Abstract

We prove some further properties of the operator 𝑇[𝑛QN] (𝑛-power quasinormal, defined in Sid Ahmed, 2011). In particular we show that the operator 𝑇[𝑛QN] satisfying the translation invariant property is normal and that the operator 𝑇[𝑛QN] is not supercyclic provided that it is not invertible. Also, we study some cases in which an operator 𝑇[2QN] is subscalar of order 𝑚; that is, it is similar to the restriction of a scalar operator of order 𝑚 to an invariant subspace.

1. Introduction

Although normality of operators (𝑇𝑇=𝑇𝑇) makes things easier, it rarely occurs and relaxing the normality condition is essential in the theory of operators on Hilbert spaces. One of the most important subclasses of the algebra of all bounded linear operators acting on a Hilbert space, the class of hyponormal operators, has been studied by many authors (see [1]). In recent years this class has been generalized, in some sense, to larger classes of the so-called 𝑝-hyponormal, log-hyponormal, posinormal, k-quasihyponormal classes, and so forth (see [26]).

In [7], Putinar showed that hyponormal operators are subscalar. This fact has led to far-reaching results, discovering deep properties of these operators. In this paper we extend that result to other generalized classes of operators.

Let 𝐻 be an infinite dimensional separable complex Hilbert space, let 𝐾 be a complex Hilbert space, and let (𝐻,𝐾) be the algebra of all bounded linear operators from 𝐻 to 𝐾. We write (𝐻) for (𝐻,𝐻). If 𝑇(𝐻,𝐾), we will write 𝑁(𝑇) and 𝑅(𝑇) for the null space (also referred to as the reducing subspace) and the range of 𝑇, respectively. The spectrum, the point spectrum, and the approximate point spectrum of an operator 𝑇 are denoted by 𝜎(𝑇),𝜎𝑝(𝑇),𝜎𝑎𝑝(𝑇), respectively. 𝑇 means the adjoint of 𝑇.

An operator 𝑇(𝐻) is(1)hyponormal if and only if 𝑇𝑇𝑇𝑇𝑇𝑥𝑇𝑥,forall𝑥𝐻,(2)posinormal if and only if 𝑅(𝑇)𝑅(𝑇), or equivalently 𝑇𝑇𝜆2𝑇𝑇 for some 𝜆>0,(3)𝑝-hyponormal if and only if (𝑇𝑇)𝑝(𝑇𝑇)𝑝,0<𝑝1,(4)𝑝-quasihyponormal if and only if 𝑇[(𝑇𝑇)𝑝(𝑇𝑇)𝑝]𝑇0,0<𝑝<1,(5)log-hyponormal if 𝑇 is invertible and log(𝑇𝑇)log(𝑇𝑇),(6)2-isometry if and only if 𝑇2𝑇22𝑇𝑇+𝐼=0 (see [8]), where 𝐼 indicates the identity operator that is, if and only if𝑇2𝑇2=2𝑇𝑇𝐼.(1.1) An operator 𝑇(𝐻) is said to be 𝑛-power quasinormal (abbreviated as 𝑛QN), 𝑛=1,2,, if𝑇𝑛𝑇𝑇=𝑇𝑇𝑇𝑛=𝑇𝑇𝑛+1.(1.2) If 𝑛=1, 𝑇 is called quasinormal. This class of operators being denoted by [𝑛QN], that is,[𝑛QN]=𝑇(𝐻)𝑇𝑛𝑇𝑇𝑇𝑇𝑇𝑛=0(1.3) was studied by the author [9].

𝑇 is called an 𝑚-partial isometry if 𝑇 satisfies𝑇𝐵𝑚(𝑇)=𝑇𝑚𝑘=0𝑚𝑘(1)𝑘𝑇𝑚𝑘𝑇𝑚𝑘=0,(1.4) where 𝐵𝑚(𝑇) is obtained formally from the binomial expansion of 𝐵𝑚(𝑇)=(𝑇𝑇𝐼)𝑚 by understanding (𝑇𝑇)𝑚𝑘=𝑇𝑚𝑘𝑇𝑚𝑘. The case when 𝑚=1 is called the partial isometries class. The class of 𝑚-partial isometries was defined by Saddi and Sid Ahmed [10] who proved some properties of the class. See Proposition 5.4.

This paper is divided into five sections. Section 2 deals with some preliminary facts concerning function spaces. Section 3 includes our main results. There we study some properties of [𝑛QN]. In particular we show that an operator 𝑇[𝑛QN] satisfying the translation invariant property is normal, and an invertible operator 𝑇[𝑛QN] and its inverse 𝑇1 have a common nontrivial invariant closed set provided that 𝑇[𝑛QN]. Also we show that some of class [2QN] satisfy an analogue of the single-valued extension for 𝑊𝑚2(𝐷,𝐻) and have scalar extension. In Section 4, we give some results about the Berberian extension. In Section 5, we shall use some properties of the approximate spectrum to obtain some results on single-valued extension (SVEP) (see Section 2) property for the 𝑚-partial isometries operators.

2. Spaces of Vector-Valued Functions

We will need the following function spaces.

Let 𝜆 be the coordinate in and let 𝑑𝜇(𝜆) denote the planar Lebesgue measure. Let 𝐷 be a bounded open subset of . We will denote by 𝐿2(𝐷,𝐻) the Hilbert space of square measurable (or summable) functions 𝑓𝐷𝐻 such that𝑓2,𝐷=𝐷𝑓(𝜆)2𝑑𝜇(𝜆)1/2<.(2.1)

Let 𝑂(𝐷,𝐻) denote the space of 𝐻-valued functions analytic on 𝐷, that is, 𝜕𝑓=𝜕𝑓/𝜕𝑧=0. Equipped with the topology of uniform convergence on compact subsets of 𝐷, 𝑂(𝐷,𝐻) is a Frechet space. Let𝐴2(𝐷,H)=𝐿2(𝐷,𝐻)𝑂(𝐷,𝐻)(2.2) denote the Bergman space for 𝐷 consisting of square measurable functions 𝑓 that are analytic on 𝐷.

We denote by 𝑃 the orthogonal projection of 𝐿2(𝐷,𝐻) onto 𝐴2(𝐷,𝐻).

Let us define now a special Sobolev type space. Let 𝑚 be a fixed nonnegative integer. The Sobolev space 𝑊𝑚2(𝐷,𝐻) of order 𝑚 of vector-valued functions with respect to 𝜕 will be the space of those functions 𝑓𝐿2(𝐷,𝐻) whose derivatives 𝜕𝑓,,𝜕𝑚𝑓 in the sense of distributions also belong to 𝐿2(𝐷,𝐻), that is,𝑊𝑚2(𝐷,𝐻)=𝑓𝐿2(𝐷,𝐻)𝜕𝑘𝑓𝐿2(𝐷,𝐻),for.𝑘=0,1,,𝑚(2.3)

Endowed with the norm𝑓2𝑊𝑚2=𝑚𝑘=0𝜕𝑘𝑓22,𝐷.(2.4)𝑊𝑚2(𝐷,𝐻) becomes a Hilbert space contained continuously in 𝐿2(D,𝐻); that is, there is a constant 0<𝐶< such that 𝑓𝐿2(𝐷,𝐻)𝐶𝑓𝑊𝑚2(𝐷,𝐻)forall𝑓𝑊𝑚2(𝐷,𝐻).

Let𝑇𝜆=𝑇𝜆=𝑇𝜆𝐼(2.5) for 𝜆 once and for all, whenever the definition is meaningful. We say that 𝑇 has the single-valued extension property at 𝜆0 (abbreviated SVEP at 𝜆0) if, for every open neighborhood 𝑈 of 𝜆0, the only analytic solution 𝑓 to the equation𝑇𝜆𝑓(𝜆)=(𝑇𝜆)𝑓(𝜆)=0(2.6) for all 𝜆 in 𝑈 is the constant function 𝑓0. We say that 𝑇 has SVEP if 𝑇 has a SVEP at every 𝜆.

It is easily seen that an operator 𝑇(𝐻) has SVEP if and only if, for each open 𝐷.

The operator 𝑇𝐷𝑂(𝐷,𝐻)𝑂(𝐷,𝐻) defined by𝑇𝐷(𝑓)(𝜆)=𝑇𝜆𝑓(𝜆)𝑓𝑂(𝐷,𝐻),𝜆𝐷.(2.7) is one to one.

Recall that, for a bounded operator 𝑇 on 𝐻, the local resolvent set 𝜌𝑇(𝑥) of 𝑇 at the point 𝑥𝐻 is defined as the union of all open subsets 𝐷 of such that there exists an analytic function 𝑓𝐷𝐻 which satisfies𝑇𝜆𝑓(𝜆)=𝑥𝜆𝐷.(2.8) The local spectrum 𝜎𝑇(𝑥) of 𝑇 at 𝑥𝐻 is the set defined by 𝜎𝑇(𝑥)=𝜌𝑇(𝑥) and obviously 𝜎𝑇(𝑥)𝜎(𝑇). It is clear from the definition that, 𝑇 has SVEP if and only if zero is the unique vector 𝑥𝐻 such that 𝜎𝑇(𝑥)= (see for more details [11]).

Recall that a bounded operator 𝑇(𝐻) is said to have the Bishop’s property (𝛽) if for every open subset 𝐷 of the complex plane and every sequence of analytic functions 𝑓𝑛𝐷𝐻 with the property that𝑇𝜆𝑓𝑛(𝜆)0as𝑛,(2.9) uniformly on all compact subsets of 𝐷, 𝑓𝑛(𝜆)0 as 𝑛 locally uniformly on 𝐷 or equivalently, for every open subset 𝐷 of , the operator 𝑇𝐷 defined in (2.7) is one to one and has the closed range [11, Proposition 3.3.5]. It is a very important notion in spectral theory. It is wellknown that every normal operator has Bishop’s property (𝛽).

A bounded operator 𝑇 on 𝐻 is called scalar of order 𝑚 if it possesses a spectral distribution of order 𝑚, that is, if there is a continuous unital morphism,Φ𝐶𝑚0()(𝐻),(2.10) such that Φ(𝑧)=𝑇, where 𝑧 stands for the identity function on and 𝐶𝑚0() for the space of compactly supported functions on , continuously differentiable of order 𝑚,0𝑚. An operator is subscalar if it is similar to the restriction of a scalar operator to an invariant subspace.

Let 𝑀𝑧 be the operator on 𝑊𝑚2(𝐷;𝐻) such that (𝑀𝑧𝑓)(𝑧)=𝑧𝑓(𝑧) for 𝑓𝑊𝑚2(𝐷;𝐻). This has a spectral distribution of order 𝑚, defined by the functional calculus Φ𝑀𝐶𝑚0()(𝑊𝑚2(𝐷,𝐻));Φ𝑀(𝑓)=𝑀𝑓. Therefore 𝑀𝑧 is a scalar operator of order 𝑚. Consider a bounded open disk 𝐷 which contains 𝜎(𝑇) and the quotient space𝑊𝐻(𝐷)=𝑚2(𝐷,𝐻)𝑇𝑧𝑊𝑚2(𝐷,𝐻)(2.11) endowed with the Hilbert space norm. We denote the class containing a vector 𝑓 or an operator A on 𝐻(𝐷) by 𝑓 or 𝐴, respectively. Let 𝑀𝑧 be the operator of multiplication by 𝑧 on 𝑊𝑚2(𝐷;𝐻). As noted above, 𝑀𝑧 is a scalar of order 𝑚 and has a spectral distribution Φ. Let 𝑀𝑆𝑧. Since 𝑇𝑧𝑊𝑚2(𝐷,𝐻) is invariant under every operator 𝑀𝑓; 𝑓𝐶𝑚0(), we infer that 𝑆 is a scalar operator of order 𝑚 with spectral distribution Φ. Consider the natural map 𝑉𝐻𝐻(𝐷) defined by 𝑉=[1], for 𝐻, where 1 denotes the constant function identically equal to . In [7], Putinar showed that if 𝑇(𝐻) is a hyponormal operator then 𝑉 is one to one and has closed range such that 𝑉𝑇=𝑆𝑉, and so 𝑇 is subscalar of order 𝑚.

3. Further Properties of the Class [𝑛QN]

We start this section with some properties of 𝑛-power quasinormal operators.

Theorem 3.1. The class [𝑛QN] has the following properties.(1)The class [𝑛QN] is closed under unitary equivalence and scalar multiplication.(2)If 𝑇 is of class [𝑛QN] and 𝑀 is a closed subspace of 𝐻 that reduces 𝑇, then 𝑇𝑀 (the restriction of𝑇 to 𝑀) is of class [𝑛QN].

Proof. (1) Let 𝑆(𝐻) be unitary equivalent to 𝑇. Then there is a unitary operator 𝑉(𝐻) such that 𝑇=𝑉𝑆𝑉 which implies that 𝑇=𝑉𝑆𝑉. Noting that 𝑇𝑛=𝑉𝑆𝑛𝑉 and inserting 𝐼=𝑉𝑉 suitably, we deduce from (1.2) that 𝑉𝑆𝑛𝑆𝑆𝑉=𝑇𝑛𝑇𝑇=𝑇𝑇𝑛+1=𝑉𝑆𝑆𝑛+1𝑉,(3.1) and (1.2) follows for 𝑆. Since (𝑇𝑀)Δ=𝑇Δ𝑀(3.2) for Δ as the 𝑛-th power or the adjoint, it follows that the left-hand side of (1.2) for (𝑇𝑀) reads 𝑇𝑛𝑇,𝑇𝑀(3.3) which is (𝑇𝑇𝑛+1𝑀)=(𝑇𝑀)(𝑇𝑀)𝑛+1, which is the right-hand side of (1.2). Thus, 𝑇𝑀 is of class [𝑛QN].

Next we characterize a matrix on a 2-dimensional complex Hilbert space which is in [𝑛QN]. Since every matrix on a finite dimensional complex Hilbert space is unitarily equivalent to an upper triangular matrix and an 𝑛-power quasinormal operator is unitarily invariant, it suffices to characterize an upper triangular matrix 𝑇. From the direct calculation, we get the following characterization.

Proposition 3.2. For 𝑛2 one has [𝑛𝑇=𝑥𝑦0𝑧QN]𝑥𝑦𝑥𝑛1+𝑧𝑥𝑛2++𝑧𝑛1𝑥=0,𝑦(𝑧𝑛𝑥𝑛)=0,𝑥𝑦(𝑧𝑛𝑥𝑛𝑥)𝑛1+𝑧𝑥𝑛2++𝑧𝑛1||𝑦||2+|𝑧|2|𝑥|2=0.(3.4)

We remark here that Proposition 3.2 offers the convenient criterion to find some examples of operators in [𝑛QN]. Also we observe that [𝑛QN] is not necessarily normal on a finite dimensional space.

Next couple of results show that [𝑛QN] does not have a translation invariant property.

Theorem 3.3 (see [9]). If 𝑇 and 𝑇𝐼 are of class [2QN], then 𝑇 is normal.

Theorem 3.4 (see [9]). If 𝑇 is of class [2QN][3QN] such that 𝑇𝐼 is of class [𝑛QN], then 𝑇 is normal.

It is natural to ask the following question: what is the operators in [𝑛QN] satisfying the translation invariant property? The answer to this question is provided by the following theorem.

Theorem 3.5. 𝑇𝜆 is of class [nQN] for every 𝜆 if and only if 𝑇 is a normal operator.

Proof. Assume that (𝑇𝜆) is of class [𝑛QN] for every 𝜆. Then (1.2) reads (𝑇𝜆)𝑛(𝑇𝜆)(𝑇𝜆)=(𝑇𝜆)(𝑇𝜆)(𝑇𝜆)𝑛, which reduces on eliminating the common factor 𝜆(𝑇𝜆)𝑛+1 to (𝑇𝜆)𝑛𝑇𝑇𝜆𝑇=𝑇𝑇𝜆𝑇(𝑇𝜆)𝑛.(3.5) By the binomial expansion, 𝑛𝑘=0(1)𝑘𝑛𝑘𝜆𝑘𝑇𝑛𝑘𝑇𝑇𝜆𝑇=𝑇𝑇𝜆T𝑛𝑘=0(1)𝑘𝑛𝑘𝜆𝑘𝑇𝑛𝑘,(3.6) whence by arranging terms suitably, the extremal terms vanishing in view of (1.2), 𝑛1𝑘=1(1)𝑘𝑛𝑘𝜆𝑘𝑇𝑛𝑘𝑇𝑇𝑇𝑇𝑇𝑛𝑘𝑛1𝑘=1(1)𝑘𝑛𝑘𝜆𝑘+1𝑇𝑛k𝑇𝑇𝑇𝑛𝑘=0.(3.7)
Now note that from the second summand in (3.7), we may extract the extremal term (1)𝑛𝑛𝜆𝑛(𝑇𝑇𝑇𝑇) and express it in terms of the remaining terms which contain 𝜆 to the power <𝑛. Hence dividing (3.7) by 𝜆𝑛 and letting 𝜆, we conclude that 𝑇𝑇𝑇𝑇0, whence the normality of 𝑇 follows.

Conversely it is known that normality is a translation invariant property; that is, if 𝑇 is normal, then (𝑇𝜆) is normal for every 𝜆, and hence (𝑇𝜆) is of class [𝑛QN].

The following proposition gives a characterization of an 𝑛-power quasinormal operator.

Proposition 3.6. Let 𝑇(𝐻), 𝐴=𝑇𝑛+𝑇𝑇, and 𝐵=𝑇𝑛𝑇𝑇. Then 𝑇 is of class [𝑛QN] if and only if 𝐴 commutes with 𝐵.

Proof. Commutativity of 𝐴 and 𝐵 is equivalent to (1.2).

Proposition 3.7. Let 𝑇,𝐴,𝐵 be as in Proposition 3.6. Then if 𝑇 is of class [𝑛QN], then 𝑇𝑛𝑇𝑇 commutes with 𝐴 and 𝐵.

Proof. By (1.2), 𝑇𝑛𝑇𝑇𝑇𝑛±𝑇𝑇=𝑇𝑛±𝑇𝑇𝑇𝑛𝑇𝑇.(3.8)
In general, the two classes [𝑛QN] and [(𝑛+1)QN] are not the same (see [9]).

Proposition 3.8. If 𝑇 is both of class [𝑛QN] and [(𝑛+1)QN], then it is of class [(𝑛+2)QN], that is, [nQN][(𝑛+1)QN][(𝑛+2)QN].

Proof. By (1.2), 𝑇𝑛+1𝑇𝑇=𝑇𝑛𝑇𝑇2(3.9) so that 𝑇𝑛+2𝑇𝑇 may be transformed into 𝑇𝑇𝑇𝑛+2.
It is known that if 𝑇 belongs to [𝑛QN] some 𝑛>0, 𝑇2 does not necessarily belong to the same class.

Theorem 3.9 (see [9]). If 𝑇 and 𝑇 are of class [𝑛QN], then 𝑇𝑛 is normal.

Proposition 3.10. If an operator 𝑇 of class [2QN] is a 2-isometry, then 𝑇2 is of class [𝑛QN] for all integers 𝑛2.

Proof. From Proposition 3.8 it suffices to prove that 𝑇2 is of class [2QN] and of class [3QN] because we may then proceed inductively.
Since 𝑇 is a 2-isometry, we have 𝑇4(𝑇2𝑇2)=𝑇4(2𝑇𝑇𝐼). Using 𝑇2𝑇𝑇=𝑇𝑇3, we may shift the power of 𝑇2 to the left, arriving at 𝑇4(𝑇2𝑇2)=𝑇2𝑇6; that is, 𝑇2 is of class [2QN].
In the same way, we may deduce that 𝑇6(𝑇2𝑇2)=𝑇2𝑇8 whence 𝑇2 is of class [3QN].

Lemma 3.11 (see [9]). If 𝑇 is of class [𝑛QN], then 𝑁(𝑇𝑛)𝑁(𝑇𝑛).

Proposition 3.12. If 𝑇 is both of class [𝑛QN] and [(𝑛+1)QN] such that 𝑇 is injective or 𝑇 is injective, then 𝑇 is quasinormal.

Proof. Since 𝑇 is of class [𝑛QN][(𝑛+1)QN], we have (3.9), which reads 𝑇𝑛(𝑇𝑇𝑇𝑇𝑇2)=0. If 𝑇 is injective, then so is 𝑇𝑛 and we have 𝑇𝑇𝑇𝑇𝑇2=0, whence 𝑇 is quasi-normal. If 𝑇 is injective, we may appeal to Lemma 3.11.

In the following theorem we prove some topological properties of the class [𝑛QN].

Theorem 3.13. The class [𝑛QN] is an arcwise-connected, closed subset of (𝐻) equipped with the uniform operator (norm) topology.

Proof. By Theorem 3.1, (1), we see that the ray 𝑎𝑇 in (𝐻) through 𝑇 is contained in [𝑛QN] for every complex number 𝑎, and therefore [𝑛QN] is arcwise-connected.
To see that [𝑛QN] is closed, we prove that any strong limit 𝑇(𝐻) of a sequence (𝑇𝑝) in [𝑛QN] also belongs to [𝑛QN]; that is, we let (𝑇𝑝) be a sequence of operators in [𝑛QN] converging to 𝑇(𝐻) in norm: 𝑇𝑝𝑥𝑇𝑥0as𝑝,foreach𝑥𝐻.(3.10) Hence it follows that 𝑇𝑝𝑥𝑇𝑥=𝑇𝑝𝑇𝑥𝑇𝑝𝑇𝑇𝑥=𝑝𝑇𝑥0,(3.11) whence (𝑇𝑝) converges strongly to 𝑇.
Since the product of operators is sequentially continuous in the strong topology (see [12, page 62]), one concludes that𝑇𝑛𝑝𝑇𝑝𝑇𝑝 converge strongly to 𝑇𝑛𝑇𝑇. Similarly 𝑇𝑝𝑇𝑝𝑛+1 converges strongly to 𝑇𝑇𝑛+1. Hence the limiting case of (1.2) shows that 𝑇 belongs to [𝑛QN], completing the proof.

Proposition 3.14. If 𝑇1,𝑇2,,𝑇𝑝 are of class [𝑛QN], then both the direct sum 𝑇1𝑇2𝑇𝑝 and the tensor product 𝑇1𝑇2𝑇𝑝 are of class [𝑛QN].

Proof. By the compatibility principle similar to (3.2), 𝑇1𝑇2Δ=𝑇Δ1𝑇Δ2,𝑇1𝑇2Δ=𝑇Δ1𝑇Δ2,(3.12) where Δ indicates either the 𝑛th power or adjoint, the proof follows.

A linear operator 𝑇 on 𝐻 is hypercyclic if there is a vector with dense orbit; that is, if there exists an 𝑥𝐻 such that orbit Orb(𝑇,𝑥)={𝑥,𝑇𝑥,𝑇2𝑥,} is dense in 𝐻, and in this case 𝑥 is called a hypercyclic vector for 𝑇.

An operator 𝑇 on 𝐻 is supercyclic if there exists a vector whose scaled orbit is dense; that is, if there exists an 𝑥𝐻 such that {𝜆𝑇𝑛𝑥,𝑛0,𝜆} is dense in 𝐻 and in this case 𝑥 is called a supercyclic vector for 𝑇.

Kitai [13] showed that hyponormal operators are not hypercyclic. We generalize Kitai’s theorem to the class [𝑛QN].

Proposition 3.15. If 𝑇 is of class [nQN] with 𝜎𝑝(𝑇𝑛), then 𝑇 is not hypercyclic.

Proof. If 𝑇 is hypercyclic, 𝑇𝑛 is hypercyclic, and hence 𝜎𝑝(𝑇𝑛)= by [13, corollary 2.4]. From Lemma 3.11 we have 𝜎𝑝(𝑇𝑛)𝜎𝑝(𝑇𝑛) and, hence, 𝜎𝑝(𝑇𝑛)=, a contradiction.

Theorem 3.16 (see [14, Theorem 2]). If 𝑇 is not hypercyclic, then 𝑇 and 𝑇1 have a common non-trivial invariant closed subset.

Proposition 3.17. If 𝑇 is of class [𝑛QN] and 0𝜎(𝑇), then 𝑇 and 𝑇1 have a common nontrivial closed invariant subset.

Proof . Since 𝑇 is of class [𝑛QN] and 0𝜎(𝑇), it follows that 𝑇𝑛 is normal, and hence (𝑇1)𝑛 is normal. By [13, Corollary 4.5] 𝑇𝑛 and (𝑇1)𝑛 have no hypercyclic vector. Thus by [15], neither 𝑇 or 𝑇1 has a hypercyclic vector. Therefore by [15] 𝑇 and 𝑇1 have a common nontrivial closed invariant subset. Hence Theorem 3.16 completes the proof.

Proposition 3.18. Operators 𝑇 that are of class [𝑛QN] such that 𝑇 is not invertible are not supercyclic.

Proof. Assume that 𝑇 is of class [𝑛QN] and supercyclic. Considering the class [𝑛QN] being closed under multiplication by nonzero scalars, we may assume that 𝑇=1. Since the supercyclic contraction 𝑇 satisfies property (𝛽), 𝜎(𝑇) is contained in the boundary 𝜕𝔻 of the unit disk 𝔻 [11, Proposition 3.3.18]. Thus 𝑇 is invertible, and we have a contradiction.

Definition 3.19. An operator 𝑇(𝐻) is algebraic if there is non-zero-polynomial 𝑝 such that 𝑝(𝑇)=0.

The following proposition shows that some quasinilpotent 𝑛-power quasi-normal operators are subscalar.

Proposition 3.20. If both 𝑇 and 𝑇 are of class [𝑛QN] such that 𝑇 is quasinilpotent, then 𝑇 is nilpotent, and hence 𝑇 is subscalar.

Proof. Since 𝑇 is quasinilpotent, 𝜎(𝑇)={0}. Hence by the spectral mapping theorem we get 𝜎(𝑇𝑛)=𝜎(𝑇)𝑛={0}. Thus 𝑇𝑛 is quasinilpotent and normal. So 𝑇𝑛=0;thatis,𝑇 is nilpotent, and 𝑇 is algebraic operator, and hence 𝑇 is subscalar.

Proposition 3.21 (see [7, Proposition 2.1]). For every bounded disk 𝐷 in , there is a constant 𝐶𝐷 such that for an arbitrary operator 𝑇(𝐻) and 𝑓𝑊𝑚2(𝐷,𝐻) we have (𝐼𝑃)𝜕𝑗𝑓2,𝐷𝐶𝐷(𝑇𝜆)𝜕𝑗+1𝑓2,𝐷+(𝑇𝜆)𝜕𝑗+2𝑓2,𝐷,𝑗=0,1,,𝑚2,(3.13) where 𝑃 is the orthogonal projection of 𝐿2(𝐷,𝐻) onto 𝐴2(𝐷,𝐻).

The next theorem is important for the proof of our main theorem, Theorem 3.27.

Theorem 3.22. Let 𝐷 be an arbitrary bounded disk in . If 𝑇 is of class [2QN] and 𝜎(𝑇)(𝜎(𝑇))=, then the operator 𝑇𝜆𝑊𝑚2(𝐷,𝐻)𝑊𝑚2(𝐷,𝐻)(3.14) is one to one.

Proof. Let 𝑔𝑊𝑚2(𝐷,𝐻) such that 𝑇𝜆𝑔=0, that is, 𝑇𝜆𝑔𝑊𝑚2=0.(3.15) Then for 𝑗=0,1,,𝑚 we have 𝑇𝜆𝜕𝑗𝑔2,𝐷=0.(3.16) Hence for 𝑗=1,,𝑚 we get 𝑇2𝜆2𝜕𝑗𝑔2,𝐷=0. Since 𝜎(𝑇)(𝜎(𝑇)) is empty then 𝑇2 is normal [9, Theorem 2.2]. Hence, (𝑇2𝜆2)𝜕𝑗𝑔2,𝐷=0.(3.17) Now we claim that (𝑇𝜆)𝜕𝑗𝑔2,𝐷=0.(3.18) Indeed, since 𝑇𝜆 is invertible for 𝜆𝐷𝜎(𝑇), (3.16) implies that 𝜕𝑗𝑔2,𝐷𝜎(𝑇)=0.(3.19) Therefore (𝑇𝜆)𝜕𝑗𝑔2,𝐷𝜎(𝑇)=0.(3.20)
Since 𝜎(𝑇)(𝜎(𝑇))= and 𝜎(𝑇)=𝜎(𝑇),(𝑇𝜆) is invertible for 𝜆𝜎(𝑇), therefore; from (3.17) we have (𝑇𝜆)𝜕𝑗𝑔2,𝜎(𝑇)=0.(3.21) It is clear form (3.20); and (3.21) that (𝑇𝜆)𝜕𝑗𝑔2,𝐷=0,for𝑗=0,1,,𝑚.(3.22) Thus Proposition 3.21 and (3.21) imply (𝐼𝑃)𝜕𝑗𝑔2,𝐷=0for𝑗=0,1,,𝑚2,(3.23) where 𝑃 denotes the orthogonal projection of 𝐿2(𝐷,𝐻) onto 𝐴2(𝐷,𝐻).
Hence (𝑇2𝜆2)𝑃𝑔=𝑇2𝜆2𝑔=0. Since 𝑇2 has SVEP, 𝑇 has SVEP. Also 𝑔=𝑃𝑔 is analytic and (𝑇𝜆)𝑔(𝜆)=0 for 𝜆𝐷. Hence 𝑔=0. Thus, 𝑇𝜆 is one to one.

Corollary 3.23. If 𝑇1 and 𝑇2 are of class [2QN] with 𝜎(𝑇𝑖)(𝜎(𝑇𝑖))=, for 𝑖=1,2 and 𝑇2𝑇1=0. Then 𝑇1+𝑇2𝜆𝑊𝑚2(𝐷,𝐻)𝑊𝑚2(𝐷,𝐻)(3.24) is one to one.

Proof. If 𝑓𝑊𝑚2(𝐷,𝐻) is such that (𝑇1+𝑇2)𝜆𝑓=0. Since 𝑇2𝑇1=0, we get (𝑇2𝜆)𝑇2𝑓=0. Since (𝑇2)𝜆 is one to one, 𝑇2𝑓=0. Hence, (𝑇1)𝜆𝑓=0. Since (𝑇1)𝜆 is one to one, 𝑓=0.

The following corollary shows that the nilpotent perturbation of operators in [2QN] satisfying SVEP satisfies SVEP.

Corollary 3.24. If an operator 𝑇(𝐻) is a nilpotent perturbation of a 2-power quasi-normal operator 𝑆, that is, 𝑇=𝑆+𝑁, where 𝑆 is of class [2QN], 𝑆 and 𝑁 commute, and 𝑁𝑚=0. If 𝜎(𝑆)(𝜎(𝑆))=, then 𝑇𝜆 is one-to-one.

Proof. If 𝑔𝑊𝑚2(𝐷,𝐻) is such that 𝑇𝜆𝑔=0, then 𝑆𝜆𝑔=𝑁𝑔.(3.25) Hence 𝑆𝜆𝑁𝑗1𝑔=𝑁𝑗𝑔 for 𝑗=1,2,,𝑚. We prove that 𝑁𝑗𝑔=0 for 𝑗=0,1,,𝑚1 by indication. Since 𝑁𝑚=0, 𝑆𝜆𝑁𝑚1𝑔=𝑁𝑚𝑔=0. Since 𝑆𝜆 is one-to-one by Theorem 3.22𝑁𝑚1𝑔=0. Assume it is true when 𝑗=𝑘, that is, 𝑁𝑘𝑔=0. From (3.25), we get 𝑆𝜆𝑁𝑘1𝑔=𝑁𝑘𝑔=0.(3.26) Since 𝑆𝜆 is one-to-one from Theorem 3.22, 𝑁𝑘1𝑔=0. By indication we have 𝑔=0. Hence 𝑇𝜆 is one-to-one.

An operator 𝑇(𝐻) is said to be the following.(1)It is left invertible if there is an operator 𝑆(𝐻) such that 𝑆𝑇=𝐼, where 𝐼 denotes the identity operator. The operator 𝑆 is called a left inverse of 𝑇.(2)It is right invertible if there is an operator𝑅(𝑋) such that 𝑇𝑅=𝐼. The operator 𝑅 is called a right inverse of 𝑇 (see [16]).

Corollary 3.25. If 𝑇 is of class [2QN] with the property 𝜎(𝑇)(𝜎(𝑇))=, and if 𝑆 is a left invertible operator with the left inverse 𝑅, then the operator (𝑆𝑇𝑅)𝜆𝑊𝑚2(𝐷,𝐻)𝑊𝑚2(𝐷,𝐻) is one-to-one.

Proof. If 𝑔𝑊𝑚2(𝐷,𝐻) is such that (STR)𝜆𝑔=0, then 𝑇𝜆𝑅𝑔=0.(3.27) Hence for 𝑗=0,1,,𝑚 we have 𝑇𝜆𝑅𝜕𝑗𝑔=0. From Theorem 3.22, we get 𝑅𝜕𝑗𝑔=0 for 𝑗=0,1,,𝑚.
Thus, STR𝜕𝑗𝑔=0 for 𝑗=0,1,,𝑚. It follows that 𝜆𝜕𝑗𝑔=0 for 𝑗=0,1,,𝑚. By application of [7, Proposition 2.1] with 𝑇=(0), we have (𝐼𝑃)𝑔2,𝐷=0,(3.28) where 𝑃 denotes the orthogonal projection of 𝐿2(𝐷,𝐻) onto the Bergman space 𝐴2(𝐷,𝐻). Hence 𝜆𝑔=𝜆𝑃𝑔=0. From [17, Corollary 10.7], there exists a constant 𝑐>0 such that 𝑐𝑃𝑔2,𝐷𝜆𝑃𝑔2,𝐷.(3.29) So 𝑔=𝑃𝑔=0. Thus, (STR)𝜆 is one-to-one.

Corollary 3.26. If 𝑇 is of class [2QN] with the property 𝜎(𝑇)(𝜎(𝑇))=, and if 𝑆 is a right invertible operator with the right inverse 𝑅 then the operator (𝑅𝑇𝑆)𝜆𝑊𝑚2(𝐷,𝐻)𝑊𝑚2(𝐷,𝐻), is one-to-one.

Now we are ready to prove our main theorem.

Theorem 3.27. If 𝑇 is of class [2QN] with the property that 𝜎(𝑇)(𝜎(𝑇))= and let 𝐷 be a bounded disk which contains 𝜎(𝑇), then the operator 𝑉𝐻𝐻(𝐷), defined by 𝑉𝑔=1𝑔+𝑇𝜆𝑊𝑚2(𝐷,𝐻)=[1𝑔],(3.30) is one to one and has closed range, where 𝐻(𝐷) is as in (2.11).

Proof. First, we will prove that if {𝑔𝑘}1𝐻 and {𝑓𝑘}1𝑊𝑚2(𝐷,𝐻) are sequences such that lim𝑘1𝑔𝑘+𝑇𝜆𝑓𝑘𝑊𝑚2=0,(3.31) then lim𝑘𝑔𝑘=0.
By the definition of the norm of a Sobolev space, (3.31) implies that lim𝑘𝑇𝜆𝜕𝑗𝑓𝑘2,𝐷=0for𝑗=1,,𝑚.(3.32) From (3.32) we get lim𝑘𝑇2𝜆2𝜕𝑗𝑓𝑘2,𝐷=0for𝑗=1,,𝑚.(3.33)
Since 𝑇2 is normal, lim𝑘(𝑇2𝜆2)𝜕𝑗𝑓𝑘2,𝐷=0for𝑗=1,,𝑚.(3.34) Since 𝑇𝜆 is invertible for 𝜆𝐷𝜎(𝑇), (3.32) implies that lim𝑘𝜕𝑗𝑓𝑘2,𝐷𝜎(𝑇)=0.(3.35) Therefore lim𝑘(𝑇𝜆)𝜕𝑗𝑓𝑘2,𝐷𝜎(𝑇)=0for𝑗=1,,𝑚.(3.36)
Since 𝜎(𝑇)(𝜎(T))= and 𝜎(𝑇)=𝜎(𝑇), it is clear that (𝑇𝜆) is invertible for 𝜆𝜎(𝑇). Therefore from (3.34), we have lim𝑘(𝑇𝜆)𝜕𝑗𝑓𝑘2,𝜎(𝑇)=0.(3.37) Hence, from (3.36) and (3.37) we get lim𝑘(𝑇𝜆)𝜕𝑗𝑓𝑘2,𝐷=0,𝑗=1,,𝑚.(3.38) Then by Proposition 3.21, we have lim𝑘(𝐼𝑃)𝜕𝑗𝑓𝑘2,𝐷=0,𝑗=1,,𝑚2.(3.39) By (3.31) and (3.39), we have lim𝑘1𝑔𝑘+𝑇𝜆𝑃𝑓𝑘2,𝐷=0.(3.40) Let Γ be a curve in 𝐷 surrounding 𝜎(𝑇). Then lim𝑘𝑃𝑓𝑘+𝑇𝜆11𝑔𝑘=0(3.41) uniformly for 𝜆Γ by (3.40). Hence by Riesz-Dunford functional calculus lim𝑘12𝜋𝑖Γ𝑃𝑓𝑘(𝑧)𝑑𝑧+𝑔𝑘=0.(3.42) But by Cauchy’s theorem Γ𝑃𝑓𝑘(𝑧)𝑑𝑧=0.(3.43) Hence lim𝑘𝑔𝑘=0. Thus the map 𝑉 is one-to-one and has closed range.

Corollary 3.28. If 𝑇 is of class [2QN] with the property that 𝜎(𝑇)(𝜎(𝑇))=, then T is subscalar of order 𝑚2.

Proof. Consider an arbitrary bounded open disk 𝐷 in that contains 𝜎(𝑇) and the quotient space given in (2.11).
Let 𝑀𝜆 be the multiplication operator by 𝜆 on 𝑊𝑚2(𝐷,𝐻). Then 𝑀𝜆 is a scalar operator of order 𝑚, and its spectral distribution is Φ𝑀𝐶𝑚0𝑊()𝑚2(𝐷;𝐻),Φ𝑀(𝑓)=𝑀𝑓,(3.44) where 𝑀𝑓 is the multiplication operator by 𝑓𝐶𝑚0(). Let 𝑀𝑆𝜆. Since 𝑇𝜆𝑊𝑚2(𝐷;𝐻) is invariant under every operator 𝑀𝑓, we infer that 𝑆 is a scalar operator of order 𝑚 with spectral distribution Φ.
Let 𝑉 be the operator 𝑉𝑔=1𝑔+𝑇𝜆𝑊𝑚2(𝐷,𝐻)(3.45) from 𝐻 into 𝐻(𝐷). Then we have the following commutative diagram 975745.equation.001(3.46) By the previous theorem the operator 𝑉 is a topological isomorphism of 𝐻 into 𝑅(𝑉). The relation 𝑉𝑇=𝑆𝑉 shows that 𝑅(𝑉) is 𝑆-invariant. Hence 𝑆 is an extension of the operator 𝑉, so this operator is subscalar. Since 𝑉 is invertible on 𝑅(𝑉), then the operator 𝑇 is subscalar of order 𝑚. On the other hand from [18, Theorem 4.3] we deduce that 𝑚2 and the theorem is proved.

Corollary 3.29. If 𝑇 is of class [2QN] with the property that 𝜎(𝑇)(𝜎(𝑇))=, then 𝑇 has Bishop’s property (𝛽).

Proof. It follows from Corollary 3.28 and [18, Lemma 2.1].

In [19] the authors study some operators with the single-valued extension property. In the following propositions we extend some of these results to operators with the Bishop’s property (𝛽).

Proposition 3.30. Let 𝑇(𝑘𝑖=1𝐻) be the following 𝑘×𝑘 triangular operator matrix 𝑇𝑇=11𝑇12𝑇13𝑇1𝑘0𝑇22𝑇23𝑇2𝑘00𝑇33𝑇3𝑘0000𝑇𝑘𝑘.(3.47) Assume that 𝑇𝑖𝑖 is of class [2QN] and satisfies 𝜎(𝑇𝑖𝑖)(𝜎(𝑇𝑖𝑖))= for 𝑖=1,2,,𝑘1 and 𝑇𝑘𝑘 is nilpotent. Then 𝑇 has Bishop’s property (𝛽).

Proof. Let 𝑓𝑝=𝑘𝑖=1𝑓𝑖𝑝𝐷𝑘𝑖=1𝐻 be a sequence of analytic functions such that 𝑇𝜆𝑓𝑝(𝜆)0 uniformly on every compact subset 𝐾 of an open set 𝐷 of ; then we have 𝑇11𝜆𝑓1𝑝(𝜆)+𝑇12𝑓2𝑝(𝜆)+𝑇13𝑓3𝑝(𝜆)+𝑇1𝑘𝑓𝑘𝑝𝑇(𝜆)0,22𝜆𝑓2𝑝(𝜆)+𝑇23𝑓3𝑝(𝜆)++𝑇2𝑘𝑓𝑘𝑝𝑇0,33𝜆𝑓3𝑝(𝜆)++𝑇3𝑘𝑓𝑘𝑝𝑇(𝜆)0,𝑘1𝑘1𝜆𝑓𝑝𝑘1(𝜆)+𝑇𝑘1𝑘𝑓𝑘𝑝(𝑇𝜆)0,𝑘𝑘𝜆𝑓𝑘𝑝(𝜆)0.(3.48) Since 𝑇𝑚𝑘𝑘=0,𝜆𝑇𝑚1𝑘𝑘𝑓𝑘𝑝(𝜆)0 and hence 𝑇𝑚1𝑘𝑘𝑓𝑘𝑝(𝜆)0 if 𝜆0. Since (𝑇𝑘𝑘)𝜆𝑓𝑘𝑝(𝜆)0 from (3.48) 𝜆𝑇𝑚2𝑘𝑘𝑓𝑘𝑝(𝜆)0. By the same reason, 𝑇𝑚3𝑘𝑘𝑓𝑘𝑝(𝜆)0. By repeating this procedure, we finally achieve 𝑓𝑘𝑝(𝜆)0,(3.49) uniformly on 𝐾. Then we obtain the following equation: (𝑇𝑘1𝑘1)𝜆𝑓𝑝𝑘10 uniformly on every compact 𝐾. Since 𝑇𝑘1𝑘1 has Bishop’s property (𝛽) from Corollary 3.29, 𝑓𝑝𝑘1(𝜆)0 uniformly on 𝐾. By repeating this process we prove that 𝑓1𝑝(𝜆)0 uniformly on 𝐾.
Hence {𝑓𝑝=𝑓1𝑝𝑓2𝑝𝑓𝑘𝑝} converge uniformly to 0 on any compact subset 𝐾 of 𝐷, and so 𝑇 has the Bishop’s property (𝛽).

Proposition 3.31. Let 𝑇 be as in Proposition 3.30. Then if 𝑇𝑖𝑖 has Bishop’s property (𝛽) for 𝑖=1,,𝑘, then 𝑇 has Bishop’s property (𝛽).

Proof. The proof is identical to the proof of Proposition 3.30.

Proposition 3.32. Let 𝑇(𝑘𝑖=1𝐻) be the following 𝑘×𝑘 triangular operator matrix: 𝑇𝑇=11𝑇12𝑇13𝑇1𝑘0𝑇22𝑇23𝑇2𝑘00𝑇33𝑇3𝑘0000𝑇𝑘𝑘.(3.50) Assume that 𝑇 is of class [2QN] and 𝜎(𝑇)(𝜎(𝑇))=. If 𝑇11𝑇𝑖𝑗=𝑇𝑖𝑗𝑇𝑗𝑗 for 𝑖=1,𝑗 and 𝑗=1,2,,𝑘, then 𝑇𝑗𝑗 has Bishop’s property (𝛽) for 𝑗=1,2,,𝑘.

Proof. Let 𝑓𝑗𝑝𝐷𝐻 be a sequence of analytic functions such that (𝑇𝑗𝑗)𝜆𝑓𝑗𝑝0 uniformly on every compact subset 𝐾 of 𝐷, then we have for 𝑗=1,2,,𝑘𝑇𝜆𝑇𝑖𝑗𝑓𝑗𝑝=𝑇(𝜆)011𝜆𝑇𝑖𝑗𝑓𝑗𝑝(𝜆)00=𝑇𝑖𝑗𝑇𝑗𝑗𝜆𝑓𝑗𝑝(𝜆)00,(3.51) for 𝑖=1,2,,𝑗. Since 𝑇 has Bishop’s property (𝛽), we get that 𝑇𝑖𝑗𝑓𝑗𝑝(𝜆)0 uniformly on 𝐾 for 𝑗=1,2,,𝑘. We have 𝑇𝜆𝑓1𝑝=𝑇011𝜆𝑓1𝑝(𝜆)00,(3.52) and for 𝑗=2,3,,𝑘, 𝑇𝜆0𝑓𝑗𝑝=0𝑇1𝑗𝑓𝑗𝑝(𝜆)𝑇𝑗1𝑗𝑓𝑗𝑝𝑇(𝜆)𝑗𝑗𝜆𝑓𝑗𝑝(𝜆)000.(3.53) Since 𝑇 has Bishop’s property (𝛽),𝑓𝑗𝑝(𝜆)0 uniformly on 𝐾 for 𝑗=1,2,,𝑘. Thus, 𝑇𝑗𝑗 has Bishop’s property (𝛽).

Proposition 3.33. Let 𝑇 be as in Proposition 3.31. Then if 𝑇 has Bishop’s property (𝛽) and 𝑇11𝑇𝑖𝑗=𝑇𝑖𝑗𝑇𝑗𝑗 for 𝑗=1,2,,𝑘 and 𝑗=1,2,,𝑘, then 𝑇𝑗𝑗 has Bishop's property (𝛽) for 𝑗=1,2,,𝑘.

Proof. The proof is identical to the proof of Proposition 3.32.

4. Berberian Extension

Denote by (𝐻) the space of all sequences (𝑥𝑛)𝑛, with 𝑥𝑛𝐻,𝑛=1,2,3, such that 𝑥𝑛 is bounded. Let 𝑐0(𝐻) denote the subspace of all null sequences of 𝐻 (those such that 𝑥𝑛0). If we set (𝑥𝑛)=sup𝑥𝑛 for every sequence (𝑥𝑛), this defines a seminorm on 𝑙(𝐻), which is zero exactly on the elements of 𝑐0(𝐻). By means of the space (𝐻) and the Banach limits, Berberian [20] constructed an extension 𝐻 of 𝐻 and obtained a homomorphism form operators 𝑇(𝐻) to operators 𝑇(𝐻) such that 𝑇 is an extension of 𝑇.

Theorem 4.1 (Berberian extention [20]). Let 𝐻 be a complex Hilbert space. Then there exists a Hilbert space 𝐻𝐻 and a map Φ(𝐻)(𝐻)𝑇𝑇,(4.1) satisfying: Φ which is an *-isometric isomorphism preserving the order such that(1)(𝑇)=(𝑇), (2)(𝜆𝑇+𝜇𝑆)=𝜆𝑇+𝜇𝑆0, (3)(𝐼𝐻)=𝐼𝐻, (4)(𝑇𝑆)=𝑇𝑆, (5)𝑇=𝑇, (6)𝑇𝑆𝑖𝑓𝑇𝑆, (7)𝜎(𝑇)=𝜎(𝑇),𝜎𝑎𝑝(𝑇)=𝜎𝑎𝑝(𝑇)=𝜎𝑝(𝑇), (8)if 𝑇 is a positive operator, then (𝑇𝛼)=|𝑇|𝛼forall𝛼>0.

An operator is said to be reducible if it has a nontrivial reducing subspace. If an operator is not reducible, then it is called irreducible.

Proposition 4.2 (see [21]). If 𝑇 is an irreducible operator, then 𝑇 is an irreducible operator.

Lemma 4.3. Let 𝐷 be a subset of , 𝑧0,𝑅>0, such that 𝐵(𝑧0,𝑅)={𝑧;|𝑧𝑧0|𝑅}𝐷, let 𝑔𝑛𝐷𝐻 be a sequence of analytic functions, and let the Taylor expansion of 𝑔𝑛 be 𝑔𝑛(𝑧)=𝑘=0𝑎𝑛𝑘𝑧𝑧0𝑘,||𝑧𝑧0||<𝑅.(4.2) If 𝑔𝑛 is uniformly bounded on 𝐵(𝑧0,𝑅)(i.e.,𝑀=sup𝑛1𝑔𝑛(𝑧)𝐵(𝑧0,𝑅)<), then 𝑔𝑛(𝑧)𝑔𝑛𝑧0𝑀𝑟𝑅𝑟,𝑧𝐵(𝑧0,𝑟),0<𝑟<𝑅.(4.3)

Proof. For all 𝑛 and 𝑧𝐵(𝑧0,𝑟) with 0<𝑟<𝑅, by Cauchy’s integral formula, we get the following inequality: 𝑔𝑛(𝑧)𝑔𝑛𝑧0=12𝑖𝜋|𝑢𝑧0|=𝑅𝑔𝑛(𝑢)1𝑢𝑧𝑑𝑢2𝑖𝜋|𝑢𝑧0|=𝑅𝑔𝑛(𝑢)𝑢𝑧01𝑑𝑢2𝜋|𝑢𝑧0|=𝑅||𝑧𝑧0||𝑔𝑛(𝑢)||𝑢𝑧𝑢𝑧0||||||𝑑𝑢𝑀𝑟.𝑅𝑟(4.4)

Remark 4.4. Let 𝐷 be an open subset of . A sequence of analytic functions 𝑔n𝐷𝐻 converges uniformly to 0 on every compact subset 𝐾 of 𝐷 if and only if for any 𝜖>0 and any 𝑧0𝐷 there exists 𝑟>0 and 𝑛0N such that 𝐵(𝑧0;𝑟)𝐷 and 𝑔𝑛𝐵(𝑧0,𝑟)<𝜖 for all 𝑛>𝑛0.

5. Single-valued Extension Property for 𝑚-Partial Isometries

In this section we examine the properties of SVEP and Bishop’s property (𝛽) for some 𝑚-partial isometries operators by using an approach which is different from that used in Section 3. We recall the definition of an 𝑚-partial isometry given by (1.4) and the operator 𝐵𝑚(𝑇).

Definition 5.1 (see [8]). An operator 𝑇(𝐻) is called an 𝑚-isometry if 𝐵𝑚(𝑇)=𝑚𝑘=0(1)𝑘𝑚𝑘𝑇𝑚𝑘𝑇𝑚𝑘=0.(5.1)

Remark 5.2. It is easy to see that 𝑇(𝐻) is an 𝑚-partial isometry if and only if 𝐵𝑚(𝑇)𝑥=𝑚𝑘=0(1)𝑘𝑚𝑘𝑇𝑚𝑘𝑇𝑚𝑘(𝑥)=0,𝑥𝑁(𝑇),(5.2) which shows that the class of 𝑚-partial isometries generalizes those of 𝑚-isometries and partial isometries.

Theorem 5.3 (see [10]). If 𝑇(𝐻) is reducible, that is, if it has a nontrivial reducing subspace 𝑁(𝑇), then the following properties are equivalent.(1)𝑇 is an 𝑚-partial isometry.(2)𝑇|𝑁(𝑇) is an 𝑚-isometry.

Proposition 5.4 (see [10]). Let 𝑇 be a reducible 𝑚-partial isometry. Then(1)𝜆𝜎ap(𝑇){0} implies 𝜆𝜎ap(𝑇), that is, if 𝑇𝜆𝑥𝑛0 for some sequence of bounded vectors {𝑥𝑛}𝐻, then (𝑇𝜆)𝑥𝑛0,(2)𝜆𝜎𝑝(𝑇){0} implies 𝜆𝜎𝑝(𝑇),(3)eigenvectors of 𝑇 corresponding to distinct eigenvalues are orthogonal, that is, 𝑁(𝑇𝜆)𝑁(𝑇𝜇) if 𝜆,𝜇𝜎𝑝(𝑇),𝜆𝜇.

Lemma 5.5. Let 𝑇 be a reducible 𝑚-partial isometry and let 𝜆,𝜇𝜎𝑎𝑝(𝑇) and (𝑥𝑛), (𝑦𝑛) be sequences of bounded vectors in 𝐻 such that 𝜆𝜇 and 𝑇𝜆𝑥𝑛𝑇0,𝜇𝑦𝑛0(𝑎𝑠𝑛).(5.3) Then we have 𝑥𝑛𝑦𝑛0(as𝑛).(5.4)

Proof. We may assume that 𝜇0. Then from Proposition 5.4(1) we have (𝑇𝜇)𝑦𝑛0 as 𝑛. Hence, (𝜆𝜇)𝑥𝑛𝑦𝑛=𝑇𝜆𝑥𝑛𝑦𝑛+𝑥𝑛𝑇𝜇𝑦𝑛0,𝑛,(5.5) which implies (5.4) in view of 𝜆𝜇 and the proof is complete.

Theorem 5.6. Any reducible 𝑚-partial isometry has SVEP.

Proof. Let 𝑈 be a bounded subset of and let 𝑓𝑈𝐻 be an analytic function such that 𝑇𝜆𝑓(𝜆)=0for𝜆𝑈.(5.6) Since 𝑁(𝑇𝜆)𝑁(𝑇𝜇),𝜆𝜇 (Proposition 5.4(3)), we have 𝑓(𝜆)2=lim𝜇𝜆𝑓(𝜆)𝑓(𝜇)=0.(5.7) This shows that 𝑓(𝜆)=0.

Lemma 5.7. If 𝑇 is an 𝑚-partial isometry, then 𝑇𝑜 is also an 𝑚-partial isometry.

Proof. It is a consequence of the properties of 𝑇𝑜 (see Theorem 4.1).

Theorem 5.8. If 𝑇 is an 𝑚-partial isometry with a nontrivial reducing space 𝑁(𝑇), then 𝑇𝑜 has the single-valued extension property (SVEP).

Proof. To prove that 𝑇 has SVEP, let 𝜆𝜎𝑎𝑝(𝑇){0}. Since 𝜎𝑎𝑝(𝑇)=𝜎𝑎𝑝(𝑇) by Theorem 4.1, 𝜆𝜎𝑎𝑝(𝑇)=𝜎𝑎𝑝((𝑇)). In particular if 𝜆𝜎𝑝(𝑇), then 𝜆𝜎𝑝((𝑇)). Hence, 𝑁((𝑇𝜆))𝑁((𝑇𝜇)) for 𝜆 and 𝜇𝜎𝑝(𝑇) with 𝜆𝜇. In a similar way as in the proof of Theorem 5.6, we can see that 𝑇 has SVEP.

Acknowledgment

The author would like to express their gratitude to the referees for their helpful and many valuable suggestions.