3.1 The Properties of Numerical Radius Isometry from $S(\mathcal{M}_n)$ onto Itself
From Lemma 2.2, the space $(\mathcal{M}_n, \omega(\cdot))$ is a Banach space, we write $\mathcal{M}_n$ instead of $(\mathcal{M}_n, \omega(\cdot))$ for convenience. A mapping $\Phi:\mathcal{M}_n\rightarrow\mathcal{M}_n$ is a numerical radius isometry if $\omega(\Phi(A)-\Phi(B))=\omega(A-B)$ for all $A, B\in\mathcal{M}_n$. Denote by $S(\mathcal{M}_n)=\{A\in\mathcal{M}_n:\omega(A)=1\}$ the unit sphere of $\mathcal{M}_n$.
Lemma 3.1 If $A, B\in S(\mathcal{M}_n)$ are real linearly independent such that $\alpha A+\beta B\in S(\mathcal{M}_n)$ holds for some $\alpha, \beta\in\Bbb{R}$ with $\alpha^2+\beta^2=1$, then condition (a) and (b) are equivalent.
(a) There exists a complex unit $\mu$ such that $(A, B)=\mu(I, \pm iI)$.
(b) For any rank one matrix $A_1\in S(\mathcal{M}_n)$, there are $\alpha, \beta\in\Bbb{R}$ with $\alpha^2+\beta^2=1$ such that $\omega(\alpha A+\beta B+A_1)=1+\omega(A_1)$.
Proof It is obvious that (a) implies (b).
Now assume $(b)$ holds. For any $x\in S(\Bbb{C}^n)$ and $\theta\in[0, 2\pi)$. Let $A_{\theta}=e^{i\theta}xx^*$, it follows from (b) that there are $\alpha_{\theta}, \beta_{\theta}\in\Bbb{R}$ with $\alpha_{\theta}^2+\beta_{\theta}^2=1$ such that
$\omega(\alpha_{\theta}A+\beta_{\theta}B+A_{\theta})=1+\omega(A_{\theta})=2, $ |
which implies that there exists $x_{\theta}\in S(\Bbb{C}^n)$ such that
$|x_{\theta}^*(\alpha_{\theta}A+\beta_{\theta}B)x_{\theta}+x_{\theta}^*A_{\theta}x_{\theta}|=2.$ |
Since $\alpha_{\theta}A+\beta_{\theta}B\in S(\mathcal{M}_n)$ and $A_{\theta}\in S(\mathcal{M}_n)$, which implies that $|x^*(\alpha_{\theta}A+\beta_{\theta}B)x|\leq1, |x^*A_{\theta}x|\leq1$ for any $x\in S(\Bbb{C}^n)$. Hence
$2=|x_{\theta}^*(\alpha_{\theta}A+\beta_{\theta}B)x_{\theta}+x_{\theta}^*A_{\theta}x_{\theta}|\leq |x_{\theta}^*(\alpha_{\theta}A+\beta_{\theta}B)x_{\theta}|+|x_{\theta}^*A_{\theta}x_{\theta}|\leq2.$ |
So we have $|x_{\theta}^*(\alpha_{\theta}A+\beta_{\theta}B)x_{\theta}|=1$ and $|x_{\theta}^*A_{\theta}x_{\theta}|=1.$ From which we obtain $x_{\theta}=x$ since $A_{\theta}=e^{i\theta}xx^*$. So $x_{\theta}^*(\alpha_{\theta}A+\beta_{\theta}B)x_{\theta}=e^{i\theta}.$
Suppose $x^*Ax=\alpha_1+i\alpha_2$ and $x^*Bx=\beta_1+i\beta_2$ with $\alpha_1, \alpha_2, \beta_1, \beta_2\in\Bbb{R}$. Let $C= \left( \begin{array}{cc} \alpha_1 \beta_1 \\ \alpha_2 \beta_2 \\ \end{array} \right)\in\mathcal{M}_2(\Bbb{R})$. By above equation, for any $\theta\in[0, 2\pi)$, we have
$\begin{aligned} x^*(\alpha_{\theta}A+\beta_{\theta}B)x&=x^*(\alpha_{\theta}A)x+x^*(\beta_{\theta}B)x\\&=\alpha_{\theta}(\alpha_1+i\alpha_2)+\beta_{\theta}(\beta_1+i\beta_2)\\&=(\alpha_1\alpha_{\theta}+\beta_1\beta_{\theta})+i(\alpha_2\alpha_{\theta}+\beta_2\beta_{\theta})\\&=e^{i\theta}. \end{aligned}$ |
Now let $u_{\theta}=(\alpha_{\theta}, \beta_{\theta})^{tr}$ then $u_{\theta}\in S(\Bbb{R}^2)$ and
$Cu_{\theta}= \left( \begin{array}{cc} \alpha_1 \beta_1 \\ \alpha_2 \beta_2 \\ \end{array} \right) \left( \begin{array}{c} \alpha_{\theta} \\ \beta_{\theta} \\ \end{array} \right) =\left( \begin{array}{c} \cos\theta \\ \sin\theta \\ \end{array} \right) =(\cos\theta, \sin\theta)^{tr}.$ |
Hence $C$ maps the unit ball in $\Bbb{R}^2$ onto itself, and thus $C$ is an isometry on $\Bbb{R}^2$ with the form
$C= \left( \begin{array}{cc} \cos t -\sin t \\ \sin t \cos t \\ \end{array} \right)\in\mathcal{M}_2(\Bbb{R}) \, \, \, {\rm or}\, \, \, C= \left( \begin{array}{cc} \cos t \sin t \\ \sin t -\cos t \\ \end{array} \right)\in\mathcal{M}_2(\Bbb{R}) .$ |
It follows that $x^*Ax=\alpha_1+i\alpha_2$ is a complex unit and $x^*Bx=\pm ix^*Ax=x^*(\pm iA)x.$
Since $x\in S(\Bbb{C}^n)$ is arbitrary, we see that $B=\pm iA$, together with the convexity of $W(A)$, implies that $A=\mu I$ for some complex unit $\mu$. Hence $(A, B)=\mu(I, \pm iI)$.
Lemma 3.2 Suppose $\Phi:S(\mathcal{M}_n)\rightarrow S(\mathcal{M}_n)$ is a surjective numerical radius isometry. If $\Phi(\alpha I+\beta iI)=\alpha\Phi(I)+\beta \Phi(iI)$ for all $\alpha, \beta\in\Bbb{R}$ with $\alpha^2+\beta^2=1$, then $(\Phi(I), \Phi(iI))=\mu(I, \pm iI))$ for some complex unit $\mu$, and consequently, $\Phi(\lambda I)=\mu\lambda I$ or $\Phi(\lambda I)=\mu\bar{\lambda} I$ for all complex unit $\lambda$.
Proof It is easy to check that real linearly independent pair $(I, iI)$ satisfying $\omega(\alpha I+\beta iI)=1, {\rm i.e.}, \alpha I+\beta iI\in S(\mathcal{M}_n)$ for all $\alpha, \beta\in\Bbb{R}$ with $\alpha^2+\beta^2=1$ and condition $(b)$ of Lemma 3.1, that is $\omega(\alpha I+\beta iI+A_1)=1+\omega(A_1).$
Since $\Phi:S(\mathcal{M}_n)\rightarrow S(\mathcal{M}_n)$ is a surjective numerical radius isometry, it follows that
$\begin{aligned} \omega(\alpha\Phi(I)+\beta\Phi(iI)+\Phi(A_1))&=\omega(\Phi(\alpha I+\beta iI)-\Phi(-A_1))\\ &=\omega(\alpha I+\beta iI+A_1)\\&=1+\omega(A_1)\\&=1+\omega(\Phi(A_1)). \end{aligned}$ |
From Lemma 3.1, condition (a) holds, that is $(\Phi(I), \Phi(iI))=\mu(I, \pm iI).$
Hence $\Phi(\lambda I)=\mu\lambda I$ or $\Phi(\lambda I)=\mu\bar{\lambda} I$ for all complex unit $\lambda$.
Theorem 3.3 Suppose $\Phi:S(\mathcal{M}_n)\rightarrow S(\mathcal{M}_n)$ is a surjective numerical radius isometry. If $\Phi(\alpha I+\beta iI)=\alpha\Phi(I)+\beta \Phi(iI)$ for all $\alpha, \beta\in\Bbb{R}$ with $\alpha^2+\beta^2=1$, then $W(\Phi(A))=W(\mu A), \forall A\in S(\mathcal{M}_n)$ for some complex unit $\mu$.
Proof By Lemma 3.2, there have two cases.
Case 1 $\Phi(\lambda I)=\mu\lambda I$.
Assume that there exists $\gamma\in\Bbb{C}$ such that $\gamma\in W(\Phi(A))\setminus W(\mu A)$. Then there is a circle with sufficient large radius and centered at a certain complex unit $\lambda$ such that $W(A)$ lies inside the circle, but $\gamma$ lies outside the circle. Hence
$\omega(\mu A-\lambda I)<|\gamma-\lambda|\leq\omega(\Phi(A)-\lambda I)=\omega(\Phi(A)-\Phi(\bar{\mu}\lambda I))=\omega(A-\bar{\mu}\lambda I) =\omega(\mu A-\lambda I), $ |
which is a contradiction. So $W(\Phi(A))\subseteq W(\mu A)$. Using the argument to $\Phi^{-1}$, we obtain that $W(\mu A)\subseteq W(\Phi(A))$.
Case 2 $\Phi(\lambda I)=\mu\bar{\lambda} I$.
Assume that there exists $\gamma\in\Bbb{C}$ such that $\gamma\in W(\Phi(A))\setminus W(\mu A)$. Then there is a circle with sufficient large radius and centered at a certain complex unit $\bar{\lambda}$ such that $W(A)$ lies inside the circle, but $\gamma$ lies outside the circle. Hence
$\omega(\mu A-\bar{\lambda} I)<|\gamma-\bar{\lambda}|\leq\omega(\Phi(A)-\bar{\lambda} I)=\omega(\Phi(A)-\Phi(\bar{\mu}\bar{\lambda} I))=\omega(A-\bar{\mu}\bar{\lambda} I) =\omega(\mu A-\bar{\lambda} I).$ |
Which is a contradiction. So $W(\Phi(A))\subseteq W(\mu A)$. Using the argument to $\Phi^{-1}$, we obtain that $W(\mu A)\subseteq W(\Phi(A))$. So $W(\Phi(A))=W(\mu A)$ for all $A\in S(\mathcal{M}_n)$.
3.2 Numerical Radius Isometric Extension from $S(\mathcal{H}_n)$ onto Itself
In this section, we turn to consider real linear space $\mathcal{H}_n$ of self-adjoint matrices over complex field $\Bbb{C}$.
Let us start using some notation and terminology that will be used throughout the section. The space $(\mathcal{H}_n, \omega(\cdot))$ is a Banach space. We write $\mathcal{H}_n$ instead of $(\mathcal{H}_n, \omega(\cdot))$ for convenience. The unit sphere of $\mathcal{H}_n$ is $S(\mathcal{H}_n)=\{H\in\mathcal{H}_n:\omega(H)=1\}$ and the unit ball of $\mathcal{H}_n$ is $B(\mathcal{H}_n)=\{H\in\mathcal{H}_n:\omega(H)\leq1\}$. The dual space of $\mathcal{H}_n$ will be denoted by $(\mathcal{H}_n)^*$. Notice that the norm on $(\mathcal{H}_n)^*$ is defined as $\| f^*\|=\sup\{|f^*(H)|:H\in S(\mathcal{H}_n)\}$. Let $H\in S(\mathcal{H}_n)$, we set $St(H)=\{G\in S(\mathcal{H}_n):\omega(H+G)=2\}$.
To discuss the numerical radius isometry of $S(\mathcal{H}_n)$, we define the following relation $\lhd$ borrowed from [21].
Definition 3.4 [21] For $H_1, H_2\in \mathcal{H}_n$, $H_1$ is said to be smaller than $H_2$ (denoted by $H_1\lhd H_2$) if $\omega(H_1+H)=\omega(H_1)+\omega(H)$ implies $\omega(H_2+H)=\omega(H_2)+\omega(H)$ for all $H\in\mathcal{H}_n$.
The relation $\lhd$ has the following properties:
Lemma 3.5 [21] For any $H_1, H_2, H\in \mathcal{H}_n$, we have
(1) $H_1\lhd H_2\Longrightarrow\forall k_1, k_2>0, k_1H_1\lhd k_2H_2$;
(2) $H_1\lhd H_2\Longrightarrow\omega(H_1+H_2)=\omega(H_1)+\omega(H_2)$;
(3) $H_1\lhd H_2\Longleftrightarrow\omega(H_1+H)=\omega(H_1)+1$ implies $\omega(H_2+H)=\omega(H_2)+1, \forall H\in S(\mathcal{H}_n)$.
Lemma 3.6 [21] Suppose $\Phi:S(\mathcal{H}_n)\rightarrow S(\mathcal{H}_n)$ is a surjective numerical radius isometry. Then for any $H_1, H_2\in S(\mathcal{H}_n)$, $H_1\lhd H_2\Longleftrightarrow\Phi(H_1)\lhd \Phi(H_2)$.
Corollary 3.7 Suppose $\Phi:S(\mathcal{H}_n)\rightarrow S(\mathcal{H}_n)$ is a surjective numerical radius isometry. Then for any $H_1, H_2\in S(\mathcal{H}_n)$, $\omega(H_1+H_2)=2\Longleftrightarrow\omega(\Phi(H_1)+\Phi(H_2))=2$.
Proof If $\omega(H_1+H_2)=2$, then there exists $x_0\in \Bbb{C}^n$ with $\|x_0\|=1$ such that
$\omega(H_1)=|x_0^*H_1x_0|=\omega(H_2)=|x_0^*H_2x_0|=1.$ |
For every $m\in\Bbb{N}$, put $G_m=(1-\frac1m)H_1+\frac1mH_2$. Then $\omega(G_m)=\omega((1-\frac1m)H_1+\frac1mH_2)\leq1$ and $|x_0^*G_mx_0|=|x_0^*((1-\frac1m)H_1+\frac1mH_2)x_0|=1$. Hence $G_m\in S(\mathcal{H}_n)$ and $G_m\rightarrow H_1$ as $m\rightarrow\infty$, i.e., $\omega(G_m-H_1)\rightarrow0$.
We assert that $G_m\lhd H_2$.
Suppose that $\omega(G_m+H)=\omega(G_m)+\omega(H)=2$ for some $H\in S(\mathcal{H}_n)$, then there exists $f^*\in S((\mathcal{H}_n)^*)$ such that $f^*(G_m+H)=\omega(G_m+H)=2$. Since $|f^*(G_m)|\leq1, |f^*(H)|\leq1$, we have $f^*(G_m)=f^*(H)=1$. It follows that $f^*(H_1)=f^*(H_2)=f^*(H)=1$. Hence $2=f^*(H_2+H)\leq\omega(H_2+H)\leq2$, which implies that $\omega(H_2+H)=2$. By Lemma 3.5 (3), we obtain $G_m\lhd H_2$.
From Lemma 3.6, we have $\Phi(G_m)\lhd\Phi(H_2)$. Hence by Lemma 3.5 (2), we get $\omega(\Phi(G_m)+\Phi(H_2))=\omega(G_m)+\omega(H_2)=2$. Because $\omega(\Phi(G_m)-\Phi(H_1))=\omega(G_m-H_1)\rightarrow0$ as $m\rightarrow\infty$, thus $\omega(\Phi(H_2)+\Phi(H_1))=\lim\limits_{m\rightarrow\infty}\omega(\Phi(H_2)+\Phi(G_m))=2$.
For the converse, we only to substitute $\Phi$ with $\Phi^{-1}$ in the above proof since $\Phi^{-1}$ also is an onto numerical radius isometry.
Definition 3.8 For $x, y\in\Bbb{C}^n$, we say $x$ is equivalent to $y$ $($denoted by $x\sim y$$)$ if $y=e^{i\theta}x$ for some $\theta$ in $[0, 2\pi)$. The equivalent class of $x$ is denoted by $[x]=\{e^{i\theta}x: \forall\theta\in[0, 2\pi)\}$.
For each $x_0\in S(\Bbb{C}^n)$ and $\theta_0\in[0, 2\pi)$, define three sets:
$ \mathcal{NRA}([x_0])=\{H\in S(\mathcal{H}_n):|x_0^*Hx_0|=1\}, \\ \mathcal{NRA}([x_0], 1)=\{H\in S(\mathcal{H}_n):x_0^*Hx_0=1\}, \\ \mathcal{NRA}_1([x_0], 1)=\{H\in S(\mathcal{H}_n):x_0^*Hx_0=1\, {\rm and} \, |x^*Hx|<1, \forall x\in S(\Bbb{C}^n)\setminus[x_0]\}. $ |
Next we give some properties of $\mathcal{NRA}([x_0], 1)$ and $\mathcal{NRA}_1([x_0], 1)$.
Remark 1 $S(\mathcal{H}_n)$ separates the points of $S(\Bbb{C}^n/\sim)$, i.e., If $x, y\in S(\Bbb{C}^n)$ and $[x]\neq[y]$, then there exists $H\in S(\mathcal{H}_n)$ such that $x^*Hx\neq y^*Hy$.
Proof If $x, y\in S(\Bbb{C}^n)$ and $[x]\neq[y]$, Then $x\neq y$. There must exists $1\leq k\leq n, k\in \Bbb{N}, x_k, y_k\in\Bbb{C}$ such that $x_k\neq y_k$ (where $x_k$ and $y_k$ is the $kth$ component of $x$ and $y$, respectively).
Case 1 If $|x_k|\neq|y_k|$, take $H={\rm diag}(0, \cdots, 0, \underset{kth}{1}, 0, \cdots, 0)$, clearly $H\in S(\mathcal{H}_n)$ and $x^*Hx=|x_k|^2, y^*Hy=|y_k|^2$. Then $x^*Hx\neq y^*Hy$;
Case 2 If $|x_k|=|y_k|$, then there exists $\lambda\in\Bbb{C}$ with $|\lambda|=1$ such that $x_k=\lambda y_k$. Together with the hypothesis of $[x]\neq[y]$, there exists $1\leq l\leq n, l\in \Bbb{N}, l\neq k, x_l, y_l\in\Bbb{C}$ such that $x_l= \mu y_l$ with $\mu\neq\lambda$ (where $x_l$ and $y_l$ is the $lth$ component of $x$ and $y$, respectively).
Subcase 2a If $|x_l|\neq|y_l|$, Take $H={\rm diag}(0, \cdots, 0, \underset{lth}{1}, 0, \cdots, 0)$;
Subcase 2b If $|x_l|=|y_l|$, then $|\mu|=1$ and $\mu\neq\lambda$ such that $x_l= \mu y_l$.
Take
$G = \left( {\begin{array}{*{20}{c}}
\ddots &{}&{}&{}&{{\rm{ }} {\mathinner{\mkern2mu\raise1pt\hbox{.}\mkern2mu
\raise4pt\hbox{.}\mkern2mu\raise7pt\hbox{.}\mkern1mu}} }\\
{}&0& \cdots &{\overline {{y_k}} {y_l}}&{}\\
{}& \vdots &{}& \vdots &{}\\
{}&{{y_k}\overline {{y_l}} }& \cdots &0&{}\\
{\mathinner{\mkern2mu\raise1pt\hbox{.}\mkern2mu
\raise4pt\hbox{.}\mkern2mu\raise7pt\hbox{.}\mkern1mu}} &{}&{}&{}& \ddots
\end{array}} \right)$ |
(where $G=(g_{ij})$ is $n\times n$ matrix with $g_{kl}=y_k\bar{y_l}, g_{lk}=y_l\bar{y_k}$ and $g_{ij}=0, (i, j)\neq(k, l), (l, k))$.
Then $H=\frac{G}{\omega(G)}\in S(\mathcal{H}_n)$ and $x^*Hx=\frac{2}{\omega(G)}|y_ky_l|^2\Re(\lambda\bar{\mu})$ (where $\Re(\lambda\bar{\mu})$ denote the real part of $\lambda\bar{\mu}\in\lambda\bar{\mu}$), $y^*Hy=\frac{2}{\omega(G)}|y_ky_l|^2$. It easy to check that $x^*Hx\neq y^*Hy$ since $\lambda, \mu\in\Bbb{C}, \lambda\neq\mu, |\lambda|=|\mu|=1$.
Hence if $[x]\neq[y]$, then there exists $H\in S(\mathcal{H}_n)$ such that $x^*Hx\neq y^*Hy$ for all $x\in[x]$ and $y\in[y]$.
Remark 2 $\mathcal{NRA}_1([x_0], 1)\neq\emptyset$.
Proof For each $x_0\in S(\Bbb{C}^n)$, $x_0x_0^*\in\mathcal{NRA}_1([x_0], 1)$.
Remark 3 Let $x_0\in S(\Bbb{C}^n)$, then for any $H_{[x_0]}\in\mathcal{NRA}_1([x_0], 1)$, we have $St(H_{[x_0]})=\mathcal{NRA}([x_0], 1)$ and $\mathcal{NRA}([x_0], 1)$ is a closed convex subset of $S(\mathcal{H}_n)$.
Proof Let $x_0\in S(\Bbb{C}^n)$ and $H_{[x_0]}\in\mathcal{NRA}_1([x_0], 1)$, then for any $H\in St(H_{[x_0]})$, we have
$\begin{aligned} 2&=\omega(H+H_{[x_0]})=\sup\{|x^*Hx+x^*H_{[x_0]}x|:x\in S(\Bbb{C}^n)\}\\&\leq\sup\{|x^*Hx|:x\in S(\Bbb{C}^n)\}+\sup\{|x^*H_{[x_0]}x|:x\in S(\Bbb{C}^n)\}=2. \end{aligned}$ |
It follows that $x_0^*Hx_0=1$ and $H\in\mathcal{NRA}([x_0], 1)$. Hence $St(H_{[x_0]})\subseteq\mathcal{NRA}([x_0], 1)$.
Conversely, for any $H\in\mathcal{NRA}([x_0], 1)$, we have $\omega(H+H_{[x_0]})=2$ for every $H_{[x_0]}\in\mathcal{NRA}_1([x_0], 1))$. It follows that $H\in St(H_{[x_0]})$. Hence $St(H_{[x_0]})\supseteq\mathcal{NRA}([x_0], 1)$.
For any $H_1, H_2\in\mathcal{NRA}([x_0], 1)$ and $t\in[0,1]$, we have $x_0^*((1-t)H_1+tH_2)x_0=1$ and $\omega((1-t)H_1+tH_2)\leq(1-t)\omega(H_1)+t\omega(H_2)=1$, which implies that $(1-t)H_1+tH_2\in\mathcal{NRA}([x_0], 1)$. Hence $\mathcal{NRA}([x_0], 1)$ is convex.
Suppose $\{G_m\}\subseteq\mathcal{NRA}([x_0], 1)(m\in\Bbb{N})$ such that $G_m\rightarrow H$ as $m\rightarrow\infty$, i.e., $\omega(G_m-H)\rightarrow0$, then $H\in S(\mathcal{H}_n)$ since $|\omega(G_m)-\omega(H)|\leq\omega(G_m-H)\rightarrow0$ and $\omega(G_m)=1(\forall m\in\Bbb{N})$. Therefore, from $|x_0^*G_mx_0-x_0^*Hx_0|\leq\sup\{|x^*G_mx-x^*Hx|:x\in S(\mathcal{H}_n)\}=\omega(G_m-H)\rightarrow0$, we have $x_0^*Hx_0=1$. Hence $H\in\mathcal{NRA}([x_0], 1)$ and $\mathcal{NRA}([x_0], 1)$ is closed.
Lemma 3.9 Suppose $\Phi:S(\mathcal{H}_n)\rightarrow S(\mathcal{H}_n)$ is a surjective numerical radius isometry. Then for any $x_0\in S(\Bbb{C}^n)$ and $H_{[x_0]}\in\mathcal{NRA}_1([x_0], 1)$, we have
$\Phi^{-1}(\mathcal{NRA}([x_0], 1))=St(\Phi^{-1}(H_{[x_0]})).$ |
Proof For any $H\in\Phi^{-1}(\mathcal{NRA}([x_0], 1))$ and $\Phi(H)\in\mathcal{NRA}([x_0], 1)$, we have
$\omega(\Phi(H)+H_{[x_0]})=2.$ |
By Corollary 3.7, we have
$\omega(H+\Phi^{-1}(H_{[x_0]}))=2, $ |
hence $H\in St(\Phi^{-1}(H_{[x_0]}))$.
Conversely, for any $H_{[x_0]}\in\mathcal{NRA}_1([x_0], 1)$ and $H\in St(\Phi^{-1}(H_{[x_0]}))$, we have $\omega(H+\Phi^{-1}(H_{[x_0]}))=2$. By Corollary 3.7, $\omega(\Phi(H)+H_{[x_0]})=2$, namely,
$\Phi(H)\in St(H_{[x_0]})=\mathcal{NRA}([x_0], 1). $ |
Therefore, $H=\Phi^{-1}(\Phi(H))\in\Phi^{-1}(\mathcal{NRA}([x_0], 1))$.
Thus $\Phi^{-1}(\mathcal{NRA}([x_0], 1))=St(\Phi^{-1}(H_{[x_0]}))$ for any $H_{[x_0]}\in\mathcal{NRA}_1([x_0], 1)$.
Lemma 3.10 Suppose $\Phi:S(\mathcal{H}_n)\rightarrow S(\mathcal{H}_n)$ is a surjective numerical radius isometry. Then for any $x_0\in S(\Bbb{C}^n)$, $\Phi^{-1}(\mathcal{NRA}([x_0], 1))$ is a closed convex subset of $S(\mathcal{H}_n)$.
Proof Let $H_{[x_0]}\in\mathcal{NRA}_1([x_0], 1)$ be fixed. By Lemma 3.9, we have
$\omega(\Phi^{-1}(H_{[x_0]})+H_1)=\omega(\Phi^{-1}(H_{[x_0]})+H_2)=2$ |
for any $H_1, H_2\in\Phi^{-1}(\mathcal{NRA}([x_0], 1))$. Take $f_1^*\in S((\mathcal{H}_n)^*)$ such that
$f_1^*(\Phi^{-1}(H_{[x_0]})+H_1)=\omega(\Phi^{-1}(H_{[x_0]})+H_1)=2.$ |
Therefore, we have $f_1^*(\Phi^{-1}(H_{[x_0]}))=f_1^*(H_1)=1$ since $|f_1^*(\Phi^{-1}(H_{[x_0]}))|\leq1$ and $|f_1^*(H_1)|\leq1$. Hence
$2=f_1^*(\Phi^{-1}(H_{[x_0]})+\frac12(\Phi^{-1}(H_{[x_0]})+H_1))\leq\omega(\Phi^{-1}(H_{[x_0]})+\frac12(\Phi^{-1}(H_{[x_0]})+H_1))\leq2.$ |
So $\omega(\Phi^{-1}(H_{[x_0]})+\frac12(\Phi^{-1}(H_{[x_0]})+H_1))=2$. This means that
$\frac12(\Phi^{-1}(H_{[x_0]})+H_1)\in St(\Phi^{-1}(H_{[x_0]})=\Phi^{-1}(\mathcal{NRA}([x_0], 1)).$ |
So $\Phi(\frac12(\Phi^{-1}(H_{[x_0]})+H_1))\in\mathcal{NRA}([x_0], 1)$. From the assumption of $\Phi(H_2)\in\mathcal{NRA}([x_0], 1)$ and the convexity of $\mathcal{NRA}([x_0], 1)$, it is easy to get $\omega(\Phi(\frac12(\Phi^{-1}(H_{[x_0]})+H_1))+\Phi(H_2))=2$. Thus from Corollary 3.7, we have
$\omega(\frac12(\Phi^{-1}(H_{[x_0]})+H_1)+H_2)=2.$ |
Choose $f_2^*\in S((\mathcal{H}_n)^*)$ such that $f_2^*(\frac12(\Phi^{-1}(H_{[x_0]})+H_1)+H_2)=2$. This implies that $f_2^*(\Phi^{-1}(H_{[x_0]}))=f_2^*(H_1)=f_2^*(H_2)=1$. Hence
$2=f_2^*(\Phi^{-1}(H_{[x_0]})+\frac12(H_1+H_2))\leq\omega(\Phi^{-1}(H_{[x_0]})+\frac12(H_1+H_2))\leq2.$ |
So $\omega(\Phi^{-1}(H_{[x_0]})+\frac12(H_1+H_2))=2$. It follows that
$\frac12(H_1+H_2)\in St(\Phi^{-1}(H_{[x_0]}))=\Phi^{-1}(\mathcal{NRA}([x_0], 1)).$ |
Therefore, $\Phi^{-1}(\mathcal{NRA}([x_0], 1))$ is a convex subset.
Since $\Phi^{-1}$ is continuous and $\mathcal{NRA}([x_0], 1)$ is closed, $\Phi^{-1}(\mathcal{NRA}([x_0], 1))$ is a closed subset.
Lemma 3.11 Suppose $\Phi:S(\mathcal{H}_n)\rightarrow S(\mathcal{H}_n)$ is a surjective numerical radius isometry. Then for any $x\in S(\Bbb{C}^n)$, there exists $f_{[x]}^*\in S((\mathcal{H}_n)^*)$ such that $f_{[x]}^*(\Phi^{-1}(\mathcal{NRA}([x], 1)))=1$ $({\rm i.e.}, f_{[x]}^*(H)=1$, if $H\in\Phi^{-1}(\mathcal{NRA}([x], 1)))$.
Proof By Lemma 3.10, $\Phi^{-1}(\mathcal{NRA}([x], 1))$ is a closed convex subset of the surface of unit ball $B(\mathcal{H}_n)$. Hence $\Phi^{-1}(\mathcal{NRA}([x], 1))$ does not meet the interior of $B(\mathcal{H}_n)$. By Eidelheit Separation Theorem [17], there exists $f_{[x]}^*\in S((\mathcal{H}_n)^*)$ such that $f_{[x]}^*(H)=1$ for all $H$ in $\Phi^{-1}(\mathcal{NRA}([x], 1))$.
Now, for any $x\in S(\Bbb{C}^n)$, take $f_{[x]}^*\in S((\mathcal{H}_n)^*)$ to be fixed as described in Lemma 3.10, then we obtain a map: $[x]\longrightarrow f_{[x]}^*, x\in S(\Bbb{C}^n)$.
Lemma 3.12 The map $[x]\longrightarrow f_{[x]}^*, x\in S(\Bbb{C}^n)$ is injective.
Proof Let $x, y\in S(\Bbb{C}^n), [x]\neq[y]$. Suppose $f_{[x]}^*=f_{[y]}^*$, then for any $H_{[x]}\in\mathcal{NRA}_1([x], 1)$ and $H_{[y]}\in\mathcal{NRA}_1([y], 1)$, we have
$1=f_{[x]}^*(\Phi^{-1}(H_{[x]}))=f_{[y]}^*(\Phi^{-1}(H_{[x]}))=f_{[y]}^*(\Phi^{-1}(H_{[y]})).$ |
Hence
$2=f_{[y]}^*(\Phi^{-1}(H_{[x]})+\Phi^{-1}(H_{[y]}))\leq\omega(\Phi^{-1}(H_{[x]})+\Phi^{-1}(H_{[y]}))\leq2.$ |
It implies that $\omega(\Phi^{-1}(H_{[x]})+\Phi^{-1}(H_{[y]}))=2$. Which contradicts with
$\omega(\Phi^{-1}(H_{[x]})+\Phi^{-1}(H_{[y]}))=\omega(H_{[x]}+H_{[y]})<2$ |
since $[x]\neq[y]$.
Lemma 3.13 Suppose $\Phi:S(\mathcal{H}_n)\rightarrow S(\mathcal{H}_n)$ is a surjective numerical radius isometry. If $H\in\mathcal{NRA}([x])$, then there exists $f_{[x]}^*\in S((\mathcal{H}_n)^*)$ such that $f_{[x]}^*(\Phi^{-1}(H))=x^*Hx$.
Proof If $x^*Hx=1$, then $H\in\mathcal{NRA}([x], 1)$. By Lemma 3.11, there exists $f_{[x]}^*\in S((\mathcal{H}_n)^*)$ such that $f_{[x]}^*(\Phi^{-1}(H))=1=x^*Ax$.
If $x^*Hx=-1$, then $-H\in\mathcal{NRA}([x], 1)$. Hence for any $H_{[x]}\in\mathcal{NRA}_1([x], 1)$, we have
$\omega(\Phi^{-1}(H)-\Phi^{-1}(H_{[x]}))=\omega(H-H_{[x]})=\omega(-H+H_{[x]})=2.$ |
It follows that $-\Phi^{-1}(H)\in St(\Phi^{-1}(H_{[x]})=\Phi^{-1}(\mathcal{NRA}([x], 1))$. By Lemma 3.11, there exists $f_{[x]}^*\in S((\mathcal{H}_n)^*)$ such that $f_{[x]}^*(-\Phi^{-1}(H))=1$. Hence $f_{[x]}^*(\Phi^{-1}(H))=-1=x^*Hx$.
Theorem 3.14 Suppose $\Phi:S(\mathcal{H}_n)\rightarrow S(\mathcal{H}_n)$ is a surjective numerical radius isometry. If $x\in S(\Bbb{C}^n)$ is an eigenvector of $H\in S(\mathcal{H}_n)$, then there exists $f_{[x]}^*\in S((\mathcal{H}_n)^*)$ such that
$f_{[x]}^*(H)=x^*(\Phi(H))x$ |
or
$f_{[x]}^*(\Phi^{-1}(H))=x^*Hx.$ |
Proof Let $\{x_1, x_2, \cdots, x_n\}\subseteq S(\Bbb{C}^n)$ be pairwise orthogonal eigenvectors of $H$ corresponding to eigenvalues $\mu_1, \mu_2, \cdots, \mu_n$ arranged in descending order. Then $\mu_i\in\Bbb{R}$ with $-1\leq\mu_i\leq1$ for all $1\leq i\leq n$ and either $\mu_1=1$ or $\mu_n=-1$ for every self-adjoint matrix $H\in S(\mathcal{H}_n)$ (see [8, 9, 18]).
Assume $x_0$ is a eigenvector corresponding to eigenvalue $\mu_k$ ($1\leq k\leq n$) of $H$, then $\mu_k=x_0^*Hx_0$. Take
$G^{\pm}=(1\mp\mu_k)x_kx_k^*\pm\sum\limits_{i=1}^n\mu_i(x_ix_i^*)$ |
and
$U=(x_1, \cdots, x_{k-1}, x_k, x_{k+1}, \cdots, x_n), $ |
where
$G^{+}=(1-\mu_k)x_kx_k^*+\sum\limits_{i=1}^n\mu_i(x_ix_i^*)$ |
and
$G^{-}=(1+\mu_k)x_kx_k^*-\sum\limits_{i=1}^n\mu_i(x_ix_i^*).$ |
It is easy to check that
$ G^{\pm}\in S(\mathcal{H}_n), \\ x_0^*G^{\pm}x_0=x_k^*G^{\pm}x_k=1, \\ U^*G^{\pm}U={\rm diag}\{\pm\mu_1, \cdots, \pm\mu_{k-1}, 1, \pm\mu_{k+1}, \cdots, \pm\mu_n\}, \\ U^*HU={\rm diag}\{\mu_1, \cdots, \mu_{k-1}, \mu_k, \mu_{k+1}, \cdots, \mu_n\}.$ |
Hence $G^{\pm}\in\mathcal{NRA}([x_0], 1)$. By Lemma 3.13, we have $f_{[x_0]}^*(\Phi^{-1}(G^{\pm}))=1$ and
$\begin{aligned} 1\mp f_{[x_0]}^*(\Phi^{-1}(H))&=f_{[x_0]}^*(\Phi^{-1}(G^{\pm}))\mp f_{[x_0]}^*(\Phi^{-1}(H))\\&=f_{[x_0]}^*(\Phi^{-1}(G^{\pm})\mp\Phi^{-1}(H))\\&\leq\omega(\Phi^{-1}(G^{\pm})\mp\Phi^{-1}(H))\\&=\omega(G^{\pm}\mp H)\\&=\omega(U^*(G^{\pm}\mp H)U)\\&=1\mp\mu_k\\&=1\mp x_0^*Hx_0. \end{aligned}$ |
The above two inequalities imply $x_0^*Hx_0=f_{[x_0]}^*(\Phi^{-1}(H))$.
Therefore $G\in\mathcal{NRA}([x_0], 1)$. By Lemma 3.13, we have
$f_{[x_0]}^*(\Phi^{-1}(G))=1$ |
and
$\begin{aligned} 1- f_{[x_0]}^*(\Phi^{-1}(H))&=f_{[x_0]}^*(\Phi^{-1}(G))-f_{[x_0]}^*(\Phi^{-1}(H))\\&=f_{[x_0]}^*(\Phi^{-1}(G)-\Phi^{-1}(H))\\&\leq\omega(\Phi^{-1}(G)-\Phi^{-1}(H))\\&=\omega(G- H)\\&=\omega(U^*(G-H)U)\\&=1- x_0^*Hx_0. \end{aligned}$ |
The above two inequalities imply $x_0^*Hx_0=f_{[x_0]}^*(\Phi^{-1}(H))$.
Lemma 3.15 Suppose $\Phi:S(\mathcal{H}_n)\rightarrow S(\mathcal{H}_n)$ is a surjective numerical radius isometry. If for any $H_1, H_2\in S(\mathcal{H}_n)$, we have $\omega(\Phi(H_1)-\alpha\Phi(H_2))\leq\omega(H_1-\alpha H_2)$ for all $\alpha\in(0, +\infty)$, then $\Phi$ can be real linearly extended to numerical radius isometry $\widetilde{\Phi}$ of $\mathcal{H}_n$ onto itself.
Proof We first show that for any $H_1, H_2\in S(\mathcal{H}_n)$ and $\alpha\in(0, 1)$, we have
$\omega(H_1-\alpha H_2)=\sup\{\omega(H_1-H)-\omega(H-\alpha H_2):H\in S(\mathcal{H}_n)\}.$ |
In fact, $\omega(H_1-\alpha H_2)\geq\omega(H_1-H)-\omega(H-\alpha H_2)$ for any $H\in S(\mathcal{H}_n)$. So
$\omega(H_1-\alpha H_2)\geq\sup\{\omega(H_1-H)-\omega(H-\alpha H_2):H\in S(\mathcal{H}_n)\}.$ |
Define $\phi(t)=\omega(tH_1+(1-t)\alpha H_2)$, $t\in(-\infty, 0].$ Clearly, $\phi(0)=\alpha<1$ and
$\phi(t)=\omega(t(H_1-\alpha H_2)+\alpha H_2)\geq|t|\omega(H_1-\alpha H_2)-\alpha\rightarrow+\infty, \, \, (t\rightarrow-\infty).$ |
Then there exists $t_0<0$ such that $\phi(t_0)=1$, i.e., $G=t_0H_1+\alpha(1-t_0)H_2\in S(\mathcal{H}_n)$. Hence
$\quad \omega(H_1-G)-\omega(G-\alpha H_2)\\ =\omega(H_1-t_0H_1+\alpha(1-t_0)H_2)-\omega(t_0H_1+\alpha(1-t_0)H_2-\alpha H_2)\\ =\omega(H_1-\alpha H_2). $ |
Thus $\omega(H_1-\alpha H_2)=\sup\{\omega(H_1-H)-\omega(H-\alpha H_2):H\in S(\mathcal{H}_n)\}.$
Since $\Phi$ is a surjective numerical radius isometry, we have
$\quad \omega(H_1-\alpha H_2)\\ =\sup\{\omega(H_1-H)-\omega(H-\alpha H_2):H\in S(\mathcal{H}_n)\}\\ \leq\sup\{\omega(\Phi(H_1)-\Phi(H))-\omega(\Phi(H)-\alpha\Phi(H_2)):H\in S(\mathcal{H}_n)\}\\ =\omega(\Phi(H_1)-\alpha\Phi(H_2)). $ |
So $\omega(\Phi(H_1)-\alpha\Phi(H_2))=\omega(H_1-\alpha H_2)$ for all $H_1, H_2\in\mathcal{H}_n, \alpha\in(0, 1).$
For any $H\in\mathcal{H}_n$, define
$\widetilde{\Phi}(H)=\left\{ \begin{array}{cc} \omega(H)\Phi(\frac{H}{\omega(H)}), \mbox{}H\neq 0; \\ 0, \mbox{}H=0. \end{array} \right.$ |
Clearly, $\omega(\widetilde{\Phi}(H))=\omega(H)$ and $\widetilde{\Phi}(\alpha^+H)=\alpha^+\widetilde{\Phi}(H), \alpha^+\in[0, +\infty)$. Thus, $\widetilde{\Phi}$ is surjective.
Finally, we show that for any $H_1, H_2\in \mathcal{H}_n$, we have $\omega(\widetilde{\Phi}(H_1)-\widetilde{\Phi}(H_2))=\omega(H_1-H_2)$.
If $H_1=0$ or $H_2=0$, it is clear that $\omega(\widetilde{\Phi}(H_1)-\widetilde{\Phi}(H_2))=\omega(H_1-H_2)$.
If $H_1, H_2\in \mathcal{H}_n$, $H_1\neq0, H_2\neq0$, without loss of generality we may assume that $\omega(H_1)\leq\omega(H_2)$, then
$\quad \omega(\widetilde{\Phi}(H_1)-\widetilde{\Phi}(H_2))\\ =\omega(\omega(H_1)\Phi(\frac{H_1}{\omega(H_1)})-\omega(H_2)\Phi(\frac{H_2}{\omega(H_2)}))\\ =\omega(H_2)\omega(\frac{\omega(H_1)}{\omega(H_2)}\Phi(\frac{H_1}{\omega(H_1)})-\Phi(\frac{H_2}{\omega(H_2)}))\\ =\omega(H_2)\omega(\frac{\omega(H_1)}{\omega(H_2)}\frac{H_1}{\omega(H_1)}-\frac{H_2}{\omega(H_2)})\\ =\omega(H_1-H_2). $ |
Hence $\widetilde{\Phi}:\mathcal{H}_n\rightarrow \mathcal{H}_n$ is a surjective numerical radius isometry. From Mazur-Ulam Theorem [16], $\widetilde{\Phi}$ is a real linear numerical radius isometry of $\mathcal{H}_n$ onto itself with $\widetilde{\Phi}|_{S(\mathcal{H}_n)}=\Phi$ since $\widetilde{\Phi}(0)=0$.
Theorem 3.16 $\Phi:S(\mathcal{H}_n)\rightarrow S(\mathcal{H}_n)$ is a surjective numerical radius isometry satisfying $\omega(\Phi(H_1)-\alpha\Phi(H_2))\leq\omega(H_1-\alpha H_2)$ for all $H_1, H_2\in S(\mathcal{H}_n)$ and $\alpha\in(0, +\infty)$ if and only if there is a unitary matrix $U\in\mathcal{M}_n$ and a real number $\mu\in\{-1, 1\}$ such that one of the following is true:
(1) $\Phi(H)=\mu UHU^*$ for every $H\in S(\mathcal{H}_n)$;
(2) $\Phi(H)=\mu UH^{tr}U^*$ for every $H\in S(\mathcal{H}_n)$.
Proof It is obvious that every map of the form (1) and (2) is a surjective numerical radius isometry satisfying $\omega(\Phi(H_1)-\alpha\Phi(H_2))\leq\omega(H_1-\alpha H_2)$ for all $H_1, H_2\in S(\mathcal{H}_n)$ and $\alpha\in(0, +\infty)$. So we only to check the "only if" part.
By Lemma 3.15, $\Phi$ can be real linearly extended to the whole space $\mathcal{H}_n$. Using Theorem 2 in [1], there is a unitary matrix $U\in\mathcal{M}_n$ and a real number $\mu\in\{-1, 1\}$ such that one of the following is true:
(1) $\Phi(H)=\mu UHU^*$ for every $H\in S(\mathcal{H}_n)$;
(2) $\Phi(H)=\mu UH^{tr}U^*$ for every $H\in S(\mathcal{H}_n)$.