Skip to main content
Springer logoLink to Springer
. 2024 Jul 13;391(1):791–817. doi: 10.1007/s00208-024-02927-4

Harmonic weak Maass forms and periods II

Claudia Alfes 1,, Jan Hendrik Bruinier 2, Markus Schwagenscheidt 3
PMCID: PMC11700918  PMID: 39776533

Abstract

In this paper we investigate the Fourier coefficients of harmonic Maass forms of negative half-integral weight. We relate the algebraicity of these coefficients to the algebraicity of the coefficients of certain canonical meromorphic modular forms of positive even weight with poles at Heegner divisors. Moreover, we give an explicit formula for the coefficients of harmonic Maass forms in terms of periods of certain meromorphic modular forms with algebraic coefficients.

Introduction

A fundamental result in the theory of modular forms is the fact that the spaces of holomorphic modular forms of fixed weight possess bases of forms with integral Fourier coefficients. In recent years, many authors have studied the Fourier coefficients of non-holomorphic generalizations of modular forms, such as harmonic Maass forms. These functions transform like modular forms but are harmonic rather than holomorphic on the upper half-plane, and they are allowed to have poles at the cusps. The most prominent examples appear in connection with mock theta functions, which are certain q-series with (typically) integral coefficients that have asymptotic expansions like modular forms as q approaches roots of unity, but are not modular. More precisely, Zwegers [36] showed that Ramanujan’s mock theta functions may be viewed as the holomorphic parts of certain harmonic Maass forms of weight 12. This raises the question for the algebraicity of the coefficients of the holomorphic parts of more general harmonic Maass forms.

As it turns out, apart from the mock theta functions (see also [6, 14]) and certain harmonic Maass forms related to newforms with complex multiplication (see [11]), these coefficients typically do not seem to be algebraic. For example, in [8], Ono and the second author related the algebraicity of the Fourier coefficients of the holomorphic parts of certain harmonic Maass forms of weight 12 to the vanishing of the central values of derivatives of L-functions of newforms of weight 2. Moreover, in [13] it was shown that the Fourier coefficients of these weight 12 harmonic Maass forms can be expressed in terms of periods of certain algebraic differentials of the third kind on modular curves, which also hints at their delicate algebraic nature. In the present work, we extend the results of [13] to harmonic Maass forms of weight 32-k with k1, and derive an expression for their Fourier coefficients in terms of periods of certain meromorphic modular forms of weight 2k.

Let us describe our main results in some more detail. To simplify the exposition we let Γ:=SL2(Z) throughout the introduction. In the body of the paper we will treat level Γ0(N) for arbitrary NN, using the language of vector-valued modular forms for the Weil representation.

Canonical meromorphic modular forms associated with divisors on modular curves

If X is a non-singular projective curve over C and D is a divisor on X whose restriction to any component of X has degree 0, then it follows from the Riemann period relations that there exists a unique meromorphic differential ηD on X with at most simple poles and residue divisor PXresP(ηD)P=D, and such that

CηD=0

for any closed path C in X avoiding the points of D. The differential ηD is called the canonical differential of the third kind associated with D. If X is a modular curve, then ηD may be viewed as a meromorphic modular form of weight 2. Our first result is a construction of certain ‘canonical’ meromorphic modular forms ηk,D of weight 2k associated with divisors D on modular curves, generalizing the notion of canonical differentials to higher weight.

For a subfield FC and an integer kZ with k2 we let D2k,F(Γ) be the F-vector space of meromorphic modular forms f(z) defined over F of weight 2k for Γ which vanish at the cusp of Γ and which have poles of order at most k on H, with expansions of the form

f(z)=af,ϱ(z-ϱ)(z-ϱ¯)ϱ-ϱ¯-k+O(1),(zϱ),

at each pole ϱ of f, with some constants af,ϱF. Note that for k=1 we have af,ϱ=resϱ(f). Hence, by a slight abuse of notation, we will call the F-linear combination

res(f):=[ϱ]Γ\Haf,ϱwϱ[ϱ]Div(Γ\H)F

of points in Γ\H the residue divisor of fD2k,F(Γ). Here wϱ:=12|Γϱ| is half the order of the stabilizer of ϱ in Γ, and Div(Γ\H)F denotes the space of all F-linear combinations of points on Γ\H. Note that the sum is well-defined, that is, af,ϱ and wϱ only depend on the class [ϱ] of ϱ in Γ\H.

It is well known that closed geodesics on Γ\H correspond to (conjugacy classes of) primitive hyperbolic matrices in Γ. More precisely, given such a primitive hyperbolic matrix γΓ, we let Sγ be the semi-circle in H connecting the two real fixed points of Γ. Then Cγ:=γ\Sγ defines a closed geodesic in Γ\H, and every closed geodesic in Γ\H is accounted for in this way. We let CR(Γ) be the real vector space spanned by all R-linear combinations of closed geodesics Cγ on Γ\H. We obtain a real-valued bilinear pairing on D2k,R(Γ)×CR(Γ) by setting, for each meromorphic modular form fD2k,R(Γ) and each closed geodesics Cγ on Γ\H,

(f,Cγ):=idγk-1z0γz0f(z)Qγ(z,1)k-1dz, 1.1

where Qγ(x,y):=cx2+(d-a)xy-by2 denotes the binary quadratic form corresponding to γ=abcd, dγ:=tr(γ)2-4 is the discriminant of Qγ, and the path of integration is any path in H (avoiding the poles of f) from some point z0H (not being a pole of f) to γz0. Using the residue theorem and the assumption that the coefficients af,ϱ are real numbers, one can show that the value of the pairing (f,Cγ) is indeed independent of the choice of z0H and the path from z0 to γz0, see Lemma 3.1 below.

We have the following higher weight analogues of canonical differentials associated with divisors on Γ\H.

Proposition 1.1

For each divisor DDiv(Γ\H)R there exists a unique meromorphic modular form ηk,DD2k,R(Γ) with residue divisor res(ηk,D)=D and (ηk,D,C)=0 for any CCR(Γ).

We call ηk,D the canonical meromorphic modular form of weight 2k associated with D. We will construct it explicitly as a linear combination of certain two-variable Poincaré series introduced by Petersson [27]. The fact that their cycle integrals are all real or all purely imaginary (depending on the parity of k) can then be shown by a direct computation, following ideas of Katok [23]. Moreover, the uniqueness of ηk,D essentially follows from the Eichler-Shimura isomorphism. We refer the reader to Sect. 3 for the details of the proof of Proposition 1.1.

In the case that D is a CM divisor, a result similar to Proposition 1.1 has been proved by Mellit [26, Theorem 1.5.3].

Remark 1.2

The above constructions may be interpreted in a more geometric way using local coefficient systems as explained in [19]. In this setting, meromorphic modular forms of weight 2k give rise to meromorphic 1-forms on Γ\H with coefficients in the local system E associated with Sym2k-2(C2), geodesics in Γ\H give rise to cycles with coefficients in the dual of E, and the pairing (1.1) may be viewed as a (co)homological pairing.

Harmonic Maass forms and normalized meromorphic modular forms

The main topic of this work is the study of canonical meromorphic modular forms of weight 2k for certain linear combinations of Heegner divisors, coming from harmonic Maass forms of half-integral weight. Let us recall the setup from [8]. For kN with k2 we let GS2k be a normalized Hecke eigenform of weight 2k for Γ, and let FG be the totally real number field generated by the Fourier coefficients of G. We let gS12+k be a Hecke eigenform of weight 12+k for Γ0(4) in the Kohnen plus space which corresponds to G under the Shimura correspondence. We may normalize g such that all its Fourier coefficients lie in FG. Then, by [8, Lemma 7.3] there exists a harmonic Maass form f of weight 32-k for Γ0(4) with principal part defined over FG such that ξ32-k(f)=g-2g, where g:=(g,g) denotes the Petersson norm of g and ξκ=2ivκτ¯¯. Every such harmonic Maass form f can be written as a sum f=f++f- of a holomorphic part f+ and a non-holomorphic part f-, with Fourier expansions of the shape

f+(τ)=nZn-cf+(n)qn,f-(τ)=nZn<0cf-(n)Γk-12,4π|n|vqn,

with coefficients cf±(n)C, where q:=e2πiτ and Γ(s,x):=xe-tts-1dt denotes the incomplete Gamma function. The condition that the principal part of f is defined over FG means that we have cf+(n)FG for all n0. We let H32-k(FG) denote the space of harmonic Maass forms of weight 32-k for Γ0(4) satisfying the Kohnen plus space condition and with principal part defined over FG. We are interested in the algebraicity properties of the coefficients cf+(n) for n>0.

Now we come to the construction of the Heegner divisor corresponding to f. For a negative discriminant d<0 we let Qd+ be the set of positive definite integral binary quadratic forms Q(x,y)=ax2+bxy+cy2 of discriminant d=b2-4ac. For QQd+ the quadratic polynomial Q(z, 1) has a unique root αQ in H, called the CM point (or Heegner point) corresponding to Q. Let ΔZ be a fundamental discriminant with (-1)kΔ<0 and let D<0 be an integer such that |Δ|D0,1(mod4). Moreover, let χΔ be the usual genus character on QD|Δ|+. Then we define the twisted Heegner divisor

ZΔ(D):=QΓ\QD|Δ|+χΔ(Q)wQ[αQ],

where wQ:=wαQ. Eventually, we let

Zk,Δ(f):=D<0cf+(D)(|DΔ|)k-12ZΔ(D)

be the twisted Heegner divisor corresponding to f, where cf+(D)FG denote the coefficients of the holomorphic part of f. We remark that for even k the Heegner divisor Zk,Δ(f) is in general not a divisor with coefficients in FG, but with coefficients in some totally real algebraic extension of FG.

The following result relates the algebraicity of the coefficients cf+(n) of the harmonic Maass form f to the algebraicity of the Fourier coefficients of the canonical meromorphic modular form of weight 2k for the Heegner divisor Zk,Δ(f). It (partly) generalizes [8, Theorem 5.5, Theorem 7.6] to higher weight.

Theorem 1.3

Let fH32-k(FG) be as above, and let ηk,Δ(f)D2k,R(Γ) be the canonical meromorphic modular form of weight 2k for Zk,Δ(f). Then the following are equivalent.

  1. We have cf+(|Δ|)FG.

  2. We have cf+(n2|Δ|)FG for all nN.

  3. All Fourier coefficients of ηk,Δ(f) are contained in iπkΔFG.

In order to prove the theorem we will construct ηk,Δ(f) as a regularized theta lift of f, extending the methods of [8] by including certain differential (weight raising) operators in the theta lift, similarly as in [4]. By computing the Fourier expansion of this lift, we obtain a formula for the coefficients of ηk,Δ(f) in terms of the coefficients cf+(n2|Δ|),nN. Invoking the action of Hecke operators on f and ηk,Δ(f) then yields the stated result. We refer to Sect. 5 for the generalization of Theorem 1.3 to higher level Γ0(N) and its proof.

We remark that a similar theta lift has been studied by Zemel [34], and it was used to show the implication (2)(3) in Theorem 1.3 (for Δ=1).

Finally, we obtain a formula for the coefficient cf+(|Δ|) in terms of cycle integrals of a certain normalized meromorphic modular form in D2k,R(Γ) with residue divisor Zk,Δ(f), generalizing [13, Theorem 1.1] to higher weight.

Theorem 1.4

For fH32-k(FG) as above, there exists a unique meromorphic modular form ζk,Δ(f)D2k,R(Γ) with the following properties.

  1. The residue divisor is given by res(ζk,Δ(f))=Zk,Δ(f).

  2. We have (ζk,Δ(f),C)=0 for every CCR(Γ) with (G,C)=0.

  3. The first Fourier coefficient of ζk,Δ(f) vanishes.

Moreover, all Fourier coefficients of ζk,Δ(f) lie in iπkΔFG, and we have

cf+(|Δ|)=-1Ck,ΔiπkΔ·(ζk,Δ(f),C)(G,C) 1.2

for every CCR(Γ) with (G,C)0, with the rational constant Ck,Δ=(-2)k|Δ|k-1(k-1)!.

Remark 1.5

The formula from Theorem 1.4 tells us that the coefficient cf+(|Δ|) can be written as a quotient of (real parts of) cycle integrals of (meromorphic) modular forms with coefficients in FG.

We call ζk,Δ(f) the normalized meromorphic modular form of weight 2k for Zk,Δ(f). We will construct ζk,Δ(f) by subtracting a suitable multiple of G from ηk,Δ(f). The fact that the coefficients ζk,Δ(f) lie in iπkΔFG and the formula (1.2) again follow by writing ηk,Δ(f) as a theta lift of f, and using the action of Hecke operators on ζk,Δ(f). The details of the proof can be found in Sect. 5.

Organization of the paper

We start with a section on the necessary preliminaries about Heegner divisors, quadratic spaces and lattices, and vector-valued harmonic Maass forms for the Weil representation. In Sect. 4 we study a regularized theta lift of harmonic Maass forms which produces canonical meromorphic modular forms for twisted Heegner divisors, and we compute its Fourier expansion. In Sect. 5 we investigate the algebraicity properties of the Fourier coefficients of the canonical and normalized meromorphic modular forms for twisted Heegner divisors, and we prove Theorem 1.3 and Theorem 1.4 for level Γ0(N). Finally, in Sect. 6 we give a short outlook on a possible future application of our results to a non-vanishing criterion for central values of derivatives of newforms of weight 2k in terms of the algebraicity of Fourier coefficient of harmonic Maass forms.

Preliminaries

Heegner divisors

Let NN. For a negative discriminant d<0 and rZ/2NZ with dr2(mod4N) we consider the set of integral binary quadratic forms

Qd,r:={Q(x,y)=aNx2+bxy+cy2:a,b,cZ,d=b2-4Nac,br(mod2N)}.

The group Γ0(N) acts on Qd,r from the right, with finitely many orbits if D0.

We let Qd,r+ be the subset of positive definite quadratic forms in Qd,r, which are given by the condition sgn(Q):=sgn(a)>0. For Q=[a,b,c]Qd,r+ we let

αQ:=-b+i|d|2Na

be the Heegner point corresponding to Q. It is the unique root of Q(z, 1) in H. The order wQ:=12|Γ0(N)Q| of the stabilizer of Q in Γ0(N) is finite.

Let ΔZ be a fundamental discriminant and let ρZ/2NZ with Δρ2(mod4N). For DZ with D<0 we let χΔ be the generalized genus character on QD|Δ|,rρ as in [20], and we define the twisted Heegner divisor

ZΔ,ρ(D,r):=QQD|Δ|,rρ+/Γ0(N)χΔ(Q)wQ[αQ],

where [αQ] denotes the class of αQ in Γ0(N)\H.

Quadratic spaces and lattices

For NN we let V be the rational quadratic space of signature (2, 1) given by the set of rational traceless 2 by 2 matrices with the quadratic form q(X):=-Ndet(X) and the associated bilinear form (X,Y):=Ntr(XY). The group SL2(Q) acts as isometries on V by γ.X:=γXγ-1.

Throughout this work we will consider the lattice

L:=bc/N-a-b:a,b,cZ 2.1

in V. The dual lattice is given by

L=b/2Nc/N-a-b/2N:a,b,cZ.

Hence the discriminant form L/L is isomorphic to Z/2NZ with the finite quadratic form xx2/4N(modZ), and we will use this identification without further notice. Note that the group Γ0(N) acts on L and fixes the classes of L/L.

For rZ/2NZ with dr2(mod4N) we let

Ld,r:={XL:q(X)=d/4NandXr(modL)},

Each element X=b/2Nc/N-a-b/2NLd,r corresponds to a binary quadratic form

QX:=[aN,b,c]Qd,r

of discriminant d=4Nq(X). The group Γ0(N) acts on both Ld,r and Qd,r, and the actions are compatible in the sense that QXγ=Qγ-1.X for γΓ0(N). In particular, we obtain a bijection between Γ0(N)\Ld,r and Qd,r/Γ0(N).

We let Gr(L) be the Grassmannian of positive definite 2-dimensional subspaces of V(R). We can identify Gr(L) with H by sending z=x+iyH to the positive definite plane U(z):=X(z), where

X(z):=12Ny-x|z|2-1x.

Note that (X(z),X(z))=-1 and γ.X(z)=X(γz) for γSL2(R). We also define the vectors

U1(z):=12Nyx-x2+y21-x,U2(z):=12Nyy-2xy0-y,

such that U1(z),U2(z),X(z) form an orthogonal basis of LR. Note that U(z) is spanned by U1(z),U2(z). For XL and zH we consider the polynomials

QX(z):=2Ny(X,U1(z)+iU2(z))=aNz2+bz+c,pz(X):=-(X,X(z))=12Ny(aN|z|2+bx+c). 2.2

Then we have the useful rules

q(Xz)=14Ny2|QX(z)|2,q(Xz)=-12pz2(X). 2.3

Harmonic Maass forms

The metaplectic extension of SL2(Z) is defined as

Γ~:=Mp2(Z):=(γ,ϕ):γ=abcdSL2(Z),ϕ:HCholomorphic,ϕ2(τ)=cτ+d.

It is generated by T:=1101,1 and S:=0-110,τ. We let Γ~ be the subgroup generated by T. Moreover, for γ=abcdSL2(Z) we let γ~:=(γ,cτ+d)Γ~, where · denotes the principal branch of the square root.

We let L be the lattice from Sect. 2.2 and we let C[L/L]C[Z/2NZ] be its group ring, which is generated by the formal basis vectors er for rZ/2NZ. The Weil representation ρL associated with L is the representation of Γ~ on C[L/L] defined by

ρL(T)(er):=er24Ner,ρL(S)(er):=e-182Nr(2N)e-rr2Ner.

The Weil representation ρ¯L associated to the lattice L-=(L,-q) is called the dual Weil representation associated to L.

Let κZ+12 and define the slash-operator by

fκ,ρL(γ,ϕ)(τ):=ϕ(τ)-2κρL-1(γ,ϕ)f(γτ),

for a function f:HC[L/L] and (γ,ϕ)Γ~. Here, and in the following we let e(z):=e2πiz. Following [5], we call a smooth function f:HC[L/L] a harmonic Maass form of weight κ with respect to ρL if it is annihilated by the weight κ Laplace operator

Δκ:=-v22u2+2v2+iκvu+iv,

if it is invariant under the slash-operator κ,ρL(γ,ϕ) for all (γ,ϕ)Γ~, and if there exists a C[L/L]-valued Fourier polynomial (the principal part of f)

Pf(τ):=r(2N)D0cf+(D,r)eDτ4Ner

such that f(τ)-Pf(τ)=O(e-εv) as v for some ε>0. We denote the vector space of harmonic Maass forms of weight κ with respect to ρL by Hκ,ρL, and we let Mκ,ρL! be the subspace of weakly holomorphic modular forms. Every fHκ,ρL can be written as a sum f=f++f- of a holomorphic and a non-holomorphic part, having Fourier expansions of the form

f+(τ)=r(2N)D-cf+(D,r)eDτ4Ner,f-(τ)=r(2N)D<0cf-(D,r)Γ1-κ,π|D|vNeDτ4Ner,

where Γ(s,x):=xts-1e-tdt denotes the incomplete Gamma function.

The antilinear differential operator ξκ:=2ivκτ¯¯ maps a harmonic Maass form fHκ,ρL to a cusp form of weight 2-κ for ρ¯L. We further require the raising operator Rκ:=2iτ+κv, which raises the weight of a smooth function transforming like a modular form of weight κ for ρL by two, and we define the iterated raising operator Rκn:=Rκ+2n-2Rκ+2Rκ for nN, and Rκ0:=id.

Maass Poincaré series

Examples of harmonic Maass forms can be constructed using Maass Poincaré series. We recall their construction from [12, Sect. 1.3]. Let κZ+12 with κ<0. For sC and v>0 we let

Ms(v):=v-κ/2M-κ/2,s-1/2(v),

with the usual M-Whittaker function. Let DZ with D<0 and rZ/2NZ with Dr2(mod4N). For sC with (s)>1 we define the C[L/L]-valued Maass Poincaré series

Pκ,D,r(τ,s):=12Γ(2s)(M,ϕ)Γ~\Mp2(Z)Msπ|D|vNeDu4Ner|κ,ρL(M,ϕ),

where Γ~ is the subgroup of Mp2(Z) generated by T=1101,1. It converges absolutely for (s)>1, it transforms like a modular form of weight κ for ρL, and it is an eigenform of the Laplace operator Δk with eigenvalue s(1-s)+(κ2-2κ)/4. The special value

Pκ,D,r(τ):=Pκ,D,rτ,1-κ2

defines a harmonic Maass form in Hκ,ρL whose Fourier expansion starts with

Pκ,D,r(τ)=q-|D|4N(er+(-1)κ-12e-r)+O(1).

The following lemma follows inductively from [9, Proposition 2.2].

Lemma 2.1

For nN0 we have that

RκnPκ,D,r(τ,s)=π|D|NnΓs+n+κ2Γs+κ2Pκ+2n,D,r(τ,s).

Maass Poincaré series for the dual Weil representation ρ¯L are defined analogously, and Lemma 2.1 remains true for them. However, in this case, we have to require that D-r2(mod4N), and the Fourier expansion starts with q-|D|/4N(er-(-1)κ-12e-r)+O(1).

W-Whittaker functions

We record some facts about derivatives and integrals of W-Whittaker functions that will be used in the computations later on. Following [12, Sect. 1.3] we put

Wκ,s(y):=|y|-κ/2Wκ2sgn(y),s-12(|y|),(κR,sC,yR\{0}),

where Wν,μ(y) is the usual W-Whittaker function. At s=1-κ2 it simplifies to

Wκ,1-κ2(y)=e-y/2,if y > 0,e-y/2Γ(1-κ,|y|),if y < 0. 2.4

Moreover, using (13.4.33) and (13.4.31) in [3], we obtain the formula

RκWκ,s(4πmv)e(mu)=-4π|m|s+κ2s-κ2-1Wκ+2,s(4πmv)e(mu),if m < 0,-4π|m|Wκ+2,s(4πmv)e(mu),if m > 0, 2.5

for u+ivH and mR\{0}. Moreover, we will need the integral formula

0vκ-2Wκ,s(αv)exp-αv2-βvdv=α14-κ2βκ2-34πW0,32-2s4αβ 2.6

for α,β>0, which follows by evaluating the integral in terms of the K-Bessel function using [17, (22) on p. 217], and then writing 2z/πKμ(z)=W0,μ(2z), see [3, (13.6.21)].

Hypergeometric series

We compute the action of the iterated raising operator on a certain hypergeometric series for later use.

Proposition 2.2

For XV(R) with q(X)=m<0 and kN we have

R0,zk(2|m|pz2(X)k22F1k2,k+12;k+12;2|m|pz2(X))=Γ(2k)Γ(k)4N|m|sgn(pz(X))QX(z)k.

Proof

For brevity, we put w:=w(z):=2|m|pz2(X). We prove the proposition by induction. To this end, for jZ and fixed kN we define the function

fj(z):=y-2QX(z)¯jwk+j22F1k+j2,k+j+12;k+12;w.

Note that we want to compute R0kf0(z). We claim that

R2jfj(z)=sgn(pz(X))4N|m|(k+j)fj+1(z). 2.7

To prove this, we first use that for every holomorphic function g:HC, every smooth function h:HC, and every κ,R we have the simple relation

R-κyκg(z)¯·h(z)=yκg(z)¯·Rh(z).

Applying this with κ=-2j,=0, and g(z)=QXj(z), we obtain

R2jfj(z)=y-2QX(z)¯jR0wk+j22F1k+j2,k+j+12;k+12;w.

Next, we write R0=2iz and use the formula z(za2F1(a,b;c;z))=aza-12F1(a+1,b;c;z) (see (15.5.3) of NIST) to obtain

R0wk+j22F1k+j2,k+j+12;k+12;w=i(k+j)wk+j2-12F1k+j2+1,k+j+12;k+12;w·zw.

A direct computation shows that

zw=-i4N|m|sgn(pz(X))y-2QX(z)¯w32.

Finally, using 2F1(a,b;c;z)=2F1(b,a;c;z) and taking everything together, we obtain (2.7).

Note that the left-hand side in the proposition equals R0kf0(z). Applying (2.7) inductively, we find that R0kf0(z) is a multiple of fk(z). More explicitly, we get

R0kf0(z)=sgn(pz(X))k4N|m|k(2k-1)!(k-1)!y-2QX(z)¯kwk2F1k,k+12;k+12;w.

We have 2F1(a,b;b;z)=(1-z)-a which implies

2F1k,k+12;k+12;w=1-2|m|pz2(X)-k=pz2(X)k|QX(z)|22Ny2-k,

where we used that m=q(X)=q(Xz)+q(Xz)=14Ny2|QX(z)|2-12pz2(X), compare (2.3). Taking everything together, we obtain the formula in the proposition.

Canonical meromorphic modular forms of weight 2k

In this section we prove Proposition 1.1 for Γ:=Γ0(N) in a series of lemmas. Let kZ with k2. For completeness, we first show that the pairing in (1.1) is well-defined.

Lemma 3.1

For fD2k,R(Γ) and any closed geodesic Cγ on Γ\H the value of the pairing (f,Cγ) is independent of the choice of the path.

Proof

We have to show that, for any pole ϱH of f and any small closed loop ϱ around ϱ, we have

ik-1ϱf(z)Qγ(z,1)k-1=0. 3.1

This essentially follows from the residue theorem. For computational convenience, we will use the fact that every function F which is meromorphic near ϱ has an elliptic expansion in weight κZ of the shape

F(z)=(z-ϱ¯)-κn-cF,ϱ(n)Xϱ(z)n,Xϱ(z):=z-ϱz-ϱ¯,

with coefficients cF,ϱ(n)C, see Proposition 17 in Zagier’s part of [15]. Here κZ can be chosen freely, but the coefficients cF,ϱ(n)C will depend on κ. In particular, we do not need to require that F transforms like a modular form of weight κ.

If G is another meromorphic function near ϱ with an elliptic expansion in weight 2-κ and coefficients cG,ϱ(n), then it follows from the residue theorem (see [1, Lemma 4.1]) that

ϱF(z)G(z)dz=π(ϱ)nZcF,ϱ(n)cG,ϱ(-n-1). 3.2

By definition of the space D2k,R(Γ), the elliptic expansion (in weight 2k) of fD2k,R(Γ) near ϱ is of the shape

f(z)=af,ϱ(2i(ϱ))k(z-ϱ¯)-2kXϱ(z)-k+O(1), 3.3

as zϱ, with af,ϱR. Since Qγ(z,1)k-1 is holomorphic at ϱ, by (3.2) we only need its elliptic expansion coefficient (in weight 2-2k) of index k-1 at ϱ. By [1, Lemma 3.1, Lemma 5.4] this coefficient is given by

(-4(ϱ))1-k(k-1)!R2-2kk-1(Qγ(z,1)k-1)|z=ϱ=(-4(ϱ))1-k2idγk-1Pk-1i(A|ϱ|2+B(ϱ)+C)(ϱ)dγ, 3.4

where dγ is the discriminant of Qγ, P(x) denotes the -th Legendre polynomial, and we wrote Qγ=[A,B,C]. Using that P is even if is even and odd if is odd, we see that (3.4) is real. Hence, putting (3.3) and (3.4) in (3.2), we see that (3.2) lies in ikR. This shows (3.1), and finishes the proof.

For ϱH and kZ with k2 we consider the Petersson Poincaré series [27]

ηk,ϱ(z):=12γΓ(z-ϱ)(z-ϱ¯)ϱ-ϱ¯-k|2kγ. 3.5

It is straightforward to check that ηk,ϱ(z) only depends on the class of ϱ in Γ\H and defines a meromorphic modular form in D2k,R(Γ) with residue divisor res(ηk,ϱ)=[ϱ]. For a given divisor D=[ϱ]cϱwϱ[ϱ]Div(Γ\H)R we put

ηk,D(z):=[ϱ]Γ\Hcϱwϱηk,ϱ(z)D2k,R(Γ).

Then res(ηk,D)=D. In order to prove Proposition 1.1, we need to show that (ηk,D,C)=0 for any CCR(Γ), and that ηk,D is uniquely determined by this condition among all forms in D2k,R(Γ) with residue divisor D.

Lemma 3.2

For any CCR(Γ) we have (ηk,D,C)=0.

Proof

It suffices to prove the lemma for D=1wϱ[ϱ] and C=Cγ being a closed geodesic on Γ\H. In this case the claim follows almost verbatim from the proof of [23, Theorem 3]: Using the unfolding argument one obtains that (ηk,ϱ,Cγ) is a finite linear combination of integrals of the shape -iizk-1dz(Az2+Bz+C)k for certain real binary quadratic forms [ABC] of negative discriminant. Since AC>0 for any such quadratic form, [23, Lemma 2] shows that these integrals all vanish.

Lemma 3.3

If fD2k,R(Γ) satisfies res(f)=D and (f,C)=0 for any CCR(Γ), then f=ηk,D.

Proof

Under the above assumptions, g:=f-ηk,D is a cusp form of weight 2k for Γ with

(g,Cγ)=idγk-1Cγg(z)Qγ(z,1)k-1dz=0

for any closed geodesic Cγ on Γ\H. In other words, either all geodesic cycle integrals of g are real, or they are all purely imaginary (depending on the parity of k). We now show that this implies g=0, using [23, Theorem 2]. We abbreviate rk(g,γ)=Cγg(z)Qγ(z,1)k-1dz and put γ=a-b-cd for γ=abcd. Then a direct computation shows that we have

rk(g,γ)=rk(gc,γ)¯, 3.6

where gc(z)=g(-z¯)¯S2k(Γ). Now, let us assume that rk(g,γ) is real for all hyperbolic γΓ. Then (3.6) says that

rk(g,γ)=rk(gc,γ) 3.7

for every hyperbolic γΓ. We can write g=g+ig with the cusp forms g=12(g+gc) and g=-i2(g-gc). Note that gc=g-ig. Now, (3.7) becomes

rk(g,γ)+irk(g,γ)=rk(g,γ)-irk(g,γ).

Since this equation holds for all hyperbolic γΓ, it also holds with γ replaced by γ. Taking the sum or the difference of the two resulting equations, we obtain that

rk(g,γ)-rk(g,γ)=0,andrk(g,γ)+rk(g,γ)=0,

for every hyperbolic γΓ. By [23, Theorem 2 ii)], this implies that g=0 and g=0, hence g=0. The case when rk(g,γ) is purely imaginary for all hyperbolic γΓ is analogous.

Alternatively, one may use the Eichler-Shimura isomorphism to deduce that g=0.

Taking together the above lemmas, we obtain the statement of Proposition 1.1. We remark that Proposition 1.1 holds more generally for discrete subgroups ΓSL2(R) with finite covolume, by the same arguments as above.

We close this section with some remarks about further properties of the meromorphic modular forms ηϱ(z).

Remark 3.4

  1. One can show that ηk,ϱ(z) is a constant multiple of R0,zkGk(z,ϱ), where Gk(z,ϱ) denotes the higher Green function on H×H, see [7, 26].

  2. It is well-known that ηk,ϱ(z) is orthogonal to cusp forms with respect to a suitable regularized Petersson inner product, see [7, 27].

  3. We have seen above that the real parts of the cycle integrals of ik-1ηk,ϱ(z) vanish. In [1, 24] it was proved that if ϱ is a CM point, then the imaginary parts of (certain linear combinations of) the cycle integrals of ik-1ηk,ϱ(z) are rational (up to some normalizing factors involving powers of π and the disciminant of ϱ).

Theta lifts and meromorphic modular forms

Twisted theta functions and theta lifts

Let L be the lattice from Sect. 2.2. Let ΔZ be a fundamental discriminant (possibly 1) and let ρZ be such that ρ2Δ(mod2N). We let χΔ be the generalized genus character on binary quadratic forms as in [20], which can also be viewed as a function on L using the identification of Ld,r with Qd,r explained in Sect. 2.2. For notational convenience we put

ρ~L:=ρL,ifΔ>0,ρ¯L,ifΔ<0.

Following [9], we define the twisted Siegel theta function by

ΘΔ,ρ(τ,z):=v12hL/LXL+ρhq(X)Δq(h)(Δ)χΔ(X)eq(Xz)|Δ|τ+q(Xz)|Δ|τ¯eh.

It transforms like a modular form of weight 12 for ρ~L in τ and is Γ0(N)-invariant in z. Similarly, following [22] we define the twisted Millson theta function by

ΘΔ,ρ(τ,z):=v32hL/LXL+ρhq(X)Δq(h)(Δ)χΔ(X)pz(X)eq(Xz)|Δ|τ+q(Xz)|Δ|τ¯eh,

with the polynomial pz(X)=-(X,X(z)) as in (2.2). The Millson theta function transforms like a modular form of weight -12 for ρ~L in τ and is Γ0(N)-invariant in z. These theta functions and the corresponding theta lifts have been studied at many places in the past years, see for example [2, 4, 16, 22].

Let kN with k2. For a harmonic Maass form fH32-k,ρ~L we define the theta lift

ΦΔ,ρk(f,z):=Ck-·R0,zkFregR32-k,τk-12f(τ),ΘΔ,ρ(τ,z)v12dμ(τ),ifk is odd,Ck+·R0,zkFregR32-k,τk2-1f(τ),ΘΔ,ρ(τ,z)v-12dμ(τ),ifk is even,

where dμ(τ):=dudvv2 denotes the usual invariant measure on H, the integrals are regularized as in [10, Sect. 6], and Ck- and Ck+ denote the normalizing constants

Ck-:=ik|Δ|k-12Nk-122k+1πk-12(k-1)!,Ck+:=ik|Δ|k2-1Nk-122k+12πk2-1(k-1)!. 4.1

We remark that Zemel [34] investigated a similar lift for lattices of more general signature (with Δ=1) and used it to prove a higher weight version of the Gross-Kohnen-Zagier Theorem. Moreover, an analogous theta lift (defined in the same way as ΦΔ,ρk(f,z) but without the raising operator R0,zk in front) was considered in [4], where it was shown to be a linear combination of higher Green functions Gk(ZΔ,ρ(D,r),z) evaluated at Heegner divisors ZΔ,ρ(D,r) in the first variable.

Here, we will show that the lift ΦΔ,ρk(f,z) defines a meromorphic modular form of weight 2k for Γ0(N) with poles of order k at a linear combination of Heegner divisors ZΔ,ρ(D,r). Moreover, we will compute the Fourier expansion of the lift to study the algebraicity of the Fourier coefficients of these meromorphic modular forms.

The theta lift as a meromorphic modular form

Let D<0 be an integer and let rZ/2NZ with Dsgn(Δ)r2(mod4N). We compute the theta lift of the Maass Poincaré series P32-k,D,r(τ), defined in Sect. 2.4, and obtain an explicit representation of the lift in terms of the meromorphic modular form

fk,D,r,Δ,ρ(z):=ik(|DΔ|)k-12QQD|Δ|,rρsgn(Q)χΔ(Q)Q(z,1)k 4.2

of weight 2k for Γ0(N). Here we put sgn(Q):=sgn(a) for Q=[aN,b,c]QD|Δ|,rρ. Let αQH be the Heegner point corresponding to QQD|Δ|,rρ. Then a short computation using

Q(z,1)=2i|DΔ|(z-αQ)(z-α¯Q)(αQ-α¯Q)

shows that

fk,D,r,Δ,ρ(z)=(|DΔ|)k-12QQD|Δ|,rρ/Γsgn(Q)χΔ(Q)ηk,αQ(z),

with the Petersson Poincaré series ηk,ϱ(z) defined in (3.5). In particular, fk,D,r,Δ,ρD2k,R(Γ) is the canonical meromorphic modular form of weight 2k for the rescaled Heegner divisor

res(fk,D,r,Δ,ρ)=|DΔ|k-12ZΔ,ρ(D,r).

Proposition 4.1

Let kZ with k2, let DZ with D<0 and rZ/2NZ with Dsgn(Δ)r2(mod4N). Then we have

ΦΔ,ρkP32-k,D,r(τ),z=fk,D,r,Δ,ρ(z).

Remark 4.2

Let f be a harmonic weak Maass form of weight 32-k. Since the Poincaré series P32-k,D,r(τ) generate the space of weight 32-k harmonic weak Maass forms, we directly deduce that the lift ΦΔ,ρk(f,z) defines a meromorphic modular form of weight 2k for Γ0(N) with poles of order k at a linear combination of Heegner divisors ZΔ,ρ(D,r).

Proof

Let k3 be odd. Using Lemma 2.1 we can write

ΦΔ,ρkP32-k,D,r(τ),z=Ck-π|D|Nk-12Γk+12R0,zkFregP12,D,rτ,k2+14,ΘΔ,ρ(τ,z)v12dμ(τ),

where Ck- is the normalizing constant defined in (4.1). The regularized integral can be computed by unfolding against the Maass Poincaré series as in the proof of [12, Theorem 2.14]. We obtain that

ΦΔ,ρkP32-k,D,r(τ),z=2Ck-π|D|Nk-12Γk+12Γk2Γk+12×XL+ρrq(X)=D|Δ|/4NχΔ(X)·R0,zkD|Δ|4Nq(Xz)k22F1k2,k+12;k+12;D|Δ|4Nq(Xz).

Recall from (2.3) that q(Xz)=-12pz2(X). By Proposition 2.2 (with m=D|Δ|4N) we find

ΦΔ,ρkP32-k,D,r(τ),z=2Ck-|D||Δ|k2π|D|Nk-12Γ(2k)Γk+12Γk2Γ(k)Γk+12×XL+ρrq(X)=D|Δ|/4Nsgn(pz(X))χΔ(X)QX(z)k.

The quotient of Gamma functions can be simplified to 2k(k-1)! using the Legendre duplication formula πΓ(2z)=22z-1Γ(z)Γ(z+12). Moreover, we have sgn(pz(X))=sgn(QX), independently of zH, where QX is the quadratic form corresponding to X under the identification of LD|Δ|,rρ with QN,D|Δ|,rρ explained in Sect. 2.2. Hence we can rewrite the sum into a multiple of fk,D,r,Δ,ρ(z). Taking into account the normalizing constant Ck- defined in (4.1) and putting everything together, we obtain the stated formula.

The proof for even k is analogous, so we omit the details.

The Fourier expansion of the theta lift

In this section we compute the Fourier expansion of the theta lift of harmonic Maass forms.

Proposition 4.3

Let kZ with k2. For y0 large enough, the theta lift ΦΔ,ρk(f,z) of fH32-k,ρ~L has the Fourier expansion

ΦΔ,ρk(f,z)=Ck,ΔiπkΔn1n2k-1dnΔdd-kcf+|Δ|n2d2,ρnde(nτ),

with the rational constant Ck,Δ:=(-2sgn(Δ))k|Δ|k-1(k-1)!.

Proof

Suppose that k3 is odd. For simplicity we assume that Δ1, but the proof for Δ=1 is very similar. We write the theta function ΘΔ,ρ(τ,z) as a Poincaré series as in [8, Theorem 4.8]. Then we use the unfolding argument as in the proof of [8, Theorem  5.3] and obtain

ΦΔ,ρk(f,z)=R0,zk(2NCk-yεXKn1Δne(sgn(Δ)n(X,μ))×0c(|Δ|q(X),ρX,v)exp-Nπn2y2|Δ|v-2πq(X)|Δ|vdvv32),

where Ck- is the normalizing constant defined in (4.1), K=m2N100-1:mZ is a positive definite sublattice of L, ε equals 1 if Δ>0 and i if Δ<0, we put μ=x-x2-1-x, and we wrote,1

R32-kk-12f(τ)=hL/Ln-c(n,h,v)e(nu)eh.

for the Fourier expansion of R32-kk-12f(τ).

We first show that the contribution for X=0 vanishes. Note that

Rκvα=(α+κ)vα-1 4.3

for any α,κR, so we have

c(0,0,v)=C0cf+(0,0)v1-k2,whereC0:=j=1k-12j-52.

Now the contribution from the X=0 summand equals

2NεCk-C0cf+(0,0)n1ΔnR0,zky0v1-k2exp-Nπn2y2|Δ|vdvv32=2N1-k2π-k2|Δ|k2εCk-C0cf+(0,0)Γk2LΔ(k)·R0,zky1-k.

It follows from (4.3) that R0ky1-k=(k-1)!R2k-2y2-2k=0. In particular, the Fourier expansion of the theta lift has no constant term.

For m>0 we have e-2πmv=W32-k,k2+14(4πmv) by (2.4), and hence

c(m,h,v)e(mu)=cf+(m,h)R32-kk-12W32-k,k2+14(4πmv)e(mu)=cf+(m,h)(-4πm)k-12W12,k2+14(4πmv)e(mu)

by (2.5). Using (2.6) (with κ=12,s=k2+14,α=4π|Δ|q(X),β=Nπn2y2|Δ|) we have for X0 the evaluation

0W12,k2+14(4π|Δ|q(X)v)exp-Nπn2y2|Δ|v-2πq(X)|Δ|vdvv32=|Δ|NnyW0,1-k4πny4Nq(X).

Note that q(X)=m24N and (X,μ)=mx for X=m2N100-1K. Using (2.5) we find

R0k(W0,1-k(4π|m|ny)e(sgn(Δ)mnx))=(-4π|m|n)kWk,1-k(4π|m|ny)e(|m|nx),sgn(Δ)m>0,0,sgn(Δ)m<0,

since for sgn(Δ)m<0 one of the factors (s+κ2) in (2.5) will be 0. By (2.4) we have Wk,1-k(4π|m|ny)=e-2π|m|ny. Taking everything together, we obtain after simplification

ΦΔ,ρk(f,z)=(-1)k+12εCk-sgn(Δ)22k+1N1-k2π3k-12|Δ|k2n1n2k-1×dnΔdd-kcf+|Δ|n2d2,ρnde(nτ),

where we also used that cf+(m,sgn(Δ)r)=sgn(Δ)cf+(m,r). Putting in the normalizing constant Ck- from (4.1), we obtain the stated formula.

The proof for even k is similar. Now we use Theorem 5.10 and Lemma 5.6 of [16] to write the theta function ΘΔ,ρ(τ,z) as a Poincaré series

ΘΔ,ρ(τ,z)=-Niy222|Δ|n1γ~Γ~\Γ~nexp-Nπn2y2|Δ|vΞ(τ,μ,-n,0)-1/2,ρ~Kγ~

where

Ξ(τ,μ,-n,0)=Δnε|Δ|hK/KXK+rhQ(λ)ΔQ(h)(Δ)eq(X)τ|Δ|-n(X,μ)|Δ|eh.

By the unfolding argument we obtain as in the proof of [8, Theorem 5.3] that

ΦΔ,ρk(f,z)=R0,zk(2Niy2ε¯XKn1nΔne(sgn(Δ)n(λ,μ))×0c(|Δ|q(X),ρX,v)exp-Nπn2y2|Δ|v-2πq(X)|Δ|vdvv52).

Now the rest of the proof proceeds as in the case of odd k, so we omit the details.

Meromorphic modular forms corresponding to Heegner divisors

In this section we investigate the canonical and normalized meromorphic modular forms corresponding to twisted Heegner divisors and prove Theorem 1.3 and Theorem 1.4.

Let NN and kN with k2, and let L be the lattice defined in (2.1). Recall from [18, 29] that the space M12+k,ρL of holomorphic vector-valued modular forms of weight 12+k for ρL is isomorphic to the space Jk+1,Nskew of skew holomorphic Jacobi forms of weight k+1 and index N. Similarly, M12+k,ρ¯L is isomomorphic to the space Jk+1,N of holomorphic Jacobi forms of weight k+1 and index N. Hence, the theory of Hecke operators for Jacobi forms developed in [18, 29, 30] carries over to vector-valued modular forms for ρL and ρ¯L. In particular, there is a newform theory for these spaces of vector-valued modular forms.

We let S2knew,+(N) be the space of cuspidal newforms of weight 2k for Γ0(N) on which the Fricke involution acts by multiplication with (-1)k.

By the Shimura correspondence, it is isomorphic as a module over the Hecke algebra to S12+k,ρLnew, compare [20, 29, 30]. Similarly, the space S2knew,-(N) of cuspidal newforms of weight 2k for Γ0(N) on which the Fricke involution acts by multiplication with -(-1)k is isomorphic as a module over the Hecke algebra to S12+k,ρ¯Lnew.

Let ρ be one of the representations ρL and ρ¯L. For every mN there is a Hecke operator Tm acting on M12+k,ρ. The action on Fourier expansions can be computed explicitly, compare [18, 30]. For example, if p is a prime with (p,N)=1, and we let f=n,rcf(n,r)qnerM12+k,ρ and f|Tp=n,rcf|Tp(n,r)qner, then we have

cf|Tp(n,r)=cf(p2n,pr)+pk-14Nσnpcf(n,r)+p2k-1cf(n/p2,r/p), 5.1

where σ=1 if ρ=ρL and σ=-1 if ρ=ρ¯L. There are similar formulas for general mN. The Hecke operators act on harmonic Maass forms in an analogous way, and the action on Fourier expansions is the same.

We let ρZ/2NZ and Δρ2(mod4N) be a fundamental discriminant, and we put ϵ:=-sgn(Δ). For a normalized newform GS2knew,ϵ(N) we let FG be the totally real number field generated by its Fourier coefficients, and we let gS12+k,ρ~¯Lnew be its Shimura correspondent, which we may normalize to have coefficients in FG, as well. Then, by [8, Lemma 7.2] there exists a harmonic Maass form fH32-k,ρ~L(FG) whose principal part has coefficients in FG, and which satisfies ξ32-kf=g-2g, where we write g=(g,g) for the Petersson norm of g. We let

Zk,Δ,ρ(f):=r(2N)D<0cf+(D,r)|DΔ|k-12ZΔ,ρ(D,r)

be the twisted Heegner divisor corresponding to f. Then the canonical meromorphic modular form with residue divisor Zk,Δ,ρ(f) is given by

ηk,Δ,ρ(f)=r(2N)D<0cf+(D,r)fk,D,r,Δ,ρ(z), 5.2

with fk,D,r,Δ,ρ defined in (4.2). On the other hand, by Proposition 4.1 the meromorphic modular form ηk,Δ,ρ(f) can be obtained as the theta lift ΦΔ,ρk(f,z) of f. In particular, by Proposition 4.3, we have the Fourier expansion

ηk,Δ,ρ(f)=Ck,ΔiπkΔn>0n2k-1dnΔdd-kcf+|Δ|n2d2,ρnde(nz) 5.3

with a rational constant Ck,ΔQ.

Theorem 5.1

Let fH32-k,ρ~L(FG) be as above, and let ηk,Δ,ρ(f)D2k,R(Γ) be the canonical meromorphic modular form of weight 2k for Zk,Δ,ρ(f) as in (5.2). Then the following are equivalent.

  1. We have cf+(|Δ|,ρ)FG.

  2. We have cf+(n2|Δ|,nρ)FG for all nN.

  3. All Fourier coefficients of ηk,Δ,ρ(f) are contained in iπkΔFG.

Proof

The implication (2) (3) follows from the Fourier expansion of ηk,Δ,ρ(f) in (5.3). Similarly, the implication (3) (1) follows from (5.3) since the first coefficient of ηk,Δ,ρ is given by Ck,ΔiπkΔcf+(|Δ|,ρ).

It remains to prove that (1) implies (2). We use the action of the Hecke operators on f. By [8, Proposition 7.1] we have ξ32-k(f|Tn)=n1-2kξ32-k(f)|Tn. Moreover, using ξ32-k(f)=g-2g and g|Tn=λn(G)g, where λn(G)FG is the Tn-eigenvalue of G, we see that

f:=n2k-1f|Tn-λn(G)fM32-k,ρ~L!(FG)

is a weakly holomorphic modular form with principal part defined over FG. Since the space of weakly holomorphic modular forms for ρ~L has a basis of forms with rational Fourier coefficients by a result of McGraw [25], we obtain that all Fourier coefficients of f lie in FG. Now using the explicit formula (5.1) for the action of Tn on the coefficients of f, we see that cf+(n2|Δ|,nρ) is a rational linear combination of coefficients cf+(m2|Δ|,mρ) for some m<n and coefficients of f, which lie in FG by the above discussion. Hence, if cf+(|Δ|,ρ)FG then it follows by induction that cf+(n2|Δ|,nρ)FG for every nN.

Theorem 5.2

For fH32-k,ρ~L(FG) as above, there exists a unique meromorphic modular form ζk,Δ,ρ(f)D2k,R(Γ) with the following properties.

  1. The residue divisor is given by res(ζk,Δ,ρ(f))=Zk,Δ,ρ(f).

  2. We have (ζk,Δ,ρ(f),C)=0 for every closed geodesic C in Γ0(N)\H with (G,C)=0.

  3. The first Fourier coefficient of ζk,Δ,ρ(f) vanishes.

Moreover, all Fourier coefficients of ζk,Δ,ρ(f) lie in iπkΔFG, and we have

cf+(|Δ|,ρ)=-1Ck,ΔπikΔ·(ζk,Δ,ρ(f),C)(G,C) 5.4

for every CCR(Γ) with (G,C)0.

Proof

By (5.3) the first coefficient of ηk,Δ,ρ(f) is given by Ck,ΔiπkΔcf+(|Δ|,ρ). Hence

ζk,Δ,ρ(f):=ηk,Δ,ρ(f)-Ck,ΔiπkΔcf+(|Δ|,ρ)G

has the desired properties. Taking the bilinear pairing with any CCR(Γ) yields

(ζk,Δ,ρ(f),C)=(ηk,Δ,ρ(f),C)-Ck,ΔiπkΔcf+(|Δ|,ρ)(G,C)=-Ck,ΔiπkΔcf+(|Δ|,ρ)(G,C),

which gives the formula (5.4).

In order to show that all Fourier coefficients of ζk,Δ,ρ(f) lie in iπkΔFG we use Hecke operators. First note that we have

ηk,Δ,ρ(f)|Tm=ηk,Δ,ρ(m2k-1f|Tm) 5.5

for any mN. This may be checked using the explicit action (5.1) of Tm on the Fourier expansion of f and the action of Tm on fk,D,r,Δ,ρ, which can be computed as in the proof of [33, Eq. (36)]. For example, if p is prime with (p,N)=1 we have

fk,D,r,Δ,ρ|Tp=fk,p2D,pr,Δ,ρ+pk-1Dσpfk,D,r,Δ,ρ+p2k-1fk,D/p2,r/D,Δ,ρ,

and there are similar formulas for any Hecke operator Tm. We leave the details of the verification of (5.5) to the reader. Now (5.5) implies that

ζk,Δ,ρ(f)|Tm-λm(G)ζk,Δ,ρ(f)=ηk,Δ,ρ(f) 5.6

with the weakly holomorphic modular form f=m2k-1f|Tm-λm(G)fM32-k,ρ~L(FG). Again, it follows from [25] that f has all coefficients in FG, which by (5.3) implies that ηk,Δ,ρ(f) has all coefficients in iπkΔFG. Using the explicit action of Tm on Fourier expansions and the fact that the first Fourier coefficient of ζk,Δ,ρ(f) vanishes, we obtain by induction from (5.6) that all coefficients of ζΔ,ρ(f) lie in iπkΔFG.

Outlook: derivatives of L-functions of newforms of higher weight

In this last section we explain a possible future application of our results to a non-vanishing criterion for central values of derivatives of L-functions of newforms of weight 2k. We keep the notation from Sect. 5 and first recall a non-vanishing criterion for the central L-derivatives of newforms of weight 2 from [8]. Let GS2new,ϵ(N) be a newform of weight 2 for Γ0(N). The functional equation of the twisted L-function L(G,χΔ,s) implies that L(G,χΔ,1)=0. Hence, it is natural to consider the (non-)vanishing of L(G,χΔ,s) at s=1. The following theorem connects this question to the algebraicity of Fourier coefficients of (the holomorphic part of) harmonic Maass forms.

Theorem 6.1

(Theorem 7.6 in [8]) Let GS2new,ϵ(N), let gS32,ρ~¯L be its Shimura correspondent with coefficients in FG, and let fH12,ρL(FG) be a harmonic Maass form with principal part defined over FG and ξ12f=g-2g. Then the following are equivalent.

  1. We have L(G,χΔ,1)=0.

  2. We have cf+(|Δ|,ρ)FG.

The proof of this theorem consists of three main steps.

  1. First, one may show that the canonical differential ηΔ,ρ(f) for the (degree 0) twisted Heegner divisor
    yΔ,ρ(f):=ZΔ,ρ(f)-deg(ZΔ,ρ(f))·
    is defined over a number field if and only if cf+(|Δ|,ρ)FG. This is done by constructing ηΔ,ρ(f) as a regularized theta lift and using Hecke operators.
  2. Secondly, by transcendence results for differentials of the third kind due to Scholl [28], Waldschmidt [31], and Wüstholz [32], we have that the canonical differential ηΔ,ρ(f) is defined over a number field if and only if some integral multiple of the Heegner divisor yΔ,ρ(f) is a principal divisor. This is in turn equivalent to saying that the Néron-Tate height of yΔ,ρ(f) vanishes.

  3. Lastly, by the Gross-Zagier Theorem [21] the Néron-Tate height of yΔ,ρ(f) is a multiple of L(G,χΔ,1), which concludes the proof of the above theorem.

It would be interesting to extend Theorem 6.1 to newforms of higher weight 2k. We may speculate that the vanishing of L(G,χΔ,k) for a newform GS2knew,ϵ(N) is equivalent to the algebraicity of the coefficient cf+(|Δ|,ρ), where fH32-k,ρ~L(FG) is a harmonic Maass form whose principal part is defined over FG and which satisfies ξ32-k(f)=g-2g, with gS12+k,ρ~¯Lnew being the Shimura correspondent of G.

Our Theorem 5.1 generalizes step (1) in the above proof sketch to higher weight 2k. Moreover, step (3) is (essentially) taken care of by Zhang’s generalization of the Gross-Zagier formula to newforms of higher weight [35]. Here the Néron-Tate height on the Jacobian of X0(N) has to be replaced with a height pairing on the Kuga-Sato variety W2k-2 of dimension 2k-1 over X0(N), and the Heegner divisor yΔ,ρ(f) has to be replaced with its corresponding Heegner cycle in the Chow group of codimension k cycles on W2k-2. Concerning step (2) of the above proof sketch, it remains to show that the Fourier coefficients of the canonical meromorphic modular form ηk,Δ,ρ(f) are contained in iπkΔFG if and only if the Heegner cycle corresponding to yΔ,ρ(f) vanishes in the Chow group.

A possible path to prove one direction of this claim is laid out in the thesis of Mellit [26]. Given a principal divisor in the Chow group of codimension k cycles in W2k-2, Mellit constructs a meromorphic modular form with algebraic Fourier coefficients. We expect that Mellit’s meromorphic modular form attached to the Heegner cycle associated with yΔ,ρ(f) (assuming that it is principal) is a multiple of ηk,Δ,ρ(f).

The converse direction seems to be more difficult. Assuming that the coefficients of ηk,Δ,ρ(f) are contained in iπkΔFG, we would need to construct a rational function F on a suitable codimension k-1 cycle U in the Kuga-Sato variety W2k-2 such that the divisor of F is the Heegner cycle corresponding to yΔ,ρ(f). For k=1, the cycle U is just the whole modular curve X0(N), and it was shown in [8] that the rational function F with divisor yΔ,ρ(f) can be constructed as the Borcherds product attached to f, but for k>1 it is not clear how to construct U and F. We plan to come back to this problem in the future.

Acknowledgements

We thank Jens Funke for helpful discussions. The first author was partially supported by the Daimler and Benz Foundation and the Klaus Tschira Boost Fund. The second author was supported by the DFG Collaborative Research Centre TRR 326 Geometry and Arithmetic of Uniformized Structures, project number 444845124. The third author was supported by SNF projects 200021_185014 and PZ00P2_202210.

Funding

Open Access funding enabled and organized by Projekt DEAL.

Data Availability

Not applicable.

Declarations

Conflict of interest statement

On behalf of all authors, the corresponding author states that there is no Conflict of interest.

Footnotes

1

note that, in contrast to [8, Theorem 5.3] we put e-2πnv into c(nhv).

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

References

  • 1.Alfes-Neumann, C., Bringmann, K., Schwagenscheidt, M.: On the rationality of cycle integrals of meromorphic modular forms. Math. Ann. 376, 243–266 (2020) [Google Scholar]
  • 2.Alfes-Neumann, C., Schwagenscheidt, M.: On a theta lift related to the Shintani lift. Adv. Math. 328, 858–889 (2018) [Google Scholar]
  • 3.Abramowitz, M., Stegun, I.A.: Handbook of mathematical functions with formulas, graphs, and mathematical tables, volume 55 of National Bureau of Standards Applied Mathematics Series. For sale by the Superintendent of Documents, U.S. Government Printing Office, Washington, D.C., (1964)
  • 4.Bruinier, J.H., Ehlen, S., Yang, T.: CM values of higher automorphic Green functions for orthogonal groups. Invent. Math. 225, 693–785 (2021) [Google Scholar]
  • 5.Bruinier, J.H., Funke, J.: On two geometric theta lifts. Duke Math. J. 125(1), 45–90 (2004) [Google Scholar]
  • 6.Bringmann, K., Folsom, A., Ono, K.: Inline graphic-series and weight Inline graphic Maass forms. Compos. Math. 145(3), 541–552 (2009) [Google Scholar]
  • 7.Bringmann, K., Kane, B., von Pippich, A.-M.: Regularized inner products of meromorphic modular forms and higher Green’s functions. Commun. Contemp. Math., 21, (2019)
  • 8.Bruinier, J.H., Ono, K.: Heegner divisors, Inline graphic-functions and harmonic weak Maass forms. Ann. of Math. (2), 172(3), 2135–2181 (2010)
  • 9.Bruinier, J.H., Ono, K.: Algebraic formulas for the coefficients of half-integral weight harmonic weak Maass forms. Adv. Math. 246, 198–219 (2013) [Google Scholar]
  • 10.Borcherds, R.E.: Automorphic forms with singularities on Grassmannians. Invent. Math. 132(3), 491–562 (1998) [Google Scholar]
  • 11.Bruinier, J.H., Ono, K., Rhoades, R.C.: Differential operators for harmonic weak Maass forms and the vanishing of Hecke eigenvalues. Math. Ann. 342(3), 673–693 (2008) [Google Scholar]
  • 12.Bruinier, J.H.: Borcherds products on O(2, Inline graphic) and Chern classes of Heegner divisors. Lecture Notes in Mathematics, vol. 1780. Springer-Verlag, Berlin (2002)
  • 13.Bruinier, J.H.: Harmonic Maass forms and periods. Math. Ann. 357(4), 1363–1387 (2013) [Google Scholar]
  • 14.Bruinier, J.H., Schwagenscheidt, M.: Algebraic formulas for the coefficients of mock theta functions and Weyl vectors of Borcherds products. J. Algebra 478, 38–57 (2017) [Google Scholar]
  • 15.Crawford, J., Funke, J.: The Shimura-Shintani correspondence via singular theta lifts and currents. Int. J. Number Theory 19(10), 2383–2417 (2023). 10.1142/s1793042123501178
  • 16.Crawford, J., Funke, J.: The Shimura-Shintani correspondence via singular theta lifts and currents. arxiv preprint arxiv:2112.11379, (2021)
  • 17.Erdélyi, A., Magnus, W., Oberhettinger, F., Tricomi, F.G.: Tables of integral transforms. Vol. I. McGraw-Hill Book Company, Inc., New York-Toronto-London, (1954). Based, in part, on notes left by Harry Bateman
  • 18.Eichler, M., Zagier, D.B.: The theory of Jacobi forms. Progress in Mathematics, vol. 55. Birkhäuser Boston Inc., Boston, MA (1985)
  • 19.Funke, J., Millson, J.: Spectacle cycles with coefficients and modular forms of half-integral weight. In Arithmetic geometry and automorphic forms, volume 19 of Adv. Lect. Math. (ALM), pages 91–154. Int. Press, Somerville, MA, (2011)
  • 20.Gross, B.H., Kohnen, W., Zagier, D.B.: Heegner points and derivatives of Inline graphic-series. II. Math. Ann. 278(1–4), 497–562 (1987) [Google Scholar]
  • 21.Gross, B.H., Zagier, D.B.: Heegner points and derivatives of Inline graphic-series. Invent. Math. 84(2), 225–320 (1986) [Google Scholar]
  • 22.Hövel, M.: Automorphe Formen mit Singularitäten auf dem hyperbolischen Raum. TU Darmstadt PhD Thesis, (2012)
  • 23.Katok, S.: Closed geodesics, periods and arithmetic of modular forms. Invent. Math. 80, 469–480 (1985) [Google Scholar]
  • 24.Löbrich, S., Schwagenscheidt, M.: Meromorphic modular forms with rational cycle integrals. Int. Math. Res, Notices (2020) [Google Scholar]
  • 25.McGraw, W.J.: The rationality of vector valued modular forms associated with the Weil representation. Math. Ann. 326(1), 105–122 (2003) [Google Scholar]
  • 26.Mellit, A.: Higher Green’s functions for modular forms. Ph. D. Thesis, Bonn, (2008)
  • 27.Petersson, H.: Über automorphe Orthogonalfunktionen und die Konstruktion der automorphen Formen von positiver reeller Dimension. Math. Ann. 127, 33–81 (1954) [Google Scholar]
  • 28.Scholl, A.J.: Fourier coefficients of Eisenstein series on noncongruence subgroups. Math. Proc. Cambridge Philos. Soc. 99(1), 11–17 (1986) [Google Scholar]
  • 29.Skoruppa, N.-P.: Developments in the theory of Jacobi forms. In Automorphic functions and their applications (Khabarovsk, 1988), pages 167–185. Acad. Sci. USSR Inst. Appl. Math., Khabarovsk, (1990)
  • 30.Skoruppa, N.-P., Zagier, D.: Jacobi forms and a certain space of modular forms. Invent. Math. 94(1), 113–146 (1988) [Google Scholar]
  • 31.Waldschmidt, M.: Nombres transcendants et groupes algébriques. Astérisque, (69-70), 218 (1987). With appendices by Daniel Bertrand and Jean-Pierre Serre
  • 32.Wüstholz, G.: Transzendenzeigenschaften von Perioden elliptischer Integrale. J. Reine Angew. Math. 354, 164–174 (1984) [Google Scholar]
  • 33.Zagier, D.B.: Eisenstein series and the Riemann zeta function. in: Automorphic Forms, Representation Theory and Arithmetic, Springer-Verlag, Berlin-Heidelberg-New York, pages 275–301 (1981)
  • 34.Zemel, S.: A Gross-Kohnen-Zagier type theorem for higher-codimensional Heegner cycles. Res. Number Theory 1(23), 1–44 (2015) [Google Scholar]
  • 35.Zhang, S.: Heights of Heegner cycles and derivatives of Inline graphic-series. Invent. Math. 130(1), 99–152 (1997) [Google Scholar]
  • 36.Zwegers, S.: Mock theta functions. Utrecht PhD thesis, (2002)

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Data Availability Statement

Not applicable.


Articles from Mathematische Annalen are provided here courtesy of Springer

RESOURCES