1 Introduction and preliminaries

We assume that \((X,\Vert \cdot \Vert _X)\) is a real Banach space. For terminology concerning Riesz spaces and function spaces, we refer the reader to [9, 13, 27].

We assume that \((\Omega ,\Sigma ,\lambda )\) is a finite measure space. Let \(L^0\) denote the corresponding space of \(\lambda \)-equivalence classes of all \(\Sigma \)-measurable real functions on \(\Omega \). Then \(L^0\) is a super Dedekind complete Riesz space, equipped with the topology \({\mathcal {T}}_o\) of convergence in measure. By \({\mathcal {S}}(\Sigma )\) we denote the space of all real \(\Sigma \)-simple functions \(s=\sum ^n_{i=1} c_i\mathbb {1}_{A_i}\), where the sets \(A_i\in \Sigma \) are pairwise disjoint.

Let \((E,\Vert \cdot \Vert _E)\) be a Banach function space, where E is an order ideal of \(L^0\) such that \(L^\infty \subset E\subset L^1\), and \(\Vert \cdot \Vert _E\) is a Riesz norm on E. By \({\mathcal {T}}_E\) we denote the \(\Vert \cdot \Vert _E\)-norm topology on E. By \(E'\) we denote the Köthe dual of E, that is,

$$\begin{aligned} E':=\bigg \{v\in L^0:\int _\Omega |u(\omega )\,v(\omega )|\,d\lambda <\infty \ \text{ for } \text{ all } u\in E\bigg \}. \end{aligned}$$

The associated norm \(\Vert \cdot \Vert _{E'}\) on \(E'\) is defined for \(v\in E'\) by

$$\begin{aligned} \Vert v\Vert _{E'}=\sup \left\{ \int _\Omega |u(\omega )\,v(\omega )|\,d\lambda :u\in E,\; \Vert u\Vert _E\le 1\right\} . \end{aligned}$$

We will assume that E is perfect, that is, \(E=E''\) and \(\Vert u\Vert _E=\Vert u\Vert _{E''}\). The order continuous dual \(E^\sim _n\) of E separates the points of E and \(E^\sim _n\) can be identified with \(E'\) through the Riesz isomorphism \(E'\ni v\mapsto F_v\in E^\sim _n\), where

$$\begin{aligned} F_v(u)=\int _\Omega u(\omega )v(\omega )\,d\lambda \ \text{ for } u\in E \ \text{ and } \ \Vert F_v\Vert =\Vert v\Vert _{E'} \end{aligned}$$

(see [13, Theorem 6.1.1]). The Mackey topology \(\tau (E,E')\) is a locally convex-solid Hausdorff topology with the Lebesgue property (see [9, Corollary 82H]). Then \(\tau (E,E')\subset {\mathcal {T}}_E\) and \(\tau (E,E')={\mathcal {T}}_E\) if the norm \(\Vert \cdot \Vert _E\) is order continuous.

The most important classes of Banach function spaces are Lebesgue spaces \(L^p\) \((1\le p\le \infty )\) and Orlicz spaces \(L^\varphi \) (see [19]).

Now we present a characterization of \((\tau (E,E'),\Vert \cdot \Vert _X)\)-continuous linear operators \(T:E\rightarrow X\) (see [17, Proposition 2.2]).

Proposition 1.1

For a bounded linear operator \(T:E\rightarrow X\) the following statements are equivalent:

  1. (i)

    T is \((\tau (E,E'),\Vert \cdot \Vert _X)\)-continuous.

  2. (ii)

    \(\Vert T(u_n)\Vert _X\rightarrow 0\) if \(\,u_n(\omega )\rightarrow 0\) \(\lambda \)-a.e. and \(|u_n(\omega )|\le |u(\omega )|\) \(\lambda \)-a.e. for some \(u\in E\) and all \(n\in {\mathbb {N}}\).

  3. (iii)

    For each \(u\in E\), \(\Vert T(u\mathbb {1}_{A_n})\Vert _X\rightarrow 0\) whenever \(\lambda (A_n)\rightarrow 0\).

For terminology and basic facts concerning vector measure, we refer the reader to [4, 6, 7, 22]. For a finitely additive measure \(m:\Sigma \rightarrow X\), by |m|(A) we denote the variation of m on \(A\in \Sigma \). A measure \(m:\Sigma \rightarrow X\) is said to be \(\lambda \) -continuous if \(\Vert m(A_n)\Vert _X\rightarrow 0\) whenever \(\lambda (A_n)\rightarrow 0\).

Let \(L^1(X)\) denote the Banach space of \(\lambda \)-equivalence classes of all X-valued Bochner integrable functions g defined on \(\Omega \), equipped with the norm \(\Vert g\Vert _1:=\int _\Omega \Vert g(\omega )\Vert _X d\lambda .\)

Recall that a \(\lambda \)-continuous measure \(m:\Sigma \rightarrow X\) of finite variation is said to have the Radon-Nikodym property with respect to \(\lambda \) if there exists a function \(g\in L^1(X)\) such that \(m(A)=\int _A g(\omega )\,d\lambda \) for all \(A\in \Sigma \). Then we write \(m=g\lambda \) and a function g is called the density of m with respect to \(\lambda \).

Assume that \(m:\Sigma \rightarrow X\) is a \(\lambda \)-continuous measure. Following [6, § 13] for \(A\in \Sigma \), we put

$$\begin{aligned} |m|_{E'}(A):=\sup \left\{ \sum ^n_{i=1} |c_i|\ \Vert m(A_i)\Vert _X\right\} , \end{aligned}$$

where the supremum is taken for all functions \(s=\sum ^n_{i=1} c_i\mathbb {1}_{A_i}\in \ {\mathcal {S}}(\Sigma )\) such that \(A_i\subset A\) for \(1\le i\le n\) and \(\Vert s\Vert _E\le 1\). The set function \(|m|_{E'}\) will be called a \(E'\) -variation of the measure m.

If, in particular, \(E=L^\infty \), then \(|m|_{L^1}(A)=|m|(A)\) for \(A\in \Sigma \).

Let \(L^0(X)\) stand for the linear space of \(\lambda \)-equivalence classes of all strongly \(\Sigma \)-measurable functions \(g:\Omega \rightarrow X\). Let

$$\begin{aligned} E(X)=\left\{ g\in L^0(X):\Vert g(\cdot )\Vert _X\in E\right\} . \end{aligned}$$

Then E(X) equipped with the norm \(\Vert g\Vert _{E(X)}:=\Vert \cdot \Vert g(\cdot )\Vert _X\Vert _{E}\) is a Banach space, called a Köthe-Bochner space (see [14]).

Definition 1.1

A bounded linear operator \(T:E\rightarrow X\) is said to be Bochner representable, if there exists \(g\in E'(X)\) such that

$$\begin{aligned} T(u)=\int _\Omega u(\omega )g(\omega )\,d\lambda \ \ \text{ for } \ \ u\in E. \end{aligned}$$

The concept of nuclear operators between Banach spaces in due to Ruston [21]. Grothendieck carried over the concept of nuclear operators to locally convex spaces [10, 11] (see also [26, p. 289], [18, 23, Chap. 3, § 7], [4, Chap. 6], [5, 22]).

Following [23, Chap. 3, § 7] (see also [2, Chap. 4], [12, 17.3, p. 379]), we have

Definition 1.2

A linear operator \(T:E\rightarrow X\) is said to be \(\tau (E,E')\) -nuclear if there exist a sequence \((v_n)\) in \(E'\) such that the family \(\{F_{v_n}:n\in {\mathbb {N}}\}\) is \(\tau (E,E')\)-equicontinuous, a bounded sequence \((x_n)\) in X and a sequence \((\alpha _n)\in \ell ^1\) such that

$$\begin{aligned} T(u)=\sum ^\infty _{n=1}\alpha _n\left( \int _\Omega u\, v_n \,d\lambda \right) x_n \ \ \text{ for } \ u\in E. \end{aligned}$$
(1.1)

Let

$$\begin{aligned} \Vert T\Vert _{nuc}:=\inf \left\{ \sum ^\infty _{n=1}|\alpha _n|\ \Vert v_n\Vert _{E'}\, \Vert x_n\Vert _X\right\} , \end{aligned}$$

where the infimum is taken over all sequences \((v_n)\) in \(E'\), \((x_n)\) in X and \((\alpha _n)\in \ell ^1\) such that T admits a representation (1.1).

It is known that a \(\tau (E,E')\)-nuclear operator \(T:E\rightarrow X\) is \((\tau (E,E'),\Vert \cdot \Vert _X)\)-continuous and \(\tau (E,E')\)-compact, that is, T(V) is relatively norm compact in X for some \(\tau (E,E')\)-neighborhood V of 0 in E (see [23, Chap. 3, § 7, Corollary 1], [12, Corollary 4, p. 379]).

In this paper we study \(\tau (E,E')\)-nuclear operators \(T:E\rightarrow X\). In Section 2 it is shown that every \(\tau (E,E')\)-nuclear operator \(T:E\rightarrow X\) is Bochner representable (see Theorem 2.3 below). In particular, we obtain that a linear operator \(T:L^\infty \rightarrow X\) is \(\tau (L^\infty ,L^1)\)-nuclear if and only if its representing measure \(m_T:\Sigma \rightarrow X\) has the Radon-Nikodym property and \(|m_T|(\Omega )=\Vert T\Vert _{nuc}\) (see Theorem 2.5 below). As an application, we obtain that some natural kernel operators on \(L^\infty \) are \(\tau (L^\infty ,L^1)\)-nuclear (see Proposition 2.9 below). In Section 3 it is shown that every \(\tau (L^\infty ,L^1)\)-nuclear operator \(T:L^\infty \rightarrow X\) admits a factorization through some Orlicz space \(L^\varphi \), that is, \(T=S\circ i_\infty \), where \(S:L^\varphi \rightarrow X\) is a Bochner representable, compact operator and \(i_\infty :L^\infty \rightarrow L^\varphi \) denotes the inclusion map (see Corollary 3.5).

2 Nuclear operators on Banach function spaces

Assume that \(T:E\rightarrow X\) is a linear operator. Then the measure \(m_T:\Sigma \rightarrow X\) defined by

$$\begin{aligned} m_T(A):=T(\mathbb {1}_A) \ \ \text{ for } \ A\in \Sigma \end{aligned}$$

is called a representing measure of T.

If, in particular, T is \((\tau (E,E'),\Vert \cdot \Vert _X)\)-continuous, then using Proposition 1.1 we obtain that \(m_T\) is countably additive. Since \(m_T(A)=0\) if \(\lambda (A)=0\), by the Pettis theorem \(m_T\) is \(\lambda \)-continuous, that is, \(m_T\ll \lambda \).

The following lemma will be useful.

Lemma 2.1

Let \(T:E\rightarrow X\) be a \((\tau (E,E'),\Vert \cdot \Vert _X)\)-continuous linear operator. If \(|m_T|_{E'}(\Omega )<\infty \) and \(m_T\) has the Radon-Nikodym property with respect to \(\lambda \) with a density \(g\in L^1(X)\), then \(g\in E'(X)\) and

$$\begin{aligned} \Vert \mathbb {1}_A g\Vert _{E'(X)}=|m_T|_{E'}(A) \ \ \text{ for } \text{ all } \ A\in \Sigma , \end{aligned}$$

and

$$\begin{aligned} T(u)=\int _\Omega u(\omega )\,g(\omega )\,d\lambda \ \ \text{ for } \text{ all } \ u\in E. \end{aligned}$$

Proof

First we shall show that for \(A\in \Sigma \),

$$\begin{aligned} |m_T|_{E'}(A)=\sup \left\{ \int _A|s(\omega )| \ \Vert g(\omega )\Vert _X \,d\lambda : s\in {\mathcal {S}}(\Sigma ), \ \Vert s\Vert _E\le 1\right\} . \end{aligned}$$
(2.1)

Note that \(|m_T|(A)=\int _A\Vert g(\omega )\Vert _X\,d\lambda \). For \(s=\sum ^k_{i=1}c_i\mathbb {1}_{A_i} \in {\mathcal {S}}(\Sigma )\), we have

$$\begin{aligned} \sum ^k_{i=1}c_i\, m_T(A\cap A_i)=\sum ^k_{i=1} c_i\int _{A\cap A_i} g(\omega )\,d\lambda = \int _A s(\omega )\, g(\omega )\, d\lambda . \end{aligned}$$

We now show that for \(s=\sum ^k_{i=1} c_i\mathbb {1}_{A_i}\in {\mathcal {S}}(\Sigma )\) and \(\Vert s\Vert _E\le 1\), we have

$$\begin{aligned} \sum ^k_{i=1}|c_i|\int _{A\cap A_i}\Vert g(\omega )\Vert _X \,d\lambda =\sum ^k_{i=1}|c_i|\ |m_T| (A\cap A_i)\le |m_T|_{E'}(A). \end{aligned}$$

Indeed, let \(\varepsilon >0\) be given. Then for each \(1\le i\le k\), there exists a \(\Sigma \)-partition \((A_{i,j})^{j_i}_{j=1}\) of \(A\cap A_i\) such that

$$\begin{aligned} |m_T|(A\cap A_i)\le \sum ^{j_i}_{j=1}\big \Vert m_T(A_{ij})\big \Vert _X+\frac{\varepsilon }{k|c_i|}. \end{aligned}$$

Hence

$$\begin{aligned} \sum ^k_{i=1}|c_i|\,|m_T|(A\cap A_i)\le \sum ^k_{i=1}\left( \sum ^{j_i}_{j=1} \Vert c_i \,m_T(A_{ij})\Vert _X\right) +\varepsilon \le |m_T|_{E'}(A)+\varepsilon , \end{aligned}$$

because \(\sum ^k_{i=1}\big (\sum ^{j_i}_{j=1}c_i\, \mathbb {1}_{A_{ij}}\big )=\sum ^k_{i=1} c_i\,\mathbb {1}_{A_i}.\)

Then we have

$$\begin{aligned} \begin{array}{l} \displaystyle \sum ^k_{i=1}|c_i|\;\Vert m_T(A\cap A_i)\Vert _X \le \sum ^k_{i=1}|c_i|\;|m_T|(A\cap A_i)\\ \quad \displaystyle =\sum ^k_{i=1}|c_i|\int _{A\cap A_i}\Vert g(\omega )\Vert _X\,d\lambda =\int _A\left( \sum ^k_{i=1}|c_i| \,\mathbb {1}_{A_i}(\omega )\right) \Vert g(\omega )\Vert _X \,d\lambda \\ \quad \displaystyle =\int _A|s(\omega )|\ \Vert g(\omega )\Vert _X\, d\lambda \le |m_T|_{E'}(A). \end{array} \end{aligned}$$

Taking supremum on the left side, we get

$$\begin{aligned} |m_T|_{E'}(A)=\sup \left\{ \int _A|s(\omega )|\ \Vert g(\omega )\Vert _X \,d\lambda :s\in {\mathcal {S}}(\Sigma ),\ \Vert s\Vert _E\le 1\right\} . \end{aligned}$$

We shall now show that \(g\in E'(X)\), that is, \(\Vert g(\cdot )\Vert _X\in E'\).

Indeed, let \(u\in E\). Then there exists a sequence \((s_n)\) in \({\mathcal {S}}(\Sigma )\) such that \(0\le s_n(\omega )\uparrow |u(\omega )|\) \(\lambda \)-a.e. (see [13, Corollary I.6]). Choose \(c>0\) such that \(\Vert cu\Vert _E\le 1\). Then by the Fatou lemma,

$$\begin{aligned} \int _\Omega c\,|u(\omega )|\ \Vert g(\omega )\Vert _X \,d\lambda \le \sup _n\int _\Omega c \,s_n(\omega )\Vert g(\omega )\Vert _X \, d\lambda \le |m_T|_{E'}(\Omega ), \end{aligned}$$

and this means that \(\Vert g(\cdot )\Vert _X\in E'\), that is, \(g\in E'(X)\). Moreover, for \(A\in \Sigma \) and \(u\in E\) with \(\Vert u\Vert _E\le 1\), using (2.1) we get

$$\begin{aligned} \int _\Omega |u(\omega )|\ \Vert \mathbb {1}_A(\omega )\,g(\omega )\Vert _X\, d\lambda \le \sup _n\int _\Omega s_n(\omega ) \Vert \mathbb {1}_A(\omega )\,g(\omega )\Vert _X\, d\lambda \le |m_T|_{E'}(A) \end{aligned}$$

and it follows that \(\Vert \mathbb {1}_A g\Vert _{E'(X)}\le |m_T|_{E'}(A)\). In view of (2.1) we get \(|m_T|_{E'}(A)\le \Vert \mathbb {1}_A g\Vert _{E'(X)}\). Hence \(|m_T|_{E'}(A)=\Vert \mathbb {1}_A g\Vert _{E'(X)}\).

Note that for \(s=\sum ^k_{i=1} c_i\mathbb {1}_{A_i}\in {\mathcal {S}}(\Sigma )\), we have

$$\begin{aligned} T(s)=\sum ^k_{i=1}c_i\, m_T(A_i)=\int _\Omega s(\omega )\,g(\omega )\,d\lambda . \end{aligned}$$

Let \(u\in E\) be given. Then there exists a sequence \((s_n)\) in \({\mathcal {S}}(\Sigma )\) such that \(|s_n(\omega )-u(\omega )|\rightarrow 0\) \(\lambda \)-a.e. and \(|s_n(\omega )|\le |u(\omega )|\) \(\lambda \)-a.e. for all \(n\in {\mathbb {N}}\). Then \(|s_n(\omega )-u(\omega )|\) \(\Vert g(\omega )\Vert _X\le 2|u(\omega )|\) \(\Vert g(\omega )\Vert _X\) \(\lambda \)-a.e., where \(u\,\Vert g(\cdot )\Vert _X\in L^1\). Hence by the Lebesgue dominated convergence theorem,

$$\begin{aligned} \bigg \Vert \int _\Omega s_n(\omega )\,g(\omega )\, d\lambda -\int _\Omega u(\omega )\,g(\omega )\, d\lambda \,\bigg \Vert _X\le \int _\Omega |s_n(\omega )-u(\omega )| \ \Vert g(\omega )\Vert _X\, d\lambda \rightarrow 0. \end{aligned}$$

On the other hand, in view of Proposition 1.1 we have

$$\begin{aligned} \Vert T(s_n)-T(u)\Vert _X=\Vert T(s_n-u)\Vert _X\rightarrow 0. \end{aligned}$$

Hence \(T(u)=\int _\Omega u(\omega ) g(\omega )\,d\lambda \). \(\square \)

As a consequence of Lemma 2.1, we have

Proposition 2.2

Assume that \(T:E\rightarrow X\) is a Bochner representable operator, that is, there exists \(g\in E'(X)\) such that

$$\begin{aligned} T(u)=\int _\Omega u(\omega )\,g(\omega )\,d\lambda \ \ for \ all \ u\in E. \end{aligned}$$

Then the following statements hold:

  1. (i)

    T is \((\tau (E,E'),\Vert \cdot \Vert _X)\)-continuous.

  2. (ii)

    For every \(A\in \Sigma \), \(|m_T|_{E'}(A)=\Vert \mathbb {1}_A g\Vert _{E'(X)}.\)

Proof

  1. (i)

    Assume that \(u_n(\omega )\rightarrow 0\) \(\lambda \)-a.e. and \(|u_n(\omega )|\le |u(\omega )|\) \(\lambda \)-a.e. for some \(u\in E\) and all \(n\in {\mathbb {N}}\). Since for \(n\in {\mathbb {N}}\),

    $$\begin{aligned} \Vert T(u_n)\Vert _X\le \int _\Omega |u_n(\omega )|\ \Vert g(\omega )\Vert _X\,d\lambda , \end{aligned}$$

    where \(u\,\Vert g(\cdot )\Vert _X\in L^1\), by the Lebesgue dominated convergence theorem, we get \(\Vert T(u_n)\Vert _X\rightarrow 0\). Hence in view of Proposition 1.1T is \((\tau (E,E'),\Vert \cdot \Vert _X)\)-continuous.

  2. (ii)

    Assume that \(s=\sum ^k_{i=1} c_i\mathbb {1}_{A_i}\in {\mathcal {S}}(\Sigma )\) and \(\Vert s\Vert _E\le 1\). Then

    $$\begin{aligned} \begin{array}{rl} \displaystyle \sum ^k_{i=1}|c_i|\ \Vert m_T(A_i)\Vert _X &{} \displaystyle =\sum ^k_{i=1}|c_i|\int _{A_i} \Vert g(\omega )\Vert _X\,d\lambda =\int _\Omega |s(\omega )|\ \Vert g(\omega )\Vert _X\,d\lambda \\ &{} \displaystyle \le \Vert s\Vert _E\ \Vert g\Vert _{E'(X)}\le \Vert g\Vert _{E'(X)}. \end{array} \end{aligned}$$

Hence \(|m_T|_{E'}(\Omega )\le \Vert g\Vert _{E'(X)}\). Using (i) and Lemma 2.1, we get \(|m_T|_{e'}(A)=\Vert \mathbb {1}_A g\Vert _{E'(X)}\) for all \(A\in \Sigma \). \(\square \)

The following result shows a relationship between \(\tau (E,E')\)-nuclear operators and Bochner representable operators \(T:E\rightarrow X\).

Theorem 2.3

Assume that \(T:E\rightarrow X\) is a \(\tau (E,E')\)-nuclear operator. Then T is Bochner representable and \(|m_T|_{E'}(\Omega )\le \Vert T\Vert _{nuc}\).

Proof

Let \(\varepsilon >0\) be given. Then there exists a bounded sequence \((v_n)\) in \(E'\), a bounded sequence \((x_n)\) in X and a sequence \((\alpha _n)\in \ell ^1\) such that

$$\begin{aligned} T(u)=\sum ^\infty _{n=1}\alpha _n\left( \int _\Omega u\,v_n\, d\lambda \right) x_n \ \ \text{ for } \ u\in E \end{aligned}$$

and

$$\begin{aligned} \sum ^\infty _{n=1}|\alpha _n| \ \Vert v_n\Vert _{E'}\ \Vert x_n\Vert _X\le \Vert T\Vert _{nuc}+\varepsilon . \end{aligned}$$
(2.2)

Hence we have

$$\begin{aligned} m_T(A)=\sum ^\infty _{n=1}\alpha _n\left( \int _A v_n\, d\lambda \right) x_n. \end{aligned}$$
(2.3)

We shall now show that \(|m_T|_{E'}(\Omega )\le \Vert T\Vert _{nuc}\). Indeed, let \(s=\sum ^k_{i=1}c_i\mathbb {1}_{A_i}\in {\mathcal {S}}(\Sigma )\) with \(\Vert s\Vert _E\le 1\). Then using (2.2) we get

$$\begin{aligned} \displaystyle \sum ^k_{i=1}|c_i|\ \Vert m_T(A_i)\Vert _X&\displaystyle =&\sum ^k_{i=1}\Vert c_i\sum ^\infty _{n=1}\alpha _n\left( \int _{A_i} v_n(\omega )\,d\lambda \right) x_n\Vert _X\\&\displaystyle \le&\sum ^\infty _{n=1}|\alpha _n|\left( \sum ^k_{i=1}|c_i|\int _{A_i}|v_n(\omega )|\,d\lambda \right) \Vert x_n\Vert _X\\&\displaystyle =&\sum ^\infty _{n=1}|\alpha _n|\ \Vert x_n\Vert _X\int _\Omega |s(\omega )| \ |v_n(\omega )|\,d\lambda \\&\displaystyle \le&\sum ^\infty _{n=1}|\alpha _n|\ \Vert x_n\Vert _X \ \Vert v_n\Vert _{E'}\le \Vert T\Vert _{nuc}+\varepsilon . \end{aligned}$$

Hence we get

$$\begin{aligned} |m_T|_{E'}(\Omega )\le \Vert T\Vert _{nuc}. \end{aligned}$$
(2.4)

For \(n\in {\mathbb {N}}\), let \(g_n:=\sum ^n_{i=1}\alpha _i v_i\otimes x_i\). Then for \(n,k\in {\mathbb {N}}\) with \(n>k\), we have

$$\begin{aligned} \begin{array}{rl} \displaystyle \int _\Omega \Vert g_n(\omega ) &{}\!\!\!\displaystyle -\,g_k(\omega )\Vert _X \,d\lambda =\int _\Omega \left\| \sum ^n_{i=k+1} \alpha _i \,v_i(\omega )x_i\right\| _X\, d\lambda \\ &{}\!\!\!\displaystyle \le \int _\Omega \left( \sum ^n_{i=k+1}|\alpha _i|\ |v_i(\omega )|\ \Vert x_i\Vert _X\right) \,d\lambda = \sum ^n_{i=k+1}|\alpha _i| \ \Vert x_i\Vert _X\int _\Omega |v_i|\,d\lambda \\ &{}\!\!\!\displaystyle \le \sum ^n_{i=k+1}|\alpha _i|\ \Vert x_i\Vert _X\ \Vert \mathbb {1}_\Omega \Vert _E\ \Vert v_i\Vert _{E'}\le \sup _{j\in {\mathbb {N}}}\Vert x_j\Vert _X \sup _{j\in {\mathbb {N}}}\Vert v_j\Vert _{E'}\Vert \mathbb {1}_\Omega \Vert _E \sum ^n_{i=k+1}|\alpha _i|. \end{array} \end{aligned}$$

This follows that \((g_n)\) is a Cauchy sequence in the Banach space \(L^1(X)\), so there exists \(g\in L^1(X)\) such that \(\Vert g_n-g\Vert _1\rightarrow 0\). One can easily show that

$$\begin{aligned} \left( \int _A v\,d\lambda \right) x=\int _A v(\omega )\,x\,d\lambda \ \text{ for } \ v\in E', \ x\in X, \ A\in \Sigma . \end{aligned}$$

Hence for \(A\in \Sigma \), we have

$$\begin{aligned} \begin{array}{l} \displaystyle \bigg \Vert \sum ^n_{i=1}\alpha _i\left( \int _A v_i \,d\lambda \right) x_i-\int _A g\,d\lambda \, \bigg \Vert _X=\bigg \Vert \int _A\sum ^n_{i=1}\alpha _i\, v_i(\omega )\,x_i \,d\lambda -\int _A g(\omega ) \,d\lambda \,\bigg \Vert _X\\ \quad \displaystyle \le \int _A\bigg \Vert \sum ^n_{i=1}\alpha _i\, v_i(\omega )\,x_i-g(\omega )\bigg \Vert _X\,d\lambda = \int _A\Vert g_n(\omega )-g(\omega )\Vert _X\,d\lambda \rightarrow 0. \end{array} \end{aligned}$$

Then in view of (2.3) for \(A\in \Sigma \), we get

$$\begin{aligned} m_T(A)=\int _A g(\omega )\, d\lambda . \end{aligned}$$

Using Lemma 2.1 and (2.4) we see that \(g\in E'(X)\) with \(\Vert g\Vert _{E'(X)}=|m_T|_{E'}(\Omega )\le \Vert T\Vert _{nuc}\) and \(T(u)=\int _\Omega u(\omega )\, g(\omega )\,d\lambda \) for all \(u\in E\). \(\square \)

Now we shall study \(\tau (L^\infty ,L^1)\)-nuclear operators \(T:L^\infty \rightarrow X\).

Making use of the Dunford–Pettis theorem (see [3, Theorem, p. 93]) we have

Theorem 2.4

For a subset H of \(L^1\) the following statements are equivalent:

  1. (i)

    H is relatively weakly compact.

  2. (ii)

    \(\sup _{v\in H}\Vert v\Vert _1<\infty \) and H is uniformly integrable.

  3. (iii)

    \(\{F_v:v\in H\}\) is \(\tau (L^\infty ,L^1)\)-equicontinuous.

Note that in view of Theorem 2.4 and Definition 1.2, a linear operator T :  \(L^\infty \rightarrow X\) is \(\tau (L^\infty ,L^1)\) -nuclear if there exist a bounded uniformly integrable sequence \((v_n)\) in \(L^1\), a bounded sequence \((x_n)\) in X and a sequence \((\alpha _n)\in \ell ^1\) such that

$$\begin{aligned} T(u)=\sum ^\infty _{n=1}\alpha _n\left( \int _\Omega u\, v_n\, d\lambda \right) x_n \ \text{ for } \text{ all } \ u\in L^\infty \end{aligned}$$
(2.5)

and then

$$\begin{aligned} \Vert T\Vert _{nuc}=\inf \left\{ \sum ^\infty _{n=1}|\alpha _n|\ \Vert v_n\Vert _1\ \Vert x_n\Vert _X\right\} , \end{aligned}$$

where the infimum is taken over all sequences \((v_n)\) in \(L^1\), \((x_n)\) in X and \((\alpha _n)\in \ell ^1\) such that T admits a representation (2.5).

Now we can state our main result.

Theorem 2.5

For a linear operator \(T:L^\infty \rightarrow X\) the following statements are equivalent:

  1. (i)

    \(m_T\) has the Radon-Nikodym property with respect to \(\lambda \).

  2. (ii)

    T is Bochner representable.

  3. (iii)

    T is a \(\tau (L^\infty ,L^1)\)-nuclear operator.

In this case \(|m_T|(\Omega )=\Vert T\Vert _{nuc}\).

Proof

(i)\(\Rightarrow \)(ii) Assume that (i) holds with the density \(g\in L^1(X)\). Note that for \(s\in {\mathcal {S}}(\Sigma )\), we have

$$\begin{aligned} \int _\Omega s\,dm_T=\int _\Omega s(\omega )\,g(\omega )\,d\lambda . \end{aligned}$$

Let \(u\in L^\infty \). Choose a sequence \((s_n)\) in \({\mathcal {S}}(\Sigma )\) such that \(\Vert u-s_n\Vert _\infty \rightarrow 0\). Then we have

$$\begin{aligned} T(u)=\int _\Omega u\,dm_T=\lim \int _\Omega s_n \,dm_T=\lim \int _\Omega s_n(\omega )\,g(\omega )\, d\lambda =\int _\Omega u(\omega )\,g(\omega )\,d\lambda . \end{aligned}$$

(ii)\(\Rightarrow \)(i) This is obvious.

(ii)\(\Rightarrow \)(iii) Assume that (ii) holds, that is, there exists \(g\in L^1(X)\) such that

$$\begin{aligned} T(u)=\int _\Omega u(\omega )\,g(\omega )\,d\lambda \ \ \text{ for } \text{ all } \ u\in L^\infty . \end{aligned}$$

Let \(L^1\hat{\otimes } X\) denote the projective tensor product of \(L^1\) and X, equipped with the norm \(\pi \) defined for \(w\in L^1\hat{\otimes } X\), by

$$\begin{aligned} \pi (w):=\inf \left\{ \sum ^\infty _{n=1}|\alpha _n|\ \Vert v_n\Vert _1\ \Vert x_n\Vert _X\right\} , \end{aligned}$$

where the infimum is taken over all sequences \((v_n)\) in \(L^1\), \((x_n)\) in X with \(\lim \Vert v_n\Vert _1=0=\lim \Vert x_n\Vert _X\) and \((\alpha _n)\in \ell ^1\) such that \(w=\sum ^\infty _{n=1}\alpha _n v_n\otimes \,x_n\) (see [22, Proposition 2.8, pp. 21–22]). It is known that \(L^1\hat{\otimes }\, X\) is isometrically isomorphic to \(L^1(X)\) through the isometry J, where

$$\begin{aligned} J(v\otimes x):=v(\cdot )\,x \ \text{ for } \ v\in L^1, \ x\in X \end{aligned}$$

(see [4, Example 10, p. 228], [22, Example 2.19, p. 29], [5, Theorem 1.1.10, p. 14]). Let \(\varepsilon >0\) be given. Then there exist sequence \((v_n)\) in \(L^1\) and \((x_n)\) in X with \(\lim _n\Vert v_n\Vert _1=0=\lim \Vert x_n\Vert _X\) and \((\alpha _n)\in \ell ^1\) such that

$$\begin{aligned} J^{-1}(g)=\sum ^\infty _{n=1}\alpha _n\, v_n\otimes \, x_n \ \text{ in } \ \big (L^1\hat{\otimes }\, X,\pi \big ) \end{aligned}$$

and

$$\begin{aligned} \sum ^\infty _{n=1}|\alpha _n|\ \Vert v_n\Vert _1 \ \Vert x_n\Vert _X\le \pi (J^{-1}(g))+\varepsilon = \Vert g\Vert _1+\varepsilon . \end{aligned}$$
(2.6)

Hence

$$\begin{aligned} g=J\left( \sum ^\infty _{n=1}\alpha _n v_n\otimes \, x_n\right) =\sum ^\infty _{n=1} \alpha _n \,v_n(\cdot )\,x_n \ \text{ in } \ \big (L^1(X),\ \Vert \cdot \Vert _1\big ). \end{aligned}$$

Note that

$$\begin{aligned} \begin{array}{l} \displaystyle \bigg \Vert \int _\Omega u(\omega )\,g(\omega )\,d\lambda -\sum ^n_{i=1}\alpha _i\left( \int _\Omega u(\omega ) \,v_i(\omega ) \,d\lambda \right) x_i\bigg \Vert _X\\ \quad \displaystyle \le \int _\Omega |u(\omega )|\ \bigg \Vert g(\omega )-\sum ^n_{i=1}\alpha _i\, v_i(\omega )\,x_i\,\bigg \Vert _X \,d\lambda \\ \quad \displaystyle \le \Vert u\Vert _{\infty }\int _\Omega \bigg \Vert g(\omega )-\sum ^n_{i=1}\alpha _i\, v_i(\omega )\,x_i\,\bigg \Vert _X\, d\lambda = \Vert u\Vert _{\infty }\cdot \bigg \Vert g-\sum ^n_{i=1}\alpha _i \,v_i\otimes x_i\bigg \Vert _{1} \end{array} \end{aligned}$$

Hence

$$\begin{aligned} T(u)=\sum ^\infty _{n=1}\alpha _n\left( \int _\Omega u\, v_n\, d\lambda \right) x_n \ \text{ for } \ u\in L^\infty . \end{aligned}$$

Since \(\lim \Vert v_n\Vert _1=0\), the set \(\{v_n:n\in {\mathbb {N}}\}\) is uniformly integrable and it follows that T is \(\tau (L^\infty ,L^1)\)-nuclear. In view of (2.6) we get

$$\begin{aligned} \Vert T\Vert _{nuc}\le \Vert g\Vert _{1}=|m_T|(\Omega ). \end{aligned}$$
(2.7)

(iii)\(\Rightarrow \)(ii) Assume that (iii) holds. Then by Theorem 2.3T is Bochner representable and

$$\begin{aligned} |m_T|(\Omega )=\Vert g\Vert _1\le \Vert T\Vert _{nuc}. \end{aligned}$$
(2.8)

Thus (i)\(\Leftrightarrow \)(ii)\(\Leftrightarrow \)(iii) hold and using (2.7) and (2.8), we get \(|m_T|(\Omega )=\Vert T\Vert _{nuc}\). \(\square \)

Let \(ba_\lambda (\Sigma )\) denote the Banach space of all bounded finitely additive real measures \(\mu \) on \(\Sigma \) such that \(\mu (A)=0\) if \(\lambda (A)=0\), equipped with norm \(\Vert \mu \Vert :=|\mu |(\Omega )\). The Banach dual \((L^\infty )^*\) of \(L^\infty \) can be identified with \(ba_\lambda (\Sigma )\) through the integration mapping \(ba_\lambda (\Sigma )\ni \mu \mapsto F_\mu \in (L^\infty )^*\), where \(F_\mu (u)=\int _\Omega u d\mu \) for all \(u\in L^\infty \) and \(|\mu |(\Omega )=\Vert F_\mu \Vert \).

Recall (see [26, p. 279]) that a linear operator \(T:L^\infty \rightarrow X\) is said to be nuclear if there exist a bounded sequence \((\mu _n)\) in \(ba_\lambda (\Sigma )\), a bounded sequence \((x_n)\) in X and a sequence \((\alpha _n)\in \ell ^1\) such that

$$\begin{aligned} T(u)=\sum ^\infty _{n=1}\alpha _n\left( \int _\Omega u\, d\mu _n\right) x_n \ \text{ for } \text{ all } \ u\in L^\infty . \end{aligned}$$

It is well known that a bounded linear operator \(T:L^\infty \rightarrow X\) is \((\tau (L^\infty ,L^1),\) \(\Vert \cdot \Vert _X)\)-continuous if and only if its representing measure \(m_T:\Sigma \rightarrow X\) is \(\lambda \)-continuous (see [17, Proposition 3.1]). Hence due to Swartz (see [24, Theorem 1 and Theorem 3]) we have

Theorem 2.6

Assume that \(T:L^\infty \rightarrow X\) is a \((\tau (L^\infty ,L^1),\Vert \cdot \Vert _X)\)-continuous linear operator. Then the following statements are equivalent:

  1. (i)

    T is nuclear.

  2. (ii)

    \(m_T\) has the Radon-Nikodym property with respect to \(\lambda \).

Combining Theorem 2.5 and Theorem 2.6, we get

Corollary 2.7

For a linear operator \(T:L^\infty \rightarrow X\) the following statements are equivalent:

  1. (i)

    T is \(\tau (L^\infty ,L^1)\)-nuclear.

  2. (ii)

    T is \((\tau (L^\infty ,L^1),\Vert \cdot \Vert _X)\)-continuous and nuclear.

As an application of Theorem 2.5 we show that some natural kernel operators on \(L^\infty \) are \(\tau (L^\infty ,L^1)\)-nuclear.

Assume that \(K_1\) and \(K_2\) are compact Hausdorff spaces and \(k(\cdot ,\cdot )\in C(K_2\times K_1)\). Let \({\mathcal {B}}o\) be the \(\sigma \)-algebra of Borel sets in \(K_1\) and \(\lambda :{\mathcal {B}}o\rightarrow [0,\infty )\) be a countably additive measure.

We will need the following lemma.

Lemma 2.8

For every \(u\in L^\infty (\lambda )\), the mapping \(\Psi _u:K_1\ni s\mapsto u(s)\,k(\cdot ,s)\in C(K_2)\) is continuous.

Proof

Let \(s_0\in K_1\) and \(\varepsilon >0\) be given. Then for every \(t\in K_2\) there exist a neighborhood \(V_t\) of t and a neighborhood \(W_t\) of \(s_0\) such that

$$\begin{aligned} |\,k(z,s)-k(t,s_0)|\le \frac{\varepsilon }{\Vert u\Vert _\infty } \ \ \text{ for } \text{ all } \ z\in V_t, \ s\in W_t. \end{aligned}$$

Hence there exist \(t_1,\dots ,t_n\in K_2\) such that \(K_2=\bigcup ^n_{i=1} V_{t_i}\). Let us put \(W:=\bigcap ^n_{i=1} W_{t_i}\). For \(t\in K_2\), choose \(i_0\) with \(1\le i_0\le n\) such that \(t\in V_{t_{i_0}}\). Then for \(s\in W\), we have \(|k(t,s)-k(t,s_0|\le \frac{\varepsilon }{\Vert u\Vert _\infty }\). Hence

$$\begin{aligned} \Vert \Psi _u(s)-\Psi _u(s_0)\Vert _\infty =\sup _{t\in K_2}|k(t,s)-k(t,s_0)|\;\Vert u\Vert _\infty \le \varepsilon . \end{aligned}$$

This means that \(\Psi _u\) is continuous. \(\square \)

Note that the Banach space \(C(K_1,C(K_2))\) can be embedded in the Banach space \(L^1(C(K_2))\) such that with each function from \(C(K_1,C(K_2))\) is associated its \(\lambda \)-equivalence class in \(L^1(C(K_2))\).

In view of Lemma 2.8 we can define a kernel operator \(T_k:L^\infty (\lambda )\rightarrow C(K_2)\) by

$$\begin{aligned} T_k(u):=\int _{K_1}u(s)\,k(\cdot ,s)\,d\lambda (s) \ \ \text{ for } \text{ all } \ u\in L^\infty (\lambda ). \end{aligned}$$

For \(t\in K_2\), let \(\delta _t(w):=w(t)\) for all \(w\in C(K_2)\). Then \(\delta _t\in C(K_2)^*\) and according to the Hille’s theorem (see [DU, Theorem 6, p. 47]), we have

$$\begin{aligned} T_k(u)(t)=\delta _t(T_k(u))=\int _{K_1}u(s)\,k(t,s)\,d\lambda (s) \ \ \text{ for } \text{ all } \ u\in L^\infty (\lambda ), \ t\in K_2. \end{aligned}$$

Then

$$\begin{aligned} m_{T_k}(A)=T_k(\mathbb {1}_A)=\int _A k(\cdot ,s)\,d\lambda (s) \ \ \text{ for } \text{ all } \ A\in {\mathcal {B}}o, \end{aligned}$$

where the mapping \(K_1\ni s\mapsto k(\cdot ,s)\in C(K_2)\) belongs to \(L^1(C(K_2))\).

Hence for \(A\in {\mathcal {B}}o\),

$$\begin{aligned} |m_{T_k}|(A)=\int _A \Vert k(\cdot ,s)\Vert _\infty \,d\lambda (s)=\int _A\sup _{t\in K_2} |k(t,s)|\,d\lambda (s). \end{aligned}$$

Hence as a consequence of Theorem 2.5 we have

Proposition 2.9

The kernel operator \(T_k:L^\infty (\lambda )\rightarrow C(K_2)\) is \(\tau (L^\infty (\lambda ),L^1(\lambda ))\)-nuclear and

$$\begin{aligned} \Vert T_k\Vert _{nuc}=\int _{K_1}\sup _{t\in K_2} |k(t,s)|\,d\lambda (s). \end{aligned}$$

3 Application of the theory of Orlicz spaces to vector measures

First we recall terminology and basic facts concerning Orlicz spaces and Orlicz-Bochner spaces (see [19] for more details). By a Young function we mean here a convex continuous mapping \(\varphi :[0,\infty )\rightarrow [0,\infty )\) that vanishes only at 0 and \(\varphi (t)/t\) \(\rightarrow 0\) as \(t\rightarrow 0\) and \(\varphi (t)/t\rightarrow \infty \) as \(t\rightarrow \infty \). By \(\varphi ^*\) we denote the complementary function of \(\varphi \) in the sense of Young, that is, \(\varphi ^*(t)=\sup \{ts-\varphi (s):s\ge 0\}\) for \(t\ge 0\). Note that \(\varphi ^{**}=\varphi \).

The corresponding Orlicz space \(L^\varphi \) is an ideal of \(L^0\) defined by

$$\begin{aligned} L^\varphi :=\bigg \{u\in L^0:\int _\Omega \varphi \,(\alpha \,|u\,(\omega )|\,)\,d\lambda <\infty \ \ \text{ for } \text{ some } \ \alpha >0\bigg \}, \end{aligned}$$

and equipped with the topology \({\mathcal {T}}_\varphi \), defined by two equivalent norms:

$$\begin{aligned} \begin{array}{l} \displaystyle \Vert u\Vert _\varphi :=\inf \bigg \{\alpha >0:\int _\Omega \varphi \,(|u(\omega )|\,/\,\alpha )\, d\lambda \le 1\bigg \},\\ \displaystyle \Vert u\Vert ^0_\varphi :=\sup \bigg \{\int _\Omega |u(\omega )\,v(\omega )|\, d\lambda :v\in L^{\varphi ^*},\int _\Omega \varphi ^* (|v(\omega )|)\,d\lambda \le 1\bigg \}, \end{array} \end{aligned}$$

called the Luxemburg norm and the Orlicz norm. Then we have:

\(\Vert u\Vert _\varphi \le 1\) if and only if \(\int _\Omega \varphi (|u(\omega )|)\,d\lambda \le 1\),

\(\Vert u_n\Vert _\varphi \rightarrow 0\) if and only if \(\int _\Omega \varphi (\alpha |u_n(\omega )|)\,d\lambda \rightarrow 0\) for every \(\alpha >0\),

$$\begin{aligned} \begin{array}{rl} E^\varphi := &{}\! \displaystyle \bigg \{u\in L^\varphi :\int _\Omega \varphi (\alpha \,|\,u(\omega )|)\,d\lambda <\infty \ \ \text{ for } \text{ all } \ \alpha >0\bigg \}\\ = &{}\! \displaystyle \big \{u\in L^\varphi :\Vert \mathbb {1}_{A_n} u\Vert _\varphi \rightarrow 0 \ \ \text{ if } \ \lambda (A_n)\rightarrow 0\big \}, \end{array} \end{aligned}$$

\((L^\varphi )'=L^{\varphi ^*}.\)

The Orlicz-Bochner space \(L^\varphi (X)\) is defined by

$$\begin{aligned} \begin{array}{rl} L^\varphi (X) \,&{}\! \displaystyle :=\bigg \{g\in L^0(X):\int _\Omega \varphi (\alpha \,\Vert \,g(\omega )\Vert _X)\, d\lambda <\infty \ \ \text{ for } \text{ some } \ \alpha >0\bigg \} \\ &{}\!\!\!\displaystyle \,=\big \{g\in L^0(X):\Vert g(\cdot )\Vert _X\in L^\varphi \big \}. \end{array} \end{aligned}$$

For \(g\in L^\varphi (X)\), let

$$\begin{aligned} \Vert g\Vert _\varphi :=\Vert \ \Vert g(\cdot )\Vert _X \ \Vert _\varphi \ \text{ and } \ \Vert g\Vert ^0_\varphi :=\Vert \ \Vert g(\cdot )\Vert _X \ \Vert ^0_\varphi . \end{aligned}$$

Then

$$\begin{aligned} E^\varphi (X):=\left\{ g\in L^0(X):\Vert g(\cdot )\Vert _X\in E^\varphi \right\} . \end{aligned}$$

Assume that \(\varphi \) is Young function. If \(m:\Sigma \rightarrow X\) is a \(\lambda \)-continuous measure, we write \(|m|_{\varphi ^*}\) instead of \(|m|_{(L^\varphi )'}\).

Lemma 3.1

Assume that \(m:\Sigma \rightarrow X\) is a \(\lambda \)-continuous measure and \(\varphi \) is Young function. Then for \(A\in \Sigma \), we have

$$\begin{aligned} |m|(A)\le \left( \varphi ^{-1}\left( \frac{1}{\lambda (A)}\right) \right) ^{-1}|m|_{\varphi ^*}(A). \end{aligned}$$

Proof

Let \(A\in \Sigma \) and \(\varepsilon >0\) be given. Then there is a \(\Sigma \)-partition \((A_i)^n_{i=1}\) of A such that

$$\begin{aligned} |m|(A)\le \sum ^n_{i=1}\Vert m(A_i)\Vert _X+\left( \varphi ^{-1}\left( \frac{1}{\lambda (A)} \right) \right) ^{-1}\varepsilon . \end{aligned}$$

It is known that \(\Vert \mathbb {1}_A\Vert _\varphi =\left( \varphi ^{-1}\left( \frac{1}{\lambda (A)} \right) \right) ^{-1}\) (see [RR, Chap. 3.4, Corollary 7, p. 79]). Hence

$$\begin{aligned} \bigg \Vert \sum ^n_{i=1}\varphi ^{-1}\left( \frac{1}{\lambda (A)}\right) \mathbb {1}_{A_i}\bigg \Vert _\varphi = \; \bigg \Vert \varphi ^{-1}\left( \frac{1}{\lambda (A)}\right) \mathbb {1}_{A}\bigg \Vert _\varphi =1 \end{aligned}$$

and

$$\begin{aligned} \varphi ^{-1}\left( \frac{1}{\lambda (A)}\right) |m|(A)\le \sum ^n_{i=1} \varphi ^{-1}\left( \frac{1}{\lambda (A)}\right) \Vert m(A_i)\Vert _X+\varepsilon . \end{aligned}$$

It follows that \(\varphi ^{-1}\left( \frac{1}{\lambda (A)}\right) |m|(A)\le |m|_{\varphi ^*}(A)\), so

$$\begin{aligned} |m|(A)\le \varphi ^{-1}\left( \frac{1}{\lambda (A)}\right) ^{-1}|m|_{\varphi ^*}(A). \end{aligned}$$

\(\square \)

By \(\gamma [{\mathcal {T}}_\varphi ,{\mathcal {T}}_0\big |_{L^\varphi }]\) (in brief, \(\gamma _\varphi \)) we denote the natural mixed topology on \(L^\varphi \) (in the sense of Wiweger), that is, \(\gamma _\varphi \) is the finest linear topology that agrees with \({\mathcal {T}}_0\) on \({\mathcal {T}}_\varphi \)-bounded sets in \(L^\varphi \) (see [1, 25] for more details). Then \(\gamma _\varphi \) is a locally convex-solid Hausdorff topology and \(\gamma _\varphi \) and \({\mathcal {T}}_\varphi \) have the same bounded sets. This means that \((L^\varphi ,\gamma _\varphi )\) is a generalized DF-space (see [20]) and it follows that \((L^\varphi ,\gamma _\varphi )\) is quasinormable (see [20, p. 422]). Hence as a consequence of the Grothendieck’s classical result (see [20, p. 429]), we have

Proposition 3.2

For a linear operator \(S:L^\varphi \rightarrow X\) the following statements are equivalent:

  1. (i)

    S is \((\gamma _\varphi ,\Vert \cdot \Vert _X)\)-continuous and compact.

  2. (ii)

    S is \(\gamma _\varphi \)-compact, that is, there exists a \(\gamma _\varphi \)-neighborhood V of \(\,0\) in \(L^\varphi \) such that T(V) is relatively norm compact in X.

We say that a Young function \(\varphi \) increases essentially more rapidly than another \(\psi \) (in symbols, \(\psi \ll \varphi \)) if for arbitrary \(c>0\), \(\psi (ct)/\varphi (t)\rightarrow 0\) as \(t\rightarrow 0\) and \(t\rightarrow \infty \). Note that \(L^\varphi \subset E^\psi \) if \(\psi \ll \varphi \).

The following result will be useful (see [16, Theorem 2.1]).

Theorem 3.3

Let \(\varphi \) be a Young function. Then

  1. (i)

    \(\gamma _\varphi \) is generated by the family of norms \(\{\Vert \cdot \Vert _\psi \big |_{L^\varphi }:\psi \ll \varphi \}\).

  2. (ii)

    \((L^\varphi ,\gamma _\varphi )^*=\{F_v:v\in E^{\varphi ^*}\}\), where \(F_v(u)=\int _\Omega u(\omega )\, v(\omega )\,d\lambda \) for all \(u\in L^\varphi .\)

According to [15, Corollary 1.6] we have the following identity:

$$\begin{aligned} L^1=\bigcup _\varphi E^\varphi , \text{ where } \varphi \text{ runs } \text{ over } \text{ the } \text{ family } \text{ of } \text{ all } \text{ Young } \text{ functions. } \end{aligned}$$
(3.1)

Now we can state the main result in this section.

Theorem 3.4

Assume that a measure \(m:\Sigma \rightarrow X\) has the Radon-Nikodym property with the density \(g\in L^1(X)\). Then there exists a Young function \(\varphi \) such that \(g\in E^{\varphi ^*}(X)\) and the following statements hold:

  1. (i)

    The operator \(S_g:L^\varphi \rightarrow X\) defined by

    $$\begin{aligned} S_g(u):=\int _\Omega u(\omega )\, g(\omega )\, d\lambda \ \ \text{ for } \text{ all } \ u\in L^\varphi \end{aligned}$$

    is \((\gamma _\varphi ,\Vert \cdot \Vert _X)\)-continuous and compact.

  2. (ii)

    \(S_g\) is \(\gamma _\varphi \)-compact, that is, there exists a Young function \(\psi \) with \(\psi \ll \varphi \) such that \(\{\int _\Omega u(\omega )g(\omega )\,d\lambda : u\in L^\varphi ,\Vert u\Vert _\psi \le 1\}\) is a relatively norm compact subset of X.

  3. (iii)

    \(|m|_{\varphi ^*}(\Omega )<\infty \) and \(|m|_{\varphi ^*}(A)=\Vert \mathbb {1}_A g\Vert ^0_{\varphi ^*}\) for every \(A\in \Sigma \).

  4. (iv)

    \(|m|_{\varphi ^*}\) is \(\lambda \)-continuous, that is, \(|m|_{\varphi ^*}(A_n)\rightarrow 0\) if \(\lambda (A_n)\rightarrow 0\).

Proof

In view of (3.1) there exists a Young function \(\psi \) such that \(g\in E^\psi (X)\), that is, \(\Vert g(\cdot )\Vert _X\in E^\psi \). Hence \(|u| \ \Vert g(\cdot )\Vert _X\in L^1\) for every \(u\in L^{\psi ^*}\). Let us put \(\varphi =\psi ^*\). Then \(\varphi ^*=\psi ^{**}=\psi \) and \((L^\varphi )'=L^{\varphi ^*}=L^\psi \) and \(L^{\psi ^*}=L^\varphi \), and so \(g\in E^{\varphi ^*}(X)\). Then for \(u\in L^\varphi \), we have

$$\begin{aligned} \Vert S_g(u)\Vert _X\le \int _\Omega |u(\omega )| \ \Vert g(\omega )\Vert _X \,d\lambda =F_{\Vert g(\cdot )\Vert _X}(|u|), \end{aligned}$$

where \(\Vert g(\cdot )\Vert _X\in E^{\varphi ^*}\). The inequality shows that \(S_g\) is \((|\sigma |(L^\varphi ,E^{\varphi ^*}),\Vert \cdot \Vert _X)\)-continuous, where \(|\sigma |\) denotes the absolute weak topology. Since \((|\sigma |(L^\varphi ,E^{\varphi ^*})\) is the coarsest locally convex-solid topology on \(L^\varphi \) with dual \(E^{\varphi ^*}\), and by Theorem 3.3(ii) \(\gamma _\varphi \) is such a topology, it follows that \(S_g\) is \((\gamma _\varphi ,\Vert \cdot \Vert _X)\)-continuous.

To show that \(S_g\) is compact, choose a sequence \((f_n)\) of X-valued simple functions on \(\Omega \) such that \(\Vert f_n(\omega )-g(\omega )\Vert _X\rightarrow 0\) and \(\Vert f_n(\omega )\Vert _X\le \Vert g(\omega )\Vert _X\) \(\lambda \)-a.e. and for all \(n\in {\mathbb {N}}\) (see [7, Theorem 6, p. 4]). Hence for every \(\alpha >0\), we have that \(\varphi ^*(\alpha \Vert f_n(\omega )-g(\omega )\Vert _X)\rightarrow 0\) \(\lambda \)-a.e. and \(\varphi ^*(\alpha \Vert f_n(\omega )-g(\omega )\Vert _X) \le \varphi ^*(2\alpha \Vert g(\omega )\Vert _X)\) \(\lambda \)-a.e. and for all \(n\in {\mathbb {N}}\), where \(\int _\Omega \varphi ^*(2\alpha \Vert g(\omega )\Vert _X)\,d\lambda <\infty \). By the Lebesgue dominated convergence theorem, \(\int _\Omega \varphi ^*(\alpha \Vert f_n(\omega )-g(\omega )\Vert _X)\,d\lambda \rightarrow 0\) and this means that \(\Vert f_n-g\Vert ^0_{\varphi ^*}\rightarrow 0\).

For each \(n\in {\mathbb {N}}\), let \(S_n:L^\varphi \rightarrow X\) be a linear operator defined by

$$\begin{aligned} S_n(u):=\int _\Omega u(\omega )f_n(\omega )\,d\lambda \ \ \text{ for } \text{ all } \ u\in L^\varphi . \end{aligned}$$

Note that the range of each \(S_n\) is contained in the span of finite set of values of \(f_n\). Therefore \(S_n\) is compact and by the Hölder’s inequality for each \(u\in L^\varphi \), we have

$$\begin{aligned} \begin{array}{l} \displaystyle \Vert (S_n-S_g)(u)\Vert _X=\bigg \Vert \int _\Omega u(\omega )(f_n(\omega )-g(\omega ))\, d\lambda \,\bigg \Vert _X \\ \quad \displaystyle \le \int _\Omega |u(\omega )|\ \Vert f_n(\omega )-g(\omega )\Vert _X\, d\lambda \le \Vert u\Vert _\varphi \ \Vert f_n-g\Vert ^0_{\varphi ^*}. \end{array} \end{aligned}$$

It follows that \(\Vert S_n-S_g\Vert \rightarrow 0\), so \(S_g\) is a compact operator.

(ii) In view of (i) and Proposition 3.2\(S_g\) is \(\gamma _\varphi \)-compact. Using Theorem 3.3 we obtain that (ii) holds.

(iii) Let \(s=\sum ^k_{i=1} c_i\mathbb {1}_{A_i}\in {\mathcal {S}}(\Sigma )\) and \(\Vert s\Vert _\varphi \le 1\). Then

$$\begin{aligned} \sum ^k_{i=1}|c_i| \ \Vert m(A_i)\Vert _X=\sum ^k_{i=1}|c_i|\int _{A_i} \Vert g(\omega )\Vert _X \,d\lambda = \int _\Omega |s(\omega )| \ \Vert g(\omega )\Vert _X \,d\lambda \le \Vert g\Vert ^0_{\varphi ^*} \end{aligned}$$

and hence \(|m|_{\varphi ^*}(\Omega )\le \Vert g\Vert ^0_{\varphi ^*}\). Note that \(m=m_{S_g}\). Hence using Lemma 2.1 we have that \(|m_{\varphi ^*}|(A)=\Vert \mathbb {1}_A g\Vert ^0_{\varphi ^*}\) for \(A\in \Sigma \).

(iv) This follows from (iii) because \(g\in E^{\varphi ^*}(X)\). \(\square \)

Let \(i_\infty :L^\infty \rightarrow L^\varphi \) denotes the inclusion map. Note that \(i_\infty \) is \((\sigma (L^\infty ,L^1),\) \(\sigma (L^\varphi ,L^{\varphi ^*}))\)-continuous, and it follows that \(i_\infty \) is \((\tau (L^\infty ,L^1),\tau (L^\varphi ,L^{\varphi ^*}))\)-continuous (see [8, Theorem 8.6.1]). Since \(\gamma _\varphi \subset \tau (L^\varphi ,L^{\varphi ^*})\), we obtain that \(i_\infty \) is \((\tau (L^\infty ,L^1),\gamma _\varphi )\)-continuous.

As a consequence of Theorem 3.4 and Theorem 2.5, we can show that every \(\tau (L^\infty ,L^1)\)-nuclear operator \(T:L^\infty \rightarrow X\) admits a factorization through some Orlicz space \(L^\varphi \).

Corollary 3.5

Assume that \(T:L^\infty \rightarrow X\) is a \(\tau (L^\infty ,L^1)\)-nuclear operator. Then there exists a Young function \(\varphi \) such that \(T=S\circ i_\infty \), where

  1. (i)

    \(S:L^\varphi \rightarrow X\) is a Bochner representable and \(\gamma _\varphi \)-compact linear operator.

  2. (ii)

    \(|m_S|_{\varphi ^*}(A_n)\rightarrow 0\) if \(\lambda (A_n)\rightarrow 0\).

Proof

  1. (i)

    In view of Theorem 2.5 the representing measure \(m_T:\Sigma \rightarrow X\) has the Radon-Nikodym property with respect to \(\lambda \), that is, \(m_T=g\lambda \), where \(g\in L^1(X)\). Hence according to Theorem 3.4 there exists a Young function \(\varphi \) such that \(g\in E^{\varphi ^*}(X)\) and an operator \(S:L^\varphi \rightarrow X\) defined by

    $$\begin{aligned} S(u):=\int _\Omega u(\omega )\,g(\omega )\,d\lambda \ \text{ for } \text{ all } \ u\in L^\varphi , \end{aligned}$$

    is \(\gamma _\varphi \)-compact. Note that for \(u\in L^\infty \), \(T(u)=\int _\Omega u\,dm_T=\int _\Omega u(\omega )g(\omega )\,d\lambda =S(u)\).

  2. (ii)

    Since \(m_S(A)=\int _A g(\omega )\,d\lambda \) for all \(A\in \Sigma \), using Theorem 3.4 we have that \(|m_S|_{\varphi ^*}(A_n)\rightarrow 0\) if \(\lambda (A_n)\rightarrow 0\).

\(\square \)