1. Introduction
Let, as usual,
,
,
,
,
and
denote the sets of all prime, positive integers, integers, rational, real and complex numbers, respectively. The main object of the analytic number theory—Riemann zeta function
,
in the half-plane
is defined by
and has a meromorphic continuation to the whole complex plane. The point
is its simple pole with residue 1.
Let
denote the Euler gamma function. The Riemann zeta function satisfies the functional equation
Suppose that
,
is the increment of the argument of the function
along the segment connecting the points
and
. Since the function
is unbounded and monotonically increases for
(it is well known that
and
), the equation
for
has the unique solution
. Gram was the first investigator of the numbers
; therefore, they are now called Gram points. The Gram points are important in the analytic number theory because they are closely related with imaginary parts of non-trivial zeros of the function
. For more information, see [
1,
2,
3,
4].
Without other important properties, the function
has the universality property obtained by Voronin [
5]. In other words, this property means that a wide class of analytic functions uniformly on compact sets can be approximated by shifts
,
. The initial Voronin theorem [
5] states that if
, and
is a continuous and non-vanishing function on the disc
and analytic in the interior of that disc, then, for all
, there exists a number
such that
A lot of authors, among them Gonek, Bagchi, Matsumoto, Pańkowski, Steuding, Laurinčikas, Garunkštis, Macaitienė, Kačinskaitė, and others, improved and extended the above Voronin theorem. Let
,
be the class of compact subsets of the strip
D with connected complements, and let
,
, denote the class of continuous non-vanishing functions on
K that are analytic in the interior of
K and
denote the Lebesgue measure of a measurable set
. Then the modern version of the Voronin theorem, see, for example, refs. [
6,
7,
8] and informative paper [
9], says that if
,
, then, for all
,
The latter inequality means that the set of shifts
approximating a given function
has a positive lower density.
Now, we will define the Dirichlet series with periodic coefficients. Let
be the periodic sequence of complex numbers with minimal period
. The Dirichlet series with periodic coefficients
, for
is defined by the series
Since the sequence
is periodic, we have
where
is the classical Hurwitz zeta function with parameter
, which has, as
, a meromorphic continuation to the whole complex plane with a unique simple pole at the point
with residue 1. Hence, the function
can be analytically continued to the whole complex plane, except for a simple pole at the point
with residue
If
,
is an entire function. If the sequence
is multiplicative (
for
, and
), then, for
, the function
has the Euler product
Universality for the function
, i.e., approximation of a wide class of analytic functions by shifts
,
, was investigated by various authors, among them, Bagchi [
10], Steuding [
8,
11], and others. The universality of the function
with a multiplicative sequence
was obtained in [
12].
Theorem 1 ([
12])
. Suppose that the sequence is multiplicative, and, for all , Let , and . Then, for all , Note that the requirement (
3) is technical and can be removed.
The universality of the Dirichlet series with periodic coefficients is a complicated problem. Kaczorowski in [
13] observed that not all Dirichlet series with periodic coefficients are universal in the Voronin sense. He obtained the necessary and sufficient conditions of the universality for
with prime period
l.
Theorem 2 ([
13])
. Let l be a prime number and let . The corresponding Dirichlet series with periodic coefficients are universal in the sense of Voronin if and only if one of the following possibilities holds:1. Not all numbers are equal;
2. We have ;
3. We have and The discrete universality for zeta functions was proposed by Reich [
14]. The first result on the approximation of analytic functions by discrete shifts
, with a fixed number
such that
is rational for all
, has been obtained in [
15] and a more general result in [
16]. Theorem 2 of [
17] with
implies the following result. Let
denote the number of elements of the set
A, and, for
,
Theorem 3 ([
17])
. Suppose that the set is multiplicative and the set is linearly independent over the field of rational numbers . Let and . Then, for all , The universality of zeta functions is a very surprising and useful phenomenon which, in some sense, reduces a study of a class of analytic functions to that of a comparatively simple one and the same zeta function. Moreover, universality theorems are applied in a lot of number-theoretical problems as the functional independence, zero distribution, denseness, and the moment problems, etc. This is the motivation to study and extend the notion of universality for zeta functions. One of the ways in this direction is to prove the universality theorems for new classes of zeta functions. The Linnik–Ibragimov conjecture (or programme), see [
8], Section 1.6, asserts that all functions given by the Dirichlet series, having analytic continuation and satisfying some natural growth conditions, are universal in the Voronin sense. On the other hand, there are examples of non-universal Dirichlet series; Theorem 2 confirms this, and there are also Dirichlet series in which universality is an open problem.
Using the generalized shifts, there is an another way to extend the universality for the Dirichlet series. This idea, for Dirichlet
L-functions, was proposed by Pańkowski in [
18] with function
with certain
. In [
19], the universality for the Dirichlet series with periodic coefficients with multiplicative sequence
using generalized shifts
, where
is a real number, and
is increasing to an
∞ continuously differentiable function with monotonic derivative
on
,
, such that
was obtained. Now, let
be a sequence of imaginary parts of non-trivial zeros of the Riemann zeta function and let the hypothesis
be satisfied. This estimate is the weak form of the Montgomery pair correlation conjecture [
20]. Then, in [
21], it was obtained that, under (
4), for a fixed number
, multiplicative sequence
,
,
and all
,
In [
22], Korolev and Laurinčikas in approximating shifts of the Riemann zeta function involved the Gram points, and in [
23], developed their result in short intervals. We notice that Gram points
are asymptotically connected to the numbers
by the equality [
22]
Moreover, using the points
for generalized shifts is more convenient than
because
has a continuous differentiable version
,
, see Lemma 4 bellow, and the analogue of (
4) is not needed. The aim of this paper is to generalize the latter result for the Dirichlet series with periodic coefficients. The main result of the paper is the following theorem.
Theorem 4. Suppose that the sequence is multiplicative. Let , , and be a fixed number. Then, for all ,Moreover, “lim inf” can be replaced by “lim” for all but at most, countably many . Theorem 4 implies that the set of shifts approximating a given function with accuracy is infinite for every .
The class of functions
is sufficiently wide; for example, it includes all Dirichlet
L-functions which are the main analytic tool for the investigation of prime numbers in arithmetic progressions. When
, we obtain the Riemann zeta function
. Therefore, Theorem 4 extends and covers the main result of [
22].
Theorem 4 is theoretical, it is not connected to specific numerical calculations, and can be considered as an impact to the Linnik–Ibragimov programme.
Theorem 4 is stated in density terms (lower density and density), the expression with respect to is a probabilistic distribution function. Therefore, for its proof, it is convenient to use a probabilistic approach. More precisely, for the proof of Theorem 4, we apply a limit theorem on weakly convergent probability measures in the space of analytic functions with an explicitly given probability limit measure.
2. Limit Theorems
The space of analytic functions on
D endowed with the topology of uniform convergence on compacta is denoted by
, and the Borel
-field of a topological space
is denoted by
. In this section, we will prove a theorem on the weak convergence of the measure
as
.
Let
. Define the set
where
for all
. With the product topology and operation of pointwise multiplication, the torus
, in view of the classical Tikhonov theorem, see, for example, ref. [
7], is a compact topological Abelian group. Therefore, on
, we can define the probability Haar measure
. Thus, we can construct the probability space
. Denote by
the
pth component of an element
,
. Now, on the probability space
, define the
-valued random element
and let
denote the distribution of
, i.e.,
We can now formulate the main theorem of this section.
Theorem 5. The measure converges weakly to as .
We divide the proof of Theorem 5 into separate lemmas. At first, for
, define
Lemma 1. The measure converges weakly to the Haar measure as .
Proof. The character
of the group
, see, for example, ref. [
7], has the representation
where only a finite number of integers
are distinct from zero and does not depend on the sequence
. Therefore, the proof of the lemma applies the Fourier transform method and coincides with the proof of Lemma 3.2 from [
22]. □
From Lemma 1, we can obtain a limit lemma in the space
for the absolutely convergent Dirichlet series. Let, for fixed
,
and
Since
with all
and
, the latter series is absolutely convergent for all
. Moreover, let
where
Then, the latter series is also absolutely convergent for all
because
. For
, define
Let
be defined by the formula
For the proof of the limit lemma for the absolutely convergent Dirichlet series, we need a lemma on the preservation of probability measures under continuous mappings. Let
P be a probability measure on
and
be a measurable mapping. Then,
is defined, for
, by
.
Lemma 2. Let P and , , be probability measures on , and be a continuous mapping. The measure converges weakly to as if the measure converges weakly to P as .
The lemma is a partial case of the Theorem 2.7 from [
24].
Lemma 3. The measure , as , converges weakly to a measure .
Proof. Seeing that the series for is absolutely convergent for , then the function is continuous. From the definitions of the measures , and mapping , we have . Therefore, from Lemmas 1 and 2, we have the assertion of the lemma. □
The next step is the approximation of by in the mean. For the proof of this fact, we need the following lemmas. The first of them is devoted to asymptotics of the function with arbitrary .
Lemma 4 ([
3])
. Suppose that , , denotes the unique solution of the equation satisfying and that . Then and Also, we recall the Gallagher lemma. This lemma connects discrete and continuous second moments of some functions. For details, see, for example, Lemma 1.4 of [
25].
Lemma 5 ([
25])
. Let , , , be a non-empty finite set in the interval , and Moreover, let the function be continuous on , which takes complex values, and have a continuous derivative on . Then On
D, there exists a sequence
of compact subsets such that
,
, and
for some
q if
is a compact set. Then, the formula
gives a metric in
. This metric induces its topology of uniform convergence on compacta. To move from the function
to
, we need the following lemma.
Lemma 6. For fixed , the following statementholds. Proof. First, we obtain some discrete second moment estimates for the function
. For
, the bounds
and
are well known. Hence, for
and
, we have
and
Actually, in view of Lemma 4, the function
is increasing. Then, by the same lemma, for
and
,
Now, taking
and summing over
, we obtain (
5). Similarly, we obtain estimate (
6).
Now, we will obtain the estimate for the discrete mean value of
involving Gram points. For this, we apply Lemma 5. Let
,
,
, and
. In this case,
. In view of (
5) and (
6), for
, we find
Let
be from the definition of
, and, for
,
Then, for
, by the Mellin formula, the function
has the expression by the contour integral
Let
be a fixed compact set and define
such that
for any point
. Then, for
, we obtain
where
Hence, we obtain the inequality
Then, shifting
to
, we have
Summing over
, we obtain
where
and
Using the well-known Stirling formula, uniformly for
, we obtain the estimate
It implies the bound
Define
. Then,
, and hence,
As above, we obtain
To obtain the estimate for the sum
, we separate it into two parts
and
: over
and over
, respectively. It is easily seen that
and
Hence,
Further, using the Cauchy inequality and (
7), we have
and
Using the above estimates for
and
, we have
Tending
, and then
, we obtain the assertion of the lemma using the definition of the metric
. □
Now, we are ready to prove Theorem 5. For the proof, we will apply the following assertion, see, for example, [
24], Theorem 3.2.
Lemma 7 ([
24])
. Suppose that is the separable space and the -valued random elements and are defined on the same probability space with measure . Let, for all l, and If for each , then Proof of Theorem 5. Let
be a certain probability space. Suppose that
is a random variable on the above probability space such that
Using
, define two
-valued random elements
and
Let
be the
-valued random element of which the distribution is
; here,
is the limit measure from Lemma 3. Then, by Lemma 3,
Using a standard method, see, for example, [
7], we obtain that the family
of probability measures is tight, i.e., for each
, there exists a compact subset
of the set
D such that
for all
. From this and the classical Prokhorov theorem, see, for example, Theorem 5.1 of [
24], we determine that
is relatively compact, i.e., each sequence of
contains a subsequence
which converges weakly to a certain probability measure
P on
as
. Thus,
Now, using Lemma 6, for all
, we have
This relation, (
9), (
10), and Lemma 7 prove that
This means that, for
,
converges weakly to
P. Additionally, (
11) shows that the measure
P in (
10) does not depend on the subsequence
. Therefore, we have that
converges weakly to
P as
.
Moreover, we need to identify the measure
P. In [
12], it was received that
as
, also converges weakly to the limit measure
P of
, and
P coincides with
. Thus,
converges weakly to
as
as well. □