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

The positivity technique and low-lying zeros of Dirichlet LL-functions

Tianyu Zhao Department of Mathematics, The Ohio State University, 231 West 18th Ave, Columbus, OH 43210, USA. zhao.3709@buckeyemail.osu.edu
Abstract.

Assuming the generalized Riemann hypothesis, we rediscover and sharpen some of the best known results regarding the distribution of low-lying zeros of Dirichlet LL-functions. This builds upon earlier work of Omar, which relies on the classical positivity technique of explicit formulas. In addition, we generalize some of our results to a larger class of LL-functions and provide effective conditional estimates for the lowest zeros of Dirichlet LL-functions.

Key words and phrases:
Dirichlet LL-functions, low-lying zeros, explicit formulas
2020 Mathematics Subject Classification:
11M06, 11M26, 11M41

1. Introduction and statement of results

1.1. Individual Dirichlet LL-functions

We investigate several important quantities that naturally arise in the study of low-lying zeros of Dirichlet LL-functions. Throughout this paper χ\chi denotes a Dirichlet character with conductor q>1q>1 unless otherwise specified. Let |γχ||\gamma_{\chi}| (resp., |γχ~||\widetilde{\gamma_{\chi}}|) be the height of the lowest (resp., lowest non-real) non-trivial zero of L(s,χ)L(s,\chi). That is,

|γχ|=\displaystyle|\gamma_{\chi}|= min{|γ|:L(β+iγ,χ)=0, 0<β<1},\displaystyle\min\{|\gamma|:L(\beta+i\gamma,\chi)=0,\>0<\beta<1\},
|γχ~|=\displaystyle|\widetilde{\gamma_{\chi}}|= min{|γ|:L(β+iγ,χ)=0, 0<β<1,|γ|0}.\displaystyle\min\{|\gamma|:L(\beta+i\gamma,\chi)=0,\>0<\beta<1,\>|\gamma|\neq 0\}.

Further, let nχ=ords=1/2L(s,χ)n_{\chi}=\underset{s=1/2}{\mathrm{ord}}L(s,\chi) be the order of vanishing of L(s,χ)L(s,\chi) at the central point. Siegel [Sie45, Theorem IV] (see also the last paragraph of [Sie45]) proved that |γχ~|0|\widetilde{\gamma_{\chi}}|\to 0 as qq\to\infty, and to be precise, that

(1) |γχ|(π2+o(1))1logloglogq,|γχ~|(π+o(1))1logloglogq.|\gamma_{\chi}|\leq\left(\frac{\pi}{2}+o(1)\right)\frac{1}{\log\log\log q},\hskip 28.45274pt|\widetilde{\gamma_{\chi}}|\leq\left(\pi+o(1)\right)\frac{1}{\log\log\log q}.

Omar [Oma08] established the bound nχ<logqn_{\chi}<\log q and verified Chowla’s conjecture, which asserts that L(12,χ)0L(\frac{1}{2},\chi)\neq 0 for all χ\chi (originally stated only for real χ\chi [Cho65]), for real primitive characters of modulus up to 101010^{10}. Conditionally on the generalized Riemann hypothesis (GRH), Omar [Oma08, Theorem 7] showed that

(2) nχlogqloglogqn_{\chi}\ll\frac{\log q}{\log\log q}

and used this to improve (1) to [Oma08, Theorem 11]

(3) |γχ~|1loglogq.|\widetilde{\gamma_{\chi}}|\ll\frac{1}{\log\log q}.

These results are achieved by the positivity technique, which involves applying a specialized form of Weil’s explicit formula to test functions with suitable properties, as will be explained in §2. This method was originally developed by Odlyzko and Serre, and subsequently improved by Poitou [Poi77], to furnish lower bounds on discriminants of number fields. We refer the reader to [Odl90] for a survey of its historical background and earlier development, and to [Mil02, Oma00, Oma14] for further applications to studying the first zeros of various families of LL-functions.

For an alternative way to arrive at (2) and (3), one may start with the Riemann–von Mangoldt formula (see, e.g., [MV07, Corollary 14.6])

(4) N0(t,χ)=tπlogqt2πe+S(t,χ)+S(t,χ¯)+O(1),t>0.N_{0}(t,\chi)=\frac{t}{\pi}\log\frac{qt}{2\pi e}+S(t,\chi)+S(t,\overline{\chi})+O(1),\hskip 14.22636ptt>0.

Here N0(t,χ)N_{0}(t,\chi) counts the number of zeros ρ=β+iγ\rho=\beta+i\gamma of L(s,χ)L(s,\chi) with 0<β<10<\beta<1 and tγt-t\leq\gamma\leq t (any zero with ordinate exactly equal to ±t\pm t is counted with weight 1/21/2), and

S(t,χ):=1πargL(12+it,χ)=1π1/2LL(σ+it,χ)dσS(t,\chi):=\frac{1}{\pi}\mathrm{arg}\>L\left(\frac{1}{2}+it,\chi\right)=-\frac{1}{\pi}\int_{1/2}^{\infty}\Im\frac{L^{\prime}}{L}(\sigma+it,\chi)\,\mathrm{d}\sigma

if tγt\neq\gamma for any γ\gamma, otherwise S(t,χ):=12(S(t+0,χ)+S(t0,χ))S(t,\chi):=\frac{1}{2}(S(t+0,\chi)+S(t-0,\chi)). Under GRH, it is known that S(t,χ)log(q(|t|+1))loglog(q(|t|+3))S(t,\chi)\ll\frac{\log(q(|t|+1))}{\log\log(q(|t|+3))} uniformly in qq and tt due to a result of Selberg [Sel46]. Taking t0t\to 0 on the right-hand side of (4) gives (2), and so N(t,χ)>nχN(t,\chi)>n_{\chi} when t(loglogq)1t\gg(\log\log q)^{-1}, proving (3). Using the theory of Beurling–Selberg type extremal functions, Carneiro, Chandee and Milinovich [CCM15, Theorem 6 & Example 8] sharpened the conditional bound on S(t,χ)S(t,\chi) to

(5) |S(t,χ)|14log(q(|t|+1))loglog(q(|t|+1))+O(log(q(|t|+1))logloglog(q(|t|+1))(loglog(q(|t|+1)))2).|S(t,\chi)|\leq\frac{1}{4}\frac{\log(q(|t|+1))}{\log\log(q(|t|+1))}+O\left(\frac{\log(q(|t|+1))\log\log\log(q(|t|+1))}{(\log\log(q(|t|+1)))^{2}}\right).

Consequently,

(6) nχ12logqloglogq+O(logqlogloglogq(loglogq)2)n_{\chi}\leq\frac{1}{2}\frac{\log q}{\log\log q}+O\left(\frac{\log q\log\log\log q}{(\log\log q)^{2}}\right)

and

(7) |γχ|π21loglogq+O(logloglogq(loglogq)2).|\gamma_{\chi}|\leq\frac{\pi}{2}\frac{1}{\log\log q}+O\left(\frac{\log\log\log q}{(\log\log q)^{2}}\right).

By the same reasoning as above, (6) implies that

(8) |γχ~|πloglogq+O(logloglogq(loglogq)2).|\widetilde{\gamma_{\chi}}|\leq\frac{\pi}{\log\log q}+O\left(\frac{\log\log\log q}{(\log\log q)^{2}}\right).

1.2. The family of Dirichlet LL-functions modulo qq

Much more work has been done towards understanding the distribution of low-lying zeros in families of Dirichlet LL-functions, for example the family of quadratic Dirichlet LL-functions and the family of all Dirichlet LL-functions of a fixed modulus. In the latter direction, Murty [Mur90] showed under GRH that

(9) 1ϕ(q)χmodqnχ12+O(1logq)\frac{1}{\phi(q)}\sum_{\chi\bmod q}n_{\chi}\leq\frac{1}{2}+O\left(\frac{1}{\log q}\right)

essentially by using the positivity technique. As a result, for any ϵ>0\epsilon>0 the proportion of χ\chi mod qq such that L(12,χ)0L(\frac{1}{2},\chi)\neq 0 is at least 12ϵ\frac{1}{2}-\epsilon for all sufficiently large qq. By constructing certain extremal functions in reproducing kernel Hilbert spaces, Carneiro, Chirre and Milinovich [CCM22] improved on the average order of vanishing at low-lying heights of the form 2πtlogq\frac{2\pi t}{\log q} for t>0t>0, but at t=0t=0 (9) has not been updated so far 111For the family of primitive Dirichlet LL-functions with conductors [Q,2Q]\in[Q,2Q], the non-vanishing proportion at the central point has been extended beyond 12\frac{1}{2} for large QQ, unconditionally by Pratt [Pra19] and conditionally by Drappeau et al. [DPR23].. The best unconditional result in this direction is due to Bui [Bui12] who proved that L(12,χ)0L(\frac{1}{2},\chi)\neq 0 for at least 34%34\% of the primitive characters modulo qq. Restricting to prime moduli, Khan, Milićevié and Ngo [KMN22] improved this proportion to 513o(1)\frac{5}{13}-o(1).

In their seminal work [HR03], Hughes and Rudnick established, assuming GRH, the one-level density for the unitary family of Dirichlet LL-functions modulo qq for test functions ff with supp(f^)[2,2]\mathrm{supp}\>(\widehat{f})\in[-2,2]. As a corollary, they [HR03, Theorem 8.1] showed that

(10) lim supqq primeminχmodqχχ0|γχ|logq2π14.\limsup_{\begin{subarray}{c}q\to\infty\\ q\text{\>prime}\end{subarray}}\min_{\begin{subarray}{c}\chi\bmod q\\ \chi\neq\chi_{0}\end{subarray}}\frac{|\gamma_{\chi}|\log q}{2\pi}\leq\frac{1}{4}.

(The restriction of qq to primes is to ensure that each non-principal χ\chi mod qq is primitive, which simplifies the discussion.) In other words, there exists zeros as low as 14+o(1)\frac{1}{4}+o(1) of the average spacing between consecutive zeros. Using the variance of the one-level density, they [HR03, Theorem 8.3] also proved that for any β0.633\beta\geq 0.633,

(11) lim infqq prime1q2#{χχ0:|γχ|logq2π<β}13+π2+72β28π2β2+48β4+16π2β412π2(4β21)2.\begin{split}&\liminf_{\begin{subarray}{c}q\to\infty\\ q\text{\>prime}\end{subarray}}\frac{1}{q-2}\#\bigg{\{}\chi\neq\chi_{0}:\frac{|\gamma_{\chi}|\log q}{2\pi}<\beta\bigg{\}}\\ &\hskip 56.9055pt\geq 1-\frac{3+\pi^{2}+72\beta^{2}-8\pi^{2}\beta^{2}+48\beta^{4}+16\pi^{2}\beta^{4}}{12\pi^{2}(4\beta^{2}-1)^{2}}.\end{split}

For instance, we see by taking β=1\beta=1 that at least 80%80\% of the characters modulo qq have low-lying zeros within the expected height, and this proportion increases to 11π2312π2=0.8913\frac{11\pi^{2}-3}{12\pi^{2}}=0.8913\cdots as β\beta\to\infty.

1.3. Statement of results.

The purpose of this work is further develop Omar’s argument in [Oma08] to reprove and refine some of the aforementioned results, in hopes of demonstrating the versatility and efficacy of the positivity technique, which is at the same time relatively simple and straightforward. For individual Dirichlet LL-functions, we refine (6), (7) and (8) by removing a logloglogq\log\log\log q factor from each error term. In particular, this argument circumvents the intricate analysis of S(t,χ)S(t,\chi) and thus has the additional advantage that it can easily be made fully explicit (see Theorem 16).

Theorem 1.

Assuming GRH,

(12) |γχ|π2loglogq+π(log4+1)2(loglogq)2+O(1(loglogq)3).|\gamma_{\chi}|\leq\frac{\pi}{2\log\log q}+\frac{\pi(\log 4+1)}{2(\log\log q)^{2}}+O\left(\frac{1}{(\log\log q)^{3}}\right).

Moreover, for any constant C>log4+1C>\log 4+1, L(s,χ)L(s,\chi) has at least (π28(Clog41)+o(1))logq(loglogq)2(\frac{\pi^{2}}{8}(C-\log 4-1)+o(1))\frac{\log q}{(\log\log q)^{2}} zeros with height at most π2(loglogq)1+π2C(loglogq)2\frac{\pi}{2}(\log\log q)^{-1}+\frac{\pi}{2}C(\log\log q)^{-2}.

This result indicates that if (12) turns out to be optimal for some primitive χmodq\chi\bmod q (although this is not the case in all likelihood), then for all C>log4+1C>\log 4+1, the average spacing between zeros with height

|γ|[π2loglogq+π(log4+1)2(loglogq)2,π2loglogq+πC2(loglogq)2]|\gamma|\in\bigg{[}\frac{\pi}{2\log\log q}+\frac{\pi(\log 4+1)}{2(\log\log q)^{2}},\>\frac{\pi}{2\log\log q}+\frac{\pi C}{2(\log\log q)^{2}}\bigg{]}

is at most (8π+o(1))1logq<2πlogq(\frac{8}{\pi}+o(1))\frac{1}{\log q}<\frac{2\pi}{\log q}.

Theorem 2.

Assuming GRH,

nχlogq2loglogq+(log4+1)logq2(loglogq)2+O(logq(loglogq)3).n_{\chi}\leq\frac{\log q}{2\log\log q}+\frac{(\log 4+1)\log q}{2(\log\log q)^{2}}+O\left(\frac{\log q}{(\log\log q)^{3}}\right).
Theorem 3.

Assuming GRH,

|γχ~|πloglogq+π(log4+1)(loglogq)2+O(1(loglogq)3).|\widetilde{\gamma_{\chi}}|\leq\frac{\pi}{\log\log q}+\frac{\pi(\log 4+1)}{(\log\log q)^{2}}+O\left(\frac{1}{(\log\log q)^{3}}\right).
Remark 1.

Despite the advantages discussed above, our approach does have the drawback that the qualities of our estimates become inferior away from s=12s=\frac{1}{2}, whereas the method of Carneiro et al. in [CCM15] applies to the order of vanishing at any height and to the distance between two arbitrary consecutive zeros. We also have to choose a different test function for each estimate, which might seem somewhat ad hoc, while (6)-(8) follow readily from (5) all at once.

In fact, (5)-(8) are special cases of results in [CCM15] that are applicable to a larger class of LL-functions including those associated to cuspidal automorphic representations of GLm\mathrm{GL}_{m} over a number field. These results sharpen what Omar obtained in this general setting [Oma14]. Thus, for completeness, we shall give extensions of Theorems 13 in §7.

Slightly more precise versions of (9) and (10) are also obtained for the family of χ\chi mod qq.

Theorem 4.

Assuming GRH,

1ϕ(q)1χmodqχχ0nχ12loglogq2logq+O(logloglogqlogq).\frac{1}{\phi(q)-1}\sum_{\begin{subarray}{c}\chi\bmod q\\ \chi\neq\chi_{0}\end{subarray}}n_{\chi}\leq\frac{1}{2}-\frac{\log\log q}{2\log q}+O\left(\frac{\log\log\log q}{\log q}\right).

Hence, the proportion of χmodq\chi\bmod q satisfying L(12,χ)0L(\frac{1}{2},\chi)\neq 0 is strictly greater than 12\frac{1}{2} for all sufficiently large qq.

Theorem 5.

Assuming GRH,

minχmodqχχ0|γχ|logq2π14loglogq4logq+O(logloglogqlogq).\min_{\begin{subarray}{c}\chi\bmod q\\ \chi\neq\chi_{0}\end{subarray}}\frac{|\gamma_{\chi}|\log q}{2\pi}\leq\frac{1}{4}-\frac{\log\log q}{4\log q}+O\left(\frac{\log\log\log q}{\log q}\right).

Notice that the restriction of qq to primes is removed from (10). A similar estimate holds for |γχ~||\widetilde{\gamma_{\chi}}| with 1/41/4 replaced by 1/21/2 on the right-hand side.

In the other direction, we prove the following result on large heights of the first zeros:

Theorem 6.

Assuming GRH,

maxχmodqχχ0|γχ|logq2π14+γ+log8π4logq+O(1(logq)2)\max_{\begin{subarray}{c}\chi\bmod q\\ \chi\neq\chi_{0}\end{subarray}}\frac{|\gamma_{\chi}|\log q}{2\pi}\geq\frac{1}{4}+\frac{\gamma+\log 8\pi}{4\log q}+O\left(\frac{1}{(\log q)^{2}}\right)

where γ=0.5772\gamma=0.5772\cdots is Euler’s constant.

Regarding the proportion of characters with low-lying zeros, we improve (11) for all 1/2<β<1/2<\beta<\infty.

Theorem 7.

Let α0=1.8652\alpha_{0}=1.8652\cdots be the unique number in (1,)(1,\infty) at which the function f(α)=(6α2+π23)(α+1)12π2(α1)f(\alpha)=\frac{(6\alpha^{2}+\pi^{2}-3)(\alpha+1)}{12\pi^{2}(\alpha-1)} is minimized. Assuming GRH,

lim infqqprime1q2\displaystyle\liminf_{\begin{subarray}{c}q\to\infty\\ q\>\text{prime}\end{subarray}}\frac{1}{q-2} #{χχ0:|γχ|logq2π<β}\displaystyle\#\left\{\chi\neq\chi_{0}:\frac{|\gamma_{\chi}|\log q}{2\pi}<\beta\right\}
{11+3+π2+72β28π2β2+48β4+16π2β412π2(4β21)2,1/2<β<β0,11+f(α0)4β2,ββ0\displaystyle\geq\begin{cases}\dfrac{1}{1+\dfrac{3+\pi^{2}+72\beta^{2}-8\pi^{2}\beta^{2}+48\beta^{4}+16\pi^{2}\beta^{4}}{12\pi^{2}(4\beta^{2}-1)^{2}}},&1/2<\beta<\beta_{0},\\ \dfrac{1}{1+\dfrac{f(\alpha_{0})}{4\beta^{2}}},&\beta\geq\beta_{0}\end{cases}

where β0=α0+1α01/2=0.9098\beta_{0}=\sqrt{\frac{\alpha_{0}+1}{\alpha_{0}-1}}/2=0.9098\cdots and f(α0)=0.7757f(\alpha_{0})=0.7757\cdots. In particular, the left-hand side converges to 1 as β\beta\to\infty.

Without any constraints on the modulus qq, we give in Remark 4 a slightly smaller proportion that is 1O(β1)1-O(\beta^{-1}) as β\beta\to\infty. These results imply that although we are unable to rule out the existence of χmodq\chi\bmod q satisfying |γχ|(loglogq)1|\gamma_{\chi}|\gg(\log\log q)^{-1} (see Theorem 1), the proportion of such characters is o(1)o(1) as qq\to\infty. A similar estimate for the family of quadratic Dirichlet LL-functions is obtained in Theorem 14.

For the proof, in contrast to Hughes and Rudnick’s approach (of fixing some test function and applying Chebyshev’s inequality using the variance of the one-level density), we pursue a more direct cardinality argument based on the Cauchy–Schwarz inequality and allow for flexibility with the choice of the test function (see §4.6).

Refer to caption
Figure 1. A plot illustrating the lower bounds provided by (11) and Theorem 7 for small β\beta.

2. Preliminaries and lemmas

2.1. Weil’s explicit formula

We begin by stating the version of Weil’s explicit formula as formulated by Barner [Bar81] and adapted for L(s,χ)L(s,\chi). We say that a function F:F:\mathbb{R}\to\mathbb{R} is admissible if it is even and satisfies F(0)=1F(0)=1 along with the following properties:

  1. (a)

    FF is continuously differentiable everywhere except at finitely many points aia_{i} for which F(ai)=12(F(ai+0)+F(ai0))F(a_{i})=\frac{1}{2}(F(a_{i}+0)+F(a_{i}-0)).

  2. (b)

    There exists ϵ>0\epsilon>0 such that F(x)F(x) and F(x)F^{\prime}(x) are both O(e(1/2+ϵ)|x|)O(e^{-(1/2+\epsilon)|x|}) as xx\to\infty.

Put

(13) Φ(F)(s):=F(x)e(s1/2)xdx.\Phi(F)(s):=\int_{-\infty}^{\infty}F(x)e^{(s-1/2)x}\,\mathrm{d}x.

Then we have the following:

Theorem 8 (Weil).

For any admissible function FF,

(14) ρΦ(F)(ρ)=logqπIχ(F)2n=1Re(χ(n))F(logn)Λ(n)n\sum_{\rho}\Phi(F)(\rho)=\log\frac{q}{\pi}-I_{\chi}(F)-2\sum_{n=1}^{\infty}\textrm{Re}(\chi(n))F(\log n)\frac{\Lambda(n)}{\sqrt{n}}

where the sum on the left-hand side runs over the non-trivial zeros ρ=β+iγ\rho=\beta+i\gamma of L(s,χ)L(s,\chi) with 0<β<10<\beta<1 (counted with multiplicity) and

Iχ(F)=0(F(x/2)e(1/4+δχ/2)x1exexx)dx,δχ=1χ(1)2.I_{\chi}(F)=\int_{0}^{\infty}\bigg{(}\frac{F(x/2)e^{-(1/4+\delta_{\chi}/2)x}}{1-e^{-x}}-\frac{e^{-x}}{x}\bigg{)}\,\mathrm{d}x,\hskip 28.45274pt\delta_{\chi}=\frac{1-\chi(-1)}{2}.

As usual, Λ(n)\Lambda(n) denotes the von Mangoldt function, which takes on the value logp\log p when n=pmn=p^{m} is a prime power, and 0 otherwise.

2.2. Overview of the positivity technique

Define the Fourier transform of an admissible function FF by

(15) F^(t):=F(x)eitxdx.\widehat{F}(t):=\int_{-\infty}^{\infty}F(x)e^{itx}\,\mathrm{d}x.

Note that if ρ=12+iγ\rho=\frac{1}{2}+i\gamma, then Φ(F)(ρ)=F^(γ)\Phi(F)(\rho)=\widehat{F}(\gamma) and Φ(FT)(ρ)=TF^(Tγ)\Phi(F_{T})(\rho)=T\widehat{F}(T\gamma) where FT(x):=F(x/T)F_{T}(x):=F(x/T). We assume GRH henceforth, so that every non-trivial zero is of this form.

To see how to bound |γχ||\gamma_{\chi}| from above, suppose that F^(t)0\widehat{F}(t)\geq 0 for |t|T0|t|\leq T_{0} and that F^(t)0\widehat{F}(t)\leq 0 for |t|T0|t|\geq T_{0}. Putting T=T0/|γχ|T=T_{0}/|\gamma_{\chi}| (we may assume that |γχ|0|\gamma_{\chi}|\neq 0), we have Φ(FT)(ρ)0\Phi(F_{T})(\rho)\leq 0 for each ρ\rho. If we can estimate the second and third terms on the right-hand side of (14) in terms of TT, then for sufficiently large qq this will lead to a lower bound on TT in terms of qq, and hence an upper bound on |γχ||\gamma_{\chi}|.

The treatment for |γχ~||\widetilde{\gamma_{\chi}}| is similar, but we get a weaker bound because the zeros at 12\frac{1}{2} do make a positive contribution to the sum. To make this precise, we first derive an upper bound on nχn_{\chi} by applying (14) to HTH_{T} for some HH with non-negative Fourier transform H^\widehat{H}. By dropping the contribution from all zeros ρ12\rho\neq\frac{1}{2}, we see that the left-hand side is at least nχH^(0)Tn_{\chi}\widehat{H}(0)T, and the problem therefore reduces to bounding the right-hand side and choosing the optimal TT.

Finally, when considering the family of all non-principal χ\chi mod qq simultaneously, we sum both sides of (14) over χ\chi and exploit the orthogonality of characters instead of trivially bounding the sum over prime powers by the triangle inequality.

2.3. Test functions

We now describe two classes of admissible functions that could be used to derive an upper bound on |γχ||\gamma_{\chi}|, for which we require that F^(t)0\widehat{F}(t)\leq 0 outside of some bounded interval. For practical reasons we also need FF to be compactly supported. The first has the form

(16) Fα(x)={(1|x|)cos(πx)+απsin(π|x|)if x[1,1],0otherwiseF^{\alpha}(x)=\begin{cases}(1-|x|)\cos(\pi x)+\dfrac{\alpha}{\pi}\sin(\pi|x|)&\text{if\>\>$x\in[-1,1]$},\\ 0&\text{otherwise}\end{cases}

where α>1\alpha>1. (Omar fixes α=3\alpha=3 in [Oma08], while our α\alpha will be a parameter depending on qq.) One can check that

(17) Fα^(t)=((α+1)(α1)t2π2)(2πcos(t/2)π2t2)2,\widehat{F^{\alpha}}(t)=\bigg{(}(\alpha+1)-(\alpha-1)\frac{t^{2}}{\pi^{2}}\bigg{)}\bigg{(}\frac{2\pi\cos(t/2)}{\pi^{2}-t^{2}}\bigg{)}^{2},

so that Fα^(t)0\widehat{F^{\alpha}}(t)\leq 0 for |t|α+1α1π|t|\geq\sqrt{\frac{\alpha+1}{\alpha-1}}\pi.

Another example is given by the Fourier transforms of the Beurling–Selberg minorants for symmetric intervals [β,β][-\beta,\beta] with β>1/2\beta>1/2. In the 1930s, Beurling studied the entire functions

B±(z)=(sin(πz)π)2(2z+n=11(zn)2n=11(z+n)2)±(sin(πz)πz)2B^{\pm}(z)=\left(\frac{\sin(\pi z)}{\pi}\right)^{2}\left(\frac{2}{z}+\sum_{n=1}^{\infty}\frac{1}{(z-n)^{2}}-\sum_{n=1}^{\infty}\frac{1}{(z+n)^{2}}\right)\pm\left(\frac{\sin(\pi z)}{\pi z}\right)^{2}

and observed that B(x)sgn(x)B+(x)B^{-}(x)\leq\mathrm{sgn}(x)\leq B^{+}(x) for all xx\in\mathbb{R}, that |B±(x)sgn(x)|dx=1\int_{-\infty}^{\infty}|B^{\pm}(x)-\mathrm{sgn}(x)|\,\mathrm{d}{x}=1, and that B±^(t)\widehat{B^{\pm}}(t) 222Here we interpret B±^(t)\widehat{B^{\pm}}(t) in the sense of distributions as B±L1()B^{\pm}\not\in L^{1}(\mathbb{R}). has compact support in [2π,2π][-2\pi,2\pi]. The Beurling–Selberg minorant for [β,β][-\beta,\beta] is then given by

(18) B[β,β](t)=12(B(t+β)+B(t+β))=12(B+(tβ)+B+(tβ)).B^{-}_{[-\beta,\beta]}(t)=\frac{1}{2}(B^{-}(t+\beta)+B^{-}(-t+\beta))=-\frac{1}{2}(B^{+}(t-\beta)+B^{+}(-t-\beta)).

One can show that B[β,β]^(0)=2β1\widehat{B^{-}_{[-\beta,\beta]}}(0)=2\beta-1 (see §4.5). Now define for β>1/2\beta>1/2

(19) Jβ(x):=12β1B[β,β]^(x).J^{\beta}(x):=\frac{1}{2\beta-1}\widehat{B^{-}_{[-\beta,\beta]}}(x).

It is not hard to verify that Jβ(0)=1J^{\beta}(0)=1, that Jβ(x)J^{\beta}(x) is even and compactly supported in [2π,2π][-2\pi,2\pi], and that Jβ^(t)𝟙[β,β](t)\widehat{J^{\beta}}(t)\leq\mathbbm{1}_{[-\beta,\beta]}(t), which denotes the characteristic function on [β,β][-\beta,\beta]. (In particular, Jβ^(t)0\widehat{J^{\beta}}(t)\leq 0 for all |t|β|t|\geq\beta.) See [Vaa85] for a detailed exposition on the constructions and some earlier applications of such functions.

2.4. Lemmas

To end this section we introduce two lemmas concerning the term IχI_{\chi} and the sum over prime powers that appear in the explicit formula (14) when applied to the test function FαF^{\alpha} as defined in (16). The proofs can easily be adapted for other test functions.

Lemma 9.
|Iχ(FTα)|αT+1|I_{\chi}(F^{\alpha}_{T})|\ll\frac{\alpha}{T}+1

where the implied constant is absolute.

Proof.

Since FTα(x)=Fα(x/T)F^{\alpha}_{T}(x)=F^{\alpha}(x/T) is compactly supported in [T,T][-T,T],

Iχ(FTα)=\displaystyle I_{\chi}(F^{\alpha}_{T})= 02TFTα(x/2)11exe(1/4+δχ/2)xdx+0(e(1/4+δχ/2)x1exexx)dx\displaystyle\int_{0}^{2T}\frac{F^{\alpha}_{T}(x/2)-1}{1-e^{-x}}e^{-(1/4+\delta_{\chi}/2)x}\,\mathrm{d}x+\int_{0}^{\infty}\bigg{(}\frac{e^{-(1/4+\delta_{\chi}/2)x}}{1-e^{-x}}-\frac{e^{-x}}{x}\bigg{)}\,\mathrm{d}x
2Te(1/4+δχ/2)x1exdx\displaystyle\hskip 28.45274pt-\int_{2T}^{\infty}\frac{e^{-(1/4+\delta_{\chi}/2)x}}{1-e^{-x}}\,\mathrm{d}x

where δχ=0\delta_{\chi}=0 or 11 according as χ\chi is even or odd. Also recall that FTα(0)=1F^{\alpha}_{T}(0)=1, so by the mean value theorem there exists x(0,1)x^{*}\in(0,1) such that the first integral can be rewritten as

12TFα(x)02Txe(1/4+δχ/2)x1exdx.\frac{1}{2T}F^{\alpha\prime}(x^{*})\int_{0}^{2T}\frac{xe^{-(1/4+\delta_{\chi}/2)x}}{1-e^{-x}}\,\mathrm{d}x.

Note that |Fα(x)|α|F^{\alpha\prime}(x)|\ll\alpha on (0,1)(0,1) and that the integral above converges. By Gauss’ Digamma theorem, the second integral equals

ΓΓ(14+δχ2)={γ+3log2+π/2if δχ=0,γ+3log2π/2if δχ=1.\frac{\Gamma^{\prime}}{\Gamma}\left(\frac{1}{4}+\frac{\delta_{\chi}}{2}\right)=\begin{cases}\gamma+3\log 2+\pi/2&\text{if $\delta_{\chi}=0$},\\ \gamma+3\log 2-\pi/2&\text{if $\delta_{\chi}=1$}.\end{cases}

The third integral is plainly O(1)O(1), and the proof is complete. ∎

Lemma 10.

For α3\alpha\geq 3 and large TT,

|n=1Re(χ(n))FTα(logn)Λ(n)n|(4+O(T1))(α1)eT/2T.\left|\sum_{n=1}^{\infty}\textrm{Re}(\chi(n))F^{\alpha}_{T}(\log n)\frac{\Lambda(n)}{\sqrt{n}}\right|\leq\left(4+O(T^{-1})\right)(\alpha-1)\frac{e^{T/2}}{T}.
Proof.

We claim that when α3\alpha\geq 3,

(20) Fα(x)(α1)(1x)forx[0,1].F^{\alpha}(x)\leq(\alpha-1)(1-x)\hskip 14.22636pt\text{for}\>\>x\in[0,1].

Since Fα(1)=(α1)F^{\alpha\prime}(1)=-(\alpha-1), it suffices to show that d2dx2Fα(x)<0\frac{\mathrm{d}^{2}}{\mathrm{d}x^{2}}F^{\alpha}(x)<0, or sin(πx)>π(1x)cos(πx)\sin(\pi x)>-\pi(1-x)\cos(\pi x), for x(0,1)x\in(0,1). This is readily seen from the Taylor expansions of both sides.

As usual, let ψ(t)=ntΛ(n)\psi(t)=\sum_{n\leq t}\Lambda(n) denote the Chebyshev function. Then, using (20), we can trivially bound the sum by

\displaystyle\leq (α1)1eTdψ(t)tα1T1eTlogttdψ(t)\displaystyle(\alpha-1)\int_{1}^{e^{T}}\frac{\,\mathrm{d}\psi(t)}{\sqrt{t}}-\frac{\alpha-1}{T}\int_{1}^{e^{T}}\frac{\log t}{\sqrt{t}}\,\mathrm{d}\psi(t)
=\displaystyle= (α1)(ψ(eT)eT/2+121eTψ(t)t3/2dt)\displaystyle(\alpha-1)\bigg{(}\frac{\psi(e^{T})}{e^{T/2}}+\frac{1}{2}\int_{1}^{e^{T}}\frac{\psi(t)}{t^{3/2}}\,\mathrm{d}t\bigg{)}
α1T(TeT/2ψ(eT)1eTψ(t)(1t3/2logt2t3/2)dt)\displaystyle\hskip 28.45274pt-\frac{\alpha-1}{T}\bigg{(}\frac{T}{e^{T/2}}\psi(e^{T})-\int_{1}^{e^{T}}\psi(t)\bigg{(}\frac{1}{t^{3/2}}-\frac{\log t}{2t^{3/2}}\bigg{)}\,\mathrm{d}t\bigg{)}
(21) =\displaystyle= α1T1eTψ(t)t3/2dt+α121eTψ(t)t3/2(1logtT)dt\displaystyle\frac{\alpha-1}{T}\int_{1}^{e^{T}}\frac{\psi(t)}{t^{3/2}}\,\mathrm{d}t+\frac{\alpha-1}{2}\int_{1}^{e^{T}}\frac{\psi(t)}{t^{3/2}}\bigg{(}1-\frac{\log t}{T}\bigg{)}\,\mathrm{d}t
=\displaystyle= (4+O(T1))(α1)eT/2T\displaystyle(4+O(T^{-1}))(\alpha-1)\frac{e^{T/2}}{T}

where we invoked the prime number theorem (in the weak form) ψ(t)tt/logt\psi(t)-t\ll t/\log t. ∎

3. Estimates for individual Dirichlet LL-functions: proofs of Theorems 13

3.1. Bounding |γχ||\gamma_{\chi}|: proof of Theorem 1

Assume that L(12,χ)0L(\frac{1}{2},\chi)\neq 0, i.e., |γχ|0|\gamma_{\chi}|\neq 0, since the theorem trivially holds otherwise. Consider the function FTα(x)=Fα(x/T)F^{\alpha}_{T}(x)=F^{\alpha}(x/T) with FαF^{\alpha} from (16) and T=α+1α1π/|γχ|T=\sqrt{\frac{\alpha+1}{\alpha-1}}\pi/|\gamma_{\chi}|. We see from (3) that TloglogqT\gg\log\log q. Note that ρΦ(FTα)(ρ)=γTFα^(Tγ)0\sum_{\rho}\Phi(F^{\alpha}_{T})(\rho)=\sum_{\gamma}T\widehat{F^{\alpha}}(T\gamma)\leq 0 because Fα^(t)0\widehat{F^{\alpha}}(t)\leq 0 for all |t|α+1α1π|t|\geq\sqrt{\frac{\alpha+1}{\alpha-1}}\pi. Thus, in view of Theorem 8 and Lemmas 9 and  10, we have

(4+O(T1))αeT/2T/2logq+O(αT+1).\displaystyle(4+O(T^{-1}))\frac{\alpha e^{T/2}}{T/2}\geq\log q+O\left(\frac{\alpha}{T}+1\right).

Set α=loglogq\alpha=\log\log q and

Δ=logq+O(αT+1)(4+O(T1))α,\Delta=\frac{\log q+O(\frac{\alpha}{T}+1)}{(4+O(T^{-1}))\alpha},

so that eT/2T/2Δ\frac{e^{T/2}}{T/2}\geq\Delta. Then

T2\displaystyle\frac{T}{2}\geq logΔ+loglogΔ+log(1+loglogΔlogΔ)\displaystyle\log\Delta+\log\log\Delta+\log\left(1+\frac{\log\log\Delta}{\log\Delta}\right)
=\displaystyle= logΔ+loglogΔ+loglogΔlogΔ+O((loglogΔlogΔ)2)\displaystyle\log\Delta+\log\log\Delta+\frac{\log\log\Delta}{\log\Delta}+O\left(\left(\frac{\log\log\Delta}{\log\Delta}\right)^{2}\right)
=\displaystyle= loglogqlog4+O(1loglogq).\displaystyle\log\log q-\log 4+O\left(\frac{1}{\log\log q}\right).

On expressing TT in terms of |γχ||\gamma_{\chi}| and α\alpha and appealing to the fact that 1x=1x/2+O(x2)\sqrt{1-x}=1-x/2+O(x^{2}) as x0x\to 0, we arrive at

π2|γχ|\displaystyle\frac{\pi}{2|\gamma_{\chi}|}\geq (11loglogq+O(1(loglogq)2))(loglogqlog4+O(1loglogq))\displaystyle\left(1-\frac{1}{\log\log q}+O\left(\frac{1}{(\log\log q)^{2}}\right)\right)\left(\log\log q-\log 4+O\left(\frac{1}{\log\log q}\right)\right)
=\displaystyle= loglogqlog41+O(1loglogq),\displaystyle\log\log q-\log 4-1+O\left(\frac{1}{\log\log q}\right),

which completes the proof of (12).

Next we prove the second assertion regarding the number of zeros below t1:=π2(loglogq)1+π2C(loglogq)2t_{1}:=\frac{\pi}{2}(\log\log q)^{-1}+\frac{\pi}{2}C(\log\log q)^{-2} with C>log4+1C>\log 4+1. Let nn be the positive integer such that |γχ,n|<t1|γχ,n+1||\gamma_{\chi,n}|<t_{1}\leq|\gamma_{\chi,n+1}| (here γχ,k\gamma_{\chi,k} denotes the kthk^{\text{th}} zero ordered by height), and our goal is to show that

(22) n(π28(Clog41)+o(1))logq(loglogq)2.n\geq\left(\frac{\pi^{2}}{8}(C-\log 4-1)+o(1)\right)\frac{\log q}{(\log\log q)^{2}}.

Take α=aloglogq\alpha=a\log\log q for some parameter a>0a>0. For T:=α+1α1π/t1=2(loglogqC+1/a+o(1))T:=\sqrt{\frac{\alpha+1}{\alpha-1}}\pi/t_{1}=2(\log\log q-C+1/a+o(1)), we have Φ(FTα)(ρk)0\Phi(F^{\alpha}_{T})(\rho_{k})\leq 0 for all k>nk>n, and then another application of (14) gives

logq+O(αT+1)(4+o(1))(α1)eT/2T/2\displaystyle\log q+O\left(\frac{\alpha}{T}+1\right)-(4+o(1))(\alpha-1)\frac{e^{T/2}}{T/2}
1knΦ(FTα)(ρk)nTFα^(0)=nT4(α+1)π2.\displaystyle\hskip 85.35826pt\leq\sum_{1\leq k\leq n}\Phi(F^{\alpha}_{T})(\rho_{k})\leq nT\widehat{F^{\alpha}}(0)=nT\frac{4(\alpha+1)}{\pi^{2}}.

After rearranging the inequality we find that

n(π2414aeC+1/a2a+o(1))logq(loglogq)2.n\geq\left(\frac{\pi^{2}}{4}\frac{1-4ae^{-C+1/a}}{2a}+o(1)\right)\frac{\log q}{(\log\log q)^{2}}.

Finally, it is a calculus exercise to determine that the optimal choice of aa is (Clog4)1(C-\log 4)^{-1}, which yields (22).

Remark 2.

By examining the the first part of the preceding proof and the proof of Lemma 10, we see that if the chosen test function FF is supported in [c,c][-c,c] and F^(t)0\widehat{F}(t)\leq 0 for |t|t0|t|\geq t_{0}, then the coefficient of the leading term 1/loglogq1/\log\log q in the bound on |γχ||\gamma_{\chi}| would be t0c/2t_{0}c/2. For our choice FαF^{\alpha}, c=1c=1 and t0π+t_{0}\to\pi^{+} as α\alpha\to\infty, which leads to the coefficient π/2\pi/2. Alternatively, we could work with JβJ^{\beta} as defined in (19) with c=2πc=2\pi and t0=βt_{0}=\beta, and we obtain the same coefficient by letting β1/2+\beta\to 1/2^{+}. However, it would be harder to determine the lower order terms in the estimate because of the more complicated nature of JβJ^{\beta}.

3.2. Bounding nχn_{\chi}: proof of Theorem 2

As in [Oma08, Lemma 9], we work with the admissible function

(23) H(x)={1|x|if |x|<1,0otherwiseH(x)=\begin{cases}1-|x|&\text{if $|x|<1$},\\ 0&\text{otherwise}\end{cases}

whose Fourier transform is

H^(t)=(sin(t/2)t/2)2.\widehat{H}(t)=\left(\frac{\sin(t/2)}{t/2}\right)^{2}.

As can be seen from the proofs of Lemmas 9 and  10, we have Iχ(HT)=O(1)I_{\chi}(H_{T})=O(1) and

n=1HT(logn)Λ(n)n(4+O(T1))eT/2T.\sum_{n=1}^{\infty}H_{T}(\log n)\frac{\Lambda(n)}{\sqrt{n}}\leq(4+O(T^{-1}))\frac{e^{T/2}}{T}.

Again by (14) and the non-negativity of H^(t)\widehat{H}(t),

nχH^(0)Tlogq+O(1)+(4+O(T1))eT/2T/2.\displaystyle n_{\chi}\widehat{H}(0)T\leq\log q+O(1)+(4+O(T^{-1}))\frac{e^{T/2}}{T/2}.

Putting T=2loglogqΔT=2\log\log q-\Delta for some constant Δ\Delta, we see that

nχ12logqloglogq+(Δ4+2eΔ/2)logq(loglogq)2+O(logq(loglogq)3),n_{\chi}\leq\frac{1}{2}\frac{\log q}{\log\log q}+\left(\frac{\Delta}{4}+\frac{2}{e^{\Delta/2}}\right)\frac{\log q}{(\log\log q)^{2}}+O\left(\frac{\log q}{(\log\log q)^{3}}\right),

and the theorem follows on choosing Δ=2log4\Delta=2\log 4.

3.3. Bounding |γχ~||\widetilde{\gamma_{\chi}}|: proof of Theorem 3

Define

(24) Gα(x):=π22(α+2)(FαH)(x)G^{\alpha}(x):=\frac{\pi^{2}}{2(\alpha+2)}(F^{\alpha}*H)(x)

where (fg)(x)=f(y)g(xy)dy(f*g)(x)=\int_{-\infty}^{\infty}f(y)g(x-y)\,\mathrm{d}y stands for the convolution of ff and gg. Note that GαG^{\alpha} is admissible, and in particular, that it is supported in [2,2][-2,2] with Gα(0)=1G^{\alpha}(0)=1. Moreover, Gα^(t)=π22(α+2)Fα^(t)H^(t)0\widehat{G^{\alpha}}(t)=\frac{\pi^{2}}{2(\alpha+2)}\widehat{F^{\alpha}}(t)\widehat{H}(t)\leq 0 for |t|α+1α1π|t|\geq\sqrt{\frac{\alpha+1}{\alpha-1}}\pi.

If |γχ~|0|\widetilde{\gamma_{\chi}}|\neq 0, consider GTα(x)G^{\alpha}_{T}(x) with T=α+1α1π/|γχ~|T=\sqrt{\frac{\alpha+1}{\alpha-1}}\pi/|\widetilde{\gamma_{\chi}}|, so that Φ(GTα)(ρ)=TGα^(Tγ)0\Phi(G^{\alpha}_{T})(\rho)=T\widehat{G^{\alpha}}(T\gamma)\\ \leq 0 except when γ=0\gamma=0. Since Fα(x)α(1|x|)F^{\alpha}(x)\ll\alpha(1-|x|) and H(x)=1|x|H(x)=1-|x| for x[1,1]x\in[-1,1], we have Gα(x)(1|x|/2)3G^{\alpha}(x)\ll(1-|x|/2)^{3}, which gives

n=1Re(χ(n))GTα(logn)Λ(n)neTT3\sum_{n=1}^{\infty}\textrm{Re}(\chi(n))G^{\alpha}_{T}(\log n)\frac{\Lambda(n)}{\sqrt{n}}\ll\frac{e^{T}}{T^{3}}

by a similar argument as in the proof of Lemma 10. Also Iχ(GTα)=O(1)I_{\chi}(G^{\alpha}_{T})=O(1) since GTα(x)=O(1)G^{\alpha}_{T}(x)=O(1), and hence

(25) nχGα^(0)Tlogq+O(eTT3)n_{\chi}\widehat{G^{\alpha}}(0)T\geq\log q+O\left(\frac{e^{T}}{T^{3}}\right)

where Gα^(0)=2(α+1)α+2\widehat{G^{\alpha}}(0)=\frac{2(\alpha+1)}{\alpha+2}. A quick proof by contradiction shows that if aT+beTT3caT+b\frac{e^{T}}{T^{3}}\geq c for T3T\geq 3 and positive numbers aa, bb and cc, then

Tmin{c(1+Δ)a,logΔc(1+Δ)b+3loglogΔc(1+Δ)b}T\geq\min\Bigg{\{}\frac{c}{(1+\Delta)a},\log\frac{\Delta c}{(1+\Delta)b}+3\log\log\frac{\Delta c}{(1+\Delta)b}\Bigg{\}}

for any Δ>0\Delta>0. In view of (25) and Theorem 2,

T\displaystyle T\geq min{(α+2)logq(α+1)(1+Δ)(logqloglogq+(log4+1)logq(loglogq)2+O(logq(loglogq)3))1,\displaystyle\min\Bigg{\{}\frac{(\alpha+2)\log q}{(\alpha+1)(1+\Delta)}\left(\frac{\log q}{\log\log q}+\frac{(\log 4+1)\log q}{(\log\log q)^{2}}+O\left(\frac{\log q}{(\log\log q)^{3}}\right)\right)^{-1},
logΔlogq(1+Δ)b+3loglogΔlogq(1+Δ)b}\displaystyle\hskip 56.9055pt\log\frac{\Delta\log q}{(1+\Delta)b}+3\log\log\frac{\Delta\log q}{(1+\Delta)b}\Bigg{\}}

where b>0b>0 is an absolute constant. Choosing α=loglogq\alpha=\log\log q and Δ=(loglogq)2\Delta=(\log\log q)^{-2}, we find that

Tloglogqlog4O(1loglogq),T\geq\log\log q-\log 4-O\left(\frac{1}{\log\log q}\right),

which yields the stated estimate for |γχ~||\widetilde{\gamma_{\chi}}|.

4. Estimates for the family of Dirichlet LL-functions modulo qq: proofs of Theorems 47

4.1. Lemmas

We start with an estimate on the average size of conductors of characters modulo qq.

Lemma 11.
logqloglogq+O(1)1ϕ(q)1χmodqχχ0log(con(χ))logq\log q-\log\log q+O(1)\leq\frac{1}{\phi(q)-1}\sum_{\begin{subarray}{c}\chi\bmod q\\ \chi\neq\chi_{0}\end{subarray}}\log(\mathrm{con}(\chi))\leq\log q

where con(χ)\mathrm{con}(\chi) stands for the conductor of χ\chi.

Proof.

The upper bound is clear with equality attained whenever q>2q>2 is a prime. On the other hand, one can show that (see [FM13, Proposition 3.3])

1ϕ(q)χmodqlog(con(χ))=logqp|qlogpp1.\frac{1}{\phi(q)}\sum_{\chi\bmod q}\log(\mathrm{con}(\chi))=\log q-\sum_{p|q}\frac{\log p}{p-1}.

Let ω(q)\omega(q) denote the number of distinct prime divisors of qq, and pkp_{k} the kthk^{\text{th}} prime. Since logxx1\frac{\log x}{x-1} is decreasing, the last sum over p|qp|q is

ppω(q)logpp1=logpω(q)+O(1)logω(q)+loglogω(q)+O(1)loglogq+O(1)\leq\sum_{p\leq p_{\omega(q)}}\frac{\log p}{p-1}=\log p_{\omega(q)}+O(1)\leq\log\omega(q)+\log\log\omega(q)+O(1)\leq\log\log q+O(1)

where we used the standard fact that ω(q)logqloglogq\omega(q)\ll\frac{\log q}{\log\log q}. ∎

Next we introduce an averaged version of Lemma 10.

Lemma 12.
|χmodqχχ0n=1Re(χ(n))FTα(logn)Λ(n)n|αeT/2T+αqlogq.\left|\sum_{\begin{subarray}{c}\chi\bmod q\\ \chi\neq\chi_{0}\end{subarray}}\sum_{n=1}^{\infty}\textrm{Re}(\chi(n))F^{\alpha}_{T}(\log n)\frac{\Lambda(n)}{\sqrt{n}}\right|\ll\frac{\alpha e^{T/2}}{T}+\alpha\sqrt{q}\log q.

where the implied constant is absolute.

Proof.

We assume that T>logqT>\log q, otherwise it will be clear how to modify (actually simplify) the argument. By (20) and the orthogonality relation of characters

(26) χmodqχχ0χ(n)={ϕ(q)1if n1modq,0if n0modq,1otherwise,\sum_{\begin{subarray}{c}\chi\bmod q\\ \chi\neq\chi_{0}\end{subarray}}\chi(n)=\begin{cases}\phi(q)-1&\text{if $n\equiv 1\bmod q$},\\ 0&\text{if $n\equiv 0\bmod q$},\\ -1&\text{otherwise},\end{cases}

we obtain

1ϕ(q)1\displaystyle\frac{1}{\phi(q)-1} χmodqχχ0pRe(χ(p))FTα(logp)logpp\displaystyle\sum_{\begin{subarray}{c}\chi\bmod q\\ \chi\neq\chi_{0}\end{subarray}}\sum_{p}\textrm{Re}(\chi(p))F^{\alpha}_{T}(\log p)\frac{\log p}{\sqrt{p}}
αp1modqpeTlogppαTp1modqpeT(logp)2p\displaystyle\leq\alpha\sum_{\begin{subarray}{c}p\equiv 1\bmod q\\ p\leq e^{T}\end{subarray}}\frac{\log p}{\sqrt{p}}-\frac{\alpha}{T}\sum_{\begin{subarray}{c}p\equiv 1\bmod q\\ p\leq e^{T}\end{subarray}}\frac{(\log p)^{2}}{\sqrt{p}}
αp1modqp2qlogpp+α22qeTπ(t;q,1)(logt2)t3/2dt\displaystyle\leq\alpha\sum_{\begin{subarray}{c}p\equiv 1\bmod q\\ p\leq 2q\end{subarray}}\frac{\log p}{\sqrt{p}}+\frac{\alpha}{2}\int_{2q}^{e^{T}}\frac{\pi(t;q,1)(\log t-2)}{t^{3/2}}\,\mathrm{d}t
α2T2qeTπ(t;q,1)((logt)24logt)t3/2dt\displaystyle\hskip 56.9055pt-\frac{\alpha}{2T}\int_{2q}^{e^{T}}\frac{\pi(t;q,1)((\log t)^{2}-4\log t)}{t^{3/2}}\,\mathrm{d}t

where π(t;q,a):=#{p:pamodq,pt}\pi(t;q,a):=\#\{p:p\equiv a\bmod q,\>p\leq t\}. The first term is plainly αlogqq\ll\frac{\alpha\log q}{\sqrt{q}}. By the Brun–Titchmarsh theorem [MV73, Theorem 2], π(t;q,a)2tϕ(q)log(t/q)\pi(t;q,a)\leq\frac{2t}{\phi(q)\log(t/q)} for t>2qt>2q, from which it follows that the sum of the two remaining terms is αeT/2ϕ(q)T\ll\frac{\alpha e^{T/2}}{\phi(q)T}. Multiplying by ϕ(q)1\phi(q)-1 gives the desired upper bound. Similarly,

χmodqχχ0pRe(χ(p))FTα(logp)logpp\displaystyle-\sum_{\begin{subarray}{c}\chi\bmod q\\ \chi\neq\chi_{0}\end{subarray}}\sum_{p}\textrm{Re}(\chi(p))F^{\alpha}_{T}(\log p)\frac{\log p}{\sqrt{p}}\leq αp0,1modqpeTlogppαTp0,1modqpeT(logp)2p\displaystyle\alpha\sum_{\begin{subarray}{c}p\not\equiv 0,1\bmod q\\ p\leq e^{T}\end{subarray}}\frac{\log p}{\sqrt{p}}-\frac{\alpha}{T}\sum_{\begin{subarray}{c}p\not\equiv 0,1\bmod q\\ p\leq e^{T}\end{subarray}}\frac{(\log p)^{2}}{\sqrt{p}}
\displaystyle\ll αeT/2T+αp2qlogpp\displaystyle\frac{\alpha e^{T/2}}{T}+\alpha\sum_{p\leq 2q}\frac{\log p}{\sqrt{p}}
\displaystyle\ll αeT/2T+αq.\displaystyle\frac{\alpha e^{T/2}}{T}+\alpha\sqrt{q}.

Finally, the contribution from higher powers of primes can be absorbed into the stated estimate. ∎

4.2. Average of nχn_{\chi}: proof of Theorem 4

We apply (14) to HTH_{T} as defined in (23) and average over all non-principal χmodq\chi\bmod q. By exploiting the orthogonality relation as in the proof of Lemma 12, we find that

χmodqχχ0n=1Re(χ(n))HT(logn)Λ(n)np0,1modqHT(logp)logppeT/2T+q.\displaystyle-\sum_{\begin{subarray}{c}\chi\bmod q\\ \chi\neq\chi_{0}\end{subarray}}\sum_{n=1}^{\infty}\textrm{Re}(\chi(n))H_{T}(\log n)\frac{\Lambda(n)}{\sqrt{n}}\ll\sum_{p\not\equiv 0,1\bmod q}H_{T}(\log p)\frac{\log p}{\sqrt{p}}\ll\frac{e^{T/2}}{T}+\sqrt{q}.

Since con(χ)q\mathrm{con}(\chi)\leq q for all χ\chi mod qq and ϕ(q)logqloglogq\phi(q)\gg\frac{\log q}{\log\log q},

1ϕ(q)1χmodqχχ0nχT\displaystyle\frac{1}{\phi(q)-1}\sum_{\begin{subarray}{c}\chi\bmod q\\ \chi\neq\chi_{0}\end{subarray}}n_{\chi}T\leq logqπ1ϕ(q)1χmodqχχ0Iχ(HT)+O(eT/2/T+qϕ(q))\displaystyle\log\frac{q}{\pi}-\frac{1}{\phi(q)-1}\sum_{\begin{subarray}{c}\chi\bmod q\\ \chi\neq\chi_{0}\end{subarray}}I_{\chi}(H_{T})+O\left(\frac{e^{T/2}/T+\sqrt{q}}{\phi(q)}\right)
=\displaystyle= logq+O((loglogq)eT/2qT+1).\displaystyle\log q+O\left(\frac{(\log\log q)e^{T/2}}{qT}+1\right).

Taking T=2(logq+loglogqlogloglogq)T=2(\log q+\log\log q-\log\log\log q), we deduce that

1ϕ(q)1χmodqχχ0nχ\displaystyle\frac{1}{\phi(q)-1}\sum_{\begin{subarray}{c}\chi\bmod q\\ \chi\neq\chi_{0}\end{subarray}}n_{\chi}\leq logq+O(1)2(logq+loglogqlogloglogq)\displaystyle\frac{\log q+O(1)}{2(\log q+\log\log q-\log\log\log q)}
=\displaystyle= 12loglogq2logq+O(logloglogqlogq),\displaystyle\frac{1}{2}-\frac{\log\log q}{2\log q}+O\left(\frac{\log\log\log q}{\log q}\right),

as desired.

4.3. Minimum of |γχ||\gamma_{\chi}|: proof of Theorem 5

Let T=α+1α1π/|γq|T=\sqrt{\frac{\alpha+1}{\alpha-1}}\pi/|\gamma_{q}| where |γq|:=minχχ0|γχ||\gamma_{q}|:=\min_{\chi\neq\chi_{0}}|\gamma_{\chi}|. Applying (14) to FTαF^{\alpha}_{T} and averaging, we find by Lemmas 11 and  12 that

logqloglogqαeT/2ϕ(q)T+αqlogqϕ(q)α(loglogq)eT/2qT.\log q-\log\log q\ll\frac{\alpha e^{T/2}}{\phi(q)T}+\frac{\alpha\sqrt{q}\log q}{\phi(q)}\ll\frac{\alpha(\log\log q)e^{T/2}}{qT}.

Now put α=alogq\alpha=a\log q for some appropriate absolute constant a>0a>0 so that eT/2T/2qloglogq\frac{e^{T/2}}{T/2}\geq\frac{q}{\log\log q}, or T/2logq+loglogqlogloglogqT/2\geq\log q+\log\log q-\log\log\log q. This in turn implies that

|γq|logq2π\displaystyle\frac{|\gamma_{q}|\log q}{2\pi}\leq logq4(logq+loglogqlogloglogq)alogq+1alogq1\displaystyle\frac{\log q}{4(\log q+\log\log q-\log\log\log q)}\sqrt{\frac{a\log q+1}{a\log q-1}}
=\displaystyle= 14loglogq4logq+O(logloglogqlogq).\displaystyle\frac{1}{4}-\frac{\log\log q}{4\log q}+O\left(\frac{\log\log\log q}{\log q}\right).

4.4. Maximum of |γχ||\gamma_{\chi}|: proof of Theorem 6

Define K(x):=F1(x)K(x):=F^{1}(x), that is,

(27) K(x)={(1|x|)cos(πx)+1πsin(π|x|)if x[1,1],0otherwise,K(x)=\begin{cases}(1-|x|)\cos(\pi x)+\dfrac{1}{\pi}\sin(\pi|x|)&\text{if\>\>$x\in[-1,1]$},\\ 0&\text{otherwise},\end{cases}

which is an admissible function with

K^(t)=8π2(cos(t/2)π2t2)2.\widehat{K}(t)=8\pi^{2}\left(\frac{\cos(t/2)}{\pi^{2}-t^{2}}\right)^{2}.

A short calculation reveals that K(x)K(x) vanishes up to the second derivative as x1x\to 1^{-}, and more precisely one has K(x)4(1|x|)3K(x)\leq 4(1-|x|)^{3}. It follows as in the proof of Lemma 12 that

χmodqχχ0n=1Re(χ(n))KT(logn)Λ(n)neT/2T3+q.-\sum_{\begin{subarray}{c}\chi\bmod q\\ \chi\neq\chi_{0}\end{subarray}}\sum_{n=1}^{\infty}\textrm{Re}(\chi(n))K_{T}(\log n)\frac{\Lambda(n)}{\sqrt{n}}\ll\frac{e^{T/2}}{T^{3}}+\sqrt{q}.

Moreover, we see from the proof of Lemma 9 that

Iχ(KT)={γ+3log2+π/2+O(T1)if χ(1)=1,γ+3log2π/2+O(T1)if χ(1)=1,I_{\chi}(K_{T})=\begin{cases}\gamma+3\log 2+\pi/2+O(T^{-1})&\text{if $\chi(-1)=1$},\\ \gamma+3\log 2-\pi/2+O(T^{-1})&\text{if $\chi(-1)=-1$},\end{cases}

and thus

1ϕ(q)1χmodqχχ0n=1Iχ(KT)=γ3log2+O(T1)+O(ϕ(q)1)-\frac{1}{\phi(q)-1}\sum_{\begin{subarray}{c}\chi\bmod q\\ \chi\neq\chi_{0}\end{subarray}}\sum_{n=1}^{\infty}I_{\chi}(K_{T})=-\gamma-3\log 2+O(T^{-1})+O(\phi(q)^{-1})

by the orthogonality relation (26).

For β>0\beta>0, denote by N(2πβlogq,χ)N(\frac{2\pi\beta}{\log q},\chi) the number of zeros of L(s,χ)L(s,\chi) with |γ|2πβlogq|\gamma|\leq\frac{2\pi\beta}{\log q}. The non-negativity of K^\widehat{K} implies that

1ϕ(q)1χmodqχχ0\displaystyle\frac{1}{\phi(q)-1}\sum_{\begin{subarray}{c}\chi\bmod q\\ \chi\neq\chi_{0}\end{subarray}} N(2πβlogq,χ)Tmin|t|2πβTlogqK^(t)\displaystyle N\left(\frac{2\pi\beta}{\log q},\chi\right)T\min_{|t|\leq\frac{2\pi\beta T}{\log q}}\widehat{K}(t)
\displaystyle\leq 1ϕ(q)1χmodqχχ0γTK^(Tγ)\displaystyle\frac{1}{\phi(q)-1}\sum_{\begin{subarray}{c}\chi\bmod q\\ \chi\neq\chi_{0}\end{subarray}}\sum_{\gamma}T\widehat{K}(T\gamma)
\displaystyle\leq logqπγ3log2+O((loglogq)eT/2qT3+loglogqq+1T).\displaystyle\log\frac{q}{\pi}-\gamma-3\log 2+O\left(\frac{(\log\log q)e^{T/2}}{qT^{3}}+\frac{\log\log q}{\sqrt{q}}+\frac{1}{T}\right).

We aim to maximize β\beta subject to the condition that

(28) logqπγ3log2+O((loglogq)eT/2qT3+loglogqq+1T)Tmin|t|2πβTlogqK^(t)<1\dfrac{\log\frac{q}{\pi}-\gamma-3\log 2+O\left(\frac{(\log\log q)e^{T/2}}{qT^{3}}+\frac{\log\log q}{\sqrt{q}}+\frac{1}{T}\right)}{T\min_{|t|\leq\frac{2\pi\beta T}{\log q}}\widehat{K}(t)}<1

for all large qq, since then the average number of zeros below 2πβlogq\frac{2\pi\beta}{\log q} will be strictly less than 1. Choose T=2logqT=2\log q, so the big-OO term in the numerator of (28) is O((logq)1)O((\log q)^{-1}). As min|t|4πβK^(t)=K^(4πβ)\min_{|t|\leq 4\pi\beta}\widehat{K}(t)=\widehat{K}(4\pi\beta) (for β<1/2\beta<1/2, say, such that K^(t)\widehat{K}(t) is monotonically decreasing on |t|4πβ|t|\leq 4\pi\beta), in view of (28) it suffices to ensure that

4π2(cos2πβπ2(4πβ)2)2>12(1logπ+γ+3log2logq+O((logq)2)).4\pi^{2}\left(\frac{\cos 2\pi\beta}{\pi^{2}-(4\pi\beta)^{2}}\right)^{2}>\frac{1}{2}\left(1-\frac{\log\pi+\gamma+3\log 2}{\log q}+O\left((\log q)^{-2}\right)\right).

Indeed, if β=(1+β~)/4\beta=(1+\tilde{\beta})/4 where β~=o(1)\tilde{\beta}=o(1), then the left-hand side becomes

8π2(sin(πβ~/2)2π2β~+π2β~2)2=8(14β~8+O(β~2))2=12(1β~+O(β~2)),8\pi^{2}\left(\frac{\sin(\pi\tilde{\beta}/2)}{2\pi^{2}\tilde{\beta}+\pi^{2}\tilde{\beta}^{2}}\right)^{2}=8\left(\frac{1}{4}-\frac{\tilde{\beta}}{8}+O(\tilde{\beta}^{2})\right)^{2}=\frac{1}{2}\left(1-\tilde{\beta}+O(\tilde{\beta}^{2})\right),

and hence we can take β~=logπ+γ+3log2logq+O((logq)2)\tilde{\beta}=\frac{\log\pi+\gamma+3\log 2}{\log q}+O((\log q)^{-2}), as required.

Remark 3.

We briefly discuss a related proportion result. For θ[0,π]\theta\in[0,\pi], consider the class of admissible functions

(29) Lθ(x):={12θsin(θ(1|x|))+12(1|x|)cos(θx)12θsinθ+12if x[1,1],0otherwiseL^{\theta}(x):=\begin{cases}\dfrac{\frac{1}{2\theta}\sin(\theta(1-|x|))+\frac{1}{2}(1-|x|)\cos(\theta x)}{\frac{1}{2\theta}\sin\theta+\frac{1}{2}}&\text{if\>\>$x\in[-1,1]$},\\ 0&\text{otherwise}\end{cases}

with non-negative Fourier transform

Lθ^(t)=(sin(12(θt))θt+sin(12(θ+t))θ+t)212θsinθ+12.\widehat{L^{\theta}}(t)=\dfrac{\left(\frac{\sin(\frac{1}{2}(\theta-t))}{\theta-t}+\frac{\sin(\frac{1}{2}(\theta+t))}{\theta+t}\right)^{2}}{\frac{1}{2\theta}\sin\theta+\frac{1}{2}}.

Note that L0(x)=H(x)L^{0}(x)=H(x) and Lπ(x)=K(x)L^{\pi}(x)=K(x) with HH as in (23) and KK as in (27). From the above argument we observe that for 0β1/40\leq\beta\leq 1/4,

lim infq1ϕ(q)1#{χχ0:|γχ|logq2π>β}\displaystyle\liminf_{\begin{subarray}{c}q\to\infty\end{subarray}}\frac{1}{\phi(q)-1}\#\left\{\chi\neq\chi_{0}:\frac{|\gamma_{\chi}|\log q}{2\pi}>\beta\right\}\geq 112maxθ[0,π]Lθ^(4πβ)\displaystyle 1-\frac{1}{2\cdot\max_{\theta\in[0,\pi]}\widehat{L^{\theta}}(4\pi\beta)}
=\displaystyle= 111+sin(4πβ)4πβ.\displaystyle 1-\frac{1}{1+\frac{\sin(4\pi\beta)}{4\pi\beta}}.

When β=0\beta=0 and 1/41/4, this corresponds to Theorems 4 and  6, respectively.

4.5. Proportion of χmodq\chi\bmod q having small zeros: A first improvement

Before proving Theorem 7, we describe a quick way of improving on (11) for large β\beta via a simple change of test function. Hughes and Rudnick’s proof of (11) relies on the following result:

Lemma 13.

Let f(x)f(x) be a real, even function with f(x)(1+|x|)1ϵf(x)\ll(1+|x|)^{-1-\epsilon} such that f^\widehat{f} is supported in [2π,2π][-2\pi,2\pi] and that f^(0)<0\widehat{f}(0)<0. If f(x)0f(x)\leq 0 for |x|β|x|\leq\beta with β>1/2\beta>1/2 and f(x)0f(x)\geq 0 otherwise, then

(30) lim infqqprime1q2#{χχ0:|γχ|logq2π<β}12π2π|t|f^(t)2dt4π2f^(0)2.\liminf_{\begin{subarray}{c}q\to\infty\\ q\>\text{prime}\end{subarray}}\frac{1}{q-2}\#\left\{\chi\neq\chi_{0}:\frac{|\gamma_{\chi}|\log q}{2\pi}<\beta\right\}\geq 1-\frac{\int_{-2\pi}^{2\pi}|t|\widehat{f}(t)^{2}\,\mathrm{d}t}{4\pi^{2}\widehat{f}(0)^{2}}.
Proof.

See the proof of [HR03, Theorem 8.3]. For consistency of notation we modified the expression on the right-hand side due to a slight difference between definitions of the Fourier transform (see (15)). ∎

The test function ff they chose to work with is essentially Fα^-\widehat{F^{\alpha}} where α\alpha is appropriately determined in terms of β\beta. Here we propose a different option. Recall from (18) that B[β,β](x)=12(B(x+β)+B(x+β))B^{-}_{[-\beta,\beta]}(x)=\frac{1}{2}(B^{-}(x+\beta)+B^{-}(-x+\beta)) and take

gβ(x):=B[β,β](x),β>0,g_{\beta}(x):=-B^{-}_{[-\beta,\beta]}(x),\hskip 28.45274pt\beta\in\mathbb{Z}_{>0},

which is given by the explicit expression

(31) gβ(x)=(sin(πx)π)2(2βx2β2n=12β11(xβ+n)2).g_{\beta}(x)=\left(\frac{\sin(\pi x)}{\pi}\right)^{2}\left(\frac{2\beta}{x^{2}-\beta^{2}}-\sum_{n=1}^{2\beta-1}\frac{1}{(x-\beta+n)^{2}}\right).

(Note that restricting β\beta to integers would not weaken the result asymptotically considering the monotonicity of the proportion on the left-hand side of (30).) See §2 for a brief discussion of some essential properties of gβ(x)g_{\beta}(x), which satisfies all the assumptions in Lemma 13. In particular,

gβ(x)12(sgn(x+β)+sgn(x+β))=0g_{\beta}(x)\geq-\frac{1}{2}(\mathrm{sgn}(x+\beta)+\mathrm{sgn}(-x+\beta))=0

for |x|>β|x|>\beta, and it is clear from (31) that gβ(x)0g_{\beta}(x)\leq 0 for |x|<β|x|<\beta (this is where we need β\beta\in\mathbb{Z}) and that gβ(x)β(1+|x|)2g_{\beta}(x)\ll_{\beta}(1+|x|)^{-2}.

Now, since (sgn(x)B(x))dx=1\int_{-\infty}^{\infty}(\mathrm{sgn}(x)-B^{-}(x))\,\mathrm{d}x=1, we have

gβ^(0)=\displaystyle\widehat{g_{\beta}}(0)= 12(sgn(x+β)B(x+β)+sgn(x+β)B(x+β)\displaystyle\frac{1}{2}\int_{-\infty}^{\infty}\big{(}\mathrm{sgn}(x+\beta)-B^{-}(x+\beta)+\mathrm{sgn}(-x+\beta)-B^{-}(-x+\beta)
2𝟙[β,β])dx\displaystyle\hskip 56.9055pt-2\cdot\mathbbm{1}_{[-\beta,\beta]}\big{)}\,\mathrm{d}x
=\displaystyle= 12β\displaystyle 1-2\beta

and

gβ^(t)=\displaystyle\widehat{g_{\beta}}(t)= ββeitxdx+(gβ(x)+𝟙[β,β](x))eitxdx=2sin(βt)t+O(1)\displaystyle-\int_{-\beta}^{\beta}e^{itx}\,\mathrm{d}x+\int_{-\infty}^{\infty}\left(g_{\beta}(x)+\mathbbm{1}_{[-\beta,\beta]}(x)\right)e^{itx}\,\mathrm{d}x=-\frac{2\sin(\beta t)}{t}+O(1)

for any t[2π,2π]t\in[-2\pi,2\pi]. Hence

2π2π|t|gβ^(t)2dt\displaystyle\int_{-2\pi}^{2\pi}|t|\widehat{g_{\beta}}(t)^{2}\,\mathrm{d}t\ll 2π2πsin(βt)2|t|dt+O(1)01/ββ2tdt+1/β2πdtt+O(1)\displaystyle\int_{-2\pi}^{2\pi}\frac{\sin(\beta t)^{2}}{|t|}\,\mathrm{d}t+O(1)\ll\int_{0}^{1/\beta}\beta^{2}t\,\mathrm{d}t+\int_{1/\beta}^{2\pi}\frac{\,\mathrm{d}t}{t}+O(1)
\displaystyle\ll logβ+1,\displaystyle\log\beta+1,

and it follows by Lemma 13 that

lim infqqprime1q2#{χχ0:|γχ|logq2π<β}1O(logβ+1β2)\liminf_{\begin{subarray}{c}q\to\infty\\ q\>\text{prime}\end{subarray}}\frac{1}{q-2}\#\left\{\chi\neq\chi_{0}:\frac{|\gamma_{\chi}|\log q}{2\pi}<\beta\right\}\geq 1-O\left(\frac{\log\beta+1}{\beta^{2}}\right)

for β1\beta\geq 1. This is already strong enough to show that the proportion converges to 1 as β\beta\to\infty.

4.6. Proportion of χmodq\chi\bmod q having small zeros: proof of Theorem 7

Let qq be a large prime. For β1/4\beta\geq 1/4, let 𝒬β={χmodq:|γχ|logq2π<β}\mathcal{Q}_{\beta}=\{\chi\bmod q:\frac{|\gamma_{\chi}|\log q}{2\pi}<\beta\} (which is non-empty by Theorem 5) and

T=α+1α1π2πβ/logq=α+1α1logq2β.T=\sqrt{\frac{\alpha+1}{\alpha-1}}\frac{\pi}{2\pi\beta/\log q}=\sqrt{\frac{\alpha+1}{\alpha-1}}\frac{\log q}{2\beta}.

In addition, we suppose that TlogqT\leq\log q. By the Cauchy–Schwarz inequality

(32) #𝒬β(χ𝒬βρχΦ(FTα)(ρχ))2χ𝒬β(ρχΦ(FTα)(ρχ))2.\#\mathcal{Q}_{\beta}\geq\frac{\left(\sum_{\chi\in\mathcal{Q}_{\beta}}\sum_{\rho_{\chi}}\Phi(F^{\alpha}_{T})(\rho_{\chi})\right)^{2}}{\sum_{\chi\in\mathcal{Q}_{\beta}}\left(\sum_{\rho_{\chi}}\Phi(F^{\alpha}_{T})(\rho_{\chi})\right)^{2}}.

We first bound the numerator from below. The key observation is again that χ𝒬βχχ0ρχΦ(FTα)(ρχ)0\sum_{\begin{subarray}{c}\chi\not\in\mathcal{Q}_{\beta}\\ \chi\neq\chi_{0}\end{subarray}}\sum_{\rho_{\chi}}\Phi(F^{\alpha}_{T})(\rho_{\chi})\leq 0, which implies that

χ𝒬β\displaystyle\sum_{\chi\in\mathcal{Q}_{\beta}} ρχΦ(FTα)(ρχ)\displaystyle\sum_{\rho_{\chi}}\Phi(F^{\alpha}_{T})(\rho_{\chi})
(q2)logqπχχ0Iχ(FTα)2χχ0n=1Re(χ(n))FTα(logn)Λ(n)n\displaystyle\geq(q-2)\log\frac{q}{\pi}-\sum_{\chi\neq\chi_{0}}I_{\chi}(F^{\alpha}_{T})-2\sum_{\chi\neq\chi_{0}}\sum_{n=1}^{\infty}\textrm{Re}(\chi(n))F^{\alpha}_{T}(\log n)\frac{\Lambda(n)}{\sqrt{n}}
=(q2)logq+O(q)\displaystyle=(q-2)\log q+O(q)

where we applied Lemma 12. On the other hand, the denominator is at most

χχ0\displaystyle\sum_{\chi\neq\chi_{0}} (ρχΦ(FTα)(ρχ))2\displaystyle\left(\sum_{\rho_{\chi}}\Phi(F^{\alpha}_{T})(\rho_{\chi})\right)^{2}
(33) =(q2)(logq+O(1))24(logq+O(1))χχ0n=1Re(χ(n))FTα(logn)Λ(n)n\displaystyle=(q-2)(\log q+O(1))^{2}-4(\log q+O(1))\sum_{\chi\neq\chi_{0}}\sum_{n=1}^{\infty}\textrm{Re}(\chi(n))F^{\alpha}_{T}(\log n)\frac{\Lambda(n)}{\sqrt{n}}
+4χχ0n1,n2Re(χ(n1))Re(χ(n2))FTα(logn1)FTα(logn2)Λ(n1)Λ(n2)n1n2.\displaystyle\hskip 28.45274pt+4\sum_{\chi\neq\chi_{0}}\sum_{n_{1},n_{2}}\textrm{Re}(\chi(n_{1}))\textrm{Re}(\chi(n_{2}))F^{\alpha}_{T}(\log n_{1})F^{\alpha}_{T}(\log n_{2})\frac{\Lambda(n_{1})\Lambda(n_{2})}{\sqrt{n_{1}n_{2}}}.

The second term on the right-hand side contributes αqlogq\ll\alpha\sqrt{q}\log q by Lemma 12 again, and according to (26) the third term is

χχ0n1,n2\displaystyle\sum_{\chi\neq\chi_{0}}\sum_{n_{1},n_{2}} (χ(n1n2)+χ(n1n2)¯+χ(n1)¯χ(n2)+χ(n1)χ(n2)¯)\displaystyle\left(\chi(n_{1}n_{2})+\overline{\chi(n_{1}n_{2})}+\overline{\chi(n_{1})}\chi(n_{2})+\chi(n_{1})\overline{\chi(n_{2})}\right)
×FTα(logn1)FTα(logn2)Λ(n1)Λ(n2)n1n2\displaystyle\hskip 28.45274pt\times F^{\alpha}_{T}(\log n_{1})F^{\alpha}_{T}(\log n_{2})\frac{\Lambda(n_{1})\Lambda(n_{2})}{\sqrt{n_{1}n_{2}}}
\displaystyle\leq 2(q2)n1n21modqFTα(logn1)FTα(logn2)Λ(n1)Λ(n2)n1n2\displaystyle 2(q-2)\sum_{n_{1}n_{2}\equiv 1\bmod q}F^{\alpha}_{T}(\log n_{1})F^{\alpha}_{T}(\log n_{2})\frac{\Lambda(n_{1})\Lambda(n_{2})}{\sqrt{n_{1}n_{2}}}
+2(q2)n1n2modqFTα(logn1)FTα(logn2)Λ(n1)Λ(n2)n1n2\displaystyle\hskip 28.45274pt+2(q-2)\sum_{n_{1}\equiv n_{2}\bmod q}F^{\alpha}_{T}(\log n_{1})F^{\alpha}_{T}(\log n_{2})\frac{\Lambda(n_{1})\Lambda(n_{2})}{\sqrt{n_{1}n_{2}}}
=:\displaystyle=: I1+I2.\displaystyle I_{1}+I_{2}.

We treat I1I_{1} and I2I_{2} as follows. First,

I1\displaystyle I_{1}\ll 2(q2)ke2T1qp1p2=kq+11p1,p2eTFTα(logp1)FTα(logp2)logp1logp2kq+1\displaystyle 2(q-2)\sum_{k\leq\frac{e^{2T}-1}{q}}\sum_{\begin{subarray}{c}p_{1}p_{2}=kq+1\\ 1\leq p_{1},p_{2}\leq e^{T}\end{subarray}}F^{\alpha}_{T}(\log p_{1})F^{\alpha}_{T}(\log p_{2})\frac{\log p_{1}\log p_{2}}{\sqrt{kq+1}}
\displaystyle\ll α2qke2Tq(logkq)2kmaxloga+logb=log(kq+1)0loga,logbT(1logaT)(1logbT)\displaystyle\alpha^{2}\sqrt{q}\sum_{k\leq\frac{e^{2T}}{q}}\frac{(\log kq)^{2}}{\sqrt{k}}\max_{\begin{subarray}{c}\log a+\log b=\log(kq+1)\\ 0\leq\log a,\log b\leq T\end{subarray}}\left(1-\frac{\log a}{T}\right)\left(1-\frac{\log b}{T}\right)
\displaystyle\ll α2qke2Tq(logkq)2(1logkq2T)2k\displaystyle\alpha^{2}\sqrt{q}\sum_{k\leq\frac{e^{2T}}{q}}\frac{(\log kq)^{2}\left(1-\frac{\log kq}{2T}\right)^{2}}{\sqrt{k}}
\displaystyle\ll α2eT\displaystyle\alpha^{2}e^{T}
\displaystyle\leq α2q.\displaystyle\alpha^{2}q.

where we used (20) in the second step. Next observe that if n1n2modqn_{1}\equiv n_{2}\bmod q and n1,n2eTqn_{1},n_{2}\leq e^{T}\leq q, then we must have n1=n2n_{1}=n_{2}. Hence

I2=\displaystyle I_{2}= 2(q2)n1eTFTα(logn1)2Λ(n1)2n1\displaystyle 2(q-2)\sum_{n_{1}\leq e^{T}}F^{\alpha}_{T}(\log n_{1})^{2}\frac{\Lambda(n_{1})^{2}}{n_{1}}
=\displaystyle= 2(q2)peTFTα(logp)2(logp)2p+2(q2)pkeTk2FTα(klogp)2(logp)2pk\displaystyle 2(q-2)\sum_{p\leq e^{T}}F^{\alpha}_{T}(\log p)^{2}\frac{(\log p)^{2}}{p}+2(q-2)\sum_{\begin{subarray}{c}p^{k}\leq e^{T}\\ k\geq 2\end{subarray}}F^{\alpha}_{T}(k\log p)^{2}\frac{(\log p)^{2}}{p^{k}}
=\displaystyle= 2(q2)2eTFTα(logx)2logxxdx+O(α2q)\displaystyle 2(q-2)\int_{2}^{e^{T}}F^{\alpha}_{T}(\log x)^{2}\frac{\log x}{x}\,\mathrm{d}x+O(\alpha^{2}q)
=\displaystyle= 2(q2)T2log2T1uFα(u)2du+O(α2q).\displaystyle 2(q-2)T^{2}\int_{\frac{\log 2}{T}}^{1}uF^{\alpha}(u)^{2}\,\mathrm{d}u+O(\alpha^{2}q).

Inserting these estimates into (32) gives

#𝒬β((q2)logq+O(q))2(q2)(logq)2+O(α2q)+(q2)T2σ(Fα)\#\mathcal{Q}_{\beta}\geq\dfrac{\left((q-2)\log q+O(q)\right)^{2}}{(q-2)(\log q)^{2}+O(\alpha^{2}q)+(q-2)T^{2}\sigma(F^{\alpha})}

where σ(Fα)=11|u|Fα(u)2du\sigma(F^{\alpha})=\int_{-1}^{1}|u|F^{\alpha}(u)^{2}\,\mathrm{d}u, provided that T=α+1α1logq2βlogqT=\sqrt{\frac{\alpha+1}{\alpha-1}}\frac{\log q}{2\beta}\leq\log q, which holds for all large enough α\alpha so long as β>1/2\beta>1/2. We therefore have

lim infqqprime#𝒬βq2\displaystyle\liminf_{\begin{subarray}{c}q\to\infty\\ q\>\text{prime}\end{subarray}}\frac{\#\mathcal{Q}_{\beta}}{q-2}\geq 11+14β2minα+1α12βα+1α1σ(Fα)\displaystyle\frac{1}{1+\frac{1}{4\beta^{2}}\min_{\sqrt{\frac{\alpha+1}{\alpha-1}}\leq 2\beta}\frac{\alpha+1}{\alpha-1}\sigma(F^{\alpha})}
=\displaystyle= 11+14β2minα4β2+14β21α+1α16α2+π2312π2.\displaystyle\frac{1}{1+\frac{1}{4\beta^{2}}\min_{\alpha\geq\frac{4\beta^{2}+1}{4\beta^{2}-1}}\frac{\alpha+1}{\alpha-1}\frac{6\alpha^{2}+\pi^{2}-3}{12\pi^{2}}}.

The function f(α)=α+1α16α2+π2312π2f(\alpha)=\frac{\alpha+1}{\alpha-1}\frac{6\alpha^{2}+\pi^{2}-3}{12\pi^{2}} attains its minimum on (1,)(1,\infty) at α0=1.8652\alpha_{0}=1.8652\cdots with f(α0)=0.7757f(\alpha_{0})=0.7757\cdots, and is increasing on [α0,)[\alpha_{0},\infty). It thus follows that

lim infqqprime#𝒬βq2{11+14β2f(4β2+14β21)if  1/2<β<α0+1α01/2,11+f(α0)4β2ifβα0+1α01/2,\liminf_{\begin{subarray}{c}q\to\infty\\ q\>\text{prime}\end{subarray}}\frac{\#\mathcal{Q}_{\beta}}{q-2}\geq\begin{cases}\dfrac{1}{1+\frac{1}{4\beta^{2}}f(\frac{4\beta^{2}+1}{4\beta^{2}-1})}&\text{if}\>\>1/2<\beta<\sqrt{\frac{\alpha_{0}+1}{\alpha_{0}-1}}/2,\\ \dfrac{1}{1+\frac{f(\alpha_{0})}{4\beta^{2}}}&\text{if}\>\>\beta\geq\sqrt{\frac{\alpha_{0}+1}{\alpha_{0}-1}}/2,\end{cases}

which concludes the proof.

Remark 4.

We can remove the restriction to prime moduli at the cost of a weaker bound. For a general qq, the second term in (4.6) has to be replaced by

4χχ0(log(con(χ))+O(1))n=1Re(χ(n))FTα(logn)Λ(n)n,-4\sum_{\chi\neq\chi_{0}}(\log(\mathrm{con}(\chi))+O(1))\sum_{n=1}^{\infty}\textrm{Re}(\chi(n))F^{\alpha}_{T}(\log n)\frac{\Lambda(n)}{\sqrt{n}},

which is not necessarily negligible since orthogonality is no longer applicable. However, this is at most

4ϕ(q)(logq+O(1))2(14ϕ(q)T2σ(Fα)+O(α2ϕ(q)))\displaystyle 4\sqrt{\phi(q)(\log q+O(1))^{2}\left(\frac{1}{4}\phi(q)T^{2}\sigma(F^{\alpha})+O(\alpha^{2}\phi(q))\right)}
(2+o(1))ϕ(q)(logq)Tσ(Fα)\displaystyle\hskip 142.26378pt\leq(2+o(1))\phi(q)(\log q)T\sqrt{\sigma(F^{\alpha})}

by the Cauchy–Schwarz equality and our treatment of the third term in (4.6). Arguing as before with minor variations, we deduce that

lim infq#𝒬βϕ(q)1\displaystyle\liminf_{\begin{subarray}{c}q\to\infty\end{subarray}}\frac{\#\mathcal{Q}_{\beta}}{\phi(q)-1}\geq 11+minα+1α12β{14β2α+1α1σ(Fα)+1βα+1α1σ(Fα)}.\displaystyle\frac{1}{1+\min_{\sqrt{\frac{\alpha+1}{\alpha-1}}\leq 2\beta}\left\{\frac{1}{4\beta^{2}}\frac{\alpha+1}{\alpha-1}\sigma(F^{\alpha})+\frac{1}{\beta}\sqrt{\frac{\alpha+1}{\alpha-1}\sigma(F^{\alpha})}\right\}}.

5. Some remarks on quadratic Dirichlet LL-functions

5.1. Non-vanishing at the central point

It is natural to consider extending our previous arguments to the family of real primitive characters. For a fundamental discriminant dd, write χd():=(d)\chi_{d}(\cdot):=\left(\frac{d}{\cdot}\right) for the associated Kronecker symbol. Unconditionally, Soundararajan [Sou00] showed that L(12,χd)0L(\frac{1}{2},\chi_{d})\neq 0 for at least 78\frac{7}{8} of the dd’s by studying the mollified moments of L(12,χd)L(\frac{1}{2},\chi_{d}); at around the same time, Özlük and Snyder [OS99] proved a proportion of 1516\frac{15}{16} assuming GRH by establishing the one-level density for the symplectic family of quadratic Dirichlet LL-functions for test functions ff with supp(f^)[2,2]\mathrm{supp}\>(\widehat{f})\in[-2,2] 333The one-level density in this setting reveals that the low-lying zeros of quadratic Dirichlet LL-functions are sparser than on average, which explains why these non-vanishing results are far better than what we can prove in the case where all Dirichlet LL-functions are included., which was independently discovered by Katz and Sarnak [KS99]. We demonstrate below a simple way of obtaining the less optimal proportion 34\frac{3}{4} under GRH (which did not seem to have ever appeared in the literature prior to the aforementioned works) that uses only the explicit formula and the following mean square estimate for character sums due to Jutila [Jut73, Corollary to Theorem 1]:

(34) nNn≠̸||d|Dχd(n)|2ND(logN)10,\sum_{\begin{subarray}{c}n\leq N\\ n\not\neq\square\end{subarray}}\bigg{|}\underset{|d|\leq D}{{\sum}^{*}}\chi_{d}(n)\bigg{|}^{2}\ll ND(\log N)^{10},
444The exponent 10 on logN\log N is not optimal since [Jut73, Theorem 1] has been improved by Armon [Arm99].

where the outer sum runs over non-square positive integers nNn\leq N and the inner sum {\sum}^{*} runs over fundamental discriminants dd with |d|D|d|\leq D, with the implied constant being absolute. On average this saves an n1/2n^{1/2} factor from the pointwise bound

(35) ||d|Dχd(n)|D1/2n1/4(logn)1/2\bigg{|}\underset{|d|\leq D}{{\sum}^{*}}\chi_{d}(n)\bigg{|}\ll D^{1/2}n^{1/4}(\log n)^{1/2}

for nn\neq\square (see, e.g., [Ayo63, pp. 325]) 555We point out a small misuse of the Pólya–Vinogradov inequality in §2.3.1 and §5.3.3 of this expository article [Con05].. Under GRH, (35) can be improved to

(36) ||d|Dχd(n)|ϵD1/2+ϵnϵ\bigg{|}\underset{|d|\leq D}{{\sum}^{*}}\chi_{d}(n)\bigg{|}\ll_{\epsilon}D^{1/2+\epsilon}n^{\epsilon}

for any ϵ>0\epsilon>0 (see [DM24, Lemma 1]), but this yields no advantage over (34) in our setting. Using (35) Özlük and Snyder showed in their earlier work [OS93] that the support for the one-level density can be taken as [23,23][-\frac{2}{3},\frac{2}{3}]. Using (34) Rubinstein [Rub01] extended the support for the nn-level density to [1,1][-1,1] for all n1n\geq 1. The key idea that allowed Özlük and Snyder to go beyond this range in [OS99] is applying the Poisson summation formula to a smoothed character sum.

We mimick the proof of Theorem 4 to obtain the proportion 34\frac{3}{4}. Denote by DD^{*} the number of fundamental discriminants dd with |d|D|d|\leq D. It is well known that D6π2DD^{*}\sim\frac{6}{\pi^{2}}D. Applying (14) to HTH_{T} and averaging over the set of χd\chi_{d}’s, we find that

TD|d|Dnχd\displaystyle\frac{T}{D^{*}}\underset{|d|\leq D}{{\sum}^{*}}n_{\chi_{d}}\leq 1D|d|Dlog|d|π2DneT|d|Dχd(n)HT(logn)Λ(n)n+O(1)\displaystyle\frac{1}{D^{*}}\underset{|d|\leq D}{{\sum}^{*}}\log\frac{|d|}{\pi}-\frac{2}{D^{*}}\sum_{n\leq e^{T}}\underset{|d|\leq D}{{\sum}^{*}}\chi_{d}(n)H_{T}(\log n)\frac{\Lambda(n)}{\sqrt{n}}+O(1)
\displaystyle\leq logD2DpeT|d|Dχd(p)HT(logp)logpp\displaystyle\log D-\frac{2}{D^{*}}\sum_{p\leq e^{T}}\underset{|d|\leq D}{{\sum}^{*}}\chi_{d}(p)H_{T}(\log p)\frac{\log p}{\sqrt{p}}
2DpeT/2|d|Dχd(p2)HT(2logp)logpp+O(1).\displaystyle\hskip 56.9055pt-\frac{2}{D^{*}}\sum_{p\leq e^{T/2}}\underset{|d|\leq D}{{\sum}^{*}}\chi_{d}(p^{2})H_{T}(2\log p)\frac{\log p}{p}+O(1).

By the Cauchy–Schwarz inequality and Jutila’s result (34), the sum over primes contributes

(37) 2DpeT||d|Dχd(p)|2peT(HT(logp)logpp)2eT/2T6D,\ll\frac{2}{D^{*}}\sqrt{\sum_{p\leq e^{T}}\bigg{|}\underset{|d|\leq D}{{\sum}^{*}}\chi_{d}(p)\bigg{|}^{2}}\sqrt{\sum_{p\leq e^{T}}\left(H_{T}(\log p)\frac{\log p}{\sqrt{p}}\right)^{2}}\ll\frac{e^{T/2}T^{6}}{\sqrt{D}},

while the sum over squares of primes is

peT/2(2+O(D1))HT(2logp)logpp=\displaystyle-\sum_{p\leq e^{T/2}}\left(2+O({D^{*}}^{-1})\right)H_{T}(2\log p)\frac{\log p}{p}= (1+O(D1))T01H(x)dx+O(1)\displaystyle-\left(1+O(D^{-1})\right)T\int_{0}^{1}H(x)\,\mathrm{d}x+O(1)
=\displaystyle= T2+O(1+TD1).\displaystyle-\frac{T}{2}+O(1+TD^{-1}).

Hence we have

1D|d|DnχdlogDT12+O(eT/2T5D+1T)12+O(loglogDlogD)\frac{1}{D^{*}}\underset{|d|\leq D}{{\sum}^{*}}n_{\chi_{d}}\leq\frac{\log D}{T}-\frac{1}{2}+O\left(\frac{e^{T/2}T^{5}}{\sqrt{D}}+\frac{1}{T}\right)\leq\frac{1}{2}+O\left(\frac{\log\log D}{\log D}\right)

by choosing T=logDcloglogDT=\log D-c\log\log D for some suitable c>0c>0. It follows that L(12,χd)=0L(\frac{1}{2},\chi_{d})=0 for at most 14+o(1)\frac{1}{4}+o(1) of the dd’s with |d|D|d|\leq D since nχdn_{\chi_{d}} is always even.

By combining the preceding argument with the proof of Theorem 6 one can deduce that

(38) lim infDmaxd[D,2D]|γχd|logD2π0.7229.\liminf_{D\to\infty}\max_{d\in[D,2D]}\frac{|\gamma_{\chi_{d}}|\log D}{2\pi}\geq 0.7229\cdots.

The task is to maximize β>0\beta>0 subject to the condition that for all large DD,

(39) logDTK^(0)2+O(eT/2T5D+1T)min|t|2πβTlogDK^(t)<2\frac{\frac{\log D}{T}-\frac{\widehat{K}(0)}{2}+O\left(\frac{e^{T/2}T^{5}}{\sqrt{D}}+\frac{1}{T}\right)}{\min_{|t|\leq\frac{2\pi\beta T}{\log D}}\widehat{K}(t)}<2
666This is analogous to the condition (28) in the proof of Theorem 6; we have 2 instead of 1 on the right-hand side because of the symmetry of zeros.

where K^(0)2=4π2\frac{\widehat{K}(0)}{2}=\frac{4}{\pi^{2}}. We find by numerical computation that β\beta can be taken as large as 0.72290.7229\cdots when T=0.83logDT=0.83\log D.

5.2. Low-lying zeros of L(s,χd)L(s,\chi_{d}) for localized dd’s

It is worth pointing out that (35) has the advantage that the shape of the estimate stays unchanged if we sum dd over an arbitrary interval of the same length, whereas it is not clear from the proofs of (34) and (36) how to remove the dependency on the initial interval. As a consequence, for D<DD^{\prime}<D we have

peT|d[DD,D]χd(p)|2min{DeTT10,De3T/2T}.\sum_{p\leq e^{T}}\bigg{|}\underset{d\in[D-D^{\prime},D]}{{\sum}^{*}}\chi_{d}(p)\bigg{|}^{2}\ll\min\left\{De^{T}T^{10},D^{\prime}e^{3T/2}T\right\}.

This observation can be useful when we average only over those χd\chi_{d}’s with |d||d| close to DD. For example, since in (39) the largest value of β\beta is achieved when T0.83logDT\approx 0.83\log D, this implies that in (38) we can shorten the interval in which the maximum occurs to d[DD,D]d\in[D-D^{\prime},D] where D=D(1+0.83)/2=D0.915D^{\prime}=D^{(1+0.83)/2}=D^{0.915}. Working with H(t)H(t) (which satisfies H^(0)=1\widehat{H}(0)=1) instead in (39) and taking T2logD5T\to\frac{2\log D}{5} and β0\beta\to 0, we see that there must exist some d[DD3/5+ϵ,D]d\in[D-D^{3/5+\epsilon},D] such that L(12,χd)0L(\frac{1}{2},\chi_{d})\neq 0. In general, for each β[0,0.7229]\beta\in[0,0.7229] one can compute an upper bound for the length of the shortest interval [DDλ(β),D][D-D^{\lambda(\beta)},D] in which there exists some dd with |γχd|logD2π>β\frac{|\gamma_{\chi_{d}}|\log D}{2\pi}>\beta. In particular, λ(0)3/5\lambda(0)\leq 3/5 and λ(0.7229)0.915\lambda(0.7229)\leq 0.915 as mentioned above. For intermediate β\beta one can work with LθL^{\theta} (see (29)) for the optimal choice of θ[0,π]\theta\in[0,\pi].

In the opposite direction, we find by applying this argument to FαF^{\alpha} that

lim supDmind[DDa,D]|γχd|logD2π{34a3π24a3π24a2,0<a3/4,14a2π24a2π24a22,3/4a1.\limsup_{D\to\infty}\min_{d\in[D-D^{a},D]}\frac{|\gamma_{\chi_{d}}|\log D}{2\pi}\leq\begin{cases}\dfrac{3}{4a}\sqrt{\dfrac{\frac{3\pi^{2}}{4a}}{\frac{3\pi^{2}}{4a}-2}},&0<a\leq 3/4,\\ \dfrac{1}{4a-2}\sqrt{\dfrac{\frac{\pi^{2}}{4a-2}}{\frac{\pi^{2}}{4a-2}-2}},&3/4\leq a\leq 1.\end{cases}

Of course this bound is not optimal for aa close to 1.

5.3. Proportion of χd\chi_{d}’s having small zeros

We now discuss an analogue of Theorem 7 for quadratic Dirichlet LL-functions by modifying the argument in §4.6. Denote by DD^{*} the number of fundamental discriminants in [D,2D][D,2D], and for β>1/2\beta>1/2 let 𝒟β={d[D,2D]:|γχd|logD2π<β}\mathcal{D}_{\beta}=\{d\in[D,2D]:\frac{|\gamma_{\chi_{d}}|\log D}{2\pi}<\beta\}. Further let T=α+1α1logD2βT=\sqrt{\frac{\alpha+1}{\alpha-1}}\frac{\log D}{2\beta} and suppose that T<logDT<\log D. We then proceed with a Cauchy–Schwarz argument as before. On the one hand, by the positivity trick we have

d𝒟βρdΦ(FTα)(ρd)\displaystyle\sum_{d\in\mathcal{D}_{\beta}}\sum_{\rho_{d}}\Phi(F^{\alpha}_{T})(\rho_{d})\geq DlogD2d[D,2D]n=1χd(n)FTα(logn)Λ(n)n+O(D)\displaystyle D^{*}\log D-2\underset{d\in[D,2D]}{{\sum}^{*}}\sum_{n=1}^{\infty}\chi_{d}(n)F^{\alpha}_{T}(\log n)\frac{\Lambda(n)}{\sqrt{n}}+O(D^{*})
=\displaystyle= DlogDDTFTα^(0)/2+O(D)\displaystyle D^{*}\log D-D^{*}T\widehat{F^{\alpha}_{T}}(0)/2+O(D^{*})
=\displaystyle= DlogD(1α+1α1(α+1)π2β+o(1)),\displaystyle D^{*}\log D\left(1-\frac{\sqrt{\frac{\alpha+1}{\alpha-1}}(\alpha+1)}{\pi^{2}\beta}+o(1)\right),

where the second term on the penultimate line corresponds to the sum over squares of primes, while the contribution from primes is negligible (see (37)). Note, however, that the right-hand side needs to be non-negative in order to keep the inequality valid upon squaring both sides. That is, we need α+1α1(α+1)<π2β\sqrt{\frac{\alpha+1}{\alpha-1}}(\alpha+1)<\pi^{2}\beta.

On the other hand,

d𝒟β(ρdΦ(FTα)(ρd))2=D(logD)2+O(DlogD)4logDd[D,2D]n=1χd(n)FTα(logn)Λ(n)n+4d[D,2D]n1,n2χd(n1n2)FTα(logn1)FTα(logn2)Λ(n1)Λ(n2)n1n2=D(logD)2D(logD)TFTα^(0)+O(DlogD)+Δ\begin{split}\sum_{d\in\mathcal{D}_{\beta}}&\left(\sum_{\rho_{d}}\Phi(F^{\alpha}_{T})(\rho_{d})\right)^{2}\\ &=D^{*}(\log D)^{2}+O(D^{*}\log D)-4\log D\underset{d\in[D,2D]}{{\sum}^{*}}\sum_{n=1}^{\infty}\chi_{d}(n)F^{\alpha}_{T}(\log n)\frac{\Lambda(n)}{\sqrt{n}}\\ &\hskip 28.45274pt+4\underset{d\in[D,2D]}{{\sum}^{*}}\sum_{n_{1},n_{2}}\chi_{d}(n_{1}n_{2})F^{\alpha}_{T}(\log n_{1})F^{\alpha}_{T}(\log n_{2})\frac{\Lambda(n_{1})\Lambda(n_{2})}{\sqrt{n_{1}n_{2}}}\\ &=D^{*}(\log D)^{2}-D^{*}(\log D)T\widehat{F^{\alpha}_{T}}(0)+O(D^{*}\log D)+\Delta\end{split}

where Δ\Delta denotes the sum over n1n_{1} and n2n_{2}. Since T<logDT<\log D, it follows from the work of Gao [Gao14] on the nn-level density (here we only need the special case n=2n=2) that

(40) limDΔDT2=211|u|Fα(u)2du\lim_{D\to\infty}\frac{\Delta}{D^{*}T^{2}}=2\int_{-1}^{1}|u|F^{\alpha}(u)^{2}\,\mathrm{d}u

(plug n=2n=2 into formula (3.13) in [Gao14, §3.6] 777Technically we need T(1cloglogDlogD)logDT\leq(1-c\frac{\log\log D}{\log D})\log D for some absolute constant c>0c>0 in order for (40) to hold, but this has no bearing on our discussion.). Roughly speaking, only the contribution from the diagonal terms p1=p2p_{1}=p_{2} matters. Thus

(5.3)D(logD)2(12α+1α1(α+1)π2β+12β2α+1α16α2+π2312π2+o(1)).\eqref{second term Cauchy-Schwarz}\leq D^{*}(\log D)^{2}\left(1-\frac{2\sqrt{\frac{\alpha+1}{\alpha-1}}(\alpha+1)}{\pi^{2}\beta}+\frac{1}{2\beta^{2}}\frac{\alpha+1}{\alpha-1}\frac{6\alpha^{2}+\pi^{2}-3}{12\pi^{2}}+o(1)\right).

The two required conditions α+1α1(α+1)<π2β\sqrt{\frac{\alpha+1}{\alpha-1}}(\alpha+1)<\pi^{2}\beta and α+1α1<2β\sqrt{\frac{\alpha+1}{\alpha-1}}<2\beta (so that T<logDT<\log D) can be simultaneously satisfied if and only if β>π2π24=0.648\beta>\frac{\pi}{2\sqrt{\pi^{2}-4}}=0.648\cdots. Now we collect the previous estimates and reach the following conclusion:

Theorem 14.

Assume GRH. For β0.649\beta\geq 0.649,

(41) lim infD#𝒟βDsupα>1α+1α1(α+1)<π2βα+1α1<2β(1α+1α1(α+1)π2β)212α+1α1(α+1)π2β+(6α2+π23)(α+1)24π2β2(α1).\liminf_{D\to\infty}\frac{\#\mathcal{D}_{\beta}}{D^{*}}\geq\sup_{\begin{subarray}{c}\alpha>1\\ \sqrt{\frac{\alpha+1}{\alpha-1}}(\alpha+1)<\pi^{2}\beta\\ \sqrt{\frac{\alpha+1}{\alpha-1}}<2\beta\end{subarray}}\dfrac{\left(1-\dfrac{\sqrt{\frac{\alpha+1}{\alpha-1}}(\alpha+1)}{\pi^{2}\beta}\right)^{2}}{1-\dfrac{2\sqrt{\frac{\alpha+1}{\alpha-1}}(\alpha+1)}{\pi^{2}\beta}+\dfrac{(6\alpha^{2}+\pi^{2}-3)(\alpha+1)}{24\pi^{2}\beta^{2}(\alpha-1)}}.

Unfortunately we are not able to find a closed form expression for the right-hand side.

Refer to caption
Figure 2. A plot of the right-hand side of (41) for small β\beta.

6. Explicit estimates on |γχ||\gamma_{\chi}|

It might be of interest to establish some effective results for the low-lying zeros of L(s,χ)L(s,\chi). All the numerical computations in this section are carried out on Mathematica. One natural question to ask is: for a given real number t0>0t_{0}>0, what is the least conductor q0q_{0} such that |γχ|t0|\gamma_{\chi}|\leq t_{0} for all qq0q\geq q_{0}? If t0>5/7t_{0}>5/7, the following estimate on the zero-counting function obtained by Bennet et al. provides an unconditional upper bound on q0q_{0}:

Theorem 15 ([Ben+21, Theorem 1.1]).

Let χ\chi be a character with conductor q>1q>1. If t5/7t\geq 5/7 and :=logq(t+2)2π>1.567\ell:=\log\frac{q(t+2)}{2\pi}>1.567, then

|N(t,χ)(tπlogqt2πeχ(1)4)|0.22737+2log(1+)0.5\left|N(t,\chi)-\left(\frac{t}{\pi}\log\frac{qt}{2\pi e}-\frac{\chi(-1)}{4}\right)\right|\leq 0.22737\ell+2\log(1+\ell)-0.5

where N(t,χ)N(t,\chi) counts the number of zeros of L(s,χ)L(s,\chi) with 0<β<10<\beta<1 and |γ|T|\gamma|\leq T.

Take t0=1t_{0}=1 for example. As noted on [Ben+21, pp. 1458], it follows from Theorem 15 that N(1,χ)1N(1,\chi)\geq 1, or |γχ|1|\gamma_{\chi}|\leq 1, when q1.3×1047q\geq 1.3\times 10^{47}. Assuming GRH, we can substantially reduce this number by employing the positivity technique. Indeed, we find that for α=2.6\alpha=2.6,

Iχ(FTα)+2neTFTα(logn)Λ(n)n20.98I_{\chi}(F^{\alpha}_{T})+2\sum_{n\leq e^{T}}F^{\alpha}_{T}(\log n)\frac{\Lambda(n)}{\sqrt{n}}\leq 20.98

for all TT:=α+1α1πT\leq T^{\prime}:=\sqrt{\frac{\alpha+1}{\alpha-1}}\pi. (The choice of α\alpha is nearly optimal.) Thus ρΦ(FTα)(ρ)>0\sum_{\rho}\Phi(F^{\alpha}_{T})(\rho)>0 for all TTT\leq T^{\prime} and qπe20.984.1×109q\geq\pi e^{20.98}\approx 4.1\times 10^{9}. As explained in §2, if |γχ|>1|\gamma_{\chi}|>1, then the sum over zeros must be negative for T=α+1α1π/|γχ|<TT=\sqrt{\frac{\alpha+1}{\alpha-1}}\pi/|\gamma_{\chi}|<T^{\prime}, from which we conclude that |γχ|1|\gamma_{\chi}|\leq 1 when q4.1×109q\geq 4.1\times 10^{9}. Similarly, for t0=5/7t_{0}=5/7, choosing α=2.9\alpha=2.9 gives |γχ|5/7|\gamma_{\chi}|\leq 5/7 for q2.1×1020q\geq 2.1\times 10^{20}.

Next we derive an effective upper bound on |γχ||\gamma_{\chi}| in terms of qq based on the proof of Theorem 1.

Theorem 16.

Assume GRH. For all q1024q\geq 10^{24},

|γχ|π/2(loglogq1.43)loglogqloglogq2.|\gamma_{\chi}|\leq\frac{\pi/2}{(\log\log q-1.43)}\sqrt{\frac{\log\log q}{\log\log q-2}}.
Proof.

We claim that for α3\alpha\geq 3,

(42) Iχ(FTα)min{1.505,8.599T}(α1)+4.228,2neTFTα(logn)Λ(n)n4.156(α1)eT/2T/22.078(α1),\begin{split}&I_{\chi}(F^{\alpha}_{T})\leq\min\bigg{\{}1.505,\frac{8.599}{T}\bigg{\}}\cdot(\alpha-1)+4.228,\\ &2\sum_{n\leq e^{T}}F^{\alpha}_{T}(\log n)\frac{\Lambda(n)}{\sqrt{n}}\leq 4.156(\alpha-1)\frac{e^{T/2}}{T/2}-2.078(\alpha-1),\end{split}

which are explicit versions of Lemmas 9 and 10, respectively. Indeed, recall from the proof of Lemma 9 that

Iχ(FTα)Fα(x)12T02Txex/41exdx+γ+3log2+π2I_{\chi}(F^{\alpha}_{T})\leq F^{\alpha\prime}(x^{*})\frac{1}{2T}\int_{0}^{2T}\frac{xe^{-x/4}}{1-e^{-x}}\,\mathrm{d}x+\gamma+3\log 2+\frac{\pi}{2}

for some x(0,1)x^{*}\in(0,1). It is not hard to show that |Fα(x)|α1|F^{\alpha\prime}(x)|\leq\alpha-1 on [0,1][0,1]. Moreover, the integral converges to 17.19717.197\cdots as TT\to\infty, and for any T>0T>0 the average of the function xex/41ex\frac{xe^{-x/4}}{1-e^{-x}} on [0,2T][0,2T] is bounded by its global maximum 1.5041.504\cdots. This proves the first assertion. The second assertion follows from (2.4) and the bound ψ(x)<1.039x\psi(x)<1.039x due to Rosser and Schoenfeld [RS62, Theorem 12].

Now set T=α+1α1π/|γχ|T=\sqrt{\frac{\alpha+1}{\alpha-1}}\pi/|\gamma_{\chi}| with α=loglogq1\alpha=\log\log q-1, which is 3\geq 3 since q1024q\geq 10^{24}. We then deduce from (42) that

eT/2T/2logqπIχ(FTα)+2.078(α1)4.156(α1)f(q):=logq+0.573(α1)5.3734.156(α1),\displaystyle\frac{e^{T/2}}{T/2}\geq\frac{\log\frac{q}{\pi}-I_{\chi}(F^{\alpha}_{T})+2.078(\alpha-1)}{4.156(\alpha-1)}\geq f(q):=\frac{\log q+0.573(\alpha-1)-5.373}{4.156(\alpha-1)},

and consequently

T2logf(q)+loglogf(q)+log(1+loglogf(q)logf(q)).\frac{T}{2}\geq\log f(q)+\log\log f(q)+\log\left(1+\frac{\log\log f(q)}{\log f(q)}\right).

Using the fact that f(q)logq4.156(loglogq2)f(q)\geq\frac{\log q}{4.156(\log\log q-2)} for loglogq12\log\log q\geq 12 and the inequality x1+xlog(1+x)x\frac{x}{1+x}\leq\log(1+x)\leq x for x>1x>-1, a quick calculation shows that T/2loglogqlog4.156>loglogq1.43T/2\geq\log\log q-\log 4.156>\log\log q-1.43 for loglogq100\log\log q\geq 100. For smaller qq we verify this computationally, thereby completing the proof. ∎

7. Extensions of Theorems 13 to general LL-functions

We conclude by reformulating Theorems 1,  2 and  3 in the more general context. We impose the same assumptions on L(s,π)L(s,\pi), the LL-function under consideration, as in [CCM15, §4]. Suppose that it has an Euler product of the form

L(s,π)=pj=1m(1αj,π(p)ps)1,Re(s)>1L(s,\pi)=\prod_{p}\prod_{j=1}^{m}\left(1-\frac{\alpha_{j,\pi}(p)}{p^{s}}\right)^{-1},\hskip 28.45274pt\textrm{Re}(s)>1

where

|αj,π(p)|pϑ|\alpha_{j,\pi}(p)|\leq p^{\vartheta}

for some constant 0ϑ10\leq\vartheta\leq 1. Define the completed LL-function by

Λ(s,π)=L(s,π)L(s,π)whereL(s,π)=Ns/2j=1mΓ(s+μj)\Lambda(s,\pi)=L_{\infty}(s,\pi)L(s,\pi)\hskip 22.76228pt\text{where}\hskip 5.69046ptL_{\infty}(s,\pi)=N^{s/2}\prod_{j=1}^{m}\Gamma_{\mathbb{R}}(s+\mu_{j})

with N>0N>0, Γ(s)=πs/2Γ(s/2)\Gamma_{\mathbb{R}}(s)=\pi^{-s/2}\Gamma(s/2) and Re(μj)>1\textrm{Re}(\mu_{j})>-1, such that it satisfies the functional equation

Λ(s,π)=ϵΛ(1s¯,π)¯,|ϵ|=1.\Lambda(s,\pi)=\epsilon\overline{\Lambda(1-\overline{s},\pi)},\hskip 28.45274pt|\epsilon|=1.

Suppose further that Λ(s,π)\Lambda(s,\pi) has r(π)r(\pi) poles at s=0s=0 with r(π)mr(\pi)\leq m, so that by the functional equation it has the same number of poles at s=1s=1 . The analytic conductor C(π)C(\pi) of Λ(s,π)\Lambda(s,\pi) is given by

C(π)=Nj=1m(|μj|+3).C(\pi)=N\prod_{j=1}^{m}(|\mu_{j}|+3).

Let FF be an admissible function. Mestre [Mes86, §I.2] established the explicit formula

(43) ρΦ(F)(ρ)=logNπm+r(π)(Φ(F)(0)+Φ(F)(1))j=1mIj(F)2p prime1jm,k1Re(αj,π(p)k)F(klogp)logppk/21<Re(μj)<1/2(Φ(F)(μj)+Φ(F)(1+μj))12Re(μj)=1/2(Φ(F)(μj)+Φ(F)(1+μj))\begin{split}\sum_{\rho}\Phi(F)(\rho)=&\log\frac{N}{\pi^{m}}+r(\pi)\left(\Phi(F)(0)+\Phi(F)(1)\right)-\sum_{j=1}^{m}I_{j}(F)\\ &\hskip 56.9055pt-2\sum_{\begin{subarray}{c}\text{$p$ prime}\\ 1\leq j\leq m,k\geq 1\end{subarray}}\textrm{Re}(\alpha_{j,\pi}(p)^{k})F(k\log p)\frac{\log p}{p^{k/2}}\\ &\hskip 56.9055pt-\sum_{-1<\textrm{Re}(\mu_{j})<-1/2}(\Phi(F)(-\mu_{j})+\Phi(F)(1+\mu_{j}))\\ &\hskip 56.9055pt-\frac{1}{2}\sum_{\textrm{Re}(\mu_{j})=-1/2}(\Phi(F)(-\mu_{j})+\Phi(F)(1+\mu_{j}))\end{split}
888The last two terms in (43) do not appear in the original formula as found in [Mes86, §I.2] because Mestre assumed that Re(μj)0\textrm{Re}(\mu_{j})\geq 0, but only those μj\mu_{j}’s with Re(μj)1/2\textrm{Re}(\mu_{j})\leq-1/2 show up when one moves the line of integration and applies the residue theorem. See also [CCM15, (4.6)] and [IK04, Theorems 5.11 & 5.12].

where the sum runs over zeros of Λ(s,π)\Lambda(s,\pi) with real part 0β10\leq\beta\leq 1, Φ(F)\Phi(F) is from (13), and

Ij(F)=0(F(x/2)e(1/4+Re(μj)/2)x1exexx)dx.I_{j}(F)=\int_{0}^{\infty}\left(\frac{F(x/2)e^{-(1/4+\textrm{Re}(\mu_{j})/2)x}}{1-e^{-x}}-\frac{e^{-x}}{x}\right)\,\mathrm{d}x.

Let |γπ||\gamma_{\pi}| (resp., |γπ~||\widetilde{\gamma_{\pi}}|) and nπn_{\pi} denote the height of the lowest (resp., lowest non-real) non-trivial zero of Λ(s,π)\Lambda(s,\pi) and its order of vanishing at s=12s=\frac{1}{2}, respectively. Assuming RH for Λ(s,π)\Lambda(s,\pi), we apply (43) to FTαF^{\alpha}_{T}, HTH_{T} and GTαG^{\alpha}_{T} as defined in (16), (23) and (24), respectively, and follow the same lines of argument as in §2 and §3. First, we can prove that

j=1mIj(FTα)m(αT+1)eT/2,\displaystyle\sum_{j=1}^{m}I_{j}(F^{\alpha}_{T})\ll m\left(\frac{\alpha}{T}+1\right)e^{T/2},
p prime1jm,k1Re(αj,π(p)k)FTα(klogp)logppk/2\displaystyle\sum_{\begin{subarray}{c}\text{$p$ prime}\\ 1\leq j\leq m,k\geq 1\end{subarray}}\textrm{Re}(\alpha_{j,\pi}(p)^{k})F^{\alpha}_{T}(k\log p)\frac{\log p}{p^{k/2}}\leq mneTFTα(logn)Λ(n)nϑ1/2\displaystyle m\sum_{n\leq e^{T}}F^{\alpha}_{T}(\log n)\Lambda(n)n^{\vartheta-1/2}
\displaystyle\ll mαe(1/2+ϑ)TT,\displaystyle\frac{m\alpha e^{(1/2+\vartheta)T}}{T},

and the last two terms in (43) are both

mTTFTα(x)ex/2dxmαTT(1|x|T)ex/2mαeT/2T.\ll m\int_{-T}^{T}F^{\alpha}_{T}(x)e^{x/2}\,\mathrm{d}x\ll m\alpha\int_{-T}^{T}\left(1-\frac{|x|}{T}\right)e^{x/2}\ll\frac{m\alpha e^{T/2}}{T}.

By modifying the proof of Theorem 1, we find that

|γπ|(12+ϑ)πloglogC(π)3/m+O(1(loglogC(π)3/m)2).|\gamma_{\pi}|\leq\left(\frac{1}{2}+\vartheta\right)\frac{\pi}{\log\log C(\pi)^{3/m}}+O\left(\frac{1}{\left(\log\log C(\pi)^{3/m}\right)^{2}}\right).

Next, for HTH_{T} we see that the second, third, and last two terms on the right-hand side of (43) are all meT/2T\ll m\frac{e^{T/2}}{T}, and the third term is me(1/2+ϑ)TT\ll m\frac{e^{(1/2+\vartheta)T}}{T}. Hence

nπ(12+ϑ)logC(π)loglogC(π)3/m+O(logC(π)(loglogC(π)3/m)2).n_{\pi}\leq\left(\frac{1}{2}+\vartheta\right)\frac{\log C(\pi)}{\log\log C(\pi)^{3/m}}+O\left(\frac{\log C(\pi)}{\left(\log\log C(\pi)^{3/m}\right)^{2}}\right).

Lastly, similar estimates for GTαG^{\alpha}_{T} imply that

|γπ~|(1+2ϑ)πloglogC(π)3/m+O(1(loglogC(π)3/m)2)|\widetilde{\gamma_{\pi}}|\leq\left(1+2\vartheta\right)\frac{\pi}{\log\log C(\pi)^{3/m}}+O\left(\frac{1}{\left(\log\log C(\pi)^{3/m}\right)^{2}}\right)

in conjunction with the preceding bound on nπn_{\pi}.

8. Acknowledgment

I would like to thank my advisor Ghaith Hiary for his comments and suggestions.

References

  • [Arm99] M. V. Armon “Averages of Real Character Sums” In J. Number Theory 77(2), 1999, pp. 209–226
  • [Ayo63] R. G. Ayoub “An introduction to the analytic theory of numbers”, Math. Surveys, No. 10 Amer. Math. Soc., Providence, R.I., 1963
  • [Bar81] K. Barner “Weil’s explicit formula” In J. reine angew. Math. 323, 1981, pp. 139–152
  • [Ben+21] M. A. Bennett, G. Martin, K. O’Bryant and A. Rechnitzer “Counting zeros of Dirichlet LL-functions” In Math. Comp. 90(329), 2021, pp. 1455–1482
  • [Bui12] H. M. Bui “Non-vanishing of Dirichlet LL-functions at the central point” In Int. J. Number Theory 8(8), 2012, pp. 1855–1881
  • [CCM15] E. Carneiro, V. Chandee and M. B. Milinovich “A note on the zeros of zeta and LL-functions” In Math. Z. 281 (1-2), 2015, pp. 315–332
  • [CCM22] E. Carneiro, A. Chirre and M. B. Milinovich “Hilbert spaces and low-lying zeros of LL-functions” In Adv. Math. 410 (108748), 2022
  • [Cho65] S. Chowla “The Riemann Hypothesis and Hilbert’s Tenth Problem” In Mathematics and Its Applications 4 GordonBreach Science Publishers, New York, 1965
  • [Con05] B. Conrey “Families of LL-functions and 1-level densities” In Recent perspectives in random matrix theory and number theory, London Math. Soc. Lecture Note Ser. Cambridge University Press, Cambridge, 2005, pp. 225–249
  • [DM24] P. Darbar and G. Maiti “Large values of quadratic Dirichlet LL-functions” Preprint, https://arxiv.org/abs/2406.01519, 2024
  • [DPR23] S. Drappeau, K. Pratt and M. Radziwiłł “One-level density estimates for Dirichlet L-functions with extended support” In Algebra Number Theory 17(4), 2023, pp. 805–830
  • [FM13] D. Fiorilli and G. Martin “Inequities in the Shanks–Rényi prime number race: An asymptotic formula for the densities” In J. reine angew. Math. 676, 2013, pp. 121–212
  • [Gao14] P. Gao nn-level density of the low-lying zeros of quadratic Dirichlet LL-functions” In Int. Math. Res. Notices 6, 2014, pp. 1699–1728
  • [HR03] C. P. Hughes and Z. Rudnick “Linear statistics of low-lying zeros of LL-functions” In Q. J. Math. 54(3), 2003, pp. 309–333
  • [IK04] H. Iwaniec and E. Kowalski “Analytic Number Theory”, AMS Colloquium Publications, Vol. 53, 2004
  • [Jut73] M. Jutila “On character sums and class numbers” In J. Number Theory 5, 1973, pp. 203–214
  • [KS99] N. M. Katz and P. Sarnak “Zeroes of zeta functions and symmetry” In Bull. Amer. Math. Soc. (N. S.) 36(1), 1999, pp. 1–26
  • [KMN22] R. Khan, D. Milićevié and H. T. Ngo “Nonvanishing of Dirichlet LL-functions, II” In Math. Z. 300(2), 2022, pp. 1603–1613
  • [Mes86] J.-F. Mestre “Formules explicites et minorations de conducteurs de variétés algébriques” In Compositio Math. 58(2), 1986, pp. 209–232
  • [Mil02] S. D. Miller “The highest lowest zero and other applications of positivity” In Duke Math. J. 112(1), 2002, pp. 83–116
  • [MV73] H. L. Montgomery and R. C. Vaughan “The large sieve” In Mathematika 20, 1973, pp. 119–134
  • [MV07] H. L. Montgomery and R. C. Vaughan “Multiplicative Number Theory I. Classical Theory”, Cambridge Studies in Advanced Mathematics 97 Cambridge University Press, Cambridge, 2007
  • [Mur90] M. R. Murty “On simple zeros of certain LL-series” In Number Theory (Banff, AB, 1988) de Gruyter, Berlin, 1990, pp. 427–439
  • [Odl90] A. M. Odlyzko “Bounds for discriminants and related estimates for class numbers, regulators and zeros of zeta functions: A survey of recent results” In Sém. Théor. Nombres Bordeaux (2) 2(1), 1990, pp. 119–141
  • [Oma00] S. Omar “Majoration du premier zéro de la fonction zêta de Dedekind” In Acta Arith. 99 (1), 2000, pp. 61–65
  • [Oma08] S. Omar “Non-vanishing of Dirichlet LL-functions at the central point” In Algorithmic number theory, Lecture Notes in Comput. Sci. 5011 Springer, Berlin, 2008, pp. 443–453
  • [Oma14] S. Omar “On small zeros of automorphic LL-functions” In C. R. Math. Acad. Sci. Paris 352 (7-8), 2014, pp. 551–556
  • [OS93] A. E. Özlük and C. Snyder “Small zeros of quadratic LL-functions” In Bull. Austral. Math. Soc. 47(2), 1993, pp. 307–319
  • [OS99] A. E. Özlük and C. Snyder “On the distribution of the nontrivial zeros of quadratic LL-functions close to the real axis” In Acta Arith. 91(3), 1999, pp. 209–228
  • [Poi77] G. Poitou “Sur les petits discriminants” In Séminaire Delange-Pisot-Poitou, 18e année: 1976/77, Théorie des nombres, Fasc. 1 Paris: Secrétariat Math., 1977, pp. 1–17
  • [Pra19] K. Pratt “Average non-vanishing of Dirichlet LL-functions at the central point” In Algebra Number Theory 13(1), 2019, pp. 227–249
  • [RS62] J. Rosser and L. Schoenfeld. “Approximate formulas for some functions of prime numbers” In Illinois J. Math. 6(1), 1962, pp. 64–94
  • [Rub01] M. Rubinstein “Low-lying zeros of LL-functions and random matrix theory” In Duke Math. J. 109(1), 2001, pp. 147–181
  • [Sel46] A. Selberg “Contributions to the theory of Dirichlet’s LL-functions” In Skr. Norske Vid.-Akad. Oslo I 3, 1946
  • [Sie45] C. L. Siegel “On the zeros of the Dirichlet LL-functions” In Ann. of Math 46 (3), 1945, pp. 409–422
  • [Sou00] K. Soundararajan “Nonvanishing of quadratic Dirichlet LL-functions at s=12s=\frac{1}{2} In Ann. of Math. (2) 152(2), 2000, pp. 447–488
  • [Vaa85] J. D. Vaaler “Some extremal functions in Fourier analysis” In Bull. Amer. Math. Soc. 12(2), 1985, pp. 183–216