume that 𝒳\mathcal{X} is stable under the |k\lvert_{k}-action of \mathcal{H}. The discussion of this subsection yields a canonical right []\mathbb{C}[\mathcal{H}]-module structure on H1(𝒢,Vk2,)H^{1}(\mathcal{G},V_{k-2,\mathbb{C}}).

Proposition 4.7.

The Eichler-Shimura homomorphism

[ESk]:𝒳H1(𝒢,Vk2,)[\emph{ES}_{k}]:\mathcal{X}\to H^{1}(\mathcal{G},V_{k-2,\mathbb{C}})

is a morphism of \mathcal{H}-sets.

Proof.

We fix a choice function for 𝒳\mathcal{X} and consider the period function and the cocycle map attached to that choice. Let f𝒳f\in\mathcal{X} and γ0\gamma_{0}\in\mathcal{H}. Then

ESk,𝒳(f)|γ0(γ)\displaystyle\text{ES}_{k,\mathcal{X}}(f)\lvert_{\gamma_{0}}(\gamma) =𝗉k,𝒳(γ0γγ01,f,X)|γ0(γ𝒢)\displaystyle=\mathsf{p}_{k,\mathcal{X}}(\gamma_{0}\gamma\gamma_{0}^{-1},f,X)\lvert_{\gamma_{0}}\hskip 14.22636pt(\begin{subarray}{c}\gamma\in\mathcal{G}\end{subarray})
=𝗉k,𝒳(γ0,f|γ0γ1,X)|γ+𝗉k,𝒳(γ,f|γ0,X)+𝗉k,𝒳(γ01,f,X)|γ0\displaystyle=\mathsf{p}_{k,\mathcal{X}}(\gamma_{0},f\lvert_{\gamma_{0}\gamma^{-1}},X)\lvert_{\gamma}+\mathsf{p}_{k,\mathcal{X}}(\gamma,f\lvert_{\gamma_{0}},X)+\mathsf{p}_{k,\mathcal{X}}(\gamma_{0}^{-1},f,X)\lvert_{\gamma_{0}}
=ESk,𝒳(f|γ0)(γ)+𝗉k,𝒳(γ0,f|γ0,X)|γ𝗉k,𝒳(γ0,f|γ0,X).\displaystyle=\text{ES}_{k,\mathcal{X}}(f\lvert_{\gamma_{0}})(\gamma)+\mathsf{p}_{k,\mathcal{X}}(\gamma_{0},f\lvert_{\gamma_{0}},X)\lvert_{\gamma}-\mathsf{p}_{k,\mathcal{X}}(\gamma_{0},f\lvert_{\gamma_{0}},X).

Therefore [ESk](f)|γ0=[ESk](f|γ0)[\text{ES}_{k}](f)\lvert_{\gamma_{0}}=[\text{ES}_{k}](f\lvert_{\gamma_{0}}) in H1(𝒢,Vk2,)H^{1}(\mathcal{G},V_{k-2,\mathbb{C}}) as desired. ∎

Remark 4.8.

With the notation above let 𝒢1\mathcal{G}_{1} be a subgroup of \mathcal{H} containing 𝒢\mathcal{G}. Set 𝒳1:=𝒳𝒢1\mathcal{X}_{1}:=\mathcal{X}^{\mathcal{G}_{1}}. Then using compatible choice functions for 𝒳\mathcal{X} and 𝒳1\mathcal{X}_{1} (cf. (2.13) and (2.14)) one discovers that the following diagram commutes:

𝒳1{\mathcal{X}_{1}}𝒳{\mathcal{X}}H1(𝒢1,Vk2,){H^{1}(\mathcal{G}_{1},V_{k-2,\mathbb{C}})}H1(𝒢,Vk2,).{H^{1}(\mathcal{G},V_{k-2,\mathbb{C}}).}[ESk]𝒢1,𝒳1\scriptstyle{{[\text{ES}_{k}]}_{\mathcal{G}_{1},\mathcal{X}_{1}}}[ESk]𝒢,𝒳\scriptstyle{{[\text{ES}_{k}]}_{\mathcal{G},\mathcal{X}}}

Here the top arrow is inclusion and the bottom arrow is the restriction map in group cohomology. Note that the bottom arrow maps H1(𝒢1,Vk2,)H^{1}(\mathcal{G}_{1},V_{k-2,\mathbb{C}}) isomorphically onto the 𝒢1\mathcal{G}_{1}-invariant subspace of H1(𝒢,Vk2,)H^{1}(\mathcal{G},V_{k-2,\mathbb{C}}).

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 PΓ\text{P}\Gamma. As per standard convention, we denote the cusp of a finite index subgroup of Γ\Gamma or PΓ\text{P}\Gamma 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 GG be a finite index subgroup of PΓ\text{P}\Gamma. Suppose that 𝒞(G)\mathcal{C}(G) is the set of cusps attached to GG. For x𝒞(G)x\in\mathcal{C}(G) let GxG_{x} denote the stablizer of xx in GG. Write Gx=πxG_{x}=\langle\pi_{x}\rangle. We also fix a set of representatives {ε1,,εr}\{\varepsilon_{1},\ldots,\varepsilon_{r}\} of the elliptic elements in GG. Let VV be a right 𝕂[G]\mathbb{K}[G]-module where 𝕂\mathbb{K} is a field of characteristic 0. For a subset QQ of GG one defines the cohomology groups vanishing on QQ, denoted HQ(G,V)H^{\ast}_{Q}(G,V), by imposing constraints on 11-cochains [16, p.224]. Suppose that 𝖯\mathsf{P} is the collection of all parabolic elements in GG. We refer to H𝖯(G,V)H^{\ast}_{\mathsf{P}}(G,V) as the parabolic cohomology of VV. However, for practical purposes, it is more convenient to work with Q={πxx𝒞(G)}𝖯Q=\{\pi_{x}\mid x\in\mathcal{C}(G)\}\subseteq\mathsf{P}. The space of 11-cocycles vanishing on QQ and 𝖯\mathsf{P} coincide and there is a canonical identification H𝖯1(G,V)HQ1(G,V)H^{1}_{\mathsf{P}}(G,V)\cong H^{1}_{Q}(G,V). This description of H𝖯1(G,V)H^{1}_{\mathsf{P}}(G,V) provides an exact sequence of 𝕂\mathbb{K}-vector spaces

(5.1) 0H𝖯1(G,V)H1(G,V)x𝒞(G)H1(Gx,V)0\to H^{1}_{\mathsf{P}}(G,V)\to H^{1}(G,V)\to\oplus_{x\in\mathcal{C}(G)}H^{1}(G_{x},V)

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.
  1. Let HGH\subseteq G be two finite index subgroups of PΓ\emph{P}\Gamma and VV be a 𝕂[G]\mathbb{K}[G]-module. Suppose that αH1(G,V)\alpha\in H^{1}(G,V). Then

    αH𝖯1(G,V)\alpha\in H^{1}_{\mathsf{P}}(G,V) if and only if resHG(α)H𝖯1(H,V)\emph{res}^{G}_{H}(\alpha)\in H^{1}_{\mathsf{P}}(H,V).
Proof.

Let x1()x\in\mathbb{P}^{1}(\mathbb{Q}). The natural inclusions give rise to a commutative diagram

H1(G,V){H^{1}(G,V)}H1(H,V){H^{1}(H,V)}H1(Gx,V){H^{1}(G_{x},V)}H1(Hx,V){H^{1}(H_{x},V)}

where all the arrows are restriction maps. By Remark 4.6 the bottom arrow in the diagram above is injective. Therefore the image of α\alpha in H1(Gx,V)H^{1}(G_{x},V) is zero if and only if the image of resHG(α)\text{res}^{G}_{H}(\alpha) in H1(Hx,V)H^{1}(H_{x},V) is zero. Since this assertion holds for each x1()x\in\mathbb{P}^{1}(\mathbb{Q}) the statement follows from (5.1). ∎

We can also compute H(G,V)H^{*}(G,V) in terms of a 22-dimensional simplicial complex 𝔎\mathfrak{K} that contains the fixed points of εj\varepsilon_{j}-s as 0-simplices and admits a simplicial GG-action [16, p.225]. With 𝔎\mathfrak{K} one associates a complex of finitely presented 𝕂[G]\mathbb{K}[G]-modules (A,)(A_{\ast},\partial) so that the complex

A(V):=Hom𝕂[G](A,V)A^{\ast}(V):=\text{Hom}_{\mathbb{K}[G]}(A_{\ast},V)

computes H(G,V)H^{\ast}(G,V). Readers may like to note that Shimura defines AA_{\ast} as a complex of left modules which we endow with the canonical right action xγ:=γ1xx\gamma:=\gamma^{-1}x. Moreover 𝔎\mathfrak{K} also provides an alternate description of HQ(G,V)H^{\ast}_{Q}(G,V). In particular, there is an isomorphism

(5.2) HQ2(G,V)V{v(γ1)vV,γG}.H^{2}_{Q}(G,V)\cong\frac{V}{\{v(\gamma-1)\mid v\in V,\gamma\in G\}}.
Lemma 5.2.

Let 𝕃\mathbb{L} be a field containing 𝕂\mathbb{K}. Then the natural change of scalar map induces isomorphisms

H1(G,V)𝕂𝕃H1(G,V𝕂𝕃),H𝖯1(G,V)𝕂𝕃H𝖯1(G,V𝕂𝕃).H^{1}(G,V)\otimes_{\mathbb{K}}\mathbb{L}\cong H^{1}(G,V\otimes_{\mathbb{K}}\mathbb{L}),\;\;H^{1}_{\mathsf{P}}(G,V)\otimes_{\mathbb{K}}\mathbb{L}\cong H^{1}_{\mathsf{P}}(G,V\otimes_{\mathbb{K}}\mathbb{L}).
Proof.

Follows from the construction above and the sequence (5.1). For details see [10, p. 168]. ∎

Now suppose VV is a finite-dimensional 𝕂\mathbb{K}-vector space. Write

n0=dim𝕂VG,n1=dim𝕂HQ2(G,V),njell=dim𝕂{vVvεj=v},(1jr)nxpar=dim𝕂{v(πx1)vV}.(x𝒞(G))\begin{gathered}n_{0}=\text{dim}_{\mathbb{K}}V^{G},\hskip 8.5359ptn_{1}=\text{dim}_{\mathbb{K}}H^{2}_{Q}(G,V),\\ n_{j}^{\text{ell}}=\text{dim}_{\mathbb{K}}\big{\{}v\in V\mid v\varepsilon_{j}=v\big{\}},\hskip 11.38092pt(\begin{subarray}{c}1\leq j\leq r\end{subarray})\\ n_{x}^{\text{par}}=\text{dim}_{\mathbb{K}}\big{\{}v(\pi_{x}-1)\mid v\in V\big{\}}.\hskip 11.38092pt(\begin{subarray}{c}x\in\mathcal{C}(G)\end{subarray})\end{gathered}
Proposition 5.3.

Let the notation be as above and gg denote the genus of the canonical compact Riemann surface attached to G\G\backslash\mathbb{H}. Then

  1. (i)

    j=02(1)jdim𝕂Hj(G,V)=(22g|𝒞(G)|)dim𝕂V\sum_{j=0}^{2}(-1)^{j}\emph{dim}_{\mathbb{K}}H^{j}(G,V)=\big{(}2-2g-\lvert\mathcal{C}(G)\rvert\big{)}\emph{dim}_{\mathbb{K}}V
    j=1r(dim𝕂Vnjell)\hskip 213.39566pt-\sum_{j=1}^{r}(\emph{dim}_{\mathbb{K}}V-n_{j}^{\emph{ell}}),

  2. (ii)

    dim𝕂H𝖯1(G,V)=(2g2)dim𝕂V+n0+n1\emph{dim}_{\mathbb{K}}H^{1}_{\mathsf{P}}(G,V)=(2g-2)\emph{dim}_{\mathbb{K}}V+n_{0}+n_{1}
    +x𝒞(G)nxpar+j=1r(dim𝕂Vnjell)\hskip 142.26378pt+\sum_{x\in\mathcal{C}(G)}n_{x}^{\emph{par}}+\sum_{j=1}^{r}(\emph{dim}_{\mathbb{K}}V-n^{\emph{ell}}_{j}).

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 mjm_{j} be the number of GG-inequivalent jj-simplices of 𝔎\mathfrak{K}. Here A1A_{1}, resp. A2A_{2}, is a free 𝕂[G]\mathbb{K}[G]-module of rank m1m_{1}, resp. m2m_{2} but the elliptic fixed points may prevent A0A_{0} from being a free module. A straightforward analysis demonstrates that

dim𝕂A0(V)=m0dim𝕂Vj=1r(dim𝕂Vnjell),dim𝕂A1(V)=m1dim𝕂V,dim𝕂A2(V)=m2dim𝕂V.\begin{gathered}\text{dim}_{\mathbb{K}}A^{0}(V)=m_{0}\text{dim}_{\mathbb{K}}V-\sum_{j=1}^{r}(\text{dim}_{\mathbb{K}}V-n_{j}^{\text{ell}}),\\ \text{dim}_{\mathbb{K}}A^{1}(V)=m_{1}\text{dim}_{\mathbb{K}}V,\hskip 28.45274pt\text{dim}_{\mathbb{K}}A^{2}(V)=m_{2}\text{dim}_{\mathbb{K}}V.\end{gathered}

But

j=02(1)jdim𝕂Hj(G,V)=j=02(1)jdim𝕂Aj(V)\sum_{j=0}^{2}(-1)^{j}\text{dim}_{\mathbb{K}}H^{j}(G,V)=\sum_{j=0}^{2}(-1)^{j}\text{dim}_{\mathbb{K}}A^{j}(V)

and m0m1+m2=22g|𝒞(G)|m_{0}-m_{1}+m_{2}=2-2g-\lvert\mathcal{C}(G)\rvert. We now substitute the values of dim𝕂Aj(V)\text{dim}_{\mathbb{K}}A^{j}(V) in the identity above to arrive at (i). ∎

Lemma 5.4.

Let GG be a finite index subgroup of PΓ\emph{P}\Gamma and VV be a right 𝕂[G]\mathbb{K}[G]-module. Then H2(G,V)=0H^{2}(G,V)=0.

Proof.

Let HH be a finite index subgroup contained in GG so that HH is torsion-free. Then the arguments given in Proposition 1 of [10, 6.1] demonstrates that H2(H,V)=0H^{2}(H,V)=0. But the natural restriction map from H2(G,V)H^{2}(G,V) to H2(H,V)H^{2}(H,V) is injective (Remark 4.6). Therefore H2(G,V)H^{2}(G,V) 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 Γ\Gamma and PΓ\text{P}\Gamma. Let 𝒢\mathcal{G} be a subgroup of Γ\Gamma and VV be a right 𝕂[𝒢]\mathbb{K}[\mathcal{G}] module. If 𝒢\mathcal{G} is a finite index subgroup and 𝖯\mathsf{P} is the subset of all parabolic elements of 𝒢\mathcal{G} then one can introduce the notion of parabolic subspace, denoted H𝖯1(𝒢,V)H^{1}_{\mathsf{P}}(\mathcal{G},V), using the same formalism mentioned before. The subspace H𝖯1(𝒢,V)H^{1}_{\mathsf{P}}(\mathcal{G},V) of H1(𝒢,V)H^{1}(\mathcal{G},V) is stable under the action of double coset operators on cohomology [16, 8.3]. We call (𝒢,V)(\mathcal{G},V) a descent module if either Id𝒢-\text{Id}\notin\mathcal{G} or Id𝒢-\text{Id}\in\mathcal{G} acts trivially on VV. For example, the familiar pair (𝒢,Vk2,𝕂)(\mathcal{G},V_{k-2,\mathbb{K}}) is a descent module if either kk is even or Id𝒢-\text{Id}\notin\mathcal{G}. Let (𝒢,V)(\mathcal{G},V) be a descent module. Then there exists a natural 𝕂[P𝒢]\mathbb{K}[\text{P}\mathcal{G}]-structure on VV so that the 𝕂[𝒢]\mathbb{K}[\mathcal{G}]-structure on VV arises from the projection 𝒢P𝒢\mathcal{G}\twoheadrightarrow\text{P}\mathcal{G}. Moreover, the pullback of the projection map identifies H1(P𝒢,V)H^{1}(\text{P}\mathcal{G},V) and H1(𝒢,V)H^{1}(\mathcal{G},V) [16, 8.2]. Now suppose 𝒢\mathcal{G} is a finite index subgroup of Γ\Gamma. Note that the image of 𝖯\mathsf{P} in P𝒢\text{P}\mathcal{G}, denoted 𝖯¯\bar{\mathsf{P}}, equals the collection of parabolic elements of P𝒢\text{P}\mathcal{G} and there is an identification H𝖯1(𝒢,V)H𝖯¯1(P𝒢,V)H^{1}_{\mathsf{P}}(\mathcal{G},V)\cong H^{1}_{\bar{\mathsf{P}}}(\text{P}\mathcal{G},V). The procedure described above allows us to rewrite (5.1) as

(5.3) 0H𝖯1(𝒢,V)H1(𝒢,V)x𝒞(𝒢)H1(𝒢x,V).0\to H^{1}_{\mathsf{P}}(\mathcal{G},V)\to H^{1}(\mathcal{G},V)\to\oplus_{x\in\mathcal{C}(\mathcal{G})}H^{1}(\mathcal{G}_{x},V).

And furthermore, the change of scalar property of Lemma 5.2 lifts to 𝒢\mathcal{G} in the following manner

(5.4) H1(𝒢,V)𝕂𝕃H1(𝒢,V𝕂𝕃),H𝖯1(𝒢,V)𝕂𝕃H𝖯1(𝒢,V𝕂𝕃)H^{1}(\mathcal{G},V)\otimes_{\mathbb{K}}\mathbb{L}\cong H^{1}(\mathcal{G},V\otimes_{\mathbb{K}}\mathbb{L}),\;H^{1}_{\mathsf{P}}(\mathcal{G},V)\otimes_{\mathbb{K}}\mathbb{L}\cong H^{1}_{\mathsf{P}}(\mathcal{G},V\otimes_{\mathbb{K}}\mathbb{L})

whenever 𝕃\mathbb{L} is a field extension of 𝕂\mathbb{K}.

Remark 5.5.

Let 𝒢\mathcal{G} be a finite index subgroup so that Id𝒢-\text{Id}\in\mathcal{G} and Id-\text{Id} acts on VV by IdV-\text{Id}_{V}. Then H1(𝒢,V)H^{1}(\mathcal{G},V) is zero (Example 4.5) and the change of coefficients isomorphisms (5.4) trivially hold for such pairs. We use this observation for the pair (𝒢,Vk2,𝕂)(\mathcal{G},V_{k-2,\mathbb{K}}) where kk is odd and Id𝒢-\text{Id}\in\mathcal{G} to obtain change of scalars isomorphisms for all (𝒢,Vk2,𝕂)(\mathcal{G},V_{k-2,\mathbb{K}}).

The final topic of this subsection is the third term appearing in (5.3). We first consider the subgroup ThΓ\langle\langle T^{h}\rangle\rangle\subseteq\Gamma for some positive integer hh where \langle\langle\cdot\rangle\rangle refers to the subgroup generated by the element. Since ThΓ\langle\langle T^{h}\rangle\rangle\subseteq\Gamma is an infinite cyclic group it follows that

(5.5) H1(Th,Vk2,𝕂)Vk2,𝕂Vk2,𝕂(Th1)𝕂.H^{1}(\langle\langle T^{h}\rangle\rangle,V_{k-2,\mathbb{K}})\cong\frac{V_{k-2,\mathbb{K}}}{V_{k-2,\mathbb{K}}(T^{h}-1)}\cong\mathbb{K}.

Here the last isomorphism stems from a linear functional Vk2,𝕂𝕂V_{k-2,\mathbb{K}}\to\mathbb{K} described as PP\mapsto the coefficient of Xk2X^{k-2} in P(Th)P(T^{h}). Let 𝒢\mathcal{G} be a finite index subgroup Γ\Gamma and x𝒞(𝒢)x\in\mathcal{C}(\mathcal{G}). Choose γΓ\gamma\in\Gamma so that x=𝒢(γ)x=\mathcal{G}(\gamma\infty). Then γ1𝒢xγ=γ1𝒢γΓ\gamma^{-1}\mathcal{G}_{x}\gamma=\gamma^{-1}\mathcal{G}\gamma\cap\Gamma_{\infty}. In more detail, there exists a positive integer hh so that

γ1𝒢xγ={Id,Th or Th,if x𝒞(𝒢);Th,if x𝒞(𝒢)\gamma^{-1}\mathcal{G}_{x}\gamma=\begin{cases}\langle\langle-\text{Id},T^{h}\rangle\rangle\text{ or }\langle\langle T^{h}\rangle\rangle,&\text{if $x\in\mathcal{C}_{\infty}(\mathcal{G})$;}\\ \langle\langle-T^{h}\rangle\rangle,&\text{if $x\notin\mathcal{C}_{\infty}(\mathcal{G})$}\end{cases}

where 𝒞(𝒢)\mathcal{C}_{\infty}(\mathcal{G}) is the collection of regular cusps. The following result is certainly well-known among the experts; cf. [10, 6.3].

Proposition 5.6.

Let 𝒢\mathcal{G} be as above and assume that Id𝒢-\emph{Id}\notin\mathcal{G} if kk is odd. Then

H1(𝒢x,Vk2,𝕂){𝕂,if k is even;𝕂,if k is odd and x𝒞(𝒢);0,if k is odd and x𝒞(𝒢).H^{1}(\mathcal{G}_{x},V_{k-2,\mathbb{K}})\cong\begin{cases}\mathbb{K},&\text{if $k$ is even;}\\ \mathbb{K},&\text{if $k$ is odd and $x\in\mathcal{C}_{\infty}(\mathcal{G})$;}\\ 0,&\text{if $k$ is odd and $x\notin\mathcal{C}_{\infty}(\mathcal{G})$.}\end{cases}
Proof.

We use the conjugation by γ\gamma isomorphism (Section 4.3) to obtain

H1(𝒢x,Vk2,𝕂)H1(γ1𝒢xγ,Vk2,𝕂).H^{1}(\mathcal{G}_{x},V_{k-2,\mathbb{K}})\cong H^{1}(\gamma^{-1}\mathcal{G}_{x}\gamma,V_{k-2,\mathbb{K}}).

If γ1𝒢xγ=Th\gamma^{-1}\mathcal{G}_{x}\gamma=\langle\langle T^{h}\rangle\rangle then H1(γ1𝒢xγ,Vk2,𝕂)𝕂H^{1}(\gamma^{-1}\mathcal{G}_{x}\gamma,V_{k-2,\mathbb{K}})\cong\mathbb{K}. Now suppose γ1𝒢xγ=Id,Th\gamma^{-1}\mathcal{G}_{x}\gamma=\langle\langle-\text{Id},T^{h}\rangle\rangle. Observe that

H1(Id,Th,Vk2,𝕂)H1(Th,Vk2,𝕂)Id,Th.H^{1}(\langle\langle-\text{Id},T^{h}\rangle\rangle,V_{k-2,\mathbb{K}})\cong H^{1}(\langle\langle T^{h}\rangle\rangle,V_{k-2,\mathbb{K}})^{\langle\langle-\text{Id},T^{h}\rangle\rangle}.

Here kk is even and Id-\text{Id} acts trivially on H1(Th,Vk2,𝕂)H^{1}(\langle\langle T^{h}\rangle\rangle,V_{k-2,\mathbb{K}}); see (4.5). As a consequence H1(Id,Th,Vk2,𝕂)𝕂H^{1}(\langle\langle-\text{Id},T^{h}\rangle\rangle,V_{k-2,\mathbb{K}})\cong\mathbb{K}. It remains to consider the case γ1𝒢xγ=Th\gamma^{-1}\mathcal{G}_{x}\gamma=\langle\langle-T^{h}\rangle\rangle. In this situation, there is an isomorphism

H1(Th,Vk2,𝕂)H1(T2h,Vk2,𝕂)Th.H^{1}(\langle\langle-T^{h}\rangle\rangle,V_{k-2,\mathbb{K}})\cong H^{1}(\langle\langle T^{2h}\rangle\rangle,V_{k-2,\mathbb{K}})^{\langle\langle-T^{h}\rangle\rangle}.

One again employs (4.5) to conclude that Th-T^{h} acts on H1(T2h,Vk2,𝕂)H^{1}(\langle\langle T^{2h}\rangle\rangle,V_{k-2,\mathbb{K}}) by (1)k2(-1)^{k-2}. Hence H1(Th,Vk2,𝕂)𝕂H^{1}(\langle\langle-T^{h}\rangle\rangle,V_{k-2,\mathbb{K}})\cong\mathbb{K} if kk is even and equals zero if kk is odd. ∎

5.2 Isomorphism theorem for cusp forms

We first describe the action of complex conjugation on our cohomology spaces. Let kk be an integer 2\geq 2 and 𝒢\mathcal{G} be a finite index subgroup of Γ\Gamma. The complex conjugation on \mathbb{C} gives rise to an conjugate linear automorphism of Z1(𝒢,Vk2,)Z^{1}\big{(}\mathcal{G},V_{k-2,\mathbb{C}}\big{)} described as P#PP\mapsto\#P where #P(γ)=P(γ)¯\#P(\gamma)=\overline{P(\gamma)}. This automorphism descends to a complex conjugation on H1(𝒢,Vk2,)H^{1}\big{(}\mathcal{G},V_{k-2,\mathbb{C}}\big{)}. Now suppose 𝒮k(𝒢)\mathcal{S}_{k}(\mathcal{G}) is the space of all weight kk cusp form on 𝒢\mathcal{G} and 𝒮ka(𝒢)={f¯f𝒮k(𝒢)}\mathcal{S}^{a}_{k}(\mathcal{G})=\{\bar{f}\mid f\in\mathcal{S}_{k}(\mathcal{G})\} is the space of antiholomorphic cusp forms on 𝒢\mathcal{G}. Let

[ESk]c,𝒢:𝒮k(𝒢)H1(𝒢,Vk2,)[\text{ES}_{k}]_{\text{c},\mathcal{G}}:\mathcal{S}_{k}(\mathcal{G})\to H^{1}(\mathcal{G},V_{k-2,\mathbb{C}})

be the Eichler-Shimura homomorphism constructed in Section 4. One defines an antiholomorphic avatar of [ESk]c,𝒢[\text{ES}_{k}]_{\text{c},\mathcal{G}} by

[ESk]c,𝒢a:𝒮ka(𝒢)H1(𝒢,Vk2,),f#[ESk]c,𝒢(f¯).[\text{ES}_{k}]^{a}_{\text{c},\mathcal{G}}:\mathcal{S}^{a}_{k}(\mathcal{G})\xrightarrow{}H^{1}(\mathcal{G},V_{k-2,\mathbb{C}}),\hskip 8.5359ptf\to\#[\text{ES}_{k}]_{\text{c},\mathcal{G}}(\bar{f}).

It is clear that [ESk]c,𝒢a[\text{ES}_{k}]^{a}_{\text{c},\mathcal{G}} is \mathbb{C}-linear. We extend the action of a double coset operator [𝒢α𝒢]k[\mathcal{G}\alpha\mathcal{G}]_{k} to the space of antiholomorphic cusp form using the formula f#[𝒢α𝒢]k(f¯)f\to\#[\mathcal{G}\alpha\mathcal{G}]_{k}(\bar{f}). Since the action of [𝒢α𝒢]k[\mathcal{G}\alpha\mathcal{G}]_{k} on a cocycle preserves the decomposition into real and imaginary parts it follows that [ESk]c,𝒢a[\text{ES}_{k}]^{a}_{\text{c},\mathcal{G}} is also Hecke equivariant. Hence [ESk]c,𝒢[\text{ES}_{k}]_{\text{c},\mathcal{G}} and [ESk]c,𝒢a[\text{ES}_{k}]^{a}_{\text{c},\mathcal{G}} together provide a Hecke-equivariant \mathbb{C}-linear map

[ESk]c,𝒢[ESk]c,𝒢a:𝒮k(𝒢)𝒮ka(𝒢)H1(𝒢,Vk2,).[\text{ES}_{k}]_{\text{c},\mathcal{G}}\oplus[\text{ES}_{k}]^{a}_{\text{c},\mathcal{G}}:\mathcal{S}_{k}(\mathcal{G})\oplus\mathcal{S}^{a}_{k}(\mathcal{G})\to H^{1}(\mathcal{G},V_{k-2,\mathbb{C}}).

The Eichler-Shimura isomorphism theorem for the cusp forms is as follows:

Theorem 5.7.

[10, p.171] The map [ESk]c,𝒢[ESk]c,𝒢a[\emph{ES}_{k}]_{\emph{c},\mathcal{G}}\oplus[\emph{ES}_{k}]^{a}_{\emph{c},\mathcal{G}} is an injection whose image equals the parabolic subspace of H1(𝒢,Vk2,)H^{1}(\mathcal{G},V_{k-2,\mathbb{C}}).

Proof.

For [P]H1(𝒢,Vk2,)[P]\in H^{1}(\mathcal{G},V_{k-2,\mathbb{C}}) write ([P])=[P]+#[P]2\mathfrak{R}([P])=\frac{[P]+\#[P]}{2} and ([P])=[P]#[P]2i\mathfrak{I}([P])=\frac{[P]-\#[P]}{2i}. The \mathbb{R}-linear version of the isomorphism theorem given in [16, 8.2] asserts that the assignment 𝒮k(𝒢)H𝖯1(𝒢,Vk2,)\mathcal{S}_{k}(\mathcal{G})\to H^{1}_{\mathsf{P}}(\mathcal{G},V_{k-2,\mathbb{R}}), f([ESk]c,𝒢(f))f\mapsto\mathfrak{R}\big{(}[\text{ES}_{k}]_{c,\mathcal{G}}(f)\big{)}, is an isomorphism of real vector spaces. One twists the map above by the \mathbb{R}-linear automorphism fiff\to if of 𝒮k(𝒢)\mathcal{S}_{k}(\mathcal{G}) to deduce that the imaginary part map f([ESk]c,𝒢(f))f\mapsto\mathfrak{I}\big{(}[\text{ES}_{k}]_{c,\mathcal{G}}(f)\big{)} also defines a real linear isomorphism between 𝒮k(𝒢)\mathcal{S}_{k}(\mathcal{G}) and H𝖯1(𝒢,Vk2,)H^{1}_{\mathsf{P}}(\mathcal{G},V_{k-2,\mathbb{R}}). 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 kk be an integer 2\geq 2 and 𝒢\mathcal{G} be a finite index subgroup of Γ\Gamma. Then [𝗉k]𝒢,Hol()[\mathsf{p}_{k}]_{\mathcal{G},\emph{Hol}(\mathbb{H})} is nonzero.

Proof.

Let 𝒢𝒢\mathcal{G}^{\prime}\subseteq\mathcal{G} be a finite index subgroup of Γ\Gamma with [𝒢:𝒢][\mathcal{G}:\mathcal{G}^{\prime}] so large that 𝒮k(𝒢)\mathcal{S}_{k}(\mathcal{G}^{\prime}) is nonzero. One uses (2.13) and (2.14) to discover that

[𝗉k]𝒢,𝒮k(𝒢)=R𝒮k(𝒢)Hol()R𝒢𝒢[𝗉k]𝒢,Hol().[\mathsf{p}_{k}]_{\mathcal{G}^{\prime},\mathcal{S}_{k}(\mathcal{G}^{\prime})}=\text{R}^{\text{Hol}(\mathbb{H})}_{\mathcal{S}_{k}(\mathcal{G}^{\prime})}\text{R}^{\mathcal{G}}_{\mathcal{G}^{\prime}}[\mathsf{p}_{k}]_{\mathcal{G},\text{Hol}(\mathbb{H})}.

But [ESk]𝒢,𝒮k(𝒢)[\text{ES}_{k}]_{\mathcal{G}^{\prime},\mathcal{S}_{k}(\mathcal{G}^{\prime})} is a nonzero injective map. Therefore by Proposition 4.2 the class [𝗉k]𝒢,𝒮k(𝒢)[\mathsf{p}_{k}]_{\mathcal{G}^{\prime},\mathcal{S}_{k}(\mathcal{G}^{\prime})} is nonzero. Thus [𝗉k]𝒢,Hol()[\mathsf{p}_{k}]_{\mathcal{G},\text{Hol}(\mathbb{H})} cannot be zero. ∎

5.3 The space of Eisenstein series

Let kk be an integer 2\geq 2 and 𝒢\mathcal{G} be a congruence subgroup of Γ\Gamma. Suppose that k(𝒢)\mathcal{M}_{k}(\mathcal{G}) is the space of weight kk modular forms on 𝒢\mathcal{G}. Then the space of Eisenstein series on 𝒢\mathcal{G}, denoted k(𝒢)\mathcal{E}_{k}(\mathcal{G}), is the unique linear complement of 𝒮k(𝒢)\mathcal{S}_{k}(\mathcal{G}) in k(𝒢)\mathcal{M}_{k}(\mathcal{G}) that is orthogonal to 𝒮k(𝒢)\mathcal{S}_{k}(\mathcal{G}) with respect to the Petersson inner product [7, 5.11]. The equivariance properties of the inner product imply that the decomposition k(𝒢)=k(𝒢)𝒮k(𝒢)\mathcal{M}_{k}(\mathcal{G})=\mathcal{E}_{k}(\mathcal{G})\oplus\mathcal{S}_{k}(\mathcal{G}) is stable under the action of a general double coset operator. Next, we explain how to use the spectral Eisenstein series on 𝒢\mathcal{G} to construct a basis for the space of Eisenstein series following the author’s recent account of the theory [15].

Let x𝒞(𝒢)x\in\mathcal{C}(\mathcal{G}) and σxΓ\sigma_{x}\in\Gamma be a scaling matrix for xx, i.e., x=σxx=\sigma_{x}\infty. If kk is odd then additionally assume that Id𝒢-\text{Id}\notin\mathcal{G} and xx is a regular cusp. The spectral Eisenstein series of weight kk attached to xx and the scaling matrix σx\sigma_{x} is

Ek,x(τ;s):=γ𝒢x\𝒢𝐣(σx1γ,τ)k(Im σx1γτ)s(Re(k+2s)>2)E_{k,x}(\tau;s):=\sum_{\gamma\in\mathcal{G}_{x}\backslash\mathcal{G}}\mathbf{j}(\sigma_{x}^{-1}\gamma,\tau)^{-k}(\text{Im }\sigma_{x}^{-1}\gamma\tau)^{s}\hskip 8.5359pt\begin{subarray}{c}(\text{Re}(k+2s)>2)\end{subarray}

where 𝐣(γ,τ)=cτ+d\mathbf{j}(\gamma,\tau)=c\tau+d for γ=(abcd)Γ\gamma=\begin{pmatrix}a&b\\ c&d\end{pmatrix}\in\Gamma. If kk is odd then our Eisenstein series may depend on the choice of scaling matrix. The series above admits analytic continuation to the larger domain Re(k+2s)>1\text{Re}(k+2s)>1 allowing us to specialize at (2,0)(2,0). For simplicity we abbreviate Ek,x(τ;0)E_{k,x}(\tau;0) as Ek,x(τ)E_{k,x}(\tau). An explicit calculation with the Fourier series shows that

Ek,x(τ){Hol(),k3,αxπIm(τ)+Hol(),k=2E_{k,x}(\tau)\in\begin{cases}\text{Hol}(\mathbb{H}),&\text{$k\geq 3$,}\\ \frac{\alpha_{x}}{\pi\text{Im}(\tau)}+\text{Hol}(\mathbb{H}),&\text{$k=2$}\end{cases}

where αx\alpha_{x} is a nonzero rational number. Moreover, we have

(5.6) π(Ek,x|γ)={(±1)k,if γ=x in 𝒞(𝒢);0,otherwise.(γΓ)\pi_{\infty}(E_{k,x}\lvert_{\gamma})=\begin{cases}(\pm 1)^{k},&\text{if $\gamma\infty=x$ in $\mathcal{C}(\mathcal{G})$;}\\ 0,&\text{otherwise.}\end{cases}\hskip 8.5359pt(\begin{subarray}{c}\forall\gamma\in\Gamma\end{subarray})

Set

𝒢,k={{Ek,xx𝒞(𝒢)},if k is even and 4;{Ek,xx𝒞(𝒢)},if k is odd and Id𝒢;{Ek,xαxαEk,x𝒞(𝒢){}},if k=2;,otherwise\mathcal{B}_{\mathcal{G},k}=\begin{cases}\{E_{k,x}\mid x\in\mathcal{C}(\mathcal{G})\},&\begin{subarray}{c}\text{if $k$ is even and $\geq 4$;}\end{subarray}\\ \{E_{k,x}\mid x\in\mathcal{C}_{\infty}(\mathcal{G})\},&\begin{subarray}{c}\text{if $k$ is odd and $-\text{Id}\notin\mathcal{G}$;}\end{subarray}\\ \{E_{k,x}-\frac{\alpha_{x}}{\alpha_{\infty}}E_{k,\infty}\mid x\in\mathcal{C}(\mathcal{G})-\{\infty\}\},&\begin{subarray}{c}\text{if $k=2$;}\end{subarray}\\ \emptyset,&\begin{subarray}{c}\text{otherwise}\end{subarray}\end{cases}

where in the second case we write the series for a fixed choice of scaling matrix for each regular cusp. Then 𝒢,k\mathcal{B}_{\mathcal{G},k} is a basis for the space of Eisenstein series k(𝒢)\mathcal{E}_{k}(\mathcal{G}). For a subfield 𝕂\mathbb{K} of \mathbb{C} we write k(𝒢,𝕂)\mathcal{E}_{k}(\mathcal{G},\mathbb{K}) for the 𝕂\mathbb{K} span of 𝒢,k\mathcal{B}_{\mathcal{G},k}. This 𝕂\mathbb{K}-structure on the space of Eisenstein series is compatible with restriction to a principal level. In more detail, if Γ(N)𝒢\Gamma(N)\subseteq\mathcal{G} then

(5.7) k(𝒢,𝕂)=k(Γ(N),𝕂)𝒢.\mathcal{E}_{k}(\mathcal{G},\mathbb{K})=\mathcal{E}_{k}\big{(}\Gamma(N),\mathbb{K}\big{)}^{\mathcal{G}}.

To write down uniform statements for all weights 2\geq 2 we introduce an extended space of Eisenstein series

k(𝒢,𝕂)={𝕂-span of {Ek,xx𝒞(𝒢)},if k=2;k(𝒢,𝕂),if k3.\mathcal{E}^{\ast}_{k}(\mathcal{G},\mathbb{K})=\begin{cases}\text{$\mathbb{K}$-span of $\{E_{k,x}\mid x\in\mathcal{C}(\mathcal{G})\}$,}&\text{if $k=2$;}\\ \mathcal{E}_{k}(\mathcal{G},\mathbb{K}),&\text{if $k\geq 3$.}\end{cases}

It is clear that 2(𝒢,𝕂)Hol()=2(𝒢,𝕂)\mathcal{E}^{\ast}_{2}(\mathcal{G},\mathbb{K})\cap\text{Hol}(\mathbb{H})=\mathcal{E}_{2}(\mathcal{G},\mathbb{K}) and 2(𝒢,𝕂)\mathcal{E}_{2}(\mathcal{G},\mathbb{K}) is a subspace of codimension 11 inside 2(𝒢,𝕂)\mathcal{E}^{\ast}_{2}(\mathcal{G},\mathbb{K}). The 𝕂\mathbb{K}-rational structure on the extended space of weight 22 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 kk for every finite index subgroup so that their specializations at s=0s=0 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 kk be an integer 2\geq 2 and 𝒢\mathcal{G} be a congruence subgroup of Γ\Gamma. Recall the decomposition k(𝒢)=k(𝒢)𝒮k(𝒢)\mathcal{M}_{k}(\mathcal{G})=\mathcal{E}_{k}(\mathcal{G})\oplus\mathcal{S}_{k}(\mathcal{G}) from Section 5.3. Suppose that

[ESk]m,𝒢:k(𝒢)H1(𝒢,Vk2,)[\text{ES}_{k}]_{\text{m},\mathcal{G}}:\mathcal{M}_{k}(\mathcal{G})\xrightarrow{}H^{1}(\mathcal{G},V_{k-2,\mathbb{C}})

is the Eichler-Shimura homomorphism attached to the space of modular forms on 𝒢\mathcal{G}. Then [ESk]m,𝒢[\text{ES}_{k}]_{\text{m},\mathcal{G}} is a Hecke equivariant extension of the map [ESk]c,𝒢[\text{ES}_{k}]_{\text{c},\mathcal{G}} on the space of cusp forms. We need to prove that

[ESk]m,𝒢[ESk]c,𝒢a:k(𝒢)𝒮ka(𝒢)H1(𝒢,Vk2,)[\text{ES}_{k}]_{\text{m},\mathcal{G}}\oplus[\text{ES}_{k}]^{a}_{\text{c},\mathcal{G}}:\mathcal{M}_{k}(\mathcal{G})\oplus\mathcal{S}^{a}_{k}(\mathcal{G})\to H^{1}(\mathcal{G},V_{k-2,\mathbb{C}})

is an isomorphism. If kk is odd and Id𝒢-\text{Id}\in\mathcal{G} then the domain and codomain of the map above are zero (Example 4.5). Assume that Id𝒢-\text{Id}\notin\mathcal{G} whenever kk is odd. Let [ESk]E,𝒢[\text{ES}_{k}]_{\text{E},\mathcal{G}} denote the restriction of [ESk]m,𝒢[\text{ES}_{k}]_{\text{m},\mathcal{G}} to the space of Eisenstein series on 𝒢\mathcal{G}. In the light of Theorem 5.7 it suffices to verify that [ESk]E,𝒢[\text{ES}_{k}]_{\text{E},\mathcal{G}} is injective and the image of [ESk]E,𝒢[\text{ES}_{k}]_{\text{E},\mathcal{G}} is a linear complement of H𝖯1(𝒢,Vk2,)H^{1}_{\mathsf{P}}(\mathcal{G},V_{k-2,\mathbb{C}}) in H1(𝒢,Vk2,)H^{1}(\mathcal{G},V_{k-2,\mathbb{C}}).

6.1 Restriction to the cusps

With each x𝒞(𝒢)x\in\mathcal{C}(\mathcal{G}) one can associate a cuspidal Eichler-Shimura map

Θx:k(𝒢)H1(𝒢x,Vk2,),Θx=Resx[ESk]m,𝒢\Theta_{x}:\mathcal{M}_{k}(\mathcal{G})\to H^{1}(\mathcal{G}_{x},V_{k-2,\mathbb{C}}),\hskip 8.5359pt\Theta_{x}=\text{Res}_{x}\circ[\text{ES}_{k}]_{\text{m},\mathcal{G}}

where Resx\text{Res}_{x} is the restriction map from H1(𝒢,Vk2,)H^{1}(\mathcal{G},V_{k-2,\mathbb{C}}) to H1(𝒢x,Vk2,)H^{1}(\mathcal{G}_{x},V_{k-2,\mathbb{C}}). Proposition 5.6 furnishes a complete description of the target of Θx\Theta_{x}. The subsequent discussion aims to find the kernel of this map.

Let x𝒞(𝒢)x\in\mathcal{C}(\mathcal{G}) and choose γ0Γ\gamma_{0}\in\Gamma so that x=γ0x=\gamma_{0}\infty. Now suppose Γ(N)𝒢\Gamma(N)\subseteq\mathcal{G} with N3N\geq 3. Note that the restriction maps fit into a commutative diagram

(6.1) H1(𝒢,Vk2,){H^{1}(\mathcal{G},V_{k-2,\mathbb{C}})}H1(Γ(N),Vk2,){H^{1}(\Gamma(N),V_{k-2,\mathbb{C}})}H1(𝒢x,Vk2,){H^{1}(\mathcal{G}_{x},V_{k-2,\mathbb{C}})}H1(Γ(N)x,Vk2,){H^{1}(\Gamma(N)_{x},V_{k-2,\mathbb{C}})}

exactly as in the proof of Lemma 5.1. Here Γ(N)\Gamma(N) is a normal subgroup of Γ\Gamma and Γ\Gamma acts by conjugation on H1(Γ(N),Vk2,)H^{1}\big{(}\Gamma(N),V_{k-2,\mathbb{C}}\big{)} (Section 4.3). The conjugation action of γ0\gamma_{0} yields another commutative diagram described by

(6.2) H1(Γ(N),Vk2,){H^{1}\big{(}\Gamma(N),V_{k-2,\mathbb{C}}\big{)}}H1(Γ(N),Vk2,){H^{1}\big{(}\Gamma(N),V_{k-2,\mathbb{C}}\big{)}}H1(Γ(N)x,Vk2,){H^{1}\big{(}\Gamma(N)_{x},V_{k-2,\mathbb{C}}\big{)}}H1(Γ(N),Vk2,){H^{1}\big{(}\Gamma(N)_{\infty},V_{k-2,\mathbb{C}}\big{)}}αα|γ0\scriptstyle{\alpha\to\alpha\lvert_{\gamma_{0}}}c(γ0)1\scriptstyle{c(\gamma_{0})^{-1}}

where the vertical arrows are restriction maps. Next, we use our equivariant machinery to verify the key property of the map Θx\Theta_{x}.

Proposition 6.1.

Suppose that fk(𝒢)f\in\mathcal{M}_{k}(\mathcal{G}). Then Θx(f)=0\Theta_{x}(f)=0 if and only if π(f|γ0)=0\pi_{\infty}(f\lvert_{\gamma_{0}})=0.

Proof.

We identify k(𝒢)\mathcal{M}_{k}(\mathcal{G}), resp. H1(𝒢,Vk2,)H^{1}(\mathcal{G},V_{k-2,\mathbb{C}}), with the 𝒢\mathcal{G}-invariant subspace of k(Γ(N))\mathcal{M}_{k}(\Gamma(N)), resp. H1(Γ(N),Vk2,)H^{1}\big{(}\Gamma(N),V_{k-2,\mathbb{C}}\big{)}, and utilize

[ESk]m,Γ(N):k(Γ(N))H1(Γ(N),Vk2,)[\text{ES}_{k}]_{\text{m},\Gamma(N)}:\mathcal{M}_{k}(\Gamma(N))\xrightarrow{}H^{1}\big{(}\Gamma(N),V_{k-2,\mathbb{C}}\big{)}

to describe [ESk]m,𝒢[\text{ES}_{k}]_{\text{m},\mathcal{G}} (Remark 4.8). First, one invokes Remark 4.6 to conclude that the bottom arrow of (6.1) is injective. Hence Θx(f)=0\Theta_{x}(f)=0 if and only if the restriction of [ESk]m,Γ(N)(f)[\text{ES}_{k}]_{\text{m},\Gamma(N)}(f) to H1(Γ(N)x,Vk2,)H^{1}\big{(}\Gamma(N)_{x},V_{k-2,\mathbb{C}}\big{)} is zero. But [ESk]m,Γ(N)[\text{ES}_{k}]_{\text{m},\Gamma(N)} is Γ\Gamma-equivariant (Proposition 4.7). Therefore [ESk]m,Γ(N)(f)|γ0=[ESk]m,Γ(N)(f|γ0)[\text{ES}_{k}]_{\text{m},\Gamma(N)}(f)\lvert_{\gamma_{0}}=[\text{ES}_{k}]_{\text{m},\Gamma(N)}(f\lvert_{\gamma_{0}}). One now employs (6.2) to conclude that Θx(f)=0\Theta_{x}(f)=0 if and only if the restriction of [ESk]m,Γ(N)(f|γ0)[\text{ES}_{k}]_{\text{m},\Gamma(N)}(f\lvert_{\gamma_{0}}) to H1(Γ(N),Vk2,)H^{1}\big{(}\Gamma(N)_{\infty},V_{k-2,\mathbb{C}}\big{)} is zero. Note that Γ(N)=TN\Gamma(N)_{\infty}=\langle\langle T^{N}\rangle\rangle. From Lemma 3.13 it follows that

𝗉k(TN,f|γ0,X)=π(f|γ0)(X+N)k1Xk1k1.\mathsf{p}^{\ast}_{k}(T^{N},f\lvert_{\gamma_{0}},X)=\pi_{\infty}(f\lvert_{\gamma_{0}})\frac{(X+N)^{k-1}-X^{k-1}}{k-1}.

Therefore, by (5.5), Θx(f)=0\Theta_{x}(f)=0 if and only if π(f|γ0)=0\pi_{\infty}(f\lvert_{\gamma_{0}})=0. ∎

One uses the theorem above to determine the cuspidal behavior of the cohomology classes attached to the basis 𝒢,k\mathcal{B}_{\mathcal{G},k} for the space of Eisenstein series. Let the notation be as in the discussion around Section 5.3. For convenience write 𝒞k(𝒢)=𝒞(𝒢)\mathcal{C}_{k}(\mathcal{G})=\mathcal{C}(\mathcal{G}) if kk is even and 𝒞(𝒢)\mathcal{C}_{\infty}(\mathcal{G}) if kk is odd.

Corollary 6.2.

Let k3k\geq 3 and assume that x,y𝒞k(𝒢)x,y\in\mathcal{C}_{k}(\mathcal{G}). Then Θx(Ek,y)0\Theta_{x}(E_{k,y})\neq 0 if and only if x=yx=y. Moreover, let k=2k=2 and x,y𝒞(𝒢){}x,y\in\mathcal{C}(\mathcal{G})-\{\infty\}. Then

Θx(E2,yαyαE2,)0\Theta_{x}\big{(}E_{2,y}-\frac{\alpha_{y}}{\alpha_{\infty}}E_{2,\infty}\big{)}\neq 0

if and only if x=yx=y.

Proof.

The assertion is an easy consequence of Proposition 6.1 and (5.6). ∎

6.2 Completion of the proof

The cuspidal Eichler-Shimura maps together give rise to a homomorphism

Θ:k(𝒢)x𝒞(𝒢)H1(𝒢x,Vk2,);Θ:=(Θx)x𝒞(𝒢).\Theta:\mathcal{E}_{k}(\mathcal{G})\to\bigoplus_{x\in\mathcal{C}(\mathcal{G})}H^{1}(\mathcal{G}_{x},V_{k-2,\mathbb{C}});\hskip 8.5359pt\Theta:=(\Theta_{x})_{x\in\mathcal{C}(\mathcal{G})}.

We use this homomorphism to extract information about the Eichler-Shimura map attached to the space of Eisenstein series.

Case I: k3k\geq 3.

Proposition 5.6 shows that H1(𝒢x,Vk2,)0H^{1}(\mathcal{G}_{x},V_{k-2,\mathbb{C}})\neq 0 if and only if x𝒞k(𝒢)x\in\mathcal{C}_{k}(\mathcal{G}). Therefore Θ\Theta maps our basis of k(𝒢)\mathcal{E}_{k}(\mathcal{G}) to a basis of x𝒞(𝒢)H1(𝒢x,Vk2,)\oplus_{x\in\mathcal{C}(\mathcal{G})}H^{1}(\mathcal{G}_{x},V_{k-2,\mathbb{C}}) (Corollary 6.2). In particular, the map Θ\Theta is injective. It follows that [ESk]E,𝒢[\text{ES}_{k}]_{\text{E},\mathcal{G}} is also injective. Moreover im([ESk]E,𝒢)H𝖯1(𝒢,Vk2,)={0}\text{im}([\text{ES}_{k}]_{\text{E},\mathcal{G}})\cap H^{1}_{\mathsf{P}}(\mathcal{G},V_{k-2,\mathbb{C}})=\{0\} where im()\text{im}(\cdot) refers to the image a map. The surjectivity of Θ\Theta demonstrates that (5.3) is exact at the right end. Therefore im([ESk]E,𝒢)\text{im}([\text{ES}_{k}]_{\text{E},\mathcal{G}}) is a linear complement of H𝖯1(𝒢,Vk2,)H^{1}_{\mathsf{P}}(\mathcal{G},V_{k-2,\mathbb{C}}) as desired.

Case II: k=2k=2.

One again uses Corollary 6.2 to conclude that Θ\Theta maps the given basis of k(𝒢)\mathcal{E}_{k}(\mathcal{G}) to a linearly independent subset of x𝒞(𝒢)H1(𝒢x,Vk2,)\oplus_{x\in\mathcal{C}(\mathcal{G})}H^{1}\big{(}\mathcal{G}_{x},V_{k-2,\mathbb{C}}\big{)}. Hence [ESk]E,𝒢[\text{ES}_{k}]_{\text{E},\mathcal{G}} is injective and im([ESk]E,𝒢)H𝖯1(𝒢,Vk2,)={0}\text{im}([\text{ES}_{k}]_{\text{E},\mathcal{G}})\cap H^{1}_{\mathsf{P}}(\mathcal{G},V_{k-2,\mathbb{C}})=\{0\}. Here Vk2,V_{k-2,\mathbb{C}} equals the trivial 𝒢\mathcal{G}-module \mathbb{C}. We combine (5.2), Proposition 5.3, and Lemma 5.4 to discover that

dimH1(𝒢,Vk2,)=dimH𝖯1(𝒢,Vk2,)+|𝒞(𝒢)|1.\text{dim}_{\mathbb{C}}H^{1}(\mathcal{G},V_{k-2,\mathbb{C}})=\text{dim}_{\mathbb{C}}H^{1}_{\mathsf{P}}(\mathcal{G},V_{k-2,\mathbb{C}})+\lvert\mathcal{C}(\mathcal{G})\rvert-1.

But dimim([ESk]E,𝒢)=dimk(𝒢)=|𝒞(𝒢)|1\text{dim}_{\mathbb{C}}\text{im}\big{(}[\text{ES}_{k}]_{\text{E},\mathcal{G}}\big{)}=\text{dim}_{\mathbb{C}}\mathcal{E}_{k}(\mathcal{G})=\lvert\mathcal{C}(\mathcal{G})\rvert-1. Thus im([ESk]E,𝒢)\text{im}([\text{ES}_{k}]_{\text{E},\mathcal{G}}) is a linear complement of H𝖯1(𝒢,Vk2,)H^{1}_{\mathsf{P}}(\mathcal{G},V_{k-2,\mathbb{C}}) in H1(𝒢,Vk2,)H^{1}(\mathcal{G},V_{k-2,\mathbb{C}}).

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 \mathcal{H} be an abstract group and 𝒢\mathcal{G} be a subgroup of \mathcal{H}. Suppose that VV is a right 𝕂[𝒢]\mathbb{K}[\mathcal{G}]-module where 𝕂\mathbb{K} is a field of characteristics zero. Write

Ind𝒢(V):=Hom𝕂[𝒢](𝕂[],V)\text{Ind}^{\mathcal{G}}_{\mathcal{H}}(V):=\text{Hom}_{\mathbb{K}[\mathcal{G}]}(\mathbb{K}[\mathcal{H}],V)

and endow it with the right \mathcal{H}-action (fh)(x)=f(hx)(fh)(x)=f(hx). If the 𝕂[𝒢]\mathbb{K}[\mathcal{G}]-module structure on VV extends to a 𝕂[]\mathbb{K}[\mathcal{H}]-structure then Ind𝒢(V)Fun(𝒢\,V)\text{Ind}^{\mathcal{G}}_{\mathcal{H}}(V)\cong\text{Fun}(\mathcal{G}\backslash\mathcal{H},V) as 𝕂[]\mathbb{K}[\mathcal{H}]-module via the assignment ff~f\to\tilde{f} where f~(𝒢x):=f(x1)x\tilde{f}(\mathcal{G}x):=f(x^{-1})x. Shapiro’s lemma states that the canonical homomorphism of 𝒢\mathcal{G}-modules Ind𝒢(V)V\text{Ind}^{\mathcal{G}}_{\mathcal{H}}(V)\to V described by ff(Id)f\mapsto f(\text{Id}) gives rise to a 𝕂\mathbb{K}-linear isomorphism

sh:H(,Ind𝒢(V))H(𝒢,V).\text{sh}^{*}:H^{\ast}\big{(}\mathcal{H},\text{Ind}^{\mathcal{G}}_{\mathcal{H}}(V)\big{)}\xrightarrow{\cong}H^{\ast}(\mathcal{G},V).

Let 𝒢\mathcal{G} be a finite index subgroup of Γ\Gamma. Suppose that VV is a right 𝕂[Γ]\mathbb{K}[\Gamma]-module. We consider the Shapiro isomorphism

sh1:H1(Γ,Fun(𝒢\Γ,V))H1(𝒢,V)\text{sh}^{1}:H^{1}\big{(}\Gamma,\text{Fun}(\mathcal{G}\backslash\Gamma,V)\big{)}\xrightarrow{\cong}H^{1}(\mathcal{G},V)

explicitly described by

(7.1) [P][ψ(P)];ψ(P)(γ):=P(γ)(𝒢Id).(γ𝒢)[P]\to[\psi(P)];\hskip 8.5359pt\psi(P)(\gamma):=P(\gamma)(\mathcal{G}\text{Id}).\hskip 8.5359pt(\begin{subarray}{c}\gamma\in\mathcal{G}\end{subarray})

In practice working with the cocycles on Γ\Gamma is more convenient since each cocycle is determined by its value on {T,S}\{T,S\}. One pulls back the action of the double coset operators on H1(𝒢,V)H^{1}(\mathcal{G},V) to obtain a 𝒢\mathcal{G}-Hecke action on H1(Γ,Fun(𝒢\Γ,V))H^{1}(\Gamma,\text{Fun}(\mathcal{G}\backslash\Gamma,V)); cf. [14, 5]. The Γ\Gamma-action on Fun(𝒢\Γ,V)\text{Fun}(\mathcal{G}\backslash\Gamma,V) need not descend to an action of PΓ\text{P}\Gamma. Define

Fun±(𝒢\Γ,V)={F|Id=±FFFun(𝒢\Γ,V)}.\text{Fun}^{\pm}(\mathcal{G}\backslash\Gamma,V)=\big{\{}F\lvert_{-\text{Id}}=\pm F\mid F\in\text{Fun}(\mathcal{G}\backslash\Gamma,V)\big{\}}.

It is clear that Fun(𝒢\Γ,V)=Fun+(𝒢\Γ,V)Fun(𝒢\Γ,V)\text{Fun}(\mathcal{G}\backslash\Gamma,V)=\text{Fun}^{+}(\mathcal{G}\backslash\Gamma,V)\oplus\text{Fun}^{-}(\mathcal{G}\backslash\Gamma,V) as 𝕂[Γ]\mathbb{K}[\Gamma]-modules. There is a natural projection

Z1(Γ,Fun(𝒢\Γ,V))Z1(Γ,Fun+(𝒢\Γ,V));FF+,F+(γ):=F(γ)+F(γ)|Id2.\begin{gathered}Z^{1}\big{(}\Gamma,\text{Fun}(\mathcal{G}\backslash\Gamma,V)\big{)}\to Z^{1}\big{(}\Gamma,\text{Fun}^{+}(\mathcal{G}\backslash\Gamma,V)\big{)};\\ F\to F^{+},\hskip 14.22636ptF^{+}(\gamma):=\frac{F(\gamma)+F(\gamma)\lvert_{-\text{Id}}}{2}.\end{gathered}

Note that (FF+)(γ)=F(Id)(1γ)2(F-F^{+})(\gamma)=\frac{F(-\text{Id})(1-\gamma)}{2} for each γΓ\gamma\in\Gamma. Therefore the natural inclusion

H1(Γ,Fun+(𝒢\Γ,V))H1(Γ,Fun(𝒢\Γ,V))H^{1}\big{(}\Gamma,\text{Fun}^{+}(\mathcal{G}\backslash\Gamma,V)\big{)}\hookrightarrow H^{1}\big{(}\Gamma,\text{Fun}(\mathcal{G}\backslash\Gamma,V)\big{)}

is an isomorphism. The action of Γ\Gamma on Fun+(𝒢\Γ,V)\text{Fun}^{+}(\mathcal{G}\backslash\Gamma,V) descends to a PΓ\text{P}\Gamma-action and one identifies H1(Γ,Fun+(𝒢\Γ,V))H^{1}\big{(}\Gamma,\text{Fun}^{+}(\mathcal{G}\backslash\Gamma,V)\big{)} and H1(PΓ,Fun+(𝒢\Γ,V))H^{1}\big{(}\text{P}\Gamma,\text{Fun}^{+}(\mathcal{G}\backslash\Gamma,V)\big{)} 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 (𝒢,V)(\mathcal{G},V) be a descent module. Then the Shapiro map sh1\emph{sh}^{1} induces an isomorphism H𝖯1(Γ,Fun(𝒢\Γ,V))H𝖯1(𝒢,V)H^{1}_{\mathsf{P}}\big{(}\Gamma,\emph{Fun}(\mathcal{G}\backslash\Gamma,V)\big{)}\cong H^{1}_{\mathsf{P}}(\mathcal{G},V). As a consequence the parabolic subspace H𝖯1(Γ,Fun(𝒢\Γ,V))H^{1}_{\mathsf{P}}\big{(}\Gamma,\emph{Fun}(\mathcal{G}\backslash\Gamma,V)\big{)} is stable under the 𝒢\mathcal{G}-Hecke action.

For a finite index subgroup 𝒢\mathcal{G} the normal core of 𝒢\mathcal{G}, denoted 𝒢nc\mathcal{G}_{\text{nc}}, is

𝒢nc:=γΓγ𝒢γ1.\mathcal{G}_{\text{nc}}:=\bigcap_{\gamma\in\Gamma}\gamma\mathcal{G}\gamma^{-1}.

This subgroup is the largest finite index normal subgroup of Γ\Gamma that is contained in 𝒢\mathcal{G}.

Proof.

Let NN be a positive integer so that TN𝒢ncT^{N}\in\mathcal{G}_{\text{nc}}. Then γTNγ1𝒢\gamma T^{N}\gamma^{-1}\in\mathcal{G} for each γΓ\gamma\in\Gamma. Suppose that FF is a cocycle on Γ\Gamma with values in Fun+(𝒢\Γ,V)\text{Fun}^{+}(\mathcal{G}\backslash\Gamma,V). A straightforward calculation yields

F(γTNγ1)(𝒢Id)F(TN)(𝒢γ)γ1+V(γTNγ11).(γΓ)F(\gamma T^{N}\gamma^{-1})(\mathcal{G}\text{Id})\in F(T^{N})(\mathcal{G}\gamma)\gamma^{-1}+V(\gamma T^{N}\gamma^{-1}-1).\hskip 8.5359pt(\begin{subarray}{c}\gamma\in\Gamma\end{subarray})

Therefore F(TN)(𝒢γ)V(TN1)F(T^{N})(\mathcal{G}\gamma)\in V(T^{N}-1) if and only if F(γTNγ1)(𝒢Id)V(γTNγ11)F(\gamma T^{N}\gamma^{-1})(\mathcal{G}\text{Id})\in V(\gamma T^{N}\gamma^{-1}-1). It follows that

F(TN)Fun+(𝒢\Γ,V)(TN1)\displaystyle F(T^{N})\in\text{Fun}^{+}(\mathcal{G}\backslash\Gamma,V)(T^{N}-1)
F(γTNγ1)(𝒢Id)V(γTNγ11),γΓ.\displaystyle\iff F(\gamma T^{N}\gamma^{-1})(\mathcal{G}\text{Id})\in V(\gamma T^{N}\gamma^{-1}-1),\;\forall\gamma\in\Gamma.

But γTNγ1\langle\langle\gamma T^{N}\gamma^{-1}\rangle\rangle, resp. TN\langle\langle T^{N}\rangle\rangle, is a finite index subgroup of the abelian group 𝒢γ\mathcal{G}_{\gamma\infty}, resp. Γ\Gamma_{\infty}. The equivalence above along with the injectivity of restriction argument and (5.3) implies that [F]H𝖯1(Γ,Fun+(𝒢\Γ,V))[F]\in H^{1}_{\mathsf{P}}\big{(}\Gamma,\text{Fun}^{+}(\mathcal{G}\backslash\Gamma,V)\big{)} if and only if sh1([F])H𝖯1(𝒢,V)\text{sh}^{1}([F])\in H^{1}_{\mathsf{P}}(\mathcal{G},V) where [][\cdot] denotes the image in cohomology. Hence sh1\text{sh}^{1} induces an isomorphism between H𝖯1(Γ,Fun(𝒢\Γ,V))H^{1}_{\mathsf{P}}\big{(}\Gamma,\text{Fun}(\mathcal{G}\backslash\Gamma,V)\big{)} and H𝖯1(𝒢,V)H^{1}_{\mathsf{P}}\big{(}\mathcal{G},V\big{)}. The second part of the assertion is clear from the definition of 𝒢\mathcal{G}-Hecke action. ∎

We next explicitly describe the preimage under sh1\text{sh}^{1} of the cohomology classes attached to modular forms. Write 𝒳Γ=the Γ-span of k(𝒢) in Hol()\mathcal{X}_{\Gamma}=\text{the $\Gamma$-span of $\mathcal{M}_{k}(\mathcal{G})$ in $\text{Hol}(\mathbb{H})$}. Note that 𝒳Γk(𝒢nc)\mathcal{X}_{\Gamma}\subseteq\mathcal{M}_{k}(\mathcal{G}_{\text{nc}}). In particular 𝒳Γ𝔉\mathcal{X}_{\Gamma}\subseteq\mathbb{C}\mathfrak{F}_{\infty} and one can invoke the theory of Eichler-Shimura integrals with base-point at infinity for this collection of functions. Let

ESk,𝒢:k(𝒢)Z1(𝒢,Vk2,)\text{ES}_{k,\mathcal{G}}^{*}:\mathcal{M}_{k}(\mathcal{G})\to Z^{1}(\mathcal{G},V_{k-2,\mathbb{C}})

be the Eichler-Shimura cocycle map attached to (𝒢,k(𝒢))(\mathcal{G},\mathcal{M}_{k}(\mathcal{G})) for the choice of base point at \infty. Define the induced Eichler-Shimura cocycle map by

(7.2) Ind-ESk,𝒢\displaystyle\text{Ind-ES}_{k,\mathcal{G}}^{*} :k(𝒢)Z1(Γ,Fun(𝒢\Γ,Vk2,)),\displaystyle:\mathcal{M}_{k}(\mathcal{G})\to Z^{1}\big{(}\Gamma,\text{Fun}(\mathcal{G}\backslash\Gamma,V_{k-2,\mathbb{C}})\big{)},
Ind-ESk,𝒢(f)(γ)(𝒢σ):=𝗉k(γ,f|σ,X).\displaystyle\text{Ind-ES}_{k,\mathcal{G}}^{*}(f)(\gamma)(\mathcal{G}\sigma):=\mathsf{p}_{k}^{*}(\gamma,f\lvert_{\sigma},X).

Our theory of equivariant period functions for the pair (Γ,𝒳Γ)(\Gamma,\mathcal{X}_{\Gamma}) directly verifies that Ind-ESk,𝒢(f)\text{Ind-ES}_{k,\mathcal{G}}^{*}(f) is indeed a cocycle on Γ\Gamma. Moreover, there is a commutative diagram

(7.3) Z1(Γ,Fun(𝒢\Γ,Vk2,)){Z^{1}\big{(}\Gamma,\text{Fun}(\mathcal{G}\backslash\Gamma,V_{k-2,\mathbb{C}})\big{)}}k(𝒢){\mathcal{M}_{k}(\mathcal{G})}Z1(𝒢,Vk2,){Z^{1}(\mathcal{G},V_{k-2,\mathbb{C}})}ψ\scriptstyle{\psi}Ind-ESk,𝒢\scriptstyle{\text{Ind-ES}_{k,\mathcal{G}}^{*}}ESk,𝒢\scriptstyle{\text{ES}_{k,\mathcal{G}}^{*}}

where ψ\psi is as in (7.1). Note that the cocycle Ind-ESk,𝒢(f)\text{Ind-ES}_{k,\mathcal{G}}^{*}(f) takes values in Fun+(𝒢\Γ,Vk2,)\text{Fun}^{+}(\mathcal{G}\backslash\Gamma,V_{k-2,\mathbb{C}}). The induced Eichler-Shimura homomorphism attached to k(𝒢)\mathcal{M}_{k}(\mathcal{G}) is

Ind-[ESk]m,𝒢:k(𝒢)H1(Γ,Fun(𝒢\Γ,Vk2,));Ind-[ESk]m,𝒢(f):=[Ind-ESk,𝒢(f)].\begin{gathered}\text{Ind-}[\text{ES}_{k}]_{\text{m},\mathcal{G}}:\mathcal{M}_{k}(\mathcal{G})\to H^{1}\big{(}\Gamma,\text{Fun}(\mathcal{G}\backslash\Gamma,V_{k-2,\mathbb{C}})\big{)};\\ \text{Ind-}[\text{ES}_{k}]_{\text{m},\mathcal{G}}(f):=[\text{Ind-}\text{ES}_{k,\mathcal{G}}^{*}(f)].\end{gathered}

We denote the restriction of Ind-[ESk]m,𝒢\text{Ind-}[\text{ES}_{k}]_{\text{m},\mathcal{G}} to 𝒮k(𝒢)\mathcal{S}_{k}(\mathcal{G}) by Ind-[ESk]c,𝒢\text{Ind-}[\text{ES}_{k}]_{\text{c},\mathcal{G}}.

Lemma 7.2.

sh1(Ind-[ESk]m,𝒢(f))=[ESk]m,𝒢(f),fk(𝒢)\emph{sh}^{1}\big{(}\emph{Ind-}[\emph{ES}_{k}]_{\emph{m},\mathcal{G}}(f)\big{)}=[\emph{ES}_{k}]_{\emph{m},\mathcal{G}}(f),\;\forall f\in\mathcal{M}_{k}(\mathcal{G}).

Proof.

The assertion is an easy consequence of (7.3). ∎

Remark 7.3.

In principle, one can start with any choice of Eichler-Shimura integrals for 𝒳Γ\mathcal{X}_{\Gamma} to obtain a lift of the corresponding Eichler-Shimura cocycle map as in (7.2). Lemma 7.2 ensures that all the lifts give rise to the same map at the level of cohomology.

The Γ\Gamma-modules Vk2,V_{k-2,\mathbb{C}} and Fun(𝒢\Γ,Vk2,)\text{Fun}(\mathcal{G}\backslash\Gamma,V_{k-2,\mathbb{C}}) inherit real structures from Vk2,V_{k-2,\mathbb{R}} and ψ\psi in (7.3) is compatible with the associated real structures. As before the real structures descend to the corresponding cohomology modules and sh1\text{sh}^{1} is compatible with complex conjugations in the source and target. Let

Ind-[ESk]c,𝒢a:𝒮ka(𝒢)H1(Γ,Fun(𝒢\Γ,Vk2,));f#Ind-[ESk]c,𝒢(f¯)\text{Ind-}[\text{ES}_{k}]^{a}_{\text{c},\mathcal{G}}:\mathcal{S}^{a}_{k}(\mathcal{G})\to H^{1}\big{(}\Gamma,\text{Fun}(\mathcal{G}\backslash\Gamma,V_{k-2,\mathbb{C}})\big{)};\hskip 5.69046ptf\mapsto\#\text{Ind-}[\text{ES}_{k}]_{\text{c},\mathcal{G}}(\bar{f})

be the induced antiholomorphic Eichler-Shimura map where #\# is the conjugation in cohomology.

Corollary 7.4.

sh1(Ind-[ESk]c,𝒢a(f))=[ESk]c,𝒢a(f);f𝒮ka(𝒢)\emph{sh}^{1}\big{(}\emph{Ind-}[\emph{ES}_{k}]^{a}_{\emph{c},\mathcal{G}}(f)\big{)}=[\emph{ES}_{k}]^{a}_{\emph{c},\mathcal{G}}(f);\;\forall f\in\mathcal{S}^{a}_{k}(\mathcal{G}).

Proof.

Follows from Lemma 7.2 and the discussion above. ∎

Lemma 7.2 and Corollary 7.4 culminate into the isomorphism theorem in the induced picture.

Theorem 7.5.

Let 𝒢\mathcal{G} be a congruence subgroup of Γ\Gamma. The map

Ind-[ESk]m,𝒢Ind-[ESk]c,𝒢a:k(𝒢)𝒮ka(𝒢)H1(Γ,Fun(𝒢\Γ,Vk2,))\emph{Ind-}[\emph{ES}_{k}]_{\emph{m},\mathcal{G}}\oplus\emph{Ind-}[\emph{ES}_{k}]^{a}_{\emph{c},\mathcal{G}}:\mathcal{M}_{k}(\mathcal{G})\oplus\mathcal{S}^{a}_{k}(\mathcal{G})\to H^{1}\big{(}\Gamma,\emph{Fun}(\mathcal{G}\backslash\Gamma,V_{k-2,\mathbb{C}})\big{)}

is a 𝒢\mathcal{G}-Hecke equivariant isomorphism. Moreover, this isomorphism maps 𝒮k(𝒢)𝒮ka(𝒢)\mathcal{S}_{k}(\mathcal{G})\oplus\mathcal{S}^{a}_{k}(\mathcal{G}) onto the parabolic subspace HP1(Γ,Fun(𝒢\Γ,Vk2,))H^{1}_{\emph{P}}\big{(}\Gamma,\emph{Fun}(\mathcal{G}\backslash\Gamma,V_{k-2,\mathbb{C}})\big{)}.

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 kk is odd and Id𝒢-\text{Id}\in\mathcal{G} 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 TT and SS. For fk(𝒢)f\in\mathcal{M}_{k}(\mathcal{G}) define the completed LL-function of ff by

(7.4) L(f,s)=(2π)sΓ(s)L(f,s):=0(f(it)π(f))ts1𝑑t.L^{*}(f,s)=(2\pi)^{-s}\Gamma(s)L(f,s):=\int^{\infty}_{0}\big{(}f(it)-\pi_{\infty}(f)\big{)}t^{s-1}dt.

The integral defining L(f,s)L^{*}(f,s) is absolutely convergent in the region Re(s)>k\text{Re}(s)>k. Moreover, we have the following identity involving entire integrals:

L(f,s)=1(f(it)π(f))ts1𝑑t(Re(s)>k)\displaystyle L^{*}(f,s)=\int_{1}^{\infty}\big{(}f(it)-\pi_{\infty}(f)\big{)}t^{s-1}dt\hskip 56.9055pt(\begin{subarray}{c}\text{Re}(s)>k\end{subarray})
+01(f(it)π(f|S1)(it)k)ts1dt+ikπ(f|S1)skπ(f)s.\displaystyle+\int_{0}^{1}\big{(}f(it)-\pi_{\infty}(f\lvert_{S^{-1}})(it)^{-k}\big{)}t^{s-1}dt+i^{-k}\frac{\pi_{\infty}(f\lvert_{S^{-1}})}{s-k}-\frac{\pi_{\infty}(f)}{s}.

In particular L(f,s)L^{*}(f,s) extends to a meromorphic on the whole complex plane with at most two simple poles supported at {0,k}\{0,k\}. It also satisfies a functional equation L(f,s)=ikL(f|S,ks)L^{\ast}(f,s)=i^{k}L^{\ast}(f\lvert_{S},k-s). The next result provides a complete description of the induced cocycle map (7.2). As before we work with the Γ\Gamma-stable set 𝒳Γ\mathcal{X}_{\Gamma} and consider it as a subset of k(𝒢nc)\mathcal{M}_{k}(\mathcal{G}_{\text{nc}}).

Proposition 7.6.

[14, 8] Let the notation be as above. Then

𝗉k(T,f|σ,X)\displaystyle\mathsf{p}_{k}^{*}(T,f\lvert_{\sigma},X) =π(f|σ)(X+1)k1Xk1k1,\displaystyle=\pi_{\infty}(f\lvert_{\sigma})\frac{(X+1)^{k-1}-X^{k-1}}{k-1},
𝗉k(S,f|σ,X)\displaystyle\mathsf{p}_{k}^{*}(S,f\lvert_{\sigma},X) =r=0k2i1r(k2r)L(f|σ,r+1)Xk2r\displaystyle=\sum_{r=0}^{k-2}i^{1-r}\binom{k-2}{r}L^{*}(f\lvert_{\sigma},r+1)X^{k-2-r}

where fk(𝒢)f\in\mathcal{M}_{k}(\mathcal{G}) and σΓ\sigma\in\Gamma.

Proof.

The first identity is a direct consequence of Lemma 3.13. We appeal to the formula given in Proposition 3.9 to prove the second identity. In our context τ0=i\tau_{0}=i and γ=S\gamma=S. For brevity write F=f|σF=f\lvert_{\sigma}. Here

Fc(S1)|S(τ)=F(τ)τkπ(F|S1).F_{c}(S^{-1})\lvert_{S}(\tau)=F(\tau)-\tau^{-k}\pi_{\infty}(F\lvert_{S^{-1}}).

A simple calculation using the identity below (7.4) yields

iFc(ξ)(Xξ)k2dξi0Fc(S1)|S(ξ)(Xξ)k2dξ\displaystyle\int_{i}^{\infty}F_{c}(\xi)(X-\xi)^{k-2}d\xi-\int_{i}^{0}F_{c}(S^{-1})\lvert_{S}(\xi)(X-\xi)^{k-2}d\xi
=r=0k2i1r(k2r)(L(F,r+1)+ikπ(F|S1)kr1+π(F)r+1)Xk2r.\displaystyle=\sum_{r=0}^{k-2}i^{1-r}\binom{k-2}{r}\big{(}L^{*}(F,r+1)+i^{-k}\frac{\pi_{\infty}(F\lvert_{S^{-1}})}{k-r-1}+\frac{\pi_{\infty}(F)}{r+1}\big{)}X^{k-2-r}.

The proof of the identity is now clear. ∎

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 kk be a positive integer 2\geq 2 and 𝒢\mathcal{G} be a finite index subgroup of Γ\Gamma. We say a character χ:𝒢×\chi:\mathcal{G}\to\mathbb{C}^{\times} is defined over a subfield 𝕂\mathbb{K} of \mathbb{C} if χ(𝒢)𝕂×\chi(\mathcal{G})\subseteq\mathbb{K}^{\times}. Let χ\chi be a character on 𝒢\mathcal{G} that is defined over 𝕂\mathbb{K}. If VV be a right 𝕂[𝒢]\mathbb{K}[\mathcal{G}]-module then VχV^{\chi} denotes a right 𝕂[𝒢]\mathbb{K}[\mathcal{G}]-module whose underlying vector space equals VV and the 𝒢\mathcal{G}-action is given by

v|χ,γ:=χ(γ)1v|γ.v\lvert_{\chi,\gamma}:=\chi(\gamma)^{-1}v\lvert_{\gamma}.

Now suppose χ\chi is an arbitrary character on 𝒢\mathcal{G}. Set

Hol()χ𝒢={fHol()f|k,γ=χ(γ)f,γ𝒢}.\text{Hol}(\mathbb{H})^{\mathcal{G}}_{\chi}=\big{\{}f\in\text{Hol}(\mathbb{H})\mid f\lvert_{k,\gamma}=\chi(\gamma)f,\;\forall\gamma\in\mathcal{G}\big{\}}.

Let 𝒳\mathcal{X} be a 𝒢\mathcal{G}-stable subset of holomorphic functions so that 𝒳Hol()χ𝒢\mathcal{X}\subseteq\text{Hol}(\mathbb{H})^{\mathcal{G}}_{\chi}. Suppose that 𝕀k,𝒳\mathbb{I}_{k,\mathcal{X}} is a choice function for 𝒳\mathcal{X}. With 𝕀k,𝒳\mathbb{I}_{k,\mathcal{X}} one associates a χ\chi-twisted Eichler-Shimura cocycle map given by

ESk,𝒳χ:𝒳Z1(𝒢,Vk2,χ);f(γ𝗉k,𝒳(γ,f,X))\text{ES}_{k,\mathcal{X}}^{\chi}:\mathcal{X}\to Z^{1}\big{(}\mathcal{G},V_{k-2,\mathbb{C}}^{\chi}\big{)};\hskip 8.5359ptf\to\big{(}\gamma\mapsto\mathsf{p}_{k,\mathcal{X}}(\gamma,f,X)\big{)}

where 𝗉k,𝒳\mathsf{p}_{k,\mathcal{X}} is the period function of 𝕀k,𝒳\mathbb{I}_{k,\mathcal{X}}. The general transformation formula (2.10) ensures that the recipe above indeed defines a cocycle with values in Vk2,χV_{k-2,\mathbb{C}}^{\chi}. As a consequence we obtain the χ\chi-twisted Eichler-Shimura homomorphism:

[ESk]𝒢,𝒳χ:𝒳H1(𝒢,Vk2,χ);f[ESk,𝒳χ(f)].[\text{ES}_{k}]_{\mathcal{G},\mathcal{X}}^{\chi}:\mathcal{X}\to H^{1}\big{(}\mathcal{G},V_{k-2,\mathbb{C}}^{\chi}\big{)};\hskip 8.5359ptf\to[\text{ES}_{k,\mathcal{X}}^{\chi}(f)].

Since the period functions attached to two distinct choice functions differ by a coboundary in Fun(𝒳,Vk2,)\text{Fun}(\mathcal{X},V_{k-2,\mathbb{C}}) a straightforward calculation verifies that [ESk]𝒢,𝒳χ[\text{ES}_{k}]_{\mathcal{G},\mathcal{X}}^{\chi} depends only on (𝒢,𝒳)(\mathcal{G},\mathcal{X}) and is independent of 𝕀k,𝒳\mathbb{I}_{k,\mathcal{X}}.

Let NN be a positive integer and consider Γ1(N)Γ0(N)\Gamma_{1}(N)\trianglelefteq\Gamma_{0}(N). Recall that the natural map

Γ0(N)(/N)×,(abcd)d(mod N)\Gamma_{0}(N)\to(\mathbb{Z}/N\mathbb{Z})^{\times},\hskip 8.5359pt\begin{pmatrix}a&b\\ c&d\end{pmatrix}\to d(\text{mod }N)

induces an isomorphism between Γ0(N)/Γ1(N)\Gamma_{0}(N)/\Gamma_{1}(N) and (/N)×(\mathbb{Z}/N\mathbb{Z})^{\times}. We use this isomorphism to identify each character on (/N)×(\mathbb{Z}/N\mathbb{Z})^{\times} with a character on Γ0(N)\Gamma_{0}(N). For simplicity write U^N:=Hom((/N)×,×)\widehat{U}_{N}:=\text{Hom}\big{(}(\mathbb{Z}/N\mathbb{Z})^{\times},\mathbb{C}^{\times}\big{)}. Our treatment of twisted Eichler-Shimura theory relies on the formalism developed in the paragraph above for the pair (Γ0(N),χ)\big{(}\Gamma_{0}(N),\chi\big{)} where χU^N\chi\in\widehat{U}_{N}. In this setting there is a notion of twisted Hecke action on the functions and the cohomology module [16, 3.5]. Put

Δ(N)={γM2()det(γ)>0,γ(a0)(mod N) with gcd(a,N)=1}\Delta(N)=\big{\{}\gamma\in M_{2}(\mathbb{Z})\mid\begin{subarray}{c}\text{det}(\gamma)>0,\hskip 2.84544pt\gamma\equiv\begin{pmatrix}a&\ast\\ 0&\ast\end{pmatrix}(\text{mod $N)$ with }\gcd(a,N)=1\end{subarray}\big{\}}

where M2()M_{2}(\cdot) refers to the ring of 2×22\times 2 matrices. Let χ\chi be a character on (/N)×(\mathbb{Z}/N\mathbb{Z})^{\times} defined over 𝕂\mathbb{K}. With χ\chi we associate a homomorphism of monoids χι:Δ(N)𝕂×\prescript{\iota}{}{\chi}:\Delta(N)\to\mathbb{\mathbb{K}}^{\times} by setting χι(γ)=χ(a)\prescript{\iota}{}{\chi}(\gamma)=\chi(a). If γΓ0(N)\gamma\in\Gamma_{0}(N) then χι(γ)=χ(γ)1\prescript{\iota}{}{\chi}(\gamma)=\chi(\gamma)^{-1}. Let αΔ(N)\alpha\in\Delta(N). Write Γ0(N)αΓ0(N)=j=1nΓ0(N)αj\Gamma_{0}(N)\alpha\Gamma_{0}(N)=\amalg_{j=1}^{n}\Gamma_{0}(N)\alpha_{j} as in Section 4.2. Define

[Γ0(N)αΓ0(N)]k,χ:Hol()χΓ0(N)Hol()χΓ0(N);fj=1nχι(αj)f|k,αj.[\Gamma_{0}(N)\alpha\Gamma_{0}(N)]_{k,\chi}:\text{Hol}(\mathbb{H})^{\Gamma_{0}(N)}_{\chi}\to\text{Hol}(\mathbb{H})^{\Gamma_{0}(N)}_{\chi};\hskip 5.69046ptf\to\sum_{j=1}^{n}\prescript{\iota}{}{\chi}(\alpha_{j})f\lvert_{k,\alpha_{j}}.

The double coset operator also acts on the cocycle model of cohomology. Let VV be a right 𝕂[GL2+()]\mathbb{K}[\text{GL}_{2}^{+}(\mathbb{Q})]-module. Set

:V,χFun(Γ0(N),Vχ)Fun(Γ0(N),Vχ),P(γj=1nχι(αj(γ))P(γj)αj(γ))\begin{gathered}{}_{V,\chi}:\text{Fun}\big{(}\Gamma_{0}(N),V^{\chi}\big{)}\to\text{Fun}\big{(}\Gamma_{0}(N),V^{\chi}\big{)},\\ P\to\big{(}\gamma\mapsto\sum_{j=1}^{n}\prescript{\iota}{}{\chi}(\alpha_{j(\gamma)})P(\gamma_{j})\alpha_{j(\gamma)}\big{)}\end{gathered}

where γj\gamma_{j}-s are as in (4.3) and αj(γ)\alpha_{j(\gamma)} acts on P(γj)P(\gamma_{j}) by the GL2+()\text{GL}_{2}^{+}(\mathbb{Q})-action on VV. This map descends to an endomorphism of H1(Γ0(N),Vχ)H^{1}\big{(}\Gamma_{0}(N),V^{\chi}\big{)} that depends only on the double coset [Γ0(N)αΓ0(N)]V,χ[\Gamma_{0}(N)\alpha\Gamma_{0}(N)]_{V,\chi}. 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 IdΓ0(N)-\text{Id}\in\Gamma_{0}(N). Therefore (Γ0(N),Vχ)(\Gamma_{0}(N),V^{\chi}) is a descent module if and only if Id-\text{Id} acts on VV by the scalar χ(1)\chi(-1). In particular (Γ0(N),Vk2,χ)(\Gamma_{0}(N),V_{k-2,\mathbb{C}}^{\chi}) is a descent module if and only if χ(1)=(1)k\chi(-1)=(-1)^{k}.

We next relate the χ\chi-eigenspace in cohomology with the cohomology of the χ\chi-twisted module using the conjugation action of Γ0(N)\Gamma_{0}(N) on the cohomology of Γ1(N)\Gamma_{1}(N). This result allows us to bypass the use of trace maps in Hida’s theory [10, p.177]. Let 𝕂\mathbb{K} be a subfield of \mathbb{C} and χU^N\chi\in\widehat{U}_{N} be a character defined over 𝕂\mathbb{K}. Suppose that VV is a 𝕂[Γ0(N)]\mathbb{K}[\Gamma_{0}(N)]-module. Since Γ1(N)\Gamma_{1}(N) is a normal subgroup of Γ0(N)\Gamma_{0}(N) the canonical restriction map induces an isomorphism

(7.5) H1(Γ0(N),Vχ)H1(Γ1(N),Vχ)Γ0(N).H^{1}\big{(}\Gamma_{0}(N),V^{\chi}\big{)}\xrightarrow{\cong}H^{1}\big{(}\Gamma_{1}(N),V^{\chi}\big{)}^{\Gamma_{0}(N)}.
Lemma 7.7.

Let H1(Γ1(N),V)[χ]H^{1}(\Gamma_{1}(N),V)[\chi] denote the χ\chi eigenspace for the Γ0(N)\Gamma_{0}(N)-action on H1(Γ1(N),V)H^{1}(\Gamma_{1}(N),V). Then

  1. (i)

    H1(Γ1(N),Vχ)Γ0(N)=H1(Γ1(N),V)[χ]H^{1}\big{(}\Gamma_{1}(N),V^{\chi}\big{)}^{\Gamma_{0}(N)}=H^{1}\big{(}\Gamma_{1}(N),V\big{)}[\chi].

  2. (ii)

    Suppose that (Γ0(N),Vχ)(\Gamma_{0}(N),V^{\chi}) is a descent module. Then the image of the parabolic subspace HP1(Γ0(N),Vχ)H^{1}_{\emph{P}}(\Gamma_{0}(N),V^{\chi}) under (7.5) equals HP1(Γ1(N),V)[χ]H^{1}_{\emph{P}}\big{(}\Gamma_{1}(N),V\big{)}[\chi].

Proof.

The assertion in part (i) follows from the explicit description of the action provided in (4.5). To prove part (ii) we identify the cohomology modules with their projective counterpart and apply Lemma 5.1 to the restriction map described in (7.5). ∎

The action of diamond operators on the spaces of modular forms on Γ1(N)\Gamma_{1}(N) yields decompositions of the form

k(Γ1(N))=χU^Nk(N,χ),𝒮k(Γ1(N))=χU^N𝒮k(N,χ)\mathcal{M}_{k}\big{(}\Gamma_{1}(N)\big{)}=\oplus_{\chi\in\widehat{U}_{N}}\mathcal{M}_{k}(N,\chi),\hskip 8.5359pt\mathcal{S}_{k}\big{(}\Gamma_{1}(N)\big{)}=\oplus_{\chi\in\widehat{U}_{N}}\mathcal{S}_{k}(N,\chi)

where the sum is over all the Dirichlet character modulo NN. The space k(N,χ)\mathcal{M}_{k}(N,\chi), resp. 𝒮k(N,χ)\mathcal{S}_{k}(N,\chi), is zero unless χ(1)=(1)k\chi(-1)=(-1)^{k}. Attached to these spaces we have Hecke-equivariant twisted Eichler-Shimura homomorphisms

χm,N\displaystyle{}_{\text{m},N}^{\chi} :k(N,χ)H1(Γ0(N),Vk2,χ),\displaystyle:\mathcal{M}_{k}(N,\chi)\to H^{1}\big{(}\Gamma_{0}(N),V_{k-2,\mathbb{C}}^{\chi}\big{)},
[ESk]c,Nχ\displaystyle[\text{ES}_{k}]_{\text{c},N}^{\chi} :𝒮k(N,χ)H1(Γ0(N),Vk2,χ)\displaystyle:\mathcal{S}_{k}(N,\chi)\to H^{1}\big{(}\Gamma_{0}(N),V_{k-2,\mathbb{C}}^{\chi}\big{)}

for each χ\chi. The action of diamond operators on the space of anti-holomorphic cusp forms yields a similar eigenspace decomposition

𝒮ka(Γ1(N))=χU^N𝒮ka(N,χ).\mathcal{S}^{a}_{k}\big{(}\Gamma_{1}(N)\big{)}=\oplus_{\chi\in\widehat{U}_{N}}\mathcal{S}^{a}_{k}(N,\chi).

It is clear that 𝒮ka(N,χ)=𝒮k(N,χ¯)¯\mathcal{S}^{a}_{k}(N,\chi)=\overline{\mathcal{S}_{k}(N,\bar{\chi})}. We transport the χ¯\bar{\chi}-Hecke action on 𝒮k(N,χ¯)\mathcal{S}_{k}(N,\bar{\chi}) via complex-conjugation to obtain a χ\chi-Hecke action on 𝒮ka(N,χ)\mathcal{S}^{a}_{k}(N,\chi). This procedure yields a well-defined χ\chi-Hecke equivariant homomorphism

[ESk]c,Na,χ:𝒮ka(N,χ)H1(Γ0(N),Vk2,χ);f#[ESk]c,Nχ¯(f¯)[\text{ES}_{k}]_{\text{c},N}^{a,\chi}:\mathcal{S}^{a}_{k}(N,\chi)\to H^{1}\big{(}\Gamma_{0}(N),V_{k-2,\mathbb{C}}^{\chi}\big{)};\hskip 5.69046ptf\to\#[\text{ES}_{k}]_{\text{c},N}^{\bar{\chi}}(\bar{f})

where #\# is the complex conjugation defined in Section 5.2.

Theorem 7.8.

Let NN be an integer 1\geq 1 and χ\chi be a character on (/N)×(\mathbb{Z}/N\mathbb{Z})^{\times}. The χ\chi-Hecke-equivariant homomorphism

(7.6) [ESk]m,Nχ[ESk]c,Na,χ:k(N,χ)𝒮ka(N,χ)H1(Γ0(N),Vk2,χ)[\emph{ES}_{k}]_{\emph{m},N}^{\chi}\oplus[\emph{ES}_{k}]_{\emph{c},N}^{a,\chi}:\mathcal{M}_{k}(N,\chi)\oplus\mathcal{S}^{a}_{k}(N,\chi)\to H^{1}\big{(}\Gamma_{0}(N),V_{k-2,\mathbb{C}}^{\chi}\big{)}

is an isomorphism. Moreover the image of 𝒮k(N,χ)𝒮ka(N,χ)\mathcal{S}_{k}(N,\chi)\oplus\mathcal{S}^{a}_{k}(N,\chi) under this map equals HP1(Γ0(N),Vk2,χ)H^{1}_{\emph{P}}\big{(}\Gamma_{0}(N),V_{k-2,\mathbb{C}}^{\chi}\big{)}.

Proof.

We start with the untwisted Eichler-Shimura isomorphism

(7.7) k(Γ1(N))𝒮ka(Γ1(N))H1(Γ1(N),Vk2,).\mathcal{M}_{k}\big{(}\Gamma_{1}(N)\big{)}\oplus\mathcal{S}^{a}_{k}\big{(}\Gamma_{1}(N)\big{)}\xrightarrow{\cong}H^{1}\big{(}\Gamma_{1}(N),V_{k-2,\mathbb{C}}\big{)}.

One considers the χ\chi-eigenspaces for the action of diamond operators to discover that the isomorphism (7.7) maps k(N,χ)𝒮ka(N,χ)\mathcal{M}_{k}(N,\chi)\oplus\mathcal{S}^{a}_{k}(N,\chi) isomorphically onto H1(Γ1(N),Vk2,)[χ]H^{1}\big{(}\Gamma_{1}(N),V_{k-2,\mathbb{C}}\big{)}[\chi]. Now the Eichler-Shimura homomorphism and its twisted avatar together yield a commutative diagram

(7.8) H1(Γ0(N),Vk2,χ){H^{1}\big{(}\Gamma_{0}(N),V_{k-2,\mathbb{C}}^{\chi}\big{)}}k(N,χ)𝒮ka(N,χ){\mathcal{M}_{k}(N,\chi)\oplus\mathcal{S}^{a}_{k}(N,\chi)}H1(Γ1(N),Vk2,){H^{1}\big{(}\Gamma_{1}(N),V_{k-2,\mathbb{C}}\big{)}}(7.7)(7.6)

where the vertical arrow is the restriction map. Note that the vertical arrow is injective and its image equals H1(Γ1(N),Vk2,)[χ]H^{1}\big{(}\Gamma_{1}(N),V_{k-2,\mathbb{C}}\big{)}[\chi] (Lemma 7.7). Therefore the twisted homomorphism (7.6) is an isomorphism. If χ(1)=(1)k+1\chi(-1)=(-1)^{k+1} then the domain and codomain of (7.6) are zero (Example 4.5). Now suppose χ(1)=(1)k\chi(-1)=(-1)^{k}. Observe that the image of 𝒮k(N,χ)𝒮ka(N,χ)\mathcal{S}_{k}(N,\chi)\oplus\mathcal{S}^{a}_{k}(N,\chi) under (7.7) is HP1(Γ1(N),Vk2,)[χ]H^{1}_{\text{P}}\big{(}\Gamma_{1}(N),V_{k-2,\mathbb{C}}\big{)}[\chi]. Hence, by Lemma 7.7(ii), the isomorphism (7.6) must map 𝒮k(N,χ)𝒮ka(N,χ)\mathcal{S}_{k}(N,\chi)\oplus\mathcal{S}^{a}_{k}(N,\chi) onto HP1(Γ0(N),Vk2,χ)H^{1}_{\text{P}}\big{(}\Gamma_{0}(N),V_{k-2,\mathbb{C}}^{\chi}\big{)}. ∎

The space of Eisenstein series with nebentypus χ\chi is given by

(7.9) k(N,χ):=k(Γ1(N))k(N,χ).\mathcal{E}_{k}(N,\chi):=\mathcal{E}_{k}\big{(}\Gamma_{1}(N)\big{)}\cap\mathcal{M}_{k}(N,\chi).

If χ\chi is trivial then k(N,χ)=k(Γ0(N))\mathcal{E}_{k}(N,\chi)=\mathcal{E}_{k}\big{(}\Gamma_{0}(N)\big{)}. Without loss of generality assume that χ\chi is a nontrivial character with χ(1)=(1)k\chi(-1)=(-1)^{k}. In this setting, the author has recently constructed an explicit basis k,χ\mathcal{B}_{k,\chi} that is parametrized by the cusps of Γ0(N)\Gamma_{0}(N) and established a weaker version of Theorem 7.8 [15, 7.3]. By construction the basis functions in k,χ\mathcal{B}_{k,\chi} lie inside k(Γ1(N),𝕂)\mathcal{E}_{k}\big{(}\Gamma_{1}(N),\mathbb{K}\big{)} if φ(N)𝕂\mathbb{Q}_{\varphi(N)}\subseteq\mathbb{K}. Here φ\varphi is Euler’s totient function and φ(N)\mathbb{Q}_{\varphi(N)} is the φ(N)\varphi(N)-th cyclotomic extension of \mathbb{Q} in \mathbb{C}. Recall the spectral basis 𝒢,k\mathcal{B}_{\mathcal{G},k} from Section 5.3. For an arbitrary character χ\chi set

(7.10) k,χ={,if χ(1)=(1)k+1;Γ0(N),k,if k is even and χ is trivial;k,χ above,if χ is nontrivial and χ(1)=(1)k.\mathcal{B}_{k,\chi}=\begin{cases}\emptyset,&\text{if $\chi(-1)=(-1)^{k+1}$};\\ \mathcal{B}_{\Gamma_{0}(N),k},&\text{if $k$ is even and $\chi$ is trivial};\\ \text{$\mathcal{B}_{k,\chi}$ above},&\text{if $\chi$ is nontrivial and $\chi(-1)=(-1)^{k}$.}\end{cases}

The compatibility with restriction property described in (5.7) shows that k(Γ0(N),𝕂)k(Γ1(N),𝕂)\mathcal{E}_{k}(\Gamma_{0}(N),\mathbb{K})\subseteq\mathcal{E}_{k}(\Gamma_{1}(N),\mathbb{K}) for each 𝕂\mathbb{K}. Therefore k,χk(Γ1(N),𝕂)\mathcal{B}_{k,\chi}\subseteq\mathcal{E}_{k}\big{(}\Gamma_{1}(N),\mathbb{K}\big{)} for all χ\chi whenever φ(N)𝕂\mathbb{Q}_{\varphi(N)}\subseteq\mathbb{K}. We study rationality of the image of this basis under [ESk]m,Nχ[\text{ES}_{k}]_{\text{m},N}^{\chi} 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 NN-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 Γ(N)\Gamma(N).

Let kk be an integer 2\geq 2 and Λ=2\Lambda=\mathbb{Z}^{2}. With each λ¯Λ/NΛ\bar{\lambda}\in\Lambda/N\Lambda one associates an Eisenstein series described as

Gk(τ,λ¯,N;s):=(c,d)λ(N),(c,d)(0,0)Im(τ)s(cτ+d)k|cτ+d|2s(τ)G_{k}(\tau,\bar{\lambda},N;s):=\sum_{\begin{subarray}{c}(c,d)\equiv\lambda(N),\\ (c,d)\neq(0,0)\end{subarray}}\frac{\text{Im}(\tau)^{s}}{(c\tau+d)^{k}\lvert c\tau+d\rvert^{2s}}\hskip 8.5359pt(\begin{subarray}{c}\tau\in\mathbb{H}\end{subarray})

where λ\lambda is lift of λ¯\bar{\lambda} and Im(τ)\text{Im}(\tau) is the imaginary part of τ\tau. This series converges uniformly on the compact subsets of Re(k+2s)>2\text{Re}(k+2s)>2 to define an analytic function of ss in this region that admits an analytic continuation to the whole ss plane. Set Gk(τ,λ¯,N):=Gk(τ,λ¯,N;0)G_{k}(\tau,\bar{\lambda},N):=G_{k}(\tau,\bar{\lambda},N;0). As a function on \mathbb{H} the Fourier expansion of the GG-series is Gk(τ,λ¯,N)=j0Ak,j(λ¯)𝐞(jτN)+Pk(τ)G_{k}(\tau,\bar{\lambda},N)=\sum_{j\geq 0}A_{k,j}(\bar{\lambda})\mathbf{e}(\frac{j\tau}{N})+P_{k}(\tau) where

(8.1) Ak,0(λ¯):=𝟙/N(λ¯1,0)n{0},nλ2(N)1nk,Pk(τ):=𝟙(k,2)πN2Im(τ)Ak,j(λ¯):=(2πi)k(k1)!Nkr{0},rj,jrλ1(N)sgn(r)rk1𝐞(rλ2N)(j1)\begin{gathered}A_{k,0}(\bar{\lambda}):=\mathbbm{1}_{\mathbb{Z}/N\mathbb{Z}}(\bar{\lambda}_{1},0)\sum_{\begin{subarray}{c}n\in\mathbb{Z}-\{0\},\\ n\equiv\lambda_{2}(N)\end{subarray}}\frac{1}{n^{k}},\;P_{k}(\tau):=-\mathbbm{1}_{\mathbb{Z}}(k,2)\frac{\pi}{N^{2}\text{Im}(\tau)}\\ A_{k,j}(\bar{\lambda}):=\frac{(-2\pi i)^{k}}{(k-1)!N^{k}}\sum_{\begin{subarray}{c}r\in\mathbb{Z}-\{0\},\\ r\mid j,\frac{j}{r}\equiv\lambda_{1}(N)\end{subarray}}\text{sgn}(r)r^{k-1}\mathbf{e}(\frac{r\lambda_{2}}{N})\hskip 11.38092pt(\begin{subarray}{c}j\geq 1\end{subarray})\end{gathered}

and (λ1,λ2)(\lambda_{1},\lambda_{2}) is a lift of λ¯\bar{\lambda} to Λ\Lambda [17, 9.1]. Recall the rational structure on the space of Eisenstein series from Section 5.3 arising from the spectral basis. Let N\mathbb{Q}_{N} denote the NN-th cyclotomic field (𝐞(1N))\mathbb{Q}\big{(}\mathbf{e}(\frac{1}{N})\big{)}. Suppose that 𝕂\mathbb{K} is a subfield of \mathbb{C} containing N\mathbb{Q}_{N}. Then the 𝕂\mathbb{K}-span of {Gk(λ¯,N)λ¯Λ/NΛ}\{G_{k}(\bar{\lambda},N)\mid\bar{\lambda}\in\Lambda/N\Lambda\} equals (2πi)kk(Γ(N),𝕂)(2\pi i)^{k}\mathcal{E}_{k}^{\ast}\big{(}\Gamma(N),\mathbb{K}\big{)} and its intersection with Hol()\text{Hol}(\mathbb{H}) is (2πi)kk(Γ(N),𝕂)(2\pi i)^{k}\mathcal{E}_{k}\big{(}\Gamma(N),\mathbb{K}\big{)} [15, 6.2]. The explicit Fourier expansion makes GG-series accessible for the computation of LL-function and LL-values. A standard approach towards this problem uses a cyclotomic linear combination of GG-series to write down a Hecke eigenfunction on Γ(N)\Gamma(N) whose LL-function is given by a product of two Dirichlet LL-functions [13, IV-39]. However, it is difficult to study the rationality properties of special values of such product LL-functions.

One defines the Eisenstein series on the universal elliptic curve as

ek:×,ek(η,τ)=wΛτ{0}χη(w)wk(k3)e_{k}:\mathbb{C}\times\mathbb{H}\to\mathbb{C},\hskip 8.5359pte_{k}(\eta,\tau)=\sum_{w\in\Lambda_{\tau}-\{0\}}\frac{\chi_{\eta}(w)}{w^{k}}\hskip 14.22636pt(\begin{subarray}{c}k\geq 3\end{subarray})

where Λτ={(cτ+d)(c,d)Λ}\Lambda_{\tau}=\{(c\tau+d)\mid(c,d)\in\Lambda\} and χη(w)=𝐞(wη¯w¯ηττ¯)\chi_{\eta}(w)=\mathbf{e}(\frac{w\bar{\eta}-\bar{w}\eta}{\tau-\bar{\tau}}) [11, 4]. This series converges uniformly on compact subsets to define a real analytic function that transforms like a Jacobi form of weight kk and index 0. In particular ek(η+cτ+d,τ)=ek(η,τ)e_{k}(\eta+c\tau+d,\tau)=e_{k}(\eta,\tau) for each (c,d)Λ(c,d)\in\Lambda. With λ¯Λ/NΛ\bar{\lambda}\in\Lambda/N\Lambda one associates the division value

ek(τ,λ¯,N):=ek(λ1τN+λ2N,τ)=(c,d)Λ,(c,d)(0,0)𝐞(cλ2dλ1N)(cτ+d)k;λ¯=(λ1,λ2)¯.e_{k}(\tau,\bar{\lambda},N):=e_{k}(\frac{\lambda_{1}\tau}{N}+\frac{\lambda_{2}}{N},\tau)=\sum_{\begin{subarray}{c}(c,d)\in\Lambda,\\ (c,d)\neq(0,0)\end{subarray}}\frac{\mathbf{e}(\frac{c\lambda_{2}-d\lambda_{1}}{N})}{(c\tau+d)^{k}};\hskip 8.5359pt\bar{\lambda}=\overline{(\lambda_{1},\lambda_{2})}.

Let 𝐛:Λ/NΛ×Λ/NΛ\mathbf{b}:\Lambda/N\Lambda\times\Lambda/N\Lambda\to\mathbb{C} be the nondegenerate bilinear form defined by

𝐛(λ¯,θ¯)=𝐞(θ1λ2λ1θ2N);λ¯=(λ1,λ2)¯ and θ¯=(θ1,θ2)¯.\mathbf{b}(\bar{\lambda},\bar{\theta})=\mathbf{e}(\frac{\theta_{1}\lambda_{2}-\lambda_{1}\theta_{2}}{N});\hskip 8.5359pt\text{$\bar{\lambda}=\overline{(\lambda_{1},\lambda_{2})}$ and $\bar{\theta}=\overline{(\theta_{1},\theta_{2})}$}.

Then the relation between ee-series and GG-series is as follows:

(8.2) ek(λ¯,N)=θ¯Λ/NΛ𝐛(λ¯,θ¯)Gk(θ¯,N).e_{k}(\bar{\lambda},N)=\sum_{\bar{\theta}\in\Lambda/N\Lambda}\mathbf{b}(\bar{\lambda},\bar{\theta})G_{k}(\bar{\theta},N).

Let kk be an integer 2\geq 2 and λ¯Λ/NΛ\bar{\lambda}\in\Lambda/N\Lambda. We define the elliptic Eisenstein series of weight kk and parameter λ¯\bar{\lambda}, denoted ek(λ¯,N)e_{k}(\bar{\lambda},N), using the identity (8.2). Note that 𝐛\mathbf{b} is compatible with the natural right action of Γ\Gamma on Λ\Lambda, i.e., 𝐛(λ¯γ,θ¯γ)=𝐛(λ¯,θ¯)\mathbf{b}(\bar{\lambda}\gamma,\bar{\theta}\gamma)=\mathbf{b}(\bar{\lambda},\bar{\theta}) for each γΓ\gamma\in\Gamma. Thus ek(λ¯,N)|γ=ek(λ¯γ,N)e_{k}(\bar{\lambda},N)\lvert_{\gamma}=e_{k}(\bar{\lambda}\gamma,N). In particular ek(λ¯,N)e_{k}(\bar{\lambda},N) is invariant under Γ(N)\Gamma(N). One uses the orthogonality of characters for Λ/NΛ\Lambda/N\Lambda to invert the linear relation (8.2) and discovers that

Gk(θ¯,N)=1N2λ¯Λ/NΛ𝐛(θ¯,λ¯)ek(λ¯,N).G_{k}(\bar{\theta},N)=\frac{1}{N^{2}}\sum_{\bar{\lambda}\in\Lambda/N\Lambda}\mathbf{b}(\bar{\theta},\bar{\lambda})e_{k}(\bar{\lambda},N).

Therefore the 𝕂\mathbb{K}-span of {ek(λ¯,N)λ¯Λ/NΛ}\{e_{k}(\bar{\lambda},N)\mid\bar{\lambda}\in\Lambda/N\Lambda\} equals the 𝕂\mathbb{K}-span of {Gk(λ¯,N)λ¯Λ/NΛ}\{G_{k}(\bar{\lambda},N)\mid\bar{\lambda}\in\Lambda/N\Lambda\} whenever N𝕂\mathbb{Q}_{N}\subseteq\mathbb{K}.

We next record the Fourier expansion formula for the elliptic Eisenstein series. Let B()B_{\bullet}(-) be the Bernoulli polynomial defined by the generating function identity

Xexp(tX)exp(X)1=n0Bn(t)n!Xn.\frac{X\exp(tX)}{\exp(X)-1}=\sum_{n\geq 0}\frac{B_{n}(t)}{n!}X^{n}.

Our calculation requires the Fourier expansion formula for the Bernoulli polynomials [17, 4.5]:

(8.3) Bk(t)=k!(2πi)kn{0}𝐞(nt)nk.(k2, 0t1)B_{k}(t)=-\frac{k!}{(2\pi i)^{k}}\sum_{n\in\mathbb{Z}-\{0\}}\frac{\mathbf{e}(nt)}{n^{k}}.\hskip 14.22636pt(\begin{subarray}{c}k\geq 2,\;0\leq t\leq 1\end{subarray})

The formula is valid even for k=1k=1 if 0<t<10<t<1 and one interprets the sum as limm0<|n|m\lim_{m\to\infty}\sum_{0<\lvert n\rvert\leq m}.

Lemma 8.1.

Let kk be an integer 2\geq 2 and λ¯Λ/NΛ\bar{\lambda}\in\Lambda/N\Lambda. Suppose that (λ1,λ2)(\lambda_{1},\lambda_{2}) is a lift of λ¯\bar{\lambda} that satisfies 0λ1<N0\leq\lambda_{1}<N. The Fourier series for ek(λ¯,N)e_{k}(\bar{\lambda},N) is

ek(λ¯,N)\displaystyle e_{k}(\bar{\lambda},N) =(2πi)kk!Bk(λ1N)𝟙(k,2)𝟙Λ/NΛ(λ¯,0)πIm(τ)\displaystyle=-\frac{(-2\pi i)^{k}}{k!}B_{k}(\frac{\lambda_{1}}{N})-\mathbbm{1}_{\mathbb{Z}}(k,2)\mathbbm{1}_{\Lambda/N\Lambda}(\bar{\lambda},0)\frac{\pi}{\emph{Im}(\tau)}
+(2πi)k(k1)!Nk1m,n1,nλ1(N)nk1𝐞(mnτN+mλ2N)\displaystyle+\frac{(-2\pi i)^{k}}{(k-1)!N^{k-1}}\sum_{\begin{subarray}{c}m,n\geq 1,\\ n\equiv\lambda_{1}(N)\end{subarray}}n^{k-1}\mathbf{e}(\frac{mn\tau}{N}+\frac{m\lambda_{2}}{N})
+(2πi)k(k1)!Nk1m,n1,nλ1(N)nk1𝐞(mnτNmλ2N).\displaystyle+\frac{(2\pi i)^{k}}{(k-1)!N^{k-1}}\sum_{\begin{subarray}{c}m,n\geq 1,\\ n\equiv-\lambda_{1}(N)\end{subarray}}n^{k-1}\mathbf{e}(\frac{mn\tau}{N}-\frac{m\lambda_{2}}{N}).
Proof.

We substitute the Fourier expansion formula for the GG-series (8.1) into the defining relation (8.2) and simplify the expression to obtain the identity above. The constant term is a consequence of (8.3) while the real analytic term follows from a straightforward calculation. ∎

Corollary 8.2.

Let λ¯Λ/NΛ\bar{\lambda}\in\Lambda/N\Lambda. Then e2(λ¯,N)e_{2}(\bar{\lambda},N) is holomorphic if and only if λ¯0\bar{\lambda}\neq 0. Moreover {e2(λ¯,N)λ¯0}\{e_{2}(\bar{\lambda},N)\mid\bar{\lambda}\neq 0\} is a spanning set for the holomorphic subspace (2πi)22(Γ(N),𝕂)(2\pi i)^{2}\mathcal{E}_{2}\big{(}\Gamma(N),\mathbb{K}\big{)} whenever N𝕂\mathbb{Q}_{N}\subseteq\mathbb{K}.

Proof.

The first part of the assertion is obvious. The extended space of GG-series (2πi)22(Γ(N),𝕂)(2\pi i)^{2}\mathcal{E}^{\ast}_{2}\big{(}\Gamma(N),\mathbb{K}\big{)} contains (2πi)22(Γ(N),𝕂)(2\pi i)^{2}\mathcal{E}_{2}\big{(}\Gamma(N),\mathbb{K}\big{)} as a subspace of codimension 11. But {e2(λ¯,N)λ¯0}\{e_{2}(\bar{\lambda},N)\mid\bar{\lambda}\neq 0\} is already a subset of holomorphic subspace. Therefore this collection spans (2πi)22(Γ(N),𝕂)(2\pi i)^{2}\mathcal{E}_{2}\big{(}\Gamma(N),\mathbb{K}\big{)}. ∎

For convenience, we introduce the notation

N,k:={Λ/NΛ,k even and 4;Λ/NΛ,k odd and N3;Λ/NΛ{0},k=2;,otherwise.\mathcal{I}_{N,k}:=\begin{cases}\Lambda/N\Lambda,&\text{$k$ even and $\geq 4$;}\\ \Lambda/N\Lambda,&\text{$k$ odd and $N\geq 3$;}\\ \Lambda/N\Lambda-\{0\},&k=2;\\ \emptyset,&\text{otherwise.}\end{cases}

8.1.2 LL-function and LL-values of the elliptic Eisenstein series

Our next job is explicitly writing down the completed LL-function (7.4) attached to the elliptic Eisenstein series and determining its values at {1,,k1}\{1,\ldots,k-1\}. Let xx\in\mathbb{R} and 0<a10<a\leq 1. One defines the Lerch zeta function [1] by

(8.4) ϕ(x,a,s)=n0𝐞(nx)(n+a)s.\phi(x,a,s)=\sum_{n\geq 0}\frac{\mathbf{e}(nx)}{(n+a)^{s}}.

The series above converges uniformly on the compact subsets of {sRe(s)>1}\{s\mid\text{Re}(s)>1\} to define an analytic function in this region. The Lerch series admits a meromorphic continuation to the whole ss-plane with a simple pole at s=1s=1 whenever xx is an integer. If xx is not an integer then ϕ(x,a,)\phi(x,a,\cdot) extends to an entire function on the ss-plane. Our treatment requires two special avatars of the Lerch function. The Lerch series reduces to the familiar Hurwitz zeta function ζ(a,s)\zeta(a,s) for x=0x=0, i.e., ζ(a,s)=ϕ(0,a,s)\zeta(a,s)=\phi(0,a,s). The value of the Hurwitz zeta function at nonnegative integers [2, 12.11] is given by

(8.5) ζ(a,n)=Bn+1(a)n+1.(0<a1,n0)\zeta(a,-n)=-\frac{B_{n+1}(a)}{n+1}.\hskip 8.5359pt(\begin{subarray}{c}0<a\leq 1,\;n\geq 0\end{subarray})

On the other hand setting a=1a=1 in (8.4) yields the polylogarithm function given by the convergent series

Lis(𝐞(x)):=n1𝐞(nx)ns=𝐞(x)ϕ(x,1,s)\text{Li}_{s}\big{(}\mathbf{e}(x)\big{)}:=\sum_{n\geq 1}\frac{\mathbf{e}(nx)}{n^{s}}=\mathbf{e}(x)\phi(x,1,s)

in the region Re(s)>1\text{Re}(s)>1. The meromorphic continuation of the Lerch function provides a meromorphic continuation for Li. If xx is not an integer then Li1(𝐞(x))\text{Li}_{1}\big{(}\mathbf{e}(x)\big{)} equals the conditionally convergent logarithmic series n1𝐞(nx)n\sum_{n\geq 1}\frac{\mathbf{e}(nx)}{n}. The Fourier expansion formula for the Bernoulli polynomials (8.3) shows that

(8.6) Lik(𝐞(x))+(1)kLik(𝐞(x))=(2πi)kk!Bk(x)(0x1)\text{Li}_{k}\big{(}\mathbf{e}(x)\big{)}+(-1)^{k}\text{Li}_{k}\big{(}\mathbf{e}(-x)\big{)}=-\frac{(2\pi i)^{k}}{k!}B_{k}(x)\hskip 8.5359pt(\begin{subarray}{c}0\leq x\leq 1\end{subarray})

for each k2k\geq 2. The identity above is valid even for k=1k=1 if 0<x<10<x<1.

Lemma 8.3.

Let k2k\geq 2 and λ¯N,k\bar{\lambda}\in\mathcal{I}_{N,k}. Suppose that (λ1,λ2)(\lambda_{1},\lambda_{2}) is a lift of λ¯\bar{\lambda} that satisfies 0λ1<N0\leq\lambda_{1}<N. Then

(8.7) (2π)sΓ(s)L(ek(λ¯,N),s)\displaystyle\frac{(2\pi)^{s}}{\Gamma(s)}L^{\ast}\big{(}e_{k}(\bar{\lambda},N),s\big{)}
=(2πi)k(k1)!{Lis(𝐞(λ2N))ζ(λ1N,sk+1)+\displaystyle=\frac{(-2\pi i)^{k}}{(k-1)!}\big{\{}\emph{Li}_{s}\big{(}\mathbf{e}(\frac{\lambda_{2}}{N})\big{)}\zeta^{\ast}(\frac{\lambda_{1}}{N},s-k+1)+
(1)kLis(𝐞(λ2N))ζ(1λ1N,sk+1)}\displaystyle\hskip 85.35826pt(-1)^{k}\emph{Li}_{s}\big{(}\mathbf{e}(-\frac{\lambda_{2}}{N})\big{)}\zeta(1-\frac{\lambda_{1}}{N},s-k+1)\big{\}}

where

ζ(λ1N,s)={ζ(λ1N,s),if 0<λ1<N;ζ(1,s),if λ1=0.\zeta^{\ast}(\frac{\lambda_{1}}{N},s)=\begin{cases}\zeta(\frac{\lambda_{1}}{N},s),&\text{if $0<\lambda_{1}<N$;}\\ \zeta(1,s),&\text{if $\lambda_{1}=0$.}\end{cases}
Proof.

We first assume that Re(s)>k\text{Re}(s)>k. Then the completed LL-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 0λ2<N0\leq\lambda_{2}<N. Suppose that 1rk11\leq r\leq k-1. Then

L(ek(λ¯,N),r)\displaystyle L^{\ast}\big{(}e_{k}(\bar{\lambda},N),r\big{)} =ir(2πi)k(k1)!Bkr(λ1N)Br(λ2N)r(kr)\displaystyle=i^{r}\frac{(-2\pi i)^{k}}{(k-1)!}\frac{B_{k-r}(\frac{\lambda_{1}}{N})B_{r}(\frac{\lambda_{2}}{N})}{r(k-r)}
+(1)k1ik𝟙/N(λ1,0)𝟙(r,k1)2πk1Lik1(𝐞(λ2N))\displaystyle+(-1)^{k-1}i^{k}\mathbbm{1}_{\mathbb{Z}/N\mathbb{Z}}(\lambda_{1},0)\mathbbm{1}_{\mathbb{Z}}(r,k-1)\frac{2\pi}{k-1}\emph{Li}_{k-1}\big{(}\mathbf{e}(\frac{\lambda_{2}}{N})\big{)}
𝟙/N(λ2,0)𝟙(r,1)2πk1Lik1(𝐞(λ1N)).\displaystyle-\mathbbm{1}_{\mathbb{Z}/N\mathbb{Z}}(\lambda_{2},0)\mathbbm{1}_{\mathbb{Z}}(r,1)\frac{2\pi}{k-1}\emph{Li}_{k-1}\big{(}\mathbf{e}(-\frac{\lambda_{1}}{N})\big{)}.

The proof of the identity involves the following standard properties of the Bernoulli polynomials [17, 4.5]:

(8.8) Bn(1t)=(1)nBn(t),(0t1,n0),Bn(1+t)=Bn(t)+ntn1.(0t1,n1)\begin{gathered}B_{n}(1-t)=(-1)^{n}B_{n}(t),\hskip 8.5359pt(\begin{subarray}{c}0\leq t\leq 1,\;n\geq 0\end{subarray}),\\ B_{n}(1+t)=B_{n}(t)+nt^{n-1}.\hskip 8.5359pt(\begin{subarray}{c}0\leq t\leq 1,\;n\geq 1\end{subarray})\end{gathered}

In particular

Bn(1)={Bn(0),if n2;Bn(0)+1,if n=1.B_{n}(1)=\begin{cases}B_{n}(0),&\text{if $n\geq 2$;}\\ B_{n}(0)+1,&\text{if $n=1$.}\end{cases}
Proof.

If 0<λ1,λ2<N0<\lambda_{1},\lambda_{2}<N then the identity is an easy consequence of the formulas given in (8.5)-(8.8):

L(ek(λ¯,N),r)\displaystyle L^{\ast}\big{(}e_{k}(\bar{\lambda},N),r\big{)}
=(r1)!(2π)r(2πi)k(k1)!(kr){Lir(𝐞(λ2N))Bkr(λ1N)+\displaystyle=-\frac{(r-1)!}{(2\pi)^{r}}\frac{(-2\pi i)^{k}}{(k-1)!(k-r)}\Big{\{}\text{Li}_{r}\big{(}\mathbf{e}(\frac{\lambda_{2}}{N})\big{)}B_{k-r}(\frac{\lambda_{1}}{N})+
(1)kLir(𝐞(λ2N))Bkr(1λ1N)}\displaystyle\hskip 170.71652pt(-1)^{k}\text{Li}_{r}\big{(}\mathbf{e}(-\frac{\lambda_{2}}{N})\big{)}B_{k-r}(1-\frac{\lambda_{1}}{N})\Big{\}}
=ir(2πi)k(k1)!Bkr(λ1N)Br(λ2N)r(kr).\displaystyle=i^{r}\frac{(-2\pi i)^{k}}{(k-1)!}\frac{B_{k-r}(\frac{\lambda_{1}}{N})B_{r}(\frac{\lambda_{2}}{N})}{r(k-r)}.

The extremal case of (8.8) together with (8.6) guarantee that the calculation for 0<λ1,λ2<N0<\lambda_{1},\lambda_{2}<N extends to the following situations without any difficulty:

  • all (λ1,λ2)(\lambda_{1},\lambda_{2}) if 2rk22\leq r\leq k-2,

  • {(0,λ2)λ20}\{(0,\lambda_{2})\mid\lambda_{2}\neq 0\} if rk1r\neq k-1, {(λ1,0)λ10}\{(\lambda_{1},0)\mid\lambda_{1}\neq 0\} if r1r\neq 1.

Suppose that r=k1r=k-1 and λ1=0\lambda_{1}=0. The hypothesis on λ¯\bar{\lambda} implies that if k=2k=2 then λ2\lambda_{2} is automatically nonzero. In this situation ζ\zeta^{\ast} contributes B1(1)B_{1}(1) instead of B1(0)B_{1}(0) and this discrepancy gives rise to the desired term involving Lik1(𝐞(λ2N))\text{Li}_{k-1}\big{(}\mathbf{e}(\frac{\lambda_{2}}{N})\big{)}. It remains to verify the identity for r=1r=1 and λ2=0\lambda_{2}=0. As before, if k=2k=2 then λ10\lambda_{1}\neq 0. The functional equation for the completed LL-function implies that

L(ek((λ1,0)¯,N),1)\displaystyle L^{\ast}\big{(}e_{k}\big{(}\overline{(\lambda_{1},0)},N\big{)},1\big{)}
=(i)kL(ek((0,λ1)¯,N),k1)\displaystyle=(-i)^{k}L^{\ast}\big{(}e_{k}\big{(}\overline{(0,\lambda_{1})},N\big{)},k-1\big{)}
=i(2πi)k(k1)!Bk1(λ1N)B1(0)k12πk1Lik1(𝐞(λ1N)).\displaystyle=i\frac{(-2\pi i)^{k}}{(k-1)!}\frac{B_{k-1}(\frac{\lambda_{1}}{N})B_{1}(0)}{k-1}-\frac{2\pi}{k-1}\text{Li}_{k-1}\big{(}\mathbf{e}(-\frac{\lambda_{1}}{N})\big{)}.

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 LL function in Proposition 8.4 motivates us to normalize the elliptic Eisenstein series as

e~k(λ¯,N):=1(2πi)kek(λ¯,N).(λ¯Λ/NΛ)\tilde{e}_{k}(\bar{\lambda},N):=\frac{1}{(-2\pi i)^{k}}e_{k}(\bar{\lambda},N).\hskip 8.5359pt(\begin{subarray}{c}\bar{\lambda}\in\Lambda/N\Lambda\end{subarray})

The following lemma computes the value of the period function attached to the Γ\Gamma-stable subset k(Γ(N))\mathcal{M}_{k}\big{(}\Gamma(N)\big{)} at the normalized elliptic Eisenstein series.

Lemma 8.5.

Let k2k\geq 2 and λ¯N,k\bar{\lambda}\in\mathcal{I}_{N,k}. Suppose that (λ1,λ2)(\lambda_{1},\lambda_{2}) is a lift of λ¯\bar{\lambda} satisfying 0λ1,λ2<N0\leq\lambda_{1},\lambda_{2}<N. Then

𝗉k(T,e~k(λ¯,N),X)=Bk(λ1N)k!(k1)((X+1)k1Xk1),\displaystyle\mathsf{p}_{k}^{\ast}\big{(}T,\tilde{e}_{k}(\bar{\lambda},N),X\big{)}=-\frac{B_{k}(\frac{\lambda_{1}}{N})}{k!(k-1)}\big{(}(X+1)^{k-1}-X^{k-1}\big{)},
𝗉k(S,e~k(λ¯,N),X)=r=0k2(kr+1)Bkr1(λ1N)Br+1(λ2N)k!(k1)Xk2r+(1)k1\displaystyle\mathsf{p}_{k}^{\ast}\big{(}S,\tilde{e}_{k}(\bar{\lambda},N),X\big{)}=-\sum_{r=0}^{k-2}\binom{k}{r+1}\frac{B_{k-r-1}(\frac{\lambda_{1}}{N})B_{r+1}(\frac{\lambda_{2}}{N})}{k!(k-1)}X^{k-2-r}+(-1)^{k-1}
𝟙/N(λ1,0)Lik1(𝐞(λ2N))(k1)(2πi)k1+𝟙/N(λ2,0)Lik1(𝐞(λ1N))(k1)(2πi)k1Xk2.\displaystyle\mathbbm{1}_{\mathbb{Z}/N\mathbb{Z}}(\lambda_{1},0)\frac{\emph{Li}_{k-1}\big{(}\mathbf{e}(\frac{\lambda_{2}}{N})\big{)}}{(k-1)(-2\pi i)^{k-1}}+\mathbbm{1}_{\mathbb{Z}/N\mathbb{Z}}(\lambda_{2},0)\frac{\emph{Li}_{k-1}\big{(}\mathbf{e}(-\frac{\lambda_{1}}{N})\big{)}}{(k-1)(-2\pi i)^{k-1}}X^{k-2}.
Proof.

The demonstration is a straightforward exercise using the formulas in Proposition 7.6, Lemma 8.1, and Proposition 8.4. ∎

Note that the value of the period polynomial at SS 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 \mathbb{Q}; cf. [20, 7]. Let

(8.9) [ESk]m,Γ(N):k(Γ(N))H1(Γ(N),Vk2,)[\text{ES}_{k}]_{\text{m},\Gamma(N)}:\mathcal{M}_{k}\big{(}\Gamma(N)\big{)}\to H^{1}\big{(}\Gamma(N),V_{k-2,\mathbb{C}}\big{)}

be the Eichler-Shimura map attached to the space of modular forms.

Proposition 8.6.

Let k2k\geq 2 and λ¯N,k\bar{\lambda}\in\mathcal{I}_{N,k}. Then the image of e~k(λ¯,N)\tilde{e}_{k}(\bar{\lambda},N) under the Eichler-Shimura map (8.9) is definable over \mathbb{Q}.

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 e~k(λ¯,N)\tilde{e}_{k}(\bar{\lambda},N) under Ind-[ESk]\text{Ind-}[\text{ES}_{k}] lies inside the rational subspace H1(Γ,Fun(Γ(N)\Γ,Vk2,))H^{1}\big{(}\Gamma,\text{Fun}(\Gamma(N)\backslash\Gamma,V_{k-2,\mathbb{Q}})\big{)}. For σΓ\sigma\in\Gamma let (λ1σ,λ2σ)(\lambda_{1}^{\sigma},\lambda_{2}^{\sigma}) denote the unique lift of λ¯σ\bar{\lambda}\sigma in [0,N)×[0,N)[0,N)\times[0,N). Note that (λ1σ,λ2σ)(\lambda_{1}^{\sigma},\lambda_{2}^{\sigma}) depends only on the coset of σ\sigma in Γ(N)\Γ\Gamma(N)\backslash\Gamma.

Case I. k3k\geq 3.

Define FFun(Γ(N)\Γ,Vk2,)F\in\text{Fun}\big{(}\Gamma(N)\backslash\Gamma,V_{k-2,\mathbb{C}}\big{)} by setting

F(Γ(N)σ)=i3kL(e~k(λ¯σ,N),k1).F\big{(}\Gamma(N)\sigma\big{)}=i^{3-k}L^{\ast}\big{(}\tilde{e}_{k}(\bar{\lambda}\sigma,N),k-1\big{)}.

A straight-forward computation using the formulas in Proposition 8.4 and Lemma 8.5 shows that the modified cocycle

ΓFun(Γ(N)\Γ,Vk2,),γInd-ESk(e~k(λ¯,N))(γ)+F|γF,\Gamma\to\text{Fun}\big{(}\Gamma(N)\backslash\Gamma,V_{k-2,\mathbb{C}}\big{)},\;\gamma\mapsto\text{Ind-}\text{ES}^{\ast}_{k}(\tilde{e}_{k}(\bar{\lambda},N))(\gamma)+F\lvert_{\gamma}-F,

takes values in Fun(Γ(N)\Γ,Vk2,)\text{Fun}\big{(}\Gamma(N)\backslash\Gamma,V_{k-2,\mathbb{Q}}\big{)} at TT and SS.

Case II. k=2k=2.

The problem, in this case, is that both the transcendental terms in the special LL-value may simultaneously contribute a term to the calculation. We construct another function FFun(Γ(N)\Γ,)F\in\text{Fun}\big{(}\Gamma(N)\backslash\Gamma,\mathbb{C}\big{)} by putting

F(Γ(N)σ)={0,if λ1σ0;iL(e~2(λ¯σ,N),1),if λ1σ=0.F\big{(}\Gamma(N)\sigma\big{)}=\begin{cases}0,&\text{if $\lambda_{1}^{\sigma}\neq 0$;}\\ iL^{\ast}\big{(}\tilde{e}_{2}(\bar{\lambda}\sigma,N),1\big{)},&\text{if $\lambda_{1}^{\sigma}=0$.}\end{cases}

Then the modified cocycle

ΓFun(Γ(N)\Γ,),γInd-ES2(e~2(λ¯,N))(γ)+F|γF,\Gamma\to\text{Fun}\big{(}\Gamma(N)\backslash\Gamma,\mathbb{C}\big{)},\;\gamma\mapsto\text{Ind-}\text{ES}^{\ast}_{2}(\tilde{e}_{2}(\bar{\lambda},N))(\gamma)+F\lvert_{\gamma}-F,

again takes values in Fun(Γ(N)\Γ,)\text{Fun}\big{(}\Gamma(N)\backslash\Gamma,\mathbb{Q}\big{)} at TT and SS. ∎

Let 𝕂\mathbb{K} be a subfield of \mathbb{C} containing N\mathbb{Q}_{N}. Recall that the 𝕂\mathbb{K}-span of {ek(λ¯,N)λ¯N,k}\{e_{k}(\bar{\lambda},N)\mid\bar{\lambda}\in\mathcal{I}_{N,k}\} equals (2πi)kk(Γ(N),𝕂)(2\pi i)^{k}\mathcal{E}_{k}\big{(}\Gamma(N),\mathbb{K}\big{)}. Therefore the collection

{e~k(λ¯,N)λ¯N,k}\{\tilde{e}_{k}(\bar{\lambda},N)\mid\bar{\lambda}\in\mathcal{I}_{N,k}\}

is a 𝕂\mathbb{K}-spanning set of k(Γ(N),𝕂)\mathcal{E}_{k}\big{(}\Gamma(N),\mathbb{K}\big{)}.

Corollary 8.7.

The homomorphism [ESk]m,Γ(N)[\emph{ES}_{k}]_{\emph{m},\Gamma(N)} maps k(Γ(N),𝕂)\mathcal{E}_{k}\big{(}\Gamma(N),\mathbb{K}\big{)} inside the canonical 𝕂\mathbb{K}-form H1(Γ(N),Vk2,𝕂)H^{1}\big{(}\Gamma(N),V_{k-2,\mathbb{K}}\big{)}.

Proof.

Proposition 8.6 demonstrates that

[ESk]m,Γ(N)(e~k(λ¯,N))H1(Γ(N),Vk2,)[\text{ES}_{k}]_{\text{m},\Gamma(N)}\big{(}\tilde{e}_{k}(\bar{\lambda},N)\big{)}\in H^{1}\big{(}\Gamma(N),V_{k-2,\mathbb{Q}}\big{)}

for each λ¯N,k\bar{\lambda}\in\mathcal{I}_{N,k}. Thus the assertion follows from the discussion above. ∎

Suppose that 𝒢\mathcal{G} is a congruence subgroup containing the principal level Γ(N)\Gamma(N). One considers the 𝒢\mathcal{G}-invariants of the domain and codomain of (8.9) to retrieve the Eichler-Shimura homomorphism for 𝒢\mathcal{G}:

(8.10) [ESk]m,𝒢:k(𝒢)H1(𝒢,Vk2,).[\text{ES}_{k}]_{\text{m},\mathcal{G}}:\mathcal{M}_{k}(\mathcal{G})\to H^{1}\big{(}\mathcal{G},V_{k-2,\mathbb{C}}\big{)}.
Theorem 8.8.

Let 𝕂\mathbb{K} be a subfield of \mathbb{C} containing N\mathbb{Q}_{N}. Then the image of k(𝒢,𝕂)\mathcal{E}_{k}(\mathcal{G},\mathbb{K}) under the Eichler-Shimura map (8.10) lies inside H1(𝒢,Vk2,𝕂)H^{1}\big{(}\mathcal{G},V_{k-2,\mathbb{K}}\big{)}.

Proof.

From Corollary 8.7 we know that the Eichler-Shimura homomorphism (8.9) maps k(Γ(N),𝕂)\mathcal{E}_{k}\big{(}\Gamma(N),\mathbb{K}\big{)} inside H1(Γ(N),Vk2,𝕂)H^{1}\big{(}\Gamma(N),V_{k-2,\mathbb{K}}\big{)}. But the 𝒢\mathcal{G}-invariant subspace of k(Γ(N),𝕂)\mathcal{E}_{k}\big{(}\Gamma(N),\mathbb{K}\big{)}, resp. H1(Γ(N),Vk2,𝕂)H^{1}\big{(}\Gamma(N),V_{k-2,\mathbb{K}}\big{)}, equals k(𝒢,𝕂)\mathcal{E}_{k}(\mathcal{G},\mathbb{K}), resp. H1(𝒢,Vk2,𝕂)H^{1}\big{(}\mathcal{G},V_{k-2,\mathbb{K}}\big{)}. Therefore the image of k(𝒢,𝕂)\mathcal{E}_{k}(\mathcal{G},\mathbb{K}) under [ESk]m,𝒢[\text{ES}_{k}]_{\text{m},\mathcal{G}} must lie in H1(𝒢,Vk2,𝕂)H^{1}\big{(}\mathcal{G},V_{k-2,\mathbb{K}}\big{)}. ∎

Proof of Theorem 1.5:

Let the notation be as in Theorem 1.5. Suppose that f𝒢,kf\in\mathcal{B}_{\mathcal{G},k}. From the definition of rational structure it is clear that fk(𝒢,N)f\in\mathcal{E}_{k}(\mathcal{G},\mathbb{Q}_{N}). Hence the assertion is a consequence of Theorem 8.8. \square

Remark 8.9.

The space of Eisenstein series k(𝒢)\mathcal{E}_{k}(\mathcal{G}) possesses a computable basis (different from 𝒢,k\mathcal{B}_{\mathcal{G},k}) so that each basis element lies in k(𝒢,N)\mathcal{E}_{k}(\mathcal{G},\mathbb{Q}_{N}) and admits Fourier expansion with coefficients in N\mathbb{Q}_{N} [15]. Theorem 8.8 shows that the image of this basis under the Eichler-Shimura map is also definable over N\mathbb{Q}_{N}.

Observe that the 𝕂\mathbb{K}-subspace H1(𝒢,Vk2,𝕂)H^{1}(\mathcal{G},V_{k-2,\mathbb{K}}) of H1(𝒢,Vk2,)H^{1}(\mathcal{G},V_{k-2,\mathbb{C}}) 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 k(𝒢,𝕂)\mathcal{E}_{k}(\mathcal{G},\mathbb{K}) 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 k(N,χ)\mathcal{E}_{k}(N,\chi). Let NN be an integer 1\geq 1 and χU^N\chi\in\widehat{U}_{N}. Suppose that χ\chi is defined over a subfield 𝕂\mathbb{K} of \mathbb{C}. Then Vk2,𝕂χV_{k-2,\mathbb{K}}^{\chi} is a 𝕂\mathbb{K}-form of Vk2,χV_{k-2,\mathbb{C}}^{\chi} and the inclusion Vk2,𝕂χVk2,χV_{k-2,\mathbb{K}}^{\chi}\hookrightarrow V_{k-2,\mathbb{C}}^{\chi} induces a change of scalar isomorphism:

H1(Γ0(N),Vk2,𝕂χ)𝕂H1(Γ0(N),Vk2,χ).H^{1}\big{(}\Gamma_{0}(N),V_{k-2,\mathbb{K}}^{\chi}\big{)}\otimes_{\mathbb{K}}\mathbb{C}\cong H^{1}\big{(}\Gamma_{0}(N),V_{k-2,\mathbb{C}}^{\chi}\big{)}.

Note that H1(Γ1(N),Vk2,𝕂)H^{1}\big{(}\Gamma_{1}(N),V_{k-2,\mathbb{K}}\big{)} is a 𝕂\mathbb{K}-form of H1(Γ1(N),Vk2,)H^{1}\big{(}\Gamma_{1}(N),V_{k-2,\mathbb{C}}\big{)}. Since χ\chi is defined over 𝕂\mathbb{K} the χ\chi-eigenspace for the Γ0(N)\Gamma_{0}(N) on H1(Γ1(N),Vk2,)H^{1}\big{(}\Gamma_{1}(N),V_{k-2,\mathbb{C}}\big{)} is also defined over 𝕂\mathbb{K}, i.e., H1(Γ1(N),Vk2,𝕂)[χ]𝕂H1(Γ1(N),Vk2,)[χ]H^{1}\big{(}\Gamma_{1}(N),V_{k-2,\mathbb{K}}\big{)}[\chi]\otimes_{\mathbb{K}}\mathbb{C}\cong H^{1}\big{(}\Gamma_{1}(N),V_{k-2,\mathbb{C}}\big{)}[\chi]. Lemma 7.7 shows that the natural restriction map induces an isomorphism

H1(Γ0(N),Vk2,χ)H1(Γ1(N),Vk2,)[χ]H^{1}\big{(}\Gamma_{0}(N),V_{k-2,\mathbb{C}}^{\chi}\big{)}\cong H^{1}\big{(}\Gamma_{1}(N),V_{k-2,\mathbb{C}}\big{)}[\chi]

that maps H1(Γ0(N),Vk2,𝕂χ)H^{1}\big{(}\Gamma_{0}(N),V_{k-2,\mathbb{K}}^{\chi}\big{)} onto H1(Γ1(N),Vk2,𝕂)[χ]H^{1}\big{(}\Gamma_{1}(N),V_{k-2,\mathbb{K}}\big{)}[\chi]. Recall the basis k,χ\mathcal{B}_{k,\chi} of k(N,χ)\mathcal{E}_{k}(N,\chi) defined in (7.10).

Theorem 8.11.

Let the notation be as above. The image of k,χ\mathcal{B}_{k,\chi} under [ESk]m,Nχ[\emph{ES}_{k}]_{\emph{m},N}^{\chi} lies inside H1(Γ0(N),Vk2,𝕂χ)H^{1}\big{(}\Gamma_{0}(N),V_{k-2,\mathbb{K}}^{\chi}\big{)} where 𝕂=φ(N)N=lcm(N,φ(N))\mathbb{K}=\mathbb{Q}_{\varphi(N)}\mathbb{Q}_{N}=\mathbb{Q}_{\emph{lcm}(N,\varphi(N))}.

Proof.

Let 𝕂\mathbb{K} be as in the statement of the theorem. We know that k,χk(Γ1(N),𝕂)\mathcal{B}_{k,\chi}\subseteq\mathcal{E}_{k}\big{(}\Gamma_{1}(N),\mathbb{K}\big{)}. Therefore, by Theorem 8.8, the image of k,χ\mathcal{B}_{k,\chi} under the Eichler-Shimura isomorphism for Γ1(N)\Gamma_{1}(N) (7.7) lies in H1(Γ1(N),Vk2,𝕂)[χ]H^{1}\big{(}\Gamma_{1}(N),V_{k-2,\mathbb{K}}\big{)}[\chi]. Thus the assertion is a consequence of the diagram in (7.8) and the discussion before the theorem. ∎

One can extract a Hecke stability statement from the theorem above.

Corollary 8.12.

Let 𝕂\mathbb{K} be a subfield of \mathbb{C} containing φ(N)N\mathbb{Q}_{\varphi(N)}\mathbb{Q}_{N}. Then the 𝕂\mathbb{K}-span of k,χ\mathcal{B}_{k,\chi} is stable under the action of twisted double coset operators defined in Section 7.2.

Proof.

Follows from the χ\chi-Hecke equivariance of the twisted Eichler-Shimura map [ESk]m,Nχ[\text{ES}_{k}]_{\text{m},N}^{\chi} 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 M1,1M_{1,1}, 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 GL2\emph{GL}_{2}, Invent. math. 89, 37-118 (1987). https://doi.org/10.1007/BF01404673
  • [10] H. Hida, Elementary theory of LL-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