Skip to main content

Quasinormed spaces generated by a quasimodular

Abstract

In this paper, we introduce the notion of a quasimodular and we prove that the respective Minkowski functional of the unit quasimodular ball becomes a quasinorm. In this way, we refer to and complete the well-known theory related to the notions of a modular and a convex modular that lead to the F-norm and to the norm, respectively. We use the obtained results to consider the basic properties of quasinormed Calderón–Lozanovskiĭ spaces \(E_{\varphi}\), where the lower Matuszewska–Orlicz index \(\alpha _{\varphi}\) plays the key role. Our studies are conducted in a full possible generality.

1 Introduction

The subject of quasinormed or F-normed spaces (which are a natural generalization of normed spaces) has been the object of research by many authors for a long time (see [1, 4, 1016, 22, 25, 27]). Recall also that the so-called Δ-norm is a generalization of both a quasinorm and an F-norm (see [12]). From a different point of view, the theory of modular spaces has also been developed (see [21, 26]).

We will try to link these threads. Namely, we will introduce a new notion of the functional ρ that we will call a quasimodular. This is an essential generalization of the concept of a convex semimodular. Furthermore, this quasimodular ρ is defined in such a way that the respective Minkowski functional of the unit quasimodular ball gives a quasinorm. We also check if the basic properties from the classical theory of modular spaces are still true for quasimodular spaces.

Then, we apply our general concept to a special class of quasimodular spaces, that is, the quasinormed Calderón–Lozanovskiĭ spaces \(E_{\varphi}\) generated by a quasimodular \(\rho _{\varphi}^{E}\). We focus mainly on the basic relations between a positive lower Matuszewska–Orlicz index \(\alpha _{\varphi}^{E}\) and the fact that \(\rho _{\varphi}^{E}\) \((\|\cdot \|_{\varphi})\) is a quasimodular (a quasinorm), respectively. We study carefully these relations in both directions. We will finish the article by characterizing the quasinormed Orlicz–Lorentz (Orlicz) spaces as spaces generated by a suitable quasimodular. We admit nonconvex, nondecreasing, and degenerated Orlicz functions, which give the full generality of studies. Some geometric properties of these spaces have been studied in [14] and, in the particular case for \(E=L^{1}\), in [16], but the authors consider them directly with a quasinorm (not as spaces generated by a “modular”).

2 Preliminaries

Definition 2.1

Given a real vector space X the functional \(x\mapsto \Vert x\Vert \) is called a quasinorm if the following three conditions are satisfied:

  1. (i)

    \(\Vert x\Vert =0\) if and only if \(x=0\);

  2. (ii)

    \(\Vert ax\Vert =|a|\Vert x\Vert \) for any \(x\in X\) and \(a\in \mathbb{R}\);

  3. (iii)

    there exists \(C=C_{X}\geq 1\) such that \(\Vert x+y\Vert \leq C(\Vert x\Vert +\Vert y\Vert )\) for all \(x,y\in X\).

For \(0< p\leq 1\), the functional \(x\mapsto \Vert x\Vert _{1}\) is called a p-norm if it satisfies the first two conditions of the quasinorm and the condition \(\Vert x+y\Vert _{1}^{p}\leq \Vert x\Vert _{1}^{p}+\Vert y\Vert _{1}^{p}\) for any \(x,y\in X\). Clearly, each p-norm is a quasinorm. By the Aoki–Rolewicz theorem (cf. [12, Theorem 1.3 on page 7], [25, page 86]), given a quasinorm \(\Vert \cdot \Vert \), if \(0< p\leq 1\) is such that \(C=2^{1/p-1}\), then there exists a p-norm \(\Vert \cdot \Vert _{1}\) which is equivalent to \(\Vert \cdot \Vert \), that is

$$ \Vert x \Vert _{1}\leq \Vert x \Vert \leq 2C \Vert x \Vert _{1} $$
(2.1)

for all \(x\in X\). The quasinorm \(\Vert \cdot \Vert \) induces a metric topology on X: in fact a metric can be defined by \(d(x,y)=\Vert x-y\Vert _{1}^{p}\). We say that \(X=(X,\Vert \cdot \Vert )\) is a quasi-Banach space if it is complete for this metric. Let us note that a lot of important information and results on quasi-Banach spaces can be found in [11], see also [12].

Recall that a quasi-Banach lattice E is called order continuous (\(E\in ({\mathrm{OC}})\)) if for each sequence \(x_{n}\downarrow 0\), that is \(x_{n}\geq x_{n+1}\) and \(\inf_{n}x_{n}=0\), we have \(\Vert x_{n}\Vert _{E}\rightarrow 0\) (see [17, 23, 28]).

3 Quasimodular spaces

In this section, we introduce the notions of a quasimodular and a quasimodular space.

Definition 3.1

Let X be a real linear space. We say that a function \(\rho :X\rightarrow {}[ 0,\infty ]\) is a quasimodular whenever for all \(x,y\in X\) the following conditions are satisfied:

  1. (i)

    \(\rho ( 0 ) =0\) and the condition \(\rho (\lambda x )\leq 1\) for all \(\lambda >0\) implies that \(x=0\);

  2. (ii)

    \(\rho ( -x ) =\rho ( x )\);

  3. (iii)

    \(\rho (\lambda x)\) is a nondecreasing function of λ, where \(\lambda \geq 0\);

  4. (iv)

    There is \(M\geq 1\) such that

    $$ \rho (\alpha x+\beta y )\leq M \bigl[\rho (x )+ \rho (y ) \bigr] $$

    provided \(\alpha ,\beta \geq 0\) and \(\alpha +\beta =1\);

  5. (v)

    There is a constant \(p>0\) such that for all \(\varepsilon >0\) and all \(A>0\) there exists \(K=K(\varepsilon ,A)\geq 1\) such that

    $$ \rho (ax )\leq Ka^{p}\rho (x )+\varepsilon $$

    for any \(0< a\leq 1\) whenever \(\rho (x)\leq A\).

Remark 3.2

\((i)\) The above definition, in particular the condition \((v)\), has been introduced in such a way as to cover the largest possible class of mappings ρ and to provide a quasinorm (see Theorem 3.4). As we will show in Theorem 4.10, condition \((v)\) can be simplified in some particular cases (more precisely, some particular modulars satisfy condition \((v)\) in a simpler or stronger form).

\((ii)\) Obviously, a convex modular (more precisely a convex semimodular) defined in [26] is in particular a quasimodular. As we will show in Example 4.15\((ii)\) and \((iii)\), the concepts of quasimodular and modular (more precisely a semimodular, which induces an F-norm) defined in [26] are incomparable.

If ρ is a quasimodular on X, then

$$ X_{\rho}:= \Bigl\{ x\in X\colon \lim_{\lambda \rightarrow 0}\rho ( \lambda x)=0 \Bigr\} $$

is called a quasimodular space. It is easy to show that \(X_{\rho}\) is a linear subspace of X. We also obtain the following:

Lemma 3.3

For any quasimodular ρ, we have

$$ X_{\rho}= \bigl\{ x\in X\colon \rho (\lambda x)< \infty \textit{ for some }\lambda >0 \bigr\} . $$

Proof

Note that in order to prove this lemma, it is enough to show that if \(\rho (\lambda _{0}x)<\infty \) for some \(\lambda _{0}>0\), then \(\lim_{\lambda \rightarrow 0}\rho (\lambda x)=0\). Let \(x\in X\) and \(\rho (\lambda _{0}x)<\infty \) for some \(\lambda _{0}>0\). We prove that for any \(\varepsilon \in (0,\rho (\lambda _{0}x))\) there exists \(\delta =\delta (\varepsilon )>0\) such that \(\rho (\lambda x)<\varepsilon \), whenever \(\lambda <\delta \). Let \(\varepsilon \in (0,\rho (\lambda _{0}x))\) be fixed and take \(K=K(\frac{\varepsilon}{2},\rho (\lambda _{0}x))\) from condition \((v)\) of Definition 3.1. Then, for \(\lambda <\delta \), where \(\delta =\lambda _{0}(\frac{\varepsilon}{2K\rho (\lambda _{0}x)})^{1/p}\), we obtain

$$ \rho (\lambda x )=\rho \biggl( \frac{\lambda}{\lambda _{0}}\lambda _{0}x \biggr)\leq K \biggl( \frac{\lambda}{\lambda _{0}} \biggr)^{p}\rho (\lambda _{0}x )+\varepsilon /2< \varepsilon . $$

 □

Theorem 3.4

Let ρ be a quasimodular on X. Then, the functional

$$ \Vert x \Vert _{\rho }=\inf \bigl\{ \lambda >0:\rho (x/\lambda )\leq 1 \bigr\} $$

is a quasinorm on \(X_{\rho}\).

Proof

The condition \(\Vert 0 \Vert _{\rho}=0\) is obvious. Suppose \(x\neq 0\). Then, by condition \((i)\), there exists \(\lambda >0\) such that \(\rho (\lambda x )>1\) and, by condition \((iii)\), \(\Vert x \Vert _{\rho }\geq 1/\lambda \).

For any \(x\in X_{\rho}\) and all \(\alpha \in \mathbb{R}\), exactly the same way as in [26], we obtain

$$ \Vert \alpha x \Vert _{\rho}=\inf \biggl\{ \lambda >0:\rho \biggl( \frac{\alpha x}{\lambda} \biggr)\leq 1 \biggr\} = \vert \alpha \vert \inf \biggl\{ \lambda / \vert \alpha \vert >0:\rho \biggl(\frac{x}{\lambda /\alpha} \biggr) \leq 1 \biggr\} = \vert \alpha \vert \Vert x \Vert _{\rho}. $$

Finally, we prove the quasitriangle inequality. Let \(0<\varepsilon <\frac{1}{2M}\) be fixed and take \(K=K(\varepsilon ,1)\) (where constants M and K arise from conditions \((iv)\) and \((v)\) of Definition 3.1). Defining

$$ C= \biggl(\frac{K}{\frac{1}{2M}-\varepsilon} \biggr)^{1/p}, $$

for all \(x,y\in X_{\rho }\) and any \(\delta >0\), we obtain

$$\begin{aligned} & \rho \biggl( \frac{x+y}{C ( \Vert x \Vert _{\rho }+ \Vert y \Vert _{\rho }+\delta ) } \biggr) \\ &\quad =\rho \biggl( \frac{ \Vert x \Vert _{\rho}+\delta /2}{ \Vert x \Vert _{\rho }+ \Vert y \Vert _{\rho}+\delta} \frac{x}{C( \Vert x \Vert _{\rho}+\delta /2)}+ \frac{ \Vert y \Vert _{\rho}+\delta /2}{ \Vert x \Vert _{\rho}+ \Vert y \Vert _{\rho }+\delta} \frac{y}{C( \Vert y \Vert _{\rho}+\delta /2)} \biggr) \\ &\quad \leq M \biggl[\rho \biggl( \frac{x}{C( \Vert x \Vert _{\rho}+\delta /2)} \biggr) + \rho \biggl( \frac{y}{C( \Vert y \Vert _{\rho}+\delta /2)} \biggr) \biggr] \\ &\quad \leq M \biggl[\frac{K}{C^{p}}\cdot \rho \biggl( \frac{x}{( \Vert x \Vert _{\rho}+\delta /2)} \biggr)+ \varepsilon +\frac{K}{C^{p}}\cdot \rho \biggl( \frac{y}{( \Vert y \Vert _{\rho}+\delta /2)} \biggr)+ \varepsilon \biggr] \\ &\quad \leq 2M \biggl(\frac{K}{C^{p}}+\varepsilon \biggr)=1, \end{aligned}$$

whence

$$ \Vert x+y \Vert _{\rho }\leq C \bigl( \Vert x \Vert _{\rho }+ \Vert y \Vert _{\rho }+\delta \bigr). $$

By the arbitrariness of δ we have \(\Vert x+y \Vert _{\rho }\leq C( \Vert x \Vert _{\rho }+ \Vert y \Vert _{\rho})\). □

By the definition of quasinorm, we obtain immediately the following:

Lemma 3.5

Let ρ be a quasimodular on X. Then, for any \(x\in X_{\rho}\) the following statements hold:

\((i)\) If \(\rho (x)\leq 1\), then \(\|x\|_{\rho}\leq 1\);

\((ii)\) If ρ is left continuous \((\lim_{\lambda \rightarrow 1-}\rho (\lambda x)=\rho (x) \textit{ for all }x\in X_{\rho})\), then \(\rho (x)\leq 1\) whenever \(\|x\|_{\rho}\leq 1\);

\((iii)\) If \(\|x\|_{\rho}<1\), then \(\rho (x)\leq 1\);

\((iv)\) If ρ is right continuous \((\lim_{\lambda \rightarrow 1+}\rho (\lambda x)=\rho (x) \textit{ for all }x\in X_{\rho})\), then \(\|x\|_{\rho}<1\) whenever \(\rho (x)<1\).

Remark 3.6

As we will show in Example 4.15\((iv)\) the implication, if \(\|x\|_{\rho}<1\) then \(\rho (x)<1\), is not always true. Recall that this implication is true for any modular as well as for any convex modular (see [26]).

Lemma 3.7

For each sequence \(( x_{n} ) \) in \(X_{\rho }\) we have \(\lim_{n\rightarrow \infty} \Vert x_{n} \Vert _{\rho }=0\) if and only if \(\lim_{n\rightarrow \infty}\rho (\lambda x_{n} )=0\) for all \(\lambda >0\).

Proof

The implication, if \(\lim_{n\rightarrow \infty}\rho (\lambda x_{n} )=0\) for all \(\lambda >0\) then \(\lim_{n\rightarrow \infty} \Vert x_{n} \Vert _{\rho }=0\) is obvious. Let now \(\|x_{n}\|_{\rho}\rightarrow 0\). Fix \(\lambda >0\) and \(\varepsilon \in (0,1)\) and let \(K=K(\varepsilon /2,1)\) be the constant from condition \((v)\) of Definition 3.1. Then, there exists \(n_{\lambda ,\varepsilon}\) such that \(\|\lambda x_{n}\|_{\rho}\leq (\varepsilon /4K)^{1/p}\) for all \(n\geq n_{\lambda ,\varepsilon}\). Hence, for each \(a\in ((\varepsilon /4K)^{1/p},(\varepsilon /2K)^{1/p})\) we obtain

$$ \rho (\lambda x_{n})=\rho \biggl(a\frac{\lambda x_{n}}{a} \biggr) \leq Ka^{p}\rho \biggl(\frac{\lambda x_{n}}{a} \biggr)+ \frac{\varepsilon}{2}\leq \varepsilon . $$

By the arbitrariness of ε, for any \(\lambda >0\) we obtain \(\lim_{n\rightarrow \infty}\rho (\lambda x_{n})=0\). □

4 Quasinormed Calderón–Lozanowskiĭ spaces

A triple \((T,\Sigma ,\mu )\) stands for a positive, complete, and σ-finite measure space and \(L^{0}=L^{0}(T,\Sigma ,\mu )\) denotes the space of all (equivalence classes of) Σ-measurable functions \(x:T\rightarrow \mathbb{R}\). For every \(x\in L^{0}\) we denote \(\operatorname{supp}x=\{ t\in T:x(t)\neq 0\}\). Moreover, for any \(x,y\in L^{0}\), we write \(x\leq y\), if \(x(t)\leq y(t)\) almost everywhere with respect to the measure μ on the set T.

A quasinormed lattice [quasi-Banach lattice] \(E=(E,\leq ,\Vert \cdot \Vert _{E})\) is called a quasinormed ideal space [quasi-Banach ideal space (or a quasi-Köthe space)] if it is a linear subspace of \(L^{0}\) satisfying the following conditions:

  1. (i)

    If \(x\in L^{0}\), \(y\in E\) and \(|x|\leq |y|\) μ-a.e., then \(x\in E \) and \(\Vert x\Vert _{E}\leq \Vert y\Vert _{E}\);

  2. (ii)

    There exists \(x\in E\) that is strictly positive on the whole T.

By \(E_{+}\) we denote the positive cone of E, that is, \(E_{+}={\{x\in E:x\geq 0\}}\). Let \(C_{E}\) be the constant from the quasitriangle inequality for E. In turn, by \(E(w)\) we denote the weighted quasinormed ideal space, that is,

$$ E(w)=\bigl\{ x\in L^{0}\colon xw\in E\bigr\} $$

with the quasinorm \(\|x\|_{E(w)}=\|xw\|_{E}\), where \(w:T\rightarrow (0,\infty )\) is a measurable weight function.

We say that a quasinormed ideal space E has the Fatou property, if for all \(x\in L^{0}\) and any \(( x_{n} ) _{n=1}^{\infty }\) in \(E_{+}\) such that \(x_{n}\uparrow |x|\) μ-a.e and \(\sup_{n\in {N}}\Vert x_{n}\Vert _{E}<\infty \), we obtain \(x\in E\) and \(\lim_{n}\Vert x_{n}\Vert _{E}=\Vert x\Vert _{E}\). It is well known that E has the Fatou property if and only if for each \(x\in L^{0}\) and all \(( x_{n} ) _{n=1}^{\infty }\) in E such that \(x_{n}\rightarrow x\) μ-a.e and \(\liminf_{n\in {N}}\Vert x_{n}\Vert _{E}<\infty \), we have \(x\in E\) and \(\Vert x\Vert _{E}\leq \liminf_{n}\Vert x_{n}\Vert _{E}\) (cf. [2, Lemma 1.5 on page 4]).

Lemma 4.1

Let E be a quasinormed ideal space.

  1. (i)

    If \(\lim_{n\rightarrow \infty}\|x-x_{n}\|_{E}=0\), where \(x\in E\) and \((x_{n})_{n=1}^{\infty}\) is a sequence in E, then \(x_{n}\rightarrow x\) locally in measure.

  2. (ii)

    For any Cauchy sequence \((x_{n})_{n=1}^{\infty}\) in E there exists \(x\in L^{0}\) such that \(x_{n}\rightarrow x\) locally in measure.

Proof

This lemma can be proved analogously as Theorem 1 on page 96 in [17]. Indeed, assuming in (2) on page 96 that \(\|x_{n}-x\|_{E}<\frac{\varepsilon}{2^{n}C_{E}^{n}}\), we obtain \(\|\chi _{B_{n}}\|_{E}<\frac{1}{2^{n}C_{E}^{n}}\) (see (5) on page 96) and, in consequence:

$$ \Vert \chi _{D_{n}} \Vert _{E}\leq \Vert \chi _{C_{ms_{m}}} \Vert _{E}\leq \Biggl\Vert \sum _{k=m+1}^{m+s_{m}}\chi _{B_{k}} \Biggr\Vert _{E}\leq \sum_{k=m+1}^{m+s_{m}}C_{E}^{(k-m)} \Vert \chi _{B_{k}} \Vert _{E}< \sum _{k=m+1}^{m+s_{m}} \frac{C_{E}^{(k-m)}}{2^{k}C_{E}^{k}}< \frac{1}{2^{m}}, $$

(see page 97, line 5). □

Lemma 4.2

[14, Lemma 2.1] A quasinormed ideal space E with the Fatou property is complete.

Proof

We recall a short proof of this lemma for the sake of completeness. By the Aoki–Rolewicz theorem, it is enough to show that for any Cauchy sequence \((x_{n})_{n=1}^{\infty}\) in E there exists \(x\in E\) such that \(\lim_{n\rightarrow \infty}\|x-x_{n}\|_{E}=0\). If \((x_{n})_{n=1}^{\infty}\) is a Cauchy sequence in E, then by Lemma 4.1\((ii)\), \(x_{n}\rightarrow x\) locally in measure for some \(x\in L^{0}\). Without loss of generality (passing to subsequences and applying the double extract convergence theorem, if necessary), we can assume \(x_{n}\rightarrow x\) μ-a.e. Hence, by the Fatou property, we obtain \(x\in E\) and \(\|x-x_{n}\|_{E}\leq \liminf_{m\rightarrow \infty}\|x_{m}-x_{n}\|_{E}\) for each \(n\in \mathbb{N}\), which completes the proof. □

The following basic fact, very well known for Banach ideal spaces (see [17, Lemma 2, p. 97]), is also true for quasi-Banach ideal spaces.

Lemma 4.3

Let \(( E, \Vert \cdot \Vert _{E} ) \) be a quasi-Banach ideal space. If \(\Vert x_{n} \Vert _{E}\rightarrow 0\), then there exists a subsequence \(( x_{n_{k}} ) _{k=1}^{\infty }\), an element \(y\in E_{+}\) and a sequence \(\varepsilon _{k}\downarrow 0\) such that \(\vert x_{n_{k}} \vert \leq \varepsilon _{k}\cdot y\) for each k.

Proof

If \(\Vert x_{n} \Vert _{E}\rightarrow 0\), then we can find a subsequence \(( x_{n_{k}} ) _{k=1}^{\infty }\) such that \(\Vert x_{n_{k}} \Vert <\frac{1}{C^{k}_{E}2^{k}}\) for any \(k\in \mathbb{N}\). Consequently,

$$ \sum_{k=1}^{\infty}C_{E}^{k} \Vert k\cdot x_{n_{k}} \Vert \leq \sum_{k=1}^{\infty}C^{k}_{E} \frac{k}{C^{k}_{E}2^{k}}= \sum_{k=1}^{\infty } \frac{k}{2^{k}}< \infty . $$

By Theorem 1.1 from [25], \(y:=\sum_{k=1}^{\infty} \vert k\cdot x_{n_{k}} \vert \in E_{+}\). Thus, \(\vert x_{n_{k}} \vert \leq \frac{y}{k}\) for all k. Taking \(\varepsilon _{k}=1/k\) we complete the proof. □

From now on, we will assume that E is a quasi-Banach ideal space with the Fatou property. During our studies we will consider three natural classes of E:

\(( 1 ) \) neither \(L_{\infty}\subset E\) nor \(E\subset L_{\infty}\);

\(( 2 ) \) \(L_{\infty}\subset E\);

\(( 3 ) \) \(E\subset L_{\infty}\).

Let \(T=[0,\gamma )\), \(0<\gamma \leq \infty \), and μ be the Lebesgue measure. Also, the space \(E=L_{p}\), \(0< p<\infty \), belongs to the class \((1)\) if \(\gamma =\infty \) and the class \((2)\) otherwise. Moreover, the space \(L_{1}\cap L_{\infty}\) belongs to the class \((3)\) whenever \(\gamma =\infty \). Let now \(T=\mathbb{N}\) and \(\mu =m\) be a counting measure. Then, the space \(l_{p}\), \(p\in (0,\infty )\), belongs to class \((3)\), the weighted sequence space \(l_{1}(w)\), \(w=(w(n))_{n=1}^{\infty}\) and \(\sum_{n=1}^{\infty}w(n)<\infty \), belongs to class \((2)\) and the Cesàro sequence space \(ces_{p}\), \(1< p<\infty \) (see [19] for the respective definition), belongs to class \((1)\).

Remark 4.4

Let \(E\subset L_{\infty }\). Then, by the closed-graph theorem that is still true for quasi-Banach spaces (see [12, Theorem. 1.6]), the inclusion is continuous, whence there exists a constant \(D_{E}>0\) such that

$$ \Vert x \Vert _{L_{\infty}}\leq D_{E} \Vert x \Vert _{E} $$
(4.1)

for each \(x\in E\). Defining

$$ a_{E}=\inf \bigl\{ \Vert \chi _{A} \Vert _{E}:\chi _{A} \in E,\mu (A )>0 \bigr\} , $$
(4.2)

by (4.1), we have \(a_{E}\geq 1/D_{E}>0\).

Definition 4.5

A function \(\varphi :[0,\infty )\rightarrow [0,\infty ]\) is called an Orlicz function if φ is nondecreasing, vanishing, and right continuous at 0, continuous on \((0,b_{\varphi})\), where

$$ b_{\varphi }=\sup \bigl\{ u\geq 0:\varphi ( u )< \infty \bigr\} $$

and left continuous at \(b_{\varphi}\). In the whole paper, excluding Remark 4.14 and Example 4.15\((iii)\), we will assume that \(\lim_{u\rightarrow \infty}\varphi (u)=\infty \).

Let

$$ a_{\varphi}=\sup \bigl\{ u\geq 0:\varphi (u ) =0 \bigr\} . $$

By \(\varphi ^{-1}\) we denote the generalized inverse of the function φ defined by

$$ \varphi ^{-1}(v)=\inf \bigl\{ u\geq 0\colon \varphi (u)>v\bigr\} \quad \text{for }v\in [0, \infty )\text{ and }\varphi ^{-1}( \infty )=\lim_{v\rightarrow \infty} \varphi ^{-1}(v) $$

(see [20]).

The following lemma is an easy exercise (see [20, Lemma 3.1], for a little different convex case).

Lemma 4.6

For any Orlicz function φ we have:

  1. (i)

    Let \(u\in [0,b_{\varphi})\). Then, \(\varphi ^{-1}(\varphi (u))>u\) if φ is constant on the interval \([u,u+\delta ) \) for some \(\delta >0\) and \(\varphi ^{-1}(\varphi (u))=u\) otherwise.

  2. (ii)

    If \(b_{\varphi}<\infty \), then \(\varphi ^{-1}(\varphi (b_{\varphi}))=b_{\varphi}\) and \(\varphi ^{-1}(\varphi (u))=b_{\varphi}< u\) for \(b_{\varphi}< u<\infty \).

  3. (iii)

    If either \(b_{\varphi}=\infty \) or \(b_{\varphi}<\infty \) with \(\varphi (b_{\varphi} )=\infty \), then \(\varphi (\varphi ^{-1}(u))=u\) for any \(u\in [0,\infty )\).

  4. (iv)

    If \(b_{\varphi}<\infty \) and \(\varphi (b_{\varphi} )<\infty \), then \(\varphi (\varphi ^{-1}(u))=u\) for \(u\in [0,\varphi (b_{\varphi})]\) and \(\varphi (\varphi ^{-1}(u))=\varphi (b_{\varphi})< u\) for \(u>\varphi (b_{\varphi})\).

From \((i)\) and \((ii)\), in particular, we obtain:

  1. (v)

    \(\varphi ^{-1}(\varphi (u))=u\) for \(a_{\varphi}\leq u< b_{\varphi}\) if either \(b_{\varphi}=\infty \) or \(b_{\varphi}<\infty \) with \(\varphi (b_{\varphi} )=\infty \) and φ is strictly increasing on \([a_{\varphi },b_{\varphi})\).

  2. (vi)

    \(\varphi ^{-1}(\varphi (u))=u\) for \(a_{\varphi}\leq u\leq b_{\varphi}\) if \(b_{\varphi}<\infty \) and \(\varphi (b_{\varphi} )<\infty \) and φ is strictly increasing on \([a_{\varphi},b_{\varphi}]\).

Finally, note that from \((iii)\) and \((iv)\) and \((i)\) and \((ii)\) we obtain

  1. (vii)

    \(\varphi (\varphi ^{-1}(u))\leq u\) for all \(u\in [0,\infty )\) and \(u\leq \varphi ^{-1}(\varphi (u))\) if \(\varphi (u)<\infty \).

Recall that for any Orlicz function φ the lower Matuszewska–Orlicz index \(\alpha _{\varphi}\) for all arguments is defined by the formula

$$\begin{aligned} \alpha ^{a}_{\varphi} =&\sup \bigl\{ p\in \mathbb{R}:\text{there exists }K \geq 1\text{ such that }\varphi (au)\leq Ka^{p} \varphi (u) \\ &\text{for any }u\geq 0\text{ and }0< a\leq 1\bigr\} . \end{aligned}$$

Analogously, the lower Matuszewska–Orlicz indices for large and for small arguments are defined as

$$\begin{aligned} \alpha ^{\infty}_{\varphi} =&\sup \bigl\{ p\in \mathbb{R}: \text{there exist }K\geq 1\text{ and }u_{0}>0\text{ such that } \varphi (u_{0})< \infty \text{ and } \\ &\varphi (au)\leq Ka^{p}\varphi (u)\text{ for any }u\geq u_{0} \text{ and }0< a\leq 1\bigr\} \end{aligned}$$

and

$$\begin{aligned} \alpha ^{0}_{\varphi} =&\sup \bigl\{ p\in \mathbb{R}:\text{there exist }K \geq 1\text{ and }u_{0}>0\text{ such that } \varphi (au)\leq Ka^{p} \varphi (u) \\ &\text{for any }0\leq u\leq u_{0}\text{ and }0< a \leq 1\bigr\} , \end{aligned}$$

respectively.

Remark 4.7

\(( i )\) If \(0< a_{\varphi }< b_{\varphi }\), then \(\alpha _{\varphi}^{0}=\infty \) and we may extend the key inequality in the definition of \(\alpha _{\varphi }^{0}\) to any \(u_{0}>a_{\varphi}\) such that \(\varphi (u_{0})<\infty \). Indeed, let \(p>0\). For every \(0\leq u\leq a_{\varphi }\) we have \(\varphi (au )=0=Ka^{p}\varphi (u )\) for each \(K>0\) and \(0< a\leq 1\). Take any \(u_{0}>a_{\varphi}\) such that \(\varphi (u_{0})<\infty \). If \(a_{\varphi }< u\leq u_{0}\) and \(0< a<\frac{a_{\varphi }}{u_{0}}\), then we have

$$ \varphi ( au ) \leq \varphi ( au_{0} ) \leq \varphi ( a_{\varphi } ) =0\leq Ka^{p}\varphi ( u ) $$

for every \(K>0\). Moreover,

$$ \sup_{a_{\varphi }< u\leq u_{0}}\sup_{\frac{a_{\varphi }}{u_{0}} \leq a\leq 1} \frac{\varphi ( au ) }{a^{p}\varphi ( u ) } \leq \sup_{ \frac{a_{\varphi }}{u_{0}}\leq a\leq 1}\frac{1}{a^{p}}= \biggl( \frac{u_{0}}{a_{\varphi }} \biggr) ^{p}. $$

Thus, for \(K= ( \frac{u_{0}}{a_{\varphi }} ) ^{p}\) we have

$$ \varphi ( au ) \leq Ka^{p}\varphi ( u ) $$

for all \(0< a\leq 1\) and \(0\leq u\leq u_{0}\).

\(( ii )\) If \(a_{\varphi }=0\) and \(\alpha _{\varphi }^{0}>0\) then we may extend the key inequality in the definition of \(\alpha _{\varphi }^{0}\) to any \(u_{1}\) such that \(\varphi ( u_{1} ) <\infty \). Indeed, suppose there is \(p>0\), \(u_{0}>0\) and \(K>0\) such that

$$ \varphi ( au ) \leq Ka^{p}\varphi ( u ) $$
(4.3)

for all \(0< a\leq 1\) and \(0\leq u\leq u_{0}\). Take \(u_{1}>u_{0}\) satisfying \(\varphi ( u_{1} ) <\infty \). Then,

$$ \sup_{u_{0}< u\leq u_{1}}\sup_{\frac{u_{0}}{u_{1}}\leq a\leq 1} \frac{\varphi ( au ) }{a^{p}\varphi ( u ) } \leq \sup_{\frac{u_{0}}{u_{1}}\leq a\leq 1}\frac{1}{a^{p}}= \biggl( \frac{u_{1}}{u_{0}} \biggr) ^{p}. $$

Set \(K_{1}= ( \frac{u_{1}}{u_{0}} ) ^{p}\). Now, we claim that

$$ K_{2}:=\sup_{u_{0}< u\leq u_{1}}\sup_{0< a< \frac{u_{0}}{u_{1}}} \frac{\varphi (au ) }{a^{p}\varphi (u )}< \infty . $$

Otherwise, for each \(n\in \mathbb{N}\) we can find \(u_{0}< u_{n}\leq u_{1}\) and \(0< a_{n}<\frac{u_{0}}{u_{1}}\) such that

$$ \varphi ( a_{n}u_{n} ) >na_{n}^{p} \varphi ( u_{n} ). $$

Denote \(b_{n}:=\frac{a_{n}u_{n}}{u_{0}}\). Then, \(b_{n}<1\) and

$$ \varphi ( b_{n}u_{0} ) =\varphi ( a_{n}u_{n} )>na_{n}^{p} \varphi ( u_{n} ) =nb_{n}^{p} \biggl( \frac{u_{0}}{u_{n}} \biggr) ^{p}\varphi ( u_{n} )\geq nb_{n}^{p} \biggl( \frac{u_{0}}{u_{1}} \biggr)^{p}\varphi (u_{0} ). $$

On the other hand, by inequality (4.3),

$$ \varphi ( b_{n}u_{0} ) \leq Kb_{n}^{p} \varphi ( u_{0} ), $$

which gives a contradiction and proves the claim. Finally, setting \(K_{3}=\max \{ K,K_{1},K_{2} \} \) we conclude that

$$ \varphi ( au ) \leq K_{3}a^{p}\varphi ( u ) $$

for all \(0< a\leq 1\) and \(0\leq u\leq u_{1}\).

\((iii )\) If \(\alpha _{\varphi}^{\infty }>0\) then we may extend the key inequality in the definition of \(\alpha _{\varphi }^{\infty }\) to any \(u_{1}>a_{\varphi }\). Indeed, suppose there is \(p>0\), \(u_{0}>0\), and \(K>0\) such that \(\varphi (u_{0})<\infty \) and

$$ \varphi ( au ) \leq Ka^{p}\varphi ( u ) $$

for all \(0< a\leq 1\) and \(u\geq u_{0}\). Take \(u_{1}< u_{0}\) satisfying \(\varphi (u_{1} )>0\). Then,

$$ \sup_{u_{1}\leq u< u_{0}}\sup_{0< a\leq 1} \frac{\varphi ( au ) }{a^{p}\varphi ( u ) } \leq \sup_{0< a\leq 1} \frac{\varphi (au_{0} ) }{a^{p}\varphi (u_{1} ) } \leq \sup_{0< a\leq 1} \frac{Ka^{p}\varphi (u_{0} ) }{a^{p}\varphi ( u_{1} ) }= \frac{K\varphi (u_{0} )}{\varphi (u_{1} )}. $$

Taking \(K_{1}=\max \{K, \frac{K\varphi (u_{0} )}{\varphi ( u_{1} ) } \} = \frac{K\varphi ( u_{0} ) }{\varphi ( u_{1} ) }\) we obtain that

$$ \varphi ( au ) \leq K_{1}a^{p}\varphi ( u ) $$

for all \(0< a\leq 1\) and \(u\geq u_{1}\).

\((iv ) \) From the above consideration, we conclude immediately that if \(\alpha _{\varphi }^{\infty }>0\) and \(\alpha _{\varphi }^{0}>0\) then \(\alpha _{\varphi }^{a}>0\).

Example 4.8

Taking \(\varphi _{1} (u )=\ln (1+u )\), for \(u\geq 0\), we easily obtain

$$ \lim_{u\rightarrow \infty} \frac{\varphi _{1} (au )}{\varphi _{1} (u )}=1 $$

for any \(a\in (0,1)\), whence \(\alpha ^{\infty}_{\varphi _{1}}=0\). Analogously, defining \(\varphi _{2}(0)=0\) and

$$ \varphi _{2}(u)=\frac{1}{\ln (1+\frac{1}{u} )} \quad \text{for } u>0, $$

we obtain

$$ \lim_{u\rightarrow 0^{+}} \frac{\varphi _{2} (au )}{\varphi _{2} (u )}=1 $$

for any \(a\in (0,1)\) and, in consequence, \(\alpha ^{0}_{\varphi _{2}}=0\).

For any pair E and φ we define the lower Matuszewska–Orlicz index \(\alpha ^{E}_{\varphi}\), by the formula

$$ \alpha ^{E}_{\varphi}:=\textstyle\begin{cases} \alpha ^{a}_{\varphi}, & \text{when neither }L_{\infty}\subset E \text{ nor }E\subset L_{\infty }, \\ \alpha ^{\infty}_{\varphi}, & \text{when }L_{\infty}\subset E, \\ \alpha ^{0}_{\varphi}, & \text{when }E\subset L_{\infty}. \end{cases} $$

Given a quasi-Banach ideal space E and an Orlicz function φ, we define on \(L^{0}\) a functional \(\rho _{\varphi }^{E}\), by

$$ \rho _{\varphi}^{E}(x):=\textstyle\begin{cases} \Vert \varphi ( \vert x \vert ) \Vert _{E} & \text{if }\varphi ( \vert x \vert ) \in E,\text{ } \vspace{1mm} \\ \infty & \text{otherwise.} \end{cases} $$

It is well known that if φ is a convex Orlicz function and E is a Banach ideal space then \(\rho _{\varphi}^{E}\) is a convex modular. The result below concerns the more general case (the index \(\alpha _{\varphi}^{E}\) plays a role of substitution of convexity of φ).

Theorem 4.9

Let E be a quasi-Banach ideal space and φ be an Orlicz function. If \(\alpha ^{E}_{\varphi}>0\), then \(\rho _{\varphi }^{E}\) is quasimodular (see Definition 3.1).

Proof

Obviously, \(\rho _{\varphi}^{E}(0)=0\) and \(\rho _{\varphi}^{E}(-x)=\rho _{\varphi}^{E}(x)\) for any \(x\in L^{0}\). Let \(x\neq 0\). Then, there exist \(A\in \Sigma \) with \(\mu (A)>0\) and \(n\in \mathbb{N}\) such that \(\frac{1}{n}\chi _{A}\leq |x|\). If \(\varphi (|x|)\notin E\), then \(\rho _{\varphi }^{E}(x)=\infty >1\), while, if \(\varphi (|x|)\in E\), by \(\varphi (\frac{1}{n})\chi _{A}\leq \varphi (|x|)\), we obtain \(\chi _{A}\in E\). Since \(\lim_{u\rightarrow \infty}\varphi (u)=\infty \), we can find \(\lambda _{A}>0\) such that \(\varphi (\lambda _{A}/n)>1/\|\chi _{A}\|_{E}\) and, in consequence, \(\rho _{\varphi }^{E}(\lambda _{A}x)=\|\varphi (\lambda _{A}|x|)\|_{E} \geq \|\varphi (\lambda _{A}/n)\chi _{A}\|_{E}>1\).

For every \(x\in L^{0}\) and all \(0\leq \lambda _{1}\leq \lambda _{2}\) we have \(\varphi (\lambda _{1}|x(t)|)\leq \varphi (\lambda _{2}|x(t)|)\) for μ-a.e. \(t\in T\), whence \(\rho _{\varphi}^{E}(\lambda _{1}x)\leq \rho _{\varphi}^{E}(\lambda _{2}x)\).

Let now \(x,y\in L^{0}\) and \(\alpha ,\beta \geq 0\), \(\alpha +\beta =1\). Then,

$$ \varphi \bigl( \bigl\vert \alpha x(t)+\beta y(t) \bigr\vert \bigr)\leq \varphi \bigl(\max \bigl( \bigl\vert x(t) \bigr\vert , \bigl\vert y(t) \bigr\vert \bigr)\bigr) \leq \varphi \bigl( \bigl\vert x(t) \bigr\vert \bigr)+ \varphi \bigl( \bigl\vert y(t) \bigr\vert \bigr) $$

for μ-a.e. \(t\in T\). Hence, if \(\varphi (|x|)\in E\) and \(\varphi (|y|)\in E\), we obtain

$$\begin{aligned} \rho _{\varphi}^{E} (\alpha x+\beta y ) =& \bigl\Vert \varphi \bigl( \vert \alpha x+\beta y \vert \bigr) \bigr\Vert _{E}\leq \bigl\Vert \varphi \bigl( \vert x \vert \bigr) + \varphi \bigl( \vert y \vert \bigr) \bigr\Vert _{E} \\ \leq & C_{E} \bigl( \bigl\Vert \varphi \bigl( \vert x \vert \bigr) \bigr\Vert _{E}+ \bigl\Vert \varphi \bigl( \vert y \vert \bigr) \bigr\Vert _{E} \bigr)=C_{E} \bigl(\rho _{ \varphi}^{E} (x )+\rho _{\varphi}^{E} (y ) \bigr). \end{aligned}$$

Obviously, the inequality \(\rho _{\varphi}^{E} (\alpha x+\beta y )\leq C_{E} ( \rho _{\varphi}^{E} (x )+\rho _{\varphi}^{E} (y ) )\) holds true, when \(\varphi (|x|)\notin E\) or \(\varphi (|y|)\notin E\).

Finally, we will prove that \(\rho _{\varphi}^{E}\) satisfies condition \((v)\). Without loss of generality, we can suppose that \(0<\rho _{\varphi}^{E}(x)<\infty \) (then, in particular, we have \(a_{\varphi}< b_{\varphi}\)). We will consider three cases.

If \(L_{\infty}\subset E\), then for any \(\varepsilon >0\) there exists \(u_{1}\in (a_{\varphi},b_{\varphi})\) such that \(C_{E}\cdot \varphi (u_{1})\|\chi _{T}\|_{E}<\varepsilon \). Since \(\alpha ^{E}_{\varphi}=\alpha ^{\infty}_{\varphi}>0\), by Remark 4.7, we obtain that there exist numbers \(p>0\) and \(K=K(\varepsilon )\geq 1\) such that \(\varphi (au)\leq Ka^{p}\varphi (u)\) for all \(u\geq u_{1}\) and \(0< a\leq 1\). Defining \(B=\{t\in T\colon |x(t)|\geq u_{1}\}\), for any \(a\in (0,1]\) we obtain

$$\begin{aligned} \rho _{\varphi}^{E}(ax) =& \bigl\Vert \varphi \bigl(a \vert x \vert \bigr) \bigr\Vert _{E}\leq C_{E} \bigl( \bigl\Vert \varphi \bigl(a \vert x \vert \bigr) \chi _{B} \bigr\Vert _{E}+ \bigl\Vert \varphi \bigl(a \vert x \vert \bigr)\chi _{T\backslash B} \bigr\Vert _{E} \bigr) \\ \leq & C_{E}Ka^{p} \bigl\Vert \varphi \bigl( \vert x \vert \bigr) \bigr\Vert _{E}+C_{E}\varphi (u_{1}) \Vert \chi _{T} \Vert _{E}\leq C_{E}Ka^{p}\rho _{\varphi}^{E}(x)+ \varepsilon . \end{aligned}$$
(4.4)

In the case when neither \(L_{\infty}\subset E\) nor \(E\subset L_{\infty}\), analogously as above, we obtain

$$ \rho _{\varphi}^{E}(ax)= \bigl\Vert \varphi \bigl(a \vert x \vert \bigr) \bigr\Vert _{E}\leq Ka^{p} \bigl\Vert \varphi \bigl( \vert x \vert \bigr) \bigr\Vert _{E}=Ka^{p} \rho _{\varphi}^{E}(x). $$
(4.5)

Let now \(E\subset L_{\infty}\), take \(A>0\) and assume that \(\rho _{\varphi}^{E}(x)=\|\varphi (|x|)\|_{E}\leq A\). By the closed-graph theorem, there exists a constant \(D_{E}>0\) such that \(\|\varphi (|x|)\|_{L^{\infty}}\leq D_{E}\|\varphi (|x|)\|_{E}\leq AD_{E}\). Thus, \(\varphi (|x(t)|)\leq \min (AD_{E},\varphi (b_{\varphi}))\) for μ-a.e. \(t\in T\), whence, by Lemma 4.6, \(|x(t)|\leq u_{2}\) for the same t, where \(u_{2}=\varphi ^{-1}(\min (AD_{E},\varphi (b_{\varphi})))\). Simultaneously, by \(\alpha ^{E}_{\varphi}=\alpha ^{0}_{\varphi}>0\) and Remark 4.7, we obtain that there exist \(p>0\) and \(K=K(A)\geq 1\) such that \(\varphi (au)\leq Ka^{p}\varphi (u)\) for any \(u\leq u_{2}\) and \(0< a\leq 1\). Therefore, \(\rho _{\varphi}^{E}(ax)\leq Ka^{p}\rho _{\varphi}^{E}(x)\). □

Now, we will show that, depending on the embeddings between E and \(L_{\infty}\), the condition \(\alpha _{\varphi}^{E}>0\) is even equivalent to a certain condition that is close to the point \((v)\) of Definition 3.1.

Theorem 4.10

\((i)\) Assume that neither \(L_{\infty}\subset E\) nor \(E\subset L_{\infty}\). Then, \(\alpha ^{a}_{\varphi}>0\) if and only if there exist constants \(p>0\) and \(K\geq 1\) such that for all \(x\in L^{0}\) and \(0< a\leq 1\) we have \(\rho _{\varphi}^{E} (ax )\leq Ka^{p}\rho _{\varphi}^{E} (x )\).

\((ii)\) Let \(L_{\infty}\subset E\). Then, \(\alpha ^{\infty}_{\varphi}>0\) if and only if there is a constant \(p>0\) such that for all \(v_{0}>0\) there exists \(K=K(v_{0})\geq 1\) such that for each \(x\in L^{0}\) satisfying \(|x(t)|\geq v_{0}\) for μ-a.e. \(t \in T\) and any \(0< a\leq 1\) we have \(\rho _{\varphi}^{E} (ax )\leq Ka^{p}\rho _{\varphi}^{E} (x )\).

\((iii)\) Let \(E\subset L_{\infty}\). Then, \(\alpha ^{0}_{\varphi}>0\) if and only if there is a constant \(p>0\) such that for every \(A>0\) there exists \(K=K(A)\geq 1\) such that for any \(x\in L^{0}\) satisfying \(\rho _{\varphi}^{E}(x)\leq A\) and every \(0< a\leq 1\) we have \(\rho _{\varphi}^{E} (ax )\leq Ka^{p}\rho _{\varphi}^{E} (x )\).

Proof

The necessity of statements \((i)\) and \((iii)\) follows from the proof of Theorem 4.9. Now, we will show the sufficiency of \((iii)\). Let \(D\in \Sigma \) be such that \(\mu (D)>0\) and \(\chi _{D}\in E\). Take \(u_{0}>0\) satisfying \(0<\varphi (u_{0})<\infty \) and define \(A:=\varphi (u_{0})\|\chi _{D}\|_{E}\). Then, for any \(u\leq u_{0}\) and any \(a\in (0,1]\) we obtain

$$\begin{aligned} \varphi (au) \Vert \chi _{D} \Vert _{E} =& \bigl\Vert \varphi (au)\chi _{D} \bigr\Vert _{E}=\rho _{ \varphi}^{E} (au\chi _{D} ) \\ \leq & K(A)a^{p}\rho _{\varphi}^{E} (u\chi _{D} )=K(A)a^{p} \bigl\Vert \varphi (u)\chi _{D} \bigr\Vert _{E}=K(A)a^{p}\varphi (u) \Vert \chi _{D} \Vert _{E}. \end{aligned}$$

Hence, by the arbitrariness of u and a, we obtain \(\alpha ^{0}_{\varphi}>0\). Analogously, we can prove the sufficiency of the condition \(\alpha ^{a}_{\varphi}>0\) in \((i)\).

Now, we will prove statement \((ii)\), that is, let \(L_{\infty}\subset E\). First, let us note that, by the proof of Theorem 4.9, we obtain the implication: if \(\alpha ^{\infty}_{\varphi}>0\), then there is a constant \(p>0\) such that for each \(\varepsilon >0\) there exists \(K=K(\varepsilon )\geq 1\) such that \(\rho _{\varphi}^{E} (ax )\leq Ka^{p}\rho _{\varphi}^{E} ( x )+\varepsilon \) for any \(x\in L^{0}\) and \(0< a\leq 1\).

Let \(\alpha ^{\infty}_{\varphi}>0\) and take \(p\in (0,\alpha ^{\infty}_{\varphi})\), \(v_{0}>0\) and \(x\in L^{0}\) such that \(\rho _{\varphi}^{E}(x)<\infty \) and \(|x(t)|\geq v_{0}\) for μ-a.e. \(t\in T\). By Remark 4.7 (especially \((iii)\) and \((iv)\)), there exists \(K=K(v_{0})\) such that

$$ \varphi \bigl(a \bigl\vert x(t) \bigr\vert \bigr)\leq Ka^{p} \varphi \bigl( \bigl\vert x(t) \bigr\vert \bigr) $$

for all \(a\in (0,1]\) and μ-a.e. \(t\in T\), whence \(\rho _{\varphi}^{E} (ax )\leq Ka^{p}\rho _{\varphi}^{E} (x )\).

Finally, we will prove the opposite implication. Take \(u_{0}>0\) satisfying \(0<\varphi (u_{0})<\infty \) and define \(v_{0}=u_{0}\). Then, for any \(u\geq u_{0}\) and any \(a\in (0,1]\) we have

$$ \varphi (au) \Vert \chi _{T} \Vert _{E}=\rho _{\varphi}^{E} (au\chi _{T} )\leq K(v_{0})a^{p}\rho _{\varphi}^{E} (u \chi _{T} )=K(v_{0})a^{p} \varphi (u) \Vert \chi _{T} \Vert _{E} $$

and, in consequence, \(\alpha ^{\infty}_{\varphi}>0\). □

Definition 4.11

Let a quasi-Banach ideal space E and an Orlicz function φ be such that \(\alpha _{\varphi}^{E}>0\). Then, the Calderón–Lozanovskiĭ space \(E_{\varphi}\) is defined by

$$ E_{\varphi}=\Bigl\{ x\in L^{0}:\lim_{\lambda \rightarrow 0} \rho _{\varphi}^{E}( \lambda x)=0\Bigr\} . $$

By Theorem 4.9 and Lemma 3.3, \(E_{\varphi}\) is a quasimodular space and

$$ E_{\varphi}=\bigl\{ x\in L^{0}:\rho _{\varphi }^{E}( \lambda x)< \infty \text{ for some }\lambda >0\bigr\} . $$

Moreover, by Theorem 3.4, the functional

$$ \Vert x \Vert _{\varphi }=\inf \bigl\{ \lambda >0:\rho _{\varphi}^{E} (x/\lambda )\leq 1 \bigr\} , $$

is a quasinorm, called a Luxemburg–Nakano quasinorm. It is easy to show that \(E_{\varphi}=(E_{\varphi},\leq ,\|\cdot \|_{\varphi})\) is a quasinormed ideal space.

Remark 4.12

It is known that there is a connection between the Banach ideal space \(E_{\varphi}\) (where φ is a convex Orlicz function and E is a Banach ideal space) and the normed Calderón–Lozanovskii interpolation construction \(\psi ( E,L_{\infty} )\) (where ψ is a homogeneous, concave function on \(\mathbb{R}_{+}^{2}\)) – see [24, Example 2, p. 178]. However, there is a similar relation if φ is a nonconvex Orlicz function, E is a quasi-Banach ideal space, and ψ is positively homogeneous, nondecreasing with respect to each variable. On the other hand, such an approach leads to quasi-Banach lattices \(\psi (E_{0},E_{1} )\) that have application in interpolation theory (see [27]).

We also obtain the following:

Lemma 4.13

For any quasi-Banach ideal space E and any Orlicz function φ the following assertions hold:

  1. (i)

    For each \(x\in E_{\varphi}\) the function \(f_{x} (\alpha ):=\rho _{\varphi}^{E}(\alpha x)\), for \(\alpha >0\), is nondecreasing and left continuous.

  2. (ii)

    For any \(x\in E_{\varphi }\) we have \(\|x\|_{\varphi}\leq 1\) if and only if \(\rho _{\varphi}^{E}(x)\leq 1\).

  3. (iii)

    The quasinormed ideal space \(E_{\varphi}\) has the Fatou property and, in consequence, \(E_{\varphi}\) is complete.

Proof

The assertion \((i)\) follows immediately from the properties of φ and E (recall that E has the Fatou property). Next, by \((i)\) and Lemma 3.5, we obtain \((ii)\). Proceeding analogously as in [6, Theorem 12] (see also [14, Lemma 2.2\((ii)\)]), we obtain that \(E_{\varphi}\) has the Fatou property. Hence, by Lemma 4.2, \(E_{\varphi}\) is complete. □

Remark 4.14

Now, we will show the naturalness of the assumption \(\lim_{u\rightarrow \infty}\varphi (u)=\infty \) in the definition of Orlicz function (see Definition 4.5). Obviously, if \(\alpha ^{a}_{\varphi}>0\) or \(\alpha ^{\infty}_{\varphi}>0\), then \(\lim_{u\rightarrow \infty}\varphi (u)=\infty \). Simultaneously, for \(\varphi (u)=\min (u^{2},1)\), we have \(\alpha ^{0}_{\varphi}>0\) and \(\lim_{u\rightarrow \infty}\varphi (u)=1\). Let \(E\subset L_{\infty}\), \(\lim_{u\rightarrow \infty}\varphi (u)<\infty \) and \(\alpha ^{E}_{\varphi}=\alpha ^{0}_{\varphi}>0\). Then, condition (i) of Definition 3.1 holds whenever \(\lim_{u\rightarrow \infty}\varphi (u)>1/a_{E}\), where \(a_{E}\) is defined by formula (4.2), and, in consequence, \(\rho _{\varphi}^{E}\) is quasimodular. Moreover, defining the new Orlicz function ψ, by \(\psi (u)=\varphi (u)\) for \(u\in [0,\varphi ^{-1}(1/a_{E})]\) and \(\psi (u)=u-(\varphi ^{-1}(1/a_{E})-1/a_{E})\) for \(u>\varphi ^{-1}(1/a_{E})\), we obtain \(\lim_{u\rightarrow \infty}\psi (u)=\infty \) and \((E_{\varphi},\|\cdot \|_{\varphi})\equiv (E_{\psi},\|\cdot \|_{\psi})\).

Now, we will show (among others) that the notion of modular and quasimodular are incomparable.

Example 4.15

Assume \(T=[0,\infty )\) and μ is the Lebesgue measure.

\((i)\) If \(E=L_{1}\) and \(\varphi (u)=u^{p}\) for \(u\geq 0\), \(p>0\), then \(\rho _{\varphi}^{E}\) is a quasimodular (a convex modular for \(p>1\)) as well as a modular (see [26]). We have

$$ \Vert x \Vert _{\varphi}= \Vert x \Vert _{p}= \biggl( \int _{0}^{\infty} \bigl\vert x(t) \bigr\vert ^{p}\,dt \biggr)^{\frac{1}{p}}. $$

Simultaneously, the F-norm is given by

$$ \big\vert \Vert x \Vert \big\vert _{\varphi}:=\inf \bigl\{ \lambda >0\colon \rho _{\varphi}^{E} (x/\lambda )\leq \lambda \bigr\} = \biggl( \int _{0}^{ \infty} \bigl\vert x(t) \bigr\vert ^{p}\,dt \biggr)^{\frac{1}{1+p}} $$

(see [26]). Note also that \(|\Vert x_{n}\Vert |_{\varphi}\rightarrow 0\) if and only if \(\|x\|_{\varphi}\rightarrow 0\), but there do not exist constants \(A,B>0\) such that \(A \|x\|_{\varphi} \leq |\Vert x\Vert |_{\varphi} \leq B \|x\|_{\varphi}\) for all \(x \in L_{\varphi}\).

\((ii)\) Let \(E=L_{(1/4)}\) and \(\varphi (u)=u^{2}\) for \(u\geq 0\). Obviously, \(\rho _{\varphi}^{E}\) is a quasimodular. Simultaneously, for \(x=\chi _{[0,1)}\) and \(y=\chi _{[1,2)}\) we obtain

$$ \rho _{\varphi}^{E} \biggl(\frac{1}{2}x+ \frac{1}{2}y \biggr)=4>2=\rho _{ \varphi}^{E}(x)+\rho _{\varphi}^{E}(y). $$

Thus, \(\rho _{\varphi}^{E}\) is not a modular.

\((iii)\) If \(E=L_{1}\) and \(\varphi (u)=\arctan (u)\) for \(u\geq 0\) (see Definition 4.5), then \(\rho _{\varphi}^{E}\) is a modular (see [26]). Now, we will show that \(\rho _{\varphi}^{E}\) is not a quasimodular, more precisely, \(\rho _{\varphi}^{E}\) does not satisfy condition \((v)\) of Definition 3.1. Let \(p>0\) and take \(A=\pi /2\) and \(\varepsilon =\pi /8\). Then, for \(x_{n}=n\chi _{[0,1)}\) and \(a_{n}=1/n\), we obtain \(\rho _{\varphi}^{E}(x_{n})\leq \pi /2\), \(\rho _{\varphi}^{E}(a_{n}x_{n})=\pi /4\), and \(\lim_{n\rightarrow \infty}(a_{n})^{p}=0\). In consequence, for any \(K\geq 1\) there exists \(n\in \mathbb{N}\) such that \(\rho _{\varphi}^{E}(a_{n}x_{n})>Ka_{n}^{p}\rho _{\varphi}^{E}(x_{n})+ \pi /8\).

\((iv)\) Let \(E=L_{1}\), \(\varphi (u)=u\) for \(u\in [0,1]\) and \(\varphi (u)=\max (1,u-1)\) for \(u>1\). For \(x=\chi _{[0,1)}\) we have \(\rho _{\varphi}^{E}(x)=\rho _{\varphi}^{E}(2x)=1\). Simultaneously, \(\rho _{\varphi}^{E}(x/\lambda )>1\) for \(\lambda <1/2\), so \(\|x\|_{\varphi}=\frac{1}{2}\) (see Remark 3.6).

Recall the notion of uniform monotonicity (see, for example, [3, 9, 22]) that plays an important role in the theory of Banach lattices. A quasi-Banach lattice \((E, \Vert \cdot \Vert _{E} )\) is said to be uniformly monotone provided for each \(\varepsilon >0\) there exists \(\delta =\delta (\varepsilon )>0\) such that for all \(x,y\in E_{+}\) with \(\Vert x \Vert _{E}=1\) we have \(\Vert x+y \Vert _{E}\geq 1+\delta \) whenever \(\Vert y \Vert _{E}\geq \varepsilon \).

Lemma 4.16

If E is uniformly monotone, then for each \(\varepsilon _{1}>0\) and \(A>0\) there exists \(\delta _{1}=\delta _{1} (\varepsilon _{1},A ) =\delta (\frac{\varepsilon _{1}}{A} )>0\) (here the function \(\delta (\cdot ) \) comes from the definition of the uniform monotonicity) such that for all \(x,y\in E_{+}\) we have \(\Vert x+y \Vert _{E}\geq \Vert x \Vert _{E} (1+\delta _{1} )\) whenever \(\Vert y \Vert _{E}\geq \varepsilon _{1}\) and \(\Vert x \Vert _{E}\leq A\).

Proof

The proof can be found in [9] (the same for the quasinorm). However, we will need the precise form of the function \(\delta _{1} (\varepsilon _{1},A )\), so we present the proof for the reader’s convenience.

Let \(\varepsilon _{1}>0\) and \(A>0\). Take \(x,y\in E_{+}\) such that \(\Vert y \Vert _{E}\geq \varepsilon _{1}\) and \(\Vert x \Vert _{E}\leq A\). Denote

$$ \widetilde{x}=\frac{x}{ \Vert x \Vert _{E}}\quad \text{and} \quad \widetilde{y}= \frac{y}{ \Vert x \Vert _{E}}. $$

Then, \(\Vert \widetilde{x} \Vert _{E}=1\) and \(\Vert \widetilde{y} \Vert _{E}\geq \frac{\varepsilon _{1}}{A}\). By the uniform monotonicity of E, there exists \(\delta _{1}=\delta _{1} (\varepsilon _{1},A )=\delta (\frac{\varepsilon _{1}}{A} )>0\) such that \(\Vert \widetilde{x}+\widetilde{y} \Vert _{E}\geq 1+ \delta _{1}\), whence

$$ \Vert x+y \Vert _{E}\geq \Vert x \Vert _{E} (1+ \delta _{1} ). $$

 □

Recall that the assumption \(\alpha _{\varphi }^{E}>0\) is important to show that \(\rho _{\varphi }^{E}\) is a quasimodular (see Theorem 4.9 and also Theorem 4.10) and, consequently, that \(\Vert \cdot \Vert _{\varphi}\) is a quasinorm. However, if we know nothing about the index \(\alpha _{\varphi }^{E}\) then we may still define a functional

$$ \rho _{\varphi }^{E}(x):=\textstyle\begin{cases} \Vert \varphi ( \vert x \vert ) \Vert _{E} & \text{if }\varphi ( \vert x \vert )\in E, \\ \infty & \text{otherwise} \end{cases} $$

and the set

$$ E_{\varphi}=\Bigl\{ x\in L^{0}:\lim_{\lambda \rightarrow 0} \rho _{\varphi}^{E}( \lambda x)=0\Bigr\} . $$

Then, it is easy to see that \(E_{\varphi }\) is a linear space and we may consider the functional

$$ \Vert x \Vert _{\varphi}=\inf \bigl\{ \lambda >0:\rho _{\varphi}^{E} (x/\lambda )\leq 1 \bigr\} ,\quad \text{for }x \in E_{\varphi}, $$

which satisfies the conditions \((i ) \) and \(( ii ) \) of the quasinorm definition. We are going to show that, under some natural assumptions, the condition \(\alpha _{\varphi }^{E}>0\) can be even necessary, that is to say, if the functional \(\Vert \cdot \Vert _{\varphi}\) is a quasinorm, then \(\alpha _{\varphi}^{E}>0\). Since the result below is only some illustration of how natural is the assumption that \(\alpha _{\varphi}^{E}>0\), we will limit ourselves only to one case of the ideal space E.

Theorem 4.17

Suppose φ is a finitely valued, strictly increasing Orlicz function. Let \((E, \Vert \cdot \Vert _{E} )\) be a uniformly monotone, p-normed ideal space over nonatomic measure space \((T,\Sigma ,\mu )\) for some \(0< p\leq 1\). Assume that neither \(L_{\infty }\subset E\) nor \(E\subset L_{\infty}\). If \((E_{\varphi},\Vert \cdot \Vert _{\varphi} )\) is a quasinormed space, then \(\alpha _{\varphi}^{a}>0\).

Proof

Denote by \(C\geq 1\) the constant from the quasitriangle inequality for \((E_{\varphi},\Vert \cdot \Vert _{\varphi} )\). Let \(\delta _{0}=\delta (1/2 )\) be the constant from the definition of the uniform monotonicity of E. Fix \(s>0\). Recall also that if E is uniformly monotone, then E is order continuous (see [22, Proposition 2.4]). Thus, by the proof of Theorem 2.4 in [14], the function ν, defined by \(\nu (A)=\|\chi _{A}\|_{E}\) for all \(A\in \Sigma \), \(\chi _{A}\in E\), is the submeasure in the sense of [5, Definition 1], whence by [5, Theorem 10], ν has the Darboux property. In consequence, we can find a set \(A\in \Sigma \), \(\chi _{A}\in E\) satisfying

$$ \varphi (s )= \frac{1}{ \Vert \chi _{A} \Vert _{E} (1+\delta _{0} )}. $$

Take a set \(B\in \Sigma \) of positive measure such that \(\chi _{B}\in E\), \(A\cap B=\emptyset \) and \(\Vert \chi _{B} \Vert _{E}=\frac{1}{2} \Vert \chi _{A} \Vert _{E}\). Applying Lemma 4.16 we conclude that

$$ \Vert \chi _{A\cup B} \Vert _{E}\geq \Vert \chi _{A} \Vert _{E} (1+\delta _{0} ). $$
(4.6)

It is well known that

$$ \Vert \chi _{A} \Vert _{\varphi}= \frac{1}{\varphi ^{-1} (\frac{1}{ \Vert \chi _{A} \Vert _{E}} )}, $$

where \(\varphi ^{-1}\) is the general right-inverse to φ. Indeed, \(\Vert \varphi (\frac{\chi _{A}}{\lambda} ) \Vert _{E}\leq 1\) if and only if \(\frac{1}{\lambda}\leq \varphi ^{-1} ( \frac{1}{ \Vert \chi _{A} \Vert _{E}} )\). In consequence,

$$ C\geq \frac{ \Vert \chi _{A}+\chi _{B} \Vert _{\varphi}}{ \Vert \chi _{A} \Vert _{\varphi}+ \Vert \chi _{B} \Vert _{\varphi}} \geq \frac{ \Vert \chi _{A}+\chi _{B} \Vert _{\varphi}}{2 \Vert \chi _{A} \Vert _{\varphi}}= \frac{ \Vert \chi _{A\cup B} \Vert _{\varphi}}{2 \Vert \chi _{A} \Vert _{\varphi}}= \frac{\varphi ^{-1} (\frac{1}{ \Vert \chi _{A} \Vert _{E}} )}{2\varphi ^{-1} (\frac{1}{ \Vert \chi _{A\cup B} \Vert _{E}} )}. $$

Moreover, also applying (4.6), we obtain

$$ 2C\varphi ^{-1} \biggl( \frac{1}{ \Vert \chi _{A} \Vert _{E} (1+\delta _{0} )} \biggr)\geq 2C\varphi ^{-1} \biggl( \frac{1}{ \Vert \chi _{A\cup B} \Vert _{E}} \biggr)\geq \varphi ^{-1} \biggl(\frac{1}{ \Vert \chi _{A} \Vert _{E}} \biggr) $$

and consequently we obtain

$$ \varphi \biggl[2C\varphi ^{-1} \biggl( \frac{1}{ \Vert \chi _{A} \Vert _{E} (1+\delta _{0} )} \biggr) \biggr]\geq \varphi \biggl[\varphi ^{-1} \biggl( \frac{1}{ \Vert \chi _{A} \Vert _{E}} \biggr) \biggr]= \frac{1}{ \Vert \chi _{A} \Vert _{E}}. $$

Taking \(C_{1}=2C\), we have

$$ (1+\delta _{0} )\varphi (s )\leq \varphi \bigl[2C\varphi ^{-1} \bigl(\varphi (s ) \bigr) \bigr]= \varphi [ 2Cs ]=\varphi [C_{1}s ] $$

for each \(s>0\). For every \(a\geq 1\) there is \(m\in \mathbb{N}\) such that \(C_{1}^{m-1}\leq a< C_{1}^{m}\). Fix \(p>0\) satisfying \(p=\frac{\ln (1+\delta _{0} )}{\ln C_{1}}\). Then, \((1+\delta _{0} )^{m-1}= (C_{1}^{m-1} )^{p}\) and

$$ \varphi (as )\geq \varphi \bigl(C_{1}^{m-1}s \bigr) \geq (1+\delta _{0} )^{m-1}\varphi (s )= \bigl(C_{1}^{m-1} \bigr)^{p}\varphi ( s )\geq \biggl( \frac{a}{C_{1}} \biggr)^{p}\varphi (s )=a^{p}C_{1}^{-p} \varphi (s ) $$

for each \(s>0\). Setting \(u:=as\) and \(b:=1/a\) we conclude that

$$ \varphi (bu )\leq b^{p}C_{1}^{p}\varphi (u ) $$

for any \(u>0\) and each \(b\in (0,1]\). This means that \(\alpha _{\varphi}^{a}>0\). □

5 The quasinormed Orlicz–Lorentz spaces and Orlicz spaces

Take \(I=[0,1]\) or \(I=[0,\infty )\) with the Lebesgue measure μ. Let \(\omega :I\rightarrow \mathbb{R_{+}}\) be a measurable function with \(\int _{0}^{t}\omega (s)\,ds<\infty \) for each \(t \in I\). We assume that there is a constant \(C>0\) such that \(\int _{0}^{2t}\omega (s)\,ds\leq C\int _{0}^{t}\omega (s)\,ds\) for each \(t\in \frac{1}{2}I\), which implies that the space

$$ \Lambda _{1,\omega}= \biggl\{ f\in L^{0}\colon \lVert f \rVert _{\omega}= \int _{I} f^{\ast}(s)\omega (s)\,ds< \infty \biggr\} , $$

where \(f^{\ast}\) is the nonincreasing rearrangement of f – see [2, 23], is the quasinormed ideal space with the Fatou property (see [13]) and it is called the Lorentz funtion space \(\Lambda _{1,\omega}\). The Lorentz sequence space \(\lambda _{1,\omega}\) over the counting measure space \((\mathbb{N},2^{\mathbb{N}},m )\) we define analogously (see [15]).

Note that,

\((1)\) neither \(L_{\infty }\subset \Lambda _{1,\omega}\) nor \(\Lambda _{1,\omega}\subset L_{\infty }\), whenever \(I=[0,\infty )\) and \(\int _{0}^{\infty}\omega (s)\,ds=\infty \);

\((2)\) \(L_{\infty}\subset \Lambda _{1,\omega}\),whenever \(I=[0,1]\) or (\(I=[0, \infty )\) and \(\int _{0}^{\infty}\omega (s)\,ds<\infty \));

\((3)\) \(\lambda _{1,\omega}\subset l_{\infty}\) (furthermore, \(\lambda _{1,\omega}=l_{\infty}\) provided \(\sum_{i=1}^{\infty}\omega (i)<\infty \)).

If E is a Lorentz function (sequence) space \(\Lambda _{1,w}\) (\(\lambda _{1,w}\)), then Calderón–Lozanovskiĭ space \(E_{\varphi }\) is the corresponding Orlicz–Lorentz function (sequence) space \(\Lambda _{\varphi ,w}\) (\(\lambda _{\varphi ,w}\)). If \(E=L_{1}\) (\(E=l_{1}\)), then the space \(E_{\varphi }\) becomes the Orlicz function (sequence) space \(L_{\varphi}\) (\(l_{\varphi}\)) (cf. [16]). On the other hand, if \(\varphi (u)=u^{p}\), \(1\leq p<\infty \) [\(0< p<1\)] then \(E_{\varphi }\) is the p-convexification [concavification] \(E^{(p)}\) of E with the quasinorm \(\Vert x\Vert _{E^{(p)}}=\Vert |x|^{p}\Vert _{E}^{1/p}\). If \(\varphi (u)=0\) for \(0\leq u\leq 1\) and \(\varphi (u)=\infty \) for \(u>1\), then \(E_{\varphi}=L_{\infty}\) (\(E_{\varphi}=l_{\infty}\)) with equality of the norms.

It is well known that Orlicz–Lorentz spaces \(\Lambda _{\varphi ,\omega}\) (in particular, the Lorentz spaces \(\Lambda _{p,\omega}\) or the Orlicz spaces \(L_{\varphi}\)) have been studied directly by many authors (see, for example, [1316] and the references therein).

Applying Theorems 3.4 and 4.9 with \(E=\Lambda _{1,w}\) or \(E=\lambda _{1,w}\) we conclude immediately:

Corollary 5.1

\((i )\) Let \(E=\Lambda _{1,w} (I,\Sigma ,\mu )\) be such that \(\mu (I)<\infty \) or \((\mu (I)=\infty \textit{ and }\int _{0}^{\infty}\omega (s)\,ds<\infty )\). If \(\alpha _{\varphi}^{\infty}>0\), then the functional \(\rho _{\varphi}^{\Lambda _{1,w}}(\cdot )\) is a quasimodular and the functional \(\Vert \cdot \Vert _{\varphi ,w}\) is a quasinorm (called a Luxemburg–Nakano quasinorm) in \(\Lambda _{\varphi ,w}\).

\((ii )\) Let \(E=\Lambda _{1,w} (I,\Sigma ,\mu )\) be such that \(\mu (I)= \infty \) and \(\int _{0}^{\infty}\omega (s)\,ds=\infty \). If \(\alpha _{\varphi}^{a}>0\), then the functional \(\rho _{\varphi}^{\Lambda _{1,w}}(\cdot )\) is a quasimodular and the functional \(\Vert \cdot \Vert _{\varphi ,w}\) is a quasinorm in \(\Lambda _{\varphi ,w}\).

\((iii )\) Let \(E=\lambda _{1,w}\) and \(\sum_{i=1}^{\infty}\omega (i)=\infty \). If \(\alpha _{\varphi}^{0}>0\), then the functional \(\rho _{\varphi}^{\lambda _{1,w}}(\cdot )\) is a quasimodular and the functional \(\Vert \cdot \Vert _{\varphi ,w}\) is a quasinorm in \(\lambda _{\varphi ,w}\).

Remark 5.2

The second conclusion in statement \((iii )\) has been proved directly in [15, Proposition 1.3], but the authors did not consider the quasimodular, only the quasinorm.

Applying the above corollary with \(\omega \equiv 1\) and Theorem 4.17 with \(E=L^{1}\) we obtain:

Corollary 5.3

\((i )\) Let \(E=L_{1} (I,\Sigma ,\mu ) \) with a finite measure μ. If \(\alpha _{\varphi}^{\infty}>0\), then the functional \(\rho _{\varphi}^{L_{1}}(\cdot )\) is a quasimodular and the functional \(\Vert \cdot \Vert _{\varphi}\) is a quasinorm (called a Luxemburg–Nakano quasinorm) in \(L_{\varphi}\).

\((ii )\) Let \(E=L_{1} (I,\Sigma ,\mu ) \) with an infinite measure μ. If \(\alpha _{\varphi}^{a}>0\), then the functional \(\rho _{\varphi}^{L_{1}}(\cdot )\) is a quasimodular and the functional \(\Vert \cdot \Vert _{\varphi}\) is a quasinorm in \(L_{\varphi}\).

\((iii )\) Let \(E=L_{1} (I,\Sigma ,\mu ) \) with an infinite measure μ. Assume that the function φ is finitely valued and strictly increasing. Then, the functional \(\Vert \cdot \Vert _{\varphi}\) is a quasinorm in \(L_{\varphi}\) if and only if \(\alpha _{\varphi}^{a}>0\).

\((iv )\) If \(E=l_{1}\) and \(\alpha _{\varphi}^{0}>0\), then the functional \(\rho _{\varphi}^{l_{1}}(\cdot )\) is a quasimodular and the functional \(\Vert \cdot \Vert _{\varphi}\) is a quasinorm in \(l_{\varphi}\).

Remark 5.4

The statement \((iii )\) has been proved directly in [16, Theorem 1.8].

6 Further research and open ends

6.1 Further research

It is well known that the relations between the modular and the norm play a crucial role in the metric geometry of normed Orlicz spaces (normed Calderón–Lozanovskiĭ spaces). The following basic results have many applications:

Lemma 6.1

Let E be a Banach ideal space and φ be a convex Orlicz function. Then,

\(( i ) \) if \(\rho _{\varphi }^{E}(x_{n})\rightarrow 1\), then \(\Vert x_{n}\Vert _{\varphi }\rightarrow 1\) for all sequences \(( x_{n} ) \) in \(E_{\varphi }\). In particular,

\(( ii ) \) if \(\rho _{\varphi }^{E}(x)=1\), then \(\Vert x\Vert _{\varphi }=1\) for every \(x\in E_{\varphi }\).

\(( iii ) \) if \(\Vert x_{n}\Vert _{\varphi }\rightarrow 0\), then \(\rho _{\varphi }^{E}(x_{n})\rightarrow 0\) for all sequences \(( x_{n} ) \) in \(E_{\varphi }\).

Suppose additionally that \(\varphi \in \Delta _{2}^{E}\). Then,

\(( iv ) \) if \(\Vert x_{n}\Vert _{\varphi }\rightarrow 1\), then \(\rho _{\varphi }^{E}(x_{n})\rightarrow 1\) for all sequences \(( x_{n} ) \) in \(E_{\varphi }\). In particular,

\(( v ) \) if \(\Vert x\Vert _{\varphi }=1\), then \(\rho _{\varphi }^{E}(x)=1\) for each \(x\in E_{\varphi }\).

Assume additionally that \(\varphi \in \Delta _{2}^{E}\) and \(a_{\varphi }=0\). Then,

\(( vi ) \) if \(\rho _{\varphi }^{E}(x_{n})\rightarrow 0\), then \(\Vert x_{n}\Vert _{\varphi }\rightarrow 0\) for all sequences \(( x_{n} ) \) in \(E_{\varphi }\).

The known proofs of properties \(( i ) \)\(( vi ) \) cannot be applied for the nonconvex Orlicz function φ. In [7] and [8] we presented new proofs of the above lemma for the quasimodular and the quasinorm. We have shown that for properties \((i ) \) and \(( ii ) \) we need additionally the condition \(\Delta _{\varepsilon }^{E}\), which is a substitute for convexity. Moreover, to show the properties \(( iv ) \) and \(( v ) \) the so-called condition \(\Delta _{2-str}^{E}\) is required (the condition \(\Delta _{2-str}^{E}\) is essentially stronger than \(\Delta _{2}^{E}\), in general). Finally, the condition \((iii)\) comes from Lemma 3.7 and the condition \(( vi ) \) is true in the same form as above (see [8]). Next, applying also some new techniques, we described order isomorphic and order linearly isometric copies of \(l^{\infty }\) in the quasinormed Calderón–Lozanovskiĭ spaces \(E_{\varphi }\) (a number of theorems describe these copies in the natural language of suitable properties of the quasinormed ideal space E and the nondecreasing Orlicz function φ – see [7]). In [8] we characterized the monotonicity properties of quasinormed Calderón–Lozanovskiĭ spaces \(E_{\varphi }\), that is, we described the strict monotonicity, the uniform monotonicity, and the respective orthogonal counterparts of the quasinormed Calderón–Lozanovskiĭ spaces \(E_{\varphi }\).

6.2 Open ends

The theory of modular spaces has been widely investigated in [26], see also [24]. Next, the modular spaces equipped with the additional measure structure (called modular function spaces) has been studied in [21]. Furthermore, the authors of [18] considered the modular function spaces from the geometry and the fixed-point theory point of view. All the monographs deal with the modulars (convex modulars) that lead to the F-normed (normed) spaces, respectively. It seems natural to study some aspects of the theory developed in the monographs [18] and [21] in the context of quasimodulars.

Data Availability

No datasets were generated or analysed during the current study.

References

  1. Bastero, J., Hudzik, H., Steinberg, A.M.: On smallest and largest spaces among rearrangement-invariant p-Banach function spaces \((0< p<1)\). Indag. Math. 2, 283–288 (1991)

    Article  MathSciNet  Google Scholar 

  2. Bennett, C., Sharpley, R.: Interpolation of Operators. Pure and Applied Mathematics, vol. 129. Academic Press, Boston (1988)

    Google Scholar 

  3. Birkhoff, G.: Lattice Theory, 3rd edn. American Mathematical Society Colloquium Publications, vol. XXV. Am. Math. Soc., Providence (1967)

    Google Scholar 

  4. Cui, Y., Hudzik, H., Kaczmarek, R., Kolwicz, P.: Uniform monotonicity of Orlicz spaces equipped with the Mazur-Orlicz F-norm and dominated best approximation in F-normed Köthe spaces. Math. Nachr. 295, 487–511 (2022)

    Article  MathSciNet  Google Scholar 

  5. Dobrakov, I.: On submeasures. I. Diss. Math. 112, 35 (1974)

    MathSciNet  Google Scholar 

  6. Foralewski, P., Hudzik, H.: Some basic properties of generalized Calderon-Lozanovskiĭ spaces. Collect. Math. 48, 523–538 (1997)

    MathSciNet  Google Scholar 

  7. Foralewski, P., Hudzik, H., Kolwicz, P.: Copies of \(l^{\infty}\) in quasi-normed Calderón–Lozanovskiĭ spaces. Submitted

  8. Foralewski, P., Kolwicz, P.: Monotonicities of quasi-normed Calderón–Lozanovskiĭ spaces with applications to approximation problems. Submitted

  9. Hudzik, H., Kamińska, A., Mastyło, M.: Monotonicity and rotundity properties in Banach lattices. Rocky Mt. J. Math. 30, 933–950 (2000)

    Article  MathSciNet  Google Scholar 

  10. Kalton, N.J.: Convexity conditions for nonlocally convex lattices. Glasg. Math. J. 25, 141–152 (1984)

    Article  MathSciNet  Google Scholar 

  11. Kalton, N.J.: Quasi-Banach spaces. In: Handbook of the Geometry of Banach Spaces, vol. 2, pp. 1099–1130. North-Holland, Amsterdam (2003)

    Chapter  Google Scholar 

  12. Kalton, N.J., Peck, N.T., Roberts, J.W.: An F-Space Sampler. London Mathematical Society Lecture Note Series, vol. 89. Cambridge University Press, Cambridge (1984)

    Book  Google Scholar 

  13. Kamińska, A., Maligranda, L.: Order convexity and concavity of Lorentz spaces \(\Lambda _{p,w}, 0< p<\infty \). Stud. Math. 160, 267–286 (2004)

    Article  MathSciNet  Google Scholar 

  14. Kamińska, A., Maligranda, L., Persson, L.E.: Indices, convexity and concavity of Calderón-Lozanovskii spaces. Math. Scand. 92, 141–160 (2003)

    Article  MathSciNet  Google Scholar 

  15. Kamińska, A., Raynaud, Y.: Copies of \(l_{p}\) and \(c_{0}\) in General Quasi-Normed Orlicz-Lorentz Sequence Spaces, Function Spaces. Contemp. Math., vol. 435, pp. 207–227. Am. Math. Soc., Providence (2007)

    Google Scholar 

  16. Kamińska, A., Żyluk, M.: Local geometric properties in quasi-normed Orlicz spaces. Banach Cent. Publ. 119, 197–221 (2019)

    Article  MathSciNet  Google Scholar 

  17. Kantorovich, L.V., Akilov, G.P.: Functional Analysis, 2nd edn. Pergamon Press, Oxford-Elmsford (1982). Translated from the Russian by, Silcock, Howard L.

    Google Scholar 

  18. Khamsi, M.A., Kozlowski, W.M.: Fixed Point Theory in Modular Function Spaces. Springer, Cham (2015). x+245

    Book  Google Scholar 

  19. Kiwerski, T., Kolwicz, P., Maligranda, L.: Isomorphic and isometric structure of the optimal domains for Hardy-type operators. Stud. Math. 260, 45–89 (2021)

    Article  MathSciNet  Google Scholar 

  20. Kolwicz, P., Leśnik, K., Maligranda, L.: Pointwise multipliers of Calderón-Lozanovskiĭ spaces. Math. Nachr. 286, 876–907 (2013)

    Article  MathSciNet  Google Scholar 

  21. Kozlowski, W.M.: Modular Function Spaces. Monographs and Textbooks in Pure and Applied Mathematics, vol. 122. Marcel Dekker, Inc., New York (1988). x+252

    Google Scholar 

  22. Lee, H.J.: Complex convexity and monotonicity in quasi-Banach lattices. Isr. J. Math. 159, 57–91 (2007)

    Article  MathSciNet  Google Scholar 

  23. Lindenstrauss, J., Tzafriri, L.: Classical Banach Spaces. II, Function Spaces. Ergebnisse der Mathematik und ihrer Grenzgebiete, vol. 97. Springer, Berlin (1979)

    Book  Google Scholar 

  24. Maligranda, L.: Orlicz Spaces and Interpolation. Seminários de Matemática [Seminars in Mathematics], vol. 5. Universidade Estadual de Campinas, Departamento de Matemática, Campinas (1989)

    Google Scholar 

  25. Maligranda, L.: Type, cotype and convexity properties of quasi-Banach spaces. In: International Symposium on Banach and Function Spaces, Kitakyushu, 2003, Banach and Function Spaces, pp. 83–120. Yokohama Publ., Yokohama (2004)

    Google Scholar 

  26. Musielak, J.: Orlicz Spaces and Modular Spaces. Lecture Notes in Mathematics, vol. 1034. Springer, Berlin (1983)

    Google Scholar 

  27. Raynaud, Y., Tradacete, P.: Calderón-Lozanovskii interpolation on quasi-Banach lattices. Banach J. Math. Anal. 12, 294–313 (2018)

    Article  MathSciNet  Google Scholar 

  28. Wnuk, W.: Banach Lattices with Order Continuous Norms. Advanced Topics in Mathematics. Polish Scientific Publishers PWN, Warszawa (1999)

    Google Scholar 

Download references

Funding

The scientific research of Paweł Kolwicz was carried out at the Poznań University of Technology as part of the Rector’s grant 2021 (research project No. 0213/SIGR/2154).

Author information

Authors and Affiliations

Authors

Contributions

HH initiated the research. PF and PK contributed equally and significantly in this manuscript, and they read and approved the final version.

Corresponding author

Correspondence to Paweł Foralewski.

Ethics declarations

Competing interests

The authors declare no competing interests.

Additional information

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Foralewski, P., Hudzik, H. & Kolwicz, P. Quasinormed spaces generated by a quasimodular. J Inequal Appl 2024, 86 (2024). https://doi.org/10.1186/s13660-024-03162-w

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s13660-024-03162-w

Mathematics Subject Classification

Keywords