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 . 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 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 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 with , 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 throughout the introduction. In the body of the paper we will treat level for arbitrary , 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 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 on X with at most simple poles and residue divisor , and such that
for any closed path C in X avoiding the points of D. The differential is called the canonical differential of the third kind associated with D. If X is a modular curve, then may be viewed as a meromorphic modular form of weight 2. Our first result is a construction of certain ‘canonical’ meromorphic modular forms of weight 2k associated with divisors D on modular curves, generalizing the notion of canonical differentials to higher weight.
For a subfield and an integer with we let be the F-vector space of meromorphic modular forms f(z) defined over F of weight 2k for which vanish at the cusp of and which have poles of order at most k on , with expansions of the form
at each pole of f, with some constants . Note that for we have . Hence, by a slight abuse of notation, we will call the F-linear combination
of points in the residue divisor of . Here is half the order of the stabilizer of in , and denotes the space of all F-linear combinations of points on . Note that the sum is well-defined, that is, and only depend on the class of in .
It is well known that closed geodesics on correspond to (conjugacy classes of) primitive hyperbolic matrices in . More precisely, given such a primitive hyperbolic matrix , we let be the semi-circle in connecting the two real fixed points of . Then defines a closed geodesic in , and every closed geodesic in is accounted for in this way. We let be the real vector space spanned by all -linear combinations of closed geodesics on . We obtain a real-valued bilinear pairing on by setting, for each meromorphic modular form and each closed geodesics on ,
| 1.1 |
where denotes the binary quadratic form corresponding to , is the discriminant of , and the path of integration is any path in (avoiding the poles of f) from some point (not being a pole of f) to . Using the residue theorem and the assumption that the coefficients are real numbers, one can show that the value of the pairing is indeed independent of the choice of and the path from to , see Lemma 3.1 below.
We have the following higher weight analogues of canonical differentials associated with divisors on .
Proposition 1.1
For each divisor there exists a unique meromorphic modular form with residue divisor and for any .
We call 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 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 with coefficients in the local system E associated with , geodesics in 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 with we let be a normalized Hecke eigenform of weight 2k for , and let be the totally real number field generated by the Fourier coefficients of G. We let be a Hecke eigenform of weight for 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 . Then, by [8, Lemma 7.3] there exists a harmonic Maass form f of weight for with principal part defined over such that , where denotes the Petersson norm of g and . Every such harmonic Maass form f can be written as a sum of a holomorphic part and a non-holomorphic part , with Fourier expansions of the shape
with coefficients , where and denotes the incomplete Gamma function. The condition that the principal part of f is defined over means that we have for all . We let denote the space of harmonic Maass forms of weight for satisfying the Kohnen plus space condition and with principal part defined over . We are interested in the algebraicity properties of the coefficients for .
Now we come to the construction of the Heegner divisor corresponding to f. For a negative discriminant we let be the set of positive definite integral binary quadratic forms of discriminant . For the quadratic polynomial Q(z, 1) has a unique root in , called the CM point (or Heegner point) corresponding to Q. Let be a fundamental discriminant with and let be an integer such that . Moreover, let be the usual genus character on . Then we define the twisted Heegner divisor
where . Eventually, we let
be the twisted Heegner divisor corresponding to f, where denote the coefficients of the holomorphic part of f. We remark that for even k the Heegner divisor is in general not a divisor with coefficients in , but with coefficients in some totally real algebraic extension of .
The following result relates the algebraicity of the coefficients 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 . It (partly) generalizes [8, Theorem 5.5, Theorem 7.6] to higher weight.
Theorem 1.3
Let be as above, and let be the canonical meromorphic modular form of weight 2k for . Then the following are equivalent.
We have .
We have for all .
All Fourier coefficients of are contained in .
In order to prove the theorem we will construct 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 in terms of the coefficients . Invoking the action of Hecke operators on f and then yields the stated result. We refer to Sect. 5 for the generalization of Theorem 1.3 to higher level and its proof.
We remark that a similar theta lift has been studied by Zemel [34], and it was used to show the implication in Theorem 1.3 (for ).
Finally, we obtain a formula for the coefficient in terms of cycle integrals of a certain normalized meromorphic modular form in with residue divisor , generalizing [13, Theorem 1.1] to higher weight.
Theorem 1.4
For as above, there exists a unique meromorphic modular form with the following properties.
The residue divisor is given by .
We have for every with .
The first Fourier coefficient of vanishes.
Moreover, all Fourier coefficients of lie in , and we have
| 1.2 |
for every with , with the rational constant .
Remark 1.5
The formula from Theorem 1.4 tells us that the coefficient can be written as a quotient of (real parts of) cycle integrals of (meromorphic) modular forms with coefficients in .
We call the normalized meromorphic modular form of weight 2k for . We will construct by subtracting a suitable multiple of G from . The fact that the coefficients lie in and the formula (1.2) again follow by writing as a theta lift of f, and using the action of Hecke operators on . 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 . 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 . For a negative discriminant and with we consider the set of integral binary quadratic forms
The group acts on from the right, with finitely many orbits if .
We let be the subset of positive definite quadratic forms in , which are given by the condition . For we let
be the Heegner point corresponding to Q. It is the unique root of Q(z, 1) in . The order of the stabilizer of Q in is finite.
Let be a fundamental discriminant and let with . For with we let be the generalized genus character on as in [20], and we define the twisted Heegner divisor
where denotes the class of in .
Quadratic spaces and lattices
For 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 and the associated bilinear form . The group acts as isometries on V by .
Throughout this work we will consider the lattice
| 2.1 |
in V. The dual lattice is given by
Hence the discriminant form is isomorphic to with the finite quadratic form , and we will use this identification without further notice. Note that the group acts on L and fixes the classes of .
For with we let
Each element corresponds to a binary quadratic form
of discriminant . The group acts on both and , and the actions are compatible in the sense that for . In particular, we obtain a bijection between and .
We let be the Grassmannian of positive definite 2-dimensional subspaces of . We can identify with by sending to the positive definite plane , where
Note that and for . We also define the vectors
such that form an orthogonal basis of . Note that U(z) is spanned by . For and we consider the polynomials
| 2.2 |
Then we have the useful rules
| 2.3 |
Harmonic Maass forms
The metaplectic extension of is defined as
It is generated by and . We let be the subgroup generated by T. Moreover, for we let , where denotes the principal branch of the square root.
We let L be the lattice from Sect. 2.2 and we let be its group ring, which is generated by the formal basis vectors for . The Weil representation associated with L is the representation of on defined by
The Weil representation associated to the lattice is called the dual Weil representation associated to L.
Let and define the slash-operator by
for a function and . Here, and in the following we let . Following [5], we call a smooth function a harmonic Maass form of weight with respect to if it is annihilated by the weight Laplace operator
if it is invariant under the slash-operator for all , and if there exists a -valued Fourier polynomial (the principal part of f)
such that as for some . We denote the vector space of harmonic Maass forms of weight with respect to by , and we let be the subspace of weakly holomorphic modular forms. Every can be written as a sum of a holomorphic and a non-holomorphic part, having Fourier expansions of the form
where denotes the incomplete Gamma function.
The antilinear differential operator maps a harmonic Maass form to a cusp form of weight for . We further require the raising operator , which raises the weight of a smooth function transforming like a modular form of weight for by two, and we define the iterated raising operator for , and .
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 with . For and we let
with the usual M-Whittaker function. Let with and with . For with we define the -valued Maass Poincaré series
where is the subgroup of generated by . It converges absolutely for , it transforms like a modular form of weight for , and it is an eigenform of the Laplace operator with eigenvalue . The special value
defines a harmonic Maass form in whose Fourier expansion starts with
The following lemma follows inductively from [9, Proposition 2.2].
Lemma 2.1
For we have that
Maass Poincaré series for the dual Weil representation are defined analogously, and Lemma 2.1 remains true for them. However, in this case, we have to require that , and the Fourier expansion starts with .
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
where is the usual W-Whittaker function. At it simplifies to
| 2.4 |
Moreover, using (13.4.33) and (13.4.31) in [3], we obtain the formula
| 2.5 |
for and . Moreover, we will need the integral formula
| 2.6 |
for , which follows by evaluating the integral in terms of the K-Bessel function using [17, (22) on p. 217], and then writing , 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 with and we have
Proof
For brevity, we put . We prove the proposition by induction. To this end, for and fixed we define the function
Note that we want to compute . We claim that
| 2.7 |
To prove this, we first use that for every holomorphic function , every smooth function , and every we have the simple relation
Applying this with and , we obtain
Next, we write and use the formula (see (15.5.3) of NIST) to obtain
A direct computation shows that
Finally, using and taking everything together, we obtain (2.7).
Note that the left-hand side in the proposition equals . Applying (2.7) inductively, we find that is a multiple of . More explicitly, we get
We have which implies
where we used that , 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 in a series of lemmas. Let with . For completeness, we first show that the pairing in (1.1) is well-defined.
Lemma 3.1
For and any closed geodesic on the value of the pairing is independent of the choice of the path.
Proof
We have to show that, for any pole of f and any small closed loop around , we have
| 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 has an elliptic expansion in weight of the shape
with coefficients , see Proposition 17 in Zagier’s part of [15]. Here can be chosen freely, but the coefficients will depend on . In particular, we do not need to require that F transforms like a modular form of weight .
If G is another meromorphic function near with an elliptic expansion in weight and coefficients , then it follows from the residue theorem (see [1, Lemma 4.1]) that
| 3.2 |
By definition of the space , the elliptic expansion (in weight 2k) of near is of the shape
| 3.3 |
as , with . Since is holomorphic at , by (3.2) we only need its elliptic expansion coefficient (in weight ) of index at . By [1, Lemma 3.1, Lemma 5.4] this coefficient is given by
| 3.4 |
where is the discriminant of , denotes the -th Legendre polynomial, and we wrote . Using that 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 . This shows (3.1), and finishes the proof.
For and with we consider the Petersson Poincaré series [27]
| 3.5 |
It is straightforward to check that only depends on the class of in and defines a meromorphic modular form in with residue divisor . For a given divisor we put
Then . In order to prove Proposition 1.1, we need to show that for any , and that is uniquely determined by this condition among all forms in with residue divisor D.
Lemma 3.2
For any we have .
Proof
It suffices to prove the lemma for and being a closed geodesic on . In this case the claim follows almost verbatim from the proof of [23, Theorem 3]: Using the unfolding argument one obtains that is a finite linear combination of integrals of the shape for certain real binary quadratic forms [A, B, C] of negative discriminant. Since for any such quadratic form, [23, Lemma 2] shows that these integrals all vanish.
Lemma 3.3
If satisfies and for any , then .
Proof
Under the above assumptions, is a cusp form of weight 2k for with
for any closed geodesic on . 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 , using [23, Theorem 2]. We abbreviate and put for . Then a direct computation shows that we have
| 3.6 |
where . Now, let us assume that is real for all hyperbolic . Then (3.6) says that
| 3.7 |
for every hyperbolic . We can write with the cusp forms and . Note that . Now, (3.7) becomes
Since this equation holds for all hyperbolic , it also holds with replaced by . Taking the sum or the difference of the two resulting equations, we obtain that
for every hyperbolic . By [23, Theorem 2 ii)], this implies that and , hence . The case when is purely imaginary for all hyperbolic is analogous.
Alternatively, one may use the Eichler-Shimura isomorphism to deduce that .
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 with finite covolume, by the same arguments as above.
We close this section with some remarks about further properties of the meromorphic modular forms .
Remark 3.4
One can show that is a constant multiple of , where denotes the higher Green function on , see [7, 26].
It is well-known that is orthogonal to cusp forms with respect to a suitable regularized Petersson inner product, see [7, 27].
We have seen above that the real parts of the cycle integrals of vanish. In [1, 24] it was proved that if is a CM point, then the imaginary parts of (certain linear combinations of) the cycle integrals of are rational (up to some normalizing factors involving powers of and the disciminant of ).
Theta lifts and meromorphic modular forms
Twisted theta functions and theta lifts
Let L be the lattice from Sect. 2.2. Let be a fundamental discriminant (possibly 1) and let be such that . We let be the generalized genus character on binary quadratic forms as in [20], which can also be viewed as a function on using the identification of with explained in Sect. 2.2. For notational convenience we put
Following [9], we define the twisted Siegel theta function by
It transforms like a modular form of weight for in and is -invariant in z. Similarly, following [22] we define the twisted Millson theta function by
with the polynomial as in (2.2). The Millson theta function transforms like a modular form of weight for in and is -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 with . For a harmonic Maass form we define the theta lift
where denotes the usual invariant measure on , the integrals are regularized as in [10, Sect. 6], and and denote the normalizing constants
| 4.1 |
We remark that Zemel [34] investigated a similar lift for lattices of more general signature (with ) 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 but without the raising operator in front) was considered in [4], where it was shown to be a linear combination of higher Green functions evaluated at Heegner divisors in the first variable.
Here, we will show that the lift defines a meromorphic modular form of weight 2k for with poles of order k at a linear combination of Heegner divisors . 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 be an integer and let with . We compute the theta lift of the Maass Poincaré series , defined in Sect. 2.4, and obtain an explicit representation of the lift in terms of the meromorphic modular form
| 4.2 |
of weight 2k for . Here we put for . Let be the Heegner point corresponding to . Then a short computation using
shows that
with the Petersson Poincaré series defined in (3.5). In particular, is the canonical meromorphic modular form of weight 2k for the rescaled Heegner divisor
Proposition 4.1
Let with , let with and with . Then we have
Remark 4.2
Let f be a harmonic weak Maass form of weight . Since the Poincaré series generate the space of weight harmonic weak Maass forms, we directly deduce that the lift defines a meromorphic modular form of weight 2k for with poles of order k at a linear combination of Heegner divisors .
Proof
Let be odd. Using Lemma 2.1 we can write
where 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
Recall from (2.3) that . By Proposition 2.2 (with ) we find
The quotient of Gamma functions can be simplified to using the Legendre duplication formula . Moreover, we have , independently of , where is the quadratic form corresponding to X under the identification of with explained in Sect. 2.2. Hence we can rewrite the sum into a multiple of . Taking into account the normalizing constant 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 with . For large enough, the theta lift of has the Fourier expansion
with the rational constant .
Proof
Suppose that is odd. For simplicity we assume that , but the proof for is very similar. We write the theta function 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
where is the normalizing constant defined in (4.1), is a positive definite sublattice of , equals 1 if and i if , we put , and we wrote,1
for the Fourier expansion of .
We first show that the contribution for vanishes. Note that
| 4.3 |
for any , so we have
Now the contribution from the summand equals
It follows from (4.3) that . In particular, the Fourier expansion of the theta lift has no constant term.
For we have by (2.4), and hence
by (2.5). Using (2.6) (with ) we have for the evaluation
Note that and for . Using (2.5) we find
since for one of the factors in (2.5) will be 0. By (2.4) we have . Taking everything together, we obtain after simplification
where we also used that . Putting in the normalizing constant 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 as a Poincaré series
where
By the unfolding argument we obtain as in the proof of [8, Theorem 5.3] that
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 and with , and let L be the lattice defined in (2.1). Recall from [18, 29] that the space of holomorphic vector-valued modular forms of weight for is isomorphic to the space of skew holomorphic Jacobi forms of weight and index N. Similarly, is isomomorphic to the space of holomorphic Jacobi forms of weight 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 and . In particular, there is a newform theory for these spaces of vector-valued modular forms.
We let be the space of cuspidal newforms of weight 2k for on which the Fricke involution acts by multiplication with .
By the Shimura correspondence, it is isomorphic as a module over the Hecke algebra to , compare [20, 29, 30]. Similarly, the space of cuspidal newforms of weight 2k for on which the Fricke involution acts by multiplication with is isomorphic as a module over the Hecke algebra to .
Let be one of the representations and . For every there is a Hecke operator acting on . The action on Fourier expansions can be computed explicitly, compare [18, 30]. For example, if p is a prime with , and we let and , then we have
| 5.1 |
where if and if . There are similar formulas for general . The Hecke operators act on harmonic Maass forms in an analogous way, and the action on Fourier expansions is the same.
We let and be a fundamental discriminant, and we put . For a normalized newform we let be the totally real number field generated by its Fourier coefficients, and we let be its Shimura correspondent, which we may normalize to have coefficients in , as well. Then, by [8, Lemma 7.2] there exists a harmonic Maass form whose principal part has coefficients in , and which satisfies , where we write for the Petersson norm of g. We let
be the twisted Heegner divisor corresponding to f. Then the canonical meromorphic modular form with residue divisor is given by
| 5.2 |
with defined in (4.2). On the other hand, by Proposition 4.1 the meromorphic modular form can be obtained as the theta lift of f. In particular, by Proposition 4.3, we have the Fourier expansion
| 5.3 |
with a rational constant .
Theorem 5.1
Let be as above, and let be the canonical meromorphic modular form of weight 2k for as in (5.2). Then the following are equivalent.
We have .
We have for all .
All Fourier coefficients of are contained in .
Proof
The implication (2) (3) follows from the Fourier expansion of in (5.3). Similarly, the implication (3) (1) follows from (5.3) since the first coefficient of is given by .
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 . Moreover, using and , where is the -eigenvalue of G, we see that
is a weakly holomorphic modular form with principal part defined over . Since the space of weakly holomorphic modular forms for has a basis of forms with rational Fourier coefficients by a result of McGraw [25], we obtain that all Fourier coefficients of lie in . Now using the explicit formula (5.1) for the action of on the coefficients of f, we see that is a rational linear combination of coefficients for some and coefficients of , which lie in by the above discussion. Hence, if then it follows by induction that for every .
Theorem 5.2
For as above, there exists a unique meromorphic modular form with the following properties.
The residue divisor is given by .
We have for every closed geodesic C in with .
The first Fourier coefficient of vanishes.
Moreover, all Fourier coefficients of lie in , and we have
| 5.4 |
for every with .
Proof
By (5.3) the first coefficient of is given by . Hence
has the desired properties. Taking the bilinear pairing with any yields
which gives the formula (5.4).
In order to show that all Fourier coefficients of lie in we use Hecke operators. First note that we have
| 5.5 |
for any . This may be checked using the explicit action (5.1) of on the Fourier expansion of f and the action of on , which can be computed as in the proof of [33, Eq. (36)]. For example, if p is prime with we have
and there are similar formulas for any Hecke operator . We leave the details of the verification of (5.5) to the reader. Now (5.5) implies that
| 5.6 |
with the weakly holomorphic modular form . Again, it follows from [25] that has all coefficients in , which by (5.3) implies that has all coefficients in . Using the explicit action of on Fourier expansions and the fact that the first Fourier coefficient of vanishes, we obtain by induction from (5.6) that all coefficients of lie in .
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 be a newform of weight 2 for . The functional equation of the twisted L-function implies that . Hence, it is natural to consider the (non-)vanishing of at . 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 , let be its Shimura correspondent with coefficients in , and let be a harmonic Maass form with principal part defined over and . Then the following are equivalent.
We have .
We have .
The proof of this theorem consists of three main steps.
- First, one may show that the canonical differential for the (degree 0) twisted Heegner divisor
is defined over a number field if and only if . This is done by constructing as a regularized theta lift and using Hecke operators. 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 is defined over a number field if and only if some integral multiple of the Heegner divisor is a principal divisor. This is in turn equivalent to saying that the Néron-Tate height of vanishes.
Lastly, by the Gross-Zagier Theorem [21] the Néron-Tate height of is a multiple of , 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 for a newform is equivalent to the algebraicity of the coefficient , where is a harmonic Maass form whose principal part is defined over and which satisfies , with 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 has to be replaced with a height pairing on the Kuga-Sato variety of dimension over , and the Heegner divisor has to be replaced with its corresponding Heegner cycle in the Chow group of codimension k cycles on . Concerning step (2) of the above proof sketch, it remains to show that the Fourier coefficients of the canonical meromorphic modular form are contained in if and only if the Heegner cycle corresponding to 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 , 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 (assuming that it is principal) is a multiple of .
The converse direction seems to be more difficult. Assuming that the coefficients of are contained in , we would need to construct a rational function F on a suitable codimension cycle U in the Kuga-Sato variety such that the divisor of F is the Heegner cycle corresponding to . For , the cycle U is just the whole modular curve , and it was shown in [8] that the rational function F with divisor can be constructed as the Borcherds product attached to f, but for 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
note that, in contrast to [8, Theorem 5.3] we put into c(n, h, v).
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.:
-series and weight
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,
-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,
) 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
-series. II. Math. Ann. 278(1–4), 497–562 (1987) [Google Scholar] -
21.Gross, B.H., Zagier, D.B.: Heegner points and derivatives of
-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
-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.
