1. Introduction
A positive integer
is called prime if it has only two divisors,
q and 1. Thus,
are prime numbers. Integer numbers
that have divisors different from
k and 1 are called composite. It is well known that the set of all primes is infinite, and this was first proved by Euclid. By the fundamental theorem of arithmetic, every integer
has a unique representation as a product of prime numbers. Thus,
and
is the
jth prime number,
, with some
.
Investigations of the number of prime numbers
were more complicated. We recall that
,
,
, means that there exists a constant
such that
. Comparatively recently, in 1896 Hadamard [
1] and de la Vallée-Poussin [
2] proved independently the asymptotic formula
For this, they applied the Riemann idea [
3] of using the function
now called the Riemann zeta-function. The distribution low of prime numbers was found.
Prime numbers have generalizations. The system
of real numbers
such that
are called generalized prime numbers. Generalized prime numbers were introduced by Beurling in [
4], and are studied by many authors. The system
generates the associated system
of generalized integers consisting of finite products of the form
with some
.
The main problem in the theory of generalized primes is the asymptotic behavior of the function
The function
is closely connected to the number of generalized integers
In these definitions, the sums are taking counting multiplicities of
p and
m. Distribution results for generalized numbers were obtained by Beurling [
4], Borel [
5], Diamond [
6,
7,
8], Malvin [
9], Nyman [
10], Ryavec [
11], Hilberdink and Lapidus [
12], Stankus [
13], Zhang [
14], and others. The important place in generalized number theory is devoted to making relations between
and
. We mention some of them. From a general Landau’s theorem for prime ideals [
15], we have the estimate
that implies
Nyman proved [
10] that the estimates
and
with arbitrary
and
are equivalent. Beurling observed [
4] that the relation
is implied by (2) with
.
It is important to stress that Beurling began to use zeta-functions for investigations of the function
. These zeta-functions
, now called Beurling zeta-functions, are defined in some half-plane
, by the Euler product
or by the Dirichlet series
where
depends on the system
.
Suppose that (1) is true. Then, the partial summation shows that the series for
is absolutely convergent for
,
the function
is analytic for
, and the equality
is valid.
Analytic continuation for the function
is not an easy problem. If (1) is true, then (3) implies
This gives analytic continuation for to the half-plane , except for the point which is a simple pole with residue a.
Beurling zeta-functions are attractive analytic objects; investigations of their properties lead to interesting results, and require new methods. Various authors put much effort into showing that the Beurling zeta-functions have similar properties to classical ones. We mention a recent paper [
16] containing deep zero-distribution results for
.
In this paper, we investigate the analytic properties of the function
. The approximation of analytic functions is one of the most important chapters of function theory. It is well known that the Riemann zeta-function
is universal in the sense of approximation of analytic functions. More precisely, this means that every non-vanishing analytic function defined on the strip
can be approximated with desired accuracy by using shifts
,
. Universality of
and other zeta-functions has deep theoretical (zero-distribution, functional independence, set denseness, moment problem, ...) and practical (approximation problem, quantum mechanics) applications. On the other hand, the universality theory of zeta-functions has some interior problems (effectivization, description of a class of universal functions, Linnik–Ibragimov conjecture, see Section 1.6 of [
17], ...); therefore, investigations of universality are continued, see [
17,
18,
19,
20,
21,
22,
23].
Our purpose is to prove the universality of the function
with a certain system
. We began studying the approximation of analytic functions by shifts
in [
24]. Suppose that the estimate (1) is valid. Let
Suppose that
and define
Here, and in the sequel, the notation
,
,
, shows that there exists a constant
such that
. Denote by
the space of analytic on
D functions equipped with the topology of uniform convergence on compacta, and by
the Lebesgue measure of a measurable set
. The main result of [
24] is the following theorem.
Theorem 1. Suppose that the system satisfies the axiom (1). Then there exists a closed non-empty subset such that, for every compact set , and , Moreover, the limitexists and is positive for all but at most countably many . Theorem 1 demonstrates good approximation properties of the function ; however, the set of approximated functions is not explicitly given. The aim of this paper, using certain additional information on system , is to identify the set .
A new approach for analytic continuation of the function
involving the generalized von Mangoldt function
and
was proposed in [
12]. Let, for
and every
,
Then, in [
12], it was obtained that the function
is analytic in the half-plane
, except for a simple pole at the point
. It turns out that estimates of type (4) are useful for the characterization of the system
. It is known [
12] that (1) does not imply the estimate
with
. Therefore, together with (1), we suppose that estimate (5) is valid.
Let
be the class of compact subsets of strip
D with the connected complement, and
with
the class of continuous functions on
K that are analytic in the interior of
K. Moreover, let
Note, that the following theorem supports the Linnik–Ibragimov conjecture.
Theorem 2. Suppose that the system satisfies the axioms (1) and (5), and is linearly independent over the field of rational numbers . Let and . Then, for every , Moreover, the limitexists and is positive for all but at most countably many at . Notice that the requirement on the set
is sufficiently strong, it shows that the numbers of the system
must be different. The simplest example is the system
where
is a transcendental number.
An example of
with a bounded mean square is given in [
25].
For the proof of Theorem 2, we will build the probabilistic theory of the function in the space of analytic functions .
The paper is organized as follows. In
Section 2, we introduce a certain probability space, and define the
valued random element.
Section 3 is devoted to the ergodicity of one group of transformations. In
Section 4, we approximate the mean of the function
by an absolutely convergent Dirichlet series.
Section 5 is the most important. In this section, we prove a probabilistic limit theorem for the function
on a weakly convergent probability measure in the space
, and identify the limit measure.
Section 6 gives the explicit form for the support of the limit measure of
Section 5. In
Section 7, the universality of the function
is proved.
2. Random Element
Define the Cartesian product
The set consists of all functions . In , the operation of pointwise multiplication and product topology can be defined, and this makes a topological group. Since the unit circle is a compact set, the group is compact. Denote by , the Borel -field of the space . Then, the compactness of implies the existence of the probability Haar measure on , and we have the probability space .
Denote the elements of by . Since the Haar measure is the product of Haar measures on unit circles, is a sequence of independent complex-valued random variables uniformly distributed on the unit circle.
Extend the functions
,
, to the generalized integers
. Let
Now, for
and
, define
Lemma 1. Under the hypotheses of Theorem 2, is an -valued random element defined on the probability space .
Proof. Fix
, and consider
Then
is a sequence of complex-valued random variables on
. Denote by
the complex conjugate of
. Suppose that
,
. Since the set
is linearly independent over
, in the product
, there exists at least one factor
,
, with integer
. Therefore, denoting by
the expectation of the random variable
, we have
because the integral includes the factor
where
is the unit circle on
, and
the Haar measure on
. This and (7) show that
is a sequence of pairwise orthogonal complex-valued random variables and the series
is convergent. Hence, by the classical Rademacher theorem, see [
26], the series
converges for almost all
with respect to the measure
. Therefore, by a property of the Dirichlet series, see [
22], the series
converges uniformly on compact sets of the half-plane
for almost all
.
Now, let
and
. Denote by the set
such that the series (8) converges uniformly on compact sets of
for almost all
. Then, by the above remark,
On the other hand, taking
we obtain from (9) that
, and the series (8) converges uniformly on compact sets of the half-plane
of the strip
D. Hence,
is the
-valued random element on
. □
Lemma 2. For almost all ω, the productconverges uniformly on compact subsets of the half-plane , and the equalityholds. Proof. The series is absolutely convergent for . Therefore, the equality of the lemma, in view of (6), is valid for . By proof of Lemma 1, the function , for almost all , is analytic in the half-plane . Therefore, by analytic continuation, it suffices to show that the product of the lemma, for almost all , converges uniformly on compact subsets of the strip D.
We observe that the convergence of product (10) follows from that of the series
Hence, the series
is convergent for all
with every
,
, thus, uniformly convergent on compact subsets of the half-plane
. To prove the convergence for the series
we apply the same arguments as in the proof of Lemma 1. For fixed
, we have
and for
,
,
Thus, the series
is convergent, and the Rademacher theorem implies that the series
converges for almost all
. Hence, this series, for almost all
, converges uniformly on compact subsets of the half-plane
. This, together with a convergence property of the series (11), shows that the series
for almost all
, converges uniformly on compact subsets of the half-plane
, and it remains to prove the same for the series
Clearly, for all
,
Hence, the series (12), for all , converges uniformly on compact subsets of the half-plane . □
5. Limit Theorems
In previous sections, we gave preparatory results for the proof of a limit theorem for
in the space of analytic functions
. In this section, we consider the weak convergence for
and
as
, where
,
.
We start with a limit lemma on
. For
, define
Lemma 7. Suppose that the set is linearly independent over . Then converges weakly to the Haar measure as .
Proof. In the proof of Lemma 3, we have seen that characters of the group
are given by (14). Therefore, the Fourier transform
,
of
is defined by
For this, we apply the linear independence of the set
. We have
if and only if
. Thus, (24),
and (25) take place. □
The next lemma is devoted to the functions
and
. For
, set
and
Lemma 8. Suppose that the set is linearly independent over . Then, on there exists a probability measure such that both the measures and converge weakly to as .
Proof. We use a property of the preservation of weak convergence under continuous mappings. Consider the mapping
given by
Since the series for
is absolutely convergent for
, the mapping
is continuous. Moreover, for
,
Thus, denoting by
the measure given by the latter equality, we obtain that
. This equality continuity of
, and the principle of preservation of weak convergence, see Theorem 5.1 of [
28], show that
converges weakly to the measure
as
.
Define one more mapping
by
Then, repeating the above arguments, we find that
converges weakly to
. Let
. Then, by invariance of the measure
, we have
Thus, and converge weakly to the same measure as . □
Next, we study the family of probability measures
. We recall some notions. A family of probability measures
on
is called tight if, for every
, there exists a compact set
such that
for all
P, and
is relatively compact if every sequence
has a subsequence
weakly convergent to a certain probability measure
P on
as
. By the classical Prokhorov theorem, see Theorem 6.1 of [
28], every tight family of probability measures is relatively compact.
Lemma 9. Under the hypotheses of Theorem 2, the family is relatively compact.
Proof. In view of the above remark, it suffices to prove the tightness of
. Let
be a compact. Then, using the Cauchy integral formula and absolute convergence of the series for
, we obtain
Suppose that
is a random variable on a certain probability space
uniformly distributed in the interval
. Define the
-valued random element
Then, denoting by
the convergence in distribution by Lemma 8, we obtain
where
is the
-valued random element with the distribution
. Since the convergence in
is uniform on compact sets, (27) implies
Now, let
, where
is a sequence of compact sets of
D from the definition of the metric
. Fix
, and set
where
. Therefore, relation (26), and the Chebyshev type inequality yield
Then
is a compact set in
. Moreover, inequality (29) implies that
for all
. Since
is the distribution of
, this shows that
for all
. The lemma is proved. □
Now, we are ready to consider the weak convergence for and . For convenience, we recall one general statement.
Proposition 1. Suppose that a metric space is separable, and the -valued random elements and , are defined on the same probability space . Suppose thatand, for every , Proof. The proposition is Theorem 4.2 of [
28], where its proof is given. □
Lemma 10. Under the hypotheses of Theorem 2, on there exists a probability measure such that both the measures and converge weakly to as .
Proof. Let
be the same random variable as in the proof of Lemma 9. By Lemma 9, there exists a sequence
and the probability measure
on
such that
converges weakly to
as
. In other words, in the notation of the proof of Lemma 9,
On
, define one more
-valued random element
Then the application of Lemma 5 gives, for
,
This, and relations (27) and (30) show that all conditions of Proposition 1 are fulfilled. Thus, we have
converges weakly to
as
. Since the family
is relatively compact, relation (31), in addition, implies that
It remains to prove weak convergence for
. On
, define the
-valued random elements
and
Lemma 8 implies the relation
while, in view of Lemma 6, for
,
This, (32), (33) and Lemma 10 yield the relation
Thus, , as , also converges weakly to . □
It remains to identify the measure
. Denote by
the distribution of the random element
, i.e.,
Theorem 3. Under hypotheses of Theorem 2, converges weakly to the measure as .
Proof. We will show that the limit measure in Lemma 10 coincides with .
We apply the equivalent of weak convergence of probability measures in terms of continuity sets, see Theorem 2.1 of [
28]. Let
A be a continuity set of the measure
, i.e.,
, where
denotes the boundary of
A. Then, Lemma 10 implies that
On
, define the random variable
Return to the group
of Lemma 3. Since, by Lemma 3, the group
is ergodic, the process
is ergodic, and application of the Birkhoff–Khintchine theorem [
27] gives
for almost all
. However, the definition of the random variable
implies that, for almost all
,
This, (35) and (36) prove that for all continuity sets A of the measure . It is well known that all continuity sets constitute a determining class. Hence, we have , and the theorem is proved. □