1 Introduction and preliminaries

Let \( {\mathcal {A}} \) be a unital \( C^{*} \)-algebra with unit \(1_{{\mathcal {A}}}.\) We denote by \({\mathcal {A}}^{\prime }\) and \({\mathcal {Z}}({\mathcal {A}})\) the topological dual space and the center of \({\mathcal {A}},\) respectively. The adjoint of any element \( x \in {\mathcal {A}} \) is denoted, as usual, by \( x^* .\) Also, \({\textrm{Re}}(x)=\dfrac{1}{2}(x+x^*)\) is reserved to indicate the real part of x. An element a of \({\mathcal {A}}\) is called positive (written by \(a\ge 0),\) if a is selfadjoint whose spectrum \(\sigma (a)\) is contained in \([0,\infty ).\) It is known that if \(a\in {\mathcal {A}}\) is positive, then there exists a unique positive element \(b\in {\mathcal {A}}\) such that \(a=b^2.\) Such an element b is called the positive square root of a and is denoted by \(a^{\frac{1}{2}}.\) The symbol \({\mathcal {A}}^+\) stands for the cone of positive elements in \( {\mathcal {A}} .\) If in addition a is invertible, then \(a^{\frac{1}{2}}\) is invertible too and its inverse is denoted by \(a^{-\frac{1}{2}}.\) A linear functional f on \({\mathcal {A}}\) is called positive if \(f(a)\ge 0\) for every positive element \(a\in {\mathcal {A}}.\) Given a positive functional f on \({\mathcal {A}},\) the following well-known version of the Cauchy–Schwarz inequality holds for every \(x,y\in {\mathcal {A}}\): \(|f(x^*y)|^2\le f(x^*x)f(y^*y).\)

A state on \({\mathcal {A}}\) is a positive linear functional whose norm is equal to one. It is well-known that a linear functional on \({\mathcal {A}}\) is positive if and only if \(f(1_{{\mathcal {A}}})=\Vert f\Vert ;\) see [17, Corollary 3.3.4]. Let \( {\mathcal {S}}({\mathcal {A}}) \) be the set of all states on \( {\mathcal {A}} .\) Then

$$\begin{aligned} {\mathcal {S}}({\mathcal {A}}) = \{ f\in {\mathcal {A}}^{'} \, : \, f(1_{{\mathcal {A}}})= \Vert f \Vert =1 \}. \end{aligned}$$

Birkhoff–James orthogonality of elements in a normed linear space was introduced by Birkhoff in [11] and developed by James [14] to generalize the concept of orthogonality in inner product spaces. If x and y are vectors of a normed linear space \((X,\Vert \cdot \Vert ),\) then x is said to be Birkhoff–James orthogonal to y,  in short \(x\perp _{BJ}y,\) if

$$\begin{aligned} \Vert x+\lambda y\Vert \ge \Vert x\Vert \quad (\forall \lambda \in {\mathbb {C}}). \end{aligned}$$

The concept of the strong Birkhoff–James orthogonality in \(C^*\)-algebras as a natural generalization of Birkhoff–James orthogonality was introduced and studied in [4, 6]. Let \(x,y\in {\mathcal {A}}.\) Then x is said to be strong Birkhoff–James orthogonal to y,  denoted by \(x\perp _{S-BJ}y,\) if

$$\begin{aligned} \Vert x+yb\Vert \ge \Vert x\Vert \quad (\forall b\in {\mathcal {A}}). \end{aligned}$$

We also recall that two elements x and y of \( {\mathcal {A}} \) are orthogonal with respect to the \( {\mathcal {A}} \)-valued inner product \( \langle x,y \rangle := x^* y \) if \( \langle x,y \rangle =0 .\) It was shown in [4] the following relation between the strong and the classical Birkhoff–James orthogonality:

$$\begin{aligned} \langle x,y\rangle =0\Rightarrow x\perp _{S-BJ}y\Rightarrow x\perp _{BJ}y\quad (\forall x,y\in {\mathcal {A}}). \end{aligned}$$

It is well-known that the Birkhoff–James orthogonality of vectors in normed linear spaces can be characterized in terms of linear functionals [14]. Over the years, the problem of finding characterizations of Birkhoff–James orthogonality of matrices and generally of the elements of \(C^*\)-algebras has been considered by many mathematicians. A complete characterization of Birkhoff–James orthogonality of bounded linear operators defined on Hilbert spaces obtained by Bhatia and S̆emrl [10] (see also, [3, 9, 18]). Some famous and useful Characterizations of the (strong) Birkhoff–James orthogonality in \(C^*\)-algebra \({\mathcal {A}}\) and in a more general setting Hilbert \(C^*\)-modules over \({\mathcal {A}}\) in terms of the elements of state space \({\mathcal {S}}({\mathcal {A}})\) have been obtained in [5, 9, 16]. The characterization of the (strong) Birkhoff–James orthogonality for elements of a \(C^*\)-algebra by means of its state space were obtained as follows:

Theorem 1.1

[5, Theorem 2.7] An element \(x\in {\mathcal {A}}\) is Birkhoff–James orthogonal to another element \(y\in {\mathcal {A}},\) if and only if there is \(f\in {\mathcal {S}}({\mathcal {A}})\) such that \(f(x^*x)=\Vert x\Vert ^2\) and \(f(x^*y)=0.\)

Theorem 1.2

[4, Theorem 2.5] An element \(x\in {\mathcal {A}}\) is strong Birkhoff–James orthogonal to another element \(y\in {\mathcal {A}}\) if and only if there is \(f\in {\mathcal {S}}({\mathcal {A}})\) such that \(f(x^*x)=\Vert x\Vert ^2\) and \(f(\langle x,y\rangle \,\langle y,x\rangle )=0\) if and only if there is \(f\in {\mathcal {S}}({\mathcal {A}})\) such that \(f(x^*x)=\Vert x\Vert ^2 \) and \( f(\langle x,y\rangle b)=0\) for all \(b\in {\mathcal {A}}.\)

The classes of \(C^*\)-algebras in which any two of these orthogonality relationships coincide have been described in [4, 6]. More precisely,

Theorem 1.3

[6, Corollary 4.10] Let \( {\mathcal {A}} \) be a nonzero \( C^* \)-algebra. Then the following statements are equivalent : 

  1. (1)

    For all \( x,y \in {\mathcal {A}} ,\) \( x \perp _{S-BJ} y \) if and only if \( \langle x, y\rangle =0 .\)

  2. (2)

    For all \( x,y \in {\mathcal {A}} ,\) \( x \perp _{BJ} y \) if and only if \( x \perp _{S-BJ} y .\)

  3. (3)

    \( {\mathcal {A}} \) is isomorphic to \( {\mathbb {C}} .\)

Let a be a nonzero positive element of \( {\mathcal {A}} .\) A generalization of state space of \({\mathcal {A}}\) was introduced in [1] as follows:

$$\begin{aligned} {\mathcal {S}}_{a}({\mathcal {A}}):= \{ \varphi \in {\mathcal {A}}^{'} \, : \, \varphi \ge 0, \, \varphi ( a ) = 1\}=\left\{ \dfrac{f}{f(a)} \, : \, f\in {\mathcal {S}}({\mathcal {A}}), \, f(a) \ne 0 \right\} . \end{aligned}$$

Observe that if \(a=1_{{\mathcal {A}}},\) then \({\mathcal {S}}_a({\mathcal {A}})={\mathcal {S}}({\mathcal {A}}).\) It has been proved in [1] that \({\mathcal {S}}_a({\mathcal {A}})\) is a nonempty convex and \(w^*\)-closed subset of \({\mathcal {A}}^{\prime }.\) But, unlike \({\mathcal {S}}({\mathcal {A}}),\) the set \({\mathcal {S}}_a({\mathcal {A}})\) may not be \(w^*\)-compact. In fact, according to the following result, \({\mathcal {S}}_a({\mathcal {A}})\) is \(w^*\)-compact if and only if a is invertible.

Proposition 1.4

[1, Proposition 2.3] Let \({\mathcal {A}}\) be a unital \(C^*\)-algebra and let \(a\in {\mathcal {A}}\) be a positive element. Then the following statement are equivalent : 

  1. (1)

    \({\mathcal {S}}_a({\mathcal {A}})\) is \(w^*\)-compact.

  2. (2)

    a is invertible.

For any element \(x\in {\mathcal {A}},\) the a-operator semi-norm \(\Vert \cdot \Vert _a :{\mathcal {A}}\rightarrow [0,\infty )\) is defined by

$$\begin{aligned} \Vert x \Vert _{a} := \sup \left\{ \sqrt{\varphi (x^* a x )} \, : \varphi \in {\mathcal {S}}_a ({\mathcal {A}}) \right\} . \end{aligned}$$

Due to the Proposition 1.4, if a is not invertible, then \({\mathcal {S}}_a({\mathcal {A}})\) is not \(w^*\)-compact, and so it may happen that \(\Vert x\Vert _a=\infty \) for some \(x\in {\mathcal {A}};\) see [1, Example 3.2]. Denote by \({\mathcal {A}}^a:=\{x\in {\mathcal {A}} : \Vert x\Vert _a<\infty \}.\) It was shown in [1] that \(\Vert \cdot \Vert _a\) is a submultiplicative semi-norm on \({\mathcal {A}}^a;\) i.e., \(\Vert xy\Vert _a\le \Vert x\Vert _a\Vert y\Vert _a\) for all \(x,y\in {\mathcal {A}}^a.\) Also, \(\Vert x\Vert _a=0\) if and only if \(ax=0.\) In addition, if a is invertible, then \( \Vert \cdot \Vert _a \) is a norm on \( {\mathcal {A}} .\) Consequently, \(\Vert \cdot \Vert _{1_{{\mathcal {A}}}}\) is equal to the \(C^*\)-norm \(\Vert \cdot \Vert \) of \({\mathcal {A}}.\)

An element \(x^\sharp \in {\mathcal {A}}\) is called an a-adjoint of \(x\in {\mathcal {A}}\) if \(ax^\sharp =x^*a.\) The set of all a-adjointable elements of \({\mathcal {A}}\) is denoted by \({\mathcal {A}}_a.\) Note that \({\mathcal {A}}_a={\mathcal {A}}\) if \({\mathcal {A}}\) is commutative.

In [1, Corollary 4.9] it was proved that if \( x \in {\mathcal {A}}_a \) and \(x^\sharp \) is an a-adjoint of it, then

$$\begin{aligned} \Vert x\Vert _{a}^2 = \Vert xx^\sharp \Vert _{a} = \Vert x^\sharp x\Vert _{a} =\Vert x^\sharp \Vert _{a}^2 . \end{aligned}$$
(1.1)

An element \(x\in {\mathcal {A}}\) is said to be a-selfadjoint if ax is selfadjoint; i.e., \(ax=x^*a.\) Moreover, any element \( x \in {\mathcal {A}}_a \) can be written as \( x=x_1 + i x_2 ,\) where \(x_1\) and \(x_2\) are a-selfadjoint. In fact, if \( x^{\sharp } \) is an a-adjoint of x,  then

$$\begin{aligned} x = \dfrac{x+x^\sharp }{2} + i \dfrac{x-x^\sharp }{2i}. \end{aligned}$$
(1.2)

This decomposition is not unique, since there might be many (or none) a-adjoints \(x^\sharp \) of x;  see e.g., [1, 8]. Note that if we assume that a is invertible, then x has the unique a-adjoint \(x^\sharp =a^{-1}x^*a,\) and, therefore, the decomposition (1.2) is unique.

The notions, \({\mathcal {S}}_a({\mathcal {A}})\) and \(\Vert \cdot \Vert _a\) were introduced in [1] to generalize algebraic numerical range and algebraic numerical radius of elements of \(C^*\)-algebra \({\mathcal {A}}.\) To study abundant results related to these concepts the reader is referred to [1, 2, 5].

In this paper, we investigate the notions of Birkhoff–James orthogonality and its strong version in an unital \(C^*\)-algebra \({\mathcal {A}}\) with respect to the norm \(\Vert \cdot \Vert _a,\) whenever \(a\in {\mathcal {A}}\) is a positive and invertible element.

In Sect. 2 first, the main properties of a-Birkhoff–James orthogonality are studied and a variety of examples in simple \(C^*\)-algebra \({\mathbb {M}}_n({\mathbb {C}})\) are presented to illustrate the relationship between a-Birkhoff–James orthogonality and Birkhoff–James orthogonality. Next, a complete characterization of a-Birkhoff–James orthogonality in terms of elements of the generalized state space \({\mathcal {S}}_a({\mathcal {A}})\) is presented. As an application, a characterization for the best approximation to elements of \({\mathcal {A}}\) in a subspace \({\mathcal {B}}\) is obtained with respect to \(\Vert \cdot \Vert _a.\) Moreover, a generalization of the well-known distance formula which obtained by Williams in [19] is given.

Section 3 is devoted to the study of strong a-Birkhoff–James orthogonality in unital \(C^*\)-algebras. The classes of unital \(C^*\)-algebras in which the a-Birkhoff–James orthogonality coincides with the strong a-Birkhoff–James orthogonality are described.

In particular, we prove that if \(x\perp ^a_{S-BJ}y\) implies \(\langle x,y\rangle _a:=x^*ay=0,\) for all \(x,y\in {\mathcal {A}},\) then the center of \({\mathcal {A}}\) is trivial, i.e., the only central elements of \({\mathcal {A}}\) are multiplies of the identity. Moreover, we prove that the right additivity (left additivity) of (strong) a-Birkhoff–James orthogonality in \({\mathcal {A}}\) concludes that \({\mathcal {Z}}({\mathcal {A}})\cong {\mathbb {C}}1_{{\mathcal {A}}}.\)

2 a-Birkhoff–James orthogonality in \(C^*\)-algebras

Let \( {\mathcal {B}}({\mathcal {H}}) \) be the \( C^* \)-algebra of all bounded linear operators on a complex Hilbert space \( {\mathcal {H}} \) with inner product \( \langle \cdot , \cdot \rangle .\) If \(\mathrm{dim {\mathcal {H}}}=n,\) then we identify \({\mathcal {B}}({\mathcal {H}}) \) with the simple \(C^*\)-algebra \({\mathbb {M}}_n({\mathbb {C}})\) of all \(n\times n\) complex matrices and denote the identity matrix by \(I_n.\) Assume that \(A\in {\mathcal {B}}({\mathcal {H}})\) is a positive operator, which induces a positive semi-definite sesquilinear form \(\langle \cdot ,\cdot \rangle _A:{\mathcal {H}}\times {\mathcal {H}}\rightarrow {\mathbb {C}}\) defined by \(\langle x,y\rangle _A=\langle Ax,y\rangle .\) The semi-norm \(\Vert \cdot \Vert _A\) induced by \(\langle \cdot ,\cdot \rangle _A\) is defined by \(\Vert x\Vert _A=\sqrt{\langle Ax,x\rangle }\) for every \(x\in {\mathcal {H}}.\) Furthermore, the set of all A-bounded operators on \({\mathcal {H}}\) is defined by

$$\begin{aligned} {\mathcal {B}}_{A^{\frac{1}{2}}} ( {\mathcal {H}}) := \{ T \in {\mathcal {B}}({\mathcal {H}}) \, : \, \exists \, c > 0, \, \Vert T x \Vert _{A} \le c \Vert x \Vert _{A}, \quad \forall x \in {\mathcal {H}} \}. \end{aligned}$$

In fact, \({\mathcal {B}}_{A^{\frac{1}{2}}} ( {\mathcal {H}})\) is a unital subalgebra of \({\mathcal {B}}({\mathcal {H}}) \) which is equipped with the semi-norm

$$\begin{aligned} \gamma _A (T) := \sup _{\Vert x\Vert _A=1} \sqrt{\langle ATx , Tx \rangle }\quad (T \in {\mathcal {B}}_{A^{\frac{1}{2}}} ( {\mathcal {H}})). \end{aligned}$$

The Birkhoff–James orthogonality with respect to the semi-norm \(\gamma _A(\cdot )\) (called A-Birkhoff–James orthogonality) was studied by Zamani in [20]. An operator \( T \in {\mathcal {B}}_{A^{\frac{1}{2}}} ( {\mathcal {H}}) \) is called A-Birkhoff–James orthogonal to the operator \( S \in {\mathcal {B}}_{A^{\frac{1}{2}}} ( {\mathcal {H}}) ,\) denoted by \( T \perp _{BJ}^{A} S ,\) if \( \gamma _{A} (T + \lambda S) \ge \gamma _{A} (T)\) for all \(\lambda \in {\mathbb {C}}.\) The following characterization of A-Birkhoff–James orthogonality which extends the Bhatia and S̆emrl Theorem for A-bounded operators has been obtained as follow:

Theorem 2.1

[20, Theorem 2.2] Let \( T,S \in {\mathcal {B}}_{A^{\frac{1}{2}}} ( {\mathcal {H}}) .\) The following conditions are equivalent : 

  1. (1)

    \( T \perp _{BJ}^{A} S .\)

  2. (2)

    There exists a sequence \( \{ h_n \} \) of A-unit vectors \((\Vert h_n\Vert _A=1)\) in \( {\mathcal {H}} \) such that

    $$\begin{aligned} \lim _{n \rightarrow \infty } \Vert T h_n \Vert _{A} = \gamma _{A} (T) \quad \text {and} \quad \lim _{n \rightarrow \infty } \langle T h_n , S h_n \rangle _A = 0 . \end{aligned}$$
    (2.1)

We recall that by the Gelfand–Naimark Theorem, any unital \(C^*\)-algebra \({\mathcal {A}}\) can be considered as a norm closed \(*\)-subalgebra of \({\mathcal {B}}({\mathcal {H}})\) for some Hilbert space \({\mathcal {H}}\) . In fact, there exists an unital faithful \(*\)-representation \(\pi :{\mathcal {A}}\rightarrow {\mathcal {B}}({\mathcal {H}})\) such that \(\Vert x\Vert =\Vert \pi (x)\Vert \) for all \(x\in {\mathcal {A}};\) see e.g., [12, 17]. It was proved that in [1], if \(\pi :{\mathcal {A}}\rightarrow {\mathcal {B}}({\mathcal {H}})\) is a unital faithful \(*\)-representation of \({\mathcal {A}},\) then

$$\begin{aligned} \Vert x \Vert _a = \gamma _{\pi (a)} ( \pi (x)) \end{aligned}$$
(2.2)

for any \(x\in {\mathcal {A}}.\) As a direct consequence of this fact, we have \({\mathcal {A}}^A={\mathcal {B}}_{A^{\frac{1}{2}}} ( {\mathcal {H}})\) for \({\mathcal {A}}={\mathcal {B}}({\mathcal {H}})\) and all positive operator \(A\in {\mathcal {B}}({\mathcal {H}}).\)

Let \({\mathcal {A}}\) be a unital \(C^*\)-algebra and let \(a\in {\mathcal {A}}\) be a nonzero positive element. It was proved in [1, Theorem 3.9] that \({\mathcal {A}}_a\subset {\mathcal {A}}^a.\) Now, if we assume that a is a positive and invertible element of \({\mathcal {A}},\) then for every \(x\in {\mathcal {A}}\) the equation \(ax^\sharp =x^*a\) has the unique solution \(x^\sharp =a^{-1}x^*a,\) and so every \(x\in {\mathcal {A}}\) is a-adjointable. Therefore \({\mathcal {A}}^a={\mathcal {A}}.\)

From now on we assume that \({\mathcal {A}}\) is a unital \(C^*\)-algebra and \(a\in {\mathcal {A}}\) is a positive and invertible element. Let us introduce the concept of a-Birkhoff–James orthogonality with respect to the norm \(\Vert \cdot \Vert _a\) in \(C^*\)-algebras.

Definition 2.2

Let \({\mathcal {A}}\) be a unital \(C^*\)-algebra and \(a\in {\mathcal {A}}\) be a positive and invertible element. We say that an element \( x \in {\mathcal {A}} \) is Birkhoff–James orthogonal with respect to the norm \( \Vert \cdot \Vert _{a} \) (a-Birkhoff–James orthogonal) to an element \( y \in {\mathcal {A}} ,\) in short \( x \perp _{BJ}^{a} y ,\) if

$$\begin{aligned} \Vert x + \lambda y \Vert _{a} \ge \Vert x \Vert _{a}\quad (\forall \lambda \in {\mathbb {C}}). \end{aligned}$$

First, note that a-Birkhoff–James orthogonality reduces to the Birkhoff–James orthogonality when \(a=1_{{\mathcal {A}}}.\) Also, it is easy to see that a-Birkhoff–James orthogonality is homogenous; i.e., if \( x \perp _{BJ}^{a} y ,\) then \( \alpha x \perp _{BJ}^{a} \beta y \) for all \( \alpha , \beta \in {\mathbb {C}} .\) It is trivial for \( \alpha =0 \) or \( \beta =0 .\) So, suppose that \( \alpha \) and \(\beta \) are nonzero complex numbers. For each \( \lambda \in {\mathbb {C}} ,\) we have

$$\begin{aligned} \Vert \alpha x + \lambda \beta y \Vert _a = \left\| \alpha \left( x + \dfrac{\beta }{\alpha } \lambda y \right) \right\| _a = |\alpha | \left\| \left( x +\dfrac{\beta }{\alpha } \lambda y\right) \right\| _a \ge | \alpha | \, \Vert x \Vert _a = \Vert \alpha x \Vert _a. \end{aligned}$$

It follows that \( \alpha x \perp _{BJ}^{a} \beta y .\)

Also, a-Birkhoff–James orthogonality is non-degenerate. Indeed, let \( 0\ne x \in {\mathcal {A}} \) and \( x \perp _{BJ}^{a} x .\) Then \( \Vert x+ \lambda x\Vert _a \ge \Vert x\Vert _a \) for all \( \lambda \in {\mathbb {C}} .\) For \( \lambda = -1 ,\) we get \( \Vert x\Vert _a =0 ,\) and so \( ax =0 .\) Therefore \( x=0 ,\) since a is invertible.

Moreover, for any two nonzero elements \(x,y\in {\mathcal {A}},\) if x is orthogonal to y in the a-Birkhoff–James sense, then x and y are linearly independent. In fact, if we assume to the contrary that there exists \( k \in {\mathbb {C}} \) such that \( y=k x ,\) then \( x \perp _{BJ}^{a} kx .\) It follows that \( x \perp _{BJ}^{a} x ,\) since a-Birkhoff–James orthogonality is homogenous. Hence \( ax =0 ,\) and so \( x=0 ,\) which is a contradiction.

Let \( f\in {\mathcal {S}}({\mathcal {A}}).\) According to [1], the linear functional defined by

$$\begin{aligned} \varphi (z)=f\left( a^{-\frac{1}{2}}za^{-\frac{1}{2}}\right) \quad (z\in {\mathcal {A}}) \end{aligned}$$
(2.3)

belongs to \({\mathcal {S}}_a({\mathcal {A}}).\) Now, let \(x \in {\mathcal {A}}\) and let \(a \in {\mathcal {A}} \) be positive and invertible such that \( ax=xa .\) Then

$$\begin{aligned} \Vert x\Vert _a^2=\sup _{\varphi \in {\mathcal {S}}_a({\mathcal {A}})}\varphi (x^*ax) =\sup _{f\in {\mathcal {S}}({\mathcal {A}})} f \left( a^{-\frac{1}{2}} x^* a x a^{-\frac{1}{2}} \right) =\sup _{f\in {\mathcal {S}}({\mathcal {A}})} f(x^*x)=\Vert x\Vert ^2. \end{aligned}$$
(2.4)

Also, note that \( x^\sharp = a^{-1} x^* a \) is a-adjoint of x,  and so it follows from (1.1) that

$$\begin{aligned} \Vert x\Vert _{a}^2 = \Vert x a^{-1} x^* a \Vert _{a} = \Vert a^{-1} x^* a x\Vert _a = \Vert a^{-1} x^* a \Vert _{a}^2. \end{aligned}$$

Hence

$$\begin{aligned} \Vert x\Vert _{a}^2 = \Vert x x^*\Vert _{a} = \Vert x^* x \Vert _{a} = \Vert x^*\Vert _{a}^2. \end{aligned}$$

Since there is at most one norm on a \(*\)-algebra making it a \(C^*\)-algebra, the following result is obtained.

Corollary 2.3

If \( {\mathcal {A}} \) is a commutative and unital \( C^* \)-algebra and \(a\in {\mathcal {A}}\) is positive and invertible,  then \(\Vert \cdot \Vert _a\) agrees with the \(C^*\)-norm of \(C^*\)-algebra \({\mathcal {A}}.\) In this case, the a-Birkhoff–James orthogonality and the Birkhoff–James orthogonality are equivalent on \({\mathcal {A}}.\)

It should be noted that \(\Vert \cdot \Vert _a\) does not satisfy to the \(C^*\)-condition in noncommutative \( C^* \)-algebra, even when a is invertible. To make this clear, we present the following example.

Example 2.4

Let \( {\mathbb {M}}_2 ({\mathbb {C}}) \) be the \( C^* \)-algebra of all \( 2 \times 2 \) complex matrices, and let \({\textrm{Tr}}\) be the usual trace functional on \( {\mathbb {M}}_2 ({\mathbb {C}}) .\) According to the Example 2.2 of [1], for any positive matrix \( h \in {\mathbb {M}}_2 ({\mathbb {C}}) ,\) let \( \varphi _h \) be the positive linear functional given by

$$\begin{aligned} \varphi _h (x) := {\textrm{Tr}}(hx), \quad (x \in {\mathbb {M}}_2 ({\mathbb {C}})). \end{aligned}$$

It is known that any state on \( {\mathbb {M}}_2 ({\mathbb {C}}) \) is of the form \( \varphi _h \) with \( {\textrm{Tr}}(h) =1 .\) For a positive matrix \( a \in {\mathbb {M}}_2 ({\mathbb {C}}) ,\) we have

$$\begin{aligned} {\mathcal {S}}_{a} ({\mathbb {M}}_{2}({\mathbb {C}})) = \{ \varphi _h \, : \, h \in {\mathbb {M}}_2 ({\mathbb {C}})^{+} \quad \text {and} \quad {\textrm{Tr}}(ha) =1\}. \end{aligned}$$

Now, let \( a=\left[ \begin{array}{cccc} 2&{}0 \\ 0&{}1 \end{array}\right] .\) Then with some simple matrix computations, we conclude that

$$\begin{aligned} {\mathcal {S}}_{a} ({\mathbb {M}}_{2}({\mathbb {C}})) =\{ \varphi _h \, : \, h\in {\mathcal {L}}_a\}, \end{aligned}$$

where

$$\begin{aligned} {\mathcal {L}}_a:=\left\{ h =\left[ \begin{array}{cccc} h_{11}&{}h_{12} \\ \overline{h}_{12}&{} h_{22} \end{array}\right] \in {\mathbb {M}}_2 ({\mathbb {C}})^{+}\, : \, h_{12} \in {\mathbb {C}}, \, h_{11}, h_{22} \ge 0 \ \text {and}\ 2h_{11} +h_{22} =1\right\} . \end{aligned}$$

Hence for \( x=\left[ \begin{array}{cccc} 0&{}2 \\ 1&{}0 \end{array}\right] ,\) we get

$$\begin{aligned} \Vert x \Vert _{a}^{2} = \sup _{\varphi _{h} \in {\mathcal {S}}_a ({\mathbb {M}}_2 ({\mathbb {C}}))} \varphi _{h} (x^* a x)&= \sup _{h\in {\mathcal {L}}_a} {\textrm{Tr}}\left( \left[ \begin{array}{cccc} h_{11}&{} 8h_{12} \\ \overline{h}_{12}&{} 8h_{22} \end{array}\right] \right) \\&=\sup _{2h_{11}+h_{22}=1, h_{11}, h_{22} \ge 0} h_{11} +8h_{22} = 8. \end{aligned}$$

But similarly, we have

$$\begin{aligned} \Vert x^* x \Vert _{a}^{2}= \sup _{\varphi _{h} \in {\mathcal {S}}_a( {\mathbb {M}}_2 ({\mathbb {C}}))} \varphi _{h}( (x^* x) a (x^* x)) =\sup _{2h_{11}+h_{22}=1, h_{11}, h_{22} \ge 0} 2h_{11} +16h_{22} = 16. \end{aligned}$$

The following provide us with examples reveal that the a-Birkhoff–James orthogonality is independent from the Birkhoff–James orthogonality in unital and noncommutative \(C^*\)-algebras, even when a is positive and invertible.

Example 2.5

In the context of the same \( a=\left[ \begin{array}{cccc} 2 &{}0 \\ 0&{}1 \end{array}\right] \) as, and similarly to the method we applied in the previous example, let \( x=\left[ \begin{array}{cccc} 0 &{}-1 \\ 0&{}1 \end{array}\right] \) and \( y=\left[ \begin{array}{cccc} 0 &{}1 \\ 0&{}1 \end{array}\right] \) be matrices in \({\mathbb {M}}_2({\mathbb {C}}).\) Then

$$\begin{aligned} \Vert x\Vert _{a}^2&= \sup _{\varphi _{h} \in {\mathcal {S}}_a ({\mathbb {M}}_2 ({\mathbb {C}}))} \varphi _{h} (x^* a x)= \sup _{h\in {\mathcal {L}}_a}{\textrm{Tr}} (h (x ^* a x)) \\&= \sup _{h\in {\mathcal {L}}_a} {\textrm{Tr}} \left( \left[ \begin{array}{cccc} h_{11}&{} h_{12} \\ \overline{h}_{12} &{} h_{22} \end{array}\right] \left[ \begin{array}{cccc} 0&{} 0 \\ 0&{} 3 \end{array}\right] \right) =\sup _{2h_{11}+h_{22}=1, h_{11}, h_{22} \ge 0} 3h_{22} = 3. \end{aligned}$$

Also, for every \( \lambda \in {\mathbb {C}} ,\) we have

$$\begin{aligned} \Vert x + \lambda y \Vert _{a}^{2}&=\sup _{h\in {\mathcal {L}}_a}\varphi _h((x+\lambda y)^*a(x+\lambda y))\\&=\sup _{2h_{11}+h_{22}=1, h_{11}, h_{22} \ge 0} \big (3(1+|\lambda |^2) -2{\textrm{Re}}(\lambda )\big ) h_{22}\\&=3(1+|\lambda |^2) -2{\textrm{Re}}(\lambda ). \end{aligned}$$

However, for \( \lambda = \dfrac{1}{3} ,\) we see that \( \Vert x+ \lambda y \Vert _{a}^2 = \dfrac{8}{3} < 3 = \Vert x\Vert _{a}^2,\) which yields that \( x \not \perp _{BJ}^a y .\) On the other hand, it can easily be seen that \( \Vert x\Vert ^2 = 2 \) and \( \Vert x+ \lambda y \Vert ^2 = 2+2|\lambda |^2 \) for all \( \lambda \in {\mathbb {C}} .\) Hence \( \Vert x+ \lambda y \Vert ^2= 2(1+|\lambda |^2) \ge 2=\Vert x\Vert ,\) for all \( \lambda \in {\mathbb {C}} .\) Thus \( x \perp _{BJ} y .\)

Now, let \( x=\left[ \begin{array}{cccc} 0&{}\frac{1}{2} \\ 0&{}-1 \end{array}\right] \) and \( y=\left[ \begin{array}{cccc} 0 &{}1 \\ 0&{}1 \end{array}\right] .\) Then \( x \perp _{BJ}^a y ,\) since \( \langle x , y\rangle _a =x^* ay =0 .\) But, for every \( \lambda \in {\mathbb {C}} ,\) we have

$$\begin{aligned} \Vert x+\lambda y\Vert ^2=\left| \dfrac{5}{4}+2|\lambda |^2-{\textrm{Re}}(\lambda ) \right| . \end{aligned}$$

So for \(\lambda = \dfrac{1}{4},\) we have \(\Vert x+\lambda y\Vert ^2 = \dfrac{9}{8} \,<\dfrac{5}{4}=\Vert x\Vert ^2 ,\) and therefore \(x \not \perp _{BJ} y .\)

Assume that \( {\mathcal {A}} \) is a unital and commutative \( C^* \)-algebra, \( x,y \in {\mathcal {A}} \) and \(a\in {\mathcal {A}}\) is positive and invertible. If \( x \perp _{BJ} y ,\) then by Theorem 1.1, there must exist \(f\in {\mathcal {S}}({\mathcal {A}})\) such that \(f(x^*x)=\Vert x\Vert ^2\) and \( f(x^*y)=0.\) Since \({\mathcal {A}}\) is commutative, by Corollary 2.3 we conclude that

$$\begin{aligned} \varphi (x^* a x) = f \left( a^{-\frac{1}{2}} x^* a x a^{-\frac{1}{2}} \right) = f (x^* x) = \Vert x\Vert ^2= \Vert x\Vert _{a}^2 , \end{aligned}$$

and

$$\begin{aligned} \varphi (x^* a y) = f \left( a^{-\frac{1}{2}} x^* a y a^{-\frac{1}{2}} \right) = f (x^* y) =0, \end{aligned}$$

where \(\varphi \in {\mathcal {S}}_a({\mathcal {A}})\) is defined in (2.3). This fact motivates us to obtain a similar characterization for a-Birkhoff–James orthogonality in unital \(C^*\)-algebras. More precisely, we shall present a characterization of a-Birkhoff–James orthogonality in a unital \(C^*\)- algebra \({\mathcal {A}}\) based on the elements of its generalized state space \({\mathcal {S}}_a({\mathcal {A}}).\) In fact, we use a simple way to obtain the next fundamental result through the standard Gelfand–Naimark representation of \({\mathcal {A}}\) as a concrete \(C^*\)-subalgebra of \({\mathcal {B}}({\mathcal {H}})\) and displayed formula (2.2). However, for completion of the subject and the convenience of the reader, we present a short proof for it. Note that this characterization is a generalization of the well-known Theorem 1.1 when we take \(a=1_{{\mathcal {A}}},\) and plays a fundamental role to achieve our forthcoming main results.

Theorem 2.6

Let \( {\mathcal {A}} \) be a unital \( C^* \)-algebra,  \(x,y\in {\mathcal {A}}\) and let a be positive and invertible element of \({\mathcal {A}}.\) Then the following statements are equivalent : 

  1. (1)

    \( x \perp _{BJ}^{a} y .\)

  2. (2)

    There is \( \varphi \in {\mathcal {S}}_{a}({\mathcal {A}} ) \) such that \( \varphi (x^* a x) = \Vert x \Vert _{a}^{2} \) and \( \varphi (y^{*} a x )= 0 \) \((\varphi (x^*ay)=0).\)

Proof

\( (1) \Rightarrow (2)\) Let \( x \perp _{BJ}^{a} y \) and let \( \pi : {\mathcal {A}} \rightarrow {\mathcal {B}}({\mathcal {H}}) \) be a unital faithful \( * \)-representation of \( {\mathcal {A}} .\) Since a is invertible, it follows from (2.2) that \( \pi (x) , \pi (y) \in {\mathcal {B}}_{\pi (a)^{\frac{1}{2}}} ( {\mathcal {H}}) ,\) and so \(\pi (x)\perp ^{\pi (a)}_{BJ} \pi (y).\) Hence Theorem 2.1, concludes that there exists a sequence of \( \pi (a) \)-unit vectors \( \{ h_n\} \in {\mathcal {H}} \) such that

$$\begin{aligned}&\lim _{n \rightarrow \infty } \Vert \pi (x ) h_n \Vert _{\pi (a)} = \gamma _{\pi (a)} (\pi (x)), \end{aligned}$$
(2.5)
$$\begin{aligned}&\lim _{n \rightarrow \infty } \langle \pi (x) h_n , \pi (y) h_n \rangle _{\pi (a)} =0. \end{aligned}$$
(2.6)

The linear functionals \( \varphi _{n} : {\mathcal {A}} \rightarrow {\mathbb {C}} \) defined by \( \varphi _n (z) = \langle \pi (z ) h_n , h_n \rangle \) belong to \( {\mathcal {S}}_a ({\mathcal {A}}) \) for all \(n\in {\mathbb {N}}\) (see [1, Theorem 3.5]). Now, (2.2) and (2.5) imply that

$$\begin{aligned} \lim _{n \rightarrow \infty } \varphi _n ( x^* a x )&= \lim _{n \rightarrow \infty } \langle \pi (x^* a x) h_n , h_n \rangle = \lim _{n \rightarrow \infty } \langle \pi (a) \pi (x) h_n , \pi (x) h_n \rangle \\&=\lim _{n \rightarrow \infty } \Vert \pi (x ) (h_n) \Vert _{\pi (a)}^{2} = \gamma ^{2 }_{\pi (a)}(\pi (x)) = \Vert x \Vert _{a}^2 . \end{aligned}$$

In addition, from (2.6), we infer that

$$\begin{aligned} \lim _{n \rightarrow \infty }\varphi _n (y^* ax)&= \lim _{n \rightarrow \infty }\langle \pi (y^* a x) h_n , h_n \rangle =\lim _{n \rightarrow \infty }\langle \pi (a) \pi (x) h_n , \pi (y) h_n \rangle \\&=\lim _{n \rightarrow \infty }\langle \pi (x) h_n , \pi (y) h_n \rangle _{\pi (a)} =0 . \end{aligned}$$

Thus

$$\begin{aligned} \lim _{n \rightarrow \infty } \varphi _n (x^* a x) = \Vert x\Vert _{a}^{2},\quad \lim _{n \rightarrow \infty } \varphi _n (y^* a x ) =0. \end{aligned}$$
(2.7)

In addition, by Proposition 1.4, \( {\mathcal {S}}_a ({\mathcal {A}}) \) is \(w^* \)-compact. So there is \( \varphi \in {\mathcal {S}}_a ({\mathcal {A}}) \) such that \( \varphi _n \overset{w^*}{\longrightarrow }\ \varphi .\) Therefore, (2.7) implies that

$$\begin{aligned} \varphi ( x^* ax) = \Vert x \Vert _{a}^{2}\quad \text{ and } \quad \varphi ( y^* ax ) =0. \end{aligned}$$

\( (2) \Rightarrow (1 )\) Assume that there is \( \varphi \in {\mathcal {S}}_{a}({\mathcal {A}} ) \) such that \( \varphi (x^* a x) = \Vert x \Vert _{a}^{2} \) and \( \varphi (y^{*} a x )= 0 .\) Then for each \( \lambda \in {\mathbb {C}} ,\) we get

$$\begin{aligned} \Vert x + \lambda y \Vert _{a}^{2}&\ge \varphi ((x + \lambda y)^* a (x + \lambda y)) \\&= \varphi ( x^* a x) + 2 {\textrm{Re}} ( {\overline{\lambda }} \varphi (y^* a x)) + | \lambda |^2 \varphi ( y^* a y) \\&=\varphi ( x^* a x) + |\lambda |^2 \varphi ( y^* a y) \ge \varphi ( x^* ax) = \Vert x \Vert _{a}^{2}. \end{aligned}$$

Therefore \(x\perp _{BJ}^ay.\) \(\square \)

As the first direct consequence of Theorem 2.6, it is easy to see that for given linearly independent vectors \(x,y\in {\mathcal {A}},\) there exists a unique \(\alpha \in {\mathbb {C}}\) such that \(x\perp ^{a}_{BJ}(\alpha x+ y).\) Indeed, we take \(\alpha =0\) if \(x\perp ^{a}_{BJ}y.\) Now, suppose that \(x\not \perp ^a_{BJ}y.\) Since \(a \in {\mathcal {A}}\) is invertible, there exists \(\varphi \in {\mathcal {S}}_a({\mathcal {A}})\) such that \(\varphi (x^*ax)=\Vert x\Vert ^2_a,\) by Proposition 1.4. Furthermore, \(\varphi (x^*ay)\ne 0,\) by Theorem 2.6. Let \(\alpha =-\dfrac{\varphi (x^*ay)}{\varphi (x^*ax)}.\) Then

$$\begin{aligned} \varphi (x^*a(\alpha x+y))=-\dfrac{\varphi (x^*ay)}{\varphi (x^*ax)}\varphi (x^*ax)+\varphi (x^*ay)=0. \end{aligned}$$

Therefore \(x\perp ^a_{BJ}(\alpha x+y).\)

Further, the next result gives us some more examples of a-Birkhoff–James orthogonality for elements of \({\mathcal {A}}\) to some appropriate elements.

Corollary 2.7

Let \( {\mathcal {A}} \) be a unital \( C^* \)-algebra and let \( a \in {\mathcal {A}} \) be positive and invertible. For each \(x, y\in {\mathcal {A}},\) we have

$$\begin{aligned} x\perp _{BJ}^a \left( \Vert x\Vert _{a}^2 y a^{\frac{1}{2}} - y a^{-\frac{1}{2}} \langle x,x\rangle _a \right) . \end{aligned}$$

Proof

For the convenience, \(x^*ax\) and \(x^*ay\) are shown with the symbols \(\langle x,x \rangle _a\) and \(\langle x,y \rangle _a,\) respectively for all \(x,y\in {\mathcal {A}}.\) For \( x=0 \) or \( y=0 ,\) the statement is trivial. Now, assume that xy are nonzero elements of \({\mathcal {A}}.\) Since a is invertible, there exists \( \varphi \in {\mathcal {S}}_a ({\mathcal {A}}) \) such that \(\varphi (\langle x,x \rangle _a ) =\Vert x\Vert _{a}^2 \) by Proposition 1.4. The Cauchy–Schwarz inequality and (1.1) tell us

$$\begin{aligned}&\big | \varphi \big ( \big \langle x , \Vert x\Vert _{a}^2\, y\, a ^{\frac{1}{2}} - y \, a ^{-\frac{1}{2}} \langle x,x \rangle _a \big \rangle _a \big )\big |^2\\&\quad =\big | \varphi \big ( \Vert x\Vert _{a}^2\, \langle x,y\rangle _a \, a ^{\frac{1}{2}} -\langle x,y\rangle _a \, a ^{-\frac{1}{2}} \langle x,x \rangle _a \big )\big |^2 \\&\quad = | \varphi \big ( \langle x,y\rangle _a ( \Vert x\Vert _{a}^2 \, a ^{\frac{1}{2}} - a ^{-\frac{1}{2}}\, \langle x,x \rangle _a \big ) |^2 \\&\quad \le \, \varphi ( \langle x,y\rangle _a \langle y,x\rangle _a ) \, \varphi (\Vert x\Vert _{a}^4 \, a -2 \Vert x\Vert _{a}^2 \, \langle x,x\rangle _a + \langle x,x\rangle _a \, a^{-1} \langle x,x \rangle _a ) \\&\quad = \varphi ( \langle x,y\rangle _a \langle y,x\rangle _a ) \, \big ( \Vert x\Vert _{a}^4 \, \varphi (a) - 2 \Vert x\Vert _{a}^2 \, \varphi ( \langle x,x\rangle _a ) + \varphi ((x^\sharp x)^* a(x^\sharp x) )\big ) \\&\quad \le \varphi ( \langle x,y\rangle _a \langle y,x\rangle _a ) \, (\Vert x\Vert _{a}^4 - 2 \Vert x\Vert _{a}^4 + \Vert x^\sharp x \Vert _{a}^2) \\&\quad = \varphi ( \langle x,y\rangle _a \langle y,x\rangle _a ) (\Vert x\Vert _{a}^4 - 2 \Vert x\Vert _{a}^4 + \Vert x \Vert _{a}^4 ) =0 . \end{aligned}$$

It follows that \( \varphi (\langle x , \Vert x\Vert _{a}^2\, y\, a ^{\frac{1}{2}} - y \, a ^{-\frac{1}{2}} \langle x,x \rangle _a \rangle _a)=0.\) Consequently, Theorem 2.6 implies that

$$\begin{aligned} x\perp _{BJ}^a \left( \Vert x\Vert _{a}^2 y a^{\frac{1}{2}} - y a^{-\frac{1}{2}} \langle x,x\rangle _a \right) . \end{aligned}$$

\(\square \)

The a-algebraic numerical range of any element \(x\in {\mathcal {A}}\) is defined by

$$\begin{aligned} V_a(x)=\{\varphi (ax) : \varphi \in {\mathcal {S}}_a({\mathcal {A}})\}. \end{aligned}$$

It has been proved in [1, Theorem 4.7] that \(V_a(x)\) is a nonempty convex and compact subset of complex numbers for all \(x\in {\mathcal {A}}^a={\mathcal {A}},\) since a is invertible. An extension of the William’s Theorem [19, Theorem 1] is obtained in [2, Theorem 2.14].

The following direct result of Theorem 2.6 gives us an alternative proof for this fact.

Corollary 2.8

Let \(x\in {\mathcal {A}}.\) Then \(0\in V_a(x)\) if and only if \(\Vert x-\lambda 1_{{\mathcal {A}}}\Vert _a\ge |\lambda |\) for all \(\lambda \in {\mathbb {C}}.\)

Proof

Since \(0\in V_a(x),\) there is \(\varphi \in {\mathcal {S}}_a({\mathcal {A}})\) such that \(\varphi (1_{{\mathcal {A}}}ax)=\varphi (ax)=0.\) Also, we have \(\varphi (1^*_{{\mathcal {A}}}a1_{{\mathcal {A}}})=\varphi (a)=1=\Vert 1_{{\mathcal {A}}}\Vert ^2_a.\) It follows from Theorem 2.6 that \(1_{{\mathcal {A}}}\perp ^a_{BJ}x,\) which implies that \(\Vert x-\lambda 1_{{\mathcal {A}}}\Vert _a\ge |\lambda |\) for all \(\lambda \in {\mathbb {C}}.\)

Now, if \(\Vert \lambda 1_{{\mathcal {A}}}-x\Vert _a=\Vert x-\lambda 1_{{\mathcal {A}}}\Vert _a\ge |\lambda |\) for all \(\lambda \in {\mathbb {C}},\) then \(1_{{\mathcal {A}}}\perp ^a_{BJ}x,\) by the homogeneity of the Birkhoff–James orthogonality. So there is \(\varphi \in {\mathcal {S}}_a({\mathcal {A}})\) such that \(\varphi (ax)=0.\) Therefore \(0\in V_a(x).\) \(\square \)

Let \({\mathcal {A}}\) be a unital \(C^*\)-algebra and let \(x\in {\mathcal {A}}.\) Suppose that \({\mathcal {B}}\) is a subspace of \({\mathcal {A}}.\) An element \(y_0\in {\mathcal {B}}\) is said to be a best approximation to x in \({\mathcal {B}}\) if

$$\begin{aligned} \Vert x-y_0\Vert = {\textrm{dist}}(x,{\mathcal {B}}):=\inf \{\Vert x-y\Vert : y\in {\mathcal {B}}\} . \end{aligned}$$

The problem of finding characterizations of orthogonality of an element to subspace \({\mathcal {B}}\) is closely related to the best approximation problems. A specific question is when is the zero vector a best approximation to x from \({\mathcal {B}}?\) This is the same as asking when is x orthogonal to \({\mathcal {B}}?\) Due to the Theorem 1.1, it has been proved in [5, 9] that for any elements x and y of \(C^*\)-algebra \({\mathcal {A}},\) 0 is a best approximation to x in \({\mathcal {B}}={\mathbb {C}}y\) if and only if there exists \(f\in {\mathcal {S}}({\mathcal {A}})\) such that \(f(x^*x)=\Vert x\Vert ^2\) and \(f(x^*y)=0.\) Moreover, a generalized version of this fact has been proved in [13] for any element x and for any subspace \({\mathcal {B}}\) of \({\mathcal {A}}.\) As an application of Theorem 2.6, we present the following characterization of the best approximation for an element of \({\mathcal {A}}\) with respect to the norm \(\Vert \cdot \Vert _a.\) To achieve this goal, we need the following nice result from [2].

Theorem 2.9

[2, Theorem 2.13] Let \({\mathcal {A}}\) be a unital \(C^*\)-algebra and let a be a positive element of \({\mathcal {A}}.\) Let \(f:a{\mathcal {A}}^a\rightarrow {\mathbb {C}}\) be a linear functional such that \(f(a)=1\) and \(| f(az) | \,\le \Vert z\Vert _a\) for all \(z\in {\mathcal {A}}^a.\) Then there exists \(\varphi \in {\mathcal {S}}_a ({\mathcal {A}})\) such that \(\varphi (az)=f(az)\) for all \(z\in {\mathcal {A}}^a.\)

Theorem 2.10

Let \( {\mathcal {A}} \) be a unital \( C^* \)-algebra,  \( a \in {\mathcal {A}} \) be a positive and invertible element and let \( {\mathcal {B}} \) be a subspace of \( {\mathcal {A}}.\) Then \( y_0 \in {\mathcal {B}} \) is a best approximation to an element \( x \in {\mathcal {A}} \) with respect to \(\Vert \cdot \Vert _a\) if and only if there exists \( \varphi \in {\mathcal {S}}_{a} ({{\mathcal {A}}}) \) such that

$$\begin{aligned} \varphi ( ( x - y_0)^{*} a (x - y_0 ) )= \Vert x - y_0 \Vert _{a}^{2} \end{aligned}$$

and

$$\begin{aligned} \varphi (x^* a y) = \varphi ( y_{0}^* a y )\quad (\forall y\in {\mathcal {B}}). \end{aligned}$$

Proof

If \( {\mathcal {A}} \) is commutative, then the desired result immediately follows from [13, Theorem 1.1] and Corollary 2.3. Now, suppose that \( {\mathcal {A}} \) is a noncommutative \( C^* \)-algebra and \( y_0 \in {\mathcal {B}} \) is a best approximation to x with respect to \(\Vert \cdot \Vert _a.\) Since

$$\begin{aligned} {\textrm{dis}}(x - y_0 , {\mathcal {B}}) = \inf _{y^{\prime }\in {\mathcal {B}}} \Vert x - y_0 - y^{\prime } \Vert _a = \inf _{y^{\prime }\in {\mathcal {B}}} \Vert x - (y_0 + y^{\prime }) \Vert _a = {\textrm{dis}} (x , {\mathcal {B}}), \end{aligned}$$

without loss of generality, we may assume that \( y_0 = 0 .\)

Now, suppose that \(x\in {\mathcal {A}}\) and there exists \( \varphi \in {\mathcal {S}}_{a} ({\mathcal {A}}) \) such that \( \varphi (x^* a x) = \Vert x\Vert _{a}^{2} \) and \( \varphi (x^* a y) = 0 \) for all \( y \in {\mathcal {B}} .\) By Theorem 2.6 and homogeneity of a-Birkhoff–James orthogonality, we conclude that \( \Vert x - \lambda y \Vert _a \, \ge \Vert x \Vert _a \) for all \( y \in {\mathcal {B}} \) and all \( \lambda \in {\mathbb {C}} .\) Hence

$$\begin{aligned} \Vert x\Vert _a = \inf _{ y \in {\mathcal {B}}} \Vert x - y \Vert _a = {\textrm{dis}} ( x , {\mathcal {B}} ). \end{aligned}$$

Therefore \( y_0 =0 \) is a best approximation to x in \( {\mathcal {B}} .\)

Conversely, suppose that \( y_0=0 \) is a best approximation to x in \( {\mathcal {B}} .\) Then we have \( \Vert x\Vert _a \, \le \Vert x+ \lambda y\Vert _a \) for all \( y \in {\mathcal {B}} \) and all \( \lambda \in {\mathbb {C}} .\) Theorem 2.6 tells us for each \( y \in {\mathcal {B}} \) there exists \( \varphi _{y} \in {\mathcal {S}}_{a} ({\mathcal {A}}) \) such that \( \varphi _{y} (x^* a x) = \Vert x\Vert _{a}^{2} \) and \( \varphi _{y} ( x^* a y ) =0 .\)

Let \( {\mathcal {M}}=\{ \alpha x^* a x + \beta a + x^* a y \, : \, \alpha , \beta \in {\mathbb {C}}, \, y \in {\mathcal {B}} \} \) be a subspace of \({\mathcal {A}}\) generated by \( x^* a x,\) a and \( x^* a {\mathcal {B}} .\) Since a is invertible, it is known that x has a unique decomposition \( x = x_1 + i x_2 \) such that \( x_1 \) and \( x_2 \) are a-selfadjoint. In fact, \( x_1 = \dfrac{x+x^\sharp }{2}\) and \(x_2= \dfrac{x-x^\sharp }{2i}.\) Hence

$$\begin{aligned} {\mathcal {M}}&=\{ \alpha (x_1+i x_2)^* a (x_1+i x_2) + \beta a + (x_1+i x_2)^* a y \, : \, \alpha , \beta \in {\mathbb {C}}, \, y \in {\mathcal {B}} \} \\&= \{ \alpha (x_1 ^* a - i x_2^* a) (x_1+i x_2) + \beta a + (x_1 ^* a - i x_2^* a) y \, : \, \alpha , \beta \in {\mathbb {C}}, \, y \in {\mathcal {B}} \} \\&=\{ \alpha (a x_1 - i a x_2) (x_1+i x_2) + \beta a + (a x_1 - i a x_2) y \, : \, \alpha , \beta \in {\mathbb {C}}, \, y \in {\mathcal {B}} \} \\&=\{ a\left( \alpha ( x_1 - i x_2) (x_1+i x_2) + \beta 1_{{\mathcal {A}}} +(x_1 - i x_2) y\right) \, : \, \alpha , \beta \in {\mathbb {C}}, \, y \in {\mathcal {B}} \}. \end{aligned}$$

Define the mapping \( \psi : {\mathcal {M}} \rightarrow {\mathbb {C}} \) by

$$\begin{aligned} \psi (\alpha x^* a x + \beta a + x^* a y) = \alpha \Vert x\Vert _{a}^{2} + \beta . \end{aligned}$$

Clearly \( \psi \) is a linear mapping. To show that \( \psi \) is well defined, it is enough to prove that if \( \alpha x^* a x + \beta a + x^* a y = 0 ,\) then \( \psi (\alpha x^* a x + \beta a + x^* a y)=0 .\) Note that for each \(\alpha ,\beta \in {\mathbb {C}}\) and any \( y \in {\mathcal {B}} ,\) we have \( \varphi _y ( \alpha x^* a x + \beta a + x^* a y) = \alpha \Vert x\Vert _{a}^{2} +\beta ,\) since \( \varphi _y (a)=1 ,\) \( \varphi _y (x^* ax)=\Vert x\Vert _{a}^2 \) and \( \varphi _y ( x^* a y)=0 .\) Now, let \(u(\alpha ,\beta ,y)=\alpha (x_1 - i x_2)(x_1 + i x_2 ) + \beta 1_{{\mathcal {A}}} + (x_1 - i x_2) y\) for all \(\alpha ,\beta \in {\mathbb {C}}\) and all \(y\in {\mathcal {B}}.\) Then by the Cauchy-Schwartz inequality, we get

$$\begin{aligned}&| \psi (\alpha x^* a x + \beta a + x^* a y) | = |\, \alpha \, \Vert x\Vert _{a}^{2} + \beta \, | =| \varphi _{y} (\alpha x^*ax + \beta a + x^*a y) \nonumber |\\&\quad =| \varphi _{y} (a u(\alpha ,\beta ,y)) |=| \varphi _{y} (a^{\frac{1}{2}}a^{\frac{1}{2}} u(\alpha ,\beta ,y)) | \nonumber \\&\quad \le \sqrt{\varphi _{y}(a)} \sqrt{\varphi _{y} (u(\alpha ,\beta ,y)^* a \, u(\alpha ,\beta ,y)) } \nonumber \\&\quad =\sqrt{\varphi _{y} (u(\alpha ,\beta ,y)^* a \, u(\alpha ,\beta ,y)) } \nonumber \\&\quad \le \Vert u(\alpha ,\beta ,y)\Vert = \Vert \alpha (x_1 - i x_2)(x_1 + i x_2) + \beta 1_{{\mathcal {A}}} + (x_1 - i x_2)y \Vert _{a} . \end{aligned}$$
(2.8)

If \(\alpha x^* a x + \beta a + x^* a y=0 ,\) then \(a(\alpha (x_1 - i x_2)(x_1 + i x_2) + \beta 1_{{\mathcal {A}}} + (x_1 - i x_2)y) =0,\) and so \( \Vert \alpha (x_1 - i x_2)(x_1 + i x_2) + \beta 1_{{\mathcal {A}}} + (x_1 - i x_2)y\Vert _a =0 .\) Thus \( \psi (\alpha x^* a x + \beta a + x^* a y) = \psi (0) = 0 ,\) by (2.8).

Define \( {\mathcal {N}}:a{\mathcal {A}}\rightarrow [0,\infty )\) by \({\mathcal {N}} (az)= \Vert z \Vert _a\) for all \(z\in {\mathcal {A}}\) and note that \( {\mathcal {N}} \) is a norm on \(a{\mathcal {A}} .\) Moreover, (2.8) follows that

$$\begin{aligned} | \psi (\alpha x^* a x + \beta a + x^* a y ) |&\le {\mathcal {N}} (a (\alpha (x_1 - i x_2)(x_1 + i x_2) + \beta 1_{{\mathcal {A}}} + (x_1 - i x_2)y) )\\&= {\mathcal {N}} ( \alpha x^* a x + \beta a + x^* a y ). \end{aligned}$$

Hence \( \Vert \psi \Vert \, \le \, 1 \) with respect to the norm \( {\mathcal {N}} ,\) and therefore \( \psi :({\mathcal {M}},{\mathcal {N}}(\cdot ) ) \subseteq a {\mathcal {A}} \rightarrow {\mathbb {C}}\) is a bounded linear functional. The Hahn–Banach Theorem tells us \( \psi \) can be extend to a linear functional \( f : a {\mathcal {A}} \rightarrow {\mathbb {C}} \) such that \( \Vert f \Vert = \Vert \psi \Vert \, \le \, 1 ,\) \( f |_{({\mathcal {M}} , {\mathcal {N}}(\cdot ) )} = \psi \) and \( f (a) =1 .\) In addition,

$$\begin{aligned} |f(az)| \, \le \, \Vert f\Vert \,{\mathcal {N}}(a z ) \, \le \, {\mathcal {N}}(a z ) = \Vert z \Vert _a \quad ( \forall z \in {\mathcal {A}}). \end{aligned}$$

Taking the above considerations into account, by Theorem 2.9 one can find \( \varphi \in {\mathcal {S}}_{a} ({\mathcal {A}}) \) such that \( \varphi ( az) =f(az)\) for all \(z\in {\mathcal {A}}.\) Therefore, there exists \( \varphi \in {\mathcal {S}}_{a} ({\mathcal {A}}) \) such that

$$\begin{aligned} \varphi (x^* ax)&= \varphi ( (x_1 + i x_2 )^* a (x_1 + i x_2) ) = \varphi (a (x_{1} - i x_{2} ) (x_1 + i x_2) ) \\&= f (a (x_{1} - i x_{2} ) (x_1 + i x_2)) = f ((x_1 + i x_2 )^* a (x_1 + i x_2)) \\&= f(x^* ax) = \psi (x^*ax)=\Vert x\Vert ^2_a , \end{aligned}$$

and

$$\begin{aligned} \varphi (x^* ay)&=\varphi ((x_1 + i x_2 )^* ay)=\varphi (a(x_1 - i x_2 ) y)=f (a(x_1 - i x_2 ) y)\\&=f ((x_1 + i x_2 )^* ay ) =f(x^*ay)=\psi (x^*ay)=0\quad (\forall y\in {\mathcal {B}}). \end{aligned}$$

\(\square \)

As a direct consequence of Theorem 2.10, we get the following characterization of a-Birkhoff–James orthogonality to a subspace in a unital \(C^*\)-algebra.

Corollary 2.11

Let \({\mathcal {B}}\) be a subspace of a unital \(C^*\)-algebra \({\mathcal {A}}\) and let x be an element of \({\mathcal {A}}.\) Then x is a-Birkhoff–James orthogonal to \({\mathcal {B}}\) if and only if there is \(\varphi \in {\mathcal {S}}_{a} ({\mathcal {A}}) \) such that \(\varphi (x^*ax)=\Vert x\Vert ^2_a\) and \(\varphi (x^*ay)=0\) for all \(y\in {\mathcal {B}}.\)

The next result present a generalization of the well-known distance formula which obtained by Williams in [19].

Corollary 2.12

Let \( {\mathcal {A}} \) be a unital \( C^* \)-algebra,  \(a\in {\mathcal {A}}\) be a positive and invertible element and let \( x \in {\mathcal {A}} .\) Then

$$\begin{aligned} {\textrm{dist}}^{2} (x, {\mathbb {C}}1_{{\mathcal {A}}})=\min _{\lambda \in {\mathbb {C}}} \Vert x - \lambda 1_{{\mathcal {A}}} \Vert _{a}^2 = \max \{ \varphi ( x^* ax ) - | \varphi (ax) |^2 \, : \, \varphi \in {\mathcal {S}}_a ({\mathcal {A}}) \}. \end{aligned}$$

Proof

Let \( \alpha \in {\mathbb {C}} \) be such that \( \Vert x - \alpha 1_{{\mathcal {A}}} \Vert _{a} = {\textrm{dist}} ( x , {\mathbb {C}} 1_{{\mathcal {A}}} ) .\) For any \( \varphi \in {\mathcal {S}}_a ({\mathcal {A}}) \) such that \( \varphi (a x)=\alpha ,\) we have

$$\begin{aligned} \varphi ( x^* ax) - | \varphi ( ax)|^2 = \varphi ((x - \alpha 1_{{\mathcal {A}}})^* a (x - \alpha 1_{{\mathcal {A}}})) \le \Vert x - \alpha 1_{{\mathcal {A}}} \Vert _{a}^2 = {\textrm{dist}}^{2} (x, {\mathbb {C}}1_{{\mathcal {A}}}). \end{aligned}$$

Hence

$$\begin{aligned} \max \{ \varphi ( x^* ax) - | \varphi ( ax)|^2 \, : \, \varphi \in {\mathcal {S}}_a ({\mathcal {A}}) \} \le {\textrm{dist}}^{2} (x, {\mathbb {C}}1_{{\mathcal {A}}}) = \min _{\lambda \in {\mathbb {C}}} \Vert x - \lambda 1_{{\mathcal {A}}} \Vert _{a}^2 . \end{aligned}$$

On the other hand, by Theorem 2.10, there is \( \varphi \in {\mathcal {S}}_a ({\mathcal {A}}) \) such that

$$\begin{aligned} \varphi ((x- \alpha 1_{{\mathcal {A}}} )^* a (x- \alpha 1_{{\mathcal {A}}})) = \Vert x- \alpha 1_{{\mathcal {A}}} \Vert _{a}^2 \quad \text {and} \quad \varphi (ax) = \alpha . \end{aligned}$$

Therefore \( \varphi ( x^* ax) - | \varphi ( ax)|^2 = {\textrm{dist}}^{2} (x, {\mathbb {C}}1_{{\mathcal {A}}}) ,\) and so

$$\begin{aligned} \max \{\varphi ( x^* ax) - | \varphi ( ax)|^2 \, : \, \varphi \in {\mathcal {S}}_a ({\mathcal {A}}) \} \ge {\textrm{dist}}^{2} (x, {\mathbb {C}}1_{{\mathcal {A}}}). \end{aligned}$$

\(\square \)

3 Strong a-Birkhoff–James orthogonality in \(C^*\)-algebras

Our main goal in this section is to introduce and study the notion of strong Birkhoff–James orthogonality with respect to the norm \(\Vert \cdot \Vert _a\) in unital \( C^* \)-algebras. It should be noted that what is obtained in this section is an extension and modification of some results of [4, 6]. We start this section with introducing the concept of strong a-Birkhoff–James orthogonality.

Definition 3.1

Let \( {\mathcal {A}} \) be a unital \( C^* \)-algebra and \( a \in {\mathcal {A}} \) be a positive and invertible element. An element \( x \in {\mathcal {A}} \) is said to be strongly a-Birkhoff–James orthogonal to an element \( y \in {\mathcal {A}} ,\) in short \( x \perp _{S-BJ}^{a} y ,\) if

$$\begin{aligned} \Vert x + y b \Vert _{a} \ge \Vert x \Vert _{a}\quad (\forall b \in {\mathcal {A}}). \end{aligned}$$

Obviously, \( x \perp _{S-BJ}^{a} y \) implies \(x \perp _{BJ}^{a} y \) for all \(x,y\in {\mathcal {A}}.\) So for every \( x,y \in {\mathcal {A}} ,\) we obtain:

$$\begin{aligned} \langle x , y \rangle _a :=x^*ay=0 \, \Rightarrow \, x \perp _{S-BJ}^{a} y \, \Rightarrow \, x \perp _{BJ}^{a} y. \end{aligned}$$
(3.1)

Indeed, if \( \langle x, y \rangle _a =0 ,\) then for each \( b \in {\mathcal {A}} ,\) we have

$$\begin{aligned} \Vert x + yb \Vert _{a}^{2}&= \sup _{\varphi \in {\mathcal {S}}_a ({\mathcal {A}})} \varphi ((x+yb)^* a (x+yb)) \\&=\sup _{\varphi \in {\mathcal {S}}_a ({\mathcal {A}})} \big ( \varphi (\langle x,x\rangle _a ) + 2{\textrm{Re}}\varphi (\langle x,y\rangle _a b ) + \varphi (\langle yb,yb \rangle _a ) \big ) \\&=\sup _{\varphi \in {\mathcal {S}}_a ({\mathcal {A}})} \big ( \varphi (\langle x,x\rangle _a ) + \varphi (\langle yb , yb \rangle _a )\big ) \, \ge \, \sup _{\varphi \in {\mathcal {S}}_a ({\mathcal {A}})} \varphi ( \langle x,x\rangle _a )= \Vert x\Vert _{a}^{2} . \end{aligned}$$

Also, note that

$$\begin{aligned} x \perp _{S-BJ}^{a} y \, \Leftrightarrow \, x \perp _{BJ}^{a} yb \quad (\forall \, b \in {\mathcal {A}}). \end{aligned}$$
(3.2)

The converses in (3.1) do not hold in general. The following example explains this fact.

Example 3.2

Let \( a=\left[ \begin{array}{cccc} 2&{}0 \\ 0&{}1 \end{array}\right] .\) If \( x=I_2 \) and \( y=\left[ \begin{array}{cccc} 0&{}1 \\ 1&{}0 \end{array}\right] ,\) then for every \( \lambda \in {\mathbb {C}} ,\) we have

$$\begin{aligned} \Vert x + \lambda y \Vert _{a}^{2}&=\sup _{h\in {\mathcal {L}}_a}\varphi _h((x+\lambda y)^*a(x+\lambda y))\\&=\sup _{h\in {\mathcal {L}}_a} \big ( (2+|\lambda |^2 ) h_{11} + 2 \textrm{Re }((2{\overline{\lambda }}+\lambda ) h_{12}) +(2|\lambda |^2 +1) h_{22} \big ) \\&\ge 1+\dfrac{|\lambda |^2}{2}\ge 1=\Vert x\Vert ^2_a, \end{aligned}$$

since \( h_0=\left[ \begin{array}{cccc} \frac{1}{2} &{}0 \\ 0&{}0 \end{array}\right] \in {\mathcal {L}}_a .\) Hence \(x\perp ^a_{BJ}y.\) But, we may easily check that \( x \not \perp _{S-BJ}^a y .\) To this end, note that for \( b=\left[ \begin{array}{cccc} \frac{1}{2}&{}-\frac{1}{2} \\ -\frac{1}{2} &{}-\frac{1}{4} \end{array}\right] ,\) we get

$$\begin{aligned} \Vert x + yb \Vert _{a}^{2}&= \sup _{h \in {\mathcal {L}}_a } {\textrm{Tr}} (h (x +yb)^* a (x+yb)) \\&=\sup _{h \in {\mathcal {L}}_a } {\textrm{Tr}} \left( \left[ \begin{array}{cccc} h_{11} &{}h_{12} \\ \overline{h}_{12}&{}h_{22} \end{array}\right] \left[ \begin{array}{cccc} \frac{3}{4} &{} 0 \\ 0 &{} \frac{3}{8} \end{array}\right] \right) \\&=\sup _{2h_{11}+h_{22}=1, h_{11},h_{22} \ge 0}\left( \dfrac{3}{4}h_{11} +\dfrac{3}{8}h_{22} \right) = \dfrac{3}{8} <1= \Vert x\Vert _{a}^2. \end{aligned}$$

Now, let \( x=I_2 \) and \( y=\left[ \begin{array}{cccc} 0&{}0 \\ 1&{}0 \end{array}\right] .\) If \( b=\left[ \begin{array}{cccc} b_{11}&{}b_{12} \\ b_{21}&{}b_{22} \end{array}\right] \in {\mathbb {M}}_2({\mathbb {C}}) \) is arbitrary, then

$$\begin{aligned} \Vert x + yb \Vert _{a}^{2}&=\sup _{h\in {\mathcal {L}}_a}\varphi _h( (x +yb)^* a (x+yb))\\&=\sup _{h \in {\mathcal {L}}_a} \big ((2+|b_{11}|^2) h_{11} +2{\textrm{Re}}( {b}_{11}(1+\overline{b}_{12}) h_{12}) +|1+b_{12} |^2 h_{22} \big ) \\&\ge 1+\dfrac{|b_{11}|^2}{2}\ge 1=\Vert x\Vert ^2_a. \end{aligned}$$

So \( x \perp _{S-BJ}^{a} y ,\) while clearly, \( \langle x,y \rangle _a = \left[ \begin{array}{cccc} 0 &{} 0 \\ 1&{} 0 \end{array}\right] \ne 0 .\)

Our next result gives us a characterization of strong a-Birkhoff–James orthogonality based on elements of generalized state space \({\mathcal {S}}_a({\mathcal {A}})\) of unital \(C^*\)-algebra \({\mathcal {A}}.\) Actually, this result extend Theorem 2.5 of [4] for the norm \(\Vert \cdot \Vert _a\) on \({\mathcal {A}}.\)

Theorem 3.3

Let \( {\mathcal {A}} \) be a unital \( C^* \)-algebra,  \( x , y \in {\mathcal {A}} \) and let \( a \in {\mathcal {A}} \) be a positive and invertible element. Then the following statements are equivalent : 

  1. (1)

    \( x \perp _{S-BJ}^{a} y ;\)

  2. (2)

    \( x \perp _{BJ}^{a} y\langle y , x \rangle _a ;\)

  3. (3)

    There is \( \varphi \in {\mathcal {S}}_a ({\mathcal {A}}) \) such that \( \varphi ( x^* ax) = \Vert x\Vert _{a}^{2} \) and \( \varphi ( \langle x, y \rangle _a \langle y , x\rangle _a ) =0 ;\)

  4. (4)

    There is \( \varphi \in {\mathcal {S}}_a ({\mathcal {A}}) \) such that \( \varphi ( x^* ax) = \Vert x\Vert _{a}^{2} \) and \( \varphi ( \langle x, y \rangle _a b ) =0 ,\) for all \( b \in {\mathcal {A}} .\)

Proof

\( (1) \Rightarrow (2 )\) If \( x \perp _{S-BJ}^{a} y,\) then \(x\perp _{BJ}^ayb\) for all \( b \in {\mathcal {A}} ,\) by (3.2). Now, let \( b=\langle y , x \rangle _a .\) Then \( x \perp _{BJ}^{a} y\langle y , x \rangle _a .\)

\( (2) \Rightarrow (3)\) If \( x \perp _{BJ}^{a} y\langle y , x \rangle _a ,\) then it follows from Theorem 2.6 that there is \( \varphi \in {\mathcal {S}}_a ({\mathcal {A}}) \) such that \( \varphi ( x^* ax) = \Vert x\Vert _{a}^{2} \) and \(\varphi ( \langle x , y \rangle _a \langle y , x\rangle _a )= \varphi (\langle x , y \langle y , x\rangle _a \rangle _a ) = 0 .\)

\( (3) \Rightarrow (4)\) If there exists \( \varphi \in {\mathcal {S}}_a ({\mathcal {A}}) \) such that \( \varphi ( x^* ax) = \Vert x\Vert _{a}^{2} \) and \( \varphi ( \langle x, y \rangle _a \langle y , x\rangle _a ) =0 ,\) then by the Cauchy–Schwarz inequality, we have

$$\begin{aligned} | \varphi ( (x^* a y ) b ) |^2 \, \le \, \varphi ( (x^* a y )(y^* a x )) \, \varphi ( b^* b) = 0, \quad (\forall b \in {\mathcal {A}}), \end{aligned}$$

which follows that \( \varphi ( \langle x, y \rangle _a b ) =0 \) for all \( b \in {\mathcal {A}} .\)

\( (4) \Rightarrow (1 )\) It follows directly from Theorem 2.6 and the definition of strong a-Birkhoff–James orthogonality. \(\square \)

Proposition 3.4

Let \( {\mathcal {A}} \) be a unital \( C^* \)-algebra, \( x , y \in {\mathcal {A}} \) and let \( a \in {\mathcal {A}} \) be a positive and invertible element. If \( \langle x , y \rangle _a \, \ge \, 0 ,\) then

$$\begin{aligned} x \perp _{S-BJ}^{a} y\Leftrightarrow x \perp _{BJ}^{a} y. \end{aligned}$$

Proof

Assume that \( x \perp _{BJ}^{a} y .\) By Theorem 2.6, there exists \( \varphi \in {\mathcal {S}}_{a} ({\mathcal {A}}) \) such that \( \varphi ( \langle x ,x \rangle _a ) = \Vert x\Vert _{a}^{2} \) and \( \varphi ( \langle x , y \rangle _a )=0 .\) Since \( \langle x , y \rangle _a \, \ge \, 0 ,\) by the Cauchy–Schwarz inequality, for every \( b \in {\mathcal {A}} ,\) we get

$$\begin{aligned} | \varphi ( \langle x ,y \rangle _a b ) |^2&= | \varphi ( \langle x ,y \rangle _{a}^{\frac{1}{2}} \langle x ,y \rangle _{a}^{\frac{1}{2} } b ) |^2 \\&\le \varphi ( \langle x ,y \rangle _{a}^{\frac{1}{2}} \langle x ,y \rangle _{a}^{\frac{1}{2}} ) \, \varphi ( b^* \langle x ,y \rangle _{a}^{\frac{1}{2}} \langle x ,y \rangle _{a}^{\frac{1}{2}} b) \\&\le \varphi ( \langle x ,y \rangle _{a} )\, \varphi ( b ^* \langle x ,y \rangle _{a} b ) =0 . \end{aligned}$$

Thus \( \varphi ( \langle x ,y \rangle _ab ) =0 \) for all \(b\in {\mathcal {A}}.\) Therefore, Theorem 3.3 shows that \( x \perp _{S-BJ}^{a} y .\) \(\square \)

Theorem 3.5

Let \( {\mathcal {A}} \) be a unital \( C^* \)-algebra and let \( a \in {\mathcal {A}} \) be positive and invertible. If

$$\begin{aligned} x \perp _{S-BJ}^{a} y \Leftrightarrow x \perp _{BJ}^{a} y\quad (\forall x,y\in {\mathcal {A}}), \end{aligned}$$

then the \(C^*\)-algebra \( {\mathcal {A}} \) is commutative.

Proof

First, note that \({\mathcal {A}}^a={\mathcal {A}},\) since a is invertible. We shall show that for every \( x, b \in {\mathcal {A}} \) there is a scalar \( 0\ne \alpha \in {\mathbb {C}} \) such that

$$\begin{aligned} x b \perp _{S-BJ}^{a} (xb^2 + \alpha x b). \end{aligned}$$
(3.3)

If \( xb =0 ,\) obviously (3.3) holds. Now, let x be an element of \({\mathcal {A}}\) such that \( xb \ne 0 .\) Then \( xb \not \perp _{BJ}^{a} x .\) Indeed, if \( xb \perp _{BJ}^{a} x ,\) then \( xb \perp _{S-BJ}^{a} x ,\) by the assumption and thus \( xb \perp _{BJ}^{a} xb ,\) by (3.2). It follows that \( xb =0 ,\) which is not possible.

Moreover, by the definition of \(\Vert \cdot \Vert _a\) and invertibility of a,  there is \(\varphi \in {\mathcal {S}}_a({\mathcal {A}})\) such that \(\varphi ( \langle xb , xb\rangle _a)=\Vert xb\Vert ^2_a.\) Hence by Theorem 2.6, we conclude that \( \varphi (\langle xb , x \rangle _a) \ne 0 .\) Now, take \( \alpha = \dfrac{- \Vert xb \Vert _a}{\varphi ( \langle xb , x\rangle _a)} .\) Thus

$$\begin{aligned} \varphi ( \langle xb , xb+ \alpha x \rangle _a ) = \Vert xb\Vert _{a}^{2} - \dfrac{\Vert xb\Vert _{a}^{2}}{\varphi (\langle xb , x \rangle _a)}\varphi (\langle xb , x \rangle _a ) =0 . \end{aligned}$$

The assumption and the Theorem 2.6 yields that \( xb \perp _{S-BJ}^{a} (xb + \alpha x) .\) Hence \( xb \perp _{BJ}^{a} (xb^2 + \alpha xb) ,\) by (3.2), and so

$$\begin{aligned} x b \perp _{S-BJ}^{a} (xb^2 + \alpha x b), \end{aligned}$$

by the hypothesis.

If \({\mathcal {A}}\) is not commutative, there will a nonzero \(b\in {\mathcal {A}}\) with \(b^2=0\) (see [12], p. 68). By (3.3), for \(x=b^*\) there is a scalar \(\alpha \ne 0\) such that \( xb \perp _{S-BJ}^{a} \alpha xb .\) Hence \(b^*b=xb=0,\) and so \(b=0.\) This contradiction shows that \({\mathcal {A}}\) is commutative. \(\square \)

The next two results are direct consequences of Theorem 1.3, Corollary 2.3 and Theorem 3.5.

Corollary 3.6

Let \( {\mathcal {A}} \) be a unital \( C^* \)-algebra and let \(a \in {\mathcal {A}}\) be positive and invertible. The following statements are equivalent : 

  1. (1)

    For all \( x,y \in {\mathcal {A}} ,\) \( x \perp ^a_{BJ} y \) if and only if \( \langle x,y \rangle _a =0 ;\)

  2. (2)

    For all \( x,y \in {\mathcal {A}} ,\) \( x\perp ^a_{BJ} y \) if and only if \( x \perp ^a_{S-BJ} y ;\)

  3. (3)

    \( {\mathcal {A}} \) is isomorphic to \( {\mathbb {C}} .\)

Corollary 3.7

Let \( {\mathcal {A}} \) be a unital \( C^* \)-algebra and let \( a \in {\mathcal {A}} \) be positive and invertible. If

$$\begin{aligned} x \perp _{S-BJ}^{a} y \Leftrightarrow \langle x,y\rangle _a =0\quad (\forall x,y \in {\mathcal {A}}), \end{aligned}$$

then \({\mathcal {Z}} ({\mathcal {A}}) \cong {\mathbb {C}} 1_{{\mathcal {A}}} .\)

Let \({\mathcal {A}}\) be a unital \(C^*\)-algebra and \(a\in {\mathcal {A}}\) be positive and invertible. If \( z \in {\mathcal {A}} \) is a noninvertible element of \({\mathcal {A}},\) then \(zz^*a\) is not invertible, and so \(0\in \sigma (zz^*a)=\sigma _a(zz^*a)\subseteq V_a(zz^*a),\) by [15, Remark 2.13 and Corollary 3.9]. Hence there exists \( \varphi \in {\mathcal {S}}_a ({\mathcal {A}}) \) such that \(\varphi (\langle 1_{{\mathcal {A}}} , z \rangle _a \langle z , 1_{{\mathcal {A}}} \rangle _a )=\varphi (azz^*a)=0 .\) Also, we have \( \varphi (1_{{\mathcal {A}}}^* a 1_{{\mathcal {A}}}) = \varphi (a)=1= \Vert 1_{\mathcal {A}}\Vert _{a}^2 .\) Consequently, Theorem 3.3 implies that

$$\begin{aligned} 1_{{\mathcal {A}}} \perp _{S-BJ}^a z,\quad \text{ and } \text{ so }\quad 1_{{\mathcal {A}}} \perp _{BJ}^a z. \end{aligned}$$
(3.4)

It has been shown in [7] that the left-additivity (right-additivity) of the (strong) Birkhoff–James orthogonality on a unital \(C^*\)-algebra implies that \({\mathcal {A}}\) is isomorphic to \({\mathbb {C}}1_{{\mathcal {A}}}.\) As a final result of this section, we will prove that if the (strong) a-Birkhoff–James orthogonality is right-additive on a unital \(C^*\)-algebra \({\mathcal {A}},\) then the center of \({\mathcal {A}}\) is trivial; i.e., \( {\mathcal {Z}}({\mathcal {A}}) \cong {\mathbb {C}}1_{{\mathcal {A}}} .\)

Theorem 3.8

Let \({\mathcal {A}}\) be a unital \(C^*\)-algebra and let a be a positive and invertible element of \({\mathcal {A}}.\) If (strong) a-Birkhoff–James orthogonality is right-additive on \({\mathcal {A}},\) then \( {\mathcal {Z}}({\mathcal {A}}) \cong {\mathbb {C}}1_{{\mathcal {A}}} .\)

Proof

First, assume that \({\mathcal {A}}\) is commutative. Then \( {\mathcal {Z}}({\mathcal {A}})={\mathcal {A}}\cong {\mathbb {C}}1_{{\mathcal {A}}},\) by Corollary 2.3 and [7, Remark 2.8]. Now, suppose that \({\mathcal {A}}\) is noncommutative and \( x \in {{\mathcal {A}}} \) is a noninvertible element of \({\mathcal {A}}.\) Then by (3.4), \( 1_{{\mathcal {A}}} \perp ^a_{BJ} x^* x .\) If we assume that \( ax=xa ,\) then (2.4) follows that \(\Vert x^* x \Vert _{a}=\Vert x^* x \Vert .\) Hence \( \Vert x^* x \Vert _{a} 1_{{\mathcal {A}}} - x^* x \) is not invertible, since

$$\begin{aligned} \Vert x^* x \Vert _{a} = \Vert x^* x \Vert \in \sigma (x^* x) . \end{aligned}$$

Thus \( 1_{{\mathcal {A}}} \perp _{BJ}^a (\Vert x^* x \Vert _{a} 1_{{\mathcal {A}}} - x^* x) .\) The right-additivity of a-Birkhoff–James orthogonality follows that \( 1_{{\mathcal {A}}} \perp _{BJ}^a \Vert x^* x \Vert _{a} 1_{{\mathcal {A}}} .\) So \( \Vert x^* x \Vert _{a} = \Vert x \Vert _{a}^{2} =0 ,\) because of the non-degeneracy of a-Birkhoff–James orthogonality. Hence \( x=0 .\) Therefore we have proved that every nonzero element of \(C^*\)-subalgebra \({\mathcal {Z}}({\mathcal {A}}) \) is invertible, and so \( {\mathcal {Z}}({\mathcal {A}}) \cong {\mathbb {C}}1_{{\mathcal {A}}} \) by the Gelfand–Mazur Theorem. A similar argument works for strong a-Birkhoff–James orthogonality. \(\square \)

Remark 3.9

Suppose that a-Birkhoff–James orthogonality is left-additive in unital \(C^*\)-algebra \({\mathcal {A}}\) and let \(x\in {\mathcal {A}}\) be positive and noninvertible such that \(xa=ax.\) Then the \(C^*\)-subalgebra, \({\mathcal {B}}:=C^*(1_{{\mathcal {A}}},a,x),\) generated by \(1_{{\mathcal {A}}},\) a and x is commutative. According to the Corollary 2.3, Birkhoff–James orthogonality is left-additive on \({\mathcal {B}}.\) Hence \(x=0,\) by [7, Remark 2.8]. It follows that every nonzero element of \({\mathcal {Z}}({\mathcal {A}})\) is invertible, and so \({\mathcal {Z}}({\mathcal {A}})\) is trivial. It should be noted that the same proof works for right-additivity of a-Birkhoff–James orthogonality. However, a different approach is presented to study right-additivity in the previous Theorem.