\documentclass[12pt,reqno]{article} \usepackage[usenames]{color} \usepackage{amssymb} \usepackage{graphicx} \usepackage{amscd} \usepackage[colorlinks=true, linkcolor=webgreen, filecolor=webbrown, citecolor=webgreen]{hyperref} \definecolor{webgreen}{rgb}{0,.5,0} \definecolor{webbrown}{rgb}{.6,0,0} \usepackage{color} \usepackage{fullpage} \usepackage{float} \usepackage{psfig} \usepackage{graphics,amsmath,amssymb} \usepackage{amsthm} \usepackage{amsfonts} \usepackage{latexsym} \usepackage{epsf} \setlength{\textwidth}{6.5in} \setlength{\oddsidemargin}{.1in} \setlength{\evensidemargin}{.1in} \setlength{\topmargin}{-.1in} \setlength{\textheight}{8.4in} \newcommand{\seqnum}[1]{\href{http://oeis.org/#1}{\underline{#1}}} \begin{document} \begin{center} \epsfxsize=4in \leavevmode\epsffile{logo129.eps} \end{center} \theoremstyle{plain} \newtheorem{theorem}{Theorem} \newtheorem{corollary}[theorem]{Corollary} \newtheorem{lemma}[theorem]{Lemma} \newtheorem{proposition}[theorem]{Proposition} \theoremstyle{definition} \newtheorem{definition}[theorem]{Definition} \newtheorem{example}[theorem]{Example} \newtheorem{conjecture}[theorem]{Conjecture} \theoremstyle{remark} \newtheorem{remark}[theorem]{Remark} \begin{center} \vskip 1cm{\LARGE\bf On the Dirichlet Convolution of Completely \\ \vskip .1in Additive Functions } \vskip 1cm \large Isao Kiuchi and Makoto Minamide \\ Department of Mathematical Sciences\\ Yamaguchi University\\ Yamaguchi 753-8512 \\ Japan\\ \href{mailto:kiuchi@yamaguchi-u.ac.jp}{\tt kiuchi@yamaguchi-u.ac.jp}\\ \href{mailto:minamide@yamaguchi-u.ac.jp}{\tt minamide@yamaguchi-u.ac.jp} \\ \end{center} \vskip .2 in \begin{abstract} Let $k$ and $l$ be non-negative integers. For two completely additive functions $f$ and $g$, we consider various identities for the Dirichlet convolution of the $k$th powers of $f$ and the $l$th powers of $g$. Furthermore, we derive some asymptotic formulas for sums of convolutions on the natural logarithms. \end{abstract} \section{Statements of results} Let $f$ and $g$ be two arithmetical functions that are completely additive. That is, these functions satisfy $f(mn)=f(m)+f(n)$ and $g(mn)=g(m)+g(n)$ for all positive integers $m$ and $n$. We shall consider the arithmetical function \begin{align}\label{def-1} D_{k,l}(n;f,g):=\sum_{d|n}f^{k}(d)g^{l}\left(\frac{n}{d}\right), \end{align} which represents the Dirichlet convolution of the $k$th power of $f$ and the $l$th power of $g$ for non-negative integers $k$ and $l$. The above function provides a certain generalization of the classical number-of-divisors function $d(n)$. In fact, $$ D_{0,0}(n;f,g)=d(n). $$ The first purpose of this study is to investigate some recurrence formulas for $D_{k,l}(n;f,g)$ with respect to $k$ and $l$. Since \begin{align} \label{f-2} \sum_{d|n}f(d)=\frac{1}{2}d(n)f(n), \end{align} where $f$ is a completely additive function, we have \begin{align} \label{def-3} D_{1,1}(n;f,g)=\frac{1}{2}d(n)f(n)g(n)-\sum_{d|n}f(d)g(d). \end{align} Similarly, as in \eqref{def-3}, we use \eqref{def-1} for $D_{k,l+1}(n;f,g)$ to obtain \begin{align*} D_{k,l+1}(n;f,g)&= \sum_{d|n}f^{k}(d)g^{l}\left(\frac{n}{d}\right)g\left(\frac{n}{d}\right) \\ &=g(n)\sum_{d|n}f^{k}(d)g^{l}\left(\frac{n}{d}\right)-\sum_{d|n}f^{k}(d)g(d)g^{l}\left(\frac{n}{d}\right). \end{align*} Hence, we deduce the following two recurrence formulas. \begin{theorem} Let $k$ and $l$ be non-negative integers, and let $f$ and $g$ be completely additive functions. Then we have \begin{align} & D_{k,l+1}(n;f,g)+\sum_{d|n}f^{k}(d)g^{l}\left(\frac{n}{d}\right)g(d)=g(n)D_{k,l}(n;f,g), \label{f-4} \\ & D_{k+1, l}(n;f,g)+\sum_{d|n}f^{k}(d)f\left(\frac{n}{d}\right)g^{l}\left(\frac{n}{d}\right)=f(n)D_{k,l}(n;f,g). \label{f-5} \end{align} \end{theorem} Now, we put $f=g$ in \eqref{f-4} (or \eqref{f-5}), and set $D_{k,l}(n;f):=D_{k,l}(n;f,f)$. Then, we deduce the following corollary. \begin{corollary}\label{coro-ichi} Using the same notation given above, we have \begin{align} D_{k,l+1}(n;f)+D_{k+1,l}(n;f)=f(n)D_{k,l}(n;f). \label{f-6} \end{align} Particularly, if $k=l$, we have \begin{align} D_{k+1, k}(n;f)=D_{k,k+1}(n;f)=\frac{1}{2}f(n)D_{k,k}(n;f). \label{f-7} \end{align} \end{corollary} Because the symmetric property $D_{k,l}(n;f)=D_{l,k}(n;f)$, we only consider the function $D_{k,k+j}(n,f)$ for $j=1,2,\ldots$. \begin{example} The formulas \eqref{f-6} and \eqref{f-7} imply that \begin{align} D_{k,k+2}(n;f)&=\frac{1}{2}f^{2}(n)D_{k,k}(n;f) - D_{k+1,k+1}(n;f), \nonumber \\ D_{k,k+3}(n;f)&=\frac{1}{2}f^{3}(n)D_{k,k}(n;f) - \frac{3}{2}f(n)D_{k+1,k+1}(n;f), \nonumber \\ D_{k,k+4}(n;f)&=\frac{1}{2}f^{4}(n)D_{k,k}(n;f) -2f^{2}(n)D_{k+1,k+1}(n;f) + D_{k+2,k+2}(n;f), \nonumber \\ D_{k,k+5}(n;f)&=\frac{1}{2}f^{5}(n)D_{k,k}(n;f) -\frac{5}{2}f^{3}(n)D_{k+1,k+1}(n;f) + \frac{5}{2}f(n)D_{k+2,k+2}(n;f), \nonumber \\ D_{k,k+6}(n;f)&=\frac{1}{2}f^{6}(n)D_{k,k}(n;f) -3f^{4}(n)D_{k+1,k+1}(n;f) +\frac{9}{2}f^{2}(n)D_{k+2,k+2}(n;f) \nonumber \\ &\quad -D_{k+3,k+3}(n;f), \nonumber \\ D_{k,k+7}(n;f)&=\frac{1}{2}f^{7}(n)D_{k,k}(n;f)- \frac{7}{2}f^{5}(n)D_{k+1,k+1}(n;f) +7f^{3}(n)D_{k+2,k+2}(n;f) \nonumber \\ &\quad -\frac{7}{2}f(n)D_{k+3,k+3}(n;f), \nonumber \\ D_{k,k+8}(n;f)&=\frac{1}{2}f^{8}(n)D_{k,k}(n;f)- 4f^{6}(n)D_{k+1,k+1}(n;f) +10f^{4}(n)D_{k+2,k+2}(n;f) \nonumber \\ &\quad -8f^{2}(n)D_{k+3,k+3}(n;f) +D_{k+4,k+4}(n;f). \nonumber \end{align} \end{example} Next, we shall demonstrate that the explicit evaluation of the function $D_{k,k+m}(n;f)$ ($m=2,3,\ldots$) can be expressed as a combination of the functions $D_{k,k}(n;f)$, $D_{k+1,k+1}(n;f)$, $D_{k+2,k+2}(n;f)$, $\ldots$, $D_{k+\lfloor \frac{m}{2}\rfloor,k+\lfloor \frac{m}{2}\rfloor}(n;f)$. Hence, we shall give a recurrence formula between $D_{k,k}(n;f), \ldots ,$ $D_{k+\lfloor\frac{m}{2}\rfloor,k+\lfloor\frac{m}{2}\rfloor}(n;f)$ and $D_{k,k+m}(n;f)$. \begin{theorem}\label{teiri-3} Let $k$ and $m$ be positive integers, and let $D_{k, k+m}(n;f)$ be the function defined by the above formula. Then we have \begin{align} D_{k,k+m}(n;f) =\sum_{j=0}^{\lfloor \frac{m}{2} \rfloor }c_{k,j}^{(m)}f^{m-2j}(n)D_{k+j,k+j}(n;f), \label{f-16} \end{align} where \begin{align*} c_{k,j}^{(m)}=\begin{cases} \displaystyle \frac{1}{2}, & \text{if $j=0$;}\\ \displaystyle -\frac{m}{2}, & \text{if $j=1$;}\\ (-1)^{j}\displaystyle \frac{m}{2\cdot j!}\displaystyle \prod_{i=1}^{j-1}\left(m-(j+i)\right), & \text{ if $2\leq j \leq \displaystyle \lfloor\frac{m}{2}\rfloor $.} \end{cases} \end{align*} \end{theorem} \begin{proof} By \eqref{f-7} in Corollary \ref{coro-ichi}, the equality \eqref{f-16} holds for $m=1$ and all $k\in\mathbb{N}$. Now, we assume that \eqref{f-16} is true for $m=1,2,\ldots, l$ and $k\in\mathbb{N}$. Using this assumption and \eqref{f-6} in Corollary \ref{coro-ichi}, we observe that \begin{align*} D_{k, k+l+1}(n;f)&=\sum_{j=0}^{\lfloor\frac{l}{2}\rfloor}c_{k,j}^{(l)}f^{l+1-2j}(n)D_{k+j,k+j}(n;f)\\ &\quad -\sum_{j=0}^{\lfloor\frac{l-1}{2}\rfloor}c_{k+1,j}^{(l-1)}f^{l-1-2j}(n)D_{k+1+j,k+1+j}(n;f). \end{align*} For even $l=2q$, we have \begin{align*} D_{k, k+2q+1}(n;f)&=\frac{1}{2}f^{2q+1}(n)D_{k,k}(n;f)\\ & \quad +\sum_{j=1}^{q}\left(c_{k,j}^{(2q)}-c_{k+1,j-1}^{(2q-1)}\right)f^{2q+1-2j}(n) D_{k+j,k+j}(n;f) \end{align*} and \begin{align*} c_{k,j}^{(2q)}-c_{k+1,j-1}^{(2q-1)}=(-1)^{j}\frac{2q+1}{2\cdot j!}\prod_{i=1}^{j-2}\left(2q-(j+i)\right)(2q-j)=c_{k,j}^{(2q+1)}. \end{align*} For odd $l=2q-1$, we observe that \begin{align*} D_{k,k+2q}(n;f) &= \frac{1}{2}f^{2q}(n)D_{k,k}(n;f)\\ &\quad +\sum_{j=1}^{q-1} \left(c_{k,j}^{(2q-1)}-c_{k+1, j-1}^{(2q-2)}\right)f^{2q-2j}(n)D_{k+j}(n;f)\\ & \quad + (-1)^{\lfloor\frac{2q}{2}\rfloor}D_{k+\lfloor \frac{2q}{2}\rfloor, k+\lfloor \frac{2q}{2}\rfloor }(n;f). \end{align*} By our assumption, since \begin{align*} c_{k,j}^{(2q-1)}-c_{k+1,j-1}^{(2q-2)} =(-1)^{j}\frac{2q}{2\cdot j!}\prod_{i=1}^{j-2}\left(2q-1-(j+i)\right) (2q-1-j)=c_{k,j}^{(2q)}, \end{align*} we obtain the assertion \eqref{f-16} for all $k$ and $m \in\mathbb{N}$. \end{proof} Now, we consider another expression for $D_{k,l}(n;f,g)$ using the arithmetical function \begin{align} \label{def-H} H_{k,m}(n;f,g):=\sum_{d|n}f^{k}(d)g^{m}(d). \end{align} If $f=g$, we set $H_{k+m}(n;f)=H_{k,m}(n;f,f)$. The right-hand side of (\ref{def-H}) implies the Dirichlet convolution of $1$ and $f^{k}g^{m}$. Since $g$ is a completely additive function, we have \begin{align*} D_{k,l}(n;f,g)&=\sum_{d|n}f^{k}(d)\left(g(n)-g(d)\right)^{l}\\ &=\sum_{d|n}f^{k}(d)\sum_{m=0}^{l}(-1)^{m}\begin{pmatrix}l\\ m\end{pmatrix}g^{l-m}(n)g^{m}(d). \end{align*} From \eqref{def-H} and the above, we obtain the following theorem. \begin{theorem}\label{teiri-1-4} Let $k$ and $l$ be non-negative integers, and let $f$ and $g$ be completely additive functions. Then we have \begin{align*} D_{k,l}(n;f,g)=\sum_{m=0}^{l}(-1)^{m}{l \choose m}g^{l-m}(n)H_{k,m}(n;f,g), \end{align*} where the function $H_{k,m}(n;f,g)$ is defined by \eqref{def-H}. \end{theorem} We immediately obtain the following corollary. \begin{corollary}Let $k$ and $l$ be non-negative integers, and let $f$ and $g$ be completely additive functions. Then we have \begin{align} D_{k,l}(n;f)=\sum_{m=0}^{l}(-1)^{m}{l \choose m}f^{l-m}(n)H_{k+m}(n;f). \label{f-19} \end{align} \end{corollary} Note that \begin{align} H_{k,m}(n;f,g)&=\sum_{d|n}f^{k}\left(\frac{n}{d}\right)g^{m}\left(\frac{n}{d}\right) \nonumber \\ &=\sum_{i=0}^{k}\sum_{j=0}^{m}(-1)^{i+j} {k \choose i}{m \choose j}f^{k-i}(n)g^{m-j}(n)\sum_{d|n}f^{i}(d)g^{j}(d). \label{f-20} \end{align} Applying \eqref{f-20} to Theorem \ref{teiri-1-4}, we have the following theorem. \begin{theorem} Let $k$ and $l$ be non-negative integers, and let $f$ and $g$ be completely additive functions. Then we have \begin{align*} &D_{k,l}(n;f,g) \nonumber \\ &=\sum_{m=0}^{l}\sum_{i=0}^{k}\sum_{j=0}^{m}(-1)^{m+i+j} {l \choose m}{k \choose i}{m \choose j} f^{k-i}(n)g^{l-j}(n)\sum_{d|n}f^{i}(d)g^{j}(d). \end{align*} \end{theorem} In the case where $f=g$, note that \begin{align*} H_{k+m}(n;f)&=\sum_{d|n}\left(f(n)-f(d)\right)^{k+m}\\ &=\sum_{j=0}^{k+m}(-1)^{j}{k+m \choose j}f^{k+m-j}(n)\sum_{d|n}f^{j}(d). \end{align*} From \eqref{f-19} and the above, we obtain the following corollary. \begin{corollary}\label{coro-17} Let $k$ and $l$ be non-negative integers, and let $f$ be a completely additive function. Then we have \begin{align} D_{k,l}(n;f)=\sum_{m=0}^{l}\sum_{j=0}^{k+m}(-1)^{m+j}{l \choose m}{k+m \choose j}f^{k+l-j}(n)\sum_{d|n}f^{j}(d). \label{f-22} \end{align} \end{corollary} \section{Recurrence formula connecting $D_{k, l}(n;f)$ with $\sum_{d|n}f^{j}(d)$} The second purpose of this study is to derive another expression for $D_{k,l}(n;f)$ that involves the divisor function $d(n)$. Before stating Theorem \ref{th-2-2}, we prepare the following lemma. \begin{lemma}\label{hodai-21} Let $f$ be a completely additive function. There exist the constants $e_{q,q}$, $e_{q,q-1}, \ldots, e_{q,1}$ ($q=1,2,\ldots$) that satisfy the equation \begin{align} \label{f-23} \sum_{d|n}f^{2q-1}(d)=e_{q,q}d(n)f^{2q-1}(n) +\sum_{j=1}^{q-1}e_{q,q-j}f^{2q-2j-1}(n)\sum_{d|n}f^{2j}(d). \end{align} Moreover, the relations among sequences $(e_{q,q-j})_{j=1}^{q}$ are as follows. \begin{align} & e_{q,q}=\frac{1}{2} \left( 1-\sum_{j=1}^{q-1}{2q-1 \choose 2j-1}e_{j,j}\right) =\frac{\left(2^{2q}-1\right)B_{2q}}{q}, \label{f-24}\\ & e_{q,q-j}=\frac{1}{2}\left( {2q-1 \choose 2j} - \sum_{i=2 \atop i-j\geq 1}^{q-1}{2q-1 \choose 2i-1}e_{i,i-j}\right ), \nonumber \end{align} where $B_{n}$ denotes the $n$th Bernoulli number, which is defined by the Taylor expansion \begin{align*} \frac{z}{e^{z}-1}=\sum_{n=1}^{\infty}\frac{B_{n}}{n!}z^{n}, \quad (|z|<2\pi). \end{align*} \end{lemma} \begin{proof} By \eqref{f-2}, the case $q=1$ in \eqref{f-23} is trivial. Assume that there exist $e_{p,p}$, $e_{p,p-1}, \ldots , e_{p,1}$ ($p\leq q$) such that \begin{align} \label{f-26} \sum_{d|n}f^{2p-1}(d)=e_{p,p}d(n)f^{2p-1}(n) +\sum_{j=1}^{p-1}e_{p,p-j}f^{2p-2j-1}(n)\sum_{d|n}f^{2j}(d). \end{align} Since \begin{align*} \sum_{d|n}f^{2q+1}(d) = \sum_{j=0}^{2q+1} (-1)^{j} {2q+1 \choose j} f^{2q+1-j}(n)\sum_{d|n}f^{j}(d), \end{align*} we have \begin{align} \sum_{d|n}f^{2q+1}(d)&=\frac{1}{2}d(n)f^{2q+1}(n) +\frac{1}{2}\sum_{j=1}^{q}{2q+1 \choose 2j}f^{2q+1-2j}(n) \sum_{d|n}f^{2j}(d) \nonumber \\ &\quad - \frac{1}{2}\sum_{j=1}^{q}{2q+1 \choose 2j-1}f^{2q+2-2j}(n) \sum_{d|n}f^{2j-1}(d). \label{f-27} \end{align} Applying \eqref{f-26} to \eqref{f-27}, we obtain \begin{align*} \sum_{d|n}f^{2q+1}(d) =&\frac{1}{2} \left( 1-\sum_{j=1}^{q}{2q+1 \choose 2j-1}e_{j,j}\right ) d(n)f^{2q+1}(n)\\ &+\frac{1}{2}\sum_{j=1}^{q}\left( {2q+1 \choose 2j} -\sum_{i=2 \atop i-j\geq 1}^{q}{2q+1 \choose 2i-1}e_{i,i-j}\right ) f^{2q-2j+1}(n)\sum_{d|n}f^{2j}(d). \end{align*} By induction, this completes the proof, except for the second term on the right-hand side of \eqref{f-24}. The first term on the right-hand side of \eqref{f-24} implies \begin{align*} e_{q,q}=1-\sum_{k=1}^{q}{2q-1 \choose 2k-1} e_{k,k} =1-\sum_{k=1}^{q}{2q \choose 2k}\frac{k}{q}e_{k,k}. \end{align*} Here we put $a(k)=k e_{k,k}$. Then we have \begin{align} a(q)=q-\sum_{k=1}^{q} {2q \choose 2k}a(k). \label{f-28} \end{align} Since $a(1)=e_{1,1}=1/2$ and $\left(2^{2}-1\right)B_{2}=1/2$, we only need to show that $\left(2^{2k}-1\right)B_{2k}$ $(k=1,\ldots , q)$ satisfies the recurrence formula \eqref{f-28}. Consider the $n$th Bernoulli polynomial $B_{n}(x)$, which is defined by the following Taylor expansion: \begin{align*} \frac{ze^{xz}}{e^{z}-1}=\sum_{n=0}^{\infty}\frac{B_{n}(x)}{n!}z^{n} \quad (|z|<2\pi). \end{align*} The following relations are known among $B_{n}(1)$, $B_{n}(1/2)$ and $B_{n}$, \begin{align*} B_{n}(1)=B_{n}, \quad \quad B_{n}\left(\frac{1}{2}\right)=-\left(1-2^{1-n}\right)B_{n}. \end{align*} By the formula \cite[Thm.\ 12.12, p.\ 264]{A} \begin{align} B_{n}(y)=\sum_{k=0}^{n}{n \choose k}B_{k}y^{n-k}, \label{f-30} \end{align} we observe that \begin{align*} B_{2q}(y)=y^{2q}-qy^{2q-1}+\sum_{k=2}^{2q}{2q \choose k}B_{k}y^{2q-k}. \end{align*} In this equation, we consider $y=1$ and $y=1/2$; then \begin{align} B_{2q}&=1-q+\sum_{k=1}^{q} {2q \choose 2k}B_{2k} \label{f-31} \end{align} and \begin{align} 2^{2q}B_{2q}\left(\frac{1}{2}\right) &=\left(2-2^{2q}\right)B_{2q} \nonumber \\ &=1-2q+\sum_{k=1}^{q}{2q \choose 2k}B_{2k}2^{2k}. \label{f-32} \end{align} Subtracting \eqref{f-32} from \eqref{f-31}, we obtain \begin{align*} \left(2^{2q}-1\right)B_{2q} =q-\sum_{k=1}^{q} {2q \choose 2k}\left(2^{2k}-1\right)B_{2k}. \end{align*} This recurrence formula for $\left(2^{2k}-1\right)B_{2k}$'s is equivalent to \eqref{f-28}. This completes the proof of \eqref{f-23}. \end{proof} Applying Lemma \ref{hodai-21} to \eqref{f-22} in Corollary \ref{coro-17}, we have \begin{align} D_{k,l}(n;f)&=\sum_{m=0}^{l}(-1)^{m}{l\choose m}\sum_{j=0}^{\lfloor \frac{k+m}{2}\rfloor }{k+m\choose 2j}f^{l+k-2j}(n)\sum_{d|n}f^{2j}(d) \nonumber \\ &\quad -\sum_{m=0}^{l}(-1)^{m}{l\choose m}\sum_{j=1}^{\lfloor \frac{k+m+1}{2}\rfloor }{k+m\choose 2j-1}f^{l+k+1-2j}(n)\sum_{d|n}f^{2j-1}(d). \label{f-33} \end{align} The second term on the right-hand side of \eqref{f-33} gives us \begin{align} &-\left( \sum_{m=0}^{l}(-1)^{m}{l\choose m}\sum_{j=1}^{\lfloor \frac{k+m}{2}\rfloor }e_{j,j}{k+m\choose 2j-1}\right ) f^{l+k}(n)d(n) \nonumber \\ &-\sum_{m=0}^{l}\sum_{j=1}^{\lfloor \frac{k+m+1}{2}\rfloor }\sum_{i=1}^{j-1}(-1)^{m}{l\choose m}{k+m\choose 2j-1}e_{j,j-i}f^{l+k-2i}(n)\sum_{d|n}f^{2i}(d) \label{f-34} \end{align} using \eqref{f-23}. From \eqref{f-24}, \eqref{f-33} and \eqref{f-34}, we have the following theorem. \begin{theorem}\label{th-2-2} Let $k$ and $l$ be non-negative integers, and let $f$ be a completely additive function. There exist the constants $e_{j,j}$, $e_{j,j-i}$ ($j=1,2,\ldots, \lfloor \frac{k+m+1}{2}\rfloor $, $1\leq j-i 2$ and all $\varepsilon >0$. Here the constants $A_{1}$ and $A_{2}$ are coefficients of the Laurent expansion of the Riemann zeta-function $\zeta(s)$ in the neighbourhood $s=1$: \begin{align*} \zeta(s)=\frac{1}{s-1}+A_{0} +A_{1}(s-1) + A_{2}(s-1)^{2} + A_{3}(s-1)^{3} + \cdots. \end{align*} We use \eqref{f-7}, \eqref{f-43} and Abel's identity \cite[Thm.~4.2, p.\ 77]{A} to obtain \begin{align*} \sum_{n\leq x}D_{2,1}(n;\log) &=\frac{1}{12}x\log^{4}x-\frac{1}{3}x\log^{3}x + (1+A_{1})x\log^{2}x \nonumber \\ &\qquad -2(1+A_{2})x\log x + 2(1+A_{2})x + O_{\varepsilon}\left(x^{\frac13 +\varepsilon}\right). \end{align*} Furthermore, a generalization of \eqref{f-43} for the partial sums of $D_{k,k}(n;\log)$ for positive integers $k$ was considered by the second author \cite[Thm.\ 1.2, \ p.\ 326]{M}, who demonstrated that there exists a polynomial $P_{2k+1}$ of degree $2k+1$ such that \begin{align} \sum_{n\leq x}D_{k,k}(n;\log) = xP_{2k+1}(\log x) + O_{k, \varepsilon}\left(x^{\frac{1}{3}+\varepsilon}\right) \label{f-46} \end{align} for every $\varepsilon>0$. Applying Theorem \ref{teiri-3} and the above formula \eqref{f-46}, we have the following theorem. \begin{theorem} There exists a polynomial $U_{2k+m+1}$ of degree $2k+m+1$ such that \begin{align*} \sum_{n\leq x}D_{k,k+m}(n;\log) =xU_{2k+m+1}(\log x) + O_{k,m,\varepsilon}\left(x^{\frac{1}{3}+\varepsilon}\right) \end{align*} for $m=2,3,\ldots$ and every $\varepsilon>0$. \end{theorem} \section{Acknowledgment} The authors deeply thank the referee for carefully reading this paper and indicating some mistakes. \begin{thebibliography}{9} \bibitem{A} T. M. Apostol. \newblock {\it Introduction to Analytic Number Theory}. \newblock Springer-Verlag, 1976. \bibitem{M} M. Minamide. \newblock The truncated Vorono{\" i} formula for the derivative of the Riemann zeta-function. \newblock {\it Indian J. Math.} {\bf 55} (2013), 325--352. \end{thebibliography} \bigskip \hrule \bigskip \noindent 2010 {\it Mathematics Subject Classification}: Primary 11A25; Secondary 11P99. \noindent \emph{Keywords: } completely additive function, Dirichlet convolution. \bigskip \hrule \bigskip \vspace*{+.1in} \noindent Received April 23 2014; revised versions received May 15 2014; July 9 2014; July 20 2014; July 31 2014. Published in {\it Journal of Integer Sequences}, August 5 2014. \bigskip \hrule \bigskip \noindent Return to \htmladdnormallink{Journal of Integer Sequences home page}{http://www.cs.uwaterloo.ca/journals/JIS/}. \vskip .1in \end{document} .