Keywords

2010 Mathematics subject classification:

1 Introduction

There are many results in the literature concerned with rational approximations \(p/q\) to irrational numbers, where \(p\) and \(q\) are restricted by additional arithmetical conditions. An important result in this direction is due to Uchiyama [11].

Theorem A

For every real irrational number \(\alpha\) and integers \(s> 1\) , \(a \geq 0\) , \(b \geq 0\) such that \(a\) and \(b\) are not simultaneously divisible by \(s\) , there are infinitely many integers \(p\) and \(q\not =0\) satisfying

$$\displaystyle{\Big\vert \alpha -\frac{p} {q}\Big\vert \, <\, \frac{s^{2}} {4q^{2}}\,,\quad p \equiv a\pmod s\,,\quad q \equiv b\pmod s\,.}$$

In [3] the author proved that the constant 1/4 in Uchiyama’s paper cannot be improved. Let \(\|\eta \|\) denote the distance of a real number \(\eta\) from the nearest integer. Then we deduce the following corollary from Theorem A:

Corollary A1

Let \(f: \mathbb{N} \rightarrow \mathbb{R}_{>0}\) be a function satisfying \(f(q) = o(q)\) for positive integers \(q\) tending to infinity. Then, for every integers \(s> 0\) , \(a \geq 0\) and every real irrational number \(\alpha\) we have

$$\displaystyle{\liminf _{\begin{array}{c} q> 0 \\ q \equiv a\pmod s \end{array} }f(q)\|q\alpha \|\, =\, 0\,.}$$

In particular cases stronger results are possible, e.g., for the number \(e =\exp (1)\) by Theorem 1.3 in [4].

Theorem B

Let \(a\) and \(s\) be arbitrary positive integers. Then

$$\displaystyle{\liminf _{\begin{array}{c} q> 0 \\ q \equiv a\pmod s\end{array} }q\|qe\|\, =\, 0\,.}$$

About 5 years later Komatsu [9, Theorem 4] showed that the result of Theorem B remains true for \(e\) replaced by every number \(e^{1/k}\) \((k \in \mathbb{N})\).

Recently, the author [5] studied the so-called error sum functions. Let

$$\displaystyle{\mathcal{E}(\alpha )\,:=\,\sum _{ m=0}^{\infty }\vert q_{ m}\alpha - p_{m}\vert \,,\qquad \mathcal{E}^{{\ast}}(\alpha )\,:=\,\sum _{ m=0}^{\infty }(q_{ m}\alpha - p_{m})\,,}$$

where for \(m \geq 0\) the fraction \(p_{m}/q_{m}\) is the \(m\)-th convergent of the real number \(\alpha\). The numbers \(p_{m}\) and \(q_{m}\) can be computed recursively from the continued fraction expansion of \(\alpha\). Various aspects of these functions have been investigated in [57], among them it is shown that \(0 \leq \mathcal{E}(\alpha ) \leq (1 + \sqrt{5})/2\) and \(0 \leq \mathcal{E}^{{\ast}}(\alpha ) \leq 1\) for all real numbers \(\alpha\). Both, \(\mathcal{E}(\alpha )\) and \(\mathcal{E}^{{\ast}}(\alpha )\), measure the average of error terms for diophantine approximations of \(\alpha\) by rationals. Moreover, \(\mathcal{E}(\alpha ) \in \mathbb{Q}(\alpha )\) holds for real numbers of algebraic degree 1 and 2. For \(e =\exp (1)\) we have the formula

$$\displaystyle{\mathcal{E}(e)\, =\, 2e\int _{0}^{1}\exp (-t^{2})\,dt - e\, =\, 1.3418751\ldots \,,}$$

which proves that \(\mathcal{E}(e)\not\in \mathbb{Q}(e)\). The function \(\mathcal{E}(\alpha )\) is continuous for every real irrational point \(\alpha\), and discontinuous for all rational numbers \(\alpha\) (see [6, Theorem 2]). Therefore, the function \(\mathcal{E}\) is Riemann-integrable on \([0,1]\). It turns out [6, Theorem 5] that

$$\displaystyle{\int _{0}^{1}\mathcal{E}(\alpha )\,d\alpha \, =\, -\frac{5} {8} + \frac{3\zeta (2)\log 2} {2\zeta (3)} \, =\, 0.79778798\ldots \,,}$$

where \(\zeta\) denotes the Riemann Zeta function. This integral represents the arithmetical mean of the function \(\mathcal{E}\) on \([0,1]\). This result can be generalized. Let n = 1, 2, 3,  and

$$\displaystyle{I_{n}\,:=\,\int _{ 0}^{1}\sum _{ m=0}^{\infty }\big\vert q_{ m}\alpha - p_{m}\big\vert ^{n}\,d\alpha \,.}$$

It can be shown [6, Sec. 4] that

$$\displaystyle{I_{n}\, =\, \frac{1} {n + 1}\left (1 - \frac{1} {2^{n+1}} -\frac{2\zeta (n + 1,-1)} {\zeta (n + 2)} \right )\,,}$$

where

$$\displaystyle{\zeta (n + 1,-1)\,:=\,\sum _{m_{2}>m_{1}>0} \frac{(-1)^{m_{1}}} {m_{1}m_{2}^{n+1}}\, =\,\sum _{ m_{2}=1}^{\infty } \frac{1} {m_{2}^{n+1}}\sum _{m_{1}=1}^{m_{2}-1}\frac{(-1)^{m_{1}}} {m_{1}} }$$

is known as multiple Zeta function. Borwein et al. [2] expressed ζ(n + 1, −1) in terms of log2, ζ(2), ζ(3), , ζ(n + 2). Thus we obtain the following results:

Theorem C

Let n ≥ 1 be an integer. Then we have

$$\displaystyle\begin{array}{rcl} & & \int _{0}^{1}\sum _{ m=0}^{\infty }\big\vert q_{ m}\alpha - p_{m}\big\vert ^{n}\,d\alpha = \frac{1} {n + 1}\left (1 - \frac{1} {2^{n+1}}\right ) - 1 + \frac{4\log (2)\,\zeta (n + 1)} {(n + 1)\zeta (n + 2)}\left (1 - \frac{1} {2^{n+1}}\right ) {}\\ & & \qquad +\, \frac{1} {(n + 1)\zeta (n + 2)}\sum _{k=1}^{n-1}\left (1 - \frac{1} {2^{k}}\right )\left (1 - \frac{1} {2^{n-k}}\right )\zeta (k + 1)\zeta (n + 1 - k)\,. {}\\ \end{array}$$

In particular, for n = 2, 3, 4 we have the identities

$$\displaystyle\begin{array}{rcl} \int _{0}^{1}\sum _{ m=0}^{\infty }\big\vert q_{ m}\alpha - p_{m}\big\vert ^{2}\,d\alpha & =& \frac{7\log 2\,\zeta (3)} {6\zeta (4)} + \frac{\zeta ^{2}(2)} {12\zeta (4)} -\frac{17} {24}\, =\, 0,39813\ldots \,, {}\\ \int _{0}^{1}\sum _{ m=0}^{\infty }\big\vert q_{ m}\alpha - p_{m}\big\vert ^{3}\,d\alpha & =& \frac{15\log 2\,\zeta (4)} {16\zeta (5)} + \frac{3\zeta (2)\zeta (3)} {16\zeta (5)} -\frac{49} {64}\, =\, 0,27019\ldots \,, {}\\ \int _{0}^{1}\sum _{ m=0}^{\infty }\big\vert q_{ m}\alpha - p_{m}\big\vert ^{4}\,d\alpha & =& \frac{31\log 2\,\zeta (5)} {40\zeta (6)} + \frac{7\zeta (2)\zeta (4)} {40\zeta (6)} + \frac{9\zeta ^{2}(3)} {80\zeta (6)} -\frac{129} {160}\, =\, 0,20731\ldots \,. {}\\ \end{array}$$

Taking \(\zeta (2s) \in \mathbb{Q}(\pi )\) into account, it follows that

$$\displaystyle\begin{array}{rcl} I_{1}& \in & \mathbb{Q}\big(\pi,\log (2),\zeta (3)\big)\,, {}\\ I_{2}& \in & \mathbb{Q}\big(\pi,\log (2),\zeta (3)\big)\,, {}\\ I_{3}& \in & \mathbb{Q}\big(\pi,\log (2),\zeta (3),\zeta (5)\big)\,, {}\\ I_{4}& \in & \mathbb{Q}\big(\pi,\log (2),\zeta (3),\zeta (5)\big)\,, {}\\ I_{5}& \in & \mathbb{Q}\big(\pi,\log (2),\zeta (3),\zeta (5),\zeta (7)\big)\,, {}\\ I_{6}& \in & \mathbb{Q}\big(\pi,\log (2),\zeta (3),\zeta (5),\zeta (7)\big)\,; {}\\ \end{array}$$

in particular we know \(I_{1},\ldots,I_{6} \in \mathbb{Q}\big(\pi,\log (2),\zeta (3),\zeta (5),\zeta (7)\big)\). This proves that I 1, , I 6 are algebraically dependent over \(\mathbb{Q}\). But indeed a stronger result holds, which can be verified using a suitable computer algebra system.

Corollary C1

The numbers I 1 ,I 2 ,I 3 ,I 4 are algebraically dependent over \(\mathbb{Q}\) . For x i = I i (i = 1,2,3,4) the algebraic equation

$$\displaystyle\begin{array}{rcl} 0& =& 10240x_{1}x_{3}x_{4} - 2976x_{1}^{2}x_{ 2} - 1488x_{1}^{2} - 5952x_{ 1}x_{2} + 5120x_{1}x_{3} + 7840x_{1}x_{4} {}\\ & & -2592x_{2}x_{3} +\, 6400x_{3}x_{4} + 944x_{1} - 4542x_{2} + 1904x_{3} + 4900x_{4} + 179 {}\\ \end{array}$$

holds.

The proof works by substituting the above expressions for I 1, I 2, I 3, I 4 into the equation given in the corollary, where additionally

$$\displaystyle{\zeta (2)\, =\, \frac{\pi ^{2}} {6}\,,\qquad \zeta (4)\, =\, \frac{\pi ^{4}} {90}\,,\qquad \zeta (6)\, =\, \frac{\pi ^{6}} {945}}$$

must be taken into account.

Note that

$$\displaystyle{I_{1},\ldots,I_{2n}\, \in \, \mathbb{Q}\big(\mathop{\underbrace{\pi,\log (2),\zeta (3),\zeta (5),\ldots,\zeta (2n + 1)}}\limits _{n+2}\big)\,.}$$

This proves

Corollary C2

For every integer n ≥ 3 any n + 3 numbers from the set {I 1 ,I 2 ,…I 2n } are algebraically dependent over \(\mathbb{Q}\) .

In this paper we focus our interest on a generalized error sum function. Let \(\alpha\) be a real number, and let \(k \geq 1\) and \(0 \leq l <k\) be integers. We define

$$\displaystyle{\mathcal{E}_{k,l}(\alpha )\,:=\,\sum _{ \begin{array}{c} m = 0 \\ q_{m} \equiv l\pmod k \end{array} }^{\infty }\vert q_{ m}\alpha -p_{m}\vert \,,}$$

in particular we set \(\mathcal{E}_{k}(\alpha ) = \mathcal{E}_{k,0}(\alpha )\). It is clear that \(\mathcal{E}_{1}(\alpha ) = \mathcal{E}(\alpha )\), and

$$\displaystyle{\sum _{l=0}^{k-1}\mathcal{E}_{ k,l}(\alpha )\, =\, \mathcal{E}(\alpha )\qquad (k \geq 1)\,.}$$

For \(k> 1\) the error sum function \(\mathcal{E}_{k}(\alpha )\) can be transformed into a more striking form. Since \(k\) does not divide \(q_{0} = 1\), the term for \(m = 0\) in \(\mathcal{E}_{k}(\alpha )\) does not occur. Moreover, for the convergents \(p_{m}/q_{m}\) \((m \geq 1\)) of \(\alpha\) satisfying \(q_{m} \equiv 0\pmod k\) we obtain the inequalities

$$\displaystyle{\vert q_{m}\alpha - p_{m}\vert \,\leq \, \frac{1} {q_{m}}\, \leq \, \frac{1} {k}\, \leq \, \frac{1} {2}\,.}$$

This proves

$$\displaystyle{ \mathcal{E}_{k}(\alpha )\, =\,\sum _{ \begin{array}{c} m = 1 \\ k\vert q_{m} \end{array} }^{\infty }\|q_{ m}\alpha \|\qquad (k> 1)\,. }$$
(1.1)

We continue to point out more basic properties of \(\mathcal{E}_{k}(\alpha )\) for \(k> 1\). Since \(q_{m}\) and \(q_{m+1}\) are coprime, at most every second term in \(\mathcal{E}_{k}(\alpha )\) does not vanish. So we obtain the following upper bound for \(\mathcal{E}_{k}(\alpha )\):

$$\displaystyle{\mathcal{E}_{k}(\alpha )\, =\,\sum _{ \begin{array}{c} m = 1 \\ k\vert q_{m} \end{array} }^{\infty }\vert q_{ m}\alpha -p_{m}\vert \,\leq \,\sum _{m=0}^{\infty }\vert q_{ 2m+1}\alpha -p_{2m+1}\vert \, =\, \frac{\mathcal{E}(\alpha ) -\mathcal{E}^{{\ast}}(\alpha )} {2} \qquad (k> 1)\,.}$$

The identities

$$\displaystyle{\mathcal{E}_{k}(\alpha )\, =\, \frac{\mathcal{E}(\alpha ) -\mathcal{E}^{{\ast}}(\alpha )} {2} \, =\,\alpha }$$

with \(k> 1\) hold for all numbers \(\alpha\) given by their continued fraction expansion

$$\displaystyle{\alpha \,=\, [0;k,1,k,1,k,1,\ldots ]\, =\, [0;\overline{k,1}]\, =\, \sqrt{\frac{1} {4} + \frac{1} {k}} -\frac{1} {2}\,,}$$

since for \(m \geq 0\) we have the congruence relations \(q_{2m+1} \equiv 0\pmod k\) and \(q_{2m} \equiv 1\pmod k\). Moreover,

$$\displaystyle{\mathcal{E}_{k}(\alpha )\, =\,\sum _{ m=0}^{\infty }\vert q_{ 2m+1}\alpha - p_{2m+1}\vert \, =\,\sum _{ m=0}^{\infty }\left (\frac{k + 2} {2} -\sqrt{\Big(\frac{k + 2} {2} \Big)^{2} - 1}\,\right )^{m+1}\, =\,\alpha \,.}$$

There exist real irrational numbers \(\alpha\) for which the series \(\mathcal{E}_{k}(\alpha )\) consists of at most finitely many terms, contrary to the series \(\mathcal{E}(\alpha )\). To prove the existence of such an irrational number, we define \(\alpha\) recursively by its continued fraction expansion \(\alpha = [0;a_{1},a_{2},\ldots ] = [0;2,1,1,2,2,4,6,\ldots ]\) as follows. We have

$$\displaystyle{q_{1} = 2\,,\quad q_{2} = 3\,,\quad q_{3} = 5\,,\quad q_{4} = 13\,,\quad q_{5} = 31\,,\quad q_{6} = 137\,,\quad q_{7} = 853\,.}$$

Now let us assume that for \(m \geq 8\) the denominators \(q_{m-1}\) and \(q_{m-2}\) are primes. Then, by the Dirichlet prime number theorem, there are infinitely many positive integers \(a\) such that \(q_{m} = aq_{m-1} + q_{m-2} \in \mathbb{P}\). The number \(a_{m}\) is uniquely defined by the smallest positive integer \(a\) satisfying this condition. Then, for every integer \(k> 1\), the series \(\mathcal{E}_{k}(\alpha )\) consists of at most one term. Furthermore, there are many situations in which \(\mathcal{E}_{k}(\alpha )\) vanishes.

Proposition 1.1

For every integer \(k> 1\) there are uncountably many irrational numbers \(\alpha\) such that \(\mathcal{E}_{k}(\alpha ) = 0\) .

To prove this proposition, let \(k> 1\) be any integer. We define an irrational number \(\alpha\) depending on \(k\) and on a sequence \((b_{n})_{n\geq 2}\) of positive integers by

$$\displaystyle{\alpha \,=\, \left [0;1,kb_{2},kb_{3},kb_{4},\ldots \right ]\, =\, \left [0;a_{1},a_{2},a_{3},\ldots \right ]\,.}$$

The denominators \(q_{m}\) of the convergents \(p_{m}/q_{m}\) of \(\alpha\) satisfy the recurrence formula

$$\displaystyle{q_{0} = 1\,,\quad q_{1} = a_{1} = 1\,,\quad q_{m+2} = a_{m+2}q_{m+1} + q_{m}\quad (m = 0,1,2,\ldots )\,.}$$

Since \(q_{m+2} \equiv q_{m}\pmod k\) for \(m = 0,1,2,\ldots\) it follows recursively that \(1 \equiv q_{0} \equiv q_{2} \equiv q_{4} \equiv \ldots \equiv q_{2m}\pmod k\) and, similarly, \(1 \equiv q_{1} \equiv q_{3} \equiv q_{5} \equiv \ldots \equiv q_{2m+1}\pmod k\) for \(m = 0,1,2,\ldots\). This proves that no denominator \(q_{m}\) is divisible by \(k\). Hence, \(\mathcal{E}_{k}(\alpha ) = 0\). By Cantor’s counting principle we have found uncountably many real numbers \(\alpha\) satisfying \(\mathcal{E}_{k}(\alpha ) = 0\).

The main goal of this paper is to study the behaviour of the numbers \(\int _{0}^{1}\mathcal{E}_{k,l}(\alpha )\,d\alpha\) depending on \(k\) and \(l\). For \(l = 0\) and \(k\) restricted to those numbers having no small prime divisors we prove the asymptotic behaviour of these integrals for \(k\) tending to infinity (Theorem 2.1 and Corollaries 2.22.5). For integers \(k\) having many small prime divisors the numbers \(\int _{0}^{1}\mathcal{E}_{k}(\alpha )\,d\alpha\) tend more quickly to zero than in the case \(k = p^{a}\) for fixed \(a \geq 1\) and primes \(p\) (Theorem 2.6 and Corollary 2.4). The integrals on the error sum functions \(\mathcal{E}_{k,l}\) with \(l> 0\) are treated in Theorem 2.7.

2 Statement of the Results

Let \(\mu: \mathbb{N} \rightarrow \{-1,0,1\}\) be the Möbius function, and let \(\zeta (s) =\sum _{ n=1}^{\infty }1/n^{s}\) for \(s \geq 2\) be the Riemann Zeta function. By \(J_{3}: \mathbb{N} \rightarrow \mathbb{N}\) we denote Jordan’s arithmetical function defined by \(J_{3}(1) = 1\) and

$$\displaystyle{ J_{3}(n)\, =\, n^{3}\prod _{ p\vert n}\Big(1 - \frac{1} {p^{3}}\Big)\qquad (n> 1)\,, }$$
(2.1)

where \(p\) runs through all prime divisors of \(n\). Moreover, for any integer \(n\) let \(\mathcal{D}_{n}\) denote the set of all positive divisors of \(n\). For every positive integer \(r\) we define the number \(T_{r}\) by

$$\displaystyle{ T_{r}\,:=\,\sum _{ n=r}^{\infty }\sum _{ 1\leq m\leq n/r}\frac{(-1)^{m+1}} {mn^{2}} \,. }$$
(2.2)

The identity from the following theorem can be considered as the main result of this paper, which contrasts with the property of the function \(\mathcal{E}_{k}(\alpha )\) given by Proposition 1.1.

Theorem 2.1

For every integer \(k> 1\) we have

$$\displaystyle{\int _{0}^{1}\mathcal{E}_{ k}(\alpha )\,d\alpha \, =\, \frac{1} {\zeta (3)}\sum _{r\in \mathcal{D}_{k}}\sum _{s\in \mathcal{D}_{r}}\frac{\mu (s)\mu (ks/r)T_{r}} {rJ_{3}(ks/r)} \,.}$$

Corollary 2.2

Let \(k> 1\) be any integer having \(t\) prime divisors, where \(P\) denotes the smallest prime divisor of \(k\) . Then we have

$$\displaystyle{\int _{0}^{1}\mathcal{E}_{ k}(\alpha )\,d\alpha \, =\, \frac{\zeta (2)} {2\zeta (3)k^{2}} + \mathcal{O}\Big( \frac{1} {k^{3}} + \frac{3^{t}} {k^{2}P}\Big)\,.}$$

Corollary 2.3

For all primes \(p\) we have

$$\displaystyle{\int _{0}^{1}\mathcal{E}_{ p}(\alpha )\,d\alpha \, =\, \frac{1} {p^{3} - 1}\Big(\frac{p^{2}T_{p}} {\zeta (3)} -\frac{3\zeta (2)\log 2} {2\zeta (3)} + \frac{1} {4}\Big)\, =\, \frac{\zeta (2)} {2\zeta (3)p^{2}} + \mathcal{O}\Big( \frac{1} {p^{3}}\Big)}$$

and

$$\displaystyle{ \frac{2} {77p^{2}}\, <\,\int _{ 0}^{1}\mathcal{E}_{ p}(\alpha )\,d\alpha \, <\, \frac{97} {109p^{2}}\qquad (\,p \geq 3)\,.}$$

Corollary 2.4

Let \(p\) be a prime and \(a\) be a positive integer. Set \(k:= p^{a}\) . Then we have

$$\displaystyle{\int _{0}^{1}\mathcal{E}_{ k}(\alpha )\,d\alpha \, =\, \frac{\zeta (2)} {2\zeta (3)k^{2}} + \mathcal{O}\Big( \frac{1} {k^{2+1/a}}\Big)\,.}$$

Corollary 2.5

Let \(k> 1\) be an integer having at most \(t\) prime divisors. The smallest prime divisor \(P\) of \(k\) satisfies \(P> k^{\varepsilon }\) for any \(0 <\varepsilon <1\) . Then we have

$$\displaystyle{\int _{0}^{1}\mathcal{E}_{ k}(\alpha )\,d\alpha \, =\, \frac{\zeta (2)} {2\zeta (3)k^{2}} + \mathcal{O}\Big( \frac{3^{t}} {k^{2+\varepsilon }}\Big)\,.}$$

To state the results in the subsequent theorems we need Euler’s totient \(\varphi\).

Theorem 2.6

For every integer \(k \geq 3\) we have

$$\displaystyle{ \frac{1} {k^{2}\log \log k}\, \ll \, \frac{\varphi (k)} {4k^{3}}\, <\,\int _{ 0}^{1}\mathcal{E}_{ k}(\alpha )\,d\alpha \, <\, \frac{\zeta (2)} {k^{2}} \,.}$$

For the numbers \(k = p_{1}p_{2}\cdots p_{r}\) given by the product on the first \(r \geq 2\) primes \(p_{1} = 2\) , \(p_{2} = 3\) , …we have

$$\displaystyle{\int _{0}^{1}\mathcal{E}_{ k}(\alpha )\,d\alpha \, \asymp \, \frac{1} {k^{2}\log \log k}\,.}$$

Theorem 2.6 shows that \(\int _{0}^{1}\mathcal{E}_{k}(\alpha )\,d\alpha \asymp k^{-2}\) does not hold for \(k \in \mathbb{N}\) tending to infinity. In the following theorem we estimate the integral on the error sum function \(\mathcal{E}_{k,l}(\alpha )\) for \(l> 0\), where the case \(l = 1\) is treated separately. By \((a,b)\) we denote the greatest common divisor of two integers \(a\) and \(b\).

Theorem 2.7

  1. (i)

    For every integer \(k \geq 2\) and \(l = 1\) we have

    $$\displaystyle{\frac{5} {8}+ \frac{\varphi (k + 1)} {4(k + 1)^{3}} <\int _{ 0}^{1}\mathcal{E}_{ k,1}(\alpha )\,d\alpha =\sum _{ \begin{array}{c} a = 1 \\ a \equiv 1\pmod k \end{array} }^{\infty }\sum _{ \begin{array}{c} b = 0 \\ (a,b) = 1\end{array} }^{a-1} \frac{1} {a(a + b)^{2}}-\frac{3} {8} \leq \frac{5} {8}+\frac{\zeta (2)} {k^{2}} \,.}$$
  2. (ii)

    For integers \(k \geq 3\) and \(2 \leq l <k\) we have

    $$\displaystyle{ \frac{\varphi (l)} {4l^{3}} <\int _{ 0}^{1}\mathcal{E}_{ k,l}(\alpha )\,d\alpha =\sum _{ \begin{array}{c} a = 1 \\ a \equiv l\pmod k \end{array} }^{\infty }\sum _{ \begin{array}{c} b = 0 \\ (a,b) = 1\end{array} }^{a-1} \frac{1} {a(a + b)^{2}} \leq \frac{1} {l^{2}}+\frac{\zeta (2)} {k^{2}} \,.}$$

For two consecutive primes \(p_{r-1}\) and \(p_{r}\) \((r \geq 2)\) it follows from (ii) in Theorem 2.7 by Bertrand’s Postulate and Theorem 9 in [8] that

$$\displaystyle{\int _{0}^{1}\mathcal{E}_{ p_{r},p_{r-1}}(\alpha )\,d\alpha \, \asymp \, \frac{1} {p_{r}^{2}}\, \asymp \, \frac{1} {r^{2}\log ^{2}r}\,.}$$

3 Auxiliary Results

Lemma 3.1

Let \(k> 1\) be an integer, and let \(r\) be any positive divisor of \(k\) . Then we have the identity

$$\displaystyle{\sum _{\begin{array}{c} d = 1 \\ (d,k) = k/r \end{array} }^{\infty }\frac{\mu (d)} {d^{3}} = \frac{1} {\zeta (3)}\sum _{s\in \mathcal{D}_{r}}\frac{\mu (s)\mu (ks/r)} {J_{3}(ks/r)} \,.}$$

Proof

We obtain

$$\displaystyle\begin{array}{rcl} S& \,:=\,& \sum _{\begin{array}{c} d = 1 \\ (d,k) = k/r \end{array} }^{\infty }\frac{\mu (d)} {d^{3}} \, =\,\sum _{ \begin{array}{c} m = 1 \\ (m,r) = 1 \end{array} }^{\infty } \frac{\mu (mk/r)} {(mk/r)^{3}}\, =\, \Big(\frac{r} {k}\Big)^{3}\sum _{ m=1}^{\infty }\sum _{ \begin{array}{c} s \geq 1 \\ s\vert (m,r) \end{array} }\frac{\mu (s)\mu (mk/r)} {m^{3}} {}\\ & \,=\,& \Big(\frac{r} {k}\Big)^{3}\sum _{ \begin{array}{c} s \geq 1 \\ s\vert r\end{array} }\mu (s)\sum _{\begin{array}{c} m = 1 \\ m \equiv 0\pmod s \end{array} }^{\infty }\frac{\mu (mk/r)} {m^{3}} \, =\, \Big(\frac{r} {k}\Big)^{3}\sum _{ s\in \mathcal{D}_{r}}\frac{\mu (s)} {s^{3}} \sum _{m=1}^{\infty }\frac{\mu (mks/r)} {m^{3}} \,. {}\\ \end{array}$$

For any positive integer \(t\) we have

$$\displaystyle{\sum _{m=1}^{\infty }\frac{\mu (mt)} {m^{3}} \, =\,\mu (t)\sum _{\begin{array}{c} m = 1 \\ (m,t) = 1 \end{array} }^{\infty }\frac{\mu (m)} {m^{3}} \, =\, \frac{\mu (t)t^{3}} {\zeta (3)J_{3}(t)}\,,}$$

where the identity on the right-hand side can be obtained by using the method explained in [10]. Substituting the last expression into \(S\) by setting \(t = ks/r\), we complete the proof of the desired identity from the lemma. \(\square\)

Lemma 3.2

For every positive integer \(r\) we have

$$\displaystyle\begin{array}{rcl} & T_{r}\, =\,\sum _{ n=1}^{\infty }\sum _{m=0}^{nr-1} \frac{1} {n(nr+m)^{2}} \,,\qquad T_{1}\, =\, \frac{3} {2}\zeta (2)\log 2 -\frac{1} {4}\zeta (3)\,, & {}\\ & T_{r}\, =\, \frac{\zeta (2)} {2r} + \mathcal{O}\Big( \frac{1} {r^{2}} \Big)\,,\qquad \mbox{ and}\qquad \frac{1} {2r}\, <\, T_{r}\, <\quad \left \{\begin{array}{lll} \frac{1} {r - 1} &\mbox{ if} \,\, &r \geq 2\,, \\ \\ 1 &\mbox{ if}&r \geq 1\,. \end{array} \right.& {}\\ \end{array}$$

In particular, we have \(T_{r} <2/r\) for all \(r \geq 1\) .

Proof

Let \(r \geq 1\) be an integer. To prove the alternative expression of \(T_{r}\), we first observe that

$$\displaystyle{ \sum _{n=1}^{\infty }\sum _{ m=0}^{nr-1} \frac{1} {n(nr + m)^{2}}\, =\,\sum _{ n=1}^{\infty }\frac{1} {n}\sum _{k=nr}^{2nr-1} \frac{1} {k^{2}}\, =\,\sum _{ k=r}^{\infty } \frac{1} {k^{2}}\sum _{n=\lfloor k/2r\rfloor +1}^{\lfloor k/r\rfloor }\frac{1} {n}\,, }$$
(3.1)

where the last identity follows by interchanging the order of summation, and where \(\lfloor \eta \rfloor\) denotes the floor function, i.e. the greatest integer not exceeding \(\eta\). Next, let \(\beta \geq 1\) be a real number, and

$$\displaystyle{\delta \,:= \quad \left \{\begin{array}{rll} 1\,,&\mbox{ if} \,\, &\lfloor \beta \rfloor \equiv 1\pmod 2\,,\\ 0\,, &\mbox{ if} \,\, &\lfloor \beta \rfloor \equiv 0\pmod 2\,. \end{array} \right.}$$

Then we have

$$\displaystyle{\sum _{n=1}^{\lfloor \beta \rfloor }\frac{1} {n}\, =\,\sum _{ m=1}^{\lfloor \beta /2\rfloor +\delta } \frac{1} {2m - 1} +\sum _{ m=1}^{\lfloor \beta /2\rfloor } \frac{1} {2m}\,,}$$

which yields, equivalently,

$$\displaystyle\begin{array}{rcl} \sum _{n=\lfloor \beta /2\rfloor +1}^{\lfloor \beta \rfloor }\frac{1} {n}& \,=\,& \sum _{m=1}^{\lfloor \beta /2\rfloor +\delta } \frac{1} {2m - 1} + \frac{1} {2}\sum _{m=1}^{\lfloor \beta /2\rfloor } \frac{1} {m} -\sum _{n=1}^{\lfloor \beta /2\rfloor }\frac{1} {n} {}\\ & \,=\,& \sum _{m=1}^{\lfloor \beta /2\rfloor +\delta } \frac{1} {2m - 1} -\frac{1} {2}\sum _{m=1}^{\lfloor \beta /2\rfloor } \frac{1} {m} {}\\ & \,=\,& \sum _{m=1}^{\lfloor \beta \rfloor }\frac{(-1)^{m+1}} {m} \,. {}\\ \end{array}$$

With \(\beta = k/r\) for \(k \geq r\) we conclude from (3.1) and (2.2) that

$$\displaystyle{\sum _{n=1}^{\infty }\sum _{ m=0}^{nr-1} \frac{1} {n(nr + m)^{2}}\, =\,\sum _{ k=r}^{\infty } \frac{1} {k^{2}}\sum _{1\leq m\leq k/r}\frac{(-1)^{m+1}} {m} \, =\, T_{r}\,.}$$

For the asymptotic expansion of \(T_{r}\) we apply Euler’s summation formula to the function \(f(x) = 1/(x + nr - 1)^{2}\): Let \(B(\eta ) =\eta -[\eta ] - 1/2\). Then,

$$\displaystyle\begin{array}{rcl} \sum _{m=1}^{nr} \frac{1} {(x + nr - 1)^{2}}& \,=\,& \int _{1}^{nr}f(x)\,dx +\int _{ 1}^{nr}B(x)f'(x)\,dx - B(1)f(1) - B(nr)f(nr) {}\\ & \,=\,& \frac{1} {nr} - \frac{1} {2nr - 1} + \mathcal{O}\Big( \frac{1} {n^{2}r^{2}}\Big)\,, {}\\ & & {}\\ \end{array}$$

which yields

$$\displaystyle\begin{array}{rcl} T_{r}& \,=\,& \sum _{n=1}^{\infty }\frac{1} {n}\sum _{m=1}^{nr} \frac{1} {(x + nr - 1)^{2}} {}\\ & \,=\,& \sum _{n=1}^{\infty } \frac{1} {n^{2}r} -\sum _{n=1}^{\infty } \frac{1} {n(2nr - 1)} + \mathcal{O}\Big(\sum _{n=1}^{\infty } \frac{1} {n^{3}r^{2}}\Big) {}\\ & \,=\,& \frac{\zeta (2)} {r} -\sum _{n=1}^{\infty }\Big( \frac{1} {2n^{2}r} + \frac{1} {2n^{2}r(2nr - 1)}\Big) + \mathcal{O}\Big( \frac{1} {r^{2}}\Big) {}\\ & \,=\,& \frac{\zeta (2)} {r} -\frac{\zeta (2)} {2r} + \mathcal{O}\Big(\sum _{n=1}^{\infty } \frac{1} {n^{2}r(2nr - 1)}\Big) + \mathcal{O}\Big( \frac{1} {r^{2}}\Big) {}\\ & \,=\,& \frac{\zeta (2)} {2r} + \mathcal{O}\Big( \frac{1} {r^{2}}\Big)\,. {}\\ \end{array}$$

\(T_{1}\) is a special case of the multivariate zeta function \(\zeta (m,n)\), see [1, Sect. 2.6]:

$$\displaystyle\begin{array}{rcl} T_{1}& \,=\,& \sum _{n=1}^{\infty }\sum _{ m=1}^{n}\frac{(-1)^{m+1}} {mn^{2}} \, =\,\sum _{ n=1}^{\infty }\sum _{ m=1}^{n-1}\frac{(-1)^{m+1}} {mn^{2}} +\sum _{ n=1}^{\infty }\frac{(-1)^{n+1}} {n^{3}} {}\\ & \,=\,& -\zeta (2,-1) + \frac{3} {4}\zeta (3)\, =\, \frac{3} {2}\zeta (2)\log 2 -\frac{1} {4}\zeta (3)\,. {}\\ \end{array}$$

The bounds for \(T_{r}\) stated in the lemma follow from (2.2) by using the inequalities

$$\displaystyle\begin{array}{rcl} & \frac{1} {2}\, \leq \,\sum _{1\leq m\leq n/r}\frac{(-1)^{m+1}} {m} \, \leq \, 1\qquad (n \geq r)\,, & {}\\ & \sum _{m=r}^{\infty } \frac{1} {m^{2}} <\int _{ r-1}^{\infty }\frac{dt} {t^{2}} = \frac{1} {r-1} \leq \frac{2} {r}\quad (r \geq 2)\,,\qquad T_{1} = 1.409757\ldots <2\,,& {}\\ \end{array}$$

and

$$\displaystyle{\sum _{m=r}^{\infty } \frac{1} {m^{2}}>\int _{ r}^{\infty }\frac{dt} {t^{2}} = \frac{1} {r}\quad (r \geq 1)\,.}$$

4 Proof of Theorem 2.1

Let \(\chi _{k,l}: \mathbb{N} \rightarrow \{ 0,1\}\) be defined by

$$\displaystyle{\chi _{k,l}(n)\,:= \quad \left \{\begin{array}{ll} 1\,&\mbox{ if $n \equiv l\pmod k$}\,,\\ 0\, &\mbox{ otherwise}\,. \end{array} \right.}$$

Note that \(\chi _{k,l}(1) = 1\) holds if and only if \(l = 1\). At the beginning of the proof of Theorem 2.1 we follow the lines in Sect. 4 in [6] and modify the arguments. Let \(m\) and \(a_{1},\ldots,a_{m}\) be positive integers. We define the rational numbers \(\xi _{1}\), \(\xi _{2}\) by their continued fraction expansion:

$$\displaystyle{\xi _{1}\, =\, \left [0;a_{1},\ldots,a_{m-1},a_{m}\right ]\qquad \mbox{ and}\qquad \xi _{2}\, =\, \left [0;a_{1},\ldots,a_{m-1},a_{m} + 1\right ]\,.}$$

We have \(\xi _{1} <\xi _{2}\) for even \(m\) and \(\xi _{2} <\xi _{1}\) for odd \(m\). We define the interval \(I_{m}\) by \(I_{m} = (\xi _{1},\xi _{2})\) for even \(m\) and \(I_{m} = (\xi _{2},\xi _{1})\) otherwise. It is well known that the intervals \(I_{m}\) are disjoint for different positive integers \(a_{1},\ldots,a_{m}\), and that for any fixed \(m\) the union of all closed intervals \(\overline{I}_{m}\) gives the interval \([0,1]\). For this decomposition of \([0,1]\) we express the integral as follows:

$$\displaystyle\begin{array}{rcl} \int _{0}^{1}\mathcal{E}_{ k,l}(\alpha )\,d\alpha & \,=\,& \int _{0}^{1}\sum _{ m=0}^{\infty }(-1)^{m}\chi _{ k,l}(q_{m})(q_{m}\alpha - p_{m})\,d\alpha {}\\ & \,=\,& \frac{\chi _{k,l}(1)} {2} +\sum _{ m=1}^{\infty }(-1)^{m}\int _{ 0}^{1}\chi _{ k,l}(q_{m})(q_{m}\alpha - p_{m})\,d\alpha {}\\ & \,=\,& \frac{\chi _{k,l}(1)} {2} +\sum _{ m=1}^{\infty }(-1)^{m}\sum _{ a_{1}=1}^{\infty }\ldots \sum _{ a_{m}=1}^{\infty }\int _{ I_{m}}\chi _{k,l}(q_{m})(q_{m}\alpha - p_{m})\,d\alpha {}\\ & \,=\,& \frac{\chi _{k,l}(1)} {2} +\sum _{ m=1}^{\infty }\sum _{ a_{1}=1}^{\infty }\ldots \sum _{ a_{m}=1}^{\infty }\chi _{ k,l}(q_{m})\int _{\xi _{1}}^{\xi _{2} }(q_{m}\alpha - p_{m})\,d\alpha \,. {}\\ \end{array}$$

Note that \(p_{m}\) and \(q_{m}\) depend on \(a_{1},\ldots,a_{m}\). The continued fraction expansion of every point \(\alpha \in I_{m}\) has the form \(\alpha = [0;a_{1},\ldots,a_{m-1},a_{m},\ldots ]\). Hence the convergents \(p_{\nu }/q_{\nu }\) for \(\nu \leq m\) depend on \(I_{m}\), but not on \(\alpha \in I_{m}\). Therefore we compute the above integral on \([\xi _{1},\xi _{2}]\) by

$$\displaystyle{\int _{\xi _{1}}^{\xi _{2} }(q_{m}\alpha - p_{m})\,d\alpha = (\xi _{2} -\xi _{1})\,\frac{(\xi _{2} +\xi _{1})q_{m} - 2p_{m}} {2} \,.}$$

Using

$$\displaystyle{\xi _{1} = \frac{p_{m}} {q_{m}}\quad \mbox{ and}\quad \xi _{2} = \frac{(a_{m} + 1)p_{m-1} + p_{m-2}} {(a_{m} + 1)q_{m-1} + q_{m-2}} }$$

we compute the expressions

$$\displaystyle{\xi _{2} -\xi _{1} = \frac{(-1)^{m}} {(q_{m} + q_{m-1})q_{m}}}$$

and

$$\displaystyle{\xi _{2} +\xi _{1} = \frac{p_{m-1}q_{m} + q_{m-1}p_{m} + 2p_{m}q_{m}} {(q_{m} + q_{m-1})q_{m}} \,,}$$

which give

$$\displaystyle{\int _{\xi _{1}}^{\xi _{2} }(q_{m}\alpha - p_{m})\,d\alpha = \frac{1} {2q_{m}(q_{m} + q_{m-1})^{2}}\,,}$$

and consequently

$$\displaystyle{ \int _{0}^{1}\mathcal{E}_{ k,l}(\alpha )\,d\alpha \, =\, \frac{\chi _{k,l}(1)} {2} +\sum _{ m=1}^{\infty }\sum _{ a_{1}=1}^{\infty }\ldots \sum _{ a_{m}=1}^{\infty } \frac{\chi _{k,l}(q_{m})} {2q_{m}(q_{m} + q_{m-1})^{2}}\,. }$$
(4.1)

For the denominators of two subsequent convergents of the continued fraction of α = 〈0; a 1, , a m , 〉 it is well known that (q m , q m−1) = 1. For fixed q m  = a we count the solutions of q m−1 = b with (a, b) = 1 and 0 ≤ b ≤ a − 1 in the multiple sum on the left-hand side of (4.1). It is necessary to distinguish the cases m ≥ 2 and m = 1.

Case 1: :

m ≥ 2. First let a 1 = 1. Then,

$$\displaystyle{\frac{q_{m-1}} {q_{m}} =\langle 0;a_{m},\ldots,a_{2},1\rangle =\langle 0;a_{m},\ldots,a_{2} + 1\rangle \,.}$$

For a 1 ≥ 2 we have

$$\displaystyle{\frac{q_{m-1}} {q_{m}} =\langle 0;a_{m},\ldots,a_{2},a_{1}\rangle =\langle 0;a_{m},\ldots,a_{2},a_{1} - 1,1\rangle \,.}$$
Case 2: :

m = 1. For a 1 = 1 we have a unique representation of the fraction

$$\displaystyle{\frac{q_{m-1}} {q_{m}} = \frac{q_{0}} {q_{1}} = \frac{1} {a_{1}} = \frac{1} {1} =\langle 0;1\rangle \,,}$$

since the integer part a 0 = 0 must not be changed. For a 1 ≥ 2 there are again two representations:

$$\displaystyle{\frac{q_{m-1}} {q_{m}} = \frac{q_{0}} {q_{1}} = \frac{1} {a_{1}} =\langle 0;a_{1}\rangle =\langle 0;a_{1} - 1,1\rangle \,.}$$

Therefore it becomes clear that for any fixed \(q_{m} = a\) every coprime integer \(b\) with \(0 \leq b \leq a - 1\) occurs exactly two times in the multiple sum on the right-hand side of (4.1), except for \(m = 1\) and \(a_{1} = 1\). For this exceptional case we separate the term

$$\displaystyle{ \frac{\chi _{k,l}(q_{1})} {2q_{1}(q_{1} + q_{0})^{2}}\, =\, \frac{\chi _{k,l}(1)} {8} }$$

from the multiple sum. Therefore we obtain

$$\displaystyle\begin{array}{rcl} & & \int _{0}^{1}\mathcal{E}_{ k,l}(\alpha )\,d\alpha \\ & \,=\,& \frac{\chi _{k,l}(1)} {2} +\sum _{ m=2}^{\infty }\sum _{ a_{1}=1}^{\infty }\ldots \sum _{ a_{m}=1}^{\infty } \frac{\chi _{k,l}(q_{m})} {2q_{m}(q_{m} + q_{m-1})^{2}} +\sum _{ a_{1}=2}^{\infty } \frac{\chi _{k,l}(q_{1})} {2q_{1}(q_{1} + 1)^{2}} + \frac{\chi _{k,l}(1)} {8} \\ & \,=\,& \frac{5\chi _{k,l}(1)} {8} +\sum _{ a=1}^{\infty }\sum _{ \begin{array}{c} b = 1 \\ (a,b) = 1 \end{array} }^{a-1} \frac{\chi _{k,l}(a)} {a(a + b)^{2}} \\ & \,=\,& -\frac{3\chi _{k,l}(1)} {8} +\sum _{ \begin{array}{c} a = 1 \\ a \equiv l\pmod k \end{array} }^{\infty }\sum _{ \begin{array}{c} b = 0 \\ (a,b) = 1 \end{array} }^{a-1} \frac{1} {a(a + b)^{2}}\,.{}\end{array}$$
(4.2)

Note that for \(b = 0\) the condition \((a,0) = 1\) holds for \(a = 1\) only. For the proof of Theorem 2.1 we now assume that \(l = 0\), so that \(\chi _{k,l}(1)\) vanishes. Then (4.2) simplifies to

$$\displaystyle{ \int _{0}^{1}\mathcal{E}_{ k}(\alpha )\,d\alpha \, =\,\sum _{ \begin{array}{c} a = 1 \\ k\vert a \end{array} }^{\infty }\sum _{ \begin{array}{c} b = 0 \\ (a,b) = 1 \end{array} }^{a-1} \frac{1} {a(a + b)^{2}}\,. }$$
(4.3)

Next, we express the arithmetic condition \((a,b) = 1\) on \(a\) and \(b\) from the inner sum by the Möbius function. Then we proceed by interchanging the order of the resulting triple sum. Here, \([d,k]\) denotes the least common multiple of \(d\) and \(k\).

$$\displaystyle\begin{array}{rcl} \int _{0}^{1}\mathcal{E}_{ k}(\alpha )\,d\alpha & \,=\,& \sum _{\begin{array}{c} a = 1\\ k\vert a \end{array} }^{\infty }\sum _{ b=0}^{a-1}\sum _{ \begin{array}{c} d> 0 \\ d\vert (a,b) \end{array} } \frac{\mu (d)} {a(a + b)^{2}}\, =\,\sum _{ d=1}^{\infty }\sum _{ \begin{array}{c} a = 1 \\ \left [d,k\right ]\vert a \end{array} }^{\infty }\sum _{ \begin{array}{c} b = 0 \\ d\vert b \end{array} }^{a-1} \frac{\mu (d)} {a(a + b)^{2}} {}\\ & \,=\,& \sum _{d=1}^{\infty }\sum _{ n=1}^{\infty }\sum _{ m=0}^{(\left [d,k\right ]n-1)/d} \frac{\mu (d)} {[d,k]n\big([d,k]n + dm\big)^{2}} {}\\ & \,=\,& \sum _{r\in \mathcal{D}_{k}}\sum _{\begin{array}{c} d = 1 \\ (d,k) = k/r \end{array} }^{\infty }\sum _{ n=1}^{\infty }\sum _{ m=0}^{\left [d,k\right ]n/d-1} \frac{\mu (d)} {[d,k]n\big([d,k]n + dm\big)^{2}}. {}\\ \end{array}$$

The condition \((d,k) = k/r\) implies that

$$\displaystyle{[d,k]\, =\, \frac{dk} {(d,k)}\, =\, dr\,.}$$

Hence the above multiple sum takes the form

$$\displaystyle\begin{array}{rcl} \int _{0}^{1}\mathcal{E}_{ k}(\alpha )\,d\alpha & \,=\,& \sum _{r\in \mathcal{D}_{k}}\sum _{\begin{array}{c} d = 1 \\ (d,k) = k/r \end{array} }^{\infty }\sum _{ n=1}^{\infty }\sum _{ m=0}^{nr-1} \frac{\mu (d)} {nrd(nrd + md)^{2}} {}\\ & \,=\,& \sum _{r\in \mathcal{D}_{k}}\frac{1} {r}\Big(\sum _{\begin{array}{c} d = 1 \\ (d,k) = k/r \end{array} }^{\infty }\frac{\mu (d)} {d^{3}} \Big)\Big(\sum _{n=1}^{\infty }\sum _{ m=0}^{nr-1} \frac{1} {n(nr + m)^{2}}\Big)\,.{}\\ \end{array}$$

Finally, we express the two terms in brackets by the identities given in Lemma 3.1 and Lemma 3.2, respectively. This completes the proof of the theorem. \(\square\)

5 Proofs of Corollaries 2.2–2.5

Proof of Corollary 2.2

From the multiple sum in Theorem 2.1 we separate the term for \(r = k\) and \(s = 1\):

$$\displaystyle\begin{array}{rcl} \int _{0}^{1}\mathcal{E}_{ k}(\alpha )\,d\alpha & \,=\,& \frac{T_{k}} {\zeta (3)k} + \frac{1} {\zeta (3)}\mathop{\sum \quad \sum }\limits _{\begin{array}{c} r \in \mathcal{D}_{k}\,\,s \in \mathcal{D}_{r} \\ r\not =k \vee s\not =1 \end{array} }\frac{\mu (s)\mu (ks/r)T_{r}} {rJ_{3}(ks/r)} \\ & \,=\,& \frac{T_{k}} {\zeta (3)k} + \mathcal{O}\Big(\mathop{\sum \quad \sum }\limits _{\begin{array}{c} r \in \mathcal{D}_{k}\,\,s \in \mathcal{D}_{r} \\ r\not =k \vee s\not =1 \end{array} }\frac{\vert \mu (s)\mu (ks/r)\vert T_{r}} {rJ_{3}(ks/r)} \Big) \\ & \,=\,& \frac{T_{k}} {\zeta (3)k} + \mathcal{O}\Big(\mathop{\sum \quad \sum }\limits _{\begin{array}{c} r \in \mathcal{D}_{k}\,\,s \in \mathcal{D}_{r} \\ r\not =k \vee s\not =1 \end{array} }\vert \mu (s)\mu (ks/r)\vert \, \frac{r} {k^{3}s^{3}}\Big)\,.{}\end{array}$$
(5.1)

Here we have applied the inequalities \(T_{r} \ll 1/r\) (Lemma 3.2) and

$$\displaystyle\begin{array}{rcl} J_{3}(n)\, =\, n^{3}\prod _{ p\vert n}\Big(1 - \frac{1} {p^{3}}\Big)\,>\, n^{3}\prod _{ p\in \mathbb{P}}\Big(1 - \frac{1} {p^{3}}\Big)\, =\, \frac{n^{3}} {\zeta (3)}\qquad (n \geq 1)& &{}\end{array}$$
(5.2)

(see [8, Theorem 280]). In order to estimate \(k^{3}s^{3}/r\) we discuss the following two cases. Recall that \(r\vert k\) and \(s\vert r\), and that the number \(P\) is the smallest prime divisor of \(k\).

Case 1: :

\(1 \leq r <k\) and \(s = 1\).

$$\displaystyle{\frac{k^{3}s^{3}} {r} \, =\, \frac{k^{3}} {r} \, \geq \, \frac{k^{3}} {k/P}\, =\, Pk^{2}\,.}$$
Case 2: :

\(1 \leq r \leq k\) and \(s \geq P\).

$$\displaystyle{\frac{k^{3}s^{3}} {r} \, \geq \, \frac{k^{3}s^{3}} {k} \, \geq \, P^{3}k^{2}\, \geq \, Pk^{2}\,.}$$

Using these bounds we estimate the error term in (5.1). This gives

$$\displaystyle\begin{array}{rcl} \int _{0}^{1}\mathcal{E}_{ k}(\alpha )\,d\alpha & \,=\,& \frac{T_{k}} {\zeta (3)k} + \mathcal{O}\Big(\mathop{\sum \quad \sum }\limits _{\begin{array}{c} r \in \mathcal{D}_{k}\,\,s \in \mathcal{D}_{r} \\ r\not =k \vee s\not =1 \end{array} }\vert \mu (s)\mu (ks/r)\vert \, \frac{1} {Pk^{2}}\Big) {}\\ & \,=\,& \frac{T_{k}} {\zeta (3)k} + \mathcal{O}\Big( \frac{1} {Pk^{2}}\sum _{r\in \mathcal{D}_{k}}\sum _{s\in \mathcal{D}_{r}}\vert \mu (s)\mu (ks/r)\vert \Big) {}\\ & \,=\,& \frac{\zeta (2)} {2\zeta (3)k^{2}} + \mathcal{O}\Big( \frac{1} {k^{3}}\Big) + \mathcal{O}\Big( \frac{1} {Pk^{2}}\sum _{r\in \mathcal{D}_{k}}\sum _{s\in \mathcal{D}_{r}}\vert \mu (s)\mu (ks/r)\vert \Big)\,, {}\\ \end{array}$$

where we have applied the asymptotic formula for \(T_{k}\) from Lemma 3.2. To complete the proof of the corollary we finally prove the identity

$$\displaystyle{q_{k}:=\sum _{r\in \mathcal{D}_{k}}\sum _{s\in \mathcal{D}_{r}}\vert \mu (s)\mu (ks/r)\vert \, =\, 3^{t}\qquad (k = p_{ 1}^{a_{1} }\cdots p_{t}^{a_{t} })}$$

by induction with respect to \(t\). For \(t = 1\) let \(k = p^{a}\). We count three pairs \([r,s] \in \mathcal{D}_{k} \times \mathcal{D}_{r}\) such that \(\vert \mu (s)\mu (ks/r)\vert = 1\) given by \([p^{a},1]\), \([p^{a},p]\), and \([p^{a-1},1]\). Now we assume that q k = 3t holds for all integers \(k'\) having \(t\) prime divisors. Let \(k = p_{1}^{a_{1}}\cdots p_{t}^{a_{t}}p_{t+1}^{a_{t+1}}\) and \(k' = p_{1}^{a_{1}}\cdots p_{t}^{a_{t}}\). While \([r',s'] \in \mathcal{D}_{k'} \times \mathcal{D}_{r'}\) runs through all \(3^{t}\) pairs which are counted for \(q_{k'}\), we obtain \(q_{k}\) by counting the \(3 \cdot 3^{t}\) pairs \([r,s] \in \mathcal{D}_{k} \times \mathcal{D}_{r}\) given by \([r'p_{t+1}^{a_{t+1}},s']\), \([r'p_{t+1}^{a_{t+1}},s'p_{t+1}]\), and \([r'p_{t+1}^{a_{t+1}-1},s']\). This completes the proof of Corollary 2.2. \(\square\)

Proof of Corollary 2.3

For \(k = p \in \mathbb{P}\) the multiple sum in Theorem 2.1 consists of three terms corresponding to \([r = 1,s = 1]\), \([r = p,s = 1]\), and \([r = p,s = p]\). Therefore, we obtain

$$\displaystyle{\int _{0}^{1}\mathcal{E}_{ p}(\alpha )\,d\alpha \, =\, \frac{1} {\zeta (3)}\Big(- \frac{T_{1}} {J_{3}(\,p)} + \frac{T_{p}} {p} + \frac{T_{p}} {pJ_{3}(\,p)}\Big)\,.}$$

With \(J_{3}(\,p) = p^{3} - 1\) and

$$\displaystyle{T_{1}\, =\, \frac{3} {2}\zeta (2)\log 2 -\frac{1} {4}\zeta (3)}$$

given in Lemma 3.2 we prove the identity stated in the corollary. Now, for \(p \geq 3\) we have with Lemma 3.2:

$$\displaystyle{\int _{0}^{1}\mathcal{E}_{ p}(\alpha )\,d\alpha \, <\, \frac{1} {p^{3} - 1}\Big( \frac{p^{2}} {(\,p - 1)\zeta (3)} -\frac{3\zeta (2)\log 2} {2\zeta (3)} + \frac{1} {4}\Big)\, <\, \frac{97} {109p^{2}}\,,}$$

and

$$\displaystyle{\int _{0}^{1}\mathcal{E}_{ p}(\alpha )\,d\alpha \,>\, \frac{1} {p^{3} - 1}\Big( \frac{p} {2\zeta (3)} -\frac{3\zeta (2)\log 2} {2\zeta (3)} + \frac{1} {4}\Big)\,>\, \frac{2} {77p^{2}}\,.}$$

This proves the inequalities stated in Corollary 2.3. \(\square\)

Proofs of Corollaries 2.4 and 2.5

We apply Corollary 2.2 with \(k = p^{a}\), \(t = 1\), and \(P = p\). Then the error term in Corollary 2.2 takes the form

$$\displaystyle{\mathcal{O}\Big( \frac{1} {k^{3}} + \frac{3^{t}} {k^{2}P}\Big)\, =\, \mathcal{O}\Big( \frac{1} {k^{3}} + \frac{1} {k^{2}p}\Big)\, =\, \mathcal{O}\Big( \frac{1} {k^{3}} + \frac{1} {k^{2}k^{1/a}}\Big)\, =\, \mathcal{O}\Big( \frac{1} {k^{2+1/a}}\Big)\,.}$$

Thus Corollary 2.4 is proven. Also Corollary 2.5 follows immediately from Corollary 2.2. \(\square\)

6 Proofs of Theorems 2.6 and 2.7

Proof of Theorem 2.6

For the upper bound we estimate the right-hand side of the identity in (4.3). Let \(k \geq 2\). Then

$$\displaystyle{\int _{0}^{1}\mathcal{E}_{ k}(\alpha )\,d\alpha \, =\,\sum _{ m=1}^{\infty }\sum _{ \begin{array}{c} b = 1 \\ (km,b) = 1 \end{array} }^{km-1} \frac{1} {km(km + b)^{2}}\, <\,\sum _{ m=1}^{\infty }\sum _{ b=1}^{km-1} \frac{1} {k^{3}m^{3}}\, <\,\sum _{ m=1}^{\infty } \frac{1} {k^{2}m^{2}}\, =\, \frac{\zeta (2)} {k^{2}} \,.}$$

On the other hand, the lower bound for the integral in Theorem 2.6 follows from the identity (4.3), too. Here we assume that \(k \geq 3\).

$$\displaystyle{ \int _{0}^{1}\mathcal{E}_{ k}(\alpha )\,d\alpha \, \geq \,\sum _{\begin{array}{c} b = 1 \\ (k,b) = 1 \end{array} }^{k-1} \frac{1} {k(k + b)^{2}}\, \geq \, \frac{\varphi (k)} {k(2k - 1)^{2}}\,>\, \frac{\varphi (k)} {4k^{3}}\, \gg \, \frac{1} {k^{2}\log \log k}\,. }$$
(6.1)

The inequality on the right-hand side involving a lower bound of Euler’s totient follows from Theorem 328 in [8].

Next, let \(k = p_{1}p_{2}\cdots p_{r}\) for some positive integer \(r \geq 2\). Then we have

$$\displaystyle{\log k\, =\,\sum _{\begin{array}{c} p \leq p_{r} \\ p \in \mathbb{P} \end{array} }\log p\, =\,\vartheta (\,p_{r})\, \ll \, p_{r}}$$

by Theorem 414 in [8]. Applying additionally Theorem 429 in [8], we obtain

$$\displaystyle{ \prod _{\begin{array}{c} p\vert k \\ p \in \mathbb{P}\end{array} }\Big(1-\frac{1} {p}\Big)\, =\,\prod _{\begin{array}{c} p \leq p_{r} \\ p \in \mathbb{P} \end{array} }\Big(1-\frac{1} {p}\Big)\, \ll \, \frac{1} {\log p_{r}}\, \ll \, \frac{1} {\log \log k}\,. }$$
(6.2)

Moreover,

$$\displaystyle\begin{array}{rcl} \int _{0}^{1}\mathcal{E}_{ k}(\alpha )\,d\alpha & \,\leq \,& \sum _{m=1}^{\infty }\sum _{ \begin{array}{c} b = 1 \\ (km,b) = 1 \end{array} }^{km-1} \frac{1} {k^{3}m^{3}}\, =\, \frac{1} {k^{3}}\sum _{m=1}^{\infty }\frac{\varphi (km)} {m^{3}} {}\\ & \,=\,& \frac{1} {k^{2}}\sum _{m=1}^{\infty } \frac{1} {m^{2}}\prod _{\begin{array}{c} p\vert km \\ p \in \mathbb{P}\end{array} }\Big(1 -\frac{1} {p}\Big)\, \leq \, \frac{\zeta (2)} {k^{2}} \prod _{\begin{array}{c} p\vert k \\ p \in \mathbb{P} \end{array} }\Big(1 -\frac{1} {p}\Big) {}\\ & \,\ll \,& \frac{1} {k^{2}\log \log k}\qquad \mbox{ (by (6.2))}\,. {}\\ \end{array}$$

Together with the lower bound (6.1) we complete the proof of the theorem. \(\square\)

Proof of Theorem 2.7

By (4.2) we have already shown the identities in the theorem. So it remains to prove the inequalities. First, we prove the upper bounds.

$$\displaystyle\begin{array}{rcl} & & \sum _{\begin{array}{c} a = 1 \\ a \equiv l\pmod k \end{array} }^{\infty }\sum _{ \begin{array}{c} b = 0 \\ (a,b) = 1 \end{array} }^{a-1} \frac{1} {a(a + b)^{2}} -\frac{3\chi _{k,l}(1)} {8} {}\\ & \,=\,& \sum _{m=0}^{\infty }\sum _{ \begin{array}{c} b = 0 \\ (km + l,b) = 1 \end{array} }^{km+l-1} \frac{1} {(km + l)(km + l + b)^{2}} -\frac{3\chi _{k,l}(1)} {8} {}\\ & \,\leq \,& \sum _{m=0}^{\infty }\sum _{ b=0}^{km+l-1} \frac{1} {(km + l)^{3}} -\frac{3\chi _{k,l}(1)} {8} \, =\, \frac{1} {l^{2}} +\sum _{ m=1}^{\infty } \frac{1} {(km + l)^{2}} -\frac{3\chi _{k,l}(1)} {8}{}\\ && {}\\ & \,\leq \,& \frac{1} {l^{2}} +\sum _{ m=1}^{\infty } \frac{1} {k^{2}m^{2}} -\frac{3\chi _{k,l}(1)} {8} \, =\, \frac{1} {l^{2}} + \frac{\zeta (2)} {k^{2}} -\frac{3\chi _{k,l}(1)} {8} {}\\ & \,=\,& \quad \left \{\begin{array}{lcl} \frac{5} {8} + \frac{\zeta (2)} {k^{2}} \, & \mbox{ if} \,\,l = 1\,,\\ \\ \frac{1} {l^{2}} + \frac{\zeta (2)} {k^{2}} \, & \mbox{ if} \,\, l> 1\,.\end{array} \right. {}\\ \end{array}$$

For the lower bounds we treat the cases \(l = 1\) and \(l> 1\) separately. First, let \(l = 1\). Then

$$\displaystyle\begin{array}{rcl} & & \sum _{\begin{array}{c} a = 1 \\ a \equiv 1\pmod k \end{array} }^{\infty }\sum _{ \begin{array}{c} b = 0 \\ (a,b) = 1 \end{array} }^{a-1} \frac{1} {a(a + b)^{2}} -\frac{3\chi _{k,1}(1)} {8} {}\\ & =& \sum _{m=0}^{\infty }\sum _{ \begin{array}{c} b = 0 \\ (km + 1,b) = 1 \end{array} }^{km} \frac{1} {(km + 1)(km + 1 + b)^{2}} -\frac{3} {8} {}\\ & =& \frac{5} {8} +\sum _{ m=1}^{\infty }\sum _{ \begin{array}{c} b = 0 \\ (km + 1,b) = 1 \end{array} }^{km} \frac{1} {(km + 1)(km + 1 + b)^{2}} {}\\ & \geq & \frac{5} {8} +\sum _{ \begin{array}{c} b = 1 \\ (k + 1,b) = 1 \end{array} }^{k} \frac{1} {(k + 1)(k + 1 + b)^{2}} {}\\ &>& \frac{5} {8} +\sum _{ \begin{array}{c} b = 1 \\ (k + 1,b) = 1 \end{array} }^{k+1} \frac{1} {4(k + 1)^{3}}\, =\, \frac{5} {8} + \frac{\varphi (k + 1)} {4(k + 1)^{3}}\,. {}\\ \end{array}$$

Next, let \(l> 1\). Then

$$\displaystyle\begin{array}{rcl} & & \sum _{\begin{array}{c} a = 1 \\ a \equiv l\pmod k \end{array} }^{\infty }\sum _{ \begin{array}{c} b = 0 \\ (a,b) = 1 \end{array} }^{a-1} \frac{1} {a(a + b)^{2}} -\frac{3\chi _{k,1}(1)} {8} {}\\ & =& \sum _{m=0}^{\infty }\sum _{ \begin{array}{c} b = 1 \\ (km + l,b) = 1 \end{array} }^{km+l-1} \frac{1} {(km + l)(km + l + b)^{2}} {}\\ & \geq & \sum _{\begin{array}{c} b = 1 \\ (l,b) = 1 \end{array} }^{l-1} \frac{1} {l(l + b)^{2}}\,>\,\sum _{ \begin{array}{c} b = 1 \\ (l,b) = 1 \end{array} }^{l} \frac{1} {4l^{3}}\, =\, \frac{\varphi (l)} {4l^{3}}\,.{}\\ \end{array}$$

This completes the proof of the theorem. \(\square\)