1 Introduction and statement of main results
1.1 Li coefficients
The Riemann hypothesis (RH) is a critical question in analytic number theory. As such, it is interesting to examine different ways to frame it, which may shed more light on its resolution. In 1997, Xian-Jin Li has discovered a new positivity criterion for the RH. In [Reference Li10], he defined the Li coefficients for the Riemann zeta function as
where $\xi $ is the completed Riemann zeta function defined by
which satisfies $\xi (s)=\xi (1-s)$ and gave a simple equivalence criterion for the RH: RH is true if and only if these coefficients are nonnegative for every positive integer n. The Li coefficients $\lambda _{n}$ can be written as follows:
where the sum runs over the nontrivial zeros of the Riemann zeta function counted with multiplicity. This criterion is generalized by Bombieri and Lagarias [Reference Bombieri and Lagarias4] for any arbitrarily multiset of numbers assuming certain convergence conditions. Voros [Reference Voros19, Section 3.3] has proved that the RH true is equivalent to the growth of $\lambda _{n}$ as $\frac {1}{2}n \log n$ determined by its archimedean part, while the RH false is equivalent to the oscillations of $\lambda _{n}$ with exponentially growing amplitude, determined by its finite part. The Li coefficients were generalized in two ways: by generalizing these coefficients to various sets of functions (the Selberg class, the class of automorphic L-functions, zeta function on function fields,…[Reference Lagarias8, Reference Mazhouda and Smajlović11, Reference Smajlović17]) and by introducing new parameter in its definition (see [Reference Mazhouda and Sodaïgui12]). The Li coefficients (and its generalizations) have generated a lot of research interest due to its applicability and simplicity.
1.2 Quadrilateral zeta function
Recall the definitions of Hurwitz and periodic zeta functions. The Hurwitz zeta function $\zeta (s,a)$ is defined by the series
The function $\zeta (s,a)$ is a meromorphic function with a simple pole at $s=1$ whose residue is $1$ (see, for example, [Reference Apostol1, Section 12]). The periodic zeta function ${\mathrm {Li}}_s (e^{2\pi ia})$ is defined by
(see, for instance, [Reference Apostol1, Exercise 12.2]). Note that the function ${\mathrm {Li}}_s (e^{2\pi ia})$ with $0<a<1$ is analytically continuable to the whole complex plane since ${\mathrm {Li}}_s (e^{2\pi ia})$ does not have any pole, that is shown by the fact that the Dirichlet series of ${\mathrm {Li}}_s (e^{2\pi ia})$ converges uniformly in each compact subset of the half-plane $\sigma>0$ when $0<a<1$ (see, for example, [Reference Laurinčikas and Garunkštis9, p. 20]). For $0 <a \le 1/2$ , we define zeta functions
We can see that $Q(s,a)$ is meromorphic functions with a simple pole at $s=1$ . In addition, we have $Q(0,a)=-1/2=\zeta (0)$ and $\xi _Q(s,a) = \xi _Q(1-s,a)$ , which is proved by
(see [Reference Nakamura13, Theorem 1.1]). Moreover, the function $Q(s,a)$ has the following properties. When $a=1/6, 1/4, 1/3$ , and $1/2$ , the RH holds true if and only if all nonreal zeros of $Q(s,a)$ are on the line ${\mathrm {Re}}(s)=1/2$ (see [Reference Nakamura14, Proposition 1.3]). Let $N_{\mathrm {Q}}^{\mathrm {CL}} (T)$ be the number of the zeros of $Q(s,a)$ on the line segment from $1/2$ to $1/2 +iT$ . In [Reference Nakamura13, Theorem 1.2], the third author proved that for any $0 < a \le 1/2$ , there exist positive constants $A(a)$ and $T_0(a)$ such that
Next, let $N_{F}(T)$ count the number of nonreal zeros of a function $F(s)$ having $|{\mathrm {Im}}(s)|<T$ . Then, for any $0<a\leq 1/2$ ,
and the third author [Reference Nakamura14, Proposition 1.8] proved that
Furthermore, he [Reference Nakamura14, Theorem 1.1] proved that there is a unique absolute $a_{0}\in {(0,1/2)}$ such that
In addition, it is proved in [Reference Nakamura14, Corollary 1.2] that all real zeros of $Q(s,a)$ are simple and are located only at the negative even integers just like $\zeta (s)$ if and only if ${a_0 < a \le 1/2}$ . Let us note by $Z_Q$ the set of all nontrivial zeros $\rho _{a}$ of $\xi _{Q}(s,a)$ . Since it is an entire function of order 1, one has
where $e^{A}=1/2,\ B=\frac {Q'}{Q}(0,a)-1-\frac {\gamma +\log \pi }{2}$ , and $\gamma $ denotes the Euler constant. Note that $Q'(0,a)$ is given explicitly in [Reference Nakamura14, Theorem 1.5].
1.3 Main results
Recall that $\zeta (1-s) = \Gamma _{\!\! {\mathrm {cos}}} (s) \zeta (s)$ and $Q(1-s,a) = \Gamma _{\!\! {\mathrm {cos}}} (s) Q(s,a)$ by (1.1). However, the function $Q(s,a)$ does not have an Euler product except for $a=1/6, 1/4, 1/3$ , and $1/2$ . Hence, the function $Q(s, a)$ is a suitable object to consider the influence of not Riemann’s functional equation but an Euler product to zeros of zeta functions. We show a criterion for nonvanishing of $Q(s,a)$ in terms of the positivity of the Li coefficients, an arithmetic and asymptotic formula for these coefficients in Theorems 1.1, 1.2, and 1.4, respectively. It should be emphasized that $\lambda _{n,a}$ defined in (1.3) are the first Li coefficients that we can explicitly give $n \in {\mathbb {N}}$ such that $\lambda _{n,a} <0$ . There is a possibility that this fact would give an idea to find negative Li coefficients for $\zeta (s)$ if they would exist.
For $n\neq 0$ , the Li coefficients attached to $Q(s,a)$ nonvanishing at zero are defined by the sum
The symmetry $\rho _{a}\longmapsto 1-\rho _{a}$ in the set $Z_Q$ of nontrivial zeros of $Q(s,a)$ implies that $\lambda _{-n,a}=\overline {\lambda _{n,a}}=\lambda _{n,a}$ for all $n\in {\mathbb {N}}$ . So, $\lambda _{n,a}$ are real. We have also
Moreover, from (1.2), we have (see [Reference Bombieri and Lagarias4, Equations (2.3) and (2.4)] or [Reference Smajlović17, Appendix A])
As an analogue of Li coefficients for the Riemann zeta function, we have the following.
Theorem 1.1 The function $Q(s,a)$ does not vanish when $\mathrm {Re} (s)>1/2$ if and only if $\lambda _{n,a} \geq 0$ for all $n \in {\mathbb {N}}$ .
An arithmetic formula for $\lambda _{n,a}$ is stated in the following theorems.
Theorem 1.2 We have
where $\gamma _{Q}(n)$ are defined as follows:
Theorem 1.3 For $a=1/2,1/3,1/4,1/6$ , under the RH, we have
For a fixed $l \in {\mathbb {N}}$ , we have the following asymptotic formula of $\lambda _{l,a}$ when $a \to +0$ . We can see that there exists $n \in {\mathbb {N}}$ such that $\lambda _{n,a} <0$ by Theorem 1.1 and the fact that $Q(s,a)$ does not satisfy an analogue of the RH when $a \in {\mathbb {Q}} \cap (0,1/2) \backslash \{1/6, 1/4, 1/3\}$ (see [Reference Nakamura14, Proposition 1.4]). Clearly, this argument gives no information on the frequency of $n \in {\mathbb {N}}$ , the smallest $n \in {\mathbb {N}}$ such that $\lambda _{n,a} <0$ and so on. However, the next theorem implies that $\lambda _{2n,a} <0$ if we fix any $n \in {\mathbb {N}}$ and then we take $a>0$ sufficiently small.
Theorem 1.4 Fix $l \in {\mathbb {N}}$ . Then it holds that
Especially, for any fixed $n \in {\mathbb {N}}$ , there are $a>0$ such that
2 Proofs
2.1 Proof of Theorem 1.1
Since $\lambda _{-n,a}=\overline {\lambda _{n,a}}=\lambda _{n,a}$ for all $n\in {\mathbb {N}}$ , then $\mathrm {Re}(\lambda _{-n,a})=\mathrm {Re}(\lambda _{n,a}) =\lambda _{n,a}$ . Using that $\xi _{Q}(s,a)$ is an entire function of order 1, and its zeros lie in the critical strip ${0< \mathrm {Re}(s)<1}$ , we obtain that the series $\sum _{\rho \in {Z_{Q}}}\frac {1+|\mathrm {Re}(\rho )|}{(1+|\rho |)^{2}}$ is convergent. Application of [Reference Bombieri and Lagarias4, Theorem 1] to the multiset $Z_{Q}$ of zeros of $Q(s,a)$ gives that $\mathrm {Re}(\rho )\leq 1/2$ if and only if $\lambda _{n,a}\geq 0$ for all $n\in {\mathbb {N}}$ . Now, the application of the same theorem to the multiset $1-Z_{Q}=Z_{Q}$ gives $\mathrm {Re}(\rho )\geq 1/2$ if and only if $\lambda _{n,a}\geq 0$ . This completes the proof.
Theorem 1.1 can also be proved by the same argument used in [Reference Brown5, Theorem 1], which is due to Oesterlé.
2.2 Proof of Theorem 1.2
From the expression of $\xi _{Q}(s,a)$ , one has
which is rewritten as
Note that $Q(s,a)$ is a meromorphic function on the whole complex plane, which is holomorphic everywhere except for a simple pole at $s = 1$ with residue 1 (see [Reference Nakamura13, Section 2.1]). Let us define the coefficients $\gamma _{Q}(n)$ and $\tau _{Q}(n)$ as follows:
and
By Equation (1.2), one has
From the functional equation for the function $\xi _{Q}(s,a)$ , in the neighborhood of $s=0$ , we have
Comparing Equations (2.1)–(2.4), we get
for $m\geq 0$ . Hence, the definition of $\lambda _{n,a}$ yields
where
using that $\psi (z)=-\gamma -\frac {1}{z}+\sum _{k=1}^{\infty }\frac {z}{k(k+z)}$ . Here, $\psi (s)=\frac {\Gamma '}{\Gamma }(s)$ is the logarithmic derivative of the Gamma function. Since $\psi (1/2)=-\gamma -2\log 2$ , we obtain
The equality above implies Theorem 1.2.
2.3 Proof of Theorem 1.3
Let us note that
where $\zeta (s, a)$ is the Hurwitz zeta function defined in Section 1.2. With the notation of Flajolet and Vespas [Reference Flajolet and Vepstas7, Lines 2–4, p. 70], this is $A_{n}(1,2)$ and which is equal to
where the o(1) error term above is exponentially small and oscillating and equal to
Then we have
It remains to prove that
To do so, we follow very closely the lines of the proof of the corresponding result in [Reference Lagarias8, Theorem 6.1] or [Reference Omar and Mazhouda16, Lemma 3.3] and it will be shortened. We use the following kernel function:
The residue theorem gives
where C is a contour enclosing the point $s = 0$ counterclockwise on a circle of small enough positive radius. The residue comes entirely from the singularity at $s = 0$ , as no other singularities lie inside the contour. Let $T=\sqrt {n}+\epsilon _{n}$ , for some $0<\epsilon _{n}<1$ . Now we follow very closely the lines in [Reference Omar and Mazhouda16, pp. 1106–1107] using that the function $\frac {Q'}{Q}(s,a)$ satisfies the propertiesFootnote 1
for $-2<\mathrm {Re}(s)<2$ and
for $-2\leq \mathrm {Re}(s)\leq 2$ , and we get
where
with $T=\sqrt {n}+\epsilon _{n}$ . For $a=1/2,1/3,1/4,1/6$ , under the RH, since $\left |1-\frac {1}{\rho _{a}}\right |=1$ and using formula of $N_{Q}(T)$ given in Section 1.2, we obtain $\lambda _{n,a,T}=O(T\log T+1)$ . Therefore, Equation (2.5) follows from that $\lambda _{-n,a,\sqrt {n}}=\lambda _{-n,a,\sqrt {n}}=O(\sqrt {n}\log n)$ .
Remark Since $2Q(s,a) := Z(s,a) + P(s,a)$ , from Corollary 2.3 below and [Reference Coffey6, Equation (1.18)], we obtain
where $\delta _n (a) = \frac {|\log a|^{n}}{an!} + O(1)$ and $l_{n}(a)$ are the coefficients in the expansion of ${\mathrm {Li}}_s (e^{2\pi ia})$ at $s = 1$ ; for $a\notin {\mathbb {Z}}$ , one has
2.4 Proof of Theorem 1.4
To show Theorem 1.4, we quote the following lemmas from [Reference Apostol2, Reference Berndt3].
Lemma 2.1 [Reference Berndt3, Theorem 1]
We set
Then it holds that
Lemma 2.2 [Reference Apostol2, Equation (26)]
Let $0 < a \le 1$ , and let n be a nonnegative integer. Then one has
where $\varphi (x) = \int _0^x (y- \lfloor y \rfloor - 1/2)dy$ is periodic with period $1$ and satisfies $2\varphi (x) = x(x-1)$ if $0 \le x \le 1$ .
By using the lemmas above, we immediately obtain the following.
Corollary 2.3 When $a>0$ is sufficiently small,
Proof The first formula and estimation are easily proved by Lemma 2.1 (see also [Reference Berndt3, Theorem 2]). For the first integral in Lemma 2.2, one has
from $x < x+a$ when $x,a>0$ . In addition, we have
Hence, we obtain
Therefore, we have $\epsilon _n (a) = O ( |\log a|^n )$ and the second formula in this corollary by the definition of $Z(s,a)$ and $Z(0,a) = \zeta (0,a) +\zeta (0,1-a) = 0$ (see [Reference Nakamura15, Equation (4.11)]).
Proof of Theorem 1.4
Recall the functional equation
(see [Reference Nakamura15, Lemma 4.11]). By using $\Gamma _{\!\! {\mathrm {cos}}} (s) \Gamma _{\!\! {\mathrm {cos}}} (1-s) =1$ , we have
Let $|s-1|$ be sufficiently small. Then, by $\lim _{s \to 1} (s-1) Q(s,a) = 1$ , the equation above, and the definitions of $Q(s,a)$ and $\xi _Q(s,a)$ , we have
where $\delta _n^{\prime } (a)$ and $\epsilon _n^{\prime } (a)$ are defined by
Clearly, the second estimation in Corollary 2.3 implies
Thus, we can see that $\epsilon _n^{\prime } (a) = O ( |\log a|^{n+1} )$ from $\lim _{s \to 1} (s-1) \Gamma _{\!\! {\mathrm {cos}}} (1-s) = -2$ and the fact that the function $(s-1) \Gamma _{\!\! {\mathrm {cos}}} (1-s)$ does not depend on a. Put $\eta _n(a) := \delta _n^{\prime } (a) + \epsilon _n^{\prime } (a)$ . Then, for $n \ge 0$ , we have
by Corollary 2.3. By virtue of
where $m \in {\mathbb {N}}$ and $a_m,x \in {\mathbb {C}}$ , the coefficient of $(s-1)^l$ in the function
is expressed as
Note that the function above is estimated by
from (2.6) when $a \to +0$ . We can find that
is analytic when $|s-1|<1$ form the poles of $Z (s,a)$ and $\Gamma _{\!\! {\mathrm {cos}}} (1-s)$ . So we can choose $|s-1|>0$ such that
Then, from (2.7), the Leibniz product rule, the definition of $\eta _n(a)$ , and the Taylor expansion of $\log (1+x)$ with $|x| <1$ , one has
Note that ( $\flat $ ) comes from $f^{(l)} (s,a)$ , ( $\natural $ ) is deduced by $f^{(l-1)} (s,a)$ , and ( $\sharp $ ) derives from $f^{(1)} (s,a)$ , $f^{(0)} (s,a)$ , and $O_l (1)$ in the left-hand side of the formula above. Therefore, by (2.8), we obtain
which implies Theorem 1.4.
At the end of the paper, we give numerical computation for $\lambda _{n,a}$ by Mathematica 13.0. Let
Then, we have the following.
For $n=1$ , we have
For $n=2$ , we have
Acknowledgements
The authors want to thank the anonymous referees for their many insightful comments and suggestions.