This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Interpolation on Convex-set Valued Lebesgue Spaces and its Applications
Yuxun Zhang, Jiang Zhou*

Abstract: In this paper, we obtain two interpolation theorems on convex-set valued Lebesgue spaces, which generalize the Marcinkiewicz interpolation theorem and Riesz-Thorin interpolation theorem on classical Lebesgue spaces, respectively. As applications, we obtain the boundedness of convex-set valued fractional averaging operators and fractional maximal operators, and prove the reverse factorization property of matrix weights.

Key Words: Interpolation; convex-set valued Lebesgue spaces; fractional maximal operators; matrix weights

Mathematics Subject Classification (2020): 26E25; 42B25; 42B35

1 Introduction

In 1972, the classical ApA_{p} weights were first introduced by Muckenhoupt[1] during the study of weighted norm inequalities for Hardy-Littlewood maximal operator. A weight (a non-negative, measurable function that satisfies 0<ω(x)<0<\omega(x)<\infty, a.e.) ω\omega is said to satisfy ωAp\omega\in A_{p} (1<p<1<p<\infty) if

[ω]Ap:=supQ(1mn(Q)Qω(x)𝑑x)(1mn(Q)Qω(x)11p𝑑x)p1<,[\omega]_{A_{p}}:=\sup_{Q}\left(\frac{1}{m_{n}(Q)}\int_{Q}\omega(x)dx\right)\left(\frac{1}{m_{n}(Q)}\int_{Q}\omega(x)^{\frac{1}{1-p}}dx\right)^{p-1}<\infty,

where the supremum is taken over all cubes QnQ\subset\mathbb{R}^{n} with sides parallel to the coordinate axes, mn(Q)m_{n}(Q) denotes the n-dimensional Lebesgue measure of QQ. A weight ω\omega is said to satisfy ωA1\omega\in A_{1} if for all xnx\in\mathbb{R}^{n},

Mω(x)Cω(x).M\omega(x)\leqslant C\omega(x).

For dd\in\mathbb{N}, given a Calderón-Zygmund singular integral operator TT, it can be extended to an operator on vector-valued functions f=(f1,f2,,fd)t\vec{f}=(f_{1},f_{2},\cdots,f_{d})^{t} by applying it to each coordinate: Tf:=(Tf1,Tf2,,Tfd)tT\vec{f}:=(Tf_{1},Tf_{2},\cdots,Tf_{d})^{t}.

Denote d\mathcal{M}_{d} to be the set of all d×dd\times d real matrices, a matrix function WW is a mapping W:ndW:\mathbb{R}^{n}\rightarrow\mathcal{M}_{d}. In a series of papers in the 1990s, Nazarov, Treil and Volberg[2, 3, 4] considered that for 1<p<1<p<\infty, what type of positive semidefinite, self-adjoint matrix functions WW make the weighted inequality

n|W1p(x)T(f)(x)|p𝑑xCn|W1p(x)f(x)|p𝑑x\int_{\mathbb{R}^{n}}|W^{\frac{1}{p}}(x)T(\vec{f})(x)|^{p}dx\leqslant C\int_{\mathbb{R}^{n}}|W^{\frac{1}{p}}(x)\vec{f}(x)|^{p}dx (1)

hold, and defined the matrix weight class Ap,matrixA_{p,matrix} to be the set of such matrix functions WW, where UvdUv\in\mathbb{R}^{d} denotes the product of UdU\in\mathcal{M}_{d} and vdv\in\mathbb{R}^{d}, |v||v| denotes the Euclidean norm of vdv\in\mathbb{R}^{d}.

In 2003, Roudenko[5] gave a characterization of matrix weights: WAp,matrixW\in A_{p,matrix} (1<p<)(1<p<\infty) if and only if

[W]Ap,matrix:=supQ(1mn(Q)Q(1mn(Q)QW1p(x)W1p(y)opp𝑑y)pp𝑑x)1p<,[W]_{A_{p,matrix}}:=\sup_{Q}\left(\frac{1}{m_{n}(Q)}\int_{Q}\left(\frac{1}{m_{n}(Q)}\int_{Q}\lVert W^{\frac{1}{p}}(x)W^{-\frac{1}{p}}(y)\rVert_{\mathrm{op}}^{p^{\prime}}dy\right)^{\frac{p}{p^{\prime}}}dx\right)^{\frac{1}{p}}<\infty, (2)

where the supremum is taken over all cubes QnQ\subset\mathbb{R}^{n} with sides parallel to the coordinate axes, Uop\lVert U\rVert_{\mathrm{op}} denotes the operator norm of the matrix UdU\in\mathcal{M}_{d}, that is,

Uop=maxvd|Uv||v|.\lVert U\rVert_{\mathrm{op}}=\max_{v\in\mathbb{R}^{d}}\frac{|Uv|}{|v|}.
Remark 1.1.

The following discussion indicates that matrix Ap,matrixA_{p,matrix} weights are the generalization of classical ApA_{p} weights: while d=1d=1, the matrix function WW will be a non-negative real valued function ω\omega, then the inequality (1) will turn into

n|T(f)(x)|pω(x)𝑑xCn|f(x)|pω(x)𝑑x,\int_{\mathbb{R}^{n}}|T(f)(x)|^{p}\omega(x)dx\leqslant C\int_{\mathbb{R}^{n}}|f(x)|^{p}\omega(x)dx,

and the condition (2) will turn into

supQ(1mn(Q)Qω(x)𝑑x)1p(1mn(Q)Qω(x)pp𝑑x)1p<,\sup_{Q}\left(\frac{1}{m_{n}(Q)}\int_{Q}\omega(x)dx\right)^{\frac{1}{p}}\left(\frac{1}{m_{n}(Q)}\int_{Q}\omega(x)^{-\frac{p^{\prime}}{p}}dx\right)^{\frac{1}{p^{\prime}}}<\infty,

which are the weighted norm inequality and the classical ApA_{p} condition, respectively.

Jones factorization theorem[6] and Rubio de Francia extrapolation[7] are two important theorems in classical ApA_{p} weight theory. In 1996, Nazarov and Treil[2] conjectured that these results can be generalized to matrix weights, which generated two longstanding open problems. In 2022, Bownik and Cruz-uribe[8] solved these problems by creatively utilizing convex analysis theory (see, for instance, [9, 10, 11]). They defined some new spaces consisted of convex-set valued functions, including L𝒦p(Ω,ρ)L_{\mathcal{K}}^{p}(\Omega,\rho) and their special form L𝒦p(Ω,||)L_{\mathcal{K}}^{p}(\Omega,|\cdot|) (1p)(1\leqslant p\leqslant\infty), and furthermore obtained that for 1<p1<p\leqslant\infty, a norm function ρ\rho satisfying 𝒜p,norm\mathcal{A}_{p,norm} condition (see[8, Definition 6.2]) ensures the boundedness of convex-set valued maximal operator on L𝒦p(Ω,ρ)L_{\mathcal{K}}^{p}(\Omega,\rho), which enables the Rubio de Francia iteration algorithm to be used and this conjecture to be ultimately proved.

Interpolation is a useful method to obtain the boundedness of several sublinear operators. In 1939, Marcinkiewicz[12] first put forward the so-called Marcinkiewicz interpolation theorem, and Zygmund[13] reintroduced it in 1956. In 1982, in order to study the commutators generated by fractional integrals and BMO functions, Chanillo [14] obtained the boundedness of fractional maximal operators first introduced in[15] by using a generalization of Marcinkiewicz interpolation theorem in[16].

Riesz[17] first proved the Riesz convexity theorem in 1927, which is the original version of Riesz-Thorin interpolation theorem. After that, Thorin [18, 19] generalized this result, and published his theses with a variety of applications. Riesz-Thorin interpolation theorem requires strong boundedness of operators at endpoints, but has the better constant and lower requirment of exponents, and is easier to be generalized. For example, in 1958, Stein and Weiss obtained the Stein-Weiss variable measure interpolation theorem[20], which can be used to obtain the reverse factorization property of ApA_{p} weights.

To be specific, given 1<p0,p1<1<p_{0},p_{1}<\infty, suppose that w0Ap0w_{0}\in A_{p_{0}} and w1Ap1w_{1}\in A_{p_{1}}, the classical ApA_{p} weight theory[1] indicates that

MfLw0p0C0fLw0p0,MfLw1p1C1fLw1p1,\lVert Mf\rVert_{L^{p_{0}}_{w_{0}}}\leqslant C_{0}\lVert f\rVert_{L^{p_{0}}_{w_{0}}},\ \ \lVert Mf\rVert_{L^{p_{1}}_{w_{1}}}\leqslant C_{1}\lVert f\rVert_{L^{p_{1}}_{w_{1}}},

where MM is the Hardy-Littlewood operator, and for r>0r>0,

gLwr=(n|g(x)|rw(x)𝑑x)1r.\lVert g\rVert_{L^{r}_{w}}=\left(\int_{\mathbb{R}^{n}}|g(x)|^{r}w(x)dx\right)^{\frac{1}{r}}.

Then by Stein-Weiss variable measure interpolation theorem[20], for

1p=1tp0+tp1,w=w0p(1t)p0w1ptp1,C=C01tC1t,\frac{1}{p}=\frac{1-t}{p_{0}}+\frac{t}{p_{1}},\ \ w=w_{0}^{\frac{p(1-t)}{p_{0}}}w_{1}^{\frac{pt}{p_{1}}},\ \ C=C_{0}^{1-t}C_{1}^{t},

there holds MfLwpCfLwp\lVert Mf\rVert_{L^{p}_{w}}\leqslant C\lVert f\rVert_{L^{p}_{w}}, which implies that wApw\in A_{p}.

In this paper, we will further investigate the properties of convex-set valued Lebesgue spaces, especially establish the interpolation theory above them, including Marcinkiewicz interpolation theorem and Riesz-Thorin interpolation theorem. These results not only enrich and develop set-valued analysis theory, but also have some applications in harmonic analysis.

The remainder of this paper is organized as follows. In Section 2, we give some necessary definitions and lemmas. In Section 3, we prove the Marcinkiewicz interpolation theorem, define the fractional averaging operator and fractional maximal operator on convex-set valued Lebesgue spaces, and obtain their boundedness. In Section 4, we prove the Riesz-Thorin interpolation theorem, and use it with some known results of norm functions to obtain the reverse factorization property of matrix weights, which is a part of Jones factorization theorem first proved in[8].

Throughout this paper we will use the following notations. We will use n\mathbb{R}^{n} to denote Euclidean space of dimension nn, which will be the domain of functions, and use dd to denote the dimension of vector and set-valued functions. We will use (Ω,𝒜,μ)(\Omega,\mathcal{A},\mu) to denote a positive, σ\sigma-infinite, and complete measurable spaces, where Ωn\Omega\subset\mathbb{R}^{n}. The open unit ball {vd:|v|<1}\{v\in\mathbb{R}^{d}:|v|<1\} will be denoted by 𝑩\boldsymbol{B} and its closure by 𝑩¯\boldsymbol{\overline{B}}. The set of all d×dd\times d, real symmetric, positive semidefinite matrices will be denoted by 𝒮d\mathcal{S}_{d}. For 1p1\leqslant p\leqslant\infty, pp^{\prime} will denote the conjugate index of pp such that 1/p+1/p=11/p+1/p^{\prime}=1, Lp(Ω)L^{p}(\Omega) will denote the Lebesgue space of real-valued functions on Ω\Omega, and Lp(Ω,d)L^{p}(\Omega,\mathbb{R}^{d}) will denote the Lebesgue space of vector-valued functions on Ω\Omega. Given two quantities AA and BB, we will write ABA\lesssim B if there exists a constant C>0C>0 such that ACBA\leqslant CB, and write ABA\approx B if ABA\lesssim B and BAB\lesssim A.

2 Preliminary

We begin with some basic definitions.

Definition 2.1.

[11] For E,FdE,F\subset\mathbb{R}^{d}, define E¯\overline{E} to be the closure of EE, E+F={x+y:xE,yF}E+F=\{x+y:x\in E,y\in F\}. For λ\lambda\in\mathbb{R}, define λE={λx:xE}\lambda E=\{\lambda x:x\in E\}.

Definition 2.2.

[11] A set EdE\subset\mathbb{R}^{d} is bounded if EBE\subset B for some ball BdB\subset\mathbb{R}^{d}; EE is convex if for all x,yEx,y\in E and 0<λ<10<\lambda<1, λx+(1λ)yE\lambda x+(1-\lambda)y\in E; EE is symmetric if E=E-E=E.

Let 𝒦(d)\mathcal{K}(\mathbb{R}^{d}) be the collection of all closed, nonempty subsets of d\mathbb{R}^{d}, and the subscripts b,c,sb,c,s appended to 𝒦\mathcal{K} will denote bounded, convex and symmetric sets, respectively. For example, 𝒦bcs(d)\mathcal{K}_{bcs}(\mathbb{R}^{d}) denotes bounded, convex, symmetric and closed subsets of d\mathbb{R}^{d}.

Definition 2.3.

[11] Given a set EdE\subset\mathbb{R}^{d}, let conv(E)\mathrm{conv}(E) denote the convex hull of EE:

conv(E)={i=1kαixi:xiE,αi0,i=1kαi=1},\mathrm{conv}(E)=\left\{\sum_{i=1}^{k}\alpha_{i}x_{i}:x_{i}\in E,\alpha_{i}\geqslant 0,\sum_{i=1}^{k}\alpha_{i}=1\right\},

and denote the closure of the convex hull of EE by conv¯(E)\overline{\mathrm{conv}}(E).

Definition 2.4.

[11] A seminorm is a function p:d[0,)p:\mathbb{R}^{d}\rightarrow[0,\infty) that satisfies the following properties: for all u,vdu,v\in\mathbb{R}^{d} and α\alpha\in\mathbb{R},

p(u+v)p(u)+p(v),p(αv)=|α|p(v).p(u+v)\leqslant p(u)+p(v),\ \ p(\alpha v)=|\alpha|p(v).

For a seminorm pp and EdE\subset\mathbb{R}^{d}, define

p(E)=supvEp(v).p(E)=\sup_{v\in E}p(v).

A seminorm is a norm if p(v)=0p(v)=0 equals v=0v=0.

Definition 2.5.

[8] A seminorm function ρ\rho on Ω\Omega is a mapping ρ:Ω×d[0,)\rho:\Omega\times\mathbb{R}^{d}\rightarrow[0,\infty) such that:

(i) for all xΩx\in\Omega, ρx()=ρ(x,)\rho_{x}(\cdot)=\rho(x,\cdot) is a seminorm on d\mathbb{R}^{d};

(ii) xρx(v)x\mapsto\rho_{x}(v) is a measurable function for any vdv\in\mathbb{R}^{d}.

In what follows, to simplify writing, we abbreviate [ρx(E)]p[\rho_{x}(E)]^{p} as ρx(E)p\rho_{x}(E)^{p}. Next, definitions of measurable set-valued functions and their integrals will be introduced.

Definition 2.6.

[9] Given a function F:Ω𝒦(d)F:\Omega\rightarrow\mathcal{K}(\mathbb{R}^{d}), FF is measurable if for every open set UdU\subset\mathbb{R}^{d}, F1(U):={xΩ:F(x)Uϕ}𝒜F^{-1}(U):=\{x\in\Omega:F(x)\cap U\neq\phi\}\in\mathcal{A}.

Definition 2.7.

[9] Given a function F:Ω𝒦(d)F:\Omega\rightarrow\mathcal{K}(\mathbb{R}^{d}), denote S0(Ω,F)S^{0}(\Omega,F) to be the set of all measurable selection functions of FF, that is, all measurable functions ff such that f(x)F(x),a.e.f(x)\in F(x),a.e.

By[9, Theorem 8.1.4], for a measurable function FF, S0(Ω,F)S^{0}(\Omega,F) is not empty.

Definition 2.8.

[9] Given a measurable function F:Ω𝒦(d)F:\Omega\rightarrow\mathcal{K}(\mathbb{R}^{d}), define the set of all integrable selection functions of FF by

S1(Ω,F):=S0(Ω,F)L1(Ω,d).S^{1}(\Omega,F):=S^{0}(\Omega,F)\cap L^{1}(\Omega,\mathbb{R}^{d}).

The Aumann integral of FF is the set of integrals of integrable selection functions of FF, i.e.

ΩF𝑑μ:={Ωf𝑑μ:fS1(Ω,F)}.\int_{\Omega}Fd\mu:=\left\{\int_{\Omega}fd\mu:f\in S^{1}(\Omega,F)\right\}.

There may not exist integrable selection function of FF, but the integrably bounded property ensures the existence.

Definition 2.9.

[9] A measurable function F:Ω𝒦(d)F:\Omega\rightarrow\mathcal{K}(\mathbb{R}^{d}) is integrably bounded if there exists a non-negative function kL1(Ω,)k\in L^{1}(\Omega,\mathbb{R}) such that

F(x)B(0,k(x))fora.e.xΩ.F(x)\subset B(0,k(x))\ \ for\ a.e.\ x\in\Omega.

If Ω\Omega is a metric space, say FF is locally integrably bounded if this holds for kLloc1(Ω,)k\in L_{loc}^{1}(\Omega,\mathbb{R}).

Remark 2.1.

If fL1(Ω,d)f\in L^{1}(\Omega,\mathbb{R}^{d}) and F(x)=conv¯({f(x),f(x)})F(x)=\overline{\mathrm{conv}}(\{f(x),-f(x)\}) for xΩx\in\Omega, then FF is integrably bounded.

Similar to Lebesgue integral of real-valued functions, Aumann integral is linear and monotone.

Lemma 2.1.

[9] Suppose that F1,F2:Ω𝒦bcs(d)F_{1},F_{2}:\Omega\rightarrow\mathcal{K}_{bcs}(\mathbb{R}^{d}) are measurable and integrably bounded. Then for any λ1,λ2\lambda_{1},\lambda_{2}\in\mathbb{R},

Ω(λ1F1+λ2F2)𝑑μ=α1ΩF1𝑑μ+α2ΩF2𝑑μ.\int_{\Omega}(\lambda_{1}F_{1}+\lambda_{2}F_{2})d\mu=\alpha_{1}\int_{\Omega}F_{1}d\mu+\alpha_{2}\int_{\Omega}F_{2}d\mu.

If F1(x)F2(x)F_{1}(x)\subset F_{2}(x) for any xΩx\in\Omega, then

ΩF1𝑑μΩF2𝑑μ.\int_{\Omega}F_{1}d\mu\subset\int_{\Omega}F_{2}d\mu.

The following lemma can be seen in the proof of[8, Proposition 5.2].

Lemma 2.2.

Suppose that F:Ω𝒦bcs(d)F:\Omega\rightarrow\mathcal{K}_{bcs}(\mathbb{R}^{d}) is measurable and integrably bounded, EΩE\subset\Omega, then

|EF𝑑μ|E|F|𝑑μ.\left|\int_{E}Fd\mu\right|\leqslant\int_{E}|F|d\mu.

This chapter ends with definitions of convex-set valued Lebesgue spaces L𝒦p(Ω,ρ){L_{\mathcal{K}}^{p}(\Omega,\rho)}, convex-set valued weak Lebesgue spaces L𝒦p,(Ω,ρ){L_{\mathcal{K}}^{p,\infty}(\Omega,\rho)} and distribution functions.

Definition 2.10.

[8] Suppose that ρ\rho is a seminorm function on Ω\Omega. For any p(0,)p\in(0,\infty), define L𝒦p(Ω,ρ)L_{\mathcal{K}}^{p}(\Omega,\rho) to be the set of all measurable functions F:Ω𝒦bcs(d)F:\Omega\rightarrow\mathcal{K}_{bcs}(\mathbb{R}^{d}) such that

FL𝒦p(Ω,ρ):=(Ωρx(F(x))p𝑑μ(x))1p<,\lVert F\rVert_{L_{\mathcal{K}}^{p}(\Omega,\rho)}:=\left(\int_{\Omega}\rho_{x}(F(x))^{p}d\mu(x)\right)^{\frac{1}{p}}<\infty,

and L𝒦p,(Ω,ρ)L_{\mathcal{K}}^{p,\infty}(\Omega,\rho) to be the set of all such F:Ω𝒦bcs(d)F:\Omega\rightarrow\mathcal{K}_{bcs}(\mathbb{R}^{d}) that satisfy

FL𝒦p,(Ω,ρ):=supλ>0λμ({xΩ:ρx(F(x))>λ})1p<.\lVert F\rVert_{L_{\mathcal{K}}^{p,\infty}(\Omega,\rho)}:=\sup_{\lambda>0}\lambda\mu(\{x\in\Omega:\rho_{x}(F(x))>\lambda\})^{\frac{1}{p}}<\infty.

When p=p=\infty, define L𝒦(Ω,ρ)L_{\mathcal{K}}^{\infty}(\Omega,\rho) to be the set of all such FF that satisfy

FL𝒦(Ω,ρ):=esssupxΩρx(F(x))<.\lVert F\rVert_{L_{\mathcal{K}}^{\infty}(\Omega,\rho)}:=\mathop{\mathrm{ess}\sup}\limits_{x\in\Omega}\rho_{x}(F(x))<\infty.
Remark 2.2.

If ρx(v)=|v|\rho_{x}(v)=|v| for any (x,v)Ω×d(x,v)\in\Omega\times\mathbb{R}^{d}, the space L𝒦p(Ω,ρ)L_{\mathcal{K}}^{p}(\Omega,\rho) will be L𝒦p(Ω,||)L_{\mathcal{K}}^{p}(\Omega,|\cdot|), which is the generalization of Lp(Ω,d)L^{p}(\Omega,\mathbb{R}^{d}). In fact, given 0<p0<p\leqslant\infty and fLp(Ω,d)f\in L^{p}(\Omega,\mathbb{R}^{d}), set F(x)=conv¯({f(x),f(x)})F(x)=\overline{\mathrm{conv}}(\{f(x),-f(x)\}) for xΩx\in\Omega, then FL𝒦p(Ω,||)F\in L_{\mathcal{K}}^{p}(\Omega,|\cdot|) and FL𝒦p(Ω,||)=fLp(Ω,d)\lVert F\rVert_{L_{\mathcal{K}}^{p}(\Omega,|\cdot|)}=\lVert f\rVert_{L^{p}(\Omega,\mathbb{R}^{d})}.

Definition 2.11.

Suppose that F:Ω𝒦bcs(d)F:\Omega\rightarrow\mathcal{K}_{bcs}(\mathbb{R}^{d}) is measurable, ρ\rho is a seminorm function on Ω\Omega. For any p(0,)p\in(0,\infty), define the distribution function of FF as

ωF(λ)=μ({xΩ:ρx(F(x))>λ})forλ>0.\omega_{F}(\lambda)=\mu(\{x\in\Omega:\rho_{x}(F(x))>\lambda\})\ \ for\ \lambda>0.

Since ρx(F(x))\rho_{x}(F(x)) is a real-valued function in Lp(Ω)L^{p}(\Omega), it’s obvious that the following integral formula holds, we omit the proof.

Lemma 2.3.

For any p(0,)p\in(0,\infty) and FL𝒦p(Ω,ρ)F\in L_{\mathcal{K}}^{p}(\Omega,\rho),

FL𝒦p(Ω,ρ)p=p0λp1ωF(λ)𝑑λ.\lVert F\rVert_{L_{\mathcal{K}}^{p}(\Omega,\rho)}^{p}=p\int_{0}^{\infty}\lambda^{p-1}\omega_{F}(\lambda)d\lambda.

3 Marcinkiewicz interpolation theorem and its application

We first prove Marcinkiewicz interpolation theorem.

Theorem 3.1.

Let 1p0q01\leqslant p_{0}\leqslant q_{0}\leqslant\infty, 1p1q11\leqslant p_{1}\leqslant q_{1}\leqslant\infty, q0q1q_{0}\neq q_{1}, ρ\rho be a seminorm function on Ω\Omega, TT be a sublinear operator defined on the space L𝒦p0(Ω,ρ)+L𝒦p1(Ω,ρ)L^{p_{0}}_{\mathcal{K}}(\Omega,\rho)+L^{p_{1}}_{\mathcal{K}}(\Omega,\rho): for all F,GL𝒦p0(Ω,ρ)+L𝒦p1(Ω,ρ)F,G\in L^{p_{0}}_{\mathcal{K}}(\Omega,\rho)+L^{p_{1}}_{\mathcal{K}}(\Omega,\rho), λ\lambda\in\mathbb{R} and xΩx\in\Omega,

T(F+G)(x)TF(x)+TG(x),T(λF)(x)=λTF(x).T(F+G)(x)\subset TF(x)+TG(x),\ \ T(\lambda F)(x)=\lambda TF(x).

Suppose that there exist two positive constants C0C_{0} and C1C_{1} such that

T(F)L𝒦q0,(Ω,ρ)C0fL𝒦p0(Ω,ρ)forallFL𝒦p0(Ω,ρ),\lVert T(F)\rVert_{L^{q_{0},\infty}_{\mathcal{K}}(\Omega,\rho)}\leqslant C_{0}\lVert f\rVert_{L^{p_{0}}_{\mathcal{K}}(\Omega,\rho)}\ \ for\ all\ F\in L^{p_{0}}_{\mathcal{K}}(\Omega,\rho),
T(F)L𝒦q1,(Ω,ρ)C1fL𝒦p1(Ω,ρ)forallFL𝒦p1(Ω,ρ),\lVert T(F)\rVert_{L^{q_{1},\infty}_{\mathcal{K}}(\Omega,\rho)}\leqslant C_{1}\lVert f\rVert_{L^{p_{1}}_{\mathcal{K}}(\Omega,\rho)}\ \ for\ all\ F\in L^{p_{1}}_{\mathcal{K}}(\Omega,\rho),

then for all 0<t<10<t<1 and FL𝒦p(Ω,ρ)F\in L^{p}_{\mathcal{K}}(\Omega,\rho), we have

T(F)L𝒦q(Ω,ρ)CtFL𝒦p(Ω,ρ),\lVert T(F)\rVert_{L^{q}_{\mathcal{K}}(\Omega,\rho)}\leqslant C_{t}\lVert F\rVert_{L^{p}_{\mathcal{K}}(\Omega,\rho)}, (3)

where

1p=1tp0+tp1,1q=1tq0+tq1,andCt=2(q|qq0|+q|qq1|)1qC01tC1t.\frac{1}{p}=\frac{1-t}{p_{0}}+\frac{t}{p_{1}},\ \ \frac{1}{q}=\frac{1-t}{q_{0}}+\frac{t}{q_{1}},\ \ and\ \ C_{t}=2\left(\frac{q}{|q-q_{0}|}+\frac{q}{|q-q_{1}|}\right)^{\frac{1}{q}}C_{0}^{1-t}C_{1}^{t}.
Proof.

Without the loss of generality, assume that p1p0p_{1}\leqslant p_{0}, and consider the following cases separately:

(a1)p1<p0,q1<q0<;(a2)p1<p0,q0<q1<;(a_{1})\ p_{1}<p_{0},\ q_{1}<q_{0}<\infty;\ \ (a_{2})\ p_{1}<p_{0},\ q_{0}<q_{1}<\infty;
(b)p1=p0,q1<q0<;(c1)p1<p0,q1<q0=;(b)\ p_{1}=p_{0},\ q_{1}<q_{0}<\infty;\ \ (c_{1})\ p_{1}<p_{0},\ q_{1}<q_{0}=\infty;
(c2)p1<p0,q0<q1=;(d)p1=p0,q1<q0=.(c_{2})\ p_{1}<p_{0},\ q_{0}<q_{1}=\infty;\ \ (d)\ p_{1}=p_{0},\ q_{1}<q_{0}=\infty.

Case (𝒂𝟏).\boldsymbol{(a_{1}).} For a fixed z>0z>0 and any FL𝒦p(Ω,ρ)F\in L^{p}_{\mathcal{K}}(\Omega,\rho), split F=F0+F1F=F_{0}+F_{1}, where

F0(x)={F(x),ρ(F(x))z,{0},ρ(F(x))>z,F1(x)={{0},ρ(F(x))z,F(x),ρ(F(x))>z.F_{0}(x)=\left\{\begin{array}[]{rcl}F(x),\ \ \ \ \rho(F(x))\leqslant z,\\ \{0\},\ \ \ \ \rho(F(x))>z,\end{array}\right.\ \ F_{1}(x)=\left\{\begin{array}[]{rcl}\{0\},\ \ \ \ \rho(F(x))\leqslant z,\\ F(x),\ \ \ \ \rho(F(x))>z.\end{array}\right.

By the definition of F0F_{0} and F1F_{1}, we have

Ωρ(F0(x))p0𝑑μ(x)=Ωρ(F0(x))p0pρ(F0(x))p𝑑μ(x)zp0pΩρ(F(x))p𝑑μ(x),\int_{\Omega}\rho(F_{0}(x))^{p_{0}}d\mu(x)=\int_{\Omega}\rho(F_{0}(x))^{p_{0}-p}\rho(F_{0}(x))^{p}d\mu(x)\leqslant z^{p_{0}-p}\int_{\Omega}\rho(F(x))^{p}d\mu(x),
Ωρ(F1(x))p1𝑑μ(x)=Ωρ(F1(x))p1pρ(F1(x))p𝑑μ(x)zp1pΩρ(F(x))p𝑑μ(x),\int_{\Omega}\rho(F_{1}(x))^{p_{1}}d\mu(x)=\int_{\Omega}\rho(F_{1}(x))^{p_{1}-p}\rho(F_{1}(x))^{p}d\mu(x)\leqslant z^{p_{1}-p}\int_{\Omega}\rho(F(x))^{p}d\mu(x),

thus F0L𝒦p0(Ω,ρ)F_{0}\in L^{p_{0}}_{\mathcal{K}}(\Omega,\rho) and F1L𝒦p1(Ω,ρ)F_{1}\in L^{p_{1}}_{\mathcal{K}}(\Omega,\rho).
By the sublinearity property of TT and ρ\rho, there holds

ρ(T(F)(x))ρ(T(F0)(x))+ρ(T(F1)(x)),\rho(T(F)(x))\leqslant\rho(T(F_{0})(x))+\rho(T(F_{1})(x)),

which implies that, for any λ>0\lambda>0,

{x:ρ(T(F)(x))>2λ}{x:ρ(T(F0)(x))>λ}{x:ρ(T(F1)(x))>λ},\left\{x:\rho(T(F)(x))>2\lambda\right\}\subset\left\{x:\rho(T(F_{0})(x))>\lambda\right\}\cup\left\{x:\rho(T(F_{1})(x))>\lambda\right\},

therefore

ωT(F)(2λ)ωT(F0)(λ)+ωT(F1)(λ).\omega_{T(F)}(2\lambda)\leqslant\omega_{T(F_{0})}(\lambda)+\omega_{T(F_{1})}(\lambda).

The weak boundedness of TT gives

ωT(F)(2λ)\displaystyle\omega_{T(F)}(2\lambda) C0q0λq0({xΩ:ρ(F(x))z}ρ(F(x))p0𝑑μ(x))k0+C1q1λq1({xΩ:ρ(F(x))>z}ρ(F(x))p1𝑑μ(x))k1,\displaystyle\leqslant\frac{C_{0}^{q_{0}}}{\lambda^{q_{0}}}\left(\int_{\{x\in\Omega:\rho(F(x))\leqslant z\}}\rho(F(x))^{p_{0}}d\mu(x)\right)^{k_{0}}+\frac{C_{1}^{q_{1}}}{\lambda^{q_{1}}}\left(\int_{\{x\in\Omega:\rho(F(x))>z\}}\rho(F(x))^{p_{1}}d\mu(x)\right)^{k_{1}},

where k0=q0/p0k_{0}=q_{0}/p_{0}, k1=q1/p1k_{1}=q_{1}/p_{1} are not less than 11. Then, by Lemma 2.3, we obtain that

T(f)L𝒦q(Ω,ρ)q\displaystyle\lVert T(f)\rVert_{L_{\mathcal{K}}^{q}(\Omega,\rho)}^{q} =q0λq1ωT(F)(λ)𝑑λ=2qq0λq1ωT(F)(2λ)𝑑λ\displaystyle=q\int_{0}^{\infty}\lambda^{q-1}\omega_{T(F)}(\lambda)d\lambda=2^{q}q\int_{0}^{\infty}\lambda^{q-1}\omega_{T(F)}(2\lambda)d\lambda
2qqC0q00λqq01(ρ(F(x))zρ(F(x))p0𝑑μ(x))k0𝑑λ\displaystyle\leqslant 2^{q}qC_{0}^{q_{0}}\int_{0}^{\infty}\lambda^{q-q_{0}-1}\left(\int_{\rho(F(x))\leqslant z}\rho(F(x))^{p_{0}}d\mu(x)\right)^{k_{0}}d\lambda
+2qqC1q10λqq11(ρ(F(x))>zρ(F(x))p1𝑑μ(x))k1𝑑λ\displaystyle\ \ \ \ +2^{q}qC_{1}^{q_{1}}\int_{0}^{\infty}\lambda^{q-q_{1}-1}\left(\int_{\rho(F(x))>z}\rho(F(x))^{p_{1}}d\mu(x)\right)^{k_{1}}d\lambda
=:2qqC0q0I0+2qqC1q1I1.\displaystyle=:2^{q}qC_{0}^{q_{0}}I_{0}+2^{q}qC_{1}^{q_{1}}I_{1}.

By the dual property of weighted Lebesgue spaces, we have

I01k0=supg00λqq01(ρ(F(x))zρ(F(x))p0𝑑μ(x))g0(λ)𝑑λ,I_{0}^{\frac{1}{k_{0}}}=\sup_{g_{0}}\int_{0}^{\infty}\lambda^{q-q_{0}-1}\left(\int_{\rho(F(x))\leqslant z}\rho(F(x))^{p_{0}}d\mu(x)\right)g_{0}(\lambda)d\lambda,
I11k1=supg10λqq11(ρ(F(x))>zρ(F(x))p1𝑑μ(x))g1(λ)𝑑λ,I_{1}^{\frac{1}{k_{1}}}=\sup_{g_{1}}\int_{0}^{\infty}\lambda^{q-q_{1}-1}\left(\int_{\rho(F(x))>z}\rho(F(x))^{p_{1}}d\mu(x)\right)g_{1}(\lambda)d\lambda,

where g0g_{0} and g1g_{1} are real-valued functions satisfying

0λqq01|g0(λ)|k0𝑑λ1,0λqq11|g1(λ)|k1𝑑λ1.\int_{0}^{\infty}\lambda^{q-q_{0}-1}|g_{0}(\lambda)|^{k^{\prime}_{0}}d\lambda\leqslant 1,\ \ \int_{0}^{\infty}\lambda^{q-q_{1}-1}|g_{1}(\lambda)|^{k^{\prime}_{1}}d\lambda\leqslant 1.

Set z=(λ/A)ξz=(\lambda/A)^{\xi}, where AA and ξ\xi will be determined later. By Fubini’s theorem, the integral under the first ’sup’ equals

Ωρ(F(x))p0(Aρ(F(x))1ξλqq01g0(λ)𝑑λ)𝑑μ(x)\displaystyle\int_{\Omega}\rho(F(x))^{p_{0}}\left(\int_{A\rho(F(x))^{\frac{1}{\xi}}}^{\infty}\lambda^{q-q_{0}-1}g_{0}(\lambda)d\lambda\right)d\mu(x)
Ωρ(F(x))p0(Aρ(F(x))1ξλqq01𝑑λ)1k0(Aρ(F(x))1ξλqq01g0(λ)1k0𝑑λ)1k0𝑑μ(x)\displaystyle\ \ \ \ \leqslant\int_{\Omega}\rho(F(x))^{p_{0}}\left(\int_{A\rho(F(x))^{\frac{1}{\xi}}}^{\infty}\lambda^{q-q_{0}-1}d\lambda\right)^{\frac{1}{k_{0}}}\left(\int_{A\rho(F(x))^{\frac{1}{\xi}}}^{\infty}\lambda^{q-q_{0}-1}g_{0}(\lambda)^{\frac{1}{k^{\prime}_{0}}}d\lambda\right)^{\frac{1}{k^{\prime}_{0}}}d\mu(x)
1(q0q)1k0Aqq0k0Ωρ(F(x))p0+qq0k0ξ𝑑μ(x),\displaystyle\ \ \ \ \leqslant\frac{1}{(q_{0}-q)^{\frac{1}{k_{0}}}}A^{\frac{q-q_{0}}{k_{0}}}\int_{\Omega}\rho(F(x))^{p_{0}+\frac{q-q_{0}}{k_{0}\xi}}d\mu(x),

and the integral under the second ’sup’ equals

Ωρ(F(x))p1(0Aρ(F(x))1ξλqq11g1(λ)𝑑λ)𝑑μ(x)\displaystyle\int_{\Omega}\rho(F(x))^{p_{1}}\left(\int_{0}^{A\rho(F(x))^{\frac{1}{\xi}}}\lambda^{q-q_{1}-1}g_{1}(\lambda)d\lambda\right)d\mu(x)
Ωρ(F(x))p1(0Aρ(F(x))1ξλqq11𝑑λ)1k1(0Aρ(F(x))1ξλqq11g1(λ)1k1𝑑λ)1k1𝑑μ(x)\displaystyle\ \ \ \ \leqslant\int_{\Omega}\rho(F(x))^{p_{1}}\left(\int_{0}^{A\rho(F(x))^{\frac{1}{\xi}}}\lambda^{q-q_{1}-1}d\lambda\right)^{\frac{1}{k_{1}}}\left(\int_{0}^{A\rho(F(x))^{\frac{1}{\xi}}}\lambda^{q-q_{1}-1}g_{1}(\lambda)^{\frac{1}{k^{\prime}_{1}}}d\lambda\right)^{\frac{1}{k^{\prime}_{1}}}d\mu(x)
1(qq1)1k1Aqq1k1Ωρ(F(x))p1+qq1k1ξ𝑑μ(x).\displaystyle\ \ \ \ \leqslant\frac{1}{(q-q_{1})^{\frac{1}{k_{1}}}}A^{\frac{q-q_{1}}{k_{1}}}\int_{\Omega}\rho(F(x))^{p_{1}+\frac{q-q_{1}}{k_{1}\xi}}d\mu(x).

Note that

qq0k0(pp0)=1p(1q01q)1q(1p01p)=tp(1q01q1)tq(1p01p1),qq1k1(pp1)=1p(1q11q)1q(1p11p)=1tp(1q11q0)1tq(1p11p0),\frac{q-q_{0}}{k_{0}(p-p_{0})}=\frac{\frac{1}{p}\left(\frac{1}{q_{0}}-\frac{1}{q}\right)}{\frac{1}{q}\left(\frac{1}{p_{0}}-\frac{1}{p}\right)}=\frac{\frac{t}{p}\left(\frac{1}{q_{0}}-\frac{1}{q_{1}}\right)}{\frac{t}{q}\left(\frac{1}{p_{0}}-\frac{1}{p_{1}}\right)},\ \ \frac{q-q_{1}}{k_{1}(p-p_{1})}=\frac{\frac{1}{p}\left(\frac{1}{q_{1}}-\frac{1}{q}\right)}{\frac{1}{q}\left(\frac{1}{p_{1}}-\frac{1}{p}\right)}=\frac{\frac{1-t}{p}\left(\frac{1}{q_{1}}-\frac{1}{q_{0}}\right)}{\frac{1-t}{q}\left(\frac{1}{p_{1}}-\frac{1}{p_{0}}\right)},

we select

ξ=qq0k0(pp0)=qq1k1(pp1),\xi=\frac{q-q_{0}}{k_{0}(p-p_{0})}=\frac{q-q_{1}}{k_{1}(p-p_{1})},

which makes

p0+qq0k0ξ=p1+qq1k1ξ=p,p_{0}+\frac{q-q_{0}}{k_{0}\xi}=p_{1}+\frac{q-q_{1}}{k_{1}\xi}=p,

therefore,

T(F)L𝒦q(Ω,ρ)q2qqC0q0Aqq0q0qFL𝒦p(Ω,ρ)pk0+2qqC1q1Aqq1qq1FL𝒦p(Ω,ρ)pk1.\lVert T(F)\rVert_{L_{\mathcal{K}}^{q}(\Omega,\rho)}^{q}\leqslant 2^{q}qC_{0}^{q_{0}}\frac{A^{q-q_{0}}}{q_{0}-q}\lVert F\rVert_{L_{\mathcal{K}}^{p}(\Omega,\rho)}^{pk_{0}}+2^{q}qC_{1}^{q_{1}}\frac{A^{q-q_{1}}}{q-q_{1}}\lVert F\rVert_{L_{\mathcal{K}}^{p}(\Omega,\rho)}^{pk_{1}}.

Next, select

A=C0q0q0q1C1q1q1q0FL𝒦p(Ω,ρ)p(k0k1)q0q1,A=C_{0}^{\frac{q_{0}}{q_{0}-q_{1}}}C_{1}^{\frac{q_{1}}{q_{1}-q_{0}}}\lVert F\rVert_{L_{\mathcal{K}}^{p}(\Omega,\rho)}^{\frac{p(k_{0}-k_{1})}{q_{0}-q_{1}}},

therefore,

T(F)L𝒦q(Ω,ρ)q2qC0q(1t)C1qt(qq0q+qqq1)FL𝒦p(Ω,ρ)q,\lVert T(F)\rVert_{L_{\mathcal{K}}^{q}(\Omega,\rho)}^{q}\leqslant 2^{q}C_{0}^{q(1-t)}C_{1}^{qt}\left(\frac{q}{q_{0}-q}+\frac{q}{q-q_{1}}\right)\lVert F\rVert_{L_{\mathcal{K}}^{p}(\Omega,\rho)}^{q},

thus we obtain (3).

Case (𝒂𝟐).\boldsymbol{(a_{2}).} The proof of this case is similar to that in case (a1)(a_{1}). We change the integration regions (Aρ(F(x))1/ξ,)(A\rho(F(x))^{1/\xi},\infty), (0,Aρ(F(x))1/ξ)(0,A\rho(F(x))^{1/\xi}) to (0,Aρ(F(x))1/ξ)(0,A\rho(F(x))^{1/\xi}), (Aρ(F(x))1/ξ,)(A\rho(F(x))^{1/\xi},\infty) respectively in the corresponding integers, and change the denominator (q0q)1/k0(q_{0}-q)^{1/k_{0}}, (qq1)1/k1(q-q_{1})^{1/k_{1}} to (qq0)1/k0(q-q_{0})^{1/k_{0}}, (q1q)1/k1(q_{1}-q)^{1/k_{1}}. Then,

T(F)L𝒦q(Ω,ρ)q2qC0q(1t)C1qt(qqq0+qq1q)FL𝒦p(Ω,ρ)q,\lVert T(F)\rVert_{L_{\mathcal{K}}^{q}(\Omega,\rho)}^{q}\leqslant 2^{q}C_{0}^{q(1-t)}C_{1}^{qt}\left(\frac{q}{q-q_{0}}+\frac{q}{q_{1}-q}\right)\lVert F\rVert_{L_{\mathcal{K}}^{p}(\Omega,\rho)}^{q},

which also implies (3).

Case (𝒃).\boldsymbol{(b).} Note that p=p1=p0p=p_{1}=p_{0}, by the weak boundedness of TT, there hold

ωT(F)(λ)(C0FL𝒦p(Ω,ρ)λ)q0,ωT(F)(λ)(C1FL𝒦p(Ω,ρ)λ)q1.\omega_{T(F)}(\lambda)\leqslant\left(\frac{C_{0}\lVert F\rVert_{L_{\mathcal{K}}^{p}(\Omega,\rho)}}{\lambda}\right)^{q_{0}},\ \ \omega_{T(F)}(\lambda)\leqslant\left(\frac{C_{1}\lVert F\rVert_{L_{\mathcal{K}}^{p}(\Omega,\rho)}}{\lambda}\right)^{q_{1}}.

Therefore, for a certain AA which will be determined later, we have

T(f)L𝒦q(Ω,ρ)q\displaystyle\lVert T(f)\rVert_{L_{\mathcal{K}}^{q}(\Omega,\rho)}^{q} =q0λq1ωT(F)(λ)𝑑λ\displaystyle=q\int_{0}^{\infty}\lambda^{q-1}\omega_{T(F)}(\lambda)d\lambda
=q(Aλq1ωT(F)(λ)𝑑λ+0Aλq1ωT(F)(λ)𝑑λ)\displaystyle=q\left(\int_{A}^{\infty}\lambda^{q-1}\omega_{T(F)}(\lambda)d\lambda+\int_{0}^{A}\lambda^{q-1}\omega_{T(F)}(\lambda)d\lambda\right)
qC0q0FL𝒦p(Ω,ρ)q0Aλqq01𝑑λ+qC1q1FL𝒦p(Ω,ρ)q10Aλqq11𝑑λ\displaystyle\leqslant qC_{0}^{q_{0}}\lVert F\rVert_{L_{\mathcal{K}}^{p}(\Omega,\rho)}^{q_{0}}\int_{A}^{\infty}\lambda^{q-q_{0}-1}d\lambda+qC_{1}^{q_{1}}\lVert F\rVert_{L_{\mathcal{K}}^{p}(\Omega,\rho)}^{q_{1}}\int_{0}^{A}\lambda^{q-q_{1}-1}d\lambda
=qC0q0Aqq0q0qFL𝒦p(Ω,ρ)q0+qC1q1Aqq1qq1FL𝒦p(Ω,ρ)q1.\displaystyle=qC_{0}^{q_{0}}\frac{A^{q-q_{0}}}{q_{0}-q}\lVert F\rVert_{L_{\mathcal{K}}^{p}(\Omega,\rho)}^{q_{0}}+qC_{1}^{q_{1}}\frac{A^{q-q_{1}}}{q-q_{1}}\lVert F\rVert_{L_{\mathcal{K}}^{p}(\Omega,\rho)}^{q_{1}}.

Select

A=C0q0q0q1C1q1q1q0FL𝒦p(Ω,ρ),A=C_{0}^{\frac{q_{0}}{q_{0}-q_{1}}}C_{1}^{\frac{q_{1}}{q_{1}-q_{0}}}\lVert F\rVert_{L_{\mathcal{K}}^{p}(\Omega,\rho)},

therefore,

T(f)L𝒦q(Ω,ρ)qC0q(1t)C1qt(qq0q+qqq1)FL𝒦p(Ω,ρ)q,\lVert T(f)\rVert_{L_{\mathcal{K}}^{q}(\Omega,\rho)}^{q}\leqslant C_{0}^{q(1-t)}C_{1}^{qt}\left(\frac{q}{q_{0}-q}+\frac{q}{q-q_{1}}\right)\lVert F\rVert_{L_{\mathcal{K}}^{p}(\Omega,\rho)}^{q},

which implies (3) with a better constant.

Case (𝒄𝟏).\boldsymbol{(c_{1}).} If p0=p_{0}=\infty, take z=δ/C0z=\delta/C_{0} in the split of FF in case (a1)(a_{1}), then

esssupρ(T(F0))C0esssupρ(F0)δ,\mathrm{ess}\sup\rho(T(F_{0}))\leqslant C_{0}\mathrm{ess}\sup\rho(F_{0})\leqslant\delta,

which follows that ωT(F0)(δ)=0\omega_{T(F_{0})}(\delta)=0. Note that A=C0A=C_{0}, according to the similar calculation as which in case (a1)(a_{1}), we have

T(F)L𝒦q(Ω,ρ)q2qC0q(1t)C1qtqqq1FL𝒦p(Ω,ρ)q.\lVert T(F)\rVert_{L_{\mathcal{K}}^{q}(\Omega,\rho)}^{q}\leqslant 2^{q}C_{0}^{q(1-t)}C_{1}^{qt}\frac{q}{q-q_{1}}\lVert F\rVert_{L_{\mathcal{K}}^{p}(\Omega,\rho)}^{q}.

If p0<p_{0}<\infty, we still select zz makes ωT(F0)(δ)=0\omega_{T(F_{0})}(\delta)=0, after which the proof proceeds as before. Suppose that

z=(δA)ξ,whereξ=p0p0p,A=ηC0FL𝒦p(Ω,ρ)pp0,z=\left(\frac{\delta}{A}\right)^{\xi},\ \ \textrm{where}\ \ \xi=\frac{p_{0}}{p_{0}-p},\ \ A=\eta C_{0}\lVert F\rVert_{L_{\mathcal{K}}^{p}(\Omega,\rho)}^{\frac{p}{p_{0}}},

and η\eta is a positive number which will be determined later. It is enough to show that for a suitable η\eta, we have esssupρ(T(F0))δ\mathrm{ess}\sup\rho(T(F_{0}))\leqslant\delta. By the weak boundedness of TT,

esssupρ(T(F0))C0FL𝒦p0(Ω,ρ)=C0(p00zλp01ωF(λ)𝑑λ)1p0.\mathrm{ess}\sup\rho(T(F_{0}))\leqslant C_{0}\lVert F\rVert_{L_{\mathcal{K}}^{p_{0}}(\Omega,\rho)}=C_{0}\left(p_{0}\int_{0}^{z}\lambda^{p_{0}-1}\omega_{F}(\lambda)d\lambda\right)^{\frac{1}{p_{0}}}.

It follows that we certainly have esssupρ(T(F0))δ\mathrm{ess}\sup\rho(T(F_{0}))\leqslant\delta if

C0p0p00zλp01ωF(λ)𝑑λ(Az1ξ)p0=Ap0zp0p,C_{0}^{p_{0}}p_{0}\int_{0}^{z}\lambda^{p_{0}-1}\omega_{F}(\lambda)d\lambda\leqslant(Az^{\frac{1}{\xi}})^{p_{0}}=A^{p_{0}}z^{p_{0}-p},

which can be implied by

C0p0p0zp0p0λp1ωF(λ)𝑑ληp0C0p0FL𝒦p(Ω,ρ)pzp0p.C_{0}^{p_{0}}p_{0}z^{p_{0}-p}\int_{0}^{\infty}\lambda^{p-1}\omega_{F}(\lambda)d\lambda\leqslant\eta^{p_{0}}C_{0}^{p_{0}}\lVert F\rVert_{L_{\mathcal{K}}^{p}(\Omega,\rho)}^{p}z^{p_{0}-p}.

Since the integral on the left is p1FL𝒦p(Ω,ρ)pp^{-1}\lVert F\rVert_{L_{\mathcal{K}}^{p}(\Omega,\rho)}^{p}, the inequality is satisfied if ηp1p1/p\eta^{p_{1}}\geqslant p_{1}/p, which completes the proof of this case.

Case (𝒄𝟐).\boldsymbol{(c_{2}).} In this case, we need only change the symbol of some exponents in case (c1)(c_{1}) as which in case (a2)(a_{2}). We omit the details here.

Case (𝒅).\boldsymbol{(d).} By the weak boundedness of TT, there hold

esssupρ(T(F))C0FL𝒦p(Ω,ρ),ωT(F)(λ)(C1FL𝒦p(Ω,ρ)λ)q1.\mathrm{ess}\sup\rho(T(F))\leqslant C_{0}\lVert F\rVert_{L_{\mathcal{K}}^{p}(\Omega,\rho)},\ \ \omega_{T(F)}(\lambda)\leqslant\left(\frac{C_{1}\lVert F\rVert_{L_{\mathcal{K}}^{p}(\Omega,\rho)}}{\lambda}\right)^{q_{1}}.

Therefore, for A=C0FL𝒦p(Ω,ρ)A=C_{0}\lVert F\rVert_{L_{\mathcal{K}}^{p}(\Omega,\rho)}, we have

T(f)L𝒦q(Ω,ρ)q\displaystyle\lVert T(f)\rVert_{L_{\mathcal{K}}^{q}(\Omega,\rho)}^{q} =q0λq1ωT(F)(λ)𝑑λ=q0Aλq1ωT(F)(λ)𝑑λ\displaystyle=q\int_{0}^{\infty}\lambda^{q-1}\omega_{T(F)}(\lambda)d\lambda=q\int_{0}^{A}\lambda^{q-1}\omega_{T(F)}(\lambda)d\lambda
qC1q1FL𝒦p(Ω,ρ)q10Aλqq11𝑑λ=qC1q1Aqq1qq1FL𝒦p(Ω,ρ)q1\displaystyle\leqslant qC_{1}^{q_{1}}\lVert F\rVert_{L_{\mathcal{K}}^{p}(\Omega,\rho)}^{q_{1}}\int_{0}^{A}\lambda^{q-q_{1}-1}d\lambda=qC_{1}^{q_{1}}\frac{A^{q-q_{1}}}{q-q_{1}}\lVert F\rVert_{L_{\mathcal{K}}^{p}(\Omega,\rho)}^{q_{1}}
=C0q(1t)C1qtqqq1FL𝒦p(Ω,ρ)q,\displaystyle=C_{0}^{q(1-t)}C_{1}^{qt}\frac{q}{q-q_{1}}\lVert F\rVert_{L_{\mathcal{K}}^{p}(\Omega,\rho)}^{q},

which implies (3) with a better constant. ∎

The following discussion shows that Theorem 3.1 can deduce Marcinkiewicz interpolation theorem on Lebesgue spaces with an additional condition of operator. Through the same method, we can obtain Marcinkiewicz interpolation theorem on vector-valued Lebesgue spaces by Theorem 3.1.

Remark 3.1.

Under the assumptions of p0,p1,q0,q1p_{0},p_{1},q_{0},q_{1} in Theorem 3.1, suppose that T0T_{0} is a sublinear operator bounded from Lp0(Ω)L^{p_{0}}(\Omega) to Lq0(Ω)L^{q_{0}}(\Omega), bounded from Lp1(Ω)L^{p_{1}}(\Omega) to Lq1(Ω)L^{q_{1}}(\Omega), and satisfying |T(f)(x)||T(|f|)(x)||T(f)(x)|\leqslant|T(|f|)(x)|. Set d=1d=1, ρ=||\rho=|\cdot|, and define an associated operator on L𝒦p0(Ω,||)+L𝒦p1(Ω,||)L^{p_{0}}_{\mathcal{K}}(\Omega,|\cdot|)+L^{p_{1}}_{\mathcal{K}}(\Omega,|\cdot|) as

T(F)(x)=[|T0(|F|)(x)|,|T0(|F|)(x)|],forxΩ.T(F)(x)=[-|T_{0}(|F|)(x)|,|T_{0}(|F|)(x)|],\ \ for\ x\in\Omega.

It’s easy to obtain that TT is a sublinear operator bounded from L𝒦p0(Ω,||)L^{p_{0}}_{\mathcal{K}}(\Omega,|\cdot|) to L𝒦q0(Ω,||)L^{q_{0}}_{\mathcal{K}}(\Omega,|\cdot|), and bounded from L𝒦p1(Ω,||)L^{p_{1}}_{\mathcal{K}}(\Omega,|\cdot|) to L𝒦q1(Ω,||)L^{q_{1}}_{\mathcal{K}}(\Omega,|\cdot|). Therefore, by Theorem 3.1, TT is bounded from L𝒦p(Ω,||)L^{p}_{\mathcal{K}}(\Omega,|\cdot|) to L𝒦q(Ω,||)L^{q}_{\mathcal{K}}(\Omega,|\cdot|).
For gLp(Ω)g\in L^{p}(\Omega), define G(x)=[|g(x)|,|g(x)|]G(x)=[-|g(x)|,|g(x)|] for xΩx\in\Omega, we have

T0gqT0(|g|)q=T0(|G|)q=TGL𝒦q(Ω,||)GL𝒦p(Ω,||)=gp,\lVert T_{0}g\rVert_{q}\leqslant\lVert T_{0}(|g|)\rVert_{q}=\lVert T_{0}(|G|)\rVert_{q}=\lVert TG\rVert_{L^{q}_{\mathcal{K}}(\Omega,|\cdot|)}\lesssim\lVert G\rVert_{L^{p}_{\mathcal{K}}(\Omega,|\cdot|)}=\lVert g\rVert_{p},

which implies that T0T_{0} is bounded from Lp(Ω)L^{p}(\Omega) to Lq(Ω)L^{q}(\Omega).

Next, we define the fractional averaging operator and fractional maximal operator of locally integrably bounded functions, and use Theorem 3.1 to obtain their boundedness.

Definition 3.1.

Let F:n𝒦bcs(d)F:\mathbb{R}^{n}\rightarrow\mathcal{K}_{bcs}(\mathbb{R}^{d}) that is locally integrably bounded, for a fixed cube QQ and α(0,1)\alpha\in(0,1), define the fractional averaging operator AQ,αA_{Q,\alpha} by

AQ,αF(x)=1|Q|1αQF(y)𝑑yχQ(x).A_{Q,\alpha}F(x)=\frac{1}{|Q|^{1-\alpha}}\int_{Q}F(y)dy\cdot\chi_{Q}(x).
Lemma 3.1.

Given any cube QQ, the fractional averaging operator is linear: if F,G:n𝒦bcs(d)F,G:\mathbb{R}^{n}\rightarrow\mathcal{K}_{bcs}(\mathbb{R}^{d}) are locally integrably bounded mappings, λ\lambda\in\mathbb{R}, then for all xnx\in\mathbb{R}^{n},

AQ,α(F+G)(x)=AQ,αF(x)+AQ,αG(x),AQ,α(λF)(x)=λAQ,αF(x).A_{Q,\alpha}(F+G)(x)=A_{Q,\alpha}F(x)+A_{Q,\alpha}G(x),\ \ A_{Q,\alpha}(\lambda F)(x)=\lambda A_{Q,\alpha}F(x).
Proof.

By Lemma 2.1, we have

AQ,α(F+G)(x)\displaystyle A_{Q,\alpha}(F+G)(x) =1|Q|1αQ(F(y)+G(y))𝑑yχQ(x)\displaystyle=\frac{1}{|Q|^{1-\alpha}}\int_{Q}(F(y)+G(y))dy\cdot\chi_{Q}(x)
=1|Q|1αQF(y)𝑑yχQ(x)+1|Q|1αQG(y)𝑑yχQ(x)\displaystyle=\frac{1}{|Q|^{1-\alpha}}\int_{Q}F(y)dy\cdot\chi_{Q}(x)+\frac{1}{|Q|^{1-\alpha}}\int_{Q}G(y)dy\cdot\chi_{Q}(x)
=AQ,αF(x)+AQ,αG(x),\displaystyle=A_{Q,\alpha}F(x)+A_{Q,\alpha}G(x),

and

AQ,α(λF)(x)\displaystyle A_{Q,\alpha}(\lambda F)(x) =1|Q|1αQλF(y)𝑑yχQ(x)=λ|Q|1αQF(y)𝑑yχQ(x)=λAQ,αF(x),\displaystyle=\frac{1}{|Q|^{1-\alpha}}\int_{Q}\lambda F(y)dy\cdot\chi_{Q}(x)=\frac{\lambda}{|Q|^{1-\alpha}}\int_{Q}F(y)dy\cdot\chi_{Q}(x)=\lambda A_{Q,\alpha}F(x),

which show the linearity. ∎

Theorem 3.2.

Given a cube QQ, for 0<α<10<\alpha<1, 1p<q1\leqslant p<q\leqslant\infty such that 1/p1/q=α1/p-1/q=\alpha, the fractional averaging operator AQ,α:L𝒦p(n,||)L𝒦q(n,||)A_{Q,\alpha}:L^{p}_{\mathcal{K}}(\mathbb{R}^{n},|\cdot|)\rightarrow L^{q}_{\mathcal{K}}(\mathbb{R}^{n},|\cdot|) is bounded.

Proof.

We begin with p=1,q=1/(1α)p=1,\ q=1/(1-\alpha). By Lemma 2.2,

AQ,αFL𝒦11α(n,||)\displaystyle\lVert A_{Q,\alpha}F\rVert_{L_{\mathcal{K}}^{\frac{1}{1-\alpha}}(\mathbb{R}^{n},|\cdot|)} =(n|1mn(Q)1αQF(y)𝑑yχQ(x)|11α𝑑x)1α\displaystyle=\left(\int_{\mathbb{R}^{n}}\left|\frac{1}{m_{n}(Q)^{1-\alpha}}\int_{Q}F(y)dy\cdot\chi_{Q}(x)\right|^{\frac{1}{1-\alpha}}dx\right)^{1-\alpha}
(1mn(Q)Q(Q|F(y)|𝑑y)11α𝑑x)1αFL𝒦1(n,||).\displaystyle\leqslant\left(\frac{1}{m_{n}(Q)}\int_{Q}\left(\int_{Q}|F(y)|dy\right)^{\frac{1}{1-\alpha}}dx\right)^{1-\alpha}\leqslant\lVert F\rVert_{L_{\mathcal{K}}^{1}(\mathbb{R}^{n},|\cdot|)}.

Then, consider the case p=1/αp=1/\alpha, q=q=\infty. By Lemma 2.2 and Hölder’s inequality, for all xnx\in\mathbb{R}^{n}, we have

|AQ,αF(x)|\displaystyle|A_{Q,\alpha}F(x)| =|1mn(Q)1αQF(y)𝑑yχQ(x)|1mn(Q)1αQ|F(y)|𝑑y\displaystyle=\left|\frac{1}{m_{n}(Q)^{1-\alpha}}\int_{Q}F(y)dy\cdot\chi_{Q}(x)\right|\leqslant\frac{1}{m_{n}(Q)^{1-\alpha}}\int_{Q}|F(y)|dy
(Q|F(y)|1α𝑑y)αFL𝒦1α(n,||),\displaystyle\leqslant\left(\int_{Q}|F(y)|^{\frac{1}{\alpha}}dy\right)^{\alpha}\leqslant\lVert F\rVert_{L_{\mathcal{K}}^{\frac{1}{\alpha}}(\mathbb{R}^{n},|\cdot|)},

which implies that

AQ,αFL𝒦(n,||)FL𝒦1α(n,||).\lVert A_{Q,\alpha}F\rVert_{L_{\mathcal{K}}^{\infty}(\mathbb{R}^{n},|\cdot|)}\leqslant\lVert F\rVert_{L_{\mathcal{K}}^{\frac{1}{\alpha}}(\mathbb{R}^{n},|\cdot|)}.

Finally, by Theorem 3.1 with T=AQ,αT=A_{Q,\alpha}, Ω=n\Omega=\mathbb{R}^{n}, ρ()=||\rho(\cdot)=|\cdot|, p0=1/αp_{0}=1/\alpha, p1=1p_{1}=1, q0=q_{0}=\infty and q1=1/(1α)q_{1}=1/(1-\alpha), we finish the proof. ∎

Definition 3.2.

Given a locally integrably bounded function F:n𝒦bcs(d)F:\mathbb{R}^{n}\rightarrow\mathcal{K}_{bcs}(\mathbb{R}^{d}) and α(0,1)\alpha\in(0,1), define the fractional maximal operator MαM_{\alpha} by

MαF(x)=conv¯(QAQ,αF(x)),M_{\alpha}F(x)=\overline{\mathrm{conv}}\left(\bigcup_{Q}A_{Q,\alpha}F(x)\right),

where the union is taken over all cubes QQ whose sides are parallel to the coordinate axes.

Lemma 3.2.

The fractional maximal operator is sublinear: if F,G:n𝒦bcs(d)F,G:\mathbb{R}^{n}\rightarrow\mathcal{K}_{bcs}(\mathbb{R}^{d}) are locally integrably bounded mappings, λ\lambda\in\mathbb{R}, then for all xnx\in\mathbb{R}^{n},

Mα(F+G)(x)MαF(x)+MαG(x),Mα(λF)(x)=λMαF(x).M_{\alpha}(F+G)(x)\subset M_{\alpha}F(x)+M_{\alpha}G(x),\ \ M_{\alpha}(\lambda F)(x)=\lambda M_{\alpha}F(x).
Proof.

By Lemma 3.1 and the linearity of the closure of convex hull,

Mα(F+G)(x)\displaystyle M_{\alpha}(F+G)(x) =conv¯(QAQ,α(F(x)+G(x)))conv¯(QAQ,αF(x)+QAQ,αG(x))\displaystyle=\overline{\mathrm{conv}}\left(\bigcup_{Q}A_{Q,\alpha}(F(x)+G(x))\right)\subset\overline{\mathrm{conv}}\left(\bigcup_{Q}A_{Q,\alpha}F(x)+\bigcup_{Q}A_{Q,\alpha}G(x)\right)
=conv¯(QAQ,αF(x))+conv¯(QAQ,αG(x))=MαF(x)+MαG(x),\displaystyle=\overline{\mathrm{conv}}\left(\bigcup_{Q}A_{Q,\alpha}F(x)\right)+\overline{\mathrm{conv}}\left(\bigcup_{Q}A_{Q,\alpha}G(x)\right)=M_{\alpha}F(x)+M_{\alpha}G(x),

and

Mα(λF)(x)\displaystyle M_{\alpha}(\lambda F)(x) =conv¯(QAQ,α(λF(x)))=λconv¯(QAQ,α(F(x)))=λMα(F)(x),\displaystyle=\overline{\mathrm{conv}}\left(\bigcup_{Q}A_{Q,\alpha}(\lambda F(x))\right)=\lambda\overline{\mathrm{conv}}\left(\bigcup_{Q}A_{Q,\alpha}(F(x))\right)=\lambda M_{\alpha}(F)(x),

which show the sublinearity. ∎

To prove the boundedness of fractional maximal operators, we need the following results.

Definition 3.3.

[21] For τ{0,±1/3}n\tau\in\{0,\pm 1/3\}^{n}, define the translated dyadic grid

𝒟τ={2k([0,1)n+m+(1)kτ):k,mn},\mathcal{D}^{\tau}=\{2^{k}([0,1)^{n}+m+(-1)^{k}\tau):k\in\mathbb{Z},m\in\mathbb{Z}^{n}\},

and define the generalized dyadic fractional maximal operator

MατF(x)=conv¯(Q𝒟τAQ,αF(x)),M^{\tau}_{\alpha}F(x)=\overline{\mathrm{conv}}\left(\bigcup_{Q\in\mathcal{D}^{\tau}}A_{Q,\alpha}F(x)\right),

which generalizes the standard dyadic grid 𝒟=𝒟0,0,,0\mathcal{D}=\mathcal{D}^{0,0,\cdots,0}, and dyadic fractional maximal operator

MαdF(x)=Mα0,0,,0F(x).M^{d}_{\alpha}F(x)=M^{0,0,\cdots,0}_{\alpha}F(x).
Definition 3.4.

[8] Given a locally integrably bounded function F:n𝒦bcs(d)F:\mathbb{R}^{n}\rightarrow\mathcal{K}_{bcs}(\mathbb{R}^{d}) and α(0,1)\alpha\in(0,1), define

M^αdF(x)=Q𝒟AQ,αF(x)¯.\widehat{M}_{\alpha}^{d}F(x)=\overline{\bigcup_{Q\in\mathcal{D}}A_{Q,\alpha}F(x)}.

Similar to [8, Lemma 5.9], the following lemma can be proved, we omit the details here.

Lemma 3.3.

Given a locally integrably bounded, convex-set valued function F:n𝒦bcs(d)F:\mathbb{R}^{n}\rightarrow\mathcal{K}_{bcs}(\mathbb{R}^{d}), we have

MαF(x)Cτ{0,±1/3}nMατF(x),M_{\alpha}F(x)\subset C\sum_{\tau\in\{0,\pm 1/3\}^{n}}M^{\tau}_{\alpha}F(x),

where the constant CC depends only on dimension nn and α\alpha.

Theorem 3.3.

For 0<α<10<\alpha<1, 1<p<q1<p<q\leqslant\infty such that 1/p1/q=α1/p-1/q=\alpha, the fractional maximal operator Mα:L𝒦p(n,||)L𝒦q(n,||)M_{\alpha}:L^{p}_{\mathcal{K}}(\mathbb{R}^{n},|\cdot|)\rightarrow L^{q}_{\mathcal{K}}(\mathbb{R}^{n},|\cdot|) is bounded. For 0<α<10<\alpha<1, p=1p=1, q=1/(1α)q=1/(1-\alpha), Mα:L𝒦1(n,||)L𝒦1/(1α),(n,||)M_{\alpha}:L^{1}_{\mathcal{K}}(\mathbb{R}^{n},|\cdot|)\rightarrow L^{1/(1-\alpha),\infty}_{\mathcal{K}}(\mathbb{R}^{n},|\cdot|) is bounded.

Proof.

First, by Lemma 3.3,

|MαF(x)|Cτ{0,±1/3}n|MατF(x)|,|M_{\alpha}F(x)|\leqslant C\sum_{\tau\in\{0,\pm 1/3\}^{n}}|M_{\alpha}^{\tau}F(x)|,

and so it will suffice to prove the strong and weak type inequalities for MατM_{\alpha}^{\tau}. In fact, given that all of the dyadic grids 𝒟τ\mathcal{D}^{\tau} have the same properties as 𝒟\mathcal{D}, it will suffice to prove them for the dyadic fractional maximal operator MαdM_{\alpha}^{d}. Moreover, arguing as the authors did in the proof of[8, Lemma 5.6], it will suffice to prove our estimates for the operator M^αd\widehat{M}^{d}_{\alpha}.

We begin with p=1p=1, q=1/(1α)q=1/(1-\alpha) by adapting the Calderón-Zygmund decomposition to convex-set valued functions. Fix λ>0\lambda>0 and define

Ωλd={xn:|M^αdF(x)|>λ}.\Omega_{\lambda}^{d}=\{x\in\mathbb{R}^{n}:|\widehat{M}_{\alpha}^{d}F(x)|>\lambda\}.

If Ωλd{\Omega_{\lambda}^{d}} is empty, there is nothing to prove. Otherwise, given xΩλdx\in\Omega_{\lambda}^{d}, there must exist a cube Q𝒟Q\in\mathcal{D} such that xQx\in Q, and

1mn(Q)1α|QF(y)𝑑y|>λ.\frac{1}{m_{n}(Q)^{1-\alpha}}\left|\int_{Q}F(y)dy\right|>\lambda.

We claim that among all the dyadic cubes containing xx, there must be a largest one with this property. By Lemma 2.2,

1mn(Q)1α|QF(y)𝑑y|1mn(Q)1αfL𝒦1(n,||).\frac{1}{m_{n}(Q)^{1-\alpha}}\left|\int_{Q}F(y)dy\right|\leqslant\frac{1}{m_{n}(Q)^{1-\alpha}}\lVert f\rVert_{L^{1}_{\mathcal{K}}(\mathbb{R}^{n},|\cdot|)}.

Since the right-hand side goes to 0 as mn(Q)m_{n}(Q)\rightarrow\infty, we see that such a maximal cube must exist, denote this cube by QxQ_{x}. Since the set of dyadic cubes is countable, we can enumerate the set {Qx:xΩλd}\{Q_{x}:x\in\Omega_{\lambda}^{d}\} by {Qj}j\{Q_{j}\}_{j\in\mathbb{N}}. The cubes {Qj}\{Q_{j}\} must be disjoint, since if one were contained in the other, it would contradict the maximality. By the choice of these cubes, ΩλdjQj\Omega_{\lambda}^{d}\subset\bigcup_{j}Q_{j}, hence

mn(Ωλd)\displaystyle m_{n}(\Omega_{\lambda}^{d}) jmn(Qj)1λ11αjmn(Qj)|1mn(Qj)1αQjF(y)𝑑y|11α(1λn|F(x)|𝑑x)11α.\displaystyle\leqslant\sum_{j}m_{n}(Q_{j})\leqslant\frac{1}{\lambda^{\frac{1}{1-\alpha}}}\sum_{j}m_{n}(Q_{j})\left|\frac{1}{m_{n}(Q_{j})^{1-\alpha}}\int_{Q_{j}}F(y)dy\right|^{\frac{1}{1-\alpha}}\leqslant\left(\frac{1}{\lambda}\int_{\mathbb{R}^{n}}|F(x)|dx\right)^{\frac{1}{1-\alpha}}.

Then we consider the case p=1/αp=1/\alpha, q=q=\infty. Since FL𝒦1/α(n,||)F\in L^{1/\alpha}_{\mathcal{K}}(\mathbb{R}^{n},|\cdot|), by the (1/α,)(1/\alpha,\infty) boundedness of AQ,αA_{Q,\alpha}, for a.e. xΩx\in\Omega, we have

|AQ,αF(x)|CFL𝒦1α(n,||),|A_{Q,\alpha}F(x)|\leqslant C\lVert F\rVert_{L^{\frac{1}{\alpha}}_{\mathcal{K}}(\mathbb{R}^{n},|\cdot|)},

which implies that

M^αdFL𝒦(n,||)CFL𝒦1α(n,||).\lVert\widehat{M}^{d}_{\alpha}F\rVert_{L^{\infty}_{\mathcal{K}}(\mathbb{R}^{n},|\cdot|)}\leqslant C\lVert F\rVert_{L^{\frac{1}{\alpha}}_{\mathcal{K}}(\mathbb{R}^{n},|\cdot|)}.

Finally, by Theorem 3.1 with T=M^αdT=\widehat{M}^{d}_{\alpha}, Ω=n\Omega=\mathbb{R}^{n}, ρ()=||\rho(\cdot)=|\cdot|, p0=1/αp_{0}=1/\alpha, p1=1p_{1}=1, q0=q_{0}=\infty and q1=1/(1α)q_{1}=1/(1-\alpha), we finish the proof. ∎

4 Riesz-Thorin interpolation theorem and its application

Complex method is often important in operator theory, see [22, 23]. In this section, we first prove Riesz-Thorin interpolation theorem, which needs some results of density, analytic functions and dual norms.

Definition 4.1.

[10, 8] Given a norm ρx\rho_{x} on d\mathbb{R}^{d} and two compact sets K1,K2dK_{1},K_{2}\subset\mathbb{R}^{d}, the corresponding Hausdorff distance function is defined by

dH,x(K1,K2)=max{supvK1infwK2ρx(vw),supvK2infwK1ρx(vw)}.d_{H,x}(K_{1},K_{2})=\max\{\sup_{v\in K_{1}}\inf_{w\in K_{2}}\rho_{x}(v-w),\sup_{v\in K_{2}}\inf_{w\in K_{1}}\rho_{x}(v-w)\}.

Hence, define the Hausdorff distance on 𝒦bcs(d)\mathcal{K}_{bcs}(\mathbb{R}^{d}) as

dp(F,G)=(ΩdH,x(F(x),G(x))p𝑑μ(x))1p.d_{p}(F,G)=\left(\int_{\Omega}d_{H,x}(F(x),G(x))^{p}d\mu(x)\right)^{\frac{1}{p}}.
Definition 4.2.

[8] A measurable function F:Ω𝒦bcs(d)F:\Omega\rightarrow\mathcal{K}_{bcs}(\mathbb{R}^{d}) is called a simple function if FF takes only finitely many values S1,S2,,Sm𝒦bcs(d)S_{1},S_{2},\cdots,S_{m}\in\mathcal{K}_{bcs}(\mathbb{R}^{d}), hence, it can be written in the form

F(x)=k=1mχAk(x)SkforxΩ,F(x)=\sum_{k=1}^{m}\chi_{A_{k}}(x)S_{k}\ \ for\ x\in\Omega,

where A1,A2,,AmA_{1},A_{2},\cdots,A_{m} are disjoint measurable sets such that

k=1mAk=Ω.\bigcup_{k=1}^{m}A_{k}=\Omega.
Lemma 4.1.

[8] Every measurable function F:Ω𝒦bcs(d)F:\Omega\rightarrow\mathcal{K}_{bcs}(\mathbb{R}^{d}) is the pointwise limit of simple measurable functions with respect to the Hausdorff distance on 𝒦bcs(d).\mathcal{K}_{bcs}(\mathbb{R}^{d}).

Lemma 4.2.

For any p>0p>0, FL𝒦p(Ω,ρ)F\in L_{\mathcal{K}}^{p}(\Omega,\rho) and ϵ>0\epsilon>0, there exists a simple measurable function GG such as dp(F,G)<ϵd_{p}(F,G)<\epsilon.

Proof.

By the proof of[8, Lemma 4.4], there exist simple measurable functions {Fn}n\{F_{n}\}_{n\in\mathbb{N}} such that

F1(x)F2(x),F(x)=nFn(x)andlimndH,x(Fn(x),F(x))=0F_{1}(x)\subset F_{2}(x)\subset\cdots,\ \ F(x)=\bigcup_{n\in\mathbb{N}}F_{n}(x)\ \ and\ \ \lim_{n\to\infty}d_{H,x}(F_{n}(x),F(x))=0

for any xΩx\in\Omega. Therefore,

dH,x(Fn(x),F(x))dH,x({0},F(x)),d_{H,x}(F_{n}(x),F(x))\leqslant d_{H,x}(\{0\},F(x)),

and

(ΩdH,x({0},F(x))p𝑑μ(x))1p=fL𝒦p(Ω,ρ)<.\left(\int_{\Omega}d_{H,x}(\{0\},F(x))^{p}d\mu(x)\right)^{\frac{1}{p}}=\lVert f\rVert_{L_{\mathcal{K}}^{p}(\Omega,\rho)}<\infty.

By Lebesgue dominated covergence theorem,

limndp(Fn,F)\displaystyle\lim_{n\to\infty}d_{p}(F_{n},F) =limn(ΩdH,x(Fn(x),F(x))p𝑑μ(x))1p\displaystyle=\lim_{n\to\infty}\left(\int_{\Omega}d_{H,x}(F_{n}(x),F(x))^{p}d\mu(x)\right)^{\frac{1}{p}}
=(ΩlimndH,x(Fn(x),F(x))pdμ(x))1p=0.\displaystyle=\left(\int_{\Omega}\lim_{n\to\infty}d_{H,x}(F_{n}(x),F(x))^{p}d\mu(x)\right)^{\frac{1}{p}}=0.

Setting G(x)=Fn(x)G(x)=F_{n}(x) for sufficiently large nn finishes the proof. ∎

Lemma 4.3.

[24] Suppose μ\mu is a complex measure on a measure space XX, Ω\Omega is an open set in the plane, ϕ\phi is a bounded function on Ω×X\Omega\times X such that ϕ(z,t)\phi(z,t) is a measurable function of tt for each zΩz\in\Omega, and ϕ(z,t)\phi(z,t) is analytic in Ω\Omega for each tXt\in X. For zΩz\in\Omega, define

f(z)=Xϕ(z,t)𝑑μ(t),f(z)=\int_{X}\phi(z,t)d\mu(t),

then ff is analytic in Ω\Omega.

Lemma 4.4.

[25] Let AA be analytic on the open strip S={z:0<Rez<1}S=\{z\in\mathbb{C}:0<\mathrm{Re}z<1\} and continuous on its closure such that

supzS¯ea|Imz|log|A(z)|A<\sup_{z\in\bar{S}}e^{-a|\mathrm{Im}z|}\log|A(z)|\leqslant A<\infty

for some fixed AA and a<πa<\pi. Then

|A(x)|exp{sin(πx)2[log|A(iy)|cosh(πy)cos(πx)+log|A(1+iy)|cosh(πy)+cos(πx)]𝑑y}|A(x)|\leqslant\exp\left\{\frac{\sin(\pi x)}{2}\int_{-\infty}^{\infty}\left[\frac{\log|A(iy)|}{\cosh(\pi y)-\cos(\pi x)}+\frac{\log|A(1+iy)|}{\cosh(\pi y)+\cos(\pi x)}\right]dy\right\}

whenever 0<x<10<x<1.

Definition 4.3.

[8] Given a seminorm pp, define p:d[0,)p^{\ast}:\mathbb{R}^{d}\rightarrow[0,\infty) by

p(v)=supwd,p(w)1|v,w|.p^{\ast}(v)=\sup_{w\in\mathbb{R}^{d},p(w)\leqslant 1}|\langle v,w\rangle|. (4)
Definition 4.4.

[8] If ρ:Ω×d[0,)\rho:\Omega\times\mathbb{R}^{d}\rightarrow[0,\infty) is a norm function, then ρ:Ω×d[0,)\rho^{\ast}:\Omega\times\mathbb{R}^{d}\rightarrow[0,\infty), defined by ρx(v)=(ρx)(v)\rho_{x}^{\ast}(v)=(\rho_{x})^{\ast}(v), is a measurable norm function.

Definition 4.5.

[26, 8] Let p0p_{0}, p1p_{1} be two norms, ρ0\rho_{0}, ρ1\rho_{1} be two norm functions. For 0<t<10<t<1, define their weighted geometric mean by

pt(v)=p0(v)1tp1(v)t,ρt(x,v)=ρ0(x,v)1tρ1(x,v)tforxΩ,vd.p_{t}(v)=p_{0}(v)^{1-t}p_{1}(v)^{t},\ \ \rho_{t}(x,v)=\rho_{0}(x,v)^{1-t}\rho_{1}(x,v)^{t}\ \ for\ x\in\Omega,\ v\in\mathbb{R}^{d}.

Let AA and BB be two symmetric positive definite matrices, for 0<t<10<t<1, define their weighted geometric mean by

A#tB=A12(A12BA12)tA12.A\#_{t}B=A^{\frac{1}{2}}(A^{-\frac{1}{2}}BA^{-\frac{1}{2}})^{t}A^{\frac{1}{2}}.

The function ptp_{t} may not be a norm. However, if we still use (4) to define ptp_{t}^{\ast}, it will be a norm.

Lemma 4.5.

[8] Let p0p_{0}, p1p_{1} be two norms, then for 0<t<10<t<1, ptp_{t}^{\ast} is a norm. Moreover, ptp_{t}^{\ast\ast} is a norm, and for all vdv\in\mathbb{R}^{d}, pt(v)pt(v)p_{t}^{\ast\ast}(v)\leqslant p_{t}(v).

Theorem 4.1.

Let 1p0,p1,q0,q1<1\leqslant p_{0},p_{1},q_{0},q_{1}<\infty, ρ0\rho_{0}, ρ1\rho_{1} be norm functions on Ω\Omega, XX be the set of all locally integrably bounded functions F:Ω𝒦bcs(d)F:\Omega\rightarrow\mathcal{K}_{bcs}(\mathbb{R}^{d}), T:XXT:X\rightarrow X be a linear, monotone operator. Suppose that

T(F)L𝒦q0(Ω,ρ0)M0FL𝒦p0(Ω,ρ0),\lVert T(F)\rVert_{L^{q_{0}}_{\mathcal{K}}(\Omega,\rho_{0})}\leqslant M_{0}\lVert F\rVert_{L^{p_{0}}_{\mathcal{K}}(\Omega,\rho_{0})},
T(F)L𝒦q1(Ω,ρ1)M1FL𝒦q0(Ω,ρ1),\lVert T(F)\rVert_{L^{q_{1}}_{\mathcal{K}}(\Omega,\rho_{1})}\leqslant M_{1}\lVert F\rVert_{L^{q_{0}}_{\mathcal{K}}(\Omega,\rho_{1})},

for all FXF\in X, and ρ0(T(χAS))\rho_{0}(T(\chi_{A}S)), ρ1(T(χAS))\rho_{1}(T(\chi_{A}S)) are bounded for all sets AnA\subset\mathbb{R}^{n}, SdS\subset\mathbb{R}^{d} that make ρ0(χAS)\rho_{0}(\chi_{A}S), ρ1(χAS)\rho_{1}(\chi_{A}S) bounded. Then for any 0<θ<10<\theta<1 and simple function FL𝒦p(Ω,ρ)F\in L^{p}_{\mathcal{K}}(\Omega,\rho), we have

T(F)L𝒦q(Ω,ρ)MθFL𝒦p(Ω,ρ),\lVert T(F)\rVert_{L^{q}_{\mathcal{K}}(\Omega,\rho)}\leqslant M_{\theta}\lVert F\rVert_{L^{p}_{\mathcal{K}}(\Omega,\rho)},

where MθM_{\theta} is a constant independent of FF, and

ρ=ρθ,1p=1θp0+θp1,1q=1θq0+θq1.\rho=\rho_{\theta}^{\ast\ast},\ \ \frac{1}{p}=\frac{1-\theta}{p_{0}}+\frac{\theta}{p_{1}},\ \ \frac{1}{q}=\frac{1-\theta}{q_{0}}+\frac{\theta}{q_{1}}. (5)

By density, TT is bounded from L𝒦p(Ω,ρ)L^{p}_{\mathcal{K}}(\Omega,\rho) to L𝒦q(Ω,ρ)L^{q}_{\mathcal{K}}(\Omega,\rho) for all ρ\rho, pp and qq as in (5).

Proof.

First, assume that FL𝒦p(Ω,ρ)=1\lVert F\rVert_{L^{p}_{\mathcal{K}}(\Omega,\rho)}=1. Let

F=k=1mχAkSkF=\sum_{k=1}^{m}\chi_{A_{k}}S_{k}

be a simple function in XX, where Sk𝒦bcsS_{k}\in\mathcal{K}_{bcs} and AkA_{k} are pairwise disjoint subsets of Ω\Omega with finite measure. We need to control

T(F)L𝒦q(Ω,ρ)=(Ωρ(T(F)(x))q𝑑x)1q=supg|Ωρ(T(F)(x))g(x)𝑑x|,\lVert T(F)\rVert_{L^{q}_{\mathcal{K}}(\Omega,\rho)}=\left(\int_{\Omega}\rho(T(F)(x))^{q}dx\right)^{\frac{1}{q}}=\sup_{g}\left|\int_{\Omega}\rho(T(F)(x))g(x)dx\right|,

where the supremum is taken over all simple functions gLq(Ω)g\in L^{q^{\prime}}(\Omega) with Lq(Ω)L^{q^{\prime}}(\Omega) norm not bigger than 1. Write

g=j=1nbjeiβjχBj,g=\sum_{j=1}^{n}b_{j}e^{i\beta_{j}}\chi_{B_{j}},

where bj>0b_{j}>0, βj\beta_{j} are real, and BjB_{j} are pairwise disjoint subsets of Ω\Omega with finite measure.
Denote the open strip S={z:0<Rez<1}S=\{z\in\mathbb{C}:0<\mathrm{Re}z<1\}, and its closure S¯={z:0Rez1}\bar{S}=\{z\in\mathbb{C}:0\leqslant\mathrm{Re}z\leqslant 1\}. For zS¯z\in\bar{S}, define

P(z)=pp0(1z)+pp1z,Q(z)=qq0(1z)+qq1z,P(z)=\frac{p}{p_{0}}(1-z)+\frac{p}{p_{1}}z,\ \ Q(z)=\frac{q^{\prime}}{q^{\prime}_{0}}(1-z)+\frac{q^{\prime}}{q^{\prime}_{1}}z,

and

A(z)=Ωρ0(T(F)(x))(1z)P(z)ρ1(T(F)(x))zP(z)gz(x)𝑑x,A(z)=\int_{\Omega}\rho_{0}(T(F)(x))^{(1-z)P(z)}\rho_{1}(T(F)(x))^{zP(z)}g_{z}(x)dx,

where

gz=j=1nbjQ(z)eiβjχBj.g_{z}=\sum_{j=1}^{n}b_{j}^{Q(z)}e^{i\beta_{j}}\chi_{B_{j}}.

We claim that F(z)F(z) is analytic on SS. By the linearity of TT,

A(z)=j=1nbjQ(z)eiβjΩρ0(k=1mT(SkχAk)(x))(1z)P(z)ρ1(k=1mT(SkχAk)(x))zP(z)χBj(x)𝑑x.A(z)=\sum_{j=1}^{n}b_{j}^{Q(z)}e^{i\beta_{j}}\int_{\Omega}\rho_{0}\left(\sum_{k=1}^{m}T(S_{k}\chi_{A_{k}})(x)\right)^{(1-z)P(z)}\rho_{1}\left(\sum_{k=1}^{m}T(S_{k}\chi_{A_{k}})(x)\right)^{zP(z)}\chi_{B_{j}}(x)dx.

Consider the open set Sλ:={z:0<Re(z)<1,|Im(z)|<λ}S_{\lambda}:=\{z\in\mathbb{C}:0<\mathrm{Re}(z)<1,\ |\mathrm{Im}(z)|<\lambda\} for given λ>0\lambda>0, for zSλz\in S_{\lambda}, by the sublinearity of ρ0\rho_{0} and ρ1\rho_{1},

|A(z)|\displaystyle|A(z)| j=1nbj|Q(z)|Ω(k=1mρ0(T(χAkSk)(x)))|1z||P(z)|\displaystyle\leqslant\sum_{j=1}^{n}b_{j}^{|Q(z)|}\int_{\Omega}\left(\sum_{k=1}^{m}\rho_{0}(T(\chi_{A_{k}}S_{k})(x))\right)^{|1-z||P(z)|}
×(k=1mρ1(T(χAkSk)(x)))|z||P(z)|χBj(x)dx\displaystyle\ \ \ \ \times\left(\sum_{k=1}^{m}\rho_{1}(T(\chi_{A_{k}}S_{k})(x))\right)^{|z||P(z)|}\chi_{B_{j}}(x)dx
=:j=1nbj|Q(z)|Ωϕ(z,x)dx.\displaystyle=:\sum_{j=1}^{n}b_{j}^{|Q(z)|}\int_{\Omega}\phi(z,x)dx.

For all xΩx\in\Omega and zSλz\in S_{\lambda}, ρ0(T(χAkSk)(x))\rho_{0}(T(\chi_{A_{k}}S_{k})(x)), ρ1(T(χAkSk)(x))\rho_{1}(T(\chi_{A_{k}}S_{k})(x)) are bounded by assumption, and |Q(z)||Q(z)|, |1z||P(z)||1-z||P(z)|, |z||P(z)||z||P(z)| are all bounded by a constant CλC_{\lambda} independent of zz, thus ϕ(z,x)\phi(z,x) is bounded on Sλ×ΩS_{\lambda}\times\Omega. By Lemma 4.3, F(z)F(z) is analytic in SλS_{\lambda}, and then F(z)F(z) is analytic in SS by the arbitrary of λ\lambda.
Besides, for all zS¯z\in\bar{S},

log|A(z)|\displaystyle\log|A(z)| log(j=1nbj|Q(z)||Bj|C1|1z||P(z)|C2|z||P(z)|)\displaystyle\leqslant\log\left(\sum_{j=1}^{n}b_{j}^{|Q(z)|}|B_{j}|C_{1}^{|1-z||P(z)|}C_{2}^{|z||P(z)|}\right)
log(C1|1z||P(z)|C2|z||P(z)|)(|1z|+|z|)|P(z)|\displaystyle\lesssim\log(C_{1}^{|1-z||P(z)|}C_{2}^{|z||P(z)|})\lesssim(|1-z|+|z|)|P(z)|
|1z|2+|z|21+|Imz|2,\displaystyle\lesssim|1-z|^{2}+|z|^{2}\lesssim 1+|\mathrm{Im}z|^{2},

thus we have

supzS¯e|Imz|log|A(z)|supzS¯e|Imz|(1+|Imz|2)=1,\sup_{z\in\bar{S}}e^{-|\mathrm{Im}z|}\log|A(z)|\lesssim\sup_{z\in\bar{S}}e^{-|\mathrm{Im}z|}(1+|\mathrm{Im}z|^{2})=1,

therefore, by Lemma 4.4, for all 0<x<10<x<1,

|A(x)|exp{sin(πx)2[log|A(iy)|cosh(πy)cos(πx)+log|A(1+iy)|cosh(πy)+cos(πx)]𝑑y}.|A(x)|\leqslant\exp\left\{\frac{\sin(\pi x)}{2}\int_{-\infty}^{\infty}\left[\frac{\log|A(iy)|}{\cosh(\pi y)-\cos(\pi x)}+\frac{\log|A(1+iy)|}{\cosh(\pi y)+\cos(\pi x)}\right]dy\right\}.

Note that

[log|A(iy)|cosh(πy)cos(πx)+log|A(1+iy)|cosh(πy)+cos(πx)]𝑑y\displaystyle\int_{-\infty}^{\infty}\left[\frac{\log|A(iy)|}{\cosh(\pi y)-\cos(\pi x)}+\frac{\log|A(1+iy)|}{\cosh(\pi y)+\cos(\pi x)}\right]dy
[1+|y|2e|y|cos(πx)+1+|y|2e|y|+cos(πx)]𝑑y<C(x)<,\displaystyle\ \ \ \ \lesssim\int_{-\infty}^{\infty}\left[\frac{1+|y|^{2}}{e^{|y|}-\cos(\pi x)}+\frac{1+|y|^{2}}{e^{|y|}+\cos(\pi x)}\right]dy<C(x)<\infty,

we have |A(x)|Mx<|A(x)|\leqslant M_{x}<\infty. Set x=θx=\theta, we obtain

|A(θ)|=Ωρ0(T(F)(x))(1θ)P(θ)ρ1(T(F)(x))θP(θ)gθ(x)𝑑xMθ.|A(\theta)|=\int_{\Omega}\rho_{0}(T(F)(x))^{(1-\theta)P(\theta)}\rho_{1}(T(F)(x))^{\theta P(\theta)}g_{\theta}(x)dx\leqslant M_{\theta}.

Note that P(θ)=Q(θ)=1P(\theta)=Q(\theta)=1, gθ(x)=g(x)g_{\theta}(x)=g(x), by Lemma 4.5, we have

Ωρ(T(F)(x))g(x)𝑑xMθ.\int_{\Omega}\rho(T(F)(x))g(x)dx\leqslant M_{\theta}.

Therefore, T(F)L𝒦q(Ω,ρ)Mθ\lVert T(F)\rVert_{L^{q}_{\mathcal{K}}(\Omega,\rho)}\leqslant M_{\theta} for all simple functions FL𝒦p(Ω,ρ)F\in L^{p}_{\mathcal{K}}(\Omega,\rho) satisfying FL𝒦p(Ω,ρ)=1\lVert F\rVert_{L^{p}_{\mathcal{K}}(\Omega,\rho)}=1. For an arbitrary simple function FL𝒦p(Ω,ρ)F\in L^{p}_{\mathcal{K}}(\Omega,\rho), by the linearity of TT and ρ\rho (with number multiplication), we obtain the same conclusion.
Finally, for any FL𝒦p(Ω,ρ)F\in L^{p}_{\mathcal{K}}(\Omega,\rho) and ϵ>0\epsilon>0, by Lemma 4.2, there exists simple measurable function GXG\in X satisfying G(x)F(x)G(x)\subset F(x) for all xΩx\in\Omega, and

dp(F,G)=(ΩdH,x(F(x),G(x))p𝑑μ(x))1p<ϵ.d_{p}(F,G)=\left(\int_{\Omega}d_{H,x}(F(x),G(x))^{p}d\mu(x)\right)^{\frac{1}{p}}<\epsilon.

Set d(x)=dH,x(F(x),G(x))d(x)=d_{H,x}(F(x),G(x)), then dLp(Ω)d\in L^{p}(\Omega) and dp<ϵ\lVert d\rVert_{p}<\epsilon. By the definition of Hausdorff distance, for all xΩx\in\Omega,

F(x)G(x)+d(x)𝑩¯.F(x)\subset G(x)+d(x)\boldsymbol{\overline{B}}.

Then, by the linearity and monotonicity of TT,

T(F)L𝒦q(Ω,ρ)\displaystyle\lVert T(F)\rVert_{L^{q}_{\mathcal{K}}(\Omega,\rho)} T(G)L𝒦q(Ω,ρ)+T(d()𝑩¯)L𝒦q(Ω,ρ)Mθ(GL𝒦p(Ω,ρ)+d()𝑩¯L𝒦p(Ω,ρ))\displaystyle\leqslant\lVert T(G)\rVert_{L^{q}_{\mathcal{K}}(\Omega,\rho)}+\lVert T(d(\cdot)\boldsymbol{\overline{B}})\rVert_{L^{q}_{\mathcal{K}}(\Omega,\rho)}\leqslant M_{\theta}(\lVert G\rVert_{L^{p}_{\mathcal{K}}(\Omega,\rho)}+\lVert d(\cdot)\boldsymbol{\overline{B}}\rVert_{L^{p}_{\mathcal{K}}(\Omega,\rho)})
=Mθ(GL𝒦p(Ω,ρ)+dp)<Mθ(FL𝒦p(Ω,ρ)+ϵ).\displaystyle=M_{\theta}(\lVert G\rVert_{L^{p}_{\mathcal{K}}(\Omega,\rho)}+\lVert d\rVert_{p})<M_{\theta}(\lVert F\rVert_{L^{p}_{\mathcal{K}}(\Omega,\rho)}+\epsilon).

By the arbitrary of ϵ\epsilon, we have T(F)L𝒦q(Ω,ρ)MθFL𝒦p(Ω,ρ)\lVert T(F)\rVert_{L^{q}_{\mathcal{K}}(\Omega,\rho)}\leqslant M_{\theta}\lVert F\rVert_{L^{p}_{\mathcal{K}}(\Omega,\rho)}, which completes the proof. ∎

By the similar argument as Remark 3.1, Theorem 4.1 can also deduce Riesz-Theoin interpolation theorem on Lebesgue spaces with some additional conditions of operator TT.

To prove the reverse factorization property, we need the 𝒜p,norm\mathcal{A}_{p,norm} condition of norm functions.

Definition 4.6.

[8] Fix 1p<1\leqslant p<\infty and suppose ρ(,v)Llocp\rho(\cdot,v)\in L^{p}_{loc} for all vdv\in\mathbb{R}^{d}, define

ρp,Q(v)=(1mn(Q)Qρx(v)p𝑑x)1p.\langle\rho\rangle_{p,Q}(v)=\left(\frac{1}{m_{n}(Q)}\int_{Q}\rho_{x}(v)^{p}dx\right)^{\frac{1}{p}}.
Definition 4.7.

[8] Given a norm function ρ:n×d[0,)\rho:\mathbb{R}^{n}\times\mathbb{R}^{d}\rightarrow[0,\infty), then for 1p<1\leqslant p<\infty, ρ𝒜p,norm\rho\in\mathcal{A}_{p,norm} if for every cube QQ and vdv\in\mathbb{R}^{d},

ρp,Q(v)ρp,Q(v).\langle\rho^{\ast}\rangle_{p^{\prime},Q}(v)\lesssim\langle\rho\rangle_{p,Q}^{\ast}(v).
Definition 4.8.

[8] A matrix mapping is called measurable if each of its components is a measurable function. Let A:ΩdA:\Omega\rightarrow\mathcal{M}_{d} be a measurable matrix mapping, define an seminorm function ρA\rho_{A} by

ρA(x,v)=|A(x)v|,xΩ,vd.\rho_{A}(x,v)=|A(x)v|,\ \ x\in\Omega,\ v\in\mathbb{R}^{d}.

For each xx, ρA(x,)\rho_{A}(x,\cdot) is a seminorm, and since AA is measurable, the map xρA(x,v)x\rightarrow\rho_{A}(x,v) is measurable for all vv. Reversely, for a given norm function, the similar result exists.

Lemma 4.6.

[8] Let ρ\rho be a norm function, then there exists an associated measurable matrix mapping W:Ω𝒮dW:\Omega\rightarrow\mathcal{S}_{d} such that for a.e. xΩx\in\Omega and every vdv\in\mathbb{R}^{d}, W(x)W(x) is positive definite, and

ρW(x,v)ρ(x,v)dρW(x,v).\rho_{W}(x,v)\leqslant\rho(x,v)\leqslant\sqrt{d}\rho_{W}(x,v).

The following lemma shows the connection between norm functions and their associated matrix weights.

Lemma 4.7.

[8] Given a norm function ρ:n×d[0,)\rho:\mathbb{R}^{n}\times\mathbb{R}^{d}\rightarrow[0,\infty) with associated matrix mapping WW and 1<p<1<p<\infty, ρ𝒜p,norm\rho\in\mathcal{A}_{p,norm} if and only if W𝒜p,matrixW\in\mathcal{A}_{p,matrix}, that is,

[W]𝒜p,matrix:=supQ(1mn(Q)Q(1mn(Q)QW(x)W1(y)opp𝑑y)pp𝑑x)1p<.[W]_{\mathcal{A}_{p,matrix}}:=\sup_{Q}\left(\frac{1}{m_{n}(Q)}\int_{Q}\left(\frac{1}{m_{n}(Q)}\int_{Q}\lVert W(x)W^{-1}(y)\rVert_{\mathrm{op}}^{p^{\prime}}dy\right)^{\frac{p}{p^{\prime}}}dx\right)^{\frac{1}{p}}<\infty.

When p=1p=1, ρ𝒜1,norm\rho\in\mathcal{A}_{1,norm} if and only if W𝒜1,matrixW\in\mathcal{A}_{1,matrix}, that is,

[W]𝒜1,matrix:=supQesssupxQ1mn(Q)QW1(x)W(y)op𝑑y<.[W]_{\mathcal{A}_{1,matrix}}:=\sup_{Q}\mathop{\mathrm{ess}\sup}\limits_{x\in Q}\frac{1}{m_{n}(Q)}\int_{Q}\lVert W^{-1}(x)W(y)\rVert_{\mathrm{op}}dy<\infty.

We need the following characterization of norm weights.

Definition 4.9.

[8] Let F:n𝒦bcs(d)F:\mathbb{R}^{n}\rightarrow\mathcal{K}_{bcs}(\mathbb{R}^{d}) that is locally integrably bounded, for a fixed cube QQ, we define the averaging operator AQA_{Q} by

AQF(x)=1|Q|QF(y)𝑑yχQ(x).A_{Q}F(x)=\frac{1}{|Q|}\int_{Q}F(y)dy\cdot\chi_{Q}(x).
Lemma 4.8.

[8] Given 1p<1\leqslant p<\infty and a norm function ρ\rho, the following propositions are equivalent:

(i) ρ𝒜p,norm\rho\in\mathcal{A}_{p,norm};

(ii) Given any cube QQ, AQ:L𝒦p(n,ρ)L𝒦p(n,ρ)A_{Q}:L^{p}_{\mathcal{K}}(\mathbb{R}^{n},\rho)\rightarrow L^{p}_{\mathcal{K}}(\mathbb{R}^{n},\rho) is bounded.

We also need the following results about the weighted geometric mean.

Lemma 4.9.

[8] Suppose that A,B𝒮dA,B\in\mathcal{S}_{d}, and the norms p0p_{0} and p1p_{1} are given by

p0(v)=|A12v|andp1(v)=|B12v|forvd,p_{0}(v)=|A^{\frac{1}{2}}v|\ \ and\ \ p_{1}(v)=|B^{\frac{1}{2}}v|\ \ for\ v\in\mathbb{R}^{d},

then the double dual of the weighted geometric mean ptp_{t} satisfies

pt(v)|(A#tB)12v|forvd.p_{t}^{\ast\ast}(v)\approx|(A\#_{t}B)^{\frac{1}{2}}v|\ \ for\ v\in\mathbb{R}^{d}.

Now, we prove the reverse factorization property.

Theorem 4.2.

Given 1p0,p1<1\leqslant p_{0},p_{1}<\infty, suppose that matrix mappings W0𝒜p0,matrixW_{0}\in\mathcal{A}_{p_{0},matrix}, W1𝒜p1,matrixW_{1}\in\mathcal{A}_{p_{1},matrix}, and sup|v|=1|W0(x)v|\sup_{|v|=1}|W_{0}(x)v|, sup|v|=1|W1(x)v|\sup_{|v|=1}|W_{1}(x)v| are bounded for xnx\in\mathbb{R}^{n}. Then for any 0<t<10<t<1, W¯=(W02#tW12)1/2𝒜p,matrix\overline{W}=(W_{0}^{2}\#_{t}W_{1}^{2})^{1/2}\in\mathcal{A}_{p,matrix}, where

1p=1tp0+1p1.\frac{1}{p}=\frac{1-t}{p_{0}}+\frac{1}{p_{1}}.
Proof.

Define the following norm functions:

ρ0(x,v)=|W0(x)v|,ρ1(x,v)=|W1(x)v|andρ(x,v)=|W¯(x)v|forxΩ,vd.\rho_{0}(x,v)=|W_{0}(x)v|,\ \ \rho_{1}(x,v)=|W_{1}(x)v|\ \ and\ \ \rho(x,v)=|\overline{W}(x)v|\ \ for\ x\in\Omega,\ v\in\mathbb{R}^{d}.

By the assumption of W0W_{0} and W1W_{1}, Lemma 4.7 and Lemma 4.8, for any cube QQ,

AQ:L𝒦p0(n,ρ0)L𝒦p0(n,ρ0),AQ:L𝒦p1(n,ρ1)L𝒦p1(n,ρ1).A_{Q}:L^{p_{0}}_{\mathcal{K}}(\mathbb{R}^{n},\rho_{0})\rightarrow L^{p_{0}}_{\mathcal{K}}(\mathbb{R}^{n},\rho_{0}),\ \ A_{Q}:L^{p_{1}}_{\mathcal{K}}(\mathbb{R}^{n},\rho_{1})\rightarrow L^{p_{1}}_{\mathcal{K}}(\mathbb{R}^{n},\rho_{1}).

Therefore, by Theorem 4.1,

AQ:L𝒦p(n,ρt)L𝒦p(n,ρt).A_{Q}:L^{p}_{\mathcal{K}}(\mathbb{R}^{n},\rho_{t}^{\ast\ast})\rightarrow L^{p}_{\mathcal{K}}(\mathbb{R}^{n},\rho_{t}^{\ast\ast}).

Using Lemma 4.8 again, we have ρt𝒜p\rho_{t}^{\ast\ast}\in\mathcal{A}_{p}. By Lemma 4.9, ρt(x,v)ρ(x,v)\rho_{t}^{\ast\ast}(x,v)\approx\rho(x,v), which implies that ρ𝒜p,norm\rho\in\mathcal{A}_{p,norm}. Finally by Lemma 4.7, we complete the proof. ∎


Acknowledgments
The authors thank the referees for their careful reading and helpful comments which indeed improved the presentation of this article.
Funding information
The research was supported by Natural Science Foundation of China (Grant No. 12061069).
Authors contributions
All authors have accepted responsibility for the entire content of this manuscript and approved its submission.
Conflict of interest
Authors state no conflict of interest.

Yuxun Zhang and Jiang Zhou
College of Mathematics and System Sciences, Xinjiang University, Urumqi 830046
E-mail : [email protected] (Yuxun Zhang); [email protected] (Jiang Zhou)

References

  • [1] B. Muckenhoupt. Weighted norm inequalities for the Hardy maximal function. Transactions of the American Mathematical Society, 165:207–226, 1972.
  • [2] S. Treil F. L. Nazarov. The hunt for a Bellman function: applications to estimates for singular integral operators and to other classical problems of harmonic analysis. Algebra i analiz, 8(5):32–162, 1996.
  • [3] A. Volberg S. Treil. Wavelets and the angle between past and future. Journal of functional analysis, 143(2):269–308, 1997.
  • [4] A. Volberg. Matrix ApA_{p} weights via S-functions. Journal of the American Mathematical Society, 10(2):445–466, 1997.
  • [5] S. Roudenko. Matrix-weighted Besov spaces. Transactions of the American Mathematical Society, 355(1):273–314, 2003.
  • [6] P. W. Jones. Factorization of ApA_{p} weights. Annals of Mathematics, 511–530, 1980.
  • [7] J. L. Rubio de Francia. Factorization theory and ApA_{p} weights. American Journal of Mathematics, 533–547, 1984.
  • [8] D. Cruz-Uribe M. Bownik. Extrapolation and factorization of matrix weights. arXiv preprint arXiv:2210.09443, 2022.
  • [9] H. Frankowska J. Aubin. Set-valued analysis. Springer Science & Business Media, 2009.
  • [10] M. Valadier C. Castaing. Measurable multifunctions. Springer, 1977.
  • [11] R. T. Rockafellar. Convex analysis, volume 11. Princeton university press, 1997.
  • [12] J. Marcinkiewicz. Sur l’interpolation d’operations. C. R. Acad. Sci., 1272–1273, 1939.
  • [13] A. Zygmund. On a theorem of Marcinkiewicz concerning interpolation of operators. J. de Math., 35:223–248, 1956.
  • [14] S. Chanillo. A note on commutators. Indiana University Mathematics Journal, 31(1):7–16, 1982.
  • [15] R. Wheeden B. Muckenhoupt. Weighted norm inequalities for fractional integrals. Transactions of the American Mathematical Society, 192:261–274, 1974.
  • [16] A. Zygmund. Trigonometric series, volume 1. Cambridge university press, 1959.
  • [17] M. Riesz. Sur les maxima des formes bilinéaires et sur les fonctionnelles linéaires. Acta mathematica, 49(3-4):465–497, 1927.
  • [18] G. Thorin. An extension of a convexity theorem due to M. Riesz. Fysiogr. Sällsk. Lund Förh, 8:166–170, 1939.
  • [19] G. Thorin. Convesity theorems generalizing those of M. Riesz and Hadamard with some applications. Comm. Sem. Math. Univ. Lund [Medd. Lunds Univ. Mat. Sem.], 1–58, 1948.
  • [20] D. Yong. Foundations of modern analysis (in Chinese). Beijing Normal University Press, 2008.
  • [21] J. Isralowitz D. Cruz-Uribe and K. Moen. Two weight bump conditions for matrix weights. Integral Equations and Operator Theory, 90:1–31, 2018.
  • [22] B. Guanlong Y. Fangqin. A Carleson type measure and a family of Möbius invariant function spaces. Indagationes Mathematicae, 34(6):1338–1354, 2023.
  • [23] C. K. Chui Z. Lefan. Order of uniform approximation by polynomial interpolation in the complex plane and beyond. Indagationes Mathematicae, 34(2):418–456, 2023.
  • [24] W. Rudin. Real and complex analysis. McGraw-Hill Book Company, 1970.
  • [25] L. Grafakos. Classical Fourier analysis. Springer, 2008.
  • [26] R. Bhatia. Positive definite matrices. Princeton University Press, 2007.