A conjugation C defined on a complex Hilbert space \(\mathcal {H}\) is an antilinear operator that is involutive \((C^{2} = I_{\mathcal{H}})\) and isometric, meaning that the following equality holds;
$$\begin{aligned} \langle C\xi, C\eta \rangle =\langle \eta, \xi \rangle\quad \text{for all $\xi, \eta \in \mathcal{H}$}. \end{aligned}$$
(3)
Thus it follows from (3) that \(\langle C\xi, C\eta \rangle =\overline{\langle \xi, \eta \rangle}\). Chō and Tanahashi [2] introduced a conjugation C on a complex Banach space \(\mathcal{B}\) as the operator satisfying the following relations;
$$\begin{aligned} C^{2}=I_{\mathcal{B}},\qquad \Vert C \Vert \leq 1,\qquad C(x+y)=Cx+Cy\quad \text{and}\quad C(\lambda x)=\overline{\lambda}Cx \end{aligned}$$
(4)
for all \(x, y \in \mathcal{B}\) and \(\lambda \in \mathbb{C}\).
Like in a Hilbert space, we will define a conjugation on a semi-inner-product space using a semi-inner-product. Throughout this section, \(\mathcal{X}\) denotes a semi-inner-product space with a semi-inner-product \([\cdot,\cdot ]\), unless specified otherwise.
Definition 3.1
An operator \(C: \mathcal{X} \rightarrow \mathcal{X}\) is a conjugation if it is involutive \((C^{2} = I_{\mathcal{X}})\) and
$$\begin{aligned}{} [ Cx, Cy ]=\overline{[ x, y ]}\quad \text{for all $x, y \in \mathcal{X}$}. \end{aligned}$$
(5)
Proposition 3.2
If C is a conjugation on \(\mathcal{X}\), then relation (4) holds for all \(x, y \in \mathcal{X}\) and \(\lambda \in \mathbb{C}\).
Proof
By the Cauchy–Schwarz inequality for a semi-inner-product, we have that \(\|Cx\|^{2}=[Cx, Cx]=\overline{[x, x]}\leq \|x\|^{2}\) for every \(x\in \mathcal{X}\), which implies that \(\|C\|\le 1\). Since a semi-inner-product is linear in the first variable, we have
$$\begin{aligned} \bigl[ C(x+y), Cz \bigr]&= \overline{[ x+y, z ]} = \overline{[ x, z ]}+ \overline{[ y, z ]} \\ &= [ Cx, Cz ]+[ Cy, Cz ] = [ Cx+Cy, Cz ] \end{aligned}$$
for all \(x, y, z \in \mathcal{X}\). Since the operator C is surjective, we can take \(z \in \mathcal{X}\) such that
$$\begin{aligned} Cz:= C(x+y)-Cx-Cy. \end{aligned}$$
Then we get that \(0=[ C(x+y)-Cx-Cy, Cz ]=[ Cz, Cz]\), so that \(Cz=0\), that is, \(C(x+y)=Cx+Cy\). To show that \(C(\lambda x)=\overline{\lambda}Cx\) for any \(x \in \mathcal{X}\) and \(\lambda \in \mathbb{C}\), take any element \(y \in \mathcal{X}\). Then we have
$$\begin{aligned} \bigl[ C(\lambda x), y \bigr]&= \overline{[ \lambda x, Cy]} = \overline{\lambda} \overline{[ x, Cy ]} \\ &= \overline{\lambda}[ Cx, y ] = [ \overline{\lambda}Cx, y ], \end{aligned}$$
which means that \(C(\lambda x)=\overline{\lambda}Cx\). Therefore C satisfies relation (4). □
Let C be a conjugation on a complex Hilbert space \(\mathcal{H}\). A bounded linear operator T on \(\mathcal{H}\) is C-symmetric if \(T= CT^{\ast}C\), where \(T^{\ast}\) is a Hilbert space adjoint of T, which is equivalent to \(\langle Tx, y \rangle =\langle x, CTCy \rangle \) for all \(x,y \in \mathcal{H}\). Chō et al. [1] have extend the notion of C-symmetric operators to Banach space operators via linear functionals in its dual space. However, we would like to extend the notion of the complex symmetry to semi-inner-product space operators without using linear functionals. Even though semi-inner-products in general are not additive in the second variables, we will use a semi-inner-product to define the C-symmetric operator on a semi-inner-product space.
Definition 3.3
Let C be a conjugation on \(\mathcal{X}\). We say that \(T\in \mathcal{L}(\mathcal{X})\) is C-symmetric if
$$\begin{aligned}{} [ Tx, y ]=[ x, CTCy]\quad \text{for all $x,y \in \mathcal{X}$}. \end{aligned}$$
(6)
Remark 3.4
In Definition 3.3, equation (6) is equivalent to
$$\begin{aligned}{} [ x, Ty ]=[ CTCx, y] \quad\text{for all $x,y \in \mathcal{X}$}. \end{aligned}$$
(7)
Indeed, by putting \(Cx, Cy\) into (6) instead of \(x, y\) we obtain that \([ TCx, Cy ]=[ Cx, CTy]\). It follows from the definition of a conjugation C that \([ CTCx, y ]=\overline{[ Cx, CTy]} = [ x, Ty]\).
Proposition 3.5
Let C be a conjugation on \(\mathcal{X}\), and let \(T \in \mathcal{L}(\mathcal{X})\) be a C-symmetric operator.
-
(i)
λT is C-symmetric for any complex number λ.
-
(ii)
If T is invertible, then \(T^{-1}\) is also C-symmetric.
-
(iii)
If \(S\in \mathcal{L}(\mathcal{X})\) is C-symmetric and commutes with T, then so is TS.
Proof
(i) For any complex number λ, we have
$$\begin{aligned} \bigl[(\lambda T)x, y\bigr]&=\lambda [Tx, y]=\lambda [x, CTCy] \\ &=[x, \overline{\lambda}CTCy]=\bigl[x, C(\lambda T)Cy\bigr], \end{aligned}$$
so that λT is C-symmetric.
(ii) For any \(y \in \mathcal{X}\), there exists \(z \in \mathcal{X}\) such that \(y=CTCz\). Indeed, since T is invertible and C is a conjugation, \(CT^{-1}C\) is also invertible. Putting \(z:=CT^{-1}Cy\), we get \(y=CTCz\). For any \(x,y \in \mathcal{X}\), we have
$$\begin{aligned} \bigl[T^{-1}x, y\bigr] =\bigl[T^{-1}x, CTCz\bigr]= \bigl[TT^{-1}x, z\bigr]=[x, z]=\bigl[x, CT^{-1}Cy\bigr], \end{aligned}$$
where the second equality follows from the C-symmetry of T. Thus \(T^{-1}\) is C-symmetric, which completes the proof.
(iii) If \(S\in \mathcal{L}(\mathcal{X})\) commutes with T and is C-symmetric, then it follows that
$$\begin{aligned} \bigl[(TS)x, y\bigr]=[Sx, CTCy]=\bigl[x,CSC (CTCy)\bigr]=\bigl[x,C(ST)Cy\bigr]= \bigl[x, C(TS)Cy\bigr]. \end{aligned}$$
Hence TS is C-symmetric. □
Let \(T \in \mathcal{L}(\mathcal{X})\) and \(y \in \mathcal{X}\). By the Riesz representation theorem in a semi-inner-product space [7], there is a unique vector \(T^{\dagger }y\) such that \([Tx, y] = [x, T^{\dagger }y]\) for all \(x \in \mathcal{X,}\) where \(T^{\dagger}\) is a generalized adjoint, which is not usually linear [11]. On the other hand, if C is a conjugation on \(\mathcal{X}\) and if \(T\in \mathcal{L}(\mathcal{X})\) is C-symmetric, then we obtain that
$$\begin{aligned} \bigl[x, T^{\dagger }y\bigr]=[Tx, y]=[x, CTCy] \quad\text{for all $x,y\in \mathcal{X}$}, \end{aligned}$$
so that \(T^{\dagger}=CTC\). Since \(CTC\) is linear, \(T^{\dagger}\) becomes a linear operator on \(\mathcal{X}\). It follows from (7) that \([x, Ty]=[CTCx, y]=[T^{\dagger }x, y]\) for all \(x,y\in \mathcal{X}\). Furthermore, \(T^{\dagger}\) is also C-symmetric. Indeed, for all \(x,y \in \mathcal{X}\),
$$\begin{aligned} \bigl[T^{\dagger }x, y\bigr]=[CTCx, y]=\overline{[TCx, Cy]}= \overline{ \bigl[Cx, T^{\dagger }Cy\bigr]}=\bigl[x, CT^{\dagger }Cy\bigr]. \end{aligned}$$
A uniform semi-inner-product space means a uniformly continuous semi-inner-product space where the induced normed vector space is complete and uniformly convex. Here the (uniform) continuity implies that
$$\begin{aligned} \operatorname{Re}\bigl\{ [y,x+ty]\bigr\} \to \operatorname{Re}\bigl\{ [y,x]\bigr\} \quad\text{(uniformly) as $t \in \mathbb{R} \to 0$}. \end{aligned}$$
Giles [7, Theorem 7] proved that for a uniform semi-inner-product space \(\mathcal{X}\), the dual space \(\mathcal{X}^{\star}\) is also a uniform complex semi-inner-product space with respect to the semi-inner-product defined by \([x^{\star}, y^{\star}]_{\star}=[y, x]\). Moreover, he proved that for every continuous linear functional \(x^{\star}\) in a dual space \(\mathcal{X}^{\star}\), there exists a unique vector \(x\in \mathcal{X}\) such that
$$\begin{aligned} x^{\star}(z)=[z, x] \quad\text{for all } z\in \mathcal{X}, \end{aligned}$$
so that the map \(x \mapsto x^{\star}=[\cdot, x]\) is a one-to-one mapping from \(\mathcal{X}\) onto \(\mathcal{X}^{\star}\). For any \(T\in \mathcal{L}(\mathcal{X})\), the dual operator \(T^{\star}\in \mathcal{L}(\mathcal{X}^{\star})\) is given by \(T^{\star }y^{\star}(z)=y^{\star}(Tz)\) for all \(y^{\star }\in \mathcal{X}^{\star}\) and \(z \in \mathcal{X}\).
If C is a conjugation on a uniform semi-inner-product space \(\mathcal{X}\), then we define the dual operator \(C^{\star}:\mathcal{X^{\star}} \to \mathcal{X^{\star}}\) by
$$\begin{aligned} \bigl(C^{\star}\bigl(x^{\star}\bigr)\bigr) (z):= \overline{x^{\star}(Cz)} \quad\text{for all $z\in \mathcal{X}$}. \end{aligned}$$
(8)
We have that \((C^{\star}(x^{\star}))(z)=\overline{x^{\star}(Cz)}= \overline{[Cz, x]}=[z, Cx]=(Cx)^{\star}(z)\), so that \(C^{\star}(x^{\star})=(Cx)^{\star}\). Thus we have the following commutative diagram:
Moreover, the dual operator \(C^{\star}\) is a conjugation on \(\mathcal{X^{\star}}\). Indeed, for any \(x^{\star}, y^{\star} \in \mathcal{X}^{\star}\), there exist unique vectors x and y in \(\mathcal{X}\) such that
$$\begin{aligned} \bigl[C^{\star}x^{\star}, C^{\star}y^{\star} \bigr]_{\star}=\bigl[(Cx)^{\star},(Cy)^{ \star} \bigr]_{\star} =[Cy, Cx]=\overline{[y, x]}= \overline{\bigl[x^{\star}, y^{\star}\bigr]_{\star}}. \end{aligned}$$
Since \((Cx)^{\star}=C^{\star}(x^{\star})\) for all \(x\in \mathcal{X}\), we observe that relation (4) implies equation (5).
Proposition 3.6
Let C be a conjugation on a uniform semi-inner-product space \(\mathcal{X}\).
-
(i)
If \(T\in \mathcal{L}(\mathcal{X})\) is C-symmetric, then \(T^{\star}\in \mathcal{L}(\mathcal{X}^{\star})\) is also \(C^{\star}\)-symmetric.
-
(ii)
If \(T\in \mathcal{L}(\mathcal{X})\) is C-symmetric, then \((T^{\dagger})^{\star}=(T^{\star})^{\dagger}\).
-
(iii)
If \(\{T_{n} \}\) is a sequence of C-symmetric operators such that \(T_{n} \to S\) in the strong topology, then S is C-symmetric.
Proof
(i) Suppose that T is a C-symmetric operator on \(\mathcal{X}\). Let f and g be arbitrary elements in the dual space \(\mathcal{X}^{\star}\). Since \(\mathcal{X}\) is a uniform semi-inner-product space, there exist unique vectors x and y in \(\mathcal{X}\) such that \(x^{\star}=f\) and \(y^{\star}=g\). First, we observe that \(T^{\star}x^{\star}=(T^{\dagger}x)^{\star}\). Indeed, for any \(z\in \mathcal{X}\), we have
$$\begin{aligned} \bigl(T^{\star}x^{\star}\bigr) (z)=x^{\star}(Tz)=[Tz, x]= \bigl[z, T^{\dagger}x\bigr]=\bigl(T^{\dagger}x\bigr)^{ \star}(z). \end{aligned}$$
Moreover, for any \(z \in \mathcal{X}\) and \(y^{\star} \in \mathcal{X}^{\star}\), we see that
$$\begin{aligned} \bigl(C^{\star}T^{\star}C^{\star}\bigr)y^{\star}(z)=y^{\star}(CTCz)=(CTC)^{ \star}y^{\star}(z), \end{aligned}$$
so \(C^{\star}T^{\star}C^{\star} = (CTC)^{\star}\). Thus we have
$$\begin{aligned} \bigl[T^{\star}x^{\star}, y^{\star}\bigr]_{\star}&= \bigl[y, T^{\dagger}x\bigr]=\bigl[CT^{\dagger}Cy, x\bigr] \\ &=\bigl[x^{\star}, \bigl(CT^{\dagger}Cy\bigr)^{\star} \bigr]_{\star}=\bigl[x^{\star}, (CTC)^{\star}y^{ \star} \bigr]_{\star} \\ &=\bigl[x^{\star}, C^{\star}T^{\star}C^{\star}y^{\star} \bigr]_{\star}, \end{aligned}$$
which means that \(T^{\star}\) is \(C^{\star}\)-symmetric.
(ii) For any \(z \in \mathcal{X}\) and \(y^{\star }\in \mathcal{X}^{\star}\), we obtain that
$$\begin{aligned} \bigl(T^{\dagger}\bigr)^{\star }y^{\star}(z)=(CTC)^{\star }y^{\star}(z)= \bigl(C^{\star }T^{ \star }C^{\star}\bigr)y^{\star }(z). \end{aligned}$$
On the other hand, it follows from (i) that \(T^{\star}\) is \(C^{\star}\)-symmetric. Hence, for all \(x^{\star }\in \mathcal{X}^{\star}\),
$$\begin{aligned} \bigl[x^{\star}, \bigl(T^{\star}\bigr)^{\dagger}y^{\star} \bigr]_{\star}=\bigl[T^{\star}x^{\star}, y^{ \star} \bigr]_{\star}=\bigl[x^{\star}, \bigl(C^{\star }T^{\star }C^{\star} \bigr)y^{\star}\bigr]_{ \star}. \end{aligned}$$
This means that \((T^{\dagger})^{\star}=(T^{\star})^{\dagger}\).
(iii) Since \(\|(S - T_{n})x\| \to 0\) for all \(x\in \mathcal{X}\), for all \(x,y\in \mathcal{X}\), we have
$$\begin{aligned}{} [CSCx, y] = \lim_{n\to \infty}[CT_{n}Cx, y] = \lim _{n\to \infty}[x, T_{n}y] = [x, Sy], \end{aligned}$$
where the third equality follows from uniform continuity. Thus S is a C-symmetric operator. □
Now we compute the numerical range of a conjugation on \(\ell _{n}^{p}(\mathbb{C})\).
Example 3.7
Let C be a complex conjugation on \(\ell _{n}^{p}(\mathbb{C})\) \((1\leq p<\infty )\) given by \(Cx=\overline{x}=(\overline{x}_{1},\ldots,\overline{x}_{n})\) for \(x\in \ell _{n}^{p}(\mathbb{C})\). Then we have:
-
(1)
\(W(C)=\{\lambda \in \mathbb{C}: |\lambda |=1 \}\) for \(n=1\),
-
(2)
\(W(C)=\{\lambda \in \mathbb{C}: |\lambda |\leq 1 \}\) for \(n\geq 2\).
It is easy to prove (1). Indeed, for any \(x\in \ell _{1}^{p}(\mathbb{C})\) with \(|x|=1\), we write \(x=e^{i\theta}\) for some real number θ. Obviously, we have \([Cx, x]_{p}=[\overline{x}, x]_{p}=e^{-2i\theta}\), and so \(W(C)=\{\lambda \in \mathbb{C}: |\lambda |=1 \}\).
To show the second statement, let \(x\in \ell _{n}^{p}(\mathbb{C})\) be any unit vector, i.e., \(\|x\|_{p}^{2} = [x, x]_{p}=1\). By the Cauchy–Schwarz inequality we have
$$\begin{aligned} \bigl\vert [Cx, x]_{p} \bigr\vert ^{2}\leq [Cx, Cx]_{p} [x, x]_{p}= \overline{[x, x]_{p}} [x, x]_{p}=1, \end{aligned}$$
which implies that \(W(C)\subseteq \{\lambda \in \mathbb{C}: |\lambda |\leq 1 \}\).
For the reverse inclusion, let λ be any complex number with \(|\lambda |\leq 1\). We write a polar form \(\lambda =|\lambda |e^{i\theta}\) for some real number θ. Now we take a unit vector \(x\in \ell _{n}^{p}(\mathbb{C})\) given by
$$\begin{aligned} x= \biggl( {\biggl(}\frac{1+ \vert \lambda \vert }{2} {\biggr)}^{\frac{1}{p}}e^{- \frac{i\theta}{2}}, {\biggl(}\frac{1- \vert \lambda \vert }{2} {\biggr)}^{ \frac{1}{p}}ie^{-\frac{i\theta}{2}}, 0, \ldots, 0 \biggr). \end{aligned}$$
Then we have
$$\begin{aligned}{} [Cx, x]_{p}=[\overline{x}, x]_{p}= \biggl( \frac{1+ \vert \lambda \vert }{2}- \frac{1- \vert \lambda \vert }{2} \biggr) e^{i\theta} = \vert \lambda \vert e^{i\theta}= \lambda, \end{aligned}$$
which implies that \(W(C)\) contains the closed unit disc. Therefore the numerical range \(W(C)\) is the closed unit disc.
Let C be the usual complex conjugation given in Example 3.7. Then we see that
$$\begin{aligned} w(C)=\sup \bigl\{ \bigl\vert [Cx, x]_{p} \bigr\vert : [x, x] =1, x \in \ell _{n}^{p} \bigr\} = 1 \quad\text{for all $n \ge 1$}, \end{aligned}$$
where \(w(C)\) is the numerical radius of C. Moreover, we can find infinitely many unit vectors x that attain the numerical radius of the complex conjugation C on \(\ell _{n}^{1}(\mathbb{C})\) \((n \geq 1)\), that is, vectors x with \(|[Cx,x]_{1}|=1\). We explicitly construct vectors attaining the numerical radius \(w(C)\) in the following example.
Example 3.8
Let \(n \geq 2\). For any \(\lambda \in \mathbb{C}\) with \(0<|\lambda |\leq \frac{1}{n-1}\), we take the vector \(x=(x_{1}, x_{2}, \ldots, x_{n})^{t} \in \ell _{n}^{1}(\mathbb{C})\) given by
$$\begin{aligned} x_{j}= \textstyle\begin{cases} \overline{\lambda} & \text{if } 1\leq j \leq n-1, \\ ( \frac {1}{ \vert \lambda \vert }-n+1 ) \overline{\lambda} & \text{if } j=n. \end{cases}\displaystyle \end{aligned}$$
(9)
Then we have that
$$\begin{aligned}{} [x, x]_{1}=1\quad \text{and}\quad \bigl\vert [Cx, x]_{1} \bigr\vert = \biggl\vert \biggl(\frac{\lambda}{ \vert \lambda \vert } \biggr) \biggr\vert ^{2}=1. \end{aligned}$$
For any complex number λ with \(|\lambda |\geq n-1\), we put \(x=(x_{1}, x_{2}, \ldots, x_{n})\in \ell _{n}^{1}(\mathbb{C})\), where
$$\begin{aligned} x_{j}= \textstyle\begin{cases} \overline{\lambda}^{-1} & \text{if } 1\leq j \leq n-1, \\ \overline{\lambda}^{-1}( \vert \lambda \vert -n+1) & \text{if } j=n. \end{cases}\displaystyle \end{aligned}$$
(10)
Then it follows that
$$\begin{aligned}{} [x, x]_{1}=1 \quad\text{and}\quad \bigl\vert [Cx, x]_{1} \bigr\vert = \biggl\vert \biggl( \frac{ \vert \lambda \vert }{\lambda} \biggr) \biggr\vert ^{2}=1. \end{aligned}$$
Similarly, we also have infinitely many numerical radius attaining vectors in the infinite-dimensional space \(\ell ^{1}(\mathbb{N})\) in the same way as (9) and (10) except for jth terms with 0 \((j>n)\).
The essential numerical range for a bounded linear operator on a Hilbert space is defined as the closure of the numerical range of the image in the Calkin algebra, and many equivalent conditions are known [5]. We now introduce the sequentially essential numerical range of T on a semi-inner-product space \(\mathcal{X}\) by
$$\begin{aligned} W_{e}(T)=\Bigl\{ z\in \mathbb{C}: \lim_{n} [Tx_{n}, x_{n}]=z \text{ for some } \{x_{n}\}\subset \mathcal{X} \text{ with } [x_{n},x_{n}]=1, x_{n} \stackrel{w}{\longrightarrow} 0 \Bigr\} . \end{aligned}$$
Theorem 3.9
Let \(T\in \mathcal{L}(\mathcal{X})\), and let C be a conjugation on \(\mathcal{X}\). Then we have
$$\begin{aligned} W(CTC)=\overline{W(T)}\quad \textit{and}\quad W_{e}(T)= \overline{W_{e}(CTC)}, \end{aligned}$$
where S̅ denotes the complex conjugation of S.
Proof
If \(z\in W(CTC)\), then there exists a vector \(x\in \mathcal{X}\) with \([x, x]=1\) such that
$$\begin{aligned} z=[CTCx, x]=\overline{[TCx, Cx]}\in \overline{W(T)}. \end{aligned}$$
This means that \(W(CTC)\subset \overline{W(T)}\). Since \(W(T)=W(C^{2}TC^{2})\subset \overline{W(CTC)}\), we get the reverse inclusion. Therefore we have \(W(CTC)=\overline{W(T)}\).
If \(z\in W_{e}(CTC)\), then there exists a sequence \(\{x_{n}\}\subset \mathcal{X}\) with \([x_{n},x_{n}]=1\) and \(x_{n} \stackrel{w}{\longrightarrow} 0\). Since \(\lim_{n} x_{n}=0\) in the weak sense, we obtain that \(\lim_{n} f(x_{n})=0\) for all \(f\in \mathcal{X}^{\star}\). Since \(C^{\star}f \in \mathcal{X}^{\star}\) for all \(f\in \mathcal{X}^{\star}\), we have \(\lim_{n} f(Cx_{n})=\lim_{n} C^{\star}f(x_{n})=0\), which implies that \(Cx_{n}\stackrel{w}{\longrightarrow}0\). Thus we have
$$\begin{aligned} z=\lim_{n} [CTCx_{n}, x_{n}]=\lim _{n} \overline{[TCx_{n}, Cx_{n}]}\in \overline{W_{e}(T)}. \end{aligned}$$
This implies that \(W_{e}(CTC)\subset \overline{W_{e}(T)}\). The reverse inclusion follows from
$$\begin{aligned} W_{e}(T)=W_{e}\bigl(C^{2}TC^{2}\bigr) \subset \overline{W_{e}(CTC)}, \end{aligned}$$
which completes the proof. □
Corollary 3.10
Let C be a conjugation on \(\mathcal{X}\), and let \(T\in \mathcal{L}(\mathcal{X})\) be C-symmetric.
-
(i)
\(W(T^{\dagger})=\overline{W(T)}\) and \(W_{e}(T^{\dagger})=\overline{W_{e}(T)}\).
-
(ii)
\(W(T)=\{\overline{[x, Tx]}: [x, x]=1, x\in \mathcal{X} \}\).
-
(iii)
If, in addition, \(\mathcal{X}\) is a uniform semi-inner-product space, then
$$\begin{aligned} W\bigl(C^{\star}T^{\star}C^{\star}\bigr)=\overline{W \bigl(T^{\star}\bigr)} =\bigl\{ \bigl[x^{\star}, T^{\star}x^{\star}\bigr]_{\star}: \bigl[x^{\star}, x^{\star}\bigr]_{\star}=1, x^{ \star}\in \mathcal{X^{\star}} \bigr\} . \end{aligned}$$
Proof
It immediately follows from Proposition 3.6 and Theorem 3.9. □
We say that \(T \in \mathcal{L}(\mathcal{X})\) is an isometry if \([Tx, Ty]=[x, y]\) for \(x, y\in \mathcal{X}\), a unitary if it is isometric and surjective, and a Hermitian operator if \(W(T)\subset \mathbb{R}\). For a conjugation C on \(\mathcal{X}\), we have that T is an isometry (a unitary or a Hermitian operator, respectively) if and only if \(CTC\) is an isometry (a unitary or a Hermitian operator, respectively). Indeed, if T is an isometry, then for \(x, y\in \mathcal{X}\),
$$\begin{aligned}{} [CTCx, CTCy]=\overline{[TCx, TCy]}=\overline{[Cx, Cy]}=[x, y], \end{aligned}$$
which implies that \(CTC\) is an isometry. Conversely, if \(CTC\) is an isometry, then for \(x, y\in \mathcal{X}\),
$$\begin{aligned}{} [Tx,Ty]=[TCz,TCw]=\overline{[CTCz, CTCw]} = \overline{[z,w]}=[Cz, Cw]=[x,y], \end{aligned}$$
where \(z=Cx\) and \(w=Cy\). Similarly, we can see that T is a unitary if and only if \(CTC\) is a unitary. It follows from Theorem 3.9 that T is Hermitian if and only if \(CTC\) is also Hermitian.
In [8, Lemma 3.1] and [6, Theorem 3.1], it has been proved that any unitary operator on a Hilbert space can be constructed by gluing together two copies of essentially the same antilinear operator. The following proposition provides a perspective on the structure of unitary operators in a semi-inner-product space.
Proposition 3.11
If C and G are conjugations on \(\mathcal{X}\), then \(U=CG\) is a unitary and is both C-symmetric and G-symmetric.
Proof
For any \(x, y \in \mathcal{X}\), we have
$$\begin{aligned}{} [Ux, Uy]=[CGx, CGy]=\overline{[Gx, Gy]}=[x, y], \end{aligned}$$
which means that U is isometric. Since C and G are conjugations on \(\mathcal{X}\), it is obvious that U is surjective, so that it is a unitary. Moreover, we have
$$\begin{aligned}{} [CUCx, y]=[GCx, y]=\overline{[Cx, Gy]}=[x, CGy]=[x, Uy] \end{aligned}$$
and
$$\begin{aligned}{} [GUGx, y]=[GCx, y]=\overline{[Cx, Gy]}=[x, CGy]=[x, Uy]. \end{aligned}$$
Thus U is both C-symmetric and G-symmetric. □
Remark 3.12
(i) In Proposition 3.11, if \(U^{\dagger}\) is a generalized adjoint of a unitary \(U=CG\), then we get from C-symmetry of U that \(U^{\dagger}=CUC=GC\). Hence we have \(UU^{\dagger}=U^{\dagger }U=I_{\mathcal{X}}\). This means that \(U^{\dagger}=U^{-1}\).
(ii) Suppose that \(\mathcal{X}\) in Proposition 3.11 is a uniform semi-inner-product space. Let \(C^{\star}\) and \(G^{\star}\) be conjugations on \(\mathcal{X}^{\star}\) corresponding to C and G, which are given by (8). By Propositions 3.6 and 3.11, \(U^{\star}=C^{\star }G^{\star}\) is a unitary on \(\mathcal{X}^{\star}\) and is both \(C^{\star}\)-symmetric and \(G^{\star}\)-symmetric. It also follows from \(C^{\star}\)-symmetry of \(U^{\star}\) that \((U^{\star})^{\dagger }=G^{\star }C^{\star}\), so that \((U^{\star})^{\dagger}=(U^{\star})^{-1}\).