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

Explicit bounds on the coefficients of modular polynomials for the elliptic jj-invariant

 and  Florian Breuer and Fabien Pazuki School of Information and Physical Sciences, The University of Newcastle, University Drive, Callaghan, NSW 2308, Australia. [email protected] Department of Mathematical Sciences, University of Copenhagen, Universitetsparken 5, 2100 Copenhagen Ø, Denmark, and Université de Bordeaux, 33405 Talence, France. [email protected]
The authors thank the IRN GandA (CNRS). The second author is supported by ANR-20-CE40-0003 Jinvariant.

Abstract. We obtain an explicit upper bound on the size of the coefficients of the elliptic modular polynomials ΦN\Phi_{N} for any N1N\geq 1. These polynomials vanish at pairs of jj-invariants of elliptic curves linked by cyclic isogenies of degree NN. The main term in the bound is asymptotically optimal as NN tends to infinity.

Keywords: Modular polynomials, elliptic curves.

Mathematics Subject Classification: 11G05.

———

1. Introduction

For any non-zero polynomial PP in one or more variables and complex coefficients we define its height to be

h(P):=logmax|c|,where c ranges over all coefficients of P.h(P):=\log\max|c|,\quad\text{where $c$ ranges over all coefficients of $P$.}

Let NN be a positive integer and denote by ΦN=ΦN(X,Y)[X,Y]\Phi_{N}=\Phi_{N}(X,Y)\in\mathbb{Z}[X,Y] the (classical) modular polynomial, which vanishes at pairs of jj-invariants of elliptic curves linked by a cyclic NN-isogeny, see [La87, Chapter 5]. Alternatively, if we view jj as the function on the complex upper half-plane where j(τ)j(\tau) is the jj-invariant of the complex elliptic curve /(+τ)\mathbb{C}/(\mathbb{Z}+\tau\mathbb{Z}), then ΦN(X,j(τ))\Phi_{N}(X,j(\tau)) is the minimal polynomial of j(Nτ)j(N\tau) over (j(τ))\mathbb{C}(j(\tau)).

Modular polynomials have important applications in cryptography and certain algorithms for computing ΦN\Phi_{N} require explicit bounds on the size of the coefficients, so one is interested in explicit bounds on h(ΦN)h(\Phi_{N}).

Paula Cohen Tretkoff [Coh84] proved that when NN tends to ++\infty

(1) h(ΦN)=6ψ(N)[logN2κN+O(1)]h(\Phi_{N})=6\psi(N)\big{[}\log N-2\kappa_{N}+O(1)\big{]}

where

ψ(N)=Np|N(1+1p)andκN=p|Nlogpp,\psi(N)=N\prod_{p|N}\left(1+\frac{1}{p}\right)\quad\text{and}\quad\kappa_{N}=\sum_{p|N}\frac{\log p}{p},

but the implied bounded function is not explicit.

In the case where N=lN=l is prime, Bröker and Sutherland [BrSu10] estimated the constants in Cohen’s argument to obtain

h(Φl)6llogl+16l+14llogl.h(\Phi_{l})\leq 6l\log l+16l+14\sqrt{l}\log l.

In the general case, the second author [Paz19] obtained in his Corollary 4.3, via a different method,

(2) h(ΦN)ψ(N)[6logN+logψ(N)+6log(12logN+2logψ(N)+25.2)+15.7].h(\Phi_{N})\leq\psi(N)\big{[}6\log N+\log\psi(N)+6\log(12\log N+2\log\psi(N)+25.2)+15.7\big{]}.

Inequality (2) has the merit of being completely explicit for all N1N\geq 1, but the main term is slightly too big when compared with the asymptotic of (1).

The goal of the present paper is to prove the following result, where we solve this issue and provide an upper bound with the correct main term for all NN. Let us first define

λN:=pnNpn1pn1(p21)logp.\lambda_{N}:=\sum_{p^{n}\|N}\frac{p^{n}-1}{p^{n-1}(p^{2}-1)}\log p.
Theorem 1.1.

Let N2N\geq 2. The height of the modular polynomial ΦN(X,Y)\Phi_{N}(X,Y) is bounded by

(3) h(ΦN)6ψ(N)[logN2λN+loglogN+4.436].h(\Phi_{N})\leq 6\psi(N)\big{[}\log N-2\lambda_{N}+\log\log N+4.436\big{]}.

We prove this theorem using a different path than the one followed in [Paz19]. The main new ingredient is a finer estimate of the Mahler measure of jj-invariants, coming from previous work of Pascal Autissier [Aut03]. We also use precise analytic estimates for the discriminant modular form on the fundamental domain of the upper half plane (under the classical action of SL2()\mathrm{SL}_{2}(\mathbb{Z})), and a classical interpolation method to help us derive bounds on the height of a polynomial in two variables, from knowledge of the height of several specializations of this polynomial.

Let us now discuss the optimality of the bound. The main term is the expected one. For lower order terms, notice that

0.385<p|Nlogpp(p+1)λNκNp|Nlogpp(p21)<0.186,-0.385<-\sum_{p|N}\frac{\log p}{p(p+1)}\leq\lambda_{N}-\kappa_{N}\leq\sum_{p|N}\frac{\log p}{p(p^{2}-1)}<0.186,

so one changes little replacing λN\lambda_{N} by κN\kappa_{N} in Theorem 1.1. On the other hand, one would like to get rid of the spurious loglogN\log\log N term, but for practical purposes this might be less useful than keeping the constant as small as possible.

It is interesting to consider the functions bλ(N)b_{\lambda}(N) and bκ(N)b_{\kappa}(N) for which

(4) h(ΦN)=6ψ(N)[logN2λN+bλ(N)]=6ψ(N)[logN2κN+bκ(N)].h(\Phi_{N})=6\psi(N)\big{[}\log N-2\lambda_{N}+b_{\lambda}(N)\big{]}=6\psi(N)\big{[}\log N-2\kappa_{N}+b_{\kappa}(N)\big{]}.

These functions are plotted in Figure 1 for N400N\leq 400, based on computations of ΦN\Phi_{N} by Andrew Sutherland [Suth] using the algorithms in [BKS12] (for prime NN) and [BOS16] (for composite NN).

The content of Cohen’s Theorem is that bκ(N)b_{\kappa}(N) and thus also bλ(N)b_{\lambda}(N) are bounded functions. Our Theorem 1.1 is equivalent to bλ(N)loglogN+4.436b_{\lambda}(N)\leq\log\log N+4.436, which is clearly seen to hold for N400N\leq 400; in fact, bλ(N)<2.1b_{\lambda}(N)<2.1 in this range.

In our proof of Theorem 1.1 we may thus assume that N>400N>400. We explain in Remark 3.2 and in Lemma 3.3 that more computations for N>400N>400 lead to minor improvements on the constant 4.436.4.436.

From Figure 1 it appears that bλ(N)b_{\lambda}(N) is bounded more tightly than bκ(N)b_{\kappa}(N), thus suggesting that λN\lambda_{N} is a more natural function to use in the bound for h(ΦN)h(\Phi_{N}) than is κN\kappa_{N}.

Acknowledgements.

The authors are grateful to Pascal Autissier for suggesting that the results in [Aut03, §2] might be fruitfully applied to estimating h(ΦN)h(\Phi_{N}). They are also grateful to Joseph Silverman for an interesting discussion around [Sil90]. The authors warmly thank Andrew Sutherland for the computations of modular polynomials he performed in record time to help them improve numerical values in the statement of Theorem 1.1. They also thank the referees for very efficient feedback. The authors thank the IRN GandA (CNRS). The second author is supported by ANR-20-CE40-0003 Jinvariant.

Refer to caption
Figure 1. The bounded functions bλ(N)b_{\lambda}(N) (bold) and bκ(N)b_{\kappa}(N) (grey) satisfying h(ΦN)=6ψ(N)[logN2λN+bλ(N)]=6ψ(N)[logN2κN+bκ(N)]h(\Phi_{N})=6\psi(N)\big{[}\log N-2\lambda_{N}+b_{\lambda}(N)\big{]}=6\psi(N)\big{[}\log N-2\kappa_{N}+b_{\kappa}(N)\big{]} for N400.N\leq 400. Notice that bλ(N)<2.1b_{\lambda}(N)<2.1 in this range. Theorem 1.1 is equivalent to bλ(N)loglogN+4.436b_{\lambda}(N)\leq\log\log N+4.436.

2. Preliminary results

Denote the complex upper half-plane by

:={z|Im(z)>0}.\mathbb{H}:=\{z\in\mathbb{C}\;|\;\mathop{\mathrm{Im}}(z)>0\}.

Every τ\tau\in\mathbb{H} defines a lattice Λτ=+τ\Lambda_{\tau}=\mathbb{Z}+\tau\mathbb{Z} in \mathbb{C}, and it is well known that every complex elliptic curve is isomorphic to /Λτ\mathbb{C}/\Lambda_{\tau} for some τ\tau\in\mathbb{H}. If we denote the jj-invariant of this elliptic curve by j(τ)j(\tau), then

j:j:\mathbb{H}\longrightarrow\mathbb{C}

defines an analytic function on \mathbb{H}.

The group SL2()\mathrm{SL}_{2}(\mathbb{Z}) acts on the upper half-plane \mathbb{H} by

γ(τ):=aτ+bcτ+d,whereγ=(abcd)SL2().\gamma(\tau):=\frac{a\tau+b}{c\tau+d},\quad\text{where}\quad\gamma=\left(\begin{matrix}a&b\\ c&d\end{matrix}\right)\in\mathrm{SL}_{2}(\mathbb{Z}).

A fundamental domain for this action is given by

={τ:|τ|1,12<Re(τ)12andRe(τ)0if|τ|=1}.\mathcal{F}=\{\tau\in\mathbb{H}\;:\;|\tau|\geq 1,\;-\frac{1}{2}<\mathrm{Re}(\tau)\leq\frac{1}{2}\;\mathrm{and}\;\mathrm{Re}(\tau)\geq 0\;\mathrm{if}\;|\tau|=1\}.

Thus every τ\tau\in\mathbb{H} is SL2()\mathrm{SL}_{2}(\mathbb{Z})-equivalent to an element τ~\tilde{\tau}\in\mathcal{F}, which we call reduced.

The modular function j:j:\mathbb{H}\rightarrow\mathbb{C} is SL2()\mathrm{SL}_{2}(\mathbb{Z})-invariant. We define

q=e2πiτ,τ,q=e^{2\pi i\tau},\quad\tau\in\mathbb{H},

then the Fourier expansion at infinity of jj can be written as a qq-expansion

j(τ)=1q+744+196884q+.j(\tau)=\frac{1}{q}+744+196884q+\ldots.

We denote by Δ\Delta the modular discriminant function

Δ:,\Delta:\mathbb{H}\longrightarrow\mathbb{C},

which is a weight 12 cusp form for SL2()\mathrm{SL}_{2}(\mathbb{Z}). We normalize Δ\Delta so that its qq-expansion is

Δ(τ)=qn=1(1qn)24=q24q2+252q3+.\Delta(\tau)=q\prod_{n=1}^{\infty}(1-q^{n})^{24}=q-24q^{2}+252q^{3}+\cdots.

We point out that the discriminant of the elliptic curve EτE_{\tau} is given by (2π)12Δ(τ)(2\pi)^{12}\Delta(\tau), which is why most sources (e.g. [La87]) normalize Δ\Delta differently, multiplying the above product by the factor (2π)12(2\pi)^{12}. We choose our normalization to be consistent with [Paz19], which contains estimates that we will use.

Let us denote, for N1N\geq 1,

CN={(ab0d):a,b,d,ad=N,a1, 0bd1,gcd(a,b,d)=1}.C_{N}=\left\{\left(\begin{matrix}a&b\\ 0&d\end{matrix}\right)\;:\;a,b,d\in\mathbb{Z},\;ad=N,\;a\geq 1,\;0\leq b\leq d-1,\;\gcd(a,b,d)=1\right\}.

We have

#CN=ψ(N)=Np|N(1+1p).\#C_{N}=\psi(N)=N\prod_{p|N}\left(1+\frac{1}{p}\right).

The elements of CNC_{N} encode cyclic NN-isogenies in the following way. Let EτE_{\tau} be an elliptic curve. For each

γ=(aγbγ0dγ)CN,\gamma=\left(\begin{matrix}a_{\gamma}&b_{\gamma}\\ 0&d_{\gamma}\end{matrix}\right)\in C_{N},

we let

τγ=γ(τ)=aγτ+bγdγ,Λγ=+τγ,andEγ=Eτγ=/Λγ.\tau_{\gamma}=\gamma(\tau)=\frac{a_{\gamma}\tau+b_{\gamma}}{d_{\gamma}},\quad\Lambda_{\gamma}=\mathbb{Z}+\tau_{\gamma}\mathbb{Z},\quad\text{and}\quad E_{\gamma}=E_{\tau_{\gamma}}=\mathbb{C}/\Lambda_{\gamma}.

Then the natural map

EEγ,(zmodΛτ)(zmodΛγ)E\longrightarrow E_{\gamma},\quad(z\bmod\Lambda_{\tau})\longmapsto(z\bmod\Lambda_{\gamma})

is a cyclic NN-isogeny.

Furthermore, up to isomorphism, every cyclic NN-isogeny with source EE arises in this way. In particular, we have the factorization

ΦN(X,j(τ))=γCN(Xj(τγ)).\Phi_{N}\big{(}X,j(\tau)\big{)}=\prod_{\gamma\in{C_{N}}}\big{(}X-j(\tau_{\gamma})\big{)}.

Our goal is to bound the coefficients of the modular polynomial ΦN(X,Y)\Phi_{N}(X,Y). By interpolation, it is enough to estimate the height of ΦN(X,j(τ))\Phi_{N}(X,j(\tau)) for several carefully chosen τ\tau\in\mathbb{H}.

By [BrZu20, Lemma 1.6] the height of ΦN(X,j(τ))\Phi_{N}(X,j(\tau)) is bounded in terms of its Mahler measure

(5) SN(τ)=γCNlogmax(1,|j(τγ)|)S_{N}(\tau)=\displaystyle{\sum_{\gamma\in{C_{N}}}\log\max\big{(}1,|j(\tau_{\gamma})|\big{)}}

by

(6) h(ΦN(X,j))SN(τ)+log(ψ(N)ψ(N)/2)SN(τ)+ψ(N)log2.h(\Phi_{N}(X,j))\leq S_{N}(\tau)+\log\binom{\psi(N)}{\psi(N)/2}\leq S_{N}(\tau)+\psi(N)\log 2.

We will concentrate on estimating SN(τ)S_{N}(\tau) for a fixed τ\tau\in\mathbb{H}.

In general, τγ\tau_{\gamma} won’t be reduced, so we choose (abcd)SL2()\left(\begin{matrix}a&b\\ c&d\end{matrix}\right)\in\mathrm{SL}_{2}(\mathbb{Z}) for which

τ~γ=aτγ+bcτγ+d\tilde{\tau}_{\gamma}=\frac{a\tau_{\gamma}+b}{c\tau_{\gamma}+d}\in\mathcal{F}

is reduced. Since

Im(τ~γ)=Im(aτγ+bcτγ+d)=Im(τγ)|cτγ+d|2,\mathop{\mathrm{Im}}(\tilde{\tau}_{\gamma})=\mathop{\mathrm{Im}}\left(\frac{a\tau_{\gamma}+b}{c\tau_{\gamma}+d}\right)=\frac{\mathop{\mathrm{Im}}(\tau_{\gamma})}{|c\tau_{\gamma}+d|^{2}},

we obtain

(7) log|cτγ+d|=12[logIm(τ~γ)logIm(τγ)].-\log|c\tau_{\gamma}+d|=\frac{1}{2}\big{[}\log\mathop{\mathrm{Im}}(\tilde{\tau}_{\gamma})-\log\mathop{\mathrm{Im}}(\tau_{\gamma})\big{]}.

Also, since Δ\Delta is a modular form of weight 1212 for SL2()\mathrm{SL}_{2}(\mathbb{Z}), we find that

Δ~γ:=Δ(τ~γ)=(cτγ+d)12Δ(τγ)=:(cτγ+d)12Δγ,\tilde{\Delta}_{\gamma}:=\Delta(\tilde{\tau}_{\gamma})=(c\tau_{\gamma}+d)^{12}\Delta(\tau_{\gamma})=:(c\tau_{\gamma}+d)^{12}\Delta_{\gamma},

so

log|Δγ|\displaystyle\log|\Delta_{\gamma}| =log|Δ~γ|12log|cτγ+d|\displaystyle=\log|\tilde{\Delta}_{\gamma}|-12\log|c\tau_{\gamma}+d|
(8) =log|Δ~γ|+6[logIm(τ~γ)logIm(τγ)].\displaystyle=\log|\tilde{\Delta}_{\gamma}|+6\big{[}\log\mathop{\mathrm{Im}}(\tilde{\tau}_{\gamma})-\log\mathop{\mathrm{Im}}(\tau_{\gamma})\big{]}.

Note that by [Paz19, Lemma 2.4] we have

(9) logIm(τ~γ)logIm(τ)logN\log\mathop{\mathrm{Im}}(\tilde{\tau}_{\gamma})-\log\mathop{\mathrm{Im}}(\tau)\leq\log N

for each γCN\gamma\in C_{N}, provided that τ\tau\in\mathcal{F}.

We need a few more preliminaries:

By [Aut03, Lemme 2.2], we have

γCNΔ(γ(τ))=[Δ(τ)]ψ(N),\prod_{\gamma\in C_{N}}\Delta(\gamma(\tau))=\big{[}-\Delta(\tau)\big{]}^{\psi(N)},

so we get

(10) γCNlog|Δγ|=ψ(N)log|Δ|.\sum_{\gamma\in C_{N}}\log|\Delta_{\gamma}|=\psi(N)\log|\Delta|.

Furthermore, [Aut03, Lemme 2.3] says

γCNlogdγaγ=ψ(N)(logN2λN),\sum_{\gamma\in C_{N}}\log\frac{d_{\gamma}}{a_{\gamma}}=\psi(N)(\log N-2\lambda_{N}),

which combined with

Im(τγ)=Im(aγτ+bγdγ)=aγdγIm(τ)\mathop{\mathrm{Im}}(\tau_{\gamma})=\mathop{\mathrm{Im}}\left(\frac{a_{\gamma}\tau+b_{\gamma}}{d_{\gamma}}\right)=\frac{a_{\gamma}}{d_{\gamma}}\mathop{\mathrm{Im}}(\tau)

gives

(11) γCNlogIm(τγ)=ψ(N)(logN2λNlogIm(τ)).-\sum_{\gamma\in C_{N}}\log\mathop{\mathrm{Im}}(\tau_{\gamma})=\psi(N)\big{(}\log N-2\lambda_{N}-\log\mathop{\mathrm{Im}}(\tau)\big{)}.

Finally, since τ~γ\tilde{\tau}_{\gamma}\in\mathcal{F}, [Paz19, (2.22)] gives us, if we denote jγ=j(τγ)j_{\gamma}=j(\tau_{\gamma}),

(12) Im(τ~γ)12πlog(|jγ|+970.8),\mathop{\mathrm{Im}}(\tilde{\tau}_{\gamma})\leq\frac{1}{2\pi}\log(|j_{\gamma}|+970.8),

whereas [Paz19, (3.18)] gives, for any γCN\gamma\in C_{N},

(13) logmax(|Δ~γ|,|jγΔ~γ|)log(9.02).\log\max(|\tilde{\Delta}_{\gamma}|,|j_{\gamma}\tilde{\Delta}_{\gamma}|)\leq\log(9.02).

This last estimate depends on our choice of normalisation of Δ(τ)\Delta(\tau).

We note that the identities (10) and (11) from [Aut03] involve the non-reduced τγ\tau_{\gamma}, whereas the estimates (9), (12) and (13) from [Paz19] depend on the reduced τ~γ\tilde{\tau}_{\gamma}. The main idea of this paper is to combine these ingredients using (8).

3. Proof of Theorem 1.1

We are now ready to start our main calculation on the sum SN(τ)S_{N}(\tau) from (5).

SN(τ)=\displaystyle S_{N}(\tau)= γCNlogmax(|Δγ|,|jγΔγ|)γCNlog|Δγ|\displaystyle\sum_{\gamma\in C_{N}}\log\max(|\Delta_{\gamma}|,|j_{\gamma}\Delta_{\gamma}|)-\sum_{\gamma\in C_{N}}\log|\Delta_{\gamma}|
=\displaystyle= γCNlogmax(|Δγ|,|jγΔγ|)ψ(N)log|Δ|(by (10))\displaystyle\sum_{\gamma\in C_{N}}\log\max(|\Delta_{\gamma}|,|j_{\gamma}\Delta_{\gamma}|)-\psi(N)\log|\Delta|\quad\text{(by (\ref{eq:Aut2.2}))}
=\displaystyle= γCNlogmax(|Δ~γ|,|jγΔ~γ|)+6γCN[logIm(τ~γ)logIm(τγ)]ψ(N)log|Δ|(by (8)),\displaystyle\sum_{\gamma\in C_{N}}\log\max(|\tilde{\Delta}_{\gamma}|,|j_{\gamma}\tilde{\Delta}_{\gamma}|)+6\sum_{\gamma\in C_{N}}\big{[}\log\mathop{\mathrm{Im}}(\tilde{\tau}_{\gamma})-\log\mathop{\mathrm{Im}}(\tau_{\gamma})\big{]}-\psi(N)\log|\Delta|\quad\text{(by (\ref{eq:cDelta})),}

hence we get

SN(τ)\displaystyle S_{N}(\tau)\leq\; ψ(N)log(9.02)+6γCN[logIm(τ~γ)logIm(τγ)]ψ(N)log|Δ|(by (13))\displaystyle\psi(N)\log(9.02)+6\sum_{\gamma\in C_{N}}\big{[}\log\mathop{\mathrm{Im}}(\tilde{\tau}_{\gamma})-\log\mathop{\mathrm{Im}}(\tau_{\gamma})\big{]}-\psi(N)\log|\Delta|\quad\text{(by (\ref{eq:Paz3.18}))}
=\displaystyle=\; ψ(N)log(9.02)+6ψ(N)(logN2λNlogImτ)\displaystyle\psi(N)\log(9.02)+6\psi(N)\big{(}\log N-2\lambda_{N}-\log\mathop{\mathrm{Im}}\tau\big{)}
+6γCNlogIm(τ~γ)ψ(N)log|Δ|(by (11))\displaystyle+6\sum_{\gamma\in C_{N}}\log\mathop{\mathrm{Im}}(\tilde{\tau}_{\gamma})-\psi(N)\log|\Delta|\quad\text{(by (\ref{eq:Aut2.3}))}
(15) \displaystyle\leq\; 6ψ(N)[logN2λN+0.367]+6γCNlogIm(τ~γ)ψ(N)log[|Δ|(Imτ)6].\displaystyle 6\psi(N)\big{[}\log N-2\lambda_{N}+0.367\big{]}+6\sum_{\gamma\in C_{N}}\log\mathop{\mathrm{Im}}(\tilde{\tau}_{\gamma})-\psi(N)\log\big{[}|\Delta|(\mathop{\mathrm{Im}}\tau)^{6}\big{]}.

At this point we record the following intermediate result. If τ\tau\in\mathcal{F} then we may apply (9) and obtain

SN(τ)\displaystyle S_{N}(\tau)\leq\; 6ψ(N)[logN2λN+0.367]+6ψ(N)[logN+logImτ]ψ(N)log[|Δ|(Imτ)6]\displaystyle 6\psi(N)\big{[}\log N-2\lambda_{N}+0.367\big{]}+6\psi(N)[\log N+\log\mathop{\mathrm{Im}}\tau]-\psi(N)\log\big{[}|\Delta|(\mathop{\mathrm{Im}}\tau)^{6}\big{]}
(16) \displaystyle\leq\; ψ(N)[12logN+2.199log|Δ|].\displaystyle\psi(N)[12\log N+2.199-\log|\Delta|].

We continue our calculation from (15).

SN(τ)\displaystyle S_{N}(\tau)\leq\; 6ψ(N)[logN2λN+0.367]ψ(N)log[|Δ|(Imτ)6]\displaystyle 6\psi(N)\big{[}\log N-2\lambda_{N}+0.367\big{]}-\psi(N)\log\big{[}|\Delta|(\mathop{\mathrm{Im}}\tau)^{6}\big{]}
+6γCNlog[12πlog(|jγ|+970.8)](by (12))\displaystyle+6\sum_{\gamma\in C_{N}}\log\Big{[}\frac{1}{2\pi}\log(|j_{\gamma}|+970.8)\Big{]}\quad\text{(by (\ref{eq:Paz2.22}))}
=\displaystyle=\; 6ψ(N)[logN2λN+0.367]ψ(N)log[|Δ|Im(τ)6]\displaystyle 6\psi(N)\big{[}\log N-2\lambda_{N}+0.367\big{]}-\psi(N)\log\big{[}|\Delta|\mathop{\mathrm{Im}}(\tau)^{6}\big{]}
+6ψ(N)logγCN[12πlog(|jγ|+970.8)]1/ψ(N)\displaystyle+6\psi(N)\log\prod_{\gamma\in C_{N}}\Big{[}\frac{1}{2\pi}\log(|j_{\gamma}|+970.8)\Big{]}^{1/\psi(N)}
\displaystyle\leq\; 6ψ(N)[logN2λN+0.367]ψ(N)log[|Δ|Im(τ)6]\displaystyle 6\psi(N)\big{[}\log N-2\lambda_{N}+0.367\big{]}-\psi(N)\log\big{[}|\Delta|\mathop{\mathrm{Im}}(\tau)^{6}\big{]}
+6ψ(N)log[12πψ(N)γCNlog(|jγ|+970.8)],\displaystyle+6\psi(N)\log\Big{[}\frac{1}{2\pi\psi(N)}\sum_{\gamma\in C_{N}}\log(|j_{\gamma}|+970.8)\Big{]},

where the last inequality follows by the arithmetic-geometric mean inequality.

For any real number xx, the inequality x+970.8971.8max{1,x}x+970.8\leq 971.8\max\{1,x\} holds, so we finally obtain

SN(τ)=\displaystyle S_{N}(\tau)=\; γCNlogmax(1,|jγ|)\displaystyle\sum_{\gamma\in C_{N}}\log\max(1,|j_{\gamma}|)
\displaystyle\leq\; 6ψ(N)[logN2λN+0.367]ψ(N)log[|Δ|Im(τ)6]\displaystyle 6\psi(N)\big{[}\log N-2\lambda_{N}+0.367\big{]}-\psi(N)\log\big{[}|\Delta|\mathop{\mathrm{Im}}(\tau)^{6}\big{]}
+6ψ(N)[logSN(τ)+loglog(971.8)logψ(N)log(2π)]\displaystyle+6\psi(N)\big{[}\log S_{N}(\tau)+\log\log(971.8)-\log\psi(N)-\log(2\pi)\big{]}
(17) \displaystyle\leq\; 6ψ(N)[logN2λN+log(SN(τ)/ψ(N))+0.458]ψ(N)log[|Δ|Im(τ)6].\displaystyle 6\psi(N)\big{[}\log N-2\lambda_{N}+\log\big{(}S_{N}(\tau)/\psi(N)\big{)}+0.458]-\psi(N)\log\big{[}|\Delta|\mathop{\mathrm{Im}}(\tau)^{6}\big{]}.

To deduce an explicit bound on SN(τ)S_{N}(\tau), we start with a crude bound on SN(τ)/ψ(N)S_{N}(\tau)/\psi(N), then strengthen our result recursively. More precisely, we prove the following technical lemma.

Lemma 3.1.

Fix τ\tau\in\mathbb{H} and let

a(τ)\displaystyle a(\tau) =0.45816log[|Δ(τ)|Im(τ)6]\displaystyle=0.458-\frac{1}{6}\log\big{[}|\Delta(\tau)|\mathop{\mathrm{Im}}(\tau)^{6}\big{]}
b(τ)\displaystyle b(\tau) =2.199log|Δ(τ)|.\displaystyle=2.199-\log|\Delta(\tau)|.

Suppose that N>N03N>N_{0}\geq 3. Consider the sequence (cn(τ))n0\big{(}c_{n}(\tau)\big{)}_{n\geq 0} defined recursively by

c0(τ)=\displaystyle c_{0}(\tau)= a(τ)+log[12+b(τ)logN0],\displaystyle\;a(\tau)+\log\left[12+\frac{b(\tau)}{\log N_{0}}\right],
cn+1(τ)=\displaystyle c_{n+1}(\tau)= a(τ)+log6+log[1+loglogN0+cn(τ)logN0],n0.\displaystyle\;a(\tau)+\log 6+\log\left[1+\frac{\log\log N_{0}+c_{n}(\tau)}{\log N_{0}}\right],\quad n\geq 0.

Then for all n0n\geq 0,

(18) SN(τ)6ψ(N)[logN2λN+loglogN+cn(τ)].S_{N}(\tau)\leq 6\psi(N)\big{[}\log N-2\lambda_{N}+\log\log N+c_{n}(\tau)\big{]}.
Proof.

The bound (16) gives

SN(τ)/ψ(N)\displaystyle S_{N}(\tau)/\psi(N) 12logN+b(τ)\displaystyle\leq 12\log N+b(\tau)
[12+b(τ)logN0]logN.\displaystyle\leq\left[12+\frac{b(\tau)}{\log N_{0}}\right]\log N.

Plugging this into (17) gives us (18) with n=0n=0.

Next, assume (18) holds for some n0n\geq 0. Since N>N0N>N_{0}, we obtain

loglogN+cn(τ)<(loglogN0+cn(τ)logN0)logN,\log\log N+c_{n}(\tau)<\left(\frac{\log\log N_{0}+c_{n}(\tau)}{\log N_{0}}\right)\log N,

so (18) gives us

SN(τ)6ψ(N)[logN+(loglogN0+cn(τ)logN0)logN]S_{N}(\tau)\leq 6\psi(N)\left[\log N+\left(\frac{\log\log N_{0}+c_{n}(\tau)}{\log N_{0}}\right)\log N\right]

and so

SN(τ)/ψ(N)6[1+loglogN0+cn(τ)logN0]logN.S_{N}(\tau)/\psi(N)\leq 6\left[1+\frac{\log\log N_{0}+c_{n}(\tau)}{\log N_{0}}\right]\log N.

Plugging this into (17) gives us

SN(τ)6ψ(N)[logN2λN+loglogN+cn+1(τ)].S_{N}(\tau)\leq 6\psi(N)\big{[}\log N-2\lambda_{N}+\log\log N+c_{n+1}(\tau)\big{]}.

The interpolation lemma [BrSu10, Lemma 20] gives, for real L>1L>1,

h(ΦN(X,Y))maxLj2Lh(ΦN(X,j))+ψ(N)(logL+1L+3log2),h(\Phi_{N}(X,Y))\leq\max_{L\leq j\leq 2L}h(\Phi_{N}(X,j))+\psi(N)\left(\frac{\log L+1}{L}+3\log 2\right),

so by (6) we get

(19) h(ΦN(X,Y))maxLj(τ)2LSN(τ)+ψ(N)(logL+1L+4log2).h(\Phi_{N}(X,Y))\leq\max_{L\leq j(\tau)\leq 2L}S_{N}(\tau)+\psi(N)\left(\frac{\log L+1}{L}+4\log 2\right).

It is well-known that the jj-function takes non-negative real values on the following path on the boundary of the fundamental domain \mathcal{F}:

Γ:={eiθ|π3θπ2}{ix|x[0,)}\Gamma:=\{e^{i\theta}\;|\;\frac{\pi}{3}\leq\theta\leq\frac{\pi}{2}\}\cup\{ix\;|\;x\in[0,\infty)\}

and the function j:Γ[0,)j:\Gamma\rightarrow[0,\infty) is a bijection.

We now define, for the values cn(τ)c_{n}(\tau) in Lemma 3.1 with N0=400N_{0}=400,

c(τ):=infn0cn(τ).c(\tau):=\inf_{n\geq 0}c_{n}(\tau).

Optimizing on the interval Lj2LL\leq j\leq 2L, we obtain

h(ΦN(X,Y))6ψ(N)[logN2λN+loglogN+cn(τ)]|j(τ)=2L+ψ(N)(logL+1L+4log2).h(\Phi_{N}(X,Y))\leq 6\psi(N)\big{[}\log N-2\lambda_{N}+\log\log N+c_{n}(\tau)\big{]}\big{|}_{j(\tau)=2L}+\psi(N)\left(\frac{\log L+1}{L}+4\log 2\right).

Optimizing c(τ)c(\tau) (using SageMath [Sage]) when L>1L>1, we obtain the strongest upper bound when we choose L=166.48L=166.48, then τ=j1(L)=ei1.257\tau=j^{-1}(L)=e^{i\cdot 1.257} and

a(τ)1.5004,b(τ)8.1532,c(τ)3.9655.a(\tau)\leq 1.5004,\quad b(\tau)\leq 8.1532,\quad c(\tau)\leq 3.9655.

Putting all of this together, we obtain

h(ΦN)6ψ(N)[logN2λN+loglogN+4.436]h(\Phi_{N})\leq 6\psi(N)\big{[}\log N-2\lambda_{N}+\log\log N+4.436\big{]}

for NN0=400N\geq N_{0}=400.

As can be seen from Figure 1, the result also holds for N400N\leq 400, thus completing the proof of Theorem 1.1.

Let us add the following remark.

Remark 3.2.

The constant in Theorem 1.1 can be further improved if we assume N>N0N>N_{0} for larger values of N0N_{0} and check the result for NN0N\leq N_{0} via direct computation. We list below the values of the constant in Theorem 1.1 obtained assuming N>N0N>N_{0} for some other values of N0N_{0}.

N0N_{0}: constant:
400400 4.4364.436
500500 4.4184.418
10001000 4.3734.373
20002000 4.3364.336
50005000 4.2924.292

The best value is always obtained when L=166.48L=166.48. The gain is somehow limited, even asymptotically, as explained in the next lemma. The next inequality is weaker numerically, but helps understand how the estimates on cnc_{n} will evolve when n+n\to+\infty and N0+N_{0}\to+\infty.

Lemma 3.3.

Suppose that N03N_{0}\geq 3. Consider the sequence (cn(τ))n0\big{(}c_{n}(\tau)\big{)}_{n\geq 0} defined in Lemma 3.1. Then for all n0n\geq 0,

(20) cn(τ)c0(τ)(logN0)n+(a(τ)+log6)logN0logN01+loglogN0logN01.c_{n}(\tau)\leq\frac{c_{0}(\tau)}{(\log N_{0})^{n}}+(a(\tau)+\log 6)\frac{\log N_{0}}{\log N_{0}-1}+\frac{\log\log N_{0}}{\log N_{0}-1}.
Proof.

Let us denote A=a(τ)+log6A=a(\tau)+\log 6, for any x0x\geq 0, we have log(1+x)x\log(1+x)\leq x, hence we get

cn+1(τ)A+loglogN0+cn(τ)logN0,c_{n+1}(\tau)\leq A+\frac{\log\log N_{0}+c_{n}(\tau)}{\log N_{0}},

which gives by induction

cn(τ)c0(τ)(logN0)n+(A+loglogN0logN0)k=0n1(logN0)kc0(τ)(logN0)n+(A+loglogN0logN0)logN0logN01,c_{n}(\tau)\leq\frac{c_{0}(\tau)}{(\log N_{0})^{n}}+\left(A+\frac{\log\log N_{0}}{\log N_{0}}\right)\sum_{k=0}^{n}\frac{1}{(\log N_{0})^{k}}\leq\frac{c_{0}(\tau)}{(\log N_{0})^{n}}+\left(A+\frac{\log\log N_{0}}{\log N_{0}}\right)\frac{\log N_{0}}{\log N_{0}-1},

which gives the conclusion. ∎

If one takes τ=ei1.257\tau=e^{i\cdot 1.257} and n+n\to+\infty in (20) we obtain the following inequality, valid for any N03N_{0}\geq 3 and any NN0N\geq N_{0}:

(21) h(ΦN)6ψ(N)[logN2λN+loglogN+3.293logN0logN01+loglogN0logN01+0.46537].h(\Phi_{N})\leq 6\psi(N)\left[\log N-2\lambda_{N}+\log\log N+3.293\frac{\log N_{0}}{\log N_{0}-1}+\frac{\log\log N_{0}}{\log N_{0}-1}+0.46537\right].

Explicit computation of cn(τ)c_{n}(\tau) will generally give better numerical values of course, but this equation (21) gives an idea of how these estimates will vary with N0N_{0}.

References

  • [Aut03] Autissier, P., Hauteur des correspondances de Hecke. Bull. Soc. Math. France 131 (2003), 421–433.
  • [BKS12] Bröker, R., Lauter, K. and Sutherland, A.V. Modular polynomials via isogeny volcanoes. Mathematics of Computation 81.278 (2012), 1201–1231.
  • [BrSu10] Bröker, R. and Sutherland, A.V., An explicit height bound for the classical modular polynomial. Ramanujan J. 22 (2010), 293–313.
  • [BOS16] Bruiner, J, Ono, K. and Sutherland, A.V. Class polynomials for nonholomorphic modular functions. J. Number Theory 161 (2016), 204–229.
  • [BrZu20] Brunault, F. and Zudilin, W., Many Variations of Mahler Measures, a Lasting Symphony. Australian Mathematical Society Lecture Series, Cambridge University Press, Cambidge, 2020.
  • [Coh84] Cohen, P., On the coefficients of the transformation polynomials for the elliptic modular function, Math. Proc. of the Cambridge Philo. Soc. 95 (1984), 389–402.
  • [La87] Lang, S. Elliptic Functions, 2nd ed. Springer-Verlag, Berlin, 1987.
  • [Paz19] Pazuki, F., Modular invariants and isogenies. Inter. J. Number Theory 15.3 (2019), 569–584.
  • [Sage] The Sage Developers, SageMath, the Sage Mathematics Software System (Version 9.1) https://www.sagemath.org, 2021.
  • [Sil90] Silverman, J.H., Hecke points on modular curves. Duke Math. J. 60.2 (1990), 401–423.
  • [Suth] Sutherland, A. V., Modular polynomials. https://math.mit.edu/~drew/ClassicalModPolys.html