1. Introduction
Throughout this paper,
stands for an open subset of
with
. We recall functions of bounded variation, which is to say functions whose weak first partial derivatives are Radon measures. Precisely, see the following definition of BV functions (cf. [
1,
2]).
Definition 1. A function is called a function of bounded variation if its total variationis finite. The class of all such functions will be denoted by , where the space is endowed with the norm Since Sobolev functions are contained within the class of BV functions of several variables, so the BV functions play an important role in some problems of variational science. For example, these function spaces have good completeness and compactness properties, consequently they are often proper settings for the applications of functional analysis, linear and nonlinear PDE theory, mathematical physics equations and fractional differential Equations (cf. [
3,
4,
5,
6]). The BV function is classically applied to the minimum area problem and the free discontinuity problem (cf. [
7]). Please see [
1,
8,
9,
10,
11] for more details. Another interesting and important aspect of the results is the analysis of sets of so-called finite perimeters. These sets have applications in a variety of settings due to their generality and utility. In [
12], Ambrosio investigated fine properties of sets of finite perimeters in doubling metric measure spaces.
It is well known that the notion of capacity is critical in describing the null sets used to handle the pointwise behavior of Sobolev functions. The functional capacities play a significant role in every branch of mathematics, such as analysis, geometry, mathematical physics, and PDEs. Refer to [
13,
14,
15] for more details. In recent years, the BV capacity has attracted the attention of many scholars. In 1989, Ziemer introduced the classical BV capacity and the related capacity inequality in [
2]. In 2010, Hakkarainen and Kinnunen [
16] studied basic properties of the BV capacity and Sobolev capacity of order one in a complete metric space equipped with a doubling measure and supporting a weak Poincaré inequality. Hakkarainen and Shanmugalingam [
17] further studied the relationship between the variational Sobolev 1-capacity and the BV capacity. In 2016, Xiao [
18] introduced the BV-type capacity in the Gaussian space
and applied the Gaussian BV capacity to trace theory of the Gaussian BV space. In 2017, Liu [
19] obtained some sharp traces and isocapacity inequalities using the BV capacity on the generalized Grushin plane
. Recently, inspired by the classical case
, Huang, Li and Liu studied the BV capacity and perimeter from the
-Hermite bounded variation (cf. [
10]). See [
17,
18,
20,
21] and the references therein for more on this topic.
Dunkl operator is found to play an increasingly important role in the study of many special functional problems with reflective symmetry. In a certain sense, the study of harmonic analysis related to the Dunkl operator is a further development of classical harmonic analysis theory and a popular research direction in the field of modern mathematics. Dunkl operator is a parameterized differential difference operator related to the finite reflection group, which operates in the Euclidean space. In recent years, these operators and their generalizations have gained considerable interest in various fields of mathematics and physics. They provide a useful tool for studying specific functions related to root systems.
Combined with the existing research results mentioned above, we know that there have been many studies of bounded variation functions in different settings due to their important roles in some problems of variational science. As we know, there are no theoretic research results in the Dunkl setting. In this paper, we will focus on the BV space, the related BV capacity and an interesting heat semigroup characterization in the context of Dunkl theory. These results may have potential applications in the theory of function spaces and PDEs in the Dunkl setting.
At first, we will present a very short introduction to Dunkl operator which is used in the following sections. The Dunkl operators were first introduced and studied by Dunkl in [
22]. See [
23,
24] for the general theory of root systems, and see [
25,
26,
27,
28] for an overview of the rational Dunkl theory.
We shall always assume that the
N-dimensional Euclidean space
is equipped with the standard Euclidean scalar product
. A root system
is a finite set if
and
for every
. The symbol
is the reflection in the hyperplane
orthogonal to
, that is,
where
and
(cf. [
24,
28]).
Note that the reflection group generated by every reflection for is finite, which is called the Weyl group, and is contained in the orthogonal group Write a root system R as the disjoint union , where and are separated by a hyperplane through the origin. Let us call a positive subsystem. Any root system can not be uniquely written as an disjoint union, but our decomposition does not affect the following definition due to the G-invariance of the coefficients k. is divided into connected open components by the set of hyperplanes which is named the Weyl chambers For convenience, we allow that R is normalized in the sense that for every .
A function
is called a multiplicity function on
R, if it is invariant under the natural action of
G on
R, that is,
for all
and
The weight function
associated with Dunkl operators on
is defined in [
22,
24,
28] as follows:
The function
is
G-invariant, that is,
for all
Furthermore, it is also a homogeneous function of degree
, where
We have
for all
due to the
G-invariance of
Hence, this definition does not depend on the special choice of
The following definitions and facts for Dunkl operator and Dunkl gradient can be seen from [
28].
For
, the Dunkl operator associated with
G on
is defined by
where
denotes the directional derivative corresponding to the
i-th standard basis vector
The Dunkl gradient is denoted by
and the Dunkl Laplacian is naturally denoted by
, more specifically,
Notice when
, the
is reduced to the usual partial derivatives. At this time,
and
indicate the usual gradient and Laplacian, respectively. Denote by
the weighted space, where
is the Dunkl measure. The following anti-symmetry of the Dunkl operator
holds for all
. Moreover, we need to know an important product rule: If
and at least one of them is
G-invariant, then
The structure of the paper is given as follows. In
Section 2, we investigate the Dunkl-bounded variation space
and obtain some basic results of BV functions belonging to
, such as the structure theorem, the lower semicontinuity, approximation with smooth functions, the compactness result, and the Gauss–Green Theorem. In
Section 3, we introduce the Dunkl BV capacity
for a set
and investigate the measure theoretic properties of
, we discuss the capacity of a Borel set by approximating with compact sets from inside and open sets from outside, furthermore, we study its connection to the Hausdorff measure of codimension one, which shows that the Dunkl BV capacity and the Hausdorff measure of codimension one have the same null sets.
Section 4 is devoted to some results concerning the behavior of the heat semigroup for Dunkl BV functions and obtaining a heat semigroup characterization of bounded variation in the Dunkl setting. Finally,
Section 5 gives a conclusion in this article.
It should be noted that, compared with the classical cases from previous works, we need to overcome some key difficulties by seeking some new methods and techniques in the proofs of our main theorems, such as, approximation with smooth functions, the Gauss–Green Theorem and the heat semigroup characterization from Dunkl-bounded variation, etc. Some key difficulties come from the difference term of the Dunkl operator. See the results in the following sections for the details.
2. Dunkl BV Space
Firstly, we introduce a suitable notion of functions of Dunkl bounded variation. The Dunkl divergence of a vector valued function
is given as follows:
The Dunkl variation of
is defined by
where
and
A
function is said to have the Dunkl-bounded variation on
if
and the set of all of these functions is denoted as
, which is a Banach space. We will prove it in detail later in Lemma 5. If we choose
, then it is the classical BV space.
A function
has locally Dunkl-bounded variation in
if for each open set
,
We use
to denote the space of such functions.
Let
be a bounded open set and
be a Borel set. We can define
In fact,
is a Radon measure in
. Let us prove the general structure theorem.
Lemma 1. (Structure Theorem for functions). Let . Then there exists a Radon measure μ on Ω such thatfor every andwhere is the total variation of the measure μ. Proof. It is easy to see that
Denote by the functional
with
where
Then using the Hahn-Banach theorem, we know that there exists a linear and continuous extension
L of
to the normed space
such that
By the Riesz representation theorem (cf. Corollary 1.55 in [
7]), there exists a unique
-valued finite Radon measure
such that
and
. Thus, we have
This completes the proof. □
As we know,
implies that each Sobolev function has bounded variation. We similarly obtain the relation in the Dunkl setting. The Dunkl Sobolev space
is the space consisting of all functions
,
and
in a weak sense. The norm of
is defined as
Lemma 2. (Local inclusion of Sobolev functions). If is an open set, then Proof. Suppose
open and let
with
Then using (
2), we have
Taking the supremum over
, we can derive the proof of the lemma. □
Next we show that for every function, the Dunkl-bounded variation boils down to the usual local Dunkl norm.
Lemma 3. ( norm on ). If , thenFurthermore, if then . Proof. If
, we know that
. For each
with
we get
By taking the supremum over
, we prove
and
Then we prove the reverse one is valid. Now define
by setting
It is easy to see that
By a standard approximation result, there exists a sequence
such that
pointwise as
, with
for all
. Considering the definition of
, after integration by parts, and then for each
we have
where
. By the dominated convergence theorem and the definition of
g, when
we obtain
so we complete the proof of the first statement.
If
, fix a compact set
with nonempty interior and define
Similarly to the previous arguments, we can find a sequence
so that
pointwise with
for all
. Thus, we get
Since
, then
. Consequently,
where we have used the dominated convergence theorem. Finally, the proof is completed using an exhaustive sequence of compacts via monotone convergence. □
Lemma 4. (Lower semicontinuity of Dunkl variation). Suppose and in Then Proof. Fix
with
. Firstly, we use the definition of
to get
Since
converges to
f in
, then via the dominated convergence theorem, we can obtain
Then, according the arbitrariness of such functions
and the definition of
, we can get the conclusion. □
Lemma 5. The space is a Banach space.
Proof. It is easy to check that
is a norm and we omit the details. In what follows, we prove the completeness of
. Let
be a Cauchy sequence, namely, for every
there exists
such that
, we have
Especially,
is a Cauchy sequence in the Banach space
, which implies that there exists
with
as
Hence, via Lemma 4, we have
which implies that
as
. This completes the proof. □
Next we will consider the approximation by smooth functions for functions of Dunkl-bounded variation.
Lemma 6. (Approximation with smooth functions). Assume , there exists a sequence of functions such that
- (i)
in ;
- (ii)
as .
Proof. We adapt the method of the proof in Theorem 5.3 in [
1] to prove this theorem. Firstly, via the lower semicontinuity property of
functions established in Lemma 4, we only need to prove that, for
and every
there exists a function
such that
Define a sequence of open sets, for
,
where
and
is an open ball of center 0 and radius
,
denotes the Euclidean distance from
x to
. Since
is a Radon measure, for
, we can choose
so large that
Note that the sequence of open domains
satisfies the following ways:
We consider another sequence of open sets
By standard results, there exists a partition of unity related to the covering
, which shows that there exists
such that
for every
and
on
. Namely, we have the following fact:
Let
be a radial nonnegative function with
and
. Given
and
, extended to zero out of
, we define the following regularization:
We can easily conclude that for each
there exists
such that
Now define
. In some neighborhood of each point
and the sum is locally finite, then we get that
and
. (
6) implies
Consequently,
Moreover, the lower semicontinuity of Dunkl variation in Lemma 4 implies that
Suppose
. We start a direct computation and get
where we have used (
5) in the fourth step and the third equal sign above holds due to the fact that if
, then
In fact, for
,
where we have used the fact
and the
G-invariance:
We note that
. Recalling that by construction every point
belongs to at most three of the sets
, then we have
where the last inequality is given by using (
4). On the other hand, we use (
6) to obtain
Then
which implies that
Now we use the above estimate and (
7) to complete the proof. □
Lemma 7. (max–min property of the Dunkl variation). If . Then Proof. Firstly, we can suppose that
otherwise, the conclusion is clearly true when
. Choose functions
such that
As
it follows that
□
Lemma 8. (Compactness for ). Suppose is an open and bounded domain with Lipschitz boundary. Let be a sequence in satisfyingThen there exists a subsequence and a function such that in when m . Proof. According to approximation with smooth functions, for
, there is a sequence
such that
In particular,
Now, we get that
and
Therefore,
is a bounded sequence in
. Since
has smooth boundary, it follows from Rellich’s compact embedding theorem that there exists
and a subsequence
such that
in
. Then from (
8), we know
in
. By lower semicontinuity of Dunkl variation, we obtain
which shows that
and this completes the proof. □
Naturally, we can give the perimeter of a set in the Dunkl setting.
Definition 2. The Dunkl perimeter of is defined as: Remark 1. It is easy to see that when , the Dunkl perimeter is reduced to the classical perimeter of , that is The following corollary can be easily obtained by replacing f in Lemma 4 with .
Corollary 1. (Lower semicontinuity of Dunkl perimeter). Suppose , then From the definition of the Dunkl perimeter and plus the above max–min inequality, we have the following lemma.
We give the following notation. For
and
, define
Lemma 10. If the mappingis Lebesgue measurable for Proof. Let
and
Firstly, we prove the claim:
Suppose
, we have
for
, and we obtain
Similarly, if
and we get
For the general case, write
, therefore, we conclude that for all
,
Thus,
□
The isoperimetric inequality for the
function is valid and it is proved in [
28].
Proposition 1. Let E be a bounded Lipschitz set of finite Dunkl perimeter on . Thenwith the sharp constant , where ϵ is any element of E such that is a Weyl chamber, and Next we show that the Gauss–Green formula is valid on sets of locally finite Dunkl perimeter.
Theorem 2. (Gauss–Green formula). Let E have locally finite perimeter. Assume E is G-invariant, that is . Thenfor , where and the unit vector is the outward normal to E. Proof. Through calculation, we obtain
where we have used the facts:
,
and the following facts for the derivatives of
:
□
4. Heat Semigroups Characterization of Dunkl Bounded Variation Functions
At first, we recall the Dunkl heat kernel and collect its properties established by R
sler (cf. [
24,
26]). One important function is the Dunkl kernel
associated with Dunkl operators. For generic multiplicities
k and each
, the system
has a unique solution on
, which generalizes the exponential functions
. For functions of non-negative multiplicity, the commutative algebra of Dunkl operators and the algebra of general partial differential operators are intertwined by a unique linear homogeneous isomorphism over polynomials. In other words, there exists a unique intertwining operator
such that
which also justifies the next formula
Note that the Dunkl kernel
is an explicit “closed” form, where it is known so far only in some particular cases. In dimension 1, the Dunkl kernel can be written as a sum of two Bessel functions (cf. [
24])
The generalized Bessel function is written as
Here
is the order of the group
G. In dimension 1,
Let us collect the main properties of Dunkl kernel in the following proposition (cf. [
24,
26]).
Proposition 2. (Properties of the Dunkl kernel).
;
;
;
;
;
Specially,
In [
26], a Dunkl transform is given by
The Dunkl heat equation
can be obtained via the Dunkl transform (
13) (see Theorem 3.12 in [
26] for more details) under suitable cases, where
and the Dunkl heat kernel is defined by
Here
is Macdonald-Mehta integral associated with the root system
R and it is represented as
Next, we collect a series of basic properties of the Dunkl heat kernel (cf. [
24,
26]).
Proposition 3. (Basic properties of the Dunkl heat kernel).
is an analytic function in ;
;
and ;
;
Lemma 11. The Dunkl semigroup is symmetric in .
Proof. We obtain
which completes the proof. □
Lemma 12. For every , we have Proof. Firstly, define
It’s clear from the calculation that
When
according to the fact: the function
is
invariant, the Formulas (
3) and (
12), we have
where denotes the Dunkl directional derivative of the variable Similarly,Thus, we get
Following the definition of
and (
14), we obtain
Letting
via the dominated convergence theorem we prove that
□
Lemma 13. The Dunkl semigroup satisfies the following properties:
- (i)
is continuous from to
- (ii)
where denotes the Dunkl gradient of the variable y,
- (iii)
Proof. The property (i) is obviously available. Next we prove (ii). Firstly, since the function
is G-invariant, we get
By the same calculation, we have
So we obtain
Next via the definition of
and the property of the Dunkl gradient we have
For (iii), it is easy to see that
and then we take the infinite norm on both sides to get
□
Theorem 6. Denote by the space of vector-valued functions with continuous partial derivatives and bounded Dunkl divergence. For each , we have Proof. Firstly, we can easily get
Next we mainly prove the opposite inequality, let
be a sequence of functions such that
- (i)
is G-invariant and for all and ;
- (ii)
there exists such that on for every compact set , ;
- (iii)
as .
If
, we have
, and
Therefore, if
and
, then we get the following inequality by the dominated convergence theorem,
Thus, the proof of Theorem 6 is completed. □
Example 1. We give an example for in the proof of Theorem 6 as follows: Note that the example we constructed implies that exists for and it satisfies the above conditions (i)–(iii) in Theorem 6. It is easy to see that (i) and (ii) hold true. Since is G-invariant, we have Thus, as , which satisfies (iii).
Theorem 7. For any , we have Proof. At first, for every functions
and
, we note that
Via the definition of
, Lemmas 12 and 13, we have
Then taking the supremum over
, we get that
Next, we prove the opposite inequality
Let
be a vector function in
such that
. Denote by
. We next explain that
where we have used the fact that
and Lemma 13 (ii). Therefore, we get
where we have used the property that Dunkl semigroup
is symmetric in
and Theorem 6. Taking the supremum over all
and
yield the desired inequality. Finally, letting
, we complete the proof. □