Skip to content
BY 4.0 license Open Access Published by De Gruyter May 9, 2021

Characterizations of variable martingale Hardy spaces via maximal functions

  • Ferenc Weisz

Abstract

We introduce a new type of dyadic maximal operators and prove that under the log-Hölder continuity condition of the variable exponent p(⋅), it is bounded on Lp(⋅) if 1 < pp+ ≤ ∞. Moreover, the space generated by the Lp(⋅)-norm (resp. the Lp(⋅), q-norm) of the maximal operator is equivalent to the Hardy space Hp(⋅) (resp. to the Hardy-Lorentz space Hp(⋅), q). As special cases, our maximal operator contains the usual dyadic maximal operator and four other maximal operators investigated in the literature.

1 Introduction

For a measurable function p(⋅), the variable Lebesgue space Lp(⋅) consists of all measurable functions f for which 01 f(x)∣p(x)dx < ∞. If p(⋅) is a constant, we get back the usual Lp space. This topic needs essentially new ideas and is investigated very intensively in the literature nowadays (see e.g. Cruz-Uribe and Fiorenza [6], Diening et al. [7], Kokilashvili et al. [16, 17], Kováčik and Rákosník [18], Cruz-Uribe et al. [4, 3, 2], Nakai and Sawano [22, 31], Kempka and Vybíral [15], Rafeiro et al. [24, 27], Samko [8, 28, 29], Jiao et al. [11, 12, 14], Yan et al. [40], Liu et al. [19, 20]). Interest in the variable Lebesgue spaces has increased since the 1990s because of their use in a variety of applications (see the references in Jiao et al. [11]).

Usually, we suppose that p(⋅) or 1/p(⋅) satisfy the log-Hölder continuity condition. The classical Hardy-Littlewood maximal operator is bounded on the variable Lp(⋅) spaces if the exponent function p(⋅) is log-Hölder continuous and 1 < pp+ < ∞, where p denotes the infimum and p+ the supremum of p(⋅) (see for example Cruz-Uribe et.al [4] and Nekvinda [23]). Moreover, under the log-Hölder continuity condition of 1/p(⋅), the maximal operator is bounded on Lp(⋅) when 1 < pp+ ≤ ∞ (see e.g. Cruz-Uribe and Fiorenza [6] and Diening et al. [7]). The fractional integral operator was investigated in Ephremidze et al. [8], Rafeiro and Samko [27, 25, 26 30] and, for martingales, in Hao et al. [9, 13].

The boundedness result for the usual dyadic (or martingale) maximal operator on the Lp(⋅) spaces was proved in [11, 12] if 1 < pp+ < ∞. In [36], we generalized this result to 1 < pp+ ≤ ∞. In [11, 35], we investigated four more dyadic maximal operators and show that they are bounded on Lp(⋅) if 1 < pp+ < ∞. The boundedness of these operators was the key point in proving the boundedness of the maximal Fejér, Cesàro and Riesz operators of the Walsh-Fourier series from the variable Hardy space Hp(⋅) to Lp(⋅) (see [11, 33]).

Nakai and Sawano [22] first introduced the Hardy space Hp(⋅)(ℝ) with a variable exponent p(⋅) and established the atomic decompositions. Independently, Cruz-Uribe and Wang [5] also investigated the variable Hardy space Hp(⋅)(ℝ). Sawano [31] improved the results in [22]. Ho [10] studied weighted Hardy spaces with variable exponents. Recently, Yan et al. [40] introduced the variable weak Hardy space Hp(⋅),∞(ℝ) and characterized these spaces via radial maximal functions. The Hardy-Lorentz spaces Hp(⋅), q(ℝ) were investigated by Jiao et al. in [14]. Similar results for the anisotropic Hardy spaces Hp(⋅)(ℝ) and Hp(⋅), q(ℝ) can be found in Liu et al. [19, 20]. Martingale Musielak–Orlicz Hardy spaces were investigated in Xie et al. [37, 38, 39]. Very recently, these results are generalized for martingale Hardy spaces with variable exponent in Jiao et al. [11].

In this paper, we introduce a common generalization of the usual dyadic and the other four maximal operators, denoted by Uγ,s, where γ and s are positive parameters. We prove that if 1/p(⋅) satisfy the log-Hölder continuity condition, 1p1p+<γ+s and 1 < pp+ ≤ ∞, then Uγ,s is bounded on Lp(⋅). We also verify that under the same conditions, ∥Uγ,sfLp(⋅) is equivalent to ∥fHp(⋅) and ∥Uγ,sfLp(⋅),q is equivalent to ∥fHp(⋅),q, where Hp(⋅) and Hp(⋅), q denote the variable Hardy and Hardy-Lorentz spaces. Moreover, we show with a counterexample that the condition 1p1p+<γ+s is important, the results do not hold without this condition.

2 Variable Lebesgue and Lorentz spaces

In this section, we recall some basic notations on variable Lebesgue spaces and give some elementary and necessary facts about these spaces. Our main references are Cruz-Uribe and Fiorenza [6] and Diening et al. [7].

For a constant p, the Lp space is equipped with the quasi-norm

fp:=01|f(x)|pdx1/p(0<p<),

with the usual modification for p = ∞. Here we integrate with respect to the Lebesgue measure λ.

We are going to generalize these spaces. A measurable function p(⋅) : [0, 1) → (0, ∞] is called a variable exponent. For a measurable set A ⊂ [0, 1), we denote

p(A):=essinfxAp(x),p+(A):=esssupxAp(x)

and for convenience

p:=p([0,1)),p+:=p+([0,1)).

Denote by 𝓟 the collection of all variable exponents p(⋅) such that

0<pp+.

In what follows, we use the symbol

p_=min{p,1}.

We define the modular functional by

ρ(f)=[0,1)Ω|f(x)|p(x)dP+fL(Ω),

where Ω = {x ∈ [0, 1) : p(x) = ∞}. The variable Lebesgue space Lp(⋅) is the collection of all measurable functions f for which there exists ν > 0 such that

ρf/ν<.

This becomes a quasi-Banach function space when it is equipped with the quasi-norm

fp():=inf{ν>0:ρ(fν1}.

If p(⋅) = p is a constant, then we get back the definition of the usual Lp spaces. For any fLp(⋅), we have ρ(f) ≤ 1 if and only if ∥fp(⋅) ≤ 1 (see Cruz-Uribe and Fiorenza [6]). It is known that ∥ν fp(⋅) = ∣ν∣∥fp(⋅),

|f|sp()=fsp()s

and

f+gp()p_fp()p_+gp()p_,

where p(⋅) ∈ 𝓟, s ∈ (0, ∞), ρ ∈ ℂ and f, gLp(⋅). Details can be found in the monographs Cruz-Uribe and Fiorenza [6] and Diening et al. [7]. The variable exponent p′(⋅) is defined pointwise by

1p(x)+1p(x)=1,x[0,1).

The next two lemmas are well known, see Cruz-Uribe and Fiorenza [6] or Diening et al. [7].

Lemma 2.1

Let p(⋅) ∈ 𝓟 with 1 ≤ pp+ ≤ ∞. For all fLp(⋅) and gLp′(⋅),

01fgdλCp()fp()gp().

Lemma 2.2

Let p(⋅) ∈ 𝓟 with 1 ≤ pp+ ≤ ∞. Then

fp()sup01fgdP,

where the supremum is taken over all gLp′(⋅) withgp′(⋅) ≤ 1.

We denote by Clog the set of all functions p(⋅) ∈ 𝓟 satisfying the so-called globally log-Hölder continuous condition, namely, there exists a positive constant Clog(p) such that, for any x, y ∈ [0, 1),

|p(x)p(y)|Clog(p)log(e+1/|xy|). (2.1)

The following two lemmas were proved in Cruze-Uribe and Fiorenza [6] (see also Hao and Jiao [9]).

Lemma 2.3

If 1/p(⋅) ∈ Clog, then there exists a constant 0 < β < 1 such that for all intervals I ⊂ [0, 1),

λ(I)1/p+(I)1/p(I)1β. (2.2)

Lemma 2.4

If 1/p(⋅) ∈ Clog, then for any interval I ⊂ [0, 1),

λ(I)1/p(I)λ(I)1/p(x)λ(I)1/p+(I)χIp()(xI),

wheredenotes the equivalence of the numbers.

Note that under the condition 0 < pp+ < ∞, p(⋅) ∈ Clog if and only if 1/p(⋅) ∈ Clog.

For general martingale Hardy spaces, instead of the log-Hölder continuity condition, we supposed in [9, 11, 13, 33, 36] the slightly more general condition (2.2) for all atoms of the σ-algebras. All results of this paper remain true if, instead of (2.1), we suppose (2.2) for all dyadic intervals I ⊂ [0, 1). By a dyadic interval, we mean one of the form [k2n, (k+1)2n) for some k, n ∈ ℕ, 0 ≤ k < 2n.

Remark 2.1

There exist a lot of functions p(⋅) satisfying (2.1) and (2.2). For concrete examples we mention the function a + cx for parameters a and c such that the function is positive (x ∈ [0, 1)). All positive Lipschitz functions with order 0 < α ≤ 1 also satisfy (2.1) and (2.2).

The following lemma can be found in [36]. A first version of this lemma was proved in Jiao et al. [12, 11].

Lemma 2.5

Let p(⋅) ∈ 𝓟 satisfy (2.2) and 1 ≤ pp+ ≤ ∞. Suppose that fLp(⋅) withfp(⋅) ≤ 1, f = f χf∣>1 and supp f Ωc . Then, for any interval I ⊂ [0, 1) with λ(I Ωc ) > 0 and for any p(I) ≤ rp+(I) (r < ∞),

βtP(I)I|f(y)|dyr1P(I)I|f(y)|p(y)dy.

The variable Lorentz spaces were introduced and investigated by Kempka and Vybíral [15]. Lp(⋅), q is defined to be the space of all measurable functions f such that

fp(),q:=0ρqχ{x[0,1):|f(x)|>ρ}p()qdρρ1/q,if 0<q<,supρ(0,)ρχ{x[0,1):|f(x)|>ρ}p(),if q=

is finite. If p(⋅) is a constant, we get back the classical Lorentz spaces (see Lorentz [21] or Bergh and Löfström [1].

In this paper the constants C are absolute constants and the constants Cp(⋅) are depending only on p(⋅) and may denote different constants in different contexts. For two positive numbers A and B, we use also the notation AB, which means that there exists a constant C such that ACB.

3 Variable martingale Hardy spaces

Let 𝓕n be the σ-algebra

Fn=σ{[k2n,(k+1)2n):0k<2n},

where σ(𝓗) denotes the σ-algebra generated by an arbitrary set system 𝓗. The conditional expectation operators relative to 𝓕n are denoted by En. An integrable sequence f = (fn)n∈ℕ is said to be a martingale if fn is 𝓕n-measurable for all n ∈ ℕ and En fm = fn in case nm. Martingales with respect to (𝓕n, n ∈ ℕ) are called dyadic martingales. It is easy to show (see e.g. Weisz [34]) that the sequence (𝓕n, n ∈ ℕ) is regular, i.e., fn ≤ Rf_{n-1} for all non-negative dyadic martingales.

For a dyadic martingale f = (fn)n∈ℕ, the maximal function is defined by

Mf:=supnNfn.

Now we can define the variable martingale Hardy spaces by

Hp():=f=fnnN:fHp():=M(f)p()<.

The variable martingale Hardy-Lorentz spaces can be defined similarly:

Hp(),q:=f=fnnN:fHp(),q:=M(f)p(),q<.

The Hardy spaces can also be defined via equivalent norms. In [11], we have shown that the Hardy spaces defined the square function and the conditional square function are equivalent. In this paper, we give other equivalent characterization of the variable Hardy spaces via new maximal functions.

The atomic decomposition is a useful characterization of the Hardy spaces. A measurable function a is called a p(⋅)-atom if there exists a stopping time τ such that

  1. En(a) = 0 for all nτ,

  2. M(a)χτ<p()1.

The atomic decomposition of the spaces Hp(⋅) and Hp(⋅), q were proved in Jiao et al. [11, 12]. The classical case can be found in [34].

Theorem 3.1

Let p(⋅) ∈ Clog, 0 < pp+ < ∞ and 0 < q ≤ ∞. Then the martingale f = (fn)n∈ℕHp(⋅) or f = (fn)n∈ℕHp(⋅), q, respectively, if and only if there exists a sequence (ak)k∈ℤ of p(⋅)-atoms such that for every n ∈ ℕ,

fn=kZμkEnakalmosteverywhere, (3.1)

where μk = 3 ⋅ 2kχ{τk<∞}p(⋅) and τk is the stopping time associated with the p(⋅)-atom ak. Moreover,

fHp()infkZμkχτk<χτ k<p()t1/tp(),fHp(),qinfkZ2kqχτk<p()q1/q,

respectively, where 0 < tp is fixed and the infimum is taken over all decompositions of the form (3.1).

4 The boundedness of maximal operators on Lp(⋅)

The following result was proved in Jiao et al. [11, 12] if p+ < ∞ and by the author [36] if p+ ≤ ∞. For the boundedness of the classical Hardy-Littlewood maximal operator see e.g. Cruz-Uribe and Fiorenza [6] and Diening et al. [7].

Theorem 4.1

If 1/p(⋅) ∈ Clog and 1 < pp+ ≤ ∞, then

Mfp()Cp()fp()(fLp()).

We generalize the preceding dyadic maximal operator. For a martingale f = (fn), let us introduce

Uγ,sf(x):=supxIm=0nj=0m2(jn)γi=jm2(ji)s1λ(Ij,i)Ij,ifndλ,

where I is a dyadic interval with length 2n, γ, s are positive constants and

Ij,i:=I+˙[0,2i)+˙2j1.

Here +̇ denotes the dyadic addition (see e.g. Schipp, Wade, Simon and Pál [32]). Of course, if fL1, then we can write in the definition f instead of fn. Let us define Ik,n := [k2n, (k+1)2n) with 0 ≤ k < 2n, n ∈ ℕ. The preceding definition can be rewritten to

Uγ,sf:=supnNk=02n1χIk,nm=0nj=0m2(jn)γi=jm2(ji)s1λ(Ik,nj,i)Ik,nj,ifndλ.

We proved the following theorem in [33].

Theorem 4.2

For all 1 < p ≤ ∞ and all 0 < γ, s < ∞, we have

Uγ,sfpCpfp(fLp).

Now we generalize this theorem to variable Lebesgue spaces.

Theorem 4.3

Let 1/p(⋅) ∈ Clog, 1 < pp+ ≤ ∞ and 0 < γ, s < ∞. If

1p1p+<γ+s, (4.1)

then

Uγ,sfp()Cp()fp()(fLp()).

Proof

By homogeneity, it is enough to show the theorem for ∥fp(⋅) = 1. We may suppose that f is non-negative. We decompose f as f1+f2, where

f1=fχ{f>1},f2=fχ{f1}.

Then ∥fip(⋅) ≤ 1 and ρ(fi) ≤ 1, i = 1,2. Since

Uγ,sf22s2s122γ(2γ1)2:=CsCγ,

we get by convexity that

ρ(αUγ,sf)12ρ(2αUγ,sf1)+12ρ(2αUγ,sf2)12Ωc2αUγ,sf1(x)p(x)dx+αUγ,sf1L(Ω)+2αCsCγ, (4.2)

where 8 α CsCγ < β. For a fixed k, n, let us denote by Λk,n those pairs (j, i) for which 0 ≤ jn, 0 ≤ in and

βλ(Ik,nj,i)Ik,nj,if1(t)dt1.

Then,

12Ωc2αUγ,sf1(x)p(x)dx14Ωc(4αsupnNk=02n1χIk,n(x)×m=0nj=0mi=jm2(jn)γ2(ji)sχΛk,n(j,i)λ(Ik,nj,i)Ik,nj,if1(t)dt)p(x)dx+14Ωc(4αsupnNk=02n1χIk,n(x)×m=0nj=0mi=jm2(jn)γ2(ji)sχΛk,nc(j,i)λ(Ik,nj,i)Ik,nj,if1(t)dt)p(x)dx=:(A1)+(A2).

It is easy to see that

(A1)14Ωc4αCsCγβp(x)dx144αCsCγβ1. (4.3)

We denote by Ik,n,j,i,1 (resp. Ik,n,j,i,2) those points xIk,n for which p(x) ≤ p(Ik,nj,i) (resp. p(x) > p(Ik,nj,i) ). Then

(A2)18l=12Ωc(supnNk=02n1χIk,n(x)m=0nj=0mi=jm2(jn)γ2(ji)s×8αχΛk,nc(j,i)χIk,n,j,i,l(x)λ(Ik,nj,i)Ik,nj,if1(t)dt)p(x)dx=:(A21)+(A22).

Let q(x) := p(x)/p0 > 1 for some 1 < p0 < p. Note that the sets Ik,n are disjoint for a fixed n. By convexity,

(A21)18ΩcsupnNk=02n1χIk,n(x)m=0nj=0mi=jm2(jn)γ2(ji)sCsCγ×8αCsCγχΛk,nc(j,i)χIk,n,j,i,1(x)λ(Ik,nj,i)Ik,nj,if1(t)dtq(x)p0dx18ΩcsupnNk=02n1χIk,n(x)m=0nj=0mi=jm2(jn)γ2(ji)sCsCγ×βχΛk,nc(j,i)χIk,n,j,i,1(x)λ(Ik,nj,i)Ik,nj,if1(t)dtq(x)p0dx. (4.4)

Using that q(x) ≤ q(Ik,nj) on Ik,n,j,K,1 and

βλ(Ik,nj,i)Ik,nj,if1(t)dt>1,

we get that

(A21)18(CsCγ)p0ΩcsupnNk=02n1χIk,n(x)m=0nj=0mi=jm2(jn)γ2(ji)s×βχΛk,nc(j,i)χIk,n,j,i,1(x)λ(Ik,nj,i)Ik,nj,if1(t)dtq(Ik,nj,i)p0dx. (4.5)

Since ρ(f1) ≤ 1 implies that supp f1 Ωc , we can apply Lemma 2.5 and Theorem 4.2 to obtain

(A21)18(CsCγ)p0ΩcsupnNk=02n1χIk,n(x)m=0nj=0mi=jm2(jn)γ2(ji)s×1λ(Ik,nj,i)Ik,nj,i|f1(t)|q(t)dtp0dx18(CsCγ)p0Uγ,s((f1)q())p0p0C1|f|q()p0p0C1. (4.6)

Choosing 0 < γ0 < γ and 0 < r < s + γ0, we obtain

(A22)18Ωc(supnNk=02n1χIk,n(x)(m=0nj=0mi=jm2(jn)(γγ0)2(ji)(γ0+sr)Cγγ0Cγ0+sr×8αCγγ0Cγ0+sr2(ji)rχΛk,nc(j,i)χIk,n,j,i,2(x)λ(Ik,nj,i)Ik,nj,if1(t)dt)q(x))p0dx18(Cγγ0Cγ0+sr)p0Ωc(supnNk=02n1χIk,n(x)×m=0nj=0mi=jm2(jn)(γγ0)2(ji)(γ0+sr)×βp2(ji)rχΛk,nc(j,i)χIk,n,j,i,2(x)λ(Ik,nj,i)Ik,nj,if1(t)dtq(x))p0dx, (4.7)

whenever 8 α Cγγ0Cγ0+srβp. By Hölder’s inequality,

(A22)18(Cγγ0Cγ0+sr)p0Ωc(supnNk=02n1χIk,n(x)×m=0nj=0mi=jm2(jn)(γγ0)2(ji)(γ0+sr)2(ji)rq(x)βpq(x)×χΛk,nc(j,i)χIk,n,j,i,2(x)λ(Ik,nj,i)Ik,nj,i|f1(t)|q(Ik,nj,i)dtq(x)/q(Ik,nj,i))p0dx18(Cγγ0Cγ0+sr)p0Ωc(supnNk=02n1χIk,n(x)×m=0nj=0mi=jm2(jn)(γγ0)2(ji)(γ0+sr)2(ji)rq(x)βpq(x)2iq(x)/q(Ik,nj,i)×χIk,n,j,i,2(x)Ik,nj,i|f1(t)|q(Ik,nj,i)dtq(x)/q(Ik,nj,i))p0dx. (4.8)

Observe that ρ(f1) ≤ 1 implies that supp f1 Ωc . Since f1 > 1 or f1 = 0, q(x) > q(Ik,nj,i) on Ik,n,j,K,2, q(Ik,nj,i) q(t) < p(t) for all t Ik,nj,i and

Ik,nj,i|f1(t)|q(Ik,nj,i)dtΩc|f1(t)|p(t)dt1, (4.9)

we can see that

(A22)18(Cγγ0Cγ0+sr)p0Ωc(supnNk=02n1χIk,n(x)×m=0nj=0mi=jm2(jn)(γγ0)2(ji)(γ0+sr)2(ji)rq(x)2iq(x)/q(Ik,nj,i)i×βpq(x)χIk,n,j,i,2(x)λ(Ik,nj,i)Ik,nj,i|f(t)|q(Ik,nj,i)dt)p0dx.

For fixed k, n let Jj denote the dyadic interval with length 2j and Ik,nJj. Then Ik,nj,i Jj +̇ 2j−1 = Jj. Inequality (2.2) implies that

2j/q(x)+j/q(Ik,nj,i)1βp01βp, (4.10)

thus

βpq(x)2jjq(x)/q(Ik,nj,i)

for xIk,n. Hence,

2(ji)rq(x)2iq(x)/q(Ik,nj,i)iβpq(x)2(ji)rq(x)2iq(x)/q(Ik,nj,i)i2jjq(x)/q(Ik,nj,i)=2(ji)(rq(x)q(x)q(Ik,nj,i)+1). (4.11)

Observe that

rq(x)q(x)q(Ik,nj)+1q(x)(r1q)+1>0,

whenever

1q1q+<r. (4.12)

In this case the expression in (4.11) can be estimated by 1. Hence

(A22)18(Cγγ0Cγ0+sr)p0Ωc(supnNk=02n1χIk,n(x)m=0nj=0mi=jm×2(jn)(γγ0)2(ji)(γ0+sr)1λ(Ik,nj,i)Ik,nj,i|f(t)|q(t)dt)p0dx18(Cγγ0Cγ0+sr)p0Uγγ0,γ0+sr(|f|q())p0p0C2|f|q()p0p0C2, (4.13)

whenever (4.12) holds. Since r can be arbitrarily near to s + γ0 and γ0 to γ, (4.12) gives (4.1).

Now, we consider the second term of (4.2). Notice that if Ω has positive measure, then p+ = ∞. We have

αUγ,sf1(x)αsupnNk=02n1χIk,n(x)×m=0nj=0mi=jm2(jn)γ2(ji)sχΛk,n(j,i)λ(Ik,nj,i)Ik,nj,if1(t)dt+αsupnNk=02n1χIk,n(x)×m=0nj=0mi=jm2(jn)γ2(ji)sχΛk,nc(j,i)λ(Ik,nj,i)Ik,nj,if1(t)dt=:(B1)(x)+(B2)(x),

where x ∈ Ω. Thus

(B1)αCsCγβ1. (4.14)

For 1 < u < ∞, we denote by Γk,n,u those pairs (j, i) for which u p(Ik,nj,i) . Then

(B2)(x)supnNk=02n1χIk,n(x)m=0nj=0mi=jm2(jn)γ2(ji)s×αχΛk,nc(j,i)χΓk,n,u(j,i)λ(Ik,nj,i)Ik,nj,if1(t)dt+supnNk=02n1χIk,n(x)m=0nj=0mi=jm2(jn)γ2(ji)s×αχΛk,nc(j,i)χΓk,n,uc(j,i)λ(Ik,nj,i)Ik,nj,if1(t)dt=:(B21)(x)+(B22)(x).

For a fixed x ∈ Ω there exist n ∈ ℕ and 0 ≤ k < 2n such that

(B21)(x)2χIk,n(x)m=0nj=0mi=jm2(jn)γ2(ji)s×αχΛk,nc(j,i)χΓk,n,u(j,i)λ(Ik,nj,i)Ik,nj,if1(t)dt=:(B21)(x).

Similarly to (4.4) and (4.5), we obtain

(B21)u(x)(m=0nj=0mi=jm2(jn)γ2(ji)sCsCγ×2αCsCγχΛk,nc(j,i)χΓk,n,u(j,i)λ(Ik,nj,i)Ik,nj,if1(t)dt)u,

and further,

(B21)u(x)1CsCγm=0nj=0mi=jm2(jn)γ2(ji)s×βχΛk,nc(j,i)χΓk,n,u(j,i)λ(Ik,nj,i)Ik,nj,if1(t)dtu1CsCγm=0nj=0mi=jm2(jn)γ2(ji)s×βχΛk,nc(j,i)χΓk,n,u(j,i)λ(Ik,nj,i)Ik,nj,if1(t)dtp(Ik,nj,i).

Note that supp f1 Ωc and we use the convention 0 = 0. So the following holds also if p(Ik,nj,i) = ∞. By Lemma 2.5 and (4.9), we get that

(B21)u(x)1CsCγm=0nj=0mi=jm2(jn)γ2(ji)s1λ(Ik,nj,i)Ik,nj,i|f1(t)|p(t)dt1CsCγm=0nj=0mi=jm2(jn)γ2(ji)s2i2n.

In other words, (B21)(x) ≤ 2n/u. Since this holds for all 1 < u < ∞, we conclude that

(B21)(x)(B21)(x)1(xΩ). (4.15)

Similarly, for a fixed x ∈ Ω there exist n ∈ ℕ and 0 ≤ k < 2n such that

(B22)(x)2χIk,n(x)m=0nj=0mi=jm2(jn)γ2(ji)s×αχΛk,nc(j,i)χΓk,n,uc(j,i)λ(Ik,nj,i)Ik,nj,if1(t)dt=:(B22)(x).

We can see as in (4.7) and (4.8) that

(B22)u(x)(m=0nj=0mi=jm2(jn)(γγ0)2(ji)(γ0+sr)Cγγ0Cγ0+sr×2αCγγ0Cγ0+sr2(ji)rχΛk,nc(j,i)χΓk,n,uc(j,i)λ(Ik,nj,i)Ik,nj,if1(t)dt)u1Cγγ0Cγ0+srm=0nj=0mi=jm2(jn)(γγ0)2(ji)(γ0+sr)×β2(ji)rχΛk,nc(j,i)χΓk,n,uc(j,i)λ(Ik,nj,i)Ik,nj,if1(t)dtu1Cγγ0Cγ0+srm=0nj=0mi=jm2(jn)(γγ0)2(ji)(γ0+sr)2(ji)ru×βuχΛk,nc(j,i)χΓk,n,uc(j,i)λ(Ik,nj,i)Ik,nj,i|f1(t)|p(Ik,nj,i)dtu/p(Ik,nj,i).

Inequality (4.9) implies

(B22)u(x)1Cγγ0Cγ0+srm=0nj=0mi=jm2(jn)(γγ0)2(ji)(γ0+sr)2(ji)ru×βu2iu/p(Ik,nj,i)χΓk,n,uc(j,i)Ik,nj,i|f1(t)|p(Ik,nj,i)dtu/p(Ik,nj,i)1Cγγ0Cγ0+srm=0nj=0mi=jm2(jn)(γγ0)2(ji)(γ0+sr)×2(ji)ruβu2iu/p(Ik,nj,i)χΓk,n,uc(j,i)Ik,nj,i|f1(t)|p(Ik,nj,i)dt.

Since x ∈ ΩIk,n, we have p+(Ik,n) = ∞. If p(Ik,nj,i) = ∞, then Ik,nj,i|f1(t)|p(Ik,nj,i)dt=0. Otherwise, we obtain

2j/u+j/p(Ik,nj,i)1βandβu2jju/p(Ik,nj,i)

for all p(Ik,nj,i) < u < ∞ as in (4.10). So

2(ji)ru2iu/p(Ik,nj,i)iβu2(ji)ru2iu/q(Ik,nj,i)i2jju/p(Ik,nj,i)=2(ji)ruup(Ik,nj,i)+11.

Indeed,

ruup(Ik,nj)+1>0

whenever

1p<r. (4.16)

Recall that p+ = ∞. This and inequality (4.10) imply

(B22)u(x)1Cγγ0Cγ0+srm=0nj=0mi=jm2(jn)(γγ0)2(ji)(γ0+sr)2i×Ik,nj,i|f1(t)|p(Ik,nj,i)dt1Cγγ0Cγ0+srm=0nj=0mi=jm2(jn)(γγ0)2(ji)(γ0+sr)2i2n

and so

(B22)(x)(B22)(x)1(xΩ). (4.17)

Since r can be arbitrarily near to s + γ0 and γ0 to γ, conditions (4.16) and (4.1) are the same. Taking into account (4.2), (4.3), (4.6), (4.13), (4.14), (4.15) and (4.17), we obtain ρ(α Uγ,sf) ≤ C, where C = 5+C1+C2. By convexity,

ρ(Uγ,sfC/α)1Cρ(αUγ,sf)1,

which means that

Uγ,sfp()Cα=Cαfp(),

whenever (4.1) is satisfied. This finishes the proof of Theorem 4.3.□

Remark 4.1

Inequality (4.1) and Theorem 4.3 hold if p > max(1/(γ+s),1).

Corollary 4.1

Let 1/p(⋅) ∈ Clog satisfy (4.1), 1 < pp+ ≤ ∞ and 0 < γ, s < ∞. Thenfp(⋅) ∼ ∥Uγ,sfp(⋅) and

Mfp()Uγ,sfp()Cp()Mfp()(fLp()).

Proof

If j = i = n, then Ij,i = I, hence MfUγ,sf. Theorem 4.3 completes the proof:

fp()Mfp()Uγ,sfp()Cp()fp()(fLp()).

In special cases, we proved in [11, 35] that the condition (4.1) is important, the theorems are not true without this condition.

5 Some special maximal operators

In this section we consider five special cases of the maximal operator Uγ,s. Theorem 4.1 holds for the operators in Examples 5.1 and 5.2 and Theorem 4.3 holds for Examples 5.3, 5.4 and 5.5. Under the additional condition p+ < ∞, Theorem 4.3 was proved in [11, 35] for the next special operators.

Example 5.1

Considering the indices j = i = n = m in the definition of Uγ,sf, only, we obtain

Uγ,s(1)f(x):=supxI1λ(I)Ifndλ=Mf(x).

Note that In,n := I.

Example 5.2

If m = 0, …, n − 1, j = i = m, then

Uγ,s(2)f(x):=supxIm=0n12(mn)γ1λ(Im,m)Im,mfndλ.

Here Im,m := I\dot+[0, 2m). It is easy to see that Uγ,s(2)fMf.

Example 5.3

If m = n, j = 0, …, n − 1 and i = n, then

Uγ,s(3)f(x):=supxIj=0n12(jn)(γ+s)1λ(Ij,n)Ij,nfndλ.

Note that Ij,n := I +̇ 2j−1.

Example 5.4

If m = n − 1, then

Uγ,s(4)f(x):=supxIj=0n12(jn)γi=jn12(ji)s1λ(Ij,i)Ij,ifndλ.

Example 5.5

If m = 0, …, n − 1, then

Uγ,s(5)f(x):=supxIm=0n1j=0m2(jn)γi=jm2(ji)s1λ(Ij,i)Ij,ifndλ.

6 The equivalence of maximal operators on Hp(⋅)

In this section, we apply the atomic characterization to prove the boundedness of Uγ,s from Hp(⋅) to Lp(⋅). We need the following result due to Jiao et al. [11].

Theorem 6.1

Let p(⋅) ∈ Clog, 0 < pp+ < ∞ and 0 < t < p. Suppose that the σ-sublinear operator T : LL is bounded and

kZμktT(ak)tχ{τk=}p()/tkZ2ktχ{τk<}p()/t, (6.1)

where μk = 3⋅ 2kχ{τk<∞}p(⋅) and τk is the stopping time associated with the p(⋅)-atom ak. Then we have

Tfp()fHp()(fHp()).

Theorem 6.2

Let 1/p(⋅) ∈ Clog, 0 < γ, s < ∞ and 1 < pp+ ≤ ∞ or 0 < pp+ < ∞. If (4.1) holds, then

Uγ,sfp()Cp()fHp()(fHp()).

Proof

In case 1 < pp+ ≤ ∞, the result follows from Theorem 4.3. So we can suppose that 0 < pp+ < ∞. Since Uγ,s is bounded on L by Theorem 4.2, it is enough to prove (6.1) for Uγ,s. For a fixed k ∈ ℤ, the sets {τk = K} are disjoint and there exist disjoint dyadic intervals Ik,K,μ ∈ 𝓕K such that

τk=K=μIk,K,μ(KN),

where the union in μ is finite and λ(Ik,K,μ) = 2K. Thus

τk<=KNμIk,K,μ,

where the dyadic intervals Ik,K,μ are disjoint for a fixed k ∈ ℤ. Then

ak=KNμakχIk,K,μ.

Since intIk,K,μ ak dλ = 0,

Ij,iakdλ=0

if iK. That is, we can suppose that i > K, thus nm > K. If xIk,K,μ, xI and jK, then Ij,iIk,K,μ = ∅. Therefore we can suppose that j < K. Similarly, if xIk,K,μ +̇ [2j−1,2j) ∖ (Ik,K,μ +̇ 2j−1), then Ij,iIk,K,μ = ∅, so we may assume that xIk,K,μ+˙2j1=Ik,K,μj,K. Therefore, for xIk,K,μ,

Uγ,s(akχIk,K,μ)(x)supn>KχI(x)m=K+1nj=0K12(jn)γi=K+1m2(ji)s1λ(Ij,i)Ij,iadλχIk,K,μj,K(x)χ{τk<}p()1supn>KχI(x)m=K+1nj=0K12(jn)γi=K+1m2(ji)sχIk,K,μj,K(x)χ{τk<}p()1supn>KχI(x)m=K+1nj=0K12(jn)γ2(jK)sχIk,K,μj,K(x)χ{τk<}p()1supn>K(nK)2(Kn)γj=0K12(jK)(γ+s)χIk,K,μj,K(x).

Since the function xx2γx is bounded, we obtain that

Uγ,s(akχIk,K,μ)(x)χ{τk<}p()1j=0K12(jK)(γ+s)χIk,K,μj,K(x).

Consequently, for x ∈ {τk = ∞},

Uγ,s(ak)(x)χ{τk<}p()1KNμj=0K12(jK)(γ+s)χIk,K,μj,K(x) (6.2)

and

kZμktUγ,s(ak)tχ{τk=}p()/tkZ2ktKNμj=0K12(jK)(γ+s)tχIk,K,μj,Kp()/t=:Z,

where 0 < t < p.

Let us choose max(1, p+) < r < ∞. By Lemma 2.2, there is gL(p()t) with norm less than 1 such that

Z01kZ2ktKNμj=0K12(jK)(γ+s)tχIk,K,μj,KgdλkZ2ktKNμj=0K12(jK)(γ+s)tχIk,K,μj,KrtχIk,K,μj,Kg(rt)kZ2ktKNμj=0K12(jK)(γ+s)t×01χIk,K,μ1λ(Ik,K,μj,K)Ik,K,μj,Kg(rt)dλ1/(rt)dλ,

because λ(Ik,K,m)=λ(Ik,K,mj,K)=2K. Then

Z01kZ2ktKNμχIk,K,μj=0K12(jK)(γ+s)t(1/(rt)+1/(rt))×1λ(Ik,K,μj,K)Ik,K,μj,Kg(rt)1/(rt)dλ,

and by Hölder’s inequality,

Z01kZ2ktKNμχIk,K,μj=0K12(jK)(γ+s)t1/(rt)×j=0K12(jK)(γ+s)t1λ(Ik,K,μj,K)Ik,K,μj,Kg(rt)1/(rt)dλ01kZ2ktKNμχIk,K,μ×j=0K12(jK)(γ+s)t1λ(Ik,K,μj,K)Ik,K,μj,Kg(rt)1/(rt)dλ.

Example 5.3 and Theorem 4.3 imply

Z01kZ2ktKNμχIk,K,μUγt,st(3)(|g|(rt))1/(rt)dλkZ2ktKNμχIk,K,μp()/tUγt,st(3)(|g|(rt))1/(rt)(p()/t)kZ2ktKNμχIk,K,μp()/tgL(p()/t)kZ2ktχ{τk<}p()/t,

whenever

1(p()/t)/(r/t)1(p()/t)/(r/t)+=r/(rt)p+/(p+t)r/(rt)p/(pt)<(γ+s)t.

Since r can be arbitrarily large, this means that

p+tp+ptp<(γ+s)t,

which is exactly (4.1). This completes the proof.□

Remark 6.1

Theorem 6.2 holds if p > 1/(γ + s).

Now we generalize Corollary 4.1 and verify that ∥⋅∥Hp(⋅)∼ ∥Uγ,s(⋅)∥p(⋅).

Corollary 6.1

Let 1/p(⋅) ∈ Clog, 0 < γ, s < ∞ and 1 < pp+ ≤ ∞ or 0 < pp+ < ∞. If (4.1) holds, then

fHp()Uγ,sfp()Cp()fHp()(fHp()).

Theorem 6.2 does not hold if (4.1) is not satisfied. More exactly, we show

Theorem 6.3

Let 1/p(⋅) ∈ Clog. If

1p(I0,n1)1p+(I0,n0,n)>γ+s (6.3)

for all n ∈ ℕ, then Uγ,s is not bounded from Hp(⋅) to Lp(⋅).

Proof

Let

an1(t)=2(n1)/p(I0,n1)(χI0,nχI1,n)

and xI0,n−1. By Lemma 2.4, an−1 is an atom for all n ≥ 1, and so ∥an−1Hp(⋅) ≤ 1. Choosing m = n = N and i = n, we can see that

Uγ,san1=supNNk=02N1χIk,Nm=0Nj=0m2(jN)γi=jm2(ji)s1λ(Ik,Nj,i)Ik,Nj,iI0,n1an1dλχJj=0n2(jn)(γ+s)1λ(Jj,n)Jj,nI0,n1an1dλ,

where J=[1/2,1/2+2n)=I0,n0,n. The terms except j = 0 are all 0, so

Uγ,san1χI0,n0,n2n(γ+s)1λ(I0,n)I0,nI0,n1an1dλχI0,n0,n(x)2n(γ+s)2(n1)/p(I0,n1).

Then

01Uγ,san1(x)p(x)dxI0,n0,n2n(γ+s)p(x)2(n1)p(x)/p(I0,n1)dxCI0,n0,n2np+(I0,n0,n)(1/p(I0,n1)(γ+s))dx=C2np+(I0,n0,n)(1/p(I0,n1)(γ+s))2n

which tends to infinity as n → ∞ if (6.3) holds.□

7 The equivalence of maximal operators on Hp(⋅), q

In this section, we extend the main results in Section 6 to the variable Hardy-Lorentz space setting. The next result was proved in Jiao et al. [11].

Theorem 7.1

Let p(⋅) ∈ Clog, 0 < pp+ < ∞ and 0 < q ≤ ∞. Suppose that the σ-sublinear operator T : LL is bounded and

|Ta|δχ{τ=}p()Cχ{τ<}p()1δ

for some 0 < δ < 1 and all p(⋅)-atoms a, where τ is the stopping time associated with a. Then we have

TfLp(),qfHp(),q(fHp(),q).

Theorem 7.2

Let p(⋅) ∈ Clog, 0 < pp+ < ∞, 0 < q ≤ ∞ and 0 < γ, s < ∞. If (4.1) holds, then

Uγ,sfp(),qCp(),qfHp(),q(fHp(),q).

Proof

Let us chose 0 < δ < 1 and 0 < ϵ < p. Instead of (4.2), we will show that

|Uγ,sa|δϵχ{τ=}p()/ϵCχ{τ<}p()/ϵχ{τ<}p()δϵ.

By (6.2),

Uγ,s(a)(x)χ{τ<}p()1KNμj=0K12(jK)(γ+s)χIK,μj,K(x)

for x ∈ {τk = ∞} and

|Uγ,sa|δϵχ{τ=}p()/ϵχ{τ<}p()δϵKNμj=0K12(jK)(γ+s)δϵχIK,μj,Kp()/ϵ=:Zχ{τ<}p()δϵ,

where

τ<=KNμIK,μ.

Choose max(1, δ p+) < r < ∞. By Lemma 2.2, there exists a function gL(p()ε) with g(p()ε)1 such that

Z=01KNμj=0K12(jK)(γ+s)δϵχIK,μj,KgdλKNμj=0K12(jK)(γ+s)δϵχIK,μj,KrδϵχIK,μj,Kg(rδϵ)KNμj=0K12(jK)(γ+s)δϵ×01χIK,μ1λ(IK,μj,K)IK,μj,Kg(rδϵ)dλ1/(rδϵ)dλ.

Then

Z01KNμχIK,μj=0K12(jK)(γ+s)δϵ(1/(rδϵ)+1/(rδϵ))×1λ(IK,μj,K)IK,μj,Kg(rδϵ)dλ1/(rδϵ)dλ,

and by Hölder’s inequality,

Z01KNμχIK,μ×j=0K12(jK)(γ+s)δϵ1λ(IK,μj,K)IK,μj,Kg(rδϵ)dλ1/(rδϵ)dλ01KNμχIK,μUγδϵ,sδϵ(3)(|g|(rδϵ))1/(rδϵ)dλKNμχIK,μp()/ϵUγδϵ,sδϵ(3)(|g|(rδϵ))1/(rδϵ)(p()/ϵ).

Since (r/δϵ)′ < (p(⋅)/ϵ)′, Theorem 4.3 implies that

Zχ{τ<}p()/ϵ,

whenever

1(p()/ϵ)/(r/δϵ)1(p()/ϵ)/(r/δϵ)+=r/(rδϵ)p+/(p+ϵ)r/(rδϵ)p/(pϵ)<(γ+s)δϵ.

This means that

p+ϵp+pϵp<(γ+s)δϵ,

in other words,

1p1p+<(γ+s)δ.

Since δ can be arbitrarily near to 1, we obtain (4.1). The proof of the theorem is complete.□

Corollary 7.1

Let p(⋅) ∈ Clog, 0 < pp+ < ∞, 0 < q ≤ ∞ and 0 < γ, s < ∞. If (4.1) holds, then

fHp(),qUγ,sfp(),qCp()fHp(),q(fHp(),q).

Dedicated to 80th anniversary of Professor Stefan Samko


Acknowledgements

This research was supported by the Hungarian National Research, Development and Innovation Office -- NKFIH, KH130426.

References

[1] J. Bergh and J. Löfström, Interpolation Spaces, an Introduction. Springer, Berlin (1976).10.1007/978-3-642-66451-9Search in Google Scholar

[2] D. Cruz-Uribe, L. Diening, and P. Hästö, The maximal operator on weighted variable Lebesgue spaces. Fract. Calc. Appl. Anal. 14, No 3 (2011), 361–374; 10.2478/s13540-011-0023-7; https://www.degruyter.com/journal/key/FCA/14/3/html.Search in Google Scholar

[3] D. Cruz-Uribe, A. Fiorenza, J. Martell, and C. Pérez, The boundedness of classical operators on variable Lp spaces. Ann. Acad. Sci. Fenn. Math. 31 (2006), 239–264.Search in Google Scholar

[4] D. Cruz-Uribe, A. Fiorenza, and C. Neugebauer, The maximal function on variable Lp spaces. Ann. Acad. Sci. Fenn. Math. 28 (2003), 223–238.Search in Google Scholar

[5] D. Cruz-Uribe and D. Wang, Variable Hardy spaces. Indiana Univ. Math. J., 63, No 2 (2014), 447–493.10.1512/iumj.2014.63.5232Search in Google Scholar

[6] D. V. Cruz-Uribe and A. Fiorenza, Variable Lebesgue Spaces. Foundations and Harmonic Analysis. Birkhäuser/Springer, New York (2013).10.1007/978-3-0348-0548-3Search in Google Scholar

[7] L. Diening, P. Harjulehto, P. Hästö, and M. Ružička, Lebesgue and Sobolev Spaces with Variable Exponents. Springer, Berlin (2011).10.1007/978-3-642-18363-8Search in Google Scholar

[8] L. Ephremidze, V. Kokilashvili, and S. Samko, Fractional, maximal and singular operators in variable exponent Lorentz spaces. Fract. Calc. Appl. Anal. 11, No 4 (2008), 407–420.Search in Google Scholar

[9] Z. Hao and Y. Jiao, Fractional integral on martingale Hardy spaces with variable exponents. Fract. Calc. Appl. Anal. 18, No 5 (2015), 1128–1145; 10.1515/fca-2015-0065; https://www.degruyter.com/journal/key/FCA/18/5/html.Search in Google Scholar

[10] K.-P. Ho, Atomic decompositions of weighted Hardy spaces with variable exponents. Tohoku Math. J. (2) 69, No 3 (2017), 383–413.10.2748/tmj/1505181623Search in Google Scholar

[11] Y. Jiao, F. Weisz, L. Wu, and D. Zhou, Variable martingale Hardy spaces and their applications in Fourier analysis. Dissertationes Math. 550 (2020), 1–67.10.4064/dm807-12-2019Search in Google Scholar

[12] Y. Jiao, D. Zhou, Z. Hao, and W. Chen, Martingale Hardy spaces with variable exponents. Banach J. Math. 10 (2016), 750–770.10.1215/17358787-3649326Search in Google Scholar

[13] Y. Jiao, D. Zhou, F. Weisz, and Z. Hao, Corrigendum: Fractional integral on martingale Hardy spaces with variable exponents. Fract. Calc. Appl. Anal. 20, No 4 (2017), 1051–1052; 10.1515/fca-2017-0055; https://www.degruyter.com/journal/key/FCA/20/4/html.Search in Google Scholar

[14] Y. Jiao, Y. Zuo, D. Zhou, and L. Wu, Variable Hardy-Lorentz spaces Hp(⋅),q(ℝn). Math. Nachr. 292 (2019), 309–349.10.1002/mana.201700331Search in Google Scholar

[15] H. Kempka and J. Vybíral, Lorentz spaces with variable exponents. Math. Nachr. 287 (2014), 938–954.10.1002/mana.201200278Search in Google Scholar

[16] V. Kokilashvili, A. Meskhi, H. Rafeiro, and S. Samko, Integral Operators in Non-standard Function Spaces. Volume 1: Variable Exponent Lebesgue and Amalgam Spaces. Birkhäuser/Springer, Basel (2016).10.1007/978-3-319-21015-5_1Search in Google Scholar

[17] V. Kokilashvili, A. Meskhi, H. Rafeiro, and S. Samko, Integral Operators in Non-standard Function Spaces. Volume 2: Variable Exponent Hölder, Morrey-Campanato and Grand Spaces. Birkhäuser/Springer, Basel (2016).Search in Google Scholar

[18] O. Kováčik and J. Rákosník, On spaces Lp(x) and Wk,p(x). Czech. Math. J. 41, No 4 (1991), 592–618.10.21136/CMJ.1991.102493Search in Google Scholar

[19] J. Liu, F. Weisz, D. Yang, and W. Yuan, Variable anisotropic Hardy spaces and their applications. Taiwanese J. Math. 22 (2018), 1173–1216.10.11650/tjm/171101Search in Google Scholar

[20] J. Liu, F. Weisz, D. Yang, and W. Yuan, Littlewood-Paley and finite atomic characterizations of anisotropic variable Hardy-Lorentz spaces and their applications. J. Fourier Anal. Appl. 25 (2019), 874–922.10.1007/s00041-018-9609-3Search in Google Scholar

[21] G. Lorentz, Some new functional spaces. Ann. of Math. 51 (1950), 37–55.10.1007/978-1-4612-5329-7_58Search in Google Scholar

[22] E. Nakai and Y. Sawano, Hardy spaces with variable exponents and generalized Campanato spaces. J. Funct. Anal. 262, No 9 (2012), 3665–3748.10.1016/j.jfa.2012.01.004Search in Google Scholar

[23] A. Nekvinda, Hardy-Littlewood maximal operator on Lp(x)(ℝ). Math. Inequal. Appl. 7 (2004), 255–265.Search in Google Scholar

[24] H. Rafeiro, N. Samko, and S. Samko, Morrey-Campanato spaces: an overview. In: Operator Theory, Pseudo-Differential Equations, and Mathematical Physics. The Vladimir Rabinovich Anniversary Volume. Birkhäuser/Springer, Basel (2013), 293–323.10.1007/978-3-0348-0537-7_15Search in Google Scholar

[25] H. Rafeiro and S. Samko, Fractional integrals and derivatives: mapping properties. Fract. Calc. Appl. Anal. 19, No 3 (2016), 580–607; 10.1515/fca-2016-0032; https://www.degruyter.com/journal/key/FCA/19/3/html.Search in Google Scholar

[26] H. Rafeiro and S. Samko, A note on vanishing Morrey → VMO result for fractional integrals of variable order. Fract. Calc. Appl. Anal. 23, No 1 (2020), 298–302; 10.1515/fca-2020-0013; https://www.degruyter.com/journal/key/FCA/23/1/html;.Search in Google Scholar

[27] H. Rafeiro and M. Yakhshiboev. The Chen-Marchaud fractional integro-differentiation in the variable exponent Lebesgue spaces. Fract. Calc. Appl. Anal. 14, No 3 (2011), 343–360; 10.2478/s13540-011-0022-8; https://www.degruyter.com/journal/key/FCA/14/3/html.Search in Google Scholar

[28] S. Samko, On a progress in the theory of Lebesgue spaces with variable exponent: maximal and singular operators. Integral Transforms Spec. Funct. 16, No 5-6 (2005), 461–482.10.1080/10652460412331320322Search in Google Scholar

[29] S. Samko, On some classical operators of variable order in variable exponent spaces. In: Analysis, Partial Differential Equations and Applications. The Vladimir Maz’ya Anniversary Volume. Selected Papers of the International Workshop, Rome, Italy, June 30–July 3, 2008. Birkhäuser, Basel (2009), 281–301.10.1007/978-3-7643-9898-9_20Search in Google Scholar

[30] S. Samko, Fractional integration and differentiation of variable order: an overview. Nonlinear Dyn. 71, No 4 (2013), 653–662.10.1007/s11071-012-0485-0Search in Google Scholar

[31] Y. Sawano. Atomic decompositions of Hardy spaces with variable exponents and its application to bounded linear operators. Integral Equ. Oper. Theory. 77 (2013), 123–148.10.1007/s00020-013-2073-1Search in Google Scholar

[32] F. Schipp, W. R. Wade, P. Simon, and J. Pál, Walsh Series: An Introduction to Dyadic Harmonic Analysis. Adam Hilger, Bristol-New York (1990).Search in Google Scholar

[33] K. Szarvas and F. Weisz, The boundedness of the Cesaro- and Riesz means in variable dyadic Hardy spaces. Banach J. Math. Anal., 13 (2019), 675–696.10.1215/17358787-2018-0037Search in Google Scholar

[34] F. Weisz, Martingale Hardy Spaces and Their Applications in Fourier Analysis, Volume 1568 of Lecture Notes in Math. Springer, Berlin (1994).10.1007/BFb0073448Search in Google Scholar

[35] F. Weisz, Boundedness of dyadic maximal operators on variable Lebesgue spaces. Adv. Oper. Theory 5 (2020), 1588–1598.10.1007/s43036-020-00071-9Search in Google Scholar

[36] F. Weisz, Doob’s and Burkholder-Davis-Gundy inequalities with variable exponent. Proc. Amer. Math. Soc. 149 (2021) 875–888.10.1090/proc/15262Search in Google Scholar

[37] G. Xie, Y. Jiao, and D. Yang, Martingale Musielak-Orlicz Hardy spaces. Sci. China, Math. 62, No 8 (2019), 1567–1584.10.1007/s11425-017-9237-3Search in Google Scholar

[38] G. Xie, F. Weisz, D. Yang, and Y. Jiao, New martingale inequalities and applications to Fourier analysis. Nonlinear Analysis 182 (2019), 143–192.10.1016/j.na.2018.12.011Search in Google Scholar

[39] G. Xie and D. Yang, Atomic characterizations of weak martingale Musielak-Orlicz Hardy spaces and their applications. Banach J. Math. Anal. 13, No 4 (2019), 884–917.10.1215/17358787-2018-0050Search in Google Scholar

[40] X. Yan, D. Yang, W. Yuan, and C. Zhuo, Variable weak Hardy spaces and their applications. J. Funct. Anal. 271 (2016), 2822–2887.10.1016/j.jfa.2016.07.006Search in Google Scholar

Received: 2020-06-12
Published Online: 2021-05-09
Published in Print: 2021-04-27

© 2021 Diogenes Co., Sofia

This work is licensed under the Creative Commons Attribution 4.0 International License.

Downloaded on 26.4.2024 from https://www.degruyter.com/document/doi/10.1515/fca-2021-0018/html
Scroll to top button