Fractional Integral Operators on Generalized Campanato Spaces with Variable Exponents ()
1. Introduction
In the last several decades, research on Morrey-Campanato-type spaces and their applications has seen a remarkable increase, with each advancing the other. Campanato spaces, as a generalization of Morrey spaces that are occasionally referred to as Morrey-Campanato spaces, were introduced by S. Campanato in 1963 [1]. These spaces also showed up in the paper by G. Stampacchia [2]. They generalize the BMO spaces that were introduced by F. John and L. Nirenberg [3] and defined as follows.
(1.1)
where
, and what follows
is the Lebesgue measure of measurable set
in
.
The interpolation features of Morrey-Campanato-type spaces were given in [2], [4] and [5]. In truth, they derived the result for a more comprehensive space that is now termed the Campanato spaces.
Let Ω be an open set in
, we denote
,
and
,
. The Morrey space
is defined as
(1.2)
This is a Banach space with respect to the norm
(1.3)
Campanato spaces are useful tools in the regularity theory of PDEs as a result of their better structures, which allow us to give an integral characterization of the spaces of Hölder continuous functions. The Campanato space
is defined as
(1.4)
The Campanato semi norm is given as
(1.5)
The importance of Campanato spaces stems from the fact that, as
, they coincide with the spaces of Hölder continuous functions, providing an integral characterization of such functions. While in the case
less then
they coincide with Morrey spaces. For some recent results for research on Morrey-Campanato-type space and their applications, see [6]-[12].
The main purpose of this paper is to discuss the boundedness of the fractional integrals and their commutators on the generalized Campanato space with variable exponents given in Section 3.
The paper is organized as follows. Some main definitions and some auxiliary results are provided in Section 2. In Section 3, we prove the main boundedness results and provide some corollaries.
2. Some Preliminaries
2.1. Fractional Integral Operator
The commutator generated by a linear operator
and a locally integrable function
is defined by
(2.1)
The investigation into the operator
commenced with the groundbreaking studied by Coifman-Rochberg-Weiss in [13]. Two primary motivations underlie the consideration of commutators in this context. First, the boundedness of commutators provides valuable insights into the characterization of function spaces [14]-[20]. Second, commutator theory is of great importance in analyzing the regularity of solutions to second-order elliptic and parabolic partial differential equations [21]-[23]. The well-posedness of solutions to many PDEs depends on the boundedness of commutators for integral operators [24]-[26]. Fractional integrals play an important role in harmonic analysis and PDEs.
Let
. Then the fractional integral is defined by
(2.2)
Let
be a locally integrable function.
is the m-th commutator generated by
and
, which is defined by
(2.3)
In [27], Chanillo studied the characterization of BMO space via boundedness of
. Paluszynski derived the characterization of the
space, which involves the boundedness of
[19]. In contrast to the extensive and profound results regarding commutators with symbol functions in BMO spaces and lipschite spaces, research on Morrey spaces has been relatively scarce. However, recent developments have emerged, providing innovative characterizations of Morrey spaces through the analysis of the boundedness of commutators associated with Calderón-Zygmund singular integral operators. These developments utilize novel methodologies, moving away from the traditional dependence on the sharp maximal function theorem [28]. This previous breakthrough marks the initial exploration of commutator studies where the symbol function belongs to Morrey spaces.
Let
be a locally integrable function on
. Hardy-Littlewood maximal operator
is defined by
(2.4)
given
,
, the fractional maximal operator
is defined by
(2.5)
where the supremum is again taken over all balls B which contain
. In the limiting case
, the fractional maximal operator reduces to the Hardy-Littlewood maximal operator.
2.2. Variable Exponent Lebesgue Spaces
Let
be a measurable function on
taking values in
, the Lebesgue space with variable exponent
is defined by
(2.6)
Then
is a Banach function space equipped with the norm
(2.7)
The space
is defined by
for all compact subsets
, where and what follows,
denotes the characteristic function of a measurable set
. Let
, we denote
,
. The set
consists of all
satisfying
and
;
consists of all
satisfying
and
.
can be similarly defined as above for
. Let
be the conjugate exponent of
, that means
.
Basic facts concerning Lebesgue spaces with variable exponent can be found in [29] [30]
Lemma 2.1 see [31] Let
belong to the Lebesgue space
and
belong to
, where
. Then, the product function
is integrable over
, and the following inequality holds.
(2.8)
where
is defined as
.
Theorem 2.2 see [32] Let Ω be an open set in
, and
,
, Let
, be such that
. Suppose further that
satisfies
(i) For any
,
, if
(2.9)
(ii) For any
,
, if
(2.10)
Define
,
, by
(2.11)
Then the fractional maximal operator
is bounded from
to
.
Theorem 2.3 see [32] Let
and
are as in Theorem 2.2, then the fractional integral operator
is bounded from
to
.
3. Boundedness of Fractional Integral Operators
3.1. Variable Exponent Generalized Morrey Spaces
The Morrey spaces
with variable exponents
and
over an open set
have recently been introduced almost simultaneously by different authors. Relevant works can be found in [33]-[37].
Let
be a measurable function on Ω with the values in
. The variable exponents Morrey space
is defined [33] as the set of measurable functions f on Ω with the finite norm
(3.1)
The Campanato space
of variable order, in the Euclidean case, are defined [38] via the norm
(3.2)
Let
and
.
satisfies the doubling condition
where
is a constant independent of
and
.
Let
be the ball of centre at
and radius
. The generalized variable Morrey space
was introduced in [39] as the space of functions equipped with the norm
(3.3)
The generalized variable Campanato space
is defined by the norm
(3.4)
The generalized Campanato spaces with variable exponents are a key link in the “space-operator-equation” chain in functional analysis. They not only generalize the theory of classical spaces but also provide a powerful functional analysis framework for dealing with problems of non-uniform regularity, such as analysis under variable coefficients and inhomogeneous media.
In [40], a characterization of Campanato spaces was provided via the boundedness of
under the assumption that
satisfies a mean value inequality. In [41], authors respectively characterized the strong and weak type boundedness of
on the generalized Morrey spaces. Inspired by these results, in this paper, we obtain a characterization of
via the boundedness of
under a suitable assumption that
satisfies the mean value inequality. A function
is said to satisfy the well-known mean value inequality if there exists a constant
such that for any ball
in
,
(3.5)
(3.5) is also called the reverse Hölder class which contains many kinds of functions. For example, if
is a polynomial and
, then
satisfies (3.5) [42]. Besides polynomial functions, the mean value equalities also characterize harmonic functions [43]. For more theories about the reverse Hölder classes see [44] and [45] for example.
Our main results can be stated as follows.
Theorem 3.1 Let
,
,
,
is a non-negative measurable function on
,
, and
satisfies the doubling condition. Then the fractional integral
is a bounded operator from
to
.
Theorem 3.2 Let
.
satisfies the doubling condition,
,
,
satisfy (3.5). Then the following statements are equivalent.
(a)
;
(b)
is a bounded operator from
to
;
(c) For any ball
and
, there exists a constant
such that
As mentioned in a prior study, solutions to a substantial category of elliptic second-order PDEs fulfill the mean value inequality (3.5). In this manner, our primary findings can provide characterizations for the spaces of solutions to some second-order elliptic PDEs. Let
be a solution to the following Laplace equation
(3.6)
where
is the Laplace operator and
is a function defined on a bounded domain
. As shown in a previous study,
satisfies (3.5) and is bounded [38]. Thus, the corresponding commutator is bounded from
to
. Then the space of solutions to (3.8) is
. The regularity of solutions to some elliptic PDEs with smooth boundaries can be credited to the boundedness of corresponding commutators with smooth kernels. The question of what occurs when the boundary conditions are relaxed can be addressed by studying the boundedness of commutators with rough kernels. For this purpose, the results are extended to the case of rough kernels.
This can be justified by simply stating that the proof follows that of Theorem 3.2 with minor adjustments to the estimates to account for the kernel
. we can claim that Theorem 3.2 also holds true for fractional integrals with homogeneous kernels, which were defined [40] by
(3.7)
where
is homogeneous of degree 0 and
(3.8)
Here
is the normalized Lebesgue measure and
. When
,
is the same as the fractional integral
. If
, then
becomes a Calderόn-Zygmund singular integral operator in the sense of the principal value Cauchy integral.
Corollary 3.3 Let
and
be as Theorem 3.2. Then the following conditions are equivalent.
(1)
;
(2)
is a bounded operator from
to
;
(3) For any ball
and
, there exists a constant
such that
3.2. Proof of Our Main Results
This section begins with some lemmas which will be used in the proof of our results.
Lemma 3.4 Let
and
be as Theorem 3.2, and
. Then
(3.9)
Proof. By applying the Hölder inequality we get
where
.
Lemma 3.5 Let
,
.Then the following inequality holds.
(3.10)
Proof. We suppose two cases.
Case 1. When
. By using Hölder inequality
Case 2. When
. Choose a sequence of nested cubes
Similarly to case 1, we have
Proof of Theorem 3.1. Let
be a function in
. For any fixed
, denote
by
.
. For any ball
, let
. We write
as
, where
and
. By the Minkowski inequality, we have
We first estimate
.
Note that
and
is bounded from
to
, we get
Thus,
Now we turn to prove
. For
and
, we have
, together with the Hölder inequality, we obtain
Therefore,
This completes the proof.
Proof of Theorem 3.2. Let
be a function in
. For any fixed
, denote
by
. It suffices to show that the fact
For any ball
, let
. We write
as
, where
and
. By using the facts
, and
([19], p. 5, [40], p. 6) imply the following.
Now we first estimate
,
where
,
.
Next we will find the estimate for
where
,
.
Now we will find the estimate for the term
,
Now we find the estimate for
, by using the boundedness of
(Theorem 2.2) and Lemma 3.4, we have.
Now we find the estimate for
. By using the boundedness of
(Theorem 2.2) and Lemma 3.5, we have.
which completes our desired results.
Now we will prove that (b)
(c). In this case, the induction can be performed on
.
is trivial, so it can be assumed that for any
,
Next, the case
can be shown. The following can be confirmed easily.
The condition (3.5) and Hölder inequality allow the user to estimate
as follows.
Similarly, the following can be shown.
This produces the following result.
(c)
(a), this proof consists of the construction of proper commutator [28]. This depends on the smoothness of the kernel function
. Choosing
and
such that
can be expressed in the neighborhood
as an absolute convergent Fourier series
where the exact form of the vectors
is irrelevant. Set
. If
,
Choose now any ball
. Set
and
. Thus, if
and
, then
Denote
, and using Hölder inequality, we have.
where
.
Hence
yields the desired results.
Acknowledgement
This work is supported by the National Natural Science Foundation of China (Grant No. 12361018).