Abstract

The paper is devoted to investigation of new Lebesgue's type differentiation theorems (LDT) in rearrangement invariant (r.i.) quasi-Banach spaces 𝐸 and in particular on Lorentz spaces Γ𝑝,𝑤={𝑓(𝑓)𝑝𝑤<} for any 0<𝑝< and a nonnegative locally integrable weight function 𝑤, where 𝑓 is a maximal function of the decreasing rearrangement 𝑓 for any measurable function 𝑓 on (0,𝛼), with 0<𝛼. The first type of LDT in the spirit of Stein (1970), characterizes the convergence of quasinorm averages of 𝑓𝐸, where 𝐸 is an order continuous r.i. quasi-Banach space. The second type of LDT establishes conditions for pointwise convergence of the best or extended best constant approximants 𝑓𝜖 of 𝑓Γ𝑝,𝑤 or 𝑓Γ𝑝1,𝑤, 1<𝑝<, respectively. In the last section it is shown that the extended best constant approximant operator assumes a unique constant value for any function 𝑓Γ𝑝1,𝑤, 1<𝑝<.

1. Introduction

The present paper is devoted to investigation of maximal inequalities and Lebesgue’s type differentiation theorems for best local approximations in r.i. quasi-Banach spaces and Lorentz spaces Γ𝑝,𝑤 for 0<𝑝<. In 1910, Henry Lebesgue has proved one of the most famous differentiation theorem, which establishes a convergence of an integral average of any locally integrable function 𝑓 on the ball 𝐵(𝑣,𝜖)𝑛 to this function 𝑓, that is, for a.a. 𝑣𝑛, 1𝜇(𝐵(𝑣,𝜖))𝐵(𝑣,𝜖)𝑓(𝑡)𝑑𝑡𝑓(𝑣)as𝜖0.(1.1) In fact, Lebesgue’s integral average coincides with a best constant approximant on the space 𝐿2(𝑛) [1]. The Lebesgue Differentiation Theorem (LDT) can be proved as a consequence of the weak maximal inequality 𝑡𝑀𝐻𝑓(𝑡)4𝑛𝑓𝐿1(1.2) for the Hardy-Littlewood maximal function 𝑀𝐻𝑓 where 𝑡𝑛 and 𝑓𝐿1(𝑛) [2]. The interesting exploration of LDT was initiated by Stein in [3], who introduced the maximal functions on 𝐿𝑝(𝑛), associated with integral average, and applied it to obtain differentiation theorem in the notation of the norm in 𝐿𝑝(𝑛) for 1𝑝<. In the spirit of this idea many authors developed new techniques of recovering functions in quasi-Banach function spaces. The first results in this subject were obtained by Bastero et al. [4] in 1999, who have investigated Hardy-Littlewood maximal functions and weak maximal inequalities in rearrangement invariant quasi-Banach function spaces. The next paper was published by Mazzone and Cuenya in 2001 [1], about generalizations of the classical Lebesgue differentiation theorem for the best local approximation by constants over balls in 𝐿𝑝+𝐿 for 0𝑝<. They evaluated maximal inequalities for the maximal function related to best constant approximants and proved convergence theorem for best constant approximants. In 2008 [5] Levis et al. extended the best constant approximant operator from Orlicz-Lorentz spaces Λ𝑤,𝜙 to the spaces Λ𝑤,𝜙 and showed monotonicity of the extended operator. In view of this result, in 2009 [6] Levis established maximal inequalities for the maximal function associated with the best constant approximation and proved Lebesgue’s type differentiation theorem for best constant approximants and for integral averages expressed in terms of the modular corresponding to these spaces. Recently, the authors have characterized properties of an expansion of the best constant approximant operator from Lorentz spaces Γ𝑝,𝑤 to the spaces Γ𝑝1,𝑤. The present paper is a continuation of the previous results and devoted to investigation of maximal inequalities and Lebesgue’s type differentiation theorems for local approximation in r.i. quasi-Banach space 𝐸 and in particular in Γ𝑝,𝑤.

The paper consists of three sections and is organized as follows.

In the preliminaries, Section 2, we establish some basic notations and definitions and also recall some auxiliary results, which will be used later.

Sections 3 and 4 consist of the main results of the paper.

We start Section 3 proving measurability of the maximal function 𝑀𝐸(𝑟)(𝑓) for 𝑓𝐸, that corresponds to the quasi-norm average of 𝑓, in r.i. quasi-Banach function spaces 𝐸. Next we establish two types of generalization of LDT in r.i. quasi-Banach function spaces 𝐸 and in Γ𝑝,𝑤. In both types of LDT we employ the assumption of upper and lower 𝜙-estimates of Γ𝑝,𝑤. The first main result, in the spirit of Stein [3], has been proved for any order continuous r.i. quasi-Banach function space. The statement is expressed in terms of quasi-norm averages. In order to show it we first prove the inequality for maximal function 𝑀𝐸(𝑟)𝑓, which corresponds to a quasi-norm average of 𝑓𝐸. In the same spirit we also provide some conditions when the LDT does not hold in 𝐸 or in Γ𝑝,𝑤. Next we continue our discussion with another type of LDT. In order to complete the second main result in this section we characterize conditions for which Lorentz space Γ𝑝,𝑤 satisfies a lower (resp., an upper) 𝜙-estimate, where 𝜙 is the fundamental function of Γ𝑝,𝑤. In view of this characterization we investigate pointwise convergence of the best constant approximants 𝑓𝜖 to 𝑓 as 𝜖0 whenever 𝑓Γ𝑝,𝑤 and 1𝑝<, as well as the convergence of the extended best constant approximants 𝑓𝜖 for any 𝑓Γ𝑝1,𝑤 and 1<𝑝<2. We also present examples showing that this assumption is fulfilled by a large class of the spaces Γ𝑞,𝑝. Finally, we investigate relations between maximal functions and the 𝐾-functional of Banach couple (Γ𝑞,𝑝,𝐿) in the spirit of the inequalities stated in [4]. We finish Section 3 with an example showing that 𝑀Γ(𝑟)𝑝,𝑤(𝑓) is not equivalent to the 𝐾-functional of the pair (Γ𝑝,𝑤,𝐿).

It is well known that the extension of the best constant approximant operator from Γ1,𝑤 to 𝐿0, or from 𝐿1 to 𝐿0, is a set valued function [1, 7]. Contrary to this in Theorem 4.5 we prove that the extended best constant approximant operator assumes a unique value for any 𝑓Γ𝑝1,𝑤 and 1<𝑝<. To show the uniqueness we need to consider strict monotonicity of the right-hand Gâteaux derivative of the norm in Γ𝑝,𝑤 at (𝑓𝑢)𝜒𝐴 in the direction 𝜒𝐴 for any 𝑓Γ𝑝1,𝑤 and 𝑢.

2. Preliminaries

Let and be the set of real and natural numbers, respectively. For any 𝐴[0,𝛼) denote 𝐴𝑐=[0,𝛼)𝐴. Let 0<𝛼 and 𝜇 be the Lebesgue measure on . We denote by 𝐿0 the space of all extended real-valued 𝜇-measurable and finite functions a.e. on [0,𝛼). Denote the outer measure on by 𝜇, the support of 𝑓𝐿0 by 𝕊(𝑓)=supp(𝑓), and the restriction of 𝑓 to the set 𝐴[0,𝛼) by 𝑓𝐴. By a simple (resp., step) function we mean a measurable function 𝑓𝐿0 with a finite measure support, which attains a finite number of values (resp., a finite number of values on a finite number of disjoint intervals). The distribution function 𝑑𝑓 of a function 𝑓𝐿0 is given by 𝑑𝑓(𝜆)=𝜇(𝑠[0,𝛼)|𝑓|(𝑠)>𝜆) for all 𝜆0. Two functions 𝑓,𝑔𝐿0 are called equimeasurable, if 𝑑𝑓(𝜆)=𝑑𝑔(𝜆) for all 𝜆0. We define the decreasing rearrangement for any 𝑓𝐿0 by 𝑓(𝑡)=inf{𝑠>0𝑑𝑓(𝑠)𝑡}, 𝑡0. For given 𝑓𝐿0 we denote the maximal function of 𝑓 by 𝑓(𝑡)=(1/𝑡)𝑡0𝑓(𝑠)𝑑𝑠. It is well known that 𝑓𝑓 and 𝑓 is decreasing and subadditive, that is, (𝑓+𝑔)𝑓+𝑔 for any 𝑓,𝑔𝐿0. For the properties of 𝑑𝑓, 𝑓, and 𝑓 see [2, 8]. A subspace 𝐸𝐿0 equipped with a quasinorm 𝐸 is called a quasinormed function space, if the following conditions are satisfied.(1)If 𝑓𝐿0, 𝑔𝐸, and |𝑓||𝑔| a.e., then 𝑓𝐸 and 𝑓𝐸𝑔𝐸.(2)There exists a strictly positive 𝑓𝐸.

If 𝐸 is complete, then it is said to be a quasi-Banach function space. We say that a quasi-Banach function space 𝐸 is rearrangement invariant (r.i. for short), if whenever 𝑓𝐿0 and 𝑔𝐸 with 𝑑𝑓=𝑑𝑔, then 𝑓𝐸 and 𝑓𝐸=𝑔𝐸 (see [2]). Throughout the paper we use the notation 𝐴𝐵, which means that the expressions 𝐴 and 𝐵 are equivalent; that is, 𝐴/𝐵 is bounded from both sides. Let 0<𝑝< and 𝑤𝐿0 be a nonnegative weight function. Lorentz space Γ𝑝,𝑤 is a subspace of 𝐿0 such that𝑓=𝑓Γ𝑝,𝑤=𝛼0𝑓𝑝𝑤1/𝑝=𝛼0𝑓𝑝(𝑡)𝑤(𝑡)𝑑𝑡1/𝑝<.(2.1) Given a measurable set 𝐴[0,𝛼) by Γ𝑝,𝑤(𝐴) we denote the set of 𝑓𝐿0 restricted to 𝐴 and satisfying the above inequality. Unless we say otherwise, throughout the paper we assume that 𝑤 belongs to the class 𝐷𝑝 (in short 𝑤𝐷𝑝), whenever it satisfies the following conditions:𝑊(𝑠)=𝑠0𝑤<,𝑊𝑝(𝑠)=𝑠𝑝𝛼𝑠𝑡𝑝𝑤(𝑡)𝑑𝑡<(2.2) for all 0<𝑠𝛼 if 𝛼< and for all 0<𝑠< otherwise. These two conditions guarantee that Γ𝑝,𝑤{0}. We also assume that𝑊𝑝(𝑠)>0for0<𝑠<𝛼,𝑊()=0𝑤=if𝛼=.(2.3) Under these assumptions (Γ𝑝,𝑤,) is a rearrangement invariant (r.i. for short) quasi-Banach function space such that it has the Fatou property and the order continuous norm. Letting 0<𝑝,𝑞< and 𝑤(𝑡)=𝑡𝑝/𝑞1, 𝑡(0,𝛼), the space Γ𝑝,𝑤 will be denoted by Γ𝑞,𝑝.

Unless we say otherwise, throughout this paper we assume that 𝜙 is the fundamental function of Γ𝑝,𝑤 defined as 𝜙(𝑡)=𝜒(0,𝑡), 𝑡(0,𝛼), and 𝜙(0)=0. It is easy to show that the fundamental function 𝜙 is strictly increasing and continuous on [0,𝛼), lim𝑡𝜙(𝑡)= and lim𝑡𝑓(𝑡)=0 for 𝑓Γ𝑝,𝑤. For more details about the properties of Γ𝑝,𝑤 see [9].

Recall that for given 0<𝑝<, classical Lorentz space Λ𝑝,𝑤 is a subspace of 𝐿0 such that𝑓Λ𝑝,𝑤=𝛼0𝑓𝑝𝑤1/𝑝=𝛼0𝑓𝑝(𝑡)𝑤(𝑡)𝑑𝑡1/𝑝<.(2.4) In case when 𝑊 satisfies the Δ2-condition, that is 𝑊(2𝑠)𝐶𝑊(𝑠) for all 𝑠>0 and some 𝐶>0, as well as 𝑊()=, the space Λ𝑝,𝑤 is a separable r.i. order continuous quasi-Banach function space [9]. The space Λ𝑝,𝑤 is a r.i. Banach function space, whenever the weight 𝑤 is decreasing and 1𝑝< [10]. Since 𝑓𝑓, we have the natural inclusion Γ𝑝,𝑤Λ𝑝,𝑤. Moreover, Γ𝑝,𝑤=Λ𝑝,𝑤 if and only if 𝑤 satisfies condition 𝐵𝑝, (𝑤𝐵𝑝 for short) which means that there is 𝐴>0 such that for all 𝑠>0 we have 𝑊𝑝(𝑠)𝐴𝑊(𝑠) [1113].

Let (Ω1,𝜇1) and (Ω2,𝜇2) be 𝜎-finite measure spaces. A map 𝛾 from Ω1 into Ω2 is said to be a measure-preserving transformation, if whenever 𝐸 is a 𝜇2-measurable subset of Ω2, the set 𝛾1(𝐸)={𝑢Ω1𝛾(𝑢)𝐸} is a 𝜇1-measurable subset of Ω1 and 𝜇1(𝛾1(𝐸))=𝜇2(𝐸). For given subsets 𝐴,𝐵+ such that 𝜇(𝐴)=𝜇(𝐵), there exists a measure-preserving transformation 𝛿𝐴𝐵 [14, Theorem 17, page 410].

Definition 2.1 (see [15]). Let 𝑓,𝐿0. Denote 𝜏(𝑓,)(𝑡)=𝑑𝑓||𝑓||||𝑓||||𝑓||||𝑓||||𝑓||(𝑡)+𝜇𝑢(𝑢)=(𝑡),(𝑢)sign(𝑓(𝑢))>(𝑡)sign(𝑓(𝑡))+𝜇𝑢(𝑢)=(𝑡),(𝑢)sign(𝑓(𝑢))=(𝑡)sign(𝑓(𝑡)),𝑢𝑡(2.5) for all 𝑡[0,𝛼).

In 1970, Ryff proved in [16] that 𝜏(𝑓[0,1],0)[0,1][0,1] is a measure-preserving transformation for any 𝑓𝐿0 and |𝑓|=𝑓𝜏(|𝑓|[0,1],0) a.e. on [0,1]. In 1993, Carothers et al. established in [15] that 𝜏(𝑓,) is a measure preserving transformation from 𝕊(𝑓) onto 𝕊(𝑓) such that |𝑓|=𝑓𝜏(𝑓,) a.e. on 𝕊(𝑓) for any 𝑓𝐿0 with 𝑑𝑓(𝜆)< for all 𝜆>0 and any 𝐿0. Notice that for any 𝑓𝐿0 with 𝑑𝑓(𝜆)< for any 𝜆>0 and 𝐿0, if 𝜇(𝑢|𝑓|(𝑢)=𝑣)=0 for every 𝑣>0 we have that 𝜏(𝑓,)(𝑡)=𝑑𝑓(|𝑓|(𝑡)) and it is the unique measure-preserving transformation up to measure zero satisfying |𝑓|=𝑓𝜏(𝑓,) a.e. on 𝕊(𝑓).

Definition 2.2. Let 𝑓,𝑔Γ𝑝,𝑤 and let 𝜏(𝑓,𝑔), 𝜏(𝑔𝕊(𝑔)𝕊(𝑓),0) be measure-preserving transformations given by Definition 2.1. Denote 𝜌(𝑓,𝑔)(𝜏𝑠)=(𝑓,𝑔)(𝑠)if𝑑𝑠𝕊(𝑓),𝑓(0)+𝜏(𝑔𝕊(𝑔)𝕊(𝑓),0)(𝑠)if𝑠𝕊(𝑔)𝕊(𝑓).(2.6)

Definition 2.3. Let 𝑓Γ𝑝,𝑤 and 𝐴[0,𝛼) with 𝜇(𝐴)<. Denote 𝐾(𝑓,𝐴)(1𝑢,𝑡)=𝑡𝐴12𝜒{𝑓<𝑢}𝜒(0,𝑡)𝜌((𝑓𝑢)𝜒𝐴,𝜒𝐴),𝑆𝑝(𝑓,𝐴)(𝑢)=𝛼0𝐾(𝑓,𝐴)(𝑢,𝑡)(𝑓𝑢)𝜒𝐴(𝑝1)(𝑡)𝑤(𝑡)𝑑𝑡(2.7) for any 𝑢 and 𝑡(0,𝛼).

Let (𝑋,𝑋) be a real normed space. Denote by 𝐵𝑋 (resp., 𝑆𝑋) the closed unit ball (resp., the unit sphere) of 𝑋. Assume that 𝑌 is a subset of 𝑋 and 𝑥 is an element of 𝑋. An element ̃𝑥𝑌 is called best approximant to 𝑥 from 𝑌 if𝑥̃𝑥𝑋=inf𝑦𝑌𝑥𝑦𝑋.(2.8) A nonempty subset 𝑌 of 𝑋 is a set of uniqueness if for any element 𝑥𝑋 there is no more than one element ̃𝑥𝑌 satisfying (2.8). The set 𝑌 is a set of existence if for every element 𝑥𝑋 there is at least one element ̃𝑥𝑌 for which condition (2.8) holds. The set 𝑌 is a Chebyshev set if for every element 𝑥𝑋 there exists exactly one element ̃𝑥𝑌 satisfying (2.8), that is, if 𝑌 is both a set of uniqueness and a set of existence (for more details see [17]). Let 2𝑌 be a collection of all subsets of 𝑌. A set value map 𝑇𝑌𝑋2𝑌 is said to be best approximant operator, if it assumes for any 𝑥𝑋 a set of all best approximant elements to 𝑥 from 𝑌, that is,𝑇𝑌(𝑥)=̃𝑥𝑌𝑥̃𝑥𝑋=inf𝑦𝑌𝑥𝑦𝑋=̃𝑥𝑌𝑥̃𝑥𝑋𝑥𝑦𝑋.,𝑦𝑌(2.9) In case when 𝑋 is a norm function space and 𝑌 is a family of constant functions, then 𝑇𝑌 is called best constant approximant operator, and each element ̃𝑥𝑇𝑌(𝑥) is called best constant approximant to 𝑥𝑋 from 𝑌. Let 𝐴[0,𝛼) with 0<𝜇(𝐴)< and 𝕂(𝐴)={𝑐𝜒𝐴𝑐}. It is well known that the set 𝑇𝕂(𝐴)(𝑓) is convex, compact, and a set of existence for all 𝑓Γ𝑝,𝑤 [18, 19]. Let’s recall some characterizations of best constant approximants over Lorentz spaces Γ𝑝,𝑤.

Theorem 2.4 (see [20, Theorem 7.5]). Let 1𝑝< and let 𝑓Γ𝑝,𝑤𝕂(𝐴). Then 𝑢𝕂(𝐴) is the best constant approximant of 𝑓 if and only if 𝑆𝑝(𝑓,𝐴)(𝑢)0,𝑆𝑝(𝑓,𝐴)(𝑢)0.(2.10)

Corollary 2.5 (see [20, Corollary 7.3]). For any 1<𝑝< and 𝑤 positive, we have that 𝑇𝕂(𝐴)(𝑓) is Chebyshev set for any 𝑓Γ𝑝,𝑤; that is, there is an unique best constant approximant 𝑢𝕂(𝐴) to 𝑓.

Recently, in [7] it has been developed an existence of extension of the best constant approximant operators from Lorentz space Γ1,𝑤 to 𝐿0, if 𝑤𝐷1, and from Γ𝑝,𝑤 to Γ𝑝1,𝑤, if 1<𝑝< and 𝑤𝐷𝑝1. Now we recall definition of the extended operator 𝑇(𝑝,𝐴) on 𝐿0, if 𝑝=1, and on Γ𝑝1,𝑤, if 1<𝑝<.

Definition 2.6 (see [7]). Let 𝐴(0,𝛼) with 0<𝜇(𝐴)< and let 𝑤𝐷𝑝1 if 1<𝑝<, and 𝑤𝐷1 if 𝑝=1. Assume that 𝑓Γ𝑝1,𝑤 if 𝑝>1, and 𝑓𝐿0 if 𝑝=1. Denote 𝑓(𝑝,𝐴)=min𝑢𝑆𝑝(𝑓,𝐴),(𝑢)0𝑓(𝑝,𝐴)=max𝑢𝑆𝑝(𝑓,𝐴).(𝑢)0(2.11) Then the extended best constant approximant operator is given by 𝑇(𝑝,𝐴)𝑓(𝑓)=(𝑝,𝐴)𝜒𝐴,𝑓(𝑝,𝐴)𝜒𝐴.(2.12)

In fact, any 𝑢𝑇(𝑝,𝐴)(𝑓) is called an extended best constant approximant of 𝑓. Notice that in view of Theorem 2.4, if 𝑓Γ𝑝,𝑤 for 1𝑝<, then any 𝑢𝑇(𝑝,𝐴)(𝑓) is a classical best constant approximant of 𝑓.

Definition 2.7 (see [4]). Let 0<𝑟<, (𝐸,𝐸) be a r.i. quasi-Banach function space, and let Φ[0,)[0,) be an increasing bijection. 𝐸 is said to satisfy an upper (resp., a lower) Φ-estimate for 𝑟𝐸, if there exists 𝐶>0 such that for all 𝑛 and (𝑓𝑖)𝑛𝑖=1𝐸 with pairwise disjoint supports we have 𝑛𝑖=1𝑓𝑖𝑟𝐸𝐶Φ𝑛𝑖=1Φ1𝑓𝑖𝑟𝐸,(2.13) respectively Φ𝑛𝑖=1Φ1𝑓𝑖𝑟𝐸𝐶𝑛𝑖=1𝑓𝑖𝑟𝐸.(2.14)

In the case when Φ(𝑡)=𝑡1/𝑝 and 𝑟=1, this definition covers the notions of the upper (resp., lower) 𝑝-estimate [21].

Let 0<𝑝< and 𝑋 be a quasi-Banach function space. We denote by 𝑋(𝑝)={𝑓|𝑓|𝑝𝑋} the 𝑝-convexification of 𝑋 equipped with the quasinorm 𝑋(𝑝)=||𝑝𝑋1/𝑝. Now we recall the definition of the maximal function for any r.i. quasi-Banach function space, that plays a crucial role in process of generalization of Lebesgue’s Differentiation Theorem in Γ𝑝,𝑤.

Definition 2.8. Let 0<𝑟<, (𝐸,𝐸) be a r.i. quasi-Banach function space. For any 𝑓𝐸 we denote 𝑓𝜖(𝑟)(𝑡)=𝑓𝜒𝐵(𝑡,𝜖)𝑟𝐸𝜒𝐵(𝑡,𝜖)𝐸(2.15) for all 𝜖>0 and 𝑡(0,𝛼). The maximal function 𝑀𝐸(𝑟)𝑓(0,𝛼) is given by 𝑀𝐸(𝑟)𝑓𝑓(𝑡)=sup𝜖(𝑟)(𝑡)𝜖>0,𝐵(𝑡,𝜖)(0,𝛼)(2.16) for any 𝑡(0,𝛼), where 𝐵(𝑡,𝜖)=(𝑡𝜖,𝑡+𝜖).

We finish the preliminaries with the following proposition needed further. It is a generalization of the well-known result, which in particular can be found in [22] for special case when 𝜑(𝑡)=𝑊(𝑡) on (0,). The proof of the proposition is quite standard and is provided for the sake of completeness.

Proposition 2.9. Let 𝛼=, 𝑓𝐿0, and let 𝜑[0,)[0,) be an increasing continuous function. If lim𝑡𝜑(𝑡)= or 𝑑𝑓(𝜆)< for all 𝜆>0, then sup𝑡>0𝜑(𝑡)𝑓(𝑡)=sup𝑡>0𝑑𝑡𝜑𝑓.(𝑡)(2.17)

Proof. Notice first if lim𝑡𝜑(𝑡)= and there exists 𝑡>0 such that 𝑑𝑓(𝑡)=, then lim𝑠𝜑(𝑠)𝑓(𝑠)=, and the conclusion follows.
Now assume that 𝑓=𝑛𝑖=1𝑎𝑖𝜒𝐸𝑖 is a nonnegative simple function, where 𝑎𝑖>𝑎𝑖+1 for any 1𝑖𝑛1, 𝑎𝑛+1=0, and 𝐸𝑖𝐸𝑗= whenever 𝑖𝑗. Then 𝑓(𝑡)=𝑛𝑖=1𝑎𝑖𝜒[𝑚𝑖1,𝑚𝑖)(𝑡) for any 𝑡>0, where 𝑚0=0 and 𝑚𝑖=𝑖𝑗=1𝜇(𝐸𝑗) for 1𝑖𝑛. We claim that sup𝑡>0𝜑(𝑡)𝑓(𝑡)=max1𝑖𝑛𝑎𝑖𝜑𝑚𝑖.(2.18) By monotonicity and continuity of 𝜑 we obtain sup𝑡>0{𝜑(𝑡)𝑓(𝑡)}max1𝑖𝑛{𝑎𝑖𝜑(𝑚𝑖)}. On the other hand we have sup𝑡>0{𝜑(𝑡)𝑓(𝑡)}sup𝑡>0{𝜑(𝑡)𝑎𝑖𝜒[𝑚𝑖1,𝑚𝑖)(𝑡)}=𝑎𝑖𝜑(𝑚𝑖). Therefore sup𝑡>0{𝜑(𝑡)𝑓(𝑡)}max1𝑖𝑛{𝑎𝑖𝜑(𝑚𝑖)}, which implies condition (2.18). Now we will show sup𝑡>0𝑑𝑡𝜑𝑓(𝑡)=max1𝑖𝑛𝑎𝑖𝜑𝑚𝑖.(2.19) For 𝑡>0 we have 𝑡𝜑(𝑑𝑓(𝑡))=𝑡𝑛𝑖=1𝜑(𝑚𝑖)𝜒[𝑎𝑖+1,𝑎𝑖)(𝑡)max1𝑖𝑛{𝑎𝑖𝜑(𝑚𝑖)}, and so sup𝑡>0{𝑡𝜑(𝑑𝑓(𝑡))}max1𝑖𝑛{𝑎𝑖𝜑(𝑚𝑖)}. On the other hand, sup𝑡>0{𝑡𝜑(𝑑𝑓(𝑡))}𝑎𝑖𝜑(𝑚𝑖) for every 1𝑖𝑛, and consequently, sup𝑡>0{𝑡𝜑(𝑑𝑓(𝑡))}max1𝑖𝑛{𝑎𝑖𝜑(𝑚𝑖)}, which provides (2.19). Both (2.18) and (2.19) show (2.17) for any nonnegative simple function 𝑓.
Now suppose that 𝑓 is a measurable function and 𝑑𝑓(𝑡)< for any 𝑡>0. Then, by standard arguments of existing a sequence of nonnegative simple functions (𝑓𝑛) such that 𝑑𝑓𝑛𝑑𝑓 and 𝑓𝑛𝑓 as 𝑛 we can show that sup𝑡>0𝜑(𝑡)𝑓(𝑡)=lim𝑛sup𝑡>0𝜑(𝑡)𝑓𝑛(𝑡)=lim𝑛sup𝑡>0𝑑𝑡𝜑𝑓𝑛(𝑡)=sup𝑡>0𝑑𝑡𝜑𝑓(𝑡)(2.20) and conclude the proof.

In fact, Proposition 2.9 describes the largest family of increasing and continuous functions 𝜑, for which (2.17) is satisfied. Indeed, let 𝛼=, 𝑓1, and 𝜑(𝑡)=2(1/(1+𝑡)) for any 𝑡[0,). Then 𝑓𝑓, 𝑑𝑓(𝑡)=0 for all 𝑡1 and 𝑑𝑓(𝑡)= for any 𝑡[0,1). Clearly, 𝜑 is increasing and continuous and also lim𝑡𝜑(𝑡)=2. Therefore, sup𝑡>0{𝜑(𝑡)𝑓(𝑡)}=2 and sup𝑡>0{𝑡𝜑(𝑑𝑓(𝑡))}=, which implies that condition (2.17) does not hold.

3. Lebesgue’s Differentiation Theorems

The intention of this section is to establish generalizations of LDT in r.i. quasi-Banach function spaces 𝐸 in terms of the formulas expressed by quasinorm averages. We also focus on convergence of the best and the extended best constant approximant of 𝑓𝐸 to 𝑓, which is another type of LDT. First we introduce the notion of the differentiation property for a quasi-Banach function space 𝐸.

Definition 3.1. Let (𝐸,𝐸) be a quasi-Banach function space on [0,𝛼). We say that 𝐸 has the Lebesgue differentiation property (LDP), whenever for any 𝑓𝐸 and for a.a. 𝑡(0,𝛼) we have lim𝜖0(𝑓𝑓(𝑡))𝜒𝐵(𝑡,𝜖)𝐸𝜒𝐵(𝑡,𝜖)𝐸=0.(3.1)

Observe that letting (𝐸,𝐸) be a quasi-Banach function space on [0,𝛼) with LDP, by the Aoki-Rolewicz theorem [23] there exist 0<𝑟1 and an equivalent 𝑟-norm ||𝐸 to 𝐸 such that for any 𝑓𝐸 and for a.a. 𝑡(0,𝛼) we getlim𝜖0||𝑓𝜒𝐵(𝑡,𝜖)||𝐸||𝜒𝐵(𝑡,𝜖)||𝐸=||||.𝑓(𝑡)(3.2) If (𝐸,𝐸) is a normed space, then the quasinorm ||𝐸 can be replaced by 𝐸.

In the next proposition we establish measurability of the maximal function 𝑀𝐸(𝑟)𝑓.

Proposition 3.2. Let 0<𝑟< and let 𝐸 be a r.i. order continuous quasi-Banach function space. If 𝑓𝐸, then the maximal function 𝑀𝐸(𝑟)𝑓 is measurable on (0,𝛼).

Proof. Let 𝜖>0 and 𝑡(0,𝛼). We first observe that 𝑓𝜖(𝑟)(𝑡) is continuous on (0,𝛼). In fact, for any 𝑡𝑛𝑡, lim𝑛𝜒𝐵(𝑡𝑛,𝜖)=𝜒𝐵(𝑡,𝜖), and by order continuity of 𝐸 we obtain that lim𝑛𝑓𝜒𝐵(𝑡𝑛,𝜖)𝑟𝐸=𝑓𝜒𝐵(𝑡,𝜖)𝑟𝐸 and lim𝑛𝜒𝐵(𝑡𝑛,𝜖)𝐸=𝜒𝐵(𝑡,𝜖)𝐸. Now by Fatou’s property of 𝐸 we have lim𝛿𝜖𝑓𝜒𝐵(𝑡,𝛿)𝐸=𝑓𝜒𝐵(𝑡,𝜖)𝐸,lim𝛿𝜖𝜒𝐵(𝑡,𝛿)𝐸=𝜒𝐵(𝑡,𝜖)𝐸.(3.3) Hence we have 𝑀𝐸(𝑟)𝑓𝑓(𝑡)=sup𝜖(𝑟),(𝑡)𝜖>0,𝜖,𝐵(𝑡,𝜖)(0,𝛼)(3.4) and thus 𝑀𝐸(𝑟)𝑓 is measurable.

Remark 3.3. If 𝑊 satisfies Δ2 condition, then we obtain measurability of the maximal function 𝑀Λ(𝑟)𝑝,𝑤𝑓 for any 𝑓Λ𝑝,𝑤, analogously as in case of the maximal function 𝑀Γ(𝑟)𝑝,𝑤𝑔 for any 𝑔Γ𝑝,𝑤 when 𝛼= and 𝑊()=.

In view of Theorem  1 in [4], we investigate the so-called weak inequality for the maximal function 𝑀𝐸(𝑟) whenever 𝐸 is a r.i. quasi-Banach function space.

Theorem 3.4. Let 0<𝑟<. If a r.i. quasi-Banach function space 𝐸 satisfies a lower 𝜙-estimate for 𝑟𝐸, then there exists 𝐶>0 such that for all 𝑓𝐸 and 𝜆>0 we have 𝑑𝜆𝜙𝑀𝐸(𝑟)𝑓(𝜆)𝐶𝑓𝑟𝐸.(3.5)

Proof. Assume that 𝜆>0. Denote Ω𝜆=𝑡(0,𝛼)𝑀𝐸(𝑟).𝑓(𝑡)>𝜆(3.6) Clearly, by Proposition 3.2 we get that Ω𝜆 is measurable for all 𝜆>0. Letting 𝑡Ω𝜆, there exists 𝜖𝑡>0 such that 𝐵(𝑡,𝜖𝑡)(0,𝛼) and 𝜆<𝑓𝜖(𝑟)𝑡(𝑡).(3.7) Let 𝑐<𝜇(Ω𝜆) and denote 𝐵=𝑡Ω𝜆𝐵(𝑡,𝜖𝑡). Since Ω𝜆𝐵, we get 𝑐<𝜇(𝐵). Hence, by regularity of the Lebesgue measure 𝜇 there is a compact set 𝐾𝐵 such that 𝑐<𝜇(𝐾). By the fact that a collection 𝐷={𝐵(𝑡,𝜖𝑡)𝑡Ω𝜆} is an open covering of the set 𝐾 and by the Vitali covering lemma [2, Lemma 3.2], there exists a pairwise disjoint finite collection {𝐵(𝑡𝑘,𝜖𝑡𝑘)1𝑘𝑛}𝐷 such that 𝜇(𝐾)4𝑛𝑘=1𝜇(𝐵(𝑡𝑘,𝜖𝑡𝑘)). Therefore, by 𝜙(4𝑠)4𝜙(𝑠) and by (3.7) we get 𝜙(𝑐)𝜙(𝜇(𝐾))4𝜙𝑛𝑘=1𝜇𝐵𝑡𝑘,𝜖𝑡𝑘4𝜙𝑛𝑘=1𝜙1𝜆1𝑓𝜒𝐵(𝑡𝑘,𝜖𝑡𝑘)𝑟𝐸.(3.8) Hence, by assumption that 𝐸 satisfies a lower 𝜙-estimate for 𝑟𝐸, there is 𝐶>0 such that for any 1𝑘𝑛 we get 𝜙(𝑐)<4𝜙𝑛𝑘=1𝜙1𝜆1𝑓𝜒𝐵(𝑡𝑘,𝜖𝑡𝑘)𝑟𝐸4𝐶𝜆𝑛𝑘=1𝑓𝜒𝐵(𝑡𝑘,𝜖𝑡𝑘)𝑟𝐸4𝐶𝜆𝑓𝑟𝐸.(3.9) Since 𝑐<𝜇(Ω𝜆) is arbitrary, we obtain 𝜙(𝜇(Ω𝜆))(4𝐶/𝜆)𝑓𝑟𝐸, which finishes the proof.

In the next theorem we present Lebesgue’s differentiation property in the space 𝐸.

Theorem 3.5. Let 𝛼= and let 𝐸 be a r.i. order continuous quasi-Banach function space 𝐸. If 𝐸 satisfies a lower 𝜙-estimate for 𝐸, then 𝐸 has LDP, that is, lim𝜖0(𝑓𝑓(𝑡))𝜒𝐵(𝑡,𝜖)𝐸𝜒𝐵(𝑡,𝜖)𝐸=0(3.10) for all 𝑓𝐸 and for a.a. 𝑡(0,). If in addition 𝐸 is normable, then lim𝜖0𝑓𝜒𝐵(𝑡,𝜖)𝐸𝜒𝐵(𝑡,𝜖)𝐸=||||.𝑓(𝑡)(3.11)

Proof. Observe first that the set of step functions with supports of finite measure is dense in 𝐸. The proof of this observation is standard, by density of the simple functions, which is equivalent to order continuity of 𝐸 and regularity of the Lebesgue measure 𝜇 on (cf. [8]).
Define an operator 𝐿(0,) by 𝐿(𝑡)=limsup𝜖0𝜖(1)(𝑡)=limsup𝜖0𝜒𝐵(𝑡,𝜖)𝐸𝜒𝐵(𝑡,𝜖)𝐸(3.12) for any 𝐸 and 𝑡(0,). Assume that 𝜖>0 and 𝐵(𝑡,𝜖)(0,) and also 𝑓𝐸, 𝑐. Let 𝑔=𝑏𝜒𝐴 be a characteristic function of an open interval 𝐴. Notice that for a.a. 𝑡(0,) there exist 𝛿𝑡>0 such that for all 0<𝜖<𝛿𝑡 we have either 𝐵(𝑡,𝜖)𝐴 or 𝐵(𝑡,𝜖)𝐴𝑐 and consequently ((𝑔𝑐)𝜒𝐵(𝑡,𝜖))=|𝑔(𝑡)𝑐|𝜒(0,2𝜖). Therefore, (𝑔𝑐)𝜖(1)(𝑡)=(𝑔𝑐)𝜒𝐵(𝑡,𝜖)𝐸𝜒𝐵(𝑡,𝜖)𝐸=(𝑔(𝑡)𝑐)𝜒(0,2𝜖)𝐸𝜒(0,2𝜖)𝐸=||||𝑔(𝑡)𝑐(3.13) for a.a. 𝑡(0,) and for any 0<𝜖<𝛿𝑡. Hence 𝐿(𝑔𝑐)(𝑡)=limsup𝜖0(𝑔𝑐)𝜖(1)=||||(𝑡)𝑔(𝑡)𝑐(3.14) for a.a. 𝑡(0,). Observe that the above equation can be proved analogously for any step function 𝑔 with support of finite measure. Let 𝜉 be a constant in the triangle inequality of the quasinorm 𝐸. Thus 𝐿𝐿||𝑔||(𝑓𝑐)(𝑡)𝜉(𝐿(𝑓𝑔)(𝑡)+𝐿(𝑔𝑐)(𝑡))=𝜉(𝑓𝑔)(𝑡)+(𝑡)𝑐(3.15) for a.a. 𝑡(0,). Clearly 𝐿(𝑓𝑔)(𝑡)𝑀𝐸(1)(𝑓𝑔)(𝑡), whence 𝑀𝐿(𝑓𝑐)(𝑡)𝜉𝐸(1)||||(𝑓𝑔)(𝑡)+𝑔(𝑡)𝑐(3.16) for a.a. 𝑡(0,). Now replacing 𝑐 by 𝑓(𝑡) we get 𝑀𝐿(𝑓𝑓(𝑡))(𝑡)𝜉𝐸(1)||||(𝑓𝑔)(𝑡)+𝑓(𝑡)𝑔(𝑡)(3.17) for a.a. 𝑡(0,). Define Ω𝑠={𝑡(0,)𝐿(𝑓𝑓(𝑡))(𝑡)>𝑠}(3.18) for any 𝑠>0. By Proposition 3.2, we have that 𝑀𝐸(1)(𝑓𝑔) is 𝜇-measurable. Recall [2, 8] that 𝑑𝑓1+𝑓2(𝑠)𝑑𝑓1(𝑠/2)+𝑑𝑓2(𝑠/2) for any 𝑓1,𝑓2𝐿0. Thus in view of (3.17) we obtain for 𝑠>0, 𝜇Ω𝑠𝜇𝑡(0,)𝑀𝐸(1)||||𝑠(𝑓𝑔)(𝑡)+𝑓𝑔(𝑡)>𝜉𝑑𝑀𝐸(1)(𝑓𝑔)𝑠2𝜉+𝑑𝑓𝑔𝑠.2𝜉(3.19) Now since 𝜙 satisfies the triangle inequality with constant 𝜉, we get 𝜙𝜇Ω𝑠𝑑𝜉𝜙𝑀𝐸(1)(𝑓𝑔)𝑠𝑑2𝜉+𝜉𝜙𝑓𝑔𝑠2𝜉(3.20) for every 𝑠>0. Observe that for any 𝐸, 𝑡(0,𝛼), we have (𝑡)𝜙(𝑡)𝜒(0,𝑡)𝐸𝐸. Thus, by Proposition 2.9 we have 𝜙𝑑𝑓𝑔𝑠2𝜉2𝜉𝑠sup𝑡>0𝜙(𝑡)(𝑓𝑔)(𝑡)2𝜉𝑠𝑓𝑔𝐸.(3.21) Furthermore, by Theorem 3.4 there exists 𝐶>0 such that 𝜙𝑑𝑀𝐸(1)(𝑓𝑔)𝑠2𝜉2𝜉𝐶𝑠𝑓𝑔𝐸.(3.22) Therefore, for any step function 𝑔 and for all 𝑠>0, 𝜙𝜇Ω𝑠2𝜉2(𝐶+1)𝑠𝑓𝑔𝐸.(3.23) Hence we have 𝜙𝜇Ω𝑠𝜇Ω=𝜙𝑠=0(3.24) for all 𝑠>0. So 𝐿(𝑓𝑓(𝑡))(𝑡)=0 for a.a. 𝑡(0,), which shows the first formula.
The second formula results from the first one since 𝐸 is a norm in 𝐸.

Now we characterize the lower and upper 𝜙-estimate of Γ𝑞,𝑝 on (0,), for 0<𝑝,𝑞<. Clearly in this case 𝑤(𝑡)=𝑡𝑝/𝑞1 satisfies 𝐵𝑝 condition. Thus Λ𝑝,𝑤=Γ𝑝,𝑤 and Λ𝑝,𝑤Γ𝑝,𝑤. Hence by Theorems 3 and  7 in [9] and by Hölder’s inequality we obtain the following corollary.

Corollary 3.6. Let 𝛼=, 0<𝑝,𝑟< and, 1<𝑞<.(i)If 𝑝𝑞, then Γ𝑞,𝑝 satisfies a lower 𝜙-estimate for the functional 𝑟Γ𝑞,𝑝, where 1𝑟, and an upper 𝜙-estimate for 𝑟Γ𝑞,𝑝, where 𝑟𝑝/𝑞.(ii)If 𝑝𝑞, then Γ𝑞,𝑝 satisfies an upper 𝜙-estimate for the functional 𝑟Γ𝑞,𝑝, where 𝑟1, and a lower 𝜙-estimate for the functional 𝑟Γ𝑞,𝑝, where 𝑟𝑝.

The immediate consequence of Theorem 3.5 is the next result, which describes under what conditions Γ𝑝,𝑤 has LDP. The second part follows from the previous corollary.

Proposition 3.7. Let 0<𝑝< and 𝛼=. If the Lorentz space Γ𝑝,𝑤 satisfies a lower 𝜙-estimate for , then Γ𝑝,𝑤 has LDP. Consequently, for 0<𝑝𝑞< and 1<𝑞, Lorentz spaces Γ𝑞,𝑝 on (0,) have LDP.

We will show next which spaces (𝐸,𝐸) do not have LDP with respect to 𝑟𝐸. Notice that the ratio 𝑓𝑟1𝐸/𝑔𝑟2𝐸 can be reduced easily to the quotient 𝑓𝑟𝐸/𝑔𝐸 for any 𝑓,𝑔𝐸, where 𝑔0. In view of this fact, we state the following theorem.

Theorem 3.8. Let 𝐸 be a r.i. quasi-Banach function space. (i) Let 1<𝑟<. If order continuous 𝐸 satisfies a lower 𝜙-estimate for 𝑟𝐸, then for any 𝑓𝐸 and for a.a. 𝑡(0,𝛼) we have lim𝜖0𝑓𝜒𝐵(𝑡,𝜖)𝑟𝐸𝜒𝐵(𝑡,𝜖)𝐸=0.(3.25)(ii) Let 0<𝑟<1. Then for any 𝑓𝐸 and for a.a. 𝑡𝕊(𝑓) we have lim𝜖0𝑓𝜒𝐵(𝑡,𝜖)𝑟𝐸𝜒𝐵(𝑡,𝜖)𝐸=.(3.26)

Proof. Let 𝐿(0,𝛼) be the operator on 𝐸 given by 𝐿𝑓(𝑡)=limsup𝜖0{𝑓𝜖(𝑟)(𝑡)}, 𝑓𝐸.
(i) Let 𝑔 be a step function with a finite measure support. Notice that for a.a. 𝑡(0,𝛼) and for small enough 𝜖>0 we have 𝑔𝜖(𝑟)(𝑡)=𝑔𝜒𝐵(𝑡,𝜖)𝑟𝐸𝜒𝐵(𝑡,𝜖)𝐸=|𝑔(𝑡)|𝑟𝜒𝐵(𝑡,𝜖)𝐸𝑟1.(3.27) Therefore 𝐿𝑔(𝑡)=limsup𝜖0{𝑔𝜖(𝑟)(𝑡)}=0 and thus 𝐿𝑓(𝑡)𝜉(𝐿(𝑓𝑔)(𝑡)+𝐿𝑔(𝑡))=𝜉𝐿(𝑓𝑔)(𝑡)𝜉𝑀𝐸(𝑟)(𝑓𝑔)(𝑡)(3.28) for a.a. 𝑡(0,𝛼). Denoting Ω𝑠={𝑡(0,𝛼)𝐿𝑓(𝑡)>𝑠}, 𝑠>0, by Proposition 3.2, 𝑀𝐸(𝑟)(𝑓𝑔) is 𝜇-measurable. Now by Theorem 3.4 there exists 𝐶>0 such that for all 𝑠>0, 𝜙𝜇Ω𝑠𝑑𝜙𝑀𝐸(𝑟)(𝑓𝑔)𝑠𝜉𝜉𝐶𝑠𝑓𝑔𝑟𝐸,(3.29) which shows that 𝜇(Ω𝑠)=0 and thus 𝐿𝑓(𝑡)=0 a.e.
(ii) Let 𝑓𝑛|𝑓| a.e., where 0𝑓𝑛 are step functions. Therefore, for a.a. 𝑡𝕊(𝑓) there is 𝑛 such that 0<𝑓𝑛(𝑡)|𝑓(𝑡)| and for small enough 𝜖>0 we have 𝑓𝑛𝜖(𝑟)𝑓(𝑡)=𝑛𝜒𝐵(𝑡,𝜖)𝑟𝐸𝜒𝐵(𝑡,𝜖)𝐸=||𝑓𝑛||(𝑡)𝑟𝜒𝐵(𝑡,𝜖)𝐸𝑟1.(3.30) Thus, by the assumption that 0<𝑟<1 we get 𝑓𝐿𝑓(𝑡)𝐿𝑛(𝑡)=liminf𝜖0𝑓𝑛𝜖(𝑟)(𝑡)=(3.31) for a.a. 𝑡𝕊(𝑓), and the proof is finished.

The next corollary follows directly from Theorem 3.8.

Corollary 3.9. (i) Let 0<𝑝<, 1<𝑟<. If Lorentz space Γ𝑝,𝑤 satisfies a lower 𝜙-estimate for 𝑟, then for any 𝑓Γ𝑝,𝑤 and for a.a. 𝑡(0,𝛼) we have lim𝜖0𝑓𝜒𝐵(𝑡,𝜖)𝑟𝜒𝐵(𝑡,𝜖)=0.(3.32)(ii) Let 0<𝑝<, 0<𝑟<1. Then for any 𝑓Γ𝑝,𝑤 and for a.a. 𝑡𝕊(𝑓) we have lim𝜖0𝑓𝜒𝐵(𝑡,𝜖)𝑟𝜒𝐵(𝑡,𝜖)=.(3.33)

The next result needed for further applications states conditions, which guarantee that Γ𝑝,𝑤 satisfies a lower 𝜙-estimate for 𝑟, where 0<𝑟<. We omit the proof of the following proposition.

Proposition 3.10. Assume that 0<𝑝,𝑟<, and Lorentz space Γ𝑝,𝑤 satisfies a lower 𝑟-estimate. If 𝜙 is concave, then Γ𝑝,𝑤 satisfies a lower 𝜙-estimate for 𝑟.

The following example shows that Theorem 3.8 is not an empty statement in case of Γ𝑝,𝑤.

Example 3.11. Let 𝛼=, 𝑝=2, and 𝑤(𝑠)=ln(𝑠+1)+1 for all 𝑠[0,). Then for any 𝑓Γ2,𝑤 and for a.a 𝑡(0,) we have lim𝜖0𝑓𝜒𝐵(𝑡,𝜖)2Γ2,𝑤𝜒𝐵(𝑡,𝜖)Γ2,𝑤=0.(3.34)

Proof. Notice that for any 𝑡(0,), 𝑊(𝑡)=(𝑡+1)ln(𝑡+1),𝑊2(𝑡)=𝑡(𝑡+1)ln(𝑡+1)𝑡2ln(𝑡)+𝑡.(3.35) Consequently, 𝜒𝜙(𝑡)=(0,𝑡)Γ2,𝑤=(𝑡+1)2ln(𝑡+1)𝑡2ln(𝑡)+𝑡)1/2(3.36) for any 𝑡(0,). Since 𝑤 is increasing, 𝑤 satisfies 𝑅𝐵2 condition. Moreover, we have 10𝑤(𝑡)𝑡2𝑑𝑡=1𝑑𝑤(𝑡)𝑑𝑡=,𝑊𝑑𝑡2(𝑡)𝑡1=ln1+𝑡>0(3.37) for any 𝑡(0,), which concludes that 𝑊2(𝑡)/𝑡 is increasing. Hence, by Theorem 3.3 [24] we obtain that Γ2,𝑤 satisfies a lower 2-estimate. Now we claim that 𝜙 is concave. By simple calculations we get the second derivative on (0,), 𝜙(𝑡)=(𝑡ln(𝑡+1))ln(𝑡)(𝑡+2)ln(𝑡+1)𝑡2.ln(1+1/𝑡)+(2𝑡+1)ln(𝑡+1)+𝑡(3.38) We observe that 𝑡21ln1+𝑡+(2𝑡+1)ln(𝑡+1)+𝑡>0,(𝑡ln(𝑡+1))ln(𝑡)<(𝑡+2)ln(𝑡+1).(3.39) Therefore, 𝜙(𝑡)<0 for all 𝑡(0,), which implies concavity of 𝜙. Hence, by Proposition 3.10 we obtain that Γ2,𝑤 satisfies a lower 𝜙-estimate for 2Γ2,𝑤. Finally, in view of Corollary 3.9 we finish the proof.

The last part of this section is devoted to a pointwise convergence of the best and the extended best constant approximant of 𝑓 to 𝑓, that is, another type of LDT.

The first result is a corollary of Proposition 3.7. In fact, let 𝑡(0,), 𝜖>0 be such that 𝐵(𝑡,𝜖)(0,). Assume that 𝑓𝜖(𝑡)𝑇(𝑝,𝐵(𝑡,𝜖))(𝑓) for 𝑓Γ𝑝,𝑤. By definition of the best constant approximant we get||𝑓(𝑡)𝑓𝜖||(𝑡)(𝑓(𝑡)𝑓)𝜒𝐵(𝑡,𝜖)Γ𝑝,𝑤𝜒𝐵(𝑡,𝜖)Γ𝑝,𝑤+𝑓𝜖𝜒(𝑡)𝑓𝐵(𝑡,𝜖)Γ𝑝,𝑤𝜒𝐵(𝑡,𝜖)Γ𝑝,𝑤2(𝑓(𝑡)𝑓)𝜒𝐵(𝑡,𝜖)Γ𝑝,𝑤𝜒𝐵(𝑡,𝜖)Γ𝑝,𝑤.(3.40) Now applying Proposition 3.7 we get that 𝑓𝜖(𝑡)𝑓(𝑡) as 𝜖0. Therefore we get the following theorem.

Theorem 3.12. Let 𝛼= and 1𝑝<. If the space Γ𝑝,𝑤 satisfies a lower 𝜙-estimate for Γ𝑝,𝑤, then for every 𝑓Γ𝑝,𝑤 we have 𝑓𝜖(𝑡)𝑓(𝑡)as𝜖0(3.41) for a.a. 𝑡(0,), where 𝑓𝜖(𝑡)𝑇(𝑝,𝐵(𝑡,𝜖))(𝑓) is a best constant approximant of 𝑓.

In order to prove the next approximation theorem let us first establish an inequality related to extended best constant approximants of any function 𝑓Γ𝑝1,𝑤 for 1<𝑝<2.

Lemma 3.13. Let 1<𝑝<2, 𝐴[0,𝛼), 0<𝜇(𝐴)<, and let 𝑓Γ𝑝1,𝑤, 𝑓0. Then for any 𝑢𝜒𝐴𝑇(𝑝,𝐴)(𝑓) extended best constant approximant of 𝑓 and for a.a. 𝑠(0,𝛼) we have ||||𝑢𝑓(𝑠)𝑝1𝜒𝐴𝑝Γ𝑝,𝑤5(𝑓𝑓(𝑠))𝜒𝐴Γ𝑝1𝑝1,𝑤.(3.42)

Proof. Let 𝑠(0,𝛼) and |𝑓(𝑠)|<. Since (𝜒𝐴)1, by subadditivity of the maximal function and the power function 𝑣𝑝1 for 1<𝑝<2 we get ||||𝑢𝑓(𝑠)𝑝1𝜒𝐴𝑝Γ𝑝,𝑤=𝛼0𝜒𝐴(𝑡)(𝑢𝑓(𝑠))𝜒𝐴(𝑝1)(𝑡)𝑤(𝑡)𝑑𝑡𝛼0𝜒𝐴(𝑡)(𝑓𝑢)𝜒𝐴(𝑝1)+(𝑡)𝑤(𝑡)𝑑𝑡(𝑓𝑓(𝑠))𝜒𝐴Γ𝑝1𝑝1,𝑤.(3.43) We will finish the proof under assumption that 𝑓(𝑠)𝑢. In the other case the proof is similar. Since 𝑢𝜒𝐴𝑇(𝑝,𝐴)(𝑓)=[𝑓(𝑝,𝐴)𝜒𝐴,𝑓(𝑝,𝐴)𝜒𝐴], by Corollary 2.11 [7] and by Hardy-Littlewood inequality we obtain 𝛼0𝜒𝐴(𝑡)(𝑓𝑢)𝜒𝐴(𝑝1)(𝑡)𝑤(𝑡)𝑑𝑡2𝛼01𝑡𝐴{𝑓𝑢}𝜒(0,𝑡)𝜌((𝑢𝑓)𝜒𝐴,𝜒𝐴)𝑑𝜇(𝑓𝑢)𝜒𝐴(𝑝1)(𝑡)𝑤(𝑡)𝑑𝑡2𝛼0𝜒𝐴{𝑓𝑢}(𝑡)(𝑓𝑢)𝜒𝐴(𝑝1)(𝑡)𝑤(𝑡)𝑑𝑡2(𝑓𝑓(𝑠))𝜒𝐴Γ𝑝1𝑝1,𝑤+2𝛼0𝜒𝐴{𝑓𝑢}(𝑡)(𝑓(𝑠)𝑢)𝜒𝐴(𝑝1)(𝑡)𝑤(𝑡)𝑑𝑡.(3.44) Moreover, by assumption that 1<𝑝<2 we have 𝜒𝐴{𝑓𝑢}(𝑡)𝜒(0,𝜇(𝐴{𝑓𝑢}))(𝑡)+𝜇(𝐴{𝑓𝑢})𝑡𝑝1𝜒[𝜇(𝐴{𝑓𝑢}),𝛼)=𝜒(𝑡)𝐴{𝑓𝑢}(𝑝1)(𝑡)(3.45) for any 𝑡(0,𝛼). Consequently, by the fact that 𝑓(𝑠)𝑢 we obtain 𝛼0𝜒𝐴{𝑓𝑢}(𝑡)(𝑓(𝑠)𝑢)𝜒𝐴(𝑝1)(𝑡)𝑤(𝑡)𝑑𝑡𝛼0(𝑓(𝑠)𝑢)𝜒𝐴{𝑓𝑢}(𝑝1)𝜒(𝑡)𝐴(𝑝1)(𝑡)𝑤(𝑡)𝑑𝑡𝛼0(𝑓(𝑠)𝑓)𝜒𝐴{𝑓𝑢}(𝑝1)𝜒(𝑡)𝐴(𝑝1)(𝑡)𝑤(𝑡)𝑑𝑡(𝑓𝑓(𝑠))𝜒𝐴Γ𝑝1𝑝1,𝑤.(3.46) Hence, by conditions (3.43) and (3.44) we finish the proof.

Corollary 3.14. Let 1<𝑝< and 𝑤𝐷𝑝1. Assume that 𝜙 and 𝜓 are the fundamental functions of Γ𝑝,𝑤 and Γ𝑝1,𝑤, respectively. If one of the following conditions(i)there exists 𝐴>0 such that 𝑊𝑝1(𝑡)𝐴𝑊𝑝(𝑡) for all 𝑡(0,𝛼),(ii)𝑤𝐵𝑝1, holds, then 𝜙𝑝𝜓𝑝1.

Proof. Notice that for all 𝑡(0,𝛼), 𝜙(𝑡)𝑝=𝑊(𝑡)+𝑡𝑝𝛼𝑡𝑠𝑝𝑤(𝑠)𝑑𝑠𝑊(𝑡)+𝑡𝑝𝛼𝑡𝑠𝑝+1𝑡1𝑤(𝑠)𝑑𝑠=𝜓(𝑡)𝑝1.(3.47) Suppose now that there is 𝐴>0 such that 𝑊𝑝1(𝑡)𝐴𝑊𝑝(𝑡) for all 𝑡(0,𝛼). Therefore, 𝜓(𝑡)𝑝1=𝑊(𝑡)+𝑊𝑝1(𝑡)𝑊(𝑡)+𝐴𝑊𝑝(𝑡)𝐴𝜙(𝑡)𝑝(3.48) for any 𝑡(0,𝛼), and so 𝜙𝑝𝜓𝑝1.
Now assume that 𝑤𝐵𝑝1. Then there exists 𝐵>0 such that for all 𝑓Γ𝑝1,𝑤 we have 𝑓Γ𝑝1𝑝1,𝑤𝐵𝑓Λ𝑝1𝑝1,𝑤. Consequently, 𝜓(𝑡)𝑝1=𝜒(0,𝑡)Γ𝑝1𝑝1,𝑤𝜒𝐵(0,𝑡)Λ𝑝1𝑝1,𝑤=𝐵𝑊(𝑡)𝐵𝜙(𝑡)𝑝(3.49) for any 𝑡(0,𝛼), and the proof is completed.

Theorem 3.15. Let 𝛼=, 1<𝑝<2, 𝑤𝐷𝑝1 and 𝑓Γ𝑝1,𝑤, 𝑓0. Assume that 𝜙 and 𝜓 are the fundamental functions of Γ𝑝,𝑤 and Γ𝑝1,𝑤, respectively. If Lorentz space Γ𝑝1,𝑤 satisfies a lower 𝜓-estimate for Γ𝑝1,𝑤 and 𝜙𝑝𝜓𝑝1, then 𝑓𝜖(𝑡)𝑓(𝑡)as𝜖0(3.50) for a.a. 𝑡(0,), where 𝑓𝜖(𝑡)𝑇(𝑝,𝐵(𝑡,𝜖))(𝑓) is an extended best constant approximant of 𝑓.

Proof. Let 𝜖>0 and 𝑡(0,), |𝑓(𝑡)|<. By Lemma 3.13 for any 𝑓𝜖(𝑡)𝑇(𝑝,𝐵(𝑡,𝜖))(𝑓) we get ||𝑓𝜖(||𝑡)𝑓(𝑡)51/(𝑝1)(𝑓𝑓(𝑡))𝜒𝐵(𝑡,𝜖)Γ𝑝1,𝑤𝜒𝐵(𝑡,𝜖)Γ𝑝/(𝑝1)𝑝,𝑤.(3.51) Since 𝜙𝑝𝜓𝑝1, there is 𝐵>0 such that for all 𝛿>0 we have 𝜒𝐵(𝑡,𝛿)Γ𝑝1,𝑤𝐵𝜒𝐵(𝑡,𝛿)Γ𝑝/(𝑝1)𝑝,𝑤. Consequently, ||𝑓𝜖(||𝑡)𝑓(𝑡)51/(𝑝1)𝐵(𝑓𝑓(𝑡))𝜒𝐵(𝑡,𝜖)Γ𝑝1,𝑤𝜒𝐵(𝑡,𝜖)Γ𝑝1,𝑤(3.52) for any 𝑡(0,) and 𝜖>0. Hence, by Proposition 3.7 and by assumption that Γ𝑝1,𝑤 satisfies a lower 𝜓-estimate for Γ𝑝1,𝑤 we obtain that 𝑓𝜖(𝑡)𝑓(𝑡) as 𝜖0 for a.a. 𝑡(0,).

Now we present a specific family of Lorentz spaces Γ𝑝,𝑤 for which Theorems 3.12 and 3.15 are fulfilled.

Corollary 3.16. Let 𝛼= and 1<𝑞<. If one of the following conditions(i)1𝑝𝑞 and 𝑓Γ𝑞,𝑝,(ii)1<𝑝<2 and 𝑓Γ𝑞,𝑝1, 𝑓0, is satisfied, then for a.a. 𝑡(0,) we have 𝑓𝜖(𝑡)𝑓(𝑡)as𝜖0,(3.53) where 𝑓𝜖(𝑡)𝑇(𝑝,𝐵(𝑡,𝜖))(𝑓) is the best constant approximant of 𝑓Γ𝑝,𝑤 and the extended best constant approximant of 𝑓Γ𝑝1,𝑤.

Proof. Suppose that condition (i) is fulfilled. Immediately, by Theorem 3.12 and Corollary 3.6 we complete the first case. Now assume that the condition (ii) is satisfied. Define 𝑤(𝑡)=𝑡(𝑝1)/𝑞1 for 𝑡(0,). Since 1<𝑝<2, we have that 𝑊𝑞(𝑡)=𝑡𝑝1(𝑝1)/𝑞,𝑊𝑝1𝑞(𝑡)=𝑡(𝑝1)𝑞𝑝+1(𝑝1)/𝑞(3.54) for all 𝑡(0,), which implies that 𝑤𝐷𝑝1𝐵𝑝1. Consequently, by Definition 2.6 there is 𝑇(𝑝,𝐵(𝑡,𝜖)) the extended best constant approximant operator on Γ𝑞,𝑝1. Moreover, by Corollary 3.14 the fundamental functions 𝜙 and 𝜓 of the spaces Γ𝑞,𝑝 and Γ𝑞,𝑝1 respectively, satisfy 𝜙𝑝𝜓𝑝1. Hence, by Corollary 3.6 we get that Γ𝑞,𝑝1 satisfies a lower 𝜓-estimate for Γ𝑞,𝑝1. Finally, by Theorem 3.15 we obtain that 𝑓𝜖(𝑡), an extended best constant approximant of 𝑓, converges to 𝑓(𝑡) as 𝜖0 for a.a. 𝑡(0,).

Observe that the 𝐾-functional of the couple (Γ𝑞,𝑝,𝐿) can be expressed equivalently by 𝑓𝜒(0,𝑡)Γ𝑞,𝑝/𝜙(𝑡) for any 𝑓Γ𝑞,𝑝 and 𝑡(0,) [25]. The next result on inequalities between maximal function 𝑀(1)𝑞,𝑝𝑓 and the 𝐾-functional of (Γ𝑞,𝑝,𝐿) follows immediately as a consequence of Theorems 1 and  2 in [4], Corollary 3.6, and Remark 3.3.

Corollary 3.17. Let 𝛼=, 𝑝[1,), and 𝑞(0,).(i)If 1<𝑝𝑞<, then there exists 𝐶>0 such that for all 𝑓Γ𝑞,𝑝 and 𝑡>0 we have𝑀Γ(1)𝑞,𝑝𝑓𝐶(𝑡)𝑓𝜙(𝑡)𝜒(0,𝑡)Γ𝑞,𝑝.(3.55)(ii)If 1<𝑞𝑝<, then there exists 𝐶>0 such that for all 𝑓Γ𝑞,𝑝 and 𝑡>0 we get𝑀Γ(1)𝑞,𝑝𝑓𝐶(𝑡)𝑓𝜙(𝑡)𝜒(0,𝑡)Γ𝑞,𝑝.(3.56)(iii)If 𝑝=1<𝑞<, then there is 𝐶>0 such that for any 𝑓Γ𝑞,𝑝 and 𝑡>0 we obtain the inequality in the condition (i).

The decreasing rearrangement of the maximal function (𝑀Γ(1)𝑞,𝑝𝑓) and 𝐾-functional of (Γ𝑞,𝑝,𝐿) are not equivalent; that is the opposite inequalities to the ones in Corollary 3.17 do not hold. It is similar as in spaces Λ𝑞,𝑝, 1𝑝<𝑞<, as we see in the next example [4].

Example 3.18. Let 1<𝑟<. There exists a function 𝑓𝐿0 such that for all 𝑡>0,𝑓𝜒(0,𝑡)Γ𝑟,1=0𝑓𝜒(0,𝑡)(𝑠)𝑠(1/𝑟)1𝑀𝑑𝑠=,Γ(1)𝑟,1𝑓(𝑡)<.(3.57)

Proof. Since 𝑤(𝑠)=𝑠(1/𝑟)1 satisfies 𝐵1 condition, there is 𝐵>0 such that 𝜒𝐵(𝑡,𝜖)Λ𝑟,1𝜒𝐵(𝑡,𝜖)Γ𝑟,1𝐵𝜒𝐵(𝑡,𝜖)Λ𝑟,1(3.58) for all Γ𝑟,1, 𝑡(0,) and suitable 𝜖>0. Hence 𝑀Λ(1)𝑟,1𝑓(1/𝐵)𝑀Γ(1)𝑟,1𝑓. Therefore, by Remark 3.3 we conclude that 𝑀Λ(1)𝑟,1𝑓1(𝑡)𝐵𝑀Γ(1)𝑟,1𝑓(𝑡)(3.59) for all 𝑡>0. Now, by Theorem  3 in [4] there exists 𝑓𝐿0 such that 𝑓𝜒(0,𝑡)Λ𝑟,1= and (𝑀Λ(1)𝑟,1𝑓)(𝑡)< for all 𝑡>0, which implies that 𝑓𝜒(0,𝑡)Γ𝑟,1= and (𝑀Γ(1)𝑟,1𝑓)(𝑡)< for any 𝑡>0.

Remark 3.19. Example 3.18 can be expanded to the case when 1𝑝<𝑞<. Indeed if 1𝑝<𝑞< and 𝑟=𝑞/𝑝 then we have Γ𝑞,𝑝=Γ(𝑝)𝑟,1 and Γ𝑞,𝑝||𝑝Γ1/𝑝𝑞,𝑝. As well as (𝑀Γ(1)𝑟,1||𝑝)1/𝑝𝑀Γ(1)𝑞,𝑝 (·) on {𝑓Γ𝑞,𝑝(𝐶)𝐶isconvex,𝜇(𝐶)<}. This can be obtained by simple observation that the weight function 𝑤(𝑡)=𝑡1/𝑟1 satisfies condition 𝐵𝑠 for any 1/𝑟<𝑠<, and then Γ𝑟,1=Λ𝑟,1 and Γ𝑞,𝑝=Λ𝑞,𝑝 as sets with equivalent norms.

4. Uniqueness of the Extended Best Constant APProximant on Γ𝑃,𝑊

In this section we prove uniqueness of the expansion of the best constant approximant for any 𝑓Γ𝑝1,𝑤, if 1<𝑝< and 𝑤>0. Namely, we show that the extended best constant approximant operator 𝑇(𝑝,𝐴) becomes the point value operator. Throughout this section we assume that 𝐴[0,𝛼) is a set of positive and finite measure. Let the truncation of any function 𝑓𝐿0 be defined as 𝑓(𝑛)(𝑠)=𝑓(𝑠) if |𝑓(𝑠)|<𝑛, and 𝑓(𝑛)(𝑠)=sign𝑓(𝑠)𝑛 if |𝑓(𝑠)|𝑛.

Lemma 4.1. Let 1<𝑝< and 𝑓Γ𝑝1,𝑤. Then 𝑓(𝑛)Γ𝑝,𝑤 for all 𝑛.

Proof. Since 𝑓Γ𝑝1,𝑤, we get that 𝑑𝑓(𝜆)< for every 𝜆>0. Let 𝜖(0,1). Then 𝑐𝑛||𝑓||||𝑓||=𝜇𝑠(𝑠)𝑛𝜇𝑠(𝑠)>𝑛𝜖=𝑑𝑓(𝑛𝜖)<(4.1) for any 𝑛. Moreover, (𝑓(𝑛))𝑓 and (𝑓(𝑛))𝑛 for any 𝑛. Therefore, 𝛼0𝑓(𝑛)𝑝(𝑡)𝑤(𝑡)𝑑𝑡𝑛𝑝𝑊𝑐𝑛+𝛼𝑐𝑛𝑓(𝑛)(𝑝1)𝑓(𝑡)(𝑛)(𝑡)𝑤(𝑡)𝑑𝑡𝑛𝑝𝑊𝑐𝑛+𝑛𝛼𝑐𝑛𝑓(𝑝1)(𝑡)𝑤(𝑡)𝑑𝑡<(4.2) for all 𝑛, which concludes the proof.

We omit the standard proof of the next lemma.

Lemma 4.2. Let 𝑓𝐿0 and 𝑢. Then for a.a. 𝑠𝐴 there exists 𝑁(𝑢,𝑠) such that for all 𝑛𝑁(𝑢,𝑠) we have the following.(i){𝑣𝐴𝑓(𝑛)(𝑣)+𝑓(𝑛)(𝑠)<2𝑢}={𝑣𝐴𝑓(𝑣)+𝑓(𝑠)<2𝑢}.(ii){𝑣𝐴𝑓(𝑛)(𝑣)=𝑓(𝑛)(𝑠),𝑣𝑠}={𝑣𝐴𝑓(𝑣)=𝑓(𝑠),𝑣𝑠}.(iii){𝑣𝐴𝑓(𝑛)(𝑣)>𝑓(𝑛)(𝑠)}={𝑣𝐴𝑓(𝑣)>𝑓(𝑠)}.

Now we discuss a convergence of a sequence of functions 𝑆𝑝(𝑓(𝑛),𝐴)(𝑢) to the function 𝑆𝑝(𝑓,𝐴)(𝑢) as 𝑛 for all 𝑢 and 1<𝑝<, whenever 𝑓Γ𝑝1,𝑤.

Theorem 4.3. Let 1<𝑝< and 𝑓Γ𝑝1,𝑤. Then for all 𝑢, lim𝑛𝑆𝑝(𝑓(𝑛),𝐴)(𝑢)=𝑆𝑝(𝑓,𝐴)(𝑢).(4.3)

Proof. Denote 𝐸=𝐴{𝑓𝑢}. Notice that there exists 𝑁𝑢 such that 𝑁𝑢>|𝑢| and for all 𝑛𝑁𝑢 we have 𝐸=𝐴{𝑓(𝑛)𝑢}. Since 𝜌((𝑓𝑢)𝜒𝐴,𝜒𝐴) is a measure-preserving transformation, by Definition 2.3 we get 𝑆𝑝(𝑓,𝐴)(𝑢)=2𝛼01𝑡𝐴{𝑓𝑢}𝜒(0,𝑡)𝜌((𝑓𝑢)𝜒𝐴,𝜒𝐴)(𝑓𝑢)𝜒𝐴(𝑝1)(𝑡)𝑤(𝑡)𝑑𝑡𝛼0𝜒𝐴(𝑡)(𝑓𝑢)𝜒𝐴(𝑝1)(𝑡)𝑤(𝑡)𝑑𝑡.(4.4) We claim that for a.a. 𝑠𝐸 there exists 𝑁(𝑢,𝑠) such that for all 𝑛𝑁(𝑢,𝑠) we have 𝜌((𝑓(𝑛)𝑢)𝜒𝐴,𝜒𝐴)(𝑠)=𝜌((𝑓𝑢)𝜒𝐴,𝜒𝐴)(𝑠).(4.5) Let 𝑠𝐸 and |𝑓(𝑠)|<. If 𝑓(𝑠)>𝑢, then by Lemma 4.2 there exists 𝑁(𝑢,𝑠) such that for any 𝑛𝑁(𝑢,𝑠) we obtain 𝜌((𝑓𝑢)𝜒𝐴,𝜒𝐴)(𝑠)=𝜇(𝑣𝐴𝑓(𝑣)>𝑓(𝑠))+𝜇(𝑣𝐴𝑓(𝑣)+𝑓(𝑠)<2𝑢)+𝜇(𝑣𝐴𝑓(𝑣)=𝑓(𝑠),𝑣𝑠)=𝜇𝑣𝐴𝑓(𝑛)(𝑣)>𝑓(𝑛)(𝑠)+𝜇𝑣𝐴𝑓(𝑛)(𝑣)+𝑓(𝑛)(𝑠)<2𝑢+𝜇𝑣𝐴𝑓(𝑛)(𝑣)=𝑓(𝑛)(𝑠),𝑣𝑠=𝜌((𝑓(𝑛)𝑢)𝜒𝐴,𝜒𝐴)(𝑠).(4.6) If 𝑓(𝑠)=𝑢, then again by Lemma 4.2 for 𝑛𝑁(𝑢,𝑠) we have 𝜌((𝑓𝑢)𝜒𝐴,𝜒𝐴)(𝑠)=𝜇(𝑣𝐴𝑓(𝑣)>𝑓(𝑠))+𝜇(𝑣𝐴𝑓(𝑣)+𝑓(𝑠)<2𝑢)+𝜇(𝑣𝐴𝑣𝑠)𝜇(𝑣𝐴𝑓(𝑣)=𝑓(𝑠),𝑣𝑠)=𝜇𝑣𝐴𝑓(𝑛)(𝑣)>𝑓(𝑛)(𝑠)+𝜇𝑣𝐴𝑓(𝑛)(𝑣)+𝑓(𝑛)(𝑠)<2𝑢+𝜇(𝑣𝐴𝑣𝑠)𝜇𝑣𝐴𝑓(𝑛)(𝑣)=𝑓(𝑛)(𝑠),𝑣𝑠=𝜌((𝑓(𝑛)𝑢)𝜒𝐴,𝜒𝐴)(𝑠).(4.7) Hence we obtain our claim. It follows lim𝑛𝐴{𝑓(𝑛)𝑢}𝜒(0,𝑡)𝜌((𝑓(𝑛)𝑢)𝜒𝐴,𝜒𝐴)=𝐴{𝑓𝑢}𝜒(0,𝑡)𝜌((𝑓𝑢)𝜒𝐴,𝜒𝐴)(4.8) for any 𝑡(0,𝛼). Moreover lim𝑛((𝑓(𝑛)𝑓)𝜒𝐴)(𝑡)=0 for all 𝑡(0,𝛼), and consequently by triangle inequality for maximal function we obtain lim𝑛𝑓(𝑛)𝜒𝑢𝐴(𝑡)=(𝑓𝑢)𝜒𝐴(𝑡)(4.9) for all 𝑡(0,𝛼). Since 𝜌((𝑓(𝑛)𝑢)𝜒𝐴,𝜒𝐴) is the measure-preserving transformation for all 𝑛, by the fact that the power function 𝑣𝑝1 for 𝑝>1 is subadditive with a constant 𝐶>0, we have 1𝑡𝐴{𝑓(𝑛)𝑢}𝜒(0,𝑡)𝜌((𝑓(𝑛)𝑢)𝜒𝐴,𝜒𝐴)𝑓(𝑛)𝜒𝑢𝐴(𝑝1)(𝜒𝑡)𝐴𝑓(𝑡)(𝑛)𝜒𝑢𝐴(𝑝1)(𝑡)𝐶𝑓𝜒𝐴(𝑝1)(𝑡)+𝑢𝜒𝐴(𝑝1)(𝑡)(4.10) for every 𝑡(0,𝛼) and 𝑛. Combining now (4.8), (4.9), and (4.10) we get lim𝑛𝛼01𝑡𝐴{𝑓(𝑛)𝑢}𝜒(0,𝑡)𝜌((𝑓(𝑛)𝑢)𝜒𝐴,𝜒𝐴)𝑓(𝑛)𝜒𝑢𝐴(𝑝1)=(𝑡)𝑤(𝑡)𝑑𝑡𝛼01𝑡𝐴{𝑓𝑢}𝜒(0,𝑡)𝜌((𝑓𝑢)𝜒𝐴,𝜒𝐴)(𝑓𝑢)𝜒𝐴(𝑝1)(𝑡)𝑤(𝑡)𝑑𝑡,lim𝑛𝛼0𝜒𝐴𝑓(𝑡)(𝑛)𝜒𝑢𝐴(𝑝1)(𝑡)𝑤(𝑡)𝑑𝑡=𝛼0𝜒𝐴(𝑡)(𝑓𝑢)𝜒𝐴(𝑝1)(𝑡)𝑤(𝑡)𝑑𝑡.(4.11) In view of (4.4) we complete the proof.

The following characterization of the function 𝑆𝑝(𝑓,𝐴)(𝑢) is an essential fact in the proof of the main result.

Proposition 4.4. Let 1<𝑝<, 𝑤 be a positive weight function and let 𝑓Γ𝑝1,𝑤. Then the function 𝑆𝑝(𝑓,𝐴)(𝑢) is strictly decreasing with respect to 𝑢.

Proof. First we will show that for any 𝑢,𝑣, 𝑢<𝑣 there exist 𝑎(𝑢,𝑣] and 𝐵[0,𝛼), 𝜇(𝐵)>0 such that for all 𝑡𝐵, 𝐾(𝑓,𝐴)(𝑢,𝑡)(𝑓+𝑢)𝜒𝐴(𝑝1)(𝑡)<𝐾(𝑓,𝐴)(𝑎,𝑡)(𝑓+𝑎)𝜒𝐴(𝑝1)(𝑡).(4.12)
Suppose for a contrary that there exist 𝑢<𝑣 such that for all 𝑎(𝑢,𝑣] and for a.a. 𝑡[0,𝛼) we have 𝐾(𝑓,𝐴)(𝑢,𝑡)(𝑓+𝑢)𝜒𝐴(𝑝1)(𝑡)𝐾(𝑓,𝐴)(𝑎,𝑡)(𝑓+𝑎)𝜒𝐴(𝑝1)(𝑡).(4.13) Since 𝐾(𝑓,𝐴)(𝑎,𝑡) and ((𝑓+𝑎)𝜒𝐴)(𝑡) are continuous functions with respect to 𝑡(0,𝛼) and 𝐾(𝑓,𝐴)(𝑎,𝑡) is right-continuous with respect to 𝑎, we get that the above inequality is fulfilled for all 𝑡(0,𝛼) and for any 𝑎[𝑢,𝑣]. By convexity of the function 𝑎((𝑓+𝑎)𝜒𝐴)𝑝(𝑡) and by Proposition 4.2 [20] we get that (𝑑+/𝑑𝑎)((𝑓+𝑎)𝜒𝐴)𝑝(𝑡)=𝑝𝐾(𝑓,𝐴)(𝑎,𝑡)((𝑓+𝑎)𝜒𝐴)(𝑝1)(𝑡) is increasing with respect to 𝑎 for any 𝑡(0,𝛼), which implies that 𝐾(𝑓,𝐴)(𝑎,𝑡)(𝑓+𝑎)𝜒𝐴(𝑝1)(𝑡)𝐾(𝑓,𝐴)(𝑏,𝑡)(𝑓+𝑏)𝜒𝐴(𝑝1)(𝑡)(4.14) for any 𝑎<𝑏 and for all 𝑡(0,𝛼). Hence 𝐾(𝑓,𝐴)(𝑢,𝑡)(𝑓+𝑢)𝜒𝐴(𝑝1)(𝑡)=𝐾(𝑓,𝐴)(𝑎,𝑡)(𝑓+𝑎)𝜒𝐴(𝑝1)(𝑡)(4.15) for all 𝑎[𝑢,𝑣] and 𝑡(0,𝛼). Pick up 𝑏(𝑢,𝑣]. Denote 𝛿=(𝑏𝑢)/2. Notice that 𝐾(𝑓,𝐴)(𝑎,𝑡)(𝑓+𝑎)𝜒𝐴(𝑝1)(𝑡)=𝐾(𝑓,𝐴)(𝑎𝛿,𝑡)(𝑓+𝑎+𝛿)𝜒𝐴(𝑝1)(𝑡)(4.16) for all 𝑎[𝑢,𝑢+𝛿] and 𝑡(0,𝛼), which yields that 𝑢𝑢+𝛿𝑑+𝑑𝑎(𝑓+𝑎)𝜒𝐴𝑝(𝑡)𝑑𝑎=𝑢𝑢+𝛿𝑑+𝑑𝑎(𝑓+𝑎+𝛿)𝜒𝐴𝑝=(𝑡)𝑑𝑎𝑏𝑢+𝛿𝑑+𝑑𝑎(𝑓+𝑎)𝜒𝐴𝑝(𝑡)𝑑𝑎(4.17) for all 𝑡(0,𝛼). Consequently, 𝑓𝜒𝐴+𝑢+𝑏2𝜒𝐴𝑝1(𝑡)=2(𝑓+𝑢)𝜒𝐴𝑝1(𝑡)+2(𝑓+𝑏)𝜒𝐴𝑝(𝑡)(4.18) for any 𝑡(0,𝛼). Moreover, by subadditivity of the maximal function and by convexity of the power function 𝑠𝑝 for 𝑝>1 we get 𝑓𝜒𝐴+𝑢+𝑏2𝜒𝐴𝑝1(𝑡)2(𝑓+𝑢)𝜒𝐴1(𝑡)+2(𝑓+𝑏)𝜒𝐴(𝑡)𝑝12(𝑓+𝑢)𝜒𝐴𝑝1(𝑡)+2(𝑓+𝑏)𝜒𝐴𝑝(𝑡)(4.19) for all 𝑡(0,𝛼). Therefore, (𝑓+𝑢)𝜒𝐴+(𝑓+𝑏)𝜒𝐴(𝑡)=(𝑓+𝑢)𝜒𝐴(𝑡)+(𝑓+𝑏)𝜒𝐴(𝑡),(4.20) which implies that 𝑡0(𝑓+𝑢)𝜒𝐴+(𝑓+𝑏)𝜒𝐴((𝑠)𝑓+𝑢)𝜒𝐴((𝑠)𝑓+𝑏)𝜒𝐴(𝑠)𝑑𝑠=0(4.21) for any 𝑡(0,𝛼). Hence (𝑓+𝑢)𝜒𝐴+(𝑓+𝑏)𝜒𝐴(𝑠)=(𝑓+𝑢)𝜒𝐴(𝑠)+(𝑓+𝑏)𝜒𝐴(𝑠)(4.22) for a.a. 𝑠(0,𝛼) and for any 𝑏[𝑢,𝑣]. By Corollary 9 [8, page 65] we obtain that the functions (𝑓+𝑢)𝜒𝐴 and (𝑓+𝑏)𝜒𝐴 for all 𝑏[𝑢,𝑣] are of constant sign almost everywhere, that is, sign(𝑓+𝑢)(𝑠)=sign(𝑓+𝑏)(𝑠)(4.23) for a.a. 𝑠𝐴 and have a common system of sets {𝐸𝑡𝑡(0,𝛼)} such that 𝜇(𝐸𝑡)=𝑡 and 𝑡0(𝑓+𝑏)𝜒𝐴(𝑠)𝑑𝑠=𝐸𝑡𝐴||||(𝑓+𝑏𝑠)𝑑𝑠(4.24) for all 𝑡(0,𝛼) and for all 𝑏[𝑢,𝑣]. We claim that there exist 𝑏(𝑢,𝑣] and 𝐵(0,𝛼), 𝜇(𝐵)>0, such that we have either for all 𝑡𝐵, (𝑓+𝑢)𝜒𝐴(𝑡)<(𝑓+𝑏)𝜒𝐴(𝑡),(4.25) or for all 𝑡𝐵, (𝑓+𝑢)𝜒𝐴(𝑡)>(𝑓+𝑏)𝜒𝐴(𝑡).(4.26) Suppose that the claim does not hold. Then for any 𝑏(𝑢,𝑣] and 𝑡(0,𝛼) we get 𝑡0(𝑓+𝑢)𝜒𝐴(𝑠)𝑑𝑠=𝑡0(𝑓+𝑏)𝜒𝐴(𝑠)𝑑𝑠.(4.27) Thus, by the condition (4.24) we obtain 𝐸𝑡𝐴||||(||||(𝑓+𝑢𝑠)𝑓+𝑏𝑠)𝑑𝑠=0(4.28) for every 𝑏(𝑢,𝑣] and 𝑡(0,𝛼). Since 𝐸𝑟𝐸𝑡 for any 𝑟<𝑡 and 𝜇(𝐸𝑡)=𝑡 for any 𝑡(0,𝛼), we get that |𝑓(𝑠)+𝑢|=|𝑓(𝑠)+𝑏| for any 𝑏(𝑢,𝑣] and for a.a. 𝑠𝐴. Hence, by the fact (4.23) we get a contradiction. Now we will consider two cases.
Case 1. Assume that there exist 𝑏(𝑢,𝑣] and 𝐵(0,𝛼), 𝜇(𝐵)>0 such that (𝑓+𝑢)𝜒𝐴(𝑡)<(𝑓+𝑏)𝜒𝐴(𝑡)(4.29) for all 𝑡𝐵. Thus and by conditions (4.23) and (4.24) we obtain 0<𝑡0(𝑓+𝑏)𝜒𝐴(𝑠)𝑑𝑠𝑡0(𝑓+𝑢)𝜒𝐴(𝑠)𝑑𝑠=𝐸𝑡𝐴||||||||𝑓(𝑠)+𝑏𝑓(𝑠)+𝑢𝑑𝑠=(𝑏𝑢)𝐸𝑡𝐴sign(𝑓(𝑠)+𝑢)𝑑𝑠(4.30) for all 𝑡𝐵. Since 𝑏𝑢>0, we have 0<𝐸𝑡𝐴sign(𝑓(𝑠)+𝑢)𝑑𝑠(4.31) for any 𝑡𝐵. The function 𝑎((𝑓+𝑎)𝜒𝐴)(𝑡) is strictly increasing on [𝑢,𝑣] and 𝑡𝐵. In fact, let 𝜉,𝜂[0,𝑣𝑢] and 𝜉<𝜂. By conditions (4.23) and (4.24) we conclude that (𝑓+𝑢+𝜉)𝜒𝐴(𝑡)(𝑓+𝑢+𝜂)𝜒𝐴=1(𝑡)𝑡𝐸𝑡𝐴1sign(𝑓(𝑠)+𝑢)(𝑓(𝑠)+𝑢+𝜉)𝑑𝑠𝑡𝐸𝑡𝐴=sign(𝑓(𝑠)+𝑢)(𝑓(𝑠)+𝑢+𝜂)𝑑𝑠𝜉𝜂𝑡𝐸𝑡𝐴sign(𝑓(𝑠)+𝑢)𝑑𝑠<0(4.32) for any 𝑡𝐵. Let 𝑎(𝑢,𝑣) and 𝛿𝑎=min{𝑣𝑎,𝑎𝑢}/2. Then, by condition (4.32) we get (𝑓+𝑢)𝜒𝐴(𝑡)<(𝑓+𝑢+𝜖)𝜒𝐴(𝑡)<(𝑓+𝑎)𝜒𝐴(𝑡)<(𝑓+𝑎+𝜖)𝜒𝐴(𝑡)(4.33) for every 0<𝜖𝛿𝑎 and for all 𝑡𝐵. Repeating calculations in condition (4.32) with 𝑎 instead of 𝑢, in view of (4.23) and (4.24) we have (𝑓+𝑢+𝜖)𝜒𝐴(𝑡)(𝑓+𝑢)𝜒𝐴(𝑡)=(𝑓+𝑎+𝜖)𝜒𝐴(𝑡)(𝑓+𝑎)𝜒𝐴(𝑡)>0(4.34) for all 0<𝜖𝛿𝑎 and 𝑡𝐵. Since the functions ((𝑓+𝑢+𝛿𝑎)𝜒𝐴)(𝑡) and ((𝑓+𝑎)𝜒𝐴)(𝑡) are continuous and decreasing with respect to 𝑡, there exists a compact interval [𝑐,𝑑]𝐵 such that 𝑓+𝑢+𝛿𝑎𝜒𝐴(𝑡)<(𝑓+𝑎)𝜒𝐴(𝑡)(4.35) for all 𝑡[𝑐,𝑑]. Define 𝐹(𝑡)=(𝑓+𝑎)𝜒𝐴(𝑡)𝑓+𝑢+𝛿𝑎𝜒𝐴(𝑡)(4.36) for any 𝑡[𝑐,𝑑]. Since 𝐹 is positive and continuous on [𝑐,𝑑], we get min𝑡[𝑐,𝑑]{𝐹(𝑡)}>0. Consequently, there is 𝛾𝑎(0,1) such that 𝑓+𝑢+𝛿𝑎𝜒𝐴(𝑡)<1𝛾𝑎(𝑓+𝑎)𝜒𝐴(𝑡)(4.37) for all 𝑡[𝑐,𝑑]. Hence, by condition (4.33) we get (𝑓+𝑢)𝜒𝐴(𝑡)<(𝑓+𝑢+𝜖)𝜒𝐴(𝑡)<1𝛾𝑎(𝑓+𝑎)𝜒𝐴<(𝑡)1𝛾𝑎(𝑓+𝑎+𝜖)𝜒𝐴(𝑡)(4.38) for every 𝑡[𝑐,𝑑] and 0<𝜖𝛿𝑎. Therefore, by (4.34) we obtain (𝑓+𝑢+𝜖)𝜒𝐴𝑝(𝑡)(𝑓+𝑢)𝜒𝐴𝑝(𝑡)𝜖1𝛾𝑎𝑝1(𝑓+𝑎+𝜖)𝜒𝐴𝑝(𝑡)(𝑓+𝑎)𝜒𝐴𝑝(𝑡)𝜖(4.39) for any 𝑡[𝑐,𝑑] and 0<𝜖𝛿𝑎. Now according to Proposition 4.2 [20] we conclude 𝐾(𝑓,𝐴)(𝑢,𝑡)(𝑓+𝑢)𝜒𝐴(𝑝1)(𝑡)<𝐾(𝑓,𝐴)(𝑎,𝑡)(𝑓+𝑎)𝜒𝐴(𝑝1)(𝑡)(4.40) for all 𝑡[𝑐,𝑑], which contradicts condition (4.15) and finishes the first case.Case 2. Suppose that there exist 𝑏(𝑢,𝑣] and 𝐵(0,𝛼) with 𝜇(𝐵)>0 such that (𝑓+𝑢)𝜒𝐴(𝑡)>(𝑓+𝑏)𝜒𝐴(𝑡)(4.41) for any 𝑡𝐵. Analogously as in the previous case we obtain that 𝐸𝑡𝐴sign(𝑓(𝑠)+𝑢)𝑑𝑠<0(4.42) for all 𝑡𝐵. We claim that the function 𝑎((𝑓+𝑎)𝜒𝐴)(𝑡) is strictly decreasing for any 𝑎[𝑢,𝑣]. Let 𝜉,𝜂[0,𝑣𝑢] and 𝜉<𝜂. The conditions (4.23) and (4.24) imply (𝑓+𝑢+𝜉)𝜒𝐴((𝑡)𝑓+𝑢+𝜂)𝜒𝐴(1𝑡)=𝑡𝐸𝑡𝐴||||||||=𝑓(𝑠)+𝑢+𝜉𝑓(𝑠)+𝑢+𝜂𝑑𝑠𝜉𝜂𝑡𝐸𝑡𝐴sign(𝑓(𝑠)+𝑢)𝑑𝑠>0(4.43) for any 𝑡𝐵. Therefore, for any 𝑎(𝑢,𝑣), 0<𝜖<𝑚𝑖𝑛{𝑣𝑎,𝑎𝑢}/2 and for all 𝑡𝐵 we have (𝑓+𝑢)𝜒𝐴(𝑡)>(𝑓+𝑢+𝜉)𝜒𝐴(𝑡)>(𝑓+𝑎)𝜒𝐴(𝑡)>(𝑓+𝑎+𝜖)𝜒𝐴(𝑡).(4.44) Repeating calculation in condition (4.43) with 𝑎 instead of 𝑢, in view of (4.23) and (4.24) we also get (𝑓+𝑢)𝜒𝐴(𝑡)(𝑓+𝑢+𝜖)𝜒𝐴(𝑡)=(𝑓+𝑎)𝜒𝐴(𝑡)(𝑓+𝑎+𝜖)𝜒𝐴(𝑡)>0(4.45) for all 𝑡𝐵. Similarly as in the previous case there exists 𝛾𝑎>0 such that (𝑓+𝑢+𝜖)𝜒𝐴𝑝(𝑡)(𝑓+𝑢)𝜒𝐴𝑝(𝑡)𝜖1+𝛾𝑎𝑝1(𝑓+𝑎+𝜖)𝜒𝐴𝑝(𝑡)(𝑓+𝑎)𝜒𝐴𝑝(𝑡)𝜖(4.46) for any 𝑡𝐵 and 0<𝜖<𝑚𝑖𝑛{𝑎𝑢,𝑣𝑎}/2. Consequently, by Proposition 4.2 [20] we get 𝐾(𝑓,𝐴)(𝑢,𝑡)(𝑓+𝑢)𝜒𝐴(𝑝1)(𝑡)<𝐾(𝑓,𝐴)(𝑎,𝑡)(𝑓+𝑎)𝜒𝐴(𝑝1)(𝑡)(4.47) for all 𝑡𝐵, which gives us a contradiction with (4.15) and completes the proof of inequality (4.12).
Let now 𝑎,𝑏 and 𝑎<𝑏. Denote 𝑣=𝑎 and 𝑢=𝑏. Clearly 𝑢<𝑣. By (4.12) there exist 𝑐(𝑢,𝑣] and 𝐵[0,𝛼), 𝜇(𝐵)>0 such that 𝐾(𝑓,𝐴)(𝑢,𝑡)(𝑓+𝑢)𝜒𝐴(𝑝1)(𝑡)<𝐾(𝑓,𝐴)(𝑐,𝑡)(𝑓+𝑐)𝜒𝐴(𝑝1)(𝑡)(4.48) for all 𝑡𝐵. Hence, by (4.14) and by Proposition  2.4 in [7] we get 𝑆𝑝(𝑓,𝐴)(𝑏)=𝑆𝑝(𝑓,𝐴)(𝑢)=𝐵𝐵𝑐𝐾(𝑓,𝐴)((𝑢,𝑡)𝑓+𝑢)𝜒𝐴(𝑝1)(<𝑡)𝑤(𝑡)𝑑𝑡𝐵𝐾(𝑓,𝐴)(𝑐,𝑡)(𝑓+𝑐)𝜒𝐴(𝑝1)+(𝑡)𝑤(𝑡)𝑑𝑡𝐵𝑐𝐾(𝑓,𝐴)(𝑢,𝑡)(𝑓+𝑢)𝜒𝐴(𝑝1)(𝑡)𝑤(𝑡)𝑑𝑡𝑆𝑝(𝑓,𝐴)(𝑐)𝑆𝑝(𝑓,𝐴)(𝑣)=𝑆𝑝(𝑓,𝐴)(𝑎),(4.49) and the proof is done.

Now we establish the main theorem of this section.

Theorem 4.5. Let 1<𝑝< and 𝑤 be a positive weight function. Then the extended best constant approximant operator 𝑇(𝑝,𝐴) assumes an unique value, that is, for any 𝑓Γ𝑝1,𝑤, 𝑇(𝑝,𝐴)(𝑓)=𝑓(𝑝,𝐴)𝜒𝐴=𝑓(𝑝,𝐴)𝜒𝐴.(4.50)

Proof. Suppose that 𝑓(𝑝,𝐴)<𝑓(𝑝,𝐴). Then there exist 𝑢,𝑣(𝑓(𝑝,𝐴),𝑓(𝑝,𝐴)) such that 𝑢<𝑣. By Theorem 2.9 [7] and by Proposition 4.4 we obtain 0𝑆(𝑓,𝐴)𝑓(𝑝,𝐴)<𝑆(𝑓,𝐴)(𝑣)<𝑆(𝑓,𝐴)(𝑢),0𝑆(𝑓,𝐴)𝑓(𝑝,𝐴)<𝑆(𝑓,𝐴)(𝑢)<𝑆(𝑓,𝐴)(𝑣).(4.51) Let (𝑓(𝑛)) be a sequence of truncations of 𝑓. By Theorem 4.3 there exists 𝑁0 such that for all 𝑛𝑁0 we have 0<𝑆(𝑓(𝑛),𝐴)(𝑣),0<𝑆(𝑓(𝑛),𝐴)(𝑣)(4.52) as well as 0<𝑆(𝑓(𝑛),𝐴)(𝑢),0<𝑆(𝑓(𝑛),𝐴)(𝑢).(4.53) Choose 𝑛𝑁0. By Lemma 4.1, 𝑓(𝑛)Γ(𝑝,𝑤) for all 𝑛. Now by conditions (4.52) and (4.53) and by Theorem 2.4 we get that 𝑢,𝑣𝑇(𝑝,𝐴)(𝑓(𝑛)) for all 𝑛𝑁0. Finally, by Corollary 2.5 we obtain that 𝑇(𝑝,𝐴)(𝑓(𝑛)) is unique for 𝑛𝑁0, which implies a contradiction and finishes the proof.