Zum Hauptinhalt springen

Bourgain–Morrey spaces meet structure of Triebel–Lizorkin spaces.

Hu, Pingxu ; Li, Yinqin ; et al.
In: Mathematische Zeitschrift, Jg. 304 (2023-05-01), Heft 1, S. 1-49
Online academicJournal

Bourgain–Morrey spaces meet structure of Triebel–Lizorkin spaces 

Let 0 < q ≤ p ≤ ∞ , r ∈ (0 , ∞ ] , and M q , r p (R n) denote the Bourgain–Morrey space which was introduced by J. Bourgain and has proved important in the study of some linear and nonlinear partial differential equations. In this article, via cleverly combining both the structure of Bourgain–Morrey spaces and the structure of Triebel–Lizorkin spaces and adding an extra exponent τ ∈ (0 , ∞ ] , the authors introduce a new class of function spaces, called Triebel–Lizorkin–Bourgain–Morrey spaces M F ˙ q , r p , τ (R n) . The authors show that M F ˙ q , r p , τ (R n) is just a bridge connecting Bourgain–Morrey spaces and global Morrey spaces. In addition, by fully using exquisite geometrical properties of cubes of Euclidean spaces, the authors also explore various fundamental real-variable properties of M F ˙ q , r p , τ (R n) as well as its relations with other Morrey type spaces, such as Besov–Bourgain–Morrey spaces and local Morrey spaces. Finally, via finding an equivalent quasi-norm of Herz spaces and making full use of both the Calderón product and the sparse family of dyadic grids, the authors obtain the sharp boundedness on M F ˙ q , r p , τ (R n) of classical operators including the Hardy–Littlewood maximal operator, the Calderón–Zygmund operator, and the fractional integral.

This project is supported by the National Natural Science Foundation of China (Grant Nos. 11971058 and 12071197) and the National Key Research and Development Program of China (Grant No. 2020YFA0712900).

Introduction

Let p(0,] and E be a measurable set of Rn . Recall that the Lebesgue space Lp(E) is defined to be the set of all the measurable functions f on E such that, when p(0,) ,

fLp(E):=E|f(x)|pdx1p<

Graph

and

fL(E):=esssupxE|f(x)|:=infFE|F|=0supxE\F|f(x)|<.

Graph

In addition, let xRn and r(0,) . Then a ballB(x, r) in Rn centered at xRn with the radius r is defined by setting

B(x,r):={yRn:|y-x|<r}.

Graph

Throughout this article, we denote by 1E the characteristic function of any set ERn and, for any p(0,] and any open set ΩRn , the symbol Llocp(Ω) is defined to be the set of all the measurable functions f on Ω such that, for any xΩ , there exists a ball B(x,r)Ω such that f1B(x,r)Lp(Ω) . In 1938, to study both the local behavior of solutions of second order elliptic partial differential equations and the calculus of variations, Morrey [[29]] introduced the classical Morrey space. Indeed, let 0<qp and, for any νZ and m:=(m1,...,mn)Zn , let

Qν,m:=j=1nmj2ν,mj+12ν

Graph

be the dyadic cube of Rn ,

Dν:=Qν,m:mZn,andD:=Qν,m:νZ,mZn.

Graph

Then the Morrey space Mqp(Rn) is defined to be the set of all the fLlocq(Rn) such that

fMqp(Rn):=supQD|Q|1p-1qf1QLq(Rn)<.

Graph

After that, function spaces based on Morrey spaces have proved very important in the study of both harmonic analysis and partial differential equations; see, for instance, [[11], [26], [28], [32]–[35], [38]–[40]] for their applications in harmonic analysis and [[6], [9], [27], [44], [49]] for their applications in partial differential equations as well as the monographs Adams [[1]], Lemarié-Rieusset [[19]], Sawano et al. [[41]], and Yuan et al. [[48]].

On the other hand, in order to study both the restriction and the Bochner–Riesz multiplier problems in R3 , Bourgain [[3]] refined the Stein–Tomas (Strichartz) estimate via a special Bourgain–Morrey space. Later on, to explore two minimization problems on non-scattering solutions of nonlinear Schrödinger equations, Masaki [[23]] introduced Bourgain–Morrey spaces for the full range of exponents. Nowadays, Bourgain–Morrey spaces have proved very important in the study related to the Strichartz estimate and the nonlinear Schrödinger equation; see, for instance, [[2], [4], [30]]. In addition, Bourgain–Morrey spaces have also been widely used in other linear and nonlinear partial differential equations, we refer the reader to [[25], [43]] for their applications in Airy equations as well as [[18], [24]] for their applications in KdV equations. We now recall the exact definition of Bourgain–Morrey spaces as follows, which is just [[23], Definition 1.8]; see also [[24], Definition 2.1] or [[16], Definition 1.1].

Definition 1.1

Let 0<qp and r(0,] . The Bourgain–Morrey space Mq,rp(Rn) is defined to be the set of all the fLlocq(Rn) such that

fMq,rp(Rn):=νZ,mZn|Qν,m|1p-1qf1Qν,mLq(Rn)r1r,

Graph

with the usual modification made when r= , is finite.

Observe that, when r= , the space Mq,rp(Rn) is just the classical Morrey space Mqp(Rn) . Recently, Hatano et al. [[16]] gave various important real-variable properties of Bourgain–Morrey spaces, including the approximation property, the interpolation, and the boundedness of several classical operators on them. The article [[16]] of Hatano et al. opens the door to study the real-variable theory of function spaces based on Bourgain–Morrey spaces. Very recently, motivated by the structure of Bourgain–Morrey spaces, amalgam-type spaces, and Besov spaces, Zhao et al. [[50]] introduced the Besov–Bourgain–Morrey space (see also Definition 3.3 below for the definition) which unifies and generalizes both Bourgain–Morrey spaces and amalgam-type spaces. Also, some important basic real-variable properties of Besov–Bourgain–Morrey spaces corresponding to Bourgain–Morrey spaces were investigated in [[50]]. Then it is a natural and meaningful question to ask whether or not it is possible to introduce the Bourgain–Morrey space with the structure of Triebel–Lizorkin spaces and to develop its corresponding real-variable theory.

The main target of this article is to give an affirmative answer to this interesting question. To be precise, via cleverly combining both the structure of Bourgain–Morrey spaces and the structure of Triebel–Lizorkin spaces and adding an extra exponent τ(0,] in the quasi-norm of Bourgain–Morrey spaces Mq,rp(Rn) , we introduce a new class of function spaces, called Triebel–Lizorkin–Bourgain–Morrey spaces MF˙q,rp,τ(Rn) . We show that the space MF˙q,rp,τ(Rn) is just a bridge connecting Bourgain–Morrey spaces and global Morrey spaces. In addition, by fully using exquisite geometrical properties of cubes of Euclidean spaces, we also explore various fundamental real-variable properties of MF˙q,rp,τ(Rn) , including the nontriviality, the approximation property, and the diversity as well as the relations with other Morrey type spaces, such as Besov–Bourgain–Morrey spaces and local Morrey spaces. Finally, via finding an equivalent quasi-norm of Herz spaces and making full use of both the Calderón product and the sparse family of dyadic grids, we obtain the sharp boundedness on MF˙q,rp,τ(Rn) of classical operators including the Hardy–Littlewood maximal operator, the Calderón–Zygmund operator, and the fractional integral. All these have laid a solid foundation for the further development of the real-variable theory of function spaces based on Bourgain–Morrey spaces.

The remainder of this article is organized as follows.

In Sect. 2, we first introduce the concept of the Triebel–Lizorkin–Bourgain–Morrey space MF˙q,rp,τ(Rn) (see Definition 2.1 below). Then we give some fundamental properties of MF˙q,rp,τ(Rn) . To be exact, we find two equivalent quasi-norms of MF˙q,rp,τ(Rn) (see Theorem 2.3 below) and obtain various inclusion relations on different indices and the completeness as well as the dilation invariance, the translation invariance, and the orthogonality invariance of MF˙q,rp,τ(Rn) . Via making full use of both the dilation invariance and the translation invariance of MF˙q,rp,τ(Rn) , we characterize the nontriviality of MF˙q,rp,τ(Rn) (see Theorem 2.9 below). Finally, when all the exponents are finite, we establish both the approximation and the density properties of MF˙q,rp,τ(Rn) using the aforementioned equivalent quasi-norms.

Section 3 is devoted to investigating the relations between Triebel–Lizorkin–Bourgain–Morrey spaces and other various Morrey type spaces and then studying the diversity of MF˙q,rp,τ(Rn) themselves. Indeed, we first prove that MF˙q,rp,τ(Rn) with some special indices goes back to Lebesgue spaces. Next, we explore the embedding between MF˙q,rp,τ(Rn) and (Besov–)Bourgain–Morrey spaces. Via constructing a concrete function which essentially characterizes the exponents r and τ (see Lemmas 3.9 and 3.11 below), we prove that the aforementioned embedding is proper and hence the Triebel–Lizorkin–Bourgain–Morrey space is indeed a new space different from the Besov–Bourgain–Morrey space. In particular, we show that

1.1 MF˙q,rp,r(Rn)=Mq,rp(Rn)

Graph

with equivalent quasi-norms. On the other hand, we give an equivalent quasi-norm of MF˙q,rp,τ(Rn) based on local Morrey spaces and then obtain

1.2 MF˙q,p,τ(Rn)=GMn(1p-1q)q,τ(Rn),

Graph

where the later space is the global Morrey space from [[15]] (see also Definition 3.12 below). Combining (1.1) and (1.2), we conclude that the Triebel–Lizorkin–Bourgain–Morrey space is just a bridge connecting Bourgain–Morrey spaces and global Morrey spaces. Finally, as an application of all the aforementioned relations, we study the diversity of MF˙q,rp,τ(Rn) about the parameters p, q, and r and also obtain the diversity of MF˙q,p,τ(Rn) about the parameter τ . However, when r(0,) , the diversity of MF˙q,rp,τ(Rn) on τ is still unclear.

The main target of Sect. 4 is to study the sharp boundedness on Triebel–Lizorkin–Bourgain–Morrey spaces of several classical operators. First, via proving the equivalence between Herz spaces and local Morrey spaces as well as using the equivalent quasi-norms of MF˙q,rp,τ(Rn) in terms of local Morrey spaces obtained in Sect. 3, we establish a boundedness criterion of operators on MF˙q,rp,τ(Rn) via Herz spaces (see Theorem 4.2 below) and then obtain the sharp boundedness of the Hardy–Littlewood maximal operator, the Fefferman–Stein vector-valued inequality, and the Calderón–Zygmund operator on MF˙q,rp,τ(Rn) . Then we also establish the sharp boundedness of the fractional integral and the fractional maximal operator on MF˙q,rp,τ(Rn) via making full use of both the Calderón product of MF˙q,rp,τ(Rn) and the sparse family of dyadic grids on Euclidean spaces.

Section 5 is devoted to presenting several further remarks. To be precise, we introduce two new distribution spaces associated with Triebel–Lizorkin–Bourgain–Morrey spaces and Besov–Bourgain–Morrey spaces, which are the generalizations of Triebel–Lizorkin spaces and Besov spaces. We will develop a real-variable theory of these new spaces in forthcoming articles.

Finally, we make some conventions on symbols. Let N:={1,2,...} , Z+:=N{0} , Z denote the set of all integers, and Zn:=Z××Z (repeated n times). For any sR , s denotes the largest integer not greater than s. We denote by C a positive constant which is independent of the main parameters, but may vary from line to line. We also use C(α,β,...) to denote a positive constant depending on the indicated parameters α , β, ... . The symbol fg means fCg . If fg and gf , we then write fg . If fCg and g=h or gh , we then write fg=h or fgh . Moreover, we use 0 to denote the origin of Rn . In this article, any cubeQ has finite edge length and all its edges parallel to the coordinate axes. For any cube Q, we denote by l(Q) the edge length of Q and by λQ for any given λ(0,) the cube with edge length λl(Q) and the same center as Q. For any yRn and t(0,) , Q(y, t) denotes a cube of Rn centered at y with l(Q)=t . In particular, for any yRn and νZ , we simply write Qν(y):=Q(y,2-ν). For a given function f on Rn , let supp(f):={xRn:f(x)0} be the support of f. Furthermore, let X and Y be two quasi-normed vector spaces equipped, respectively, with the quasi-norms ·X and ·Y . Then we use XY to denote XY and there exists a positive constant C such that, for any fX , fYCfX. In the end, when we prove a theorem or the like, we always use the same symbols in the wanted proved theorem or the like.

Triebel–Lizorkin–Bourgain–Morrey spaces

In this section, we first introduce the concept of Triebel–Lizorkin–Bourgain–Morrey spaces. Then, in Sect. 2.1, we give two equivalent quasi-norms of MF˙q,rp,τ(Rn) and obtain their inclusion relations on different indices, the dilation invariance, the translation invariance, and the orthogonality invariance of MF˙q,rp,τ(Rn) , and also the completeness of MF˙q,rp,τ(Rn) . Next, in Sect. 2.2, we characterize the nontriviality of Triebel–Lizorkin–Bourgain–Morrey spaces. Finally, in Sect. 2.3, when all the exponents are finite, we establish both the approximation and the density properties of MF˙q,rp,τ(Rn) .

First, we introduce Triebel–Lizorkin–Bourgain–Morrey spaces as follows.

Definition 2.1

Let p,q,r,τ(0,]. The Triebel–Lizorkin–Bourgain–Morrey space (for short, TLBM space) MF˙q,rp,τ(Rn) is defined by setting

MF˙q,rp,τ(Rn):=fLlocq(Rn):fMF˙q,rp,τ(Rn)<,

Graph

where

fMF˙q,rp,τ(Rn):=0|B(·,t)|1p-1q-1rf1B(·,t)Lq(Rn)τdtt1τLr(Rn)

Graph

when τ(0,) and where

fMF˙q,rp,τ(Rn):=supt(0,)|B(·,t)|1p-1q-1rf1B(·,t)Lq(Rn)Lr(Rn)

Graph

when τ= .

Remark 2.2

  • We should point out that, in Definition 2.1, if τ=r , then, in this case, the quasi-norm ·MF˙q,rp,τ(Rn) is just an equivalent quasi-norm of the Bourgain–Morrey space Mq,rp(Rn) , which was established in [[50], Theorem 2.9] (see also Lemma 3.5 below). This further implies that MF˙q,rp,r(Rn)=Mq,rp(Rn) and hence MF˙q,rp,τ(Rn) is just a generalization of Mq,rp(Rn) [see Proposition 3.6(iii) below for the details].
  • The quasi-norm of MF˙q,rp,τ(Rn) essentially contains the structure of Triebel–Lizorkin spaces. Indeed, in the definition of ·MF˙q,rp,τ(Rn) , if we replace f therein by the Littlewood–Paley functions {Δtφ(f)}t(0,) [see (5.3) below for the definition], then we obtain a new function space which is just a generalization of Triebel–Lizorkin spaces (see Definition 5.1 and Remark 5.2 below for more details).
Basic properties

The main target of this subsection is to investigate some fundamental properties of TLBM spaces. All these results play important roles throughout this article.

First, we give the following two equivalent quasi-norms of TLBM spaces.

Theorem 2.3

Let p,q,r,τ(0,].

  • Then fMF˙q,rp,τ(Rn) if and only if fLlocq(Rn) and
  • ‖|f|‖MF˙q,rp,τ(Rn):=∑ν∈Z2νn(1p-1q-1r)f1B(·,2ν)Lq(Rn)τ1τLr(Rn),

Graph

  • with the usual modification made when τ= , is finite. Moreover, there exist two positive constants C1 and C2 such that, for any fMF˙q,rp,τ(Rn) ,
  • C1‖f‖MF˙q,rp,τ(Rn)≤‖|f|‖MF˙q,rp,τ(Rn)≤C2‖f‖MF˙q,rp,τ(Rn).

Graph

  • For any fLlocq(Rn) , let
  • [f]MF˙q,rp,τ(Rn):=∑ν∈Z2-νn(1p-1q-1r)f1Qν(·)Lq(Rn)τ1τLr(Rn)

Graph

  • with the usual modification made when τ= . Then the same conclusions of (i) hold true with |·|MF˙q,rp,τ(Rn) therein replaced by [·]MF˙q,rp,τ(Rn) .
Proof

We only show the case that both r and τ are finite because the proofs of the other cases that r= or τ= are quite similar and hence we omit the details.

First, we prove (i). For the sufficiency, let fLlocq(Rn) be such that |f|MF˙q,rp,τ(Rn)< . Observe that, for any νZ and t[2ν-1,2ν] , we have

2.1 tn(1p-1q-1r)2νn(1p-1q-1r).

Graph

By this, we find that

2.2 fMF˙q,rp,τ(Rn)Rn0tn(1p-1q-1r)f1B(y,t)Lq(Rn)τdttrτdy1r=RnνZ2ν-12νtn(1p-1q-1r)f1B(y,t)Lq(Rn)τdttrτdy1rRnνZ2νn(1p-1q-1r)f1B(y,2ν)Lq(Rn)τrτdy1r=|f|MF˙q,rp,τ(Rn)<.

Graph

This implies that fMF˙q,rp,τ(Rn) and hence the sufficiency of (i) holds true.

Conversely, we next prove the necessity of (i). Indeed, from (2.1) again, it follows that, for any fMF˙q,rp,τ(Rn) ,

2.3 |f|MF˙q,rp,τ(Rn)RnνZ2ν2ν+1dtt2(ν+1)n(1p-1q-1r)f1B(y,2ν)Lq(Rn)τrτdy1rRnνZ2ν2ν+1tn(1p-1q-1r)f1B(y,t)Lq(Rn)τdttrτdy1r=Rn0tn(1p-1q-1r)f1B(y,t)Lq(Rn)τdttrτdy1r=fMF˙q,rp,τ(Rn)<,

Graph

which completes the proof of the necessity of (i). Moreover, combining this estimate and (2.2), we further obtain, for any fMF˙q,rp,τ(Rn) ,

fMF˙q,rp,τ(Rn)|f|MF˙q,rp,τ(Rn).

Graph

This then finishes the proof of (i).

Now, we turn to show (ii). Indeed, from (i), we deduce that, to prove (ii), it suffices to show that, for any fLlocq(Rn) , [f]MF˙q,rp,τ(Rn)|f|MF˙q,rp,τ(Rn) . Notice that, for any yRn and νZ ,

By,2-ν-1Qν(y)=Qy,2-νBy,n2-ν-1By,2-ν+ν0,

Graph

where ν0:=log2n . Applying this and repeating the arguments similar to those used in both (2.2) and (2.3), we conclude that, for any fLlocq(Rn) ,

[f]MF˙q,rp,τ(Rn)|f|MF˙q,rp,τ(Rn).

Graph

This finishes the proof of (ii) and hence Theorem 2.3.

Now, we consider the inclusion properties of TLBM spaces as follows. We point out that Proposition 2.4(i) follows from both Theorem 2.3 and the monotonicity of τ and that Proposition 2.4(ii) is a direct corollary of the Hölder inequality; we omit the details.

Proposition 2.4

Let p,q,r,τ(0,].

  • If 0<τ1τ2 , then MF˙q,rp,τ1(Rn)MF˙q,rp,τ2(Rn) .
  • If 0<q1q2 , then MF˙q2,rp,τ(Rn)MF˙q1,rp,τ(Rn) . Moreover, for any fMF˙q2,rp,τ(Rn) ,
  • ‖f‖MF˙q1,rp,τ(Rn)≤‖f‖MF˙q2,rp,τ(Rn).

Graph

Recall that a real n×n matrix O is said to be orthogonal if OtO=In , where Ot denotes the transpose of O and where In denotes the n×n identity matrix. We next give the dilation invariance, the translation invariance, and the orthogonality invariance of MF˙q,rp,τ(Rn) as follows.

Proposition 2.5

Let p,q,r,τ(0,]. Then

  • for any fMF˙q,rp,τ(Rn) and β(0,) ,
  • ‖f(β·)‖MF˙q,rp,τ(Rn)=β-np‖f‖MF˙q,rp,τ(Rn);

Graph

  • for any fMF˙q,rp,τ(Rn) and ξRn ,
  • ‖f(·-ξ)‖MF˙q,rp,τ(Rn)=‖f‖MF˙q,rp,τ(Rn);

Graph

  • for any fMF˙q,rp,τ(Rn) and any n×n orthogonal matrix O ,
  • fO·MF˙q,rp,τ(Rn)=‖f‖MF˙q,rp,τ(Rn).

Graph

Proof

Let fMF˙q,rp,τ(Rn) . Then, using the definition of ·MF˙q,rp,τ(Rn) and the change of variables, we easily find that, for any β(0,) and ξRn and for any n×n orthogonal matrix O ,

f(β·)MF˙q,rp,τ(Rn)=β-npfMF˙q,rp,τ(Rn),f(·-ξ)MF˙q,rp,τ(Rn)=fMF˙q,rp,τ(Rn),

Graph

and

fO·MF˙q,rp,τ(Rn)=fMF˙q,rp,τ(Rn).

Graph

These then finish the proof of Proposition 2.5.

At the end of this subsection, we consider the completeness of MF˙q,rp,τ(Rn) . In what follows,we use the symbol M(Rn) to denote the set of all measurable functions on Rn . Then we have the following lemma about the completeness of general quasi-normed linear spaces, which is just [[20], Proposition 1.2.36] (see also [[8], Theorem 2]).

Lemma 2.6

Let XM(Rn) be a quasi-normed linear space, equipped with a quasi-norm ·X which makes sense for all functions in M(Rn) . Further assume that X satisfies

  • for any f,gM(Rn) , |g||f| almost everywhere implies that gXfX ;
  • for any {f}mNM(Rn) and fM(Rn) , 0fmf as m almost everywhere implies that fmf as m .

Then X is complete.

The following completeness of TLBM spaces is just a direct corollary of both Lemma 2.6 and the monotone convergence theorem; we omit the details.

Theorem 2.7

If p,q,r,τ(0,] , then MF˙q,rp,τ(Rn) is complete.

Nontriviality

In this subsection, via making use of the characteristic function of B(0,1) and both the dilation invariance and the translation invariance of MF˙q,rp,τ(Rn) , we give a sufficient and necessary condition for the nontriviality of TLBM spaces.

We first consider the sufficient and necessary condition for the characteristic function 1B(0,1) belonging to TLBM spaces as follows.

Proposition 2.8

Let p,q,r,τ(0,] . Then 1B(0,1)MF˙q,rp,τ(Rn) if and only if one of the following three statements holds true:

  • 0 τ(0,) ;
  • 0

  • 0
Proof

For any yRn , let

Ap,q,r,τ(y):=0|B(y,t)|1p-1q-1r1B(0,1)1B(y,t)Lq(Rn)τdtt1τ

Graph

when τ(0,) and let

Ap,q,r,τ(y):=supt(0,)|B(y,t)|1p-1q-1r1B(0,1)1B(y,t)Lq(Rn)

Graph

when τ= . Then 1B(0,1)MF˙q,rp,τ(Rn)=Ap,q,r,τLr(Rn) . Therefore, to prove the present proposition, we only need to estimate Ap,q,r,τLr(Rn) . For this purpose, we consider the following three cases on y.

Case (1) yB(0,1) . In this case, via some simple computations, we have

2.4 |B(y,t)B(0,1)|=|B(y,t)|tnwhent[0,1-|y|),

Graph

2.5 |B(y,t)B(0,1)||B(y,t)|tnwhent[1-|y|,1+|y|],

Graph

and

2.6 |B(y,t)B(0,1)|=|B(0,1)|1whent(1+|y|,).

Graph

By these, we conclude that, if τ(0,) and 1p-1rR\(0,1q) , then

Ap,q,r,τ(y)01-|y|tn(1p-1r)τdtt1τ+1+|y|tn(1p-1q-1r)τdtt1τ=;

Graph

if τ(0,) and 1p-1r(0,1q) , then

Ap,q,r,τ(y)01+|y|tn(1p-1r)τdtt1τ+1+|y|tn(1p-1q-1r)τdtt1τ(1+|y|)n(1p-1r)+(1+|y|)n(1p-1q-1r)1.

Graph

From these, we deduce that, if τ(0,) , then Ap,q,r,τ1B(0,1)Lr(Rn) if and only if

2.7 1p-1r0,1q.

Graph

On the other hand, by (2.4), (2.5), and (2.6) again, we find that, if τ= and 1p-1rR\[0,1q] , then

Ap,q,r,τ(y)supt(0,1-|y|)tn(1p-1r)+supt[1-|y|,)tn(1p-1q-1r)=;

Graph

if τ= and 1p-1r[0,1q] , then

Ap,q,r,τ(y)sup(0,1+|y|]tn(1p-1r)+sup(1+|y|,)tn(1p-1q-1r)(1+|y|)n(1p-1r)+(1+|y|)n(1p-1q-1r)1.

Graph

Therefore, if τ= , then Ap,q,r,τ1B(0,1)Lr(Rn) if and only if

2.8 1p-1r0,1q.

Graph

Case (2) yB(0,2)\B(0,1) . In this case, by some simple computations, we have

|B(y,t)B(0,1)|=0whent[0,|y|-1],

Graph

|B(y,t)B(0,1)||B(y,t)|tnwhent(|y|-1,|y|+1),

Graph

and

|B(y,t)B(0,1)|=|B(0,1)|1whent[|y|+1,).

Graph

Similarly to the proof of Case (1), we easily find that Ap,q,r,τ1B(0,2)\B(0,1)Lr(Rn) if and only if

2.9 1p-1r-,1qandτ(0,)

Graph

or

2.10 1p-1r-,1qandτ=.

Graph

Case (3) yRn\B(0,2) . In this case, observe that

|B(y,t)B(0,1)|=0whent[0,|y|-1],

Graph

|B(y,t)B(0,1)||B(0,1)|1whent(|y|-1,|y|+1),

Graph

and

|B(y,t)B(0,1)|=|B(0,1)|1whent[|y|+1,).

Graph

Similarly to the proof of Case (1) again, we conclude that Ap,q,r,τ1Rn\B(0,2)Lr(Rn) if and only if

2.11 1p-1q(-,0)andmin{r,τ}<

Graph

or

2.12 1p-1q(-,0]andτ=r=.

Graph

Therefore, combining (2.7), (2.8), (2.9), (2.10), (2.11), and (2.12), we conclude that 1B(0,1)MF˙q,rp,τ(Rn) if and only if (i), (ii), or (iii) of the present proposition holds true. This then finishes the proof of Proposition 2.8.

Now, we have the following characterization of the nontriviality of TLBM spaces.

Theorem 2.9

If p,q,r,τ(0,] , then MF˙q,rp,τ(Rn) is nontrivial if and only if (i), (ii), or (iii) of Proposition 2.8 holds true.

To show Theorem 2.9, we need the following Young inequality.

Lemma 2.10

If p,q,r,τ[1,] , then, for any fMF˙q,rp,τ(Rn) and gL1(Rn) , fgMF˙q,rp,τ(Rn) and

fgMF˙q,rp,τ(Rn)fMF˙q,rp,τ(Rn)gL1(Rn).

Graph

Proof

From the Minkowski integral inequality and Proposition 2.5(ii), we infer that, for any fMF˙q,rp,τ(Rn) and gL1(Rn) ,

fgMF˙q,rp,τ(Rn)Rn0{Rn|B(y,t)|1p-1q-1r×{f(·-ξ)1B(y,t)Lq(Rn)|g(ξ)|dξ}τdttrτdy1rRnf(·-ξ)MF˙q,rp,τ(Rn)|g(ξ)|dξ=fMF˙q,rp,τ(Rn)gL1(Rn).

Graph

This finishes the proof of Lemma 2.10.

Based on this Young inequality, the estimate of 1B(0,1) , and both the dilation invariance and the translation invariance of MF˙q,rp,τ(Rn) , we show Theorem 2.9 as follows.

Proof of Theorem 2.9

We first show the sufficiency. To do so, assume that (i), (ii), or (iii) of the present theorem holds true. Then, by Proposition 2.8, we find that 1B(0,1)MF˙q,rp,τ(Rn) . This further implies that the TLBM space under consideration is nontrivial and hence finishes the proof of the sufficiency.

To prove the necessity, let fMF˙q,rp,τ(Rn)\{0} . Then notice that, for any u(0,min{p,τ,q,r}) ,

|f|uMF˙q/u,r/up/u,τ/u(Rn)1u=fMF˙q,rp,τ(Rn)<.

Graph

Thus, we may assume that q,p,r,τ(1,] . Without loss of generality, we may also assume that f0 and fL(Rn) via replacing f by min{1,|f|} . Then, using Lemma 2.10, we conclude that 1B(0,1)fMF˙q,rp,τ(Rn)\{0} . Since 1B(0,1)f is a non-negative and non-zero continuous function, it follows that there exist x0Rn , ϵ(0,) , and s(0,) such that

ϵ1B(x0,s)1B(0,1)f.

Graph

Thus, 1B(x0,s)MF˙q,rp,τ(Rn) . From this and both (i) and (ii) of Proposition 2.5, it follows that

1B(0,1)MF˙q,rp,τ(Rn)=1B(0,s)(s·)MF˙q,rp,τ(Rn)=s-np1B(0,s)MF˙q,rp,τ(Rn)=s-np1B(x0,s)MF˙q,rp,τ(Rn)<,

Graph

which, together with Proposition 2.8, further implies that one of the above four statements in the present theorem holds true. Therefore, the necessity holds true. This finishes the proof of Theorem 2.9.

Remark 2.11

By Theorem 2.9, we find that, only when 0<qpr and τ(0,] , the TLBM space MF˙q,rp,τ(Rn) may be nontrivial. Thus, in what follows, we always assume that 0<qpr and τ(0,] when we consider the TLBM space MF˙q,rp,τ(Rn) .

Approximation and density properties

In this subsection, we investigate both the approximation and the density properties of TLBM spaces when all the exponents are finite. First, combining the equivalent quasi-norms established in Theorem 2.3(ii) and the exquisite geometrical properties of cubes of Euclidean spaces, we obtain the approximation property of MF˙q,rp,τ(Rn) . Then, due to the Lebesgue dominated convergence theorem, we also give the density property of MF˙q,rp,τ(Rn) .

We first give some preliminary concepts. Let Q be a cube of Rn . The average operator mQ is defined by setting, for any fLloc1(Rn) ,

2.13 mQ(f):=1|Q|Qf(x)dx.

Graph

Moreover, for any q(0,) , any cube Q, any fLlocq(Rn) , and any kZ , let

mQ(q)(f):=mQ|f|q1qandEk(q)(f):=QDkmQ(q)(f)1Q.

Graph

Then, via the operators {Ek(q)}k=1 , we have the following approximation property of TLBM spaces.

Theorem 2.12

If 0<q<p<r< and τ(0,) , then, for any non-negative function fMF˙q,rp,τ(Rn) , {Ek(q)(f)}k=1 converges to f in MF˙q,rp,τ(Rn) .

To show this theorem, we need two auxiliary lemmas. First, repeating the proof of [[50], (2.23)] with Qj,m therein replaced by Q, we obtain the following locally approximation property of Lq(Rn) via the operators {Ek(q)}k=1 , which plays a vital role in the proof of Theorem 2.12; we omit the details.

Lemma 2.13

If q(0,) , then, for any cube Q and any fLlocq(Rn)

limkf-Ek(q)(f)1QLq(Rn)=0.

Graph

In addition, we also require the following estimate of operators {Ek(q)}k=1 .

Lemma 2.14

Let 0<q<p<r< and τ(0,) . Then there exists a positive constant C, depending only on p, q, r, τ , and n, such that, for any kN , yRn , and fMF˙q,rp,τ(Rn) ,

2.14 νZ2-νn(1p-1q-1r)τEk(q)(f)1Qν(y)Lq(Rn)τ1τCνZ2-νn(1p-1q-1r)τf1Qν(y)Lq(Rn)τ1τ.

Graph

Proof

Let kN , yRn , and fMF˙q,rp,τ(Rn) . Then we have

2.15 νZ2-νn(1p-1q-1r)τEk(q)(f)1Qν(y)Lq(Rn)τ1τν=-k2-νn(1p-1q-1r)τEk(q)(f)1Qν(y)Lq(Rn)τ1τ+ν=k+11τ=:Gk(1)(y)+Gk(2)(y).

Graph

We first estimate Gk(1)(y) . Indeed, for any νZ , we have

2.16 Ek(q)(f)1Qν(y)Lq(Rn)=Qv(y)QDkmQ(q)(f)q1Q(x)dx1q=QDk,QQν(y)|QQν(y)||Q|Q|f(z)|qdz1q.

Graph

For any νZ(-,k] and QDk such that QQν(y) , we find that QQν-2(y). From this and (2.16), we deduce that, for any νZ(-,k] ,

Ek(q)(f)1Qν(y)Lq(Rn)QDk,QQν(y)Q|f(z)|qdz1qQDk,QQν-2(y)Q|f(z)|qdz1qQν-2(y)|f(z)|qdz1q=f1Qν-2(y)Lq(Rn).

Graph

This, combined with the change of variables, further implies that

2.17 Gk(1)(y)ν=-k2-νn(1p-1q-1r)τf1Qν-2(y)Lq(Rn)τ1τν=-k-22-νn(1p-1q-1r)τf1Qν(y)Lq(Rn)τ1τνZ2-νn(1p-1q-1r)τf1Qν(y)Lq(Rn)τ1τ,

Graph

which is just the desired estimate of Gk(1)(y) .

Then we estimate Gk(2)(y) . Indeed, for any νZ[k+1,) and QDk such that QQν(y) , it holds true that QQk-1(y) , which further implies that

QDk,QQν(y)QQk-1(y).

Graph

From this and (2.16) again, it follows that, for any νZ[k+1,) ,

Ek(q)(f)1Qν(y)Lq(Rn)QDk,QQν(y)2-(ν-k)nQ|f(z)|qdz1q2-(ν-k)nqf1Qk-1(y)Lq(Rn),

Graph

which, together with 0<p<r , further implies that

Gk(2)(y)ν=k+12-νn(1p-1q-1r)τ2-(ν-k)nτqf1Qk-1(y)Lq(Rn)τ1τ2knτqf1Qk-1(y)Lq(Rn)ν=k+12-νn(1p-1r)τ1τ2-(k-1)n(1p-1q-1r)f1Qk-1(y)Lq(Rn)νZ2-νn(1p-1q-1r)τf1Qν(y)Lq(Rn)τ1τ,

Graph

which completes the estimation of Gk(2)(y) . Using this estimate, (2.15), and (2.17), we then conclude that (2.14) holds true. This finishes the proof of Lemma 2.14.

Via both Lemmas 2.13 and 2.14 and the equivalent quasi-norms of MF˙q,rp,τ(Rn) established in Theorem 2.3(ii), we now establish the approximation property of TLBM spaces as follows.

Proof of Theorem 2.12

Let fMF˙q,rp,τ(Rn) be a non-negative function. Then, by Theorem 2.3, we find that

2.18 fMF˙q,rp,τ(Rn)[f]MF˙q,rp,τ(Rn)=RnνZ2-νn(1p-1q-1r)f1Qν(y)Lq(Rn)τrτdy1r=:Rn[F(y)]rdy1r.

Graph

Now, for any kN and yRn , let

Hk(y):=νZ2-νn(1p-1q-1r)τf-Ek(q)(f)1Qν(y)Lq(Rn)τ1τ.

Graph

Then, using Lemma 2.14, we conclude that, for any kN and yRn ,

2.19 Hk(y)νZ2-νn(1p-1q-1r)τf1Qν(y)Lq(Rn)τ1τ+νZ2-νn(1p-1q-1r)τEk(q)(f)1Qν(y)Lq(Rn)τ1τF(y).

Graph

Let D:={yRn:F(y)=} . From (2.18), we deduce that FLr(Rn) and hence |D|=0 . We next show that, for any yRn\D ,

2.20 limkHk(y)=0.

Graph

To this end, fix an ε(0,1) and a yRn\D . Then, by (2.19), we conclude that, for any kN ,

Hk(y)F(y)<,

Graph

which, combined with the assumption τ(0,) , further implies that there exists a JN such that

2.21 Hk(1)(y):=νZ\[-J,J]2-νn(1p-1q-1r)τf-Ek(q)(f)1Qν(y)Lq(Rn)τ1τνZ\[-J,J]2-νn(1p-1q-1r)τf1Qν(y)Lq(Rn)τ1τ<ε.

Graph

On the other hand, for any kN and yRn , let

Hk(2)(y):=ν=-JJ2-νn(1p-1q-1r)τf-Ek(q)(f)1Qν(y)Lq(Rn)τ1τ.

Graph

Then, for any kN , we have

2.22 Hk(y)Hk(1)(y)+Hk(2)(y).

Graph

By Lemma 2.13, we find that, for any νZ[-J,J] ,

limkf-Ek(q)(f)1Qν(y)Lq(Rn)=0.

Graph

This further implies that limkHk(2)(y)=0 , and hence there exists a K1N such that, for any kN(K1,) ,

Hk(2)(y)<ε.

Graph

By this, (2.21), and (2.22), we conclude that (2.20) holds true.

From Theorem 2.3, (2.19), the Lebesgue dominated convergence theorem, and (2.20), it follows that

limkf-Ek(q)(f)MF˙q,rp,τ(Rn)limkf-Ek(q)(f)MF˙q,rp,τ(Rn)=limkRn[Hk(y)]rdy1r=Rnlimk[Hk(y)]rdy1r=0.

Graph

This finishes the proof of Theorem 2.12.

In what follows, for any real function fM(Rn) , let f+:=max{f,0} and f-:=max{-f,0} . For any zC , we denote by z the real part of z and by z the imaginary part of z. As a simple application of Theorem 2.12 and the quasi-triangle inequalities of both lp and Lp(Rn) , we have the following approximation properties; we omit the details.

Corollary 2.15

If 0<q<p<r< and τ(0,) , then, for any fMF˙q,rp,τ(Rn) ,

  • {Ek(q)((ℜf)+)+iEk(q)((ℑf)+)-[Ek(q)((ℜf)-)+iEk(q)((ℑf)-)]}k=1∞ converges to f in MF˙q,rp,τ(Rn) , where i:=-1 ;
  • {Ek(q)(f)}k=1∞ converges to |f| in MF˙q,rp,τ(Rn) .

Let Lc(Rn) denote the set of all the bounded functions on Rn with compact support and Cc(Rn) the set of all the infinitely differentiable functions on Rn with compact support. Now, with the help of the Lebesgue dominated convergence theorem, we show the following density property, that is, Lc(Rn) and Cc(Rn) are both dense in MF˙q,rp,τ(Rn) when all the exponents are finite.

Proposition 2.16

If 0<q<p<r< and τ(0,) , then

  • Lc∞(Rn) is dense in MF˙q,rp,τ(Rn) ;
  • Cc∞(Rn) is dense in MF˙q,rp,τ(Rn) .
Proof

We first show (i). To do this, let fMF˙q,rp,τ(Rn) and, for any mN , let

2.23 fm:=f1[0,m](|f|)1B(0,m).

Graph

Obviously, for any mN , fmLc(Rn) . Moreover, it is easy to show that fm converges to f almost everywhere in Rn as m and that, for any mN , |f-fm|2|f|MF˙q,rp,τ(Rn) . From these and the Lebesgue dominated convergence theorem, we deduce that

2.24 limmf-fmMF˙q,rp,τ(Rn)=0,

Graph

which completes the proof of (i).

Next, we show (ii). For this purpose, let fMF˙q,rp,τ(Rn) and fix any ϵ(0,) . Then, by (i), we find that there exists a gϵLc(Rn) , depending on both f and ϵ , such that

2.25 f-gϵMF˙q,rp,τ(Rn)<ϵ.

Graph

Then there exists an M(0,) , depending on gϵ , such that |gϵ|1B(0,M) almost everywhere in Rn with the implicit positive constant depending on both ϵ and M. Moreover, let ηCc(Rn) be such that 0η1B(0,1) and Rnη(x)dx=1 . For any kN , let ηk(·):=knη(k·) . Then, by this, supp(gϵ)B(0,M) , and [[12], Proposition 8.6 and Theorem 8.15], we find that gϵηkCc(Rn) for any kN and

2.26 gϵηkgϵ

Graph

as k almost everywhere in Rn . In addition, from the definition of ηk , 0η1B(0,1) , and |gϵ|1B(0,M) almost everywhere in Rn with the implicit positive constant depending on both ϵ and M, we deduce that, for any kN and xRn ,

|gϵηk(x)-gϵ(x)|B(x,k-1)ηk(x-z)gϵ(z)dz+|gϵ(x)|knB(x,k-1)η(k(x-z))|gϵ(z)|dz+|gϵ(x)|1|B(x,k-1)|B(x,k-1)|gϵ(z)|dz+|gϵ(x)|1B(0,M+1)(x)

Graph

with the implicit positive constants depending on both ϵ and M, but independent of both k and x. This, combined with the Lebesgue dominated convergence theorem, (2.26), Propositions 2.8 and 2.5(i), and Theorem 2.9, further implies that

gϵηk-gϵMF˙q,rp,τ(Rn)0

Graph

as k . By this and (2.25), we conclude that there exists a kϵN such that

gϵηkϵ-fMF˙q,rp,τ(Rn)gϵηkϵ-gϵMF˙q,rp,τ(Rn)+gϵ-fMF˙q,rp,τ(Rn)ϵ

Graph

with the implicit positive constants independent of ϵ . Therefore, from this and gϵηkϵCc(Rn) , it further follows that Cc(Rn) is dense in MF˙q,rp,τ(Rn) . This then finishes the proof of (ii) and hence Proposition 2.16.

Relations with other Morrey type spaces

The targets of this section are fourfold. The first one is to establish the coincidence between special TLBM spaces and Lebesgue spaces. The second one is to study the relation between TLBM spaces and (Besov–)Bourgain–Morrey spaces. In particular, in Sect. 3.2, we show that the Bourgain–Morrey space is a special case of the TLBM space as well as the TLBM space is indeed a new space different from the (Besov–)Bourgain–Morrey space. The third target is to establish an equivalent quasi-norm of TLBM spaces based on local Morrey spaces and then prove that the global Morrey space is also a special TLBM space, which is presented in Sect. 3.3. Thus, we find that the TLBM space is just a bridge connecting Bourgain–Morrey spaces and global Morrey spaces. Finally, the last target is to study the diversity of TLBM spaces via making full use of the above relations, which is the main content of Sect. 3.4.

Coincidence with Lebesgue spaces

In this subsection, we show that a special case of the TLBM space MF˙q,rp,τ(Rn) is just the Lebesgue space. To do so, recall that the Hardy–Littlewood maximal operator M is defined by setting, for any fLloc1(Rn) and xRn ,

M(f)(x):=supBx1|B|B|f(y)|dy,

Graph

where the supremum is taken over all the balls B containing x. Then, as an application of the boundedness of M on Lebesgue spaces, we establish the following coincidence between special TLBM spaces and Lebesgue spaces.

Proposition 3.1

Let 0<q<p . Then MF˙q,pp,(Rn)=Lp(Rn) with equivalent quasi-norms.

Proof

We first show MF˙q,pp,(Rn)Lp(Rn) . To do this, let fMF˙q,pp,(Rn) . Then, from Definition 2.1, we infer that

3.1 fMF˙q,pp,(Rn)=supt(0,)|B(·,t)|-1qf1B(·,t)Lq(Rn)Lp(Rn)=supt(0,)1|B(·,t)|B(·,t)|f(z)|qdz1qLp(Rn)M(|f|q)1qLp(Rn).

Graph

Notice that |f(x)|qM(|f|q)(x) for almost every xRn . Thus, using (3.1), we have

3.2 fLp(Rn)M(|f|q)1qLp(Rn)fMF˙q,pp,(Rn),

Graph

which implies that fLp(Rn) and hence MF˙q,pp,(Rn)Lp(Rn) .

Next, we prove Lp(Rn)MF˙q,pp,(Rn) . To this end, let fLp(Rn) . Then, by (3.1) again, the assumption q<p , and the boundedness of the Hardy–Littlewood maximal operator M on Lp/q(Rn) , we find that

3.3 fMF˙q,pp,(Rn)M(|f|q)Lp/q(Rn)1p|f|qLp/q(Rn)1p=fLp(Rn)<,

Graph

which further implies that fMF˙q,pp,(Rn) and hence Lp(Rn)MF˙q,pp,(Rn) . Therefore, from this, (3.2), and (3.3), we deduce that Lp(Rn)=MF˙q,pp,(Rn) with equivalent quasi-norms, which completes the proof of Proposition 3.1.

Remark 3.2

Let q(0,) . Then, applying Proposition 3.1 with p= , we obtain

MF˙q,,(Rn)=L(Rn).

Graph

On the other hand, by Remark 2.2(i) [see also Proposition 3.6(iii) below], we find that

MF˙q,,(Rn)=Mq(Rn).

Graph

Therefore, when p= , Proposition 3.1 goes back to the well-known coincidence between the Morrey space Mq(Rn) and the Lebesgue space L(Rn) (see, for instance, [[42], p. 13]).

Relations with (Besov–)Bourgain–Morrey spaces

In this subsection, we explore the (proper) embedding relations between TLBM spaces and Besov–Bourgain–Morrey spaces. To do this, we first recall the following concept of Besov–Bourgain–Morrey spaces introduced in [[50], Definition 1.2].

Definition 3.3

Let 0<qpr and τ(0,] . The Besov–Bourgain–Morrey space MB˙q,rp,τ(Rn) is defined to be the set of all the fLlocq(Rn) such that

fMB˙q,rp,τ(Rn):=νZmZn|Qν,m|1p-1qf1Qν,mLq(Rn)rτr1τ,

Graph

with the usual modifications made when q= , r= , or τ= , is finite.

Remark 3.4

  • As was pointed out in [[50], Remark 1.3], the assumptions that 0<qpr and τ(0,] in Definition 3.3 are reasonable;
  • Observe that, in Definition 3.3, if τ=r , then
  • MB˙q,rp,r(Rn)=Mq,rp(Rn).

Graph

In order to investigate the relation between MF˙q,rp,τ(Rn) and MB˙q,rp,τ(Rn) , we require the following equivalent quasi-norm of Besov–Bourgain–Morrey spaces, which is just [[50], Theorem 2.9].

Lemma 3.5

Let 0<qpr and τ(0,] and, for any hLlocq(Rn) , let

[h]MB˙q,rp,τ(Rn):=0Rn|B(y,t)|1p-1q-1rh1B(y,t)Lq(Rn)rdyτrdtt1τ

Graph

with the usual modifications made when q= , r= , or τ= . Then fMB˙q,rp,τ(Rn) if and only if fLlocq(Rn) and [f]MB˙q,rp,τ(Rn)< . Moreover, there exist two positive constants C1 and C2 such that, for any fMB˙q,rp,τ(Rn) ,

C1fMB˙q,rp,τ(Rn)[f]MB˙q,rp,τ(Rn)C2fMB˙q,rp,τ(Rn).

Graph

Applying this lemma and the Minkowski integral inequality, we immediately conclude the following embedding relations between TLBM spaces and Besov–Bourgain–Morrey spaces; we omit the details.

Proposition 3.6

Let 0<qpr and τ(0,] . Then,

  • when τ(0,r] , MB˙q,rp,τ(Rn)MF˙q,rp,τ(Rn) ;
  • when τ[r,] , MF˙q,rp,τ(Rn)MB˙q,rp,τ(Rn) ;
  • MF˙q,rp,r(Rn)=MB˙q,rp,r(Rn)=Mq,rp(Rn) with equivalent quasi-norms.

When all the exponents are finite, the above inclusions (i) and (ii) are proper, that is, the following conclusions hold true.

Proposition 3.7

Let 0<q<p<r< and τ(0,) .

  • If τ(0,r) , then MB˙q,rp,τ(Rn)MF˙q,rp,τ(Rn) .
  • If τ(r,) , then MF˙q,rp,τ(Rn)MB˙q,rp,τ(Rn) .

To prove this proposition, we require some preliminary lemmas. First, we give both the dilation invariance and the orthogonality invariance of Besov–Bourgain–Morrey spaces as follows.

Lemma 3.8

Let 0<qpr and τ(0,] . Then

  • there exist two positive constants C1 and C2 such that, for any fMB˙q,rp,τ(Rn) and γ(0,) ,
  • C1γ-np‖f‖MB˙q,rp,τ(Rn)≤‖f(γ·)‖MB˙q,rp,τ(Rn)≤C2γ-np‖f‖MB˙q,rp,τ(Rn);

Graph

  • there exist two positive constants C1 and C2 such that, for any fMB˙q,rp,τ(Rn) and any n×n orthogonal matrix O ,
  • C1‖f‖MB˙q,rp,τ(Rn)≤‖f(O·)‖MB˙q,rp,τ(Rn)≤C2‖f‖MB˙q,rp,τ(Rn).

Graph

Proof

(i) is just [[50], Proposition 2.11]. Thus, we only need to show (ii) here. Indeed, by the definition of [·]MB˙q,rp,τ(Rn) and the change of variables, we have, for any fMB˙q,rp,τ(Rn) and any n×n orthogonal matrix O ,

[f(O·)]MB˙q,rp,τ(Rn)=[f]MB˙q,rp,τ(Rn),

Graph

which, together with Lemma 3.5, further implies that

f(O·)MB˙q,rp,τ(Rn)[f(O·)]MB˙q,rp,τ(Rn)=[f]MB˙q,rp,τ(Rn)fMB˙q,rp,τ(Rn).

Graph

This finishes the proof of (ii) and hence Lemma 3.8.

In what follows, for any kN , let Γk:=B(0,2-k+1)\B(0,2-k). Then the following specific function essentially characterizes the exponent τ of MB˙q,rp,τ(Rn) , which plays an important role in the proof of Proposition 3.7.

Lemma 3.9

Let 0<q<p<r< and a,τ(0,) and let

fa(p):=kN2knpk-a1Γk.

Graph

Then fa(p)MB˙q,rp,τ(Rn) if and only if aτ(1,) .

In what follows, let l:={x=(x1,,xn)Rn:x1==xn} and, for any x=(x1,...,xn)Rn , let

3.4 x~:=1ni=1nxi

Graph

and |x|:=maxi{1,,n}|xi| . To prove the last lemma, we need the following geometrical property about cubes.

Lemma 3.10

Let α:=1516n12 , β:=130n-12 , and

E:={x=(x1,...,xn)Rn:x10,|x|<α,and|x-(x~,...,x~)|β|x|},

Graph

where x~ is the same as in (3.4). For any kN , let

3.5 Rk,(1,,1):=Q92k+3,...,92k+3,32k+2.

Graph

Then one has

EkNQk,(1,...,1)kNRk,(1,...,1)

Graph

(see Fig. 1 above for the case that n=2) .

Graph: Fig. 1This figure reflects the result of Lemma 3.10 when n=2 , where A1:=(α,0) , A2:=(α/2,0) , B1:=(0,α) , and B2:=(0,α/2)

Proof

For any kN , let

Ek:={x=(x1,...,xn)Rn:x10,2-kα|x|<2-k+1α,and|x-(x~,...,x~)|β|x|.

Graph

Then E=kNEk . This implies that, to show the present lemma, it suffices to prove that, for any kN ,

3.6 EkQk,(1,,1)Rk,(1,,1).

Graph

Observe that, for any kN , Ek=2-k+1E1:={2-k+1x:xE1} . Combining this and the homogeneity of |·| , we find that, in order to show (3.6), we only need to prove (3.6) with k=1 . To do this, let xE1 . Then, via an obvious geometrical observation, we have x~[716,1516) , where x~ is defined in (3.4). Next, we show (3.6) with k=1 by considering the following two cases on x~ .

Case (i) x~[716,2132) . In this case, from |y||y| for any yRn , |x|<α , and both α=1516n12 and β=130n-12 (hence αβ=132<116 ), we deduce that

3.7 x-916,...,916x-x~,...,x~+x~,...,x~-916,...,916x-x~,...,x~+x~,...,x~-916,...,916<β|x|+18<αβ+18<316,

Graph

which further implies that xR1,(1,,1) and hence completes the proof of (3.6) with k=1 in this case.

Case (ii) x~[2132,1516) . In this case, similarly to (3.7), we have

x-34,...,34x-x~,...,x~+x~,...,x~-34,...,34x-x~,...,x~+x~,...,x~-34,...,34<β|x|+316<αβ+316<14.

Graph

This then implies that xQ1,(1,,1) and hence finishes the proof of (3.6) with k=1 . Thus, (3.6) holds true for any kN , which further implies that

EkNEkkNQk,(1,...,1)kNRk,(1,...,1).

Graph

This finishes the proof of Lemma 3.10.

Applying this lemma and both the dilation invariance and the orthogonality invariance of Besov–Bourgain–Morrey spaces, we now show Lemma 3.9.

Proof of Lemma 3.9

Let

ga(p):=kN2knpk-a1Qk,(1,,1).

Graph

We first claim that

3.8 fa(p)MB˙q,rp,τ(Rn)ga(p)MB˙q,rp,τ(Rn).

Graph

Indeed, observe that, for any xRn ,

ga(p)(x)fa(p)n-12x.

Graph

This, combined with Lemma 3.8(i), further implies that

3.9 ga(p)MB˙q,rp,τ(Rn)fa(p)n-12·MB˙q,rp,τ(Rn)fa(p)MB˙q,rp,τ(Rn).

Graph

Conversely, notice that, for any xRn ,

ga(p)43x=kN2knpk-a1Qk,(1,,1)43x=kN2knpk-a1Rk,(1,,1)x,

Graph

where Rk,(1,...,1) is the same as in (3.5). Therefore, from Lemma 3.10, we deduce that there exist two positive constants α and β and a cone

E:={x=(x1,...,xn)Rn:x10,|x|<α,and|x-(x~,...,x~)|β|x|}

Graph

such that

3.10 EkNQk,(1,...,1)kNRk,(1,...,1)=suppga(p)43·suppga(p).

Graph

On the other hand, via the compactness of the unit sphere of Rn , we find that there exists an NN and a finite sequence {Oj}j=0N of orthogonal matrices such that O0=In and

B(0,α)=j=0NOjE,

Graph

where In denotes the n×n identity matrix. This, together with (3.10), further implies that, for any xRn ,

fa(p)α-1xj=0Nga(p)Oj43x+ga(p)Ojx.

Graph

From this and both (i) and (ii) of Lemma 3.8, it follows that

fa(p)MB˙q,rp,τ(Rn)fa(p)α-1·MB˙q,rp,τ(Rn)j=0Nga(p)Oj43·+ga(p)Oj·MB˙q,rp,τ(Rn)j=0Nga(p)Oj43·MB˙q,rp,τ(Rn)+ga(p)Oj·MB˙q,rp,τ(Rn)(N+1)ga(p)43·MB˙q,rp,τ(Rn)+ga(p)MB˙q,rp,τ(Rn)ga(p)MB˙q,rp,τ(Rn),

Graph

which, combined with (3.9), further implies that fa(p)MB˙q,rp,τ(Rn)ga(p)MB˙q,rp,τ(Rn) and hence the above claim (3.8) holds true.

Next, by [[50], Lemma 2.17(ii)], we find that ga(p)MB˙q,rp,τ(Rn) if and only of aτ(1,) . Combining this and the above claim (3.8), we further conclude that fa(p)MB˙q,rp,τ(Rn) if and only if aτ(1,) , which completes the proof of Lemma 3.9.

In addition, the function fa(p) defined as in Lemma 3.9 characterizes the exponent r of TLBM spaces MF˙q,rp,τ(Rn) as follows, which also plays an essential tool in the proof of Proposition 3.7.

Lemma 3.11

Let 0<q<p<r< and a,τ(0,) . Assume that fa(p) is the same as in Lemma 3.9. Then fa(p)MF˙q,rp,τ(Rn) if and only if ar(1,) .

Proof

For any yRn\{0} , let s:=-log2|y| . Then, from Theorem 2.3(i) and the change of variables, we deduce that

3.11 fa(p)MF˙q,rp,τ(Rn)rfa(p)MF˙q,rp,τ(Rn)r=RnνZ2ν(1p-1q-1r)τkN2knqpk-aqΓkB(y,2ν)τqrτdy=RnνZ2-ν(1p-1q-1r)τ×kN2knqpk-aqΓkB|y|e1,2-ντqrτdyR2-snνZ2-νn(1p-1q-1r)τ×kN2knqpk-aqΓkB2-se1,2-ντqrτds=:R2-snG(s)rτds,

Graph

here and thereafter, e1:=(1,0,...,0) .

Now, we prove the sufficiency. For this purpose, assume ar(1,) . Observe that, for any kN , sR , and νZ ,

3.12 ΓkB(2-se1,2-ν)=0

Graph

if and only if B(2-se1,2-ν)B(0,2-k) or B(2-se1,2-ν)Rn\B(0,2-k+1) , which then further implies that (3.12) holds true if and only if 2-s+2-ν2-k or 2-s-2-ν2-k+1 . Therefore, for any kN , sR , and νZ ,

ΓkB(2-se1,2-ν)0

Graph

if and only if

3.13 2-k<2-s+2-ν

Graph

and

3.14 2-k+1>2-s-2-ν.

Graph

We next estimate (3.11) by considering the following three cases on s.

Case (i) s(-,-1) . In this case, 2-s>2 . Using this and (3.14), we find that

2-ν>2-s-1>1

Graph

and hence ν-1 . If ν>s , then 2-s2-ν+1 . On the other hand, from (3.14) again, we infer that

2-ν>2-s-2-k+12-s-1,

Graph

which contradicts the fact 2-s2-ν+1 . Therefore, νs and hence νs . In this case, applying the assumption 0<q<p<r , we conclude that

3.15 G(s)=ν=-s2-νn(1p-1q-1r)τkN2knqpk-aqΓkB(2-se1,2-ν)τqν=-s2-νn(1p-1q-1r)τkN2knqpk-aq2-knτqν=-s2-νn(1p-1q-1r)τ2-sn(1p-1q-1r)τ.

Graph

This, together with ss and 0<q<p , further implies that

3.16 --12-snG(s)rτds--12-sn(1p-1q-1r)τrτ2-snds--12-sn(1p-1q)rds<,

Graph

which is the desired estimate in this case.

Case (ii) s[-1,1] . In this case, 2-s[2-1,2] . If ν2 , then 2-s-2-ν2-2 . From this and (3.14), we deduce that 2-k+1>2-2 and hence k2 . In this case, by this and the assumption 0<q<p<r< , we have

3.17 G(s)=ν=-12-νn(1p-1q-1r)τkN2knqpk-aqΓkB(2-se1,2-ν)τq+ν=2ν=-12-νn(1p-1q-1r)τkN2knqpk-aq2-nkτq+ν=22-νn(1p-1q-1r)τ2-νnτqk=122knqpk-aqτqν=-12-νn(1p-1q-1r)τ+ν=22-νn(1p-1r)τ1,

Graph

which further implies that

3.18 -112-snG(s)rτds-112-snds<.

Graph

This is the desired estimate in this case.

Case (iii) s(1,) . In this case, 2-s(0,2-1) . Notice that

3.19 G(s)=ν=-02-νn(1p-1q-1r)τkN2knqpk-aqΓkB(2-se1,2-ν)τq+ν=1s+1+ν=s+2=:I1+I2+I3.

Graph

For I1 , applying the assumption 0<q<p<r< , we find that

3.20 I1ν=-02-νn(1p-1q-1r)τkN2knqpk-aq2-knτqν=-02-νn(1p-1q-1r)τ1.

Graph

Now, we deal with I2 . Indeed, for any νN[1,s+1] , 2-s+2-ν3·2-ν<2-ν+2 . From this and (3.13), we deduce that kν-1 . Thus, by the assumption q<p , we obtain

3.21 I2=ν=1s+12-νn(1p-1q-1r)τk=max{1,ν-1}2knqpk-aqΓkB(2-se1,2-ν)τqν=1s+12-νn(1p-1q-1r)τk=max{1,ν-1}2knqpk-aq2-knτqν=1s+12-νn(1p-1q-1r)τ2νn(1p-1q)τν-aτ2nsτr(s)-aτ,

Graph

which completes the estimation of I2 . Finally, for I3 , observe that, for any νZ[s+2,) , νs+2>s+1 and hence 2-s-2-ν>2-s-1 . This, combined with (3.14), further implies that 2-k+1>2-s-1 and hence ks+2 . In addition, by ν>s+1 , we have 2-s+2-ν<2-s+1 . Using this and (3.13), we obtain ks . From this, ks+2 , and the assumption 0<q<p<r again, it follows that

I3ν=s+22-νn(1p-1q-1r)τk=ss+22knqpk-aq2-nντq2snqp(s)-aqτqν=s+22-νn(1p-1r)τ2nsτr(s)-aτ,

Graph

which is the desired estimate of I3 . Combining this estimate, (3.19), (3.20), and (3.21), we further obtain

3.22 G(s)1+2nsτr(s)-aτ.

Graph

Using this and the assumption ar(1,) , we find that

3.23 12-snG(s)rτds12-sn1+2nsτr(s)-aτrτds12-sn+2(s-s)n(s)-ards12-sn+s-ards<.

Graph

This is the desired estimate in this case. Therefore, combining (3.11), (3.16), (3.18), and (3.23), we have fa(p)MF˙q,rp,τ(Rn)r< and hence fa(p)MF˙q,rp,τ(Rn) . This then finishes the proof of the sufficiency.

Next, we show the necessity. To do this, let ar(-,1] . It suffices to prove fa(p)MF˙q,rp,τ(Rn) . Indeed, for any sj=1(j+1/3,j+2/3) , we have 2-se1Γs+1 . In addition, for any νZ[s+3,) , it holds true that

2-s-2-ν2-s-1and2-s+2-ν2-s,

Graph

which imply that B(2-se1,2-ν)Γs+1 . From this and the assumption 0<q<p<r< , we infer that

G(s)ν=s+32-νn(1p-1q-1r)τkN2knqpk-aqΓkB(2-se1,2-ν)τqν=s+32-νn(1p-1q-1r)τ2snqp(s)-aq2-νnτq2snτr(s)-aτ.

Graph

By this, (3.11), and the assumption ar(-,1] , we conclude that

fa(p)MF˙q,rp,τ(Rn)rR2-snG(s)rτds12-sn[G(s)]rτdsjNj+13j+232-sn[G(s)]rτdsjNj+13j+232-sn2snτr(s)-aτrτds=jNj+13j+23(s)-ar2(s-s)ndsjNj-ar=,

Graph

which further implies that fa(p)MF˙q,rp,τ(Rn) . This finishes the proof of the necessity and hence Lemma 3.11.

Based on above preparations, we now show Proposition 3.7.

Proof of Proposition 3.7

We first show (i). To this end, assume τ(0,r) and let a(1r,1τ) . Then aτ<1<ar, which, together with both Lemmas 3.9 and 3.11, further implies that

fa(p)MF˙q,rp,τ(Rn)\MB˙q,rp,τ(Rn).

Graph

This then finishes the proof of (i).

For (ii), assume τ(r,) and let a(1τ,1r) . Then, from Lemmas 3.9 and 3.11 again, it follows that ar<1<aτ and hence

fa(p)MB˙q,rp,τ(Rn)\MF˙q,rp,τ(Rn).

Graph

This implies that (ii) holds true and then finishes the proof of Proposition 3.7.

Relations with local and global Morrey spaces

This subsection is devoted to investigating the relation between TLBM spaces and local or global Morrey spaces. First, we give an equivalent quasi-norm of TLBM spaces via local Morrey spaces. Then, as an application, we show that the global Morrey space is just a special case of TLBM spaces. We now recall the following concepts of local and global Morrey spaces which are originally introduced in [[15]] (see also [[41], p. 214, Definition 36]).

Definition 3.12

Let q,τ(0,] , αR , and ξRn .

  • The local Morrey space LMαq,τ,ξ(Rn) is defined to be the set of all the fLlocq(Rn) such that
  • ‖f‖LMαq,τ,ξ(Rn):=∫0∞tατf1B(ξ,t)Lq(Rn)τdtt1τ<∞.

Graph

  • The global Morrey space GMαq,τ(Rn) is defined to be the set of all the fLlocq(Rn) such that
  • ‖f‖GMαq,τ(Rn):=supξ∈Rn∫0∞tατf1B(ξ,t)Lq(Rn)τdtt1τ<∞.

Graph

Remark 3.13

Let q, τ , α , and ξ be the same as in Definition 3.12. We point out that, in some sense, ξ is the center of the local Morrey space LMαq,τ,ξ(Rn) . Indeed, for any fLlocq(Rn) , by the change of variables and the fact that, for any xRn , |(2ξ-x)-ξ|=|x-ξ| , we obtain

f(2ξ-·)LMαq,τ,ξ(Rn)=fLMαq,τ,ξ(Rn).

Graph

Observe that x and 2ξ-x are symmetrical about ξ . Therefore, in this sense, ξ can be regarded as the center of LMαq,τ,ξ(Rn) .

Via local Morrey spaces, we have an equivalent quasi-norm of MF˙q,rp,τ(Rn) as follows.

Proposition 3.14

Let 0<qpr and τ(0,] . Then there exist two positive constants C1 and C2 such that, for any fLlocq(Rn) ,

C1fMF˙q,rp,τ(Rn)RnfLMα(p,q,r)q,τ,y(Rn)rdy1rC2fMF˙q,rp,τ(Rn),

Graph

here and thereafter,

3.24 α(p,q,r):=n1p-1q-1r.

Graph

Moreover, if r= , then

3.25 GMα(p,q)q,τ(Rn)MF˙q,p,τ(Rn),

Graph

here and thereafter,

3.26 α(p,q):=n1p-1q.

Graph

Proof

Notice that, for any fLlocq(Rn) ,

fMF˙q,rp,τ(Rn)Rn0tn(1p-1q-1r)f1B(y,t)Lq(Rn)τdttrτdy1r=RnfLMα(p,q,r)q,τ,y(Rn)rdy1r.

Graph

Using this and letting r= , we further conclude that, for any fLlocq(Rn) ,

3.27 fMF˙q,p,τ(Rn)esssupyRnfLMα(p,q)q,τ,y(Rn)supyRnfLMα(p,q)q,τ,y(Rn)=fGMα(p,q)p,q(Rn),

Graph

which implies that (3.25) holds true and hence completes the proof of Proposition 3.14.

In particular, the following conclusion shows that the TLBM space MF˙q,p,τ(Rn) is just the global Morrey space.

Theorem 3.15

Let 0<q<p< and τ(0,) , or let 0<qp and τ= . Then

MF˙q,p,τ(Rn)=GMα(p,q)q,τ(Rn)

Graph

with equivalent quasi-norms, where α(p,q) is the same as in (3.26).

Proof

We first consider the case τ= . Indeed, in this case, by Proposition 3.6(iii), we have

MF˙q,p,(Rn)=Mq,p(Rn)=Mqp(Rn).

Graph

On the other hand, from [[41], p. 214, Example 68], we infer that

GMα(p,q)q,(Rn)=Mqp(Rn).

Graph

Thus, we have

MF˙q,p,(Rn)=Mqp(Rn)=GMα(p,q)q,(Rn).

Graph

This finishes the proof of the present theorem in this case.

In what follows, we consider the case τ(0,) . In this case, due to (3.27), it suffices to show that, for any fLlocq(Rn) ,

fGMα(p,q)q,τ(Rn)esssupyRnfLMα(p,q)q,τ,y(Rn).

Graph

We first claim that, if |f|1B for some ball B, then, for any ξRn ,

3.28 fLMα(p,q)q,τ,η(Rn)fLMα(p,q)q,τ,ξ(Rn)

Graph

as ηξ . To do this, fix an f and a ball B such that |f|1B and let ξRn . Then, for any ηRn ,

3.29 fLMα(p,q)q,τ,ξ(Rn)τ-fLMα(p,q)q,τ,η(Rn)τ0tα(p,q)f1B(ξ,t)Lq(Rn)-f1B(η,t)Lq(Rn)τdtt=01tα(p,q)f1B(ξ,t)Lq(Rn)-f1B(η,t)Lq(Rn)τdtt+1=:I(ξ,η)+II(ξ,η).

Graph

By |f|1B , we find that, for any ηRn and t(0,) ,

f1B(η,t)q-f1B(ξ,t)q1BL1(Rn)

Graph

with the implicit positive constant independent of both ξ and η . This, combined with the Lebesgue dominated convergence theorem and the fact that |f1B(η,t)| converges to |f1B(ξ,t)| almost everywhere in Rn as ηξ , further implies that

3.30 limηξf1B(η,t)Lq(Rn)=f1B(ξ,t)Lq(Rn).

Graph

Moreover, using the Hölder inequality and |f|1B , we have, for any t(0,) and ηRn ,

f1B(ξ,t)Lq(Rn)-f1B(η,t)Lq(Rn)fL(Rn)|B(ξ,t)|1q+|B(η,t)|1qtnq,

Graph

which further implies that

3.31 I(ξ,η)01tα(p,q)tnqτdtt1τ=01tnτpdtt1τ<.

Graph

On the other hand, from |f|1B again, we deduce that, for any t(0,) and ηRn ,

f1B(ξ,t)Lq(Rn)-f1B(η,t)Lq(Rn)fLq(Rn)1,

Graph

which further implies that

II(ξ,η)1tα(p,q)τdtt1τ<.

Graph

Using this, (3.31), (3.30), and the Lebesgue dominated convergence theorem, we conclude that I(ξ,η)0 and II(ξ,η)0 as ηξ . This, together with (3.29), further implies that the above claim (3.28) holds true.

Next, for any kN and xRn , let

fk(x):=f(x)1B(0,k)(x)when|f(x)|k;0otherwise.

Graph

Then, obviously, for any kN , |fk|k1B(0,k) and |fk||f| as k . Thus, from the monotone convergence theorem, we deduce that, for any ξRn ,

3.32 fkLMα(p,q)q,τ,ξ(Rn)fLMα(p,q)q,τ,ξ(Rn)

Graph

as k . In addition, for any ϵ(0,fGMα(p,q)p,τ(Rn)) , by the definition of ·GMα(p,q)p,τ(Rn) , we conclude that there exists a ξ0 such that ϵ(0,fLMα(p,q)p,τ,ξ0(Rn)) . This, combined with (3.32), further implies that there exists a k0N such that

3.33 ϵ0,fk0LMα(p,q)p,τ,ξ0(Rn).

Graph

From the definition of the essential supremum, we deduce that there exists a set E with |E|=0 such that

3.34 esssupyRnfLMα(p,q)q,τ,y(Rn)=supRn\EfLMα(p,q)q,τ,ξ(Rn).

Graph

Moreover, notice that |fk0(·+ξ0)|k01B(-ξ0,k0) . Therefore, applying the above claim (3.28) with f:=fk0(·+ξ0) , (3.33), and the density of Rn\E in Rn , we find that there exists an η0Rn\E such that ϵ(0,fk0LMα(p,q)q,τ,η0(Rn)) . This, together with |fk0||f| and (3.34), further implies that

ϵ<fk0LMα(p,q)q,τ,η0(Rn)fLMα(p,q)q,τ,η0(Rn)esssupyRnfLMα(p,q)q,τ,y(Rn).

Graph

Letting ϵfGMα(p,q)p,τ(Rn) , we obtain

fGMα(p,q)p,τ(Rn)esssupyRnfLMα(p,q)q,τ,y(Rn),

Graph

which completes the proof of Theorem 3.15.

Remark 3.16

  • Combining Theorem 3.15 and Propositions 3.1 and 3.6(iii), we conclude that MF˙q,rp,τ(Rn) is a new space different from MB˙q,rp,τ(Rn) , when 0<q<p<r< and τ(0,] , and
  • MF˙q,rp,τ(Rn)=GMn(1p-1q)q,τ(Rn)if0

Graph

  • Applying these and Proposition 3.6(iii) again, we conclude that the TLBM space MF˙q,rp,τ(Rn) is indeed a new space and a bridge connecting Bourgain–Morrey spaces and global Morrey spaces.
  • Let 0<q<p=r< and τ(0,) . Then, by Theorem 2.9 and Proposition 3.1, we find that
  • MF˙q,pp,τ(Rn)={0}⫋Lp(Rn)=MF˙q,pp,∞(Rn).

Graph

  • Thus, the index τ indeed plays a role in MF˙q,rp,τ(Rn) .
Diversity of MF˙q,rp,τ(Rn)

The main target of this subsection is to show the following diversity of MF˙q,rp,τ(Rn) based on its relations with both (Besov–)Bourgain–Morrey spaces and global Morrey spaces.

Theorem 3.17

Let 0<qi<pi<ri for any i{1,2} and let τ1,τ2(0,) .

  • If MF˙q1,r1p1,τ1(Rn)=MF˙q2,r2p2,τ2(Rn) , then p1=p2 , q1=q2 , and r1=r2 .
  • MF˙q1,∞p1,τ1(Rn)=MF˙q2,∞p2,τ2(Rn) if and only if p1=p2 , q1=q2 , and τ1=τ2 .
Remark 3.18

Theorem 3.17 gives the diversity of MF˙q,rp,τ(Rn) on the exponents p, q, and r and, when r= , also on τ . However, we should point out that, when r(0,) , the diversity of MF˙q,rp,τ(Rn) on τ is still unclear.

To prove Theorem 3.17, we need the following auxiliary lemma of Besov–Bourgain–Morrey spaces, which is just [[50], Theorem 2.22(i)].

Lemma 3.19

Let 0<q1<p1<r1 , 0<q2<p2<r2 , and τ1,τ2(0,) . Then MB˙q1,r1p1,τ1(Rn)MB˙q2,r2p2,τ2(Rn) if and only if q1q2 , p1=p2 , r1r2 , and τ1τ2 .

We also require the following technical conclusion.

Proposition 3.20

Let 0<q<p< and a,τ(0,) . Assume that fa(p) is the same as in Lemma 3.9. Then fa(p)MF˙q,p,τ(Rn) if and only if aτ(1,) .

Proof

Repeating the arguments similar to those used in the estimations of both (2.2) and (2.3) with the norm ·Lr(Rn) therein replaced by supξRn , we have, for any ξRn and fLlocq(Rn) ,

3.35 0tα(p,q)τf1B(ξ,t)Lq(Rn)τdtt1τνZ2-να(p,q)τf1B(ξ,2-ν)Lq(Rn)τ1τ.

Graph

From this, it follows that

3.36 fa(p)GMα(p,q)q,τ(Rn)τ=supξRn0tα(p,q)τfa(p)1B(ξ,t)Lq(Rn)τdttsupξRnνZ2-να(p,q)τfa(p)1B(ξ,2-ν)Lq(Rn)τ=supξRnνZ2-να(p,q)τ×kN2knqpk-aqΓkB(ξ,2-ν)τq=supξRnνZ2-να(p,q)τ×kN2knqpk-aqΓkB(|ξ|e1,2-ν)τq=:supξRnF(ξ).

Graph

Next, we claim that

3.37 F(0)<if and only ifaτ(1,).

Graph

Indeed, it is easy to show that

ΓkB(0,2-ν)0if and only ifkν+1.

Graph

From this, it follows that

νZ2-να(p,q)τkN2knqpk-aqΓkB(0,2-ν)τqνZ2-να(p,q)τk=ν+12kn(qp-1)k-aqτq=ν=02-να(p,q)τk=ν+12kn(qp-1)k-aqτq+ν=--12-να(p,q)τk=12kn(qp-1)k-aqτqνZ+ν-aτ+1,

Graph

which implies that the above claim (3.37) holds true.

Now, we show the necessity of the present proposition. For this purpose, assume that fa(p)MF˙q,p,τ(Rn) . Then, from Theorem 3.15, (3.36), and (3.37), we deduce that

F(0)fGMα(p,q)q,τ(Rn)fMF˙q,p,τ(Rn)<

Graph

and hence aτ(1,) . This finishes the proof of the necessity of the present proposition.

Next, we prove the sufficiency of the present proposition. To do this, assume aτ(1,) . For any ξRn\{0} , let s:=-log2|ξ| and

G(s):=νZ2-να(p,q)τk=12knqpk-aqΓkB(2-se1,2-ν)τq.

Graph

Then, repeating the arguments similar to those used in the estimations of (3.15), (3.17), and (3.22) with r replaced by , we have

G(s)2-sα(p,q)τwhens(-,-1),1whens[-1,1],1+(s)-aτwhens(1,).

Graph

By these and the assumption aτ(1,) , we find that supsRG(s)< . This, combined with Theorem 3.15, (3.36), (3.37), and the assumption aτ(1,) again, further implies that

fMF˙q,p,τ(Rn)fGMα(p,q)q,τ(Rn)=maxF(0),supξRn\{0}F(ξ)=maxF(0),supsRG(s)<

Graph

and hence fMF˙q,p,τ(Rn) . This finishes the proof of the sufficiency and hence Proposition 3.20.

Now, we show Theorem 3.17 as follows.

Proof of Theorem 3.17

We first show (i). Indeed, from Proposition 3.6 and Proposition 2.4(i), it follows that

MB˙q1,r1p1,min{τ1,r1}(Rn)MF˙q1,r1p1,τ1(Rn)=MF˙q2,r2p2,τ2(Rn)MB˙q2,r2p2,max{τ2,r2}(Rn)

Graph

and

MB˙q2,r2p2,min{τ2,r2}(Rn)MF˙q2,r2p2,τ2(Rn)=MF˙q1,r1p1,τ1(Rn)MB˙q1,r1p1,max{τ1,r1}(Rn).

Graph

Using these and Lemma 3.19, we conclude that q1=q2 , p1=p2 , and r1=r2 . This finishes the proof of (i).

Next, we prove (ii). Indeed, by Proposition 2.4(i), we find that, it is enough to show that, if τ1<τ2 , then

MF˙q,p,τ1(Rn)MF˙q,p,τ2(Rn).

Graph

To this end, let fa(p) be the same as in Proposition 3.20, τ1<τ2 , and a(1τ2,1τ1) . Then, from Proposition 3.20, it follows that

fa(p)MF˙q,p,τ2(Rn)\MF˙q,p,τ1(Rn),

Graph

which, together with Proposition 2.4(i) again, further implies that MF˙q,p,τ1(Rn)MF˙q,p,τ2(Rn) . This then finishes the proof of (ii) and hence Theorem 3.17.

Boundedness of operators

In this section, we consider the sharp boundedness of several classical operators on TLBM spaces. To be precise, in Sect. 4.1, we first establish a boundedness criterion of operators on TLBM spaces via Herz spaces (see Theorem 4.2 below) and then give the sharp boundedness of the Hardy–Littlewood maximal operator, the Fefferman–Stein vector-valued inequality, and the Calderón–Zygmund operator. Next, in Sect. 4.2, we obtain the sharp boundedness of both the fractional integral and the fractional maximal operator on TLBM spaces.

Boundedness of operators via Herz spaces

In this subsection, via establishing the equivalence between Herz spaces and local Morrey spaces, we obtain a boundedness criterion of operators on TLBM spaces. As applications, we establish the boundedness of several classical operators on MF˙q,rp,τ(Rn) , including the Hardy–Littlewood maximal operator and the Calderón–Zygmund operator as well as giving the Fefferman–Stein vector-valued inequality on MF˙q,rp,τ(Rn) .

First, we recall the following definition of Herz spaces which were originally introduced in [[21], Definition 1.1(a)] (see also [[22], Chapter 1]).

Definition 4.1

Let q,τ(0,] , αR , and ξRn . The homogeneous Herz space K˙αq,τ,ξ(Rn) is defined to be the set of all the fLlocq(Rn\{ξ}) such that

fK˙αq,τ,ξ(Rn):=νZ2νατf1B(ξ,2ν)\B(ξ,2ν-1)Lq(Rn)τ1τ<.

Graph

The main result of this subsection is the following boundedness criterion of operators on TLBM spaces via Herz spaces.

Theorem 4.2

Let 0<qpr and τ(0,] . Let T be an operator mapping Llocq(Rn) into M(Rn) . Assume that there exists a positive constant C such that, for almost every ξRn and for any fLlocq(Rn) ,

4.1 T(f)K˙α(p,q,r)q,τ,ξ(Rn)CfK˙α(p,q,r)q,τ,ξ(Rn),

Graph

where α(p,q,r) is the same as in (3.24). Then T is bounded on MF˙q,rp,τ(Rn) .

To prove this theorem, we first establish the following technique lemma about the equivalence between Herz spaces and local Morrey spaces.

Lemma 4.3

Let q,τ(0,] , α(-,0) , and ξRn . Then

LMαq,τ,ξ(Rn)=K˙αq,τ,ξ(Rn)

Graph

with equivalent quasi-norms and the positive equivalence constants independent of ξ .

Proof

We first show LMαq,τ,ξ(Rn)K˙αq,τ,ξ(Rn) . Indeed, let fLMαq,τ,ξ(Rn) . Then, from (3.35), we deduce that

4.2 fK˙αq,τ,ξ(Rn)νZ2νατf1B(ξ,2ν)Lq(Rn)τ1τfLMαq,τ,ξ(Rn)<.

Graph

This implies that fK˙αq,τ,ξ(Rn) and hence LMαq,τ,ξ(Rn)K˙αq,τ,ξ(Rn) .

Conversely, we now prove that K˙αq,τ,ξ(Rn)LMαq,τ,ξ(Rn) . To do this, assume that fK˙αq,τ,ξ(Rn) and consider the following four cases on both q and τ .

Case (i) q[1,] and τ(0,1) . In this case, by q[1,] and the Minkowski inequality of Lq(Rn) , we conclude that, for any νZ ,

f1B(ξ,2ν)Lq(Rn)i=-νf1B(ξ,2i)\B(ξ,2i-1)Lq(Rn),

Graph

which, combined with (3.35), the well-known inequality that, for any p(0,1] and {aj}jNC ,

4.3 jN|aj|pjN|aj|p,

Graph

the Tonelli theorem, and α(-,0) , further implies that

4.4 fLMαq,τ,ξ(Rn)νZ2νατf1B(ξ,2ν)Lq(Rn)τ1τνZ2νατi=-νf1B(ξ,2i)\B(ξ,2i-1)Lq(Rn)τ1τνZi=-ν2νατf1B(ξ,2i)\B(ξ,2i-1)Lq(Rn)τ1τ=iZ2iατf1B(ξ,2i)\B(ξ,2i-1)Lq(Rn)τν=i2(ν-i)ατ1τiZ2iατf1B(ξ,2i)\B(ξ,2i-1)Lq(Rn)τ1τ=fK˙αq,τ,ξ(Rn)<.

Graph

Thus, fLMαq,τ,ξ(Rn) . This then finishes the proof of K˙αq,τ,ξ(Rn)LMαq,τ,ξ(Rn) in this case.

Case (ii) q[1,] and τ[1,] . In this case, observe that, for any νZ ,

4.5 f1B(ξ,2ν)=k=0f1B(ξ,2ν-k)\B(ξ,2ν-k-1)

Graph

almost everywhere in Rn . Using this, (3.35), the Minkowski inequality, the change of variables, and α(-,0) , we have

4.6 fLMαq,τ,ξ(Rn)νZ2νατk=0f1B(ξ,2ν-k)\B(ξ,2ν-k-1)Lq(Rn)τ1τk=0νZ2νατf1B(ξ,2ν-k)\B(ξ,2ν-k-1)Lq(Rn)τ1τ=k=02kανZ2(ν-k)ατf1B(ξ,2ν-k)\B(ξ,2ν-k-1)Lq(Rn)τ1τfK˙αq,τ,ξ(Rn)<,

Graph

which further implies that fK˙αq,τ,ξ(Rn) . Therefore, K˙αq,τ,ξ(Rn)LMαq,τ,ξ(Rn) in this case.

Case (iii) q(0,1) and τ(0,q) . In this case, by the assumption q(0,1) and (4.3), we find that, for any νZ ,

4.7 f1B(ξ,2ν)Lq(Rn)τRni=-νf(x)1B(ξ,2i)\B(ξ,2i-1)(x)qdxτq=i=-νf1B(ξ,2i)\B(ξ,2i-1)Lq(Rn)qτq,

Graph

which, together with the assumption τq(0,1) and (4.3) again, further implies that

f1B(ξ,2ν)Lq(Rn)τi=-νf1B(ξ,2i)\B(ξ,2i-1)Lq(Rn)τ.

Graph

From this, (3.35), the Tonelli theorem, and α(-,0) , it follows that

4.8 fLMαq,τ,ξ(Rn)νZ2νατf1B(ξ,2ν)Lq(Rn)τ1τνZ2νατi=-νf1B(ξ,2i)\B(ξ,2i-1)Lq(Rn)τ1τ=iZ2iατf1B(ξ,2i)\B(ξ,2i-1)Lq(Rn)τν=i2(ν-i)ατ1τiZ2iατf1B(ξ,2i)\B(ξ,2i-1)Lq(Rn)τ1τ=fK˙αq,τ,ξ(Rn)<.

Graph

Thus, fLMαq,τ,ξ(Rn) and hence K˙αq,τ,ξ(Rn)LMαq,τ,ξ(Rn) in this case.

Case (iv) q(0,1) and τ[q,) . In this case, applying (4.5) and repeating an argument similar to that used in the estimation of (4.7), we conclude that

f1B(ξ,2ν)Lq(Rn)τk=0f1B(ξ,2ν-k)\B(ξ,2ν-k-1)Lq(Rn)qτq.

Graph

This, combined with (3.35), τq[1,) , the Minkowski inequality, the change of variables, and α(-,0) , further implies that

4.9 fLMαq,τ,ξ(Rn)qνZ2νατf1B(ξ,2ν)Lq(Rn)τqτνZ2νατk=0f1B(ξ,2ν-k)\B(ξ,2ν-k-1)Lq(Rn)qτqqτk=0νZ2νατf1B(ξ,2ν-k)\B(ξ,2ν-k-1)Lq(Rn)τqτ=k=02kαqνZ2(ν-k)ατf1B(ξ,2ν-k)\B(ξ,2ν-k-1)Lq(Rn)τqτνZ2νατf1B(ξ,2ν)\B(ξ,2ν-1)Lq(Rn)τqτ=fK˙αq,τ,ξ(Rn)q<.

Graph

Therefore, fK˙αq,τ,ξ(Rn) . This then finishes the proof of K˙αq,τ,ξ(Rn)LMαq,τ,ξ(Rn) in this case.

Combining (4.2), (4.4), (4.6), (4.8), and (4.9), we further conclude that LMαq,τ,ξ(Rn)=K˙αq,τ,ξ(Rn) and, for any fLlocq(Rn) , fK˙αq,τ,ξ(Rn)fLMαq,τ,ξ(Rn) with the positive equivalence constants independent of ξ . This then finishes the proof of Lemma 4.3.

Remark 4.4

We should point out that, in Lemma 4.3, if τ= , then the conclusion of Lemma 4.3 was established in [[42], Theorem 26].

We now show Theorem 4.2.

Proof of Theorem 4.2

Observe that α(p,q,r)(-,0] . We now show the present theorem by considering the following two cases on α(p,q,r) .

Case (i) α(p,q,r)=0 . In this case, from the assumption qp , we infer that p=q and r= . If τ(0,) , then, by Theorem 2.9, we find that MF˙q,rp,τ(Rn) is trivial and hence the present theorem obviously holds true. If τ= , then, applying Proposition 3.6(iii), we have

MF˙p,p,(Rn)=Mp,p(Rn)=Mpp(Rn)=Lp(Rn)

Graph

with equivalent quasi-norms. This, together with (4.1), further implies that, for any fLlocq(Rn) ,

T(f)MF˙p,p,(Rn)T(f)Lp(Rn)=supξRn,kZT(f)1B(ξ,2k)\B(ξ,2k-1)Lp(Rn)=supξRnT(f)K˙0p,,ξ(Rn)supξRnfK˙0p,,ξ(Rn)=fLp(Rn)fMF˙p,p,(Rn),

Graph

which completes the proof of the present theorem in this case.

Case (ii) α(p,q,r)(-,0) . In this case, by Proposition 3.14, Lemma 4.3, and (4.1), we conclude that, for any fLlocq(Rn) ,

T(f)MF˙q,rp,τ(Rn)RnT(f)LMα(p,q,r)q,τ,ξ(Rn)rdξ1rRnT(f)K˙α(p,q,r)q,τ,ξ(Rn)rdξ1rRnfK˙α(p,q,r)q,τ,ξ(Rn)rdξ1rRnfLMα(p,q,r)q,τ,ξ(Rn)rdξ1rfMF˙q,rp,τ(Rn).

Graph

This further implies that, in this case, T is bounded on MF˙q,rp,τ(Rn) and hence finishes the proof Theorem 4.2.

As applications of Theorem 4.2, we now establish the boundedness on MF˙q,rp,τ(Rn) of several important operators from harmonic analysis. First, we have the following boundedness of the Hardy–Littlewood maximal operator on MF˙q,rp,τ(Rn) .

Theorem 4.5

If 1<qpr and τ(0,] , then the Hardy–Littlewood maximal operator M is bounded on MF˙q,rp,τ(Rn) .

Proof

We consider the following two cases on both p and r.

Case (i) p<r . In this case, applying the assumptions on p, q, r, and τ of the present theorem, we have α(p,q,r)(-nq,0] . From this and the boundedness of M on Herz spaces (see, for instance, [[22], Theorem 5.1.1 and Remark 5.1.3] or [[20], Corollary 1.5.4]), we deduce that, for any ξRn and fLloc1(Rn) ,

M(f)K˙α(p,q,r)q,τ,ξ(Rn)fK˙α(p,q,r)q,τ,ξ(Rn).

Graph

Applying this and Theorem 4.2, we further find that, for any fLloc1(Rn) ,

M(f)MF˙q,rp,τ(Rn)fMF˙q,rp,τ(Rn),

Graph

that is, M is bounded on MF˙q,rp,τ(Rn) in this case.

Case (ii) p=r . In this case, by Theorem 2.9 and Remark 3.16, we obtain

MF˙q,pp,τ(Rn)={0}whenτ(0,),Lp(Rn)whenτ=andp(1,),Mq(Rn)=L(Rn)whenτ=andp=.

Graph

It is well known that M is bounded on Lp(Rn) when p(1,] . This finishes the proof of Theorem 4.5.

Using [[20], Theorem 1.6.1] and repeating an argument similar to that used in the proof of Theorem 4.5, we obtain the following Fefferman–Stein vector-valued inequality on MF˙q,rp,τ(Rn) ; we omit the details.

Theorem 4.6

Let u(1,] . If 1<qp<r and τ(0,] , or if 1<q<p=r< and τ= , then there exists a positive constant C such that, for any {fj}jNLloc1(Rn) ,

jNM(fj)u1uMF˙q,rp,τ(Rn)CjN|fj|u1uMF˙q,rp,τ(Rn).

Graph

Remark 4.7

  • In Theorems 4.5 and 4.6, if τ=r , then, in this case, by Proposition 3.6(iii), we find that MF˙q,rp,τ(Rn)=Mq,rp(Rn) and hence Theorems 4.5 and 4.6 go back, respectively, to [[16], Lemma 4.1 and Theorem 4.3].
  • We should point out that the assumptions on p, q, r, and τ of Theorems 4.5 and 4.6 are sharp. Indeed, let p=q=1 and r=τ= . Then, in this case, the TLBM space MF˙1,1,(Rn) does not satisfy the assumptions of Theorem 4.5; by Remark 3.16(i), we have MF˙1,1,(Rn)=L1(Rn) and hence M is not bounded on MF˙1,1,(Rn) (see, for instance, [[13], p. 87]). Thus, the assumption q>1 of Theorem 4.5 is sharp. On the other hand, let u(1,) and p=q=r=τ= . Then, in this case, the TLBM space MF˙,,(Rn) does not satisfy the assumptions of Theorem 4.6; from Remark 3.16(i) again, we infer that MF˙,,(Rn)=L(Rn) and hence the Fefferman–Stein vector-valued inequality does not hold true on MF˙,,(Rn) (see, for instance, [[13], Exercise 5.6.4]). This further implies that the assumptions of Theorems 4.5 and 4.6 are sharp.

At the end of this subsection, we establish the boundedness of Calderón–Zygmund operators on MF˙q,rp,τ(Rn) . To this end, we first give some symbols. Let

Δ:=(x,y)Rn×Rn:x=y.

Graph

We now present the following definition of standard kernels; see, for instance, [[13], Definition 7.4.1].

Definition 4.8

Let δ(0,1] . Then a measurable function K on (Rn×Rn)\Δ is called a δ -standard kernel if there exists a positive constant C such that

  • for any x,yRn with xy ,
  • |K(x,y)|≤C|x-y|n;

Graph

  • for any x,y,zRn with |x-y|>2|y-z| ,
  • |K(x,y)-K(x,z)|≤C|y-z|δ|x-y|n+δ

Graph

• and

  • |K(y,x)-K(z,x)|≤C|y-z|δ|x-y|n+δ.

Graph

The following concept of Calderón–Zygmund operators is just [[13], Definition 7.4.2].

Definition 4.9

Let δ(0,1] and K be a δ -standard kernel the same as in Definition 4.8. A linear operator T is called a δ -Calderón–Zygmund operator with a δ -standard kernel K if T is bounded on L2(Rn) and, for any fL2(Rn) with compact support and for almost every xsupp(f)¯ , T has an integral representation

T(f)(x)=RnK(x,y)f(y)dy.

Graph

Next, we show that Calderón–Zygmund operators are bounded on MF˙q,rp,τ(Rn) as follows.

Theorem 4.10

Let δ(0,1] . If 1<qp<r and τ(0,] , or if 1<q<p=r< and τ= , then any δ -Calderón–Zygmund operator T can be extended into a bounded linear operator on MF˙q,rp,τ(Rn) .

Proof

We first show that T is well defined on MF˙q,rp,τ(Rn) . Indeed, let fMF˙q,rp,τ(Rn) . Then, combining Lemma 4.3 and Proposition 3.14, we find that

RnfK˙α(p,q,r)q,τ,ξ(Rn)rdξ1rRnfLMα(p,q,r)q,τ,ξ(Rn)rdξ1rfMF˙q,rp,τ(Rn)<,

Graph

which further implies that there exists a ξ0Rn such that fK˙α(p,q,r)q,τ,ξ0(Rn) . From this and the boundedness of T on the homogeneous Herz space (see, for instance [[22], Theorem 5.1.1 and Remark 5.1.3]), we infer that T(f)K˙α(p,q,r)q,τ,ξ0(Rn) . This then further implies that T is well defined on MF˙q,rp,τ(Rn) .

Now, we prove that T is bounded on MF˙q,rp,τ(Rn) . Indeed, if 1<qp<r and τ(0,] , then, by α(p,q,r)(-nq,0] and [[22], Theorem 5.1.1 and Remark 5.1.3] again, we conclude that, for any ξRn and fK˙α(p,q,r)q,τ,ξ(Rn) ,

T(f)K˙α(p,q,r)q,τ,ξ(Rn)fK˙α(p,q,r)q,τ,ξ(Rn).

Graph

This, combined with Theorem 4.2, further implies that T is bounded on MF˙q,rp,τ(Rn) in this case. If 1<q<p=r<τ= , then, from Proposition 3.1 and [[14], Theorem 4.2.2], we infer that T is bounded on MF˙q,pp,(Rn) . This finishes the proof Theorem 4.10.

Remark 4.11

  • In Theorem 4.10, if τ=r , then, in this case, applying Proposition 3.6(iii), we have MF˙q,rp,τ(Rn)=Mq,rp(Rn) and hence Theorem 4.10 coincides with [[50], Theorem 5.9]. In particular, if further assume that δ=1 and K(x,y)K1(x-y) for any x,yRn and some K1Lloc1(Rn\{0}) , then Theorem 4.10 just goes back to [[16], Theorem 4.7].
  • We should point out that the assumptions on p, q, r, and τ of Theorem 4.10 are sharp. Indeed, the Calderón–Zygmund operator is not bounded on both L1(Rn) and L(Rn) (see, for instance, [[13], Example 5.1.3]). On the other hand, similarly to Remark 4.7(ii), we have MF˙1,1,(Rn)=L1(Rn) and MF˙,,(Rn)=L(Rn) do not satisfy the assumptions of Theorem 4.10. Thus, in this sense, the assumptions of Theorem 4.10 are sharp.
Fractional integrals

The main target of this subsection is to explore the boundedness of both the fractional integral and the fractional maximal operator on TLBM spaces.

Let α(0,n) . Recall that the fractional integral Iα of order α is defined by setting, for any fLloc1(Rn) and xRn ,

Iα(f)(x):=Rnf(y)|x-y|n-αdy

Graph

if the right-hand side makes sense; the fractional maximal operator Mα is defined by setting, for any fLloc1(Rn) and xRn ,

4.10 Mαf(x):=supQx1|Q|1-αnQ|f(y)|dy,

Graph

where the supremum is taken over all the cubes Q of Rn containing x.

Applying Proposition 3.6 and repeating an argument similar to that used in the proof of [[16], Theorem 4.4], we conclude the following boundedness of both Iα and Mα on MF˙q,rp,τ(Rn) , which simultaneously elevates all the four exponents of MF˙q,rp,τ(Rn) ; we omit the details.

Theorem 4.12

Let α(0,n) , 1<q1<p1<r1< , and τ1(0,) . Assume that either 1<q2<p2<r2 and τ2(0,) , or 1<q2p2r2 and τ2= satisfy

1p1-1p2=αnandτ1τ2=r1r2=q1q2=p1p2.

Graph

Then there exists a positive constant C such that, for any fMF˙q1,r1p1,τ1(Rn) ,

Iα(f)MF˙q2,r2p2,τ2(Rn)CfMF˙q1,r1p1,τ1(Rn)

Graph

and

Mα(f)MF˙q2,r2p2,τ2(Rn)CfMF˙q1,r1p1,τ1(Rn).

Graph

Remark 4.13

In Theorem 4.12, if τi=ri for any i{1,2} , then, in this case, by Proposition 3.6(iii), we have

MF˙qi,ripi,τi(Rn)=Mqi,ripi(Rn)

Graph

for any i{1,2} , and hence Theorem 4.12 goes back to [[16], Theorem 4.4 and Corollary 4.5]. Moreover, the other cases of Theorem 4.12 are new.

Next, we are devoted to improving Theorem 4.12 via Calderón products. To do this, we first recall several basic concepts. Recall that a quasi-Banach space X of complex-valued measurable functions is called a quasi-Banach lattice if, for any f,gX , |g||f| implies that gXfX . Assume that X0 and X1 are two quasi-Banach lattices and θ[0,1] . Then their Calderón product, denoted by X01-θX1θ , is defined to be the set of all the measurable functions f satisfying

4.11 |f|f(0)1-θf(1)θ,

Graph

for some f(0)X0 and f(1)X1 , and equipped with the quasi-norm

4.12 fX01-θX1θ:=inff(0)X01-θf(1)X1θ,

Graph

where the infimum is taken over all the f(0)X0 and f(1)X1 satisfying (4.11); see, for instance, [[5], [17]]. Obviously, the TLBM space MF˙q,rp,τ(Rn) is a quasi-Banach lattice.

Via borrowing some ideas from the proofs of [[50], Theorems 5.3 and 5.4], we now establish the following improved boundedness of both Iα and Mα on TLBM spaces via Calderón products.

Theorem 4.14

Let α(0,n) and, for any i{1,2} , pi,qi,ri,τi(1,] be such that

1p1-1p2=αnandτ1τ2=r1r2=q1q2=p1p2.

Graph

  • Assume that either
  • τi∈(1,∞)and1

Graph

• or

  • τi=∞and1

Graph

  • Then there exists a positive constant C such that, for any
  • f∈MF˙q1,r1p1,τ1(Rn)p1p2M1p1(Rn)1-p1p2,‖Mα(f)‖MF˙q2,r2p2,τ2(Rn)≤C‖f‖[MF˙q1,r1p1,τ1(Rn)]p1/p2[M1p1(Rn)]1-p1/p2.

Graph

  • Assume that τi(1,) and 1<qi<pi<ri< for any i{1,2} . Then there exists a positive constant C such that, for any f[MF˙q1,r1p1,τ1(Rn)]p1p2[M1p1(Rn)]1-p1p2 ,
  • 4.13 Iα(f)MF˙q2,r2p2,τ2(Rn)Cf[MF˙q1,r1p1,τ1(Rn)]p1/p2[M1p1(Rn)]1-p1/p2.

Graph

Remark 4.15

  • In Theorem 4.14(i), let τi=ri= and pi[qi,) for any i{1,2} . Then, in this case, MF˙q2,r2p2,τ2(Rn)=Mq2p2(Rn) and
  • [MF˙q1,r1p1,τ1(Rn)]p1p2[M1p1(Rn)]1-p1p2=[Mq1p1(Rn)]p1p2[M1p1(Rn)]1-p1p2

Graph

  • and hence Theorem 4.14(i) just coincides with [[37], Theorem 1.1]. Furthermore, in this case, Sawano and Sugano [[37]] pointed out that the interpolation index θ:=1-p1/p2 is optimal. Therefore, in this sense, the results obtained in Theorem 4.14 are sharp.
  • In Theorem 4.14, when τi=ri(1,) for any i{1,2} , then MF˙q2,r2p2,τ2(Rn)=Mq2,r2p2(Rn) and
  • [MF˙q1,r1p1,τ1(Rn)]p1p2[M1p1(Rn)]1-p1p2=[Mq1,r1p1(Rn)]p1p2[M1p1(Rn)]1-p1p2.

Graph

  • In this case, Theorem 4.14 goes back to [[50], Theorems 5.3 and 5.4 and Remark 5.5]. Moreover, by [[50], Proposition 4.11], we have
  • Mq1,r1p1(Rn)⫋[Mq1,r1p1(Rn)]p1p2[M1p1(Rn)]1-p1p2.

Graph

  • In this sense, Theorem 4.14 indeed improves Theorem 4.12.

Via the boundedness of the Hardy–Littlewood maximal operator on TLBM spaces, we next show Theorem 4.14(i).

Proof of Theorem 4.14(i)

We only show the present theorem under the assumption that τi(1,) and, for any i{1,2} , 1<qi<pi<ri< because the cases that τi= or ri= are quite similar and hence we omit the details. Let f[MF˙q1,r1p1,τ1(Rn)]p1p2[M1p1(Rn)]1-p1p2 , f0MF˙q1,r1p1,τ1(Rn) , and f1M1p1(Rn) satisfy |f||f0|p1p2|f1|1-p1p2 almost everywhere on Rn . Then, from the Hölder inequality and the assumption that 1/p1-1/p2=α/n , we deduce that, for any R(0,) and xRn ,

RαmQ(x,R)(|f|)RαmQ(x,R)(|f0|)p1p2mQ(x,R)(|f1|)1-p1p2M(f0)(x)p1p2f1M1p1(Rn)1-p1p2.

Graph

This, together with (4.10), further implies that, for any xRn ,

Mα(f)(x)M(f0)(x)p1p2f1M1p1(Rn)1-p1p2,

Graph

where the implicit positive constant depends only on both n and α . By this, r1(1,) , and Theorem 4.5, we conclude that

Mα(f)MF˙q2,r2p2,τ2(Rn)M(f0)p1p2MF˙q2,r2p2,τ2(Rn)f1M1p1(Rn)1-p1p2M(f0)MF˙q1,r1p1,τ1(Rn)p1p2f1M1p1(Rn)1-p1p2f0MF˙q1,r1p1,τ1(Rn)p1p2f1M1p1(Rn)1-p1p2,

Graph

which, combined with (4.12) and the choice of both f0 and f1 , further implies that

Mα(f)MF˙q2,r2p2,τ2(Rn)f[MF˙q1,r1p1,τ1(Rn)]p1/p2[M1p1(Rn)]1-p1/p2.

Graph

This finishes the proof of Theorem 4.14(i).

Next, we turn to show Theorem 4.14(ii). To achieve this, we require some concepts about sparse families on Euclidean spaces. Recall that a dyadic grid D is defined to be set of countable cubes, which has the following three properties:

  • if QD , then l(Q)=2k for some kZ ;
  • if P,QD , then PQ{,P,Q} ;
  • for any kZ , one has
  • Rn=⋃Q∈D,l(Q)=2kQ.

Graph

For any dyadic grid D , a set SD is said to be sparse if, for any QS ,

QS,QQQ12|Q|.

Graph

Let S be a sparse family. Then, as was proved in [[7], (13)], there exist {EQ}QS of disjoint measurable sets such that, for any QS ,

4.14 EQQand|EQ|12|Q|.

Graph

The following pointwise estimate about the fractional integral and the sparse family is a simple corollary of [[7], Propositions 2.2 and 2.3], which plays an essential role in the proof of Theorem 4.14(ii); we omit the details.

Lemma 4.16

Let α(0,n) and fLc(Rn) be a non-negative function. Then there exist {Di}i=12n of dyadic grids and {Si}i=12n of sparse families satisfying that, for any i{1,...,2n} , SiDi and there exists a positive constant C, independent of f, such that

Iα(f)Cmaxi{1,...,2n}QSi[l(Q)]αmQ(f)1Q,

Graph

where mQ(·) is the same as in (2.13).

Based on this estimate, the boundedness of the fractional maximal operator established in Theorem 4.14(i), and the Fefferman–Stein vector-valued inequality on TLBM spaces, we now show Theorem 4.14(ii).

Proof of Theorem 4.14(ii)

We first show the present theorem for any non-negative function fLc(Rn) . Indeed, by Lemma 4.16, we conclude that it suffices to prove that, for any sparse family S and any non-negative function fLc(Rn) ,

4.15 QS[l(Q)]αmQ(f)1QMF˙q2,r2p2,τ2(Rn)f[MF˙q1,r1p1,τ1(Rn)]p1/p2[M1p1(Rn)]1-p1/p2.

Graph

To do this, let {EQ}QS be the same as in (4.14). Then, using the fact that 1Q2M(1EQ) for any QS , r(1,), and Theorem 4.6, we find that

QS[l(Q)]αmQ(f)1QMF˙q2,r2p2,τ2(Rn)QSM[l(Q)]α2[mQ(f)]121EQ2MF˙q2,r2p2,τ2(Rn)=QSM[l(Q)]α2[mQ(f)]121EQ212MF˙2q2,2r22p2,2τ2(Rn)2QS[l(Q)]αmQ(f)1EQ12MF˙2q2,2r22p2,2τ2(Rn)2=QS[l(Q)]αmQ(f)1EQMF˙q2,r2p2,τ2(Rn).

Graph

Therefore, to show (4.15), we only need to prove that, for any non-negative function fLc(Rn) ,

QS[l(Q)]αmQ(f)1EQMF˙q2,r2p2,τ2(Rn)f[MF˙q1,r1p1,τ1(Rn)]p1/p2[M1p1(Rn)]1-p1/p2.

Graph

Indeed, from the disjointness of {EQ}QS , we infer that, for any non-negative function fLc(Rn) ,

QS[l(Q)]αmQ(f)1EQMα(f).

Graph

This, together with Theorem 4.14(i), further implies that (4.15) holds true and hence (4.13) also holds true for any non-negative function fLc(Rn) .

Next, we show that (4.13) holds true for any f[MF˙q1,r1p1,τ1(Rn)]p1/p2[M1p1(Rn)]1-p1/p2 . To do this, let f[MF˙q1,r1p1,τ1(Rn)]p1/p2[M1p1(Rn)]1-p1/p2 . Then f=(f)+-(f)-+i[(f)+-(f)-] and hence

4.16 Iα(f)MF˙q2,r2p2,τ2(Rn)Iα((f)+)MF˙q2,r2p2,τ2(Rn)+Iα((f)-)MF˙q2,r2p2,τ2(Rn)+Iα((f)+)MF˙q2,r2p2,τ2(Rn)+Iα((f)-)MF˙q2,r2p2,τ2(Rn).

Graph

For any mN , let (f)+,m be the same as in (2.23) with f replaced by (f)+ . Then, for any mN , (f)+,mLc(Rn) is non-negative. This, combined with the Fatou lemma, r2,τ2(1,) , and (2.24), further implies that

4.17 Iα((f)+)MF˙q2,r2p2,τ2(Rn)lim̲mIα((f)+,m)MF˙q2,r2p2,τ2(Rn)limm(f)+,m[MF˙q1,r1p1,τ1(Rn)]p1/p2[M1p1(Rn)]1-p1/p2=(f)+[MF˙q1,r1p1,τ1(Rn)]p1/p2[M1p1(Rn)]1-p1/p2f[MF˙q1,r1p1,τ1(Rn)]p1/p2[M1p1(Rn)]1-p1/p2.

Graph

Similarly, we have

Iα((f)-)MF˙q2,r2p2,τ2(Rn)f[MF˙q1,r1p1,τ1(Rn)]p1/p2[M1p1(Rn)]1-p1/p2,Iα((f)+)MF˙q2,r2p2,τ2(Rn)f[MF˙q1,r1p1,τ1(Rn)]p1/p2[M1p1(Rn)]1-p1/p2,

Graph

and

Iα((f)-)MF˙q2,r2p2,τ2(Rn)f[MF˙q1,r1p1,τ1(Rn)]p1/p2[M1p1(Rn)]1-p1/p2.

Graph

These, combined with both (4.16) and (4.17), further imply that

Iα(f)MF˙q2,r2p2,τ2(Rn)f[MF˙q1,r1p1,τ1(Rn)]p1/p2[M1p1(Rn)]1-p1/p2,

Graph

which completes the proof of Theorem 4.14(ii) and hence Theorem 4.14.

Further remarks

In this section, we give some remarks.

Let S(Rn) be the space of all Schwartz functions on Rn , whose topology is determined by a family of norms, {·SM(Rn)}MZ+ , where, for any MZ+ and φS(Rn) ,

φSM(Rn):=sup{αZ+n,|α|M}supxRn1+|x|n+M+|α|αφ(x)

Graph

with |α|:=α1++αn for any α:=(α1,...,αn)Z+n . As in [[45], [47]], let

S(Rn):=φS(Rn):Rnφ(x)xγdx=0for anyγZ+n,

Graph

regarded as a subspace of S(Rn) with the same topology. We denote by S(Rn) the space of all continuous linear functionals on S(Rn) , equipped with the weak- topology. Let φS(Rn) be such that

5.1 suppφ^xRn:12|x|2

Graph

and there exists a positive constant C satisfying that

5.2 φ^(x)C>0for anyxRnwith35|x|53,

Graph

where, for any ξRn ,

φ^(ξ):=Rnf(x)e-2πix·ξdx.

Graph

For any t(0,) , let φt(·):=t-nφ(·/t) . The Littlewood–Paley operators {Δjφ}jZ and {Δtφ}t(0,) of φ are defined, respectively, by setting, for any jZ , t(0,) , and fS(Rn) ,

5.3 Δjφ(f):=φ2-jfandΔtφ(f):=φtf.

Graph

Combining the structure of both MF˙q,rp,τ(Rn) and MB˙q,rp,τ(Rn) and the Littlewood–Paley operators, we next introduce the following two new function spaces.

Definition 5.1

Let sR , q,r,τ(0,] , and φS(Rn) satisfy both (5.1) and (5.2).

  • The Bourgain–Morrey–Triebel–Lizorkin space F˙Mq,rs,τ(Rn;φ) is defined to be the set of all the fS(Rn) such that fF˙Mq,rs,τ(Rn;φ)< , where
  • ‖f‖F˙Mq,rs,τ(Rn;φ):=∫0∞ts-nqΔtφ(f)1B(·,t)Lq(Rn)τdtt1τLr(Rn)

Graph

  • when τ(0,) and where
  • ‖f‖F˙Mq,rs,τ(Rn;φ):=supt∈(0,∞)ts-nqΔtφ(f)1B(·,t)Lq(Rn)Lr(Rn)

Graph

  • when τ= .
  • The Bourgain–Morrey–Besov space B˙Mq,rs,τ(Rn;φ) is defined to be the set of all the fS(Rn) such that fB˙Mq,rs,τ(Rn;φ)< , where
  • ‖f‖B˙Mq,rs,τ(Rn;φ):=∑j∈Z2jsτ∑m∈ZnΔjφ(f)1Qj,mLq(Rn)rτr1τ

Graph

  • with the usual modifications made when q= , r= , or τ= .
Remark 5.2

We should point out that, in Definition 5.1(i), if τ=q and r(0,) , then, in this case, by [[46], Theorem 2.8], we find that F˙Mq,rs,q(Rn;φ) goes back to the Triebel–Lizorkin space F˙q,rs(Rn) (see also [[45]]). In addition, if r=q in Definition 5.1(ii), then, in this case, B˙Mq,qs,τ(Rn;φ) is just the Besov space B˙q,τs(Rn) ; see, for instance, [[36], [45]].

We will develop a real-variable theory of the above two new function spaces in a forthcoming article.

Acknowledgements

Pingxu Hu and Yinqin Li would like to thank Jin Tao and Yirui Zhao for many helpful discussions on the subject of this article. The authors would also like to thank the referee for his/her carefully reading and several useful and helpful comments which definitely improve the presentation of this article.

Data availability

Data sharing is not applicable to this article as no data sets were generated or analysed.

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

References 1 Adams DR. Morrey Spaces. Lecture Notes in Applied and Numerical Harmonic Analysis. 2015: Cham; Birkhäuser/Springer 2 Bégout P, Vargas A. Mass concentration phenomena for the L2\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$L^2$$\end{document}-critical nonlinear Schrödinger equation. Trans. Am. Math. Soc.. 2007; 359: 5257-5282. 10.1090/S0002-9947-07-04250-X. 1171.35109 3 Bourgain, J.: On the restriction and multiplier problems in R3\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$${{\mathbb{R}}}^3$$\end{document}. In: Geometric Aspects of Functional Analysis (1989–90), pp. 179–191, Lecture Notes in Mathematics, vol. 1469. Springer, Berlin (1991) 4 Bourgain J. Refinements of Strichartz' inequality and applications to 2D-NLS with critical nonlinearity. Int. Math. Res. Notices. 1998; 1998; 5: 253-283. 10.1155/S1073792898000191. 0917.35126 5 Calderón A-P. Intermediate spaces and interpolation, the complex method. Studia Math.. 1964; 24: 113-190. 167830. 10.4064/sm-24-2-113-190. 0204.13703 6 Chen D, Chen X, Sun L. Well-posedness of the Euler equation in Triebel–Lizorkin–Morrey spaces. Appl. Anal.. 2020; 99: 772-795. 4075474. 10.1080/00036811.2018.1510491. 1439.35379 7 Cruz-Uribe D, Moen K. One and two weight norm inequalities for Riesz potentials. Ill. J. Math.. 2013; 57: 295-323. 3224572. 1297.42022 8 del Campo R, Fernández A, Mayoral F, Naranjo F. Orlicz spaces associated to a quasi-Banach function space: applications to vector measures and interpolation. Collect. Math.. 2021; 72: 481-499. 4297141. 10.1007/s13348-020-00295-1. 1482.46031 9 Di Fazio G, Nguyen T. Regularity estimates in weighted Morrey spaces for quasilinear elliptic equations. Rev. Mat. Iberoam.. 2020; 36: 1627-1658. 4163979. 10.4171/rmi/1178. 1467.35156 Di Fazio G, Ragusa MA. Interior estimates in Morrey spaces for strong solutions to nondivergence form equations with discontinuous coefficients. J. Funct. Anal.. 1993; 112: 241-256. 1213138. 10.1006/jfan.1993.1032. 0822.35036 Feuto J. Norm inequalities in generalized Morrey spaces. J. Fourier Anal. Appl.. 2014; 20: 896-909. 3232590. 10.1007/s00041-014-9337-2. 1320.47048 Folland, G.B.: Real Analysis. Modern Techniques and Their Applications, 2nd edn, Pure and Applied Mathematics (New York). Wiley, New York (1999) Grafakos, L.: Classical Fourier Analysis, 3rd edn, Graduate Texts in Mathematics, vol. 249. Springer, New York (2014) Grafakos, L.: Modern Fourier Analysis, 3rd edn, Graduate Texts in Math., vol. 250. Springer, New York (2014) Guliev VS, Mustafaev RCh. Fractional integrals in spaces of functions defined on spaces of homogeneous type. Anal. Math.. 1998; 24: 181-200. 1639211. 0917.42019 Hatano, N., Nogayama, T., Sawano, Y., Hakim, D.I.: Bourgain–Morrey spaces and their applications to boundedness of operators. J. Funct. Anal. 284, Paper No. 109720 (2023) Kalton, N., Mayboroda, S., Mitrea, M.: Interpolation of Hardy–Sobolev–Besov–Triebel–Lizorkin spaces and applications to problems in partial differential equations. In: Interpolation Theory and Applications, pp. 121–177, Contemp. Math., vol. 445. Amer. Math. Soc., Providence (2007) Kenig, C.E., Ponce, G., Vega, L.: On the concentration of blow up solutions for the generalized KdV equation critical in L2\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$L^2$$\end{document}. In: Nonlinear Wave Equations (Providence, RI, 1998), pp. 131–156, Contemp. Math., vol. 263. Amer. Math. Soc., Providence (2000) Lemarié-Rieusset PG. The Navier–Stokes Problem in the 21st Century. 2016: Boca Raton; CRC Press. 10.1201/b19556. 1342.76029 Li, Y., Yang, D., Huang, L.: Real-Variable Theory of Hardy Spaces Associated with Generalized Herz Spaces of Rafeiro and Samko. Lecture Notes in Mathematics, vol. 2320. Springer, Cham (2022) Lu S, Yang D. The local versions of Hp(Rn)\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$H^p({{\mathbb{R} }^n})$$\end{document} spaces at the origin. Studia Math.. 1995; 116: 103-131. 1354135. 10.4064/sm-116-2-103-131. 0935.42012 Lu S, Yang D, Hu G. Herz Type Spaces and Their Applications. 2008: Beijing; Science Press Masaki, S.: Two minimization problems on non-scattering solutions to mass-subcritical nonlinear Schrödinger equation. arXiv:1605.09234 Masaki, S., Segata, J.: Existence of a minimal non-scattering solution to the mass-subcritical generalized Korteweg–de Vries equation. Ann. Inst. H. Poincaré C Anal. Non Linéaire 35, 283–326 (2018) Masaki S, Segata J. Refinement of Strichartz estimates for Airy equation in nondiagonal case and its application. SIAM J. Math. Anal.. 2018; 50: 2839-2866. 3809534. 10.1137/17M1153893. 1392.35265 Mastyło M, Sawano Y. Complex interpolation and Calderón–Mityagin couples of Morrey spaces. Anal. PDE. 2019; 12: 1711-1740. 3986539. 10.2140/apde.2019.12.1711. 1435.46018 Mastyło M, Sawano Y. Applications of interpolation methods and Morrey spaces to elliptic PDEs. Ann. Sc. Norm. Super. Pisa Cl. Sci. (5). 2020; 21: 999-1021. 4288625. 1480.46032 Mastyło M, Sawano Y, Tanaka H. Morrey-type space and its Köthe dual space. Bull. Malays. Math. Sci. Soc.. 2018; 41: 1181-1198. 3818410. 10.1007/s40840-016-0382-7. 1393.42007 Morrey CB. On the solutions of quasi-linear elliptic partial differential equations. Trans. Am. Math. Soc.. 1938; 43: 126-166. 1501936. 10.1090/S0002-9947-1938-1501936-8. 0018.40501 Moyua, A., Vargas, A., Vega, L.: Schrödinger maximal function and restriction properties of the Fourier transform. Internat. Math. Res. Notices, pp. 793–815 (1996) Moyua A, Vargas A, Vega L. Restriction theorems and maximal operators related to oscillatory integrals in R3\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$${{\mathbb{R} }}^3$$\end{document}. Duke Math. J.. 1999; 96: 547-574. 1671214. 10.1215/S0012-7094-99-09617-5. 0946.42011 Nakai E. Hardy–Littlewood maximal operator, singular integral operators and the Riesz potentials on generalized Morrey spaces. Math. Nachr.. 1994; 166: 95-103. 1273325. 10.1002/mana.19941660108. 0837.42008 Nakai E. Orlicz–Morrey spaces and the Hardy–Littlewood maximal function. Studia Math.. 2008; 188: 193-221. 2429821. 10.4064/sm188-3-1. 1163.46020 Nakai, E.: Pointwise multipliers on Musielak–Orlicz–Morrey spaces. In: Function Spaces and Inequalities, pp. 257–281, Springer Proc. Math. Stat., vol. 206. Springer, Singapore (2017) Nakai E, Sadasue G. Commutators of fractional integrals on martingale Morrey spaces. Math. Inequal. Appl.. 2019; 22: 631-655. 3937732. 1488.60115 Sawano, Y.: Theory of Besov Spaces, Dev. Math., vol. 56. Springer, Singapore (2018) Sawano Y, Sugano S. Complex interpolation and the Adams theorem. Potential Anal.. 2021; 54: 299-305. 4202741. 10.1007/s11118-020-09827-7. 1477.42025 Sawano Y, Tanaka H. Morrey spaces for non-doubling measures. Acta Math. Sin. (Engl. Ser.). 2005; 21: 1535-1544. 2190025. 10.1007/s10114-005-0660-z. 1129.42403 Sawano Y, Sugano S, Tanaka H. Generalized fractional integral operators and fractional maximal operators in the framework of Morrey spaces. Trans. Am. Math. Soc.. 2011; 363: 6481-6503. 2833565. 10.1090/S0002-9947-2011-05294-3. 1229.42024 Sawano Y, Sugano S, Tanaka H. Orlicz–Morrey spaces and fractional operators. Potential Anal.. 2012; 36: 517-556. 2904632. 10.1007/s11118-011-9239-8. 1242.42017 Sawano Y, Di Fazio G, Hakim D. Morrey Spaces: Introduction and Applications to Integral Operators and PDE's. 2020: New York; Chapman and Hall/CRC. 10.1201/9780429085925. 1482.42002 Sawano Y, Di Fazio G, Hakim D. Morrey Spaces: Introduction and Applications to Integral Operators and PDE's. 2020: New York; Chapman and Hall/CRC. 10.1201/9780429085925. 1482.42002 Shao S. The linear profile decomposition for the airy equation and the existence of maximizers for the airy Strichartz inequality. Anal. PDE. 2009; 2: 83-117. 2561172. 10.2140/apde.2009.2.83. 1185.35239 Shen Z. Boundary value problems in Morrey spaces for elliptic systems on Lipschitz domains. Am. J. Math.. 2003; 125: 1079-1115. 2004429. 10.1353/ajm.2003.0035. 1046.35029 Triebel H. Theory of Function Spaces, Monographs in Mathematics. 1983: Basel; Birkhuser Verlag. 10.1007/978-3-0346-0416-1 Ullrich, T.: Continuous characterizations of Besov–Lizorkin–Triebel spaces and new interpretations as coorbits. J. Funct. Spaces Appl., Art. ID 163213 (2012) Yang D, Yuan W. A new class of function spaces connecting Triebel–Lizorkin spaces and Q\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$Q$$\end{document} spaces. J. Funct. Anal.. 2008; 255: 2760-2809. 2464191. 10.1016/j.jfa.2008.09.005. 1169.46016 Yuan, W., Sickel, W., Yang, D.: Morrey and Campanato Meet Besov, Lizorkin and Triebel. Lecture Notes in Mathematics, vol. 2005. Springer, Berlin (2010) Zhang, J., Yang, Y., Zhang, Q.: On the stability to Keller–Segel system coupled with Navier–Stokes equations in Besov–Morrey spaces. Nonlinear Anal. Real World Appl. 71, Paper No. 103828 (2023) Zhao, Y., Sawano, Y., Tao, J., Yang, D., Yuan, W.: Bourgain–Morrey spaces mixed with structure of Besov spaces. Proc. Steklov Inst. Math. (2023) (submitted)

Reported by Author; Author; Author

Titel:
Bourgain–Morrey spaces meet structure of Triebel–Lizorkin spaces.
Autor/in / Beteiligte Person: Hu, Pingxu ; Li, Yinqin ; Yang, Dachun
Link:
Zeitschrift: Mathematische Zeitschrift, Jg. 304 (2023-05-01), Heft 1, S. 1-49
Veröffentlichung: 2023
Medientyp: academicJournal
ISSN: 0025-5874 (print)
DOI: 10.1007/s00209-023-03282-x
Schlagwort:
  • CALDERON-Zygmund operator
  • NONLINEAR differential equations
  • PARTIAL differential equations
  • FRACTIONAL integrals
  • FUNCTION spaces
  • MAXIMAL functions
  • Subjects: CALDERON-Zygmund operator NONLINEAR differential equations PARTIAL differential equations FRACTIONAL integrals FUNCTION spaces MAXIMAL functions
  • (Triebel–Lizorkin–)Bourgain–Morrey space
  • 42B20
  • 42B25
  • 42B35
  • Cube
  • Global Morrey space
  • Herz space
  • Maximal operator
  • Primary 46E35
  • Secondary 46B70
Sonstiges:
  • Nachgewiesen in: DACH Information
  • Sprachen: English
  • Document Type: Article
  • Author Affiliations: 1 = Laboratory of Mathematics and Complex Systems (Ministry of Education of China), School of Mathematical Sciences, Beijing Normal University, 100875, Beijing, People’s Republic of China
  • Full Text Word Count: 26006

Klicken Sie ein Format an und speichern Sie dann die Daten oder geben Sie eine Empfänger-Adresse ein und lassen Sie sich per Email zusenden.

oder
oder

Wählen Sie das für Sie passende Zitationsformat und kopieren Sie es dann in die Zwischenablage, lassen es sich per Mail zusenden oder speichern es als PDF-Datei.

oder
oder

Bitte prüfen Sie, ob die Zitation formal korrekt ist, bevor Sie sie in einer Arbeit verwenden. Benutzen Sie gegebenenfalls den "Exportieren"-Dialog, wenn Sie ein Literaturverwaltungsprogramm verwenden und die Zitat-Angaben selbst formatieren wollen.

xs 0 - 576
sm 576 - 768
md 768 - 992
lg 992 - 1200
xl 1200 - 1366
xxl 1366 -