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 H which satisfy modularity properties under SL2(Z), are annihilated by the associated hyperbolic weight k 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 H 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 i, but also an elliptic expansion
(1.1)
around each point ϱH in terms of powers of Xϱ(z)zϱzϱ¯, where cf,ϱ(n)C. Polar harmonic Maass forms have a more general elliptic expansion where the coefficients cf,ϱ(n) may additionally depend on rϱ(z)Xϱ(z); 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 Greens functions are modular functions on H×H 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
    (1.2)
    where A is an SL2(Z)-equivalence class of integral binary quadratic forms of negative discriminant−D.
  • Functions related to (1.2) (namely summing up all classes A of discriminant D) 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 SL2(Z)H and in particular found a formula for the reciprocal of the Eisenstein series E6. 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 2κ and 22κ are interconnected by certain differential operators which naturally occur in the theory of harmonic Maass forms. Here and throughout, κ is assumed to be an arbitrary integer, and we use k instead if there is some restriction on the weight (for instance, if we require kN). For κZ and kN, set
(1.3)
where z=x+iyH. If f satisfies weight 2κ modularity, then ξ2κ(f) is modular of weight 22κ, while if f satisfies weight 22k modularity, then D2k1(f) satisfies weight 2k modularity.
Given a polar harmonic Maass form of weight 2κ, one may eliminate the singularity at i and, if κ0, 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 2κpolar harmonic cusp form, a weight 2κ polar harmonic Maass form which vanishes at i if κN0 and is bounded at i if κN. We denote the subspace of such forms by H2κ. A canonical basis for this space may be defined by specifying the growth behavior near singularities in SL2(Z)H, which is given via principal parts at z; 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 kN>1 they indeed span H22k. Specifically, for each nN and zH, in (4.5) we construct the unique weight 22k polar harmonic cusp form P22k,nz with principal part
(1.4)
Here ωz is the size of the stabilizer of z in PSL2(Z) and C2k1,n is the constant defined in (2.7) below and explicitly computed as a quotient of factorials in (4.11). For nN0, the functions P22k,nz have non-meromorphic principal parts. We describe these explicitly in Theorem 4.3 below.
To understand the behavior of the basis elements P22k,nzH22k under the differential operators defined in (1.3), we define the subspace S2κH2κ consisting of meromorphic modular forms without poles at i, which we call meromorphic cusp forms. For each point z in the fundamental domain SL2(Z)H, we let H2κz (resp. S2κz) be the subspace of forms in H2κ (resp. S2κ) with singularities allowed only at z. For kN>1, Petersson (cf. [25, equation (5c.3)] or [26, Equation (21)]) defined a family of meromorphic Poincaré series Ψ2k,nz which form a natural canonical basis for the space S2kz. Specifically, for nN and a point zH, the function Ψ2k,nz 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
(1.5)

As shown in the next theorem, the action of the differential operators ξ22k and D2k1 give an additional natural splitting of the space S2kz into three subspaces, which we denote by D2kz, E2kz, and the space of cusp forms S2k. Again using the regularization [26, Equation (3)], or its extension [12, Equation (3.3)] to arbitrary meromorphic cusp forms, the subspace D2kz (resp. E2kz) consists of those forms in S2kz which are orthogonal to cusp forms and whose principal parts are linear combinations of (1.5) with n2k (resp. 2k<n<0). The families Ψ2k,mz of meromorphic Poincaré series with m2k or with 2k<m<0 form bases for D2kz and E2kz, respectively. 

Theorem 1.1.

Suppose thatkN>1.

  1. EveryFH22kis a linear combination of the functions from{P22k,nz:zH,nZ}. Moreover, if the only poles ofFinHoccur at points equivalent tozunder the action ofSL2(Z), thenFis a linear combination of functions from{P22k,nz:nZ}.

  2. IfFH22kz, thenξ22k(F)D2kz (resp. ξ22k(F)S2k) if and only ifD2k1(F)S2k (resp. D2k1(F)D2kz).

  3. IfFH22kz, thenξ22k(F)E2kzif and only ifD2k1(F)E2kz.

  4. FornZandz=x+iyH, we have

 
Remark.

As alluded to before the theorem, Theorems 1.1 (2) and 1.1 (3) naturally split S2k into three subspaces D2kz, E2kz and S2k which are paired up as images under ξ22k and D2k1 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 CE2k spanned by the Eisenstein series is paired with itself in the same way as the space E2kz, while the space S2k 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 D2kz.

For z1,z2H, our next result gives a duality-type relationship between the coefficients of Ψ2k,mz1 and those of P22k,z2. To state it, for mN, let c2k,z2z1(m,n) denote the nth coefficient in the elliptic expansion (1.1) around ϱ=z2 of Ψ2k,mz1. Similarly, c22k,z1z2,+(m,n) is the nth coefficient in the meromorphic part of the elliptic expansion (see (2.11)) around ϱ=z1 of C2k1,m1P22k,mz2; in other words, by (1.4) these are the coefficients of the unique weight 22k polar harmonic Maass forms with principal parts
which closely resemble the principal parts (1.5) in positive weight. 
Theorem 1.2.
Form,nN0, we have
 
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 m,nN, take z1=z2=i and let c2k(m,n) denote the nth coefficient of the weight 2k weakly holomorphic modular form which grows toward i like e2πimz and let c22k+(m,n) be the nth coefficient of the holomorphic part of the weight 22k harmonic Maass form which grows towards i like e2πimz. 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 2k1 times. If kZ, 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 M=(acbd)SL2(Z), κZ, and f:HC, we define the usual slash operator
 
Definition.

For κZ, a polar harmonic Maass form of weight2κ is a function F:HC which is real analytic outside a discrete set of points and satisfies the following conditions:

  1. For every MSL2(Z), we have F2κM=F.

  2. We have Δ2κ(F)=0, with Δ2κ the weight2κhyperbolic Laplace operator
    (2.1)
  3. For every zH, there exists nN0 such that (zz)nF(z) is bounded in some neighborhood of z.

  4. The function F has at most linear exponential growth at i; that is, F(z)=O(eCy) for some constant CR+ (uniform in x for y sufficiently large) as y.

If (2) is replaced by Δ2κ(F)=λF, then F is called a polar Maass form with eigenvalue λ.

Denote by H2κ the space of polar harmonic Maass forms of weight 2κ. The subspace of H2κ consisting of forms that map under ξ2κ to cusp forms is denoted by H2κcusp; more generally, we add the superscript ‘cusp’ to any subspace of H2κ to indicate the space formed by taking the intersection of the subspace with H2κcusp. We also use the superscript z to indicate the subspace of forms whose only singularity in SL2(Z)H appears at z.

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 i. 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 F of weight 22k<0 and yF1, we have
where, for some cF±(n)C, we define the holomorphic partF+ (resp. non-holomorphic partF) of F at i as
(2.2)
with the incomplete gamma function Γ(α,w)wettα1dt. The sum of all of the terms which grow towards i is called the principal part of F.
We next consider elliptic expansions of polar harmonic Maass forms. Rather than expansions in e2πiz, the natural expansions of polar harmonic Maass forms around ϱ are given in terms of Xϱ(z). We further write
(2.3)
with d(z,ϱ) the hyperbolic distance between z and ϱ. The second identity in the definition of rϱ(z) follows by the well-known formula (see [1, p. 131])
(2.4)
where throughout the paper ηIm(ϱ). From (2.4), for MSL2(Z) one also immediately obtains the invariance
(2.5)
For 0Z<1 and aN and bZ, we also require the function
(2.6)
where
is the incomplete beta function and
(2.7)
Note that by [23, 8.17.7], we have
(2.8)
where F12 is the Gauss hypergeometric function defined by
with (a)na(a+1)(a+n1). We often use the fact that
(2.9)
We also require the Euler transformation (see 15.8.1 of [23])
(2.10)

The modified incomplete β-function β0 may also be written in special cases as a hypergeometric function, as can be seen by a direct calculation. 

Lemma 2.1.

Assume that0Z<1, aN, andbZ.

  1. We have
    Here and throughout we use the notationδS=1if some propertySis true and0otherwise.
  2. Ifb>0, then

We have the following elliptic expansion of weight 22k harmonic functions, whose proof is deferred to Section 3. 

Proposition 2.2.

Suppose thatkNandϱH.

  1. IfFsatisfiesΔ22k(F)=0and for somen0Nthe functionrϱn0(z)F(z)is bounded in some neighborhoodNaroundϱ, then there existcF,ϱ±(n)C, such that forzNandnkmin(2k2,n0), we have
    (2.11)
  2. IfFH22k, then the sum in (2.11) only runs over thosenwhich satisfynk1(modωϱ). IfFH22kcusp, then the second sum is empty and the third sum only runs overn<0.

 
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 ξ22k nor D2k1. 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 i, which is a constant multiple of y2k1, is annihilated by neither operator, and also exhibits a logarithmic singularity.

For F annihilated by Δ22k (with kN), we define the meromorphic part of the elliptic expansion (2.11) around ϱ 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 ξ22k and D2k1. 

Proposition 2.3.
ForkNandF:HCsatisfyingΔ22k(F)=0, we have
and

In addition to the operators ξ2κ and D2k1 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 2. Moreover,
(2.12)
We also require iterated raising
For kN, the raising operator and D2k1 are related by Bols identity
(2.13)

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 g:HC transforming for M=(acbd)SL2(Z) as
Then f(z)yκg(z) transforms as
Define the operator
By [16, p. 144], for M=(acbd)SL2(R), we have
Let Fκ,s denote the space of g:HC satisfying the following conditions:
  1. g(Mz)=(cz+dcz¯+d)κg(z);

  2. Dκ(g)=s(s1)g;

  3. g has at most finitely many singularities of finite order in SL2(Z)H¯, where H¯HQ{i}.

Functions in Fκ,s are closely related to polar Maass forms. In order to study the relationship between Dκ acting on Fay’s functions and Δ2κ acting on polar Maass forms, we require the following variants of the Maass raising and lowering operators (see [16, (3)]):
Note that Kκ sends Fκ,s to Fκ+1,s and Lκ sends Fκ,s to Fκ1,s. Moreover (see [16, (7)])
(2.14)
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.

  1. FornN0, we have
    (2.15)
    (2.16)
    (2.17)
    IfgFκ,s, then
    (2.18)
    In particular, fis harmonic if and only ifgFκ,κorgFκ,1κ.
  2. The function gFκ,s if and only if the function f is a polar Maass form of weight 2κ with eigenvalue (sκ)(1sκ). In particular, if gFκ,κ or gFκ,1κ and grows at most like yκ for y, then fH2κ.

 
Proof
(1) First, it is not hard to see that
(2.19)
Iterating (2.19) yields (2.15). Similarly, to prove (2.16), one first shows that
(2.20)
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 gFκ,s. Then, by (2.17), we have

(2) Part (1) implies that the eigenfunction properties of f and g 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 einθz(z) with θz(z)R satisfying Xz(z)=rz(z)eiθz(z). For sC, κR, and z,zH, these are given by (see [16, p. 147], slightly modified)
with
(2.21)
(2.22)
where for nZ we set sgn*(n)sgn(n+1/2).
These functions are meromorphic in s with at most simple poles at s±κN0 and satisfy certain useful relations. Directly from the definitions (2.21) and (2.22), one obtains
Moreover, for tR, we have
The special values of P and Q in the cases s=κ and s=1κ play an important role in our investigation. To describe these, we set
(2.23)
In the next section, we prove the following lemma. 
Lemma 2.5.

  1. ForκN0, we have
    (2.24)
    (2.25)
  2. ForκN and nN, we have
    (2.26)
  3. ForκNandnN0, we have
    (2.27)

We next define certain Poincaré series considered by Fay. For this, we set (see [16, (44)]), slightly modified)
(2.28)
with
(2.29)
where min(m,n) and
(2.30)
 
Remarks.

  1. Note that if cs,κ2m2,n2=0, then we multiply both sides of (2.28) by an appropriate factor to cancel the simple poles occurring in the Γ-factors and then take the limit, as in Lemma 2.5 (3).

  2. The functions Gs,κm,n satisfy the symmetry relations
    (2.31)
    (2.32)

Fay related these functions to the resolvent kernel Gs,κGs,κ0,0. 

Theorem 2.6.
(Fay [16, Theorem 2.1]). ForRe(s)>1,zGs,κm,n(z,z)Fκ+m,sandzGs,κm,n(z,z)Fκ+n,s. We have, form,nN0,
Ifmorn<0, then we replaceK,wj by L,wj.

Fay also considered elliptic expansions of functions in Fκ,s. 

Proposition 2.7.
(Fay [16, Theorem 1.1]). IfDκ(g)=s(s1)gin some annulusA:r1<d(z,ϱ)<r2aroundϱH, thenghas an elliptic expansion of the shape

The proof of Proposition 2.2, which we give in the next section, mostly relies on rewriting Fay’s functions Ps,κn and Qs,κn.

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 rrz(z), we have
(3.1)
For n0, Equation (2.24) follows from the definition, using (2.10), (2.9) and (3.1). If n<0 then, using (2.10), (3.1), and abbreviating XXz(z)=reiθ with θθz(z), the left-hand side of (2.24) equals
Since 2κ0 and n>0, the claim follows from Lemma 2.1 (2).

We next prove (2.25). The claim for n0 follows from the definition using (2.10), (2.8) and (3.1). For n<0, the claim follows by (2.8) and (3.1).

(2) From the definition of Qs,κn(z,z), the left-hand side of (2.26) equals
Once again using reiθ=X and (2.9), we obtain
By (3.1), we then obtain the claim.
(3) The left-hand side of (2.27) equals
Canceling Γ(sκ), using X=reiθ, taking the limit, employing (3.1), and plugging in the definition of the F12, we obtain
(3.2)
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 G(z)yκF(z), we have, by (2.17),
since Δ2κ(F)=0 by assumption. Thus, by Proposition 2.7 with s=1κ,
By Lemmas 2.5 (1) and (2.6), this gives
for some constants αϱ(n),γϱ(n)C. Rewriting yields the expansion (2.11) up to the restrictions on n in each of the sums. It thus remains to show that cF,ϱ+(n)=0 for n<n0 and cF,ϱ(n)=0 for n>n0. To do so, we investigate the asymptotic growth of each term in the sum as zϱ. We repeatedly use the fact that, as zϱ, Xϱ(z)ηzϱ, where by G1(z,ϱ)ηG2(z,ϱ) we mean that there is a constant Cη0 depending only on η such that limzϱG1(z,ϱ)CηG2(z,ϱ)=1. This gives that nn0 for the first summand in (2.11).
Moreover, by Lemma 2.1,
Thus, again using Xϱn(z)η(zϱ)n, we have, as zϱ,
(3.3)
Furthermore, for 0<n<nκ, since Xϱn(z)0 as zϱ, the asymptotic in (3.3) implies that we also have
This gives the claimed bounds for n0. Finally, the n=0 term behaves like Log(zϱ) by (3.3). This growth is canceled upon multiplying by rϱn0(z).
(2) By (2.3) and (2.5), for M in the stabilizer ΓϱSL2(Z) of M
(3.4)
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),
(3.5)
We rewrite the right-hand side of (3.5) in terms of the iterated operators (for N0 and κZ)
where (see [16, after formula (18)])
Namely, using (see [16, (14)]) that for f:R0+C
and iteratively carrying out the differentiation on θ yields
(3.6)
By (3.5) and (2.24), we thus have, for n0,
(3.7)
By [16, (18)], we know that
(3.8)
(3.9)
where
Plugging (3.8) into the right-hand side of (3.7) simplifies to
(3.10)
We split into the cases n2k1 and n<2k1.
For n2k1 we have, using (2.9),
Thus (3.10) becomes, using that rϱ(z)eiθϱ(z)=Xϱ(z) and (3.1),
Explicitly computing the constant then finishes the claim for n2k1. For 0n2k2, we have ek,k1(n+2k2)=0, giving the claim in this range.
We next act by Dz2k1 on the non-meromorphic part of F. First assume that n[0,2k2]. By Lemma 2.1 (1), we then have, using that rϱ2(z)=Xϱ(z)Xϱ(z)¯,
This is a polynomial in z of degree at most 2k2 (with antiholomorphic coefficients depending on z¯). Differentiating 2k1 times hence annihilates these terms.
It remains to determine the image of D2k1 on the terms in the non-meromorphic part with 0n2k2. Using (2.25), (3.5), (3.6), (3.9) and (2.26), we obtain that
Computing
and plugging in (2.23), noting that n0, yields the claimed formula.□

4. Poincaré series and the proof of Theorem 1.1

For zH, nZ and k>1, we define the meromorphic Poincaré series, due to Petersson,
(4.1)
where
We use the convention that z appears as a superscript in the notation if we consider zH as a fixed point and we write it as a two-variable function if we consider the properties in the z-variable. The main properties of Ψ2k,nz 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{Ψ2k,nz:zH,nZ} (resp. {Ψ2k,nz:nZ})spanS2k (resp. S2kz). Forn0, they are cusp forms. Forn<0they are orthogonal to cusp forms and have the principal part2ωzψ2k,n(z,z)aroundz=z.

 
Remarks.

  1. By Proposition 4.1, the elliptic expansion of Ψ2k,nz around ϱH may be written
    (4.2)
    where [z] denotes the SL2(Z)-equivalence class of z.
  2. As pointed out in [25, p. 72], zy2k+nΨ2k,nz(z) is modular of weight 2k2n. Moreover, it is an eigenfunction under Δ2k2n,z with eigenvalue (2k+n)(n+1).

We next write Ψ as a special case of Fay’s function G. We set
 
Lemma 4.2.

  1. We have
  2. IfnN0, then
    IfnN, then

 
Proof
(1) By inspecting the definitions (2.28) and (4.1), the claim follows once we show that
(4.3)
By definition (2.30), we have
If n<2k1, then we may plug in s=k directly and then use Lemma 2.5 (2) to obtain
For n0, we have ck,1kn,2k1=(1) implying (4.3) in this case.
For 0<n<2k1, we obtain (4.3) for n<2k1, computing

If n2k1, then =2k1 in (2.29) and we use Lemma 2.5 (3) (replacing n by n2k+1) to obtain the desired formula.

(2) By (2.31), we have
(4.4)
For n0, we may then directly plug in s=k and use (1) to obtain the claim by simplifying, with nk2n,
For n<0, 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 nZ the following polar harmonic Maass Poincaré series
(4.5)
with, using (2.25),
(4.6)

The following more precise version of Theorem 1.1 shows how the functions P22k,nz are related to the functions Ψ2k,nz via differential operators. 

Theorem 4.3.
AssumekN>1. The functions{P22k,nz:zH,nZ} (resp. {P22k,nz:nZ}) span the spaceH22k (resp. H22kz). Moreover,
(4.7)
(4.8)
The functionsP22k,nzvanish unlessnk1(modωz), in which case their principal parts equal
 
Remarks.
  1. It is also natural to ask about the properties of zP22k,nz(z) for fixed z. To investigate this, note that by a comparison of Definitions (4.5) and (2.28), we find that
    (4.9)
    with a1k,n given in (2.23), and we evaluate ck,1kn,0=1 via (2.29). Combining this with Theorem 2.6, one can conclude that the function zyn+22kP22k,nz(z) has weight 2k22n and eigenvalue (n+1)(n+22k) under Δ2k22n,z.
  2. By Theorem 4.3 and Proposition 2.2, one may write the elliptic expansion of P22k,nz around ϱ for n<0 and zH as
    (4.10)
Furthermore, using (2.7), Lemma 2.1 (1) and then Lemma 2.1 (2) yields
(4.11)
giving the constant in (1.4) in terms of factorials.
 
Proof of Theorem 4.3

Using (2.28), we conclude that P22k,nz satisfies weight 22k modularity.

The principal part of P22k,nz around z comes from the terms MΓz in (4.5). Vanishing of the principal part for nk1(modωz) follows from (3.4) together with [25, (2a.16)]. For nk1(modωz), this yields 2ωzφ22k,nz. For 0n2k2, this directly yields the principal part. For n>2k2 or n<0, we rewrite the incomplete beta function using (2.6) and note that for n>2k2 only the non-meromorphic part grows as z approaches z, while for n<0 only the meromorphic part grows.

Since every possible principal part in the elliptic expansion of an element of H22kz is obtained as a linear combination of the Poincaré series P22k,nz(nZ), these span the space H22kz. Moreover, eliminating the principal parts at different points in H one at a time implies that the space H22k is spanned by {P22k,nz:zH,nZ}.

We next compute the image of the Poincaré series under D2k1. Using (4.9) and (2.15), we obtain
(4.12)
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 Ck,n and a1k,n and use (2.13) to conclude (4.7).
It remains to compute the image under ξ22k. First, by Proposition 2.4 (1), with f:HC, 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 n0, we then use Lemma 4.2 (2) to obtain
We then simplify the factor in front using (2.23) to obtain the claim for n0. For n<0, 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 Ψ2k,nzS2k if and only if n0 and Ψ2k,nzD2kz if and only if n2k. These claims about Ψ2k,nz 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 Ψ2k,nzE2kz if and only if 2k<n<0.

(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 f,gS2k in SL2(Z)H are at the points z1,...,zr, where we abuse notation to allow z to denote both the coset [z] is SL2(Z)H as well as its representative zH. Petersson constructed a punctured fundamental domain (ε>0)
where F* is a fundamental domain with z in the interior of ΓzF* and Bε(z) is the ball around z of hyperbolic radius ε (see (2.4)). He then defined the regularized inner product betweenfandg
and explicitly determined (see [26, (6)]) that this regularization converges if and only if for all n<0 and ϱH

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, Ψ2k,nz is orthogonal to cusp forms for n<0 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 z1,z2H and n<0m, we use Theorem 4.3 to compute, using Stokes’ Theorem,
(5.1)
Rewriting
(5.1) becomes
Stokes’ theorem together with invariance of the integrand under the action of SL2(Z) then yields
where Bε(z) denotes the boundary of Bε(z). The differential Ψ2k,nz1(z)P22k,m1z2(z)dz is invariant under Γzj, and hence we may extend the integrals to precisely one copy of Bεj(zj), obtaining
(5.2)
where
Note that rϱ(z)=ε for zBε(ϱ). Hence, plugging in the elliptic expansions (4.2) around ϱ=zj, of Ψ2k,nz1 and (4.10) of P22k,m1z2, we evaluate
where we abbreviate X=Xϱ(z) and r=rϱ(z). The integral gives 2πi times the residue of the integrand at z=ϱ, yielding
However, Lemma 2.1 (1) implies that as, ε0,
so that for +1>0
Therefore,
Plugging back into (5.2) yields
This gives the claim after the change of variables nn1.□
 
Remark.

The orthogonality to cusp forms shown by Petersson can also be reproven directly either by rewriting Ψ2k,nz1 as a constant multiple of ξ22k(P22k,n1z1) or rewriting Ψ2k,mz2 as a constant multiple of D2k1(P22k,m+2k1).

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

1

A.
Beardon
,
The geometry of discrete groups
, Grad. Texts in Math.
91
,
Springer
,
New York
,
1995
.

2

B.
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
.

3

P.
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.

4

R.
Borcherds
,
Automorphic forms with singularities on Grassmannians
,
Invent. Math.
132
(
1998
),
491
562
.

5

K.
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
.

6

K.
Bringmann
,
P.
Guerzhoy
and
B.
Kane
,
Shintani lifts and fractional derivatives of harmonic weak Maass forms
,
Adv. Math.
255
(
2014
),
641
671
.

7

K.
Bringmann
and
B.
Kane
,
A problem of Petersson about weight 0 meromorphic modular forms
,
Res. Math. Sci.
3
(
2016
),
31
.

8

K.
Bringmann
and
B.
Kane
, Ramanujan-like formulas for Fourier coefficients of all meromorphic cusp forms, submitted for publication.

9

K.
Bringmann
and
B.
Kane
,
Ramanujan and coefficients of meromorphic modular forms
,
J. Math. Pures. Appl.
107
(
2017
),
100
122
.

10

K.
Bringmann
,
B.
Kane
and
W
Kohnen
,
Locally harmonic Maass forms and the kernel of the Shintani lift
,
Int. Math. Res. Not.
(
2015
),
3185
3224
.

11

K.
Bringmann
,
B.
Kane
,
S.
Löbrich
,
L.
Rolen
and
K.
Ono
,
On divisors of modular forms
,
Adv. Math.
329
(
2018
),
541
554
.

12

K.
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.

13

J.
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
.

14

J.
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
.

15

W.
Duke
and
P.
Jenkins
,
On the zeros and coefficients of certain weakly holomorphic modular forms
,
Pure Appl. Math. Q
4
(
2008
),
1327
1340
.

16

J.
Fay
,
Fourier coefficients of the resolvent for a Fuchsian group
,
J. Reine Angew. Math.
293–294
(
1977
),
143
203
.

17

H.
Gross
and
D.
Zagier
,
Heegner points and derivatives of L-series
,
Invent. Math.
84
(
1986
),
225
320
.

18

P.
Guerzhoy
,
On weak harmonic Maass—modular grids of even integral weights
,
Math. Res. Lett.
16
(
2009
),
59
65
.

19

G.
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.

20

D.
Hejhal
, The Selberg trace formula for PSL2(R), Lecture Notes in Mathematics 1001,
1983
.

21

M.
Hövel
, Automorphe Formen mit Singularitäten auf dem hyperbolischen Raum, Ph.D. thesis, Technische Universität Darmstadt,
2012
.

22

A.
Mellit
, Higher Green’s functions for modular forms, Ph.D thesis. Rheinische Friedrich-Wilhelms-Universität Bonn,
2008
.

23

NIST Digital Library of Mathematical functions. http://dlmf.nist.gov.

24

H.
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
.

25

H.
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
.

26

H.
Petersson
,
Über automorphe Orthogonalfunktionen und die Konstruktion der automorphen Formen von positiver reeller Dimension
,
Math. Ann.
127
(
1954
),
33
81
.

27

B.
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.

28

M.
Viazovska
, CM values of higher Green's functions, preprint.

29

D.
Zagier
, Traces of singular moduli in ‘Motives, polylogarithms and Hodge Theory’, Lecture series 3, Intl. Press (
2002
), 209–244.

This article is published and distributed under the terms of the Oxford University Press, Standard Journals Publication Model (https://academic.oup.com/journals/pages/open_access/funder_policies/chorus/standard_publication_model)