Meromorphic quasi-modular forms and their -functions
Abstract.
We investigate the meromorphic quasi-modular forms and their -functions. We study the space of meromorphic quasi-modular forms. Then we define their -functions by using the technique of regularized integral. Moreover, we give an explicit formula for the -functions. As an application, we obtain some vanishing results of special -values of meromorphic quasi-modular forms.
1. Introduction
Given a holomorphic cusp form of weight on , where , the Dirichlet -function associated to is the series
It is well-known that satisfies a functional equation under and admits a meromorphic continuation to the complex plane. We can extend the -function to holomorphic quasi-modular forms by Fourier expansions naturally. Due to the work of Kaneko and Zagier [12], holomorphic quasi-modular form is always a linear combination of iterated derivatives of modular forms and . So their -function comes from shifts of the -functions of modular forms and the -function of . One can also find more details of holomorphic quasi-modular form and their -functions in [1, 5, 24].
However, everything goes differently if we consider meromorphic quasi-modular forms. If is a weakly holomorphic modular form, then it has exponential growth at infinity. So in this case, the Dirichlet series associated to the Fourier coefficients of never converges. To overcome this problem, we need to introduce the regularized integrals. The regularized integrals and -functions of weakly holomorphic modular forms have been studied in [3]. Löbrich and Schwagenscheidt studied the -values of some special meromorphic modular forms using Cauchy principal valued integrals in [16]. McGady defined and investigated the -functions for meromorphic modular forms which are holomorphic at infinity in [17].
In this paper, we will study the structure of meromorphic quasi-modular forms and their Dirichlet -functions. Unlike the classical case, the depth of meromorphic quasi-modular forms could be larger than , so a meromorphic quasi-modular form may not become a linear combination of iterated derivatives of meromorphic modular forms. In fact, we have the following structure theorem.
Theorem 1.1.
We have the following decomposition of -vector space of meromorphic quasi-modular forms
We next generalize the Rankin–Cohen bracket to meromorphic modular forms. Cohen [4] proved that the Rankin–Cohen bracket of two holomorphic modular forms is again a modular form. We prove that the Rankin–Cohen bracket of two meromorphic modular forms is also a meromorphic modular form. Together with Theorem 1.1, we prove
Theorem 1.2.
Let , be two meoromorphic quasi-modular forms in the space or of weight , and depth , respectively. Then
where is the subalgebra of consisting of meoromorphic cusp forms, i.e. those with zero constant terms in their Fourier expansions.
To state our main result, we need to generalize the regularized integral to meromorphic quasi-modular form . For the meromorphic quasi-modular form , we consider its regularized integral
We will give an explicit formula for in Theorem 8.5. The -function admits some classical properties as usual. More precisely, we have
Theorem 1.3.
Let be a meromorphic quasi-modular form of weight and depth . Let be the component functions corresponding to . Then we have
-
(i)
The complete -function extends to a meromorphic function for all with the functional equation
-
(ii)
The complete -function has only possibly simple poles either at or at all integers within , Moreover, the residue of at an integer is
-
(iii)
In general, the -functions of component functions satisfy the functional equations
The paper is organized as follows. In Section 2, we introduce the notations and basic properties of meromorphic quasi-modular forms. In Section 3, we recall the Maass–Shimura derivative and Serre derivative. In Section 4 and Section 5, we prove the structure theorem of meromorphic quasi-modular forms and generalize the Rankin–Cohen bracket to meromorphic quasi-modular forms. In Section 6 and Section 7, we introduce the regularized integral for meromorphic functions. In Section 8, we define the -function of meromorphic quasi-modular form through regularized integral and give an explicit formula for the -function. Finally, in Section 9, we give some special values of -functions.
2. meromorphic quasi-modular forms
Let be the Poincaré upper half-plane. A meromorphic modular form of weight is a meromorphic function on which satisfies
and which is also meromorphic at infinity, that is, having a Laurent (Fourier) expansion
If is holomorphic on but meromorphic at infinity, we say that is a weakly holomorphic modular form. If is further holomorphic at infinity, i.e. for all , we say that is a holomorphic modular form. Note that a meromorphic modular form usually has poles on and has exponential growth at infinity.
We denote by (resp. , ) the -vector space of meromorphic (resp. weakly holomorphic, holomorphic) modular forms of weight , we denote also by
the graded -algebra of meromorphic (resp. weakly holomorphic, holomorphic) modular forms.
As usual, we define for integer the Eisenstein series
where is the -th Bernoulli number and
The Eisenstein series are holomorphic modular forms of weight for . In particular, the Eisenstein series and are algebraically independent and generate the whole graded ring of meromorphic modular forms. To be specific, as graded -algebras, one has
and
where is the unique normalized cusp form of weight and denotes the homogeneous localization at the prime ideal .
However in the case , the Eisenstein series is no longer modular. In fact, it verifies the following transformation rule
(1) |
for any . In general, we define
Definition 2.1.
A meromorphic quasi-modular form of weight is a meromorphic function on with a collection of component functions over , such that
-
(i)
each is meromorphic on and is also meromorphic at infinity,
-
(ii)
the function verifies the transformation rule
(2)
If , the number is called the depth of . We write also for the -th component of a meromorphic quasi-modular form .
Remark.
It can be seen that the Eisenstein series is of weight and depth . A meromorphic quasi-modular form of depth is nothing but a meromorphic modular form.
We will denote by (resp. , ) for the set of meromorphic (resp. weakly holomorphic, holomorphic) quasi-modular forms of weight and depth and (resp. , ) the -vector space of meromorphic (resp. weakly holomorphic, holomorphic) quasi-modular forms of weight and at most depth . Analogously, write
for the -algebra of meromorphic (resp. weakly holomorphic, holomorphic) quasi-modular forms.
Similar to holomorphic quasi-modular forms, each component function of a meromorphic quasi-modular form is again a meromorphic quasi-modular form.
Lemma 2.2.
Let be a meromorphic quasi-modular form of weight and depth with component functions . Then for any , the function is meromorphic quasi-modular form of weight and depth . More precisely, for any ,
In particular, we have that and is indeed a meromorphic modular form of weight .
Proof.
Remark.
Since there are no holomorphic quasi-modular forms of negative weights, we know a holomorphic quasi-modular form always has weight and depth . However, a meromorphic quasi-modular form can has arbitrary weight and depth . This may be the first glimpse of how a meromorphic quasi-modular form differs from a holomorphic quasi-modular form.
Meromorphic quasi-modular forms, especially those with nonpositive weights, enjoy certain same properties like Lemma 2.2 compared with holomorphic modular quasi-forms. For the reader’s convenience, we will sketch the proofs and reassure the reader that some arguments of holomorphic quasi-modular forms still work here.
3. Differential operators
This section gives an introduction to the differential operators we will be using.
Let and be integers. Let be a meromorphic function on the upper-half plane. For later use, we recall that the Maass–Shimura derivative of is
where and . We recall also the Serre derivative of
We will occasionally abbreviate by and in an abuse of notations. We will often use the following identities found by Ramanujan.
Lemma 3.1.
Let be an integer and be a meromorphic function on the upper-half plane. For any , we have
Moreover, we have the following explicit formula for times Maass–Shimura derivative.
(3) |
where and for , we set and to be the identity operator.
Proof.
The first and second identities are immediate calculations, so we just check the last one. We prove it by induction on . When , the identity is just the definition. Applying Maass–Shimura derivatives on , we find that equals
In particular, when and the coefficients in vanish for , in that case we obtain the Bol’s identity (cf. [15])
(4) |
This indicates that is in fact an -invariant differential operator on . This features, as we will see later, will significantly change the behaviors of meromorphic quasi-modular forms.
Lemma 3.2.
Given a meromorphic quasi-modular form with the component functions , we have
where by convention .
Proof.
Since , by applying the differential operator to both sides of , we get
We complete the proof by applying Lemma 3.1. ∎
The following lemma implies that the differential operator usually increases the depth of a meromorphic quasi-modular form by (note that in holomorphic case it always does).
Proposition 3.3.
The space of meromorphic quasi-modular forms is stable under the derivation . It acts on by increasing the weight by and increasing the depth by at most
More precisely, for a meromorphic modular form , we have if and if .
Proof.
The Serre derivative, however, usually preserves the depth of a quasi-modular form. We now state an analogous result.
Proposition 3.4.
The space of meromorphic quasi-modular forms is stable under the Serre derivative . It acts on by increasing the weight by ,
In particular, if is a meromorphic modular form of weight , then is a meromorphic modular form of weight .
Proof.
Lemma 3.5.
Let be a meromorphic modular form of weight . Then
-
(i)
If , we have ,
-
(ii)
If and , then we have ,
-
(iii)
If and , then we have .
Proof.
Applying iteratively Lemma 3.2, we obtain
(5) |
which is always nonvanishing when or . This proves the assertions and .
For the third one, keep in mind that is a meromorphic modular form of weight by Bol’s identity. Therefore by the assertion , we have
Lemma 3.6.
Let be a nonnegative integer. Then the sequence
of graded -vector space is left exact where when and equals to when .
Proof.
Let be a meromorphic quasi-modular form in the kernel of the Serre derivative . As , we have the identity of logarithmic derivative . Therefore is a power of with and . ∎
4. Structure of meromorphic quasi-modular forms
In this section, we study the structure of the graded ring of meromorphic quasi-modular forms.
Throughout this section, we set . The -algebra of meromorphic quasi-modular forms is graded by the weight and filtered by the depth
Lemma 4.1.
Let and be integers with . Then we have the following split exact sequence
Proof.
Let be a meromorphic quasi-modular form. We recall that by Lemma 2.2, the last component is in fact a meromorphic modular form of weight . Thus the above sequence is exact. From the modular transformation (1) of , we know that for any . Thus the map is a section of the map , so the above exact sequence is also split. ∎
Theorem 4.2.
The graded -algebra of meromorphic quasi-modular forms is generated by meromorphic modular forms and
where the depth of a meromorphic quasi-modular forms is exactly the degree of within it.
Proof.
Induction on the depth of . The statement is straightforward when the depth is . As explained in Lemma 4.1, we see that the form has depth . Therefore by induction, for any there exist meromorphic modular forms of weight such that . ∎
Let be the associated graded -algebra with respect to the depth . Then is a bigraded ring
Then the induced derivative on is homogeneous, increasing the weight by and depth by .
Proposition 4.3.
We have the following left exact sequence of bigraded -vector spaces
In fact, the induced derivative
is a bijection when and is a zero map if .
Proof.
It follows from Lemma 3.5 that the map is injective when and is a zero map when . We next show the surjectivity of for . Using the fundamental identity , we observe that for a meromorphic modular form of weight ,
(6) |
Here we recall that is the Serre derivative of , which is modular of weight . Hence
Since , the map is surjective. ∎
Likewise, we have an induced homogeneous Serre derivative on , which increases the weight by and preserves the depth.
Proposition 4.4.
We have the following left exact sequence of bigraded -vector spaces
Proof.
A direct computation
yields that . Therefore the kernel of is the graded subalgebra generated by powers of and . ∎
It is well-known that a holomorphic quasi-modular form is always a linear combination of iterated derivatives of holomorphic modular forms and . The theorem below shows this fails for meromorphic quasi-modular forms. This is one of the major differences between the holomorphic case and meromorphic case.
Theorem 4.5.
Let be a meromorphic quasi-modular form with depth . Then is never a linear combination of iterated derivatives of meromorphic modular forms.
Proof.
Suppose that admits a decomposition , where is a meromorphic modular form of weight . Lemma 3.5 shows that when or , the depth of is exactly . So we are left to consider only . If , the depth of is , which is strictly smaller than , and if the depth of is , also strictly smaller than . This implies that the depth of is always , which leads to a contradiction. ∎
Finally, we give the proof of the decomposition of the space of meromorphic quasi-modular forms.
Proof of Theorem 1.1.
Lemma 3.5 indicates that each component in the first and last parts should have different depth and respectively. In the middle part, applying repeatedly Proposition 3.3 we find that each component has depth for . So the above sum runs through all depths and must be a direct sum of -vector spaces. It remains to show every meromorphic quasi-modular form has such decomposition. We divided the proof into three parts.
Part 1.
When , the result is direct. For , on account of the computation in (5) (or using Proposition 4.3), the -th component function of
is zero, so we complete the proof by induction on .
Part 2.
In this case, since , the induction argument still works unless the depth goes to less than . This implies that we can find where and such that
So we reduce the case to the case .
Part 3.
We claim that for any such there exists a meromorphic quasi-modular form so that . Then the proof will be converted to the first part. We will prove the claim by induction on .
Starting from , the only possible case is , . In this case , the result is direct since it corresponds to . We assume that . If , then the result is also direct since .
Then we can assume that . By Theorem 4.2, there exists a meromorphic modular form of weight such that for some . The same calculation as in equation (6) yields that can be represented as the linear combination
On the right-hand side, the first term comes from the derivative of , which has exactly weight and depth . So by induction assumption, we can find a weight meromorphic quasi-modular form such that
Then by applying the operator , we get
where
Besides, the second term and the last term on the right-hand side and and the function all have weight and depth . So we find that
By induction assumption, we know there exists so that
which proves the previous claim. ∎
Remarks.
-
(i)
Theorem 1.1 shows that every meromorphic quasi-modular form can be written uniquely (up to Bol’s identity) as a linear combination of iterated derivatives of meromorphic modular forms and iterated derivatives of quasi-modular forms of weight and depth .
-
(ii)
When , we get only the first part, thus every meromorphic quasi-modular form of nonpositive weight is just a linear combination of iterated derivatives of meromorphic modular forms of nonpositive weights.
-
(iii)
For holomorphic quasi-modular form this reduces to the well-known (see Zagier [24, Prop. 20])
where the depth comes from the iterated derivatives of the Eisenstein series which generates .
-
(iv)
According to the work of Paşol and Zudilin [19], it is reasonable to conjecture that all magnetic meromorphic quasi-modular forms come from iterated derivatives with of (quasi-)modular forms with Fourier expansion in .
5. Rankin–Cohen brackets of meromorphic modular forms
In this section, we introduce the Rankin–Cohen brackets of meromorphic modular forms. The Rankin–Cohen brackets of holomorphic modular forms has been studied in many literature. The reader can find details in [5].
We first introduce the Cohen–Kuznetsov series associated with a meromorphic modular form. For holomorphic modular forms these series were originally introduced by Cohen [4] and Kuznetsov [13]. When is a positive integer, our series is the same as Cohen and Kuznetsov. When is a negative integer, we will define a minus series and a plus series.
Definition 5.1.
Let be a meromorphic modular form of weight . We define its Cohen–Kuznetsov series by
Similarly, we can define by replacing by . When , by convention, , and starts from the term . Moreover, we define the slash operator on by
where .
We first give some basic properties of Cohen–Kuznetsov series. When has positive weight, similar results can also be found in [5].
Proposition 5.2.
Let be a meromorphic function of weight . Suppose is a nonpositive integer. Then
-
(i)
We have
where denotes the operator or .
-
(ii)
The functions and are linked by
-
(iii)
The series and commutes with the slash operator up to a factor. Namely, for any , we have
Proof.
. Since is a modular form of weight , one has
The proof for is similar.
. We note that if is a meromorphic modular form of positive weight , then by Lemma 3.1, we have
So when the weight of is nonpositive, we can apply the result above to and get
. Similar to the proof above, we start from the positive weight case. By Lemma 3.1, we have
Then
Finally, we can complete the proof by applying
As for the series , we just need to apply . ∎
Then we focus on the minus Cohen–Kuznetsov series.
Proposition 5.3.
Let be a meromorphic modular form of weight . Suppose is a nonpositive integer. Then for the minus part of Cohen–Kuznetsov series, we have the following relations
-
(i)
The functions and are linked by
-
(ii)
The function also commutes with the slash operator, i.e. for any , we have
-
(iii)
The function has the transformation
Proof.
. The proof is similar to . By definition, we have
By changing , we get
The proof of and are similar to in Proposition 5.2, we omit the details here. ∎
Definition 5.4.
Let be two meromorphic modular forms of weight respectively. Then the -th Rankin–Cohen bracket of and is defined by
Theorem 5.5.
Let be two meromorphic modular form of weight respectively. Then the -th Rankin–Cohen bracket is a meromorphic modular form of weight .
Proof.
The case of is a result of Cohen [5]. So we assume that at least one of is nonpositive. By symmetry , we may assume that . There are exactly two possibilities, either one of and is positive or none of and is positive. For these two parts, we further separate them into the several cases depending on .
Part 1. ,
Case 1. . We consider the product , it is equal to
where
Here we use the identity for . Thus we have
On the other hand, by Proposition 5.2 and 5.3, for any , we have
This implies that is a meromorphic modular form of weight .
Case 2. . Then the terms where in the Rankin–Cohen bracket vanish since the binomial coefficients become zero. Put , the Rankin–Cohen bracket of and turns out to be
(7) |
Therefore,
Notice that is a meromorphic modular form of positive weight , so is a multiple of , thus also a meromorphic modular form of weight .
Part 2. ,
Case 1. . we consider the product . With the same calculation as in Part 1, the product is
So the Rankin–Cohen bracket is again a meromorphic modular form of weight .
Case 2. . Put , the same calculation as in shows that
Thus
is a multiple of , which reduces to Case 1 in Part 1, since is a meromorphic modular form of positive weight .
Case 3. . We note that is non-vanishing if and only if and is non-vanishing if and only if . So when , at least one of these two binomials vanishes, this implies that the Rankin–Cohen bracket
is always vanishing in this case.
Case 4. . Similar to equation , letting we find that
Thus
is still a meromorphic modular form of weight . ∎
The following theorem by Lanphier [14] and El Gradechi [6], originally stated for positive weights modular forms, can be also extended to negative weights modular forms.
Theorem 5.6.
Let be two meromorphic modular forms of weight , respectively. Let be a positive integer. Suppose that or . Then we have
(8) |
and
(9) |
where
and
Proof.
Proposition 4.6 in [6] shows that for positive weights , the constants and are mutually inverse
(10) |
Define now the following two -subalgebras of with
Then the space is generated by positive weight derivatives of positive weight meromorphic modular forms and the space is generated by nonpositive weight derivatives of nonpositive weight meromorphic quasi-modular forms. It can be deduced from Laphier–El Gradechi formula (9) that any product on each space can be rewritten as linear combination of iterated derivatives of Rankin–Cohen brackets. In particular, we get
Proof.
Remark.
In fact, the Rankin–Cohen brackets can be extended to meoromorphic quasi-modular forms using and . Let and be two meoromorphic quasi-modular forms. Their (Serre–)Rankin–Cohen brackets can be defined as
where and are meromorphic quasi-modular forms in .
6. Fourier coefficients of meromorphic quasi-modular forms
Let us denote by
the standard fundamental domain of the translation in . We set also the trimmed fundamental domain for by
Let be a meromorphic function on . For any pole of with order , let
(11) |
be the principle part of the Fourier expansion of at and
(12) |
be the principle part of the Laurent expansion of at . We put also
and
where . Note that the expansion in Appendix shows that has the same principle part at as does.
We have the following estimation on Fourier coefficients of meromorphic quasi-modular form
Proposition 6.1.
Let be a meromorphic quasi-modular form of weight with poles and principle parts as described above. Then we have
Proof.
Let be all the poles of in . Removing all the principle parts with all these kinds of , we have finally a holomorphic function in
So is bounded in , say by . Note the function is holomorphic within the closure of the domain . Then using Cauchy integral formula at infinity, the -th coefficient of is bounded by
(13) |
On the other hand, the -th coefficient of the polylogarithm function is
Combining with the estimation , we get the desired result. ∎
Weakly holomorphic quasi-modular forms have much less growth than meromorphic quasi-modular forms. One has
Proposition 6.2.
Let be a weakly holomorphic quasi-modular form with . Then for any , we have
Proof.
When is a weakly holomorphic modular form, then we have the following estimation [20] on the Fourier coefficients of
(14) |
Theorem 4.2 shows that every weakly holomorphic quasi-modular form is of the form
Note that the order of at infinity is at most for any . Combining the estimation , we get
since for any . ∎
Remark.
In fact, we can get a more accurate estimation by using the Circle Method to weakly holomorphic quasi-modular forms
7. Regularized integrals of meromorphic functions
At the start of this section, we recall the -functions of modular forms. For a cusp form of weight on , the completed -function of is just the Mellin transform of
This completed -function is connected with the Dirichlet series of in the following way
However in general, a meromorphic quasi-modular form may have exponential growth at cusps and polynomial growth at poles. To overcome this, we now construct the regularized integrals of meromorphic functions. This regularization procedure can be divided into two parts, the regularization at infinity (and hence at ) and the regularization at positive real numbers.
Following [2], under the assumption that has at most linear exponential growth at infinity, we give the definition of regularized integral of .
Definition 7.1.
Let be an analytic function with at most linear exponential growth for large . If the integral
has a continuation to , then the regularized integral of is defined to be
We use the notation to indicate a regularized integral.
Similarly, if has at most linear exponential growth at the cusp , then we can define the regularized integral of at using the reflection . This is to say,
For integrals of meromorphic functions near real positive poles, we will use Hadamard regularization. The idea of regularizing a divergent integral can be traced back to Cauchy. Precise definition of such regularized integrals was firstly introduced by Hadamard [9] in his study of Cauchy problem for differential equations of hyperbolic type. The interpretation of Hadamard regularization using meromorphic continuation was due to Riesz [21, 22]. Various theories and generalization of Hadamard regularization can be found in the later literature. Gelfand–Shilov [8] formalized Hadamard regularization in the framework of generalized functions. Afterwards, the concept of generalized functions was extended to hyperfunctions by Sato [23]. Due to space constraints we will neglect the technical and theoretical details in this paper. The reader can find more precise presentations in the previous mentioned articles and the books [10, 11].
We can use meromorphic continuation to give the following definition.
Definition 7.2.
Let be a meromorphic function in a neighbourhood of with only one real positive pole , then the regularized integral of from to is defined as
where the suffix indicates the constant term in the Laurent expansion of at .
As already mentioned, there are different approaches of Hadamard regularization. The following proposition explains why they are actually equivalent.
Proposition 7.3.
Let be a meromorphic function in a neighbourhood of with only one real positive pole of order . Let . The the following different approaches of regularization coincides
-
(i)
The meromorphic continuation of the integral by Riesz
-
(ii)
The integral of in the sense of Sato’s hyperfunction. Equivalently, the Cauchy principal valued integral by
where (resp. ) is a path from to above (resp. below) the real axis.
-
(iii)
The Hadamard finite part integral of
where stands for the constant term in the Laurent expansion with respect to .
-
(iv)
The following integral in the sense of pairing the Schwartz distribution with
where stands for the Cauchy principal value of the integral.
-
(v)
The following Cauchy principle value given by Sokhotski–Plemelj formula, i.e.
where tends to on both sides of real axis.
Proof.
. If is holomorphic then the implication is immediate. So it suffices to check the function . We may set the paths to be the paths agreed with the real axis but modified with small upper (resp. lower) semi-circles at of radius .
For , the meromorphic continuation gives us
Hence, the constant term in the Laurent expansion at is
(15) |
Meanwhile, the integrations along the two small semi-circles give
An elementary calculation shows that this integral coincides with . This implies that the meromorphic continuation yields the same result as hyperfunction.
Observe that the above integrals have always finite part with respect to . It follows that the Hadamard finite part integral has the same value as the integral of hyperfunction. These prove that , and are equal.
. Integrating by parts, for any testing function we get
The last term is a convergent integral with Cauchy principal value as . Thus
is the finite part with respect to . This shows .
In general, let be a function which has a finite number of positive real poles and has at most linear exponential growth at and infinity. Consider finitely many open intervals
Suppose that all poles are contained in these intervals and each interval contains exactly one isolated pole. On every interval, we use the previous approaches from Proposition 7.3 to get a regularized integral. Moreover, on the intervals
we assign to them the regularized integrals from Definition 7.1. At last, we sum up them all. This gives the regularized integral of on , written again as
It is clear that the above definition is independent of the choice of intervals.
8. -function of meromorphic quasi-modular forms
We first define the -function of a meromorphic quasi-modular form through the regularized integral defined in Section 7.
Definition 8.1.
Let be a meromorphic quasi-modular form. Then we define its complete -function by
The Dirichlet -function associated to is defined as
In the following, we will give an explicit formula for for any meromorphic quasi-modular form . We first consider the regularized integral of meromorphic function at infinity.
Lemma 8.2.
Let be a meromorphic function in a neighbourhood of the half-strip with only pole at infinity. Suppose its Fourier expansion at infinity is given as . Then the regularized integral of exists and defines a meromorphic function in . More precisely, we have
Proof.
To avoid problems on negative real axis, following [2], we take only one branch of the incomplete gamma function with the branch cut to be the ray , where is a fixed angle. It is easy to see that when , the integral
is absolutely convergent for any .
Since is holomorphic in a neighbourhood of the half-strip , its Fourier coefficients satisfy . This ensures that the value of the above integral is the absolutely convergent sum
in view of .
We can see that the partial sum defines a holomorphic function of with and . The sum is a finite sum, it can be continued to a holomorphic function of in the open domain
Hence both parts can be extend to a holomorphic function of in a neighbourhood of .
We only need to deal with the term . When and , the integral has well-defined value (cf. [2, Remark after Prop. 3.3]). This extends to a meromorphic function to the whole complex plane in .
At last we remark that the above evaluation is independent of the choice of . ∎
To give the precise formula of regularized integral at positive reals, we will follow the method from McGady [17], whose idea is to remove all the poles with polylogarithm functions. The succeeding calculation deals with the regularized integrals of polylogarithm functions first.
Let be an integer. For with , we define the regularized integral
Lemma 8.3.
Let and be a positive integer. If , then we have
If , we have
(16) |
Proof.
Suppose that is given with , then , so we have the convergent integral
(17) |
Note both sides extend to a holomorphic function of except only when on the imaginary axis. Thus it holds for all .
When , if not on the imaginary axis, the formula just follows from rewriting (17) with the reflection formula of polylogarithm (26) in the Appendix.
The difficulty arises as is on the imaginary axis where , where we encounter a Hadamard regularized integral. In this case we may rewrite the integral as
We recall that when is a positive integer, the polylogarithm is rational function. So the integrand is in fact a rational function of . By the Sokhotski–Plemelj formula in Proposition 7.3, the value of on imaginary axis should be the mean value of limits as tends to on left side and right side of imaginary axis. Moreover, when , by the reflection formula (26), we get
(18) |
So by combining and , we have
Using the reflection formula again, we get
∎
More generally, we will encounter the following integral. Let be an integer, be a real number and , we define
Lemma 8.4.
Let be an integer and be a positive real number. Then the function extends to an entire function for all and
When , we have
When , we have
and
Proof.
When , we can integrate term-wisely. It is immediate that for any
Here the absolute convergence is guaranteed by .
When , we may assume that first. The integral is defined by Hadamard regularization and can not be computed directly. We first evaluate the following convergent integral
When , by the reflection formula , this integral equals to
Hence we have
(19) |
where each term has exponential decay since as grows to infinity. When , we have , so . In this case, the formula becomes
To finish the proof, by combining with Lemma 8.3, we have to show that
(20) |
From , we know the identity is equivalent to
But this is exactly the Euler’s reflection formula.
For general , we consider analytic continuation on both sides and thus obtain the same formula. Indeed, the function is meromorphic only when . It has a unique single pole at with residue
However, this pole cancels with the term in , giving us an entire function . ∎
Now we are able to give the explicit formula for the -function. Choose any real positive number . Suppose has poles in . Put
We define also
Theorem 8.5.
Let be any real positive number. Let be a meromorphic quasi-modular form with prescribed poles and principle parts as above. Let be the component functions of . Suppose that , then we have
Proof.
We divide into two parts:
We first deal with the second part.
Since is holomorphic, by Lemma 8.2, we get
The integral of is shown in Lemma 8.4 which gives
For the first part, by changing the variable , we get
Since is a meromorphic quasi-modular form, we have the transformation
(21) |
by applying equation with the inversion . So we have
Finally, we complete the proof by noting that . ∎
We are now ready to give the proof of Theorem 1.3.
Proof of Theorem 1.3.
The meromorphic continuation and residues follow directly from Theorem 8.5, since is entire in . So we only need to prove the functional equations.
Lemma 2.2 shows that is also a meromorphic quasi-modular form of weight and depth with components . It is thus enough for us to show the functional equation of in . The completed -function of is
(22) |
Here we use the transformation formula again.
On the other hand, under the transform the first integral becomes
(23) |
Combining equation , we get
This proves the functional equation of . ∎
Proposition 8.6.
Let be a meromorphic quasi-modular form. Then its Dirichlet -function is a meromorphic function for all . It has only possible simple poles at positive integers within .
Proof.
We note that has a simple pole at nonpositive integers, so this proposition follows directly from Theorem 1.3. ∎
The operator acts on the Fourier expansion of holomorphic quasi-modular form by
This implies that the Dirichlet -series of is exactly the the shift of the original Dirichlet -series of . Actually, for meromorphic quasi-modular form, we can obtain the same result.
Theorem 8.7.
Let be a meromorphic quasi-modular form. Then we have
and
Proof.
The above identities are nothing but integration by parts. Evidently it is enough for us to prove the case . With integration by parts we have
When large enough, the first term equals to . Clearly, it has a holomorphic continuation to the whole plane in and its value at is just . For the integral, by Lemma 8.2, it has a holomorphic continuation to a neighbourhood of . So we get
Another way to see this is using the precise formula in Lemma 8.2 and the recurrence relation (24). We can deal with regularized integrals at positive real poles and in the same way with integration by parts. At last we get
This gives the identity for . The identity for then follows directly after . ∎
If is a meromorphic modular form, the formula of its -function is much simpler.
Corollary 8.8.
Let be a meromorphic modular form of weight . Then the -function of is
Moreover, it satisfies the following functional equation
Remark.
In particular, when is a weakly holomorphic cusp form, we obtain
This computation coincides with Theorem in [3].
9. Vanishing -values of meromorphic quasi-modular forms
In this section, we give some vanishing results of certain special -values of meromorphic quasi-modular forms.
Proposition 9.1.
Let be a meromorphic quasi-modular form of weight . Let be a negative integer, then
-
(i)
If , then the Dirichlet -function is always entire in . Moreover, when either or , we have
-
(ii)
If , when , we have
Proof.
If , the function has only possibly simple poles at by Theorem 1.3. Meanwhile, also has a simple pole at negative integers. So the Dirichlet -function is entire and whenever or .
If , the function has no poles for but has pole at negative integer , so always vanishes if . ∎
Corollary 9.2.
Let be a meromorphic quasi-modular form of weight . If is a positive integer with , then we have
Remark.
This corollary shows that some periods of the meromorphic quasi-modular form vanish. When is a weakly holomorphic modular form, this recovers Theorem in [3].
Proposition 9.3.
The period of a meromorphic modular form of weight is always vanishing, i.e. for any meromorphic modular form ,
Proof.
Appendix A Special functions
A.1. Incomplete Gamma functions
We give some basic properties of incomplete gamma functions. Let with and , the upper gamma function is defined by
whereas the lower incomplete gamma function is defined by
The two functions are connected by
Integration by parts yields the following recurrence relation for incomplete gamma function
(24) |
Both incomplete gamma functions can be extended with respect to both and . The function extends to a multi-valued function in with a branch point at , and is holomorphic in each sector. If , then is always an entire function in .
For , the incomplete gamma function has exponential growth asymptotic expansion
as (see [18, Section 8.11 (i)]).
A.2. Hurwitz zeta function
Let with and . The Hurwitz function is defined by series expansion
The function has a meromorphic continuation to the whole -plane. It has only a simple pole at with residue .
A.3. Polylogarithm
Let with . The polylogarithm function is defined to be
It has an analytic continuation to a multi-valued holomorphic function in and an entire function in . If is not a nonnegative integer, then the principle branch is holomorphic in with branch cut along .
When , the function is a rational function in and has pole only at . More precisely, we have
Let be a positive integer, then the Laurent series of at is given by
(25) |
The polylogarithm is related to the Hurwitz zeta function by
When , we have the reflection formula of polylogarithm
(26) |
Acknowledgements
The research of the second author was supported by Fundamental Research Funds for the Central Universities (Grant No. 531118010622).
References
- [1] A. Bhand, K. DeoShankhadhar, On Dirichlet series attached to quasimodular forms, Journal of Number Theory (2019), Volume 202, pp. 91-106.
- [2] K. Bringmann, N. Diamantis, S. Ehlen, Regularized inner products and errors of modularity, International Mathematics Research Notices (2017), Issue 24, pp. 7420-7458.
- [3] K. Bringmann, K. Fricke, Z. A. Kent, Special L-values and periods of weakly holomorphic modular forms, Proceedings of the American Mathematical Society 142 (2014), pp. 3425-3439.
- [4] H. Cohen, Sums involving the values at negative integers of L-functions of quadratic characters, Mathematische Annalen (1975), Volume 217 No. 3, pp. 271–285.
- [5] H. Cohen, F. Strömberg, Modular Forms: A Classical Approach, Graduate Studies in Mathematics 179 (2017).
- [6] A. M. El Gradechi, The Lie theory of certain modular form and arithmetic identities, Ramanujan Journal 31, pp. 397–433.
- [7] C. Fox, A Generalization of the Cauchy Principal Value, Canadian Journal of Mathematics (1957), Volume 9, pp. 110-117.
- [8] I. M. Gel’fand, G. E. Shilov, Generalized functions, Volume 1, Moscow Fismatgiz (1962), [English translation, Academic Press (1964)].
- [9] J. Hadamard, Le problème de Cauchy et les équations aux dérivées partielles linéaires hyperboliques (in French), Paris: Hermann & Cie. (1932), p. 542
- [10] A. Kaneko, Introduction to hyperfunctions, Mathematics and its applications, Kluwer Academic Publishers (1988).
- [11] Ram P. Kanwal, Generalized Functions: Theory and Technique, Birkhäuser Boston (1998).
- [12] M. Kaneko, D. Zagier, A generalized Jacobi theta function and quasimodular forms, in: The Moduli Space of Curves, Texel Island, 1994, in: Progr. Math., vol.129, Birkhäuser Boston, Boston, MA, 1995, pp.165–172.
- [13] N. V. Kuznetsov, A new class of identities for the Fourier coefficients of modular forms (in Russian), Collection of articles in memory of Juriǐ Vladimirovič Linnik, Acta Arithmetica 27 (1975), pp. 505–519.
- [14] D. Lanphier, Combinatorics of Maass–Shimura operators, Journal of Number Theory (2008), Volume 128, Issue 8 , pp. 2467-2487.
- [15] J. Lewis, D. Zagier, Period functions for Maass wave forms. I, Annals of Mathematics (2001), Volume 153, Issue 1, pp. 191-258.
- [16] S. Löbrich, M. Schwagenscheidt, Locally harmonic maass forms and periods of meromorphic modular forms, Transactions of the American Mathematical Society 375 (2022), 501-524.
- [17] D. A. McGady, L-functions for Meromorphic Modular Forms and Sum Rules in Conformal Field Theory, Journal of High Energy Physics (2019).
- [18] F. W. J. Olver et al., eds. NIST Handbook of Mathematical Functions, Cambridge University Press (2010), pp. 173-191.
- [19] V. Paşol, W. Zudilin, Magnetic (quasi-)modular forms, arXiv:2009.14609 (2020).
- [20] H. Rademacher and H. S. Zuckerman, On the Fourier coefficients of certain modular forms of positive dimension, Annals of Mathematics (2) 39 (1938), No. 2, pp. 433–462.
- [21] M. Riesz, Intégrales de Riemann-Liouville et potentiels (in French), Acta Scientiarum Mathematicarum 9 (1938), pp. 1–42.
- [22] M. Riesz, L’intégrale de Riemann-Liouville et le problème de Cauchy (in French), Acta Mathematica 81(1949), pp. 1–223.
- [23] M. Sato, Theory of Hyperfunctions, Journal of the Faculty of Science, University of Tokyo, Section 1, Mathematics, Astronomy, Physics, Chemistry (1959).
- [24] D. Zagier, Elliptic Modular Forms and Their Applications, in The 1-2-3 of Modular Forms, Springer-Verlag (2008), pp. 181–245.