Keywords

1 Introduction

Since the non-trivial zeros of the Riemann zeta function \(\zeta (z) \), until now found, lie on the line \(\mathfrak {R}z=1/2\) (the assertion that all them are situated on that line is the Riemann Hypothesis) and the trivial ones are on the real axis (they are the negative even numbers [9, p. 8]), it seems that the zeros of \(\zeta (z)\) are situated on those two perpendicular lines. However that is not so for the zeros of the partial sums \(\zeta _{n}(z):=\sum _{j=1}^{n}j^{-z}\) of the series \(\sum _{j=1}^{\infty }j^{-z}\) that defines the Riemann zeta function \(\zeta (z)\) on the half-plane \(\mathfrak {R}z>1\). Indeed, except for \(\zeta _{2}(z)\) whose zeros all are imaginary (it is immediate to check that the zeros of \(\zeta _{2}(z)\) are \(z_{2,j}=\frac{(2j+1)\pi i}{\log 2}\), \(j\in \mathbb {Z}\)), so aligned, the zeros of each \(\zeta _{n}(z)\) for any \(n>2\) are dispersed in a vertical strip forming a sort of cloud, more or less uniform, that extends up, down and left as n increases, whereas at the right the cloud of zeros is upper bounded (essentially) by the line \(\mathfrak {R}z=1\) (see Fig. 1).

Fig. 1
figure 1

Graphs of the zeros of \(\zeta _{n}(z)\) for some values of n, with \(\mathfrak {R}z\in [-3,1]\) and \(\mathfrak {I}z\in [0,5000]\)

An explanation grosso modo why the zeros of the \(\zeta _{n}(z)\)’s are distributed of such a form is supported by the following facts:

(a) Any exponential polynomial (EP for short) of the form

$$\begin{aligned} P(z):=1+\sum _{j=1}^{N}a_{j}e^{-z\lambda _{j}},\quad z\in \mathbb {C},\quad a_{j}\in \mathbb {C}\setminus \{ 0\} ,\quad 0<\lambda _{1}<\ldots <\lambda _{N},\quad N\ge 1, \end{aligned}$$
(1)

has zeros as a consequence of Hadamard’s Factorization Theorem or from Pólya’s Theorem [13, p. 71]. For \(N=1\), it is immediate that an EP of the form (1) has its zeros aligned. For \(N>1\), noticing that for any y,

$$ \lim _{x\rightarrow +\infty }P(z)=\lim _{x\rightarrow -\infty }Q(z)=1, $$

where \(Q(z):=a_{N}^{-1}e^{z\lambda _{N}}P(z)\) (observe that P(z) and Q(z) have exactly the same zeros), it follows that the zeros of P(z) are situated in a vertical strip. Therefore, for every EP P(z) of the form (1), there exist two real numbers

$$\begin{aligned} a_{P(z)}:=\inf \{ \mathfrak {R}z:P(z)=0\} ,\quad b_{P(z)}:=\sup \{ \mathfrak {R}z:P(z)=0\} , \end{aligned}$$
(2)

that define an interval \([ a_{P(z)},b_{P(z)}] \), called critical interval associated with P(z). Therefore the set \([a_{P(z)},b_{P(z)}] \times \mathbb {R}\), called critical strip associated with P(z), is the minimal vertical strip that contains all the zeros of P(z).

It is immediate that any partial sum \(\zeta _{n}(z):=\sum _{j=1}^{n}j^{-z}\), \(n\ge 2\), is an EP of the form (1). Therefore the zeros of each \(\zeta _{n}(z)\) are situated on its critical strip \([ a_{\zeta _{n}(z)},b_{\zeta _{n}(z)}] \times \mathbb {R}\) (a detailed proof on the existence of the zeros of \(\zeta _{n}(z)\) and their distribution with respect to the imaginary axis can be found in [14, Prop. 1, 2, 3]). Regarding the bounds \(a_{\zeta _{n}(z)}\), \(b_{\zeta _{n}(z)}\), taking into account that all the zeros of \(\zeta _{2}(z)\) lie on the imaginary axis, we get the property

$$\begin{aligned} a_{\zeta _{2}(z)}=b_{\zeta _{2}(z)}=0;\quad a_{\zeta _{n}(z)}<0<b_{\zeta _{n}(z)},\quad n>2, \end{aligned}$$
(3)

that will be proved below in Lemma 2.3, Part (ii). A much more precise estimation of such bounds is given by the formulas:

$$\begin{aligned} b_{\zeta _{n}(z)}=1+\left( \frac{4}{\pi }-1+o(1)\right) \frac{\log \log n}{\log n},\quad n\rightarrow \infty , \end{aligned}$$
(4)

obtained by Montgomery and Vaughan [12] in 2001, by completing a previous work of Montgomery [11] of 1983, and

$$\begin{aligned} a_{\zeta _{n}(z)}=-\frac{\log 2}{\log ( \frac{n-1}{n-2}) }+\varDelta _{n},\quad \lim \sup _{n\rightarrow \infty }| \varDelta _{n}| \le \log 2, \end{aligned}$$
(5)

found by Mora [17] in 2015. Consequently, from (5) and (4), we have

$$ \lim _{n\rightarrow \infty }a_{\zeta _{n}(z)}=-\infty ,\quad \lim _{n\rightarrow \infty }b_{\zeta _{n}(z)}=1, $$

what justifies the fact of the cloud of zeros of \(\zeta _{n}(z)\) moves to the left as n increases but not to the right, where the cloud is upper bounded (essentially) by the line \(\mathfrak {R}z=1\) (it does not mean that some \(\zeta _{n}(z)\) can have zeros with real part greater than 1; in fact, many works prove the existence of such zeros [10, 22, 23, 25], among others).

(b) Since the zeros of an analytic function are isolated, and all the \(\zeta _{n}(z)\)’s are entire functions, by taking into account the real parts of the zeros of each \(\zeta _{n}(z)\) are bounded (the real parts are contained in the critical interval \([ a_{\zeta _{n}(z)},b_{\zeta _{n}(z)}] \) for every fixed n), their imaginary parts cannot be. Furthermore, as the coefficients of every \(\zeta _{n}(z)\) are real, its zeros are conjugate. Consequently the zeros of the \(\zeta _{n}(z)\)’s are located up and down, symmetrically with respect to the real axis.

(c) From (3) we deduce that, for any \(n>2\), \(\zeta _{n}(z)\) has zeros with positive and negative real parts.

With the aim to understand what law controls the distribution of the real projections of the zeros of \(\zeta _{n}(z)\), we introduce a Fixed Point Theory focused on two real functions, \(f_{n}\) and \(g_{n}\), for every \(n>2\). Firstly, such functions, by virtue of a recent result [19, Theorem 3], allow us to have an easy characterization of the sets

$$\begin{aligned} R_{\zeta _{n}(z)}:=\overline{\{ \mathfrak {R}z:\zeta _{n}(z)=0\} }. \end{aligned}$$
(6)

Secondly, among others relevant results deduced from the fixed point properties of \(f_{n}\) and \(g_{n}\), we stress those that characterize some notable arithmetic sets such as \(\mathscr {P}^{*}\) and \(\mathscr {C}^{*}\), the set of primes greater than 2 and the set of composite numbers greater than 2, respectively. In this way, we show how close the arithmetic sets \(\mathscr {P}^{*}\) and \(\mathscr {C}^{*}\) from the law of the distribution of the zeros of the partial sums of the Riemann zeta function are. Furthermore, our point fixed theory proves the existence of a minimal density interval for each \(\zeta _{n}(z)\), that is, a closed interval \([ A_{n},b_{\zeta _{n}(z)}] \), with \(a_{\zeta _{n}(z)}\le A_{n}<b_{\zeta _{n}(z)}\) contained in the set \(R_{\zeta _{n}(z)} \), for any integer \(n>2\), which means that there is no vertical sub-trip contained in \([ A_{n},b_{\zeta _{n}(z)}] \times \mathbb {R}\) zero-free for \(\zeta _{n}(z)\). Then, since it is always true that \(R_{\zeta _{n}(z)}\subset [ a_{\zeta _{n}(z)},b_{\zeta _{n}(z)}] \), when the bound \(A_{n}\) coincides with \(a_{\zeta _{n}(z)}\) it follows that \(R_{\zeta _{n}(z)}=[ a_{\zeta _{n}(z)},b_{\zeta _{n}(z)}] \). In this case we will say that \(\zeta _{n}(z)\) has a maximum density interval, and it is exactly the critical interval \([ a_{\zeta _{n}(z)},b_{\zeta _{n}(z)}] \). Finally, we will give a sufficient condition in terms of the quantity of fixed points of \(f_{n}\) for \(\zeta _{n}(z)\) have a maximum density interval.

2 The Functions \(f_{n}\) and \(g_{n}\)

The functions \(f_{n}\) and \(g_{n}\) that we are going to introduce below, are directly linked to the interval of variation of the variable x of the Cartesian equation of an analytic variety associated with the \(n\mathrm{th}\)-partial sum \(\zeta _{n}(z):=\sum _{j=1}^{n}j^{-z}\), \(n>2\). First we consider the EP

$$\begin{aligned} \zeta _{n}^{*}(z):=\zeta _{n}(z)-p_{k_{n}}^{-z},\quad n>2, \end{aligned}$$
(7)

where \(p_{k_{n}}\) is the last prime not exceeding n. The bounds \(a_{\zeta _{n}^{*}(z)}\), \(b_{\zeta _{n}^{*}(z)}\) defined in (2) corresponding to \(\zeta _{n}^{*}(z)\) satisfy the following crucial property (for details see [16, Theorem 15]) :

$$\begin{aligned} a_{\zeta _{n}^{*}(z)}=b_{\zeta _{n}^{*}(z)}=0,\quad \text {for }n=3,4;\quad a_{\zeta _{n}^{*}(z)}<0<b_{\zeta _{n}^{*}(z)},\quad \text {for all }n>4. \end{aligned}$$
(8)

Now our objective is to analyse the behavior of the end-points of the interval of variation of the variable x of the analytic variety, or level curve [24, p. 121], of equation

$$\begin{aligned} | \zeta _{n}^{*}(z)| =p_{k_{n}}^{-c},\quad n>2,\quad c\in \mathbb {R}. \end{aligned}$$
(9)

To do it, we square (9) and by using elementary formulas of trigonometry we obtain the Cartesian equation of \(| \zeta _{n}^{*}(z)| =p_{k_{n}}^{-c}\), namely

$$\begin{aligned} \begin{array}{l} \displaystyle \sum _{j=1,j\ne p_{k_{n}}}^{n}j^{-2x}+2\cdot 1^{-x}\sum _{j=2,j\ne p_{k_{n}}}^{n}j^{-x}\cos ( y\log (\frac{j}{1})) + \\ \\ \displaystyle 2\cdot 2^{-x}\sum _{j=3,j\ne p_{k_{n}}}^{n}j^{-x}\cos ( y\log (\frac{j}{2})) +\cdots + \\ \\ \displaystyle 2(n-1)^{-x}\sum _{j=n,j\ne p_{k_{n}}}^{n}j^{-x}\cos ( y\log (\frac{j}{n-1})) =p_{k_{n}}^{-2c}. \end{array} \end{aligned}$$
(10)

It is immediate to see that for any value of y, the left-hand side of (10) tends to \(+\infty \) as \(x\rightarrow -\infty \). Then, as the right-hand side of (10) is a constant, the variation of x is always lower bounded by a number denoted by \(a_{n,c}\). On the other hand, the limit of the left-hand side of (10) is 1 when \(x\rightarrow +\infty \). Then, if \(c\ne 0\), the variation of x is upper bounded by a number denoted by \(b_{n,c}\). Therefore, fixed an integer \(n>2\), we have:

If \(c\ne 0\), the variable x in the Eq. (10) varies on an open interval \(( a_{n,c},b_{n,c}) \) satisfying the properties: (a) Given \(x\in ( a_{n,c},b_{n,c}) \), there is at least a point of the level curve \(| \zeta _{n}^{*}(z)| =p_{k_{n}}^{-c}\) with abscissa x. Exceptionally \(| \zeta _{n}^{*}(z)| =p_{k_{n}}^{-c}\) could have points of abscissas \(a_{n,c}\), \(b_{n,c}\). In this case we will say that \(a_{n,c}\), \(b_{n,c}\) are accessible. Otherwise the lines \(x=a_{n,c}\), \(x=b_{n,c}\) are asymptotes of the variety. (b) For \(x<a_{n,c}\) or \(x>b_{n,c}\) there is no point of the variety \(| \zeta _{n}^{*}(z)| =p_{k_{n}}^{-c}\).

If \(c=0\), x varies on \(( a_{n,0},+\infty ) \), so \(b_{n,0} \) can be defined as \(+\infty \), satisfying: (c) Given \(x\in ( a_{n,0},+\infty ) \), there is at least a point of the variety \(| \zeta _{n}^{*}(z)| =1\) with abscissa x. If there is a point of \(| \zeta _{n}^{*}(z)| =1\) with abscissa \(a_{n,0}\), we will say that \(a_{n,0}\) is accessible. Otherwise the line \(x=a_{n,0}\) is an asymptote of the variety. (d) For \(x<a_{n,0}\) there is no point of \(| \zeta _{n}^{*}(z)| =1\).

We show in Fig. 2 the varieties \(| \zeta _{n}^{*}(z)|=p_{k_{n}}^{-c}\) for \(n=3\) and some values of c.

Fig. 2
figure 2

Graphs of the varieties \(| \zeta _{3}^{*}(z)|=3^{-c}\) for some values of c

The end-points \(a_{3,c}\), \(b_{3,c}\) corresponding to the variety \(|\zeta _{3}^{*}(z)| =p_{k_{3}}^{-c}\) can be easily determined by a completely similar way to those of the variety \(| \zeta _{3}^{*}(-z)| =p_{k_{3}}^{c}\) (see [8, p. 49]). Each bound \(a_{3,c}\), \(b_{3,c}\) as a function of c is given by the formulas

$$\begin{aligned} a_{3,c}=-\frac{\log (1+3^{-c})}{\log 2},\quad c\in \mathbb {R}; \quad b_{3,c}=\left\{ \begin{array}{c} -\frac{\log (3^{-c}-1)}{\log 2},\quad \text {if }c<0 \\ \\ -\frac{\log (1-3^{-c})}{\log 2},\quad \text {if }c>0 \end{array} \right. . \end{aligned}$$
(11)

By virtue of above considerations (a), (b), (c), (d), and by using an elementary geometric reasoning, similar to that it was used to find the graphs of \(| \zeta _{n}^{*}(-z)| =p_{k_{n}}^{c}\) (see [16, Proposition 8]), the graphs of the varieties \(| \zeta _{n}^{*}(z)| =p_{k_{n}}^{-c}\) are described in the next result.

Proposition 2.1

Fixed an integer \(n>2\), we have:

  1. (i)

    If \(c>0\), \(| \zeta _{n}^{*}(z)| =p_{k_{n}}^{-c}\) has infinitely many arc-connected components which are closed curves and x varies on a finite interval \(( a_{n,c},b_{n,c}) \), where \(a_{n,c},b_{n,c}\) could be accessible.

  2. (ii)

    If \(c=0\), \(| \zeta _{n}^{*}(z)| =1\) has infinitely many arc-connected components which are open curves with horizontal asymptotes of equations \(y=( 2j+1) \frac{\pi }{2\log 2}\), \(j\in \mathbb {Z}\), and x varies on the infinite interval \(( a_{n,0},+\infty ) \), where \(a_{n,0}\) could be accessible.

  3. (iii)

    If \(c<0\), \(| \zeta _{n}^{*}(z)| =p_{k_{n}}^{-c}\) has only one arc-connected component which is an open curve; x varies on a finite interval \((a_{n,c},b_{n,c}) \), where \(a_{n,c},b_{n,c}\) could be accessible. The variable y takes all real values. Furthermore, \(| \zeta _{n}^{*}(z)| =p_{k_{n}}^{-c}\) intersects the real axis at a unique point of abscissa \(b_{n,c}\), so \(b_{n,c}\) is always accessible when \(c<0\).

In Fig. 3 we show the graph of \(| \zeta _{n}^{*}(z)|=p_{k_{n}}^{-c}\) for some values of \(n>3\) and c.

Fig. 3
figure 3

Graphs of the varieties \(| \zeta _{7}^{*}(z)|=7^{-c}\) and \(| \zeta _{12}^{*}(z)|=11^{-c}\) for some values of c

From Proposition 2.1, a simple consequence is deduced:

Corollary 2.1

Fixed an integer \(n>2\), if \(u\in \mathbb {C}\) satisfies \(| \zeta _{n}^{*}(u)| <p_{k_{n}}^{-c}\) (in this case we will say that u is an interior point of the variety \(| \zeta _{n}^{*}(z)| =p_{k_{n}}^{-c}\)), then there exists a point w of \(| \zeta _{n}^{*}(z)|=p_{k_{n}}^{-c}\), so \(a_{n,c}\le \mathfrak {R}w\le \) \(b_{n,c}\), such that \(\mathfrak {R}w<\mathfrak {R}u\).

Definition 2.1

Given an integer \(n>2\), we define the real functions

$$\begin{aligned} f_{n}(c):=a_{n,c},\quad c\in \mathbb {R};\quad \quad g_{n}(c):=b_{n,c},\quad c\in \mathbb {R}\setminus \{ 0\} , \end{aligned}$$
(12)

where \(a_{n,c}\), \(b_{n,c}\) are the end-points of the interval of variation of the variable x in the Eq. (10).

We show in Fig. 4 the graph of the functions \(f_{3}(c)\) and \(g_{3}(c)\), defined by the Eq. (11), and the function \(f_{4}(c)\).

Fig. 4
figure 4

Left: Graph of the functions \(f_{3}(c)\) (blue), \(g_{3}(c)\) (red) and \(y=x\) (plotted). Right: Graph of the function \(f_{4}(c)\) (blue) and \(y=x\) (plotted)

Since \(| \zeta _{n}^{*}(z)| =p_{k_{n}}^{-d}\) tends to \(| \zeta _{n}^{*}(z)| =p_{k_{n}}^{-c}\) as d tends to c, it is immediate that \(f_{n}\), \(g_{n}\) are both continuous on \(\mathbb {R}\setminus \{ 0\} \), and \(f_{n}\) is continuous on whole of \(\mathbb {R}\). For \(c=0\), by Part (ii) of Proposition 1 we can agree \(b_{n,0}=+\infty \), and then we should define \(g_{n}(0):=\) \(+\infty \).

Now we are ready to give a characterization of the set \(R_{\zeta _{n}(z)}\), defined in (6), by using the functions \(f_{n}\) and \(g_{n}\).

Theorem 2.1

Let \(n>2\) be a fixed integer. A real number \(c\in R_{\zeta _{n}(z)}\) if and only if

$$\begin{aligned} f_{n}(c)\le c\le g_{n}(c). \end{aligned}$$
(13)

Proof

If c \(\in \) \(R_{\zeta _{n}(z)}\), there exists a sequence \((z_{m}) _{m=1,2,\ldots }\) of zeros of \(\zeta _{n}(z)\) such that \(\lim _{m\rightarrow \infty }\mathfrak {R}z_{m}=c\). From (7), \(\zeta _{n}^{*}(z_{m})=-p_{k_{n}}^{-z_{m}}\) for each \(m=1,2,\ldots \). By taking the modulus, we have \(| \zeta _{n}^{*}(z_{m})| =p_{k_{n}}^{-x_{m}}\), where \(x_{m}:=\mathfrak {R}z_{m}\). This means that each \(z_{m}\) is a point of the variety \(| \zeta _{n}^{*}(z)| =p_{k_{n}}^{-x_{m}}\), so \(x_{m}\in [ a_{n,x_{m}},b_{n,x_{m}}] \) and then we get

$$ f_{n}(x_{m})=a_{n,x_{m}}\le x_{m}\le b_{n,x_{m}}=g_{n}(x_{m}),\quad \text {for all }m. $$

Now by taking the limit when \(m\rightarrow \infty \), noticing that \(\lim _{m\rightarrow \infty }x_{m}=c\), because of the continuity of \(f_{n}\) and \(g_{n}\), the inequalities (13) follow. Conversely, if \(f_{n}(c)<c<g_{n}(c)\), by taking into account the definitions of \(f_{n}\), \(g_{n}\), the value c is in the interval of variation of x of the variety \(| \zeta _{n}^{*}(z)| =p_{k_{n}}^{-c}\) and then the line \(x=c\) intersects the variety. Therefore, by applying [16, Theorem 3], \(c\in \) \(R_{\zeta _{n}(z)}\). If \(f_{n}(c)=c\) or \(g_{n}(c)=c\), the line \(x=c\) intersects the variety \(| \zeta _{n}^{*}(z)| =p_{k_{n}}^{-c}\) provided that \(a_{n,c}\) or \(b_{n,c}\) be accessible. Otherwise the line \(x=c\) is an asymptote of \(| \zeta _{n}^{*}(z)| =p_{k_{n}}^{-c}\). Therefore, in both cases, again by [19, Theorem 3], the point \(c\in \) \(R_{\zeta _{n}(z)}\).       \(\square \)

As we can easily check, the function \(f_{3}(c):=a_{3,c}\), with \(a_{3,c}\) given in (11), is strictly increasing; this property is true for all the functions \(f_{n}(c)\), \(n>2\), defined in (12), as we prove below.

Lemma 2.1

For every integer \(n>2\), \(f_{n}\) is a strictly increasing function on \(\mathbb {R}\).

Proof

Firstly, for each fixed c \(\in \mathbb {R}\), we claim that \(f_{n}\) satisfies

$$\begin{aligned} \inf \{ | \zeta _{n}^{*}(f_{n}(c)+iy)| :y\in \mathbb {R}\} =p_{k_{n}}^{-c}. \end{aligned}$$
(14)

Indeed, we put \(\lambda _{n,c}:=\inf \{ | \zeta _{n}^{*}(f_{n}(c)+iy)| :y\in \mathbb {R}\} \). By assuming \(\lambda _{n,c}<p_{k_{n}}^{-c}\), there exists a point \(z_{c}:=f_{n}(c)+iy_{c}\) such that

$$ \lambda _{n,c}\le | \zeta _{n}^{*}(z_{c})|<p_{k_{n}}^{-c} , $$

and then it means that \(z_{c}\) is an interior point of \(| \zeta _{n}^{*}(z)| =p_{k_{n}}^{-c}\). By Corollary 2.1 there exists w belonging to the variety \(| \zeta _{n}^{*}(z)|=p_{k_{n}}^{-c}\), so \(a_{n,c}\le \mathfrak {R}w\le \) \(b_{n,c}\), such that \(\mathfrak {R}w<\mathfrak {R}z_{c}=f_{n}(c)=a_{n,c}\). But this is a contradiction, and then necessarily

$$\begin{aligned} \lambda _{n,c}\ge p_{k_{n}}^{-c}. \end{aligned}$$
(15)

For \(\varepsilon >0\) sufficiently small, we consider the strip

$$ S_{\varepsilon }:=\{ z\in \mathbb {C}:a_{n,c}\le \mathfrak {R}z<a_{n,c}+\varepsilon \} , $$

and put

$$ \lambda _{n,c,\varepsilon }:=\inf \{ | \zeta _{n}^{*}(z)| :z\in S_{\varepsilon }\} . $$

From the definition of \(a_{n,c}\), the set \(S_{\varepsilon }\) contains infinitely many points of the variety \(| \zeta _{n}^{*}(z)| =p_{k_{n}}^{-c}\). Then \(\lambda _{n,c,\varepsilon }\) \(\le p_{k_{n}}^{-c}\) for all \(\varepsilon >0\), so \(\lambda _{n,c}\le p_{k_{n}}^{-c}\). Therefore, according to (15), \(\lambda _{n,c}=p_{k_{n}}^{-c}\) and then (14) follows. Let d be a real number such that \(d<c\), so \(p_{k_{n}}^{-d}>p_{k_{n}}^{-c}\). Let \(\eta \) be such that \(0<\eta <p_{k_{n}}^{-d}-p_{k_{n}}^{-c}\). From (14), there exists some point \(z_{\eta }:=f_{n}(c)+iy_{\eta }\) such that

$$ p_{k_{n}}^{-c}\le | \zeta _{n}^{*}(z_{\eta })|<p_{k_{n}}^{-c}+\eta <p_{k_{n}}^{-d}, $$

so \(z_{\eta }\) is interior of \(| \zeta _{n}^{*}(z)|=p_{k_{n}}^{-d}\). By Corollary 2.1, there exists a point \(w_{\eta }\) of \(| \zeta _{n}^{*}(z)| =p_{k_{n}}^{-d}\), so \(a_{n,d}\le \mathfrak {R}w_{\eta }\le b_{n,d}\), such that \(\mathfrak {R}w_{\eta }<\mathfrak {R}z_{\eta }\). Then

$$ f_{n}(d)=a_{n,d}\le \mathfrak {R}w_{\eta }<\mathfrak {R}z_{\eta }=f_{n}(c), $$

which definitely proves the lemma.       \(\square \)

In the next result we prove that \(f_{n}\) is upper bounded by the number \(a_{\zeta _{n}^{*}(z)}\) defined in (2) corresponding to the EP \(\zeta _{n}^{*}(z)\), defined in (7).

Lemma 2.2

For every \(n>2\), the function \(f_{n}\) satisfies

$$ f_{n}(c)<a_{\zeta _{n}^{*}(z)}\text { for any }c\in \mathbb {R}. $$

Proof

Let c be an arbitrary real number. By taking into account the definition of \(a_{\zeta _{n}^{*}(z)}\), there exists a sequence \(( z_{m})_{m=1,2,\ldots }\)of zeros of \(\zeta _{n}^{*}(z)\), with \(\mathfrak {R}z_{m}\ge a_{\zeta _{n}^{*}(z)}\), such that

$$\begin{aligned} \lim _{m\rightarrow \infty }\mathfrak {R}z_{m}=a_{\zeta _{n}^{*}(z)}. \end{aligned}$$
(16)

Since \(\zeta _{n}^{*}(z_{m})=0\), we get \(| \zeta _{n}^{*}(z_{m})| <p_{k_{n}}^{-c}\), for all m. Then, from Corollary 2.1, there exists a sequence \(( w_{m}) _{m=1,2,\ldots }\) of points of \(| \zeta _{n}^{*}(z)| =p_{k_{n}}^{-c}\), so \(a_{n,c}\le \mathfrak {R}w_{m}\le b_{n,c}\), such that \(\mathfrak {R}w_{m}<\mathfrak {R}z_{m}\), for all m. Therefore, since \(f_{n}(c)=a_{n,c}\), we have

$$ f_{n}(c)\le \mathfrak {R}w_{m}<\mathfrak {R}z_{m},\quad \text {for all }m. $$

Now, by taking the limit in the above inequality when \(m\rightarrow \infty \), by (16), we get

$$ f_{n}(c)\le a_{\zeta _{n}^{*}(z)}\quad \text {for any }c\in \mathbb {R}, $$

implying, noticing that by Lemma 2.1 \(f_{n}\) is strictly increasing, that \(f_{n}(c)<a_{\zeta _{n}^{*}(z)}\) for any \(c\in \mathbb {R}\).       \(\square \)

For every \(n>2\), let \(a_{\zeta _{n}(z)}\), \(b_{\zeta _{n}(z)}\) be the bounds, defined in (2), corresponding to the EP \(\zeta _{n}(z)\). The function \( g_{n} \), defined in (12), has the following properties.

Lemma 2.3

For every \(n>2\), the function \(g_{n}\) satisfies:

  1. (i)

    \(g_{n}\) is strictly increasing on \(( -\infty ,0) \) and decreasing on \((0,+\infty ) \).

  2. (ii)

    If n is composite, then \(c\le g_{n}(c)\) for any \(c\in ( -\infty ,b_{\zeta _{n}(z)}] \setminus \{ 0\} \) and the inequality is strict for all \(c\in ( -\infty ,b_{\zeta _{n}(z)}) \setminus \{ 0\} \); if \(c\in ( b_{\zeta _{n}(z)},+\infty ) \), then \(c>g_{n}(c)\).

  3. (iii)

    If n is prime, then \(c\le g_{n}(c)\) for any \(c\in [ a_{\zeta _{n}(z)},b_{\zeta _{n}(z)}] \setminus \{ 0\} \) and the inequality is strict for all \(c\in ( a_{\zeta _{n}(z)},b_{\zeta _{n}(z)}) \setminus \{0\} \); if \(c\in ( -\infty ,a_{\zeta _{n}(z)}) \cup (b_{\zeta _{n}(z)},+\infty ) \), then \(c>g_{n}(c)\).

Proof

Part (i). Let c, d be real numbers such that \(c<d<0\). From Proposition 2.1, \(b_{n,c}\) and \(b_{n,d}\) are the unique points of \(| \zeta _{n}^{*}(z)| =p_{k_{n}}^{-c}\) and \(| \zeta _{n}^{*}(z)| =p_{k_{n}}^{-d}\) that intersect the real axis, respectively. Therefore \(b_{n,c}\) and \(b_{n,d}\) satisfy the equations

$$\begin{aligned} \sum _{\begin{array}{c} m=1\\ m\ne p_{k_{n}} \end{array}}^{n}m^{-x}=p_{k_{n}}^{-c},\quad \sum _{\begin{array}{c} m=1\\ m\ne p_{k_{n}} \end{array}}^{n}m^{-x}=p_{k_{n}}^{-d}, \end{aligned}$$
(17)

respectively. Each equation of (17) has only one real solution by virtue of [20, p. 46] and then, since \(p_{k_{n}}^{-c}>p_{k_{n}}^{-d}\), the real solution of the first equation is obviously greater than the second one. Therefore \(-b_{n,c}>-b_{n,d}\), equivalently, \(b_{n,c}<b_{n,d}\). Consequently, \(g_{n}(c)<g_{n}(d)\) and then \(g_{n}\) is strictly increasing in \(( -\infty ,0) \). Let c, d be such that \(c>d>0\). From Proposition 2.1, \(| \zeta _{n}^{*}(z)| =p_{k_{n}}^{-c}\) and \(| \zeta _{n}^{*}(z)| =p_{k_{n}}^{-d}\) have infinitely many arc-connected components which are closed curves. Since \(p_{k_{n}}^{-c}<p_{k_{n}}^{-d}\), any point of \(| \zeta _{n}^{*}(z)| =p_{k_{n}}^{-c}\) is interior of \(| \zeta _{n}^{*}(z)| =p_{k_{n}}^{-d}\), so \(b_{n,c}\le b_{n,d}\). That is, \(g_{n}(c)\le g_{n}(d)\), which means that \(g_{n}\) is decreasing on \(( 0,+\infty ) \).

Part (ii). We firstly demonstrate that the bounds \(a_{\zeta _{n}(z)}\), \(b_{\zeta _{n}(z)}\) defined in (2) corresponding to \(\zeta _{n}(z)\) satisfy the second inequality of (3), that is

$$\begin{aligned} a_{\zeta _{n}(z)}<0<b_{\zeta _{n}(z)}\quad \text {for all }n>2. \end{aligned}$$
(18)

Indeed, we introduce the EP

$$\begin{aligned} G_{n}(z):=\zeta _{n}(-z). \end{aligned}$$
(19)

In [7, Chap. 3, Theorem 3.20] was shown that

$$ b_{G_{n}(z)}:=\sup \{ \mathfrak {R}z:G_{n}(z)=0\}>0\quad \text {for all }n>2, $$

now we claim that

$$ a_{G_{n}(z)}:=\inf \{ \mathfrak {R}z:G_{n}(z)=0\} <0\quad \text {for all }n>2. $$

Otherwise, if all the zeros of \(G_{n}(z)\), say \((z_{n,k})_{k=1,2,\ldots }\), satisfy \(\mathfrak {R}z_{n,k}\ge 0\), since \(b_{G_{n}(z)}>0\), there is at least a zero \(z_{n,k_{0}}\) with \(\mathfrak {R}z_{n,k_{0}}>0\). Then, as \(G_{n}(z)\) is almost-periodic (see for instance [4, 5] and [10, Chap. VI]), \(G_{n}(z)\) has infinitely many zeros in the strip

$$ S_{\varepsilon }:=\{ z:\mathfrak {R}z_{n,k_{0}}-\varepsilon<\mathfrak {R}z<\mathfrak {R}z_{n,k_{0}}+\varepsilon \} ,\quad 0<\varepsilon <\mathfrak {R}z_{n,k_{0}}, $$

and consequently

$$\begin{aligned} \sum _{k=1}^{\infty }\mathfrak {R}z_{n,k}=+\infty . \end{aligned}$$
(20)

However, as all the coefficients of \(G_{n}(z)\) are equal to 1, [21, formula (9)] applies and then we get \(\sum _{k=1}^{\infty }\mathfrak {R}z_{n,k}=O(1)\), contradicting (20). Therefore the claim follows, that is, \(a_{G_{n}(s)}<0\) for all \(n>2\). By (19) we have \(a_{\zeta _{n}(z)}=-b_{G_{n}(z)}\) and \(b_{\zeta _{n}(z)}=-a_{G_{n}(z)}\), so (18) follows.

We now consider the point \(b_{\zeta _{n}(z)}\). It is immediate that \(b_{\zeta _{n}(z)}\) belongs to the set \(R_{\zeta _{n}(z)}\) defined in (6). Then from Theorem 2.1 we have \(b_{\zeta _{n}(z)}\le g_{n}(b_{\zeta _{n}(z)})\), so the property \(c\le g_{n}(c)\) is true for \(c=b_{\zeta _{n}(z)}\). From (18) and by using that \(g_{n}\) is decreasing on \(( 0,\infty ) \) by virtue of Part (i), for any c \(\in ( 0,b_{\zeta _{n}(z)}) \) we have

$$\begin{aligned} 0<c<b_{\zeta _{n}(z)}\le g_{n}(b_{\zeta _{n}(z)})\le g_{n}(c). \end{aligned}$$
(21)

Consequently, Part (ii) follows for \(c\in ( 0,b_{\zeta _{n}(z)}] \). We now assume \(c<0\) and n composite, so \(p_{k_{n}}<n\). If \(b_{n,c}\ge 0\), then \(c<b_{n,c}=g_{n}(c)\) and again Part (ii) is true. Finally, we suppose \(b_{n,c}<0\). Since \(c<0\), \(b_{n,c}\) satisfies the first equation of (17) and then \(p_{k_{n}}^{-c}>n^{-b_{n,c}}\). Consequently \(-c>-b_{n,c}\), so \(c<b_{n,c}\) and then Part (ii) follows for \(c\in ( -\infty ,b_{\zeta _{n}(z)}] \setminus \{ 0\} \). Finally, we claim that \(c>g_{n}(c)\) when \(c>b_{\zeta _{n}(z)}\). Indeed, because of Lemma 2.2 and (8), we have \(f_{n}(c)<a_{\zeta _{n}^{*}(z)}\le 0\) for any c. Therefore, since \(c>b_{\zeta _{n}(z)}\), by (18) c is positive and then \(f_{n}(c)<c\). Assume \(c>g_{n}(c)\) is not true. Then we would have \(f_{n}(c)<c\le g_{n}(c)\) and by Theorem 2.1, \(c\in R_{\zeta _{n}(z)}\subset [ a_{\zeta _{n}(z)},b_{\zeta _{n}(z)}] \) which means that \(c\le b_{\zeta _{n}(z)}\). This is a contradiction because \(c>b_{\zeta _{n}(z)}\), so the claim follows. This definitely proves Part (ii).

Part (iii). We first note that, since n is prime, \(p_{k_{n}}=n\). Therefore the first equation in (17) becomes \(\sum _{m=1}^{n-1}m^{-x}=n^{-c}\). By assuming \(c<0\), \(b_{n,c}\) satisfies the above equation and then we have

$$\begin{aligned} \sum _{m=1}^{n-1}m^{-b_{n,c}}=n^{-c}. \end{aligned}$$
(22)

For every \(n\ge 2\), we consider the number \(\beta _{G_{n}(z)}\), defined as the unique real solution of the equation \(\sum _{m=1}^{n-1}m^{x}=n^{x}\) (see [20, p. 46]). By [6, Proposition 5], \(\beta _{G_{n}(z)}\ge b_{G_{n}(z)}\) and the equality is attained for n prime. Therefore the set \(\mathbb {R}\) of real numbers is partitioned in two sets:

$$\begin{aligned} ( -\infty ,\beta _{G_{n}(z)}] =\{ x\in \mathbb {R}:\sum _{m=1}^{n-1}m^{x}\ge n^{x}\} , \end{aligned}$$
(23)

and

$$\begin{aligned} ( \beta _{G_{n}(z)},\infty ) =\{ x\in \mathbb {R}:\sum _{m=1}^{n-1}m^{x}<n^{x}\} . \end{aligned}$$
(24)

Now we claim that \(c\le g_{n}(c)\) when \(a_{\zeta _{n}(z)}\le c<0\). Indeed, by (19), \(b_{G_{n}(z)}=-a_{\zeta _{n}(z)}\), so c is such that \(0<-c\le b_{G_{n}(z)}=\beta _{G_{n}(z)}\). Then, according to (23), we have

$$\begin{aligned} \sum _{m=1}^{n-1}m^{-c}\ge n^{-c}. \end{aligned}$$
(25)

Therefore, if we assume \(c>g_{n}(c)=b_{n,c}\), by applying (25) and taking into account (22), we get

$$ n^{-c}\le \sum _{m=1}^{n-1}m^{-c}<\sum _{m=1}^{n-1}m^{-b_{n,c}}=n^{-c}, $$

which is a contradiction. Therefore \(c\le g_{n}(c)\) is true for c such that \(a_{\zeta _{n}(z)}\le c<0\). Consequently, taking into account (21), it follows

$$ c\le g_{n}(c)\text {, for any }c\in [ a_{\zeta _{n}(z)},b_{\zeta _{n}(z)}] \setminus \{ 0\} , $$

where the inequality is strict for all c of \(( a_{\zeta _{n}(z)},b_{\zeta _{n}(z)}) \setminus \{ 0\} \). Now suppose \(c\in ( -\infty ,a_{\zeta _{n}(z)}) \). Then, since \(-c >-a_{\zeta _{n}(z)}=b_{G_{n}(z)}=\beta _{G_{n}(z)}\), by applying (24) we have

$$\begin{aligned} \sum _{m=1}^{n-1}m^{-c}<n^{-c}. \end{aligned}$$
(26)

It implies that \(c>g_{n}(c)\). Indeed, by supposing \(c\le g_{n}(c)=b_{n,c}\), from (22) and (26) we are led to the following contradiction:

$$ n^{-c}=\sum _{m=1}^{n-1}m^{-b_{n,c}}\le \sum _{m=1}^{n-1}m^{-c}<n^{-c}. $$

Therefore \(c>g_{n}(c)\) if \(c\in ( -\infty ,a_{\zeta _{n}(z)}) \). Finally, if \(c\in ( b_{\zeta _{n}(z)},+\infty ) \), the reasoning used to demonstrate the end of Part (ii) of the lemma proves that \(c>g_{n}(c) \).       \(\square \)

As a consequence of Lemma 2.3 we find the fixed points of the function \(g_{n}\).

Corollary 2.2

For every composite number \(n>2\), \(b_{\zeta _{n}(z)}\) is the fixed point of the function \(g_{n}\). If \(n>2\) is prime, \(a_{\zeta _{n}(z)}\), \(b_{\zeta _{n}(z)}\) are the fixed points of \(g_{n}\).

Proof

Fixed an integer \(n>2\), by (18) \(a_{\zeta _{n}(z)}\), \(b_{\zeta _{n}(z)}\ne 0\), so \(g_{n}\) is well defined at \(a_{\zeta _{n}(z)}\) and \(b_{\zeta _{n}(z)}\). By applying Part (ii) of Lemma 2.3 for \(n>2\) composite, it is immediate, by the continuity of \(g_{n}\), that the unique fixed point of \(g_{n}\) is \(b_{\zeta _{n}(z)}\). If \(n>2\) is prime, by Part (iii) of Lemma 2.3, we get \(g_{n}(a_{\zeta _{n}(z)})=a_{\zeta _{n}(z)}\) and \(g_{n}(b_{\zeta _{n}(z)})=b_{\zeta _{n}(z)}\). Furthermore, Part (iii) of Lemma 2.3 also proves that \(a_{\zeta _{n}(z)}\), \(b_{\zeta _{n}(z)}\) are the unique fixed points of \(g_{n}\).       \(\square \)

In the next result we obtain a characterization of \(\mathscr {P}^{*}\), the set of prime numbers greater than 2.

Theorem 2.2

An integer \(n>2\) belongs to \(\mathscr {P}^{*}\) if and only if \(a_{\zeta _{n}(z)}\) is a fixed point of the function \(g_{n}\).

Proof

Assume \(n>2\) is prime, from Corollary 2.2, \(a_{\zeta _{n}(z)}\) is a fixed point of \(g_{n}\). Conversely, if

$$\begin{aligned} g_{n}(a_{\zeta _{n}(z)})=a_{\zeta _{n}(z)}, \end{aligned}$$
(27)

by supposing n composite, from Part (ii) of Lemma 2.3, we have \(c<g_{n}(c)\) for all \(c\in ( -\infty ,b_{\zeta _{n}(z)}) \setminus \{0\} \). From (18), \(a_{\zeta _{n}(z)}\in ( -\infty ,b_{\zeta _{n}(z)}) \setminus \{ 0\} \). Then, \(a_{\zeta _{n}(z)}<g_{n}(a_{\zeta _{n}(z)})\). This contradicts (27). Consequently n is a prime number and then the theorem follows.       \(\square \)

3 The Fixed Points of \(f_{n}\) and the Sets \(R_{\zeta _{n}(z)}\)

For every integer \(n>2\), the function \(f_{n}\) defined in (12) allows us to give a sufficient condition to have points of the set \(R_{\zeta _{n}(z)}\), defined in (6).

Theorem 3.1

For every integer \(n>2\), if a point c \(\in \) \([ a_{\zeta _{n}(z)},b_{\zeta _{n}(z)}] \) satisfies \(f_{n}(c)\le c\), then \(c\in R_{\zeta _{n}(z)}\).

Proof

We first claim that

$$\begin{aligned} a_{\zeta _{n}(z)}, 0, b_{\zeta _{n}(z)}\in R_{\zeta _{n}(z)}\quad \text {for every }n\ge 2. \end{aligned}$$
(28)

Indeed, for \(n=2\), the claim trivially follows because as we have seen in Introduction all the zeros of \(\zeta _{2}(z)\) are imaginary, so \(a_{\zeta _{2}(z)}=\) \(b_{\zeta _{2}(z)}=0\) and then \(R_{\zeta _{2}(z)}=\{0\} \). Therefore we assume \(n>2\). By taking into account the definitions of \(a_{\zeta _{n}(z)}\), \(b_{\zeta _{n}(z)}\), both numbers obviously belong to \(R_{\zeta _{n}(z)}\). Regarding the fact that \(0\in R_{\zeta _{n}(z)}\)for all \(n>2\), it was proved in [18, (3.7)]. Then (28) is true. Hence it only remains to prove the theorem for c \(\in \) \((a_{G_{n}(z)},b_{G_{n}(z)})\setminus \{ 0\} \). But in this case,since by hypothesis \(f_{n}(c)\le c\), by using Parts (ii) and (iii) of Lemma 2.3 we are lead to \(f_{n}(c)\le c<g_{n}(c)\) and then, by Theorem 2.1, \(c\in R_{\zeta _{n}(z)}\).       \(\square \)

An important conclusion is deduced from the above theorem.

Theorem 3.2

For every integer \(n>2\), if c belongs to \(R_{\zeta _{n}(z)}\) then

$$\begin{aligned}{}[ f_{n}(c),c] \cap [ a_{\zeta _{n}(z)},b_{\zeta _{n}(z)}] \subset R_{\zeta _{n}(z)}. \end{aligned}$$
(29)

If \(n>2\) is composite and c belongs to \(R_{\zeta _{n}(z)}\), then

$$\begin{aligned}{}[ f_{n}(c),c] \subset R_{\zeta _{n}(z)}. \end{aligned}$$
(30)

Proof

Assume \(c\in \) \(R_{\zeta _{n}(z)}\). Then, by Theorem 2.1, \(f_{n}(c)\le c\le g_{n}(c)\). Therefore the interval \([ f_{n}(c),c] \) is well defined. If \(f_{n}(c)=c\) the theorem trivially follows. Suppose \(f_{n}(c)<c\). Let t be a point of \([ a_{\zeta _{n}(z)},b_{\zeta _{n}(z)}] \) such that \(f_{n}(c)<t<c\). By Lemma 2.1, \(f_{n}(t)<f_{n}(c)\). Therefore we have

$$ f_{n}(t)<f_{n}(c)<t<c, $$

and then, by applying Theorem 3.1, \(t\in R_{\zeta _{n}(z)}\). Consequently

$$ (f_{n}(c),c)\cap [ a_{\zeta _{n}(z)},b_{\zeta _{n}(z)}] \subset R_{\zeta _{n}(z)}, $$

and from the closedness of \(R_{\zeta _{n}(z)}\), (29) follows.

Assume \(n>2\) is composite. Since \(c\in \) \(R_{\zeta _{n}(z)}\) and

$$ R_{\zeta _{n}(z)}\subset [ a_{\zeta _{n}(z)},b_{\zeta _{n}(z)}] , $$

we have \(c\le b_{\zeta _{n}(z)}\). Furthermore, from Theorem 2.1, \(f_{n}(c)\le c\le g_{n}(c)\). Then, if \(f_{n}(c)\) \(=c\), (30) is obviously true. Suppose \(f_{n}(c)<c\). Consider a number t such that \(f_{n}(c)\le t<c \). Then, we get

$$\begin{aligned} f_{n}(c)\le t<c\le b_{\zeta _{n}(z)}. \end{aligned}$$
(31)

If \(t=0\), by virtue of (28), \(t\in R_{\zeta _{n}(z)}\). If \(t\ne 0\), from (31), \(t\in ( -\infty ,b_{\zeta _{n}(z)}) \setminus \{0\} \). Then, as n is composite, by Part (ii) of Lemma 2.3, \(t<g_{n}(t)\). On the other hand, since \(t<c\), from Lemma 2.1, \(f_{n}(t)<f_{n}(c)\) and then, again by (31), we have

$$ f_{n}(t)<f_{n}(c)\le t<g_{n}(t). $$

Now, by applying Theorem 2.1, \(t\in R_{\zeta _{n}(z)}\). Consequently \([f_{n}(c),c) \) \(\subset R_{\zeta _{n}(z)}\) and then, since by hypothesis \(c\in \) \(R_{\zeta _{n}(z)}\), we get \([ f_{n}(c),c]\subset R_{\zeta _{n}(z)}\). The proof is now completed.       \(\square \)

As a consequence of the two preceding results we characterize the set \(\mathscr {C}^{*}\) of composite numbers \(n>2\).

Corollary 3.1

For every \(n\in \mathscr {C}^{*}\), \(a_{\zeta _{n}(z)}\) is a fixed point of the function \(f_{n}\).

Proof

Assume \(n\in \mathscr {C}^{*}\). From (28), \(a_{\zeta _{n}(z)}\in R_{\zeta _{n}(z)}\). Since n is composite and greater than 2, by (30) we have \([ f_{n}(a_{\zeta _{n}(z)}),a_{\zeta _{n}(z)}] \subset R_{\zeta _{n}(z)}\). Noticing \(R_{\zeta _{n}(z)}\) \(\subset \) \([ a_{\zeta _{n}(z)},b_{\zeta _{n}(z)}] \), necessarily \(f_{n}(a_{\zeta _{n}(z)})=a_{\zeta _{n}(z)}\).       \(\square \)

In the next result we prove that \(a_{\zeta _{n}(z)}\) is not a fixed point of \(f_{n}\) for any \(n\in \mathscr {P}^{*}\).

Corollary 3.2

For every \(n\in \mathscr {P}^{*}\), \(f_{n}(a_{\zeta _{n}(z)})<a_{\zeta _{n}(z)}\).

Proof

For every \(n>2\), the variety \(| \zeta _{n}^{*}(z)|=p_{k_{n}\text { }}^{-c}\), for arbitrary \(c\in \mathbb {R}\), by virtue of equation (10) is not contained in a vertical line, so the interval of the variation of the variable x in the variety \(|\zeta _{n}^{*}(z)| =p_{k_{n}\text { }}^{-c}\) is not degenerate. Therefore, taking into account (12), we have

$$\begin{aligned} f_{n}(c)<g_{n}(c)\text { for every integer }n>2,\quad \text {for all }c\in \mathbb {R} . \end{aligned}$$
(32)

Assume \(n>2\) prime. By Corollary 2.2, \(g_{n}(a_{\zeta _{n}(z)})=a_{\zeta _{n}(z)}\). Then, by taking \(c=a_{\zeta _{n}(z)}\) in (32), the corollary follows.       \(\square \)

As a simple consequence from Corollary 3.2 we obtain a characterization of \(\mathscr {C}^{*}\).

Theorem 3.3

An integer \(n>2\) belongs to \(\mathscr {C}^{*}\) if and only if \(a_{\zeta _{n}(z)}\) is a fixed point of the function \(f_{n}\).

Proof

From Corollary 3.1, if \(n>2\) is composite, \(a_{\zeta _{n}(z)}\) is a fixed point of \(f_{n}\). Reciprocally, if \(a_{\zeta _{n}(z)}\) is a fixed point of \(f_{n}\), by assuming \(n>2\) is not composite, by applying Corollary 3.2 we are led to a contradiction. Therefore, the theorem follows.       \(\square \)

The bounds \(a_{\zeta _{n}(z)}\), \(a_{\zeta _{n}^{*}(z)}\) satisfy the following inequality.

Proposition 3.1

For every integer \(n>2\), \(a_{\zeta _{n}(z)}<a_{\zeta _{n}^{*}(z)}\).

Proof

By taking \(c=a_{\zeta _{n}^{*}(z)}\) in Lemma 2.2 we have

$$\begin{aligned} f_{n}(a_{\zeta _{n}(z)}^{*})<a_{\zeta _{n}(z)}^{*}\text { for all }n>2 . \end{aligned}$$
(33)

Again from Lemma 2.2, for \(c=a_{\zeta _{n}(z)}\), we get \(f_{n}(a_{\zeta _{n}(z)})<a_{\zeta _{n}(z)}^{*}\). If n is composite, by Corollary 3.1 \( f_{n}(a_{\zeta _{n}(z)})=a_{\zeta _{n}(z)}\) and from (33) we then deduce that \(a_{\zeta _{n}(z)}<a_{\zeta _{n}(z)}^{*}\). This proves the proposition for n composite.

Assume n is prime. Then \(p_{k_{n}}=n\) and, from (7), \(\zeta _{n}^{*}(z)=\zeta _{n-1}(z)\), so \(a_{\zeta _{n}^{*}(z)}=a_{\zeta _{n-1}(z)}\). Now we consider the function \(G_{n}(z)\) defined in (19). As we have seen in the proof of Lemma 2.3, because of [6, Proposition 5] we have \(b_{G_{n}(z)}\le \beta _{G_{n}(z)}\) for all \(n\ge 2\) and the equality is attained for n prime. Noticing [17, Lemma 1], \(\beta _{G_{n-1}(z)}<\beta _{G_{n}(z)}\) for all \(n>2\). Then we get

$$\begin{aligned} b_{G_{n-1}(z)}\le \beta _{G_{n-1}(z)}<\beta _{G_{n}(z)}=b_{G_{n}(z)},\quad \text {for all prime }n>2, \end{aligned}$$
(34)

or equivalently

$$ -b_{G_{n-1}(z)}\ge -\beta _{G_{n-1}(z)}>-\beta _{G_{n}(z)}=-b_{G_{n}(z)},\quad \text {for all prime }n>2. $$

Now, since from (19) \(a_{\zeta _{n}(z)}=-b_{G_{n}(z)}\) for all \(n\ge 2\), from the above chain of inequalities we deduce

$$ a_{\zeta _{n}^{*}(z)}=a_{\zeta _{n-1}(z)}=-b_{G_{n-1}(z)}>-b_{G_{n}(z)}=a_{\zeta _{n}(z)},\quad \text {for all prime }n>2. $$

The proof is now completed.       \(\square \)

Corollary 3.3

For every integer \(n>2\), \(a_{\zeta _{n}(z)}^{*}\in R_{\zeta _{n}(z)}\).

Proof

For \(n=3,4\), because of (8) we have \(a_{\zeta _{n}(z)}^{*}=0\). Therefore, from (28), \(a_{\zeta _{n}(z)}^{*}\in R_{\zeta _{n}(z)}\) for \(n=3\), 4. Assume \(n>4\). By Proposition 3.1, \(a_{\zeta _{n}(z)}<a_{\zeta _{n}^{*}(z)}\) for all \(n>2\). Then, from (8) and (18), \(a_{\zeta _{n}(z)}^{*}\in [ a_{\zeta _{n}(z)},b_{\zeta _{n}(z)}] \) for all \(n>4\). Therefore, by using (33) and applying Theorem 3.1, \(a_{\zeta _{n}(z)}^{*}\in R_{\zeta _{n}(z)}\) for all \(n>4\). This proves the corollary.       \(\square \)

In the next result we prove the existence of a minimal density interval for every \(\zeta _{n}(z)\), \(n>2\).

Theorem 3.4

For every integer \(n>2\) there exists a number \(A_{n}\in [ a_{\zeta _{n}(z)},a_{\zeta _{n}^{*}(z)}) \) such that \([ A_{n},b_{\zeta _{n}(z)}] \subset R_{\zeta _{n}(z)}\).

Proof

Firstly we note that, by Proposition 3.1, the interval \([ a_{\zeta _{n}(z)},a_{\zeta _{n}^{*}(z)}) \) is well defined. On the other hand, by (18) \(b_{\zeta _{n}(z)}>0\) and, by (8) \(a_{\zeta _{n}^{*}(z)}\le 0\) for all \(n>2\), so by Proposition 3.1 we have

$$\begin{aligned} a_{\zeta _{n}(z)}<a_{\zeta _{n}^{*}(z)}\le 0<b_{\zeta _{n}(z)},\quad \text {for all }n>2. \end{aligned}$$
(35)

This means that \([ a_{\zeta _{n}^{*}(z)},b_{\zeta _{n}(z)}] \) is a non-degenerate sub-interval of \([ a_{\zeta _{n}(z)},b_{\zeta _{n}(z)}] \) for any \(n>2\). By Lemma 2.2, we have \(f_{n}(b_{\zeta _{n}(z)})<a_{\zeta _{n}^{*}(z)}\). Then, according to (35), we get

$$ f_{n}(b_{\zeta _{n}(z)})\le a_{\zeta _{n}^{*}(z)}<b_{\zeta _{n}(z)}, $$

so

$$ [ a_{\zeta _{n}^{*}(z)},b_{\zeta _{n}(z)}] \subset [f_{n}(b_{\zeta _{n}(z)}),b_{\zeta _{n}(z)}] . $$

Now, since \(b_{\zeta _{n}(z)}\in R_{\zeta _{n}(z)}\), because of Theorem 3.2 we obtain

$$\begin{aligned}{}[ a_{\zeta _{n}^{*}(z)},b_{\zeta _{n}(z)}] \subset [f_{n}(b_{\zeta _{n}(z)}),b_{\zeta _{n}(z)}] \cap [ a_{\zeta _{n}(z)},b_{\zeta _{n}(z)}] \subset R_{\zeta _{n}(z)} . \end{aligned}$$
(36)

This implies that \(a_{\zeta _{n}^{*}(z)}\in R_{\zeta _{n}(z)}\) (observe that from Corollary 3.3 we already knew that \(a_{\zeta _{n}^{*}(z)}\in R_{\zeta _{n}(z)})\)) so, again by Theorem 3.2, we have

$$\begin{aligned}{}[ f_{n}(a_{\zeta _{n}^{*}(z)}),a_{\zeta _{n}^{*}(z)} \cap [ a_{\zeta _{n}(z)},b_{\zeta _{n}(z)}] \subset R_{\zeta _{n}(z)}. \end{aligned}$$
(37)

If \(f_{n}(a_{\zeta _{n}^{*}(z)})\le a_{\zeta _{n}(z)}\), from (37) we deduce that \([ a_{\zeta _{n}(z)},a_{\zeta _{n}^{*}(z)}] \subset R_{\zeta _{n}(z)}\) and then, by (36) we get \([ a_{\zeta _{n}(z)},b_{\zeta _{n}(z)}] \subset R_{\zeta _{n}(z)}\). In this case by taking \(A_{n}=a_{\zeta _{n}(z)}\), the theorem follows. Moreover, \(\zeta _{n}(z)\) has a maximum density interval and it coincides with its critical interval \([ a_{\zeta _{n}(z)},b_{\zeta _{n}(z)}] \).

If \(f_{n}(a_{\zeta _{n}^{*}(z)})>a_{\zeta _{n}(z)}\), from (37) we deduce

$$\begin{aligned}{}[ f_{n}(a_{\zeta _{n}^{*}(z)}),a_{\zeta _{n}^{*}(z)}]\subset R_{\zeta _{n}(z)}. \end{aligned}$$
(38)

Therefore \(f_{n}(a_{\zeta _{n}^{*}(z)})\) \(\in R_{\zeta _{n}(z)}\) and, again by Theorem 3.2, we have

$$\begin{aligned}{}[ f_{n}^{(2)}(a_{\zeta _{n}^{*}(z)}),f_{n}(a_{\zeta _{n}^{*}(z)})] \cap [ a_{\zeta _{n}(z)},b_{\zeta _{n}(z)}] \subset R_{\zeta _{n}(z)}, \end{aligned}$$
(39)

where \(f_{n}^{(2)}\) denotes \(f_{n}\) composed with itself. Then, if \(f_{n}^{(2)}(a_{\zeta _{n}^{*}(z)})\le a_{\zeta _{n}(z)}\), from (39), we have \([ a_{\zeta _{n}(z)},f_{n}(a_{\zeta _{n}^{*}(z)})]\subset R_{\zeta _{n}(z)}\) and by (38), we get \([ a_{\zeta _{n}(z)},a_{\zeta _{n}^{*}(z)}] \subset R_{\zeta _{n}(z)}\). Therefore taking into account (36) we obtain \([ a_{\zeta _{n}(z)},b_{\zeta _{n}(z)}] \subset R_{\zeta _{n}(z)}\). Consequently, by taking \(A_{n}=a_{\zeta _{n}(z)}\), the theorem follows and \(\zeta _{n}(z)\) has a maximum density interval that coincides with its critical interval \([ a_{\zeta _{n}(z)},b_{\zeta _{n}(z)}] \). If \(f_{n}^{(2)}(a_{\zeta _{n}^{*}(z)})>a_{\zeta _{n}(z)}\), from (39), we get

$$ [ f_{n}^{(2)}(a_{\zeta _{n}^{*}(z)}),f_{n}(a_{\zeta _{n}^{*}(z)})] \subset R_{\zeta _{n}(z)} . $$

Therefore \(f_{n}^{(2)}(a_{\zeta _{n}^{*}(z)})\in R_{\zeta _{n}(z)}\) and, again by Theorem 3.2, we have

$$ [ f_{n}^{(3)}(a_{\zeta _{n}^{*}(z)}),f_{n}^{(2)}(a_{\zeta _{n}^{*}(z)})] \cap [ a_{\zeta _{n}(z)},b_{\zeta _{n}(z)}] \subset R_{\zeta _{n}(z)}, $$

and so on. Therefore, by denoting \(f_{n}^{(k)}=f_{n}^{(k-1)}\circ f_{n}\) for \(k\ge 2\) and repeating the process above, we are led to one of the two cases:

(i) There is some \(k\ge 1\) such that \(f_{n}^{(k)}(a_{\zeta _{n}^{*}(z)})\le a_{\zeta _{n}(z)}\). In this case, as we have seen \(A_{n}=a_{\zeta _{n}(z)}\) and then \(\zeta _{n}(z)\) has a maximum density interval that coincides with its critical interval \([ a_{\zeta _{n}(z)},b_{\zeta _{n}(z)}] \).

(ii) For all k, \(f_{n}^{(k)}(a_{\zeta _{n}^{*}(z)})>a_{\zeta _{n}(z)}\) and then, by virtue of Lemma 2.1 and (33), we have

$$ a_{\zeta _{n}(z)}<\cdots<f_{n}^{(k)}(a_{\zeta _{n}^{*}(z)})<\cdots<f_{n}^{(2)}(a_{\zeta _{n}^{*}(z)})<f_{n}(a_{\zeta _{n}^{*}(z)})<a_{\zeta _{n}^{*}(z)}. $$

Consequently there exists \(\lim _{k\rightarrow \infty }f_{n}^{(k)}(a_{\zeta _{n}^{*}(z)})\) and then, by defining

$$ A_{n}:=\lim _{k\rightarrow \infty }f_{n}^{(k)}(a_{\zeta _{n}^{*}(z)}), $$

we have \(a_{\zeta _{n}(z)}\le A_{n}<a_{\zeta _{n}^{*}(z)}\). On the other hand, by reiterating Theorem 3.2, we get

$$\begin{aligned}{}[ f_{n}^{(k)}(a_{\zeta _{n}^{*}(z)}),f_{n}^{(k-1)}(a_{\zeta _{n}^{*}(z)})] \subset R_{\zeta _{n}(z)}\text {, for all }k\ge 2 . \end{aligned}$$
(40)

Then taking into account (36) and (38), by (40) we deduce that \([A_{n},b_{\zeta _{n}(z)}] \subset R_{\zeta _{n}(z)}\). This definitely proves the theorem.       \(\square \)

Remark 3.5

Observe that if the case (ii) of above theorem holds, \(A_{n}\) will be a fixed point of \(f_{n}\) by virtue of the continuity of \(f_{n}\). Then if \(n\in \mathscr {C}^{*}\), by Theorem 14, the point \(A_{n}\) could be \(a_{\zeta _{n}(z)}\). But if \(n\in \mathscr {P}^{*}\), from Corollary 3.2, \(A_{n}\) can not be equal to \(a_{\zeta _{n}(z)}\).

In the next result we prove that the number of fixed points of \(f_{n}\) influences on the existence of a maximum density interval of \(\zeta _{n}(z)\).

Theorem 3.6

For every integer \(n>2\), if \(f_{n}\) has at most a fixed point in the interval \((a_{\zeta _{n}(z)},a_{\zeta _{n}^{*}(z)})\) then \(\zeta _{n}(z) \) has a maximum density interval that coincides with the critical interval \([ a_{\zeta _{n}(z)},b_{\zeta _{n}(z)}] \) associated with \(\zeta _{n}(z)\).

Proof

We first assume \(f_{n}\) has no fixed point in \((a_{\zeta _{n}(z)},a_{\zeta _{n}^{*}(z)})\). Then we claim that \(f_{n}(c)<c\) for all \(c\in (a_{\zeta _{n}(z)},a_{\zeta _{n}^{*}(z)}] \). Indeed, we define the function \(h_{n}(c):=f_{n}(c)-c\). Then \(h_{n}\) is continuous on \(\mathbb {R}\), and by virtue of Lemma 2.2 and (33), \(h_{n}\) is negative on \([a_{\zeta _{n}^{*}(z)},\infty ) \). Then, since \(f_{n}\) by hypothesis has no fixed point on \((a_{\zeta _{n}(z)},a_{\zeta _{n}^{*}(z)})\), \(h_{n}(c)\) has no zero on \(( a_{\zeta _{n}(z)},\infty ) \). Consequently, \(h_{n}(c)<0\) for any \(c\in ( a_{\zeta _{n}(z)},\infty ) \) and in particular we have

$$\begin{aligned} f_{n}(c)<c\text { for all }c\in ( a_{\zeta _{n}(z)},a_{\zeta _{n}^{*}(z)}] . \end{aligned}$$
(41)

Hence the claim follows. On the other hand, by Corollary 3.3 \(a_{\zeta _{n}^{*}(z)}\in \) \(R_{\zeta _{n}(z)}\subset [ a_{\zeta _{n}(z)},b_{\zeta _{n}(z)}] \), so

$$ ( a_{\zeta _{n}(z)},a_{\zeta _{n}^{*}(z)}] \subset [a_{\zeta _{n}(z)},b_{\zeta _{n}(z)}] . $$

Consequently, by taking into account (41) and by applying Theorem 3.1 we have

$$ ( a_{\zeta _{n}(z)},a_{\zeta _{n}^{*}(z)}] \subset R_{\zeta _{n}(z)} . $$

Therefore, since from (28) \(a_{\zeta _{n}(z)}\in \) \(R_{\zeta _{n}(z)}\), we get \([ a_{\zeta _{n}(z)},a_{\zeta _{n}^{*}(z)}] \subset R_{\zeta _{n}(z)}\) and then by (36) it follows that \([ a_{\zeta _{n}(z)},b_{\zeta _{n}(z)}] \subset R_{\zeta _{n}(z)}\). As always is true that \(R_{\zeta _{n}(z)}\subset [ a_{\zeta _{n}(z)},b_{\zeta _{n}(z)}] \) we deduce that \(R_{\zeta _{n}(z)}=[ a_{\zeta _{n}(z)},b_{\zeta _{n}(z)}] \), i.e. \(\zeta _{n}(z)\) has a maximum density interval. Then the theorem follows in this case.

We now suppose \(f_{n}\) has only one fixed point, say \(c_{1}\), in \((a_{\zeta _{n}(z)},a_{\zeta _{n}^{*}(z)})\). Then the function \(h_{n}(c):=f_{n}(c)-c \), continuous on \(\mathbb {R}\), is non-positive on \([ c_{1},+\infty ) \) by virtue of Lemma 2.2. Therefore, in particular, \(f_{n}(c)\le c\) for all \(c\in [c_{1},a_{\zeta _{n}^{*}(z)}] \). Since \([ c_{1},a_{\zeta _{n}^{*}(z)}] \subset [ a_{\zeta _{n}(z)},b_{\zeta _{n}(z)}] \), by applying the Theorem 3.1 at any \(c\in [ c_{1},a_{\zeta _{n}^{*}(z)}] \) we have

$$\begin{aligned}{}[ c_{1},a_{\zeta _{n}^{*}(z)}] \subset R_{\zeta _{n}(z)}. \end{aligned}$$
(42)

Now we claim that \(h_{n}\) is negative on \(( a_{\zeta _{n}(z)},c_{1}) \). Indeed, if we assume that \(h_{n}\) is non-negative on \(( a_{\zeta _{n}(z)},c_{1}) \), since \(c_{1}\) is the unique fixed point of \(f_{n}\) in \((a_{\zeta _{n}(z)},a_{\zeta _{n}^{*}(z)})\), then \(f_{n}(c)>c\) for all \(c\in ( a_{\zeta _{n}(z)},c_{1}) \). Then, by Theorem 2.1, \(c\notin R_{\zeta _{n}(z)}\) for all \(c\in \) \((a_{\zeta _{n}(z)},c_{1}) \). This means that \(\zeta _{n}(z)\) has no zero on the strip \(( a_{\zeta _{n}(z)},c_{1}) \times \mathbb {R}\). But, taking into account that \(a_{\zeta _{n}(z)}\in R_{\zeta _{n}(z)}\), \(a_{\zeta _{n}(z)}\) would be an isolated point of \(R_{\zeta _{n}(z)} \) and it contradicts [2, Corollary 3.2]. Therefore the claim follows. Consequently, \(f_{n}(c)<c\) for all \(c\in ( a_{\zeta _{n}(z)},c_{1}) \) and then, by Theorem 3.1, \(( a_{\zeta _{n}(z)},c_{1}) \subset R_{\zeta _{n}(z)}\). From the closedness of \(R_{\zeta _{n}(z)}\), we have

$$\begin{aligned}{}[ a_{\zeta _{n}(z)},c_{1}] \subset R_{\zeta _{n}(z)}. \end{aligned}$$
(43)

Then, from (43), (42) and (36) we deduce that \([ a_{\zeta _{n}(z)},b_{\zeta _{n}(z)}] \subset R_{\zeta _{n}(z)}\). Consequently, \(\zeta _{n}(z)\) has a maximum density interval and it coincides with its critical interval \([ a_{\zeta _{n}(z)},b_{\zeta _{n}(z)}] \).       \(\square \)

As a first application of the usefulness of Theorem 3.6 we prove a result on \(\zeta _{3}(z)\) (the same result can be also deduced from others methods as we can see in [13, 15]).

Corollary 3.4

\(\zeta _{3}(z)\) has a maximum density interval and it coincides with its critical interval \([ a_{\zeta _{3}(z)},b_{\zeta _{3}(z)}] \).

Proof

The function \(f_{3}(c):=a_{3,c}\) is explicitly given by the formula (11). Then it is immediate to check that \(f_{3}(c)<c\) for all \(c\in \mathbb {R}\). Therefore \(f_{3}(c)\) has no fixed point and then, from Theorem 3.6, \(\zeta _{3}(z)\) has a maximum density interval and it coincides with \([a_{\zeta _{3}(z)},b_{\zeta _{3}(z)}] \).       \(\square \)

4 The Fixed Point Theory and the Maximum Density Interval for \(\zeta _{n}(z)\)

In this section our aim is to give a very useful result (see below Lemma 4.1) based on Kronecker Theorem [8, Theorem 444] that allows us to apply our fixed point theory to prove the existence of a maximum density interval.

Let \(\mathscr {P}:=\) \(\{ p_{j}:j=1,2,3,\ldots \} \) be the set of prime numbers and \(U:=\{ 1,-1\} \). For every map \(\delta :\mathscr {P}\rightarrow U\), we define the function \(\omega _{\delta }: \mathbb {N}\rightarrow U\) as

$$\begin{aligned} \omega _{\delta }(1):=1,\quad \omega _{\delta }(m):=(\delta (p_{k_{1}}))^{\alpha _{1}}\ldots (\delta (p_{k_{l(m)}}))^{\alpha _{l(m)}},\quad m>1, \end{aligned}$$
(44)

where \((p_{k_{1}})^{\alpha _{1}}\ldots (p_{k_{l(m)}})^{\alpha _{l(m)}}\), with \(\alpha _{1}\), ..., \(\alpha _{l(m)}\in \mathbb {N}\), is the decomposition of m in prime factors. Let \(\varOmega \) be the set of all the \(\omega _{\delta }\)’s defined in (44). Observe that all functions of \(\varOmega \) are completely multiplicative (see for instance [1, p. 138]).

Lemma 4.1

Let \(n>2\) a fixed integer, \(p_{k_{n}}\) the last prime not exceeding n and \(f_{n}\) defined in (12). Given an arbitrary \(\omega _{\delta }\in \varOmega \), the inequality

$$\begin{aligned} p_{k_{n}}^{-c}\le | \sum _{ \begin{array}{c} m=1\\ m\ne p_{k_{n}} \end{array}}^{n}\omega _{\delta }(m)m^{-f_{n}(c)}| , \end{aligned}$$
(45)

holds for all \(c\in \mathbb {R}\).

Proof

Because of (7), \(\zeta _{n}^{*}(z):=\) \(\sum _{m=1,m\ne p_{k_{n}}}^{n}m^{-z}\). Therefore, given \(c\in \mathbb {R}\) we have

$$ \zeta _{n}^{*}(f_{n}(c)+iy)=\sum _{\begin{array}{c} m=1\\ m\ne p_{k_{n}} \end{array}}^{n}m^{-f_{n}(c)}(\cos (y\log m)-i\sin (y\log m)) . $$

Then taking into account (14),

$$\begin{aligned} p_{k_{n}}^{-c}\le | \sum _{\begin{array}{c} m=1\\ m\ne p_{k_{n}} \end{array}}^{n}m^{-f_{n}(c)}(\cos (y\log m)-i\sin (y\log m))| \text {, for all }y\in \mathbb {R}. \end{aligned}$$
(46)

Given \(n>2\), we define \(J_{n}:=\{ 1,2,3,\ldots \pi (n)\} \), where \(\pi (n)\) denotes the number of prime numbers not exceeding n. As the set \(\{ \log p_{j}:j\in J_{n}\} \) is rationally independent, the set \(\{ \frac{\log p_{j}}{2\pi }:j\in J_{n}\} \) is also rationally independent. Then by Kronecker Theorem [8, Theorem 444] fixed an arbitrary set of real numbers \(\{ \gamma _{j}:j\in J_{n}\} \) and given an integer \(N\ge 1\), there exists a real number \(y_{N}>N\) and integers \(m_{j,N}\), such that

$$\begin{aligned} | y_{N}\frac{\log p_{j}}{2\pi }-m_{j,N}-\gamma _{j}| <\frac{1}{N},\quad \text { for all }j\in J_{n}. \end{aligned}$$
(47)

For each \(n>2\), we define the set \(\mathscr {P}_{n}:=\{ p_{j}\in \mathscr {P}:p_{j}\le n\} \). Then, given a mapping \(\delta :=\mathscr {P}_{n}\rightarrow U\), we consider the set \(\{ \gamma _{j}:j\in J_{n}\} \) where \(\gamma _{j}=1\) for those j such that \(\delta (p_{j})=1\) and \(\gamma _{j}=1/2\) for those j such that \(\delta (p_{j})=-1\). Then by applying the aforementioned Kronecker Theorem for \(N=1,2 \ldots \), we can determine a sequence \((y_{N})_{N}\) satisfying, by virtue of (47), that

$$ \cos (y_{N}\log p_{j})\rightarrow 1,\quad \sin (y_{N}\log p_{j})\rightarrow 0\text { as }N\rightarrow \infty \text {, for }p_{j}\text { with }\delta (p_{j})=1 , $$

and

$$\begin{aligned} \cos (y_{N}\log p_{j})\rightarrow -1,\quad \sin (y_{N}\log p_{j})\rightarrow 0\text { as }N\rightarrow \infty \text {, for }p_{j}\text { with }\delta (p_{j})=-1. \end{aligned}$$

Therefore for each m such that \(1\le m\le n\) we get

$$\begin{aligned} \cos (y_{N}\log m)\rightarrow \omega _{\delta }(m),\quad \sin (y_{N}\log m)\rightarrow 0\quad \text {as }N\rightarrow \infty . \end{aligned}$$
(48)

Now, we substitute y by \(y_{N}\) in (46) and we take the limit as \(N\rightarrow \infty \). Then, according to (48), the inequality (45) follows.       \(\square \)

Theorem 4.1

For all prime numbers \(n>2\) except at most for a finite quantity, \(f_{n}\) has no fixed point in the interval \((a_{\zeta _{n}(z)},a_{\zeta _{n}^{*}(z)})\).

Proof

Corollary 3.4 proves the theorem for \(n=3\). Assume \(n>3\) prime. The numbers \(n-2\) and \(n-1\) are relatively primes and both cannot be perfect squares, so there exists \(\omega _{\delta }\in \varOmega \) such that \(\omega _{\delta }(n-2)\omega _{\delta }(n-1)=-1\). Since n is prime, \(a_{\zeta _{n}^{*}(z)}=a_{\zeta _{n-1}(z)}\) and \(p_{k_{n}}=n\). By supposing the existence of a fixed point \(c_{n}\in (a_{\zeta _{n}(z)},a_{\zeta _{n-1}(z)})\) for the function \(f_{n}\) for infinitely many prime \(n>3\), we are led to the following contradiction:

By (45) we have

$$\begin{aligned} n^{-c_{n}}\le | \pm ((n-1)^{-c_{n}}-(n-2)^{-c_{n}})+\sum _{m\in P_{n-3,\omega _{\delta }}}m^{-c_{n}}-\sum _{m\notin P_{n-3,\omega _{\delta }}}m^{-c_{n}}| , \end{aligned}$$
(49)

where, for a fixed integer \(n>2\) and \(\omega _{\delta }\in \varOmega \), the set \(P_{n,\omega _{\delta }}\) is defined as

$$ P_{n,\omega _{\delta }}:=\{ m:1\le m\le n\text { such that }\omega _{\delta }(m)=1\} . $$

On the other hand, \(\lim _{n\rightarrow \infty }\frac{a_{\zeta _{n}(z)}}{n}=-\log 2\) (see [3, Theorem 1] and [17, Theorem 2]). Then noticing that \(a_{\zeta _{n}(z)}<c_{n}<a_{\zeta _{n-1}(z)}\), we get

$$ \lim _{\begin{array}{c} n\text { prime} \\ n\rightarrow \infty \end{array}} \frac{c_{n}}{n-1}=-\log 2. $$

Therefore, for each fixed \(j\ge 0\), it follows

$$\begin{aligned} \lim _{\begin{array}{c} n\text { prime} \\ n\rightarrow \infty \end{array}} \left( \frac{n-j}{n-1}\right) ^{-c_{n}}=2^{-j+1}. \end{aligned}$$
(50)

Now, dividing by \((n-1)^{-c_{n}}\) the inequality (49), we have

$$\begin{aligned} \left( \frac{n}{n-1}\right) ^{-c_{n}}&\le \Bigg | \pm \Bigg ( 1-\left( \frac{n-2}{n-1}\right) ^{-c_{n}}\Bigg ) \nonumber \\&\qquad +\sum _{m\in P_{n-3,\omega _{\delta }}}\left( \frac{m}{n-1}\right) ^{-c_{n}} -\sum _{m\notin P_{n-3,\omega _{\delta }}}\left( \frac{m}{n-1}\right) ^{-c_{n}}\Bigg | \nonumber \\&\le \Bigg | \pm \Bigg ( 1-\left( \frac{n-2}{n-1}\right) ^{-c_{n}}\Bigg )\Bigg | \\&\qquad + \Bigg | \sum _{m\in P_{n-3,\omega _{\delta }}} \left( \frac{m}{n-1}\right) ^{-c_{n}} -\sum _{m\notin P_{n-3,\omega _{\delta }}}\left( \frac{m}{n-1}\right) ^{-c_{n}} \Bigg |\nonumber \\&\le \left( 1-\left( \frac{n-2}{n-1}\right) ^{-c_{n}}\right) +\sum _{j=3}^{n-1}\left( \frac{n-j}{n-1}\right) ^{-c_{n}}.\nonumber \end{aligned}$$
(51)

According to (50), by taking the limit in (51) for n prime, \(n\rightarrow \infty \), it follows that the limit of the left-hand side of (51) is 2 whereas the limit of the right-hand side one is \(1/2+\sum _{j=3}^{\infty }2^{-j+1}=1\). This is the contradiction desired. Hence the theorem follows.       \(\square \)

As a consequence from Theorem 4.1, an important property of the partial sums of order n prime can be deduced.

Theorem 4.2

For all prime numbers \(n>2\) except at most for a finite quantity, \(\zeta _{n}(z)\) has a maximum density interval and it coincides with its critical interval \([ a_{\zeta _{n}(z)},b_{\zeta _{n}(z)}] \).

Proof

It is enough to apply Theorems 3.6 and 4.1.       \(\square \)

5 Numerical Experiences

Simple numerical experiences carried out for some values of n in inequality (45) joint with the application of Theorem 3.6 and Lemma 4.1, allows us to prove the existence of a maximum density interval of \(\zeta _{n}(z)\) for all \(2\le n\le 8\). Indeed: For \(n=2\), we have already seen in the Introduction section that the zeros of \(\zeta _{2}(z)\) are all imaginary, so the set \(R_{\zeta _{2}(z)}=\{ 0\} \) and then \(a_{\zeta _{2}(z)}=b_{G_{2}(z)}=0\) which means that we trivially have

$$\begin{aligned} R_{\zeta _{2}(z)}=[ a_{\zeta _{2}(z)},b_{\zeta _{2}(z)}] . \end{aligned}$$

Therefore \(\zeta _{2}(z)\) has a maximum density interval (in this case degenerate).

For \(n=3\), Corollary 3.4 proves that

$$\begin{aligned} R_{\zeta _{3}(z)}=[ a_{\zeta _{3}(z)},b_{\zeta _{3}(z)}] \end{aligned}$$

and then \(\zeta _{3}(z)\) has a maximum density interval. In this case the end-points \(a_{\zeta _{3}(z)}\), \(b_{\zeta _{3}(z)}\) can be easily computed, being \(a_{\zeta _{3}(z)}=-1\) and \(b_{\zeta _{3}(z)}\approx 0.79\). Thus, \(R_{\zeta _{3}(z)}\approx [ -1,0.79] \).

For \(n=4\), we firstly claim that \(f_{4}\) has no fixed point in the interval \((a_{\zeta _{4}(z)},a_{\zeta _{4}^{*}(z)})\). Indeed, by (8), \(a_{\zeta _{4}^{*}(z)}=0\) and from (18), \(a_{\zeta _{4}(z)}<0\). Therefore we only study the behavior of \(f_{4}(c)\) for \(c<0\). We recall that from (12) \(f_{4}(c)=a_{4,c}\), where \(a_{4,c}\) is the left end-point of the interval of variation of the variable x in the Cartesian equation of the variety \(| \zeta _{4}^{*}(z)| =p_{k_{4}}^{-c}\). By taking into account formula (10) for \(n=4\), the equation of that variety is

$$\begin{aligned} 1+2^{-2x}+4^{-2x}+2\cdot 2^{-x}(1+4^{-x})\cos (y\log 2)+2\cdot 4^{-x}\cos ( y\log 4)) =3^{-2c}. \end{aligned}$$
(52)

By putting \(\cos ( y\log 4)) =2\cos ^{2}(y\log 2)-1\) in (52) and solving it for \(\cos (y\log 2)\) we have

$$\begin{aligned} \cos (y\log 2)=\frac{-(1+4^{-x})\pm \sqrt{(2\cdot 3^{-c})^{2}-(\sqrt{3}(4^{-x}-1))^{2}}}{4\cdot 2^{-x}}. \end{aligned}$$

Then the variable x must satisfy the inequality \((\sqrt{3}(4^{-x}-1))^{2}\le (2\cdot 3^{-c})^{2}\) which is equivalent to say that

$$\begin{aligned} 4^{-x}\in [ 1-2\cdot 3^{-c-\frac{1}{2}},1+2\cdot 3^{-c-\frac{1}{2}}]. \end{aligned}$$
(53)

Since \(1-2\cdot 3^{-c-\frac{1}{2}}<0\) for all \(c<0\), by noting that \(4^{-x}>0\) for any x, (53) is in turn equivalent to

$$\begin{aligned} -\frac{\log ( 1+2\cdot 3^{-c-\frac{1}{2}}) }{\log 4}\le x. \end{aligned}$$

Hence the minimum value for x is \(-\frac{\log ( 1+2\cdot 3^{-c-\frac{1}{2}}) }{\log 4}\), so \(a_{4,c}=-\frac{\log ( 1+2\cdot 3^{-c-\frac{1}{2}}) }{\log 4}\) and consequently for \(c<0\) the function \(f_{4}(c)\) is given by the formula

$$ f_{4}(c)=-\frac{\log ( 1+2\cdot 3^{-c-\frac{1}{2}}) }{\log 4}. $$

Then the fixed points of \(f_{4}(c)\) are the solutions of the equation \(f_{4}(c)=c\), that is

$$\begin{aligned} 1+2\cdot 3^{-c-1/2}=4^{-c}. \end{aligned}$$
(54)

According to [20, p. 46] Eq. (54) has a unique real solution, say \(c_{0}\), whose approached value is \(-1.21\). On the other hand, since \(n=4\) belongs to \(\mathscr {C}^{*}\), by Theorem 3.3 \(a_{\zeta _{4}(z)}\) is a fixed point of the function \(f_{4}\). Since \(c_{0}\) is the unique solution of \(f_{4}(c)=c\), necessarily \(a_{\zeta _{4}(z)}=c_{0}\approx -1.21\) and then \(f_{4}\) has no fixed point in \((a_{\zeta _{4}(z)},a_{\zeta _{4}^{*}(z)})\). Hence the claim follows. Then, by applying Theorem 3.6, \(\zeta _{4}(z)\) has a maximum density interval and consequently

$$ R_{\zeta _{4}(z)}=[ a_{\zeta _{4}(z)},b_{\zeta _{4}(z)}] . $$

For \(n=5\) we take a mapping \(\delta :\mathscr {P}\) \(\rightarrow U\) satisfying \(\delta (2)=\delta (3)=-1\) and consider its corresponding \(\omega _{\delta }:\mathbb {N}\rightarrow U\) defined in (44). Assume \(f_{5}\) has some fixed point, say \(c_{0}\), in the interval \((a_{\zeta _{5}(z)},a_{\zeta _{5}^{*}(z)})\). By (8) \(a_{\zeta _{5}^{*}(z)}<0\) and then \((a_{\zeta _{5}(z)},a_{\zeta _{5}^{*}(z)})\) contains only negative numbers, so \(c_{0}<0\). By applying (45) for \(n=5\), \(f_{5}\) and the above defined \(\omega _{\delta }\), under the assumption \(f_{5}(c_{0})=c_{0}\), we have

$$ 5^{-c_{0}}\le | 1-2^{-c_{0}}-3^{-c_{0}}+4^{-c_{0}}| . $$

But this inequality is clearly impossible for any \(c_{0}<0\). Hence \(f_{5}\) has no fixed point in \((a_{\zeta _{5}(z)},a_{\zeta _{5}^{*}(z)})\). Then, by applying Theorem 3.6, \(\zeta _{5}(z)\) has a maximum density interval and consequently

$$ R_{\zeta _{5}(z)}=[ a_{\zeta _{5}(z)},b_{\zeta _{5}(z)}] . $$

For \(n=6\), we take a mapping \(\delta :\mathscr {P}\) \(\rightarrow U\) satisfying \(\delta (2)=-1\), \(\delta (3)=1\) and consider its corresponding \(\omega _{\delta }:\mathbb {N}\rightarrow U\) defined in (44). Assume \(f_{6}\) has some fixed point, say \(c_{0}\), in the interval \((a_{\zeta _{6}(z)},a_{\zeta _{6}^{*}(z)})\). By (8) \(a_{\zeta _{6}^{*}(z)}<0\) and then \((a_{\zeta _{6}(z)},a_{\zeta _{6}^{*}(z)})\) contains only negative numbers, so \(c_{0}<0\). By applying (45) for \(n=6\), \(f_{6}\) and the above defined \(\omega _{\delta }\), under the assumption \(f_{6}(c_{0})=c_{0}\), we have

$$\begin{aligned} 5^{-c_{0}}\le | 1-2^{-c_{0}}+3^{-c_{0}}+4^{-c_{0}}-6^{-c_{0}}| . \end{aligned}$$
(55)

Regarding inequality (55) we consider the two possible cases: (a) \(1-2^{-c_{0}}+3^{-c_{0}}+4^{-c_{0}}-6^{-c_{0}}\ge 0\), (b) \(1-2^{-c_{0}}+3^{-c_{0}}+4^{-c_{0}}-6^{-c_{0}}<0\). In (a), according to (55), we have the inequality

$$ 1+3^{-c_{0}}+4^{-c_{0}}\ge 2^{-c_{0}}+5^{-c_{0}}+6^{-c_{0}}, $$

that as we easily can check is not possible for any \(c_{0}<0\). In (b), because of (55), we get

$$\begin{aligned} 1+3^{-c_{0}}+4^{-c_{0}}+5^{-c_{0}}\le 2^{-c_{0}}+6^{-c_{0}}. \end{aligned}$$
(56)

By a direct computation we see that (56) is only true for \(c_{0}\le a_{\zeta _{6}(z)}\approx -2.8\) (observe that for \(c_{0}\approx -2.8\), inequality (56) becomes an equality and since \(n=6\) belongs to \(C^{*}\), by Theorem 3.3, \(a_{\zeta _{6}(z)}\) is a fixed point of the function \(f_{6}\)). Therefore for \(c_{0}>a_{\zeta _{6}(z)}\), (56) is not possible. Hence \(f_{6}\) has no fixed point in \((a_{\zeta _{6}(z)},a_{\zeta _{6}^{*}(z)})\). Then, by applying Theorem 3.6, \(\zeta _{6}(z)\) has a maximum density interval and consequently

$$ R_{\zeta _{6}(z)}=[ a_{\zeta _{6}(z)},b_{\zeta _{6}(z)}] . $$

For \(n=7\), we take a mapping \(\delta :\mathscr {P}\) \(\rightarrow U\) satisfying \(\delta (2)=\) \(\delta (3)=\delta (5)=-1\) and consider its corresponding \(\omega _{\delta }:\mathbb {N} \rightarrow U\) defined in (44). Assume \(f_{7}\) has some fixed point, say \(c_{0}\), in the interval \((a_{\zeta _{7}(z)},a_{\zeta _{7}^{*}(z)})\). By (8) \(a_{\zeta _{7}^{*}(z)}<0\) and then \((a_{\zeta _{7}(z)},a_{\zeta _{7}^{*}(z)})\) contains only negative numbers, so \(c_{0}<0\). By applying (45) for \(n=7\), \(f_{7}\) and the above defined \(\omega _{\delta }\), under the assumption \(f_{7}(c_{0})=c_{0}\), we have

$$\begin{aligned} 7^{-c_{0}}\le |1-2^{-c_{0}}-3^{-c_{0}}+4^{-c_{0}}-5^{-c_{0}}+6^{-c_{0}}| . \end{aligned}$$
(57)

We consider the two possible cases: (a) \(1-2^{-c_{0}}-3^{-c_{0}}+4^{-c_{0}}-5^{-c_{0}}+6^{-c_{0}}\ge 0\), (b) \(1-2^{-c_{0}}-3^{-c_{0}}+4^{-c_{0}}-5^{-c_{0}}+6^{-c_{0}}<0\). In (a), according to (57), we have the inequality

$$ 1+4^{-c_{0}}+6^{-c_{0}}\ge 2^{-c_{0}}+3^{-c_{0}}+5^{-c_{0}}+7^{-c_{0}}, $$

that is clearly impossible for any \(c_{0}<0\). In (b), because of (57), we get

$$\begin{aligned} 1+4^{-c_{0}}+6^{-c_{0}}+7^{-c_{0}}\le 2^{-c_{0}}+3^{-c_{0}}+5^{-c_{0}}. \end{aligned}$$
(58)

It is immediate to check that inequality (58) is false for any \(c_{0}<0\). Hence \(f_{7}\) has no fixed point in \((a_{\zeta _{7}(z)},a_{\zeta _{7}^{*}(z)})\). Then, by applying Theorem 3.6, \(\zeta _{7}(z)\) has a maximum density interval and consequently

$$ R_{\zeta _{7}(z)}=[ a_{\zeta _{7}(z)},b_{\zeta _{7}(z)}] . $$

For \(n=8\), we take a mapping \(\delta :\mathscr {P}\) \(\rightarrow U\) satisfying \(\delta (2)=1\), \(\delta (3)=\delta (5)=-1\) and consider its corresponding \(\omega _{\delta }:=\mathbb {N}\rightarrow U\) defined in (44). Assume \(f_{8}\) has some fixed point, say \(c_{0}\), in the interval \((a_{\zeta _{8}(z)},a_{\zeta _{8}^{*}(z)})\). By (8) \(a_{\zeta _{8}^{*}(z)}<0\) and then \((a_{\zeta _{8}(z)},a_{\zeta _{8}^{*}(z)})\) contains only negative numbers, so \(c_{0}<0\). By applying (45) for \(n=8\), \(f_{8}\) and the above defined \(\omega _{\delta }\), under the assumption \(f_{8}(c_{0})=c_{0}\), we have

$$\begin{aligned} 7^{-c_{0}}\le |1+2^{-c_{0}}-3^{-c_{0}}+4^{-c_{0}}-5^{-c_{0}}-6^{-c_{0}}+8^{-c_{0}}|. \end{aligned}$$
(59)

Regarding inequality (59) we consider the two possible cases: (a) \(1+2^{-c_{0}}-3^{-c_{0}}+4^{-c_{0}}-5^{-c_{0}}-6^{-c_{0}}+8^{-c_{0}}<0\), (b) \(1+2^{-c_{0}}-3^{-c_{0}}+4^{-c_{0}}-5^{-c_{0}}-6^{-c_{0}}+8^{-c_{0}}\ge 0\). In case (a), according to (59), we have the inequality

$$ 3^{-c_{0}}+5^{-c_{0}}+6^{-c_{0}}\ge 1+2^{-c_{0}}+4^{-c_{0}}+7^{-c_{0}}+8^{-c_{0}}, $$

which is clearly impossible for any \(c_{0}<0\). In case (b), because of (59), we get

$$\begin{aligned} 1+2^{-c_{0}}+4^{-c_{0}}+8^{-c_{0}}\ge 3^{-c_{0}}+5^{-c_{0}}+6^{-c_{0}}+7^{-c_{0}}. \end{aligned}$$
(60)

By an elementary analysis we can see that (60) is only true for \(c_{0}\le a_{\zeta _{8}(z)}\approx -4.1\) (observe that for \(c_{0}\approx -4.1\) inequality (60) becomes an equality and since \(n=8\) belongs to \(C^{*}\), by Theorem 3.3, \(a_{\zeta _{8}(z)}\approx -4.1\) is a fixed point of the function \(f_{8}\)). Therefore for \(c_{0}\in ( a_{\zeta _{8}(z)},0)\), (60) is not possible. Then, since by (8) \(a_{\zeta _{8}^{*}(z)}<0\), in particular (60) is not possible in \((a_{\zeta _{8}(z)},a_{\zeta _{8}^{*}(z)})\). Hence \(f_{8}\) has no fixed point in the interval \((a_{\zeta _{8}(z)},a_{\zeta _{8}^{*}(z)})\). Then, by applying Theorem 3.6, \(\zeta _{8}(z)\) has a maximum density interval and consequently

$$ R_{\zeta _{8}(z)}=[ a_{\zeta _{8}(z)},b_{\zeta _{8}(z)}] . $$