1 Introduction

Let \(C({\overline{{\mathbb {R}}}})\) be the \(C^*\)-algebra of all continuous functions on the two-point compactification of the real line \({\overline{{\mathbb {R}}}}=[-\infty ,+\infty ]\) and

$$\begin{aligned} C({{\dot{{\mathbb {R}}}}})=\{f\in C({\overline{{\mathbb {R}}}}):f(-\infty )=f(+\infty )\}, \end{aligned}$$

where \({{\dot{{\mathbb {R}}}}}={\mathbb {R}}\cup \{\infty \}\) is the one-point compactification of the real line. Let APP denote the set of all almost periodic polynomials, that is, finite sums of the form \(\sum _{\lambda \in \varLambda }c_\lambda e_\lambda \), where

$$\begin{aligned} e_\lambda (x):=e^{i\lambda x}, \quad x\in {\mathbb {R}}, \end{aligned}$$

\(c_\lambda \in {\mathbb {C}}\) and \(\varLambda \subset {\mathbb {R}}\) is a finite subset of \({\mathbb {R}}\). The smallest closed subalgebra of \(L^\infty ({\mathbb {R}})\) that contains APP is denoted by AP and called the algebra of (uniformly) almost periodic functions. Sarason [36] introduced the algebra of semi-almost periodic functions as the smallest closed subalgebra of \(L^\infty ({\mathbb {R}})\) that contains AP and \(C({\overline{{\mathbb {R}}}})\):

$$\begin{aligned} SAP:={\text {alg}}_{L^\infty ({\mathbb {R}})}\{AP,C({\overline{{\mathbb {R}}}})\}. \end{aligned}$$

It is not difficult to see that AP and SAP are \(C^*\)-subalgebras of \(L^\infty ({\mathbb {R}})\).

Theorem 1.1

(Sarason [36], see also [10, Theorem 1.21]) Let \(u\in C({\overline{{\mathbb {R}}}})\) be any function for which \(u(-\infty )=0\) and \(u(+\infty )=1\). If \(a\in SAP\), then there exist \(a_l,a_r\in AP\) and \(a_0\in C({{\dot{{\mathbb {R}}}}})\) such that \(a_0(\infty )=0\) and

$$\begin{aligned} a=(1-u)a_l+ua_r+a_0. \end{aligned}$$
(1.1)

The functions \(a_l,a_r\) are uniquely determined by a and independent of the particular choice of u. The maps \(a\mapsto a_l\) and \(a\mapsto a_r\) are \(C^*\)-algebra homomorphisms of SAP onto AP.

The uniquely determined function \(a_l\) (resp. \(a_r\)) is called the left (resp. right) almost periodic representative of the semi-almost periodic function a.

Let \(\mathcal{F}:L^2({\mathbb {R}})\rightarrow L^2({\mathbb {R}})\) denote the Fourier transform:

$$\begin{aligned} ( \mathcal{F}f)(x):={\widehat{f}}(x):=\int _{\mathbb {R}}f(t)e^{itx}\,\mathrm{d}t, \quad x\in {\mathbb {R}}, \end{aligned}$$

and let \(\mathcal{F}^{-1}:L^2({\mathbb {R}})\rightarrow L^2({\mathbb {R}})\) be the inverse of \(\mathcal{F}\),

$$\begin{aligned} ( \mathcal{F}^{-1}g)(t)=\frac{1}{2\pi }\int _{\mathbb {R}}g(x)e^{-itx}\,\mathrm{d} x, \quad t\in {\mathbb {R}}. \end{aligned}$$

It is well known that the Fourier convolution operator

$$\begin{aligned} W^0(a):=\mathcal{F}^{-1}a\mathcal{F} \end{aligned}$$
(1.2)

is bounded on the space \(L^2({\mathbb {R}})\) for every \(a\in L^\infty ({\mathbb {R}})\).

Let \(X({\mathbb {R}})\) be a separable Banach function space (see Sect. 2.1 for the definition and some properties of Banach function spaces). Then \(L^2({\mathbb {R}})\cap X({\mathbb {R}})\) is dense in \(X({\mathbb {R}})\) (see, e.g., [15, Lemma 2.1]). A function \(a\in L^\infty ({\mathbb {R}})\) is called a Fourier multiplier on \(X({\mathbb {R}})\) if the convolution operator \(W^0(a)\) defined by (1.2) maps the set \(L^2({\mathbb {R}})\cap X({\mathbb {R}})\) into the space \(X({\mathbb {R}})\) and extends to a bounded linear operator on \(X({\mathbb {R}})\). The function a is called the symbol of the Fourier convolution operator \(W^0(a)\). The set \({\mathcal {M}}_{X({\mathbb {R}})}\) of all Fourier multipliers on \(X({\mathbb {R}})\) is a unital normed algebra under pointwise operations and the norm:

$$\begin{aligned} \left\| a\right\| _{{\mathcal {M}}_{X({\mathbb {R}})}}:=\left\| W^0(a)\right\| _{{\mathcal {B}}(X({\mathbb {R}}))}, \end{aligned}$$

where \({\mathcal {B}}(X({\mathbb {R}}))\) denotes the Banach algebra of all bounded linear operators on the space \(X({\mathbb {R}})\).

Note that the Lebesgue spaces \(L^p({\mathbb {R}})\), \(1\le p\le \infty \), constitute the simplest example of Banach function spaces. Motivated by the work of Duduchava and Saginashvili [14], Karlovich and Spitkovsky [29] (see also [10, Section 19.1]) introduced the algebra \(SAP_{L^p({\mathbb {R}})}\) of semi-almost periodic Fourier multipliers on the Lebesgue spaces \(L^p({\mathbb {R}})\), \(1<p<\infty \), and proved an analogue of Sarason’s Theorem 1.1 for \(SAP_{L^p({\mathbb {R}})}\) (see [29, Lemma 3.1(iv)] and [10, Proposition 19.3]).

We should mention that, after Sarason’s pioneering paper [36], various classes of Toeplitz and convolution type operators involving semi-almost periodic functions were studied on various function spaces, for instance, by Saginashvili [35], Grudsky [19]; Böttcher et al. [3,4,5,6, 8,9,10]; Nolasco and Castro [32, 33]; Bogveradze and Castro [2]; the second author and Spitkovsky [25].

Let \({\mathfrak {M}}({\mathbb {R}})\) denote the set of all measurable complex-valued Lebesgue measurable functions on \({\mathbb {R}}\). As usual, we identify two functions on \({\mathbb {R}}\) which are equal almost everywhere. A measurable function \(w:{\mathbb {R}}\rightarrow [0,\infty ]\) is called a weight if the set \(w^{-1}(\{0,\infty \})\) has measure zero. For \(1<p<\infty \), the Muckenhoupt class \(A_p({\mathbb {R}})\) is defined as the class of all weights \(w:{\mathbb {R}}\rightarrow [0,\infty ]\) such that \(w\in L_{\mathrm{loc}}^p({\mathbb {R}})\), \(w^{-1}\in L_{\mathrm{loc}}^{p'}({\mathbb {R}})\) and

$$\begin{aligned} \sup _I \left( \frac{1}{|I|}\int _I w^p(x)\,\mathrm{d}x\right) ^{1/p} \left( \frac{1}{|I|}\int _I w^{-p'}(x)\,\mathrm{d}x\right) ^{1/p'} <\infty , \end{aligned}$$
(1.3)

where \(1/p+1/p'=1\) and the supremum is taken over all intervals \(I\subset {\mathbb {R}}\) of finite length |I|. Since \(w\in L_{\mathrm{loc}}^p({\mathbb {R}})\) and \(w^{-1}\in L_{\mathrm{loc}}^{p'}({\mathbb {R}})\), the weighted Lebesgue space

$$\begin{aligned} L^p({\mathbb {R}},w):=\{f\in {\mathfrak {M}}({\mathbb {R}}):fw\in L^p({\mathbb {R}})\} \end{aligned}$$

is a separable Banach function space (see, e.g., [26, Lemma 2.4]) with the norm:

$$\begin{aligned} \Vert f\Vert _{L^p({\mathbb {R}},w)}:=\left( \int _{\mathbb {R}}|f(x)|^p w^p(x)\,\mathrm{d}x\right) ^{1/p}. \end{aligned}$$

Note that if \(w\in A_p({\mathbb {R}})\), then it may happen that the function \(e_\lambda \) does not belong to \({\mathcal {M}}_{L^p({\mathbb {R}},w)}\) for some \(\lambda \in {\mathbb {R}}\). Hence,order to generalize Theorem 1.1 to the setting of weighted Lebesgue spaces \(L^p({\mathbb {R}},w)\), one has to restrict the study to a narrower class of weights. For \(1<p<\infty \), let

$$\begin{aligned} A_p^0({\mathbb {R}}):=\left\{ w\in A_p({\mathbb {R}})\ :\ v_\lambda =\frac{w(\cdot +\lambda )}{w(\cdot )}\in L^\infty ({\mathbb {R}}) \text{ for } \text{ all } \lambda \in {\mathbb {R}}\right\} . \end{aligned}$$

For a weight \(w\in A_p^0({\mathbb {R}})\), Karlovich and Loreto Hernández defined the algebra \(SAP_{L^p({\mathbb {R}},w)}\) of semi-almost periodic Fourier multipliers on the weighted Lebesgue space \(L^p({\mathbb {R}},w)\) and proved an analogue of Theorem 1.1 in this setting (see [27, Theorem 3.1]). The aim of this paper is to extend this result to the setting of separable rearrangement-invariant Banach function spaces with suitable Muckenhoupt weights.

It is well known that the Lebesgue spaces \(L^p({\mathbb {R}})\), \(1\le p\le \infty \), fall in the class of rearrangement-invariant Banach function spaces. Other classical examples of rearrangement-invariant Banach function spaces are Orlicz spaces \(L^\varPhi ({\mathbb {R}})\) and Lorentz spaces \(L^{p,q}({\mathbb {R}})\), \(1\le p,q\le \infty \). For a rearrangement-invariant Banach function space \(X({\mathbb {R}})\), its Boyd indices \(\alpha _X,\beta _X\) are important interpolation characteristics. In particular, \(\alpha _{L^p}=\beta _{L^p}=1/p\) for \(1\le p\le \infty \). In general, \(0\le \alpha _X\le \beta _X\le 1\) and it may happen that \(\alpha _X<\beta _X\). We postpone formal definitions of rearrangement-invariant Banach function spaces and their Boyd indices until Sects. 2.22.3 and refer to [1, Chap. 3] and [30, Chap. 2] for the detailed study of these concepts.

Let \(X({\mathbb {R}})\) be a separable rearrangement-invariant Banach function space with the Boyd indices \(\alpha _X,\beta _X\) satisfying \(0<\alpha _X,\beta _X<1\). Suppose that a weight w belongs to \(A_{1/\alpha _X}({\mathbb {R}})\cap A_{1/\beta _X}({\mathbb {R}})\). Then

$$\begin{aligned} X({\mathbb {R}},w):=\{f\in {\mathfrak {M}}({\mathbb {R}}):fw\in X({\mathbb {R}})\} \end{aligned}$$

is a separable Banach function space (see Lemma 2.3(b) below). Suppose that \(a:{\mathbb {R}}\rightarrow {\mathbb {C}}\) is a function of finite total variation V(a) given by

$$\begin{aligned} V(a):=\sup \sum _{k=1}^n |a(x_k)-a(x_{k-1})|, \end{aligned}$$

where the supremum is taken over all partitions of \({\mathbb {R}}\) of the form

$$\begin{aligned} -\infty<x_0<x_1<\cdots<x_n<+\infty \end{aligned}$$

with \(n\in {\mathbb {N}}\). The set \(V({\mathbb {R}})\) of all functions of finite total variation on \({\mathbb {R}}\) with the norm

$$\begin{aligned} \Vert a\Vert _V:=\Vert a\Vert _{L^\infty ({\mathbb {R}})}+V(a) \end{aligned}$$

is a unital non-separable Banach algebra. It follows from [21, Corollary 2.2] that there exists a constant \(c_{X({\mathbb {R}},w)}\in (0,\infty )\) such that for all \(a\in V({\mathbb {R}})\),

$$\begin{aligned} \Vert a\Vert _{{\mathcal {M}}_{X({\mathbb {R}},w)}}\le c_{X({\mathbb {R}},w)}\Vert a\Vert _{V({\mathbb {R}})}. \end{aligned}$$
(1.4)

This inequality is usually called a Stechkin-type inequality (see, e.g., [13, Theorem 2.11] and [10, Theorem 17.1] for the case of Lebesgue spaces and Lebesgue spaces with Muckenhoupt weights, respectively). Let \(C_{X({\mathbb {R}},w)}({{\dot{{\mathbb {R}}}}})\) and \(C_{X({\mathbb {R}},w)}({\overline{{\mathbb {R}}}})\) denote the closures of \(C({{\dot{{\mathbb {R}}}}})\cap V({\mathbb {R}})\) and \(C({\overline{{\mathbb {R}}}})\cap V({\mathbb {R}})\) with respect to the norm of \({\mathcal {M}}_{X({\mathbb {R}},w)}\), respectively.

If \(w\in A_{1/\alpha _X}^0({\mathbb {R}})\cap A_{1/\beta _X}^0({\mathbb {R}})\), then \(APP\subset {\mathcal {M}}_{X({\mathbb {R}},w)}\) (see Corollary 5.2 below). Because of this observation, we will refer to \(A_{1/\alpha _X}^0({\mathbb {R}})\cap A_{1/\beta _X}^0({\mathbb {R}})\) as the class of suitable Muckenhoupt weights. By \(AP_{X({\mathbb {R}},w)}\) we denote the closure of APP with respect to the norm of \({\mathcal {M}}_{X({\mathbb {R}},w)}\). Finally, let \(SAP_{X({\mathbb {R}},w)}\) be the smallest closed subalgbera of \({\mathcal {M}}_{X({\mathbb {R}},w)}\) that contains the algebras \(AP_{X({\mathbb {R}},w)}\) and \(C_{X({\mathbb {R}},w)}({\overline{{\mathbb {R}}}})\):

$$\begin{aligned} SAP_{X({\mathbb {R}},w)}={\text {alg}}_{{\mathcal {M}}_{X({\mathbb {R}},w)}} \left\{ AP_{X({\mathbb {R}},w)},C_{X({\mathbb {R}},w)}({\overline{{\mathbb {R}}}})\right\} . \end{aligned}$$

In this paper we present a self-contained proof of the following result.

Theorem 1.2

(Main result) Let \(X({\mathbb {R}})\) be a separable rearrangement-invariant Banach function space with the Boyd indices satisfying \(0<\alpha _X\), \(\beta _X<1\). Suppose that \(w\in A_{1/\alpha _X}^0({\mathbb {R}})\cap A_{1/\beta _X}^0({\mathbb {R}})\). Let \(u\in C({\overline{{\mathbb {R}}}})\) be any real-valued monotonically increasing function such that \(u(-\infty )=0\) and \(u(+\infty )=1\). Then for every function \(a\in SAP_{X({\mathbb {R}},w)}\) there exist functions \(a_l,a_r\in AP_{X({\mathbb {R}},w)}\) and a function \(a_0\in C_{X({\mathbb {R}},w)}({{\dot{{\mathbb {R}}}}})\) such that \(a_0(\infty )=0\) and (1.1) holds. The functions \(a_l\), \(a_r\) are uniquely determined by the function a and are independent of the particular choice of the function u. The maps \(a\mapsto a_l\) and \(a\mapsto a_r\) are continuous Banach algebra homomorphisms of \(SAP_{X({\mathbb {R}},w)}\) onto \(AP_{X({\mathbb {R}},w)}\) of norm 1.

The paper is organized as follows. In Sect. 2, we collect definitions and properties of rearrangement-invariant Banach functions spaces and their Boyd indices \(\alpha _X,\beta _X\). Further, we discuss properties of weighted rearrangement-invariant spaces \(X({\mathbb {R}},w)\) and state several results about general Fourier multipliers on \(X({\mathbb {R}},w)\) for weights w belonging to the intersection of the Muckenhoupt classes \(A_{1/\alpha _X}({\mathbb {R}})\cap A_{1/\beta _X}({\mathbb {R}})\).

In Sect. 3, we show that, under the assumption \(w\in A_{1/\alpha _X}({\mathbb {R}})\cap A_{1/\beta _X}({\mathbb {R}})\), the set of continuous Fourier multipliers vanishing at infinity on the space \(X({\mathbb {R}},w)\) coincides with the closure of the set of all smooth compactly supported functions with respect to the norm of \({\mathcal {M}}_{X({\mathbb {R}},w)}\).

Relying on the results of the previous section, in Sect. 4, we show that \(C_{X({\mathbb {R}},w)}({{\dot{{\mathbb {R}}}}})=C_{X({\mathbb {R}},w)}({\overline{{\mathbb {R}}}})\cap C({{\dot{{\mathbb {R}}}}})\) and that the algebra \(C_{X({\mathbb {R}},w)}({{\dot{{\mathbb {R}}}}})\) is contained in the algebra \(SO_{X({\mathbb {R}},w)}\) of slowly oscillating Fourier multipliers (see [21]).

In Sect. 5, we show that if \(w\in A_{1/\alpha _X}^0({\mathbb {R}})\cap A_{1/\beta _X}^0({\mathbb {R}})\), then the set of almost periodic polynomials APP is contained in \({\mathcal {M}}_{X({\mathbb {R}},w)}\). We give an example of a nontrivial weight in \(A_{1/\alpha _X}^0({\mathbb {R}})\cap A_{1/\beta _X}^0({\mathbb {R}})\) (based on an example from [27]). Further, we show that the product of an almost periodic Fourier multiplier and a continuous Fourier multiplier vanishing at infinity is a continuous Fourier multiplier vanishing at infinity.

Section 6 is devoted to the proof of the main result. We show that the set \({\mathcal {A}}_u\) of functions of the form (1.1) with \(a_l,a_r\in AP_{X({\mathbb {R}},w)}\) and \(a_0\in C_{X({\mathbb {R}},w)}({{\dot{{\mathbb {R}}}}})\) such that \(a_0(\infty )=0\) forms an algebra, and that the mappings \(a\mapsto a_l\) and \(a\mapsto a_r\) are algebraic homomorphisms of \({\mathcal {A}}_u\) onto \(AP_{X({\mathbb {R}},w)}\). We prove that

$$\begin{aligned} \Vert a_l\Vert _{{\mathcal {M}}_{X({\mathbb {R}},w)}}\le \Vert a\Vert _{{\mathcal {M}}_{X({\mathbb {R}},w)}}, \quad \Vert a_r\Vert _{{\mathcal {M}}_{X({\mathbb {R}},w)}}\le \Vert a\Vert _{{\mathcal {M}}_{X({\mathbb {R}},w)}}, \quad a\in {\mathcal {A}}_u, \end{aligned}$$
(1.5)

which implies that the algebra \({\mathcal {A}}_u\) is closed. Since the closure of \({\mathcal {A}}_u\) with respect to the norm of \({\mathcal {M}}_{X({\mathbb {R}},w)}\) coincides with \(SAP_{X({\mathbb {R}},w)}\), we conclude that \({\mathcal {A}}_u\) is equal to \(SAP_{X({\mathbb {R}},w)}\). Moreover, inequalities (1.5) mean that \(a\mapsto a_l\) and \(a\mapsto a_r\) are Banach algebra homomorphisms of \(SAP_{X(|R,w)}\) onto \(AP_{X({\mathbb {R}},w)}\) of norm 1.

2 Preliminaries

2.1 Banach function spaces

Let \({\mathbb {R}}_+:=(0,\infty )\) and \({\mathbb {S}}\in \{{\mathbb {R}}_+,{\mathbb {R}}\}\). The set of all Lebesgue measurable complex-valued functions on \({\mathbb {S}}\) is denoted by \({\mathfrak {M}}({\mathbb {S}})\). Let \({\mathfrak {M}}^+({\mathbb {S}})\) be the subset of functions in \({\mathfrak {M}}({\mathbb {S}})\) whose values lie in \([0,\infty ]\). The Lebesgue measure of a measurable set \(E\subset {\mathbb {S}}\) is denoted by |E| and its characteristic function is denoted by \(\chi _E\). Following [1, Chap. 1, Definition 1.1], a mapping \(\rho :{\mathfrak {M}}^+({\mathbb {S}})\rightarrow [0,\infty ]\) is called a Banach function norm if, for all functions \(f,g, f_n \ (n\in {\mathbb {N}})\) in \({\mathfrak {M}}^+({\mathbb {S}})\), for all constants \(a\ge 0\), and for all measurable subsets E of \({\mathbb {S}}\), the following properties hold:

$$\begin{aligned} \mathrm{(A1)}\;&\rho (f)=0 \Leftrightarrow f=0\ \text{ a.e.}, \quad \rho (af)=a\rho (f), \quad \rho (f+g) \le \rho (f)+\rho (g),\\ \mathrm{(A2)}\;&0\le g \le f \ \text{ a.e. } \ \Rightarrow \ \rho (g) \le \rho (f) \quad \text{(the } \text{ lattice } \text{ property)}, \\ \mathrm{(A3)}\;&0\le f_n \uparrow f \ \text{ a.e. } \ \Rightarrow \ \rho (f_n) \uparrow \rho (f)\quad \text{(the } \text{ Fatou } \text{ property)},\\ \mathrm{(A4)}\;&|E|<\infty \Rightarrow \rho (\chi _E)<\infty ,\\ \mathrm{(A5)}\;&|E|<\infty \Rightarrow \int _E f(x)\,\mathrm{d}x \le C_E\rho (f) \end{aligned}$$

with \(C_E \in (0,\infty )\) which may depend on E and \(\rho \) but is independent of f. When functions differing only on a set of measure zero are identified, the set \(X({\mathbb {S}})\) of all functions \(f\in {\mathfrak {M}}({\mathbb {S}})\) for which \(\rho (|f|)<\infty \) is called a Banach function space. For each \(f\in X({\mathbb {S}})\), the norm of f is defined by

$$\begin{aligned} \left\| f\right\| _{X({\mathbb {S}})} :=\rho (|f|). \end{aligned}$$

Under the natural linear space operations and under this norm, the set \(X({\mathbb {S}})\) becomes a Banach space (see [1, Chap. 1, Theorems 1.4 and 1.6]). If \(\rho \) is a Banach function norm, its associate norm \(\rho '\) is defined on \({\mathfrak {M}}^+({\mathbb {S}})\) by

$$\begin{aligned} \rho '(g):=\sup \left\{ \int _{{\mathbb {S}}} f(x)g(x)\,dx \ : \ f\in {\mathfrak {M}}^+({\mathbb {S}}), \ \rho (f) \le 1 \right\} , \quad g\in {\mathfrak {M}}^+({\mathbb {S}}). \end{aligned}$$

It is a Banach function norm itself [1, Chap. 1, Theorem 2.2]. The Banach function space \(X'({\mathbb {R}})\) determined by the Banach function norm \(\rho '\) is called the associate space (Köthe dual) of \(X({\mathbb {S}})\). The associate space \(X'({\mathbb {S}})\) is naturally identified with a subspace of the (Banach) dual space \([X({\mathbb {S}})]^*\).

2.2 Rearrangement-invariant Banach function spaces

Suppose that \({\mathbb {S}}\in \{{\mathbb {R}},{\mathbb {R}}_+\}\). Let \({\mathfrak {M}}_0({\mathbb {S}})\) and \({\mathfrak {M}}_0^+({\mathbb {S}})\) be the classes of a.e. finite functions in \({\mathfrak {M}}({\mathbb {S}})\) and \({\mathfrak {M}}^+({\mathbb {S}})\), respectively. The distribution function \(\mu _f\) of a function \(f\in {\mathfrak {M}}_0({\mathbb {S}})\) is given by

$$\begin{aligned} \mu _f(\lambda ):= \big |\{x\in {\mathbb {S}}:|f(x)|>\lambda \}\big |, \quad \lambda \ge 0. \end{aligned}$$

Two functions \(f,g\in {\mathfrak {M}}_0({\mathbb {S}})\) are said to be equimeasurable if \(\mu _f(\lambda )=\mu _g(\lambda )\) for all \(\lambda \ge 0\). The non-increasing rearrangement of \(f\in {\mathfrak {M}}_0({\mathbb {S}})\) is the function defined by

$$\begin{aligned} f^*(t):=\inf \{\lambda :\mu _f(\lambda )\le t\},\quad t\ge 0. \end{aligned}$$

We here use the standard convention that \(\inf \emptyset =+\infty \).

A Banach function norm \(\rho :{\mathfrak {M}}^+({\mathbb {S}}) \rightarrow [0,\infty ]\) is called rearrangement-invariant if for every pair of equimeasurable functions \(f,g \in {\mathfrak {M}}^+_0({\mathbb {S}})\) the equality \(\rho (f)=\rho (g)\) holds. In that case, the Banach function space \(X({\mathbb {S}})\) generated by \(\rho \) is said to be a rearrangement-invariant Banach function space (or simply rearrangement-invariant space). Lebesgue, Orlicz, and Lorentz spaces are classical examples of rearrangement-invariant Banach function spaces (see, e.g., [1] and the references therein). By [1, Chap. 2, Proposition 4.2], if a Banach function space \(X({\mathbb {S}})\) is rearrangement-invariant, then its associate space \(X'({\mathbb {S}})\) is rearrangement-invariant, too.

2.3 Boyd indices

Suppose \(X({\mathbb {R}})\) is a rearrangement-invariant Banach function space generated by a rearrangement-invariant Banach function norm \(\rho \). In this case, the Luxemburg representation theorem [1, Chap. 2, Theorem 4.10] provides a unique rearrangement-invariant Banach function norm \({\overline{\rho }}\) over the half-line \({\mathbb {R}}_+\) equipped with the Lebesgue measure, defined by

$$\begin{aligned} {\overline{\rho }}(h):= \sup \left\{ \int _{{\mathbb {R}}_+} g^*(t) h^*(t)\,\mathrm{d}t: \ \rho '(g)\le 1\right\} , \end{aligned}$$

and such that \(\rho (f)={\overline{\rho }}(f^*)\) for all \(f\in {\mathfrak {M}}_0^+({\mathbb {R}})\). The rearrangement-invariant Banach function space generated by \({\overline{\rho }}\) is denoted by \({\overline{X}}({\mathbb {R}}_+)\).

For each \(t>0\), let \(E_t\) denote the dilation operator defined on \({\mathfrak {M}}({\mathbb {R}}_+)\) by

$$\begin{aligned} (E_tf)(s)=f(st),\quad 0<s<\infty . \end{aligned}$$

With \(X({\mathbb {R}})\) and \({\overline{X}}({\mathbb {R}}_+)\) as above, let \(h_X(t)\) denote the operator norm of \(E_{1/t}\) as an operator on \({\overline{X}}({\mathbb {R}}_+)\). By [1, Chap. 3, Proposition 5.11], for each \(t>0\), the operator \(E_t\) is bounded on \({\overline{X}}({\mathbb {R}}_+)\) and the function \(h_X\) is increasing and submultiplicative on \((0,\infty )\). The Boyd indices of \(X({\mathbb {R}})\) are the numbers \(\alpha _X\) and \(\beta _X\) defined by

$$\begin{aligned} \alpha _X:=\sup _{t\in (0,1)}\frac{\log h_X(t)}{\log t}, \quad \beta _X:=\inf _{t\in (1,\infty )}\frac{\log h_X(t)}{\log t}. \end{aligned}$$

By [1, Chap. 3, Proposition 5.13], \(0\le \alpha _X\le \beta _X\le 1\). The Boyd indices are said to be nontrivial if \(\alpha _X,\beta _X\in (0,1)\). The Boyd indices of the Lebesgue space \(L^p({\mathbb {R}})\), \(1\le p\le \infty \), are both equal to 1/p. Note that the Boyd indices of a rearrangement-invariant space may be different [1, Chap. 3, Exercises 6, 13].

The next theorem follows from the Boyd interpolation theorem [11, Theorem 1] for quasi-linear operators of weak types (pp) and (qq). Its proof can also be found in [1, Chap. 3, Theorem 5.16] and [30, Theorem 2.b.11].

Theorem 2.1

Let \(1\le q<p\le \infty \) and \(X({\mathbb {R}})\) be a rearrangement-invariant Banach function space with the Boyd indices \(\alpha _X,\beta _X\) satisfying \(1/p<\alpha _X\), \(\beta _X<1/q\). Then there exists a constant \(C_{p,q}\in (0,\infty )\) such that if a linear operator \(T:{\mathfrak {M}}({\mathbb {R}})\rightarrow {\mathfrak {M}}({\mathbb {R}})\) is bounded on the Lebesgue spaces \(L^p({\mathbb {R}})\) and \(L^q({\mathbb {R}})\), then it is also bounded on the rearrangement-invariant Banach function space \(X({\mathbb {R}})\) and

$$\begin{aligned} \Vert T\Vert _{{\mathcal {B}}(X({\mathbb {R}}))}\le C_{p,q}\max \big \{ \Vert T\Vert _{{\mathcal {B}}(L^p({\mathbb {R}}))},\Vert T\Vert _{{\mathcal {B}}(L^q({\mathbb {R}}))} \big \}. \end{aligned}$$
(2.1)

Notice that estimate (2.1) is not stated explicitly in [1, 11, 30]. However, it can be extracted from the proof of the Boyd interpolation theorem.

2.4 Weighted Banach function spaces

Let \(X({\mathbb {R}})\) be a Banach function space generated by a Banach function norm \(\rho \). We say that \(f\in X_{\mathrm{loc}}({\mathbb {R}})\) if \(f\chi _E\in X({\mathbb {R}})\) for any measurable set \(E\subset {\mathbb {R}}\) of finite measure.

Lemma 2.2

[26, Lemma 2.4] Let \(X({\mathbb {R}})\) be a Banach function space generated by a Banach function norm \(\rho \), let \(X'({\mathbb {R}})\) be its associate space, and let \(w:{\mathbb {R}}\rightarrow [0,\infty ]\) be a weight. Suppose that \(w\in X_{\mathrm{loc}}({\mathbb {R}})\) and \(1/w\in X_{\mathrm{loc}}'({\mathbb {R}})\). Then

$$\begin{aligned} \rho _w(f):=\rho (fw),\quad f\in {\mathfrak {M}}^+({\mathbb {R}}), \end{aligned}$$

is a Banach function norm and

$$\begin{aligned} X({\mathbb {R}},w):=\{f\in {\mathfrak {M}}({\mathbb {R}}):fw\in X({\mathbb {R}})\} \end{aligned}$$

is a Banach function space generated by the Banach function norm \(\rho _w\). The space \(X'({\mathbb {R}},w^{-1})\) is the associate space of \(X({\mathbb {R}},w)\).

2.5 Density of nice functions in separable rearrangement-invariant Banach function spaces with Muckenhoupt weights

Recall that the (noncentered) Hardy–Littlewood maximal function Mf of a function \(f\in L_{\mathrm{loc}}^1({\mathbb {R}})\) is defined by

$$\begin{aligned} (Mf)(x):=\sup _{I\ni x}\frac{1}{|I|}\int _I|f(y)|\,\mathrm{d}y, \quad x\in {\mathbb {R}}, \end{aligned}$$

where the supremum is taken over all intervals \(I\subset {\mathbb {R}}\) of finite length containing the point x.

Let \({\mathcal {S}}({\mathbb {R}})\) be the Schwartz space of rapidly decreasing smooth functions and let us denote by \({\mathcal {S}}_0({\mathbb {R}})\) the set of all functions \(f\in {\mathcal {S}}({\mathbb {R}})\) such that their Fourier transforms \({\mathcal {F}}f\) have compact supports.

Lemma 2.3

Let \(X({\mathbb {R}})\) be a separable rearrangement-invariant Banach function space and \(X'({\mathbb {R}})\) be its associate space. Suppose that the Boyd indices of \(X({\mathbb {R}})\) satisfy \(0<\alpha _X\), \(\beta _X<1\) and \(w\in A_{1/\alpha _X}({\mathbb {R}})\cap A_{1/\beta _X}({\mathbb {R}})\). Then

  1. (a)

    \(w\in X_{\mathrm{loc}}({\mathbb {R}})\) and \(1/w\in X_{\mathrm{loc}}'({\mathbb {R}})\);

  2. (b)

    the Banach function space space \(X({\mathbb {R}},w)\) is separable;

  3. (c)

    the Hardy-Littlewood maximal operator M is bounded on the Banach function space \(X({\mathbb {R}},w)\) and on its associate space \(X'({\mathbb {R}},w^{-1})\);

  4. (d)

    the set \({\mathcal {S}}_0({\mathbb {R}})\) is dense in the Banach function space \(X({\mathbb {R}},w)\).

Proof

Parts (a) and (c) are proved in [21, Section 4.3]. Part (b) follows from part (a), Lemma 2.2 and [26, Lemmas 2.7 and 2.11]. Part (d) is a consequence of parts (b), (c) and [16, Theorem 4]. \(\square \)

2.6 The Banach algebra \({\mathcal {M}}_{X({\mathbb {R}},w)}\) of Fourier multipliers

The following result plays an important role in this paper.

Theorem 2.4

Let \(X({\mathbb {R}})\) be a separable rearrangement-invariant Banach function space with the Boyd indices satisfying \(0<\alpha _X\), \(\beta _X<1\). Suppose that a weight w belongs to \(A_{1/\alpha _X}({\mathbb {R}})\cap A_{1/\beta _X}({\mathbb {R}})\). If \(a\in {\mathcal {M}}_{X({\mathbb {R}},w)}\), then

$$\begin{aligned} \Vert a\Vert _{L^\infty ({\mathbb {R}})}\le \Vert a\Vert _{{\mathcal {M}}_{X({\mathbb {R}},w)}}. \end{aligned}$$
(2.2)

The constant 1 on the right-hand side of (2.2) is best possible.

This theorem follows from Lemma 2.3(b) and [15, Theorem 2.4] (which was deduced from [24, Corollary 4.2]).

Inequality (2.2) was established earlier in [22, Theorem 1] with some constant on the right-hand side that depends on the space \(X({\mathbb {R}},w)\).

Since (2.2) is available, an easy adaptation of the proof of [18, Proposition 2.5.13] leads to the following (we refer to the proof of [22, Corollary 1] for details).

Corollary 2.5

Let \(X({\mathbb {R}})\) be a separable rearrangement-invariant Banach function space with the Boyd indices satisfying \(0<\alpha _X\), \(\beta _X<1\). Suppose that a weight w belongs to \(A_{1/\alpha _X}({\mathbb {R}})\cap A_{1/\beta _X}({\mathbb {R}})\). Then the set of the Fourier multipliers \({\mathcal {M}}_{X({\mathbb {R}},w)}\) is a Banach algebra under pointwise operations and the norm \(\Vert \cdot \Vert _{{\mathcal {M}}_{X({\mathbb {R}},w)}}\).

As usual, we denote by \(C_c^\infty ({\mathbb {R}})\) the set of all infinitely differentiable functions with compact support.

Theorem 2.6

Suppose that a non-negative even function \(\varphi \in C_c^\infty ({\mathbb {R}})\) satisfies the condition

$$\begin{aligned} \int _{\mathbb {R}}\varphi (x)\,\mathrm{d}x=1 \end{aligned}$$
(2.3)

and the function \(\varphi _\delta \) is defined for \(\delta >0\) by

$$\begin{aligned} \varphi _\delta (x):=\delta ^{-1}\varphi (x/\delta ), \quad x\in {\mathbb {R}}. \end{aligned}$$
(2.4)

Let \(X({\mathbb {R}})\) be a separable rearrangement-invariant Banach function space with the Boyd indices satisfying \(0<\alpha _X\), \(\beta _X<1\). Suppose that a weight w belongs to \(A_{1/\alpha _X}({\mathbb {R}})\cap A_{1/\beta _X}({\mathbb {R}})\). If \(a\in {\mathcal {M}}_{X({\mathbb {R}},w)}\), then for every \(\delta >0\),

$$\begin{aligned} \Vert a*\varphi _\delta \Vert _{{\mathcal {M}}_{X({\mathbb {R}},w)}}\le \Vert a\Vert _{{\mathcal {M}}_{X({\mathbb {R}},w)}}. \end{aligned}$$
(2.5)

Proof

The proof is analogous to the proof of [23, Theorem 2.6]. It follows from Lemma 2.3(c) and [26, Theorems 3.8(a) and 3.9(c)] that if the weight w belongs to \(A_{1/\alpha _X}({\mathbb {R}})\cap A_{1/\beta _X}({\mathbb {R}})\), then

$$\begin{aligned} \sup _{-\infty<a<b<\infty }\frac{1}{b-a} \Vert \chi _{(a,b)}\Vert _{X({\mathbb {R}},w)}\Vert \chi _{(a,b)}\Vert _{X'({\mathbb {R}},w^{-1})}<\infty . \end{aligned}$$

Therefore, by [24, Lemma 1.3], the Banach function space \(X({\mathbb {R}},w)\) satisfies the hypotheses of [24, Theorem 1.3]. It is shown in its proof (see [24, Section 4.2]) that for every \(\delta >0\) and every \(f\in {\mathcal {S}}({\mathbb {R}})\cap X({\mathbb {R}},w)\),

$$\begin{aligned} \Vert {\mathcal {F}}^{-1}(a*\varphi _\delta ){\mathcal {F}}f\Vert _{X({\mathbb {R}},w)} \le \sup \left\{ \frac{\Vert {\mathcal {F}}^{-1}a{\mathcal {F}}f\Vert _{X({\mathbb {R}},w)}}{\Vert f\Vert _{X({\mathbb {R}},w)}}: f\in X_{\mathcal {S}}({\mathbb {R}},w) \right\} \Vert f\Vert _{X({\mathbb {R}},w)}, \end{aligned}$$

where

$$\begin{aligned} X_{\mathcal {S}}({\mathbb {R}},w):=({\mathcal {S}}({\mathbb {R}})\cap X({\mathbb {R}},w)){\setminus }\{0\}. \end{aligned}$$

Then, for every \(\delta >0\),

$$\begin{aligned} \sup \left\{ \frac{\Vert {\mathcal {F}}^{-1}(a*\varphi _\delta ){\mathcal {F}}f\Vert _{X({\mathbb {R}},w)}}{\Vert f\Vert _{X({\mathbb {R}},w)}}: f\in X_{\mathcal {S}}({\mathbb {R}},w) \right\} \le \Vert a\Vert _{{\mathcal {M}}_{X({\mathbb {R}},w)}}. \end{aligned}$$
(2.6)

By Lemma 2.3(b), the Banach function space \(X({\mathbb {R}},w)\) is separable. Then it follows from [1, Chap. 1, Corollary 5.6] and [24, Theorems 2.3 and 6.1] that for every \(\delta >0\), the left-hand side of inequality (2.6) coincides with the multiplier norm \(\Vert a*\varphi _\delta \Vert _{{\mathcal {M}}_{X({\mathbb {R}},w)}}\), which completes the proof of inequality (2.5). \(\square \)

3 Continuous Fourier multipliers vanishing at infinity

3.1 The case of Lebesgue spaces with Muckenhoupt weights

The closure of a subset \(\mathfrak{S}\) of a Banach space \(\mathcal{E}\) in the norm of \(\mathcal{E}\) will be denoted by \({\text {clos}}_{\mathcal {E}}({\mathfrak {S}})\).

Let \(C_0({\mathbb {R}})\) be the set of all functions \(f\in C({{\dot{{\mathbb {R}}}}})\) such that \(f(\infty )=0\).

Lemma 3.1

Let \(1<p<\infty \) and \(w\in A_p({\mathbb {R}})\). Then

$$\begin{aligned} C_0({\mathbb {R}})\cap V({\mathbb {R}}) \subset {\text {clos}}_{{\mathcal {M}}_{L^p({\mathbb {R}},w)}}\big (C_c^\infty ({\mathbb {R}})\big ). \end{aligned}$$

Proof

The idea of the proof is borrowed from [20, Theorem 1.16] (see also [23, Theorem 3.1]). If \(w\in A_p({\mathbb {R}})\), then \(w^{1+\delta _2}\in A_{p(1+\delta _1)}({\mathbb {R}})\) whenever \(|\delta _1|\) and \(|\delta _2|\) are sufficiently small (see, e.g., [7, Theorem 2.31]). If \(p\ge 2\), then one can find sufficiently small \(\delta _1,\delta _2>0\) and a number \(\theta \in (0,1)\) such that

$$\begin{aligned} \frac{1}{p}=\frac{1-\theta }{2}+\frac{\theta }{p(1+\delta _1)}, \quad w=1^{1-\theta }w^{(1+\delta _2)\theta }, \quad w^{1+\delta _2}\in A_{p(1+\delta _1)}({\mathbb {R}}). \end{aligned}$$
(3.1)

If \(1<p<2\), then one can find a sufficiently small number \(\delta _2>0\), a number \(\delta _1<0\) with sufficiently small \(|\delta _1|\), and a number \(\theta \in (0,1)\) such that all conditions in (3.1) are fulfilled.

Let us use the following abbreviations:

$$\begin{aligned} {\mathcal {M}}_p:= & \, {\mathcal {M}}_{L^p({\mathbb {R}},w)}, \quad {\mathcal {M}}_{p_\theta }:={\mathcal {M}}_{L^{p(1+\delta _1)}({\mathbb {R}},w^{1+\delta _2})},\\ {\mathcal {B}}_p:= & \, {\mathcal {B}}(L^p({\mathbb {R}},w)), \quad {\mathcal {B}}_{p_\theta }:={\mathcal {B}}(L^{p(1+\delta _1)}({\mathbb {R}},w^{1+\delta _2})). \end{aligned}$$

For \(n\in {\mathbb {N}}\), let

$$\begin{aligned} \psi _n(x):= \left\{ \begin{array}{lll} 1 &{} \quad \text{ if }& |x|\le n,\\ n+1-|x| & \quad \text{ if }&{} n<|x|<n+1,\\ 0 & \quad \text{ if }& |x|\ge n+1. \end{array}\right. \end{aligned}$$
(3.2)

Then \(\psi _n\) has compact support and \(\Vert \psi _n\Vert _{V({\mathbb {R}})}=3\). By the Stechkin-type inequality (1.4),

$$\begin{aligned} \Vert \psi _n\Vert _{{\mathcal {M}}_{p_\theta }}\le c_\theta , \end{aligned}$$

where \(c_\theta \) is three times \(c_{L^{p(1+\delta _1)}({\mathbb {R}},w^{1+\delta _2})}\), and the latter constant is the constant from (1.4).

Let \(a\in \ C_0({\mathbb {R}})\cap V({\mathbb {R}})\). Fix \(\varepsilon >0\). For \(n\in {\mathbb {N}}\), take \(b_n:=a\psi _n\). Then

$$\begin{aligned} \lim _{n\rightarrow \infty }\Vert a-b_n\Vert _{L^\infty ({\mathbb {R}})}=0 \end{aligned}$$
(3.3)

and \(b_n\in C_0({\mathbb {R}})\) has compact support. Taking into account the Stechkin-type inequality (1.4), we get

$$\begin{aligned} \Vert a-b_n\Vert _{{\mathcal {M}}_{p_\theta }} \le \Vert a\Vert _{{\mathcal {M}}_{p_\theta }}(1+\Vert \psi _n\Vert _{{\mathcal {M}}_{p_\theta }}) \le (1+c_\theta )c_\theta \Vert a\Vert _{V({\mathbb {R}})} \end{aligned}$$
(3.4)

and

$$\begin{aligned} \Vert b_n\Vert _{{\mathcal {M}}_{p_\theta }} \le \Vert a\Vert _{{\mathcal {M}}_{p_\theta }}\Vert \psi _n\Vert _{{\mathcal {M}}_{p_\theta }} \le c_\theta ^2\Vert a\Vert _{V({\mathbb {R}})}. \end{aligned}$$
(3.5)

It follows from (3.1) and the Stein–Weiss interpolation theorem (see, e.g., [1, Chap. 3, Theorem 3.6]) that

$$\begin{aligned} \Vert a-b_n\Vert _{{\mathcal {M}}_p}&= \Vert W^0(a-b_n)\Vert _{{\mathcal {B}}_p} \nonumber \\&\le \Vert W^0(a-b_n)\Vert _{{\mathcal {B}}(L^2({\mathbb {R}}))}^{1-\theta } \Vert W^0(a-b_n)\Vert _{{\mathcal {B}}_{p_\theta }}^\theta \nonumber \\&= \Vert a-b_n\Vert _{L^\infty ({\mathbb {R}})}^{1-\theta } \Vert a-b_n\Vert _{{\mathcal {M}}_{p_\theta }}^\theta . \end{aligned}$$
(3.6)

Combining (3.3), (3.4) and (3.6), we see that there exists \(n_0\in {\mathbb {N}}\) such that

$$\begin{aligned} \Vert a-b_{n_0}\Vert _{{\mathcal {M}}_p}<\varepsilon /2. \end{aligned}$$
(3.7)

Let \(\varphi \in C_c^\infty ({\mathbb {R}})\) be a non-negative even function satisfying (2.3) and for \(\delta >0\) let the function \(\varphi _\delta \) be defined by (2.4). By Theorem 2.6 and inequality (3.5), for every \(\delta >0\),

$$\begin{aligned} \Vert b_{n_0}*\varphi _\delta \Vert _{{\mathcal {M}}_{p_\theta }} \le \Vert b_{n_0}\Vert _{{\mathcal {M}}_{p_\theta }} \le c_\theta ^2\Vert a\Vert _{V({\mathbb {R}})}. \end{aligned}$$
(3.8)

It follows from [12, Propositions 4.18, 4.20–4.21] that \(b_{n_0}*\varphi _\delta \in C_c^\infty ({\mathbb {R}})\) and

$$\begin{aligned} \lim _{\delta \rightarrow 0^+}\Vert b_{n_0}*\varphi _\delta -b_{n_0}\Vert _{L^\infty ({\mathbb {R}})}=0. \end{aligned}$$
(3.9)

In view of (3.1) and the Stein-Weiss interpolation theorem (see, e.g., [1, Chap. 3, Theorem 3.6]), we see that

$$\begin{aligned}&\Vert b_{n_0}*\varphi _\delta -b_{n_0}\Vert _{{\mathcal {M}}_p} \nonumber \\&\quad = \Vert W^0(b_{n_0}*\varphi _\delta -b_{n_0})\Vert _{{\mathcal {B}}_p} \nonumber \\&\quad \le \Vert W^0(b_{n_0}*\varphi _\delta -b_{n_0})\Vert _{{\mathcal {B}}(L^2({\mathbb {R}}))}^{1-\theta } \Vert W^0(b_{n_0}*\varphi _\delta -b_{n_0})\Vert _{{\mathcal {B}}_{p_\theta }}^\theta \nonumber \\&\quad = \Vert b_{n_0}*\varphi _\delta -b_{n_0}\Vert _{L^\infty ({\mathbb {R}})}^{1-\theta } \Vert b_{n_0}*\varphi _\delta -b_{n_0}\Vert _{{\mathcal {M}}_{p_\theta }}^\theta \nonumber \\&\quad \le \Vert b_{n_0}*\varphi _\delta -b_{n_0}\Vert _{L^\infty ({\mathbb {R}})}^{1-\theta } ( \Vert b_{n_0}*\varphi _\delta \Vert _{{\mathcal {M}}_{p_\theta }}+\Vert b_{n_0}\Vert _{{\mathcal {M}}_{p_\theta }} )^\theta . \end{aligned}$$
(3.10)

Combining (3.8)–(3.10), we conclude that there exists \(\delta _0>0\) such that

$$\begin{aligned} \Vert b_{n_0}*\varphi _{\delta _0}-b_{n_0}\Vert _{{\mathcal {M}}_p}<\varepsilon /2. \end{aligned}$$
(3.11)

Hence, it follows from (3.7) and (3.11) that for every function a in the intersection \(C_0({\mathbb {R}})\cap V({\mathbb {R}})\) and every \(\varepsilon >0\) there exists a function \(b_{n_0}*\varphi _{\delta _0}\in C_c^\infty ({\mathbb {R}})\) such that \(\Vert a-b_{n_0}*\varphi _{\delta _0}\Vert _{{\mathcal {M}}_p}<\varepsilon \). Therefore, \(a\in {\text {clos}}_{{\mathcal {M}}_p}\big (C_c^\infty ({\mathbb {R}})\big )\). \(\square \)

3.2 The case of rearrangement-invariant spaces with Muckenhoupt weights

The following lemma is an extension of the previous result to the case of rearrangement-invariant Banach function spaces with Muckenhoupt weights.

Lemma 3.2

Let \(X({\mathbb {R}})\) be a separable rearrangement-invariant Banach function space with the Boyd indices satisfying \(0<\alpha _X\), \(\beta _X<1\). Suppose that a weight w belongs to \(A_{1/\alpha _X}({\mathbb {R}})\cap A_{1/\beta _X}({\mathbb {R}})\). Then

$$\begin{aligned} C_0({\mathbb {R}})\cap V({\mathbb {R}}) \subset {\text {clos}}_{{\mathcal {M}}_{X({\mathbb {R}},w)}}\big (C_c^\infty ({\mathbb {R}})\big ). \end{aligned}$$

Proof

Since \(\alpha _X,\beta _X\in (0,1)\) and \(w\in A_{1/\alpha _X}({\mathbb {R}})\cap A_{1/\beta _X}({\mathbb {R}})\), it follows from [7, Theorem 2.31] that there exist p and q such that

$$\begin{aligned} 1<q<1/\beta _X\le 1/\alpha _X<p<\infty , \quad w\in A_p({\mathbb {R}})\cap A_q({\mathbb {R}}). \end{aligned}$$
(3.12)

Let \(C_{p,q}\in (0,\infty )\) be the constant from estimate (2.1). Fix \(\varepsilon >0\) and take a function \(a\in C_0({\mathbb {R}})\cap V({\mathbb {R}})\). As in the proof of inequality (3.7) (see the proof of Lemma 3.1), it can be shown that there exists \(n_0\in {\mathbb {N}}\) such that

$$\begin{aligned} \Vert a-b_{n_0}\Vert _{{\mathcal {M}}_{L^p({\mathbb {R}},w)}}<\frac{\varepsilon }{2C_{p,q}}, \quad \Vert a-b_{n_0}\Vert _{{\mathcal {M}}_{L^q({\mathbb {R}},w)}}<\frac{\varepsilon }{2C_{p,q}}, \end{aligned}$$
(3.13)

where \(b_n=a\psi _n\) and \(\psi _n\) is given by (3.2) for every \(n\in {\mathbb {N}}\). It follows from (3.12), (3.13) and Theorem 2.1 that

$$\begin{aligned}&\Vert a-b_{n_0}\Vert _{{\mathcal {M}}_{X({\mathbb {R}},w)}} = \Vert W^0(a-b_{n_0})\Vert _{{\mathcal {B}}(X({\mathbb {R}},w))} \nonumber \\&\quad = \Vert wW^0(a-b_{n_0})w^{-1}I\Vert _{{\mathcal {B}}(X({\mathbb {R}}))} \nonumber \\&\quad \le C_{p,q}\max \left\{ \Vert wW^0(a-b_{n_0})w^{-1}I\Vert _{{\mathcal {B}}(L^p({\mathbb {R}}))}, \Vert wW^0(a-b_{n_0})w^{-1}I\Vert _{{\mathcal {B}}(L^q({\mathbb {R}}))} \right\} \nonumber \\&\quad = C_{p,q}\max \left\{ \Vert W^0(a-b_{n_0})\Vert _{{\mathcal {B}}(L^p({\mathbb {R}},w))}, \Vert W^0(a-b_{n_0})\Vert _{{\mathcal {B}}(L^q({\mathbb {R}},w))} \right\} \nonumber \\&\quad = C_{p,q} \max \left\{ \Vert a-b_{n_0}\Vert _{{\mathcal {M}}_{L^p({\mathbb {R}},w)}}, \Vert a-b_{n_0}\Vert _{{\mathcal {M}}_{L^q({\mathbb {R}},w)}} \right\} <\varepsilon /2. \end{aligned}$$
(3.14)

As in the proof of inequality (3.11) (see the proof of Lemma 3.1), it can be shown that there exists \(\delta _0>0\) such that

$$\begin{aligned} \Vert b_{n_0}*\varphi _{\delta _0}-b_{n_0}\Vert _{{\mathcal {M}}_{L^p({\mathbb {R}},w)}}<\frac{\varepsilon }{2C_{p,q}}, \quad \Vert b_{n_0}*\varphi _{\delta _0}-b_{n_0}\Vert _{{\mathcal {M}}_{L^q({\mathbb {R}},w)}}<\frac{\varepsilon }{2C_{p,q}}, \end{aligned}$$
(3.15)

where \(\varphi \in C_c^\infty ({\mathbb {R}})\) is a non-negative even function satisfying (2.3) and the functions \(\varphi _\delta \) are defined for all \(\delta >0\) by (2.4). Arguing as in the proof of (3.14), we deduce from (3.12), (3.15) and Theorem 2.1 that

$$\begin{aligned} \Vert b_{n_0}*\varphi _{\delta _0}-b_{n_0}\Vert _{{\mathcal {M}}_{X({\mathbb {R}},w)}}<\varepsilon /2. \end{aligned}$$
(3.16)

It follows from (3.14) and (3.16) that for every function a in the intersection \(C_0({\mathbb {R}})\cap V({\mathbb {R}})\) and every \(\varepsilon >0\) there exists a function \(b_{n_0}*\varphi _{\delta _0}\in C_c^\infty ({\mathbb {R}})\) such that \(\Vert a-b_{n_0}*\varphi _{\delta _0}\Vert _{{\mathcal {M}}_{X({\mathbb {R}},w)}}<\varepsilon \). Therefore, \(a\in {\text {clos}}_{{\mathcal {M}}_{X(R,w)}}\big (C_c^\infty ({\mathbb {R}})\big )\). \(\square \)

Now we are in a position to prove the main result of this section.

Theorem 3.3

Let \(X({\mathbb {R}})\) be a separable rearrangement-invariant Banach function space with the Boyd indices satisfying \(0<\alpha _X\), \(\beta _X<1\). Suppose that a weight w belongs to \(A_{1/\alpha _X}({\mathbb {R}})\cap A_{1/\beta _X}({\mathbb {R}})\). Consider the set

$$\begin{aligned} C_{0,X({\mathbb {R}},w)}({{\dot{{\mathbb {R}}}}}):=\left\{ a\in C_{X({\mathbb {R}},w)}({{\dot{{\mathbb {R}}}}})\ :\ a(\infty )=0\right\} . \end{aligned}$$
(3.17)

Then

$$\begin{aligned} C_{0,X({\mathbb {R}},w)}({{\dot{{\mathbb {R}}}}}) = {\text {clos}}_{{\mathcal {M}}_{X({\mathbb {R}},w)}}\big (C_c^\infty ({\mathbb {R}})\big ). \end{aligned}$$
(3.18)

Proof

Let \(a\in C_{X({\mathbb {R}},w)}({{\dot{{\mathbb {R}}}}})\) be such that \(a(\infty )=0\). Fix \(\varepsilon >0\). By the definition of the algebra \(C_{X({\mathbb {R}},w)}({{\dot{{\mathbb {R}}}}})\), there exists a function \(b\in C({{\dot{{\mathbb {R}}}}})\cap V({\mathbb {R}})\) such that

$$\begin{aligned} \Vert a-b\Vert _{{\mathcal {M}}_{X({\mathbb {R}},w)}}<\varepsilon /3. \end{aligned}$$
(3.19)

It follows from this observation and the continuous embedding of \({\mathcal {M}}_{X({\mathbb {R}},w)}\) into \(L^\infty ({\mathbb {R}})\) (see Theorem 2.4) that

$$\begin{aligned} |b(\infty )|=|a(\infty )-b(\infty )| \le \Vert a-b\Vert _{L^\infty ({\mathbb {R}})}\le \Vert a-b\Vert _{{\mathcal {M}}_{X({\mathbb {R}},w)}}<\varepsilon /3. \end{aligned}$$
(3.20)

Take \(c=b-b(\infty )\in C_0({\mathbb {R}})\cap V({\mathbb {R}})\). By Lemma 3.2, there exists a function \(d\in C_c^\infty ({\mathbb {R}})\subset {\mathcal {M}}_{X({\mathbb {R}},w)}\) such that

$$\begin{aligned} \Vert c-d\Vert _{{\mathcal {M}}_{X({\mathbb {R}},w)}}<\varepsilon /3. \end{aligned}$$
(3.21)

Combining inequalities (3.19)–(3.21), we see that

$$\begin{aligned} \Vert a-d\Vert _{{\mathcal {M}}_{X({\mathbb {R}},w)}} \le \Vert a-b\Vert _{{\mathcal {M}}_{X({\mathbb {R}},w)}} + |b(\infty )| + \Vert c-d\Vert _{{\mathcal {M}}_{X({\mathbb {R}},w)}}<\varepsilon . \end{aligned}$$

Hence

$$\begin{aligned} C_{0,X({\mathbb {R}},w)}({{\dot{{\mathbb {R}}}}}) \subset {\text {clos}}_{{\mathcal {M}}_{X({\mathbb {R}},w)}}\big (C_c^\infty ({\mathbb {R}})\big ). \end{aligned}$$
(3.22)

Let us prove the reverse embedding. Take \(a\in {\text {clos}}_{{\mathcal {M}}_{X({\mathbb {R}},w)}}\big (C_c^\infty ({\mathbb {R}})\big )\). Then there exists a sequence \(\{a_n\}_{n\in {\mathbb {N}}}\subset C_c^\infty ({\mathbb {R}})\) such that

$$\begin{aligned} \lim _{n\rightarrow \infty }\Vert a_n-a\Vert _{{\mathcal {M}}_{X({\mathbb {R}},w)}}=0. \end{aligned}$$

Since \(C_c^\infty ({\mathbb {R}})\subset C({{\dot{{\mathbb {R}}}}})\cap V({\mathbb {R}})\), the above equality and the continuous embedding of the algebra \({\mathcal {M}}_{X({\mathbb {R}},w)}\) into the algebra \(L^\infty ({\mathbb {R}})\) (see Theorem 2.4) imply that \(a\in C_{X({\mathbb {R}},w)}({{\dot{{\mathbb {R}}}}})\) and

$$\begin{aligned} |a(\infty )|&= \lim _{n\rightarrow \infty }|a_n(\infty )-a(\infty )| \le \lim _{n\rightarrow \infty }\Vert a_n-a\Vert _{L^\infty ({\mathbb {R}})} \\&\le \lim _{n\rightarrow \infty }\Vert a_n-a\Vert _{{\mathcal {M}}_{X(R,w)}}=0. \end{aligned}$$

Thus

$$\begin{aligned} {\text {clos}}_{{\mathcal {M}}_{X({\mathbb {R}},w)}}\big (C_c^\infty ({\mathbb {R}})\big ) \subset C_{0,X({\mathbb {R}},w)}({{\dot{{\mathbb {R}}}}}). \end{aligned}$$
(3.23)

Combining (3.22) and (3.23), we arrive at (3.18). \(\square \)

4 Continuous and slowly oscillating Fourier multipliers

4.1 Continuous Fourier multipliers on one and two-point compactifications of the real line

For a function \(f\in C({\overline{{\mathbb {R}}}})\), let

$$\begin{aligned} J_f(x):=\left\{ \begin{array}{lll} f(-\infty ) &{} \quad \text{ if }&{} x\in (-\infty ,-1), \\ \frac{1}{2}\big [f(-\infty )(1-x)+f(+\infty )(1+x)\big ] &{} \quad \text{ if } &{} x\in [-1,1], \\ f(+\infty ) &{} \quad \text{ if }&{} x\in (1,+\infty ). \end{array}\right. \end{aligned}$$
(4.1)

It is easy to see that

$$\begin{aligned} \Vert J_f\Vert _{V({\mathbb {R}})}=\max \big \{|f(-\infty )|,|f(+\infty )|\big \}+ |f(+\infty )-f(-\infty )|. \end{aligned}$$
(4.2)

Therefore \(J_f\in C({\overline{{\mathbb {R}}}})\cap V({\mathbb {R}})\) and \(f-J_f\in C_0({\mathbb {R}})\).

The next lemma extends [29, Lemma 3.1(i)] from the setting of Lebesgue spaces to the setting of rearrangement-invariant Banach function spaces with Muckenhoupt weights.

Lemma 4.1

Let \(X({\mathbb {R}})\) be a separable rearrangement-invariant Banach function space with the Boyd indices satisfying \(0<\alpha _X\), \(\beta _X<1\). Suppose that \(w\in A_{1/\alpha _X}({\mathbb {R}})\cap A_{1/\beta _X}({\mathbb {R}})\). Then

$$\begin{aligned} C_{X({\mathbb {R}},w)}({{\dot{{\mathbb {R}}}}})=C_{X({\mathbb {R}},w)}({\overline{{\mathbb {R}}}})\cap C({{\dot{{\mathbb {R}}}}}). \end{aligned}$$
(4.3)

Proof

The proof is analogous to the proof of [29, Lemma 3.1(i)] (see also [23, Lemma 3.2]). It is obvious that \(C_{X({\mathbb {R}},w)}({{\dot{{\mathbb {R}}}}})\subset C_{X({\mathbb {R}},w)}({\overline{{\mathbb {R}}}})\). On the other hand, it follows from Theorem 2.4 that \(C_{X({\mathbb {R}},w)}({{\dot{{\mathbb {R}}}}})\subset C({{\dot{{\mathbb {R}}}}})\). Therefore,

$$\begin{aligned} C_{X({\mathbb {R}},w)}({{\dot{{\mathbb {R}}}}})\subset C_{X({\mathbb {R}},w)}({\overline{{\mathbb {R}}}})\cap C({{\dot{{\mathbb {R}}}}}). \end{aligned}$$
(4.4)

To prove the opposite embedding, let us consider an arbitrary function \(a\in C_{X({\mathbb {R}},w)}({\overline{{\mathbb {R}}}})\) such that \(a(+\infty )=a(-\infty )\). Let \(\{a_n\}_{n\in {\mathbb {N}}}\subset C({\overline{{\mathbb {R}}}})\cap V({\mathbb {R}})\) be a sequence such that \(\Vert a_n-a\Vert _{{\mathcal {M}}_{X({\mathbb {R}},w)}}\rightarrow 0\) as \(n\rightarrow \infty \). According to Theorem 2.4, the sequence \(\{a_n\}_{n\in {\mathbb {N}}}\) converges to a uniformly on \({\mathbb {R}}\). Hence, in particular, \(a_n(\pm \infty )\rightarrow a(\infty )\) as \(n\rightarrow \infty \). Let the functions \(b_n:=J_{a_n-a(\infty )}\) be defined by (4.1) with \(a_n-a(\infty )\) in place of f. By the Stechkin-type inequality (1.4) and equality (4.2), we have

$$\begin{aligned} \Vert b_n\Vert _{{\mathcal {M}}_{X({\mathbb {R}},w)}}&\le c_{X({\mathbb {R}},w)}\Vert J_{a_n-a(\infty )}\Vert _{V({\mathbb {R}})} \\&= c_{X({\mathbb {R}},w)}\max \big \{|a_n(-\infty )-a(\infty )|,|a_n(+\infty )-a(\infty )|\big \} \\&\quad +c_{X({\mathbb {R}},w)}|a_n(+\infty )-a_n(-\infty )|. \end{aligned}$$

Therefore, \(\Vert b_n\Vert _{{\mathcal {M}}_{X({\mathbb {R}},w)}}\rightarrow 0\) as \(n\rightarrow \infty \) and thus,

$$\begin{aligned} \lim _{n\rightarrow \infty }\Vert a_n-b_n-a\Vert _{{\mathcal {M}}_{X({\mathbb {R}},w)}}=0. \end{aligned}$$

Since \(a_n-b_n\in C({{\dot{{\mathbb {R}}}}})\cap V({\mathbb {R}})\), the latter equality implies that \(a\in C_{X({\mathbb {R}},w)}({{\dot{{\mathbb {R}}}}})\). Thus

$$\begin{aligned} C_{X({\mathbb {R}},w)}({\overline{{\mathbb {R}}}})\cap C({{\dot{{\mathbb {R}}}}})\subset C_{X({\mathbb {R}},w)}({{\dot{{\mathbb {R}}}}}). \end{aligned}$$
(4.5)

Combining embeddings (4.4)–(4.5), we arrive at equality (4.3). \(\square \)

4.2 Embedding of the algebra \(C_{X({\mathbb {R}},w)}({{\dot{{\mathbb {R}}}}})\) into the algebra \(SO_{X({\mathbb {R}},w)}\) of slowly oscillating Fourier multipliers

Let \(C_b({\mathbb {R}}):=C({\mathbb {R}})\cap L^\infty ({\mathbb {R}})\). For a bounded measurable function \(f:{\mathbb {R}}\rightarrow {\mathbb {C}}\) and a set \(J\subset {\mathbb {R}}\), let

$$\begin{aligned} {\text {osc}}(f,J):= \mathop {\text {ess sup}}\limits _{x,y\in J}|f(x)-f(y)|. \end{aligned}$$

Let SO be the \(C^*\)-algebra of all slowly oscillating functions at \(\infty \) defined by

$$\begin{aligned} SO:=\left\{ f\in C_b({\mathbb {R}}): \lim _{x\rightarrow +\infty } {\text {osc}}(f,[-x,-x/2]\cup [x/2,x])=0\right\} . \end{aligned}$$

Consider the differential operator \((Df)(x)=xf'(x)\) and its iterations defined by \(D^0f=f\) and \(D^jf=D(D^{j-1}f)\) for \(j\in {\mathbb {N}}\). Let

$$\begin{aligned} SO^3:=\left\{ a\in SO\cap C^3({\mathbb {R}}):\lim _{x\rightarrow \infty } (D^ja)(x)=0,\ j=1,2,3 \right\} , \end{aligned}$$

where \(C^3({\mathbb {R}})\) denotes the set of all three times continuously differentiable functions. It is easy to see that \(SO^3\) is a commutative Banach algebra under pointwise operations and the norm

$$\begin{aligned} \Vert a\Vert _{SO^3}:=\sum _{j=0}^3\frac{1}{j!}\Vert D^ja\Vert _{L^\infty ({\mathbb {R}})}. \end{aligned}$$

It follows from [21, Corollary 2.6] that if \(X({\mathbb {R}})\) is a separable rearrangement-invariant Banach function space with the Boyd indices \(\alpha _X,\beta _X\) such that \(0<\alpha _X\), \(\beta _X<1\) and \(w\in A_{1/\alpha _X}({\mathbb {R}})\cap A_{1/\beta _X}({\mathbb {R}})\), then there exists a constant \(c_{X({\mathbb {R}},w)}\in (0,\infty )\) such that for all \(a\in SO^3\),

$$\begin{aligned} \Vert a\Vert _{{\mathcal {M}}_{X({\mathbb {R}},w)}}\le c_{X({\mathbb {R}},w)}\Vert a\Vert _{SO^3}. \end{aligned}$$

The continuous embedding \(SO^3\subset {\mathcal {M}}_{X({\mathbb {R}},w)}\) allows us to define the algebra \(SO_{X({\mathbb {R}},w)}\) of slowly oscillating Fourier multipliers as the closure of \(SO^3\) with respect to the multiplier norm:

$$\begin{aligned} SO_{X({\mathbb {R}},w)}:={\text {clos}}_{{\mathcal {M}}_{X({\mathbb {R}},w)}}\big (SO^3\big ). \end{aligned}$$

The following result is analogous to [28, Lemma 3.6].

Lemma 4.2

Let \(X({\mathbb {R}})\) be a separable rearrangement-invariant Banach function space with the Boyd indices satisfying \(0<\alpha _X\), \(\beta _X<1\). Suppose that a weight w belongs to \(A_{1/\alpha _X}({\mathbb {R}})\cap A_{1/\beta _X}({\mathbb {R}})\). Then \(C_{X({\mathbb {R}},w)}({{\dot{{\mathbb {R}}}}})\subset SO_{X({\mathbb {R}},w)}\).

Proof

Let \(a\in C_{X({\mathbb {R}},w)}({{\dot{{\mathbb {R}}}}})\). Fix \(\varepsilon >0\). Then there exists \(b\in C({{\dot{{\mathbb {R}}}}})\cap V({\mathbb {R}})\) such that

$$\begin{aligned} \Vert a-b\Vert _{{\mathcal {M}}_{X({\mathbb {R}},w)}}<\varepsilon /2. \end{aligned}$$
(4.6)

Then \(b-b(\infty )\in C_0({\mathbb {R}})\cap V({\mathbb {R}})\). By Lemma 3.2,

$$\begin{aligned} b-b(\infty )\in {\text {clos}}_{{\mathcal {M}}_{X({\mathbb {R}},w)}} \big (C_c^\infty ({\mathbb {R}})\big ). \end{aligned}$$

Then there exists \(c\in C_c^\infty ({\mathbb {R}})\) such that

$$\begin{aligned} \Vert b-b(\infty )-c\Vert _{{\mathcal {M}}_{X({\mathbb {R}},w)}}<\varepsilon /2. \end{aligned}$$
(4.7)

It follows from inequalities (4.6) and (4.7) that

$$\begin{aligned} \Vert a-(c+b(\infty ))\Vert _{{\mathcal {M}}_{X({\mathbb {R}},w)}}<\varepsilon . \end{aligned}$$

Since \(c+b(\infty )\in C_c^\infty ({\mathbb {R}})\dot{+}{\mathbb {C}}\subset SO^3\), we get \(a\in {\text {clos}}_{{\mathcal {M}}_{X({\mathbb {R}},w)}}(SO^3)=SO_{X({\mathbb {R}},w)}\). \(\square \)

5 Almost periodic Fourier multipliers and their products with continuous Fourier multipliers vanishing at infinity

5.1 The algebra \(AP_{X({\mathbb {R}},w)}\) of almost periodic Fourier multipliers

For \(\lambda \in {\mathbb {R}}\), let \(T_\lambda \) denote the translation operator defined by

$$\begin{aligned} (T_\lambda f)(x)=f(x-\lambda ),\quad x\in {\mathbb {R}}. \end{aligned}$$

Lemma 5.1

Let \(X({\mathbb {R}})\) be a rearrangement-invariant Banach function space and \(w:{\mathbb {R}}\rightarrow [0,\infty ]\) be a weight such that \(w\in X_{\mathrm{loc}}({\mathbb {R}})\) and \(1/w\in X_{\mathrm{loc}}'({\mathbb {R}})\). Suppose that \(\lambda \in {\mathbb {R}}\). Then the translation operator \(T_\lambda \) is bounded on the Banach function space \(X({\mathbb {R}},w)\) if and only if the function

$$\begin{aligned} v_\lambda (x):=\frac{w(x+\lambda )}{w(x)},\quad x\in {\mathbb {R}}, \end{aligned}$$

belongs to the space \(L^\infty ({\mathbb {R}})\). In that case \(\Vert T_\lambda \Vert _{{\mathcal {B}}(X({\mathbb {R}},w))}=\Vert v_\lambda \Vert _{L^\infty ({\mathbb {R}})}\).

Proof

The operator \(T_\lambda \) is bounded on the space \(X({\mathbb {R}},w)\) if and only if the operator \(wT_\lambda w^{-1}I=T_\lambda (v_\lambda I)\) is bounded on the space \(X({\mathbb {R}})\). Moreover, their norms coincide. It is easy to see that for every \(f\in X({\mathbb {R}})\), the function \(T_\lambda f\) is equimeasurable with f, whence \(\Vert T_\lambda f\Vert _{X({\mathbb {R}})}=\Vert f\Vert _{X({\mathbb {R}})}\). Therefore,

$$\begin{aligned} \Vert T_\lambda \Vert _{{\mathcal {B}}(X({\mathbb {R}},w))} = \Vert T_\lambda (v_\lambda I)\Vert _{{\mathcal {B}}(X({\mathbb {R}}))} = \Vert v_\lambda I\Vert _{{\mathcal {B}}(X({\mathbb {R}}))}. \end{aligned}$$

By [31, Theorem 1], the multiplication operator \(v_\lambda I\) is bounded on the space \(X({\mathbb {R}})\) if and only if \(v_\lambda \in L^\infty ({\mathbb {R}})\) and \(\Vert v_\lambda I\Vert _{{\mathcal {B}}(X({\mathbb {R}}))}=\Vert v_\lambda \Vert _{L^\infty ({\mathbb {R}})}\). Thus, \(\Vert T_\lambda \Vert _{{\mathcal {B}}(X({\mathbb {R}},w))}=\Vert v_\lambda \Vert _{L^\infty ({\mathbb {R}})}\). \(\square \)

As a consequence of the previous result, we show that for all \(\lambda \in {\mathbb {R}}\), the exponential functions \(e_\lambda (x)=e^{i\lambda x}\), \(x\in {\mathbb {R}}\), are Fourier multipliers on separable rearrangement-invariant Banach function spaces with weights in the sublclass \(A_{1/\alpha _X}^0({\mathbb {R}})\cap A_{1/\beta _X}^0({\mathbb {R}})\) of the class of Muckenhoupt weights.

Corollary 5.2

Let \(X({\mathbb {R}})\) be a separable rearrangement-invariant Banach function space with the Boyd indices satisfying \(0<\alpha _X\le \beta _X<1\). Suppose that a weight w belongs to \(A_{1/\alpha _X}^0({\mathbb {R}})\cap A_{1/\beta _X}^0({\mathbb {R}})\). Then for every \(\lambda \in {\mathbb {R}}\), the function \(e_\lambda \) belongs to \({\mathcal {M}}_{X({\mathbb {R}},w)}\) and \(\Vert e_\lambda \Vert _{{\mathcal {M}}_{X({\mathbb {R}},w)}}=\Vert v_\lambda \Vert _{L^\infty ({\mathbb {R}})}\).

Proof

It follows from the definition of the classes \(A_{1/\alpha _X}^0({\mathbb {R}})\) and \(A_{1/\beta _X}^0({\mathbb {R}})\) that the function \(v_\lambda (x)=\frac{w(x+\lambda )}{w(x)}\), \(x\in {\mathbb {R}}\), is bounded for every \(\lambda \in {\mathbb {R}}\). By Lemma 2.3(a), \(w\in X_{\mathrm{loc}}({\mathbb {R}})\) and \(1/w\in X_{\mathrm{loc}}'({\mathbb {R}})\). Then, by Lemma 5.1, the operator \(T_\lambda \) is bounded on the Banach function space \(X({\mathbb {R}},w)\) and

$$\begin{aligned} \Vert T_\lambda \Vert _{{\mathcal {B}}(X({\mathbb {R}},w))}=\Vert v_\lambda \Vert _{L^\infty ({\mathbb {R}})}, \quad \lambda \in {\mathbb {R}}. \end{aligned}$$

It remains to observe that \(T_\lambda =W^0(e_\lambda )\). Thus \(e_\lambda \in {\mathcal {M}}_{X({\mathbb {R}},w)}\) and

$$\begin{aligned} \Vert e_\lambda \Vert _{{\mathcal {M}}_{X({\mathbb {R}},w)}} = \Vert W^0(e_\lambda )\Vert _{{\mathcal {B}}(X({\mathbb {R}},w))} = \Vert v_\lambda \Vert _{L^\infty ({\mathbb {R}})}, \quad \lambda \in {\mathbb {R}}, \end{aligned}$$

which completes the proof. \(\square \)

Corollary 5.2 implies that if \(X({\mathbb {R}})\) is a separable rearrangement-invariant Banach function spaces and \(w\in A_{1/\alpha _X}^0({\mathbb {R}})\cap A_{1/\beta _X}^0({\mathbb {R}})\), then \(APP\subset {\mathcal {M}}_{X({\mathbb {R}},w)}\). We define the algebra \(AP_{X({\mathbb {R}},w)}\) of almost periodic Fourier multipliers by

$$\begin{aligned} AP_{X({\mathbb {R}},w)}:={\text {clos}}_{{\mathcal {M}}_{X({\mathbb {R}},w)}}\big (APP\big ). \end{aligned}$$

It is natural to refer to the weights in \(A_{1/\alpha _X}^0\cap A_{1/\beta _X}^0\) as suitable Muckenhoupt weights. The class of suitable Muckenhoput weights contains many nontrivial weights as the following example shows.

For \(\delta ,\nu ,\eta \in {\mathbb {R}}\), consider the weight

$$\begin{aligned} w(x):=\left\{ \begin{array}{lll} \exp \big ( \delta +\nu \sin (\eta \log (\log |x|)) \big ) &{} \quad \text{ if }&{} |x|\ge e, \\ \exp (\delta ) &{} \quad \text{ if }&{} |x|<e. \end{array}\right. \end{aligned}$$

Let \(r\in (1,\infty )\). It was shown in [27, Example 4.2] that if

$$\begin{aligned} -1/r<\delta -|\nu |\sqrt{\eta ^2+1} \le \delta +|\nu |\sqrt{\eta ^2+1}<1-1/r, \end{aligned}$$

then \(w\in A_r^0({\mathbb {R}})\). Hence if \(0<\alpha _X\le \beta _X<1\) and

$$\begin{aligned} -\alpha _X<\delta -|\nu |\sqrt{\eta ^2+1} \le \delta +|\nu |\sqrt{\eta ^2+1}<1-\beta _X, \end{aligned}$$

then \(w\in A_{1/\alpha _X}^0({\mathbb {R}})\cap A_{1/\beta _X}^0({\mathbb {R}})\).

5.2 Products of almost periodic Fourier multipliers and continuous Fourier multipliers vanishing at infinity

The next lemma generalizes [29, Lemma 3.1(iii)] from the setting of Lebesgue spaces to the setting of rearrangement-invariant Banach function spaces with suitable Muckenhoupt weights.

Lemma 5.3

Let \(X({\mathbb {R}})\) be a separable rearrangement-invariant Banach function space with the Boyd indices satisfying \(0<\alpha _X\), \(\beta _X<1\). Suppose that w belongs to \(A_{1/\alpha _X}^0({\mathbb {R}})\cap A_{1/\beta _X}^0({\mathbb {R}})\) and \(C_{0,X({\mathbb {R}},w)}({{\dot{{\mathbb {R}}}}})\) is defined by (3.17). If \(a\in AP_{X({\mathbb {R}},w)}\) and \(\psi \in C_{0,X({\mathbb {R}},w)}({{\dot{{\mathbb {R}}}}})\), then \(a\psi \in C_{0,X({\mathbb {R}},w)}({{\dot{{\mathbb {R}}}}})\).

Proof

By Theorem 3.3, there exists a sequence \(\{\psi _n\}_{n\in {\mathbb {N}}}\subset C_c^\infty ({\mathbb {R}})\) such that

$$\begin{aligned} \lim _{n\rightarrow \infty }\Vert \psi _n-\psi \Vert _{{\mathcal {M}}_{X({\mathbb {R}},w)}}=0. \end{aligned}$$
(5.1)

By the definition of the algebra \(AP_{X({\mathbb {R}},w)}\), there exists a sequence \(a_n\in APP\) such that

$$\begin{aligned} \lim _{n\rightarrow \infty }\Vert a_n-a\Vert _{{\mathcal {M}}_{X({\mathbb {R}},w)}}=0. \end{aligned}$$
(5.2)

Then \(a_n\psi _n\in C_c^\infty ({\mathbb {R}})\subset C({{\dot{{\mathbb {R}}}}})\cap V({\mathbb {R}})\) for every \(n\in {\mathbb {N}}\). Moreover, (5.1)–(5.2) imply that

$$\begin{aligned} \lim _{n\rightarrow \infty }\Vert a_n\psi _n-a\psi \Vert _{{\mathcal {M}}_{X({\mathbb {R}},w)}}=0. \end{aligned}$$

Hence \(a\psi \in C_{X({\mathbb {R}},w)}({{\dot{{\mathbb {R}}}}})\). In view of the continuous embedding of \({\mathcal {M}}_{X({\mathbb {R}},w)}\) into \(L^\infty ({\mathbb {R}})\) (see Theorem 2.4) and the above equality, we obtain

$$\begin{aligned} |(a\psi )(\infty )|&= \lim _{n\rightarrow \infty }|(a_n\psi _n)(\infty )-(a\psi )(\infty )| \le \lim _{n\rightarrow \infty }\Vert a_n\psi _n-a\psi \Vert _{L^\infty ({\mathbb {R}})} \\&\le \lim _{n\rightarrow \infty }\Vert a_n\psi _n-a\psi \Vert _{{\mathcal {M}}_{X(R,w)}}=0. \end{aligned}$$

Thus \((a\psi )(\infty )=0\) and \(a\psi \in C_{0,X({\mathbb {R}},w)}({{\dot{{\mathbb {R}}}}})\). \(\square \)

6 Proof of the main result

6.1 The algebra \({\mathcal {A}}_u\)

For a real-valued monotonically increasing function \(u\in C({\overline{{\mathbb {R}}}})\) such that

$$\begin{aligned} u(-\infty )=0 \quad u(+\infty )=1, \end{aligned}$$
(6.1)

consider the set

$$\begin{aligned} {\mathcal {A}}_u:= \left\{ a=(1-u)a_l+ua_r+a_0\ :\ a_l,a_r\in AP_{X({\mathbb {R}},w)},\ a_0\in C_{0,X({\mathbb {R}},w)}({{\dot{{\mathbb {R}}}}}) \right\} . \end{aligned}$$

Lemma 6.1

Let \(X({\mathbb {R}})\) be a separable rearrangement-invariant Banach function space with the Boyd indices satisfying \(0<\alpha _X\), \(\beta _X<1\). Suppose that \(w\in A_{1/\alpha _X}^0({\mathbb {R}})\cap A_{1/\beta _X}^0({\mathbb {R}})\). If \(u\in C({\overline{{\mathbb {R}}}})\) is a real-valued monotonically increasing function such that \(u(-\infty )=0\) and \(u(+\infty )=1\), then the set \({\mathcal {A}}_u\) is an algebra and the mappings \(a\mapsto a_l\) and \(a\mapsto a_r\) are algebraic homomorphisms of \({\mathcal {A}}_u\) onto \(AP_{X({\mathbb {R}},w)}\).

Proof

If \(a,b\in {\mathcal {A}}_u\), then

$$\begin{aligned} a=(1-u)a_l+ua_r+a_0, \quad b=(1-u)b_l+ub_r+b_0 \end{aligned}$$

with some \(a_l,a_r,b_l,b_r\in AP_{X({\mathbb {R}},w)}\) and \(a_0,b_0\in C_{0,X({\mathbb {R}},w)}({{\dot{{\mathbb {R}}}}})\). Therefore

$$\begin{aligned} a+b=(1-u)(a_l+b_l)+u(a_r+b_r)+(a_0+b_0) \end{aligned}$$
(6.2)

and

$$\begin{aligned} ab&=(1-u)^2a_lb_l+u^2a_rb_r+(1-u)u(a_lb_r+a_rb_l) \nonumber \\&\quad + \big ((1-u)a_l+ua_r\big )b_0+ \big ((1-u)b_l+ub_r\big )a_0+a_0b_0 \nonumber \\&= (1-u)a_lb_l+ua_rb_r+c_0, \end{aligned}$$
(6.3)

where

$$\begin{aligned} c_0&= (u-u^2)\big [(a_lb_r+a_rb_l)-(a_lb_l+a_rb_r)\big ] \nonumber \\&\quad + \big ((1-u)a_l+ua_r\big )b_0+ \big ((1-u)b_l+ub_r\big )a_0+a_0b_0. \end{aligned}$$
(6.4)

Since \(1-u,u\in C({\overline{{\mathbb {R}}}})\cap V({\mathbb {R}})\subset C_{X({\mathbb {R}},w)}({\overline{{\mathbb {R}}}})\) and \(a_0,b_0\in C_{0,X({\mathbb {R}},w)}({\mathbb {R}})\), it follows from Lemma 4.1 that

$$\begin{aligned} (1-u)a_0, ua_0,(1-u)b_0,ub_0\in C_{0,X({\mathbb {R}},w)}({{\dot{{\mathbb {R}}}}}). \end{aligned}$$

Then, by Lemma 5.3,

$$\begin{aligned} (1-u)a_lb_0,ua_rb_0,(1-u)b_la_0,ub_ra_0\in C_{0,X({\mathbb {R}},w)}({{\dot{{\mathbb {R}}}}}). \end{aligned}$$
(6.5)

Since \(u-u^2\in C({\overline{{\mathbb {R}}}})\cap V({\mathbb {R}})\subset C_{X({\mathbb {R}},w)}({\overline{{\mathbb {R}}}})\) and \(u(\pm \infty )-u^2(\pm \infty )=0\), by Lemma 4.1, \(u-u^2\in C_{0,X({\mathbb {R}},w)}({{\dot{{\mathbb {R}}}}})\). Then, in view of Lemma 5.3, we also conclude that

$$\begin{aligned} (u-u^2)\big [(a_lb_r+a_rb_l)-(a_lb_l+a_rb_r)\big ] \in C_{0,X({\mathbb {R}},w)}({{\dot{{\mathbb {R}}}}}). \end{aligned}$$
(6.6)

It follows from (6.4) to (6.6) that \(c_0\in C_{0,X({\mathbb {R}},w)}({{\dot{{\mathbb {R}}}}})\). In view of this observation and equalities (6.2)–(6.3), we see that \(a+b,ab\in {\mathcal {A}}_u\). Therefore, \({\mathcal {A}}_u\) is an algebra. It is clear that the mappings \(a\mapsto a_l\) and \(a\mapsto a_r\) are algebraic homomorphisms of \({\mathcal {A}}_u\) onto \(AP_{X({\mathbb {R}},w)}\). \(\square \)

6.2 The multiplier norm of \(a=(1-u)a_r +ua_r+a_0\in {\mathcal {A}}_u\) dominates the multiplier norms of \(a_r\) and \(a_l\)

In this section we will prepare the proof of the fact that the algebraic homomorphisms \({\mathcal {A}}_u\rightarrow AP_{X({\mathbb {R}},w)}\) given by \(a\mapsto a_l\) and \(a\mapsto a_r\) are actually Banach algebra homomorphisms of norm 1. To this end, we will show that for \(a\in {\mathcal {A}}_u\),

$$\begin{aligned} \Vert a_r\Vert _{{\mathcal {M}}_{X({\mathbb {R}},w)}}\le \Vert a\Vert _{{\mathcal {M}}_{X({\mathbb {R}},w)}}, \quad \Vert a_l\Vert _{{\mathcal {M}}_{X({\mathbb {R}},w)}}\le \Vert a\Vert _{{\mathcal {M}}_{X({\mathbb {R}},w)}}. \end{aligned}$$
(6.7)

For \(a\in L^\infty ({\mathbb {R}})\) and \(h\in {\mathbb {R}}\), we define

$$\begin{aligned} a^h(x):=a(x+h), \quad x\in {\mathbb {R}}. \end{aligned}$$

The following consequence of Kronecker’s theorem (see, e.g., [10, Theorem 1.12]) plays a crucial role in the proof of inequalities (6.7).

Lemma 6.2

If \(a_1,\ldots ,a_k\in APP\) is a finite collection of almost periodic polynomials, then there exists a sequence \(\{h_n\}_{n\in {\mathbb {N}}}\) of real numbers such that \(h_n\rightarrow +\infty \) as \(n\rightarrow \infty \) and

$$\begin{aligned} \lim _{n\rightarrow \infty }\Vert a_m^{\pm h_n}-a_m\Vert _{L^\infty ({\mathbb {R}})}=0 \end{aligned}$$

for each \(m\in \{1,\ldots ,k\}\).

For the sign “+”, the proof of the above lemma is given in [10, Lemma 10.2], for the sign “–”, the proof is analogous.

We start the proof of inequalities (6.7) for \(a=(1-v)a_l+va_r+a_0\) with a nice function v in place of u and nice functions \(a_l,a_r\) and \(a_0\).

Lemma 6.3

Let \(X({\mathbb {R}})\) be a separable rearrangement-invariant Banach function space with the Boyd indices satisfying \(0<\alpha _X\), \(\beta _X<1\). Suppose that \(w\in A_{1/\alpha _X}^0({\mathbb {R}})\cap A_{1/\beta _X}^0({\mathbb {R}})\). Let \(v\in C({\overline{{\mathbb {R}}}})\) be any real-valued monotonically increasing function such that there exists a point \(x_0>0\) such that \(v(x)=0\) for \(x<-x_0\) and \(v(x)=1\) for \(x>x_0\). If \(a_l,a_r\in APP\), \(a_0\in C_c^\infty ({\mathbb {R}})\), and

$$\begin{aligned} a=(1-v)a_l+va_r+a_0, \end{aligned}$$
(6.8)

then inequalities (6.7) hold.

Proof

The idea of the proof is borrowed from the proof of [27, Theorem 3.1]. By Lemma 6.2, there is a sequence \(\{h_n\}_{n\in {\mathbb {N}}}\) of real numbers such that \(h_n\rightarrow +\infty \) as \(n\rightarrow \infty \) and

$$\begin{aligned}&\lim _{n\rightarrow \infty }\Vert a_r^{h_n}-a_r\Vert _{L^\infty ({\mathbb {R}})}=0, \quad \lim _{n\rightarrow \infty }\Vert (a_r')^{h_n}-a_r'\Vert _{L^\infty ({\mathbb {R}})}=0, \end{aligned}$$
(6.9)
$$\begin{aligned}&\lim _{n\rightarrow \infty }\Vert a_l^{-h_n}-a_l\Vert _{L^\infty ({\mathbb {R}})}=0, \quad \lim _{n\rightarrow \infty }\Vert (a_l')^{-h_n}-a_l'\Vert _{L^\infty ({\mathbb {R}})}=0. \end{aligned}$$
(6.10)

Let us show that

$$\begin{aligned} \mathop {\text {s-lim}}\limits _{n\rightarrow \infty }e_{h_n}W^0(a)e_{-h_n}I=W^0(a_r), \quad \mathop {\text {s-lim}}\limits _{n\rightarrow \infty }e_{-h_n}W^0(a)e_{h_n}I=W^0(a_l) \end{aligned}$$
(6.11)

on the space \(X({\mathbb {R}},w)\). As

$$\begin{aligned} e_{\pm h_n}W^0(a)e_{\mp h_n}I=W^0(a^{\pm h_n}), \end{aligned}$$

we have to prove that for every \(f\in X({\mathbb {R}},w)\),

$$\begin{aligned}&\lim _{n\rightarrow \infty }\left\| W^0(a^{h_n}-a_r)f\right\| _{X({\mathbb {R}},w)}=0, \end{aligned}$$
(6.12)
$$\begin{aligned}&\lim _{n\rightarrow \infty }\left\| W^0(a^{-h_n}-a_l)f\right\| _{X({\mathbb {R}},w)}=0. \end{aligned}$$
(6.13)

Since the operators \(W^0(a^{h_n}-a_r)\) and \(W^0(a^{-h_n}-a_l)\) are uniformly bounded in \(n\in {\mathbb {N}}\) and the set \({\mathcal {S}}_0({\mathbb {R}})\) is dense in the space \(X({\mathbb {R}},w)\) in view of Lemma 2.3, applying [34, Lemma 1.4.1], we conclude that it is enough to prove equalities (6.12)–(6.13) for all \(f\in {\mathcal {S}}_0({\mathbb {R}})\).

Fix \(f\in {\mathcal {S}}_0({\mathbb {R}})\). Then, by a smooth version of Urysohn’s lemma (see, e.g., [17, Proposition 6.5]), there is a function \(\psi \in C_c^\infty ({\mathbb {R}})\) such that \(0\le \psi \le 1\), \({\text {supp}}{\mathcal {F}}f\subset {\text {supp}}\psi \) and \(\psi |_{{\text {supp}}{\mathcal {F}}f}=1\). Therefore, for all \(n\in {\mathbb {N}}\),

$$\begin{aligned} W^0(a^{h_n}-a_r)f={\mathcal {F}}^{-1}(a^{h_n}-a_r)\psi {\mathcal {F}}f, \quad W^0(a^{-h_n}-a_l)f={\mathcal {F}}^{-1}(a^{-h_n}-a_l)\psi {\mathcal {F}}f \end{aligned}$$

and

$$\begin{aligned}&\left\| W^0(a^{h_n}-a_r)f\right\| _{X({\mathbb {R}},w)} \le \Vert (a^{h_n}-a_r)\psi \Vert _{{\mathcal {M}}_{X({\mathbb {R}},w)}}\Vert f\Vert _{X({\mathbb {R}},w)}, \end{aligned}$$
(6.14)
$$\begin{aligned}&\left\| W^0(a^{-h_n}-a_l)f\right\| _{X({\mathbb {R}},w)} \le \Vert (a^{-h_n}-a_l)\psi \Vert _{{\mathcal {M}}_{X({\mathbb {R}},w)}}\Vert f\Vert _{X({\mathbb {R}},w)}. \end{aligned}$$
(6.15)

Since \(v(x)=1\) for \(x>x_0\) and \(v(x)=0\) for \(x<-x_0\) and \(a_0\in C_c^\infty ({\mathbb {R}})\), there exists \(N\in {\mathbb {N}}\) such that for all \(x\in {\text {supp}}\psi \) and \(n>N\),

$$\begin{aligned} v(x+h_n)=1, \quad v(x-h_n)=0, \quad a_0(x\pm h_n)=0. \end{aligned}$$

Hence, for all \(n>N\) and \(x\in {\mathbb {R}}\),

$$\begin{aligned} \big (a^{h_n}(x)-a_r(x)\big )\psi (x)&= \big (a_r^{h_n}(x)-a_r(x)\big )\psi (x), \end{aligned}$$
(6.16)
$$\begin{aligned} \big (a^{-h_n}(x)-a_l(x)\big )\psi (x)&= \big (a_l^{-h_n}(x)-a_l(x)\big )\psi (x). \end{aligned}$$
(6.17)

It is clear that the functions on the right-hand sides of (6.16)–(6.17) belong to \(C_c^\infty ({\mathbb {R}})\). Therefore, by the Stechkin-type inequality (1.4), for all \(n>N\),

$$\begin{aligned} \big \Vert \big (a^{h_n}-a_r\big )\psi \big \Vert _{{\mathcal {M}}_{X({\mathbb {R}},w)}}&= \big \Vert \big (a_r^{h_n}-a_r\big )\psi \big \Vert _{{\mathcal {M}}_{X({\mathbb {R}},w)}} \nonumber \\&\le c_{X({\mathbb {R}},w)}\big \Vert \big (a_r^{h_n}-a_r\big )\psi \big \Vert _{V({\mathbb {R}})} \nonumber \\&= c_{X({\mathbb {R}},w)}\big \Vert \big (a_r^{h_n}-a_r\big )\psi \big \Vert _{L^\infty ({\mathbb {R}})} \nonumber \\&\quad + c_{X({\mathbb {R}},w)}\int _{\mathbb {R}}\left| (a_r^{h_n})'(x)-a_r'(x)\right| \,|\psi (x)|\,dx \nonumber \\&\quad + c_{X({\mathbb {R}},w)}\int _{\mathbb {R}}\left| a_r^{h_n}(x)-a_r(x)\right| \,|\psi '(x)|\,dx \nonumber \\&\le c_{X({\mathbb {R}},w)}\big (\Vert \psi \Vert _{L^\infty ({\mathbb {R}})}+\Vert \psi '\Vert _{L^1({\mathbb {R}})}\big ) \big \Vert a_r^{h_n}-a_r\big \Vert _{L^\infty ({\mathbb {R}})} \nonumber \\&\quad + c_{X({\mathbb {R}},w)}\Vert \psi \Vert _{L^1({\mathbb {R}})} \big \Vert (a_r^{h_n})'-a_r'\big \Vert _{L^\infty ({\mathbb {R}})} \end{aligned}$$
(6.18)

and, analogously,

$$\begin{aligned} \big \Vert \big (a^{-h_n}-a_l\big )\psi \big \Vert _{{\mathcal {M}}_{X({\mathbb {R}},w)}}&\le c_{X({\mathbb {R}},w)}\big (\Vert \psi \Vert _{L^\infty ({\mathbb {R}})}+\Vert \psi '\Vert _{L^1({\mathbb {R}})}\big ) \big \Vert a_l^{-h_n}-a_l\big \Vert _{L^\infty ({\mathbb {R}})} \nonumber \\&\quad + c_{X({\mathbb {R}},w)}\Vert \psi \Vert _{L^1({\mathbb {R}})} \big \Vert (a_l^{-h_n})'-a_l'\big \Vert _{L^\infty ({\mathbb {R}})}. \end{aligned}$$
(6.19)

Combining (6.14)–(6.15) and (6.18)–(6.19) with (6.9)–(6.10), we see that equalities (6.12)–(6.13) hold for every \(f\in {\mathcal {S}}_0({\mathbb {R}})\). Therefore, (6.11) are fulfilled for every \(f\in X({\mathbb {R}},w)\). Hence, by the Banach-Steinhaus theorem (see, e.g., [34, Theorem 1.4.2]),

$$\begin{aligned} \Vert a_r\Vert _{{\mathcal {M}}_{X({\mathbb {R}},w)}}&= \Vert W^0(a_r)\Vert _{{\mathcal {B}}(X({\mathbb {R}},w))} \le \liminf _{n\rightarrow \infty }\Vert e_{h_n}W^0(a)e_{-h_n}I\Vert _{{\mathcal {B}}(X({\mathbb {R}},w))} \\&\le \Vert W^0(a)\Vert _{{\mathcal {B}}(X({\mathbb {R}},w))}=\Vert a\Vert _{{\mathcal {M}}_{X({\mathbb {R}},w)}} \end{aligned}$$

and, analogously,

$$\begin{aligned} \Vert a_l\Vert _{{\mathcal {M}}_{X({\mathbb {R}},w)}}&= \Vert W^0(a_l)\Vert _{{\mathcal {B}}(X({\mathbb {R}},w))} \le \liminf _{n\rightarrow \infty }\Vert e_{-h_n}W^0(a)e_{h_n}I\Vert _{{\mathcal {B}}(X({\mathbb {R}},w))} \\&\le \Vert W^0(a)\Vert _{{\mathcal {B}}(X({\mathbb {R}},w))}=\Vert a\Vert _{{\mathcal {M}}_{X({\mathbb {R}},w)}}, \end{aligned}$$

which completes the proof of (6.7). \(\square \)

Now we extend the previous result for functions a of the form (6.8) with general \(a_l,a_r\in AP_{X({\mathbb {R}},w)}\) and \(a_0\in C_{0,X({\mathbb {R}},w)}({{\dot{{\mathbb {R}}}}})\), keeping the nice function v as above.

Lemma 6.4

Let \(X({\mathbb {R}})\) be a separable rearrangement-invariant Banach function space with the Boyd indices satisfying \(0<\alpha _X\), \(\beta _X<1\). Suppose that \(w\in A_{1/\alpha _X}^0({\mathbb {R}})\cap A_{1/\beta _X}^0({\mathbb {R}})\). Let \(v\in C({\overline{{\mathbb {R}}}})\) be any real-valued monotonically increasing function such that there exists a point \(x_0>0\) such that \(v(x)=0\) for \(x<-x_0\) and \(v(x)=1\) for \(x>x_0\). If \(a_l,a_r\in AP_{X({\mathbb {R}},w)}\), \(a_0\in C_{0,X({\mathbb {R}},w)}({{\dot{{\mathbb {R}}}}})\), where \(C_{0,X({\mathbb {R}},w)}({{\dot{{\mathbb {R}}}}})\) is defined by (3.17), and a is given by equality (6.8), then inequalities (6.7) hold.

Proof

By the definition of \(AP_{X({\mathbb {R}},w)}\), there are sequences \(\{a_l^{(n)}\}_{n\in {\mathbb {N}}},\{a_r^{(n)}\}_{n\in {\mathbb {N}}}\) in APP such that

$$\begin{aligned} \lim _{n\rightarrow \infty }\big \Vert a_l^{(n)}-a_l\big \Vert _{{\mathcal {M}}_{X({\mathbb {R}},w)}}=0, \quad \lim _{n\rightarrow \infty }\big \Vert a_r^{(n)}-a_r\big \Vert _{{\mathcal {M}}_{X({\mathbb {R}},w)}}=0. \end{aligned}$$
(6.20)

On the other hand, by Theorem 3.3, there is a sequence \(\{a_0^{(n)}\}_{n\in {\mathbb {N}}}\) in \(C_c^\infty ({\mathbb {R}})\) such that

$$\begin{aligned} \lim _{n\rightarrow \infty }\big \Vert a_0^{(n)}-a_0\big \Vert _{{\mathcal {M}}_{X({\mathbb {R}},w)}}=0. \end{aligned}$$
(6.21)

For \(n\in {\mathbb {N}}\), consider the functions

$$\begin{aligned} a^{(n)}:=(1-v)a_l^{(n)}+va_r^{(n)}+a_0^{(n)}. \end{aligned}$$
(6.22)

It follows from equalities (6.20)–(6.22) and Lemma 6.3 that

$$\begin{aligned} \Vert a_l\Vert _{{\mathcal {M}}_{X({\mathbb {R}},w)}}&= \lim _{n\rightarrow \infty }\big \Vert a_l^{(n)}\big \Vert _{{\mathcal {M}}_{X({\mathbb {R}},w)}} \le \lim _{n\rightarrow \infty }\big \Vert a^{(n)}\big \Vert _{{\mathcal {M}}_{X({\mathbb {R}},w)}} = \Vert a\Vert _{{\mathcal {M}}_{X({\mathbb {R}},w)}}, \\ \Vert a_r\Vert _{{\mathcal {M}}_{X({\mathbb {R}},w)}}&= \lim _{n\rightarrow \infty }\big \Vert a_r^{(n)}\big \Vert _{{\mathcal {M}}_{X({\mathbb {R}},w)}} \le \lim _{n\rightarrow \infty }\big \Vert a^{(n)}\big \Vert _{{\mathcal {M}}_{X({\mathbb {R}},w)}} = \Vert a\Vert _{{\mathcal {M}}_{X({\mathbb {R}},w)}}, \end{aligned}$$

which completes the proof of inequalities (6.7). \(\square \)

Now we observe that the algebra \({\mathcal {A}}_u\) does not depend on the particular choice of a real-valued monotonically increasing function \(u\in C({\overline{{\mathbb {R}}}})\) satisfying conditions (6.1).

Lemma 6.5

Let \(X({\mathbb {R}})\) be a separable rearrangement-invariant Banach function space with the Boyd indices satisfying \(0<\alpha _X\), \(\beta _X<1\). Suppose that \(w\in A_{1/\alpha _X}^0({\mathbb {R}})\cap A_{1/\beta _X}^0({\mathbb {R}})\). Let \(u,v\in C({\overline{{\mathbb {R}}}})\) be two real-valued monotonically increasing functions such that

$$\begin{aligned} u(-\infty )=v(-\infty )=0, \quad u(+\infty )=v(+\infty )=1. \end{aligned}$$

Then \({\mathcal {A}}_u={\mathcal {A}}_v\).

Proof

If \(a\in {\mathcal {A}}_u\), then \(a=(1-u)a_l+ua_r+a_0\) for some \(a_l,a_r\in AP_{X({\mathbb {R}},w)}\) and \(a_0\in C_{0,X({\mathbb {R}},w)}({{\dot{{\mathbb {R}}}}})\). On the other hand, \(a=(1-v)a_l+va_r+b_0\) with

$$\begin{aligned} b_0=(v-u)a_l+(u-v)a_r+a_0=(u-v)(a_r-a_l)+a_0. \end{aligned}$$

Since the functions uv are monotonically increasing, we have \(u,v\in V({\mathbb {R}})\). Hence \(u-v\in V({\mathbb {R}})\cap C({\overline{{\mathbb {R}}}})\) and

$$\begin{aligned} u(+\infty )-v(+\infty )=u(-\infty )-v(-\infty )=0. \end{aligned}$$

Thus \(u-v\in C({{\dot{{\mathbb {R}}}}})\cap V({\mathbb {R}})\subset C_{X({\mathbb {R}},w)}({{\dot{{\mathbb {R}}}}})\) and \((u-v)(\infty )=0\). Since the function \(a_r-a_l\) belongs to \(AP_{X({\mathbb {R}},w)}\), it follows from Lemma 5.3 that

$$\begin{aligned} (u-v)(a_r-a_l)\in C_{0,X({\mathbb {R}},w)}({{\dot{{\mathbb {R}}}}}). \end{aligned}$$

Then \(b_0\in C_{0,X({\mathbb {R}},w)}({{\dot{{\mathbb {R}}}}})\) and \(a\in {\mathcal {A}}_v\). Therefore \({\mathcal {A}}_u\subset {\mathcal {A}}_v\). It can be shown analogously that \({\mathcal {A}}_v\subset {\mathcal {A}}_u\). Thus \({\mathcal {A}}_u={\mathcal {A}}_v\). \(\square \)

Combining Lemmas 6.46.5, we arrive at the main result of this subsection.

Theorem 6.6

Let \(X({\mathbb {R}})\) be a separable rearrangement-invariant Banach function space with the Boyd indices satisfying \(0<\alpha _X\), \(\beta _X<1\). Suppose that \(w\in A_{1/\alpha _X}^0({\mathbb {R}})\cap A_{1/\beta _X}^0({\mathbb {R}})\). Let \(u\in C({\overline{{\mathbb {R}}}})\) be a real-valued monotonically increasing function such that \(u(-\infty )=0\) and \(u(+\infty )=1\). If \(a\in {\mathcal {A}}_u\), that is,

$$\begin{aligned} a=(1-u)a_l+ua_r+a_0 \quad \text{ with }\quad a_l,a_r\in AP_{X({\mathbb {R}},w)},a_0\in C_{0,X({\mathbb {R}},w)}({{\dot{{\mathbb {R}}}}}), \end{aligned}$$

where \(C_{0,X({\mathbb {R}},w)}({{\dot{{\mathbb {R}}}}})\) is defined by (3.17), then inequalities (6.7) hold.

6.3 Proof of Theorem 1.2

The idea of the proof is borrowed from the proof of [10, Theorem 1.21]. If \(a\in AP_{X({\mathbb {R}},w)}\), then \(a=(1-u)a+ua+0\), whence \(a\in {\mathcal {A}}_u\). If \(f\in C_{X({\mathbb {R}},w)}({\overline{{\mathbb {R}}}})\), then the function \(f_0=f-(1-u)f(-\infty )-uf(+\infty )\) belongs to \(C_{0,X({\mathbb {R}},w)}({{\dot{{\mathbb {R}}}}})\). Therefore \(f=(1-u)f(-\infty )+uf(+\infty )+f_0\in {\mathcal {A}}_u\). These observations imply that

$$\begin{aligned} SAP_{X({\mathbb {R}},w)}\subset {\text {clos}}_{{\mathcal {M}}_{X({\mathbb {R}},w)}}({\mathcal {A}}_u). \end{aligned}$$
(6.23)

On the other hand, it is obvious that

$$\begin{aligned} {\text {clos}}_{{\mathcal {M}}_{X({\mathbb {R}},w)}}({\mathcal {A}}_u)\subset SAP_{X({\mathbb {R}},w)}. \end{aligned}$$
(6.24)

Combining (6.23)–(6.24), we arrive at the equality

$$\begin{aligned} SAP_{X({\mathbb {R}},w)}={\text {clos}}_{{\mathcal {M}}_{X({\mathbb {R}},w)}}({\mathcal {A}}_u). \end{aligned}$$
(6.25)

By Theorem 6.6, for every \(a=(1-u)a_r+ua_r+a_0\in {\mathcal {A}}_u\) with \(a_l,a_r\in AP_{X({\mathbb {R}},w)}\) and \(a_0\in C_{0,X({\mathbb {R}},w)}({{\dot{{\mathbb {R}}}}})\), one has

$$\begin{aligned} \Vert a_r\Vert _{{\mathcal {M}}_{X({\mathbb {R}},w)}}\le \Vert a\Vert _{{\mathcal {M}}_{X({\mathbb {R}},w)}}, \quad \Vert a_l\Vert _{{\mathcal {M}}_{X({\mathbb {R}},w)}}\le \Vert a\Vert _{{\mathcal {M}}_{X({\mathbb {R}},w)}}. \end{aligned}$$
(6.26)

Consequently, if \(\big \{(1-u)a_l^{(n)}+ua_r^{(n)}+a_0^{(n)}\big \}_{n\in {\mathbb {N}}}\) is a Cauchy sequence in \({\mathcal {A}}_u\), where \(\big \{a_l^{(n)}\big \}_{n\in {\mathbb {N}}}, \big \{a_r^{(n)}\big \}_{n\in {\mathbb {N}}}\) are sequences in \(AP_{X({\mathbb {R}},w)}\) and \(\big \{a_0^{(n)}\big \}_{n\in {\mathbb {N}}}\) is a sequence in \(C_{0,X({\mathbb {R}},w)}({{\dot{{\mathbb {R}}}}})\), then \(\big \{a_l^{(n)}\big \}_{n\in {\mathbb {N}}}\) and \(\big \{a_r^{(n)}\big \}_{n\in {\mathbb {N}}}\) are Cauchy sequences in \(AP_{X({\mathbb {R}},w)}\). Consequently, \(\big \{a_0^{(n)}\big \}_{n\in {\mathbb {N}}}\) is a Cauchy sequence in \(C_{0,X({\mathbb {R}},w)}({{\dot{{\mathbb {R}}}}})\). Since \(AP_{X({\mathbb {R}},w)}\) is closed by its definition and \(C_{0,X({\mathbb {R}},w)}({{\dot{{\mathbb {R}}}}})\) is closed in view of Theorem 3.3, we conclude that the limits

$$\begin{aligned} a_l:=\lim _{n\rightarrow \infty }a_l^{(n)}, \quad a_r=\lim _{n\rightarrow \infty }a_r^{(n)} \end{aligned}$$

belong to \(AP_{X({\mathbb {R}},w)}\) and that the limit

$$\begin{aligned} a_0:=\lim _{n\rightarrow \infty }a_0^{(n)} \end{aligned}$$

belongs to \(C_{0,X({\mathbb {R}},w)}({{\dot{{\mathbb {R}}}}})\). Therefore, the limit

$$\begin{aligned} \lim _{n\rightarrow \infty } \big ((1-u)a_l^{(n)}+ua_r^{(n)}+a_0^{(n)}\big ) \end{aligned}$$

belongs to \({\mathcal {A}}_u\). Thus

$$\begin{aligned} {\text {clos}}_{{\mathcal {M}}_{X({\mathbb {R}},w)}}({\mathcal {A}}_u)={\mathcal {A}}_u. \end{aligned}$$
(6.27)

It follows from (6.25) and (6.27) that \({\mathcal {A}}_u=SAP_{X({\mathbb {R}},w)}\). In particular, every function \(a\in SAP_{X({\mathbb {R}},w)}\) is of the form

$$\begin{aligned} a=(1-u)a_l+ua_r+a_0 \end{aligned}$$
(6.28)

with \(a_l,a_r\in AP_{X({\mathbb {R}},w)}\) and \(a_0\in C_{0,X({\mathbb {R}},w)}\). We infer from (6.26) that the representation (6.28) is unique for the function u. Moreover, the proof of Lemma 6.5 shows that \(a_l,a_r\in AP_{X({\mathbb {R}},w)}\) are independent of the particular choice of the function u. By Lemma 6.1, the mappings \(a\mapsto a_l\) and \(a\mapsto a_r\) are algebraic homomorphisms of \({\mathcal {A}}_u=SAP_{X({\mathbb {R}},w)}\) onto \(AP_{X({\mathbb {R}},w)}\). In view of (6.26), these homomorphisms are Banach algebra homomorphisms of the Banach algebra \(SAP_{X({\mathbb {R}},w)}\) onto the Banach algebra \(AP_{X({\mathbb {R}},w)}\) and the norms of these homomorphisms are not greater than one. For every function \(a\in AP_{X({\mathbb {R}},w)}\), we have equalities in (6.26) because

$$\begin{aligned} a=(1-u)a+ua+0=a_l=a_r. \end{aligned}$$

Thus, the norms of the homomorphisms \(a\mapsto a_l\) and \(a\mapsto a_l\) are equal to one. \(\square \)