Hostname: page-component-7bb8b95d7b-wpx69 Total loading time: 0 Render date: 2024-09-28T07:22:25.130Z Has data issue: false hasContentIssue false

ON SUMS INVOLVING THE EULER TOTIENT FUNCTION

Published online by Cambridge University Press:  24 August 2023

ISAO KIUCHI
Affiliation:
Department of Mathematical Sciences, Faculty of Science, Yamaguchi University, Yoshida 1677-1, Yamaguchi 753-8512, Japan e-mail: kiuchi@yamaguchi-u.ac.jp
YUKI TSURUTA*
Affiliation:
Graduate School of Engineering, Oita University, 700 Dannoharu, Oita 870–1192, Japan
Rights & Permissions [Opens in a new window]

Abstract

Let $\gcd (n_{1},\ldots ,n_{k})$ denote the greatest common divisor of positive integers $n_{1},\ldots ,n_{k}$ and let $\phi $ be the Euler totient function. For any real number $x>3$ and any integer $k\geq 2$, we investigate the asymptotic behaviour of $\sum _{n_{1}\ldots n_{k}\leq x}\phi (\gcd (n_{1},\ldots ,n_{k})). $

Type
Research Article
Copyright
© The Author(s), 2023. Published by Cambridge University Press on behalf of Australian Mathematical Publishing Association Inc.

1 Introduction and main results

Let $s=\sigma +it$ be the complex variable and let $\zeta (s)$ denote the Riemann zeta-function. For any positive integer $k\geq 2$ , let $\tau _{k}$ denote the k-factors divisor function defined by ${\textbf {1}*\textbf {1}*\cdots *\textbf {1}}_{}$ and $\tau =\tau _{2}$ . Here $*$ denotes the Dirichlet convolution of arithmetic functions and $\textbf {1}$ is given by $\textbf {1}(n)=1$ for any positive integer n. We define the error term $\Delta _{k}(x)$ in the generalised divisor problem by

(1.1) $$ \begin{align} \sum_{n\leq x}\tau_{k}(n) = Q_{k}(\log x) x + \Delta_{k}(x), \end{align} $$

where $Q_{k}(\log x)={\mathrm {Res}}_{s=1}\,\zeta ^{k}(s){x^{s-1}}/{s}$ is a polynomial in $\log x$ of degree $k-1$ . The order of magnitude of $\Delta _{k}(x)$ as $x\to \infty $ is an open problem called the Piltz divisor problem and it has attracted much interest in analytic number theory. It has been conjectured that

(1.2) $$ \begin{align} \Delta_{k}(x) =O(x^{{(k-1)}/{2k}+\varepsilon}) \end{align} $$

for any integer $k\geq 2$ and any $\varepsilon>0$ (see Ivić [Reference Ivić7, Chapter 13] or Titchmarsh [Reference Titchmarsh14]). Let  $\mu $ denote the Möbius function defined by

$$ \begin{align*} \mu(n)=\left\{\begin{array}{ll} \displaystyle 1 &\ \ \ \mathrm{if} \ n=1, \ \ \mbox{} \\ \displaystyle (-1)^k &\ \ \ \mathrm{if} \ n\ \mathrm{is\ squarefree\ and}\ n=p_1p_2\ldots p_k, \ \ \mbox{} \\ \displaystyle 0 &\ \ \ \mathrm{if} \ n\ \mathrm{is\ not\ squarefree}, \\ \end{array} \right. \end{align*} $$

and let $\gcd (n_{1},\ldots ,n_{k})$ denote the greatest common divisor of the positive integers $n_{1},\ldots ,n_{k}$ for any integer $k\geq 2$ . For a real number $x>3$ , let $S_{f,k}(x)$ denote the summatory function

(1.3) $$ \begin{align} S_{f,k}(x) := \sum_{n_{1}\ldots n_{k}\leq x}f(\gcd(n_{1},\ldots,n_{k})), \end{align} $$

where f is any arithmetic function, by summing over the hyperbolic region $\{(n_{1},\ldots ,n_{k}) \in {\mathbb N}^{k}:\ n_{1}\ldots n_{k}\leq x \}$ . In 2012, Krätzel et al. [Reference Krätzel, Nowak and Tóth10] showed that

$$ \begin{align*} S_{f,k}(x) = \sum_{n\leq x}g_{f,k}(n), \end{align*} $$

where

(1.4) $$ \begin{align} g_{f,k}(n) = \sum_{n=m^{k}l}(\mu*f)(m)\tau_{k}(l) \end{align} $$

(see also Heyman and Tóth [Reference Heyman and Tóth3], Kiuchi and Saad Eddin [Reference Kiuchi and Saad Eddin8]). If f is multiplicative, then (1.4) is multiplicative. We use (1.4) to get the formal Dirichlet series

(1.5) $$ \begin{align} \sum_{n=1}^{\infty}\frac{g_{f,k}(n) }{n^s} = \frac{\zeta^{k}(s)}{\zeta(ks)}\sum_{n=1}^{\infty}\frac{f(n)}{n^{ks}}, \end{align} $$

which converges absolutely in the half-plane $\sigma>\sigma _{0}$ , where $\sigma _{0}$ depends on f and k.

When $f={\mathrm {id}}$ , it follows from (1.5) that

$$ \begin{align*}\sum_{n=1}^{\infty}\frac{g_{{\mathrm{id}},k}(n) }{n^s} = \frac{\zeta^{k}(s)\zeta(ks-1)}{\zeta(ks)} \quad\mbox{for Re } s>1. \end{align*} $$

Here the symbol $\mathrm {id}$ is given by $\mathrm {id}(n)=n$ for any positive integer n. For $k=2$ , Krätzel et al. [Reference Krätzel, Nowak and Tóth10] used the following three methods:

  1. (1) the complex integration approach (see [Reference Ivić7, Reference Titchmarsh14]);

  2. (2) a combination of fractional part sums and the theory of exponent pairs (see [Reference Graham and Kolesnik2, Reference Krätzel9]); and

  3. (3) Huxley’s method (see [Reference Huxley4Reference Huxley and Watt6]),

to prove

(1.6) $$ \begin{align} \sum_{ab\leq x}\gcd(a,b) &= P_{2}(\log x) x + O(x^{\theta}(\log x)^{\theta'}). \end{align} $$

Here $\theta $ satisfies $\tfrac 12 < \theta <1$ , $\theta '$ is some real number and $P_{2}$ is a certain quadratic polynomial with

$$ \begin{align*} P_{2}(\log{x})= \underset{s=1}{\mathrm{Res}}~\frac{\zeta^{2}(s)\zeta(2s-1)}{\zeta(2s)}\frac{x^{s-1}}{s}. \end{align*} $$

They showed that methods (1), (2) and (3) imply the results $\theta =\tfrac 23$ and $\theta '={16}/{9}$ , $\theta ={925}/{1392}$ and $\theta '=0$ , and $\theta ={547}/{832}$ and $\theta '={26947}/{8320}$ , respectively. Let $\phi $ denote the Euler totient function defined by $ \phi ={\mathrm {id}}*\mu. $ The Dirichlet series (1.5) with $f=\phi $ implies that

(1.7) $$ \begin{align} \sum_{n=1}^{\infty}\frac{g_{\phi,k}(n)}{n^{s}} = \frac{\zeta^{k}(s) \zeta(ks-1)}{\zeta^{2}(ks)} \quad\mbox{for Re } s>1. \end{align} $$

We consider some properties of the hyperbolic summation for the Euler totient function involving the gcd. The first purpose of this paper is to investigate the asymptotic behaviour of (1.3) with $f=\phi $ for $k=2$ . Applying fractional part sums and the theory of exponent pairs, we obtain the following result.

Theorem 1.1. For any real number $x>3$ ,

(1.8) $$ \begin{align} &\sum_{ab\leq x}\phi(\gcd(a,b)) = \frac{1}{4\zeta^{2}(2)} x \log^{2} x + \frac{1}{\zeta^{2}(2)}\bigg(2\gamma - \frac12 - 2\frac{\zeta'(2)}{\zeta(2)}\bigg) x \log x \nonumber\\ &\quad + \frac{1}{2\zeta^{2}(2)}\bigg({5}\gamma^{2}+6\gamma_{1}-4\gamma + 1 -4(4\gamma -1)\frac{\zeta'(2)}{\zeta(2)} - 4\frac{\zeta"(2)}{\zeta(2)} +12\bigg(\frac{\zeta'(2)}{\zeta(2)}\bigg)^{2}\bigg) x\nonumber \\ &\quad + O(x^{{55}/{84}+\varepsilon}), \end{align} $$

where $\gamma $ and $\gamma _{1}$ are the Euler constant and the first Stieltjes constant, respectively.

We note that the main term of (1.8) is given by (7.2) below.

Remark 1.2. Note that $\tfrac 12 < {55}/{84}=\tfrac 12+ {13}/{84}<{547}/{832}= \tfrac 12 +{131}/{832}$ .

The summation for the arithmetic functions $h(n)$ in the Dirichlet series

$$ \begin{align*} F_{h}(s)= \sum_{n=1}^{\infty}\frac{h_{}(n)}{n^s} = \zeta^{2}(s)\zeta(2s-1)\zeta^{M}(2s) \end{align*} $$

for $\mathrm {Re}~s>1$ and any fixed integer M was considered by Kühleitner and Nowak [Reference Kühleitner and Nowak11] in 2013. We use their results to show that the error term on the right-hand side of (1.8) is $\Omega (x^{1/2}{\log ^{2}x}/{\log \log x})$ as $x\to \infty $ . This suggests the following conjecture.

Conjecture 1.3. The order of magnitude of the error term on the right-hand side of (1.8) is $ O(x^{1/2}(\log x)^{A}) $ with $A>2$ .

When $k=3$ , Krätzel et al. [Reference Krätzel, Nowak and Tóth10] also derived the formula

$$ \begin{align*} \sum_{abc\leq x}\gcd(a,b,c) &= M_{3}(x) + O(x^{1/2}(\log x)^{5}), \end{align*} $$

where

$$ \begin{align*} M_{3}(x) & = \sum_{s_{0}=1,2/3} \underset{s=s_{0}}{\mathrm{Res}} \bigg(\frac{\zeta^{3}(s)\zeta(3s-1)}{\zeta(3s)}\frac{x^s}{s}\bigg) \\ &= x(0.6842 \ldots \log^{2}x - 0.6620\ldots \log x + 4.845\ldots) - 4.4569\ldots x^{2/3}. \end{align*} $$

For $k=3$ , we derive an asymptotic formula for (1.3) with $f=\phi $ by using the complex integration approach.

Theorem 1.4. For any real number $x>3$ ,

(1.9) $$ \begin{align} \sum_{abc\leq x} & \phi(\gcd(a,b,c)) \nonumber \\ &= \frac{\zeta(2)}{2\zeta^{2}(3)} x \log^{2} x + \frac{\zeta(2)}{\zeta^{2}(3)}\bigg(3\gamma - 1 + 3\frac{\zeta'(2)}{\zeta(2)} - 6\frac{\zeta'(3)}{\zeta(3)} \bigg) x \log x \nonumber \\ &\quad + \frac{\zeta(2)}{\zeta^{2}(3)}\bigg(3\gamma^{2}+3\gamma_{1}-3\gamma + 1 +3(3\gamma -1)\frac{\zeta'(2)}{\zeta(2)} -6(3\gamma -1) \frac{\zeta'(3)}{\zeta(3)} \bigg) x \nonumber \\ &\quad + \frac{\zeta(2)}{\zeta^{2}(3)}\bigg(27\bigg(\frac{\zeta'(3)}{\zeta(3)}\bigg)^{2} +\frac{9}{2}\frac{\zeta"(2)}{\zeta(2)} -9\frac{\zeta"(3)}{\zeta(3)}-18\frac{\zeta'(2)}{\zeta(2)}\frac{\zeta'(3)}{\zeta(3)} \bigg) x \nonumber \\ &\quad + \frac{\zeta^{3}(\frac{2}{3})}{2\zeta^{2}(2)}x^{{2}/{3}} + O(x^{{1}/{2}}\log^{5}x), \end{align} $$

where $\gamma $ and $\gamma _{1}$ are the Euler constant and the first Stieltjes constant, respectively.

We note that the main term of (1.9) is given by (7.3) below.

For $k=4$ , we use the complex integration approach to calculate the asymptotic formula for (1.3) with $f=\phi $ .

Theorem 1.5. For any real number $x>3$ ,

(1.10) $$ \begin{align} \sum_{abcd\leq x}\phi(\gcd(a,b,c,d)) &=x {P}_{\phi,4}(\log x) + O(x^{1/2} \log^{{17}/{3}} x ), \end{align} $$

where ${P}_{\phi ,4}(u)$ is a polynomial in u of degree three depending on $\phi $ .

For $k=5$ , from Ivić [Reference Ivić7, Theorem 13.2], the error term $\Delta _{5}(x)$ is estimated by

(1.11) $$ \begin{align} \Delta_{5}(x) =O(x^{{11}/{20}+\varepsilon}) \end{align} $$

for any $\varepsilon>0$ . We use an elementary method and (1.1) to obtain the following result.

Theorem 1.6. For any real number $x>3$ ,

(1.12) $$ \begin{align} \sum_{abcde\leq x}\phi(\gcd(a,b,c,d,e)) &= x {P}_{\phi,5}(\log x) + \sum_{n\leq x^{1/5}}(\mu*\mu)(n)\sum_{m\leq x^{1/5}/n}m\Delta_{5}\bigg(\frac{x}{m^{5}n^{5}}\bigg), \end{align} $$

where ${P}_{\phi ,5}(u)$ is a polynomial in u of degree four depending on $\phi $ . In particular, it follows from (1.11) that the error term on the right-hand side of (1.12) is $O(x^{{11}/{20}+\varepsilon }).$

Remark 1.7. If we can use Conjecture (1.2) with $k=5$ , then the error term on the right-hand side of (1.12) becomes $O(x^{{2}/{5}+\varepsilon })$ .

Assuming Conjecture (1.2), it is easy to obtain an asymptotic formula for $S_{\phi ,k}(x)$ for any integer $k \geq 5$ .

Proposition 1.8. Assume Conjecture (1.2). With the previous notation,

$$ \begin{align*} \sum_{n_{1}n_{2}\ldots n_{k}\leq x}\phi(\gcd(n_{1},n_{2},\ldots,n_{k})) &= x {P}_{\phi,k}(\log x) + O(x^{{(k-1)}/{2k}+\varepsilon}) \end{align*} $$

for any real number $x>3$ , where ${P}_{\phi ,k}(u)\ (k \geq 5)$ is a polynomial in u of degree $k-1$ depending on $\phi $ .

Notation

We denote by $\varepsilon $ an arbitrary small positive number which may be different at each occurrence.

2 Auxiliary results

We will need the following lemma.

Lemma 2.1. For $t\geq t_{0}>0$ , uniformly in $\sigma $ ,

$$ \begin{align*} \zeta(\sigma+it) &\ll \left\{\begin{array}{ll} t^{(3-4\sigma)/6}\log t & \text{if } 0\leq \sigma \leq 1/2, \\ t^{(1-\sigma)/3} \log t & \text{if } 1/2 \leq \sigma \leq 1, \\ \log t & \text{if } 1\leq \sigma < 2, \\ 1 & \text{if } \sigma \geq 2. \end{array} \right. \end{align*} $$

Proof. The lemma follows from Tenenbaum [Reference Tenenbaum13, Theorem II.3.8]; see also Ivić [Reference Ivić7] or Titchmarsh [Reference Titchmarsh14].

3 Proof of Theorem 1.1

Our main work is to evaluate the sum $ A(x)=\sum _{mnl^{2}\leq x,\, m,n,l>0}l $ . We utilise [Reference Krätzel, Nowak and Tóth10, Section 3.2] to derive the formula

$$ \begin{align*}A(x) = M_{1}(x) + \Delta(x), \end{align*} $$

where the error term $\Delta (x)$ is estimated by $O(x^{1/4+(\alpha+\beta)/2})$ . Here $(\alpha ,\beta )$ is an exponent pair (see [Reference Graham and Kolesnik2, Reference Ivić7]) and $M_{1}(x)$ is the main term given by

$$ \begin{align*}M_{1}(x)=\underset{s=1}{\mathrm{Res}}~\zeta^{2}(s)\zeta(2s-1)\frac{x^{s}}{s}. \end{align*} $$

From (1.7), this gives

(3.1) $$ \begin{align} \sum_{ab\leq x}\phi(\gcd(a,b)) = \sum_{l\leq \sqrt{x}}(\mu*\mu)(l)A\bigg(\frac{x}{l^2}\bigg) = M_{2}(x) + O(x^{1/4+(\alpha+\beta)/2+\varepsilon}), \end{align} $$

where the main term $M_{2}(x)$ is given by

$$ \begin{align*} M_{2}(x) & =\underset{s=1}{\mathrm{Res}}~\frac{\zeta^{2}(s)\zeta(2s-1)}{\zeta^{2}(2s)}\frac{x^{s}}{s} \\ &= \frac{1}{4\zeta^{2}(2)} x \log^{2} x + \frac{1}{\zeta^{2}(2)}\bigg(2\gamma - \frac12 - 2\frac{\zeta'(2)}{\zeta(2)}\bigg) x \log x \\ & \quad + \frac{1}{2\zeta^{2}(2)}\bigg({5}\gamma^{2}+6\gamma_{1}-4\gamma + 1 -4(4\gamma -1)\frac{\zeta'(2)}{\zeta(2)} - 4\frac{\zeta"(2)}{\zeta(2)} +12\bigg(\frac{\zeta'(2)}{\zeta(2)}\bigg)^{2}\bigg) x \end{align*} $$

by (7.2) below. Choosing, in particular, the exponent pair

$$ \begin{align*}(\alpha,\beta)=(\tfrac{13}{84}+\varepsilon,\tfrac{55}{84}+\varepsilon), \end{align*} $$

discovered by Bourgain [Reference Bourgain1, Theorem 6], we obtain the order of magnitude $ O(x^{{55}/{84}+\varepsilon }) $ of the error term on the right-hand side of (3.1). This completes the proof of Theorem 1.1.

4 Preparations for the proof of Theorems 1.4 and 1.5

In order to derive the formulas (1.9) and (1.10), we use the following notation. Let k be any integer such that $k \geq 3$ and let $ \sigma _{0} =1+{1}/{k}+\varepsilon. $ Consider the estimation of the error terms of Perron’s formula (see [Reference Montgomery and Vaughan12, Theorem 5.2 and Corollary 5.3]) for (1.4) with $f=\phi $ . The estimation of $g_{\phi ,k}(n)$ is given by

$$ \begin{align*} g_{\phi,k}(n) = \sum_{n=m^{k}l}({\mathrm{id}} *\mu*\mu)(m) \tau_{k}(l) & = \sum_{n=d^{k}m^{k}l}d (\mu*\mu)(m) \tau_{k}(l) \\ & \ll n^{1/k} \sum_{n=d^{k}m^{k}l} \tau_{}(m)\tau_{k}(l) \ll n^{{1}/{k}+\varepsilon}. \end{align*} $$

In Perron’s formula,

$$ \begin{align*} R \ll x^{{1}/{k} +\varepsilon} \bigg(1+\frac{x}{T}\sum_{1\leq k \leq x}\frac{1}{k}\bigg) + \frac{(4x)^{\sigma_0}}{T} \bigg|\frac{\zeta^{k}(\sigma_{0})\zeta(k\sigma_{0}-1)}{\zeta^{2}(k\sigma_{0})}\bigg| \ll \frac{x^{\sigma_0}}{T} \end{align*} $$

for $T\leq x$ . Hence, from Perron’s formula and (1.7),

$$ \begin{align*}S_{\phi,k}(x) = \frac{1}{2\pi i}\int_{\sigma_0-iT}^{\sigma_0+iT}\frac{\zeta^{k}(s)\zeta(ks-1)}{\zeta^{2}(ks)} \frac{x^s}{s} \,ds + O\bigg(\frac{x^{\sigma_0}}{T}\bigg) \end{align*} $$

for any real number $x>3$ . When $k=3$ , we move the line of integration to $\mathrm {Re}~s=\tfrac 12$ and consider the rectangular contour formed by the line segments joining the points $c_{0}-iT$ , $c_{0}+iT$ , $\tfrac 12+iT$ , $\tfrac 12-iT$ and $c_{0}-iT$ in the anticlockwise sense. We observe that the integrand has a triple pole at $s=1$ and a simple pole at $s=\tfrac 23$ . Thus, we obtain the main term from the sum of the residues coming from the poles at $s=1$ and $\tfrac 23$ . Hence, using the Cauchy residue theorem,

(4.1) $$ \begin{align} S_{\phi,3}(x) &= J_{3}(x,T) + I_{3,1}(x,T) + I_{3,2}(x,T) - I_{3,3}(x,T) + O\bigg(\frac{x^{{4}/{3}+\varepsilon}}{T}\bigg), \end{align} $$

where

(4.2) $$ \begin{align} J_{3}(x,T) &= \Big(\underset{s=1}{\mathrm{Res}} + \underset{s=\frac23}{\mathrm{Res}}\Big)~\frac{\zeta^{3}(s)\zeta(3s-1)}{\zeta^{2}(3s)}\frac{x^{s}}{s}. \end{align} $$

Here the integrals are given by

(4.3) $$ \begin{align} I_{3,1}(x,T) &= \frac{1}{2\pi i}\int_{{1}/{2}+iT}^{4/3+\varepsilon +iT} \frac{\zeta^{3}(s)\zeta(3s-1)}{\zeta^{2}(3s)}\frac{x^{s}}{s} \,ds, \end{align} $$
(4.4) $$ \begin{align} I_{3,2}(x,T) &= \frac{1}{2\pi i}\int_{{1}/{2}-iT}^{{1}/{2}+iT} \frac{\zeta^{3}(s)\zeta(3s-1)}{\zeta^{2}(3s)}\frac{x^{s}}{s}\,ds, \end{align} $$
$$ \begin{align*} I_{3,3}(x,T) &= \frac{1}{2\pi i}\int_{{1}/{2}-iT}^{4/3+\varepsilon -iT} \frac{\zeta^{3}(s)\zeta(3s-1)}{\zeta^{2}(3s)}\frac{x^{s}}{s} \,ds. \end{align*} $$

Similarly,

$$ \begin{align*} S_{\phi,4}(x) &= J_{4}(x,T) + I_{4,1}(x,T) + I_{4,2}(x,T) - I_{4,3}(x,T) + O\bigg(\frac{x^{{5}/{4}+\varepsilon}}{T}\bigg), \end{align*} $$

where

(4.5) $$ \begin{align} J_{4}(x,T) &= \underset{s=1}{\mathrm{Res}}~\frac{\zeta^{4}(s)\zeta(4s-1)}{\zeta^{2}(4s)}\frac{x^{s}}{s}, \end{align} $$
(4.6) $$ \begin{align} I_{4,1}(x,T) &= \frac{1}{2\pi i}\int_{{1}/{2}+a+iT}^{5/4+\varepsilon +iT} \frac{\zeta^{4}(s)\zeta(4s-1)}{\zeta^{2}(4s)}\frac{x^{s}}{s} \,ds, \end{align} $$
$$ \begin{align*} I_{4,2}(x,T) &= \frac{1}{2\pi i} \int_{{1}/{2}+a-iT}^{{1}/{2}+a+iT} \frac{\zeta^{4}(s)\zeta(4s-1)}{\zeta^{2}(4s)}\frac{x^{s}}{s}\,ds, \end{align*} $$
$$ \begin{align*} I_{4,3}(x,T) &= \frac{1}{2\pi i}\int_{{1}/{2}+a-iT}^{5/4+\varepsilon -iT} \frac{\zeta^{4}(s)\zeta(4s-1)}{\zeta^{2}(4s)}\frac{x^{s}}{s} \,ds \end{align*} $$

with $ a={1}/{\log T}$ for any large number $T(>5)$ .

5 Proofs of Theorems 1.4 and 1.5

5.1 Proof of the formula (1.9)

Consider the estimate $I_{3,1}(x,T)$ . We use Lemma 2.1 and (4.3) to deduce the estimation

$$ \begin{align*} I_{3,1}(x;T) &= \frac{1}{2\pi i} \int_{1/2}^{4/3+\varepsilon} \frac{\zeta^{}(\sigma+iT)^3 \zeta(3\sigma-1+3iT)}{\zeta^{}(3\sigma+3iT)^2 (\sigma+iT)} x^{\sigma+iT} \,d\sigma \\ &\ll \frac{1}{T} \bigg(\int_{1/2}^{2/3}+\int_{2/3}^{1} + \int_{1}^{4/3 +\varepsilon}\bigg) |\zeta^{}(\sigma+iT)|^3 |\zeta(3\sigma-1+3iT)| x^{\sigma} \,d\sigma \\ &\ll T^{2/3} \log^{4}T \int_{1/2}^{2/3} \bigg(\frac{x}{T^{2}}\bigg)^{\sigma}\,d\sigma + \log^{4}T \int_{2/3}^{1} \bigg(\frac{x}{T^{}}\bigg)^{\sigma} \,d\sigma + \frac{\log^{4}T}{T} \int_{1}^{4/3+\varepsilon} {x}^{\sigma}\,d\sigma \\ &\ll \frac{x^{{4}/{3}+\varepsilon}}{T^{}}\log^{4}T. \end{align*} $$

Similarly, the estimation of $ I_{3,3}(x,T) $ is of the same order. Hence, taking $T=x^{}$ in the estimations of $I_{3,1}(x,T)$ and $I_{3,3}(x,T)$ , we find that the total contribution of the horizontal lines in absolute value is

(5.1) $$ \begin{align} \ll x^{1/3 +\varepsilon}. \end{align} $$

Now we estimate $I_{3,2}(x,T)$ . We use (4.4), the estimate $ \zeta (\tfrac 32+it) \asymp 1 $ for $t\geq 1$ , the well-known estimate

(5.2) $$ \begin{align} \int_{1}^{T}\frac{|\zeta(\frac12 +iu)|^{4}}{u}\,du \ll \log^{5}T \end{align} $$

for any large T and the Hölder inequality to obtain the estimate

(5.3) $$ \begin{align} I_{3,2}(x,T) & = \frac{1}{2\pi}\int_{-T}^{T}\frac{\zeta^{3}(\frac{1}{2}+it)\zeta(\frac{1}{2}+3it)}{\zeta^{2}(\frac32+3it)} \frac{x^{1/2 +it}}{\frac12 +it} \,dt \nonumber \\ &\ll x^{1/2} + x^{1/2} \int_{1}^{T} \frac{|\zeta(\frac12 +it)|^3}{|\zeta(\frac32 +3it)|^2}\cdot \frac{|\zeta(\frac12+3it)|}{t}\,dt \nonumber \\ &\ll x^{1/2} \bigg(\int_{1}^{T} \frac{|\zeta(\frac12 +it)|^4}{t} \,dt\bigg)^{3/4} \bigg(\int_{1}^{T} \frac{|\zeta(\frac12 +3it)|^4}{3t} \,dt\bigg)^{1/4} \nonumber \\ &\ll x^{1/2} \log^{5}T. \end{align} $$

Taking $T=x^{}$ in (5.1), (5.3) and (4.1) with $k=3$ , and substituting the above and the residue (4.2) into (4.1) with $k=3$ , we obtain the formula (1.9).

5.2 Proof of the formula (1.10)

Let $a={1}/{\log T}\ (T\geq 5)$ . From (4.5) with $k=4$ ,

(5.4) $$ \begin{align} J_{4}(x,T) = x {P}_{\phi,4}(\log x), \end{align} $$

since $s=1$ is a pole of $\zeta ^{4}(s)$ of order four, where $ {P}_{\phi ,4}(u)$ is a polynomial in u of degree three depending on $\phi $ . Consider the estimate $I_{4,1}(x,T)$ . From (4.6) and Lemma 2.1,

$$ \begin{align*} I_{4,1}(x,T) &\ll \frac{1}{T} \bigg(\int_{1/2+a}^{1}+\int_{1}^{5/4 +\varepsilon}\bigg) |\zeta^{}(\sigma+iT)|^4 |\zeta(4\sigma-1+4iT)| x^{\sigma} \,d\sigma \\ &\ll T^{1/3}\log^{5}T \int_{1/2+a}^{1} \bigg(\frac{x}{T^{4/3}}\bigg)^{\sigma}\,d\sigma + \frac{\log^{5}T}{T} \int_{1}^{5/4 +\varepsilon} {x}^{\sigma} \,d\sigma \\ &\ll x^{1/2+a}\frac{\log^{5}T}{T^{1/3}} + x^{5/4 +\varepsilon} \frac{\log^{5}T}{T}. \end{align*} $$

Similarly, the estimation of $ I_{4,3}(x,T) $ is of the same order. Hence, taking $T=x^{}$ in the estimations $I_{4,1}(x,T)$ and $I_{4,3}(x,T)$ , we find that the total contribution of the horizontal lines in absolute value is

(5.5) $$ \begin{align} \ll x^{1/4+\varepsilon}. \end{align} $$

We use (5.2) and the estimation $ \zeta (1+it) \ll {\log ^{2/3} t} $ for $t\geq t_{0}$ (see [Reference Ivić7, Theorem 6.3]) to obtain the estimation

(5.6) $$ \begin{align} I_{4,2}(x,T) \ll x^{1/2+a} + x^{1/2+a}\log^{2/3}T \int_{1}^{T} \frac{|\zeta(\frac12 +it)|^4}{t} \,dt \ll x^{1/2} \log^{{17}/{3}}T. \end{align} $$

We take $T=x^{}$ in (5.4), (5.5), (5.6) and (4.1) with $k=4$ to complete the proof of the formula (1.10).

6 Proof of Theorem 1.6

We use (1.4) with $k=5$ to deduce that

(6.1) $$ \begin{align} \sum_{abcde\leq x}\phi(\gcd(a,b,c,d,e)) =\sum_{lm^{5}n^{5}\leq x}(\mu*\mu)(n)m\tau_{5}(l) =\sum_{n\leq x^{1/5}}(\mu*\mu)(n)B\bigg(\frac{x}{n^5}\bigg), \end{align} $$

where $ B(x):=\sum _{lm^{5}\leq x}m\tau _{5}(l). $ From (1.1),

(6.2) $$ \begin{align} B(x) &= \sum_{m\leq x^{1/5}}m\sum_{n\leq {x}/{m^5}} \tau_{5}(n) \nonumber \\ &= \sum_{m\leq x^{1/5}}m\bigg(A_{1}\frac{x}{m^5}\log^{4}\frac{x}{m^5} + \cdots + A_{5}\frac{x}{m^5} + \Delta_{5}\bigg(\frac{x}{m^5}\bigg) \bigg) \nonumber \\ &= \widetilde{Q_{4}}(\log x) x + \sum_{m\leq x^{1/5}}m\Delta_{5}\bigg(\frac{x}{m^5}\bigg), \end{align} $$

where $\widetilde {{Q}_{4}}(u)$ is a polynomial in u of degree four and $A_{1},A_{2},\ldots , A_{5}$ are computable constants. Inserting (6.2) into (6.1),

$$ \begin{align*} \sum_{abcde\leq x}\phi(\gcd(a,b,c,d,e)) = x {P}_{\phi,5}(\log x) + \sum_{n\leq x^{1/5}}(\mu*\mu)(n)\sum_{m\leq x^{1/5}/n}m\Delta_{5}\bigg(\frac{x}{m^{5}n^{5}}\bigg). \end{align*} $$

Hence, we obtain the formula (1.12).□

7 Appendix

To calculate the main terms of Theorems 1.1 and 1.4, we use the Laurent expansion of the Riemann zeta-function at $s=1$ : that is,

(7.1) $$ \begin{align} \zeta(s) = \frac{1}{s-1} + \gamma + \gamma_{1}(s-1) +\gamma_{2}(s-1)^2 + \gamma_{3}(s-1)^3 + \cdots \end{align} $$

as $s\to \infty $ , where $\gamma $ is the Euler constant and $\gamma _{k}\ (k=1,2,3,\ldots )$ are the Stieltjes constants,

$$ \begin{align*}\gamma_{k}:=\frac{(-1)^k}{k!}\lim_{N\to \infty}\bigg(\sum_{m\leq N}\frac{\log^{k}m}{m}-\frac{\log^{k+1}N}{k+1}\bigg). \end{align*} $$

We need the following residues.

(7.2) $$ \begin{align} M_{2}(x) &:=\underset{s=1}{\mathrm{Res}}~\frac{\zeta^{2}(s)\zeta(2s-1)}{\zeta^{2}(2s)}\frac{x^{s}}{s} \nonumber\\ &= \frac{1}{4\zeta^{2}(2)} x \log^{2} x + \frac{1}{\zeta^{2}(2)}\bigg(2\gamma - \frac12 - 2\frac{\zeta'(2)}{\zeta(2)}\bigg) x \log x \nonumber\\ &\quad + \frac{1}{2\zeta^{2}(2)}\bigg({5}\gamma^{2}+6\gamma_{1}-4\gamma + 1 -4(4\gamma -1)\frac{\zeta'(2)}{\zeta(2)} - 4\frac{\zeta"(2)}{\zeta(2)} +12\bigg(\frac{\zeta'(2)}{\zeta(2)}\bigg)^{2}\bigg) x, \end{align} $$

and

(7.3) $$ \begin{align} J_{3}(x,T) &:= \Big(\underset{s=1}{\mathrm{Res}} + \underset{s=2/3}{\mathrm{Res}}\Big)~\frac{\zeta^{3}(s)\zeta(3s-1)}{\zeta^{2}(3s)}\frac{x^{s}}{s} \nonumber\\ & = \frac{\zeta(2)}{2\zeta^{2}(3)} x \log^{2} x + \frac{\zeta(2)}{\zeta^{2}(3)}\bigg(3\gamma - 1 + 3\frac{\zeta'(2)}{\zeta(2)} - 6\frac{\zeta'(3)}{\zeta(3)} \bigg) x \log x \nonumber \\ &\quad+ \frac{\zeta(2)}{\zeta^{2}(3)}\bigg(3\gamma^{2}+3\gamma_{1}-3\gamma + 1 +3(3\gamma -1)\frac{\zeta'(2)}{\zeta(2)} -6(3\gamma -1) \frac{\zeta'(3)}{\zeta(3)} \bigg) x \nonumber \\ &\quad+ \frac{\zeta(2)}{\zeta^{2}(3)}\bigg(27\bigg(\frac{\zeta'(3)}{\zeta(3)}\bigg)^{2} +\frac{9}{2}\frac{\zeta"(2)}{\zeta(2)} -9\frac{\zeta"(3)}{\zeta(3)}-18\frac{\zeta'(2)}{\zeta(2)}\frac{\zeta'(3)}{\zeta(3)} \bigg) x +\frac{\zeta(\frac{2}{3})^3}{2\zeta^{2}(2)}x^{{2}/{3}}. \end{align} $$

Proof. Suppose that $g(s)$ is regular in the neighbourhood of $s=1$ and $f(s)$ has only a triple pole at $s=1$ . Then the Laurent expansion of $f(s)$ implies that

$$ \begin{align*}f(s) := \frac{a}{(s-1)^3} + \frac{b}{(s-1)^2} + \frac{c}{s-1} + h(s), \end{align*} $$

where $h(s)$ is regular in the neighbourhood of $s=1$ and $a,b,c$ are computable constants. We use the residue calculation to deduce that

$$ \begin{align*} &\underset{s=1}{\mathrm{Res}} f(s)g(s) =\frac{a}{2} g"(1) + b g'(1) + c g(1). \end{align*} $$

To prove (7.3), we use (7.1) to deduce that

$$ \begin{align*} \zeta^{3}(s)&= \frac{1}{(s-1)^3} + \frac{3\gamma}{(s-1)^2} + \frac{3\gamma^{2}+3\gamma_{1}}{s-1} + O(1) \quad\mbox{as } s\to 1. \end{align*} $$

Setting

$$ \begin{align*}g(s):=\frac{\zeta(3s-1)}{\zeta^{2}(3s)}\cdot \frac{x^s}{s}, \end{align*} $$

we have

$$ \begin{align*}g(1) =\frac{\zeta(2)}{\zeta^{2}(3)}x, \quad g'(1) =\frac{\zeta(2)}{\zeta^{2}(3)}\bigg(\log x + 3\frac{\zeta'(2)}{\zeta(2)} -6\frac{\zeta'(3)}{\zeta(3)}-1\bigg)x, \end{align*} $$
$$ \begin{align*} g"(1)&= \frac{\zeta(2)}{\zeta^{2}(3)}x \log^{2} x + \frac{2\zeta(2)}{\zeta^{2}(3)}\bigg(3\frac{\zeta'(2)}{\zeta(2)} - 6\frac{\zeta'(3)}{\zeta(3)}-1 \bigg)x\log x \\ &\quad+ \frac{2\zeta(2)}{\zeta^{2}(3)}\bigg(1+6\frac{\zeta'(3)}{\zeta(3)} - 3\frac{\zeta'(2)}{\zeta(2)} + 27\bigg(\frac{\zeta'(3)}{\zeta(3)}\bigg)^2 \bigg)x \\ &\quad+ \frac{2\zeta(2)}{\zeta^{2}(3)}\bigg(\frac{9}{2}\frac{\zeta"(2)}{\zeta(2)} -9\frac{\zeta"(3)}{\zeta(3)} - 18\frac{\zeta'(2)}{\zeta(2)} \frac{\zeta'(3)}{\zeta(3)} \bigg)x. \end{align*} $$

Hence,

$$ \begin{align*} &\Big(\underset{s=1}{\mathrm{Res}} + \underset{s=2/3}{\mathrm{Res}}\Big)~\frac{\zeta^{3}(s)\zeta(3s-1)}{\zeta^{2}(3s)}\frac{x^{s}}{s} \\ &\quad=\frac{1}{2}g"(1) +3\gamma g'(1) + 3(\gamma_{1}+\gamma^2)g(1) + \frac{\zeta^{3}(\frac{2}{3})}{2\zeta^{2}(2)}x^{2/3}. \end{align*} $$

Hence, we obtain the stated identity. To prove (7.2), we use

$$ \begin{align*} \zeta^{2}(s)\zeta(2s-1)&= \frac{\frac12}{(s-1)^3} + \frac{2\gamma}{(s-1)^2} + \frac{\frac52 \gamma^{2}+3\gamma_{1}}{s-1} + O(1) \quad\mbox{as } s\to 1. \end{align*} $$

The proof of (7.2) is similar to that of (7.3).

Acknowledgement

The authors would like to thank the referee for many valuable suggestions and remarks.

Footnotes

The first author is supported by JSPS Grant-in-Aid for Scientific Research (C)(21K03205).

References

Bourgain, J., ‘Decoupling, exponential sums and the Riemann zeta-function’, J. Amer. Math. Soc. 30 (2017), 205224.CrossRefGoogle Scholar
Graham, S. W. and Kolesnik, G., Van der Corput’s Method of Exponential Sums, London Mathematical Society Lecture Note Series, 126 (Cambridge University Press, Cambridge, 1991).CrossRefGoogle Scholar
Heyman, R. and Tóth, L., ‘Hyperbolic summation for functions of the GCD and LCM of several integers’, Ramanujan J., to appear. Published online (28 December 2022).CrossRefGoogle Scholar
Huxley, M. N., ‘Exponential sums and lattice points. III’, Proc. Lond. Math. Soc. (3) 87 (2003), 591609.CrossRefGoogle Scholar
Huxley, M. N., ‘Exponential sums and the Riemann zeta-function. V’, Proc. Lond. Math. Soc. (3) 90 (2005), 140.CrossRefGoogle Scholar
Huxley, M. N. and Watt, N., ‘Exponential sums and the Riemann zeta-function’, Proc. Lond. Math. Soc. (3) 57 (1988), 124.CrossRefGoogle Scholar
Ivić, A., The Riemann Zeta-Function: Theory and Applications (Dover, Mineola, NY, 1985).Google Scholar
Kiuchi, I. and Saad Eddin, S., ‘On sums of arithmetic functions involving the greatest common divisor’, Preprint, 2021, arxiv:2102.03714v1.Google Scholar
Krätzel, E., Lattice Points, Mathematics and its Applications. East European Series, 33 (Kluwer, Dordrecht, 1988).Google Scholar
Krätzel, E., Nowak, W. G. and Tóth, L., ‘On certain arithmetic functions involving the greatest common divisor’, Cent. Eur. J. Math. 10 (2012), 761774.CrossRefGoogle Scholar
Kühleitner, M. and Nowak, W. G., ‘On a question of A. Schinzel: omega estimate for a special type of arithmetic functions’, Cent. Eur. J. Math. 11 (2013), 477486.Google Scholar
Montgomery, H. L. and Vaughan, R. C., Multiplicative Number Theory I. Classical Theory, Cambridge Studies in Advanced Mathematics, 97 (Cambridge University Press, Cambridge, 2007).Google Scholar
Tenenbaum, G., Introduction to Analytic and Probabilistic Number Theory, Graduate Studies in Mathematics, 163 (American Mathematical Society, Providence, RI, 2008).Google Scholar
Titchmarsh, E. C., The Theory of the Riemann Zeta-function, revised by D. R. Heath-Brown (Oxford University Press, Oxford, 1986).Google Scholar