1 Introduction

Let X be a Banach space with (topological) dual \(X^*\), let \((\Omega ,\Sigma )\) be a measurable space and let \(\nu :\Sigma \rightarrow X\) be a (countably additive) vector measure. A \(\Sigma \)-measurable function \(f:\Omega \rightarrow \mathbb {R}\) is called \(\nu \)-integrable if it is \(|x^*\nu |\)-integrable for all \(x^*\in X^*\) and, for each \(A\in \Sigma \), there is \(\int _A f \, d\nu \in X\) such that

$$\begin{aligned} x^*\left( \int _A f \, d\nu \right) =\int _A f\,d(x^*\nu ) \quad \text {for all } x^*\in X^*. \end{aligned}$$

Here \(x^*\nu \) is the signed measure obtained as the composition of \(\nu \) with \(x^*\) and \(|x^*\nu |\) denotes its variation. By identifying functions which coincide except to a \(\nu \)-null set (where \(A\in \Sigma \) is said to be \(\nu \)-null if \(\nu (B)=0\) for every \(B\in \Sigma \) with \(B\subseteq A\)), the set \(L_1(\nu )\) of all (equivalence classes of) \(\nu \)-integrable functions is a Banach lattice with the \(\nu \)-a.e. order and the norm

$$\begin{aligned} \Vert f\Vert _{L_1(\nu )}:=\sup _{x^*\in B_{X^*}}\int _\Omega |f|\,d|x^*\nu |. \end{aligned}$$

Here \(B_{X^*}\) denotes the closed unit ball of \(X^*\). Let us agree to say that \(L_1(\nu )\) is the \(L_1\) space of the vector measure \(\nu \).

Every Banach lattice with order continuous norm and a weak unit is lattice-isometric to the \(L_1\) space of a vector measure, [11, Theorem 8] (cf. [19, Proposition 2.4]). Such a representation is not unique. For instance, the usual space \(L_1[0,1]\) is equal to \(L_1(\nu _i)\) for each one of the following \(X_i\)-valued vector measures \(\nu _i\) defined on the Borel \(\sigma \)-algebra of [0, 1]:

  • \(X_1:={\mathbb {R}}\) and \(\nu _1(A):=\lambda (A)\) (the Lebesgue measure of A);

  • \(X_2:=L_1[0,1]\) and \(\nu _2(A):=\chi _A\) (the characteristic function of A);

  • \(X_3:=c_0\) and \(\nu _3(A):=(\int _A r_n \, d\lambda )_{n\in \mathbb {N}}\), where \((r_n)_{n\in \mathbb {N}}\) is the sequence of Rademacher functions.

The structure of the space \(L_1(\nu )\) can be greatly conditioned by certain properties of \(\nu \). For complete information on these spaces and their important role in Banach lattices and operator theory, we refer the reader to the monograph [33] and the papers [6, 14, 15, 28, 34, 36, 38].

The inclusion map

$$\begin{aligned} \iota _\nu : L_1(|\nu |) \rightarrow L_1(\nu ) \end{aligned}$$

is a well-defined injective lattice-homomorphism, where \(|\nu |\) is the variation of \(\nu \) (see, e.g., [33, Lemma 3.14]). If \(\iota _\nu \) is surjective, then it is a lattice-isomorphism and, moreover, we have \(|\nu |(\Omega )<\infty \). Curbera [12] addressed the question of when the \(L_1\) space of a vector measure is lattice-isomorphic to an AL-space. Recall that a Banach lattice E is said to be an AL-space if its norm satisfies \(\Vert x+y\Vert =\Vert x\Vert +\Vert y\Vert \) whenever \(x,y\in E\) are disjoint, which is equivalent to saying that E is lattice-isometric to the usual space \(L_1(\mu )\) of a non-negative measure \(\mu \) (see, e.g., [3, Theorem 4.27]). It turns out that \(L_1(\nu )\) is lattice-isomorphic to an AL-space if and only if \(\iota _\nu \) is surjective, [12, Proposition 2]. This is also equivalent to the fact that the integration operator of \(\nu \), that is, the norm 1 operator

$$\begin{aligned} I_\nu :L_1(\nu )\rightarrow X,\quad I_\nu (f):=\int _\Omega f \, d\nu \quad \text {for all } f\in L_1(\nu ), \end{aligned}$$

is cone absolutely summing (i.e., the series \(\sum _{n=1}^\infty I_\nu (f_n)\) is absolutely convergent whenever \(\sum _{n\in \mathbb {N}}f_n\) is unconditionally convergent and \(f_n \in L_1(\nu )^+\) for all \(n\in \mathbb {N}\)), [10, Proposition 3.1]. As usual, given a Banach lattice E, we denote by \(E^+\) its positive cone, that is, \(E^+:=\{x\in E: x\ge 0\}\). At this point we should stress that if a Banach lattice is isomorphic (just as a Banach space) to an AL-space, then it is lattice-isomorphic to an AL-space [1] (cf. [16, Proposition 2.1]).

An operator between Banach spaces is said to be Dunford-Pettis (or completely continuous) if it maps weakly null sequences to norm null ones. The space \(L_1(\mu )\) of a non-negative measure \(\mu \) has the Dunford-Pettis property, that is, every weakly compact operator from \(L_1(\mu )\) to an arbitrary Banach space is Dunford-Pettis (see, e.g., [2, Theorem 5.4.5] or [3, Theorem 5.85]). In general, this is not true for the \(L_1\) space of a vector measure. Indeed, reflexive infinite-dimensional Banach spaces fail the Dunford-Pettis property and, as we have already mentioned, spaces like \(\ell _p\) and \(L_p[0,1]\) for \(1<p<\infty \) can be seen as \(L_1\) spaces of a vector measure. On the other side, there are \(L_1\) spaces of a vector measure having the Dunford-Pettis property which are not lattice-isomorphic to an AL-space, like \(c_0\). Curbera showed in [13, Theorem 4] that \(L_1(\nu )\) has the Dunford-Pettis property if \(\nu \) has \(\sigma \)-finite variation and X has the Schur property (i.e., every weakly null sequence in X is norm null). In fact, he proved that:

  1. (i)

    \(L_1(\nu )\) has the positive Schur property whenever X has the Schur property.

  2. (ii)

    If \(L_1(\nu )\) has the positive Schur property and \(\nu \) has \(\sigma \)-finite variation, then \(L_1(\nu )\) has the Dunford-Pettis property (cf. [6, Section 3.2]).

Recall that a Banach lattice E is said to have the positive Schur property if every weakly null sequence in \(E^+\) is norm null. Note that statement (i) can be deduced at once from the fact that \(L_1(\nu )\) has the positive Schur property if and only if the integration operator \(I_\nu \) is almost Dunford-Pettis (i.e., \((I_\nu (f_n))_{n\in \mathbb {N}}\) is norm null for every weakly null sequence \((f_n)_{n\in \mathbb {N}}\) in \(L_1(\nu )^+\)), see [6, Theorem 5.12].

The integration operator is undoubtedly a key point in the theory of \(L_1\) spaces of a vector measure. Note that its properties depend on \(\nu \) rather than on the space \(L_1(\nu )\) itself. For instance, going back to the example at the beginning, we have:

  • \(I_{\nu _1}\) is the functional given by \(I_{\nu _1}(f)=\int _{[0,1]} f \, d\lambda \);

  • \(I_{\nu _2}\) is the identity operator on \(L_1[0,1]\);

  • \(I_{\nu _3}: L_1[0,1]\rightarrow c_0\) is the operator given by \(I_{\nu _3}(f)=(\int _{[0,1]} r_n f \, d\lambda )_{n\in \mathbb {N}}\), which is strictly singular but fails to be weakly compact.

It is known that \(L_1(\nu )\) is lattice-isomorphic to an AL-space whenever \(I_\nu \) is compact (see [30, Theorem 1], cf. [32, Theorem 2.2] and [7, Theorem 3.3]), absolutely p-summing for \(1\le p <\infty \) (see [31, Theorem 2.2]) or, more generally, Dunford-Pettis and Asplund (see [35, Theorem 3.3]). Recall that an operator between Banach spaces is said to be Asplund if it factors through a Banach space which is Asplund (i.e., all of its separable subspaces have separable dual). In particular, \(L_1(\nu )\) is lattice-isomorphic to an AL-space if \(I_\nu \) is Dunford-Pettis and X is Asplund, [7, Theorem 1.3]. This is a partial answer to the following question posed by Okada, Ricker and Rodríguez-Piazza [31]:

Question 1.1

Suppose that \(I_\nu \) is Dunford-Pettis and that X contains no subspace isomorphic to \(\ell _1\). Is \(L_1(\nu )\) lattice-isomorphic to an AL-space?

They showed that this is the case if, in addition, X has an unconditional Schauder basis, [31, Theorem 1.2]. Note that any Banach space with an unconditional Schauder basis and no subspace isomorphic to \(\ell _1\) has separable dual (see, e.g., [2, Theorem 3.3.1]). To the best of our knowledge, Question 1.1 remains open.

In this paper we deal with \(L_1\) spaces of a vector measure with focus on the property of being isomorphic to an AL-space, the positive Schur property and the Dunford-Pettis property. Our aim is twofold: we include new results and we also present alternative proofs of some already known ones which hopefully might led to a better understanding of the theory. The structure of the paper is as follows.

In Sect. 2 we collect some known preliminary facts on \(L_1\) spaces of a vector measure that will be needed later.

In Sect. 3 we revisit the aforementioned positive answer to Question 1.1 for Asplund spaces (Corollary 3.8) and the related result for integration operators which are Dunford-Pettis and Asplund (Corollary 3.11).

In Sect. 4 we show that the positive Schur property of \(L_1(\nu )\) can be characterized by means of a Dunford-Pettis type property with respect to the so-called “vector duality” induced by the integration operator, that is, the continuous bilinear map

$$\begin{aligned} L_1(\nu )\times L_\infty (\nu ) \rightarrow X, \qquad (f,g) \mapsto I_\nu (fg)=\int _\Omega fg \, d\nu \end{aligned}$$

(Theorem 4.3). We also give another proof of the aforementioned result of [13] stating that \(L_1(\nu )\) has the Dunford-Pettis property if it has the positive Schur property and \(\nu \) has \(\sigma \)-finite variation (Corollary 4.5). It seems to be an open question whether the assumption on the variation can be dropped, namely:

Question 1.2

Suppose that \(L_1(\nu )\) has the positive Schur property. Does \(L_1(\nu )\) have the Dunford-Pettis property?

Finally, in Example 4.6 we discuss a class of vector measures \(\nu \) such that \(L_1(\nu )\) has the positive Schur property and the Dunford-Pettis property, but fails to be lattice-isomorphic to an AL-space, among other interesting properties.

2 Preliminaries

All Banach spaces considered in this paper are real. An operator is a continuous linear map between Banach spaces. Given an operator T, its adjoint is denoted by \(T^*\). By a subspace of a Banach space we mean a norm closed linear subspace. Let Z be a Banach space. The norm of Z is denoted by \(\Vert \cdot \Vert _Z\), or simply \(\Vert \cdot \Vert \), and we write \(B_Z:=\{z\in Z: \Vert z\Vert \le 1\}\) (the closed unit ball of Z). The evaluation of \(z^*\in Z^*\) at \(z\in Z\) is denoted by either \(z^*(z)\) or \(\langle z^*,z\rangle \). By a projection from Z onto a subspace \(Y \subseteq Z\) we mean an operator \(P:Z\rightarrow Z\) such that \(P(Z)=Y\) and P is the identity when restricted to Y. The subspace of Z generated by a set \(H \subseteq Z\) is denoted by \(\overline{\textrm{span}}(H)\).

In this section we gather, for the reader’s convenience, some known facts on \(L_1\) spaces of a vector measure. A basic reference on this topic is [33, Chapter 3].

Throughout this section X is a Banach space, \((\Omega ,\Sigma )\) is a measurable space and \(\nu \in \textrm{ca}(\Sigma ,X)\). As usual, we denote by \(\textrm{ca}(\Sigma ,X)\) the set of all countably additive X-valued vector measures defined on \(\Sigma \). The range of \(\nu \) is the set

$$\begin{aligned} {\mathcal {R}}(\nu ):=\{\nu (A): \, A\in \Sigma \}\subseteq X. \end{aligned}$$

The variation and semivariation of \(\nu \) are denoted by \(|\nu |\) and \(\Vert \nu \Vert \), respectively. The family of all \(\nu \)-null sets is denoted by \({\mathcal {N}}(\nu )\). By a Rybakov control measure of \(\nu \) we mean a finite non-negative measure of the form \(\mu =|x^*\nu |\) for some \(x^*\in X^*\) such that \({\mathcal {N}}(\mu ) = {\mathcal {N}}(\nu )\) (see, e.g., [18, p. 268, Theorem 2]). Throughout this section \(\mu \) is a fixed Rybakov control measure of \(\nu \).

2.1 \(L_\infty \) of a vector measure

A function \(f:\Omega \rightarrow \mathbb {R}\) is called \(\Sigma \)-simple if it is a linear combination of functions of the form \(\chi _A\), where \(A\in \Sigma \). Clearly, all \(\Sigma \)-simple functions are \(\nu \)-integrable. The set of all \(\Sigma \)-simple functions is norm dense in \(L_1(\nu )\) (see, e.g., [33, Theorem 3.7(ii)]), so one has

$$\begin{aligned} \overline{I_\nu (L_1(\nu ))} = \overline{\textrm{span}}({\mathcal {R}}(\nu )). \end{aligned}$$
(2.1)

More generally, every \(\nu \)-essentially bounded \(\Sigma \)-measurable function \(f:\Omega \rightarrow \mathbb {R}\) is \(\nu \)-integrable. By identifying functions which coincide \(\nu \)-a.e., the set \(L_\infty (\nu )\) of all (equivalence classes of) \(\nu \)-essentially bounded \(\Sigma \)-measurable functions is a Banach lattice with the \(\nu \)-a.e. order and the \(\nu \)-essential supremum norm \(\Vert \cdot \Vert _{L_\infty (\nu )}\). Of course, \(L_\infty (\nu )\) is equal to the usual spaces \(L_\infty (|\nu |)\) and \(L_\infty (\mu )\). The inclusion map

$$\begin{aligned} j_\nu : L_\infty (\nu ) \rightarrow L_1(\nu ) \end{aligned}$$

is an injective operator. Moreover, it is weakly compact. Indeed, \(j_\nu (B_{L_\infty (\nu )})\) coincides with the order interval \([-\chi _\Omega ,\chi _\Omega ]\) in \(L_1(\nu )\), so it is weakly compact as \(L_1(\nu )\) has order continuous norm (see, e.g., [3, Theorem 4.9]). Hence, \(I_\nu (j_\nu (B_{L_\infty (\nu )}))\) is weakly compact in X. We have the following characterization of relative norm compactness of \({\mathcal {R}}(\nu )\) (see, e.g., [33, Proposition 2.41]):

Proposition 2.1

The following statements are equivalent:

  1. (i)

    \({\mathcal {R}}(\nu )\) is relatively norm compact.

  2. (ii)

    \(I_\nu (j_\nu (B_{L_\infty (\nu )}))\) is norm compact.

2.2 Composition of a vector measure with an operator

We will use several times the following fact (see, e.g., [33, Lemma 3.27]):

Proposition 2.2

Let \(T:X\rightarrow Y\) be an operator between Banach spaces. Then:

  1. (i)

    The composition \({\tilde{\nu }}:=T\circ \nu :\Sigma \rightarrow Y\) is a countably additive vector measure.

  2. (ii)

    Every \(\nu \)-integrable function is \({\tilde{\nu }}\)-integrable.

  3. (iii)

    The inclusion map \(u: L_1(\nu ) \rightarrow L_1({\tilde{\nu }})\) is an operator and \(I_{{\tilde{\nu }}}\circ u=T\circ I_\nu \).

2.3 L-weakly compact sets and the positive Schur property

Let E be a Banach lattice. Given a set \(W \subseteq E\), we denote by \(\textrm{Sol}(W)\) its solid hull, that is, the set of all \(x\in E\) such that \(|x|\le |y|\) for some \(y\in W\). It is known that if W is relatively weakly compact, then every disjoint sequence in \(\textrm{Sol}(W)\) is weakly null (see, e.g., [3, Theorem 4.34]). The set W is said to be L-weakly compact if it is bounded and every disjoint sequence in \(\textrm{Sol}(W)\) is norm null. Every L-weakly compact set is relatively weakly compact (see, e.g., [3, Theorem 5.55]), but the converse does not hold in general. The following result is well-known (see [26, Corollaries 2.3.5 and 3.6.8], [39, Theorem 1.16] and [41, Lemma 3]):

Proposition 2.3

Let E be a Banach lattice. The following statements are equivalent:

  1. (i)

    E has the positive Schur property.

  2. (ii)

    Every disjoint weakly null sequence in E is norm null.

  3. (iii)

    Every disjoint weakly null sequence in \(E^+\) is norm null.

  4. (iv)

    Every relatively weakly compact subset of E is L-weakly compact.

Proposition 2.4 below characterizes L-weakly compact sets in the \(L_1\) space of a vector measure. We first need to introduce some terminology. Given \(f\in L_1(\nu )\), the map \(\nu _f:\Sigma \rightarrow X\) defined by

$$\begin{aligned} \quad \nu _f(A):=I_\nu (f\chi _A)=\int _A f \, d\nu \quad \text {for all } A\in \Sigma \end{aligned}$$

is a countably additive vector measure by the Orlicz-Pettis theorem (see, e.g., [18, p. 22, Corollary 4]). Note that \(\Vert \nu _f\Vert (A)=\Vert f\chi _A\Vert _{L_1(\nu )}\) for all \(A\in \Sigma \). Moreover, \(\nu _f\) is \(\mu \)-continuous, that is, for every \(\varepsilon >0\) there is \(\delta >0\) such that \(\Vert \nu _f(A)\Vert \le \varepsilon \) for every \(A\in \Sigma \) with \(\mu (A)\le \delta \) (see, e.g., [18, p. 10, Theorem 1]). A set \(F \subseteq L_1(\nu )\) is said to be equi-integrable if the set \(\{\nu _f:f\in F\}\) is uniformly \(\mu \)-continuous, that is, for every \(\varepsilon >0\) there is \(\delta >0\) such that

$$\begin{aligned} \sup _{f\in F}\Vert \nu _f(A)\Vert \le \varepsilon \quad \text {for every } A\in \Sigma \text {with }\mu (A)\le \delta . \end{aligned}$$

The following result can be found in [26, Proposition 3.6.2] and [33, Lemma 2.37] within a more general framework.

Proposition 2.4

Let \(F \subseteq L_1(\nu )\) be a set. The following statements are equivalent:

  1. (i)

    F is L-weakly compact.

  2. (ii)

    F is bounded and equi-integrable.

  3. (iii)

    F is approximately order bounded, i.e., for every \(\varepsilon >0\) there is \(\rho >0\) such that \(F \subseteq j_\nu (\rho B_{L_\infty (\nu )})+\varepsilon B_{L_1(\nu )}\).

We refer the reader to the recent works [4, 9, 22] for further results related to the positive Schur property in Banach lattices.

2.4 Characterization of Dunford-Pettis integration operators

Proposition 2.6 below was first proved in [6, Theorem 5.8]. We include here a more direct proof for the reader’s convenience. One part follows the argument used in [13, Theorem 4] to show that \(L_1(\nu )\) has the positive Schur property if X has the Schur property. The following auxiliary lemma will also be used later.

Lemma 2.5

Let \((f_n)_{n\in \mathbb {N}}\) be a sequence in \(L_1(\nu )\) such that the sequence \((\nu _{f_n}(A))_{n\in \mathbb {N}}\) is norm convergent for every \(A\in \Sigma \). Then \((f_n)_{n\in \mathbb {N}}\) is equi-integrable.

Proof

This follows from the Vitali-Hahn-Saks theorem (see, e.g., [18, p. 24, Corollary 10]) applied to the sequence of \(\mu \)-continuous vector measures \((\nu _{f_n})_{n\in \mathbb {N}}\). \(\square \)

Proposition 2.6

The following statements are equivalent:

  1. (i)

    \(L_1(\nu )\) has the positive Schur property and \({\mathcal {R}}(\nu )\) is relatively norm compact.

  2. (ii)

    \(I_\nu \) is Dunford-Pettis.

Proof

(i)\(\Rightarrow \)(ii): Let \(F \subseteq L_1(\nu )\) be a relatively weakly compact set. We will show that \(I_\nu (F)\) is relatively norm compact by checking that for each \(\varepsilon >0\) there is a norm compact set \(K_\varepsilon \subseteq X\) such that \(I_\nu (F) \subseteq K_\varepsilon +\varepsilon B_X\). Fix \(\varepsilon >0\). Since F is approximately order bounded (by Propositions 2.3 and 2.4), there is \(\rho >0\) such that \(F \subseteq j_\nu (\rho B_{L_\infty (\nu )})+\varepsilon B_{L_1(\nu )}\). Therefore, \(I_\nu (F) \subseteq K_\varepsilon +\varepsilon B_{X}\), where \(K_\varepsilon := I_\nu (j_\nu (\rho B_{L_\infty (\nu )}))\) is norm compact by Proposition 2.1.

(ii)\(\Rightarrow \)(i): Since \(j_\nu (B_{L_\infty (\nu )})\) is weakly compact in \(L_1(\nu )\) (see the paragraph preceding Proposition 2.1) and \(I_\nu \) is Dunford-Pettis, the set \(I_\nu (j_\nu (B_{L_\infty (\nu )}))\) is norm compact and so \({\mathcal {R}}(\nu )\) is relatively norm compact (Proposition 2.1). To prove that \(L_1(\nu )\) has the positive Schur property it suffices to check that every weakly convergent sequence is equi-integrable (Propositions 2.3 and 2.4). Let \((f_n)_{n\in \mathbb {N}}\) be a sequence in \(L_1(\nu )\) which converges weakly to some \(f\in L_1(\nu )\). Then for each \(A\in \Sigma \) the sequence \((f_n\chi _A)_{n\in \mathbb {N}}\) converges weakly to \(f\chi _A\) in \(L_1(\nu )\) (bear in mind that the map \(h \mapsto h\chi _A\) is an operator on \(L_1(\nu )\)). Since \(I_\nu \) is Dunford-Pettis, for each \(A\in \Sigma \) the sequence \((I_{\nu }(f_n\chi _A))_{n\in \mathbb {N}}=(\nu _{f_n}(A))_{n\in \mathbb {N}}\) converges in norm to \(I_{\nu }(f\chi _A)=\nu _f(A)\). Now, Lemma 2.5 applies to conclude that \((f_n)_{n\in \mathbb {N}}\) is equi-integrable. \(\square \)

2.5 The “vector duality” induced by the integration operator

The following result (see, e.g., [33, Proposition 3.31]) shows, in particular, that we have a continuous bilinear map

$$\begin{aligned} L_1(\nu )\times L_\infty (\nu )\rightarrow X \end{aligned}$$

defined by

$$\begin{aligned} \quad (f,g) \mapsto I_\nu (fg)=\int _\Omega fg \, d\nu . \end{aligned}$$

Proposition 2.7

Let \(f\in L_1(\nu )\). Then:

  1. (i)

    For every \(g\in L_\infty (\nu )\) the product \(fg\in L_1(\nu )\) and

    $$\begin{aligned} \Vert fg\Vert _{L_1(\nu )}\le \Vert f\Vert _{L_1(\nu )}\Vert g\Vert _{L_\infty (\nu )}. \end{aligned}$$
  2. (ii)

    The norm of f in \(L_1(\nu )\) is

    $$\begin{aligned} \Vert f\Vert _{L_1(\nu )}=\sup _{g\in B_{L_\infty (\nu )}}\Vert I_\nu (fg)\Vert _X. \end{aligned}$$

There are some elements of \(L_1(\nu )^*\) which admit a simple description and are helpful for dealing with the weak topology of \(L_1(\nu )\). By Proposition 2.7, for each \((g,x^*)\in B_{L_\infty (\nu )}\times B_{X^*}\) we can define a functional \(\gamma _{(g,x^*)} \in B_{L_1(\nu )^*}\) by the formula

$$\begin{aligned} \gamma _{(g,x^*)}(f):=x^*(I_\nu (fg))=\int _\Omega fg \, d(x^*\nu ) \quad \text {for all } f\in L_1(\nu ), \end{aligned}$$

and the set

$$\begin{aligned} \Gamma _\nu :=\{\gamma _{(g,x^*)}: \, (g,x^*)\in B_{L_\infty (\nu )}\times B_{X^*}\} \subseteq B_{L_1(\nu )^*} \end{aligned}$$

is norming for \(L_1(\nu )\), that is,

$$\begin{aligned} \Vert f\Vert _{L_1(\nu )}=\sup _{\gamma \in \Gamma _\nu } \gamma (f) \quad \text {for all } f\in L_1(\nu ). \end{aligned}$$
(2.2)

Let \(\sigma (L_1(\nu ),\Gamma _\nu )\) be the (locally convex Hausdorff) topology on \(L_1(\nu )\) of pointwise convergence on \(\Gamma _\nu \), which is weaker than the weak topology. Proposition 2.8 below was first proved in [29, Proposition 17]. It was pointed out in [25, Section 4.7] that it can also be seen as a corollary of the Rainwater-Simons theorem (see, e.g., [20, Theorem 3.134]) and the fact that \(\Gamma _\nu \) is a James boundary for \(L_1(\nu )\) (i.e., the supremum in (2.2) is a maximum) whenever \({\mathcal {R}}(\nu )\) is relatively norm compact. We refer the reader to [6, Section 4] and [8, Section 2] for more information on this topic.

Proposition 2.8

Suppose that \({\mathcal {R}}(\nu )\) is relatively norm compact. Then every bounded and \(\sigma (L_1(\nu ),\Gamma _\nu )\)-convergent sequence in \(L_1(\nu )\) is weakly convergent.

3 Dunford-Pettis integration operators

Let \({\mathcal {A}}\) be an operator ideal. Following [31], a Banach space X is said to be \({\mathcal {A}}\)-variation admissible if for every measurable space \((\Omega ,\Sigma )\) and for every \(\nu \in \textrm{ca}(\Sigma ,X)\) such that \(I_\nu \in {\mathcal {A}}\), we have \(|\nu |(\Omega )<\infty \). The interest of this concept is based on the following result, [31, Proposition 1.1], which provides a tool for proving that \(L_1\) is lattice-isomorphic to an AL-space under some additional assumptions.

Proposition 3.1

Let \({\mathcal {A}}\) be an operator ideal and let X be a Banach space. If X is \({\mathcal {A}}\)-variation admissible, then for every measurable space \((\Omega ,\Sigma )\) and for every \(\nu \in \textrm{ca}(\Sigma ,X)\) such that \(I_\nu \in {\mathcal {A}}\), the inclusion map \(\iota _\nu : L_1(|\nu |) \rightarrow L_1(\nu )\) is a lattice-isomorphism.

The next proposition is elementary.

Proposition 3.2

Let \({\mathcal {A}}\) be an operator ideal and let X and Y be Banach spaces.

  1. (i)

    If X and Y are isomorphic, then X is \({\mathcal {A}}\)-variation admissible if and only if Y is \({\mathcal {A}}\)-variation admissible.

  2. (ii)

    If X is \({\mathcal {A}}\)-variation admissible, then every subspace of X is \({\mathcal {A}}\)-variation admissible.

Proof

Let \((\Omega ,\Sigma )\) be a measurable space.

(i) If \(T:X\rightarrow Y\) is an isomorphism and \(\nu \in \textrm{ca}(\Sigma ,X)\), then we can apply Proposition 2.2 to deduce that a function is \(\nu \)-integrable if and only if it is \({\tilde{\nu }}\)-integrable, where we write \({\tilde{\nu }}:=T\circ \nu \in \textrm{ca}(\Sigma ,Y)\), and that the identity map \(u: L_1(\nu )\rightarrow L_1({\tilde{\nu }})\) is an isomorphism satisfying \(I_{{\tilde{\nu }}}\circ u=T\circ I_\nu \). Hence, \(I_\nu \in {\mathcal {A}}\) if and only if \(I_{{\tilde{\nu }}}\in {\mathcal {A}}\). Moreover, we have \(\Vert T^{-1}\Vert ^{-1}\cdot |\nu |(\Omega ) \le |{\tilde{\nu }}|(\Omega )\le \Vert T\Vert \cdot |\nu |(\Omega )\).

(ii) Let \(Z\subseteq X\) be a subspace and let \(i:Z\rightarrow X\) be the inclusion operator. Fix \(\nu \in \textrm{ca}(\Sigma ,Z)\) such that \(I_\nu \in {\mathcal {A}}\) and define \({\tilde{\nu }}:=i\circ \nu \in \textrm{ca}(\Sigma ,X)\). Then a function is \(\nu \)-integrable if and only if it is \({\tilde{\nu }}\)-integrable, the identity map \(u: L_1(\nu )\rightarrow L_1({\tilde{\nu }})\) is an isometry and \(I_{{\tilde{\nu }}}=i\circ I_\nu \circ u^{-1} \in {\mathcal {A}}\) (Proposition 2.2). Since X is \({\mathcal {A}}\)-variation admissible, we have \(|\nu |(\Omega )=|{\tilde{\nu }}|(\Omega )<\infty \). \(\square \)

In this section we focus on the operator ideal \({\mathcal {A}}_{cc}\) of all Dunford-Pettis operators. Our main goal is to provide a somehow simpler proof of the fact that Asplund spaces are \({\mathcal {A}}_{cc}\)-variation admissible, [7, Theorem 1.3], see Corollary 3.8 below.

Part (i) of the following result was already pointed out in [31]:

Proposition 3.3

Let X be a Banach space.

  1. (i)

    If X is \({\mathcal {A}}_{cc}\)-variation admissible, then X contains no subspace isomorphic to \(\ell _1\).

  2. (ii)

    If every separable subspace of X is \({\mathcal {A}}_{cc}\)-variation admissible, then X is \({\mathcal {A}}_{cc}\)-variation admissible.

Proof

(i) By Proposition 3.2, its suffices to check that \(\ell _1\) is not \({\mathcal {A}}_{cc}\)-variation admissible. Since \(\ell _1\) is infinite-dimensional, the Dvoretzky-Rogers theorem (see, e.g., [17, Theorem 1.2]) ensures the existence of an unconditionally convergent series \(\sum _{n=1}^\infty x_n\) in \(\ell _1\) which is not absolutely convergent. Now, define \(\nu :{\mathcal {P}}(\mathbb {N})\rightarrow \ell _1\) by \(\nu (A):=\sum _{n\in A}x_n\) for all \(A \in {\mathcal {P}}(\mathbb {N})\) (the power set of \(\mathbb {N}\)). Then \(\nu \in \textrm{ca}({\mathcal {P}}(\mathbb {N}),\ell _1)\) satisfies \(|\nu |({\mathcal {P}}(\mathbb {N}))=\sum _{n=1}^\infty \Vert x_n\Vert =\infty \), while \(I_\nu \) is Dunford-Pettis by the Schur property of \(\ell _1\) (see, e.g., [2, Theorem 2.3.6]). Hence, \(\ell _1\) is not \({\mathcal {A}}_{cc}\)-variation admissible.

(ii) Let \((\Omega ,\Sigma )\) be a measurable space and let \(\nu \in \textrm{ca}(\Sigma ,X)\) such that \(I_\nu \) is Dunford-Pettis. Then \({\mathcal {R}}(\nu )\) is relatively norm compact (Proposition 2.6), so it is separable. Hence, the subspace \(Z:=\overline{\textrm{span}}({\mathcal {R}}(\nu ))=\overline{I_\nu (L_1(\nu ))}\) (see (2.1) at page 4) is separable. Then Z is \({\mathcal {A}}_{cc}\)-variation admissible by assumption. Since \(\nu \) takes values in Z and \(I_\nu \) is Dunford-Pettis, we deduce that \(|\nu |(\Omega )<\infty \). \(\square \)

3.1 Schauder decompositions and the variation of a vector measure

Let X be a Banach space. A Schauder decomposition of X is a sequence \((X_n)_{n\in \mathbb {N}}\) of (non-zero) subspaces of X such that each \(x\in X\) can be written in a unique way as a convergent series of the form \(x=\sum _{n=1}^\infty x_n\), where \(x_n\in X_n\) for all \(n\in \mathbb {N}\). In this case, for each \(n\in \mathbb {N}\) there is a projection \(S_n\) from X onto \(X_n\) such that \(x=\sum _{n=1}^\infty S_n(x)\) for all \(x\in X\). For each \(k\in \mathbb {N}\), the operator \(P_k:=\sum _{n=1}^k S_n\) is a projection from X onto the subspace \(\bigoplus _{n=1}^k X_n\), we have \(\sup _{k\in \mathbb {N}}\Vert P_k\Vert <\infty \) and the formula \(|||x|||=\sup _{k\in \mathbb {N}}\Vert P_k(x)\Vert \) defines an equivalent norm on X. Note that \(P_{k'}\circ P_{k}=P_{k}\circ P_{k'}=P_{\min \{k,k'\}}\) for all \(k,k'\in \mathbb {N}\) and \(\Vert P_k(x)-x\Vert \rightarrow 0\) as \(k\rightarrow \infty \) for every \(x\in X\). Of course, if X has a Schauder basis \((e_n)_{n\in \mathbb {N}}\), then the sequence of 1-dimensional subspaces generated by each \(e_n\) is a Schauder decomposition of X.

The following lemma will be a key tool for proving Theorem 3.6 below.

Lemma 3.4

Let X be a Banach space, let \((\Omega ,\Sigma )\) be a measurable space and let \(\nu \in \textrm{ca}(\Sigma ,X)\). Suppose that:

  • \(|\nu |(\Omega )=\infty \) and \({\mathcal {R}}(\nu )\) is relatively norm compact.

  • X has a Schauder decomposition \((X_n)_{n\in \mathbb {N}}\) such that \(|P_k\circ \nu |(\Omega )<\infty \) for all \(k\in \mathbb {N}\), where \(P_k\) is the associated projection from X onto \(\bigoplus _{n=1}^k X_n\).

Then there exist a sequence \((B_j)_{j\in \mathbb {N}}\) of pairwise disjoint elements of \(\Sigma \setminus {\mathcal {N}}(\nu )\), a strictly increasing sequence \((k_j)_{j\in \mathbb {N}}\) in \(\mathbb {N}\) and \(\varepsilon >0\) such that the functions \(f_j:=\frac{1}{\Vert \nu \Vert (B_j)}\chi _{B_j} \in L_1(\nu )\) and the projections \(R_j:=P_{k_{j+1}}-P_{k_{j}}\) satisfy:

  1. (i)

    \(\Vert I_\nu (f_jg)-R_{j}(I_\nu (f_jg))\Vert \le 2^{-j}\) for all \(g\in B_{L_\infty (\nu )}\) and \(j\in \mathbb {N}\).

  2. (ii)

    There is \(j_0\in \mathbb {N}\) such that \(\Vert I_\nu (f_j)-I_\nu (f_{j'})\Vert \ge \varepsilon \) for all distinct \(j,j' \ge j_0\).

Proof

Write \(Q_k:=\textrm{id}_X-P_k\) for all \(k\in \mathbb {N}\) (where \(\textrm{id}_X\) stands for the identity operator on X). Since \(\sup _{k\in \mathbb {N}}\Vert Q_k\Vert <\infty \) and \(\Vert Q_k(x)\Vert \rightarrow 0\) as \(k\rightarrow \infty \) for every \(x\in X\), the sequence \((Q_k)_{k\in \mathbb {N}}\) converges to 0 uniformly on each norm compact subset of X. By renorming, we can assume without loss of generality that \(\Vert P_k\Vert =1\) for all \(k\in \mathbb {N}\).

Since \(|\nu |(\Omega )=\infty \), there is a sequence \((C_l)_{l\in \mathbb {N}}\) of pairwise disjoint elements of \(\Sigma \setminus {\mathcal {N}}(\nu )\) such that \(\sum _{l=1}^\infty \Vert \nu (C_l)\Vert =\infty \), [27, Corollary 2]. Fix \(\rho >2\) and, for each \(l\in \mathbb {N}\), take \(A_l \in \Sigma {\setminus } {\mathcal {N}}(\nu )\) such that \(A_l \subseteq C_l\) and \(\rho \Vert \nu (A_l)\Vert \ge \Vert \nu \Vert (C_l)\). Then \(\sum _{l=1}^\infty \Vert \nu (A_l)\Vert =\infty \) and

$$\begin{aligned} \Vert \nu (A_l)\Vert \ge \rho ^{-1}\Vert \nu \Vert (A_l) \quad \text {for all } l\in \mathbb {N}. \end{aligned}$$
(3.1)

Claim. There exist two strictly increasing sequences \((k_j)_{j\in \mathbb {N}}\) and \((l_j)_{j\in \mathbb {N}}\) in \(\mathbb {N}\) such that for every \(j\in \mathbb {N}\) we have:

(\(\alpha \)):

\(\Vert P_{k_{j}}(I_\nu (\chi _{A_{l_{j+1}}}g))\Vert \le 2^{-j-1}\Vert \nu \Vert (A_{l_{j+1}})\) for all \(g\in B_{L_\infty (\nu )}\);

(\(\beta \)):

\(\Vert Q_{k_{j}}(I_\nu (\chi _{A_{l_j}}g))\Vert \le 2^{-j} \Vert \nu \Vert (A_{l_j})\) for all \(g\in B_{L_\infty (\nu )}\).

Indeed, we proceed by induction. Set \(l_1:=1\) and consider

$$\begin{aligned} K_1:=\big \{I_\nu (\chi _{A_{1}}g): \, g \in B_{L_\infty (\nu )}\big \} \subseteq X. \end{aligned}$$

Since \({\mathcal {R}}(\nu )\) is relatively norm compact, so is \(K_1\) (Proposition 2.1) and therefore we can pick \(k_1\in \mathbb {N}\) such that

$$\begin{aligned} \sup _{x \in K_{1}}\Vert Q_{k_1}(x)\Vert \le \frac{\Vert \nu \Vert (A_{1})}{2}. \end{aligned}$$

Hence, (\(\beta \)) holds for \(j=1\). Suppose now that \(k_N,l_N\in \mathbb {N}\) are already chosen for some \(N\in \mathbb {N}\). Since \({\tilde{\nu }}:=P_{k_N} \circ \nu \) satisfies \(|{\tilde{\nu }}|(\Omega )<\infty \) and \(\sum _{l=1}^\infty \Vert \nu \Vert (A_l)=\infty \), there is \(l_{N+1}\in \mathbb {N}\) with \(l_{N+1}>l_N\) such that

$$\begin{aligned} |{\tilde{\nu }}|(A_{l_{N+1}}) \le 2^{-N-1}\Vert \nu \Vert (A_{l_{N+1}}). \end{aligned}$$
(3.2)

Observe that for each \(g\in B_{L_\infty (\nu )}\) we have

$$\begin{aligned} \Vert P_{k_N}(I_\nu (\chi _{A_{l_{N+1}}}g))\Vert{} & {} {\mathop {=}\limits ^{(*)}} \Vert I_{{\tilde{\nu }}}(\chi _{A_{l_{N+1}}}g)\Vert \le \Vert \chi _{A_{l_{N+1}}}g\Vert _{L_1({\tilde{\nu }})} {\mathop {\le }\limits ^{(**)}} \Vert \chi _{A_{l_{N+1}}}\Vert _{L_1({\tilde{\nu }})} \\{} & {} = \Vert {\tilde{\nu }}\Vert (A_{l_{N+1}}) \le |{\tilde{\nu }}|(A_{l_{N+1}}) {\mathop {\le }\limits ^{(3.2)}} 2^{-N-1}\Vert \nu \Vert (A_{l_{N+1}}), \end{aligned}$$

where \((*)\) and \((**)\) follow from Propositions 2.2 and 2.7(i), respectively. Hence, (\(\alpha \)) holds for \(j=N\). Now, we consider the relatively norm compact subset of X defined by

$$\begin{aligned} K_{N+1}:=\big \{I_\nu (\chi _{A_{l_{N+1}}}g): \, g \in B_{L_\infty (\nu )}\big \} \end{aligned}$$

(apply Proposition 2.1 again) and we choose \(k_{N+1}\in \mathbb {N}\) with \(k_{N+1}>k_N\) such that

$$\begin{aligned} \sup _{x \in K_{N+1}}\Vert Q_{k_{N+1}}(x)\Vert \le 2^{-N-1}\Vert \nu \Vert (A_{l_{N+1}}). \end{aligned}$$

Therefore, (\(\beta \)) holds for \(j=N+1\). This finishes the proof of the Claim.

For each \(j\in \mathbb {N}\), define \(B_j:=A_{l_{j+1}}\) and let \(f_j\) and \(R_j\) be as in the statement. To check property (i), take \(j\in \mathbb {N}\) and \(g\in B_{L_\infty (\nu )}\). Then (\(\alpha \)) and (\(\beta \)) imply

$$\begin{aligned} \Vert I_\nu (f_jg)-R_{j}(I_\nu (f_jg))\Vert= & {} \Vert Q_{k_{j+1}}(I_\nu (f_jg)) + P_{k_{j}}(I_\nu (f_j g))\Vert \\\le & {} \Vert Q_{k_{j+1}}(I_\nu (f_jg))\Vert + \Vert P_{k_{j}}(I_\nu (f_j g))\Vert \\ {}\le & {} \frac{1}{2^{j+1}}+\frac{1}{2^{j+1}}=\frac{1}{2^j}. \end{aligned}$$

Finally, we will check that (ii) holds for an arbitrary \(0<\varepsilon <\rho ^{-1}\). Choose \(j_0\in \mathbb {N}\) large enough such that \(\rho ^{-1}-2^{-j_0}\ge \varepsilon \). Take \(j'>j\ge j_0\) in \(\mathbb {N}\). Then

$$\begin{aligned} \Vert I_\nu (f_j)-I_\nu (f_{j'})\Vert\ge & {} \Vert P_{k_{j+1}}(I_\nu (f_j)-I_\nu (f_{j'}))\Vert \\{} & {} (\text {because } \Vert P_{k_{j+1}}\Vert =1) \\= & {} \Vert I_\nu (f_j)-Q_{k_{j+1}}(I_\nu (f_j))-P_{k_{j+1}}(I_\nu (f_{j'}))\Vert \\\ge & {} \Vert I_\nu (f_j)\Vert -\Vert Q_{k_{j+1}}(I_\nu (f_j))\Vert -\Vert P_{k_{j+1}}(I_\nu (f_{j'}))\Vert \\= & {} \Vert I_\nu (f_j)\Vert -\Vert Q_{k_{j+1}}(I_\nu (f_j))\Vert -\Vert P_{k_{j+1}}(P_{k_{j'}}(I_\nu (f_{j'})))\Vert \\{} & {} (\text {because } P_{k_{j+1}}\circ P_{k_{j'}}=P_{k_{j+1}}) \\\ge & {} \rho ^{-1}-\Vert Q_{k_{j+1}}(I_\nu (f_j))\Vert -\Vert P_{k_{j'}}(I_\nu (f_{j'}))\Vert \\{} & {} (\text {by } (3.1 \text {and } \Vert P_{k_{j+1}}\Vert =1) \\\ge & {} \rho ^{-1}-\frac{1}{2^{j+1}} -\frac{1}{2^{j'+1}} \\{} & {} (\text {by } (\alpha ) \text { and } (\beta ) \text { with } g=\chi _\Omega ) \\> & {} \rho ^{-1}-\frac{1}{2^{j_0}} \ge \varepsilon . \end{aligned}$$

The proof is finished. \(\square \)

3.2 Asplund spaces are \({\mathcal {A}}_{cc}\)-variation admissible

Let \((X_n)_{n\in \mathbb {N}}\) be a Schauder decomposition of a Banach space X. By a block sequence with respect to \((X_n)_{n\in \mathbb {N}}\) we mean a sequence \((x_j)_{j\in \mathbb {N}}\) in X for which there is a sequence \((I_j)_{j\in \mathbb {N}}\) of non-empty finite subsets of \(\mathbb {N}\) such that \(\max (I_j)<\min (I_{j+1})\) and \(x_j\in \bigoplus _{n\in I_j}X_n\) for all \(j\in \mathbb {N}\). We say that \((X_n)_{n\in \mathbb {N}}\) is shrinking if \(\Vert P_k^*(x^*)-x^*\Vert \rightarrow 0\) as \(k\rightarrow \infty \) for every \(x^*\in X^*\), where \(P_k\) is the associated projection from X onto \(\bigoplus _{n=1}^k X_n\). When X has a Schauder basis \((e_n)_{n\in \mathbb {N}}\) and each \(X_n\) is the subspace generated by \(e_n\), then \((X_n)_{n\in \mathbb {N}}\) is shrinking if and only if \((e_n)_{n\in \mathbb {N}}\) is shrinking in the usual sense.

The following fact belongs to the folklore and can be proved as in the case of Schauder bases (see, e.g., [2, Proposition 3.2.7]).

Proposition 3.5

Let \((X_n)_{n\in \mathbb {N}}\) be a Schauder decomposition of a Banach space X. The following statements are equivalent:

  1. (i)

    \((X_n)_{n\in \mathbb {N}}\) is shrinking.

  2. (ii)

    Every bounded block sequence with respect to \((X_n)_{n\in \mathbb {N}}\) is weakly null.

Theorem 3.6

Let X be a Banach space having a shrinking Schauder decomposition \((X_n)_{n\in \mathbb {N}}\) such that \(X_n\) is finite-dimensional for all \(n\in \mathbb {N}\). Then X is \({\mathcal {A}}_{cc}\)-variation admissible. In particular, every Banach space having a shrinking Schauder basis is \({\mathcal {A}}_{cc}\)-variation admissible.

Proof

Let \((\Omega ,\Sigma )\) be a measurable space and let \(\nu \in \textrm{ca}(\Sigma ,X)\) such that \(I_\nu \) is Dunford-Pettis. Then \({\mathcal {R}}(\nu )\) is relatively norm compact (Proposition 2.6). Fix \(k\in \mathbb {N}\) and denote by \(P_k\) the associated projection from X onto \(\bigoplus _{n=1}^k X_n\). Since \(\bigoplus _{n=1}^k X_n\) is finite-dimensional, we have \(|P_k \circ \nu |(\Omega )<\infty \). By renorming, we can assume that \(\Vert P_k\Vert =1\) for all \(k\in \mathbb {N}\).

Suppose, by contradiction, that \(|\nu |(\Omega )=\infty \). Let \((f_j)_{j\in \mathbb {N}}\) and \((R_j)_{j\in \mathbb {N}}\) be as in Lemma 3.4. Since \((I_\nu (f_j))_{j\in \mathbb {N}}\) is not norm convergent (by property (ii) in Lemma 3.4) and \(I_\nu \) is Dunford-Pettis, the sequence \((f_j)_{j\in \mathbb {N}}\) is not weakly convergent in \(L_1(\nu )\). In addition, \(\Vert f_j\Vert _{L_1(\nu )}=1\) for all \(j\in \mathbb {N}\). By Proposition 2.8, there is \(g\in B_{L_\infty (\nu )}\) such that the sequence \((I_\nu (f_j g))_{j\in \mathbb {N}}\) is not weakly null in X. Then \((R_j(I_\nu (f_j g)))_{j\in \mathbb {N}}\) is a bounded block sequence with respect to \((X_n)_{n\in \mathbb {N}}\) which cannot be weakly null, by property (i) in Lemma 3.4. This contradicts that \((X_n)_{n\in \mathbb {N}}\) is shrinking (Proposition 3.5). \(\square \)

The last ingredient of our proof that Asplund spaces are \({\mathcal {A}}_{cc}\)-variation admissible is the following deep result of Zippin [43] (cf. [21, Theorem III.1] and [40]):

Theorem 3.7

(Zippin) Every Banach space having separable dual is isomorphic to a subspace of a Banach space having a shrinking Schauder basis.

Corollary 3.8

Every Asplund space is \({\mathcal {A}}_{cc}\)-variation admissible.

Proof

By Proposition 3.3(ii), it suffices to prove that every Banach space having separable dual is \({\mathcal {A}}_{cc}\)-admissible. Since every Banach space having a shrinking Schauder basis is \({\mathcal {A}}_{cc}\)-variation admissible (Theorem 3.6), the conclusion follows from Theorem 3.7 and Proposition 3.2(ii). \(\square \)

3.3 An application of the Davis-Figiel-Johnson-Pełczyński factorization

We begin by recalling the refinement of the DFJP factorization developed by Lima, Nygaard and Oja in [23]. Let Z and X be Banach spaces, let \(T:Z\rightarrow X\) be a (non-zero) operator and consider the set

$$\begin{aligned} K:=\frac{1}{\Vert T\Vert }\overline{T(B_Z)} \subseteq B_X. \end{aligned}$$

Fix \(a \in (1,\infty )\) and write

$$\begin{aligned} f(a):=\left( \sum _{n=1}^\infty \frac{a^{n}}{(a^{n}+1)^2}\right) ^{1/2}. \end{aligned}$$

For each \(n\in \mathbb {N}\), let \(\Vert \cdot \Vert _n\) be the Minkowski functional of \(K_n:=a^{n/2} K +a^{-n/2} B_X\), that is,

$$\begin{aligned} \Vert x\Vert _n:=\inf \{t>0: x\in tK_n\} \quad \text {for all } x\in X. \end{aligned}$$

The following theorem can be found in [23, Lemmas 1.1 and 2.1, Theorem 2.2], with the exception of part (vi), which can be obtained similarly as for the usual DFJP factorization (see, e.g., [5, §3]).

Theorem 3.9

Under the previous assumptions, the following statements hold:

  1. (i)

    \(Y:=\{x\in X: \ \sum _{n=1}^\infty \Vert x\Vert _n^2 <\infty \}\) is a Banach space with the norm

    $$\begin{aligned} \Vert x\Vert _Y:=\left( \sum _{n=1}^\infty \Vert x\Vert _n^2\right) ^{1/2}. \end{aligned}$$
  2. (ii)

    \(K \subseteq f(a) B_{Y}\) and the identity map \(J: Y \rightarrow X\) is an operator.

  3. (iii)

    T factors as

    (3.3)

    where S is an operator.

  4. (iv)

    J is a norm-to-norm homeomorphism when restricted to K. In fact:

    $$\begin{aligned} \Vert x\Vert _Y^2 \le \Big (\frac{1}{4}+\frac{1}{\ln a}\Big ) \Vert x\Vert \quad \text {for all } x\in K. \end{aligned}$$

    Therefore, if T is Dunford-Pettis, then S is Dunford-Pettis as well.

  5. (v)

    If T is weakly compact, then Y is reflexive.

  6. (vi)

    If T is Asplund, then Y is Asplund.

  7. (vii)

    If a is the unique element of \((1,\infty )\) satisfying \(f(a)=1\), then \(\Vert S\Vert =\Vert T\Vert \) and \(\Vert J\Vert =1\). In this case, (3.3) is called the DFJP-LNO factorization of T.

In [28] the DFJP-LNO factorization was applied to the integration operator of a vector measure. Our next proposition gathers some of the results obtained in [28, Theorems 3.7 and 4.5]:

Proposition 3.10

Let X be a Banach space, let \((\Omega ,\Sigma )\) be a measurable space, let \(\nu \in \textrm{ca}(\Sigma ,X)\) and let

be the DFJP-LNO factorization of \(I_\nu \). Define \({\tilde{\nu }}:\Sigma \rightarrow Y\) by \({\tilde{\nu }}(A):=S(\chi _A)\) for all \(A\in \Sigma \). Then:

  1. (i)

    \({\tilde{\nu }} \in \textrm{ca}(\Sigma ,Y)\), \(\nu =J\circ {\tilde{\nu }}\) and \({\mathcal {N}}(\nu )={\mathcal {N}}({\tilde{\nu }})\).

  2. (ii)

    \(L_1({\tilde{\nu }})=L_1(\nu )\), with \(\Vert f\Vert _{L_1(\nu )}=\Vert f\Vert _{L_1({\tilde{\nu }})}\) for all \(f\in L_1(\nu )\), and \(S=I_{{\tilde{\nu }}}\).

  3. (iii)

    \({\tilde{\nu }}\) has finite (resp., \(\sigma \)-finite) variation whenever \(\nu \) does.

Corollary 3.11

Let X be a Banach space, let \((\Omega ,\Sigma )\) be a measurable space and let \(\nu \in \textrm{ca}(\Sigma ,X)\). If \(I_\nu \) is Asplund and Dunford-Pettis, then \(|\nu |(\Omega )<\infty \) and the inclusion map \(\iota _\nu : L_1(|\nu |) \rightarrow L_1(\nu )\) is a lattice-isomorphism.

Proof

Let Y, J and \({\tilde{\nu }}\) be as in Proposition 3.10. Since \(I_{{\tilde{\nu }}}\) is Dunford-Pettis and Y is Asplund (Theorem 3.9, parts (iv) and (vi)), we can apply Corollary 3.8 to get \(|{\tilde{\nu }}|(\Omega )<\infty \), hence \(|\nu |(\Omega )=|J\circ {\tilde{\nu }}|(\Omega ) \le |{\tilde{\nu }}|(\Omega )<\infty \). This shows that every Banach space is \({\mathcal {A}}\)-variation admissible, where \({\mathcal {A}}\) denotes the operator ideal of all Asplund and Dunford-Pettis operators. The last statement follows from Proposition 3.1. \(\square \)

4 Dunford-Pettis type properties

4.1 A remark on equimeasurability

Let \((\Omega ,\Sigma ,\mu )\) be a finite measure space. A set \(H \subseteq L_\infty (\mu )\) is said to be equimeasurable if for every \(\varepsilon >0\) there is \(A\in \Sigma \) with \(\mu (\Omega {\setminus } A)\le \varepsilon \) such that \(\{h\chi _A:h\in H\}\) is relatively norm compact in \(L_\infty (\mu )\). Theorem 4.1 below is a particular case of [5, Theorem 5.5.4]. We include a direct proof for the sake of completeness.

Theorem 4.1

Let \((\Omega ,\Sigma ,\mu )\) be a finite measure space. If \(H \subseteq L_\infty (\mu )\) is relatively weakly compact, then it is equimeasurable.

Proof

By the Davis-Figiel-Johnson-Pełczyński factorization (see, e.g., [3, Theorem 5.37]), there exist a reflexive Banach space Y and an operator \(T:Y \rightarrow L_\infty (\mu )\) such that \(T(B_Y) \supseteq H\). Let \(i: L_1(\mu ) \rightarrow L_\infty (\mu )^*\) be the inclusion operator and let \(S:=T^*\circ i: L_1(\mu ) \rightarrow Y^*\). Since \(Y^*\) is reflexive, S is representable, that is, there is \(g\in L_\infty (\mu ,Y^*)\) such that

$$\begin{aligned} S(f)=\text {(Bochner)-}\int _\Omega fg \, d\mu \quad \text {for all } f\in L_1(\mu ) \end{aligned}$$

(see, e.g., [18, p. 75, Theorem 12]).

Fix \(\varepsilon >0\). Since g is strongly \(\mu \)-measurable, Egorov’s theorem ensures the existence of \(A\in \Sigma \) with \(\mu (\Omega {\setminus } A)\le \varepsilon \) and a sequence \(g_n:\Omega \rightarrow Y^*\) of \(\Sigma \)-simple \(Y^*\)-valued functions such that

$$\begin{aligned} \Vert g(t)-g_n(t)\Vert \le \frac{1}{n} \quad \text {for every }t\in A \text { and for every } n\in \mathbb {N}. \end{aligned}$$
(4.1)

For each \(n\in \mathbb {N}\), let us consider the operator \(S_n:L_1(\mu )\rightarrow Y^*\) defined by

$$\begin{aligned} S_n(f)=\text {(Bochner)-}\int _A fg_n \, d\mu \quad \text {for all } f\in L_1(\mu ). \end{aligned}$$

Note that \(S_n\) is a finite-rank operator, because \(g_n\) is the sum of finitely many functions of the form \(y^*\chi _B\), where \(y^*\in Y^*\) and \(B\in \Sigma \). Hence, \(S_n\) is compact. Moreover, if \(P_A:L_1(\mu )\rightarrow L_1(\mu )\) is the projection defined by \(P_A(f):=f\chi _A\) for all \(f\in L_1(\mu )\), then the operator \(S\circ P_A: L_1(\mu )\rightarrow Y^*\) satisfies

$$\begin{aligned} \Vert S\circ P_A-S_n\Vert =\sup _{f\in B_{L_1(\mu )}}\left\| \text {(Bochner)-}\int _A f (g-g_n) \, d\mu \right\| {\mathop {\le }\limits ^{(4.1)}} \frac{1}{n}. \end{aligned}$$

It follows that \((S_n)_{n\in \mathbb {N}}\) converges to \(S\circ P_A\) in the operator norm. In particular, \(S\circ P_A\) is compact and, therefore, \((S\circ P_A)^*: Y \rightarrow L_\infty (\mu )\) is compact as well (by Schauder’s theorem).

For every \(y\in Y\) and for every \(f\in L_1(\mu )\) we have

$$\begin{aligned} \langle (S\circ P_A)^*(y),f \rangle = \langle y,(S\circ P_A)(f) \rangle = \langle y,T^*(i(f\chi _A)) \rangle \\ = \langle T(y),f\chi _A \rangle = \int _AfT(y)\, dy=\langle T(y)\chi _A,f \rangle . \end{aligned}$$

Therefore \((S\circ P_A)^*(y)=T(y)\chi _A\) for all \(y \in Y\). It follows that

$$\begin{aligned} \{h\chi _A: \, h\in H\} \subseteq \{T(y)\chi _A: \, y\in B_Y\}= (S\circ P_A)^*(B_Y) \end{aligned}$$

and so \(\{h\chi _A: \, h\in H\}\) is relatively norm compact in \(L_\infty (\mu )\). \(\square \)

4.2 A Dunford-Pettis type property for \(L_1\) of a vector measure

Recall that a Banach space Z has the Dunford-Pettis property if and only if \(z_n^*(z_n)\rightarrow 0\) as \(n\rightarrow \infty \) for all weakly null sequences \((z_n)_{n\in \mathbb {N}}\) and \((z_n^*)_{n\in \mathbb {N}}\) in Z and \(Z^*\), respectively (see, e.g., [2, Theorem 5.4.4]).

We next show that the \(L_1\) space of an arbitrary vector measure enjoys a Dunford-Pettis type property with respect to the “vector duality” induced by the integration operator (Subsection 2.5).

Theorem 4.2

Let X be a Banach space, let \((\Omega ,\Sigma )\) be a measurable space and let \(\nu \in \textrm{ca}(\Sigma ,X)\). Let \((f_n)_{n\in \mathbb {N}}\) be a sequence in \(L_1(\nu )\) and let \((g_n)_{n\in \mathbb {N}}\) be a weakly null sequence in \(L_\infty (\nu )\).

  1. (i)

    If \((f_n)_{n\in \mathbb {N}}\) is weakly null, then \((I_\nu (f_ng_n))_{n\in \mathbb {N}}\) is weakly null.

  2. (ii)

    If \((f_n)_{n\in \mathbb {N}}\) is bounded and equi-integrable, then \((I_\nu (f_ng_n))_{n\in \mathbb {N}}\) is norm null.

Proof

(i) Fix \(x^*\in X^*\). Let \(h_{x^*} \in L_\infty (|x^* \nu |)\) be the Radon-Nikodým derivative of \(x^*\nu \) with respect to \(|x^*\nu |\). For each \(n\in \mathbb {N}\) we have

$$\begin{aligned} x^*\big (I_\nu (f_n g_n)\big )=\int _\Omega f_n g_n \, d(x^*\nu )=\int _\Omega f_n h_{x^*} g_n \, d|x^*\nu |. \end{aligned}$$
(4.2)

Since \((f_n)_{n\in \mathbb {N}}\) is weakly null in \(L_1(\nu )\) and the inclusion map \(L_1(\nu )\rightarrow L_1(|x^*\nu |)\) is an operator, \((f_n)_{n\in \mathbb {N}}\) is weakly null in \(L_1(|x^*\nu |)\) and so the same holds for \((f_n h_{x^*})_{n\in \mathbb {N}}\). In the same way, \((g_n)_{n\in \mathbb {N}}\) is weakly null in \(L_\infty (|x^* \nu |)\), so we can apply the Dunford-Pettis property of \(L_1(|x^*\nu |)\) and (4.2) to conclude that \(x^*(I_\nu (f_n g_n))\rightarrow 0\) as \(n\rightarrow \infty \). Since \(x^*\in X^*\) is arbitrary, \((I_\nu (f_ng_n))_{n\in \mathbb {N}}\) is weakly null.

(ii) Define \(\alpha :=\sup _{n\in \mathbb {N}}\Vert f_n\Vert _{L_1(\nu )}\) and \(\beta :=\sup _{n\in \mathbb {N}}\Vert g_n\Vert _{L_\infty (\nu )}\). Let \(\mu \) be a Rybakov control measure of \(\nu \). Fix \(\varepsilon >0\). Since \(F:=\{f_n: n\in \mathbb {N}\}\) is equi-integrable, we can choose \(\delta >0\) such that

$$\begin{aligned} \sup _{f\in F}\Vert f\chi _B\Vert _{L_1(\nu )}\le \varepsilon \quad \text {for every } B\in \Sigma \text {with } \mu (B)\le \delta . \end{aligned}$$
(4.3)

By Theorem 4.1, the set \(\{g_n:n\in \mathbb {N}\}\) is equimeasurable, so there is \(A\in \Sigma \) with \(\mu (\Omega {\setminus } A)\le \delta \) such that \(\{g_n \chi _A: n\in \mathbb {N}\}\) is relatively norm compact in \(L_\infty (\nu )\). Since the sequence \((g_n \chi _A)_{n\in \mathbb {N}}\) is weakly null in \(L_\infty (\nu )\) (bear in mind that the map \(g \mapsto g\chi _A\) is an operator on \(L_\infty (\nu )\)), we conclude that \((g_n \chi _A)_{n\in \mathbb {N}}\) is norm null in \(L_\infty (\nu )\). Choose \(n_0\in \mathbb {N}\) such that

$$\begin{aligned} \sup _{n\ge n_0} \Vert g_n\chi _A\Vert _{L_\infty (\nu )} \le \varepsilon . \end{aligned}$$
(4.4)

Now, for every \(f\in F\) and for every \(n\in \mathbb {N}\) with \(n\ge n_0\) we have

$$ \begin{aligned} \big \Vert I_\nu (fg_n)\big \Vert{} & {} \le \big \Vert I_\nu (fg_n\chi _{\Omega \setminus A})\big \Vert +\big \Vert I_\nu (f g_n\chi _A)\big \Vert \\{} & {} {\mathop {\le }\limits ^{(\text {Prop.}~2.7(i))}} \Vert f\chi _{\Omega \setminus A}\Vert _{L_1(\nu )} \Vert g_n\Vert _{L_\infty (\nu )}+\Vert f\Vert _{L_1(\nu )} \Vert g_n\chi _A\Vert _{L_\infty (\nu )}\\{} & {} {\mathop {\le }\limits ^{(4.3) \& (4.4)}} (\beta +\alpha )\varepsilon . \end{aligned}$$

As \(\varepsilon >0\) is arbitrary, the sequence \((I_\nu (f_ng_n))_{n\in \mathbb {N}}\) is norm null. \(\square \)

4.3 The positive Schur property as a Dunford-Pettis type property

As a natural outcome of our previous work we get the following characterization:

Theorem 4.3

Let X be a Banach space, let \((\Omega ,\Sigma )\) be a measurable space and let \(\nu \in \textrm{ca}(\Sigma ,X)\). The following statements are equivalent:

  1. (i)

    \(L_1(\nu )\) has the positive Schur property.

  2. (ii)

    For all weakly null sequences \((f_n)_{n\in \mathbb {N}}\) and \((g_n)_{n\in \mathbb {N}}\) in \(L_1(\nu )\) and \(L_\infty (\nu )\), respectively, the sequence \((I_\nu (f_ng_n))_{n\in \mathbb {N}}\) is norm null.

Proof

(i)\(\Rightarrow \)(ii): This follows from Theorem 4.2, because the positive Schur property of \(L_1(\nu )\) is equivalent to the fact that every relatively weakly compact subset of \(L_1(\nu )\) is equi-integrable (Propositions 2.3 and 2.4).

(ii)\(\Rightarrow \)(i): By Propositions 2.3 and 2.4, it suffices to prove that every disjoint weakly null sequence \((f_n)_{n\in \mathbb {N}}\) in \(L_1(\nu )\) is equi-integrable. Let \((A_n)_{n\in \mathbb {N}}\) be a sequence of pairwise disjoint elements of \(\Sigma \) such that \(f_n\chi _{A_n}=f_n\) for all \(n\in \mathbb {N}\). Observe that \((\chi _{A_n})_{n\in \mathbb {N}}\) is weakly null in \(L_\infty (\nu )\). Indeed, we can assume without loss of generality that \(A_n\not \in {\mathcal {N}}(\nu )\) for all \(n\in \mathbb {N}\). Then \((\chi _{A_n})_{n\in \mathbb {N}}\) is a basic sequence in \(L_\infty (\nu )\) which is equivalent to the usual basis of \(c_0\). In particular, \((\chi _{A_n})_{n\in \mathbb {N}}\) is weakly null in \(L_\infty (\nu )\).

Fix \(A\in \Sigma \). Define \({\tilde{f}}_n:=f_n\chi _A\) for all \(n\in \mathbb {N}\). Note that

$$\begin{aligned} I_\nu ({\tilde{f}}_n\chi _{A_n})=I_\nu ({\tilde{f}}_n)=\nu _{f_n}(A) \quad \text {for all } n\in \mathbb {N}. \end{aligned}$$
(4.5)

Since \(({\tilde{f}}_n)_{n\in \mathbb {N}}\) is weakly null in \(L_1(\nu )\) (because \((f_n)_{n\in \mathbb {N}}\) is weakly null and the map \(h \mapsto h\chi _A\) is an operator on \(L_1(\nu )\)) and \((\chi _{A_n})_{n\in \mathbb {N}}\) is weakly null in \(L_\infty (\nu )\), condition (ii) and (4.5) imply that the sequence \((\nu _{f_n}(A))_{n\in \mathbb {N}}\) is norm null. As \(A\in \Sigma \) is arbitrary, we can apply Lemma 2.5 to conclude that \((f_n)_{n\in \mathbb {N}}\) is equi-integrable. \(\square \)

Of course, Theorems 4.2 and 4.3 provide another point of view for the positive Schur property of the \(L_1\) space of a vector measure taking values in a Banach space with the Schur property, [13, proof of Theorem 4].

4.4 Vector measures with \(\sigma \)-finite variation

The analysis of the Dunford-Pettis property is simpler for \(L_1\) spaces of a vector measure with \(\sigma \)-finite variation.

Proposition 4.4

Let X be a Banach space, let \((\Omega ,\Sigma )\) be a measurable space and let \(\nu \in \textrm{ca}(\Sigma ,X)\) with \(\sigma \)-finite variation. If \((f_n)_{n\in \mathbb {N}}\) is a bounded and equi-integrable sequence in \(L_1(\nu )\) and \((\varphi _n)_{n\in \mathbb {N}}\) is a weakly null sequence in \(L_1(\nu )^*\), then \(\varphi _n(f_n)\rightarrow 0\) as \(n\rightarrow \infty \).

Proof

The sequence \((f_n)_{n\in \mathbb {N}}\) is approximately order bounded (Proposition 2.4). Hence, we can assume without loss of generality that \(f_n\in j_\nu (B_{L_\infty (\nu )})\) for all \(n\in \mathbb {N}\). Define \(\alpha :=\sup _{n\in \mathbb {N}}\Vert \varphi _n\Vert _{L_1(\nu )^*}\). Let \((A_m)_{m\in \mathbb {N}}\) be an increasing sequence in \(\Sigma \) such that \(\Omega =\bigcup _{m\in \mathbb {N}}A_m\) and \(|\nu |(A_m)<\infty \) for all \(m\in \mathbb {N}\). Fix \(\varepsilon >0\). Choose \(m\in \mathbb {N}\) large enough such that

$$\begin{aligned} \Vert \nu \Vert (\Omega \setminus A_m)\le \varepsilon . \end{aligned}$$
(4.6)

Define \(\mu (A):=|\nu |(A\cap A_m)\) for all \(A\in \Sigma \), so that \(\mu \) is a finite non-negative measure. Consider the inclusion operator \(\iota : L_1(\mu ) \rightarrow L_1(\nu )\) (see, e.g., [33, Lemma 3.14]) and \(\iota ^*:L_1(\nu )^*\rightarrow L_\infty (\mu )\). Define \(g_n:=\iota ^*(\varphi _n)\in L_\infty (\mu )\) for all \(n\in \mathbb {N}\), so that \((g_n)_{n\in \mathbb {N}}\) is weakly null in \(L_\infty (\mu )\).

The sequence \((f_n\chi _{A_m})_{n\in \mathbb {N}}\) is bounded and equi-integrable in \(L_1(\mu )\) and

$$\begin{aligned} \langle g_n,f_n \chi _{A_m}\rangle = \int _{A_m} f_n g_n \, d\mu =\varphi _n(f_n \chi _{A_m}) \quad \text {for all } n\in \mathbb {N}. \end{aligned}$$

Therefore, the Dunford-Pettis property of \(L_1(\mu )\) (cf. Theorem 4.2(ii)) ensures that \(\varphi _n(f_n \chi _{A_m})\rightarrow 0\) as \(n\rightarrow \infty \). Take \(n_0\in \mathbb {N}\) such that

$$\begin{aligned} |\varphi _n(f_n \chi _{A_m})|\le \varepsilon \quad \text {whenever } n\ge n_0. \end{aligned}$$

Since

$$\begin{aligned} |\varphi _n(f_n\chi _{\Omega \setminus A_m})| \le \alpha \Vert f_n\chi _{\Omega \setminus A_m}\Vert _{L_1(\nu )} \le \alpha \Vert \nu \Vert (\Omega \setminus A_m)\le \alpha \varepsilon \quad \text {for all } n\in \mathbb {N}\end{aligned}$$

(by Proposition 2.7(i) and (4.6)), we have

$$\begin{aligned} | \varphi _n(f_n)|\le |\varphi _n(f_n\chi _{A_m})| + |\varphi _n(f_n\chi _{\Omega \setminus A_m})| \le (1+\alpha )\varepsilon \quad \text {whenever } n\ge n_0. \end{aligned}$$

This shows that \(\varphi _n(f_n)\rightarrow 0\) as \(n\rightarrow \infty \). \(\square \)

By putting together Propositions 2.3, 2.4 and 4.4, we get the already mentioned result from [13]:

Corollary 4.5

Let X be a Banach space, let \((\Omega ,\Sigma )\) be a measurable space and let \(\nu \in \textrm{ca}(\Sigma ,X)\) with \(\sigma \)-finite variation. If \(L_1(\nu )\) has the positive Schur property, then it has the Dunford-Pettis property.

Let E be a Banach space with a normalized 1-unconditional Schauder basis, say \((e_n)_{n\in \mathbb {N}}\). The E-sum of countably many copies of \(L_1[0,1]\) is the Banach lattice Z of all sequences \((h_n)_{n\in \mathbb {N}}\) in \(L_1[0,1]\) such that the series \(\sum _{n=1}^\infty \Vert h_n\Vert _{L_1[0,1]} \, e_n\) converges unconditionally in E, equipped with the norm

$$\begin{aligned} \left\| (h_n)_{n\in \mathbb {N}}\right\| _{Z}:= \left\| \sum _{n=1}^\infty \Vert h_n\Vert _{L_1[0,1]} \, e_n\right\| _E \end{aligned}$$

and the coordinatewise order. If E has the the Schur property, then Z has the positive Schur property, but it is not lattice-isomorphic to an AL-space unless E is isomorphic to \(\ell _1\), [42, Section 3].

The following construction provides more examples of Banach lattices with such features:

Example 4.6

Let X be a Banach space and let \(\sum _{n=1}^\infty x_n\) be an unconditionally convergent series in X with \(x_n\ne 0\) for all \(n\in \mathbb {N}\). Let \(\lambda \) be the Lebesgue measure on the \(\sigma \)-algebra \(\Sigma \) of all Borel subsets of [0, 1]. Write \(I_n:=(2^{-n},2^{-n+1}]\) for all \(n\in \mathbb {N}\). Then:

  1. (i)

    The formula

    $$\begin{aligned} \nu (A):=\sum _{n=1}^\infty 2^n\lambda (A\cap I_n)x_n, \quad A\in \Sigma , \end{aligned}$$

    defines a vector measure \(\nu \in \textrm{ca}(\Sigma ,X)\).

  2. (ii)

    \({\mathcal {N}}(\nu )={\mathcal {N}}(\lambda )\). Hence, \(\nu \) is atomless and \(L_1(\nu )\) is separable.

  3. (iii)

    \({\mathcal {R}}(\nu )\) is relatively norm compact.

  4. (iv)

    \(|\nu |\) is \(\sigma \)-finite and \(|\nu |([0,1])=\sum _{n=1}^\infty \Vert x_n\Vert \).

  5. (v)

    If \(\sum _{n=1}^\infty x_n\) is not absolutely convergent, then \(L_1(\nu )\) is not lattice-isomorphic to an AL-space.

  6. (vi)

    If X has the Schur property, then \(L_1(\nu )\) has the positive Schur property and the Dunford-Pettis property.

  7. (vii)

    If \(\sum _{n=1}^\infty x_n\) is not absolutely convergent and X has the Schur property, then \(L_1(\nu )\) is not lattice-isomorphic to \(L_1({\tilde{\nu }})\) for any \(\sigma \)-algebra \({\tilde{\Sigma }}\) and any \({\tilde{\nu }} \in \textrm{ca}({\tilde{\Sigma }},c_0)\) such that \({\mathcal {R}}({\tilde{\nu }})\) is relatively norm compact.

Proof

Since \(\sum _{n=1}^\infty x_n\) is unconditionally convergent, for every \((a_n)_{n\in \mathbb {N}}\in \ell _\infty \) the series \(\sum _{n=1}^\infty a_nx_n\) is unconditionally convergent and the map

$$\begin{aligned} T:\ell _\infty \rightarrow X, \quad T((a_n)_{n\in \mathbb {N}}):=\sum _{n=1}^\infty a_n x_n, \end{aligned}$$

is a compact operator (see, e.g., [17, Theorem 1.9]). This shows that the map \(\nu \) is well-defined and has relatively norm compact range (note that \(2^n \lambda (A\cap I_n) \le 1\) for all \(n\in \mathbb {N}\)). Since the map \(\Sigma \ni A\mapsto 2^n\lambda (A\cap I_n)x_n \in X\) is countably additive for each \(n\in \mathbb {N}\), the Vitali-Hahn-Saks theorem (see, e.g., [18, p. 24, Corollary 10]) ensures that \(\nu \) is countably additive. This proves parts (i) and (iii).

(ii) The equality \({\mathcal {N}}(\nu )={\mathcal {N}}(\lambda )\) is obvious. Since \(\lambda \) is atomless, so is \(\nu \). Let \({\mathcal {C}} \subseteq \Sigma \) be a countable set such that for every \(A\in \Sigma \) we have \(\inf _{C\in {\mathcal {C}}}\lambda (A\triangle C)=0\). Then for every \(A\in \Sigma \) we also have \(\inf _{C\in {\mathcal {C}}}\Vert \nu \Vert (A\triangle C)=0\) (notice that \(\nu \) is \(\lambda \)-continuous). This implies that \(L_1(\nu )\) is separable, because the set of all \(\Sigma \)-simple functions is norm dense in \(L_1(\nu )\).

(iv) It is easy to check that \(|\nu |(A)=\sum _{n=1}^\infty 2^n\lambda (A\cap I_n)\Vert x_n\Vert \) for every \(A\in \Sigma \).

(v) This follows from [12, Proposition 2] and (iv).

(vi) We already know that the Schur property of X implies that \(L_1(\nu )\) has the positive Schur property, [13, proof of Theorem 4]. Now, (iv) and Corollary 4.5 ensure that \(L_1(\nu )\) has the Dunford-Pettis property.

(vii) Suppose, by contradiction, that there exist a \(\sigma \)-algebra \({\tilde{\Sigma }}\) and \({\tilde{\nu }}\in \textrm{ca}({\tilde{\Sigma }},c_0)\) such that \({\mathcal {R}}({\tilde{\nu }})\) is relatively norm compact and \(L_1(\nu )\) is lattice-isomorphic to \(L_1({\tilde{\nu }})\). Then \(L_1({\tilde{\nu }})\) has the positive Schur property (by (vi)) and we can apply Proposition 2.6 to infer that the integration operator \(I_{{\tilde{\nu }}}:L_1({\tilde{\nu }}) \rightarrow c_0\) is Dunford-Pettis. Now, Proposition 3.1 and Theorem 3.6 (the usual basis of \(c_0\) is shrinking) imply that \(L_1({\tilde{\nu }})\) is lattice-isomorphic to an AL-space, which contradicts (v). \(\square \)

Remark 4.7

Part (vii) of Example 4.6 should be compared with [12, Theorem 1]. That result states that if X is a Banach space, \((\Omega ,\Sigma )\) is a measurable space, the vector measure \(\nu \in \textrm{ca}(\Sigma ,X)\) is atomless and \(L_1(\nu )\) is separable, then there is \({\tilde{\nu }}\in \textrm{ca}(\Sigma ,c_0)\) such that \(L_1(\nu )\) and \(L_1({\tilde{\nu }})\) are lattice-isometric (cf. [24, Theorem 5] for another proof). For variants in the non-separable setting, see [36] and [37]. In [24, Theorem 5] it was claimed that if, in addition, \({\mathcal {R}}(\nu )\) is relatively norm compact, then \({\tilde{\nu }}\) can be chosen so that \({\mathcal {R}}({\tilde{\nu }})\) is relatively norm compact as well. Unfortunately, this turns out to be false in general, as shown in Example 4.6(vii).