Let H→pA(Rn) be the anisotropic mixed-norm Hardy space, where →p∈(0,∞)n and A is a general expansive matrix on Rn. In this paper, a general summability method, the so-called θ-summability is considered for multi-dimensional Fourier transforms in H→pA(Rn). Precisely, the author establishes the boundedness of maximal operators, induced by the so-called θ-means, from H→pA(Rn) to the mixed-norm Lebesgue space L→p(Rn). As applications, some norm and almost everywhere convergence results of the θ-means are presented. Finally, the corresponding conclusions of two well-known specific summability methods, namely, Bochner–Riesz and Weierstrass means, are also obtained.
Citation: Nan Li. Summability in anisotropic mixed-norm Hardy spaces[J]. Electronic Research Archive, 2022, 30(9): 3362-3376. doi: 10.3934/era.2022171
[1] | Mohammed Shehu Shagari, Faryad Ali, Trad Alotaibi, Akbar Azam . Fixed point of Hardy-Rogers-type contractions on metric spaces with graph. Electronic Research Archive, 2023, 31(2): 675-690. doi: 10.3934/era.2023033 |
[2] | Xintao Li, Rongrui Lin, Lianbing She . Periodic measures for a neural field lattice model with state dependent superlinear noise. Electronic Research Archive, 2024, 32(6): 4011-4024. doi: 10.3934/era.2024180 |
[3] | Francisco Javier García-Pacheco, María de los Ángeles Moreno-Frías, Marina Murillo-Arcila . On absolutely invertibles. Electronic Research Archive, 2024, 32(12): 6578-6592. doi: 10.3934/era.2024307 |
[4] | Long Huang, Xiaofeng Wang . Schur's test, Bergman-type operators and Gleason's problem on radial-angular mixed spaces. Electronic Research Archive, 2023, 31(10): 6027-6044. doi: 10.3934/era.2023307 |
[5] | Wenya Qi, Padmanabhan Seshaiyer, Junping Wang . A four-field mixed finite element method for Biot's consolidation problems. Electronic Research Archive, 2021, 29(3): 2517-2532. doi: 10.3934/era.2020127 |
[6] | Lili Li, Boya Zhou, Huiqin Wei, Fengyan Wu . Analysis of a fourth-order compact $ \theta $-method for delay parabolic equations. Electronic Research Archive, 2024, 32(4): 2805-2823. doi: 10.3934/era.2024127 |
[7] | Fang Wang, Weiguo Li, Wendi Bao, Li Liu . Greedy randomized and maximal weighted residual Kaczmarz methods with oblique projection. Electronic Research Archive, 2022, 30(4): 1158-1186. doi: 10.3934/era.2022062 |
[8] | Jia-Min Luo, Hou-Biao Li, Wei-Bo Wei . Block splitting preconditioner for time-space fractional diffusion equations. Electronic Research Archive, 2022, 30(3): 780-797. doi: 10.3934/era.2022041 |
[9] | Jon Johnsen . Well-posed final value problems and Duhamel's formula for coercive Lax–Milgram operators. Electronic Research Archive, 2019, 27(0): 20-36. doi: 10.3934/era.2019008 |
[10] | Xinfeng Ge, Keqin Su . Stability of thermoelastic Timoshenko system with variable delay in the internal feedback. Electronic Research Archive, 2024, 32(5): 3457-3476. doi: 10.3934/era.2024160 |
Let H→pA(Rn) be the anisotropic mixed-norm Hardy space, where →p∈(0,∞)n and A is a general expansive matrix on Rn. In this paper, a general summability method, the so-called θ-summability is considered for multi-dimensional Fourier transforms in H→pA(Rn). Precisely, the author establishes the boundedness of maximal operators, induced by the so-called θ-means, from H→pA(Rn) to the mixed-norm Lebesgue space L→p(Rn). As applications, some norm and almost everywhere convergence results of the θ-means are presented. Finally, the corresponding conclusions of two well-known specific summability methods, namely, Bochner–Riesz and Weierstrass means, are also obtained.
Let A be a general expansive matrix on Rn. In 2003, Bownik [1] investigated the anisotropic Hardy space HpA(Rn) with p∈(0,∞), which includes both the classical Hardy space and the parabolic Hardy space of Calderón and Torchinsky [2] as special cases. Recently, Huang et al. [3] introduced the anisotropic mixed-norm Hardy space H→pA(Rn) with respect to →p∈(0,∞)n and a general expansive matrix A, and established its various real-variable characterizations. This extends the real-variable theory of the Hardy space HpA(Rn) from [1]. For more information on mixed-norm function spaces, we refer the reader to [4,5,6,7,8,9,10,11].
On the other hand, it is well known that Stein, Taibleson and Weiss [12] proved for the Bochner–Riesz summability that the maximal operator σθ∗ of the θ-means is bounded from the classical Hardy Hp(Rn) to the Lebesgue space Lp(Rn) with the index p greater than some constant p0. This result has been extended to many other Hardy-type and other summability methods. For more progress about this topic, we refer the reader to [13,14,15,16,17,18] and references therein. In particular, Weisz [18] proved that the maximal operator, induced by the so-called θ-means, is bounded from the isotropic mixed-norm Hardy space H→p(Rn) to the mixed-norm Lebesgue space L→p(Rn). However, the corresponding conclusion of summability in anisotropic mixed-norm Hardy space H→pA(Rn) is still unknown.
In this paper, under some conditions on θ and →p, we prove that the maximal operator σθ∗ is bounded from H→pA(Rn) to L→p(Rn). As a consequence, we prove some norm and almost everywhere convergence results for the θ-means. Moreover, sa special cases of the θ-means, we consider the well-known Bochner–Riesz and Weierstrass summations. This paper is organized as follows: As a preliminary, in Section 2, we recall some definitions of expansive matrices, mixed-norm Lebesgue spaces L→p(Rn) and anisotropic mixed-norm Hardy spaces H→pA(Rn). In Section 3, via borrowing some ideas from [18,Theorem 3] and [13,Theorem 7.4] as well as [14,Theorem 2.17], we prove our main result by using the known finite atomic characterization of H→pA(Rn) and a criterion on the boundedness of sublinear operators from H→pA(Rn) into L→p(Rn). Section 4 is aimed to consider two special summability methods, namely, the Bochner–Riesz and Weierstrass summations.
Finally, we make some conventions on notation. Let N:={1,2,…}, Z+:={0}∪N and 0 be the origin of Rn. For any γ:=(γ1,…,γn)∈(Z+)n=:Zn+, let |γ|:=γ1+⋯+γn and ∂γ:=(∂∂x1)γ1⋯(∂∂xn)γn. We use C to denote a positive constant which is independent of the main parameters, but its value may change from line to line. In addition, we use f≲g to denote f≤Cg and, if f≲g≲f, we then write f∼g. Moreover, for a given set Ω⊂Rn, we denote its characteristic function by 1Ω, the set Rn∖Ω by Ω∁ and its n-dimensional Lebesgue measure by |Ω|. For any t∈R, The symbol ⌊t⌋ denotes the largest integer not greater than t. For each r∈[1,∞], we denote by r′ its conjugate index, namely, 1/r+1/r′=1. Moreover, if →r:=(r1,…,rn)∈[1,∞]n, we denote by →r′:=(r′1,…,r′n) its conjugate index.
In this section, we recall the notions of expansive matrices, mixed-norm Lebesgue spaces and anisotropic mixed-norm Hardy spaces.
We begin with the following notion of mixed-norm Lebesgue spaces from [19].
Definition 2.1. Let →p:=(p1,…,pn)∈(0,∞]n. The mixed-norm Lebesgue space L→p(Rn) is defined to be the set of all measurable functions f such that
‖f‖L→p(Rn):={∫R⋯[∫R|f(x1,…,xn)|p1dx1]p2p1⋯dxn}1pn<∞ |
with the usual modifications made when pi=∞ for some i∈{1,…,n}.
Recall also that the notions of expansive matrices and homogeneous quasi-norms were originally introduced by Bownik in [1].
Definition 2.2. A real n×n matrix A is called an expansive matrix (shortly, a dilation) if
minλ∈σ(A)|λ|>1, |
here and thereafter, σ(A) denotes the collection of all eigenvalues of A.
Definition 2.3. Let A be a dilation. A measurable mapping ρ: Rn→[0,∞) is called a homogeneous quasi-norm, associated with A, if
(i) x≠0 implies that ρ(x)∈(0,∞);
(ii) for each x∈Rn, ρ(Ax)=bρ(x), here and below, b:=|detA|;
(iii) for any x, y∈Rn, ρ(x+y)≤c[ρ(x)+ρ(y)], where c is a positive constant independent of x and y.
For any given dilation A, it was proved in [1,Lemma 2.2] that there exists an open set Δ⊂Rn which has the following property: |Δ|=1, and we can find a constant τ∈(1,∞) such that Δ⊂τΔ⊂AΔ. For any i∈Z, we define Bi:=AiΔ. It is easy to check that {Bi}i∈Z is a family of open sets around the origin, Bi⊂τBi⊂Bi+1 and |Bi|=bi. For any given dilation A, we use the symbol B to denote the set of all dilated balls, namely,
B:={x+Bi: x∈Rn, i∈Z} | (2.1) |
and
ω:=inf{k∈Z: τk≥2}. | (2.2) |
By [1,Lemma 2.4], we know that any two homogeneous quasi-norms associated with the same fixed dilation A are equivalent. Thus, in what follows, we always use the step homogeneous quasi-norm defined by setting, for each x∈Rn,
ρ(x):={biwhenx∈Bi+1∖Bi0whenx=0 |
for convenience. Let λ−, λ+∈(1,∞) be two numbers such that
λ−≤min{|λ|: λ∈σ(A)}≤max{|λ|: λ∈σ(A)}≤λ+. | (2.3) |
Throughout this article, the symbol S(Rn) denotes the space of all Schwartz functions, namely, the set of all C∞(Rn) functions ϕ satisfying that, for any k∈Z+ and multi-index β∈Zn+,
‖ϕ‖β,k:=supx∈Rn[ρ(x)]k|∂βϕ(x)|<∞. |
The topology of S(Rn) is determined by {‖⋅‖β,k}β∈Zn+,k∈Z+. Moreover, we use S′(Rn) to denote the dual space of S(Rn), namely, the space of all tempered distributions on Rn equipped with the weak-∗ topology. For any N∈Z+, let SN(Rn) denote the following set:
{ϕ∈S(Rn):‖ϕ‖SN(Rn):=supβ∈Zn+,|β|≤Nsupx∈Rn[|∂βϕ(x)|max{1,[ρ(x)]N}]≤1}. |
For an n-dimensional vector →p:=(p1,…,pn)∈(0,∞]n, let
p−:=min{p1,…,pn},p+:=max{p1,…,pn}andp_∈(0,min{p−,1}). | (2.4) |
The following definition of anisotropic mixed-norm Hardy spaces was first introduced by Huang et. al [3].
Definition 2.4. (i) Let ϕ∈S(Rn) and f∈S′(Rn). The non-tangential maximal function Mϕ(f) with respect to ϕ is defined by setting, for any x∈Rn,
Mϕ(f)(x):=supy∈x+Bk,k∈Z|f∗ϕk(y)|, |
here and thereafter, for any ϕ∈S(Rn) and k∈Z, let ϕk(⋅):=bkϕ(Ak⋅). Moreover, for any given N∈N, the non-tangential grand maximal function MN(f) of f∈S′(Rn) is defined by setting, for any x∈Rn,
MN(f)(x):=supϕ∈SN(Rn)Mϕ(f)(x). |
(ii) Let →p∈(0,∞)n and N∈N∩[⌊(1min{1,p−}−1)lnblnλ−⌋+2,∞), where p− is as in (2.4). The anisotropic mixed-norm Hardy space H→pA(Rn) is defined by setting
H→pA(Rn):={f∈S′(Rn): MN(f)∈L→p(Rn)} |
and, for any f∈H→pA(Rn), let ‖f‖H→pA(Rn):=‖MN(f)‖L→p(Rn).
Remark 1. (i) Observe that, in [3,Theorem 4.7], it was proved that the space H→pA(Rn) is independent of the choice of N as in Definition 2.4(ⅱ).
(ii) Note that Cleanthous et al. [20] investigated another kind of anisotropic mixed-norm Hardy space H→p→a(Rn), where →a∈[1,∞)n and →p∈(0,∞)n; see [20,Definition 3.3]. We should point out that [3,Proposition 4] shows that, if
A:=(2a10⋯002a2⋯0⋮⋮⋮00⋯2an) | (2.5) |
with →a:=(a1,…,an)∈[1,∞)n, then the Hardy space H→pA(Rn) in Definition 2.4(ⅱ) and the anisotropic mixed-norm Hardy space H→p→a(Rn) from [20] coincide with equivalent quasi-norms.
In this section, we study the so-called θ-summability for multi-dimensional Fourier transforms in the anisotropic mixed-norm Hardy space H→pA(Rn).
We always use L1loc(Rn) to denote the set of all locally integrable functions on Rn. To present our main result, we need several technical lemmas as follows. First, the following Lemmas 3.1, 3.2 and 3.3 are just, respectively, [3,Lemmas 3.4,3.2 and 4.4], which show some properties of the L→p(Rn) quasi-norm, the boundedness of the anisotropic Hardy–Littlewood maximal operator MHL on the space L→p(Rn) and the anisotropic Fefferman–Stein vector-valued inequality on L→p(Rn).
Lemma 3.1. Let →p∈(0,∞]n. Then, for any s∈(0,∞) and f∈L→p(Rn),
‖|f|s‖L→p(Rn)=‖f‖sLs→p(Rn), |
here and below, for any γ∈R, γ→p:=(γp1,…,γpn).In addition, for any λ∈C, r∈[0,min{1,p−}]with p− as in (2.4) and f, h∈L→p(Rn), ‖λf‖L→p(Rn)=|λ|‖f‖L→p(Rn) and
‖f+h‖rL→p(Rn)≤‖f‖rL→p(Rn)+‖h‖rL→p(Rn). |
Lemma 3.2. Let →p∈(1,∞)n. Then there exists a positiveconstant C such that, for any f∈L1loc(Rn),
‖MHL(f)‖L→p(Rn)≤C‖f‖L→p(Rn), |
where MHL denotes the anisotropic Hardy–Littlewood maximal operatordefined by setting, for any f∈L1loc(Rn) and x∈Rn,
MHL(f)(x):=supk∈Zsupy∈x+Bk1|Bk|∫y+Bk|f(z)|dz=supx∈B∈B1|B|∫B|f(z)|dz |
with B as in (2.1).
Lemma 3.3. Let →p∈(1,∞)n and v∈(1,∞]. Then, for any sequence {fk}k∈N of measurable functions,
‖{∑k∈N[MHL(fk)]v}1/v‖L→p(Rn)≤C‖(∑k∈N|fk|v)1/v‖L→p(Rn) |
with the usual modification made when v=∞, where C is a positive constant independent of {fk}k∈N.
Let →p∈(0,∞)n and k∈Z+. Then, by the fact that, for any dilated ball B∈B and q∈(0,p_), 1AkB≤bk/q[MHL(1B)]1/q as well as Lemmas 3.1 and 3.3, we conclude that there exists a positive constant C such that, for any sequence {B(i)}i∈N⊂B,
‖∑i∈N1AkB(i)‖L→p(Rn)≤C‖∑i∈Nbk/q[MHL(1B(i))]1/q‖L→p(Rn)=Cbk/q‖{∑i∈N[MHL(1B(i))]1/q}q‖1/qL→p/q(Rn)≤Cbk/q‖∑i∈N1B(i)‖L→p(Rn). | (3.1) |
The last inequality used Lemma 3.3 with the fact that →p/q∈(1,∞)n.
For any given p∈(0,∞] and measurable set Ω⊂Rn. The Lebesgue space Lp(Ω) is defined to be the set of all measurable functions f on Ω such that, when p∈(0,∞),
‖f‖Lp(Ω):=[∫Ω|f(x)|pdx]1/p<∞ |
and
‖f‖L∞(Ω):=esssupx∈Ω|f(x)|<∞. |
We also need the following notions of anisotropic mixed-norm (→p,r,s)-atoms and anisotropic mixed-norm finite atomic Hardy spaces H→p,r,sA,fin(Rn) from [3].
Definition 3.4. Let →p∈(0,∞)n, r∈(1,∞] and
s∈[⌊(1p−−1)lnblnλ−⌋,∞)∩Z+ | (3.2) |
with p− as in (2.4).
(I) A measurable function a on Rn is called an anisotropic mixed-norm (→p,r,s)-atom if
(I)1 suppa⊂B with some B∈B, where B is as in (2.1);
(I)2 ‖a‖Lr(Rn)≤|B|1/r‖1B‖L→p(Rn);
(I)3 for any α∈Zn+ with |α|≤s, ∫Rna(x)xαdx=0.
(II) The anisotropic mixed-norm finite atomic Hardy space H→p,r,sA,fin(Rn) is defined to be the set of all f∈S′(Rn) satisfying that there exist I∈N, {λi}i∈[1,I]∩N⊂C and a finite sequence of (→p,r,s)-atoms, {ai}i∈[1,I]∩N, supported, respectively, in {B(i)}i∈[1,I]∩N⊂B such that
f=I∑i=1λiaiinS′(Rn). |
Moreover, for any f∈H→p,r,sA,fin(Rn), let
‖f‖H→p,r,sA,fin(Rn):=inf‖{I∑i=1[|λi|1B(i)‖1B(i)‖L→p(Rn)]p_}1/p_‖L→p(Rn), |
where p_ is as in (2.4) and the infimum is taken over all decompositions of f as above.
Then, from [3,Theorem 8.1(ⅰ) and Remark 13], we immediately deduce the succeeding criterion on the boundedness of sublinear operators from H→pA(Rn) into L→p(Rn), which plays a key role in the proof of our main result.
Lemma 3.5. Let →p∈(0,∞)n, r∈(max{p+,1},∞)with p+ as in (2.4) and s be as in (3.2).Assume that T: H→p,r,sA,fin(Rn)→L→p(Rn) is a sublinear operator satisfying thatthere exists a positive constant C such that, for any f∈H→p,r,sA,fin(Rn),
‖T(f)‖L→p(Rn)≤C‖f‖H→p,r,sA,fin(Rn). |
Then T uniquely extends to a bounded sublinearoperator from H→pA(Rn) into L→p(Rn).
Recall that, for any given p∈[1,2] and any f∈Lp(Rn), the Fourier inversion formula, namely,
f(x)=∫Rnˆf(t)e2πıx⋅tdt,∀ x∈Rn, |
holds true if ˆf∈L1(Rn), here and below, ı:=√−1, x⋅t:=∑nk=1xktk for any x:=(x1,…,xn), t:=(t1,…,tn)∈Rn, and ˆf denotes the Fourier transform of f, which is defined by setting, for any t∈Rn,
ˆf(t):=∫Rnf(x)e−2πıx⋅tdx. |
This motivates the following definition of θ-summability of Fourier transforms; see, for instance, [15,16,17,18,21] for the classical case and [13,14] for the anisotropic case. We always suppose that
θ∈C0(R),θ(|⋅|)∈L1(Rn),θ(0)=1andθ is even, | (3.3) |
where the symbol C0(R) denotes the set of all continuous functions f satisfying that lim|x|→∞|f(x)|=0. Let A be a given dilation, m∈Z and p∈[1,2]. The m-th anisotropic θ-mean, denoted by σθm, is defined by setting, for any f∈Lp(Rn) and x∈Rn,
σθmf(x):=∫Rnθ(|(A∗)−mu|)ˆf(u)e2πıx⋅udu, | (3.4) |
where A∗ be the transposed matrix of A. This integral is well defined because θ∈Lp(R) with p∈[1,2] and ˆf∈Lp′(Rn). Let θ0(x):=θ(|x|) for any x∈Rn and assume that
^θ0∈L1(Rn). | (3.5) |
Moreover, by [14,(2.17)], we can rewrite σθmf(x) in (3.4) as
σθmf(x)=bm∫Rnf(t)^θ0(Am(x−t))dt. |
This definition of anisotropic θ-means can be extended to any f∈L→p(Rn) with p−∈[1,∞) by setting, for any x∈Rn,
σθmf(x):=bm∫Rnf(x−t)^θ0(Amt)dt, | (3.6) |
where m∈Z. Furthermore, (3.6) induces the definition of maximal θ-operators σθ∗ as follows: for any f∈L→p(Rn) with p−∈[1,∞),
σθ∗f:=supm∈Z|σθmf|. |
Now we state the main result of this paper as follows.
Theorem 3.6. Let θ and θ0 be, respectively, as in (3.3) and (3.5) satisfying that there exists a positive constant β∈(1,∞) such that, for anyγ∈Zn+ and x∈Rn∖{0},
|∂γ^θ0(x)|≤C|x|−β, | (3.7) |
where C is a positive constant independent of x.If →p∈(0,∞)n,
β∈(lnblnλ−,∞)andp−∈(lnbβlnλ−,∞) | (3.8) |
with λ− as in (2.3), then there exists a positive constant C(p−,p+), with p− and p+ as in (2.4), such that, for any f∈H→pA(Rn),
‖σθ∗f‖L→p(Rn)≤C(p−,p+)‖f‖H→pA(Rn). |
Proof. By Lemma 3.5, to prove Theorem 3.6, we only need to show that, for any f∈H→p,r,sA,fin(Rn),
‖σθ∗f‖Lp(⋅)(Rn)≲‖f‖H→p,r,sA,fin(Rn) | (3.9) |
with s being as in (3.2) large enough and r∈(max{p+,1},∞) to be chosen later, where p+ is as in (2.4).
For this purpose, suppose now f∈H→p,r,sA,fin(Rn). Then, by Definition 3.4(Ⅱ), we find that there exist some I∈N, {λi}i∈[1,I]∩N⊂C and a finite sequence of (→p,r,s)-atoms, {ai}i∈[1,I]∩N, supported, respectively, in {B(i)}i∈[1,I]∩N⊂B such that f=∑Ii=1λiai in S′(Rn) and
‖f‖H→p,r,sA,fin(Rn)∼‖{I∑i=1[|λi|1B(i)‖1B(i)‖L→p(Rn)]p_}1/p_‖L→p(Rn), | (3.10) |
where p_ is as in (2.4). Take two sequences {xi}i∈[1,I]∩N⊂Rn and {ki}i∈[1,I]∩N⊂Z such that, for any i∈[1,I]∩N, xi+Bki=B(i). Then
‖σθ∗f‖L→p(Rn)≲‖I∑i=1|λi|σθ∗(ai)1xi+AωBki‖L→p(Rn)+‖I∑i=1|λi|σθ∗(ai)1(xi+AωBki)∁‖L→p(Rn)=:J1+J2, | (3.11) |
where ω is as in (2.2).
For J1, take u∈L(→p/p_)′(Rn) satisfying that ‖u‖L(→p/p_)′(Rn)≤1 and
‖I∑i=1|λi|p_[σθ∗(ai)]p_1xi+AωBki‖L→p/p_(Rn)=∫RnI∑i=1|λi|p_[σθ∗(ai)(x)]p_1xi+AωBki(x)u(x)dx. |
From this and the Hölder inequality, it follows that, for any t∈(1,∞) with p+<tp_<r,
(J1)p_≲‖I∑i=1|λi|p_[σθ∗(ai)]p_1xi+AωBki‖L→p/p_(Rn)∼∫RnI∑i=1|λi|p_[σθ∗(ai)(x)]p_1xi+AωBki(x)u(x)dx.≲I∑i=1|λi|p_‖[σθ∗(ai)]p_1xi+AωBki‖Lt(Rn)‖1xi+AωBkiu‖Lt′(Rn)≲I∑i=1|λi|p_‖σθ∗(ai)‖p_Lr(Rn)‖1xi+AωBki‖1/tLrr−tp_(Rn)‖1xi+AωBkiu‖Lt′(Rn). |
By this, the fact that σθ∗ is bounded on Lq(Rn) for any 1<q<∞, and Definition 3.4(Ⅰ), we further conclude that
(J1)p_≲I∑i=1|λi|p_‖1xi+Bki‖−p_L→p(Rn)|AωBki|p_/r|AωBki|r−tp_rt‖1xi+AωBkiu‖Lt′(Rn)∼I∑i=1|λi|p_‖1xi+Bki‖−p_L→p(Rn)|AωBki|1/t‖1xi+AωBkiu‖Lt′(Rn)∼I∑i=1|λi|p_‖1xi+Bki‖−p_L→p(Rn)|AωBki|[1|AωBki|∫xi+AωBki[u(x)]t′dx]1/t′≲I∑i=1|λi|p_‖1xi+Bki‖−p_L→p(Rn)∫Rn1xi+AωBki(x)[MHL(ut′)(x)]1/t′dx≲‖I∑i=1|λi|p_‖1xi+Bki‖−p_L→p(Rn)1xi+AωBki‖L→p/p_(Rn)‖[MHL(ut′)]1/t′‖L(→p/p_)′(Rn). |
Note that the assumption 0<p+/p_<t implies that t′<→p/p_)′≤∞. From this, Lemmas 3.2 and 3.1, the fact that ‖u‖L(→p/p_)′(Rn)≤1, (3.1) and (3.10), we deduce that
J1≲‖I∑i=1|λi|p_‖1xi+Bki‖−p_L→p(Rn)1xi+Bki‖1/p_L→p/p_(Rn)‖u‖1/p_L(→p/p_)′(Rn)∼‖{I∑i=1[|λi|1xi+Bki‖1xi+Bki‖L→p(Rn)]p_}1/p_‖L→p(Rn)∼‖f‖H→p,r,sA,fin(Rn). | (3.12) |
We next deal with J2. To do this, we first claim that, for any i∈[1,I]∩N and x∈(xi+AωBki)∁,
σθ∗(ai)(x)≲‖1xi+Bki‖−1L→p(Rn)[MHL(1xi+Bki)(x)]βlnλ−/lnb, | (3.13) |
where β is as in (3.7) and (3.8). Assume that (3.13) holds true for the time being. Then, by (3.8), Lemmas 3.1 and 3.3 as well as (3.10), we find that
J2≲‖I∑i=1|λi|‖1xi+Bki‖−1L→p(Rn)[MHL(1xi+Bki)(x)]βlnλ−/lnb1(xi+AωBki)∁‖L→p(Rn)≲‖I∑i=1[|λi|lnb/(βlnλ−)‖1xi+Bki‖−lnb/(βlnλ−)L→p(Rn)MHL(1xi+Bki)]βlnλ−/lnb‖L→p(Rn)≲‖[I∑i=1|λi|‖1xi+Bki‖−1L→p(Rn)1xi+Bki]lnb/(βlnλ−)‖βlnλ−/lnbL→pβlnλ−/lnb(Rn)≲‖{I∑i=1[|λi|1xi+Bki‖1xi+Bki‖L→p(Rn)]p_}1/p_‖L→p(Rn)∼‖f‖H→p,r,sA,fin(Rn). |
This, together with (3.11) and (3.12), further implies that (3.9) holds true.
Thus, to complete the proof of Theorem 3.6, it suffices to verify (3.13). To this end, let a be any (→p,r,s)-atom supported in x0+Bk∈B, where x0∈Rn, k∈Z and B is as in (2.1). Without loss of generality, we may assume that x0=0. Suppose that P is a polynomial of degree not more than s. Then, by (3.6), Definition 3.4(Ⅰ) and the Hölder inequality, we obtain that, for any m∈Z and x∈(Bk+ω)∁,
|σθma(x)|=bm|∫Bka(t)^θ0(Am(x−t))dt|=bm|∫Bka(t)[^θ0(Am(x−t))−P(Am(x−t))]dt|≤bm‖a‖Lr(Rn)[∫Bk|^θ0(Am(x−t))−P(Am(x−t))|r′dt]1/r′≤bmbk/r‖1Bk‖−1L→p(Rn)b−m/r′[∫Amx+Bk+m|^θ0(y)−P(y)|r′dy]1/r′≤bmbk/r‖1Bk‖−1L→p(Rn)b−m/r′b(k+m)/r′supy∈Amx+Bk+m|^θ0(y)−P(y)|≤bk+m‖1Bk‖−1L→p(Rn)supy∈Amx+Bk+m|^θ0(y)−P(y)|. | (3.14) |
Assume that x∈Bk+ω+ν+1∖Bk+ω+ν for some ν∈Z+. Then, using [1,(2.11)], we have
Amx+Bk+m⊂Ak+m(Bω+ν+1∖Bω+ν+B0)⊂Ak+m(B2ω+ν+1∖Bν). | (3.15) |
In addition, the Taylor remainder theorem implies that
supy∈Amx+Bk+m|^θ0(y)−P(y)|≲supt,˜t∈Bk+msupγ∈Zn+,|γ|≤N|∂γ^θ0(Amx+˜t)||t|N, | (3.16) |
where N∈[0,s+1]∩Z+.
For the case when k+m∈Z∖Z+ and k+m+ν∈Z+, by (3.16), [1,(2.2)] and (3.7), it is easy to see that
supy∈Amx+Bk+m|^θ0(y)−P(y)|≲λN(k+m)−supz∈Amx+Bk+msupγ∈Zn+,|γ|≤N|∂γ^θ0(z)|≲λN(k+m)−supz∈Amx+Bk+m|z|−β |
with β as in (3.7) and (3.8). From this, (3.15) and [1,(3.2)], it follows that
supy∈Amx+Bk+m|^θ0(y)−P(y)|≲λN(k+m)−supz∈Amx+Bk+mρ(z)−βlnλ−/lnb≲λN(k+m)−b−β(k+m+ν)lnλ−/lnb. |
Therefore,
|σθma(x)|≲‖1Bk‖−1L→p(Rn)bk+mbN(k+m)lnλ−/lnbb−β(k+m+ν)lnλ−/lnb∼‖1Bk‖−1L→p(Rn)b(k+m)[1+(N−β)lnλ−/lnb]b−βνlnλ−/lnb. | (3.17) |
Choosing N larger than β, we have
σθ∗a(x)≲‖1Bk‖−1L→p(Rn)[MHL(1Bk)(x)]βlnλ−/lnb. | (3.18) |
For the case when k+m∈Z∖Z+ and k+m+ν∈Z∖Z+, similarly to (3.17), it is easy to check that, for any x∈Bk+ω+ν+1∖Bk+ω+ν,
|σθma(x)|≲‖1Bk‖−1L→p(Rn)bk+mλN(k+m)−supz∈Amx+Bk+mρ(z)−βlnλ+/lnb≲‖1Bk‖−1L→p(Rn)b(k+m)(1+Nlnλ−/lnb−βlnλ+/lnb)b−βνlnλ+/lnb. |
This further implies that (3.18) holds true for this case by choosing N large enough such that
1+Nlnλ−lnb−βlnλ+lnb>0. |
Finally, for the case when k+m∈Z+, we choose P≡0. Then, by (3.14), (3.15) and the fact that β∈(lnb/lnλ−,∞), we conclude that, for any x∈Bk+ω+ν+1∖Bk+ω+ν,
|σθma(x)|≲‖1Bk‖−1L→p(Rn)bk+msupz∈Amx+Bk+mρ(z)−βlnλ−/lnb≲‖1Bk‖−1L→p(Rn)b(k+m)(1−βlnλ−/lnb)b−βνlnλ−/lnb≲‖1Bk‖−1L→p(Rn)b−βνlnλ−/lnb |
and hence (3.18) also holds true for this case. This finishes the proof of (3.13) and hence of Theorem 3.6.
Remark 2. (i) If A:=dIn×n for some d∈R with |d|∈(1,∞), where In×n denotes the n×n unit matrix, then lnblnλ−=n and the Hardy space H→pA(Rn) goes back to the isotropic mixed-norm Hardy space H→p(Rn). In this case, Theorem 3.6 was obtained by Weisz in [18,Theorem 3]. Moreover, if →p:=(n times⏞p,…,p) with some p∈(0,∞), then Theorem 3.6 implies the well-known result, with β∈(n,∞) and p∈(n/β,∞), for the classical Hardy space Hp(Rn) (see Weisz [16]). This classical result was also proved in a special case, namely, for the Bochner–Riesz means, in Stein et al. [12]. For the same case, a counterexample was also given in [12] to show that the same conclusion is not true for p∈(0,n/β].
(ii) When A is as in (2.5), the space H→pA(Rn) is just the anisotropic mixed-norm Hardy space H→p→a(Rn); see Remark 1(ⅱ). We should point out that Theorem 3.6 is new even for this case.
As applications of Theorem 3.6, we next give some consequences about the convergence of σθmf as follows.
Corollary 3.7. With the same assumptions as in Theorem 3.6, if f∈H→pA(Rn), then {σθmf}m∈N converges almost everywhereas well as in the L→p(Rn)-norm as m→∞.
Proof. Let h be a continuous function with compact support on Rn. Then, by (3.6), we know that, for any m∈N and x∈Rn,
σθmh(x)=∫Rnh(x−A−mt)^θ0(t)dt. |
Note that, for any t∈Rn, limm→∞A−mt=0. By (3.5), the fact that h is bounded and the Lebesgue dominated convergence theorem, we find that, for almost every x∈Rn,
limm→∞σθmh(x)=∫Rnh(x)^θ0(t)dt=h(x)θ0(0)=h(x). |
This convergence also holds true in the L→p(Rn) quasi–norm since h∈H→pA(Rn) which implies σθ∗h∈L→p(Rn).
On the other hand, from [22,Lemma 9(ⅰ)], we deduce that, for any given f∈H→pA(Rn) and ϵ∈(0,∞), there exists a continuous function with compact support h such that
‖f−h‖H→pA(Rn)<ϵ. | (3.19) |
For any K∈N and x∈Rn, let
PK(x):=supm,k∈[K,∞)∩N|σθmf(x)−σθkf(x)|andP(x):=limK→∞PK(x). |
To show Corollary 3.7, we only need to prove that P=0 almost everywhere. To do this, observe that, for any K∈N and x∈Rn,
PK(x)≤supm∈[K,∞)∩N|σθm(f−h)(x)|+supm,k∈[K,∞)∩N|σθmh(x)−σθkh(x)|+supk∈[K,∞)∩N|σθk(h−f)(x)|. |
Then we have
P(x)≤2σθ∗(f−h)(x),∀x∈Rn. |
Combining this, Theorem 3.6 and (3.19), we conclude that
‖P‖L→p(Rn)≤2‖σθ∗(f−h)‖L→p(Rn)≲‖f−h‖H→pA(Rn)≲ϵ. |
Note that ϵ∈(0,∞) is arbitrary. Then we immediately obtain P=0 almost everywhere, which completes the proof of Corollary 3.7.
The following Corollary 3.8 can be easily verified by Theorem 3.6 and an argument same as that used in the proof of [14,Corollary 2.20]; the details are omitted.
Corollary 3.8. With the same assumptions as in Theorem 3.6, if f∈H→pA(Rn)and there exists a subset I⊂Rn such that the restriction f|I∈L→q(I) with q−∈[1,∞), then
limm→∞σθmf(x)=f(x)for almost every x∈I as well as in the L→p(I) quasi-norm. |
Remark 3. Note that, if p−∈(1,∞), then H→pA(Rn)=L→p(Rn) with equivalent quasi-norm (see [3,Proposition 2]). Therefore, Corollary 3.8 further implies the following result: Under the same assumptions as in Theorem 3.6, if f∈L→p(Rn) with p−∈(1,∞), then
limm→∞σθmf(x)=f(x)for almost every x∈Rn as well as in the L→p(Rn) norm. |
As special cases, we consider two specific summability methods.
For any α∈(0,∞) and γ∈N, the Bochner-Riesz summation is defined by setting, for any t∈Rn,
θ0(t):={(1−|t|γ)αwhen|t|∈[0,1),0when|t|∈[1,∞). | (4.1) |
The next lemma comes from [16].
Lemma 4.1. Let θ0 be as in (4.1). If α∈(n−12,∞), then (3.3) and (3.5) hold trueand, for any β∈Zn+, there exists a positive constantC(α,β), depending on α and β, such that, for any x∈Rn∖{0},
|∂β^θ0(x)|≤C(α,β)|x|−n/2−α−1/2. |
By Lemma 4.1 and Theorem 3.6, we have the following conclusion for the Bochner-Riesz summation; the details are omitted.
Theorem 4.2. Let θ0 be as in (4.1) and →p∈(0,∞)n.Assume that
α∈(max{n−12,lnblnλ−−n+12},∞)andp−∈(lnblnλ−(n/2+α+1/2),∞). |
Then, for any f∈H→pA(Rn),
‖σθ∗f‖L→p(Rn)≤C(p−,p+)‖f‖H→pA(Rn), |
where the positive constant C(p−,p+), with p− and p+ as in (2.4), is independent of f.
Remark 4. Let θ0 be as in (4.1). Then, in this special case, the corresponding conclusions in Corollaries 3.7 and 3.8 hold true as well.
The Weierstrass summation is defined by setting, for any t∈Rn,
θ0(t):=e−|t|2/2. | (4.2) |
The succeeding Lemma 4.3 is just [14,Lemma 2.27].
Lemma 4.3. Let θ0 be as in (4.2). Then (3.3) and (3.5) hold true and, for any β∈(1,∞) and α∈Zn+, there exists a positive constant˜C(α,β), depending on α and β, such that, for any x∈Rn∖{0},
|∂α^θ0(x)|≤˜C(α,β)|x|−β. |
By Lemma 4.3 and Theorem 3.6, we obtain the following result for the Weierstrass summation; the details are omitted.
Theorem 4.4. Let θ0 be as in (4.2) and →p∈(0,∞)n.Then, for any f∈H→pA(Rn),
‖σθ∗f‖L→p(Rn)≤˜C(p−,p+)‖f‖H→pA(Rn), |
where the positive constant ˜C(p−,p+), with p− and p+ as in (2.4), is independent of f.Moreover, the corresponding conclusions in Corollaries 3.7 and 3.8 hold true as well.
The author would like to thank the referees for their careful reading and helpful comments which indeed improved the presentation of this article.
The author declares that there is no conflict of interests in this manuscript.
[1] | M. Bownik, Anisotropic Hardy Spaces and Wavelets, American Mathematical Society, Rhode Island, 2003. https://doi.org/http://dx.doi.org/10.1090/memo/0781 |
[2] |
A. P. Calderón, A. Torchinsky, Parabolic maximal functions associated with a distribution, Adv. Math., 16 (1975), 1–64. https://doi.org/10.1016/0001-8708(75)90099-7 doi: 10.1016/0001-8708(75)90099-7
![]() |
[3] |
L. Huang, J. Liu, D. Yang, W. Yuan, Real-variable characterizations of new anisotropic mixed-norm Hardy spaces, Commun. Pure Appl. Anal., 19 (2020), 3033–3082. https://doi.org/10.3934/cpaa.2020132 doi: 10.3934/cpaa.2020132
![]() |
[4] |
T. Chen, W. Sun, Iterated and mixed weak norms with applications to geometric inequalities, J. Geom. Anal., 30 (2020), 4268–4323. https://doi.org/10.1007/s12220-019-00243-x doi: 10.1007/s12220-019-00243-x
![]() |
[5] |
T. Chen, W. Sun, Extension of multilinear fractional integral operators to linear operators on mixed-norm Lebesgue spaces, Math. Ann., 379 (2021), 1089–1172. https://doi.org/10.1007/s00208-020-02105-2 doi: 10.1007/s00208-020-02105-2
![]() |
[6] |
T. Chen, W. Sun, Hardy–Littlewood–Sobolev inequality on mixed-norm Lebesgue spaces, J. Geom. Anal., 32 (2022), 43. https://doi.org/10.1007/s12220-021-00855-2 doi: 10.1007/s12220-021-00855-2
![]() |
[7] |
G. Cleanthous, A. G. Georgiadis, Mixed-norm α-modulation spaces, Trans. Amer. Math. Soc., 373 (2020), 3323–3356. https://doi.org/10.1090/tran/8023 doi: 10.1090/tran/8023
![]() |
[8] |
G. Cleanthous, A. G. Georgiadis, M. Nielsen, Molecular decomposition of anisotropic homogeneous mixed-norm spaces with applications to the boundedness of operators, Appl. Comput. Harmon. Anal., 47 (2019), 447–480. https://doi.org/10.1016/j.acha.2017.10.001 doi: 10.1016/j.acha.2017.10.001
![]() |
[9] |
J. Johnsen, S. Munch Hansen, W. Sickel, Characterisation by local means of anisotropic Lizorkin–Triebel spaces with mixed norms, Z. Anal. Anwend., 32 (2013), 257–277. https://doi.org/10.4171/ZAA/1484 doi: 10.4171/ZAA/1484
![]() |
[10] |
J. Johnsen, S. Munch Hansen, W. Sickel, Anisotropic Lizorkin–Triebel spaces with mixed norms–traces on smooth boundaries, Math. Nachr., 288 (2015), 1327–1359. https://doi.org/10.1002/mana.201300313 doi: 10.1002/mana.201300313
![]() |
[11] |
T. Nogayama, T. Ono, D. Salim, Y. Sawano, Atomic decomposition for mixed Morrey spaces, J. Geom. Anal., 31 (2021), 9338–9365. https://doi.org/10.1007/s12220-020-00513-z doi: 10.1007/s12220-020-00513-z
![]() |
[12] | E. M. Stein, M. H. Taibleson, G. Weiss, Weak type estimates for maximal operators on certain Hp classes, Rend. Circ. Mat. Palermo, 1 (1981), 81–97. |
[13] |
J. Liu, F. Weisz, D. Yang, W. Yuan, Variable anisotropic Hardy spaces and their applications, Taiwanese J. Math., 22 (2018), 1173–1216. https://doi.org/10.11650/tjm/171101 doi: 10.11650/tjm/171101
![]() |
[14] |
J. Liu, F. Weisz, D. Yang, 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. https://doi.org/10.1007/s00041-018-9609-3 doi: 10.1007/s00041-018-9609-3
![]() |
[15] | F. Weisz, Summability of Multi-dimensional Fourier Series and Hardy Spaces, Dordrecht: Kluwer Academic Publishers, 2002. https://doi.org/10.1007/978-94-017-3183-6 |
[16] | F. Weisz, Convergence and Summability of Fourier Transforms and Hardy Spaces, Basel, Birkhäuser, 2017. |
[17] | F. Weisz, Summability of Fourier transforms in variable Hardy and Hardy–Lorentz spaces, Jaen J. Approx., 10 (2018), 101–131. |
[18] | F. Weisz, Summability in mixed-norm Hardy spaces, Ann. Univ. Sci. Budapest. Sect. Comput., 48 (2018), 233–246. |
[19] |
A. Benedek, R. Panzone, The space Lp, with mixed norm, Duke Math. J., 28 (1961), 301–324. https://doi.org/10.1215/S0012-7094-61-02828-9 doi: 10.1215/S0012-7094-61-02828-9
![]() |
[20] |
G. Cleanthous, A. G. Georgiadis, M. Nielsen, Anisotropic mixed-norm Hardy spaces, J. Geom. Anal., 27 (2017), 2758–2787. https://doi.org/10.1007/s12220-017-9781-8 doi: 10.1007/s12220-017-9781-8
![]() |
[21] | F. Weisz, Summability of multi-dimensional trigonometric Fourier series, Surv. Approx. Theory, 7 (2012), 1–179. |
[22] |
J. Liu, L. Huang, C. Yue, Molecular characterizations of anisotropic mixed-norm Hardy spaces and their applications, Mathematics, 9 (2021), 2216. https://doi.org/10.3390/math9182216 doi: 10.3390/math9182216
![]() |