This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Meromorphic quasi-modular forms and their LL-functions

Weijia Wang  and  Hao Zhang Yanqi Lake Beijing Institute of Mathematical Sciences and Applications &\& Yau Mathematical Sciences Center, Tsinghua University, Beijing 101408, P. R. China School of Mathematics, Hunan University, Changsha 410082, P. R. China
Abstract.

We investigate the meromorphic quasi-modular forms and their LL-functions. We study the space of meromorphic quasi-modular forms. Then we define their LL-functions by using the technique of regularized integral. Moreover, we give an explicit formula for the LL-functions. As an application, we obtain some vanishing results of special LL-values of meromorphic quasi-modular forms.

1. Introduction

Given a holomorphic cusp form f(τ)=n>0af(n)qnf(\tau)=\sum_{n>0}a_{f}(n)q^{n} of weight kk on SL2()\operatorname{SL}_{2}(\mathbb{Z}), where q=e2πiτq=e^{2\pi i\tau}, the Dirichlet LL-function associated to ff is the series

L(f,s)=n1af(n)ns=(2π)sΓ(s)0f(it)ts1𝑑t.L(f,s)=\sum_{n\geq 1}\frac{a_{f}(n)}{n^{s}}=\frac{(2\pi)^{s}}{\Gamma(s)}\int_{0}^{\infty}f(it)t^{s-1}dt.

It is well-known that L(f,s)L(f,s) satisfies a functional equation under skss\mapsto k-s and admits a meromorphic continuation to the complex plane. We can extend the LL-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 E2E_{2}. So their LL-function comes from shifts of the LL-functions of modular forms and the LL-function of E2E_{2}. One can also find more details of holomorphic quasi-modular form and their LL-functions in [1, 5, 24].

However, everything goes differently if we consider meromorphic quasi-modular forms. If ff is a weakly holomorphic modular form, then it has exponential growth at infinity. So in this case, the Dirichlet series L(f,s)L(f,s) associated to the Fourier coefficients of ff never converges. To overcome this problem, we need to introduce the regularized integrals. The regularized integrals and LL-functions of weakly holomorphic modular forms have been studied in [3]. Löbrich and Schwagenscheidt studied the LL-values of some special meromorphic modular forms using Cauchy principal valued integrals in [16]. McGady defined and investigated the LL-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 LL-functions. Unlike the classical case, the depth of meromorphic quasi-modular forms could be larger than k/2k/2, 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 \mathbb{C}-vector space of meromorphic quasi-modular forms

𝒬k=(l=0k21Dlk2l)(l=0k21Dl𝒬k2l,k2l1)(l=kDlk2l).\mathcal{Q}\mathcal{M}_{k}^{*}=\bigg{(}\bigoplus_{l=0}^{\frac{k}{2}-1}D^{l}\mathcal{M}_{k-2l}^{*}\bigg{)}\bigoplus\bigg{(}\bigoplus_{l=0}^{\frac{k}{2}-1}D^{l}\mathcal{Q}\mathcal{M}_{k-2l}^{*,\,k-2l-1}\bigg{)}\bigoplus\bigg{(}\bigoplus_{l=k}^{\infty}D^{l}\mathcal{M}_{k-2l}^{*}\bigg{)}.

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 ff, gg be two meoromorphic quasi-modular forms in the space 𝒬+\mathcal{Q}\mathcal{M}_{+}^{*} or 𝒬\mathcal{Q}\mathcal{M}_{-}^{*} of weight kk, ll and depth ss, tt respectively. Then

fgk+l(i=0s+tDs+ti𝒮k+l2s2t+2i),fg\in\mathcal{M}_{k+l}^{*}\bigoplus\bigg{(}\bigoplus_{i=0}^{s+t}D^{s+t-i}\mathcal{S}_{k+l-2s-2t+2i}^{*}\bigg{)},

where 𝒮\mathcal{S}^{*} is the subalgebra of \mathcal{M}^{*} 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 ff. For the meromorphic quasi-modular form ff, we consider its regularized integral

Λ(f,s)=0,f(it)ts1𝑑t.\Lambda(f,s)=\int_{0}^{\infty,\ast}f(it)t^{s-1}dt.

We will give an explicit formula for Λ(f,s)\Lambda(f,s) in Theorem 8.5. The LL-function Λ(f,s)\Lambda(f,s) admits some classical properties as usual. More precisely, we have

Theorem 1.3.

Let f𝒬kmero,pf\in\mathcal{Q}\mathcal{M}_{k}^{\text{mero},\,p} be a meromorphic quasi-modular form of weight kk and depth pp. Let f0,,fpf_{0},\cdots,f_{p} be the component functions corresponding to ff. Then we have

  1. (i)

    The complete LL-function Λ(f,s)\Lambda(f,s) extends to a meromorphic function for all ss\in\mathbb{C} with the functional equation

    Λ(f,s)=r=0pikrΛ(fr,krs).\Lambda(f,s)=\sum_{r=0}^{p}i^{k-r}\Lambda(f_{r},k-r-s).
  2. (ii)

    The complete LL-function Λ(f,s)\Lambda(f,s) has only possibly simple poles either at s=0s=0 or at all integers within kpskk-p\leq s\leq k, Moreover, the residue of Λ(f,s)\Lambda(f,s) at an integer nn is

    Ress=nΛ(f,s)={af(0)+afk(0)if n=0inafkn(0)if n0 and kpnk.\operatorname*{Res}_{s=n}\,\Lambda(f,s)=\begin{cases*}-a_{f}(0)+a_{f_{k}}(0)&if $n=0$\\ i^{n}\,a_{f_{k-n}}(0)&if $n\neq 0$ and $k-p\leq n\leq k$\end{cases*}.
  3. (iii)

    In general, the LL-functions of component functions Λ(fm,s)\Lambda(f_{m},s) satisfy the functional equations

    Λ(fm,s)=r=0pmik2mr(m+rr)Λ(fm+r,k2mrs).\Lambda(f_{m},s)=\sum_{r=0}^{p-m}i^{k-2m-r}\binom{m+r}{r}\Lambda(f_{m+r},k-2m-r-s).

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 LL-function of meromorphic quasi-modular form through regularized integral and give an explicit formula for the LL-function. Finally, in Section 9, we give some special values of LL-functions.

2. meromorphic quasi-modular forms

Let ={τ|Im(τ)>0}\mathcal{H}=\{\tau\in\mathbb{C}\,|\,\operatorname{Im}(\tau)>0\} be the Poincaré upper half-plane. A meromorphic modular form of weight kk\in\mathbb{Z} is a meromorphic function on \mathcal{H} which satisfies

f|kγ(τ)(cτ+d)kf(aτ+bcτ+d)=f(τ),where γ=(abcd)SL2(),f|_{k}\gamma(\tau)\coloneqq(c\tau+d)^{-k}f\left(\frac{a\tau+b}{c\tau+d}\right)=f(\tau),\quad\text{where }\gamma=\begin{pmatrix}a&b\\ c&d\end{pmatrix}\in\operatorname{SL}_{2}(\mathbb{Z}),

and which is also meromorphic at infinity, that is, having a Laurent (Fourier) expansion

f(τ)=naf(n)qn,where q=e2πiτ.f(\tau)=\sum_{n\gg-\infty}a_{f}(n)q^{n},\quad\text{where }q=e^{2\pi i\tau}.

If ff is holomorphic on \mathcal{H} but meromorphic at infinity, we say that ff is a weakly holomorphic modular form. If ff is further holomorphic at infinity, i.e. af(n)=0a_{f}(n)=0 for all n<0n<0, we say that ff is a holomorphic modular form. Note that a meromorphic modular form usually has poles on \mathcal{H} and has exponential growth at infinity.

We denote by kmero\mathcal{M}_{k}^{\text{mero}} (resp. k!\mathcal{M}_{k}^{!}, k\mathcal{M}_{k}) the \mathbb{C}-vector space of meromorphic (resp. weakly holomorphic, holomorphic) modular forms of weight kk, we denote also by

=k2k,{mero,!,}\mathcal{M}^{*}=\bigoplus_{k\in 2\mathbb{Z}}\mathcal{M}_{k}^{*},\quad*\in\{\text{\text{mero}},!,-\}

the graded \mathbb{C}-algebra of meromorphic (resp. weakly holomorphic, holomorphic) modular forms.

As usual, we define for integer k2k\geq 2 the Eisenstein series

Ek(τ)=12kBkn=1σk1(n)qn,E_{k}(\tau)=1-\frac{2k}{B_{k}}\sum_{n=1}^{\infty}\sigma_{k-1}(n)q^{n},

where BkB_{k} is the kk-th Bernoulli number and

σk1(n)=d|ndk1.\sigma_{k-1}(n)=\sum_{d|n}d^{k-1}.

The Eisenstein series EkE_{k} are holomorphic modular forms of weight kk for k4k\geq 4. In particular, the Eisenstein series E4E_{4} and E6E_{6} are algebraically independent and generate the whole graded ring of meromorphic modular forms. To be specific, as graded \mathbb{C}-algebras, one has

\displaystyle\mathcal{M} =[E4,E6]=k(a,b)24a+6b=kE4aE6b,\displaystyle=\mathbb{C}[E_{4},E_{6}]=\bigoplus_{k\in\mathbb{N}}\bigoplus_{\begin{subarray}{c}(a,b)\in\mathbb{N}^{2}\\ 4a+6b=k\end{subarray}}\mathbb{C}E_{4}^{a}E_{6}^{b},
!\displaystyle\mathcal{M}^{!} =[Δ1,E4,E6]=k(a,b,n)34a+6b12n=kE4aE6bΔn\displaystyle=\mathbb{C}[\Delta^{-1},E_{4},E_{6}]=\bigoplus_{k\in\mathbb{Z}}\bigoplus_{\begin{subarray}{c}(a,b,n)\in\mathbb{N}^{3}\\ 4a+6b-12n=k\end{subarray}}\mathbb{C}\frac{E_{4}^{a}E_{6}^{b}}{\Delta^{n}}

and

mero=[E4,E6]((0)),\mathcal{M}^{\text{mero}}=\mathbb{C}[E_{4},E_{6}]_{((0))},

where Δ=11728(E43E62)\Delta=\frac{1}{1728}(E_{4}^{3}-E_{6}^{2}) is the unique normalized cusp form of weight 1212 and ((0))((0)) denotes the homogeneous localization at the prime ideal (0)(0).

However in the case k=2k=2, the Eisenstein series E2E_{2} is no longer modular. In fact, it verifies the following transformation rule

(1) E2|2γ(τ)=E2(τ)+6πi(ccτ+d).E_{2}|_{2}\gamma(\tau)=E_{2}(\tau)+\frac{6}{\pi i}\left(\frac{c}{c\tau+d}\right).

for any γ=(abcd)SL2()\gamma=\begin{pmatrix}a&b\\ c&d\end{pmatrix}\in\operatorname{SL}_{2}(\mathbb{Z}). In general, we define

Definition 2.1.

A meromorphic quasi-modular form of weight kk\in\mathbb{Z} is a meromorphic function ff on \mathcal{H} with a collection of component functions f0,f1,,fpf_{0},f_{1},\cdots,f_{p} over \mathcal{H}, such that

  1. (i)

    each fif_{i} is meromorphic on \mathcal{H} and is also meromorphic at infinity,

  2. (ii)

    the function ff verifies the transformation rule

    (2) (f|kγ)(τ)=r=0pfr(τ)(ccτ+d)r,for any γSL2().(f|_{k}\gamma)(\tau)=\sum_{r=0}^{p}f_{r}(\tau)\left(\frac{c}{c\tau+d}\right)^{r},\quad\text{for any }\gamma\in\operatorname{SL}_{2}(\mathbb{Z}).

If fp0f_{p}\neq 0, the number pp is called the depth of ff. We write also fr=Qr(f)f_{r}=Q_{r}(f) for the rr-th component of a meromorphic quasi-modular form ff.

Remark.

It can be seen that the Eisenstein series E2E_{2} is of weight 22 and depth 11. A meromorphic quasi-modular form of depth 0 is nothing but a meromorphic modular form.

We will denote by 𝒬kmero,p\mathcal{Q}\mathcal{M}_{k}^{\text{mero},\,p} (resp. 𝒬k!,p\mathcal{Q}\mathcal{M}_{k}^{!,\,p}, 𝒬kp\mathcal{Q}\mathcal{M}_{k}^{p}) for the set of meromorphic (resp. weakly holomorphic, holomorphic) quasi-modular forms of weight kk and depth pp and 𝒬kmero,p\mathcal{Q}\mathcal{M}_{k}^{\text{mero},\,\leq p} (resp. 𝒬k!,p\mathcal{Q}\mathcal{M}_{k}^{!,\,\leq p}, 𝒬kp\mathcal{Q}\mathcal{M}_{k}^{\leq p}) the \mathbb{C}-vector space of meromorphic (resp. weakly holomorphic, holomorphic) quasi-modular forms of weight kk and at most depth pp. Analogously, write

𝒬=p,k𝒬k,p,{mero,!,},\mathcal{Q}\mathcal{M}^{*}=\bigcup_{p\in\mathbb{N},k\in\mathbb{Z}}\mathcal{Q}\mathcal{M}_{k}^{*,\,p},\quad*\in\{\text{\text{mero}},\,!\,,-\},

for the \mathbb{C}-algebra of meromorphic (resp. weakly holomorphic, holomorphic) quasi-modular forms.

Similar to holomorphic quasi-modular forms, each component function fjf_{j} of a meromorphic quasi-modular form ff is again a meromorphic quasi-modular form.

Lemma 2.2.

Let f𝒬kmero,pf\in\mathcal{Q}\mathcal{M}_{k}^{\text{mero},\,p} be a meromorphic quasi-modular form of weight kk and depth pp with component functions f0,,fpf_{0},\cdots,f_{p}. Then for any 0jp0\leq j\leq p, the function fjf_{j} is meromorphic quasi-modular form of weight k2jk-2j and depth pjp-j. More precisely, for any γSL2()\gamma\in\operatorname{SL}_{2}(\mathbb{Z}),

(fj|k2jγ)(τ)=r=0pj(j+rr)fj+r(τ)(ccτ+d)r.(f_{j}|_{k-2j}\gamma)(\tau)=\sum_{r=0}^{p-j}\binom{j+r}{r}f_{j+r}(\tau)\left(\frac{c}{c\tau+d}\right)^{r}.

In particular, we have that f0=ff_{0}=f and fpf_{p} is indeed a meromorphic modular form of weight k2pk-2p.

Proof.

The details of the proof is omitted since it is completely identical with the proof for the holomorphic case given by [24] (see also [5, Thm 5.1.22]). ∎

Remark.

Since there are no holomorphic quasi-modular forms of negative weights, we know a holomorphic quasi-modular form always has weight k0k\geq 0 and depth pk/2p\leq k/2. However, a meromorphic quasi-modular form can has arbitrary weight kk and depth pp. 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 kk and p0p\geq 0 be integers. Let ff be a meromorphic function on the upper-half plane. For later use, we recall that the Maass–Shimura derivative of ff is

δkf=DfkYf\delta_{k}f=Df-k\,Y\!f

where D=12πiddτD=\frac{1}{2\pi i}\frac{d}{d\tau} and Y=14πyY=-\frac{1}{4\pi y}. We recall also the Serre derivative of ff

ϑk,pf=Dfkp12E2f.\vartheta_{k,p}f=Df-\frac{k-p}{12}\,E_{2}f.

We will occasionally abbreviate by δ\delta and ϑ\vartheta in an abuse of notations. We will often use the following identities found by Ramanujan.

DE2=112(E22E4),DE4=13(E2E4E6),DE6=12(E2E6E42).DE_{2}=\frac{1}{12}(E_{2}^{2}-E_{4}),\quad DE_{4}=\frac{1}{3}(E_{2}E_{4}-E_{6}),\quad DE_{6}=\frac{1}{2}(E_{2}E_{6}-E_{4}^{2}).
Lemma 3.1.

Let kk be an integer and ff be a meromorphic function on the upper-half plane. For any γSL2()\gamma\in\operatorname{SL}_{2}(\mathbb{R}), we have

(Df)|k+2γ\displaystyle(Df)|_{k+2}\gamma =D(f|kγ)+k2πiccτ+df|kγ\displaystyle=D(f|_{k}\gamma)+\frac{k}{2\pi i}\frac{c}{c\tau+d}f|_{k}\gamma
(δkf)|k+2γ\displaystyle(\delta_{k}f)|_{k+2}\gamma =δk(f|kγ).\displaystyle=\delta_{k}(f|_{k}\gamma).

Moreover, we have the following explicit formula for nn times Maass–Shimura derivative.

(3) δknf=j=0n(nj)(k+j)njYnjDjf,\delta_{k}^{n}f=\sum_{j=0}^{n}\binom{n}{j}(k+j)_{n-j}Y^{n-j}D^{j}f,

where (a)n=a(a+1)(a+n1)(a)_{n}=a(a+1)\cdots(a+n-1) and for n>0n>0, we set δkn=δk+2n2δk+2n4δk\delta_{k}^{n}=\delta_{k+2n-2}\circ\delta_{k+2n-4}\circ\dots\circ\delta_{k} and δk0\delta_{k}^{0} 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 nn. When n=1n=1, the identity (3)\eqref{eq:deltan} is just the definition. Applying Maass–Shimura derivatives on (3)\eqref{eq:deltan}, we find that δn+1f\delta^{n+1}f equals

D(δnf)\displaystyle D(\delta^{n}f) +(k+2n2)Yδnf\displaystyle+(k+2n-2)Y\delta^{n}f
=\displaystyle= j=0n+1((nj1)(k+j1)nj+1Yn+1jDjf\displaystyle\sum_{j=0}^{n+1}\bigg{(}\binom{n}{j-1}(k+j-1)_{n-j+1}Y^{n+1-j}D^{j}f
+(nj)(k+j)nj(n+k+j)Yn+1jDjf)\displaystyle+\binom{n}{j}(k+j)_{n-j}(n+k+j)Y^{n+1-j}D^{j}f\bigg{)}
=\displaystyle= j=0n+1(n+1j)(k+j)n+1jYn+1jDjf.\displaystyle\sum_{j=0}^{n+1}\binom{n+1}{j}(k+j)_{n+1-j}Y^{n+1-j}D^{j}f.\qed

In particular, when k0k\leq 0 and n=1kn=1-k the coefficients in (3)\eqref{eq:deltan} vanish for j0j\neq 0, in that case we obtain the Bol’s identity (cf. [15])

(4) δk1kf=D1kffor any fkmero.\delta_{k}^{1-k}f=D^{1-k}f\quad\text{for any }f\in\mathcal{M}_{k}^{\text{mero}}.

This indicates that D1kD^{1-k} is in fact an SL2()\operatorname{SL}_{2}(\mathbb{Z})-invariant differential operator on kmero\mathcal{M}_{k}^{\text{mero}}. 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 f𝒬kmero,pf\in\mathcal{Q}\mathcal{M}_{k}^{\text{mero},\,p} with the component functions f0,f1,,fpf_{0},f_{1},\cdots,f_{p}, we have

(Df)|kγ(τ)=r=0p+1(D(fr)+kr+12πifr1)(ccτ+d)r,(Df)|_{k}\gamma(\tau)=\sum_{r=0}^{p+1}\left(D(f_{r})+\frac{k-r+1}{2\pi i}f_{r-1}\right)\left(\frac{c}{c\tau+d}\right)^{r},

where by convention f1=fp+1=0f_{-1}=f_{p+1}=0.

Proof.

Since D(ccτ+d)=12πi(ccτ+d)2D(\frac{c}{c\tau+d})=-\frac{1}{2\pi i}(\frac{c}{c\tau+d})^{2}, by applying the differential operator DD to both sides of (2)\eqref{eq:slashf}, we get

D(f|kγ)=r=0pD(fr)(ccτ+d)rr2πifr(τ)(ccτ+d)r+1.D(f|_{k}\gamma)=\sum_{r=0}^{p}D(f_{r})\left(\frac{c}{c\tau+d}\right)^{r}-\frac{r}{2\pi i}f_{r}(\tau)\left(\frac{c}{c\tau+d}\right)^{r+1}.

We complete the proof by applying Lemma 3.1. ∎

The following lemma implies that the differential operator DD usually increases the depth of a meromorphic quasi-modular form by 11 (note that in holomorphic case it always does).

Proposition 3.3.

The space of meromorphic quasi-modular forms 𝒬mero\mathcal{Q}\mathcal{M}^{\text{mero}} is stable under the derivation DD. It acts on 𝒬mero\mathcal{Q}\mathcal{M}^{\text{mero}} by increasing the weight by 22 and increasing the depth by at most 11

D:𝒬kmero,p𝒬k+2mero,p+1.D:\mathcal{Q}\mathcal{M}_{k}^{\text{mero},\,\leq p}\rightarrow\mathcal{Q}\mathcal{M}_{k+2}^{\text{mero},\,\leq p+1}.

More precisely, for a meromorphic modular form f𝒬kmero,pf\in\mathcal{Q}\mathcal{M}_{k}^{\text{mero},\,p}, we have Df𝒬k+2mero,p+1Df\in\mathcal{Q}\mathcal{M}_{k+2}^{\text{mero},\,p+1} if kpk\neq p and Df𝒬k+2mero,pDf\in\mathcal{Q}\mathcal{M}_{k+2}^{\text{mero},\,\leq p} if k=pk=p.

Proof.

This follows immediately from Lemma 3.2. Note that one has

Qp+1(Df)=kp2πiQp(f).Q_{p+1}(Df)=\frac{k-p}{2\pi i}Q_{p}(f).

So DfDf will have exactly depth p+1p+1 unless k=pk=p. ∎

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 𝒬mero\mathcal{Q}\mathcal{M}^{\text{mero}} is stable under the Serre derivative ϑ\vartheta. It acts on 𝒬mero\mathcal{Q}\mathcal{M}^{\text{mero}} by increasing the weight by 22,

ϑk,p:𝒬kmero,p𝒬k+2mero,p.\vartheta_{k,p}:\mathcal{Q}\mathcal{M}_{k}^{\text{mero},\,\leq p}\rightarrow\mathcal{Q}\mathcal{M}_{k+2}^{\text{mero},\,\leq p}.

In particular, if ff is a meromorphic modular form of weight kk, then ϑk,pf\vartheta_{k,p}f is a meromorphic modular form of weight k+2k+2.

Proof.

Using the modular transformation of E2E_{2} in equation (1) and modular transformation of DfDf in Lemma 3.2, we get

Qp+1(Df)=kp2πiQp(f),Qp+1(E2f)=6πiQp(f).Q_{p+1}(Df)=\frac{k-p}{2\pi i}Q_{p}(f),\quad Q_{p+1}(E_{2}f)=\frac{6}{\pi i}Q_{p}(f).

Hence we always have Qp+1(ϑk,pf)=0Q_{p+1}(\vartheta_{k,p}f)=0. Thus ϑk,pf\vartheta_{k,p}f has at most depth pp. ∎

Lemma 3.5.

Let fkmerof\in\mathcal{M}_{k}^{\text{mero}} be a meromorphic modular form of weight kk. Then

  1. (i)

    If k>0k>0, we have Dpf𝒬k+2pmero,pD^{p}f\in\mathcal{Q}\mathcal{M}_{k+2p}^{\text{mero},\,p},

  2. (ii)

    If k0k\leq 0 and pkp\leq-k, then we have Dpf𝒬k+2pmero,pD^{p}f\in\mathcal{Q}\mathcal{M}_{k+2p}^{\text{mero},\,p},

  3. (iii)

    If k0k\leq 0 and p1kp\geq 1-k, then we have Dpf𝒬k+2pmero,p+k1D^{p}f\in\mathcal{Q}\mathcal{M}_{k+2p}^{\text{mero},\,p+k-1}.

Proof.

Applying iteratively Lemma 3.2, we obtain

(5) Qp(Dpf)=p!(2πi)p(k+p1p)f,Q_{p}(D^{p}f)=\frac{p!}{(2\pi i)^{p}}\binom{k+p-1}{p}\,f,

which is always nonvanishing when k>0k>0 or p+k0p+k\leq 0. This proves the assertions (i)(i) and (ii)(ii).

For the third one, keep in mind that D1kfD^{1-k}f is a meromorphic modular form of weight 2k2-k by Bol’s identity. Therefore by the assertion (i)(i), we have

Dpf=Dp+k1(D1kf)𝒬k+2pmero,p+k1.D^{p}f=D^{p+k-1}(D^{1-k}f)\in\mathcal{Q}\mathcal{M}_{k+2p}^{\text{mero},\,p+k-1}.\qed
Lemma 3.6.

Let pp be a nonnegative integer. Then the sequence

0δp=0[Δ±]𝒬mero,pϑ𝒬mero,p0\longrightarrow\delta_{p=0}\,\mathbb{C}[\Delta^{\pm}]\longrightarrow\mathcal{Q}\mathcal{M}^{\text{mero},\,\leq p}\stackrel{{\scriptstyle\vartheta}}{{\longrightarrow}}\mathcal{Q}\mathcal{M}^{\text{mero},\,\leq p}

of graded \mathbb{C}-vector space is left exact where δp=0=1\delta_{p=0}=1 when p=0p=0 and equals to 0 when p0p\neq 0.

Proof.

Let f𝒬kmero,pf\in\mathcal{Q}\mathcal{M}_{k}^{\text{mero},\,p} be a meromorphic quasi-modular form in the kernel of the Serre derivative ϑ\vartheta. As E2=DΔ/ΔE_{2}=D\Delta/\Delta, we have the identity of logarithmic derivative 12Df/f=(kp)DΔ/Δ12Df/f=(k-p)D\Delta/\Delta. Therefore ff is a power of Δ\Delta with k12k\in 12\mathbb{Z} and p=0p=0. ∎

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 {mero,!}*\in\{\text{mero},!\}. The \mathbb{C}-algebra of meromorphic quasi-modular forms is graded by the weight kk and filtered by the depth pp

𝒬=kp𝒬k,p.\mathcal{Q}\mathcal{M}^{*}=\bigoplus_{k\in\mathbb{Z}}\bigcup_{p\in\mathbb{N}}\mathcal{Q}\mathcal{M}_{k}^{*,\,\leq p}.
Lemma 4.1.

Let kk and pp be integers with p0p\geq 0. Then we have the following split exact sequence

0𝒬k,p1𝒬k,pQpk2p0.0\longrightarrow\mathcal{Q}\mathcal{M}_{k}^{*,\,\leq p-1}\longrightarrow\mathcal{Q}\mathcal{M}_{k}^{*,\,\leq p}\stackrel{{\scriptstyle Q_{p}}}{{\longrightarrow}}\mathcal{M}_{k-2p}^{*}\longrightarrow 0.
Proof.

Let f𝒬k,pf\in\mathcal{Q}\mathcal{M}_{k}^{*,\,\leq p} be a meromorphic quasi-modular form. We recall that by Lemma 2.2, the last component Qp(f)k2pQ_{p}(f)\in\mathcal{M}_{k-2p}^{*} is in fact a meromorphic modular form of weight k2pk-2p. Thus the above sequence is exact. From the modular transformation (1) of E2E_{2}, we know that Qp(gE2p)=(6/πi)pgQ_{p}(gE_{2}^{p})=\left(6/\pi i\right)^{p}g for any gk2pg\in\mathcal{M}_{k-2p}^{*}. Thus the map g(2πi12)pgE2pg\mapsto(\frac{2\pi i}{12})^{p}gE_{2}^{p} is a section of the map QpQ_{p}, so the above exact sequence is also split. ∎

Theorem 4.2.

The graded \mathbb{C}-algebra of meromorphic quasi-modular forms is generated by meromorphic modular forms and E2E_{2}

𝒬=[E2]=p0E2p,{mero,!},\mathcal{Q}\mathcal{M}^{*}=\mathcal{M}^{*}[E_{2}]=\bigoplus_{p\geq 0}\mathcal{M}^{*}E_{2}^{p},\quad*\in\{\text{mero},\,!\},

where the depth of a meromorphic quasi-modular forms is exactly the degree of E2E_{2} within it.

Proof.

Induction on the depth of f𝒬f\in\mathcal{Q}\mathcal{M}^{*}. The statement is straightforward when the depth is 0. As explained in Lemma 4.1, we see that the form f(2πi12)pQp(f)E2pf-(\frac{2\pi i}{12})^{p}Q_{p}(f)E_{2}^{p} has depth p1\leq p-1. Therefore by induction, for any f𝒬k,pf\in\mathcal{Q}\mathcal{M}_{k}^{*,\,p} there exist meromorphic modular forms gig_{i} of weight k2ik-2i such that f=i=0pgiE2if=\sum_{i=0}^{p}g_{i}E_{2}^{i}. ∎

Let grp𝒬\operatorname{gr}_{p}\mathcal{Q}\mathcal{M}^{*} be the associated graded \mathbb{C}-algebra with respect to the depth pp. Then grp𝒬\operatorname{gr}_{p}\mathcal{Q}\mathcal{M}^{*} is a bigraded ring

grp𝒬kpk2pE2p.\operatorname{gr}_{p}\mathcal{Q}\mathcal{M}^{*}\simeq\bigoplus_{k\in\mathbb{Z}}\bigoplus_{p\in\mathbb{N}}\mathcal{M}_{k-2p}^{*}E_{2}^{p}.

Then the induced derivative grpD\operatorname{gr}_{p}D on grp𝒬\operatorname{gr}_{p}\mathcal{Q}\mathcal{M}^{*} is homogeneous, increasing the weight by 22 and depth by 11.

Proposition 4.3.

We have the following left exact sequence of bigraded \mathbb{C}-vector spaces

0ppE2pgrp𝒬grpDgrp𝒬.0\longrightarrow\bigoplus_{p\in\mathbb{N}}\mathcal{M}_{-p}^{*}\,E_{2}^{p}\longrightarrow\operatorname{gr}_{p}\mathcal{Q}\mathcal{M}^{*}\xrightarrow{\operatorname{gr}_{p}D}\operatorname{gr}_{p}\mathcal{Q}\mathcal{M}^{*}.

In fact, the induced derivative grpD\operatorname{gr}_{p}D

grpD:𝒬k,p/𝒬k,p1𝒬k+2,p+1/𝒬k+2,p,\operatorname{gr}_{p}D:\mathcal{Q}\mathcal{M}_{k}^{*,\,\leq p}/\mathcal{Q}\mathcal{M}_{k}^{*,\,\leq p-1}\to\mathcal{Q}\mathcal{M}_{k+2}^{*,\,\leq p+1}/\mathcal{Q}\mathcal{M}_{k+2}^{*,\,\leq p},

is a bijection when pkp\neq k and is a zero map if p=kp=k.

Proof.

It follows from Lemma 3.5 that the map grpD\operatorname{gr}_{p}D is injective when pkp\neq k and is a zero map when p=kp=k. We next show the surjectivity of grpD\operatorname{gr}_{p}D for pkp\neq k. Using the fundamental identity DE2=112(E22E4)DE_{2}=\frac{1}{12}(E_{2}^{2}-E_{4}), we observe that for a meromorphic modular form gg of weight k2pk-2p,

(6) D(gE2p)=DgE2p+p12g(E2p+1E2p1E4)=ϑgE2p+kp12gE2p+1p12gE4E2p1.\begin{split}&D(gE_{2}^{p})=Dg\,E_{2}^{p}+\frac{p}{12}g(E_{2}^{p+1}-E_{2}^{p-1}E_{4})\\ =&\vartheta g\,E_{2}^{p}+\frac{k-p}{12}gE_{2}^{p+1}-\frac{p}{12}gE_{4}E_{2}^{p-1}.\end{split}

Here we recall that ϑg=Dgk2p12E2g\vartheta g=Dg-\frac{k-2p}{12}E_{2}g is the Serre derivative of gg, which is modular of weight k2p+2k-2p+2. Hence

D(gE2p)kp12gE2p+1+𝒬k+2mero,p.D(gE_{2}^{p})\in\frac{k-p}{12}gE_{2}^{p+1}+\mathcal{Q}\mathcal{M}_{k+2}^{\text{mero},\,\leq p}.

Since kpk\neq p, the map grpD\operatorname{gr}_{p}D is surjective. ∎

Likewise, we have an induced homogeneous Serre derivative grpϑ\operatorname{gr}_{p}\vartheta on grp𝒬\operatorname{gr}_{p}\mathcal{Q}\mathcal{M}^{*}, which increases the weight by 22 and preserves the depth.

Proposition 4.4.

We have the following left exact sequence of bigraded \mathbb{C}-vector spaces

0[Δ±,E2]grp𝒬grpϑgrp𝒬.0\longrightarrow\mathbb{C}[\Delta^{\pm},\,E_{2}]\longrightarrow\operatorname{gr}_{p}\mathcal{Q}\mathcal{M}^{*}\xrightarrow{\operatorname{gr}_{p}\vartheta}\operatorname{gr}_{p}\mathcal{Q}\mathcal{M}^{*}.
Proof.

A direct computation

ϑ(gE2p)\displaystyle\vartheta(gE_{2}^{p}) =pϑ(E2)E2p1g+ϑ(g)E2p\displaystyle=p\vartheta(E_{2})\,E_{2}^{p-1}g+\vartheta(g)E_{2}^{p}
=p12gE4E2p1+ϑgE2p\displaystyle=-\frac{p}{12}gE_{4}E_{2}^{p-1}+\vartheta g\,E_{2}^{p}

yields that ϑ(gE2p)ϑgE2p+𝒬k+2mero,p1\vartheta(gE_{2}^{p})\in\vartheta g\,E_{2}^{p}+\mathcal{Q}\mathcal{M}_{k+2}^{\text{mero},\,\leq p-1}. Therefore the kernel of grpϑ\operatorname{gr}_{p}\vartheta is the graded subalgebra generated by powers of Δ\Delta and E2E_{2}. ∎

It is well-known that a holomorphic quasi-modular form is always a linear combination of iterated derivatives of holomorphic modular forms and E2E_{2}. 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 f𝒬kmero,pf\in\mathcal{Q}\mathcal{M}_{k}^{\text{mero},\,p} be a meromorphic quasi-modular form with depth k/2p<kk/2\leq p<k. Then ff is never a linear combination of iterated derivatives of meromorphic modular forms.

Proof.

Suppose that ff admits a decomposition f=l=0l0DlFk2lf=\sum_{l=0}^{l_{0}}D^{l}F_{k-2l}, where Fk2lF_{k-2l} is a meromorphic modular form of weight k2lk-2l. Lemma 3.5 shows that when l0<k/2l_{0}<k/2 or l0kl_{0}\geq k, the depth of ff is exactly l0l_{0}. So we are left to consider only k/2l0<kk/2\leq l_{0}<k. If k/2ll0k/2\leq l\leq l_{0}, the depth of DlFk2lD^{l}F_{k-2l} is kl1k-l-1, which is strictly smaller than k/2k/2, and if l<k/2l<k/2 the depth of DlFk2lD^{l}F_{k-2l} is ll, also strictly smaller than k/2k/2. This implies that the depth of ff is always <k/2<k/2, 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 lk/21l\leq k/2-1 and lkl\geq k respectively. In the middle part, applying repeatedly Proposition 3.3 we find that each component has depth kl1k-l-1 for 0lk/210\leq l\leq k/2-1. So the above sum runs through all depths and must be a direct sum of \mathbb{C}-vector spaces. It remains to show every meromorphic quasi-modular form f𝒬k,pf\in\mathcal{Q}\mathcal{M}_{k}^{*,\,p} has such decomposition. We divided the proof into three parts.

Part 1. pk/21p\leq k/2-1

\@afterheading

When p=0p=0, the result is direct. For 0<pk/210<p\leq k/2-1, on account of the computation in (5) (or using Proposition 4.3), the pp-th component function of

f(2πi)pp!(kp1p)DpQp(f)f-\frac{(2\pi i)^{p}}{p!\binom{k-p-1}{p}}D^{p}Q_{p}(f)

is zero, so we complete the proof by induction on pp.

Part 2. pkp\geq k

\@afterheading

In this case, since (kp1p)0\binom{k-p-1}{p}\neq 0, the induction argument still works unless the depth goes to less than kk. This implies that we can find Fk2lk2lF_{k-2l}\in\mathcal{M}_{k-2l}^{*} where l=k,,pl=k,\dots,p and F𝒬k,k1F\in\mathcal{Q}\mathcal{M}_{k}^{*,\,\leq k-1} such that

f=DpFk2p++DkFk+F.f=D^{p}F_{k-2p}+\cdots+D^{k}F_{-k}+F.

So we reduce the case pkp\geq k to the case p<kp<k.

Part 3. k/2p<kk/2\leq p<k

\@afterheading

We claim that for any such ff there exists a meromorphic quasi-modular form hl=0k21Dl𝒬k2l,k2l1h\in\bigoplus_{l=0}^{\frac{k}{2}-1}D^{l}\mathcal{Q}\mathcal{M}_{k-2l}^{*,\,k-2l-1} so that fh𝒬k,k/21f-h\in\mathcal{Q}\mathcal{M}_{k}^{*,\,\leq k/2-1}. Then the proof will be converted to the first part. We will prove the claim by induction on k+pk+p.

Starting from k+p=3k+p=3, the only possible case is k=2k=2, p=1p=1. In this case f𝒬2, 1f\in\mathcal{Q}\mathcal{M}_{2}^{*,\,1}, the result is direct since it corresponds to l=0l=0. We assume that k+p>3k+p>3. If p=k1p=k-1, then the result is also direct since f𝒬k,k1f\in\mathcal{Q}\mathcal{M}_{k}^{*,\,k-1}.

Then we can assume that p<k1p<k-1. By Theorem 4.2, there exists a meromorphic modular form Fk2pF_{k-2p} of weight k2pk-2p such that f=g+Fk2pE2pf=g+F_{k-2p}\,E_{2}^{p} for some g𝒬k,p1g\in\mathcal{Q}\mathcal{M}_{k}^{*,\,\leq p-1}. The same calculation as in equation (6) yields that Fk2pE2pF_{k-2p}\,E_{2}^{p} can be represented as the linear combination

kp112Fk2pE2p=D(Fk2pE2p1)ϑ(Fk2p)E2p1+p112Fk2pE4E2p2.\frac{k-p-1}{12}F_{k-2p}\,E_{2}^{p}=D(F_{k-2p}\,E_{2}^{p-1})-\vartheta(F_{k-2p})E_{2}^{p-1}+\frac{p-1}{12}F_{k-2p}\,E_{4}E_{2}^{p-2}.

On the right-hand side, the first term comes from the derivative of Fk2pE2p1F_{k-2p}\,E_{2}^{p-1}, which has exactly weight k2k-2 and depth p1p-1. So by induction assumption, we can find a weight k2k-2 meromorphic quasi-modular form Gk2G_{k-2} such that

Fk2pE2p1Gk2𝒬k2,k221with Gk2l=0k221Dl𝒬k2l2,k2l3.F_{k-2p}\,E_{2}^{p-1}-G_{k-2}\in\mathcal{Q}\mathcal{M}_{k-2}^{*,\,\leq\frac{k-2}{2}-1}\hskip 5.0pt\text{with }\,G_{k-2}\in\!\!\bigoplus_{l=0}^{\frac{k-2}{2}-1}D^{l}\mathcal{Q}\mathcal{M}_{k-2l-2}^{*,\,k-2l-3}.

Then by applying the operator DD, we get

D(Fk2pE2p1)D(Gk2)𝒬k,k21,D(F_{k-2p}\,E_{2}^{p-1})-D(G_{k-2})\in\mathcal{Q}\mathcal{M}_{k}^{*,\,\leq\frac{k}{2}-1},

where

D(Gk2)l=1k21Dl𝒬k2l,k2l1.D(G_{k-2})\in\bigoplus_{l=1}^{\frac{k}{2}-1}D^{l}\mathcal{Q}\mathcal{M}_{k-2l}^{*,\,k-2l-1}.

Besides, the second term ϑ(Fk2p)E2p1\vartheta(F_{k-2p})\,E_{2}^{p-1} and the last term Fk2pE4E2p2F_{k-2p}\,E_{4}E_{2}^{p-2} on the right-hand side and and the function gg all have weight kk and depth p1\leq p-1. So we find that

f12kp1D(Gk2)𝒬k,p1.f-\frac{12}{k-p-1}D(G_{k-2})\in\mathcal{Q}\mathcal{M}_{k}^{*,\,\leq p-1}.

By induction assumption, we know there exists Hl=0k21Dl𝒬k2l,k2l1H\in\bigoplus_{l=0}^{\frac{k}{2}-1}D^{l}\mathcal{Q}\mathcal{M}_{k-2l}^{*,\,k-2l-1} so that

f12kp1D(Gk2)H𝒬k,k21,f-\frac{12}{k-p-1}D(G_{k-2})-H\in\mathcal{Q}\mathcal{M}_{k}^{*,\,\leq\frac{k}{2}-1},

which proves the previous claim. ∎

Remarks.
  1. (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 ll and depth l1l-1.

  2. (ii)

    When k0k\leq 0, 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.

  3. (iii)

    For holomorphic quasi-modular form this reduces to the well-known (see Zagier [24, Prop. 20])

    𝒬k=(l=0k22Dlk2l)Dk21𝒬21,\mathcal{Q}\mathcal{M}_{k}=\bigg{(}\bigoplus_{l=0}^{\frac{k}{2}-2}D^{l}\mathcal{M}_{k-2l}\bigg{)}\bigoplus D^{\frac{k}{2}-1}\mathcal{Q}\mathcal{M}_{2}^{1},

    where the depth p=k/2p=k/2 comes from the iterated derivatives of the Eisenstein series E2E_{2} which generates 𝒬21\mathcal{Q}\mathcal{M}_{2}^{1}.

  4. (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 l>0l>0 of (quasi-)modular forms with Fourier expansion in [[q]]\mathbb{Q}\otimes_{\mathbb{Z}}\mathbb{Z}[[q]].

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 kk is a positive integer, our series is the same as Cohen and Kuznetsov. When kk is a negative integer, we will define a minus series and a plus series.

Definition 5.1.

Let ff be a meromorphic modular form of weight kk\in\mathbb{Z}. We define its Cohen–Kuznetsov series by

CKD(f;τ,T)\displaystyle CK_{D}^{-}(f;\tau,T) =n=0k(1)n+k(kn)!n!DnfTn,\displaystyle=\sum_{n=0}^{-k}(-1)^{n+k}\frac{(-k-n)!}{n!}D^{n}f\,T^{n},
CKD+(f;τ,T)\displaystyle CK_{D}^{+}(f;\tau,T) =n1kDnfn!(n+k1)!Tn.\displaystyle=\sum_{n\geq 1-k}\frac{D^{n}f}{n!(n+k-1)!}T^{n}.

Similarly, we can define CKδ±(f;τ,T)CK_{\delta}^{\pm}(f;\tau,T) by replacing DD by δ\delta. When k>0k>0, by convention, CK=0CK^{-}=0, and CK+CK^{+} starts from the term n=0n=0. Moreover, we define the slash operator on CK±CK^{\pm} by

(CK±|kγ)(f;τ,T)=(cτ+d)kCK±(f;aτ+bcτ+d,T(cτ+d)2)(CK^{\pm}|_{k}\gamma)(f;\tau,T)=(c\tau+d)^{-k}CK^{\pm}\left(f;\frac{a\tau+b}{c\tau+d},\frac{T}{(c\tau+d)^{2}}\right)

where γSL2()\gamma\in\operatorname{SL}_{2}(\mathbb{R}).

We first give some basic properties of Cohen–Kuznetsov series. When ff has positive weight, similar results can also be found in [5].

Proposition 5.2.

Let ff be a meromorphic function of weight kk. Suppose kk is a nonpositive integer. Then

  1. (i)

    We have

    CK+(f;τ,T)=T1kCK+(D1kf;τ,T),CK_{*}^{+}(f;\tau,T)=T^{1-k}CK_{*}^{+}(D^{1-k}f;\tau,T),

    where * denotes the operator DD or δ\delta.

  2. (ii)

    The functions CKδ+CK_{\delta}^{+} and CKD+CK_{D}^{+} are linked by

    CKδ+(f;τ,T)=eTYCKD+(f;τ,T).CK_{\delta}^{+}(f;\tau,T)=e^{TY}CK_{D}^{+}(f;\tau,T).
  3. (iii)

    The series CKδ+CK_{\delta}^{+} and CKD+CK_{D}^{+} commutes with the slash operator up to a factor. Namely, for any γSL2()\gamma\in\operatorname{SL}_{2}(\mathbb{R}), we have

    (CKδ+)|kγ(f;τ,T)\displaystyle(CK_{\delta}^{+})|_{k}\gamma(f;\tau,T) =CKδ+(f|kγ;τ,T),\displaystyle=CK_{\delta}^{+}(f|_{k}\gamma;\tau,T),
    (CKD+)|kγ(f;τ,T)\displaystyle(CK_{D}^{+})|_{k}\gamma(f;\tau,T) =eT2πiccτ+dCKD+(f|kγ;τ,T).\displaystyle=e^{\frac{T}{2\pi i}\frac{c}{c\tau+d}}CK_{D}^{+}(f|_{k}\gamma;\tau,T).
Proof.

(i)(i). Since D1kfD^{1-k}f is a modular form of weight 2k>02-k>0, one has

T1kCKD+(D1kf;τ,T)=n0Dn(D1kf)n!(n+1k)!Tn+1k=CKD+(f;τ,T).T^{1-k}CK_{D}^{+}(D^{1-k}f;\tau,T)=\sum_{n\geq 0}\frac{D^{n}(D^{1-k}f)}{n!(n+1-k)!}T^{n+1-k}=CK_{D}^{+}(f;\tau,T).

The proof for CKδ+CK_{\delta}^{+} is similar.

(ii)(ii). We note that if ff is a meromorphic modular form of positive weight kk, then by Lemma 3.1, we have

CKδ+(f;τ,T)=\displaystyle CK_{\delta}^{+}(f;\tau,T)= n0j=0n(nj)(k+j)njYnjDjfn!(n+k1)!Tn\displaystyle\sum_{n\geq 0}\sum_{j=0}^{n}\frac{\binom{n}{j}(k+j)_{n-j}Y^{n-j}D^{j}f}{n!(n+k-1)!}T^{n}
=\displaystyle= j,l0YlDjf(k+j1)!l!j!Tl+j\displaystyle\sum_{j,\,l\geq 0}\frac{Y^{l}D^{j}f}{(k+j-1)!l!j!}T^{l+j}
=\displaystyle= eYTCKD+(f;τ,T).\displaystyle e^{YT}CK_{D}^{+}(f;\tau,T).

So when the weight of ff is nonpositive, we can apply the result above to D1kfD^{1-k}f and get

CKδ+\displaystyle CK_{\delta}^{+} (f;τ,T)=T1kCKδ+(D1kf;τ,T)\displaystyle(f;\tau,T)=T^{1-k}CK_{\delta}^{+}(D^{1-k}f;\tau,T)
=T1keTYCKδ+(D1kf;τ,T)=eTYCKD+(f;τ,T).\displaystyle=T^{1-k}e^{TY}CK_{\delta}^{+}(D^{1-k}f;\tau,T)=e^{TY}CK_{D}^{+}(f;\tau,T).

(iii)(iii). Similar to the proof above, we start from the positive weight case. By Lemma 3.1, we have

(cτ+d)2n(δnf)(aτ+bcτ+d)=δn(f|kγ).(c\tau+d)^{-2n}(\delta^{n}f)\left(\frac{a\tau+b}{c\tau+d}\right)=\delta^{n}(f|_{k}\gamma).

Then

(CKδ+)|kγ(f;τ,T)=n0(δnf)(aτ+bcτ+d)n!(n+k1)!Tn(cτ+d)2n+k\displaystyle(CK_{\delta}^{+})|_{k}\gamma(f;\tau,T)=\sum_{n\geq 0}\frac{(\delta^{n}f)(\frac{a\tau+b}{c\tau+d})}{n!(n+k-1)!}\frac{T^{n}}{(c\tau+d)^{2n+k}}
=\displaystyle= n0δnf(τ)n!(n+k1)!Tn(cτ+d)k=CKδ+(f|kγ;τ,T).\displaystyle\sum_{n\geq 0}\frac{\delta^{n}f(\tau)}{n!(n+k-1)!}\frac{T^{n}}{(c\tau+d)^{k}}=CK_{\delta}^{+}(f|_{k}\gamma;\tau,T).

Finally, we can complete the proof by applying (i)(i)

(CKδ+)\displaystyle(CK_{\delta}^{+}) |kγ(f;τ,T)=(cτ+d)kCKδ+(f;aτ+bcτ+d,T(cτ+d)2)\displaystyle|_{k}\gamma(f;\tau,T)=(c\tau+d)^{-k}CK_{\delta}^{+}\left(f;\frac{a\tau+b}{c\tau+d},\frac{T}{(c\tau+d)^{2}}\right)
=\displaystyle= (cτ+d)k2CKδ+(D1kf;aτ+bcτ+d,T(cτ+d)2)T1k\displaystyle(c\tau+d)^{k-2}CK_{\delta}^{+}\left(D^{1-k}f;\frac{a\tau+b}{c\tau+d},\frac{T}{(c\tau+d)^{2}}\right)T^{1-k}
=\displaystyle= T1kCKδ+((D1kf)|2kγ;τ,T)=CKδ+(f|kγ;τ,T).\displaystyle T^{1-k}CK_{\delta}^{+}((D^{1-k}f)|_{2-k}\gamma;\tau,T)=CK_{\delta}^{+}(f|_{k}\gamma;\tau,T).

As for the series CKD+CK_{D}^{+}, we just need to apply (ii)(ii). ∎

Then we focus on the minus Cohen–Kuznetsov series.

Proposition 5.3.

Let ff be a meromorphic modular form of weight kk. Suppose kk is a nonpositive integer. Then for the minus part of Cohen–Kuznetsov series, we have the following relations

  1. (i)

    The functions CKδCK_{\delta}^{-} and CKDCK_{D}^{-} are linked by

    CKδ(f;τ,T)=eTYCKD(f;τ,T)+O(T1k).CK_{\delta}^{-}(f;\tau,T)=e^{TY}CK_{D}^{-}(f;\tau,T)+O(T^{1-k}).
  2. (ii)

    The function CKδCK_{\delta}^{-} also commutes with the slash operator, i.e. for any γSL2()\gamma\in\operatorname{SL}_{2}(\mathbb{R}), we have

    (CKδ)|kγ(f;τ,T)=CKδ(f|kγ;τ,T).(CK_{\delta}^{-})|_{k}\gamma(f;\tau,T)=CK_{\delta}^{-}(f|_{k}\gamma;\tau,T).
  3. (iii)

    The function CKDCK_{D}^{-} has the transformation

    (CKD)|kγ(f;τ,T)=eT2πiccτ+dCKD(f|kγ;τ,T)+O(T1k).(CK_{D}^{-})|_{k}\gamma(f;\tau,T)=e^{\frac{T}{2\pi i}\frac{c}{c\tau+d}}CK_{D}^{-}(f|_{k}\gamma;\tau,T)+O(T^{1-k}).
Proof.

(i)(i). The proof is similar to CKδ+CK_{\delta}^{+}. By definition, we have

CKδ(f;τ,T)=n=0kj=0n(nj)(k+j)nj(1)n+k(nk)!n!YnjDjfTn\displaystyle CK_{\delta}^{-}(f;\tau,T)=\sum_{n=0}^{-k}\sum_{j=0}^{n}\binom{n}{j}(k+j)_{n-j}\frac{(-1)^{n+k}(-n-k)!}{n!}Y^{n-j}D^{j}f\,T^{n}

By changing n=j+ln=j+l, we get

j=0k(1)kj(kj)!j!DjfTjl=0kjYll!Tl\displaystyle\sum_{j=0}^{-k}\frac{(-1)^{k-j}(-k-j)!}{j!}D^{j}f\,T^{j}\sum_{l=0}^{-k-j}\frac{Y^{l}}{l!}T^{l}
=\displaystyle= j=0k(1)kj(kj)!j!DjfTjl=0Yll!Tl+O(T1k)\displaystyle\sum_{j=0}^{-k}\frac{(-1)^{k-j}(-k-j)!}{j!}D^{j}f\,T^{j}\sum_{l=0}^{\infty}\frac{Y^{l}}{l!}T^{l}+O(T^{1-k})
=\displaystyle= eTYCKD(f;τ,T)+O(T1k).\displaystyle e^{TY}CK_{D}^{-}(f;\tau,T)+O(T^{1-k}).

The proof of (ii)(ii) and (iii)(iii) are similar to (iii)(iii) in Proposition 5.2, we omit the details here. ∎

Definition 5.4.

Let f,gf,g be two meromorphic modular forms of weight k,lk,l respectively. Then the nn-th Rankin–Cohen bracket of ff and gg is defined by

[f,g]nj=0n(1)j(n+k1j)(n+l1nj)DnjfDjg.[f,g]_{n}\coloneqq\sum_{j=0}^{n}(-1)^{j}\binom{n+k-1}{j}\binom{n+l-1}{n-j}D^{n-j}fD^{j}g.
Theorem 5.5.

Let f,gf,g be two meromorphic modular form of weight k,lk,l respectively. Then the nn-th Rankin–Cohen bracket [f,g]n[f,g]_{n} is a meromorphic modular form of weight k+l+2nk+l+2n.

Proof.

The case of k,l>0k,l>0 is a result of Cohen [5]. So we assume that at least one of k,lk,l is nonpositive. By symmetry [f,g]n=(1)n[g,f]n[f,g]_{n}=(-1)^{n}[g,f]_{n}, we may assume that klk\geq l. There are exactly two possibilities, either one of kk and ll is positive or none of kk and ll is positive. For these two parts, we further separate them into the several cases depending on nn.

Part 1. k>0k>0, l0l\leq 0

\@afterheading

Case 1. 0nl0\leq n\leq-l. We consider the product CKD+(f;τ,T)CKD(g;τ,T)CK_{D}^{+}(f;\tau,T)CK_{D}^{-}(g;\tau,-T), it is equal to

CKD+(f;τ,T)CKD(g;τ,T)=n=0lj=0n(1)jAjk,l,nDnjfDjg+O(T1l),CK_{D}^{+}(f;\tau,T)CK_{D}^{-}(g;\tau,-T)=\sum_{n=0}^{-l}\sum_{j=0}^{n}(-1)^{j}A_{j}^{k,l,n}D^{n-j}fD^{j}g+O(T^{1-l}),

where

Ajk,l,n\displaystyle A_{j}^{k,l,n} =(1)l+j(lj)!j!(nj)!(n+kj1)!\displaystyle=(-1)^{l+j}\frac{(-l-j)!}{j!(n-j)!(n+k-j-1)!}
=(1)n+l(ln)!(k+n1)!(k+n1j)(n+l1nj).\displaystyle=(-1)^{n+l}\frac{(-l-n)!}{(k+n-1)!}\binom{k+n-1}{j}\binom{n+l-1}{n-j}.

Here we use the identity (nj)=(1)j(n+j1j)\binom{-n}{j}=(-1)^{j}\binom{n+j-1}{j} for n0n\geq 0. Thus we have

CKD+(f;τ,T)CKD(g;τ,T)=n=0l(1)n+l(ln)!(k+n1)![f,g]nTn+O(T1l).CK_{D}^{+}(f;\tau,T)CK_{D}^{-}(g;\tau,-T)=\sum_{n=0}^{-l}\frac{(-1)^{n+l}(-l-n)!}{(k+n-1)!}[f,g]_{n}T^{n}+O(T^{1-l}).

On the other hand, by Proposition 5.2 and 5.3, for any γSL2()\gamma\in\operatorname{SL}_{2}(\mathbb{Z}), we have

CKD+(f;τ,T)CKD(g;τ,T)|k+lγ=CKD+(f;τ,T)CKD(g;τ,T)+O(T1l).CK_{D}^{+}(f;\tau,T)CK_{D}^{-}(g;\tau,-T)|_{k+l}\gamma=CK_{D}^{+}(f;\tau,T)CK_{D}^{-}(g;\tau,-T)+O(T^{1-l}).

This implies that [f,g]n[f,g]_{n} is a meromorphic modular form of weight k+l+2nk+l+2n.

Case 2. n1ln\geq 1-l. Then the terms where j<1lj<1-l in the Rankin–Cohen bracket vanish since the binomial coefficients become zero. Put n=n+l1n^{\prime}=n+l-1, the Rankin–Cohen bracket of ff and gg turns out to be

(7) [f,g]n=j=1ln(1)j(n+k1nj)(n+l1j)DnjfDjg=(2n+k+l2n)(2n+k+l2n)j=0n(1)l+1+j(n+k1nj)(n+1lj)DnjfDj(D1lg)\begin{split}[f&,g]_{n}=\sum_{j=1-l}^{n}(-1)^{j}\binom{n+k-1}{n-j}\binom{n+l-1}{j}D^{n-j}fD^{j}g\\ &=\frac{\binom{2n+k+l-2}{n}}{\binom{2n+k+l-2}{n^{\prime}}}\sum_{j=0}^{n^{\prime}}(-1)^{l+1+j}\binom{n^{\prime}+k-1}{n^{\prime}-j}\binom{n^{\prime}+1-l}{j}D^{n^{\prime}-j}fD^{j}(D^{1-l}g)\end{split}

Therefore,

[f,g]n=(1)l1(2n+k+l2n)(2n+k+l2n)[f,D1lg]n.[f,g]_{n}=(-1)^{l-1}\frac{\binom{2n+k+l-2}{n}}{\binom{2n+k+l-2}{n^{\prime}}}\,[f,D^{1-l}g]_{n^{\prime}}.

Notice that D1lgD^{1-l}g is a meromorphic modular form of positive weight 2l2-l, so [f,g]n[f,g]_{n} is a multiple of [f,D1lg]n[f,D^{1-l}g]_{n^{\prime}}, thus also a meromorphic modular form of weight k+l+2nk+l+2n.

Part 2. kk, l0l\leq 0

\@afterheading

Case 1. 0nk0\leq n\leq-k. we consider the product CKD(f;τ,T)CKD(g;τ,T)CK_{D}^{-}(f;\tau,T)CK_{D}^{-}(g;\tau,-T). With the same calculation as in Part 1, the product is

CKD(f;\displaystyle CK_{D}^{-}(f; τ,T)CKD(g;τ,T)\displaystyle\tau,T)CK_{D}^{-}(g;\tau,-T)
=m=0k(1)k+l(km)!(lm)![f,g]nTm+O(T1k).\displaystyle=\sum_{m=0}^{-k}(-1)^{k+l}(-k-m)!(-l-m)![f,g]_{n}T^{m}+O(T^{1-k}).

So the Rankin–Cohen bracket [f,g]n[f,g]_{n} is again a meromorphic modular form of weight k+l+2nk+l+2n.

Case 2. 1knl1-k\leq n\leq-l. Put n=n+k1n^{\prime}=n+k-1, the same calculation as in (7)\eqref{eq:dfgn} shows that

[f\displaystyle[f ,g]n=j=0n+k1(1)j(n+k1j)(n+l1nj)DnjfDjg\displaystyle,g]_{n}=\sum_{j=0}^{n+k-1}(-1)^{j}\binom{n+k-1}{j}\binom{n+l-1}{n-j}D^{n-j}fD^{j}g
=(ln)(ln)j=0n(1)j(n+1kj)(n+l1nj)Dnj(D1kf)Djg\displaystyle=\frac{\binom{-l}{n}}{\binom{-l}{n^{\prime}}}\,\sum_{j=0}^{n^{\prime}}(-1)^{j}\binom{n^{\prime}+1-k}{j}\binom{n^{\prime}+l-1}{n^{\prime}-j}D^{n^{\prime}-j}(D^{1-k}f)D^{j}g

Thus

[f,g]n=(ln)(ln)[D1kf,g]n[f,g]_{n}=\frac{\binom{-l}{n}}{\binom{-l}{n^{\prime}}}\,[D^{1-k}f,g]_{n^{\prime}}

is a multiple of [D1kf,g]n[D^{1-k}f,g]_{n^{\prime}}, which reduces to Case 1 in Part 1, since D1kfD^{1-k}f is a meromorphic modular form of positive weight 2k2-k.

Case 3. 1ln1kl1-l\leq n\leq 1-k-l. We note that (n+k1j)\binom{n+k-1}{j} is non-vanishing if and only if jn+k1j\leq n+k-1 and (n+l1nj)\binom{n+l-1}{n-j} is non-vanishing if and only if 1lj1-l\leq j. So when n1kln\leq 1-k-l, at least one of these two binomials vanishes, this implies that the Rankin–Cohen bracket

[f,g]n=j=0n(1)j(n+k1j)(n+l1nj)DnjfDjg0[f,g]_{n}=\sum_{j=0}^{n}(-1)^{j}\binom{n+k-1}{j}\binom{n+l-1}{n-j}D^{n-j}fD^{j}g\equiv 0

is always vanishing in this case.

Case 4. n2kln\geq 2-k-l. Similar to equation (7)\eqref{eq:dfgn}, letting n′′=n+k+l2n^{\prime\prime}=n+k+l-2 we find that

[\displaystyle[ f,g]n=j=1ln+k1(1)j(n+k1j)(n+l1nj)DnjfDjg\displaystyle f,g]_{n}=\sum_{j=1-l}^{n+k-1}(-1)^{j}\binom{n+k-1}{j}\binom{n+l-1}{n-j}D^{n-j}fD^{j}g
=j=0n′′(1)j+l1(n′′+1kj)(n′′+1ln′′j)Dn′′j(D1kf)Dj(D1lg)\displaystyle=\sum_{j=0}^{n^{\prime\prime}}(-1)^{j+l-1}\binom{n^{\prime\prime}+1-k}{j}\binom{n^{\prime\prime}+1-l}{n^{\prime\prime}-j}D^{n^{\prime\prime}-j}(D^{1-k}f)D^{j}(D^{1-l}g)

Thus

[f,g]n=(1)l1[D1kf,D1lg]n′′[f,g]_{n}=(-1)^{l-1}[D^{1-k}f,D^{1-l}g]_{n^{\prime\prime}}

is still a meromorphic modular form of weight k+l+2nk+l+2n. ∎

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 f,gf,g be two meromorphic modular forms of weight kk, ll respectively. Let nn be a positive integer. Suppose that k+l2k+l\geq 2 or k+l+2n0k+l+2n\leq 0. Then we have

(8) Dni[f,g]i=j=0nci,jk,l,n(Dnjf)(Djg),D^{n-i}[f,g]_{i}=\sum_{j=0}^{n}c_{i,j}^{k,l,n}(D^{n-j}f)(D^{j}g),

and

(9) (Dnif)(Dig)=j=0nbi,jk,l,nDnj[f,g]j,(D^{n-i}f)(D^{i}g)=\sum_{j=0}^{n}b_{i,j}^{k,l,n}D^{n-j}[f,g]_{j},

where

ci,jk,l,n=r=0i(1)r(ninjr)(k+i1ir)(l+i1r),c_{i,j}^{k,l,n}=\sum_{r=0}^{i}(-1)^{r}\binom{n-i}{n-j-r}\binom{k+i-1}{i-r}\binom{l+i-1}{r},

and

bi,jk,l,n=(nj)r=0j(1)r(jr)(k+ni1nir)(l+i1r+ij)(ni)(k+l+n+j1nj)(k+l+2j2j).b_{i,j}^{k,l,n}=\frac{\binom{n}{j}\sum_{r=0}^{j}(-1)^{r}\binom{j}{r}\binom{k+n-i-1}{n-i-r}\binom{l+i-1}{r+i-j}}{\binom{n}{i}\binom{k+l+n+j-1}{n-j}\binom{k+l+2j-2}{j}}.
Proof.

Proposition 4.6 in [6] shows that for positive weights kk, ll the constants bi,jk,l,nb_{i,j}^{k,l,n} and ci,jk,l,nc_{i,j}^{k,l,n} are mutually inverse

(10) r=0nbi,rk,l,ncr,jk,l,n=δi,j=r=0nci,rk,l,nbr,jk,l,n.\displaystyle\sum_{r=0}^{n}b_{i,r}^{k,l,n}c_{r,j}^{k,l,n}=\delta_{i,j}=\sum_{r=0}^{n}c_{i,r}^{k,l,n}b_{r,j}^{k,l,n}.

We note that for fixed ii, jj and nn, the coefficients ci,jk,l,nc_{i,j}^{k,l,n} and bi,jk,l,nb_{i,j}^{k,l,n} are actually polynomials in k,lk,l. Since (10)\eqref{eq:b&c} holds for all positive integers kk, ll, it still holds for all integers kk, ll. As long as the denominator of bi,rk,l,nb_{i,r}^{k,l,n} is non-zero, the identities (8) and (9) remain valid. ∎

Define now the following two \mathbb{C}-subalgebras of 𝒬\mathcal{Q}\mathcal{M}^{*} with {mero,!}*\in\{\text{mero},!\}

𝒬+=(k>0p=0k21𝒬k,p),𝒬=k0p0𝒬k,p.\mathcal{Q}\mathcal{M}^{*}_{+}=\bigg{(}\bigoplus_{k>0}\bigoplus_{p=0}^{\frac{k}{2}-1}\mathcal{Q}\mathcal{M}_{k}^{*,\,p}\bigg{)}\bigoplus\mathbb{C},\quad\mathcal{Q}\mathcal{M}^{*}_{-}=\bigoplus_{k\leq 0}\bigoplus_{p\geq 0}\mathcal{Q}\mathcal{M}_{k}^{*,\,p}.

Then the space 𝒬+\mathcal{Q}\mathcal{M}^{*}_{+} is generated by positive weight derivatives of positive weight meromorphic modular forms and the space 𝒬\mathcal{Q}\mathcal{M}^{*}_{-} 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.

Proof of Theorem 1.2 The result follows immediately from Laphier–El Gradechi formula (9). The condition k+l2k+l\geq 2 or k+l+2n0k+l+2n\leq 0 is automatically satisfied in 𝒬+\mathcal{Q}\mathcal{M}_{+}^{*} and 𝒬\mathcal{Q}\mathcal{M}_{-}^{*} respectively. ∎

Remark.

In fact, the Rankin–Cohen brackets can be extended to meoromorphic quasi-modular forms using DD and ϑ\vartheta. Let f𝒬k,sf\in\mathcal{Q}\mathcal{M}_{k}^{*,\,\leq s} and g𝒬l,tg\in\mathcal{Q}\mathcal{M}_{l}^{*,\,\leq t} be two meoromorphic quasi-modular forms. Their (Serre–)Rankin–Cohen brackets can be defined as

[f,g]n=j=0n(1)j(n+k+s1j)(n+l+t1nj)DnjfDjg,[f,g]_{n}=\sum_{j=0}^{n}(-1)^{j}\binom{n+k+s-1}{j}\binom{n+l+t-1}{n-j}D^{n-j}fD^{j}g,
Se[f,g]n=j=0n(1)j(n+k1j)(n+l1nj)ϑnjfϑjg,\textrm{Se}\,[f,g]_{n}=\sum_{j=0}^{n}(-1)^{j}\binom{n+k-1}{j}\binom{n+l-1}{n-j}\vartheta^{n-j}f\vartheta^{j}g,

where [f,g]n[f,g]_{n} and Se[f,g]n\textrm{Se}\,[f,g]_{n} are meromorphic quasi-modular forms in 𝒬k+l+2n,s+t\mathcal{Q}\mathcal{M}_{k+l+2n}^{*,\,\leq s+t}.

6. Fourier coefficients of meromorphic quasi-modular forms

Let us denote by

={τ{}|1/2Re(τ)<1/2}\mathcal{F}=\{\tau\in\mathcal{H}\cup\{\infty\}\,|\,-1/2\leq\operatorname{Re}(\tau)<1/2\}

the standard fundamental domain of the translation ττ+1\tau\mapsto\tau+1 in {}\mathcal{H}\cup\{\infty\}. We set also the trimmed fundamental domain for t0>0t_{0}>0 by

t0={τ|Imτt0}.\mathcal{F}_{t_{0}}=\{\tau\in\mathcal{F}\,|\,\operatorname{Im}\tau\geq t_{0}\}.

Let ff be a meromorphic function on \mathcal{F}. For any pole α\alpha of ff with order ordf(α)\operatorname{ord}_{f}(\alpha), let

(11) PP(f)(τ)=n<0af(n)qn\operatorname{PP}_{\infty}(f)(\tau)=\sum_{n<0}a_{f}(n)q^{n}

be the principle part of the Fourier expansion of ff at α=\alpha=\infty and

(12) PPα(f)(τ)=m=1ordf(α)cf,α(m)1(τα)m\operatorname{PP}_{\alpha}(f)(\tau)=\sum_{m=1}^{\operatorname{ord}_{f}(\alpha)}c_{f,\alpha}(m)\frac{1}{(\tau-\alpha)^{m}}

be the principle part of the Laurent expansion of ff at τ=α\tau=\alpha. We put also

P(f)=n<0af(n)qn,P_{\infty}(f)=\sum_{n<0}a_{f}(n)q^{n},

and

Pα(f)=m=1ordf(α)(2πi)m(m1)!cf,α(m)Li1m(e(τα)),P_{\alpha}(f)=\sum_{m=1}^{\operatorname{ord}_{f}(\alpha)}\frac{(-2\pi i)^{m}}{(m-1)!}c_{f,\alpha}(m)\operatorname{Li}_{1-m}(\textbf{e}(\tau-\alpha)),

where e(τ)=e2πiτ\textbf{e}(\tau)=e^{2\pi i\tau}. Note that the expansion (25)\eqref{eq:limz} in Appendix shows that Pα(f)P_{\alpha}(f) has the same principle part at τ=α\tau=\alpha as ff does.

We have the following estimation on Fourier coefficients of meromorphic quasi-modular form

Proposition 6.1.

Let f=naf(n)qnf=\sum_{n\gg-\infty}a_{f}(n)q^{n} be a meromorphic quasi-modular form of weight kk with poles and principle parts as described above. Then we have

af(n)=αt0,1mordf(α)(2πi)m(m1)!cf,α(m)nm1e2πinα+O(e2πnt0).a_{f}(n)=\sum_{\begin{subarray}{c}\alpha\in\mathcal{F}_{t_{0}},\\ 1\leq m\leq\operatorname{ord}_{f}(\alpha)\end{subarray}}\frac{(-2\pi i)^{m}}{(m-1)!}c_{f,\alpha}(m)n^{m-1}e^{-2\pi in\alpha}+O(e^{2\pi nt_{0}}).
Proof.

Let α1,,αl\alpha_{1},\dots,\alpha_{l} be all the poles of ff in t0\mathcal{F}_{t_{0}}. Removing all the principle parts with all these kinds of Pα(f)P_{\alpha}(f), we have finally a holomorphic function in t0\mathcal{F}_{t_{0}}

f~(τ)=f(τ)Pα1(f)Pαl(f).\tilde{f}(\tau)=f(\tau)-P_{\alpha_{1}}(f)-\dots-P_{\alpha_{l}}(f).

So f~\tilde{f} is bounded in t0\mathcal{F}_{t_{0}}, say by CC. Note the function f~(τ)\tilde{f}(\tau) is holomorphic within the closure of the domain t0\mathcal{F}_{t_{0}}. Then using Cauchy integral formula at infinity, the nn-th coefficient of f~(τ)\tilde{f}(\tau) is bounded by

(13) |it0it0+1f~(τ)e2πinτ𝑑τ|e2πint001|f~(t+it0)|𝑑tCe2πnt0.\left|\int_{it_{0}}^{it_{0}+1}\tilde{f}(\tau)e^{-2\pi in\tau}d\tau\right|\leq e^{2\pi int_{0}}\int_{0}^{1}\left|\tilde{f}(t+it_{0})\right|dt\leq Ce^{2\pi nt_{0}}.

On the other hand, the nn-th coefficient of the polylogarithm function Pα(f)P_{\alpha}(f) is

m=1ordf(α)(2πi)m(m1)!cf,α(m)nm1e2πinα.\sum_{m=1}^{\operatorname{ord}_{f}(\alpha)}\frac{(-2\pi i)^{m}}{(m-1)!}c_{f,\alpha}(m)n^{m-1}e^{-2\pi in\alpha}.

Combining with the estimation (13)\eqref{eq:esti}, 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 f=nn0af(n)qn𝒬k!,pf=\sum_{n\geq-n_{0}}a_{f}(n)q^{n}\in\mathcal{Q}\mathcal{M}_{k}^{!,\,p} be a weakly holomorphic quasi-modular form with n0>0n_{0}>0. Then for any ε>0\varepsilon>0, we have

af(n)e(4π+ε)n0n.a_{f}(n)\ll e^{(4\pi+\varepsilon)\sqrt{n_{0}n}}.
Proof.

When FF is a weakly holomorphic modular form, then we have the following estimation [20] on the Fourier coefficients of FF

(14) aF(n)e(4π+ϵ)n0n.a_{F}(n)\ll e^{(4\pi+\epsilon)\sqrt{n_{0}n}}.

Theorem 4.2 shows that every weakly holomorphic quasi-modular form is of the form

f=F0+F1E2++FpE2p.f=F_{0}+F_{1}E_{2}+\cdots+F_{p}E_{2}^{p}.

Note that the order of FiF_{i} at infinity is at most n0n_{0} for any i=0,1,,pi=0,1,\cdots,p. Combining the estimation (14)\eqref{eq:afn}, we get

af(n)e(4π+ε)n0n,a_{f}(n)\ll e^{(4\pi+\varepsilon)\sqrt{n_{0}n}},

since σ(n)n1+ϵ\sigma(n)\ll n^{1+\epsilon} for any ϵ>0\epsilon>0. ∎

Remark.

In fact, we can get a more accurate estimation by using the Circle Method to weakly holomorphic quasi-modular forms

af(n)n2k34e4πn0n.a_{f}(n)\ll n^{\frac{2k-3}{4}}e^{4\pi\sqrt{n_{0}n}}.

7. Regularized integrals of meromorphic functions

At the start of this section, we recall the LL-functions of modular forms. For a cusp form ff of weight kk on \mathcal{H}, the completed LL-function of ff is just the Mellin transform of ff

Λ(f,s)=0f(it)ts1𝑑t.\Lambda(f,s)=\int_{0}^{\infty}f(it)t^{s-1}dt.

This completed LL-function is connected with the Dirichlet series of ff in the following way

Λ(f,s)=Γ(s)(2π)sn=1af(n)ns.\Lambda(f,s)=\frac{\Gamma(s)}{(2\pi)^{s}}\sum_{n=1}^{\infty}\frac{a_{f}(n)}{n^{s}}.

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 0) and the regularization at positive real numbers.

Following [2], under the assumption that ff has at most linear exponential growth at infinity, we give the definition of regularized integral of ff.

Definition 7.1.

Let f(t)f(t) be an analytic function with at most linear exponential growth for large t>0t\in\mathbb{R}_{>0}. If the integral

t0ewtf(t)𝑑t\int_{t_{0}}^{\infty}e^{-wt}f(t)dt

has a continuation to w=0w=0, then the regularized integral of ff is defined to be

t0,f(t)𝑑t[t0ewtf(t)𝑑t]w=0.\int_{t_{0}}^{\infty,\ast}f(t)dt\coloneqq\left[\int_{t_{0}}^{\infty}e^{-wt}f(t)dt\right]_{w=0}.

We use the notation * to indicate a regularized integral.

Similarly, if f(1/t)f(1/t) has at most linear exponential growth at the cusp 0, then we can define the regularized integral of ff at 0 using the reflection t1/tt\mapsto 1/t. This is to say,

0t0,f(t)𝑑tt01,1t2f(1t)𝑑t.\int_{0}^{t_{0},\ast}f(t)dt\coloneqq\int_{t_{0}^{-1}}^{\infty,\ast}\frac{1}{t^{2}}f\left(\frac{1}{t}\right)dt.

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 f(t)f(t) be a meromorphic function in a neighbourhood of [a,b][a,b] with only one real positive pole a<c<ba<c<b, then the regularized integral of ff from aa to bb is defined as

ab,f(t)𝑑t[ab|tc|sf(t)𝑑t]s=0,\int_{a}^{b,*}f(t)dt\coloneqq\left[\int_{a}^{b}|t-c|^{s}f(t)dt\right]_{s=0},

where the suffix indicates the constant term in the Laurent expansion of ss at 0.

As already mentioned, there are different approaches of Hadamard regularization. The following proposition explains why they are actually equivalent.

Proposition 7.3.

Let f(t)f(t) be a meromorphic function in a neighbourhood of [a,b][a,b] with only one real positive pole a<c<ba<c<b of order nn. Let F(t)=f(t)(tc)nF(t)=f(t)(t-c)^{n}. The the following different approaches of regularization coincides

  1. (i)

    The meromorphic continuation of the integral by Riesz

    [ab|tc|sf(t)𝑑t]s=0.\left[\int_{a}^{b}|t-c|^{s}f(t)dt\right]_{s=0}.
  2. (ii)

    The integral of f(t)f(t) in the sense of Sato’s hyperfunction. Equivalently, the Cauchy principal valued integral by

    12(C++C)f(t)dt,\frac{1}{2}\biggl{(}\int_{C^{+}}+\int_{C^{-}}\biggr{)}f(t)dt,

    where C+C^{+} (resp. CC^{-}) is a path from aa to bb above (resp. below) the real axis.

  3. (iii)

    The Hadamard finite part integral of f(t)f(t)

    FPε=0(acε+c+εb)f(t)dt,\operatorname*{FP}_{\varepsilon=0}\,\biggl{(}\int_{a}^{c-\varepsilon}+\int_{c+\varepsilon}^{b}\biggr{)}f(t)dt,

    where FP\,\operatorname*{FP} stands for the constant term in the Laurent expansion with respect to ε\varepsilon.

  4. (iv)

    The following integral in the sense of pairing the Schwartz distribution FP(xc)n\operatorname*{FP}\,(x-c)^{-n} with F(t)F(t)

    (FP1(xc)n,F(t))\displaystyle\Bigl{(}\operatorname*{FP}\,\frac{1}{(x-c)^{n}},F(t)\Bigr{)} i=0n1(ni1)!n!(F(i)(a)(ac)niF(i)(b)(bc)ni)\displaystyle\coloneqq\sum_{i=0}^{n-1}\frac{(n-i-1)!}{n!}\biggl{(}\frac{F^{(i)}(a)}{(a-c)^{n-i}}-\frac{F^{(i)}(b)}{(b-c)^{n-i}}\biggr{)}
    +1n!PVabF(n)(t)tc𝑑t,\displaystyle\quad\quad+\frac{1}{n!}\operatorname{PV}\int_{a}^{b}\frac{F^{(n)}(t)}{t-c}dt,

    where PV\operatorname{PV} stands for the Cauchy principal value of the integral.

  5. (v)

    The following Cauchy principle value given by Sokhotski–Plemelj formula, i.e.

    12±(limuc±i0abF(t)(tu)n𝑑t),\frac{1}{2}\sum_{\pm}\biggl{(}\lim_{u\to c\pm i0}\int_{a}^{b}\frac{F(t)}{(t-u)^{n}}dt\biggr{)},

    where uu tends to cc on both sides of real axis.

Proof.

(i)(ii)(iii)(i)\Leftrightarrow(ii)\Leftrightarrow(iii). If f(t)f(t) is holomorphic then the implication is immediate. So it suffices to check the function f(t)=(tc)nf(t)=(t-c)^{-n}. We may set the paths C±C^{\pm} to be the paths agreed with the real axis but modified with small upper (resp. lower) semi-circles Sε±S_{\varepsilon}^{\pm} at cc of radius ε\varepsilon.

C+C^{+}CC^{-}cc

For Re(s)0\operatorname{Re}(s)\ll 0, the meromorphic continuation gives us

cεc+ε|tc|sf(t)𝑑t\displaystyle\int_{c-\varepsilon}^{c+\varepsilon}|t-c|^{s}f(t)dt =(1)ncεc(ct)sn𝑑t+cc+ε(tc)sn𝑑t\displaystyle=(-1)^{n}\int_{c-\varepsilon}^{c}(c-t)^{s-n}dt+\int_{c}^{c+\varepsilon}(t-c)^{s-n}dt
=(1+(1)n)εs+1ns+1n.\displaystyle=(1+(-1)^{n})\frac{\varepsilon^{s+1-n}}{s+1-n}.

Hence, the constant term in the Laurent expansion at s=0s=0 is

(15) [cεc+ε|tc|sf(t)𝑑t]s=0={2ε1n/(1n)n even0n odd.\left[\int_{c-\varepsilon}^{c+\varepsilon}|t-c|^{s}f(t)dt\right]_{s=0}=\begin{cases}2\,\varepsilon^{1-n}/(1-n)&n\text{ even}\\ 0&n\text{ odd}\end{cases}.

Meanwhile, the integrations along the two small semi-circles Sε±S_{\varepsilon}^{\pm} give

12(Sε++Sε)1(tc)ndt\displaystyle\frac{1}{2}\left(\int_{S^{+}_{\varepsilon}}+\int_{S^{-}_{\varepsilon}}\right)\frac{1}{(t-c)^{n}}dt =i2(π0ε1ndθei(n1)θ+π0ε1ndθei(n1)θ)\displaystyle=\frac{i}{2}\left(\int_{\pi}^{0}\varepsilon^{1-n}\frac{d\theta}{e^{i(n-1)\theta}}+\int_{-\pi}^{0}\varepsilon^{1-n}\frac{d\theta}{e^{i(n-1)\theta}}\right)

An elementary calculation shows that this integral coincides with (15)\eqref{eq:epsinodd}. This implies that the meromorphic continuation yields the same result as hyperfunction.

Observe that the above integrals have always finite part 0 with respect to ε\varepsilon. It follows that the Hadamard finite part integral has the same value as the integral of hyperfunction. These prove that (i)(i), (ii)(ii) and (iii)(iii) are equal.

(iii)(iv)(iii)\Leftrightarrow(iv). Integrating by parts, for any testing function ϕ\phi we get

abϕ(t)(tc)n𝑑t=\displaystyle\int_{a}^{b}\frac{\phi(t)}{(t-c)^{n}}dt= i=0n1(ni1)!n!(ϕ(i)(a)(ac)niϕ(i)(b)(bc)ni)\displaystyle\sum_{i=0}^{n-1}\frac{(n-i-1)!}{n!}\biggl{(}\frac{\phi^{(i)}(a)}{(a-c)^{n-i}}-\frac{\phi^{(i)}(b)}{(b-c)^{n-i}}\biggr{)}
+i=0n1ϕ(i)(c)i!1(1)ni(ni)εni+1n!(acε+c+εb)ϕ(n)(t)tcdt.\displaystyle+\sum_{i=0}^{n-1}\frac{\phi^{(i)}(c)}{i!}\frac{1-(-1)^{n-i}}{(n-i)\varepsilon^{n-i}}+\frac{1}{n!}\biggl{(}\int_{a}^{c-\varepsilon}+\int_{c+\varepsilon}^{b}\biggr{)}\frac{\phi^{(n)}(t)}{t-c}dt.

The last term is a convergent integral with Cauchy principal value as ε0\varepsilon\to 0. Thus

(FP1(xc)n,ϕ(t))=\displaystyle\Bigl{(}\operatorname*{FP}\,\frac{1}{(x-c)^{n}},\phi(t)\Bigr{)}= i=0n1(ni1)!n!(ϕ(i)(a)(ac)niϕ(i)(b)(bc)ni)\displaystyle\sum_{i=0}^{n-1}\frac{(n-i-1)!}{n!}\biggl{(}\frac{\phi^{(i)}(a)}{(a-c)^{n-i}}-\frac{\phi^{(i)}(b)}{(b-c)^{n-i}}\biggr{)}
+1n!limε0(acε+c+εb)ϕ(n)(t)tcdt\displaystyle+\frac{1}{n!}\lim_{\varepsilon\to 0}\biggl{(}\int_{a}^{c-\varepsilon}+\int_{c+\varepsilon}^{b}\biggr{)}\frac{\phi^{(n)}(t)}{t-c}dt

is the finite part with respect to ε\varepsilon. This shows (iii)(iv)(iii)\Leftrightarrow(iv).

(iv)(v)(iv)\Leftrightarrow(v). This parts follows closely with Fox [7] and Gelfand–Shilov [8]. Again using integration by parts, for any uu not in [a,b][a,b], one has

abF(t)(tu)n𝑑t=\displaystyle\int_{a}^{b}\frac{F(t)}{(t-u)^{n}}dt= i=0n1(ni1)!n!(F(i)(a)(au)niF(i)(b)(bu)ni)\displaystyle\sum_{i=0}^{n-1}\frac{(n-i-1)!}{n!}\biggl{(}\frac{F^{(i)}(a)}{(a-u)^{n-i}}-\frac{F^{(i)}(b)}{(b-u)^{n-i}}\biggr{)}
+1n!abF(n)(t)tu𝑑t.\displaystyle+\frac{1}{n!}\int_{a}^{b}\frac{F^{(n)}(t)}{t-u}dt.

Let uc±i0u\to c\pm i0 from both sides of real axis, by the Sokhotski–Plemelj formula of Cauchy principal valued integral we obtain (iv)(v)(iv)\Leftrightarrow(v). ∎

In general, let f(t)f(t) be a function which has a finite number of positive real poles and has at most linear exponential growth at 0 and infinity. Consider finitely many open intervals

(a1,a2),(a2,a3),(an1,an).(a_{1},a_{2}),(a_{2},a_{3})\dots,(a_{n-1},a_{n}).

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

(0,a1),(an,),(0,a_{1}),(a_{n},\infty),

we assign to them the regularized integrals from Definition 7.1. At last, we sum up them all. This gives the regularized integral of ff on (0,)(0,\infty), written again as

0,f(t)𝑑t.\int_{0}^{\infty,*}f(t)dt.

It is clear that the above definition is independent of the choice of intervals.

8. LL-function of meromorphic quasi-modular forms

We first define the LL-function of a meromorphic quasi-modular form through the regularized integral defined in Section 7.

Definition 8.1.

Let ff be a meromorphic quasi-modular form. Then we define its complete LL-function by

Λ(f,s)=0,f(it)ts1𝑑t.\Lambda(f,s)=\int_{0}^{\infty,\ast}f(it)t^{s-1}dt.

The Dirichlet LL-function associated to ff is defined as

L(f,s)=(2π)sΓ(s)Λ(f,s).L(f,s)=\frac{(2\pi)^{s}}{\Gamma(s)}\Lambda(f,s).

In the following, we will give an explicit formula for Λ(f,s)\Lambda(f,s) for any meromorphic quasi-modular form ff. We first consider the regularized integral of meromorphic function ff at infinity.

Lemma 8.2.

Let ff be a meromorphic function in a neighbourhood of the half-strip t0\mathcal{F}_{t_{0}} with only pole at infinity. Suppose its Fourier expansion at infinity is given as f(τ)=nn0af(n)qnf(\tau)=\sum_{n\geq-n_{0}}a_{f}(n)q^{n}. Then the regularized integral of f(it)ts1f(it)t^{s-1} exists and defines a meromorphic function in ss. More precisely, we have

t0,f(it)ts1𝑑t=af(0)t0ss+n0af(n)Γ(s,2πnt0)(2πn)s.\int_{t_{0}}^{\infty,\ast}f(it)t^{s-1}dt=-\frac{a_{f}(0)t_{0}^{s}}{s}+\sum_{n\neq 0}\frac{a_{f}(n)\Gamma(s,2\pi nt_{0})}{(2\pi n)^{s}}.
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 {reiθ|r>0}\{re^{i\theta}\,|\,r\in\mathbb{R}_{>0}\} , where θ(π,32π)\theta\in(\pi,\frac{3}{2}\pi) is a fixed angle. It is easy to see that when Re(w)>2πn0\operatorname{Re}(w)>2\pi n_{0}, the integral

t0n0af(n)e(w+2πn)tts1dt\int_{t_{0}}^{\infty}\sum_{n\neq 0}a_{f}(n)e^{-(w+2\pi n)t}t^{s-1}dt

is absolutely convergent for any ss\in\mathbb{C}.

Since ff is holomorphic in a neighbourhood of the half-strip t0\mathcal{F}_{t_{0}}, its Fourier coefficients satisfy af(n)=O(e2πnt0)a_{f}(n)=O(e^{2\pi nt_{0}}). This ensures that the value of the above integral is the absolutely convergent sum

n0af(n)Γ(s,(2πn+w)t0)(2πn+w)s=n>0+n<0,\sum_{n\neq 0}\frac{a_{f}(n)\Gamma(s,(2\pi n+w)t_{0})}{(2\pi n+w)^{s}}=\sum_{n>0}+\sum_{n<0}\,,

in view of Γ(s,2πnt0)(2πnt0)s1e2πnt0\Gamma(s,2\pi nt_{0})\sim(2\pi nt_{0})^{s-1}e^{-2\pi nt_{0}}.

We can see that the partial sum n>0\sum_{n>0} defines a holomorphic function of (w,s)(w,s) with Re(w)>2π\operatorname{Re}(w)>-2\pi and ss\in\mathbb{C}. The sum n<0\sum_{n<0} is a finite sum, it can be continued to a holomorphic function of (ω,s)(\omega,s) in the open domain

\m=1n0{m+reiθ|r0}×.\mathbb{C}\,\backslash\bigcup_{m=1}^{n_{0}}\{m+re^{i\theta}\,|\,r\in\mathbb{R}_{\geq 0}\}\times\mathbb{C}.

Hence both parts can be extend to a holomorphic function of ss in a neighbourhood of w=0w=0.

We only need to deal with the term n=0n=0. When w=0w=0 and Re(s)<0\operatorname{Re}(s)<0, the integral has well-defined value t0s/s-t_{0}^{s}/s (cf. [2, Remark after Prop. 3.3]). This extends to a meromorphic function to the whole complex plane in ss.

At last we remark that the above evaluation is independent of the choice of θ\theta. ∎

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 m0m\geq 0 be an integer. For s,αs,\alpha\in\mathbb{C} with 1/2Re(α)<1/2-1/2\leq\operatorname{Re}(\alpha)<1/2, we define the regularized integral

Jm(s,α)=0,Lim(e(itα))tsdtt.J_{m}(s,\alpha)=\int_{0}^{\infty,\ast}\operatorname{Li}_{-m}(\textbf{e}(it-\alpha))t^{s}\frac{dt}{t}.
Lemma 8.3.

Let Re(s)>0\operatorname{Re}(s)>0 and mm be a positive integer. If Im(α)<0\operatorname{Im}(\alpha)<0, then we have

Jm(s,α)=Γ(s)(2π)sLism(e2πiα).J_{m}(s,\alpha)=\frac{\Gamma(s)}{(2\pi)^{s}}\operatorname{Li}_{s-m}(e^{-2\pi i\alpha}).

If Im(α)>0\operatorname{Im}(\alpha)>0, we have

(16) Jm(s,α)=eiπ(sm)(2π)mΓ(s)Γ(sm)ζ(1s+m,Reα+1α)Γ(s)eiπ(sm)(2π)sLism(e2πiα)+δReα=0Γ(s)Γ(sm)is(α)sm12(2πi)m.\begin{split}J_{m}(s,&\,\alpha)=\frac{e^{i\pi(s-m)}}{(2\pi)^{m}}\frac{\Gamma(s)}{\Gamma(s-m)}\zeta(1-s+m,\lfloor\operatorname{Re}\alpha\rfloor+1-\alpha)\\ &-\frac{\Gamma(s)e^{i\pi(s-m)}}{(2\pi)^{s}}\operatorname{Li}_{s-m}(e^{2\pi i\alpha})+\delta_{\operatorname{Re}\alpha=0}\frac{\Gamma(s)}{\Gamma(s-m)}\frac{i^{s}(-\alpha)^{s-m-1}}{2(2\pi i)^{m}}.\end{split}
Proof.

Suppose that α\alpha is given with Im(α)<0\operatorname{Im}(\alpha)<0, then |e(itα)|<1|\textbf{e}(it-\alpha)|<1, so we have the convergent integral

(17) Jm(s,α)=0n=1e2πn(y+iα)nmtsdtt=Γ(s)(2π)sn=1e2πinαnsm=Γ(s)(2π)sLism(e2πiα).\begin{split}&J_{m}(s,\alpha)=\int_{0}^{\infty}\sum_{n=1}^{\infty}\frac{e^{-2\pi n(y+i\alpha)}}{n^{-m}}t^{s}\frac{dt}{t}\\ =&\frac{\Gamma(s)}{(2\pi)^{s}}\sum_{n=1}^{\infty}\frac{e^{-2\pi in\alpha}}{n^{s-m}}=\frac{\Gamma(s)}{(2\pi)^{s}}\operatorname{Li}_{s-m}(e^{-2\pi i\alpha}).\end{split}

Note both sides extend to a holomorphic function of α\alpha except only when α\alpha on the imaginary axis. Thus it holds for all Re(α)0\operatorname{Re}(\alpha)\neq 0.

When Im(α)>0\operatorname{Im}(\alpha)>0, if α\alpha not on the imaginary axis, the formula (16)\eqref{eq:j1mam} just follows from rewriting (17) with the reflection formula of polylogarithm (26) in the Appendix.

The difficulty arises as α=ai\alpha=ai is on the imaginary axis where a>0a\in\mathbb{R}_{>0}, where we encounter a Hadamard regularized integral. In this case we may rewrite the integral as

Jm(s,α)=0,Lim(e(τα))(τi)sdττ.J_{m}(s,\alpha)=\int_{0}^{\infty,*}\operatorname{Li}_{-m}(\textbf{e}(\tau-\alpha))\left(\frac{\tau}{i}\right)^{s}\frac{d\tau}{\tau}.

We recall that when mm is a positive integer, the polylogarithm Lim(z)\operatorname{Li}_{-m}(z) is rational function. So the integrand is in fact a rational function of e2πiαe^{2\pi i\alpha}. By the Sokhotski–Plemelj formula (v)(v) in Proposition 7.3, the value of Jm(s,α)J_{m}(s,\alpha) on imaginary axis should be the mean value of limits as Re(α)\operatorname{Re}(\alpha) tends to 0 on left side and right side of imaginary axis. Moreover, when z[1,)z\in[1,\infty), by the reflection formula (26), we get

(18) limϵ0+Lis(ze2πiϵ)Lis(ze2πiϵ)=2πiΓ(s)(lnz)s1.\lim_{\epsilon\to 0^{+}}\operatorname{Li}_{s}(ze^{2\pi i\epsilon})-\operatorname{Li}_{s}(ze^{-2\pi i\epsilon})=\frac{2\pi i}{\Gamma(s)}(\ln z)^{s-1}.

So by combining (17)\eqref{eq:jnsa} and (18)\eqref{eq:refep}, we have

Jm(s,α)\displaystyle J_{m}(s,\alpha) =12Γ(s)(2π)slimε0+(Lism(e2πi(α+ε))+Lism(e2πi(αε)))\displaystyle=\frac{1}{2}\frac{\Gamma(s)}{(2\pi)^{s}}\lim\limits_{\varepsilon\to 0^{+}}\left(\operatorname{Li}_{s-m}(e^{-2\pi i(\alpha+\varepsilon)})+\operatorname{Li}_{s-m}(e^{-2\pi i(\alpha-\varepsilon)})\right)
=12Γ(s)(2π)s(2Lism(e2πa)+2πiΓ(sm)(2πa)sm1).\displaystyle=\frac{1}{2}\frac{\Gamma(s)}{(2\pi)^{s}}\left(2\operatorname{Li}_{s-m}(e^{2\pi a})+\frac{2\pi i}{\Gamma(s-m)}(2\pi a)^{s-m-1}\right).

Using the reflection formula again, we get

Jm(s,α)\displaystyle J_{m}(s,\alpha) =eiπ(sm)Γ(s)(2π)sLism(e2πa)+is(2πi)mΓ(s)Γ(sm)ζ(1s+m,1ia)\displaystyle=-\frac{e^{i\pi(s-m)}\Gamma(s)}{(2\pi)^{s}}\operatorname{Li}_{s-m}(e^{-2\pi a})+\frac{i^{s}}{(2\pi i)^{m}}\frac{\Gamma(s)}{\Gamma(s-m)}\zeta(1-s+m,1-ia)
+is2(2πi)mΓ(s)Γ(sm)(ia)sm1.\displaystyle+\frac{i^{s}}{2(2\pi i)^{m}}\frac{\Gamma(s)}{\Gamma(s-m)}(-ia)^{s-m-1}.

More generally, we will encounter the following integral. Let m0m\geq 0 be an integer, t0>0t_{0}>0 be a real number and s,αs,\alpha\in\mathbb{C}, we define

Gm(s,α,t0)=t0,Lim(e(itα))tsdtt.G_{m}(s,\alpha,t_{0})=\int_{t_{0}}^{\infty,*}\operatorname{Li}_{-m}(\textbf{e}(it-\alpha))t^{s}\frac{dt}{t}.
Lemma 8.4.

Let m0m\geq 0 be an integer and t0t_{0} be a positive real number. Then the function Gm(s,α,t0)G_{m}(s,\alpha,t_{0}) extends to an entire function for all ss\in\mathbb{C} and

Gm(s,α,t0)=G1,m(s,α,t0)+G2,m(s,α,t0).G_{m}(s,\alpha,t_{0})=G_{1,m}(s,\alpha,t_{0})+G_{2,m}(s,\alpha,t_{0}).

When Im(α)<t0\operatorname{Im}(\alpha)<t_{0}, we have

G1,m(s,α,t0)=1(2π)mn=0e2πinαΓ(s,2πnt0)(2πn)sm and G2,m(s,α,t0)=0.G_{1,m}(s,\alpha,t_{0})=\frac{1}{(2\pi)^{m}}\sum_{n=0}^{\infty}\frac{e^{-2\pi in\alpha}\Gamma(s,2\pi nt_{0})}{(2\pi n)^{s-m}}\text{ and }G_{2,m}(s,\alpha,t_{0})=0.

When Im(α)>t0\operatorname{Im}(\alpha)>t_{0}, we have

G1,m(s,α,t0)=eiπ(sm)(2π)m(δm=0t0ss+n=1e2πinαΓ(s,2πnt0)(2πn)sm)G_{1,m}(s,\alpha,t_{0})=-\frac{e^{-i\pi(s-m)}}{(2\pi)^{m}}\left(\delta_{m=0}\frac{t_{0}^{s}}{s}+\sum_{n=1}^{\infty}\frac{e^{2\pi in\alpha}\Gamma(s,-2\pi nt_{0})}{(2\pi n)^{s-m}}\right)

and

G2,m\displaystyle G_{2,m} (s,α,t0)=is(2πi)mΓ(s)Γ(sm)ζ(1s+m,Reα+1α)\displaystyle(s,\alpha,t_{0})=\frac{i^{s}}{(2\pi i)^{m}}\frac{\Gamma(s)}{\Gamma(s-m)}\zeta(1-s+m,\lfloor\operatorname{Re}\alpha\rfloor+1-\alpha)
+(1)m1(2π)s2πiΓ(1s)Lism(e2πiα)+δReα=0is2(2πi)mΓ(s)Γ(sm)(α)sm1.\displaystyle+\frac{(-1)^{m-1}}{(2\pi)^{s}}\frac{2\pi i}{\Gamma(1-s)}\operatorname{Li}_{s-m}(e^{2\pi i\alpha})+\delta_{\operatorname{Re}\alpha=0}\frac{i^{s}}{2(2\pi i)^{m}}\frac{\Gamma(s)}{\Gamma(s-m)}(-\alpha)^{s-m-1}.
Proof.

When Imα<t0\operatorname{Im}\alpha<t_{0}, we can integrate term-wisely. It is immediate that for any ss\in\mathbb{C}

Gm(s,α,t0)=n=1t0e2πn(t+iα)nmtsdtt=1(2π)mn=1e2πinαΓ(s,2πnt0)(2πn)sm.G_{m}(s,\alpha,t_{0})=\sum_{n=1}^{\infty}\int_{t_{0}}^{\infty}\frac{e^{-2\pi n(t+i\alpha)}}{n^{-m}}t^{s}\frac{dt}{t}=\frac{1}{(2\pi)^{m}}\sum_{n=1}^{\infty}\frac{e^{-2\pi in\alpha}\Gamma(s,2\pi nt_{0})}{(2\pi n)^{s-m}}.

Here the absolute convergence is guaranteed by Γ(s,2πnt0)(2πnt0)s1e2πnt0\Gamma(s,2\pi nt_{0})\sim(2\pi nt_{0})^{s-1}e^{-2\pi nt_{0}}.

When Im(α)>t0\operatorname{Im}(\alpha)>t_{0}, we may assume that Re(s)>0\operatorname{Re}(s)>0 first. The integral is defined by Hadamard regularization and can not be computed directly. We first evaluate the following convergent integral

0t0Lim(e(itα))tsdtt.\int_{0}^{t_{0}}\operatorname{Li}_{-m}(\textbf{e}(it-\alpha))t^{s}\frac{dt}{t}.

When m>0m>0, by the reflection formula (26)\eqref{eq:reflcli}, this integral equals to

(1)m10t0Lim(e(αit))tsdtt.(-1)^{m-1}\int_{0}^{t_{0}}\operatorname{Li}_{-m}(\textbf{e}(\alpha-it))t^{s}\frac{dt}{t}.

Hence we have

(19) 0t0Lim(e(τα))tsdtt=(1)m1n=10t0tse2πn(t+iα)nmdtt=eiπ(sm)1(2π)mn=1e2πinαγ(s,2πnt0)(2πn)sm\begin{split}&\int_{0}^{t_{0}}\operatorname{Li}_{-m}(\textbf{e}(\tau-\alpha))t^{s}\frac{dt}{t}\\ =&(-1)^{m-1}\sum_{n=1}^{\infty}\int_{0}^{t_{0}}\frac{t^{s}e^{2\pi n(t+i\alpha)}}{n^{-m}}\frac{dt}{t}\\ =&-e^{-i\pi(s-m)}\frac{1}{(2\pi)^{m}}\sum_{n=1}^{\infty}\frac{e^{2\pi in\alpha}\gamma(s,-2\pi nt_{0})}{(2\pi n)^{s-m}}\end{split}

where each term has exponential decay since γ(s,2πnt0)(2πnt0)s1e2πnt0\gamma(s,-2\pi nt_{0})\sim(-2\pi nt_{0})^{s-1}e^{2\pi nt_{0}} as nn grows to infinity. When m=0m=0, we have Li0(x)=x/(1x)\operatorname{Li}_{0}(x)=x/(1-x), so Li0(x)=Li0(1/x)1\operatorname{Li}_{0}(x)=-\operatorname{Li}_{0}(1/x)-1. In this case, the formula becomes

0t0Li0(e(τα))tsdtt=eiπs(t0ss+n=1e2πnaγ(s,2πnt0)(2πn)s)\int_{0}^{t_{0}}\operatorname{Li}_{0}(\textbf{e}(\tau-\alpha))t^{s}\frac{dt}{t}=-e^{-i\pi s}\left(\frac{t_{0}^{s}}{s}+\sum_{n=1}^{\infty}\frac{e^{-2\pi na}\gamma(s,-2\pi nt_{0})}{(2\pi n)^{s}}\right)

To finish the proof, by combining with Lemma 8.3, we have to show that

(20) eiπ(sm)n1e2πinαγ(s,2πnt0)nsmΓ(s)eiπ(sm)Lism(e2πiα)=(1)m12πiΓ(1s)Lism(e2πiα)eiπ(sm)n1e2πinαΓ(s,2πnt0)nsm.\begin{split}&e^{-i\pi(s-m)}\sum_{n\geq 1}\frac{e^{2\pi in\alpha}\gamma(s,-2\pi nt_{0})}{n^{s-m}}-\Gamma(s)e^{i\pi(s-m)}\operatorname{Li}_{s-m}(e^{2\pi i\alpha})\\ =&\frac{(-1)^{m-1}2\pi i}{\Gamma(1-s)}\operatorname{Li}_{s-m}(e^{2\pi i\alpha})-e^{-i\pi(s-m)}\sum_{n\geq 1}\frac{e^{2\pi in\alpha}\Gamma(s,-2\pi nt_{0})}{n^{s-m}}.\end{split}

From Lism(e2πiα)=n1nmse2πinα\operatorname{Li}_{s-m}(e^{2\pi i\alpha})=\sum_{n\geq 1}n^{m-s}e^{2\pi in\alpha}, we know the identity (20)\eqref{eq:eulerref} is equivalent to

eiπ(sm)Γ(s)eiπ(sm)Γ(s)=(1)m12πiΓ(1s).e^{-i\pi(s-m)}\Gamma(s)-e^{i\pi(s-m)}\Gamma(s)=\frac{(-1)^{m-1}2\pi i}{\Gamma(1-s)}.

But this is exactly the Euler’s reflection formula.

For general ss\in\mathbb{C}, we consider analytic continuation on both sides and thus obtain the same formula. Indeed, the function G2,m(s,α,t0)G_{2,m}(s,\alpha,t_{0}) is meromorphic only when m=0m=0. It has a unique single pole at s=0s=0 with residue

Ress=0ζ(1s,Reα+1α)=1.\operatorname*{Res}_{s=0}\,\zeta(1-s,\lfloor\operatorname{Re}\alpha\rfloor+1-\alpha)=-1.

However, this pole cancels with the term δm=0t0s/s\delta_{m=0}\,t_{0}^{s}/s in G1,m(s,α,t0)G_{1,m}(s,\alpha,t_{0}), giving us an entire function Gm(s,α,t0)G_{m}(s,\alpha,t_{0}). ∎

Now we are able to give the explicit formula for the LL-function. Choose any real positive number t0t_{0}. Suppose ff has poles α1,,αl\alpha_{1},\cdots,\alpha_{l} in t0{}\mathcal{F}_{t_{0}}-\{\infty\}. Put

f~(τ)=f(τ)Pα1(f)Pαl(f).\tilde{f}(\tau)=f(\tau)-P_{\alpha_{1}}(f)-\dots-P_{\alpha_{l}}(f).

We define also

I(f,s,t0)j=1lm=1ordfαj(2πi)m(m1)!cf,αj(m)Gm1(s,αj,t0).I(f,s,t_{0})\coloneqq\sum_{j=1}^{l}\sum_{m=1}^{\operatorname{ord}_{f}\alpha_{j}}\frac{(-2\pi i)^{m}}{(m-1)!}c_{f,\alpha_{j}}(m)G_{m-1}(s,\alpha_{j},t_{0}).
Theorem 8.5.

Let t0t_{0} be any real positive number. Let f𝒬kmero,pf\in\mathcal{Q}\mathcal{M}_{k}^{\text{mero},\,p} be a meromorphic quasi-modular form with prescribed poles and principle parts as above. Let f1,,fpf_{1},\cdots,f_{p} be the component functions of ff. Suppose that f~(τ)=na~f(n)qn\tilde{f}(\tau)=\sum_{n\gg-\infty}\tilde{a}_{f}(n)q^{n}, then we have

Λ(f,s)=\displaystyle\Lambda(f,s)= a~f(0)(t0ss+ikt0skks)+n0a~f(n)(Γ(s,2πnt0)(2πn)s+ikΓ(ks,2πn/t0)(2πn)ks)\displaystyle-\tilde{a}_{f}(0)\left(\frac{t_{0}^{s}}{s}+\frac{i^{k}t_{0}^{s-k}}{k-s}\right)+\sum_{n\neq 0}\tilde{a}_{f}(n)\left(\frac{\Gamma(s,2\pi nt_{0})}{(2\pi n)^{s}}+\frac{i^{k}\Gamma(k-s,2\pi n/t_{0})}{(2\pi n)^{k-s}}\right)
+I(f,s,t0)+ikI(f,ks,t01)+r=1p(ikra~fr(0)t0krskrs\displaystyle+I(f,s,t_{0})+i^{k}I(f,k-s,t_{0}^{-1})+\sum_{r=1}^{p}\left(-\frac{i^{k-r}\tilde{a}_{f_{r}}(0)t_{0}^{k-r-s}}{k-r-s}\right.
+n0ikra~fr(n)Γ(krs,2πnt0)(2πn)krs+ikrI(fr,krs,t01)).\displaystyle+\left.\sum_{n\neq 0}\frac{i^{k-r}\tilde{a}_{f_{r}}(n)\Gamma(k-r-s,\frac{2\pi n}{t_{0}})}{(2\pi n)^{k-r-s}}+i^{k-r}I(f_{r},k-r-s,t_{0}^{-1})\right).
Proof.

We divide Λ(f,s)\Lambda(f,s) into two parts:

Λ(f,s)=0t0,f(it)ts1𝑑t+t0,f(it)ts1𝑑t.\Lambda(f,s)=\int_{0}^{t_{0},\ast}f(it)t^{s-1}dt+\int_{t_{0}}^{\infty,\ast}f(it)t^{s-1}dt.

We first deal with the second part.

t0,f(it)ts1𝑑t=t0,f~(it)ts1𝑑t+j=1lt0,Pαj(f)(it)ts1𝑑t.\displaystyle\int_{t_{0}}^{\infty,\ast}f(it)t^{s-1}dt=\int_{t_{0}}^{\infty,\ast}\tilde{f}(it)t^{s-1}dt+\sum_{j=1}^{l}\int_{t_{0}}^{\infty,\ast}P_{\alpha_{j}}(f)(it)t^{s-1}dt.

Since f~\tilde{f} is holomorphic, by Lemma 8.2, we get

t0,f~(it)ts1𝑑t=a~f(0)t0ss+n0a~f(n)Γ(s,2πnt0)(2πn)s.\int_{t_{0}}^{\infty,\ast}\tilde{f}(it)t^{s-1}dt=-\frac{\tilde{a}_{f}(0)t_{0}^{s}}{s}+\sum_{n\neq 0}\frac{\tilde{a}_{f}(n)\Gamma(s,2\pi nt_{0})}{(2\pi n)^{s}}.

The integral of Pαj(f)P_{\alpha_{j}}(f) is shown in Lemma 8.4 which gives

t0,Pαj(f)(it)ts1𝑑t=m=1ordfαj(2πi)m(m1)!cf,αj(m)Gm1(s,αj,t0).\int_{t_{0}}^{\infty,\ast}P_{\alpha_{j}}(f)(it)t^{s-1}dt=\sum_{m=1}^{\operatorname{ord}_{f}\alpha_{j}}\frac{(-2\pi i)^{m}}{(m-1)!}c_{f,\alpha_{j}}(m)G_{m-1}(s,\alpha_{j},t_{0}).

For the first part, by changing the variable t1/tt\to 1/t, we get

0t0,f(it)ts1𝑑t=t01,f(i/t)ts1𝑑t.\int_{0}^{t_{0},\ast}f(it)t^{s-1}dt=\int_{t_{0}^{-1}}^{\infty,\ast}f(i/t)t^{-s-1}dt.

Since ff is a meromorphic quasi-modular form, we have the transformation

(21) f(i/t)=r=0pfr(it)(it)kr,f(i/t)=\sum_{r=0}^{p}f_{r}(it)(it)^{k-r},

by applying equation (2)\eqref{eq:slashf} with the inversion γ=(0110)\gamma=\begin{pmatrix}0&-1\\ 1&0\end{pmatrix}. So we have

0t0,f(it)ts1𝑑t=r=0pikrt01,fr(it)tkrs1𝑑t\displaystyle\int_{0}^{t_{0},\ast}f(it)t^{s-1}dt=\sum_{r=0}^{p}i^{k-r}\int_{t_{0}^{-1}}^{\infty,\ast}f_{r}(it)t^{k-r-s-1}dt
=r=0p(ikra~fr(0)t0krskrs+n0ikra~fr(n)Γ(krs,2πnt0)(2πn)krs\displaystyle\quad=\sum_{r=0}^{p}\left(-\frac{i^{k-r}\tilde{a}_{f_{r}}(0)t_{0}^{k-r-s}}{k-r-s}+\sum_{n\neq 0}\frac{i^{k-r}\tilde{a}_{f_{r}}(n)\Gamma(k-r-s,\frac{2\pi n}{t_{0}})}{(2\pi n)^{k-r-s}}\right.
+ikrI(fr,krs,t01)).\displaystyle\qquad+i^{k-r}I(f_{r},k-r-s,t_{0}^{-1})\Bigg{)}.

Finally, we complete the proof by noting that f0=ff_{0}=f. ∎

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 I(fr,s,t0)I(f_{r},s,t_{0}) is entire in ss. So we only need to prove the functional equations.

Lemma 2.2 shows that fmf_{m} is also a meromorphic quasi-modular form of weight k2mk-2m and depth pmp-m with components fm,(m+11)fm+1,,(ppm)fpf_{m},\binom{m+1}{1}f_{m+1},\cdots,\binom{p}{p-m}f_{p}. It is thus enough for us to show the functional equation of ff in (i)(i). The completed LL-function of ff is

(22) Λ(f,s)=t0,f(it)ts1𝑑t+t01,f(i/t)ts1𝑑t=t0,f(it)ts1𝑑t+t01,r=0pikrfr(it)tkrs1dt.\begin{split}&\Lambda(f,s)=\int_{t_{0}}^{\infty,\ast}f(it)t^{s-1}dt+\int_{t_{0}^{-1}}^{\infty,\ast}f(i/t)t^{-s-1}dt\\ &\quad=\int_{t_{0}}^{\infty,\ast}f(it)t^{s-1}dt+\int_{t_{0}^{-1}}^{\infty,\ast}\sum_{r=0}^{p}i^{k-r}f_{r}(it)t^{k-r-s-1}dt.\end{split}

Here we use the transformation formula (21)\eqref{eq:fit} again.

On the other hand, under the transform t1/tt\mapsto 1/t the first integral becomes

(23) t0,r=0pikrfr(i/t)tr+sk1dt=0t01,r=0pikrfr(it)tkrs1dt.\int_{t_{0}}^{\infty,\ast}\,\sum_{r=0}^{p}i^{k-r}f_{r}(i/t)t^{r+s-k-1}dt=\int_{0}^{t_{0}^{-1},\ast}\,\sum_{r=0}^{p}i^{k-r}f_{r}(it)t^{k-r-s-1}dt.

Combining equation (22)\eqref{eq:fms}, we get

Λ(f,s)=\displaystyle\Lambda(f,s)= (0t01,+t01,)r=0pikrfr(it)tkrs1dt\displaystyle\left(\int_{0}^{t_{0}^{-1},\ast}+\int_{t_{0}^{-1}}^{\infty,\ast}\right)\sum_{r=0}^{p}i^{k-r}f_{r}(it)t^{k-r-s-1}dt
=\displaystyle= r=0pikrΛ(fr,krs).\displaystyle\sum_{r=0}^{p}i^{k-r}\Lambda(f_{r},k-r-s).

This proves the functional equation of ff. ∎

Proposition 8.6.

Let f𝒬kmero,pf\in\mathcal{Q}\mathcal{M}_{k}^{\text{mero},\,p} be a meromorphic quasi-modular form. Then its Dirichlet LL-function L(f,s)L(f,s) is a meromorphic function for all ss\in\mathbb{C}. It has only possible simple poles at positive integers within kpskk-p\leq s\leq k.

Proof.

We note that Γ(s)\Gamma(s) has a simple pole at nonpositive integers, so this proposition follows directly from Theorem 1.3. ∎

The operator DD acts on the Fourier expansion of holomorphic quasi-modular form ff by

D=qddq:n0af(n)qnn0naf(n)qn.D=q\frac{d}{dq}:\sum_{n\geq 0}a_{f}(n)q^{n}\mapsto\sum_{n\geq 0}na_{f}(n)q^{n}.

This implies that the Dirichlet LL-series of DfDf is exactly the the shift of the original Dirichlet LL-series of ff. Actually, for meromorphic quasi-modular form, we can obtain the same result.

Theorem 8.7.

Let ff be a meromorphic quasi-modular form. Then we have

Λ(Dlf,s)=(sl)l(2π)lΛ(f,sl),\Lambda(D^{l}f,s)=\frac{(s-l)_{l}}{(2\pi)^{l}}\Lambda(f,s-l),

and

L(Dlf,s)=L(f,sl).L(D^{l}f,s)=L(f,s-l).
Proof.

The above identities are nothing but integration by parts. Evidently it is enough for us to prove the case l=1l=1. With integration by parts we have

t0(Df)(it)ewtts1𝑑t\displaystyle\int_{t_{0}}^{\infty}(Df)(it)e^{-wt}t^{s-1}dt
=\displaystyle= 12πf(it)ewtts1|t012πit0iwewtf(it)ts1+s1iewtf(it)ts2dt.\displaystyle-\frac{1}{2\pi}f(it)e^{-wt}t^{s-1}\big{|}_{t_{0}}^{\infty}-\frac{1}{2\pi i}\int_{t_{0}}^{\infty}iwe^{-wt}f(it)t^{s-1}+\frac{s-1}{i}e^{-wt}f(it)t^{s-2}dt.

When ww large enough, the first term equals to 12πf(it0)t0s1ewt0\frac{1}{2\pi}f(it_{0})t_{0}^{s-1}e^{-wt_{0}}. Clearly, it has a holomorphic continuation to the whole plane in ww and its value at w=0w=0 is just 12πf(it0)t0s1\frac{1}{2\pi}f(it_{0})t_{0}^{s-1}. For the integral, by Lemma 8.2, it has a holomorphic continuation to a neighbourhood of w=0w=0. So we get

t0,(Df)(it)ts1𝑑t=12πf(it0)t0s1+s12πt0,f(it)ts2𝑑t.\int_{t_{0}}^{\infty,\ast}(Df)(it)t^{s-1}dt=\frac{1}{2\pi}f(it_{0})t_{0}^{s-1}+\frac{s-1}{2\pi}\int_{t_{0}}^{\infty,\ast}f(it)t^{s-2}dt.

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 0 in the same way with integration by parts. At last we get

0,(Df)(it)ts1𝑑t=s12π0,f(it)ts2𝑑t.\int_{0}^{\infty,\ast}(Df)(it)t^{s-1}dt=\frac{s-1}{2\pi}\int_{0}^{\infty,\ast}f(it)t^{s-2}dt.

This gives the identity for Λ(f,s)\Lambda(f,s). The identity for L(f,s)L(f,s) then follows directly after Γ(z+1)=zΓ(z)\Gamma(z+1)=z\Gamma(z). ∎

If ff is a meromorphic modular form, the formula of its LL-function is much simpler.

Corollary 8.8.

Let fkmerof\in\mathcal{M}_{k}^{\text{mero}} be a meromorphic modular form of weight kk. Then the LL-function of ff is

Λ(f,s)=\displaystyle\Lambda(f,s)= a~f(0)(t0ss+ikt0skks)+I(f,s,t0)+ikI(f,ks,t01)\displaystyle-\tilde{a}_{f}(0)\left(\frac{t_{0}^{s}}{s}+\frac{i^{k}t_{0}^{s-k}}{k-s}\right)+I(f,s,t_{0})+i^{k}I(f,k-s,t_{0}^{-1})
+\displaystyle+ n0a~f(n)(Γ(s,2πnt0)(2πn)s+ikΓ(ks,2πn/t0)(2πn)ks)\displaystyle\sum_{n\neq 0}\tilde{a}_{f}(n)\left(\frac{\Gamma(s,2\pi nt_{0})}{(2\pi n)^{s}}+\frac{i^{k}\Gamma(k-s,2\pi n/t_{0})}{(2\pi n)^{k-s}}\right)

Moreover, it satisfies the following functional equation

Λ(f,s)=ikΛ(f,ks).\Lambda(f,s)=i^{k}\Lambda(f,k-s).
Remark.

In particular, when fSk!f\in S_{k}^{!} is a weakly holomorphic cusp form, we obtain

Λ(f,s)=n0af(n)(Γ(s,2πnt0)(2πn)s+ikΓ(ks,2πn/t0)(2πn)ks).\Lambda(f,s)=\sum_{n\neq 0}a_{f}(n)\left(\frac{\Gamma(s,2\pi nt_{0})}{(2\pi n)^{s}}+\frac{i^{k}\Gamma(k-s,2\pi n/t_{0})}{(2\pi n)^{k-s}}\right).

This computation coincides with Theorem 2.22.2 in [3].

9. Vanishing LL-values of meromorphic quasi-modular forms

In this section, we give some vanishing results of certain special LL-values of meromorphic quasi-modular forms.

Proposition 9.1.

Let f𝒬kmero,pf\in\mathcal{Q}\mathcal{M}_{k}^{\text{mero},\,p} be a meromorphic quasi-modular form of weight kk. Let ss be a negative integer, then

  1. (i)

    If k0k\leq 0, then the Dirichlet LL-function L(f,s)L(f,s) is always entire in ss. Moreover, when either s<kps<k-p or k<s<0k<s<0, we have

    L(f,s)=0.L(f,s)=0.
  2. (ii)

    If k2k\geq 2, when s<kps<k-p, we have

    L(f,s)=0.L(f,s)=0.
Proof.

If k0k\leq 0, the function Λ(f,s)\Lambda(f,s) has only possibly simple poles at kpsk0k-p\leq s\leq k\leq 0 by Theorem 1.3. Meanwhile, Γ(s)\Gamma(s) also has a simple pole at negative integers. So the Dirichlet LL-function L(f,s)L(f,s) is entire and L(f,s)=0L(f,s)=0 whenever s<kps<k-p or k<s<0k<s<0.

If k2k\geq 2, the function Λ(f,s)\Lambda(f,s) has no poles for s<kps<k-p but Γ(s)\Gamma(s) has pole at negative integer ss, so L(f,s)L(f,s) always vanishes if s<kps<k-p. ∎

Corollary 9.2.

Let f𝒬kmero,pf\in\mathcal{Q}\mathcal{M}_{k}^{\text{mero},\,p} be a meromorphic quasi-modular form of weight k0k\leq 0. If ss is a positive integer with 2s|k|2\leq s\leq|k|, then we have

L(D1kf,s)=0.L(D^{1-k}f,s)=0.
Proof.

By Theorem 8.7 and Proposition 9.1, when 2sk2\leq s\leq-k, we have

L(D1kf,s)=L(f,s+k1)=0.L(D^{1-k}f,s)=L(f,s+k-1)=0.\qed
Remark.

This corollary shows that some periods of the meromorphic quasi-modular form D1kfD^{1-k}f vanish. When ff is a weakly holomorphic modular form, this recovers Theorem 2.52.5 in [3].

Proposition 9.3.

The period of a meromorphic modular form of weight 22 is always vanishing, i.e. for any meromorphic modular form f2merof\in\mathcal{M}_{2}^{\text{mero}},

L(f,1)=Λ(f,1)=0.L(f,1)=\Lambda(f,1)=0.
Proof.

Applying the functional equation in Corollary 8.8 at weight 22, we get

Λ(f,1)=Λ(f,1).\Lambda(f,1)=-\Lambda(f,1).

So the period L(f,1)=Λ(f,1)=0L(f,1)=\Lambda(f,1)=0. ∎

Appendix A Special functions

A.1. Incomplete Gamma functions

We give some basic properties of incomplete gamma functions. Let s,zs,z\in\mathbb{C} with Re(s)>0\operatorname{Re}(s)>0 and |arg(z)|<π|\arg(z)|<\pi, the upper gamma function is defined by

Γ(s,z)=zts1et𝑑t,\Gamma(s,z)=\int_{z}^{\infty}t^{s-1}e^{-t}dt,

whereas the lower incomplete gamma function is defined by

γ(s,z)=0zts1et𝑑t.\gamma(s,z)=\int_{0}^{z}t^{s-1}e^{-t}dt.

The two functions are connected by

Γ(s)=Γ(s,z)+γ(s,z).\Gamma(s)=\Gamma(s,z)+\gamma(s,z).

Integration by parts yields the following recurrence relation for incomplete gamma function

(24) Γ(s+1,z)=sΓ(s,z)+zsez.\Gamma(s+1,z)=s\,\Gamma(s,z)+z^{s}\,e^{-z}.

Both incomplete gamma functions can be extended with respect to both ss and zz. The function Γ(s,z)\Gamma(s,z) extends to a multi-valued function in zz with a branch point at z=0z=0, and is holomorphic in each sector. If z0z\neq 0, then Γ(s,z)\Gamma(s,z) is always an entire function in ss.

For (s,z)2(s,z)\in\mathbb{C}^{2}, the incomplete gamma function has exponential growth asymptotic expansion

Γ(s,z)zs1ez(1+s1z+(s1)(s2)z2+)\Gamma(s,z)\sim z^{s-1}e^{-z}\left(1+\frac{s-1}{z}+\frac{(s-1)(s-2)}{z^{2}}+\dots\right)

as |z||z|\to\infty (see [18, Section 8.11 (i)]).

A.2. Hurwitz zeta function

Let s,as,a\in\mathbb{C} with Re(s)>1\operatorname{Re}(s)>1 and a0,1,2,a\neq 0,-1,-2,\cdots. The Hurwitz function ζ(s,a)\zeta(s,a) is defined by series expansion

ζ(s,a)=n=01(n+a)s.\zeta(s,a)=\sum_{n=0}^{\infty}\frac{1}{(n+a)^{s}}.

The function ζ(s,a)\zeta(s,a) has a meromorphic continuation to the whole ss-plane. It has only a simple pole at s=1s=1 with residue 11.

A.3. Polylogarithm

Let s,zs,z\in\mathbb{C} with |z|<1|z|<1. The polylogarithm function is defined to be

Lis(z)=k=1zkks.\operatorname{Li}_{s}(z)=\sum_{k=1}^{\infty}\frac{z^{k}}{k^{s}}.

It has an analytic continuation to a multi-valued holomorphic function in zz and an entire function in ss. If ss is not a nonnegative integer, then the principle branch Lis(z)\operatorname{Li}_{s}(z) is holomorphic in zz with branch cut along [1,)[1,\infty).

When nn\in\mathbb{N}, the function Lin(z)\operatorname{Li}_{-n}(z) is a rational function in zz and has pole only at z=1z=1. More precisely, we have

Lin(z)=(zz)nz1z.\operatorname{Li}_{-n}(z)=\left(z\frac{\partial}{\partial z}\right)^{n}\frac{z}{1-z}.

Let mm\in\mathbb{N} be a positive integer, then the Laurent series of Li1m(e(z))\operatorname{Li}_{1-m}(\textbf{e}(z)) at z=0z=0 is given by

(25) Li1m(e(z))=(m1)!(2πiz)m+k=0ζ(1mk)k!(2πiz)k.\operatorname{Li}_{1-m}(\textbf{e}(z))=\frac{(m-1)!}{(-2\pi iz)^{m}}+\sum_{k=0}^{\infty}\frac{\zeta(1-m-k)}{k!}(2\pi iz)^{k}.

The polylogarithm is related to the Hurwitz zeta function by

Lis(z)=Γ(1s)(2π)1s(i1sζ(1s,12+ln(z)2πi)+is1ζ(1s,12ln(z)2πi)).\operatorname{Li}_{s}(z)=\frac{\Gamma(1-s)}{(2\pi)^{1-s}}\left(i^{1-s}\zeta\left(1-s,\frac{1}{2}+\frac{\ln(-z)}{2\pi i}\right)+i^{s-1}\zeta\left(1-s,\frac{1}{2}-\frac{\ln(-z)}{2\pi i}\right)\right).

When z(0,1]z\notin(0,1], we have the reflection formula of polylogarithm

(26) Lis(z)+Lis(1/z)=(2πi)sΓ(s)ζ(1s,12+ln(z)2πi).\operatorname{Li}_{s}(z)+\operatorname{Li}_{s}(1/z)=\frac{(2\pi i)^{s}}{\Gamma(s)}\zeta\left(1-s,\,\frac{1}{2}+\frac{\ln(-z)}{2\pi i}\right).

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.