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

Factorization of measures and applications to the weak Goldfeld conjecture

Merrick Cai
(Date: August 2021)
Abstract.

Extending Gross’s result, we prove that a certain factorizaton of measures holds for all pp and any finite even Dirichlet character χ\chi of any conductor, rather than only for split pp and χ\chi with conductor a power of pp. Using this generalization, we find lower bounds on the proportion of imaginary quadratic fields KK for which (under certain assumptions on the elliptic curve) a chosen quadratic twist of an elliptic curve EE over KK has rank 11. We also find lower and upper bounds for the proportion of quadratic twists with rank 11 when we vary DD, the factor we twist by, under the assumption that ω\omega (the prime factor counting function) is sufficiently close to a Gaussian distribution, as described by Erdös-Kac. We apply similar methods to cubic twists, and then derive analogous lower bounds for the proportion of imaginary quadratic fields for which a sextic twist has rank 11. Lastly, for elliptic curves over \mathbb{Q} satisfying certain assumptions, we find positive lower bounds on the proportion of quadratic twists (over \mathbb{Q}) which have rank 0 and rank 11, which yields examples of elliptic curves satisfying the weak Goldfeld conjecture.

1. Introduction

1.1. Algebraic and analytic rank

Let EE be an elliptic curve over \mathbb{Q}. The \mathbb{Q}-points of EE form an abelian group E(E(\mathbb{Q}) called the Mordell-Weil group. Mordell’s theorem states that E()E(\mathbb{Q}) is finitely generated, and thus the rank of E()E(\mathbb{Q}) is a well-defined, nonnegative integer. We call the rank of E()E(\mathbb{Q}) the algebraic rank of EE, and denote it as ralg(E)r_{alg}(E).

However, the algebraic rank is rather difficult to handle. Instead, we may attach the following LL-function to the elliptic curve E/E/\mathbb{Q}:

L(E/,s)=pLp(E/,s),L(E/\mathbb{Q},s)=\prod_{p}L_{p}(E/\mathbb{Q},s),

where

Lp(E/,s)={(1apps+p12s)1p has good reduction,(1±apps)1p has multiplicative reduction,1p has additive reduction,L_{p}(E/\mathbb{Q},s)=\begin{cases}\left(1-a_{p}\cdot p^{-s}+p^{1-2s}\right)^{-1}&p\text{ has good reduction},\\ \left(1\pm a_{p}\cdot p^{-s}\right)^{-1}&p\text{ has multiplicative reduction},\\ 1&p\text{ has additive reduction},\end{cases}

and ap=p+1|E(𝔽p)|a_{p}=p+1-|E(\mathbb{F}_{p})| is the trace of the Frobenius element associated to pp. (In the multiplicative reduction case, the type of reduction determines the sign of the plus/minus.) This LL-function satisfies a functional equation relating its values at ss and 2s2-s, and thus its order of vanishing at s=1s=1 is of interest. We call the order of vanishing of L(E/,s)L(E/\mathbb{Q},s) at s=1s=1 the analytic rank of EE, and denote it as ran(E)r_{an}(E).

Although the notions of analytic and algebraic rank may seem unrelated, they are not. The famous Birch and Swinnerton-Dyer conjecture [6] posits that they are in fact equal.

Conjecture 1.1 (Birch and Swinnerton-Dyer).

The algebraic rank is the same as the analytic rank: ralg(E)=ran(E)r_{alg}(E)=r_{an}(E).

The BSD conjecture is still wide open, although significant advances have been made. Some of the strongest known results are due to [25], [27], [2], [13], [15], and [16], and they relate the algebraic and analytic ranks in low rank cases.

Theorem 1.2.

If ran(E){0,1}r_{an}(E)\in\{0,1\}, then ran(E)=ralg(E)r_{an}(E)=r_{alg}(E).

However, it’s still unproven as to whether ralg(E){0,1}r_{alg}(E)\in\{0,1\} implies that ralg(E)=ran(E)r_{alg}(E)=r_{an}(E).

1.2. Goldfeld’s conjecture

Elliptic curves can be ordered by a property called height. This property is useful when studying statistics of elliptic curves, since it allows us to formalize the notion of an average: to measure the average of a quantity over all elliptic curves, we can calculate the average over the finitely many elliptic curves with height at most XX, and then take a limit as XX\to\infty. The analytic rank of an elliptic curve is one particularly important property that can be studied in this way. Originating from [11] and [18], it is widely believed that among all elliptic curves over \mathbb{Q}, the elliptic curves with analytic rank 0 or 11 should each have density 50%50\%, while elliptic curves with analytic rank greater than 11 should have density 0. Recent developments by [5], [7], [4], and others have placed increasingly tighter bounds on the average, putting it closer and closer to the conjectured value of 0.50.5; for example, the average rank is bounded below by 0.20680.2068 and bounded above by 0.8850.885.

Understanding the average rank over all elliptic curves is rather difficult. We can instead look at one particular family of elliptic curves: the quadratic twists EDE_{D} of a fixed elliptic curve EE. In [11], Goldfeld postulated that the average rank of a family of quadratic twists should behave in the same way as the set of elliptic curves over \mathbb{Q}.

Conjecture 1.3 (Goldfeld).

Let EE be an elliptic curve and let {ED}\{E_{D}\} be the family of quadratic twists of EE as DD varies over the set of fundamental domains. Then 50%50\% of the EDE_{D} have analytic rank 0, 50%50\% have analytic rank 11, and 0%0\% have analytic rank greater than 11.

However, Goldfeld’s conjecture is still very open. There is no elliptic curve which has been shown to satisfy Goldfeld’s conjecture. We will instead study the following weaker version of Goldfeld’s conjecture (see, for example, [14, Conjecture 1.2]).

Conjecture 1.4 (Weak Goldfeld).

As in Conjecture 1.3, fix EE and let {ED}\{E_{D}\} be the family of quadratic twists of EE. A positive proportion of the EDE_{D} have rank 0 and a positive proportion of the EDE_{D} have rank 11.

In the last section of this paper, we will prove a result which, given certain conditions on the elliptic curve, guarantee that a positive proportion of its quadratic twists will have rank 0 and 11; in addition, we give lower bounds for these proportions.

1.3. Measures on profinite groups

We follow the exposition in [12, §1]. Let pp\in\mathbb{Z} be a prime, p\mathbb{Z}_{p} the ring of pp-adic integers, p\mathbb{Q}_{p} the field of pp-adic numbers, and p\mathbb{C}_{p} the algebraic closure of p\mathbb{Q}_{p}. Now let 𝔻p\mathbb{D}_{p} be the ring of integral elements in p\mathbb{C}_{p}. For a commutative profinite group GG, we consider its completed group algebra over 𝔻p\mathbb{D}_{p}, ΛG𝔻p[[G]]=limHG open𝔻p[G/H]\Lambda_{G}\coloneqq\mathbb{D}_{p}[[G]]=\varprojlim_{H\subset G\text{ open}}\mathbb{D}_{p}[G/H]. The elements of ΛG\Lambda_{G} are called measures on GG. We also define ΛG\Lambda_{G}^{\prime}, the total ring of fractions of ΛG\Lambda_{G}, as the ring whose elements are α/β\alpha/\beta for α,βΛG\alpha,\beta\in\Lambda_{G} and β\beta is not a zero-divisor.

We define a bilinear pairing between continuous functions GpG\to\mathbb{C}_{p} and measures in ΛG\Lambda_{G} by approximating ff by locally constant functions and taking a limit, as in [22]:

f,λ=Gf𝑑λ.\langle f,\lambda\rangle=\int_{G}f\,d\lambda.

For λ=α/βΛG\lambda=\alpha/\beta\in\Lambda_{G}^{\prime}, we extend this pairing by f,λf,α/f,β\langle f,\lambda\rangle\coloneqq\langle f,\alpha\rangle/\langle f,\beta\rangle. This construction is well-defined since it does not depend on the representation of λ\lambda, and agrees with our previous definition for λΛG\lambda\in\Lambda_{G}.

Let KK be an imaginary quadratic field. We will primarily consider the case where G=Gal(K(μp)/)G=\text{Gal}(K(\mu_{p^{\infty}})/\mathbb{Q}) or G=Gal(K(μp)/)/σG=\text{Gal}(K(\mu_{p^{\infty}})/\mathbb{Q})/\sigma where σ\sigma is complex conjugation, and f=χf=\chi is a (continuous) character from GG to 𝔻p×\mathbb{D}_{p}^{\times}.

1.4. Structure of the paper and main results

Let pp be a prime, K=(C)K=\mathbb{Q}(\sqrt{-C}) an imaginary quadratic field where pp splits, and χ\chi a continuous pp-adic character of Gal(K(μp)/)\text{Gal}(K(\mu_{p^{\infty}})/\mathbb{Q}) which is trivial on complex conjugation. Let χK\chi_{K} be the restriction of χ\chi to Gal(K(μp)/K)\text{Gal}(K(\mu_{p^{\infty}})/K), ϵ\epsilon the quadratic character modulo CC, and ω\omega the Teichmüller character. As in [12], define the measures λ1,λ2,λ3\lambda_{1},\lambda_{2},\lambda_{3} by χ,λ1=Lp(0,χK)\langle\chi,\lambda_{1}\rangle=L_{p}(0,\chi_{K}), χ,λ2=Lp(0,χϵω)\langle\chi,\lambda_{2}\rangle=L_{p}(0,\chi\epsilon\omega), and χ,λ3=Lp(1,χ1)\langle\chi,\lambda_{3}\rangle=L_{p}(1,\chi^{-1}). Motivated by the classical factorization of LL-series L(s,(χK))=L(s,χϵ)L(s,χ)L(s,(\chi_{K})_{\infty})=L(s,\chi_{\infty}\epsilon)L(s,\chi_{\infty}), Gross [12, Theorem 3.1] derives the factorization of measures

λ1=λ2λ3,\lambda_{1}=\lambda_{2}\cdot\lambda_{3},

when pp is split in KK and χ\chi is a finite even Dirichlet character whose conductor is a power of pp. In §2, we extend this result in Theorem 2.13 to all pp (not just split pp) and any finite even Dirichlet character χ\chi (with any conductor).

We then turn our attention to elliptic curves. In §3, we introduce assumptions on the elliptic curve which will hold for the remainder of the paper. We will assume that EE is residually reducible modulo 33 (Assumption 3.1), and we will work with integers DD and imaginary quadratic fields KK satisfying various divisibility and congruence conditions relating DD, the conductor of EE, and the discriminant of KK (Assumptions 3.2 and 3.3).

In §4, we discuss general congruences of LL-series and Eisenstein series, especially those associated with quadratic characters. We obtain some auxiliary results concerning the congruence of certain modular forms with Eisenstein series, and calculate the Euler factor at pp after pp-depleting (see §4.3).

In §5.1, we use the factorization in Theorem 2.13 to arrive at the two key technical results, Theorem 5.6 and Theorem 5.7. Under Assumption 3.3 and the assumption that DD satisfies the nonvanishing of a certain class number modulo 33 (for D>0D>0, we need 3h(3D)3\nmid h_{\mathbb{Q}(\sqrt{-3D})} and for D<0D<0, we need 3h(D)3\nmid h_{\mathbb{Q}(\sqrt{D})}), we find a lower bound on the proportion of imaginary quadratic fields KK for which ED/KE_{D}/K has rank 11. In §5.2, we vary DD instead. Assuming that ω(n)\omega(n) is sufficiently close to a Gaussian distribution, we find bounds on the proportion of DD such that ED/KE_{D}/K has rank 11; these are given in Theorem 5.11.

In §6, we address cubic and sextic twists. In §6.1, we obtain results similar to §4 but for cubic twists. Since sextic twists are a composition of a cubic twist and a quadratic twist, we apply our results from §5.1 to obtain similar results on sextic twists in §6.2. The results, paralleling Theorem 5.6 and Theorem 5.7, are given by Theorem 6.6 and Theorem 6.7.

Finally, in §7, we positive lower bounds on the proportion of DD for which ED/E_{D}/\mathbb{Q} has rank 0 and 11, under similar assumptions. Given the assumptions before, plus the additional assumption that 3Ncond(E)3\nmid N\coloneqq\text{cond}(E), in Theorem 7.1 we find that ED/E_{D}/\mathbb{Q} has rank 0 for at least ϕ(N)4N\frac{\phi(N)}{4N} of all such DD, and rank 11 for at least ϕ(N)4N\frac{\phi(N)}{4N} of all such DD. As an easy corollary, we conclude Conjecture 1.4 for certain elliptic curves.

Acknowledgements

The author would like to thank Daniel Kriz for supervising this project, mentoring the author, and providing much needed guidance. The author also thanks Jonathan Love and Professor Andrew Sutherland for many helpful discussions and feedback.

2. Factorization of measures

We follow the notation in [12]. Let pp be a prime, K=(C)K=\mathbb{Q}(\sqrt{-C}) an imaginary quadratic field where pp splits, χ\chi a finite even Dirichlet character on Gal(K(μp)/)\text{Gal}(K(\mu_{p^{\infty}})/\mathbb{Q}), χK\chi_{K} the restriction of χ\chi to Gal(K(μp)/K)\text{Gal}(K(\mu_{p^{\infty}})/K), and χ\chi_{\infty} the composition of χ\chi with some fixed injection ¯\overline{\mathbb{Q}}\hookrightarrow\mathbb{C}. Let ϵ\epsilon be the quadratic character modulo CC and ω\omega the Teichmüller character. We define λ1,λ2,λ3\lambda_{1},\lambda_{2},\lambda_{3} as in [12, p. 92], and obtain the formulas χ,λ1=Lp(0,χK)\langle\chi,\lambda_{1}\rangle=L_{p}(0,\chi_{K}), χ,λ2=Lp(0,χϵω)\langle\chi,\lambda_{2}\rangle=L_{p}(0,\chi\epsilon\omega), and χ,λ3=Lp(1,χ1)\langle\chi,\lambda_{3}\rangle=L_{p}(1,\chi^{-1}), as in [12, p. 93].

2.1. Dirichlet characters with conductors a prime power

We start with the classical factorization L(s,(χK))=L(s,χϵ)L(s,χ)L(s,(\chi_{K})_{\infty})=L(s,\chi_{\infty}\epsilon)L(s,\chi_{\infty}) and the functional equation for L(s,χ)L(s,\chi):

L(s,χ)=L(1s,χ¯)Γ(1s+a2)Γ(s+a2)(kπ)12sτ(χ)iak,L(s,\chi)=L(1-s,\overline{\chi})\frac{\Gamma\left(\frac{1-s+a}{2}\right)}{\Gamma\left(\frac{s+a}{2}\right)}\left(\frac{k}{\pi}\right)^{\frac{1}{2}-s}\frac{\tau(\chi)}{i^{a}\sqrt{k}},

where Γ\Gamma is the gamma function, kk is the conductor of χ\chi, τ=n=1kχ(n)e2πin/k\tau=\sum_{n=1}^{k}\chi(n)e^{2\pi in/k} is the Gauss sum, and a=0a=0 if χ(1)=1\chi(-1)=1 while a=1a=1 if χ(1)=1\chi(-1)=-1.

Proposition 2.1.

L(0,(χK))=L(0,χϵ)L(0,χ)L^{\prime}(0,(\chi_{K})_{\infty})=L(0,\chi_{\infty}\epsilon)L^{\prime}(0,\chi_{\infty}).

Proof.

By differentiating,

L(s,(χK))=L(s,χϵ)L(s,χ)+L(s,χϵ)L(s,χ).L^{\prime}(s,(\chi_{K})_{\infty})=L^{\prime}(s,\chi_{\infty}\epsilon)L(s,\chi_{\infty})+L(s,\chi_{\infty}\epsilon)L^{\prime}(s,\chi_{\infty}).

Setting s=0s=0 yields

L(0,(χK))=L(0,χϵ)L(0,χ)+L(0,χϵ)L(0,χ).L^{\prime}(0,(\chi_{K})_{\infty})=L^{\prime}(0,\chi_{\infty}\epsilon)L(0,\chi_{\infty})+L(0,\chi_{\infty}\epsilon)L^{\prime}(0,\chi_{\infty}).

But since χ(1)=1a=0\chi_{\infty}(-1)=1\implies a=0, we have

L(0,χ)=L(1,χ¯)Γ(12)Γ(0)(kπ)1/2τ(χ)k.L(0,\chi_{\infty})=L(1,\overline{\chi_{\infty}})\frac{\Gamma\left(\frac{1}{2}\right)}{\Gamma(0)}\left(\frac{k}{\pi}\right)^{1/2}\frac{\tau(\chi)}{\sqrt{k}}.

Notably, 1Γ(0)=0\frac{1}{\Gamma(0)}=0, which concludes the result. ∎

The explicit formulas of Dirichlet and Kronecker are as follows:

  • L(0,(χK))=16prAχ(a)logF+(a)L^{\prime}(0,(\chi_{K})_{\infty})=-\frac{1}{6p^{r}}\sum_{A}\chi_{\infty}(a)\log F^{+}(a)_{\infty} [12, p. 91],

  • L(0,χ)=a=1fafχ(a)=B1,χL(0,\chi)=-\sum_{a=1}^{f}\frac{a}{f}\chi(a)=-B_{1,\chi} for ff the conductor of χ\chi [12, p. 88],

  • L(1,χ)=g(χ)Aχ1(a)logC+(a)L(1,\chi_{\infty})=-g(\chi_{\infty})\sum_{A}\chi_{\infty}^{-1}(a)\log C^{+}(a)_{\infty} [12, p. 91] where g(χ)=1fa=1fχ(a)e2πia/fg(\chi)=\frac{1}{f}\sum_{a=1}^{f}\chi(a)e^{2\pi ia/f} for ff the conductor of χ\chi [12, p. 88].

Proposition 2.2.

L(0,χ)=g(χ)2L(1,χ¯)L^{\prime}(0,\chi_{\infty})=-\frac{g(\chi_{\infty})}{2}L(1,\overline{\chi_{\infty}}).

Proof.

By differentiating, we have

L(s,χ)=\displaystyle L^{\prime}(s,\chi)= L(1s,χ¯)Γ(1s+a2)Γ(s+a2)(kπ)12sg(χ)iak,\displaystyle L^{\prime}(1-s,\overline{\chi})\frac{\Gamma\left(\frac{1-s+a}{2}\right)}{\Gamma\left(\frac{s+a}{2}\right)}\left(\frac{k}{\pi}\right)^{\frac{1}{2}-s}\frac{g(\chi)}{i^{a}\sqrt{k}},
12L(1s,χ¯)Γ(1s+a2)Γ(s+a2)(kπ)12sg(χ)iak,\displaystyle-\frac{1}{2}L(1-s,\overline{\chi})\frac{\Gamma^{\prime}\left(\frac{1-s+a}{2}\right)}{\Gamma\left(\frac{s+a}{2}\right)}\left(\frac{k}{\pi}\right)^{\frac{1}{2}-s}\frac{g(\chi)}{i^{a}\sqrt{k}},
Γ(s+a2)2L(1s,χ¯)Γ(1s+a2)Γ(s+a2)2(kπ)12sg(χ)iak,\displaystyle-\frac{\Gamma^{\prime}\left(\frac{s+a}{2}\right)}{2}L(1-s,\overline{\chi})\frac{\Gamma\left(\frac{1-s+a}{2}\right)}{\Gamma\left(\frac{s+a}{2}\right)^{2}}\left(\frac{k}{\pi}\right)^{\frac{1}{2}-s}\frac{g(\chi)}{i^{a}\sqrt{k}},
log(k/π)L(1s,χ¯)Γ(1s+a2)Γ(s+a2)(kπ)12sg(χ)iak.\displaystyle-\log(k/\pi)L(1-s,\overline{\chi})\frac{\Gamma\left(\frac{1-s+a}{2}\right)}{\Gamma\left(\frac{s+a}{2}\right)}\left(\frac{k}{\pi}\right)^{-\frac{1}{2}-s}\frac{g(\chi)}{i^{a}\sqrt{k}}.

Since χ(1)=1\chi_{\infty}(-1)=1, we have a=0a=0. Setting s=0s=0 yields

L(0,χ)=\displaystyle L^{\prime}(0,\chi_{\infty})= L(1,χ¯)Γ(12)Γ(0)(kπ)12g(χ)k,\displaystyle L^{\prime}(1,\overline{\chi_{\infty}})\frac{\Gamma\left(\frac{1}{2}\right)}{\Gamma(0)}\left(\frac{k}{\pi}\right)^{\frac{1}{2}}\frac{g(\chi_{\infty})}{\sqrt{k}},
12L(1,χ¯)Γ(12)Γ(0)(kπ)12g(χ)k,\displaystyle-\frac{1}{2}L(1,\overline{\chi_{\infty}})\frac{\Gamma^{\prime}\left(\frac{1}{2}\right)}{\Gamma(0)}\left(\frac{k}{\pi}\right)^{\frac{1}{2}}\frac{g(\chi_{\infty})}{\sqrt{k}},
Γ(0)2L(1,χ¯)Γ(12)Γ(0)2(kπ)12g(χ)k,\displaystyle-\frac{\Gamma^{\prime}(0)}{2}L(1,\overline{\chi_{\infty}})\frac{\Gamma\left(\frac{1}{2}\right)}{\Gamma(0)^{2}}\left(\frac{k}{\pi}\right)^{\frac{1}{2}}\frac{g(\chi_{\infty})}{\sqrt{k}},
log(k/π)L(1,χ¯)Γ(12)Γ(0)(kπ)12g(χ)k.\displaystyle-\log(k/\pi)L(1,\overline{\chi_{\infty}})\frac{\Gamma\left(\frac{1}{2}\right)}{\Gamma(0)}\left(\frac{k}{\pi}\right)^{-\frac{1}{2}}\frac{g(\chi_{\infty})}{\sqrt{k}}.

Note that the Laurent series of Γ(s)\Gamma(s) is Γ(s)=1s+a0+a1s+\Gamma(s)=\frac{1}{s}+a_{0}+a_{1}s+\dots which implies that

1Γ(0)=0,Γ(0)Γ(0)2=(1s2+a1+1s2+2a0s+)|s=0=1.\frac{1}{\Gamma(0)}=0,\hskip 14.22636pt\frac{\Gamma^{\prime}(0)}{\Gamma(0)^{2}}=\left(\frac{\frac{-1}{s^{2}}+a_{1}+\dots}{\frac{1}{s^{2}}+\frac{2a_{0}}{s}+\dots}\right)\bigg{|}_{s=0}=-1.

Furthermore, Γ(1/2)=π\Gamma(1/2)=\sqrt{\pi}. Combining these, we find that three of the terms cancel, which yields L(0,χ)=g(χ)2L(1,χ¯).L^{\prime}(0,\chi_{\infty})=-\frac{g(\chi_{\infty})}{2}L(1,\overline{\chi_{\infty}}).

Using the identity L(1,χ)=g(χ)Aχ¯(a)logC+(a)L(1,\chi_{\infty})=-g(\chi_{\infty})\sum_{A}\overline{\chi_{\infty}}(a)\log C^{+}(a)_{\infty}, and the fact that τ(χ¯)g(χ)=τ(χ)¯τ(χ)f=|f|2f=1\tau(\overline{\chi_{\infty}})g(\chi_{\infty})=\frac{\overline{\tau(\chi_{\infty})}\tau(\chi_{\infty})}{f}=\frac{|\sqrt{f}|^{2}}{f}=1, we find that

L(0,χ)=12Aχ(a)logC+(a).L^{\prime}(0,\chi_{\infty})=\frac{1}{2}\sum_{A}\chi_{\infty}(a)\log C^{+}(a)_{\infty}.

Now, combining these with the fact that L(0,χϵ)=B1,χϵL(0,\chi_{\infty}\epsilon)=-B_{1,\chi_{\infty}\epsilon}, we find that

16prAχ(a)logF+(a)=(B1,χϵ)(12Aχ(a)logC+(a)),-\frac{1}{6p^{r}}\sum_{A}\chi_{\infty}(a)\log F^{+}(a)_{\infty}=(-B_{1,\chi_{\infty}\epsilon})\left(-\frac{1}{2}\sum_{A}\chi_{\infty}(a)\log C^{+}(a)_{\infty}\right),

or equivalently

Proposition 2.3.

The equation

3prB1,χϵAχ(a)logC+(a)=Aχ(a)logF+(a)-3p^{r}B_{1,\chi_{\infty}\epsilon}\sum_{A}\chi_{\infty}(a)\log C^{+}(a)_{\infty}=\sum_{A}\chi_{\infty}(a)\log F^{+}(a)_{\infty}

holds for all pp.

Remark.

This is [12, (3.5)], but he only proves it for split pp.

Now recall that C+(a)C^{+}(a)_{\infty} and F+(a)F^{+}(a)_{\infty} are pp-units in the field Mpr=(cos2πpr)M_{p^{r}}=\mathbb{Q}\left(\cos\frac{2\pi}{p^{r}}\right). Let E(Mpr)E(M_{p^{r}}) denote the group of all pp-units. It is a finitely generated subgroup of ×\mathbb{R}^{\times}. Now consider the complex vector space E(Mpr)\mathbb{C}\otimes_{\mathbb{Z}}E(M_{p^{r}}). This is isomorphic to the regular representation of A=Gal(Mpr/)A=\text{Gal}(M_{p^{r}}/\mathbb{Q}). Now note that for all σA\sigma\in A, due to transport of structure, we have that

σ(Aχ(a)C+(a))\displaystyle\sigma\left(\sum_{A}\chi_{\infty}(a)\otimes_{\mathbb{Z}}C^{+}(a)_{\infty}\right) =Aχ(a)C+(σa)=χ1(σ)Aχ(σa)C+(σa),\displaystyle=\sum_{A}\chi_{\infty}(a)\otimes_{\mathbb{Z}}C^{+}(\sigma a)_{\infty}=\chi_{\infty}^{-1}(\sigma)\sum_{A}\chi_{\infty}(\sigma a)\otimes_{\mathbb{Z}}C^{+}(\sigma a)_{\infty},
=χ1(σ)(Aχ(a)C+(a)),\displaystyle=\chi_{\infty}^{-1}(\sigma)\left(\sum_{A}\chi_{\infty}(a)\otimes_{\mathbb{Z}}C^{+}(a)_{\infty}\right),
σ(Aχ(a)F+(a))\displaystyle\sigma\left(\sum_{A}\chi_{\infty}(a)\otimes_{\mathbb{Z}}F^{+}(a)_{\infty}\right) =Aχ(a)F+(σa)=χ1(σ)Aχ(σa)F+(σa),\displaystyle=\sum_{A}\chi_{\infty}(a)\otimes_{\mathbb{Z}}F^{+}(\sigma a)_{\infty}=\chi_{\infty}^{-1}(\sigma)\sum_{A}\chi_{\infty}(\sigma a)\otimes_{\mathbb{Z}}F^{+}(\sigma a)_{\infty},
=χ1(σ)(Aχ(a)F+(a)).\displaystyle=\chi_{\infty}^{-1}(\sigma)\left(\sum_{A}\chi_{\infty}(a)\otimes_{\mathbb{Z}}F^{+}(a)_{\infty}\right).

This implies that both Aχ(a)C+(a)\sum_{A}\chi_{\infty}(a)\otimes_{\mathbb{Z}}C^{+}(a)_{\infty} and Aχ(a)F+(a)\sum_{A}\chi_{\infty}(a)\otimes_{\mathbb{Z}}F^{+}(a)_{\infty} lie in the χ1\chi_{\infty}^{-1}-eigenspace of E(Mpr)\mathbb{C}\otimes_{\mathbb{Z}}E(M_{p^{r}}), which is one-dimensional. Therefore

c~Aχ(a)C+(a)=Aχ(a)F+(a)\tilde{c}\sum_{A}\chi_{\infty}(a)\otimes_{\mathbb{Z}}C^{+}(a)_{\infty}=\sum_{A}\chi_{\infty}(a)\otimes_{\mathbb{Z}}F^{+}(a)_{\infty}

for some c~\tilde{c}\in\mathbb{C}. Consider the map

γ:E(Mpr),\gamma:\mathbb{C}\otimes_{\mathbb{Z}}E(M_{p^{r}})\rightarrow\mathbb{C},

defined by γ(ca)=cloga\gamma(c\otimes a)=c\log a. This map is clearly \mathbb{C}-linear, so

γ(c~Aχ(a)C+(a))=c~γ(Aχ(a)C+(a)).\gamma(\tilde{c}\sum_{A}\chi_{\infty}(a)\otimes_{\mathbb{Z}}C^{+}(a)_{\infty})=\tilde{c}\gamma(\sum_{A}\chi_{\infty}(a)\otimes_{\mathbb{Z}}C^{+}(a)_{\infty}).

Applying γ\gamma to both sides of c~Aχ(a)C+(a)=Aχ(a)F+(a)\tilde{c}\sum_{A}\chi_{\infty}(a)\otimes_{\mathbb{Z}}C^{+}(a)_{\infty}=\sum_{A}\chi_{\infty}(a)\otimes_{\mathbb{Z}}F^{+}(a)_{\infty} yields that

c~=3prB1,χϵ.\tilde{c}=-3p^{r}B_{1,\chi_{\infty}\epsilon}.

In particular, note that E(Mpr)¯E(M_{p^{r}})\subset\overline{\mathbb{Q}}. We can actually say that

Proposition 2.4.

As elements of ¯E(Mpr)\overline{\mathbb{Q}}\otimes_{\mathbb{Z}}E(M_{p^{r}}), we have

3prB1,χϵAχ(a)C+(a)=Aχ(a)F+(a).-3p^{r}B_{1,\chi_{\infty}\epsilon}\sum_{A}\chi_{\infty}(a)\otimes_{\mathbb{Z}}C^{+}(a)_{\infty}=\sum_{A}\chi_{\infty}(a)\otimes_{\mathbb{Z}}F^{+}(a)_{\infty}.

Now take some φ:¯p\varphi:\overline{\mathbb{Q}}\xhookrightarrow{}\mathbb{C}_{p}. Applying (1logp)φ(1\otimes_{\mathbb{Z}}\log_{p})\circ\varphi to both sides, we find the following equality in p\mathbb{C}_{p}:

3prB1,χϵAχ(a)logpC+(a)p=Aχ(a)logpF+(a)p.-3p^{r}B_{1,\chi\epsilon}\sum_{A}\chi(a)\log_{p}C^{+}(a)_{p}=\sum_{A}\chi(a)\log_{p}F^{+}(a)_{p}.

Now consider the explicit formulas provided by [12, p. 93]:

Lp(0,χK)\displaystyle L_{p}(0,\chi_{K}) =13prg(χ1)Aχ(a)logpF+(a)p,\displaystyle=-\frac{1}{3p^{r}}g(\chi^{-1})\sum_{A}\chi(a)\log_{p}F^{+}(a)_{p},
Lp(0,χϵω)\displaystyle L_{p}(0,\chi\epsilon\omega) =B1,χϵ,\displaystyle=-B_{1,\chi\epsilon},
Lp(1,χ1)\displaystyle L_{p}(1,\chi^{-1}) =g(χ1)Aχ(a)logpC+(a)p.\displaystyle=-g(\chi^{-1})\sum_{A}\chi(a)\log_{p}C^{+}(a)_{p}.

Putting these together yields the pp-adic identity

Proposition 2.5.
Lp(0,χK)=Lp(0,χϵω)Lp(1,χ1).L_{p}(0,\chi_{K})=L_{p}(0,\chi\epsilon\omega)L_{p}(1,\chi^{-1}).

Now, following [12, p. 93], there exist measures λ2,λ3\lambda_{2},\lambda_{3} such that for any finite even Dirichlet character χ\chi of conductor prp^{r}, we have

χ,λ2\displaystyle\langle\chi,\lambda_{2}\rangle =Lp(0,χϵω),\displaystyle=L_{p}(0,\chi\epsilon\omega),
χ,λ3\displaystyle\langle\chi,\lambda_{3}\rangle =Lp(1,χ1).\displaystyle=L_{p}(1,\chi^{-1}).

Now define a measure λ1\lambda_{1} given by

χ,λ1=Lp(0,χK).\langle\chi,\lambda_{1}\rangle=L_{p}(0,\chi_{K}).

Then we have the equality

χ,λ1=χ,λ2χ,λ3\langle\chi,\lambda_{1}\rangle=\langle\chi,\lambda_{2}\rangle\langle\chi,\lambda_{3}\rangle

for all finite even Dirichlet characters χ\chi with conductor prp^{r}. Thus we have that

Theorem 2.6.

λ1=λ2λ3\lambda_{1}=\lambda_{2}\cdot\lambda_{3}.

2.2. Generalization to conductor not a prime power

We will now work more generally and extend to the remaining cases. We fix ff to be some positive integer with at least two distinct prime divisors. We again start with the classical factorization L(s,(χK))=L(s,χϵ)L(s,χ)L(s,(\chi_{K})_{\infty})=L(s,\chi_{\infty}\epsilon)L(s,\chi_{\infty}) and the functional equation for L(s,χ)L(s,\chi):

L(s,χ)=L(1s,χ¯)Γ(1s+a2)Γ(s+a2)(fπ)12sτ(χ)iaf,L(s,\chi)=L(1-s,\overline{\chi})\frac{\Gamma\left(\frac{1-s+a}{2}\right)}{\Gamma\left(\frac{s+a}{2}\right)}\left(\frac{f}{\pi}\right)^{\frac{1}{2}-s}\frac{\tau(\chi)}{i^{a}\sqrt{f}},

where Γ\Gamma is the gamma function, ff is the conductor of χ\chi, τ=n=1fχ(n)e2πin/f\tau=\sum_{n=1}^{f}\chi(n)e^{2\pi in/f} is the Gauss sum, and a=0a=0 if χ(1)=1\chi(-1)=1 while a=1a=1 if χ(1)=1\chi(-1)=-1.

Proposition 2.7.

L(0,(χK))=L(0,χϵ)L(0,χ)L^{\prime}(0,(\chi_{K})_{\infty})=L(0,\chi_{\infty}\epsilon)L^{\prime}(0,\chi_{\infty}).

Proof.

By differentiating,

L(s,(χK))=L(s,χϵ)L(s,χ)+L(s,χϵ)L(s,χ).L^{\prime}(s,(\chi_{K})_{\infty})=L^{\prime}(s,\chi_{\infty}\epsilon)L(s,\chi_{\infty})+L(s,\chi_{\infty}\epsilon)L^{\prime}(s,\chi_{\infty}).

Setting s=0s=0 yields

L(0,(χK))=L(0,χϵ)L(0,χ)+L(0,χϵ)L(0,χ).L^{\prime}(0,(\chi_{K})_{\infty})=L^{\prime}(0,\chi_{\infty}\epsilon)L(0,\chi_{\infty})+L(0,\chi_{\infty}\epsilon)L^{\prime}(0,\chi_{\infty}).

But since χ(1)=1a=0\chi_{\infty}(-1)=1\implies a=0, we have

L(0,χ)=L(1,χ¯)Γ(12)Γ(0)(fπ)1/2τ(χ)f.L(0,\chi_{\infty})=L(1,\overline{\chi_{\infty}})\frac{\Gamma\left(\frac{1}{2}\right)}{\Gamma(0)}\left(\frac{f}{\pi}\right)^{1/2}\frac{\tau(\chi)}{\sqrt{f}}.

Since 1Γ(0)=0\frac{1}{\Gamma(0)}=0, which concludes the result. ∎

We have the following explicit formulas:

  • L(0,χ)=a=1fafχ(a)=B1,χL(0,\chi)=-\sum_{a=1}^{f}\frac{a}{f}\chi(a)=-B_{1,\chi} [12, p. 88],

  • L(1,χ)=τ(χ)fAχ1(a)logC+(a)L(1,\chi_{\infty})=-\frac{\tau(\chi_{\infty})}{f}\sum_{A}\chi_{\infty}^{-1}(a)\log C^{+}(a)_{\infty} [12, p. 88] (note that this is for general conductors ff),

  • L(0,χ)=16fw(𝔣)c𝒞χ(c)log|E(c)|L^{\prime}(0,\chi)=-\frac{1}{6fw(\mathfrak{f})}\sum_{c\in\mathcal{C}}\chi(c)\log|E(c)| [24, p. 281], where χ\chi is a ray class character modulo 𝔣\mathfrak{f}, w(𝔣)w(\mathfrak{f}) is the number of roots of unity equivalent to 11 mod 𝔣\mathfrak{f}, and ff is the conductor of χ\chi.

Note that Cl(f)=(/f)×Cl(f)=(\mathbb{Z}/f\mathbb{Z})^{\times}, but A=Gal((cos2πf)/)=(/f)×/±1A=\text{Gal}(\mathbb{Q}(\cos\frac{2\pi}{f})/\mathbb{Q})=(\mathbb{Z}/f\mathbb{Z})^{\times}/\pm 1.

Proposition 2.8.

L(0,χ)=12Aχ(a)logC+(a)L^{\prime}(0,\chi_{\infty})=-\frac{1}{2}\sum_{A}\chi_{\infty}(a)\log C^{+}(a)_{\infty}.

Proof.

First note that the classical factorization (with value a=0a=0) yields

L(s,χ)Γ(s/2)=L(1s,χ1)Γ((1s)/2)(fπ)12sτ(χ)f.L(s,\chi_{\infty})\Gamma(s/2)=L(1-s,\chi_{\infty}^{-1})\Gamma((1-s)/2)\left(\frac{f}{\pi}\right)^{\frac{1}{2}-s}\frac{\tau(\chi_{\infty})}{\sqrt{f}}.

In particular, consider the left hand side’s power series expansion around s=0s=0: although L(s,χ)L(s,\chi_{\infty}) vanishes at s=0s=0, Γ(0)\Gamma(0) has a pole of order 11. But since the residue of Γ(s/2)\Gamma(s/2) is 22, we have that

2L(0,χ)=L(1,χ1)Γ(1/2)f1/πτ(χ)f=τ(χ)L(1,χ1).2L^{\prime}(0,\chi_{\infty})=L(1,\chi_{\infty}^{-1})\Gamma(1/2)\sqrt{f}\sqrt{1/\pi}\frac{\tau(\chi_{\infty})}{\sqrt{f}}=\tau(\chi_{\infty})L(1,\chi_{\infty}^{-1}).

Now using Gross’s formula, we find that

L(0,χ)=12Aχ(a)logC+(a).L^{\prime}(0,\chi_{\infty})=-\frac{1}{2}\sum_{A}\chi_{\infty}(a)\log C^{+}(a).

Proposition 2.9.

L(0,(χK))=16fAχ(a)logF+(a)L^{\prime}(0,(\chi_{K})_{\infty})=-\frac{1}{6f}\sum_{A}\chi(a)\log F^{+}(a).

Proof.

Note that we have a quotient homomorphism between the ray class group 𝒞\mathcal{C} modulo 𝔣\mathfrak{f}, isomorphic to (/f)×(\mathbb{Z}/f\mathbb{Z})^{\times}, with A=Gal(K/)=(/f)×/±1A=\text{Gal}(K/\mathbb{Q})=(\mathbb{Z}/f\mathbb{Z})^{\times}/\pm 1. Furthermore, w(𝔣)=1w(\mathfrak{f})=1 since KK is totally real. Then [24] gives the result

L(0,(χK))=16f𝒞χ(c)log|E(c)|.L^{\prime}(0,(\chi_{K})_{\infty})=-\frac{1}{6f}\sum_{\mathcal{C}}\chi_{\infty}(c)\log|E(c)|.

But note that

|E(c)|2=E(c)E(c)=Ff(a)Ff(a)¯=Ff(a)Ff(a)|E(c)|^{2}=E(c)E(-c)=F_{f}(a)\overline{F_{f}(a)}=F_{f}(a)F_{f}(-a)
logF+(a)=logFf(a)Ff(a)=logE(a)E(a)¯=log|E(a)|+log|E(a)|,\implies\log F^{+}(a)=\log F_{f}(a)F_{f}(-a)=\log E(a)\overline{E(a)}=\log|E(a)|+\log|E(-a)|,

so we have

L(0,(χK))=16fAχ(a)logF+(a).L^{\prime}(0,(\chi_{K})_{\infty})=-\frac{1}{6f}\sum_{A}\chi_{\infty}(a)\log F^{+}(a).

Now, combining these with the fact that L(0,χϵ)=B1,χϵL(0,\chi_{\infty}\epsilon)=-B_{1,\chi_{\infty}\epsilon}, we find that

16fAχ(a)logF+(a)=(B1,χϵ)(12Aχ(a)logC+(a)),-\frac{1}{6f}\sum_{A}\chi_{\infty}(a)\log F^{+}(a)_{\infty}=(-B_{1,\chi_{\infty}\epsilon})\left(-\frac{1}{2}\sum_{A}\chi_{\infty}(a)\log C^{+}(a)_{\infty}\right),

or equivalently

Proposition 2.10.

The equation

3fB1,χϵAχ(a)logC+(a)=Aχ(a)logF+(a)-3fB_{1,\chi_{\infty}\epsilon}\sum_{A}\chi_{\infty}(a)\log C^{+}(a)_{\infty}=\sum_{A}\chi_{\infty}(a)\log F^{+}(a)_{\infty}

holds for all ff.

Remark.

This is [12, (3.5)], but he only proves it for split pp and f=prf=p^{r}. Note that χ\chi_{\infty} depends on ff.

Now note that C+(a)C^{+}(a)_{\infty} and F+(a)F^{+}(a)_{\infty} are units in the field Mf=(cos2πf)M_{f}=\mathbb{Q}\left(\cos\frac{2\pi}{f}\right). (In particular,

a(/f)×/±1(1e2πia/f)=Φf(1)=1,\prod_{a\in(\mathbb{Z}/f\mathbb{Z})^{\times}/\pm 1}(1-e^{2\pi ia/f})=\Phi_{f}(1)=1,

which holds whenever ff has at least two distinct prime divisors.) Let E(Mf)E(M_{f}) denote the group of all units. It is a finitely generated subgroup of ×\mathbb{R}^{\times} of rank |A|1|A|-1, by Dirichlet’s unit theorem. Now consider the complex vector space E(Mf)\mathbb{C}\otimes_{\mathbb{Z}}E(M_{f}). This is isomorphic to the quotient of the regular representation of A=Gal(Mf/)A=\text{Gal}(M_{f}/\mathbb{Q}) by the subspace spanned by (1,1,1,)(1,1,1,\dots). For all σA\sigma\in A, due to transport of structure, we have that

σ(Aχ(a)C+(a))\displaystyle\sigma\left(\sum_{A}\chi_{\infty}(a)\otimes_{\mathbb{Z}}C^{+}(a)_{\infty}\right) =Aχ(a)C+(σa)=χ1(σ)Aχ(σa)C+(σa),\displaystyle=\sum_{A}\chi_{\infty}(a)\otimes_{\mathbb{Z}}C^{+}(\sigma a)_{\infty}=\chi_{\infty}^{-1}(\sigma)\sum_{A}\chi_{\infty}(\sigma a)\otimes_{\mathbb{Z}}C^{+}(\sigma a)_{\infty},
=χ1(σ)(Aχ(a)C+(a)),\displaystyle=\chi_{\infty}^{-1}(\sigma)\left(\sum_{A}\chi_{\infty}(a)\otimes_{\mathbb{Z}}C^{+}(a)_{\infty}\right),
σ(Aχ(a)F+(a))\displaystyle\sigma\left(\sum_{A}\chi_{\infty}(a)\otimes_{\mathbb{Z}}F^{+}(a)_{\infty}\right) =Aχ(a)F+(σa)=χ1(σ)Aχ(σa)F+(σa),\displaystyle=\sum_{A}\chi_{\infty}(a)\otimes_{\mathbb{Z}}F^{+}(\sigma a)_{\infty}=\chi_{\infty}^{-1}(\sigma)\sum_{A}\chi_{\infty}(\sigma a)\otimes_{\mathbb{Z}}F^{+}(\sigma a)_{\infty},
=χ1(σ)(Aχ(a)F+(a)).\displaystyle=\chi_{\infty}^{-1}(\sigma)\left(\sum_{A}\chi_{\infty}(a)\otimes_{\mathbb{Z}}F^{+}(a)_{\infty}\right).

This implies that both Aχ(a)C+(a)\sum_{A}\chi_{\infty}(a)\otimes_{\mathbb{Z}}C^{+}(a)_{\infty} and Aχ(a)F+(a)\sum_{A}\chi_{\infty}(a)\otimes_{\mathbb{Z}}F^{+}(a)_{\infty} lie in the χ1\chi_{\infty}^{-1}-eigenspace of E(Mf)\mathbb{C}\otimes_{\mathbb{Z}}E(M_{f}), which is one-dimensional since AA is abelian. Therefore

c~Aχ(a)C+(a)=Aχ(a)F+(a)\tilde{c}\sum_{A}\chi_{\infty}(a)\otimes_{\mathbb{Z}}C^{+}(a)_{\infty}=\sum_{A}\chi_{\infty}(a)\otimes_{\mathbb{Z}}F^{+}(a)_{\infty}

for some c~\tilde{c}\in\mathbb{C}. Consider the map

γ:E(Mf),\gamma:\mathbb{C}\otimes_{\mathbb{Z}}E(M_{f})\rightarrow\mathbb{C},

defined by γ(ca)=cloga\gamma(c\otimes a)=c\log a. This map is clearly \mathbb{C}-linear, so

γ(c~Aχ(a)C+(a))=c~γ(Aχ(a)C+(a)).\gamma(\tilde{c}\sum_{A}\chi_{\infty}(a)\otimes_{\mathbb{Z}}C^{+}(a)_{\infty})=\tilde{c}\gamma(\sum_{A}\chi_{\infty}(a)\otimes_{\mathbb{Z}}C^{+}(a)_{\infty}).

Applying γ\gamma to both sides of c~Aχ(a)C+(a)=Aχ(a)F+(a)\tilde{c}\sum_{A}\chi_{\infty}(a)\otimes_{\mathbb{Z}}C^{+}(a)_{\infty}=\sum_{A}\chi_{\infty}(a)\otimes_{\mathbb{Z}}F^{+}(a)_{\infty} yields that

c~=3fB1,χϵ.\tilde{c}=-3fB_{1,\chi_{\infty}\epsilon}.

In particular, note that E(Mf)¯E(M_{f})\subset\overline{\mathbb{Q}}. We can actually say that

Proposition 2.11.

As elements of ¯E(Mf)\overline{\mathbb{Q}}\otimes_{\mathbb{Z}}E(M_{f}), we have

3fB1,χϵAχ(a)C+(a)=Aχ(a)F+(a).-3fB_{1,\chi_{\infty}\epsilon}\sum_{A}\chi_{\infty}(a)\otimes_{\mathbb{Z}}C^{+}(a)_{\infty}=\sum_{A}\chi_{\infty}(a)\otimes_{\mathbb{Z}}F^{+}(a)_{\infty}.

Now take some φ:¯p\varphi:\overline{\mathbb{Q}}\xhookrightarrow{}\mathbb{C}_{p}. Applying (1logp)φ(1\otimes_{\mathbb{Z}}\log_{p})\circ\varphi to both sides, we find the following equality in p\mathbb{C}_{p}:

3fB1,χϵAχ(a)logpC+(a)p=Aχ(a)logpF+(a)p.-3fB_{1,\chi\epsilon}\sum_{A}\chi(a)\log_{p}C^{+}(a)_{p}=\sum_{A}\chi(a)\log_{p}F^{+}(a)_{p}.

Now consider the explicit formulas provided by [12, p. 93]:

Lp(0,χK)\displaystyle L_{p}(0,\chi_{K}) =13fg(χ1)Aχ(a)logpF+(a)p,\displaystyle=-\frac{1}{3f}g(\chi^{-1})\sum_{A}\chi(a)\log_{p}F^{+}(a)_{p},
Lp(0,χϵω)\displaystyle L_{p}(0,\chi\epsilon\omega) =B1,χϵ,\displaystyle=-B_{1,\chi\epsilon},
Lp(1,χ1)\displaystyle L_{p}(1,\chi^{-1}) =g(χ1)Aχ(a)logpC+(a)p.\displaystyle=-g(\chi^{-1})\sum_{A}\chi(a)\log_{p}C^{+}(a)_{p}.

Putting these together yields the pp-adic identity

Proposition 2.12.
Lp(0,χK)=Lp(0,χϵω)Lp(1,χ1).L_{p}(0,\chi_{K})=L_{p}(0,\chi\epsilon\omega)L_{p}(1,\chi^{-1}).

Now, following [12, p. 93] there exist measures λ2,λ3\lambda_{2},\lambda_{3} such that for any finite even Dirichlet character χ\chi of conductor ff, we have

χ,λ2\displaystyle\langle\chi,\lambda_{2}\rangle =Lp(0,χϵω),\displaystyle=L_{p}(0,\chi\epsilon\omega),
χ,λ3\displaystyle\langle\chi,\lambda_{3}\rangle =Lp(1,χ1).\displaystyle=L_{p}(1,\chi^{-1}).

Now define a measure λ1\lambda_{1} given by

χ,λ1=Lp(0,χK).\langle\chi,\lambda_{1}\rangle=L_{p}(0,\chi_{K}).

Then we have the equality

χ,λ1=χ,λ2χ,λ3\langle\chi,\lambda_{1}\rangle=\langle\chi,\lambda_{2}\rangle\langle\chi,\lambda_{3}\rangle

for all finite even Dirichlet characters χ\chi with conductor ff. Combining with Theorem 2.6, we have the following factorization:

Theorem 2.13.

For any finite even Dirichlet character χ\chi, with χ,λ1=Lp(0,χK)\langle\chi,\lambda_{1}\rangle=L_{p}(0,\chi_{K}), χ,λ2=Lp(0,χϵω)\langle\chi,\lambda_{2}\rangle=L_{p}(0,\chi\epsilon\omega), and χ,λ3=Lp(1,χ1)\langle\chi,\lambda_{3}\rangle=L_{p}(1,\chi^{-1}), we have that λ1=λ2λ3\lambda_{1}=\lambda_{2}\cdot\lambda_{3}.

3. Assumptions and conventions

For the remainder of the article, we fix several assumptions. We will restate them throughout the article, but organize them here for convenience. We will let KK denote an imaginary quadratic field and ΔK\Delta_{K} the discriminant of KK. We will let E/FE/F denote an elliptic curve over a number field FF (usually either KK or \mathbb{Q}) with conductor cond(E)=N\text{cond}(E)=N. Let f(q)f(q) be the modular form associated to EE. We assume that EE will be residually reducible modulo 33:

Assumption 3.1 (Residually reducible).

All elliptic curves EE will be residually reducible modulo 33. In other words, the 33-adic Galois representation ρ3:Gal(F¯/F)Aut(T3(E))GL2(3)\rho_{3}:\text{Gal}(\overline{F}/F)\rightarrow\text{Aut}(T_{3}(E))\cong GL_{2}(\mathbb{Z}_{3}) reduced modulo 33 to ρ3:Gal(F¯/F)GL2(𝔽3)\rho_{3}:\text{Gal}(\overline{F}/F)\rightarrow GL_{2}(\mathbb{F}_{3}) is reducible.

We will also require that EE satisfies the Heegner hypothesis relative to KK in many situations, as found in [1, p. 8]:

Assumption 3.2 (Heegner hypothesis).

For every prime |N\ell|N, then ΔK\Delta_{K} is a quadratic residue modulo \ell.

We will focus a great deal of attention to quadratic twists of E/FE/F. Let EE be an elliptic curve given by y2=x3+ax+by^{2}=x^{3}+ax+b. Then for DFD^{\prime}\in F such that F(D)FF(\sqrt{D^{\prime}})\supsetneq F, the quadratic twist of E/FE/F by DD^{\prime} is given by Dy2=x3+ax+bD^{\prime}y^{2}=x^{3}+ax+b, and denoted E(D)E^{(D^{\prime})} with modular form f(D)(q)f^{(D^{\prime})}(q). We will primarily focus on DD^{\prime}\in\mathbb{Z}. We will later see that when E/FE/F is residually reducible, then f(D)(q)E2χD,χD(q)(mod3)f^{(D^{\prime})}(q)\equiv E_{2}^{\chi_{D},\chi_{D}}(q)\pmod{3} for some Eisenstein series EE and integer DD, and therefore we will denote EDE(D)E_{D}\coloneqq E^{(D^{\prime})}.

In a similar manner, we will denote Ed,3E_{d,3} the cubic twist of E/FE/F by dd, where the cubic twist is given by y2=x3+cy2=x3+dcy^{2}=x^{3}+c\mapsto y^{2}=x^{3}+dc.

Finally, we will denote by Assumption 3.3 the following series of assumptions on (N,D)(N,D).

Assumption 3.3.

We make the following assumptions:

  • For all primes >3\ell>3, if v(N)=1v_{\ell}(N)=1, then 2(mod3)\ell\equiv 2\pmod{3},

  • gcd(N,D)=1\gcd(N,D)=1,

  • 2ND2\nmid ND,

  • v3(N)1v_{3}(N)\neq 1.

4. Congruences modulo pp

4.1. Congruences of LL-series and Eisenstein series

Let

f(q)=n0anqnf(q)=\sum_{n\geq 0}a_{n}q^{n}

be the modular form attached to an elliptic curve EE. Let the 33-adic Galois representation be ρE\rho_{E}; we will assume that ρE\rho_{E} is always residually reducible modulo 33. Let

E2λ,ψ(q)=L(1k,χ)+nσλ,ψ(n)qnE_{2}^{\lambda,\psi}(q)=L(1-k,\chi)+\sum_{n}\sigma^{\lambda,\psi}(n)q^{n}

be an Eisenstein series, where

σλ,ψ(n)=d|nλ(n/d)ψ(d)d.\sigma^{\lambda,\psi}(n)=\sum_{d|n}\lambda(n/d)\psi(d)d.
Proposition 4.1.

The Galois representation of E2ψ,ψ¯E_{2}^{\psi,\overline{\psi}} is isomorphic up to semisimplification to ψψ¯χ\psi\oplus\overline{\psi}\chi where χ\chi is the cyclotomic character.

Proof.

The Brauer-Nesbitt theorem implies that up to semisimplication, ρE2\rho_{E_{2}} is determined by its characteristic polynomial, or equivalently, trace and determinant. Furthermore, the Cebotarev density function implies that the Frobenius elements are dense in the Galois group. Since ρE2\rho_{E_{2}} is a continuous function, it suffices to check that trace and determinant match on the Frobenius elements \ell for each prime. We have

tr()=[q]E2ψ,ψ¯(q)=σψ,ψ¯()=ψ()+ψ¯(),\text{tr}(\ell)=[q^{\ell}]E_{2}^{\psi,\overline{\psi}}(q)=\sigma^{\psi,\overline{\psi}}(\ell)=\psi(\ell)+\overline{\psi}(\ell)\ell,

which confirms that the trace function matches. The determinant yields

det()==ψ()ψ¯()χ(),\det(\ell)=\ell=\psi(\ell)\overline{\psi}(\ell)\chi(\ell),

and both functions match. ∎

Proposition 4.2.

Suppose ρE\rho_{E} is residually reducible, i.e. the representation mod3\bmod\hskip 2.84526pt3 is isomorphic to χ1χ2\chi_{1}\oplus\chi_{2} up to semisimplification. Then ρEρE2χM,χM(mod3)\rho_{E}\cong\rho_{E_{2}^{\chi_{M},\chi_{M}}}\pmod{3}, where E2χM,χME_{2}^{\chi_{M},\chi_{M}} is the Eisenstein series with χM\chi_{M} a quadratic character, and χ1=χM\chi_{1}=\chi_{M} and χ2=χMχ\chi_{2}=\chi_{M}\chi for χ\chi the cyclotomic character. Furthermore, f(q)E2χM,χM(q)(mod3)f(q)\equiv E_{2}^{\chi_{M},\chi_{M}}(q)\pmod{3}.

Proof.

Since 𝔽3×{±1}\mathbb{F}_{3}^{\times}\cong\{\pm 1\}, it follows that χ1,χ2\chi_{1},\chi_{2} are quadratic characters. By Brauer-Nesbitt, it suffices to check that the trace and determinant functions agree on Frobenius elements, which are dense in the Galois group by the Cebotarev density theorem. Checking the determinant function, we have that

χ()==det()=χ1()χ2().\chi(\ell)=\ell=\det(\ell)=\chi_{1}(\ell)\chi_{2}(\ell).

Thus

χ2=χ11χ=χ1χ.\chi_{2}=\chi_{1}^{-1}\chi=\chi_{1}\chi.

Letting χ1=χM\chi_{1}=\chi_{M} for some quadratic character modulo MM, we have that

χ1=χM,χ2=χMχ.\chi_{1}=\chi_{M},\hskip 5.69054pt\chi_{2}=\chi_{M}\chi.

Now, the trace functions yield that

a=χM()+χM()χ()=χM()+χM()=d|χM(/d)χM()d=σχM,χM()=[q]E2χM,χM(q).a_{\ell}=\chi_{M}(\ell)+\chi_{M}(\ell)\chi(\ell)=\chi_{M}(\ell)+\chi_{M}(\ell)\ell=\sum_{d|\ell}\chi_{M}(\ell/d)\chi_{M}(\ell)d=\sigma^{\chi_{M},\chi_{M}}(\ell)=[q^{\ell}]E_{2}^{\chi_{M},\chi_{M}}(q).

Since the coefficients of the two modular forms agree on prime indices, they agree on all nonconstant terms. Thus f(q)E2χM,χM(q)c(mod3)f(q)-E_{2}^{\chi_{M},\chi_{M}}(q)\equiv c\pmod{3}, where the left hand side is a modular form of weight 22, and thus the right hand side must also be a modular form of weight 22. By [20], c=0c=0, and we have that f(q)E2χM,χM(q)(mod3)f(q)\equiv E_{2}^{\chi_{M},\chi_{M}}(q)\pmod{3}. ∎

4.2. Congruences of quadratic twists

Consider some arbitrary squarefree DD^{\prime}\in\mathbb{Z}. We will study the quadratic twist of ff by DD^{\prime} and write it as f(D)(q)f^{(D^{\prime})}(q), with the elliptic curve Ey2=x3+ax+bE\coloneqq y^{2}=x^{3}+ax+b becoming E(D)Dy2=x3+ax+bE^{(D^{\prime})}\coloneqq D^{\prime}y^{2}=x^{3}+ax+b. We will assume that ρE\rho_{E} is residually reducible. By Proposition 4.2, we have f(q)E2χM,χM(q)(mod3)f(q)\equiv E_{2}^{\chi_{M},\chi_{M}}(q)\pmod{3} for some quadratic character χM\chi_{M}. Since fD(q)=n0χD(n)anqnf_{D^{\prime}}(q)=\sum_{n\geq 0}\chi_{D^{\prime}}(n)a_{n}q^{n}, it follows that fD(q)E2χMχD,χMχD(q)(mod3)f_{D^{\prime}}(q)\equiv E_{2}^{\chi_{M}\chi_{D^{\prime}},\chi_{M}\chi_{D^{\prime}}}(q)\pmod{3}. Since χD\chi_{D^{\prime}} is again a quadratic character, we may write χD=χMχD\chi_{D}=\chi_{M}\chi_{D^{\prime}} for some squarefree DD\in\mathbb{Z}. Thus every quadratic twist of an elliptic curve whose Galois representation is residually reducible mod3\bmod\hskip 2.84526pt3 is congruent to E2χD,χD(q)E_{2}^{\chi_{D},\chi_{D}}(q) modulo 33, where χD\chi_{D} is some quadratic character. From now on, we will write the quadratic twist of the elliptic curve E(D)E^{(D^{\prime})} as EDE_{D} and the associated modular form as fDf_{D}, where χD=χMχD\chi_{D}=\chi_{M}\chi_{D^{\prime}}.

4.3. Stabilizations

We follow [3] and describe the pp-depletions/stabilizations. Let f(q)=nanqnf(q)=\sum_{n}a_{n}q^{n} be the modular form of an elliptic curve E/FE/F. Then the pp-depletion is given by

f(q)=pnanqn.f^{\flat}(q)=\sum_{p\nmid n}a_{n}q^{n}.

Suppose pp is a good prime; then f=f|VUUVf^{\flat}=f|VU-UV, where UU and VV are given by [3, p. 1085].

Let NN be the conductor of E/FE/F. If a prime 2|N\ell^{2}|N, then a=0a_{\ell}=0, hence there is no need to change the value through \ell-depletion. If a prime |N\ell|N with v(N)=1v_{\ell}(N)=1, then a=±1a_{\ell}=\pm 1, so we may use 1V1\mp V_{\ell} instead of VUUVVU-UV. In particular, 1aV=1TV1-a_{\ell}V=1-T_{\ell}V suffices.

We assume the Heegner hypothesis [1, p. 8], Assumption 3.2: for every |N\ell|N, the ideal ()(\ell) splits in 𝒪F\mathcal{O}_{F} as 𝔩𝔩¯\mathfrak{l}\cdot\overline{\mathfrak{l}}.

Proposition 4.3.

The Euler factor at \ell of SχS_{\chi}^{\flat_{\ell}} for some |N\ell|N is 1χ1(𝔩¯)1-\chi^{-1}(\overline{\mathfrak{l}}).

Proof.

Following [3, p. 1135], we set

Sχ=[𝔞]χj1(𝔞)θjf(𝔞(A0,t0,ω0)).S_{\chi}^{\flat_{\ell}}=\sum_{[\mathfrak{a}]}\chi_{j}^{-1}(\mathfrak{a})\cdot\theta^{j}f^{\flat_{\ell}}(\mathfrak{a}*(A_{0},t_{0},\omega_{0})).

Let a=±1a_{\ell}=\pm 1. Then

θjf(𝔞(A0,t0,ω0))\displaystyle\theta^{j}f^{\flat_{\ell}}(\mathfrak{a}*(A_{0},t_{0},\omega_{0})) ={θjf|(1TV)}(𝔞(A0,t0,ω0)),\displaystyle=\{\theta^{j}f|(1\mp T_{\ell}V)\}(\mathfrak{a}*(A_{0},t_{0},\omega_{0})),
=θjf(𝔞(A0,t0,ω0))jaθjf(𝔩¯1𝔞(A0,t0,ω0)),\displaystyle=\theta^{j}f(\mathfrak{a}*(A_{0},t_{0},\omega_{0}))\mp\ell^{j}a_{\ell}\theta^{j}f(\overline{\mathfrak{l}}^{-1}\mathfrak{a}*(A_{0},t_{0},\omega_{0})),

so

Sχ=(1aχ1(𝔩¯))Sχ=(1χ1(𝔩¯))Sχ.S_{\chi}^{\flat_{\ell}}=(1\mp a_{\ell}\chi^{-1}(\overline{\mathfrak{l}}))S_{\chi}=(1-\chi^{-1}(\overline{\mathfrak{l}}))S_{\chi}.

Corollary 4.3.1.

Assuming the Heegner hypothesis, the Euler factor at p\ell\neq p does not vanish modulo pp when 1(modp)\ell\neq 1\pmod{p}.

Proof.

Since χ(𝔩¯)=\chi(\overline{\mathfrak{l}})=\ell, it suffices to have 110(modp)1(modp)1-\ell^{-1}\not\equiv 0\pmod{p}\iff\ell\not\equiv 1\pmod{p}. ∎

If f1(q)f_{1}(q) and f2(q)f_{2}(q) agree on all coefficients [qn]fi(q)[q^{n}]f_{i}(q) whenever pinp_{i}\nmid n for all piXp_{i}\in X where XX is a finite set of primes, then we may take the pip_{i}-depletions to force them to be equal. In particular,

f1X(q)=f2X(q),f_{1}^{\flat_{X}}(q)=f_{2}^{\flat_{X}}(q),

where fiX(q)f_{i}^{\flat_{X}}(q) indicates the modular form fi(q)f_{i}(q) after pjp_{j} depletions for each pjp_{j}.

5. Congruence of modular forms

5.1. Varying KK

We follow the discussion from section 33. For the remainder of this section, we set p=3p=3. Suppose we have an LL-series attached to an elliptic curve EE whose Fourier expansion is f(q)f(q) whose Galois representation ρE\rho_{E} modulo 33 is residually reducible. Then proposition 1515 implies that ρEρE2χM,χM(mod3)\rho_{E}\cong\rho_{E_{2}^{\chi_{M},\chi_{M}}}\pmod{3}, the Galois representation modulo 33 of the Eisenstein series E2χM,χME_{2}^{\chi_{M},\chi_{M}}. Now consider the quadratic twist by DD^{\prime}, so that f(D)(q)=χD(n)anqnf^{(D^{\prime})}(q)=\sum\chi_{D^{\prime}}(n)a_{n}q^{n}, where χD\chi_{D^{\prime}} is the Kronecker character. It follows that ρE(D)ρE2χD,χD(mod3)\rho_{E^{(D^{\prime})}}\cong\rho_{E_{2}^{\chi_{D},\chi_{D}}}\pmod{3} where χD=χDχM\chi_{D}=\chi_{D^{\prime}}\chi_{M} is some quadratic character. Now denote EDE(D)E_{D}\coloneqq E^{(D^{\prime})}, and fD(q)f(D)(q)f_{D}(q)\coloneqq f^{(D^{\prime})}(q), so that we parametrize the twists by the corresponding Eisenstein series. In particular, [q]f(q)[q]E2χD,χD(q)(mod3)[q^{\ell}]f(q)\equiv[q^{\ell}]E_{2}^{\chi_{D},\chi_{D}}(q)\pmod{3} for all cond(ED)=ND2\ell\nmid\text{cond}(E_{D})=ND^{2}, where N=cond(E)N=\text{cond}(E). We thus have that fD(q)f_{D}(q) and E2χD,χD(q)E_{2}^{\chi_{D},\chi_{D}}(q) are congruent modulo 33 everywhere except possibly at indices divisible by some bad prime \ell.

Let N=cond(E)N=\text{cond}(E). Assume pNDp\nmid ND. If gcd(N,D)=1\gcd(N,D)=1, then cond(ED)=level(fD)=ND2\text{cond}(E_{D})=\text{level}(f_{D})=ND^{2}. On the other hand, D=cond(χD)D=\text{cond}(\chi_{D}), so level(E2χD,χD)=D2\text{level}(E_{2}^{\chi_{D},\chi_{D}})=D^{2}. We need only stabilize (\ell-deplete) at primes \ell such that v(ND2)=1v_{\ell}(ND^{2})=1. Due to Corollary 4.3.1, we will require all such \ell to satisfy 1(mod3)\ell\not\equiv 1\pmod{3} for the rest of the paper. This is noted in the section on assumptions. For each of these \ell, we have D\ell\nmid D, so \ell is a good prime for E2χD,χDE_{2}^{\chi_{D},\chi_{D}}. On the other hand, for |D2\ell|D^{2}, it immediately follows that 2|D2\ell^{2}|D^{2} so [q]E2χD,χD=0[q^{\ell}]E_{2}^{\chi_{D},\chi_{D}}=0, and since 2|ND2\ell^{2}|ND^{2}, then [q]fD=0[q^{\ell}]f_{D}=0. Hence there is no need to \ell-deplete EkχD,χD(q)E_{k}^{\chi_{D},\chi_{D}}(q) at such primes (it is already zero), and we only need to consider the primes \ell for which v(N)=1v_{\ell}(N)=1.

Denote this set by XX. Then XX is a set of bad primes for fD(q)f_{D}(q), but good primes for E2χD,χD(q)E_{2}^{\chi_{D},\chi_{D}}(q). Take ZZ to be the product of X\ell\in X.

Note that Sχ(f)Sχ(f[])(modp)S_{\chi}^{\flat_{\ell}}(f)\equiv S_{\chi}(f^{[\ell]})\pmod{p} and fD(q)E2χD,χD(q)(modp)f_{D}(q)\equiv E_{2}^{\chi_{D},\chi_{D}}(q)\pmod{p}, so by the qq-expansion principle, we have

Proposition 5.1.

For infinity types χ\chi of type (k+j,j)(k+j,-j) with j0j\geq 0, we have Sχ(fD[pZ])Sχ(E2χD,χD[pZ])(modp)S_{\chi}^{\flat}(f_{D}^{[pZ]})\equiv S_{\chi}^{\flat}(E_{2}^{\chi_{D},\chi_{D}[pZ]})\pmod{p}.

Setting k=2k=2 and taking the limit of jm=pm1j_{m}=p^{m}-1 as mm\rightarrow\infty gives, by continuity,

Proposition 5.2.

For K\mathbb{N}_{K} the norm character of type (1,1)(1,1), we have

SKχD(fD[pZ])SKχD(E2χD,χD[pZ])(modp).S_{\mathbb{N}_{K}\chi_{D}}^{\flat}(f_{D}^{[pZ]})\equiv S_{\mathbb{N}_{K}\chi_{D}}^{\flat}(E_{2}^{\chi_{D},\chi_{D}[pZ]})\pmod{p}.

In fact, more generally:

Proposition 5.3.

Suppose we have two modular forms ff and gg with Galois representations ρf\rho_{f} and ρg\rho_{g} such that ρfρg(modp)\rho_{f}\equiv\rho_{g}\pmod{p}. Then Sχ(f[N])Sχ(g[N])(modp)S_{\chi}^{\flat}(f^{[N]})\equiv S_{\chi}^{\flat}(g^{[N]})\pmod{p}, where NN is such that [q]f[q]g(modp)[q^{\ell}]f\equiv[q^{\ell}]g\pmod{p} whenever N\ell\nmid N, and χ\chi is type (k+j,j)(k+j,-j) for j0j\geq 0. Furthermore, this is true for χ=K\chi=\mathbb{N}_{K}, the norm character of infinity type (1,1)(1,1).

Proof.

The stabilization at \ell yields [q]f[]=0[q^{\ell}]f^{[\ell]}=0, and thus after stabilizing at all |N\ell|N, it follows that

|N[q]f[N][q]g[N]0(modp)\ell|N\implies[q^{\ell}]f^{[N]}\equiv[q^{\ell}]g^{[N]}\equiv 0\pmod{p}

and by hypothesis, they are already congruent modulo pp when N\ell\nmid N. The qq-expansion principle then implies that

Sχ(f[N])Sχ(g[N])(modp).S_{\chi}^{\flat}(f^{[N]})\equiv S_{\chi}^{\flat}(g^{[N]})\pmod{p}.

To see that this holds for χ=K\chi=\mathbb{N}_{K}, take χ\chi of type (2+pm1,1pm)(2+p^{m}-1,1-p^{m}) as mm\rightarrow\infty. Since SχS_{\chi}^{\flat} is continuous in χ\chi, it follows that the limit is (21,1)=(1,1)(2-1,1)=(1,1) and the congruence for K\mathbb{N}_{K} holds. ∎

Define Lp,α(w,χ)=Sχ(w)L_{p,\alpha}(w,\chi)=S_{\chi}^{\flat}(w) as in [17, Definition 8.8]; we’ll write Lp(w,χ)L_{p}(w,\chi) as shorthand.

Proposition 5.4.

If Sχ(E2χD,χD[pZ])0(modp)S_{\chi}^{\flat}(E_{2}^{\chi_{D},\chi_{D}[pZ]})\not\equiv 0\pmod{p}, then ED/KE_{D}/K has rank 11.

Proof.

We have SKχD(E2χD,χD[pZ])=Lp(E2χD,χD[pZ],KχD)S_{\mathbb{N}_{K}\chi_{D}}^{\flat}(E_{2}^{\chi_{D},\chi_{D}[pZ]})=L_{p}(E_{2}^{\chi_{D},\chi_{D}[pZ]},\mathbb{N}_{K}\chi_{D}) and SKχD(fD[pZ])=Lp(fD[pZ],KχD)S_{\mathbb{N}_{K}\chi_{D}}^{\flat}(f_{D}^{[pZ]})=L_{p}(f_{D}^{[pZ]},\mathbb{N}_{K}\chi_{D}). From Proposition 5.2, Lp(E2χD,χD[pZ],KχD)Lp(fD[pZ],KχD)(modp)L_{p}(E_{2}^{\chi_{D},\chi_{D}[pZ]},\mathbb{N}_{K}\chi_{D})\equiv L_{p}(f_{D}^{[pZ]},\mathbb{N}_{K}\chi_{D})\pmod{p}. By [17, Theorem 9.10], Lp(fD,KχD)=Ω(A,t)Ξp(fD,KχD)logED(PK)L_{p}(f_{D},\mathbb{N}_{K}\chi_{D})=\Omega(A,t)\Xi_{p}(f_{D},\mathbb{N}_{K}\chi_{D})\log_{E_{D}}(P_{K}), where PKP_{K} is a Heegner point. Now suppose Sχ(E2χD,χD[pZ])0(modp)S_{\chi}^{\flat}(E_{2}^{\chi_{D},\chi_{D}[pZ]})\not\equiv 0\pmod{p}. By Corollary 4.3.1, due to the Heegner hypothesis, none of the Euler factors vanish, and thus Lp(w,KχD)0(modp)Lp(w,KχD)0L_{p}(w,\mathbb{N}_{K}\chi_{D})\not\equiv 0\pmod{p}\implies L_{p}(w,\mathbb{N}_{K}\chi_{D})\neq 0. It follows that logED(PK)0\log_{E_{D}}(P_{K})\neq 0, and hence PKP_{K} is not a torsion point, and it follows that ED/KE_{D}/K has positive rank. A theorem due to Kolyvagin [15] (for example, see [9, Theorem 2.9]) implies that in fact ED/KE_{D}/K has rank exactly 11. ∎

Adopting the notation from [17, Theorem 9.11], for D>0D>0 we have that

SχD(E2χD,χD)\displaystyle S_{\mathbb{N}\chi_{D}}(E_{2}^{\chi_{D},\chi_{D}}) =Lp,α(0,(χD)K),\displaystyle=L_{p,\alpha}(0,(\chi_{D})_{K}),
=Ω(A,t)Ξp(0,(χD)K)g(χD)𝔞𝒞(𝒪K)(χ1K)(𝔞)a=1N1χD1(a)logpga(𝔞(A,t)).\displaystyle=\Omega(A,t)\frac{\Xi_{p}(0,(\chi_{D})_{K})}{g(\chi_{D})}\sum_{\mathfrak{a}\in\mathcal{C}\ell(\mathcal{O}_{K})}(\chi^{-1}\mathbb{N}_{K})(\mathfrak{a})\sum_{a=1}^{N-1}\chi_{D}^{-1}(a)\log_{p}g_{a}(\mathfrak{a}\star(A,t)).

By Theorem 2.13, this sum is equal to p(0,χDχKω)p(1,χD)\mathcal{L}_{p}(0,\chi_{D}\chi_{K}\omega)\mathcal{L}_{p}(1,\chi_{D}), where p\mathcal{L}_{p} is the Katz pp-adic LL-function. By [26, Theorem 5.11], we have that

p(0,χDχKω)p(1,χD)=B1,χDχKp(1,χD)B1,χDχKB1,χDω1(modp),\mathcal{L}_{p}(0,\chi_{D}\chi_{K}\omega)\mathcal{L}_{p}(1,\chi_{D})=-B_{1,\chi_{D}\chi_{K}}\mathcal{L}_{p}(1,\chi_{D})\equiv B_{1,\chi_{D}\chi_{K}}B_{1,\chi_{D}\omega^{-1}}\pmod{p},

so we conclude that

SχD(E2χD,χD)B1,χDχKB1,χDω1(modp).S_{\mathbb{N}\chi_{D}}(E_{2}^{\chi_{D},\chi_{D}})\equiv B_{1,\chi_{D}\chi_{K}}B_{1,\chi_{D}\omega^{-1}}\pmod{p}.

If D<0D<0 then χD\chi_{D} is odd, so

Ω(A,t)Ξp(0,(χD)K)g(χD)𝔞𝒞(𝒪K)(χ1K)(𝔞)a=1N1χD1(a)logpga(𝔞(A,t))=p(0,χDω)p(1,χDχK)\Omega(A,t)\frac{\Xi_{p}(0,(\chi_{D})_{K})}{g(\chi_{D})}\sum_{\mathfrak{a}\in\mathcal{C}\ell(\mathcal{O}_{K})}(\chi^{-1}\mathbb{N}_{K})(\mathfrak{a})\sum_{a=1}^{N-1}\chi_{D}^{-1}(a)\log_{p}g_{a}(\mathfrak{a}\star(A,t))=\mathcal{L}_{p}(0,\chi_{D}\omega)\mathcal{L}_{p}(1,\chi_{D}\chi_{K})

instead. By [26, Theorem 5.11] we have that

p(0,χDω)p(1,χDχK)=B1,χDp(1,χDχK)B1,χDB1,χDχKω1(modp).\mathcal{L}_{p}(0,\chi_{D}\omega)\mathcal{L}_{p}(1,\chi_{D}\chi_{K})=-B_{1,\chi_{D}}\mathcal{L}_{p}(1,\chi_{D}\chi_{K})\equiv B_{1,\chi_{D}}B_{1,\chi_{D}\chi_{K}\omega^{-1}}\pmod{p}.

This implies that for D<0D<0, we have

SχD(E2χD,χD)B1,χDB1,χDχKω1(modp).S_{\mathbb{N}\chi_{D}}(E_{2}^{\chi_{D},\chi_{D}})\equiv B_{1,\chi_{D}}B_{1,\chi_{D}\chi_{K}\omega^{-1}}\pmod{p}.

Hence

SχD(E2χD,χD){B1,χDω1B1,χDχK(modp)D>0,B1,χDχKω1B1,χD(modp)D<0.S_{\mathbb{N}\chi_{D}}(E_{2}^{\chi_{D},\chi_{D}})\equiv\begin{cases}B_{1,\chi_{D}\omega^{-1}}B_{1,\chi_{D}\chi_{K}}\pmod{p}&D>0,\\ B_{1,\chi_{D}\chi_{K}\omega^{-1}}B_{1,\chi_{D}}\pmod{p}&D<0.\\ \end{cases}

For D>0D>0, this turns out to be h(3D)h(DΔK)h_{\mathbb{Q}(\sqrt{-3D})}h_{\mathbb{Q}(\sqrt{D\Delta_{K}})}. For D<0D<0, this turns out to be h(3DDK)h(D)h_{\mathbb{Q}(\sqrt{-3DD_{K}})}h_{\mathbb{Q}(\sqrt{D})}.

Proposition 5.5.
SχD(E2χD,χD){h(3D)h(DΔK)(modp)D>0,h(3DΔK)h(D)(modp)D<0.S_{\mathbb{N}\chi_{D}}(E_{2}^{\chi_{D},\chi_{D}})\equiv\begin{cases}h_{\mathbb{Q}(\sqrt{-3D})}h_{\mathbb{Q}(\sqrt{D\Delta_{K}})}\pmod{p}&D>0,\\ h_{\mathbb{Q}(\sqrt{-3D\Delta_{K}})}h_{\mathbb{Q}(\sqrt{D})}\pmod{p}&D<0.\\ \end{cases}

We now turn to calculating the proportion of ΔK\Delta_{K} such that p=3h(3D)h(DΔK)p=3\nmid h_{\mathbb{Q}(\sqrt{-3D})}h_{\mathbb{Q}(\sqrt{D\Delta_{K}})}; we will address the other case shortly after.

We assume the Heegner hypothesis. This requires that ΔK\Delta_{K} is a quadratic residue modulo all primes dividing cond(ED)\text{cond}(E_{D}), except for 33. Letting N=cond(E)N=\text{cond}(E) and gcd(N,D)=1\gcd(N,D)=1 such that 2ND2\nmid ND, then cond(ED)=ND2\text{cond}(E_{D})=ND^{2}. Furthermore, v3(N)1v_{3}(N)\neq 1 (due to the conditions provided by Nakagawa-Horie in [19, p. 21] or [8, Lemma 2.2]). Recall that Assumption 3.3 denotes the follow conditions on (N,D)(N,D):

  • For all primes >3\ell>3, if v(N)=1v_{\ell}(N)=1, then 2(mod3)\ell\equiv 2\pmod{3},

  • gcd(N,D)=1\gcd(N,D)=1,

  • 2ND2\nmid ND,

  • v3(N)1v_{3}(N)\neq 1.

Theorem 5.6.

For fixed N,DN,D with D>0D>0 satisfying Assumption 3.3 and 3h(3D)3\nmid h_{\mathbb{Q}(\sqrt{-3D})}, the proportion of ΔK\Delta_{K} such that 3h(DΔK)3\nmid h_{\mathbb{Q}(\sqrt{D\Delta_{K}})} is at least 21ω(ND/3v3(ND))2^{-1-\omega(ND/3^{v_{3}(ND)})}.

Proof.

For each 3|ND3\neq\ell|ND, the proportion of ΔK\Delta_{K} which are quadratic residues mod\bmod\ell is +12>12\frac{\ell+1}{2\ell}>\frac{1}{2}. The number of such primes is ω(ND/3v3(ND))\omega(ND/3^{v_{3}(ND)}), so the proportion of such ΔK\Delta_{K} is greater than 12ω(ND/3v3(ND))\frac{1}{2^{\omega(ND/3^{v_{3}(ND)})}}. Of this set XX, [8, Lemma 2.2] implies that

|{ΔKX and 3h(DΔK)}||X|12.\frac{|\{\Delta_{K}\in X\text{ and }3\nmid h_{\mathbb{Q}(\sqrt{D\Delta_{K}})}\}|}{|X|}\geq\frac{1}{2}.

Hence for a fixed DD, the proportion of ΔK\Delta_{K} which are quadratic residues modulo all |D\ell|D and 3h(DΔK)3\nmid h_{\mathbb{Q}(\sqrt{D\Delta_{K}})} is at least 21ω(ND/3v3(ND))2^{-1-\omega(ND/3^{v_{3}(ND)})}. ∎

We have the immediate

Corollary 5.6.1.

For fixed N=cond(E)N=\text{cond}(E) and D>0D>0 satisfying Assumption 3.3 and 3h(3D)3\nmid h_{\mathbb{Q}(\sqrt{-3D})}, the proportion of imaginary quadratic fields KK which admit a quadratic twist of EE by the fixed DD is positive.

We also immediately obtain information about the rank of EDE_{D} over KK.

Corollary 5.6.2.

For fixed N,DN,D with D>0D>0 satisfying Assumption 3.3 and 3h(3D)3\nmid h_{\mathbb{Q}(\sqrt{-3D})}, the proportion of imaginary quadratic fields KK such that ED/KE_{D}/K has rank 11 is at least 21ω(ND/3v3(ND))2^{-1-\omega(ND/3^{v_{3}(ND)})}.

Proof.

Theorem 5.6 implies that the proportion of KK with 3h(3D)h(DΔK)3\nmid h_{\mathbb{Q}(\sqrt{-3D})}h_{\mathbb{Q}(\sqrt{D\Delta_{K}})} is at least 21ω(ND/3v3(ND))2^{-1-\omega(ND/3^{v_{3}(ND)})}. Combining with the fact that

SχD(E2χD,χD)h(3D)h(DΔK)(mod3),S_{\mathbb{N}_{\chi_{D}}}(E_{2}^{\chi_{D},\chi_{D}})\equiv h_{\mathbb{Q}(\sqrt{-3D})}h_{\mathbb{Q}(\sqrt{D\Delta_{K}})}\pmod{3},

we have that 3SχD(E2χD,χD)3\nmid S_{\mathbb{N}_{\chi_{D}}}(E_{2}^{\chi_{D},\chi_{D}}). Now applying Proposition 5.4, we find that every such KK also satisfies that ED/KE_{D}/K has rank 11. ∎

We also address the D<0D<0 case. Once again, let N=cond(E)N=\text{cond}(E) and gcd(N,D)=1\gcd(N,D)=1 with 2ND2\nmid ND, and v3(N)1v_{3}(N)\neq 1.

Theorem 5.7.

For fixed N,DN,D with D<0D<0 satisfying Assumption 3.3 and 3h(D)3\nmid h_{\mathbb{Q}(\sqrt{D})}, the proportion of ΔK\Delta_{K} such that 3h(3DΔK)3\nmid h_{\mathbb{Q}(\sqrt{-3D\Delta_{K}})} is at least 21ω(ND/3v3(ND))2^{-1-\omega(ND/3^{v_{3}(ND)})}.

Proof.

For each 3|ND3\neq\ell|ND, the proportion of DKD_{K} which are quadratic residues mod\bmod{\ell} is +12>12\frac{\ell+1}{2\ell}>\frac{1}{2}. The number of such primes is ω(ND/3v3(ND))\omega(ND/3^{v_{3}(ND)}), so the proportion of such ΔK\Delta_{K} is greater than 2ω(ND/3v3(ND))2^{-\omega(ND/3^{v_{3}(ND)})}. This set XX is given by a system of congruence conditions, modulo all 3|ND3\neq\ell|ND. Of this set XX, [8, Lemma 2.2] implies that

|ΔKX and 3h(3DΔK)||X|12.\frac{\left|\Delta_{K}\in X\text{ and }3\nmid h_{\mathbb{Q}(\sqrt{-3D\Delta_{K}})}\right|}{|X|}\geq\frac{1}{2}.

Hence for fixed DD, the proportion of ΔK\Delta_{K} which satisfy the Heegner hypothesis and eh(3DΔK)e\nmid h_{\mathbb{Q}(\sqrt{-3D\Delta_{K}})} is at least 21ω(ND/3v3(ND))2^{-1-\omega(ND/3^{v_{3}(ND)})}. ∎

Once again, we find two corollaries.

Corollary 5.7.1.

For fixed N=cond(E)N=\text{cond}(E) and D<0D<0 satisfying Assumption 3.3 and 3h(D)3\nmid h_{\mathbb{Q}(\sqrt{D})}, the proportion of imaginary quadratic fields KK which admit a quadratic twist of EE by the fixed DD is positive.

Corollary 5.7.2.

For fixed N,DN,D with D<0D<0 satisfying Assumption 3.3 and 3h(D)3\nmid h_{\mathbb{Q}(\sqrt{D})}, the proportion of imaginary quadratic fields KK such that ED/KE_{D}/K has rank 11 is at least 21ω(ND/3v3(ND))2^{-1-\omega(ND/3^{v_{3}(ND)})}.

Proof.

Theorem 5.7 implies that the proportion of KK with 3h(D)h(3DΔK)3\nmid h_{\mathbb{Q}(\sqrt{D})}h_{\mathbb{Q}(\sqrt{-3D\Delta_{K}})} is at least 21ω(ND/3v3(ND))2^{-1-\omega(ND/3^{v_{3}(ND)})}. Combining with the fact that

SχD(E2χD,χD)h(D)h(3DΔK)(mod3),S_{\mathbb{N}_{\chi_{D}}}(E_{2}^{\chi_{D},\chi_{D}})\equiv h_{\mathbb{Q}(\sqrt{D})}h_{\mathbb{Q}(\sqrt{-3D\Delta_{K}})}\pmod{3},

we have that 3SχD(E2χD,χD)3\nmid S_{\mathbb{N}_{\chi_{D}}}(E_{2}^{\chi_{D},\chi_{D}}). Now applying Proposition 5.4, we find that every such KK also satisfies that ED/KE_{D}/K has rank 11. ∎

5.2. Varying DD

In this paper we will usually fix DD and vary KK. Let us now fix NN and KK and vary DD. The first result we have is considering the proportion of DD which satisfy the Heegner hypothesis.

Theorem 5.8.

For fixed N,KN,K, the number of 0<D<X0<D<X satisfying the Heegner hypothesis is asymptotic to X/(logX)1/2X/(\log X)^{1/2}.

Proof.

For each 3|ND3\neq\ell|ND, we need ΔK\Delta_{K} to be a quadratic residue mod\bmod \ell. Let ΔK=2ei=1Mpi\Delta_{K}=-2^{e}\prod_{i=1}^{M}p_{i} where the pip_{i} are distinct odd primes and e{0,2}e\in\{0,2\}. We need

(1)i(pi)=1,\left(\frac{-1}{\ell}\right)\prod_{i}\left(\frac{p_{i}}{\ell}\right)=1,

since the factor of 22 is always a square. Quadratic reciprocity implies that (pi)=(1)pi1212(pi)\left(\frac{p_{i}}{\ell}\right)=(-1)^{\frac{p_{i}-1}{2}\cdot\frac{\ell-1}{2}}\left(\frac{\ell}{p_{i}}\right) for each pip_{i}. In particular, when \ell is fixed, the sign depends only on pip_{i}, and is also fixed. Then (ΔK)=1\left(\frac{\Delta_{K}}{\ell}\right)=1 is equivalent to the condition that an even number of the (1)pi1212(pi)(-1)^{\frac{p_{i}-1}{2}\cdot\frac{\ell-1}{2}}\left(\frac{\ell}{p_{i}}\right) are 1-1, and thus in the product

(ΔK)=i(1)pi1212(pi),\left(\frac{\Delta_{K}}{\ell}\right)=\prod_{i}(-1)^{\frac{p_{i}-1}{2}\cdot\frac{\ell-1}{2}}\left(\frac{\ell}{p_{i}}\right),

it suffices to allow anything in the first M1M-1 indices, and the last index is determined in order to yield a product of 11. Note that if =pi\ell=p_{i} for any ii, then it is always a quadratic residue. Thus at most pi12pi\frac{p_{i}-1}{2p_{i}} of \ell have ΔK\Delta_{K} not a quadratic residue (mod)\pmod{\ell}, and it follows that the proportion of \ell with ΔK\Delta_{K} a quadratic residue (mod)\pmod{\ell} is at least 12\frac{1}{2}. Even stronger, ΔK\Delta_{K} is a quadratic residue (mod)\pmod{\ell} whenever ¯S/ΔK\bar{\ell}\in S\subset\mathbb{Z}/\Delta_{K}^{\prime}\mathbb{Z} where ΔK=ΔK\Delta_{K}^{\prime}=\Delta_{K} or 4ΔK4\Delta_{K}, and |S|>|ΔK|/2|S|>|\Delta_{K}^{\prime}|/2. It follows that DD can only be constructed from such primes.

From the Wiener-Ikehara Tauberian theorem [21, Theorem 2.4], we find that the proportion of D<XD<X such that DD is constructed from this set of primes is asymptotic to X/(logX)1/2X/(\log X)^{1/2}.

Suppose we let DD vary and count the proportion of pairs (D,K)(D,K) (equivalently pairs (D,ΔK)(D,\Delta_{K})) such that ED/KE_{D}/K has rank 11. Corollary 5.6.1 and Corollary 5.6.2 imply that the proportion of ΔK\Delta_{K} for fixed DD depends only on DD, and in particular, on the number of prime factors of DD. Thus it suffices to consider only positive DD, as the sign does not matter. Consider the interval 0<D<X0<D<X. Then we seek to measure

F(X)=1XD=1X2ω(D).F(X)=\frac{1}{X}\sum_{D=1}^{X}2^{-\omega(D)}.

By summing over values of ω(n)\omega(n) instead, we have

F(X)=1Xn=0logX2n#{DX|ω(D)=n}.F(X)=\frac{1}{X}\sum_{n=0}^{\log X}2^{-n}\cdot\#\{D\leq X|\omega(D)=n\}.

The final proportion will be at least 12F(X)\frac{1}{2}F(X). We will now find asymptotic bounds for F(X)F(X).

By Erdös-Kac [10], for 1nX1\leq n\leq X, ω(n)\omega(n) follows a Gaussian distribution with σ=loglogX\sigma=\sqrt{\log\log X} and μ=loglogX\mu=\log\log X. Therefore we may assume that for sufficiently large XX, ω\omega may be approximated by a continuous distribution; we will assume that this continuous distribution is sufficiently accurate and measure

F(X)T(X)=0X(YloglogX)/loglogX2x1σ2πe12(xμσ)2𝑑x,F(X)\sim T(X)=\int_{0}^{X\cdot(Y-\log\log X)/\sqrt{\log\log X}}2^{-x}\cdot\frac{1}{\sigma\sqrt{2\pi}}e^{-\frac{1}{2}\left(\frac{x-\mu}{\sigma}\right)^{2}}\,dx,

where Y=max{ω(D)|DX}Y=\max\{\omega(D)|D\leq X\}. Since for large XX, it’s clear that the upper bound exceeds 2X2X, we may take

S(X)=02X2x1σ2πe12(xμσ)2𝑑x<T(X).S(X)=\int_{0}^{2X}2^{-x}\cdot\frac{1}{\sigma\sqrt{2\pi}}e^{-\frac{1}{2}\left(\frac{x-\mu}{\sigma}\right)^{2}}\,dx<T(X).

Substituting y=2xy=2^{-x}, we transform the following integral, which is S(X)S(X) but extended from -\infty to \infty:

2x1σ2πe12(xμσ)2𝑑x=e(log2)μ+(log2)2σ22>S(X).\int_{-\infty}^{\infty}2^{-x}\cdot\frac{1}{\sigma\sqrt{2\pi}}e^{-\frac{1}{2}\left(\frac{x-\mu}{\sigma}\right)^{2}}\,dx=e^{-(\log 2)\mu+\frac{(\log 2)^{2}\sigma^{2}}{2}}>S(X).

Let κ=(log2)22+log20.45\kappa=-\frac{(\log 2)^{2}}{2}+\log 2\approx 0.45. Then substituting μ=σ2=loglogX\mu=\sigma^{2}=\log\log X, we have

S(X)=1(logX)κA(X)B(X),S(X)=\frac{1}{(\log X)^{\kappa}}-A(X)-B(X),

where

A(X)\displaystyle A(X) =02x1σ2πe12(xμσ)2𝑑x,\displaystyle=\int_{-\infty}^{0}2^{-x}\cdot\frac{1}{\sigma\sqrt{2\pi}}e^{-\frac{1}{2}\left(\frac{x-\mu}{\sigma}\right)^{2}}\,dx,
B(X)\displaystyle B(X) =2X2x1σ2πe12(xμσ)2𝑑x.\displaystyle=\int_{2X}^{\infty}2^{-x}\cdot\frac{1}{\sigma\sqrt{2\pi}}e^{-\frac{1}{2}\left(\frac{x-\mu}{\sigma}\right)^{2}}\,dx.

We will now bound A(X)A(X) and B(X)B(X) (both of which are positive values, since the integrand is strictly positive).

Lemma 5.9.

A(X)<1(logX)1/2XA(X)<\frac{1}{(\log X)^{1/2}\sqrt{X}}.

Proof.

Take

A(X)=02x1σ2πe12(x+μσ)2𝑑x.A(X)=\int_{0}^{\infty}2^{x}\cdot\frac{1}{\sigma\sqrt{2\pi}}e^{-\frac{1}{2}\left(\frac{x+\mu}{\sigma}\right)^{2}}\,dx.

Let N=μ=σ2=loglogXN=\mu=\sigma^{2}=\log\log X. Then we can write

A(X)=1N2π0exp(xlog2x2+2xN+N22N)𝑑x.A(X)=\frac{1}{N\sqrt{2\pi}}\int_{0}^{\infty}\text{exp}\left(x\log 2-\frac{x^{2}+2xN+N^{2}}{2N}\right)\,dx.

Now consider that

xlog2x2+2xN+N22N=x22NN2x(1log2)N<x22NN2.x\log 2-\frac{x^{2}+2xN+N^{2}}{2N}=-\frac{x^{2}}{2N}-\frac{N}{2}-\frac{x(1-\log 2)}{N}<-\frac{x^{2}}{2N}-\frac{N}{2}.

As a result, we have that A(X)<eN/2N2πI(X)A(X)<\frac{e^{-N/2}}{N\sqrt{2\pi}}I(X), where I(X)=0ex22N𝑑xI(X)=\int_{0}^{\infty}e^{-\frac{x^{2}}{2N}}\,dx. But I(X)I(X) can be solved using the well-known Poisson trick, which yields that I(X)2=2πNI(X)=2πNI(X)^{2}=2\pi N\implies I(X)=\sqrt{2\pi N}, so we have that A(X)<eN/2/NA(X)<e^{-N/2}/\sqrt{N}. ∎

With B(X)B(X), we can calculate it almost exactly.

Lemma 5.10.

B(X)<e2X2/loglogXB(X)<e^{-2X^{2}/\log\log X}.

Proof.

We have

B(X)=1N2π2Xexlog2(xN)22N𝑑x.B(X)=\frac{1}{N\sqrt{2\pi}}\int_{2X}^{\infty}e^{-x\log 2-\frac{(x-N)^{2}}{2N}}\,dx.

The exponent rearranges to

N2(1(1log2)2)(xN(1log2))22NB(X)=eN2(1(1log2)2)N2πJ(X),-\frac{N}{2}(1-(1-\log 2)^{2})-\frac{(x-N(1-\log 2))^{2}}{2N}\implies B(X)=\frac{e^{-\frac{N}{2}(1-(1-\log 2)^{2})}}{N\sqrt{2\pi}}J(X),

where

J(X)=2Xe(x(1log2)N)22N𝑑x=2X(1log2)Nex2/2N𝑑x.J(X)=\int_{2X}^{\infty}e^{-\frac{(x-(1-\log 2)N)^{2}}{2N}}\,dx=\int_{2X-(1-\log 2)N}^{\infty}e^{-x^{2}/2N}\,dx.

Using Poisson’s trick once again, we find

J(X)2<2π2X(1log2)Nrer2/2N𝑑r=2πNe(2X(1log2)N)22N.J(X)^{2}<2\pi\int_{2X-(1-\log 2)N}^{\infty}re^{-r^{2}/2N}\,dr=2\pi Ne^{-\frac{(2X-(1-\log 2)N)^{2}}{2N}}.

As a result, we conclude that

logB(X)<N2(1(1log2)2)(2X(1log2)N)22N=2X2N+2X(1log2)N2<2X2N.\log B(X)<-\frac{N}{2}(1-(1-\log 2)^{2})-\frac{(2X-(1-\log 2)N)^{2}}{2N}=-\frac{2X^{2}}{N}+2X(1-\log 2)-\frac{N}{2}<-\frac{2X^{2}}{N}.

Putting the above two lemmas together, we conclude that

1(logX)κ1(logX)1/2X1e2X2/loglogX<S(X)<1(logX)κ,\frac{1}{(\log X)^{\kappa}}-\frac{1}{(\log X)^{1/2}\sqrt{X}}-\frac{1}{e^{2X^{2}/\log\log X}}<S(X)<\frac{1}{(\log X)^{\kappa}},

which implies that

Theorem 5.11.

Assuming that ω\omega is approximated by a (continuous) Gaussian distribution sufficiently well, in the set 𝒟X={(D,ΔK)||D|<X,K an imaginary quadratic field}\mathcal{D}_{X}=\{(D,\Delta_{K})|\hskip 2.84526pt|D|<X,K\text{ an imaginary quadratic field}\}, the proportion P(X)P(X) of 𝒟X\mathcal{D}_{X} (for X0X\gg 0) which yield a quadratic twist with rank 11 over KK satisfies

12(1(logX)κ1(logX)1/2X1e2X2/loglogX)<P(X)<12(logX)κ,\frac{1}{2}\left(\frac{1}{(\log X)^{\kappa}}-\frac{1}{(\log X)^{1/2}\sqrt{X}}-\frac{1}{e^{2X^{2}/\log\log X}}\right)<P(X)<\frac{1}{2(\log X)^{\kappa}},

where κ=log2(log2)220.45\kappa=\log 2-\frac{(\log 2)^{2}}{2}\approx 0.45.

6. Higher twists

6.1. Cubic twists

Let ff be the modular form associated to EE, an elliptic curve over L=(3)L=\mathbb{Q}(\sqrt{-3}). We have that f=θψf=\theta_{\psi} (of weight 22) for some Hecke character ψ\psi of type (1,0)(1,0), where

θψ=𝔞𝒪Lψ(𝔞)qN(𝔞)θψ[[q]].\theta_{\psi}=\sum_{\mathfrak{a}\subset\mathcal{O}_{L}}\psi(\mathfrak{a})q^{N(\mathfrak{a})}\in\theta_{\psi}\in\mathbb{Z}[[q]].

Let χd:Gal(L(d3)/L)ζ3\chi_{d}:\text{Gal}(L(\sqrt[3]{d})/L)\rightarrow\langle\zeta_{3}\rangle be the associated cubic twist character, where ζ3=1+32\zeta_{3}=\frac{-1+\sqrt{-3}}{2} is a primitive third root of unity. Let fdf_{d} be the modular form associated to Ed,3E_{d,3}, the cubic twist of EE by dd, such that

fd=θψχd=𝔞𝒪Lψχd(𝔞)qN(𝔞)[[q]].f_{d}=\theta_{\psi\chi_{d}}=\sum_{\mathfrak{a}\subset\mathcal{O}_{L}}\psi\chi_{d}(\mathfrak{a})q^{N(\mathfrak{a})}\in\mathbb{Z}[[q]].
Proposition 6.1.

The modular forms ff and fdf_{d} are equivalent modulo 33 at all coefficients except those which are not relatively prime to NdNd.

Proof.

For all nn\in\mathbb{Z} with gcd(n,cond(χd))=1\gcd(n,\text{cond}(\chi_{d}))=1, we have χd(n)1(modζ31)\chi_{d}(n)\equiv 1\pmod{\zeta_{3}-1}. As a result, for all nn coprime to cond(χd)cond(E)=Nd\text{cond}(\chi_{d})\cdot\text{cond}(E)=Nd, we have

ψ(n)ψχd(n)mod(ζ31)𝒪L[[q]],\psi(n)\equiv\psi\chi_{d}(n)\bmod{(\zeta_{3}-1)\mathcal{O}_{L}[[q]]},

and thus we have ffdmod(ζ31)𝒪L[[q]]f\equiv f_{d}\bmod{(\zeta_{3}-1)\mathcal{O}_{L}[[q]]} except at the coefficients of qnq^{n} for gcd(n,Nd)1\gcd(n,Nd)\neq 1. Since (ζ31)𝒪L=(3)(\zeta_{3}-1)\mathcal{O}_{L}\cap\mathbb{Z}=(3)\mathbb{Z} and f,fd[[q]]f,f_{d}\in\mathbb{Z}[[q]], it follows that ffd[[q]]f-f_{d}\in\mathbb{Z}[[q]] and ffdG(q)(ζ31)𝒪L[[q]]f-f_{d}-G(q)\in(\zeta_{3}-1)\mathcal{O}_{L}[[q]] for some G(q)[[q]]G(q)\in\mathbb{Z}[[q]] with G(q)=gcd(n,Nd)>1gnqnG(q)=\sum_{\gcd(n,Nd)>1}g_{n}q^{n}. Therefore ffdG(q)[[q]]f-f_{d}-G(q)\in\mathbb{Z}[[q]], and hence

ffdG(q)(ζ31)𝒪L[[q]][[q]]=(3)[[q]].f-f_{d}-G(q)\in(\zeta_{3}-1)\mathcal{O}_{L}[[q]]\cap\mathbb{Z}[[q]]=(3)\mathbb{Z}[[q]].

As a result, ffdG(q)(mod3)f-f_{d}\equiv G(q)\pmod{3}, and therefore ffd(mod3)f\equiv f_{d}\pmod{3} except at coefficients of qnq^{n} for nn not relatively prime to NdNd. ∎

Proposition 6.2.

The Galois representations of ff and fdf_{d} are isomorphicmod3\mod 3.

Proof.

Let ρE:Gal(¯/)GL(limE[n])pGL2(p)\rho_{E}:\text{Gal}(\bar{\mathbb{Q}}/\mathbb{Q})\rightarrow GL(\varprojlim E[n])\cong\prod_{p}GL_{2}(\mathbb{Z}_{p}) be the Galois representation of EE. Let ρd\rho_{d} be the Galois representation of Ed,3E_{d,3}, the cubic twist of EE by dd. Let NN be the conductor of EE, so that NdNd is the conductor of Ed,3E_{d,3}. Then Neron-Ogg-Shafarevich implies that ρ\rho and ρd\rho_{d} are unramified outside of NN and NdNd, respectively. As a result, the Galois representations factor through Gal((N)/)\text{Gal}(\mathbb{Q}^{(N)}/\mathbb{Q}) and Gal((Nd)/\text{Gal}(\mathbb{Q}^{(Nd)}/\mathbb{Q}, respectively, where (n)\mathbb{Q}^{(n)} is the maximal unramified extension of \mathbb{Q} outside of nn. Now for all primes Nd\ell\nmid Nd, the Artin map gives a Frobenius element Frob\text{Frob}_{\ell} such that tr Frob,E=[q]f\text{tr }\text{Frob}_{\ell,E}=[q^{\ell}]f and tr Frob,Ed,3[q]fd\text{tr }\text{Frob}_{\ell,E_{d,3}}[q^{\ell}]f_{d}. The prior discussion shows that

ffd(mod3)tr Frob,Etr Frob,Ed,3(mod3).f\equiv f_{d}\pmod{3}\implies\text{tr }\text{Frob}_{\ell,E}\equiv\text{tr }\text{Frob}_{\ell,E_{d,3}}\pmod{3}.

Furthermore, the Frobenius elements always satisfy detFrob=\det\text{Frob}_{\ell}=\ell. The Brauer-Nesbitt theorem applied to ρ\rho implies that ρ\rho and ρd\rho_{d} are characterized (up to isomorphism) by their characteristic polynomials, and thus by the trace and determinant functions. We showed that ρ\rho and ρd\rho_{d} agree on the trace and determinant functions modulo 33 for all Frobenius elements Frob\text{Frob}_{\ell} with Nd\ell\nmid Nd. By the Cebotarev density function, the Frobenius elements have density 11 in the Galois groups, and therefore all but finitely many of the Frobenius elements are dense in the Galois group. Since trace and determinant are continuous functions, this implies that ρ\rho and ρd\rho_{d} agree modulo 33 on trace and determinant on the entire Galois group, and thus they agree modulo 33 everywhere (by Brauer-Nesbitt). As a result, we find that ρρd(mod3)\rho\equiv\rho_{d}\pmod{3}. ∎

Proposition 6.3.

If SK(f[Nd])0(mod3)S_{\mathbb{N}_{K}}^{\flat}(f^{[Nd]})\not\equiv 0\pmod{3}, then Ed,3/KE_{d,3}/K has rank 11.

Proof.

By Proposition 5.2, we have that

SK(f[Nd])SK(fd[Nd])(mod3).S_{\mathbb{N}_{K}}^{\flat}(f^{[Nd]})\equiv S_{\mathbb{N}_{K}}^{\flat}(f_{d}^{[Nd]})\pmod{3}.

Following [17, Definition 8.8], we have SK(f[Nd])=Lp(f[Nd],K)S_{\mathbb{N}_{K}}^{\flat}(f^{[Nd]})=L_{p}(f^{[Nd]},\mathbb{N}_{K}) and SK(fd[Nd])=Lp(fd[Nd],K)S_{\mathbb{N}_{K}}^{\flat}(f_{d}^{[Nd]})=L_{p}(f_{d}^{[Nd]},\mathbb{N}_{K}). From [17, Theorem 9.10], we have

Lp(fd[Nd],K)=ω(A,t)Ξp(fd[Nd],K)logEd,3(PK),L_{p}(f_{d}^{[Nd]},\mathbb{N}_{K})=\omega(A,t)\Xi_{p}(f_{d}^{[Nd]},\mathbb{N}_{K})\log_{E_{d,3}}(P_{K}),

where PKP_{K} is a Heegner point. As a result, if SK(f[Nd])0(mod3)S_{\mathbb{N}_{K}}^{\flat}(f^{[Nd]})\not\equiv 0\pmod{3}, then this implies that

SK(fd[Nd])SK(f[Nd])0(mod3)SK(fd[Nd])0logEd,3(PK)0.S_{\mathbb{N}_{K}}^{\flat}(f_{d}^{[Nd]})\equiv S_{\mathbb{N}_{K}}^{\flat}(f^{[Nd]})\not\equiv 0\pmod{3}\implies S_{\mathbb{N}_{K}}^{\flat}(f_{d}^{[Nd]})\neq 0\implies\log_{E_{d,3}}(P_{K})\neq 0.

Now applying Proposition 5.4, we find that Ed,3/KE_{d,3}/K has rank exactly 11. ∎

We assume the Galois representation ρE\rho_{E} modulo 33 is reducible. By Proposition 4.2, we have that ρEχMχMχ\rho_{E}\cong\chi_{M}\oplus\chi_{M}\chi up to semisimplification.

6.2. Sextic twists

Consider the family of elliptic curves y2=x3+cy^{2}=x^{3}+c over \mathbb{Q} for cc\in\mathbb{Q} up to isomorphism; denote this by 𝒞c\mathcal{C}_{c}. This family of elliptic curves has jj-invariant j=0j=0. The sextic twist by DD, g6,D(𝒞c)g_{6,D}(\mathcal{C}_{c}), is the elliptic curve given by

y2=x3+cD(D3y)2=(D2x)3+cD7y2=x3+cD7,y^{2}=x^{3}+cD\cong_{\mathbb{Q}}(D^{3}y)^{2}=(D^{2}x)^{3}+cD^{7}\cong_{\mathbb{Q}}y^{2}=x^{3}+cD^{7},

where E1FE2E_{1}\cong_{F}E_{2} denotes that E1E_{1} is isomorphic to E2E_{2} over FF. Thus the sextic twist by DD is a function

g6,D:𝒞c𝒞cD7.g_{6,D}:\mathcal{C}_{c}\mapsto\mathcal{C}_{cD^{7}}.

The quadratic twist by DD on 𝒞c\mathcal{C}_{c}, denoted by g2,D(𝒞c)g_{2,D}(\mathcal{C}_{c}), is the curve

Dy2=x3+c(D2y)2=(Dx)3+cD3y2=x3+cD3,Dy^{2}=x^{3}+c\cong_{\mathbb{Q}}(D^{2}y)^{2}=(Dx)^{3}+cD^{3}\cong_{\mathbb{Q}}y^{2}=x^{3}+cD^{3},

so

g2,D(𝒞c)=𝒞cD3.g_{2,D}(\mathcal{C}_{c})=\mathcal{C}_{cD^{3}}.

The cubic twist by DD on 𝒞c\mathcal{C}_{c}, denoted by g3,D(𝒞c)g_{3,D}(\mathcal{C}_{c}), is the curve

y2=Dx3+c(Dy)2=(Dx)3+cD2y2=x3+cD2,y^{2}=Dx^{3}+c\cong_{\mathbb{Q}}(Dy)^{2}=(Dx)^{3}+cD^{2}\cong_{\mathbb{Q}}y^{2}=x^{3}+cD^{2},

so

g3,D(𝒞c)=𝒞cD2.g_{3,D}(\mathcal{C}_{c})=\mathcal{C}_{cD^{2}}.

We easily check that g6,D(𝒞c)=𝒞cD7=g2,D(𝒞cD4)=g2,D(g3,D2(𝒞c))=g23,Dg3,Dg3,D(𝒞c)g_{6,D}(\mathcal{C}_{c})=\mathcal{C}_{cD^{7}}=g_{2,D}(\mathcal{C}_{cD^{4}})=g_{2,D}(g_{3,D^{2}}(\mathcal{C}_{c}))=g_{23,D}\circ g_{3,D}\circ g_{3,D}(\mathcal{C}_{c}).

As a result, we have that

g6,D=g2,Dg3,Dg3,D=g3,D2g2,D,g_{6,D}=g_{2,D}\circ g_{3,D}\circ g_{3,D}=g_{3,D^{2}}\circ g_{2,D},

and it’s clear that these functions commute.

Proposition 6.4.

The family of curves 𝒞c\mathcal{C}_{c} are exactly the elliptic curves which admit cubic twists.

Proof.

This family is precisely the family of elliptic curves with Weierstrass form y2=x3+ax+by^{2}=x^{3}+ax+b with a=0a=0; in particular, this is exactly the family of elliptic curves with jj-invariant 0, since j(E)=1728(4a)3Δj(E)=-1728\frac{(4a)^{3}}{\Delta}, where Δ(E)=16(4a3+27b2)\Delta(E)=-16(4a^{3}+27b^{2}) (see [23, p. 45]).

On the other hand, an elliptic curve EE admits a cubic twist iff it has CM by (3)=(ζ3)\mathbb{Q}(\sqrt{-3})=\mathbb{Q}(\zeta_{3}). Since the underlying field has characteristic 0, by [23, cor. III.10.2], this is equivalent to Aut(E)=μ6=ζ6j(E)=0\text{Aut}(E)=\mu_{6}=\langle\zeta_{6}\rangle\iff j(E)=0. ∎

Due to this, the family 𝒞c\mathcal{C}_{c} are the only elliptic curves which interest us.

Proposition 6.5.

The Galois representation of any of the curves 𝒞c\mathcal{C}_{c} is residually reducible modulo 33.

Proof.

Since this family of curves admits cubic twists, they have CM by L=(3)L=\mathbb{Q}(\sqrt{-3}). The prime (3)(3) is ramified in L/L/\mathbb{Q}, so let (3)𝒪L(\sqrt{-3})\subset\mathcal{O}_{L} be the prime lying over (3)(3). Then E[(3)]{xE|[a]x=0a(3)}={xE|[3]x=0}E[(\sqrt{-3})]\coloneqq\{x\in E|[a]x=0\forall a\in(\sqrt{-3})\}=\{x\in E|[\sqrt{-3}]x=0\} is a group of order 33. This group is defined over the Hilbert class field of LL, which is LL since the class number of LL is 11. Now, the group Gal(L/)\text{Gal}(L/\mathbb{Q}) is generated by σ\sigma, the automorphism given by complex conjugation. Since σ((3))=(3)\sigma((\sqrt{-3}))=(\sqrt{-3}), it follows that E[(3)]E[(\sqrt{-3})] is defined over L=L\cap\mathbb{R}=\mathbb{Q}. As a result, E[(3)]E[3]E[(\sqrt{-3})]\subsetneq E[3] is a subgroup preserved by Gal(¯/)\text{Gal}(\overline{\mathbb{Q}}/\mathbb{Q}), and thus ρ𝒞c\rho_{\mathcal{C}_{c}} is residually reducible modulo 33. ∎

Let N=cond(𝒞c)N=\text{cond}(\mathcal{C}_{c}) and let DD be some positive integer satisfying Assumption 3.3.

Theorem 6.6.

For a fixed cc (and thus NN) and D>0D>0 satisfying Assumption 3.3 and 3h(3D)3\nmid h_{\mathbb{Q}(\sqrt{-3D})}, the proportion of imaginary quadratic fields KK such that g6,D(𝒞c)/Kg_{6,D}(\mathcal{C}_{c})/K has rank 11 is at least 21ω(ND/3v3(ND))2^{-1-\omega\left(ND/3^{v_{3}(ND)}\right)}.

Proof.

We have cond(g6,D(𝒞c))=NDt\text{cond}(g_{6,D}(\mathcal{C}_{c}))=ND^{t} for some nonnegative integer tt. Since cond(g2,D(𝒞c))=ND2\text{cond}(g_{2,D}(\mathcal{C}_{c}))=ND^{2}, it follows that the set of primes dividing cond(g6,D(𝒞c))\text{cond}(g_{6,D}(\mathcal{C}_{c})) is a subset of the set of primes dividing cond(g2,D(𝒞c))\text{cond}(g_{2,D}(\mathcal{C}_{c})). As a result, the subsequent cubic twist by D2D^{2} yields an elliptic curve whose conductor does not have any new primes dividing it (compared to ND2ND^{2}), and therefore does not require any more \ell-depletions.

As a result, any KK which admits a quadratic twist of 𝒞c\mathcal{C}_{c} by DD will also admit a cubic twist by D2D^{2}. This occurs when ΔK\Delta_{K} is a quadratic residue modulo all primes |ND\ell|ND, and by Theorem 5.6, occurs for at least (1/2)1+ω(ND/3v3(ND))(1/2)^{1+\omega\left(ND/3^{v_{3}(ND)}\right)} of the ΔK\Delta_{K}. Now applying Proposition 5.4, we conclude that every such KK also satisfies the property that g6,D(𝒞c)/Kg_{6,D}(\mathcal{C}_{c})/K has rank 11. ∎

In particular, since every 𝒞c\mathcal{C}_{c} is isomorphic to the sextic twist of 𝒞1\mathcal{C}_{1} by cc (over a sufficient KK), it is of particular interest to study 𝒞1y2=x3+1\mathcal{C}_{1}\coloneqq y^{2}=x^{3}+1. Thus specializing Theorem 6.6, we have

Corollary 6.6.1.

For fixed D>0D>0 with 3D3\nmid D and 3h(3D)3\nmid h_{\mathbb{Q}(\sqrt{-3D})}, the proportion of imaginary quadratic fields KK such that g6,D(𝒞1)/Kg_{6,D}(\mathcal{C}_{1})/K has rank 11 is at least 21ω(D)2^{-1-\omega(D)}.

Proof.

By applying Theorem 6.6 with N=cond(𝒞1)=27N=\text{cond}(\mathcal{C}_{1})=27, the result follows. ∎

Addressing the D<0D<0 case, we have the analogous results.

Theorem 6.7.

For a fixed cc (and thus NN) and D<0D<0 satisfying Assumption 3.3 and 3h(D)3\nmid h_{\mathbb{Q}(\sqrt{D})}, the proportion of imaginary quadratic fields KK such that g6,D(𝒞c)/Kg_{6,D}(\mathcal{C}_{c})/K has rank 11 is at least 21ω(ND/3v3(ND))2^{-1-\omega\left(ND/3^{v_{3}(ND)}\right)}.

Proof.

We have cond(g6,D(𝒞c))=±NDt\text{cond}(g_{6,D}(\mathcal{C}_{c}))=\pm ND^{t} for some nonnegative integer tt. Since cond(g2,D(𝒞c))=ND2\text{cond}(g_{2,D}(\mathcal{C}_{c}))=ND^{2}, it follows that the set of primes dividing cond(g6,D(𝒞c))\text{cond}(g_{6,D}(\mathcal{C}_{c})) is a subset of the set of primes dividing cond(g2,D(𝒞c))\text{cond}(g_{2,D}(\mathcal{C}_{c})). As a result, the subsequent cubic twist by D2D^{2} yields an elliptic curve whose conductor does not have any new primes dividing it (compared to ND2ND^{2}), and therefore does not require any more \ell-depletions.

As a result, any KK which admits a quadratic twist of 𝒞c\mathcal{C}_{c} by DD will also admit a cubic twist by D2D^{2}. This occurs when ΔK\Delta_{K} is a quadratic residue modulo all primes |ND\ell|ND, and by Theorem 5.7, occurs for at least 21ω(ND/3v3(ND))2^{-1-\omega\left(ND/3^{v_{3}(ND)}\right)} of the ΔK\Delta_{K}. Now applying Proposition 5.4, we conclude that every such KK also satisfies the property that g6,D(𝒞c)/Kg_{6,D}(\mathcal{C}_{c})/K has rank 11. ∎

Once again specializing to 𝒞1\mathcal{C}_{1}, we have:

Corollary 6.7.1.

For fixed D<0D<0 with 3D3\nmid D and 3h(D)3\nmid h_{\mathbb{Q}(\sqrt{D})}, the proportion of imaginary quadratic fields KK such that g6,D(𝒞1)/Kg_{6,D}(\mathcal{C}_{1})/K has rank 11 is at least 21ω(D)2^{-1-\omega(D)}.

Proof.

By applying Theorem 6.7 with N=cond(𝒞1)=27N=\text{cond}(\mathcal{C}_{1})=27, the result follows. ∎

7. Ranks of twists over \mathbb{Q}

For some suitable elliptic curve E/E/\mathbb{Q}, we have discussed the proportion of imaginary quadratic fields KK with ED/KE_{D}/K yielding elliptic curves of either rank 11 or rank 0. We will now consider the ranks over \mathbb{Q} instead.

We will need the concept of a root number. The root number wE/Kw_{E/K} of an elliptic curve E/KE/K is the value wE/K{1,1}w_{E/K}\in\{-1,1\} such that LE/K(s)=wE/KLE/K(2s)L_{E/K}(s)=w_{E/K}L_{E/K}(2-s).

Theorem 7.1.

Fix EE with N=cond(E)N=\text{cond}(E) and DD satisfying gcd(N,D)=gcd(N,6)=1\gcd(N,D)=\gcd(N,6)=1. Then ED/E_{D}/\mathbb{Q} has rank 11 for at least ϕ(N)4N\frac{\phi(N)}{4N} of all such DD, and rank 0 for at least ϕ(N)4N\frac{\phi(N)}{4N} of all such DD.

Proof.

By [15], Heegner points in E/KE/K exist iff wE/K=1w_{E/K}=-1. Furthermore, if E/KE/K has rank 11, then E/E/\mathbb{Q} has rank 1wE/2\frac{1-w_{E/\mathbb{Q}}}{2}, and wED/=(DN)wE/w_{E_{D}/\mathbb{Q}}=\left(\frac{D}{-N}\right)w_{E/\mathbb{Q}}. It follows that if ED/KE_{D}/K has rank 11, then ED/E_{D}/\mathbb{Q} has rank 11 if wE/=1w_{E/\mathbb{Q}}=-1.

For D>0D>0, Corollary 5.6.1 shows that when 3h(3D)3\nmid h_{\mathbb{Q}(\sqrt{-3D})}, there exists some imaginary quadratic field KK (in fact, a positive density) such that ED/KE_{D}/K has rank 11, and thus it suffices to check when wE/=1w_{E/\mathbb{Q}}=-1. Since wE/w_{E/\mathbb{Q}} is fixed, we check the proportion of D>0D>0 such that (DN)=±1\left(\frac{D}{-N}\right)=\pm 1 in each case. We have (DN)=(DN)\left(\frac{D}{-N}\right)=\left(\frac{D}{N}\right) which depends only on the residue of DD modulo NN. There are exactly ϕ(N)/2\phi(N)/2 quadratic residues and quadratic nonresidues, and thus the proportion of DD (assuming 3h(3D)3\nmid h_{\mathbb{Q}(\sqrt{-3D})}) is exactly ϕ(N)2N\frac{\phi(N)}{2N}. Now [8, Lemma 2.2] implies that for every mm such that Dm(modN)(DN)=wE/D\equiv m\pmod{N}\implies\left(\frac{D}{-N}\right)=-w_{E/\mathbb{Q}}, then the proportion

S(X,3m,N)|D>0|3D3m(modN),h(3D)0(mod3)||D>0|3D3m(modN)|12.S_{-}(X,-3m,N)\coloneqq\frac{\left|D>0\hskip 2.84526pt|\hskip 2.84526pt-3D\equiv-3m\pmod{N},\hskip 2.84526pth_{\mathbb{Q}(\sqrt{-3D})}\not\equiv 0\pmod{3}\right|}{\left|D>0\hskip 2.84526pt|\hskip 2.84526pt-3D\equiv-3m\pmod{N}\right|}\geq\frac{1}{2}.

Since this holds for every such mm, it follows that the proportion of such DD satisfying (DN)=wE/\left(\frac{D}{-N}\right)=-w_{E/\mathbb{Q}} is at least 12\frac{1}{2}.

For D<0D<0, Corollary 5.6.2 shows that when 3h(D)3\nmid h_{\mathbb{Q}(\sqrt{D})}, there exists some imaginary quadratic field KK (in fact, a positive density) such that ED/KE_{D}/K has rank 11, and thus it suffices to check when wE/=1w_{E/\mathbb{Q}}=-1. Since wE/w_{E/\mathbb{Q}} is fixed, we check the proportion of D<0D<0 such that (DN)=±1\left(\frac{D}{-N}\right)=\pm 1 in each case. We have (DN)=(DN)(D1)=(DN)\left(\frac{D}{-N}\right)=\left(\frac{D}{N}\right)\left(\frac{D}{-1}\right)=-\left(\frac{D}{N}\right) which depends only on the residue of DD modulo NN. There are exactly ϕ(N)/2\phi(N)/2 quadratic residues and quadratic nonresidues, and thus the proportion of DD (assuming 3h(D)3\nmid h_{\mathbb{Q}(\sqrt{D})}) is exactly ϕ(N)2N\frac{\phi(N)}{2N}. Now [8, Lemma 2.2] implies that for every mm such that Dm(modN)(DN)=wE/D\equiv m\pmod{N}\implies\left(\frac{D}{-N}\right)=-w_{E/\mathbb{Q}}, then the proportion

S(X,m,N)|D<0|Dm(modN),h(D)0(mod3)||D<0|Dm(modN)|12.S_{-}(X,m,N)\coloneqq\frac{\left|D<0\hskip 2.84526pt|\hskip 2.84526ptD\equiv m\pmod{N},\hskip 2.84526pth_{\mathbb{Q}(\sqrt{D})}\not\equiv 0\pmod{3}\right|}{\left|D<0\hskip 2.84526pt|\hskip 2.84526ptD\equiv m\pmod{N}\right|}\geq\frac{1}{2}.

Since this holds for every such mm, it follows that the proportion of such DD satisfying (DN)=wE/\left(\frac{D}{-N}\right)=-w_{E/\mathbb{Q}} is at least 12\frac{1}{2}.

We conclude that in either case, the proportion of DD with ED/E_{D}/\mathbb{Q} having rank 11 is at least ϕ(N)2N12=ϕ(N)4N\frac{\phi(N)}{2N}\cdot\frac{1}{2}=\frac{\phi(N)}{4N}. Analogously, when (DN)=wE/\left(\frac{D}{-N}\right)=w_{E/\mathbb{Q}}, we find that ED/E_{D}/\mathbb{Q} has rank 0, and the same result holds. ∎

Noting that the assumptions hold for a fixed (positive) proportion of DD, we conclude that for elliptic curves satisfying the above assumptions, Conjecture 1.4 holds.

Corollary 7.1.1.

For such EE satisfying the assumptions of Theorem 7.1, the weak Goldfeld conjecture holds.

References

  • BCD+ [14] Massimo Bertolini, Francesc Castella, Henri Darmon, Samit Dasgupta, Kartik Prasanna, and Victor Rotger. pp-adic L-functions and euler systems: a tale in two trilogies. Automorphic forms and Galois representations, 1:52–102, 2014.
  • BCDT [01] Christophe Breuil, Brian Conrad, Fred Diamond, and Richard Taylor. On the modularity of elliptic curves over \mathbb{Q}: wild 3-adic exercises. Journal of the American Mathematical Society, pages 843–939, 2001.
  • BDP+ [13] Massimo Bertolini, Henri Darmon, Kartik Prasanna, et al. Generalized heegner cycles and pp-adic rankin L-series. Duke Mathematical Journal, 162(6):1033–1148, 2013.
  • BS [13] Manjul Bhargava and Arul Shankar. The average size of the 5-selmer group of elliptic curves is 6, and the average rank is less than 1. arXiv preprint arXiv:1312.7859, 2013.
  • BS [15] Manjul Bhargava and Arul Shankar. Ternary cubic forms having bounded invariants, and the existence of a positive proportion of elliptic curves having rank 0. Annals of Mathematics, pages 587–621, 2015.
  • BSD [65] Bryan John Birch and Henry Peter Francis Swinnerton-Dyer. Notes on elliptic curves. ii. Journal für die reine und angewandte Mathematik, 1965.
  • BSZ [14] Manjul Bhargava, Christopher Skinner, and Wei Zhang. A majority of elliptic curves over \mathbb{Q} satisfy the birch and swinnerton-dyer conjecture. arXiv preprint arXiv:1407.1826, 2014.
  • Bye [04] Dongho Byeon. Class numbers of quadratic fields (D)\mathbb{Q}(\sqrt{D}) and (tD)\mathbb{Q}(\sqrt{tD}). Proceedings of the American Mathematical Society, 132(11):3137–3140, 2004.
  • Dar [06] Henri Darmon. Heegner points, stark-heegner points, and values of L-series. In International congress of mathematicians, volume 2, pages 313–345, 2006.
  • EK [40] Paul Erdös and Mark Kac. The gaussian law of errors in the theory of additive number theoretic functions. American Journal of Mathematics, 62(1):738–742, 1940.
  • Gol [79] Dorian Goldfeld. Conjectures on elliptic curves over quadratic fields. In Number Theory Carbondale 1979, pages 108–118. Springer, 1979.
  • Gro [80] Benedict H Gross. On the factorization of pp-adic L-series. Inventiones mathematicae, 57(1):83–95, 1980.
  • GZ [86] Benedict H Gross and Don B Zagier. Heegner points and derivatives of L-series. Inventiones mathematicae, 84(2):225–320, 1986.
  • KL [19] Daniel Kriz and Chao Li. Goldfeld’s conjecture and congruences between heegner points. In Forum of Mathematics, Sigma, volume 7. Cambridge University Press, 2019.
  • Kol [89] Viktor Alexandrovich Kolyvagin. Finiteness of E(Q) and (E,Q) for a subclass of weil curves. Mathematics of the USSR-Izvestiya, 32(3):523, 1989.
  • Kol [07] VA Kolyvagin. Euler systems. In The Grothendieck Festschrift, pages 435–483. Springer, 2007.
  • Kri [21] Daniel Kriz. Supersingular pp-adic L-functions, maass-shimura operators and waldspurger formulas. In Supersingular p-adic L-functions, Maass-Shimura Operators and Waldspurger Formulas. Princeton University Press, 2021.
  • KS [99] Nicholas M Katz and Peter Sarnak. Random matrices, Frobenius eigenvalues, and monodromy, volume 45. American Mathematical Soc., 1999.
  • NH [88] Jin Nakagawa and Kuniaki Horie. Elliptic curves with no rational points. Proceedings of the American Mathematical Society, 104(1):20–24, 1988.
  • Ser [73] Jean-Pierre Serre. Formes modulaires et fonctions zêta p-adiques. In Modular functions of one variable III, pages 191–268. Springer, 1973.
  • Ser [74] Jean-Pierre Serre. Divisibilité de certaines fonctions arithmétiques. Séminaire Delange-Pisot-Poitou. Théorie des nombres, 16(1):1–28, 1974.
  • Ser [78] Jean-Pierre Serre. Sur le résidu de la fonction zêta pp-adique d’un corps de nombres. CR Acad. Sci. Paris, 278:183–188, 1978.
  • Sil [09] Joseph H Silverman. The arithmetic of elliptic curves, volume 106. Springer Science & Business Media, 2009.
  • Sta [77] Harold M Stark. Class fields and modular forms of weight one. In Modular Functions of one Variable V, pages 277–287. Springer, 1977.
  • TW [95] Richard Taylor and Andrew Wiles. Ring-theoretic properties of certain hecke algebras. Annals of Mathematics, 141(3):553–572, 1995.
  • Was [97] Lawrence C Washington. Introduction to cyclotomic fields, volume 83. Springer Science & Business Media, 1997.
  • Wil [95] Andrew Wiles. Modular elliptic curves and fermat’s last theorem. Annals of mathematics, pages 443–551, 1995.