1,639
Views
11
CrossRef citations to date
0
Altmetric
Original Articles

Kronecker’s Limit Formula, Holomorphic Modular Functions, and q-Expansions on Certain Arithmetic Groups

, &

ABSTRACT

For any square-free integer N such that the “moonshine group” Γ0(N)+ has genus zero, the Monstrous Moonshine Conjectures relate the Hauptmodul of Γ0(N)+ to certain McKay–Thompson series associated to the representation theory of the Fischer–Griess monster group. In particular, the Hauptmoduli admits a q-expansion which has integer coefficients. In this article, we study the holomorphic function theory associated to higher genus groups Γ0(N)+. For all such arithmetic groups of genus up to and including three, we prove that the corresponding function field admits two generators whose q-expansions have integer coefficients, has lead coefficient equal to one, and has minimal order of pole at infinity. As corollary, we derive a polynomial relation which defines the underlying projective curve, and we deduce whether i∞ is a Weierstrass point. Our method of proof is based on modular forms and includes extensive computer assistance, which, at times, applied Gauss elimination to matrices with thousands of entries, each one of which was a rational number whose numerator and denominator were thousands of digits in length.

2010 AMS Subject Classification:

1. Introduction

1.1. Classical aspects of the j-Invariant

The action of the discrete group on the hyperbolic upper half plane yields a quotient space which has genus zero and one cusp. By identifying with a well-known fundamental domain for the action of on , we will, as is conventional, identify the cusp of with the point denoted by i∞. From the uniformization theorem, we have the existence of a single-valued meromorphic function on which has a simple pole at i∞ and which maps the one-point compactification of onto the one-dimensional projective space . Let z denote the global coordinate on , and set q = eiz, which is a local coordinate in a neighborhood of i∞ in the compactification of . The bi-holomorphic map f from the compactification of onto is uniquely determined by specifying constants c−1 ≠ 0 and c0 such that the local expansion of f near i∞ is of the form f(q) = c−1q−1 + c0 + O(q) as q → 0. For reasons coming from the theory of automorphic forms, one chooses c−1 = 1 and c0 = 744. The unique function obtained by setting of c−1 = 1 and c0 = 744 is known as the j-invariant, which we denote by j(z).

Let One defines the holomorphic Eisenstein series Ek of even weight k ≥ 4 by the series: (1-1) Classically, it is known that (1-2) which yields an important explicit expression for the j-invariant. As noted on page 90 of [CitationSerre 73], which is a point we will emphasize in the next subsection, the j-invariant admits the series expansion: (1-3) An explicit evaluation of the coefficients in the expansion (Equation1–3) was established by Rademacher in [CitationRademacher 38], and we refer the reader to the fascinating article [CitationKnopp 90] for an excellent exposition on the history of the j-invariant in the setting of automorphic forms. More recently, there has been some attention on the computational aspects of the Fourier coefficients in (Equation1–3); see, for example, [CitationBaier and Köhler 03] and, for the sake of completeness, we mention the monumental work in [CitationEdixhoven and Couveignes 11] with its far reaching vision.

For quite some time it has been known that the j-invariant has importance far beyond the setting of automorphic forms and the uniformization theorem as applied to . For example, in 1937 T. Schneider initiated the study of the transcendence properties of j(z). Specifically, Schneider proved that if z is a quadratic irrational number in the upper half plane then j(z) is an algebraic integer, and if z is an algebraic number but not imaginary quadratic then j(z) is transcendental. More specifically, if z is any element of an imaginary quadratic extension of , then j(z) is an algebraic integer, and the field extension is abelian. In other words, special values of the j-invariant provide the beginning of explicit class field theory. The seminal work of Gross–Zagier [CitationGross and Zagier 85] studies the factorization of the j-function evaluated at imaginary quadratic integers, yielding numbers known as singular moduli, from which we have an abundance of current research which reaches in various directions of algebraic and arithmetic number theory.

There is so much richness in the arithmetic properties of the j-invariant that we are unable to provide an exhaustive list, but rather ask the reader to accept the above-mentioned examples as indicating the important role played by the j-invariant in number theory and algebraic geometry.

1.2. Monstrous moonshine

The j-invariant attained another realm of importance with the discovery of “Monstrous Moonshine.” We refer to the article [CitationGannon 06a] for a fascinating survey of the history of monstrous moonshine, as well as the monograph [CitationGannon 06b] for a thorough account of the underlying mathematics and physics surrounding the moonshine conjectures. At this point, we offer a cursory overview in order to provide motivation for this article.

In the mid-1900s, there was considerable research focused toward the completion of the list of sporadic finite simple groups. In 1973, B. Fischer and R. L. Griess discovered independently, but simultaneously, certain evidence suggesting the existence of the largest of all sporadic simple groups; the group itself was first constructed by R. L. Griess in [CitationGriess 82] and is now known as “the monster,” or “the friendly giant,” or “the Fischer–Griess monster,” and we denote the group by . Prior to [CitationGriess 82], certain properties of were deduced assuming its existence, such as its order and various aspects of its character table. At that time, there were two very striking observations. On one hand, A. Ogg observed that the set of primes which appear in the factorization of the order of is the same set of primes such that the discrete group Γ0(p)+ has genus zero. On the other hand, J. McKay observed that the linear-term coefficient in (Equation1–3) is the sum of the two smallest irreducible character degrees of . J. Thompson further investigated McKay’s observation in [CitationThompson 79a] and [CitationThompson 79b], which led to the conjectures asserting all coefficients in the expansion (Equation1–3) are related to the dimensions of the components of a graded module admitting action by . Building on this work, J. Conway and S. Norton established the “monstrous moonshine” conjectures in [CitationConway and Norton 79] which more precisely formulated relations between and the j-invariant (Equation1–3). A graded representation for was explicitly constructed by I. Frenkel, J. Lepowsky, and A. Meurman in [CitationFrenkel et al. 88] thus proving aspects of the McKay–Thompson conjectures from [CitationThompson 79a] and [CitationThompson 79b]. Building on this work, R. Borcherds proved a significant portion of the Conway–Norton “monstrous moonshine” conjectures in his celebrated work [CitationBorcherds 92]. More recently, additional work by many authors (too numerous to list here) has extended “moonshine” to other simple groups and other j-invariants associated to certain genus zero Fuchsian groups.

Still, there is a considerable amount yet to be understood within the framework of “monstrous moonshine.” Specifically, we call attention to the following statement by T. Gannon from [CitationGannon 06a]:

In genus >0, two functions are needed to generate the function field. A complication facing the development of a higher-genus Moonshine is that, unlike the situation in genus 0 considered here, there is no canonical choice for these generators.

In other words, one does not know the analogue of the j-invariant for moonshine groups of genus greater than zero from which one can begin the quest for “higher genus moonshine.”

1.3. Our main result

The motivation behind this article is to address the above statement by T. Gannon. The methodology we developed yields the following result, which is the main theorem of the present paper.

Theorem 1.

Let N be a square-free integer such that the genus g of Γ0(N)+ satisfies the bounds 1 ≤ g ≤ 3. Then the function field associated to Γ0(N)+ admits two generators whose q-expansions have integer coefficients after the lead coefficient has been normalized to equal one. Moreover, the orders of poles of the generators at i∞ are at most g + 2.

In all cases, the generators we compute have the minimal poles possible, as can be shown by the Weierstrass gap theorem. Finally, as an indication of the explicit nature of our results, we compute a polynomial relation associated to the underlying projective curve. For all groups Γ0(N)+ of genus two and genus three, we deduce whether i∞ is a Weierstrass point.

In brief, our analysis involves four steps. First, we establish an integrality theorem which proves that if a holomorphic modular function f on Γ0(N)+ admits a q-expansion of the form f(z) = qa + ∑k > −ackqk, then there is an explicitly computable κ such that if for k ≤ κ then for all k. Second, we prove the analogue of Kronecker’s limit formula, resulting in the construction of a non-vanishing holomorphic modular form ΔN on Γ0(N)+; we refer to ΔN as the Kronecker limit function. Third, we construct spaces of holomorphic modular functions on Γ0(N)+ by taking ratios of holomorphic modular forms whose numerators are holomorphic Eisenstein series on Γ0(N)+ and whose denominator is a power of the Kronecker limit function. Finally, we employ considerable computer assistance in order to implement an algorithm, based on the Weierstrass gap theorem and Gauss elimination, to derive generators of the function fields from our spaces of holomorphic functions. By computing the q-expansions of the generators out to order qκ, the integrality theorem from the first step completes our main theorem.

1.4. Additional aspects of the main theorem

There are ways in which one can construct generators of the function fields associated to the groups Γ0(N)+, two of which we now describe. However, the methodologies do not provide all the information which we developed in the proof of our main theorem.

Classically, one can use Galois theory and elementary aspects of Hecke congruence groups Γ0(N) in order to form modular functions using the j-invariant. In particular, the functions generate all holomorphic functions which are Γ0(N)+ invariant and, from (Equation1–3), admit q-expansions with lead coefficients equal to one and all other coefficients are integers. However, the order of poles at i∞ are much larger than g + 2.

From modern arithmetic algebraic geometry, we have another approach toward our main result. We shall first present the argument, which was first given to us by an anonymous reader of a previous draft of the present article, and then discuss how our main theorem goes beyond the given argument.

The modular curve X0(N) exists over , is nodal, and its cusps are disjoint -valued points. For all primes p dividing N, the Atkin–Lehner involutions wp of X0(N) commute with each other and generate a group GN of order 2r. Let X0(N)+ = X0(N)/GN, and let c denote the unique cusp of X0(N)+. The cusp can be shown to be smooth over ; moreover, all the fibers of X0(N)+ over are geometrically irreducible. Therefore, the complement U of in X0(N)+ is affine. The coordinate ring is a finitely generated -algebra, with an increasing filtration by sub--modules consisting of the functions with a pole of order at most k along the cusp c. The successive quotients are free -modules, each with rank either zero or one. The k for which the rank is zero form the gap-sequence of , the generic fiber of X0(N)+. If k is not a gap, there is a non-zero element , unique up to sign and addition of elements from . The q-parameter is obtained by considering the formal parameter at c, which is the projective limit of the sequence of quotients where I is the ideal of c. The projective limit is naturally isomorphic to the formal power series ring . With all this, if f is any element of , then there is an integer n such that nf has q-expansion in . In other words, the function field of can be generated by two elements whose q-expansions are in .

Returning to our main result, we can discuss the information contained in our main theorem which goes beyond the above general argument. The above approach from arithmetic algebraic geometry yields generators of the function field with poles of small order with integral q-expansion; however, it was necessary to “clear denominators” by multiplying through by the integer n, thus allowing for the possibility that the lead coefficient is not equal to one. In the case that X0(N)+ has genus one, then results which are well-known to experts in the field (or so we have been told) may be used to show that the generators satisfy an integral Weierstrass equation. However, such arguments do not apply to the genus two and genus three cases which we exhaustively analyzed. Furthermore, we found that our result holds in all cases when N is square-free, independent of the structure of the gap sequence at i∞; see, [CitationKohnen 03]. As we found, for genus two i∞ was not a Weierstrass point for any N; however, for genus three, there were various types of gap sequences for different levels N.

Finally, given all of the above discussion, one naturally is led to the following conjecture.

Conjecture 2.

For any square-free N, the function field associated to any positive genus g group Γ0(N)+ admits two generators whose q-expansions have integer coefficients after the lead coefficient has been normalized to equal one. Moreover, the orders of poles of the generators at i∞ are at most g + 2.

1.5. Our method of proof

Let us now describe our theoretical results and computational investigations in greater detail.

For any square-free integer N, the subset of defined by is an arithmetic subgroup of . As shown in [CitationCummins 04], there are precisely 44 such groups which have genus zero and 38, 39, and 31 which have genus one, two, and three, respectively. These groups of genus up to three will form a considerable portion of the focus in this article. There are two reasons for focusing on the groups Γ0(N)+ for square-free N. First, for any positive integer N, the groups Γ0(N)+ are of moonshine-type (see, e.g., Definition 1 from [CitationGannon 06a]). Second, should N not be square-free, then there exist genus zero groups Γ0(N)+, namely when N = 25, 49, and 50, but those groups correspond to “ghost” classes of the monster. In summary, we are letting known results from “monstrous moonshine” serve as a guide that, perhaps, the information derived in this paper may someday find meaning elsewhere.

For arbitrary square-free N, the discrete group Γ0(N)+ has one-cusp, which we denote by i∞. Associated to the cusp of Γ0(N)+ one has, from spectral theory and harmonic analysis, a well-defined non-holomorphic Eisenstein series denoted by . The real analytic Eisenstein series is defined for and by (1-4) where is the parabolic subgroup of Γ0(N)+. Our first result is to determine the Fourier coefficients of in terms of elementary arithmetic functions, from which one obtains the meromorphic continuation of the real analytic Eisenstein series to all .

As a corollary of these computations, it is immediate that has no pole in the interval (1/2, 1). Consequently, we prove that groups Γ0(N)+ for all square-free N have no residual spectrum besides the obvious one at s = 1.

Using our explicit formulas for the Fourier coefficients of , we are able to study the special values at s = 1 and s = 0, which, of course, are related by the functional equation for . As a result, we arrive at the following generalization of Kronecker’s limit formula. For any square-free N which has r prime factors, the real analytic Eisenstein series admits a Taylor series expansion of the form where η(z) is Dedekind’s eta function associated to . From the modularity of , one concludes that is invariant with respect to the action by Γ0(N)+. Consequently, there exists a multiplicative character ϵN(γ) on Γ0(N)+ such that one has the identity (1-5) We study the order of the character, and we define ℓN to be the minimal positive integer so that ϵNN = 1. A priori, it is not immediate that ℓN is finite for general N. Classically, when N = 1, the study of the character ϵ1 on is the beginning of the theory of Dedekind sums; see [CitationLang 76]. For general N, we prove that ℓN is finite and, furthermore, can be evaluated by the expression where denotes the least common multiple function and σ(N) stands for the sum of divisors of N.

With the above notation, we define the Kronecker limit function ΔN(z) associated to Γ0(N)+ to be (1-6) ΔN(z) is a cusp form and we let kN denote its weight. By combining (Equation1–5) and (Equation1–6), we get that kN = 2r − 1N. In summary, ΔN(z) is a non-vanishing, weight kN holomorphic modular form with respect to Γ0(N)+.

Analogous to the setting of , one defines the holomorphic Eisenstein series of even weight k ≥ 4 associated to Γ0(N)+ by the series (1-7) We show that one can express E(N)2m in terms of E2m, the holomorphic Eisenstein series associated to ; namely, for m ≥ 2 one has the relation (1-8)

With the above analysis, we are now able to construct holomorphic modular functions on the space . For any non-negative integer M, the function (1-9) is a holomorphic modular function on , meaning a weight zero modular form with exponential growth in z as zi∞. The vector b = (bν) can be viewed as a weighted partition of the integer kNM with weights m = (mν). For considerations to be described below, we let denote the set of all possible rational functions defined in (Equation1–9) by varying the vectors b and m yet keeping M fixed. Trivially, consists of the constant function {1}, which is convenient to include in our computations.

Dedekind’s eta function can be expressed by the well-known product formula, namely and the holomorphic Eisenstein series associated to admits the q-expansion: (1-10) where Bk denotes the kth Bernoulli number and σ is the generalized divisor function:

It is immediate that each function Fb in (Equation1–9) admits a q-expansion with rational coefficients. However, it is clear that such coefficients are certainly not integers and, actually, can have very large numerators and denominators. Indeed, when combining (Equation1–8) and (Equation1–10), one gets that which, evidentially, can have a large denominator when m and N are large. In addition, note that the function in (Equation1–9) is defined with a product of holomorphic Eisenstein series in the numerators, so the rational coefficients in the q-expansion of (Equation1–9) are even farther removed from being easily described.

With the above theoretical background material, we implement the algorithm described in Section 7.3 to find a set of generators for the function field associated to the genus g group Γ0(N)+. In addition to the theoretical results outlined above, we obtain the following results which are based on our computational investigations:

(1)

For all genus zero groups Γ0(N)+, the algorithm concludes successfully, thus yielding the q-expansion of the Hauptmoduli. In all such cases, the q-expansions have positive integer coefficients as far as computed. The computations were completed for all 44 different genus zero groups Γ0(N)+.

(2)

For all genus one groups Γ0(N)+, the algorithm concludes successfully, thus yielding the q-expansions for two generators of the associated function fields. In all such cases, each q-expansion has integer coefficients as far as computed. The computations were completed for all 38 different genus one groups Γ0(N)+. The function whose q-expansion begins with q−2 is the Weierstrass ℘-function in the coordinate for the underlying elliptic curve.

(3)

For all 38 different genus one groups Γ0(N)+, we computed a cubic relation satisfied by the two generators of the function fields.

(4)

In addition, we consider all groups Γ0(N)+ of genus two and three. In every instance, the algorithm concludes successfully, yielding generators for the function fields whose q-expansions admit integer coefficients as far as computed. Only for the generators of the function field associated to the group Γ0(510)+, the coefficients are half-integers. For the latter case, we present in Section 10 an additional base change such that the coefficients get integers.

(5)

We extend all q-expansions out to order qκ, where κ is given in and or evaluated according to Remark 6, thus showing that the field generators have integer q-expansions.

Table 1. Number of expansion coefficients that need to be integer in order that all coefficients in the q-expansion of the Hauptmodul are integer. Listed is the level N and the value of κ according to Remark 6 for the genus zero groups Γ0(N)+.

Table 2. Number of expansion coefficients that need to be integer in order that all coefficients in the q-expansions of the field generators are integer. Listed is the level N and the value of κ according to Remark 6 for the genus one groups Γ0(N)+.

The fact that all of the q-expansions which we uncovered have integer coefficients is not at all obvious and leads us to believe there is deeper, so-far hidden, arithmetic structure which perhaps can be described as “higher genus moonshine.”

In some instances, the computations from our algorithm were elementary and could have been completed without computer assistance. For instance, when N = 5 or N = 6, the first iteration of the algorithm used a set with only two functions to conclude successfully. However, as N grew, the complexity of the computations became quite large. As an example, for N = 71, which is genus zero and appears in “monstrous moonshine,” the smallest non-zero weight for a denominator in (Equation1–9) was 4, but we needed to consider all functions whose numerators had weight up to 40, resulting in 362 functions whose largest pole had order 120. The most computationally extensive genus one example was N = 79 where the smallest non-zero weight denominator in (Equation1–9) was 12, but we needed to consider all functions whose numerator had weight up to 84, resulting in 13158 functions whose largest pole had order 280.

As one can imagine, the data associated to the q-expansions we considered is massive. In some instances, we encountered rational numbers whose numerators and denominators each occupied a whole printed page. In addition, in the cases where the algorithm required several iterations, the input data of q-expansions of all functions were stored in computer files which if printed would occupy hundreds of thousands of pages. As an example of the size of the problem we considered, it was necessary to write computer programs to search the output from the Gauss elimination to determine if all coefficients of all q-expansions were integers since the output itself, if printed, would occupy thousands of pages.

The computational results are summarized in Sections 810. In , , and , we list the q-expansions of the two generators of function fields associated to each genus one, genus two, and genus three group Γ0(N)+. As T. Gannon’s comment suggests, the information summarized in those tables does not exist elsewhere. In , , and , we list the polynomial relations satisfied by the generators of , , and , respectively. The stated results, in particular the q-expansions, were limited solely by space considerations; a thorough documentation of our findings will be given in forthcoming articles.

All input and output information associated to the computational investigations undertaken in the present article is made available at http://www.efsa.unsa.ba/∼lejla.smajlovic/.

1.6. Further studies

As stated above, the j-invariant can be written in terms of holomorphic Eisenstein series associated to . By storing all information from the computations from Gauss elimination, we obtain similar expressions for the Hauptmoduli for all genus zero groups Γ0(N)+. As one would imagine based upon the above discussion, some of the expressions will be rather large. In [CitationJorgenson et al. preprint-a], we will report of this investigation, which in many instances yields new relations for the j-invariants and for holomorphic Eisenstein series themselves. For the genus one groups, the computations from Gauss elimination produce expressions for the Weierstrass ℘-function, in the coordinate on , in terms of holomorphic Eisenstein series; these computations will be presented in [CitationJorgenson et al. in preparation-b]. Finally, in [CitationJorgenson et al. in preparation-c], we compute values of the Hauptmoduli jN associated to all genus zero groups Γ0(N)+ at elliptic fixed points using differential equation satisfied by the Schwarzian derivative of jN and prove that all values are algebraic integers.

1.7. Outline of the paper

In Section 2, we will establish notation and recall various background material. In Section 3, we prove that if a certain number of coefficients in the q-expansions of the generators are integers, then all coefficients are integers. In Section 4, we will compute the Fourier expansion of the non-holomorphic Eisenstein series associated to Γ0(N)+. The generalization of Kronecker’s limit formula for groups Γ0(N)+ is proven in Section 5, including an investigation of its weight and the order of the associated Dedekind sums. In Section 6, we relate the holomorphic Eisenstein series associated to Γ0(N)+ to holomorphic Eisenstein series on , as cited above. Further details regarding the algorithm we develop and implement are given in Section 7. Results of our computational investigations are given in Section 8 for genus zero, Section 9 for genus one, and Section 10 for genus two and genus three. Finally, in Section 11, we offer concluding remarks and discuss directions for future study, most notably our forthcoming articles [CitationJorgenson et al. preprint-a], [CitationJorgenson et al. in preparation-b], and [CitationJorgenson et al. in preparation-c].

1.8. Closing comment

The quote we presented above from [CitationGannon 06a] indicates that “higher genus moonshine” has yet to have the input from which one can search for the type of mathematical clues that are found in McKay’s observation involving the coefficients of the j-invariant or in Ogg’s computation of the levels of all genus zero moonshine groups Γ0(N)+ and their appearance in the prime factorization of the order of the Fischer–Griess monster . It is our hope that someone will recognize some patterns in the q-expansions we present in this article, as well as in [CitationJorgenson et al. preprint-a] and [CitationJorgenson et al. in preparation-b], and then, perhaps, higher genus moonshine will manifest itself.

2. Background material

2.1. Preliminary notation

Throughout we will employ the standard notation for several arithmetic quantities and functions, including: the generalized divisor function σa, Bernoulli numbers Bk, the Möbius function μ, and the Euler totient function ϕ, the Jacobi symbol , the greatest common divisor function ( ·, ·), and the least common multiple function . Throughout the paper we denote by {pi}, i = 1, …, r, a set of distinct primes and by N = p1⋅⋅⋅pr a square-free, positive integer.

The convention we employ for the Bernoulli numbers follows [CitationZagier 08] which is slightly different than [CitationSerre 73] although, of course, numerical evaluations agree when following either set of notation. For a precise discussion, we refer to the footnote on page 90 of [CitationSerre 73].

2.2. Certain arithmetic groups

As stated above, denotes the hyperbolic upper half plane with global variable with z = x + iy and y > 0, and we set q = eiz = e(z).

The subset of , defined by (2-11) is an arithmetic subgroup of . We denote by the corresponding subgroup of .

Basic properties of Γ0(N)+, for square-free N are derived in [CitationJorgenson et al. 14] and references therein. In particular, we use that the non-compact surface has exactly one cusp, which can be taken to be at i∞.

2.3. Function fields and modular forms

The set of meromorphic functions on XN is a function field, meaning a degree one transcendental extension of . Let gN denote the genus of XN. If gN = 0, then the function field of XN is isomorphic to where jN is a single-valued meromorphic function on XN. If gN > 0, then the function field of XN is generated by two elements which satisfy a polynomial relation.

A meromorphic function f on is a weight 2k meromorphic modular form if we have the relation In other language, the differential is Γ0(N)+ invariant.

The product, resp. quotient, of weight k1 and weight k2 meromorphic forms is a weight k1 + k2, resp. k1k2, meromorphic form. Since , each holomorphic modular form admits a Laurent series expansion which, when setting q = eiz, can be written as As is common in the mathematical literature, we consider forms such that cm = 0 whenever m < m0 for some . If , the function f frequently will be re-scaled so that when cm = 0 for m < m0; the constant m0 is the order of the pole of f at i∞.

If XN has genus zero, then by the Hauptmoduli jN for XN one means the weight zero holomorphic form which is the generator of the function field on XN. If gN > 0, then we will use the notation j1; N and j2; N to denote two generators of the function field.

Unfortunately, the notation of hyperbolic geometry writes the local coordinate on as x + iy, and the notation of the algebraic geometry of curves uses x and y to denote the generators of the function field under consideration; see, for example, page 31 of [CitationLang 82]. We will follow both conventions and provide ample discussion in order to prevent confusion.

3. Integrality of the coefficients in the q-expansion

In this section we prove that integrality of all coefficients in the q-expansion of the Hauptmodul jN when gN = 0 and the generators j1; N and j2; N when gN > 0 can be deduced from integrality of a certain finite number of coefficients. Also, our proof yields an effective bound on the number of coefficients needed to test for integrality. First, we present a proof in the case when genus is zero, followed by a proof for the higher genus setting which is a generalization of the genus zero case. The proof is based on the property of Hecke operators and Atkin–Lehner involutions. We begin with a simple lemma.

Lemma 3.

For any prime p which is relatively prime to N, let Tp denote the unscaled Hecke operator which acts on Γ0(N) invariant functions f, If f is a holomorphic modular form on Γ0(N)+, then Tp(f) is also a holomorphic modular form on Γ0(N)+.

Proof.

The form f is Γ0(N)+ invariant, hence f is Γ0(N) invariant and so is Tp(f). By Lemma 11 of [CitationAtkin and Lehner 70], if W is any coset representative of the quotient group Γ0(N)\Γ0(N)+, then Tp(f) is also W invariant since f is W invariant. Therefore, Tp(f) is a Γ0(N)+ invariant holomorphic form.

Theorem 4.

Let jN be the Hauptmodul for a genus zero group Γ0(N)+. Let p1 and p2 with p2 > p1 be distinct primes which are relatively prime to N. If the q-expansion of jN contains integer coefficients out to order qκ for some κ ≥ (p2/(p2 − 1)) · (p1p2)2, then all further coefficients in the q-expansion of jN are integers.

Proof.

Let ck denote the coefficient of qk in the q-expansion of jN. Let m > κ be arbitrary integer. Assume that ck is an integer for k < m, and let us now prove that cm is an integer. To do so, let us consider the Γ0(N)+ invariant functions for specific integers r1 and r2 depending on m which are chosen as follows. First, take r1 to be the unique integer such that We now wish to choose r2 to be the smallest positive integer such that Let us argue in general the existence of r2 and establish a bound in terms of p1 and p2, noting that for any specific example one may be able to choose r2 much smaller than determined by the bound below.

Since p1 and p2 are primes and 1 ≤ r1p1 < p2, we have that (p2, r1p1) = 1. By applying the Euclidean algorithm and Bezout’s identity, there exists an r ≥ 1 such that rp2 ≡ 1modr1p1. Choose A be the smallest positive integer such that and Clearly, and, furthermore In particular, r1p1 and r2p2 are relatively prime, and we have the bounds r1p1p21 and r2p2p21p22.

With the above choices of r1 and r2, set f1 = jr1N and f2 = jr2N. The function f1 is a Γ0(N)+ invariant modular function with pole of order r1 at i∞. Therefore, there is a polynomial of degree r1p1 such that (3-12) If we write the q-expansion of f1 as then The coefficients bk are determined by the binomial theorem and the coefficients of jN, namely by By assumption, c1, …, cm − 1 are integers, and m − 1 ≥ (p2/(p2 − 1))(p1p2)2 > r1. From this, we conclude that are integers. In particular, all coefficients of out to order q0 are integers, from which we can compute the coefficients of and conclude that the polynomial has integer coefficients, in particular the lead coefficient is one.

Let us determine the first appearance of the coefficient cm in . First, the smallest k where cm appears in the formula for bk is when k = m + 1 − r1, and then we have that By our choice of r1, the index m + 1 − r1 is positive and divisible by p1. As a result, the first appearance of cm in the expansion of is within the coefficient of qd where d = (m + 1 − r1)/p1. Going further, we have that the coefficient of qd for d = (m + 1 − r1)/p1 in the expansion of is of the form p1r1cm plus an integer which can be determined by binomial coefficients and ck for k < m.

Let us now determine the first appearance of cm in . Again, by the binomial theorem, we have that cm first appears as a coefficient of qe where e = mr1p1 + 1. By the choice of m and r1, since p1 < p2, one has that The above inequality is equivalent to in other words, d < e. As a consequence, we have that the coefficient of qd in can be written as a polynomial expression involving binomial coefficients and ck for k < m. By induction, the coefficient of qd in is an integer.

In summary, by equating the coefficients of qd where d = (m + 1 − r1)/p1 in the formula (Equation3–12), on the left-hand side we get an expression of the formula r1p1cm plus an integer, and on the right-hand side we get an integer. Therefore, cm is a rational number whose denominator is a divisor of r1p1.

Let us now consider . Since m + 1 − r2 is divisible by p2, we consider the coefficient of where d′ = (m + 1 − r2)/p2. By the choice of m and r2, recalling that r2p21p2, we have and this is equivalent to or d′ < e′ = mr2p2 + 1. With all this, we conclude, analogously as in the first case that cm is a rational number whose denominator is a divisor of r2p2. However, r1p1 and r2p2 are relatively prime, hence cm is an integer.

By induction on m, the proof of the theorem is complete.

Theorem 5.

Let Γ0(N)+ have genus greater than zero, and assume there exists two holomorphic modular functions j1; N and j2; N which generate the function field associated to Γ0(N)+. Furthermore, assume that the q-expansions of j1; N and j2; N are normalized in the form: Let p1 and p2 with p2 > p1 be distinct primes which are relatively prime to a1a2N. Assume the q-expansions of j1; N and j2; N contain integer coefficients out to order qκ with κ ≥ a2(p2/(p2 − 1)) · (p1p2)2. Then all further coefficients in the q-expansions of j1; N and j2; N are integers.

Proof.

The argument is very similar to the proof of Theorem 4. Let cl; k denote the coefficient of qk in the q-expansion of jl; N, l ∈ {1, 2}. Assume that cl; k is an integer for l ∈ {1, 2} and k < m, and let us prove that c1; m and c2; m are integers.

For i, l ∈ {1, 2}, we study the expression (3-13) for certain polynomials Qi, l of two variables. The left-hand side of the above equation has a pole at i∞ of order alpiri. Hence The existence of Qi, l(x, y) follows from the assumption that j1; N and j2; N generate the function field associated to Γ0(N)+ and the observation that is Γ0(N)+ invariant and has a pole of order less than alripi.

Of course, the polynomials Qi, l are not unique since j1; N and j2; N satisfy a polynomial relation. This does not matter. We introduce the following canonical choice in order to uniquely determine the polynomials. Consider the coefficient . If there exist non-negative integers n1 and n2 such that a1n1 + a2n2 = a1n1 + a2n2 with n1 < n1 then we set equal to zero.

Integrality of coefficients of j1; N and j2; N out to order qκ with κ ≥ a2(p2/(p2 − 1)) · (p1p2)2 implies that all coefficients of out to order q0 are integers. The coefficient of the polynomial Qi, l(x, y) first appears on the right-hand side of (Equation3–13) as a coefficient multiplying . The canonical choice of coefficients enables us to deduce, inductively in the degree ranging from −alripi + 1 to zero that all coefficients are integers.

Having established that the coefficients of the polynomials Qi, l are integers, now we wish to determine two values r1 and r2 so that the coefficient c1; m first appears in as a factor of qd for the smallest possible d, which leads to determining ri from the equation . Such a solution exists provided p1 and p2 are relatively prime to a1. Without loss of generality, we may assume that 1 ≤ r1p1. As in the proof of Theorem 4, we impose the further condition that (r1p1, r2p2) = 1. For at least one value of r2 in the range 1 ≤ r2r1p1p2, we have that and , so, in particular, (r1p1, r2p2) = 1.

With the above choices of r1, we determine the first appearance of c1; m in the equation: On the left-hand side, c1; m appears as a coefficient of qd where d = (m − (r1 − 1)a1)/p1, and, in fact, the coefficient of qd is equal to an integer plus r1p1c1; m. On the right-hand side, c1; m first appears as a coefficient of qe where e = ma1(r1p1 − 1). We have that d < e when m > m′ where The resulting expression shows that c1; m can be expressed as a fraction, where the numerator is a finite sum involving integer multiples of positive powers of c1; k and c2; k for k < m and denominator equal to r1p1. Similarly, by studying the first appearance of c1; m in the expression: we obtain an expression showing that c1; m can be expressed as a fraction, where the numerator is a finite sum involving integer multiples of positive powers of c1; k and c2; k for k < m and denominator equal to r2p2. By induction on m, and that (r1p1, r2p2) = 1, we conclude that c1; m is an integer.

Analogously, one studies c2; m. In the four different equations studied, the minimum number of coefficients needed to be integers in order to initiate the induction is the smallest integer that is larger than or equal to a2(p2/(p2 − 1))(p1p2)2, which is assumed to hold in the premise of the theorem.

Remark 6.

Let us now describe, then implement, an algorithm which will reduce the number of coefficients which need to be tested for integrality. For this remark alone, let us set a1 = a2 = 1 if gN = 0.

Let m be the lower bound determined in Theorem 4 and Theorem 5. Let p1, …, pk be the set of primes less than m which are relatively prime to a1a2N. For each prime, let ri, l be the integer in the range 1 ≤ ri, lpi satisfying , l ∈ {1, 2}. Let Sm, l denote the set of triples: Note that if ri, l is such that al(ri, l(pi + 1) − 1) ≥ m, then the set Sm, l does not contain a triple whose prime is pi.

Now let Assume that the greatest common divisor of all elements in Rm, 1 is one and that of Rm, 2 is also one. Then, by using all the Hecke operators applied to the functions jri, 1 + jpi1; N and jri, 2 + jpi2; N and the arguments from Theorem 4 and Theorem 5, we can determine c1; m and c2; m from lower indexed coefficients and show that c1; m and c2; m are integers if all lower indexed coefficients are integers.

Although the above observation is too cumbersome to employ theoretically, it does lead to the following algorithm which can be implemented:

(1)

Let m be the lower bound given in Theorem 4 for g = 0 or Theorem 5 for g > 0.

(2)

Construct the sets Sm, l and Rm, l, as described above.

(3)

If the greatest common divisor of all elements in Rm, 1 is one and that of Rm, 2 is one, too, then replace m by m − 1 and repeat with Step 2.

(4)

If the greatest common divisor of all elements of Rm, 1 or Rm, 2 is greater than one, let κ = m.

The outcome of this algorithm yields a reduced number κ of coefficients which need to be tested for integrality in order to conclude that all coefficients are integers.

In and we list the level N and improved bounds on κ which were determined by the above algorithm for all genus zero and genus one groups Γ0(N)+.

4. The Fourier coefficients of the real analytic Eisenstein series

In this section we derive formulas for coefficients in the Fourier expansion of the real analytic Eisenstein series . First, we derive an intermediate result, expressing the Fourier coefficients in terms of Ramanujan sums. Then, we compute coefficients in a closed form.

4.1. The Fourier expansion of real analytic Eisenstein series on Γ0(N)+

Lemma 7.

The real analytic Eisenstein series , defined by (Equation1–4) for z = x + iy and , admits a Fourier series expansion: (4-14) where Ws(mz) is the Whittaker function given by and Ks − 1/2 is the Bessel function. Furthermore, coefficients of the Fourier expansion (Equation4–14) are given by where is the completed Riemann zeta function and (4-15) where we put

Proof.

The Fourier expansion (Equation4–14) of the real analytic Eisenstein series for is a special case of [CitationIwaniec 02], Theorem 3.4., when the surface has only one cusp at i∞ with identity scaling matrix. By ϕN(s) we denote the scattering matrix, evaluated in [CitationJorgenson et al. 14] as where

If CN denotes the set of positive left-lower entries of matrices from Γ0(N)+, then, by [CitationIwaniec 02], Theorem 3.4. The Fourier coefficients of (Equation4–14) are given by (4-16) where SN denotes the Kloosterman sum (see [CitationIwaniec 02], formula (2.23)) defined, for cCN, as (4-17) In [CitationJorgenson et al. 14] we proved that For a fixed , with vN and arbitrary, we can take e = v in the definition of Γ0(N)+ to deduce that matrices from Γ0(N)+ with left lower entry c are given by for some integers a, b, and d such that vad − (N/v)bn = 1. Since N is square-free, this equation has a solution if and only if (v, n) = 1 and (d, (N/v)n) = 1.

Therefore, for , and fixed equation (Equation4–17) becomes This is a Ramanujan sum, which can be evaluated using formula (3.3) from [CitationIwaniec and Kowalski 04], page 44, to get For element belongs to CN if and only if vN and (n, v) = 1, hence equation (Equation4–16) becomes which proves (Equation4–15).

4.2. Computation of Fourier coefficients

In this subsection we will compute coefficients (Equation4–15) in closed form. We will assume that m > 0 and incorporate negative terms via the identity ϕN( − m, s) = ϕN(m, s).

Let p and q denote prime numbers, and let denote the highest power of p that divides m, i.e., the number αp is such that and . For any prime p, set and

Theorem 8.

Assume N is square-free and let ϕN(m, s) denote the mth coefficient in the Fourier series expansion (Equation4–14) of the Eisenstein series. Then the coefficients ϕN(m, s) can be meromorphically continued from the half plane to the whole complex plane by the equation:

Proof.

We employ equation (Equation4–15) and observe that coefficients am(n) are multiplicative. Therefore, it is natural to investigate the associated L-series: defined for . The series Lm(s) is defined in [CitationKoyama 09], Lemma 2.2, for , where its analytic continuation to the whole complex plane is given by the formula Lm(s) = σ1 − s(m)/ζ(s).

The multiplicativity of coefficients am(n) implies that for one has the Euler product: where Ep(m, s) stands for the pth Euler factor, i.e., (4-18) By the computations presented in [CitationKoyama 09], page 1137, one has hence, Ep(m, s) = Fp(m, s). Therefore, (4-19)

Let vN be fixed. In order to compute the sum: using equation (Equation4–19), we use the fact that N is square-free, so (v, N/v) = 1, hence every integer n coprime to v can be represented as (4-20) Since (k, v) = 1 and (k, q) = 1 for every q∣(N/v), we deduce that actually (k, N) = 1. Therefore (4-21) Let q1, …, ql denote the set of all prime divisors of N/v. The multiplicativity of coefficients am implies that for n given by (Equation4–20) one has hence, by (Equation4–21) one gets that The identity (Equation4–19) is now applied to the inner sum in the above equation, from which we deduce that for .

For all ν = 1, …, l one then has by (Equation4–18). Therefore, (4-22)

Substituting (Equation4–22) into (Equation4–15) completes the proof.

Proposition 9.

For a square-free integer N = ∏rν = 1pν, DN(m, s) is multiplicative in N, Furthermore, for a prime p and : where is the highest power of p dividing m.

Proof.

We apply induction on number r of prime factors of N. For r = 1 the statement is immediate.

Assume that the statement is true for all numbers N with r distinct prime factors and let N = ∏r + 1ν = 1pν. Since {v: vN} = {v: v∣∏rν = 1pν}∪{v · pr + 1: v∣∏rν = 1pν}, the statement is easily deduced from the definition of DN(m, s) and the inductive assumption.

The second statement follows trivially from the properties of the function σs(m) and the definition of the Euler factor Fp.

Corollary 10.

For any square-free N, the groups Γ0(N)+ have no residual eigenvalues λ ∈ [0, 1/4) besides the obvious one λ0 = 0.

Proof.

From Lemma 7 with m = 0, we have a formula for the constant term ϕN(s) in the Fourier expansion of the real analytic Eisenstein series, namely see also Lemma 5 of [CitationJorgenson et al. 14]. It is elementary to see that ϕN(s) has no poles in the interval (1/2, 1). From the spectral theory, one has that the residual eigenvalues λ of the hyperbolic Laplacian such that λ ∈ (0, 1/4) correspond to poles s of the real analytic Eisenstein series with s ∈ (1/2, 1) with the relation s(1 − s) = λ. Since poles of real analytic Eisenstein series are exactly the poles of ϕN(s), the statement is proved.

5. Kronecker’s limit formula for Γ0(N)+

In this section we derive Kronecker’s limit formula for the Laurent series expansion of the real analytic Eisenstein series. First, we will prove the following, intermediate result.

Proposition 11.

For a square-free integer N = ∏rν = 1pν, we have the Laurent expansion: (5-23) as s → 1, where and

Proof.

We start with the formula (Equation4–14) for the Fourier expansion of real analytic Eisenstein series at the cusp i∞.

Let ϕ denote the scattering matrix for . Then, one has ϕN(s) = ϕ(s)DN(s). We will use this fact in order to deduce the first two terms in the Laurent series expression of around s = 1.

By the classical Kronecker limit formula for the Laurent series expansion of ϕ(s)y1 − s at s = 1 is given by (see, e.g., [CitationKühn 01], page 228) where A = 2 − 24ζ′( − 1) − 2log 4π.

The function DN(s) is holomorphic in any neighborhood of s = 1, hence DN(s) = DN(1) + DN(1)(s − 1) + O(s − 1)2, as s → 1. Therefore, as s → 1.

In order to get the residue and the constant term stemming from ϕN it is left to compute values of DN and DN at s = 1, This, together with (Equation4–14) and Theorem 8 completes the proof.

Since the series always has a pole at s = 1 with residue , from the above proposition, we easily deduce a simple formula for the volume of the surface XN in terms of the level of the group, namely

Theorem 12.

For a square-free integer N = ∏rν = 1pν, we have the asymptotic expansion as s → 1, where C−1, Nand C0, N are defined in Proposition 11 and η is the Dedekind eta function defined for by

Proof.

From the definition of Euler factors Fp(m, s), using the fact that , we get that Therefore, a simple computation implies that hence where we put P(i1, …, ik) = 1 for k = 0, and for k = 1, …, r Every m ≥ 1 can be written as m = pα11⋅⋅⋅pαrr · l, where αi ≥ 0 and (l, N) = 1. Therefore, we may write the sum over m in (Equation5–23) as (5-24) When k = 0 the above sum obviously reduces to (5-25) as deduced from the proof of the classical Kronecker limit formula.

Now, we will take k ∈ {1, …, r} and compute one term in the sum (Equation5–24). Without loss of generality, in order to ease the notation, we may assume that i1 = 1, …, ik = k and compute the term with P(1, …, k). Later, we will take the sum over all indices 1 ≤ i1 < ⋅⋅⋅ < ikr. First, we observe that P(1, …, k) = 0 if αi = 0, for any i ∈ 1, …, k and if all αi ≥ 1, for i = 1, …, k. Furthermore, for αi ≥ 1, i = 1, …, k one has Therefore, Since every integer m ≥ 1 can be represented as m = pα1 − 11⋅⋅⋅pαk − 1k · pαk + 1k + 1⋅⋅⋅pαrr · l, where α1, …, αk ≥ 1; αk + 1, …, αr ≥ 0 and (l, N) = 1 we immediately deduce that by (Equation5–25).

Let us now sum over all k = 0, …, r and sum over all indices 1 ≤ i1 < … < ikr, which is equivalent to taking the sum over all vN. The sum (Equation5–24) then becomes since ∑vNv = σ(N). Therefore, we get that The proof is complete.

Remark 13.

Since the number of divisors of N = p1⋅⋅⋅pr is 2r, the expression is a geometric mean.

When N = 1, Theorem 12 amounts to the classical Kronecker limit formula.

From Theorem 12 and the functional equation for real analytic Eisenstein series, we can reformulate the Kronecker limit formula as asserting an expansion for as s → 0.

Proposition 14.

For , the real analytic Eisenstein series admits a Taylor series expansion of the form

An immediate consequence of Theorem 12 and the above proposition is the fact that the function (5-26) is invariant under the action of the group Γ0(N)+. Using the fact that the eta function is non-vanishing on , we may deduce the stronger statement.

Proposition 15.

There exists a character ϵN(γ) on Γ0(N)+ such that (5-27) for all γ ∈ Γ0(N)+.

Proof.

Let Since the expression (Equation5–26) is a group invariant we have Dividing by and raising the expression to the 2r/4 power, we get that hence for some function f of absolute value 1.

Since the function η is a holomorphic function, non-vanishing on , for a fixed γ ∈ Γ0(N)+ the function f(γ, z) is a holomorphic function in z on of absolute value 1 on , hence it is constant, as a function of z. Therefore, f(z, γ) = ϵN(γ), for all where |ϵN(γ)| = 1.

It is left to prove that ϵN1γ2) = ϵN1N2), for all γ1, γ2 ∈ Γ0(N)+. This is an immediate consequence of formula (Equation5–27).

The character ϵN(γ) is the analogue of the classical Dedekind sum.

In the case when the genus of the surface XN is equal to zero, the group Γ0(N)+ is generated by a finite number of elliptic elements and one parabolic element. The character ϵN(γ) is necessarily finite on each elliptic element. The fundamental group of XN contains a relation which expresses the generator of the parabolic group as a product of finite, cyclic elements; hence ϵN(γ) is finite on the parabolic element as well. Therefore, there exists an integer ℓN such that for all γ ∈ Γ0(N)+.

Irrespectiveof thegenus of the surface XN, we will prove that the character ϵN(γ) is a certain 24th root of unity.

Theorem 16.

Let N = p1⋅⋅⋅pr. Let us define the constant ℓN by Then, the function is a weight kN = 2r − 1N holomorphic form on Γ0(N)+ vanishing only at the cusp.

Proof.

The vanishing of the function ΔN(z) at the cusp i∞ only is an immediate consequence of properties of the Dedekind eta function. Therefore, it is left to prove that ΔN(z) is a weight kN = 2r − 1N holomorphic form on Γ0(N)+.

We begin with the decomposition , where see [CitationMaclachlan 81], page 147, with a slightly different notation; elements and are interchanged.

For each vN and v > 1, from [CitationAtkin and Lehner 70], Lemma 8 and Lemma 9 (see also [CitationMaclachlan 81], page 147) it follows that Γ0(N, v) is the coset under the action of the congruence group Γ0(N) of a product of at most r non-trivial elements , jAv⊆{1, …, r} such that . Therefore, every element of Γ0(N)+ can be written as a finite product of an element from Γ0(N) and elements from Γ0(N)+ whose square is in Γ0(N). Since the character ϵN defined in Proposition 15 is multiplicative, it is sufficient to prove that for all γ ∈ Γ0(N) and for all elements τ of Γ0(N)+ such that τ2 ∈ Γ0(N).

Therefore, it is actually sufficient to prove that for for an arbitrary .

For this, we apply results of [CitationRaji 06], chapter 2.2.3, pages 21–23, with f1 = ΔN; g = 2r; δl = v, , for all δl = v and k = kN. By the definition of ℓN, we see that , therefore, conditions (2.11) and (2.12) of [CitationRaji 06] are fulfilled, so then where denotes the Jacobi symbol. The multiplicativity of the Jacobi symbol implies that hence Therefore, since kN is divisible by 4. The proof is complete.

In and we list the values of degree ℓN and weight kN for all genus zero and genus one groups Γ0(N)+.

Table 3. The degree ℓN and the weight kN of the modular form ΔN on Γ0(N)+ for all groups Γ0(N)+ of genus zero. Listed are the level N, and the values of degree ℓN and weight kN.

Table 4. The degree ℓN and the weight kN of the modular form ΔN on Γ0(N)+ for all groups Γ0(N)+ of genus one. Listed are the level N and the values of degree ℓN and weight kN.

6. Holomorphic Eisenstein series on Γ0(N)+

In the classical case, holomorphic Eisenstein series of even weight k ≥ 4 are defined by formula (Equation1–1). Modularity is an immediate consequence of the definition. We proceed analogously in the case of the arithmetic groups Γ0(N)+ and define for even k ≥ 4 modular forms by (Equation1–7). The following proposition relates E(N)k(z) to Ek(z).

Proposition 17.

Let E(N)k(z) be the modular form defined by (Equation1–7). Then, for all even k = 2m ≥ 4 and all one has

 

(6-28) where Ek is defined by (Equation1–1).

Proof.

Taking e ∈ {v: vN} in the definition (Equation2–11) of the arithmetic group Γ0(N)+ we see that where

Now, we use the fact that N is square-free to deduce that for a fixed v1N we have the disjoint union decomposition Therefore, for a fixed v1N we have that Multiplying the above equation by vk/21 and taking the sum over all v1N we get Once we fix v1N and put , it is easy to see that v ranges over all divisors of N, as u1 runs through divisors of v1 and u2 runs through divisors of N/v1.

Furthermore, for fixed v1, vN there exists a unique pair (u1, u2) of positive integers such that u1v1, u2∣(N/v1) and . Reasoning in this way, we see that By the definition of the function σm(N) the above equation is equivalent to (Equation6–28), which completes the proof.

7. Searching for generators of function fields

With the above analysis, we are now able to define holomorphic modular functions on the space .

7.1. q-expansions

According to Theorem 16, the form ΔN has weight kN, and its q-expansion is easily obtained from inserting the product formula of the classical Δ function: in This allows us to get the q-expansion for any power of ΔN, including negative powers.

Similarly, the q-expansion of the holomorphic Eisenstein series E(N)k is obtained from the q-expansion of the classical Eisenstein series Ek. Using the notation of [CitationZagier 08], we write where Bk stands for the kth Bernoulli number (e.g., −8/B4 = 240; −12/B6 = −504, etc.). From (Equation6–28) we deduce

7.2. Subspaces of rational functions

Now it is evident that for any positive integer M, the function: (Equation1–9) is a holomorphic modular function on , meaning a weight zero modular form with exponential growth in z as zi∞. Its q-expansion follows from substituting E(N)k and ΔN by their q-expansions.

Let denote the set of all possible rational functions defined in (Equation1–9) for all possible vectors b = (bν) and m = (mν) with fixed M. Our rationale for finding a set of generators for the function field of the smooth, compact algebraic curve associated to XN is based on the assumption that a finite span of contains the set of generators for the function field. Subject to this assumption, the set of generators follows from a base change. The base change is performed as follows.

7.3. Our algorithm

Choose a non-negative integer κ. Let M = 1 and set . Let gN denote the genus of XN.

(1)

Form the matrix of coefficients from the q-expansions of all elements of , where each element in is expanded along a row with each column containing the coefficient of a power, negative, zero, or positive, of q. The expansion is recorded out to order qκ.

(2)

Apply Gauss elimination to , thus producing a matrix which is in row-reduced echelon form.

(3)

Implement the following decision to determine whether the algorithm has completed: If the gN lowest non-trivial rows at the bottom of correspond to q-expansions whose lead terms have precisely gN gaps in the set {q−1, …, q−2gN}, then the algorithm is completed. If the indicator to stop fails, then replace M by M + 1, by and reiterate the algorithm.

We described the algorithm with the choice of an arbitrary κ. For reasons of efficiency, we initially selected κ to be zero, so that all coefficients for qν for ν ≤ κ are included in , but finally increased κ to its desired value according to Remark 6.

The rationale for the stopping decision in Step 3 above is based on two ideas, one factual and one hopeful. First, the Weierstrass gap theorem states that for any point P on a compact Riemann surface there are precisely gN gaps in the set of possible orders from 1 to 2gN of functions whose only pole is at P. For the main considerations of this paper, we study gN ≤ 3. For instance, when gN = 1 there occurs exactly one gap which for topological reasons is always {q−1}. Second, for any genus, the assumption which is hopeful is that the function field is generated by the set of holomorphic modular functions defined in (Equation1–9). The latter assumption is not obvious, but it has turned out to be true for all groups Γ0(N)+ that we have studied so far. This includes in particular all groups Γ0(N)+ of genus zero, genus one, genus two, and genus three.

7.4. Implementation notes

We have implemented C code that generates in integer arithmetic the set of all possible rational functions defined in (Equation1–9) for any given positive integer K and any square-free level N. The set is passed in symbolic notation to PARI/GP [CitationThe PARI Group 11] which computes in rational arithmetic the q-expansions of all elements of up to any given positive order κ and forms the matrix . Theoretically, we could have used PARI/GP for the Gauss elimination of . However, for reasons of efficiency, we have written our own C code, linked against the GMP MP library [CitationGranlund and the GMP development team 12], to compute the Gauss elimination in rational arithmetic. The result is a matrix whose rows correspond to q-expansions with rational coefficients. Inspecting the rational coefficients of , we find that the denominators are 1, i.e., after Gauss elimination the coefficients turn out to be integers. All computations are in rational arithmetic, hence are exact.

After the fact, we select κ as in and in case when genus is zero or one, or compute κ as explained in Remark 6 when genus is bigger than one, and extend the q-expansions of the field generators out to order qκ. Inspecting the coefficients we find that they are integer which proves that the field generators have integer q-expansions.

It is important that we compute the Gauss elimination with a small value of κ first, and extend the q-expansions for the field generators out to the larger number κ according to Remark 6, later. There are two reasons for this: (i) It would be a waste of resources to start with a large value of κ. Once the Gauss elimination is done with a small value of κ, it is easy to extend the q-expansions out to a larger value of κ using reverse engineering of the Gauss elimination. (ii) If the genus is larger than one, we do not know the gap sequence for the field generators in advance. We can determine a1 and a2, and hence κ according to Remark 6 only if we have completed the Gauss elimination with some other zero or positive value of κ before.

8. Genus zero

The genus zero case, due to its simplicity, is suitable for demonstration of the algorithm. We do that at the example of holomorphic modular functions on XN for the level N = 17. After four iterations, the set consists of 15 functions. With the above notation, the functions are as follows, with their associated q-expansions: After Gauss elimination, we derive a set of functions with the following q-expansions: Every function fk with k = 1, …, 11 can be written as a polynomial in f12. For example, f11 = f212 − 14. Further relations are easily derived.

As stated, the j-invariant, also called the Hauptmodul, is the unique holomorphic modular function with simple pole at z = i∞ with residue equal to 1 and constant term equal to zero. For Γ0(17)+, the Hauptmodul is given by j17 = f12(z). It is remarkable that the q-expansion for j17 has integer coefficients, especially when considering how we obtained j17.

In the case when the surface XN has genus equal to zero, q-expansions of Hauptmoduls jN coincide with McKay–Thompson series for certain classes of the monster . Therefore, we were able to compare results obtained using the algorithm explained above with well known expansions that may be found, e.g., at [CitationSloane 10]. All expansions were computed out to order qκ and they match with the corresponding expansions of McKay–Thompson series available at [CitationSloane 10].

The successful execution of our algorithm in the genus zero case, besides direct check of the correctness of the algorithm provides two additional corollaries. First, the j-invariant is expressed as a linear combination of elements from the set , for some M, thus showing that the Hauptmoduli can be written as a rational function in holomorphic Eisenstein series and the Kronecker limit function, thus generalizing (Equation1–2). Second, since the function field associated to the underlying algebraic curve is generated by the j-invariant, we conclude that all holomorphic modular forms on XN are generated by a finite set of holomorphic Eisenstein series and a power of the Kronecker limit function, again generalizing a classical result for . A complete analysis of these two results for all genus zero groups Γ0(N)+ will be given in [CitationJorgenson et al. preprint-a].

9. Genus one

The smallest N such that Γ0(N)+ has genus one is N = 37. In that example, the Kronecker limit function is , which has weight 12. After four iterations, the algorithm completes successfully. However, consists of 434 functions, and this set contains functions whose pole at i∞ has order as large as 76. For instance, the function F(48)(z) = E(37)48(z)/(Δ37(z))4 is in , and the first few terms in the q-expansion of F(48) are where a−75 and a−74 are the following rational numbers:

In the setting of genus one, there is no holomorphic modular function with a simple pole at i∞. However, there does exist a holomorphic modular function with an order two pole at i∞, which is equal to the Weierstrass ℘-function. We let j1; N denote the ℘-function for Γ0(N)+. Also, let j2; N denote the holomorphic modular function with a third order pole at i∞. In the coordinates of the complex plane, j2; N is the derivative of j1; N. However, in the coordinate of the upper half plane, either z or q, one will not have that j2; N is the derivative of j1; N, as can be seen by the chain rule from one variable calculus. In all cases, the functions j1; N and j2; N generate the function field of the underlying elliptic curve.

Our algorithm yields the q-expansions of the functions j1; N and j2; N. In the case N = 37, we obtain the expansions: and

As we have stated, the coefficients in the q-expansions of j1; 37 and j2; 37 are integers. Again, noting the size of the denominators of the coefficients in the q-expansions of elements of , it is striking that the generators of the function fields would have integer coefficients.

In , we list the generators of the function field of the underlying algebraic curve for all 38 genus one groups Γ0(N)+. In some instances, the computations behind were quite intensive. For N = 79, the Kronecker limit function has weight 12, and we needed seven iterations of the algorithm. When considering all functions whose numerator had weight up to 84, the set had 13158 functions, and the orders of the pole at i∞ was as large as 280. As N became larger, the size of the denominators of the rational coefficients in the q-expansions for grew very large, in some cases becoming thousands of digits in length. Nonetheless, we have computed q-expansions of generators of the function field in all 38 genus one cases out to order qκ where κ is listed in . The exact arithmetic of the program and successful completion of the algorithm, together with Theorem 5, yields the following result.

Table 5. q-expansion of y = j2; N(z), and q-expansion of x = j1; N(z) for the genus one groups Γ0(N)+.

For each genus one group Γ0(N)+, the two generators j1; N and j2; N of the function field have integer q-expansions.

If we set x = j1;37(z) and y = j2;37(z), then it can be shown that j1; 37 and j2; 37 satisfy the cubic relation: In we present the list of cubic relations satisfied by j1; N and j2; N for all genus one groups Γ0(N)+.

Table 6. Cubic relation satisfied by x = j1;N(z) and y = j2;N(z) for the genus one groups Γ0(N)+.

As in the genus zero setting, there are two important consequences of the successful implementation of our algorithm. First, the functions j1; N and j2; N can be expressed as linear combinations of elements in for some M. In particular, the Weierstrass ℘-function j1; N can be written in terms of certain holomorphic Eisenstein series and powers of the Kronecker limit function. The same assertion holds for j2; N. Second, since j1; N and j2; N generate the function field, we conclude that all holomorphic modular forms can be written in terms of certain holomorphic Eisenstein series and the Kronecker limit functions. A full documentation of these results will be given in [CitationJorgenson et al. in preparation-b].

10. Higher genus examples

To conclude the presentation of results, we describe some of the information we obtained for genus two and genus three groups Γ0(N)+. In all instances, we derived q-expansions for the generators of the function fields out to qκ, where κ is determined according to Remark 6. With the exception of j1;510 and j2;510, the q-expansions have integer coefficients. The q-expansions of j1;510 and j2;510 have half-integer coefficients. With a further base change, namely v = j2;510 + j1;510 and w = j2;510j1;510, the q-expansions for the generators v and w have integer coefficients for N = 510, also. This together with Theorem 5 yields the following result and finally proves Theorem 1.

For each group Γ0(N)+ of genus up to three, the two generators j1; N and j2; N for N ≠ 510 and j2; N + j1; N and j2; Nj1; N for N = 510 have integer q-expansions. Moreover, the leading coefficients are one and the orders of the poles at i∞ are at most gN + 2.

As it turned out, for all groups Γ0(N)+ of genus two and for all but four groups Γ0(N)+ of genus three, the gap sequence consists of {q−1, …, qgN}, and i∞ was not a Weierstrass point. For Γ0(109)+, Γ0(151)+, and Γ0(179)+, the gap sequence consists of {q−1, q−2, q−4} and for Γ0(282)+ the gap sequence consists of {q−1, q−2, q−5}. For the latter four groups, i∞ is a Weierstras point.

We list in the q-expansions of the generators as well as the algebraic relation satisfied by the generators.

Table 7. q-expansion of y = j2; N(z), and q-expansion of x = j1; N(z), and κ for the genus two groups Γ0(N)+.

Table 8. Polynomial relation satisfied by x = j1;N(z) and y = j2;N(z) for the genus two groups Γ0(N)+.

Table 9. q-Expansion of y = j2; N(z), and q-expansion of x = j1; N(z), and κ for the genus three groups Γ0(N)+. For N = 510 the value of κ is for the generators v = y + x and w = yx.

Table 10. Polynomial relation satisfied by x = j1;N(z) and y = j2;N(z) for the genus three groups Γ0(N)+. For N = 510, we give the polynomial relation satisfied by v = y + x and w = yx, instead.

11. Concluding remarks

In a few words, we can express a response to the problem posed by T. Gannon which we quoted in Section 1.

The generators of the function fields associated to XN which one should consider are given by the two invariant holomorphic functions whose poles at i∞ have the smallest possible orders and with initial q-expansions which are normalized by the criteria of row-reduced echelon form.

The above statement when applied to all the groups Γ0(N)+ we considered produced generators with integer q-expansions, except for the level N = 510 where a further base change was necessary. The cases we considered include all groups Γ0(N)+ with square-free N of genus up to three.

In other articles, we will present further results related to the material presented here. Specifically, in [CitationJorgenson et al. in preprint-a], we study further details regarding genus zero groups Γ0(N)+. In particular, for each level, we list the generators of the ring of holomorphic forms as well as relations for the Hauptmodul and Kronecker limit functions in terms of holomorphic Eisenstein series. Similar information for genus one groups Γ0(N)+ is studied in [CitationJorgenson et al. in preparation-b], as well as additional aspects of the “arithmetic of the moonshine elliptic curves.”

Finally, in [CitationJorgenson et al. in preparation-c], we compute values of the Hauptmoduli jN associated to all genus zero groups Γ0(N)+ at elliptic fixed points and prove that all values are algebraic integers.

Funding

J. J. acknowledges grant support from NSF and PSC-CUNY grants, and H. T. acknowledges support from EPSRC grant EP/H005188/1.

Acknowledgments

The first named author (J.J.) discussed a preliminary version of this paper with Peter Sarnak which included the findings from Section 4 onward. During the discussion, Professor Sarnak suggested that we seek a proof of the integrality of all coefficients in the q-expansion once we have shown that a certain number of the first coefficients are integers. His suggestion led us to formulate and prove the material in Section 3. We thank Professor Sarnak for generously sharing with us his mathematical insight. The classical material in Section 1.4 came from a discussion between the third named author (H.T.) and Abhishek Saha. The approach using arithmetic algebraic geometry stemmed from a discussion between the first named author (J.J.) and an anonymous reader of a previous draft of this article. We thank these individuals for allowing us to include their remarks. Finally, the numerical computations for some levels were quite resource demanding. We thank the Public Enterprise Electric Utility of Bosnia and Herzegovina for generously granting us full access to one of their new 256 GB RAM computers, which enabled us to compute for all groups Γ0(N)+ of genus up to three. We are very grateful for their support of our academic research.

References

  • [Atkin and Lehner 70] A. O. L. Atkin, J. Lehner. “Hecke Operators on Γ0(m).” Math. Ann. 185 (1970), 134–160.
  • [Baier and Köhler 03] H. Baier and G. Köhler. “How to Compute the Coefficients of the Elliptic Modular Function j(z).” Exp. Math. 12 (2003), 115–121.
  • [Borcherds 92] R. E. Borcherds. “Monstrous Moonshine and Monstrous Lie Superalgebras.” Invent. Math. 109 (1992), 405–444.
  • [Conway and Norton 79] J. H. Conway and S. P. Norton. “Monstrous Moonshine.” Bull. Lond. Math. Soc. 11 (1979), 308–339.
  • [Cummins 04] C. J. Cummins. “Congruence Subgroups of Groups Commensurable with of Genus 0 and 1.” Exp. Math. 13 (2004), 361–382.
  • [Edixhoven and Couveignes 11] B. Edixhoven and J.-M. Couveignes (eds.). Computational Aspects of Modular Forms and Galois Representations. Annals of Mathematics Studies, vol. 176. Princeton, NJ: Princeton University Press, 2011.
  • [Frenkel et al. 88] I. Frenkel, J. Lepowsky, and A. Meurman. Vertex Operator Algebras and the Monster. San Diego, CA: Academic Press, 1988.
  • [Gannon 06a] T. Gannon. “Monstrous Moonshine: The First Twenty-Five Years.” Bull. Lond. Math. Soc. 38 (2006), 1–33.
  • [Gannon 06b] T. Gannon. Moonshine Beyond the Monster. The Bridge Connecting Algebra, Modular Forms and Physics. Cambridge: Cambridge Monographs on Mathematical Physics, Cambridge University Press, 2006.
  • [Granlund and the GMP development team 12] T. Granlund and the GMP development team. “GNU MP: The GNU Multiple Precision Arithmetic Library, Version 5.0.5.′ Free Software Foundation . Available at http://gmplib.org/, 2012.
  • [Gross and Zagier 85] B. H. Gross and D. B. Zagier. “On Singular Moduli.” J. Reine Angew. Math. 355 (1985), 191–220.
  • [Griess 82] R. L. Griess, Jr,. “The Friendly Giant.” Invent. Math. 69 (1982), 1–102.
  • [Iwaniec 02] H. Iwaniec. Spectral Methods of Automorphic Forms. Graduate Studies in Mathematics, vol. 53. Providence, RI: AMS, 2002.
  • [Iwaniec and Kowalski 04] H. Iwaniec and E. Kowalski. Analytic Number Theory. AMS Colloquium Publications, vol. 53. Providence, RI: AMS, 2004.
  • [Jorgenson et al. 14] J. Jorgenson, L. Smajlović, and H. Then. “On the Distribution of Eigenvalues of Maass Forms on Certain Moonshine Groups.” Math. Comp. 83 (2014), 3039–3070.
  • [Jorgenson et al. preprint-a] J. Jorgenson, L. Smajlović, and H. Then. “Certain Aspects of Holomorphic Function Theory on Genus Zero Arithmetic Groups.” submitted, arXiv: 1505.06042.
  • [Jorgenson et al. in preparation-b] J. Jorgenson, L. Smajlović, and H. Then. “Relating Weierstrass ℘-Functions to Holomorphic Eisenstein Series on Certain Arithmetic Groups of Genus One.” in preparation.
  • [Jorgenson et al. in preparation-c] J. Jorgenson, L. Smajlović, and H. Then. “On the Algebraicity of Hauptmoduli Associated to some Genus Zero Arithmetic Groups at Elliptic Fixed Points.” in preparation.
  • [Knopp 90] M. I. Knopp. “Rademacher on J(τ), Poincaré Series of Nonpositive Weights and the Eichler Cohomology.” Notices Amer. Math. Soc. 37 (1990), 385–393.
  • [Kohnen 03] W. Kohnen. “Weierstrass Points at Cusps on Special Modular Curves.” Abh. Math. Sem. Univ. Hamburg 73 (2003), 241–251.
  • [Koyama 09] S. Koyama. “Equidistribution of Eisenstein Series in the Level Aspect.” Commun. Math. Phys. 289 (2009), 1131–1150.
  • [Kühn 01] U. Kühn. “Generalized Arithmetic Intersection Numbers.” J. Reine Angew. Math. 534 (2001), 209–236.
  • [Lang 76] S. Lang. Introduction to Modular Forms. Grundlehren der mathematischen Wissenschaften, vol. 222. New York, NY: Springer-Verlag, 1976.
  • [Lang 82] S. Lang. Introduction to Algebraic and Abelian Functions. Graduate Texts in Mathematics, vol. 89. New York, NY: Springer-Verlag, 1982.
  • [Maclachlan 81] C. Maclachlan. “Groups of Units of Zero Ternary Quadratic Forms.” Proc. R. Soc. Edinburgh 88A (1981), 141–157.
  • [Rademacher 38] H. Rademacher. “The Fourier Coefficients of the Modular Invariant J(τ). Amer. J. Math. 60 (1938), 501–512.
  • [Raji 06] W. Raji. “Modular Forms Representable as Eta Products.” PhD thesis, Temple University, 2006.
  • [Serre 73] J.-P. Serre. A Course in Arithmetic. Graduate Texts in Mathematics, vol. 7. New York, NY: Springer-Verlag, 1973.
  • [Sloane 10] N. J. A. Sloane. “The On-Line Encyclopedia of Integer Sequences.” Available online http://oeis.org/, 2010.
  • [The PARI Group 11] The PARI Group. “PARI/GP, Version 2.5.0.” Bordeaux. Available at http://pari.math.u-bordeaux.fr/, 2011.
  • [Thompson 79a] J. G. Thompson. “Finite Groups and Modular Functions.” Bull. Lond. Math. Soc. 11 (1979), 347–351.
  • [Thompson 79b] J. G. Thompson. “Some Numerology Between the Fischer-Griess Monster and the Elliptic Modular Function.” Bull. Lond. Math. Soc. 11 (1979), 352–353.
  • [Zagier 08] D. B. Zagier. “Elliptic Modular Forms and Their Applications.” In K. Ranestad (Ed.) The 1-2-3 of Modular Forms, 1–103. Berlin: Universitext, Springer, 2008.