1. Introduction
One of the most important branches of the function theory is the approximation of analytic functions, and is widely used not only in mathematics but also in other natural sciences. In the 1980s, it was discovered that there exist analytic objects that approximate large classes of analytic functions. S.M. Voronin found [
1] that the first such object as the Riemann zeta-function
,
, given by
where
is the set of all prime numbers. As is well-known,
has the meromorphic continuation of the whole complex plane with
. Voronin proved [
1] (see also [
2]) that if
, the function
is continuous and non-vanishing on the disc
, and analytic in the interior of that disc, then there exists a real number
such that
for any
.
Thus, Voronin reported that all non-vanishing analytic functions on the strip
, and uniformly on discs can be approximated by shifts
of one and the same function
. The Bohr–Courant theorem [
3] claims that the set
is dense everywhere on a complex plane for every fixed
. From here, it follows that the set of values of the function
is very rich. Thus, in terms of approximation, the function
is universal, and this might be natural in view of the remark above.
We denote by the space of the analytic on D functions equipped with the topology of uniform convergence on the compacta. Since the space has an infinite-dimension, the Voronin theorem is a infinite-dimensional extension of the Bohr–Courant denseness theorem.
The above-mentioned Voronin universality theorem has a more general statement which follows the Mergelyan theorem on the approximation of analytic functions by polynomials [
4]. We denote by
the set of compact subsets of the strip
D with connected complements, and by
the class of continuous non-vanishing functions on
that are analytic in the interior of
K. Moreover, we let
stand for the Lebesgue measure of a measurable set
. Then the following statement on the
’s universality is known, see, for example, [
5,
6,
7,
8,
9].
Theorem 1. Suppose that and . Then, for every , The inequality of the theorem shows the infinitude of shifts of approximating a given function .
The statement of Theorem 1 was influenced by a probabilistic method proposed in [
6]. The initial Voronin method based on the Riemann-type rearrangement theorem in the Hilbert space was developed in [
7,
8].
Since
in the shifts
of Theorem 1 is an arbitrary real number, Theorem 1 is called a continuous universality theorem. Parallel to continuous universality theorems for zeta-functions, there are discrete universality theorems when
takes values from a certain discrete set. These were proposed by A. Reich [
10] for Dedekind zeta-functions of algebraic number fields
. If
, we deal with a discrete universality for the Riemann zeta-function. As an example, we now state a classical result in the following (see [
6]).
Theorem 2. Suppose that , and . Then, for every ,
Here
denotes the number of elements of the set
, and
N runs over the set
.
Note that discrete universality theorems were also investigated in [
6,
7,
8].
Some other functions given by a Dirichlet series also fulfil the property of universality in the Voronin sense. For example, Dirichlet
L-functions
with arbitrary Dirichlet character
,
are universal, as was mentioned by Voronin in [
2]. Let
be a periodic sequence. Then the periodic zeta-function
also has the universal approximation property [
11]. For values of the parameters
and
, the Hurwitz zeta-function
and Lerch zeta-function
, for
, respectively given by
are universal (see [
12]). In other words, they approximate analytic functions from the class
considered continuous on
K and analytic in the interior of
K functions. This observation leads to certain conjectures. For example, by the Linnik–Ibragimov conjecture (or programme), see [
8], all functions in a certain half-plane defined by a Dirichlet series, with analytic continuation left of the absolute convergence abscissa and satisfying some natural growth hypotheses are universal in the Voronin sense. However, currently there are Dirichlet series which their universality is not known, for example, the function
with an algebraic irrational parameter. Results in this direction for the Hurwitz zeta-function
, as in [
13], are presented.
To obtain more general results, the universality of separate functions and some classes of functions are considered. One such class was introduced by A. Selberg (see [
14,
15]), known as the Selberg class
. The structure of the class
was studied by various authors, see [
8,
16,
17,
18,
19,
20], but until now its structure was not completely known. However, the class includes all main zeta- and
L-functions, for example,
,
, the zeta-functions of certain cusp forms, etc. The Selberg class
is defined axiomatically, with its functions
satisfying four axioms. Recall that the notation
,
, means that there is a positive constant
such that
, and that
denotes the Euler gamma-function. The axioms of the class
have the names:
- (1)
(Ramanujan conjecture). The estimate is valid with any .
- (2)
(Analytic continuation). For some , in an entire function of finite order.
- (3)
(Functional equation). Let
where
q,
, and
such that
. Then the functional equation of the form
is valid. Here,
, and, as usual, by
we denote the conjugate of
s.
- (4)
(Euler product). Let
with coefficients
such that
. Then the representation
holds.
Axioms (1)–(4) of the class
are insufficient to prove universality as they do not include the analogue of the prime number theorem. Therefore, J. Steuding, who was first to study the class
with an emphasis on universality [
8], introduced the following axioms.
- (5)
There exists
such that
where function
counts the number of primes up to
x. Moreover, in [
8] the Euler product of the type
- (6)
was required with some complex
.
For the universality for the above functions, we need one important ingredient of the class
. For
, the quantity
is called the degree of the function
L. The degree is an deep characteristic of the class
. If
, then
coincides with
or
with some
. For
, let
We denote by
,
the class of compact subsets of the strip
with connected complements, and
the class of continuous non-vanishing functions on
K that are analytic in the interior of
K. Then, in [
8], the following universality theorem has been proved.
Theorem 3. Suppose that satisfies Axioms (2), (3), (5) and (6). Let and . Then, for every , the inequalityholds. In [
21], Axiom (6) was removed. Thus, Theorem 3 holds for the so-called Selberg–Steuding class
; more precisely, for the functions belonging to the Selberg class and satisfying Axiom (5).
The discrete version of Theorem 3 has been obtained in [
22].
Theorem 4. Suppose that , K and are the same as in Theorem 3. Then, for every and , We can consider a simultaneous approximation of a tuple of analytic functions by a tuple of shifts of zeta- or
L-functions. This type of universality is called joint universality. This phenomenon of a Dirichlet series was also introduced by Voronin. In [
23], he studied the joint functional independence of Dirichlet
L-functions using the joint universality. Of course, the joint universality is more complicated, but, on the other hand, it is more interesting. Obviously, in the case of joint universality, the approximating shifts require some independence conditions. For example, Voronin used Dirichlet
L-functions with pairwise non-equivalent Dirichlet characters. Later, the joint universality theorems were proven for zeta-functions defined by a Dirichlet series with periodic coefficients, Matsumoto zeta-functions, and automorphic
L-functions. For these proofs, see the very informative paper [
9].
This paper deals with the discrete joint universality property for
L-functions for the class
. Let
be fixed positive numbers, and
. We define the multiset
and then we prove the following theorem.
Theorem 5. Suppose that , and the set is linearly independent over the field of rational numbers . For , let and . Then, for every and , Moreover, for all but at most countably many , the limitexists and is positive. In [
24], a joint continuous universality theorem for a function
on the approximation of analytic functions by shifts
with linear independence over
real algebraic numbers
was obtained.
For example, for , we can take , , and in Theorem 5.
We denote by
the Borel
-field of the space
, and let
P and
, where
, be probability measures on
. We report that
converges weakly to
P as
, and write
, if, for all bounded continuous functions
on
,
We derive Theorem 5 from a probabilistic joint discrete limit theorem on weakly convergent probability measures in the space of analytic functions. For proof of the latter theorem, we consider the weak convergence of probability measures on the infinitedimensional torus, and in the space of analytic functions for certain absolutely convergent Dirichlet series. After this, we show a comparison in the mean between the initial L-function and functions defined by an absolutely convergent Dirichlet series. This will give the desired joint discrete limit theorem for the tuple of functions we are interested in.
2. Case of the Torus
We define the infinite-dimensional torus as
where
is the infinite Cartesian product over prime numbers of unit circles. Since each circle is a compact set, by the Tikhonov theorem,
with the product topology and operation of pairwise multiplication is a compact topological abelian group. Now, we construct the set
where
,
. Then, the Tikhonov theorem again shows that
is a compact topological group. We denote by
,
,
,
, the elements of
.
For
, we set
In this section, we consider the weak convergence for as .
Proposition 1. Suppose that the set is linearly independent over . Then, , where is the probability Haar measure on .
Proof. The characters of the
are of the form
with integers
, where the star indicates that only a finite number of
are not zeroes. Therefore, the Fourier transform
,
,
, can be represented by
By a continuity theorem on the compact groups, for the proof of Proposition 1, it is sufficient to show that the Fourier transform
converges, as
, to the Fourier transform
of the Haar measure
. Here,
.
Equality (
1), obviously, gives
Thus, it remains to consider only the case
. Since the set
is linearly independent over
, we have, in this case,
Actually, if (
3) is false, then
for some
and the integers
. However, this contradicts the assumption that the set
is linearly independent. Now, using (3) and the formula for the sum of geometric progressions, we deduce from (
1) that, for
,
Hence,
for
. This, together with (
4), shows that
thus proving the Proposition 1.
□
We apply Proposition 1 for the proof of weak convergence for the measures defined by means of certain absolutely convergent Dirichlet series connected to the function
. We fix a number
, and
We define the functions
and
where, for
,
If
, then
with arbitrary
. Obviously,
decreases exponentially with respect to
m. Therefore, the series for
and
are absolutely convergent for
with arbitrary finite
and fixed
. Let
and
Moreover, let
stand for the space of analytic on
functions endowed with the topology of uniform convergence on compact sets, and let
For
, we set
Proposition 2. On , a probability measure exists such that .
Proof. Let the mapping
be given by
. The absolute convergence of the series for
,
, implies the continuity of
. Hence,
is
-measurable. Therefore, every probability measure
P on
induces the unique probability measure
on
given by
Let
be from Proposition 1. Then, for every
,
Hence, we have
. Therefore, Proposition 1, the continuity of
and Theorem 5.1 in [
25] show that
, where
. □
We see that the measure
is independent of
. This allows us to obtain the weak convergence of
as
, and identify the limit measure. Let
It is known [
8] that the Dirichlet series for
, for almost all
, is uniformly convergent on compact subsets of the strip
. Thus,
, for
, is a
-valued random element. The probability Haar measure
on
is the product of the Haar measure
on
, i.e., for
,
The above remarks show that
is a
-valued random element defined on the probability space
. We denote by
the distribution of
.
The measure
coincides with that studied in the continuous case in [
24]. Therefore, we have the following proposition.
Lemma 1. The relation holds. Moreover, the support of the measure is set as Proof. The first assertion of the lemma is contained in Lemma 7 in [
24], while the second one is in Lemma 9 in [
24]. □