Abstract
In this paper, we study polar harmonic Maass forms of negative integral weight. Using work of Fay, we construct Poincaré series which span the space of such forms and show that their elliptic coefficients exhibit duality properties which are similar to the properties known for Fourier coefficients of harmonic Maass forms and weakly holomorphic modular forms.
1. Introduction and statement of results
Harmonic Maass forms are smooth functions on the upper half-plane which satisfy modularity properties under , are annihilated by the associated hyperbolic weight Laplacian and have at most linear exponential growth at the cusps. These naturally generalize weakly holomorphic modular forms, which are meromorphic modular forms whose only poles appear at cusps; in particular, holomorphicity in is replaced with annihilation by the hyperbolic Laplacian defined in (2.1). Harmonic Maass forms play a major role in work on mock theta functions, singular moduli and their real quadratic analogs, and many other applications.
Functions with properties similar to harmonic Maass forms, but with poles in the upper half-plane, appear in a number of recent results, including work on the resolvent kernel [16], outputs of theta lifts [4], cycle integrals [12] and Fourier coefficients of meromorphic modular forms [8, 25]. In considering functions with poles in the upper half-plane rather than solely at the cusps, we might expect that the behavior of such functions is similar to the behavior of harmonic Maass forms. In this paper, we study spaces of such polar harmonic Maass forms, which generalize harmonic Maass forms in the same way that meromorphic modular forms generalize weakly holomorphic modular forms.
From another perspective, the subspace of polar harmonic Maass forms consisting of meromorphic modular forms is analogous to the subspace of harmonic Maass forms consisting of weakly holomorphic modular forms. Meromorphic modular forms have not only a Fourier expansion at the cusp
, but also an
elliptic expansion around each point
in terms of powers of
, where
. Polar harmonic Maass forms have a more general elliptic expansion where the coefficients
may additionally depend on
; we give this expansion in Proposition
2.2 below.
Polar harmonic Maass forms are not the only class of generalizations of harmonic Maass forms. Locally harmonic Maass forms are obtained by relaxing the harmonicity condition to allow jump discontinuities on the upper half-plane. For instance, these discontinuities may appear along geodesics associated with quadratic forms of positive discriminant, just as polar harmonic Maass forms may have poles at CM points coming from quadratic forms of negative discriminant. Locally harmonic Maass forms have been independently studied by Hövel [21] as outputs of theta lifts in weight zero, by Pioline [27] in the setting of string theory, and by Kohnen and two of the authors [10] as ‘generating functions’ of indefinite integral binary quadratic forms. Additional applications of polar harmonic Maass forms include the following:
In level one, divisors (sets of zeros and poles, with their orders) of modular forms can be studied using only ordinary modular forms [13]. The obstacle in higher level is that the required meromorphic modular forms no longer exist; however, polar harmonic analogs can take their place [7, 11]. These functions have weight zero and are analogous to the functions investigated in this paper.
Higher Green’
s functions are modular functions on
which are smooth away from the diagonal, where they have logarithmic singularities, and they are annihilated by the Laplace operator in both variables. These higher Green’s functions appear as special cases of the resolvent kernel studied by Fay [
16] (see also [
20] for a detailed study of these functions). Gross and Zagier [
17] conjectured that their evaluations at CM points are essentially logarithms of algebraic numbers; there has been progress on this conjecture, see for example [
22,
28]. In [
12], two of the authors showed jointly with von Pippich that the value of the Green’s function evaluated at CM points is basically given by a (regularized) inner product of the functions
where
is an
-equivalence class of integral binary quadratic forms of negative discriminant−
.
Functions related to (1.2) (namely summing up all classes of discriminant ) naturally occur as outputs of theta lifts and also closely resemble the locally harmonic Maass forms appearing in [10], except that the integral binary quadratic forms upon which one takes ‘generating functions’ instead have negative discriminants.
Hardy and Ramanujan [19] considered modular forms with simple poles in and in particular found a formula for the reciprocal of the Eisenstein series . General formulas for functions with simple poles were investigated by Bialek [3]. Berndt et al. [2] then investigated functions with poles of order two. All of these proofs use the Hardy–Ramanujan circle method, but the calculation rapidly becomes more complicated with rising pole order. Using polar harmonic Maass forms gives a powerful tool to systematically study such coefficients by writing down explicit bases of the space and then specializing the resulting formulas to meromorphic modular forms in special cases, as done by two of the authors [9, 12].
In addition to having similar elliptic expansions, polar harmonic Maass forms and meromorphic modular forms of weights
and
are interconnected by certain differential operators which naturally occur in the theory of harmonic Maass forms. Here and throughout,
is assumed to be an arbitrary integer, and we use
instead if there is some restriction on the weight (for instance, if we require
). For
and
, set
where
. If
satisfies weight
modularity, then
is modular of weight
, while if
satisfies weight
modularity, then
satisfies weight
modularity.
Given a polar harmonic Maass form of weight
, one may eliminate the singularity at
and, if
, also the constant term in the Fourier expansion, by subtracting an appropriate harmonic Maass form (cf. [
5, Theorem 6.10] for the existence of forms with arbitrary principal parts). This yields a weight
polar harmonic cusp form, a weight
polar harmonic Maass form which vanishes at
if
and is bounded at
if
. We denote the subspace of such forms by
. A canonical basis for this space may be defined by specifying the growth behavior near singularities in
, which is given via principal parts at
; see (
2.11) for further details on the principal parts which may occur. This basis is defined in (
4.5) below, and we show in Theorem
1.1 that for
they indeed span
. Specifically, for each
and
, in (
4.5) we construct the unique weight
polar harmonic cusp form
with principal part
Here
is the size of the stabilizer of
in
and
is the constant defined in (
2.7) below and explicitly computed as a quotient of factorials in (
4.11). For
, the functions
have non-meromorphic principal parts. We describe these explicitly in Theorem
4.3 below.
To understand the behavior of the basis elements
under the differential operators defined in (
1.3), we define the subspace
consisting of meromorphic modular forms without poles at
, which we call
meromorphic cusp forms. For each point
in the fundamental domain
, we let
(resp.
) be the subspace of forms in
(resp.
) with singularities allowed only at
. For
, Petersson (cf. [
25, equation (5c.3)] or [
26, Equation (21)]) defined a family of meromorphic Poincaré series
which form a natural canonical basis for the space
. Specifically, for
and a point
, the function
is the unique meromorphic cusp form which is orthogonal to cusp forms (see [
26, Satz 8]) under a regularized inner product defined in [
26, equation (3)] and whose principal part is
As shown in the next theorem, the action of the differential operators and give an additional natural splitting of the space into three subspaces, which we denote by , , and the space of cusp forms . Again using the regularization [26, Equation (3)], or its extension [12, Equation (3.3)] to arbitrary meromorphic cusp forms, the subspace (resp. ) consists of those forms in which are orthogonal to cusp forms and whose principal parts are linear combinations of (1.5) with (resp. ). The families of meromorphic Poincaré series with or with form bases for and , respectively.
Theorem 1.1.
Suppose that.
Everyis a linear combination of the functions from. Moreover, if the only poles ofinoccur at points equivalent tounder the action of, thenis a linear combination of functions from.
If, then (resp. ) if and only if (resp. ).
If, thenif and only if.
Forand,
we have
Remark.
As alluded to before the theorem, Theorems 1.1 (2) and 1.1 (3) naturally split into three subspaces , and which are paired up as images under and of the same subspace. There are analogous subspaces of weakly holomorphic modular forms which are connected via these differential operators in the same manner. Specifically, the space spanned by the Eisenstein series is paired with itself in the same way as the space , while the space of cusp forms is associated with its orthogonal complement inside the subspace of weakly holomorphic modular forms which have vanishing constant terms in their Fourier expansion. This latter space mirrors .
For
, our next result gives a duality-type relationship between the coefficients of
and those of
. To state it, for
, let
denote the
th coefficient in the elliptic expansion (
1.1) around
of
. Similarly,
is the
th coefficient in the meromorphic part of the elliptic expansion (see (
2.11)) around
of
; in other words, by (
1.4) these are the coefficients of the unique weight
polar harmonic Maass forms with principal parts
which closely resemble the principal parts (
1.5) in positive weight.
Remark.
Similar duality results for Fourier coefficients of harmonic Maass forms and weakly holomorphic modular forms are well known in both integral and half-integral weight. Petersson used such identities in his construction of a basis of meromorphic modular forms (see [
25, (3a.9)]), while a systematic study of them originated from Zagier’s work on singular moduli [
29]. Since then, results have been obtained by a number of authors, including the second author and Duke in [
15], and Guerzhoy [
18], among others. To give one such result, for
, take
and let
denote the
th coefficient of the weight
weakly holomorphic modular form which grows toward
like
and let
be the
th coefficient of the holomorphic part of the weight
harmonic Maass form which grows towards
like
. Then the duality
holds. A natural question for future investigation is whether similar duality results hold for elliptic coefficients of harmonic Maass forms, and what they are dual to. The techniques of this paper may be a useful starting point.
Remark.
Another interesting question for future study is whether results similar to those in this paper can also be obtained in half-integral weight. The Poincaré series may be constructed in the same way, but to fully prove Theorem 1.1, we must differentiate times. If , this leaves the challenging question of finding a half-derivative. For harmonic Maass forms, such derivatives were constructed in [6].
The paper is organized as follows. In Section 2, we introduce polar harmonic Maass forms and recall results from Fay, who studied related functions in [16]. In Section 3, we relate Fay’s functions to polar harmonic Maass forms and compute the elliptic expansions of polar harmonic Maass forms, and, in Section 4, we investigate Poincaré series and prove Theorem 1.1. We conclude the paper by proving Theorem 1.2 in Section 5.
2. Preliminaries
2.1. Basic definitions
For
,
, and
, we define the usual slash operator
Definition.
For , a polar harmonic Maass form of weight is a function which is real analytic outside a discrete set of points and satisfies the following conditions:
For every , we have .
We have
, with
the
weighthyperbolic Laplace operatorFor every , there exists such that is bounded in some neighborhood of .
The function has at most linear exponential growth at ; that is, for some constant (uniform in for sufficiently large) as .
If
is replaced by
, then
is called a
polar Maass form with eigenvalue
.
Denote by the space of polar harmonic Maass forms of weight . The subspace of consisting of forms that map under to cusp forms is denoted by ; more generally, we add the superscript ‘cusp’ to any subspace of to indicate the space formed by taking the intersection of the subspace with . We also use the superscript to indicate the subspace of forms whose only singularity in appears at .
Although in this paper we are primarily interested in expansions of polar harmonic Maass forms around points in the upper half-plane, for completeness and for later comparison we next recall some properties about the Fourier expansions of polar harmonic Maass forms around
. These expansions yield natural decompositions of polar harmonic Maass forms into holomorphic and non-holomorphic parts (cf. [
20, Proposition 4.3]). Namely, for a polar harmonic Maass form
of weight
and
, we have
where, for some
, we define the
holomorphic part (resp.
non-holomorphic part) of
at
as
with the incomplete gamma function
. The sum of all of the terms which grow towards
is called the
principal part of
.
We next consider elliptic expansions of polar harmonic Maass forms. Rather than expansions in
, the natural expansions of polar harmonic Maass forms around
are given in terms of
. We further write
with
the hyperbolic distance between
and
. The second identity in the definition of
follows by the well-known formula (see [
1, p. 131])
where throughout the paper
. From (
2.4), for
one also immediately obtains the invariance
For
and
and
, we also require the function
where
is the
incomplete beta function and
Note that by [
23, 8.17.7], we have
where
is the
Gauss hypergeometric function defined by
with
. We often use the fact that
We also require the
Euler transformation (see 15.8.1 of [
23])
The modified incomplete -function may also be written in special cases as a hypergeometric function, as can be seen by a direct calculation.
Lemma 2.1.
Assume that, , and.
We haveHere and throughout we use the notationif some propertyis true andotherwise.
We have the following elliptic expansion of weight harmonic functions, whose proof is deferred to Section 3.
Proposition 2.2.
Suppose thatand.
Ifsatisfiesand for somethe functionis bounded in some neighborhoodaround,
then there exist,
such that forand,
we have If, then the sum in (2.11) only runs over thosewhich satisfy. If, then the second sum is empty and the third sum only runs over.
Remark.
Instead of the expansion given in (2.11), one could rewrite the second sum in the shape of the third to get a seemingly more uniform expansion. However, it is natural to split off these terms because they have logarithmic singularities. They are also special, as we shall see in Proposition 2.3, in that they are annihilated neither by nor . Thus, they may be viewed in a sense both as both meromorphic and non-meromorphic parts. This emulates the constant term of the non-holomorphic part (2.2) of the expansion at , which is a constant multiple of , is annihilated by neither operator, and also exhibits a logarithmic singularity.
For
annihilated by
(with
), we define the
meromorphic part of the elliptic expansion (
2.11) around
by
and its
non-meromorphic part by
The next proposition, proven in Section 3, explicitly gives the elliptic expansion under the action of the operators and .
Proposition 2.3.
Forandsatisfying,
we haveand In addition to the operators
and
given in (
1.3), we require the classical
Maass raising and
lowering operators:
The raising operator (resp. lowering operator) increases (resp. decreases) the weight by
. Moreover,
We also require iterated raising
For
, the raising operator and
are related by
Bol’
s identity 2.2. Work of Fay
In this section, we recall work of Fay [
16] and rewrite some of his statements in the notation used in this paper. Fay considered functions
transforming for
as
Then
transforms as
Define the operator
By [
16, p. 144], for
, we have
Let
denote the space of
satisfying the following conditions:
;
;
has at most finitely many singularities of finite order in , where .
Functions in
are closely related to polar Maass forms. In order to study the relationship between
acting on Fay’s functions and
acting on polar Maass forms, we require the following variants of the Maass raising and lowering operators (see [
16, (3)]):
Note that
sends
to
and
sends
to
. Moreover (see [
16, (7)])
We also require iterated raising and lowering
We next translate these operators into the notation used in this paper and compare eigenfunctions under these operators.
Proposition 2.4.
For,
we haveIf,
thenIn particular,
is harmonic if and only ifor.
The function if and only if the function is a polar Maass form of weight with eigenvalue . In particular, if or and grows at most like for , then .
Proof
(1) First, it is not hard to see that
Iterating (
2.19) yields (
2.15). Similarly, to prove (
2.16), one first shows that
One then obtains (
2.16) inductively. The eigenfunction property (
2.17) then follows using (
2.14), (
2.20), (
2.19) and (
2.12). To prove (
2.18), suppose that
. Then, by (
2.17), we have
(2) Part (1) implies that the eigenfunction properties of and are equivalent. Comparing the singularities of both functions then yields the claim.□
Fay then considered a natural family of functions which behave well under his differential operators when multiplied by
with
satisfying
. For
,
, and
, these are given by (see [
16, p. 147], slightly modified)
with
where for
we set
.
These functions are meromorphic in
with at most simple poles at
and satisfy certain useful relations. Directly from the definitions (
2.21) and (
2.22), one obtains
Moreover, for
, we have
The special values of
and
in the cases
and
play an important role in our investigation. To describe these, we set
In the next section, we prove the following lemma.
Lemma 2.5.
For and
,
we have
We next define certain Poincaré series considered by Fay. For this, we set (see [
16, (44)]), slightly modified)
with
where
and
Remarks.
Note that if , then we multiply both sides of (2.28) by an appropriate factor to cancel the simple poles occurring in the -factors and then take the limit, as in Lemma 2.5 (3).
The functions
satisfy the symmetry relations
Fay related these functions to the resolvent kernel .
Theorem 2.6.
(Fay [
16, Theorem 2.1]).
Forand.
We have,
for,
If,
then we replace by
.
Fay also considered elliptic expansions of functions in .
Proposition 2.7.
(Fay [
16, Theorem 1.1]).
Ifin some annulusaround,
thenhas an elliptic expansion of the shapeThe proof of Proposition 2.2, which we give in the next section, mostly relies on rewriting Fay’s functions and .
3. Special functions and elliptic expansions
To prove Proposition 2.2 we write the elliptic expansion in terms of Fay’s, which is done in Lemma 2.5 (1).
Proof of Lemma 2.5
(1) Throughout, we use the fact that, with
, we have
For
, Equation (
2.24) follows from the definition, using (
2.10), (
2.9) and (
3.1). If
then, using (
2.10), (
3.1), and abbreviating
with
, the left-hand side of (
2.24) equals
Since
and
, the claim follows from Lemma
2.1 (2).
We next prove (2.25). The claim for follows from the definition using (2.10), (2.8) and (3.1). For , the claim follows by (2.8) and (3.1).
(2) From the definition of
, the left-hand side of (
2.26) equals
Once again using
and (
2.9), we obtain
By (
3.1), we then obtain the claim.
(3) The left-hand side of (
2.27) equals
Canceling
, using
, taking the limit, employing (
3.1), and plugging in the definition of the
, we obtain
We obtain the desired identity by using 15.4.6 of [
23] to evaluate
□
We next combine Lemma 2.5 (1) with Fay’s elliptic expansion in Proposition 2.7 to obtain Proposition 2.2.
Proof of Proposition 2.2
With
, we have, by (
2.17),
since
by assumption. Thus, by Proposition
2.7 with
,
By Lemmas
2.5 (1) and (
2.6), this gives
for some constants
. Rewriting yields the expansion (
2.11) up to the restrictions on
in each of the sums. It thus remains to show that
for
and
for
. To do so, we investigate the asymptotic growth of each term in the sum as
. We repeatedly use the fact that, as
,
, where by
we mean that there is a constant
depending only on
such that
. This gives that
for the first summand in (
2.11).
Moreover, by Lemma
2.1,
Thus, again using
, we have, as
,
Furthermore, for
, since
as
, the asymptotic in (
3.3) implies that we also have
This gives the claimed bounds for
. Finally, the
term behaves like
by (
3.3). This growth is canceled upon multiplying by
.
(2) By (
2.3) and (
2.5), for
in the stabilizer
of
One concludes the claim by [
25, (2a.16)].□
We next compute the action of differential operators on elliptic expansions in Proposition 2.2.
Proof of Proposition 2.3
We first note that, by Proposition
2.4 (1),
We rewrite the right-hand side of (
3.5) in terms of the iterated operators (for
and
)
where (see [
16, after formula (18)])
Namely, using (see [
16, (14)]) that for
and iteratively carrying out the differentiation on
yields
By (
3.5) and (
2.24), we thus have, for
,
By [
16, (18)], we know that
where
Plugging (
3.8) into the right-hand side of (
3.7) simplifies to
We split into the cases
and
.
For
we have, using (
2.9),
Thus (
3.10) becomes, using that
and (
3.1),
Explicitly computing the constant then finishes the claim for
. For
, we have
, giving the claim in this range.
We next act by
on the non-meromorphic part of
. First assume that
. By Lemma
2.1 (1), we then have, using that
,
This is a polynomial in
of degree at most
(with antiholomorphic coefficients depending on
). Differentiating
times hence annihilates these terms.
It remains to determine the image of
on the terms in the non-meromorphic part with
. Using (
2.25), (
3.5), (
3.6), (
3.9) and (
2.26), we obtain that
Computing
and plugging in (
2.23), noting that
, yields the claimed formula.□
4. Poincaré series and the proof of Theorem 1.1
For
,
and
, we define the meromorphic Poincaré series, due to Petersson,
where
We use the convention that
appears as a superscript in the notation if we consider
as a fixed point and we write it as a two-variable function if we consider the properties in the
-variable. The main properties of
needed for this paper are given in the proposition below.
Proposition 4.1
(Petersson [26, Sätze 7 and 8] and [24, Satz 7]). The functions (resp. span (resp. . For, they are cusp forms. Forthey are orthogonal to cusp forms and have the principal partaround.
Remarks.
By Proposition
4.1, the elliptic expansion of
around
may be written
where
denotes the
-equivalence class of
.
As pointed out in [25, p. 72], is modular of weight . Moreover, it is an eigenfunction under with eigenvalue .
We next write
as a special case of Fay’s function
. We set
Lemma 4.2.
If,
thenIf,
then
Proof
(1) By inspecting the definitions (
2.28) and (
4.1), the claim follows once we show that
By definition (
2.30), we have
If
, then we may plug in
directly and then use Lemma
2.5 (2) to obtain
For
, we have
implying (
4.3) in this case.
For
, we obtain (
4.3) for
, computing
If , then in (2.29) and we use Lemma 2.5 (3) (replacing by ) to obtain the desired formula.
(2) By (
2.31), we have
For
, we may then directly plug in
and use (1) to obtain the claim by simplifying, with
,
For
, we use (1) to obtain, by (
4.4),
We then plug in (
2.29) and take the limit to obtain the claim.□
Next define for
the following polar harmonic Maass Poincaré series
with, using (
2.25),
The following more precise version of Theorem 1.1 shows how the functions are related to the functions via differential operators.
Theorem 4.3.
Assume.
The functions (
resp.
)
span the space (
resp.
).
Moreover,The functionsvanish unless,
in which case their principal parts equal Remarks.
It is also natural to ask about the properties of
for fixed
. To investigate this, note that by a comparison of Definitions (
4.5) and (
2.28), we find that
with
given in (
2.23), and we evaluate
via (
2.29). Combining this with Theorem
2.6, one can conclude that the function
has weight
and eigenvalue
under
.
By Theorem
4.3 and Proposition
2.2, one may write the elliptic expansion of
around
for
and
as
Furthermore, using (
2.7), Lemma
2.1 (1) and then Lemma
2.1 (2) yields
giving the constant in (
1.4) in terms of factorials.
Proof of Theorem 4.3
Using (2.28), we conclude that satisfies weight modularity.
The principal part of around comes from the terms in (4.5). Vanishing of the principal part for follows from (3.4) together with [25, (2a.16)]. For , this yields . For , this directly yields the principal part. For or , we rewrite the incomplete beta function using (2.6) and note that for only the non-meromorphic part grows as approaches , while for only the meromorphic part grows.
Since every possible principal part in the elliptic expansion of an element of is obtained as a linear combination of the Poincaré series , these span the space . Moreover, eliminating the principal parts at different points in one at a time implies that the space is spanned by .
We next compute the image of the Poincaré series under
. Using (
4.9) and (
2.15), we obtain
Using Theorem
2.6 twice, we then find that
We next employ Lemma
4.2 (1) and plug back into (
4.12), yielding
We then plug in the definitions of
and
and use (
2.13) to conclude (
4.7).
It remains to compute the image under
. First, by Proposition
2.4 (1), with
, we have
Using (
4.9) and (
2.29) thus gives
Now by Theorem
2.6, we have
Again using Theorem
2.6 and then applying (
2.32) gives
For
, we then use Lemma
4.2 (2) to obtain
We then simplify the factor in front using (
2.23) to obtain the claim for
. For
, we use Lemma
4.2 (2) to obtain that
Simplifying the constant yields the claim.□
We may now combine the results in this section to obtain Theorem 1.1.
Proof of Theorem 1.1
(1) Part (1) is the first statement in Theorem 4.3.
(2) The claim follows from (4.8) and (4.7) together with the fact that if and only if and if and only if . These claims about follow in turn from the principal parts and orthogonality given in Proposition 4.1.
(3) This follows by (4.8), (4.7) and Proposition 4.1, since if and only if .
(4) The statements given here are precisely (4.8) and (4.7).□
5. Duality, orthogonality, and the proof of Theorem 1.2
5.1. Definition of the inner product
Petersson defined a regularized inner product (see [
26, p. 34]) for meromorphic modular forms by taking the Cauchy principal value of the naive definition. More precisely, suppose that all of the poles of
in
are at the points
, where we abuse notation to allow
to denote both the coset
is
as well as its representative
. Petersson constructed a punctured fundamental domain (
)
where
is a fundamental domain with
in the interior of
and
is the ball around
of hyperbolic radius
(see (
2.4)). He then defined the
regularized inner product betweenand and explicitly determined (see [
26, (6)]) that this regularization converges if and only if for all
and
5.2. Proof of Theorem 1.2: duality for meromorphic cusp forms
Proof of Theorem 1.2
The basic idea is to use the fact that, by Proposition 4.1, is orthogonal to cusp forms for and then compute the inner product in a second way, evaluating it as the sum of two elliptic coefficients. This method was used by Guerzhoy [18] to obtain duality results for Fourier coefficients of weakly holomorphic modular forms. In order to evaluate the inner product of meromorphic cusp forms as a sum of elliptic coefficients, we mimic calculations given in [14, Theorem 4.1] and [12].
To begin, for
and
, we use Theorem
4.3 to compute, using Stokes’ Theorem,
Rewriting
(
5.1) becomes
Stokes’ theorem together with invariance of the integrand under the action of
then yields
where
denotes the boundary of
. The differential
is invariant under
, and hence we may extend the integrals to precisely one copy of
, obtaining
where
Note that
for
. Hence, plugging in the elliptic expansions (
4.2) around
, of
and (
4.10) of
, we evaluate
where we abbreviate
and
. The integral gives
times the residue of the integrand at
, yielding
However, Lemma
2.1 (1) implies that as,
,
so that for
Therefore,
Plugging back into (
5.2) yields
This gives the claim after the change of variables
.□
Remark.
The orthogonality to cusp forms shown by Petersson can also be reproven directly either by rewriting as a constant multiple of or rewriting as a constant multiple of .
Funding
The research of the first author is supported by the Alfried Krupp Prize for Young University Teachers of the Krupp foundation, and the research leading to these results receives funding from the European Research Council under the European Union’s Seventh Framework Programme (FP/2007–2013)/ERC Grant agreement no. 335220—AQSER. This work was partially supported by a grant from the Simons Foundation (#281876 to P.J.). The research of the third author was supported by grants from the Research Grants Council of the Hong Kong SAR, China (Project nos. HKU 27300314, 17302515 and 17316416).
Acknowledgements
We thank the referees for many helpful comments which helped to improve this paper.
References
1A.
Beardon
,
The geometry of discrete groups
, Grad. Texts in Math.
91
,
Springer
,
New York
,
1995
.
2B.
Berndt
, P.
Bialek
and A.
Yee
,
Formulas of Ramanujan for the power series coefficients of certain quotients of Eisenstein series
,
Int. Math. Res. Not.
(
2002
),
1077
–
1109
.
3P.
Bialek
, Ramanujan’s formulas for the coefficients in the power series expansions of certain modular forms, ProQuest LLC, Ann Arbor, MI,
1995
, Ph.D. thesis, University of Illinois at Urbana-Champaign.
4R.
Borcherds
,
Automorphic forms with singularities on Grassmannians
,
Invent. Math.
132
(
1998
),
491
–
562
.
5K.
Bringmann
, A.
Folsom
, K.
Ono
and L.
Rolen
,
Harmonic Maass forms and mock modular forms: theory and applications
,
American Mathematical Society Colloquium Publications
,
Providence, R.I., USA
,
2017
.
6K.
Bringmann
, P.
Guerzhoy
and B.
Kane
,
Shintani lifts and fractional derivatives of harmonic weak Maass forms
,
Adv. Math.
255
(
2014
),
641
–
671
.
7K.
Bringmann
and B.
Kane
,
A problem of Petersson about weight 0 meromorphic modular forms
,
Res. Math. Sci.
3
(
2016
),
31
.
8K.
Bringmann
and B.
Kane
, Ramanujan-like formulas for Fourier coefficients of all meromorphic cusp forms, submitted for publication.
9K.
Bringmann
and B.
Kane
,
Ramanujan and coefficients of meromorphic modular forms
,
J. Math. Pures. Appl.
107
(
2017
),
100
–
122
.
10K.
Bringmann
, B.
Kane
and W
Kohnen
,
Locally harmonic Maass forms and the kernel of the Shintani lift
,
Int. Math. Res. Not.
(
2015
),
3185
–
3224
.
11K.
Bringmann
, B.
Kane
, S.
Löbrich
, L.
Rolen
and K.
Ono
,
On divisors of modular forms
,
Adv. Math.
329
(
2018
),
541
–
554
.
12K.
Bringmann
, B.
Kane
and A.
von Pippich
, Regularized inner products of meromorphic modular forms and higher Green’s functions, Commun. Contemp. Math., accepted for publication.
13J.
Bruinier
, W.
Kohnen
and K.
Ono
,
The arithmetic of the values of modular functions and the divisors of modular forms
,
Compos. Math.
140
(
2004
),
552
–
566
.
14J.
Bruinier
, K.
Ono
and R.
Rhoades
,
Differential operators for harmonic weak Maass forms and the vanishing of Hecke eigenvalues
,
Math. Ann.
342
(
2008
),
673
–
693
.
15W.
Duke
and P.
Jenkins
,
On the zeros and coefficients of certain weakly holomorphic modular forms
,
Pure Appl. Math. Q
4
(
2008
),
1327
–
1340
.
16J.
Fay
,
Fourier coefficients of the resolvent for a Fuchsian group
,
J. Reine Angew. Math.
293–294
(
1977
),
143
–
203
.
17H.
Gross
and D.
Zagier
,
Heegner points and derivatives of L-series
,
Invent. Math.
84
(
1986
),
225
–
320
.
18P.
Guerzhoy
,
On weak harmonic Maass—modular grids of even integral weights
,
Math. Res. Lett.
16
(
2009
),
59
–
65
.
19G.
Hardy
and S.
Ramanujan
,
On the coefficients in the expansions of certain modular functions
,
Proc. Roy. Soc. A
95
(
1919
),
144
–
155
. Collected papers of Srinivasa Ramanujan, AMS Chelsea Publ., Providence, RI, 2000, 310–321.
20D.
Hejhal
, The Selberg trace formula for PSL2(R), Lecture Notes in Mathematics 1001,
1983
.
21M.
Hövel
, Automorphe Formen mit Singularitäten auf dem hyperbolischen Raum, Ph.D. thesis, Technische Universität Darmstadt,
2012
.
22A.
Mellit
, Higher Green’s functions for modular forms, Ph.D thesis. Rheinische Friedrich-Wilhelms-Universität Bonn,
2008
.
24H.
Petersson
,
Einheitliche Begründung der Vollständigkeitssätze für die Poincaréschen Reihen von reeller Dimension bei beliebigen Grenzkreisgruppen von erster Art
,
Abh. Math. Sem. Univ. Hmbg.
14
(
1941
),
22
–
60
.
25H.
Petersson
,
Konstruktion der Modulformen und der zu gewissen Grenzkreisgruppen gehörigen automorphen Formen von positiver reeller Dimension und die vollständige Bestimmung ihrer Fourierkoeffzienten, S.-B. Heidelberger
,
Akad. Wiss. Math. Nat. Kl.
(
1950
),
415
–
474
.
26H.
Petersson
,
Über automorphe Orthogonalfunktionen und die Konstruktion der automorphen Formen von positiver reeller Dimension
,
Math. Ann.
127
(
1954
),
33
–
81
.
27B.
Pioline
, Rankin-Selberg methods for closed string amplitudes, String-Math. 2013, Proc. Sympos. Pure Math. 88, Amer. Math. Soc., Providence, RI,
2014
, pp. 119–143.
28M.
Viazovska
, CM values of higher Green's functions, preprint.
29D.
Zagier
, Traces of singular moduli in ‘Motives, polylogarithms and Hodge Theory’, Lecture series 3, Intl. Press (
2002
), 209–244.
© The Author(s) 2019. Published by Oxford University Press. All rights reserved. For permissions, please e-mail:
[email protected]