Home Mathematics Modular iterated integrals associated with cusp forms
Article Open Access

Modular iterated integrals associated with cusp forms

  • Nikolaos Diamantis ORCID logo EMAIL logo
Published/Copyright: December 1, 2021

Abstract

We construct an explicit family of modular iterated integrals which involves cusp forms. This leads to a new method of producing modular invariant functions based on iterated integrals of modular forms. The construction will be based on an extension of higher-order modular forms which, in contrast to the standard higher-order forms, applies to general Fuchsian groups of the first kind and, as such, is of independent interest.

MSC 2010: 11F37

1 Introduction

This paper deals with two classes of functions that have not been previously studied together, namely modular iterated integrals and higher-order modular forms. We show that they are interrelated in a way that key features of one of them can be elucidated through constructions in the other.

The first class of objects, modular iterated integrals, were introduced recently [2, 4, 3] by Brown in the context of the theory of real-analytic modular forms. Those are real-analytic functions f on the upper half-plane , characterized by the following (cf. Section 2.1 for the precise definition):

  1. A transformation law of the form

    f(γz)=(cz+d)r(cz¯+d)sf(z)for all z

    and for all γ=(**cd) in a suitable group Γ.

  2. An expansion of the form

    f(z)=|j|Myj(m,n0am,n(j)qmq¯n),

    where y=Im(z) and q=e2πiz.

The motivation for their introduction included their possible use towards arithmetic questions involving periods and evidence that the modular graph functions of String Theory are real-analytic modular forms.

A special subclass of the class of real-analytic modular forms consists of the spaces of modular iterated integrals of length . Their defining relation is

(1.1)+M[y]×-1and¯+M¯[y]×-1,

where M[y] (resp. M¯[y]) denotes polynomials in y=Im(z) with coefficients in the space of standard holomorphic (resp. anti-holomorphic) modular forms, and ,¯ are certain differential operators that will again be defined precisely in Section 2.1.

A reason for the special interest in this class is that the modular graph functions are expected to belong to it (cf. [2, Sections 1 and 9], where explicit evidence of this is provided). A second reason is that its structure seems to have arithmetic significance, as indicated by evidence provided by Brown (in [2, 4]) for the motivic nature of the space and by the association to it of classical number theoretic invariants, such as L-functions and period polynomials, by Drewitt and the author [10].

In this paper, we will address two questions that arise from the works mentioned above.

Question 1.

The elements of almost exclusively studied in the papers above are those that satisfy a condition more specific than (1.1), namely

+E[y]×-1and¯+E¯[y]×-1,

where E is the subspace of M generated by Eisenstein series.

What can be said about the remaining modular iterated integrals, i.e. the part originating in cusp forms? This is important not only because an answer describes more completely the structure of , but, especially, because arithmetic information is normally expected to be encapsulated by forms that are cuspidal. This is particularly relevant in view of the evidence for the arithmetic significance of mentioned above.

We will provide an answer to this question by constructing (in Section 2.2.2) an explicit family of such functions originating in cusp forms. We denote those functions by ϕh;r,s±. To this end, we will first restate (in Section 2.2.1) the question in a concrete and precise form, a task of independent interest.

Question 2.

A more explicit characterization of the space can be given in terms of Γ-invariant linear combinations of real and imaginary parts of iterated integrals of modular forms. This is proven in the case of elements of originating in Eisenstein series [4] and is conjectured to hold in general. Constructing such “modular invariant versions” of iterated integrals of modular forms is one of the important themes of [2], especially in relation to the applications to the theory of modular graph functions.

Here we discuss a new approach to this problem: We construct explicit “real-analytic iterated integrals”, denoted by ψh;r,s±, which naturally decompose into a modular invariant piece and a piece that can be thought of as known because it is expressed in terms of lower length modular iterated integrals. The modular invariant piece can immediately be read off of the formula for the function ϕh;r,s± that was the basis for the answer to Question 1 above. At the same time, this modular invariant piece belongs to (an extension of) . The construction is carried out, in the case of =2, in Section 3.2, where it is shown how ψh;r,s± yield polynomials whose coefficients are Γ-invariant elements of (an extension of) .

The tool with which we achieve both of those two aims is based on higher-order forms, the second object we deal with in this work.

The characterizing feature of higher-order modular forms, in the special case of order 2 and weight k, for instance, is the transformation law

f|kγδ-f|kγ-f|kδ+f=0for all γ,δΓ,

where the action of the group on the function is given by

g|kγ(z)=g(γz)(cz+d)-k.

The precise definition will be given in Section 3, where the original notion of higher-order forms will be generalized. Higher-order modular forms have been studied from various perspectives (analytic, adelic, algebraic, spectral etc.) and led to applications to modular symbols, mathematical physics etc. In all these cases, the theory had to be developed on congruence subgroups of level higher than 1, because such forms were parametrized by cusp forms of weight 2, which are trivial in level 1. This was a unnatural constraint because it excluded integrals of higher weight forms from consideration and it also prevented availing oneself of simplifications occurring in SL2().

In this paper, we resolve this problem too, by proposing a very general framework within which to consider higher-order forms. The class obtained includes several known and new objects, including higher-order forms for all levels and iterated integrals.

It turns out that the solution to this problem allows us to realize the constructions behind our answers to Questions 1 and 2 above. Firstly, it allows us to produce a family of (extended) modular iterated integrals of length 2 originating in cusp forms (Question 1). These modular iterated integrals, in turn, by their very construction, are obtained from a class of second-order modular forms whose prototypes are exactly the iterated integrals of cusp forms (Question 2). At the same time, these second-order modular forms are not of an ad hoc nature. They form a basis of the class of second-order modular forms they belong to (Theorem 3.8). This suggests a deeper relation between the two objects that are the subject of this paper.

2 Subclasses of the space of real-analytic modular forms

2.1 Review of definitions and notation

We start by introducing some of the notation we will be using and by recalling the definitions of real-analytic modular forms and of modular iterated integrals.

2.1.1 Basic spaces and actions

Let Γ=SL2() and set

S=(0-110),T=(1101),R=ST=(0-111).

If denotes the upper half-plane and z=x+iy, set

:={real analytic f:f(z)=O(yC) as y, uniformly in x for some C>0},
c:={ffor all c>0,f(z)=O(e-cy) as y, uniformly in x},
𝒪:={holomorphic f},
𝒪c:={holomorphic fc}.

For r,s, f and γΓ, define a function f|r,sγ by

f|r,sγ(z)=j(γ,z)-rj(γ,z¯)-sf(γz)for all z.

Here

j(γ,z)=cγz+dγ,where γ=(aγbγcγdγ).

We extend the action to [Γ] by linearity.

We can now give the definition of a real-analytic modular form.

We call an f (resp. fc) a real-analytic modular (resp. cusp) form of weights (r,s) for Γ if the following conditions hold:

  1. For all γΓ and z, we have f|r,sγ=f, i.e.

    f(γz)=j(γ,z)rj(γ,z¯)sf(z)for all z.
  2. For some M and am,n(j),

    (2.1)f(z)=|j|Myj(m,n0am,n(j)qmq¯n),q:=exp(2πiz).

We denote the space of real analytic modular (resp. cusp) forms of weights (r,s) for Γ by r,s (resp. 𝒮r,s). We set =r,sr,s (resp. 𝒮=r,s𝒮r,s.)

For s=0, upon restriction to holomorphic functions, we retrieve the space of standard holomorphic modular (resp. cusp) forms denoted by Mr (resp. Sr). We also set M=rMr (resp. S=rSr.)

2.1.2 Lie structure

The Lie algebra 𝔰𝔩2 acts on via the Maass operators r:r,sr+1,s-1 and ¯s:r,sr-1,s+1 given by

r=2iyz+rand¯s=-2iyz¯+s

(for proofs and further details on this and the rest of this subsection, see [2, Section 2.2]). These operators induce bigraded derivations on denoted by and ¯, respectively. We set

(m):=:r,sr+m,s-m,

with a similar definition for ¯(m).

Two pairs of identities we will be using often, taken from [2, Lemma 2.5 and (2.13)], are

(2.2)r(g|r,sγ)=(rg)|r+1,s-1γand¯r(g|r,sγ)=(¯rg)|r-1,s+1γ,
(2.3)r(ykg)=ykr+k(g)and¯s(ykg)=yk¯s+k(g).

To simplify notation, we will omit the index of r (resp. ¯s) when this is implied by the context.

2.1.3 Modular iterated integrals

We can now recall the definition of the space of modular iterated integrals of length for Γ=SL2(). Set recursively:

  1. -1=0.

  2. For each integer 0, we let be the largest subspace of r,s0r,s which satisfies

    +M[y]×-1,
    ¯+M¯[y]×-1,

    where M¯ is the ring of anti-holomorphic modular forms.

In [2, Lemma 3.10] it is proved that 0=[y-1] and, in [2, Corollary 4.4], the following statement is shown.

Proposition 2.1.

Let r,sZ be of the same parity and such that r+s>2. Let Er,s be the real-analytic Eisenstein series given, for zH, by

r,s(z)=γBΓ1j(γ,z)rj(γ,z¯)s

(where B={±TnnZ}). Then Er,sMr,s and

1=[y-1]r,s1,r+s4yr,s.

Note that our normalization of r,s is different from that of [2] and that the indices have been shifted relative to [2].

2.2 The space of extended modular iterated integrals

2.2.1 Motivating remarks and definition

With Proposition 2.1 and the definition of , we see that the space 2 is defined as the largest subspace of r,s0r,s which satisfies

(2.4)22+jr,s1,r+s4yjr,sM,
(2.5)¯22+jr,s1,r+s4yjr,sM¯.

Now, to the best of our knowledge, all explicit examples of elements of 2 considered in the literature correspond to elements of M in the right-hand side of (2.4) that are of a specific kind, namely (holomorphic) Eisenstein series. This, for instance, is the case for the important examples studied in [2, Section 9]. It is natural then to ask what is the nature of the functions corresponding to the remaining “cuspidal piece”. Specifically, we would like to investigate the largest subspace 𝒩 of satisfying inclusions of the form

𝒩𝒩+yjr,sSand¯𝒩𝒩+yjr,sS¯

for a suitable choice of indices. The motivation for that, apart from the general aim of understanding more fully the space 2, is to study modular iterated integrals whose L-functions are more likely to have classical arithmetic significance than those originating in Eisenstein series.

Before we state our definition, we note that, in contrast to the definition of (Section 2.1.3), it does not seem possible to define it so that the space is contained in the “first quadrant” r,s0r,s only.

We provide a heuristic argument why this should not be possible. Assume, for contradiction, that the “cuspidal part” is indeed restricted to the “first quadrant” and that fr,s belongs to the space. Since F:=(s)f satisfies s+rFr+s+1,-1, our assumption means that we should have

s+rF=yjm,lg,

where the sum ranges over a finite number of gS, j and m,l1 with m+l4. With (2.3), this gives 0(yr+sF)=yr+s+jm,lg, and hence, since g is cuspidal,

F=y-s-r2yj0(y+t)r+s+j-1m,l(z+it)g(z+it)𝑑t+y-r-sh(z)¯

for some holomorphic function h(z). Since, again according to our assumption, F belongs to a space contained in the “first quadrant”, ¯(r+1)F should belong to yjr,sS¯. On the other hand, the recursive relations of [2, Proposition 4.1], combined with (2.3), show that ¯(r+s+1)F is a linear combination of elements of the form

yj0tjm,l(z+it)g(z+it)𝑑tandy-r-s(r+s+1)h(z)¯,

where, for compactness of notation, we have taken 0,m to stand for y𝔾m+2 in accordance to [2, (4.1)]. Such linear combinations do not seem to belong to yjr,sS¯, but we have not been able to show that rigorously. The above argument therefore remains heuristic and only serves to suggest that it may not be possible to avoid a key difference between 2 and the following space (we will further comment on this difference after the statement of the definition).

Definition 2.2.

Let the space 2 of extended modular iterated integrals of length 2 be the largest subspace of which satisfies

(2.6)22+jr,s,r+s4yjr,sSand¯22+jr,s,r+s4yjr,sS¯.

There are two differences between 2 and 2: Firstly, compared with (2.4) and (2.5), the space M in the right-hand side has been replaced by S in (2.6). It is in this sense that the elements of the space 2 we will construct in the sequel are said to originate in cusp forms.

Secondly, the restriction of the space belonging to the “first quadrant” present in the definition of 2 is no longer required. A consequence of this is that arguments such as those leading to the full description of 1 in [2] are no longer possible because they rely on the finiteness of chains of functions generated by repeated applications of the operators and ¯. In the case of 1, the finiteness is guaranteed by the “first quadrant” condition. However, as suggested by the heuristics above, it seems unlikely that there exist a space with such a condition that, at the same time, satisfies the properties we want to encode in 2.

Finally, it should be stressed that we do not claim that 2 is the direct sum of 2 and the space of the functions corresponding to Eisenstein series previously studied in the literature.

2.2.2 An explicit sub-class of 2

We will now define an explicit family of elements of 2 that will give an answer to Question 1. In later sections, we will show that it originates in second-order modular forms. We will first introduce some preparatory constructions and results.

Let Pk-2 denote the space of polynomials in [X] of degree less than or equal to k-2, acted upon by |2-k,0. Denote the tensor product of the representations

(,|r,s)and(Pk-2,|2-k,0)

by |r,s,2-k. The group Γ then acts on Pk-2 as

(f|r,s,2-kγ)(z,X)=f(γz,γX)j(γ,z)-rj(γ,z¯)-sj(γ,X)k-2.

We use the same notation for the sub-representations corresponding to c,𝒪 and 𝒪c.

Let now

f(z)=n=1a(n)e2πinz

be a cusp form of weight k for Γ and consider its Eichler integrals

Ff+(z,X)=izf(w)(w-X)k-2𝑑wandFf-(z,X)=izf(w)(w-X)k-2𝑑w¯,

where the bar means complex conjugation (acting trivially on X).

Let r,s be integers of the same parity and such that r+s>k. We set

rf(γ;X):=rf+(γ;X)=γ-1iif(w)(w-X)k-2𝑑w,rf-(γ;X)=γ-1iif(w)(w-X)k-2𝑑w¯

and

(2.7)ϕr,s±(f;z,X):=γBΓFf±|r,s,2-kγ=γBΓFf±(γz,γX)j(γ,z)rj(γ,z¯)sj(γ,X)k-2.

We have the following proposition.

Proposition 2.3.

Suppose that r,sZ have the same parity and satisfy r+s>k. Then, for each zH, the series ϕr,s±(f;z,X) converges absolutely and it is invariant under the action of |r,s,2-k of Γ. Viewed as a polynomial in X, its coefficients are functions of (at most) polynomial growth at infinity.

Proof.

We show it for ϕ+, the proof for ϕ- being deduced upon conjugating ϕ+.

By the first equality of (2.7), ϕr,s+ is invariant under the action |r,s,2-k of Γ, for those r,s for which it converges.

To prove the statement about absolute convergence, we first note that the change of variables wγw, the transformation law of f and the identity

(2.8)(γz-γX)j(γ,z)j(γ,X)=z-X

imply that Ff+(γz,γX)j(γ,X)k-2 equals

(2.9)γ-1izf(w)(w-X)k-2𝑑w=rf(γ;X)+F+(z,X)
(2.10)=j=0k-2(-1)j(k-2j)γ-1iif(w)(w-γ-1)j𝑑w(X-γ-1)k-2-j+F+(z,X).

By applying this decomposition to the defining series for ϕr,s+, we get a sum of two terms.

To analyze the part corresponding to the first term of (2.10), we note that each of the integrals appearing in the sum is (up to a power of i) the value at s=l+1 of the “completed” L-function with additive twists. Specifically,

(2.11)Λf(s,pq):=0f(pq+ix)xs-1𝑑x=Γ(s)(2π)-sn=1a(n)e2πinp/qns.

It is well-known that Λf(s,p/q) has a functional equation (see [11] for a general version of the functional equation) and, with convexity, this implies that, for each j=0,,k-2,

(2.12)qj+1Λf(j+1,p/q)qk+12+ϵ.

Also,

(2.13)X-γ-1i=(X-z)+(z-g-1i)=(X-z)+j(γ,z)/cγ.

Upon applying the binomial formula to (2.13) and substituting into the polynomial

γBΓγ-1iif(w)(w-γ-1)j𝑑w(X-γ-1)k-2-jj(γ,z)rj(γ,z¯)s,

we get an expansion in (X-z)k-2-j-m (0mk-2-j). In this expansion, the coefficient of (X-z)k-2-j-m equals

(k-2-jm)ij+1γBΓΛf(j+1,γ-1)cγmj(γ,z)r-mj(γ,z¯)sγBΓcγk-12+ϵ-j-m|j(γ,z)|r+s-m.

For the last estimate we used (2.12). The elementary inequality |cγ||j(γ,z)|Im(z)-1 implies that the sum is

(2.14)y1-k2-ϵ+j+mγBΓ|j(γ,z)|k-12-r-s-j+ϵ,

which converges uniformly for z in a compact set since -k-12+r+s+j-ϵ>2.

The second term of (2.10) gives the polynomial

r,s(z)izf(w)(w-X)k-2𝑑w,

which, by Proposition 2.1, has real-analytic functions as coefficients since r+s>2.

Therefore, both pieces of ϕr,s+ induced by (2.10) will converge to a polynomial in X with coefficients that are real-analytic functions if -k-12+r+s+j-ϵ>2 for all j=0,,k-2 and r+s>2. This is indeed the case if r+s>k.

The bound (2.14) and the polynomial growth of Er,s(z) show that the coefficients of (X-z)j (and of Xj) in ϕr,s+ are of, at most, polynomial growth as y. ∎

The series ϕr,s±(f;z,X) can be decomposed in terms of elements of . Specifically, let

ϕr,s±(f;i,z),i=0,,k-2,

be functions such that

(2.15)ϕr,s±(f;z,X)=i=0k-2ϕr,s±(f;i,z)(X-z)i(X-z¯)k-2-i.

From [2, Proposition 7.1], we know that ϕr,s±(f;i,z) is |r+i,s+k-2-i-invariant. To show that it actually belongs to

r+i,s+k-2-i,

we need to show that it has an expansion of the form (2.1). This is part of the content of the next proposition.

Proposition 2.4.

Suppose r,s,k are as in Proposition 2.3. Then, for each j=0,,k-2, we have

ϕr,s+(f;j,z)=(-1)j(k-2j)y2-k(izf(w)(w-z¯)j(w-z)k-2-j𝑑w)r,s
(2.16)+m=0jn=0k-2-jαm,ny2-k1γBΓΛf(m+n+1,γ-1())cγm+n-k+2j(γ,z)r+j+n+2-kj(γ,z¯)s+m-j

and

ϕr,s-(f;j,z)=(-1)j(k-2j)y2-k(izf(w)(w-z¯)k-2-j(w-z)j𝑑w¯)r,s
(2.17)+m=0jn=0k-2-jαm,ny2-k1γBΓΛf(m+n+1,γ-1())¯cγm+n-k+2j(γ,z)2-k+r+m+jj(γ,z¯)s-j+n,

where

αm,n:=i1-2j-m-n(k-2j)(jm)(k-2-jn).

Further, each ϕr,s±(f;j,z) is in M.

Proof.

Replacing w-X in Ff+(z,X) according to the identity

w-X=((w-z)(X-z¯)+(z¯-w)(X-z))/Imz

and expanding with the binomial theorem, we see that the coefficients in the right-hand side of (2.15) can be written as

ϕr,s+(f;j,z)=(-1)j(k-2j)γBΓ((Imz)2-kizf(w)(w-z¯)j(w-z)k-2-j𝑑w)|r+j,k-2+s-j,0γ,

respectively

ϕr,s-(f;j,z)=(-1)j(k-2j)γBΓ((Imz)2-kizf(w)(w-z¯)k-2-j(w-z)j𝑑w¯)|r+j,k-2+s-j,0γ.

Upon unraveling the definition of the action |, the sum in ϕr,s+(f;j,z) equals

γBΓy2-kiγzf(w)(w-γz¯)jj(γ,z¯)j(w-γz)k-2-jj(γ,z)k-2-j𝑑wj(γ,z)rj(γ,z¯)s.

The change of variables wγw and the transformation law of f imply that the integral equals

(2.18)γ-1izf(w)(w-z¯)j(w-z)k-2-j𝑑w=(γ-1ii+iz)f(w)(w-z¯)j(w-z)k-2-jdw.

The first integral in the right-hand side is 0 for γ=1. For γ1, the binomial theorem leads to

m=0jn=0k-2-j(jm)(k-2-jn)j(γ,z¯)j-mj(γ,z)k-2-j-n(-cγ)k-2-m-nγ-1iif(w)(w-γ-1i)m+n𝑑w.

Equation (2.16) follows from this and (2.11) combined with (2.18). Equation (2.17) can be deduced from equation (2.16) upon a conjugation.

To show that ϕr,s+(f;j,z) has an expansion of the form (2.1), we apply the usual double coset decomposition to the series

1γBΓΛf(m,γ-1())cγnj(γ,z)pj(γ,z¯)t,where m,n,p,t are integers.

Then this becomes

c>0dmodclΛf(m,-dc)cn(c(z+l)+d)p(c(z¯+l)+d)t
(2.19)=c>0cn-p-tdmodcΛf(m,-dc)le2πil(x+dc)e-2πilt1dt1(t1+iy)p(t1-iy)t,

where we used the Poisson formula followed by a change of variables. With [12, Section 3.2, equation (12)], combined with [14, equation 13.14.9], the integral equals Pl(y)e-2π|l|y for some polynomial Pl(y) of degree less than or equal to |p-2| in y±. This implies that (2.19) can be written as

l0qlPl(y)c>0dmodcΛf(m,-dc)e2πildccp+t-n+l<0q¯-lPl(y)c>0dmodcΛf(m,-dc)e2πildccp+t-n.

We can apply this with m, n, p, t replaced by m+n+1, m+n-k+2,, r+j+n+2-k, s+m-j, respectively, because, by (2.12), the inner series converge with r+s>k. We deduce that ϕr,s+(f;j,z) has an expansion of the form (2.1). The analogous assertion for ϕr,s-(f;j,z) can be deduced from this after a conjugation. ∎

We can now verify an identity for ϕr,s±(f;z,X) that will allow us to show that ϕr,s±(f;j,z)2.

Proposition 2.5.

Let r,s be integers of the same parity and such that r+s>k. We have

r(ϕr,s+)=rϕr+1,s-1++2iyf(z)(X-z)k-2r,s𝑎𝑛𝑑¯s(ϕr,s+)=sϕr-1,s+1+,
¯s(ϕr,s-)=sϕr-1,s+1--2iyf(z)¯(X-z¯)k-2r,s𝑎𝑛𝑑r(ϕr,s-)=rϕr+1,s-1-.

Proof.

With (2.2), we have

(2.20)r(ϕr,s±)=γBΓr(Ff±)|r+1,s-1,2-kγ.

The definition of r and the identity Im(γz)=Im(z)/(j(γ,z)j(γ,z¯)) imply that, in the plus-case, this equals

rϕr+1,s-1++2iyγBΓf(γz)(γz-γX)k-2j(γ,X)k-2j(γ,z)rj(γ,z¯)s.

With the transformation law for f(z) and (2.8), this implies the statement in this case.

In the minus-case, (2.20) equals rϕr+1,s-1- as required. The identities involving ¯ are proved upon conjugating those we just proved. ∎

This proposition will enable us to define a subset of 2 which, in turn, will be the basis for the solution of one of our motivating problems. In fact, the equations proved in Proposition 2.5 and their difference from the equations of the analogous proposition [2, Proposition 4.1] help explain one of the differences between the definitions of 2 and 2. Specifically, in [2, Proposition 4.1], the application of to functions with indices (r,s)=(k-2,0) gives (up to a factor of y) only a classical Eisenstein series. By contrast, in our case, the outcome involves other ϕ-,-+ as well. Therefore, the “chain” obtained in this way cannot be finite.

We can now define our sub-class of 2 as the vector space 𝒜 generated over by ϕr,s±(f;i,-) for all fSk (k12), all integers r,s such that r+s is even and greater than k, and 0ik-2. With this definition, we can state our answer to Question 1 as follows.

Theorem 2.6.

The space A is a subspace of the space MI2 of extended modular iterated integrals of length 2.

Proof.

An elementary computation implies that for all real-analytic fj we have

(2.21)m(j=0k-2fj(z)(X-z)j(X-z¯)k-2-j)=j=0k-2(m+jfj(z)-(j+1)fj+1(z))(X-z)j(X-z¯)k-2-j,

where, for convenience, fk-1 is set to equal 0.

Proposition 2.5, combined with (2.21), implies that each rϕr,s±(f;i,-) is a linear combination of ϕr,s±(f;i,-) (for varying r,s,i) and an element of S[y]r,sr,s. Therefore, 𝒜 satisfies the first of the inclusions (2.6).

In the same way, we verify the analogous statement for ¯sϕr,s±(f;i,-) with S¯ in place of S.

Finally, by Proposition 2.4, 𝒜 is a subspace of . Since 2 is, by definition, the largest subspace of satisfying the inclusions (2.6), 𝒜 is contained in 2. ∎

3 The space of iterated invariants

To define our extended higher-order modular forms, it will be necessary to describe a general framework involving a family of representations; see [8, 9] for two alternative general definitions of higher-order objects, which are built on only one representation and which use the formalism of the augmentation ideal.

Exceptionally, we will give the next definition for general Fuchsian groups Γ of the first kind acting on with non-compact quotient Γ. The reason is that we want to compare it with the standard higher-order modular forms, as defined in [6], which are trivial in level 1.

Let V=(ρi,Vi)i0 be a sequence of representations of Γ, where the right-action on each Vi is denoted by “.”. Assume further that the -vector spaces Vi are finite-dimensional when i1. For each n, we consider the tensor representation i=0n-1Vi. To ease notation, we will generally denote the action on it also by “.”. It will generally be clear which representation it refers to in each case, but, in cases of potential ambiguity, it will be explained separately.

In the following definition, if V is a Γ-module, we view H0(Γ,V) as a subset of V.

Definition 3.1.

Set M(0):={0} and define, inductively, M(n)=M(n)(V) to be the subspace of i=0n-1Vi given by

M(n)=pr-1H0(Γ,(i=0n-1Vi)/(M(n-1)Vn-1)),

where the implied action is induced by that of Γ on i=0n-1Vi and pr is the canonical projection of i=0n-1Vi onto i=0n-1Vi/(M(n-1)Vn-1). We then set

Mc(n)=M(n)parabolic πH0(π,i=0n-1Vi),

where π is the subgroup generated by π.

We call the elements of M(n)(V)iterated invariants of order n.

In the next proposition, we show that this definition is well-founded and we give an equivalent formulation of it.

Proposition 3.2.

  1. For each n, M(n) and M(n-1)Vn-1are closed under the action of Γ.

  2. We have

    M(n-1)Vn-1M(n).

    Hence, composing with a map M(n-1)M(n-1)Vn-1, induced by vvv0 (for some v00), we have M(n-1)M(n) and Mc(n-1)Mc(n).

  3. The space Mc(n) is isomorphic to the space of fVi such that, for each γΓ,

    f.(γ-1)M(n-1)Vn-1

    and, for each parabolic πΓ,

    f.(π-1)=0.

Proof.

(i) We show this by induction over n. Let f be an element of M(n) (n1). Then, by definition, f.(ε-1)M(n-1)Vn-1 for each εΓ. Let γΓ. Then, for each δΓ,

(f.γ).(δ-1)=g.γfor g:=f.(γδγ-1-1)M(n-1)Vn-1.

Suppose that g=j=0kn-1fjvj for some fjM(n-1), where {vi}j=0kn-1 is a basis of Vn-1. Then

(3.1)(f.γ).(δ-1)=g.γ=j=0kn-1fj.γvj.γ,

which, by induction hypothesis, belongs to M(n-1)Vn-1. Therefore, f.γ belongs to M(n).

Because (3.1) holds for all gM(n-1)Vn-1, the Γ-invariance of M(n-1) implies the Γ-invariance of M(n-1)Vn-1.

(ii) We have M(n-1)Vn-1=pr-1({0})M(n).

(iii) This is seen by unraveling the definition, which, as shown in (i), is well-founded. ∎

A first result on the structure of the space of iterated invariants is provided by the following lemma. To state it, we introduce some additional notation for each Γ-module M:

C1(Γ,M)={α:ΓM},Cc1(Γ,M)={αC1(Γ,M):α(π)=0 for all parabolic πΓ},
Z1(Γ,M)={1-cocycles of Γ in M},Zc1(Γ,M)=Z1(Γ,M)Cc1(Γ,M),
B1(Γ,M)={1-coboundaries of Γ in M},Bc1(Γ,M)=B1(Γ,M)Cc1(Γ,M),
H1(Γ,M)=Z1(Γ,M)/B1(Γ,M),Hc1(Γ,M)=Zc1(Γ,M)/Bc1(Γ,M).

With this notation, we have the following lemma.

Lemma 3.3.

Let nN. There is a map ψ such that the following sequence is exact:

0H0(Γ,i=0n-1Vi)𝜄Mc(n)𝜓M(n-1)Cc1(Γ,Vn-1).

In particular, for n=2 we have the exact sequence

0H0(Γ,V0V1)𝜄Mc(2)𝜓H0(Γ,V0)Zc1(Γ,V1).

Proof.

Fix a basis {ui} of M(n-1). Then, for every fMc(n) and every γΓ, we have

f.(γ-1)=ψif(γ)ui

for some ψif(γ)Vn-1. By definition, each map γψif(γ) gives an element of Cc1(Γ,Vn-1). Therefore, the assignment fψifui induces the map ψ of the proposition.

For the case n=2, we note, with Proposition 3.2 (iii), that M(1)=Mc(1)=H0(Γ,V0). Therefore, the 1-cocycle condition satisfied by γf.(γ-1) is inherited by each ψifCc1(Γ,V1). ∎

Corollary 3.4.

Let ψ¯ be induced by ψ and the natural projection Zc1(Γ,V1)Hc1(Γ,V1). Then we have the following exact sequence:

0H0(Γ,V0V1)/(H0(Γ,V0)H0(Γ,V1))ι¯Mc(2)/(M(1)V1c)ψ¯H0(Γ,V0)Hc1(Γ,V1),

where ι¯ is induced by ι and V1c consists of vV1 invariant under all parabolic πΓ.

Proof.

This is deduced directly from Lemma 3.3. One can also use the long exact sequence associated with

0M(1)V1M(2)M(2)/(M(1)V1)0

to deduce the corollary. ∎

Since the maps ψ and ψ¯ constructed in Lemma 3.3 and Corollary 3.4 will play an important role in the sequel, we restate their definition separately: Let n and let {ui} be a basis of M(n-1). Then the map

ψ:Mc(n)M(n-1)Cc1(Γ,Vn-1)

is given, for each fMc(n), by

ψ(f)=ψifui,

where the ψifCc1(Γ,Vn-1) are such that

f.(γ-1)=ψif(γ)uifor all γΓ.

For n=2, let π denote the natural projection

M(1)Zc1(Γ,V1)M(1)Hc1(Γ,V1).

Then

ψ¯:=πψ.

3.1 Extended higher-order modular forms

For k0 and positive even integers k1,k2,, let

V=𝔒=(|2-ki,0,Vi)i0,

where V0=𝒪 and Vi=Pki-2[Xi] (i1) is the space of polynomials in Xi of degree less than or equal to ki-2. We call the elements of Mc(n)(𝔒)extended modular forms of order n. With Proposition 3.2 (iii), we see that this is the space of f(z;X1,,Xn-1)𝒪[X1,,Xn-1] such that

f.(γ-1){{0}(n=1),M(n-1)Pkn-1-2[Xn-1](n2),

and, for all parabolic πΓ,

f.π=f,

where the action of Γ is induced by

(3.2)(f.γ)(z;X1,,Xn-1):=f(γz;γX1,γXn-1)j(γ,z)k0-2j(γ,X1)k1-2j(γ,z)kn-1-2.

In particular,

(3.3)Mc(1)(𝔒)=M2-k0(Γ)={weight 2-k0 holomorphic modular forms for Γ}.

Let

V=𝔒c=(|2-ki,0,Vi)i0,

where V0=𝒪c and Vi=Pki-2[Xi]. Then we obtain the space Mc(n)(𝔒c) of extended cusp forms of order n.

Remark.

The adjective “extended” in the previous examples aims to distinguish them from the class of (standard) higher-order modular forms (see, e.g., [6]). We can retrieve the standard higher-order modular forms by setting k1==kn=2. Then “.” is simply

|2-k0,0for all n

and the space Mc(n)(𝔒) consists of all f𝒪 such that, for all γΓ0(N) and for all parabolic πΓ0(N),

f|2-k0,0(γ-1)Mc(n-1)(𝔒)andf|2-k0,0π=f.

The standard higher-order modular forms become trivial in Γ0(1) because, as was shown in [6], they are parametrized by weight 2 cusp forms which are trivial in SL2(). Finally, note that, in contrast to general iterated invariants, Mc(n-1)(𝔒) can be legitimately used instead of M(n-1)(𝔒) in the last displayed equation because of the identity (γ-1)(π-1)=(γπγ-1-1)γ-(π-1)).

3.1.1 Iterated Eichler integrals

Important examples and, indeed, some of the prototypes, of the standard higher-order forms mentioned in the closing remark of the last section are the antiderivatives of weight 2 cusp forms and their higher iterated analogues:

izf(w)𝑑w,izf(w1)iw1g(w2)𝑑w2𝑑w1  for weight 2 cusp forms f,g,.

In Lemma 3.5 we show that, more generally, Eichler integrals and their iterated counterparts belong to Mc(n)(𝔒). Its content is not essentially new (e.g., it is implied by the more general arguments of [1, Section 5.3]), but we revisit it from the perspective of higher-order forms in order to motivate the term “real-analytic iterated integrals” introduced in the next subsection.

Lemma 3.5.

Let k0=2 and k1,,kn-12N. Suppose that, for i=1,,n-1, fi is a weight ki cusp form for SL2(Z). Let FnO[X1,,Xn-1] be defined by F1=1 and, for n2,

Fn(w;X1,,Xn-1):=iwf1(w1)(w1-X1)k1-2iw1f2(w2)(w2-X2)k2-2𝑑wn-1𝑑w1.

Then FnMc(n)(O).

Remark.

Though not “visible” in the statement, k0 is necessary for the definition of Mc(n)(𝔒). It reflects the fact that, in terms of the variable w, Γ acts like a regular representation.

Proof.

We first show the assertion for n=2.

The action “.” is given explicitly by (3.2). We then see that F2.γ is

iγwf1(w1)(w1-γX1)k1-2j(γ,X1)k1-2𝑑w1=γ-1iwf1(w1)(w1-X1)k1-2𝑑w1,

where the last integral is obtained by a change of variables and (2.8). Therefore, with (3.3),

F2.(γ-1)=γ-1iif1(w1)(w1-X1)k1-2𝑑w1Pk1-2[X1]M(1)(𝔒)Pk1-2[X1].

The same identity shows that F2.T=F2, and hence F2Mc(2)(𝔒).

The proof for general n is an application of the shuffle product formula for iterated integrals, but we give a direct proof by induction. As before, by the definition of the action “.” on 𝒪[X1,,Xn-1], the changes of variables wiγwi and (2.8), we deduce that, for n>2, Fn.γ equals

γ-1iwf1(w1)(w1-X1)k1-2(γ-1iw1f2(w2)(w2-X2)k2-2𝑑w2)𝑑w1
=(iw+γ-1ii)f1(w1)(w1-X1)k1-2((iw1+γ-1ii)f2(w2)(w2-X2)k2-2dw2)dw1,

where the sum of integral signs indicates that they are both applied to the integrand following them. Therefore, Fn.(γ-1) is a sum of iterated integrals such that each iterated integral includes at least one constituent integral with limits i and γ-1i, e.g.,

iwf1(w1)(w1-X1)k1-2γ-1iif2(w2)(w2-X2)k2-2iw2f3(w3)(w3-X3)k3-2𝑑w3𝑑w2𝑑w1.

This, on the one hand, implies that Fn.(T-1)=0 and, on the other, that Fn.(γ-1) is a sum of products of the form Fi-1Pi-1(Xi-1,,Xn-1) (i=2,n), where the polynomials Pi-1 are of degree less than or equal to kj-2 in Xj (j=i-1,,n-1) and independent of w. By induction, each product

Fi-1Pi-1(Xi-1,,Xn-1)

belongs to

M(i-1)(𝔒)Pki-1[Xi]Pkn-1-2[Xn-1]M(n-1)(𝔒)Pkn-1-2[Xn-1].

This completes the proof of the statement. ∎

Upon tensoring with a space of cusp forms, we obtain the following corollary.

Corollary 3.6.

Let k0,,kn-12Z. Suppose that, for i=1,,n-1, fi is a weight ki cusp form for SL2(Z) and that f0 is a cusp form of weight 2-k0. Let FnO[X1,,Xn-1] be defined by

f0(w)iwf1(w1)(w1-X1)k1-2iw1f2(w2)(w2-X2)k2-2𝑑wn-1𝑑w1.

Then FnMc(n)(Oc).

3.2 The space of real-analytic iterated integrals

As mentioned in Section 1, modular iterated integrals are conjectured to be given in terms of Γ-invariant linear combinations of iterated integrals of modular forms (see [7] for a somewhat analogous phenomenon, whereby linear combinations of multiple Hecke L-values are expressed in terms of usual Hecke L-series). In view of this conjectured connection, we discuss here a way to obtain invariant objects from iterated integrals associated with modular forms. To this end, we first associate iterated invariants to real analytic modular forms.

Let V=(Vi,ρi)i0, where V0= (resp. V0=c) with Γ=SL2() acting through |r,s, and Vi=Pki-2[Xi] (i1) with the usual action |2-ki,0 on polynomials. Then we have that, for i>0,

(3.4)Vic=H0(Γ,Vi)=H0(Γ,Pki-2)=0if ki>2 (by translation invariance),
(3.5)Hc1(Γ,Vi)SkiS¯ki(by Eichler–Shimura combined with [16, Lemma 1 of VI, Section 5]).

Notice that, although H0(Γ,) is very similar to r,s, they are not the same, because the former includes functions that do not have an expansion of the form (2.1).

We also consider the one-dimensional subspace ~r,s of r,sH0(Γ,) generated by r,s. With the map ψ¯ defined in Corollary 3.4, we set

M~c(2)():=ψ¯-1(~r,sHc1(Γ,V1)).

Explicit examples of elements of M~c(2)() are certain real-analytic analogues of iterated Eichler integrals in the case of r,s. We will make this more specific in the case n=2 with the following corollary of Lemma 3.5.

Corollary 3.7.

Suppose that k12N and that f1 is a weight k1 cusp form for SL2(Z). Let F2 be defined by

F2(w,X1):=r,s(w)iwf1(w1)(w1-X1)k1-2𝑑w1.

Then F2M~c(2)(R).

Proof.

This follows immediately from Lemma 3.5 combined with the identity F2.(γ-1)=r,srf(γ;X1) and the definition of ψ¯. ∎

The functions of this corollary have been the prototypes for the elements of M~c(2)(). At the same time, these functions are, up to multiplication with real-analytic r,s, special cases of the iterated integrals of Lemma 3.5. Therefore, we refer to M~c(2)() as the space of real-analytic iterated integrals, even though its elements are not necessarily representable by iterated integrals in the strict sense.

Now, with the definition of M~c(2)() and with (3.4) and (3.5), Corollary 3.4 becomes

0H0(Γ,Pk1-2[X1])ι¯M~c(2)()ψ¯~r,s(Sk1S¯k1).

We will show that this can be completed to a right exact sequence.

Theorem 3.8.

Suppose that r+s>k1. The sequence of maps

0H0(Γ,Pk1-2[X1])ι¯M~c(2)()ψ¯~r,s(Sk1S¯k1)0

is exact.

Proof.

The only part remaining to be proved is the surjectivity of ψ¯. Let r,s(f,g¯) be an arbitrary basis element of ~r,s(Sk1S¯k1). With the notation of Section 2.2.2, assign to each hSk1 a function ψh;r,s± given by

ψh;r,s±(z,X):=ϕr,s±(h;z,X)-Fh±(z,X)r,s=γBΓrh±(γ;X)j(γ,z)rj(γ,z¯)s.

By Proposition 2.3, this is absolutely convergent and its coefficients are of polynomial growth at infinity, and thus they belong to .

The image of ψf;r,s++ψg;r,s- under ψ is induced by the mapping

γ(ψf;r,s++ψg;r,s-)|r,s,2-k1(γ-1)=-(Ff++Fg-)|0,0,2-k1(γ-1)r,s(z)
=-(rf(γ;X)+rg(γ;X)¯)r,s(z).

For the two equalities, we have used Proposition 2.3 and (2.9). By the explicit formula for the Eichler–Shimura map, we deduce that

ψ¯(-ψf;r,s+-ψg;r,s-)=r,s(f,g¯).

This shows that -ψf;r,s+-ψg;r,s-M~c(2)() and that its image is the element r,s(f,g¯). ∎

Remark.

The theorem could be stated in more general form so that the real-analytic analogues of both Eisenstein and Poincaré series are captured. That would have the advantage of accounting for the full space Mc(2) instead of M~c(2), but we would need to enlarge our investigations to objects that do not satisfy (2.1). This is because the “real-analytic Poincaré series” do not satisfy (2.1). However, they are clearly interesting objects, worthwhile studying, and they are the subject of work in progress with F. Strömberg.

The family {ψh;r,s±} constructed in Theorem 3.8 allows us to describe our approach to Question 2 of Section 1. Specifically, for hSk1, set

ψh;r,s±(z,X):=γBΓrh±(γ;X)j(γ,z)rj(γ,z¯)s.

The family addresses Question 2, inasmuch as it satisfies the following three properties:

  1. Firstly, by Theorem 3.8, the ψh;r,s± belong to the space M~c(2)() of real-analytic iterated integrals.

  2. Secondly, this family is “canonical” in the sense that it induces a generating set for ~r,s(Sk1S¯k1).

  3. Thirdly, it is possible to obtain, by a simple process, explicit Γ-equivariant versions of the real-analytic iterated integrals ψh;r,s±.

This process is given in the following proposition, which also formalizes a link between the two main themes of this note, namely second-order modular forms and iterated integrals.

Proposition 3.9.

Let r+s>k1. There is a well-defined linear map from the subspace of Mc(2)(R) generated by the family {ψh;r,s±} to i=0k1-2MI2.

Proof.

For each ψh;r,s±(z,X), consider

ϕr,s±(h;z,X)=ψh;r,s±(z,X)+Fh±(z,X)r,s.

By Theorem 2.6, the coefficients ϕr,s±(h;i,z,X) of (X-z)i(X-z¯)k1-2-i in ϕr,s±(h;z,X) belong to 2. Therefore, the assignment

ψh;r,s±(z,X)(ϕr,s±(h;0,z,X),,ϕr,s±(h;k-2,z,X))

defines the sought map. ∎

3.2.1 A classification of holomorphic second-order forms

We close with another implication of Corollary 3.4 to holomorphic extended second-order modular forms. The resulting theorem is the analogue of [6, Theorem 2.3] to the class of extended second-order modular forms.

Let ρ be a complex representation of Γ of dimension k1-1 induced by

ρ(-I2)=Ik1-1,ρ(S)=((-1)iδj,k1-2-i)i,j=0k1-2,ρ(T)=((-1)i+j(ji))i,j=0k1-2.

For an even k>0, we denote by Mk(ρ) the space of vector-valued modular forms for the representation ρ.

Proposition 3.10.

Suppose that k>k1>2. The sequence of maps

0Mk(ρ)Mc(2)(𝔒)ψ¯Mk(Sk1S¯k1)0

is exact. In particular,

dimMc(2)(𝔒)=2dim(Mk)dim(Sk1)+5+k12(k1-1)+ik+k1-24-13(k1-13)(k-13).

Proof.

We apply Corollary 3.4 to V=𝔒 with k0:=2-k to deduce, in the first instance,

0H0(Γ,𝒪Pk1-2)ι¯Mc(2)(𝔒)ψ¯Mk(Sk1Sk1).

Here we used (3.4) and (3.5).

The isomorphism H0(Γ,𝒪Pk1-2)Mk(ρ) is induced by the mapping

g=j=0k1-2fj(z)Xj(f0,f1,,fk1-2)T.

By a direct computation, we see that the invariance of g under S and T is equivalent to the invariance of the associated vector under S and T in terms of the representation ρ.

To prove the surjectivity, let

f(g,h¯)Mk(Sk1S¯k1).

Assume that f=n0λnPn, where Pn is the classical Poincaré series of weight k given by

Pn(z)=BΓe2πinγzj(γ,z)k

(here λn=0 for all but finitely many integer n0). For n0 and hSk1, set

Gn,h±(z,X):=γBΓ(Ff±e2πin-)|k,0,2-k1γ-Fh±(z,X)Pn=γBΓrh±(γ;X)e2πinγzj(γ,z)k.

Since |e2πinγz|1, the absolute convergence of this series, for k>k1, follows from the absolute convergence of ϕk,0± proved in Proposition 2.3. The same proposition implies the polynomial growth of the polynomial coefficients of Gn,h±.

We then have, for all γΓ,

-n0λn(Gn,g++Gn,h-)|k,0,2-k1(γ-1)=n0λn(rg+(γ;X)+rh-(γ;X))Pn=(rg+(γ;X)+rh-(γ;X))f.

By the definition of the map ψ¯, we deduce that the image of λn(Gn,g++Gn,h-) is f(g,h¯).

To deduce the dimension formula, we use [5, Theorem 6.3 and Remark 6.4] to compute the dimension of Mk(ρ). Indeed, ρ is an even representation, dimρ=k1-1 and a direct computation gives

Tr(ρ(ST)2)=Tr(ρ(ST))=Tr(ρ(S)ρ(T))=i=0k1-2(-1)i(ik1-2-i)(k1-13),

where () stands for the Legendre symbol. This is seen by noting that the sequence given by

an=i,j,i+j=n(-1)i(ij),where (ij)=0, unless ij0,

satisfies the recurrence relation an+an-1+an-2=0, and therefore, by induction, an=n+13

We can also see, by induction, that, for ξ=eπi/3,

ξk1-ξ2+x2k1-ξ-2=-(k-13).

Since the only eigenvalue of ρ(T) is 1 and the corresponding eigenspace has dimension 1, there is one Jordan block with 1 in the diagonal. By [5, Theorem 3.4], we deduce that the trace for a standard choice of exponents for ρ(T) is 0. This, together with [5, Remark 6.3], implies that, since k>k1, the dimension of Mk(ρ) is given by the formula of [5, Theorem 6.3], which, by the preceding remarks, is

5+k12(k1-1)+ik+k1-24-13(k1-13)(k-13).

Communicated by Jan Bruinier


Award Identifier / Grant number: EP/S032460/1

Funding statement: Research on this work was supported in part by EPSRC grant EP/S032460/1.

Acknowledgements

The author is grateful to F. Brown, L. Candelori, C. Franc, G. Mason and F. Strömberg for many helpful comments and suggestions. He also thanks the anonymous referee for a very careful reading of the manuscript and many useful suggestions.

References

[1] F. Brown, Multiple modular values and the relative completion of the fundamental group of M1,1, preprint (2014), https://arxiv.org/abs/1407.5167. Search in Google Scholar

[2] F. Brown, A class of non-holomorphic modular forms I, Res. Math. Sci. 5 (2018), no. 1, Paper No. 7. 10.1007/s40687-018-0130-8Search in Google Scholar

[3] F. Brown, A class of non-holomorphic modular forms III: Real analytic cusp forms for SL2(), Res. Math. Sci. 5 (2018), no. 3, Paper No. 34. 10.1007/s40687-018-0151-3Search in Google Scholar

[4] F. Brown, A class of nonholomorphic modular forms II: Equivariant iterated Eisenstein integrals, Forum Math. Sigma 8 (2020), Paper No. e31. 10.1017/fms.2020.24Search in Google Scholar

[5] L. Candelori and C. Franc, Vector-valued modular forms and the modular orbifold of elliptic curves, Int. J. Number Theory 13 (2017), no. 1, 39–63. 10.1142/S179304211750004XSearch in Google Scholar

[6] G. Chinta, N. Diamantis and C. O’Sullivan, Second order modular forms, Acta Arith. 103 (2002), no. 3, 209–223. 10.4064/aa103-3-2Search in Google Scholar

[7] Y. Choie, Parabolic cohomology and multiple Hecke L-values, Ramanujan J. 41 (2016), no. 1–3, 543–561. 10.1007/s11139-016-9844-7Search in Google Scholar

[8] A. Deitmar, Higher order group cohomology and the Eichler-Shimura map, J. Reine Angew. Math. 629 (2009), 221–235. 10.1515/CRELLE.2009.032Search in Google Scholar

[9] A. Deitmar, Invariants, cohomology, and automorphic forms of higher order, Selecta Math. (N. S.) 18 (2012), no. 4, 855–883. 10.1007/s00029-012-0087-1Search in Google Scholar

[10] N. Diamantis and J. Drewitt, Period functions associated to real-analytic modular forms, Res. Math. Sci. 7 (2020), no. 3, Paper No. 21. 10.1007/s40687-020-00221-8Search in Google Scholar

[11] N. Diamantis, J. Hoffstein, E. M. Kıral and M. Lee, Additive twists and a conjecture by Mazur, Rubin and Stein, J. Number Theory 209 (2020), 1–36. 10.1016/j.jnt.2019.11.016Search in Google Scholar

[12] A. Erdélyi, W. Magnus, F. Oberhettinger and F. G. Tricomi, Tables of Integral Transforms. Vol. I, McGraw-Hill, New York, 1954. Search in Google Scholar

[13] S. Lang, Introduction to Modular Forms, Grundlehren Math. Wiss. 222, Springer, Berlin, 1976. 10.1007/978-3-642-51447-0_1Search in Google Scholar

[14] F. Olver, D. Lozier, R. Boisvert and C. Clark, NIST Handbook of Mathematical Functions, Cambridge University, Cambridge, 2010. Search in Google Scholar

Received: 2021-08-31
Revised: 2021-10-01
Published Online: 2021-12-01
Published in Print: 2022-01-01

© 2022 Walter de Gruyter GmbH, Berlin/Boston

This work is licensed under the Creative Commons Attribution 4.0 International License.

Downloaded on 11.1.2026 from https://www.degruyterbrill.com/document/doi/10.1515/forum-2021-0224/html?lang=en
Scroll to top button