1 Introduction

Let

$$\begin{aligned} F(s)=\sum _{n=1}^\infty \frac{a(n)}{n^s} \end{aligned}$$

be a Dirichlet series which converges somewhere in the complex plane. It is well known that there are four classical abscissae associated with F(s): the abscissa of convergence \(\sigma _c(F)\), of uniform convergence \(\sigma _u(F)\), of absolute convergence \(\sigma _a(F)\) and of boundedness \(\sigma _b(F)\). It may well be, in general, that \(\sigma _c(F)=-\infty \), in which case the other three abscissae equal \(-\infty \) as well. From the theory of Dirichlet series we know that

$$\begin{aligned} \sigma _c(F)\le \sigma _b(F)= \sigma _u(F)\le \sigma _a(F), \end{aligned}$$

and in general this is best possible, i.e. inequalities cannot be replaced by equalities. We refer to Maurizi and Queffélec [15] for a modern reference for this sort of problems.

Our first result is that \(\sigma _b(F)=\sigma _a(F)\) for an important class of Dirichlet series, namely those defining the L-functions of the Selberg class \({\mathcal S}\). We recall that the axiomatic class \({\mathcal S}\) contains, at least conjecturally, most L-functions from number theory and automorphic forms theory, and that \(\sigma _b(F)=\sigma _a(F)\) is known in some special cases like the Riemann or the Dedekind zeta functions. The Selberg class \({\mathcal S}\) is defined, roughly, as the class of Dirichlet series absolutely convergent for \(\sigma >1\), having analytic continuation to \(\mathbb C\) with at most a pole at \(s=1\), satisfying a functional equation of Riemann type and having an Euler product representation. Moreover, their coefficients satisfy the Ramanujan condition \(a(n)\ll n^\varepsilon \) for any \(\varepsilon >0\). We also recall that the extended Selberg class \({\mathcal S}^\sharp \) is the larger class obtained by dropping the Euler product and Ramanujan condition requirements in the definition of \({\mathcal S}\). We refer to our survey papers [7, 9, 1719] and to the forthcoming book [12] for definitions, examples and the basic theory of the Selberg classes \({\mathcal S}\) and \({\mathcal S}^\sharp \). In particular, we refer to these surveys for the definition of degree \(d_F\), conductor \(q_F\) and standard twist of F(s).

Theorem 1

Suppose that F(s) belongs to the Selberg class. Then

$$\begin{aligned} \sigma _b(F)= \sigma _u(F) = \sigma _a(F). \end{aligned}$$
(1)

Several months after submitting this result, the note by Brevig and Heap [3] appeared, where the authors prove the same theorem in the much more general framework of Dirichlet series with multiplicative coefficients. Trying to understand Brevig-Heap’s proof, based on Bohr’s theory, we noticed that their result was already known to Bohr himself in 1913 (see [1], Satz XI, p. 480); incidentally, Bohr’s paper [1] appears as item [5] of the reference list in Brevig-Heap [3]. We wish to thank Dr. Mattia Righetti for bringing [1, 3] to our attention and for his advice concerning these papers. We decided to keep Theorem 1 since our proof is different, easier and more direct; moreover, some points in the proof will be useful for the other results in the paper.

We expect that actually \(\sigma _a(F)=1\) for all \(F\in {\mathcal S}\). This is known for most classical L-functions and, in the general case of the class \({\mathcal S}\), under the assumption of the Selberg orthonormality conjecture; however, an unconditional proof is missing at present. See again the above quoted references for definitions and results about such a conjecture.

Note that the abscissa of convergence \(\sigma _c(F)\) can be smaller than 1 for functions in \({\mathcal S}\). For example, the Dirichlet L-functions \(L(s,\chi )\) with a primitive non-principal character \(\chi \) are convergent in the half-plane \(\sigma >0\). Actually, several general results are known about the abscissa \(\sigma _c(F)\) for functions F(s) in the extended Selberg class \({\mathcal S}^\sharp \). First of all

$$\begin{aligned} {{ if} F\in {\mathcal S}^\sharp \,\, { is} \,\, { entire}\,\, { with} \,\, { degree} \,\, d\ge 1, { then} \ \frac{1}{2} - \frac{1}{2d} \le \sigma _c(F) \le 1-\frac{2}{d+1}} \end{aligned}$$
(2)

(recall that there exist no functions \(F\in {\mathcal S}^\sharp \) with degree \(0<d<1\), see [8] and Conrey and Ghosh [5]). Indeed, the first inequality in (2) is Corollary 3 in [11] and is based on the properties of the standard twist, while the second inequality follows from a well known theorem of Landau [13]. Moreover, in accordance with classical degree 2 conjectures and with the general \(\Omega \)-theorem in Corollary 2 of [11], we expect that equality holds in the left inequality in (2). Further

$$\begin{aligned} \sigma _c(F)=-\infty \quad \,{ if} \,\,{ and} \,\,{ only} \,\,{ if} \quad \,d_F=0, \end{aligned}$$

since the degree 0 functions of \({\mathcal S}^\sharp \) are Dirichlet polynomials (see [8]). From (2) we also deduce that

$$\begin{aligned} \sigma _c(F)=1\,\, \textit{if}\, \textit{and} \,\textit{only} \,\textit{if} \, F(s) \, \textit{has} \,\,a \,\,\textit{pole} \,\,\textit{at} \,s=1. \end{aligned}$$

We also remark that if the Lindelöf Hypothesis holds for \(F\in {\mathcal S}^\sharp \), then \(\sigma _c(F)\le 1/2\).

The behavior of \(\sigma _a(F)\) in the extended class \({\mathcal S}^\sharp \) is different from the expected behavior in \({\mathcal S}\). Indeed, in the next section, which is also of independent interest, we show that

$$\begin{aligned} \textit{there} \, \textit{exist} \,\textit{functions} \,F\in {\mathcal S}^\sharp \quad \textit{with}\, \sigma _a(F) \, \textit{arbitrarily}\, \textit{close} \,\textit{to} \, 1/2. \end{aligned}$$

We conclude this section with the following

Question

Does (1) hold for the functions in the extended Selberg class ? \(\square \)

A variant of the question is: does (1) hold for linear combinations

$$\begin{aligned} F(s) = \sum _{j=1}^N c_jF_j(s) \end{aligned}$$

with \(F_j\in {\mathcal S}\) and \(c_j\in \mathbb C\)? If needed, one may assume that F(s) belongs to \({\mathcal S}^\sharp \).

Since \(\sigma _a(F)=1\) for most classical L-functions F(s), Theorem 1 prevents the possibility of getting information on the non-trivial zeros exploiting the properties of the abscissa of uniform convergence. On the other hand, if \(F\in {\mathcal S}\) is bounded for \(\sigma >1-\delta \) for some \(\delta >0\), then its Dirichlet series is absolutely convergent for \(\sigma >1-\delta \) and hence \(F(s)\ne 0\) by Euler’s identity. In the next theorems we replace boundedness by more general majorants and deduce some consequences.

Let \(F\in {\mathcal S}\) be of degree d, \(N_F(\sigma ,T)\) be the number of zeros \(\rho =\beta +i\gamma \) with \(\beta >\sigma \) and \(|\gamma |\le T\), and denote the density abscissa \(\sigma _D(F)\) by

$$\begin{aligned} \sigma _D(F) = \inf \{\sigma : N_F(\sigma ,T)=o(T)\}. \end{aligned}$$

An inspection of the proof of Lemma 3 in [10], obtained by a rudimentary version of Montgomery’s zero-detecting method, shows that

$$\begin{aligned} N_F(\sigma ,T) \ll T^{4(d+3)(1-\sigma )+\varepsilon }. \end{aligned}$$

Hence in general

$$\begin{aligned} \frac{1}{2} \le \sigma _D(F) \le 1-\frac{1}{4(d+3)}, \end{aligned}$$

although it is well known that the classical L-functions F(s) of degree 1 and 2 have \(\sigma _D(F)=1/2\), see e.g. Luo [14]. Actually, one can prove that \(\sigma _D(F)=1/2\) for all \(F\in {\mathcal S}\) with degree \(0<d\le 2\). Further, let f(s) be holomorphic in \(\sigma >1-\delta \) for some \(\delta >0\) and almost periodic on the line \(\sigma =A\) for some \(A>1\). We say that f(s) is a \(\delta \)-almost periodic majorant of F(s) if

$$\begin{aligned} |F(s)|\le c(\sigma )|f(s)| \end{aligned}$$
(3)

in the half-plane \(\sigma >1-\delta \), where \(c(\sigma )>0\) is a continuous function for \(\sigma >1-\delta \).

Theorem 2

Let \(F\in {\mathcal S}\) and f(s) be a \(\delta \)-almost periodic majorant of F(s). Then F(s) and f(s) have the same zeros, with the same multiplicity, in the half-plane \(\sigma >\max (1-\delta ,\sigma _D(F))\).

Remark

Clearly, in view of (3) each zero of f(s) is also a zero of F(s); the non-trivial part of Theorem 2 says that the opposite assertion holds true as well. Note that we do not require that f(s) is almost periodic for \(\sigma >1-\delta \), but only on some vertical line far on the right. We already noticed that, as a consequence of Theorem 1, \(F(s)\ne 0\) in every right half-plane where it is bounded. An immediate consequence of Theorem 2 is that \(F(s)\ne 0\) for \(\sigma >\max (1-\delta ,\sigma _D(F))\) if f(s) is a non-vanishing \(\delta \)-almost periodic majorant. In particular, from the density estimates reported above when \(d\le 2\), if \(\delta =1/2\) then the Riemann Hypothesis holds for such F(s). \(\square \)

Our final result is a kind of new independence statement for L-functions from the Selberg class. Several forms of independence are known in \({\mathcal S}\), such as the linear independence, the multiplicity one property and the orthogonality conjecture and some of its consequences; see our above quoted surveys on the Selberg class. The new independence result is expressed in terms of majorants as follows.

Theorem 3

Let \(F,G\in {\mathcal S}\) be such that \(F(s)\ll |G(s)|\) for \(\sigma >1/2\). Then \(F(s)=G(s)\).

The special nature of the majorant is very important here. Indeed, suppose that G(s) is entire; then Theorem 2 gives only that F(s) and G(s) have the same zeros for \(\sigma >\sigma _D(F)\). Instead, exploiting the information that \(G\in {\mathcal S}\), Theorem 3 shows that actually \(F(s)=G(s)\). In other words, no function from \({\mathcal S}\) can dominate in \(\sigma >1/2\) another function from \({\mathcal S}\). We may regard this as a weak form of a well known result obtained, under stronger assumptions, by Selberg [20] and Bombieri and Hejhal [2] about the statistical independence of the values of L-functions.

2 The lift operator

Let \(Q>0\), \({\varvec{\lambda }}=(\lambda _1,\ldots ,\lambda _r)\) with \(\lambda _j>0\), \({\varvec{\mu }}=(\mu _1,\ldots ,\mu _r)\) with \(\mu _j\in \mathbb C\) and \(|\omega |=1\). We denote by \(W(Q,{\varvec{\lambda }},{\varvec{\mu }},\omega )\) the \(\mathbb R\)-linear space of the Dirichlet series F(s) solutions of the functional equation

$$\begin{aligned} Q^s \prod _{j=1}^r\Gamma (\lambda _js+\mu _j)F(s) = \omega Q^{1-s} \prod _{j=1}^r\Gamma (\lambda _j(1-s)+\overline{\mu _j})\overline{F(1-\overline{s})}. \end{aligned}$$
(4)

Given an integer \(k\ge 1\), we define the k-lift operator by

$$\begin{aligned} F(s)\longmapsto F_k(s)=F\left( ks+\frac{1-k}{2}\right) ; \end{aligned}$$

clearly, the operator is trivial for \(k=1\). A simple computation shows that

$$\begin{aligned} if F\in W(Q,{\varvec{\lambda }},{\varvec{\mu }},\omega ) \quad then \quad F_k\in W\left( Q^k,k{\varvec{\lambda }},{\varvec{\mu }}+ \frac{1-k}{2}\mathbf {\varvec{\lambda }},\omega \right) . \end{aligned}$$
(5)

In particular, from (5) we have that degree \(d_{F_k}\) and conductor \(q_{F_k}\) of \(F_k(s)\) satisfy

$$\begin{aligned} d_{F_k} = kd_F \quad q_{F_k} = q_F^k k^{kd_F}. \end{aligned}$$
(6)

We recall (see the above references) that the class \({\mathcal S}^\sharp \) consists of the Dirichlet series satisfying a functional equation of type (4), where now \(\mathfrak {R}{\mu _j}\ge 0\), with the following properties: F(s) is absolutely convergent for \(\sigma >1\) and \((s-1)^mF(s)\) is entire of finite order for some integer \(m\ge 0\). Therefore we consider

$$\begin{aligned} B_F= 2\min _{1\le j\le r} \frac{\mathfrak {R}{\mu _j}}{\lambda _j}+1, \end{aligned}$$

which is an invariant of \({\mathcal S}^\sharp \) (see again the above references) since a simple computation shows that

$$\begin{aligned} B_F=-2\max _{\rho } \mathfrak {R}{\rho } + 1, \end{aligned}$$

where \(\rho \) runs over the trivial zeros of F(s). From the definition of the k-lift operator and (5) we see that, given \(F\in {\mathcal S}^\sharp \), the lifted function \(F_k(s)\) also belongs to \({\mathcal S}^\sharp \) provided \(1\le k\le B_F\) and, if \(B_F\ge 2\), F(s) is entire. Indeed, if \(k\ge 2\), F(s) has to be holomorphic at \(s=1\) otherwise the pole of \(F_k(s)\) is not at \(s=1\), and the bound \(k\le B_F\) is needed to have non-negative real part of the \(\mu \)’s data of \(F_k(s)\). Therefore, defining \(V(Q,{\varvec{\lambda }},{\varvec{\mu }},\omega )\) to be the \(\mathbb R\)-linear space of the entire functions \(F\in {\mathcal S}^\sharp \) satisfying (4) (again with \(\mathfrak {R}{\mu _j}\ge 0\)), we have that

$$\begin{aligned} for\,\, 1\le & {} k\le B_F, the\,\,k-lift \,\,operator \,\,maps\,\, V(Q,{\varvec{\lambda }},{\varvec{\mu }},\omega )\,\, into \,\, \\&\quad V\left( Q^k,k{\varvec{\lambda }},{\varvec{\mu }}+\frac{1-k}{2}\mathbf {\varvec{\lambda }},\omega \right) . \end{aligned}$$

Note that \(B_F\) depends only on \({\varvec{\lambda }}\) and \({\varvec{\mu }}\), so it is the same for all functions in \(V(Q,{\varvec{\lambda }},{\varvec{\mu }},\omega )\). Note also that the Selberg class \({\mathcal S}\) is not preserved under the above mappings since the Ramanujan condition is not (necessarily) satisfied by \(F_k(s)\) even if F(s) does; see the examples below. Further, a simple computation shows that the k-lift operator commutes with the map sending F(s) to its standard twist. We also remark that the requirement \(\mathfrak {R}{\mu _j}\ge 0\) in the definition of \({\mathcal S}^\sharp \), which is responsible for the limitation \(k\le B_F\) in (6), is apparently not of primary importance in the theory of the Selberg class. Hence, although formally not belonging to \({\mathcal S}^\sharp \), the lifts \(F_k(s)\) of entire \(F\in {\mathcal S}^\sharp \) with \(k>B_F\) are further examples of Dirichlet series with continuation over \(\mathbb C\) and functional equation. A similar remark applies to the other condition in the definition of \(V(Q,{\varvec{\lambda }},{\varvec{\mu }},\omega )\), namely the holomorphy at \(s=1\).

Example

The Riemann zeta function \(\zeta (s)\) cannot be lifted inside \({\mathcal S}^\sharp \) since it has \(B_\zeta =1\). The same holds for the Dirichlet L-functions with even primitive characters, while those associated with odd primitive characters may be lifted inside \({\mathcal S}^\sharp \) for \(k=2\) and \(k=3\). However, after lifting their Dirichlet coefficients do not satisfy the Ramanujan condition, hence the lifted Dirichlet L-functions do not belong to \({\mathcal S}\). Note that, once suitably normalized, the lifts with \(k=2\) become the L-functions associated with half-integral weight modular forms; see the books by Hecke [6] and Ogg [16]. Concerning degree 2, we consider the L-functions associated with holomorphic eigenforms of level N and integral weight K; see Ogg [16]. Denoting by F(s) their normalization satisfying a functional equation reflecting \(s\mapsto 1-s\) (instead of the original \(s\mapsto K-s\)), we have that

$$\begin{aligned} B_F=K. \end{aligned}$$

Hence the normalized L-functions of eigenforms of weight K may be lifted inside \({\mathcal S}^\sharp \) with k up to their weight. Here we consider only eigenforms since in general the L-functions of modular forms of level N satisfy a slightly different functional equation, not of \({\mathcal S}^\sharp \) type.

\(\square \)

We finally turn to the problem of the absolute convergence abscissa in \({\mathcal S}^\sharp \). Let \(F\in {\mathcal S}^\sharp \) be of degree \(d\ge 1\). Then, thanks again to the properties of the standard twist, we know that

$$\begin{aligned} \sigma _a(F) \ge \frac{1}{2} + \frac{1}{2d}; \end{aligned}$$
(7)

this folows from Theorem 1 of [11]. On the other hand, if \(F\in {\mathcal S}^\sharp \) we have that the series

$$\begin{aligned} \sum _{n=1}^\infty \frac{|a(n)|}{n^{k\sigma +(1-k)/2}} \end{aligned}$$

converges for \(\sigma >1/2 + 1/(2k)\). Hence from (6) and (7) we obtain that if both F(s) of degree \(d\ge 1\) and \(F_k(s)\) belong to \({\mathcal S}^\sharp \), then

$$\begin{aligned} \frac{1}{2} + \frac{1}{2kd} \le \sigma _a(F_k) \le \frac{1}{2} +\frac{1}{2k}. \end{aligned}$$
(8)

Since the above examples show that there exist functions \(F\in {\mathcal S}^\sharp \) with arbitrarily large \(B_F\) (e.g. the holomorphic eigenforms with arbitrarily large weight K), (8) shows that \(\sigma _a(F)\) can be arbitrarily close to 1 / 2 inside \({\mathcal S}^\sharp \). Hence the behavior of \(\sigma _a(F)\) in the extended class \({\mathcal S}^\sharp \) is definitely different from its expected behavior in the class \({\mathcal S}\).

3 Proof of Theorem 1

Observe that the case \(d=0\) is trivial, since F(s) is identically 1; see Conrey and Ghosh [5]. For d positive we have \(\sigma _b(F)\ge 1/2\), since F(s) is unbounded for \(\sigma <1/2\) by the functional equation and the properties of the \(\Gamma \) function. Therefore, to prove the assertion it suffices to show the following fact: if for a certain \(1/2<\sigma _0\le 1\) the function F(s) is bounded for \(\sigma >\sigma _0\), then \(\sigma _a(F)\le \sigma _0\).

Let us fix an \(\varepsilon \in (0,\sigma _0-1/2)\), and let \(c_0=c_0(\varepsilon )\) be such that \(|a(n)|\le c_0n^{\varepsilon \slash 2}\) for all \(n\ge 1\). Without loss of generality we may assume that \(c_0\ge 3\). Consider the finite set of primes

$$\begin{aligned} S_{\varepsilon }=\{p: |a(p)|>p^{\varepsilon \slash 2} \ \ \text {or}\ \ p< c_0^{2\slash \varepsilon }\}. \end{aligned}$$

Let

$$\begin{aligned} F_p(s) = \sum _{m=0}^\infty \frac{a(p^m)}{p^{ms}} \end{aligned}$$
(9)

denote the pth Euler factor of F(s). We split the Euler product as

$$\begin{aligned} \begin{aligned} F(s)&=\prod _{p\not \in S_{\varepsilon }}\left( 1+\frac{a_F(p)}{p^s}\right) \prod _{p\in S_{\varepsilon }}F_p(s) \prod _{p\not \in S_{\varepsilon }}\left( F_p(s)\left( 1+\frac{a_F(p)}{p^s}\right) ^{-1}\right) \\&=P_1(s)P_2(s)P_3(s), \end{aligned} \end{aligned}$$
(10)

say. Both \(P_2(s)\) and its inverse \(1\slash P_2(s)\) have Dirichlet series representations which converge absolutely for \(\sigma >\theta \) for some \(\theta <1\slash 2\). This is a simple consequence of the definition of the Selberg class; see the above quoted references. Therefore, \(P_2(s)\) and \(1\slash P_2(s)\) are bounded for \(\sigma >\sigma _0\).

In view of (9) we have

$$\begin{aligned} P_3(s)=\prod _{p\not \in S_{\varepsilon }}\left( 1+\sum _{m=2}^{\infty }\frac{b(p^m)}{p^{ms}}\right) \end{aligned}$$

with

$$\begin{aligned} b(p^m)=\sum _{l=0}^m(-1)^la(p)^la(p^{m-l}). \end{aligned}$$

Hence, recalling that \(p\not \in S_{\varepsilon }\), \(m\ge 2\) and \(c_0\ge 3\), we have

$$\begin{aligned} |b(p^m)|\le \sum _{l=0}^m|a(p)|^l|a(p^{m-l})|\le c_0mp^{m\varepsilon \slash 2}\le p^{m\varepsilon }. \end{aligned}$$

Thus for \(\sigma >1/2+\varepsilon \) and \(p\not \in S_{\varepsilon }\) we have

$$\begin{aligned} \sum _{m=2}^{\infty }\frac{|b(p^m)|}{p^{m\sigma }} <1 \quad \text {and} \quad \sum _{p\not \in S_{\varepsilon }} \sum _{m=2}^{\infty }\frac{|b(p^m)|}{p^{m\sigma }}\ll 1. \end{aligned}$$

Hence both \(P_3(s)\) and \(1\slash P_3(s)\) are bounded and have Dirichlet series representations which converge absolutely for \(\sigma >\sigma _0\) (recall that \(\sigma _0>1/2+\varepsilon \)).

We therefore see that \(P_1(s)=F(s)\slash (P_2(s)P_3(s))\) is bounded for \(\sigma >\sigma _0\). Let us write

$$\begin{aligned} P_1(s)=\sum _{n=1}^{\infty }\frac{c(n)}{n^s}. \end{aligned}$$

The coefficients c(n) are completely multiplicative, and the series converges for \(\sigma >\sigma _0\). Fix such a \(\sigma \), and a positive \(\delta <\sigma -\sigma _0\). Consider the following familiar Mellin’s transform

$$\begin{aligned} \sum _{n=1}^{\infty }\frac{c(n)}{n^{\sigma +it}}e^{-n\slash Y} = \frac{1}{2\pi i}\int _{1-i\infty }^{1+i\infty } \frac{F(w+\sigma +it)}{P_2(w+\sigma +it)P_3(w+\sigma +it)} \Gamma (w) Y^w \text {d}w. \end{aligned}$$

We shift the line of integration to \(\mathfrak {R}(w)=-\delta \) and obtain

$$\begin{aligned} \sum _{n=1}^{\infty }\frac{c(n)}{n^{\sigma +it}}e^{-n\slash Y} = \frac{F(\sigma +it)}{P_2(\sigma +it)P_3(\sigma +it)} + O(Y^{-\delta })\ll 1 \end{aligned}$$

uniformly in \(t\in {\mathbb R}\) and \(Y\ge 1\). Since \(|c(n)|\le n^{\varepsilon /2}\), due to the decay of the exponential we may cut the sum on the left hand side to \(n\le 3Y\log Y\), say, producing an extra error term of size \(O(1\slash Y)\). Thus

$$\begin{aligned} \sum _{n\le 3Y\log Y}\frac{c(n)}{n^{\sigma +it}}e^{-n\slash Y} \ll 1 \end{aligned}$$
(11)

uniformly in \(t\in \mathbb R\) and \(Y\ge 1\).

Now we apply Kronecker’s theorem in the following form, see Theorem 8 of Ch. VIII of Chandrasekharan [4]. If \(\theta _1,\ldots ,\theta _k\in \mathbb R\) are linearly independent over \(\mathbb Z\), \(\beta _1,\ldots ,\beta _k\in \mathbb R\) and \(T,\eta >0\), then there exist \(t>T\) and \(n_1,\ldots ,n_k\in \mathbb Z\) such that

$$\begin{aligned} |t\theta _\ell - n_\ell -\beta _\ell |<\eta \qquad \ell =1,\ldots ,k. \end{aligned}$$
(12)

We choose the \(\theta \)’s as \(-\frac{1}{2\pi }\log p\) with the primes \(p\le 3Y\log Y\) not in \(S_{\varepsilon }\) and, correspondingly, the \(\beta \)’s such that \(|c(p)|=c(p)e^{2\pi i\beta _p}\) for each such p. Hence by (12) there exists a sequence of real numbers \(t_\nu \rightarrow +\infty \) such that

$$\begin{aligned} c(p)p^{-it_\nu } \rightarrow |c(p)| \qquad \nu \rightarrow \infty \end{aligned}$$

uniformly for the primes \(p\le 3Y\log Y\) not in \(S_\varepsilon \). By the complete multiplicativity of c(n) we infer that

$$\begin{aligned} c(n)n^{-it_{\nu }} \rightarrow |c(n)| \quad \quad \nu \rightarrow \infty \end{aligned}$$

uniformly for \(n\le 3Y\log Y\). Thus putting \(t=t_{\nu }\) in (11) and making \(\nu \rightarrow \infty \) we obtain

$$\begin{aligned} \sum _{n\le Y}\frac{|c(n)|}{n^{\sigma }}\le e\sum _{n\le 3Y\log Y}\frac{|c(n)|}{n^{\sigma }}e^{-n\slash Y} = e\lim _{\nu \rightarrow \infty } \sum _{n\le 3Y\log Y}\frac{c(n)}{n^{\sigma +it_{\nu }}}e^{-n\slash Y} \ll 1 \end{aligned}$$

uniformly for \(Y\ge 1\). Letting \(Y\rightarrow \infty \), we see that the Dirichlet series of \(P_1(s)\) converges absolutely for \(\sigma >\sigma _0\).

Summarizing, we have shown that the Dirichlet series of \(P_1(s)\), \(P_2(s)\) and \(P_3(s)\) are absolutely convergent for \(\sigma >\sigma _0\), hence the Dirichlet series of F(s) is also absolutely convergent for \(\sigma >\sigma _0\) thanks to (10), and the result follows.

4 Proof of Theorem 2

As in Theorem 1 the case \(d=0\) is trivial, hence we assume \(d>0\) and consider the function

$$\begin{aligned} h(s) = \frac{F(s)}{f(s)} \end{aligned}$$

for \(\sigma >1-\delta \). From (3) we have that h(s) is holomorphic for \(\sigma >1-\delta \), bounded on every closed vertical strip inside \(\sigma >1-\delta \) and almost periodic on the line \(\sigma =A\). For a given \(\varepsilon >0\), let \(\tau \) be an \(\varepsilon \)-almost period of \(h(A+it)\), namely for every \(t\in \mathbb R\)

$$\begin{aligned} |h(A+i(t+\tau )) - h(A+it)| < \varepsilon . \end{aligned}$$

Then, by the convexity following from Phragmén-Lindelöf’s theorem applied to \(h(s+i\tau )-h(s)\), given \(\eta >1-\delta \) and any \(\eta<\sigma <A\) we have

$$\begin{aligned} \begin{aligned} \sup _{t\in \mathbb R} |h(\sigma +i(t+\tau )) - h(\sigma +it)|&\le \big (\sup _{t\in \mathbb R} |h(\eta +i(t+\tau )) - h(\eta +it)|\big )^{\frac{A-\sigma }{A-\eta }} \\&\quad \times \,\big (\sup _{t\in \mathbb R} |h(A+i(t+\tau )) - h(A+it)|\big )^{\frac{\sigma -\eta }{A-\eta }}. \end{aligned} \end{aligned}$$

Hence we obtain that

$$\begin{aligned} \sup _{t\in \mathbb R} |h(\sigma +i(t+\tau )) - h(\sigma +it)| \ll \varepsilon ^c \end{aligned}$$

uniformly in any closed strip contained in \(\eta<\sigma <A\), where \(c>0\) depends on the strip. Since \(\varepsilon \) is arbitrarily small, h(s) is uniformly almost periodic in such strips. Suppose now that \(h(\rho )=0\) for some \(\rho \) with \(\mathfrak {R}{\rho }>1-\delta \). Then by a well known argument based on Rouché’s theorem we have that for any \(1-\delta<\eta <\mathfrak {R}{\rho }\)

$$\begin{aligned} T \ll N_h(\eta ,T) \le N_F(\eta ,T) =o(T) \end{aligned}$$

if \(\eta >\sigma _D(F)\), a contradiction. Thus \(h(s)\ne 0\) for \(\sigma >\max (1-\delta ,\sigma _D(F))\), hence every zero of F(s) in this half-plane is a zero of f(s). Theorem 2 is therefore proved, since the opposite implication is a trivial consequence of (3).

5 Proof of Theorem 3

Again the case \(d=0\) is trivial, since in this case \(F(s)\equiv 1\) and so G(s) does not vanish inside the critical strip, thus its degree is 0 and hence \(G(s)\equiv 1\) as well. Let \(F,G\in {\mathcal S}\) be with positive degrees and coefficients \(a_F(n)\) and \(a_G(n)\), respectively, and consider the function

$$\begin{aligned} H(s)=\frac{F(s)}{G(s)} = \sum _{n=1}^\infty \frac{h(n)}{n^s}, \end{aligned}$$

say. By our hypothesis H(s) is bounded, and hence holomorphic, for \(\sigma >1/2\). We modify the proof of Theorem 1 at several points. By Lemma 1 of [10] we have that for every \(\varepsilon >0\) there exists an integer \(K=K(\varepsilon )\) such that the coefficients \(a^{-1}_G(n)\) of 1 / G(s) satisfy

$$\begin{aligned} a^{-1}_G(n) \ll n^\varepsilon \quad \quad \quad (n,K)=1, \end{aligned}$$

and hence

$$\begin{aligned} h(n)\ll n^{2\varepsilon } \quad \quad \quad (n,K)=1. \end{aligned}$$

Therefore the set

$$\begin{aligned} S = \{p: |h(p^m)|>p^{m/10} \quad \text {for some} \quad m\ge 1 \text { or } p\le 10^4\} \end{aligned}$$

is finite and we write

$$\begin{aligned} \begin{aligned} H(s)&= \prod _p \frac{F_p(s)}{G_p(s)} = \prod _p H_p(s) \\&=\prod _{p\not \in S}\left( 1+\frac{h(p)}{p^s} + \frac{h(p^2)}{p^{2s}}\right) \prod _{p\in S}H_p(s) \prod _{p\not \in S}\left( H_p(s)\left( 1+\frac{h(p)}{p^s} + \frac{h(p^2)}{p^{2s}}\right) ^{-1}\right) \\&=Q_1(s)Q_2(s)Q_3(s), \end{aligned} \end{aligned}$$
(13)

say. As in the prof of Theorem 1, \(Q_2(s)\) and \(1/Q_2(s)\) are holomorphic and bounded for \(\sigma \ge 1/2\). Moreover we have

$$\begin{aligned} Q_3(s) = \prod _{p\not \in S}\left( 1 + \frac{\sum _{m=3}^\infty \frac{h(p^m)}{p^{ms}}}{1+\frac{h(p)}{p^s} + \frac{h(p^2)}{p^{2s}}}\right) = \prod _{p\not \in S}\left( 1+ \sum _{m=3}^\infty \frac{k(p^m)}{p^{ms}}\right) , \end{aligned}$$

say, and a computation shows that for \(\sigma \ge 1/2\)

$$\begin{aligned} \sum _{m=3}^\infty \frac{|k(p^m)|}{p^{m\sigma }} \le \frac{1}{3} \ \ \quad \text {for every }\quad p\not \in S \quad {\text{ and }} \quad \sum _{p\not \in S} \sum _{m=3}^\infty \frac{|k(p^m)|}{p^{m\sigma }} \ll 1. \end{aligned}$$

Therefore, no factor of the product vanishes, and \(Q_3(s)\) and \(1/Q_3(s)\) are holomorphic and bounded for \(\sigma \ge 1/2\) as well.

In order the treat \(Q_1(s)\) we need the following elementary lemma.

Lemma

For every \(a,b\in \mathbb C\) there exists \(\theta \in \mathbb C\) with \(|\theta |=1\) such that

$$\begin{aligned} |1+\theta a + \theta ^2b| \ge 1 + \frac{1}{24}(|a|+|b|). \end{aligned}$$

Proof

Suppose first that \(|a|\le |b|/2\). Then

$$\begin{aligned} \max _{|\theta |=1} |1+\theta a + \theta ^2b| \ge 1+|b|-|a| \ge 1+\frac{1}{2}|b| \ge 1+\frac{1}{3}(|a|+|b|), \end{aligned}$$

and the result follows in this case. In the opposite case \(|a|> |b|/2\) we apply the maximum modulus principle to the function \(f(z) = 1 +az+bz^2\), thus obtaining

$$\begin{aligned} \begin{aligned} \max _{|\theta |=1} |1+\theta a + \theta ^2b|&\ge \max _{|\theta |=1} \left| 1+\frac{1}{4}\theta a +\frac{1}{16} \theta ^2b\right| \\&\ge 1+\frac{1}{4}|a| -\frac{1}{16}|b| \ge 1+ \frac{1}{24} (|a|+|b|), \end{aligned} \end{aligned}$$

and the Lemma follows. Note that the constant 1 / 24 is neither optimal nor important in what follows; moreover, in general it cannot be made arbitrarily close to 1. \(\square \)

From (13), our hypothesis and the above information on \(Q_2(s)\) and \(Q_3(s)\) we deduce that there exists \(M>0\) such that for \(\sigma >1/2\)

$$\begin{aligned} |Q_1(s)| = \prod _{p\not \in S}\left| 1+p^{-it}\frac{h(p)}{p^\sigma } +p^{-2it} \frac{h(p^2)}{p^{2\sigma }}\right| \le M. \end{aligned}$$

By the Lemma, for every \(\sigma \) and p there exists \(|\theta _{p,\sigma }|=1\) such that

$$\begin{aligned} \left| 1+\theta _{p,\sigma }\frac{h(p)}{p^\sigma } +\theta _{p,\sigma }^2\frac{h(p^2)}{p^{2\sigma }}\right| \ge 1 + \frac{1}{24}\left( \frac{|h(p)|}{p^\sigma } +\frac{|h(p^2)|}{p^{2\sigma }}\right) . \end{aligned}$$

Assuming that \(\sigma >1/2\) and \(p\not \in S\), applying Kronecker’s theorem as in the last part of the proof of Theorem 1 we find that

$$\begin{aligned} \prod _{p\not \in S} \left( 1 + \frac{1}{24}\left( \frac{|h(p)|}{p^\sigma } +\frac{|h(p^2)|}{p^{2\sigma }} \right) \right) \le M. \end{aligned}$$

Then, letting \(\sigma \rightarrow 1/2^+\), we deduce that the product

$$\begin{aligned} \prod _{p\not \in S} \left( 1 + \frac{1}{24}\left( \frac{|h(p)|}{p^{1/2}} +\frac{|h(p^2)|}{p} \right) \right) \end{aligned}$$

is convergent. Thus the series

$$\begin{aligned} \sum _{p\not \in S} \left( \frac{|h(p)|}{p^{1/2}} +\frac{|h(p^2)|}{p} \right) \end{aligned}$$

is convergent as well and, in turn, the product

$$\begin{aligned} \prod _{p\not \in S} \left( 1 + \left( \frac{|h(p)|}{p^{1/2}} +\frac{|h(p^2)|}{p} \right) \right) \end{aligned}$$

converges. Hence \(Q_1(s)\) and \(Q_1(s)^{-1}\) are non-vanishing for \(\sigma \ge 1/2\).

From (13) and the above properties of \(Q_j(s)\), \(j=1,2,3\), we immediately see that H(s) is holomorphic and non-vanishing for \(\sigma \ge 1/2\). Denoting by \(\gamma _F(s)\) and \(\gamma _G(s)\) the \(\gamma \)-factors of F(s) and G(s), thanks to the functional equation we deduce that

$$\begin{aligned} \frac{\gamma _F(s)}{\gamma _G(s)}H(s) \end{aligned}$$

is a non-vanishing entire function of order \(\le 1\), and hence by Hadamard’s theory we have

$$\begin{aligned} H(s) = \frac{\gamma _G(s)}{\gamma _F(s)} e^{as+b} \end{aligned}$$
(14)

with some \(a,b\in \mathbb C\). Now we can conclude by means of the almost periodicity argument that we used in our proof of the multiplicity one property of \({\mathcal S}\). For this we refer to Lemma 2.1 of [9] and to Theorem 2.3.2 of [7]; in particular, (14) is exactly the last displayed formula of p. 167 of [7]. This way we get that \(H(s)\equiv 1\), hence Theorem 3 is proved.