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

Sums of singular series along arithmetic progressions and with smooth weights

Vivian Kuperberg
Abstract.

Sums of the singular series constants that appear in the Hardy–Littlewood kk-tuples conjectures have long been studied in connection to the distribution of primes. We study constrained sums of singular series, where the sum is taken over sets whose elements are specified modulo rr or weighted by smooth functions. We show that the value of the sum is governed by incidences modulo rr of elements of the set in the case of arithmetic progressions and by pairings of the smooth functions in the case of weights. These sums shed light on sums of singular series in other formats.

The author is supported by NSF GRFP grant DGE-1656518 as well as the NSF Mathematical Sciences Research Program through the grant DMS-2202128, and would like to thank Kannan Soundararajan for many helpful comments and discussions.

1. Introduction

The Hardy–Littlewood kk-tuples conjectures, and the constants known as singular series that appear within them, have long been studied in connection to the distribution of primes. These conjectures state that for any kk-tuple ={h1,,hk}\mathcal{H}=\{h_{1},\dots,h_{k}\} of distinct integers, the number of kk-tuples of primes of the form (n+h1,,n+hk)(n+h_{1},\dots,n+h_{k}), with nxn\leq x, is given by

(1) nxi=1kΛ(n+hi)=𝔖()x+o(1)x,\sum_{n\leq x}\prod_{i=1}^{k}\Lambda(n+h_{i})=\mathfrak{S}(\mathcal{H})x+o(1)x,

where 𝔖()\mathfrak{S}(\mathcal{H}) is the singular series

(2) 𝔖()=p prime1ν(p)/p(11/p)k,\mathfrak{S}(\mathcal{H})=\prod_{p\text{ prime}}\frac{1-\nu_{\mathcal{H}}(p)/p}{(1-1/p)^{k}},

and ν(p)\nu_{\mathcal{H}}(p) denotes the number of distinct residue classes modulo pp occupied by the elements of \mathcal{H}.

In [1], Gallagher showed that the Hardy–Littlewood conjectures imply that the distribution of primes in intervals of size h=λlogxh=\lambda\log x is Poissonian for fixed λ\lambda, by showing that the singular series is 11 on average. In particular, he showed that for any fixed kk,

[1,h]||=k𝔖()[1,h]||=k1.\sum_{\begin{subarray}{c}\mathcal{H}\subset[1,h]\\ |\mathcal{H}|=k\end{subarray}}\mathfrak{S}(\mathcal{H})\sim\sum_{\begin{subarray}{c}\mathcal{H}\subset[1,h]\\ |\mathcal{H}|=k\end{subarray}}1.

In 2004, Montgomery and Soundararajan [7] used a more refined estimate of sums of singular series to show that when hh is in a larger regime, with h/logxh/\log x\to\infty but h=o(x)h=o(x), the Hardy–Littlewood conjectures imply that the distribution of primes, counted with von Mangoldt weights, becomes Gaussian with mean h\sim h and variance hlogxh\sim h\log\frac{x}{h}, matching numerical data. Instead of the singular series itself, they considered alternating sums of singular series, defining 𝔖0():=𝒥(1)|𝒥|𝔖(𝒥).\mathfrak{S}_{0}(\mathcal{H}):=\sum_{\mathcal{J}\subset\mathcal{H}}(-1)^{|\mathcal{H}\setminus\mathcal{J}|}\mathfrak{S}(\mathcal{J}). These sums have the effect of subtracting the main term from the outset, making it easier to understand lower-order contributions. The analogous Hardy–Littlewood conjecture says that

nxi=1k(Λ(n+hi)1)=𝔖0()x+o(x),\sum_{n\leq x}\prod_{i=1}^{k}\left(\Lambda(n+h_{i})-1\right)=\mathfrak{S}_{0}(\mathcal{H})x+o(x),

so that each term on the left-hand side has expectation 0.

Montgomery and Soundararajan [7] showed that for fixed kk,

(3) [1,h]||=k𝔖0()=μk(hlogh+A)k/2+Ok(hk/21/(7k)+ε),\sum_{\begin{subarray}{c}\mathcal{H}\subset[1,h]\\ |\mathcal{H}|=k\end{subarray}}\mathfrak{S}_{0}(\mathcal{H})=\mu_{k}(-h\log h+A)^{k/2}+O_{k}(h^{k/2-1/(7k)+\varepsilon}),

where μk=13(k1)\mu_{k}=1\cdot 3\cdots(k-1) if kk is even, and 0 if kk is odd, and A=2γ0log2πA=2-\gamma_{0}-\log 2\pi, with γ0\gamma_{0} the Euler–Mascheroni constant. Their results depend crucially on work of Montgomery and Vaughan [8] on the distribution of reduced residues mod qq in short intervals.

Here we develop analogs of the work of Montgomery and Vaughan as well as that of Montgomery and Soundararajan by studying sums of singular series with added conditions on the set \mathcal{H}. First, instead of summing over all subsets of [1,h][1,h], we restrict to sets whose elements lie in arithmetic progressions. For a fixed modulus rr and congruence classes c1,,ckc_{1},\dots,c_{k} mod rr, we study the sum

(4) Rk(h;r,c1,,ck):=={h1,,hk}[1,h]||=khicimodr𝔖0,r(),R_{k}(h;r,c_{1},\dots,c_{k}):=\sum_{\begin{subarray}{c}\mathcal{H}=\{h_{1},\dots,h_{k}\}\subset[1,h]\\ |\mathcal{H}|=k\\ h_{i}\equiv c_{i}\bmod{r}\end{subarray}}\mathfrak{S}_{0,r}(\mathcal{H}),

where 𝔖0,r()=𝒥𝔖r(𝒥)(1)|𝒥|\mathfrak{S}_{0,r}(\mathcal{H})=\sum_{\mathcal{J}\subset\mathcal{H}}\mathfrak{S}_{r}(\mathcal{J})(-1)^{|\mathcal{H}\setminus\mathcal{J}|} and 𝔖r()\mathfrak{S}_{r}(\mathcal{H}) is the singular series of \mathcal{H} away from rr, given by

(5) 𝔖r():=p primepr1ν(p)/p(11/p)k.\mathfrak{S}_{r}(\mathcal{H}):=\prod_{\begin{subarray}{c}p\text{ prime}\\ p\nmid r\end{subarray}}\frac{1-\nu_{\mathcal{H}}(p)/p}{(1-1/p)^{k}}.

The study of these sums is of interest for two reasons. First, they appear in the work of Lemke Oliver and Soundararajan in [5] on bias in the distribution of consecutive primes in arithmetic progressions. Specifically, Lemke Oliver and Soundararajan conjecture that if π(x;q,(a,b))\pi(x;q,(a,b)) is defined as the number of primes pxp\leq x with pamodqp\equiv a\bmod q and pnextbmodqp_{\text{next}}\equiv b\bmod q, then

π(x;q,(a,b))1q2xα(y)ϵq(a,b)(qϕ(q)logy)2𝒟(a,b;y)dy,\pi(x;q,(a,b))\sim\frac{1}{q}\int_{2}^{x}\alpha(y)^{\epsilon_{q}(a,b)}\left(\frac{q}{\phi(q)\log y}\right)^{2}\mathcal{D}(a,b;y)\mathrm{d}y,

where α(y):1qϕ(q)logy\alpha(y):-1-\frac{q}{\phi(q)\log y}, ϵq(a,b)\epsilon_{q}(a,b) is a constant defined only in terms of q,a,q,a, and bb, and

𝒟(a,b;y):=h>0hbamodq𝒜{0,h}𝒯[1,h1](t+a,q)=1t𝒯(1)|𝒯|𝔖0,q(𝒜𝒯)(qϕ(q)α(y)logy)|𝒯|α(y)hϕ(q)/q.\mathcal{D}(a,b;y):=\sum_{\begin{subarray}{c}h>0\\ h\equiv b-a\bmod q\end{subarray}}\sum_{\mathcal{A}\subset\{0,h\}}\sum_{\begin{subarray}{c}\mathcal{T}\subset[1,h-1]\\ (t+a,q)=1\forall t\in\mathcal{T}\end{subarray}}(-1)^{|\mathcal{T}|}\mathfrak{S}_{0,q}(\mathcal{A}\cup\mathcal{T})\left(\frac{q}{\phi(q)\alpha(y)\log y}\right)^{|\mathcal{T}|}\alpha(y)^{h\phi(q)/q}.

They proceed to heuristically estimate 𝒟(a,b;y)\mathcal{D}(a,b;y), and thus π(x;q,(a,b))\pi(x;q,(a,b)), by estimating a weighted version of R2(h;r,c1,c2)R_{2}(h;r,c_{1},c_{2}). Our results generalize these arguments by estimating Rk(h;r,c1,c2)R_{k}(h;r,c_{1},c_{2}) for all kk, which is necessary for understanding some of the error terms appearing in Lemke Oliver and Soundararajan’s heuristic. Secondly, restricting sums of singular series in this manner may shed light on other questions about sums of singular series. We show in Theorem 1.2 that the asymptotics for (4) are governed by incidences among the cic_{i}’s mod rr. As discussed in [4], we do not yet know the asymptotic average size of sums of 𝔖0()\mathfrak{S}_{0}(\mathcal{H}) when |||\mathcal{H}| is odd. The results of these more refined averages of singular series may clarify where the main term should be coming from for sums of singular series with an odd number of terms.

The sums over 𝔖0,r\mathfrak{S}_{0,r} admit the expansion

1h1,,hkhhicimodridistinct𝔖0,r({h1,,hk})=q1,,qkqi>1(qi,r)=1μ(qi)ϕ(qi)a1,,ak1aiqi(ai,qi)=1iai/qi1h1,,hkhhicimodrdistincti=1ke(aihiqi).\sum_{\begin{subarray}{c}1\leq h_{1},\dots,h_{k}\leq h\\ h_{i}\equiv c_{i}\bmod r\>\forall i\\ \text{distinct}\end{subarray}}\mathfrak{S}_{0,r}(\{h_{1},\dots,h_{k}\})=\sum_{\begin{subarray}{c}q_{1},\dots,q_{k}\\ q_{i}>1\\ (q_{i},r)=1\end{subarray}}\frac{\mu(q_{i})}{\phi(q_{i})}\sum_{\begin{subarray}{c}a_{1},\dots,a_{k}\\ 1\leq a_{i}\leq q_{i}\\ (a_{i},q_{i})=1\\ \sum_{i}a_{i}/q_{i}\in{\mathbb{Z}}\end{subarray}}\sum_{\begin{subarray}{c}1\leq h_{1},\dots,h_{k}\leq h\\ h_{i}\equiv c_{i}\bmod r\text{distinct}\end{subarray}}\prod_{i=1}^{k}e\left(\frac{a_{i}h_{i}}{q_{i}}\right).

Following [7] and [8], we first consider a related quantity, where the summands qiq_{i} are restricted to divide a secondary modulus q>1q>1 and the hih_{i}’s are not necessarily distinct, to get

(6) Vk(q,h;r,c1,,ck):=q1,,qk1<qi|q(qi,r)=1μ(qi)ϕ(qi)a1,,ak1aiqi(ai,qi)=1iai/qii=1kEr,ci(aihiqi),V_{k}(q,h;r,c_{1},\dots,c_{k}):=\sum_{\begin{subarray}{c}q_{1},\dots,q_{k}\\ 1<q_{i}|q\\ (q_{i},r)=1\end{subarray}}\frac{\mu(q_{i})}{\phi(q_{i})}\sum_{\begin{subarray}{c}a_{1},\dots,a_{k}\\ 1\leq a_{i}\leq q_{i}\\ (a_{i},q_{i})=1\\ \sum_{i}a_{i}/q_{i}\in{\mathbb{Z}}\end{subarray}}\prod_{i=1}^{k}E_{r,c_{i}}\left(\frac{a_{i}h_{i}}{q_{i}}\right),

where for α\alpha\in{\mathbb{R}},

(7) Er,ci(α):=mhmcimodre(mα).E_{r,c_{i}}(\alpha):=\sum_{\begin{subarray}{c}m\leq h\\ m\equiv c_{i}\bmod r\end{subarray}}e(m\alpha).

In order to state our result, we will need to fix some notation concerning perfect matchings of [1,k][1,k]. For k1k\geq 1, a perfect matching σ\sigma of [1,k][1,k] is a set σ={(i,j)}\sigma=\{(i,j)\} of unordered pairs in [1,k][1,k] so that each element is paired with exactly one other element, i.e. each ii appears in exactly one pair (i,j)(i,j), and iji\neq j. Since each pair (i,j)σ(i,j)\in\sigma is unordered, we will generally choose to write the representative with i<ji<j, so that σ{(i,j)}\sigma\{(i,j)\} with i<ji<j. Let k\mathcal{B}_{k} denote the set of perfect matchings of [1,k][1,k], so that

(8) k:={σ={(i,j)}:(i,j)σ,1i<jki[1,k],!j[1,k] with (i,j)σ or (j,i)σ}.\mathcal{B}_{k}:=\left\{\sigma=\{(i,j)\}:\begin{array}[]{l}\forall(i,j)\in\sigma,1\leq i<j\leq k\\ \forall i\in[1,k],\exists!j\in[1,k]\text{ with }(i,j)\in\sigma\text{ or }(j,i)\in\sigma\end{array}\right\}.

Note that when kk is odd, k=\mathcal{B}_{k}=\varnothing. Moreover, for a set of integers {a1,,ak}\{a_{1},\dots,a_{k}\}, we will denote by (a1,,ak)\mathcal{B}(a_{1},\dots,a_{k}) the set of matchings of {a1,,ak}\{a_{1},\dots,a_{k}\} into pairs, so that k=([1,k])\mathcal{B}_{k}=\mathcal{B}([1,k]).

In Section 2, we prove the following result, which mirrors Theorem 1 of [7].

Theorem 1.1.

Fix a modulus r1r\geq 1, an integer k1k\geq 1, and kk congruence classes c1,,ckc_{1},\dots,c_{k} modulo rr. Define k\mathcal{B}_{k} as in (8). Let q1q\geq 1 be a squarefree integer with (r,q)=1(r,q)=1, and define Vk(q,h;r,c1,,ck)V_{k}(q,h;r,c_{1},\dots,c_{k}) as in (6). For h3h\geq 3,

Vk(q,h;r,c1,,ck)=σk(i,j)σV2(q,h;r,ci,cj)+Or,k(hk/21/(7k)(qϕ(q))2k+k/2).V_{k}(q,h;r,c_{1},\dots,c_{k})=\sum_{\begin{subarray}{c}\sigma\in\mathcal{B}_{k}\end{subarray}}\prod_{(i,j)\in\sigma}V_{2}(q,h;r,c_{i},c_{j})+O_{r,k}\left(h^{k/2-1/(7k)}\left(\frac{q}{\phi(q)}\right)^{2^{k}+k/2}\right).

In order to state our main result on the asymptotics of Rk(h;r,c1,,ck)R_{k}(h;r,c_{1},\dots,c_{k}), we define some further notation. For 1r1\leq\ell\leq r, define

𝒞:={i:cimodr}.\mathcal{C}_{\ell}:=\left\{i:c_{i}\equiv\ell\bmod r\right\}.

Note that some of the sets 𝒞\mathcal{C}_{\ell} may be empty, and that =1r𝒞=[1,k]\bigcup_{\ell=1}^{r}\mathcal{C}_{\ell}=[1,k]. We will say that a partition 𝒫={𝒮1,,𝒮M}\mathcal{P}=\{\mathcal{S}_{1},\dots,\mathcal{S}_{M}\} of [1,k][1,k] refines {𝒞}[1,k]\{\mathcal{C}_{\ell}\}_{\ell\in[1,k]} if for each 𝒮m𝒫\mathcal{S}_{m}\in\mathcal{P}, there exists some \ell with 𝒮m𝒞\mathcal{S}_{m}\subset\mathcal{C}_{\ell}; note that \ell is then unique. For such a partition, write 𝒫{𝒞}[1,k]\mathcal{P}\preceq\{\mathcal{C}_{\ell}\}_{\ell\in[1,k]} and define c(𝒮m)c(\mathcal{S}_{m}) to be the value \ell with 𝒮m𝒞\mathcal{S}_{m}\subset\mathcal{C}_{\ell}.

Theorem 1.2.

Fix a modulus r1r\geq 1 and an integer k1k\geq 1, as well as kk congruence classes c1,,ckc_{1},\dots,c_{k} modulo rr. Define Rk(h;r,c1,,ck)R_{k}(h;r,c_{1},\dots,c_{k}) by (4). Then for h3h\geq 3,

Rk(h;r,c1,,ck)=\displaystyle R_{k}(h;r,c_{1},\dots,c_{k})= 0jk/2(1)j𝒫{𝒞}[1,k]𝒫={𝒮1,,𝒮kj}|𝒮m|=21mj|𝒮m|=1j<mkj(hrd|Qd>1μ(d)2ϕ(d))j\displaystyle\sum_{0\leq j\leq k/2}(-1)^{j}\sum_{\begin{subarray}{c}\mathcal{P}\preceq\{\mathcal{C}_{\ell}\}_{\ell\in[1,k]}\\ \mathcal{P}=\{\mathcal{S}_{1},\dots,\mathcal{S}_{k-j}\}\\ |\mathcal{S}_{m}|=2\>\forall 1\leq m\leq j\\ |\mathcal{S}_{m}|=1\>\forall j<m\leq k-j\end{subarray}}\left(\frac{h}{r}\sum_{\begin{subarray}{c}d|Q\\ d>1\end{subarray}}\frac{\mu(d)^{2}}{\phi(d)}\right)^{j}
(9) σ(j+1,,kj)(i1,i2)σV2(Q,h;r,c(𝒮i1),c(𝒮i2)),+Or,k(hk/21/(7k)+ε).\displaystyle\sum_{\sigma\in\mathcal{B}(j+1,\dots,k-j)}\prod_{(i_{1},i_{2})\in\sigma}V_{2}(Q,h;r,c(\mathcal{S}_{i_{1}}),c(\mathcal{S}_{i_{2}})),+O_{r,k}(h^{k/2-1/(7k)+\varepsilon}).

In particular, if #~(c1,,ck)\#\widetilde{\mathcal{B}}(c_{1},\dots,c_{k}) is the number of ways to pair the cic_{i}’s such that every pair has equal values, then

Rk(h;r,c1,,ck)=#~(c1,,ck)(hϕ(r)rlogh+C0(r)h)k/2+Or,k(hk/2(logh)k/21).R_{k}(h;r,c_{1},\dots,c_{k})=\#\widetilde{\mathcal{B}}(c_{1},\dots,c_{k})\left(-h\frac{\phi(r)}{r}\log h+C_{0}(r)h\right)^{k/2}+O_{r,k}(h^{k/2}(\log h)^{k/2-1}).
Remark.

When all the cic_{i}’s are congruent mod rr, then #~(c1,,ck)=μk\#\widetilde{\mathcal{B}}(c_{1},\dots,c_{k})=\mu_{k} and this theorem implies Theorem 2 of [7]. When the cic_{i}’s are not all congruent, the estimate depends crucially on the precise arrangement of the cic_{i}’s. Nevertheless, while the value of the main term is dependent on the cic_{i}’s, we always have the upper bound Rk(h;r,c1,,ck)r,khk/2(logh)k/2R_{k}(h;r,c_{1},\dots,c_{k})\ll_{r,k}h^{k/2}(\log h)^{k/2}.

The framework of Theorems 1.1 and 1.2 also applies to sums of singular series weighted by smooth functions. Let f1,,fk:)0f_{1},\dots,f_{k}:{\mathbb{R}}){\geq 0}\to{\mathbb{C}} be functions with compact supports contained in (0,)(0,\infty) and such that |fi^(ξ)|ξ2|\hat{f_{i}}(\xi)|\ll\xi^{-2}, where the Fourier transform f^\hat{f} is defined as

f^(ξ):=f(x)e2πixξdx.\hat{f}(\xi):=-\int_{\infty}^{\infty}f(x)e^{-2\pi ix\xi}\mathrm{d}x.

For hh\in{\mathbb{N}}, we are interested in the quantity

(10) Rk(h;f1,,fk):=h1,,hkdistincti=1kfi(hih)𝔖0({h1,,hk}).R_{k}(h;f_{1},\dots,f_{k}):=\sum_{\begin{subarray}{c}h_{1},\dots,h_{k}\in{\mathbb{Z}}\\ \text{distinct}\end{subarray}}\prod_{i=1}^{k}f_{i}\left(\frac{h_{i}}{h}\right)\mathfrak{S}_{0}(\{h_{1},\dots,h_{k}\}).

The size of Rk(h;f1,,fk)R_{k}(h;f_{1},\dots,f_{k}) as hh\to\infty is similar in structure to the analogous result for sums of singular series along arithmetic progressions: the main term is a sum over perfect matchings of [1,k][1,k], and for each matching σ\sigma the contribution is determined by interactions of fif_{i} and fjf_{j} for each pair (i,j)σ(i,j)\in\sigma.

We set some notation before stating our results. Define

(11) V(q,h;f1,,fk):=1<q1,,qk|qi=1kμ(qi)ϕ(qi)a1,,ak1aiqi(ai,qi)=1iai/qiEf1,h(a1q1)Efk,h(akqk),V(q,h;f_{1},\dots,f_{k}):=\sum_{1<q_{1},\dots,q_{k}|q}\prod_{i=1}^{k}\frac{\mu(q_{i})}{\phi(q_{i})}\sum_{\begin{subarray}{c}a_{1},\dots,a_{k}\\ 1\leq a_{i}\leq q_{i}\\ (a_{i},q_{i})=1\\ \sum_{i}a_{i}/q_{i}\in{\mathbb{Z}}\end{subarray}}E_{f_{1},h}\left(\frac{a_{1}}{q_{1}}\right)\cdots E_{f_{k},h}\left(\frac{a_{k}}{q_{k}}\right),

where for α\alpha\in{\mathbb{R}} and ff smooth, with compact support, and such that |f^(ξ)|=O(|ξ|2)|\hat{f}(\xi)|=O(|\xi|^{-2}),

(12) Ef,h(α):=m=f(mh)e(mα).E_{f,h}(\alpha):=\sum_{m=-\infty}^{\infty}f\left(\frac{m}{h}\right)e(m\alpha).

Since ff is smooth, Ef,h(α)E_{f,h}(\alpha) is a smoother indicator function of values of α\alpha that are close to 0. In particular, by Poisson summation,

Ef,h(α)=hn=f^(h(nα)).E_{f,h}(\alpha)=h\sum_{n=-\infty}^{\infty}\hat{f}(h(n-\alpha)).

By assumption, f^(ξ)=O(|ξ|2)\hat{f}(\xi)=O(|\xi|^{-2}). For any real number α\alpha, only one value of αn\alpha-n will be in the interval [1/2,1/2)[-1/2,1/2); let α¯\overline{\alpha} denote this value, so that α¯\overline{\alpha} is the representative of αmod1\alpha\bmod 1 satisfying 1/2α¯<1/2-1/2\leq\overline{\alpha}<1/2. Then

(13) Ef,h(α)=hf^(hα¯)+hn=1O((hn)2)=hf^(hα¯)+O(h1).E_{f,h}(\alpha)=h\hat{f}(-h\overline{\alpha})+h\sum_{n=1}^{\infty}O((hn)^{-2})=h\hat{f}(-h\overline{\alpha})+O(h^{-1}).

The following results about Vk(q,h;f1,,fk)V_{k}(q,h;f_{1},\dots,f_{k}) hold, analogous to Theorem 1.1 and Lemma 3.2.

Theorem 1.3.

Fix k1k\geq 1 and kk smooth functions f1,,fk:0f_{1},\dots,f_{k}:{\mathbb{R}}_{\geq 0}\to{\mathbb{C}} with compact supports supp(fi)(0,)\mathrm{supp}(f_{i})\subset(0,\infty) and such that |f^i(ξ)|=O(|ξ|2)|\hat{f}_{i}(\xi)|=O(|\xi|^{-2}) for all 1ik1\leq i\leq k. Define k\mathcal{B}_{k} as in (8). Let q1q\geq 1 be a squarefree integer, and define Vk(q,h;f1,,fk)V_{k}(q,h;f_{1},\dots,f_{k}) as in (11). For h3h\geq 3,

Vk(q,h;f1,,fk)=σk(i,j)σV2(q,h;fi,fj)+Of1,,fk(hk/21/(7k)(qϕ(q))2k+k/2).V_{k}(q,h;f_{1},\dots,f_{k})=\sum_{\begin{subarray}{c}\sigma\in\mathcal{B}_{k}\end{subarray}}\prod_{(i,j)\in\sigma}V_{2}(q,h;f_{i},f_{j})+O_{f_{1},\dots,f_{k}}\left(h^{k/2-1/(7k)}\left(\frac{q}{\phi(q)}\right)^{2^{k}+k/2}\right).
Lemma 1.4.

Fix h1h\geq 1 and let f1,f2:0f_{1},f_{2}:{\mathbb{R}}_{\geq 0}\to{\mathbb{C}} be smooth functions with compact supports supp(fi)(0,)\mathrm{supp}(f_{i})\subset(0,\infty) such that |fi^(ξ)|O(|ξ|2)|\hat{f_{i}}(\xi)|\ll O(|\xi|^{-2}). Define Q:=ph2pQ:=\prod_{p\leq h^{2}}p, and define V2(q,h;f1,f2)V_{2}(q,h;f_{1},f_{2}) via (11). Then

(14) V2(Q,h;f1,f2)=(f1^(0)f2^(0)+{f}(2))h2{f}(1)2hlogh+Of1,f2(h),V_{2}(Q,h;f_{1},f_{2})=(-\hat{f_{1}}(0)\hat{f_{2}}(0)+\{\mathcal{M}f\}(2))h^{2}-\frac{\{\mathcal{M}f^{\prime}\}(1)}{2}h\log h+O_{f_{1},f_{2}}(h),

where {f}(s)\{\mathcal{M}f\}(s) is the Mellin transform of ff, defined by

(15) {f}(s):=0xs1f(x)dx.\{\mathcal{M}f\}(s):=\int_{0}^{\infty}x^{s-1}f(x)\mathrm{d}x.

Using precisely the same techniques as in the case of arithmetic progressions, an analog to Theorem 1.2 also holds for the smooth functions setting, where the input to each summand is clarified by Lemma 1.4.

Theorem 1.5.

Fix an integer kk and smooth functions f1,,fk:0f_{1},\dots,f_{k}:{\mathbb{R}}_{\geq 0}\to{\mathbb{C}} with compact support such that supp(fi)(0,)\mathrm{supp}(f_{i})\subset(0,\infty) and such that |fi^(ξ)|=O(|ξ|2)|\hat{f_{i}}(\xi)|=O(|\xi|^{-2}) for each 1ik1\leq i\leq k. Define Rk(h;f1,,fk)R_{k}(h;f_{1},\dots,f_{k}) via (10). Then for h3h\geq 3,

Rk(h;f1,,fk)=\displaystyle R_{k}(h;f_{1},\dots,f_{k})= 0jk/2(1)j𝒫={𝒮1,,𝒮kj}|𝒮m|=21mj|𝒮m|=1j<mkj(hd|Qd>1μ(d)2ϕ(d))j\displaystyle\sum_{0\leq j\leq k/2}(-1)^{j}\sum_{\begin{subarray}{c}\mathcal{P}=\{\mathcal{S}_{1},\dots,\mathcal{S}_{k-j}\}\\ |\mathcal{S}_{m}|=2\>\forall 1\leq m\leq j\\ |\mathcal{S}_{m}|=1\>\forall j<m\leq k-j\end{subarray}}\left(h\sum_{\begin{subarray}{c}d|Q\\ d>1\end{subarray}}\frac{\mu(d)^{2}}{\phi(d)}\right)^{j}
(16) σ(j+1,,kj)(i1,i2)σV2(Q,h;f𝒮i1,f𝒮i2),+Or,k(hk/21/(7k)+ε),\displaystyle\sum_{\sigma\in\mathcal{B}(j+1,\dots,k-j)}\prod_{(i_{1},i_{2})\in\sigma}V_{2}(Q,h;f_{\mathcal{S}_{i_{1}}},f_{\mathcal{S}_{i_{2}}}),+O_{r,k}(h^{k/2-1/(7k)+\varepsilon}),

where the sum is taken over partitions of [1,k][1,k] where each part has either 11 or 22 elements, and for |𝒮m|=1|\mathcal{S}_{m}|=1, f𝒮mf_{\mathcal{S}_{m}} denotes fjf_{j} where j𝒮mj\in\mathcal{S}_{m}.

The proofs of Theorems 1.3 and 1.5 are identical to the proofs of Theorems 1.1 and 1.2, so we omit them. However, Theorems 1.3 and 1.1, which provide asymptotics for VkV_{k} in the cases of smooth weighting and arithmetic progressions, rely on lemmas about sums of Ef,h(α)E_{f,h}(\alpha). The results are fundamentally the same, but the flavor of several of these lemmas is somewhat different in the smooth case, so we record them in Section 5. We also provide the proof of Lemma 1.4, which is similar to the proof of Lemma 3.2.

The organization of this paper is as follows. In Section 2, we prove Theorem 1.1. In Section 3, we compute certain sums of 2-term singular series which will then be used in the proof of Theorem 1.2. In Section 4, we prove Theorem 1.2. Finally, in Section 5 we discuss the smooth case.

2. Proof of Theorem 1.1

The arguments in this section closely follow the arguments in the proof of Lemma 8 in [8] and Theorem 1 in [7]. Here we outline the argument; where it differs, we present detailed explanations, but many steps are cited to [8] and [7].

The functions Er,ci(α)E_{r,c_{i}}(\alpha), defined in (7), are modular versions of sums E(α):=mhe(mα)E(\alpha):=\sum_{m\leq h}e(m\alpha). The sums E(α)E(\alpha) are large if α\alpha is close to an integer and otherwise exhibit large amounts of cancellation. Since Er,ci(α)E_{r,c_{i}}(\alpha) is summed only over mm in a set congruence class modulo rr, it will also take large values when α\alpha is close to a multiple of rr. For any c,rc,r, write

Er,c(α)\displaystyle E_{r,c}(\alpha) =mh1rn=1re((mc)nr)e(mα)\displaystyle=\sum_{\begin{subarray}{c}m\leq h\end{subarray}}\frac{1}{r}\sum_{\begin{subarray}{c}n=1\end{subarray}}^{r}e\left(\frac{(m-c)n}{r}\right)e(m\alpha)
=1rn=1re(cn/r)mhe(mα+mn/r),\displaystyle=\frac{1}{r}\sum_{\begin{subarray}{c}n=1\end{subarray}}^{r}e(-cn/r)\sum_{m\leq h}e(m\alpha+mn/r),

so that

|Er,c(α)|1rn=1rmin{h,1α+n/r}.|E_{r,c}(\alpha)|\leq\frac{1}{r}\sum_{n=1}^{r}\min\left\{h,\frac{1}{\|\alpha+n/r\|}\right\}.

Define

(17) Fr(α):=1rn=1rmin{h,1α+n/r},F_{r}(\alpha):=\frac{1}{r}\sum_{n=1}^{r}\min\left\{h,\frac{1}{\|\alpha+n/r\|}\right\},

so that |Er,c(α)|Fr(α).|E_{r,c}(\alpha)|\leq F_{r}(\alpha). We can then closely follow the analysis of Montgomery and Vaughan in [8] and Montgomery and Soundararajan in [7]. In particular, Fr(α)F_{r}(\alpha) can be bounded, up to some dependence on rr, by the same bounds as appear in Lemmas 4,5,6, and 7 of [8] and using the same arguments. We arrive at the following result.

Theorem 2.1.

Fix r,k,q1r,k,q\geq 1. Let 𝒮(q)\mathcal{S}(q) denote the set of kk-tuples (q1,,qk)(q_{1},\dots,q_{k}) such that for all ii, qi|qq_{i}|q, qi>1q_{i}>1, and the qiq_{i}’s are not equal in pairs and otherwise distinct; that is to say, there is no reordering permutation σ\sigma of the kk indices with qσ(i)=qσ(i+k/2)q_{\sigma(i)}=q_{\sigma(i+k/2)} for all 1ik/21\leq i\leq k/2, and no other equalities. Then

(q1,,qk)𝒮(q)i=1kμ(qi)2ϕ(qi)a1,,ak1aiqi(ai,qi)=1iai/qii=1kFr(aiqi)rqhk/21/7k(ϕ(q)q)k/22k.\displaystyle\sum_{\begin{subarray}{c}(q_{1},\dots,q_{k})\in\mathcal{S}(q)\end{subarray}}\prod_{i=1}^{k}\frac{\mu(q_{i})^{2}}{\phi(q_{i})}\sum_{\begin{subarray}{c}a_{1},\dots,a_{k}\\ 1\leq a_{i}\leq q_{i}\\ (a_{i},q_{i})=1\\ \sum_{i}a_{i}/q_{i}\in{\mathbb{Z}}\end{subarray}}\prod_{i=1}^{k}F_{r}\left(\frac{a_{i}}{q_{i}}\right)\ll_{r}qh^{k/2-1/7k}\left(\frac{\phi(q)}{q}\right)^{k/2-2^{k}}.

Theorem 2.1 accounts for all terms in Vk(q,h;r,c1,,ck)V_{k}(q,h;r,c_{1},\dots,c_{k}) except for those terms where the qiq_{i}’s are equal in pairs and otherwise distinct. When kk is odd, there are no such terms, so assume that kk is even. By the same argument as in [7], we can drop the assumption that the pairs are distinct, so that the terms left to be estimated are given by

σk(qi)(i,j)k1<qi|q(i,j)kμ(qi)2ϕ(qi)2(bi)(i,j)k1biqiibi/qi(i,j)kJr,ci,cj(bi,qi),\sum_{\begin{subarray}{c}\sigma\in\mathcal{B}_{k}\end{subarray}}\sum_{\begin{subarray}{c}(q_{i})_{(i,j)\in\mathcal{B}_{k}}\\ 1<q_{i}|q\end{subarray}}\prod_{(i,j)\in\mathcal{B}_{k}}\frac{\mu(q_{i})^{2}}{\phi(q_{i})^{2}}\sum_{\begin{subarray}{c}(b_{i})_{(i,j)\in\mathcal{B}_{k}}\\ 1\leq b_{i}\leq q_{i}\\ \sum_{i}b_{i}/q_{i}\in{\mathbb{Z}}\end{subarray}}\prod_{(i,j)\in\mathcal{B}_{k}}J_{r,c_{i},c_{j}}(b_{i},q_{i}),

where for each ii,

Jr,ci,cj(bi,qi):=1aiqi(ai,qi)=1(biai,qi)=1Er,ci(aiqi)Er,cj(biaiqi).J_{r,c_{i},c_{j}}(b_{i},q_{i}):=\sum_{\begin{subarray}{c}1\leq a_{i}\leq q_{i}\\ (a_{i},q_{i})=1\\ (b_{i}-a_{i},q_{i})=1\end{subarray}}E_{r,c_{i}}\left(\frac{a_{i}}{q_{i}}\right)E_{r,c_{j}}\left(\frac{b_{i}-a_{i}}{q_{i}}\right).

Each term σ\sigma in our sum is identical up to changing labels, so without loss of generality we will work with the term σ={(i,k/2+i):1ik/2}\sigma=\{(i,k/2+i):1\leq i\leq k/2\}, so that qi=qi+k/2q_{i}=q_{i+k/2} for all 1ik/21\leq i\leq k/2. This term is given by

(18) q1,,qk/21<qi|qi=1k/2μ(qi)2ϕ(qi)2b1,,bk/21biqiibi/qii=1k/2Jr,ci,ck/2+i(bi,qi).\sum_{\begin{subarray}{c}q_{1},\dots,q_{k/2}\\ 1<q_{i}|q\end{subarray}}\prod_{i=1}^{k/2}\frac{\mu(q_{i})^{2}}{\phi(q_{i})^{2}}\sum_{\begin{subarray}{c}b_{1},\dots,b_{k/2}\\ 1\leq b_{i}\leq q_{i}\\ \sum_{i}b_{i}/q_{i}\in{\mathbb{Z}}\end{subarray}}\prod_{i=1}^{k/2}J_{r,c_{i},c_{k/2+i}}(b_{i},q_{i}).

Let 𝒥[1,k/2]\mathcal{J}\subset[1,k/2] be the subset of ii’s with 0<bi<qi0<b_{i}<q_{i} instead of bi=qib_{i}=q_{i}. For i𝒥i\not\in\mathcal{J}, we have

Ji,k/2+i(qi,qi)=1aiqi(ai,qi)=1Er,ci(aiqi)Er,ck/2+i(aiqi),J_{i,k/2+i}(q_{i},q_{i})=\sum_{\begin{subarray}{c}1\leq a_{i}\leq q_{i}\\ (a_{i},q_{i})=1\end{subarray}}E_{r,c_{i}}\left(\frac{a_{i}}{q_{i}}\right)E_{r,c_{k/2+i}}\left(-\frac{a_{i}}{q_{i}}\right),

which, when summed over 1<qi|q1<q_{i}|q with the weight μ(qi)2ϕ(qi)2\frac{\mu(q_{i})^{2}}{\phi(q_{i})^{2}}, is equal to V2(q,h;r,ci,ck/2+i)V_{2}(q,h;r,c_{i},c_{k/2+i}). Thus, (18) is equal to

(19) 𝒥[1,k/2]i𝒥V2(q,h;r,ci,ck/2+i)W𝒥(q,h;r,(bi)i𝒥),\sum_{\begin{subarray}{c}\mathcal{J}\subset[1,k/2]\end{subarray}}\prod_{i\not\in\mathcal{J}}V_{2}(q,h;r,c_{i},c_{k/2+i})W_{\mathcal{J}}(q,h;r,(b_{i})_{i\in\mathcal{J}}),

where

W𝒥(q,h;r,(bi)i𝒥):=(qi)i𝒥1<qi|qi𝒥μ(qi)2ϕ(qi)2(bi)i𝒥0<bi<qii𝒥bi/qii𝒥Jci,ck/2+i(bi,qi).W_{\mathcal{J}}(q,h;r,(b_{i})_{i\in\mathcal{J}}):=\sum_{\begin{subarray}{c}(q_{i})_{i\in\mathcal{J}}\\ 1<q_{i}|q\end{subarray}}\prod_{i\in\mathcal{J}}\frac{\mu(q_{i})^{2}}{\phi(q_{i})^{2}}\sum_{\begin{subarray}{c}(b_{i})_{i\in\mathcal{J}}\\ 0<b_{i}<q_{i}\\ \sum_{i\in\mathcal{J}}b_{i}/q_{i}\in{\mathbb{Z}}\end{subarray}}\prod_{i\in\mathcal{J}}J_{c_{i},c_{k/2+i}}(b_{i},q_{i}).

The term 𝒥=\mathcal{J}=\varnothing gives the desired main term, so it remains to show that the terms with 𝒥\mathcal{J}\neq\varnothing are smaller. By following the reasoning on page 11 of [7], we get that

W𝒥(q,h;r,(bi)i𝒥)rh|𝒥|/2(qϕ(q))2|𝒥|,W_{\mathcal{J}}(q,h;r,(b_{i})_{i\in\mathcal{J}})\ll_{r}h^{|\mathcal{J}|/2}\left(\frac{q}{\phi(q)}\right)^{2^{|\mathcal{J}|}},

and that for any c1,c2modrc_{1},c_{2}\bmod r,

V2(q,h;r,c1,c2)rhqϕ(q).V_{2}(q,h;r,c_{1},c_{2})\ll_{r}h\frac{q}{\phi(q)}.

Applying these estimates to (19) completes the proof of Theorem 1.1.

3. Auxiliary lemmas: two-term computations

In order to prove Theorem 1.2, we will ultimately invoke Theorem 1.1, which will then relate quantities involving kk-term singular series to quantities involving 22-term singular series. In this section, we state two lemmas which compute, respectively, sums of two-term singular series in arithmetic progressions, and the quantity V2(Q,h;r,c1,c2)V_{2}(Q,h;r,c_{1},c_{2}) when QQ is the product of all primes below h2h^{2}, which will be applied in Section 4.

The following lemma is a computation of sums of two-term singular series to the modulus qq. Its proof is nearly identical to the proof of Proposition 2.1 in [5], so we omit it. Similar quantities were previously studied in [2] and [6].

Lemma 3.1.

Fix r,h1r,h\geq 1, and let vmodrv\bmod r be any residue class. Define

S(r,v,h):=1mhmvmodr(hm)𝔖r({0,m}).S(r,v,h):=\sum_{\begin{subarray}{c}1\leq m\leq h\\ m\equiv v\bmod r\end{subarray}}(h-m)\mathfrak{S}_{r}(\{0,m\}).

Then when v=0v=0,

S(r,v,h)=h22rh2ϕ(r)(loghr+log2π+γ0+p|rlogpp1)+Or(h1/2+ε),S(r,v,h)=\frac{h^{2}}{2r}-\frac{h}{2}\phi(r)\left(\log\frac{h}{r}+\log 2\pi+\gamma_{0}+\sum_{p|r}\frac{\log p}{p-1}\right)+O_{r}(h^{1/2+\varepsilon}),

where γ0\gamma_{0} denotes the Euler–Mascheroni constant. Meanwhile, if v0v\neq 0, then if d=(v,r)d=(v,r),

S(r,v,h)=h22r\displaystyle S(r,v,h)=\frac{h^{2}}{2r} h2ϕ(r)r1ϕ(r/d)Λ(r/d)\displaystyle-\frac{h}{2}\frac{\phi(r)}{r}\frac{1}{\phi(r/d)}\Lambda(r/d)
+hϕ(r/d)χχ0modr/dχ¯(v/d)L(0,χ)L(1,χ)Ar,χ+Or(h1/2+ε),\displaystyle+\frac{h}{\phi(r/d)}\sum_{\chi\neq\chi_{0}\bmod{r/d}}\overline{\chi}(v/d)L(0,\chi)L(1,\chi)A_{r,\chi}+O_{r}(h^{1/2+\varepsilon}),

with

(20) Ar,χ=p|r(1χ(p)p)pr(1(1χ(p))2(p1)2).A_{r,\chi}=\prod_{p|r}\left(1-\frac{\chi(p)}{p}\right)\prod_{p\nmid r}\left(1-\frac{(1-\chi(p))^{2}}{(p-1)^{2}}\right).
Lemma 3.2.

Fix integers r,h1r,h\geq 1 and two congruence classes c1,c2modrc_{1},c_{2}\bmod r. Define

Q=ph2prp,Q=\prod_{\begin{subarray}{c}p\leq h^{2}\\ p\nmid r\end{subarray}}p,

so that QQ is the product of all primes below h2h^{2} that do not divide rr. Let V2(Q,h;r,c1,c2)V_{2}(Q,h;r,c_{1},c_{2}) be defined as in (6). Then if c1c2modrc_{1}\equiv c_{2}\bmod r,

V2(Q,h;r,c1,c1)=hrd|Qd>1μ(d)2ϕ(d)hϕ(r)rlogh+C0(r)h+Or(h1/2+ε),V_{2}(Q,h;r,c_{1},c_{1})=\frac{h}{r}\sum_{\begin{subarray}{c}d|Q\\ d>1\end{subarray}}\frac{\mu(d)^{2}}{\phi(d)}-h\frac{\phi(r)}{r}\log h+C_{0}(r)h+O_{r}(h^{1/2+\varepsilon}),

where

(21) C0(r)=ϕ(r)r(logr2πγ0p|rlogpp1),C_{0}(r)=\frac{\phi(r)}{r}\left(\log\frac{r}{2\pi}-\gamma_{0}-\sum_{p|r}\frac{\log p}{p-1}\right),

and γ0\gamma_{0} is the Euler–Mascheroni constant. If c1c2modrc_{1}\not\equiv c_{2}\bmod r, define d=(c1c2,r)d=(c_{1}-c_{2},r), with d<rd<r. Then

V2(Q,h;r,c1,c2)=\displaystyle V_{2}(Q,h;r,c_{1},c_{2})=- hϕ(r)r2ϕ(r/d)Λ(r/d)\displaystyle h\frac{\phi(r)}{r^{2}\phi(r/d)}\Lambda(r/d)
(22) +2hrϕ(r/d)χχ0modr/dχ¯(c1c2d)L(0,χ)L(1,χ)Ar,χ+Or(h1/2+ε),\displaystyle+\frac{2h}{r\phi(r/d)}\sum_{\chi\neq\chi_{0}\bmod{r/d}}\overline{\chi}\left(\frac{c_{1}-c_{2}}{d}\right)L(0,\chi)L(1,\chi)A_{r,\chi}+O_{r}(h^{1/2+\varepsilon}),

where Ar,χA_{r,\chi} is defined in (20).

Proof.

Begin by expanding

V2(Q,h;r,c1,c2)\displaystyle V_{2}(Q,h;r,c_{1},c_{2}) =Er,c1(1)Er,c2(1)+q|Qμ(q)2ϕ(q)21aq(a,q)=1Er,c1(aq)Er,c2(aq)\displaystyle=-E_{r,c_{1}}(1)E_{r,c_{2}}(1)+\sum_{q|Q}\frac{\mu(q)^{2}}{\phi(q)^{2}}\sum_{\begin{subarray}{c}1\leq a\leq q\\ (a,q)=1\end{subarray}}E_{r,c_{1}}\left(\frac{a}{q}\right)E_{r,c_{2}}\left(-\frac{a}{q}\right)
=h2r2+q|Qμ(q)2ϕ(q)21aq(a,q)=1m1,m2hmicimodre((m1m2)aq).\displaystyle=-\frac{h^{2}}{r^{2}}+\sum_{q|Q}\frac{\mu(q)^{2}}{\phi(q)^{2}}\sum_{\begin{subarray}{c}1\leq a\leq q\\ (a,q)=1\end{subarray}}\sum_{\begin{subarray}{c}m_{1},m_{2}\leq h\\ m_{i}\equiv c_{i}\bmod r\end{subarray}}e\left((m_{1}-m_{2})\frac{a}{q}\right).

Assume first that c1c2modrc_{1}\equiv c_{2}\bmod r. Then

(23) V2(Q,h;r,c1,c1)=h2r2+q|Qμ(q)2ϕ(q)21aq(a,q)=1|m|hr|m1r(h|m|)e(maq).V_{2}(Q,h;r,c_{1},c_{1})=-\frac{h^{2}}{r^{2}}+\sum_{q|Q}\frac{\mu(q)^{2}}{\phi(q)^{2}}\sum_{\begin{subarray}{c}1\leq a\leq q\\ (a,q)=1\end{subarray}}\sum_{\begin{subarray}{c}|m|\leq h\\ r|m\end{subarray}}\frac{1}{r}(h-|m|)e\left(m\frac{a}{q}\right).

Let 𝐜q(m)\mathbf{c}_{q}(m) denote the Ramanujan sum 1aq(a,q)=1e(maq)\sum_{\begin{subarray}{c}1\leq a\leq q\\ (a,q)=1\end{subarray}}e\left(m\frac{a}{q}\right). The expression (23) is then

V2(Q,h;r,c1,c1)\displaystyle V_{2}(Q,h;r,c_{1},c_{1}) =h2r2+1rq|Qμ(q)2ϕ(q)2(h𝐜q(0)+21mh/r(hrm)𝐜q(rm))\displaystyle=-\frac{h^{2}}{r^{2}}+\frac{1}{r}\sum_{q|Q}\frac{\mu(q)^{2}}{\phi(q)^{2}}\left(h\mathbf{c}_{q}(0)+2\sum_{1\leq m\leq h/r}(h-rm)\mathbf{c}_{q}(rm)\right)
=h2r2+hrq|Qμ(q)2ϕ(q)2+2r1mh/r(hrm)q|Qμ(q)2ϕ(q)2𝐜q(rm).\displaystyle=-\frac{h^{2}}{r^{2}}+\frac{h}{r}\sum_{q|Q}\frac{\mu(q)^{2}}{\phi(q)^{2}}+\frac{2}{r}\sum_{1\leq m\leq h/r}(h-rm)\sum_{q|Q}\frac{\mu(q)^{2}}{\phi(q)^{2}}\mathbf{c}_{q}(rm).

The inside sum over q|Qq|Q is multiplicative and, since QQ is the product of primes ph2p\leq h^{2} not dividing rr, it is given by

q|Qμ(q)2ϕ(q)2𝐜q(rm)=p|mpr(1+1p1)pmpr(11(p1)2)+Or(h2)=𝔖r({0,m})+Or(h2),\sum_{q|Q}\frac{\mu(q)^{2}}{\phi(q)^{2}}\mathbf{c}_{q}(rm)=\prod_{\begin{subarray}{c}p|m\\ p\nmid r\end{subarray}}\left(1+\frac{1}{p-1}\right)\prod_{\begin{subarray}{c}p\nmid m\\ p\nmid r\end{subarray}}\left(1-\frac{1}{(p-1)^{2}}\right)+O_{r}(h^{-2})=\mathfrak{S}_{r}(\{0,m\})+O_{r}(h^{-2}),

by the definition of 𝔖r\mathfrak{S}_{r} in (5).

Then

V2(Q,h;r,c1,c1)=h2r2+hrq|Qμ(q)2ϕ(q)2+2r1mh/r(hrm)𝔖r({0,m})+Or(1),V_{2}(Q,h;r,c_{1},c_{1})=-\frac{h^{2}}{r^{2}}+\frac{h}{r}\sum_{q|Q}\frac{\mu(q)^{2}}{\phi(q)^{2}}+\frac{2}{r}\sum_{1\leq m\leq h/r}(h-rm)\mathfrak{S}_{r}(\{0,m\})+O_{r}(1),

which by Lemma 3.1 is equal to

h2r2+hrq|Qμ(q)2ϕ(q)2+h2r2hϕ(r)r(loghr+log2π+γ0+p|rlogpp1)+Or(h1/2+ε).-\frac{h^{2}}{r^{2}}+\frac{h}{r}\sum_{q|Q}\frac{\mu(q)^{2}}{\phi(q)^{2}}+\frac{h^{2}}{r^{2}}-h\frac{\phi(r)}{r}\left(\log\frac{h}{r}+\log 2\pi+\gamma_{0}+\sum_{p|r}\frac{\log p}{p-1}\right)+O_{r}(h^{1/2+\varepsilon}).

After a little rearranging this gives the desired result.

Now assume c1c2modrc_{1}\not\equiv c_{2}\bmod r. In this case,

V2(Q,h;r,c1,c2)\displaystyle V_{2}(Q,h;r,c_{1},c_{2}) =h2r2+q|Qμ(q)2ϕ(q)21aq(a,q)=1|m|hmc1c2modr1r(h|m|)e(maq)\displaystyle=-\frac{h^{2}}{r^{2}}+\sum_{q|Q}\frac{\mu(q)^{2}}{\phi(q)^{2}}\sum_{\begin{subarray}{c}1\leq a\leq q\\ (a,q)=1\end{subarray}}\sum_{\begin{subarray}{c}|m|\leq h\\ m\equiv c_{1}-c_{2}\bmod r\end{subarray}}\frac{1}{r}(h-|m|)e\left(m\frac{a}{q}\right)
=h2r2+|m|hmc1c2modr1r(h|m|)q|Qμ(q)2ϕ(q)2𝐜q(m)\displaystyle=-\frac{h^{2}}{r^{2}}+\sum_{\begin{subarray}{c}|m|\leq h\\ m\equiv c_{1}-c_{2}\bmod r\end{subarray}}\frac{1}{r}(h-|m|)\sum_{q|Q}\frac{\mu(q)^{2}}{\phi(q)^{2}}\mathbf{c}_{q}(m)
=h2r2+|m|hmc1c2modr1r(h|m|)𝔖r({0,m})+Or(1),\displaystyle=-\frac{h^{2}}{r^{2}}+\sum_{\begin{subarray}{c}|m|\leq h\\ m\equiv c_{1}-c_{2}\bmod r\end{subarray}}\frac{1}{r}(h-|m|)\mathfrak{S}_{r}(\{0,m\})+O_{r}(1),

where if mc1c2modrm\equiv c_{1}-c_{2}\bmod r, then m0m\neq 0.

Applying Lemma 3.1 completes the proof. ∎

4. Proof of Theorem 1.2

Throughout this section, fix r,k,h1r,k,h\geq 1 and set Q=pyprpQ=\prod_{\begin{subarray}{c}p\leq y\\ p\nmid r\end{subarray}}p, where y=hk+1y=h^{k+1}.

Begin with the expansion

(24) Rk(h;r,c1,,ck)=q1,,qk1<qi|Qi=1kμ(qi)ϕ(qi)a1,,ak1aiqi(ai,qi)=1iai/qid1,,dk1dihdistinctdicimodre(i=1kaidiqi)+Or(1),R_{k}(h;r,c_{1},\dots,c_{k})=\sum_{\begin{subarray}{c}q_{1},\dots,q_{k}\\ 1<q_{i}|Q\end{subarray}}\prod_{i=1}^{k}\frac{\mu(q_{i})}{\phi(q_{i})}\sum_{\begin{subarray}{c}a_{1},\dots,a_{k}\\ 1\leq a_{i}\leq q_{i}\\ (a_{i},q_{i})=1\\ \sum_{i}a_{i}/q_{i}\in{\mathbb{Z}}\end{subarray}}\sum_{\begin{subarray}{c}d_{1},\dots,d_{k}\\ 1\leq d_{i}\leq h\\ \text{distinct}\\ d_{i}\equiv c_{i}\bmod r\end{subarray}}e\left(\sum_{i=1}^{k}\frac{a_{i}d_{i}}{q_{i}}\right)+O_{r}(1),

where the error term is due to our choice of QQ. The expression on the right-hand side of (24) is very close to Vk(Q,h;r,c1,,ck)V_{k}(Q,h;r,c_{1},\dots,c_{k}), but in order to apply Theorem 1.1, we will need to remove the distinctness condition on the did_{i}’s. As in the proof of Theorem 2 from [7], removing this condition will be the bulk of our work.

This distinctness condition is heavily dependent on the congruence classes c1,,ckc_{1},\dots,c_{k}; in particular, if cicjmodrc_{i}\not\equiv c_{j}\bmod r, then did_{i} and djd_{j} never coincide and the distinctness condition is immaterial. Our arguments follow those of [7] closely, but with additional bookkeeping in order to account for the congruence classes c1,,ckmodrc_{1},\dots,c_{k}\bmod r.

For a given tuple (d1,,dk)(d_{1},\dots,d_{k}) with 1dih1\leq d_{i}\leq h and dicimodrd_{i}\equiv c_{i}\bmod r for all ii, put δij=1\delta_{ij}=1 if di=djd_{i}=d_{j} and δij=0\delta_{ij}=0 otherwise. Then

=1ri,j𝒞i<j(1δij)={1 if the di are distinct 0 otherwise.\prod_{\ell=1}^{r}\prod_{\begin{subarray}{c}i,j\in\mathcal{C}_{\ell}\\ i<j\end{subarray}}(1-\delta_{ij})=\begin{cases}1&\text{ if the }d_{i}\text{ are distinct }\\ 0&\text{ otherwise.}\end{cases}

When the left-hand side above is expanded, it is a linear combination of products of the δ\delta symbols. Let Δ\Delta denote one such product, and let |Δ||\Delta| denote the number of δij\delta_{ij} in the product. As in [7], define an equivalence relation on these δ\delta-products by setting Δ1Δ2\Delta_{1}\sim\Delta_{2} if Δ1\Delta_{1} and Δ2\Delta_{2} have the same value for all choices of did_{i}’s; for example, δ12δ23δ12δ13δ12δ23δ13.\delta_{12}\delta_{23}\sim\delta_{12}\delta_{13}\sim\delta_{12}\delta_{23}\delta_{13}.

Recall that a partition 𝒫={𝒮1,,𝒮M}\mathcal{P}=\{\mathcal{S}_{1},\dots,\mathcal{S}_{M}\} of [1,k][1,k] refines {𝒞}[1,k]\{\mathcal{C}_{\ell}\}_{\ell\in[1,k]} if for each 𝒮m𝒫\mathcal{S}_{m}\in\mathcal{P}, there exists some \ell with 𝒮m𝒞\mathcal{S}_{m}\subset\mathcal{C}_{\ell}; note that \ell is then unique. For such a partition, write 𝒫{𝒞}[1,k]\mathcal{P}\preceq\{\mathcal{C}_{\ell}\}_{\ell\in[1,k]} and define c(𝒮m)c(\mathcal{S}_{m}) to be the value \ell with 𝒮m𝒞\mathcal{S}_{m}\subset\mathcal{C}_{\ell}. Given a partition 𝒫\mathcal{P} refining {𝒞}[1,k]\{\mathcal{C}_{\ell}\}_{\ell\in[1,k]}, let

Δ𝒫=m=1Mi<ji,j𝒮mδij.\Delta_{\mathcal{P}}=\prod_{m=1}^{M}\prod_{\begin{subarray}{c}i<j\\ i,j\in\mathcal{S}_{m}\end{subarray}}\delta_{ij}.

Every equivalence class of δ\delta-products contains a unique Δ𝒫\Delta_{\mathcal{P}}, where the condition that 𝒫\mathcal{P} refines {𝒞}[1,k]\{\mathcal{C}_{\ell}\}_{\ell\in[1,k]} corresponds precisely to the fact that we are only considering δij\delta_{ij} when cicjmodrc_{i}\equiv c_{j}\bmod r. Equivalence classes of δ\delta-products are thus in bijection with partitions of [1,k][1,k] that refine {𝒞}[1,k]\{\mathcal{C}_{\ell}\}_{\ell\in[1,k]}. For a partition 𝒫\mathcal{P}, put

w(𝒫)=ΔΔ𝒫(1)|Δ|,w(\mathcal{P})=\sum_{\Delta\sim\Delta_{\mathcal{P}}}(-1)^{|\Delta|},

so that

=1ri,j𝒞i<j(1δij)=𝒫{𝒞}[1,k]w(𝒫)Δ𝒫,\prod_{\ell=1}^{r}\prod_{\begin{subarray}{c}i,j\in\mathcal{C}_{\ell}\\ i<j\end{subarray}}(1-\delta_{ij})=\prod_{\mathcal{P}\preceq\{\mathcal{C}_{\ell}\}_{\ell\in[1,k]}}w(\mathcal{P})\Delta_{\mathcal{P}},

and the sum over aia_{i}’s in (24) is equal to

𝒫{𝒞}[1,k]𝒫={𝒮1,,𝒮M}w(𝒫)a1,,ak1aiqi(ai,qi)=1iai/qim=1MEr,c(𝒮m)(i𝒮maiqi).\sum_{\begin{subarray}{c}\mathcal{P}\preceq\{\mathcal{C}_{\ell}\}_{\ell\in[1,k]}\\ \mathcal{P}=\{\mathcal{S}_{1},\dots,\mathcal{S}_{M}\}\end{subarray}}w(\mathcal{P})\sum_{\begin{subarray}{c}a_{1},\dots,a_{k}\\ 1\leq a_{i}\leq q_{i}\\ (a_{i},q_{i})=1\\ \sum_{i}a_{i}/q_{i}\in{\mathbb{Z}}\end{subarray}}\prod_{m=1}^{M}E_{r,c(\mathcal{S}_{m})}\left(\sum_{i\in\mathcal{S}_{m}}\frac{a_{i}}{q_{i}}\right).

By the same reasoning as in [7], the contribution to Rk(h;r,c1,,ck)R_{k}(h;r,c_{1},\dots,c_{k}) from terms where |𝒮m|3|\mathcal{S}_{m}|\geq 3 for some mm is Or(h(k1)/2+ε)O_{r}(h^{(k-1)/2+\varepsilon}), so that

Rk(h;r,c1,,ck)=𝒫{𝒞}[1,k]𝒫={𝒮1,,𝒮M}|𝒮m|2m\displaystyle R_{k}(h;r,c_{1},\dots,c_{k})=\sum_{\begin{subarray}{c}\mathcal{P}\preceq\{\mathcal{C}_{\ell}\}_{\ell\in[1,k]}\\ \mathcal{P}=\{\mathcal{S}_{1},\dots,\mathcal{S}_{M}\}\\ |\mathcal{S}_{m}|\leq 2\>\forall m\end{subarray}} w(𝒫)q1,,qk1<qi|Qi=1kμ(qi)ϕ(qi)\displaystyle w(\mathcal{P})\sum_{\begin{subarray}{c}q_{1},\dots,q_{k}\\ 1<q_{i}|Q\end{subarray}}\prod_{i=1}^{k}\frac{\mu(q_{i})}{\phi(q_{i})}
(25) a1,,ak1aiqi(ai,qi)=1iai/qim=1MEr,c(𝒮m)(i𝒮maiqi)+Or,k(h(k1)/2+ε).\displaystyle\cdot\sum_{\begin{subarray}{c}a_{1},\dots,a_{k}\\ 1\leq a_{i}\leq q_{i}\\ (a_{i},q_{i})=1\\ \sum_{i}a_{i}/q_{i}\in{\mathbb{Z}}\end{subarray}}\prod_{m=1}^{M}E_{r,c(\mathcal{S}_{m})}\left(\sum_{i\in\mathcal{S}_{m}}\frac{a_{i}}{q_{i}}\right)+O_{r,k}(h^{(k-1)/2+\varepsilon}).

Suppose that 𝒫\mathcal{P} consists of jj doubleton sets 𝒮1,,𝒮j\mathcal{S}_{1},\dots,\mathcal{S}_{j} and k2jk-2j singleton sets 𝒮j+1,,𝒮kj\mathcal{S}_{j+1},\dots,\mathcal{S}_{k-j}. Note that the number of these partitions depends on the partition {𝒞}\{\mathcal{C}_{\ell}\}, because of the constraint that 𝒫{𝒞}[1,k]\mathcal{P}\preceq\{\mathcal{C}_{\ell}\}_{\ell\in[1,k]}. The term in Rk(h;r,c1,,ck)R_{k}(h;r,c_{1},\dots,c_{k}) corresponding to a fixed such partition 𝒫\mathcal{P} is

(26) (1)jq1,,qk1<qi|Qi=1kμ(qi)ϕ(qi)a1,,ak1aiqi(ai,qi)=1iai/qim=1jEr,c(𝒮m)(i𝒮maiqi)m=j+1kjEr,c(𝒮m)(a𝒮mq𝒮m),(-1)^{j}\sum_{\begin{subarray}{c}q_{1},\dots,q_{k}\\ 1<q_{i}|Q\end{subarray}}\prod_{i=1}^{k}\frac{\mu(q_{i})}{\phi(q_{i})}\sum_{\begin{subarray}{c}a_{1},\dots,a_{k}\\ 1\leq a_{i}\leq q_{i}\\ (a_{i},q_{i})=1\\ \sum_{i}a_{i}/q_{i}\in{\mathbb{Z}}\end{subarray}}\prod_{m=1}^{j}E_{r,c(\mathcal{S}_{m})}\left(\sum_{i\in\mathcal{S}_{m}}\frac{a_{i}}{q_{i}}\right)\prod_{m=j+1}^{k-j}E_{r,c(\mathcal{S}_{m})}\left(\frac{a_{\mathcal{S}_{m}}}{q_{\mathcal{S}_{m}}}\right),

where we are slightly abusing notation in the final product by identifying the singletons 𝒮m\mathcal{S}_{m} with their unique element.

For 1mj1\leq m\leq j, define bmb_{m} and sms_{m} by the relations

bmsm=i𝒮maiqimod1,1bmsm,(bm,sm)=1,\frac{b_{m}}{s_{m}}=\sum_{i\in\mathcal{S}_{m}}\frac{a_{i}}{q_{i}}\bmod 1,\qquad 1\leq b_{m}\leq s_{m},\qquad(b_{m},s_{m})=1,

and define

Hm(bs)=Er,c(𝒮m)(bs)d1,d2|Q1<diμ(d1)μ(d2)ϕ(d1)ϕ(d2)e1,e21eidi(ei,di)=1e1d1+e2d2=bsmod11.H_{m}\left(\frac{b}{s}\right)=E_{r,c(\mathcal{S}_{m})}\left(\frac{b}{s}\right)\sum_{\begin{subarray}{c}d_{1},d_{2}|Q\\ 1<d_{i}\end{subarray}}\frac{\mu(d_{1})\mu(d_{2})}{\phi(d_{1})\phi(d_{2})}\sum_{\begin{subarray}{c}e_{1},e_{2}\\ 1\leq e_{i}\leq d_{i}\\ (e_{i},d_{i})=1\\ \tfrac{e_{1}}{d_{1}}+\tfrac{e_{2}}{d_{2}}=\tfrac{b}{s}\bmod 1\end{subarray}}1.

Then (26) is equal to

(27) s1,,sjsi|Qb1,,bj1bisi(bi,si)=1i=1jHi(bisi)qj+1,,qkj1<qi|Qaj+1,,akj1aiqi(ai,qi)=1iai/qi+ibi/sii=j+1kjμ(qi)ϕ(qi)Er,c(𝒮i)(aiqi).\sum_{\begin{subarray}{c}s_{1},\dots,s_{j}\\ s_{i}|Q\end{subarray}}\sum_{\begin{subarray}{c}b_{1},\dots,b_{j}\\ 1\leq b_{i}\leq s_{i}\\ (b_{i},s_{i})=1\end{subarray}}\prod_{i=1}^{j}H_{i}\left(\frac{b_{i}}{s_{i}}\right)\sum_{\begin{subarray}{c}q_{j+1},\dots,q_{k-j}\\ 1<q_{i}|Q\end{subarray}}\sum_{\begin{subarray}{c}a_{j+1},\dots,a_{k-j}\\ 1\leq a_{i}\leq q_{i}\\ (a_{i},q_{i})=1\\ \sum_{i}a_{i}/q_{i}+\sum_{i}b_{i}/s_{i}\in{\mathbb{Z}}\end{subarray}}\prod_{i=j+1}^{k-j}\frac{\mu(q_{i})}{\phi(q_{i})}E_{r,c(\mathcal{S}_{i})}\left(\frac{a_{i}}{q_{i}}\right).

Now separate the indices ii with si=1s_{i}=1. To do so, let ={i:si>1}\mathcal{L}=\{i:s_{i}>1\}. We can again rewrite (27) as

(28) [1,j]M()iHi(1),\sum_{\mathcal{L}\subset[1,j]}M(\mathcal{L})\prod_{i\not\in\mathcal{L}}H_{i}(1),

where

M()=(si)i1<si|Q(bi)i1bisi(bi,si)=1iHi(bisi)qj+1,,qkj1<qi|Qaj+1,,akj1aiqi(ai,qi)=1iai/qi+ibi/sii=j+1kjμ(qi)ϕ(qi)Er,c(𝒮i)(aiqi).M(\mathcal{L})=\sum_{\begin{subarray}{c}(s_{i})_{i\in\mathcal{L}}\\ 1<s_{i}|Q\end{subarray}}\sum_{\begin{subarray}{c}(b_{i})_{i\in\mathcal{L}}\\ 1\leq b_{i}\leq s_{i}\\ (b_{i},s_{i})=1\end{subarray}}\prod_{i\in\mathcal{L}}H_{i}\left(\frac{b_{i}}{s_{i}}\right)\sum_{\begin{subarray}{c}q_{j+1},\dots,q_{k-j}\\ 1<q_{i}|Q\end{subarray}}\sum_{\begin{subarray}{c}a_{j+1},\dots,a_{k-j}\\ 1\leq a_{i}\leq q_{i}\\ (a_{i},q_{i})=1\\ \sum_{i}a_{i}/q_{i}+\sum_{i}b_{i}/s_{i}\in{\mathbb{Z}}\end{subarray}}\prod_{i=j+1}^{k-j}\frac{\mu(q_{i})}{\phi(q_{i})}E_{r,c(\mathcal{S}_{i})}\left(\frac{a_{i}}{q_{i}}\right).

Note that M()=V2kj(Q,h;r,c(𝒮j+1),,c(𝒮kj))M(\varnothing)=V_{2k-j}(Q,h;r,c(\mathcal{S}_{j+1}),\dots,c(\mathcal{S}_{k-j})).

By precisely the same arguments as in [7], the contributions when ||1|\mathcal{L}|\geq 1 can be absorbed into the error term. Moreover,

Hi(1)\displaystyle H_{i}(1) =Er,c(𝒮i)(1)d|Qd>1μ(d)2ϕ(d)\displaystyle=E_{r,c(\mathcal{S}_{i})}(1)\sum_{\begin{subarray}{c}d|Q\\ d>1\end{subarray}}\frac{\mu(d)^{2}}{\phi(d)}
=(hr+O(1))d|Qd>1μ(d)2ϕ(d).\displaystyle=\left(\frac{h}{r}+O(1)\right)\sum_{\begin{subarray}{c}d|Q\\ d>1\end{subarray}}\frac{\mu(d)^{2}}{\phi(d)}.

Thus the expression (28) is equal to

(hrd|Qd>1μ(d)2ϕ(d))jVk2j(Q,h;r,c(𝒮j+1),,c(𝒮kj))+Or,k(h(k1)/2+ε).\left(\frac{h}{r}\sum_{\begin{subarray}{c}d|Q\\ d>1\end{subarray}}\frac{\mu(d)^{2}}{\phi(d)}\right)^{j}V_{k-2j}(Q,h;r,c(\mathcal{S}_{j+1}),\dots,c(\mathcal{S}_{k-j}))+O_{r,k}(h^{(k-1)/2+\varepsilon}).

Inserting this back into (25) yields

Rk(Q,h;r,c1,,ck)=0jk/2(1)j\displaystyle R_{k}(Q,h;r,c_{1},\dots,c_{k})=\sum_{0\leq j\leq k/2}(-1)^{j}
𝒫{𝒞}[1,k]𝒫={𝒮1,,𝒮M}|𝒮m|2m|𝒫|=kj(hrd|Qd>1μ(d)2ϕ(d))jVk2j(Q,h;r,c(𝒮j+1),,c(𝒮kj))+Or,k(h(k1)/2+ε).\displaystyle\sum_{\begin{subarray}{c}\mathcal{P}\preceq\{\mathcal{C}_{\ell}\}_{\ell\in[1,k]}\\ \mathcal{P}=\{\mathcal{S}_{1},\dots,\mathcal{S}_{M}\}\\ |\mathcal{S}_{m}|\leq 2\>\forall m\\ |\mathcal{P}|=k-j\end{subarray}}\left(\frac{h}{r}\sum_{\begin{subarray}{c}d|Q\\ d>1\end{subarray}}\frac{\mu(d)^{2}}{\phi(d)}\right)^{j}V_{k-2j}(Q,h;r,c(\mathcal{S}_{j+1}),\dots,c(\mathcal{S}_{k-j}))+O_{r,k}(h^{(k-1)/2+\varepsilon}).

We are finally prepared to appeal to Theorem 1.1. If kk is odd, then so is k2jk-2j, so there is no main term. Suppose that kk is even. Recall that (j+1,,kj)\mathcal{B}(j+1,\dots,k-j) denotes the set of perfect matchings of the set {j+1,,kj}\{j+1,\dots,k-j\}. Then the main term is

(29) 0jk/2(1)j𝒫{𝒞}[1,k]𝒫={𝒮1,,𝒮M}|𝒮m|2m|𝒫|=kj(hrd|Qd>1μ(d)2ϕ(d))jσ(j+1,,kj)(i1,i2)σV2(Q,h;r,c(𝒮i1),c(𝒮i2)),\displaystyle\sum_{0\leq j\leq k/2}(-1)^{j}\sum_{\begin{subarray}{c}\mathcal{P}\preceq\{\mathcal{C}_{\ell}\}_{\ell\in[1,k]}\\ \mathcal{P}=\{\mathcal{S}_{1},\dots,\mathcal{S}_{M}\}\\ |\mathcal{S}_{m}|\leq 2\>\forall m\\ |\mathcal{P}|=k-j\end{subarray}}\left(\frac{h}{r}\sum_{\begin{subarray}{c}d|Q\\ d>1\end{subarray}}\frac{\mu(d)^{2}}{\phi(d)}\right)^{j}\sum_{\sigma\in\mathcal{B}(j+1,\dots,k-j)}\prod_{(i_{1},i_{2})\in\sigma}V_{2}(Q,h;r,c(\mathcal{S}_{i_{1}}),c(\mathcal{S}_{i_{2}})),

which proves the first claim in Theorem 1.2.

By Lemma 3.2, V2(Q,h;r,c(𝒮i1),c(𝒮i2))=Or,k(h)V_{2}(Q,h;r,c(\mathcal{S}_{i_{1}}),c(\mathcal{S}_{i_{2}}))=O_{r,k}(h) unless c(𝒮i1)=c(𝒮i2)c(\mathcal{S}_{i_{1}})=c(\mathcal{S}_{i_{2}}). So, the largest term comes from those σ\sigma with c(𝒮i1)=c(𝒮i2)c(\mathcal{S}_{i_{1}})=c(\mathcal{S}_{i_{2}}) for all (i1,i2)σ(i_{1},i_{2})\in\sigma. Note that the error term is then quite large; it is only smaller by a factor of (logh)1(\log h)^{-1}.

If there exists some σ\sigma with c(𝒮i1)=c(𝒮i2)c(\mathcal{S}_{i_{1}})=c(\mathcal{S}_{i_{2}}) for all (i1,i2)σ(i_{1},i_{2})\in\sigma, then it must be that |𝒞||\mathcal{C}_{\ell}| is even for all \ell. Moreover, each term in this sum corresponds to a perfect pairing of [1,k][1,k] such that for each pair (i1,i2)(i_{1},i_{2}), ci1=ci2c_{i_{1}}=c_{i_{2}}; either two indices are paired by lying in the same 𝒮m\mathcal{S}_{m}, or by lying in the same element of σ\sigma. The choice of 𝒫\mathcal{P} then corresponds to choosing jj of these pairs. Note also that V2(Q,h;r,c,c)=V2(Q,h;r,c,c)+Or(h1/2+ε)V_{2}(Q,h;r,c,c)=V_{2}(Q,h;r,c^{\prime},c^{\prime})+O_{r}(h^{1/2+\varepsilon}) for any c,cmodrc,c^{\prime}\bmod r, which allows us to simplify the main term in this case to get

#~(c1,,ck)0jk/2(1)j(k/2j)(hrd|Qd>1μ(d)2ϕ(d))jV2(Q,h;r,0,0)k/2j,\#\widetilde{\mathcal{B}}(c_{1},\dots,c_{k})\sum_{0\leq j\leq k/2}(-1)^{j}\binom{k/2}{j}\left(\frac{h}{r}\sum_{\begin{subarray}{c}d|Q\\ d>1\end{subarray}}\frac{\mu(d)^{2}}{\phi(d)}\right)^{j}V_{2}(Q,h;r,0,0)^{k/2-j},

where #~(c1,,ck)\#\widetilde{\mathcal{B}}(c_{1},\dots,c_{k}) is the number of ways to pair the cic_{i}’s such that every pair has equal values. By the binomial theorem, this is

#~(c1,,ck)(V2(Q,h;r,0,0)hrd|Qd>1μ(d)2ϕ(d))k/2.\#\widetilde{\mathcal{B}}(c_{1},\dots,c_{k})\left(V_{2}(Q,h;r,0,0)-\frac{h}{r}\sum_{\begin{subarray}{c}d|Q\\ d>1\end{subarray}}\frac{\mu(d)^{2}}{\phi(d)}\right)^{k/2}.

By Lemma 3.2, this is

#~(c1,,ck)(hϕ(r)rlogh+C0(r)h)k/2+Or,k(h(k1)/2+ε),\#\widetilde{\mathcal{B}}(c_{1},\dots,c_{k})\left(-h\frac{\phi(r)}{r}\log h+C_{0}(r)h\right)^{k/2}+O_{r,k}(h^{(k-1)/2+\varepsilon}),

for C0(r)C_{0}(r) defined in (21), which gives the result.

5. Weighting by smooth functions

We now consider sums of singular series weighted by smooth functions and the proofs of Theorems 1.3 and 1.5. Theorem 1.3 follows arguments identical to those in the proof of Theorem 1.1 as well as Theorem 1 of [7]. In particular, all estimates used in bounding E(α)E(\alpha) in the proof of Theorem 1 of [7] hold for the sums Efi,h(α)E_{f_{i},h}(\alpha) that we consider in the smooth setting, and the remainder of the proof is identical.

Accordingly, for the proof of Theorem 1.3 we restrict our attention to relevant estimates of the sums Efi,h(α)E_{f_{i},h}(\alpha), which is the only place where the proof differs. These estimates, the equivalents of Lemmas 4 and 6 from [8], are contained in Section 5.1.

Similarly for Theorem 1.5, the proof is identical to that of Theorem 1.2 presented in Section 4, so we omit it. In Section 5.2, we prove Lemma 1.4, whose proof follows similar lines as the proofs in Section 3.

5.1. Exponential sums weighted by smooth functions

Lemma 5.1.

Let m,h1m,h\geq 1, let f1,f2:f_{1},f_{2}:{\mathbb{R}}\to{\mathbb{R}} be a smooth functions with compact support such that |fi^(ξ)|=O(|ξ|2)|\hat{f_{i}}(\xi)|=O(|\xi|^{-2}), and define Efi,hE_{f_{i},h} by (12). Then

(30) μmodmμ0Ef1,h(μm)Ef2,h(μm)f1,f2mh2min{m3,h3}.\sum_{\begin{subarray}{c}\mu\bmod m\\ \mu\neq 0\end{subarray}}E_{f_{1},h}\left(\frac{\mu}{m}\right)E_{f_{2},h}\left(\frac{\mu}{m}\right)\ll_{f_{1},f_{2}}mh^{-2}\min\{m^{3},h^{3}\}.

Moreover, for any α\alpha\in{\mathbb{R}},

(31) μmodmEf1,h(μm+α)Ef2,h(μm+α)=h2f^(hm(mα¯))2+Of(mh2min{m3,h3}).\sum_{\begin{subarray}{c}\mu\bmod m\end{subarray}}E_{f_{1},h}\left(\frac{\mu}{m}+\alpha\right)E_{f_{2},h}\left(\frac{\mu}{m}+\alpha\right)=h^{2}\hat{f}\left(-\frac{h}{m}(\overline{m\alpha})\right)^{2}+O_{f}(mh^{-2}\min\{m^{3},h^{3}\}).
Proof.

Begin with the second statement. Expand via (13) to get

μmodm\displaystyle\sum_{\begin{subarray}{c}\mu\bmod m\end{subarray}} Ef1,h(μm+α)Ef2,h(μm+α)\displaystyle E_{f_{1},h}\left(\frac{\mu}{m}+\alpha\right)E_{f_{2},h}\left(\frac{\mu}{m}+\alpha\right)
=h2μmodmf1^(h(μ/m+α¯))f2^(h(μ/m+α¯))\displaystyle=h^{2}\sum_{\begin{subarray}{c}\mu\bmod m\end{subarray}}\hat{f_{1}}(-h(\overline{\mu/m+\alpha}))\hat{f_{2}}(-h(\overline{\mu/m+\alpha}))
+O(mh2+μmodmf^1(h(μ/m+α¯))+f^2(h(μ/m+α¯)))\displaystyle\qquad+O\Big{(}mh^{-2}+\sum_{\begin{subarray}{c}\mu\bmod m\end{subarray}}\hat{f}_{1}(-h(\overline{\mu/m+\alpha}))+\hat{f}_{2}(-h(\overline{\mu/m+\alpha}))\Big{)}
=h2μmodmf1^(h(μ/m+α)¯)f2^(h(μ/m+α)¯)+Of1,f2(min{m,m2h2}).\displaystyle=h^{2}\sum_{\begin{subarray}{c}\mu\bmod m\end{subarray}}\hat{f_{1}}(-h\overline{(\mu/m+\alpha)})\hat{f_{2}}(-h\overline{(\mu/m+\alpha)})+O_{f_{1},f_{2}}\left(\min\{m,m^{2}h^{-2}\}\right).

Let μ0\mu_{0} be the value of μ\mu such that |μm+α¯|\left|\overline{\frac{\mu}{m}+\alpha}\right| is minimized; then μm+α¯=mα¯/m\overline{\frac{\mu}{m}+\alpha}=\overline{m\alpha}/m. If mhm\leq h, then

h2μmodmf^1\displaystyle h^{2}\sum_{\begin{subarray}{c}\mu\bmod m\end{subarray}}\hat{f}_{1} (h(μ/m+α¯))f^2(h(μ/m+α¯))\displaystyle(-h(\overline{\mu/m+\alpha}))\hat{f}_{2}(-h(\overline{\mu/m+\alpha}))
=h2f^1(hmmα¯)f^2(hmmα¯)+h2μmodmμμ0f^1(h(μ/m+α¯))f^2(h(μ/m+α¯))\displaystyle=h^{2}\hat{f}_{1}\left(-\frac{h}{m}\overline{m\alpha}\right)\hat{f}_{2}\left(-\frac{h}{m}\overline{m\alpha}\right)+h^{2}\sum_{\begin{subarray}{c}\mu\bmod m\\ \mu\neq\mu_{0}\end{subarray}}\hat{f}_{1}(-h(\overline{\mu/m+\alpha}))\hat{f}_{2}(-h(\overline{\mu/m+\alpha}))
f1,f2h2+O(m4h2),\displaystyle\ll_{f_{1},f_{2}}h^{2}+O(m^{4}h^{-2}),

using the fact that |f^i(ξ)|=O(|ξ|2)|\hat{f}_{i}(\xi)|=O(|\xi|^{-2}). On the other hand, if m>hm>h, then

h2μmodmf^1\displaystyle h^{2}\sum_{\begin{subarray}{c}\mu\bmod m\end{subarray}}\hat{f}_{1} (h(μ/m+α¯))f^2(h(μ/m+α¯))\displaystyle(-h(\overline{\mu/m+\alpha}))\hat{f}_{2}(-h(\overline{\mu/m+\alpha}))
=h2μmodm|μ/m+α¯|m/hf^1(h(μ/m+α¯))f^2(h(μ/m+α¯))\displaystyle=h^{2}\sum_{\begin{subarray}{c}\mu\bmod m\\ |\overline{\mu/m+\alpha}|\leq m/h\end{subarray}}\hat{f}_{1}(-h(\overline{\mu/m+\alpha}))\hat{f}_{2}(-h(\overline{\mu/m+\alpha}))
+h2μmodm|μ/m+α¯|>m/hf^1(h(μ/m+α¯))f^2(h(μ/m+α¯))\displaystyle\qquad+h^{2}\sum_{\begin{subarray}{c}\mu\bmod m\\ |\overline{\mu/m+\alpha}|>m/h\end{subarray}}\hat{f}_{1}(-h(\overline{\mu/m+\alpha}))\hat{f}_{2}(-h(\overline{\mu/m+\alpha}))
f1,f2h2mhf^(0)2+h2mh\displaystyle\ll_{f_{1},f_{2}}h^{2}\frac{m}{h}\hat{f}(0)^{2}+h^{2}\frac{m}{h}
f1,f2mh,\displaystyle\ll_{f_{1},f_{2}}mh,

as desired.

Equation (30) follows from the proof of (31) for α=0\alpha=0 upon excluding the term m=0m=0. ∎

Remark.

In the case of analogous exponential sums in the function field case of polynomials over 𝐅q[t]\mathbf{F}_{q}[t], as explored in [4], the sum analogous to μmodmE(μm+α)2\sum_{\mu\bmod m}E\left(\frac{\mu}{m}+\alpha\right)^{2} can be bounded by the analog of mhmh whenever α\alpha is large, i.e. whenever α\alpha is a rational function of sufficiently large degree (see Lemma 3.4 of [4]). The equivalent statement here would be that μmodmE(μm+α)2\sum_{\mu\bmod m}E\left(\frac{\mu}{m}+\alpha\right)^{2} is bounded by mhmh whenever α\alpha is far away from an integer. However, this is not true: if mm is even and α\alpha is very close to 12\frac{1}{2}, say, then this sum will still have a contribution of size h2h^{2}.

In the function field case of [4], the simplified and stronger bounds correspondingly yield a simplified proof of the analog of Theorem 1.3. The smooth weights fif_{i} make the exponential sums here cleaner, but because the function field-style bounds are not available for the sums in Lemma 5.1, the simplified proof of the analog of Theorem 1.3 in the function field case also fails to apply.

The following lemma corresponds to Lemma 6 of Montgomery and Vaughan’s work.

Lemma 5.2.

Let f1,f2:f_{1},f_{2}:{\mathbb{R}}\to{\mathbb{R}} be smooth functions with compact support such that |fi^(ξ)|=O(|ξ|2)|\hat{f_{i}}(\xi)|=O(|\xi|^{-2}), and define Efi,hE_{f_{i},h} by (12). Fix α1,α2\alpha_{1},\alpha_{2}\in{\mathbb{R}}. Then

(32) μmodmEf1,h(μm+α1)Ef2,h(μm+α2)(m+h)Ef1f2¯(α1α2)+O(m).\sum_{\begin{subarray}{c}\mu\bmod m\end{subarray}}E_{f_{1},h}\left(\frac{\mu}{m}+\alpha_{1}\right)E_{f_{2},h}\left(\frac{\mu}{m}+\alpha_{2}\right)\ll(m+h)E_{f_{1}\overline{f_{2}}}(\alpha_{1}-\alpha_{2})+O(m).
Proof.

By Lemma 3 from [8],

μmodm\displaystyle\sum_{\mu\bmod m} Ef1,h(μm+α1)Ef2,h(μm+α2)\displaystyle E_{f_{1},h}\left(\frac{\mu}{m}+\alpha_{1}\right)E_{f_{2},h}\left(\frac{\mu}{m}+\alpha_{2}\right)
=μmodmh2f1^(hμmhα1)f2^(hμmhα2)+O(m)\displaystyle=\sum_{\mu\bmod m}h^{2}\hat{f_{1}}\left(-h\frac{\mu}{m}-h\alpha_{1}\right)\hat{f_{2}}\left(-h\frac{\mu}{m}-h\alpha_{2}\right)+O(m)
(m+h)f1^(hthα1)f2^(hthα2)h2dt+O(m),\displaystyle\ll(m+h)\int_{-\infty}^{\infty}\hat{f_{1}}(-ht-h\alpha_{1})\hat{f_{2}}(-ht-h\alpha_{2})h^{2}\mathrm{d}t+O(m),

keeping in mind that fi(x)fi(y)f_{i}(x)\asymp f_{i}(y) whenever |xy|1/h|x-y|\leq 1/h. The integral is the convolution of f1^\hat{f_{1}} and f2^\hat{f_{2}}:

(m+h)f1^(hthα1)f2^(hthα2)h2dt\displaystyle(m+h)\int_{-\infty}^{\infty}\hat{f_{1}}(-ht-h\alpha_{1})\hat{f_{2}}(-ht-h\alpha_{2})h^{2}\mathrm{d}t =(m+h)hf1^(u)f2^(u+hα1hα2)du\displaystyle=(m+h)h\int_{-\infty}^{\infty}\hat{f_{1}}(u)\hat{f_{2}}(u+h\alpha_{1}-h\alpha_{2})\mathrm{d}u
=(m+h)hf1^f2¯^(hα2hα1)k\displaystyle=(m+h)h\hat{f_{1}}\ast\hat{\overline{f_{2}}}(h\alpha_{2}-h\alpha_{1})k
=(m+h)Ef1f2¯,h(α1α2)+O(m/h),\displaystyle=(m+h)E_{f_{1}\overline{f_{2}},h}(\alpha_{1}-\alpha_{2})+O(m/h),

which yields the desired bound. ∎

Remark.

The bound in Lemma 5.2 improves on the analogous lemma in [8] by a factor of logh\log h. However, the improvement in this lemma does not affect the error term in Theorem 1.3.

5.2. Proof of Lemma 1.4

Lemma 5.3.

Fix h1h\geq 1, and let f:0f:{\mathbb{R}}_{\geq 0}\to{\mathbb{C}} be a smooth function with compact support supp(f)(0,)\mathrm{supp}(f)\subset(0,\infty) and such that |f^(ξ)|O(|ξ|2)|\hat{f}(\xi)|\ll O(|\xi|^{-2}). Define

S(f,h):=m=1f(mh)𝔖({0,m}).S(f,h):=\sum_{m=1}^{\infty}f\left(\frac{m}{h}\right)\mathfrak{S}(\{0,m\}).

Then

S(f,h)={f}(2)h{f}(1)2logh+{f}(1)2(γ0log2π{f′′}{f}(1))+Of(h1/2+ε),S(f,h)=\{\mathcal{M}f\}(2)h-\frac{\{\mathcal{M}f^{\prime}\}(1)}{2}\log h+\frac{\{\mathcal{M}f^{\prime}\}(1)}{2}\left(\gamma_{0}-\log 2\pi-\frac{\{\mathcal{M}f^{\prime\prime}\}}{\{\mathcal{M}f^{\prime}\}(1)}\right)+O_{f}(h^{-1/2+\varepsilon}),

where γ0\gamma_{0} denotes the Euler–Mascheroni constant and for a function gg, {g}(s)\{\mathcal{M}g\}(s) is the Mellin transform of gg defined in (15).

Proof.

The proof proceeds almost entirely along the lines of Proposition 2.1 in [5], so we will be brief. Define for Re(s)>1\mathrm{Re}(s)>1

F(s):=n1𝔖({0,n})ns,F(s):=\sum_{n\geq 1}\frac{\mathfrak{S}(\{0,n\})}{n^{s}},

so that

(33) S(f,h)=12πi(2)F(s)hs{f}(s)ds.S(f,h)=\frac{1}{2\pi i}\int_{(2)}F(s)h^{s}\{\mathcal{M}f\}(s)\mathrm{d}s.

As noted in [5] and [7], F(s)F(s) admits a meromorphic continuation to Re(s)>1/2\mathrm{Re}(s)>-1/2 via

F(s)=ζ(s)ζ(s+1)p prime(111/ps)2(p1)2).F(s)=\zeta(s)\zeta(s+1)\prod_{p\text{ prime}}\left(1-\frac{1-1/p^{s})^{2}}{(p-1)^{2}}\right).

Since ff is smooth and has compact support, {f}(s)\{\mathcal{M}f\}(s) is analytic in Re(s)>0\mathrm{Re}(s)>0. It can be extended meromorphically to the complex plane via the identity

{f}=1s{f},\{\mathcal{M}f\}=-\frac{1}{s}\{\mathcal{M}f^{\prime}\},

which follows from integration by parts. Thus {f}(s)\{\mathcal{M}f\}(s) has simple poles at all nonpositive integers and no other poles.

The result follows from moving the line of integration in (33) to Re(s)=1/2+ε\mathrm{Re}(s)=-1/2+\varepsilon and recording the contributions from the simple pole at s=1s=1 and the double pole at s=0s=0. ∎

We are now ready to prove Lemma 1.4.

Proof.

Begin by expanding

V2\displaystyle V_{2} (Q,h;f1,f2)\displaystyle(Q,h;f_{1},f_{2})
=Ef1,h(1)Ef2,h(1)+q|Qμ(q)2ϕ(q)21aq(a,q)=1Ef1,h(aq)Ef2,h(aq)\displaystyle=-E_{f_{1},h}(1)E_{f_{2},h}(1)+\sum_{q|Q}\frac{\mu(q)^{2}}{\phi(q)^{2}}\sum_{\begin{subarray}{c}1\leq a\leq q\\ (a,q)=1\end{subarray}}E_{f_{1},h}\left(\frac{a}{q}\right)E_{f_{2},h}\left(-\frac{a}{q}\right)
=h2f1^(0)f2^(0)+Of1,f2(1)\displaystyle=-h^{2}\hat{f_{1}}(0)\hat{f_{2}}(0)+O_{f_{1},f_{2}}(1)
+q|Qμ(q)2ϕ(q)21aq(a,q)=1m1,m2=1f1(m1h)f2(m2h)e((m1m2)aq)\displaystyle\qquad+\sum_{q|Q}\frac{\mu(q)^{2}}{\phi(q)^{2}}\sum_{\begin{subarray}{c}1\leq a\leq q\\ (a,q)=1\end{subarray}}\sum_{m_{1},m_{2}=1}^{\infty}f_{1}\left(\frac{m_{1}}{h}\right)f_{2}\left(-\frac{m_{2}}{h}\right)e\left((m_{1}-m_{2})\frac{a}{q}\right)
=h2f1^(0)f2^(0)+Of1,f2(1)+m1,m2=1f1(m1h)f2(m2h)q|Qμ(q)2ϕ(q)2𝐜q(m1m2),\displaystyle=-h^{2}\hat{f_{1}}(0)\hat{f_{2}}(0)+O_{f_{1},f_{2}}(1)+\sum_{m_{1},m_{2}=1}^{\infty}f_{1}\left(\frac{m_{1}}{h}\right)f_{2}\left(-\frac{m_{2}}{h}\right)\sum_{q|Q}\frac{\mu(q)^{2}}{\phi(q)^{2}}\mathbf{c}_{q}(m_{1}-m-2),

where 𝐜q(m)\mathbf{c}_{q}(m) is a Ramanujan sum. Just as for the arithmetic progressions case, this simplifies to

=h2f1^(0)f2^(0)+m1,m2=1m1m2f1(m1h)f2(m2h)𝔖({0,m1m2})\displaystyle=-h^{2}\hat{f_{1}}(0)\hat{f_{2}}(0)+\sum_{\begin{subarray}{c}m_{1},m_{2}=1\\ m_{1}\neq m_{2}\end{subarray}}^{\infty}f_{1}\left(\frac{m_{1}}{h}\right)f_{2}\left(-\frac{m_{2}}{h}\right)\mathfrak{S}(\{0,m_{1}-m_{2}\})
+m=1f1(mh)f2(mh)+Of1,f2(1),\displaystyle\qquad+\sum_{m=1}^{\infty}f_{1}\left(\frac{m}{h}\right)f_{2}\left(-\frac{m}{h}\right)+O_{f_{1},f_{2}}(1),

since for any m0m\neq 0, by our choice of QQ,

q|Qμ(q)2ϕ(q)2𝐜q(m)=p|m(1+1p1)pm(11(p1)2)=𝔖({0,m})+O(h2).\sum_{q|Q}\frac{\mu(q)^{2}}{\phi(q)^{2}}\mathbf{c}_{q}(m)=\prod_{p|m}\left(1+\frac{1}{p-1}\right)\prod_{p\nmid m}\left(1-\frac{1}{(p-1)^{2}}\right)=\mathfrak{S}(\{0,m\})+O(h^{-2}).

The sum over mm can be interpreted as a Riemann sum as hh\to\infty, yielding

m=1f1(mh)f2(mh)\displaystyle\sum_{m=1}^{\infty}f_{1}\left(\frac{m}{h}\right)f_{2}\left(-\frac{m}{h}\right) =h0f1(x)f2(x)dx+Of1,f2(1)\displaystyle=h\int_{0}^{\infty}f_{1}(x)f_{2}(-x)\mathrm{d}x+O_{f_{1},f_{2}}(1)
=h(f1f2)(0)+Of1,f2(1).\displaystyle=h(f_{1}\ast f_{2})(0)+O_{f_{1},f_{2}}(1).

Similarly, the sum over m1m_{1} and m2m_{2} can also be interpreted as a Riemann integral (and also as the convolution of f1f_{1} and f2f_{2}), yielding

m1,m2=1m1m2\displaystyle\sum_{\begin{subarray}{c}m_{1},m_{2}=1\\ m_{1}\neq m_{2}\end{subarray}}^{\infty} f1(m1h)f2(m2h)𝔖({0,m1m2})\displaystyle f_{1}\left(\frac{m_{1}}{h}\right)f_{2}\left(-\frac{m_{2}}{h}\right)\mathfrak{S}(\{0,m_{1}-m_{2}\})
=m=1𝔖({0,m})n=1(f1(nh)f2(mnh)+f1(nh)f2(mnh))\displaystyle=\sum_{m=1}^{\infty}\mathfrak{S}(\{0,m\})\sum_{n=1}^{\infty}\left(f_{1}\left(\frac{n}{h}\right)f_{2}\left(\frac{m-n}{h}\right)+f_{1}\left(\frac{n}{h}\right)f_{2}\left(\frac{-m-n}{h}\right)\right)
=m=1𝔖({0,m})h((f1f2)(mh))+Of1,f2(m=1Ch𝔖({0,m})),\displaystyle=\sum_{m=1}^{\infty}\mathfrak{S}(\{0,m\})h\left((f_{1}\ast f_{2})\left(\frac{m}{h}\right)\right)+O_{f_{1},f_{2}}\left(\sum_{m=1}^{Ch}\mathfrak{S}(\{0,m\})\right),

where CC is a constant large enough that f1(x)=f2(x)=0f_{1}(x)=f_{2}(x)=0 for any |x|C/2|x|\geq C/2; note that CC depends only on f1f_{1} and f2f_{2}. By the results of [1] and [7], the error term is Of1,f2(h)O_{f_{1},f_{2}}(h). For the inside sum, apply Lemma 5.3 with f=f1f2f=f_{1}\ast f_{2} to get

m1,m2=1m1m2\displaystyle\sum_{\begin{subarray}{c}m_{1},m_{2}=1\\ m_{1}\neq m_{2}\end{subarray}}^{\infty} f1(m1h)f2(m2h)𝔖({0,m1m2})\displaystyle f_{1}\left(\frac{m_{1}}{h}\right)f_{2}\left(-\frac{m_{2}}{h}\right)\mathfrak{S}(\{0,m_{1}-m_{2}\})
={f}(2)h2{f}(1)2hlogh+Of1,f2(h),\displaystyle=\{\mathcal{M}f\}(2)h^{2}-\frac{\{\mathcal{M}f^{\prime}\}(1)}{2}h\log h+O_{f_{1},f_{2}}(h),

which, after collecting terms, implies the result. ∎

References

  • [1] P. X. Gallagher, On the distribution of primes in short intervals, Mathematika 23 (1976), no. 1, 4–9. MR 409385
  • [2] D. A. Goldston, Linnik’s theorem on Goldbach numbers in short intervals, Glasgow Math. J. 32 (1990), no. 3, 285–297. MR 1073669
  • [3] A. Granville and K. Soundararajan, Sieving and the Erdős-Kac theorem, Equidistribution in number theory, an introduction, NATO Sci. Ser. II Math. Phys. Chem., vol. 237, Springer, Dordrecht, 2007, pp. 15–27. MR 2290492
  • [4] V. Kuperberg, Odd moments in the distribution of primes, arXiv:2109.03767, 2021.
  • [5] R. J. Lemke Oliver and K. Soundararajan, Unexpected biases in the distribution of consecutive primes, Proc. Natl. Acad. Sci. USA 113 (2016), no. 31, E4446–E4454. MR 3624386
  • [6] H. L. Montgomery and K. Soundararajan, Beyond pair correlation, Paul Erdős and his mathematics, I (Budapest, 1999), Bolyai Soc. Math. Stud., vol. 11, János Bolyai Math. Soc., Budapest, 2002, pp. 507–514. MR 1954710
  • [7] by same author, Primes in short intervals, Comm. Math. Phys. 252 (2004), no. 1-3, 589–617. MR 2104891
  • [8] H. L. Montgomery and R. C. Vaughan, On the distribution of reduced residues, Ann. of Math. (2) 123 (1986), no. 2, 311–333. MR 835765