Let X and Y be Banach spaces. We provide the representation of the dual space of compact operators K(X,Y) as a subspace of bounded linear operators L(X,Y). The main results are: (1) If Y is separable, then the dual forms of K(X,Y) can be represented by the integral operator and the elements of C[0,1]. (2) If X∗∗ has the weak Radon-Nikodym property, then the dual forms of K(X,Y) can be represented by the trace of some tensor products.
Citation: Keun Young Lee, Gwanghyun Jo. The dual of a space of compact operators[J]. AIMS Mathematics, 2024, 9(4): 9682-9691. doi: 10.3934/math.2024473
[1] | Sezer Erdem . Compact operators on the new Motzkin sequence spaces. AIMS Mathematics, 2024, 9(9): 24193-24212. doi: 10.3934/math.20241177 |
[2] | Heng Yang, Jiang Zhou . Compactness of commutators of fractional integral operators on ball Banach function spaces. AIMS Mathematics, 2024, 9(2): 3126-3149. doi: 10.3934/math.2024152 |
[3] | Carlos F. Álvarez, Javier Henríquez-Amador, John Millán G., Eiver Rodríguez . On the composition operator with variable integrability. AIMS Mathematics, 2025, 10(2): 2021-2041. doi: 10.3934/math.2025095 |
[4] | Saak Gabriyelyan . Limited type subsets of locally convex spaces. AIMS Mathematics, 2024, 9(11): 31414-31443. doi: 10.3934/math.20241513 |
[5] | Aydah Mohammed Ayed Al-Ahmadi . Differences weighted composition operators in several variables between some spaces of analytic functions. AIMS Mathematics, 2023, 8(11): 27363-27375. doi: 10.3934/math.20231400 |
[6] | Alaa M. Abd El-latif . Specific types of Lindelöfness and compactness based on novel supra soft operator. AIMS Mathematics, 2025, 10(4): 8144-8164. doi: 10.3934/math.2025374 |
[7] | Ateq Alsaadi, Manochehr Kazemi, Mohamed M. A. Metwali . On generalization of Petryshyn's fixed point theorem and its application to the product of n-nonlinear integral equations. AIMS Mathematics, 2023, 8(12): 30562-30573. doi: 10.3934/math.20231562 |
[8] | Yan Ling Fu, Wei Zhang . Some results on frames by pre-frame operators in Q-Hilbert spaces. AIMS Mathematics, 2023, 8(12): 28878-28896. doi: 10.3934/math.20231480 |
[9] | Lian Hu, Rong Yang, Songxiao Li . Embedding and Volterra integral operators on a class of Dirichlet-Morrey spaces. AIMS Mathematics, 2021, 6(7): 7782-7797. doi: 10.3934/math.2021453 |
[10] | Stevo Stević . Note on a new class of operators between some spaces of holomorphic functions. AIMS Mathematics, 2023, 8(2): 4153-4167. doi: 10.3934/math.2023207 |
Let X and Y be Banach spaces. We provide the representation of the dual space of compact operators K(X,Y) as a subspace of bounded linear operators L(X,Y). The main results are: (1) If Y is separable, then the dual forms of K(X,Y) can be represented by the integral operator and the elements of C[0,1]. (2) If X∗∗ has the weak Radon-Nikodym property, then the dual forms of K(X,Y) can be represented by the trace of some tensor products.
Let X and Y be Banach spaces. The spaces of compact operators K(X,Y) is the one of the most important spaces of Banach space theory. Many mathematicians have investigated the nature of K(X,Y) and their duals [1,2,3,4,5]. In particular, Feder and Saphar proved the following theorem.
Theorem 1.1. Suppose X∗∗ or Y∗ has the Radon-Nikodˊym property. For every ϕ∈K(X,Y)∗ and ε>0, there are (x∗∗n)⊂X∗∗ and (y∗n)⊂Y∗ such that ϕ(T)=∑∞n=1x∗∗nT∗(y∗n) for all T∈K(X,Y) where ∑∞n=1‖x∗∗n‖‖y∗n‖<‖ϕ‖+ε.
Lima and Oja applied this theorem to solve the famous metric approximation problem [6]. The first author of this work provided the generalized representation of K(X,Y)∗ concerning Feder and Saphar's theorem and the topological property of K(X,Y)∗, in the case that X∗∗ or Y∗ has the weak Radon-Nikodym property (weak RNP) [7].
Theorem 1.2. Let X and Y be Banach spaces such that X∗∗ or Y∗ has the weak Radon-Nikodˊym property. Then, for all ϕ∈K(X,Y)∗, there exist a sequence (((xni)∗∗)mni=1)∞n=1 in X∗∗ and a sequence (((yni)∗)mni=1)∞n=1 in Y∗ such that
ϕ(T)=limn→∞mn∑i=1(xni)∗∗(T∗((yni)∗)) |
for all T∈K(X,Y). Moreover, we have
lim supnmn∑i=1‖(xni)∗∗‖‖(yni)∗‖≤‖ϕ‖. |
Our aim in this paper is to provide the more general representation of K(X,Y)∗ concerning Feder and Saphar's theorem and the topological property of K(X,Y)∗. The main results are :
(1) If Y is separable, K(X,Y)∗ can be represented by the integral operator and the elements of C[0,1].
(2) If X∗∗ has the weak RNP, K(X,Y)∗ can be represented by the trace of some tensor products.
To prove the main theorems, we shall use the technique of two-trunk tree in Banach space which Lima and Oja developed [8]. Our paper is organized as follows : In Section 2, we introduce the concepts of two-trunk tree in Banach space. Then, for given Banach spaces X and Y, we provide the representation of K(X,Y)∗ in the case that Y is separable (see Theorem 2.4). In Section 3, we provide the new characterization of the weak RNP (see Theorem 3.2). By using this, we present the improved form of the representation of K(X,Y)∗ compared to Theorem B. Our conclusion follows in Section 4.
In this section, we provide the representation of K(X,Y)∗ when Y is separable. We need some lemmas and definitions. Let F(X,Y) be the space of finite rank operators. For S∈F(X,Y), let
‖S‖N0:=inf{l∑n=1‖x∗n‖‖yn‖:S=l∑n=1x∗n⊗yn}. |
As usual, L(X,Y) is the space of bounded linear operators, and I is the space of integral operators. The space of Pietsch integral operators from X into Y with Pietsch integral norm ‖⋅‖P is denoted by P(X,Y). Note that if T is a Pietsch integral operator, then T is an integral operator. Also, it is known that I(X,Y)=P(X,Y) if Y is a dual space. We denote by Smτc⟶R as m→∞ in L(X,Y) if for every compact subset K0 of X and every ε>0, there exists N∈N such that for all m⩾, \sup_{x\in K_0}\|S_{m}(x)-R(x)\| \leq \varepsilon .
Lemma 2.1. ([9], p. 102, Lemma 1) Let X , Y , and Z be Banach spaces. For every S \in \mathcal F(X, Y) and every R \in \mathcal I(Y, Z) , we have
\|RS\|_{\mathcal N^{0}}\leq \|R\|_{\mathcal I}\|S\|. |
Definition 2.1 Let X be a Banach space. We say that a system ((x_{k, 2^{n}})_{k = 0}^{2^{n}})_{n = 0}^{\infty} of elements of X is a two-trunk tree in X if for all n = 0, 1, \ldots and k = 1, 2, \ldots, 2^{n}-1
x_{k, 2^{n}} = \frac{1}{2}x_{2k-1, 2^{n+1}}+x_{2k, 2^{n+1}}+\frac{1}{2}x_{2k+1, 2^{n+1}}, |
x_{0, 2^{n}} = x_{0, 2^{n+1}}+x_{1, 2^{n+1}}, |
x_{2^{n}, 2^{n}} = \frac{1}{2}x_{2^{n+1}-1, 2^{n+1}}+x_{2^{n+1}, 2^{n+1}}. |
The \ell_{1} -tree space \ell_{1}^{tree}(X) consists of all two-trunk tree x = (x_{k, 2^{n}}) = ((x_{k, 2^{n}})_{k = 0}^{2^{n}})_{n = 0}^{\infty} in X such that
\|x\|: = \sup\limits_{n}\sum\limits_{k = 0}^{2^{n}}\|x_{k, 2^{n}}\| < \infty. |
By [8], \ell_{1}^{tree}(X) is a Banach space.
The scaling function \varphi: \mathbb{R} \rightarrow \mathbb{R} is defined by \varphi(t) = 1+t for t\in [-1, 0] , \varphi(t) = 1-t for t\in [0, 1] , and \varphi(t) = 0 for t\notin [-1, 1] . The functions g_{k, 2^{n}}, n = 0, 1, \ldots, k = 0, 1, \ldots, 2^{n} are defined by \varphi(2^{n}t-k), t\in[0, 1] . Then, g_{k, 2^{n}} satisfies the following:
g_{k, 2^{n}}(\frac{k}{2^{n}}) = 1, \quad g_{k, 2^{n}}(\frac{j}{2^{n}}) = 0, \quad j\neq k, |
g_{k, 2^{n}}(\frac{j}{2^{n+1}}) = 0, \quad j\notin \{2k-1, 2k, 2k+1\}, |
g_{k, 2^{n}}(\frac{j}{2^{n+1}}) = 1, \quad j = 2k, |
g_{k, 2^{n}}(\frac{j}{2^{n+1}}) = \frac{1}{2}, \quad j\in \{2k-1, 2k+1\}. |
Then, it is clear that ((g_{k, 2^{n}})_{k = 0}^{2^{n}})_{n = 0}^{\infty} is a two-trunk tree in C[0, 1] .
Theorem 2.1. ([8], Theorem 3.2) Let X be a Banach space. Then, \mathcal{P}(C[0, 1], X) is isometrically isomorphic to \ell_{1}^{tree}(X) , by mapping
T\rightarrow ((Tg_{k, 2^{n}})_{k = 0}^{2^{n}})_{n = 0}^{\infty}, T\in \mathcal{P}(C[0, 1], X). |
The inverse mapping
((x_{k, 2^{n}})_{k = 0}^{2^{n}})_{n = 0}^{\infty}\rightarrow T |
is given by
Tf = \lim\limits_{n}\sum\limits_{k = 0}^{2^{n}}f(\frac{k}{2^{n}})x_{k, 2^{n}}, f\in C[0, 1]. |
Now, we are in a position to state our main theorem.
Theorem 2.2. Suppose that Y is separable. If \phi \in (\mathcal K(X, Y), \|\cdot\|)^* , then there exists R\in \mathcal I(C[0, 1], X^{**}) such that
\phi(U) = \lim\limits_m\sum\limits_{k = 0}^{2^{m}}R(g_{k, 2^{m}})(U^{*}i^{*}(\delta_{k, 2^{m}})), \quad {for~~ every}~~ \ U \in \mathcal K(X, Y), |
where i: Y\rightarrow C[0, 1] is an isometry and \sup_{m}\sum_{k = 0}^{2^{m}}\|R(g_{k, 2^{m}})\|_{X^{**}} = \|\phi\| and \delta_{k, 2^{m}}\in C[0, 1]^{*} is given by \delta_{k, 2^{m}}(f) = f(\frac{k}{2^{m}}) for all f\in C[0, 1] .
Proof. Since Y is separable, there exists an isometry i : Y \rightarrow C[0, 1] . Thus, the map J_{1}: \mathcal K(X, Y) \rightarrow \mathcal K(X, C[0, 1]) defined by J_{1}(T) = iT is an isometry. Let J_2: \mathcal K(X, C[0, 1]) \rightarrow \mathcal K_{adj}(C[0, 1]^*, X^*): = \{T^* : T \in \mathcal K(X, C[0, 1])\} be the isometry via J_2(T) = T^* . Since C[0, 1] has the metric approximation property, \mathcal K(X, C[0, 1]) = \overline{\mathcal F(X, C[0, 1])}^{\|\cdot\|} , and so \mathcal K_{adj}(X, C[0, 1]) = \overline{\mathcal F_{adj}(X, C[0, 1])}^{\|\cdot\|} , which is isometrically isomorphic to the injective tensor product C[0, 1] \hat{\otimes}_{\varepsilon}X^* via the canonical isometry J_{3} .
\mathcal K(X, Y) \stackrel{J_{1}}{\longrightarrow}\mathcal K(X, C[0, 1]) \stackrel{J_{2}}{\longrightarrow} \mathcal K_{adj}(C[0, 1]^*, X^*) \stackrel{J_{3}}{\longrightarrow} C[0, 1] \hat{\otimes}_{\varepsilon}X^*. |
Let J: = J_3J_{2}J_{1} . Now, suppose that \phi \in (\mathcal{K}(X, Y), \|\cdot\|)^* . Then, \phi J^{-1}\in (J(\mathcal K(X, Y)), \|\cdot\|_{\varepsilon})^* . Choose a Hahn-Banach extension \widehat{\phi J^{-1}} \in (C[0, 1] \hat{\otimes}_{\varepsilon}X^*)^* of \phi J^{-1} . Let
\psi : \mathcal I(C[0, 1], X^{**}) \rightarrow (C[0, 1] \hat{\otimes}_{\varepsilon}X^*)^* |
be the canonical isometry ([10], Section 3). Let R: = \psi^{-1}(\widehat{\phi J^{-1}}) \in \mathcal I(C[0, 1], X^{**}) . By the well-known results of Grothendieck ([2], p.99), we have \mathcal P(C[0, 1], X^{**}) = \mathcal I(C[0, 1], X^{**}) as Banach spaces. Then, by Theorem 2.3, we have
R(f) = \lim\limits_{m}\sum\limits_{k = 0}^{2^{m}}\delta_{k, 2^{m}}(f)R(g_{k, 2^{m}}), f\in C[0, 1], |
and \| \phi \| = \|R\|_{\mathcal{I}} = \|R\|_{\mathcal{P}} = \sup_{m}\sum_{k = 0}^{2^{m}} \|R(g_{k, 2^{m}})\|_{X^{**}} . For each m\in \mathbb{N} , let S_{m} = \sum_{k = 0}^{2^{m}}\delta_{k, 2^{m}}\otimes R(g_{k, 2^{m}}) . From Theorem 2.3, it is clear that S_{m} converges to R pointwisely in \mathcal{I}(C[0, 1], X^{**}) . Moreover, by ([8], proof of Theorem 3.2), we have
\|S_{m}\|_{\mathcal{I}} = \|S_{m}\|_{\mathcal{P}} = \sum\limits_{k = 0}^{2^{m}}\|R(g_{k, 2^{m}})\|_{X^{**}} \leq \|R\|_{\mathcal{I}}. |
Or, put P_{m} = \sum_{k = 0}^{2^{m}}\delta_{k, 2^{m}}\otimes g_{k, 2^{m}} . Then, P_{m} is a projection from C[0, 1] into C[0, 1] with \|P_{m}\| = 1 . Since S_{m} = RP_{m} for all m , by Lemma 2.1 we have \|S_{m}\|_{\mathcal I} \leqslant \|S_{m}\|_{\mathcal N^{0}} \|R\|_{I} .
For every V \in \mathcal I(C[0, 1], X^{**}) , we denote by [\psi(V), \ \cdot ] the dual action on C[0, 1] \hat{\otimes}_{\varepsilon}X^{*} . Now, let U \in \mathcal K(X, Y) . Since \psi is an isometric isomorphism and S_{m} converges to R pointwisely, we have
\lim\limits_m [\psi(S_m), J(U)] = [\psi(R), J(U)]. |
For every m \in \mathbb{N} , by the definition of J and the dual action [\psi(\cdot), \ \cdot \] on C[0, 1] \hat{\otimes}_{\varepsilon}X^{*} , we have
[\psi(S_m), J(U)] = \sum\limits_{k = 0}^{2^{m}}R(g_{k, 2^{n}})((iU)^*\delta_{k, 2^{m}}). |
By the above arguments, we have the desired equation that
\begin{equation*} \begin{split} \phi (U) & = \phi J^{-1}(J(U)) \\ & = \psi \psi^{-1}(\widehat{ \phi J^{-1}})(J(U)) \\ & = [\psi(R), J(U)] \\ & = \lim\limits_m [\psi(S_m), J(U)] \\ & = \lim\limits_m \sum\limits_{k = 0}^{2^{m}}R(g_{k, 2^{n}})((iU)^*\delta_{k, 2^{m}}) \\ & = \lim\limits_m\sum\limits_{k = 0}^{2^{m}}R(g_{k, 2^{n}})(U^{*}i^{*}\delta_{k, 2^{m}}). \end{split} \end{equation*} |
Theorem 2.3. Suppose that Y is separable. If \phi \in (\mathcal K_{w^{*}}(X^{*}, Y), \|\cdot\|)^* , then there exists R\in \mathcal I(C[0, 1], X^{*}) such that
\phi(U) = \lim\limits_m\sum\limits_{k = 0}^{2^{m}}i^{*}(\delta_{k, 2^{m}})(UR(g_{k, 2^{m}})), \quad \forall U \in \mathcal K_{w^{*}}(X^{*}, Y), |
where i: Y\rightarrow C[0, 1] is an isometry and \sup_{m}\sum_{k = 0}^{2^{m}}\|R(g_{k, 2^{m}})\|_{X^*} = \|\phi \| and \delta_{k, 2^{m}}\in C[0, 1]^{*} is given by \delta_{k, 2^{m}}(f) = f(\frac{k}{2^{m}}) for all f\in C[0, 1] .
Proof. Since Y is separable, there exists an isometry i : Y \rightarrow C[0, 1] . Thus, the map J_{1} : \mathcal K(X, Y) \rightarrow \mathcal K(X, C[0, 1]) defined by J_{1}(T) = iT is an isometry. Recall that if a Banach space B has the approximation property (AP), then for every Banach space Z , we have \mathcal K_{w^{*}}(Z^{*}, B) = \overline{\mathcal F_{w^{*}}(Z^{*}, B)}^{\|\cdot\|} . Since C[0, 1] has AP, we have \mathcal K_{w^{*}}(X^{*}, C[0, 1]) = \overline{\mathcal F_{w^{*}}(X^{*}, C[0, 1])}^{\|\cdot\|} , which is isometric to the injective tensor product C[0, 1] \hat{\otimes}_{\varepsilon}X via the isometry J_{2} .
\mathcal K_{w^{*}}(X^{*}, Y) \stackrel{J_{1}}{\longrightarrow} \mathcal K_{w^{*}}(X^{*}, C[0, 1])\stackrel{J_{2}}{\longrightarrow} C[0, 1] \hat{\otimes}_{\varepsilon}X. |
Let J: = J_{2}J_{1} . Then, \phi J^{-1}\in (J(\mathcal K_{w^{*}}(X^{*}, Y)), \|\cdot\|_{\varepsilon})^* . Choose a Hahn-Banach extension \widehat{\phi J^{-1}} \in (C[0, 1] \hat{\otimes}_{\varepsilon}X)^* of \phi J^{-1} . Let \psi : \mathcal I(C[0, 1], X^*) \rightarrow (C[0, 1] \hat{\otimes}_{\varepsilon}X)^* be the canonical isometry ([10], Section 3). Let R: = \psi^{-1}(\widehat{\phi J^{-1}}) \in \mathcal I(C[0, 1], X^{*}) . By the well-known results of Grothendieck ([2], p. 99), we have \mathcal P(C[0, 1], X^{*}) = \mathcal I(C[0, 1], X^{*}) as Banach spaces. Then, by Theorem 2.3, we have
R(f) = \lim\limits_{m}\sum\limits_{k = 0}^{2^{m}}\delta_{k, 2^{m}}(f)R(g_{k, 2^{m}}), \quad f\in C[0, 1], |
and \| \phi \| = \|R\|_{\mathcal{I}} = \|R\|_{\mathcal{P}} = \sup_{m}\sum_{k = 0}^{2^{m}}\|R(g_{k, 2^{m}})\|_{X^*} . For each m\in \mathbb{N} , let S_{m} = \sum_{k = 0}^{2^{m}}\delta_{k, 2^{m}}\otimes R(g_{k, 2^{m}}) . From Theorem 2.3, it is clear that S_{m} converges to R pointwisely in \mathcal{I}(C[0, 1], X^{*}) and \|S_{m}\|\leq \|R\|_{I} . Then we obtain that S_m \stackrel{\tau_{c}}\longrightarrow R as m \rightarrow \infty . Moreover, by ([8], proof of Theorem 3.2), we have
\|S_{m}\|_{\mathcal{I}} = \|S_{m}\|_{\mathcal{P}} = \sum\limits_{k = 0}^{2^{m}}\|R(g_{k, 2^{m}})\|_{X^*}\leq \|R\|_{\mathcal{I}}. |
For every V \in \mathcal I(C[0, 1], X^{*}) , denote by [\psi(V), \ \cdot ] the dual action on C[0, 1] \hat{\otimes}_{\varepsilon}X . Now, let U \in \mathcal K(X, Y) . Since \psi is isometric isomorphism and S_{m} converges to R pointwisely, we have
\lim\limits_m [\psi(S_m), J(U)] = [\psi(R), J(U)]. |
For every m , by the definition of J and the dual action [\psi(\cdot), \ \cdot ] on C[0, 1] \hat{\otimes}_{\varepsilon}X^{*} ,
[\psi(S_m), J(U)] = \sum\limits_{k = 0}^{2^{m}}\delta_{k, 2^{m}}(iUR(g_{k, 2^{m}})). |
By the above arguments, we have
\begin{equation*} \begin{split} \phi (U) & = \phi J^{-1}(J(U)) \\ & = \psi\psi^{-1}(\widehat{\phi J^{-1}})(J(U)) \\ & = [\psi(R), J(U)] \\ & = \lim\limits_m [\psi(S_m), J(U)] \\ & = \lim\limits_m \sum\limits_{k = 0}^{2^{m}}\delta_{k, 2^{m}}(iUR(g_{k, 2^{m}})) \\ & = \lim\limits_m\sum\limits_{k = 0}^{2^{m}}i^{*}(\delta_{k, 2^{m}})(UR(g_{k, 2^{m}})). \end{split} \end{equation*} |
Now, suppose X is separable. Using the proof (b) of ([7], Corollary 4.4), we obtain the following.
Proposition 2.1. Suppose that X is separable. If \phi \in (\mathcal K_{w^{*}}(X^{*}, Y), \|\cdot\|)^* , then there exists R\in \mathcal I(C[0, 1], Y^{*}) such that
\phi(U) = \lim\limits_m\sum\limits_{k = 0}^{2^{m}}R(g_{k, 2^{m}})(Ui^{*}(\delta_{k, 2^{m}})), \quad \forall U \in \mathcal K_{w^{*}}(X^{*}, Y), |
where i: X\rightarrow C[0, 1] is an isometry, \sup_{m}\sum_{k = 0}^{2^{m}}\|R(g_{k, 2^{m}})\|_{Y^*} = \|\phi\| , and \delta_{k, 2^{m}}\in C[0, 1]^{*} is given by \delta_{k, 2^{m}}(f) = f(\frac{k}{2^{m}}) for all f\in C[0, 1] .
In this section, we provide the representation of \mathcal{K}(X, Y)^{*} when X^{**} has the weak RNP. We need some definitions. First, we introduce the definition of weak RNP. To do this, we have to define weakly integrable functions. Let (\Omega, \Sigma, \mu) be a finite complete measure space. We recall that a \mu -measurable function f: \Omega \rightarrow X is said to be weakly integrable with respect to \mu if the function x^*f is integrable for every x^* \in X^* . We say that the weakly integrable function f is Pettis integrable if the Dunford integral \int_{E}f d\mu belongs to X for every E\in \Sigma . A Banach space X has the weak RNP if for each finite complete measure space (\Omega, \Sigma, \mu) and each \mu -continuous X -valued countably additive vector measure m: \Sigma\rightarrow X of bounded variation, there exists a Pettis integrable function f: \Omega\rightarrow X such that
m(E) = \int_{E}f d\mu, \quad \forall E\in \Sigma. |
Musial proved that a Banach space X contains no isomorphic copy of \ell_{1} if and only if X^* has the weak RNP [11].
Now, we provide the definition of p -integral and p -nuclear operators. Let K be a compact topological space and \mathcal B_{K} be the \sigma -algebra of Borel sets on K . Recall that if T: C(K)\rightarrow X is a bounded linear operator, then \mu: \mathcal B_{K}\rightarrow X^{**} is called the representing measure for the operator T if \mu(E) = T^{**}(\chi_{E}). The vector sequence \{x_n \} in X is weakly p -summable if the scalar sequences \{x^*(x_n)\} are in \ell_{p} for every x^* \in X^* . We denote by
\ell_{p}^{w}(X) |
the set of all such sequences in X . A Banach space operator u: X \rightarrow Y is called p -integral ( 1\leqslant p \leqslant \infty ) if there are a probability measure \mu and bounded linear operators a: L_{p}(\mu) \rightarrow Y^{**} and b: X \rightarrow L_{\infty}(\mu) such that Q_{Y}u = ai_{p}b where i_{p}: L_{\infty}(\mu) \rightarrow L_{p}(\mu) and Q_{Y}: Y \rightarrow Y^{**} is the canonical isometric embedding. A Banach space operator u: X \rightarrow Y is called strictly p -integral ( 1\leqslant p \leqslant \infty ) if there is a probability measure \mu and bounded linear operators a: L_{p}(\mu) \rightarrow Y and b: X \rightarrow L_{\infty}(\mu) such that u = ai_{p}b . A Banach space operator u: X \rightarrow Y is called p -nuclear ( 1\leqslant p < \infty ) if there are operators a\in \mathcal{L}(\ell_{p}, Y) , b\in \mathcal{L}(X, \ell_{\infty}) and a sequence \lambda\in \ell_{p} such that
u = aM_{\lambda}b, |
where M_{\lambda}: \ell_{\infty} \rightarrow \ell_{p}: (\xi_{n}) \rightarrow (\lambda_{n}\xi_{n}) . We define the p -nuclear norm as
\nu_{p}(u): = \inf \|a\| \cdot \|M_{\lambda}\| \cdot \|b\|, |
the infimum being extended over all factorizations above. We denote by
\mathcal{N}_{p}(X, Y) |
the space of all p -nuclear operators from X into Y with the norm \nu_{p} . Throughout the rest of the paper, p^* is the Hölder's conjugate of p , i.e., 1/p+1/p^* = 1 .
Now, we provide a new characterization of the weak RNP property for the dual version.
Theorem 3.1. X^{*} has the weak RNP if and only if R\in \mathcal I(C(K), X^{*}) , then R\in \mathcal{N}_{p}(C(K), X^{*}) for all p > 1 . That is, there exist (z_{n}^{*})\in \ell_{p}(C(K)^{*}) and (x_{n}^{*})\in \ell_{p^*}^{w}(X^{*}) such that
R = \sum\limits_{n = 1}^{\infty}z_{n}^{*}\otimes x_{n}^{*} |
and the series converges in \mathcal{L}(C(K), X^{*}) .
Proof. Suppose that if R\in \mathcal I(C(K), X^{*}) , then R\in \mathcal{N}_{p}(C(K), X^{*}) for all p > 1 . Take any R\in \mathcal I(C(K), X^{*}) and p > 1 . Then, R is p -nuclear. It is clear that R is compact. By ([7], Lemma 3.1), X^{*} has the weak RNP.
Conversely, suppose that X^{*} has the weak RNP. Let R: C(K)\rightarrow X^{*} be an integral operator and p > 1 . By ([10], Proposition 5.28), the representing measure \mu is a vector measure of bounded variation with values in X^{*} and R = S\circ J where J: C(K)\rightarrow L_{1}(|\mu|_{1}) is the natural injection and S: L_{1}(|\mu|_{1})\rightarrow X^{*} is a bounded linear operator via
S(f) = \int_{K} f d|\mu|, \quad \forall f\in L_{1}(|\mu|_{1}). |
By the assumption, we have \ell_{1}\nsubseteq X , and \mu has a relatively compact range by Lemma 3.53 in [12]. By ([12], Proposition 3.56), S is completely continuous. Also, since J is a Piesch integral operator, J is strictly p -integral. By ([13], p.124), R is p -nuclear.
Now, we are in a position to state our main theorem.
Theorem 3.2. Suppose that X^{**} has the weak RNP. If \phi \in (\mathcal K(X, Y), \|\cdot\|)^* , then there exists (z_{n}^{*})\in \ell_{p}(C(K)^{*}) and (x_{n}^{**}) \in \ell_{p^{*}}^{w}(X^{**}) such that
\phi (U) = \sum\limits_{n = 1}^{\infty} x_{n}^{**}((U^* i^*(z_{n}^*)) \ \ {for ~~every}~~ \ U \in \mathcal K(X, Y), |
where K = B_{Y^{*}} and i: Y\rightarrow C(K) by
i(y)(y^*) = \delta_{y}(y^{*}) = y^{*}(y), \forall y^{*}\in B_{Y^{*}} |
is an isometry and the series
\sum\limits_{n = 1}^{\infty}z_{n}^{*}\otimes x_{n}^{**} |
converges in \mathcal{L} (C(K), X^{**}) .
Proof. Suppose that X^{**} has the weak RNP. Let K be B_{Y^{*}} . Then K is weak ^* compact subset of Y^* . Define i: Y\rightarrow C(K) by
i(y) = \delta_{y}(y^{*}) = y^{*}(y), \forall y^{*}\in B_{Y^{*}}. |
Then, i is the canonical isometry of Y into C(K) . Thus the map J_{1} : \mathcal K(X, Y) \rightarrow \mathcal K(X, C(K)) defined by J_{1}(T) = iT is an isometry. Let J_2 : \mathcal K(X, C(K)) \rightarrow \mathcal K_{adj}(C(K)^*, X^*): = \{T^* : T \in \mathcal K(X, C(K))\} be the isometry via J_2(T) = T^* . Since C(K) has the metric approximation property, \mathcal K(X, C(K))) = \overline{\mathcal F(X, C(K)))}^{\|\cdot\|} and so \mathcal K_{adj}(X, C(K))) = \overline{\mathcal F_{adj}(X, C(K)))}^{\|\cdot\|} which is isometrically isomorphic to the injective tensor product C(K) \hat{\otimes}_{\varepsilon}X^* via the canonical isometry J_{3} .
\mathcal K(X, Y) \stackrel{J_{1}}{\longrightarrow}\mathcal K(X, C(K)) \stackrel{J_{2}}{\longrightarrow} \mathcal K_{adj}(C(K)^*, X^*) \stackrel{J_{3}}{\longrightarrow} C(K) \hat{\otimes}_{\varepsilon}X^*. |
Let J: = J_3J_{2}J_{1} . Now, suppose that \phi \in (\mathcal K(X), \|\cdot\|)^* . Then \phi J^{-1}\in (J(\mathcal K(X, Y)), \|\cdot\|_{\varepsilon})^* . Choose a Hahn-Banach extension \widehat{\phi J^{-1}} \in (C(K) \hat{\otimes}_{\varepsilon}X^*)^* of \phi J^{-1} . Let
\psi : \mathcal I(C(K), X^{**}) \rightarrow (C(K) \hat{\otimes}_{\varepsilon}X^*)^* |
be the canonical isometry ([10], Section 3). Let R: = \psi^{-1}(\widehat{\phi J^{-1}}) \in \mathcal I(C(K), X^{**}) . By the well-known results of Grothendieck ([2], p. 99), we have \mathcal P(C(K), X^{**}) = \mathcal I(C(K), X^{**}) as Banach spaces. Then, by Theorem 3.1, there exist (z_{n}^{*})\in \ell_{p}(C(K)^{*}) and (x_{n}^{**})\in \ell_{p^{*}}^{w}(X^{**}) such that
R = \sum\limits_{n = 1}^{\infty}z_{n}^{*}\otimes x_{n}^{**} |
converges in \mathcal{L}(C(K), X^{**}) . Then, we have
R(f) = \sum\limits_{n = 1}^{\infty}z_{n}^{*}(f) x_{n}^{**}, \quad \forall f\in C(K). |
For each m\in \mathbb{N} , let S_{m} = \sum_{n = 1}^{m}z_{n}^{*}\otimes x_{n}^{**} . From Theorem 2.3, it is clear that S_{m} converges to R pointwisely in \mathcal{I}(C(K), X^{*}) and \|S_{m}\|\leq \|R\|_{I} . Then, we obtain that S_m \stackrel{\tau_{c}}\longrightarrow R as m \rightarrow \infty .
For every V \in \mathcal I(C(K), X^{**}) , denote by [\psi(V), \ \cdot ] the dual action on C(K) \hat{\otimes}_{\varepsilon}X^{*} . Now, let U \in \mathcal K(X, Y) . Since \psi is an isometric isomorphism and S_{m} converges to R pointwisely, we have
\lim\limits_m [\psi(S_m), J(U)] = [\psi(R), J(U)]. |
For every m \in \mathbb{N} , by the definition of J and the dual action [\psi(\cdot), \ \cdot ] on C(K) \hat{\otimes}_{\varepsilon}X^{*} ,
[\psi(S_m), J(U)] = \sum\limits_{n = 1}^{m}x_{n}^{**}((iU)^*(z_n^{*})). |
Hence, we have
\begin{equation*} \begin{split} \phi(U) & = \phi J^{-1}(J(U)) \\ & = \psi\psi^{-1}(\widehat{\phi J^{-1}})(J(U)) \\ & = [\psi(R), J(U)] \\ & = \lim\limits_m [\psi(S_m), J(U)] \\ & = \sum\limits_{n = 1}^{\infty} x_{n}^{**}((iU)^*(z_{n}^*)). \end{split} \end{equation*} |
In this work, we provide new representations of the dual of a space of compact operators K(X, Y) under the separable condition of Y or weak RNP condition of X^{**} . The dual of K(X, Y) can be represented by the integral operator and the elements of C[0, 1] if Y is separable. On the other hand, the dual of K(X, Y) can be represented by the trace of some tensor products if X^{**} has the weak RNP.
The authors declare they have not used Artificial Intelligence (AI) tools in the creation of this article.
This work was supported by the National Research Foundation of Korea (NRF) grant funded by the Korea government (MSIT) (No. 2020R1C1C1A01005396).
All authors declare no conflicts of interest in this paper.
[1] |
M. Feder, P. Saphar, Spaces of compact operators and their dual spaces, Israel J. Math., 21 (1975), 38–49. https://doi.org/10.1007/BF02757132 doi: 10.1007/BF02757132
![]() |
[2] | A. Grothendieck, Produits tensoriels topologiques et espaces nucléaires, Providence: American Mathematical Society, 16 (1955), 193–200. |
[3] |
J. Johnson, Remarks on Banach spaces of compact operators, J. Funct. Anal., 32 (1979), 304–311. https://doi.org/10.1016/0022-1236(79)90042-9 doi: 10.1016/0022-1236(79)90042-9
![]() |
[4] |
N. J. Kalton, Spaces of compact operators, Math. Ann., 208 (1974), 267–278. https://doi.org/10.1007/BF01432152 doi: 10.1007/BF01432152
![]() |
[5] | W. Ruess, Duality and geometry of spaces of compact operators, In: North-Holland Mathematics Studies, 1984. |
[6] |
Å. Lima, E. Oja, The weak metric approximation property, Math. Ann., 333 (2005), 471–484. https://doi.org/10.1007/s00208-005-0656-0 doi: 10.1007/s00208-005-0656-0
![]() |
[7] |
K. Y. Lee, Dual spaces of compact operator spaces and the weak Radon-Nikod\acute{ \rm y }m property, Stud. Math., 210 (2012), 247–260. https://doi.org/10.4064/sm210-3-5 doi: 10.4064/sm210-3-5
![]() |
[8] |
Å. Lima, V. Lima, E. Oja, Absolutely summing operators on C[0, 1] as a tree space and the bounded approixmation property, J. Funct. Anal., 259 (2010), 2886–2901. https://doi.org/10.1016/j.jfa.2010.07.017 doi: 10.1016/j.jfa.2010.07.017
![]() |
[9] | A. Pietsch, Operator ideals, North-Holland Publishing Company, 1980. |
[10] | R. A. Ryan, Introduction to tensor products of Banach spaces, London: Springer, 2002. https://doi.org/10.1007/978-1-4471-3903-4 |
[11] | K. Musiał, The weak Radon-Nikod\acute{y}m property in Banach space, Stud. Math., 64 (1979), 151–174. |
[12] | S. Okada, W. Ricker, E. A. Sánchez Pérez, Optimal domain and integral extension of operators, In: Operator theory: Advances and applications, Springer, 2008. https://doi.org/10.1007/978-3-7643-8648-1 |
[13] | J. Distel, H. Jarchow, A. Tonge, Absolutely summing operators, Cambridge university press, 1995. |