692
Views
0
CrossRef citations to date
0
Altmetric
Articles

A lower estimate for weak-type Fourier multipliers

ORCID Icon &
Pages 642-660 | Received 28 Nov 2020, Accepted 18 Jul 2021, Published online: 01 Aug 2021

Abstract

Asmar et al. [Note on norm convergence in the space of weak type multipliers. J Operator Theory. 1998;39(1):139–149] proved that the space of weak-type Fourier multipliers acting from Lp into Lp, is continuously embedded into L. We obtain a sharper result in the setting of abstract Lorentz spaces Λq(X) with 0<q built upon a Banach function space X on Rn. We consider a source space S and a target space T in the class of admissible spaces A:={X,Λq(X):0<q}. Let MS,T0 denote the space of Fourier multipliers acting from S to T. We show that if the space X satisfies the weak doubling property, then the space MΛq(X),Λ(X)0 is continuously embedded into L for every 0<q. This implies that MS,T0 is a quasi-Banach space for all choices of source and target spaces S,TA.

AMS Subject Classifications:

1. Introduction

Let S and S denote the Schwartz spaces of rapidly decaying functions and of tempered distributions on Rn, respectively. The action of a distribution aS on a function uS is denoted by a,u:=a(u). A Fourier multiplier on Rn with symbol aS is defined as the operator (1) uF1aFu,(1) where (Fu)(ξ):=Rnu(x)eixξdx is the Fourier transform of uS, F1 denotes the inverse Fourier transform, and xξ denotes the scalar product of x,ξRn. We observe that since uS and aS, the function Fu belongs to the space S and aFu is a tempered distribution. Thus F1aFu is well defined and it belongs to S. In fact, we have F1aFu=(F1a)u, and therefore, F1aFuCpoly (see, e.g. [Citation1, Theorem 2.3.20]), where Cpoly denotes the set of all infinitely differentiable functions f:RnC such that for every αZ+n there exist mαZ+:={0,1,2,} and Cα>0 satisfying |xαf(x)|Cα(1+|x|)mα for all xRn. Thus, if uS and aS, then F1aFu is a regular tempered distribution, whose action on vS is evaluated as follows: F1aFu,v=Rn(F1aFu)(x)v(x)dxfor allvS. The aim of this paper is to study the Fourier multiplier operator (Equation1) as an operator acting from a source space S to a target space T, where S,T belong to the class of admissible spaces A consisting of a given Banach function space X on Rn and all abstract Lorentz spaces Λq(X), 0<q, built upon X. Our paper is inspired by the work by Asmar et al. [Citation2], where the operator (Equation1) was considered as acting from the Lebesgue space X=Lp(G), 1p<, to the Marcinkiewicz space Λ(X)=Lp,(G) over a locally compact abelian group G. It is closely related to our recent work [Citation3], where we treated the operator (Equation1) as acting from a Banach function space X on Rn into itself. To formulate our results, we need several definitions.

Let M denote the set of all Lebesgue measurable extended complex-valued functions on Rn, that is, the functions of the form f=f1+if2, where f1 and f2 are Lebesgue measurable extended real-valued functions (see, e.g. [Citation4, Definition 16.1]). Let M+ be the subset of functions in M whose values lie in [0,]. The characteristic function of a measurable set ERn is denoted by χE and the Lebesgue measure of E is denoted by |E|. Following [Citation5, p. 3] (see also [Citation6, Chap. 1, Definition 1.1]), a mapping ρ:M+[0,] is called a Banach function norm if, for all functions f,g,fj (jN) in M+, for all constants a0, and for all measurable subsets E of Rn, the following properties hold: (A1)ρ(f)=0f=0a.e.,ρ(af)=aρ(f),ρ(f+g)ρ(f)+ρ(g),(A2)0gfa.e.ρ(g)ρ(f)(the lattice property),(A3)0fjfa.e.ρ(fj)ρ(f)(the Fatou property),(A4)E is boundedρ(χE)<,(A5)E is boundedEf(x)dxCEρ(f) with CE(0,) that may depend on E and ρ but is independent of f. When functions differing only on a set of measure zero are identified, the set X of all functions fM for which ρ(|f|)< becomes a Banach space under the norm fX:=ρ(|f|) and under the natural linear space operations (see [Citation5, Chap. 1, § 1, Theorem 1] or [Citation6, Chap. 1, Theorems 1.4 and 1.6]). It is called a Banach function space. The class of Banach function spaces is very large. It includes, for example, classical Lebesgue spaces Lp, 1p [Citation6, p. 3], Orlicz spaces LΦ [Citation6, Chap. 4, Theorem 8.9], variable Lebesgue spaces Lp() [Citation7, Section 2.10.3].

We note that our definition of a Banach function space is slightly different from that found in [Citation6, Chap. 1, Definition 1.1]. In particular, in Axioms (A4) and (A5), we assume that E is a bounded set, whereas it is sometimes assumed that E merely satisfies |E|<. We do this so that the weighted Lebesgue spaces with Muckenhoupt weights satisfy Axioms (A4) and (A5). Moreover, it is well known that all main elements of the general theory of Banach function spaces work with (A4) and (A5) stated for bounded sets [Citation5] (see also the discussion at the beginning of Chapter 1 on page 2 of [Citation6]). Unfortunately, we overlooked that the definition of a Banach function space in our previous work [Citation3] had to be changed by replacing in Axioms (A4) and (A5) the requirement of |E|< by the requirement that E is a bounded set to include weighted Lebesgue spaces in our considerations. However, the results proved in the above paper remain true.

Let M0 (resp., M0+) denote the set of all a.e. finite functions in M (resp., in M+). For 0<p,q, the classical Lorentz space Lp,q consists of all functions fM0 such that the quantity (2) fLp,q:={(0(t1/pf(t))qdtt)1/q,0<q<,supt>0t1/pf(t),q=,(2) is finite, where f denotes the non-increasing rearrangement of f (see, e.g. [Citation6, Chap. 2, Section 1] or [Citation1, Section 1.4.1]). Note that if 1qp<, then Lp,q is a Banach function space with respect to the Banach function norm Lp,q. On the other hand, L,q={0} for 0<q<.

For fM and λ>0, let χλf:=χ{xRn:|f(x)|>λ}. The quantity fLp,q for 0<p< can also be written as (3) fLp,q={p1/q(0(λχλfLp)qdλλ)1/q,0<q<,supλ>0λχλfLpq=,(3) (see, e.g. [Citation1, Proposition 1.4.9] or [Citation8, Theorem 6.6]).

Bearing in mind formula (Equation3), for a given Banach function space X on Rn and 0<q, we define the abstract Lorentz space Λq(X) built upon X as the set of all functions fM such that (4) fΛq(X):={q1/q(0(λχλfX)qdλλ)1/q,0<q<,supλ>0λχλfX,q=(4) is finite. The normalising factor q1/q in the above definition is taken different from p1/q in (Equation3) to guarantee that for every bounded measurable set ERn, one has χEΛq(X)=χEX for 0<q (see Lemma 2.4 below).

It follows from (Equation3)–(Equation4) that for 0<q and 1p<, one has Λq(Lp)=Lp,q and fΛq(Lp)=(q/p)1/qfLp,q. Further, Λ(L)=L and fΛ(L)=fL. Finally, Λq(L)={0} whenever 0<q<.

The abstract Lorentz space Λ1(X) built upon a rearrangement-invariant Banach function space X is considered, e.g. in [Citation6, Chap. 2, Section 5] (see Definition 5.12 there and Appendix below). For variable Lebesgue spaces Lp(), the spaces Λq(Lp()) were introduced by Kempka and Vybíral in [Citation9, Definition 2.2]. For an arbitrary Banach function space X, the space Λ(X) was considered by Ho in [Citation10, Section 2].

We collect basic properties of abstract Lorentz spaces Λq(X), 0<q, built upon a Banach function space X in the following statement.

Theorem 1.1.

Let X be a Banach function space on Rn.

(a)

If fX, then fΛ(X) and fΛ(X)fX.

(b)

If fΛ1(X), then fX and fXfΛ1(X).

(c)

If 0<qr<, then for all fM, fΛ(X)fΛr(X)(r/q)1/rfΛq(X).

(d)

If 0<q, then Λq(X) is a quasi-Banach space such that for all f,gM, |f||g|a.e.fΛq(X)gΛq(X).

Part (a) and part (d) for q= were proved in [Citation10, Theorem 2.7]. Other statements will be proved in Section 2.

For a given Banach function space X on Rn, denote by A:={X,Λq(X):0<q} the collection of admissible spaces. Suppose that a source space S and a target space T belong to the collection A. We say that a distribution aS belongs to the set MS,T of Fourier multipliers from S to T if aMS,T:=sup{F1aFuTuS:u(SS){0}}<. A function aL is said to belong to the set MS,T0 of Fourier multipliers from S to T if aMS,T0:=sup{F1aFuTuS:u(L2S){0}}<. Here, F±1 are understood as mappings on L2. Since SL2, it is clear that (5) MS,T0MS,TLMS,T(5) and (6) aMS,TaMS,T0forallaMS,T0.(6) Since T is a quasi-norm, it is not difficult to see that the sets MS,T and MS,T0 are quasi-normed linear spaces with respect to the quasi-norms MS,T and MS,T0, respectively.

The quasi-normed space MLp(G),Lp,(G)0 for 1p< and a locally compact abelian group G was studied in [Citation2], where it was shown that it is continuously embedded into L(Γ), where Γ is the dual group of G. As a consequence of this continuous embedding, it was shown there that MLp(G),Lp,(G)0 is a quasi-Banach space.

For yRn and R>0, let B(y,R):={xRn:|xy|<R} be the open ball of radius R centred at y. Following [Citation3, Definition 1.2], we say that a Banach function space X on Rn satisfies the weak doubling property if there exists a number τ>1 such that lim infR(infyRnχB(y,τR)XχB(y,R)X)<. This condition is satisfied, for instance, if X is translation-invariant [Citation3, Corollary 3.5]. Another sufficient condition for the weak doubling property can be stated in terms of the Hardy-Littlewood maximal operator given for fLloc1 by (Mf)(x):=supQx1|Q|Q|f(y)|dy,xRn, where the supremum is taken over all cubes with sides parallel to the coordinate axes that contain x. If M is bounded from X to Λ(X), then X satisfies the weak doubling property. This fact follows from the combination of [Citation3, Lemma 3.3] and [Citation10, Lemma 2.9]. For further discussion of the weak doubling property, see [Citation3, Sections 3.2 and 3.5].

For σR, we will say that a function fM belongs to the space L1,σ if Rn(1+|ξ|)σ|f(ξ)|dξ<. Let L:=σRL1,σ.

Theorem 1.2

Main result

Let X be a Banach function space satisfying the weak doubling property and 0<q. If aLMΛq(X),Λ(X), then (7) aLaMΛq(X),Λ(X).(7) The constant 1 on the right-hand side of inequality (Equation7) is best possible.

Since LL1,σ for σ>n, Theorem 1.2 and inequality (Equation6) imply the following generalisation and refinement of [Citation2, Theorem 1.1] for G=Γ=Rn.

Corollary 1.3.

Let X be a Banach function space satisfying the weak doubling property and 0<q. If aMΛq(X),Λ(X)0, then (8) aLaMΛq(X),Λ(X)0.(8) The constant 1 on the right-hand side of inequality (Equation8) is best possible.

Inequality (Equation8) and Theorem 1.1(c) imply the following.

Corollary 1.4.

Let X be a Banach function space satisfying the weak doubling property. Suppose that S,TA. Then MS,T0 and MS,TL are quasi-Banach spaces with respect to the quasi-norms MS,T0 and MS,T, respectively. Moreover, MS,X0 and MS,XL are Banach spaces with respect to the norms MS,X0 and MS,X, respectively.

The paper is organized as follows. In Section 2, we collect basic properties of abstract Lorentz spaces built upon Banach function spaces and prove Theorem 1.1. In Section 3, we recall some auxiliary results proved in our recent paper [Citation3] and then prove Theorem 1.2 and Corollary 1.4. We conclude the paper stating two open problems in Section 4. In Appendix, we provide a formula for the quasi-norm fΛq(X) in the case of a rearrangement-invariant Banach function space X in terms of the non-increasing rearrangement f, which generalises (Equation2).

It is our pleasure to dedicate this paper to Professor Vladimir Rabinovich on his eightieth birthday.

2. Abstract Lorentz spaces built upon Banach function spaces

2.1. Inclusion Λq(X)M0

Lemma 2.1.

Let X be a Banach function space and 0<q. If fΛq(X), then fM0.

Proof.

If fMM0, then there exists a measurable set ERn of positive measure such that |f(x)|=+ for a.e. xE. For every λ>0, one has χλfχE. Hence fΛq(X){q1/q(0(λχEX)qdλλ)1/q,q<,supλ>0λχEX,q=. Since χEX>0, the right-hand side is infinite, which completes the proof.

2.2. Proof of Theorem 1.1(a),(b)

Since λχλf|f| for all λ>0, one has fΛ(X)fX, which completes the proof of part (a).

Suppose now that fΛ1(X). For λ>0, one has χλf(y)={1,if|f(y)|>λ,0,if|f(y)|λ, and hence 0χλf(y)dλ=0λ<|f(y)|1dλ=|f(y)|,yRn. To proceed further, we need the notion of the associate space X of the Banach function space X defined by a Banach function norm ρ. Its associate norm ρ is defined on M+ by ρ(g):=sup{Rnf(x)g(x)dx:fM+,ρ(f)1}. It is a Banach function norm itself (see [Citation5, Chap. 1, § 1] or [Citation6, Chap. 1, Theorem 2.2]). The Banach function space X determined by the Banach function norm ρ is called the associate space (Köthe dual) of X. The space X is defined similarly.

It follows from the Lorentz-Luxemburg theorem (see [Citation5, Chap. 1, Theorem 4] or [Citation6, Chap. 1, Theorem 2.9]) and Tonelli's theorem (see, e.g. [Citation11, Theorem 5.28]) that fX=fX=suphX,hX1Rn|f(y)||h(y)|dy=suphX,hX1Rn(0χλf(y)dλ)|h(y)|dy=suphX,hX10(Rnχλf(y)|h(y)|dy)dλ0(suphX,hX1Rnχλf(y)|h(y)|dy)dλ=0χλfXdλ=0χλfXdλ=fΛ1(X), which completes the proof of part (b).

2.3. Proof of Theorem 1.1(c)

The proof is almost identical to that of [Citation6, Chap. 4, Proposition 4.2]. Since χλfX is a non-increasing function of λ, one has λχλfX=r1/r(0λ(τχλfX)rdττ)1/rr1/r(0λ(τχτfX)rdττ)1/rfΛr(X). Taking the supremum over all λ>0, one obtains (9) fΛ(X)fΛr(X).(9) If r<, then using (Equation9) with q in place of r, one gets fΛr(X)=r1/r(0(λχλfX)rq+qdλλ)1/rfΛ(X)1q/rr1/r(0(λχλfX)qdλλ)1/rfΛq(X)1q/r(rq)1/rfΛq(X)q/r=(rq)1/rfΛq(X), which completes the proof.

2.4. Quasi-triangle inequality

The quasi-triangle inequality for Λ(X) with the constant 2 can be found in the proof of [Citation10, Theorem 2.7]. We will need a slightly stronger version of this inequality for Λ(X), which is also true for Λq(X) with 0<q<.

Lemma 2.2.

Let X be a Banach function space, 0<q, and let q:=min{1,q}. Then for every f,gΛq(X) and every κ(0,1), one has f+gΛq(X)q1κqfΛq(X)q+1(1κ)qfΛq(X)q.

Proof.

For λ>0 and f,gM0, we have {xRn:|f(x)+g(x)|>λ}{xRn:|f(x)|>κλ}{xRn:|g(x)|>(1κ)λ}. Therefore, χλf+gχκλf+χ(1κ)λg. This inequality and Lemma 2.1 imply that f+gΛ(X)supλ>0λχκλf+χ(1κ)λgXsupλ>0λχκλfX+supλ>0λχ(1κ)λgX=1κsupτ>0τχτfX+11κsupτ>0τχτgX=1κfΛ(X)+11κgΛ(X), as well as f+gΛq(X)q=qq/q(0(λχλf+gX)qdλλ)q/qqq/q(0(λχκλf+χ(1κ)λgX)qdλλ)q/qqq/q(0(λχκλfX+λχ(1κ)λgX)qdλλ)q/q. Using the Minkowski inequality for 1<q< and the subadditivity of the concave function φ(t)=tq for 0<q1 and t0, we further get f+gΛq(X)qqq/q(0(λχκλfX)qdλλ)q/q+qq/q(0(λχ(1κ)λgX)qdλλ)q/q=qq/qκq(0(τχτfX)qdττ)q/q+qq/q(1κ)q(0(τχτgX)qdττ)q/q=1κqfΛq(X)q+1(1κ)qgΛq(X)q, which completes the proof.

2.5. Fatou's lemma

We will need Fatou's lemma for abstract Lorentz spaces Λq(X), 0<q, built upon a Banach function space X.

Lemma 2.3.

Let X be a Banach function space and 0<q. If fkf a.e. as k and lim infkfkΛq(X)<, then fΛq(X) and fΛq(X)lim infkfkΛq(X).

Proof.

For q=, this lemma is proved in [Citation10, Lemma 2.5]. Although for 0<q< the proof is analogous, we provide details here for the reader's convenience. Write hj:=infkj|fk|. Then hj0 for all jN and hj|f| a.e. as j. Therefore, χλhjχλf a.e. as j for every λ>0. By the Fatou property of X (Axiom (A3)), χλfX=limjχλhjXlimjinfkjχλfkX=lim infjχλfjX. Then χλfXqlim infjχλfjXq. Integrating this inequality over (0,) and applying the classical Fatou lemma (see, e.g. [Citation12, Lemma 1.7]), we get (10) fΛq(X)=q1/q(0(λχλfX)qdλλ)1/qq1/q(0(λqlim infjχλfjXq)dλλ)1/qq1/q(lim infj0(λχλfjX)qdλλ)1/q=lim infjfjΛq(X).(10) But f is certainly measurable (being the pointwise limit of a sequence of measurable functions), so (Equation10) shows that f belongs to Λq(X), which completes the proof.

2.6. Proof of Theorem 1.1(d)

For q=, the proof is given in [Citation10, Theorem 2.7]. For 0<q< the proof is similar. Since the proof of completeness of Λ(X) in [Citation10] contains a minor inaccuracy, we provide details for the reader's convenience. It follows immediately from the definition of the quantity Λq(X) that (i) fΛq(X)=0 if and only if f = 0 a.e.; (ii) for every μC, λ>0 and fΛq(X), one has χλμf=χλ/|μ|f. Therefore, making the change of variables τ=λ/|μ|, we get μfΛq(X)=q1/q(0(λχλμfX)qdλλ)1/q=q1/q(0(λχλ/|μ|fX)qdλλ)1/q=q1/q(0(|μ|τχτfX)qdττ)1/q=|μ|fΛq(X). (iii) The quasi-triangle inequality for Λq(X) follows from Lemma 2.2 with κ=1/2. Thus Λq(X) is a quasi-normed space.

If |f||g| a.e., then for all λ>0 one has χλfχλg. By Axiom (A2) for the space X, one gets χλfXχλgX, whence (λχλfX)q(λχλgX)q. Integrating this inequality over (0,), we easily get fΛq(X)gΛq(X).

It remains to show that the quasi-normed space Λq(X) is complete with respect to the quasi-norm Λq(X). Let {fj}j=1 be a Cauchy sequence in Λq(X). It follows from Theorem 1.1(c) that {fj}j=1 is a Cauchy sequence in Λ(X) as well.

For every kN, let Bk={xRn:|x|k}. Axiom (A5) implies that for every kN, there exists Ck(0,) such that for all j,mN and all λ>0, CkχλfjfmXBkχλfjfm(x)dx=|{xRn:|fj(x)fm(x)|>λ}Bk|. Since {fj}j=1 is a Cauchy sequence in Λ(X), it follows from the above inequality and the definition of quasi-norm in Λ(X) that |{xRn:|fj(x)fm(x)|>λ}Bk|0 as j,m for each kN and λ>0. Thus the sequence {fj}j=1 is locally Cauchy in measure. It follows from [Citation13, Proposition 1.2.2(ii)] that there exists a function fM and a subsequence {fjk}k=1 such that fjkf a.e. as k.

Since Λq(X) is a quasi-norm, by the Aoki-Rolewicz theorem (see, e.g. [Citation14, Theorem 1.3]), there exists p>0 such that for any g,hΛq(X), one has g+hΛq(X)pgΛq(X)p+hΛq(X)p. This inequality implies that for all j,mN, |fjΛq(X)pfmΛq(X)p|fjfmΛq(X)p. Therefore, {fjΛq(X)p}j=1 is a Cauchy sequence in R. Hence, the limit limjfjΛq(X) exists and it is finite. Since fjkf a.e. as k, in view of Lemma 2.3, we conclude that fΛq(X).

Fix ε>0. Since {fj}j=1 is a Cauchy sequence in Λq(X), there exists NN such that for all j, m>N, one has fjfmΛq(X)<ε. We have fjkfmffm a.e. as k for all m>N. By Lemma 2.3, for m>N, ffmΛq(X)lim infkfjkfmΛq(X)ε, that is, fmf in Λq(X) as m. Thus, the quasi-normed space Λq(X) is complete with respect to the quasi-norm Λq(X).

2.7. Quasi-norms of characteristic functions of bounded measurable sets

We observe that the quasi-norms of the characteristic function of a bounded measurable set in Λq(X) for all 0<q coincide with its norm in X.

Lemma 2.4.

Let X be a Banach function space and 0<q. For a bounded measurable set ERn, one has χEΛq(X) and χEΛq(X)=χEX.

Proof.

The proof for q= can be found in [Citation10, Lemma 2.6]. The proof for q< is quite similar. We include the details here for the reader's convenience. Since {xRn:|χE(x)|>λ}={E,0<λ<1,,λ1, one has χλχE={χE,0<λ<1,0,λ1. Hence χEΛq(X)=q1/q(0(λχλχEX)qdλλ)1/q=q1/q(01(λχEX)qdλλ)1/q=χEXq1/q(01λq1dλ)1/q=χEX, which completes the proof.

3. Proofs of the main results

3.1. Lemma on approximation at Lebesgue points

Given δ>0 and a function ψ on Rn, we define the function ψδ by ψδ(ξ):=δnψ(ξ/δ),ξRn. Recall that a point xRn is said to be a Lebesgue point of a function fLloc1 if limR0+1|B(x,R)|B(x,R)|f(y)f(x)|dy=0.

Lemma 3.1

[Citation3, Lemma 2.16].

Let σ1,σ2R be such that σ2σ1 and σ2>n. Suppose ψ is a measurable function on Rn satisfying (11) |ψ(ξ)|C(1+|ξ|)σ2for almost allξRn(11) with some constant C(0,). Then for every Lebesgue point ηRn of a function a belonging to the space L1,σ1, one has Rn|a(ξ)a(η)||ψδ(ηξ)|dξ0asδ0.

3.2. The infimum of the doubling constants

For a Banach function space X and τ>1, consider the doubling constant (12) DX,τ:=lim infR(infyRnχB(y,τR)XχB(y,R)X).(12)

Lemma 3.2

[Citation3, Lemma 3.1].

If a Banach function space X satisfies the weak doubling property, then infτ>1DX,τ=1.

3.3. Proof of Theorem 1.2

We follow the scheme of the proof of [Citation3, Theorem 4.1]. Let DX,ϱ be defined for all ϱ>1 by (Equation12). If, for some ϱ>1, the quantity DX,ϱ is infinite, then it is obvious that (13) aLDX,ϱaMΛq(X),Λ(X).(13) Since X satisfies the weak doubling property, there exists ϱ>1 such that DX,ϱ is finite.

Take an arbitrary Lebesgue point ηRn of the function a. Let an even function φC0(Rn) satisfy the following conditions: 0φ1,φ(x)=1 for |x|1,φ(x)=0 for |x|ϱ. Let fδ,η(x):=eiηxφ(δx),xRn,δ>0, and fδ,η,y(x):=fδ,η(xy),yRn. Then (Ffδ,η,y)(ξ)=eiξy(Ffδ,η)(ξ)=eiξyδn(Fφ)(ξηδ)=eiξy(Fφ)δ(ξη)=eiξy(Fφ)δ(ηξ) and (F1aFfδ,η,y)(x)=1(2π)nRnei(xy)ξa(ξ)(Fφ)δ(ηξ)dξ,a(η)fδ,η,y(x)=1(2π)nRnei(xy)ξa(η)(Fφ)δ(ηξ)dξ. Hence, for all x,yRn and δ>0, |(F1aFfδ,η,y)(x)a(η)fδ,η,y(x)|=1(2π)n|Rnei(xy)ξ(a(ξ)a(η))(Fφ)δ(ηξ)dξ|1(2π)nRn|a(ξ)a(η)||(Fφ)δ(ηξ)|dξ. Since FφS and η is a Lebesgue point of a, it follows from Lemma 3.1 that for any ε>0 there exists δε>0 such that for all x,yRn and all δ(0,δε), |(F1aFfδ,η,y)(x)a(η)fδ,η,y(x)|<ε. It is clear that |fδ,η,y|χB(y,1/δ)=χB(y,1/δ). Then the above inequality implies that for all yRn and δ(0,δε), |a(η)|χB(y,1/δ)|F1aFfδ,η,y|+εχB(y,1/δ). Hence, it follows from Lemma 2.4, Theorem 1.1(d) for Λ(X) and Lemma 2.2 that for all κ(0,1), |a(η)|χB(y,1/δ)X=|a(η)|χB(y,1/δ)Λ(X)|F1aFfδ,η,y|+εχB(y,1/δ)Λ(X)1κF1aFfδ,η,yΛ(X)+ε1κχB(y,1/δ)Λ(X)1κaMΛq(X),Λ(X)fδ,η,yΛq(X)+ε1κχB(y,1/δ)X. Taking into account that |fδ,y,η|χB(y,ϱ/δ), it follows from the above inequality, Theorem 1.1(d) for Λq(X) and Lemma 2.4 that for all κ(0,1), (14) |a(η)|χB(y,1/δ)X1κaMΛq(X),Λ(X)χB(y,ϱ/δ)Λq(X)+ε1κχB(y,1/δ)X=1κaMΛq(X),Λ(X)χB(y,ϱ/δ)X+ε1κχB(y,1/δ)X.(14) Since DX,ϱ<, the definition of DX,ϱ given in (Equation12) implies that there exist δ(0,δε) and yRn such that χB(y,ϱ/δ)XχB(y,1/δ)XDX,ϱ+ε. Choosing these δ and y, and dividing both sides of inequality (Equation14) by χB(y,1/δ)X, we get |a(η)|1κ(DX,ϱ+ε)aMΛq(X),Λ(X)+ε1κfor allε>0,κ(0,1). Passing to the limit as ε0, we obtain for all Lebesgue points ηRn of the function a, |a(η)|1κDX,ϱaMΛq(X),Λ(X)for allκ(0,1). Passing to the limit as κ1, we get |a(η)|DX,ϱaMΛq(X),Λ(X). Since aL1,σLloc1, almost all points ηRn are Lebesgue points of the function a in view of the Lebesgue differentiation theorem (see, e.g. [Citation1, Corollary 2.1.16 and Exercise 2.1.10]). Therefore aL and inequality (Equation13) holds for all ϱ>1. It is now left to apply Lemma 3.2.

Suppose that there is a constant DX>0 such that aLDXaMΛq(X),Λ(X) for all aLMΛq(X),Λ(X). It follows from Theorem 1.1(c) that a01 belongs to MΛq(X),Λ(X) and a0MΛq(X),Λ(X)1. Since LL, we conclude that 1=a0LDXa0MΛq(X),Λ(X)DX. So, the constant DX=1 in the estimate aLaMΛq(X),Λ(X) is best possible.

3.4. Proof of Corollary 1.4

Let S,TA. We already know that MS,T0 and MS,T are quasi-normed spaces with respect to the quasi-norms MS,T0 and MS,T, respectively. Moreover, MS,X0 and MS,X are norms. So, it remains only to show that the spaces MS,T0 and MS,TL are complete with respect to MS,T0 and MS,T, respectively.

It follows from Theorem 1.1(a),(c) that (15) aMS,Λ(X)0aMS,T0for allS,TA.(15) By definition, every SA is equal to either X or Λq(X) for some q(0,]. Taking q = 1 in the former case and using Theorem 1.1(b), we conclude that (16) aMΛq(X),Λ(X)0aMS,Λ(X)0for someq(0,].(16) It follows from (Equation15), (Equation16) that for every SA there exists q(0,] such that (17) aMΛq(X),Λ(X)0aMS,T0(17) for all TA.

Suppose that {am}m=1 is a Cauchy sequence in MS,T0. It follows from Corollary 1.3 and inequality (Equation17) that {am}m=1 is a Cauchy sequence in L. Since L is complete, there exists aL such that akaL0 as k. Hence, in view of [Citation1, Theorem 2.5.10], for each uL2X, F1akFuF1aFuL2akaLuL20ask. In view of the standard fact on Lp spaces (see, e.g. [Citation11, Section 7.23]), there is a subsequence {aks}s=1 such that F1aksFuF1aFu a.e. as s.

Fix ε>0. Since {am}m=1 is a Cauchy sequence in MS,T0, there exists NN such that for all m, k>N and all uL2S, F1amFuF1akFuTεuS. Hence, for all m>N, all sN such that ks>N, and all uL2S, (18) F1amFuF1aksFuTεuS.(18) Since for all m>N and uL2S, F1amFuF1aksFuF1amFuF1aFua.e. ass, applying the Fatou lemma for T (Lemma 2.3 if T is one of the spaces Λq(X) with 0<q or [Citation5, Chap. 1, Lemma 1(e)], [Citation6, Chap. 1, Lemma 1.5] if T=X) to inequality (Equation18), we get for all m>N and uL2S, F1amFuF1aFuTlim infsF1amFuF1aksFuTεuS. Hence for m>N we have amaMS,T0ε, which completes the proof in the case of MS,T0.

The proof of the completeness of the space MS,TL is analogous.

4. Concluding remarks and open problems

Recall that if 1p<, then Λ1(Lp)=Lp,1 and Λ(Lp)=Lp,. The Marcinkiewicz space L1, is not normable (see, e.g. [Citation15, Chap. V, 5.12]). On the other hand, one can equip the Lorentz space Lp,, 1<p<, with an equivalent norm and turn it into a Banach function space (see, e.g. [Citation6, Chap. 4, Theorem 4.6]). The problem of normability of more general Lorentz-type spaces was considered, for instance, in [Citation16, Section 2.5].

Problem 4.1

Describe Banach function spaces X for which the abstract Lorentz spaces Λq(X) with 1q can be equipped with equivalent Banach function norms.

The answer does not seem to be known even for Lebesgue spaces with variable exponents Lp() (see [Citation9, Section 2.3]).

If the space Λq(X) admits an equivalent Banach function norm, one can apply to Λq(X) results from [Citation3] (it follows from Lemma 2.4 that Λq(X) satisfies the weak doubling property if and only if X does). This would provide results on Fourier multipliers acting from SΛq(X) to Λq(X), while Theorem 1.2 is concerned with an a priori wider class of Fourier multipliers acting from SΛq(X) to Λ(X) because, in view of Theorem 1.1(c), one has aMΛq(X),Λ(X)aMΛq(X),Λr(X) for 0<q,r.

Note that for Fourier multipliers acting from SX to X, an analogue of Theorem 1.2 holds without the a priori assumption aL (see [Citation3, Theorem 1.3 and Section 4.2]). Unfortunately, we don't know whether or not it can also be removed from Theorem 1.2.

Problem 4.2

Prove (or disprove) Theorem 1.2 for arbitrary aMΛq(X),Λ(X)S.

Disclosure statement

No potential conflict of interest was reported by the author(s).

Additional information

Funding

This work was supported by national funds through the FCT – Fundação para a Ciência e a Tecnologia, I.P. (Portuguese Foundation for Science and Technology) within the scope of the project UIDB/00297/2020 (Centro de Matemática e Aplicações).

References

Appendix. Abstract Lorentz spaces built upon rearrangement-invariant Banach function spaces

Following [Citation6, Chap. 2, Definitions 1.1 and 1.2], the distribution function Df of a function fM0 is given by Df(λ):=|{xRn:|f(x)|>λ}|,λ0. Two functions f,gM0 are said to be equimeasurable if Df(λ)=Dg(λ) for all λ0. The non-increasing rearrangement of a function fM0 is the function f defined by f(t):=inf{λ0:Df(λ)t},t0, with the convention that inf= (see, e.g. [Citation6, Chap. 2, Definition 1.5]).

A Banach function norm ρ:M[0,] is said to be rearrangement-invariant if ρ(f)=ρ(g) for every pair of equimeasurable functions f,gM0+. In that case, the Banach function space X=X(ρ) generated by ρ is said to be a rearrangement-invariant Banach function space (see [Citation6, Chap. 2, Definition 4.1]).

The fundamental function φX of a rearrangement-invariant Banach function space X is defined by φX(t):=χEX,t[0,), where ERn is a measurable set with |E|=t (see, e.g. [Citation6, Chap. 2, Definition 5.1]).

Lemma A.1.

Let 0<q and X be a rearrangement-invariant Banach function space. For every function fΛq(X), one has (A1) fΛq(X)={(0(f(t))qdφXq(t))1/q,q<,supt>0(f(t)φX(t)),q=.(A1)

Proof.

It follows from the definition of φX that (A2) χ{xRn:|f(x)|>λ}X=φX(Df(λ)),λ0.(A2) Take any t>0 and suppose that f(t)>0. It follows from [Citation6, Chap. 2, formula (1.10)] that for every ε>0 there exists λ>f(t)1+ε such that Df(λ)>t. Since φX is non-decreasing (see [Citation6, Chap. 2, Corollary 5.3]), it follows from (EquationA2) that f(t)φX(t)(1+ε)λφX(Df(λ))(1+ε)supλ>0(λφX(Df(λ)))=(1+ε)fΛ(X). Since ε>0 is arbitrary, one gets f(t)φX(t)fΛ(X), and it is clear that this inequality holds in the case f(t)=0 as well. So, (A3) supt>0(f(t)φX(t))fΛ(X).(A3) Now, fix any λ>0 and suppose that Df(λ)>0. Take any ε(0,Df(λ)/2] and set t:=Df(λ)ε. Since Df(λ)>t, it follows from [Citation6, Chap. 2, formula (1.10)] that λf(t) and hence λφX(Df(λ))f(t)φX(t+ε)f(t)(φX(t+ε)φX(t))+supt>0(f(t)φX(t))f(Df(λ)/2)sups[Df(λ)/2,Df(λ)](φX(s+ε)φX(s))+supt>0(f(t)φX(t)). Sending ε to 0 and using the fact that φX is uniformly continuous on the segment [Df(λ)/2,3Df(λ)/2] (see [Citation6, Chap. 2, Corollary 5.3]), one gets (A4) λφX(Df(λ))supt>0(f(t)φX(t)).(A4) If Df(λ)=0, then φX(Df(λ))=φX(0)=0 (see [Citation6, Chap. 2, Corollary 5.3]) and (EquationA4) remains true. It follows from (EquationA2) and (EquationA4) that (A5) fΛ(X)=supλ>0(λφX(Df(λ)))supt>0(f(t)φX(t)).(A5) Combining (EquationA3) and (EquationA5), we arrive at (EquationA1) in the case q=.

Now, suppose that q<. Using (EquationA2), the equality Df=Df (see, e.g. [Citation6, Chap. 2, formula (1.19)]), monotonicity of f, the equality φX(0)=0, and continuity of φX (see [Citation6, Chap. 2, Corollary 5.3]), one gets 0(f(s))qdφXq(s)=0(0(f(s))q1dτ)dφXq(s)=0(0τ<(f(s))q1dτ)dφXq(s)=0({s0:f(s)>τ1/q}1dφXq(s))dτ=0(0Df(τ1/q)1dφXq(s))dτ=0φXq(Df(τ1/q))dτ=0φXq(Df(λ))dλq=q0(λφX(Df(λ)))qdλλ=fΛq(X)q, which immediately implies (EquationA1) for q<.