1. Introduction
Let
be a complex variable. The term “zeta-function” steams from the Riemann zeta-function
Thus, zeta-functions in the classical sense are analytic functions of a complex variable in some half-plane
defined by Dirichlet series
with coefficients
of certain arithmetical nature, and have meromorphic continuation to the region
. In other words, zeta-functions are various generalizations of the function
.
The function
appears to be quite simple; however, its value distribution is especially complicated. This is clearly illustrated by the Riemann hypothesis on the location of the nontrivial zeros of
. Roughly speaking, the value distribution of zeta-functions is chaotic, and information on their majority concrete values is insufficient. On the other hand, in the context of certain problems, precise information on values of
is not necessary: certain average data are sufficient. This situation led H. Bohr to the idea that statistical methods could be applied for investigation of
(see ref. [
1]). By examining values of
on vertical lines
with a fixed
, one can take measurable sets
and consider the frequency of
t such that
. The first result of such a kind, for fixed
, was obtained in [
2]. Denote by J the Jordan measure on the real line, and let
R be a rectangle with edges parallel to the coordinate axis. Then, the main result of [
2] states that the limit
exists and depends only on
R and
. The case for
is more complicated than that of
, as it depends on the possible zeros of
. Therefore, in [
3], the existence of the limit for the frequency
as
was considered. Here,
where the union is taken over all zeros
of
located in the strip
. The proofs of these results are based on a theory of convex curves developed by the authors themselves [
4].
Bohr–Jessen theorems were further developed in [
5,
6], leading to probabilistic limit theorems on weakly convergent probability measures. Let
be a topological space with Borel
-field
, and let
P and
,
, be probability measures on
. Recall that
converges weakly to
P as
, denoted
, if for every bounded continuous real-valued function
g on
,
In this terminology, Bohr–Jessen theorems can be rephrased as follows: on the space
, there exists a probability measure
such that, for fixed
, and
the relation
holds (see, for example, refs. [
7,
8]). Here and throughout,
denotes the Lebesgue measure on the real line.
A new stage in the development of the probabilistic theory of zeta-functions was initiated by B. Bagchi. In his thesis [
9], he created a theory of probabilistic limit theorems in the space of analytic functions and applied it to prove universality property of zeta-functions on approximation of classes of analytic functions. In his theorems, Bagchi proposed a new way for identification of the limit measure. In [
10], we applied Bagchi’s method to prove a limit theorem for the Epstein zeta-function.
Suppose that
Q is a positive-definite quadratic matrix of order
. The Epstein zeta-function
is defined, for
, by the series
where
is the transpose of
. Moreover,
has analytic continuation to the whole complex plane, except for the point
, which is a simple pole with residue
, where
is the Euler gamma-function. The function
was introduced and studied in [
11]. Its author aimed to define the most general zeta-function with the functional equation of the Riemann type, i.e.,
The attempt was successful: Epstein proved the following functional equation
for
, where
denotes the inverse of
Q.
We observe that in a particular case, the function
can be expressed as an ordinary Dirichlet series. Actually, suppose that
for all
, and, for
,
is the number of
such that
. Then, we have
For example, for the
n-dimensional unit matrix
and
,
,
, where
is the Dirichlet
L-function, and
is the non-principal Dirichlet character modulo 4.
Note that the Epstein zeta-function is not only an object of interest in pure mathematics, particularly in algebra and algebraic number theory, but it also has practical applications. In fact, the function
appears in crystallography [
12], as well as in physics, including quantum-field theory [
13], and in studies related to energy and temperature [
14,
15].
The function is, in general, significantly more complicated than the Riemann zeta-function. For example, the analytic continuation and functional equation of depend essentially on the signature and rank of the real quadratic form Q on : when Q is positive-definite, the function admits a meromorphic continuation to the whole complex plane and satisfies a classical functional equation. In contrast, when Q is indefinite, the analytic behavior of becomes more intricate due to the presence of a continuous spectrum and its connections to non-holomorphic automorphic forms, such as Maass forms. Therefore, it is difficult to expect any general results about the value distribution characteristics of for all matrices Q.
As previously mentioned, a Bohr–Jessen type theorem for
was obtained under the following restrictions in [
10]. Based on (
1), it was shown in [
16] that
where
and
are zeta-functions of a certain Eisenstein series and a cusp form, respectively, related to the coefficients
. Moreover, even for
, it is known that the mentioned Eisenstein series is a modular form of weight
and level
such that
is an integral matrix [
17], and then the zeta-function
is a certain combination of Dirichlet
L-functions [
18]. This leads to the representation
where
k and
l run over positive divisors of the level
q, and
K and
L are finite integers; the characters
,
, and
,
are the pairwise nonequivalent Dirichlet characters modulo
and
, respectively, and
and
are the corresponding Dirichlet
L-functions. The coefficients
are certain complex numbers. Furthermore, the series with coefficients
is absolutely convergent for
.
The formulation of the main limit theorem relies on the following object
where
is the set of all prime numbers. With the product topology and operation of pairwise multiplication,
is a compact topological group, and this leads to the probability space
, where
is the probability Haar measure on
. Denote by
the elements of
, and extend
to the set
by using the formula
Now, on the probability space
, define the complex-valued random element
where
Denote by
the distribution
of the random element
. Then, the main result of [
10] is the following.
Proposition 1 (Theorem 2 of [
10])
. Suppose that is fixed. Then, the probability measureconverges weakly to the measure as . In [
19], a discrete version of Proposition 1 is presented. To state it, some definitions are needed. A number
h is called type 1 if the number
is irrational for all
. In the opposite case,
h is type 2. Then, in this case, there exists the smallest
such that
Let
be a subset of
:
Then,
, where
denotes the cardinality of the set
A.
We now return to the group
. For
, let
be the closed subgroup of
generated by
. Then,
, as
, is a compact group. Therefore, on
, the probability Haar measure
can be defined. It is known (see Lemma 4.2.2 of [
9] and Lemma 1 of [
20]) that
Let
, for
, be the complex-valued random element on the probability space
given by the Representation (
3), and
denote its distribution. Let
N run over the set
. Then, in [
19], the following statement has been obtained.
Proposition 2 (Theorem 1 of [
19])
. Suppose that and are fixed. Then, the probability measureconverges weakly to the measure as . In Propositions 1 and 2, the weak convergence of probability measures defined by frequencies in the intervals
and
is discussed. It is well known that frequencies in short intervals contain more information on the discussed objects. This motivates the refinement of the limit theorems in Propositions 1 and 2, leading to limit theorems in short intervals, i.e., intervals of length shorter than
T and
N. A result of this type, corresponding to Proposition 1, was obtained in [
21].
Proposition 3 (Theorem 2 of [
21])
. Suppose that is fixed, and . Then, the probability measureconverges weakly to the measure as . The purpose of this paper is to develop a more complex version of Proposition 2 in short intervals. For brevity, we use the notation
where the dots indicate a condition to be satisfied by
k. The main result is the following theorem.
Theorem 1. Suppose that and are fixed, and . Then, the probability measureconverges weakly to the measure as . Theorem 1 is theoretical, it extends and develops Bohr–Jessen’s ideas. Recall that even the eminent mathematician Atle Selberg, who was awarded the Fields Medal, devoted much attention to probabilistic limit theorems for zeta-functions; see [
22,
23]. Note that results concerning short intervals are highly valued in analytic number theory, especially those related to the distribution of zeros of zeta-functions and prime numbers. Moreover, problems in short intervals are more complicated. Theorem 1 continues investigations in this direction, providing new results that confirm the chaotic behaviour of the function
. The results presented here could be useful for researchers studying this function in applied mathematics.
Theorem 1 will be proved in
Section 3; the cases where
h is type 1 and type 2 will be considered separately. Before that, we prove limit lemmas (Lemmas 7 and 8) for probability measures defined on the one-dimensional torus
, the structure of which depends on the arithmetic nature of the number
h. Using Lemmas 7 and 8, we obtain limit lemmas (Lemmas 11 and 13) for probability measures defined in terms of
involving absolutely convergent Dirichlet series.
Section 2 is devoted to certain discrete mean estimates for
in short intervals. Lemma 6 occupies a central place with respect to short intervals. It shows that the functions
and
are close in the mean in such intervals.
3. Proof of the Theorem 1
For the proof of Theorem 1, we will apply the method of Fourier transforms, the preservation of weak convergence of probability measures under certain mappings, as well as a connection of weak convergence and convergence in distribution. For the identification of the limit measure, we will apply the results of [
10,
19].
We start with a limit lemma for probability measures on
. For
, set
Lemma 7. Suppose that is type 1, and as . Then, converges weakly to the Haar measure μ as .
Proof. The group
is compact, as the product of compact sets. Therefore, it provides a reason to study the Fourier transform of the measure
, and, from its convergence, derive the weak convergence for
. Denote by
the dual group, or character group, of
. It is a well-known and widely used fact that
is isomorphic to the group
where
for all
, and an element
, where only a finite number of integers
are non-zero, acts on
by the formula
From this, it follows that the characters of
are given by the product
where the sign “*” shows that only a finite number of
. This implies that the Fourier transform
of the measure
is
Hence, the definition of the measure
yields
For brevity, let
We observe that the requirement that
h is type 1 is equivalent to the linear independence over the field of rational numbers
for the multi-set
. Actually, if the set
is linearly dependent over
, then there exist the numbers
, and
that
and
and this contradicts the definition of type 1. Thus,
h of type 1 implies the linear independence of
. Similarly, if
h is not type 1, then there exists
such that
is a rational number. Hence, the set
is linearly dependent over
. This shows that linear independence of
implies the type 1 for
h.
The above remarks imply that
with
if and only if
and
. Hence, in view of Expression (
15),
if and only if
. If
, then
, and (
15) gives
Hence, for
,
Let
Then, Equalities (
16) and (
17) show that
Since
is the Fourier transform of the Haar measure
on
, and
, as a compact group, is the Lévy group, in view of Theorem 1.4.2 of [
29] and Equation (
18), we determine that the measure
converges weakly to
as
. □
Now consider the case with h of type 2.
Lemma 8. Suppose that is type 2, and as . Then, converges weakly to the Haar measure as .
Proof. It is known, see the proof of Lemma 4 of [
19], that in this case, characters of
are of the form
and only a finite number of integers
are not zero. Here,
is a finite subset of
. Hence, the Fourier transform
of
is
due to the equality
since
for
. Taking into account (
19) and (
20), we find that
for
,
, and
,
.
Now, suppose that
for some
, or there is
such that
for all
. Then, we have
Actually, if (
22) is not true, then
with some integer
. If
is a multiple of
(
is from the definition of type 2), then
with
. Hence, by (
23),
with some integer
, and this contradicts the linear independence over
Q of logarithms of prime numbers. If
is not a multiple of
, then the number
is irrational, and this contradicts (
23) since
is a rational number. These contradictions show that inequality (
22) is true. Let, for brevity,
Then, (
19) and (
22) yield that
This and Equality (
21) show that
and the proof of the lemma is complete because the right-hand side of the latter equality is the Fourier transform of
. □
Lemma 7 serves as an important component for the proof of a limit lemma for the function . For this, we recall one simple but useful property of weak convergence of probability measures.
Let
and
be two topological spaces, and
a continuous mapping. Then,
g is
-measurable, i.e.,
Then, every probability measure
P on
induces the unique probability measure
on
defined by
where
is the preimage of
A.
Lemma 9 (Section 5 of [
30])
. Let , , and P be probability measures on , and be a continuous mapping. Suppose that . Then, . For
, define
and, for
, set
Lemma 10. Suppose that is type 1, , and as . Then, on , there exists a probability measure such that .
Proof. Consider the mapping
given by the formula
All functions involved in the definition of
are absolutely convergent, hence, uniform in
. Therefore,
is a continuous mapping. Moreover, we have
Therefore,
for all
. Thus, we have
. This, continuity of
and Lemmas 7 and 9 proves that
. □
A separate analysis of weak convergence for the measure
as
is unnecessary, as it has already been investigated in [
19].
Lemma 11. Suppose that is fixed and is type 1. Then, .
We notice that the statement of the lemma has been obtained in the proof of Theorem 2 of [
10]. For identification of the limit measure of
as
, elements of the ergodic theory have been involved; more precisely, the classical Birkhoff–Khintchine ergodic theorem (see, ref. [
31]) has been applied.
Now, we resume the analysis of the case where h is type 2. Repeating the proof of Lemma 10 and using Lemma 8 leads to the following statement.
Lemma 12. Suppose that is type 2. Then, on , there exists a probability measure such that .
Proof. We notice that, in the case of type 2, the mapping depends on h. Thus, . □
The measure
does not depend on
M. Thus, we may use the results of [
19].
Lemma 13 ([
19], pp. 12–14)
. Suppose that is fixed and is type 2. Then . To complete the proof of Theorem 1, one statement on convergence in distribution is needed. Let , , and X be -valued random elements on a certain probability space , and and , be the corresponding distributions. We say that converges to X in distribution if .
Lemma 14 (Theorem 4.2 of [
30])
.Let and , , be -valued random elements on the probability space , and the space is separable. Suppose that, for every ,Moreover, let, for every ,Then, . Proof of Theorem 1. Case for h of type 1. On a certain probability space
, define a random variable
having the distribution
For
, set
and
In virtue of Lemma 10, it follows that
where
is the
-valued random element with the distribution
. Moreover, in view of Lemma 11,
Now, we apply Lemma 6 and find that, for fixed
and
,
The later equality, relations (
24) and (
25) together with Lemma 14 show that
i.e.,
.
Case for h of type 2. We repeat the proof of Theorem 1 in the case of h of type preserving the notations with minor changes.
By Lemma 12, we have
where
is the
-valued random element with the distribution
, and from Lemma 13 it follows that
Using Lemma 6 yields, for
and
,
Now, (
26)–(
28) and Lemma 14 prove the theorem in the case of
h of type 2, i.e.,
The proof of Theorem 1 is complete. □