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

Sifting for small split primes
of an imaginary quadratic field
in a given ideal class

Louis M. Gaudet
Abstract.

Let D>3D>3, D3(4)D\equiv 3\;(4) be a prime, and let 𝒞\mathcal{C} be an ideal class in the field 𝐐(D)\mathbf{Q}(\sqrt{-D}). In this article, we give a new proof that p(D,𝒞)p(D,\mathcal{C}), the smallest norm of a split prime 𝔭𝒞\mathfrak{p}\in\mathcal{C}, satisfies p(D,𝒞)DLp(D,\mathcal{C})\ll D^{L} for some absolute constant LL. Our proof is sieve theoretic. In particular, this allows us to avoid the use of log-free zero-density estimates (for class group LL-functions) and the repulsion properties of exceptional zeros, two crucial inputs to previous proofs of this result.

Keywords: sieve, primes, binary quadratic forms, Linnik theorem.

1.   Introduction

For integers q2q\geqslant 2 and (a,q)=1(a,q)=1, let p(q,a)p(q,a) denote the least prime pa(q)p\equiv a\;(q). In 1944, Linnik [29] showed that

(1) p(q,a)qL,p(q,a)\;\ll\;q^{L},

where both LL and the implied constant are absolute. Since then, there have been many improvements on this result. Building on the work of Heath-Brown [19], Xylouris [37] showed that unconditionally one can take L=5L=5 in (1), which is the current record. This comes quite close to the bound

p(q,a)(qlogq)2,p(q,a)\;\ll\;(q\log q)^{2},

which is what follows assuming that the Riemann hypothesis holds for the Dirichlet LL-functions L(s,χ)L(s,\chi).

Such results are difficult to establish unconditionally, and have traditionally (following Linnik) depended on deep results on the zeros of these LL-functions, namely a log-free zero-density estimate and a quantitative version of the Deuring-Heilbronn phenomenon (exceptional zero repulsion effect).

Linnik’s theorem has been generalized in the setting of the Chebotarev density theorem: given a Galois extension of number fields L/KL/K with Galois group GG, each prime 𝔭\mathfrak{p} of KK (unramified in LL) can be associated to a conjugacy class CGC\subset G by the Artin symbol [L/K𝔭]\big{[}\frac{L/K}{\mathfrak{p}}\big{]}. The analogue of Linnik’s theorem is a bound (in terms of the various number field parameters involved) on the least norm NK/𝐐𝔭\text{N}_{K/\mathbf{Q}}\mathfrak{p} of a prime 𝔭\mathfrak{p} with prescribed Artin symbol. There have been many works in this direction, both conditional (see [28] and [2]) and unconditional (see [8], [27], [36], [26], [39], and [35]). These unconditional results proceed by establishing analogues of Linnik’s log-free zero-density estimate and quantitative Deuring-Heilbronn phenomenon for the Hecke LL-functions.

In a different direction, there has been a growing interest in finding new proofs of Linnik’s theorem that avoid using input about the zeros of LL-functions—see for instance [7], [16], [25], [32], [11], [12], [31], [30], and Chapter 24 in [10]. Often these works combine sieve theoretic techniques with “pretentious methods” and/or with techniques coming from additive combinatorics. While such proofs are not usually (as of yet) as numerically strong as those that use the zeros of LL-functions, they are very interesting from a conceptual point of view. Similar to the “elementary proof” of the prime number theorem, such proofs show how challenging results in arithmetic can be achieved without (or with minimal) use of results on the zeros of LL-functions.

In this article, we are interested in a particular analogue of Linnik’s theorem for primes in imaginary quadratic fields, which we prove without using zero-density theorems or exceptional zero repulsion results. Let D>3D>3 be a prime, D3(4)D\equiv 3\;(4), so that D-D is a negative fundamental discriminant. (We work with prime DD for simplicity, though we expect that our methods could be adapted without major modifications to work for general fundamental discriminants D-D.) For an integral ideal 𝔞𝒪K\mathfrak{a}\subseteq\mathcal{O}_{K}, we denote by N𝔞=|𝒪K/𝔞|\text{N}\mathfrak{a}=|\mathcal{O}_{K}/\mathfrak{a}| its (absolute) norm. The class group \mathcal{H} of KK is a finite abelian group of order h=||h=|\mathcal{H}|, the class number.

In analogy to Dirichlet’s theorem on primes in arithmetic progression, one can show using class group characters χ^\chi\in\widehat{\mathcal{H}} that for any specified ideal class 𝒞\mathcal{C}\in\mathcal{H}, there are infinitely many split primes 𝔭\mathfrak{p} (i.e., unramified primes 𝔭\mathfrak{p} with N𝔭=p\text{N}\mathfrak{p}=p, a rational prime) in the class 𝒞\mathcal{C}. Therefore, in analogy to Linnik’s theorem, we are interested in bounding

p(D,𝒞)min{N𝔭:𝔭𝒞 is a split prime of 𝒪K},p(D,\mathcal{C})\;\coloneqq\;\min\{\text{N}\mathfrak{p}:\mathfrak{p}\in\mathcal{C}\text{ is a split prime of }\mathcal{O}_{K}\},

the least norm of a split prime in the class 𝒞\mathcal{C}. Our main result is

Theorem 1.1:

There is an absolute constant L>0L>0 such that

p(D,𝒞)DL,p(D,\mathcal{C})\;\ll\;D^{L},

and the implied constant is absolute.

This result is a special case of the results for the Chebotarev theorem, so it has been established several times before in those works listed above. In particular, Thorner and Zaman [35] showed

p(D,𝒞)D694p(D,\mathcal{C})\;\ll\;D^{694}

for all negative fundamental discriminants D-D. However, our work is novel in that it is the first time such a result has been established without the use of zero-density theorems and quantitative Deuring-Heilbronn results. We also handle new sieve-theoretic challenges (in comparison with works on Linnik’s theorem) in the sifting dimension aspect; see Section 3 for details.

Remarks:

In Zaman’s thesis [38] it is shown that

p(D,𝒞)D455,p(D,\mathcal{C})\;\ll\;D^{455},

which is the current record to our knowledge.

Ditchen [6] has shown that except for a density zero subset of negative fundamental discriminants D0(8)-D\not\equiv 0\;(8), one has p(D,𝒞)D20/3+εp(D,\mathcal{C})\ll D^{20/3+\varepsilon}. For this result, they establish a large sieve inequality for class group characters on average over discriminants, as well as an analogue of the Bombieri-Vinogradov theorem for primes in ideal classes.

Given the correspondence between imaginary quadratic fields and binary quadratic forms (see [5], for instance), for the principal class 𝒞0\mathcal{C}_{0}, the quantity p(D,𝒞0)p(D,\mathcal{C}_{0}) is the also the least prime of the form p=x2+Dy2p=x^{2}+Dy^{2} when D1(4)D\equiv 1\;(4). In our case with D3(4)D\equiv 3\;(4), p(D,𝒞0)p(D,\mathcal{C}_{0}) is the least prime pp of the form 4p=x2+Dy24p=x^{2}+Dy^{2}. The distribution of such primes has been studied by Fouvry and Iwaniec [9] in connection with low-lying zeros of dihedral LL-functions.

2.   Statement of results

Theorem 1.1 is the result of combining Theorems 2.2 and 2.3 below; we now establish our notations and state these theorems.

Given an ideal class 𝒞\mathcal{C} in the class group \mathcal{H} of K=𝐐(D)K=\mathbf{Q}(\sqrt{-D}), we put

λ𝒞(n)=#{𝔞𝒞;N𝔞=n}.\lambda_{\mathcal{C}}(n)\;=\;\#\{\mathfrak{a}\in\mathcal{C};\;\text{N}\mathfrak{a}=n\}.

Given a character χ^\chi\in\widehat{\mathcal{H}} of the class group, we define

(2) λχ(n)=N𝔞=nχ(𝔞),\lambda_{\chi}(n)\;=\;\sum_{\text{N}\mathfrak{a}=n}\chi(\mathfrak{a}),

the sum being taken over integral ideals 𝔞𝒪K\mathfrak{a}\subseteq\mathcal{O}_{K} of norm nn. Then by the orthogonality of the class group characters we have

(3) λ𝒞(n)=1hχ^χ¯(𝒞)λχ(n).\lambda_{\mathcal{C}}(n)\;=\;\frac{1}{h}\sum_{\chi\in\widehat{\mathcal{H}}}\overline{\chi}(\mathcal{C})\lambda_{\chi}(n).

For the trivial character χ0^\chi_{0}\in\widehat{\mathcal{H}} we have

(4) λχ0(n)=(1χD)(n)=dnχD(d),\lambda_{\chi_{0}}(n)\;=\;(1*\chi_{D})(n)\;=\;\sum_{d\mid n}\chi_{D}(d),

which is the number of ideals in 𝒪K\mathcal{O}_{K} of norm nn, and

χD(n)=(Dn)\chi_{D}(n)\;=\;\Big{(}\frac{-D}{n}\Big{)}

is the Kronecker symbol. Indeed, χD\chi_{D} is a primitive real Dirichlet character of conductor DD. By the Dirichlet class number formula, we have

h=1πDL(1,χD).h\;=\;\frac{1}{\pi}\sqrt{D}\;L(1,\chi_{D}).

Our main object of study is the sequence

(5) an=1nλ𝒞(n)f(lognlogx),a_{n}\;=\;\frac{1}{n}\lambda_{\mathcal{C}}(n)f\Big{(}\frac{\log n}{\log x}\Big{)},

f(u)0f(u)\geqslant 0 a smooth function with f^(0)>0\widehat{f}(0)>0, f^\widehat{f} denoting the Fourier transform,

f^(ξ)f(u)e(ξu)du,\widehat{f}(\xi)\;\coloneqq\;\int_{-\infty}^{\infty}f(u)\text{e}(-\xi u)\mathop{}\!\mathrm{d}u,

and e(z)e2πiz\text{e}(z)\coloneqq e^{2\pi iz}. We assume that f(u)f(u) is supported in the segment

1νu1,1-\nu\leqslant u\leqslant 1,

where ν>0\nu>0 is a small concrete number whose value can be determined in the course of our arguments (though the exact value is not important to us). Here, xx is a large parameter going to infinity, and our goal is to estimate

Spap=p1pλ𝒞(p)f(logplogx).S\;\coloneqq\;\sum_{p}a_{p}\;=\;\sum_{p}\frac{1}{p}\lambda_{\mathcal{C}}(p)f\Big{(}\frac{\log p}{\log x}\Big{)}.

By the prime number theorem, we have

p1pf(logplogx)f~(0)> 0,\sum_{p}\frac{1}{p}f\Big{(}\frac{\log p}{\log x}\Big{)}\;\sim\;\widetilde{f}(0)\;>\;0,

where f~\widetilde{f} denotes the Mellin transform of ff,

f~(s)=0us1f(u)du.\widetilde{f}(s)\;=\;\int_{0}^{\infty}u^{s-1}f(u)\mathop{}\!\mathrm{d}u.

One expects prime ideals 𝔭\mathfrak{p} to equidistribute among the hh ideal classes 𝒞\mathcal{C} in \mathcal{H} even when the discriminant DD is comparable in size (in the logarithmic scale) to the norm N𝔭\text{N}\mathfrak{p}. Thus we expect the asymptotic formula

S1hf~(0)S\;\sim\;\frac{1}{h}\;\widetilde{f}(0)

to hold for xDAx\geqslant D^{A} for some absolute constant A>0A>0. Indeed, the Riemann hypothesis for the class group LL-functions LK(s,χ)L_{K}(s,\chi) implies that the above asymptotic formula holds with xD2(logD)4x\geqslant D^{2}(\log D)^{4}. Here we establish the bound

(6) S1hf~(0)S\;\gg\;\frac{1}{h}\;\widetilde{f}(0)

uniformly for xDLx\gg D^{L} for some absolute L>0L>0. (In actuality, we prove a slightly weaker lower bound in the case of an exceptional character; see the precise statement in Theorem 2.2 and the remarks that follow.) In particular, it follows from this lower bound that for all prime D3(4)D\equiv 3\;(4), we have

(7) λ𝒞(p)>0for some pDL.\lambda_{\mathcal{C}}(p)>0\quad\text{for some }p\ll D^{L}.

In other words, every class 𝒞\mathcal{C} contains a prime ideal 𝔭\mathfrak{p} with p=N𝔭DLp=\text{N}\mathfrak{p}\ll D^{L}.

To establish the bound (6) unconditionally, we split our argument into two cases depending on the non/existence of real zeros of the Dirichlet LL-function L(s,χD)L(s,\chi_{D}). We will use assumptions about such zeros in several places in the work (and also for other LL-functions), so to clarify this, we make the following

Definition 2.1:

Let L(s,f)L(s,f) be an LL-function of conductor Δ3\Delta\geqslant 3 (see Appendix A for definitions). For a real number c>0c>0, we say that “Hypothesis H(c)\text{H}(c)” holds for L(s,f)L(s,f) if every zero ρ=β+iγ\rho=\beta+i\gamma of L(s,f)L(s,f) with |γ|1|\gamma|\leqslant 1 satisfies

(8) β 1clogΔ.\beta\;\leqslant\;1-\frac{c}{\log\Delta}.

Now we state our main two theorems, which together prove Theorem 1.1.

Theorem 2.2:

There exists an absolute constant c>0c>0 such that if the Dirichlet LL-function L(s,χD)L(s,\chi_{D}) has a real zero β\beta that satisfies

β> 1clogD,\beta\;>\;1-\frac{c}{\log D},

then we have

(9) Sf^(0)hL(1,χD)logx.S\;\gg\;\frac{\widehat{f}(0)}{h}\frac{L(1,\chi_{D})}{\log x}.
Theorem 2.3:

Let c>0c>0 be the constant from the theorem above, and suppose that Hypothesis H(c)\text{H}(c) holds for L(s,χD)L(s,\chi_{D}). Then we have

(10) Sνh.S\;\gg\;\frac{\nu}{h}.
Remarks:

In both theorems above, the implied constants are absolute and effective, though we make no attempt at computing them here. See [13], where they compute an explicit admissible value for the exponent LL in (7).

We have f~(0)ν\widetilde{f}(0)\ll\nu, so the bound (10) implies (6). On the other hand, the lower bound in (9) is somewhat smaller than the true order of magnitude. This is because in the proof we have used the crude bound V(z)(logx)2V(z)\gg(\log x)^{-2} that does not involve cancellation in sums over χD(p)\chi_{D}(p). We would recover the correct order of magnitude XV(z)f^(0)XV(z)\gg\widehat{f}(0) if we showed that L(1,χD)L(1,\chi_{D}) is well-approximated by the product p<z(1χD(p)/p)1\prod_{p<z}(1-\chi_{D}(p)/p)^{-1}. This is of course expected to be the case, and we prove it in Appendix A under the complementary assumption that χD\chi_{D} is not exceptional. In any case, the bound (9) still shows that there are many primes pDLp\ll D^{L} with λ𝒞(p)>0\lambda_{\mathcal{C}}(p)>0.

3.   Outline of the arguments

We follow the general approach of Friedlander and Iwaniec’s proof of Linnik’s theorem in Chapter 24 of [10]—see also their recent related articles [11] and [12]. Our work differs from theirs in a few key aspects, which we now explain.

The goal is to give a positive lower bound for the sum pap\sum_{p}a_{p}, where for them 𝒜=(an)\mathcal{A}=(a_{n}) is the characteristic function of the arithmetic progression na(q)n\equiv a\;(q), x/2<nxx/2<n\leqslant x, and for us (an)(a_{n}) is defined by (5). They begin with an application of Buchstab’s identity,

(11) S(𝒜,x)=S(𝒜,z)zp<xS(𝒜p,p),S(\mathcal{A},\sqrt{x})\;=\;S(\mathcal{A},z)\;-\sum_{z\leqslant p<\sqrt{x}}S(\mathcal{A}_{p},p),

where S(𝒜,w)S(\mathcal{A},w) denotes the sum of (an)(a_{n}) over nn having no prime factor less than ww, and 𝒜p=(apm)\mathcal{A}_{p}=(a_{pm}) denotes the subsequence of 𝒜\mathcal{A} over multiples of pp. Here we take z=x1/rz=x^{1/r} with rr taken to be as large as necessary.

First they treat the case where there is an exceptional character χ(q)\chi\;(q). In this case, they apply (11) to the “twisted” sequence a~n=λ(n)an\widetilde{a}_{n}=\lambda(n)a_{n}, where

λ(n)=dnχ(d).\lambda(n)=\sum_{d\mid n}\chi(d).

They show using the Fundamental Lemma of Sieve Theory that the two terms on the right-hand side of (11) are (asymptotically as rr becomes large) of the same size, up to a factor of

δ(z,x)zp<xλ(p)p=zp<x1+χ(p)p\delta(z,x)\;\coloneqq\;\sum_{z\leqslant p<x}\frac{\lambda(p)}{p}\;=\;\sum_{z\leqslant p<x}\frac{1+\chi(p)}{p}

present in the second term. Assuming that χ\chi is exceptional, δ(z,x)\delta(z,x) is very small, and a positive lower bound for S(𝒜,x)=papS(\mathcal{A},\sqrt{x})=\sum_{p}a_{p} follows. Friedlander and Iwaniec also give in [11] an alternative approach via Selberg’s sieve that works on similar principles and gives comparable results.

We follow their approach in [10] to prove our Theorem 2.2, which is under the assumption of an exceptional character. This is taken up in Section 9. We require little modification of their arguments, since their method does not require very specific properties of the sequence 𝒜=(an)\mathcal{A}=(a_{n}) beyond some basic sieve assumptions that also apply in our case. In fact, it is even simpler for us, since we have no need to twist our sequence (an)(a_{n}) by the weights λ(n)\lambda(n) above—such a factor naturally appears in this particular sequence already; see Proposition 7.1 for a precise statement.

For the non-exceptional case, Friedlander and Iwaniec work with a combinatorial sieve identity that leads to

(12) S(𝒜,z)S(𝒜,z)+124Q(𝒜),S(\mathcal{A},z)\;\geqslant\;S^{-}(\mathcal{A},z)\;+\;\frac{1}{24}Q(\mathcal{A}),

where

S(𝒜,z)=dP(z)λdAdS^{-}(\mathcal{A},z)\;=\;\sum_{d\mid P(z)}\lambda_{d}^{-}A_{d}

is the lower bound coming from the beta-sieve (so that (λd)(\lambda_{d}^{-}) are the lower-bound beta-sieve weights, and AdA_{d} are the congruence sums for the sequence 𝒜\mathcal{A}—see Section 5.1 for details), and

Q(𝒜)p0p1p2p3p4ap0p1p2p3p4,Q(\mathcal{A})\;\coloneqq\;\sum_{p_{0}}\sum_{p_{1}}\sum_{p_{2}}\sum_{p_{3}}\sum_{p_{4}}a_{p_{0}p_{1}p_{2}p_{3}p_{4}},

where the variables pjp_{j} run over specific segments xαjpjxβjx^{\alpha_{j}}\leqslant p_{j}\leqslant x^{\beta_{j}}.

Their sifting problem is linear (i.e. of sieve dimension κ=1\kappa=1; again, see 5.1 for details), which means that one can show that S(𝒜,z)S^{-}(\mathcal{A},z) is negligible (relatively very small) for zz close to x\sqrt{x}. The upshot is that they show that

S(𝒜,x)124Q(𝒜)S(\mathcal{A},\sqrt{x})\;\gtrsim\;\frac{1}{24}Q(\mathcal{A})

up to some comparably negligible contributions. This reduces the problem of counting primes to finding a lower bound for Q(𝒜)Q(\mathcal{A}), which counts products of prime quintuplets in arithmetic progression. Indeed, the common parity here (products of 1 and 5 primes, respectively) is an artifact of the sieve process.

By contrast, in this work we cannot so readily work with (12). The sifting density function g(d)g(d) for our sequence 𝒜=(an)\mathcal{A}=(a_{n}) is given on primes pp by

g(p)=1+χD(p)p+O(1p2),g(p)\;=\;\frac{1+\chi_{D}(p)}{p}\;+\;O\Big{(}\frac{1}{p^{2}}\Big{)},

and the presence of the character χD(p)\chi_{D}(p) causes fluctuations that hinder one from easily claiming the one-dimensionality of the sieve problem. One possible approach would be to use the fact that we are working in the non-exceptional case (i.e., assuming Hypothesis H(c)\text{H}(c) for L(s,χD)L(s,\chi_{D}), say) to effectively bound the sum p<zg(p)\sum_{p<z}g(p) and hence control the sieve dimension.

However, here we choose to proceed differently: by the trivial bound g(p)2/p+O(1/p2)g(p)\leqslant 2/p+O(1/p^{2}), we can work with a κ=2\kappa=2-dimensional sieve. We can no longer show that S(𝒜,z)S^{-}(\mathcal{A},z) is negligible for zz so close to x\sqrt{x} (only for zx1/β(2)=x1/4.8339z\leqslant x^{1/\beta(2)}=x^{1/4.8339\dots}; see Section 5.1), and so we employ a different combinatorial identity than in [12]. This identity comes from applying a second iteration of the Buchstab formula to each term S(𝒜p,p)S(\mathcal{A}_{p},p) in (11), which gives

(13) S(𝒜,x)=S(𝒜,z)zp<xS(𝒜p,z)+zp2<p1<xS(𝒜p1p2,p2).S(\mathcal{A},\sqrt{x})\;=\;S(\mathcal{A},z)-\sum_{z\leqslant p<\sqrt{x}}S(\mathcal{A}_{p},z)+\mathop{\sum\sum}_{z\leqslant p_{2}<p_{1}<\sqrt{x}}S(\mathcal{A}_{p_{1}p_{2}},p_{2}).

Rather than work with an inequality, we evaluate (nearly asymptotically) each of the three terms on the right-hand side of (13) and show that the result is positive. The first two terms are readily handled via the Fundamental Lemma, so we reduce the problem to analyzing the third term, which is

(14) zp2<p1<xS(𝒜p1p2,p2)=zp2<p1<x(b,P(p2))=1ap1p2b.\mathop{\sum\sum}_{z\leqslant p_{2}<p_{1}<\sqrt{x}}S(\mathcal{A}_{p_{1}p_{2}},p_{2})\;=\mathop{\sum\sum\sum}_{\begin{subarray}{c}z\leqslant p_{2}<p_{1}<\sqrt{x}\\ (b,P(p_{2}))=1\end{subarray}}a_{p_{1}p_{2}b}.

This sum is our analogue of Q(𝒜)Q(\mathcal{A})—note that it is supported on integers which are products of three (almost-) primes, the same parity as in Q(𝒜)Q(\mathcal{A}).

Friedlander and Iwaniec handle Q(𝒜)Q(\mathcal{A}) via a multiplicative analogue of a additive ternary problem treated by the classical circle method. They use Dirichlet characters χ(q)\chi\;(q) to decouple the prime variables pjp_{j}. After removing the contribution of the principal character (the “major arc”), they use two of the prime variables and the orthogonality of the characters to recover the cost of opening the sum with the characters. The remaining three prime variables are used to obtain a nontrivial cancellation in the character sums over primes—importantly, they do not have need for any zero density bounds or repulsion properties of the exceptional zeros.

We handle the sum (14) in a similar manner, here using the class group characters χ^\chi\in\widehat{\mathcal{H}} instead of Dirichlet characters to decouple our variables. Just as above, two variables (p1p_{1} and bb) and the orthogonality of the class group characters are used to recover the cost in using these characters. This involves a type of large sieve inequality for these characters (over integers free from small prime factors) that we develop in Section 8. For nontrivial cancellation in a character sum over the final prime variable p2p_{2}, we apply the explicit formula and use a zero-free region for the class group LL-functions. It is a technical reason that we do not use three prime variables for this as they do in [12]. While their sequence (an)(a_{n}) is localized dyadically, x/2<nxx/2<n\leqslant x, ours is supported in a longer segment x1νnxx^{1-\nu}\leqslant n\leqslant x. This means that we work with a longer sum over the prime variable p2p_{2}, which effectively localizes the dual sum over zeros in the explicit formula to essentially be supported on zeros within the classical zero-free region. It is in this way that we do not make use of any zero density estimates or repulsion effects of exceptional zeros.

4.   Acknowledgments

This work was completed as part of the author’s PhD thesis. He is deeply grateful to his advisor, Henryk Iwaniec, who provided constant support and guidance throughout this project, and who provided very insightful and helpful feedback during the writing of this article.

5.   Preliminaries

5.1.   The beta-sieve

For a nonnegative sequence of real numbers 𝒜=(an)\mathcal{A}=(a_{n}) we define

S(𝒜,z)(n,P(z))=1anS(\mathcal{A},z)\;\coloneqq\sum_{(n,P(z))=1}a_{n}

for z2z\geqslant 2, where P(z)=p<zpP(z)\;=\;\prod_{p<z}p. The congruence sums for 𝒜\mathcal{A} are

Adn0(d)an,A_{d}\;\coloneqq\sum_{n\equiv 0\;(d)}a_{n},

which we will evaluate in the form

(15) Ad=g(d)X+rd,A_{d}\;=\;g(d)X\;+\;r_{d},

where g(d)g(d) is a multiplicative function with 0g(p)< 10\;\leqslant\;g(p)\;<\;1 for prime pp, XX is a smooth approximation to A1A_{1}, and rdr_{d} is a remainder term that is small (on average over dd) in comparison to g(d)Xg(d)X. The range of the modulus dd for which (15) holds is called the level of distribution of the sequence 𝒜\mathcal{A}.

By the inclusion-exclusion principle, one expects that

(16) S(𝒜,z)XV(z),S(\mathcal{A},z)\;\asymp\;XV(z),

where

V(z)p<z(1g(p)).V(z)\;\coloneqq\;\prod_{p<z}(1-g(p)).

To establish the estimates (16), we use a sequence of sieve weights ξ=(ξd)\xi=(\xi_{d}), which are real numbers ξd\xi_{d} supported on squarefree integers dd satisfying

dP(z),dy,d\mid P(z),\qquad d\leqslant y,

and we call yy the level of the sieve. We assume that they satisfy

(17) |ξd| 1for all d.|\xi_{d}|\;\leqslant\;1\quad\text{for all }d.

For sieve weights (ξd)(\xi_{d}), we put θ= 1ξ\theta\;=\;1*\xi; that is,

θ=(θn),θn=dnξd.\theta=(\theta_{n}),\qquad\theta_{n}\;=\;\sum_{d\mid n}\xi_{d}.

To achieve lower- and upper-bounds as in (16), we use two sets of weights (ξd)(\xi_{d}^{-}) and (ξd+)(\xi_{d}^{+}), called lower- and upper-bound sieve weights. We put

(18) S±(𝒜,z)=nanθn±,whereθ±= 1ξ±,S^{\pm}(\mathcal{A},z)\;=\;\sum_{n}a_{n}\theta_{n}^{\pm},\quad\text{where}\quad\theta^{\pm}\;=\;1*\xi^{\pm},

and we require that

θnd(n,P(z))μ(d)θn+for all n,\theta_{n}^{-}\;\leqslant\;\sum_{d\mid(n,P(z))}\mu(d)\;\leqslant\;\theta_{n}^{+}\qquad\text{for all }n,

which implies that

S(𝒜,z)S(𝒜,z)S+(𝒜,z).S^{-}(\mathcal{A},z)\;\leqslant\;S(\mathcal{A},z)\;\leqslant\;S^{+}(\mathcal{A},z).

Finally, we say that our sifting problem has dimension at most κ0\kappa\geqslant 0 if

(19) wp<z(1g(p))1K(logzlogw)κ\prod_{w\leqslant p<z}(1-g(p))^{-1}\;\leqslant\;K\Big{(}\frac{\log z}{\log w}\Big{)}^{\kappa}

for every 2w<z2\leqslant w<z, for some constant K>1K>1.

While many choices of sieve weights would suffice for our purposes (any that furnish a strong-enough “fundamental lemma” result), for concreteness we will from here on work with a specific construction of sieve weights known as the beta-sieve. These weights were first constructed by Iwaniec [21] and also appear in unpublished work of Rosser. They are of combinatorial type, and they satisfy all of the general properties discussed above, including (17)—see Chapter 11 in [10] for a comprehensive treatment.

The main result we require about the beta-sieve weights is

Proposition 5.1 (see Theorem 11.13 in [10]):

Let ξ±\xi^{\pm} be the upper- and lower-bound beta-sieve weights of level yy. Let 𝒜=(an)\mathcal{A}=(a_{n}) be a sequence of nonnegative reals, let rdr_{d} be defined by (15), and assume that g(d)g(d) satisfies (19) with κ0\kappa\geqslant 0.

Let z2z\geqslant 2 and put s=logy/logzs=\log y/\log z. Define S±(𝒜,z)S^{\pm}(\mathcal{A},z) by (18), and put

R±(y,z)=dP(z)ξd±rd.R^{\pm}(y,z)\;=\;\sum_{d\mid P(z)}\xi_{d}^{\pm}r_{d}.

Then we have

S+(𝒜,z)XV(z){𝔉(s)+O((logy)1/6)}+R+(y,z)S^{+}(\mathcal{A},z)\;\leqslant\;XV(z)\Big{\{}\mathfrak{F}(s)+O((\log y)^{-1/6})\Big{\}}\;+\;R^{+}(y,z)

for sβ1s\geqslant\beta-1, and

S(𝒜,z)XV(z){𝔣(s)+O((logy)1/6)}+R(y,z)S^{-}(\mathcal{A},z)\;\geqslant\;XV(z)\Big{\{}\mathfrak{f}(s)+O((\log y)^{-1/6})\Big{\}}\;+\;R^{-}(y,z)

for sβs\geqslant\beta, where β=β(κ)\beta=\beta(\kappa) is a specific absolute constant that depends only on the sifting dimension κ\kappa, 𝔉(s)\mathfrak{F}(s) and 𝔣(s)\mathfrak{f}(s) are the continuous solutions to the following system of differential-difference equations,

{sκ𝔉(s)=Aif β1sβ+1,sκ𝔣(s)=Bat s=β,\displaystyle\begin{cases}s^{\kappa}\mathfrak{F}(s)=A&\text{if }\beta-1\leqslant s\leqslant\beta+1,\\ s^{\kappa}\mathfrak{f}(s)=B&\text{at }s=\beta,\end{cases}
{(sκ𝔉(s))=κsκ1𝔣(s1)if s>β1,(sκ𝔣(s))=κsκ1𝔉(s1)if s>β,\displaystyle\begin{cases}(s^{\kappa}\mathfrak{F}(s))^{\prime}=\kappa s^{\kappa-1}\mathfrak{f}(s-1)&\text{if }s>\beta-1,\\ (s^{\kappa}\mathfrak{f}(s))^{\prime}=\kappa s^{\kappa-1}\mathfrak{F}(s-1)&\text{if }s>\beta,\end{cases}

and A=A(κ)A=A(\kappa) and B=B(κ)B=B(\kappa) are specific absolute constants that depend only on the sifting dimension κ\kappa. As s+s\to+\infty, we have

𝔉(s)= 1+O(es)and𝔣(s)= 1+O(es).\mathfrak{F}(s)\;=\;1+O(e^{-s})\qquad\text{and}\qquad\mathfrak{f}(s)\;=\;1+O(e^{-s}).

We will only apply Proposition 5.1 in the case κ=2\kappa=2. In this case we have B=0B=0; in fact, B(κ)=0B(\kappa)=0 if κ1/2\kappa\geqslant 1/2. On the other hand, it is nontrivial to compute the values of β(κ)\beta(\kappa) and A(κ)A(\kappa) for k1/2k\geqslant 1/2; see Chapter §11.19 in [10] for a discussion of this and a number of useful inequalities. In particular, they provide a table of numerical values of β\beta and AA for specific values of κ1/2\kappa\geqslant 1/2 that were computed by Sara Blight in 2009 and confirmed by Alastair J. Irving in 2014 (who also corrected one value in the table). When κ=2\kappa=2, we have

β(2)= 4.8339865967andA(2)= 43.4968874616.\beta(2)\;=\;4.8339865967\dots\qquad\text{and}\qquad A(2)\;=\;43.4968874616\dots.
Remark:

Since |ξd±|1|\xi_{d}^{\pm}|\leqslant 1 for the beta-sieve weights, R±R^{\pm} are bounded by

R(y,z)=dP(z)dy|rd|.R(y,z)\;=\sum_{\begin{subarray}{c}d\mid P(z)\\ d\leqslant y\end{subarray}}|r_{d}|.

We will always bound the sieve remainder terms absolutely in this work; we have no need to extract additional cancellation from among these terms.

5.2.   Class group LL-functions

Given a character χ\chi of the class group \mathcal{H}, we define the associated LL-function by

LK(s,χ)=𝔞χ(𝔞)(N𝔞)s=n1λχ(n)ns,L_{K}(s,\chi)\;=\;\sum_{\mathfrak{a}}\chi(\mathfrak{a})(\text{N}\mathfrak{a})^{-s}\;=\;\sum_{n\geqslant 1}\lambda_{\chi}(n)n^{-s},

the first sum being taken over all nonzero integral ideals 𝔞\mathfrak{a} of 𝒪K\mathcal{O}_{K}. These functions are entire except in the case that the character is the trivial one, χ=χ0\chi=\chi_{0}; in this case, the above LL-function is the Dedekind zeta function associated to the field KK,

LK(s,χ0)=ζK(s)=ζ(s)L(s,χD),L_{K}(s,\chi_{0})\;=\;\zeta_{K}(s)\;=\;\zeta(s)L(s,\chi_{D}),

where L(s,χD)L(s,\chi_{D}) is the Dirichlet LL-function associated to the Kronecker symbol χD\chi_{D}. Note that χD\chi_{D} is primitive, since D-D is a fundamental discriminant.

The functions

fχ(z)=hδ(χ)+n1λχ(n)e(nz)f_{\chi}(z)\;=\;h\delta(\chi)+\sum_{n\geqslant 1}\lambda_{\chi}(n)\text{e}(nz)

are modular forms of weight 1 for the group Γ0(D)\Gamma_{0}(D) with nebentypus χD\chi_{D}. When χχ0\chi\neq\chi_{0}, they are Hecke eigencuspforms, and in fact they are newforms because the character χD\chi_{D} is primitive. Thus it follows that the coefficients satisfy the Hecke relations

(20) λχ(dm)=q(d,m)μ(q)χD(q)λχ(dq)λχ(mq)for all integers d,m1.\lambda_{\chi}(dm)\;=\sum_{q\mid(d,m)}\mu(q)\chi_{D}(q)\lambda_{\chi}\Big{(}\frac{d}{q}\Big{)}\lambda_{\chi}\Big{(}\frac{m}{q}\Big{)}\qquad\text{for all integers }d,m\geqslant 1.

A convenient reference for these facts is [22]; see in particular §6.6. The class group LL-functions are self-dual in the sense that

LK(s,χ¯)=LK(s,χ),L_{K}(s,\overline{\chi})=L_{K}(s,\chi),

since for all n1n\geqslant 1 we have

λχ¯(n)=λχ(n)\overline{\lambda_{\chi}}(n)=\lambda_{\chi}(n)

even though the character χ^\chi\in\widehat{\mathcal{H}} need not be real. The completed LL-functions

ΛK(s,χ)=γ(s)Ds/2LK(s,χ),γ(s)(2π)sΓ(s),\Lambda_{K}(s,\chi)=\gamma(s)D^{s/2}L_{K}(s,\chi),\qquad\gamma(s)\coloneqq(2\pi)^{-s}\Gamma(s),

satisfy the functional equation (with root number ε=1\varepsilon=1)

(21) ΛK(s,χ)=ΛK(1s,χ).\Lambda_{K}(s,\chi)=\Lambda_{K}(1-s,\chi).

5.3.   The explicit formula and zeros of LK(s,χ)L_{K}(s,\chi)

By logarithmic differentiation of the Euler products

LK(s,χ)=𝔭(1χ(𝔭)(N𝔭)s)1=p(1λχ(p)ps+χD(p)p2s)1,L_{K}(s,\chi)=\prod_{\mathfrak{p}}(1-\chi(\mathfrak{p})(\text{N}\mathfrak{p})^{-s})^{-1}=\prod_{p}(1-\lambda_{\chi}(p)p^{-s}+\chi_{D}(p)p^{-2s})^{-1},

we get

LKLK(s,χ)=𝔞Λχ(𝔞)(N𝔞)s=n1Λχ(n)ns,-\frac{L_{K}^{\prime}}{L_{K}}(s,\chi)=\sum_{\mathfrak{a}}\Lambda_{\chi}(\mathfrak{a})(\text{N}\mathfrak{a})^{-s}=\sum_{n\geqslant 1}\Lambda_{\chi}(n)n^{-s},

where (by a slight abuse of notation)

Λχ(𝔞)={χ(𝔞)logN𝔭 if 𝔞=𝔭k,0otherwise,andΛχ(n)=N𝔞=nΛχ(𝔞).\Lambda_{\chi}(\mathfrak{a})=\begin{cases}\chi(\mathfrak{a})\log\text{N}\mathfrak{p}&\text{ if }\mathfrak{a}=\mathfrak{p}^{k},\\ 0&\text{otherwise,}\end{cases}\qquad\text{and}\qquad\Lambda_{\chi}(n)=\sum_{\text{N}\mathfrak{a}=n}\Lambda_{\chi}(\mathfrak{a}).

Thus Λχ(𝔞)\Lambda_{\chi}(\mathfrak{a}) is supported on powers of prime ideals 𝔭\mathfrak{p}, and Λχ(n)\Lambda_{\chi}(n) is supported on powers of (rational) primes pp. Note that for a rational prime pp we have

Λχ(p)=λχ(p)logp.\Lambda_{\chi}(p)=\lambda_{\chi}(p)\log p.

Using standard arguments (see Theorem 5.11 in [23], for instance), we have

Lemma 5.2:

Let Φ\Phi be a smooth, compactly supported function on 𝐑+\mathbf{R}^{+} with Mellin transform Φ~\widetilde{\Phi}. Then we have

n1Λχ(n)\displaystyle\sum_{n\geqslant 1}\Lambda_{\chi}(n) Φ(n)=Φ~(1)δ(χ)ρΦ~(ρ)+Φ(1)logD\displaystyle\Phi(n)\;=\;\widetilde{\Phi}(1)\delta(\chi)-\sum_{\rho}\widetilde{\Phi}(\rho)+\Phi(1)\log D
(22) +n1Λχ(n)nΦ(1n)+12πi(3/2)Φ~(1s)(γγ(s)+γγ(1s))ds,\displaystyle+\sum_{n\geqslant 1}\frac{\Lambda_{\chi}(n)}{n}\Phi\Big{(}\frac{1}{n}\Big{)}+\frac{1}{2\pi i}\int_{(3/2)}\widetilde{\Phi}(1-s)\Big{(}\frac{\gamma^{\prime}}{\gamma}(s)+\frac{\gamma^{\prime}}{\gamma}(1-s)\Big{)}\mathop{}\!\mathrm{d}s,

where the sum over ρ\rho is taken over all nontrivial zeros of LK(s,χ)L_{K}(s,\chi).

We will apply this formula with test functions Φ\Phi that have the form

(23) Φ(t)=1tlogtϕ(logtlogx),\Phi(t)=\frac{1}{t\log t}\phi\Big{(}\frac{\log t}{\log x}\Big{)},

where ϕ(u)\phi(u) is a compactly supported function on 𝐑+\mathbf{R}^{+}. In this case we have

Proposition 5.3:

Let ϕ\phi be a smooth function supported on [α1,α2][0,1][\alpha_{1},\alpha_{2}]\subset[0,1]. Putting θ=logD/logx\theta=\log D/\log x, suppose that θ<2α1α2\theta<2\alpha_{1}-\alpha_{2}. Then we have

(24) pλχ(p)pϕ(logplogx)=Φ~(1)δ(χ)ρΦ~(ρ)+O(1hlogx),\sum_{p}\frac{\lambda_{\chi}(p)}{p}\phi\Big{(}\frac{\log p}{\log x}\Big{)}\;=\;\widetilde{\Phi}(1)\delta(\chi)\;-\;\sum_{\rho}\widetilde{\Phi}(\rho)\;+\;O\Big{(}\frac{1}{h\log x}\Big{)},

where h=h(D)h=h(D) is the class number, Φ\Phi is defined as in (23), and ρ\rho runs over all nontrivial zeros of LK(s,χ)L_{K}(s,\chi). The implied constant depends only on ϕ\phi.

Proof.

Since ϕ(u)=0\phi(u)=0 for u<α1u<\alpha_{1}, we see that Φ(1/n)=0\Phi(1/n)=0 for all n1n\geqslant 1, and so the third and fourth terms on the right-hand side of (22) vanish. The rest of the proof follows in a standard way: integrate by parts to estimate the integral, and estimate trivially (using the assumption θ<2α1α2\theta<2\alpha_{1}-\alpha_{2}) the contribution of prime powers to the left-hand side of (22). ∎

Finally, we record the following result that we will use in Subsection 12.4 to estimate a sum over the zeros ρ\rho of LK(s,χ)L_{K}(s,\chi).

Proposition 5.4:

Suppose that the Dirichlet LL-function L(s,χD)L(s,\chi_{D}) satisfies Hypothesis H(c)\text{H}(c) with c1/12c\leqslant 1/12. Then each of the class group LL-functions LK(s,χ)L_{K}(s,\chi) satisfy Hypothesis H(c)\text{H}(c) as well.

Proof.

Zaman [40] has shown that for DD sufficiently large, the product of LL-functions χLK(s,χ)\prod_{\chi\in\mathcal{H}}L_{K}(s,\chi) has at most one zero in the region

(25) σ 10.0875logD+3,|t|1,\sigma\;\geqslant\;1-\frac{0.0875}{\log D+3},\qquad|t|\leqslant 1,

and that if such a zero exists, then both it and the associated class group character χ\chi are real. By the genus theory, the only real class group character χ\chi\in\mathcal{H} is the trivial character χ=χ0\chi=\chi_{0} because DD is prime, and in this case the Kronecker factorization theorem gives us

LK(s,χ0)=ζ(s)L(s,χD).L_{K}(s,\chi_{0})\;=\;\zeta(s)L(s,\chi_{D}).

Assuming now that L(s,χD)L(s,\chi_{D}) satisfies Hypothesis H(c)\text{H}(c), we deduce from the above that each LK(s,χ)L_{K}(s,\chi) does as well (assuming that c1/12<0.0875c\leqslant 1/12<0.0875). ∎

Remarks:

An explicit value for cc as in (25) is not necessary for our result in this article. Without an explicit value of the constant, the above result is due to Fogels [8]. Additionally, in [40] they give many other explicit results for more general Hecke LL-functions. See also [24] and [1].

The above proof is precisely the moment where we make use of the fact that DD is prime. To work with general fundamental discriminants D-D, one may adjust the hypothesis of Proposition 5.4 to read “Suppose that for every divisor D1DD_{1}\mid D, the Dirichlet LL-function L(s,χD1)L(s,\chi_{D_{1}}) satisfies Hypothesis H(c)\text{H}(c) with c1/12c\leqslant 1/12,” and the conclusion of the Proposition would still be true. In this case, one would have to correspondingly prove a version of Theorem 2.2 with a different hypothesis (i.e., “There exists a constant c>0c>0 and a divisor D1DD_{1}\mid D such that if L(s,χD1)L(s,\chi_{D_{1}}) has a real zero β\beta…”), which we have chosen not to do here for the sake of a cleaner exposition.

6.   The congruence sums

In this section, we consider the sequence 𝒜=(an)\mathcal{A}=(a_{n}) defined in (5) and evaluate the associated congruence sums,

Ad=n0(d)an=n0(d)1nλ𝒞(n)f(lognlogx).A_{d}\;=\;\sum_{n\equiv 0\;(d)}a_{n}\;=\;\sum_{n\equiv 0\;(d)}\frac{1}{n}\lambda_{\mathcal{C}}(n)f\Big{(}\frac{\log n}{\log x}\Big{)}.

Expressing λ𝒞\lambda_{\mathcal{C}} in terms of λχ\lambda_{\chi} by (3), this is accomplished via

Proposition 6.1:

Let χ^\chi\in\widehat{\mathcal{H}} be a class group character. Let ϕ\phi be a smooth function supported on [α1,α2]𝐑+[\alpha_{1},\alpha_{2}]\subset\mathbf{R}^{+}, and let dxα1/D3/2(logx)A+2d\leqslant x^{\alpha_{1}}/D^{3/2}(\log x)^{A+2} for some A>0A>0. Then

n0(d)1nλχ(n)ϕ(lognlogx)=g(d)(L(1,\displaystyle\sum_{n\equiv 0\;(d)}\frac{1}{n}\lambda_{\chi}(n)\phi\Big{(}\frac{\log n}{\log x}\Big{)}\;=\;g(d)\cdot(L(1, χD)ϕ^(0)logx)δ(χ)\displaystyle\chi_{D})\widehat{\phi}(0)\log x)\cdot\delta(\chi)
+O(τ(d)2dD1/2(logx)A),\displaystyle+\;O\Big{(}\frac{\tau(d)^{2}}{d}D^{-1/2}(\log x)^{-A}\Big{)},

where g(d)g(d) is the multiplicative function given by

(26) g(d)=1dqdμ(q)qχD(q)λχ0(dq),g(d)=\frac{1}{d}\sum_{q\mid d}\frac{\mu(q)}{q}\chi_{D}(q)\lambda_{\chi_{0}}\Big{(}\frac{d}{q}\Big{)},

and δ(χ)=1\delta(\chi)=1 if χ=χ0\chi=\chi_{0}, and δ(χ)=0\delta(\chi)=0 otherwise.

Remark:

The evaluation of the congruence sums AdA_{d} follows directly from the proposition above using (3). In the sieve terminology, this shows that the sequence 𝒜=(an)\mathcal{A}=(a_{n}) has level of distribution y=x1ν/D3/2(logx)A+2y=x^{1-\nu}/D^{3/2}(\log x)^{A+2}.

To prove Proposition 6.1, we use Lemma 6.2, a summation formula for the harmonics λχ\lambda_{\chi} (see for instance (5.16) in [23]). We omit its proof, which is standard—it follows essentially from the functional equation (21).

Lemma 6.2:

Let χ\chi be a class group character. Then for a smooth, compactly supported function Φ\Phi on 𝐑+\mathbf{R}^{+} with Mellin transform Φ~\widetilde{\Phi}, we have

n1λχ(n)Φ(n)=L(1,χD)Φ~(1)δ(χ)+1Dn1λχ(n)H(nD),\sum_{n\geqslant 1}\lambda_{\chi}(n)\Phi(n)\;=\;L(1,\chi_{D})\widetilde{\Phi}(1)\delta(\chi)+\frac{1}{\sqrt{D}}\sum_{n\geqslant 1}\lambda_{\chi}(n)H\Big{(}\frac{n}{D}\Big{)},

where δ(χ)=1\delta(\chi)=1 if χ=χ0\chi=\chi_{0} the trivial character, δ(χ)=0\delta(\chi)=0 if χχ0\chi\neq\chi_{0}, and

(27) H(y)=12πi(3)Φ~(1s)γ(s)γ(1s)ysds,H(y)=\frac{1}{2\pi i}\int_{(3)}\widetilde{\Phi}(1-s)\;\frac{\gamma(s)}{\gamma(1-s)}\;y^{-s}\mathop{}\!\mathrm{d}s,

with γ(s)=(2π)sΓ(s)\gamma(s)=(2\pi)^{-s}\Gamma(s).

Next, we establish a version of the above formula in the logarithmic scale.

Lemma 6.3:

Let ϕ\phi be a smooth function with support contained in [α1,α2]𝐑+[\alpha_{1},\alpha_{2}]\subset\mathbf{R}^{+}. Then for any kxα1/Dk\leqslant x^{\alpha_{1}}/D, we have

11λχ()ϕ(logklogx)=(L(1,χD)ϕ^(0)logx)δ(χ)+O(D1/2(logx)A)\sum_{\ell\geqslant 1}\frac{1}{\ell}\lambda_{\chi}(\ell)\phi\Big{(}\frac{\log k\ell}{\log x}\Big{)}\;=\;(L(1,\chi_{D})\widehat{\phi}(0)\log x)\cdot\delta(\chi)+O(D^{-1/2}(\log x)^{-A})

for any A>0A>0, where the implied constant depends only on ϕ\phi and AA.

Proof.

We apply Lemma 6.2 with the choice

Φ()=1ϕ(logklogx).\Phi(\ell)=\frac{1}{\ell}\phi\Big{(}\frac{\log k\ell}{\log x}\Big{)}.

It is straightforward to verify that

Φ~(1)=ϕ^(0)logx,\widetilde{\Phi}(1)=\widehat{\phi}(0)\log x,

and by partial integration mm times we derive

Φ~(1s)(logx)1m(kxα1)σ(1+|s|)m,\widetilde{\Phi}(1-s)\;\ll\;(\log x)^{1-m}\Big{(}\frac{k}{x^{\alpha_{1}}}\Big{)}^{\sigma}(1+|s|)^{-m},

where σ=Re(s)\sigma=\operatorname{Re}(s), and mm is at our disposal. By Stirling’s formula, we have

γ(s)γ(1s)t2σ1for fixed σ.\frac{\gamma(s)}{\gamma(1-s)}\ll t^{2\sigma-1}\qquad\text{for fixed }\sigma.

Now we estimate the function H(y)H(y) given by (27): we move the line of integration to Re(s)=σ0\operatorname{Re}(s)=\sigma_{0} and take m>2σ0m>2\sigma_{0} to get

H(y)(logx)12σ0(kyxα1)σ0.H(y)\ll(\log x)^{1-2\sigma_{0}}\Big{(}\frac{k}{yx^{\alpha_{1}}}\Big{)}^{\sigma_{0}}.

Now |λχ(n)|τ(n)|\lambda_{\chi}(n)|\leqslant\tau(n), so as long as kxα1/Dk\leqslant x^{\alpha_{1}}/D, we get

1Dn1λχ(n)H(nD)\displaystyle\frac{1}{\sqrt{D}}\sum_{n\geqslant 1}\lambda_{\chi}(n)H\Big{(}\frac{n}{D}\Big{)}\; D1/2(logx)12σ0(kDxα1)σ0n1τ(n)nσ0\displaystyle\ll\;D^{-1/2}(\log x)^{1-2\sigma_{0}}\Big{(}\frac{kD}{x^{\alpha_{1}}}\Big{)}^{\sigma_{0}}\sum_{n\geqslant 1}\tau(n)n^{-\sigma_{0}}
D1/2(logx)A\displaystyle\ll\;D^{-1/2}(\log x)^{-A}

for any given A>0A>0 by taking σ02\sigma_{0}\geqslant 2 sufficiently large. ∎

Note that the main term in the above lemma does not depend on kk, which is a convenient feature in the forthcoming transformations. Using the lemma above, we prove Proposition 6.1.

Proof of Proposition 6.1.

We write n=dmn=dm and use the Hecke relations (20), and then we put m=qm=q\ell to get

(28) n0(d)1nλχ(n)\displaystyle\sum_{n\equiv 0\;(d)}\frac{1}{n}\lambda_{\chi}(n) ϕ(lognlogx)\displaystyle\phi\Big{(}\frac{\log n}{\log x}\Big{)}
=1dqdμ(q)qχD(q)λχ(dq)11λχ()ϕ(logdqlogx).\displaystyle=\;\frac{1}{d}\sum_{q\mid d}\frac{\mu(q)}{q}\chi_{D}(q)\lambda_{\chi}\Big{(}\frac{d}{q}\Big{)}\sum_{\ell\geqslant 1}\frac{1}{\ell}\lambda_{\chi}(\ell)\phi\Big{(}\frac{\log dq\ell}{\log x}\Big{)}.

Next we split the qq-sum according to whether qQq\leqslant Q or q>Qq>Q, with QQ to be chosen later. For the latter range where q>Qq>Q, we estimate the \ell-sum trivially using |λχ()|τ()|\lambda_{\chi}(\ell)|\leqslant\tau(\ell), which gives us

11τ()ϕ(logqdlogx)(logx)2,\sum_{\ell\geqslant 1}\frac{1}{\ell}\tau(\ell)\phi\Big{(}\frac{\log qd\ell}{\log x}\Big{)}\quad\ll\quad(\log x)^{2},

whose contribution to (28) is

(29) 1dqdq>Q|μ(q)qχD(q)λχ(dq)|11τ()ϕ(logqdlogx)τ(d)2d(logx)2Q.\frac{1}{d}\sum_{\begin{subarray}{c}q\mid d\\ q>Q\end{subarray}}\Big{|}\frac{\mu(q)}{q}\chi_{D}(q)\lambda_{\chi}\Big{(}\frac{d}{q}\Big{)}\Big{|}\sum_{\ell\geqslant 1}\frac{1}{\ell}\tau(\ell)\phi\Big{(}\frac{\log qd\ell}{\log x}\Big{)}\quad\ll\quad\frac{\tau(d)^{2}}{d}\frac{(\log x)^{2}}{Q}.

In the other range where qQq\leqslant Q, we apply Proposition 6.3, which is applicable as long as Qdxα1/DQd\leqslant x^{\alpha_{1}}/D, which we will arrange for with our later choice of QQ. This gives

11λχ()ϕ(logqdlogx)=(L(1,χD)ϕ^(0)logx)δ(χ)+O(D1/2(logx)A1).\sum_{\ell\geqslant 1}\frac{1}{\ell}\lambda_{\chi}(\ell)\phi\Big{(}\frac{\log qd\ell}{\log x}\Big{)}\;=\;(L(1,\chi_{D})\widehat{\phi}(0)\log x)\cdot\delta(\chi)+O(D^{-1/2}(\log x)^{-A-1}).

Plugging the above into (28) essentially gives the expression for g(d)g(d) in (26), except that the qq-sum here is restricted to qQq\leqslant Q. The range is easily extended to all qq up to the same error term in (29). Thus in total we have now shown

n0(d)1nλχ(n)ϕ(lognlogx)=g(d)\displaystyle\sum_{n\equiv 0\;(d)}\frac{1}{n}\lambda_{\chi}(n)\phi\Big{(}\frac{\log n}{\log x}\Big{)}\;=\;g(d)\cdot (L(1,χD)ϕ^(0)logx)δ(χ)\displaystyle(L(1,\chi_{D})\widehat{\phi}(0)\log x)\cdot\delta(\chi)
+O(τ(d)2d((logx)2Q+logQD1/2(logx)A+1)).\displaystyle+O\Big{(}\frac{\tau(d)^{2}}{d}\Big{(}\frac{(\log x)^{2}}{Q}+\frac{\log Q}{D^{1/2}(\log x)^{A+1}}\Big{)}\Big{)}.

Finally, choosing Q=D1/2(logx)A+2Q=D^{1/2}(\log x)^{A+2} gives the result. ∎

7.   Sums of λχ\lambda_{\chi} twisted by sieve weights

In this section, let [α1,α2]𝐑+[\alpha_{1},\alpha_{2}]\subset\mathbf{R}^{+}, and let ξ±=(ξd±)\xi^{\pm}=(\xi^{\pm}_{d}) be the beta-sieve weights (upper- or lower-bound) of level yxα1/D3/2(logx)9y\leqslant x^{\alpha_{1}}/D^{3/2}(\log x)^{9}; put θ±=1ξ±=(θn±)\theta^{\pm}=1*\xi^{\pm}=(\theta_{n}^{\pm}).

Proposition 7.1:

Let ϕ\phi be a smooth function supported on [α1,α2][\alpha_{1},\alpha_{2}], and let β=β(2)=4.83398\beta=\beta(2)=4.83398\dots be the constant from Proposition 5.1. Let z2z\geqslant 2 and put s=logy/logzs=\log y/\log z. Then for χχ0\chi\neq\chi_{0} we have

n1θn±nλχ(n)ϕ(lognlogx)h1(logx)2.\sum_{n\geqslant 1}\frac{\theta_{n}^{\pm}}{n}\lambda_{\chi}(n)\phi\Big{(}\frac{\log n}{\log x}\Big{)}\;\ll\;h^{-1}(\log x)^{-2}.

For χ=χ0\chi=\chi_{0} we have

n1θn+nλχ0(n)ϕ(lognlogx)XV(z){𝔉(s)+O((logy)1/6)}\sum_{n\geqslant 1}\frac{\theta_{n}^{+}}{n}\lambda_{\chi_{0}}(n)\phi\Big{(}\frac{\log n}{\log x}\Big{)}\;\leqslant\;XV(z)\Big{\{}\mathfrak{F}(s)+O((\log y)^{-1/6})\Big{\}}

for sβ1s\geqslant\beta-1, and

n1θnnλχ0(n)ϕ(lognlogx)XV(z){𝔣(s)+O((logy)1/6)}\sum_{n\geqslant 1}\frac{\theta_{n}^{-}}{n}\lambda_{\chi_{0}}(n)\phi\Big{(}\frac{\log n}{\log x}\Big{)}\;\geqslant\;XV(z)\Big{\{}\mathfrak{f}(s)+O((\log y)^{-1/6})\Big{\}}

for sβs\geqslant\beta, where

(30) X=L(1,χD)ϕ^(0)(logx),V(z)=p<z(1g(p)),X\;=\;L(1,\chi_{D})\widehat{\phi}(0)(\log x),\qquad V(z)\;=\;\prod_{p<z}(1-g(p)),

and g(d)g(d) is given by (26).

Proof.

We have

n1θn±nλχ(n)ϕ(lognlogx)=1dyξdn0(d)1nλχ(n)ϕ(lognlogx).\sum_{n\geqslant 1}\frac{\theta_{n}^{\pm}}{n}\lambda_{\chi}(n)\phi\Big{(}\frac{\log n}{\log x}\Big{)}\;=\;\sum_{1\leqslant d\leqslant y}\xi_{d}\sum_{n\equiv 0\;(d)}\frac{1}{n}\lambda_{\chi}(n)\phi\Big{(}\frac{\log n}{\log x}\Big{)}.

Applying Proposition 6.1 (with A=7A=7) for the nn-sum on the right-hand side then shows that when χχ0\chi\neq\chi_{0}, the above is bounded by (recall that |ξd|1|\xi_{d}|\leqslant 1)

(31) D1/2(logx)71dy|ξd|τ(d)2dD1/2(logx)3h1(logx)2,D^{-1/2}(\log x)^{-7}\sum_{1\leqslant d\leqslant y}|\xi_{d}|\frac{\tau(d)^{2}}{d}\;\;\ll\;\;D^{-1/2}(\log x)^{-3}\;\;\ll\;\;h^{-1}(\log x)^{-2},

after using the bound h=h(D)D1/2logDh=h(D)\ll D^{1/2}\log D. For χ=χ0\chi=\chi_{0}, we observe that

S±(𝒜,z)=n1θn±nλχ0(n)ϕ(lognlogx)S^{\pm}(\mathcal{A},z)\;=\;\sum_{n\geqslant 1}\frac{\theta_{n}^{\pm}}{n}\lambda_{\chi_{0}}(n)\phi\Big{(}\frac{\log n}{\log x}\Big{)}

is the sifted sum for the sequence 𝒜=(an)\mathcal{A}=(a_{n}) given by

an=1nλχ0(n)ϕ(lognlogx).a_{n}\;=\;\frac{1}{n}\lambda_{\chi_{0}}(n)\phi\Big{(}\frac{\log n}{\log x}\Big{)}.

By Proposition 6.1, the congruence sums for this sequence are

Ad(x)=n0(d)1nλχ0(n)ϕ(lognlogx)=g(d)X+rd(x),A_{d}(x)\;=\;\sum_{n\equiv 0\;(d)}\frac{1}{n}\lambda_{\chi_{0}}(n)\phi\Big{(}\frac{\log n}{\log x}\Big{)}\;=\;g(d)X+r_{d}(x),

where

X=L(1,χD)ϕ^(0)logx,rd(x)τ(d)2dD1/2(logx)7,X\;=\;L(1,\chi_{D})\widehat{\phi}(0)\log x,\qquad r_{d}(x)\;\ll\;\frac{\tau(d)^{2}}{d}D^{-1/2}(\log x)^{-7},

and g(d)g(d) is given by (26). We have

(32) 1g(p)=(11p)(1χD(p)p)(11p)2,1-g(p)\;=\;\Big{(}1-\frac{1}{p}\Big{)}\Big{(}1-\frac{\chi_{D}(p)}{p}\Big{)}\;\geqslant\;\Big{(}1-\frac{1}{p}\Big{)}^{2},

so we can take κ=2\kappa=2 in (19) for some K>1K>1. Applying Proposition 5.1 gives

S+(𝒜,z)\displaystyle S^{+}(\mathcal{A},z) XV(z){𝔉(s)+O((logy)1/6)}+R+(y,z)for sβ1,\displaystyle\leqslant XV(z)\Big{\{}\mathfrak{F}(s)+O((\log y)^{-1/6})\Big{\}}\;+\;R^{+}(y,z)\qquad\text{for }s\geqslant\beta-1,
S(𝒜,z)\displaystyle S^{-}(\mathcal{A},z) XV(z){𝔣(s)+O((logy)1/6)}+R(y,z)for sβ.\displaystyle\geqslant XV(z)\Big{\{}\mathfrak{f}(s)+O((\log y)^{-1/6})\Big{\}}\;+\;\;R^{-}(y,z)\qquad\text{for }s\geqslant\beta.

Just as in (31), the remainder terms R±R^{\pm} are each bounded by

R(y,z)=dP(z)d<y|rd(x)|D1/2(logx)3,R(y,z)\;=\;\sum_{\begin{subarray}{c}d\mid P(z)\\ d<y\end{subarray}}|r_{d}(x)|\;\ll\;D^{-1/2}(\log x)^{-3},

which is covered by O((logy)1/6)O((\log y)^{-1/6}), and the proof is complete. ∎

Here we state a corollary of the above proposition that is ready-to-use for the applications below. From here on we take D=xθD=x^{\theta}, 0<θ<10<\theta<1, and z=x1/rz=x^{1/r}.

Corollary 7.2:

Let 0<ε<1/10r0<\varepsilon<1/10r, and let ϕ\phi be a smooth function supported on [α1ε,α2]𝐑+[\alpha_{1}-\varepsilon,\alpha_{2}]\subset\mathbf{R}^{+}. Let ξ±\xi^{\pm} be the beta-sieve weights of level y=xα1ε/D3/2y=x^{\alpha_{1}-\varepsilon}/D^{3/2}. Suppose that

s(α132θ)r 5.s\;\coloneqq\;(\alpha_{1}-\tfrac{3}{2}\theta)r\;\geqslant\;5.

Then for xx sufficiently large we have

n1θn±nλχ(n)ϕ(lognlogx)\displaystyle\sum_{n\geqslant 1}\frac{\theta_{n}^{\pm}}{n}\lambda_{\chi}(n)\phi\Big{(}\frac{\log n}{\log x}\Big{)}\; h1(logx)2ifχχ0,\displaystyle\ll\;h^{-1}(\log x)^{-2}\qquad\text{if}\quad\chi\neq\chi_{0},
n1θn+nλχ0(n)ϕ(lognlogx)\displaystyle\sum_{n\geqslant 1}\frac{\theta_{n}^{+}}{n}\lambda_{\chi_{0}}(n)\phi\Big{(}\frac{\log n}{\log x}\Big{)}\; XV(z){1+O(es)},and\displaystyle\leqslant\;XV(z)\Big{\{}1+O(e^{-s})\Big{\}},\quad\text{and}
n1θnnλχ0(n)ϕ(lognlogx)\displaystyle\sum_{n\geqslant 1}\frac{\theta_{n}^{-}}{n}\lambda_{\chi_{0}}(n)\phi\Big{(}\frac{\log n}{\log x}\Big{)}\; XV(z){1+O(es)},\displaystyle\geqslant\;XV(z)\Big{\{}1+O(e^{-s})\Big{\}},

where XX and V(z)V(z) are given by (30). Furthermore, if Hypothesis H(c)\text{H}(c) holds for L(s,χD)L(s,\chi_{D}), then we can replace XV(z)XV(z) in the above with

XV(z)=eγϕ^(0)r{1+O(ec/3rθ)}.XV(z)\;=\;e^{-\gamma}\widehat{\phi}(0)r\Big{\{}1+O(e^{-c/3r\theta})\Big{\}}.
Proof.

This follows directly from Proposition 7.1, with s=logy/logzs=\log y/\log z there taken to be sεr=(α1ε32θ)rs-\varepsilon r=(\alpha_{1}-\varepsilon-\frac{3}{2}\theta)r. The condition s5s\geqslant 5 implies sεrβs-\varepsilon r\geqslant\beta (and β1\beta-1 for the lower bound), since ε<1/10r\varepsilon<1/10r.

If Hypothesis H(c)\text{H}(c) holds for L(s,χD)L(s,\chi_{D}), we may use the estimate (74); this together with Mertens’ theorem shows that

XV(z)\displaystyle XV(z)\; =eγϕ^(0)logxlogz{1+O(exp(clogz3logD)+1logz)}\displaystyle=\;e^{-\gamma}\widehat{\phi}(0)\frac{\log x}{\log z}\Big{\{}1+O\Big{(}\exp\Big{(}\frac{-c\log z}{3\log D}\Big{)}+\frac{1}{\log z}\Big{)}\Big{\}}
=eγϕ^(0)r{1+O(ec/3rθ)}.\displaystyle=\;e^{-\gamma}\widehat{\phi}(0)r\Big{\{}1+O(e^{-c/3r\theta})\Big{\}}.

8.   A large sieve-type inequality for λχ\lambda_{\chi} over almost-primes

In this section we give a type of large sieve inequality for the harmonics λχ(n)\lambda_{\chi}(n) where nn runs over integers with no small prime factors. General large sieve inequalities for Hecke characters in number fields were given by Huxley [20] and Schaal [33]; results of the same analytic strength but with explicit constants were later established by Schumer [34].

However, these results are not sufficient for our particular application here. The above results (after specializing to our case) imply a bound

(33) χ^|nNcnnλχ(n)|2(logN)2\sum_{\chi\in\widehat{\mathcal{H}}}\Big{|}\sum_{n\sim N}\frac{c_{n}}{n}\lambda_{\chi}(n)\Big{|}^{2}\;\ll\;(\log N)^{2}

for any complex numbers cnc_{n} satisfying |cn|1|c_{n}|\leqslant 1, as long as DND\leqslant N. In our case, the variable nn has no small prime factors, say (n,P(z))=1(n,P(z))=1. Rather than applying the bound (33) by restriction of the coefficients cnc_{n}, we apply a sieve that allows us to gain two factors of logz\log z, which is crucial for our application here. Our precise result is

Proposition 8.1:

Put z=x1/rz=x^{1/r} and D=xθD=x^{\theta}, and suppose that Hypothesis H(c)\text{H}(c) holds for L(s,χD)L(s,\chi_{D}). Then as long as (α132θ)r5(\alpha_{1}-\frac{3}{2}\theta)r\geqslant 5 and rθ1r\theta\ll 1, we have

(34) χ^|xα1nxα2(n,P(z))=1cnnλχ(n)|2(α2α1)2r2,\sum_{\chi\in\widehat{\mathcal{H}}}\Big{|}\sum_{\begin{subarray}{c}x^{\alpha_{1}}\leqslant n\leqslant x^{\alpha_{2}}\\ (n,P(z))=1\end{subarray}}\frac{c_{n}}{n}\lambda_{\chi}(n)\Big{|}^{2}\;\ll\;(\alpha_{2}-\alpha_{1})^{2}r^{2},

where cnc_{n} are any complex numbers satisfying |cn|1|c_{n}|\leqslant 1.

Proof.

Let 0<ε<1/10r0<\varepsilon<1/10r. Let 0ϕ(u)10\leqslant\phi(u)\leqslant 1 be a smooth function supported on [α1ε,α2+ε][\alpha_{1}-\varepsilon,\alpha_{2}+\varepsilon] such that ϕ(u)=1\phi(u)=1 for α1uα2\alpha_{1}\leqslant u\leqslant\alpha_{2}, and let ξ+=(ξd+)\xi^{+}=(\xi_{d}^{+}) be the upper-bound beta-sieve weights of level y=xα1ε/D3/2y=x^{\alpha_{1}-\varepsilon}/D^{3/2}.

Expanding the square and rearranging, the left-hand side of (34) is

xα1n1,n2xα2(n1n2,P(z))=1cn1cn2¯n1n2χ^λχ(n1)λχ¯(n2).\mathop{\sum\sum}_{\begin{subarray}{c}x^{\alpha_{1}}\leqslant n_{1},n_{2}\leqslant x^{\alpha_{2}}\\ (n_{1}n_{2},P(z))=1\end{subarray}}\frac{c_{n_{1}}\overline{c_{n_{2}}}}{n_{1}n_{2}}\sum_{\chi\in\widehat{\mathcal{H}}}\lambda_{\chi}(n_{1})\overline{\lambda_{\chi}}(n_{2}).

Using the definition (2) of λχ\lambda_{\chi}, the above sum over χ\chi is equal to

χ^λχ(n1)λχ¯(n2)=N𝔞1=n1N𝔞2=n2χχ(𝔞1)χ¯(𝔞2)=hN𝔞1=n1N𝔞2=n2𝔞1𝔞21,\sum_{\chi\in\widehat{\mathcal{H}}}\lambda_{\chi}(n_{1})\overline{\lambda_{\chi}}(n_{2})=\sum_{\text{N}\mathfrak{a}_{1}=n_{1}}\sum_{\text{N}\mathfrak{a}_{2}=n_{2}}\sum_{\chi}\chi(\mathfrak{a}_{1})\overline{\chi}(\mathfrak{a}_{2})=h\mathop{\sum_{\text{N}\mathfrak{a}_{1}=n_{1}}\sum_{\text{N}\mathfrak{a}_{2}=n_{2}}}_{\mathfrak{a}_{1}\sim\mathfrak{a}_{2}}1,

where 𝔞1𝔞2\mathfrak{a}_{1}\sim\mathfrak{a}_{2} means 𝔞1\mathfrak{a}_{1} and 𝔞2\mathfrak{a}_{2} are in the same ideal class. In particular, the above sum is real and nonnegative. Therefore, taking absolute values, we have

χ^|xα1nxα2(n,P(z))=1cnnλχ(n)|2xα1n1,n2xα2(n1n2,P(z))=11n1n2χ^λχ(n1)λχ¯(n2).\sum_{\chi\in\widehat{\mathcal{H}}}\Big{|}\sum_{\begin{subarray}{c}x^{\alpha_{1}}\leqslant n\leqslant x^{\alpha_{2}}\\ (n,P(z))=1\end{subarray}}\frac{c_{n}}{n}\lambda_{\chi}(n)\Big{|}^{2}\;\leqslant\;\mathop{\sum\sum}_{\begin{subarray}{c}x^{\alpha_{1}}\leqslant n_{1},n_{2}\leqslant x^{\alpha_{2}}\\ (n_{1}n_{2},P(z))=1\end{subarray}}\frac{1}{n_{1}n_{2}}\sum_{\chi\in\widehat{\mathcal{H}}}\lambda_{\chi}(n_{1})\overline{\lambda_{\chi}}(n_{2}).

The right-hand side above is majorized by

n1n2θn1+θn2+n1n2ϕ(logn1logx)ϕ(logn1logx)χ^λχ(n1)λχ¯(n2),\sum_{n_{1}}\sum_{n_{2}}\frac{\theta_{n_{1}}^{+}\theta_{n_{2}}^{+}}{n_{1}n_{2}}\phi\Big{(}\frac{\log n_{1}}{\log x}\Big{)}\phi\Big{(}\frac{\log n_{1}}{\log x}\Big{)}\sum_{\chi\in\widehat{\mathcal{H}}}\lambda_{\chi}(n_{1})\overline{\lambda_{\chi}}(n_{2}),

and so we have shown that

(35) χ^|xα1nxα2(n,P(z))=1cnnλχ(n)|2χ^|n1θn+nλχ(n)ϕ(lognlogx)|2.\sum_{\chi\in\widehat{\mathcal{H}}}\Big{|}\sum_{\begin{subarray}{c}x^{\alpha_{1}}\leqslant n\leqslant x^{\alpha_{2}}\\ (n,P(z))=1\end{subarray}}\frac{c_{n}}{n}\lambda_{\chi}(n)\Big{|}^{2}\;\leqslant\;\sum_{\chi\in\widehat{\mathcal{H}}}\Big{|}\sum_{n\geqslant 1}\frac{\theta_{n}^{+}}{n}\lambda_{\chi}(n)\phi\Big{(}\frac{\log n}{\log x}\Big{)}\Big{|}^{2}.

Next we apply Corollary 7.2 to evaluate the nn-sum on the right-hand side of the above, which gives

n1θn+nλχ(n)ϕ(lognlogx)eγϕ^(0)r{1+O(ec/3rθ)}δ(χ)+O(h1(logx)2),\sum_{n\geqslant 1}\frac{\theta_{n}^{+}}{n}\lambda_{\chi}(n)\phi\Big{(}\frac{\log n}{\log x}\Big{)}\;\leqslant\;e^{-\gamma}\widehat{\phi}(0)r\Big{\{}1+O(e^{-c/3r\theta})\Big{\}}\cdot\delta(\chi)+O(h^{-1}(\log x)^{-2}),

as long as (α132θ)r5(\alpha_{1}-\frac{3}{2}\theta)r\geqslant 5. Plugging this into (35), we have shown

χ^|xα1nxα2(n,P(z))=1cnnλχ(n)|2e2γϕ^(0)2r2{1+O(ec/3rθ)}.\sum_{\chi\in\widehat{\mathcal{H}}}\Big{|}\sum_{\begin{subarray}{c}x^{\alpha_{1}}\leqslant n\leqslant x^{\alpha_{2}}\\ (n,P(z))=1\end{subarray}}\frac{c_{n}}{n}\lambda_{\chi}(n)\Big{|}^{2}\;\leqslant\;e^{-2\gamma}\widehat{\phi}(0)^{2}r^{2}\Big{\{}1+O(e^{-c/3r\theta})\Big{\}}.

We have ϕ^(0)(α2α1)\widehat{\phi}(0)\ll(\alpha_{2}-\alpha_{1}), and the O(ec/3rθ)O(e^{-c/3r\theta}) term is superfluous since we assume rθ1r\theta\ll 1. This completes the proof. ∎

9.   The exceptional case: proof of Theorem 2.2

Proof of Theorem 2.2.

We begin by applying the Buchstab formula for the sequence 𝒜=(an)\mathcal{A}=(a_{n}) given by (5). This gives

(36) S(𝒜,x)=S(𝒜,z)zp<xS(𝒜p,p),S(\mathcal{A},\sqrt{x})\;=\;S(\mathcal{A},z)\;-\sum_{z\leqslant p<\sqrt{x}}S(\mathcal{A}_{p},p),

where z=x1/rz=x^{1/r}, with r>0r>0 to be chosen later. Let ξ=(ξd)\xi^{-}=(\xi_{d}^{-}) be the lower-bound beta-sieve weights of level y=x1ν/D3/2y=x^{1-\nu}/D^{3/2}, and put θ=1ξ=(θn)\theta^{-}=1*\xi^{-}=(\theta_{n}^{-}). The orthogonality (3) of the class group characters χ^\chi\in\widehat{\mathcal{H}} implies

S(𝒜,z)1hχχ¯(𝒞)n1θnnλχ(n)f(lognlogx).S(\mathcal{A},z)\;\geqslant\;\frac{1}{h}\sum_{\chi}\overline{\chi}(\mathcal{C})\sum_{n\geqslant 1}\frac{\theta_{n}^{-}}{n}\lambda_{\chi}(n)f\Big{(}\frac{\log n}{\log x}\Big{)}.

Now we apply Corollary 7.2 to get

S(𝒜,z)1hXV(z){1+O(es)},S(\mathcal{A},z)\;\geqslant\;\frac{1}{h}XV(z)\Big{\{}1+O(e^{-s})\Big{\}},

where XX and V(z)V(z) are given by (30), as long as

(37) s=(132θ)r 5.s\;=\;(1-\tfrac{3}{2}\theta)r\;\geqslant\;5.

Now we consider the second term in (36). For each pp, we use the upper bound

S(𝒜p,p)S(𝒜p,z).S(\mathcal{A}_{p},p)\;\leqslant\;S(\mathcal{A}_{p},z).

We have

(38) S(𝒜p,z)=(n,P(z))=1(n,p)=1apn+O((logx)2pz),S(\mathcal{A}_{p},z)\;=\;\sum_{\begin{subarray}{c}(n,P(z))=1\\ (n,p)=1\end{subarray}}a_{pn}\;+\;O\Big{(}\frac{(\log x)^{2}}{pz}\Big{)},

the error term appearing from the terms apna_{pn} with pnp\mid n. Now we let ξ+=(ξd+)\xi^{+}=(\xi_{d}^{+}) be the upper-bound beta-sieve weights supported on dx1/2ν/D3/2d\leqslant x^{1/2-\nu}/D^{3/2}, putting θ+=1ξ+=(θn+)\theta^{+}=1*\xi^{+}=(\theta^{+}_{n}). Expanding in terms of the characters χ\chi, we have

S(𝒜p,z)1hχχ¯(𝒞)λχ(p)pn1(n,p)=1θn+nλχ(n)f(logpnlogx).S(\mathcal{A}_{p},z)\;\leqslant\;\frac{1}{h}\sum_{\chi}\overline{\chi}(\mathcal{C})\frac{\lambda_{\chi}(p)}{p}\!\!\sum_{\begin{subarray}{c}n\geqslant 1\\ (n,p)=1\end{subarray}}\frac{\theta_{n}^{+}}{n}\lambda_{\chi}(n)f\Big{(}\frac{\log pn}{\log x}\Big{)}.

We remove the condition (n,p)=1(n,p)=1 up to the same error term in (38). Note that f(logpn/logx)f(\log pn/\log x) is supported on nx1/2νn\geqslant x^{1/2-\nu} for any pp since p<xp<\sqrt{x}. So Corollary 7.2 gives

S(𝒜p,z)λχ0(p)p1hXV(z){1+O(es+r/2)}+O((logx)2pz),S(\mathcal{A}_{p},z)\;\leqslant\;\frac{\lambda_{\chi_{0}}(p)}{p}\cdot\frac{1}{h}XV(z)\Big{\{}1+O(e^{-s+r/2})\Big{\}}\;+\;O\Big{(}\frac{(\log x)^{2}}{pz}\Big{)},

as long as

(39) sr/2=(1232θ)r 5.s-r/2\;=\;(\tfrac{1}{2}-\tfrac{3}{2}\theta)r\;\geqslant\;5.

The second error term above is subsumed by the first after summing over pp, and hence we get

zp<xS(𝒜p,p)δ(z,x)1hXV(z){1+O(es+r/2)},\sum_{z\leqslant p<\sqrt{x}}S(\mathcal{A}_{p},p)\;\leqslant\;\delta(z,\sqrt{x})\cdot\frac{1}{h}XV(z)\Big{\{}1+O(e^{-s+r/2})\Big{\}},

where we have put

δ(z,w)=zp<wλχ0(p)p.\delta(z,w)\;=\;\sum_{z\leqslant p<w}\frac{\lambda_{\chi_{0}}(p)}{p}.

One can show (see for instance §24.2 of [10]) that

δ(z,w) 2(1β)[logw+O(z1/4)]if w>zD2,\delta(z,w)\;\leqslant\;2(1-\beta)[\log w+O(z^{-1/4})]\qquad\text{if }w>z\geqslant D^{2},

where β\beta is any real zero of L(s,χD)L(s,\chi_{D}). Assuming that L(s,χD)L(s,\chi_{D}) has a real zero β>1c/logD\beta>1-c/\log D, we get

δ(z,x)cθ1+O(z1/4).\delta(z,\sqrt{x})\;\leqslant\;c\theta^{-1}\;+\;O(z^{-1/4}).

Putting everything together, we have shown that

S(𝒜,x)1hXV(z){(1cθ1)+O(e(1/23θ/2)r)}.S(\mathcal{A},\sqrt{x})\;\geqslant\;\frac{1}{h}XV(z)\Big{\{}(1-c\theta^{-1})+O(e^{-(1/2-3\theta/2)r})\Big{\}}.

We choose values for the parameters, each in terms of rr. We take

θ= 1/r2andc= 1/r3.\theta\;=\;1/r^{2}\quad\text{and}\quad c\;=\;1/r^{3}.

With these choices, we see that both (37) and (39) hold for rr sufficiently large, and that S(𝒜,x)h1XV(z)S(\mathcal{A},\sqrt{x})\gg h^{-1}XV(z). Finally, by (32) we have

V(z)p<x(11p)2(logx)2,V(z)\;\geqslant\;\prod_{p<x}\Big{(}1-\frac{1}{p}\Big{)}^{2}\;\gg\;(\log x)^{-2},

and hence

1hXV(z)f^(0)hL(1,χD)logx,\frac{1}{h}XV(z)\;\gg\;\frac{\widehat{f}(0)}{h}\frac{L(1,\chi_{D})}{\log x},

which completes the proof. ∎

10.   The non-exceptional case: proof of Theorem 2.3

In this section we prove Theorem 2.3 under the assumption of the three propositions below, whose proofs we give in the final section. Throughout this entire section and the following two, we assume that Hypothesis H(c)\text{H}(c) holds for L(s,χD)L(s,\chi_{D}).

We begin by applying the Buchstab formula

S(𝒜,x)=S(𝒜,z)zp<xS(𝒜p,p),S(\mathcal{A},\sqrt{x})\;=S(\mathcal{A},z)\;-\;\sum_{z\leqslant p<\sqrt{x}}S(\mathcal{A}_{p},p),

where z=x1/rz=x^{1/r}, with rr to be chosen later. (Note that this r>0r>0 is independent from the one in the previous section, and may be chosen to have a different numerical value.) We then apply a second iteration of the Buchstab formula to each term S(𝒜p,p)S(\mathcal{A}_{p},p), which gives

(40) S(𝒜,x)=S1(𝒜)+S2(𝒜)+S3(𝒜),S(\mathcal{A},\sqrt{x})\;=\;S_{1}(\mathcal{A})\;+\;S_{2}(\mathcal{A})\;+\;S_{3}(\mathcal{A}),

where we have put

S1(𝒜)=S(𝒜,z),S2(𝒜)=zp<xS(𝒜p,z),S_{1}(\mathcal{A})\;=\;S(\mathcal{A},z),\qquad S_{2}(\mathcal{A})\;=\;-\sum_{z\leqslant p<\sqrt{x}}S(\mathcal{A}_{p},z),

and

S3(𝒜)=zp2<p1<xS(𝒜p1p2,p2).S_{3}(\mathcal{A})\;=\mathop{\sum\sum}_{z\leqslant p_{2}<p_{1}<\sqrt{x}}S(\mathcal{A}_{p_{1}p_{2}},p_{2}).

We evaluate each term Si(𝒜)S_{i}(\mathcal{A}) in the following three propositions.

Proposition 10.1:

With z=x1/rz=x^{1/r} and D=xθD=x^{\theta}, suppose that

(41) s(132θ)r 5.s\;\coloneqq\;(1-\tfrac{3}{2}\theta)r\;\geqslant\;5.

Then we have

(42) S1(𝒜)=eγf^(0)rh{1+O(ec/3rθ+es)},S_{1}(\mathcal{A})\;=\;\frac{e^{-\gamma}\widehat{f}(0)r}{h}\Big{\{}1+O(e^{-c/3r\theta}+e^{-s})\Big{\}},

where the implied constant is absolute.

Proposition 10.2:

With z=x1/rz=x^{1/r} and D=xθD=x^{\theta}, suppose that

(43) sr/2=(1232θ)r 5.s-r/2\;=\;(\tfrac{1}{2}-\tfrac{3}{2}\theta)r\;\geqslant\;5.

Then we have

(44) S2(𝒜)=eγf^(0)rlog(r/2)h{1+O(ec/3rθ+es+r/2)},S_{2}(\mathcal{A})\;=\;-\frac{e^{-\gamma}\widehat{f}(0)r\log(r/2)}{h}\Big{\{}1+O(e^{-c/3r\theta}+e^{-s+r/2})\Big{\}},

where the implied constant is absolute.

Proposition 10.3:

Take z=x1/rz=x^{1/r} and D=xθD=x^{\theta}; suppose r10r\geqslant 10 and ν1/20\nu\leqslant 1/20. Let k1k\geqslant 1 be such that

(45) (12r32θ)kr 5andkrθ 1.(\tfrac{1}{2r}-\tfrac{3}{2}\theta)kr\;\geqslant\;5\qquad\text{and}\qquad kr\theta\;\ll\;1.

Then we have

(46) S3(𝒜)=W+O(1h{r2ν2+kr(logr)2ec/18rθ+k2r6ν7/2e5c/18rθ}),S_{3}(\mathcal{A})\;=\;W\;+\;O\Big{(}\frac{1}{h}\Big{\{}r^{2}\nu^{2}+kr(\log r)^{2}e^{-c/18r\theta}+k^{2}r^{6}\nu^{-7/2}e^{-5c/18r\theta}\Big{\}}\Big{)},

where WW is an explicit quantity defined in Section 12, and the implied constant is absolute. Importantly, WW does not depend on the ideal class 𝒞\mathcal{C} in the definition of the sequence 𝒜=(an)\mathcal{A}=(a_{n}).

Assuming these propositions for now, we prove Theorem 2.3.

Proof of Theorem 2.3.

The sequence 𝒜=(an)\mathcal{A}=(a_{n}) defined by

an=1nλ𝒞(n)f(lognlogx)a_{n}\;=\;\frac{1}{n}\lambda_{\mathcal{C}}(n)f\Big{(}\frac{\log n}{\log x}\Big{)}

depends on the ideal class 𝒞\mathcal{C}, so for the moment we write an=an(𝒞)a_{n}=a_{n}(\mathcal{C}) to emphasize this dependence. We now define the sequence =(bn)\mathcal{B}=(b_{n}) by

bn1h𝒞an(𝒞)=1h1nλχ0(n)f(lognlogx),b_{n}\;\coloneqq\;\frac{1}{h}\sum_{\mathcal{C}\in\mathcal{H}}a_{n}(\mathcal{C})\;=\;\frac{1}{h}\cdot\frac{1}{n}\lambda_{\chi_{0}}(n)f\Big{(}\frac{\log n}{\log x}\Big{)},

the second inequality following from the orthogonality (3) of the class group characters. Applying (40) to 𝒜\mathcal{A} and \mathcal{B} and taking the difference, we get

S(𝒜,x)=S(,x)+1i3(Si(𝒜)Si()).S(\mathcal{A},\sqrt{x})\;=\;S(\mathcal{B},\sqrt{x})\;+\;\sum_{1\leqslant i\leqslant 3}\Big{(}S_{i}(\mathcal{A})-S_{i}(\mathcal{B})\Big{)}.

For each ii we have Si()=h1𝒞Si(𝒜)S_{i}(\mathcal{B})=h^{-1}\sum_{\mathcal{C}\in\mathcal{H}}S_{i}(\mathcal{A}). Since each of the right-hand sides of (42), (44), and (46) does not depend on the class 𝒞\mathcal{C}, it follows that each of those equations holds with Si(𝒜)S_{i}(\mathcal{A}) replaced by Si()S_{i}(\mathcal{B}). Therefore we get

S1(𝒜)S1()\displaystyle S_{1}(\mathcal{A})-S_{1}(\mathcal{B})\; rνh{ec/3rθ+es},\displaystyle\ll\;\frac{r\nu}{h}\Big{\{}e^{-c/3r\theta}+e^{-s}\Big{\}},
S2(𝒜)S2()\displaystyle S_{2}(\mathcal{A})-S_{2}(\mathcal{B})\; (logr)rνh{ec/3rθ+es+r/2},and\displaystyle\ll\;\frac{(\log r)r\nu}{h}\Big{\{}e^{-c/3r\theta}+e^{-s+r/2}\Big{\}},\qquad\text{and}
S3(𝒜)S3()\displaystyle S_{3}(\mathcal{A})-S_{3}(\mathcal{B})\; 1h{r2ν2+kr(logr)2ec/18rθ+k2r6ν7/2e5c/18rθ}.\displaystyle\ll\;\frac{1}{h}\Big{\{}r^{2}\nu^{2}+kr(\log r)^{2}e^{-c/18r\theta}+k^{2}r^{6}\nu^{-7/2}e^{-5c/18r\theta}\Big{\}}.

Now we choose our parameters, each in terms of rr. (NB: the rr and θ\theta here are chosen independently from those in Section 9, and cc here is fixed.) We take

(47) k=r1/2,θ=c/r2,andν=er/20.k\;=\;r^{1/2},\quad\theta\;=\;c/r^{2},\quad\text{and}\quad\nu\;=\;e^{-r/20}.

With these choices, one verifies that the conditions (41), (43), and (45) are verified for rr sufficiently large, and that we have

1i3(Si(𝒜)Si())1hr2er/18.\sum_{1\leqslant i\leqslant 3}\Big{(}S_{i}(\mathcal{A})-S_{i}(\mathcal{B})\Big{)}\;\ll\;\frac{1}{h}r^{2}e^{-r/18}.

On the other hand, from (4) we have

S(,x)=1hxp<x1+χD(p)pf(logplogx).S(\mathcal{B},\sqrt{x})\;=\;\frac{1}{h}\sum_{\sqrt{x}\leqslant p<x}\frac{1+\chi_{D}(p)}{p}f\Big{(}\frac{\log p}{\log x}\Big{)}.

From the prime number theorem we have

p1pf(logplogx)=f~(0)+O(1logx)ν,\sum_{p}\frac{1}{p}f\Big{(}\frac{\log p}{\log x}\Big{)}\;=\;\widetilde{f}(0)+O\Big{(}\frac{1}{\log x}\Big{)}\;\gg\;\nu,

and from (76) (which assumes Hypothesis H(c)\text{H}(c) holds for L(s,χD)L(s,\chi_{D})) we have

pχD(p)pf(logplogx)ν2e(1ν)c/θ.\sum_{p}\frac{\chi_{D}(p)}{p}f\Big{(}\frac{\log p}{\log x}\Big{)}\;\ll\;\nu^{-2}e^{-(1-\nu)c/\theta}.

Therefore for our choices of parameters (47), we get

S(,x)νh=1her/20.S(\mathcal{B},\sqrt{x})\;\gg\;\frac{\nu}{h}\;=\;\frac{1}{h}e^{-r/20}.

From (40) this implies S(𝒜,x)ν/hS(\mathcal{A},\sqrt{x})\gg\nu/h for rr sufficiently large, which completes the proof. ∎

11.   Proofs of Propositions 10.1 and 10.2

11.1.   Evaluating S1(𝒜)S_{1}(\mathcal{A})

Proof of Proposition 10.1.

Using the nonnegativity of the terms ana_{n}, we have

n1θnλ𝒞(n)f(lognlogx)S1(𝒜)n1θ+nλ𝒞(n)f(lognlogx),\sum_{n\geqslant 1}\frac{\theta^{-}}{n}\lambda_{\mathcal{C}}(n)f\Big{(}\frac{\log n}{\log x}\Big{)}\;\leqslant\;S_{1}(\mathcal{A})\;\leqslant\;\sum_{n\geqslant 1}\frac{\theta^{+}}{n}\lambda_{\mathcal{C}}(n)f\Big{(}\frac{\log n}{\log x}\Big{)},

where θ±\theta^{\pm} are the beta-sieve weights of level x1ν/D3/2x^{1-\nu}/D^{3/2}. We expand λ𝒞(n)\lambda_{\mathcal{C}}(n) using the orthogonality (3) of the class group characters,

n1θ±nλ𝒞(n)f(lognlogx)=1hχχ¯(𝒞)n1θ±nλχ(n)f(lognlogx),\sum_{n\geqslant 1}\frac{\theta^{\pm}}{n}\lambda_{\mathcal{C}}(n)f\Big{(}\frac{\log n}{\log x}\Big{)}\;=\;\frac{1}{h}\sum_{\chi}\overline{\chi}(\mathcal{C})\sum_{n\geqslant 1}\frac{\theta^{\pm}}{n}\lambda_{\chi}(n)f\Big{(}\frac{\log n}{\log x}\Big{)},

then we apply Corollary 7.2 to each nn-sum on the right-hand side. We get

n1θnλ𝒞(n)f(lognlogx)\displaystyle\sum_{n\geqslant 1}\frac{\theta^{-}}{n}\lambda_{\mathcal{C}}(n)f\Big{(}\frac{\log n}{\log x}\Big{)}\; eγf^(0)rh{1+O(ec/3rθ+es)},\displaystyle\geqslant\;\frac{e^{-\gamma}\widehat{f}(0)r}{h}\Big{\{}1+O(e^{-c/3r\theta}+e^{-s})\Big{\}},
n1θ+nλ𝒞(n)f(lognlogx)\displaystyle\sum_{n\geqslant 1}\frac{\theta^{+}}{n}\lambda_{\mathcal{C}}(n)f\Big{(}\frac{\log n}{\log x}\Big{)}\; eγf^(0)rh{1+O(ec/3rθ+es)},\displaystyle\leqslant\;\frac{e^{-\gamma}\widehat{f}(0)r}{h}\Big{\{}1+O(e^{-c/3r\theta}+e^{-s})\Big{\}},

where s=(132θ)r5s=(1-\tfrac{3}{2}\theta)r\geqslant 5. This gives the result. ∎

11.2.   Evaluating S2(𝒜)S_{2}(\mathcal{A})

Proof of Proposition 10.2.

We evaluate the negative of S2(𝒜)S_{2}(\mathcal{A}),

S2(𝒜)=zp<xS(𝒜p,z).-S_{2}(\mathcal{A})\;=\sum_{z\leqslant p<\sqrt{x}}S(\mathcal{A}_{p},z).

To evaluate the terms

S(𝒜p,z)=(n,P(z))=11pnλ𝒞(pn)f(logpnlogx),S(\mathcal{A}_{p},z)\;=\sum_{(n,P(z))=1}\frac{1}{pn}\lambda_{\mathcal{C}}(pn)f\Big{(}\frac{\log pn}{\log x}\Big{)},

we first attach sieve weights, putting

S±(𝒜p,z)=n1θn±pnλ𝒞(pn)f(logpnlogx),S^{\pm}(\mathcal{A}_{p},z)\;=\;\sum_{n\geqslant 1}\frac{\theta_{n}^{\pm}}{pn}\lambda_{\mathcal{C}}(pn)f\Big{(}\frac{\log pn}{\log x}\Big{)},

where θ±=1ξ±\theta^{\pm}=1*\xi^{\pm}, and ξ±=(ξd±)\xi^{\pm}=(\xi_{d}^{\pm}) are upper- and lower-bound beta-sieve weights of level y=x1/2ν/D3/2y=x^{1/2-\nu}/D^{3/2}. Next, using (3) (note that (p,P(z))=1(p,P(z))=1, since pzp\geqslant z), we have

S±(𝒜p,z)=1hχχ¯(𝒞)S±(𝒜p,z;χ),S^{\pm}(\mathcal{A}_{p},z)\;=\;\frac{1}{h}\sum_{\chi}\overline{\chi}(\mathcal{C})S^{\pm}(\mathcal{A}_{p},z;\chi),

where we have put

S±(𝒜p,z;χ)=n1θn±pnλχ(pn)f(logpnlogx).S^{\pm}(\mathcal{A}_{p},z;\chi)\;=\;\sum_{n\geqslant 1}\frac{\theta_{n}^{\pm}}{pn}\lambda_{\chi}(pn)f\Big{(}\frac{\log pn}{\log x}\Big{)}.

To use the multiplicativity of λχ\lambda_{\chi}, we first remove the terms where pnp\mid n. Such terms contribute to the above sum at most

n1pnτ(n)pnτ(pn)f(logpnlogx)m1τ(p)3p2τ(m)2mf(logp2mlogx)1p2(logx)4,\sum_{\begin{subarray}{c}n\geqslant 1\\ p\mid n\end{subarray}}\frac{\tau(n)}{pn}\tau(pn)f\Big{(}\frac{\log pn}{\log x}\Big{)}\;\leqslant\;\sum_{m\geqslant 1}\frac{\tau(p)^{3}}{p^{2}}\frac{\tau(m)^{2}}{m}f\Big{(}\frac{\log p^{2}m}{\log x}\Big{)}\;\ll\;\frac{1}{p^{2}}(\log x)^{4},

and hence we get

S±(𝒜p,z;χ)=λχ(p)pn1θn±nλχ(n)f(logpnlogx)+O(1p2(logx)4).S^{\pm}(\mathcal{A}_{p},z;\chi)\;=\;\frac{\lambda_{\chi}(p)}{p}\sum_{n\geqslant 1}\frac{\theta_{n}^{\pm}}{n}\lambda_{\chi}(n)f\Big{(}\frac{\log pn}{\log x}\Big{)}+O\Big{(}\frac{1}{p^{2}}(\log x)^{4}\Big{)}.

Now for the nn-sum, we apply Corollary 7.2 with ϕ(u)=f(u+u0)\phi(u)=f(u+u_{0}), where u0=logp/logxu_{0}=\log p/\log x. Note that ϕ\phi is supported on 1/2νu1/21/2-\nu\leqslant u\leqslant 1/2, which accounts for the condition (43). Following the same lines as in the proof of Proposition 10.1, we get

S(𝒜p,z)=λχ0(p)peγf^(0)rh{1+O(ec/3rθ+es+r/2)}+O(1p2(logx)4).S(\mathcal{A}_{p},z)\;=\;\frac{\lambda_{\chi_{0}}(p)}{p}\cdot\frac{e^{-\gamma}\widehat{f}(0)r}{h}\Big{\{}1+O(e^{-c/3r\theta}+e^{-s+r/2})\Big{\}}+O\Big{(}\frac{1}{p^{2}}(\log x)^{4}\Big{)}.

Next, we sum over pp. The contribution of the second OO-term above is at most

1z(logx)4zp<x1pz1(logx)5,\frac{1}{z}(\log x)^{4}\sum_{z\leqslant p<\sqrt{x}}\frac{1}{p}\;\ll\;z^{-1}(\log x)^{5},

which is negligible. For the main term, we evaluate

zp<xλχ0(p)p=zp<x1+χD(p)p=log(r/2)+O(ec/3rθ),\sum_{z\leqslant p<\sqrt{x}}\frac{\lambda_{\chi_{0}}(p)}{p}\;=\sum_{z\leqslant p<\sqrt{x}}\frac{1+\chi_{D}(p)}{p}\;=\;\log(r/2)+O(e^{-c/3r\theta}),

which follows from Mertens’ theorem and Corollary A.2 (using that Hypothesis H(c)\text{H}(c) holds for L(s,χD)L(s,\chi_{D})). Putting everything together completes the proof. ∎

12.   Proof of Proposition 10.3

In this section we prove Proposition 10.3, where we evaluate the sum

(48) S3(𝒜)=zp2<p1<xS(𝒜p1p2,p2)=zp2<p1<x(b,P(p2))=1ap1p2b.S_{3}(\mathcal{A})\;=\mathop{\sum\sum}_{z\leqslant p_{2}<p_{1}<\sqrt{x}}S(\mathcal{A}_{p_{1}p_{2}},p_{2})\;=\mathop{\sum\sum\sum}_{\begin{subarray}{c}z\leqslant p_{2}<p_{1}<\sqrt{x}\\ (b,P(p_{2}))=1\end{subarray}}a_{p_{1}p_{2}b}.
Proof of Proposition 10.3.

We define the quantity

W12(W+(χ0)+W(χ0)),W\;\coloneqq\;\frac{1}{2}\Big{(}W^{+}(\chi_{0})+W^{-}(\chi_{0})\Big{)},

where W±(χ0)W^{\pm}(\chi_{0}) are defined in (56). From the definitions of W±(χ0)W^{\pm}(\chi_{0}) in Section 12.2, it is apparent that WW does not depend on the ideal class in the definition of the sequence 𝒜=(an)\mathcal{A}=(a_{n}). Combining the results of Lemmas 12.1, 12.2, 12.3, and 12.6 below, we see that (46) holds, subject to the conditions (45). ∎

In the remainder of this section, we state and prove Lemmas 12.1, 12.2, 12.3, and 12.6.

12.1.   First arrangements

We separate from (48) the terms where either b=1b=1 or (b,p1p2)>1(b,p_{1}p_{2})>1, putting

S3(𝒜)=V+V+V′′,S_{3}(\mathcal{A})\;=\;V\;+\;V^{\prime}\;+\;V^{\prime\prime},

where

Vzp2<p1<x(b,P(p2))=1,b1(b,p1p2)=1ap1p2bV\;\coloneqq\;\mathop{\sum\sum\sum}_{\begin{subarray}{c}z\leqslant p_{2}<p_{1}<\sqrt{x}\\ (b,P(p_{2}))=1,\;b\neq 1\\ (b,p_{1}p_{2})=1\end{subarray}}a_{p_{1}p_{2}b}

gives the main contribution, and we will show that

Vzp2<p1<xap1p2andV′′zp2<p1<x(b,P(p2))=1(b,p1p2)>1ap1p2bV^{\prime}\;\coloneqq\;\mathop{\sum\sum}_{z\leqslant p_{2}<p_{1}<\sqrt{x}}a_{p_{1}p_{2}}\quad\text{and}\quad V^{\prime\prime}\;\coloneqq\;\mathop{\sum\sum\sum}_{\begin{subarray}{c}z\leqslant p_{2}<p_{1}<\sqrt{x}\\ (b,P(p_{2}))=1\\ (b,p_{1}p_{2})>1\end{subarray}}a_{p_{1}p_{2}b}

give lesser contributions to S3(𝒜)S_{3}(\mathcal{A}). First we estimate V′′V^{\prime\prime}: using |λ𝒞(mn)|τ(mn)τ(m)τ(n)|\lambda_{\mathcal{C}}(mn)|\leqslant\tau(mn)\leqslant\tau(m)\tau(n), we have

V′′zp2<xτ(p2)p2zp1<xτ(p1)p1zb<x(b,P(p2))=1(b,p1p2)>1τ(b)b.V^{\prime\prime}\;\leqslant\sum_{z\leqslant p_{2}<\sqrt{x}}\frac{\tau(p_{2})}{p_{2}}\sum_{z\leqslant p_{1}<\sqrt{x}}\frac{\tau(p_{1})}{p_{1}}\sum_{\begin{subarray}{c}z\leqslant b<\sqrt{x}\\ (b,P(p_{2}))=1\\ (b,p_{1}p_{2})>1\end{subarray}}\frac{\tau(b)}{b}.

Write b=pibb=p_{i}b^{\prime} with i=1i=1 or 22. In either case, the bb-sum above is bounded by

(49) τ(pi)zbxτ(b)b(logx)2z,\frac{\tau(p_{i})}{z}\sum_{b^{\prime}\leqslant x}\frac{\tau(b^{\prime})}{b^{\prime}}\;\ll\;\frac{(\log x)^{2}}{z},

and summing over p1,p2p_{1},p_{2} shows that the same bound holds for V′′V^{\prime\prime}.

For VV^{\prime}, we have

V=zp2<p1<x1p1p2λ𝒞(p1p2)f(logp1p2logx).V^{\prime}\;=\mathop{\sum\sum}_{z\leqslant p_{2}<p_{1}<\sqrt{x}}\frac{1}{p_{1}p_{2}}\lambda_{\mathcal{C}}(p_{1}p_{2})f\Big{(}\frac{\log p_{1}p_{2}}{\log x}\Big{)}.

The support of ff forces p1p2x1νp_{1}p_{2}\geqslant x^{1-\nu}, hence p1x1ν/p2x1/2νp_{1}\geqslant x^{1-\nu}/p_{2}\geqslant x^{1/2-\nu}, and the same for p2p_{2}. We now open λ𝒞\lambda_{\mathcal{C}} using the class group characters, getting

V\displaystyle V^{\prime} 12x1/2νpix1/2p1p21p1p2λ𝒞(p1p2)=12hχχ¯(𝒞)x1/2νpix1/2p1p2λχ(p1p2)p1p2\displaystyle\;\leqslant\;\frac{1}{2}\mathop{\sum\sum}_{\begin{subarray}{c}x^{1/2-\nu}\leqslant p_{i}\leqslant x^{1/2}\\ p_{1}\neq p_{2}\end{subarray}}\frac{1}{p_{1}p_{2}}\lambda_{\mathcal{C}}(p_{1}p_{2})\;=\;\frac{1}{2h}\sum_{\chi}\overline{\chi}(\mathcal{C})\!\!\!\!\!\mathop{\sum\sum}_{\begin{subarray}{c}x^{1/2-\nu}\leqslant p_{i}\leqslant x^{1/2}\\ p_{1}\neq p_{2}\end{subarray}}\frac{\lambda_{\chi}(p_{1}p_{2})}{p_{1}p_{2}}
12hχ|x1/2νpx1/2λχ(p)p|2+O(1x1/2ν).\displaystyle\leqslant\;\frac{1}{2h}\sum_{\chi}\Big{|}\sum_{x^{1/2-\nu}\leqslant p\leqslant x^{1/2}}\frac{\lambda_{\chi}(p)}{p}\Big{|}^{2}\;+\;O\Big{(}\frac{1}{x^{1/2-\nu}}\Big{)}.

Now we apply Proposition 8.1: we choose the coefficients |cn|1|c_{n}|\leqslant 1 appropriately so that the summation

x1/2νnx1/2(n,P(z))=1cnnλχ(n)\sum_{\begin{subarray}{c}x^{1/2-\nu}\leqslant n\leqslant x^{1/2}\\ (n,P(z))=1\end{subarray}}\frac{c_{n}}{n}\lambda_{\chi}(n)

is supported on prime nn. Then as long as

sr/2=(1232θ)r 5,s-r/2\;=\;(\tfrac{1}{2}-\tfrac{3}{2}\theta)r\;\geqslant\;5,

the bound (34) gives

12hχ|x1/2νpx1/2λχ(p)p|2ν2r2h.\frac{1}{2h}\sum_{\chi}\Big{|}\sum_{x^{1/2-\nu}\leqslant p\leqslant x^{1/2}}\frac{\lambda_{\chi}(p)}{p}\Big{|}^{2}\;\ll\;\frac{\nu^{2}r^{2}}{h}.

Thus we have now established

Lemma 12.1:

As long as (1232θ)r5(\frac{1}{2}-\frac{3}{2}\theta)r\geqslant 5 and rθ1r\theta\ll 1, we have

S3(𝒜)=V+O(ν2r2h).S_{3}(\mathcal{A})\;=\;V\;+\;O\Big{(}\frac{\nu^{2}r^{2}}{h}\Big{)}.
Remarks:

Here we have used our large sieve-type inequality (Proposition 8.1) to not lose the factor hh (or any logarithmic factors) after expanding via the class group characters χ^\chi\in\widehat{\mathcal{H}}.

Lemma 12.1 indicates that the sum VV^{\prime} does in fact contribute to a positive proportion of the sum S3(𝒜)S_{3}(\mathcal{A}). However, this proportion is (so to speak) of a lower order of magnitude (proportional to ν2\nu^{2}) than the full sum S3(𝒜)S_{3}(\mathcal{A}) (proportional to ν\nu), which is due to the short range of the variables p1,p2p_{1},p_{2}.

12.2.   A smooth decoupling

For the remaining terms VV from S3(𝒜)S_{3}(\mathcal{A}), we have bp2b\geqslant p_{2}, hence

xp1p2bp1p22>p23,x\;\geqslant\;p_{1}p_{2}b\;\geqslant\;p_{1}p_{2}^{2}\;>\;p_{2}^{3},

and so p2<x1/3p_{2}<x^{1/3}. We make a smooth partition of the variable p2p_{2} into segments that are geometric in the logarithmic scale. First, we partition the range

zp2<x1/3z\;\leqslant\;p_{2}\;<\;x^{1/3}

into segments zj1p2zjz_{j-1}\leqslant p_{2}\leqslant z_{j} with j=1,2,,Jj=1,2,\dots,J, where zjz_{j} are given by

zj=zαj=xαj/r,zαJ=x1/3,z_{j}=z^{\alpha^{j}}=x^{\alpha^{j}/r},\qquad z^{\alpha^{J}}=x^{1/3},

so J1J\geqslant 1 is at our disposal, and it determines α>1\alpha>1. Now we make a smooth partition from these points zjz_{j} in the following way:

  • we now let the index jj run over half-integers 12,1,32,,J,J+12\frac{1}{2},1,\frac{3}{2},\dots,J,J+\frac{1}{2};

  • for each such jj, let 0hj(t)10\leqslant h_{j}(t)\leqslant 1 be a smooth bump function supported on [αj1/r,αj/r][\alpha^{j-1}/r,\alpha^{j}/r], such that for j=12,1,32,,Jj=\frac{1}{2},1,\frac{3}{2},\dots,J we have

    hj(t)+hj+1/2(t)=1fort[αj/r,αj+1/2/r].h_{j}(t)+h_{j+1/2}(t)=1\qquad\text{for}\qquad t\in[\alpha^{j}/r,\alpha^{j+1/2}/r].

We put

h(t)=1jJhj(t)andh+(t)=12jJ+12hj(t),h^{-}(t)\;=\sum_{1\leqslant j\leqslant J}h_{j}(t)\qquad\text{and}\qquad h^{+}(t)\;=\sum_{\frac{1}{2}\leqslant j\leqslant J+\frac{1}{2}}h_{j}(t),

and from the above properties we get

(50) h(t) 1[1/r, 1/3](t)h+(t)\displaystyle h^{-}(t)\;\leqslant\;\mathbf{1}_{[1/r,\;1/3]}(t)\;\leqslant\;h^{+}(t)\qquad for all t,\displaystyle\text{for all }t,
h+(t)= 1\displaystyle h^{+}(t)\;=\;1\qquad for   1/rt1/3,\displaystyle\text{for }\;\;1/r\leqslant t\leqslant 1/3,
h(t)= 1\displaystyle h^{-}(t)\;=\;1\qquad for α1/2/rtαJ+1/2/r.\displaystyle\text{for }\;\;\alpha^{1/2}/r\leqslant t\leqslant\alpha^{J+1/2}/r.

We now use this smooth partition of unity to decouple the variables p1,p2,bp_{1},p_{2},b. From (50) we have

V\displaystyle V 1jJp2hj(logp2logx)p2<p1<xp2b(b,P(p2))=1(b,p1p2)=1ap1p2b,and\displaystyle\;\geqslant\sum_{1\leqslant j\leqslant J}\sum_{p_{2}}h_{j}\Big{(}\frac{\log p_{2}}{\log x}\Big{)}\sum_{p_{2}<p_{1}<\sqrt{x}}\sum_{\begin{subarray}{c}p_{2}\leqslant b\\ (b,P(p_{2}))=1\\ (b,p_{1}p_{2})=1\end{subarray}}a_{p_{1}p_{2}b},\qquad\text{and}
V\displaystyle V 12jJ+12p2hj(logp2logx)p2<p1<xp2b(b,P(p2))=1(b,p1p2)=1ap1p2b.\displaystyle\;\leqslant\sum_{\frac{1}{2}\leqslant j\leqslant J+\frac{1}{2}}\sum_{p_{2}}h_{j}\Big{(}\frac{\log p_{2}}{\log x}\Big{)}\sum_{p_{2}<p_{1}<\sqrt{x}}\sum_{\begin{subarray}{c}p_{2}\leqslant b\\ (b,P(p_{2}))=1\\ (b,p_{1}p_{2})=1\end{subarray}}a_{p_{1}p_{2}b}.

The conditions

(51) p1>p2,bp2,and(b,P(p2))=1p_{1}>p_{2},\qquad b\geqslant p_{2},\qquad\text{and}\qquad(b,P(p_{2}))=1

entangle p1p_{1} and bb with p2p_{2}, so we adjust them to decouple these variables. The variable p2p_{2} lies in the restricted range zj1p2zjz_{j-1}\leqslant p_{2}\leqslant z_{j} by the support of hjh_{j}, and so (by positivity) we replace the three conditions (51) respectively by

p1>zj,bzj,and(b,P(zj))=1p_{1}>z_{j},\qquad b\geqslant z_{j},\qquad\text{and}\qquad(b,P(z_{j}))=1

in the lower bound for VV, and with

p1>zj1,bzj1,and(b,P(zj1))=1p_{1}>z_{j-1},\qquad b\geqslant z_{j-1},\qquad\text{and}\qquad(b,P(z_{j-1}))=1

in the upper bound for VV. After these adjustments we have

(52) V\displaystyle V W1jJp2hj(logp2logx)zj<p1<xzjb(b,P(zj))=1(b,p1p2)=1ap1p2b,and\displaystyle\;\geqslant\;W^{-}\;\coloneqq\sum_{1\leqslant j\leqslant J}\sum_{p_{2}}h_{j}\Big{(}\frac{\log p_{2}}{\log x}\Big{)}\sum_{z_{j}<p_{1}<\sqrt{x}}\sum_{\begin{subarray}{c}z_{j}\leqslant b\\ (b,P(z_{j}))=1\\ (b,p_{1}p_{2})=1\end{subarray}}a_{p_{1}p_{2}b},\qquad\text{and}
(53) V\displaystyle V W+12jJ+12p2hj(logp2logx)zj1<p1<xzj1b(b,P(zj1))=1(b,p1p2)=1ap1p2b.\displaystyle\;\leqslant\;W^{+}\;\coloneqq\sum_{\frac{1}{2}\leqslant j\leqslant J+\frac{1}{2}}\sum_{p_{2}}h_{j}\Big{(}\frac{\log p_{2}}{\log x}\Big{)}\sum_{z_{j-1}<p_{1}<\sqrt{x}}\sum_{\begin{subarray}{c}z_{j-1}\leqslant b\\ (b,P(z_{j-1}))=1\\ (b,p_{1}p_{2})=1\end{subarray}}a_{p_{1}p_{2}b}.

Now we open ap1p2ba_{p_{1}p_{2}b} using characters: by (5) and (3) we have

an=1hχ^χ¯(𝒞)λχ(n)nf(lognlogx),a_{n}=\frac{1}{h}\sum_{\chi\in\widehat{\mathcal{H}}}\overline{\chi}(\mathcal{C})\frac{\lambda_{\chi}(n)}{n}f\Big{(}\frac{\log n}{\log x}\Big{)},

which we put into (52) and (53). By our adjustments above, we have arranged that p1,p2,bp_{1},p_{2},b are automatically pairwise coprime, so the multiplicativity of λχ\lambda_{\chi} now gives

W\displaystyle W^{-} =1h1jJχχ¯(𝒞)zj<p1<xλχ(p1)p1\displaystyle\;=\;\frac{1}{h}\sum_{1\leqslant j\leqslant J}\sum_{\chi}\overline{\chi}(\mathcal{C})\sum_{z_{j}<p_{1}<\sqrt{x}}\frac{\lambda_{\chi}(p_{1})}{p_{1}}
zjb(b,P(zj))=1(b,p1p2)=1λχ(b)bp2λχ(p2)p2hj(logp2logx)f(logp1p2blogx),\displaystyle\phantom{\frac{1}{h}\sum_{1\leqslant j\leqslant J}\sum_{\chi}\overline{\chi}(\mathcal{C})}\cdot\sum_{\begin{subarray}{c}z_{j}\leqslant b\\ (b,P(z_{j}))=1\\ (b,p_{1}p_{2})=1\end{subarray}}\frac{\lambda_{\chi}(b)}{b}\sum_{p_{2}}\frac{\lambda_{\chi}(p_{2})}{p_{2}}h_{j}\Big{(}\frac{\log p_{2}}{\log x}\Big{)}f\Big{(}\frac{\log p_{1}p_{2}b}{\log x}\Big{)},

and

W+\displaystyle W^{+} =1h12jJ+12χχ¯(𝒞)zj1<p1<xλχ(p1)p1\displaystyle\;=\;\frac{1}{h}\sum_{\frac{1}{2}\leqslant j\leqslant J+\frac{1}{2}}\sum_{\chi}\overline{\chi}(\mathcal{C})\sum_{z_{j-1}<p_{1}<\sqrt{x}}\frac{\lambda_{\chi}(p_{1})}{p_{1}}
zj1b(b,P(zj1))=1(b,p1p2)=1λχ(b)bp2λχ(p2)p2hj(logp2logx)f(logp1p2blogx).\displaystyle\phantom{\frac{1}{h}\sum_{1\leqslant j\leqslant J}\sum_{\chi}\overline{\chi}(\mathcal{C})}\cdot\sum_{\begin{subarray}{c}z_{j-1}\leqslant b\\ (b,P(z_{j-1}))=1\\ (b,p_{1}p_{2})=1\end{subarray}}\frac{\lambda_{\chi}(b)}{b}\sum_{p_{2}}\frac{\lambda_{\chi}(p_{2})}{p_{2}}h_{j}\Big{(}\frac{\log p_{2}}{\log x}\Big{)}f\Big{(}\frac{\log p_{1}p_{2}b}{\log x}\Big{)}.

Finally, we wish to remove the conditions (b,p1p2)=1(b,p_{1}p_{2})=1 now that we have made use of them to decouple λχ(p1p2b)\lambda_{\chi}(p_{1}p_{2}b). To do so, we add back the missing terms, which are bounded by the same error term in (49) from before. Putting

Wj(χ)\displaystyle W^{-}_{j}(\chi) =zj<p1<xλχ(p1)p1zjb(b,P(zj))=1λχ(b)bp2λχ(p2)p2hj(logp2logx)f(logp1p2blogx),and\displaystyle\;=\sum_{z_{j}<p_{1}<\sqrt{x}}\!\!\!\!\!\frac{\lambda_{\chi}(p_{1})}{p_{1}}\!\!\!\!\!\!\sum_{\begin{subarray}{c}z_{j}\leqslant b\\ (b,P(z_{j}))=1\end{subarray}}\!\!\!\!\!\!\!\frac{\lambda_{\chi}(b)}{b}\sum_{p_{2}}\frac{\lambda_{\chi}(p_{2})}{p_{2}}h_{j}\Big{(}\frac{\log p_{2}}{\log x}\Big{)}f\Big{(}\frac{\log p_{1}p_{2}b}{\log x}\Big{)},\quad\text{and}
Wj+(χ)\displaystyle W^{+}_{j}(\chi) =zj1<p1<xλχ(p1)p1zj1b(b,P(zj1))=1λχ(b)bp2λχ(p2)p2hj(logp2logx)f(logp1p2blogx),\displaystyle\;=\sum_{z_{j-1}<p_{1}<\sqrt{x}}\!\!\!\!\!\frac{\lambda_{\chi}(p_{1})}{p_{1}}\!\!\!\!\!\!\!\!\sum_{\begin{subarray}{c}z_{j-1}\leqslant b\\ (b,P(z_{j-1}))=1\end{subarray}}\!\!\!\!\!\!\!\!\!\frac{\lambda_{\chi}(b)}{b}\sum_{p_{2}}\frac{\lambda_{\chi}(p_{2})}{p_{2}}h_{j}\Big{(}\frac{\log p_{2}}{\log x}\Big{)}f\Big{(}\frac{\log p_{1}p_{2}b}{\log x}\Big{)},

we have shown the following

Lemma 12.2:

We have

(54) V\displaystyle V\; 1h1jJχχ¯(𝒞)Wj(χ),and\displaystyle\gg\;\frac{1}{h}\sum_{1\leqslant j\leqslant J}\sum_{\chi}\overline{\chi}(\mathcal{C})W_{j}^{-}(\chi),\quad\text{and}
(55) V\displaystyle V\; 1h12jJ+12χχ¯(𝒞)Wj+(χ).\displaystyle\ll\;\frac{1}{h}\sum_{\frac{1}{2}\leqslant j\leqslant J+\frac{1}{2}}\sum_{\chi}\overline{\chi}(\mathcal{C})W_{j}^{+}(\chi).

12.3.   The contribution of the principal character

Now we extract from (54) and (55) the contribution from the principal character χ=χ0\chi=\chi_{0}, which constitutes the main part of VV. Accordingly we put

(56) W(χ0)=1h1jJWj(χ0)andW+(χ0)=1h12jJ+12Wj+(χ0),W^{-}(\chi_{0})=\frac{1}{h}\sum_{1\leqslant j\leqslant J}W^{-}_{j}(\chi_{0})\qquad\text{and}\qquad W^{+}(\chi_{0})=\frac{1}{h}\sum_{\frac{1}{2}\leqslant j\leqslant J+\frac{1}{2}}W^{+}_{j}(\chi_{0}),

and we will show that the difference

W+(χ0)W(χ0)=1h(W1/2+(χ0)+WJ+1/2+(χ0)+1jJ(Wj+(χ0)Wj(χ0)))W^{+}(\chi_{0})-W^{-}(\chi_{0})\;=\;\frac{1}{h}\Big{(}W_{1/2}^{+}(\chi_{0})+W_{J+1/2}^{+}(\chi_{0})+\sum_{1\leqslant j\leqslant J}(W_{j}^{+}(\chi_{0})-W_{j}^{-}(\chi_{0}))\Big{)}

is comparably small.

Lemma 12.3:

Let k1k\geqslant 1, and suppose that ν1/20\nu\leqslant 1/20, r10r\geqslant 10, and

(57) (12r32θ)kr 5.(\tfrac{1}{2r}-\tfrac{3}{2}\theta)kr\;\geqslant\;5.

Then we have

W+(χ0)W(χ0)1hkr(logr)2ec/18rθ.W^{+}(\chi_{0})-W^{-}(\chi_{0})\;\ll\;\frac{1}{h}kr(\log r)^{2}e^{-c/18r\theta}.
Remark:

The variable k1k\geqslant 1 plays no essential theoretical role, but it must be present for technical reasons. On a mechanical level, it is a parameter that can be taken larger to ensure that Corollary 7.2 is applicable even when the range of the involved summation includes (relatively) very small integers.

To prove this lemma, we will use the following couple of estimates.

Lemma 12.4:

Let 0<α1<α2<10<\alpha_{1}<\alpha_{2}<1, and suppose that L(s,χD)L(s,\chi_{D}) satisfies Hypothesis H(c)\text{H}(c). Then we have

xα1pxα2λχ0(p)plog(α2/α1)+ecα1/3θ.\sum_{x^{\alpha_{1}}\leqslant p\leqslant x^{\alpha_{2}}}\frac{\lambda_{\chi_{0}}(p)}{p}\;\ll\;\log(\alpha_{2}/\alpha_{1})\;+\;e^{-c\alpha_{1}/3\theta}.
Proof.

We have λχ0(p)=1+χD(p)\lambda_{\chi_{0}}(p)=1+\chi_{D}(p), so the result follows directly from Mertens’ theorem and Corollary A.2. ∎

Lemma 12.5:

Let 0<α1<α2<10<\alpha_{1}<\alpha_{2}<1. Let k1k\geqslant 1, and suppose that

(58) (α132θ)kr 5.(\alpha_{1}-\tfrac{3}{2}\theta)kr\;\geqslant\;5.

Then for any wz1/k=x1/krw\geqslant z^{1/k}=x^{1/kr}, we have

(59) xα1bxα2(b,P(w))=1λχ0(b)bkr.\sum_{\begin{subarray}{c}x^{\alpha_{1}}\leqslant b\leqslant x^{\alpha_{2}}\\ (b,P(w))=1\end{subarray}}\frac{\lambda_{\chi_{0}}(b)}{b}\;\ll\;kr.
Proof.

Since wz1/kw\geqslant z^{1/k}, by positivity the left hand side of (59) is bounded by

xα1bxα2(b,P(z1/k))=1λχ0(b)bϕ(logblogx),\sum_{\begin{subarray}{c}x^{\alpha_{1}}\leqslant b\leqslant x^{\alpha_{2}}\\ (b,P(z^{1/k}))=1\end{subarray}}\frac{\lambda_{\chi_{0}}(b)}{b}\phi\Big{(}\frac{\log b}{\log x}\Big{)},

where 0ϕ(u)10\leqslant\phi(u)\leqslant 1 is a smooth function supported on [α1ε,α2+ε][\alpha_{1}-\varepsilon,\alpha_{2}+\varepsilon] with 0<ε<1/10kr0<\varepsilon<1/10kr, and ϕ(u)=1\phi(u)=1 when α1uα2\alpha_{1}\leqslant u\leqslant\alpha_{2}. We apply Corollary 7.2. ∎

Proof of Lemma 12.3.

In the following, we will apply Lemma 12.5 in situations where the variable bb always satisfies bz1/2b\geqslant z_{-1/2}, i.e., α11/α1/2r1/2r\alpha_{1}\geqslant 1/\alpha^{1/2}r\geqslant 1/2r. Therefore we assume throughout that k1k\geqslant 1 is chosen so that the condition (57) holds. This ensures that (58) holds any time we apply Lemma 12.5.

First we handle W1/2+(χ0)W^{+}_{1/2}(\chi_{0}). For these terms we have z1/2p2z1/2z_{-1/2}\leqslant p_{2}\leqslant z_{1/2}, or

1rα1/2logp2logx1rα1/2,\frac{1}{r}\alpha^{-1/2}\leqslant\frac{\log p_{2}}{\log x}\leqslant\frac{1}{r}\alpha^{1/2},

as well as z1/2p1xz_{-1/2}\leqslant p_{1}\leqslant\sqrt{x}, or

(60) 1rα1/2logp1logx12.\frac{1}{r}\alpha^{-1/2}\leqslant\frac{\log p_{1}}{\log x}\leqslant\frac{1}{2}.

Using x1νp1p2bxx^{1-\nu}\leqslant p_{1}p_{2}b\leqslant x and the above bounds, we get

(61) 12ν1rα1/2logblogx12rα1/2.\frac{1}{2}-\nu-\frac{1}{r}\alpha^{1/2}\leqslant\frac{\log b}{\log x}\leqslant 1-\frac{2}{r}\alpha^{-1/2}.

Assuming that α2\alpha\leqslant 2, r10r\geqslant 10, and ν1/20\nu\leqslant 1/20, we replace (by positivity) the inequalities (60) and (61) by the simpler ones

12rlogp1logx12and14logblogx1.\frac{1}{2r}\leqslant\frac{\log p_{1}}{\log x}\leqslant\frac{1}{2}\qquad\text{and}\qquad\frac{1}{4}\leqslant\frac{\log b}{\log x}\leqslant 1.

Now applying Lemmas 12.4 and 12.5, we get (using α1/2<2\alpha^{1/2}<2)

(62) W1/2+(χ0)kr(logr)(logα+ec/6rθ).W^{+}_{1/2}(\chi_{0})\;\ll\;kr(\log r)\Big{(}\log\alpha+e^{-c/6r\theta}\Big{)}.

Next we analyze WJ+1/2+(χ0)W^{+}_{J+1/2}(\chi_{0}). We have αJ/r=1/3\alpha^{J}/r=1/3, so for these terms we have

13α1/2logp2logx13α1/2.\frac{1}{3}\alpha^{-1/2}\leqslant\frac{\log p_{2}}{\log x}\leqslant\frac{1}{3}\alpha^{1/2}.

The same lower bounds in (60) and (61) hold, and hence p1p2bxp_{1}p_{2}b\leqslant x gives

13α1/2logp1logx,logblogx123α1/2.\frac{1}{3}\alpha^{-1/2}\quad\leqslant\quad\frac{\log p_{1}}{\log x}\;,\;\frac{\log b}{\log x}\quad\leqslant\quad 1-\frac{2}{3}\alpha^{-1/2}.

We apply Lemmas 12.4 and 12.5 again, and we find that WJ+1/2+(χ0)W_{J+1/2}^{+}(\chi_{0}) satisfies the same bound as W1/2+(χ0)W_{1/2}^{+}(\chi_{0}),

(63) WJ+1/2+(χ0)kr(logr)(logα+ec/6rθ).W_{J+1/2}^{+}(\chi_{0})\;\ll\;kr(\log r)\Big{(}\log\alpha+e^{-c/6r\theta}\Big{)}.

The differences Wj+(χ0)Wj(χ0)W^{+}_{j}(\chi_{0})-W^{-}_{j}(\chi_{0}) are more complicated, but using positivity we can majorize them by two simpler sums,

Wj+(χ0)Wj(χ0)U1+U2,W^{+}_{j}(\chi_{0})-W^{-}_{j}(\chi_{0})\;\leqslant\;U_{1}+U_{2},

where

U1=zj1p1<zjλχ0(p1)p1zj1p2<zjλχ0(p2)p2zj1bx(b,P(zj1))=1λχ0(b)bf(logp1p2blogx)U_{1}\;=\;\sum_{z_{j-1}\leqslant p_{1}<z_{j}}\frac{\lambda_{\chi_{0}}(p_{1})}{p_{1}}\sum_{z_{j-1}\leqslant p_{2}<z_{j}}\frac{\lambda_{\chi_{0}}(p_{2})}{p_{2}}\sum_{\begin{subarray}{c}z_{j-1}\leqslant b\leqslant x\\ (b,P(z_{j-1}))=1\end{subarray}}\frac{\lambda_{\chi_{0}}(b)}{b}f\Big{(}\frac{\log p_{1}p_{2}b}{\log x}\Big{)}

and

U2=zj1p1<xλχ0(p1)p1zj1p2<zjλχ0(p2)p2bBjλχ0(b)bf(logp1p2blogx),U_{2}\;=\;\sum_{z_{j-1}\leqslant p_{1}<\sqrt{x}}\frac{\lambda_{\chi_{0}}(p_{1})}{p_{1}}\sum_{z_{j-1}\leqslant p_{2}<z_{j}}\frac{\lambda_{\chi_{0}}(p_{2})}{p_{2}}\sum_{b\in B_{j}}\frac{\lambda_{\chi_{0}}(b)}{b}f\Big{(}\frac{\log p_{1}p_{2}b}{\log x}\Big{)},

and the set BjB_{j} is given by the difference of sets

Bj={bzj1;(b,P(zj1))=1}{bzj;(b,P(zj))=1}.B_{j}\;=\;\{b\geqslant z_{j-1}\;;\;(b,P(z_{j-1}))=1\}\;\setminus\;\{b\geqslant z_{j}\;;\;(b,P(z_{j}))=1\}.

For U1U_{1}, the condition p1p2bxp_{1}p_{2}b\leqslant x implies

logblogx 12rαj1 12rfor 1jJ.\frac{\log b}{\log x}\;\leqslant\;1-\frac{2}{r}\alpha^{j-1}\;\leqslant\;1-\frac{2}{r}\qquad\text{for }1\leqslant j\leqslant J.

Applying Lemmas 12.4 and 12.5 then gives

(64) U1kr(logα+ec/3rθ)2.U_{1}\;\ll\;kr\Big{(}\log\alpha+e^{-c/3r\theta}\Big{)}^{2}.

For U2U_{2}, we observe that if bBjb\in B_{j} and b<zjb<z_{j}, then bb must be prime, since zj1>zj1/2z_{j-1}>z_{j}^{1/2} (assuming α<2\alpha<2). Otherwise, the elements of BjB_{j} are b=p3bb=p_{3}b^{\prime}, where zj1p3<zjz_{j-1}\leqslant p_{3}<z_{j}, bzj1b^{\prime}\geqslant z_{j-1}, and (b,P(zj1))=1(b^{\prime},P(z_{j-1}))=1. In other words,

Bj{b=p3b;zj1p3<zj,\displaystyle B_{j}\;\subseteq\;\Big{\{}b=p_{3}b^{\prime}\;;\;z_{j-1}\leqslant p_{3}<z_{j},\;\; and either b=1\displaystyle\text{and either }b^{\prime}=1
or bzj1 and (b,P(zj1))=1}.\displaystyle\text{ or }b^{\prime}\geqslant z_{j-1}\text{ and }(b^{\prime},P(z_{j-1}))=1\Big{\}}.

Therefore, using λχ0(p3b)λχ0(p3)λχ0(b)\lambda_{\chi_{0}}(p_{3}b^{\prime})\leqslant\lambda_{\chi_{0}}(p_{3})\lambda_{\chi_{0}}(b^{\prime}), we get

bBjλχ0(b)b(zj1p3<zjλχ0(p3)p3)(1+bzj1(b,P(zj1))=1λχ0(b)b).\sum_{b\in B_{j}}\frac{\lambda_{\chi_{0}}(b)}{b}\quad\leqslant\quad\Big{(}\sum_{z_{j-1}\leqslant p_{3}<z_{j}}\frac{\lambda_{\chi_{0}}(p_{3})}{p_{3}}\Big{)}\Big{(}1\;\;+\!\!\!\!\sum_{\begin{subarray}{c}b^{\prime}\geqslant z_{j-1}\\ (b^{\prime},P(z_{j-1}))=1\end{subarray}}\frac{\lambda_{\chi_{0}}(b^{\prime})}{b^{\prime}}\Big{)}.

Using the condition p1p2p3bxp_{1}p_{2}p_{3}b^{\prime}\leqslant x, we apply Lemmas 12.4 and 12.5 to get

U2kr(logr)(logα+ec/3rθ)2,U_{2}\;\ll\;kr(\log r)\Big{(}\log\alpha+e^{-c/3r\theta}\Big{)}^{2},

which we combine with (64) and sum over jj to get

1jJ(Wj+(χ0)Wj(χ0))Jkr(logr)(logα+ec/3rθ)2.\sum_{1\leqslant j\leqslant J}(W_{j}^{+}(\chi_{0})-W_{j}^{-}(\chi_{0}))\;\ll\;Jkr(\log r)\Big{(}\log\alpha+e^{-c/3r\theta}\Big{)}^{2}.

Now we make a choice for the parameter JJ: we take

(65) J=[(logr)ec/18rθ]+1,J\;=\;\Big{[}(\log r)e^{c/18r\theta}\Big{]}+1,

where [][\;\cdot\;] denotes the integer part. This determines α\alpha via the relation

αJ=r/3,orlogα=J1log(r/3),\alpha^{J}\;=\;r/3,\quad\text{or}\quad\log\alpha=J^{-1}\log(r/3),

and hence our choice of JJ implies

(66) logαec/18rθandJ(logα+ec/3rθ)logr.\log\alpha\;\ll\;e^{-c/18r\theta}\qquad\text{and}\qquad J\Big{(}\log\alpha+e^{-c/3r\theta}\Big{)}\;\ll\;\log r.

Therefore we have

1jJ(Wj+(χ0)Wj(χ0))kr(logr)2ec/18rθ.\sum_{1\leqslant j\leqslant J}(W_{j}^{+}(\chi_{0})-W_{j}^{-}(\chi_{0}))\;\ll\;kr(\log r)^{2}e^{-c/18r\theta}.

Similarly, from (62) and (63) and the bound (66), we see that both W1/2+(χ0)W_{1/2}^{+}(\chi_{0}) and WJ+1/2+(χ0)W_{J+1/2}^{+}(\chi_{0}) satisfy the same bound. This gives the result. ∎

12.4.   The contribution of the other characters

In this section we estimate the contributions of the nonprincipal characters χχ0\chi\neq\chi_{0} to the lower and upper bounds (54) and (55).

Lemma 12.6:

Let k1k\geqslant 1 be such that (57) holds and krθ1kr\theta\ll 1. Then

1hjχχ0χ¯(𝒞)Wj±(χ)1hk2r6ν7/2e5c/18rθ,\displaystyle\frac{1}{h}\sum_{j}\sum_{\chi\neq\chi_{0}}\overline{\chi}(\mathcal{C})W^{\pm}_{j}(\chi)\;\ll\;\frac{1}{h}k^{2}r^{6}\nu^{-7/2}e^{-5c/18r\theta},

where the jj-sum runs over 1jJ1\leqslant j\leqslant J for WW^{-} and 12jJ+12\frac{1}{2}\leqslant j\leqslant J+\frac{1}{2} for W+W^{+}.

Proof.

We treat both W±W^{\pm} at the same time because our arguments do not depend on the specific ranges of the variables p1p_{1} and bb, only that the inequalities

p1,bz1/2p_{1},b\;\geqslant\;z_{-1/2}

hold for every jj. First, we evaluate the sum over p2p_{2} using the explicit formula. Specifically, we apply Proposition 5.3 with the choice

ϕ(u)=hj(u)f(u+logp1logx+logblogx),\phi(u)\;=\;h_{j}(u)f\Big{(}u+\frac{\log p_{1}}{\log x}+\frac{\log b}{\log x}\Big{)},

which gives us

Wj±(χ)=p1λχ(p1)p1bλχ(b)bρχΦ~(ρχ)+O(1hlogx),W^{\pm}_{j}(\chi)\;=\sum_{p_{1}}\frac{\lambda_{\chi}(p_{1})}{p_{1}}\sum_{b}\frac{\lambda_{\chi}(b)}{b}\;\sum_{\rho_{\chi}}\widetilde{\Phi}(\rho_{\chi})\;+\;O\Big{(}\frac{1}{h\log x}\Big{)},

where

Φ~(s)=xu(s1)u1hj(u)f(u+logp1logx+logblogx)du\widetilde{\Phi}(s)\;=\;\int_{-\infty}^{\infty}x^{u(s-1)}u^{-1}h_{j}(u)f\Big{(}u+\frac{\log p_{1}}{\log x}+\frac{\log b}{\log x}\Big{)}\mathop{}\!\mathrm{d}u

and the sum ρχ\sum_{\rho_{\chi}} runs over the nontrivial zeros of LK(s,χ)L_{K}(s,\chi). (We don’t write the conditions for p1,bp_{1},b explicitly, since they differ for W±W^{\pm}, but of course we keep them in our minds as necessary.) Note that there is no polar contribution as in (24) since here we treat the nonprincipal characters χχ0\chi\neq\chi_{0}. We decouple the variables p1,bp_{1},b from the above integral by applying the Fourier inversion

f(u+δ)=f^(w)e((u+δ)w)dw,f(u+\delta)=\int_{-\infty}^{\infty}\widehat{f}(w)\text{e}((u+\delta)w)\mathop{}\!\mathrm{d}w,

which gives us

Wj±(χ)=f^(w)p1\displaystyle W^{\pm}_{j}(\chi)\;=\int_{-\infty}^{\infty}\widehat{f}(w)\sum_{p_{1}} e(wlogp1logx)λχ(p1)p1\displaystyle\text{e}\Big{(}w\frac{\log p_{1}}{\log x}\Big{)}\frac{\lambda_{\chi}(p_{1})}{p_{1}}
be(wlogblogx)λχ(b)bρχH(ρχ,w)dw+O(1hlogx),\displaystyle\sum_{b}\text{e}\Big{(}w\frac{\log b}{\log x}\Big{)}\frac{\lambda_{\chi}(b)}{b}\sum_{\rho_{\chi}}H(\rho_{\chi},w)\mathop{}\!\mathrm{d}w\;+\;O\Big{(}\frac{1}{h\log x}\Big{)},

where

H(s,w)=xu(s1)u1hj(u)e(wu)du.H(s,w)\;=\;\int_{-\infty}^{\infty}x^{u(s-1)}u^{-1}h_{j}(u)\text{e}(wu)\mathop{}\!\mathrm{d}u.

Summing over χχ0\chi\neq\chi_{0} and taking the absolute value, we get

|χχ0χ¯(𝒞)Wj±(χ)||f^(w)|H(w)K(w)dw,\Big{|}\sum_{\chi\neq\chi_{0}}\overline{\chi}(\mathcal{C})W^{\pm}_{j}(\chi)\Big{|}\;\leqslant\;\int_{-\infty}^{\infty}|\widehat{f}(w)|H(w)K(w)\mathop{}\!\mathrm{d}w,

where we have put

(67) H(w)=maxχχ0|ρχH(ρχ,w)|H(w)\;=\;\max_{\chi\neq\chi_{0}}\Big{|}\sum_{\rho_{\chi}}H(\rho_{\chi},w)\Big{|}

and

(68) K(w)=χχ0|p1e(wlogp1logx)λχ(p1)p1||be(wlogblogx)λχ(b)b|.K(w)\;=\;\sum_{\chi\neq\chi_{0}}\Big{|}\sum_{p_{1}}\text{e}\Big{(}w\frac{\log p_{1}}{\log x}\Big{)}\frac{\lambda_{\chi}(p_{1})}{p_{1}}\Big{|}\Big{|}\sum_{b}\text{e}\Big{(}w\frac{\log b}{\log x}\Big{)}\frac{\lambda_{\chi}(b)}{b}\Big{|}.

To estimate H(w)H(w), we integrate by parts three times (after borrowing/returning a factor eue^{u}) to get

H(s,w)=(1+(1s)logx)3euxu(s1)(euu1hj(u)e(wu))′′′du.H(s,w)\;=\;(1+(1-s)\log x)^{-3}\int_{-\infty}^{\infty}e^{-u}x^{u(s-1)}(e^{u}u^{-1}h_{j}(u)\text{e}(wu))^{\prime\prime\prime}\mathop{}\!\mathrm{d}u.

For k=1,2,3k=1,2,3 we have

hj(k)(u)(αjrαj1r)k(r/αjlogα)k(r/logα)kh_{j}^{(k)}(u)\;\ll\;\Big{(}\frac{\alpha^{j}}{r}-\frac{\alpha^{j-1}}{r}\Big{)}^{-k}\;\ll\;(r/\alpha^{j}\log\alpha)^{k}\;\leqslant\;(r/\log\alpha)^{k}

for every jj, with αj1/ruαj/r\alpha^{j-1}/r\leqslant u\leqslant\alpha^{j}/r, and thus we get

(euu1hj(u)e(wu))′′′eu(r/logα)3(1+|w|3),(e^{u}u^{-1}h_{j}(u)\text{e}(wu))^{\prime\prime\prime}\;\ll\;e^{u}(r/\log\alpha)^{3}(1+|w|^{3}),

which holds for all ww. Using this to estimate the above integral, we derive

H(s,w)|1+(1s)logx|3x(σ1)/2r(r/logα)3(1+|w|3).H(s,w)\;\ll\;|1+(1-s)\log x|^{-3}x^{(\sigma-1)/2r}(r/\log\alpha)^{3}(1+|w|^{3}).

We use this bound to estimate (67). The number of zeros ρχ=βχ+iγχ\rho_{\chi}=\beta_{\chi}+i\gamma_{\chi} of LK(s,χ)L_{K}(s,\chi) with 0<βχ<10<\beta_{\chi}<1 and t<|γχ|t+1t<|\gamma_{\chi}|\leqslant t+1 is O(logD(|t|+1))O(\log D(|t|+1)), so the above bound implies

k1k<|γχ|k+1|H(ρχ,w)|\displaystyle\sum_{k\geqslant 1}\sum_{k<|\gamma_{\chi}|\leqslant k+1}|H(\rho_{\chi},w)|\; (r/logα)3(1+|w|3)k1(logDk)k3(logx)3\displaystyle\ll\;(r/\log\alpha)^{3}(1+|w|^{3})\sum_{k\geqslant 1}(\log Dk)k^{-3}(\log x)^{-3}
(r/logα)3(1+|w|3)(logx)2.\displaystyle\ll\;(r/\log\alpha)^{3}(1+|w|^{3})(\log x)^{-2}.

By Proposition 5.4, the remaining zeros ρχ\rho_{\chi} of LK(s,χ)L_{K}(s,\chi) fall in the range of Hypothesis H(c)\text{H}(c), so they satisfy

βχ 1clogD,\beta_{\chi}\;\leqslant\;1-\frac{c}{\log D},

and hence x(1βχ)/2rec/2rθx^{(1-\beta_{\chi})/2r}\leqslant e^{-c/2r\theta}. Applying Lemma A.3 now gives

(69) ρχ=βχ+iγχ|γχ|1|H(ρχ,w)|(r/logα)3ec/2rθ(1+|w|3).\sum_{\begin{subarray}{c}\rho_{\chi}=\beta_{\chi}+i\gamma_{\chi}\\ |\gamma_{\chi}|\leqslant 1\end{subarray}}|H(\rho_{\chi},w)|\;\ll\;(r/\log\alpha)^{3}e^{-c/2r\theta}(1+|w|^{3}).

This bound covers the one above for |γχ|>1|\gamma_{\chi}|>1, and it is uniform in χ\chi, so we deduce that H(w)H(w) is also bounded by the right-hand side of (69).

For K(w)K(w), we apply the Cauchy inequality to the χ\chi-sum in (68) to get

(70) K(w)(χ|p1\displaystyle K(w)\;\leqslant\;\Big{(}\sum_{\chi}\Big{|}\sum_{p_{1}} e(wlogp1logx)λχ(p1)p1|2)1/2\displaystyle\text{e}\Big{(}w\frac{\log p_{1}}{\log x}\Big{)}\frac{\lambda_{\chi}(p_{1})}{p_{1}}\Big{|}^{2}\Big{)}^{1/2}
(χ|be(wlogblogx)λχ(b)b|2)1/2.\displaystyle\cdot\Big{(}\sum_{\chi}\Big{|}\sum_{b}\text{e}\Big{(}w\frac{\log b}{\log x}\Big{)}\frac{\lambda_{\chi}(b)}{b}\Big{|}^{2}\Big{)}^{1/2}.

Here we have (by positivity) added back the principal character to the χ\chi-sums. We apply Proposition 8.1: with k1k\geqslant 1 chosen so that (57) holds, we have

K(w)(χ|m1(m,P(x1/kr))=1cmλχ(m)m|2)1/2(χ|n1(n,P(x1/kr))=1cnλχ(n)n|2)1/2,K(w)\;\leqslant\;\Big{(}\sum_{\chi}\Big{|}\!\!\!\!\!\!\sum_{\begin{subarray}{c}m\geqslant 1\\ (m,P(x^{1/kr}))=1\end{subarray}}\!\!\!\!\!\!c_{m}\frac{\lambda_{\chi}(m)}{m}\Big{|}^{2}\Big{)}^{1/2}\Big{(}\sum_{\chi}\Big{|}\!\!\!\!\!\!\sum_{\begin{subarray}{c}n\geqslant 1\\ (n,P(x^{1/kr}))=1\end{subarray}}\!\!\!\!\!\!c_{n}^{\prime}\frac{\lambda_{\chi}(n)}{n}\Big{|}^{2}\Big{)}^{1/2},

where we choose the coefficients cm,cnc_{m},c_{n}^{\prime} to agree with those of p1,bp_{1},b in (70) when m=p1,n=bm=p_{1},n=b, and we choose them to be 0 otherwise. This gives

K(w)k2r2.K(w)\;\ll\;k^{2}r^{2}.

For the integration over ww, we have

|f^(w)|(1\displaystyle\int_{-\infty}^{\infty}|\widehat{f}(w)|(1 +|w|3)dw\displaystyle+|w|^{3})\mathop{}\!\mathrm{d}w
(|f^(w)|2(1+|w|4)2dw)1/2((1+|w|31+|w|4)2dw)1/2\displaystyle\leqslant\;\Big{(}\int_{-\infty}^{\infty}|\widehat{f}(w)|^{2}(1+|w|^{4})^{2}\mathop{}\!\mathrm{d}w\Big{)}^{1/2}\Big{(}\int_{-\infty}^{\infty}\Big{(}\frac{1+|w|^{3}}{1+|w|^{4}}\Big{)}^{2}\mathop{}\!\mathrm{d}w\Big{)}^{1/2}
((|f(u)|2+|f(2)(u)|2+|f(4)(u)|2)du)1/2.\displaystyle\ll\;\Big{(}\int_{-\infty}^{\infty}(|f(u)|^{2}+|f^{(2)}(u)|^{2}+|f^{(4)}(u)|^{2})\mathop{}\!\mathrm{d}u\Big{)}^{1/2}.

The derivatives f(k)(u)f^{(k)}(u) are supported on 1νu11-\nu\leqslant u\leqslant 1 and satisfy

maxu|f(k)(u)|νk,\max_{u}|f^{(k)}(u)|\;\ll\;\nu^{-k},

and hence we derive

|f^(w)|(1+|w|3)dwν7/2.\int_{-\infty}^{\infty}|\widehat{f}(w)|(1+|w|^{3})\mathop{}\!\mathrm{d}w\;\ll\;\nu^{-7/2}.

Therefore

|f^(w)|H(w)K(w)dwk2r2(r/logα)3ν7/2ec/2rθ.\int_{-\infty}^{\infty}|\widehat{f}(w)|H(w)K(w)\mathop{}\!\mathrm{d}w\;\ll\;k^{2}r^{2}(r/\log\alpha)^{3}\nu^{-7/2}e^{-c/2r\theta}.

Finally we sum over jJ=log(r/3)/logαj\ll J=\log(r/3)/\log\alpha, which gives

jχχ0χ¯(𝒞)Wj±(χ)k2r6(logα)4ν7/2ec/2rθ.\sum_{j}\sum_{\chi\neq\chi_{0}}\overline{\chi}(\mathcal{C})W^{\pm}_{j}(\chi)\;\ll\;k^{2}r^{6}(\log\alpha)^{-4}\nu^{-7/2}e^{-c/2r\theta}.

From (65) we have (logα)4e2/9rθ(\log\alpha)^{-4}\ll e^{2/9r\theta}, which completes the proof. ∎

Appendix A Approximating LL-functions by finite Euler products

Let L(s,f)L(s,f) be an entire LL-function of degree d1d\geqslant 1, where we think of ff as some interesting arithmetic object to which L(s,f)L(s,f) is attached. It is natural to try to approximate L(1,f)L(1,f) by a partial Euler product. This question and similar ones have been addressed by many works in the literature—see for instance [18], [17], [15], [3], [4], and [14], where such approximations are both developed and used for interesting arithmetic applications.

The main result in this section, Proposition A.1, would follow from results in the cited works above (for L(s,χD)L(s,\chi_{D}), at least, from a slight modification of the results in [3]) after using an appropriate zero-density estimate for L(s,f)L(s,f). However, we wish to give here a self-contained proof of the result that does not rely on any zero-density estimates.

We assume that L(s,f)L(s,f) is given by a Dirichlet series and Euler product,

L(s,f)\displaystyle L(s,f) =n1λf(n)ns=pLp(s,f),\displaystyle\;=\;\sum_{n\geqslant 1}\lambda_{f}(n)n^{-s}\;=\;\prod_{p}L_{p}(s,f),
Lp(s,f)\displaystyle L_{p}(s,f) =1jd(1αj(p)ps)1,\displaystyle\;=\;\prod_{1\leqslant j\leqslant d}(1-\alpha_{j}(p)p^{-s})^{-1},

if σ=Res>1\sigma=\operatorname{Re}s>1, where |αj(p)|1|\alpha_{j}(p)|\leqslant 1. Further we assume that L(s,f)L(s,f) is entire and that it satisfies a functional equation of conductor Δ3\Delta\geqslant 3, where

γ(s,f)=πds/21jdΓ(s+κj2),Reκj0,\gamma(s,f)\;=\;\pi^{-ds/2}\prod_{1\leqslant j\leqslant d}\Gamma\Big{(}\frac{s+\kappa_{j}}{2}\Big{)},\qquad\operatorname{Re}\kappa_{j}\geqslant 0,

is its gamma factor. This means (see Chapter 5 of [23])

Λ(s,f)=Δs/2γ(s,f)L(s,f)=εΛ(1s,f¯)\Lambda(s,f)\;=\;\Delta^{s/2}\gamma(s,f)L(s,f)\;=\;\varepsilon\Lambda(1-s,\overline{f})

where ε\varepsilon denotes the root number, |ε|=1|\varepsilon|=1.

Our goal is to approximate L(1,f)L(1,f) by the finite product

(71) E(x)=p<xLp(1,f)E(x)\;=\;\prod_{p<x}L_{p}(1,f)

when xx comparable to Δ\Delta in the logarithmic scale. Indeed, assuming the Riemann hypothesis for L(s,f)L(s,f) shows that

L(1,f)E(x)if x(ΔlogΔ)2L(1,f)\;\sim\;E(x)\qquad\text{if }\;x\;\gg\;(\Delta\log\Delta)^{2}

as Δ\Delta\to\infty. By comparison, our result will be unconditional. We do not require any zero-density estimate for L(s,f)L(s,f), only that it have a zero-free region of “classical” type; that is, we assume that Hypothesis H(c)\text{H}(c) holds for L(s,f)L(s,f).

Proposition A.1:

Suppose L(s,f)L(s,f) is entire and that it satisfies Hypothesis H(c)\text{H}(c). Let Δ=xθ\Delta=x^{\theta} for some 0<θ<10<\theta<1. Then

(72) L(1,f)=E(x)eη(x),L(1,f)\;=\;E(x)e^{\eta(x)},

where

(73) η(x)exp(c3θ)+1logx.\eta(x)\;\ll\;\exp\Big{(}\frac{-c}{3\theta}\Big{)}\;+\;\frac{1}{\log x}.

The implied constant above depends on the degree dd and parameters κ1,,κd\kappa_{1},\dots,\kappa_{d}.

Remark:

A version of (73) with explicit numerical constants is given in [13].

In this article, we use the above result only for the quadratic Dirichlet LL-function L(s,χD)L(s,\chi_{D}). In this case the result reads: assuming Hypothesis H(c)\text{H}(c) holds for L(s,χD)L(s,\chi_{D}), then for z>Dz>D we have

(74) p<z(1χD(p)p)=L(1,χD)1{1+O(exp(clogz3logD)+1logz)}.\prod_{p<z}\Big{(}1-\frac{\chi_{D}(p)}{p}\Big{)}\;=\;L(1,\chi_{D})^{-1}\Big{\{}1+O\Big{(}\exp\Big{(}\frac{-c\log z}{3\log D}\Big{)}+\frac{1}{\log z}\Big{)}\Big{\}}.

Before giving the proof of Proposition A.1, we give a corollary that we need at other points in the work.

Corollary A.2:

Assume Hypothesis H(c)\text{H}(c) holds for L(s,χD)L(s,\chi_{D}), and that D=xθD=x^{\theta} where 0<θ<10<\theta<1. Let 0<α1<α2<10<\alpha_{1}<\alpha_{2}<1. Then for xx sufficiently large we have

(75) xα1p<xα2χD(p)pexp(cα13θ).\sum_{x^{\alpha_{1}}\leqslant p<x^{\alpha_{2}}}\frac{\chi_{D}(p)}{p}\;\ll\;\exp\Big{(}\frac{-c\alpha_{1}}{3\theta}\Big{)}.

If ϕ\phi is a smooth function supported on [α1,α2][\alpha_{1},\alpha_{2}], then we have

(76) pχD(p)pϕ(logplogx)(α2α1)2exp(cα1θ).\sum_{p}\frac{\chi_{D}(p)}{p}\phi\Big{(}\frac{\log p}{\log x}\Big{)}\;\ll\;(\alpha_{2}-\alpha_{1})^{-2}\exp\Big{(}\frac{-c\alpha_{1}}{\theta}\Big{)}.
Proof of Corollary A.2.

For the function L(s,χD)L(s,\chi_{D}), we see from (72) that

η(x)=pxlog(1χD(p)p1)=pxχD(p)p+O(1x).\eta(x)\;=\;-\sum_{p\geqslant x}\log(1-\chi_{D}(p)p^{-1})\;=\;\sum_{p\geqslant x}\frac{\chi_{D}(p)}{p}\;+\;O\Big{(}\frac{1}{x}\Big{)}.

Then (75) follows from (73). The bound (76) follows from minor adaptations of the proof of Proposition A.1. ∎

To prove Proposition A.1, we begin with

Lemma A.3:

Let ρ=β+iγ\rho=\beta+i\gamma denote a zero of L(s,f)L(s,f). Then for all x3x\geqslant 3,

ρ(1+(1β)logx)1(1+(|γ|logx)2)1d+θ2+O(1/logx),\sum_{\rho}(1+(1-\beta)\log x)^{-1}(1+(|\gamma|\log x)^{2})^{-1}\;\leqslant\;d+\frac{\theta}{2}+O(1/\log x),

where the sum ρ\sum_{\rho} is taken over nontrivial zeros of L(s,f)L(s,f), θ=logΔ/logx\theta=\log\Delta/\log x, and the implied constant depends only on the gamma factor parameters κ1,,κd\kappa_{1},\dots,\kappa_{d}.

Proof of Lemma A.3.

For σ=Res>1\sigma=\operatorname{Re}s>1 we have

LL(s,f)=n1Λf(n)ns-\frac{L}{L}(s,f)\;=\;\sum_{n\geqslant 1}\Lambda_{f}(n)n^{-s}

with Λf(n)\Lambda_{f}(n) supported on prime powers,

Λf(pk)\displaystyle\Lambda_{f}(p^{k}) =1jdαj(p)klogp,\displaystyle\;=\;\sum_{1\leqslant j\leqslant d}\alpha_{j}(p)^{k}\log p,
Λf(p)\displaystyle\Lambda_{f}(p) =(α1(p)++αd(p))logp=λf(p)logp.\displaystyle\;=\;(\alpha_{1}(p)+\cdots+\alpha_{d}(p))\log p=\lambda_{f}(p)\log p.

Hence |Λf(n)|dΛ(n)|\Lambda_{f}(n)|\leqslant d\Lambda(n) and

(77) |LL(σ,f)|d|ζζ(σ)|=dσ1+O(1).\Big{|}\frac{L^{\prime}}{L}(\sigma,f)\Big{|}\;\leqslant\;d\Big{|}\frac{\zeta^{\prime}}{\zeta}(\sigma)\Big{|}\;=\;\frac{d}{\sigma-1}+O(1).

On the other hand the Hadamard product yields

LL(s,f)=12logΔ+γγ(s,f)bρ0,1(1sρ+1ρ).-\frac{L^{\prime}}{L}(s,f)\;=\;\frac{1}{2}\log\Delta+\frac{\gamma^{\prime}}{\gamma}(s,f)-b-\sum_{\rho\neq 0,1}\Big{(}\frac{1}{s-\rho}+\frac{1}{\rho}\Big{)}.

Note the contribution of the trivial zeros ρ\rho to the above sum is O(1)O(1). For the constant bb, we have

Reb=ρ0Re1ρ.\operatorname{Re}b\;=\;-\sum_{\rho\neq 0}\operatorname{Re}\frac{1}{\rho}.

Therefore for s=σs=\sigma with 1<σ21<\sigma\leqslant 2 we have

(78) Reρ1σρ=12logΔ+ReLL(σ,f)+O(1),\operatorname{Re}\sum_{\rho}\frac{1}{\sigma-\rho}\;=\;\frac{1}{2}\log\Delta+\operatorname{Re}\frac{L^{\prime}}{L}(\sigma,f)+O(1),

where the ρ\rho summation now runs over all nontrivial zeros of L(s,f)L(s,f). With s=σ>1s=\sigma>1 and ρ=β+iγ\rho=\beta+i\gamma we have

(79) Re1σρ=σβ(σβ)2+γ21σ1(1+1βσ1)1(1+(|γ|σ1)2)1.\operatorname{Re}\frac{1}{\sigma-\rho}=\frac{\sigma-\beta}{(\sigma-\beta)^{2}+\gamma^{2}}\;\geqslant\;\frac{1}{\sigma-1}\Big{(}1+\frac{1-\beta}{\sigma-1}\Big{)}^{-1}\Big{(}1+\Big{(}\frac{|\gamma|}{\sigma-1}\Big{)}^{2}\Big{)}^{-1}.

Thus now we choose s=σ=1+1/logxs=\sigma=1+1/\log x (where x3x\geqslant 3) and combine (77), (78), and (79) to prove the lemma. ∎

Proof of Proposition A.1.

We have

L(1,f)=E(x)eη(x),L(1,f)\;=\;E(x)e^{\eta(x)},

where E(x)E(x) is defined in (71), and

η(x)=log(pxLp(1,f)).\eta(x)\;=\;\log\Big{(}\prod_{p\geqslant x}L_{p}(1,f)\Big{)}.

We put

K(x)=nxΛf(n)nlogn.K(x)\;=\;\sum_{n\geqslant x}\frac{\Lambda_{f}(n)}{n\log n}.

Note that this sum converges by virtue of Hypothesis H(c)\text{H}(c)—it does not converge absolutely. We then check that

|K(x)η(x)|\displaystyle|K(x)-\eta(x)|\; p<xklogxlogpdkpkdp<xlogplogxklogxlogp1pk\displaystyle\leqslant\;\sum_{p<x}\sum_{k\geqslant\frac{\log x}{\log p}}\frac{d}{kp^{k}}\;\leqslant\;d\sum_{p<x}\frac{\log p}{\log x}\sum_{k\geqslant\frac{\log x}{\log p}}\frac{1}{p^{k}}
dp<xlogplogx2x1logx,\displaystyle\leqslant\;d\sum_{p<x}\frac{\log p}{\log x}\cdot\frac{2}{x}\;\ll\;\frac{1}{\log x},

and therefore we have

η(x)=K(x)+O(1logx),\eta(x)\;=\;K(x)\;+\;O\Big{(}\frac{1}{\log x}\Big{)},

where the implied constant depends on dd. To estimate K(x)K(x), we put

Kϕ(x)=n1ϕ(lognlogx)Λf(n)nlogn,K_{\phi}(x)\;=\;\sum_{n\geqslant 1}\phi\Big{(}\frac{\log n}{\log x}\Big{)}\frac{\Lambda_{f}(n)}{n\log n},

where ϕ(u)\phi(u) is a smooth function satisfying ϕ(u)=0\phi(u)=0 if u1u\leqslant 1, 0ϕ(u)10\leqslant\phi(u)\leqslant 1 if 1u1+ε1\leqslant u\leqslant 1+\varepsilon, and ϕ(u)=1\phi(u)=1 if u1+εu\geqslant 1+\varepsilon, with 0<ε<10<\varepsilon<1. Then we have

|K(x)Kϕ(x)|\displaystyle|K(x)-K_{\phi}(x)|\; dxnx1+εΛ(n)nlogndxpx1+ε1p+dpxkmax(2,logxlogp)1kpk\displaystyle\leqslant\;d\sum_{x\leqslant n\leqslant x^{1+\varepsilon}}\frac{\Lambda(n)}{n\log n}\;\leqslant\;d\sum_{x\leqslant p\leqslant x^{1+\varepsilon}}\frac{1}{p}\;+\;d\sum_{p\leqslant x}\sum_{k\geqslant\max(2,\frac{\log x}{\log p})}\frac{1}{kp^{k}}
dlog(1+ε)+dpxlogplogxk21pkε+1logx,\displaystyle\leqslant\;d\log(1+\varepsilon)\;+\;d\sum_{p\leqslant x}\frac{\log p}{\log x}\sum_{k\geqslant 2}\frac{1}{p^{k}}\;\ll\;\varepsilon\;+\;\frac{1}{\log x},

where again the implied constant depends on dd. Thus now we have

(80) η(x)=Kϕ(x)+O(ε+1logx).\eta(x)\;=\;K_{\phi}(x)\;+\;O\Big{(}\varepsilon+\frac{1}{\log x}\Big{)}.

To estimate Kϕ(x)K_{\phi}(x), we apply the explicit formula for L(s,f)L(s,f),

(81) Kϕ(x)=ρΦ~(ρ),K_{\phi}(x)\;=\;-\sum_{\rho}\widetilde{\Phi}(\rho),

where Φ~(s)\widetilde{\Phi}(s) is the Mellin transform of

Φ(y)=ϕ(logylogx)1ylogy.\Phi(y)\;=\;\phi\Big{(}\frac{\log y}{\log x}\Big{)}\frac{1}{y\log y}.

We have

Φ~(s)\displaystyle\widetilde{\Phi}(s) =0Φ(y)ys1dy=ϕ(u)xu(s1)u1du\displaystyle\;=\;\int_{0}^{\infty}\Phi(y)y^{s-1}\mathop{}\!\mathrm{d}y\;=\;\int_{-\infty}^{\infty}\phi(u)x^{u(s-1)}u^{-1}\mathop{}\!\mathrm{d}u
(82) =(1+(1s)logx)31(euu1ϕ(u))′′′euxu(s1)du\displaystyle\;=\;(1+(1-s)\log x)^{-3}\int_{1}^{\infty}(e^{u}u^{-1}\phi(u))^{\prime\prime\prime}e^{-u}x^{u(s-1)}\mathop{}\!\mathrm{d}u

by partial integration three times. We can choose ϕ(u)\phi(u) so that for 1u1+ε1\leqslant u\leqslant 1+\varepsilon,

(euu1ϕ(u))′′′euε3,(e^{u}u^{-1}\phi(u))^{\prime\prime\prime}\;\ll\;e^{u}\varepsilon^{-3},

and for u1+εu\geqslant 1+\varepsilon we have

(euu1ϕ(u))′′′eu.(e^{u}u^{-1}\phi(u))^{\prime\prime\prime}\;\ll\;e^{u}.

Using these bounds in (82), we get

(83) Φ~(s)|1+(1s)logx|3(ε2+((1σ)logx)1)xσ1.\widetilde{\Phi}(s)\;\ll\;|1+(1-s)\log x|^{-3}\Big{(}\varepsilon^{-2}+((1-\sigma)\log x)^{-1}\Big{)}x^{\sigma-1}.

We also have the bound

(84) |Φ~(s)|ε2|(1s)logx|3,|\widetilde{\Phi}(s)|\;\ll\;\varepsilon^{-2}|(1-s)\log x|^{-3},

which is derived in a similar way, using the identity

Φ~(s)=((1s)logx)31(u1ϕ(u))′′′xu(s1)du\widetilde{\Phi}(s)\;=\;((1-s)\log x)^{-3}\int_{1}^{\infty}(u^{-1}\phi(u))^{\prime\prime\prime}x^{u(s-1)}\mathop{}\!\mathrm{d}u

and the bound

1(u1ϕ(u))′′′duε2.\int_{1}^{\infty}(u^{-1}\phi(u))^{\prime\prime\prime}\mathop{}\!\mathrm{d}u\;\ll\;\varepsilon^{-2}.

Recall that the number of zeros ρ=β+iγ\rho=\beta+i\gamma with 0<β<10<\beta<1 and t<|γ|t+1t<|\gamma|\leqslant t+1 is O(logΔ(|t|+1))O(\log\Delta(|t|+1)) for all tt. Using this with (84), we see that the contribution of ρ=β+iγ\rho=\beta+i\gamma with |γ|>1|\gamma|>1 to the explicit formula (81) is bounded by

(85) k1k<|γ|k+1Φ~(ρ)ε2(logΔ)k1k2(logx)3ε2(logx)2.\sum_{k\geqslant 1}\sum_{k<|\gamma|\leqslant k+1}\widetilde{\Phi}(\rho)\;\ll\;\varepsilon^{-2}(\log\Delta)\sum_{k\geqslant 1}k^{-2}(\log x)^{-3}\;\ll\;\varepsilon^{-2}(\log x)^{-2}.

The remaining zeros ρ=β+iγ\rho=\beta+i\gamma satisfy (8), so for these zeros we have

(86) |xρ1|exp(clogxlogΔ)=exp(cθ),and(1β)logxcθ.|x^{\rho-1}|\;\leqslant\;\exp\Big{(}\frac{-c\log x}{\log\Delta}\Big{)}\;=\;\exp\Big{(}\frac{-c}{\theta}\Big{)},\quad\text{and}\quad(1-\beta)\log x\;\geqslant\;\frac{c}{\theta}.

Noting that

|1+(1ρ)logx|3(1+(1β)logx)1(1+(|γ|logx)2)1,|1+(1-\rho)\log x|^{-3}\;\leqslant\;(1+(1-\beta)\log x)^{-1}(1+(|\gamma|\log x)^{2})^{-1},

we apply Lemma A.3 to see that the contribution of these zeros to (81) is

(87) ρ=β+iγ|γ|1|Φ~(ρ)|(d+θ2)(ε2+θc)exp(cθ)\sum_{\begin{subarray}{c}\rho=\beta+i\gamma\\ |\gamma|\leqslant 1\end{subarray}}|\widetilde{\Phi}(\rho)|\;\ll\;\Big{(}d+\frac{\theta}{2}\Big{)}\Big{(}\varepsilon^{-2}+\frac{\theta}{c}\Big{)}\exp\Big{(}\frac{-c}{\theta}\Big{)}

by (83) and (86). The bound above is larger than the one in (85), and hence Kϕ(x)K_{\phi}(x) is also bounded by the right-hand side of (87). Combining this bound with (80) then shows

η(x)ε+ε2exp(cθ)+1logx.\eta(x)\;\ll\;\varepsilon+\varepsilon^{-2}\exp\Big{(}\frac{-c}{\theta}\Big{)}+\frac{1}{\log x}.

Finally we choose ε=exp(c/3θ)\varepsilon=\exp(-c/3\theta), which gives (73). ∎

References

  • [1] J.-H. Ahn and S.-H. Kwon “Some explicit zero-free regions for Hecke LL-functions” In J. Number Theory 145 Elsevier, 2014, pp. 433–473
  • [2] E. Bach and J. Sorenson “Explicit bounds for primes in residue classes” In Math. Comp. 65.216, 1996, pp. 1717–1735
  • [3] H.. Bui and J.. Keating “On the mean values of Dirichlet LL-functions” In Proc. London Math. Soc. 95.2 Oxford University Press, 2007, pp. 273–298
  • [4] H.M. Bui and J.P. Keating “On the mean values of LL-functions in orthogonal and symplectic families” In Proc. London Math. Soc. 96.2 Oxford University Press, 2008, pp. 335–366
  • [5] D.. Cox “Primes of the Form x2+ny2x^{2}+ny^{2}: Fermat, Class Field Theory, and Complex Multiplication” Amer. Math. Soc., 2022
  • [6] J. Ditchen “Primes of the shape x2+ny2x^{2}+ny^{2}: the distribution on average and prime number races”, 2013
  • [7] P…. Elliot “The least prime primitive root and Linnik’s theorem” In Number Theory for the Millenium I, 2002, pp. 393–418
  • [8] E. Fogels “On the distribution of prime ideals” In Acta Arith. 7, 1962, pp. 255–269
  • [9] É. Fouvry and H. Iwaniec “Low-lying zeros of dihedral LL-functions” In Duke Math. J. 116, 2013, pp. 189–217
  • [10] J.. Friedlander and H. Iwaniec “Opera de Cribro” Amer. Math. Soc., 2010
  • [11] J.. Friedlander and H. Iwaniec “Selberg’s sieve of irregular density” In Acta Arith. 209, 2023, pp. 385–396
  • [12] J.. Friedlander and H. Iwaniec “Sifting for small primes from an arithmetic progression” In Sci. China Math. 66.12, 2023, pp. 2715–2730
  • [13] L.. Gaudet “On the least prime represented by a positive-definite binary quadratic form”, 2023
  • [14] S.. Gonek “Finite Euler products and the Riemann hypothesis” In Trans. Amer. Math. Soc. 364.4, 2012, pp. 2157–2191
  • [15] S.. Gonek, C.. Hughes and J.. Keating “A hybrid Euler-Hadamard product for the Riemann zeta function” In Duke Math. J. 136.3, 2007
  • [16] A. Granville, A. Harper and K. Soundararajan “A new proof of Hálasz’s theorem, and its consequences” In Compositio Math. 155.1 London Mathematical Society, 2019, pp. 126–163
  • [17] A. Granville and K. Soundararajan “Extreme values of |ζ(1+it)||\zeta(1+it)| In The Riemann zeta function and related themes: papers in honour of Professor K. Ramachandra, Ramanujan Math. Soc. Lect. Notes Ser., vol. 2 Ramanujan Math. Soc., Mysore, 2006, pp. 65–80
  • [18] A. Granville and K. Soundararajan “The distribution of values of L(1,χd)L(1,\chi_{d}) In Geom. Func. Anal. 5.13, 2003, pp. 992–1028
  • [19] D.. Heath-Brown “Zero-free regions for Dirichlet LL-functions, and the least prime in an arithmetic progression” In Proc. London Math. Soc. 64, 1992, pp. 265–338
  • [20] M.. Huxley “The large sieve inequality for algebraic number fields” In Mathematika 15.2 London Mathematical Society, 1968, pp. 178–187
  • [21] H. Iwaniec “Rosser’s sieve” In Acta Arith. 36.2, 1980, pp. 171–202
  • [22] H. Iwaniec “Topics in Classical Automorphic Forms” Amer. Math. Soc., 1997
  • [23] H. Iwaniec and E. Kowalski “Analytic Number Theory” Amer. Math. Soc., 2004
  • [24] H. Kadiri “Explicit zero-free regions for Dedekind zeta functions” In Int. J. Number Theory 8.1 World Scientific, 2012, pp. 125–147
  • [25] D. Koukoulopoulos “The Distribution of Prime Numbers” Amer. Math. Soc., 2019
  • [26] E. Kowalski and P. Michel “Zeros of families of automorphic LL-functions close to 1” In Pacific J. Math. 207.2, 2002, pp. 411–431
  • [27] J.. Lagarias, H.. Montgomery and A.. Odlyzko “A bound for the least prime ideal in the Chebotarev density theorem” In Invent. Math. 54.3, 1979, pp. 271–296
  • [28] J.. Lagarias and A.. Odlyzko “Effective versions of the Chebotarev density theorem” In Algebraic number fields: LL-functions and Galois properties (Proc. Sympos., Univ. Durham, Durham, 1975) 7, 1977, pp. 409–464
  • [29] Y.. Linnik “On the least prime in an arithmetic progression. I. The basic theorem” In Rec. Math. [Mat. Sbornik] N.S. 15.2, 1944, pp. 139–178
  • [30] K. Matomäki, J. Merikoski and J. Teräväinen “Primes in arithmetic progressions and short intervals without LL-functions” In arXiv:2401.17570 [math.NT], 2024
  • [31] K. Matomäki and J. Teräväinen “Products of primes in arithmetic progressions” In J. reine angew. Math. 2024.808, 2024, pp. 193–240
  • [32] S. Sachpazis “A pretentious proof of Linnik’s estimate for primes in arithmetic progressions” In Mathematika 69.3, 2023, pp. 879–902
  • [33] W. Schaal “On the large sieve method in algebraic number fields” In J. Number Theory 2.3 Elsevier, 1970, pp. 249–270
  • [34] P.. Schumer “On the large sieve inequality in an algebraic number field” In Mathematika 33.1 London Mathematical Society, 1986, pp. 31–54
  • [35] J. Thorner and A. Zaman “An explicit bound for the least prime ideal in the Chebotarev density theorem” In Algebra & Number Theory 11.5, 2017, pp. 1135–1197
  • [36] A. Weiss “The least prime ideal” In J. reine angew. Math. 338, 1983, pp. 56–94
  • [37] T. Xylouris “On Linnik’s constant” In Acta Arith. 1.150, 2011, pp. 65–91
  • [38] A. Zaman “Analytic estimates for the Chebotarev Density Theorem and their applications”, 2017
  • [39] A. Zaman “Bounding the least prime ideal in the Chebotarev density theorem” In Funct. Approx. Comment. Math. 57.1, 2017, pp. 115–142
  • [40] A. Zaman “Explicit estimates for the zeros of Hecke LL-functions” In J. Number Theory 162 Elsevier, 2016, pp. 312–375