1 Introduction
Let X be a Banach space and $B(X)$ the algebra of all bounded linear operators on X. Suppose that S is a subset of $B(X)$ . Following [Reference Jiménez-Vargas, Li, Peralta, Wang and Wang4, Reference Li, Liu and Ren6], a map $\phi : X \rightarrow X$ (which is not assumed to be linear) is called a 2-local S-map if for any $a, b \in X$ , there exists $\phi _{a, b} \in S$ , depending on a and b, such that
Here, X is said to be 2-S-reflexive if every 2-local S-map belongs to S.
The concept of a 2-local S-map dates back to the paper [Reference Šemrl13], where Šemrl investigated 2-local automorphisms and 2-local derivations, motivated by Kowalski and Słodkowski [Reference Kowalski and Słodkowski5]. Then in [Reference Molnár8], the earliest investigation of 2-local $\operatorname {Iso}(X)$ -maps (also called 2-local isometries in some papers) was carried out by Molnár, where $\operatorname {Iso}(X)$ denotes the set of all surjective linear isometries of X. By an isometry of X, we mean a function $\varphi : X \rightarrow X$ such that $\|\varphi (a)-\varphi (b)\|=\|a-b\|$ for all $a, b \in X$ . In [Reference Molnár8], Molnár proved that $B(H)$ is 2- $\operatorname {Iso}(B(H))$ -reflexive, where H is an infinite-dimensional separable Hilbert space. Recently, there has been a growing interest in 2- $\operatorname {Iso}(X)$ -reflexive problems for several operator algebras and function algebras (see, for example, [Reference Al-Halees and Fleming1, Reference Molnár9, Reference Mori12]). However, the 2- $\operatorname {Iso}(X)$ -reflexivity in the context of nest algebras has not yet been considered. In this paper, we study 2- $\operatorname {Iso}(X)$ -reflexivity in some nest algebras.
Throughout, H will denote a separable Hilbert space over $\mathbb {C}$ with $\dim H \geq 2$ , along with its dual space $H^*$ . For a subset $S \subseteq H$ , we set $S^\perp := \{f \in H^* : f(S)=0\}$ .
By a subspace lattice on H, we mean a collection $\mathcal {L}$ of closed subspaces of H with $(0)$ and H in $\mathcal {L}$ such that, for every family $\{E_r\}$ of elements of $\mathcal {L}$ , both $\bigvee \{E_r\}$ and $\bigwedge \{E_r\}$ belong to $\mathcal {L}$ , where $\bigvee \{E_r\}$ denotes the closed linear span of $\{E_r\}$ and $\bigwedge \{E_r\}$ denotes the intersection of $\{E_r\}$ . We say a subspace lattice is a nest if it is totally ordered with respect to inclusion. When there is no confusion, we identify the closed subspace and the orthogonal projection on it.
Let $\mathcal {L}$ be a subspace lattice on H and $E \in \mathcal {L}$ . Define
If $\mathcal {N}$ is a nest on H, then it is not difficult to verify that
It follows that the subspaces $\bigcup \{E: E \in \mathcal {J}(\mathcal {N})\}$ and $\bigcup \{E_{-}^{\perp }: E \in \mathcal {J}(\mathcal {N})\}$ are both dense in H and $H^*$ , respectively, where $E_{-}^{\perp }=(E_{-})^{\perp }$ .
Denote by $B(H)$ , $K(H)$ and $F(H)$ the algebra of all bounded linear operators on H, the algebra of all compact operators on H and the algebra of all bounded finite rank operators on H, respectively.
By a nest algebra $\operatorname {Alg}\mathcal {N}$ , we mean the set of all operators in $B(H)$ leaving each element in $\mathcal {N}$ invariant, that is, $\operatorname {Alg}\mathcal {N}=\{T \in B(H): T E \subseteq E \text { for all } E \in \mathcal {N}\}$ . Denote $F(\mathcal {N})=\operatorname {Alg}\mathcal {N}\cap F(H)$ and $K(\mathcal {N})=\operatorname {Alg}\mathcal {N}\cap K(H)$ .
For $x \in H$ and $f \in H^*$ , the rank-one operator $x \otimes f$ is defined as the map $z \mapsto f(z)x$ . The following well-known result about rank-one operators will be repeatedly used.
Proposition 1.1 [Reference Longstaff7].
If $\mathcal {L}$ is a subspace lattice, then $x \otimes y \in \operatorname {Alg}\mathcal {L}$ if and only if there exists an element $E \in \mathcal {L}$ such that $x \in E$ and $y \in E_-^\perp $ .
2 Main result
Our main result is the following theorem.
Theorem 2.1. Let $\mathcal {N}$ be a nest on H such that $E_+ \neq E$ for any $E \neq H, E \in \mathcal {N}$ . If $\phi $ is a 2-local isometry of $\operatorname {Alg}\mathcal {N}$ , then $\phi $ is a surjective linear isometry.
The proof of Theorem 2.1 will be organised in a series of lemmas. In what follows, $\mathcal {N}$ is a nest on H such that $E_+ \neq E$ for any $E \neq H, E \in \mathcal {N}$ and $\phi $ is a 2-local isometry of $\operatorname {Alg}\mathcal {N}$ . For $A, B \in \operatorname {Alg}\mathcal {N}$ , the symbol $\phi _{A, B}$ stands for a surjective linear isometry from $\operatorname {Alg}\mathcal {N}$ to itself such that $\phi _{A, B}(A)=\phi (A)$ and $\phi _{A, B}(B)=\phi (B)$ . For a nest $\mathcal {M}$ , we denote by $\mathcal {M}^{\perp }$ the nest $\{I-E: E \in \mathcal {M}\}$ . A conjugation is a conjugate linear map on H such that $J^2=I$ and $ \langle J x, y\rangle = \langle J y, x\rangle $ for all $x, y \in H$ .
Proposition 2.2 below is cited from the paper by Moore and Trent [Reference Moore and Trent11] where they summarise the results in [Reference Arazy and Solel2, Reference Moore and Trent10] and characterise the surjective linear isometries on nest algebras.
Proposition 2.2. Let $\mathcal {M}$ be a nest on H and $\rho : \operatorname {Alg}\mathcal {M} \rightarrow \operatorname {Alg}\mathcal {M}$ be a surjective linear isometry. Then there are unitary operators U and V in $B(H)$ such that U and $U^*$ lie in $\operatorname {Alg}\mathcal {M}$ . Moreover, one of the following cases holds:
-
(1) $\rho (A)=U V^* A V$ for every $A \in \operatorname {Alg} \mathcal {M}$ and the map $E \mapsto V^* E V$ is an order isomorphism of $\mathcal {M}$ ;
-
(2) $\rho (A)=U V^* J A^* J V$ for every $A \in \operatorname {Alg}\mathcal {M}$ , where J is a conjugation on H such that $JE=EJ$ for each $E \in \mathcal {M}$ and the map $E \mapsto V^* J E J V$ is an order isomorphism from $\mathcal {M}$ onto $\mathcal {M}^{\perp }$ .
Remark 2.3. (1) It can be easily verified that the map $T \mapsto JT^*J$ is a *-anti-isomorphism of $B(H)$ and J maps an orthonormal basis onto another orthonormal basis.
(2) For any $a, b \in H$ ,
so $(Jf \otimes Jx)^*=J(x \otimes f)J$ .
(3) If $\rho $ is a surjective linear isometry of $\operatorname {Alg}\mathcal {M}$ , then according to Proposition 2.2, for any rank-one operator $x \otimes f \in \operatorname {Alg} \mathcal {M}$ , $\rho $ maps $x \otimes f$ to either $UV^*x \otimes V^*f$ or $UV^*Jf \otimes V^*Jx$ , both of which are also rank-one operators. Since every finite rank operator in $\operatorname {Alg}\mathcal {M}$ can be written as a sum of finitely many rank-one operators in $\operatorname {Alg}\mathcal {M}$ and $\rho $ preserves linear independence, it follows that $\rho $ preserves the rank of a finite rank operator. Since $\rho ^{-1}$ is also a surjective linear isometry, $\rho $ preserves the rank in both directions.
Lemma 2.4. $\phi $ is rank preserving and $\phi |_{F(\mathcal {N})}$ is linear.
Proof. It follows from Remark 2.3 that $\phi $ is rank preserving. According to Proposition 2.2, $\phi _{A, B}(X)=U_{A, B}V_{A, B}^*XV_{A, B}$ or $\phi _{A, B}(X)=U_{A, B}V_{A, B}^*JX^*JV_{A, B}$ , where $U_{A, B}$ and $V_{A, B}$ are unitary operators in $B(H)$ depending on $A, B$ and $U_{A, B}, U_{A, B}^*$ lie in $\operatorname {Alg}\mathcal {N}$ .
First, we show that $\phi $ is complex homogeneous. For any $A \in \operatorname {Alg}\mathcal {N}$ and $\lambda \in \mathbb {C}$ , $\phi (\lambda A)=\phi _{A, \lambda A}(\lambda A)=\lambda \phi _{A, \lambda A}(A)=\lambda \phi (A)$ .
Next, we prove that $\phi $ is additive on $F(\mathcal {N})$ . For any $A, B \in F(\mathcal {N})$ , since $\phi $ is rank preserving, $\phi (A)$ and $\phi (B)$ are in $F(\mathcal {N})$ . We claim that $\operatorname {tr}(\phi (A)\phi (B)^*)=\operatorname {tr}(AB^*)$ . Indeed, if $\phi _{A, B}(X)=U_{A, B}V_{A, B}^*XV_{A, B}$ , then
If $\phi _{A, B}(X)=U_{A, B}V_{A, B}^*JX^*JV_{A, B}$ , then
Thus, for any $A, A^\prime \in F(\mathcal {N})$ , by the linearity of $\operatorname {tr}$ ,
By replacing B with $A+A^\prime $ , A and $A^\prime $ , we obtain
It follows that $\phi (A+A^{\prime })-\phi (A)-\phi (A^{\prime })=0$ , which means that $\phi $ is additive on $F(\mathcal {N})$ .
By Lemma 2.4 and [Reference Hou and Cui3, Corollary 2.2] where Hou and Cui characterise rank-1 preserving linear maps between nest algebras acting on Banach spaces, we can easily prove Lemma 2.5.
Lemma 2.5. One of the following statements holds.
-
(1) There exist injective linear transformations
$$ \begin{align*} D: \bigcup\{E: E \in \mathcal{J}(\mathcal{N})\} \rightarrow H \quad\text{and}\quad C: \bigcup\{E_{-}^{\perp}: E \in \mathcal{J}(\mathcal{N})\} \rightarrow H^* \end{align*} $$such that $\phi (x \otimes f)=D x \otimes C f$ for every $x \otimes f \in F(\mathcal {N})$ . -
(2) There exist injective linear transformations
$$ \begin{align*} D: \bigcup\{E_{-}^{\perp}: E \in \mathcal{J}(\mathcal{N})\} \rightarrow H \quad\text{and}\quad C: \bigcup\{E: E \in \mathcal{J}(\mathcal{N})\} \rightarrow H^* \end{align*} $$such that $\phi (x \otimes f)=D f \otimes C x$ for every $x \otimes f \in F(\mathcal {N})$ .
By categorising and discussing the two cases in Lemma 2.5, we can obtain the following result.
Lemma 2.6. One of the following statements holds.
-
(1) There exist unitary operators $C, D \in B(H)$ such that $\phi (A)=DAC^*$ for any $A \in K(\mathcal {N})$ .
-
(2) There exist bounded conjugate linear operators $C, D$ such that $CJ, DJ \in B(H)$ are unitary operators and $\phi (A)=(DJ)JA^*J(CJ)^*$ for any $A \in K(\mathcal {N})$ .
Proof. We consider two cases.
Case 1. If Lemma 2.5(1) holds, then based on the assumption on $\mathcal {N}$ , there exist injective linear transformations $D: \bigcup \{E : E \in \mathcal {J}(\mathcal {N})\}\rightarrow H $ and $C: H^* \rightarrow H^*$ such that $\phi (x \otimes f)=D x \otimes C f$ for every $x \otimes f \in F(\mathcal {N})$ . Thus, for any $x \otimes f \in \operatorname {Alg}\mathcal {N}$ ,
Fix $x_0 \neq 0 \in (0)_+$ . Then $x_0 \otimes f$ is in $\operatorname {Alg}\mathcal {N}$ for any $f \neq 0, f \in ((0)_+)_-^\perp =H^*$ . It follows that $\|Dx_0\|\,\|Cf\| =\|x_0\|\,\|f\|$ . So ${\|Cf\|}/{\|f\|}={\|x_0\|}/{\|Dx_0\|}$ for any $f \neq 0, f \in H^*$ , which means that $C \in B(H^*)$ and $\|C\|={\|x_0\|}/{\|Dx_0\|}$ .
For any $E \in \mathcal {J}(\mathcal {N})$ , fix $f_0 \neq 0, f \in E_-^\perp $ . Then $x \otimes f_0 \in \operatorname {Alg}\mathcal {N}$ for any $x \neq 0, x \in E$ . It follows that $\|Dx\|\,\|Cf_0\| =\|x\|\,\|f_0\|$ . Therefore, ${\|Dx\|}/{\|x\|} ={\|f_0\|}/{\|Cf_0\|} ={\|Dx_0\|}/{\|x_0\|}$ , which means that $\|D|_E\| = {\|Dx_0\|}/{\|x_0\|}$ . Since $\bigcup \{E : E \in \mathcal {J}(\mathcal {N})\}$ is dense in H, we can extend D to an operator in $B(H)$ also denoted by D such that ${\|Dx\|}/{\|x\|}= {\|Dx_0\|}/{\|x_0\|}$ for any $x \neq 0, x \in H$ . So we can assume that $C, D$ are isometries. Since $\phi $ is an isometry, by the linearity of $\phi |_{F(\mathcal {N})}$ and the continuity of $\phi $ , we have $\phi (A)=DAC^*$ for all $A \in K(\mathcal {N})$ .
By the Riesz–Frechet theorem, $H^*$ can be identified with H through a conjugate linear surjective isometry. For any $E \neq H, E \in \mathcal {N}$ , we have $(E_+)_-=E$ by the hypothesis on $\mathcal {N}$ . Thus, x is in $(E_+)_-^{\perp }$ for any $x \in E_+ \ominus E$ , and so $x \otimes x \in \operatorname {Alg}\mathcal {N}$ . Let $\mathcal {N}=\{E_j:j \in \Omega \}$ and $\{e_i^j : i \in \Lambda _j\}$ be an orthonormal basis of $(E_j)_+ \ominus E_j$ . Then $K:=\sum _{i,j} e_i^j \otimes e_i^j/(i\cdot j)$ is a compact operator in $\operatorname {Alg}\mathcal {N}$ . Moreover, K is an injective operator with dense range. We claim that $\phi (K)$ is also an injective operator with dense range.
For the case when $\phi (K)=U_{K, 0}V_{K, 0}^*KV_{K, 0}$ , since $U_{K, 0}, V_{K, 0}$ are unitary operators, $\phi (K)$ is also an injective operator with dense range.
For the case when $\phi (K)=U_{K, 0}V_{K, 0}^*JK^*JV_{K, 0}$ , since $\operatorname {Ker} K= (\operatorname {Ran} K^*)^\perp $ , $K^*$ is an injective operator with dense range. As J is a conjugate linear isometry, it follows that $\phi (K)$ is also an injective operator with dense range.
Therefore, $\phi (K)=\sum _{i,j} De_i^j \otimes Ce_i^j/(i\cdot j)$ is an injective operator with dense range, which implies D and C have dense ranges. Consequently, D and C are surjective isometries (unitary operators).
Case 2. If Lemma 2.5(2) holds, then there exist injective linear transformations $D: H^* \rightarrow H$ and $C: \bigcup \{E \in \mathcal {N} \mid E_{-} \neq H\} \rightarrow H^*$ such that $\phi (x \otimes f)=D f \otimes C x$ for every $x \otimes f \in F(\mathcal {N})$ .
According to the Riesz–Frechet theorem, we can consider D as an injective conjugate linear transformation from H to H, and C as an injective conjugate linear transformation from $\bigcup \{E \in \mathcal {N} \mid E_{-} \neq H\}$ to H. Similarly to Case 1, we can conclude that $DJ$ and $CJ$ are unitary operators. By Remark 2.3,
for any $x \otimes f \in \operatorname {Alg}\mathcal {N}$ . By the linearity of $\phi |_{F(\mathcal {N})}$ and the continuity of $\phi $ , we have $\phi (A)=(DJ)(JA^*J)(CJ)^*$ for any $A \in K(\mathcal {N})$ .
Lemma 2.7. $\phi (P)\phi (T)^*\phi (P)=\phi (PT^*P)$ for any $T \in \operatorname {Alg}\mathcal {N}$ and any $P=x \otimes f \in \operatorname {Alg}\mathcal {N}$ .
Proof. By Lemma 2.2, $\phi _{P, T}(X)=U_{P, T}V_{P, T}^*XV_{P, T}$ or $\phi _{P, T}(X)=U_{P, T}V_{P, T}^*JX^*JV_{P, T}$ . To simplify the notation, denote $U_{P, T}, V_{P, T}$ by $U, V$ , respectively. For $\phi _{P, T}(X)=UV^*XV$ ,
For $\phi _{P, T}(X)=UV^*JX^*JV$ , using Remark 2.3,
Furthermore, if $\phi $ is the form in Lemma 2.6(1), then $DPC^* \phi (T)^* DPC^*=DPT^*PC^*$ , which implies that
for any $T \in \operatorname {Alg}\mathcal {N}$ and $P=x \otimes f \in \operatorname {Alg}\mathcal {N}$ .
If $\phi $ is the form in Lemma 2.6(2), then it follows that
which implies that
for any $T \in \operatorname {Alg}\mathcal {N}$ and any $P=x \otimes f \in \operatorname {Alg}\mathcal {N}$ .
Under the assumption on $\mathcal {N}$ , Lemmas 2.8 and 2.9 follow from Proposition 2.2.
Lemma 2.8. Let $\rho : \operatorname {Alg}\mathcal {N} \rightarrow \operatorname {Alg}\mathcal {N}$ be a surjective linear isometry. If Case (1) in Proposition 2.2 holds for $\rho $ , then $V, V^*$ are in $\operatorname {Alg}\mathcal {N}$ .
Proof. It is sufficient to show that $V^* E V=E$ for all $E \in \mathcal {N}$ . We prove it by the principle of transfinite induction.
It is evident that $V^*(0)V=(0)$ . Moreover, for any given $F \in \mathcal {N}$ , if the equation $V^* G V=G$ holds for all $G \in \mathcal {N}$ such that $G < F$ , then because $E \mapsto V^* E V$ is an order isomorphism from $\mathcal {N}$ onto $\mathcal {N}$ , it follows that $V^* F V = F$ .
Lemma 2.9. Let $\rho : \operatorname {Alg}\mathcal {N} \rightarrow \operatorname {Alg}\mathcal {N}$ be a surjective linear isometry. If Case (2) in Proposition 2.2 holds for $\rho $ , then the following statements hold.
-
(1) $E_- \neq E$ for any $E \neq (0), E \in \mathcal {N}$ .
-
(2) $\mathcal {N}$ is finite.
-
(3) We denote $\mathcal {N}=\{E_0, E_1, \ldots , E_n\}$ where $(0)=E_0 < E_1 < \cdots <E_n = H$ . Then $V^*$ and V both map $E_i$ onto $I-E_{n-i}$ for $0 \leq i \leq n$ .
Proof. (1) In the nest $\mathcal {N^\perp }$ , we denote $E_+^{\mathcal {N^\perp }}=\bigwedge \{F \in \mathcal {N^\perp }: F \nsubseteq E\}$ for any $E \neq H, E \in \mathcal {N^\perp }$ , and $E_-^{\mathcal {N^\perp }}=\bigvee \{F \in \mathcal {N^\perp }: F \nsupseteq E\}$ for any $E \neq (0), E \in \mathcal {N^\perp }$ .
Since the map $\pi : E \mapsto V^* E V$ is an order isomorphism from $\mathcal {N}$ onto $\mathcal {N}^{\perp }$ , we have $(I-E)_+^{\mathcal {N^\perp }}\neq (I-E)$ for any $I-E \neq H, I-E \in \mathcal {N^\perp }$ . So
for any $I-E \neq H, I-E \in \mathcal {N^\perp }$ . It follows that $E_- \neq E$ for any $E \neq (0) \in \mathcal {N}$ .
(2) Suppose that $\mathcal {N}$ is infinite, then there is a sequence $\{E_i: i \in \mathbb {N}\} \subseteq \mathcal {N}$ such that $E_i \neq (0)$ or H for any $i \in \mathbb {N}$ and $E_i < E_j$ when $i < j$ . Let $G=\bigvee \{E_i: i \in \mathbb {N}\}$ . Then $G_-=\bigvee \{F \in \mathcal {N}: F < G\}\supseteq \bigvee \{E_i: i \in \mathbb {N}\}=G$ which contradicts $G_- \neq G$ . This implies that $\mathcal {N}$ is finite.
(3) Since $E \mapsto V^* J E J V$ is an order isomorphism from $\mathcal {N}$ onto $\mathcal {N}^{\perp }$ and $EJ=JE$ for any $E \in \mathcal {N}$ , we obtain $E_i \mapsto V^*E_iV=I-E_{n-i}$ for $0 \leq i \leq n$ . Since V is a unitary operator, it follows that $V^*$ and V both map $E_i$ onto $I-E_{n-i}$ for $0 \leq i \leq n$ .
Using the characterisation of the $\phi _{A, B}$ provided by Proposition 2.2, we divide the proof of Theorem 2.1 into two lemmas based on whether $\mathcal {N}$ is isomorphic to $\mathcal {N}^\perp $ .
Lemma 2.10. If $\mathcal {N}$ is not order isomorphic to $\mathcal {N}^\perp $ , then $\phi $ is a surjective linear isometry.
Proof. Since $\mathcal {N}$ is not order isomorphic to $\mathcal {N}^\perp $ , every surjective linear isometry of $\operatorname {Alg}\mathcal {N}$ is of the form in Proposition 2.2(1). We distinguish two cases according to Lemma 2.6.
Case 1. Suppose that Lemma 2.6(1) holds, that is, $\phi (A)=DAC^*$ for every $A \in K(\mathcal {N})$ where $C, D$ are unitary operators. We claim that C and D are both in $\operatorname {Alg}\mathcal {N} \cap \operatorname {Alg}\mathcal {N}^\perp $ .
For any fixed $E \in \mathcal {N}$ , if $x \neq 0, x \in E$ and $f \neq 0, f \in E_-^\perp $ , then it follows from $\phi (x \otimes f)=Dx \otimes Cf=U_{T, x \otimes f}V^*_{T, x \otimes f}(x \otimes f)V_{T, x \otimes f}$ that
where $\lambda _{T, x \otimes f} \in \mathbb {C}$ is on the unit circle.
By Proposition 2.2 and Lemma 2.8, $U_{T, x \otimes f}, V_{T, x \otimes f}$ are both in $\operatorname {Alg}\mathcal {N} \cap \operatorname {Alg}\mathcal {N}^\perp $ . Fix $x_0 \neq 0, x_0 \in (0)_+$ . Then $x_0 \otimes f$ is in $\operatorname {Alg}\mathcal {N}$ for any $f \neq 0, f \in H$ . Thus, for any $E \neq (0), E \in \mathcal {N}$ , we have $Cf=V_{T, x_0 \otimes f}^*f/{\overline {\lambda }_{T, x_0 \otimes f}} \in E$ for any $f \neq 0, f \in E$ . Also, for any $E \neq H, E \in \mathcal {N}$ , we have $Cf=V_{T, x_0 \otimes f}^*f/{\overline {\lambda }_{T, x_0 \otimes f}} \in E^\perp $ for any $f \neq 0, f \in E^\perp $ . This shows that C is in $\operatorname {Alg}\mathcal {N} \cap \operatorname {Alg}\mathcal {N}^\perp $ .
For any fixed $E \in \mathcal {J}(\mathcal {N})$ , there exists an $f_0 \neq 0, f_0 \in E_-^\perp $ . It follows that $Dx=\lambda _{T, x \otimes f_0} U_{T, x \otimes f_0}V_{T, x \otimes f_0}^*x \in E$ for any $x \neq 0, x \in E$ , which means that $D \in \operatorname {Alg}\mathcal {N}$ .
Fix $E \in \mathcal {J}(\mathcal {N})$ . Then, for any $y \in E$ and any $x \in E^\perp \cap (\bigcup \{F: F \in \mathcal {J}(\mathcal {N})\})$ ,
So $D^*E \perp (E^\perp \cap (\bigcup \{F: F \in \mathcal {J}(\mathcal {N})\}))$ . Since $E^\perp \cap (\bigcup \{F: F \in \mathcal {J}(\mathcal {N})\})$ is dense in $E^\perp $ , it follows that $D^* \in \operatorname {Alg}\mathcal {N}$ . This completes the claim.
For any $T \in \operatorname {Alg}\mathcal {N}$ , denote $G:=C^* \phi (T)^* D-T^*$ . By (2.1), $f(Gx)x \otimes f=0$ for any $P=x \otimes f \in \operatorname {Alg}\mathcal {N}$ . Thus, G maps $E_+$ into E for any $E \neq H, E \in \mathcal {N}$ . It is clear that G is in $\operatorname {Alg}\mathcal {N}^\perp $ , and hence G maps every $E^\perp \in \mathcal {N}^\perp $ into $E^\perp $ . It follows that G maps $E_+ \ominus E=E_+ \cap E^\perp $ into $E \cap E^\perp $ for any $E \neq H, E \in \mathcal {N}$ which yields $G=0$ and $\phi (T)=DTC^*$ .
Case 2. Suppose that Lemma 2.6(2) holds, that is, $\phi (x \otimes f)=D f \otimes C x$ for every $x \otimes f \in \operatorname {Alg}\mathcal {N}$ where $C, D$ are conjugate linear operators such that $CJ, DJ \in B(H)$ are unitary operators.
Then for $x_0 \neq 0, x_0 \in (0)_+$ and linear independent $f_1, f_2 \in H$ ,
and
It follows that $Df_1$ and $Df_2$ are linearly dependent which leads to a contradiction.
In conclusion, $\phi (T)=DTC^*$ for any $T \in \operatorname {Alg}\mathcal {N}$ and it is clear that $\phi $ is a surjective linear isometry of $\operatorname {Alg}\mathcal {N}$ .
Lemma 2.11. If $\mathcal {N}$ is order isomorphic to $\mathcal {N}^\perp $ , then $\phi $ is a surjective linear isometry.
Proof. According to Lemma 2.9, $\mathcal {N}$ is finite; denote $\mathcal {N}=\{E_0, E_1, \ldots , E_n\}$ where $(0)=E_0 < E_1 < \cdots <E_n = H$ . We distinguish two cases according to Lemma 2.6.
Case 1. Suppose that Lemma 2.6(1) holds, that is, $\phi (A)=DAC^*$ for every $A \in K(\mathcal {N})$ where $C, D$ are unitary operators. In this case, for any $E \in \mathcal {J}(\mathcal {N})$ satisfying $\dim E_-^\perp> 1$ , fix $x_0 \neq 0, x_0 \in E$ . For any linearly independent $f_1, f_2 \in E_-^\perp $ , we have $x_0 \otimes f_1, x_0 \otimes f_2 \in \operatorname {Alg}\mathcal {N}$ .
We claim that $\phi _{x_0 \otimes f_1, x_0 \otimes f_2}$ is not of the form in Proposition 2.2(2). Otherwise,
and
It follows that $f_1$ and $f_2$ are linear dependent, leading to a contradiction.
Thus, for any $f_1 \neq 0, f_1 \in H$ , there exist $x_0 \neq 0, x_0 \in (0)_+$ and $f_2 \neq 0, f_2 \in H$ such that
Hence, $Dx_0=\lambda _{x_0 \otimes f_1, x_0 \otimes f_2} U_{x_0 \otimes f_1, x_0 \otimes f_2}V_{x_0 \otimes f_1, x_0 \otimes f_2}^*x_0$ and $Cf_1=V_{x_0 \otimes f_1, x_0 \otimes f_2}^*f_1/{\overline {\lambda }_{x_0 \otimes f_1, x_0 \otimes f_2}}$ for some $\lambda _{x_0 \otimes f_1, x_0 \otimes f_2} \in \mathbb {C}$ on the unit circle. By the arbitrariness of $f_1$ and $V_{x_0 \otimes f_1, x_0 \otimes f_2}^* \in \operatorname {Alg}\mathcal {N} \cap \operatorname {Alg}\mathcal {N}^\perp $ , we obtain $C \in \operatorname {Alg}\mathcal {N} \cap \operatorname {Alg}\mathcal {N}^\perp $ .
Similarly, for any $E \in \mathcal {N}$ with $\dim E> 1$ , fix $f_0 \in E_-^\perp $ . Let $x_1, x_2 \in E$ be any linearly independent elements. It is impossible for $\phi _{x_1 \otimes f_0, x_2 \otimes f_0}$ to be in the form of Lemma 2.2(2). Thus, for any $x_1 \neq 0, x_1 \in H$ , there exist $f_0 \neq 0, f_0 \in H_-^\perp $ and $x_2 \neq 0, x_2 \in H$ such that
It follows that $D \in \operatorname {Alg}\mathcal {N} \cap \operatorname {Alg}\mathcal {N}^\perp $ .
For any $T \in \operatorname {Alg}\mathcal {N}$ , denote $G:=C^* \phi (T)^* D-T^*$ . Using a similar method to that in Lemma 2.10, we see that G maps $E_+ \ominus E=E_+ \cap E^\perp $ into $E \cap E^\perp $ for any $E \neq H, E \in \mathcal {N}$ , which yields $G=0$ and $\phi (T)=DTC^*$ .
Case 2. Suppose that Lemma 2.6(2) holds, that is, $\phi (x \otimes f)=D f \otimes C x$ for every $x \otimes f \in \operatorname {Alg}\mathcal {N}$ where $C, D$ are conjugate linear operators such that $CJ, DJ \in B(H)$ are unitary operators.
In this case, for any $E \in \mathcal {J}(\mathcal {N})$ with $\dim E_-^\perp> 1$ , fix $x_0 \in E$ . For any linearly independent $f_1, f_2 \in E_-^\perp $ , $x_0 \otimes f_1, x_0 \otimes f_2$ are in $\operatorname {Alg}\mathcal {N}$ . It is impossible for $\phi _{x_0 \otimes f_1, x_0 \otimes f_2}$ to be in the form of Proposition 2.2(1). Otherwise,
and
implying that $f_1, f_2$ are linear dependent, which leads to a contradiction.
Thus, for any $f_1 \neq 0, f_1 \in H$ , there exist $x_0 \neq 0, x_0 \in (0)_+$ and $f_2 \neq 0, f_2 \in H$ such that
So $Df_1=\lambda _{x_0 \otimes f_1, x_0 \otimes f_2} U_{x_0 \otimes f_1, x_0 \otimes f_2}V_{x_0 \otimes f_1, x_0 \otimes f_2}^*Jf_1$ and $Cx_0=V_{x_0 \otimes f_1, x_0 \otimes f_2}^*Jx_0/{\overline {\lambda }_{x_0 \otimes f_1, x_0 \otimes f_2}}$ for some $\lambda _{x_0 \otimes f_1, x_0 \otimes f_2} \in \mathbb {C}$ on the unit circle. According to Lemma 2.9, $V_{x_0 \otimes f_1, x_0 \otimes f_2}^*$ and $V_{x_0 \otimes f_1, x_0 \otimes f_2}$ both map $E_i$ onto $I-E_{n-i}$ for $0 \leq i \leq n$ . Since $EJ=JE$ for any $E \in \mathcal {N}$ , by the arbitrariness of $f_1$ and $U_{x_0 \otimes f_1, x_0 \otimes f_2} \in \operatorname {Alg}\mathcal {N} \cap \operatorname {Alg}\mathcal {N}^\perp $ , we see that D maps $E_i$ into $I-E_{n-i}$ and $I-E_i$ into $E_{n-i}$ for $0\leq i\leq n$ , respectively.
Similarly, for any $E \in \mathcal {N}$ with $\dim E> 1$ , fix $f_0 \in E_-^\perp $ . For any linearly independent $x_1, x_2 \in E$ , $x_1 \otimes f_0, x_2 \otimes f_0$ are in $\operatorname {Alg}\mathcal {N}$ . It is impossible for $\phi _{x_1 \otimes f_0, x_2 \otimes f_0}$ to be in the form of Proposition 2.2(1). Thus, for any $x_1 \neq 0, x_1 \in H$ , there exist $f_0 \neq 0, f_0 \in H_-^\perp $ and $x_2 \neq 0, x_2 \in H$ such that
So $Df_0=\lambda _{x_1 \otimes f_0, x_2 \otimes f_0} U_{x_1 \otimes f_0, x_2 \otimes f_0}V_{x_1 \otimes f_0, x_2 \otimes f_0}^*Jf_0$ and $Cx_1=V_{x_1 \otimes f_0, x_2 \otimes f_0}^*Jx_1/{\overline {\lambda }_{x_1 \otimes f_0, x_2 \otimes f_0}}$ for some $\lambda _{x_1 \otimes f_0, x_2 \otimes f_0} \in \mathbb {C}$ on the unit circle. Since $V_{x_1 \otimes f_0, x_2 \otimes f_0}^*, V_{x_1 \otimes f_0, x_2 \otimes f_0}$ both map $E_i$ onto $I-E_{n-i}$ for any $0 \leq i \leq n$ and $EJ=JE$ for any $E \in \mathcal {N}$ , by the arbitrariness of $x_1$ , we see that C maps $E_i$ into $I-E_{n-i}$ and $I-E_i$ into $E_{n-i}$ for all $0\leq i\leq n$ , respectively.
By (2.2), $ (JP^*J)((CJ)^* \phi (T)^* (DJ)-(JTJ))(JP^*J)=0 $ for any $T \in \operatorname {Alg}\mathcal {N}$ and any $P=x \otimes f \in \operatorname {Alg}\mathcal {N}$ . So $\langle ((CJ)^* \phi (T)^* (DJ)-JTJ)Jf , Jx \rangle =0$ for all $P=x \otimes f \in \operatorname {Alg}\mathcal {N}$ which means that $((CJ)^* \phi (T)^* (DJ)-JTJ)$ maps $(E_i)_-^\perp $ into $(E_i)^\perp $ .
Moreover, for any $E_i \in \mathcal {N}$ ,
and $JTJ$ maps $E_i$ into $E_i$ . It follows that $((CJ)^* \phi (T)^* (DJ)-JTJ)$ maps $E_i \cap (E_i)_-^\perp $ into $E_i \cap E_i^\perp = \{0\}$ . So $((CJ)^* \phi (T)^* (DJ)-JTJ)=0$ , which implies that $\phi (T)=(DJ) J T^* J (CJ)^*$ for any $T \in \operatorname {Alg}\mathcal {N}$ . It is easy to check that $\phi (T)$ is a surjective linear isometry.
Combining Lemmas 2.10 and 2.11 completes the proof of Theorem 2.1.