Hostname: page-component-8448b6f56d-m8qmq Total loading time: 0 Render date: 2024-04-23T09:47:18.065Z Has data issue: false hasContentIssue false

Mean square values of L-functions over subgroups for nonprimitive characters, Dedekind sums and bounds on relative class numbers

Published online by Cambridge University Press:  30 January 2023

Stéphane R. Louboutin
Affiliation:
Centrale Marseille, Aix Marseille Université, CNRS, I2M, Marseille, France e-mail: stephane.louboutin@univ-amu.fr
Marc Munsch*
Affiliation:
Dipartimento di Matematica, Unige, Genova, Italia
Rights & Permissions [Opens in a new window]

Abstract

An explicit formula forthe mean value of $\vert L(1,\chi )\vert ^2$ is known, where $\chi $ runs over all odd primitive Dirichlet characters of prime conductors p. Bounds on the relative class number of the cyclotomic field ${\mathbb Q}(\zeta _p)$ follow. Lately, the authors obtained that the mean value of $\vert L(1,\chi )\vert ^2$ is asymptotic to $\pi ^2/6$, where $\chi $ runs over all odd primitive Dirichlet characters of prime conductors $p\equiv 1\ \ \pmod {2d}$ which are trivial on a subgroup H of odd order d of the multiplicative group $({\mathbb Z}/p{\mathbb Z})^*$, provided that $d\ll \frac {\log p}{\log \log p}$. Bounds on the relative class number of the subfield of degree $\frac {p-1}{2d}$ of the cyclotomic field ${\mathbb Q}(\zeta _p)$ follow. Here, for a given integer $d_0>1$, we consider the same questions for the nonprimitive odd Dirichlet characters $\chi '$ modulo $d_0p$ induced by the odd primitive characters $\chi $ modulo p. We obtain new estimates for Dedekind sums and deduce that the mean value of $\vert L(1,\chi ')\vert ^2$ is asymptotic to $\frac {\pi ^2}{6}\prod _{q\mid d_0}\left (1-\frac {1}{q^2}\right )$, where $\chi $ runs over all odd primitive Dirichlet characters of prime conductors p which are trivial on a subgroup H of odd order $d\ll \frac {\log p}{\log \log p}$. As a consequence, we improve the previous bounds on the relative class number of the subfield of degree $\frac {p-1}{2d}$ of the cyclotomic field ${\mathbb Q}(\zeta _p)$. Moreover, we give a method to obtain explicit formulas and use Mersenne primes to show that our restriction on d is essentially sharp.

Type
Article
Creative Commons
Creative Common License - CCCreative Common License - BY
This is an Open Access article, distributed under the terms of the Creative Commons Attribution licence (https://creativecommons.org/licenses/by/4.0/), which permits unrestricted re-use, distribution, and reproduction in any medium, provided the original work is properly cited.
Copyright
© The Author(s), 2023. Published by Cambridge University Press on behalf of The Canadian Mathematical Society

1 Introduction

Let $X_f$ be the multiplicative group of the $\phi (f)$ Dirichlet characters modulo $f>2$ . Let $X_f^- =\{\chi \in X_f;\ \chi (-1)=-1\}$ be the set of the $\phi (f)/2$ odd Dirichlet characters modulo f. Let $L(s,\chi )$ be the Dirichlet L-function associated with $\chi \in X_f$ . Let H denote a subgroup of index m in the multiplicative group $G:=({\mathbb Z}/f{\mathbb Z})^*$ . We assume that $-1\not \in H$ . Hence, m is even. We set $X_f(H) =\{\chi \in X_f;\ \chi _{/H}=1\}$ , a subgroup of order m of $X_f$ isomorphic to the group of Dirichlet characters of the abelian quotient group $G/H$ of order m. Define $X_f^-(H)=\{\chi \in X_f^-;\ \chi _{/H}=1\}$ , a set of cardinal $m/2$ . Let K be an abelian number field of degree m and prime conductor $p\geq 3$ , i.e., let K be a subfield of the cyclotomic number field ${\mathbb Q}(\zeta _p)$ (Kronecker–Weber’s theorem). The Galois group $\mathrm {Gal}({\mathbb Q}(\zeta _p)/{\mathbb Q})$ is canonically isomorphic to the multiplicative cyclic group $({\mathbb Z}/p{\mathbb Z})^*$ and $H :=\mathrm {Gal}({\mathbb Q}(\zeta _p)/K)$ is a subgroup of $({\mathbb Z}/p{\mathbb Z})^*$ of index m and order

$$ \begin{align*}d=(p-1)/m.\end{align*} $$

Now, assume that K is imaginary. Then d is odd, m is even, $-1\not \in H$ and the set

$$ \begin{align*}X_K^- :=X_p^-(H) :=\{\chi\in X_p^-;\mbox{ and } \chi_{/H}=1\}\end{align*} $$

is of cardinal $(p-1)/(2d) =m/2$ . Let $K^+$ be the maximal real subfield of K of degree $m/2$ fixed by the complex conjugation. The class number $h_{K^+}$ of $K^+$ divides the class number $h_K$ of K. The relative class number of K is defined by $h_K^- =h_K/h_{K^+}$ . We refer the reader to [Reference SerreSer, Was] for such basic knowledge. The mean square value of $L(1,\chi )$ as $\chi $ ranges in $X_f^-(H)$ is defined by

(1.1) $$ \begin{align} M(f,H) :={1\over\# X_f^-(H)}\sum_{\chi\in X_f^-(H)}\vert L(1,\chi)\vert^2. \end{align} $$

The analytic class number formula and the arithmetic–geometric mean inequality give

(1.2) $$ \begin{align} h_K^- =w_K\left (\frac{p}{4\pi^2}\right )^{m/4}\prod_{\chi\in X_K^-} L(1,\chi) \leq w_K\left ({pM(p,H)\over 4\pi^2}\right )^{m/4}, \end{align} $$

where $w_K$ is the number of complex roots of unity in K. Hence, $w_K=2p$ for $K ={\mathbb Q}(\zeta _p)$ and $w_K=2$ otherwise. In [Reference Louboutin and MunschLM21, Theorem 1.1], we proved that

(1.3) $$ \begin{align} M(p,H) =\frac{\pi^2}{6}+o(1) \end{align} $$

as p tends to infinity uniformly over subgroups H of $({\mathbb Z}/p{\mathbb Z})^*$ of odd order $d\leq \frac {\log p}{3(\log \log p)}$ Footnote 1 . Hence, by (1.2), we have

(1.4) $$ \begin{align} h_{K}^{-} \leq w_K\left(\frac{(1+o(1))p}{24}\right)^{(p-1)/4d}. \end{align} $$

In some situations, it is even possible to give an explicit formula for $M(p,H)$ implying a completely explicit bound for $h_K^-$ . Indeed, by [Reference WalumMet, Wal] (see also (4.2)), we have

(1.5) $$ \begin{align} M(p,\{1\}) ={\pi^2\over 6}\left (1-{1\over p}\right )\left (1-{2\over p}\right ) \leq\frac{\pi^2}{6} \ \ \ \ \ (p\geq 3). \end{align} $$

Hence,

(1.6) $$ \begin{align} h_{{\mathbb Q}(\zeta_p)}^- \leq 2p\left ({pM(p,\{1\})\over 4\pi^2}\right )^{(p-1)/4} \leq 2p\left ({p\over 24}\right )^{(p-1)/4}. \end{align} $$

We refer the reader to [Reference GranvilleGra] for more information about the expected size of $h_{{\mathbb Q}(\zeta _p)}^-$ . The only other situation where a similar explicit result is known is the following one (see Theorem 6.6 for a new proof).

Theorem (SeeFootnote 2 [Reference LouboutinLou16, Theorem 1])

Let $p\equiv 1\ \ \pmod 6$ be a prime integer. Let K be the imaginary subfield of degree $(p-1)/3$ of the cyclotomic number field ${\mathbb Q}(\zeta _p)$ . Let H be the subgroup of order $3$ of the multiplicative group $({\mathbb Z}/p{\mathbb Z})^*$ . We have (compare with ( 1.5 ) and ( 1.6 ))

(1.7) $$ \begin{align} M(p,H) ={\pi^2\over 6}\left (1-{1\over p}\right ) \leq\frac{\pi^2}{6} \mbox{ and } h_K^-\leq 2\left (\frac{p}{24}\right )^{(p-1)/12}.\end{align} $$

In [Reference LouboutinLou94] (see also [Reference LouboutinLou11]), the following simple argument allowed to improve on (1.6). Let $d_0>1$ be a given integer. Assume that $\gcd (d_0,f)=1$ . For $\chi $ modulo f, let $\chi '$ be the character modulo $d_0f$ induced by $\chi $ . Then

(1.8) $$ \begin{align} L(1,\chi) =L(1,\chi')\prod_{q\mid d_0}\left (1-\frac{\chi(q)}{q}\right )^{-1} \end{align} $$

(throughout the paper, this notation means that q runs over the distinct prime divisors of $d_0$ ). Let H be a subgroup of order d of the multiplicative group $({\mathbb Z}/f{\mathbb Z})^*$ , with $-1\not \in H$ . We define

(1.9) $$ \begin{align} M_{d_0}(f,H) :={1\over\# X_f^-(H)}\sum_{\chi\in X_f^-(H)}\vert L(1,\chi')\vert^2 \end{align} $$

andFootnote 3

(1.10) $$ \begin{align} \Pi_{d_0}(f,H) :=\prod_{q\mid d_0}\prod_{\chi\in X_f^-(H)}\left (1-\frac{\chi(q)}{q}\right ) \text{ and } D_{d_0}(f,H) :=\Pi_{d_0}(f,H)^{4/m}. \end{align} $$

Clearly, there is no restriction in assuming from now on that $d_0$ is square- free. Let now H be of odd order d in the multiplicative group $({\mathbb Z}/p{\mathbb Z})^*$ . Using (1.8), we obtain (compare with (1.2))

(1.11) $$ \begin{align} h_K^- =\frac{w_K}{\Pi_{d_0}(p,H)} \left (\frac{p}{4\pi^2}\right )^{m/4}\prod_{\chi\in X_K^-}L(1,\chi') \leq w_K\left (\frac{pM_{d_0}(p,H)}{4\pi^2D_{d_0}(p,H)}\right )^{m/4}. \end{align} $$

Let $d=o(\log p)$ as $p\rightarrow \infty $ . Then, by Corollary 2.4, we have

$$ \begin{align*}D_{d_0}(p,H) =1+o(1)\end{align*} $$

and we expect that

(1.12) $$ \begin{align} M_{d_0}(p,H) \sim\left\{\prod_{q\mid d_0}\left (1-\frac{1}{q^2}\right )\right\} \times M(p,H). \end{align} $$

Hence, (1.11) should indeed improve on (1.2). The aim of this paper is twofold. First, in Theorem 1.1, we give an asymptotic formula for $M_{d_0}(p,H)$ when d satisfies the same restriction as in (1.3) allowing us to improve on the bound (1.4). Second, we treat the case of groups of orders $1$ and $3$ for small $d_0$ ’s as well as the case of Mersenne primes and groups of size $\approx \log p$ . In both cases, an explicit description of these subgroups allows us to obtain explicit formulas for $M_{d_0}(p,H)$ . Our main result is the following.

Theorem 1.1 Let $d_0\geq 1$ be a given square-free integer. As $p\rightarrow +\infty $ , we have the following asymptotic formula:

$$ \begin{align*}M_{d_0}(p,H) =\frac{\pi^2}{6}\prod_{q\mid d_0}\left (1-{1\over q^2}\right ) +O(d(\log p)^2 p^{-\frac{1}{d-1}}) =\frac{\pi^2}{6}\prod_{q\mid d_0}\left (1-{1\over q^2}\right ) +o(1)\end{align*} $$

uniformly over subgroups H of $({\mathbb Z}/p{\mathbb Z})^*$ of odd order $d\leq \frac {\log p}{3(\log \log p)}.$ Moreover, let K be an imaginary abelian number field of prime conductor p and of degree $m=(p-1)/d$ . Let $C<4\pi ^2=39.478..$ be any positive constant. If p is sufficiently large and $m\geq 3\frac {(p-1) \log \log p} {\log p}$ , then we have

(1.13) $$ \begin{align} h_K^- \leq w_K\left (\frac{p}{C}\right )^{(p-1)/4d}. \end{align} $$

Remark 1.2 The second result in Theorem 1.1 improves on (1.4), (1.6), and (1.7). It follows from the first result in Theorem 1.1, and by using (1.11) and (2.2), where we take $d_0$ as the product of sufficiently many consecutive first primes.

The special case $d_0=1$ was proved in [Reference Louboutin and MunschLM21, Theorem $1.1$ ]. Note that the restriction on d cannot be extended further to the range $d=O(\log p)$ as shown by Theorem 5.2. Moreover, the constant C in (1.13) cannot be taken larger than $4\pi ^2$ (see the discussion about Kummer’s conjecture in [Reference Murty and PetridisMP01].

In the first part of the paper, the presentation goes as follows:

  • In Section 2, we explain the condition about the prime divisors of $d_0$ and prove that $D_{d_0}(p,H) =1+o(1)$ .

  • In Section 3, we review some results on Dedekind sums and prove a new bound of independent interest for Dedekind sums $s(h,f)$ with h being of small order modulo f (see Theorem 3.1). To do so, we use techniques from uniform distribution and discrepancy theory. Then we relate $M_{d_0}(p,H)$ to twisted moments of L- functions which we further express in terms of Dedekind sums. For the sake of clarity, we first treat separately the case $H=\{1\}$ . Note that we found that this case is related to elementary sums of maxima that we could not estimate directly (see Section 3.4). Using our estimates on Dedekind sums, we deduce the asymptotic formula of Theorem 1.1 and the related class number bounds.

In the second part of the paper, we focus on the explicit aspects. Let us describe briefly our presentation:

  • In Section 4.1, we establish a formula for $M_{d_0}(f,\{1\})$ , $d_0>2$ , provided that all the prime factors q of f satisfy $q\equiv \pm 1\ \ \pmod {d_0}$ . In particular, we get formulas for $M_{d_0}(f,\{1\})$ for $d_0\in \{1,2,3,6\}$ and $\gcd (d_0,f)=1$ (such formulae become harder to come by as $d_0$ gets larger). For example, for $p\geq 5$ and $d_0=6$ , using Theorem 4.1, we obtain the following formula for $M_6(p,\{1\})$ :

    $$ \begin{align*}M_6(p,\{1\}) ={\pi^2\over 9}\left (1-\frac{c_p}{p}\right ) \leq\frac{\pi^2}{9}, \mbox{ where } c_p =\begin{cases} 1,&\mbox{if } p\equiv 1\ \ \pmod 3,\\ 0,&\mbox{if } p\equiv 2\ \ \pmod 3,\\ \end{cases}\end{align*} $$
    which by (1.11) and Corollary 2.4 give improvements on (1.6) (see also [Reference FengFeng])
    $$ \begin{align*}h_{{\mathbb Q}(\zeta_p)}^- \leq 3p\left ({p\over 36}\right )^{(p-1)/4}.\end{align*} $$
    See also [Reference LouboutinLou23, Theorem 5.2] for even better bounds. In Section 4.3, we obtain an explicit formula of the form
    (1.14) $$ \begin{align} M_{d_0}(p,H) =\frac{\pi^2}{6} \left\{\prod_{q\mid d_0}\left (1-\frac{1}{q^2}\right )\right\} \left (1+\frac{N_{d_0}(p,H)}{p}\right ){\kern-1.5pt}, \end{align} $$
    where $N_{d_0}(p,H)$ defined in (4.5) is an explicit average of Dedekind sums. In Proposition 4.6, we prove that $N_{d_0}(p,\{1\})\in {\mathbb Q}$ depends only on p modulo $d_0$ and is easily computable.
  • For $H\neq \{1\}$ explicit formulae for $M_{d_0}(p,H)$ seem difficult to come by. In Section 5, we focus on Mersenne primes $p=2^d-1$ , with d odd. We take $H=\{2^k;\ 0\leq k\leq d-1\}$ , a subgroup of odd order d of the multiplicative group $({\mathbb Z}/p{\mathbb Z})^*$ . For $d_0\in \{1,3,15\}$ , we prove in Theorem 5.4 that

    $$ \begin{align*}M_{d_0}(p,H) =\frac{\pi^2}{2}\left\{\prod_{q\mid d_0}\left (1-\frac{1}{q^2}\right )\right\} \left (1+\frac{N_{d_0}'(p,H)}{p}\right ){\kern-1.5pt},\end{align*} $$
    where $N_{d_0}'(p,H)=a_1(p)d+a_0(p)$ with $a_1(p),a_0(p)\in {\mathbb Q}$ depending only on $p=2^d-1$ modulo $d_0$ and easily computable. In the range $d \gg \log p$ , we see that $M_{d_0}(p,H)$ has a different asymptotic behavior than the one in Theorem 1.1.
  • In Section 6, we turn to the specific case of subgroups of order $3$ . Writing $f=a^2+ab+b^2$ not necessarily prime, and taking $H=\{1,a/b,b/a\}$ , the subgroup of order $3$ of the multiplicative group $({\mathbb Z}/f{\mathbb Z})^*$ , we prove in Proposition 6.4 that $N_{d_0}(f,H) =O(\sqrt f)$ in (1.14) for $d_0\in \{1,2,3,6\}$ . To do so, we obtain bounds for the Dedekind sums stronger than the one in Theorem 3.1. Note that this cannot be expected in general for subgroups of order $3$ modulo composite f (see Remarks 3.4 and 6.2). Furthermore, we show that these bounds are sharp in the case of primes $p=a^2+a+1$ , in accordance with Conjecture 7.1.

2 Preliminaries

2.1 Algebraic considerations

Take $a\in {\mathbb Z}$ with $\gcd (a,f)=1$ . There are infinitely many prime integers in the arithmetic progressions $a+f{\mathbb Z}$ . Taking a prime $p\in a+f{\mathbb Z}$ with $p>d_0f$ , we have $s_{d_0}(p)=a$ , where $s_{d_0}:({\mathbb Z}/d_0f{\mathbb Z})^*\longrightarrow ({\mathbb Z}/f{\mathbb Z})^*$ is the canonical morphism. Therefore, $s_{d_0}$ surjective and its kernel is of order $\phi (d_0)$ . Let H be a subgroup of $({\mathbb Z}/f{\mathbb Z})^*$ of order d. Then $H_{d_0} =s_{d_0}^{-1}(H)$ is a subgroup of order $\phi (d_0)d$ of $({\mathbb Z}/d_0f{\mathbb Z})^*$ and as $\chi $ runs over $X_f^-(H)$ the $\chi '$ ’s run over $X_{d_0f}^-(H_{d_0}),$ and by (1.1) and (1.9), we have

(2.1) $$ \begin{align} M_{d_0}(f,H) =M(d_0f,H_{d_0}). \end{align} $$

The following Lemma is probably well known but we found no reference in the literature.

Lemma 2.1 Let $f>2$ . Let H be a subgroup of index $m =(G:H)$ in the multiplicative group $G:=({\mathbb Z}/f{\mathbb Z})^*$ . Then $\# X_f(H) =m$ and $H=\cap _{\chi \in X_f(H)}\ker \chi $ . Moreover, if $-1\not \in H$ , then m is even, $\# X_f^-(H)=m/2$ and $H=\cap _{\chi \in X_f^-(H)}\ker \chi $ .

Proof Since $X_f(H)$ is isomorphic to the group of Dirichlet characters of the abelian quotient group $G/H$ , it is of order m, by [Reference SerreSer, Chapter VI, Proposition 2]. Clearly, $H\subseteq \cap _{\chi \in X_f(H)}\ker \chi $ . Conversely, take $g\not \in H$ , of order $n\geq 2$ in the abelian quotient group $G/H$ . Define a character $\chi $ of the subgroup $\langle g,H\rangle $ of G generated by g and H by $\chi (g^kh) =\exp (2\pi ik/n)$ , $(k,h)\in {\mathbb Z}\times H$ . It extends to a character of G still denoted $\chi $ , by [Reference SerreSer, Chapter VI, Proposition 1]. Since $g\not \in \ker \chi $ and $\chi \in X_f(H)$ , we have $g\not \in \cap _{\chi \in X_f(H)}\ker \chi $ , i.e., $ \cap _{\chi \in X_f(H)}\ker \chi \subseteq H$ .

Now, assume that $-1\not \in H$ . Set $H'=\langle -1,H\rangle $ , of index $m/2$ in G. Then $X_f^-(H) =X_f(H)\setminus X_f(H')$ is indeed of order $m-m/2 =m/2$ , by the first assertion. Clearly, $H\subseteq \cap _{\chi \in X_f^-(H)}\ker \chi $ . Conversely, take $g\not \in H$ . Set $H" :=\langle g,H\rangle =\{g^kh;\ k\in {\mathbb Z},\ h\in H\}$ , of index $m"$ in G, with $m>m"$ . If $-1=g^kh\in H"$ , then clearly $\chi (g)\neq 1$ for $\chi \in X_f^-(H)$ ; hence, $g\not \in \cap _{\chi \in X_f^-(H)}\ker \chi $ . If $-1\not \in H"$ and $\chi \in X_f^-(H)\setminus X_f^-(H")$ , a nonempty set or cardinal $m/2-m"/2 =(H":H)/2\geq 1$ , then clearly $\chi (g)\neq 1$ ; hence, $g\not \in \cap _{\chi \in X_f^-(H)}\ker \chi $ . Therefore, $\cap _{\chi \in X_f^-(H)}\ker \chi \subseteq H$ .

Remark 2.2 We have $M_{d_0}(p,H)/D_{d_0}(p,H) =M_{d_0/q}(p,H)/D_{d_0/q}(p,H)$ whenever a prime q dividing $d_0$ is in $\cap _{\chi \in X_p^-(H)}\ker \chi $ . Hence, by Lemma 2.1, when applying (1.11) we may assume that no prime divisor of $d_0$ is in H.

2.2 On the size of $\Pi _{d_0}(f,H)$ and $D_{d_0}(f,H)$ defined in (1.10)

Lemma 2.3 Let H be a subgroup of order $d\geq 1$ of the multiplicative group $({\mathbb Z}/f{\mathbb Z})^*$ , where $f>2$ . Assume that $-1\not \in H$ . Let g be the order of a given prime integer q in the multiplicative quotient group $({\mathbb Z}/f{\mathbb Z})^*/H$ . Let $X_f(H)$ be the multiplicative group of the $\phi (f)/d$ Dirichlet characters modulo f for which $\chi _{/H} =1$ . Define $X_f^-(H) = \{\chi \in X_f(H);\ \chi (-1)=-1\}$ , a set of cardinal $\phi (f)/(2d)$ . Then

$$ \begin{align*}\Pi_{q}(f,H) :=\prod_{\chi\in X_f^-(H)}\left (1-\frac{\chi (q)}{q}\right ) =\begin{cases} \left (1+\frac{1}{q^{g/2}}\right )^{\frac{\phi(f)}{dg}},&\mbox{if } g \mbox{ is even and } -q^{g/2}\in H,\\ \left (1-\frac{1}{q^g}\right )^{\frac{\phi(f)}{2dg}},&\mbox{otherwise.} \end{cases}\end{align*} $$

Proof Let $\alpha $ be of order g in an abelian group A of order n. Let $B=\langle \alpha \rangle $ be the cyclic group generated by $\alpha $ . Let $\hat B$ be the group of the g characters of B. Then $P_{B}(X) :=\prod _{\chi \in \hat B} (X-\chi (\alpha )) =X^g-1$ . Now, the restriction map $\chi \in \hat A\rightarrow \chi _{/B}\in \hat B$ is surjective, by [Reference SerreSer, Proposition 1], and of kernel isomorphic to $\widehat {A/B}$ of order $n/g$ , by [Reference SerreSer, Proposition 2]. Therefore, $P_{A}(X) :=\prod _{\chi \in \hat A} (X-\chi (\alpha )) =P_B(X)^{n/g} =(X^g-1)^{n/g}$ . With $A=({\mathbb Z}/f{\mathbb Z})^*/H$ of order $n=\phi (f)/d$ , we have $\hat A =X_f(H)$ and

$$ \begin{align*}\prod_{\chi\in X_f(H)}(X-\chi (q)) =(X^g-1)^{\frac{\phi(f)}{dg}}.\end{align*} $$

Let $H'$ be the subgroup of order $2d$ generated by $-1$ and H. With $A'=({\mathbb Z}/f{\mathbb Z})^*/H'$ of order $n'=\phi (f)/(2d)$ , we have $\hat A' =X_f(H') =X_f^+(H):=\{\chi \in X_f(H);\ \chi (-1)=+1\}$ and

$$ \begin{align*}\prod_{\chi\in X_f^+(H)}(X-\chi (q)) =(X^{g'}-1)^{\frac{\phi(f)}{2dg'}}, \end{align*} $$

where q is of order $g'$ in $A'$ .

Since $X_f^-(H) =X_f(H)\setminus X_f^+(H)$ , it follows that

$$ \begin{align*}\prod_{\chi\in X_f^-(H)} (X-\chi (q)) =\frac{(X^g-1)^{\frac{\phi(f)}{dg}}}{(X^{g'}-1)^{\frac{\phi(f)}{2dg'}}}.\end{align*} $$

Since $q^g\in H$ , we have $q^g\in H'$ and $g'$ divides g. Since $q^{g'}\in H'=\{\pm h;\ h\in H\}$ , we have $q^{2g'}\in H$ and g divides $2g'$ . Hence, $g=g'$ or $g =2g'$ and $g=2g'$ if and only if g is even and $q^{g/2}=q^{g'}\in H'\setminus H=\{-h;\ h\in H\}$ . The assertion follows.

Corollary 2.4 Fix $d_0>1$ square-free. Let $p\geq 3$ run over the prime integers that do not divide $d_0$ . Let H a subgroup of odd order d of the multiplicative group $({\mathbb Z}/p{\mathbb Z})^*$ . Then

(2.2) $$ \begin{align} D_{d_0}(p,H)\ =1+O(\omega(d_0)p^{-1/2(d-1)}), \end{align} $$

where $\omega (d_0)$ stands for the number of prime divisors of $d_0$ . In particular when $d =o(\log p)$ , we have

(2.3) $$ \begin{align} D_{d_0}(p,H)=1+o(1). \end{align} $$

Moreover,

$$ \begin{align*}\Pi _{d_0}(p,\{1\}) \geq\exp\left (\frac{\log d_0}{2}F(p+1)\right ), \mbox{ where } F(x):=\frac{(x-2)\log\left (1-\frac{1}{x}\right )}{\log x},\ (x>1).\end{align*} $$

In particular, $\Pi _{6}(p,\{1\}) \geq 2/3 \,\, \mbox {for } p\geq 5.$

Proof Let q be a prime divisor of $d_0$ . Let g be the order of q in the multiplicative quotient group $({\mathbb Z}/p{\mathbb Z})^*/H$ . Then

$$ \begin{align*}\left (1-\frac{1}{q^g}\right )^{\frac{2}{g}} \leq D_q(p,H) =\Pi_{q}(p,H)^{\frac{4d}{p-1}} \leq\left (1+\frac{1}{q^{g/2}}\right )^{\frac{4}{g}},\end{align*} $$

by (1.10) and Lemma 2.3, with $f=p$ , $\phi (f)=p-1$ and $m =(p-1)/d$ . Either $q^g\equiv 1\ \ \pmod p$ , in which case $q^g\geq p+1$ , or $q^g\equiv h\ \ \pmod p$ for some $h\in \{2,\ldots ,p-1\}\cap H$ , in which case p divides $S:=1+h+\cdots +h^{d-1}$ which satisfies $p\leq S\leq 2h^{d-1}$ . Therefore, in both cases, we have $q^{g}\geq (p/2)^{\frac {1}{d-1}}$ . Hence,

$$ \begin{align*}\log D_q(p,H) \geq \frac{2}{g} \log(1-q^{-g}) \geq \frac{2}{g}(-2\log 2)q^{-g} \geq -4(\log 2)(p/2)^{-1/(d-1)},\end{align*} $$

where we used for $x=q^{-g}$ the fact that $\log (1-x) \geq -2(\log 2)x$ in $\left [0,1/2\right ]$ ,

$$ \begin{align*}D_q(p,H) \geq 1-4(\log 2)(p/2)^{-1/(d-1)},\end{align*} $$

where we used the fact that $e^{-x} \geq 1-x$ . Therefore, we have

$$ \begin{align*}D_{d_0}(p,H) =\prod_{q \mid d_0}D_q(p,H) \geq 1-4(\log 2)\omega(d_0)\left(\frac{p}{2}\right)^{-1/(d-1)},\end{align*} $$

where we used the inequality $(1-x)^n \geq 1-nx$ for $x \leq 1$ and $n\in \mathbb {N}$ . A similar reasoning gives an explicit upper bound $D_{d_0}(p,H) \leq 1+ c\omega (d_0)p^{-1/2(d-1)}$ for some constant $c>0$ . Therefore, we do get (2.2). Finally, $p^{1/(d-1)}$ tends to infinity in the range $d =o(\log p)$ and (2.3) follows.

Notice that if $p=2^d-1$ runs over the Mersenne primes and $H=\langle 2\rangle $ , we have $d=O(\log p)$ but $D_2(p,H) =\left (1-\frac {1}{2}\right )^2$ does not satisfy (2.3).

Now, assume that $H=\{1\}$ . Then $K={\mathbb Q}(\zeta _p)$ and $q^g\geq p+1$ . Hence,

$$ \begin{align*}\Pi_q(p,\{1\}) \geq\left (1-\frac{1}{p+1}\right )^{\frac{p-1}{2g}} \geq\left (1-\frac{1}{p+1}\right )^{\frac{(p-1)\log q}{2\log (p+1)}} =\exp\left (\frac{\log q}{2}F(p+1)\right ){\kern-1.5pt}.\end{align*} $$

The desired lower bound easily follows.

3 Dedekind sums and mean square values of L-functions

3.1 Dedekind sums and Dedekind–Rademacher sums

The Dedekind sums is the rational number defined by

(3.1) $$ \begin{align} s(c,d) ={1\over 4d}\sum_{n=1}^{\vert d\vert -1}\cot\left ({\pi n\over d}\right )\cot\left ({\pi nc\over d}\right ) \ \ \ \ \ (c\in {\mathbb Z},\ d\in {\mathbb Z}\setminus\{0\},\ \gcd (c,d)=1){\kern-1.5pt}, \end{align} $$

with the convention $s(c,-1)=s(c,1)=0$ for $c\in {\mathbb Z}$ (see [Reference ApostolApo] or [Reference Rademacher and GrosswaldRG] where it is, however, assumed that $d>1$ ). It depends only on c mod $\vert d\vert $ and $c\mapsto s(c,d)$ can therefore be seen as a mapping from $({\mathbb Z}/\vert d\vert {\mathbb Z})^*$ to ${\mathbb Q}$ . Notice that

(3.2) $$ \begin{align} s(c^*,d)=s(c,d) \text{ whenever } cc^*\equiv 1\ \ \pmod d \end{align} $$

(make the change of variables $n\mapsto nc$ in $s(c^*,d)$ ). Recall the reciprocity law for Dedekind sums

(3.3) $$ \begin{align} s(c,d)+s(d,c) ={c^2+d^2-3\vert cd\vert +1\over 12cd}, \ \ \ \ \ (c,d\in {\mathbb Z}\setminus\{0\},\ \gcd(c,d)=1). \end{align} $$

In particular,

(3.4) $$ \begin{align} s(1,d) =\frac{d^2-3\vert d\vert +2}{12d} \text{ and } s(2,d) =\frac{d^2-6\vert d\vert +5}{24d} \ \ \ \ \ (d\in {\mathbb Z}\setminus\{0\}). \end{align} $$

For $b,c\in {\mathbb Z},\ d\in {\mathbb Z}\setminus \{-1,0,1\}$ such that $\gcd (b,d) =\gcd (c,d)=1$ , the Dedekind–Rademacher sum is the rational number defined by

$$ \begin{align*} s(b,c,d) ={1\over 4d}\sum_{n=1}^{\vert d\vert -1}\cot\left ({\pi nb\over d}\right )\cot\left ({\pi nc\over d}\right ){\kern-1.5pt},\end{align*} $$

with the convention $s(b,c,-1)=s(b,c,1)=0$ for $b,c\in {\mathbb Z}$ . Hence, $s(c,d) =s(1,c,d)$ , if $\alpha \in ({\mathbb Z}/\vert d\vert {\mathbb Z})^*$ is represented as $\alpha =b/c$ with $\gcd (b,d)=\gcd (c,d) =1$ , then $s(\alpha ,d) =s(b,c,d)$ , and

(3.5) $$ \begin{align} s(b,c,d) =s(ab,ac,d) \mbox{ for any } a\in {\mathbb Z} \mbox{ with } \gcd (a,d)=1. \end{align} $$

For $\gcd ( b,c) =\gcd (c,d) =\gcd (d,b) =1$ , we have a reciprocity law for Dedekind–Rademacher sums (see [Reference RademacherRad] or [Reference Bayad and RaoujBR])

(3.6) $$ \begin{align} s(b,c,d)+s(d,b,c)+s(c,d,b) ={b^2+c^2+d^2-3\vert bcd\vert\over 12bcd}. \end{align} $$

The Cauchy–Schwarz inequality and (3.4) yield

(3.7) $$ \begin{align} \vert s(c,d)\vert \leq s(1,\vert d\vert) \leq\vert d\vert/12 \mbox { and } \vert s(b,c,d)\vert \leq s(1,\vert d\vert) \leq\vert d\vert/12. \end{align} $$

3.2 Nontrivial bounds on Dedekind sums

In this section, we will use the alternative definition of the Dedekind sums given by

$$ \begin{align*}s(c,d) =\sum_{a=1}^{d-1} \left(\left(\frac{a}{d}\right)\right) \left(\left(\frac{ac}{d}\right)\right) \ \ \ \ \ (c\in {\mathbb Z},\ d\geq 1,\ \gcd (c,d)=1),\end{align*} $$

where $ \left (\left (\right )\right ):\mathbb {R} \rightarrow \mathbb {R}$ stands for the sawtooth function defined by

$$ \begin{align*}\left(\left(x\right)\right) := \begin{cases} x-\lfloor x\rfloor - 1/2,& \text{ if } x \in \mathbb{R} \backslash \mathbb{Z},\\ 0, &\text{ if } x \in \mathbb{Z}. \end{cases} \end{align*} $$

In order to prove Theorem 1.1, we need general bounds on Dedekind sums depending on the multiplicative order of the argument. This is a new type of bounds for Dedekind sums and the following result that improves upon (3.7) when the order is $o\left (\frac {\log p}{\log \log p}\right )$ might be of independent interest (see also Conjecture 7.1 for further discussions).

Theorem 3.1 Let $p>1$ be a prime integer and assume that h has order $k\geq 3$ in the multiplicative group $({\mathbb Z}/p{\mathbb Z})^*$ We have

$$ \begin{align*}\vert s(h,p)\vert \ll (\log p)^2 p^{1-\frac{1}{\phi(k)}}.\end{align*} $$

Remark 3.2 Let us notice that by a result of Vardi [Reference VardiVar], for any function f such that $\lim _{n\to +\infty } f(n)=+\infty $ we have $s(c,d) \ll f(d)\log d$ for almost all $(c,d)$ with $\gcd (c,d)=1$ . However, Dedekind sums take also very large values (see, for instance, [Reference Conrey, Fransen and KleinCEK, Reference GirstmairGir03] for more information).

Our proof builds from ideas of the proof of [Reference Louboutin and MunschLM21, Theorem $4.1$ ] where some tools from equidistribution theory and the theory of pseudo-random generators were used. We refer for more information to [Reference KorobovKor, Nied77], or the book of Konyagin and Shparlinski [Reference Konyagin and ShparlinskiKS, Chapter $12$ ] (see [Reference Louboutin and MunschLM21, Section $4$ ] for more details and references). Let us recall some notations. For any fixed integer s, we consider the s-dimensional cube $I_s=\left [0,1\right ]^s$ equipped with its s-dimensional Lebesgue measure $\lambda _s$ . We denote by $\mathcal {B}$ the set of rectangular boxes of the form

$$ \begin{align*}\prod_{i=1}^{s}[\alpha_i,\beta_i) =\left\{x\in I_s, \alpha_i\leq x_i <\beta_i\right\}{\kern-1.5pt},\end{align*} $$

where $0\leq \alpha _i<\beta _i\leq 1.$ If S is a finite subset of $I^s$ , we define the discrepancy $D(S)$ by

$$ \begin{align*}D(S) =\sup_{B \in \mathcal{B}}\left\vert \frac{\# (B\cap S)}{\# S}-\lambda_s(B)\right\vert{\kern-1.5pt}.\end{align*} $$

Let us introduce the following set of points:

$$ \begin{align*}S_{h,p} =\left\{\left(\frac{x}{p},\frac{x h}{p}\right) \in I_2, x \bmod p\right\}{\kern-1.5pt}.\end{align*} $$

For good choice of h, the points are equidistributed and we expect for “nice” functions f

$$ \begin{align*}\lim_{p\to \infty} \frac{1}{p}\sum_{x \bmod p} f\left(\frac{x}{p},\frac{hx}{p}\right) = \int_{I_2} f(x,y)dx dy.\end{align*} $$

Lemma 3.3 For any h of order $k\geq 3$ , we have the following discrepancy bound:

$$ \begin{align*}D(S_{h,p}) \leq (\log p)^{2}p^{-1/\phi(k)}.\end{align*} $$

Proof It follows from the proof of [Reference Louboutin and MunschLM21, Theorem 4.1] where the bound was obtained as a consequence of Erdős–Turan inequality and tools from pseudo-random generators theory.

3.2.1 Proof of Theorem 3.1

Observe that

$$ \begin{align*}s(h,p)= \sum_{x \bmod p} f\left(\frac{x}{p},\frac{hx}{p}\right){\kern-1.5pt},\end{align*} $$

where $f(x,y)= ((x))((y))$ . By Koksma–Hlawka inequality [Reference Drmota and TichyDT, Theorem $1.14$ ], we have

$$ \begin{align*}\left\vert \frac{1}{p} \sum_{ x \bmod p} f\left(\frac{x}{p},\frac{x h}{p}\right) -\int_{I_{2}} f(u,v)dudv \right\vert \leq V(f)D(S_{h,p}),\end{align*} $$

where $V(f)$ is the Hardy–Krause variation of f. Moreover, we have

$$ \begin{align*}\int_{I_{2}} f(u,v)dudv =0.\end{align*} $$

The readers can easily convince themselves that $V(f) \ll 1$ . Hence, the result follows from Lemma 3.3.

Remark 3.4 The same method used to bound the discrepancy leads to a similar bound for composite f. Indeed, for $h \in ({\mathbb Z}/f{\mathbb Z})^*$ of order $k\geq 3$ , we have $s(h,f)= O\left ((\log f)^2 f/E(f)\right )$ with $E(f)=\max \{P^{+}(f)^{1/\phi (k^*)},\textrm {rad}(f)^{1/k}\}$ where $P^{+}(f)$ is the largest prime factor of f, $k^*$ is the order of h modulo $P^+(f)$ and $\displaystyle {\mathrm{rad}(f)=\prod _{\ell \mid f \atop \ell \textrm {prime}}\ell }$ is the radical of f. If $f=h^3-1$ is square-free, then we have $E(f) = f^{1/3}$ and $s(h,f)=O\left ((\log f)^2 f^{2/3}\right )$ which is close to the truth by a logarithmic factor (see Remark 6.2).

For $\gcd (b,p)=\gcd (c,p)=1$ , we recall the other definition of Dedekind–Rademacher sums

$$ \begin{align*}s(b,c,p)= \sum_{a=1}^{p-1} \left(\left(\frac{ab}{p}\right)\right) \left(\left(\frac{ac}{p}\right)\right){\kern-1.5pt}.\end{align*} $$

A similar argument as in the proof of Theorem 3.1 leads to a bound on these generalized sums.

Theorem 3.5 Let $q_1$ , $q_2$ and $k\geq 3$ be given natural integers. Let p run over the primes and h over the elements of order k in the multiplicative group $({\mathbb Z}/p{\mathbb Z})^*$ . Then, we have

$$ \begin{align*}\vert s(q_1,q_2 h,p)\vert \ll (\log p)^2 p^{1-\frac{1}{\phi(k)}}.\end{align*} $$

Proof The proof follows exactly the same lines as the proof of Theorem 3.1 except for the fact that the function f is replaced by the function $g(x,y)= ((q_1 x))((q_2 y))$ . Hence, we have

$$ \begin{align*}s(q_1,q_2 h,p)= g\left(\frac{x}{p},\frac{hx}{p}\right)\end{align*} $$

and by symmetry we remark that

$$ \begin{align*}\int_{I_{2}} g(u,v)dudv =0.\end{align*} $$

Again, $V(g) \ll 1$ and the result follows from Lemma 3.3 and Koksma–Hlawka inequality.

3.3 Twisted second moment of L- functions and Dedekind sums

We illustrate the link between Dedekind sums and twisted moments of L- functions by first proving Theorem 1.1 in the case $H=\{1\}$ with a stronger error term. For any integers $q_1,q_2 \geq 1$ and any prime $p\geq 3$ , we define the twisted moment

(3.8) $$ \begin{align} M_{q_1,q_2}(p) :=\frac{2}{\phi(p)} \sum_{\chi \in X_p^-}\chi(q_1)\overline{\chi}(q_2)\vert L(1,\chi)\vert^2. \end{align} $$

The following formula (see [Reference LouboutinLou94, Proposition 1]) will help us to relate L- functions to Dedekind sums:

(3.9) $$ \begin{align} L(1,\chi) =\frac{\pi}{2f}\sum_{a=1}^{f-1}\chi (a)\cot\left (\frac{\pi a}{f}\right ) \ \ \ \ \ (\chi\in X_f^-).\end{align} $$

Theorem 3.6 Let $q_1$ and $q_2$ be given coprime integers. Then, when p goes to infinity

$$ \begin{align*}M_{q_1,q_2}(p)= \frac{\pi^2}{6q_1q_2}+O_{q_1,q_2}(1/p).\end{align*} $$

Remark 3.7 It is worth to notice that in the case $q_2=1$ , explicit formulas are known by [Reference LouboutinLou15, Theorem $4$ ] (see also [Reference Lee and LeeLee17]). This also gives a new and simpler proof of [Reference Lee and LeeLee19, Theorem $1.1$ ] in a special case.

Proof Let us define

$$ \begin{align*}\epsilon(a,b) :=\frac{2}{\phi(p)}\sum_{\chi \in X_p^-}\chi(a)\overline{\chi}(b) =\begin{cases} 1,&\text{ if } p\nmid ab \text{ and } a=b \bmod p, \\ -1,&\text{ if } p\nmid ab \text{ and } a=-b \bmod p, \\ 0, &\text{ otherwise.} \end{cases}\end{align*} $$

For p large enough, we have $\gcd (q_1,p)=\gcd (q_2,p)=1$ . Hence, using orthogonality relations and (3.9), we arrive at

$$ \begin{align*} M_{q_1,q_2}(p) &=\frac{\pi^2}{4p^2} \sum_{a=1}^{p-1} \sum_{b=1}^{p-1}\epsilon(q_1 a,q_2 b)\cot\left(\frac{\pi a}{p}\right)\cot\left(\frac{\pi b}{p}\right) \\ & = \frac{\pi^2}{2p^2} \sum_{a=1}^{p-1}\cot\left(\frac{\pi q_1a}{p}\right)\cot\left(\frac{\pi q_2 a}{p}\right) = \frac{2\pi^2}{p}s(q_1,q_2,p). \end{align*} $$

When $q_1$ and $q_2$ are fixed coprime integers and p goes to infinity, we infer from (3.6) and (3.7) that

$$ \begin{align*}s(q_1,q_2,p)= \frac{p}{12q_1 q_2}+O(1).\end{align*} $$

The result follows immediately.

Corollary 3.8 Let $q_1$ and $q_2$ be given natural integers. Then, when p goes to infinity

$$ \begin{align*}M_{q_1,q_2}(p )=\frac{\pi^2}{6}\frac{ \gcd(q_1,q_2)^2}{q_1q_2}+O_{q_1,q_2}(1/p).\end{align*} $$

Proof Let $\delta = \gcd (q_1,q_2)$ . We clearly have $M_{q_1,q_2}(p)=M_{q_1/\delta ,q_2/\delta }(p)$ and the result follows from Theorem 3.6.

The proof of Theorem 1.1 in the case of the trivial subgroup follows easily.

Corollary 3.9 Let $d_0$ be a given square-free integer. When p goes to infinity, we have the following asymptotic formula:

$$ \begin{align*}M_{d_0}(p,\{1\}) =\frac{\pi^2}{6}\prod_{q \mid d_0} \left(1-\frac{1}{q^2}\right)+ O(1/p).\end{align*} $$

Proof For $\chi $ modulo p, let $\chi '$ be the character modulo $d_0p$ induced by $\chi $ . By (1.8) and Corollary 3.8, we have

$$ \begin{align*} M_{d_0}(p,\{1\})& =\frac{2}{\# X_p^-} \sum_{\chi \in X_p^-} |L(1,\chi')|^2 =\sum_{\delta_1 \mid d_0} \sum_{\delta_2 \mid d_0} \frac{\mu(\delta_1)}{\delta_1}\frac{\mu(\delta_2)}{\delta_2} M_{\delta_1,\delta_2}(p) \\ &=\frac{\pi^2}{6}\sum_{\delta_1 \mid d_0} \sum_{\delta_2 \mid d_0} \frac{\mu(\delta_1)}{\delta_1^2} \frac{\mu(\delta_2)}{\delta_2^2}\gcd(\delta_1,\delta_2)^2 + O(1/p) \\ &=\frac{\pi^2}{6} \prod_{q \mid d_0} \left(1-\frac{1}{q^2}\right)+O(1/p). \end{align*} $$

3.4 An interesting link with sums of maxima

Before turning to the general case of Theorem 1.1, we explain how to use Theorem 3.6 to estimate the seemingly innocuous sumFootnote 4 defined for any integers $q_1,q_2 \geq 1$ by

$$ \begin{align*}\text{Ma}_{q_1,q_2,p}:=\sum_{x \bmod p}\max(q_1 x,q_2 x),\end{align*} $$

where, here and below, $q_1 x, q_2 x$ denote the representatives modulo p taken in $\left [1,p\right ]$ .

Theorem 3.10 Let $q_1$ and $q_2$ be natural integers such that $q_1 \neq q_2$ . Then, we have the following asymptotic formula:

$$ \begin{align*}\text{Ma}_{q_1,q_2,p} = p^2\left(\frac{2}{3}-\frac{\gcd(q_1,q_2)^2}{12q_1q_2}\right)(1+o(1)).\end{align*} $$

Remark 3.11 In the special case $q_1=1$ , we are able to evaluate the sum directly without the need of Dedekind sums and L- functions. However, we could not prove Theorem 3.10 in the general case using elementary counting methods.

Remark 3.12 Let us notice that $\int _{0}^{1}\int _{0}^{1} \max (x,y) dx dy = 2/3$ . Hence, using the same method as in Section 3.2, we can show that if the points $\left (\left \{\frac {x}{p}\right \},\left \{\frac {qx}{p}\right \}\right )$ are equidistributed in the square $[0,1]^2,$ then

$$ \begin{align*}\sum_{x \bmod p} \max(x,qx) \sim \frac{2}{3}p^2.\end{align*} $$

For q fixed and $p \rightarrow +\infty $ , the points are not equidistributed in the square and we see that the correcting factor $\frac {\gcd (q_1,q_2)^2}{12q_1q_2}$ from equidistribution is related to the Dedekind sum $s(q_1,q_2,p)$ .

We need the following result of [Reference Louboutin and MunschLM21, Theorem $2.1$ ].

Proposition 3.13 Let $\chi $ be a primitive Dirichlet character modulo $f>2$ , its conductor. Set $\displaystyle {S(k,\chi ) =\sum _{l=0}^k\chi (l)}$ . Then

$$ \begin{align*}\sum_{k=1}^{f-1}\vert S(k,\chi)\vert^2 =\frac{f^2}{12}\prod_{p\mid f}\left (1-\frac{1}{p^2}\right ) +a_\chi\frac{f^2}{\pi^2}\vert L(1,\chi)\vert^2, \mbox { where } a_\chi :=\begin{cases} 0,&\mbox{if } \chi (-1)=+1,\\ 1,&\mbox{if } \chi (-1)=-1. \end{cases}\end{align*} $$

3.4.1 Proof of Theorem 3.10

We follow a strategy similar to the proof of [Reference Louboutin and MunschLM21, Corollary $2.2$ ]. We denote by $\chi _0$ the trivial character. Using Proposition 3.13 and recalling the definition (3.8), we arrive at

$$ \begin{align*}\sum_{\chi \in X_p \backslash \chi_0}\chi(q_1)\overline{\chi}(q_2)\sum_{k=1}^{p-1}\vert S(k,\chi)\vert^2 = \sum_{\chi \in X_p\backslash \chi_0}\chi(q_1)\overline{\chi}(q_2)\frac{p^2-1}{12} +\frac{p^3}{2\pi^2}M_{q_1,q_2}(p).\end{align*} $$

Adding the contribution of the trivial character

$$ \begin{align*}\chi_0(q_1)\overline{\chi_0}(q_2)\sum_{k=1}^{p-1}\left\vert\sum_{l=1}^k 1\right\vert{}^2 =\sum_{k=1}^{p-1} k^2 =\frac{(p-1)p(2p-1)}{6},\end{align*} $$

we obtain

(3.10) $$ \begin{align} \sum_{\chi \in X_p}\chi(q_1)\overline{\chi}(q_2)\sum_{k=1}^{p-1}\vert S(k,\chi)\vert^2 =&\sum_{\chi \in X_p}\chi(q_1)\overline{\chi}(q_2)\frac{p^2-1}{12} + \frac{(p-1)p(2p-1)}{6} \nonumber \\ &+\frac{p^3}{2\pi^2}M_{q_1,q_2}(p) +O(p^2). \end{align} $$

For sufficiently large p, using the fact that $q_1 \neq q_2 \bmod p$ and the orthogonality relations, we have

$$ \begin{align*}\sum_{\chi \in X_p}\chi(q_1)\overline{\chi}(q_2)\frac{p^2-1}{12}=0.\end{align*} $$

We now follow the method used in the proof of [Reference Louboutin and MunschLM21, Theorem $4.1$ ] (see also [Reference ElmaElma]) with some needed changes to treat the left-hand side of (3.10). Again by orthogonality, we obtain

$$ \begin{align*} \sum_{\chi \in X_p}\chi(q_1)\overline{\chi}(q_2)\sum_{k=1}^{p-1}\vert S(k,\chi)\vert^2 =\sum_{\chi \in X_p}\chi(q_1)\overline{\chi}(q_2)\sum_{k=1}^{p-1}\left\vert\sum_{l=1}^k\chi (l)\right\vert{}^2 \\ =\sum_{\chi \in X_p}\sum_{k=1}^{p-1}\sum_{1\leq l_1,l_2\leq k} \chi(q_1 l_1)\overline{\chi(q_2 l_2)} =(p-1)^2{\mathcal {A}}(q_1,q_2,p), \end{align*} $$

where

$$ \begin{align*} {\cal A}(q_1,q_2,p) =\frac{1}{p-1} \sum_{N=1}^{p-1}\left (\sum_{1 \leq n_1,n_2 \leq N \atop q_1 n_1=q_2 n_2 \bmod p} 1\right ){\kern-1.5pt}. \end{align*} $$

Changing the order of summation and making the change of variables $n_1=q_2m_1$ , we arrive at

$$ \begin{align*}(p-1){\cal A}(q_1,q_2,p) = \sum_{1\leq m_1 \leq p}(p- \max(q_1 m_1,q_2 m_1)) =p^2 - \sum_{x \bmod p}\max(q_1 x,q_2 x).\end{align*} $$

By symmetry, injecting this into (3.10), we arrive at

(3.11) $$ \begin{align} p^3 -p\sum_{x \bmod p}\max(q_1 x,q_2 x) =\frac{(p-1)p(2p-1)}{6}+ \frac{p^3}{2\pi^2}M_{q_1,q_2}(p) +o(p^3). \end{align} $$

Hence, comparing the terms of order $p^3$ in the above formula (3.11) and using Corollary 3.8, we have

$$ \begin{align*}\sum_{x \bmod p}\max(q_1 x,q_2 x) =c_{q_1,q_2}(p^2 +o(p^2)),\end{align*} $$

where

$$ \begin{align*}1-c_{q_1,q_2} =\frac{1}{3}+\frac{1}{12}\frac{\gcd(q_1,q_2)^2}{q_1 q_2}.\end{align*} $$

This concludes the proof.

We know turn to the general case of Theorem 1.1. Let $d_0$ be a given square-free integer such that $\gcd (d_0,p)=1$ . For $\chi $ modulo p, let $\chi '$ be the character modulo $d_0p$ induced by $\chi $ . Recall that we want to show for H a subgroup of $\left (\mathbb {Z}/p\mathbb {Z}\right )^*$ of odd order $d \ll \frac {\log p}{\log \log p}$ that

$$ \begin{align*} M_{d_0}(p,H) &=\frac{1}{\# X_p^-(H)} \sum_{\chi \in X_p^-(H)} |L(1,\chi')|^2 =(1+o(1))\frac{\pi^2}{6}\prod_{q \mid d_0}\left(1-\frac{1}{q^2}\right){\kern-1.5pt}. \end{align*} $$

3.5 Twisted average of L- functions over subgroups

For any integers $q_1,q_2 \geq 1$ and any prime $p\geq 3$ , we define

$$ \begin{align*}M_{q_1,q_2}(p,H) :=\frac{1}{\# X_p^-(H)} \sum_{\chi \in X_p^-(H)} \chi(q_1)\overline{\chi}(q_2)|L(1,\chi)|^2.\end{align*} $$

Our main result is the following.

Theorem 3.14 Let $q_1$ and $q_2$ be given coprime integers. When H runs over the subgroups of $\left (\mathbb {Z}/p\mathbb {Z}\right )^*$ of odd order d, we have the following asymptotic formula:

$$ \begin{align*}M_{q_1,q_2}(p,H) =\frac{\pi^2}{6q_1 q_2} + O\left( d (\log p)^2p^{-\frac{1}{\phi(d)}} \right){\kern-1.5pt}.\end{align*} $$

Proof The proof follows the same lines as the proof of Theorem 3.6. Let us define

$$ \begin{align*}\epsilon_H(a,b) :=\frac{1}{\# X_p^-(H)}\sum_{\chi \in X_p^-(H)}\chi(a)\overline{\chi}(b) =\begin{cases} 1, &\text{ if } p\nmid ab \text{ and } a \in bH, \\ -1, &\text{ if } p\nmid ab \text{ and } a \in -bH, \\ 0, &\text{ otherwise.} \end{cases}\end{align*} $$

Hence, we obtain similarly

$$ \begin{align*} M_{q_1,q_2}(p,H) &=\frac{\pi^2}{4p^2} \sum_{a=1}^{p-1} \sum_{b=1}^{p-1}\epsilon_H(q_1 a,q_2 b)\cot\left(\frac{\pi a}{p}\right)\cot\left(\frac{\pi b}{p}\right) \\& = \frac{\pi^2}{2p^2}\sum_{h\in H} \sum_{a=1}^{p-1}\cot\left(\frac{\pi q_1a}{p}\right)\cot\left(\frac{\pi q_2h a}{p}\right) \\& = \frac{2\pi^2}{p}s(q_1,q_2,p)+ O\left(p^{-1}\sum_{1\neq h\in H} s(q_1,q_2h,p)\right) \\& = \frac{\pi^2}{6q_1q_2} + O(1/p) + O\left( \vert H\vert (\log p)^2p^{-\frac{1}{\phi(d)}} \right) \\& =\frac{\pi^2}{6q_1q_2} + O\left( d(\log p)^2p^{-\frac{1}{\phi(d)}} \right){\kern-1.5pt}, \end{align*} $$

where we used Theorem 3.5 in the last line and noticed that $\phi (k)$ divides $\phi (d)$ whenever k divides d.

Remark 3.15 The error term is negligible as soon as $d\leq \frac {\log p}{3(\log \log p)}$ .

Corollary 3.16 Let $q_1$ and $ q_2$ be given integers. When H runs over the subgroups of $\left (\mathbb {Z}/p\mathbb {Z}\right )^*$ of odd order d, we have the following asymptotic formula:

$$ \begin{align*}M_{q_1,q_2}(p,H) = \frac{\pi^2}{6}\frac{\gcd(q_1,q_2)^2}{q_1 q_2} + O\left( d (\log p)^2p^{-\frac{1}{\phi(d)}} \right){\kern-1.5pt}.\end{align*} $$

3.6 Proof of Theorem 1.1

As in the proof of Corollary 3.9 and using Corollary 3.16,

$$ \begin{align*} M_{d_0}(p,H) & =\frac{1}{\# X_p^-(H)} \sum_{\chi \in X_p^-(H)} |L(1,\chi')|^2 =\sum_{\delta_1 \mid d_0} \sum_{\delta_2 \mid d_0} \frac{\mu(\delta_1)}{\delta_1}\frac{\mu(\delta_2)}{\delta_2} M_{\delta_1,\delta_2}(p,H) \\ &= \frac{\pi^2}{6} \sum_{\delta_1 \mid d_0} \sum_{\delta_2 \mid d_0} \frac{\mu(\delta_1)}{\delta_1^2}\frac{\mu(\delta_2)}{\delta_2^2}\gcd(\delta_1,\delta_2)^2 + O\left( d (\log p)^2p^{-\frac{1}{\phi(d)}} \right) \\ & = \frac{\pi^2}{6} \prod_{q \mid d_0} \left(1-\frac{1}{q^2}\right) + O\left( d (\log p)^2p^{-\frac{1}{\phi(d)}} \right) = (1+o(1))\frac{\pi^2}{6} \prod_{q \mid d_0}\left(1-\frac{1}{q^2}\right) \end{align*} $$

using the condition on d.

4 Explicit formulas for $M_{d_0}(f,H)$

Recall that by (3.9)

$$ \begin{align*} L(1,\chi) =\frac{\pi}{2f}\sum_{a=1}^{f-1}\chi (a)\cot\left (\frac{\pi a}{f}\right ) \ \ \ \ \ (\chi\in X_f^-). \end{align*} $$

Hence, using the definition of Dedekind sums, we obtain (see [Reference LouboutinLou16, Proof of Theorem 2])

(4.1) $$ \begin{align} M(f,H) ={2\pi^2\over f}\sum_{\delta\mid f}\frac{\mu (\delta)}{\delta}\sum_{h\in H}s(h,f/\delta). \end{align} $$

4.1 A formula for $M_{d_0}(f,\{1\})$ for $d_0=1,2,3,6$

The first consequence of (4.1) is a short proof of [Reference LouboutinLou94, Théorèmes 2 and 3] by taking $H=\{1\}$ , the trivial subgroup of the multiplicative group $({\mathbb Z}/f{\mathbb Z}^*)$ . Indeed, (4.1) and (3.4) give

$$ \begin{align*}M(f,\{1\}) ={2\pi^2\over f}\sum_{\delta\mid f}\frac{\mu (\delta)}{\delta}s(1,f/\delta) =\frac{\pi^2}{6}\sum_{\delta\mid f}\mu (\delta)\left (\frac{1}{\delta^2}-\frac{3}{\delta f}+\frac{2}{f^2}\right ){\kern-1.5pt}.\end{align*} $$

The arithmetic functions $f\mapsto \sum _{\delta \mid f}\mu (\delta )\delta ^k$ being multiplicative, we obtain (see also [Reference QiQi])

(4.2) $$ \begin{align} M(f,\{1\}) ={\pi^2\over 6} \times\left\{ \prod_{q\mid f}\left (1-{1\over q^2}\right )-{3\over f}\prod_{q\mid f}\left (1-{1\over q}\right ) \right\} \ \ \ \ \ (f>2). \end{align} $$

Now, it is clear by (2.1) that for $d_0$ odd and square-free and f odd, we have

$$ \begin{align*} M_{2d_0}(f,\{1\}) =M_{d_0}(2f,\{1\}). \end{align*} $$

Hence, on applying (4.2) to $2f$ instead of f, we therefore obtain

$$ \begin{align*} M_2(f,\{1\}) ={\pi^2\over 8} \times\left\{ \prod_{q\mid f}\left (1-{1\over q^2}\right )-{1\over f}\prod_{q\mid f}\left (1-{1\over q}\right ) \right\} \ \ \ \ \ (f>2 \mbox{ odd}). \end{align*} $$

For $d_0\in \{3,6\}$ , the following explicit formula holds true for any f coprime with $d_0$ . It generalizes [Reference LouboutinLou94, Théorème 4] to composite moduli.

Theorem 4.1 Let $d_0>2$ be a given square-free integer. Set

$$ \begin{align*}\kappa_{d_0} :=\frac{\pi^2}{6} \prod_{q\mid d_0}\left (1-\frac{1}{q^2}\right ) \mbox{ and } c :=3\prod_{q\mid d_0}\frac{q-1}{q+1}.\end{align*} $$

For $n\in {\mathbb Z}$ , set $\varepsilon (n) =+1$ if $n\equiv +1\ \ \pmod {d_0}$ and $\varepsilon (n) =-1$ if $n\equiv -1\ \ \pmod {d_0}$ . Then, for $f>2$ such that all its prime divisors q satisfy $q\equiv \pm 1\ \ \pmod {d_0}$ , we have

$$ \begin{align*}M_{d_0}^{}(f,\{1\}) =\kappa_{d_0} \times\left\{ \prod_{q\mid f}\left (1-\frac{1}{q^2}\right ) -\frac{c}{f}\prod_{q\mid f}\left (1-\frac{1}{q}\right ) +\varepsilon(f) \frac{c-1}{f}\prod_{q\mid f}\left (1-\frac{\varepsilon(q)}{q}\right ) \right\}{\kern-1.5pt}.\end{align*} $$

In particular, for $f>2$ such that all its prime divisors q satisfy $q\equiv 1\ \ \pmod {d_0}$ , we have

$$ \begin{align*}M_{d_0}^{}(f,\{1\}) =\kappa_{d_0} \times\left\{ \prod_{q\mid f}\left (1-\frac{1}{q^2}\right ) -\frac{1}{f}\prod_{q\mid f}\left (1-\frac{1}{q}\right ) \right\}{\kern-1.5pt}.\end{align*} $$

Proof With the notation of [Reference LouboutinLou11, Lemma 2], we have $M_{d_0}(f,\{1\})) =4\pi ^2S(d_0,f)$ . Hence, by [Reference LouboutinLou11, Lemmas 3 and 6], we have

$$ \begin{align*}M_{d_0}(f,\{1\}) =\frac{\pi^2}{6}\prod_{q\mid d_0f}\left (1-\frac{1}{q^2}\right ) -\frac{\pi^2}{2}\frac{\phi(d_0)^2\phi(f)}{d_0^2f^2} +\frac{\pi^2}{2d_0^2f}\sum_{d\mid f}\frac{\mu(d)}{d}A(d_0,f/d),\end{align*} $$

where the $A(d_0,f/d)$ ’s are rational numbers such that $A(d_0,f/d) =\varepsilon A(d_0,1)$ if $f/d\equiv \varepsilon \ \ \pmod {d_0}$ with $\varepsilon \in \{\pm 1\}$ (see (4.13)). If all the prime divisors q of f satisfy $q\equiv \pm 1\ \ \pmod {d_0},$ then $f/d\equiv \varepsilon (f/d)\ \ \pmod {d_0}$ and $A(d_0,f/d) =\varepsilon (f/d)A(d_0,1) =\varepsilon (f)A(d_0,1)\varepsilon (d)$ and

$$ \begin{align*}\sum_{d\mid f}\frac{\mu(d)}{d}A(d_0,f/d) =\varepsilon(f)A(d_0,1)\prod_{q\mid f}\left (1-\frac{\varepsilon(q)}{q}\right ){\kern-1.5pt}.\end{align*} $$

Hence, we finally get

$$ \begin{align*}M_{d_0}(f,\{1\}) =\frac{\pi^2}{6}\prod_{q\mid d_0f}\left (1-\frac{1}{q^2}\right ) -\frac{\pi^2}{2}\frac{\phi(d_0)^2\phi(f)}{d_0^2f^2} +\frac{\pi^2}{2d_0^2f}\varepsilon(f)A(d_0,1)\prod_{q\mid f}\left (1-\frac{\varepsilon(q)}{q}\right ){\kern-1.5pt}.\end{align*} $$

The desired formula for $M_{d_0}(f,\{1\})$ follows by using the explicit formula

$$ \begin{align*}A(d_0,1) =\phi(d_0)^2 -\frac{d_0^2}{3}\prod_{q\mid d_0}\left (1-\frac{1}{q^2}\right )\end{align*} $$

given in [Reference LouboutinLou11, Lemma 6].

4.2 A formula for $M(p,H)$

The second immediate consequence of (4.1) and (3.4) is:

Proposition 4.2 For $f>2$ and H a subgroup of the multiplicative group $({\mathbb Z}/f{\mathbb Z})^*$ , set

(4.3) $$ \begin{align} S'(H,f)=\sum_{1\neq h\in H}s(h,f) \mbox{ and } N(f,H) :=-3+\frac{2}{f}+12S'(H,f). \end{align} $$

Then, for $p\geq 3$ a prime and H a subgroup of odd order of the multiplicative group $({\mathbb Z}/p{\mathbb Z})^*$ , we have

(4.4) $$ \begin{align} M(p,H) =\frac{\pi^2}{6} \left (1+\frac{N(p,H)}{p}\right ) =\frac{\pi^2}{6}\left (\left (1-\frac{1}{p}\right )\left (1-\frac{2}{p}\right )+\frac{12S'(H,p)}{p}\right ). \end{align} $$

Remark 4.3 In particular, $N(f,\{1\}) =-3+2/f$ and (4.4) implies (1.5). Notice also that $N(p,H)\in {\mathbb Z}$ for $H\neq \{1\}$ , by [Reference LouboutinLou19, Theorem 6]. Moreover, by [Reference Louboutin and MunschLM21, Theorem $1.1$ ], the asymptotic formula $M(p,H) =\frac {\pi ^2}{6}+o(1)$ holds as p tends to infinity and H runs over the subgroup of $({\mathbb Z}/p{\mathbb Z})^*$ of odd order $d\leq \frac {\log p}{\log \log p}$ . Hence, we have $N(p,H)=o(p)$ under this restriction.

4.3 A formula for $M_{d_0}(p,H)$

We will now derive a third consequence of (4.1): a formula for the mean square value $M_{d_0}(f,H)$ defined in (1.9) when f is prime.

Theorem 4.4 Let $d_0>1$ be a square-free integer. Let $f>2$ be coprime with $d_0$ . Let H be a subgroup of the multiplicative group $({\mathbb Z}/f{\mathbb Z})^*$ . Whenever $\delta $ divides $d_0$ , let $s_\delta :({\mathbb Z}/\delta f{\mathbb Z})^*\longrightarrow ({\mathbb Z}/f{\mathbb Z})^*$ be the canonical surjective morphism and set $H_{\delta }=s_\delta ^{-1}(H)$ and $H_{\delta }' =s_\delta ^{-1}(H\setminus \{1\})$ . Define the rational number

(4.5) $$ \begin{align} N_{d_0}(f,H) =-f +\frac{12\mu(d_0)}{\prod_{q\mid d_0}(q^2-1)} \sum_{\delta\mid d_0}\delta\mu (\delta)\sum_{h\in H_{d_0}}s(h,\delta f). \end{align} $$

Then, for $p\geq 3$ a prime which does not divide $d_0$ and H a subgroup of odd order of the multiplicative group $({\mathbb Z}/p{\mathbb Z})^*$ , we have

(4.6) $$ \begin{align} M_{d_0}(p,H) =\frac{2\pi^2\mu(d_0)\phi(d_0)}{d_0^2p} \sum_{\delta\mid d_0}\frac{\delta\mu(\delta)}{\phi(\delta)} S(H_\delta,\delta p),\end{align} $$

where

$$ \begin{align*} S(H_\delta,\delta f) =\sum_{h\in H_\delta}s(h,\delta f), \end{align*} $$

and

(4.7) $$ \begin{align} M_{d_0}(p,H) =\kappa_{d_0} \times\left ( 1 +\frac{N_{d_0}(p,H)}{p} \right ), \mbox{ where } \kappa_{d_0} :=\frac{\pi^2}{6} \prod_{q\mid d_0}\left (1-\frac{1}{q^2}\right ). \end{align} $$

Moreover,

(4.8) $$ \begin{align} N_{d_0}(f,H) &=-f +\frac{12\mu(d_0)}{\prod_{q\mid d_0}(q+1)} \sum_{\delta\mid d_0}\frac{\delta\mu(\delta)}{\phi(\delta)}S(H_\delta,\delta f) \end{align} $$
(4.9) $$ \begin{align} \kern66pt &=N_{d_0}(f,\{1\}) +\frac{12\mu(d_0)}{\prod_{q\mid d_0}(q+1)} \sum_{\delta\mid d_0}\frac{\delta\mu(\delta)}{\phi(\delta)} S'(H_{\delta},\delta f), \end{align} $$

where

$$ \begin{align*} S'(H_{\delta},\delta f) :=\sum_{h\in H_{\delta}'}s(h,\delta f). \end{align*} $$

Proof Using (2.1) and by making the change of variables $\delta \mapsto d_0f/\delta $ in (4.1), we obtain

(4.10) $$ \begin{align} M_{d_0}(f,H) =M(d_0f,H_{d_0}) ={2\pi^2\over d_0^2f^2}\sum_{\delta\mid d_0f}\delta\mu (d_0f/\delta)\sum_{h\in H_{d_0}}s(h,\delta). \end{align} $$

Since $\{\delta ;\ \delta \mid d_0p\}$ is the disjoint union of $\{\delta ;\ \delta \mid d_0\}$ and $\{\delta p;\ \delta \mid d_0\}$ , by (4.10), we obtain

$$ \begin{align*}M_{d_0}(p,H) =-{2\pi^2\mu(d_0)\over d_0^2p^2}\sum_{\delta\mid d_0}\delta\mu (\delta)\sum_{h\in H_{d_0}}s(h,\delta) +{2\pi^2\mu(d_0)\over d_0^2p}\sum_{\delta\mid d_0}\delta\mu (\delta)\sum_{h\in H_{d_0}}s(h,\delta p).\end{align*} $$

Now, $S:=\sum _{h\in H_{d_0}}s(h,\delta )=0$ whenever $\delta \mid d_0$ , which gives

(4.11) $$ \begin{align} M_{d_0}(p,H) ={2\pi^2\mu(d_0)\over d_0^2p}\sum_{\delta\mid d_0}\delta\mu (\delta)\sum_{h\in H_{d_0}}s(h,\delta p) \end{align} $$

and implies (4.7). Indeed, let $\sigma :({\mathbb Z}/d_0f{\mathbb Z})^*\longrightarrow ({\mathbb Z}/\delta {\mathbb Z})^*$ be the canonical surjective morphism. Its restriction $\tau $ to the subgroup $H_{d_0}$ is surjective, by the Chinese reminder theorem. Hence, $S =(H_{d_0}:\ker \tau ) \times S'$ , where $S':=\sum _{c\in ({\mathbb Z}/\delta {\mathbb Z})^*}s(c,\delta ) =\sum _{c\in ({\mathbb Z}/\delta {\mathbb Z})^*}s(-c,\delta ) =-S'$ yields $S'=0$ . In the same way, whenever $\delta \mid d_0$ , the kernel of the canonical surjective morphism $s:({\mathbb Z}/d_0f{\mathbb Z})^*\longrightarrow ({\mathbb Z}/\delta f{\mathbb Z})^*$ being a subgroup of order $\phi (d_0f)/\phi (\delta f) =\phi (d_0)/\phi (\delta )$ , we have

(4.12) $$ \begin{align} \sum_{h\in H_{d_0}}s(h,\delta f) =\frac{\phi(d_0)}{\phi(\delta)}\sum_{h\in H_{\delta}}s(h,\delta f) \end{align} $$

and (4.6) follows from (4.11) and (4.12).

Then, (4.7) is a direct consequence of (4.6) and (4.5). Finally, (4.9) is an immediate consequence of (4.5) and (4.12).

4.3.1 A new proof of Theorem 1.1

We split the sum in (4.11) into two cases depending whether $h=1$ or not. By (3.4), we have $s(1,\delta p)=\frac {p\delta }{12}+O(1)$ giving a contribution to the sum of order

$$ \begin{align*}\frac{\pi^2 \mu(d_0)}{6d_0^2} \sum_{\delta \mid d_0}\delta^2\mu(\delta) + O(1/p) = \frac{\pi^2}{6} \prod_{q \mid d_0}\left(1-\frac{1}{q^2}\right)+O(1/p).\end{align*} $$

When $h\neq 1$ and $h\in H_{d_0}$ , it is clear that the order of h modulo p is between $3$ and d. Hence, it follows from Theorem 3.1 (see the Remark after) that $s(h,\delta p)=O((\log p)^2 p^{1-\frac {1}{\phi (d)}})$ . The integer $d_0$ being fixed, we can sum up these error terms and the proof is finished.

4.4 An explicit way to compute $N_{d_0}(f,\{1\})$

Lemma 4.5 Let $d_0>1$ be a square-free integer. Let $f>2$ be coprime with $d_0$ . Recall that $H_{d_0}(f) =\{h\in ({\mathbb Z}/d_0f{\mathbb Z})^*,\ h\equiv 1\ \ \pmod f\}$ and set

$$ \begin{align*}U(d_0,f) :=\sum_{1\neq h\in H_{d_0}(f)} \sum_{n=1\atop\gcd(d_0,n)=1}^{d_0f -1} \left (1+\cot\left ({\pi n\over d_0f}\right )\cot\left ({\pi nh\over d_0f}\right )\right )\end{align*} $$

and

(4.13) $$ \begin{align} A(d_0,f) =\sum_{a\in ({\mathbb Z}/d_0{\mathbb Z})^*} \sum_{b\in ({\mathbb Z}/d_0{\mathbb Z})^*\atop b\neq a} \cot\left (\frac{\pi (b-a)}{d_0}\right ) \left ( \cot\left (\frac{\pi fa}{d_0}\right ) -\cot\left (\frac{\pi fb}{d_0}\right ) \right ){\kern-1.5pt}, \end{align} $$

a rational number depending only on f modulo $d_0$ . Then $U(d_0,f) =fA(d_0,f).$

Proof As in [Reference LouboutinLou11, Lemma 3], set

$$ \begin{align*}T(d_0,f) :=\sum_{1\neq h\in H_{d_0}(f)} \sum_{n=1\atop\gcd(d_0f,n)=1}^{d_0f -1} F\left (\frac{n}{d_0f},\frac{nh}{d_0f}\right ){\kern-1.5pt},\end{align*} $$

where $F(x,y)=1+\cot (\pi x)\cot (\pi y)$ . On the one hand, since $\gcd (d_0f,n)=1$ if and only if $\gcd (d_0,n)=\gcd (f,n)=1$ and $\displaystyle {\sum _{d\mid f\atop d\mid n}\mu (d)}$ is equal to $1$ if $\gcd (f,n)=1$ and is equal to $0$ otherwise, we have

$$ \begin{align*}T(d_0,f) =\sum_{d\mid f}\mu(d) \sum_{1\neq h\in H_{d_0}(f)} \sum_{n=1\atop\gcd(d_0,n)=1}^{d_0(f/d) -1}F\left (\frac{n}{d_0(f/d)},\frac{nh}{d_0(f/d)}\right ){\kern-1.5pt}.\end{align*} $$

On the other hand, the canonical morphism $\sigma :H_{d_0}(f) \rightarrow H_{d_0}(f/d)$ is surjective and both groups have order $\phi (d_0f)/\phi (f) =\phi (d_0(f/d))/\phi (f/d) =\phi (d_0)$ . Hence, $\sigma $ is bijective and

$$ \begin{align*}T(d_0,f) =\sum_{d\mid f}\mu(d)U(d_0,f/d).\end{align*} $$

Using [Reference LouboutinLou11, Lemma 6] and Möbius’ inversion formula, we finally do obtain

$$ \begin{align*} U(d_0,f) =\sum_{d\mid f}T(d_0,d) &=\sum_{d\mid f}d\sum_{\delta\mid d}\frac{\mu(\delta)}{\delta}A(d_0,d/\delta)\\ &=\sum_{\delta'\mid f} \delta'\left (\sum_{\delta\mid f/\delta'}\mu(\delta)\right )A(d_0,\delta') =fA(d_0,f), \end{align*} $$

where we set $\delta '=d/\delta $ .

Proposition 4.6 Let $d_0>1$ be a square-free integer. Set $B=\prod _{q\mid d_0}(q^2-1)$ . For $f>2$ and $\gcd (d_0,f)=1,$ we have

$$ \begin{align*}N_{d_0}(f,\{1\}) =\frac{3}{B} \left (A(d_0,f)-\phi(d_0)^2\right ){\kern-1.5pt}.\end{align*} $$

Consequently, $N_{d_0}(f,\{1\})$ is a rational number depending only on f modulo $d_0$ .

Proof Set $H=H_{d_0}(f):=\{h\in ({\mathbb Z}/d_0f{\mathbb Z})^*,\ h\equiv 1\ \ \pmod f\}$ . By (4.5), we have

$$ \begin{align*}N_{d_0}(f,\{1\}) =-f+\frac{12\mu(d_0)}{B}\sum_{\delta\mid d_0}\delta\mu (\delta) \sum_{h\in H}s(h,\delta f).\end{align*} $$

Using (3.4) to evaluate the contribution of $h=1$ in this expression and $\sum _{\delta \mid d_0}\mu (\delta ) =0$ , we get

$$ \begin{align*}N_{d_0}(f,\{1\}) =-\frac{3\phi(d_0)}{B} +\frac{12\mu(d_0)}{B}\sum_{\delta\mid d_0}\delta\mu (\delta) \sum_{1\neq h\in H}s(h,\delta f)\end{align*} $$

and

$$ \begin{align*}N_{d_0}(f,\{1\}) =-\frac{3\phi(d_0)^2}{B} +\frac{3\mu(d_0)}{Bf} \sum_{1\neq h\in H} \sum_{\delta\mid d_0}\mu (\delta) \sum_{n=1}^{\delta f -1}\left (1+\cot\left ({\pi n\over\delta f}\right )\cot\left ({\pi nh\over\delta f}\right )\right ){\kern-1.5pt},\end{align*} $$

by (3.1) and by noticing that $\# H =\phi (d_0)$ . Therefore,

(4.14) $$ \begin{align} N_{d_0}(f,\{1\}) =-\frac{3\phi(d_0)^2}{B} +\frac{3}{Bf}S(d_0,f) \end{align} $$

(make the change of variable $\delta \mapsto d_0/\delta $ ). Lemma 4.5 gives the desired result.

Remark 4.7 As a consequence, we obtain $M_{d_0}(p,\{1\}) =\frac {\pi ^2}{6}\prod _{q\mid d_0}\left (1-\frac {1}{q^2}\right ) +O(p^{-1})$ , using (4.7) and the fact that $N_{d_0}(p,\{1\})$ depends only on p modulo $d_0$ . This gives in this extreme situation another proof of Theorem 1.1 with a better error term. Moreover, in that situation, we have $K ={\mathbb Q}(\zeta _p)$ and in (1.11) the term $\Pi _{d_0}(p,\{1\})$ is bounded from below by a constant independent of p, by Corollary 2.4.

5 The case where $f=a^{d-1}+\cdots +a^2+a+1$

In this specific case, we are able to obtain explicit formulas for $M_{d_0}(f,H)$ when the subgroup H is defined in terms of the parameter a defining the modulus. For a general subgroup H, it seems unrealistic to be more explicit than the formula involving Dedekind sums given in Theorem 4.4. It might be interesting to explore formulas involving continued fraction expansions in view of their link to Dedekind sums [Reference HickersonHic].

5.1 Explicit formulas for $d_0=1,2$

Lemma 5.1 Let $f>1$ be a rational integer of the form $f=(a^d-1)/(a-1)$ for some $a\neq -1,0,1$ and some odd integer $d\geq 3$ . Hence, f is odd. Set $H=\{a^k;\ 0\leq k\leq d-1\}$ , a subgroup of order d of the multiplicative group $({\mathbb Z}/f{\mathbb Z})^*$ . Then

$$ \begin{align*}S(H,f) =\frac{a+1}{a-1}\times\frac{f-(d-1)a-1}{12}\end{align*} $$

and

$$ \begin{align*}S(H_2,2f) =\begin{cases} \frac{a+1}{a-1}\times\frac{4f-(d-1)a-3d-1}{24},&\mbox{if } a \mbox{ is odd,}\\ \frac{2a-1}{a-1}\times\frac{f-(d-1)a-1}{12},&\mbox{if } a \mbox{ is even.} \end{cases}\end{align*} $$

Proof We have $S(H,f) =\sum _{k=0}^{d-1}s(a^k,f)$ . Moreover, $S(H_2,2f) =\sum _{k=0}^{d-1}s(a^k,2f)$ if a is odd and $S(H_2,2f) =s(1,2f)+\sum _{k=1}^{d-1}s(a^k+f,2f)$ if a is even. Now, we claim that for $0\leq k\leq d-1$ , we have

$$ \begin{align*}s(a^k,f) =\frac{a^k}{12f}+\frac{(f^2+1)a^{-k}}{12f} +\frac{a^k+a^{-k}(a^2-2a+2)}{12(a-1)} -\frac{a(a+1)}{12(a-1)} \mbox{ whatever the parity of } a, \end{align*} $$
$$ \begin{align*}s(a^k,2f) =\frac{a^k}{24f}+\frac{(4f^2+1)a^{-k}}{24f} +\frac{4a^k+a^{-k}(a^2-2a+5)}{24(a-1)} -\frac{(a+1)(a+3)}{24(a-1)} \mbox{ if } a \mbox{ is odd,} \end{align*} $$

and that for $1\leq k\leq d-1$ , we have

$$ \begin{align*}s(a^k+f,2f) =\frac{a^k}{24f}+\frac{(f^2+1)a^{-k}}{24f} +\frac{a^k+a^{-k}(a^2-2a+2)}{24(a-1)} -\frac{a(2a-1)}{12(a-1)} \mbox{ if } a \mbox{ is even.} \end{align*} $$

Noticing that $\sum _{k=1}^{d-1}a^k =f-1$ and $\sum _{k=1}^{d-1}a^{-k} =\frac {f-1}{(a-1)f+1}$ , we then get the assertions on $S(H,f)$ and $S(H_2,2f)$ . Now, let us, for example, prove the third claim. Hence, assume that a is even and that $1\leq k\leq d-1$ . Then $f_k :=(a^k-1)/(a-1)$ is odd, $\mathrm {sign}(f_k)=\mathrm {sign}(a)^k$ and $a^k+f>0$ . First, since $2f\equiv -2a^{k}\ \ \pmod {a^k+f}$ , using (3.3), we have

$$ \begin{align*}s(a^k+f,2f) =\frac{a^k+f}{24f} +\frac{f}{6(a^k+f)} -\frac{1}{4} +\frac{1}{24(a^k+f)f} +s(2a^{k},a^k+f).\end{align*} $$

Second, noticing that $a^k+f\equiv f_k\ \ \pmod {2a^{k}}$ and using (3.3), we have

$$ \begin{align*}s(2a^{k},a^k+f) =\frac{a^{k}}{6(a^k+f)} +\frac{a^k+f}{24a^{k}} -\frac{\mathrm{sign}(a)^k}{4} +\frac{1}{24a^{k}(a^k+f)} -s(f_k,2a^{k}).\end{align*} $$

Finally, noticing that $2a^{k}\equiv 2\ \ \pmod {f_k}$ and using (3.3) and (3.4), we have

$$ \begin{align*} s(f_k,2a^{k}) &=\frac{f_k}{24a^{k}} +\frac{a^{k}}{6f_k} -\frac{\mathrm{sign}(a)^k}{4} +\frac{1}{24f_ka^{k}} -s(2,f_k)\\ &=\frac{f_k}{24a^{k}} +\frac{a^{k}}{6f_k} -\frac{\mathrm{sign}(a)^k}{4} +\frac{1}{24f_ka^{k}} -\frac{f_k^2-6f_k+5}{24f_k}. \end{align*} $$

After some simplifications, we obtain the desired formula for $s(a^k+f,2f)$ .

Notice that for $d=3$ , we obtain $S(H,f) =\frac {f-1}{12}$ , in accordance with (6.1).

Using (4.6) and Lemma 5.1, we readily obtain:

Theorem 5.2 Let $d\geq 3$ be a prime integer. Let $p\equiv 1\ \ \pmod {2d}$ be a prime integer of the form $p=(a^d-1)/(a-1)$ for some $a\neq -1,0,1$ . Let K be the imaginary subfield of degree $(p-1)/d$ of the cyclotomic field ${\mathbb Q}(\zeta _p)$ . Set $H=\{a^k;\ 0\leq k\leq d-1\}$ , a subgroup of order d of the multiplicative group $({\mathbb Z}/p{\mathbb Z})^*$ . We have the mean square value formulas

(5.1) $$ \begin{align} M(p,H) =\frac{\pi^2}{6}\times\frac{a+1}{a-1}\times\left (1-\frac{(d-1)a+1}{p}\right ) \end{align} $$

and

(5.2) $$ \begin{align} M_2(p,H) =\frac{\pi^2}{8}\times \begin{cases} \frac{a+1}{a-1}\times\left (1-\frac{d}{p}\right ),&\mbox{ if } a \mbox{ is odd,}\\ 1-\frac{(d-1)a+1}{p},&\mbox{ if } a \mbox{ is even.}\\ \end{cases} \end{align} $$

Consequently, for a given d, as $p\rightarrow \infty $ , we have

$$ \begin{align*} M(p,H) =\frac{\pi^2}{6}+o(1) \mbox{ and } M_2(p,H) =\frac{\pi^2}{8}+o(1). \end{align*} $$

On the other hand, for a given a, as $p\rightarrow \infty $ , we have

$$ \begin{align*} M(p,H) =\frac{\pi^2}{6}\times\frac{a+1}{a-1}+o(1) \mbox{ and } M_2(p,H) =\begin{cases} \frac{\pi^2}{8}\times\frac{a+1}{a-1}+o(1), &\mbox{if } a \mbox{ is odd,}\\ \frac{\pi^2}{8}+o(1), &\mbox{if } a \mbox{ is even.} \end{cases} \end{align*} $$

Remark 5.3 Assertion (5.1) was initially provedFootnote 5 in [Reference LouboutinLou16, Theorem 5] for $d=5$ and then generalized in [Reference Louboutin and MunschLM21, Proposition 3.1] to any $d\geq 3$ . However, (5.1) is much simpler than [Reference Louboutin and MunschLM21, equation (22)]. Notice that if p runs over the prime of the form $p=(a^d-1)/(a-1)$ with $a\neq 0,2$ even then $M_2(p,H) =\frac {6}{8}\times \frac {a-1}{a+1}\times M(p,H)$ and the asymptotic (1.12) is not satisfied.

5.2 The case where p is a Mersenne prime and $d_0=1,3,15$

In the setting of Theorem 5.4, we have $2\in H$ . Hence, by Remark 2.2, we assume that $d_0$ is odd.

Theorem 5.4 Let $p=2^d-1>3$ be a Mersenne prime. Hence, d is odd and $H=\{2^k;\ 0\leq k\leq d-1\}$ is a subgroup of odd order d of the multiplicative group $({\mathbb Z}/p{\mathbb Z})^*$ . Let K be the imaginary subfield of degree $m =(p-1)/d$ of ${\mathbb Q}(\zeta _p)$ . Then

$$ \begin{align*}M(p,H) =\frac{\pi^2}{2}\left (1-\frac{2d-1}{p}\right ) \leq\frac{\pi^2}{2} \text{ and } h_K^-\leq 2\left (\frac{p}{8}\right )^{m/4},\end{align*} $$
$$ \begin{align*}M_3(p,H) =\frac{4\pi^2}{9}\left (1-\frac{d}{p}\right ) \leq\frac{4\pi^2}{9} \text{ and } h_K^-\leq 2\left (\frac{p}{9}\right )^{m/4},\end{align*} $$

and

$$ \begin{align*}M_{15}(p,H) =\frac{32\pi^2}{75}\left (1-\frac{c_d}{48p}\right ) \leq\frac{32\pi^2}{75}, \mbox{ where } c_d =\begin{cases} 47d+1,&\mbox{if } d\equiv 1\ \ \pmod 4,\\ 17d-3,&\mbox{if } d\equiv 3\ \ \pmod 4. \end{cases}\end{align*} $$

In particular, for $d\equiv 3\ \ \pmod 4$ , we have $h_K^-\leq 2\left (\frac {8p}{75}\right )^{m/4}.$

Proof By (4.6), we have

(5.3) $$ \begin{align} M_{d_0}(p,H) =\frac{\pi^2}{2}\left\{\prod_{q\mid d_0}\left (1-\frac{1}{q^2}\right )\right\} \left (1+\frac{N_{d_0}'(p,H)}{p}\right ), \end{align} $$

where for H a subgroup of odd order of the multiplicative group $({\mathbb Z}/f{\mathbb Z})^*$ , we set

(5.4) $$ \begin{align} N_{d_0}'(f,H) :=-f +\frac{4\mu(d_0)}{\prod_{q\mid d_0}(q+1)} \sum_{\delta\mid d_0}\frac{\delta\mu(\delta)}{\phi(\delta)}S(H_\delta,\delta f). \end{align} $$

The formulas for $M(p,H),M_3(p,H)$ , and $M_{15}(p,H)$ follow from (5.3)) and Lemma 5.5. The upper bounds on $h_K^-$ follow from (1.11) and Lemma 2.3 according to which $\Pi _q(p,H)\geq 1$ if q is of even order in the quotient group $G/H$ , where $G=({\mathbb Z}/p{\mathbb Z})^*$ , hence, if q is of even order in the group G. Now, since $p\equiv 3\ \ \pmod 4$ the group G is of order $p-1=2N$ with N odd and q is of even order in G if and only $q^N=-1$ in G, i.e., if and only if the Legendre symbol $\left (\frac {q}{p}\right )$ is equal to $-1$ . Now, since $p=2^d-1\equiv -1\equiv 3 \ \ \pmod 4$ for $d\geq 3$ , the law of quadratic reciprocity gives $\left (\frac {3}{p}\right ) =-\left (\frac {p}{3}\right ) =-\left (\frac {1}{3}\right ) =-1$ , as $p\equiv (-1)^d-1\equiv -2\equiv 1\ \ \pmod 3$ . Hence, $\Pi _3(p,H)\geq 1$ . In the same way, if $d\equiv 3\ \ \pmod 4$ then $p=2^d-1 =2\cdot 4^{\frac {d-1}{2}}-1\equiv 2\cdot (-1)^{\frac {d-1}{2}}-1\equiv -3\equiv 2\ \ \pmod 5$ and $\left (\frac {5}{p}\right ) =\left (\frac {p}{5}\right ) =\left (\frac {2}{5}\right ) =-1$ and $\Pi _5(p,H)\geq 1$ .

Lemma 5.5 Set $f=2^d-1$ and $\varepsilon _d =(-1)^{(d-1)/2}$ with $d\geq 2$ odd. Hence, $\gcd (f,15)=1$ . Set $H =\{2^k;\ 0\leq k\leq d-1\}$ , a subgroup of order d of the multiplicative group $({\mathbb Z}/f{\mathbb Z})^*$ . Then,

(5.5) $$ \begin{align} S(H,f) =\frac{f-2d+1}{4} \mbox{ and } N'(f,H) =-2d+1, \end{align} $$
(5.6) $$ \begin{align} S(H_3,3f) =\frac{5f-6d+1}{6} \mbox{ and } N_3'(f,H) =-d, \end{align} $$
(5.7) $$ \begin{align} S(H_5,5f) =\frac{7f-10d+2+\varepsilon_d}{5} \mbox{ and } N_5'(f,H) =-\frac{4}{3}d+\frac{1+\varepsilon_d}{6}, \end{align} $$
(5.8) $$ \begin{align} S(H_{15},15f) =\frac{14f-\left (12+3\varepsilon_d\right )d+1}{3} \mbox{ and } N_{15}'(f,H) =-\frac{32+15\varepsilon_d}{48}d +\frac{1-2\varepsilon_d}{48}. \end{align} $$

Proof The first assertion is the special case $a=2$ of Lemma 5.1. Let us now deal with the second assertion. Here, $H_3 =\{2^k;\ 0\leq k\leq d-1\} \cup \{2^k+(-1)^kf;\ 0\leq k\leq d-1\}$ . We assume that $0\leq k\leq d-1$ . Hence, $\mathrm {sign}(2^k+(-1)^kf)=(-1)^k$ .

1. Noticing that $3f\equiv -3\ \ \pmod {2^k}$ , by (3.3), we obtain

$$ \begin{align*}s(2^k,3f) =\frac{4^k+9f^2-9\cdot 2^k\cdot f+1}{36\cdot 2^k\cdot f}+s(3,2^k).\end{align*} $$

Noticing that $2^k\equiv (-1)^k\ \ \pmod 3$ , by (3.3) and (3.4), we obtain

$$ \begin{align*}s(3,2^k) =\frac{9+4^k-9\cdot 2^k+1}{36\cdot 2^k}-(-1)^ks(1,3) =\frac{9+4^k-9\cdot 2^k+1}{36\cdot 2^k}-\frac{(-1)^k}{18}.\end{align*} $$

Hence,

$$ \begin{align*}s(2^k,3f) =\frac{f+1}{36f}2^k +\frac{(f+1)(9f+1)}{36f}2^{-k} -\frac{1}{2} -\frac{(-1)^k}{18}.\end{align*} $$

2. Noticing that $3f\equiv -3\cdot (-1)^k2^k\ \ \pmod {2^k+(-1)^kf}$ , by (3.3), we obtain

$$ \begin{align*} s(2^k+(-1)^kf,3f) =&\frac{2^k+(-1)^kf}{36f} +\frac{f}{4(2^k+(-1)^kf)} -\frac{(-1)^k}{4} +\frac{1}{36(2^k+(-1)^kf)f}\\ &+(-1)^ks(3\cdot 2^k,2^k+(-1)^kf), \end{align*} $$

and noticing that $2^k+(-1)^kf\equiv (-1)^{k-1}\ \ \pmod {3\cdot 2^k}$ , by (3.3), we obtain

$$ \begin{align*} s(3\cdot 2^k,2^k+(-1)^kf) =&\frac{3\cdot 2^k}{12(2^k+(-1)^kf)} +\frac{2^k+(-1)^kf}{36\cdot 2^k} -\frac{(-1)^k}{4} \\ &+\frac{1}{36\cdot 2^k\cdot (2^k+(-1)^kf)}+(-1)^ks(1,3\cdot 2^k). \end{align*} $$

Using (3.4), we finally obtain

$$ \begin{align*}s(2^k+(-1)^kf,3f) =\frac{9f+1}{36f}2^k +\frac{(f+1)^2}{36f}2^{-k} -\frac{1}{2} +\frac{(-1)^k}{18}.\end{align*} $$

3. Using $\sum _{k=0}^{d-1}2^k =f$ , $\sum _{k=0}^{d-1}2^{-k}=\frac {2f}{f+1}$ and $\sum _{k=0}^{d-1}(-1)^k=1$ , we obtain

$$ \begin{align*}\sum_{k=0}^{d-1}s(2^k,3f) =\frac{19f-18d+1}{36} \mbox{ and } \sum_{k=0}^{d-1}s(2^k+(-1)^kf,3f) =\frac{11f-18d+5}{36}.\end{align*} $$

Hence, we do obtain

$$ \begin{align*}S(H_3,3f) =\sum_{h\in H_3}s(h,3f) =\frac{19f-18d+1}{36} +\frac{11f-18d+5}{36} =\frac{5f-6d+1}{6}\end{align*} $$

and $N_3'(f,H) =-d$ , by (5.4).

Let us finally deal with the third and fourth assertions. The proof involves tedious and repetitive computations. For this reason, we will restrict ourselves to a specific case. Let us, for example, give some details for the proof of (5.8) in the case that $d\equiv 1\ \ \pmod 4$ . We have $f=2^d-1\equiv 1\ \ \pmod {30}$ and $H_{15}=\cup _{l=0}^{14}E_l$ , where $E_l := \{2^k+lf;\ 0\leq k\leq d-1,\ \gcd (2^k+l,15)=1\}$ for $0\leq l\leq 14$ . We have to compute the sums $s_l:=\sum _{n\in E_l}s(n,15f)$ . Let us, for example, give some details in the case that $l=1$ . We have $\gcd (2^k+1,15)=1$ if and only if $k\equiv 0\ \ \pmod 4$ . Hence, $s_1=\sum _{k=0}^{(d-1)/4}s(16^{k}+f,15f)$ . Using (3.3) and (3.4), we obtain

$$ \begin{align*}s(16^k+f,15f) =\frac{9f+1}{180f}16^k +\frac{14}{45} +\frac{(f+1)^2}{180f}16^{-k}.\end{align*} $$

Finally, using $\sum _{k=0}^{(d-1)/4}16^k =\frac {8f+7}{15}$ and $\sum _{k=0}^{(d-1)/4}16^{-k} =\frac {2(8f+7)}{15(f+1)}$ , we obtain

$$ \begin{align*}s_1 =\sum_{k=0}^{(d-1)/4}s(16^k+f,15f) =\frac{88f^2+(210d+731)f+21}{2700f}.\end{align*} $$

Finally, using (5.4)–(5.7), we get (5.8).

We conclude this section with the following result for $d_0 =3\times 5\times 7 =105$ , whose long proof we omit:Footnote 6

Lemma 5.6 Set $f =2^d-1$ with $d>1$ odd. Assume $\gcd (f,105)=1$ , i.e., that $d\equiv 1, 5, 7, 11\ \ \pmod {12}$ . Set $H =\{2^k;\ 0\leq k\leq d-1\}$ , a subgroup of order d of the multiplicative group $({\mathbb Z}/f{\mathbb Z})^*$ . Then

$$ \begin{align*} N_{105}'(f,H) =-\frac{1}{576} \times\begin{cases} 437d+139,&\mbox{if } d\equiv 1\ \ \pmod {12},\\ 535d-644,&\mbox{if } d\equiv 5\ \ \pmod{12},\\ 97d-324,&\mbox{if } d\equiv 7\ \ \pmod{12},\\ 195d+13,&\mbox{if } d\equiv 11\ \ \pmod {12}. \end{cases} \end{align*} $$

Lemmas 5.5 and 5.6 show that the following conjecture holds true for $d_0\in \{1,3,5,15,105\}$ :

Conjecture 5.7 Let $d_0\geq 1$ be odd and square-free. Let N be the order of $2$ in the multiplicative group $({\mathbb Z}/d_0{\mathbb Z})^*$ . Set $f =2^d-1$ with $d>1$ odd and $H =\{2^k;\ 0\leq k\leq d-1\}$ , a subgroup of order d of the multiplicative group $({\mathbb Z}/f{\mathbb Z})^*$ . Assume $\gcd (f,d_0)=1$ . Then $N_{d_0}'(f,H) =A_1(d)d+A_0(d)$ , where $A_1(d)$ and $A_0(d)$ are rational numbers which depend only on d modulo N, i.e., only on f modulo $d_0$ . Hence, for a prime $p\geq 3$ , we expect

$$ \begin{align*}M_{d_0}(p,H) =\frac{\pi^2}{2}\left\{\prod_{q \mid d_0} \left(1-\frac{1}{q^2}\right)\right\}\left(1+\frac{A_1(d)d}{p}+\frac{A_0(d)}{p}\right),\end{align*} $$

confirming again that the restriction on d in Theorem 1.1 should be sharp.

There is apparently no theoretical obstruction preventing us to prove Conjecture 5.7. Indeed, for a fixed $d_0$ , the formulas for $A_0(d)$ and $A_1(d)$ could be guessed using numerous examples on a computer algebra system. However, for large $d_0$ ’s, the set of cases to consider grows linearly and a more unified approach seems to be required to give a complete proof.

6 The case of subgroups of order $d=3$

6.1 Formulas for $d_0=1,2,6$

Let $p\equiv 1\ \ \pmod 6$ be a prime integer. Let K be the imaginary subfield of degree $m=(p-1)/3$ of the cyclotomic field ${\mathbb Q}(\zeta _p)$ . Since p splits completely in the quadratic field ${\mathbb Q}(\sqrt {-3})$ of class number one, there exists an algebraic integer $\alpha =a+b\frac {1+\sqrt {-3}}{2}$ with $a,b\in {\mathbb Z}$ such that $p =N_{{\mathbb Q}(\sqrt {-3})/{\mathbb Q}}(\alpha ) =a^2+ab+b^2$ . Then, $H=\{1,a/b,b/a\}$ is the unique subgroup of order $3$ of the cyclic multiplicative group $({\mathbb Z}/p{\mathbb Z})^*$ . So we consider the integers $f>3$ of the form $f=a^2+ab+b^2$ , with $a,b\in {\mathbb Z}\setminus \{0\}$ and $\gcd (a,b)=1$ , which implies $\gcd (a,f) =\gcd (b,f)=1$ and the oddness of f. We have the following explicit formula.

Lemma 6.1 Let $f>3$ be of the form $f=a^2+ab+b^2$ , with $a,b\in {\mathbb Z}$ and $\gcd (a,b)=1$ . Set $H=\left \{1,a/b,b/a\right \}$ , a subgroup of order $3$ of the multiplicative group $({\mathbb Z}/f{\mathbb Z})^*$ . Then

(6.1) $$ \begin{align} s(a,b,f) =\frac{f-1}{12f}, \ S(H,f) =\frac{f-1}{12} \mbox{ and } N(f,H) =-1+12S(H,f) =-1. \end{align} $$

Proof Noticing that $s(b,f,a) =s(b,b^2,a) =s(1,b,a) =s(b,a)$ , by (3.5), and $s(f,a,b) =s(a^2,a,b) =s(a,1,b) =s(a,b)$ , and using (3.3), we obtain

$$ \begin{align*} s(a,b,f) &=\frac{a^2+b^2+f^2-3\vert ab\vert f}{12abf}-s(b,f,a)-s(f,a,b) \ \ \ \ \ \text{(by (3.6))}\\ &=\frac{a^2+b^2+f^2-3\vert ab\vert f}{12abf} -s(b,a)-s(a,b)\\ &=\frac{a^2+b^2+f^2-3\vert ab\vert f}{12abf} -\frac{a^2+b^2-3\vert ab\vert+1}{12ab} =\frac{f-1}{12f}. \end{align*} $$

Finally, $S(H,f)=s(1,f)+s(a,b,f)+s(b,a,f) =s(1,f)+2s(a,b,f)$ and use (3.4) and (4.9).

Remark 6.2 Take $f_1=A^2+AB+B^2>0$ , where $3\nmid f_1$ and $\gcd (A,B)=1$ . Set $f=(f_1+1)^3-1$ . Then $f=a^2+ab+b^2$ , where $a=Af_1+A-B$ , $b=Bf_1+A+2B$ and $\gcd (a,b)=1$ . By Lemmas 6.1, we have an infinite family of moduli f for which the multiplicative group $({\mathbb Z}/f{\mathbb Z})^*$ contains at the same time an element $h=a/b$ of order $d=3$ for which $s(h,f)$ is asymptotic to $1/12$ and an element $h'=f_1+1$ of order $d=3$ for which $s(h',f)$ is asymptotic to $f^{2/3}/12$ . Indeed, by (3.3) and (3.4), for $f=(f_1+1)^3-1$ , we have $s(h',f) =\frac {h^{\prime 5}+h^{\prime 4} - 6h^{\prime 3}+6}{12f}$ .

To deal with the case $d_0>1$ , we notice that by (4.9), we have the following proposition.

Proposition 6.3 Let $d_0\geq 1$ be a given square-free integer. Take $f>3$ odd of the form $f =a^2+ab+b^2$ , where $\gcd (a,b)=1$ and $\gcd (d_0,f)=1$ . Set $H=\{1,a/b,b/a\}$ , a subgroup of order $3$ of the multiplicative group $({\mathbb Z}/f{\mathbb Z})^*$ . Let $N_{d_0}(f,H)$ be the rational number defined in ( 4.5 ). Then

$$ \begin{align*}N_{d_0}(f,H) =N_{d_0}(f,\{1\}) +\frac{24\mu(d_0)}{\prod_{q\mid d_0}(q+1)} \sum_{\delta\mid d_0}\frac{\delta\mu(\delta)}{\phi(\delta)}S(a,b,\delta f),\end{align*} $$

where $N_{d_0}(f,\{1\})$ is a rational number which depends only on f modulo $d_0$ , by Proposition 4.6, and where

$$ \begin{align*}S(a,b,\delta f) =\sum_{h\in ({\mathbb Z}/\delta f{\mathbb Z})^*\atop h\equiv a/b\ \ \pmod f} s(h,\delta f) =\sum_{h\in ({\mathbb Z}/\delta f{\mathbb Z})^*\atop h\equiv b/a\ \ \pmod f} s(h,\delta f).\end{align*} $$

It seems that there are no explicit formulas for $S(a,b,\delta f)$ , $S(H_\delta ,\delta f)$ , or $N_\delta (f,H)$ for $\delta>1$ (however, assuming that $b=1$ , we will obtain such formulas in Section 6.2 for $\delta \in \{2,3,6\}$ ). Instead, our aim is to prove in Proposition 6.4 that $N_\delta (f,H) =O(\sqrt f)$ for $\delta \in \{2,3,6\}$ .

Let $f>3$ be of the form $f=a^2+ab+b^2$ , with $a,b\in {\mathbb Z}$ and $\gcd (a,b)=1$ . Hence, a or b is odd. Since $a^2+ab+b^2 =a^{\prime 2} + a'b' + b^{\prime 2} = a^{\prime \prime 2} + a" b" + b^{\prime \prime 2}$ and $a'/b' =a/b$ and $a"/b" =a/b$ in $({\mathbb Z}/f{\mathbb Z})^*$ , where $(a',b') =(-b,a+b)$ and $(a",b") =(-a-b,a)$ , we may assume that both a and b are odd. Moreover, assume that $\gcd (3,f)=1$ . If $3\nmid ab$ , by swapping a and b as needed, which does not change neither H nor $S(a,b,H)$ , we may assume that $a\equiv -1\ \ \pmod 6$ and $b\equiv 1\ \ \pmod 6$ . If $3\mid ab$ , by swapping a and b and then changing both a and b to their opposites as needed, which does not change neither H nor $S(a,b,H)$ , we may assume that $a\equiv 3\ \ \pmod 6$ and $b\equiv 1\ \ \pmod 6$ . So in Proposition 6.3, we may restrict ourselves to the integers of the form

(6.2) $$ \begin{align} f> &3 \mbox{ is odd of the form } f=a^2+ab+b^2, \mbox{ with } a,b\in {\mathbb Z} \mbox{ odd } \mbox{ and } \gcd(a,b)=1\notag\\ &\mbox{and if } \gcd(3,f)=1 \mbox{ then } a\equiv -1\mbox{ or }3\ \ \pmod 6 \mbox{ and } b\equiv 1\ \ \pmod 6. \end{align} $$

Proposition 6.4 Let $\delta \in \{2,3,6\}$ be given. Let f be as in ( 6.2 ), with $\gcd (f,\delta )=1$ . Then, $s(h,\delta f) =O(\sqrt f)$ for any $h\in ({\mathbb Z}/\delta f{\mathbb Z})^*$ such that $h\equiv a/b\ \ \pmod f$ . Consequently, for a given $d_0\in \{1,2,3,6\}$ , in Proposition 6.3, we have $N_{d_0}(f,H) =O(\sqrt f)$ , and we cannot expect great improvements on these bounds, by ( 6.11 ), ( 6.13 ), and ( 6.15 ).

Proof First, by (6.1), we have

$$ \begin{align*} S(a,b,f) =s(a,b,f) =\frac{f-1}{12f}. \end{align*} $$

Second, f being odd, recalling (4.13), we have $A(2,f) =A(2,1) =0$ , $N_2(f,\{1\}) =-1$ ,

(6.3) $$ \begin{align} S(a,b,2f) =s(a,b,2f), \end{align} $$

and

$$ \begin{align*} N_2(f,H) =-1 -8S(a,b,f) +16S(a,b,2f). \end{align*} $$

Third, assume that $d_0\in \{3,6\}$ . Then $\gcd (f,3)=1$ . Hence, $f\equiv 1\ \ \pmod 6$ . Therefore, $A(3,f) =A(3,1) =4/3$ , $A(6,f) =A(6,1) =-4$ , $N_3(f,\{1\}) =N_6(f,\{1\}) =-1$ ,

$$ \begin{align*}N_3(f,H) =-1-6S(a,b,f) +9S(a,b,3f),\end{align*} $$

and

$$ \begin{align*}N_6(f,H) =-1+2S(a,b,f)-4S(a,b,2f)-3S(a,b,3f)+6S(a,b,6f).\end{align*} $$

If $a\equiv -1\ \ \pmod 6$ , $b\equiv 1\ \ \pmod 6,$ and $\delta \in \{1,2\}$ , then $\{h\in ({\mathbb Z}/3\delta f{\mathbb Z})^*;\ h\equiv a/b\ \ \pmod f\} =\{a/b,(a+2f)/b\}$ and

(6.4) $$ \begin{align} S(a,b,3\delta f) =s(a,b,3\delta f) +s(a+2f,b,3\delta f). \end{align} $$

If $a\equiv 3\ \ \pmod 6$ , $b\equiv 1\ \ \pmod 6,$ and $\delta \in \{1,2\}$ , then $\{h\in ({\mathbb Z}/3\delta f{\mathbb Z})^*;\ h\equiv a/b\ \ \pmod f\} =\{(a-\delta f)/b,(a+\delta f)/b\}$ and

(6.5) $$ \begin{align} S(a,b,3\delta f) =s(a-\delta f,b,3\delta f) +s(a+\delta f,b,3\delta f). \end{align} $$

Let us now bound the Dedekind–Rademacher sums in (6.3)–(6.5). We will need the bounds,

(6.6) $$ \begin{align} \mbox{if } f=a^2+ab+b^2, \mbox{ then } \vert a\vert+\vert b\vert\leq\sqrt{4f} \mbox{ and } \vert ab\vert\geq\sqrt{f/3}. \end{align} $$

Indeed, $4f-(\vert a\vert +\vert b\vert )^2 \geq 3(\vert a\vert -\vert b\vert )^2\geq 0$ and $f\leq a^2+a^2b^2+b^2 =3a^2b^2$ .

First, we deal with the Dedekind–Rademacher sums $s(a,b,\delta f)$ in (6.3) and (6.4), where $\delta \in \{2,3,6\}$ . Here, $\gcd (a,b) =\gcd (a,\delta f)=\gcd (b,\delta f) =1$ . Then (3.7) and (6.6) enable us to write (3.6) as follows:

$$ \begin{align*}s(a,b,\delta f)+O(\sqrt f)+O(\sqrt f) =O(\sqrt f).\end{align*} $$

Hence, in (6.3) and (6.4), we have $s(a,b,2f),\ s(a,b,3f),\ s(a,b,6f)=O(\sqrt f)$ .

Second, the remaining and more complicated Dedekind–Rademacher sums in (6.4) and (6.5) are of the form $s(a+\varepsilon \delta f,b,3\delta f)$ , where $\varepsilon \in \{\pm 1\}$ , $\delta \in \{1,2\}$ and $\gcd (a+\varepsilon \delta f,3\delta f) =\gcd (b,3\delta f)=1$ . Set $\delta ' =\gcd (a+\varepsilon \delta f,b)$ . Then $\gcd (\delta ',3\delta f)=1$ . Thus, $s(a+\varepsilon \delta f,b,3\delta f)$ $=s((a+\varepsilon \delta f)/\delta ',b/\delta ',3\delta f)$ , where now the three terms in this latter Dedekind–Rademacher are pairwise coprime. Then (3.7) and (6.6) enable us to write (3.6) as follows:

$$ \begin{align*} s((a+\varepsilon\delta f)/\delta',b/\delta',3\delta f) +O(\sqrt f) &+s(b/\delta',3\delta f,(a+\varepsilon\delta f)/\delta')\\ &\qquad\qquad =O(\delta^{\prime 2}/b) =O(b) =O(\sqrt f). \end{align*} $$

Now, $3\delta f\equiv -3\varepsilon a\ \ \pmod {a+\varepsilon \delta f}$ gives $s(b/\delta ',3\delta f,(a+\varepsilon \delta f)/\delta ') =-\varepsilon s(b/\delta ',3a,(a+\varepsilon \delta f)/\delta ')$ . Since the three rational integers in this latter Dedekind–Rademacher are pairwise coprime, the bounds (6.6) and (3.7) enable us to write (3.6) as follows:

$$ \begin{align*}s(b/\delta',3a,(a+\varepsilon\delta f)/\delta')+O(\sqrt f)+O(\sqrt f) =O(\sqrt f).\end{align*} $$

It follows that $s(a+\varepsilon \delta f,b,3\delta f) =s((a+\varepsilon \delta f)/\delta ',b/\delta ',3\delta f) =O(\sqrt f)$ , i.e., in (6.4) and (6.5), we have $s(a+2f,b,6f),\ s(a-2f,b,6f),\ s(a+2f,b,3f), s(a-f,b,3f), s(a+f,b,3f) =O(\sqrt f)$ .

Conjecture 6.5 Let $\delta $ be a given square-free integer. Let $f>3$ run over the odd integers of the form $f=a^2+ab+b^2$ with $\gcd (a,b)=1$ and $\gcd (\delta ,f)=1$ . Then $s(h,\delta f) =O(\sqrt f)$ for any $h\in ({\mathbb Z}/\delta f{\mathbb Z})^*$ such that $h\equiv a/b\ \ \pmod f$ . Consequently, for a given square-free integer $d_0$ , in Proposition 6.3, we would have $N_{d_0}(f,H) =O(\sqrt f)$ for $\gcd (d_0,f)=1$ .

Putting everything together, we obtain:

Theorem 6.6 Let $p\equiv 1\ \ \pmod 6$ be a prime integer. Let K be the imaginary subfield of degree $(p-1)/3$ of the cyclotomic field ${\mathbb Q}(\zeta _p)$ . Let H be the subgroup of order $3$ of the multiplicative group $({\mathbb Z}/p{\mathbb Z})^*$ . We have

$$ \begin{align*} M(p,H) ={\pi^2\over 6}\left (1+\frac{N(p,H)}{p}\right ) ={\pi^2\over 6}\left (1-\frac{1}{p}\right ) \mbox{ and } h_K^- \leq 2\left (\frac{p}{24}\right )^{(p-1)/12}, \end{align*} $$

and the following effective asymptotics and upper bounds

(6.7) $$ \begin{align} M_2(p,H) ={\pi^2\over 8}\left (1+\frac{N_2(p,H)}{p}\right ) ={\pi^2\over 8}\left (1+O(p^{-1/2})\right ) \mbox{ and } h_K^- \leq 2\left (\frac{p+o(p)}{32}\right )^{\frac{(p-1)}{12}}, \end{align} $$
$$ \begin{align*} M_6(p,H) ={\pi^2\over 9}\left (1+\frac{N_6(p,H)}{p}\right ) ={\pi^2\over 9}\left (1+O(p^{-1/2})\right ) \mbox{ and } h_K^- \leq 2\left (\frac{p+o(p)}{36}\right )^{\frac{(p-1)}{12}}. \end{align*} $$

Proof The formulas for $M(p,H)$ , $M_2(p,H)$ , and $M_6(p,H)$ follow from (4.4), (4.7), (6.1), and Proposition 6.4. The inequalities on $h_K^-$ are consequences as usual of (1.11) and Corollary 2.4.

6.2 The special case $p=a^2+a+1$ and $d_0=1,2,6$

Let $f>3$ be of the form $f=a^2+a+1$ , $a\in {\mathbb Z}$ . Then $\gcd (f,6)=1$ if and only if $a\equiv 0,2,3,5\ \ \pmod 6$ . We define $c_a^{\prime }$ , $c_a^{\prime \prime }$ , $c_a^{\prime \prime \prime }$ and $c_a =(-1-2c_a^{\prime } -c_a^{\prime \prime }+2c_a^{\prime \prime \prime })/12$ , as follows:

Theorem 6.7 Let $p\equiv 1\ \ \pmod 6$ be a prime integer of the form $p=a^2+a+1$ with $a\in {\mathbb Z}$ . Let K be the imaginary subfield of degree $(p-1)/3$ of the cyclotomic field ${\mathbb Q}(\zeta _p)$ . Let H be the subgroup of order $3$ of the multiplicative group $({\mathbb Z}/p{\mathbb Z})^*$ . We have

(6.8) $$ \begin{align} M_2(p,H) ={\pi^2\over 8}\left (1-(-1)^{a}\frac{2a+1}{p}\right ) \end{align} $$

and

(6.9) $$ \begin{align} M_6(p,H) ={\pi^2\over 9}\left (1+\frac{c_a}{p}\right ), \end{align} $$

showing that the error term in ( 6.7 ) is optimal.

Proof The formula (6.8) is a special case of (5.2) for $d=3$ . By (4.7), we have

$$ \begin{align*}M_6(p,H) ={\pi^2\over 9}\left (1+\frac{N_6(p,H)}{p}\right ).\end{align*} $$

Hence, (6.9) follows from Lemma 6.9.

Lemma 6.8 Let $f>3$ be of the form $f=a^2+a+1$ , $a\in {\mathbb Z}$ . Set $H=\{1,a,a^2\}$ , a subgroup of order $3$ of the multiplicative group $({\mathbb Z}/f{\mathbb Z})^*$ . We have

(6.10) $$ \begin{align} S(H,f) =\frac{f-1}{12}, \ S(H_2,2f) =\frac{2f+c_a^{\prime}}{12} \end{align} $$

and

(6.11) $$ \begin{align} N_2(f,H) =(-1)^{a-1}(2a+1). \end{align} $$

Proof Apply Lemma 5.1 with $d=3$ and $f=a^2+a+1$ to get (6.10). Then, using (4.8), we get $N_2(f,H) =-f-4S(H,f)+8S(H_2,2f) =\frac {2c_a^{\prime }+1}{3} =(-1)^{a-1}(2a+1)$ .

Lemma 6.9 Let $f>3$ be of the form $f=a^2+a+1$ , $a\in {\mathbb Z}$ . Assume that $\gcd (f,6)=1$ , i.e., that $a\equiv 0, 2, 3, 5\ \ \pmod 6$ . Set $H=\{1,a,a^2\}$ , a subgroup of order $3$ of the multiplicative group $({\mathbb Z}/f{\mathbb Z})^*$ . Then

(6.12) $$ \begin{align} S(H_3,3f) =\frac{5f+c_a^{\prime\prime}}{18}, \end{align} $$
(6.13) $$ \begin{align} N_3(f,H) =\begin{cases} -2a-1,&\mbox{if } a\equiv 0\ \ \pmod 3,\\ 2a+1,&\mbox{if } a\equiv 2 \ \ \pmod 3, \end{cases} \end{align} $$
(6.14) $$ \begin{align} S(H_6,6f) =\frac{10f+c_a^{\prime\prime\prime}}{18}, \end{align} $$

and

(6.15) $$ \begin{align} N_6(f,H) =\begin{cases} -2a-1,&\mbox{if } a\equiv 0\ \ \pmod 6,\\ -3,&\mbox{if } a\equiv 2,3\ \ \pmod 6,\\ 2a+1,&\mbox{if } a\equiv 5\ \ \pmod 6. \end{cases} \end{align} $$

Proof Let us, for example, detail the computation of $S(H_6,6f)$ in the case that $a\equiv 0\ \ \pmod 6$ . We have $f\equiv 1\ \ \pmod 6$ and $H_6=\{1,1+4f,a+f,a+5f,a^2+f,a^2+5f\}$ . Since $a^2+f =(a+f)^{-1}$ and $a^2+5f=(a+5f)^{-1}$ in $({\mathbb Z}/f{\mathbb Z})^*$ , we have $S(H_6,6f) =s(1,6f)+s(1+4f,6f)+2s(a+f,6f)+2s(a+5f,6f)$ , by (3.2). Using (3.3) and (3.4), we obtain $s(1,6f) =\frac {18f^2-9f+1}{36f}$ , $s(1+4f,6f) =\frac {2f^2-13f+1}{36f}$ , $s(a+f,6f) =-\frac {(3a-21)f+1}{72f}$ , and $s(a+5f,6f) =-\frac {(35a+19)f+1}{72f}$ . Formula (6.14) follows.

By (4.8), we have

$$ \begin{align*}N_3(f,H) =-f-3S(H,f)+\frac{9}{2}S(H_3,3f)\end{align*} $$

and

$$ \begin{align*}N_6(f,H) =-f +S(H,f)-2S(H_2,2f)-\frac{3}{2}S(H_3,3f)+3S(H_6,6f).\end{align*} $$

Using (6.1), (6.10), and (6.12), we obtain $N_3(f,H) =\frac {c_a^{\prime \prime }+1}{4}$ and (6.13). Using (6.1), (6.10), (6.12), and (6.14), we obtain $N_6(f,H) =\frac {-1-2c_a^{\prime }-c_a^{\prime \prime }+2c_a^{\prime \prime \prime }}{12} =c_a$ and (6.15).

7 Conclusion and a conjecture

The proof of Lemma 5.1 gives for $d\geq 3$ odd and $a\neq 0,\pm 1$

(7.1) $$ \begin{align} s\left (a,\frac{a^d-1}{a-1}\right ) =\frac{(f-1)(f-a^2-1)}{12af} =O\left (f^{1-\frac{1}{d-1}}\right ). \end{align} $$

Our numerical computations suggest the following stronger version of Theorem 3.1:

Conjecture 7.1 There exists $C>0$ such that for any $d>2$ dividing $p-1$ and any h of order d in the multiplicative group $({\mathbb Z}/p{\mathbb Z})^*$ , we have

(7.2) $$ \begin{align} \vert s(h,p)\vert \leq Cp^{1-\frac{1}{\phi(d)}}. \end{align} $$

Indeed, for $p\leq 2\cdot 10^5$ , we checked on a desk computer that any $d>2$ dividing $p-1$ and any h of order d in the multiplicative group $({\mathbb Z}/p{\mathbb Z})^*$ , we have

$$ \begin{align*}Q(h,p) :=\frac{\vert s(h,p)\vert}{p^{1-\frac{1}{\phi(d)}}} \leq Q(2,2^7-1) =0.08903\ldots.\end{align*} $$

The estimate (7.2) would allow to slightly extend the range of validity of Theorem 1.1 to $d \leq (1-\varepsilon )\frac {\log p}{\log \log p}$ . Moreover, the choice $a=2$ in (7.1) for which $s(2,f)$ is asymptotic to $\frac {1}{24}f$ with $f=2^d-1$ shows that $s(h,p) =o(p)$ cannot hold true in the range $d \asymp \log p$ . Notice that we cannot expect a better bound than (7.2), by (7.1). Finally, the restriction that p be prime in (7.2) is paramount by Remark 6.2 where $s(a,f) \sim f^{2/3}/12$ for a of order $3$ in $({\mathbb Z}/(a^3-1){\mathbb Z})^*$ .

Footnotes

This work was supported by the Ministero della Istruzione e della Ricerca Young Researchers Program Rita Levi Montalcini (M.M.).

1 This restriction on d is probably optimal, by (5.1).

2 Note the misprint in the exponent in [Reference LouboutinLou16, equation (8)].

3 Note that $\Pi _{d_0}(f,H)\in {\mathbb Q}_+^*$ , by Lemma 2.3.

4 In [Reference SunSun], the author uses lattice point interpretation to study sums with a similar flavor.

5 Note the misprint in the exponent in [Reference LouboutinLou16, Theorem 5].

6 The formulas can be and have been checked on numerous examples using a computer algebra system. Indeed, by (3.3) and (3.4), any Dedekind sum $s(c,d)\in {\mathbb Q}$ with $c,d\geq 1$ can be easily computed by successive euclidean divisions of c by d and exchanges of c and d, until we reach $c=1$ .

References

Apostol, M. T., Modular functions and Dirichlet series in number theory, Graduate Texts in Mathematics, 41, Springer, New York, 1976.CrossRefGoogle Scholar
Bayad, A. and Raouj, A., Reciprocity formulae for multiple Dedekind–Rademacher sums . C. R. Math. Acad. Sci. Paris 349(2011), 131136.CrossRefGoogle Scholar
Conrey, J. B., Fransen, E., and Klein, R., Mean values of Dedekind sums . J. Number Theory 56(1996), 214226.CrossRefGoogle Scholar
Drmota, M. and Tichy, R. F., Sequences, discrepancies and applications, Lecture Notes in Mathematics, 1651, Springer, Berlin, 1997.CrossRefGoogle Scholar
Elma, E., On a problem related to discrete mean values of Dirichlet  $L$ -functions . J. Number Theory 217(2020), 3643.CrossRefGoogle Scholar
Feng, K. Q., On the first factor of the class number of a cyclotomic field . Proc. Amer. Math. Soc. 84(1974), 479482.CrossRefGoogle Scholar
Girstmair, K., Zones of large and small values for Dedekind sums . Acta Arith. 109(2003), 299308.CrossRefGoogle Scholar
Granville, A., On the size of the first factor of the class number of a cyclotomic field . Invent. Math. 100(1990), 321338.CrossRefGoogle Scholar
Hickerson, D., Continued fractions and density results for Dedekind sums . J. Reine Angew. Math. 290(1977), 113116.Google Scholar
Korobov, N. M., Some problems in the theory of Diophantine approximation . Russian Math. Surveys 22(1967), no. 3, 80118.CrossRefGoogle Scholar
Konyagin, S. V. and Shparlinski, I. E., Character sums with exponential functions and their applications, Cambridge Tracts in Mathematics, 136, Cambridge University Press, Cambridge, 1999.CrossRefGoogle Scholar
Lee, S. H. and Lee, S., On the twisted quadratic moment for Dirichlet L-functions . J. Number Theory 174(2017), 427435.CrossRefGoogle Scholar
Lee, S. H. and Lee, S., The twisted moments and distribution of values of Dirichlet L-functions at 1 . J. Number Theory 197(2019), 168184.CrossRefGoogle Scholar
Louboutin, S. and Munsch, M., Second moment of Dirichlet  $L$ -functions, character sums over subgroups and upper bounds on relative class numbers . Quart. J. Math. 72(2021), 13791399.CrossRefGoogle Scholar
Louboutin, S., Quelques formules exactes pour des moyennes de fonctions $L$ de Dirichlet. Canad. Math. Bull . 36(1993), 190196. Addendum. Canad. Math. Bull. 37(1994), 89.CrossRefGoogle Scholar
Louboutin, S., Mean values of  $L$ -functions and relative class numbers of cyclotomic fields . Publ. Math. Debrecen 78(2011), 647658.CrossRefGoogle Scholar
Louboutin, S., Twisted quadratic moments for Dirichlet  $L$ -functions . Bull. Korean Math. Soc. 52(2015), 20952105.CrossRefGoogle Scholar
Louboutin, S., Dedekind sums, mean square value of $L$ -functions at $s=1$ and upper bounds on relative class numbers. Bull. Pol. Acad. Sci. Math. 64(2016), 165174.CrossRefGoogle Scholar
Louboutin, S., On the denominator of Dedekind sums . Bull. Korean Math. Soc. 56(2019), 815827.Google Scholar
Louboutin, S., Mean square value of $L$ -functions at $s=1$ for non-primitive characters, Dedekind sums and bounds on relative class numbers. Funct. Approx. Comment. Math. (2022), 112. 10.7169/facm/2027 Google Scholar
Metsänkylä, T., Class numbers and $\mu$ -invariants of cyclotomic fields . Proc. Amer. Math. Soc. 43(1974), 299300.Google Scholar
Murty, R. and Petridis, Y., On Kummer’s conjecture . J. Number Theory 90(2001), 294303.CrossRefGoogle Scholar
Niederreiter, H., Pseudo-random numbers and optimal coefficients . Adv. Math. 26(1977), no. 2, 99181.CrossRefGoogle Scholar
Qi, M.-G., A class of mean square formulas for  $L$ -functions (Chinese English summary) . J. Tsinghua Univ. 31(1991), no. 3, 3441.Google Scholar
Rademacher, H., Generalization of the reciprocity formula for Dedekind sums . Duke Math. J. 21(1954), 391397.CrossRefGoogle Scholar
Rademacher, H. and Grosswald, E., Dedekind sums, The Carus Mathematical Monographs, 16, The Mathematical Association of America, Washington, DC, 1972.CrossRefGoogle Scholar
Serre, J.-P., A course in arithmetic, Graduate Texts in Mathematics, 7, Springer, New York and Heidelberg, 1973.CrossRefGoogle Scholar
Sun, Z.-W., Sums of minima and maxima . Discrete Math. 257(2002), 143159.CrossRefGoogle Scholar
Vardi, I., Dedekind sums have a limiting distribution . Int. Math. Res. Not. 1993(1993), 112.CrossRefGoogle Scholar
Walum, H., An exact formula for an average of  $L$ -series . Illinois J. Math. 26(1982), 13.CrossRefGoogle Scholar
Washington, L. C., Introduction to cyclotomic fields, 2nd ed., Graduate Texts in Mathematics, 83, Springer, New York, 1997.CrossRefGoogle Scholar