1. Introduction
Throughout the paper,
is the main complex variable. We consider the approximation of analytic functions by Hurwitz zeta-functions. Let
be a fixed parameter. The Hurwitz zeta-function
was introduced in [
1] and, for
, is defined by the Dirichlet series
Moreover,
has an analytic continuation to
, and the point
is its simple pole and
. Clearly,
coincides with the Riemann zeta-function
and
The latter observations show that the function
is an extension of the famous Riemann zeta-function. Unlike
, the function
, except for values
and
, has no Euler product over prime numbers. Hence, the value distribution of
differs from that of
. For example, it is well known that
for
, while
, where
,
, has infinitely many zeros in the latter half-plane [
2,
3,
4].
On the other hand, the function
has an indirect connection to the distribution of prime numbers in arithmetic progressions. The main tool for investigations of the asymptotics for
is Dirichlet
L-functions. Let
be a Dirichlet character modulo
q (where
is periodic with period
q, completely multiplicative,
if
, and
for
). The Dirichlet
L-function
with character
, for
, is given by
and has meromorphic continuation to the whole
. From the periodicity of
, it follows that
Thus, properties of
with rational
can be applied for investigations of Dirichlet
L-functions, and consequently for
. Nevertheless, applications of the Hurwitz zeta-function are not limited by the distribution of prime numbers;
plays an important role in special function theory, algebraic number theory, probability theory, and even quantum mechanics. The classical theory of
can be found in [
5,
6,
7]. One significant feature of
is connected to the approximation of analytic functions by shifts
,
. This approximation is of a novel type in function theory, and is called universality: shifts of one and the same function
approximate the whole class of analytic functions. The universality of the Riemann zeta-function
was discovered by S.M. Voronin in [
8,
9,
10,
11,
12,
13]. After Voronin, the universality of
was studied by many authors (see [
14,
15,
16,
17,
18]). We recall some universality results for
. For
, denote by
the class of compact subsets of the strip
with connected complements, and by
,
, the set of continuous functions on
K that are analytic inside of
K. Let
be the Lebesgue measure of measurable set
. Then, the following result is known [
14,
15,
16,
18,
19].
Proposition 1. Suppose that α is rational or , or a transcendental number, and , . Then, for every ,Moreover, the limitexists and is positive for all but at most countably many . If on K, then the proposition remains valid for or as well.
The case of algebraic irrational
(
is a root of a polynomial
with rational coefficients) has been considered in [
20]. Denote by
d the degree of
, and put
and
. Then the universality of
with algebraic irrational
is contained in the following statement.
Proposition 2 (see [
20])
. Suppose that , , , and is a continuous function on the disc , , and analytic inside of that disc. Let and . Then, for all but finitely many of degree at most withthere are and such thatHere, is given explicitly, the set of exceptional α can be described effectively, and κ can be effectively computable as well. Propositions 1 and 2 are of the so-called continuous type because
in shifts
can take arbitrary values in the interval. In parallel to continuous universality theorems for zeta-functions, theorems of discrete universality are studied when
takes values from certain discrete sets. The first discrete universality theorem for zeta-functions has been obtained by A. Reich. In [
21], he proved the discrete universality of Dedekind zeta-functions
of algebraic number fields
on the approximation of analytic functions by shifts
,
, where
h is a fixed real number.
For statements of discrete universality theorems, we introduce some notation. Denote by
the cardinality of the set
A, and, for
, put
where in place of dots a condition satisfied for
k is to be written. The first discrete universality theorem for
has been obtained by B. Bagchi.
Proposition 3 (see [
15], Corollary 5.3.7)
. Suppose that α is a rational number , , and , , . Then, for every , Discrete universality of
with non-rational
involves the set
which can be a multiset.
Proposition 4 (see [
22,
23])
. Let the set be linearly independent over . Then, for every , and , the inequality (
1)
is valid. Moreover, the lower limit in (
1)
can be replaced by the limit for all but at most countably many . The second assertion of Proposition 4 has been obtained in [
23,
24]. As was noted in [
22], one can take
and
in Proposition 4.
In [
25], A. Sourmelidis proved that continuous universality for
implies a discrete one with shifts
,
. Hence, Proposition 1 implies Proposition 3 not only with rational but also with transcendental
. On the other hand, Proposition 4 may be true with algebraic irrational
; however, examples of such
are not known.
Also, a joint universality of Hurwitz zeta-functions is considered. In this case, a tuple
of analytic functions is approximated simultaneously by shifts
with both continuous and discrete
. Obviously, for this, the functions
must be independent in a certain sense. This independence may be described in terms of parameters
, for example, that
are algebraically independent over
, i.e., there is no polynomial
with coefficients in
such that
. A more general case involves the set
. The following statement is known [
26].
Proposition 5. Suppose that the set is linearly independent over . For , let and . Then, for every , The first joint discrete universality theorem was obtained for Hurwitz zeta-functions with rational parameters. For
, denote the Euler totient function as
; let
be pairwise non-equivalent modulo Dirichlet characters
q, and let
be the quadratic matrix of order
. For some functions
,
,
, define the matrix
where
means the transpose of a matrix
B. Then in [
27], we find the following result.
Proposition 6. Suppose that , for each , ; let , and all components of be non-vanishing on K. Then, for every , The most general joint discrete universality theorem for Hurwitz zeta-functions uses the set
Proposition 7 (see [
28], Theorem 1.7)
. Suppose that the set is linearly independent over . For , let , . Then, for every , the inequalityholds. Propositions 1 and 3–7 imply that there are infinitely many shifts of the Hurwitz zeta-function approximating a given analytic function or a tuple of analytic functions; however, any concrete shift is not known. In this sense, the mentioned results are ineffective. Proposition 2 has effectivity features because it indicates the explicit interval containing values with approximating property.
Another way towards effectivisation of universality for zeta-functions consists of shortening of intervals with approximating values
. This idea leads to extension of universality theorems for zeta-functions in short intervals. The first result in this direction for the Riemann zeta-function has been given in [
29], and improved in [
30,
31]. We recall that some universality results for Hurwitz zeta-function in short intervals. The main theorem of [
32] is stated as follows.
Proposition 8 (see [
32], Theorem 4)
. Suppose that the numbers are algebraically independent over , and . For , let and . Then, for every ,Moreover, the limitexists, is explicitly given, and positive for all but at most countably many . The case with
for transcendental
was obtained in [
33].
For
and
, set
where in place of dots a condition satisfied by
k is to be written. A version of Proposition 4 in short intervals has been proved in [
34].
Proposition 9 (see [
34], Theorem 1.5)
. Suppose that the set is linearly independent over , and . Then, for every , and ,Moreover, the lower limit can be replaced by the limit for all but at most countably many . The purpose of this paper is to connect Propositions 8 and 9, i.e., to obtain joint discrete universality for Hurwitz zeta-functions in short intervals.
Theorem 1. Suppose that the set is linearly independent over , and . For , let and . Then, for every ,Moreover, the limitexists and is positive for all but at most countably many . Denote by
the space of analytic functions on the strip
endowed with the topology of uniform convergence on compacta, and let
is considered with the product topology.
Theorem 2. Suppose that the parameter , , and the positive numbers are arbitrary, and . Then there exists a closed non-empty set such that, for compact sets , and every ,Moreover, the limitexists and is positive for all but at most countably many . We observe that .
Remark 1. We suppose in Theorem 2 that and because, in these cases, coincides with the Riemann zeta-function , or differs from it by a simple multiple . The function has the Euler product, and, for studying its universality, another scheme is used. Moreover, the above restriction for , in the case , removes confusion because the universality for in short intervals is known, and Theorem 2 then becomes meaningless.
Theorems 1 and 2 will be proved in
Section 4.
Section 2 is devoted to mean value estimates for Hurwitz zeta-functions in short intervals. In
Section 3, we will prove limit theorems on weakly convergent probability measures in the space of analytic functions
.
2. Estimates in Short Intervals
Throughout the paper, we will often use the notation , , , which means that there exists a constant such that . Thus, is an equivalent of .
Lemma 1. Suppose that and are fixed, and . Then, uniformly in H, the estimateholds. Proof. In the proof of Theorem 2 from [
35], the bound of Lemma 1 was obtained for
and fixed
. For this, the exponent pair
for the estimation of mean squares of Dirichlet polynomials has been applied. Using the exponent pair
in place of
gives Lemma 1. □
Since the present paper is devoted to discrete value distribution problems of Hurwitz zeta-functions, we need a discrete version of Lemma 1. To pass from Lemma 1 to its discrete analogue, we will apply the following Gallagher lemma which connects continuous and discrete mean squares.
Lemma 2 (see [
36], Lemma 1.4)
. Suppose that , , is a finite non-empty set, , andLet a complex-valuable function be continuous on the interval , and have a continuous derivative inside this interval. Then the inequalityis valid. Lemma 3. Suppose that , and fixed , , and . Then the estimateis valid. Proof. We apply Lemma 2, with
,
,
,
and
. Obviously,
Therefore, in virtue of Lemma 2,
Clearly, for large
N,
We have
for
, and
for
and large
N. Hence, Lemma 1 gives
Observe that
. Therefore, a standard application of the Cauchy integral formula and (
3) leads to the estimate
This, together with (
3) and (
2), yields the estimate of the lemma. □
Let, for brevity,
, and
. For the investigation of
, we introduce an auxiliary object. Let
be a fixed number, and, for
, and
,
where
. Define the series
which is absolutely convergent in any half-plane
with a finite
. Let
. Our aim is to replace the investigation of
by a simpler one of
. We will show that
and
coincide in the mean. To describe this, we need the metric in the space
.
It is well known (see, for example, [
37]) that there exists a sequence of compact subsets
such that
for all
,
and every compact set
lies in some
. For
, put
Then,
is a metric in
, which induces its topology of uniform convergence on compact sets.
Now, let
,
. Then
is the metric in
, inducing its product topology.
We now state the main lemma of this section. Let .
Lemma 4. Suppose that for , and . Then Proof. The definitions of the metrics
and
show that it suffices to prove the equalities
for every compact set
.
We will fix the parameter
, the number
, and a compact set
, and recall the integral representation for
. Let, as usual,
stand for the Euler gamma-function, and
where
is the number from the definition of
. Then the integral representation
holds. It follows easily from the classical Mellin formula
and the definition of
.
There exists
such that
for all
. Now, we choose more precisely
, and introduce
. Clearly,
but
. Therefore, the integrand in (
5) has a simple pole
of the function
with residue
, and a simple pole
of
with residue
, both of which lie in the strip
. Hence, taking into account the exponential decreasing of the gamma-function,
and applying the residue theorem, we obtain, for
,
Hence,
For simplification of further estimations, we present some elementary results. For
and
(see, for example, [
38]), the estimate
holds, and, in view of (
6), we have, for
,
and
Therefore,
and
This and (
7) lead to
It is easily seen that, for large
N,
For the estimation of
, we will apply Lemma 3. Thus, for
,
Hence, by (
8),
Now, summarising the results (
9)–(
11), we obtain
We notice that the implied constant in (
12) depends on
; however, this is omitted because
depends on
K.
Taking
in (
12), and then
, we obtain (
4) with
and
. The proof of the lemma is complete. □
3. Results on Weak Convergence
This section is devoted to discrete limit theorems on weakly convergent probability measures in the space
. Recall the definition of weak convergence. As usual, denote by
the Borel
-field of a topological space
. Let
,
, and
Q be probability measures on the measurable space
. Then
converges weakly to
Q as
(
) if, for every real continuous bounded function
f on
, the equality
holds. The theory of weak convergence of probability measures is presented in the monograph [
39].
In this section, we will deal with the probability measure
We will consider the asymptotic behaviour of
as
by using some auxiliary spaces and probability measures on them. We start with analysing probability measures on a certain group. Weak convergence on locally compact groups is developed in [
40].
Denote by
the Cartesian product of unit circles over
, i.e.,
With the product topology and pointwise multiplication,
is a compact Abelian group, which is the product of compact sets (Tikhonov theorem [
41,
42]). From this, the existence of the probability Haar measure
follows [
43]. Denote by
the elements of
.
Introduce one more group
where
for all
. Then, again,
is a compact topological Abelian group, and, on
, the probability Haar measure
can be defined. Let
,
,
, be elements of
. Notice that the Haar measure
is the product of Haar measures
,
, i.e., for
,
,
, we have
For
, define the probability measure
and consider its weak convergence as
.
Lemma 5. Suppose that and , where , are arbitrary, and is as . Then, on , there exists a probability measure such that .
Proof. As it is mentioned in [
40], for the proof of weak convergence on groups, it is convenient to use Fourier transforms. Since characters of the group
have the representation
where the star ∗ indicates that the integers
only for a finite number of
, the Fourier transform
,
,
, of the measure
is defined by
Hence, we find
For
with
, equality (
13) yields
Otherwise, we have
in view of the formula for a sum of geometric progression. Therefore, taking into account (
13) and (
14), we obtain
because, in (
14), the numerator is bounded and the denominator tends to ∞ as
. Since the group
is compact, it is the Lévy group. Hence, the convergence of Fourier transforms implies weak convergence of the corresponding probability measures. Let the probability measure
on
be given by the Fourier transform
Then, by (
15), we have
. □
Lemma 6. Suppose that the set is linearly independent over , and as . Then the relation holds.
Proof. The lemma is a corollary of Lemma 5. Actually, since the set
is linearly independent over
, the equality
holds if and only if
for
, and
. Let
denote a collection consisting from zeros. Thus, in virtue of (
16), in this case,
and this proves the lemma because the right-hand side of the latter equality corresponds to the Fourier transform of the Haar measure
on
. □
The next limit lemma concerns
. For
, set
Lemma 7. Suppose that and , , are arbitrary numbers, and as . Then on , there exists a probability measure such that .
Proof. We will apply the principle of preservation of weak convergence under certain mappings; see 5.1 of [
39].
Define
and put
Consider the mapping
given by
The series for
are absolutely convergent in any half-plane
; therefore, the mapping
is continuous. Moreover, it follows that
and thus, for
,
where
is the measure from Lemma 5, and
denotes the preimage of the set
A. Since
is continuous, it is
-measurable [
39]. Hence, the limit measure
in Lemma 5 defines a new measure
in
given by
These observations, Lemma 5, and Theorem 5.1 of [
39] show that
converges weakly to the measure
as
. Thus, denoting
, we have
. □
Lemma 8. Suppose that the set is linearly independent over and as . Then, the relation holds.
Proof. We repeat the proof of Lemma 7 using Lemma 6 in place of Lemma 5. □
Lemma 7 is a key for the proof of weak convergence for in the general case. Additionally, we need one classical result on convergence in distribution. Let , , and X be -valued random elements with distributions and P, respectively. We say that converges to X as in distribution () if and only if .
Lemma 9 (see [
39], Theorem 4.2)
. Suppose that the metric space is separable, -valued random elements and , , are defined on the same probability space with a measure ν, and the relationsandhold. If, for every ,then . Theorem 3. Suppose that and , , are arbitrary, and . Then, on , there exists a probability measure such that .
Proof. First we will prove that the limit measure
is tight, i.e., that, for every
, there is a compact set
such that
for all
. We observe that it suffices to show the tightness for the marginal measures of
,
Actually, if
are tight, then, for
, there exists compact sets
such that
for all
. Let
. Then
for all
. Hence, inequality (
17) holds. Thus, it is sufficient to consider
with arbitrary
and
.
Let
be a compact set of
from the definition of the metric
. There exists
such that
for
. Then, by Lemma 3, we have on the hypothesis for
M,
Hence,
This, together with the Cauchy integral formula, yields
Therefore, in view of (
4),
Suppose that
is a random variable defined on a certain probability space
and having the distribution
Introduce the
-valued random element
and denote by
the random element having the distribution
. Then the assertion of Lemma 7 implies
Since convergence in
is uniform on compact sets, from this, we get
Now, we fix
, and put
. Then (
19) and (
20) give
Let
Then
K is a uniformly bounded set in
, and hence it is compact; by (
21),
for all
. Hence, by the definition of
,
for all
, i.e.,
is tight. □
Now, we continue the direct proof of the theorem preserving the notation for
. Since the measure
is tight, by the classical Prokhorov theorem (see [
39], Theorem 6.1), it is relatively compact, i.e., every sequence of
contains a subsequence weakly convergent to a certain probability measure on
. Thus, let
be a subsequence such that
with a probability measure on
.
Denoting
the
-valued random element with distribution
, we may rewrite this in the form
Moreover, setting
in virtue of Lemma 7, we have
Introduce one more
-valued random element
Then, application of Lemma 4, for
, yields
Therefore, the latter equality, (
22), and (
23) show that Lemma 9 is applicable for the random elements
,
, and
, which corresponds to the measure
. This leads to the relation
and the definition of
proves the theorem.
The case of Theorem 3 with the linearly independent over
set
is simpler, and we have a full answer. On the probability space
, define the
-valued random element
where
and denote by
its distribution. Thus,
Theorem 4. Suppose that the set is linearly independent over , and . Then .
Proof. Denote by
the limit measure in Lemma 8, i.e.,
. This measure is independent of
M and
. In other words, we have the same situation as considered in [
28]; see proofs of Lemma 2.6 and Theorem 2.1. Thus, we have the relation
Let
be the
-valued random element with distribution
. Then, by Lemma 8, the relation
holds. Moreover, Lemma 4 yields
for every
. Here, we have used the notations of the proof of Theorem 3. Now, Lemma 9 and (
24)–(
26) give the assumption of the theorem. □