1 Introduction

Given f an \(L_{1}({\mathbb {R}}^n )\) function, its Fourier transform, defined by

$$\begin{aligned} ( {\mathscr {F}}f )(\xi )= {\hat{f}}(\xi ) = \int _{{\mathbb {R}}^{n}} f(x) e^{- 2 \pi i x \cdot \xi } \, dx,\quad \xi \in {\mathbb {R}}^n, \end{aligned}$$

satisfies the inequality

$$\begin{aligned} \Vert {\hat{f}} \Vert _{\infty } \le \Vert f\Vert _{1}. \end{aligned}$$

Plancherel, in 1910, proved the n-dimensional version of the Riesz–Fischer theorem, namely

$$\begin{aligned} \Vert {\hat{f}} \Vert _2=\Vert f\Vert _2. \end{aligned}$$

Standard interpolation theorems yield that \(L_{p'} ({\mathbb {R}}^n),\) \(p'= \frac{p}{p-1},\) is an interpolation space (defined in Sect. 2) between \(L_{2} ({\mathbb {R}}^n)\) and \(L_{\infty } ({\mathbb {R}}^n)\) for \(1<p<2,\) leading to the Hausdorff–Young inequality (1926),

$$\begin{aligned} \Vert {\hat{f}} \Vert _{p'} \le C_p\Vert f\Vert _p, \end{aligned}$$

in this case.

Inspired by the work of Jodeit and Torchinsky [13], in which the authors have generalized the Hausdorff–Young inequality, replacing the \(L_p\) spaces with Orlicz spaces, we prove the following theorem which is central to the rest of the results in this paper.

Theorem 1.1

Let \(\rho (f) = {\bar{\rho }}(f^{*}) \) be an r.i. norm such that the Banach space \(L_{\rho } ({\mathbb {R}}^n)\) is an interpolation space between \(L_{2}( {\mathbb {R}}^{n} )\) and \(L_{\infty }( {\mathbb {R}}^{n} )\). Then,

$$\begin{aligned} \rho ({\hat{f}}) \le C \sigma (f), \end{aligned}$$
(1.1)

for any r.i. norm \(\sigma \) if and only if

$$\begin{aligned} {\bar{\rho }}( U f^{*}) \le C \, {\bar{\sigma }}( f^{*}), \end{aligned}$$
(1.2)

where \(C>0\) is independent of \(f \in L_{\sigma }({\mathbb {R}}^n)\).

For r.i norms \(\rho =\rho _{p'}\) and \(\sigma = \rho _{p},\) where \(\rho _{p}(f) =\Vert f\Vert _p \), \(1< p < \infty \), \(\textstyle {\frac{1}{p} + \frac{1}{p'}=1}\), the space \(L_{\rho _{p'}} ({\mathbb {R}}^n) = L_{p'} ({\mathbb {R}}^n)\) is an interpolation space between \(L_{2} ({\mathbb {R}}^n)\) and \(L_{\infty } ({\mathbb {R}}^n)\) when \(1<p<2\) and the inequality (1.2), amounts to

$$\begin{aligned} \begin{aligned} C \, {\bar{\rho }}_{p} (f^{*}) \ge {\bar{\rho }}_{p'} ( U f^{*}) = \left[ \int _{ {\mathbb {R}}_+ } \left( \int _{0}^{1/t} f^{*} \right) ^{p'} dt \right] ^{\frac{1}{p'}} = \left[ \int _{ {\mathbb {R}}_+ } \left( \int _{0}^{t} f^{*} \right) ^{p'} \frac{dt}{t^2} \right] ^{\frac{1}{p'}}, \end{aligned} \end{aligned}$$

which is a special case of Hardy’s inequality; see [9, p. 124]. Therefore, Theorem 1.1 leads to the Hausdorff–Young inequality in this case.

The Orlicz spaces, \(L_{\rho _{\Phi }}({\mathbb {R}}^n)\), are defined in terms of a nondecreasing convex (Orlicz) function \(\Phi \) mapping \({\mathbb {R}}_{+}\) onto itself with the norm being given by

$$\begin{aligned} \rho _{\Phi }(f) = \inf \left\{ \lambda >0: \int _{{\mathbb {R}}^{n}} \Phi \left( \frac{|f(x)|}{\lambda } \right) dx \le 1 \right\} . \end{aligned}$$

Our reformulation of the result in [13] asserts that, given an Orlicz function \(\Phi \), one has

$$\begin{aligned} \rho _{\Phi _{1}} ({\hat{f}}) \le C \rho _{\Phi _{2}} (f), \end{aligned}$$

in which

$$\begin{aligned} \rho _{\Phi _{1}}(f) = \rho _{\Phi }(f^{2})^{1/2} \end{aligned}$$

and \(\rho _{\Phi _{2}}\) defined in terms of \( {\widetilde{\Phi }}_{2}\), with

$$\begin{aligned} {\widetilde{\Phi }}_{2}(t) = \frac{1}{ \Phi _{1}(t^{-1})}, \ \ t \in {\mathbb {R}}_+. \end{aligned}$$

We discuss this and related results on Orlicz spaces in detail in Sect. 4. Theorem 1.1 tells us that, for \(1< p \le 2\), the smallest r.i. norm \(\sigma \) for which

$$\begin{aligned} \rho _{p'} ( {\hat{f}} ) \le C \sigma (f), \end{aligned}$$

is given by

$$\begin{aligned} \sigma (f)= {\bar{\sigma }}(f^{*})= \rho _{p'} \left( U f^{*} \right) = \left[ \int _{ {\mathbb {R}}_+ } \left( \int _{0}^{t} f^{*} \right) ^{p'} \frac{dt}{t^2} \right] ^{\frac{1}{p'}} = \left[ \int _{ {\mathbb {R}}_+ } \left( t^{1/p} f^{**}(t) \right) ^{p'} \frac{dt}{t} \right] ^{\frac{1}{p'}}, \end{aligned}$$

the so-called Lorentz norm \(\rho _{p, p'}\), which is smaller than \(\rho _{p}\).

In the next section we provide material on r.i. spaces and interpolation theory. Theorem 1.1 and some of its consequences are proved in Sect. 3. Section 4 deals with the Fourier transform in the context of Orlicz spaces and Sect. 5 considers the boundedness of the Fourier transform between Lorentz Gamma spaces. Section 6 concludes with some remarks on other related work.

Throughout this article, we write \(A \simeq B \) to abbreviate \(C_1 A\le B \le C_2 A\) for some constants \(C_1, C_2 > 0 \) independent of A and B.

2 Rearrangement Invariant Spaces and the K-Functional

Definition 2.1

A rearrangement-invariant (r.i.) Banach function norm \(\rho \) on \(M(\Omega )\), \(\Omega = {\mathbb {R}}^n\) or \(\mathbb {R_+}\), satisfies

  1. (1)

    \(\rho (f) \ge 0\), with \(\rho (f)=0\) if and only if \(f=0\) a.e.;

  2. (2)

    \(\rho (cf)=c \, \rho (f)\), \(c>0\);

  3. (3)

    \(\rho (f+g) \le \rho (f) + \rho (g)\);

  4. (4)

    \(0 \le f_n \nearrow f\) implies \(\rho (f_n) \nearrow \rho (f)\);

  5. (5)

    \(\rho (\chi _{E})<\infty \) for all measurable \(E \subset \Omega \) such that \(|E|<\infty \);

  6. (6)

    \(\int _{E} f \le C_{E} \, \rho (f)\), with \(E \subset \Omega \), \(|E| < \infty \) and \(C_{E}>0\) independent of \(f \in M(\Omega )\);

  7. (7)

    \(\rho (f)= \rho (g)\) whenever \(\mu _{f} = \mu _{g}\). Here, \(\mu _{h}\), for \(h \in M(\Omega )\), denotes the distribution function of h defined as \(\mu _{h}(\lambda ) = |\{ x \in \Omega : |h(x)| > \lambda \}|\), \(\lambda \in {\mathbb {R}}_+\).

Corresponding to an r.i. norm \(\rho \) on \(M(\Omega )\) is the class

$$\begin{aligned} L_{\rho }(\Omega ):= \left\{ f \in M(\Omega ): \rho (f)< \infty \right\} , \end{aligned}$$

which becomes a Banach space of Lebesgue measurable functions under the norm \(\rho (f)\), \(f \in L_{\rho }(\Omega )\). The space \(L_{\rho }(\Omega )\) is then a rearrangement-invariant space.

According to a fundamental result of Luxemberg [9, Chapter 2, Theorem 4.10], there corresponds to every r.i. norm \(\rho \) on \(M({\mathbb {R}}^n)\) an r.i. norm \({\bar{\rho }}\) on \(M(\mathbb {R_+})\) such that

$$\begin{aligned} \rho (f) = {\bar{\rho }}(f^{*}), \ \ \ f \in M({\mathbb {R}}^n). \end{aligned}$$
(2.1)

Here,

$$\begin{aligned} f^{*}(t) = \mu _{f}^{-1}(t) = \inf \left\{ \lambda \in {\mathbb {R}}_+: \mu _{f}(\lambda ) \le t\right\} , \, t \in {\mathbb {R}}_+. \end{aligned}$$

There is only one such \({\bar{\rho }}\) since both \({\mathbb {R}}^{n}\) and \({\mathbb {R}}_+\) are nonatomic and have infinite Lebesgue measure, see [9, p. 64].

A theorem of Hardy and Littlewood asserts that

$$\begin{aligned} \int _{{\mathbb {R}}^n} |f(x)g(x)| \, dx \le \int _{\mathbb {R_+}}f^*(t)g^*(t)dt,\quad \ f,g \in M({\mathbb {R}}^n). \end{aligned}$$
(2.2)

The operation of rearrangement, though not sublinear itself, is sublinear in the average, namely,

$$\begin{aligned} (f+g)^{**}(t) \le f^{**}(t) + g^{**}(t),\quad \ f,g \in M({\mathbb {R}}^n),\quad \ t \in \mathbb {R_+}, \end{aligned}$$
(2.3)

in which

$$\begin{aligned} h^{**}(t)={t}^{-1} \int _0^t h^{*}, \ \ \ 0 \le h \in M({\mathbb {R}}_+),\quad \ t \in \mathbb {R_+}. \end{aligned}$$

A basic technique for working with r.i. norms involves the Hardy–Littlewood–Polya (HLP) Principle which asserts that

$$\begin{aligned} f^{**} \le g^{**}\quad \text {implies}\quad \rho (f) \le \rho (g); \end{aligned}$$

see [9, Chapter 3, Proposition 4.6]. This principle is based on a result of Hardy, a generalized form of which reads

$$\begin{aligned} \int _{0}^{t} f \le \int _{0}^{t} g \end{aligned}$$
(2.4)

implies

$$\begin{aligned} \int _{0}^{t} f h^{*} \le \int _{0}^{t} g h^{*},\quad \ t \in \mathbb {R_+}, \end{aligned}$$

for all \(0 \le f, g \in M( {\mathbb {R}}_+)\) and \(h \in M({\mathbb {R}}_+)\). The Köthe dual of an r.i. norm \(\rho \) on \(M(\Omega )\) is another such norm, \(\rho '\), with

$$\begin{aligned} \rho '(g):= \sup _{\rho (f) \le 1} \int _{\Omega } |f(x)g(x)| dx,\quad \ f, g \in M(\Omega ). \end{aligned}$$

It obeys the Principle of Duality

$$\begin{aligned} \rho '' = (\rho ')' = \rho . \end{aligned}$$
(2.5)

Further, one has the Hölder inequality

$$\begin{aligned} \int _{\Omega } |f(x) g(x)| dx \le \rho (f) \rho '(g),\quad \ f, g \in M(\Omega ). \end{aligned}$$

Finally,

$$\begin{aligned} \bar{\rho '} = ({\bar{\rho }})'. \end{aligned}$$

The Orlicz and Lorentz Gamma spaces studied in sections 5 and 6, respectively, are examples of such r.i. spaces.

The dilation operator \(E_{s}\), \(s \in \mathbb {R_+}\), is defined at \(f \in M(\mathbb {R_+})\), \(t \in \mathbb {R_+}\), by

$$\begin{aligned} (E_s f)(t)= f(st),\quad \ s, t \in \mathbb {R_+}. \end{aligned}$$

The operator \(E_{s}\) is bounded on any r.i. space \(L_{\rho }(\mathbb {R_+})\). We denote its norm by \(h_{\rho }(s)\). Using \(h_{\rho }\) we define the lower and upper indices of \(L_{\rho }(\mathbb {R_+})\) as

$$\begin{aligned} i_{\rho }= \sup \limits _{s > 1} \frac{- \log h_{\rho }(s)}{\log s}\quad \text {and}\quad \ I_{\rho }= \inf \limits _{0<s < 1} \frac{- \log h_{\rho }(s)}{\log s}, \end{aligned}$$
(2.6)

respectively. One has

$$\begin{aligned} i_{\rho } = \lim \limits _{s \rightarrow \infty } \frac{- \log h_{\rho }(s)}{\log s} \ \ \ \text {and} \ \ \ I_{\rho }= \lim \limits _{s \rightarrow 0^+} \frac{- \log h_{\rho }(s)}{\log s}. \end{aligned}$$

Further, \(0 \le i_{\rho } \le I_{\rho } \le 1\) and, moreover,

$$\begin{aligned} i_{\rho '}= 1 - I_{\rho } \ \ \ \text {and} \ \ \ I_{\rho '} = 1 - i_{\rho }. \end{aligned}$$

For all this, see [8, pp. 1250–1252].

If we denote by \(k_{\rho }(s)\) the norm of \(E_{s}\) on the characteristic functions \(\chi _{F}\), \(F \subset \mathbb {R_+}\), \(|F|< \infty \), and define \(j_{\rho }\) and \(J_{\rho }\) by replacing \(h_{\rho }(s)\) in (2.6) by \(k_{\rho }(s)\), we obtain the fundamental indices of \(L_{\rho }(\mathbb {R_+})\). It turns out that when \(L_{\rho }(\mathbb {R_+})\) is an Orlicz space or Lorentz Gamma space \(i_{\rho }= j_{\rho }\) and \(I_{\rho }= J_{\rho }\). For \(\rho \) an Orlicz norm see [9]; for \(\rho \) a Lorentz Gamma norm see [10].

Finally, we describe that part of Interpolation Theory which is relevant to this paper.

Let \(X_{1}\) and \(X_{2}\) be Banach spaces compatible in the sense that both are continuously imbedded in the same Hausdorff topological space H, written

$$\begin{aligned} X_{i} \hookrightarrow H,\quad \ i = 1,2. \end{aligned}$$

The spaces \(X_{1} \cap X_{2}\) and \(X_{1} + X_{2}\) are the sets

$$\begin{aligned} X_{1} \cap X_{2}:= \left\{ x: x \in X_{1} \ \ \text {and} \ \ x \in X_{2} \right\} \end{aligned}$$

and

$$\begin{aligned} X_{1} + X_{2}:= \left\{ x: x= x_{1} + x_{2}, \ \ \text {for some} \ x_{1} \in X_{1}, \ x_{2} \in X_{2} \right\} , \end{aligned}$$

with norms

$$\begin{aligned} \Vert x \Vert _{X_{1} \cap X_{2} } = \max \left[ \Vert x \Vert _{X_{1}}, \Vert x \Vert _{X_{2}} \right] \end{aligned}$$

and

$$\begin{aligned} \Vert x \Vert _{X_{1} + X_{2} } = \inf \left\{ \, \Vert x_{1} \Vert _{X_{1}} + \Vert x_{2} \Vert _{X_{2}}: x= x_{1} + x_{2}, \, x_{1} \in X_{1}, \ x_{2} \in X_{2} \right\} . \end{aligned}$$

Recall that given Banach spaces \(X_{1}\) and \(X_{2}\) imbedded in a common Hausdorff topological vector space, their Peetre K-functional is defined for \(x \in X_{1}+ X_{2}\), \(t>0\), by

$$\begin{aligned} K(t,x;X_{1},X_{2}) = \inf _{x \, = \, x_{1} + x_{2}} \left[ \Vert x_{1} \Vert _{X_{1}} + t \, \Vert x_{2} \Vert _{X_{2}} \right] . \end{aligned}$$

We observe that, for \(\Omega = {\mathbb {R}}^n\) or \(\mathbb {R_+}\), \(p \in [1, \infty )\), \(L_{p}(\Omega )\) and \(L_{\infty }(\Omega )\) are compatible, each being continuously imbedded in the Hausdorff topological space \(M(\Omega )\) equipped with the topology of convergence in measure. One has

$$\begin{aligned} K(t, f; L_{p}(\Omega ), L_{\infty }(\Omega )) \simeq \left[ \int _{0}^{t^{p}} f^{*}(s)^{p} ds \right] ^{1/p}, \ t>0, \end{aligned}$$
(2.7)

\(f \in \left( L_{p} + L_{\infty } \right) (\Omega )\), see [12].

The inequality

$$\begin{aligned} \int _0^t {({\hat{f}})}^{*}(s)^2ds \le C_1 \int _0^t (U{f}^{*})(s)^2ds,\quad t \in {\mathbb {R}}_+, \end{aligned}$$
(2.8)

from [13] reads

$$\begin{aligned} K \left( t, ({\hat{f}})^{*}; L_{2}(\mathbb {R_+}), L_{\infty }(\mathbb {R_+}) \right) \le K \left( t, C_{2} \, U f^{*}; L_{2}(\mathbb {R_+}), L_{\infty }(\mathbb {R_+}) \right) . \end{aligned}$$
(2.9)

Definition 2.2

A Banach space Y is said to be intermediate between \(X_{1}\) and \(X_{2}\) if

$$\begin{aligned} X_{1} \cap X_{2} \hookrightarrow Y \hookrightarrow X_{1} + X_{2}. \end{aligned}$$

Definition 2.3

A Banach space Y intermediate between the compatible spaces \(X_{1}\) and \(X_{2}\) is said to be an interpolation space between \(X_{1}\) and \(X_{2}\) if every linear operator T on \(X_{1} + X_{2}\) satisfying

$$\begin{aligned} T: X_{i} \rightarrow X_{i},\quad i =1,2, \end{aligned}$$

also satisfies \(T: Y \rightarrow Y\).

Suppose now that \(\mu \) is an r.i. norm on \(M(\mathbb {R_+})\) satisfying \(\mu \left( \frac{1}{1+t} \right) < \infty \). Denote by \(X_{\mu }\) the set of all \(x \in X_{1}+ X_{2}\) for which

$$\begin{aligned} \rho _{\mu } (x)= \mu \left( \frac{ K(t,x;X_{1},X_{2}) }{t} \right) < \infty . \end{aligned}$$

Then, \(X_{\mu }\), with the norm \(\rho _{\mu }\), is an interpolation space between \(X_{1}\) and \(X_{2}\), see [4]. Therefore, from (2.7), we have that the space \(X_{\rho _{\mu , p}}\), with the norm

$$\begin{aligned} \rho _{\mu , p} (f) = \overline{\rho _{\mu , p}} (f^{*}) = {\bar{\mu }} \left( t^{-1} \left[ \int _{0}^{t^{p}} f^{*}(s)^{p} ds \right] ^{1/p} \right) , \ \ \ f \in M(\Omega ), \end{aligned}$$
(2.10)

is an interpolation space between \(L_{p}(\Omega )\) and \(L_{\infty }(\Omega )\).

Definition 2.4

A Banach space Y intermediate between the compatible spaces \(X_{1}\) and \(X_{2}\) is said to be monotone if, given \(x, y \in X_{1} + X_{2}\), with

$$\begin{aligned} K(t, x; X_{1}, X_{2}) \le K(t, y; X_{1}, X_{2}), \ \ \ t \in \mathbb {R_+}, \end{aligned}$$
(2.11)

one has \(y \in Y\) implies \(x \in Y\) and \( \Vert x \Vert _{Y} \le \Vert y \Vert _{Y}\).

The result of Lorentz–Shimogaki in [17, Theorem 2 and Lemma 3] asserts that the r.i. interpolation spaces between \(L_{p}(\Omega )\) and \(L_{\infty }(\Omega )\) are precisely the monotone spaces in that context. Further, the inequality (2.9) is a special case of (2.11). Thus, for \(L_{\rho }({\mathbb {R}}^{n})\) between \(L_{2}({\mathbb {R}}^{n})\) and \(L_{\infty }({\mathbb {R}}^{n})\), there holds

$$\begin{aligned} \begin{aligned} \rho ({\hat{f}}) = {\bar{\rho }} \left( ({\hat{f}})^{*} \right)&\le C_{2} {\bar{\rho }} \left( U f^{*} \right) \\&\le M C_{2} {\bar{\sigma }} ( f^{*} )\\&= M C_{2} \sigma ( f ), \end{aligned} \end{aligned}$$

whenever the r.i. norms \(\rho \) and \(\sigma \) on \(M({\mathbb {R}}^{n})\) satisfy

$$\begin{aligned} {\bar{\rho }}(U f^{*}) \le M \, {\bar{\sigma }}(f^{*}),\quad f \in M({\mathbb {R}}^{n}). \end{aligned}$$
(2.12)

Remark 2.1

We have, for simplicity, chosen to restrict attention to functions \(f \in L_{1}({\mathbb {R}}^n)\), since then \({\hat{f}}\) is defined as a classical Lebesgue integral. Again, it is well known that for \(f \in L_{2}({\mathbb {R}}^n)\)

$$\begin{aligned} \lim _{R \rightarrow \infty } \int _{ \left\{ x \in {\mathbb {R}}^n \,: \, |x| \le R \right\} } f(x) \, e^{ - 2 \pi i x \cdot \xi } \, dx, \ \ \ \xi \in {\mathbb {R}}^n, \end{aligned}$$

exists in the norm of \(L_{2}({\mathbb {R}}^n)\), which can be used to define \({\hat{f}}\). Thus, the Fourier transform can be defined as a function for all \(f \in \left( L_{1} + L_{2}\right) ({\mathbb {R}}^n)\). Indeed, it is shown in [3] that \( \left( L_{1} + L_{2}\right) ({\mathbb {R}}^n)\) is the largest r.i. space of functions that is mapped by \({\mathscr {F}}\) into a space of locally integrable functions.

The Editor has referred us to the paper [28], among others, where it is shown that, essentially the set of functions f for which \({\hat{f}}\) is defined as a function is the amalgam space \(\ell _{2} \left( L_{1} ({\mathbb {R}}^n) \right) \), which in the case \(n=1\) has the norm

$$\begin{aligned} \left( \sum _{k= - \infty }^{\infty } \left( \int _{k}^{k+1} |f(x)| \, dx \, \right) ^{2} \right) ^{1/2}. \end{aligned}$$

This is a Banach function norm on \(M({\mathbb {R}}^n)\) that is not rearrangement-invariant, namely, it satisfies (1)–(6) in Definition 2.1, but not (7). Thus, we need spaces other than the r.i. ones to study the Fourier transform in the context of this space.

3 Proof of Theorem 1.1

Proof

The “if” part was proved towards the end of the Sect. 2. For the “only if” part, let B be the unit ball in \({\mathbb {R}}^n\) centered at the origin. Then \( \widehat{ \chi _{B} }\) is real-valued, radial and continuous, with \( \widehat{\chi _{B}}(0)=|B|.\) Also, \(0 \le \chi _B *\chi _{B} \le |B| \chi _{2B} \in L^1\) and \(\widehat{(\chi _B *\chi _{B})}=({{\widehat{\chi }}}_{B})^2.\)

Choose \(r > 0\) such that \(\widehat{\chi _B} \ge |B|/2 \) on rB. Let \(0 \le f \in L^1({\mathbb {R}}^n)\) be radial and radially decreasing. For \(t \in \mathbb {R_+}\), choose \(s>0\) such that \(|sB|=t^{-1}.\) Then

$$\begin{aligned} \begin{aligned} (U f^{*})(t)&= \int _{0}^{1/t} f^{*} = \int _{sB}f(y)dy \\&= \left( r^{-1}s\right) ^{n} \int _{rB} f\left( r^{-1}sy\right) dy\\&\le \left( r^{-1}s\right) ^{n} \int _{rB} \frac{4}{|B|^2} \left( \widehat{\chi _{B}}(y)\right) ^2 f\left( r^{-1}sy\right) dy \\&\le \left( r^{-1}s\right) ^{n} \frac{4}{|B|^2} \int _{{\mathbb {R}}^n} \widehat{\left( \chi _B *\chi _{B}\right) } (y) f\left( r^{-1}sy\right) dy \\&= \left( r^{-1}s\right) ^{n} \frac{4}{|B|^2} (rs^{-1})^n\int _{{\mathbb {R}}^n} {\left( \chi _B *\chi _{B}\right) } (\xi ){\hat{f}}\left( rs^{-1}\xi \right) d\xi \\&\le \frac{4}{|B|^2} |B| \int _{2B}|{\hat{f}}\left( rs^{-1}\xi \right) | d\xi \\&=\frac{4}{|B|} \left( r^{-1}s\right) ^{n} \int _{|\eta |\le 2 r s^{-1}}|{\hat{f}}(\eta )| d\eta \\&= \frac{4}{|B|} \left( r^{-1}s\right) ^{n} \int _0^{ {\left( 2r s^{-1}\right) }^n |B|} \left( {\hat{f}}\right) ^* \\&= \frac{4}{|B|^2} r^{-n} \frac{1}{t} \int _0^{2^n{r}^n |B|^2 t} \left( {\hat{f}}\right) ^* \\&\le \frac{4}{|B|^2} r^{-n} \frac{1}{t} \int _0^{t} \left( {\hat{f}}\right) ^* = C_n \, \frac{1}{t} \int _0^{t} \left( {\hat{f}}\right) ^*, \end{aligned} \end{aligned}$$

where we further shrink r to be such that \( 2^nr^n|B|^2<1\) and the constant \(C_n>0\) depends only on n.

Therefore, for \(f \in (L_{1} \cap L_{\sigma } )({\mathbb {R}}^n) \) such that \(f(x)= g(|x|)\), \(x \in {\mathbb {R}}^n\), with \(g \downarrow \) on \(\mathbb {R_+},\)

$$\begin{aligned} \begin{aligned} {\bar{\rho }} (U f^{*})&\le C_{n} \, {\bar{\rho }} \left( \frac{1}{t} \int _{0}^{t} \left( {\hat{f}}\right) ^{*} \right) \\&\le C_{n}' \, {\bar{\rho }} \left( \left( {\hat{f}}\right) ^{*}\right) \\&= C_{n}' \, \rho \left( {\hat{f}}\right) \\&\le C \, C_{n}' \, \sigma (f), \ \ \text {by assumption,}\\&= C \, C_{n}' \, {\bar{\sigma }}\left( f^{*}\right) , \end{aligned} \end{aligned}$$

where the second inequality is the boundedness of the averaging operator, \(P: g \mapsto \frac{1}{t} \int _{0}^{t} g\), on \(L_{{\bar{\rho }}}({\mathbb {R}}_+)\), which follows from our hypothesis on \(L_{{\bar{\rho }}}({\mathbb {R}}_+)\) that it is the interpolation space between \(L_{2}({\mathbb {R}}_+)\) and \(L_{\infty }({\mathbb {R}}_+)\), and the Hardy’s inequality.

Given \(h \in \left( L_{1} \cap L_{ {\bar{\sigma }} } \right) (\mathbb {R_+})\), let \(g(t) = h^{*} ( |B| t^{n} )\) and set \(f(x)= g(|x|)\). Then, the rearrangement of f with respect to n-dimensional Lebesgue measure is equal to the rearrangement of h with respect to 1-dimensional Lebesgue measure. The foregoing argument then yields

$$\begin{aligned} {\bar{\rho }} (U h) \le {\bar{\rho }} \left( U h^{*} \right) = {\bar{\rho }} \left( U f^{*} \right) \le C {\bar{\sigma }} (f^{*}) = C {\bar{\sigma }} (h^{*})= C {\bar{\sigma }} (h). \ \ \end{aligned}$$

The space \( \left( L_{1} \cap L_{ {\bar{\sigma }} } \right) (\mathbb {R_+})\) includes all bounded functions of compact support whence the monotone convergence theorem and the Fatou property of \({\bar{\rho }}\) and \({\bar{\sigma }}\) completes the proof. \(\square \)

Boyd in [7, pp. 92–98] associates to each r.i. norm \(\rho \) on \(M(\Omega )\), \(\Omega = {\mathbb {R}}^{n}\) or \({\mathbb {R}}_+\), and each \(p>1\) the functional

$$\begin{aligned} \rho ^{(p)} (f) = \rho (|f|^{p})^{\frac{1}{p}}, \ \ \ f \in M(\Omega ). \end{aligned}$$
(3.1)

He shows that \(\rho ^{(p)}\) is an r.i. norm on \(M(\Omega )\) and that \({\bar{\rho }}(f^{**}) \le C {\bar{\rho }}(f^{*})\) holds with \({\bar{\rho }} = \overline{\rho ^{(p)}} = {\bar{\rho }}^{(p)}\).

The space defined by the norm \(\rho ^{(p)}\) is now referred to as the p-convexification of \(L_{\rho }({\mathbb {R}}^n)\). It was studied in a series of papers by G. Lozanovskiĭ about the time Boyd, independently, introduced his spaces. See the references to G. Lozanovskiĭ’s work in [19]. This latter paper treats the K-functional of p-convexifications, as does the paper [1]. These papers should shed light on the work involving \(\rho ^{(2)}\) in this and the next two sections.

Theorem 3.1

Let \(\rho \) be an r.i. norm on \(M(\Omega )\). For fixed \(p>1\), define \(\rho ^{(p)}\) as in (3.1). Then, \(L_{\rho ^{(p)}}(\Omega )\) is an interpolation space between \(L_{p}(\Omega )\) and \(L_{\infty }(\Omega )\).

Proof

Suppose the linear operator T satisfies

$$\begin{aligned} T: L_{p}(\Omega ) \rightarrow L_{p}(\Omega ) \ \ \ \text {and} \ \ \ T: L_{\infty }(\Omega ) \rightarrow L_{\infty }(\Omega ). \end{aligned}$$

Then, according to [9, Theorem 1.11, pp. 301–304], there exists \(C>0\), such that

$$\begin{aligned} \int _{0}^{t} (Tf)^{*}(s)^{p} ds \le C \int _{0}^{t} f^{*}(s)^{p} ds, \ \ \ f \in (L_{p} + L_{\infty })(\Omega ), \ t \in \mathbb {R_+}. \end{aligned}$$
(3.2)

The HLP Principle involving \({\bar{\rho }}\) yields

$$\begin{aligned} \begin{aligned} \rho \left( (Tf)^{p} \right)&= {\bar{\rho }} \left( \left[ (Tf)^{*} \right] ^{p} \right) \\&\le C {\bar{\rho }} \left( \left[ f^{*} \right] ^{p} \right) \\&= C \rho \left( |f|^{p} \right) , \ \ \ f \in L_{\rho ^{(p)}}(\Omega ), \end{aligned} \end{aligned}$$

and hence

$$\begin{aligned} \rho ^{(p)} (Tf) \le C^{1/p} \rho ^{(p)} (f), \ \ \ f \in L_{\rho ^{(p)}}(\Omega ). \end{aligned}$$

\(\square \)

Theorem 3.2

Let \(\rho \) and \(\sigma \) be r.i. norms on \(M({\mathbb {R}}^n)\) determined, respectively, by the r.i. norms \({\bar{\rho }}\) and \({\bar{\sigma }}\) on \(M(\mathbb {R_+})\) by \(\rho (f) = {\bar{\rho }}(f^{*})\) and \(\sigma (f) = {\bar{\sigma }}(f^{*})\), \(f \in M({\mathbb {R}}^n)\). Then,

$$\begin{aligned} \rho ^{(2)} ( {\hat{f}}) \le C \sigma (f), \ \ \ f \in (L_{\sigma } \, \cap \, L_{1})({\mathbb {R}}^n), \end{aligned}$$

if and only if

$$\begin{aligned} \overline{\rho ^{(2)}} (Ug) \le C {\bar{\sigma }} (g), \ \ \ g \in M(\mathbb {R_+}). \end{aligned}$$

Proof

The result is a consequence of Theorems 3.1 and 1.1. \(\square \)

From our discussion on the spaces \(X_{\rho _{\mu , p}}\), with the norm \(\rho _{\mu , p}\) given by (2.10), Theorem 1.1 guarantees

Theorem 3.3

Let \(\mu \) and \(\sigma \) be r.i. norms on \(M({\mathbb {R}}^n)\) determined, respectively, by the r.i. norms \({\bar{\mu }}\) and \({\bar{\sigma }}\) on \(M(\mathbb {R_+})\). Suppose \({\bar{\mu }} \left( \frac{1}{1+t} \right) < \infty \). Set

$$\begin{aligned} \rho _{\mu , 2} (f) = \overline{\rho _{\mu , 2}} (f^{*}) = {\bar{\mu }} \left( t^{-1} \left[ \int _{0}^{t^{2}} f^{*}(s)^{2} ds \right] ^{1/2} \right) . \end{aligned}$$

Then,

$$\begin{aligned} \rho _{\mu , 2} ( {\hat{f}}) \le C \sigma (f), \ \ \ f \in (L_{\sigma } \, \cap \, L_{1})({\mathbb {R}}^n), \end{aligned}$$

if and only if

$$\begin{aligned} \overline{ \rho _{\mu , 2} } (Ug) \le C {\bar{\sigma }} (g), \ \ \ g \in M(\mathbb {R_+}). \end{aligned}$$

Finally, consider an r.i. norm \(\rho \) on \(M({\mathbb {R}}^n)\) determined by the r.i. norm \(\bar{\rho }\) on \(M(\mathbb {R_+})\) and set

$$\begin{aligned} \rho _{U} (f):= \left( \bar{\rho } \, \circ U \right) (f^{*}) = {\bar{\rho }} (U f^{*}), \ \ \ f \in M({\mathbb {R}}^n). \end{aligned}$$

One has \(\rho _{U}\) an r.i. norm if \(\left( \bar{\rho } \, \circ U \right) ( \chi _{(0,t)} ) < \infty \) for all \(t>0\), or, equivalently, \({\bar{\rho }} \left( \frac{1}{1+t} \right) < \infty \). In that case, \(L_{{\bar{\rho }} \, \circ U } (\mathbb {R_+})\) is the largest r.i. space to be mapped into \(L_{{\bar{\rho }}}(\mathbb {R_+})\) by U.

With this background we now have

Theorem 3.4

Let \(\rho \) be an r.i. norm on \(M({\mathbb {R}}^n)\) defined in terms of an r.i. norm \({\bar{\rho }}\) on \(M(\mathbb {R_+})\) such that

$$\begin{aligned} {\bar{\rho }} \left( \frac{1}{1+t} \right) < \infty . \end{aligned}$$

Assuming \(L_{ {\bar{\rho }}}(\mathbb {R_+})\) is an interpolation space between \(L_{2}(\mathbb {R_+})\) and \(L_{\infty }(\mathbb {R_+})\), one has that \(L_{\rho _{U}} ({\mathbb {R}}^n)\) is the largest r.i. space of functions on \({\mathbb {R}}^n\) to be mapped into \(L_{\rho }({\mathbb {R}}^n)\) by \({\mathscr {F}}\).

4 \({\mathscr {F}}\) in the Context of Orlicz Spaces

An Orlicz gauge norm is given in terms of an N-function

$$\begin{aligned} \Phi (x)= \int _{0}^{x} \phi ,\quad x \in \mathbb {R_+}; \end{aligned}$$

here \(\phi \) is a nondecreasing function mapping \(\mathbb {R_+}\) onto itself. These N-functions are convex functions of the type from [13] referred to in the Introduction. Specifically, the gauge norm \(\rho _{\Phi }\) is defined at \(f \in M(\Omega )\), \(\Omega = {\mathbb {R}}^n \) or \({\mathbb {R}}_+\), by

$$\begin{aligned} \rho _{\Phi }(f) = \inf \left\{ \lambda >0: \int _{\Omega } \Phi \left( \frac{|f(x)|}{\lambda } \right) dx \le 1 \right\} . \end{aligned}$$

One can show \(\rho _{\Phi }(f)= {\bar{\rho }}_{\Phi }(f^{*})\), so that the Orlicz space

$$\begin{aligned} L_{\Phi }(\Omega )= \left\{ f \in M(\Omega ): \rho _{\Phi }(f) < \infty \right\} \end{aligned}$$

is an r.i. space. The norm \(\left( \rho _{\Phi } \right) '\) dual to \(\rho _{\Phi }\) is equivalent to the gauge norm \(\rho _{{\tilde{\Phi }}}\), where \({\tilde{\Phi }}(t) = \int _{0}^{t} \phi ^{-1}\), \(t \in {\mathbb {R}}_+\), see [9].

The definitive work on \({\mathscr {F}}\) between Orlicz spaces is due to Jodeit and Torchinsky. See, in particular, [13, Theorem 2.16]. This theorem asserts that if A and B are N-functions with \(L_{A}({\mathbb {R}}^n) \subset (L_{1} + L_{2})({\mathbb {R}}^n)\), \(L_{B}({\mathbb {R}}^n) \subset (L_{2} + L_{\infty })({\mathbb {R}}^n)\) and \({\mathscr {F}}: L_{A}({\mathbb {R}}^n) \rightarrow L_{B}({\mathbb {R}}^n)\), then there exist N-functions \(A_{1}\) and \(B_{1}\) with \(L_{A_{1}}({\mathbb {R}}^n) \supset L_{A}({\mathbb {R}}^n)\) and \(L_{B_{1}}({\mathbb {R}}^n) \subset L_{B}({\mathbb {R}}^n)\) for which \({\mathscr {F}}: L_{A_{1}}({\mathbb {R}}^n) \rightarrow L_{B_{1}}({\mathbb {R}}^n)\). Moreover, \(B_{1}(t)= 1 / \tilde{A_{1}}(t^{-1})\); \(A_{1}(t) / t^{2} \downarrow \) on \(\mathbb {R_+}\) and so \(B_{1}(t) / t^{2}\) \(\uparrow \) on \(\mathbb {R_+}\).

Using the results in the previous sections we now show \(L_{A_{1}}(\mathbb {R_+})\) is an interpolation space between \(L_{1}(\mathbb {R_+})\) and \(L_{2}(\mathbb {R_+})\), while \(L_{B_{1}}(\mathbb {R_+})\) is an interpolation space between \(L_{2}(\mathbb {R_+})\) and \(L_{\infty }(\mathbb {R_+})\).

To begin, we observe that \(B_{1}(t) / t^{2}\) \(\uparrow \) is equivalent to \(B_{1}(t) = \Phi (t^{2})\) for some N- function \(\Phi \). Indeed, given the latter, one has \(B_{1}(t) / t^{2} = \Phi (t^{2}) / t^{2} \uparrow \) on \(\mathbb {R_+}\). Again \(B_{1}(t) / t^{2} \uparrow \) implies \(B_{1}(t^{1/2} ) / t \uparrow \), so that \(\Phi (t)=B_{1}(t^{1/2})\) is such that \(B_{1}(t)= \Phi (t^{2})\). Next,

$$\begin{aligned} \rho _{\Phi }^{(2)} (f) = \rho _{\Phi } (f^{2})^{\frac{1}{2}}= \rho _{B_{1}} (f). \end{aligned}$$

According to Theorem 3.1, then, \(L_{B_{1}}({\mathbb {R}}^n)\) is an interpolation space between \(L_{2}({\mathbb {R}}^n)\) and \(L_{\infty }({\mathbb {R}}^n)\).

Now, \(B_{1}(t)= 1 / \tilde{A_{1}}(t^{-1})\) is equivalent to \( \tilde{A_{1}}(t)= 1 / B_{1}(t^{-1})\), whence \(\tilde{A_{1}}(t) / t^{2}\) \(= (t^{-1})^{2} / B_{1}(t^{-1})\) and so \(B_{1}(t) / t^{2} \uparrow \) amounts to \(\tilde{A_{1}}(t) / t^{2} \uparrow \), that is, \(L_{\tilde{A_{1}}}({\mathbb {R}}^n)\) is an interpolation space between \(L_{2}({\mathbb {R}}^n)\) and \(L_{\infty }({\mathbb {R}}^n)\). Since \(L_{\tilde{A_{1}}}({\mathbb {R}}^n)\) is the Köthe dual of \(L_{A_{1}}({\mathbb {R}}^n)\) we conclude that \(L_{A_{1}}({\mathbb {R}}^n)\) is an interpolation space between \(L_{1}({\mathbb {R}}^n)\) and \(L_{2}({\mathbb {R}}^n)\).

The monotonicity conditions on \(A_{1}\) and \(B_{1}\) translate into conditions on their associated fundamental functions. For example, \(L_{B_{1}}({\mathbb {R}}^n)\) has fundamental function \(\phi _{B_{1}}(t)= \rho _{B_{1}} \left( \chi _{(0,t)} \right) = 1 / B_{1}^{-1}(t^{-1})\), \(t \in \mathbb {R_+}\). Thus setting \(t = 1 / B_{1}(y)\) in \(\frac{\phi _{B_{1}}(t)}{t^{1/2}} = \frac{1}{B_{1}^{-1}(t^{-1}) t^{1/2}}\) we arrive at \(\left( \frac{B_{1}(y)}{ y^{2}} \right) ^{1/2}\), which increases in y and therefore decreases in t, so \(\frac{\phi _{B_{1}} (t) }{ t^{1/2}} \downarrow \).

We observe that \(L_{A_{1}} ({\mathbb {R}}^n)\) is not the largest r.i. space that \({\mathscr {F}}\) maps into \(L_{B_{1}} ({\mathbb {R}}^n)\); that space has norm \(\rho _{B_{1}} (U f^{*})\). In the Lebesgue context, in which, say, \(B_{1}(t)= t^{p'}\), \(1<p<2\),

$$\begin{aligned} \rho _{p'} (U f^{*}) = \left[ \int _{\mathbb {R_+}} (U f^{*})(t)^{p'} dt \right] ^{1/p'} \simeq \left[ \int _{\mathbb {R_+}} \left[ t^{\frac{1}{p}} f^{**}(t) \right] ^{p'} \frac{dt}{t} \right] ^{1/p'}, \end{aligned}$$

which is the so-called Lorentz norm \(\rho _{p,p'}\). This norm is smaller than \(\rho _{A_{1}} = \rho _{p}\). For more details see the next section.

The foregoing argument can be used to associate a pair of N-functions (AB) to a given N-function \(\Phi \) such that \({\mathscr {F}}: L_{A} ({\mathbb {R}}^{n}) \rightarrow L_{B} ({\mathbb {R}}^{n})\). Moreover, \(L_{A} ({\mathbb {R}}^{n})\) is an interpolation space between \(L_{1}({\mathbb {R}}^{n})\) and \(L_{2}({\mathbb {R}}^{n})\), while \(L_{B} ({\mathbb {R}}^{n})\) is an interpolation space between \(L_{2}({\mathbb {R}}^{n})\) and \(L_{\infty }({\mathbb {R}}^{n})\). Indeed, we have

Theorem 4.1

Let \(\Phi \) be an N-function. Set \(B(t)= \Phi (t^2)\) and \({\tilde{A}}(t) = 1 / B(t^{-1})\). Then, essentially, \(\frac{A(t)}{t^2} \downarrow \),

$$\begin{aligned} {\mathscr {F}}: L_{A} (\mathbb {{\mathbb {R}}}^n) \rightarrow L_{B} (\mathbb {{\mathbb {R}}}^n) \end{aligned}$$

or, equivalently,

$$\begin{aligned} {\mathscr {F}}: L_{ {\tilde{B}}} (\mathbb {{\mathbb {R}}}^n) \rightarrow L_{{\tilde{A}}} (\mathbb {{\mathbb {R}}}^n) \end{aligned}$$

with \(L_{A} ({\mathbb {R}}^n)\) an interpolation space between \(L_{1}({\mathbb {R}}^n)\) and \(L_{2}({\mathbb {R}}^n)\), while \(L_{B}({\mathbb {R}}^n)\) is an interpolation space between \(L_{2}({\mathbb {R}}^n)\) and \(L_{\infty }({\mathbb {R}}^n)\).

Proof

The preceding discussion shows \(\rho _{B} = \rho _{\Phi ^{(2)}}\) whence \(L_{B} ({\mathbb {R}}^n)\) is between \(L_{2}({\mathbb {R}}^n)\) and \(L_{\infty }({\mathbb {R}}^n)\). Again \(B(t) / t^2 \uparrow \) and \({\tilde{A}} (t) / t^2 \uparrow \) which means \(L_{{\tilde{A}}}\) is an interpolation space between \(L_{2}({\mathbb {R}}^n)\) and \(L_{\infty }({\mathbb {R}}^n)\), whence \(L_{A}({\mathbb {R}}^n)\) is an interpolation space between \(L_{1}({\mathbb {R}}^n)\) and \(L_{2}({\mathbb {R}}^n)\). Finally, Theorem 3.10 in [13] ensures

$$\begin{aligned} {\mathscr {F}}: L_{A}({\mathbb {R}}^n) \rightarrow L_{B}({\mathbb {R}}^n), \end{aligned}$$

since \(B(t) = 1 / {\tilde{A}}(t^{-1})\) is equivalent to \({\tilde{A}}(t) = 1 / B(t^{-1})\). \(\square \)

5 \({\mathscr {F}}\) Between Lorentz Gamma Spaces

In this section, we make use of the operators P and Q defined by

$$\begin{aligned} (P f) (t) = \frac{1}{t} \int _{0}^{t} f\quad \text {and}\quad (Q g )(t) = \int _{t}^{\infty } g(s) \frac{ds}{s}, \ \ \ f,g \in M(\mathbb {R_+}), \ t \in \mathbb {R_+}. \end{aligned}$$

These operators satisfy the equations

$$\begin{aligned} \int _{\mathbb {R_+}} g \, Pf = \int _{\mathbb {R_+}} f \, Qg, \ \ \ f,g \in M(\mathbb {R_+}), \end{aligned}$$

and

$$\begin{aligned} PQ= QP = P + Q. \end{aligned}$$

Fix an index \(p \in (1, \infty )\) and a weight \(0 \le u \in M({\mathbb {R}}_+)\). The Lorentz Gamma norm \(\rho _{p, u}\) defined in terms of the Lorentz norm \(\lambda _{p,u}(f) = \lambda _{p,u}(f^{*}) = \left( \int _{{\mathbb {R}}_+} f^{*}(t)^p dt \right) ^{1/p}\) by

$$\begin{aligned} \rho _{p,u} (f) = \lambda _{p,u} (f^{**}), \ \ \ f \in M(\Omega ), \end{aligned}$$

where, once again, \(\Omega = {\mathbb {R}}^{n}\) or \({\mathbb {R}}_{+}\).

To guarantee \(\rho _{p, u}(\chi _{E}) < \infty \) for all measurable sets \(E \subset \Omega \) with \(|E|< \infty \), we require

$$\begin{aligned} \int _{{\mathbb {R}}_+} \frac{u(t)}{ 1+ t^p } dt < \infty . \end{aligned}$$
(5.1)

The Lorentz Gamma space

$$\begin{aligned} \Gamma _{p,u}(\Omega )=\left\{ f \in M(\Omega ): \rho _{p,u}(f) <\infty \right\} , \end{aligned}$$

is then an r.i. space. The norm, \(\rho _{p',v'}\), dual to \(\rho _{p,u}\) is given by

$$\begin{aligned} \rho _{p', v'}(g) = \left( \int _{{\mathbb {R}}_+} g^{**}(t)^p v'(t) dt \right) ^{1/p'}, \ \ \ g \in M(\Omega ) \end{aligned}$$

where, \(p' = p / (p-1)\) and

$$\begin{aligned} v'(t)= \frac{t^{p'+p-1}\int _0^t u \ \int _t^\infty u(s)s^{-p} ds}{{\left[ \int _0^t v + t^p \int _t^\infty u(s)s^{-p} ds\right] }^{p'+1}}. \end{aligned}$$

This is shown, for example, in [10].

In this section we study the inequality

$$\begin{aligned} \rho _{p,u}({\hat{f}}) \le C \, \rho _{q,v}(f), \ \ \ f \in (L_1\cap \Gamma _{q,v})({\mathbb {R}}^n). \end{aligned}$$
(5.2)

We begin by assuming \(\Gamma _{p,u}(\mathbb {R_+})\) is an interpolation space between \(L_{2}(\mathbb {R_+})\) and \(L_{\infty }(\mathbb {R_+}),\) then address the question of when this is the case later in the section.

Recall that Theorem 1.1 ensures that (5.2) holds if and only if

$$\begin{aligned} {{\bar{\rho }}}_ {p,u}(Uf^*) \le C {{\bar{\rho }}}_ {q,v}(f^*), \ \ \ f \in M(\mathbb {R_+}). \end{aligned}$$
(5.3)

Theorem 5.1

Let the indices pq and weights uv be as described above. Then, given that \(\Gamma _{p,u}(\mathbb {R_+})\) is an interpolation space between \(L_{2}(\mathbb {R_+})\) and \(L_{\infty }(\mathbb {R_+}),\) one has (5.2) if and only if

$$\begin{aligned} {{\bar{\rho }}}_{q',v'}(g^{**}) \le C {{\bar{\rho }}}_{p',{u_p}'}(g^{*}), \ \ \ g \in M(\mathbb {R_+}). \end{aligned}$$
(5.4)

where, as usual, \(p'=\frac{p}{p-1}\), \( q'=\frac{q}{q-1}\), \(\int _{\mathbb {R_+}} v=\infty \),

$$\begin{aligned}{} & {} v'(t)= \frac{t^{q'+q-1}\int _0^t v\int _t^\infty v(s)s^{-q} ds}{{\left[ \int _0^t v+t^q \int _t^\infty v(s)s^{-q} ds\right] }^{q'+1}}\\{} & {} u_p(t)= u(t^{-1})t^{p-2}, \ \ \ \int _{\mathbb {R_+}} u_{p} = \infty , \end{aligned}$$

and

$$\begin{aligned} {u_p}'(t)= \frac{ t^{p'+p-1} \int _{0}^{t} u_{p} \int _{t}^{\infty } u_{p}(s) s^{-p} ds }{ \left[ \int _{0}^{t} u_{p} + t^{p} \int _{t}^{\infty } u_{p}(s)s^{-p} ds \right] ^{p'+1 } }, \ \ \ t \in {{\mathbb {R}}}_{+}. \end{aligned}$$

Proof

The inequality (5.3) tells that the space determined by the r.i. norm \({{\bar{\rho }}}_{p,u} (Uf^*)\) is the largest one mapped into \(\Gamma _{p,u}({\mathbb {R}}_+)\) by U. Now,

$$\begin{aligned}{} & {} {\bar{\rho }}_{p,u}(Uf^*)= \left[ \int _{ {\mathbb {R}}_+ } P(Uf^{*})(t)^{p} u(t) \, dt \right] ^{\frac{1}{p}},\\{} & {} \begin{aligned} P \left( Uf^* \right) (t)&= t^{-1} \left( P (Q f^{*}) \right) ( t^{-1}). \end{aligned} \end{aligned}$$

Further,

$$\begin{aligned} \begin{aligned} \int _{\mathbb {R_+}} \left[ t^{-1} \left( P (Q f^{*}) \right) ( t^{-1}) \right] ^{p} u(t) dt&= \int _{\mathbb {R_+}} \left[ t \left( P (Q f^{*}) \right) ( t) \right] ^{p} u(t^{-1}) t^{-2} dt. \end{aligned} \end{aligned}$$

We have shown that

$$\begin{aligned} \begin{aligned} {{\bar{\rho }}}_{p,u}(Uf^*)&=\left[ \int _{\mathbb {R_+}} \left( P (Q f^{*}) \right) (t)^p u_p(t) dt\right] ^{\frac{1}{p}}\\&={{\bar{\rho }}}_{p,u_p}(Qf^*). \end{aligned} \end{aligned}$$

Therefore, any \( \Gamma _{q,v}(\mathbb {R_+})\) mapped into \( \Gamma _{p,u}(\mathbb {R_+})\) by U must be embedded into this largest domain; that is,

$$\begin{aligned} {{\bar{\rho }}}_ {p,u_p}(Qf^*) \le C {{\bar{\rho }}}_ {q,v}(f^*), \ \ \ f \in M({\mathbb {R}}_+). \end{aligned}$$
(5.5)

But, since (5.5) is equivalent to (5.4), its dual inequality, (5.5) may be tested over any \(0 \le f \in M({\mathbb {R}}_+),\) as is seen in

$$\begin{aligned} \int _{\mathbb {R_+}} g^{*} Qf = \int _{\mathbb {R_+}} f P g^{*} \le \int _{\mathbb {R_+}} f^{*} P g^{*}= \int _{\mathbb {R_+}} g^{*} Q f^{*}. \end{aligned}$$

\(\square \)

The inequality (5.4), and hence (5.2), amounts to

$$\begin{aligned} \left[ \int _{\mathbb {R_+}} (Ph)^{q'}v' \right] ^{\frac{1}{q'}} \le C \left[ \int _{\mathbb {R_+}} h^{p'}{u_p}' \right] ^{\frac{1}{p'}}, \end{aligned}$$
(5.6)

with \(h=g^{**}\) belonging to

$$\begin{aligned} \Omega _{0,1}(\mathbb {R_+})= \{ 0 \le h \in M(\mathbb {R_+}): h(t) \downarrow \ \text {and} \ t \, h(t) \uparrow \text {on} \ \mathbb {R_+}\}. \end{aligned}$$

Such inequalities are shown in Theorem 4.4 of [10] to be equivalent to a pair of weighted norm inequalities involving general non-negative measurable functions. In the case of (5.6) this leads to

Theorem 5.2

Let the indices pq and weights uv be as in Theorem 5.1. Then, (5.6) holds if and only if

$$\begin{aligned} \left[ \int _{\mathbb {R_+}} \left[ (P+Q)g \right] ^{q'} v'\right] ^{\frac{1}{q'}} \le C \left[ \int _{\mathbb {R_+}} g^{p'}{u_p}^{1-p'}\right] ^{\frac{1}{p'}} \end{aligned}$$
(5.7)

and

$$\begin{aligned} \left[ \int _{\mathbb {R_+}} (P^2g)^{q'} v'\right] ^{\frac{1}{q'}} \le C \left[ \int _{\mathbb {R_+}} g^{p'}{u_p}^{1-p'}\right] ^{\frac{1}{p'}}, \ \ \ 0 \le g \in M( \mathbb {R_+}). \end{aligned}$$

Proof

According to Theorem 4.4 in [10], one has (5.6) if and only if

$$\begin{aligned} \left[ \int _{\mathbb {R_+}} \left[ (P+Q)Qg \right] ^{p} u_p\right] ^{\frac{1}{p}} \le C \left[ \int _{\mathbb {R_+}} g^{q}{v'}^{1-q}\right] ^{\frac{1}{q}} \end{aligned}$$

and

$$\begin{aligned} \left[ \int _{\mathbb {R_+}} \left[ P(P+Q)g \right] ^{q'} v'\right] ^{\frac{1}{q'}} \le C \left[ \int _{\mathbb {R_+}} g^{p'}{u_p}^{1-p'}\right] ^{\frac{1}{p'}}, \end{aligned}$$

holds for all \( 0 \le g \in M(\mathbb {R_+}).\)

These are dual inequalities. We choose the second one, which easily reduces to (5.7). \(\square \)

To deal with the case \(q \le p\) we will use special instances of the following combination of Theorems 1.7 and 4.1 from [6].

Theorem 5.3

Consider \(0 \le K(x,y) \in M \left( \mathbb {R_+}\times \mathbb {R_+} \right) \), which, for fixed \(y \in \mathbb {R_+}\), increases in x and, for fixed \(x \in \mathbb {R_+},\) decreases in y and which, moreover, satisfies the growth condition

$$\begin{aligned} K(x,y) \le K(x,z)+K(z,y),\quad 0<y<z<x. \end{aligned}$$

Let tuv and w be nonnegative, measurable (weight) functions on \(\mathbb {R_+}\) and suppose \(\Phi _1(x)= \int _0^x \phi _1\) and \(\Phi _2(x)= \int _0^x \phi _2\) are N-functions having complementary functions \(\Psi _1(x)= \int _0^x {\phi _1}^{-1} \) and \(\Psi _2(x)= \int _0^x {\phi _2}^{-1}\), respectively, with \(\Phi _1\circ {\Phi _2}^{-1}\) convex. Then there exists \(c>0\) such that

$$\begin{aligned} {\Phi _1}^{-1} \left( \int _{\mathbb {R_+}} \Phi _1 \left( c w(x) \int _0^x K(x,y) f(y) dy \right) t(x) dx \right) \\ \le {\Phi _2}^{-1} \left( \int _{\mathbb {R_+}} \Phi _2 \left( \, u(y) f(y) \, \right) v(y) dy \right) , \end{aligned}$$

\(0 \le f \in M({\mathbb {R_+}})\), if and only if

$$\begin{aligned} \int _0^x \frac{K(x,y)}{u(y)} {\phi _2}^{-1} \left( \frac{c \alpha (\lambda ,x) K(x,y)}{\lambda u(y)v(y)} \right) dy \le c^{-1}\lambda \end{aligned}$$

and

$$\begin{aligned} \int _0^x \frac{1}{u(y)}\phi ^{-1}_2 \left( \frac{ c \beta (\lambda ,x)}{\lambda u(y)v(y)} \right) {dy}\le c^{-1} \lambda , \end{aligned}$$

where

$$\begin{aligned} \begin{aligned} \alpha (\lambda ,x)&= \Phi _2\circ {\Phi _1}^{-1}\left( \int _x^{\infty } \Phi _1 \left( \lambda w(y) \right) t(y)dy \right) \end{aligned} \end{aligned}$$

and

$$\begin{aligned} \begin{aligned} \beta (\lambda ,x)&= \Phi _2\circ {\Phi _1}^{-1}\left( \int _x^{\infty }\Phi _1 \left( \lambda w(y)K(y,x) \right) t(y)dy \right) . \end{aligned} \end{aligned}$$

Theorem 5.4

Let \(p,q,u,u',u_p,v,v'\) be as in the Theorem 5.1, with \(1<q \le p<\infty \). Then, given that \(\Gamma _{p,u}(\mathbb {R_+})\) is an interpolation space between \(L_{2}(\mathbb {R_+})\) and \(L_{\infty }(\mathbb {R_+})\), one has (5.2) if and only if

  1. (1)
    $$\begin{aligned} {\left( \int _0^x u_p(y)dy\right) }^{\frac{1}{p}} {\left( \int _x^{\infty }v'(y)y^{-q'} dy \right) }^{\frac{1}{q'}} \le C \end{aligned}$$
  2. (2)
    $$\begin{aligned} {\left( \int _0^x v'(y) dy \right) }^{\frac{1}{q'}} {\left( \int _x^{\infty }u_p(y)y^{-p} dy \right) }^{\frac{1}{p}}\le C \end{aligned}$$
  3. (3)
    $$\begin{aligned} {\left( \int _0^x {\left( \log \frac{x}{y} \right) }^p u_p(y) dy \right) }^{\frac{1}{p}} {\left( \int _x^{\infty }v'(y)y^{-q'} dy \right) }^{\frac{1}{q'}} \le C \end{aligned}$$
  4. (4)
    $$\begin{aligned} {\left( \int _0^x u_p(y)dy\right) }^{\frac{1}{p}} {\left( \int _x^{\infty } v'(y){\left( \frac{1}{y} \log \frac{y}{x} \right) }^{q'} dy \right) }^{\frac{1}{q'}}\le C. \end{aligned}$$

Indeed (1) and (3) can be combined into

$$\begin{aligned} {\left( \int _0^\infty { \left( \log \left( 1+\frac{x}{y} \right) \right) }^p u_p(y) dy \right) }^{\frac{1}{p}} {\left( \int _x^{\infty }v'(y)y^{-q'} dy \right) }^{\frac{1}{q'}} \le C. \end{aligned}$$

Proof

The first inequality in (5.7) amounts to

$$\begin{aligned} \left[ \int _{\mathbb {R_+}}(P g )^{q'} v'\right] ^{\frac{1}{q'}} \le C \left[ \int _{\mathbb {R_+}} g^{p'} u_{p}^{1-p'}\right] ^{\frac{1}{p'}} \end{aligned}$$

and

$$\begin{aligned} \left[ \int _{\mathbb {R_+}}(Q g )^{q'} v'\right] ^{\frac{1}{q'}} \le C \left[ \int _{\mathbb {R_+}} g^{p'} u_{p}^{1-p'}\right] ^{\frac{1}{p'}}, \ \ \ 0 \le g \in M(\mathbb {R_+}), \end{aligned}$$

the latter inequality being, by duality, equivalent to

$$\begin{aligned} \left[ \int _{\mathbb {R_+}}(Pf)^{p} \, u_p\right] ^{\frac{1}{p}}\le C \left[ \int _{\mathbb {R_+}}{f}^{q} \, {v'}^{1-q}\right] ^{\frac{1}{q}}, \ \ \ 0 \le f \in M(\mathbb {R_+}). \end{aligned}$$

We illustrate the method of proof with the second inequality in (6.7) involving

$$\begin{aligned} (P^2g)(x)= \frac{1}{x} \int _0^x \log \left( \frac{x}{y} \right) g(y) dy. \end{aligned}$$

Thus, taking, in Theorem 6.3, \(K(x,y)={\log }_+ \frac{x}{y}\), \(\Phi _1(x)=x^{q'}\), \(\Phi _2(x)=x^{p'}\) (observe that \((\Phi _1 \circ {\Phi _2}^{-1})(x)= x^{\frac{q'}{p'}},\) which is convex when \(q \le p\)), \(w(y)=y^{-1}\), \(t(y)=v'(y)\), \( u(y)={u_p(y)}^{-1}\), \(v(y)=u_p(y)\) we get

$$\begin{aligned} \alpha (\lambda ,x) = \lambda ^{p'} {\left( \int _{x}^\infty v'(y) y^{-q'} dy \right) }^{\frac{p'}{q'}} \end{aligned}$$

and

$$\begin{aligned} \beta (\lambda ,x) = \lambda ^{p'} {\left( \int _{x}^\infty v'(y) \left( y^{-1}\log \frac{y}{x} \right) ^{q'} dy \right) }^{\frac{p'}{q'}}, \end{aligned}$$

from which the conditions in Theorem 5.3 yields (3) and (4). We point out that \(\lambda \) cancels. \(\square \)

The inequality (5.2) is much easier to deal with when

$$\begin{aligned} \rho _{p,u} (f) \simeq \lambda _{p,u} (f) = \left( \int _{\mathbb {R_+}} f^{*}(t)^{p} \, u(t) \, dt \right) ^{\frac{1}{p}}, \end{aligned}$$
(5.8)

which equivalence is not all that uncommon, as we will see later in this section. Indeed, given (5.8),

$$\begin{aligned} \begin{aligned} \rho _{p,u} (U f^{*})&\simeq \left( \int _{\mathbb {R_+}} \left( U f^{*} \right) (t)^{p} \, u(t) \, dt \right) ^{\frac{1}{p}} \\&= \left( \int _{\mathbb {R_+}} f^{**} (t)^{p} \, u_{p}(t) \, dt \right) ^{\frac{1}{p}}\\&= \rho _{p,u_{p}} ( f ). \end{aligned} \end{aligned}$$

We thus have

Theorem 5.5

Let p, q, u, \(u_{p}\) and v be as in Theorem 5.1. Then, given that \(\Gamma _{p,u} (\mathbb {R_+})\) is an interpolation space between \(L_{2} (\mathbb {R_+})\) and \(L_{\infty } (\mathbb {R_+})\), with \(\rho _{p,u}\) satisfying (5.8), one has

$$\begin{aligned} \rho _{ p, u} ( {\hat{f}} ) \le C \, \rho _{ p, u_{p}} (f). \end{aligned}$$
(5.9)

Moreover, there is no essentially smaller r.i.-norm that can replace \(\rho _{p, u_{p}}\) in (5.9).

Finally, there is a relatively simple condition sufficient to guarantee (5.2). It comes out of working with the inequality (5.5) and involves the norm of the dilation operator \(E_s\) as a mapping from \(\Gamma _{q,v}(\mathbb {R_+})\) to \(\Gamma _{p,u_p}(\mathbb {R_+}),\) namely,

$$\begin{aligned} h(\Gamma _{q,v},\Gamma _{p,u_p})(t)\!=\! \inf \left\{ M>0: \bar{ \rho }_{p, u_{p}}(f(ts))\!=\! {\bar{\rho }}_{p, u_{p}} \!\left( (E_tf)(s) \right) \! \le \! M {\bar{\rho }}_{{q, v}}(f) \!<\! \infty \!\right\} . \end{aligned}$$

The argument in the proof of Theorem 4.1 of [16] ensures (5.5) provided

$$\begin{aligned} \int _1^ \infty h(\Gamma _{q,v},\Gamma _{p,u_p})(t)\, \frac{dt}{t} < \infty . \end{aligned}$$

Again the argument in the proof of Theorem 5.2 in [10] yields

$$\begin{aligned} h(\Gamma _{q,v},\Gamma _{p,u_p})(t)= \sup _{s>0} \frac{ \left[ \int _0^{s/t} u_p + (s/t)^p\int _{s/t}^\infty u_p(y)y^{-p} dy \right] ^{\frac{1}{p}} }{{\left[ \int _0^s v+s^q \int _s^\infty v(y)y^{-q} dy\right] }^{\frac{1}{q}}}, \end{aligned}$$

when \(1<q\le p<\infty .\)

Altogether, we have

Theorem 5.6

Let p, q, u, \(u_{p}\) and v be as in Theorem 5.1, with \(1<q\le p<\infty \). Then, given that \(\Gamma _{p,u} (\mathbb {R_+})\) is an interpolation space between \(L_{2} (\mathbb {R_+})\) and \(L_{\infty } (\mathbb {R_+})\), one has (5.9) provided

$$\begin{aligned} \int _{1}^{\infty } \sup _{s>0} \frac{ \left[ \int _0^{s/t} u_p + (s/t)^p\int _{s/t}^\infty u_p(y)y^{-p} dy \right] ^{\frac{1}{p}} }{{\left[ \int _0^s v+s^q \int _s^\infty v(y)y^{-q} dy\right] }^{\frac{1}{q}}} \frac{dt}{t} < \infty . \end{aligned}$$

Proof

The result follows from the preceeding discussion, since (5.9) and (5.5) are equivalent when \(\Gamma _{p,u} (\mathbb {R_+})\) is an interpolation space between \(L_{2} (\mathbb {R_+})\) and \(L_{\infty } (\mathbb {R_+}).\) \(\square \)

We now consider the question of when \(\Gamma _{p,u} (\mathbb {R_+})\) is an interpolation space between \(L_{2} (\mathbb {R_+})\) and \(L_{\infty } (\mathbb {R_+}).\)

To begin, recall that \(\rho _{p,u} \simeq \lambda _{p,u} \) was shown in [2] to be equivalent to the \(B_{p}\) condition

$$\begin{aligned} t^{p} \int _{t}^{\infty } u(s) \frac{ds}{s^{p}} \le C \int _{0}^{t} u, \ \ \ t \in \mathbb {R_+}. \end{aligned}$$
(5.10)

We have

Theorem 5.7

Fix \(p \in (2, \infty )\) and a weight \(0 \le u \in M(\mathbb {R_+})\). Suppose u satisfies the \(B_{p/2}\) condition. Then, \(\Gamma _{p,u}(\mathbb {R_+})\) is an interpolation space between \(L_{2}(\mathbb {R_+})\) and \(L_{\infty }(\mathbb {R_+})\).

Proof

The \(B_{p/2}\) condition is necessary and sufficient in order that \(\rho _{p/2,u} \simeq \lambda _{p/2,u} \). Thus,

$$\begin{aligned} \begin{aligned} \rho _{p/2, u}^{(2)} (f^{*})&\simeq \left[ \int _{\mathbb {R_+}} ( f^{*}(t)^2 )^{\frac{p}{2}} \, u(t) \, dt \right] ^{ \frac{2}{p} \cdot \frac{1}{2} }\\&= \lambda _{p,u} (f^{*}). \end{aligned} \end{aligned}$$

But, \(B_{p/2}\) condition implies

$$\begin{aligned} t^p \int _{t}^{\infty } u(s) \frac{ds}{s^p} = \int _{t}^{\infty } u(s) \left( \frac{t}{s} \right) ^{p} ds \le \int _{t}^{\infty } u(s) \left( \frac{t}{s} \right) ^{\frac{p}{2}} ds \le C \int _{0}^{t} u, \ \ \ t \in \mathbb {R_+}, \end{aligned}$$

and so

$$\begin{aligned} \lambda _{p,u} (f^{*}) \simeq \rho _{p,u} (f^{*}), \ \ \ f \in M(\mathbb {R_+}). \end{aligned}$$

We conclude \(\Gamma _{p,u}(\mathbb {R_+}) = L_{ \rho _{p/2, u}^{(2)} } (\mathbb {R_+})\) and hence, in view of Theorem 3.1, \(\Gamma _{p,u}(\mathbb {R_+})\) is an interpolation space between \(L_{2}(\mathbb {R_+})\) and \(L_{\infty }(\mathbb {R_+})\). \(\square \)

Remark 5.1

G. Sinnamon in [27] proved that, given \(u \in B_{p/2}\) and provided \(0<q \le 2 \le p < \infty \), one has (5.2) if and only if

$$\begin{aligned} {\bar{\rho }}_{p, u} (\chi _{(0,t)}) \le C {\bar{\rho }}_{q, v_{q}} (\chi _{(0,t)}), \ \ \ t \in {\mathbb {R}}_+, \end{aligned}$$

with \(v_{q}(t)= v(t^{-1}) \, t^{q-2}\), \(t \in {\mathbb {R}}_+\). Theorem 5.7 and the fact that \({\bar{\rho }}_{p, u} (U f^{*}) \simeq {\bar{\rho }}_{p, u_{p}}(f^{*})\), ensures that, for \(p \in [2, \infty )\) and any \(q \in (1, \infty )\), one has (5.2) if and only if

$$\begin{aligned} {\bar{\rho }}_{p, u_{p}} (\chi _{(0,t)}) \le C {\bar{\rho }}_{q, v} (\chi _{(0,t)}), \ \ \ t \in {\mathbb {R}}_+. \end{aligned}$$

In the proof of Theorem 5.10 below we require a corollary of the following result of R. Sharpley from [25, Lemma 3.1, Corollary 3.2]

Theorem 5.8

Let \(\rho \) be an r.i. norm on \(M({\mathbb {R}}^{n})\). Suppose the fundamental indices of \(L_{ {\bar{\rho }}} (\mathbb {R_+})\) lie in (0, 1). Given \(p \in (1, \infty )\), set \(\mu _{p}(t) = \frac{ {\bar{\rho }} \left( \chi _{(0,t)}\right) ^{p} }{t}\), \(t \in \mathbb {R_+}\). Then, \({\bar{\rho }}_{p, \mu _{p}} \left( \chi _{(0,t)}\right) = {\bar{\rho }} \left( \chi _{(0,t)}\right) \), \(t \in \mathbb {R_+}\). Moreover,

$$\begin{aligned} {\bar{\rho }}_{p, \mu _{p}}(f^{*}) \simeq {\bar{\lambda }}_{p, \mu _{p}}(f^{*}), \ \ \ f \in M({\mathbb {R}}^n). \end{aligned}$$

Corollary 5.9

Let \(\rho = \rho _{p, u}\) be as in Theorem 5.8. Then, \(\rho = \rho _{p, \mu _{p}}\), where \(\mu _{p}(t) = \frac{ {\bar{\rho }} \left( \chi _{(0,t)}\right) ^{p} }{t},\) \(t \in \mathbb {R_+}\).

Proof

The spaces \(\Gamma _{p, u}({\mathbb {R}}^n)\) and \(\Gamma _{p, \mu _{p}}({\mathbb {R}}^n)\) have \({\bar{\rho }}_{p, \mu _{p}} \left( \chi _{(0,t)}\right) = {\bar{\rho }}_{p,u} \left( \chi _{(0,t)}\right) \), \(t \in \mathbb {R_+}\). As such, the spaces are identical, in view of [10, Theorem 5.1]. \(\square \)

The principal result of this section is

Theorem 5.10

Fix \(p \in [2, \infty )\) and \(0 \le u \in M(\mathbb {R_+})\), with \(\int _{R_+} \frac{u(t)}{1+t^{p}}dt < \infty \). Suppose the fundamental indices of \(\Gamma _{p,u}({\mathbb {R}}^{n})\) lie in (0, 1). Then, \(\Gamma _{p,u}({\mathbb {R}}^n)\) is an interpolation space between \(L_{p}({\mathbb {R}}^n)\) and \(L_{\infty }({\mathbb {R}}^n)\) (and hence between \(L_{2}({\mathbb {R}}^n)\) and \(L_{\infty }({\mathbb {R}}^n)\)) if and only if

$$\begin{aligned} \sup _{s \ge t} \frac{{\bar{\rho }}_{p,u} \left( \chi _{(0,s)}\right) ^{p}}{s} \le C \, \frac{{\bar{\rho }}_{p,u} \left( \chi _{(0,t)}\right) ^{p}}{t}, \end{aligned}$$
(5.11)

for some \(C>0\) independent of \(t \in \mathbb {R_+}\). Moreover, the optimal r.i. domain for \({\mathscr {F}}\) corresponding to \(\Gamma _{p,u}({\mathbb {R}}^n)\) has the norm

$$\begin{aligned} {\bar{\rho }}_{p,u} (U f^{*}) \simeq {\bar{\lambda }}_{p, u} (U f^{*}) = {\bar{\rho }}_{p, u_{p}}(f^{*}). \end{aligned}$$

Proof

Suppose first that \(p=2\). Given \( T: L_{2}({\mathbb {R}}^n), L_{\infty }({\mathbb {R}}^n) \rightarrow L_{2}({\mathbb {R}}^n), L_{\infty }({\mathbb {R}}^n)\) one has, according to [9, Theorem 1.11, p. 301] and [12],

$$\begin{aligned} \int _{0}^{t} (T f)^{*} (s)^{2} ds \le C' M_{2}^{2} \int _{0}^{Mt} f^{*} (s)^{2} ds = C' M_{2}M_{\infty } \int _{0}^{t} f^{*} (Ms)^{2} ds, \end{aligned}$$

\(f \in \left( L_{2} + L_{\infty } \right) (\mathbb {R_+})\), in which \(M= M_{\infty } / M_{2}\), \(M_{k}\) being the norm of T on \(L_{k}(\mathbb {R_+})\), \(k=2, \infty \). In view of (5.11), HLP yields

$$\begin{aligned} \begin{aligned} \int _{\mathbb {R_+}} (Tf)^{*}(t)^{2} \, \frac{{\bar{\rho }}_{2,u} \left( \chi _{(0,t)}\right) ^{2}}{t} dt&\le \int _{\mathbb {R_+}} (Tf)^{*}(t)^{2} \sup _{s \ge t} \frac{{\bar{\rho }}_{2,u} \left( \chi _{(0,s)}\right) ^{2}}{s} dt\\&\le C' M_{2} M_{\infty } \int _{\mathbb {R_+}} f^{*}(Mt)^{2} \sup _{s \ge t} \frac{{\bar{\rho }}_{2,u} \left( \chi _{(0,s)}\right) ^{2}}{s} dt\\&\le C C' M_{2} M_{\infty } \int _{\mathbb {R_+}} f^{*}(Mt)^{2} \, \frac{{\bar{\rho }}_{2,u} \left( \chi _{(0,t)}\right) ^{2}}{t} dt, \end{aligned} \end{aligned}$$

\(f \in \left( L_{2} + L_{\infty } \right) (\mathbb {R_+})\). Theorem 5.8 now ensures the latter is equivalent to

$$\begin{aligned} \int _{\mathbb {R_+}} (Tf)^{**}(t)^{2} \, \frac{{\bar{\rho }}_{2,u} \left( \chi _{(0,t)}\right) ^{2}}{t} dt \le C C' M_{2} M_{\infty } h(M)^2 \int _{\mathbb {R_+}} f^{**}(t)^{2} \, \frac{{\bar{\rho }}_{2,u} \left( \chi _{(0,t)}\right) ^{2}}{t} dt, \end{aligned}$$

where h(t) is the norm of the dilation operator \(E_{t}\) on \(\Gamma _{ 2, \mu _{2} } (\mathbb {R_+}) = \Gamma _{2,u}(\mathbb {R_+})\), \(\mu _{2}(s)= \frac{{\bar{\rho }}_{2,u} \left( \chi _{(0,s)}\right) ^{2}}{s}\), by Corollary 5.9, that is, \(T: \Gamma _{2,u}({\mathbb {R}}^n) \rightarrow \Gamma _{2,u}({\mathbb {R}}^n)\). Thus, \(\Gamma _{2,u}({\mathbb {R}}^n)\) is between \(L_{2}({\mathbb {R}}^n)\) and \(L_{ \infty }({\mathbb {R}}^n)\).

Suppose, next, \(p>2\). The “if” part of our theorem will follow in this case if we can show (5.11) implies the weight \(w(t)= \frac{{\bar{\rho }}_{p,u} \left( \chi _{(0,t)}\right) ^{p}}{t}\) satisfies \(B_{p/2}\) condition. But,

$$\begin{aligned} \begin{aligned} t^{p/2} \int _{t}^{\infty } \frac{{\bar{\rho }}_{p,u} \left( \chi _{(0,s)}\right) ^{p}}{s} \frac{ds}{s^{p/2}}&\le t^{p/2} \int _{t}^{\infty } \sup _{y \ge s} \frac{{\bar{\rho }}_{p,u} \left( \chi _{(0,y)}\right) ^{p}}{y} \frac{ds}{s^{p/2}}\\&\le t^{p/2} \sup _{y \ge t} \frac{{\bar{\rho }}_{p,u} \left( \chi _{(0,y)}\right) ^{p}}{y} \int _{t}^{\infty }\frac{ds}{s^{p/2}}\\&\le C \, t \, \frac{{\bar{\rho }}_{p,u} \left( \chi _{(0,t)}\right) ^{p}}{t} \\&\le C^2 \int _{0}^{t} \frac{{\bar{\rho }}_{p,u} \left( \chi _{(0,s)}\right) ^{p}}{s} ds, \ \ \ t \in \mathbb {R_+}. \end{aligned} \end{aligned}$$

This completes the proof of “if” part.

As for the “only if” part we rely on a result of L. Maligranda [18] asserting that if \(L_{\rho }({\mathbb {R}}^n)\) is an interpolation space between \(L_{p}({\mathbb {R}}^n)\) and \(L_{\infty }({\mathbb {R}}^n)\), then

$$\begin{aligned} \frac{ {\bar{\rho }} (\chi _{(0,s)}) }{ {\bar{\rho }} (\chi _{(0,t)}) } \le C \max \left[ \left( \frac{s}{t} \right) ^{\frac{1}{p}}, 1 \right] . \end{aligned}$$
(5.12)

Indeed, for \(t \le s\), (5.12) yields

$$\begin{aligned} \frac{ {\bar{\rho }} (\chi _{(0,s)}) }{ {\bar{\rho }} (\chi _{(0,t)}) } \le C \left( \frac{s}{t} \right) ^{\frac{1}{p}} \end{aligned}$$

or

$$\begin{aligned} \frac{ {\bar{\rho }} (\chi _{(0,s)})^{p} }{ s } \le C \, \frac{ {\bar{\rho }} (\chi _{(0,t)})^{p} }{t}, \end{aligned}$$

from which (5.11) follows. \(\square \)

To this point the Lorentz Gamma range norms have been equivalent to functionals of the form

$$\begin{aligned} \lambda _{p,u} (f) = \left[ \int _{\mathbb {R_+}} f^{*}(s)^{p} u(t) dt \right] ^{\frac{1}{p}}. \end{aligned}$$

This need not be the case for the \(\rho _{ 2p, u} \) in Theorem 5.12 below.

Lemma 5.11

Fix \(p \in (1, \infty )\) and \(0 \le u \in M(\mathbb {R_+})\), with

$$\begin{aligned} \int _{\mathbb {R_+}} \frac{ u(t) }{ 1+t^{p} } dt < \infty . \end{aligned}$$

Then,

$$\begin{aligned} \left( \int _{\mathbb {R_+}} f^{**}(t)^{p} \, u(t) \, dt \right) ^{\frac{1}{p}} \le \left( \int _{\mathbb {R_+}} f^{*}(t)^{p} \, u^{(p)}(t) \, dt \right) ^{\frac{1}{p}}, \ \ \ f \in M(\mathbb {R_+}), \end{aligned}$$
(5.13)

where

$$\begin{aligned} u^{(p)}(t) = p \,t^{p-1} \int _{t}^{\infty } u(s) s^{-p} ds, \ \ \ t \in \mathbb {R_+}; \end{aligned}$$

moreover, \(u^{(p)}\) is essentially the smallest weight for which (5.13) holds.

Proof

It is shown in [21] that

$$\begin{aligned} \left( \int _{\mathbb {R_+}} f^{**}(t)^{p} \, u(t) \, dt \right) ^{\frac{1}{p}} \le \left( \int _{\mathbb {R_+}} f^{*}(t)^{p} \, v (t) \, dt \right) ^{\frac{1}{p}}, \ \ \ f \in M(\mathbb {R_+}), \end{aligned}$$

if and only if

$$\begin{aligned} \int _{0}^{t} u(s) ds + t^{p} \int _{t}^{\infty } u(s) s^{-p} ds \le C \int _{0}^{t} v, \ \ \ t \in \mathbb {R_+}. \end{aligned}$$

But,

$$\begin{aligned} \begin{aligned} \int _{0}^{t} u^{(p)} (s) ds&= \int _{0}^{t} p \, s^{p-1} \int _{s}^{\infty } u(y) y^{-p} dy \, ds\\&= \int _{0}^{t} p \, s^{p-1} \int _{s}^{t} u(y) y^{-p} dy \, ds + \left[ \int _{0}^{t} p \, s^{p-1} ds \right] \left[ \int _{t}^{\infty } u(s) s^{-p} ds \right] \\&= \int _{0}^{t} \left( \int _{0}^{y} p \, s^{p-1} ds \right) u(y) y^{-p} dy + t^{p} \int _{t}^{\infty } u(s) s^{-p} ds \\&= \int _{0}^{t} u + t^{p} \int _{t}^{\infty } u(s) s^{-p} ds. \end{aligned} \end{aligned}$$

We conclude that

$$\begin{aligned} \left( \int _{\mathbb {R_+}} f^{**}(t)^{p} \, u(t) \, dt \right) ^{\frac{1}{p}} \le C \left( \int _{\mathbb {R_+}} f^{*}(t)^{p} \, u^{(p)}(t) \, dt \right) ^{\frac{1}{p}}, \ \ \ f \in M(\mathbb {R_+}). \end{aligned}$$

\(\square \)

Theorem 5.12

Let p and u be as in Lemma 5.11. Then,

$$\begin{aligned} \rho _{ 2p, u } ( {\widehat{f}} \, ) = {\bar{\rho }}_{ 2p, u } ( ( {\widehat{f}} )^{*} ) \le C {\bar{\rho }}_{ 2p, u^{(p)}_{2p} } ( f^{*} ) = \rho _{ 2p, u^{(p)}_{2p} } ( f ), \end{aligned}$$
(5.14)

where

$$\begin{aligned} u^{(p)}_{2p} (t)\!=\! u^{(p)} ( t^{-1} ) \, t^{2p -2} \!=\! p(t^{-1})^{p-1}\! \left( \int _{t^{-1}}^{\infty } u(s) s^{-p} ds \right) t^{2p -2} \!=\! p \, t^{p-1}\! \int _{t^{-1}}^{\infty }\! u(s)s^{-p} ds.\end{aligned}$$

Proof

Applying the construction in (3.1) to the functionals in (5.13) yields

$$\begin{aligned} \left( \int _{\mathbb {R_+}} \left( t^{-1} \int _{0}^{t} f^{*}(s)^{2} ds \right) ^{p} \, u(t) \, dt \right) ^{\frac{1}{2p}} \le \left( \int _{\mathbb {R_+}} f^{*}(t)^{2p} \, u^{(p)}(t) \, dt \right) ^{\frac{1}{2p}} = \lambda _{2p, u^{(p)}} (f). \end{aligned}$$

Again,

$$\begin{aligned} \left( t^{-1} \int _{0}^{t} f^{*}(s) ds \right) ^{2p} \le \left( t^{-1} \int _{0}^{t} f^{*}(s)^{2} ds \right) ^{p} \end{aligned}$$

by Hölder’s inequality.

Hence, using HLP in (2.8), yields

$$\begin{aligned} \begin{aligned} \rho _{ 2p, u } ( {\widehat{f}} \, )&\le \rho _{ p, u } ( \, |{\widehat{f}} \, |^{ \ 2} \, )^{1/2} \\&\le C \rho _{ p, u } ( (U f^{*})^{ 2} \, )^{1/2} \\&\le C \lambda _{ 2p, u^{(p)} } ( U f^{*} ) \\&= C \left( \int _{\mathbb {R_+}} (U f^{*})(t)^{2p} \, u^{(p)}(t) \, dt \right) ^{\frac{1}{2p}} \\&= C \left( \int _{\mathbb {R_+}} f^{**}(t)^{2p} \, u^{(p)}_{2p}(t) \, dt \right) ^{\frac{1}{2p}} \\&= C \rho _{ 2p, u^{(p)}_{2p} } ( f ). \end{aligned} \end{aligned}$$

\(\square \)

Example 5.1

Fix p, \(1<p< \infty \), and set

$$\begin{aligned} u(t)= {\left\{ \begin{array}{ll} t^{2p-1} \left( \log \textstyle \frac{1}{t} \right) ^{- \alpha }, &{} 0<t<1,\\ t^{p-1- \alpha }, &{} t>1, \end{array}\right. } \end{aligned}$$

with \(0<\alpha <1\). Then, one has

$$\begin{aligned} \rho _{ 2p, u } (f) \not \simeq \lambda _{2p, u} (f), \ \ \ f \in M(\mathbb {R_+}), \end{aligned}$$

or, equivalently,

$$\begin{aligned} t^{2p} \int _{t}^{\infty } u(s) s^{-2p} ds \le C \int _{0}^{t} u, \ \ \ t \in \mathbb {R_+}, \end{aligned}$$
(5.15)

does not hold. Indeed, the left hand side of (5.15) is equal to \(C t^{2p} \left( \log \textstyle \frac{1}{t} \right) ^{- \alpha + 1}\), while the right hand side is

$$\begin{aligned} \int _{0}^{t} u = \int _{0}^{t} s^{2p-1} \left( \log \textstyle \frac{1}{s} \right) ^{- \alpha } \simeq t^{2p} \left( \log \textstyle \frac{1}{t} \right) ^{- \alpha }, \ \ \ 0<t<1, \end{aligned}$$

in view of L’Hôspital rule. The ratio of the left side to the right side in (5.15) is, essentially, \(\log \textstyle \frac{1}{t}\) which \(\rightarrow \infty \) as \(t \rightarrow 0^{+}\).

6 Other Work

Inequalities involving Fourier transform other than those considered in this paper are weighted Lebesgue inequalities

$$\begin{aligned} \left( \int _{{\mathbb {R}}^n} |{\hat{f}}(x) \, w(x) |^{q} \, dx \right) ^{\frac{1}{q}} \le C \left( \int _{{\mathbb {R}}^n} |f(x) \, v(x) |^{p} \, dt \right) ^{\frac{1}{p}} \end{aligned}$$

and weighted Lorentz inequalities

$$\begin{aligned} \left( \int _{\mathbb {R_+}} ({\hat{f}} \, )^{*}(t)^{q} \, w(t) \, dt \right) ^{\frac{1}{q}} \le C \left( \int _{\mathbb {R_+}} \left( \int _{0}^{1/t} f^{*} \, \right) ^{p} \, v(t) \, dt \right) ^{\frac{1}{p}}, \end{aligned}$$

in which \(0 \le v, w \in M({\mathbb {R}}^n)\) and \(1< p, q < \infty \).

In both [15, 20] conditions are for the Lebesgue inequalities that apply not just to w and v but to all weights equimeasurable with them. The extreme cases of these are the decreasing rearrangement, W, of w and the increasing rearrangement, V, of v. This reduces the considerations to the case \(w \downarrow \) and \(v \uparrow \).

The weighted Lebesgue inequalities are shown in [15, 20] to be equivalent to inequalities of the form, for example when \(1 \le p \le q \le \infty \),

$$\begin{aligned} \left( \int _{0}^{t^{-1}} w \right) ^{\frac{1}{q}} \left( \int _{0}^{t} w^{- \frac{1}{p-1}} \right) ^{\frac{q}{p'}} \le B, \ \ \ t \in {\mathbb {R}}_+. \end{aligned}$$
(6.1)

In [15] the sufficiency is proved using the inequality (2.8) from [13]. The necessity comes out of the inequality

$$\begin{aligned} \rho _{q} \left( ({\hat{f}} )^{*} w \right) \simeq \rho _{2} \left( U f^{*} w \right) \end{aligned}$$

from [15]. The proofs in [20] are more complicated. The conditions for the weighted Lorentz inequalities are similar to (6.1).

A brief survey of papers on these inequalities, from the pioneering work of Benedetto and Heinig [5] through that of G. Sinnamon [26] and Rastegari and Sinnamon [24], is given in the paper [22] of Nursultanov and Tikhonov.

In this paper we have seen the behaviour of \({\mathscr {F}}\) on r.i. spaces depends on its action on radially decreasing functions. But what about the size of f if \({\hat{f}}\) is radially decreasing? This question is taken up in [11] in the context of Fourier series where functions with a cosine series having decreasing coefficients as \(|n| \rightarrow \infty \) are studied.