1 Introduction

The purpose of the paper was to explore the orthogonality and the norm attainment of bounded linear operators in the context of semi-Hilbertian structure induced by positive operators on a Hilbert space. Such a study was initiated by Krein in [10] and it remains an active and productive area of research till date. We refer the readers to [2, 3, 8, 18] and the references therein for more information on this. Let us now mention the relevant notations and the terminologies to be used in the article.

We use the symbol \( {\mathbb {H}} \) to denote a Hilbert space. Finite-dimensional Hilbert spaces are also known as Euclidean spaces. Unless mentioned specifically, we work with both real and complex Hilbert spaces. The scalar field is denoted by \( {\mathbb {K}}, \) which can be either \( {\mathbb {R}} \) or \( {\mathbb {C}}. \) The underlying inner product and the corresponding norm on \( {\mathbb {H}} \) are denoted by \( \langle ~,~ \rangle \) and \( \Vert \cdot \Vert , \) respectively. In general, inner products on \( {\mathbb {H}} \) are defined as positive definite, conjugate symmetric forms which are linear in the first argument. It should be noted that apart from the underlying inner product \( \langle ~,~\rangle \) on \( {\mathbb {H}}, \) there may be many other inner products defined on \( {\mathbb {H}}, \) generating different norms. In order to avoid any confusion, whenever we talk of a topological concept on \({\mathbb {H}},\) we explicitly mention the norm that generates the corresponding topology. Let \(B_{{\mathbb {H}}} = \left\{ x \in {\mathbb {H}} : \Vert x\Vert \le 1\right\} \) and \(S_{{\mathbb {H}}} = \left\{ x \in {\mathbb {H}} : \Vert x\Vert = 1\right\} \) be the unit ball and the unit sphere of \({\mathbb {H}},\) respectively. We use the symbol \(\theta \) to denote the zero vector of any Hilbert space other than the scalar fields \({\mathbb {R}}\) and \({\mathbb {C}}.\) For any complex number zRe(z) and Im(z) denote the real part and the complex part of z,  respectively. For any set \(G \subset {\mathbb {H}},\) \({\overline{G}}\) denotes the norm closure of G. Let \({\mathbb {L}}({\mathbb {H}})({\mathbb {K}}({\mathbb {H}}))\) denote the Banach space of all bounded (compact) linear operators on \( {\mathbb {H}} \), endowed with the usual operator norm. Given any \( A \in {\mathbb {L}}({\mathbb {H}}), \) we denote the null space of A by N(A) and the range space of A by R(A). The symbol I is used to denote the identity operator on \({\mathbb {H}}.\) For \(A \in {\mathbb {L}}({\mathbb {H}}),\) \(A^*\) denotes the Hilbert adjoint of A. An operator \(A \in {\mathbb {L}}({\mathbb {H}})\) can be represented as \(A = Re A + i Im A,\) where \(Re A = \frac{1}{2}(A + A^*)\) and \(Im A =\frac{1}{2i}(A - A^*).\) Recall that \(A \in {\mathbb {L}}({\mathbb {H}})\) is said to be a positive operator if \(A = A^*\) and \(\langle Ax,x\rangle \ge 0 \) for all \(x \in {\mathbb {H}}\). A positive operator A is said to be positive definite if \(\langle Ax,x\rangle > 0\) for all \(x \in {\mathbb {H}} \setminus \{\theta \}\). It is well known [2] that any positive operator \(A \in {\mathbb {L}}({\mathbb {H}})\) induces a positive semi-definite sesquilinear form \(\langle ~,~\rangle _A\) on \({\mathbb {H}},\) given by \(\langle x,y\rangle _A = \langle Ax,y\rangle ,\) where \(x , y \in {\mathbb {H}}.\) It is easy to see that \(\langle ~,~\rangle _A\) induces a semi-norm \( \Vert \cdot \Vert _A \) on \( {\mathbb {H}}, \) given by \( \Vert x \Vert _A = \sqrt{\langle Ax,x\rangle }.\) Moreover, when A is positive definite, it can be verified that \( \langle ~,~\rangle _A \) is an inner product on \( {\mathbb {H}} \) and \( \Vert \cdot \Vert _A \) is a norm on \( {\mathbb {H}}. \) In fact, given any \( A \in {\mathbb {L}}({\mathbb {H}}), \) it is natural to ask when the functions \(\langle ~,~\rangle _A\) and \( \Vert \cdot \Vert _A, \) defined as above, are an inner product and a norm on \( {\mathbb {H}}, \) respectively. We explore this question and some related topics in the first part of our main results. We refer the readers to [1, 4, 7, 11] for some more interesting results in this direction.

Given a Hilbert space \(({\mathbb {H}}, \Vert \cdot \Vert )\) and a positive \( A \in {\mathbb {L}}({\mathbb {H}}),\) it is clear that \(ker \Vert \cdot \Vert _A = \{x \in {\mathbb {H}} : \Vert x\Vert _A = 0\}\) is a closed linear subspace of \({\mathbb {H}}\). Then there is a closed linear subspace \(W \subseteq {\mathbb {H}}\) such that \(W\bot ker \Vert \cdot \Vert _A\) and \({\mathbb {H}} = W + ker \Vert \cdot \Vert _A.\) Let P be the linear projection on W such that \(ker P = ker \Vert \cdot \Vert _A.\) Then it follows from [17] that \(\Vert x\Vert _A = \Vert Px\Vert _A.\) In other words, the restriction of \(\Vert \cdot \Vert _A\) to the subspace W is indeed a norm which satisfies the parallelogram property and so \((W, \Vert \cdot \Vert _A)\) is an inner product space. The investigations for the space \({\mathbb {H}}\) equipped with the seminorm \(\Vert \cdot \Vert _A\) are very closely connected to the investigations for the inner product space \((W, \Vert \cdot \Vert _A).\) Furthermore, we consider \(A-\)bounded linear operator \(T : {\mathbb {H}} \longrightarrow {\mathbb {H}}.\) Next, we define linear operator \({\hat{T}} : W \longrightarrow W\) by \( {\hat{T}}(w) := T(w).\) Now, it is very easy to see that we can think of the \(A-\)norm on \({\mathbb {L}}({\mathbb {H}})\) as the classical operator norm in the operator space \({\mathbb {L}}(W)\). Of course, in this case, W is equipped with the norm \(\Vert \cdot \Vert _A : W \longrightarrow [0,\infty ).\) Recently, Zamani [18] investigated the orthogonality relation induced by a positive linear operator on a Hilbert space and obtained some interesting results. In particular, he generalized Theorem 1.1 of [5], also known as the Bhatia-\( \breve{S} \)emrl Theorem, that characterizes the Birkhoff-James orthogonality of matrices on Euclidean spaces. Let us now recall some relevant definitions from [2] and [18].

Definition 1.1

Let \( {\mathbb {H}} \) be a Hilbert space. Let \( A \in {\mathbb {L}}({\mathbb {H}})\) be positive. An element \(x \in {\mathbb {H}}\) is said to be \(A-\)orthogonal to an element \(y \in {\mathbb {H}},\) denoted by \(x \bot _A y,\) if \(\langle x, y\rangle _A = 0.\)

Note that if \(A=I\), then the above definition coincides with the usual notion of orthogonality in Hilbert spaces.

Let \(B_{A^{1/2}}({\mathbb {H}} )= \big \{ T \in {\mathbb {L}}({\mathbb {H}}) :~ \exists ~ c > 0 ~\text {such that}~\Vert Tx\Vert _A \le c\Vert x\Vert _A ~\forall x\in {\mathbb {H}} \big \}\). The \( A- \)norm of \( T \in B_{A^{1/2}}({\mathbb {H}}) \) is given as follows:

$$\begin{aligned} \Vert T\Vert _A = \sup _{x\in {\mathbb {H}}, \Vert x\Vert _A =1} \Vert Tx\Vert _A = \sup \left\{ |\langle Tx,y\rangle _A|: x,y \in {\mathbb {H}}, \Vert x\Vert _A= \Vert y\Vert _A =1\right\} . \end{aligned}$$

An operator \( T \in {\mathbb {L}}({\mathbb {H}}) \) is said to be \(A-\)bounded if \( T \in B_{A^{1/2}}({\mathbb {H}} ).\)

Definition 1.2

\(T \in B_{A^{1/2}}({\mathbb {H}})\) is said to be \(A-\)Birkhoff–James orthogonal to \(S \in B_{A^{1/2}}({\mathbb {H}}),\) denoted by \(T\bot _A^B S, \) if \(\Vert T+\gamma S\Vert _A \ge \Vert T\Vert _A~\text {for all }\gamma \in {\mathbb {C}}.\)

Note that the above definition gives a generalization of the Birkhoff–James orthogonality of bounded linear operators on a Hilbert space. For more information on Birkhoff–James orthogonality in normed linear spaces, we refer the readers to the pioneering articles [6, 9]. Birkhoff–James orthogonality of bounded linear operators and some related applications have been explored in recent times in [5, 12, 13, 15, 16]. We also make use of the following notations:

Given a positive operator \(A \in {\mathbb {L}}({\mathbb {H}}),\) let \(B_{{\mathbb {H}}(A)}\) and \(S_{{\mathbb {H}}(A)}\) denote the \(A-\)unit ball and the \(A-\)unit sphere of \( {\mathbb {H}}, \) respectively, i.e., \(B_{{\mathbb {H}}(A)} = \left\{ x \in {\mathbb {H}} : \Vert x\Vert _A \le 1\right\} \) and \(S_{{\mathbb {H}}(A)} = \left\{ x \in {\mathbb {H}} : \Vert x\Vert _A = 1\right\} \). For any \(T \in B_{A^{1/2}}({\mathbb {H}}),\) the \(A-\)norm attainment set \(M^T_A\) of T was considered in [18]:

$$\begin{aligned} M^T_A = \left\{ x \in {\mathbb {H}} : \Vert x\Vert _A = 1,~\Vert Tx\Vert _A = \Vert T\Vert _A \right\} . \end{aligned}$$

We study the structure of the \(A-\)norm attainment set of an \( A- \)bounded operator \( T \in {\mathbb {L}}({\mathbb {H}}) \) and also explore the corresponding compactness property. As the most important result of the present article, we obtain a complete characterization of the \( A- \)Birkhoff–James orthogonality of compact and \( A- \)bounded operators on \( {\mathbb {H}} \) under an additional condition. This extends the Bhatia–\( \breve{S} \)emrl Theorem to the setting of semi-Hilbertian spaces, induced by a positive operator.

2 Main Results

We begin this section with a characterization of the norm-generating operators on a Hilbert space.

Theorem 2.1

Let \({\mathbb {H}}\) be a Hilbert space and let \(A \in {\mathbb {L}}({\mathbb {H}}).\) Then \(\Vert \cdot \Vert _A\) is a norm on \({\mathbb {H}}\) if and only if \(\langle Ax,x \rangle > 0\) for all \(x \in {\mathbb {H}} \setminus \{ \theta \}. \)

Proof

As the necessary part of the theorem follows trivially, we only prove the sufficient part.

Clearly, \(\Vert x+y\Vert _A ^2 = \Vert x\Vert _A^2 + \Vert y\Vert _A^2 + \langle Ax,y \rangle +\langle Ay,x \rangle .\) This shows that \(\langle Ax,y \rangle +\langle Ay,x \rangle \) is real. It is easy to see that \(Re \langle Ax,y \rangle + Re \langle Ay,x \rangle =\langle (Re A)x,y \rangle + \langle (Re A)y,x \rangle , \) where \(Re A = \frac{1}{2}(A + A^*).\)

Clearly, \(\Vert \cdot \Vert _A \) trivially satisfies all the properties for being a norm, except possibly the triangle inequality. The triangle inequality is satisfied if for all \(x,y \in {\mathbb {H}},\)

$$\begin{aligned} \Vert x+y\Vert _A\le & {} \Vert x\Vert _A + \Vert y\Vert _A \\ i.e., \ if, \langle A(x+y), x+y \rangle\le & {} \langle Ax,x \rangle + \langle Ay,y \rangle + 2 \langle Ax,x \rangle ^{1/2} \langle Ay,y \rangle ^{1/2}\\ i.e., \ if, \langle Ax,y \rangle + \langle Ay,x \rangle\le & {} 2 \langle Ax,x \rangle ^{1/2} \langle Ay,y \rangle ^{1/2}\\ i.e., \ if, Re \langle Ax,y \rangle + Re \langle Ay,x \rangle\le & {} 2 \langle Ax,x \rangle ^{1/2} \langle Ay,y \rangle ^{1/2}\\ i.e., \ if, \langle (Re A)x,y \rangle + \langle (Re A)y,x \rangle\le & {} 2 \langle Ax,x \rangle ^{1/2} \langle Ay,y \rangle ^{1/2}. \end{aligned}$$

Note that for all \(x \in {\mathbb {H}}, \ \langle Re Ax, x \rangle = \frac{1}{2} ( \langle Ax, x \rangle + \langle A^*x,x \rangle \rangle ) = \frac{1}{2} ( \langle Ax, x \rangle +\overline{\langle Ax, x \rangle } ) = \langle Ax, x \rangle \). This proves that ReA is positive definite and so there exists a unique positive operator B on \( {\mathbb {H}} \) such that \(Re A = B^2.\) Now, we have

$$\begin{aligned} |\langle (Re A)x,y \rangle |= & {} |\langle B^2x,y \rangle | = |\langle Bx,By \rangle | = \Vert Bx\Vert \Vert By\Vert \\= & {} \langle B^2x,x \rangle ^{1/2} \langle B^2y,y \rangle ^{1/2}= \langle (Re A)x,x \rangle ^{1/2} \langle (Re A)y,y \rangle ^{1/2}\\= & {} \langle Ax,x \rangle ^{1/2} \langle Ay,y \rangle ^{1/2}. \end{aligned}$$

Similarly, we can show that \( |\langle (Re A)y,x \rangle | \le \langle Ax,x \rangle ^{1/2} \langle Ay,y \rangle ^{1/2}.\) Therefore,

$$\begin{aligned} \langle (Re A)x,y \rangle + \langle (Re A)y,x \rangle \le |\langle (Re A)x,y \rangle + \langle (Re A)y,x \rangle | \le 2 \langle Ax,x \rangle ^{1/2} \langle Ay,y \rangle ^{1/2}. \end{aligned}$$

This completes the proof of the fact that \(\Vert \cdot \Vert _A\) is a norm on \({\mathbb {H}}.\) \(\square \)

As mentioned in the introduction, if A is a positive definite operator on a Hilbert space \({\mathbb {H}},\) then A generates an inner product \( \langle ~ , ~\rangle _A \) on \({\mathbb {H}}\) defined as \( \langle x, y \rangle _A = \langle Ax, y \rangle \) for all \(x, y \in {\mathbb {H}}.\) On the other hand, suppose that \( A \in {\mathbb {L}}({\mathbb {H}}) \) is such that \( \langle x, y \rangle _A \) is an inner product on \({\mathbb {H}}.\) From the conjugate-symmetry of inner product, it follows that A must be self adjoint and from the positive definiteness of inner product, it follows that A must be positive definite. This is mentioned in the following proposition:

Proposition 2.1

Let \({\mathbb {H}}\) be a Hilbert space and let \(A \in {\mathbb {L}}({\mathbb {H}}).\) Then \(\langle ~,~ \rangle _A\) is an inner product on \({\mathbb {H}}\) if and only if A is positive definite.

Remark 2.1

In view of the above theorem, there is a subtle difference in the description of the norm generating operators, depending on whether the underlying Hilbert space is complex or real. This is illustrated in the following two points:

  1. 1.

    If \({\mathbb {H}} \) is a complex Hilbert space then \( \langle ~,~\rangle _A \) and \( \Vert \cdot \Vert _A\) are inner product and norm on \({\mathbb {H}}\), respectively, if and only if A is a positive definite operator on \({\mathbb {H}}\). This is because of the well-known fact that in case of a complex Hilbert space \({\mathbb {H}},\) if \(A \in {\mathbb {L}}({\mathbb {H}})\) is such that \(\langle Ax,x\rangle \ge 0\) for all \(x \in {\mathbb {H}},\) then \(A=A^*.\)

  2. 2.

    If \({\mathbb {H}} \) is real, then there may exist \( A \in {\mathbb {L}}({\mathbb {H}}) \) such that \( A \ne A^{*} \) (and consequently, A is not positive definite) but \( \Vert \cdot \Vert _A\) is a norm on \({\mathbb {H}}. \) As for example, consider the operator A on the Hilbert space \(\ell _2^2({\mathbb {R}})\) defined as \(A(x,y)=(x-y,x+y)\) for all \((x,y) \in {\mathbb {R}}^2.\) Then it is easy to see that \( \langle Ax, x \rangle > 0\) for all \(x \ne \theta \) but \( A \ne A^{*}\). A generates a norm given by \( \Vert x \Vert _A = \langle Ax, x \rangle ^{1/2} \) on \(\ell _2^2({\mathbb {R}})\) but \( \langle x, y \rangle _A = \langle Ax, y \rangle \) is not an inner product on \(\ell _2^2({\mathbb {R}})\). The inner product that induces the norm \(\Vert \cdot \Vert _A \) is given by \(\langle (Re A) x, y \rangle \). In fact, given any \(A \in {\mathbb {L}}({\mathbb {H}})\) with \( \langle Ax, x \rangle > 0\) for all \(x \ne \theta , \) the positive definite operator ReA always generates an inner product \( \langle x, y \rangle _{Re A} = \langle (Re A) x, y \rangle \) which induces the norm \(\Vert \cdot \Vert _A.\)

Our next theorem guarantees that under a suitable condition, given any inner product on an infinite-dimensional separable Hilbert space \({\mathbb {H}}\), there exists a unique positive definite operator that generates the given inner product.

Theorem 2.2

Let \(({\mathbb {H}}, \langle ~,~\rangle )\) be a separable Hilbert space. Let \(\langle ~,~ \rangle _1\) be another inner product on \({\mathbb {H}}.\) Then the following two conditions are equivalent:

  1. (i)

    there exists a positive definite operator A on \({\mathbb {H}}\) such that \(\langle ~,~ \rangle _1 = \langle ~,~ \rangle _A.\)

  2. (ii)

    there exists \(M >0\) such that \(\Vert x\Vert _1 \le M \Vert x\Vert \) for all \(x \in {\mathbb {H}},\) where \(\Vert \cdot \Vert _1\) is the norm induced by the inner product \(\langle ~,~ \rangle _1\) on \( {\mathbb {H}}. \)

Proof

(i\() \Rightarrow (\)ii) :  Clearly, \(\Vert x\Vert _1^2 = \langle x,x \rangle _1 = \langle x,x \rangle _A = \langle Ax,x \rangle \le \Vert A\Vert \Vert x\Vert ^2.\)

(ii\() \Rightarrow (\)i) :  Since \(\Vert x\Vert _1 \le M\Vert x\Vert \) for all \(x \in {\mathbb {H}}\), it follows that \({\mathbb {H}}\) is a separable inner product space with respect to \(\langle ~,~ \rangle _1.\) Let \(({\mathcal {H}}, \langle ~,~\rangle _{{\mathcal {H}}})\) be the completion of \(({\mathbb {H}}, \langle ~,~ \rangle _1).\) Clearly, \( \langle x,y\rangle _{{\mathcal {H}}}= \langle x,y\rangle _1\) for all \(x,y \in {\mathbb {H}}.\) Since \({\mathbb {H}}\) is separable with respect to \(\langle ~,~\rangle _1,\) it is easy to deduce that \({\mathcal {H}}\) is separable with respect to \(\langle ~,~\rangle _{{\mathcal {H}}}.\) Let \(B = \{e_1,e_2,e_3,\ldots \}\) be an orthonormal basis of \({\mathbb {H}}\) with respect to \(\langle ~,~ \rangle \) and let \( B_1 = \{f_1,f_2,f_3,\ldots \}\) be an orthonormal basis of \({\mathcal {H}}\) with respect to \(\langle ~,~\rangle _{{\mathcal {H}}}\). Consider the map \({\widetilde{T}} : ({\mathcal {H}},\langle ~,~\rangle _{{\mathcal {H}}}) \rightarrow ({\mathbb {H}}, \langle ~,~ \rangle )\) defined by \({\widetilde{T}}(\sum _{i=1}^{\infty }a_i f_i) =\sum _{i=1}^{\infty }a_i e_i\) where \(a_i \in {\mathbb {K}}(={\mathbb {R}},{\mathbb {C}})\) for all \(i \in {\mathbb {N}}.\) It can be verified easily that \({\widetilde{T}}\) is well-defined and linear. Let \( T ={\widetilde{T}}\mid _{({\mathbb {H}}, \langle ~,~\rangle _1)}.\) It is easy to see that \(\langle x, y\rangle _1 = \langle Tx, Ty\rangle \) for all \(x,y \in {\mathbb {H}}\). Thus \(\Vert Tx\Vert ^2 = \langle x, x\rangle _1 \le M^2 \Vert x\Vert ^2.\) In particular, T is bounded and, therefore, the adjoint operator \(T^* : ({\mathbb {H}},\langle ~,~\rangle ) \longrightarrow ({\mathbb {H}},\langle ~,~\rangle _1) \) exists. Let \( A = T^*T\). Then it is easy to see that A is a positive definite operator on \(({\mathbb {H}}, \langle ~,~\rangle )\) such that \( \langle x, y \rangle _1 = \langle Ax,y \rangle \) for all \(x,y \in {\mathbb {H}}.\)

The uniqueness of A follows from the fact that if B is any positive definite operator that generates the inner product \( \langle ~,~ \rangle _1 \) then \( \langle Ax, y \rangle = \langle Bx, y \rangle \) for all \(x, y \in {\mathbb {H}} \) and so \(A=B.\) \(\square \)

In light of the above theorem, let us make the following two remarks:

Remark 2.2

In case \( {\mathbb {H}} \) is finite-dimensional, Condition (ii) of the above theorem holds true automatically. Therefore, we obtain a complete description of the set of all inner products defined on an Euclidean space, in terms of positive definite operators on \( {\mathbb {H}}. \) Following the usual matricial representation of linear operators on Euclidean spaces, it seems convenient to say that every positive definite matrix defines an inner product on \( {\mathbb {K}}^{n} \) and conversely.

Remark 2.3

We note that if \(\langle ~, ~\rangle _1\) is an inner product on \({\mathbb {H}}\) such that Condition (ii) of the above theorem is satisfied, it is not necessarily true that \(({\mathbb {H}}, \langle ~,~\rangle _1)\) is complete. Such an example will be constructed explicitly in the proof of Theorem 2.3 (iv).

The unit ball \(B_{{\mathbb {H}}}\) is convex and bounded with respect to \( \Vert \cdot \Vert \). Also, it is compact (in the topology induced by \( \Vert \cdot \Vert \)) if and only \({\mathbb {H}} \) is finite-dimensional. We next study some analogous geometric and topological properties of the \(A-\)unit ball \(B_{{\mathbb {H}}(A)}\) with respect to the norm \( \Vert \cdot \Vert \). We begin with the following proposition, the proof of which is omitted as it follows rather trivially from the convexity of the \(A-\)norm and the continuity of the inner product.

Proposition 2.2

Let \( {\mathbb {H}} \) be a Hilbert space and let \(A \in {\mathbb {L}}({\mathbb {H}})\) be positive. Then \(B_{{\mathbb {H}}(A)}\) is convex and closed with respect to \(\Vert \cdot \Vert \).

We would like to describe the boundedness properties of the \(A-\)unit ball and the \(A-\)unit sphere with respect to the norm \( \Vert \cdot \Vert . \) We require the following proposition which is particularly useful in our study. The proof is omitted, as it can be obtained quite easily.

Proposition 2.3

Let \({\mathbb {H}}\) be a Hilbert space. Let \(A \in {\mathbb {L}}({\mathbb {H}})\) be positive. Then \({\mathbb {H}} = N(A)\oplus \overline{R(A)}.\)

We describe the boundedness properties of the \(A-\)unit ball and the \(A-\)unit sphere in the next theorem.

Theorem 2.3

Let \({\mathbb {H}}\) be a Hilbert space and let \(A \in {\mathbb {L}}({\mathbb {H}})\) be positive. Then the following hold true:

  1. (i)

    If \(N(A) \ne \{\theta \}\), then both \(S_{{\mathbb {H}}(A)}\) and \(B_{{\mathbb {H}}(A)}\) are unbounded with respect to \(\Vert \cdot \Vert \).

  2. (ii)

    If \({\mathbb {H}}\) is finite-dimensional, then \(B_{{\mathbb {H}}(A)}\cap \overline{R(A)}(=B_{{\mathbb {H}}(A)}\cap {R(A)})\) is bounded with respect to \(\Vert \cdot \Vert \).

  3. (iii)

    If H is finite-dimensional, then \(B_{{\mathbb {H}}(A)}\) is bounded with respect to \(\Vert \cdot \Vert \) if and only if \(N(A)=\{\theta \}.\)

  4. (iv)

    Both (ii) and (iii) fail to hold if \({\mathbb {H}}\) is infinite-dimensional.

Proof

We first observe that

$$\begin{aligned}\Vert Ax\Vert ^2 =\langle Ax,Ax\rangle =\langle A^2x,x\rangle =\langle Ax,x\rangle _A \end{aligned}$$

and so it follows that \(\Vert x\Vert _A =0\) if and only if \(x \in N(A)\):

  1. (i)

    Let \(x \in N(A) \) be such that \(x \ne \theta .\) Then \(\Vert x\Vert _A = 0\) and so \(\Vert \lambda x\Vert _A = 0\) for all \(\lambda \in {\mathbb {K}}(={\mathbb {R}},{\mathbb {C}}).\) Next we claim that if \(z \in S_{{\mathbb {H}}(A)},\) then \(z+\lambda x \in S_{{\mathbb {H}}(A)}\) for all \(\lambda \in {\mathbb {K}}(={\mathbb {R}},{\mathbb {C}}).\) Clearly, \(\Vert z+\lambda x\Vert _A \le \Vert z\Vert _A + |\lambda | \Vert x\Vert _A =1.\) Again, \(\Vert z+\lambda x\Vert _A \ge \Vert z\Vert _A - |\lambda | \Vert x\Vert _A =1.\) Thus \(z+\lambda x \in S_{{\mathbb {H}}(A)}\) for all \(\lambda \in {\mathbb {K}}(={\mathbb {R}},{\mathbb {C}}).\) Therefore, \(S_{{\mathbb {H}}(A)}\) is unbounded with respect to \(\Vert \cdot \Vert \) and so \(B_{{\mathbb {H}}(A)}\) is also unbounded with respect to \(\Vert \cdot \Vert \).

  2. (ii)

    Suppose on the contrary that \(B_{{\mathbb {H}}(A)}\cap {R(A)}\) is unbounded with respect to \(\Vert \cdot \Vert \). Then for each \(n\in {\mathbb {N}}\), there exists \(v_n\in B_{{\mathbb {H}}(A)}\cap {R(A)}\) such that \(\Vert v_n\Vert \ge n\). Let \(w_n=\frac{v_n}{\Vert v_n\Vert }\). Then \(\Vert w_n\Vert =1\) and \(\Vert w_n\Vert _A\le \frac{1}{n}\). Clearly, \(\{w_n\}\subseteq S_{\mathbb {H}}\). Since \({\mathbb {H}}\) is finite-dimensional, \(S_{{\mathbb {H}}}\) is compact. Without loss of generality we may assume that \(w_n\longrightarrow w,\) where \(w\in S_{{\mathbb {H}}}\). By Proposition 2.2, \(B_{{\mathbb {H}}(A)}\cap {R(A)}\) is a closed set with respect to \(\Vert \cdot \Vert \) and hence \(w\in B_{{\mathbb {H}}(A)}\cap {R(A)}\). It is easy to check that \(\Vert w_n\Vert _A\longrightarrow \Vert w\Vert _A\). Therefore, \(\Vert w\Vert _A=0\) and so \(w\in N(A).\) This shows that \(w \in N(A) \cap R(A) \) and so \(w=\theta \), a contradiction to our assumption that \(w\in S_{{\mathbb {H}}}\). Therefore, \(B_{{\mathbb {H}}(A)}\cap {R(A)}\) is bounded with respect to \(\Vert \cdot \Vert \).

  3. (iii)

    As \({\mathbb {H}}\) is finite-dimensional, \({\mathbb {H}} = N(A)\oplus R(A).\) Therefore, any \( x \in B_{{\mathbb {H}}(A)}\) can be uniquely written as \(x = u+v,\) where \(u \in B_{{\mathbb {H}}(A)}\cap {N(A)}\) and \(v \in B_{{\mathbb {H}}(A)}\cap R(A).\) From (ii), it follows that \(B_{{\mathbb {H}}(A)}\) is bounded with respect to \(\Vert \cdot \Vert \) if and only if \(B_{{\mathbb {H}}(A)}\cap {N(A)}\) is bounded with respect to \(\Vert \cdot \Vert \). From (i), it follows that \(N(A)=\{\theta \}\) if \(B_{{\mathbb {H}}(A)} \) is bounded with respect to \(\Vert \cdot \Vert \). On the other hand, if \(N(A)=\{\theta \}\) then \( B_{{\mathbb {H}}(A)} = B_{{\mathbb {H}}(A)}\cap R(A) \) is bounded with respect to \(\Vert \cdot \Vert \) by applying (ii).

  4. (iv)

    Consider the Hilbert space \( \ell _2.\) Let \(A \in {\mathbb {L}}(\ell _2)\) be defined by \(A(x_1,x_2,x_3,\ldots ) = (x_1, \frac{x_2}{2},\frac{x_3}{3},\ldots ),\) where \((x_1,x_2,x_3,\ldots ) \in \ell _2.\) It is easy to check that A is positive definite and \(N(A)=\{\theta \}.\) Therefore, \(\ell _2=\overline{R(A)}(\ne R(A))\). Consider the sequence \(\{v_n\} \subseteq \ell _2\) where \(\{v_n\} = \{\sqrt{n}e_n\},\) where \(\{e_n\}\) is the usual orthonormal basis of \(\ell _2.\) Clearly, \(\Vert v_n\Vert _A^2 = \langle Av_n,v_n\rangle = 1\) for each \(n \in {\mathbb {N}}\) but \(\Vert v_n\Vert = \sqrt{n}\) for each \(n \in {\mathbb {N}}\).

\(\square \)

In view of the above theorem, we make the following remark on the geometry of semi-Hilbertian spaces.

Remark 2.4

Let A be a positive operator on a Hilbert space \( {\mathbb {H}}. \) If \(\Vert x\Vert _A =0\) for some \(x \ne \theta ,\) then by (i) of Theorem 2.3, the \(A-\)unit sphere of \( {\mathbb {H}} \) contains a straight line. In other words, the semi-normed space \( ({\mathbb {H}}, \Vert \cdot \Vert _A) \) is not strictly convex whenever A is not positive definite.

There is another nice way to obtain Remark 2.4. Namely, now suppose that A is positive, but not positive definite. Since \(\Vert \cdot \Vert _A\) is a seminorm, it follows from Theorem 3.2 from [17] that any \(x \in B_{{\mathbb {H}}(A)}\) can be uniquely written as \(x = u+v\), where \(u \in B_{W}\) and \(v \in ker \Vert \cdot \Vert _A.\) Note that \(B_{W}\) is the closed unit ball in the inner product space \((W, \Vert \cdot \Vert _A)\), and \(ker \Vert \cdot \Vert _A\) is a linear subspace. Therefore, it is easy to see that \(A-\)unit sphere of \({\mathbb {H}}\) contains a straight line and the seminormed space \(({\mathbb {H}}, \Vert \cdot \Vert _A)\) is not strictly convex whenever A is not positive definite.

In Theorem 2.2 of [14], the authors studied the norm attainment sets of bounded linear operators on a Hilbert space. In particular, it was proved that in an inner product space \({\mathbb {H}}\), for any operator \(T \in {\mathbb {L}}({\mathbb {H}}),\) the norm attainment set \(M_T\) is either the empty set \( \phi , \) or, \( M_T \) is the unit sphere of some subspace of \({\mathbb {H}}.\) Our next result generalizes this, in case of \(A-\)bounded operators.

Theorem 2.4

Let \({\mathbb {H}}\) be a Hilbert space. Let \(A \in {\mathbb {L}}({\mathbb {H}})\) be positive and let \(T \in B_{A^{1/2}}({\mathbb {H}}).\) Then either \(M_A^T = \phi \) or \(M_A^T \cap \overline{R(A)}\) is the \(A-\)unit sphere of some subspace of \({\mathbb {H}}.\)

Proof

If \(M_A^T = \phi ,\) then we have nothing to prove. Let us assume that \(M_A^T \ne \phi .\) Let \(x \in M_A^T.\) As \({\mathbb {H}} = N(A)\oplus \overline{R(A)},\) x can be uniquely written as \(x=u+v,\) where \(u \in N(A)\) and \(v \in \overline{R(A)}.\) Hence \( \Vert u \Vert _A = 0 \) and \( \Vert x \Vert _A = \Vert v \Vert _A. \) As \(T \in B_{A^{1/2}}({\mathbb {H}}),\) it follows that \( \Vert Tu\Vert _A = 0 \) and, therefore, \(\Vert Tx\Vert _A = \Vert Tv\Vert _A = \Vert T\Vert _A.\) This proves that \(M_A^T \cap \overline{R(A)} \ne \phi .\)

To prove that \(M_A^T \cap \overline{R(A)}\) is the \(A-\) unit sphere of some subspace of \({\mathbb {H}},\) it is enough to show that \(\frac{\lambda _1 e_1 \pm \lambda _2 e_2}{\Vert \lambda _1 e_1 \pm \lambda _2 e_2\Vert _A} \in M_A^T \cap \overline{R(A)}\), whenever \(e_1,~e_2 \in M_A^T \cap \overline{R(A)}\) and \(\lambda _1,~\lambda _2 \in {\mathbb {K}} (={\mathbb {R}},~{\mathbb {C}}).\) Let \(e_1,~ e_2 \in M_A^T \cap \overline{R(A)},\) then \(\Vert Te_1\Vert _A = \Vert Te_2\Vert _A = \Vert T\Vert _A\) and \(\Vert e_1\Vert _A = \Vert e_2\Vert _A = 1.\) First we claim that \(\Vert \cdot \Vert _A\) satisfies the parallelogram law for all \(x,~y \in {\mathbb {H}}.\) Let \(x,y \in {\mathbb {H}}.\) Then we have

$$\begin{aligned} \Vert x+y\Vert _A^2 +\Vert x-y\Vert _A^2= & {} \langle x+y,x+y\rangle _A + \langle x-y,x-y\rangle _A \\= & {} \langle A(x+y),x+y\rangle + \langle A(x-y),x-y\rangle \\= & {} 2(\langle Ax,x\rangle + \langle Ay,y\rangle ) \\= & {} 2(\Vert x\Vert _A^2 +\Vert y\Vert _A^2). \end{aligned}$$

This proves our claim. Therefore,

$$\begin{aligned} 2(|\lambda _1|^2 +|\lambda _2|^2)\Vert T\Vert _A^2= & {} 2(\Vert \lambda _1 Te_1\Vert _A^2 + \Vert \lambda _2 Te_2\Vert _A^2) \\= & {} \Vert \lambda _1Te_1 + \lambda _2 Te_2\Vert _A^2 + \Vert \lambda _1 Te_1 - \lambda _2 Te_2\Vert _A^2 \\= & {} \Vert T(\lambda _1 e_1 + \lambda _2 e_2)\Vert _A^2 + \Vert T(\lambda _1 e_1 - \lambda _2 e_2)\Vert _A^2\\\le & {} \Vert T\Vert _A^2(\Vert \lambda _1 e_1 + \lambda _2 e_2\Vert _A^2 + \Vert \lambda _1 e_1 - \lambda _2 e_2\Vert _A^2)\\= & {} 2(|\lambda _1|^2 +|\lambda _2|^2)\Vert T\Vert _A^2.\\ \end{aligned}$$

Hence the above inequality is actually an equality. Since \(\Vert T(\lambda _1 e_1\pm \lambda _2 e_2)\Vert _A \le \Vert T\Vert _A\Vert \lambda _1 e_1 \pm \lambda _2 e_2\Vert _A, \) it follows that

$$\begin{aligned} \Vert T(\lambda _1 e_1\pm \lambda _2 e_2)\Vert _A = \Vert T\Vert _A\Vert \lambda _1 e_1 \pm \lambda _2 e_2\Vert _A. \end{aligned}$$

This establishes the theorem. \(\square \)

In the next theorem we study the compactness property of \(M_A^T \cap \overline{R(A)}.\)

Theorem 2.5

Let \({\mathbb {H}} \) be a Hilbert space and let \(A \in {\mathbb {L}}({\mathbb {H}})\) be a positive operator such that \(B_{{\mathbb {H}}(A)} \cap \overline{R(A)}\) is bounded with respect to \(\Vert \cdot \Vert \). Let \(T \in {\mathbb {K}}({\mathbb {H}}) \cap B_{A^{1/2}}({\mathbb {H}}).\) Then \(M_A^T \cap \overline{R(A)}\) is compact with respect to \(\Vert \cdot \Vert \).

Proof

Clearly, \(\overline{R(A)}\) is a Hilbert space with respect to \(\Vert \cdot \Vert .\) It is easy to see that A is positive definite on \(\overline{R(A)}.\) Therefore, \(\Vert \cdot \Vert _A\) is a norm on \(\overline{R(A)}.\) We claim that \(\Vert \cdot \Vert _A\) and \(\Vert \cdot \Vert \) are equivalent norms on \(\overline{R(A)}.\) Clearly, \(\frac{1}{\sqrt{\Vert A\Vert }}\Vert x\Vert _A \le \Vert x\Vert \) for all \(x\in {\mathbb {H}}\). Let \(x\in \overline{R(A)}\). Then \(\frac{x}{\Vert x\Vert _A}\in B_{{\mathbb {H}}(A)} \cap \overline{R(A)}\). Since \(B_{{\mathbb {H}}(A)} \cap \overline{R(A)}\) is bounded, there exists \(M >0 \) such that \(\Vert z\Vert \le M\) for all \(z \in B_{{\mathbb {H}}(A)} \cap \overline{R(A)}\). Therefore, \(\frac{\Vert x\Vert }{\Vert x\Vert _A} \le M\). Thus \(\frac{1}{\sqrt{\Vert A\Vert }}\Vert x\Vert _A \le \Vert x\Vert \le M\Vert x\Vert _A\) for all \(x\in \overline{R(A)}\). Thus our claim is established. Therefore, \(\overline{R(A)}\) is a Hilbert space with respect to \(\Vert \cdot \Vert _A.\) Next, let \(\{v_n\}\) be a sequence in \(M_A^T\cap \overline{R(A)}\). We show that \(\{v_n\}\) has a convergent subsequence in \(M_A^T\cap \overline{R(A)}\) with respect to \(\Vert \cdot \Vert \). Since \({\mathbb {H}}\) is reflexive and \(B_{{\mathbb {H}}(A)} \cap \overline{R(A)}\) is closed, convex and bounded with respect to \(\Vert \cdot \Vert \), it follows that \(B_{{\mathbb {H}}(A)} \cap \overline{R(A)}\) is weakly compact with respect to \(\Vert \cdot \Vert \). Thus the sequence \(\{v_n\}\) has a weakly convergent subsequence \( \{ v_{n_k}\}\) with respect to \(\Vert \cdot \Vert .\) Suppose \(v_{n_k} \rightharpoonup v\) for some \(v \in B_{{\mathbb {H}}(A)} \cap \overline{R(A)}\) with respect to \(\Vert \cdot \Vert \). Since \(T \in {\mathbb {K}}({\mathbb {H}}),\) it follows that \(T v_{n_k} \longrightarrow Tv\) with respect to \(\Vert \cdot \Vert \). It is easy to see that

$$\begin{aligned} \Vert T\Vert _A^2 = \lim _{k \rightarrow \infty }\Vert T v_{n_k}\Vert _A^2 =\lim _{k \rightarrow \infty }\langle AT v_{n_k},T v_{n_k}\rangle = \langle AT v,T v\rangle =\Vert Tv\Vert _A^2. \end{aligned}$$

As \(\Vert v\Vert _A \le 1,\) we conclude that \(v \in M_A^T \cap \overline{R(A)}\) and \( 1 = \Vert v_{n_k}\Vert _A \longrightarrow \Vert v\Vert _A = 1.\) As \(v_{n_k} \rightharpoonup v\) with respect to \(\Vert \cdot \Vert ,\) clearly, \(v_{n_k} \rightharpoonup v\) with respect to \(\Vert \cdot \Vert _A.\) Since \( (\overline{R(A)}, \langle ~,~ \rangle _A) \) is a Hilbert space, it follows that \( v_{n_k} \longrightarrow v \) with respect to \( \Vert \cdot \Vert _A. \) As \(\Vert \cdot \Vert _A \) and \(\Vert \cdot \Vert \) are equivalent norms on \(\overline{R(A)} \), therefore, \( v_{n_k} \longrightarrow v \) with respect to \( \Vert \cdot \Vert . \) This establishes the theorem. \(\square \)

Remark 2.5

Note that, \(M_A^T \cap \overline{R(A)}\) is also compact with respect to \(\Vert \cdot \Vert _A\) in \(\overline{R(A)}\), due to the fact that \(\Vert \cdot \Vert _A\) and \(\Vert \cdot \Vert \) are equivalent norms on \(\overline{R(A)}.\)

In [18], the author has characterized the \(A-\)Birkhoff–James orthogonality of \(A-\)bounded operators on a Hilbert space with the help of \(A-\)norming sequences. In the finite-dimensional case, the Bhatia-\( \breve{S} \)emrl Theorem follows from the said characterization, as shown in Theorem 2.4 of [18]. The main difference between the characterizations of \( A- \)Birkhoff–James orthogonality of operators in the infinite-dimensional case and the finite-dimensional case is that the approximate orthogonality of the images of norming sequences in the former case can be strengthened to the exact orthogonality of the images of a norming vector in the later case. For the convenience of the readers, let us mention the relevant results from [18] and [5].

Theorem 2.6

(Zamani, Theorem 2.2 of [18]). Let \(T,S \in B_{A^{1/2}}({\mathbb {H}}).\) Then the following conditions are equivalent:

  1. (i)

    there exists a sequence of \(A-\)unit vectors \(\{x_n\}\) in \({\mathbb {H}}\) such that

    \(\lim _{n \rightarrow \infty }\Vert Tx_n\Vert _A =\Vert T\Vert _A\) and \(\lim _{n \rightarrow \infty }\langle Tx_n,Sx_n \rangle _A = 0.\)

  2. (ii)

    \(T \bot _A^B S \).

Theorem 2.7

(Bhatia and \(\breve{S}\)emrl, Theorem 1.1 of [5]) A matrix A is orthogonal to a matrix B if and only if there exists a unit vector \(x \in {\mathbb {H}}\) such that \(\Vert Ax\Vert =\Vert A\Vert \) and \(\langle Ax,Bx\rangle = 0.\)

In our next theorem, we show that under certain additional conditions, the said strengthening of the \( A- \)Birkhoff–James orthogonality of \( A- \)bounded operators can be preserved even in the infinite-dimensional case.

Theorem 2.8

Let \({\mathbb {H}}\) be a Hilbert space and let \(A \in {\mathbb {L}}({\mathbb {H}})\) be positive such that \(B_{{\mathbb {H}}(A)} \cap \overline{R(A)}\) is bounded with respect to \(\Vert \cdot \Vert .\) Let \(T,S \in {\mathbb {K}}({\mathbb {H}}) \cap B_{A^{1/2}}({\mathbb {H}}).\) Then \(T \bot _A^BS \) if and only if there exists \(v \in M^T_A\) such that \(Tv\bot _ASv.\)

Proof

The sufficient part of the theorem follows easily. Indeed, suppose that there exists \(v \in M^T_A\) such that \(Tv\bot _ASv.\) Then

$$\begin{aligned} \Vert T+\lambda S\Vert _A\ge & {} \Vert Tv+\lambda Sv\Vert _A \\\ge & {} \Vert Tv\Vert _A \\= & {} \Vert T\Vert _A~\text {for all }\lambda \in {\mathbb {K}} (={\mathbb {R}},~{\mathbb {C}}). \end{aligned}$$

Let us prove the necessary part of the theorem. By Theorem 2.2 of [18], there exists a sequence \(\{x_n\} \subseteq S_{{\mathbb {H}}(A)}\) such that

$$\begin{aligned} \lim _{n \rightarrow \infty }\Vert Tx_n\Vert _A =\Vert T\Vert _A~\text {and} \lim _{n \rightarrow \infty }\langle Tx_n,Sx_n \rangle _A = 0.\end{aligned}$$

Since \({\mathbb {H}} = N(A)\oplus \overline{R(A)}\), it follows that \(x_n = u_n + v_n\) for each \(n \in {\mathbb {N}}\), where \(u_n \in N(A)\) and \(v_n \in \overline{R(A)}.\) Clearly, \(\Vert u_n\Vert _A =0 \) for all \(n \in {\mathbb {N}}\). Thus \(\Vert x_n\Vert _A = \Vert v_n + u_n\Vert _A \le \Vert v_n\Vert _A + \Vert u_n\Vert _A=\Vert v_n\Vert _A\). Again, \(\Vert x_n\Vert _A = \Vert v_n + u_n\Vert _A \ge \Vert v_n\Vert _A - \Vert u_n\Vert _A=\Vert v_n\Vert _A\). Therefore, \(\Vert x_n\Vert _A = \Vert v_n\Vert _A\) for each \(n \in {\mathbb {N}}.\) As \(\{x_n\} \subseteq S_{{\mathbb {H}}(A)}\), we conclude that \(\{v_n\} \subseteq S_{{\mathbb {H}}(A)} \cap \overline{R(A)}.\) Since \(T,S \in B_{A^{1/2}}({\mathbb {H}})\), \(\Vert Tu_n\Vert _A = \Vert Su_n\Vert _A = 0\) for all \(n \in {\mathbb {N}}\). Hence \(\Vert Tx_n\Vert _A = \Vert Tv_n\Vert _A\) and \(\Vert Sx_n\Vert _A = \Vert Sv_n\Vert _A\) for each \(n \in {\mathbb {N}}.\) Since \({\mathbb {H}}\) is reflexive and \(B_{{\mathbb {H}}(A)} \cap \overline{R(A)}\) is closed, convex and bounded with respect to \(\Vert \cdot \Vert ,\) therefore, \(B_{{\mathbb {H}}(A)} \cap \overline{R(A)}\) is weakly compact with respect to \(\Vert \cdot \Vert .\) Thus the sequence \(\{v_n\}\) has a weakly convergent subsequence. Without loss of generality we may assume that \(v_n \rightharpoonup v\) with respect to \(\Vert \cdot \Vert \) on \({\mathbb {H}}\), for some \(v \in B_{{\mathbb {H}}(A)} \cap \overline{R(A)}.\) Since \(T,S \in {\mathbb {K}}({\mathbb {H}}),\) it follows that \(Tv_n \longrightarrow Tv\) and \(Sv_n \longrightarrow Sv\) with respect to \(\Vert \cdot \Vert \) in \({\mathbb {H}}\). Therefore,

$$\begin{aligned} \Vert T\Vert ^2_A= & {} \lim _{n \rightarrow \infty }\Vert Tx_n\Vert ^2_A = \lim _{n \rightarrow \infty }\Vert Tv_n\Vert ^2_A\\= & {} \lim _{n \rightarrow \infty } \langle ATv_n, Tv_n \rangle = \Vert Tv\Vert ^2_A. \end{aligned}$$

As \(\Vert v\Vert _A \le 1,\) we conclude that \(v \in M_A^T \cap \overline{R(A)}.\)

Next we show that \(Tv\bot _A Sv.\) As \(\Vert Tu_n\Vert _A = \Vert Su_n\Vert _A = 0,\) it is immediate that \(Tu_n,Su_n \in N(A)\) for all \(n \in {\mathbb {N}}.\) Since A is positive, it follows that \(N(A)=N(A^{1/2}).\) Hence \(A^{1/2}(Tu_n)=A^{1/2}(Su_n)= \theta\) for all \(n \in {\mathbb {N}}.\) Therefore, we have

$$\begin{aligned} 0= & {} \lim _{n \rightarrow \infty }\langle Tx_n,Sx_n \rangle _A = \lim _{n \rightarrow \infty }\langle Tu_n + Tv_n,Su_n + Sv_n \rangle _A \\= & {} \lim _{n \rightarrow \infty }\langle A(Tu_n + Tv_n),Su_n + Sv_n \rangle \\= & {} \lim _{n \rightarrow \infty }\langle A^{1/2}(Tu_n + Tv_n), A^{1/2}(Su_n + Sv_n) \rangle \\= & {} \lim _{n \rightarrow \infty }\langle A^{1/2}Tv_n, A^{1/2}Sv_n \rangle = \lim _{n \rightarrow \infty }\langle ATv_n,Sv_n \rangle = \langle ATv,Sv \rangle = \langle Tv,Sv \rangle _A. \end{aligned}$$

Thus \(Tv\bot _ASv.\) This completes the proof of the theorem. \(\square \)

We end this article with the following closing remark:

Remark 2.6

Note that in Theorem 2.8, if \({\mathbb {H}}\) is finite-dimensional and \(A = I,\) then the Bhatia-\( \breve{S} \)emrl Theorem (Theorem 1.1 of [5]) follows immediately. In particular, the finite-dimensional Bhatia-\( \breve{S} \)emrl Theorem can be extended verbatim to the infinite-dimensional setting of semi-Hilbertian spaces, provided certain additional conditions are satisfied. We further observe that Theorem 2.4 of [18] follows as a corollary to Theorem 2.8, since in a finite-dimensional Hilbert space, \(B_{{\mathbb {H}}(A)} \cap \overline{R(A)}\) is bounded with respect to \( \Vert \cdot \Vert \) and every linear operator is compact.