Hostname: page-component-745bb68f8f-lrblm Total loading time: 0 Render date: 2025-02-05T18:55:47.543Z Has data issue: false hasContentIssue false

Differentiability of the operator norm on $\ell _p$ spaces

Published online by Cambridge University Press:  16 December 2024

Sreejith Siju*
Affiliation:
Kerala School of Mathematics, Kozhikode, 673 571, India
Rights & Permissions [Opens in a new window]

Abstract

In this paper, we present a characterization of strong subdifferentiability of the norm of bounded linear operators on $\ell _p$ spaces, $1\leq p<\infty $. Furthermore, we prove that the set of all bounded linear operators in ${B}(\ell _p, \ell _q)$ for which the norm of ${B}(\ell _p, \ell _q)$ is strongly subdifferentiable is dense in ${B}(\ell _p, \ell _q)$. Additionally, we present a characterization of Fréchet differentiability of the norm of bounded linear operators from $\ell _p$ to $\ell _q$, where $1 < p, q < \infty $. Applying this result, we will show that the Fréchet differentiability and the Gateaux differentiability of the norm of bounded linear operators on $\ell _p$ spaces coincide, extending a known theorem regarding the operator norm on Hilbert spaces.

Type
Article
Copyright
© The Author(s), 2024. Published by Cambridge University Press on behalf of Canadian Mathematical Society

1 Introduction

The differentiability of the norm is a very useful tool in understanding the geometry of the underlying Banach space. Various notions of differentiability of the norm of Banach spaces include Fréchet, Gateaux, and strong subdifferentiability (referred to as $SSD$ ) (see Definitions 2.1 and 3.2). It is well known that a point x in a Banach space X is a smooth point if and only if the norm of X is Gateaux differentiable at x. We employ the terms Gateaux differentiability and smoothness interchangeably. The Fréchet differentiability of the norm of a Banach space at point x is equivalent to the existence of the unique support functional at x which is strongly exposed by x. Strong subdifferentiability provides a non-smooth extension of Fréchet differentiability, relating Fréchet and Gateaux differentiability. That is, the norm of a Banach space is Fréchet differentiable at a point x if and only if it is Gateaux differentiable and strongly subdifferentiable at x [Reference Franchetti and Payá10]. The monographs [Reference Diestel9] and [Reference Deville, Godefroy and Zizler8] provide a detailed discussion of the concepts Fréchet and Gateaux differentiability in Banach spaces. For a detailed study of strong subdifferentiability of the norm, we recommend referring to the discussions found in [Reference Franchetti and Payá10Reference Godefroy, Montesinos and Zizler12], and the references therein.

This article primarily investigates the strong subdifferentiability of operator norm over the sequence spaces $\ell _p$ , where $1<p<\infty $ . Additionally, we study the Fréchet and Gateaux differentiability within the realm of operators acting on $\ell _p$ spaces.

The strong subdifferentiability of Banach space norms has been studied extensively in the literature. We refer the reader [Reference Franchetti and Payá10] for the basic theory of strong subdifferentiability. Strong subdifferentiability has been also studied in the context of dual spaces, operator norm, and $JB^*$ -triples (see [Reference Guerrero and Palacios2, Reference Contreras, Payá and Werner4, Reference Godefroy, Indumathi and Lust-Piquard11, Reference Rao21]). The connection between strong subdifferentiability and Bishop–Phelps Bollobas properties has been explored in [Reference Dantas, Kim, Lee and Mazzitelli6]. To see connections between strong subdifferentiability and other notions in Banach space theory, we refer [Reference Contreras3, Reference Guerrero and Peralta13, Reference Martínez-Moreno, Mena-Jurado, Payá-Albert and Rodríguez-Palacios18, Reference Rao22].

An interesting characterization of the strong subdifferentiability of the dual norm of a Banach space X at a bounded linear functional f has been obtained in [Reference Godefroy, Indumathi and Lust-Piquard11] by G. Godefroy, V. Indumathi, and F. Lust-Piquard. The characterizing condition for strong subdifferentiability is established in terms of the set of norm-attaining points of f (see Theorem 2.1). Moreover, in [Reference Contreras, Payá and Werner4], M. Contreras, R. Paya, and W. Werner characterized the strong subdifferentiability of the operator norm of Hilbert spaces (see Theorem 2.3), and showed that the collection of all strongly subdifferentiable points is dense in the space of bounded linear operators on a Hilbert space. Subsequently, J. Guerrero and A. Palacios broadened this characterization of strong subdifferentiability to encompass $JB^*$ -triples [Reference Guerrero and Palacios2].

In Theorem 2.4, we will show that the characterizing conditions of strong subdifferentiability obtained in [Reference Godefroy, Indumathi and Lust-Piquard11, Theorem 2.1] and [Reference Contreras, Payá and Werner4, Theorem 2.3] are essentially the same if we view from the perspective of maximizing sequence (see Definition 2.3). This equivalence gives rise to two natural questions in this context.

Question 1 Is it possible to extend Theorem 2.4 to bounded linear operators on Banach spaces?

Question 2 Is the set of strongly subdifferentiable points of ${B}(X)$ , the space of bounded linear operators on X, dense in ${B}(X)$ ?

In Section 2, we will show that Question 1 has an affirmative answer in the case of $\ell _p$ spaces, $1<p<\infty $ . We obtain a characterization for the strong subdifferentiability of operator norm on $\ell _p$ spaces in terms of the maximizing sequence of an operator, as observed in the case of dual norm and operator norm on Hilbert space. We also prove that such operators attain their norm. Our idea here is to first formulate a condition that is equivalent to the characterization of strong subdifferentiability of operator norm Hilbert spaces found in [Reference Contreras, Payá and Werner4] (see Theorem 2.4). Then we will prove the equivalent condition characterizes the strong subdifferentiability of operator norm on $\ell _p$ spaces. We will also address the borderline cases $p=1,\infty $ toward the end of this section.

In Section 3, we will give an affirmative answer to Question 2 in the case of $\ell _p$ spaces, $1\leq p <\infty $ . We will exhibit that, similar to the case of operator norm on Hilbert spaces, the set of all strongly subdifferentiable points is norm dense in the space of bounded linear operators on $\ell _p$ spaces. This will also extend, in $\ell _p$ space situation, a result of J. Lindenstrauss [Reference Lindenstrauss17] regarding norm-attaining operators to the class of strongly subdifferentiable points of the space of bounded linear operator.

In [Reference Kittaneh and Younis16], F. Kittaneh and R. Younis presented a characterization of smooth points in the space of bounded linear operators on a Hilbert space, ${B}(H)$ , using the essential norm of an operator. Subsequently, W. Deeb and R. Khalil extended this characterization to include operators on $\ell _p$ spaces [Reference Deeb and Khalil7]. In another work [Reference Taylor and Werner26] by W. Werner and K. F. Taylor, it was exhibited that for a bounded linear operator T on a Hilbert space H, the assumption made by F. Kittaneh and R. Younis regarding the essential norm of T goes beyond characterizing smoothness and serves as a characterization for the Fréchet differentiability of the norm at T. Thereby, proving that Fréchet and Gateaux differentiability coincide in ${B}(H)$ . Furthermore, it has been independently proved in [Reference Abatzoglou1] that the equivalence between Fréchet and Gateaux differentiability holds in ${B}(H)$ . So, it is natural to inquire whether the coincidence exists between the Fréchet and Gateaux differentiability of the norm of the space of bounded linear operators on $\ell _p$ spaces.

Toward the conclusion of Section 3, we establish, as a consequence of the results on strong subdifferentiability, the equivalence between Fréchet and Gateaux differentiability of the operator norm on $\ell _p$ spaces. Further, we will show that the assumption made by F. Kittaneh and R. Younis regarding the essential norm of T serves as a characterization for the Fréchet differentiability of operator norm on $\ell _p$ spaces (see Proposition 3.9).

Notations: In this article, we will use X and Y to represent infinite dimensional Banach spaces, H to denote an infinite dimensional Hilbert space, and the scalar fields are either real or complex, denoted by $\mathbb {F}$ . When referring to a subspace of a Banach space, we will always assume it to be a closed subspace. The closed unit ball and the unit sphere of a Banach space X will be denoted by $B_X$ and $S_X$ , respectively. We consider every Banach space X as a subspace of its bidual $X^{**}$ through the canonical embedding. For a subset Y of a Banach space X, the annihilator of Y in the dual space $X^*$ is denoted by $Y^\perp $ , defined as $Y^\perp =\{f\in X^*: f(y)=0 \quad \forall y\in Y\}$ . The convex hull of a set S is the smallest convex set containing S, denoted by $\operatorname {conv}(S)$ . The closure of $\operatorname {conv}(S)$ is represented by $\overline {\operatorname {conv}}(S)$ . For a Banach space X, the extreme points of $B_X$ will be denoted by $\operatorname {ext} B_{X}$ . For a bounded subset S of X, the diameter of S will be denoted by $\operatorname {diam}S$ .

Let X and Y be Banach spaces. We denote by ${B}(X, Y)$ , the space of all bounded linear operators from X to Y endowed with the usual operator norm. The space of all compact operators from X to Y is denoted by ${K}(X, Y)$ . We denote $B(X\times Y)$ to be the space of all bounded bilinear functional from $X\times Y$ to $\mathbb {F}$ . When $X=Y$ , we abbreviate ${B}(X, Y)$ as ${B}(X)$ , and ${K}(X, Y)$ as ${K}(X)$ . The essential norm of a bounded linear operator $T\in {B}(X, Y)$ is defined as the distance from T to the space of compact operators, denoted by $\|T\|_e$ . For an operator $T\in {B}(X, Y)$ , we define the set $M_T$ to be the set $M_T=\{x\in S_X: \|Tx\|=\|T\|\}$ . We call the set $M_T$ , the norm-attaining set of T. For, $1\leq p<\infty $ , the direct sum of X and Y endowed with the norm $\|(x, y)\| = (\|x\|^p+\|y\|^p)^{\frac {1}{p}}$ is denoted by $X\oplus _p Y$ . We denote by $X\hat {\otimes }_\pi Y$ , and $X\check {\otimes }_\varepsilon Y$ , the projective and injective tensor product of X and Y, respectively.

2 Strong subdifferentiability of operator norm

In this section, we will characterize strong subdifferentiability of the norm of ${B}(\ell _p,\ell _q)$ at an operator $T\in {B}(\ell _p,\ell _q)$ in terms of maximizing sequence of T, where $1<p,q<\infty $ . To begin, we will review some definitions. We will first recall the definition of strong subdifferentiability.

Definition 2.1 The norm of a Banach space X is strongly subdifferentiable at a point $u\in X$ if

(2.1) $$ \begin{align} \lim_{t \rightarrow 0^{+}} \frac{\|u+t x\|-\|u\|}{t}= \operatorname{Max}\{\operatorname{Re} f(x) : f \in J_{X^*}(u)\} \end{align} $$

uniformly for $x \in B_X$ , where $J_{X^*}(u)=\left \{f \in S_{X^*} : f(u)=\|u\|\right \}$ .

If this happens, we say that u is an $SSD$ point of X or simply u is an $SSD$ point when the Banach space under consideration is already known. It is evident that an element $u \in X$ is an $SSD$ point of X if and only if $\frac {u}{\|u\|}$ is an $SSD$ point of X. Hence, in the subsequent discussion, we will consider strong subdifferentiability at points on the unit sphere of a Banach space.

We need the following two definitions in the sequel.

Definition 2.2 Let X and Y be Banach spaces. A bounded linear operator $T : X \rightarrow Y$ is said to attain its norm if there exists an element $x \in S_X$ such that

$$ \begin{align*}\|Tx\|=\|T\|. \end{align*} $$

Definition 2.3 A maximizing sequence for $T \in {B}(X, Y)$ is a sequence $\left \{x_n\right \}$ in X with $x_n\in S_X$ for all n so that $\{\left \|Tx_n\right \|\}$ converges to $\|T\|$ as $n \rightarrow \infty $ .

We know, from [Reference Franchetti and Payá10], an element u is an $SSD$ point of X if and only if the face $J_{X^*}(u)$ is strongly exposed by u, in the sense that the distance $d\left (f_n, J_{X^*}(u)\right )$ tends to zero for any sequence $\{f_n\}$ in $B_{X^*}$ satisfying $f_n(u) \rightarrow 1$ . In [Reference Godefroy, Indumathi and Lust-Piquard11], it is observed that for the dual norm of a Banach space X to be strongly subdifferentiable at a bounded linear functional $f\in S_{X^*}$ , it is enough to work with the set $ J_{X}(f)=\left \{x \in B_X:\|f\|=f(x)=1\right \}$ instead of $J_{X^{**}}(f)$ . We will now recall this result from [Reference Godefroy, Indumathi and Lust-Piquard11].

Theorem 2.1 [Reference Godefroy, Indumathi and Lust-Piquard11, Proposition 2.2]

Let X be a Banach space and $f \in S_{X^*}$ . Then the following are equivalent.

  1. (i) The norm of $X^*$ is strongly subdifferentiable at f.

  2. (ii) $J_{X}(f)\neq \emptyset $ , and for every sequence $\{x_n\}$ in $B_X$ satisfying $f(x_n)\rightarrow 1$ , there exists a subsequence $\{x_{n_i}\}$ of $\{x_n\}$ such that $d(x_{n_i}, J_{X}(f))\rightarrow 0.$

Remark 2.2 It is easy to see that condition (ii) above is equivalent to the statement, $J_{X}(f)\neq\emptyset $ , and for each $\varepsilon>0$ there exists $\delta>0$ such that whenever $x\in S_X$ satisfies $f(x)>1-\delta $ , we have $d(x, J_{X}(f))<\varepsilon $ . In the subsequent sections of this paper, we use these two equivalent conditions interchangeably.

Similarly, it is observed in [Reference Contreras, Payá and Werner4, Theorem 1] that an analogous characterization holds for the norm of $C^*$ -algebra, in particular for the operator norm on Hilbert spaces. We will now recall this result.

Theorem 2.3 [Reference Contreras, Payá and Werner4, Theorem 1]

Let $\mathcal {A}$ be a $C^*$ -algebra and $a \in S_{\mathcal {A}}$ . The following assertions are equivalent.

(i) The norm of $\mathcal {A}$ is strongly subdifferentiable at a.

(ii) 1 is an isolated point in the spectrum of $|a|$ .

Assume that in the above theorem the $C^*$ -algebra under consideration is ${B}(H)$ and the element $a\in S_{\mathcal {A}}$ corresponds to an operator $T\in S_{{B}(H)}$ . Then we have the following equivalent characterization of strong subdifferentiability of operator norm on Hilbert spaces in terms of maximizing sequences/norm-attaining set of T.

Theorem 2.4 Let H be a Hilbert space and $T\in S_{{B}(H)}$ . Then the following are equivalent.

  1. (i) The norm of ${B}(H)$ is strongly subdifferentiable at T.

  2. (ii) $M_T\neq \emptyset $ and for every maximizing sequence $\{x_n\}$ of T, there exists a subsequence $\{x_{n_i}\}$ of $\{x_n\}$ such that $d(x_{n_i}, M_T)\rightarrow 0$ .

  3. (iii) T attains its norm and $\|P_T|_{{M_T}^\perp }\|<1$ , where $P_T=(T^*T)^{1/2}.$

Proof It is enough to show that $(ii)$ and $(iii)$ are equivalent, since $(i)\Leftrightarrow (iii)$ follows from [Reference Contreras, Payá and Werner4, Theorem 1, proof of $(ii)\Rightarrow \left( ii i\right)$ ].

Assume that $(iii)$ holds and write $M=\overline {\operatorname *{\mathrm {span}}}M_T$ , recall that

$$ \begin{align*}M_T=\{x\in S_H: \|Tx\|=\|T\|=1\}.\end{align*} $$

Let $\{x_n\}$ in $S_H$ be a maximizing sequence for T, that is, $\|Tx_n\|\rightarrow 1$ . We can write

$$ \begin{align*}x_n = z_n+w_n,\end{align*} $$

where $z_n\in M$ and $w_n\in M^\perp $ with

(2.2) $$ \begin{align} 1=\|x_n\|^2=\|z_n\|^2+\|w_n\|^2. \end{align} $$

We also have $\langle P_Tz_n, P_Tw_n\rangle = 0$ . Therefore,

$$ \begin{align*}\|P_Tx_n\|^2=\|P_Tz_n\|^2+\|P_Tw_n\|^2.\end{align*} $$

Since $P_T$ is a positive operator and $z_n\in \overline {\operatorname *{\mathrm {span}}}M_T = \overline {\operatorname *{\mathrm {span}}}M_{P_T}$ , we have

(2.3) $$ \begin{align} P_Tz_n =z_n. \end{align} $$

Together with Equation (2.2) and the fact that $\|Tx\| =\|P_Tx\|$ for all $x\in H$ , we get

(2.4) $$ \begin{align} \lim \|P_Tw_n\|^2 & =1- \lim\|P_Tz_n\|^2\\\nonumber & = 1-\lim\|z_n\|^2\\\nonumber & =\|w_n\|^2. \end{align} $$

Now, if $\lim \|w_n\| \neq 0$ , then we have $\lim \frac { \|P_Tw_n\|^2}{\|w_n\|^2} = 1$ . Since $w_n\in M^\bot $ , we will then have $\|P_T|_{M^\bot }\|=1$ .

Observe that, $M^\bot =M_T^\bot $ . Therefore, we will have $\|P_T|_{M_T^\bot }\|=1$ . But, by the assumption, $\|P_T|_{M^\perp }\|<1$ . Therefore, $\lim \|w_n\| = 0$ . That is, $\lim \|x_n -z_n\| =0$ and $\|z_n\|\rightarrow 1$ .

Take the sequence $\{y_n\}$ to be $y_n =\frac {z_n}{\|z_n\|}$ . Then, by Equation (2.3), $y_n\in M_T$ . Also, since $\|z_n\|\rightarrow 1$ , we get

$$ \begin{align*}\lim\|x_n-y_n\|= \lim\|x_n-z_n\| = 0.\end{align*} $$

This implies $(ii)$ .

Conversely assume that $(ii)$ holds. If possible, assume that $\|P_T|_{{M_T}^\perp }\|=\|T\|.$ Then there exists a sequence $\{w_n\} $ in $S_{M_T^\perp }$ such that $\lim \|Tw_n\|=\lim \|P_Tw_n\| = 1$ . Since ${w_n\in M_T^\perp }$ , we have $d(w_n, M_T)\geq \|w_n\|=1$ for all n, which contradicts $(ii)$ .

From the condition $(ii)$ of Theorems 2.1 and 2.4, it is evident that a shared property of maximizing sequence of both bounded linear functionals on Banach spaces and operators on a Hilbert space characterize the strong subdifferentiability of the norm of respective spaces.

The equivalent characterization (condition $(ii)$ ) given in Theorem 2.4 does not rely on the orthogonality properties specific to Hilbert spaces. Consequently, this property can be verified for more general Banach spaces. So, it is worth considering whether this new reformulation characterizes the strong subdifferentiability of the norm of the space of bounded linear operators on a Banach space which is not necessarily a Hilbert space.

In the remaining part of this section, we will prove that Theorem 2.4 can be extended to $\ell _p$ spaces, $1<p<\infty $ . To prove the sufficient part, we need the following lemma, which is similar to [Reference Ruess and Stegall24, Theorem 1].

Lemma 2.5 Let X and Y be Banach spaces. Let $B\in {B}({X\times Y})$ with $\|B\|=1$ . Then the norm of $B(X\times Y)$ is strongly subdifferentiable at B if the following condition holds. For each $\varepsilon> 0$ there exists $\delta>0$ such that, whenever $(x,y)\in S_X\times S_Y$ satisfies $B(x,y)>1-\delta $ , there exists $(x_0,y_0)\in S_X\times S_Y$ satisfying $B(x_0,y_0)=1$ , $\|x-x_0\|<\varepsilon $ and $\|y-y_0\|<\varepsilon $ .

Proof The proof of this lemma utilizes a similar idea employed in the proof of [Reference Ruess and Stegall24, Theorem 1].

Suppose $B\in {B}({X\times Y})$ with $\|B\|=1$ and the given condition holds. Assume that the norm of $B(X\times Y)$ is not strongly subdifferentiable at B. Then, by the definition of strong subdifferentiability, there exists $\varepsilon>0$ and bounded bilinear functionals $C_n$ on $X\times Y$ with $\|C_n\|< \frac {1}{n}$ such that

(2.5) $$ \begin{align} \left|\left\|B+C_n\right\|-\left\|B\right\|-\operatorname{Re}\phi\left(C_n\right)\right| \geq \varepsilon\left\|C_n\right\| \text { for all } \phi\in J_{{B}(X\times Y)^*}(B). \end{align} $$

Using the assumed condition, choose a number $\delta>0$ which satisfies the following.

If $\operatorname {Re} B(x, y)> 1-\delta $ for some $(x, y)\in S_X\times S_{Y}$ , then there exists $(x_0, y_0)\in S_X\times S_{Y}$ satisfying

  1. (i) $B(x_0, y_0) =1$ ,

  2. (ii) $\|x-x_0\|<\frac {\varepsilon }{4}$ , $\|y-y_0\|<\frac {\varepsilon }{4}.$

We can assume, w. l. o. g, that $\delta <\frac {\varepsilon }{2}$ . From Equation (2.5), we can find $C_n$ such that $\|C_n\|<\frac {\delta }{2+\delta }$ which satisfies the Inequality (2.5). Let $C = C_n$ . Choose an element $\left (x, y\right ) \in B_X \times B_{Y}$ such that

$$ \begin{align*}\left(B+C\right)\left(x, y\right)>\left\|B+C\right\|-\delta\left\|C\right\|.\end{align*} $$

Then we have:

$$ \begin{align*} \left\|B\right\| \geq \operatorname{Re} B\left(x, y\right) & =\left(B+C\right)\left(x, y\right)-\operatorname{Re} C\left(x, y\right) \geq\left(B+C\right)\left(x, y\right)-\left\|C\right\| \\ & \geq\left\|B+C\right\|-(1+\delta)\left\|C\right\|\\ &\geq\|B\|-(2+\delta)\left\|C\right\|\\ &> 1-\delta. \end{align*} $$

Hence, $\operatorname {Re} B\left (x, y\right )>1-\delta $ . By choice of the number $\delta $ , there exists $(x_0, y_0)\in S_X\times S_{Y}$ such that $B(x_0, y_0) =1$ and $\|x-x_0\|<\frac {\varepsilon }{4}$ and $\|y-y_0\|<\frac {\varepsilon }{4}$ .

Recall the isometric isomorphism $B(X\times Y) \cong (X \tilde {\otimes }_\pi Y)^*$ . Let c be the bounded linear functional on $X \tilde {\otimes }_\pi Y$ corresponding to the bilinear functional C. Then

(2.6) $$ \begin{align} \left|C\left(x, y\right)-C\left(x_0, y_0\right)\right| & =\left|c\left(x \otimes y-x_0 \otimes y_0\right)\right| \\\nonumber & \leq\left\|C\right\|\left\|x \otimes y-x_0 \otimes y_0\right\|_\pi\\\nonumber & \leq \left\|C\right\|\left\|(x-x_0)\otimes y + x_0\otimes (y-y_0)\right\|_\pi\\\nonumber & \leq \|C\| \frac{\varepsilon}{2}. \end{align} $$

We also have:

(2.7) $$ \begin{align} \left\|B+C\right\| \geq\operatorname{Re} \left(B+C\right)\left(x_0, y_0\right)=\left\|B\right\|+\operatorname{Re} C\left(x_0, y_0\right), \end{align} $$

From the Inequalities (2.6) and (2.7), we have

$$ \begin{align*}\begin{aligned} 0 & \leq\left\|B+C\right\|-\left\|B\right\|-\operatorname{Re} C\left(x_0, y_0\right) \\ & <\left(B+C\right)\left(x, y\right)+\delta\left\|C\right\|-\left\|B\right\|-\operatorname{Re} C\left(x_{0}, y_0\right) \\ & <\delta\left\|C\right\|+\operatorname{Re} C(x, y)-\operatorname{Re} C(x_0, y_0) \\ &\leq \delta\left\|C\right\|+ \left|C\left(x, y\right)-C\left(x_0, y_0\right)\right| \\ & < \left\|C\right\|(\delta+\frac{\varepsilon}{2})\\ &< \|C\|\varepsilon, \end{aligned} \end{align*} $$

which is a contradiction to the Inequality (2.5), since $(x_0, y_0)\in J_{{B}(X\times Y)^*}(B)$ . Therefore, the norm of $B(X\times Y)$ is strongly subdifferentiable at B.

We will next recall a theorem from [Reference Kim and Lee15].

Theorem 2.6 [Reference Kim and Lee15, Corollary 2.2]

A reflexive Banach space X is uniformly smooth if and only if for every $\varepsilon>0$ there is $0<\eta (\varepsilon )<1$ such that, for all $f \in B_{X^*}$ and all $x \in S_X$ satisfying $|f(x)|>1-\eta (\varepsilon )$ , there exists $f_0 \in S_{X^*}$ satisfying $\left |f_0(x)\right |=1$ and $\left \|f-f_0\right \|<\varepsilon $ .

We will now give a sufficient condition for strong subdifferentiability of operator norm on Banach spaces when the range is uniformly smooth.

Proposition 2.7 Let X and Y be Banach spaces and $T\in S_{{B}(X, Y)}$ . Then the norm of ${B}(X, Y)$ is strongly subdifferentiable at T if the following conditions hold.

  1. (i) $M_T\neq \emptyset $ and for every maximizing sequence $\{x_n\}$ of T, there exists a subsequence $\{x_{n_i}\}$ of $\{x_n\}$ such that $d(x_{n_i}, M_T)\rightarrow 0$ .

  2. (ii) Y is uniformly smooth.

Proof Assume that an operator $T\in S_{{B}(X, Y)}$ and the conditions $(i), (ii)$ hold. Since Y is uniformly smooth, we have the isometric isomorphism between the following Banach spaces.

$$ \begin{align*}{B}(X, Y)\cong{B}(X, Y^{**}) \cong(X\hat{\otimes}_\pi Y^*)^* \cong {B}(X\times Y^*).\end{align*} $$

So it is enough to prove that the norm of ${B}(X\times Y^*)$ is strongly subdifferentiable at T when viewed as a bounded bilinear functional on $(X\times Y^*)$ . Using Lemma 2.5, it is enough to prove the following property (P).

(P): For each $\varepsilon>0$ , there exists $\delta>0$ such that, whenever $(x, y^*)\in S_{X}\times S_{Y^*}$ satisfies $T(x, y^*)>1-\delta $ , there exists $x_0\otimes y_0^*\in J_{B(X\times Y^*)^*}(T)$ satisfying $\|x-x_0\|<{\varepsilon }$ and $\|y-y_0^*\|<{\varepsilon }$ .

To see this, let $\varepsilon> 0$ . Since Y is uniformly smooth, by Theorem 2.6, we have the following property.

(P1): We can find $\delta _1> 0$ such that whenever an element $(y, y^*)\in S_Y\times S_{Y^*}$ satisfies $|y^*(y)|> 1-\delta _1$ , there exists an element $y_0^*\in J_{Y^*}{(y)}$ with $\|y^*-y_0^*\|<\frac {\varepsilon }{2}$ .

From $(i)$ , we also have the following.

(P2): There exists $\delta _2>0$ such that whenever an element $x\in B_X$ satisfies $\|Tx\|>1-\delta _2$ , we have $d(x, M_T)<\frac {\delta _1}{2}$ .

Define

$$ \begin{align*}\delta = \min\left\{\frac{\delta_1}{2},\delta_2,\varepsilon\right\}.\end{align*} $$

Choose an element $(x, y^*)\in S_X\times S_{Y^*}$ such that $T(x, y^*)>1-\delta $ . Then

$$ \begin{align*}\|Tx\|\geq T(x, y^*)> 1-\delta.\end{align*} $$

Then, by property (P2), there exists $x_0\in M_T$ such that

$$ \begin{align*} \|x-x_0\|<\frac{\delta_1}{2}. \end{align*} $$

Therefore,

$$ \begin{align*} |y^*(Tx_0)|= |y^*(Tx_0-Tx+Tx)|> \operatorname{Re}y^*(Tx_0-Tx+Tx)>\frac{-\delta_1}{2}+1-\delta>1-\delta_1. \end{align*} $$

Since $Tx_0\in S_Y$ , by property (P1), there exists $y_0^*\in J_{Y^*}{(Tx_0)}$ such that $\|y^*-y_0^*\|<{\varepsilon }.$

Thus, we have $(x_0, y_0^*)\in J_{{B}(X\times Y^*)^*}(T)$ which satisfies the requirements of property (P) above. Therefore, the norm of ${B}(X\times Y^*) $ is strongly subdifferentiable at T.

We are now in a position to give the main theorem of this section. We will next obtain a characterization of strong subdifferentiability of the norm of bounded linear operators on $\ell _p$ spaces, analogous to that of operators on Hilbert space. Toward this, we recall a few definitions and results.

Definition 2.4 Let X be a Banach space. For $f\in S_{X^*}$ and $\delta>0$ , the slice of $B_X$ corresponding to f and $\delta $ is defined to be $S(f, \delta , B_X) = \{x\in B_X:\operatorname {Re}f(x)>1-\delta \}$ .

Remark 2.8 It is easy to see that if X is a uniformly convex Banach space, then for an $f\in S_{X^*}$ and $\varepsilon>0$ we have $\operatorname {diam}S(f, \delta (\varepsilon ), B_X)<\varepsilon $ , where $\delta (\varepsilon )$ is the modulus of convexity of X corresponding to $\varepsilon $ .

We will now recall the definition of M-ideals in Banach spaces.

Definition 2.5 A closed subspace J of a Banach space X is called an M-ideal if there exists a closed subspace $J'$ of $X^*$ such that $X^*=J^\bot \oplus _1 J'$ , where $J^\bot = \{f\in X^*: f|_J =0\}$ .

We need the following property of M-ideals in the proof of our main theorem.

Lemma 2.9 [Reference Harmand, Werner and Werner14]

Let J be an M-ideal in a Banach space X. Then

$$ \begin{align*}\operatorname{ext} B_{X^*} = \operatorname{ext} B_{J^\bot}\cup \operatorname{ext} B_{J^*}.\end{align*} $$

We will next recall a theorem that characterizes the extreme points in the duals of operator spaces.

Theorem 2.10 [Reference Ruess and Stegall23]

Let X and Y be Banach spaces. Then we have:

$$ \begin{align*}\operatorname{ext} B_{{K}(X, Y)^*} =\operatorname{ext} B_{X^{**}}\otimes \operatorname{ext} B_{Y^*}.\end{align*} $$

We may replace ${K}(X, Y)$ by any linear subspace containing $X^* \otimes Y$ .

The following is the main theorem of this section, in which we give an affirmative answer, in the case of $\ell _p$ space, to Question 1 proposed in the introduction. This theorem has many interesting applications including the denseness of strongly subdifferentiable operators and a characterization of Fréchet differentiability, both of which are discussed in Section 3.

Theorem 2.11 Let $1< p, q<\infty $ and an operator $T\in {B}(\ell _p, \ell _q)$ with $\|T\|=1$ . Then the following are equivalent.

  1. (i) Norm of ${B}(\ell _p, \ell _q)$ is strongly subdifferentiable at T.

  2. (ii) $M_T\neq \emptyset $ and for every maximizing sequence $\{x_n\}$ of T, there exists a subsequence $\{x_{n_i}\}$ of $\{x_n\}$ such that $d(x_{n_i}, M_T)\rightarrow 0$ .

Proof The implication $(ii)\Rightarrow (i)$ follows from Proposition 2.7.

We will now prove the implication $(i)\Rightarrow (ii)$ . For, it is enough to prove that, for each $\varepsilon>0$ , there exists $\delta>0$ such that

(2.8) $$ \begin{align} d(x, M_T)\leq \varepsilon \text{ whenever } \|T(x)\|>1-\delta \text{ for some } x\in B_{\ell_p}. \end{align} $$

To begin, assume that the norm of ${B}(\ell _p, \ell _q)$ is strongly subdifferentiable at an operator $T\in S_{{B}(\ell _p, \ell _q)}$ . Observe that, we have the following isometric identification,

$$ \begin{align*}{B}(\ell_p, \ell_q) \cong (\ell_p\hat{\otimes}_\pi \ell_q^*)^*.\end{align*} $$

under the map

(2.9) $$ \begin{align} {B}(\ell_p, \ell_q)\ni T \rightarrow \tilde{T}\in (\ell_p\hat{\otimes}_\pi \ell_q^*)^*, \end{align} $$

where the action of $\tilde {T}$ on $\ell _p\hat {\otimes }_\pi \ell _q^*$ is given by

$$ \begin{align*}\tilde{T}\left(\sum_{i=1}^nx_i\otimes y_i^*\right) = \sum_{i=1}^ny_i^*(Tx_i).\end{align*} $$

Therefore, the norm of the dual space $(\ell _p\hat {\otimes }_\pi \ell _q^*)^*$ is strongly subdifferentiable at $\tilde {T}$ .

Fix $\varepsilon>0$ and let $\delta _1>0$ be the modulus convexity of the space $\ell _p$ corresponding to $\varepsilon $ . Apply Remark 2.2 to obtain $\delta>0 $ such that

(2.10) $$ \begin{align} d(v,J_{\ell_p\hat{\otimes}_\pi \ell_q^*}(\tilde{T}))< \delta_1 \text{ whenever } \tilde{T}(v)>1-\delta \text{ for some } v\in B_{\ell_p\hat{\otimes}_\pi \ell_q^*}. \end{align} $$

We will now prove that T satisfies the requirements of (2.8) corresponding to the $\delta>0$ found in (2.10). For, let $x\in S_{\ell _p}$ such that $\|Tx\|>1-\delta $ . Then there exists a bounded linear functional $y^*\in S_{\ell _q^*}$ such that $ y^*(Tx) = \|Tx\|$ . But

$$ \begin{align*}\tilde{T}(x\otimes y^*) = y^*(Tx) = \|Tx\|>1-\delta.\end{align*} $$

Therefore, from the expression (2.10), it follows that there exists an element ${u\in J_{\ell _p\hat {\otimes }_\pi \ell _q^*}(\tilde {T})}$ such that

$$ \begin{align*}\|x\otimes y^*-u\|_\pi<\delta_1.\end{align*} $$

At this stage, we will try to get a representation of u. For, we will look at the set $J_{{B}(\ell _p, \ell _q)^*}(T)$ . It is clear that

$$ \begin{align*}u\in J_{\ell_p\hat{\otimes}_\pi \ell_q^*}(\tilde{T})\subseteq J_{{B}(\ell_p, \ell_q)^*}(T).\end{align*} $$

Observe that $J_{{B}(\ell _p, \ell _q)^*}(T)$ is a $w^*$ -closed extremal subset of the closed unit ball of $ {{B}(\ell _p, \ell _q)^*}$ , where a subset E of a convex set C is called extremal if $x,y\in E$ whenever $x,y\in C$ and $tx+(1-t)y\in E$ , $t\in (0,1)$ .

Hence, $J_{{B}(\ell _p, \ell _q)^*}(T)$ is the $w^*$ -closed convex hull of its extreme points. In particular,

$$ \begin{align*}u\in J_{{B}(\ell_p, \ell_q)^*}(T) = \overline{\operatorname*{\mathrm{conv}}}^{w^*}(\operatorname{ext}J_{{B}(\ell_p, \ell_q)^*}(T)).\end{align*} $$

That is,

$$ \begin{align*}u=w^*-\lim_{k \rightarrow \infty} \sum_n^{m_k} t_n^k \varphi_n^k,\end{align*} $$

where $ \sum _n^{m_k} t_n^k =1$ for each k, and $\varphi _n^k\in \operatorname {ext}J_{{B}(\ell _p, \ell _q)^*}(T)$ for all $n,k$ , consequently, $(\varphi _n^k)(T) =1$ .

Since $J_{{B}(\ell _p, \ell _q)^*}(T)$ is extremal, we have

(2.11) $$ \begin{align} \operatorname{ext}J_{{B}(\ell_p, \ell_q)^*}(T)\subseteq \operatorname{ext}B_{{B}(\ell_p, \ell_q)^*}. \end{align} $$

It is well known that ${K}(\ell _p, \ell _q)$ is an M-ideal in ${B}(\ell _p, \ell _q)$ [Reference Harmand, Werner and Werner14]. Therefore, by Lemma 2.9, we have

$$ \begin{align*}\operatorname{ext}B_{{B}(\ell_p, \ell_q)^*} = \operatorname{ext}B_{{K}(\ell_p, \ell_q)^\perp}\cup \operatorname{ext}B_{{K}(\ell_p, \ell_q)^*}.\end{align*} $$

From the elementary theory of projective tensor products, we know that

$$ \begin{align*}{K}(\ell_p, \ell_q)^* = \ell_p\hat{\otimes}_\pi\ell_q^*.\end{align*} $$

Hence

$$ \begin{align*}\operatorname{ext}B_{{B}(\ell_p, \ell_q)^*} = \operatorname{ext}B_{{K}(\ell_p, \ell_q)^\perp}\cup \operatorname{ext}B_{\ell_p\hat{\otimes}_\pi\ell_q^*}.\end{align*} $$

From Equation (2.11), we conclude that

$$ \begin{align*}\operatorname{ext}J_{{B}(\ell_p, \ell_q)^*}(T)\subset\operatorname{ext}B_{{K}(\ell_p, \ell_q)^\perp}\cup \operatorname{ext}B_{\ell_p\hat{\otimes}_\pi\ell_q^*}.\end{align*} $$

Therefore, either $\varphi _n^k\in \operatorname {ext}B_{{K}(\ell _p, \ell _q)^\perp }$ or $\varphi _n^k\in \operatorname {ext}B_{\ell _p\hat {\otimes }_\pi \ell _q^*}$ for each n and k. Define the set,

$$ \begin{align*}L^k=\{n \in \mathbb{N}: \varphi_n^k\in \operatorname{ext}B_{{K}(\ell_p, \ell_q)^\perp}\}.\end{align*} $$

Define

$$ \begin{align*}z_n^k =\sum_{n\in L^k}t_n^k\varphi_n^k.\end{align*} $$

Then, for each k, the element $z_n^k\in {K}(\ell _p, \ell _q)^\perp $ . Since the net $\{z_n^k\}$ is bounded, passing onto a subnet if necessary, we assume that $\{z_n^k\}$ converges ( $w^*$ -topology) to some x as k goes to $\infty $ . Thus, $x\in {K}(\ell _p, \ell _q)^\perp $ , since ${K}(\ell _p, \ell _q)^\perp $ is $w^*$ -closed.

We also have,

$$ \begin{align*}\lim_{k\rightarrow\infty}z_n^k = u-\lim_{k\rightarrow\infty}\sum_{n\notin L^k}t_n^k\varphi_n^k,\end{align*} $$

where the above limit is taken in the $w^*$ -topology on ${B}(\ell _p, \ell _q)^*$ . But then, since the elements u and $\sum _{n\notin L^k}t_n^k\varphi _n^k$ are in ${\ell _p\hat {\otimes }_\pi \ell _q^*}$ , we must have $x\in {\ell _p\hat {\otimes }_\pi \ell _q^*}$ . Therefore, $x=0$ . That is, we have

(2.12) $$ \begin{align} w^*-\lim_{k\rightarrow\infty}\sum_{n\in L^k}t_n^k\varphi_n^k = 0, \text{ and } w^*-\lim_{k\rightarrow\infty}\sum_{n\notin L^k}t_n^k\varphi_n^k = u. \end{align} $$

For each k, let $t^k = \sum _{n\in L^k}t_n^k$ . Then the net $\{t^k\}$ in $\mathbb {R}$ is bounded, passing onto a subnet if necessary, we assume that $\{t^k\}$ converges to some $t\in \mathbb {R}$ as k goes to $\infty $ .

Suppose, $t\neq 0$ . Then we can find $\varepsilon _1>0$ such that $t^k>\varepsilon _1$ for all k. This would imply,

$$ \begin{align*}\sum_{n\notin L^k}t_n^k=1-t^k<1-\varepsilon_1.\end{align*} $$

But then, since the norm is $w^*$ -lower semi-continuous,

$$ \begin{align*}\|u\|\leq \liminf_{k\rightarrow\infty}\left\|\sum_{n\notin L^k}t_n^k\varphi_n^k\right\|\leq \sum_{n\notin L^k}t_n^k <1-\varepsilon_1<1.\end{align*} $$

Which is not possible, since $\|u\|=1$ . Therefore, $t= 0$ and $L^k \subsetneq \{1,2, \ldots , m_k\}$ for all but finitely many k. Therefore,

$$ \begin{align*}\operatorname{ext}B_{\ell_p\hat{\otimes}_\pi\ell_q^*}\cap J_{{B}(\ell_p, \ell_q)^*}(T)\neq \emptyset.\end{align*} $$

From Theorem 2.10, we know that

$$ \begin{align*}\operatorname{ext}B_{(\ell_p\hat{\otimes}_\pi \ell_q^*)} =\operatorname{ext}B_{{K}(\ell_p, \ell_q)^*} = \{x\otimes y: x\in S_{\ell_p}, y\in S_{\ell_q^*}\}.\end{align*} $$

Now, for each $n\notin L^k$ , we have $\varphi _n^k\in \operatorname {ext}B_{\ell _p\hat {\otimes }_\pi \ell _q^*}$ and hence

$$ \begin{align*}\varphi_n^k= u_n^k \otimes v_n^{*k}, \text{ where } u_n^k\in S_{\ell_p}, v_n^{*k}\in S_{\ell_q^*}.\end{align*} $$

Consequently, by Equation (2.12),

(2.13) $$ \begin{align} u=w^*-\lim_{k \rightarrow \infty} \sum_{n\notin L^k} t_n^k u_n^k \otimes v_n^{*k}. \end{align} $$

Now, fix $\varphi _{n_0}^{k_0}\in \operatorname {ext}B_{\ell _p\hat {\otimes }_\pi \ell _q^*}\cap J_{{B}(\ell _p, \ell _q)^*}(T)\neq \emptyset .$ Since $(\varphi _{n_0}^{k_0})(T) =1$ , we have $v_{n_0}^{*k_0}(T u_{n_0}^{k_0}) =1$ . Hence $\|Tu_{n_0}^{k_0}\|=1$ and thus $M_T\neq \emptyset $ .

Let elements $x_0^*\in S_{{\ell _p}^*}$ and $y_0\in S_{\ell _q}$ be such that $x_0^*(x) =1$ and $y^*(y_0)=1$ , recall from the beginning that $y^*(Tx)=1$ . Then $x_0^*\otimes y_0$ defines a compact operator from $\ell _p$ to $\ell _q$ or in other words $x_0^*\otimes y_0\in \ell _{p}^*\check {\otimes }_\varepsilon \ell _q$ . Therefore, we can find k such that

$$ \begin{align*}\left|x_0^*\otimes y_0\left(u-\sum_{n\notin L^k} t_n^k u_n^k \otimes v_n^{*k} \right)\right|<{\delta_1-\|x\otimes y^*-u\|_\pi}.\end{align*} $$

Thus for the above k, we have

(2.14) $$ \begin{align} \left|x_0^*\otimes y_0\left(x\otimes y^*-\sum_{n\notin L^k} t_n^k u_n^k \otimes v_n^{*k}\right)\right|<{\delta_1}. \end{align} $$

It remains to prove that $d(x, M_T)<\varepsilon $ . If possible, suppose that $d(x, M_T)>\varepsilon $ .

Recall that the elements $x_0^*\in S_{\ell _p^*}$ and $y_0\in S_{\ell _q}$ satisfies $x_0^*(x) =1$ and $y^*(y_0)=1$ . Then, by Remark 2.8, we get

$$ \begin{align*}\operatorname{diam}S(x_0^*, \delta_1, B_{\ell_p})<\varepsilon.\end{align*} $$

Since we are assuming that $d(x, M_T)>\varepsilon $ and $x\in S(x_0^*, \delta _1, B_{\ell _p})$ , we can conclude that

$$ \begin{align*}y\notin S(x_0^*, \delta_1, B_X) \quad\text{ for all } y\in M_T.\end{align*} $$

Therefore, by the definition of the slice.

$$ \begin{align*}\operatorname{Re}x_0^*(y)<1-\delta_1 \quad\text{ for all }y\in M_T.\end{align*} $$

Since $M_T = -M_T$ , we have

(2.15) $$ \begin{align} |\operatorname{Re}x_0^*(y)|<1-\delta_1 \quad\text{ for all } y\in M_T. \end{align} $$

We can assume, w. l. o. g., the elements $u_n^k$ in Equation (2.13) satisfies $x_0^*(u_n^k)\in \mathbb {R}$ . To see this, note that, if $x_0^*(u_n^k) = r_n^ke^{i\theta _n^k}$ , where $0\leq r_n^k\leq 1$ and $0\leq \theta _n^k< 2 \pi $ , then take $\varphi _n^k = e^{-i\theta _n^k}u_n^k\otimes e^{i\theta _n^k}v_n^{*k}$ . Also, note that $\sum _{n\notin L^k} t_n^k\leq 1$ .

So, for the same k above, since the elements $u_n^k$ are in $M_T$ for all $n\notin L^k$ ,

$$ \begin{align*} |\operatorname{Re}x_0^*\otimes y_0(u_n^k \otimes v_n^{*k})| & = |\operatorname{Re}x_0^*(u_n^k)v_n^{*k}(y_0)|\\ & < 1-\delta_1. \end{align*} $$

Therefore,

$$ \begin{align*} \left|\operatorname{Re}x_0^*\otimes y_0\left(x\otimes y^*-\sum_{n\notin L^k} t_n^k u_n^k \otimes v_n^{*k}\right)\right| &=\left|\operatorname{Re}x_0^*(x)y^*(y_0)- \sum_{n\notin L^k} t_n^k \operatorname{Re} x_0^*\otimes y_0(u_n^k \otimes v_n^{*k})\right|\\ &\geq 1-\left| \sum_{n\notin L^k} t_n^k\operatorname{Re}x_0^*(u_n^k)v_n^{*k}(y_0)\right|\\ &\geq 1-\sum_{n\notin L^k}t_n^k\left|\operatorname{Re}x_0^*(u_n^k)v_n^{*k}(y_0)\right|\\ &\geq 1-\sum_{n\notin L^k}t_n^k(1-\delta_1)\\ &\geq\delta_1. \end{align*} $$

Which is a contradiction to the Inequality (2.14). Hence $d(x, M_T)<\varepsilon $ .

Remark 2.12 In the proof of Theorem 2.11, we have utilized the uniform convexity and uniform smoothness of $\ell _p$ spaces. By employing analogous arguments to those used in the proof of Theorem 2.11, it can be shown that if Banach spaces X and Y satisfies the following properties: X is uniformly convex, Y is uniformly smooth, ${K}(X, Y)$ is an M-ideal in ${B}(X, Y)$ , $X^*$ has Radon–Nikodym property, and either $X^*$ or $Y^*$ has approximation property, then the condition $(ii)$ of Theorem 2.11 characterizes the strong subdifferentiability of the norm of ${B}(X, Y)$ .

The following is an example of an operator at which the norm of ${B}(\ell _p)$ is strongly subdifferentiable.

Example 2.13 Let $U\in {B}(\ell _p)$ , $1< p<\infty $ , be an isometry. Then $M_U = S_{\ell _p}$ . Therefore, by Theorem 2.11, U is an $SSD$ point of ${B}(\ell _p)$ .

We will next show that for certain Banach spaces X and Y, condition $(ii)$ of Theorem 2.11 is necessary for the operator norm on ${B}(X, Y)$ to be strongly subdifferentiable at a bounded linear operator $T\in S_{{B}(X, Y)}$ . Consequently, we will address the borderline case $p=\infty $ in Corollary 2.15.

Observe that for all Banach spaces Y, the pair $(Y, \ell _1)$ has the property that every element u of $Y \hat {\otimes }_\pi \ell _1 $ has a representation $u= \sum _{n=1}^\infty \lambda _n u_n\otimes v_n$ , where $\lambda _n \geq 0$ , $\sum _{n=1}^\infty \lambda _n=\|u\|$ , $\|u_n\|=\|v_n\|=1$ [Reference Ryan25].

Motivated by the above property of $\ell _1$ , we define the following for our current purpose.

Definition 2.6 A pair of Banach spaces $(X, Y)$ is said to have property $(N)$ if the following hold. Whenever $f\in (X\hat {\otimes }_\pi Y)^*$ is an $SSD$ point and satisfies $f(u)=\|f\|$ for some $u\in B_{X\hat {\otimes }_\pi Y}$ , then u has the form, $u= \sum _{i=1}^\infty \lambda _i u_i\otimes v_i$ with $\|u_i\|=1$ , $\|v_i\|=1$ , and $\sum _{i=1}^\infty \lambda _i =1$ .

Examples of Banach spaces satisfying the property $(N)$ include $(Y, \ell _1)$ for every Banach space Y, the pair $(H, H)$ where H is a Hilbert space, and the pair $(X, Y)$ for finite-dimensional Banach spaces X and Y. Besides these examples, we do not know a pair of Banach spaces satisfying property $(N)$ . We refer the reader [Reference Dantas, Jung, Roldán and Zoca5], where a stronger notion of property $(N)$ has been studied.

We will next prove a result wherein the condition $(ii)$ of Theorem 2.11 is necessary for the operator norm to be strongly subdifferentiable.

Theorem 2.14 Let X and Y be Banach spaces such that X is uniformly convex and the pair $(X, Y^*)$ has property $(N)$ . If the norm of ${B}(X, Y^{**})$ is strongly subdifferentiable at $T\in S_{{B}(X, Y^{**})}$ , then $M_T\neq \emptyset $ and for every maximizing sequence $\{x_n\}$ of T, there exists a subsequence $\{x_{n_i}\}$ of $\{x_n\}$ such that $d(x_{n_i}, M_T)\rightarrow 0$ .

Proof Suppose the norm of ${B}(X, Y^{**})$ is strongly subdifferentiable at T. We have $(X\hat {\otimes }_\pi Y^*)^* \cong {B}(X, Y^{**})$ . Hence the norm of $(X\hat {\otimes }_\pi Y^*)^*$ is strongly subdifferentiable at $\tilde {T}$ (see Expression (2.9) in Theorem 2.11). Let $\varepsilon>0$ and $\delta _1$ be the modulus convexity of X corresponding to $\varepsilon $ . Using Theorem 2.1, choose $\delta>0 $ such that whenever an element $u\in B_{X\hat {\otimes }_\pi Y^*}$ satisfies $\tilde {T}(u)>1-\delta $ , we have $d(u,J_{X\hat {\otimes }_\pi Y^*}(T))< \delta _1$ .

Let $x\in S_X$ such that $1-\delta <\|Tx\|$ . Then there exists a bounded linear functional $y^*\in S_{Y^*}$ such that $ y^*(Tx) = \|Tx\|$ .

Since the element $x\otimes y^*\in B_{X\hat {\otimes }_\pi Y^*} $ satisfies $\tilde {T}(x\otimes y^*)>1-\delta $ , there exists ${u\in J_{X\hat {\otimes }_\pi Y^*}(T)}$ such that $\|x\otimes y^*-u\|_\pi <\delta _1$ .

By our assumption, the pair has $(X, Y^*)$ has the property $(N)$ . Therefore, we can write

$$ \begin{align*} u =\sum_{i=1}^\infty\lambda_ix_i\otimes y_i^*, \text{ where } \sum_{i=1}^\infty\lambda_i =1, \quad x_i\in S_X,\text{ and } y_i^*\in S_{Y^*}. \end{align*} $$

Since $\tilde {T}(u) = 1$ , we get $\tilde {T}(x_i\otimes y_i^*) = 1$ . Consequently, $\|T(x_i)\|= \|y_i^*\|= 1$ for all i and thus $M_T\neq \emptyset $ . Moreover,

(2.16) $$ \begin{align} \left\|x\otimes y^*-\sum_{i=1}^\infty\lambda_ix_i\otimes y_i^*\right\|_{X\hat{\otimes}_\pi Y^*}<\delta_1. \end{align} $$

We will now prove that $d(x, M_T)<\varepsilon $ . If possible, suppose that $d(x, M_T)>\varepsilon $ . Let the element $x_0^*\in S_{X^*}$ be such that $x_0^*(x) = 1$ . An argument, as in the proof of Theorem 2.11, using the slice of $B_X$ corresponding to the above $\delta>0$ and $x_0^*$ will obtain that

(2.17) $$ \begin{align} \left\|x_0^*(x)y^*-\sum_{i=1}^\infty\lambda_{i}\operatorname{Re}x_0^*(x_i)y_i^*\right\|_{Y^*}> \delta_1. \end{align} $$

Since the projective tensor norm is larger than the injective tensor norm, we get, using Inequality (2.17)

$$ \begin{align*} \left\|x\otimes y^*-\sum_{i=1}^\infty\lambda_ix_i\otimes y_i^*\right\|_{X\hat{\otimes}_\pi Y^*}&\geq \left\|x\otimes y^*-\sum_{i=1}^\infty\lambda_ix_i\otimes y_i^*\right\|_{X\check{\otimes}_\varepsilon Y^*}\\ &=\left\| x\otimes y^*-\sum_{i=1}^\infty\lambda_ix_i\otimes y_i^*\right\|_{{B}(X^*, Y^*)}\\ &\geq \left\|x_0^*(x)y^*-\sum_{i=1}^\infty\lambda_i\operatorname{Re}x_0^*(x_i)y_i^*\right\|_{Y^*} \\ &> \delta_1. \end{align*} $$

Which is a contradiction to Inequality (2.16). Hence $d(x, M_T)<\varepsilon $ .

In the following corollary, we will give a necessary condition for the operator norm to be strongly subdifferentiable in the borderline case $p=\infty $ .

Corollary 2.15 Let X be a uniformly convex Banach space and $T\in {B}({X, \ell _\infty })$ such that $\|T\|=1$ . If the norm of ${B}({X, \ell _\infty })$ is strongly subdifferentiable at T, then $M_T\neq \emptyset $ and for every maximizing sequence $\{x_n\}$ of T, there exists a subsequence $\{x_{n_i}\}$ of $\{x_n\}$ such that $d(x_{n_i}, M_T)\rightarrow 0$ .

Proof We have ${B}({X, \ell _\infty }) = {B}(X, c_0^{**})$ . Since the pair $(X, c_0^*)=(X, \ell _1)$ has property $(N)$ , the result follows from Theorem 2.14.

We will next consider the other extreme value $p=1$ . The strong subdifferentiability of the operator norm $\ell _1$ can be deduced from [Reference Franchetti and Payá10, Theorem 2.5] and Theorem 2.1, for completeness, we will recall it here.

Proposition 2.16 Let $T\in {B(\ell _1)}$ be an operator such that $\|T\|=1$ and $\{e_j\}$ denotes the canonical basis of $\ell _1$ . Then T is an $SSD$ point if and only if the following two conditions hold.

  1. (i) The set, $\{j\in \mathbb {N} : \|Te_j\| =\|T\|\}\neq \emptyset $ and $\sup \{\|Te_j\| : \|Te_j\| <1\}<1.$

  2. (ii) $Te_j\in c_{00}$ whenever $\|Te_j\|=1$ .

Proof We have $B(\ell _1) = \left (\bigoplus _{n=1}^\infty \ell _1\right )_{\infty }$ under the isometric identification $T \mapsto (Te_1, Te_2, \ldots )$ with $\|T\| =\sup _j\{\|Te_j\|\}$ .

Therefore, T is an $SSD$ point of $B(\ell _1)$ if and only if $(Te_1, Te_2, \ldots )$ is an $SSD$ point of $\left (\bigoplus _{n=1}^\infty \ell _1\right )_{\infty }$ .

From [Reference Franchetti and Payá10, Theorem 2.5], the element $(Te_1, Te_2, \ldots )$ is an $SSD$ point of $\left (\bigoplus _{n=1}^\infty \ell _1\right )_{\infty }$ if and only if $Te_j$ is an $SSD$ point of $\ell _1$ for all j satisfying $\|Te_j\| = 1$ and the element $(\|Te_j\|)\in \ell _\infty $ is an $SSD$ point of $\ell _\infty $ .

Since $\ell _1 =c_0^*$ , it can be easily verified using Theorem 2.1 that $Te_j$ is an $SSD$ point of $\ell _1$ if and only if $Te_j\in c_{00}$ . Since $\ell _\infty =\left (\oplus _{n=1}^\infty \mathbb {F}\right )_\infty $ , using [Reference Franchetti and Payá10, Theorem 2.5] again, we get that $(\|Te_j\|)\in \ell _\infty $ is an $SSD$ point of $\ell _\infty $ if and only if $\{j\in \mathbb {N} : \|Te_j\| =1\}\neq \emptyset $ and $\sup \{\|Te_j\| : \|Te_j\| <1\}<1.$

3 Denseness of $SSD$ points of ${B}(\ell _p, \ell _q)$

We know that the set of all $SSD$ points in ${B}(H)$ is dense in ${B}(H)$ [Reference Contreras, Payá and Werner4]. In this section, we will extend this result to the operator norm on $\ell _p$ spaces. Additionally, we will show that the set of $SSD$ points of ${B}(\ell _p, \ell _q)$ contains an important class of operators, namely, those operators with essential norm strictly less than its operator norm. Toward the end of this section, we prove that the Fréchet and Gateaux differentiability coincide in ${B}(\ell _p, \ell _q)$ . We will also show that the assumption made by F. Kittaneh and R. Younis in [Reference Kittaneh and Younis16] regarding the smoothness of operator norm characterizes Fréchet differentiability in ${B}(\ell _p, \ell _q)$ .

For $1<p,q<\infty $ , let $\operatorname {SSD}\{{B}(\ell _p, \ell _q)\}$ denotes the set of all operators in ${B}(\ell _p, \ell _q)$ at which the norm of ${B}(\ell _p, \ell _q)$ is strongly subdifferentiable.

We will next provide a theorem which has very interesting consequences.

Theorem 3.1 The following inclusion holds for $1< p, q<\infty $ .

(3.1) $$ \begin{align} \left\{T \in {B}(\ell_p, \ell_q): \|T\|_e<\|T\|\right\}\subset \operatorname{SSD}\{{B}(\ell_p, \ell_q)\}. \end{align} $$

In particular, $\operatorname {SSD}\{{B}(\ell _p, \ell _q)\}$ is a dense subset of ${B}(\ell _p, \ell _q)$ .

To establish Theorem 3.1, we have to integrate Theorem 2.11 with a few findings from [Reference Harmand, Werner and Werner14] and Lemma 3.4. We will first recall the required results.

Lemma 3.2 [Reference Harmand, Werner and Werner14, Proposition 4.8, Proof of (b)]

Let $1<p,q<\infty $ . Then the set ${K}^\theta (\ell _p, \ell _q)=\left \{T \in {B}(\ell _p, \ell _q): \| T\|=\| T \|_e\right \}$ is nowhere dense in ${B}(\ell _p, \ell _q)$ .

We will next recall a formula for computing the essential norm of an operator.

Lemma 3.3 [Reference Werner27, Lemma 5]

Suppose that ${K}(X, Y)$ is an M-ideal in ${B}(X, Y)$ and an operator $T\in {B}(X, Y)$ . Then

$$ \begin{align*}\|T\|_e=\max \left\{w(T), w^*(T)\right\}, \end{align*} $$

where

$$ \begin{align*}w(T):=\sup \left\{\limsup _\alpha\left\|T x_\alpha\right\| :\left\|x_\alpha\right\|=1, x_\alpha \xrightarrow{w} 0\right\} \end{align*} $$

and, in the same way,

$$ \begin{align*}w^*(T):=\sup \left\{\limsup _\alpha\left\|T^* x_\alpha\right\|: \| x_\alpha \|=1, x_\alpha \xrightarrow{w^*} 0\right\} \end{align*} $$

and at least one of the involved suprema is achieved.

We will now recall the definition of $(M_p)$ -space.

Definition 3.1 [Reference Oja and Werner19]

Let $1 \leq p <\infty $ . We say that a Banach space X has property $\left (M_p\right )$ or is an $\left (M_p\right )$ -space if ${K}\left (X \oplus _p X\right )$ is an M-ideal of ${B}\left (X \oplus _p X\right )$ .

We refer the reader [Reference Harmand, Werner and Werner14, Reference Oja and Werner19] for more details on $(M_p)$ -spaces. Note that the space $\ell _p$ belongs to the class of $(M_p)$ -space for $1 < p < \infty $ . From [Reference Oja and Werner19], we know that if X is an $\left (M_p\right )$ -space, then X is reflexive and ${K}(X)$ is an M-ideal in ${B}(X)$ . Moreover, from [Reference Harmand, Werner and Werner14, chapter VI, Theorem 6.6] we get, if X is an $(M_p)$ -space, then for any weakly null sequence $\{u_n\}$ in X, the following equality holds

(3.2) $$ \begin{align} \lim\|u+u_n\|^p = \|u\|^p+\lim\|u_n\|^p \quad (u\in X). \end{align} $$

The following lemma is an extension of [Reference Pellegrino and Teixeira20, Theorem 1, 2] to the broader class of Banach spaces known as $(M_p)$ -spaces, under the additional assumption that $\|T\|_e<\|T\|$ .

Lemma 3.4 Let X and Y be separable Banach spaces which belongs to the classes $(M_p)$ and $(M_q)$ , respectively, $1<p,q<\infty $ . Suppose $T\in {B}(X, Y)$ such that $\|T\|_e<\|T\|$ and the sequence $\{x_n\}$ in $B_X$ is a maximizing sequence for T. If a subsequence $\{x_{n_i}\}$ of $\{x_n\}$ converges weakly to $x\neq 0$ , then $\|Tx\|=\|T\|$ and $\|x\|=1$ .

Proof For $q< p$ , we have ${K}(X, Y) = {B}(X, Y)$ [Reference Harmand, Werner and Werner14, Chapter VI, Corollary 5.14]. Therefore, the results hold since compact operators are weak-to-norm sequentially continuous.

Therefore, we can assume that $1<p\leq q<\infty $ . Let $T\in {B}(X, Y)$ , we can take $\|T\|=1$ . Let the sequence $\{x_n\}$ be a maximizing sequence for T and $\{x_n\}$ converges weakly to $x\neq 0$ (we are denoting subsequence of $\{x_n\}$ by $\{x_n\}$ itself). Since X is $(M_p)$ and Y is $(M_q)$ , it follows from Equation (3.2) that

(3.3) $$ \begin{align} 1-\|x\|^p =\lim\|x_n-x\|^p \end{align} $$

and

(3.4) $$ \begin{align} 1=\lim\|Tx_n\|^q =\|Tx\|^q+\lim\|Tx_n-Tx\|^q. \end{align} $$

Note that $\|Tx\|\leq 1$ and $\lim \|Tx_n-Tx\|\leq 1$ , consequently, since $q\geq p\geq 1$ ,

$$ \begin{align*} 1=\lim\left\|Tx_n\right\|^{q} &=\|Tx\|^q+\lim\|Tx_n-Tx\|^q\\ &\leq \|Tx\|^p+\lim\|Tx_n-Tx\|^p\\ &\leq \|Tx\|^p+\lim\|x_n-x\|^p\\ &= \|Tx\|^p+ 1-\|x\|^p. \end{align*} $$

Therefore,

(3.5) $$ \begin{align} \|Tx\|= \|x\|. \end{align} $$

Since $\|x\|\leq 1$ , using Equations (3.3) and (3.4) together with the fact that $q\geq p> 1$ , we get

(3.6) $$ \begin{align} 1-\|x\|^q\geq 1-\|x\|^p. \end{align} $$

and

(3.7) $$ \begin{align} \lim \|x_n-x\|^q = \lim (\|x_n-x\|^p)^{q/p} = (1-\|x\|^p)^{q/p}\leq 1-\|x\|^p. \end{align} $$

If possible, assume that the sequence $\{\|x_n-x\|\}$ does not converge to $0$ . Then there exists $\varepsilon>0$ and a subsequence $\{x_{n_k}\}$ of $\{x_n\}$ such that $\|x_{n_k}-x\|>\varepsilon $ for all k. Then, by Equations (3.6) and (3.7),

$$ \begin{align*} \lim \left\|T\left(\frac{x_{n_k}-x}{\|x_{n_k}-x\|}\right)\right\|^q & =\frac{1-\|Tx\|^q}{\lim\|x_{n_k}-x\|^q}\\ &=\frac{1-\|x\|^q}{\lim\|x_{n_k}-x\|^q}\\ &\geq\frac{1-\|x\|^p}{1-\|x\|^p}\\ &=1. \end{align*} $$

Since the sequence $\left \{\frac {x_{n_k}-x}{\|x_{n_k}-x\|}\right \}$ converges weakly to $0$ , by Lemma 3.3,

$$ \begin{align*} \|T\|_e\geq \lim \left\|T\left(\frac{x_{n_k}-x}{\|x_{n_k}-x\|}\right)\right\| \geq 1 =\|T\|. \end{align*} $$

Which is not possible by our assumption. Therefore, the sequence $(\|x_n-x\|)$ converges to $0$ . Hence, by Equations (3.5) and (3.3), $\|Tx\| =\|x\|=1 =\|T\|$ .

We will next give a proof of Theorem 3.1.

Proof Proof of Theorem 3.1

Let $T \in {B}(\ell _p, \ell _q) $ such that $\|T\|_e<\|T\|$ . To see that the operator $T\in \operatorname {SSD}\{{B}(\ell _p, \ell _q)\}$ , it is enough to verify, by Theorem 2.11, $\liminf d(x_n, M_T) = 0$ for every maximizing sequence $\{x_n\}$ of T.

Let the sequence $\{x_n\}$ be a maximizing sequence for T, then there exists a subsequence of $\{x_n\}$ (we will denote the subsequence by $\{x_n\}$ itself) converges weakly to some $x\in \ell _p$ by the reflexivity of $\ell _p$ .

If $x=0$ , then by Lemma 3.3, $\|T\|=\|T\|_e$ , which is not possible.

Therefore, $x\neq 0$ , by Lemma 3.4, $\|Tx\| =\|T\|$ , and $\|x\|=1$ . Thus $x\in M_T$ .

Since the sequence $\{x_n\}$ is a maximizing sequence for T, from the proof of Lemma 3.4, the sequence $\{\|x_n-x\|\}$ converges to $0$ . Thus $d(x_n, M_T)\rightarrow 0$ . Hence, by Theorem 2.11, the operator $T\in \operatorname {SSD}\{{B}(\ell _p, \ell _q)\}$ .

The denseness of the set $\operatorname {SSD}\{{B}(\ell _p, \ell _q)\}$ is now follows from Lemma 3.2.

We know that, for a Hilbert space H, the norm of ${B}(H)$ is strongly subdifferentiable at every compact operator on H [Reference Contreras3]. We will next obtain a corollary which extends this result to $\ell _p$ spaces.

Corollary 3.5 Suppose $1<p,q<\infty $ and $K\in {B}(\ell _p, \ell _q)$ be a compact operator. Then the norm of ${B}(\ell _p, \ell _q)$ is strongly subdifferentiable at K.

Proof The proof follows directly from the inclusion, $\left \{T \in {B}(\ell _p, \ell _q): \|T\|_e<\|T\|\right \}\subseteq \operatorname {SSD}\{{B}(\ell _p, \ell _q)\}.$

Observe from Theorem 2.11 that if an operator $T\in \operatorname {SSD}\{{B}(\ell _p, \ell _q)\}$ , then T attains its norm, the converse is not true as the following example suggest.

Example 3.6 Let $1<p, q<\infty $ and $\{e_n\}$ denotes the canonical basis of $\ell _p$ . Let ${T\in {B}(\ell _p, \ell _q)}$ be an operator defined as follows, for each $x=(\zeta _1, \zeta _2, \ldots )\in \ell _p$

$$ \begin{align*}T\left(\sum_{i=1}^\infty \zeta_ne_n\right)= \zeta_1e_1+\sum_{i=2}^\infty \left(1-\frac{1}{n}\right)\zeta_ne_n.\end{align*} $$

Clearly, $M_T=\{\alpha e_1 : \alpha \in \mathbb {F}, |\alpha |=1\}$ . Let the sequence $\{x_n\}$ be a maximizing sequence for T. Then, for some subsequence $\{x_{n_i}\}$ , $d(x_{n_i}, M_T)\rightarrow 0$ if and only if $x_{n_i}\rightarrow e_1$ . But the sequence $\{e_n\}$ is a maximizing sequence for T and does not contain any subsequence converging to $e_1$ . Hence $(ii)$ of Theorem 2.11 fails to hold. Therefore, $T\notin \operatorname {SSD}\{{B}(\ell _p, \ell _q)\}$ .

Remark 3.7 It is well known that the set of all norm-attaining bounded linear operators on a reflexive Banach space X is dense in $B(X)$ [Reference Lindenstrauss17]. From Example 3.6 and Theorem 2.11 we get $\operatorname {SSD}\{{B}(\ell _p)\}$ is a proper subset of norm-attaining bounded linear operators on $\ell _p$ . Moreover, from Theorem 3.1, $\operatorname {SSD}\{{B}(\ell _p)\}$ is dense in ${B}(\ell _p)$ .

As an application of Theorem 3.1, we will next obtain a characterization of Fréchet differentiability. Consequently, we will show the equivalence between Fréchet differentiability and Gateaux differentiability of the norm of ${B}(\ell _p, \ell _q)$ (see Theorem 3.9), just as observed in the case of operator norm on Hilbert spaces (see [Reference Abatzoglou1, Theorem 3.1] and [Reference Taylor and Werner26, Theorem]). We begin by recalling a few definitions and results.

Definition 3.2 The norm of a Banach space X is Fréchet differentiable at a point u in the unit sphere $S_X$ if and only if there is a bounded linear functional $f \in X^*$ (unique) such that

(3.8) $$ \begin{align} \lim_{t \rightarrow 0} \frac{\|u+t x\|-1}{t}=\operatorname{Re} f(x) \end{align} $$

uniformly on $ B_X$ .

If we drop the uniformity assumption for x in the above definition, we have the definition of Gateaux differentiability of the norm at u. That is, Gateaux differentiability demands only the existence of the limit in Equation (3.8) at each point $x\in B_X$ . When $x= u$ , we see that the unique bounded linear functional f in Definition 3.2 satisfies $\|f\|=f(u)=1$ . Indeed, the functional f, which we call the gradient of the norm at u, is then uniquely determined by the condition

$$ \begin{align*}\|f\|=f(u)=1. \end{align*} $$

If the norm of a Banach space X is Gateaux differentiable at a point $u\in S_X$ , then it is usually said that u is a smooth point of $S_X$ .

We will next recall a theorem from [Reference Deeb and Khalil7], which combined with Theorem 3.1 of the present paper will give a characterization for Fréchet differentiability in ${B}\left (\ell _p,\ell _q\right )$ in terms of the essential norm of an operator.

Theorem 3.8 [Reference Deeb and Khalil7, Theorem 2.1] and [Reference Werner27, Theorem 1]

Let $T \in {B}\left (\ell _p,\ell _q\right ), 1<p,q<\infty $ , and $\|T\|=1$ . Then the following are equivalent:

  1. (i) T is smooth.

  2. (ii) $\|T\|_e <1$ and if $\left \|T x_1\right \|=\left \|T x_2\right \|=1$ for $x_1, x_2 \in B_{\ell _p}$ then $x_1= \alpha x _2$ , where $\alpha \in \mathbb {F}$ and $|\alpha | =1$ .

In the following theorem, we will see that the condition $(ii)$ of Theorem 3.8 characterizes Fréchet differentiability rather than smoothness. Moreover, the following theorem is an extension of [Reference Abatzoglou1, Theorem 3.1] and [Reference Taylor and Werner26, Theorem] to the class of $\ell _p$ spaces. At this stage, the proof of the following corollary looks simple, but Theorem 3.1 and consequently Theorem 2.11 are very crucial in the proof.

Proposition 3.9 Let $T \in {B}\left (\ell _p,\ell _q\right ), 1<p<\infty $ , and $\|T\|=1$ . Then the following are equivalent:

  1. (i) The norm of ${B}\left (\ell _p\right )$ is Fréchet differentiable at T.

  2. (ii) $\|T\|_e <1$ and if $\left \|T x_1\right \|=\left \|T x_2\right \|=1$ for $x_1, x_2 \in B_{\ell _p}$ then $x_1= \alpha x _2$ , where $\alpha \in \mathbb {F}$ and $|\alpha | =1$ .

Consequently, the norm of ${B}(\ell _p, \ell _q)$ is Fréchet differentiable at T if and only if norm of ${B}(\ell _p, \ell _q)$ is Gateaux differentiable at T.

Proof Fréchet differentiability implies Gateaux differentiability (smoothness). Hence, by Theorem 3.8, $(i)\Rightarrow (ii)$ .

Conversely, if $(ii)$ holds, then by Theorem 3.8, T is smooth. Also, by Theorem 3.1, T is an $SSD$ point of ${B}(\ell _p, \ell _q)$ . Since Gateaux differentiability together with strong subdifferentiability implies Fréchet differentiability, the norm of ${B}(\ell _p, \ell _q)$ is Fréchet differentiable at T.

Acknowledgments

A part of this research was completed during the author’s tenure as a Visiting Scientist at ISI Bangalore and ISI Kolkata. The author extends his heartfelt gratitude to ISI Bangalore and ISI Kolkata, especially to Prof. Jaydeb Sarkar of ISI Bangalore and Prof. Debashish Goswami of ISI Kolkata, for their generous hospitality and unwavering support during the period.

Footnotes

A portion of this research has been supported through the JC Bose National Fellowship and grant of Prof. Debashish Goswami, sponsored by the Science and Engineering Research Board (SERB) under the Government of India.

References

Abatzoglou, T. J., Norm derivatives on spaces of operators . Math. Ann. 239(1979), no. 2, 129135. MR519008Google Scholar
Guerrero, J. B. and Palacios, A. R., Strong subdifferentiability of the norm on ${\mathrm{JB}}^{\ast }$ -triples . Q. J. Math. 54(2003), no. 4, 381390. MR2031171Google Scholar
Contreras, M. D., Strong subdifferentiability in spaces of vector-valued continuous functions . Quart. J. Math. Oxford Ser. (2), 47(1996), no. 186, 147155. MR1397934Google Scholar
Contreras, M. D., Payá, R. and Werner, W., ${C}^{\ast }$ -algebras that are $I$ -rings . J. Math. Anal. Appl. 198(1996), no. 1, 227236. MR1373537Google Scholar
Dantas, S., Jung, M., Roldán, Ó., and Zoca, A. R., Norm-attaining tensors and nuclear operators . Mediterr. J. Math. 19(2022), no. 1, 38, 27. MR4371175Google Scholar
Dantas, S., Kim, S. K., Lee, H. J., and Mazzitelli, M., Strong subdifferentiability and local Bishop-Phelps-Bollobás properties . Rev. R. Acad. Cienc. Exactas Fís. Nat. Ser. A Mat. RACSAM 114(2020), no. 2, 47, 16. MR4049872Google Scholar
Deeb, W. M. and Khalil, R. R., Exposed and smooth points of some classes of operation in $L({l}^p)$ . J. Funct. Anal, 103(1992), no. 2, 217228. MR1151547Google Scholar
Deville, R., Godefroy, G., and Zizler, V., Smoothness and renormings in Banach spaces. Pitman Monographs and Surveys in Pure and Applied Mathematics, vol. 64, Longman Scientific & Technical, Harlow; copublished in the United States with John Wiley & Sons, Inc., New York, 1993. MR1211634Google Scholar
Diestel, J., Geometry of Banach spaces—selected topics, Springer-Verlag, Berlin, 1975. MR461094Google Scholar
Franchetti, C. and Payá, R., Banach spaces with strongly subdifferentiable norm , Boll. Un. Mat. Ital. B (7), 7(1993), no. 1, 4570. MR1216708Google Scholar
Godefroy, G., Indumathi, V., and Lust-Piquard, F., Strong subdifferentiability of convex functionals and proximinality , J. Approx. Theory 116(2002), no. 2, 397415. MR1911087Google Scholar
Godefroy, G., Montesinos, V., and Zizler, V., Strong subdifferentiability of norms and geometry of Banach spaces . Comment. Math. Univ. Carolin. 36(1995), no. 3, 493502. MR1364490Google Scholar
Guerrero, J. B. and Peralta, A. M., Subdifferentiability of the norm and the Banach-Stone theorem for real and complex ${JB}^{\ast }$ -triples . Manuscripta Math. 114(2004), no. 4, 503516. MR2081949Google Scholar
Harmand, P., Werner, D., and Werner, W., $M$ -ideals in Banach spaces and Banach algebras, Lecture Notes in Mathematics, vol. 1547, Springer-Verlag, Berlin, 1993. MR1238713Google Scholar
Kim, S. Kwang and Lee, H. J., Uniform convexity and Bishop-Phelps-Bollobás property . Canad. J. Math. 66(2014), no. 2, 373386. MR3176146Google Scholar
Kittaneh, F. and Younis, R., Smooth points of certain operator spaces . Integral Equ. Oper. Theory 13(1990), no. 6, 849855. MR1073855Google Scholar
Lindenstrauss, J., On operators which attain their norm . Israel J. Math. 1(1963), 139148. MR160094Google Scholar
Martínez-Moreno, J., Mena-Jurado, J. F., Payá-Albert, R., and Rodríguez-Palacios, Á. An approach to numerical ranges without Banach algebra theory . Illinois J. Math. 29(1985), no. 4, 609626. MR806469Google Scholar
Oja, E. and Werner, D., Remarks on $M$ -ideals of compact operators on $X{\oplus}_pX$ . Math. Nachr. 152(1991), 101111. MR1121227Google Scholar
Pellegrino, D. and Teixeira, E. V., Norm optimization problem for linear operators in classical Banach spaces . Bull. Braz. Math. Soc. (N.S.) 40(2009), no. 3, 417431. MR2540517Google Scholar
Rao, T. S. S. R. K., Subdifferential set of an operator . Monatsh. Math. 199(2022), no. 4, 891898. MR4497173Google Scholar
Rao, T. S. S. R. K., Subdifferentiability and polyhedrality of the norm . Boll. Unione Mat. Ital. 16(2023), no. 4, 741746. MR4653431Google Scholar
Ruess, W. M. and Stegall, C. P., Extreme points in duals of operator spaces . Math. Ann. 261(1982), no. 4, 535546. MR682665Google Scholar
Ruess, W. M. and Stegall, C. P., Exposed and denting points in duals of operator spaces . Israel J. Math. 53(1986), no. 2, 163190. MR845870Google Scholar
Ryan, R. A., Introduction to tensor products of Banach spaces, Springer Monographs in Mathematics, Springer-Verlag, London, 2002. MR1888309Google Scholar
Taylor, K. F. and Werner, W., Differentiability of the norm in von Neumann algebras . Proc. Amer. Math. Soc. 119(1993), no. 2, 475480. MR1149980Google Scholar
Werner, W., Smooth points in some spaces of bounded operators . Integral Equ. Oper. Theory 15(1992), no. 3, 496502. MR1155717Google Scholar