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

A note on transcendence of special values
of functions related to modularity

Tapas Bhowmik  and  Siddhi Pathak Department of Mathematics, University of South Carolina, 1523 Greene Street, LeConte College, Columbia, South Carolina, USA 29208. Chennai Mathematical Institute, H-1 SIPCOT IT Park, Siruseri, Kelambakkam, Tamil Nadu, India 603103. [email protected] [email protected]
Abstract.

In this note, we study the arithmetic nature of values of modular functions, meromorphic modular forms and meromorphic quasi-modular forms with respect to arbitrary congruence subgroups, that have algebraic Fourier coefficients. This approach unifies many of the known results, and leads to generalizations of the theorems of Schneider, Nesterenko and others.

Key words and phrases:
Values of modular functions, values of modular forms, Nesterenko’s theorem, Schneider’s theorem
2010 Mathematics Subject Classification:
11J91, 11F03, 11F11
Research of the second author was partially supported by an INSPIRE Faculty fellowship.

1. Introduction


The study of the arithmetic nature of values of special transcendental functions at algebraic arguments has been a well-established theme in number theory. Continuing in this spirit, this note focuses on the transcendental nature and algebraic independence of values of functions arising in the modular world, such as modular functions, modular forms and quasi-modular forms (see Section 2 for definition of the functions appearing below). The genesis of this study can be traced back to a 1937 theorem of Schneider [15], namely,

Theorem 1.1 (Schneider).

If τ\tau\in\mathbb{H} is algebraic but not imaginary quadratic, then j(τ)j(\tau) is transcendental.

It is known from the theory of complex multiplication that if τ\tau\in\mathbb{H} generates an imaginary quadratic field (τ\tau is a CM point), then j(τ)j(\tau) is an algebraic number. Therefore, Schneider’s theorem translates to the statement: if τ\tau is algebraic, then j(τ)j(\tau) is algebraic if and only if τ\tau is CM. A further conjecture by Mahler [9] in this regard, proved by Barré-Sirieix, Diaz, Gramain and Philibert [2] states that

Theorem 1.2 (Barré-Sirieix, Diaz, Gramain, Philibert).

For any τ\tau\in\mathbb{H}, at least one of the two numbers e2πiτe^{2\pi i\tau} and j(τ)j(\tau) is transcendental.

This can be derived as a consequence of a remarkable theorem of Nesterenko [13] with several applications.

Theorem 1.3 (Nesterenko).

If τ\tau\in\mathbb{H}, then at least three of the numbers

e2πiτ,E2(τ),E4(τ),E6(τ)e^{2\pi i\tau},\quad E_{2}(\tau),\quad E_{4}(\tau),\quad E_{6}(\tau)

are algebraically independent over ¯\overline{\mathbb{Q}}.

Although the above theorems are about specific functions of “level 11”, it is the aim of this note to highlight that they are sufficient to deduce the corresponding results for functions associated with arbitrary congruence subgroups. The authors believe that this fact may be known to experts, but is not well-documented. Often, the congruence subgroup in question is restricted to be the group Γ0(N)\Gamma_{0}(N). The results in this paper apply to functions satisfying appropriate modularity properties with respect to Γ(N)\Gamma(N), and hence, arbitrary congruence subgroups.

A detailed investigation of algebraic independence of values of modular forms and quasi-modular forms was carried out by S. Gun, M. R. Murty and P. Rath [12] in 2011. Their results on values of modular forms were further elaborated upon for higher level in [7] by A. Hamieh and M. R. Murty111A small correction in their statement of Theorem 1.2 is required. The conclusion should read as (π/ωτ)kLq(k,χ){(\pi/\omega_{\tau})}^{k}\,L_{q}(k,\chi) is algebraic, without the L(1k,χ)L(1-k,\chi) term.. These theorems will follow from our discussion later. In the context of quasimodular forms, it was proven independently by Gun, Murty and Rath [12, Theorem 7] and C. Y. Chang [3] that

Theorem 1.4 (Gun-Murty-Rath and Chang).

If ff is a quasi-modular form of non-zero weight for SL2()SL_{2}(\mathbb{Z}) with algebraic Fourier coefficients and τ\tau\in\mathbb{H} is such that j(τ)¯j(\tau)\in\overline{\mathbb{Q}}, then either f(τ)=0f(\tau)=0 or f(τ)f(\tau) is transcendental.

The above statement is also proved for quasimodular forms with respect to Γ0(N)\Gamma_{0}(N) in [12].

Another instance of investigation is a recent paper of D. Jeon, S.-Y. Kang and C. H. Kim [5, Theorem 2.4], where they prove the following. Let NN\in\mathbb{N}, 𝔤:=𝔤0(N)=\mathfrak{g}:=\mathfrak{g}_{0}(N)= genus of X0(N)X_{0}(N), the modular curve obtained as the quotient of the extended upper half plane by Γ0(N)\Gamma_{0}(N). Suppose that 𝔤>0\mathfrak{g}>0 and m𝔤+1m\geq\mathfrak{g}+1 is an integer. Let 𝔣N,m\mathfrak{f}_{N,m} denote the unique modular function with respect to Γ0(N)\Gamma_{0}(N) constructed in [6] such that 𝔣N,m\mathfrak{f}_{N,m} is holomorphic on the upper half plane and

𝔣N,m(q)=1qm+l=1𝔤aN(m,l)1ql+O(q),\mathfrak{f}_{N,m}(q)=\frac{1}{q^{m}}+\sum_{l=1}^{\mathfrak{g}}a_{N}(m,-l)\frac{1}{q^{l}}+O(q),

with the coefficients of powers of qq being algebraic numbers. Then, they show the following.

Theorem 1.5 (Jeon, Kang, Kim).

Let ff be a non-zero meromorphic modular form with respect to Γ0(N)\Gamma_{0}(N) with algebraic Fourier coefficients. If τ\tau is either a zero or a pole of ff, then 𝔣N,m(τ)\mathfrak{f}_{N,m}(\tau) is algebraic for all mm.

As a corollary, they deduce that any zero or pole of ff should be either CM or transcendental.

In this note, we first give an exposition of the algebraic structure of modular functions of higher level, following Shimura [17]. Building upon this and using Schneider’s theorem, we prove

Theorem 1.6.

Let gg be a non-constant modular function with respect to a congruence subgroup Γ\Gamma of level NN, with algebraic Fourier coefficients at ii\infty.

  1. (a)

    If τ\tau\in\mathbb{H} is either a zero or a pole of gg, then j(τ)j(\tau) is algebraic, and hence, τ\tau is either CM or transcendental.

  2. (b)

    If τ\tau\in\mathbb{H} is such that τ\tau is not a pole of gg, then j(τ)¯g(τ)¯j(\tau)\in\overline{\mathbb{Q}}\Leftrightarrow g(\tau)\in\overline{\mathbb{Q}}. Thus, at least one of g(τ)g(\tau) and e2πiτe^{2\pi i\tau} is transcendental for any τ\tau\in\mathbb{H}.

In the context of meromorphic modular forms, we establish a generalization of Theorem 1.5 to arbitrary modular forms and arbitrary modular functions.

Theorem 1.7.

Let ff be a non-constant meromorphic modular form of weight kk\in\mathbb{Z} with respect to a congruence subgroup of level NN and gg be a non-constant modular function with respect to a congruence subgroup of level MM. Suppose that both ff and gg have algebraic Fourier coefficients at ii\infty. Let τ\tau\in\mathbb{H} be a zero or a pole of ff. Then either g(z)g(z) has a pole at z=τz=\tau or g(τ)g(\tau) is algebraic.

Furthermore, we generalize the theorems in [12] and [7] to the setting of meromorphic modular forms. In particular, we show the following.

Theorem 1.8.

Let ff be a non-constant meromorphic modular form with respect to a congruence subgroup Γ\Gamma of level NN. Suppose that ff has algebraic Fourier coefficients at ii\infty. Suppose that τ\tau is not a pole of ff.

  1. (a)

    If τ\tau\in\mathbb{H} is such that e2πiτe^{2\pi i\tau} is algebraic, then f(τ)f(\tau) is transcendental.

  2. (b)

    If τ\tau\in\mathbb{H} is such that j(τ)¯j(\tau)\in\overline{\mathbb{Q}}, then there exists a transcendental number ωτ\omega_{\tau} which depends only on τ\tau and is ¯\overline{\mathbb{Q}}-linearly independent with π\pi, such that (πωτ)kf(τ)¯\left(\dfrac{\pi}{\omega_{\tau}}\right)^{k}f(\tau)\in\overline{\mathbb{Q}}. Therefore, f(τ)f(\tau) is either zero or transcendental.

With regard to quasi-modular forms, we extend the previously known results to meromorphic quasi-modular forms and prove

Theorem 1.9.

Let f~\widetilde{f} be a non-constant meromorphic quasi-modular form with depth p1p\geq 1, with respect to a congruence subgroup Γ\Gamma of level NN. Suppose that f~\widetilde{f} has algebraic Fourier coefficients at ii\infty and that τ\tau is not a pole of f~\widetilde{f}.

  1. (a)

    If τ\tau\in\mathbb{H} is such that e2πiτe^{2\pi i\tau} is algebraic, then f~(τ)\widetilde{f}(\tau) is transcendental.

  2. (b)

    Let

    f~=r=0pfrE2r with frMk2r,N,¯m.\widetilde{f}=\sum_{r=0}^{p}f_{r}\,E_{2}^{r}\qquad\text{ with }\qquad f_{r}\in M^{m}_{k-2r,N,\overline{\mathbb{Q}}}.

    Here Mj,N,¯mM^{m}_{j,N,\overline{\mathbb{Q}}} denotes the space of meromorphic modular forms of weight jj, level NN and algebraic Fourier coefficients at ii\infty. If τ\tau\in\mathbb{H} is such that j(τ)¯j(\tau)\in\overline{\mathbb{Q}}, then

    f~(τ)=0fr(τ)=0 for all 0rp.\widetilde{f}(\tau)=0\qquad\iff\qquad f_{r}(\tau)=0\text{ for all }0\leq r\leq p.

    Moreover, if f~(τ)0\widetilde{f}(\tau)\neq 0, then f~(τ)\widetilde{f}(\tau) is transcendental.

This generalizes Theorem 1.4 as well as [1, Theorem 1.8].

Let 𝒵j,¯:={τ:j(τ)¯}\mathcal{Z}_{j,\overline{\mathbb{Q}}}:=\left\{\tau\in\mathbb{H}\,:\,j(\tau)\in\overline{\mathbb{Q}}\right\}. From the above results, we deduce the following interesting corollary.

Corollary 1.10.

Let 𝒵mdfn\mathcal{Z}_{\text{mdfn}} be the set of zeros and poles of modular functions of arbitrary level with algebraic Fourier coefficients, 𝒵mdfrm\mathcal{Z}_{\text{mdfrm}} be the set of zeros and poles of meromorphic modular forms of arbitrary level with algebraic Fourier coefficients and 𝒵quasi-mdf\mathcal{Z}_{\text{quasi-mdf}} be the set of zeros and poles of meromorphic quasi-modular forms of arbitrary level with algebraic Fourier coefficients. Then

𝒵quasi-mdf𝒵mdfrm𝒵mdfn𝒵j,¯.\mathcal{Z}_{\text{quasi-mdf}}\subseteq\mathcal{Z}_{\text{mdfrm}}\subseteq\mathcal{Z}_{\text{mdfn}}\subseteq\mathcal{Z}_{j,\overline{\mathbb{Q}}}.

In particular, zeros and poles of quasi-modular forms, modular forms and modular functions are either CM or transcendental.

Finally, we have the following generalization of Theorem 1.3.

Theorem 1.11.

Let ff be a non-constant meromorphic modular form of weight kk\in\mathbb{Z} with respect to a congruence subgroup of level NN, gg be a non-constant modular function with respect to a congruence subgroup of level MM and f~\widetilde{f} be a non-constant meromorphic quasi-modular function of depth at least 11, with respect to a subgroup of level N~\widetilde{N}. Suppose that ff, gg and f~\widetilde{f} have algebraic Fourier coefficients at ii\infty. If τ\tau\in\mathbb{H} is such that f(τ)0f(\tau)\neq 0, f~(τ)0\widetilde{f}(\tau)\neq 0 and τ\tau is not a pole of ff, gg and f~\widetilde{f}, then

trdeg¯(e2πiτ,f(τ),g(τ),f~(τ))3.\operatorname{trdeg}_{\,\mathbb{Q}}\,\,\overline{\mathbb{Q}}\left(e^{2\pi i\tau},\,f(\tau),\,g(\tau),\,\widetilde{f}(\tau)\right)\geq 3.

This theorem is in the same spirit as [19, Theorem 1.2], where W. Wang considers the algebraic independence of the values of three algebraically independent quasi-modular forms. However, Theorem 1.11 allows one to also compare values of modular functions with those of quasi-modular forms. It can be shown that a modular function, a modular form of non-zero weight and a quasi-modular form of positive depth are algebraically independent. We include a proof of this assertion in Theorem 2.14 for completeness.

We also remark that in Theorem 1.7 and Theorem 1.11, one can replace a meromorphic modular form by a half-integer weight modular form with algebraic Fourier coefficients since the square of a half-integer weight modular form is an integer weight modular form.


2. Preliminaries


The aim of this section is to study the algebraic structure of the field of modular functions and to record the required results from transcendental number theory. For the sake of completeness and clarity of exposition, a brief account of the proofs is included, and appropriate references are given.

2.1. Modular and quasi-modular forms

For each NN\in\mathbb{N},

Γ(N):={(abcd)SL2():(abcd)(1001)(modN)},\Gamma(N):=\left\{\left(\begin{array}[]{cc}a&b\\ c&d\end{array}\right)\in SL_{2}(\mathbb{Z}):\left(\begin{array}[]{cc}a&b\\ c&d\end{array}\right)\equiv\left(\begin{array}[]{cc}1&0\\ 0&1\end{array}\right)(\operatorname{mod}N)\right\},

with Γ(1)=SL2()\Gamma(1)=SL_{2}(\mathbb{Z}). A subgroup Γ\Gamma of SL2()SL_{2}(\mathbb{Z}) is said to be congruence subgroup if there exists NN\in\mathbb{N} such that Γ(N)Γ\Gamma(N)\subseteq\Gamma. The smallest such NN is said to be the level of Γ\Gamma.

A holomorphic modular form of integer weight k0k\geq 0 with respect to a congruence subgroup Γ\Gamma is a holomorphic function on the upper half-plane \mathbb{H} which satisfies

  1. (i)

    f|kγ=ff\big{|}_{k}\gamma=f for all γΓ\gamma\in\Gamma

  2. (ii)

    f|kαf\big{|}_{k}\alpha is holomorphic at ii\infty for all αSL2()\alpha\in SL_{2}(\mathbb{Z}).

The function ff is said to be a weakly holomorphic modular form if f|kαf\big{|}_{k}\alpha is allowed to have poles at ii\infty, that is, ff is meromorphic at the cusps. More generally, ff is called a meromorphic modular form if it is meromorphic on \mathbb{H} and also at cusps.

We will say that ff has algebraic Fourier coefficients if the Fourier coefficients of f|kαf\big{|}_{k}\alpha at ii\infty are algebraic numbers. Denote the space of all holomorphic, weakly holomorphic and meromorphic modular forms, with algebraic Fourier coefficients by Mk,¯(Γ)M_{k,\overline{\mathbb{Q}}}(\Gamma), Mk,¯w(Γ)M_{k,\overline{\mathbb{Q}}}^{w}(\Gamma) and Mk,¯m(Γ)M_{k,\overline{\mathbb{Q}}}^{m}(\Gamma) respectively. Clearly, Mk,¯(Γ)Mk,¯w(Γ)Mk,¯m(Γ)M_{k,\overline{\mathbb{Q}}}(\Gamma)\subset M_{k,\overline{\mathbb{Q}}}^{w}(\Gamma)\subset M_{k,\overline{\mathbb{Q}}}^{m}(\Gamma).

For even integer k2k\geq 2, define the normalized Eisenstein series of weight kk for SL2()SL_{2}(\mathbb{Z}) by

Ek(τ)=12kBkn=1σk1(n)qn, where q=e2πiτ and σs(n)=d|nd>0ds.E_{k}(\tau)=1-\frac{2k}{B_{k}}\sum_{n=1}^{\infty}\sigma_{k-1}(n)q^{n},\text{ where }q=e^{2\pi i\tau}\text{ and }\sigma_{s}(n)=\sum_{\begin{subarray}{c}d|n\\ d>0\end{subarray}}d^{s}.

Here, BkB_{k} is the kk-th Bernoulli number. Define

Δ(τ):=E4(τ)3E6(τ)21728.\Delta(\tau):=\frac{E_{4}(\tau)^{3}-E_{6}(\tau)^{2}}{1728}.

For k4k\geq 4, the function EkMk,¯(SL2())E_{k}\in M_{k,\overline{\mathbb{Q}}}(SL_{2}(\mathbb{Z})) and ΔM12,¯(SL2())\Delta\in M_{12,\overline{\mathbb{Q}}}(SL_{2}(\mathbb{Z})). But E2(τ)E_{2}(\tau) is not a modular form (see [10], Chapter 55), as

E2(1τ)=τ2E2(τ)+6iπτ.E_{2}\left(\frac{-1}{\tau}\right)={\tau}^{2}E_{2}(\tau)+\frac{6}{i\pi}\tau.

This motivates the definition of a quasi-modular form, of which there are several equivalent formulations. We use the following characterization, which was established for holomorphic quasi-modular forms in [20, Proposition 20] and meromorphic quasi-modular forms in [18, Theorem 4.2].

Theorem 2.1.

Every meromorphic quasi-modular form for a congruence subgroup Γ\Gamma is a polynomial in E2E_{2} with modular coefficients. More precisely, if f~\widetilde{f} is a meromorphic quasi-modular form of weight kk and depth pp with respect to Γ\Gamma, then f~\widetilde{f} can be uniquely written as f~=r=0pfrE2r\widetilde{f}=\sum_{r=0}^{p}f_{r}\,E_{2}^{r}, where frf_{r} is a meromorphic modular form with respect to Γ\Gamma of weight k2rk-2r for all 0rp0\leq r\leq p and fp0f_{p}\neq 0.

A quasi-modular form f~\widetilde{f} is said to have algebraic Fourier coefficients if all the modular coefficients in the above expression of ff have algebraic Fourier coefficients.

2.2. The Weierstrass \wp-function

Let L=ω1ω2L=\omega_{1}\mathbb{Z}\,\oplus\,\omega_{2}\mathbb{Z} be a two-dimensional lattice in \mathbb{C} with ω1/ω2\omega_{1}/\omega_{2}\in\mathbb{H}. The Weierstrass \wp-function associated to LL, given by

(z)=1z2+ωLω0(1(zω)21ω2),for zL,\wp(z)=\frac{1}{z^{2}}+\sum_{\begin{subarray}{c}\omega\in L\\ \omega\neq 0\end{subarray}}\left(\frac{1}{(z-\omega)^{2}}-\frac{1}{\omega^{2}}\right),\quad\text{for }z\in\mathbb{C}\setminus L, (1)

defines a meromorphic function on \mathbb{C}. It satisfies the differential equation

(z)2=4(z)3g2(L)(z)g3(L),\wp^{\prime}(z)^{2}=4\wp(z)^{3}-g_{2}(L)\wp(z)-g_{3}(L), (2)

where

g2(L)=60ωLω01ω4andg3(L)=140ωLω01ω6.g_{2}(L)=60\,\sum_{\begin{subarray}{c}\omega\in L\\ \omega\neq 0\end{subarray}}\frac{1}{\omega^{4}}\qquad\text{and}\qquad g_{3}(L)=140\,\sum_{\begin{subarray}{c}\omega\in L\\ \omega\neq 0\end{subarray}}\frac{1}{\omega^{6}}.

For the lattice Lτ=τL_{\tau}=\tau\,\mathbb{Z}\oplus\mathbb{Z} with τ\tau\in\mathbb{H},

g2(Lτ)=4π43E4(τ)andg3(Lτ)=8π627E6(τ).g_{2}(L_{\tau})=\frac{4\pi^{4}}{3}E_{4}(\tau)\qquad\text{and}\qquad g_{3}(L_{\tau})=\frac{8\pi^{6}}{27}E_{6}(\tau).

Define the discriminant of a lattice,

Δ0(L):=g2(L)327g3(L)20, for any two dimensional lattice L.\Delta_{0}(L):=g_{2}(L)^{3}-27g_{3}(L)^{2}\neq 0,\text{ for any two dimensional lattice }L.

In particular, we have Δ0(Lτ)=(2π)12Δ(τ)\Delta_{0}(L_{\tau})=(2\pi)^{12}\,\Delta(\tau) for all τ\tau\in\mathbb{H}.

The Weierstrass zeta-function associated to L=ω1ω2L=\omega_{1}\,\mathbb{Z}\oplus\omega_{2}\,\mathbb{Z} is defined by

ζL(z)=1z+ωLω0(1zω+1ω+zω2)for zL.\zeta_{L}(z)=\frac{1}{z}+\sum_{\begin{subarray}{c}\omega\in L\\ \omega\neq 0\end{subarray}}\left(\frac{1}{z-\omega}+\frac{1}{\omega}+\frac{z}{\omega^{2}}\right)\quad\text{for }z\in\mathbb{C}\setminus L.

Note that ζL(z)=L(z)\zeta_{L}^{\prime}(z)=-\wp_{L}(z) is a periodic function with each point of LL as a period. Hence, the functions

ζL(z+ω1)ζL(z) and ζL(z+ω2)ζL(z)\zeta_{L}(z+\omega_{1})-\zeta_{L}(z)\text{ and }\zeta_{L}(z+\omega_{2})-\zeta_{L}(z)

are constants. These constants are denoted by η1(L)\eta_{1}(L) and η2(L)\eta_{2}(L) respectively and are called quasi-periods. Moreover, for ω1=τ\omega_{1}=\tau and ω2=1\omega_{2}=1, it is known that

η2(Lτ)=G2(τ)=π23E2(τ) for all τ.\eta_{2}(L_{\tau})=G_{2}(\tau)=\frac{\pi^{2}}{3}\,E_{2}(\tau)\text{ for all }\tau\in\mathbb{H}. (3)

The reader is referred to ([8], chapter 18.318.3) for proof of these results.

The Weierstrass \wp-function satisfies the following addition formula ([11], chapter 1111).

Theorem 2.2.

For z1,z2z_{1},z_{2}\in\mathbb{C} such that z1±z2Lz_{1}\pm z_{2}\notin L we have

(z1+z2)=(z1)(z2)+14((z1)(z2)(z1)(z2))2.\wp\left(z_{1}+z_{2}\right)=-\wp\left(z_{1}\right)-\wp\left(z_{2}\right)+\frac{1}{4}\left(\frac{\wp^{\prime}\left(z_{1}\right)-\wp^{\prime}\left(z_{2}\right)}{\wp\left(z_{1}\right)-\wp\left(z_{2}\right)}\right)^{2}.

Using the addition formula of the Weierstrass \wp-function, Schneider [16] proved that

Theorem 2.3.

Let L=ω1ω2L=\omega_{1}\,\mathbb{Z}\oplus\omega_{2}\,\mathbb{Z} be a lattice such that both g2(L)g_{2}(L), g3(L)g_{3}(L) are algebraic. If α\alpha is an algebraic number with αL\alpha\notin L, then (α)\wp(\alpha) is transcendental.

The addition formula also implies the following important proposition (see [11]). Two generators ω1\omega_{1} and ω2\omega_{2} of a lattice LL are said to be primitive if both have minimal absolute value among all generators of LL.

Proposition 2.4.

Let L=ω1ω2L=\omega_{1}\,\mathbb{Z}\oplus\omega_{2}\,\mathbb{Z} be such that both g2(L)g_{2}(L) and g3(L)g_{3}(L) are algebraic. Assume that ω1\omega_{1} and ω2\omega_{2} are primitive generators of LL. Then, for any natural number n>1n>1, the numbers (ω1n)\wp(\frac{\omega_{1}}{n}) and (ω2n)\wp(\frac{\omega_{2}}{n}) are algebraic. Moreover, any non-zero period of LL is necessarily transcendental.

2.3. Modular functions

A meromorphic function gg on \mathbb{H} is said to be a modular function if it satisfies

g(aτ+bcτ+d)=g(τ)for all γ=(abcd)Γ,g\left(\frac{a\,\tau+b}{c\,\tau+d}\right)=g(\tau)\quad\text{for all }\gamma=\left(\begin{smallmatrix}a&b\\ c&d\end{smallmatrix}\right)\in\Gamma, (4)

that is, g|0γ=gg\big{|}_{0}\gamma=g, and is also meromorphic at all the cusps. In particular, we call a modular function on the congruence subgroup Γ(N)\Gamma(N) to a be modular function of level NN. Note that if gg is a modular function with respect to Γ\Gamma, which is of level NN, then gg is a modular function of level NN.

An example of modular function of level one is given by

j(τ):=E43(τ)Δ(τ) for all τ,j(\tau):=\frac{E_{4}^{3}(\tau)}{\Delta(\tau)}\text{ for all }\tau\in\mathbb{H},

which has the following Fourier expansion at ii\infty

j(τ)=1q+744+196884q+,j(\tau)=\frac{1}{q}+744+196884q+\cdots,

where q=e2πiτq=e^{2\pi i\tau}.

More specifically, for any modular function (or form) ff, define

(f):=({Fourier coefficients of f at all Γ-in-equivalent cusps})\mathbb{Q}(f):=\mathbb{Q}\left(\left\{\text{Fourier coefficients of $f$ at all $\Gamma$-in-equivalent cusps}\right\}\right)

The jj-function is a canonical example of a modular function of level one, and governs properties of modular functions of higher levels as well. This is made precise in the following series of propositions.

Proposition 2.5.

Let gg be a non-constant modular function of level one. Let \mathcal{F} denote the standard fundamental domain for the action of SL2()SL_{2}(\mathbb{Z}) on \mathbb{H}. Suppose that the poles of gg in \mathcal{F} are τ1,τ2,,τm\tau_{1},\tau_{2},\cdots,\tau_{m}. Then g(τ)g(\tau) is a rational function in j(τ)j(\tau) with coefficients in the field (g)(j(τ1),,j(τm))\mathbb{Q}(g)\left(j(\tau_{1}),\cdots,j(\tau_{m})\right).

Proof.

By the compactness of i\mathcal{F}\cup i\infty, we know that gg has only finitely many poles in \mathcal{F}. Consider the function

h(τ):=(j=1m(j(τ)j(τj))ordτj(g))g(τ),h(\tau):=\left(\prod_{j=1}^{m}\left(j(\tau)-j(\tau_{j})\right)^{-\operatorname{ord}_{\tau_{j}}(g)}\right)g(\tau),

where ordτj(g)=\operatorname{ord}_{\tau_{j}}(g)=- order of the pole of gg at τj\tau_{j}. Then hh is holomorphic on \mathbb{H}.

Suppose hh has pole of order M1M\geq 1 at ii\infty. Then the Fourier expansion of h(τ)h(\tau) at ii\infty has the form

h(τ)=n=Mcnqn, where cM0.h(\tau)=\sum_{n=-M}^{\infty}c_{n}q^{n},\,\text{ where }c_{-M}\neq 0.

Note that the modular function h(τ)cMj(τ)Mh(\tau)-c_{-M}\,j(\tau)^{M} is holomorphic on \mathbb{H} and its Fourier expansion starts with at most a polar term of order M1M-1. Iterating this process, we can subtract a polynomial in j(τ)j(\tau) to get a holomorphic modular function that vanishes at ii\infty, and hence is identically zero. Thus, h(τ)h(\tau) is a polynomial in j(τ)j(\tau) over (g)\mathbb{Q}(g), and so g(τ)g(\tau) is a rational function of j(τ)j(\tau) over (g)(j(τ1),,j(τm))\mathbb{Q}(g)(j(\tau_{1}),\cdots,j(\tau_{m})). ∎

One can also conclude the following important fact from the above proof.

Corollary 2.6.

If gg is a modular function of level one which is holomorphic on \mathbb{H} with a pole of order MM at ii\infty, then g(τ)g(\tau) is a polynomial in j(τ)j(\tau) of degree MM with coefficients in (g)\mathbb{Q}(g).

The jj-function is sufficient to ‘generate’ all higher level modular functions as well. This is proved below.

Theorem 2.7.

Let gg be a modular function with respect to a congruence subgroup Γ\Gamma and let τ1,τ2,,τm\tau_{1},\,\tau_{2},\cdots,\,\tau_{m} be the poles of gg in \mathcal{F}. Set

K:=(g)(j(τ1),j(τ2),,j(τm)).K:=\mathbb{Q}(g)\left(j(\tau_{1}),\,j(\tau_{2}),\cdots,\,j(\tau_{m})\right).

Then there exists a monic polynomial Pg(X)K(j)[X]P_{g}(X)\in K(j)[X] such that Pg(g)=0P_{g}(g)=0.

Proof.

Let [SL2():Γ]=r[SL_{2}(\mathbb{Z}):\Gamma]=r and {γ1=I,γ2,,γr}\{\gamma_{1}=I,\gamma_{2},\cdots,\gamma_{r}\} be a complete set of right coset representatives so that

SL2()=i=1rΓγi.SL_{2}(\mathbb{Z})=\bigsqcup_{i=1}^{r}\Gamma\gamma_{i}.

For all 1ir1\leq i\leq r, define the functions

gi(τ):=g(γiτ)for τ.g_{i}(\tau):=g(\gamma_{i}\tau)\quad\text{for }\tau\in\mathbb{H}.

Each gig_{i} is independent of the choice of coset representatives as gg is Γ\Gamma-invariant. Moreover, each gig_{i} is a modular function with respect to Γ\Gamma and the Fourier expansion of gig_{i} at ii\infty is precisely the expansion of gg at the cusp γi(i)\gamma_{i}(i\infty). For any γSL2()\gamma\in SL_{2}(\mathbb{Z}), we have gi(γτ)=g(γiγτ)=gj(τ)g_{i}(\gamma\tau)=g(\gamma_{i}\gamma\tau)=g_{j}(\tau) for some jj with 1jr1\leq j\leq r such that γiγΓγj\gamma_{i}\gamma\in\Gamma\gamma_{j}, i.e., the set {g1,g2,,gr}\{g_{1},g_{2},...,g_{r}\} gets permuted under the action of SL2().SL_{2}(\mathbb{Z}). This observation implies that any elementary symmetric polynomial in g1,g2,,grg_{1},g_{2},\cdots,g_{r} is a modular function of level one, and is in K(j)K(j) by Theorem 2.5. Note that the polynomial

P(X)=i=1r(Xgi)P(X)=\prod_{i=1}^{r}(X-g_{i})

is satisfied by gg as g=g1g=g_{1} and has coefficients that are elementary symmetric polynomials in g1,g2,,grg_{1},g_{2},\cdots,g_{r}. This proves the theorem. ∎

Corollary 2.8.

Let gg be a modular function with respect to a congruence subgroup Γ\Gamma which is holomorphic on \mathbb{H} and K=(g)K=\mathbb{Q}(g). Then the monic polynomial Pg(X)P_{g}(X) satisfied by gg has coefficients in K[j][X]K[j][X].

Proof.

The coefficients of Pg(X)P_{g}(X) constructed in the proof above are modular functions of level one, holomorphic on \mathbb{H} with Fourier coefficients in KK. The result now follows from Corollary 2.6. ∎

Remark.

An important point to note here is that if gg is a modular function with respect Γ\Gamma that has algebraic Fourier coefficients, Theorem 2.7 does not guarantee that the coefficients of PgP_{g} are algebraic, unless gg is holomorphic in \mathbb{H}.

To that effect, we now study the structure of the field of modular functions of a fixed level N>1N>1. For any field 𝔽\mathbb{F}\subseteq\mathbb{C}, let

N,𝔽:={Modular functions of level N whose Fourier coefficients at i are in 𝔽}.\mathcal{F}_{N,\mathbb{F}}:=\left\{\text{Modular functions of level $N$ whose Fourier coefficients at $i\infty$ are in $\mathbb{F}$}\right\}.

Theorem 2.7 implies that N,\mathcal{F}_{N,\mathbb{C}} is an algebraic extension of (j)\mathbb{C}(j).

Following [17, Chapter 6], we consider explicit modular function of level NN whose Fourier coefficients have good rationality properties. Let 𝒂=(a1,a2)1N22\boldsymbol{a}=(a_{1},a_{2})\in\frac{1}{N}\mathbb{Z}^{2}\,\setminus\,\mathbb{Z}^{2}, consider the function

f𝒂(τ):=g2(Lτ)g3(Lτ)Δ0(Lτ)Lτ(a1τ+a2),f_{\boldsymbol{a}}(\tau):=\frac{g_{2}(L_{\tau})\,g_{3}(L_{\tau})}{\Delta_{0}(L_{\tau})}\,\wp_{L_{\tau}}(a_{1}\tau+a_{2}),

which is holomorphic on \mathbb{H}. The following properties can be checked routinely, and we leave the proof to the reader.

Proposition 2.9.

Let SS denotes the set {(rN,sN):0r,sN1 and (r,s)(0,0)}\{\left(\frac{r}{N},\frac{s}{N}\right):0\leq r,s\leq N-1\text{ and }(r,s)\neq(0,0)\}. Let 𝐚\boldsymbol{a}, 𝐛1N22\boldsymbol{b}\in\frac{1}{N}\mathbb{Z}^{2}\,\setminus\,\mathbb{Z}^{2} and f𝐚f_{\boldsymbol{a}} be as defined above. For γ=(pqrs)\gamma=\big{(}\begin{smallmatrix}p&q\\ r&s\end{smallmatrix}\big{)} and 𝐚=(a1,a2)\boldsymbol{a}=(a_{1},a_{2}), let 𝐚𝛄=(pa1+ra2,qa1+sa2){\boldsymbol{a\gamma}=(pa_{1}+ra_{2},\,qa_{1}+sa_{2})}. Then

  1. (a)

    f𝒂=f𝒃f_{\boldsymbol{a}}=f_{\boldsymbol{b}} if and only if 𝒂𝒃mod2\boldsymbol{a}\equiv\boldsymbol{b}\bmod\mathbb{Z}^{2},

  2. (b)

    for γSL2()\gamma\in SL_{2}(\mathbb{Z}), f𝒂(γτ)=f𝒂𝜸(τ)f_{\boldsymbol{a}}(\gamma\tau)=f_{\boldsymbol{a\gamma}}(\tau) for all τ\tau\in\mathbb{H}.

Therefore, all elements in {f𝐚:𝐚S}\{f_{\boldsymbol{a}}:\boldsymbol{a}\in S\} satisfy modularity with respect to Γ(N)\Gamma(N).

In order to conclude that the functions f𝒂f_{\boldsymbol{a}} are modular functions, we study their Fourier expansions at the cusp. But first, we need to understand the behaviour of the Weierstrass \wp-function.

Lemma.

Fix τ\tau\in\mathbb{H} and denote τ=Lτ\wp_{\tau}=\wp_{L_{\tau}}. For zLτz\not\in L_{\tau}, we have

(2πi)2τ(z)=112+1ξ12+ξ+n=1cnqn,(2\pi i)^{-2}\wp_{\tau}(z)=\frac{1}{12}+\frac{1}{\xi^{-1}-2+\xi}+\sum_{n=1}^{\infty}c_{n}q^{n},

where

ξ=e2πiz,q=e2πiτ and cn=d|nd(ξd2+ξd)n1.\displaystyle\xi=e^{2\pi iz},\,q=e^{2\pi i\tau}\text{ and }c_{n}=\sum_{d|n}d\left(\xi^{-d}-2+\xi^{d}\right)\ \forall\ n\geq 1.
Proof.

From the definition of the \wp-function, we have

τ(z)\displaystyle\wp_{\tau}(z) =1z2+(m,n)2(m,n)(0,0)[1(zmτn)21(mτ+n)2]\displaystyle=\frac{1}{z^{2}}+\sum_{\begin{subarray}{c}(m,n)\in\mathbb{Z}^{2}\\ (m,n)\neq(0,0)\end{subarray}}\left[\frac{1}{(z-m\tau-n)^{2}}-\frac{1}{(m\tau+n)^{2}}\right]
=1z2+n{0}1(z+n)2n{0}1n2+m0n1(zmτn)2m0n1(mτ+n)2\displaystyle=\frac{1}{z^{2}}+\sum_{n\in\mathbb{Z}\setminus\{0\}}\frac{1}{(z+n)^{2}}-\sum_{n\in\mathbb{Z}\setminus\{0\}}\frac{1}{n^{2}}+\sum_{m\neq 0}\sum_{n\in\mathbb{Z}}\frac{1}{(z-m\tau-n)^{2}}-\sum_{m\neq 0}\sum_{n\in\mathbb{Z}}\frac{1}{(m\tau+n)^{2}}
=2ζ(2)+n1(z+n)22m=1n1(mτ+n)2\displaystyle=-2\zeta(2)+\sum_{n\in\mathbb{Z}}\frac{1}{(z+n)^{2}}-2\sum_{m=1}^{\infty}\sum_{n\in\mathbb{Z}}\frac{1}{(m\tau+n)^{2}}
+m=1n[1(z+mτ+n)2+1(z+mτ+n)2].\displaystyle\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad+\sum_{m=1}^{\infty}\sum_{n\in\mathbb{Z}}\left[\frac{1}{(-z+m\tau+n)^{2}}+\frac{1}{(z+m\tau+n)^{2}}\right].

Recall the Lipschitz summation formula, which states

n1(z+n)k=(2πi)k(k1)!n=1nk1qn.\sum_{n\in\mathbb{Z}}\frac{1}{(z+n)^{k}}=\frac{(-2\pi i)^{k}}{(k-1)!}\sum_{n=1}^{\infty}n^{k-1}q^{n}.

For a proof, see [10, Theorem 4.2.2]. Applying this, we obtain

τ(z)=π23+(2πi)2n=1ne2πinz\displaystyle\wp_{\tau}(z)=-\frac{\pi^{2}}{3}+(-2\pi i)^{2}\sum_{n=1}^{\infty}ne^{2\pi inz}
+(2πi)2m=1n=1n[e2πin(z+mτ)+e2πin(z+mτ)2e2πimnτ]\displaystyle\quad\quad\quad\quad\quad\quad\quad\quad+(-2\pi i)^{2}\sum_{m=1}^{\infty}\sum_{n=1}^{\infty}n\left[e^{2\pi in(-z+m\tau)}+e^{2\pi in(z+m\tau)}-2e^{2\pi imn\tau}\right]
(2πi)2τ(z)=112+n=1ne2πinz+m,n=1nqmn[e2πinz+e2πinz2]\displaystyle\Rightarrow(2\pi i)^{-2}\wp_{\tau}(z)=\frac{1}{12}+\sum_{n=1}^{\infty}ne^{2\pi inz}+\sum_{m,n=1}^{\infty}nq^{mn}\left[e^{2\pi inz}+e^{-2\pi inz}-2\right]

This implies the result. ∎

Using the above expansion, the interpretation of g2(Lτ)g_{2}(L_{\tau)}, g3(Lτ)g_{3}(L_{\tau}) and Δ0(Lτ)\Delta_{0}(L_{\tau}) in terms of Eisenstein series, we get for 𝒂=(rN,sN)S\boldsymbol{a}=\left(\frac{r}{N},\frac{s}{N}\right)\in S,

f𝒂(τ)\displaystyle f_{\boldsymbol{a}}(\tau)
=12592E4(τ)E6(τ)Δ(τ)[112n=1ne2πinsNqNnr+m,n=1nqmn[e2πinsNqNnr+e2πinsNqNnr2]].\displaystyle=\frac{-1}{2592}\cdot\frac{E_{4}(\tau)E_{6}(\tau)}{\Delta(\tau)}\left[\frac{1}{12}-\sum_{n=1}^{\infty}ne^{\frac{2\pi ins}{N}}q_{N}^{nr}+\sum_{m,n=1}^{\infty}nq^{mn}\left[e^{\frac{2\pi ins}{N}}q_{N}^{nr}+e^{\frac{-2\pi ins}{N}}q_{N}^{-nr}-2\right]\right]. (5)

Here qN=e2πiτ/Nq_{N}=e^{2\pi i\tau/N}. Since the Fourier series of Δ(τ)\Delta(\tau) begins with q=qNNq=q_{N}^{N}, the Fourier expansion of f𝒂(τ)f_{\boldsymbol{a}}(\tau) begins with a rational multiple of qNNq_{N}^{-N}. Thus, f𝒂f_{\boldsymbol{a}} has pole of order NN at ii\infty for all 𝒂S\boldsymbol{a}\in S. If ss is any other cusp, then there exists γSL2()\gamma\in SL_{2}(\mathbb{Z}) such that s=γ(i)s=\gamma(i\infty). Since f𝒂(γτ)=f𝒂𝜸(τ)f_{\boldsymbol{a}}(\gamma\tau)=f_{\boldsymbol{a\gamma}}(\tau), which has pole of order NN at ii\infty, we conclude that f𝒂f_{\boldsymbol{a}} is meromorphic at all cusps, with pole of order NN. Thus, f𝒂f_{\boldsymbol{a}} is a modular function of level NN for all 𝒂S\boldsymbol{a}\in S.

This helps us to conclude the following.

Theorem 2.10.

For all 𝐚S\boldsymbol{a}\in S, the Fourier coefficients of f𝐚f_{\boldsymbol{a}} with respect to all cusps belong to (μN)\mathbb{Q}(\mu_{N}), where μN=e2πi/N\mu_{N}=e^{2\pi i/N}.

Proof.

Recall that Fourier coefficients of E4(τ),E6(τ)E_{4}(\tau),E_{6}(\tau) and Δ(τ)\Delta(\tau) are integers. Hence, from (5), it follows that the Fourier coefficients of f𝒂f_{\boldsymbol{a}} at ii\infty lie in (μN)\mathbb{Q}(\mu_{N}) for all 𝒂S\boldsymbol{a}\in S. If ss is any other cusp, then there exists γSL2()\gamma\in SL_{2}(\mathbb{Z}) such s=γ(i)s=\gamma(i\infty). But f𝒂(γτ)=f𝒂𝜸(τ)f_{\boldsymbol{a}}(\gamma\tau)=f_{\boldsymbol{a\gamma}}(\tau) also has Fourier coefficients in (μN)\mathbb{Q}(\mu_{N}) with respect to ii\infty. This completes the proof. ∎

These modular functions, together with the jj-function serve to generate all modular functions of level NN. That is,

Theorem 2.11.

We have N,¯=¯(j,{f𝐚|𝐚S})\mathcal{F}_{N,\overline{\mathbb{Q}}}=\overline{\mathbb{Q}}\left(j,\,\{f_{\boldsymbol{a}}|{\boldsymbol{a}}\in S\}\right) and N,¯\mathcal{F}_{N,\overline{\mathbb{Q}}} is a finite Galois extension of ¯(j)\overline{\mathbb{Q}}(j).

Proof.

Let EN,¯:=¯(j,{f𝒂|𝒂S})E_{N,\overline{\mathbb{Q}}}:=\overline{\mathbb{Q}}\left(j,\,\{f_{\boldsymbol{a}}|{\boldsymbol{a}}\in S\}\right). Then EN,¯E_{N,\overline{\mathbb{Q}}} is a Galois extension of ¯(j)\overline{\mathbb{Q}}(j). Indeed, for each 𝒂S{\boldsymbol{a}}\in S, the modular function f𝒂f_{\boldsymbol{a}} is holomorphic on \mathbb{H} with algebraic Fourier coefficients at all cusps. Hence, by Corollary 2.8 we get a polynomial P(X)¯(j)[X]P(X)\in\overline{\mathbb{Q}}(j)[X], which is satisfied by f𝒂f_{\boldsymbol{a}}. Thus, EN,¯/¯(j)E_{N,\overline{\mathbb{Q}}}\,/\,\overline{\mathbb{Q}}(j) is an algebraic extension. Moreover, this extension is normal because each conjugate of f𝒂f_{\boldsymbol{a}} is of the form f𝒂𝜸𝒊f_{\boldsymbol{a\gamma_{i}}}, which is equal to f𝒃f_{\boldsymbol{b}} for some 𝒃S\boldsymbol{b}\in S. Since EN,¯E_{N,\overline{\mathbb{Q}}} is obtained from ¯(j)\overline{\mathbb{Q}}(j) by adjoining finite number of elements, EN,¯/¯(j)E_{N,\overline{\mathbb{Q}}}\,/\,\overline{\mathbb{Q}}(j) is a finite Galois extension.

Now we show that EN,¯=N,¯E_{N,\overline{\mathbb{Q}}}=\mathcal{F}_{N,\overline{\mathbb{Q}}}. To begin with, observe that \mathbb{C} and N,¯\mathcal{F}_{N,\overline{\mathbb{Q}}} are linearly disjoint over ¯\overline{\mathbb{Q}}. This can be seen as follows. Suppose {c1,c2,,cr}\{c_{1},c_{2},\ldots,c_{r}\}\subset\mathbb{C} is an arbitrary ¯\overline{\mathbb{Q}}-linearly independent subset of \mathbb{C}. If there exists giN,¯g_{i}\in\mathcal{F}_{N,\overline{\mathbb{Q}}} such that i=1rgi(τ)ci=0\sum_{i=1}^{r}g_{i}(\tau)\,c_{i}=0 for all τ\tau\in\mathbb{H} with

gi(τ)=ndinqNn,din¯for 0ir,g_{i}(\tau)=\sum_{n}d_{in}\,\,q_{N}^{n},\qquad d_{in}\in\overline{\mathbb{Q}}\quad\text{for }0\leq i\leq r,

then

i=1rcidin=0 for all n.\sum_{i=1}^{r}c_{i}\,d_{in}=0\quad\text{ for all }n.

The linear independence of cic_{i} over ¯\overline{\mathbb{Q}} implies that all din=0d_{in}=0 for all 0ir0\leq i\leq r, and hence, f1=f2==fr=0f_{1}=f_{2}=\cdots=f_{r}=0. Thus, we have EN,¯N,¯EN,¯.E_{N,\overline{\mathbb{Q}}}\subseteq\mathcal{F}_{N,\overline{\mathbb{Q}}}\subseteq\mathbb{C}E_{N,\overline{\mathbb{Q}}}. Suppose there exists fN,¯EN,¯f\in\mathcal{F}_{N,\overline{\mathbb{Q}}}\setminus E_{N,\overline{\mathbb{Q}}}. Since fEN,¯f\in\mathbb{C}E_{N,\overline{\mathbb{Q}}}, we get a ¯\overline{\mathbb{Q}}-linearly independent subset {f1,f2,,fm}EN,¯\{f_{1},f_{2},\ldots,f_{m}\}\subseteq E_{N,\overline{\mathbb{Q}}} such that

f=i=1mαifi,αi,f=\sum_{i=1}^{m}\alpha_{i}f_{i},\quad\alpha_{i}\in\mathbb{C}, (6)

with at least one of the αi¯\alpha_{i}\in\mathbb{C}\setminus\overline{\mathbb{Q}}. Since EN,¯N,¯E_{N,\overline{\mathbb{Q}}}\subseteq\mathcal{F}_{N,\overline{\mathbb{Q}}}, the set {f,f1,f2,,fm}\{f,\,f_{1},f_{2},\ldots,f_{m}\} is ¯\overline{\mathbb{Q}}-linearly independent subset of N,¯\mathcal{F}_{N,\overline{\mathbb{Q}}} and hence, \mathbb{C}-linearly independent. This contradicts (6). Therefore, EN,¯=N,¯E_{N,\overline{\mathbb{Q}}}=\mathcal{F}_{N,\overline{\mathbb{Q}}} and N,¯/¯(j)\mathcal{F}_{N,\overline{\mathbb{Q}}}\,/\,\overline{\mathbb{Q}}(j) is a finite Galois extension. ∎

Thus, Theorem 2.10 and Theorem 2.11 imply the following crucial fact.

Corollary 2.12.

Let gN,¯g\in\mathcal{F}_{N,\overline{\mathbb{Q}}}. Then the Fourier coefficients of gg with respect to all cusps are algebraic numbers and gg satisfies a polynomial over ¯(j)\overline{\mathbb{Q}}(j).

2.4. Algebraic independence of modular, quasi-modular forms and modular functions

The aim of the discussion below is to establish the algebraic independence (as functions) of the three functions arising from modular considerations - modular functions, modular forms and quasi-modular forms. For the basic theory of quasi-modular forms, we refer the reader to [14]. To begin with, we prove the following lemma.

Lemma 2.13.

A sum of meromorphic quasi-modular forms of distinct weights is not identically zero, unless, each of these quasi-modular forms is identically zero.

Proof.

Suppose f1,f2,,frf_{1},f_{2},\ldots,f_{r} are non-identically zero meromorphic quasi-modular forms of weights k1<k2<<krk_{1}<k_{2}<\cdots<k_{r} with respect to the congruence subgroups Γ1,Γ2,,Γr\Gamma_{1},\Gamma_{2},\ldots,\Gamma_{r} of level N1,N2,,NrN_{1},N_{2},\ldots,N_{r} respectively. Let pp be the greatest depth of these quasi-modular forms. Suppose that for each i{1,2,,r}i\in\{1,2,\ldots,r\}, they have the following transformation formulae:

fi|ki(abcd)(τ)=j=0pfi,j(τ)(ccτ+d)j\displaystyle f_{i}\big{|}_{k_{i}}\left(\begin{smallmatrix}a&b\\ c&d\end{smallmatrix}\right)(\tau)=\sum_{j=0}^{p}f_{i,j}(\tau)\left(\frac{c}{c\tau+d}\right)^{j}
i.e., fi(aτ+bcτ+d)=j=0pfi,j(τ)cj(cτ+d)kij\displaystyle f_{i}\left(\frac{a\tau+b}{c\tau+d}\right)=\sum_{j=0}^{p}f_{i,j}(\tau)\,c^{j}\,(c\tau+d)^{k_{i}-j}

for every (abcd)Γi\left(\begin{smallmatrix}a&b\\ c&d\end{smallmatrix}\right)\in\Gamma_{i} and τ\tau\in\mathbb{H}. Here the functions fi,0,fi,1,,fi,pf_{i,0},f_{i,1},\ldots,f_{i,p} are the components of fif_{i}, and in particular, fi,0=fif_{i,0}=f_{i} for all 1ir1\leq i\leq r. Then we consider N:=i=1rNiN:=\prod_{i=1}^{r}N_{i} and S:={bN2+1:b}S:=\{bN^{2}+1:b\in\mathbb{N}\} such that for each b=Nb2+1Sb^{\prime}=Nb^{2}+1\in S, the matrix (1bNNb)i=1rΓi\left(\begin{smallmatrix}1&bN\\ N&b^{\prime}\end{smallmatrix}\right)\in\cap_{i=1}^{r}\Gamma_{i}. For all such matrices we have

fi(τ+bNNτ+b)=j=0pfi,j(τ)Nj(Nτ+b)kij for all 1ir.f_{i}\left(\frac{\tau+bN}{N\tau+b^{\prime}}\right)=\sum_{j=0}^{p}f_{i,j}(\tau)\,N^{j}\,\left(N\tau+b^{\prime}\right)^{k_{i}-j}\quad\text{ for all }1\leq i\leq r.

Suppose that i=1rfi=0\sum_{i=1}^{r}f_{i}=0. Then for τ\tau\in\mathbb{H} and bSb^{\prime}\in S, we have

i=1rfi(τ+bNNτ+b)=0.\sum_{i=1}^{r}f_{i}\left(\frac{\tau+bN}{N\tau+b^{\prime}}\right)=0.

From the above transformation formula for each fif_{i}, we obtain

i=1rj=0pfi,j(τ)Nj(Nτ+b)kij=0.\sum_{i=1}^{r}\sum_{j=0}^{p}f_{i,j}(\tau)\,N^{j}\,(N\tau+b^{\prime})^{k_{i}-j}=0. (7)

For a fixed τ\tau\in\mathbb{H}, multiplying (7) by (Nτ+b)p(N\tau+b^{\prime})^{p^{\prime}}, where p=2max{p,|k1|,,|kr|}p^{\prime}=2\cdot\max\{p,|k_{1}|,\ldots,|k_{r}|\}, we get

0\displaystyle 0 =i=1rj=0pfi,j(τ)Nj(Nτ+b)ki+pj\displaystyle=\sum_{i=1}^{r}\sum_{j=0}^{p}f_{i,j}(\tau)\,N^{j}\,(N\tau+b^{\prime})^{k_{i}+p^{\prime}-j}
=i=1rNki+pj=0pfi,j(τ)(τ+bN)ki+pj,\displaystyle=\sum_{i=1}^{r}N^{k_{i}+p^{\prime}}\sum_{j=0}^{p}f_{i,j}(\tau)\left(\tau+\frac{b^{\prime}}{N}\right)^{k_{i}+p^{\prime}-j},

which holds for each bSb^{\prime}\in S. This shows that for any fixed τ\tau\in\mathbb{H}, the polynomial

P(X)=i=1rNki+pj=0pfi,j(τ)(τ+X)ki+pj[X]P(X)=\sum_{i=1}^{r}N^{k_{i}+p^{\prime}}\sum_{j=0}^{p}f_{i,j}(\tau)\left(\tau+X\right)^{k_{i}+p^{\prime}-j}\in\mathbb{C}[X]

has infinitely many roots, and hence P(X)P(X) is identically zero. Note that even if any of the kjk_{j}’s are negative, the maximum exponent of XX occurring in P(X)P(X) is p+krp^{\prime}+k_{r}. Thus, the leading coefficient of P(X)P(X) is Nkr+pfr,0(τ)N^{k_{r}+p^{\prime}}\,f_{r,0}(\tau). Since P(X)=0P(X)=0, we get fr,0(τ)=fr(τ)=0f_{r,0}(\tau)=f_{r}(\tau)=0. This is true for any τ\tau\in\mathbb{H}. Hence, fr=0f_{r}=0, which is a contradiction, proving the lemma. ∎

We are now ready to prove the main theorem in this context.

Theorem 2.14.

Let ff a non-constant meromorphic modular form of non-zero weight. Let gg and f~\widetilde{f} be a modular function and a meromorphic quasi-modular form of positive depth and non-zero weight, respectively. Assume that ff, gg and f~\widetilde{f} are of arbitrary level. Then the functions ff, gg and f~\widetilde{f} are algebraically independent.

Proof.

Let ff be of weight mm and level NN, gg be of level MM and f~\widetilde{f} be of weight nn and level N~\widetilde{N}. Suppose that P[X,Y,Z]P\in\mathbb{C}[X,Y,Z] is such that P(f,g,f~)=0P(f,g,\widetilde{f})=0. Recall that a product of a modular form of weight mm and level NN and a quasimodular form of weight nn and level N~\widetilde{N} is a quasimodular form of weight mnmn and level NN~N\widetilde{N}. Denote 0=0\mathbb{N}_{0}=\mathbb{Z}_{\geq 0}. By grouping the terms of P(f,g,f~)P(f,g,\widetilde{f}) of the same weight, we can rewrite it as

P(f,g,f~)=k=KK(r,s,t)03mr+nt=kpr,s,tfrgsf~t,with pr,s,t.P(f,g,\widetilde{f})=\sum_{k=-K^{\prime}}^{K}\,\sum_{\begin{subarray}{c}{(r,s,t)\in\mathbb{N}_{0}^{3}}\\ {mr+nt=k}\end{subarray}}p_{r,s,t}\,\,f^{r}\,g^{s}\,{\widetilde{f}}^{t},\qquad\text{with }p_{r,s,t}\in\mathbb{C}. (8)

As P(X,Y,Z)P(X,Y,Z) is a polynomial, pr,s,t=0p_{r,s,t}=0 for all but finitely many rr, ss and tt. As noted above, for each k{K,,K}k\in\{-K^{\prime},\ldots,K\} the inner sum in (8) is a meromorphic quasi-modular form of weight kk and level NMN~NM\widetilde{N}. Thus, Lemma 2.13 implies that

(r,s,t)03mr+nt=kpr,s,tfrgsf~t=0.\sum_{\begin{subarray}{c}{(r,s,t)\in\mathbb{N}_{0}^{3}}\\ {mr+nt=k}\end{subarray}}p_{r,s,t}\,\,f^{r}\,g^{s}\,{\widetilde{f}}^{t}=0. (9)

for each KkK-K^{\prime}\leq k\leq K.

Fix a k{K,,K}k\in\{-K^{\prime},\ldots,K\} and consider the corresponding inner sum from (9). If there exists an integer t00t_{0}\neq 0 such that pr,s,t00p_{r,s,t_{0}}\neq 0 for at least one tuple (r,s,t0)03(r,s,t_{0})\in\mathbb{N}_{0}^{3} satisfying mr+nt0=kmr+nt_{0}=k, then the term on the left in (9) is a meromorphic quasi-modular form of depth at least t0t_{0}. But the uniqueness of depth implies that the term on the left in (9) is of depth 0. Hence, the coefficient pr,s,t=0p_{r,s,t}=0 when t0t\neq 0. If t=0t=0, then mr=kmr=k, or in other words, r=k/mr=k/m\in\mathbb{Z}. Thus, for all 0k{K,,K}0\neq k\in\{-K^{\prime},\ldots,K\} such that km\frac{k}{m}\in\mathbb{Z}, the relation in (9) takes the form

s0pkm,s,0gs=0,\sum_{s\in\mathbb{N}_{0}}p_{\frac{k}{m},s,0}\,\,g^{s}=0, (10)

as ff is not identically zero. If k=0k=0, then clearly, r=0r=0 and we get

s0p0,s,0gs=0.\sum_{s\in\mathbb{N}_{0}}p_{0,s,0}\,\,g^{s}=0. (11)

Since gg is a non-constant modular function, gg must have either a zero or a pole in \mathbb{H}. Say that the order of gg at τ0\tau_{0} is b0b\neq 0. Then any function r=0Rcrgr\sum_{r=0}^{R}c_{r}\,g^{r} with not all cr=0c_{r}=0 will have order bR0bR\neq 0 at τ0\tau_{0}. Thus, any combination of the form r=0Rcrgr\sum_{r=0}^{R}c_{r}\,g^{r} cannot be identically zero. Thus, relations (10) and (11) imply that p0,s,0=pkm,s,0=0p_{0,s,0}=p_{\frac{k}{m},s,0}=0 for all kk and ss. Hence P=0P=0 and the theorem is proved. ∎


3. Proof of Main Results


3.1. Values of modular functions when j(τ)¯j(\tau)\not\in\overline{\mathbb{Q}}

Before proceeding with the proof of the main theorems, we establish the following intermediary observation.

Lemma 3.1.

Let gN,¯g\in\mathcal{F}_{N,\overline{\mathbb{Q}}}. If τ\tau\in\mathbb{H} such that j(τ)¯j(\tau)\notin\overline{\mathbb{Q}}, then g(τ)g(\tau) is non-zero.

Proof.

By Theorem 2.11, we know that gg satisfies a non-trivial polynomial over ¯(j)\overline{\mathbb{Q}}(j). On clearing denominators, we can assume that gg satisfies the irreducible polynomial

P(X)=r=0mcr(j)Xr=r=0m(s=0drcrsjs)Xr¯[j][X],m=degP(X),dr=degcr(j),P(X)=\sum_{r=0}^{m}c_{r}(j)\,X^{r}=\sum_{r=0}^{m}\bigg{(}\sum_{s=0}^{d_{r}}c_{rs}\,j^{s}\bigg{)}\,X^{r}\in\overline{\mathbb{Q}}[j][X],\quad m=\deg P(X),\,\,d_{r}=\deg c_{r}(j),

Note that c0(j)¯[j]c_{0}(j)\in\overline{\mathbb{Q}}[j] is not the identically zero polynomial. Now suppose τ\tau\in\mathbb{H} is such that j(τ)j(\tau) is transcendental. If g(τ)g(\tau) is zero, then we get that c0(j(τ))=0c_{0}(j(\tau))=0 which contradicts the transcendence of j(τ)j(\tau). This completes the proof. ∎

Proof of Theorem 1.6(a).

If g(τ)=0g(\tau)=0, then it follows from Lemma 3.1 that j(τ)¯j(\tau)\in\overline{\mathbb{Q}}. If τ\tau is a pole of gg, we consider the modular function

h(z):=1g(z)N,¯,h(z):=\frac{1}{g(z)}\in\mathcal{F}_{N,\overline{\mathbb{Q}}},

which vanishes at z=τz=\tau, and so j(τ)¯j(\tau)\in\overline{\mathbb{Q}} by Lemma 3.1. ∎

The proof of Theorem 1.6(b) is established in two parts. We first show that if j(τ)j(\tau) is transcendental, then g(τ)g(\tau) is as well. The converse implication is proved later.

Proposition 3.2.

Let gN,¯g\in\mathcal{F}_{N,\overline{\mathbb{Q}}} and τ\tau\in\mathbb{H} be such that it is not a pole of gg. If j(τ)j(\tau) is transcendental, then g(τ)g(\tau) is transcendental.

Proof.

Suppose that g(τ)¯g(\tau)\in\overline{\mathbb{Q}}. Then the modular function

h(z):=g(z)g(τ)h(z):=g(z)-g(\tau)

belongs to N,¯\mathcal{F}_{N,\overline{\mathbb{Q}}} and vanishes at τ\tau. This contradicts Lemma 3.1. Therefore, g(τ)g(\tau) is transcendental. ∎

3.2. Values of modular functions when j(τ)¯j(\tau)\in\overline{\mathbb{Q}}

From Theorem 2.11, we have

N,¯=¯({j,f𝒂:𝒂S}), where S={(rN,sN):0r,sN1 and (r,s)(0,0)}.\mathcal{F}_{N,\overline{\mathbb{Q}}}=\overline{\mathbb{Q}}\left(\{j,f_{\boldsymbol{a}}:{\boldsymbol{a}}\in S\}\right),\text{ where }S=\left\{\left(\frac{r}{N},\frac{s}{N}\right):0\leq r,s\leq N-1\text{ and }(r,s)\neq(0,0)\right\}.

Hence, for any gN,¯g\in\mathcal{F}_{N,\overline{\mathbb{Q}}}, the algebraic nature of g(τ)g(\tau) is determined by the numbers in the set {f𝒂(τ):𝒂S}\{f_{\boldsymbol{a}}(\tau):{\boldsymbol{a}}\in S\}.

For N=1N=1, 1,¯=¯(j)\mathcal{F}_{1,\overline{\mathbb{Q}}}=\overline{\mathbb{Q}}(j). Hence, for any g1,¯g\in\mathcal{F}_{1,\overline{\mathbb{Q}}}, j(τ)¯j(\tau)\in\overline{\mathbb{Q}} implies that g(τ)¯g(\tau)\in\overline{\mathbb{Q}}. For N>1N>1, we have to study the nature of the values f𝒂(τ)f_{\boldsymbol{a}}(\tau) for all 𝒂S{\boldsymbol{a}}\in S. We start by proving the following important lemma.

Lemma 3.3.

If τ\tau\in\mathbb{H} such that j(τ)j(\tau) is algebraic, then there exists a unique transcendental number (up to an algebraic multiples) ωτ\omega_{\tau} for which g2(ωτLτ)g_{2}(\omega_{\tau}L_{\tau}) and g3(ωτLτ)g_{3}(\omega_{\tau}L_{\tau}) are both algebraic numbers.

Proof.

Note that for any α×\alpha\in\mathbb{C}^{\times},

j(τ)=E4(τ)3Δ(τ)=1728g2(Lτ)3Δ0(Lτ)=1728g2(αLτ)3Δ0(αLτ),j(\tau)=\frac{E_{4}(\tau)^{3}}{\Delta(\tau)}=\frac{1728\,g_{2}(L_{\tau})^{3}}{\Delta_{0}(L_{\tau})}=\frac{1728\,g_{2}\left(\alpha L_{\tau}\right)^{3}}{\Delta_{0}\left(\alpha L_{\tau}\right)}, (12)

because of the homogeneity properties of the g2g_{2}, g3g_{3} functions. Here Lτ=τL_{\tau}=\tau\,\mathbb{Z}\oplus\mathbb{Z} and αLτ=ατα\alpha L_{\tau}=\alpha\,\tau\,\mathbb{Z}\oplus\alpha\,\mathbb{Z}. If we choose ωτ\omega_{\tau} such that ωτ4=g2(Lτ)\omega_{\tau}^{4}=g_{2}(L_{\tau}), then g2(ωτLτ)=1g_{2}(\omega_{\tau}L_{\tau})=1. Given that j(τ)j(\tau) is algebraic, from (12), we get that Δ0(ωτLτ)\Delta_{0}(\omega_{\tau}L_{\tau}) is also algebraic. Thus, the numbers g2(ωτLτ)g_{2}(\omega_{\tau}L_{\tau}) and g3(ωτLτ)g_{3}(\omega_{\tau}L_{\tau}) are both algebraic. Moreover, ωτ\omega_{\tau} is a period of ωτLτ\omega_{\tau}L_{\tau} and both g2(ωτLτ)g_{2}(\omega_{\tau}L_{\tau}) and g3(ωτLτ)g_{3}(\omega_{\tau}L_{\tau}) are algebraic. By Corollary 2.4 we conclude that ωτ\omega_{\tau} is transcendental.

To prove uniqueness, consider an arbitrary ωτ×\omega_{\tau}^{\prime}\in\mathbb{C}^{\times} such that both g2(ωτLτ)g_{2}(\omega_{\tau}^{\prime}L_{\tau}) and g3(ωτLτ)g_{3}(\omega_{\tau}^{\prime}L_{\tau}) are algebraic. By homogeneity of g2g_{2}, we have

g2(ωτLτ)=(ωτ)4g2(Lτ)=:β¯.\quad g_{2}(\omega_{\tau}^{\prime}L_{\tau})=(\omega_{\tau}^{\prime})^{-4}\,g_{2}(L_{\tau})=:\beta\in\overline{\mathbb{Q}}.

Then, we get g2(Lτ)=β(ωτ)4g_{2}(L_{\tau})=\beta\cdot(\omega_{\tau}^{\prime})^{4} and hence ωτ4=β(ωτ)4.\omega_{\tau}^{4}=\beta\cdot(\omega_{\tau}^{\prime})^{4}. This completes the proof that ωτ\omega_{\tau} is unique up to algebraic multiples. ∎

Now recall that

f𝒂(τ):=g2(Lτ)g3(Lτ)Δ0(Lτ)Lτ(a1τ+a2),f_{\boldsymbol{a}}(\tau):=\frac{g_{2}(L_{\tau})\,g_{3}(L_{\tau})}{\Delta_{0}(L_{\tau})}\,\wp_{L_{\tau}}(a_{1}\tau+a_{2}),

where Lτ=τL_{\tau}=\tau\,\mathbb{Z}\oplus\mathbb{Z} and 𝒂=(a1,a2)S{\boldsymbol{a}}=(a_{1},a_{2})\in S. Using homogeneity properties of the functions involved, we can rewrite this as follows:

f𝒂(τ)=g2(ωτLτ)g3(ωτLτ)Δ0(ωτLτ)ωτLτ(a1ωττ+a2ωτ).f_{\boldsymbol{a}}(\tau)=\frac{g_{2}(\omega_{\tau}L_{\tau})\,g_{3}(\omega_{\tau}L_{\tau})}{\Delta_{0}(\omega_{\tau}L_{\tau})}\,\wp_{\omega_{\tau}L_{\tau}}(a_{1}\omega_{\tau}\tau+a_{2}\omega_{\tau}).
Lemma 3.4.

If τ\tau\in\mathbb{H} is such that j(τ)¯j(\tau)\in\overline{\mathbb{Q}}, then f𝐚(τ)¯f_{\boldsymbol{a}}(\tau)\in\overline{\mathbb{Q}} for all 𝐚S{\boldsymbol{a}}\in S.

Proof.

Let 𝒂=(rN,sN)S{\boldsymbol{a}}=\left(\frac{r}{N},\frac{s}{N}\right)\in S. Then we have

f𝒂(τ)=g2(ωτLτ)g3(ωτLτ)Δ0(ωτLτ)ωτLτ(rωττN+sωτN),f_{\boldsymbol{a}}(\tau)=\frac{g_{2}(\omega_{\tau}L_{\tau})\,g_{3}(\omega_{\tau}L_{\tau})}{\Delta_{0}(\omega_{\tau}L_{\tau})}\,\wp_{\omega_{\tau}L_{\tau}}\left(\frac{r\,\omega_{\tau}\,\tau}{N}+\frac{s\,\omega_{\tau}}{N}\right),

where ωτ\omega_{\tau} is chosen as in Lemma 3.3 so that g2(ωτLτ),g3(ωτLτ)¯g_{2}(\omega_{\tau}L_{\tau}),\,g_{3}(\omega_{\tau}L_{\tau})\in\overline{\mathbb{Q}}.

Recall that both ωτLτ(ωττN)\wp_{\omega_{\tau}L_{\tau}}\left(\frac{\omega_{\tau}\,\tau}{N}\right) and ωτLτ(ωτN)\wp_{\omega_{\tau}L_{\tau}}\left(\frac{\omega_{\tau}}{N}\right) are algebraic by Corollary 2.4. Moreover, by the addition formula for Weierstrass \wp-function in Theorem 2.2, we get that

ωτLτ(rωττN+sωτN)¯\wp_{\omega_{\tau}L_{\tau}}\left(\frac{r\,\omega_{\tau}\,\tau}{N}+\frac{s\,\omega_{\tau}}{N}\right)\in\overline{\mathbb{Q}}

for all 𝒂=(rN,sN)S{\boldsymbol{a}}=\left(\frac{r}{N},\frac{s}{N}\right)\in S. Thus, the number f𝒂(τ)f_{\boldsymbol{a}}(\tau) is algebraic for all 𝒂S{\boldsymbol{a}}\in S. ∎

Proposition 3.5.

Let gN,¯g\in\mathcal{F}_{N,\overline{\mathbb{{Q}}}} and τ\tau\in\mathbb{H} such that j(τ)¯j(\tau)\in\overline{\mathbb{Q}}. If τ\tau is not a pole of gg, then g(τ)g(\tau) is algebraic.

Proof.

From Lemma 3.4, we have f𝒂(τ)¯f_{\boldsymbol{a}}(\tau)\in\overline{\mathbb{Q}} for all 𝒂S{\boldsymbol{a}}\in S. Since

gN,¯=¯({j,f𝒂|𝒂S})g\in\mathcal{F}_{N,\overline{\mathbb{Q}}}=\overline{\mathbb{Q}}\left(\{j,f_{\boldsymbol{a}}|{\boldsymbol{a}}\in S\}\right)

by Theorem 2.11, g(τ)g(\tau) is also algebraic. ∎


Propositions 3.2 and 3.5 together complete the proof of Theorem 1.6(b). We now prove the generalization of [5, Theorem 2.4].

Proof of Theorem 1.7.

Let ff be a meromorphic modular form of weight kk\in\mathbb{Z}. Define

h(τ):=f(τ)12Δ(τ)kτ.h(\tau):=\frac{f(\tau)^{12}}{\Delta(\tau)^{k}}\quad\forall\tau\in\mathbb{H}.

The function hN,¯h\in\mathcal{F}_{N,\overline{\mathbb{Q}}} because the Fourier coefficients of ff and Δ\Delta are algebraic. Moreover, any τ\tau\in\mathbb{H} is a zero or a pole of hh if and only if it is a zero or a pole of ff because Δ\Delta is non-vanishing and holomorphic on \mathbb{H}. Thus, if τ\tau\in\mathbb{H} is a zero or pole of ff, then hh is non-constant and by Theorem 1.6(a), we deduce that j(τ)¯j(\tau)\in\overline{\mathbb{Q}}, and by Theorem 1.6(b), that either g(z)g(z) has a pole at z=τz=\tau or that g(τ)¯g(\tau)\in\overline{\mathbb{Q}}. ∎


3.3. Values of modular forms

Recall that for any integer kk and NN, we denote Mk,N,¯M_{k,N,\overline{\mathbb{Q}}}, Mk,N,¯wM_{k,N,\overline{\mathbb{Q}}}^{w} and Mk,N,¯mM_{k,N,\overline{\mathbb{Q}}}^{m} to be the set of holomorphic, weakly holomorphic and meromorphic modular forms respectively of level NN with algebraic Fourier coefficients at ii\infty. By Corollary 2.12, all elements of these sets have algebraic Fourier coefficients with respect to all cusps as well.

Proof of Theorem 1.8(a) .

Let fMk,N,¯mf\in M^{m}_{k,N,\overline{\mathbb{Q}}}. Consider the modular function

g(τ):=f(τ)12Δ(τ)k.g(\tau):=\frac{{f(\tau)}^{12}}{\Delta(\tau)^{k}}.

If gg is a non-zero constant, then f(τ)=cΔ(τ)k/12f(\tau)=c\Delta(\tau)^{k/12} for some c¯×c\in\overline{\mathbb{Q}}^{\times} and all τ\tau\in\mathbb{H}. Since 1728Δ(τ)=E43(τ)E62(τ)1728\Delta(\tau)=E_{4}^{3}(\tau)-E_{6}^{2}(\tau), and E4(τ)E_{4}(\tau) is algebraically independent with E6(τ)E_{6}(\tau) for τ\tau\in\mathbb{H} such that e2πiτ¯e^{2\pi i\tau}\in\overline{\mathbb{Q}} by Nesterenko’s theorem 1.3, we deduce that f(τ)f(\tau) is transcendental.

Now suppose that gg is non-constant. Since fMk,N,¯f\in M_{k,N,\overline{\mathbb{Q}}}, gN,¯g\in\mathcal{F}_{N,\overline{\mathbb{Q}}}. Note that for τ0\tau_{0}\in\mathbb{H} such that e2πiτ0e^{2\pi i\tau_{0}} is algebraic, j(τ0)j(\tau_{0}) is transcendental by Theorem 1.2. Such a τ0\tau_{0} cannot be a pole of ff. For if τ0\tau_{0} is a pole of ff, then τ0\tau_{0} is a pole of gg and by Theorem 1.6(a), j(τ0)j(\tau_{0}) would be algebraic, leading to a contradiction. As τ0\tau_{0} is not a pole of gg, Theorem 1.6(b) implies that g(τ0)g(\tau_{0}) is transcendental. Moreover, by Theorem 2.11, g(τ0)g(\tau_{0}) is algebraic over ¯(j(τ0))\overline{\mathbb{Q}}(j(\tau_{0})). By Nesterenko’s theorem,

Δ(τ0)=E43(τ0)E62(τ0)1728 and j(τ0)=E43(τ0)Δ(τ0)\Delta(\tau_{0})=\frac{E_{4}^{3}(\tau_{0})-E_{6}^{2}(\tau_{0})}{1728}\qquad\text{ and }\qquad j(\tau_{0})=\frac{E_{4}^{3}(\tau_{0})}{\Delta(\tau_{0})}

are algebraically independent. Therefore,

trdeg¯(g(τ0),Δ(τ0))=trdeg¯(j(τ0),Δ(τ0))=2.\text{trdeg}_{\,\mathbb{Q}}\,\,\overline{\mathbb{Q}}\left(g(\tau_{0}),\,\Delta(\tau_{0})\right)=\text{trdeg}_{\,\mathbb{Q}}\,\,\overline{\mathbb{Q}}\left(j(\tau_{0}),\,\Delta(\tau_{0})\right)=2.

Hence, f(τ0)12=g(τ0)Δ(τ0)kf(\tau_{0})^{12}=g(\tau_{0})\,\Delta(\tau_{0})^{k} is transcendental, proving the claim. ∎

We remark here that the proof of Theorem 1.8(a) only requires Nesterenko’s theorem and the structure of modular functions of higher level. One also immediately deduces the following.

Proposition 3.6.

If fMk,N,¯mf\in M^{m}_{k,N,\overline{\mathbb{Q}}}, then ff is algebraically dependent with E4E_{4} and E6E_{6} over ¯\overline{\mathbb{Q}}. In particular, If fMk,N,¯mf\in M_{k,N,\overline{\mathbb{Q}}}^{m} and τ\tau\in\mathbb{H} is not a pole of ff, then f(τ)f(\tau) is algebraic over ¯(E4(τ),E6(τ))\overline{\mathbb{Q}}(E_{4}(\tau),E_{6}(\tau)).

Proof.

Let fMk,N,¯mf\in M^{m}_{k,N,\overline{\mathbb{Q}}} and consider

g(τ):=f12(τ)Δk(τ)N,¯.g(\tau):=\frac{f^{12}(\tau)}{\Delta^{k}(\tau)}\in\mathcal{F}_{N,\overline{\mathbb{Q}}}.

By Corollary 2.12, there exists a polynomial P(X)¯[j](X)P(X)\in\overline{\mathbb{Q}}[j](X) such that P(g)=0P(g)=0. More specifically,

r=0ms=0drcr,sj(τ)sg(τ)r=0for allτ.\sum_{r=0}^{m}\sum_{s=0}^{d_{r}}c_{r,s}\,{j(\tau)}^{s}\,{g(\tau)}^{r}=0\qquad\text{for all}\qquad\tau\in\mathbb{H}.

Here cr,s¯c_{r,s}\in\overline{\mathbb{Q}} for all 0sdr0\leq s\leq d_{r} and 0rm0\leq r\leq m. Multiplying by Δl(τ)\Delta^{l}(\tau) for any positive integer l>kml>km and substituting j(τ)j(\tau) and Δ(τ)\Delta(\tau) in terms of E4(τ)E_{4}(\tau) and E6(τ)E_{6}(\tau) gives

r=0ms=0drt=0lkrs(1)tcr,s1728lkrs(lkrst)f(τ)12rE4(τ)3(lkrt)E6(τ)2t=0.\sum_{r=0}^{m}\sum_{s=0}^{d_{r}}\sum_{t=0}^{l-kr-s}\frac{(-1)^{t}c_{r,s}}{1728^{l-kr-s}}\left(\begin{array}[]{c}l-k\,r-s\\ t\end{array}\right)f(\tau)^{12\,r}E_{4}(\tau)^{3(l-k\,r-t)}E_{6}(\tau)^{2\,t}=0.

This proves the proposition. ∎

We now consider the complementary case, namely, points τ\tau\in\mathbb{H} such that j(τ)¯j(\tau)\in\overline{\mathbb{Q}}.

Proof of Theorem 1.8(b).

Fix a τ\tau\in\mathbb{H} such that j(τ)¯j(\tau)\in\overline{\mathbb{Q}}. From Lemma 3.3, we get a transcendental number ωτ\omega_{\tau} such that g2(ωτLτ)g_{2}(\omega_{\tau}L_{\tau}) and g3(ωτLτ)g_{3}(\omega_{\tau}L_{\tau}) are both algebraic. Moreover, we have the formulae

E4(τ)\displaystyle E_{4}(\tau) =34π4g2(Lτ)=34(ωτπ)4g2(ωτLτ),\displaystyle=\frac{3}{4\pi^{4}}\,g_{2}(L_{\tau})=\frac{3}{4}\left(\frac{\omega_{\tau}}{\pi}\right)^{4}g_{2}(\omega_{\tau}L_{\tau}),
E6(τ)\displaystyle E_{6}(\tau) =278π6g3(Lτ)=278(ωτπ)6g3(ωτLτ).\displaystyle=\frac{27}{8\pi^{6}}\,g_{3}(L_{\tau})=\frac{27}{8}\left(\frac{\omega_{\tau}}{\pi}\right)^{6}g_{3}(\omega_{\tau}L_{\tau}).

From Nesterenko’s Theorem 1.3, we know that at most one of E4(τ)E_{4}(\tau) and E6(τ)E_{6}(\tau) is algebraic. The above formulae imply that if ωτ/π¯{\omega_{\tau}}/{\pi}\in\overline{\mathbb{Q}}, then both E4(τ)E_{4}(\tau) and E6(τ)E_{6}(\tau) are algebraic, which is a contradiction. Hence, ωτ/π{\omega_{\tau}}/{\pi} is a transcendental number. Besides, we have

Δ(τ)=E4(τ)3E6(τ)21728\displaystyle\Delta(\tau)=\frac{E_{4}(\tau)^{3}-E_{6}(\tau)^{2}}{1728} =146(ωτπ)12(g2(ωτLτ)327g3(ωτLτ)2)\displaystyle=\frac{1}{4^{6}}\left(\frac{\omega_{\tau}}{\pi}\right)^{12}\bigg{(}g_{2}(\omega_{\tau}L_{\tau})^{3}-27g_{3}(\omega_{\tau}L_{\tau})^{2}\bigg{)}
=146(ωτπ)12Δ0(ωτLτ).\displaystyle=\frac{1}{4^{6}}\left(\frac{\omega_{\tau}}{\pi}\right)^{12}\Delta_{0}(\omega_{\tau}L_{\tau}).

As g2(ωτLτ)g_{2}(\omega_{\tau}L_{\tau}) and g3(ωτLτ)g_{3}(\omega_{\tau}L_{\tau}) both are algebraic, the number Δ0(ωτLτ)¯{0}\Delta_{0}(\omega_{\tau}L_{\tau})\in\overline{\mathbb{Q}}\setminus\{0\}. Since ωτ/π{\omega_{\tau}}/{\pi} is transcendental and Δ(τ)\Delta(\tau) is a non-zero algebraic multiple of (ωτ/π)12\left({\omega_{\tau}}/{\pi}\right)^{12}, we deduce that Δ(τ)\Delta(\tau) is transcendental.

Consider the modular function

g(τ):=f(τ)12Δ(τ)kτ,g(\tau):=\frac{f(\tau)^{12}}{\Delta(\tau)^{k}}\quad\forall\tau\in\mathbb{H},

which lies in N,¯\mathcal{F}_{N,\overline{\mathbb{Q}}}. Since j(τ)¯j(\tau)\in\overline{\mathbb{Q}}, Theorem 3.5 implies that g(τ)g(\tau) is algebraic, say α\alpha. Thus, we get that

f(τ)12=αΔ(τ)k=α46Δ0(ωτLτ)k(ωτπ)12k.f(\tau)^{12}=\alpha\,\cdot\,\Delta(\tau)^{k}=\frac{\alpha}{4^{6}}\,\cdot\,\Delta_{0}(\omega_{\tau}L_{\tau})^{k}\,\left(\frac{\omega_{\tau}}{\pi}\right)^{12k}. (13)

This shows that if f(τ)0f(\tau)\neq 0, then it is a non-zero algebraic multiple of (ωτ/π)k\left({\omega_{\tau}}/{\pi}\right)^{k} and hence, is transcendental. ∎


3.4. Values of quasi-modular forms

Let M~k,¯(p)(Γ)\widetilde{M}_{k,\overline{\mathbb{Q}}}^{(p)}(\Gamma) denote the set of all meromorphic quasi-modular forms of weight kk and depth p(>0)p(>0) for Γ\Gamma with algebraic Fourier coefficients. We study their values at the points τ\tau\in\mathbb{H}, where exactly one of e2πiτe^{2\pi i\tau} and j(τ)j(\tau) is algebraic.

Proof of Theorem 1.9(a).

By Theorem 2.1, we can write f~\widetilde{f} in the form

f~=r=0pfrE2r,frMk,N,¯.\widetilde{f}=\sum_{r=0}^{p}f_{r}E_{2}^{r},\quad f_{r}\in M_{k,N,\overline{\mathbb{Q}}}.

Suppose that both e2πiτ,f~(τ)¯e^{2\pi i\tau},\widetilde{f}(\tau)\in\overline{\mathbb{Q}}. By Corollary 3.6, we know that each number fr(τ)f_{r}(\tau) is algebraically dependent with E4(τ),E6(τ)E_{4}(\tau),E_{6}(\tau). Thus, we get that E2(τ)E_{2}(\tau) is algebraic over ¯(E4(τ),E6(τ))\overline{\mathbb{Q}}(E_{4}(\tau),E_{6}(\tau)). This implies that

trdeg¯(e2πiτ,E2(τ),E4(τ),E6(τ))2,\text{trdeg}_{\,\mathbb{Q}}\,\,\overline{\mathbb{Q}}\bigg{(}e^{2\pi i\tau},E_{2}(\tau),E_{4}(\tau),E_{6}(\tau)\bigg{)}\leq 2,

contradicting Nesterenko’s theorem 1.3. This proves the claim. ∎

To study values of quasi-modular forms in the complementary case (i.e., for j(τ)¯j(\tau)\in\overline{\mathbb{Q}}), we need the following lemma.

Lemma 3.7.

Let τ\tau\in\mathbb{H} be such that j(τ)j(\tau) is algebraic and ωτ\omega_{\tau} be the transcendental number determined in Lemma 3.3 such that g2(ωτLτ)g_{2}(\omega_{\tau}L_{\tau}) and g3(ωτLτ)g_{3}(\omega_{\tau}L_{\tau)} are algebraic. Let η2(ωτ):=η2(ωτLτ)\eta_{2}(\omega_{\tau}):=\eta_{2}(\omega_{\tau}L_{\tau}) be the quasi-period

η2(ωτLτ)=ζωτLτ(z+ωτ)ζωτLτ(z).\eta_{2}(\omega_{\tau}L_{\tau})=\zeta_{\omega_{\tau}L_{\tau}}(z+\omega_{\tau})-\zeta_{\omega_{\tau}L_{\tau}}(z).

Then ωτπ\frac{\omega_{\tau}}{\pi} and η2π\frac{\eta_{2}}{\pi} are algebraically independent over ¯\overline{\mathbb{Q}}.

Proof.

From the definition of the Weierstrass zeta-function, one gets ζωτLτ(z)=1ωτζLτ(zωτ).{\zeta}_{\omega_{\tau}L_{\tau}}(z)=\frac{1}{\omega_{\tau}}\cdot\zeta_{L_{\tau}}\left(\frac{z}{\omega_{\tau}}\right). Using the definition of a quasi-period and the identity (3), we obtain

η2=ζωτLτ(z+ωτ)ζωτLτ(z)=1ωτη2(Lτ)=1ωτG2(τ)=13π2ωτE2(τ).\eta_{2}=\zeta_{\omega_{\tau}L_{\tau}}(z+\omega_{\tau})-\zeta_{\omega_{\tau}L_{\tau}}(z)=\frac{1}{\omega_{\tau}}\,\eta_{2}(L_{\tau})=\frac{1}{\omega_{\tau}}\,G_{2}(\tau)=\frac{1}{3}\cdot\frac{\pi^{2}}{\omega_{\tau}}\,E_{2}(\tau).

Thus, we get the following formulae

E2(τ)=3ωτπη2π,E4(τ)=34(ωτπ)4g2(ωτLτ),E6(τ)=278(ωτπ)6g3(ωτLτ).E_{2}(\tau)=3\,\frac{\omega_{\tau}}{\pi}\frac{\eta_{2}}{\pi},\quad E_{4}(\tau)=\frac{3}{4}\left(\frac{\omega_{\tau}}{\pi}\right)^{4}g_{2}(\omega_{\tau}L_{\tau}),\quad E_{6}(\tau)=\frac{27}{8}\left(\frac{\omega_{\tau}}{\pi}\right)^{6}g_{3}(\omega_{\tau}L_{\tau}).

The above formulae imply that E2(τ)E_{2}(\tau), E4(τ)E_{4}(\tau) and E6(τ)E_{6}(\tau) are algebraic over ¯(ωτπ,η2π)\overline{\mathbb{Q}}\left(\frac{\omega_{\tau}}{\pi},\frac{\eta_{2}}{\pi}\right). Suppose that ωτπ,η2π\frac{\omega_{\tau}}{\pi},\frac{\eta_{2}}{\pi} are algebraically dependent. This implies that

trdeg¯(E2(τ),E4(τ),E6(τ))=1,\text{trdeg}_{\,\mathbb{Q}}\,\,\overline{\mathbb{Q}}\bigg{(}E_{2}(\tau),E_{4}(\tau),E_{6}(\tau)\bigg{)}=1,

which contradicts Nesterenko’s Theorem 1.3. ∎

Lemma 3.7 and Theorem 2.1 together allow us to describe values of quasi-modular forms explicitly at the points τ\tau\in\mathbb{H} where the jj-function is algebraic.

Proof of Theorem 1.9(b).

From Theorem 2.1, we have the expression

f~(τ)=r=0pfr(τ)E2(τ)r,where frMk,N,¯m.\widetilde{f}(\tau)=\sum_{r=0}^{p}f_{r}(\tau)E_{2}(\tau)^{r},\quad\text{where }f_{r}\in M^{m}_{k,N,\overline{\mathbb{Q}}}.

Writing the value fr(τ)f_{r}(\tau) as in (13) for each coefficient frf_{r}, and using the above formula for E2E_{2}, we obtain

f~(τ)=r=0pcr(ωτπ)k2r(ωτπη2π)r=r=0pcr(ωτπ)kr(η2π)r,\widetilde{f}(\tau)=\sum_{r=0}^{p}c_{r}\left(\frac{\omega_{\tau}}{\pi}\right)^{k-2r}\left(\frac{\omega_{\tau}}{\pi}\cdot\frac{\eta_{2}}{\pi}\right)^{r}=\sum_{r=0}^{p}\,c_{r}\,\left(\frac{\omega_{\tau}}{\pi}\right)^{k-r}\,\left(\frac{\eta_{2}}{\pi}\right)^{r}, (14)

where each crc_{r} is an algebraic number. If cr0c_{r}\neq 0 for some rr satisfying 0rp0\leq r\leq p, then (14) gives a non-trivial algebraic relation among ωτπ\frac{\omega_{\tau}}{\pi} and η2π\frac{\eta_{2}}{\pi}. This contradicts Lemma 3.7. Therefore, the number f~(τ)\widetilde{f}(\tau) is either zero, precisely when each f~(τ)¯{0}\widetilde{f}(\tau)\in\overline{\mathbb{Q}}\setminus\{0\}, or is transcendental. ∎


3.5. Algebraic independence of special values

Proof of Theorem 1.11.

From Theorem 2.11, we know that for any τ\tau\in\mathbb{H} which is not a pole of gg, g(τ)g(\tau) is algebraic over ¯(j(τ))\overline{\mathbb{Q}}(j(\tau)). Moreover, Corollary 3.6 gives that f(τ)f(\tau) is algebraic over ¯(E4(τ),E6(τ))\overline{\mathbb{Q}}(E_{4}(\tau),E_{6}(\tau)). Since f~(τ)=r=0pfr(τ)E2(τ)r\widetilde{f}(\tau)=\sum_{r=0}^{p}f_{r}(\tau)E_{2}(\tau)^{r}, we have that f~(τ)\widetilde{f}(\tau) is algebraic over

¯(E2(τ),E4(τ),E6(τ)).{\overline{\mathbb{Q}}(E_{2}(\tau),E_{4}(\tau),E_{6}(\tau)).}

Hence, we get that

trdeg¯(e2πiτ,g(τ),f(τ),f~(τ))\displaystyle\text{trdeg}_{\,\mathbb{Q}}\,\,\overline{\mathbb{Q}}\bigg{(}e^{2\pi i\tau},g(\tau),f(\tau),\widetilde{f}(\tau)\bigg{)} =trdeg¯(e2πiτ,j(τ),E2(τ),E4(τ),E6(τ))\displaystyle=\text{trdeg}_{\,\mathbb{Q}}\,\,\overline{\mathbb{Q}}\bigg{(}e^{2\pi i\tau},j(\tau),E_{2}(\tau),E_{4}(\tau),E_{6}(\tau)\bigg{)}
=trdeg¯(e2πiτ,E2(τ),E4(τ),E6(τ))3,\displaystyle=\text{trdeg}_{\,\mathbb{Q}}\,\,\overline{\mathbb{Q}}\bigg{(}e^{2\pi i\tau},E_{2}(\tau),E_{4}(\tau),E_{6}(\tau)\bigg{)}\geq 3,

by Nesterenko’s theorem 1.3. This establishes the claim. ∎

Acknowledgment

The authors thank Prof. Michel Waldschmidt and Dr. Veekesh Kumar for insightful suggestions on an earlier version of the paper. They are also grateful to the referee for helpful comments.

References

  • [1] R. Balasubramanian, S. Gun, On zeros of quasi-modular forms, Journal of Number Theory, Vol. 132 (2012) 2228-2241.
  • [2] K. Barré-Sirieix, G. Diaz, F. Gramain, G. Philibert, Une preuve de la conjecture de Mahler-Manin, Invent. Math., 124 (1996), 1-9.
  • [3] C. Y. Chang, Transcendence of special values of quasi-modular forms, Forum Math. 24 (2012) 539-551.
  • [4] H. Cohen and F. Stromberg, Modular Forms, A Classical Approach, Graduate Studies In Mathematics 179.
  • [5] D. Jeon, S.-Y. Kang, C.H. Kim, On values of weakly holomorphic modular functions at divisors of meromorphic modular forms, J. Number Theory 239, 183-206.
  • [6] D. Jeon, S.-Y. Kang, C.H. Kim, Hecke system of harmonic Maass functions and its applications to modular curves of higher genera, Ramanujan J., Vol. 62, (2023) 675-717.
  • [7] Alia Hamieh and M. Ram Murty, A note on qq-analogues of Dirichlet L-functions, International Journal of Number Theory, Vol. 12 (2016) 765-773.
  • [8] Serge Lang, Elliptic Functions, Addision-Wesley Publishing Company, Inc.
  • [9] Kurt Mahler, Remarks on a paper by W. Schwarz., Journal of Number Theory, 1 (1969) 512–521.
  • [10] M. R. Murty, M. Dewar and H. Graves, Problems in the Theory of Modular Forms, Hindustan Book Agency.
  • [11] M. Ram Murty and P. Rath, Transcendental Numbers Springer-Verlag, (2014).
  • [12] S. Gun, M. Ram Murty, P. Rath, Algebraic independence of values of modular forms. Int. J. Number Theory 7(4), 1065–1074 (2011).
  • [13] Y. Nesterenko, Modular functions and transcendence questions, Mat. Sb. 187 (9) (1996) 65–96 (in Russian); translation in: Sb. Math. 187 (9) (1996) 1319–1348.
  • [14] E. Royer, Quasimodular forms: an introduction, Annales Mathématiques Blaise Pascal, Vol. 19 no. 2 (2012) 297-306.
  • [15] T. Schneider, Arithmetische Untersuchungen elliptischer Integrale, Math. Ann. 113 (1937) 1–13.
  • [16] T. Schneider, Ein Satz über ganzwertige Funktionen als Prinzip für Transzendenzbeweise, Mathematische Annalen, 121 (1949): 131–140.
  • [17] G. Shimura, Introduction to the Arithmetic Theory of Automorphic Functions, (Princeton University Press, 19711971).
  • [18] W. Wang and H. Zhang, Meromorphic quasi-modular forms and their LL-functions, Journal of Number Theory, Vol. 241 (2022) 465-503.
  • [19] W. Wang, Algebraic independence of values of quasi-modular forms, Journal of Number Theory, Vol. 255 (2024) 85-97.
  • [20] D. Zagier, Elliptic Modular Forms and Their Applications, Lectures at a Summer School in Nordfjordeid, Norway.