1. Introduction
The history of continuous-time martingale theory dates back to the 1930s. Lévy [
1] investigated the properties and convergence conditions of integrals composed of independent random variables. This work laid a foundation for the research on continuous-time martingale theory. Later, Ville [
2] introduced the term martingale and extended its definition to continuous-time martingales. Doob [
3] systematically explored continuous-time martingales and their connections to subharmonic functions. Burkholder [
4] showed the connection between the
spaces and harmonic functions. For a local continuous martingale
X, Weisz [
5] considered the maximal function
, the quadratic variation
and the conditional quadratic variation
. He established corresponding atomic decomposition theorems, key martingale inequalities and duality theorems. Afterward, Weisz [
6] researched a new decomposition theorem for local martingales and studied the interpolation spaces between Hardy and
spaces and between martingale Hardy spaces. The
spaces
and
were introduced, and their relationships with
,
and Hardy spaces were investigated in [
7]. Hong et al. [
8] studied asymmetric Doob inequalities in continuous time. Lu and Peng [
9] introduced several martingale Orlicz–Hardy spaces and established some martingale inequalities. We refer the readers to [
10,
11,
12,
13,
14,
15] for additional contributions to the theory of continuous martingales.
In this paper, we investigate martingale operators as studied by Burkholder and Gundy in [
16]. They introduced quasi-operators and proved some integral inequalities in discrete time. Weisz [
17] further explored the martingale Hardy space generated by an operator
T and established a corresponding atomic decomposition theorem, interpolation theorems, some martingale inequalities and dual theorems based on the work of [
16]. The operator
T maps the set of martingales into the set of non-negative
-measurable functions. In particular, he pointed out that in discrete time, the maximal function
M, the
p-variation operator
and the conditional
p-variation operator
belong to the class of the preceding operators. Subsequently, Fan [
18] proved some inequalities concerning
and
. Hou and Ren [
19] established some inequalities about sublinear operators in weak Hardy spaces. Liu [
20] introduced various types of Hardy martingale spaces and Lorentz Hardy martingale spaces and their corresponding results. Bakas et al. [
21] investigated continuous bilinear decompositions of the multiplication between elements in martingale Hardy spaces and their dual spaces. Besides, for a local continuous martingale
X, Pisier and Xu [
22] studied the
p-variation
and derived some important inequalities associated with
. Peter and Victoir [
23] further established quantitative bounds of the
p-variation norm in the form of a
inequality. Motivated by these works, we naturally consider the general form of the
T operator in continuous time. Here, the operators
M,
,
and
also belong to our class of operators. We focus on the Hardy and
spaces generated by an arbitrary operator
T. More precisely, our results generalize the related results in [
5,
6,
7]. Several new results are proved, and many known results are generalized for arbitrary operators
T in continuous time.
This paper is organized as follows.
Section 2 provides essential definitions and preliminary concepts. In
Section 3, we provide the definition of the atoms and establish the atomic decomposition theorem for the
space generated by a predictable operator
T. If
in Theorem 1, our result can lead back to Theorem 1 in [
5].
Section 4 introduces the sharp operator
of an operator
T and considers the
spaces in continuous time. We study the
space, which is defined by the
-norm of
and prove that all
spaces are equivalent for
(see Theorem 2). Additionally, we demonstrate the equivalence between the
-norm of
and the
-norm of
X for
(see Theorem 3).
In
Section 5, we investigate the interpolation spaces between the Hardy and
spaces in continuous time by the real method. To be precise, we prove that if
T is predictable,
,
and
, then
If
T is predictable,
,
and
, then
In particular, setting
in (
1) and (
2), the above statements can reduce to Theorem 4 and Colollary 2 in [
6].
In the last section, with the help of atomic decomposition and interpolation theorems, we prove that if a continuous-time martingale inequality holds for a parameter p, then it also holds for all parameters less than p.
Before we proceed further, let us fix some notations. We use , and to denote the set of nonnegative integers, the set of integers and the set of positive real numbers, respectively. The letter C denotes a positive constant, which may vary from line to line. The symbol stands for the inequality . If we write , then it means . We denote by the indicator function of a measurable set I.
2. Preliminaries and Notations
Let
be a complete probability space endowed with a right-continuous filtration
satisfying the usual conditions, where each
is a sub-
-algebra of
and
for
. Define
with
. For
, we define
where
. Throughout, we assume
for all
.
A stochastic process is a mapping X: such that for every , the function is -measurable. A stochastic process X is adapted if is -measurable for all . X is called a regular process if X is adapted and all the functions have a left and a right limit for every . Denote left and right limits by and , respectively. The process is right-continuous (resp. left-continuous) if (resp. ) for all .
The optional -algebra is defined as the -algebra on generated by real-valued, regular and right-continuous processes. A stochastic process X is optional if it is -measurable. Similarly, the predictable -algebra is generated by adapted continuous real processes. A process X is predictable if it is -measurable.
A map
is said to be a stopping time with respect to
if
for all
. The set of stopping times is denoted by
. According to [
24],
is a stopping time if and only if
is predictable or, equivalently, if
for all
t. The graph of
is defined as
A stopping time
is predictable if
and the set of all predictable stopping times is denoted by
.
For any stopping time
, the
-algebras
and
are defined as
and
For a constant stopping time
, we have
and
.
Denote the expectation operator by , the conditional expectation operator with respect to , , and by , , and , respectively. We assume for all .
A stochastic process
X is called a martingale if
X is adapted, satisfies
for all
and
whenever
. For simplicity, we assume
for all martingales
X. Denote by
the set of all martingales.
denotes the set of all
-bounded martingales, i.e., those for which
In martingale theory, it is a classical result that if the martingale
X is uniformly integrable (resp.
with
), then there exists
such that
a.e. and in
(resp.
) as
. Consequently, the map
establishes an isomorphism between the space of uniformly integrable martingales, which is a subspace of
and
, and additionally between
and
for
(see [
25]).
Let
X be a real stochastic process and
be a stopping time. The stopped process
is defined as
An adapted process X is said to be a local martingale when there exists a non-decreasing sequence of stopping times satisfying a.e. and such that each stopped process is a uniformly integrable martingale. is said to be a fundamental sequence of X. Indeed, every martingale is a local martingale (take ), but the converse is not valid. Let denote the set of all local martingales. A local martingale X is called locally -bounded if there exists a non-decreasing sequence of stopping times with a.e. such that each stopped process is an -bounded martingale. Let denote the set of such local martingales.
A stochastic process
X is of class (D) if the set
is uniformly integrable. From [
25], local martingales of class (D) are martingales. We further define,
For
, there exists a unique predictable, right-continuous and increasing process
, called conditional quadratic variation (or sharp bracket) such that
is a local martingale (see [
25]). Similarly, if
X is a local martingale, then there exists a unique right-continuous and increasing process
, called quadratic variation (or square bracket) such that
is a local martingale and
(
).
A martingale operator T is defined as an operator that maps the set of martingales to the set of non-negative -measurable functions. Throughout this paper, we assume that T satisfies the following conditions:
- (B1)
Subadditivity: For any sequence of martingales
and
(in the sense of
a.e. for all
), we have
- (B2)
Homogeneity: for .
- (B3)
Locality: on .
- (B4)
Symmetry: .
-
Define and with . Then,
- (1)
;
- (2)
on .
Furthermore, for a finite stopping time
, if
on
, then
. Notably, the operator
also satisfies (B1)–(B4). An operator
T is adapted (resp. predictable) if
is
-measurable (resp.
-measurable). If
, then
and
. It is easy to see that the operators such as
M,
,
and
satisfy (B1)–(B4), where
denotes the
p-variation of
X, i.e.,
Now, we introduce the operator
. Let
be a non-negative, non-decreasing and predictable sequence of functions for which
Define
Clearly,
satisfies (B1)–(B4), and
is predictable and non-decreasing in
t.
Now, for
, the martingale Hardy space
generated by
T is defined as
Especially, if
,
and
in the definition above, respectively, we get the martingale Hardy spaces associated with the maximal function
M, the conditional quadratic variation
and the quadratic variation
, respectively. For the definitions of these martingale Hardy spaces, we refer the reader to [
5].
4. BMO Spaces and Sharp Operators
In this section, we introduce
spaces associated with operator
T satisfying (B1)–(B4), which are essential for establishing our main theorems. The classical
spaces (see [
7]) contain all martingales
for which
These spaces satisfy the following inequalities:
Weisz [
7] proved that all
spaces
are equivalent and
For an operator
T satisfying (B1)–(B4), we define the sharp operator
as
Building on these foundations, we define the
space generated by
T as
In order to better characterize the
space, we assume that
T satisfies the following condition
where
are martingales. Obviously,
, the quadratic variation
and the conditional quadratic variation
also satisfy (
8) (see [
7] (pp. 70–72)).
To obtain the main conclusion of this section, we give the following lemmas.
Lemma 1 ([
5]).
Let and A be a non-negative, non-decreasing process. If A is predictable with and satisfiesthen Lemma 2. Let be a local martingale, T be a predictable operator satisfying – and (
8)
. Suppose that and there exists a function γ such thatThen, for any , Proof. Define the stopping times:
On the set
, by the monotonicity of
, we get
Thus,
In virtue of (B1), we obtain
Next, we show that for any predictable stopping time
W,
Let
be a sequence of predictable stopping times with
on
. By Fatou’s lemma and (
8), we get
From (
9), for any predictable stopping time
U, we have
Assume that
Obviously,
is an elementary stopping time,
and
. As a consequence of Fatou’s lemma, (
8) and
, we derive
Applying this to
, we obtain (
11). Finally, since
S is predictable, and
is
-measurable (see [
24,
25]), from (
10) and (
11), we obtain
The proof is complete. □
Theorem 2. Let T be a predictable operator satisfying – and (
8)
. For any , the spaces generated by T are all equivalent. Proof. Let
and denote
. Define
. By conditions (B1)–(B4), we know that
is a non-decreasing, predictable process (see [
26]). Let
be fixed and
. By
and (B1), we get that
From (
12) and (B3), we conclude that
Thus, by Lemma 1, we obtain
Consider the martingale
, substituting
into (
13), we have
The case
follows from standard techniques in Garsia [
26]. The proof is finished. □
Theorem 3. Let be a local martingale, T be a predictable operator satisfying – and (
8)
. Then, Proof. Set
. Lemma 2 yields that
Multiplying (
14) by
and integrating over
, we get
By Hölder’s inequality, we obtain
Let
. Therefore,
On the other hand, estimating
by
Applying Doob’s inequality, we have
This completes the proof. □
5. Interpolation of Martingale Spaces
In this section, we explore the interpolation spaces between
spaces and martingale Hardy spaces generated by
. Fundamental definitions of interpolation theory are introduced below; we refer to Weisz [
6] for further details.
Definition 2. For a measurable function f, the non-increasing rearrangement function of f is defined as Definition 3. Let and . The Lorentz space consists of those measurable functions f with , where Definition 4. The martingale Hardy–Lorentz space (, ) consists of all local martingales X satisfying
If
in Definition 4, we obtain martingale Hardy spaces
. Moreover, taking
in the definition above, we obtain the martingale Hardy–Lorentz space associated with the conditional quadratic variation
.
Let
be a compatible couple of quasi-normed spaces continuously embedded in a topological vector space
A. The following definitions are standard in interpolation theory (see [
6,
27]).
Definition 5. For , the K-functional is defined by Definition 6. For and , the real interpolation space is defined asBy convention,
and
. From [
17], the
K-functional for
satisfies
Let
be another compatible couple in a topological vector space
B. A map
is called quasi-linear if for any
and decomposition
, there exists
such that
The following well-known lemmas are useful to prove our main results in this section.
Lemma 3 ([
28]).
Let , and . If and , thenIf and are complete and , thenwhere Lemma 4 ([
28]).
Let , and . Then,In particular,Furthermore, for , Lemma 5 ([
17,
29]).
Suppose has the lattice property, i.e., a.e. implies . If U is a subadditive operator bounded from to with norms , then for and , we have Next, a new decomposition theorem for the interpolation spaces generated by an operator T is given.
Theorem 4. Let T be a predictable operator satisfying –, , and be a local martingale. Then, X can be decomposed as , where R and Q satisfyand Proof. Suppose that
. Let
satisfy
. For each
, define the same predictable stopping time as in Theorem 1:
From Theorem 1,
X has the following decomposition
where
are
-atoms and
. Define
Evidently,
for all
and
. By the definition of the stopping time
, we have
Thus, (
15) holds. For
Q, by Theorem 1, we obtain that
Applying the Abel rearrangement, we get that
which means (
16) holds. This completes the proof. □
Now, we characterize the interpolation spaces between the martingale Hardy spaces generated by the predictable operator T.
Theorem 5. Let T be a predictable operator satisfying –. For , and , we have In order to prove this theorem, we first need to present two lemmas as follows.
Lemma 6. Let T be a predictable operator satisfying –. If , thenwhere . Proof. We pick
such that
, where
for fixed
. Let
and
denote the martingales from Theorem 4 associated with
y. We apply the definition of
K to obtain
In view of Theorem 4, we deduce
Moreover,
which completes the proof. □
Lemma 7 ([
29]).
For a non-increasing function and , we have Proof of Theorem 5. We deduce from Lemma 6 and the definition of real interpolation that
Applying Lemma 7, we get that
Clearly,
and
are bounded. Therefore, by Theorem 4 and Lemma 5, we get
is bounded. Therefore,
which finishes the proof of Theorem 5 for
. By appropriately modifying the previous proof, the theorem can also be proven in the case of
. □
Extending the application of Lemma 3 and Theorem 5, we establish the following result.
Corollary 1. Let T be a predictable operator satisfying –. If , and , then We proceed to characterize the interpolation spaces between and , where denotes one of the spaces .
Theorem 6. Let T be a predictable operator satisfying – and (
8)
. If , and , then Proof. It is straightforward to verify that
Corollary 1 and (
17) show that
To prove the converse part, fix
for the operator
. Theorem 3 establishes the boundedness:
. Moreover, the operator
is also bounded. Applying Lemmas 4 and 5 with
, we obtain
By Theorem 3, we get
which establishes the result for
, namely,
Utilizing Lemma 3, the result could be proven by a standard argument (see [
30]). This completes the proof. □
Applying Lemma 3 and Theorem 6, we establish the following interpolation result between the martingale Hardy–Lorentz spaces and spaces generated by T.
Corollary 2. Let T be a predictable operator satisfying – and (
8)
. If , and , then Setting in Corollaries 1 and 2, the following statements hold.
Corollary 3. If , and , then Corollary 4. If , and , then