1 Introduction

Let \(\Omega \) be a set and \(\mathcal {R}\) a \(\delta \)-ring consisting of subsets of \(\Omega \). Given a vector measure \(\nu : \mathcal {R}\rightarrow X\), where X is a (real or complex) Banach space, we obtain the Banach space of weakly-integrable functions \(L^1_w(\nu )\), which has the space of \(\nu \)-integrable functions \(L^1 (\nu )\) as a closed subspace. With the order given by \(f \ge g\) if \(f \ge g\) outside a \(\nu \)-null set, we have that \(L^1_w (\nu )\) is a \(\sigma \)-Fatou Banach lattice and \(L^1 (\nu )\) is an order continuous Banach lattice. (See Sect. 2 for definitions.) The integration theory with respect to vector measures defined on \(\delta \)-rings was developed mainly by Lewis [18], Masani and Niemi [21, 22] and Delgado [8]. It extends the well known theory for vector measures defined on \(\sigma \)-algebras [25, Ch. 3].

If E is a real or complex order continuous Banach lattice, Curbera [5, p. 22], [10, p. 246] showed that there is a set \(\Omega \), a \(\delta \)-ring \(\mathcal {R}\) consisting of subsets of \(\Omega \), and a vector measure \(\nu : \mathcal {R}\rightarrow E\), such that E and \(L^1 (\nu )\) are order and isometrically isomorphic . It follows that the dual spaces \(L^1 (\nu )^*\) and \(E^*\), both of which are Banach lattices, are also order and isometrically isomorphic. Hence by studying a dual space of the form \(L^1 (\nu )^*\) we are implicitly analyzing the dual space of the most general order continuous Banach lattice.

The study of \(L^1 (\nu )^*\) can be done through the associate space theory, systematically developed by Luxemburg and Zaanen [28, Ch. 15], [29, Ch. 16, Sect. 112]. This theory begins with an arbitrary positive measure space \((\Omega , \Sigma , \mu )\) and applies to a \(\mu \)-Banach function space (\(\mu \)-B.f.s. for short) E which is saturated. Then also its associate space \(E^\times \) is a \(\mu \)-Banach function space and the operator \(R: E^\times \rightarrow E^*\), defined by \(R(g) := R_g\), where \(R_g (f) := \int _\Omega fg d \mu \), is a linear isometry. This operator, called the canonical isometry, allows us to consider the associate space \(E^\times \) as a closed subspace of the dual space \(E^*\). When the involved measure \(\mu \) is \(\sigma \)-finite and the saturated B.f.s. E is \(\sigma \)-order continuous, then the canonical isometry R is onto. Moreover, R also preserves the lattice structure and so we write \(E^\times =E^*\). In the following we will maintain this notation to indicate that R is onto.

To apply the above theory for studying \(L^1(\nu )^*\), in the first place we have to find a positive measure \(\mu \) which is also a local control measure for \(\nu \), that is, such that \(\mu \) and \(\nu \) have the same null sets. In this situation \(L^1 (\nu )\) is a \(\mu \)-B.f.s. Assume the \(\delta \)-ring \(\mathcal {R}\) on which the vector measure \(\nu \) is defined is a \(\sigma \)-algebra. Then it is well known that \(\nu \) has a finite local control measure \(\mu \) with respect to which \(L^1(\nu )\) is a saturated \(\mu \)-B.f.s. [25, pp. 107–108]. It follows that in this case we have \(L^1 (\nu )^\times = L^1 ( \nu )^*\).

If the \(\delta \)-ring \(\mathcal {R}\) is not a \(\sigma \)-algebra, we cannot proceed directly as above to obtain for \(L^1 (\nu )\) similar results to those we have just mentioned, since in this situation \(\nu \) may not have a \(\sigma \)-finite local control measure. This is a problem, since a key fact for the associate space theory to work is that when the measure \(\mu \) is \(\sigma \)-finite, then the saturation of a \(\mu \)-B.f.s. E implies that of \(E^\times \).

However, relying on a result of Brooks and Dinculeanu [3], it has recently been pointed out by Jiménez et al. [16] the fact that any vector measure defined on a \(\delta \)-ring \(\mathcal {R}\), always has a local control measure \(\mu \) that is also \(\sigma \)-finite on any set \(B \in \mathcal {R}\). We will see that a measure of this kind, which we have called Brooks–Dinculeanu measure, is appropriate for our objectives.

In this paper we first study the associate space \(E^\times \) of a saturated \(\mu \)-B.f.s., for a locally \(\sigma \)-finite positive measure \(\mu \). Thereafter we consider a vector measure \(\nu \) and apply the results to the \(\mu \)-Banach function spaces of p integrable functions \(L^p (\nu )\) and of weakly integrable functions \(L ^p_w (\nu ), 1 \le p < \infty \), where \(\mu \) is a Brooks–Dinculeanu measure for \(\nu \). The main question we discussed was that of the validity of the equality \(L^1(\nu )^\times = L^1(\nu )^*\).

We divided our work in five sections, including this Introduction. In Sect. 2 we present the notation, definitions and basic results that we have needed.

The general theory for the associate space \(E^\times \) of a saturated \(\mu \)-B.f.s E is given in Sect. 3. As in the well known \(\sigma \)-finite case, it turns out that the associate space \(E^\times \) always has the \(\sigma \)-Fatou property. When \(\mu \) is a \(\sigma \)-Fatou property, we show that the \(\sigma \)-Fatou property and the Fatou properties are equivalent for E. Based on general properties of the dual space of an arbitrary Banach lattice [29, Ch. 14, Ch. 15], we establish that if \(E^\times = E^*\), then E is order continuous and \(E^\times \) has the Fatou property.

In Sect. 4 we restrict our considerations to a locally \(\sigma \)-finite measure \(\mu \). In this setting we obtain results that are well known in the \(\sigma \)-finite case. Particularly we show that \(E^\times \) is also a saturated \(\mu \)-B.f.s., that \(E^{\times \times } = E\) when E has the Fatou property, and that the factorization \(L^1 (\mu ) = E E ^\times \) holds. When the saturated \(\mu \)-B.f.s. is \(\sigma \)-order continuous, we prove that \(E^\times = E^*\) if, and only if, \(E^\times \) has the Fatou property. In a forthcoming work we plan to show that this last is not always the case.

In the last section we apply our development to the spaces \(L^p(\nu )\) and \(L^p_w(\nu )\), \(1\le p <\infty \), obtained from a vector measure \(\nu \). These are Banach function spaces with respect to any Brooks–Dinculeanu measure for \(\nu \). Thus we show that \(L^p (\nu )^{\times \times } = L^p_w (\nu )\) and \(L^p(\nu )^\times = L^p_w (\nu )^\times \) for \(1 \le p < \infty \). When the vector measure \(\nu \) is defined on a \(\sigma \)-algebra the first of these equalities was established by Curbera and Ricker [6, Prop. 2.4], [7, Prop. 1].

We give several situations where \(L^1(\nu )^\times =L^1(\nu )^*\) holds, one of them being the case of a decomposable vector measure. This turns out to be important, since the vector measure \(\nu \) that Calabuig et al. used to represent an order continuous Banach lattice as \(L^1(\nu )\), is a decomposable vector measure [4].

Finally, we establish that \(L^p(\nu )^\times =L^p(\nu )^*\) if \(1<p<\infty \) and that \(L^1(\nu )^\times =L^1(\nu )^*\) when \(L^1(\nu )\) is reflexive. We also verify that a reflexivity criterion proven by Fernández et al. [12, Cor. 3.10] for a vector measure defined on a \(\sigma \)-algebra is still valid in the \(\delta \)-ring case.

To complete this introduction, we want to note that Okada was the first to obtain a description of \(L^1(\nu )^*\) for a classical vector measure \(\nu \) [24]. Later Galaz-Fontes gave a representation for \(L^p(\nu )^*\) when \(1 < p < \infty \) [14] and recently Mastylo and Sánchez Pérez have established representations of these kind for a dual Banach space in a more general context [23].

2 Notation and basic results

2.1 Banach lattices

Throughout this paper all vector spaces considered will be with respect to \(\mathbb {K}\), where \(\mathbb {K}=\mathbb {C}\), the field of complex numbers or \(\mathbb {K}=\mathbb {R}\), the field of real numbers. Let X be a normed space. By \(B_X\) we will indicate its unit closed ball and by \(X^*\) its dual space. We will represent by \(\langle \cdot ,\cdot \rangle \) the duality pairing, i.e. \(\langle x,x^*\rangle :=x^*(x), \forall \ x\in X \text{ and } \ x^*\in X^*\). If Y is other normed space, to express that \(X=Y\) as sets and with equal norms, we will write \(X\equiv Y\).

Let X be a real vector lattice with order \(\le \). For \(A\subset X\) we will be denote by \(A^+\) the subset of X consisting of all \(f\in A\) such that \(0\le f\). Given \(f\in X\), we indicate by \(f^+\), \(f^-\) and |f| its positive part, negative part and modulus, respectively. Finally X is called Dedekind \(\sigma \) -complete if every non-empty countable subset which is bounded from above has a supremum.

Let J be a directed set. A net \(\{f_\tau \}_{\tau \in J}\subset X\) is an upwards directed system if for \(\tau _1\) and \(\tau _2\) in J, there exists \(\tau _3\in J\) such that \(f_{\tau _1}\le f_{\tau _3}\) and \(f_{\tau _2}\le f_{\tau _3}\). In this case we will use the notation \(f_\tau \uparrow \). If additionally there exists \(f=\sup _\tau f_\tau \in X\) we write \(f_\tau \uparrow f\). Similarly, if the sequence \(\{f_n\}\subset X\) is such that \(f_n\le f_{n+1}\), \(\forall \ n\in \mathbb {N}\), we will indicate \(f_n\uparrow \) and if additionally \(f=\sup _n f_n\) then we will write \(f_n\uparrow f\). Analogously, we define a downwards directed system \(\{f_\tau \}\) and the notations \(f_\tau \downarrow \) and \(f_\tau \downarrow f\).

A real normed vector lattice X is a real normed space that is a vector lattice and whose norm \(\Vert \cdot \Vert _X\) has the lattice property, that is

$$\begin{aligned} \mathrm{if } f,g\in Y\, \mathrm{satisfy },\ |f|\le |g|, \text{ then } \Vert f\Vert _X\le \Vert g\Vert _X. \end{aligned}$$
(2.1)

If in addition the space is complete, we will say that X is real Banach lattice.

Let X be a real normed vector lattice. Then X has the weak Fatou property if for each upwards directed system \(\{f_\tau \}\subset X^+\) with \(\sup _\tau \Vert f_\tau \Vert _X<\infty \), there exists \(f=\sup _\tau f_\tau \in X\); if additionally \(\Vert f\Vert _X=\sup _n \Vert f_n\Vert _X\), then X has the Fatou property. Similarly, X has the weak \(\sigma \) -Fatou property, if given \(\{f_n\}\subset X^+\) such that \(f_n\uparrow \) and \(\sup _n \Vert f_n\Vert _X<\infty \), then there exists \(f=\sup _n f_n\in X\), and if additionally \(\Vert f\Vert _X=\sup _n \Vert f_n\Vert _X\), then X is said to have the \(\sigma \) -Fatou property. We say that X is order continuous, if for any system \(\{f_\tau \}\subset X\) satisfying \(f_\tau \downarrow 0\) it follows that \(\Vert f_\tau \Vert _X\downarrow 0\). Analogously, X is \(\sigma \) -order continuous if for any sequence \(\{f_n\}\subset X\) satisfying \(f_n\downarrow 0\) we have that \(\Vert f_n\Vert _X\downarrow 0\).

Take a real Banach lattice X. Then in \(Z:=X+iX\), the complexification of X, the modulus is defined by \(|h|:=\sup \{|(\cos \theta ) f+(\sin \theta ) g|:0\le \theta <2\pi \}, \ \forall \ h:=f+ig\in Z\) [29, Ch. 14; Thm. 91.2], the norm by \(\Vert h\Vert _Z=\Vert \ |h| \ \Vert _X, \ \forall \ h\in Z\), and the order is given by \(f\le g\) in Z, if \(f,g\in X\) and \(f\le g\). In this case Z is called a complex Banach lattice, X is its real part and we write \(X = Z_\mathbb {R}\). Observe that \(Z^+=X^+\). We will say that a complex Banach lattice has one of the properties we introduced above if its real part has it. Henceforth we will say only Banach lattice (normed vector lattice) to refer to a complex or real Banach lattice (normed vector lattice).

Let X be a Banach lattice. An ideal Y of X is a vector subspace of X if \(f\in X\) with \(|f|\le |g|\) for some \(g\in Y\) implies \(f\in Y\).

Let \(T : X \rightarrow Y\) be a linear operator between Banach lattices. Then T is said to be positive if for each \(f \in X^+\) we have that \(T(f) \in Y^+\). In this case \(T(X_\mathbb {R})\subset Y_\mathbb {R}\) and T is bounded [1, Lemma 3.22]. We will say that T is an order isometry if T is an isometry, T is onto and both T and \(T^{-1}\) are positive operators. This last condition is equivalent to

$$\begin{aligned} Tf \ge 0 \text{ if } \text{ and } \text{ only } \text{ if } f \ge 0, \quad \forall f \in X. \end{aligned}$$

In this case we have that \(T(X_\mathbb {R}) = Y_\mathbb {R}\), \(T(\sup \{f, g\}) = \sup \{Tf, Tg\}, \forall f, g \in X_\mathbb {R}\) and \(T|f| = |Tf|, \ \forall f \in X\).

Let X be a real Banach lattice. Then its dual space \(X^*\) is a Banach lattice with the order given by

$$\begin{aligned} \varphi \le \psi , \text{ if } \varphi (f)\le \psi (f), \quad \forall \ f\in X^+, \ \varphi ,\psi \in X^*. \end{aligned}$$
(2.2)

In this case the supremum and infimum are uniquely determined by

$$\begin{aligned} \begin{array}{c} \sup \{\varphi ,\psi \}(f):=\sup \{\varphi (g)+\psi (h): f=g+h, g\ge 0, \ h\ge 0 \}, \\ \inf \{\varphi ,\psi \}(f):=\inf \{\varphi (g)+\psi (h): f=g+h, g\ge 0, \ h\ge 0 \}, \end{array} \end{aligned}$$
(2.3)

for each \(\varphi ,\psi \in X^*\) and \(f\in X^+\) [27][Chap II, Props. 4.2, 5.5] .

Now assume that X is a complex Banach lattice. Given \(\varphi \in X_\mathbb {R}^{ \ *}\), we will indicate by \(\widetilde{\varphi }:X\rightarrow \mathbb {C}\) its canonical extension, that is, \(\widetilde{\varphi }(x+iy)=\varphi (x)+i\varphi (y)\). If \(\Phi :X\rightarrow \mathbb {C}\) is a bounded linear functional, then \(\Phi \) has the form \(\Phi =\widetilde{\varphi }+i\widetilde{\psi }\), where \(\widetilde{\varphi }\) and \(\widetilde{\psi }\) are the canonical extensions of linear functionals \(\varphi ,\psi \in X_\mathbb {R}^{ \ *}\). Identifying \(X_\mathbb {R}^{ \ *}\) with \(\widetilde{X}^*_{ \ \mathbb {R}} \subset X^*\), we have that \(X^*=X_\mathbb {R}^{ \ *}+iX_\mathbb {R}^{ \ *}\) is a Banach lattice. As can be seen in [27, §11], in this case

$$\begin{aligned} |\Phi |(f)=\sup _{|g|\le f}|\Phi (g)|, \quad \forall \ f\in X^+. \end{aligned}$$
(2.4)

2.2 \(\mu \)-Banach function spaces

Given a measurable space \((\Omega ,\Sigma )\) we will denote by \(L^0(\Sigma )\) the space formed by the \(\Sigma \)-measurable functions \(f:\Omega \rightarrow \mathbb {K}\). If additionally we have a positive measure \(\mu \) defined on \(\Sigma \), we indicate by \(\mathcal {N}_0(\mu )\) the family of \(\mu \) -null subsets, i. e., the sets \(A\in \Sigma \) such that \(\mu (A)=0\). As usual a property holds \(\mu \) -almost everywhere (briefly \(\mu \)-a.e.) if it holds except on a \(\mu \)-null set. We indicate by \(L^0(\mu )\) the space of equivalence classes of functions in \(L^0(\Sigma )\), where two functions are identified when they are equal \(\mu \)-a.e.

Note that, when \(\mathbb {K}=\mathbb {C}\), the space \(L^0(\mu )\) is the complexification of the real space \( L^0(\mu )_\mathbb {R}:=\{f\in L^0(\mu ): f \text{ take } \text{ its } \text{ values } \text{ in } \ \mathbb {R}\ \mu \text{-a.e. } \}\).

In \(L^0(\mu )_\mathbb {R}\) we will always consider the \(\mu \)-a.e. pointwise order. Let \(f\in L^0(\mu )\). So \(\text{ Re }f,\text{ Im }f\in L^0(\mu )_\mathbb {R}\) and \(f=\text{ Re }f+i\text{ Im }f\). Moreover,

$$\begin{aligned} {\sup _{0\le \theta <2\pi }|(\cos \theta )\text{ Re }f +(\sin \theta ) \text{ Im }f|=\sqrt{(\text{ Re }f)^2+(\text{ Im }f)^2}=|f|.} \end{aligned}$$
(2.5)

We will say that a normed space \(E\subset L^0(\mu )\) is a normed function space related to \(\mu \) (briefly \(\mu \)-n.f.s.) if E is a vector subspace of \(L^0(\mu )\) such that \(f\in L^0(\mu )\) with \(|f|\le |g|\) for some \(g\in E\), implies \(f\in E\) and the lattice property (2.1) holds (with E instead of X). If additionally E is complete we will call it Banach function space related to \(\mu \) (briefly \(\mu \)-B.f.s.). We must note that in the literature there appear other definitions with the same name, such as in [19, Def. 1.b.17] and in [2, Def. I.1.3].

Let E be a \(\mu \)-B.f.s. Then E is Dedekind \(\sigma \)-complete. So, E is \(\sigma \)-order continuous if and only if E is order continuous [29, 103.9]. Now if E is a \(\mu \)-n.f.s. with the \(\sigma \)-Fatou property, then E is complete [28, Ch. 15, §65, Thm. 1]. Thus, E is a \(\mu \)-B.f.s. with the \(\sigma \) -Fatou property.

Let E be a complex \(\mu \)-B.f.s. Note that \(E_\mathbb {R}=E\cap L^0(\mu )_\mathbb {R}\), with the \(\mu \)-a.e. pointwise order, is a real Banach lattice and \(E=E_\mathbb {R}+iE_\mathbb {R}\). It follows from (2.5) that E is a complex Banach lattice. Moreover, \(f\in E\) if and only if \((\text{ Re }f)^+,(\text{ Re }f)^-,(\text{ Im }f)^+,(\text{ Im }f)^-\in E^+\).

2.3 Integration with respect to measures defined on \(\delta \)-rings

Let \(\Omega \) be a set. A family \(\mathcal {R}\) of subsets of \(\Omega \) is a \(\delta \) -ring if \(\mathcal {R}\) is a ring which is closed under countable intersections. From now on in this paper \(\mathcal {R}\) will be a \(\delta \)-ring. We denote by \(\mathcal {R}^{loc}\) the \(\sigma \)-algebra of all sets \(A\subset \Omega \) such that \(A\cap B\in \mathcal {R}\), \(\forall \ B\in \mathcal {R}\). Given \(A\in \mathcal {R}^{loc}\) we indicate by \(\mathcal {R}_A\) the \(\delta \)-ring \(\{B\subset A:B\in \mathcal {R}\}\) and by \(\pi _A\) the collection of finite families of pairwise disjoint sets in \(\mathcal {R}_A\). Note that if \(\Omega \in \mathcal {R}\), then \(\mathcal {R}\) is a \(\sigma \)-algebra, and in this case we have that \(\mathcal {R}^{loc}=\mathcal {R}\). Moreover, for each \(B\in \mathcal {R}\) it turns out that \(\mathcal {R}_B\) is a \(\sigma \)-algebra.

A scalar measure (positive measure) is a function \(\lambda :\mathcal {R}\rightarrow \mathbb {K}\) (\(\lambda :\mathcal {R}\rightarrow [0,\infty ]\)) satisfying that if \(\{B_n\}\subset \mathcal {R}\), is a family of pairwise disjoint sets such that \(\bigcup _{n=1}^{\infty }B_n\in \mathcal {R}\), then \(\sum _{n=1}^{\infty } \lambda (B_n)=\lambda \left( \bigcup _{n=1}^{\infty }B_n\right) \). The variation of \(\lambda \) is the countably additive measure \(|\lambda |:\mathcal {R}^{loc}\rightarrow [0,\infty ]\) defined by

$$\begin{aligned} |\lambda |(A):=\sup \left\{ \sum _{j=1}^{n}|\lambda (A_j)| :\{A_j\}\in \pi _A\right\} . \end{aligned}$$

A function \(f\in L^0(\mathcal {R}^{loc})\) is \(\lambda \) -integrable if \(f\in L^1(|\lambda |)\). We denote by \(L^1(\lambda )\) the subspace of \(L^0(\lambda )\) formed by the \(\lambda \)-integrable functions. Then \(L^1(\lambda )\) with norm given by \(|f|_{1,\lambda }:=\int _\Omega |f|d|\lambda |\) is a \(\sigma \)-order continuous \(|\lambda |\)-B.f.s. with the \(\sigma \)-Fatou property. The following result is basic in the theory; when \(\lambda \) is a scalar measure it was established by Masani and Niemi [21, Lemma 2.30, Thm. 2.32], for the case of a positive measure we can proceed similarly.

Proposition 2.1

If \(f\in L^0(\mathcal {R}^{loc})\), then

$$\begin{aligned} {\int _A |f|d|\lambda |=\sup _{B\in \mathcal {R}_A}\int _B|f|d|\lambda |,} \quad \forall \ A\in \mathcal {R}^{loc}. \end{aligned}$$
(2.6)

Therefore, \(f\in L^1(|\lambda |)\) if and only if \({\sup _{B\in \mathcal {R}}\int _B|f|d|\lambda |<\infty }\).

Let X be a Banach space. A function \(\nu :\mathcal {R}\rightarrow X\) is a vector measure if \(\sum _{n=1}^\infty \nu (B_n)=\nu \left( \bigcup _{n=1}^\infty B_n\right) \), for any collection \(\{B_n\}\subset \mathcal {R}\) of pairwise disjoint sets such that \(\bigcup _{n=1}^\infty B_n \in \mathcal {R}\). The variation of \(\nu \) is the positive measure \(|\nu |\) defined in \(\mathcal {R}^{loc}\) by

$$\begin{aligned} |\nu |(A):=\sup \left\{ \sum _j \Vert \nu (A_j)\Vert _X : \{A_j\}\in \pi _A \right\} . \end{aligned}$$

The semivariation of \(\nu \) is the function \(\Vert \nu \Vert :\mathcal {R}^{loc}\rightarrow [0,\infty ]\) given by

$$\begin{aligned} {\Vert \nu \Vert (A):=\sup \{|\langle \nu ,x^*\rangle |(A):x^*\in B_{X^*}\}}, \end{aligned}$$

where \(|\langle \nu ,x^*\rangle |\) is the variation of the scalar measure \(\langle \nu ,x^*\rangle :\mathcal {R}\rightarrow \mathbb {K}\), where

$$\begin{aligned} \langle \nu ,x^*\rangle (B):=\langle \nu (B),x^*\rangle , \quad \forall \ B\in \mathcal {R}. \end{aligned}$$

The semivariation of \(\nu \) is finite in \(\mathcal {R}\) and for any \(A\in \mathcal {R}^{loc}\) satisfies \(\Vert \nu \Vert (A)\le |\nu |(A)\). A set \(A\in \mathcal {R}^{loc}\) is said to be \(\nu \)-null if \(\Vert \nu \Vert (A)=0\). We will denote by \(\mathcal {N}_0(\nu )\) the collection of \(\nu \)-null sets. It turns out that \(\mathcal {N}_0(\nu )=\mathcal {N}_0(|\nu |)\). Moreover \(A\in \mathcal {N}_0(\nu )\) if and only if \(\nu (B)=0\), \(\forall \ B\in \mathcal {R}_A\). We say that a positive measure \(\lambda :\mathcal {R}\rightarrow [0,\infty ]\) is a local control measure for \(\nu \), if \(\mathcal {N}_0(|\lambda |)=\mathcal {N}_0(\nu )\) [8, p. 437]. Then, \(|\nu |\) is a local control measure for \(\nu \). We define \(L^0(\nu )\) as the space of equivalence classes of functions in \(L^0(\mathcal {R}^{loc})\), where two functions are identified when they are equal \(\nu \)-a.e. So, \(L^0(\nu )=L^0(|\lambda |)\), where \(\lambda \) is any local control measure for \(\nu \).

A function \(f\in L^0(\mathcal {R}^{loc})\) is weakly \(\nu \) -integrable, if \(f\in L^1(\langle \nu ,x^*\rangle )\), for each \(x^*\in X^*\). We will denote by \(L^1_w(\nu )\) the subspace of \(L^0(\nu )\) of all weakly \(\nu \)-integrable functions. With the norm \({\Vert f\Vert _\nu :=\sup \{\int _\Omega |f| d|\langle \nu ,x^*\rangle |:x^*\in B_{X^*} \}}\), \(L^1_w(\nu )\) is a \(|\lambda |\)-B.f.s. with the \(\sigma \)-Fatou property, where \(\lambda \) is a local control measure for \(\nu \).

A function \(f\in L^1_w(\nu )\) is \(\nu \) -integrable, if for each \(A\in \mathcal {R}^{loc}\) there exists a vector \(x_A\in X\), such that \({ \langle x_A,x^*\rangle =\int _A f d\langle \nu ,x^*\rangle , \ \forall \ x^*\in X^*}\). The subset of all \(\nu \)-integrable functions is a closed subspace of \(L^1_w(\nu )\) and it will be denoted by \(L^1(\nu )\). We indicate by \(S(\mathcal {R})\) the collection of simple functions in \(L^0(\mathcal {R}^{loc})\) which have support in \(\mathcal {R}\). It turns out that \(L^1(\nu )\), with norm \(\Vert \cdot \Vert _\nu \) is a \(\sigma \)-order continuous \(\mu \)-B.f.s. where \(S(\mathcal {R})\) is a dense subspace.

We also notice that \(L^1(\nu )=L^1_w(\nu )\) if X does not contain a copy of \(c_0\) [18, Thm. 5.1].

3 Associate space

Let \((\Omega ,\Sigma ,\mu )\) be a positive measure space and let us consider a \(\mu \)-B.f.s. E. We will show that several of the basic results about the associate space of E when \(\mu \) is a \(\sigma \)-finite measure, can also be established in the case that \(\mu \) is not \(\sigma \)-finite. We will begin by just enunciating some of these results; they can be proven as in the \(\sigma \)-finite case [28, Ch. 15], [2, Ch. 1].

The vector space defined by \( E^\times :=\{g\in L^0(\mu ): gf\in L^1(\mu ), \ \forall \ f\in E \}\) is called the associate space of E and the function

$$\begin{aligned} \Vert g\Vert _{E^\times }:=\sup \left\{ \int _\Omega |gf|d\mu : f\in B_E\right\} , \quad \forall \ g\in E^\times , \end{aligned}$$
(3.1)

is a seminorm. For each \(f\in E\) and \(g\in E^\times \) the Hölder inequality is satisfied:

$$\begin{aligned} {\int _\Omega |gf|d\mu \le \Vert g\Vert _{E^\times }\Vert f\Vert _E}. \end{aligned}$$

We also have that for the function \(\Vert \cdot \Vert _{E^\times }\) to be a norm, it is necessary and sufficient that E be saturated, that is, for each \(A\in \Sigma \) with positive measure there exists \(B\in \Sigma _A\) such that \(\mu (B)>0\) and \(\chi _B\in E\). Next we give an equivalent condition for saturation. For this, let us first recall that a \(\mu \)-B.f.s. Y is order dense in \(L^0(\mu )\) if for any \(f\in L^0(\mu )^+\) there exists an upwards directed system \(\{f_\tau \}\subset Y^+\) such that \(f_\tau \uparrow f\).

Lemma 3.1

Let E a \(\mu \)-B.f.s. The following statements are equivalent:

  1. (i)

    The space E is saturated.

  2. (ii)

    The space E is order dense in \(L^0(\mu )\).

  3. (iii)

    The seminorm \(\Vert \cdot \Vert _{E^\times }\) is a norm.

Proof

The equivalence (i) \(\Leftrightarrow \) (iii) is proved as in the \(\sigma \)-finite case [28, Ch. 15, §69, Thm. 4]. Let us prove (i) \(\Leftrightarrow \) (ii).

Since E is a Banach lattice we have that E is archimedean. Then, it is enough to prove that E is saturated if and only if for each \(0\ne f\in L^0(\mu )\) there exists \(g\in E\) such that \(0<|g|\le |f|\) [20, Thm. 22.3(vi)]. Assume that E is saturated. Consider \(0\ne f\in L^0(\mu )\) and define

$$\begin{aligned} A_n:=\left\{ x\in \Omega :\frac{1}{n}\le |f(x)|\right\} , \quad \forall \ n\in \mathbb {N}. \end{aligned}$$

Let us fix \(N\in \mathbb {N}\) such that \(\mu (A_N)>0\). Since E is saturated there exists \(B\subset A_N\) with \(\mu (B)>0\) and \(\chi _B\in E\). Then, \(g:=\frac{1}{N}\chi _B\in E\) and \(0<|g|\le |f|\).

To establish the other implication, take \(A\in \Sigma \) such that \(\mu (A)>0\). Hence \(0\ne \chi _A\in L^0(\mu )\). Thus there exists \(g\in E\) satisfying \(0<|g|\le \chi _A\). As \(0\ne g\in L^0(\mu )\) we can take \(\varphi \in S(\Sigma )\) with \(0<\varphi \le |g|\). Therefore \(\varphi \in E\) and so, there exists \(B\in \Sigma \) such that \(B\subset \mathrm{supp}\varphi \), \(\mu (B)>0\) and \(\chi _B\in E\). \(\square \)

Henceforth we will assume that E is a saturated \(\mu \)-B.f.s. Then as in the \(\sigma \)-finite case we have:

Proposition 3.2

The space \(E^\times \) is a \(\mu \)-B.f.s. with the \(\sigma \)-Fatou property.

For our next result, let us recall that a Banach lattice E is super Dedekind complete if every non-empty subset D of E which is bounded from above has a supremum and it contains an at most countable subset possessing the same supremum as D.

Proposition 3.3

If there exists a \(\sigma \)-order continuous \(\mu \)-B.f.s. F with the \(\sigma \)-Fatou property, such that \(E^\times \subset F\), then \(E^\times \) has the Fatou property.

Proof

Consider \(\{g_\tau \}\subset E^\times \) such that \(0\le g_\tau \uparrow \) and \(\sup _\tau \Vert g_\tau \Vert _{E^\times }<\infty \). Then \(\{g_\tau \}\subset F^+\) is an upwards directed system such that \(\sup _\tau \Vert g_\tau \Vert _F<\infty \). Since F is \(\sigma \)-order continuous and has the \(\sigma \)-Fatou property, then F has the Fatou property and is super Dedekind complete [29, Thm. 113.4]. Hence \(g:=\sup _\tau g_\tau \in F\) and there exists a sequence \(\{g_{\tau _n}\}\subset \{g_\tau \}\) such that \(g_{\tau _n}\uparrow g\) [20, Thm. 23.2.(iii)]. From the \(\sigma \)-Fatou property in \(E^\times \) we obtain that \(g\in E^\times \). And since \(\Vert \cdot \Vert _E^\times \) is a lattice norm, \(\sup _\tau \Vert g_\tau \Vert _{E^\times }\le \Vert g\Vert _{E^\times }\).

Now take \(f\in B_E\). Using the monotone convergence theorem and the Hölder inequality we have

$$\begin{aligned} \int _\Omega |gf|d\mu =\sup _n\int _\Omega |g_{\tau _n} f|d\mu \le \sup _n\Vert g_{\tau _n}\Vert _{E^\times }\Vert f\Vert _E\le \sup _n\Vert g_\tau \Vert _{E^\times }. \end{aligned}$$

Thus, \(\Vert g\Vert _{E^\times }\le \sup _\tau \Vert g_\tau \Vert _{E^\times }<\infty \). Hence, \(E^\times \) has the Fatou property. \(\square \)

Since \(L^1(\mu )\) is \(\sigma \)-order continuous and has the \(\sigma \)-Fatou property, we obtain:

Corollary 3.4

If \(\chi _\Omega \in E\), then \(E^\times \) has the Fatou property.

Next we will show that when \(\mu \) is \(\sigma \)-finite, it turns out that the associate space always has the Fatou property. For this it is necessary to make before a brief discussion.

Let A be in \(\Sigma \). We denote by \(\mu _A\) the restriction of the measure \(\mu \) to the \(\sigma \)-algebra \(\Sigma _A\) formed by the measurable subsets of A. Thus \((A,\Sigma _A,\mu _A)\) is a measure space.

For each \(f\in L^0(\Sigma _A)\) define the function \(f^\Omega :\Omega \rightarrow \mathbb {K}\) by \(f^\Omega (x)=f(x)\), if \(x\in A\) and \(f^\Omega (x)=0\) otherwise. Then \(f^\Omega \) is a \(\Sigma \)-measurable function which is called canonical extension of f. Now, the set \(E_A\) defined by

$$\begin{aligned} E_A:=\left\{ f\in L^0(\mu _A): f^\Omega \in E\right\} \end{aligned}$$
(3.2)

is a vector space. If \(h\in E\), then \((h_A)^\Omega =h\chi _A\in E\), where \(h_A\) is the restriction of h to A. Thus, \(h_A\in E_A\). In \(E_A\) we define the norm \(\Vert \cdot \Vert _A\) by

$$\begin{aligned} \Vert f\Vert _A:=\Vert f^\Omega \Vert _E, \quad \forall \ f\in E_A. \end{aligned}$$
(3.3)

Since E is a saturated \(\mu \)-B.f.s., it follows that also \(E_A\) is a saturated \(\mu _A\)-B.f.s. On the other hand if \({A=\bigcup _{n=1}^\infty A_n}\), where \(A_n\in \Sigma \) and \(\mu (A_n)<\infty \), \(\forall \ n\in \mathbb {N}\), we obtain that \(\mu _A\) is \(\sigma \)-finite. In this case it is well known that \(E_A^\times :=(E_A)^\times \) is saturated [28, Ch. 15, §71, Thm. 4]. Furthermore \(E^\times _A= (E^\times )_A\) and

$$\begin{aligned} {\Vert g\Vert _{E_A^\times }=\sup \left\{ \int _A|gf|d\mu _A:f\in B_{E_A}\right\} }=\Vert g^\Omega \Vert _{E^\times }, \quad \forall \ g\in E_A^\times . \end{aligned}$$
(3.4)

Theorem 3.5

If the measure \(\mu \) is \(\sigma \)-finite, then \(E^\times \) has the Fatou property.

Proof

We will assume that \(\mu (\Omega )>0\). Let us take an upwards directed system \(\{g_\tau \}_{\tau \in I}\subset E^\times \) such that \(g_\tau \ge 0\), \(\forall \ \tau \in I\) and \(\sup _{\tau }\Vert g_\tau \Vert _{E^\times }<\infty \). Now since \(\mu \) is \(\sigma \)-finite and E is a saturated \(\mu \)-B-f.s., there exist \(\{\Omega _n\}\subset \Sigma \) and \(N\in \mathcal {N}_0(\mu )\) such that \(\Omega _n\subset \Omega _{n+1}\), \(0\ne \chi _{\Omega _n}\in E\), \(\forall \ n\in \mathbb {N}\) and \({\Omega =\bigcup _{n=1}^\infty \Omega _n \cup N}\) [28, Ch. 15, §67, Thm. 4].

Fix \(n\in \mathbb {N}\). Let us denote by \(\Sigma _n\) the \(\sigma \)-algebra \(\Sigma _{\Omega _n}\) and by \(\mu _n\) the restriction of \(\mu \) to \(\Sigma _n\). Thus \((\Omega _n,\Sigma _n,\mu _n)\) is a finite measure space. Then the space \(E_n:=E_{\Omega _n}\) with the norm \(\Vert \cdot \Vert _n:=\Vert \cdot \Vert _{\Omega _n}\) is a saturated \(\mu _n\)-B.f.s. such that \(\chi _{\Omega _n}\in E_n\). By the above corollary we have that \(E^\times _n\) has the Fatou property.

Let \(g_{\tau ,n}\) be the restriction of \(g_\tau \) to the set \(\Omega _n\). Thus \(\{g_{\tau ,n}\}_{\tau \in I}\subset E^\times _n\) is an upwards directed system with \(\sup _{\tau }\Vert g_{\tau ,n}\Vert _{E^\times _n}\le \sup _{\tau }\Vert g_\tau \Vert _{E^\times }<\infty \). Therefore \(g_{(n)}:=\sup _{\tau }g_{\tau ,n}\in E^\times _n\) and \(\Vert g_{(n)}\Vert _{E^\times _n}=\sup _{\tau }\Vert g_{\tau ,n}\Vert _{E^\times _n}\).

Now let \(g_n:=(g_{(n)})^\Omega \) be the canonical extension of \(g_{(n)}\). Then \(\{g_n\}\subset E^\times \) is an increasing sequence. By (3.4),

$$\begin{aligned} {\sup _n\Vert g_n\Vert _{E^\times }=\sup _n\Vert g_{(n)}\Vert _{E^\times _n}\le \sup _{\tau }\Vert g_\tau \Vert _{E^\times }}. \end{aligned}$$
(3.5)

From the \(\sigma \)-Fatou property in \(E^\times \) and (3.5), it follows that

$$\begin{aligned} {g:=\sup _n g_n\in E^\times } \text{ and } {\Vert g\Vert _{E^\times }\le \sup _{\tau }\Vert g_\tau \Vert _{E^\times }}. \end{aligned}$$
(3.6)

Let us prove that \(g=\sup _\tau g_\tau \). Fix \(\tau \in I\). Since \(g_\tau \chi _{\Omega _n}\le g_n\le g\) for each \(n\in \mathbb {N}\), we have that \(g_\tau \le g\). Suppose that \(g'\in E^\times \) satisfies \(g_\tau \le g'\), \(\forall \ \tau \in I\). Now fix \( n\in \mathbb {N}\). Then \(g_\tau \chi _{\Omega _n}\le g'\chi _{\Omega _n}\), \(\forall \ \tau \), so \(g_n\le g'\). Therefore, \(g\le g'\) and we obtain that \(g=\sup _\tau g_\tau \). Finally \(\Vert \cdot \Vert _E^\times \) is a lattice norm and so the conclusion follows from (3.6). \(\square \)

Since we are assuming that E is a saturated \(\mu \)-B.f.s., then \(E^\times \) is a \(\mu \)-B.f.s. and we can consider its associate space, which is called second associate space of E and it is denoted by \(E^{\times \times }\). Thus

$$\begin{aligned} E^{\times \times }:=(E^\times )^\times =\left\{ h\in L^0(\mu ): hg\in L^1(\mu )\quad \forall \ g\in E^\times \right\} \end{aligned}$$

and the seminorm \(\Vert \cdot \Vert _{E^{\times \times }}:E^{\times \times }\rightarrow [0,\infty )\) is given by

$$\begin{aligned} \Vert h\Vert _{E^{\times \times }}:=\sup \left\{ \int _\Omega |hg|d\mu : g\in B_{E^\times }\right\} . \end{aligned}$$

Unlike the \(\sigma \)-finite case, we will see in Example 4.5 that it can happen that \(E^\times \) is not saturated. Nevertheless, we have that \(E\subset E^{\times \times }\) and

$$\begin{aligned} \Vert f\Vert _{E^{\times \times }}\le \Vert f\Vert _E \quad \forall \ f\in E. \end{aligned}$$
(3.7)

Corollary 3.6

Let E be a saturated \(\mu \)-B.f.s. If \(\mu \) is \(\sigma \)-finite, then

  1. (i)

    \(E^{\times \times }\) is a \(\mu \)-B.f.s. with the Fatou property.

  2. (ii)

    E has the \(\sigma \)-Fatou property if and only if E has the Fatou property.

Proof

Since \(\mu \) is \(\sigma \)-finite we have that \(E^\times \) is a saturated \(\mu \)-B.f.s. [28, Ch. 15, §71, Thm. 4]. Thus, from Theorem 3.5 we obtain (i).

Now from the \(\sigma \)-Fatou property in E it follows that \(E\equiv E^{\times \times }\) [28, Ch. 15, §71, Thm. 1]. Therefore from (i) we have (ii). \(\square \)

Let us fix \(g\in E^\times \). Then the function \(\varphi _g:E\rightarrow \mathbb {K}\) defined by

$$\begin{aligned} {\varphi _g(f):=\int _\Omega gfd\mu }, \end{aligned}$$
(3.8)

is a linear and bounded functional such that \(\Vert \varphi _g\Vert =\Vert g\Vert _{E^\times }\). Thus we consider the operator

$$\begin{aligned} R:E^\times \rightarrow E^* \text{ defined } \text{ by } R(g):=\varphi _g. \end{aligned}$$
(3.9)

Clearly R is a linear isometry, called canonical isometry. Accordingly, the associate space \(E^\times \) can be identified with a certain closed subspace of \(E^*\). The canonical isometry also preserves the order in the sense that

\(g \ge 0\) if, and only if, \(\varphi _g \ge 0\).

In the case \(\mathbb {K}=\mathbb {C}\) we also have that g is real if, and only if, \(\varphi _g\) is real. Therefore if R is onto, then R is an order isometry. Hence in what follows we will write \(E^\times = E^*\) to mean that the canonical isometry R is onto.

Next we distinguish two necessary conditions for \(E^\times =E^*\) to hold. We will need the following result, which is obtained from [29, Thm. 102.3, p. 415].

Lemma 3.7

If E is a Banach lattice, then \(E^*\) has the Fatou property.

We also need to recall that a functional \(\varphi \in E^*\) is \(\sigma \) -order continuous whenever \(f_n\downarrow 0\) implies \(\varphi (f_n)\rightarrow 0\).

Proposition 3.8

Let E be a saturated \(\mu \)-B.f.s. If \(E^*=E^\times \), then E is order continuous and \(E^\times \) has the Fatou property.

Proof

Since E is a Dedekind \(\sigma \)-complete Banach lattice, we only have to show that E is \(\sigma \)-order continuous. And so, by [29, Lemma 84.1, Thm. 102.7] it is enough to establish that \(\varphi \) is \(\sigma \)-order continuous for any \(\varphi \in (E^*)^+\). So, take \(\varphi \in E^*\) such that \(\varphi \ge 0\) and consider \(\{f_n\}\subset E\) satisfying that \(f_n\downarrow 0\). Since \(\varphi \) is positive, there exists \(g\in (E^\times )^+\) such that

$$\begin{aligned} \varphi (f)=\int _\Omega gf d\mu , \quad \forall \ f\in E. \end{aligned}$$

Then \(\{gf_n\}\subset L^1(\mu )\) is a decreasing sequence such that \(0\le gf_n\). Since the space \(L^1(\mu )\) is Dedekind \(\sigma \)-complete, there exists \(0\le h=\inf _n gf_n\). Let \(A:=\mathrm{supp}g\). It is clear that \(h\chi _{\Omega \setminus A}=0\). Taking \(\frac{0}{0}:=0\), we have that

$$\begin{aligned} \frac{h\chi _A}{g}\le f_n\chi _A\le f_n, \quad \forall \ n\in \mathbb {N}. \end{aligned}$$

It follows that \(\frac{h\chi _A}{g}\in E\) and from \(f_n\downarrow 0\) we have that \(\frac{h\chi _A}{g}\le 0\). Hence \(h\chi _A=0\) and then \(gf_n\downarrow 0\). Therefore \( { \varphi (f_n)=\int _\Omega gf_n d\mu \downarrow 0}\). So \(\varphi \) is \(\sigma \)-order continuous.

The other affirmation follows from the above lemma. \(\square \)

4 Locally \(\sigma \)-finite measure defined on a \(\delta \)-ring

Let us assume now that the \(\sigma \)-algebra \(\Sigma \) that we have been considering is given as \(\Sigma =\mathcal {R}^{loc}\), where \(\mathcal {R}\) is a \(\delta \)-ring and \(\mu = |\lambda |\), where \(\lambda :\mathcal {R}\rightarrow [0,\infty ]\) is a measure on \(\mathcal {R}\). Noting that for any \(A\in \mathcal {R}^{loc}\) with \(|\lambda |(A)>0\) we can find \(B\in \mathcal {R}_A\) with \(\lambda (B)>0\), next we give a simple sufficient condition for E to be saturated.

Lemma 4.1

Let E be a \(|\lambda |\)-B.f.s. If \(S(\mathcal {R})\subset E\), then E is saturated.

Remark 4.2

When \(S(\mathcal {R})\subset E\), the space E is a B.f.s. with respect to \((\Omega ,\mathcal {R},\lambda )\) in the sense introduced by Delgado in [9, Def. 3.1]. Thus these class of spaces are always saturated.

Let \(\nu :\mathcal {R}\rightarrow X\) be a vector measure having \(\lambda :\mathcal {R}\rightarrow [0,\infty ]\) as a local control measure. For \(1\le p<\infty \), the spaces \(L^p_w(\nu )\) and \(L^p(\nu )\) are defined by

$$\begin{aligned} L^p_w(\nu ):=\left\{ f\in L^0(\nu ):|f|^p\in L^1_w(\nu )\right\} \quad \text{ and }\quad L^p(\nu ) :=\left\{ f\in L^0(\nu ): |f|^p\in L^1(\nu )\right\} . \end{aligned}$$

Each function in \(L^p_w(\nu )\) is called weakly p -integrable with respect to \(\nu \) and each function in \(L^p(\nu )\) is called p -integrable with respect to \(\nu \). Note that \(L^p(\nu )\subset L^p_w(\nu )\). Moreover, \(L^p_w(\nu )\) and \(L^p(\nu )\) are \(|\lambda |\)-B.f.s. with norm

$$\begin{aligned} \Vert f\Vert _{p,\nu }:=\Vert |f|^p\Vert _\nu ^\frac{1}{p}=\sup _{x^*\in B_{X^*}}\left( \int _\Omega |f|^pd|\langle \nu ,x^*\rangle |\right) ^{\frac{1}{p}}, \quad \forall \ f\in L^p_w(\nu ). \end{aligned}$$

Also \(S(\mathcal {R})\) is a dense subspace of \(L^p(\nu )\), the space \(L^p(\nu )\) is \(\sigma \)-order continuous and \(L^p_w(\nu )\) has the \(\sigma \)-Fatou property [17, p. 37].

From the above lemma and Theorem 3.5 we obtain:

Proposition 4.3

Let \(1\le p<\infty \). Then \(L^p(\nu )\) is saturated. Thus, \(L^p(\nu )^\times \), with norm \(\Vert \cdot \Vert _{\nu ^\times }:=\Vert \cdot \Vert _{L^p(\nu )^\times }\), is a \(|\lambda |\)-B.f.s. with the \(\sigma \)-Fatou property. If in addition the measure \(\lambda \) is \(\sigma \)-finite, then \(L^p(\nu )^\times \) has the Fatou property.

Remark 4.4

Since \(L^1(\nu )\) is always saturated with respect to any local control measure for \(\nu \), by Lemma 3.1 we have that \(L^1(\nu )\) is order dense in \(L^0(|\lambda |)\). Using other methods, this result was established by Calabuig et al. [4, 4.2].

Example 4.5

Given a vector measure \(\nu \), let us consider its variation \(|\nu |\) as a local control measure. Then \(L^1(\nu )^\times \) is a \(|\nu |\)-B.f.s. with the \(\sigma \)-Fatou property. It may happen that the range of \(|\nu |\) is \(\{0,\infty \}\). For instance, if \(\Sigma \) is the Lebesgue \(\sigma \)-algebra on [0, 1], then the function \(\nu :\Sigma \rightarrow L^2([0,1])\) defined by \(\nu (A):=\chi _A\) is a vector measure whose range is \(\{0,\infty \}\) [5, p. 57]. In this case \(L^1(|\nu |)=\{0\}\) and so \( L^1(\nu )^\times =\{g\in L^0(|\nu |): gf=0, \ \forall \ f\in L^1(\nu )\}=\{0\}\). Thus clearly the space \(L^1(\nu )^\times \) is not saturated. Then in this situation the study of the associate space will not give interesting information.

As we have just seen, when the measure involved is not \(\sigma \)-finite the associate space is not necessarily saturated. This motivates to look for a class of measures for which this problem does not occur. In this direction, let us recall the following definition, introduced by Brooks and Dinculeanu [3, p. 162].

Definition 4.6

A measure \(\lambda :\mathcal {R}\rightarrow [0,\infty ]\) is locally \(\sigma \) -finite, if for each \(B\in \mathcal {R}\), there exists \(\{B_n\}_{n\in \mathbb {N}}\subset \mathcal {R}\) such that \(B={\bigcup _{n=1}^\infty B_n}\) and \(\lambda (B_n)<\infty \), \(\forall \) \(n\in \mathbb {N}\).

Clearly any positive \(\sigma \)-finite measure on a \(\sigma \)-algebra is locally \(\sigma \)-finite.

Example 4.7

Let us consider an uncountable set \(\Gamma \). Let \(\mathcal {R}:=\{B\subset \Gamma : B \text{ is } \text{ finite }\}\) and \(\lambda :\mathcal {R}\rightarrow [0,\infty ]\) be the counting measure. Then \(\mathcal {R}\) is a \(\delta \)-ring and \(\lambda \) is a locally \(\sigma \)-finite measure which is not \(\sigma \)-finite.

Remark 4.8

Consider a locally \(\sigma \)-finite measure \(\lambda :\mathcal {R}\rightarrow [0,\infty ]\) and a \(|\lambda |\)-B.f.s. E. Let us take \(A\in \mathcal {R}^{loc}\) and assume that

$$\begin{aligned} {A:=\bigcup _{n=1}^\infty B_n\cup N} \text{ with } \{B_n\}\subset \mathcal {R}, \ \lambda (B_n)<\infty , \quad \forall \ n\in \mathbb {N} \text{ and } N\in \mathcal {N}_0(|\lambda |). \end{aligned}$$
(4.1)

Hence \((A,(\mathcal {R}_A)^{loc},|\lambda _A|)\) is a \(\sigma \)-finite measure space, where \((\mathcal {R}_A)^{loc}\) is the \(\sigma \)-algebra related to \(\mathcal {R}_A\) and \(\lambda _A\) is the restriction of \(\lambda \) to \(\mathcal {R}_A\). It follows that the space \(E_A\) defined in (3.2), with norm \(\Vert \cdot \Vert _A\), is a saturated \(|\lambda _A|\)-B.f.s. As in this case \(|\lambda _A|\) is \(\sigma \)-finite, we have that \(E_A^\times \), with norm \(\Vert \cdot \Vert _{E^\times _A}\), is a saturated \(|\lambda _A|\)-B.f.s.

Note that if \(B\in \mathcal {R}\), then B has the form (4.1) and in this case \(\mathcal {R}_B=(\mathcal {R}_B)^{loc}\). Hence \((B,\mathcal {R}_B,\lambda _B)\) is a \(\sigma \)-finite measure space.

We now show that the problem of having nonsaturated associate spaces does not appear when we work with a locally \(\sigma \)-finite measure.

Theorem 4.9

Let E be a saturated \(|\lambda |\)-B.f.s. If the measure \(\lambda \) is locally \(\sigma \)-finite, then \(E^\times \) is a saturated \(|\lambda |\)-B.f.s.

Proof

Let \(A\in \mathcal {R}^{loc}\) be such that \(|\lambda |(A)>0\) and consider \(B\in \mathcal {R}_A\) satisfying \(\lambda (B)>0\). Since \(E_B\), defined in (3.2), is a saturated \(\lambda _B\)-B.f.s. and \(\lambda _B(B)>0\) there exists \(C\in \mathcal {R}_B\) with \(0<\lambda _B(C)=\lambda (C)\) and \(\chi _C\in E_B^\times \). Now take \(f\in E\). Then \(f_B\in E_B\) and \( {\int _\Omega |f|\chi _Cd|\lambda |=\int _B|f_B|\chi _Cd|\lambda _B|<\infty }\). Hence \(\chi _C\in E^\times \). \(\square \)

Hereafter, to the condition that E be a saturated \(|\lambda |\)-B.f.s. we will add that of \(\lambda :\mathcal {R}\rightarrow [0,\infty ]\) being always a locally \(\sigma \)-finite measure and sometimes we will omit it explicitly. In the \(\sigma \)-finite case it is well known that \(E\equiv E^{\times \times }\) when E has the \(\sigma \)-Fatou property. By using this fact we will establish the corresponding result in the more general context we are discussing.

Theorem 4.10

If E has the Fatou property, then \(E\equiv E^{\times \times }\).

Proof

As we have that \(E\subset E^{\times \times }\) and \(\Vert f\Vert _{E^{\times \times }}\le \Vert f\Vert _E\), \(\forall \ f\in E\), it only rests to prove the another contention and the other norm inequality. For this, it is enough to establish the conclusion only for non-negative functions.

Let \(0\le f\in E^{\times \times }\). Fix \(B\in \mathcal {R}\). Since E has the Fatou property we have that \(E_B\) is also a \(\lambda _B\)-B.f.s. with the Fatou property. Moreover, as established in Remark 4.8, \(\lambda _B\) is a \(\sigma \)-finite measure. Then we obtain \(E_B\equiv E^{\times \times }_B\) [28, Ch. 15, §71, Thm. 1]. Denote by \(f_B\) the restriction of f to B. Hence \(f_B\in E_B\) and \(\Vert f_B\Vert _B=\Vert f_B\Vert _{E^{\times \times }_B}\). Noting that \((f_B)^\Omega =f\chi _B\), we have \(f\chi _B\in E\) and \(\Vert f\chi _B\Vert _E=\Vert f\chi _B\Vert _{E^{\times \times }}\).

On the other hand, as \(\mathcal {R}\) is a directed set with the order given by \( B\le C \ \text{ if, } \ B\subset C, \ \forall B,C\in \mathcal {R}\), we can consider the net \(\{f\chi _B\}_{B\in \mathcal {R}}\subset E\). Then \(\{f\chi _B\}_{B\in \mathcal {R}}\) is an upwards directed system and \(\Vert f\chi _B\Vert _E=\Vert f\chi _B\Vert _{E^{\times \times }}\le \Vert f\Vert _{E^{\times \times }}\), \(\forall \ B\in \mathcal {R}\). Since E has the Fatou property, there exists \(h\in E\subset E^{\times \times }\) with \({h=\sup _{B\in \mathcal {R}} f\chi _B}\) and

$$\begin{aligned} {\Vert h\Vert _E=\sup _{B\in \mathcal {R}}\Vert f\chi _B\Vert _E=\sup _{B\in \mathcal {R}}\Vert f\chi _B\Vert _{E^{\times \times }}\le \Vert f\Vert _{E^{\times \times }}}. \end{aligned}$$
(4.2)

Assume that there exists \(A\in \mathcal {R}^{loc}\) such that \(h\chi _A<f\chi _A\) and \(|\lambda |(A)>0\). Then for some \(B\in \mathcal {R}_A\) with positive measure we have that \(h\chi _B<f\chi _B\), which is a contradiction and it follows that \(f\in E\). Moreover, as \(f\chi _B\le f\), \(\forall \ B\in \mathcal {R}\) we have that \(h\le f\), \(\lambda \)-a.e. Therefore \(h=f\), \(|\lambda |\)-a.e. The remaining inequality between the norms follows from (4.2). \(\square \)

It is well known that if \(\mu \) is a \(\sigma \)-finite measure and E is a \(\mu \)-B.f.s., then we can write \(L^1 (\mu ) = \{f g: f \in E, g \in E^\times \}\). Next we will show that this result remains valid when we consider a locally \(\sigma \)-finite measure.

Proposition 4.11

Let \(\lambda \) be a locally \(\sigma \)-finite measure and E be a saturated \(|\lambda |\)-B.f.s. If \(h\in L^1(\lambda )\):

  1. (i)

    then for each \(\varepsilon >0\) there exist \(f\in E\) and \(g\in E^\times \) such that

    $$\begin{aligned} h=fg \quad \text{ and } \quad {\Vert f\Vert _E\Vert g\Vert _{E^\times }\le (1+\varepsilon )\int _\Omega |h|d|\lambda |.} \end{aligned}$$
  2. (ii)

    if in addition E has the \(\sigma \)-Fatou property, then there exist \(f\in E\) and \(g\in E^\times \) such that

    $$\begin{aligned} h=fg \quad \text{ and } \quad {\Vert f\Vert _E\Vert g\Vert _{E^\times }=\int _\Omega |h|d|\lambda |.} \end{aligned}$$

Proof

If \(h=0\), the conclusion is clear. Assume that \(h\ne 0\). Since \(h\in L^1(\lambda )\), we have that \({A:=\mathrm{supp}h=\bigcup _{n=1}^\infty B_n\cup N}\), where \(\{B_n\}\subset \mathcal {R}\) and \(N\in \mathcal {N}_0(\lambda )\). As \(\lambda \) is a locally \(\sigma \)-finite measure we can assume that \(\lambda (B_n)<\infty \), \(\forall \ n\in \mathbb {N}\). Then \(\lambda _A\) is \(\sigma \)-finite and \(h_A\in L^1(\lambda _A)\).

(i) Given \(\varepsilon >0\), [15, Thm. 1, (ii)] there exist \(\tilde{f}\in E_A\) and \(\tilde{g}\in E^\times _A\) such that

$$\begin{aligned} h_A=\tilde{f}\tilde{g} \quad \text{ and } \quad {\Vert \tilde{f}\Vert _{E_A}\Vert \tilde{g}\Vert _{E^\times _A}\le (1+\varepsilon )\int _A|h_A|d|\lambda _A|.} \end{aligned}$$

(ii) Since E has the \(\sigma \)-Fatou property it follows that \(E_A\) also has it. From [15, Thm. 1i)] we get \(\tilde{f}\in E_A\) and \(\tilde{g}\in E^\times _A\) such that

$$\begin{aligned} h_A=\tilde{f}\tilde{g} \quad \text{ and } \quad {\Vert \tilde{f}\Vert _{E_A}\Vert \tilde{g}\Vert _{E^\times _A}=\int _A|h_A|d|\lambda _A|.} \end{aligned}$$

As \(h=h_A\chi _A\), by taking \(f:=\tilde{f}^\Omega \in E\) and \(g:=\tilde{g}^\Omega \in E^\times \), the conclusion follows. \(\square \)

In the \(\sigma \)-finite case we know that \(E^\times =E^*\) if, and only if, E is \(\sigma \)-order continuous [28, Ch. 15, §72, Thm. 5]. We proved in Proposition 3.8 that if \(E^\times =E^*\), then E is \(\sigma \)-order continuous and \(E^\times \) has the Fatou property. Now we will show that the converse also is valid in our context. For this let us recall that an ideal Y of a Banach lattice X is a band whenever, for every subset D of Y possessing a supremum in X, this supremum is already in Y.

Theorem 4.12

If E is \(\sigma \)-order continuous, then the following properties are equivalent:

  1. (i)

    \(E^\times =E^*\).

  2. (ii)

    \(E^\times \) is a band of \(E^*\).

  3. (iii)

    \(E^\times \) has the Fatou property.

  4. (iv)

    \(E^\times \) has the weak Fatou property.

Proof

The implications (i) \(\Rightarrow \) (ii) and (iii) \(\Rightarrow \) (iv) are clear. (ii) \(\Rightarrow \) (iii) Let \(\{g_\tau \}\subset E^\times \) be an upwards directed system such that \(\sup _\tau \Vert g_\tau \Vert _{E^\times }<\infty \). Let us take \(\varphi _\tau :=R(g_\tau )\). Then \(\{\varphi _\tau \}\subset E^*\) is an upwards directed system such that \(\sup _\tau \Vert \varphi _\tau \Vert <\infty \). From Lemma 3.7 we have that \(E^*\) has the Fatou property. Thus there exists \(\varphi \in E^*\) such that \(\varphi _\tau \uparrow \varphi \) and \(\Vert \varphi \Vert =\sup _\tau \Vert \varphi _\tau \Vert \). Now since \(\{\varphi _\tau \}\subset R(E^\times )\) and \(R(E^\times )\) is a band of \(E^*\) we have that \(\varphi \in R(E^\times )\). Let \(g\in E^\times \) be such that \(\varphi =R(g)\). Since R is an order isometry \(g_\tau \uparrow g\) and \(\Vert g\Vert _{E^\times }=\sup _\tau \Vert g_\tau \Vert _{E^\times }\).

(iv) \(\Rightarrow \) (i) Now let us assume that \(E^\times \) has the weak Fatou property. First note that since \(E^*\) is a Banach lattice and R is a linear operator it is enough to represent only the positive functionals. Take \(0\le \varphi \in E^*\). Consider the \(\delta \)-ring \(\mathcal {R}(E):=\{B\in \mathcal {R}: \chi _B\in E\}\) and define \(m:\mathcal {R}(E)\rightarrow [0,\infty )\) by \( m(B):=\varphi (\chi _B), \ \forall \ B\in \mathcal {R}(E)\). Since E is \(\sigma \)-order continuous and \(\varphi \) is a positive linear functional we have that m is a positive measure.

Fix \(B\in \mathcal {R}(E)\). Then \(\mathcal {R}_B=\mathcal {R}(E)_B\). Let us denote by \(m_B\) the restriction of m to \(\mathcal {R}_B\).Then \(m_B\) is bounded. On the other hand as \(\lambda \) is a locally \(\sigma \)-finite measure, its restriction \(\lambda _B\) to \(\mathcal {R}_B\) is a positive \(\sigma \)-finite measure.

Now let \(A\in \mathcal {R}_B\) with \(\lambda _B(A)=|\lambda |(A)=0\), then \(\chi _A=0\), \(\lambda \)-a.e. Hence \(m_B(A)=0\). By the Radon–Nikodym Theorem [26, p. 121] there exists a unique \(h_B\in L^0(\mathcal {R}_B)\) such that

$$\begin{aligned} \varphi (\chi _A)=m_B(A)=\int _A h_Bd\lambda _B=\int _B h_B\chi _A d\lambda _B, \quad \forall \ A\in \mathcal {R}_B. \end{aligned}$$

Using a standard procedure it follows now that

$$\begin{aligned} \varphi (f)=\int _B h_Bfd\lambda _B, \quad \forall \ f\in E_B. \end{aligned}$$
(4.3)

Let us denote by \(H_B\) the canonical extension of \(h_B\), then \(H_B\in E^\times \), \(\varphi _B(f):=\varphi (f\chi _B)=\int _\Omega fH_Bd|\lambda |\), \(\forall \ f\in E\) and, by the uniqueness of \(h_B\) it turns out that \(H_B\chi _C=H_{B\cap C}, \ \forall \ C\in \mathcal {R}\). Noting that \(\varphi _B\le \varphi \) we have that

$$\begin{aligned} \Vert H_B\Vert _{E^\times }=\Vert \varphi _B\Vert \le \Vert \varphi \Vert <\infty . \end{aligned}$$

This shows that \(\sup _{B\in \mathcal {R}(E)}\Vert H_B\Vert _{E^\times }<\infty \). It turns out that \(\{H_B\}_{B\in \mathcal {R}(E)}\subset E^\times \) is an upwards directed system. Now since \(E^\times \) has the weak Fatou property, there exists \({h=\sup _{B\in \mathcal {R}(E)} H_B\in E^\times }\).

Let \(f\in E^+\). To prove that \(\varphi (f)=\int _\Omega f h d|\lambda |\) first we will show that \(H_B=h\chi _B, \ \forall \ B\in \mathcal {R}(E)\). Fix \(B\in \mathcal {R}(E)\). It is clear that \(H_B\le h\chi _B\). Let us assume that \(H_B<h\chi _B\), so we can take \(C\in \mathcal {R}_B\) with positive measure such that \(H_B(t)<h\chi _B(f), \ \forall \ t\in C\). Define \(k=h\chi _{\Omega \setminus C}+H_B\chi _C\). Now let \(D\in \mathcal {R}(E)\), then

$$\begin{aligned} H_D=H_D\chi _{\Omega \setminus C}+H_{D\cap C}\le h\chi _{\Omega \setminus C}+H_B\chi _C=k, \end{aligned}$$

that is, k is an upper bound of \(\{H_B\}_{B\in \mathcal {R}(E)}\) which contradicts that h is the supremum. We conclude that \(H_B=h\chi _B\).

Since \(fh\in L^1(|\lambda |)\), it follows that \(A:=\mathrm{supp}fh=\bigcup _{n=1}^\infty B_n\cup N\) where \(\{B_n\}\) is a disjoint family of subsets of \(\mathcal {R}\) and N is a \(|\lambda |\)-null set. Observe that since \(\lambda _B\) is \(\sigma \)-finite and \(E_B\) is saturated, \(\forall \ B\in \mathcal {R}\), by [28, Ch. 15, §67, Thm. 4] we can take \(B_n\in \mathcal {R}(E)\). Since E is order continuous, it follows that \(f \chi _A = \sum _{n=1}^\infty f \chi _{B_n}\) in E. Assume that \(\varphi (f\chi _{\Omega \setminus A})>0\). Then \(f\chi _{\Omega \setminus A}>0\), so we can choose \(B\in \mathcal {R}(E)_{\Omega \setminus A}\) such that \(|\lambda |(B)>0\) and \(f\chi _B>0\). Take its corresponding \(H_B\in E^\times \). So, \(0<\varphi (f\chi _B)=\int _\Omega fH_B d|\lambda |=\int _\Omega fh\chi _B d|\lambda |\), but \(B\subset \Omega \setminus \mathrm{supp}fh\), thus \(\int _\Omega fh\chi _B d|\lambda |=0\), which is a contradiction. Therefore \(\varphi (f\chi _{\Omega \setminus A})=0\) and so

$$\begin{aligned} \varphi (f)= & {} {\sum _{n=1}^\infty \varphi (f\chi _{B_n})=\sum _{n=1}^\infty \int _\Omega fH_Bd|\lambda |} \\= & {} {\sum _{n=1}^\infty \int _\Omega fh\chi _Bd|\lambda |=\int _\Omega \sum _{n=1}^\infty f\chi _Bhd|\lambda |=\int _\Omega fhd|\lambda |}. \end{aligned}$$

This show that \(\varphi =\varphi _h\in R(L^1(\nu )^\times )\) and hence the conclusion follows. \(\square \)

Although the above theorem characterizes when \(E^\times =E^*\), up to now we do not know if \(E^\times \) always has the Fatou property. Next we present a situation where this holds. The proof follows from Theorem 4.12 and Proposition 3.3.

Corollary 4.13

Let \(\lambda :\mathcal {R}\rightarrow [0,\infty ]\) a locally \(\sigma \)-finite measure and E be a saturated \(\lambda \)-B.f.s. If E and \(E^\times \) are \(\sigma \)-order continuous, then \(E^\times =E^*\).

Using Theorem 4.12 we can give a characterization of reflexivity as follows.

Theorem 4.14

Let \(\lambda \) be a locally \(\sigma \)-finite measure and E a saturated \(\lambda \)-B.f.s. If E is order continuous then the space E is reflexive if, and only if, \(E=E^{\times \times }\) and \(E^\times \) is \(\sigma \)-order continuous.

Proof

Let us denote by \(R_1:E^\times \rightarrow E^*\) and \(R_2:E^{\times \times }\rightarrow {E^\times }^*\) the corresponding canonical isometries. Then for the adjoint operator of \(R_1\) we have \(R_1^*:E^{**}\rightarrow {E^\times }^*\). First assume that \(E^\times \) is \(\sigma \)-order continuous and \(E=E^{\times \times }\). Then by Corollary 4.13 \(R_1\) is onto and by hypothesis \(E^{\times \times }=E\). So \(E^{\times \times }\) is \(\sigma \)-order continuous and has the \(\sigma \)-Fatou property. It follows that \(E^{\times \times }\) has the Fatou property and so we can apply Theorem 4.12 to obtain that \(R_2\) is onto. Let us see that \(R_1^*j=R_2\) where \(j:E\rightarrow E^{**}\) is the canonical injection. Take \(f\in E\) and \(g\in E^\times \), then

$$\begin{aligned} \langle g,R_1^*j(f)\rangle =\langle R_1(g),j(f)\rangle =\langle f,R_1(g)\rangle ={\int _\Omega fgd|\lambda |}=\langle g,R_2(f)\rangle . \end{aligned}$$

Therefore j is onto, that is, E is reflexive.

Now assume that E is reflexive. To establish that \(E=E^{\times \times }\) it only rests to prove that \(E^{\times \times }\subset E\). Take \(h\in E^{\times \times }\). Since \(R_1\) is an injective linear operator with closed range, it follows that \(R_1^*\) is onto [11, Thm. VI.6.2]. Hence there exists \(\varphi \in E^{**}\) such that \(R_1^*(\varphi )=R_2(h)\). Let \(f\in E\) satisfy \(j(f)=\varphi \). Thus for \(g\in E^\times \) we have

$$\begin{aligned} \langle g,R_1^*(\varphi )\rangle =\langle R_1(g), \varphi \rangle =\langle R_1(g),j(f)\rangle =\langle f,R_1(g)\rangle ={\int _\Omega fgd|\lambda |}. \end{aligned}$$

Hence, \({\int _\Omega fgd|\lambda |=\langle g,R_2(h)\rangle =\int _\Omega hgd|\lambda |}\). Then \(h=f\), \(|\lambda |\)-c.t.p. and \(h \in E\).

We will now prove that \(R_2\) is onto, then by Proposition 3.8 we will obtain that \(E^\times \) is order continuous. Consider \(\varphi \in L^1(\nu )^{**}\), then \(\varphi \circ R_1^{-1}:R_1(E^\times )\rightarrow \mathbb {K}\) is linear and bounded. By the Hahn–Banach Theorem there exists \(\widetilde{\varphi }\in E^{**}\) such that \(\langle \psi ,\widetilde{\varphi }\rangle =\langle \psi ,\varphi \circ R_1^{-1}\rangle \), \(\forall \ \varphi \in R_1(E^\times )\). Let \(f\in E=E^{\times \times }\) be satisfy \(j(f)=\widetilde{\varphi }\). Then for each \(g\in E^\times \) we have

$$\begin{aligned} \langle g,R_2(f)\rangle= & {} \int _\Omega fgd|\lambda |=\langle f,R_1(g)\rangle = \langle R_1(g),j(f)\rangle \\= & {} \langle R_1(g),\widetilde{\varphi }\rangle =\left\langle R_1(g), \varphi \circ R_1^{-1}\right\rangle = \langle g,\varphi \rangle \end{aligned}$$

It follows that \(\varphi =R_2(f)\) and we conclude that \(R_2\) is onto. \(\square \)

Proposition 4.15

Let \(\lambda \) be a locally \(\sigma \)-finite measure and E a saturated \(|\lambda |\)-B.f.s. If \(E^\times \) has the weak Fatou property, then \(E^\times \) is a band of \(E^*\).

Proof

To prove that \(R(E^\times )\) is an ideal of \(E^*\) we can proceed as in implication (iv) \(\Rightarrow \) (i) of Theorem 4.12 only observing that if \(0\le \varphi \le \varphi _g\in R(E^\times )\), then the set function \(m_\varphi :\mathcal {R}(E)\rightarrow [0,\infty )\), defined by \(m_\varphi (B)=\varphi (\chi _B)\) is a positive measure. Now let \(A\subset E^\times \) be a non empty set such that there exists \({\varphi :=\sup _{g\in A} \varphi _g\in E^*}\). We have to prove that \(\varphi \in R(E^\times )\).

Let us note that \(\mathcal {F}:=\{F\subset A: F~\mathrm{is~finite}\}\) is a directed set with the order given by \(F_1\le F_2\) if \(F_1\subset F_2\). For each \(F\in \mathcal {F}\) define \({\varphi _F:=\max _{g\in F} \varphi _g}\). Let us fix \(F_0\in \mathcal {F}\) and take \(\mathcal {F}_0:=\{F\in \mathcal {F}: F_0\subset F\}\). It turns out that \({\sup _{F\in \mathcal {F}_0}\varphi _F=\varphi }\). Then \(0\le \varphi _F-\varphi _{F_0}\le \varphi -\varphi _{F_0}\), \(\forall \ F\in \mathcal {F}_0\). Thus \(\{\varphi _F-\varphi _{F_0}\}_{\mathcal {F}_0}\) is an upwards directed system such that \({\sup _{F\in \mathcal {F}_0}\Vert \varphi _F-\varphi _{F_0}\Vert <\infty }\). Since \(\{\varphi _F-\varphi _{F_0}\}_{\mathcal {F}_0}\subset R(E^\times )\) and R is an order isometry we obtain an upwards directed system \(\{g_F\}_{\mathcal {F}_0}\subset {E^\times }^+\) with \({\sup _{F\in \mathcal {F}_0}\Vert g_F\Vert _{E^\times }}<\infty \). By the weak Fatou property in \(E^\times \), there exists \(g\in E^\times \) such that \(g_F\uparrow g\). Using again that R is an order isometry we have that \(\varphi _F-\varphi _{F_0}\uparrow \varphi _g\). Since \({\sup _{F\in \mathcal {F}_0}\varphi _F=\varphi }\) we have that \(\varphi -\varphi _{F_0}=\varphi _g\). Therefore \(\varphi \in R(E^\times )\). \(\square \)

The following result was established in [29, p. 418]. We obtain it as consequence of the above proposition and Theorem 3.5.

Corollary 4.16

Let E be a saturated \(|\lambda |\)-B.f.s. If \(|\lambda |\) is \(\sigma \)-finite, then \(E^\times \) is a band of \(E^*\).

5 Brooks–Dinculeanu measure

Let \(\nu :\mathcal {R}\rightarrow X\) be a vector measure defined on a \(\delta \)-ring. Since we are interested in providing a representation of the dual space of \(L^1(\nu )\) as its associate space, it is important to know if \(\nu \) has a local control measure which is locally \(\sigma \)-finite. Then, by Theorem 4.9, the associate space of \(L^1(\nu )\), with respect to this local control measure, will be saturated. Let us distinguish this kind of measures.

Definition 5.1

A measure \(\lambda :\mathcal {R}\rightarrow [0,\infty ]\) is a Brooks–Dinculeanu measure for \(\nu \), if \(\lambda \) is a local control measure for \(\nu \) which is locally \(\sigma \)-finite.

Example 5.2

  1. 1.

    Let \(\nu :\Sigma \rightarrow X\) be a vector measure defined on a \(\sigma \)-algebra. If \(\mu :\Sigma \rightarrow [0,\infty )\) is a Rybakov control measure for \(\nu \), then \(\mu \) is a Brooks–Dinculeanu measure for \(\nu \).

  2. 2.

    Let \(\nu :\mathcal {R}\rightarrow X\) be a \(\sigma \)-finite vector measure. Then \(\nu \) has a bounded local control measure \(\lambda :\mathcal {R}\rightarrow [0,\infty )\) [8, Thm. 3.3]. Hence, \(\lambda \) is a Brooks–Dinculeanu for \(\nu \).

Fortunately, it turns out that each vector measure defined in a \(\delta \)-ring has a Brooks–Dinculeanu measure. This result was established by Jiménez Fernández et al. in [16, p. 3]. Given its importance, we will state it below.

Theorem 5.3

If \(\nu :\mathcal {R}\rightarrow X\) is a vector measure, then \(\nu \) has a Brooks–Dinculeanu measure.

Let us define

$$\begin{aligned} \widehat{\mathcal {R}}:=\{B\in \mathcal {R}: \lambda (B)<\infty \}. \end{aligned}$$

It is clear that \(\widehat{\mathcal {R}}\) is a \(\delta \)-ring satisfying that \(\widehat{\mathcal {R}}\subset \mathcal {R}\). Moreover, it turns out that \(\mathcal {R}^{loc}=\widehat{\mathcal {R}}^{loc}\). Now let us show that

$$\begin{aligned} |\langle \nu ,x^*\rangle |=|\langle \widehat{\nu },x^*\rangle |, \quad \forall \ x^*\in X^*, \end{aligned}$$
(5.1)

where \(\widehat{\nu }\) is the restriction of \(\nu \) to \(\widehat{\mathcal {R}}\). By definition we obtain that \(|\langle \widehat{\nu },x^*\rangle |\le |\langle \nu ,x^*\rangle |\). To establish the other inequality let us fix \(x^*\in X^*\) and consider \(A\in \mathcal {R}^{loc}\). Take \(B\in \mathcal {R}_A\). Since \(\lambda \) is locally \(\sigma \)-finite, there exists an increasing sequence, \(\{B_n\}\subset \widehat{\mathcal {R}}\) such that \({B=\bigcup _{n=1}^\infty B_n}\). So,

$$\begin{aligned} {|\langle \nu ,x^*\rangle |(B)}= & {} {\sup _n|\langle \nu ,x^*\rangle |(B_n) =\sup _n|\langle \widehat{\nu },x^*\rangle |(B_n)} \\\le & {} {\sup _{C\in \widehat{\mathcal {R}}_A}|\langle \widehat{\nu },x^*\rangle |(C)=|\langle \widehat{\nu },x^*\rangle |(A).} \end{aligned}$$

It follows that \(|\langle \nu ,x^*\rangle |(A)\le |\langle \widehat{\nu },x^*\rangle |(A)\). Hence we have established (5.1).

From (5.1) we have that

$$\begin{aligned} \Vert f\Vert _\nu =\Vert f\Vert _{\widehat{\nu }}, \quad \forall \ f\in L^0(\mathcal {R}^{loc}). \end{aligned}$$
(5.2)

Thus \(L^1_w(\nu )\equiv L^1_w(\widehat{\nu })\). Since (5.2) is valid, from the density of \(S(\mathcal {R})\) in \(L^1(\nu )\) and of \(S(\widehat{\mathcal {R}})\) in \(L^1(\widehat{\nu })\), to prove that \(L^1(\nu )\equiv L^1(\widehat{\nu })\), it is sufficient to check that \(S(\mathcal {R})\subset L^1(\widehat{\nu })\) and that \(S(\widehat{\mathcal {R}})\subset L^1(\nu )\). Noting that \(\widehat{\mathcal {R}}\subset \mathcal {R}\) the second contention is clear. Now consider \(B\in \mathcal {R}\) and take \(\{B_n\}\subset \widehat{\mathcal {R}}\) satisfying that \(B_n\subset B_{n+1}\) and \({B=\bigcup _{n=1}^\infty B_n}\). Then \(\chi _{B_n}\rightarrow \chi _B\), moreover

$$\begin{aligned} \int _A \chi _{B_n}d\widehat{\nu }=\widehat{\nu }(B_n\cap A)=\nu (B_n\cap A)=\int _A\chi _{B_n}d\nu \rightarrow \int _A\chi _Bd\nu , \quad \forall \ A\in \mathcal {R}^{loc}. \end{aligned}$$

From [8, Prop. 2.3] we have that \(\chi _B\in L^1(\widehat{\nu })\). It follows that \(S(\mathcal {R})\subset L^1(\widehat{\nu })\) and \(I_\nu (s)=I_{\widehat{\nu }}(s)\), \(\forall \ s\in S(\mathcal {R})\). By the continuity of the integration operators we have \(I_\nu =I_{\widehat{\nu }}\). Therefore, we have proven the following result and so, whenever we find it convenient we can work on the \(\delta \)-ring \(\widehat{\mathcal {R}}\) instead of \(\mathcal {R}\).

Lemma 5.4

If \(\lambda :\mathcal {R}\rightarrow [0,\infty ]\) is a Brooks–Dinculeanu measure for a given vector measure \(\nu \), then

  1. (i)

    for each \(x^*\in X^*\) we have that \(|\langle \nu ,x^*\rangle |=|\langle \widehat{\nu },x^*\rangle |\),

  2. (ii)

    \(L^1_w(\nu )\equiv L^1_w(\widehat{\nu })\), \(L^1(\nu )\equiv L^1(\widehat{\nu })\) and \({\int _\Omega fd\nu =\int _\Omega fd\widehat{\nu }}\), \(\forall \ f\in L^1(\nu )\).

Curbera and Ricker established that \(L^p(\nu )^{\times \times }\equiv L^p_w(\nu )\) when a vector measure defined on a \(\sigma \)-algebra and a Rybakov control measure are considered [7, Prop. 2]. We will show that this equality remains true if we consider instead a vector measure defined on a \(\delta \)-ring and a Brooks–Dinculeanu measure. Before it is necessary to establish a useful characterization for the functions in \(L^1_w(\nu )\).

Lemma 5.5

Let \(f\in L^0(\mathcal {R}^{loc})\). Then \(f\in L^1_w(\nu )\) if and only if for each \(B\in \mathcal {R}\), \(f\chi _B\in L^1_w(\nu )\) and \({\sup _{B\in \mathcal {R}}\Vert f\chi _B\Vert _\nu <\infty }\). In this case \({\Vert f\Vert _\nu =\sup _{B\in \mathcal {R}}\Vert f\chi _B\Vert _\nu }\).

Proof

First let us assume that \(f\in L^1_w(\nu )\). Since \(L^1_w(\nu )\) is a Banach lattice we have that \(f\chi _B\in L^1_w(\nu )\) and \(\Vert f\chi _B\Vert _\nu \le \Vert f\Vert _\nu \), \(\forall \) \(B\in \mathcal {R}\). Then, \({\sup _{B\in \mathcal {R}}\Vert f\chi _B\Vert _\nu \le \Vert f\Vert _\nu }<\infty \).

Now assume that \(f\chi _B\in L^1_w(\nu )\), \(\forall \) \(B\in \mathcal {R}\) and \({M:=\sup _{B\in \mathcal {R}}\Vert f\chi _B\Vert _\nu <\infty }\). Let \(x^*\in B_{X^*}\), then

$$\begin{aligned} {\sup _{B\in \mathcal {R}}\int _B |f|d|\langle x^*,\nu \rangle |\le M}. \end{aligned}$$

From Proposition 2.1 we obtain that \(f\in L^1(|\langle x^*,\nu \rangle |)\), \(\forall \) \(x^*\in B_{X^*}\). Thus \(f\in L^1_w(\nu )\) and \(\Vert f\Vert _\nu \le M\). \(\square \)

Although most of the time we will not state it explicitly, in what follows \(\lambda :\mathcal {R}\rightarrow [0,\infty ]\) will be a Brooks–Dinculeanu measure for a given vector measure \(\nu \) and we will consider \(L^p_w(\nu )\) and \(L^p(\nu )\) as Banach function spaces with respect to \(|\lambda |\).

Theorem 5.6

Let \(1\le p<\infty \). Then \(L^p(\nu )^{\times \times }\equiv L^p_w(\nu )\).

Proof

First we prove that \(L^p_w(\nu )\subset L^p(\nu )^{\times \times }\) and \(\Vert f\Vert _{p,\nu ^{\times \times }}\le \Vert f\Vert _{p,\nu }\), \(\forall \ f\in L^p_w(\nu )\). Let \(\varphi \in S(\mathcal {R}^{loc})\) be such that \(\varphi \in L^p_w(\nu )\) and \(B\in \widehat{\mathcal {R}}\). Then \(\varphi \chi _B\in S(\mathcal {R})\subset L^p(\nu )\). By the Hölder inequality, for each \(g\in L^p(\nu )^\times \)

$$\begin{aligned} {\int _B|g\varphi |d|\lambda |\le \Vert g\Vert _{p,\nu ^\times }\Vert \varphi \chi _B\Vert _{p,\nu }\le \Vert g\Vert _{p,\nu ^\times }\Vert \varphi \Vert _{p,\nu }}. \end{aligned}$$

From Proposition 2.1 we have

$$\begin{aligned} {\int _\Omega |g\varphi |d|\lambda |=\sup _{B\in \widehat{\mathcal {R}}}\int _B|g\varphi |d|\lambda |\le \Vert g\Vert _{p,\nu ^\times }\Vert \varphi \Vert _{p,\nu }}. \end{aligned}$$

It follows that \(\varphi \in L^p(\nu )^{\times \times }\) and

$$\begin{aligned} \Vert \varphi \Vert _{p,\nu ^{\times \times }}\le \Vert \varphi \Vert _{p,\nu }. \end{aligned}$$
(5.3)

Now consider \(f\in L^p_w(\nu )\) and take \(\{\varphi _n\}\subset S(\mathcal {R}^{loc})\) with \(0\le \varphi _n\uparrow |f|\). Then, \(\{\varphi _n\}\subset L^p_w(\nu )\). By (5.3), \(\Vert \varphi _n\Vert _{p,\nu ^{\times \times }}\le \Vert \varphi _n\Vert _{p,\nu }\le \Vert f\Vert _{p,\nu }\), \(\forall \) \(n\in \mathbb {N}\). Since \(L^p(\nu )^{\times \times }\) has the \(\sigma \)-Fatou property it turns out that \(f\in L^p(\nu )^{\times \times }\) and \(\Vert f\Vert _{p,\nu ^{\times \times }}\le \Vert f\Vert _{p,\nu }\).

For the other contention let us fix \(B\in \mathcal {R}\) and let \(\nu _B\) be the restriction of \(\nu \) to the \(\sigma \)-algebra \(\mathcal {R}_B\). As \(L^1(\nu _B)\equiv L^1(\nu )_B\), it follows that \(L^p(\nu _B)\equiv L^p(\nu )_B\). Hence we obtain that \(L^p(\nu _B)^{\times \times }\equiv L^p_w(\nu _B)\) [7, Prop. 2].

Take \(f\in L^p(\nu )^{\times \times }\) and let us denote by \(f_B\) its restriction to B, then \(f_B\in L^p_w(\nu _B)\) and \(\Vert f_B\Vert _{p,\nu _B}=\Vert f_B\Vert _{p,\nu ^{\times \times }_B}\le \Vert f\Vert _{p,\nu ^{\times \times }}\). And so, for each \(x^*\in B_{X^*}\)

$$\begin{aligned} {\int _B|f|^pd|\langle \nu ,x^*\rangle |\le \Vert f\Vert _{p,\nu ^{\times \times }}^p}. \end{aligned}$$

From the above proposition we have \(|f|^p\in L^1_w(\nu )\) and \(\Vert |f|^p\Vert _\nu \le \Vert f\Vert ^p_{p,\nu ^{\times \times }}\). Hence \(f\in L^p_w(\nu )\) and \(\Vert f\Vert _{p,\nu }\le \Vert f\Vert _{p,\nu ^{\times \times }}\). \(\square \)

The following result was established in [4, p. 77] by other methods, we obtain it as consequence of the above proposition and Corollary 3.6.

Corollary 5.7

Let \(1\le p<\infty \). If \(\nu :\mathcal {R}\rightarrow X\) is a \(\sigma \)-finite vector measure, then \(L^p_w(\nu )\) has the Fatou property.

From Theorem 5.6 and Proposition 3.3 we obtain:

Corollary 5.8

If \(L^p_w(\nu )\subset L^1(\lambda )\), \(1\le p<\infty \), then \(L^p_w(\nu )\) has the Fatou property.

The sufficiency in the next result was proven in [4, Prop. 5.4]. Since \(L^1(\nu )\) is \(\sigma \)-order continuous we obtain it from Theorems 4.10 and 5.6.

Corollary 5.9

\(L^1(\nu )\) has the Fatou property if, and only if, \(L^1(\nu )= L^1_w(\nu )\).

As in the \(\sigma \)-finite case, we have the next result.

Lemma 5.10

If E and F are \(\mu \)-B.f.s such that \(E\subset F\) and there exists \(a>0\) with \(\Vert f\Vert _F\le a\Vert f\Vert _E\), \(\forall \ f\in E\), then \(F^\times \subset E^\times \) and

$$\begin{aligned} \Vert g\Vert _{E^\times }\le a\Vert g\Vert _{F^\times }, \quad \forall \ g\in F^\times . \end{aligned}$$

Corollary 5.11

\(L^p(\nu )^\times \equiv L^p_w(\nu )^\times \) and \(L^p_w(\nu )^{\times \times } \equiv L^p_w (\nu )\), \(1\le p<\infty \).

Proof

From the above lemma, we have \(L^p_w(\nu )^\times \subset L^p(\nu )^\times \) and \(\Vert g\Vert _{p,\nu ^\times }\le \Vert g\Vert _{L^p_w(\nu )^\times }\), \(\forall \ g\in L^p_w(\nu )^\times \). Now consider \(f\in L^p_w(\nu )\) and \(g\in L^p(\nu )^\times \), from the Hölder inequality and Theorem 5.6 we have that

$$\begin{aligned} {\int _\Omega |gf|d|\lambda |\le \Vert g\Vert _{p,\nu ^\times }\Vert f\Vert _{p,\nu ^{\times \times }}=\Vert g\Vert _{p,\nu ^\times }\Vert f\Vert _{p,\nu }.} \end{aligned}$$

Hence \(g\in L^p_w(\nu )^\times \) and \(\Vert g\Vert _{L^p_w(\nu )^\times }\le \Vert g\Vert _{p,\nu ^\times }\).

The second equality follows from Theorem 5.6. \(\square \)

In Proposition 3.5 we have seen that the associate space of a \(\mu \)-B.f.s. has the Fatou property when \(\mu \) is \(\sigma \)-finite. We will show that this result remains true for certain Brooks–Dinculeanu measures, introduced by Calabuig et al. [4, p. 77].

Definition 5.12

A vector measure \(\nu \) is \(\mathcal {R}\) -decomposable if we can write \(\Omega =\bigcup _{\alpha \in \Delta }\Omega _\alpha \cup N\), where \(N\in \mathcal {N}_0(\nu )\) and \({\{\Omega _\alpha \}_{\alpha \in \Delta }\subset \mathcal {R}}\) is a family of pairwise disjoint sets satisfying that

  1. (a)

    if \(A_\alpha \in \mathcal {R}_{\Omega _\alpha }\), \(\forall \) \(\alpha \in \Delta \), then \({\bigcup _{\alpha \in \Delta }A_\alpha \in \mathcal {R}^{loc}}\), and

  2. (b)

    if \(x^*\in X^*\) and \(N_\alpha \in \mathcal {N}_0(\langle \nu ,x^*\rangle )\), \(\forall \) \(\alpha \in \Delta \), then \({\bigcup _{\alpha \in \Delta }}N_\alpha \in \mathcal {N}_0(\langle \nu ,x^*\rangle )\).

Note that if \(\nu \) is an \(\mathcal {R}\)-decomposable vector measure and \(A\in \mathcal {R}^{loc}\) is such that \(A\cap \Omega _\alpha \in \mathcal {N}_0(\nu )\), \(\forall \ \alpha \in \Delta \), then \(A\cap \Omega _\alpha \) is \(\langle \nu ,x^*\rangle \)-null, \(\forall \ \alpha \in \Delta \) and \(\forall \ x^*\in B_{X^*}\). From b) in the above definition it follows that A is \(\nu \)-null.

Some examples of \(\mathcal {R}\)-decomposable measures are the \(\sigma \)-finite vector measures and the discrete vector measures [4, Lemma 4.6, p. 77]. However there are \(\mathcal {R}\)-decomposable measures which are neither \(\sigma \)-finite nor discrete [4, p. 85].

Proposition 5.13

Let \(\nu :\mathcal {R}\rightarrow X\) be a vector measure, \(\lambda :\mathcal {R}\rightarrow [0,\infty ]\) be a Brooks–Dinculeanu measure for \(\nu \) and E be \(|\lambda |\)-B.f.s. If \(S(\mathcal {R})\subset E\) and \(\nu \) is \(\mathcal {R}\)-decomposable, then \(E^\times \) has the Fatou property.

Proof

Since \(\nu \) is \(\mathcal {R}\)-decomposable, \({\Omega =\bigcup _{\alpha \in \Delta }\Omega _\alpha \cup N}\), where \(N\in \mathcal {N}_0(\nu )\) and \({\{\Omega _\alpha \}_{\alpha \in \Delta }\subset \mathcal {R}}\) is a family of pairwise disjoint sets satisfying (a) and (b) in Definition 5.12. Moreover, since \(\lambda \) is locally \(\sigma \)-finite we can consider that \(\lambda (\Omega _\alpha )<\infty \), \(\forall \alpha \in \Delta \). Let us note that by Lemma 4.1, E is a saturated \(|\lambda |\)-e.f.B.

Let \(I\subset \Delta \) be a countable set and take \({\Omega _I:=\bigcup _{\alpha \in I}\Omega _\alpha }\), \(\mathcal {R}_I:=\mathcal {R}_{\Omega _I}\) and \(\lambda _I\) the restriction of \(\lambda \) to the \(\delta \)-ring \(\mathcal {R}_I\). Then \((\Omega _I,(\mathcal {R}_I)^{loc},|\lambda _I|)\) is a \(\sigma \)-finite measure space and \(E_I:=E_{\Omega _I}\) is a \(|\lambda _I|\)-B.f.s. Then, from Theorem 3.5, we obtain that \(E^\times _I\) is a \(|\lambda _I|\)-B.f.s. whit the Fatou property.

Now let us consider an upwards directed system \(\{g_\tau \}_{\tau \in K}\subset E^\times \) such that \(g_\tau \ge 0\), \(\forall \ \tau \in K\) and \(M:=\sup _{\tau }\Vert g_\tau \Vert _{E^\times }<\infty \). Denoting by \(g_{\tau ,I}\) to the restriction of \(g_\tau \) to \(\Omega _I\), we have that \(\{g_{\tau ,I}\}_{\tau \in K}\subset E^\times _I\) is an upwards directed system, and from (3.4), \(\sup _\tau {\Vert g_{\tau ,I}\Vert _{E^\times _I}\le M<\infty }\). Since \(E^\times _I\) has the Fatou property, it turns out that \(g_I:=\sup _{\tau }g_{\tau ,I}\in E^\times _I\) and

$$\begin{aligned} {\Vert g_I\Vert _{E^\times _I}=\sup _{\tau }\Vert g_{\tau ,I}\Vert _{E^\times _I}\le M}. \end{aligned}$$
(5.4)

In particular, for each \(\alpha \in \Delta \) exists \(g_{\{\alpha \}}\in E^\times _{\{\alpha \}}\) such that \({g_{\{\alpha \}}=\sup _{\tau }g_{\tau ,\{\alpha \}}}\) and

$$\begin{aligned} {\Vert g_{\{\alpha \}}\Vert _{E^\times _{\{\alpha \}}}=\sup _{\tau }\Vert g_{\tau ,\{\alpha \}}\Vert _{E^\times _{\{\alpha \}}}}. \end{aligned}$$

Let us denote by \(g^\alpha \) the canonical extension of \(g_{\{\alpha \}}\) and define \({g:=\sum _{\alpha \in \Delta }g^\alpha }\). As \(\mathcal {N}_0(\nu )=\mathcal {N}_0(|\lambda |)\) and from (a) in Definition 5.12 we have that \(g\in L^0(|\lambda |)\).

Let us prove that \(g=\sup _\tau g_\tau \). Consider \(\alpha \in \Delta \) and \(\tau \in K\). Since \(g_\tau \chi _{\Omega _\alpha }\le g^\alpha \le g\), it follows that \(g_\tau \le g\). Let us assume that \(g'\in L^0(|\lambda |)\) is such that \(g_\tau \le g'\). Then \(g_\tau \chi _{\Omega _\alpha }\le g'\chi _{\Omega _\alpha }\). So, \(g^\alpha \le g'\). Hence \(g\le g'\) and \(g=\sup _\tau g_\tau \).

Finally we will establish that \(g\in E^\times \). Let us fix \(f\in B_E\) and let \(I\subset \Delta \) be a countable set. Note that the canonical extension of \(g_I=\sup _{\tau }g_{\tau ,I}\) is given by \({g^I=\sum _{\alpha \in I} g^\alpha }\) and \(f_I\in B_{E_I}\), where \(f_I\) is the restriction of f to \(\Omega _I\). Moreover, \({\int _\Omega |g^I f|d|\lambda |=\int _{\Omega _I}|g_I f_I|d|\lambda _I|}\). By using the monotone convergence theorem

$$\begin{aligned} {\sum _{\alpha \in I}\int _\Omega |g^\alpha f|d|\lambda |=\int _\Omega |g^If|d|\lambda |=\int _{\Omega _I}|g_I f_I|d|\lambda _I|\le \Vert g_I\Vert _{E^\times _I}}. \end{aligned}$$

From this and (5.4) we obtain that \({\sum _{\alpha \in I}\int _\Omega |g^\alpha f|d\lambda |\le M}\), for each finite subset I of \(\Delta \). Then, there exists a countable set \(J\subset \Delta \) such that \({\int _\Omega |g^\alpha f|d|\lambda |=0}\), \(\forall \ \alpha \in \Delta \setminus J\). This implies that

$$\begin{aligned} {\int _\Omega |gf|d|\lambda |=\sum _{\alpha \in J}\int _\Omega |g^\alpha f|d|\lambda |=\int _\Omega |g^Jf|d|\lambda |\le M}. \end{aligned}$$
(5.5)

We conclude that \(g\in E^\times \); moreover, from lattice property of the norm in \(E^\times \) and from (5.5), we have that \(\Vert g\Vert _{E^\times }=\sup _\tau \Vert g_\tau \Vert _{E^\times }\). \(\square \)

Let us consider the canonical isometry R between \(L^1(\nu )^\times \) and \(L^1(\nu )^*\). When \(\nu \) is a \(\sigma \)-finite vector measure, then \(\nu \) has a bounded local control measure \(\lambda :\mathcal {R}\rightarrow [0,\infty )\) [8, Thm. 3.3]. Since \(L^1(\nu )\) is a \(\sigma \)-order continuous \(|\lambda |\)-B.f.s., then we have \(L^1(\nu )^*=L^1(\nu )^\times \) [28, Ch. 15, §72, Thm. 5]. In what follows we will present other situations where this holds.

Since \(L^1(\nu )\) is an order continuous \(|\lambda |\)-B.f.s. from Theorem 4.12 we obtain the following result.

Corollary 5.14

The following properties are equivalent:

  1. (i)

    \(L^1(\nu )^\times =L^1(\nu )^*\).

  2. (ii)

    \(L^1(\nu )^\times \) is a band of \(L^1(\nu )^*\).

  3. (iii)

    \(L^1(\nu )^\times \) has the Fatou property.

  4. (iv)

    \(L^1(\nu )^\times \) has the weak Fatou property.

The next result is a consequence of Proposition 5.13 and the previous result.

Corollary 5.15

If \(\nu \) is \(\mathcal {R}\)-decomposable, then \(L^1(\nu )^\times =L^1(\nu )^*\).

If E is a real \(\sigma \)-order continuous Banach lattice it is well known that there exists an \(\mathcal {R}\)-decomposable vector measure \(\nu :\mathcal {R}\rightarrow E\) such that E is order isometric to the space \(L^1(\nu )\) [10, Thm. 5]. Then from the above corollary we obtain the following result.

Corollary 5.16

If E is a \(\sigma \)-order continuous Banach lattice, then there exist an \(\mathcal {R}\)- decomposable vector measure \(\nu :\mathcal {R}\rightarrow E\) and a order isometry from \(L^1(\nu )^\times \) onto \(E^*\). More precisely, if T is a lattice isometry from E onto \(L^1(\nu )\), \(\lambda :\mathcal {R}\rightarrow [0,\infty ]\) is a Brooks–Dinculeanu measure for \(\nu \) and \(\varphi \in E^*\), then there exists \(g\in L^1(\nu )^\times \) such that

$$\begin{aligned} \varphi (f)=\int _\Omega (Tf)gd|\lambda |, \quad \forall \ f\in E. \end{aligned}$$

Proof

Note that it only rests to verify that the result mentioned before remains valid in the complex case. Since E is a \(\sigma \)-order continuous Banach lattice, then \(E_\mathbb {R}\) is also a \(\sigma \)-order continuous Banach lattice. Thus, there exist an \(\mathcal {R}\)-decomposable vector measure \(\tilde{\nu }:\mathcal {R}\rightarrow E_\mathbb {R}\) and an onto lattice isometry \(S:L^1(\nu )\rightarrow E_\mathbb {R}\). Now let us define \(\nu :\mathcal {R}\rightarrow E\), by \( \nu (B)=\tilde{\nu }(B), \ \forall \ B\in \mathcal {R}\). It turns out that \(\nu \) is an \(\mathcal {R}\)-decomposable vector measure and \(L^1(\nu )_\mathbb {R}=L^1(\tilde{\nu })\). Let \(T:L^1(\nu )\rightarrow E\) be the canonical extension of S, then T is an onto lattice isometry [25, Lemma 3.8]. \(\square \)

As a consequence of Theorem 4.14 we obtain the following result.

Corollary 5.17

\(L^1(\nu )\) is reflexive if, and only if, \(L^1(\nu )=L^1_w(\nu )\) and \(L^1(\nu )^\times \) is \(\sigma \)-order continuous.

Now from the previous result and Corollary 4.13 we have:

Corollary 5.18

If \(L^1(\nu )\) is reflexive, then \(L^1(\nu )^\times =L^1(\nu )^*\).

If \(1<p<\infty \), Ferrando and Rodríguez established that \(L^p(\nu )^*\) is order continuous when \(\nu \) is defined on a \(\sigma \)-algebra [13, Thm 3.1]. Using the same arguments, it follows that in our context we also have that \(L^p(\nu )^\times \) is order continuous. Then from Corollary 4.13 we have the following result.

Corollary 5.19

\(L^p(\nu )^\times =L^p(\nu )^*\), \(1<p<\infty \).

Since \(L^p(\nu )^\times \) is order continuous and \(L^p(\nu )=L^p_w(\nu )\) if, and only if, \(L^1(\nu )=L^1_w(\nu )\) [17, Prop. 3.1.6], the next result follows from Theorems 4.14 and 5.6. It was proven when \(\nu \) is defined in a \(\sigma \)-algebra by Fernández et al. [12, Cor. 3.10].

Corollary 5.20

Let \(1 < p < \infty \). Then \(L^p(\nu )\) is reflexive if, and only if, \(L^1(\nu )= L^1_w(\nu )\).

Let us fix \(1<p<\infty \). Then Corollary 5.19 implies that each functional in \(L^p(\nu )^*\) has the form \(\varphi _g\), \(g\in L^p(\nu )^\times \). So we can define \(S:L^p(\nu )^{\times \times }\rightarrow L^p(\nu )^{**}\) by

$$\begin{aligned} \langle \varphi _g, S(h)\rangle :=\int _\Omega ghd|\lambda |. \end{aligned}$$

It turns out that S is a linear isometry and we will write \(L^p(\nu )^{\times \times }=L^p(\nu )^{**}\) to indicate that is onto. Let \(R_1:L^p(\nu )^\times \rightarrow L^p(\nu )^*\) and \(R_2:L^p(\nu )^{\times \times }\rightarrow L^p(\nu )^{\times *}\) be the corresponding canonical isometries, then \(S=(R_1^*)^{-1}\circ R_2\). Thus S is onto if, and only if, \(R_2\) is it. So from Theorems 5.6 and 4.12 we have:

Corollary 5.21

Let \(1 < p < \infty \). Then \(L_w^p(\nu )\) has the Fatou property if, and only if, \(L^p_w(\nu )=L^p(\nu )^{**}\).

Remark 5.22

Calabuig, Delgado, Juan and Sánchez-Pérez asked if in general \(L^1_w(\nu )\) always has the Fatou property [4, pp. 77–78]. With respect to this question we have the following. Let \(1<p<\infty \) and notice that \(L^1_w(\nu )\) has the Fatou property if, and only if, \(L^p_w(\nu )\) has it. Then from the previous result we obtain that

\(L^1_w(\nu )\) has the Fatou property if, and only if, \(L^p_w(\nu )=L^p(\nu )^{**}\).