Skip to main content

BMO-Lorentz martingale spaces

Abstract

In this paper BMO-Lorentz martingale spaces are investigated. We give the characterization of BMO-Lorentz martingale spaces. Moreover, we discuss the relationship between the Carleson measure and BMO-Lorentz martingales. As a consequence, we find a new way to characterize the geometrical properties of a Banach space.

1 Introduction and preliminaries

Since 1951 when they were first introduced by Lorentz in [1], Lorentz spaces have attracted more and more attention. A lot of results were obtained such as normability, duality, interpolation, and so on [2–7].

We know that martingale theory is intimately related to harmonic analysis. In martingale case, Weisz [8] and Long [9] considered the spaces H p , q and the interpolations between them, respectively. It is well known that the validity of a classical (scalar-valued) result in the vector-valued setting, i.e., for functions or martingales with values in a Banach space X, depends on the geometrical or topological properties of X.

It was exactly with this in mind that Xu [10] developed the vector-valued Littlewood-Paley theory, which was inspired by Pisier’s celebrated work [8] on martingale inequalities in uniformly convex spaces. Very recently, Ouyang and Xu [11] studied the endpoint case of the main results of [10] by means of the classical relationship between BMO functions and Carleson measures. Jiao [12] discussed the relationship between Carleson measures and vector-valued martingales.

Let (Ω,μ) be a nonatomic σ-finite measure space. Suppose that f is a measurable function on a measure space (Ω,μ). We define its distribution function

λ f (t)=μ { ω : ∥ f ( ω ) ∥ > t } ,t≥0,

and its decreasing rearrangement function

f ∗ (t)=inf { s > 0 : λ f ( s ) ≤ t } .

Given a measurable function f on a measure space (Ω,μ) and 0<p,q≤∞, define

∥ f ∥ L p , q ={ ( ∫ 0 ∞ ( t 1 p f ∗ ( t ) ) q d t t ) 1 q if  q < ∞ , sup t > 0 t 1 p f ∗ ( t ) if  q = ∞ .

Remark 1.1 Observe that for all 0<p,r<∞ and 0<q≤∞ we have

∥ | g | r ∥ L p , q = ∥ g ∥ L p r , q r r .
(1.1)

Unfortunately, the functions ∥ ⋅ ∥ L p , q do not satisfy the triangle inequality. However, since for all t>0, ( f + g ) ∗ (t)≤ f ∗ (t/2)+ g ∗ (t/2), we have

∥ f + g ∥ L p , q ≤ c p , q ( ∥ f ∥ L p , q + ∥ g ∥ L p , q ) ,
(1.2)

where c p , q = 2 1 / p max{1, 2 ( 1 − q ) / q }.

The set of all f with ∥ f ∥ L p , q <∞ is denoted by L p , q (X,μ) and is called the Lorentz space with indices p and q. Observe that the definition implies that L ∞ , ∞ = L ∞ .

Let (Ω,Σ,P) be a complete probability space, and let ( Σ n ) n ≥ 0 be a nondecreasing sequence of sub-σ-algebras of Σ with Σ=σ( ⋃ n ≥ 0 Σ n ). We denote by E and E n the expectation and conditional expectation with respect to Σ and Σ n , respectively. For a martingale f= ( f n ) n ≥ 0 with martingale difference d f n = f n − f n − 1 , n≥0, f − 1 ≡0, we define its maximal function, p-square function, respectively, as usual:

Mf= sup n ∥ f n ∥, S ( p ) (f)= ( ∑ n = 1 ∞ ∥ d f n ∥ p ) 1 / p .

We say that an X-valued martingale f= ( f n ) n ≥ 0 ∈ L p , q (X) if sup n ∥ f n ∥ L p , q <∞.

The space BMO L p , q ( X ) a (a≥1, 1<p,q≤∞) consists of all martingale f∈ L p , q (X) such that

∥ f ∥ BMO L p , q ( X ) a = sup n ∥ ( E ( ∥ f − f n − 1 ∥ a | Σ n ) ) 1 / a ∥ L p , q ( X ) <∞.

Remark 1.2 The spaces BMO L p , q ( X ) a are independent of a and all corresponding norms are equivalent. This allows us to denote any of them by BMO L p , q ( X ) .

Proof If a≥1, φ(x)= x a is a convex function, by Jensen’s inequality, we have E(∥f− f n − 1 ∥| Σ n )≤ ( E ( ∥ f − f n − 1 ∥ a | Σ n ) ) 1 / a , which implies E ( ∥ f − f n − 1 ∥ | Σ n ) ∗ (t)≤ ( ( E ( ∥ f − f n − 1 ∥ a | Σ n ) ) 1 / a ) ∗ (t), i.e., BMO L p , q ( X ) a ⊂ BMO L p , q ( X ) 1 .

On the contrary, let g n =E(∥f− f n − 1 ∥| Σ n ), Mg= sup n ∥ g n ∥, h n =E ( ∥ f − f n − 1 ∥ a | Σ n ) 1 / a , Mh= sup n ∥ h n ∥. Now we set A t ={ω:Mh>t}. Then Mg>t a.e. on A t . (Factually, if there is a B⊂ A t with μ(B)>0 such that Mg≤t on B, then E(∥f− f n − 1 ∥| Σ n )≤t=E(t| Σ n ) a.e. on B, which implies ∥f− f n − 1 ∥≤t a.e. on B. This is a contradiction for A t .) So, P{ω:Mh>t}≤cP{ω:Mg>t}. Then

∥ f ∥ BMO L p , q ( X ) a ≤ ( q ∫ 0 ∞ [ t P ( M h ( ω ) > t ) 1 / p ] q d t t ) 1 / q ≤ c ( q ∫ 0 ∞ [ t P ( M g ( ω ) > t ) 1 / p ] q d t t ) 1 / q ≤ c ∥ f ∥ BMO L p , q ( X ) 1 ,

i.e., BMO L p , q ( X ) 1 ⊂ BMO L p , q ( X ) a . Thus we complete the proof. □

Remark 1.3 If p=q=∞, BMO L p , q ( X ) is the classical BMO space.

Remark 1.4 For the classical BMO space, we have the following statement (see [12]):

∥ f ∥ BMO ∼ sup τ μ ( τ < ∞ ) − 1 / p ∥ f − f τ − 1 ∥ L p ,1≤p<∞.
(1.3)

It is well known that L p , q is a subspace of L p , r for 0<p≤∞, 0<q<r≤∞ and L p 2 , q 2 is a subspace of L p 1 , q 1 for 1< p 1 ≤ p 2 ≤∞, 1< q 1 , q 2 ≤∞. Thus we have the following proposition.

Proposition 1.5 (1) If 0<p≤∞, 0<q<r≤∞, BMO L p , q ⊆ BMO L p , r .

  1. (2)

    If 1< p 1 ≤ p 2 ≤∞, 1< q 1 , q 2 ≤∞, BMO L p 2 , q 2 ⊆ BMO L p 1 , q 1 .

Theorem 1.6 Let f= ( f n ) n ≥ 0 be an X-valued martingale in L p , q (X), 1<p,q≤∞. Then f∈ BMO L p , q ( X ) if and only if there exists an adapted process θ= ( θ ) n ≥ 0 such that

C θ = sup n ∥ E ( ∥ f − θ n − 1 ∥ | Σ n ) ∥ L p , q ( X ) <∞.

And, in any case, we have

⦀f ⦀ BMO L p , q ( X ) := inf θ C θ ≤ ∥ f ∥ BMO L p , q ( X ) ≤c⦀f ⦀ BMO L p , q ( X ) .

Proof Assume f∈ BMO L p , q ( X ) . Then, obviously,

⦀f ⦀ BMO L p , q ( X ) ≤ ∥ f ∥ BMO L p , q ( X ) .

Now let ⦀f ⦀ BMO L p , q ( X ) <∞ and θ= ( θ ) n ≥ 0 be any one such that C θ <∞. Then we have

E ( ∥ f − f n − 1 ∥ | Σ n ) ≤ E ( ∥ f − θ n − 1 ∥ | Σ n ) + ∥ θ n − 1 − f n − 1 ∥ = E ( ∥ f − θ n − 1 ∥ | Σ n ) + ∥ E ( ( f − θ n − 1 ) | Σ n − 1 ) ∥ ≤ E ( ∥ f − θ n − 1 ∥ | Σ n ) + E ( ∥ f − θ n − 1 ∥ | Σ n − 1 ) = E ( ∥ f − θ n − 1 ∥ | Σ n ) + E ( E ( ∥ f − θ n − 1 ∥ | Σ n ) | Σ n − 1 ) .

Taking ‘inf’ over all possible θ and (1.2), we get the desired inequality. □

Theorem 1.7 Let f= ( f n ) n ≥ 0 be an X-valued martingale in BMO L p , q ( X ) , where 1 p + 1 s 1 = 1 r and 1 q + 1 s 2 = 1 s , 1<p,q,r, s 1 , s 2 ≤∞. Then

∥ f ∥ BMO L p , q ( X ) ∼ sup τ μ ( { τ < ∞ } ) − 1 s 1 ∥ f − f τ − 1 ∥ L r , s ( X ) ,

where ‘sup’ is taken over all stopping times τ.

Proof Assume that ∥ f ∥ BMO L p , q ( X ) <∞, τ is any stopping time. Then, by Hölder’s inequality, we have

∥ f − f τ − 1 ∥ L r , s ( X ) = ∥ ( f − f τ − 1 ) χ { τ < ∞ } ∥ L r , s ( X ) ≤ c ∥ f − f τ − 1 ∥ L p , q ( X ) ∥ χ { τ < ∞ } ∥ L s 1 , s 2 ( X ) ≤ c sup ∥ g ∥ ( L p , q ) ∗ ≤ 1 | ∫ { τ < ∞ } ( ∥ f − f τ − 1 ∥ ) g d μ | ⋅ ∥ χ { τ < ∞ } ∥ L s 1 , s 2 ( X ) = c ( q p ) 1 / q sup ∥ g ∥ ( L p , q ) ∗ ≤ 1 | ∫ { τ < ∞ } E ( ∥ f − f τ − 1 ∥ ⋅ g | Σ τ ) d μ | ⋅ μ ( τ < ∞ ) 1 s 1 ≤ c ( q p ) 1 / q ∥ f ∥ BMO L p , q ⋅ μ ( τ < ∞ ) 1 s 1 .

This proves one half of the assertion. Conversely, assume that β= sup τ μ ( { τ < ∞ } ) − 1 s 1 ∥ f − f τ − 1 ∥ L r , s ( X ) <∞, and τ is any stopping time, F∈ Σ τ , F⊂{τ<∞}. Define

Ï„ F ={ Ï„ if  ω ∈ F , ∞ if  ω ∉ F .

Thus we get

β ≥ μ ( { τ < ∞ } ) − 1 s 1 ∥ f − f τ − 1 ∥ L r , s ( X ) = 1 μ ( F ) 1 / s 1 ∥ f − f τ F − 1 ∥ L r , s ( X ) ≥ 1 μ ( F ) 1 / s 1 ∥ f − f τ F − 1 ∥ L r ∧ 1 , r ∧ 1 ( X ) ≥ 1 μ ( F ) ∥ f − f τ F − 1 ∥ L 1 , 1 ( X ) = 1 μ ( F ) ∫ F ∥ f − f τ F − 1 ∥ d u .

That is, E(∥f− f τ F − 1 ∥| Σ τ F )≤β. By Remark 1.2 we have

∥ f ∥ BMO L p , q ( X ) ≤cβ.

Thus we complete the proof of the theorem. □

Proposition 1.8 Particularly, if s=r and p=q=∞, we get Remark 1.4.

By Theorem 1.6 and Theorem 1.7, we have the following proposition.

Proposition 1.9 Let f= ( f n ) n ≥ 0 be an X-valued martingale in BMO L p , q ( X ) , where 1 p + 1 s 1 = 1 r and 1 q + 1 s 2 = 1 s , 1<p,q,r, s 1 , s 2 ≤∞. Then

∥ f ∥ BMO L p , q ( X ) ∼ inf θ sup τ μ ( { τ < ∞ } ) − 1 s 1 ∥ f − θ τ − 1 ∥ L r , s ( X ) ,

where ‘sup’ is taken over all stopping times τ and ‘inf’ is taken over all adapted process θ= ( θ ) n ≥ 0 .

2 Carleson measure and BMO-Lorentz martingale spaces

Definition 2.1 Let ν be a nonnegative measure on Ω×N, where N is equipped with the counting measure dm. Let τ ˆ denote the tent over τ:

τ ˆ = { ( ω , k ) : k ≥ τ ( ω ) , τ ( ω ) < ∞ } .

ν is said to be an s-Carleson measure if

⦀ν⦀= sup τ ν ( τ ˆ ) μ ( { τ < ∞ } ) s <∞,

where Ï„ runs through all stopping times.

Theorem 2.2 Let 1<p,q≤∞, and g= ( g n ) n ≥ 0 be a real-valued martingale and dν= | Δ k g | 2 δ k dμ, where δ k is the Dirac measure centered at k. So, if g∈ BMO L 2 p , 2 q 2 , ν is a 1/ p ′ -Carleson measure. Moreover, if 1<p,q<∞, the converse is also true, where 1 p ′ + 1 p =1, 1 q ′ + 1 q =1.

Proof Let g= ( g n ) n ≥ 0 be a real-valued martingale, let ν be generated by g as above, and let τ be any stopping time. Then, for 1<q<∞,

ν ( τ ˆ ) = E ( ∑ k = 0 ∞ | Δ k g | 2 χ { τ ( ω ) ≤ k } ) = E ( E ( ∑ k = τ ( ω ) ∞ | Δ k g | 2 | Σ τ ) χ { τ ( ω ) ≤ k } ) = E ( E ( | g − g τ − 1 | 2 | Σ τ ) χ { τ ( ω ) ≤ k } ) ≤ c ∥ E ( | g − g τ − 1 | 2 | Σ τ ) ∥ L p , q ⋅ ∥ χ { τ ( ω ) ≤ k } ∥ L p ′ , q ′ = c ∥ E ( | g − g τ − 1 | 2 | Σ τ ) 1 2 ⋅ 2 ∥ L p , q ⋅ ∥ χ { τ ( ω ) ≤ k } ∥ L p ′ , q ′ ≤ c ∥ E ( | g − g τ − 1 | 2 | Σ τ ) 1 / 2 ∥ L 2 p , 2 q 2 ⋅ μ ( { τ ( ω ) ≤ k } ) 1 / p ′ ≤ c ∥ g ∥ BMO L 2 p , 2 q 2 2 ⋅ μ ( { τ ( ω ) ≤ k } ) 1 / p ′ .

So, g∈ BMO L 2 p , 2 q 2 implies that ν is a 1/ p ′ -Carleson measure and ⦀ν⦀≤ ∥ g ∥ BMO L 2 p , 2 q 2 2 .

Conversely, for any n and any F∈ Σ n , we define

Ï„={ n if  ω ∈ F , ∞ if  ω ∉ F .

Since ν is a 1/ p ′ -Carleson measure, we have

⦀ ν ⦀ ≥ 1 μ ( { τ < ∞ } ) 1 / p ′ ν ( { ( ω , k ) : k ≥ τ ( ω ) , τ ( ω ) < ∞ } ) = 1 μ ( F ) 1 / p ′ ∫ F ∑ k = n ∞ | Δ k g | 2 d μ ≥ 1 μ ( F ) ∫ F ∑ k = n ∞ | Δ k g | 2 d μ = 1 μ ( F ) ∫ F | g − g n − 1 | 2 d μ .

That is, ⦀ν⦀≥E( | g − g n − 1 | 2 | Σ n ). Then we have ∥ g ∥ BMO L 2 p , 2 q 2 2 ≤c⦀ν⦀. Thus, we complete the proof of the theorem. □

3 The characterization of Banach space’s geometrical properties

Let 1<q<∞. Then a Banach space X has an equivalent q-uniformly convex norm if and only if for one 1<p<∞ (or equivalently, for every 1<p<∞) there exists a positive constant c such that

∥ S ( q ) ( f ) ∥ p ≤c sup n ∥ f n ∥ p

for all finite L p -martingales f with values in X. Again, the validity of the converse inequality amounts to saying that X has an equivalent q-uniformly smooth norm.

Definition 3.1 Let X 1 and X 2 be two Banach spaces. Let L( X 1 , X 2 ) denote the space of all bounded linear operators from X 1 to X 2 . Let v= ( v n ) n ≥ 1 be an adapted sequence such that v n ∈ L ∞ (L( X 1 , X 2 )) and sup n ≥ 1 ∥ v n ∥ L ∞ ( L ( X 1 , X 2 ) ) ≤1. Then the martingale transform T associated to v is defined as follows. For any X 1 -valued martingale f= ( f n ) n ≥ 1 ,

( T f ) n = ∑ k = 1 n v k d f k .

We get the following results from [13, 14].

Lemma 3.2 With the assumptions above, the following statements are equivalent:

  1. (1)

    There exists a positive constant c such that

    ∥ T f ∥ BMO ( X 2 ) ≤c ∥ f ∥ BMO ( X 1 ) ,∀f∈BMO( X 1 ).
  2. (2)

    There exists a positive constant c such that

    ∥ ( T f ) ∗ ∥ BMO ( X 2 ) ≤c ∥ f ∥ BMO ( X 1 ) ,∀f∈BMO( X 1 ).
  3. (3)

    For some 1≤p<∞ (or equivalently, for every 1≤p<∞), there exists a positive constant c such that

    ∥ T f ∥ L p ( X ) ≤c ∥ f ∥ L p ( X ) ,∀f∈ L p (X).

Theorem 3.3 Let X be a Banach space, 2≤q<∞, 1<p<∞. Then the following statements are equivalent:

  1. (1)

    There exists a positive constant c such that for any finite X-valued martingale,

    ∥ S ( q ) ( f ) ∥ BMO L p , q ( X ) ≤c ∥ f ∥ BMO L p , q ( X ) .
  2. (2)

    X has an equivalent norm which is q-uniformly convex.

Proof (1)⇒(2) Let 1 p + 1 s 1 = 1 r , 1 q + 1 s 2 = 1 r , 1<p,q,r, s 1 , s 2 ≤∞ By Theorem 1.7 we have

∥ S ( q ) ( f ) ∥ BMO L p , q ( X ) ∼ sup τ μ ( { τ < ∞ } ) − 1 s 1 ∥ S ( q ) ( f ) − S τ − 1 ( q ) ( f ) ∥ L r , r ( X ) ,
(3.1)
∥ f ∥ BMO L p , q ( X ) a ∼ sup τ μ ( { τ < ∞ } ) − 1 s 1 ∥ f − f τ − 1 ∥ L r , r ( X ) .
(3.2)

So, if (1) holds, we have

∥ S ( q ) ( f ) − S τ − 1 ( q ) ( f ) ∥ L r , r ( X ) ≤c ∥ f − f τ − 1 ∥ L r , r ( X ) .

By Remark 1.4 we have

∥ S ( q ) ( f ) ∥ BMO ≤c ∥ f ∥ BMO .
(3.3)

We now consider a martingale transform operator Q from the family of X-valued martingales to that of l q (X)-valued martingales. Let v∈L(X, l q (X)) be the operator defined by v k (x)= { x j } j = 1 ∞ for x∈X, where x j =x if j=k and x j =0 otherwise. Q is the martingale transform associated to the sequence ( v k ):

( Q f ) n = ∑ k = 1 n v k d f k =(d f 1 ,d f 2 ,…,d f n ,0,…).

Then

( Q f ) ∗ = S ( q ) (f).
(3.4)

By (3.3) and (3.4) we have

∥ ( Q f ) ∗ ∥ BMO ≤c ∥ f ∥ BMO .

By Lemma 3.2 we have

∥ S ( q ) ( f ) ∥ q = ∥ ( Q f ) ∗ ∥ q ≤c ∥ f ∥ q .

Thus, by Pisier’s theorem, X has an equivalent q-uniformly convex norm.

(2)⇒(1) Suppose that X has an equivalent q-uniformly convex norm. By Pisier’s theorem [15], we find, for any 1≤n<∞,

E ( ∑ i = n ∞ ∥ d f i ∥ q | Σ n ) ≤cE ( ∥ f − f n − 1 ∥ q | Σ n ) .

Since E( | S ( q ) ( f ) − S n − 1 ( q ) ( f ) | q | Σ n )≤cE(| ( S ( q ) ( f ) ) q − ( S n − 1 ( q ) ( f ) ) q || Σ n ), we have

E ( | S ( q ) ( f ) − S n − 1 ( q ) ( f ) | q | Σ n ) ≤cE ( ∥ f − f n − 1 ∥ q | Σ n ) .

Thus

∥ S ( q ) ( f ) ∥ BMO L p , q ( X ) ≤c ∥ f ∥ BMO L p , q ( X ) .

We complete the proof. □

Theorem 3.4 Let X be a Banach space and 1<p≤2, 1<q<∞. If there exists a positive constant c such that for any finite X-valued martingale

∥ f ∥ BMO L p ( X ) ≤c ∥ S ( p ) ( f ) ∥ BMO L p ( X ) ,
(3.5)

then X has an equivalent norm which is p-uniformly smooth.

On the contrary, if X has an equivalent norm which is p-uniformly smooth, then

∥ f ∥ BMO L p , q ( X ) a ≤c ∥ S ( p ) ( f ) ∥ BMO L p , q ( X ) a

for every martingale f.

Proof Let X ∗ be the dual space of X. It suffices to prove that X ∗ has an equivalent p ′ -uniformly convex norm, where p ′ is the conjugate index of p. By Pisier’s theorem, this is equivalent to showing that

∥ S ( p ′ ) ( g ) ∥ L p ′ ≤c ∥ g ∗ ∥ L p ′ =c ∥ g ∥ H p ′ ∗ ( X ∗ ) .
(3.6)

Recall that H p ′ ∗ ( X ∗ ) is defined by

H p ′ ∗ ( X ∗ ) = { X ∗ -valued martingale  g = ( g n ) : g ∗ ∈ L p ′ } .

It is well known that BMO L p ( X ) can be identified as a subspace of H p ′ ∗ ( X ∗ ). Thus, for any finite martingale, f= ( f n ) n ≥ 0 ∈ BMO L p ( X ) and g= ( g n ) n ≥ 0 ∈ H p ′ ∗ ( X ∗ ).

| 〈 g , f 〉 | = ∫ Ω 〈 g ( ω ) , f ( ω ) 〉 dP≤c ∥ f ∥ BMO L p ( X ) ⋅ ∥ g ∥ H p ′ ∗ ( X ∗ ) .

On the other hand, ∥ S ( p ′ ) ( g ) ∥ L p ′ is the norm of the difference sequence (d g n ) in L p ′ ( l p ′ ( X ∗ )).

Thus

∥ S ( p ′ ) ( g ) ∥ L p ′ = sup ( a k ) { | ∑ 〈 d g k , a k 〉 | : ∥ ( a k ) ∥ L p ( l p ( X ) ) ≤ 1 } = sup ( a k ) { | ∑ 〈 d g k , E k ( a k ) − E k − 1 ( a k ) 〉 | : ∥ ( a k ) ∥ L p ( l p ( X ) ) ≤ 1 } .

Set d f k = E k ( a k )− E k − 1 ( a k ) and f=∑d f k . Then f is an X-valued martingale. We have

∥ S ( p ′ ) ( g ) ∥ L p ′ = sup ( a k ) { | ∑ 〈 d g k , d f k 〉 | } ≤c ∥ f ∥ BMO L p ( X ) ⋅ ∥ g ∥ H p ′ ∗ ( X ∗ ) .
(3.7)

It remains to estimate ∥ f ∥ BMO L p ( X ) . Since ∥ ( a k ) ∥ L p ( l p ( X ) ) ≤1, we can also get the conditional case E(( ∑ k = n ∞ ∥ a k ∥ p )| Σ n )≤1. Then by (3.5)

∥ f ∥ BMO L p ( X ) ≤ c ∥ S ( p ) ( f ) ∥ BMO L p ( X ) ≤ c sup n ∥ E ( ∥ S ( p ) ( f ) − S n − 1 ( p ) ( f ) ∥ p | Σ n ) ∥ L p ≤ c sup n ∥ E ( ∥ S ( p ) ( f ) p − S n − 1 ( p ) ( f ) p ∥ | Σ n ) ∥ L p ≤ c sup n ∥ E ( ∑ k = n ∞ ∥ E k ( a k ) − E k − 1 ( a k ) ∥ p | Σ n ) ∥ L p ≤ c ∥ E ( ( ∑ k = n ∞ ∥ a k ∥ p ) | Σ n ) ∥ L p ≤ c .

On the contrary, we define f ˆ τ =( f ˆ i τ ,), where Σ ˆ i = Σ τ + i , f ˆ i τ = f τ + i − f τ , i≥0. So, we have

( S ( p ) ( f ˆ τ ) ) p = S ( p ) ( f ) p − S τ ( p ) ( f ) p .

Suppose that X has an equivalent p-uniformly smooth norm. Then by Pisier’s theorem, we have

∥ f ˆ τ ∥ p ≤c ∥ ( S ( p ) ( f ˆ τ ) ∥ p ,

i.e.,

E ( ∥ f − f τ ∥ p ) ≤E ( S ( p ) ( f ) p − S τ ( p ) ( f ) p ) .

Moreover, we conditionalize it, we will get

E ( ∥ f − f τ ∥ p | Σ τ + 1 ) ≤E ( S ( p ) ( f ) p − S τ ( p ) ( f ) p | Σ τ + 1 ) .

By the definition and Theorem 1.7, we get

∥ f ∥ BMO L p , q ( X ) a ≤c ∥ S ( p ) ( f ) ∥ BMO L p , q ( X ) a .

Thus we complete the proof. □

References

  1. Lorentz GG: On the theory of spaces. Pac. J. Math. 1951, 1: 411–429. 10.2140/pjm.1951.1.411

    Article  MathSciNet  MATH  Google Scholar 

  2. Neugebauer CJ: Some classical operators on Lorentz spaces. Forum Math. 1992, 4: 135–146.

    Article  MathSciNet  MATH  Google Scholar 

  3. Soria J: Lorentz spaces of weak-type. Q. J. Math. 1998, 49(2):93–103.

    Article  MathSciNet  MATH  Google Scholar 

  4. Munckenhoupt AMA: Maximal functions on classical Lorentz spaces and Hardy’s inequality with weights for nonincreasing functions. Trans. Am. Math. Soc. 1990, 320: 727–735.

    Google Scholar 

  5. Cerdà J, Martín J: Interpolation restricted to decreasing function and Lorentz spaces. Proc. Edinb. Math. Soc. 1999, 42: 243–256. 10.1017/S0013091500020228

    Article  MATH  Google Scholar 

  6. Boza S, Martín J: Equality of some classical Lorentz spaces. Positivity 2005, 9: 225–232. 10.1007/s11117-003-2712-x

    Article  MathSciNet  MATH  Google Scholar 

  7. Carro MJ, Raposo JA, Soria J: Recent developments in the theory of Lorentz spaces and weighted inequalities. Mem. Am. Math. Soc. 2007., 187: Article ID 877

    Google Scholar 

  8. Weisz F: Martingale Hardy Spaces and Their Applications in Fourier Analysis. Springer, Berlin; 1994.

    MATH  Google Scholar 

  9. Long RL: Martingale Spaces and Inequalities. Peking University Press, Beijing; 1993.

    Book  MATH  Google Scholar 

  10. Xu Q: Littlewood-Paley theory for functions with values in uniformly convex spaces. J. Reine Angew. Math. 1998, 504: 195–226.

    MathSciNet  MATH  Google Scholar 

  11. Ouyang C, Xu Q: BMO functions and Carleson measures with values in uniformly convex spaces. Can. J. Math. 2010, 62: 827–844. 10.4153/CJM-2010-043-6

    Article  MathSciNet  MATH  Google Scholar 

  12. Yong J: Carleson measures and vector-valued BMO martingales. Probab. Theory Relat. Fields 2009, 145: 421–434. 10.1007/s00440-008-0173-7

    Article  MathSciNet  MATH  Google Scholar 

  13. Martínez T:Uniform convexity in terms of martingale H 1 and BMO spaces. Houst. J. Math. 2001, 27: 461–478.

    MathSciNet  Google Scholar 

  14. Martínez T, Torrea JL: Operator-valued martingale transforms. Tohoku Math. J. 2000, 52(2):449–474.

    Article  MathSciNet  MATH  Google Scholar 

  15. Pisier G: Martingales with values in uniformly convex spaces. Isr. J. Math. 1975, 20: 326–350. 10.1007/BF02760337

    Article  MathSciNet  MATH  Google Scholar 

Download references

Acknowledgements

This work was supported by the National Natural Science Foundation of China (Grant No. 11201354), by Hubei Province Key Laboratory of Systems Science in Metallurgical Process (Wuhan University of Science and Technology) (Y201121), (Y201321) and by the National Natural Science Foundation of Pre-Research Item (2011XG005). The authors would like to deeply thank all the reviewers for their insightful and constructive comments.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Xueying Zhang.

Additional information

Competing interests

The authors declare that they have no competing interests.

Authors’ contributions

All authors contributed equally to the manuscript and read and approved the final manuscript.

Rights and permissions

Open Access This article is distributed under the terms of the Creative Commons Attribution 2.0 International License (https://creativecommons.org/licenses/by/2.0), which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited.

Reprints and permissions

About this article

Cite this article

Zhang, X., Yi, X. & Zhang, C. BMO-Lorentz martingale spaces. J Inequal Appl 2013, 371 (2013). https://doi.org/10.1186/1029-242X-2013-371

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/1029-242X-2013-371

Keywords