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

Improved bounds for Serre’s open image theorem

Imin Chen Department of Mathematics, Simon Fraser University
Burnaby, BC V5A 1S6, Canada.
ichen@sfu.ca
 and  Joshua Swidinsky Department of Mathematics, Simon Fraser University
Burnaby, BC V5A 1S6, Canada.
joshua_swidinsky@sfu.ca
Abstract.

Let EE be an elliptic curve over the rationals which does not have complex multiplication. Serre showed that the adelic representation attached to E/E/\mathbb{Q} has open image, and in particular there is a minimal natural number CEC_{E} such that the mod \ell representation ρ¯E,{\bar{\rho}}_{E,\ell} is surjective for any prime >CE\ell>C_{E}. Assuming the Generalized Riemann Hypothesis, Mayle-Wang gave explicit bounds for CEC_{E} which are logarithmic in the conductor of EE and have explicit constants. The method is based on using effective forms of the Chebotarev density theorem together with the Faltings-Serre method, in particular, using the ‘deviation group’ of the 22-adic representations attached to two elliptic curves.

By considering quotients of the deviation group and a characterization of the images of the 22-adic representation ρE,2\rho_{E,2} by Rouse and Zureick-Brown, we show in this paper how to further reduce the constants in Mayle-Wang’s results. Another result of independent interest are improved effective isogeny theorems for elliptic curves over the rationals.

2020 Mathematics Subject Classification:
11G05, 11F80

1. Introduction

Let EE be an elliptic curve over \mathbb{Q} without complex multiplication. Serre showed in [14] that the adelic representation attached to E/E/\mathbb{Q} has open image, in particular, there is a minimal natural number CEC_{E} such that the mod \ell representation ρ¯E,{\bar{\rho}}_{E,\ell} is surjective for any prime >CE\ell>C_{E}.

In determining effective bounds on CEC_{E}, one typically uses effective versions of the Chebotarev density theorem under the assumption of the Generalized Riemann Hypothesis (GRH) as was first done by Serre. The bounds on CEC_{E} usually depend on the radical rad(NE)\operatorname{rad}(N_{E}) of the conductor NEN_{E} of EE over \mathbb{Q}. In Serre’s original treatment [15], the following theorem was shown. By GRH, we mean the conjecture which applies to the Artin LL-functions of Galois extensions L/KL/K; unless otherwise stated, K=K=\mathbb{Q}.

Theorem 1.1.

[15, Theorem 21] Assume GRH. Let EE and EE^{\prime} be two elliptic curves defined over \mathbb{Q}. Suppose that EE and EE^{\prime} are not \mathbb{Q}-isogenous. Then there exists a prime pp of good reduction for EE and EE^{\prime} such that ap(E)ap(E)a_{p}(E)\neq a_{p}(E^{\prime}) and satisfying the inequality

(1.1) pC1(lograd(NENE))2(loglograd(2NENE))12,p\leq C_{1}(\log\operatorname{rad}(N_{E}N_{E^{\prime}}))^{2}(\log\log\operatorname{rad}(2N_{E}N_{E^{\prime}}))^{12},

where C1C_{1} is an absolute constant.

Based on the method used, the constant C1C_{1} here is unfortunately rather large. Recent work of Mayle-Wang [9] has given an explicit result on the smallest prime which achieves ap(E)ap(E)a_{p}(E)\neq a_{p}(E^{\prime}). The constants are quite small, and like Serre, depend on only knowledge of the primes of bad reduction of the two elliptic curves EE and EE^{\prime}.

Theorem 1.2.

[9, Theorem 2] Assume GRH. Let EE and EE^{\prime} be two elliptic curves over \mathbb{Q} without complex multiplication. Suppose EE and EE^{\prime} are not \mathbb{Q}-isogenous. Then there exists a prime pp of good reduction for EE and EE^{\prime} such that ap(E)ap(E)a_{p}(E)\neq a_{p}(E^{\prime}) and satisfying the inequality

(1.2) p(482lograd(2NENE)+2880)2,p\leq(482\log\operatorname{rad}(2N_{E}N_{E^{\prime}})+2880)^{2},

where NEN_{E} and NEN_{E^{\prime}} denote the conductors of EE and EE^{\prime}, respectively.

The method is based on using effective forms of the Chebotarev density theorem together with the Faltings-Serre method, in particular, by studying the ‘deviation group’ δ(G)\delta(G) of the 22-adic representations attached to two elliptic curves.

In our work, we explain how to replace δ(G)\delta(G) with smaller quotients in Mayle-Wang’s original arguments. Using these smaller quotients allows us to prove an improved effective isogeny theorem for elliptic curves over \mathbb{Q} with a certain condition on the mod 22 representations.

Theorem 1.3.

Assume GRH. Let EE and EE^{\prime} be two elliptic curves over \mathbb{Q}. Suppose EE and EE^{\prime} are not \mathbb{Q}-isogenous. Assume the mod 2 representations ρ¯E,2\bar{\rho}_{E,2} and ρ¯E,2\bar{\rho}_{E^{\prime},2} are not isomorphic, or if they are isomorphic that they are absolutely irreducible. Then there exists a prime pp of good reduction for EE and EE^{\prime} such that ap(E)ap(E)a_{p}(E)\neq a_{p}(E^{\prime}) and satisfying the inequality

(1.3) p(124lograd(2NENE)+561)2.p\leq(124\log\operatorname{rad}(2N_{E}N_{E^{\prime}})+561)^{2}.
Remark 1.4.

Mayle-Wang, in Theorem 1.2, include a hypothesis that the elliptic curves EE and EE^{\prime} be without complex multiplication; in Proposition 5.1 and Theorem 1.3, we have dropped this assumption. All we require here is the existence of a prime of good reduction such that ap(E)ap(E)a_{p}(E)\neq a_{p}(E^{\prime}), which is satisfied once we assume the two elliptic curves are not \mathbb{Q}-isogenous, a consequence of Faltings’ Theorem [7] (see translation in [6]).

We also prove another improved effective isogeny theorem which applies for elliptic curves over \mathbb{Q} which are quadratic twists of each other and do not have complex multiplication.

Theorem 1.5.

Assume GRH. Let EE and EE^{\prime} be two elliptic curves over \mathbb{Q} which are quadratic twists of each other and do not have complex multiplication. Suppose EE and EE^{\prime} are not \mathbb{Q}-isogenous. Then there exists a prime pp of good reduction for EE and EE^{\prime} such that ap(E)ap(E)a_{p}(E)\neq a_{p}(E^{\prime}) and satisfying the inequality

(1.4) p(223lograd(2NENE)+1127)2.p\leq(223\log\operatorname{rad}(2N_{E}N_{E^{\prime}})+1127)^{2}.

This version requires the results of Rouse and Zureick-Brown [12] which characterizes the images of the 22-adic representations attached to an elliptic curve over \mathbb{Q}.

Using Theorem 1.2, Mayle-Wang [9, Theorem 1] prove the following bound for Serre’s open image theorem.

Theorem 1.6.

Assume GRH. Let EE an elliptic curve over \mathbb{Q} without complex multiplication. Then

CE964lograd(2NE)+5760.C_{E}\leq 964\log\operatorname{rad}(2N_{E})+5760.

A consequence of our improved effective isogeny Theorems 1.3 and 1.5 is an improvement in the constants for Serre’s open image theorem.

Theorem 1.7.

Assume GRH. Let EE an elliptic curve over \mathbb{Q} without complex multiplication. Then

CE446lograd(2NE)+2254.C_{E}\leq 446\log\operatorname{rad}(2N_{E})+2254.

The Magma computational algebra system [2] was used for verifying assertions in this paper. The electronic resources are available from

https://github.com/ichensfuca/ChenSwidinsky.

2. Explicit forms of the Chebotarev density theorem

Let L/KL/K be a finite Galois extension with Galois group GG. Define the counting function πC(x,L/K)\pi_{C}(x,L/K), for a conjugacy class CC of the Galois group GG of L/KL/K, to be the function

πC(x,L/K)=|{𝔭𝔭 unramified in K,(L/K𝔭)=C,NK/(𝔭)x}|.\pi_{C}(x,L/K)=\left|\left\{\mathfrak{p}\mid\mathfrak{p}\text{ unramified in $K$},\left(\frac{L/K}{\mathfrak{p}}\right)=C,\,\text{N}_{K/\mathbb{Q}}(\mathfrak{p})\leq x\right\}\right|.
Theorem 2.1 (Chebotarev Density Theorem).

Let πC(x,L/K)\pi_{C}(x,L/K) be as above. Then,

πC(x,L/K)|C||G|Li(x).\pi_{C}(x,L/K)\sim\frac{|C|}{|G|}\operatorname{Li}(x).

Effective versions of Chebotarev’s Density Theorem exist as well and we shall be applying results in which the constants are explicitly computable in terms of the discriminants of LL and KK, as well as the degree of each extension over \mathbb{Q}. The first of these was given by Lagarias and Odlyzko [8]. Their result relies on the validity of GRH.

We now state the first explicit form of Theorem 2.1.

Theorem 2.2.

[8, Theorem 1.1] There exists an effectively computable positive absolute constant c1c_{1} such that if GRH holds for the Dedekind zeta function of LL, then for every x2x\geq 2,

|πC(x,L/K)|C||G|Li(x)|c1(|C||G|x1/2log(|dL|xnL)+log|dL|).|\pi_{C}(x,L/K)-\frac{|C|}{|G|}\operatorname{Li}(x)|\leq c_{1}\left(\frac{|C|}{|G|}x^{1/2}\log(|d_{L}|x^{n_{L}})+\log|d_{L}|\right).

An important corollary, one that we shall make use of, is finding an x0x_{0} such that πC(x0,L/K)>0\pi_{C}(x_{0},L/K)>0.

Corollary 2.3.

[8, Corollary 1.2] There exists an effectively computable positive absolute constant c2c_{2} such that if GRH holds for the Artin LL-functions of L/L/\mathbb{Q}, LL\neq\mathbb{Q}, then for every conjugacy class CC of GG there exists an unramified prime ideal 𝔭\mathfrak{p} of KK such that

(L/K𝔭)=C\left(\frac{L/K}{\mathfrak{p}}\right)=C

and

NK/(𝔭)c2(log|dL|)2(loglog|dL|)4.\text{N}_{K/\mathbb{Q}}(\mathfrak{p})\leq c_{2}(\log|d_{L}|)^{2}(\log\log|d_{L}|)^{4}.

If L=L=\mathbb{Q}, then 𝔭=(2)\mathfrak{p}=(2) is a solution.

The above is a non-nullity result about πC(x,L/K)\pi_{C}(x,L/K); it asserts the size of xx we must take to ensure that πC(x,L/K)\pi_{C}(x,L/K) is nonzero, that is, there is some unramified prime ideal 𝔭\mathfrak{p} of KK whose Artin symbol hits CC and with norm smaller than xx.

Theorem 2.4.

[10, Théorème 4] There exists an effectively computable positive absolute constant c3c_{3} such that if GRH and Artin’s Conjecture (AC) hold for the Artin L-functions of L/L/\mathbb{Q}, LL\neq\mathbb{Q}, then for every conjugacy class CC of GG, we have that

πC(x,L/K)1\pi_{C}(x,L/K)\geq 1

for all x2x\geq 2 such that xc3(log|dL|)2x\geq c_{3}(\log|d_{L}|)^{2}.

Oesterlé [10, Théorème 4] finds that c3=70c_{3}=70, although his proof was seemingly never published. An improvement to Lagarias and Odlyzko is given by Bach-Sorenson [1]:

Theorem 2.5.

[1, Theorem 5.1] Assume GRH. Let K/K/\mathbb{Q} be a Galois extension of number fields, with KK\neq\mathbb{Q}. Let dKd_{K} denote the discriminant of KK. Let nKn_{K} denote the degree of KK. Let CGal(K/)C\subseteq\operatorname{Gal}(K/\mathbb{Q}) be a nonempty subset closed under conjugation. Then, there is a prime pp of \mathbb{Q} unramified in KK with

(K/p)C,\left(\frac{K/\mathbb{Q}}{p}\right)\subseteq C,

satisfying

p(alog|dK|+bnK+c)2p\leq(a\log|d_{K}|+bn_{K}+c)^{2}

for some triple (a,b,c)(a,b,c) taken from [1, Table 3] according to the quantities log|dK|\log|d_{K}| and nKn_{K}. We may take a=4a=4, b=2.5b=2.5, and c=5c=5 to cover all cases of log|dK|\log|d_{K}| and nK=[K:]n_{K}=[K:\mathbb{Q}].

A corollary to the above is given in Mayle-Wang [9, Corollary 6] when we need to pick the prime pp to be coprime to a given positive integer mm.

Corollary 2.6.

[9, Corollary 6] Assume GRH. Let K/K/\mathbb{Q} be a Galois extension of number fields, with KK\neq\mathbb{Q}. Let mm be a positive integer, and set K~=K(m)\tilde{K}=K(\sqrt{m}). Denote dK~d_{\tilde{K}} to be the absolute value of the discriminant of K~\tilde{K}. Let nK~n_{\tilde{K}} denote the degree of K~\tilde{K}. Let CGal(K/)C\subseteq\operatorname{Gal}(K/\mathbb{Q}) be a nonempty subset that is closed under conjugation. Then there exists a prime number pp not dividing mm that is unramified in K/K/\mathbb{Q} with (K/p)C\left(\frac{K/\mathbb{Q}}{p}\right)\subseteq C and satisfying

p(a~log|dK~|+b~nK~+c~)2,p\leq(\tilde{a}\log|d_{\tilde{K}}|+\tilde{b}n_{\tilde{K}}+\tilde{c})^{2},

for some triple (a~,b~,c~)(\tilde{a},\tilde{b},\tilde{c}) taken from [1, Table 3] according to the quantities log|dK~|\log|d_{\tilde{K}}| and nK~n_{\tilde{K}}. We may take a~=4\tilde{a}=4, b~=2.5\tilde{b}=2.5, and c~=5\tilde{c}=5 to cover all cases of log|dK~|\log|d_{\tilde{K}}| and nK~=[K~:]n_{\tilde{K}}=[\tilde{K}:\mathbb{Q}].

For a fixed nKn_{K}, we say a triple (a,b,c)(a,b,c) is bigger than a triple (a,b,c)(a^{\prime},b^{\prime},c^{\prime}) (resp. a triple (a,b,c)(a^{\prime},b^{\prime},c^{\prime}) is smaller than a triple (a,b,c)(a,b,c)) if

(alog|dK|+bnK+c)2(alog|dK|+bnK+c)2(a^{\prime}\log|d_{K}|+b^{\prime}n_{K}+c^{\prime})^{2}\leq(a\log|d_{K}|+bn_{K}+c)^{2}

for all values of log|dK|\log|d_{K}| in a row of [1, Table 3].

We give our own version of Theorem 2.5 and Corollary 2.6. The idea is to collapse [1, Table 3] into a 1-dimensional table, removing the condition on log|dK~|\log|d_{\tilde{K}}| so that each triple is valid for a range of nK~n_{\tilde{K}}. We do this by picking a “pivot” triple for each column, for which all triples appearing before the pivot are absorbed into a special constant p0(nK~)p_{0}(n_{\tilde{K}}), and all triples appearing after are checked to be smaller than the pivot triple.

The pivot triple in each column of [1, Table 3] is chosen to be the first one so that (2.3) holds.

nK~n_{\tilde{K}} (a¯,b¯,c¯)(\bar{a},\bar{b},\bar{c})
2 (1.446,0.23,6.8)(1.446,0.23,6.8)
3-4 (1.527,0.17,6.4)(1.527,0.17,6.4)
5-9 (1.629,0.11,6.1)(1.629,0.11,6.1)
10-14 (1.667,0.09,6.0)(1.667,0.09,6.0)
15-49 (1.745,0.04,5.8)(1.745,0.04,5.8)
50-128 (1.755,0,5.7)(1.755,0,5.7)
Table 1. Triples (a¯,b¯,c¯)(\bar{a},\bar{b},\bar{c}) as appearing in Proposition 2.7.
Proposition 2.7.

Assume GRH. Let K/K/\mathbb{Q} be a Galois extension of number fields with KK\neq\mathbb{Q}. Let mm be a positive integer, and set K~=K(m)\tilde{K}=K(\sqrt{m}). Denote dK~d_{\tilde{K}} to be the discriminant of K~\tilde{K}. Let nK~n_{\tilde{K}} denote the degree of K~/\tilde{K}/\mathbb{Q}. Let CGal(K/)C\subseteq\operatorname{Gal}(K/\mathbb{Q}) be a nonempty subset that is closed under conjugation. Then there exists a triple (a¯,b¯,c¯)(\bar{a},\bar{b},\bar{c}) taken from Table 1, a special constant p0(nK~)p_{0}(n_{\tilde{K}}), and a prime number pp not dividing mm that is unramified in K/K/\mathbb{Q} with (K/p)C\left(\frac{K/\mathbb{Q}}{p}\right)\subseteq C and satisfying

(2.1) p\displaystyle p max((a¯log|dK~|+b¯nK~+c¯)2,p0(nK~))\displaystyle\leq\max((\bar{a}\log|d_{\tilde{K}}|+\bar{b}n_{\tilde{K}}+\bar{c})^{2},p_{0}(n_{\tilde{K}}))
(2.2) (a¯((n01)lograd(dK~)+n0logn0)+b¯n0+c¯)2,\displaystyle\leq(\bar{a}\cdot((n_{0}-1)\log\operatorname{rad}(d_{\tilde{K}})+n_{0}\log n_{0})+\bar{b}\cdot n_{0}+\bar{c})^{2},

where n0=max(72,nK~)n_{0}=\max(72,n_{\tilde{K}}). If we only have an upper bound for nK~n1n_{\tilde{K}}\leq n_{1}, then we have to replace each of a¯,b¯,c¯\bar{a},\bar{b},\bar{c} with the maximum of their values over entries in Table 1 with nK~n1n_{\tilde{K}}\leq n_{1}, respectively, and n0n_{0} with max(72,n1)\max(72,n_{1}).

Proof.

We have written a program in Magma which verifies the required inequalities (2.1). For inequality (2.2), there are two parts to check. The program checks that

(2.3) p0(nK~)(a¯((n01)log2+n0logn0)+b¯n0+c¯)2,p_{0}(n_{\tilde{K}})\leq(\bar{a}\cdot((n_{0}-1)\log 2+n_{0}\log n_{0})+\bar{b}\cdot n_{0}+\bar{c})^{2},

and the inequality

(2.4) (a¯log|dK~|+b¯nK~+c¯)2(a¯((n01)lograd(dK~)+n0logn0)+b¯n0+c¯)2(\bar{a}\log|d_{\tilde{K}}|+\bar{b}n_{\tilde{K}}+\bar{c})^{2}\leq(\bar{a}\cdot((n_{0}-1)\log\operatorname{rad}(d_{\tilde{K}})+n_{0}\log n_{0})+\bar{b}\cdot n_{0}+\bar{c})^{2}

follows from Lemma 2.8. ∎

Lemma 2.8.

[9, Lemma 7] If K/K/\mathbb{Q} is a nontrivial finite Galois extension, then

(12log3)[K:]log|dK|([K:]1)lograd(dK)+[K:]log([K:]),\left(\frac{1}{2}\log 3\right)[K:\mathbb{Q}]\leq\log|d_{K}|\leq([K:\mathbb{Q}]-1)\log\operatorname{rad}(d_{K})+[K:\mathbb{Q}]\log([K:\mathbb{Q}]),

where dKd_{K} is the absolute value of the discriminant of KK.

3. The deviation group δ(G)\delta(G)

In this section, we wish to construct a finite group, called the deviation group, denoted δ(G)\delta(G), from which we can find a finite subset that will determine if the two representations are isomorphic or not.

Our treatment of the deviation group will follow the exposition given in Ignasi’s thesis [11]. We note that Ignasi’s exposition is, itself, taken from Chênevert’s thesis [5], whose work follows the work of Serre [16] (the propositions and lemmas which appear here, with the exception of Lemma 3.8, can also be found in [5, Chapter 5]).

Let GG be a group, and LL be a finite extension of \mathbb{Q}_{\ell}, for \ell prime, with ring of integers 𝒪λ\mathcal{O}_{\lambda}, maximal ideal λ\lambda, and residue field k=𝒪λ/λ𝒪λk=\mathcal{O}_{\lambda}/\lambda\mathcal{O}_{\lambda}. We let π\pi be a uniformizer, so λ=π𝒪λ\lambda=\pi\mathcal{O}_{\lambda}. Let ρ1,ρ2:GGLn(𝒪λ)\rho_{1},\rho_{2}:G\to\operatorname{GL}_{n}(\mathcal{O}_{\lambda}) be two λ\lambda-adic representations. We begin by extending the map ρ1×ρ2:GGLn(𝒪λ)×GLn(𝒪λ)\rho_{1}\times\rho_{2}:G\to\operatorname{GL}_{n}(\mathcal{O}_{\lambda})\times\operatorname{GL}_{n}(\mathcal{O}_{\lambda}) from GG to the group ring 𝒪λ[G]\mathcal{O}_{\lambda}[G].

We define the map ρ:𝒪λ[G]Mn(𝒪λ)Mn(𝒪λ)\rho:\mathcal{O}_{\lambda}[G]\to M_{n}(\mathcal{O}_{\lambda})\oplus M_{n}(\mathcal{O}_{\lambda}) to be

ρ(aigi)=(aiρ1(gi),aiρ2(gi)).\rho\left(\sum a_{i}g_{i}\right)=\left(\sum a_{i}\rho_{1}(g_{i}),\sum a_{i}\rho_{2}(g_{i})\right).

Let MM be the full image of ρ\rho inside Mn(𝒪λ)Mn(𝒪λ)M_{n}(\mathcal{O}_{\lambda})\oplus M_{n}(\mathcal{O}_{\lambda}), and consider the composition map δ:G𝜌M×(M/λM)×\delta:G\xrightarrow{\rho}M^{\times}\to(M/\lambda M)^{\times}.

Definition 3.1.

[11, Definition 2.1.1] The image δ(G)\delta(G) of GG inside (M/λM)×(M/\lambda M)^{\times} is called the deviation group of the pair of representations ρ1,ρ2\rho_{1},\rho_{2}.

Remark 3.2.

Since MM is a subalgebra of R=Mn(𝒪λ)×Mn(𝒪λ)R=M_{n}(\mathcal{O}_{\lambda})\times M_{n}(\mathcal{O}_{\lambda}), it might be tempting to think δ(G)\delta(G) is a subgroup of (R/λR)×=GL2(k)×GL2(k)(R/\lambda R)^{\times}=\operatorname{GL}_{2}(k)\times\operatorname{GL}_{2}(k) but this may not be the case. See the remark after [11, Definition 2.1.1].

The deviation group turns out to be finite, as described by the following proposition.

Proposition 3.3.

[11, Proposition 2.1.2] The group δ(G)\delta(G) is finite, and in particular we have |δ(G)||k|2n2|\delta(G)|\leq|k|^{2n^{2}}.

Proof.

MM is a submodule of the free 𝒪λ\mathcal{O}_{\lambda}-module Mn(𝒪λ)Mn(𝒪λ)M_{n}(\mathcal{O}_{\lambda})\oplus M_{n}(\mathcal{O}_{\lambda}). Since 𝒪λ\mathcal{O}_{\lambda} is a local ring, MM is free and is of rank rr, where rr satisfies

rrank(Mn(𝒪λ)Mn(𝒪λ))=2n2.r\leq\operatorname{rank}(M_{n}(\mathcal{O}_{\lambda})\oplus M_{n}(\mathcal{O}_{\lambda}))=2n^{2}.

Given MM is a 𝒪λ\mathcal{O}_{\lambda}-module, M/λMM/\lambda M is a kk-algebra of dimension rr. Hence,

|δ(G)||(M/λM)×||k|r|k|2n2|\delta(G)|\leq|(M/\lambda M)^{\times}|\leq|k|^{r}\leq|k|^{2n^{2}}

as claimed. ∎

Remark 3.4.

A similar bound on |δ(G)||\delta(G)| is employed by Mayle-Wang in their proof of Theorem 1.2, although they do not explicitly mention the deviation group. See the proof in [9, Theorem 2].

Let us turn our attention now to the practical use of δ(G)\delta(G), that being its ability to help us determine when two representations are isomorphic.

Proposition 3.5.

[11, Proposition 2.1.3] Let ΣG\Sigma\subseteq G be a subset that surjects onto δ(G)\delta(G). Then, ρ1ρ2\rho_{1}\sim\rho_{2} if and only if tr(ρ1(g))=tr(ρ2(g))\operatorname{tr}(\rho_{1}(g))=\operatorname{tr}(\rho_{2}(g)) for all gΣg\in\Sigma.

Before we introduce the next proposition, some further explanations are needed (following [11]). We assume now the representations ρ1,ρ2:GGLn(𝒪λ)\rho_{1},\rho_{2}:G\to\operatorname{GL}_{n}(\mathcal{O}_{\lambda}) are not isomorphic, that is, they are not conjugate in GLn(𝒪λ)\operatorname{GL}_{n}(\mathcal{O}_{\lambda}), but that the residual representations ρ¯1\bar{\rho}_{1} and ρ¯2\bar{\rho}_{2} obtained from ρ1\rho_{1} and ρ2\rho_{2} by reduction modulo λ\lambda are isomorphic. We then have an equality ρ¯1=Pρ¯2P1\bar{\rho}_{1}=P\bar{\rho}_{2}P^{-1} for some matrix PMn(k)P\in M_{n}(k).

Define β\beta to be the largest integer such that ρ1\rho_{1} and ρ2\rho_{2} are conjugated modulo λβ\lambda^{\beta}, that is, there is a matrix PGLn(𝒪λ)P\in\operatorname{GL}_{n}(\mathcal{O}_{\lambda}) such that ρ1Pρ2P1(modλβ)\rho_{1}\equiv P\rho_{2}P^{-1}\ (\mathrm{mod}\ \lambda^{\beta}); we then have β1\beta\geq 1, since ρ¯1ρ¯2\bar{\rho}_{1}\cong\bar{\rho}_{2}. In addition, there is an integer α1\alpha\geq 1 such that tr(ρ1)tr(ρ2)(modλα)\operatorname{tr}(\rho_{1})\equiv\operatorname{tr}(\rho_{2})\ (\mathrm{mod}\ \lambda^{\alpha}) and tr(ρ1)tr(ρ2)(modλα+1)\operatorname{tr}(\rho_{1})\not\equiv\operatorname{tr}(\rho_{2})\ (\mathrm{mod}\ \lambda^{\alpha+1}); in particular, ρ1\rho_{1} and ρ2\rho_{2} are not conjugate modulo λα+1\lambda^{\alpha+1}, so βα\beta\leq\alpha. Given that ρ1\rho_{1} and ρ2\rho_{2} are conjugate modulo λβ\lambda^{\beta} but not conjugate modulo λβ+1\lambda^{\beta+1}, if we replace ρ2\rho_{2} with a conjugate we may assume ρ1ρ2(modλβ)\rho_{1}\equiv\rho_{2}\ (\mathrm{mod}\ \lambda^{\beta}) but ρ1ρ2(modλβ+1)\rho_{1}\not\equiv\rho_{2}\ (\mathrm{mod}\ \lambda^{\beta+1}).

Hence, for any gGg\in G, we have

(3.1) ρ2(g)ρ1(g)0(modλβ)ρ2(g)ρ1(g)=θgπβ\rho_{2}(g)-\rho_{1}(g)\equiv 0\ (\mathrm{mod}\ \lambda^{\beta})\Rightarrow\rho_{2}(g)-\rho_{1}(g)=\theta_{g}\pi^{\beta}

for some θgMn(𝒪λ)\theta_{g}\in M_{n}(\mathcal{O}_{\lambda}) and π\pi a uniformizer of λ\lambda. Rearranging, we get an equation for ρ2(g)\rho_{2}(g) of the form

(3.2) ρ2(g)=(In+θgπβρ1(g)1)ρ1(g),\rho_{2}(g)=(I_{n}+\theta_{g}\pi^{\beta}\rho_{1}(g)^{-1})\rho_{1}(g),

where InI_{n} is the n×nn\times n identity matrix. Let θ:GMn(𝒪λ)\theta:G\to M_{n}(\mathcal{O}_{\lambda}) be the map gθgρ1(g)1g\to\theta_{g}\rho_{1}(g)^{-1}, and notice that (3.2) becomes

(3.3) ρ2(g)=(In+πβθ(g))ρ1(g).\rho_{2}(g)=(I_{n}+\pi^{\beta}\theta(g))\rho_{1}(g).
Proposition 3.6.

[11, Proposition 2.2.1] Let ρ1,ρ2:GGLn(𝒪λ)\rho_{1},\rho_{2}:G\to\operatorname{GL}_{n}(\mathcal{O}_{\lambda}) be representations that are not isomorphic, and suppose ρ¯1,ρ¯2:GGLn(k)\bar{\rho}_{1},\bar{\rho}_{2}:G\to\operatorname{GL}_{n}(k) are isomorphic. Let β\beta be the largest integer such that ρ1\rho_{1} and ρ2\rho_{2} are conjugate modulo λβ\lambda^{\beta}, and as above, assume ρ2\rho_{2} has been replaced by a conjugate such that ρ1ρ2(modλβ)\rho_{1}\equiv\rho_{2}\ (\mathrm{mod}\ \lambda^{\beta}). Let

(3.4) φ:GMn(k)GLn(k)g(θ(g)(modλ),ρ1(g)(modλ))\begin{split}\varphi:G&\to M_{n}(k)\rtimes\operatorname{GL}_{n}(k)\\ &g\mapsto(\theta(g)\ (\mathrm{mod}\ \lambda),\rho_{1}(g)\ (\mathrm{mod}\ \lambda))\end{split}

where the semidirect product is with respect to the action of GLn(k)\operatorname{GL}_{n}(k) on Mn(k)M_{n}(k) by conjugation, that is multiplication is given by

(A,B)(C,D)=(A+BCB1,BD).(A,B)\cdot(C,D)=(A+BCB^{-1},BD).

Then φ\varphi is a group homomorphism which factors through the deviation group δ(G)\delta(G).

Remark 3.7.

The homomorphism δ(G)φ(G)\delta(G)\twoheadrightarrow\varphi(G) may not be injective. See [11, Remark 2.2.2].

Lastly, we state a general lemma regarding determinants of matrices that we shall employ later.

Lemma 3.8.

[11, Lemma 2.2.3] Let RR be a discrete valuation ring with uniformizer π\pi, and FF its field of fractions. For any AGLn(F)A\in\operatorname{GL}_{n}(F),

det(In+πA)=1+πtr(A)+O(π2).\det(I_{n}+\pi A)=1+\pi\operatorname{tr}(A)+O(\pi^{2}).

It can be difficult to compute the exact size of δ(G)\delta(G), or find a tighter upper bound for it. We will, in the following section, work to replace δ(G)\delta(G) with φ(G)\varphi(G) in the case of 22-adic representations. The codomain of φ\varphi is easily understood, and hence a bound for |φ(G)||\varphi(G)| is easily computable. This is what allows us to prove Theorem 1.3.

4. The tools of Mayle-Wang

The methodology of Mayle-Wang relies on the following proposition that is due to Serre (a proof of which can be found in [3, Theorem 4.7]). The proposition which follows is a refined version of Serre’s original argument due to Mayle-Wang [9, Proposition 12] in which we have reworked the statement and proof to follow the work and notation done in Section 3. We note that the statement is similar to that of Proposition 3.5: here, we show that if the representations are not isomorphic, then their traces must disagree on some finite set. While the proofs are very similar, the advantage of the following proposition is that it is in a form to which we may readily apply Chebotarev.

Proposition 4.1.

[9, Proposition 12] Let nn a positive integer. Let GG be a group and ρ1,ρ2:GGLn(𝒪λ)\rho_{1},\rho_{2}:G\rightarrow\operatorname{GL}_{n}(\mathcal{O}_{\lambda}) be representations, and δ(G)\delta(G) the deviation group of GG with respect to the two representations ρ1\rho_{1} and ρ2\rho_{2}. Suppose that there exists an element gGg\in G such that trρ1(g)trρ2(g)\operatorname{tr}\rho_{1}(g)\not=\operatorname{tr}\rho_{2}(g). Then there exists a subset Cδ(G)C\subseteq\delta(G) for which

  1. (1)

    the set CC is non-empty and closed under conjugation by δ(G)\delta(G), and

  2. (2)

    if the image in δ(G)\delta(G) of an element gGg\in G belongs to CC, then trρ1(g)trρ2(g)\operatorname{tr}\rho_{1}(g)\not=\operatorname{tr}\rho_{2}(g).

Proof.

Let R:=Mn(𝒪λ)×Mn(𝒪λ)R:=M_{n}(\mathcal{O}_{\lambda})\times M_{n}(\mathcal{O}_{\lambda}). Let MM denote the 𝒪λ\mathcal{O}_{\lambda}-subalgebra of RR generated by the image of GG under the product map

ρ1×ρ2:GGLn(𝒪λ)×GLn(𝒪λ).\rho_{1}\times\rho_{2}:G\to\operatorname{GL}_{n}(\mathcal{O}_{\lambda})\times\operatorname{GL}_{n}(\mathcal{O}_{\lambda}).

Recall that δ(G)\delta(G) is the image of GG under ρ1×ρ2\rho_{1}\times\rho_{2} in M/λMM/\lambda M.

Let α\alpha be the largest nonnegative integer such that for each gGg\in G, one has that

tr(ρ1(g))tr(ρ2(g))(modλα).\operatorname{tr}(\rho_{1}(g))\equiv\operatorname{tr}(\rho_{2}(g))\ (\mathrm{mod}\ \lambda^{\alpha}).

As MM is a 𝒪λ\mathcal{O}_{\lambda}-subalgebra generated by the image of GG under ρ1×ρ2\rho_{1}\times\rho_{2}, it follows that the congruence trx1trx2(modλα)\operatorname{tr}x_{1}\equiv\operatorname{tr}x_{2}\ (\mathrm{mod}\ \lambda^{\alpha}) holds for each pair (x1,x2)M(x_{1},x_{2})\in M. We obtain the 𝒪λ\mathcal{O}_{\lambda}-module homomorphism ϕ:M𝒪λ\phi:M\to\mathcal{O}_{\lambda} given by

ϕ(x1,x2)=λα(tr(x2)tr(x1)).\phi(x_{1},x_{2})=\lambda^{-\alpha}(\operatorname{tr}(x_{2})-\operatorname{tr}(x_{1})).

Since ϕ(λM)λ𝒪λ\phi(\lambda M)\subseteq\lambda\mathcal{O}_{\lambda}, we may consider the induced 𝒪λ/λ𝒪λ\mathcal{O}_{\lambda}/\lambda\mathcal{O}_{\lambda}-module homomorphism ϕ¯:M/λM𝒪λ/λ𝒪λ\bar{\phi}:M/\lambda M\to\mathcal{O}_{\lambda}/\lambda\mathcal{O}_{\lambda}.

Let C=δ(G)kerϕ¯C=\delta(G)\setminus\ker\bar{\phi} be the set of elements in δ(G)\delta(G) whose image under ϕ¯\bar{\phi} in M/λMM/\lambda M all are nonzero. From the definition of α\alpha and the linearity of the trace map, there exists g0Gg_{0}\in G such that

tr(ρ1(g0))tr(ρ2(g0))(modλα+1).\operatorname{tr}(\rho_{1}(g_{0}))\not\equiv\operatorname{tr}(\rho_{2}(g_{0}))\ (\mathrm{mod}\ \lambda^{\alpha+1}).

Note that the image of (ρ1×ρ2)(g0)(\rho_{1}\times\rho_{2})(g_{0}) in δ(G)\delta(G) is contained in CC, so CC is nonempty. Also, CC is closed under conjugation since the trace map is invariant under conjugation.

Finally, suppose that gGg\in G is such that the image of gg in δ(G)\delta(G) is contained in CC. Then, ϕ(ρ1×ρ2(g))λ𝒪λ\phi(\rho_{1}\times\rho_{2}(g))\not\in\lambda\mathcal{O}_{\lambda}, and in particular trρ1(g)trρ2(g)\operatorname{tr}\rho_{1}(g)\neq\operatorname{tr}\rho_{2}(g). ∎

Remark 4.2.

In the notation δ(G)\delta(G) and φ(G)\varphi(G) we suppress the dependence on the representations ρ1\rho_{1} and ρ2\rho_{2} as they are usually fixed in the context.

We now give an analogous version of Proposition 4.1 in the case where the mod 2 representations are isomorphic and absolutely irreducible. This allows us to replace δ(G)\delta(G) in Proposition 4.1 with φ(G)\varphi(G) from Proposition 3.6, a set which is easier to estimate the size of. The idea to replace δ(G)\delta(G) with φ(G)\varphi(G) comes from Chênevert [5, pg. 114], in which he gives a remark that, in the 22-adic case, Serre [16] implies that δ(G)φ(G)\delta(G)\cong\varphi(G). However, in a conversation with Chênevert, Serre mentions he might not have proven the map δ(G)φ(G)\delta(G)\to\varphi(G) was an isomorphism, but, in an unpublished letter to Tate, that the α\alpha in the proof of Proposition 3.5 is equal to the β\beta coming from the construction of the function φ\varphi if the residual representation is surjective. We show, in the 2-adic case, that α=β\alpha=\beta, and that we can replace δ(G)\delta(G) in Proposition 4.1 with φ(G)\varphi(G) and get the same conclusion, that is, there is a subset Cφ(G)C\subseteq\varphi(G) that is a conjugacy class, and if gGg\in G is such that φ(g)C\varphi(g)\in C, then trρ1(g)trρ2(g)\operatorname{tr}\rho_{1}(g)\neq\operatorname{tr}\rho_{2}(g).

In order to prove this special case, we require a theorem of Carayol [4].

Theorem 4.3.

[4, Theorem 1] Let AA be a local ring, RR an AA-algebra, and let ρ1,ρ2:RMn(A)\rho_{1},\rho_{2}:R\to M_{n}(A) be two representations of RR of the same dimension nn. Suppose that the residual representation ρ¯:RAFMn(F)\bar{\rho}:R\otimes_{A}F\to M_{n}(F), where FF is the residue field of AA, is absolutely irreducible. Suppose that the traces for ρ1\rho_{1} and ρ2\rho_{2} are the same for every rRr\in R. Then, ρ1\rho_{1} and ρ2\rho_{2} are isomorphic as representations, that is, there exists a matrix QGLn(A)Q\in\operatorname{GL}_{n}(A) such that ρ1(r)=Qρ2(r)Q1\rho_{1}(r)=Q\rho_{2}(r)Q^{-1} for all rRr\in R.

We now prove the special case.

Proposition 4.4.

Let nn be a positive integer. Let GG be a group and ρ1,ρ2:GGLn(2)\rho_{1},\rho_{2}:G\rightarrow\operatorname{GL}_{n}(\mathbb{Z}_{2}) be representations, and suppose their reductions ρ¯1,ρ¯2\bar{\rho}_{1},\bar{\rho}_{2} are isomorphic and absolutely irreducible. Suppose that there exists an element gGg\in G such that trρ1(g)trρ2(g)\operatorname{tr}\rho_{1}(g)\not=\operatorname{tr}\rho_{2}(g). Then there exists a subset Cφ(G)C\subseteq\varphi(G) for which

  1. (1)

    the set CC is non-empty and closed under conjugation by φ(G)\varphi(G), and

  2. (2)

    if the image in φ(G)\varphi(G) of an element gGg\in G belongs to CC, then trρ1(g)trρ2(g)\operatorname{tr}\rho_{1}(g)\not=\operatorname{tr}\rho_{2}(g).

Proof.

Our setup begins, as it did, in Section 3. Let α\alpha be the largest nonnegative integer such that for each gGg\in G, we have

tr(ρ1(g))tr(ρ2(g))(mod 2α)andtr(ρ1(g))tr(ρ2(g))(mod 2α+1).\operatorname{tr}(\rho_{1}(g))\equiv\operatorname{tr}(\rho_{2}(g))\ (\mathrm{mod}\ 2^{\alpha})\quad\text{and}\quad\operatorname{tr}(\rho_{1}(g))\not\equiv\operatorname{tr}(\rho_{2}(g))\ (\mathrm{mod}\ 2^{\alpha+1}).

In addition, we let β\beta be the largest integer such that ρ1\rho_{1} and ρ2\rho_{2} are conjugated modulo 2β2^{\beta}, that is, there is a matrix PGLn(2)P\in\operatorname{GL}_{n}(\mathbb{Z}_{2}) such that ρ1Pρ2P1(mod 2β)\rho_{1}\equiv P\rho_{2}P^{-1}\ (\mathrm{mod}\ 2^{\beta}). As demonstrated before, we have βα\beta\leq\alpha. Also, given that ρ1\rho_{1} and ρ2\rho_{2} are conjugate modulo 2β2^{\beta} but not conjugate modulo λβ+1\lambda^{\beta+1}, if we replace ρ2\rho_{2} with a conjugate Pρ2P1P\rho_{2}P^{-1} for PGLn(2)P\in\operatorname{GL}_{n}(\mathbb{Z}_{2}), we may assume

(4.1) ρ1Pρ2P1(mod 2β) and ρ1Pρ2P1(mod 2β+1).\rho_{1}\equiv P\rho_{2}P^{-1}\ (\mathrm{mod}\ 2^{\beta})\text{ and }\rho_{1}\not\equiv P\rho_{2}P^{-1}\ (\mathrm{mod}\ 2^{\beta+1}).

This implies Pρ2(g)P1ρ1(g)0(mod 2β)P\rho_{2}(g)P^{-1}-\rho_{1}(g)\equiv 0\ (\mathrm{mod}\ 2^{\beta}) for any gGg\in G. In particular, we get Pρ2(g)P1ρ1(g)=θg2βP\rho_{2}(g)P^{-1}-\rho_{1}(g)=\theta_{g}2^{\beta} for some θgMn(2)\theta_{g}\in M_{n}(\mathbb{Z}_{2}), which we can write as

(4.2) θg=Pρ2(g)P1ρ1(g)2β.\theta_{g}=\frac{P\rho_{2}(g)P^{-1}-\rho_{1}(g)}{2^{\beta}}.

In particular, note that

(4.3) tr(θg)=2β(tr(Pρ2(g)P1)tr(ρ1(g)))=2β(tr(ρ2(g))tr(ρ1(g)))\operatorname{tr}(\theta_{g})=2^{-\beta}(\operatorname{tr}(P\rho_{2}(g)P^{-1})-\operatorname{tr}(\rho_{1}(g)))=2^{-\beta}(\operatorname{tr}(\rho_{2}(g))-\operatorname{tr}(\rho_{1}(g)))

by the invariance of trace under conjugation.

We now show α=β\alpha=\beta. Extend the maps ρ1,ρ2\rho_{1},\rho_{2} to the group ring /2α[G]\mathbb{Z}/2^{\alpha}\mathbb{Z}[G] by ρi(ajgj)=ajρi(gj)\rho_{i}(\sum a_{j}g_{j})=\sum a_{j}\rho_{i}(g_{j}), for i=1,2i=1,2 and aj/2αa_{j}\in\mathbb{Z}/2^{\alpha}\mathbb{Z} and gjGg_{j}\in G. Then, notice that

(4.4) tr(ρ1(ajgj))(mod 2α)tr(ajρ1(gj))(mod 2α)ajtr(ρ1(gj))(mod 2α)ajtr(ρ2(gj))(mod 2α)tr(ρ2(ajgj))(mod 2α).\begin{split}\operatorname{tr}(\rho_{1}(\sum a_{j}g_{j}))\ (\mathrm{mod}\ 2^{\alpha})&\equiv\operatorname{tr}(\sum a_{j}\rho_{1}(g_{j}))\ (\mathrm{mod}\ 2^{\alpha})\\ &\equiv\sum a_{j}\operatorname{tr}(\rho_{1}(g_{j}))\ (\mathrm{mod}\ 2^{\alpha})\\ &\equiv\sum a_{j}\operatorname{tr}(\rho_{2}(g_{j}))\ (\mathrm{mod}\ 2^{\alpha})\\ &\equiv\operatorname{tr}(\rho_{2}(\sum a_{j}g_{j}))\ (\mathrm{mod}\ 2^{\alpha}).\end{split}

Since we satisfy the hypotheses of Theorem 4.3 with A=/2αA=\mathbb{Z}/2^{\alpha}\mathbb{Z} and R=/2α[G]R=\mathbb{Z}/2^{\alpha}\mathbb{Z}[G], we can find a matrix QGLn(/2α)Q\in\operatorname{GL}_{n}(\mathbb{Z}/2^{\alpha}\mathbb{Z}) such that ρ1(g)Qρ2(g)Q1(mod 2α)\rho_{1}(g)\equiv Q\rho_{2}(g)Q^{-1}\ (\mathrm{mod}\ 2^{\alpha}) for all gGg\in G. However, β\beta is the largest integer such that ρ1\rho_{1} and ρ2\rho_{2} are conjugate modulo 2β2^{\beta}, so αβ\alpha\leq\beta, implying α=β\alpha=\beta.

Recall, from (3.4), the map φ:GMn(𝔽2)GLn(𝔽2)\varphi:G\to M_{n}(\mathbb{F}_{2})\rtimes\operatorname{GL}_{n}(\mathbb{F}_{2}) is defined by

(4.5) φ(g)=(θ(g)(mod 2),ρ1(g)(mod 2)))=([θgρ1(g)1]2,[ρ1(g)]2).\varphi(g)=(\theta(g)\ (\mathrm{mod}\ 2),\rho_{1}(g)\ (\mathrm{mod}\ 2)))=([\theta_{g}\rho_{1}(g)^{-1}]_{2},[\rho_{1}(g)]_{2}).

We note our use of the notation [N]2[N]_{2}, for NMn(2)N\in M_{n}(\mathbb{Z}_{2}), to denote the residue class of NN modulo 22.

Define the map ϕ:Mn(𝔽2)GLn(𝔽2)𝔽2\phi^{\prime}:M_{n}(\mathbb{F}_{2})\rtimes\operatorname{GL}_{n}(\mathbb{F}_{2})\to\mathbb{F}_{2} by

(4.6) ϕ((A,B))=tr(AB)\phi^{\prime}((A,B))=\operatorname{tr}(AB)

where the product of matrices is taken to be the action of GLn(𝔽2)\operatorname{GL}_{n}(\mathbb{F}_{2}) on Mn(𝔽2)M_{n}(\mathbb{F}_{2}), and the trace is considered to be modulo 22. By (4.5), notice

(4.7) ϕ(φ(g))=tr([θgρ1(g)1]2[ρ1(g)]2)=tr([θgρ1(g)1ρ1(g)]2)=tr([θg]2)=[tr(θg)]2=[2α(tr(ρ2(g))tr(ρ1(g)))]2\begin{split}\phi^{\prime}(\varphi(g))&=\operatorname{tr}([\theta_{g}\rho_{1}(g)^{-1}]_{2}[\rho_{1}(g)]_{2})\\ &=\operatorname{tr}([\theta_{g}\rho_{1}(g)^{-1}\rho_{1}(g)]_{2})\\ &=\operatorname{tr}([\theta_{g}]_{2})\\ &=[\operatorname{tr}(\theta_{g})]_{2}\\ &=[2^{-\alpha}(\operatorname{tr}(\rho_{2}(g))-\operatorname{tr}(\rho_{1}(g)))]_{2}\end{split}

where we have used (4.3) above (with β\beta replaced with α\alpha, since α=β\alpha=\beta) and the fact that, for a matrix NMn(2)N\in M_{n}(\mathbb{Z}_{2}) with tr(N)=i=1naii\operatorname{tr}(N)=\sum_{i=1}^{n}a_{ii} for entries aii2a_{ii}\in\mathbb{Z}_{2} along the diagonal, we have

(4.8) tr([N]2)=i=1n[aii]2=[i=1naii]2=[tr(N)]2\begin{split}\operatorname{tr}([N]_{2})&=\sum_{i=1}^{n}[a_{ii}]_{2}\\ &=\left[\sum_{i=1}^{n}a_{ii}\right]_{2}\\ &=\left[\operatorname{tr}(N)\right]_{2}\end{split}

which shows the final equality (noting our use of [x]2[x]_{2} in (4.8) to denote the residue class of a 22-adic integer x2x\in\mathbb{Z}_{2}).

Let CC be the set of elements in φ(G)\varphi(G) that take a nonzero value under the map ϕ\phi^{\prime}. From the definition of α\alpha and the linearity of the trace map, there exists g0Gg_{0}\in G such that

tr(ρ1(g0))tr(ρ2(g0))(mod 2α+1).\operatorname{tr}(\rho_{1}(g_{0}))\not\equiv\operatorname{tr}(\rho_{2}(g_{0}))\ (\mathrm{mod}\ 2^{\alpha+1}).

Note that the image of g0g_{0} in φ(G)\varphi(G) is inside CC, so CC is nonempty. In addition, let ϕ(h)ϕ(G)\phi(h)\in\phi(G) for some hGh\in G; then, given φ\varphi is a homomorphism (Proposition 3.6) and by (4.7) and the invariance under conjugation of the trace map,

(4.9) ϕ(φ(h)φ(g)φ(h)1)=ϕ(φ(hgh1))=[2α(tr(ρ2(hgh1))tr(ρ1(hgh1)))]2=[2α(tr(ρ2(h)ρ2(g)ρ2(h)1)tr(ρ1(h)ρ1(g)ρ1(h)1))]2=[2α(tr(ρ2(g))tr(ρ1(g)))]2=ϕ(φ(g))0,\begin{split}\phi^{\prime}(\varphi(h)\varphi(g)\varphi(h)^{-1})&=\phi^{\prime}(\varphi(hgh^{-1}))\\ &=[2^{-\alpha}(\operatorname{tr}(\rho_{2}(hgh^{-1}))-\operatorname{tr}(\rho_{1}(hgh^{-1})))]_{2}\\ &=[2^{-\alpha}(\operatorname{tr}(\rho_{2}(h)\rho_{2}(g)\rho_{2}(h)^{-1})-\operatorname{tr}(\rho_{1}(h)\rho_{1}(g)\rho_{1}(h)^{-1}))]_{2}\\ &=[2^{-\alpha}(\operatorname{tr}(\rho_{2}(g))-\operatorname{tr}(\rho_{1}(g)))]_{2}\\ &=\phi^{\prime}(\varphi(g))\\ &\neq 0,\end{split}

so CC is closed under conjugation. Finally, suppose that gGg\in G is such that the image of gg in φ(G)\varphi(G) is contained in CC. Then, ϕ(φ(g))0\phi^{\prime}(\varphi(g))\not=0, in particular trρ1(g)trρ2(g)\operatorname{tr}\rho_{1}(g)\neq\operatorname{tr}\rho_{2}(g). ∎

5. Improved bounds for the effective isogeny theorem

For a prime \ell and an elliptic curve EE, we define

(E[])=k=1(E[k]).\mathbb{Q}(E[\ell^{\infty}])=\bigcup_{k=1}^{\infty}\mathbb{Q}(E[\ell^{k}]).

We begin with the proof of Proposition 5.1. We note that the work which follows is the same as that of Mayle-Wang [9], except for our use of Proposition 2.7 and Table 1.

Let EE and EE^{\prime} be two elliptic curves over \mathbb{Q} and let A=E×EA=E\times E^{\prime}. Let G=Gal((A[2])/))G=\operatorname{Gal}(\mathbb{Q}(A[2^{\infty}])/\mathbb{Q})). We may regard the 22-adic representations ρE,2\rho_{E,2} and ρE,2\rho_{E^{\prime},2} as representations of GG instead of GG_{\mathbb{Q}} since (E[2])\mathbb{Q}(E[2^{\infty}]) and (E[2])\mathbb{Q}(E^{\prime}[2^{\infty}]) are subfields of (A[2])/\mathbb{Q}(A[2^{\infty}])/\mathbb{Q}.

The representation ρA,2=ρE,2×ρE,2\rho_{A,2}=\rho_{E,2}\times\rho_{E^{\prime},2} and is a continuous homomorphism GGL2(2)×GL2(2)G\rightarrow\operatorname{GL}_{2}(\mathbb{Z}_{2})\times\operatorname{GL}_{2}(\mathbb{Z}_{2}) with image being MM as defined in the proof of Proposition 4.1. Since 2M2M is closed inside MM, we see that δ(G)\delta(G) is a closed subgroup of GG and hence by the fundamental theorem of infinite Galois theory corresponds to a finite Galois extension K/K/\mathbb{Q} with K(A[2])/K\subseteq\mathbb{Q}(A[2^{\infty}])/\mathbb{Q}.

Since K(A[2n])K\subseteq\mathbb{Q}(A[2^{n}]) for some nn\in\mathbb{N}, we have that

(5.1) |δ(G)|=[K:][(A[2n]):]|GL2(/2n)|2=(616n1)2,|\delta(G)|=[K:\mathbb{Q}]\mid[\mathbb{Q}(A[2^{n}]):\mathbb{Q}]\mid|\operatorname{GL}_{2}(\mathbb{Z}/2^{n}\mathbb{Z})|^{2}=(6\cdot 16^{n-1})^{2},

for some nn\in\mathbb{N} and by [9, Proposition 12],

(5.2) |δ(G)|255.|\delta(G)|\leq 255.

The set φ(G)\varphi(G) (defined in Proposition 3.6) is a subset of a very explicit semi-direct product, and estimating |φ(G)||\varphi(G)| gives a smaller bound.

Proposition 5.1.

Assume GRH. Let EE and EE^{\prime} be two elliptic curves over \mathbb{Q}. Suppose EE and EE^{\prime} are not \mathbb{Q}-isogenous. Let δ(G)\delta(G) be the deviation group of GG with respect to the 22-adic representations ρE,2\rho_{E,2} and ρE,2\rho_{E^{\prime},2}.

Choose the triple (a¯,b¯,c¯)(\bar{a},\bar{b},\bar{c}) from Table 1 for n0=max(72,2|δ(G)|)n_{0}=\max(72,2|\delta(G)|). Then there exists a prime pp of good reduction for EE and EE^{\prime} such that ap(E)ap(E)a_{p}(E)\neq a_{p}(E^{\prime}) such that

(5.3) p(a¯((2|δ(G)|1)lograd(2NENE)+2|δ(G)|log(2|δ(G)|))+2|δ(G)|b¯+c¯)2.p\leq(\bar{a}((2|\delta(G)|-1)\log\operatorname{rad}(2N_{E}N_{E^{\prime}})+2|\delta(G)|\log(2|\delta(G)|))+2|\delta(G)|\bar{b}+\bar{c})^{2}.

Furthermore, if EE and EE^{\prime} are such that their mod 2 representations are isomorphic and absolutely irreducible, then we may replace |δ(G)||\delta(G)| with |φ(G)||\varphi(G)|.

Proof.

Let 𝒪λ=2\mathcal{O}_{\lambda}=\mathbb{Z}_{2}. Let A=E×EA=E\times E^{\prime} and apply Proposition 4.1 with =2\ell=2, n=2n=2, G=Gal((A[2])/))G=\operatorname{Gal}(\mathbb{Q}(A[2^{\infty}])/\mathbb{Q})), and the 22-adic representations ρ1=ρE,2\rho_{1}=\rho_{E,2} and ρ2=ρE,2\rho_{2}=\rho_{E^{\prime},2}. By Faltings theorem [7], since the two elliptic curves EE and EE^{\prime} are not \mathbb{Q}-isogenous, ρ1\rho_{1} and ρ2\rho_{2} are not isomorphic; therefore, by Serre [13, pg. IV-15], there is some prime pp such that ap(E)ap(E)a_{p}(E)\neq a_{p}(E^{\prime}). By Proposition 4.1, there exists a conjugacy class Cδ(G)C\subseteq\delta(G) obeying the stated conclusion of this proposition.

Let KK be the subfield of (A[2])\mathbb{Q}(A[2^{\infty}]) for which Gal(K/)=δ(G)\operatorname{Gal}(K/\mathbb{Q})=\delta(G). Choosing m=rad(NENE)m=\operatorname{rad}(N_{E}N_{E^{\prime}}), Proposition 2.7 produces a prime pp not dividing mm such that (K/p)C\left(\frac{K/\mathbb{Q}}{p}\right)\subseteq C and

(5.4) p\displaystyle p (a¯((n01)lograd(dK~)+n0logn0)+b¯n0+c¯)2,\displaystyle\leq(\bar{a}\cdot((n_{0}-1)\log\operatorname{rad}(d_{\tilde{K}})+n_{0}\log n_{0})+\bar{b}\cdot n_{0}+\bar{c})^{2},

where n0=max(72,2|δ(G)|)n_{0}=\max(72,2|\delta(G)|) and a¯,b¯,c¯\bar{a},\bar{b},\bar{c} are the maximum of their values over entries in Table 1 with nK~2|δ(G)|n_{\tilde{K}}\leq 2|\delta(G)|, respectively. It follows from Proposition 4.1 that

trρE,2(Frobp)trρE,2(Frobp),\operatorname{tr}\rho_{E,2}(\text{Frob}_{p})\neq\operatorname{tr}\rho_{E^{\prime},2}(\text{Frob}_{p}),

and consequently ap(E)ap(E)a_{p}(E)\neq a_{p}(E^{\prime}).

The abelian variety AA has good reduction at some prime qq if and only if both EE and EE^{\prime} have good reduction at qq. Thus, K/K/\mathbb{Q} is unramified outside of the prime divisors of m=NENEm=N_{E}N_{E^{\prime}}. As K~\tilde{K} is the compositum of KK and (m)\mathbb{Q}(\sqrt{m}), the primes that ramify in K~\tilde{K} are precisely those that ramify in KK or in (m)\mathbb{Q}(\sqrt{m}). Since rad(d(m))rad(2m)=rad(2NENE)\operatorname{rad}(d_{\mathbb{Q}(\sqrt{m})})\mid\operatorname{rad}(2m)=\operatorname{rad}(2N_{E}N_{E^{\prime}}), and rad(dK)(d_{K})\mid rad(2NENE)(2N_{E}N_{E^{\prime}}), we have that

(5.5) rad(dK~)=rad(dKd(m))rad(2NENE).\operatorname{rad}(d_{\tilde{K}})=\text{rad}(d_{K}d_{\mathbb{Q}(\sqrt{m})})\mid\text{rad}(2N_{E}N_{E^{\prime}}).

Now, applying Lemma 2.8 to (5.4) gives us

p\displaystyle p (a¯((2|δ(G)|1)lograd(2NENE)+2|δ(G)|log(2|δ(G)|))+2|δ(G)|b¯+c¯)2\displaystyle\leq(\bar{a}((2|\delta(G)|-1)\log\operatorname{rad}(2N_{E}N_{E^{\prime}})+2|\delta(G)|\log(2|\delta(G)|))+2|\delta(G)|\bar{b}+\bar{c})^{2}

which matches (5.3).

To prove the final statement, note if ρ¯E,2\bar{\rho}_{E,2} and ρ¯E,2\bar{\rho}_{E^{\prime},2} are isomorphic and absolutely irreducible, then we instead apply Proposition 4.4 over Proposition 4.1, in which case δ(G)\delta(G) is replaced with φ(G)\varphi(G); in particular, |δ(G)||\delta(G)| can be replaced with |φ(G)||\varphi(G)| in (5.3). ∎

Now we give a proof of Theorem 1.3.

Proof of Theorem 1.3.

We split our analysis into two cases. If the mod 22 representations are not isomorphic, then mod 22 already distinguishes the traces. Define

(5.6) ρ¯2:GGL2(/2)×GL2(/2)\bar{\rho}_{2}:G_{\mathbb{Q}}\to\operatorname{GL}_{2}(\mathbb{Z}/2\mathbb{Z})\times\operatorname{GL}_{2}(\mathbb{Z}/2\mathbb{Z})

by ρ¯2(x)=(ρ¯E,2(x),ρ¯E,2(x))\bar{\rho}_{2}(x)=(\bar{\rho}_{E,2}(x),\bar{\rho}_{E^{\prime},2}(x)). Let G2=ρ¯2(G)GL2(/2)×GL2(/2)G_{2}=\bar{\rho}_{2}(G_{\mathbb{Q}})\subset\operatorname{GL}_{2}(\mathbb{Z}/2\mathbb{Z})\times\operatorname{GL}_{2}(\mathbb{Z}/2\mathbb{Z}) be the image of the map ρ¯2\bar{\rho}_{2}. Let

(5.7) C2={(s,s)G2tr(s)tr(s)}.C_{2}=\{(s,s^{\prime})\in G_{2}\mid\operatorname{tr}(s)\neq\operatorname{tr}(s^{\prime})\}.

Apply Proposition 2.7 to the field L=(E[2],E[2])L=\mathbb{Q}(E[2],E^{\prime}[2]), whose Galois group is G2G_{2}, the conjugacy class C2C_{2} given in (5.7), and m=rad(NENE)m=\operatorname{rad}(N_{E}N_{E^{\prime}}), so that we get a prime pp unramified in LL and pmp\nmid m such that ap(E)ap(E)(mod 2)a_{p}(E)\not\equiv a_{p}(E^{\prime})\ (\mathrm{mod}\ 2) (implying ap(E)ap(E)a_{p}(E)\neq a_{p}(E^{\prime})) and satisfying

(5.8) p\displaystyle p max((a¯log|dL~|+b¯nL~+c¯)2,p0(nL~))\displaystyle\leq\max((\bar{a}\log|d_{\tilde{L}}|+\bar{b}n_{\tilde{L}}+\bar{c})^{2},p_{0}(n_{\tilde{L}}))
(5.9) (1.755((2nL1)lograd(dL~)+2nLlog(2nL))+0.232nL+6.8).\displaystyle\leq\left(1.755\left((2n_{L}-1)\log\operatorname{rad}(d_{\tilde{L}})+2n_{L}\log(2n_{L})\right)+0.23\cdot 2n_{L}+6.8\right).

Taking [L:]|GL2(𝔽2)|2=62=36[L:\mathbb{Q}]\leq|\operatorname{GL}_{2}(\mathbb{F}_{2})|^{2}=6^{2}=36 gives us

(5.10) p(1.745((71)lograd(dL~)+72log(72))+720.23+6.8)2(124lograd(2NENE)+561)2\begin{split}p&\leq\left(1.745\left((71)\log\operatorname{rad}(d_{\tilde{L}})+72\log(72)\right)+72\cdot 0.23+6.8\right)^{2}\\ &\leq\left(124\log\operatorname{rad}(2N_{E}N_{E^{\prime}})+561\right)^{2}\\ \end{split}

Next, assume that the mod 2 representations ρ1=ρ¯E,2\rho_{1}=\bar{\rho}_{E,2} and ρ2=ρ¯E,2\rho_{2}=\bar{\rho}_{E^{\prime},2} are isomorphic and absolutely irreducible. Apply Proposition 5.1, and replace δ(G)\delta(G) with φ(G)\varphi(G) (since the mod 2 representations are isomorphic and absolutely irreducible) to get a prime pp such that ap(E)ap(E)a_{p}(E)\neq a_{p}(E^{\prime}) and satisfying

p(a¯((2|φ(G)|1)lograd(2NENE)+2|φ(G)|log(2|φ(G)|))+2|φ(G)|b¯+c¯)2.p\leq(\bar{a}((2|\varphi(G)|-1)\log\operatorname{rad}(2N_{E}N_{E^{\prime}})+2|\varphi(G)|\log(2|\varphi(G)|))+2|\varphi(G)|\bar{b}+\bar{c})^{2}.

From (3.2) and Lemma 3.8, we have for any gGg\in G,

(5.11) det(ρ2(g))=det((I2+2βθ(g))ρ1(g))=(1+2βtr(θ(g))+O(22β))det(ρ1(g)).\begin{split}\det(\rho_{2}(g))&=\det((I_{2}+2^{\beta}\theta(g))\rho_{1}(g))\\ &=(1+2^{\beta}\operatorname{tr}(\theta(g))+O(2^{2\beta}))\det(\rho_{1}(g)).\end{split}

As we have detρ1=detρ2\det\rho_{1}=\det\rho_{2} being the cyclotomic character, so the above can be rewritten as 0=2βtr(θ(g))+O(22β)0=2^{\beta}\operatorname{tr}(\theta(g))+O(2^{2\beta}), which, after multiplying through by 2β2^{-\beta} implies

tr(θ)0(mod 2).\operatorname{tr}(\theta)\equiv 0\ (\mathrm{mod}\ 2).

In particular, the map φ\varphi from Proposition 3.6 takes values in M20(𝔽2)GL2(𝔽2)M_{2}^{0}(\mathbb{F}_{2})\rtimes\operatorname{GL}_{2}(\mathbb{F}_{2}), where M20(𝔽2)M_{2}^{0}(\mathbb{F}_{2}) denotes the matrices with trace 0 with entries in 𝔽2\mathbb{F}_{2}. Therefore, we have |φ(G)||M20(𝔽2)GL2(𝔽2)|=86=48|\varphi(G)|\leq|M_{2}^{0}(\mathbb{F}_{2})\rtimes\operatorname{GL}_{2}(\mathbb{F}_{2})|=8\cdot 6=48. We find from Proposition 2.7 that

(5.12) p(1.745(95lograd(2NENE)+96log(96))+960.23+6.8)2(166lograd(2NENE)+794)2\begin{split}p&\leq(1.745(95\log\operatorname{rad}(2N_{E}N_{E^{\prime}})+96\log(96))+96\cdot 0.23+6.8)^{2}\\ &\leq(166\log\operatorname{rad}(2N_{E}N_{E^{\prime}})+794)^{2}\\ \end{split}

We now improve the bound in (5.12) by using the following two results.

Proposition 5.2.

Let GG be a group and ρ1,ρ2:GGL2(2)\rho_{1},\rho_{2}:G\rightarrow\operatorname{GL}_{2}(\mathbb{Z}_{2}) be a group homomorphism, and suppose the mod 2 representations ρ¯1,ρ¯2\bar{\rho}_{1},\bar{\rho}_{2} are isomorphic. Let Ξ\Xi be the elements gGg\in G such that the characteristic polynomials of ρ1(g)\rho_{1}(g) and ρ2(g)\rho_{2}(g) coincide. Then for gΞg\in\Xi, the order of δ(g)\delta(g) in δ(G)\delta(G) is 3\leq 3.

Proof.

See [5, Proposition 5.5.6]. ∎

Corollary 5.3.

Let GG be a group and ρ1,ρ2:GGL2(2)\rho_{1},\rho_{2}:G\rightarrow\operatorname{GL}_{2}(\mathbb{Z}_{2}) be a group homomorphism, and suppose the mod 2 representations ρ¯1,ρ¯2\bar{\rho}_{1},\bar{\rho}_{2} are isomorphic. Let π:δ(G)G¯\pi:\delta(G)\twoheadrightarrow\bar{G} be a quotient having a conjugacy class CG¯C\subseteq\bar{G} of order >3>3. If gGg\in G is such that π(δ(g))C\pi(\delta(g))\in C, then gΞg\not\in\Xi.

From (5.1) and (5.2), the possible sizes of δ(G)\delta(G) are

(5.13) 1,2,3,4,6,8,9,12,16,18,24,32,36,48,64,72,96,128,144,192.1,2,3,4,6,8,9,12,16,18,24,32,36,48,64,72,96,128,144,192.

As the possible orders are small, it is possible to enumerate in Magma all isomorphism classes of groups of these sizes. The group δ(G)\delta(G) therefore lies in an explicit finite list which can be computed.

By Corollary 5.3, if δ(G)\delta(G) has a quotient G¯\bar{G} with an element of order >3>3, then we may replace δ(G)\delta(G) by G¯\bar{G}. We check this in Magma for each value of |δ(G)||\delta(G)| and find that either δ(G)\delta(G) has such a quotient with strictly smaller size in the given list of (5.13) or it is in a small list of problematic groups. We list for each size |δ(G)||\delta(G)|, the Magma labels of the isomorphism classes of the problematic groups of that order.

|δ(G)||\delta(G)| # of isomorphism classes problematic groups
192 1543 1023, 1025, 1541
144 197 none
128 2328 2326, 2327, 2328
96 231 204
72 50 none
64 267 266, 267
48 52 3, 50
36 14 11
32 51 49, 50, 51
Table 2. List of problematic groups

Consider the homomorphism φ:δ(G)M20(𝔽2)GL2(𝔽2)\varphi:\delta(G)\rightarrow M_{2}^{0}(\mathbb{F}_{2})\rtimes\operatorname{GL}_{2}(\mathbb{F}_{2}). The possible orders of the image are:

(5.14) 1,2,3,4,6,8,12,16,24,48.1,2,3,4,6,8,12,16,24,48.

Using Magma, we check if |δ(G)|32|\delta(G)|\geq 32, there is no homomorphism from a problematic group to a subgroup of order 2424 or 4848 in the codomain of φ\varphi. Hence, either |δ(G)|24|\delta(G)|\leq 24 or the image of φ\varphi is 16\leq 16. In either case, we can replace δ(G)\delta(G) by a quotient of order 24\leq 24. Hence, (5.12) is improved to the same bound in (5.10).

6. The results of Rouse and Zureick-Brown on 22-adic images

The 22-adic representation ρE,2:GGL2(2)\rho_{E,2}:G_{\mathbb{Q}}\rightarrow\operatorname{GL}_{2}(\mathbb{Z}_{2}) of an elliptic curve EE over \mathbb{Q} has open image in GL2(2)\operatorname{GL}_{2}(\mathbb{Z}_{2}) and the properties that detρE,2:G2×\det\rho_{E,2}:G_{\mathbb{Q}}\rightarrow\mathbb{Z}_{2}^{\times} is surjective and ρE,2(c)\rho_{E,2}(c) is an element with determinant 1-1 and trace 0 for a complex conjugation. With this in mind, the authors in [12] make the following definition:

Definition 6.1.

An open subgroup HGL2(2)H\subseteq\operatorname{GL}_{2}(\mathbb{Z}_{2}) is arithmetically maximal if

  1. (1)

    det:H2×\det:H\rightarrow\mathbb{Z}_{2}^{\times} is surjective,

  2. (2)

    there is an element of HH with determinant 1-1 and trace 0, and

  3. (3)

    the is no subgroup KK satisfying (1) and (2) with HKH\subsetneq K and so that the genus of XKX_{K} is 2\geq 2.

The idea behind this definition is that arithmetically maximal subgroups HGL2(2)H\subseteq\operatorname{GL}_{2}(\mathbb{Z}_{2}) are maximal among the subgroups HH satisfying (1) and (2), except possibly when HH is contained in a subgroup KK such that XKX_{K} has genus 1\leq 1. For instance, if HKH\subsetneq K and XKX_{K} has genus 2\geq 2, it would be easier and sufficient to determine the \mathbb{Q}-rationals point XKX_{K} rather than XHX_{H}.

For an arithmetically maximal subgroup HH, either XHX_{H} has infinitely many \mathbb{Q}-rational points (hence has genus 1\leq 1) or XHX_{H} has finitely many \mathbb{Q}-rational points. The union of the latter cases leads to a finite list of jj-invariants.

In [12], it is shown there are 727727 arithmetically maximal subgroups HGL2(2)H\subseteq\operatorname{GL}_{2}(\mathbb{Z}_{2}) up to conjugation such that IH-I\in H. There are an additional 10061006 arithmetically maximal subgroups up to conjugation such that IH-I\notin H. Thus, there are a total of 17331733 arithmetically maximal subgroups HGL2(2)H\subseteq\operatorname{GL}_{2}(\mathbb{Z}_{2}). Among these, there are 14141414 which have genus 1\leq 1 and 12081208 of these are such that XHX_{H} has infinitely many \mathbb{Q}-rational points.

In [12, Theorem 1.1], the possible images of ρE,2\rho_{E,2} are determined in the following sense:

Theorem 6.2.

Let HGL2(2)H\subseteq\operatorname{GL}_{2}(\mathbb{Z}_{2}) be a subgroup and let E/E/\mathbb{Q} be an elliptic curve such that the image of ρE,2\rho_{E,2} is contained in a conjugate of HH. Then one of the following holds:

  1. (1)

    The modular curve XHX_{H} has infinitely many \mathbb{Q}-rational points (hence has genus 1\leq 1).

  2. (2)

    The elliptic curve EE has complex multiplication.

  3. (3)

    The jj-invariant of EE is one of

    (6.1) 211,24173,40973/24,2573/28,8579853/628,9194253/4964,\displaystyle 2^{11},2^{4}\cdot 17^{3},4097^{3}/2^{4},257^{3}/2^{8},-857985^{3}/62^{8},919425^{3}/496^{4},
    3182499203/1716,717231878060803/7916.\displaystyle-3\cdot 18249920^{3}/17^{16},7\cdot 1723187806080^{3}/79^{16}.

We rephrase the above theorem for the purposes of proving the main results of this paper.

Theorem 6.3.

Let E/E/\mathbb{Q} be an elliptic curve without complex multiplication. Then ρE,2(G)\rho_{E,2}(G) is one of the 12081208 arithmetically maximal groups listed in [12] which have infinitely many \mathbb{Q}-rational points, or the jj-invariant of EE appears in (6.1).

7. Improved bounds for Serre’s open image theorem

In Theorem 1.3, there is a hypothesis that ρ¯E,2ρ¯E,2{\bar{\rho}}_{E,2}\simeq{\bar{\rho}}_{E^{\prime},2} is absolutely irreducible. While we do not have an argument to remove this condition and achieve better bounds than Theorem 1.2, we are able to do so in the case when EE and EE^{\prime} are quadratic twists of each other and do not have complex multiplication.

Proof of Theorem 1.5.

Suppose EE^{\prime} is a twist of EE by a quadratic character χ\chi associated to the extension (d)\mathbb{Q}(\sqrt{d}). Then ρE,2=ρE,2χ\rho_{E^{\prime},2}=\rho_{E,2}\otimes\chi. Since EE does not have complex multiplication, EE and EE^{\prime} are not \mathbb{Q}-isogenous.

Assume ρE,2\rho_{E,2} and hence ρE,2\rho_{E^{\prime},2} are not absolutely irreducible.

If (d)\mathbb{Q}(\sqrt{d}) is not a subfield of L=(E[2])L=\mathbb{Q}(E[2^{\infty}]), then L(d)L(\sqrt{d}) is a degree 22 extension of LL. Hence, the identity automorphism of LL extends to an automorphism σ\sigma of L(d)L(\sqrt{d}) such that χ(σ)=1\chi(\sigma)=-1. It follows that ρE,2(σ)=I\rho_{E,2}(\sigma)=I and ρE,2(σ)=I\rho_{E^{\prime},2}(\sigma)=-I where II is the identity element. This means that α=β=1\alpha=\beta=1 so we may apply the proofs of Theorem 1.3 and Proposition 4.4 (no need for Theorem 4.3) to get the desired conclusion.

Otherwise (d)\mathbb{Q}(\sqrt{d}) lies in the field LL. In [12] (see Theorem 6.3), the possible 22-adic images ρE,2(G)\rho_{E,2}(G) of an elliptic curve E/E/\mathbb{Q} are determined up to conjugacy. Every such image contains the principal congruence subgroup of level 3232 and can be regarded as a subgroup of GL2(/2t)\operatorname{GL}_{2}(\mathbb{Z}/2^{t}\mathbb{Z}) for 0t50\leq t\leq 5. In order to apply [12, Lemma 3.3], without loss of generality we take t2t\geq 2.

Let Γ=1+2t+1M2(2)N=1+2tM2(2)\Gamma=1+2^{t+1}M_{2}(\mathbb{Z}_{2})\subseteq N=1+2^{t}M_{2}(\mathbb{Z}_{2}). The character χ\chi corresponds to a subgroup HH of index 22 inside ρE,2(G)\rho_{E,2}(G). Either ΓNH\Gamma\subseteq N\subseteq H or

(7.1) N/NHNH/HG/HN/N\cap H\cong NH/H\cong G/H

so N/NHN/N\cap H has order 22. In the latter case, NHN\cap H is a maximal subgroup of NN, hence by [12, Lemma 3.3], we obtain again that ΓNHH\Gamma\subseteq N\cap H\subseteq H.

It follows that χ\chi factors through

(7.2) ρE,2(G)/Γ.\rho_{E,2}(G)/\Gamma.

Consider the product representation

(7.3) ρ2=ρE,2×ρE,2:GGL2(2)×GL2(2).\rho_{2}=\rho_{E,2}\times\rho_{E^{\prime},2}:G\rightarrow\operatorname{GL}_{2}(\mathbb{Z}_{2})\times\operatorname{GL}_{2}(\mathbb{Z}_{2}).

We note that

(7.4) ρE,2(g)={ρE,2(g) if ρE,2(g)HρE,2(g) if ρE,2(g)H.\rho_{E^{\prime},2}(g)=\begin{cases}\rho_{E,2}(g)&\text{ if }\rho_{E,2}(g)\in H\\ -\rho_{E,2}(g)&\text{ if }\rho_{E,2}(g)\notin H.\end{cases}

Let θ\theta be defined as:

(7.5) θ:ρE,2(G)\displaystyle\theta:\rho_{E,2}(G) ρE,2(G)\displaystyle\rightarrow\rho_{E^{\prime},2}(G)
ρE,2(g)\displaystyle\rho_{E,2}(g) {ρE,2(g) if ρE,2(g)HρE,2(g) if ρE,2(g)H.\displaystyle\mapsto\begin{cases}\rho_{E,2}(g)&\text{ if }\rho_{E,2}(g)\in H\\ -\rho_{E,2}(g)&\text{ if }\rho_{E,2}(g)\notin H.\end{cases}

The map θ\theta is an isomorphism which is the identity when restricted to ΓH\Gamma\subseteq H and

(7.6) ρ2(G)=(ρE,2×ρE,2)(G)\rho_{2}(G)=(\rho_{E,2}\times\rho_{E^{\prime},2})(G)

is the subgroup of GL2(2)×GL2(2)\operatorname{GL}_{2}(\mathbb{Z}_{2})\times\operatorname{GL}_{2}(\mathbb{Z}_{2}) given by the graph of θ\theta, namely,

(7.7) {(ρE,2(g),θ(ρE,2(g))):gG}.\left\{(\rho_{E,2}(g),\theta(\rho_{E,2}(g))):g\in G\right\}.

Hence, it follows that the inclusion

(7.8) ι:ρE,2(G)\displaystyle\iota:\rho_{E,2}(G) ρ(G)\displaystyle\rightarrow\rho(G)
(7.9) ρE,2(g)\displaystyle\rho_{E,2}(g) (ρE,2(g),θ(ρE,2(g)))\displaystyle\mapsto(\rho_{E,2}(g),\theta(\rho_{E,2}(g)))

is an isomorphism. Composing with the homomorphism ρ2(G)(M/2M)×\rho_{2}(G)\rightarrow(M/2M)^{\times}, we obtain a homomorphism

(7.10) ϕ:ρE,2(G)ρ2(G)δ(G)(M/2M)×.\phi:\rho_{E,2}(G)\rightarrow\rho_{2}(G)\twoheadrightarrow\delta(G)\subseteq(M/2M)^{\times}.

Since ρE,2(G)1+2tM2(2)\rho_{E,2}(G)\supseteq 1+2^{t}M_{2}(\mathbb{Z}_{2}), we see that M=2[ρ2(G)]2tΔ(M2(2))M=\mathbb{Z}_{2}[\rho_{2}(G)]\supseteq 2^{t}\Delta(M_{2}(\mathbb{Z}_{2})) where Δ:M2(2)M2(2)×M2(2)\Delta:M_{2}(\mathbb{Z}_{2})\rightarrow M_{2}(\mathbb{Z}_{2})\times M_{2}(\mathbb{Z}_{2}) is the diagonal homomorphism. Thus, 2M2t+1Δ(M2(2))2M\supseteq 2^{t+1}\Delta(M_{2}(\mathbb{Z}_{2})). Thus, we see that the kernel of the map ϕ\phi contains Γ=1+2t+1M2(2)\Gamma=1+2^{t+1}M_{2}(\mathbb{Z}_{2}) and hence ϕ\phi factors through

(7.11) ρE,2(G)/Γ.\rho_{E,2}(G)/\Gamma.

Using Magma, we check that each one of the 12081208 possibilities for ρE,2(G)\rho_{E,2}(G) does not have a quotient which factors through ρE,2(G)/Γ\rho_{E,2}(G)/\Gamma and is isomorphic to one of the problematic groups of order 192192, 128128, or 9696 in Table 2, if the image of ρ¯E,2{\bar{\rho}}_{E,2} is not absolutely irreducible. We conclude that it is possible to replace δ(G)\delta(G) with a quotient of order 64\leq 64. It is also checked that the elliptic curves EE over \mathbb{Q} with jj-invariant in the finite list of Theorem 6.2 (3) have CE2C_{E}\leq 2.

Hence, we obtain the bound

(7.12) p(1.755(127lograd(2NENE)+128log(128))+1280.23+6.8)2(223lograd(2NENE)+1127)2.\begin{split}p&\leq(1.755(127\log\operatorname{rad}(2N_{E}N_{E^{\prime}})+128\log(128))+128\cdot 0.23+6.8)^{2}\\ &\leq(223\log\operatorname{rad}(2N_{E}N_{E^{\prime}})+1127)^{2}.\end{split}

Proof of Theorem 1.7.

We combine Theorems 1.3 and 1.5 with the arguments in [9]. ∎

References

  • [1] Bach, E., and Sorenson, J. Explicit bounds for primes in residue classes. Mathematics of Computation 65, 216 (1996), 1717–1735.
  • [2] Bosma, W., Cannon, J., and Playoust, C. The Magma algebra system. I. The user language. J. Symbolic Comput. 24, 3-4 (1997), 235–265. Computational algebra and number theory (London, 1993).
  • [3] Bucur, A., and Kedlaya, K. S. An application of the effective sato-tate conjecture. Contemporary Mathematics 663 (2016), 45–56.
  • [4] Carayol, H. Formes modulaires et représentations galoisiennes à valeurs dans un anneau local complet. Contemprary Mathematics 165 (1994), 213–237.
  • [5] Chênevert, G. Exponential sums, hypersurfaces with many symmetries and Galois representations. PhD thesis, McGill University, 2008.
  • [6] Cornell, G., and Silverman, J. H. Arithmetic Geometry. Springer-Verlag, 1986.
  • [7] Faltings, G. Endlichkeitssätze für abelsche varietäten über zahlkörpern. Inventiones Mathematicae 73 (1983), 349–366.
  • [8] Lagarias, J. C., and Odlyzko, A. M. Effective versions of the Chebotarev density theorem. Algebraic Number Fields 54 (1979), 271–296.
  • [9] Mayle, J., and Wang, T. On the effective version of Serre’s open image theorem. Bulletin of the London Mathematical Society (2024), to appear.
  • [10] Oesterlé, J. Versions effectives du théorème de Chebotarev sous l’hypothèse de Riemann généralisée. Astérisque 61 (1979), 165–167.
  • [11] Rodríguez, I. S. Comparing galois representations and the Falting-Serre-Livné method. Master’s thesis, Universitat de Barcelona, 2020.
  • [12] Rouse, J., and Zureick-Brown, D. Elliptic curves over \mathbb{Q} and 2-adic images of Galois. Res. Number Theory 1 (2015), Paper No. 12, 34.
  • [13] Serre, J.-P. Abelian \ell-Adic Representations and Elliptic Curves. W. A. Benjamin, Inc., 1968.
  • [14] Serre, J.-P. Propriétés galoisiennes des points d’ordre fini des courbes elliptiques. Invent. Math. 15, 4 (1972), 259–331.
  • [15] Serre, J.-P. Quelques applications du théorème de densité de Chebotarev. Publications Mathématiques de l’IHÉS 54 (1981), 123–201.
  • [16] Serre, J.-P. Résumé des cours de 1984-1985. In Annuaire du Collège de France. 1985, pp. 85–90.