Brought to you by:
Paper

On small values of the Riemann zeta-function at Gram points

© 2014 Russian Academy of Sciences (DoM), London Mathematical Society, Turpion Ltd.
, , Citation M. A. Korolev 2014 Sb. Math. 205 63 DOI 10.1070/SM2014v205n01ABEH004367

1064-5616/205/1/63

Abstract

In this paper, we prove the existence of a large set of Gram points $t_{n}$ such that the values $\zeta(0.5+it_{n})$ are 'anomalously' close to zero. A lower bound for the negative 'discrete' moment of the Riemann zeta-function on the critical line is also given.

Bibliography: 13 titles.

Export citation and abstract BibTeX RIS

§ 1. Introduction

The questions considered in the present paper belong essentially to the discrete theory of the Riemann zeta-function $\zeta(s)$. This is the name of a whole class of problems where the behaviour of $\zeta(s)$ and related functions on different discrete sets of points $\{s_{n}\}$ is studied.

Let $\vartheta(t)$ be the increment of a continuous branch of the argument of the function $\pi^{-s/2}\Gamma\bigl(s/2\bigr)$ along the line segment joining the points $s = 0.5$ and $s = 0.5+it$ $(t>0)$. Then Hardy's $Z$-function, defined by the formula $Z(t) = e^{i\vartheta(t)}\zeta(0.5+it)$, is real for real $t$, and its real zeros coincides with the ordinates of zeros of $\zeta(s)$ lying on the critical line $\Re s = 0.5$ (see, for example, [1], Ch. III, § 4).

Let $\varrho_{n} = \beta_{n} + i\gamma_{n}, n = 1,2,\dots$, be complex zeros of $\zeta(s)$ lying in the upper half-plane and indexed in ascending order of their ordinates (and arbitrarily in the case of the coincidence of the latter): $0< \gamma_{1}< \gamma_{2}< \dots\leqslant \gamma_{n}\leqslant\gamma_{n+1}\leqslant \dotsb$. Further, for any $n\geqslant 0$, we define $t_{n}$ as the unique solution of the equation

that satisfies the condition $\vartheta'(t_{n}) > 0$. The values $t_{n}$ are called Gram points. Some characteristics of the sequences $\{\gamma_{n}\}$ and $\{t_{n}\}$ appear to be very close to one another, which is the reason for the interest of researchers in Gram points. In particular, it is known that the asymptotic relations

hold when $n$ increases unboundedly, and that the number of ordinates $\gamma_{n}$ lying on the sufficiently long segment $(0,T]$, and expressed by the Riemann-von Mangoldt formula as

coincides with the number of Gram points $t_{n}$ lying on the same segment to within the term $O(\ln T)$ (see [2], Ch. X, § 6 and [3], § 7).

The equations

and the definition of $t_{n}$ imply that the values $\zeta(0.5+it_{n})$ are real and coincide with the $x$-coordinates of the points of intersection of the curve

and the real axis. The zeros of Hardy's function $Z(t)$ also correspond to the intersections of $\mathscr{C}$ and the real axis, but such points coincide with the origin.

It is natural to ask how these points of intersection are distributed. For example, with increasing $n$, can they run arbitrarily far to the left or to the right of the origin, or, conversely, approach it arbitrarily closely?

Figure 1.

Figure 1. A piece of the curve $\mathscr{C}$ that corresponds to $0\leqslant t\leqslant 40$.

Standard image

A positive answer to the first question was given in the papers [4] and [5]. It appears that in the case of sufficiently large $N$ there exist indices $m$ and $n$ such that $1\leqslant m,n\leqslant N$ and

for some absolute constant $c > 0$. Moreover, for any $0 < \varepsilon < 0.5$ and $N\geqslant N_{0}(\varepsilon)$, there exists a number $k$ lying on the same segment such that

The aim of this paper is to answer the second question.

On the one hand, the coincidence of a Gram point $t_{n}$ with any ordinate $\gamma_{m}$ of a zero of $\zeta(s)$ seems absolutely improbable. This allows one to suppose that

Equation (1.1)

for any 1 $n\geqslant 0$. At the same time, the numerical data provides examples of Gram points with values $\zeta(0.5+it_{n})$ that are 'anomalously' close to zero. The table below contains all the indices $n$ that do not exceed $10^{8}$ and are such that these values do not exceed $10^{-7}$ in modulus.

no.$n$$\zeta(0.5+it_{n})$no.$n$$\zeta(0.5+it_{n})$
1368 383$8.908459\cdot 10^{-8}$755 785 549$3.751899\cdot 10^{-8}$
212 984 109${-}2.052298\cdot 10^{-8}$861 769 885$8.026449\cdot 10^{-8}$
321 567 185${-}6.709404\cdot 10^{-8}$965 463 721$1.083783\cdot 10^{-8}$
445 898 152$8.302024\cdot 10^{-8}$1069 612 841$9.033652\cdot 10^{-8}$
550 550 325${-}6.397503\cdot 10^{-8}$1174 201 244$2.875945\cdot 10^{-8}$
652 220 649${-}6.529425\cdot 10^{-8}$

The same segment contains 60 values of $n$ with the condition $|\zeta(0.5+it_{n})| < 10^{-6}$.

In what follows, the theoretical explanation of this phenomenon is given. Namely, the following assertion holds.

Theorem 1. Suppose that $0< \varepsilon < 10^{-3}$ is fixed, $N\geqslant N_{0}(\varepsilon)>0,$ $M=[N^{\alpha+\varepsilon}]$, and $\alpha=27/82$. Further, let $\varphi(x)$ be any unbounded and monotonically increasing function that satisfies the conditions

where $a$ is a sufficiently large positive constant. Then the inequality

holds with $\varphi=\varphi(N)$ for at least

Gram points $t_{n},$ $N < n\leqslant N+M$.

The lower bound for the negative 'discrete' moment of $\zeta(s)$ is an immediate consequence of the above assertion.

Theorem 2. Suppose that $\varepsilon$ and $\delta$ are any fixed numbers that satisfy the conditions $0 < \varepsilon < 10^{-3}$ and $\delta>0$. Then the estimate

holds for any $N\geqslant N_{0}(\varepsilon,\delta)$ and $M=[N^{\alpha+\varepsilon}],$ $\alpha=27/82$.

In proving these theorems, we essentially use the ideas and methods of the classical paper [7] of Selberg and the recent paper [8] of Radziwiłł. In particular, this relates to the notation of the number of solutions of the equation (3.10) proposed in [8] by means of the multiplicative function $f(n)$ (see Lemma 7). This also relates to the trick of excluding from consideration the 'bad' set where the function under study is 'anomalously' large (the set $\mathscr{A}_{3}$ in Lemma 7).

In what follows, we use the following notations: $|\mathscr{A}|$ denotes the number of elements of a finite set $\mathscr{A}$; $G(x)$ denotes the Gaussian distribution, that is,

is the Bessel function of the first kind; the symbols $\sum_{n}$ and $\sum_{n\in \mathscr{A}}$ stand for the sum over all $n$ from the interval $(N,N+M]$ and for the sum over all $n$ under the condition $n\in \mathscr{A}$ lying in the same interval, correspondingly; $\mathbf{E}(\xi)$ denotes the mean value of the random variable $\xi$ and $\mathbf{P}\{\xi\leqslant x\}$ means the probability of the event 'the random variable $\xi$ does not exceed $x$'; finally, $\theta, \theta_{1}, \theta_{2},\dots$ are complex numbers (generally speaking, different in different formulae) whose absolute value does not exceed 1.

§ 2. Auxiliary assertions

This section contains auxiliary lemmas of a 'technical' nature.

Lemma 1. Let $\{\xi_{N}\}$ be a sequence of random variables. Suppose that there exist some fixed $r>0$ and some positive parameters $\sigma(N)$ and $\kappa(N)$ that increase unboundedly as $N\to+\infty$ and such that the generating function $K_{N}(s) =\mathbf{E}(e^{s\xi_{N}})$ for the moments of the random variable $\xi_{N}$ satisfies the relation

in the disc $|s|\leqslant r$, where the implied constant in the $O$-symbol is absolute and the function $\Phi(s),$ $\Phi(0) = 1$, is analytic in the disc $|s|\leqslant r$ and independent of $N$. Then the distribution

of the normalized random variable $\xi_{N}$ satisfies the following relations for any $u=o(\min{(\kappa(N),\sqrt{\sigma(N)})}),$ $u>1$:

where

and the constants in $O$ are absolute.

This is a particular case of Hwang's theorem in [9], where $u(s)=s^{2},$ $v(s)=\ln\Phi(s),$ $\phi(N)=\sigma(N)/4$.

Lemma 2. Let $k, \mu, \nu$ be integers, $\mu,\nu\geqslant 0,$ $k\geqslant 1,$ $\mu+\nu =k,$ $y_{0}\leqslant y\leqslant N^{1/(2k)}$. Further, suppose that $p_{1},\dots, p_{\mu}, q_{1},\dots, q_{\nu}$ run through the prime numbers from the interval $(1,y]$ that satisfy the condition $p_{1}\dotsb p_{\mu}\ne q_{1}\dotsb q_{\nu}$. Finally, suppose that all the terms of a sequence $a(p)$ for $p\leqslant y$ satisfy the condition $|a(p)|\leqslant \delta$. Then the sum

satisfies the bound $|S|< (\delta x)^{k}\ln{N}$.

The proof repeats almost verbatim that of Lemma 6 in [10].

§ 3. Basic lemmas

First we introduce some additional notation. Suppose that $2\leqslant x\leqslant t^{2}$. Following Selberg (see [7]), we define

where the maximum is taken over all the zeros $\varrho = \beta + i\gamma$ of the Riemann zeta-function that satisfy the condition

Further, let $n\geqslant 2$ be an integer. We set

The proof, going back to Selberg's theorem, stating that the function

has a distribution tending to the Gaussian as $T\to +\infty$, is based on the following 'explicit' formula for the real part of the logarithm of the Riemann zeta-function:

Equation (3.1)

here $\eta(t)=\min_{\varrho=\beta+i\gamma}|t-\gamma|$ is the distance between $t$ and the closest ordinate of a zero of $\zeta(s)$, so $\chi(\gamma_{m})=+\infty$ (see [11], Theorem 5.2).

However, this formula is unsuitable for handling the random variable with the values

The reason is that the proof of Selberg's theorem is based on estimates for the integrals of the positive powers of the $O$-term in the right-hand side of (3.1) and, in particular, on estimates for integrals of the following type:

Equation (3.2)

All these integrals are finite because, for example, the integrals

are finite. In the discrete case, one should handle the sums

Equation (3.3)

instead of the integrals (3.2). The absence of precise information about the relative location of Gram points and ordinates of zeros of $\zeta(s)$ does not enable us even to assert the finiteness of the quantities (3.3).

The next lemma partially eliminates this inconvenience and enables us to estimate the logarithm of $|\zeta(0.5+it)|$ above by some function $C(t)$ for any $t$. Roughly speaking, this function can be obtained by omitting the factor $\ln{\chi(t)}$ from the right-hand side of (3.1).

Lemma 3. Suppose that $10\leqslant x \leqslant t^{\delta},$ $800\delta=1$. Then the inequality

holds with

$a_{1}=4.01,$ $a_{2}=7.52$.

Proof. If $0.5+it$ is a zero of $\zeta(s)$ then the inequality of the lemma holds. So in what follows we assume that $\zeta(0.5+it)\ne 0$. In this case, we have

where

To estimate $I_{1}$ we use the inequality

Equation (3.4)

proved in [7] (the calculation of the constants is contained in Lemma 7 of the survey [12]). Thus we get

Equation (3.5)

Using the estimate $0.5(\sigma_{x,t}-0.5)\ln{x}\geqslant 1$ we rewrite (3.5) as follows:

where $a_{1}^{(1)}=0.4,$ $a_{2}^{(1)}=1.3,$ $a_{3}^{(1)}=2.41$.

To calculate $I_{2}$ we apply the formula

where $s=\sigma+it,$ $\sigma\geqslant\sigma_{x,t}$ (see [7] and [12], Lemma 3)2. Thus we find

where $a_{1}^{(2)}=13/(12e),$ $a_{2}^{(2)}=13/(6e),$ $a_{3}^{(2)}=26/(3e)$.

Finally, using the equality

(see [13], §§ 76, 77 and also [12], Lemma 2), for $\zeta(\sigma+it)\ne 0$ we obtain

where

We also put

Then

Since the estimates

are valid on the whole segment $0.5\leqslant \sigma\leqslant \sigma_{x,t}$, for any zero $\varrho = \beta+i\gamma$ we have

To estimate $\kappa_{2}$ and $j_{2}$, we note first that the numerator of $\kappa_{2}$ is nonpositive on the whole segment of integration in both the cases $\beta\leqslant 0.5$ and $\beta\geqslant \sigma_{x,t}$. Hence, for all such zeros, we have $j_{2}\leqslant 0$. So, in what follows we assume that $0.5< \beta< \sigma_{x,t}$. Let us put $\tau = |t-\gamma|$ and consider two cases.

  • Case 1:  
    $\tau > (\ln{x})^{-1}$. Transforming $\kappa_{2}$, we find
    Further, noting that the last fraction is negative for $\sigma > \beta$ we get
    Equation (3.6)
  • Case 2:  
    $0 \leqslant \tau\leqslant (\ln{x})^{-1}$. First let us assume that $\sigma_{x,t}=0.5+2(\beta_{0}-0.5)$, where $\beta_{0}>0.5$ is the abscissa of some zero $\varrho_{0}=\beta_{0}+i\gamma_{0}$ that satisfies the condition

Then the definition of $\sigma_{x,t}$ implies that $\beta_{0}$ is the largest among the abscissas of the zeros $\varrho^{*} =\beta^{*}+i\gamma^{*}$ satisfying the condition

Equation (3.7)

But the estimate

holds for any $\beta^{*}$. Since the zero $\varrho = \beta + i\gamma$ under consideration satisfies the inequality $\tau =|t-\gamma|\leqslant (\ln{x})^{-1}$, this zero $\varrho$ is certainly contained in the set of zeros $\varrho^{*}$ satisfying (3.7). Hence its abscissa $\beta$ cannot exceed $\beta_{0}$:

The last inequality implies that

Equation (3.8)

Now let us assume that $\sigma_{x,t}=0.5+2(\ln{x})^{-1}$. Then for all the zeros under condition (3.7) (and, in particular, for the zero $\varrho = \beta + i\gamma$) we have

Thus, in this case the abscissa $\beta$ satisfies (3.8), too. Passing to the estimate of $j_{2}(\varrho)$, we find

where

Setting $u=\beta-\sigma$ and $v=\sigma-\beta$ in $j_{3}$ and $j_{4}$ respectively, we obtain

and hence

By (3.8), the fraction under the logarithm sign does not exceed 1. Therefore,

Equation (3.9)

Comparing (3.6) and (3.9) we conclude that inequality (3.6) holds for any $\beta$. Hence,

Now let us use the inequality

from [7] (see also [12], Lemma 6), together with the estimate (3.4). Thus we get

where $a_{1}^{(3)}=77/24,$ $a_{2}^{(3)}=13/4,$ $a_{3}^{(3)}=19.3$. Therefore,

where $a_{1}^{(4)} < 4.00687,$ $a_{2}^{(4)} < 7.51374,$ $a_{3}^{(4)} < 24.89829$. Further, setting

and noting that

for $10\leqslant x\leqslant t^{\delta},$ $\delta=800^{-1}$, we obtain

The lemma is proved.

For what follows, we set $X=N^{0.1\varepsilon}$ and $L=\ln\ln N$.

Lemma 4. Let $k$ be an integer, $1\leqslant k\leqslant c\ln{X}$, where $c>0$ is a sufficiently small absolute constant, and $x=X^{1/(8k+3)}$. Then the following inequality holds:

where $c_{0}=a_{0}\varepsilon^{-1},$ $a_{0}=5132.8$.

The proof of this assertion is similar to the proof of Theorem 4 in [7] and to the proof of Lemma 2 in [10].

Let $\varphi(x)$ be an unbounded and monotonically increasing function for $x\to +\infty$ and suppose that the inequalities $1\leqslant \varphi(x)\leqslant c_{0}L(3\ln{L})^{-1}$ hold for $N\leqslant x\leqslant N+M$, where the constant $c_{0}$ is defined in Lemma 4. We set

Lemma 5. The inequality

holds for any $n$ under the condition $N < n\leqslant N+M$, except for a set of at most $M\Delta_{1}$ indices, where

Proof. We put $k_{1}=[(3ec_{0}\varphi)^{-1}L],$ $x=X^{1/(8k_{1}+3)}$. Let $\mathscr{A}_{1}$ be the set of $n, N< n\leqslant N+M$, such that $|C(t_{n})-V_{x}(t_{n})| > (3\varphi)^{-1}L$. Then by Lemma 4 we find

and hence

Further, it is easy to note that

so $x > y^{10^{5}}$. Denoting the set of $n$ under the condition $|V_{x}(t_{n})-V_{y}(t_{n})| > L (3\varphi)^{-1}$ by $\mathscr{A}_{2}$ and applying Lemma 2, we find

and therefore

Thus, excluding from consideration at most

Gram points $t_{n}$, for the remaining points we get

The lemma is proved.

Lemma 6. The inequality

holds for any $n$ under the condition $N< n\leqslant N+M$, except for a set of at most $M\Delta_{2}$ indices, where $\Delta_{2}=9e\sqrt{\mathfrak{S}}e^{-9\mathfrak{S}}$.

Proof. Let $\mathscr{A}_{3}$ be the set of $n$ satisfying the condition $|V_{y}(t_{n})|> 3 \mathfrak{S}$. Setting $k_{2}=[9\mathfrak{S}]$ we therefore get $y^{2k_{2}}\ln{N}< M^{1/3}$. Using Lemma 4, we obtain the inequalities

which lead us to the desired bound.

Denote by $\mathscr{A}$ the complement of $\mathscr{A}_{3}$ in the set of all integers in the segment $N< n\leqslant N+M$.

Lemma 7. Let $K(s)=\sum_{n\in\mathscr{A}}\exp{(sV_{y}(t_{n}))}$. Then the formula

holds with $F_{y}(s)=\prod_{p\leqslant y}J_{0}(is/\sqrt{p})$ for any complex $s$ satisfying the condition $|s|\leqslant 1$.

Proof. We put $K=[3e^{2}\mathfrak{S}]+1$. Then

and, moreover,

Further, using the notation of Lemma 6 we represent $K_{1}(s)$ as follows:

and note that

Since $y^{2K}\ln{N}< M^{2/3}$, the Cauchy inequality and Lemma 6 imply

Hence

To transform $K_{2}(s)$, we set

If $k\leqslant K$ is odd, then Lemma 2 implies

The contribution to the sum $v_{k}$ of the terms satisfying the condition $p_{1}\dotsb p_{\nu} = q_{1}\dotsb q_{k-\nu}$ for $k = 2m,$ $\nu = m$ is estimated in the same way. Therefore, in this case

where

Thus,

Noting that

we find

Now let $\varpi_{1}^{\alpha_{1}}\dotsb\varpi_{r}^{\alpha_{r}}$ be the canonical factorization of $n$ satisfying the condition $\Omega(n)= \alpha_{1}+\dots+\alpha_{r}=m$. Then the number of solutions of the equation

Equation (3.10)

in primes $p_{1},\dots, q_{m}$ not exceeding $y$ is equal to

or 0 according to whether or not all the prime divisors $\varpi_{1},\dots, \varpi_{r}$ of $n$ are less than or equal to $y$. Defining the multiplicative function $f(n)$ for the powers of primes by the relations

we conclude that the number of solutions of (3.10) is represented as $(m!)^{2}f(n)$. Hence,

Summing the bounds obtained above for the quantities $R_{j}(s)$, we find

It remains to estimate the quantity $\min_{|s|\leqslant 1}|F_{y}(s)|$ below. We represent $F_{y}(s)$ in the form

where

Expanding $\psi(v)$ in a series we find

In particular, $q_{0}=1,$ $q_{1}=0,$ $q_{2}=-1/4$, and $q_{3}=1/9$. These identities, together with the obvious estimate

Equation (3.11)

imply that $|q_{m}|\leqslant 1$ for any $m\geqslant 1$. Therefore,

where $\Phi_{0}=1$ and

for $n\geqslant 1$. Since $q_{1}=0$, a nontrivial contribution to $\Phi_{n}$ for $n\geqslant 2$ comes only from the tuples $(m_{1},\dots, m_{r})$ with $m_{1},\dots, m_{r}\geqslant 2$ (which is possible only in the case when $1\leqslant r\leqslant n/2$); moreover, the number of such tuples is equal to

for given $n$ and $r$. Therefore, for any $n\geqslant 2$, we have

It is easy to check that the last sum coincides with the coefficient of $x^{[n/2]-1}$ in the expansion of the fraction

in a power series; here $j$ denotes the remainder occurring when $n$ is divided by 2. Expanding this expression into partial fractions, we get the estimate $|\Phi_{n}|< \phi^{n}$ where $\phi = (1+\sqrt{5})/2$. Thus, for $|s|\leqslant 1$ we have

and hence

Finally, noting that $62\mathfrak{S}e^{-3.75\mathfrak{S}}< (\ln{N})^{-3}$, we obtain the result of the lemma.

§ 4. The proofs of Theorems 1 and 2

Let $K_{N}(s)$ be the generating function for the moments of the discrete random quantity $\xi_{N}$ with the values $V_{y}(t_{n}),$ $N< n\leqslant N+M,$ $n\in \mathcal{A}$. Then, using the notations and the results of the Lemmas 6 and 7, we easily conclude that the relations

hold in the disc $|s|\leqslant 1$. Denoting the products

by $\Phi(s)$ and $\Phi_{y}(s)$, respectively, we obtain

Suppose that $p > y$. Using inequality (3.11), we get

and hence

Taking

in Lemma 1, we conclude that the inequality $V(t_{n})\leqslant -u\sqrt{\mathfrak{S}/2}$ holds for at least

Gram points. Using Lemma 5 and excluding from the above quantity at most

points $t_{n}$, in the case $\varphi\geqslant 1.3\cdot 10^{5}\varepsilon^{-1}>9ec_{0}$, for at least

remaining points we have the inequalities

Theorem 1 is proved.

Now taking $\varphi(x)=0.5(\ln\ln\ln{x})^{\delta}$, we get $\varphi=0.5(\ln{L})^{\delta}$ and hence

Theorem 2 is proved.

Footnotes

  • It is necessary to note that Selberg's paper [6] contains an assertion that implies that inequality (1.1) holds for a positive proportion of indices $n$.

  • There is a misprint in the statement of Lemma 3 in [12]: it should be $1/4-\sigma/2$ in place of $1/2-\sigma/2$.

Please wait… references are loading.
10.1070/SM2014v205n01ABEH004367