Hostname: page-component-848d4c4894-8bljj Total loading time: 0 Render date: 2024-06-26T01:10:15.950Z Has data issue: false hasContentIssue false

Range inclusion and diagonalization of complex symmetric operators

Published online by Cambridge University Press:  04 April 2024

Cun Wang
Affiliation:
School of Mathematics and Statistics, Beijing Institute of Technology, Beijing 102488, P. R. China e-mail: [email protected]
Jiayi Zhao
Affiliation:
Department of Mathematics, Jilin University, Changchun 130012, P. R. China e-mail: [email protected]
Sen Zhu*
Affiliation:
Department of Mathematics, Jilin University, Changchun 130012, P. R. China e-mail: [email protected]
Rights & Permissions [Opens in a new window]

Abstract

We consider the range inclusion and the diagonalization in the Jordan algebra $\mathcal {S}_C$ of C-symmetric operators, that are, bounded linear operators T satisfying $CTC =T^{*}$, where C is a conjugation on a separable complex Hilbert space $\mathcal H$. For $T\in \mathcal {S}_C$, we aim to describe the set $C_{\mathcal {R}(T)}$ of those operators $A\in \mathcal {S}_C$ satisfying the range inclusion $\mathcal {R}(A)\subset \mathcal {R}(T)$. It is proved that (i) $C_{\mathcal {R}(T)}=T\mathcal {S}_C T$ if and only if $\mathcal {R}(T)$ is closed, (ii) $\overline {C_{\mathcal {R}(T)}}=\overline {T\mathcal {S}_C T}$, and (iii) $C_{\overline {\mathcal {R}(T)}}$ is the closure of $C_{\mathcal {R}(T)}$ in the strong operator topology. Also, we extend the classical Weyl–von Neumann Theorem to $\mathcal {S}_C$, showing that every self-adjoint operator in $\mathcal {S}_C$ is the sum of a diagonal operator in $\mathcal {S}_C$ and a compact operator with arbitrarily small Schatten p-norm for $p\in (1,\infty )$.

Type
Article
Copyright
© The Author(s), 2024. Published by Cambridge University Press on behalf of Canadian Mathematical Society

1 Introduction

This paper is a continuation of a recent paper [Reference Wang and Zhu44], and aims to study the range inclusion and the diagonalization of complex symmetric operators. We start by recalling some terminology and basic facts.

1.1 Preliminaries

Throughout the following, ${\mathcal {H}}$ will always denote a separable, complex Hilbert space with an inner product $\langle \cdot ,\cdot \rangle $ . We let $\mathcal {B}({\mathcal {H}}_1,{\mathcal {H}}_2)$ denote the Banach space of all bounded linear operators from ${\mathcal {H}}_1$ to ${\mathcal {H}}_2$ . We shall write $\mathcal {B(H)}$ instead of $\mathcal {B}({\mathcal {H}},{\mathcal {H}})$ . A map $C:{\mathcal {H}} \rightarrow {\mathcal {H}}$ is called a conjugation if:

  1. (i) C is antilinear, i.e., $C(\alpha x+y)=\overline {\alpha } C x+ Cy$ for $x,y\in {\mathcal {H}}$ and $\alpha \in \mathbb {C}$ ,

  2. (ii) C is invertible with $C^{-1}=C$ , and

  3. (iii) $\langle Cx, Cy\rangle =\langle y,x\rangle $ for all $x, y\in {\mathcal {H}}$ .

An operator $T\in \mathcal {B(H)}$ is said to be complex symmetric (c.s.) if $CTC=T^{*}$ for some conjugation C on ${\mathcal {H}}$ ; in this case, T is called C-symmetric. The general definition of a C-symmetric operator was first given in I. M. Glazman’s paper [Reference Glazman18], which combined with [Reference Glazman19] marks the foundation of the extension theory of C-symmetric differential operators. Glazman’s work was complemented in a series of papers such as [Reference Kalinina26, Reference Knowles28, Reference Li30, Reference Makarova32, Reference Race35, Reference Rayh and Tsekanovskii37], most of which were devoted to the existence and concrete descriptions of extensions of C-symmetric differential operators. The reader is referred to [Reference Zagorodnyuk47] and the references cited therein for more historical comments. Throughout the following, we let C denote a conjugation on ${\mathcal {H}}$ unless otherwise stated, and let $\mathcal {S}_C$ denote the collection of all C-symmetric operators. The term “symmetric” stems from the fact that an operator T is complex symmetric if and only if there is an orthonormal basis $\{e_i\}$ of ${\mathcal {H}}$ such that T can be written as a (complex) symmetric matrix relative to $\{e_i\}$ .

Now let us introduce the background of the general study of c.s. operators, initiated in [Reference Garcia and Putinar14, Reference Garcia and Putinar15]. In the finite-dimensional case, the class of c.s. operators, containing those operators induced by Toeplitz matrices and Hankel matrices, has been studied for many years. In fact, the study of c.s. operators has classical roots in the work on automorphic functions [Reference Hua24], projective geometry [Reference Jacobson25], quadratic forms [Reference Schur39], symplectic geometry [Reference Siegel40], and function theory [Reference Takagi42]. In the infinite-dimensional case, it is shown that the class of c.s. operators contains many important special operators, such as normal operators, binormal operators, truncated Toeplitz operators [Reference Sarason38], and many integration operators. People’s current interests in c.s. operators are greatly inspired by many results of S. Garcia, M. Putinar, and W. Wogen [Reference Garcia and Putinar14Reference Garcia and Wogen16] as well as their connections to concrete operators [Reference Garcia10Reference Garcia12] and applications to mathematical physics [Reference Garcia, Prodan and Putinar13, Reference Hai and Putinar20, Reference Prodan, Garcia and Putinar34].

Also, c.s. operators play an important role in the study of JB $^{*}$ -triples, a class of complex Banach spaces with well-behaved algebraic, geometric and holomorphic properties. This is due to the Jordan structure of $\mathcal {S}_C$ . In fact, $\mathcal {S}_C$ is a weak operator closed subspace of $\mathcal {B(H)}$ , closed under the Jordan product $\circ $ , defined by

$$ \begin{align*}A\circ B=\frac{1}{2}(AB+BA),\ \ \ \forall A,B\in\mathcal{B(H)}. \end{align*} $$

$\mathcal {S}_C$ has been studied under the name of Hermitian type Cartan factors for many years. There are six types of Cartan factors, namely rectangular type, Hermitian type, symplectic type, triple spin factors, and two finite-dimensional exceptional Cartan factors. They originally arose in É. Cartan’s classification of finite-dimensional bounded symmetric domains (see [Reference Cartan2] or [Reference Chu3, Theorem 2.5.9]) and play an important role in the proof of the Gelfand–Naimark theorem for JB $^{*}$ -triples [Reference Friedman and Russo9].

The present study is inspired by several interesting results concerning $\mathcal {S}_C$ , which suggest a rich structure of $\mathcal {S}_C$ . In [Reference Garcia11], it was proved that each contraction $T\in \mathcal {S}_C$ is the mean of two unitary operators in $\mathcal {S}_C$ . In [Reference Wang and Zhu44], the authors classified Jordan ideals of $\mathcal {S}_C$ , showing that Jordan ideals of $\mathcal {S}_C$ arise by intersection from associative ideals of $\mathcal {B(H)}$ and hence are self-adjoint. Moreover, Jordan automorphisms of $\mathcal {S}_C$ are shown to be induced by certain unitary operators. Also, it is proved that those Jordan invertible ones constitute a dense, path connected subset of $\mathcal {S}_C$ . These results provide interesting contrasts between $\mathcal {S}_C$ and $\mathcal {B(H)}$ .

1.2 Range inclusion in $\mathcal {S}_C$

The first aim of this paper is to characterize the range inclusion of operators in $\mathcal {S}_C$ , that is, given an operator $T\in \mathcal {S}_C$ , to characterize all those operators $A\in \mathcal {S}_C$ satisfying $\mathcal {R}(A)\subseteq \mathcal {R}(T)$ , where $\mathcal {R}(A)$ denotes the range of A.

In [Reference Douglas6], R. Douglas proved the following classical result on the range inclusion for bounded linear operators on Hilbert spaces.

Theorem 1.1 [Reference Douglas6]

Let A and B be two bounded linear operators on ${\mathcal {H}}$ . Then the following statements are equivalent:

  1. (i) $\mathcal {R}(A)\subseteq \mathcal {R} (B)$ ;

  2. (ii) $AA^{*}\leq \lambda BB^{*}$ for some $\lambda \geq 0$ ;

  3. (iii) $A=BX$ for some operator $X\in \mathcal {B(H)}$ .

The preceding result exhibits a close relationship among the notions of range inclusion, majorization, and factorization for bounded linear operators. In [Reference Embry7], M. R. Embry extended this result to Banach spaces and obtained that $A=XB$ for some bounded operator X on $\mathcal {R} (B)$ if and only if $\mathcal {R} (A')\subseteq \mathcal {R}(B')$ . Here, $A'$ and $B'$ denote, respectively, the adjoint of A and the adjoint of B. P. H. Wang and X. Zhang [Reference Wang and Zhang45] established some range inclusion theorems for non-archimedean Banach spaces over general valued fields.

It is natural to ask whether there is a C-symmetric analogue of Douglas’ range inclusion theorem. Given a linear subspace ${\mathcal {M}}$ of ${\mathcal {H}}$ , we denote

$$ \begin{align*}C_{\mathcal{M}}=\{X\in\mathcal{S}_C: \mathcal{R}(X)\subset{\mathcal{M}}\}.\end{align*} $$

Let $T\in \mathcal {S}_C$ . We attempt to characterize $C_{\mathcal {R}(T)}$ . Note that $\mathcal {S}_C$ is not closed in the usual operator multiplication, that is, given $A,B\in \mathcal {S}_C$ , $\mathcal {R}(A)\subseteq \mathcal {R}(B)$ does not imply $B=AX$ for some $X\in \mathcal {S}_C$ (see Example 2.5). So we turn to the quadratic product in $\mathcal {S}_C$ given by $(X,Y)\longmapsto XYX$ for $X,Y\in \mathcal {S}_C.$ For $T\in \mathcal {S}_C$ , it is easy to check that

$$ \begin{align*}T\mathcal{S}_C T:=\{TXT: X\in\mathcal{S}_C\}\subseteq C_{\mathcal{R}(T)}.\end{align*} $$

It is natural to ask whether the converse inclusion holds.

The first result of this paper is the following theorem, which provides a complete answer to the preceding question.

Theorem 1.2 If $T\in \mathcal {S}_C$ , then $C_{\mathcal {R}(T)}=T\mathcal {S}_C T$ if and only if $\mathcal {R}(T)$ is closed.

The preceding result shows that if $T\in \mathcal {S}_C$ with non-closed range, then $T\mathcal {S}_C T$ is properly contained in $C_{\mathcal {R}(T)}$ . However, by the following result, they always have the same norm closure.

Theorem 1.3 If $T\in \mathcal {S}_C$ , then $\overline {C_{\mathcal {R}(T)}}=\overline {T\mathcal {S}_C T}$ .

By the preceding result, each $A\in \mathcal {S}_C$ with $\mathcal {R}(A)\subseteq {\mathcal {R}(T)}$ is a norm limit of operators with the form $TXT$ for $X\in \mathcal {S}_C$ . Thus Theorem 1.3 can be viewed as an approximate range inclusion theorem for $\mathcal {S}_C$ .

Given $A,B\in \mathcal {S}_C$ , it is interesting to compare the closure of $C_{\mathcal {R}(A)}$ with that of $C_{\mathcal {R}(B)}$ . To state our result, we introduce a notation. Let $\mathcal {M},\mathcal {N}$ be two linear subspaces of ${\mathcal {H}}$ . We write $\mathcal {M}\prec \mathcal {N}$ if for each closed subspace $\mathcal {M}_1$ of ${\mathcal {H}}$ contained in $\mathcal {M}$ and $\varepsilon>0$ there exists a closed subspace $\mathcal {N}_1$ of ${\mathcal {H}}$ contained in $\mathcal {N}$ such that $\|P_{\mathcal {M}_1}-P_{\mathcal {N}_1}\|<\varepsilon $ , where $P_{\mathcal {M}_1}$ denotes the orthogonal projection of ${\mathcal {H}}$ onto $\mathcal {M}_1$ .

Theorem 1.4 If $A,B\in \mathcal {S}_C$ , then $\overline {C_{\mathcal {R}(A)}}\subset \overline {C_{\mathcal {R}(B)}}$ if and only if $\mathcal {R}(A)\prec \mathcal {R}(B)$ .

Hence, $\overline {C_{\mathcal {R}(A)}}=\overline {C_{\mathcal {R}(B)}}$ if and only if $\mathcal {R}(A)\prec \mathcal {R}(B)$ and $\mathcal {R}(B)\prec \mathcal {R}(A)$ . Clearly, $\mathcal {R}(A)=\mathcal {R}(B)$ implies $\overline {C_{\mathcal {R}(A)}}=\overline {C_{\mathcal {R}(B)}}$ . In general, the converse does not hold (see Example 2.12).

Inspired by Theorem 1.3, we are interested in characterizing the closure of $T\mathcal {S}_C T$ in several usual topologies for $T\in \mathcal {S}_C$ , such as the weak* topology, the strong operator topology (sot) and the weak operator topology (wot). Given a subset $\mathcal {E}$ of $\mathcal {B(H)}$ , we write $\overline {\mathcal {E}}^{w*}$ , and to denote the the weak*-closure, the sot-closure and the wot-closure of $\mathcal {E}$ , respectively.

Theorem 1.5 If $T\in \mathcal {S}_C$ , then

In view of the preceding results, it is natural to classify the inclusion relations among $T\mathcal {S}_C T, C_{\mathcal {R}(T)}$ , $\overline {C_{\mathcal {R}(T)}}$ and $C_{\overline {\mathcal {R}(T)}}$ for $T\in \mathcal {S}_C$ . It is obvious that

$$ \begin{align*}T\mathcal{S}_C T\subseteq C_{\mathcal{R}(T)} \subseteq \overline{C_{\mathcal{R}(T)}}\subseteq C_{\overline{\mathcal{R}(T)}}.\end{align*} $$

In Section 2, we shall show that any two of these sets coincide if and only if T has a closed range (see Proposition 2.15).

1.3 Diagonalization in $\mathcal {S}_C$

The other aim of this paper is to study the extension of the Weyl–von Neumann Theorem to $\mathcal {S}_C$ .

The Weyl–von Neumann Theorem, due to H. Weyl and J. von Neumann [Reference von Neumann33, Reference Weyl46], states that, after the addition of a compact (or even Hilbert–Schmidt) operator of arbitrarily small norm, a self-adjoint operator becomes a diagonal operator. Recall that an operator $T\in \mathcal {B(H)}$ is diagonal if there is an orthonormal basis for ${\mathcal {H}}$ consisting of eigenvectors for T. In 1958, S. T. Kuroda [Reference Kuroda29] strengthened the result by proving that every self-adjoint operator is the sum of a diagonal operator and a compact operator with arbitrarily small Schatten p-norm for $p\in (1,\infty )$ .

We shall prove in Section 3 the following result, which can be viewed as an $\mathcal {S}_C$ -analogue of the Weyl–von Neumann Theorem.

Theorem 1.6 Let $T\in \mathcal {S}_C$ be a self-adjoint operator. If $p>1$ and $\varepsilon>0$ , then there is a diagonal, self-adjoint operator $D\in \mathcal {S}_C$ such that $T-D$ lies in the Schatten p-class $\mathcal {B}_p\mathcal {(H)}$ and $\|T-D\|_p<\varepsilon $ .

Remark 1.7

  1. (i) The proof of Theorem 1.6 is inspired by [Reference Conway4, Theorem 38.1]. The main difficulty stems from the fact that we always have to deal with not only the self-adjoint operator T but also the conjugation C at the same time.

  2. (ii) By [Reference Kato27, Theorem 1], if A is a self-adjoint operator not purely singular, then A is not a trace class perturbation of any diagonal operator. Thus the preceding result does not hold for $p=1$ .

  3. (iii) I. D. Berg [Reference Berg1] and W. Sikonia [Reference Sikonia41] independently proved that each normal operator T on ${\mathcal {H}}$ is the sum of a diagonal operator and a compact one with arbitrarily small norm. The preceding result is sometimes called the Weyl–von Neumann–Berg Theorem. In [Reference Halmos22], P. R. Halmos gave a more perspicuous proof that reduces the normal case to the Hermitian one. As an application of Theorem 1.6, we shall prove that finite commuting normal operators in $\mathcal {S}_C$ can be simultaneously diagonalized (see Theorem 3.3).

  4. (iv) In [Reference Voiculescu43], D. Voiculescu considered the simultaneous diagonalization of commuting self-adjoint operators and showed that if $T_1,\dots ,T_n (n\geq 2)$ are commuting self-adjoint operators and $\varepsilon>0$ , then there exist n commuting self-adjoint diagonal operators $D_1,\dots ,D_n$ such that $\|T_i-D_i\|_n<\varepsilon $ for all $i=1,\dots , n$ , where $\|\cdot \|_n$ is the Schatten n-norm. We do not know whether Voiculescu’s result can be extended to the $\mathcal {S}_C$ -setting.

As another application of Theorem 1.6, we shall provide a result concerning the irreducible approximation in $\mathcal {S}_C$ , which shows that those irreducible ones in $\mathcal {S}_C$ constitute a dense subset of $\mathcal {S}_C$ (see Corollary 3.6). This is an $\mathcal {S}_C$ -analogue of Halmos’ irreducible approximation theorem [Reference Halmos21].

The proofs of Theorems 1.21.5 will be given in Section 2. Section 3 is devoted to the proof of Theorem 1.6.

2 The range inclusion in $\mathcal {S}_C$

The aim of this section is to give the proofs of Theorems 1.21.5.

2.1 Proof of Theorem 1.2

Before giving the proof of Theorem 1.2, we first make some preparations.

Given a bounded linear operator A, we denote by $\ker A$ the kernel of A.

Lemma 2.1 Let $A, B\in \mathcal {S}_C$ . If $\mathcal {R}(A)\subset \mathcal {R}(B)$ , then

$$ \begin{align*}\mathcal{R}(A^{*})\subset \mathcal{R}(B^{*}),\ \ \ \ker{B}\subset\ker{A},\ \ \ \ker{B^{*}}\subset\ker{A^{*}}.\end{align*} $$

Proof Since $CAC=A^{*}$ and $CA^{*}C=A$ , one can see that $C(\mathcal {R}(A))=\mathcal {R}(A^{*})$ . Likewise, we have $C(\mathcal {R}(B))=\mathcal {R}(B^{*})$ . It follows immediately that $\mathcal {R}(A^{*})\subset \mathcal {R}(B^{*})$ . Thus $\mathcal {R}(B^{*})^\bot \subset \mathcal {R}(A^{*})^\bot $ , that is, $\ker {B}\subset \ker {A}$ .

On the other hand, $\mathcal {R}(A)\subset \mathcal {R}(B)$ implies $\mathcal {R}(B)^{\bot }\subset \mathcal {R}(A)^{\bot }$ and, equivalently, $\ker {B^{*}}\subset \ker {A^{*}}.$

Throughout the following, given a closed subspace $\mathcal {M}$ of ${\mathcal {H}}$ , we denote by $P_{\mathcal {M}}$ the orthogonal projection of ${\mathcal {H}}$ onto ${\mathcal {M}}$ .

Proof of Theorem 1.2

$\Longrightarrow $ ”. Assume that $C_{\mathcal {R}(T)}=T\mathcal {S}_C T$ . We shall show that $\mathcal {R}(T)$ is closed.

Assume that $\{x_n\}_{n=1}^\infty \subset {\mathcal {H}}$ and $Tx_n\rightarrow y_0\in {\mathcal {H}}$ . It suffices to show that $y_0\in \mathcal {R}(T)$ . Note that $T\in C_{\mathcal {R}(T)}$ . Thus there exists $A\in \mathcal {S}_C$ such that $T=TAT$ . Hence, $TATx_n\rightarrow y_0$ and $TATx_n\rightarrow TAy_0$ , which implies $y_0=TAy_0\in \mathcal {R}(T).$

$\Longleftarrow $ ”. Now assume that $\mathcal {R}(T)$ is closed. Since $T\mathcal {S}_C T \subseteq C_{\mathcal {R}(T)}$ is clear, it remains to show $ C_{\mathcal {R}(T)} \subseteq T\mathcal {S}_C T$ .

Denote $\mathcal {N}=(\ker {T})^{\bot }$ and $\mathcal {M}=\mathcal {R}(T)$ . For $x\in \mathcal {N}$ , define $Fx=Tx$ . Since $\mathcal {M}$ is norm-closed, F is an invertible operator from $\mathcal {N}$ into $\mathcal {M}$ . Define $S=P_{\mathcal {N}}F^{-1}P_{\mathcal {M}}$ . One can verify that $TS=P_{\mathcal {M}}$ and $ST=P_{\mathcal {N}}$ .

Since $T\in \mathcal {S}_C$ , one can see that $C(\ker T)=\ker T^{*}$ , that is, $C(\mathcal {N}^\bot )=\mathcal {M}^\bot $ . Thus $C(\mathcal {N})=\mathcal {M}$ and $C(\mathcal {M})=\mathcal {N}$ . Denote $C_1=C|_{\mathcal {M}}$ . Then $C_1:\mathcal {M}\longrightarrow \mathcal {N} $ is conjugate-linear, invertible, and isometric. One can check that $Cx=C^{-1}x= C_1^{-1}x$ for all $x\in \mathcal {N}.$

Claim 1 $C_1{(F^{-1})}^{*}=F^{-1}C_1^{-1}$ .

In fact, for any $x\in \mathcal { M} $ , note that

$$ \begin{align*}TCx=TC_1x=FC_1x, \ \ \ CT^{*}x=CF^{*}x=C_1^{-1}F^{*}x. \end{align*} $$

Thus $FC_1=C_1^{-1}F^{*}$ , ${(F^{-1})}^{*}C_1=C_1^{-1} F^{-1}$ and $C_1{(F^{-1})}^{*}= F^{-1}C_1^{-1}$ .

Claim 2 $S\in \mathcal {S}_C.$

For any $x\in {\mathcal {H}}$ , there is the unique decomposition $x=x_1+x_2$ , where $x_1\in \mathcal {N}$ and $x_2\in \mathcal {N}^{\bot }$ . So $Cx_1\in \mathcal {M}$ and $Cx_2\in \mathcal {M}^{\bot }$ . By Claim 1, we have

$$ \begin{align*} CSCx&=C(P_{\mathcal{N}}F^{-1}P_{\mathcal{M}})(Cx_1+Cx_2)\\ &=C(P_{\mathcal{N}}F^{-1}P_{\mathcal{M}})Cx_1\\ &=C(P_{\mathcal{N}}F^{-1}P_{\mathcal{M}})C_1^{-1}x_1\\ &=C(P_{\mathcal{N}}F^{-1} C_1^{-1}x_1)\\ &=C_1^{-1} (F^{-1} C_1^{-1}x_1) \\ &=(C_1^{-1}F^{-1} C_1^{-1})x_1= {(F^{-1})}^{*} x_1\\ &= P_{\mathcal{M}}{(F^{-1})}^{*} P_{\mathcal{N}}x =S^{*}x. \end{align*} $$

This proves $S\in \mathcal {S}_C.$

Set $X=SAS$ . Thus one can easily check that $X\in \mathcal {S}_C$ . We shall prove that $TXT=A$ . Note that $\mathcal {R}(A)\subset \mathcal {R} (T)= \mathcal {M}$ . It follows that

$$ \begin{align*}TXT=TSAST=P_{\mathcal{M}}AP_{\mathcal{N}}=AP_{\mathcal{N}}.\end{align*} $$

It suffices to prove that $AP_{\mathcal {N}}=A$ , that is, $A(I-P_{\mathcal {N}})=0$ .

Since $\mathcal {R}(A)\subset \mathcal {R}(T)$ , it follows from Lemma 2.1 that $\ker {T}\subset \ker {A}$ . That is, $\mathcal {N}^{\bot }\subset \ker A$ . Therefore we conclude that $A(I-P_{\mathcal {N}})=0$ .

Corollary 2.2 Let $T,A\in \mathcal {S}_C$ . If either $\mathcal {R}(T)$ or $\mathcal {R}(A)$ is closed, then $\mathcal {R}(A)\subset \mathcal {R}(T)$ if and only if $A=TXT$ for some $X\in \mathcal {S}_C$ .

Proof It suffices to prove the necessity.

$\Longrightarrow $ ”. By Theorem 1.2, we need only deal with the case that $\mathcal {R}(A)$ is closed.

Using Theorem 1.2 again, we can find $Z\in \mathcal {S}_C$ such that $A=AZA$ . Since $\mathcal {R} (A)\subset \mathcal {R} (T)$ , by Douglas’ range inclusion theorem, $A=TY$ for some $Y\in \mathcal {B(H)}$ . Noting that $A,T\in \mathcal {S}_C$ , we have

$$ \begin{align*} A&=CA^{*}C=C(TY)^{*}C=CY^{*}T^{*}C\\ &=(CY^{*}C)(CT^{*}C)=(CY^{*}C)T. \end{align*} $$

Then

$$ \begin{align*}A=AZA=(TY)Z(CY^{*}C)T=T[YZ(CY^{*}C)]T.\end{align*} $$

One can verify that $X:=YZ(CY^{*}C)\in \mathcal {S}_C$ , which completes the proof.

Remark 2.3 The result of Corollary 2.2 is sharp. In fact, given any $T\in \mathcal {S}_C$ , it holds that $\mathcal {R}(T)\subset \mathcal {R}(T)$ ; if, in addition, $\mathcal {R}(T)$ is not closed, then, by the proof for the necessity of Theorem 1.2, there exists no $X\in \mathcal {B(H)}$ such that $TXT=T$ . Here, we provide a concrete example of complex symmetric operator with non-closed range.

Let $\{e_n\}_{n=1}^\infty $ be an orthonormal basis of ${\mathcal {H}}$ and

$$\begin{align*}T=\sum_{n=1}^\infty \frac{e_n\otimes e_n}{n}.\end{align*}$$

For $x=\sum _{n=1}^\infty \alpha _i e_i$ , define $Cx=\sum _{n=1}^\infty \overline {\alpha _i} e_i$ . Then C is a conjugation on ${\mathcal {H}}$ and one can check that $T\in \mathcal {S}_C$ . Clearly, $\mathcal {R}(T)$ is not closed. If $X\in \mathcal {S}_C$ and $T=TXT$ , then

$$ \begin{align*}\frac{1}{n}=\langle Te_n, e_n \rangle=\langle TXTe_n, e_n \rangle=\langle XTe_n, Te_n \rangle=\frac{1}{n^2}\langle Xe_n, e_n \rangle,\end{align*} $$

that is, $\langle Xe_n, e_n \rangle = n$ . Since n is arbitrary, we deduce that X is not bounded, a contradiction.

For $e,f\in {\mathcal {H}}$ , we define an operator $e\otimes f$ on ${\mathcal {H}}$ as

$$\begin{align*}(e\otimes f)(x)=\langle x,f\rangle e, \ \ \ \forall x\in{\mathcal{H}}.\end{align*}$$

Corollary 2.4 Let $T\in \mathcal {S}_C$ . If $x,y\in \mathcal {R}(T)$ , then

$$ \begin{align*}x\otimes(Cx)\in T\mathcal{S}_C T \ \ \ \text{and}\ \ \ x\otimes(Cy)+y\otimes(Cx)\in T\mathcal{S}_C T.\end{align*} $$

Proof Denote $X=x\otimes (Cx)$ and $Y=x\otimes (Cy)+y\otimes (Cx)$ . It is easy to verify that $X\in \mathcal {S}_C$ and $\overline {\mathcal {R}(X)}=\mathcal {R}(X)\subset \mathcal {R}(T)$ . By Corollary 2.2, we have $X\in T\mathcal {S}_C T$ . Likewise we have $Y\in T\mathcal {S}_C T$ .

At the end of this subsection, we provide several illustrating examples.

The following example shows that Douglas’ range inclusion theorem does not hold for $\mathcal {S}_C$ .

Example 2.5 Let D be a conjugation on ${\mathcal {H}}$ and

$$\begin{align*}A=\begin{bmatrix} 0&I\\ I&0 \end{bmatrix}\begin{matrix} {\mathcal{H}}\\ {\mathcal{H}} \end{matrix}~,\ \ \ T=\begin{bmatrix} I&0\\ 0&-I \end{bmatrix}\begin{matrix} {\mathcal{H}}\\ {\mathcal{H}} \end{matrix}~, \ \ \ C=\begin{bmatrix} D&0\\ 0&D \end{bmatrix}\begin{matrix} {\mathcal{H}}\\ {\mathcal{H}} \end{matrix}~, \end{align*}$$

where I is the identity operator on ${\mathcal {H}}$ . Then it is easy to check that C is a conjugation on ${\mathcal {H}}\oplus {\mathcal {H}}$ , $A, T\in \mathcal {S}_C$ and $\mathcal {R}(A)\subset \mathcal {R}(T)$ .

If $X\in \mathcal {B}({\mathcal {H}}\oplus {\mathcal {H}})$ and $A=TX$ , then one can verify that

$$\begin{align*}X=T^{-1}A=\begin{bmatrix} 0&I\\ -I&0 \end{bmatrix}.\end{align*}$$

Clearly, $X\notin \mathcal {S}_C$ .

The following example shows that Theorem 1.2 can not be extended to $\mathcal {B(H)}$ or general Jordan operator algebras.

Example 2.6 Let $\mathcal {M}$ be a closed subspace of ${\mathcal {H}}$ with $\dim \mathcal {M}=\dim \mathcal {M}^{\bot }$ . Set

$$\begin{align*}T=\begin{bmatrix} I_0&0\\ 0&0 \end{bmatrix}\begin{matrix} \mathcal{M} \\ \mathcal{M}^{\bot} \end{matrix}\ \ \ \text{and} \ \ \ A=\begin{bmatrix} 0&B\\ 0&0 \end{bmatrix}\begin{matrix} \mathcal{M} \\ \mathcal{M}^{\bot} \end{matrix}, \end{align*}$$

where $I_0$ is the identity operator on $\mathcal {M}$ and B is an invertible operator from $\mathcal {M}^{\bot }$ onto $\mathcal {M}$ . Then $\mathcal {R}(A)\subset \mathcal {R}(T)$ .

For any $X\in \mathcal {B(H)}$ with

$$\begin{align*}X= \begin{bmatrix} X_{1,1}&X_{1,2}\\ X_{2,1}&X_{2,2} \end{bmatrix}\begin{matrix} \mathcal{M} \\ \mathcal{M}^{\bot} \end{matrix},\end{align*}$$

one can check that

$$\begin{align*}TXT= \begin{bmatrix} X_{1,1}&0\\ 0&0 \end{bmatrix}\begin{matrix} \mathcal{M} \\ \mathcal{M}^{\bot} \end{matrix} \ne A.\end{align*}$$

Hence, $A\ne TXT$ for any $X\in \mathcal {B(H)}$ .

2.2 Proof of Theorem 1.3

Let $\mathcal {M}$ be a closed subspace of ${\mathcal {H}}$ . A conjugate-linear map J on ${\mathcal {H}}$ is called a partial conjugation supported on $\mathcal {M}$ if $\ker J = \mathcal {M}^\bot $ reduces J and $J|_{\mathcal {M}}$ is a conjugation.

Lemma 2.7 [Reference Garcia and Putinar15, Theorem 2]

Let $T\in \mathcal {B(H)}$ and C be a conjugation on ${\mathcal {H}}$ . Then $T\in \mathcal {S}_C$ if and only if $T=CJ|T|$ for some partial conjugation J supported on $\mathcal {R}(|T|)$ and commuting with $|T|$ .

Proof of Theorem 1.3

Since it is clear that $T\mathcal {S}_C T\subset C_{\mathcal {R}(T)}$ , it suffices to prove that $C_{\mathcal {R}(T)}\subset \overline {T\mathcal {S}_C T}$ . Fix an operator A in $C_{\mathcal {R}(T)}$ . We shall prove that $A\in \overline {T \mathcal {S}_C T}$ . This follows readily from the following claim.

Claim Given $\varepsilon>0$ , there exists $B\in \mathcal {S}_C$ with $\mathcal {R}(B)=\overline {\mathcal {R}(B)}\subset \mathcal {R}(A)$ such that $\|A-B\|<\varepsilon $ .

In fact, if the preceding claim holds, then, by Corollary 2.2, $B\in T\mathcal {S}_C T$ and $\text {dist}(A, T\mathcal {S}_C T)<\varepsilon $ . Since $\varepsilon>0$ was arbitrary, we deduce that $A\in \overline {T\mathcal {S}_C T}$ .

By Lemma 2.7, $A=CJP$ , where $P=|A|$ and J is a partial conjugation supported on $\overline {\mathcal {R}(P)}$ with $JP=PJ$ .

Now we fix an $\varepsilon>0$ and define three functions $f, g, h $ on $[0,\|A\|]$ as

$$ \begin{align*}f(t)=\begin{cases} 0,& t\in[0,\varepsilon/2), \\ t,& t\in[\varepsilon/2, \|A\|],\end{cases} \ \ h(t)=\begin{cases} 0,& t\in[0,\varepsilon/2), \\ 1/t,& t\in[\varepsilon/2, \|A\|], \end{cases} \ \ \ g=fh.\end{align*} $$

It is easy to see that:

  1. (i) $f(P), g(P),$ and $h(P)$ are positive operators commuting with J,

  2. (ii) $g(P)$ is an orthogonal projection with $\mathcal {R}(f(P))=\mathcal {R}(g(P))=\mathcal {R}(h(P))$ , and

  3. (iii) $\mathcal {R}(g(P))$ reduces J.

Define $\tilde {J}=Jg(P)$ . Then $\tilde {J}$ is a partial anti-conjugation supported on $\mathcal {R}(h(P))$ commuting with $h(P)$ . Set $B=CJf(P)$ . Note that $Jf(P)=\tilde {J}P$ . It follows that ${B=C\tilde {J}f(P)}$ . By Lemma 2.7, we have $B\in \mathcal {S}_C$ . Note that

$$ \begin{align*}\overline{\mathcal{R}(f(P))}=\mathcal{R}(f(P))\subset \mathcal{R}(P).\end{align*} $$

Since J is supported on $\overline {\mathcal {R}(P)}$ , it follows that

$$ \begin{align*}\overline{\mathcal{R}(B)}= \mathcal{R}(B)\subset\mathcal{R}(A).\end{align*} $$

Note that

$$ \begin{align*}\|A-B\|=\|CJP-CJf(P)\|\leq \|P-f(P)\|<\varepsilon.\end{align*} $$

This completes the proof.

The following result can be seen from the proof of Theorem 1.3.

Corollary 2.8 Let $T\in \mathcal {S}_C$ . Then, given $\varepsilon>0$ , there exists $A\in \mathcal {S}_C$ with $\|T-A\|<\varepsilon $ such that $\mathcal {R}(A)=\overline {\mathcal {R}(A)}\subset \mathcal {R}(T)$ .

2.3 Proof of Theorem 1.4

The aim of this subsection is to prove Theorem 1.4.

A subset $\mathcal {E}$ of ${\mathcal {H}}$ is called an operator range if $\mathcal {E}=\mathcal {R}(A)$ for some $A\in \mathcal {B(H)}$ . We denote by $\text {Ran}(\mathcal {H})$ the set of all operator ranges in ${\mathcal {H}}$ . Then, by [Reference Fillmore and Williams8, Section 2], $\text {Ran}(\mathcal {H})$ is a lattice.

The following lemma shows that each operator range can be attained in $\mathcal {S}_C$ .

Lemma 2.9 Given $T\in \mathcal {B(H)}$ , there exists $A\in \mathcal {S}_C$ with $\mathcal {R}(A)=\mathcal {R}(T)$ .

Proof Let $T=U|T|$ be the polar decomposition of T, where U is a partial isometry. So $\mathcal {R}(T)=\mathcal {R} (U|T|U^{*})$ .

Denote $P=CU|T|U^{*}C$ . Clearly, P is positive and hence a complex symmetric operator. Then we can find a conjugation J on ${\mathcal {H}}$ so that $JP=PJ$ . Set $A=CJP$ . Then, by [Reference Gilbreath and Wogen17, Theorem 3.1], $A\in \mathcal {S}_C$ and

$$ \begin{align*}\mathcal{R}(A)=\mathcal{R}(CJP)=C(\mathcal{R} (JP))=C(\mathcal{R} (P))=\mathcal{R} (U|T|U^{*}C)=\mathcal{R}(T).\end{align*} $$

This ends the proof.

For $A\in \mathcal {B(H)}$ , we denote $A^t=CA^{*}C$ . Then it is easy to check the following.

  1. (a) $A\in \mathcal {S}_C$ if and only if $A=A^t$ .

  2. (b) $A\mathcal {S}_C A^t\subset \mathcal {S}_C$ .

Proposition 2.10 If $A,B\in \mathcal {B(H)}$ , then the following are equivalent:

  1. (i) $\mathcal {R}(A)\subset \mathcal {R}(B)$ ;

  2. (ii) $C_{\mathcal {R}(A)}\subset C_{\mathcal {R}(B)}$ ;

  3. (iii) $A\mathcal {S}_C A^t\subset B\mathcal {S}_C B^t$ .

Proof (i) $\Longrightarrow $ (ii). This is obvious.

(ii) $\Longrightarrow $ (i). For any $x\in \mathcal {R}(A)$ , by Corollary 2.4, we have

$$ \begin{align*}x\otimes (Cx)\in C_{\mathcal{R}(A)}\subset C_{\mathcal{R}(B)},\end{align*} $$

which implies $x\in \mathcal {R}(B)$ . Hence $\mathcal {R}(A)\subset \mathcal {R}(B)$ .

(i) $\Longrightarrow $ (iii). By Douglas’ range inclusion theorem, $\mathcal {R}(A)\subset \mathcal {R}(B)$ implies $A=BZ$ for some $Z\in \mathcal {B(H)}$ . Note that

$$ \begin{align*}A^t=CA^{*}C=C(BZ)^{*}C=CZ^{*}B^{*}C=(CZ^{*}C)(CB^{*}C)=Z^tB^t.\end{align*} $$

Thus, for any $X\in \mathcal {S}_C$ , we have

$$\begin{align*}AXA^t=(BZ)X(Z^tB^t) =B(ZXZ^t)B^t.\end{align*}$$

One can verify that $ZXZ^t\in \mathcal {S}_C$ . Thus $AXA^t\in B\mathcal {S}_C B^t$ .

(iii) $\Longrightarrow $ (i). For $x\in {\mathcal {H}}$ , denote $y=Ax$ . We shall prove that $y\in \mathcal {R}(B)$ . Note that

$$ \begin{align*} y\otimes (Cy)&=(Ax)\otimes(CAx)=(Ax)\otimes[(CAC)(Cx)]\\ &=A[x\otimes (Cx)](CAC)^{*} =A[x\otimes (Cx)]A^t\in A\mathcal{S}_C A^t. \end{align*} $$

So $y\otimes (Cy)\in B\mathcal {S}_C B^t$ . Then $y\otimes (Cy)=B X B^t$ for some $X\in \mathcal {S}_C$ . It follows that $y\in \mathcal {R}(B)$ .

Corollary 2.11 If $T\in \mathcal {B(H)}$ , then $C_{\mathcal {R}(T)}=T\mathcal {S}_C T^t$ if and only if $\mathcal {R}(T)=\overline {\mathcal {R}(T)}$ .

Proof By Lemma 2.9, there exists $A\in \mathcal {S}_C$ with $\mathcal {R}(A)=\mathcal {R}(T)$ . Then $C_{\mathcal {R}(A)}=C_{\mathcal {R}(T)}$ and, by Proposition 2.10, we have

$$ \begin{align*}T\mathcal{S}_C T^t=A\mathcal{S}_C A^t=A\mathcal{S}_C A.\end{align*} $$

Then $C_{\mathcal {R}(T)}=T\mathcal {S}_C T^t$ if and only if $C_{\mathcal {R}(A)}=A\mathcal {S}_C A$ . The desired result follows readily from Theorem 1.2.

Proof of Theorem 1.4

Denote ${\mathcal {M}}=\mathcal {R}(A)$ and $\mathcal {N}=\mathcal {R}(B)$ . By Theorem 1.3, $\overline {A\mathcal {S}_C A}=\overline {C_{{\mathcal {M}}}}$ and $\overline {B\mathcal {S}_C B}=\overline {C_{\mathcal {N}}}$ .

$\Longrightarrow $ ”. Choose a closed subspace $\mathcal {M}_1$ of ${\mathcal {M}}$ and an orthonormal basis $\{e_i\}_{i\in \Lambda }$ of $\mathcal {M}_1$ . Define $U=\sum _{i\in \Lambda } e_i\otimes (ce:i)$ . Then one can check that $U\in \mathcal {S}_C$ is a partial isometry with $\mathcal {R}(U)=\mathcal {M}_1$ , which implies $U\in C_{\mathcal {M}}\subset \overline {C_{\mathcal {N}}}$ . By the hypothesis, there exist $X_n\in C_{\mathcal {N}}$ , $n\geq 1$ , such that $X_n\rightarrow U$ .

By Corollary 2.8, we can find $\{Y_n\}\subset \mathcal {S}_C$ with $\mathcal {R}(Y_n)=\overline {\mathcal {R}(Y_n)}\subset \mathcal {N}$ such that $\|Y_n-X_n\|<\frac {1}{n}, n\geq 1$ . Then $Y_n\in C_{\mathcal {N}}$ and $Y_n\rightarrow U$ . Furthermore, we have $Y_nY_n^{*}\rightarrow UU^{*}$ .

Note that $UU^{*}=P_{\mathcal {M}_1}$ and $\mathcal {R}(Y_nY_n^{*})=\mathcal {R}(Y_n)$ is closed, $n\geq 1$ . There exists $\delta>0$ such that $\sigma (Y_nY_n^{*})\subset \{0\}\cup [\delta , \infty )$ . Define a continuous function f on $\{0\}\cup (\delta /2, \infty )$ as

$$ \begin{align*}f(t)=\begin{cases} 0,& t=0, \\ 1,& t\in(\delta/2, \infty).\end{cases}\end{align*} $$

Then $f(Y_nY_n^{*})\rightarrow f(UU^{*})=P_{\mathcal {M}_1}$ . Note that $f(Y_nY_n^{*})$ is an orthogonal projection with

$$ \begin{align*}\mathcal{R} (f(Y_nY_n^{*}))=\mathcal{R} (Y_nY_n^{*})=\mathcal{R}(Y_n)\subset\mathcal{N}.\end{align*} $$

This proves the necessity.

$\Longleftarrow $ ”. Clearly, it suffices to prove $C_{{\mathcal {M}}}\subset \overline {C_{\mathcal {N}}}$ . By Corollary 2.8, it suffices to prove that any $X\in \mathcal {S}_C$ with $\mathcal {R}(X)=\overline {\mathcal {R}(X)}\subset {\mathcal {M}}$ satisfies $X\in \overline {C_{\mathcal {N}}}$ .

Denote $\mathcal {M}_1=\mathcal {R}(X)$ and, for convenience, we write P for $P_{\mathcal {M}_1}$ . By Corollary 2.11, we have $X\in C_{\mathcal {M}_1}=P\mathcal {S}_C P^t$ . So $X=PYP^t$ for some $Y\in \mathcal {S}_C$ . On the other hand, by the hypothesis, we can find orthogonal projections $\{P_n\}$ with $\mathcal {R}(P_n)\subset \mathcal {N}$ such that $P_n\rightarrow P$ . So $P_nYP_n^t\rightarrow PYP^t=X$ . Noting that $P_nYP_n^t \in \mathcal {S}_C$ and $\mathcal {R}(P_nYP_n^t)\subset \mathcal {R}(P_n)\subset \mathcal {N}$ , we conclude that $P_nYP_n^t \in C_{\mathcal {N}}$ and $X\in \overline {C_{\mathcal {N}}}$ .

Next, we determine $\overline {T\mathcal {S}_C T}$ for certain compact operators T.

Example 2.12 Let $\{e_n\}_{n=1}^\infty $ be an orthonormal basis of ${\mathcal {H}}$ such that $Ce_n=e_n, n=1,2,\dots $ . We define two diagonal operators on ${\mathcal {H}}$ as

$$\begin{align*}A=\begin{bmatrix} 1&&&&\\ &\frac{1}{2^2}&&&\\ &&\frac{1}{3^2}&&\\ &&&\frac{1}{4^2}&\\ &&&&\ddots \end{bmatrix}\begin{matrix} e_1\\ e_2\\ e_3\\ e_4\\ \vdots \end{matrix} , \ \ \ \ B= \begin{bmatrix} 1&&&&\\ &\frac{1}{2 }&&&\\ &&\frac{1}{3 }&&\\ &&&\frac{1}{4 }&\\ &&&&\ddots \end{bmatrix}\begin{matrix} e_1\\ e_2\\ e_3\\ e_4\\ \vdots \end{matrix} .\end{align*}$$

It is clear that $A,B\in \mathcal {S}_C$ and $A=B^2$ . Thus $\mathcal {R}(A)\subset \mathcal {R}(B)$ . Noting that

$$ \begin{align*}\sum_n\frac{e_n}{n^2}=B(\sum_n\frac{e_n}{n})\in\mathcal{R}(B), \ \ \ \ \ \sum_n\frac{e_n}{n^2}\notin\mathcal{R}(A),\end{align*} $$

we obtain $\mathcal {R}(A)\subsetneq \mathcal {R}(B).$ Next, we shall show that $\overline {C_{\mathcal {R}(A)}}=\overline {C_{\mathcal {R}(B)}}$ or, equivalently, $\overline {A\mathcal {S}_C A}=\overline {B\mathcal {S}_C B}$ .

Since it is clear that $[{A\mathcal {S}_C A}\cup {B\mathcal {S}_C B}]\subset [\mathcal {S}_C\cap \mathcal {K(H)}],$ where $\mathcal {K(H)}$ is the ideal of operators in $\mathcal {B(H)}$ , it suffices to show $[\mathcal {S}_C\cap \mathcal {K(H)}]\subset [\overline {{A\mathcal {S}_C A}}\cap {\overline {B\mathcal {S}_C B}}]$ .

Arbitrarily choose an operator $T\in \mathcal {S}_C\cap \mathcal {K(H)}$ . For each $n\geq 1$ , we let $P_n$ denote the orthogonal projection of ${\mathcal {H}}$ onto $\vee \{e_1,\dots ,e_n\}$ . Note that $P_nTP_n\in \mathcal {S}_C$ with

$$ \begin{align*}\mathcal{R}(P_nTP_n)=\overline{\mathcal{R}(P_nTP_n)}\subset\mathcal{R}(A)\cap \mathcal{R}(B).\end{align*} $$

By Corollary 2.2, $P_nTP_n\in C_{\mathcal {R}(A)}\cap C_{\mathcal {R}(B)}$ . Note that $P_nTP_n\rightarrow T$ . We obtain $T\in \overline {{A\mathcal {S}_C A}}\cap {\overline {B\mathcal {S}_C B}}.$ Thus we have shown

$$\begin{align*}\overline{A\mathcal{S}_C A}=\overline{B\mathcal{S}_C B}=\mathcal{S}_C\cap\mathcal{K(H)}.\end{align*}$$

Remark 2.13 Let $A,B\in \mathcal {B(H)}$ and C be defined as in the preceding example. Set

$$ \begin{align*}\widetilde{C}=C\oplus C, \ \ \ R=A\oplus B, \ \ \ T=B\oplus A.\end{align*} $$

Then one can see that $R,T\in {\mathcal S}_{\widetilde {C}}$ with $\mathcal {R}(R)\nsubseteq \mathcal {R}(T) $ and $\mathcal {R}(T)\nsubseteq \mathcal {R}(R)$ . However, one can show that $\widetilde {C}_{\mathcal {R}(R)}$ and $\widetilde {C}_{\mathcal {R}(T)}$ have the same norm closure.

2.4 Proof of Theorem 1.5

We first make some preparations.

Lemma 2.14 Let $\mathcal {M}$ be a subspace of ${\mathcal {H}}$ . Then

$$ \begin{align*}C_{\overline{\mathcal{M}}}\cap\mathcal{K(H)}=\overline{C_{\mathcal{M}}}\cap\mathcal{K(H)}=\overline{C_{\mathcal{M}}\cap\mathcal{K(H)}}.\end{align*} $$

Proof One can easily check that $C_{\overline {\mathcal {M}}}$ is closed in the weak operator topology. It is clear that

$$ \begin{align*}[C_{\mathcal{M}}\cap\mathcal{K(H)}] \subset[\overline{C_{\mathcal{M}}}\cap\mathcal{K(H)}]\subset [C_{\overline{\mathcal{M}}}\cap\mathcal{K(H)}].\end{align*} $$

So it suffices to prove $[C_{\overline {\mathcal {M}}}\cap \mathcal {K(H)}] \subset \overline { C_{\mathcal {M}}\cap \mathcal {K(H)} }$ .

Now assume that $A\in C_{\overline {\mathcal {M}}}\cap \mathcal {K(H)}$ . We shall prove $A\in \overline {C_{\mathcal {M}}\cap \mathcal {K(H)}}$ . By [Reference Garcia and Putinar15, Theorem 3], A can be written as

$$ \begin{align*}A=\sum_{n=1}^\infty \lambda_n(Ce_n)\otimes e_n,\end{align*} $$

where $e_n$ are certain orthonormal eigenvectors of $|A| =|AA^{*}|^{1/2}$ and the $\lambda _n$ are the nonzero eigenvalues of $|A|$ , repeated according to multiplicity. Clearly, $\lambda _n\rightarrow 0$ and $Ce_n\in \mathcal {R}(A)\subset \overline {\mathcal {M}}$ . Then

$$ \begin{align*}\sum_{n=1}^m \lambda_n(Ce_n)\otimes e_n\longrightarrow A\ \ \ \ \ \ (m\rightarrow\infty).\end{align*} $$

So it remains to prove that $(Ce_n)\otimes e_n\in \overline {C_{\mathcal {M}}\cap \mathcal {K(H)}}$ for each $n\geq 1$ .

Since $Ce_n\in \overline {{\mathcal {M}}}$ , there exists a sequence $\{x_k\}_{k=1}^\infty \subset {\mathcal {M}}$ converging to $Ce_n$ . Clearly, $x_k\otimes (Cx_k)\in C_{{\mathcal {M}}}\cap \mathcal {K(H)}$ for all $k\geq 1$ . Note that $\{x_k\otimes (Cx_k)\}$ converges to ${(Ce_n)\otimes e_n}$ in the norm topology. We deduce that $(Ce_n)\otimes e_n \in \overline {C_{\mathcal {M}}\cap \mathcal {K(H)}}$ . Thus we complete the proof.

Proof of Theorem 1.5

We denote $\mathcal {M}={\mathcal {R}(T)}$ . It is trivial to see that $C_{\overline {{\mathcal {M}}}}$ is closed in both the weak operator topology and the weak* topology; moreover,

On the other hand, since $C_{{\mathcal {M}}}$ is a subspace of $\mathcal {B(H)}$ , it is clear that . Then it suffices to show $C_{\overline {{\mathcal {M}}}}\subset \overline {C_{{\mathcal {M}}}}^{w*}$ .

We need only consider the case that $\mathcal {M}$ is not closed. In this case, we can choose an orthonormal basis $\{e_i\}_{i=1}^\infty $ of $\overline {\mathcal {M}}$ . Define $U\in \mathcal {B(H)}$ as

$$\begin{align*}U=\sum_{i=1}^\infty e_{i}\otimes (ce_{i}).\end{align*}$$

Thus U is a partial isometry lying in $\mathcal {S}_C$ , $\mathcal {R}(U)=\overline {\mathcal {M}}$ and $(\ker U)^\bot =C(\overline {\mathcal {M}})$ . So $U\in C_{\overline {{\mathcal {M}}}}$ and it follows from Theorem 1.2 that $C_{\overline {{\mathcal {M}}}}=U\mathcal {S}_C U$ .

Fix an operator $A\in \mathcal {S}_C$ . Then it suffices to prove $UAU\in \overline {C_{{\mathcal {M}}}}^{w*}$ .

For the conjugation C, there is an orthonormal basis $\{f_i\}$ of ${\mathcal {H}}$ such that $Cf_i=f_i$ for all i (see [Reference Garcia, Prodan and Putinar13, Lemma 2.11]). For each $n\geq 1$ , set $P_n=\sum _{i=1}^nf_i\otimes f_i$ . It is easy to see $P_n\in \mathcal {S}_C$ . Noting that $P_nAP_n\in \mathcal {S}_C$ and, as $n\rightarrow \infty $ , $P_nAP_n$ converges to A in the . So $UP_nAP_nU$ converges to $UAU$ in the . Now it remains to prove the following claim.

Claim $UP_nAP_nU\in \overline {C_{{\mathcal {M}}}}^{w*}$ for each $n=1,2,\dots $ .

In fact, if the preceding claim holds, then, by [Reference Conway4, Proposition 20.3], $UAU\in \overline {C_{{\mathcal {M}}}}^{w*}$ .

Set $a_{j,i}=\langle Af_i, f_j\rangle $ for all $i,j\in \{1,2,\dots \}$ . Since $A\in \mathcal {S}_C$ , we have $a_{i,j}=a_{j,i}.$ Then

$$ \begin{align*}P_nAP_n=\sum_{i=1}^n a_{i,i} f_i\otimes f_i +\sum_{1\leq i<j\leq n} a_{i,j}(f_i\otimes f_j+f_j\otimes f_i) \end{align*} $$

and

(1) $$ \begin{align} &UP_nAP_nU \end{align} $$
(2) $$ \begin{align} &= \sum_{i=1}^n a_{i,i} (Uf_i)\otimes (U^{*}f_i) +\sum_{1\leq i<j\leq n} a_{i,j}[(Uf_i)\otimes (U^{*}f_j)+(Uf_j)\otimes (U^{*}f_i)] \end{align} $$
(3) $$ \begin{align} =& \sum_{i=1}^n a_{i,i} (Uf_i)\otimes (U^{*}Cf_i) +\sum_{1\leq i<j\leq n} a_{i,j}[(Uf_i)\otimes (U^{*}Cf_j)+(Uf_j)\otimes (U^{*}Cf_i)] \end{align} $$
(4) $$ \begin{align} =& \sum_{i=1}^n a_{i,i} (Uf_i)\otimes (CUf_i) +\sum_{1\leq i<j\leq n} a_{i,j}[(Uf_i)\otimes (CUf_j)+(Uf_j)\otimes (CUf_i)]. \end{align} $$

Note that $Uf_i, Uf_j\in \mathcal {R}(U)=\overline {{\mathcal {M}}}$ .

For any $x,y\in \overline {{\mathcal {M}}}$ , it follows from Lemma 2.14 that $x\otimes (Cy)+y\otimes (Cx)\in \overline {C_{{\mathcal {M}}}}.$ In view of (1), we obtain $UP_nAP_nU\in \overline {C_{{\mathcal {M}}}}\subset \overline {C_{{\mathcal {M}}}}^{w*}$ . This proves Claim and completes the proof.

The following result classifies the inclusion relations among $T\mathcal {S}_C T, {C_{\mathcal {R}(T)}},\overline {C_{\mathcal {R}(T)}}$ and $C_{\overline {\mathcal {R}(T)}}$ for $T\in \mathcal {S}_C$ .

Proposition 2.15 For $T\in \mathcal {S}_C$ , the following are equivalent:

  1. (i) $\mathcal {R}(T)=\overline {\mathcal {R}(T)}$ ;

  2. (ii) $C_{\mathcal {R}(T)}=\overline {C_{\mathcal {R}(T)}}$ ;

  3. (iii) ;

  4. (iv) $\overline {C_{\mathcal {R}(T)}}=C_{\overline {\mathcal {R}(T)}}$ ;

  5. (v) $C_{\mathcal {R}(T)}=A\mathcal {S}_C A$ for some $A\in \mathcal {S}_C$ ;

  6. (vi) $T\mathcal {S}_C T=C_{\mathcal {M}}$ for some subspace $\mathcal {M}$ of ${\mathcal {H}}$ ;

  7. (vii) $T\mathcal {S}_C T=\overline {T\mathcal {S}_C T}$ ;

  8. (viii) .

Proof “(i) $\Longrightarrow $ (iii), (iv) and (viii)”. Since $\mathcal {R}(T)$ is norm-closed, one can easily verify that $C_{\mathcal {R}(T)}$ is closed in the wot and, by Theorem 1.2, we have $C_{\mathcal {R}(T)}=T\mathcal {S}_C T$ . Hence $T\mathcal {S}_C T$ is closed in the wot.

“(viii) $\Longrightarrow $ (vii)” and “(iii) $\Longrightarrow $ (ii)”. Both are obvious.

“(vii) $\Longrightarrow $ (i)”. Since $T\mathcal {S}_C T$ is norm closed, by Theorem 1.3, $ T\mathcal {S}_C T= \overline {C_{\mathcal {R}(T)}} \supseteq C_{\mathcal {R}(T)} $ . Hence, $T\mathcal {S}_C T=C_{\mathcal {R}(T)}$ and, by Theorem 1.2, $\mathcal {R}(T)$ is closed.

“(ii) $\Longrightarrow $ (i)”. Choose an $x\in \overline {\mathcal {R}(T)}$ . Then there exists a sequence $\{x_n\}\in \mathcal {R}(T)$ converging to x. Clearly, $x_n\otimes (Cx_n)\in C_{\mathcal {R}(T)}$ for all n. Note that $\{x_n\otimes (Cx_n)\}$ converges to $x\otimes (Cx)$ in norm. Since $C_{\mathcal {R}(T)}$ is norm-closed, we deduce that $x\otimes (Cx)\in C_{\mathcal {R}(T)}$ . Thus $x\in \mathcal {R}(T)$ . This shows that ${\mathcal {R}(T)}=\overline {\mathcal {R}(T)}.$

“(i) $\Longrightarrow $ (vi)”. By Theorem 1.2, we have $T\mathcal {S}_C T=C_{\mathcal {R}(T)}$ .

“(vi) $\Longrightarrow $ (v)”. Assume that $T\mathcal {S}_C T=C_{\mathcal {M}}$ for some subspace $\mathcal {M}$ of ${\mathcal {H}}$ . It is clear that $T\mathcal {S}_C T\subset C_{\mathcal {R}(T)}$ . So $C_{\mathcal {M}}\subset C_{\mathcal {R}(T)}$ . Then, for any $x\in \mathcal {M},$ we have $x\otimes (Cx)\in C_{\mathcal {M}}\subset C_{\mathcal {R}(T)}$ and $x\in \mathcal {R}(T)$ . This shows $\mathcal {M}\subset \mathcal {R}(T)$ .

On the other hand, if $x\in \mathcal {R}(T)$ , then, by Corollary 2.4, $x\otimes (Cx)\in T\mathcal {S}_C T=C_{\mathcal {M}}$ , which implies $x\in \mathcal {M}$ . Hence $\mathcal {R}(T)\subset \mathcal {M}$ . We conclude that $\mathcal {R}(T)=\mathcal {M}$ and $T\mathcal {S}_C T=C_{\mathcal {R}(T)}$ .

“(v) $\Longrightarrow $ (i)”. Assume that $C_{\mathcal {R}(T)}=A\mathcal {S}_C A$ . Since $A\mathcal {S}_C A\subset C_{\mathcal {R}(A)}$ , it follows that $C_{\mathcal {R}(T)} \subset C_{\mathcal {R}(A)}$ . By Proposition 2.10, we have $ {\mathcal {R}(T)} \subset {\mathcal {R}(A)}$ .

On the other hand, if $x\in \mathcal {R}(A)$ , then, by Corollary 2.4, $x\otimes (Cx)\in A\mathcal {S}_C A=C_{\mathcal {R}(T)}$ , which implies $x\in \mathcal {R}(T)$ . Hence $ {\mathcal {R}(A)} \subset {\mathcal {R}(T)}$ and ${\mathcal {R}(A)} ={\mathcal {R}(T)}$ . By Proposition 2.10, we have $A\mathcal {S}_C A=T\mathcal {S}_C T$ . So $C_{\mathcal {R}(T)}=T\mathcal {S}_C T$ . By Theorem 1.2, we deduce that $\mathcal {R}(T)$ is closed.

“(iv) $\Longrightarrow $ (i)”. Denote ${\mathcal {M}}=\mathcal {R}(T)$ . We choose an orthonormal basis $\{e_i\}_{i=1}^\infty $ of $\overline {\mathcal {M}}$ . Define $U\in \mathcal {B(H)}$ as

$$\begin{align*}U=\sum_{i=1}^\infty e_{i}\otimes (ce_{i}).\end{align*}$$

Thus $U\in \mathcal {B(H)}$ is a partial isometry, $\mathcal {R}(U)=\overline {\mathcal {M}}$ and $ \mathcal {R}(U^{*})=(\ker U)^\bot =C(\overline {\mathcal {M}})$ . It is easy to check that $U\in \mathcal {S}_C$ . So $U\in C_{\overline {\mathcal {M}}}$ .

By the hypothesis, we can find $\{T_n\}\subset C_{\mathcal {M}}$ such that $T_n\longrightarrow U$ .

Since $\mathcal {R}(T_n)\subset \mathcal {M} \subset \overline {\mathcal {M}}=\mathcal {R}(U)$ , by Lemma 2.1, we have

$$ \begin{align*}(\ker T_n)^\bot=\overline{{\mathcal{R}(T_n^{*})}}\subset \mathcal{R}(U^{*})=(\ker U)^\bot=C(\overline{\mathcal{M}}).\end{align*} $$

Denote $\widetilde {U}=U|_{C(\overline {\mathcal {M}})}$ and $\widetilde {T}_n=(T_n)|_{C(\overline {\mathcal {M}})}$ . Thus $\widetilde {U},\widetilde {T}_n: C(\overline {\mathcal {M}})\rightarrow \overline {\mathcal {M}}$ are bounded linear operators and $\widetilde {U}$ is invertible.

Since $T_n\rightarrow U$ , it follows that $\widetilde {T}_n\rightarrow \widetilde {U}$ . By the stability of invertibility, we deduce that $\widetilde {T}_n$ is invertible for n large enough. This implies that $\mathcal {R}(T_n)=\mathcal {R}(\widetilde {T}_n)=\overline {\mathcal {M}}$ for n large enough. Since $\mathcal {R}(T_n)\subset \mathcal {M}$ , we obtain $\overline {\mathcal {M}}=\mathcal {M}$ .

Remark 2.16 By the preceding result and Theorem 1.2, if $T\in \mathcal {S}_C$ has a non-closed range, then

$$\begin{align*}T\mathcal{S}_C T\subsetneq C_{\mathcal{R}(T)}\subsetneq\overline{C_{\mathcal{R}(T)}} \subsetneq C_{\overline{\mathcal{R}(T)}}. \end{align*}$$

On the other hand, if T has a closed range, then all these sets coincide.

Corollary 2.17 Let $\mathcal {M}$ be a subspace of ${\mathcal {H}}$ . Then the following are equivalent:

  1. (i) $C_{\mathcal {M}}=T\mathcal {S}_C T$ for some $T\in \mathcal {S}_C$ ;

  2. (ii) $C_{\mathcal {M}}$ is norm closed;

  3. (iii) $C_{\mathcal {M}}$ is wot closed;

  4. (iv) $\mathcal {M}=\overline {\mathcal {M}}$ .

Proof The implications (iv) $\Longrightarrow $ (iii) $\Longrightarrow $ (ii) is obvious.

“(ii) $\Longrightarrow $ (iv)”. Using a similar argument as in the proof for (ii) $\Longrightarrow $ (i) of Proposition 2.15, one can see that $\mathcal {M}$ is closed.

“(i) $\Longrightarrow $ (iv)”. Assume that $C_{\mathcal {M}}=T\mathcal {S}_C T$ for some $T\in \mathcal {S}_C$ . From the proof for (vi) $\Longrightarrow $ (v) in Proposition 2.15, one can see that $T\mathcal {S}_C T=C_{\mathcal {R}(T)}$ and $\mathcal {M}=\mathcal {R}(T)$ . By Theorem 1.2, we deduce that ${\mathcal {M}}=\mathcal {R}(T)$ is closed.

“(iv) $\Longrightarrow $ (i)”. By Lemma 2.9, we can choose an operator $T\in \mathcal {S}_C$ with $\mathcal {R}(T)={\mathcal {M}}$ . Then, by Theorem 1.2, $T\mathcal {S}_C T=C_{\mathcal {R}(T)}=C_{\mathcal {M}}$ .

3 The diagonalization in $\mathcal {S}_C$

The aim of this section is to study the diagonalization of normal operators in $\mathcal {S}_C$ and give the proof of Theorem 1.6.

3.1 Proof of Theorem 1.6

We first give a key lemma.

Lemma 3.1 Let $T \in \mathcal {S}_C$ be a self-adjoint operator and $e\in {\mathcal {H}}$ . Then, given $\varepsilon>0$ and $p>1$ , there exist a finite-rank orthogonal projection $P\in \mathcal {S}_C$ with $e\in \mathcal {R}(P)$ and a finite-rank self-adjoint operator $K\in \mathcal {S}_C$ with $\|K\|_p<\varepsilon $ such that $P(T+K)=(T+K)P$ .

Proof Let E be the projection-valued spectral measure for T and assume that $\sigma (T)\subset [a,b]$ . We fix a positive integer n. Then there exist pairwise disjoint Borel subsets $\Delta _1,\Delta _2,\dots ,\Delta _n$ and $a_1,a_2,\dots ,a_n\in [a,b]$ such that

$$ \begin{align*}[a,b]=\cup_{i=1}^n \Delta_i \ \ \ \text{and}\ \ \ \Delta_i\subset[a_i-\frac{b-a}{2n}, a_i+\frac{b-a}{2n}].\end{align*} $$

For $i=1,2,\dots ,n$ , denote ${\mathcal {H}}_i=\mathcal {R} (E(\Delta _i))$ . Then each ${\mathcal {H}}_i$ reduces T and ${\mathcal {H}}=\oplus _{i=1}^n{\mathcal {H}}_i$ . Put $T_i=T|_{{\mathcal {H}}_i}$ . Then $T=\oplus _{i=1}^nT_i$ . For each i, choose $\lambda _i\in \Delta _i$ . Since $\sigma (T_i)\subset \Delta _i^-$ , it follows that $\|T_i-\lambda _i\|\leq (b-a)/n$ .

Claim For each i with $1\leq i\leq n$ , ${\mathcal {H}}_i$ reduces C.

Fix an i with $i=1,\dots ,n$ . It suffices to prove $CE(\Delta _i)C=E(\Delta _i)$ . Note that there exists a sequence $\{p_k\}_{k=1}^\infty $ of polynomials with real coefficients such that $p_k(T)\rightarrow E(\Delta _i)$ in the strong operator topology. Since $Cp_k(T)C=p_k(T)$ for all k, it follows immediately that $CE(\Delta _i)C=E(\Delta _i)$ . This proves the claim.

For each i, denote $C_i=C|_{{\mathcal {H}}_i}$ . Then $C_i$ is a conjugation on ${\mathcal {H}}_i$ and, from $CTC=T^{*}=T$ , we obtain $C_iT_iC_i=T_i$ .

Since ${\mathcal {H}}=\oplus _{i=1}^n{\mathcal {H}}_i$ , we may assume that $e=\sum _{i=1}^n e_i$ with $e_i\in {\mathcal {H}}_i$ . Then, by Claim, $Ce_i\in {\mathcal {H}}_i$ . For each i with $1\leq i\leq n$ , put ${\mathcal {M}}_i=\vee \{e_i, Ce_i\}$ . Then ${\mathcal {M}}_i$ is a subspace of ${\mathcal {H}}_i$ , reducing C, and $1\leq \dim {\mathcal {M}}_i\leq 2$ . Then, relative to the decomposition ${\mathcal {H}}_i={\mathcal {M}}_i\oplus ({\mathcal {H}}_i\ominus {\mathcal {M}}_i)$ , $T_i$ can be written as

$$ \begin{align*}T_i=\begin{bmatrix} A_i&F_i\\ G_i&B_i \end{bmatrix}\begin{matrix} {\mathcal{M}}_i\\ {\mathcal{H}}_i\ominus {\mathcal{M}}_i \end{matrix}.\end{align*} $$

Note that $G_i=F_i^{*}$ , since $T_i$ is self-adjoint.

Since $C_i({\mathcal {M}}_i)={\mathcal {M}}_i$ , relative to the decomposition ${\mathcal {H}}_i={\mathcal {M}}_i\oplus ({\mathcal {H}}_i\ominus {\mathcal {M}}_i)$ , $C_i$ can be written as

$$ \begin{align*}C_i=\begin{bmatrix} C_{i,1}&0\\ 0&C_{i,2} \end{bmatrix}\begin{matrix} {\mathcal{M}}_i\\ {\mathcal{H}}_i\ominus {\mathcal{M}}_i \end{matrix}.\end{align*} $$

Then both $C_{i,1}$ and $C_{i,2}$ are conjugations. It follows readily from $C_iT_iC_i=T_i$ that $C_iK_iC_i=K_i$ , where

$$ \begin{align*}K_i:=\begin{bmatrix} 0&F_i\\ G_i&0 \end{bmatrix}\begin{matrix} {\mathcal{M}}_i\\ {\mathcal{H}}_i\ominus {\mathcal{M}}_i \end{matrix}.\end{align*} $$

Clearly, $K_i$ is self-adjoint, $\text {rank} K_i\leq 4$ and

$$ \begin{align*} \|K_i\|&=\max\{ \|F_i\|, \|G_i\| \} \leq \left\|\begin{bmatrix} A_i-\lambda_i&F_i\\ G_i&B_i-\lambda_i \end{bmatrix}\right\| =\|T_i-\lambda_i\|\leq(b-a)/n. \end{align*} $$

Set $K=-\oplus _{i=1}^n K_i$ . Thus K is a self-adjoint, finite-rank operator, $CKC=K$ ,

$$ \begin{align*}T+K=\oplus_{i=1}^n (T_i-K_i)=\oplus_{i=1}^n (A_i\oplus B_i)\end{align*} $$

and

$$ \begin{align*}\|K\|_p\leq (\text{rank} K)^{1/p} \|K\| \leq (4n)^{1/p}\frac{b-a}{n}\leq \frac{4(b-a)}{n^{1/q}},\end{align*} $$

where $q=1/(1-1/p)$ . Thus, for any $\varepsilon>0$ , there exists n large enough such that $\|K\|_p<\varepsilon $ .

Denote by P the orthogonal projection of ${\mathcal {H}}$ onto $\oplus _{i=1}^n {\mathcal {M}}_i$ . Thus $e\in \mathcal {R}(P)$ , $\text {rank} P\leq 2n$ and $(T+K)P=P(T+K)$ ; indeed,

$$ \begin{align*}(T+K)|_{\mathcal{R}(P)}=\oplus_{i=1}^n A_i.\end{align*} $$

Noting that $CPC=P$ and $CKC=K$ , that are, $P,K\in \mathcal {S}_C$ , we complete the proof.

Now we are going to prove Theorem 1.6.

Proof of Theorem 1.6

Since C is a conjugation, we can find an orthonormal basis $\{e_n\}_{n=1}^\infty $ of ${\mathcal {H}}$ such that $Ce_n=e_n$ for all n.

Now fix an $\varepsilon>0$ and $p>1$ . By Lemma 3.1, we can find a finite-rank, self-adjoint operator $K_1\in \mathcal {S}_C$ with $\|K_1\|_p<\varepsilon /2$ and a finite-rank orthogonal projection $P_1\in \mathcal {S}_C$ such that $e_1\in \mathcal {R}(P_1)$ and $P_1(T+K_1)=(T+K_1)P_1$ . Denote ${\mathcal {H}}_1=\mathcal {R}(P_1)$ and $\widetilde {{\mathcal {H}}_1}={\mathcal {H}}\ominus {\mathcal {H}}_1$ . Then, with respect to the decomposition ${\mathcal {H}}={\mathcal {H}}_1\oplus \widetilde {{\mathcal {H}}_1}$ ,

(5) $$ \begin{align} T+K_1=\begin{bmatrix} T_1&0\\ 0& \widetilde{T_1} \end{bmatrix},\ \ \ C=\begin{bmatrix} C_1&0\\ 0& \widetilde{C_1} \end{bmatrix}, \end{align} $$

where $C_1T_1C_1=T_1$ and $\widetilde {C_1}\widetilde {T_1}\widetilde {C_1}=\widetilde {T_1}$ .

Now apply Lemma 3.1 again to the self-adjoint operator $\widetilde {T_1}$ and the vector $(I-P_1)e_2$ to get a finite-rank, self-adjoint operator $\widetilde {K_2}\in \mathcal {S}_{\widetilde {C_1}}$ and two subspaces ${\mathcal {H}}_2,\widetilde {{\mathcal {H}}_2}$ of $\widetilde {{\mathcal {H}}_1}$ such that

$$ \begin{align*}\|\widetilde{K_2}\|_p<\varepsilon/4, \ \ (I-P_1)e_2\in{\mathcal{H}}_2, \ \ \dim{\mathcal{H}}_2<\infty, \ \ \widetilde{{\mathcal{H}}_1}={\mathcal{H}}_2\oplus\widetilde{{\mathcal{H}}_2}\end{align*} $$

and with respect to which

$$\begin{align*}\widetilde{T_1}+\widetilde{K_2}=\begin{bmatrix} T_2&0\\ 0& \widetilde{T_2} \end{bmatrix},\ \ \ \widetilde{C_1}=\begin{bmatrix} C_2&0\\ 0& \widetilde{C_2} \end{bmatrix},\end{align*}$$

where $C_2T_2C_2=T_2$ and $\widetilde {C_2}\widetilde {T_2}\widetilde {C_2}=\widetilde {T_2}$ . In view of (5), we can find a finite-rank, self-adjoint operator $K_2\in \mathcal {S}_C$ with $\|K_2\|_p<\varepsilon /4$ such that

(6) $$ \begin{align} T+K_1+K_2=\begin{bmatrix} T_1&0&0\\ 0&T_2&0\\ 0& 0&\widetilde{T_2} \end{bmatrix}\begin{matrix} {\mathcal{H}}_1\\ {\mathcal{H}}_2\\ \widetilde{{\mathcal{H}}_2} \end{matrix},\ \ \ C=\begin{bmatrix} C_1&0&0\\ 0&C_2&0\\ 0&0 &\widetilde{C_2} \end{bmatrix}\begin{matrix} {\mathcal{H}}_1\\ {\mathcal{H}}_2\\ \widetilde{{\mathcal{H}}_2} \end{matrix}. \end{align} $$

Note that $e_1,e_2\in {\mathcal {H}}_1\oplus {\mathcal {H}}_2$ , $C_iT_iC_i=T_i$ $(i=1,2),$ and $\widetilde {C_2}\widetilde {T_2}\widetilde {C_2}=\widetilde {T_2}$ .

By induction, we can find a sequence of self-adjoint, finite-rank operators $\{K_i\}_{i=1}^\infty \subset \mathcal {S}_C$ and a sequence of pairwise orthogonal, finite-dimensional subspaces $\{{\mathcal {H}}_i\}$ such that for each $n\geq 1,$

  1. (i) $e_1,\dots ,e_n\in \oplus _{i=1}^n{\mathcal {H}}_i$ ,

  2. (ii) $\|K_n\|_p<\varepsilon /2^n$ , and

  3. (iii) relative to the decomposition ${\mathcal {H}}=(\oplus _{i=1}^n {\mathcal {H}}_i)\oplus \widetilde {{\mathcal {H}}_n}$ ,

    (7) $$ \begin{align} T+\sum_{i=1}^nK_i=(\oplus_{i=1}^n T_i)\oplus \widetilde{T_n},\ \ \ \ C=(\oplus_{i=1}^n C_i)\oplus \widetilde{C_n}, \end{align} $$
    where $\widetilde {{\mathcal {H}}_n}={\mathcal {H}}\ominus (\oplus _{i=1}^n {\mathcal {H}}_i)$ and
    $$ \begin{align*}\widetilde{C_n}\widetilde{T_n}\widetilde{C_n}=\widetilde{T_n}, \ \ \ C_iT_iC_i=T_i, \ \ \ i=1,2,\dots, n. \end{align*} $$

Since $\{e_i\}$ is an orthonormal basis of ${\mathcal {H}}$ , it follows from statement (i) that ${\mathcal {H}}=\oplus _{i=1}^\infty {\mathcal {H}}_i$ and hence $C=\oplus _{i=1}^\infty C_i$ . Set $K=\sum _{i=1}^\infty K_i$ . Then $K\in \mathcal {B}_p\mathcal {(H)}$ is self-adjoint, $\|K\|_p<\varepsilon $ and $T+K=\oplus _{i=1}^\infty T_i$ . Note that each $T_i$ is self-adjoint and acting on a finite-dimensional space. Thus $T+K$ is diagonal. Since $C_iT_iC_i=T_i$ for all i, it follows that $C(T+K)C=T+K$ . Set $D=T+K$ . This completes the proof.

Corollary 3.2 Let T be a self-adjoint operator in $\mathcal {S}_C$ . Then, given $\varepsilon>0$ , there exists a compact operator $K\in \mathcal {S}_C$ with $\|K\|<\varepsilon $ such that $T+K$ is diagonal with $\sigma (T+K)=\sigma (T)$ .

Proof Assume that $\sigma _p(T)=\{\lambda _i: i\in \Lambda \}$ . Set

$$ \begin{align*}{\mathcal{H}}_0=\vee_{i\in\Lambda}\ker(T-\lambda_i)\ \ \ \text{and}\ \ \ {\mathcal{H}}_1={\mathcal{H}}_0^\bot.\end{align*} $$

Since T is self-adjoint, we deduce that ${\mathcal {H}}_0$ reduces T. Then

$$ \begin{align*}T=\begin{bmatrix} T_0&0 \\ 0&T_1 \end{bmatrix}\begin{matrix} {\mathcal{H}}_0\\ {\mathcal{H}}_1 \end{matrix}.\end{align*} $$

Clearly, $T_0$ is diagonal and $\sigma _p(T_1)=\emptyset $ . Since $CTC=C$ , one can see $C({\mathcal {H}}_0)={\mathcal {H}}_0$ , which implies $C({\mathcal {H}}_1)={\mathcal {H}}_1$ and

$$ \begin{align*}C=\begin{bmatrix} C_0&0 \\ 0&C_1 \end{bmatrix}\begin{matrix} {\mathcal{H}}_0\\ {\mathcal{H}}_1 \end{matrix}.\end{align*} $$

It is easy to check that each $C_i$ is a conjugation and $T_i\in {\mathcal S}_{C_i}$ .

By Theorem 1.6, we can find $K_1\in {\mathcal S}_{C_1}$ with $\|K_1\|<\frac {\varepsilon }{2}$ such that $T_1+K_1$ is diagonal. Moreover, by the upper semi-continuity of spectrum, it can be required that $\sigma (T_1+K_1)\subset \sigma (T_1)+B(0,\frac {\varepsilon }{2})$ .

Since $T_1+K_1$ is diagonal, without loss of generality, we may assume that $\sigma _p(T_1+K_1)=\{z_i: i\geq 1\}$ . Denote $\mathcal {K}_i=\ker (T_1+K_1-z_i)$ for $i\geq 1$ . Then

$$ \begin{align*}T_1+K_1=\oplus_{i=1}^\infty z_i I_i\end{align*} $$

relative to the decomposition ${\mathcal {H}}_1=\oplus _{i=1}^\infty \mathcal {K}_i$ , where $I_i$ is the identity operator on $\mathcal {K}_i$ . Since $T_1+K_1\in \mathcal {S}_C$ is diagonal, it follows that

$$ \begin{align*}C_1(\ker(T_1+K_1-z_i))=\ker(T_1+K_1-z_i),\ \ \ \ i=1,2,3,\dots.\end{align*} $$

Then each $\mathcal {K}_i$ reduces $C_1$ . We may assume that $C_1=\oplus _{i=1}^\infty C_i'$ , where $C_i'=C_1|_{\mathcal {K}_i}$ is a conjugation on $\mathcal {K}_i$ .

Fix an $i\geq 1$ . If $z_i\in \sigma (T_1)$ , then we set $\widetilde {z_i}=z_i$ . If $z_i\notin \sigma (T_1)$ , then $z_i\notin \sigma _e(T_1)=\sigma _e(T_1+K_1)$ and hence $\dim \mathcal {K}_i<\infty $ ; noting that $z_i\in \sigma (T_1)+B(0,\frac {\varepsilon }{2})$ , we can find $\widetilde {z_i}\in \sigma (T_1)$ such that $|z_i-\widetilde {z_i}|=\text {dist}(z_i,\sigma (T_1))<\varepsilon /2$ .

Set $K_2=\oplus _{i=1}^\infty (\widetilde {z_i}-z_i) I_i $ . Then $K_2\in {\mathcal S}_{C_1}$ , $\|K_2\|\leq \varepsilon /2$ and

$$ \begin{align*}T_1+K_1+K_2=\oplus_{i=1}^\infty \widetilde{z_i} I_i.\end{align*} $$

Clearly, $\sigma (T_1+K_1+K_2)=\{\widetilde {z_i}: i\geq 1 \}^-\subseteq \sigma (T_1)$ .

Claim $K_2$ is compact.

It suffices to show $\lim _i|z_i-\widetilde {z_i}|=0.$ Otherwise, we can choose a subsequence $\{i_k\}_{k=1}^\infty $ of $\mathbb {N}$ such that $z_{i_k}\notin \sigma (T_1)$ , $z_{i_k}\rightarrow z_0$ and

$$ \begin{align*}\inf_k \text{dist}(z_{i_k},\sigma(T_1))>0.\end{align*} $$

Hence, $\text {dist}(z_{0},\sigma (T_1))>0$ and

$$ \begin{align*}z_0\in\sigma_e(T_1+K_1)=\sigma_e(T_1)\subset \sigma(T_1),\end{align*} $$

a contradiction.

Set $K=0\oplus (K_1+K_2)$ . Then $K\in \mathcal {S}_C$ is compact with $\|K\|<\varepsilon $ and

$$ \begin{align*}T+K=T_0\oplus (T_1+K_1+K_2).\end{align*} $$

By the preceding discussion, we have

$$ \begin{align*}\sigma(T+K)=\sigma(T_0)\cup \sigma(T_1+K_1+K_2)\subset \sigma(T_0)\cup \sigma(T_1)= \sigma(T).\end{align*} $$

It remains to show $\sigma (T)\subset \sigma (T+K)$ .

Note that K is compact. Hence

$$ \begin{align*}\sigma_e(T)=\sigma_e(T+K)\subset \sigma(T+K).\end{align*} $$

If $z\in \sigma (T)\setminus \sigma _e(T)$ , then $z\in \sigma _p(T)$ and $z=\lambda _i\in \sigma (T_0)\subset \sigma (T+K)$ for some $i\in \Lambda $ . Thus we have proved that $\sigma (T)\subset \sigma (T+K)$ , which implies $\sigma (T)= \sigma (T+K)$ .

Theorem 3.3 If $N_1, \dots ,N_n\in \mathcal {S}_C$ are commuting normal operators and $\varepsilon>0$ , then there are commuting, diagonal operators $D_1$ , $\dots $ , $D_n\in \mathcal {S}_C$ such that $N_i-D_i\in \mathcal {K(H)}$ and $\|N_i-D_i\|<\varepsilon $ for all $ i=1,2,\dots ,n$ .

Proof Denote by ${\mathcal {A}}$ the von Neumann algebra generated by $N_1$ , $\dots $ , $N_n$ . It is easy to see $\mathcal {A}\subset \mathcal {S}_C$ . By [Reference Davidson5, Lemma II.2.8], we can find a self-adjoint operator $A\in \mathcal {B(H)}$ generating ${\mathcal {A}}$ and $N_1$ , $\dots $ , $N_n\in C^{*}(A)$ . Clearly, $A\in \mathcal {S}_C$ . We can find continuous functions $f_1,\dots ,f_n$ on $\mathbb {R}$ such that $N_i=f_i(A)$ , $i=1,\dots ,n.$

Now fix an $\varepsilon>0$ . In view of [Reference Conway4, Lemma 39.5], there exists $\delta>0$ such that $\sup _{1\leq i\leq n}\|f_i(X)-f_i(Y)\|<\varepsilon $ for all self-adjoint operators $X,Y$ satisfying $\|X\|\leq \|A\|, \|Y\|\leq \|A\|$ and $\|X-Y\|<\delta $ .

By Corollary 3.2, we can find self-adjoint, diagonal $D\in \mathcal {S}_C$ with $\sigma (D)=\sigma (A)$ such that $D-A\in \mathcal {K(H)}$ and $\|D-A\|< \delta $ . Then $\|D\|=\|A\|$ and

$$ \begin{align*}\sup_{1\leq i\leq n}\|N_i-f_i(D)\|=\sup_{1\leq i \leq n}\|f_i(A)-f_i(D)\|<\varepsilon.\end{align*} $$

Note that $f_1(D)$ , $f_2(D)$ , $\dots $ , $f_n(D)$ are commuting diagonal operators. Also, since $D\in \mathcal {S}_C$ , it follows that $f_1(D),f_2(D)$ , $\dots $ , $f_n(D)$ $\in $ $\mathcal {S}_C$ .

Note that $A-D\in \mathcal {K(H)}$ . Using the functional calculus for self-adjoint operators, one can easily prove that $f_i(A)-f_i(D)\in \mathcal {K(H)}$ , $i=1,2,\dots ,n$ . This completes the proof.

Using a similar argument as in the proof of Corollary 3.2, one can prove the following.

Corollary 3.4 If $T\in \mathcal {S}_C$ is normal and $\varepsilon>0$ , then there exists a diagonal operator $N\in \mathcal {S}_C$ such that $\sigma (N)=\sigma (T)$ and $\|T-N\|<\varepsilon $ .

3.2 Irreducible approximation

We conclude this paper with an application of Theorem 1.6 to the irreducible approximation in $\mathcal {S}_C$ .

By a classical approximation result of P. R. Halmos [Reference Halmos21], the set of irreducible operators is a dense $G_\delta $ set and hence a topologically large subset of $\mathcal {B(H)}$ . H. Radjavi and P. Rosenthal [Reference Radjavi and Rosenthal36] gave a short proof of the density of irreducible operators in $\mathcal {B(H)}$ . It can be seen from their proof that each operator has an arbitrarily small compact perturbation which is irreducible (see [Reference Herrero23, Lemma 4.33]).

In a recent paper [Reference Liu, Zhao and Zhu31], T. Liu, J. Y. Zhao, and the last author studied irreducible approximation of c.s. operators and obtained the following result.

Theorem 3.5 [Reference Liu, Zhao and Zhu31, Theorem 2.1]

Let $T\in \mathcal {S}_C$ and $T=A+\text {i}B$ , where $A,B$ are self-adjoint. If A or B is diagonal, then, given $\varepsilon>0$ , there exists $K\in \mathcal {S}_C$ with $\|K\|<\varepsilon $ such that $T+K$ is irreducible.

In addition, it can be required that K lies in the Schatten p-class and $\|K\|_p<\varepsilon $ . Using the preceding result, it was proved in [Reference Liu, Zhao and Zhu31] that each c.s. operator has a small perturbation being irreducible. Moreover, the following question was raised:

Is every complex symmetric operator a compact perturbation or a small compact perturbation of irreducible ones?

Using Theorems 1.6 and 3.5, one can see the following result, which combining Halmos’ result shows that those irreducible ones in $\mathcal {S}_C$ constitute a dense $G_\delta $ subset of $\mathcal {S}_C$ .

Corollary 3.6 Given $T\in \mathcal {S}_C$ , $p\in (1,\infty )$ and $\varepsilon>0$ , there exists a compact operator $K\in \mathcal {S}_C$ with $\|K\|_p<\varepsilon $ such that $T+K$ is irreducible.

Acknowledgments

The authors are grateful to the referee for helpful comments and constructive suggestions concerning the manuscript.

Footnotes

The third author is the corresponding author and was partially supported by the National Natural Science Foundation of China (Grant No. 12171195)

References

Berg, I. D., An extension of the Weyl–von Neumann theorem to normal operators. Trans. Amer. Math. Soc. 160(1971), 365371.CrossRefGoogle Scholar
Cartan, É., Sur les domaines bornés homog‘enes de l’espace de $n$ variables complexes. Abh. Math. Semin. Univ. Hamburg 11(1935), 116162.CrossRefGoogle Scholar
Chu, C.-H., Jordan structures in geometry and analysis, Cambridge Tracts in Mathematics, 190, Cambridge University Press, Cambridge, 2012.Google Scholar
Conway, J. B., A course in operator theory, Graduate Studies in Mathematics, 21, American Mathematical Society, Providence, RI, 2000.Google Scholar
Davidson, K. R., ${C}^{\ast }$ -algebras by example, Fields Institute Monographs, 6, American Mathematical Society, Providence, RI, 1996.Google Scholar
Douglas, R. G., On majorization, factorization, and range inclusion of operators on Hilbert space. Proc. Amer. Math. Soc. 17(1966), 413415.CrossRefGoogle Scholar
Embry, M. R., Factorization of operators on Banach space. Proc. Amer. Math. Soc. 38(1973), 587590.CrossRefGoogle Scholar
Fillmore, P. A. and Williams, J. P., On operator ranges. Adv. Math. 7(1971), 254281.CrossRefGoogle Scholar
Friedman, Y. and Russo, B., The Gelfand–Naimark theorem for JB ${}^{\ast }$ -triples. Duke Math. J. 53(1986), no. 1, 139148.CrossRefGoogle Scholar
Garcia, S. R., Conjugation and Clark operators , In: A. L. Matheson, M. I. Stessin and R. M. Timoney (eds.), Recent advances in operator-related function theory, Proceedings of the conference held at Trinity College, Dublin, Ireland, August 4–6, 2004, Contemporary Mathematics, 393, American Mathematical Society, Providence, RI, 2006, pp. 67111.Google Scholar
Garcia, S. R., Means of unitaries, conjugations, and the Friedrichs operator. J. Math. Anal. Appl. 335(2007), 941947.CrossRefGoogle Scholar
Garcia, S. R., The norm and modulus of a Foguel operator. Indiana Univ. Math. J. 58(2009), no. 5, 23052315.CrossRefGoogle Scholar
Garcia, S. R., Prodan, E., and Putinar, M., Mathematical and physical aspects of complex symmetric operators. J. Phys. A.: Math. Gen. 47(2014), 353001.CrossRefGoogle Scholar
Garcia, S. R. and Putinar, M., Complex symmetric operators and applications. Trans. Amer. Math. Soc. 358(2006), 12851315.CrossRefGoogle Scholar
Garcia, S. R. and Putinar, M., Complex symmetric operators and applications, II. Trans. Amer. Math. Soc. 359(2007), 39133931.CrossRefGoogle Scholar
Garcia, S. R. and Wogen, W. R., Complex symmetric partial isometries. J. Funct. Anal. 257(2009), 12511260.CrossRefGoogle Scholar
Gilbreath, T. M. and Wogen, W. R., Remarks on the structure of complex symmetric operators. Integral Equations Operator Theory 59(2007), no. 4, 585590.CrossRefGoogle Scholar
Glazman, I. M., An analogue of the extension theory of Hermitian operators and a non-symmetric one-dimensional boundary problem on a half-axis. Dokl. Akad. Nauk SSSR 115(1957), 214216.Google Scholar
Glazman, I. M., Direct methods of qualitative spectral analysis of singular differential operators . In: Israel program for scientific translations, Jerusalem, 1965, Daniel Davey, New York, 1966, Translated from the Russian by the IPST staff.Google Scholar
Hai, P. V. and Putinar, M., Complex symmetric evolution equations. Anal. Math. Phys. 10(2020), no. 1, Article no. 14, 36 pp.CrossRefGoogle Scholar
Halmos, P. R., Irreducible operators. Michigan Math. J. 15(1968), 215223.CrossRefGoogle Scholar
Halmos, P. R., Continuous functions of Hermitian operators. Proc. Amer. Math. Soc. 31(1972), 130132.CrossRefGoogle Scholar
Herrero, D. A., Approximation of Hilbert space operators. Vol. 1, 2nd ed., Pitman Research Notes in Mathematics Series, 224, Longman Scientific & Technical, Harlow, 1989.Google Scholar
Hua, L.-K., On the theory of automorphic functions of a matrix level I. Geometrical basis. Amer. J. Math. 66(1944), 470488.CrossRefGoogle Scholar
Jacobson, N., Normal semi-linear transformations. Amer. J. Math. 61(1939), 4558.CrossRefGoogle Scholar
Kalinina, T. B., On the theory of extensions of $K$ -symmetric operators. Funkcionalniy analiz (Ulyanovsk) 7(1976), 7885 (Russian).Google Scholar
Kato, T., Perturbation of continuous spectra by trace class operators. Proc. Japan Acad. 33(1957), 260264.Google Scholar
Knowles, I., On the boundary conditions characterizing $J$ -selfadjoint extensions of $J$ -symmetric operators. J. Differential Equations 40(1981), 193216.CrossRefGoogle Scholar
Kuroda, S. T., On a theorem of Weyl-von Neumann. Proc. Japan Acad. 34(1958), 1115.Google Scholar
Li, V. P., On the theory of J-symmetric operators. Funkcionalniy analiz (Ulyanovsk) 3(1974), 8491 (Russian).Google Scholar
Liu, T., Zhao, J. Y., and Zhu, S., Reducible and irreducible approximation of complex symmetric operators. J. Lond. Math. Soc. 100(2019), 341360.CrossRefGoogle Scholar
Makarova, A. D., Extensions of $J$ -symmetric operators with a non-dense domain. Funkcionalniy analiz (Ulyanovsk), 8(1977), 102112 (Russian).Google Scholar
von Neumann, J., Charakterisierung des Spektrums eines Integraloperators. Actualités Sci. Indust. 229(1935), 120.Google Scholar
Prodan, E., Garcia, S. R., and Putinar, M., Norm estimates of complex symmetric operators applied to quantum systems. J. Phys. A 39(2006), 389400.CrossRefGoogle Scholar
Race, D., The theory of $J$ -selfadjoint extensions of $J$ -symmetric operators. J. Differential Equations 57(1985), 258274.CrossRefGoogle Scholar
Radjavi, H. and Rosenthal, P., The set of irreducible operators is dense. Proc. Amer. Math. Soc. 21(1969), 256.Google Scholar
Rayh, L. M. and Tsekanovskii, E. R., Biinvolutive self-adjoint biextensions of $J$ -symmetric operators. Teoriya funkciy, funcionalniy analiz i ih prilozheniya (Kharkov) 23(1975), 7993.Google Scholar
Sarason, D., Algebraic properties of truncated Toeplitz operators. Oper. Matrices 1(2007), 491526.CrossRefGoogle Scholar
Schur, I., Ein Satz ueber quadratische Formen mit komplexen Koeffizienten. Amer. J. Math. 67(1945), 472480.CrossRefGoogle Scholar
Siegel, C. L., Symplectic geometry. Amer. J. Math. 65(1943), 186.CrossRefGoogle Scholar
Sikonia, W., Essential, singular, and absolutely continuous spectra. Ph.D. thesis, University of Colorado at Boulder, 1970.Google Scholar
Takagi, T., On an algebraic problem related to an analytic theorem of Carathéodory and Fejér and on an allied theorem of Landau. Jpn. J. Math. 1(1925), 8393.CrossRefGoogle Scholar
Voiculescu, D., Some results on norm-ideal perturbations of Hilbert space operators. J. Operator Theory 2(1979), 337.Google Scholar
Wang, C. and Zhu, S., The Jordan algebra of complex symmetric operators. To appear in Chinese Annals of Mathematics (Series B). arXiv:1912.10391v2 Google Scholar
Wang, P. H. and Zhang, X., Range inclusion of operators on non-archimedean Banach space. Sci. China Math. 53(2010), no. 12, 32153224.CrossRefGoogle Scholar
Weyl, H., Über beschränkte quadratische formen deren differenz vollstetig ist. Rend. Circ. Mat. Palermo 27(1909), 373392.CrossRefGoogle Scholar
Zagorodnyuk, S. M., On a $J$ -polar decomposition of a bounded operator and matrices of $J$ -symmetric and $J$ -skew-symmetric operators. Banach J. Math. Anal. 4(2010), no. 2, 1136.CrossRefGoogle Scholar